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

The prescribed Ricci curvature problem for naturally reductive metrics on non-compact simple Lie groups

Romina M. Arroyo  
[email protected]
FaMAF &\& CIEM, Universidad Nacional de Córdoba, Córdoba, ArgentinaSchool of Mathematics and Physics, The University of Queensland, St Lucia, QLD, AustraliaResearch supported by the Australian Government through the Australian Research Council’s Discovery Projects funding scheme (DP180102185).
   Mark D. Gould22footnotemark: 2 33footnotemark: 3
[email protected]
   Artem Pulemotov22footnotemark: 2 33footnotemark: 3
[email protected]
Abstract

We investigate the prescribed Ricci curvature problem in the class of left-invariant naturally reductive Riemannian metrics on a non-compact simple Lie group. We obtain a number of conditions for the solvability of the underlying equations and discuss several examples.

1 Introduction

The study of the prescribed Ricci curvature problem is an important part of modern geometry with ties to flows, relativity and other subjects. The first wave of interest in this problem came in the 1980s; see [Bes87, Chapter 5] and [Aub98, Section 6.5]. Particularly extensive contributions were made at that time by DeTurck and his collaborators. For a discussion of the subsequent advances, including the recent progress in the framework on homogeneous spaces, see the survey [BP19].

Let MM be a smooth manifold. In its original interpretation, the prescribed Ricci curvature problem comes down to the equation

Ric(g)=T,\displaystyle\operatorname{Ric}(g)=T, (1.1)

where the Riemannian metric gg on MM is the unknown and the (0,2)-tensor field TT is given. The paper [DeT81] proved, for nondegenerate TT, the existence of gg satisfying this equation in a neighbourhood of a point on MM; see also [DG99, Pul13, Pul16a]. However, subsequent research into the solvability of (1.1) on all of MM revealed the need for a more nuanced interpretation of the prescribed Ricci curvature problem. Specifically, suppose MM is closed and TT is positive-definite. The results of [Ham84, DeT85, Delan03] and other papers suggest that there exists at most one cc\in\mathbb{R} such that the equation

Ric(g)=cT\displaystyle\operatorname{Ric}(g)=cT (1.2)

can be solved for gg on all of MM. This is certainly the case if MM is the 2-dimensional sphere and TT is positive-definite; see [WW70, DeT85, Ham84]. Thus, on a closed manifold, one customarily interprets the prescribed Ricci curvature problem as the question of finding gg and cc such that (1.2) holds. This paradigm was originally proposed by DeTurck and Hamilton in [Ham84, DeT85]. As it turns out, (1.2) arises in applications, such as the construction of the Ricci iteration; see [PR19, BPRZ19] and also [BP19, Section 3.10]. On the other hand, if MM is open, it may be possible to obtain compelling existence theorems for (1.1) without the additional constant cc. We refer to [Delay02, Delay18] for examples of such theorems.

In recent years, the third-named author and his collaborators produced a series of results, surveyed in [BP19], on the prescribed Ricci curvature problem in the class of homogeneous metrics. More precisely, suppose GG is a connected Lie group. Let MM be a homogeneous space with respect to GG. Assume that the metric gg and the tensor field TT are GG-invariant. Then (1.1) reduces to an overdetermined system of algebraic equations, whereas (1.2) reduces to a determined one. For compact MM and positive-semidefinite TT, the third-named author showed in [Pul16b] that homogeneous metrics satisfying (1.2) are, up to scaling, critical points of the scalar curvature functional subject to the constraint trgT=1\operatorname{tr}_{g}T=1. This observation led to the discovery of several sufficient conditions for the solvability of (1.2) in [Pul16b, GP17, Pul20]. It parallels the well-known variational approach to the Einstein equation; see, e.g., [WZ86, §1]. In the case where MM is compact but TT is not positive-semidefinite, the question of solvability of (1.2) remains largely open. We hope that the present paper will stimulate its investigation; see Remark 5.1.

As for the prescribed Ricci curvature problem for homogeneous metrics on non-compact spaces, progress has been scarce so far. Buttsworth conducted in [But19] a comprehensive study of this problem on unimodular three-dimensional Lie groups. In most of the cases he considered, there is at most one constant cc\in\mathbb{R} such that a metric gg satisfying (1.2) exists. Several questions related to, but distinct from, the solvability of (1.1) and (1.2) on non-compact Lie groups have been studied by Milnor, Kowalski–Nikcevic, Eberlein, Kremlev–Nikonorov, Ha–Lee, Pina–dos Santos, the first-named author in collaboration with Lafuente (forthcoming work), and others. For a discussion of those results and a collection of references, see [BP19, Sections 2 and 4.1].

Left-invariant naturally reductive metrics on a Lie group form an important family nested between the set of all left-invariant metrics and the set of bi-invariant ones. The investigation of this family has led to several significant advances in geometry. For instance, it yielded new solutions to the Einstein equation and new insights into the spectral theory of the Laplacian on manifolds; see [DZ79, GS10, Lau19]. In the recent paper [APZ20], Ziller and two of the authors studied (1.2) for naturally reductive metrics on a compact Lie group using variational methods. That work exposed several interesting and previously unseen patterns of behaviour of the scalar curvature functional. For instance, one of the main theorems of [APZ20] is a necessary condition for the existence of a critical point subject to the constraint trgT=1\operatorname{tr}_{g}T=1. No results of this kind had appeared in the literature before.

The present paper studies the prescribed Ricci curvature problem for naturally reductive metrics on a non-compact Lie group GG. We assume that GG is simple. The more general case of semisimple GG seems to be much more difficult analytically—we intend to consider it elsewhere. Since GG is an open manifold, it is reasonable for us to view the prescribed Ricci curvature problem as the question of finding solutions to (1.1). On the other hand, the fact that naturally reductive metrics are homogeneous suggests that a “better” interpretation of this problem may be given by (1.2). The present paper studies both equations. We show that (1.1) reduces to an overdetermined algebraic system. Even so, we are able to obtain a comprehensive existence and uniqueness theorem. Equation (1.2) reduces to a determined system. In order to find conditions for solvability, we characterise metrics satisfying (1.2) as critical points of the scalar curvature functional subject to one of three TT-dependent constraints. While this characterisation is similar in spirit to the one obtained for compact Lie groups in [APZ20], it bears some conceptual distinctions and requires a different proof. We obtain existence theorems for global maxima and classify some of the other critical points. The development and application of our variational methods presents many interesting analytical challenges and provides a wealth of insight for the investigation of (1.2) on compact homogeneous spaces in the case of mixed-signature TT (see, e.g., Remark 5.1).

The paper is organised as follows. In Section 2, we recall the characterisation, originally obtained by Gordon, of naturally reductive metrics on a non-compact simple Lie group. This characterisation underpins all of our results. In Section 3, we compute the Ricci curvature of a naturally reductive metric on GG. We believe that the formulas we obtain are of independent interest. Section 4 is devoted to equation (1.1). We produce a necessary and sufficient condition for the existence of a solution. We also establish uniqueness up to scaling. Section 5 focuses on (1.2). We develop the variational approach to this equation and describe several types of critical points of the scalar curvature functional. At the end, we summarise the implications for the existence and the number of solutions. Section 6 examines the case where the metrics we consider are naturally reductive with respect to G×KG\times K for a simple subgroup K<GK<G. Here, we find conditions for the solvability of (1.2) that are both necessary and sufficient. We also determine the precise number of solutions. Finally, Section 7 offers a series of examples.

2 Naturally reductive metrics on non-compact simple Lie groups

Consider a connected non-compact simple Lie group GG with Lie algebra 𝔤\mathfrak{g}. The results of [Gor85, Section 5] yield a convenient characterisation of left-invariant naturally reductive metrics on GG. We present this characterisation in Theorem 2.1 below. For the basic theory of naturally reductive metrics, see [DZ79, Section 1] and [Gor85, Section 2]. In what follows, we identify every left-invariant (0,2)-tensor field on GG with the bilinear form it induces on 𝔤\mathfrak{g}.

Let KK be a maximal compact subgroup of GG with Lie algebra 𝔨\mathfrak{k}. Suppose BB is the Killing form of 𝔤\mathfrak{g}. Denote by 𝔭\mathfrak{p} the BB-orthogonal complement of 𝔨\mathfrak{k} in 𝔤\mathfrak{g}. Then

𝔤=𝔭𝔨\displaystyle\mathfrak{g}=\mathfrak{p}\oplus\mathfrak{k}

is a Cartan decomposition. We have the inclusions

[𝔨,𝔨]𝔨,[𝔨,𝔭]𝔭,[𝔭,𝔭]𝔨.\displaystyle[\mathfrak{k},\mathfrak{k}]\subset\mathfrak{k},\qquad[\mathfrak{k},\mathfrak{p}]\subset\mathfrak{p},\qquad[\mathfrak{p},\mathfrak{p}]\subset\mathfrak{k}.

The Killing form BB is positive-definite on 𝔭\mathfrak{p} and negative-definite on 𝔨\mathfrak{k}. Thus,

Q=B|𝔭B|𝔨\displaystyle Q=B|_{\mathfrak{p}}-B|_{\mathfrak{k}}

is an inner product on 𝔤\mathfrak{g}. Clearly, QQ is ad(𝔨)\operatorname{ad}(\mathfrak{k})-invariant, and

Q([X,Y],Z)=Q(X,[Y,Z]),X,Y𝔭,Z𝔨.Q([X,Y],Z)=-Q(X,[Y,Z]),\qquad X,Y\in\mathfrak{p},~{}Z\in\mathfrak{k}. (2.1)

The quotient G/KG/K is a symmetric space. Because GG is simple, this space is irreducible. Consequently, the pair (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}) must appear in Table 3 or 4 of [Bes87, Section 7.H]. Let 𝔨1,,𝔨r\mathfrak{k}_{1},\ldots,\mathfrak{k}_{r} be the simple ideals of [𝔨,𝔨][\mathfrak{k},\mathfrak{k}]. Denote by 𝔨r+1\mathfrak{k}_{r+1} the centre of 𝔨\mathfrak{k}. Then

𝔨=𝔨1𝔨r+s,\displaystyle\mathfrak{k}=\mathfrak{k}_{1}\oplus\cdots\oplus\mathfrak{k}_{r+s}, (2.2)

where s=0s=0 if 𝔨r+1\mathfrak{k}_{r+1} is trivial and s=1s=1 otherwise. Analysing the tables in [Bes87, Section 7.H], we conclude that 𝔨r+1\mathfrak{k}_{r+1} is at most 1-dimensional.

The direct product G×KG\times K acts on GG in accordance with the formula

(x,k)y=xyk1,x,yG,kK.\displaystyle(x,k)\,y=xyk^{-1},\qquad x,y\in G,~{}k\in K.

The isotropy subgroup at the identity element of GG is

{(k,k)G×K|kK}.\displaystyle\{(k,k)\in G\times K\,|\,k\in K\}.

Denote by K\mathcal{M}_{K} the set of left-invariant metrics on GG that are naturally reductive with respect to G×KG\times K and some decomposition of the Lie algebra of G×KG\times K. The main purpose of this paper is to study the prescribed Ricci curvature problem in K\mathcal{M}_{K}. Gordon showed in [Gor85, Section 5] that every left-invariant naturally reductive metric on GG must lie in K\mathcal{M}_{K} for some choice of KK. Moreover, she obtained the following characterisation result.

Theorem 2.1 (Gordon).

A left-invariant metric gg on the simple group GG lies in K\mathcal{M}_{K} if and only if

g=βQ|𝔭+α1Q|𝔨1++αr+sQ|𝔨r+sg=\beta Q|_{\mathfrak{p}}+\alpha_{1}Q|_{\mathfrak{k}_{1}}+\cdots+\alpha_{r+s}Q|_{\mathfrak{k}_{r+s}} (2.3)

for some β,α1,,αr+s>0\beta,\alpha_{1},\ldots,\alpha_{r+s}>0.

Remark 2.2.

In [Gor85], Gordon studied naturally reductive metrics on non-compact homogeneous spaces, not just on Lie groups. She obtained a version of Theorem 2.1 in this more general framework.

3 The Ricci curvature

Our main objective in this section is to produce formulas for the Ricci curvature and the scalar curvature of a metric gg given by (2.3). To do so, we need to introduce an array of constants, κ1,,κr+s\kappa_{1},\ldots,\kappa_{r+s}, associated with the pair (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}). Throughout the paper,

n=dim𝔭,di=dim𝔨i,i=1,,r+s.\displaystyle n=\dim\mathfrak{p},\qquad d_{i}=\dim\mathfrak{k}_{i},\qquad i=1,\ldots,r+s.

As we explained above, dr+1=1d_{r+1}=1 if the centre of 𝔨\mathfrak{k} is non-trivial.

Suppose BiB_{i} is the Killing form of 𝔨i\mathfrak{k}_{i} for i=1,,r+si=1,\ldots,r+s. There exists κi\kappa_{i}\in\mathbb{R} such that

Bi=κiB|𝔨i.\displaystyle B_{i}=\kappa_{i}B|_{\mathfrak{k}_{i}}.

Using the assumption that GG is simple, one can easily check that 0<κi<10<\kappa_{i}<1 for i=1,,ri=1,\ldots,r. If the centre of 𝔨\mathfrak{k} is non-trivial, then κr+1=0\kappa_{r+1}=0.

Next, we state a proposition that provides a way of computing κi\kappa_{i} for a specific pair (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}) and i=1,,ri=1,\ldots,r. In what follows, superscript \mathbb{C} means complexification. Clearly, the algebra 𝔤\mathfrak{g}^{\mathbb{C}} is semisimple. We preserve the notation ad\operatorname{ad} for the adjoint representation of 𝔤\mathfrak{g}^{\mathbb{C}}. Choose Cartan subalgebras HH in 𝔤\mathfrak{g}^{\mathbb{C}} and HiH_{i} in 𝔨i\mathfrak{k}_{i}^{\mathbb{C}}. It is easy to verify that 𝔤\mathfrak{g}^{\mathbb{C}} is a completely reducible 𝔨i\mathfrak{k}_{i}^{\mathbb{C}}-module under the action given by ad\operatorname{ad}. This observation implies that every element of HiH_{i} must be semisimple in 𝔤\mathfrak{g}^{\mathbb{C}}. Consequently, we may assume that HH contains HiH_{i}. Let Φ+\Phi^{+} and Φi+\Phi_{i}^{+} be sets of positive roots of 𝔤\mathfrak{g}^{\mathbb{C}} and 𝔨i\mathfrak{k}_{i}^{\mathbb{C}}. The notation tr\operatorname{tr} stands for the trace of a linear operator.

Proposition 3.1.

Given i=1,,ri=1,\ldots,r and ZHiZ\in H_{i}, the constant κi\kappa_{i} satisfies

νΦi+(ν(Z))2=κiνΦ+(ν(Z))2.\displaystyle\sum_{\nu\in\Phi_{i}^{+}}(\nu(Z))^{2}=\kappa_{i}\sum_{\nu\in\Phi^{+}}(\nu(Z))^{2}.
Proof.

We preserve the notation BB and BiB_{i} for the Killing forms of 𝔤\mathfrak{g}^{\mathbb{C}} and 𝔨i\mathfrak{k}_{i}^{\mathbb{C}}. Because KK is compact, 𝔨i\mathfrak{k}_{i}^{\mathbb{C}} is a simple subalgebra of 𝔤\mathfrak{g}^{\mathbb{C}}. Therefore,

Bi(Z,Z)=κiB(Z,Z).\displaystyle B_{i}(Z,Z)=\kappa_{i}B(Z,Z).

Using basic properties of root systems, we find

Bi(Z,Z)\displaystyle B_{i}(Z,Z) =tr(ad(Z)ad(Z)|𝔨i)=2νΦi+(ν(Z))2,\displaystyle=\operatorname{tr}\!\big{(}\!\operatorname{ad}(Z)\operatorname{ad}(Z)|_{\mathfrak{k}_{i}^{\mathbb{C}}}\big{)}=2\sum_{\nu\in\Phi_{i}^{+}}(\nu(Z))^{2},
B(Z,Z)\displaystyle B(Z,Z) =tr(ad(Z)ad(Z))=2νΦ+(ν(Z))2.\displaystyle=\operatorname{tr}(\operatorname{ad}(Z)\operatorname{ad}(Z))=2\sum_{\nu\in\Phi^{+}}(\nu(Z))^{2}.

Remark 3.2.

The assertion of Proposition 3.1 holds even if GG is not simple but merely semisimple.

Remark 3.3.

One can use properties of Casimir elements to produce another formula for κi\kappa_{i}. More precisely, given i=1,,ri=1,\ldots,r, there exists a decomposition

𝔭=𝔭1i𝔭rii\displaystyle\mathfrak{p}^{\mathbb{C}}=\mathfrak{p}_{1}^{i}\oplus\cdots\oplus\mathfrak{p}_{r_{i}}^{i}

such that every 𝔭ji\mathfrak{p}_{j}^{i} is a non-trivial irreducible 𝔨i\mathfrak{k}_{i}^{\mathbb{C}}-module under the action defined by ad\operatorname{ad}. Let ψji\psi_{j}^{i} be the highest weight of 𝔭ji\mathfrak{p}_{j}^{i}. Denote by ρi\rho_{i} the half-sum of positive roots of 𝔨i\mathfrak{k}_{i}^{\mathbb{C}}. Using classical results on eigenvalues of Casimir elements, one can show that

κi=didi+j=1riBi(ψji,ψji+2ρi)dim𝔭ji,\displaystyle\kappa_{i}=\frac{d_{i}}{d_{i}+\sum_{j=1}^{r_{i}}B_{i}(\psi_{j}^{i},\psi_{j}^{i}+2\rho_{i})\dim\mathfrak{p}_{j}^{i}},

where we preserve the notation BiB_{i} for the bilinear form induced on HiH_{i}^{*} by the Killing form of 𝔨i\mathfrak{k}_{i}^{\mathbb{C}}. Related formulas can be found in Dynkin’s work; see [Dyn57].

Example 3.4.

Assume G=𝖲𝖴(p,q)G=\mathsf{SU}(p,q) and K=𝖲𝖴(p)×U(q)K=\mathsf{SU}(p)\times U(q) with 2pq2\leq p\leq q. Then r=2r=2 and s=1s=1 in the decomposition (2.2). Clearly,

𝔤=𝔰𝔩(p+q,),𝔨1=𝔰𝔩(p,),𝔨2=𝔰𝔩(q,),𝔨3=.\displaystyle\mathfrak{g}^{\mathbb{C}}=\mathfrak{sl}(p+q,\mathbb{C}),\qquad\mathfrak{k}_{1}^{\mathbb{C}}=\mathfrak{sl}(p,\mathbb{C}),\qquad\mathfrak{k}_{2}^{\mathbb{C}}=\mathfrak{sl}(q,\mathbb{C}),\qquad\mathfrak{k}_{3}^{\mathbb{C}}=\mathbb{C}.

Denote by EijE_{i}^{j} the matrix of size (p+q)×(p+q)(p+q)\times(p+q) that has 1 in the (i,j)(i,j)th slot and 0 elsewhere. Suppose

H={i=1p+q1λi(EiiEi+1i+1)|λi},\displaystyle H=\bigg{\{}\sum_{i=1}^{p+q-1}\lambda_{i}\big{(}E_{i}^{i}-E_{i+1}^{i+1}\big{)}\,\bigg{|}\,\lambda_{i}\in\mathbb{C}\bigg{\}}, H1={i=1p1λi(EiiEi+1i+1)|λi},\displaystyle H_{1}=\bigg{\{}\sum_{i=1}^{p-1}\lambda_{i}\big{(}E_{i}^{i}-E_{i+1}^{i+1}\big{)}\,\bigg{|}\,\lambda_{i}\in\mathbb{C}\bigg{\}},
Φ+={ϵiϵj| 1i<jp+q},\displaystyle\Phi^{+}=\{\epsilon_{i}-\epsilon_{j}\,|\,1\leq i<j\leq p+q\}, Φ1+={ϵiϵj| 1i<jp},\displaystyle\Phi_{1}^{+}=\{\epsilon_{i}-\epsilon_{j}\,|\,1\leq i<j\leq p\},

where ϵi\epsilon_{i} is the linear functional on HH such that ϵi(EjjEj+1j+1)\epsilon_{i}\big{(}E_{j}^{j}-E_{j+1}^{j+1}\big{)} is the difference of Kronecker deltas δijδij+1\delta_{i}^{j}-\delta_{i}^{j+1}. Choosing Z=E11E22Z=E_{1}^{1}-E_{2}^{2}, we find

νΦ+(ν(Z))2=\displaystyle\sum_{\nu\in\Phi^{+}}(\nu(Z))^{2}= ((ϵ1ϵ2)(E11E22))2+i=3p+q((ϵ1ϵi)(E11E22))2\displaystyle\big{(}(\epsilon_{1}-\epsilon_{2})\big{(}E_{1}^{1}-E_{2}^{2}\big{)}\big{)}^{2}+\sum_{i=3}^{p+q}\big{(}(\epsilon_{1}-\epsilon_{i})\big{(}E_{1}^{1}-E_{2}^{2}\big{)}\big{)}^{2}
+i=3p+q((ϵ2ϵi)(E11E22))2=2(p+q),\displaystyle+\sum_{i=3}^{p+q}\big{(}(\epsilon_{2}-\epsilon_{i})\big{(}E_{1}^{1}-E_{2}^{2}\big{)}\big{)}^{2}=2(p+q),
νΦ1+(ν(Z))2=\displaystyle\sum_{\nu\in\Phi_{1}^{+}}(\nu(Z))^{2}= 2p.\displaystyle 2p.

Proposition 3.1 implies κ1=pp+q\kappa_{1}=\frac{p}{p+q}. A similar argument with Z=Ep+1p+1Ep+2p+2Z=E_{p+1}^{p+1}-E_{p+2}^{p+2} yields κ2=qp+q\kappa_{2}=\frac{q}{p+q}. Since 𝔨3\mathfrak{k}_{3} is abelian, κ3=0\kappa_{3}=0.

If r=1r=1, one can calculate κi\kappa_{i} using formula (3.3) below; see Examples 7.1 and 7.2.

Remark 3.5.

The work [DZ79] computes a range of constants analogous to κi\kappa_{i} in the framework of compact Lie groups. One can find κi\kappa_{i} for a specific pair (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}) using those results along with duality for symmetric spaces; see, e.g., [Bes87, Sections 7.82–7.83]. In the present paper, we choose a more direct approach.

We are now ready to state the main result of this section.

Theorem 3.6.

Suppose gKg\in\mathcal{M}_{K} is a naturally reductive metric on the simple group GG satisfying (2.3). The Ricci curvature of gg is given by the formulas

Ric(g)|𝔭=i=1r+s(αi2β+1)di(1κi)nQ|𝔭,\displaystyle\operatorname{Ric}(g)|_{\mathfrak{p}}=-\sum_{i=1}^{r+s}\Big{(}\frac{\alpha_{i}}{2\beta}+1\Big{)}\frac{d_{i}(1-\kappa_{i})}{n}Q|_{\mathfrak{p}},
Ric(g)|𝔨j=14(αj2β2(1κj)+κj)Q|𝔨j,\displaystyle\operatorname{Ric}(g)|_{\mathfrak{k}_{j}}=\tfrac{1}{4}\Big{(}\frac{\alpha_{j}^{2}}{\beta^{2}}(1-\kappa_{j})+\kappa_{j}\Big{)}Q|_{\mathfrak{k}_{j}},
Ric(g)(𝔭,𝔨j)=Ric(g)(𝔨j,𝔨k)=0,\displaystyle\operatorname{Ric}(g)(\mathfrak{p},\mathfrak{k}_{j})=\operatorname{Ric}(g)(\mathfrak{k}_{j},\mathfrak{k}_{k})=0, j,k=1,,r+s,jk.\displaystyle j,k=1,\ldots,r+s,~{}j\neq k.

To prove Theorem 3.6, we apply the strategy developed in [DZ79, Section 5]. In what follows, trh\operatorname{tr}_{h} stands for the trace of a bilinear form with respect to an inner product hh. The notation π𝔲\pi_{\mathfrak{u}} is used for the QQ-orthogonal projection onto 𝔲𝔤\mathfrak{u}\subset\mathfrak{g}. For j=1,,r+sj=1,\ldots,r+s, define a bilinear form AjA_{j} on 𝔭\mathfrak{p} by setting

Aj(X,Y)=tr(π𝔨jad(X)ad(Y)).\displaystyle A_{j}(X,Y)=\operatorname{tr}(\pi_{\mathfrak{k}_{j}}\operatorname{ad}(X)\operatorname{ad}(Y)).

Fix QQ-orthonormal bases (vi)i=1n(v_{i})_{i=1}^{n} of 𝔭\mathfrak{p} and (vkj)k=1dj\big{(}v_{k}^{j}\big{)}_{k=1}^{d_{j}} of 𝔨j\mathfrak{k}_{j}. We need the following auxiliary result; cf. [Jen73, pages 609–610] and [DZ79, pages 32–34].

Lemma 3.7.

Given j=1,,r+sj=1,\ldots,r+s,

trQ|𝔭Aj=dj(1κj),i=1r+sAi=12Q|𝔭.\displaystyle\operatorname{tr}_{Q|_{\mathfrak{p}}}A_{j}=d_{j}(1-\kappa_{j}),\qquad\sum_{i=1}^{r+s}A_{i}=\tfrac{1}{2}Q|_{\mathfrak{p}}.
Proof.

Invoking (2.1), we compute

trQ|𝔭Aj\displaystyle\operatorname{tr}_{Q|_{\mathfrak{p}}}A_{j} =i=1nk=1djQ([vi,[vi,vkj]],vkj)=i=1nk=1djQ(vi,[[vi,vkj],vkj])\displaystyle=\sum_{i=1}^{n}\sum_{k=1}^{d_{j}}Q\big{(}[v_{i},[v_{i},v_{k}^{j}]],v_{k}^{j}\big{)}=-\sum_{i=1}^{n}\sum_{k=1}^{d_{j}}Q\big{(}v_{i},[[v_{i},v_{k}^{j}],v_{k}^{j}]\big{)}
=trQ|𝔨jB|𝔨j+k=1djl=1r+sm=1dlQ([vkj,[vkj,vml]],vml).\displaystyle=-\operatorname{tr}_{Q|_{\mathfrak{k}_{j}}}B|_{\mathfrak{k}_{j}}+\sum_{k=1}^{d_{j}}\sum_{l=1}^{r+s}\sum_{m=1}^{d_{l}}Q\big{(}[v_{k}^{j},[v_{k}^{j},v_{m}^{l}]],v_{m}^{l}\big{)}.

Since 𝔨j\mathfrak{k}_{j} is an ideal of 𝔨\mathfrak{k},

l=1r+sm=1dlk=1djQ([vkj,[vkj,vml]],vml)\displaystyle\sum_{l=1}^{r+s}\sum_{m=1}^{d_{l}}\sum_{k=1}^{d_{j}}Q\big{(}[v_{k}^{j},[v_{k}^{j},v_{m}^{l}]],v_{m}^{l}\big{)} =m=1djk=1djQ([vkj,[vkj,vmj]],vmj)=trQ|𝔨jBj=κjtrQ|𝔨jB|𝔨j.\displaystyle=\sum_{m=1}^{d_{j}}\sum_{k=1}^{d_{j}}Q\big{(}[v_{k}^{j},[v_{k}^{j},v_{m}^{j}]],v_{m}^{j}\big{)}=\operatorname{tr}_{Q|_{\mathfrak{k}_{j}}}B_{j}=\kappa_{j}\operatorname{tr}_{Q|_{\mathfrak{k}_{j}}}B|_{\mathfrak{k}_{j}}.

We conclude that

trQ|𝔭Aj=trQ|𝔨jB|𝔨j+κjtrQ|𝔨jB|𝔨j=dj(1κj).\displaystyle\operatorname{tr}_{Q|_{\mathfrak{p}}}A_{j}=-\operatorname{tr}_{Q|_{\mathfrak{k}_{j}}}B|_{\mathfrak{k}_{j}}+\kappa_{j}\operatorname{tr}_{Q|_{\mathfrak{k}_{j}}}B|_{\mathfrak{k}_{j}}=d_{j}(1-\kappa_{j}).

This proves the first equality in the statement of the lemma.

For X,Y𝔭X,Y\in\mathfrak{p},

i=1r+sAi(X,Y)\displaystyle\sum_{i=1}^{r+s}A_{i}(X,Y) =i=1r+sk=1diQ([X,[Y,vki]],vki)=B(X,Y)l=1nQ([X,[Y,vl]],vl).\displaystyle=\sum_{i=1}^{r+s}\sum_{k=1}^{d_{i}}Q\big{(}[X,[Y,v_{k}^{i}]],v_{k}^{i}\big{)}=B(X,Y)-\sum_{l=1}^{n}Q([X,[Y,v_{l}]],v_{l}).

It is easy to see that the forms AiA_{i} are symmetric. Consequently,

l=1nQ([X,[Y,vl]],vl)=tr(ad(X)π𝔨ad(Y))=tr(π𝔨ad(Y)ad(X))=i=1r+sAi(X,Y).\displaystyle\sum_{l=1}^{n}Q([X,[Y,v_{l}]],v_{l})=\operatorname{tr}(\operatorname{ad}(X)\pi_{\mathfrak{k}}\operatorname{ad}(Y))=\operatorname{tr}(\pi_{\mathfrak{k}}\operatorname{ad}(Y)\operatorname{ad}(X))=\sum_{i=1}^{r+s}A_{i}(X,Y).

We conclude that

i=1r+sAi(X,Y)=B(X,Y)i=1r+sAi(X,Y),\displaystyle\sum_{i=1}^{r+s}A_{i}(X,Y)=B(X,Y)-\sum_{i=1}^{r+s}A_{i}(X,Y),

which implies the second inequality. ∎

Proof of Theorem 3.6.

Let \nabla be the Levi-Civita connection of the metric gg. The Koszul formula yields

XY={12[X,Y]ifX,Y𝔭orX,Y𝔨,αi2β[X,Y]ifX𝔭andY𝔨ifor somei=1,,r+s,(αi2β+1)[X,Y]ifX𝔨ifor somei=1,,r+sandY𝔭;\displaystyle\nabla_{X}Y=\begin{cases}\tfrac{1}{2}[X,Y]&\mbox{if}~{}X,Y\in\mathfrak{p}~{}\mbox{or}~{}X,Y\in\mathfrak{k},\\ -\frac{\alpha_{i}}{2\beta}[X,Y]&\mbox{if}~{}X\in\mathfrak{p}~{}\mbox{and}~{}Y\in\mathfrak{k}_{i}~{}\mbox{for some}~{}i=1,\ldots,r+s,\\ \big{(}\frac{\alpha_{i}}{2\beta}+1\big{)}[X,Y]&\mbox{if}~{}X\in\mathfrak{k}_{i}~{}\mbox{for some}~{}i=1,\ldots,r+s~{}\mbox{and}~{}Y\in\mathfrak{p};\end{cases} (3.1)

see [Gor85, page 485]. The Ricci curvature of gg satisfies

Ric(g)(X,Y)=trYX,X,Y𝔤.\displaystyle\operatorname{Ric}(g)(X,Y)=-\operatorname{tr}\nabla_{\nabla_{\cdot}Y}X,\qquad X,Y\in\mathfrak{g}. (3.2)

This fact goes back to [Sag70]; a simpler proof appeared in [DZ79, Section 5]. Substituting (3.1) into (3.2), we easily obtain the required identities for Ric(g)|𝔨j\operatorname{Ric}(g)|_{\mathfrak{k}_{j}}, Ric(g)(𝔭,𝔨j)\operatorname{Ric}(g)(\mathfrak{p},\mathfrak{k}_{j}) and Ric(g)(𝔨j,𝔨k)\operatorname{Ric}(g)(\mathfrak{k}_{j},\mathfrak{k}_{k}) with jkj\neq k; cf. [DZ79, page 33].

Because gg lies in K\mathcal{M}_{K}, it is ad(𝔨)\operatorname{ad}(\mathfrak{k})-invariant. Consequently, there exists τ\tau\in\mathbb{R} such that Ric(g)|𝔭=τQ|𝔭\operatorname{Ric}(g)|_{\mathfrak{p}}=\tau Q|_{\mathfrak{p}}. Taking the trace with respect to Q|𝔭Q|_{\mathfrak{p}} on both sides and exploiting Lemma 3.7, we find

nτ\displaystyle n\tau =i=1nRic(g)(vi,vi)=i=1n(l=1nQ(vlvivi,vl)+j=1r+sm=1djQ(vmjvivi,vmj))\displaystyle=\sum_{i=1}^{n}\operatorname{Ric}(g)(v_{i},v_{i})=-\sum_{i=1}^{n}\bigg{(}\sum_{l=1}^{n}Q(\nabla_{\nabla_{v_{l}}v_{i}}v_{i},v_{l})+\sum_{j=1}^{r+s}\sum_{m=1}^{d_{j}}Q\big{(}\nabla_{\nabla_{v_{m}^{j}}v_{i}}v_{i},v_{m}^{j}\big{)}\bigg{)}
=12i=1nj=1r+s(αj2β+1)m=1dj(l=1nQ([vl,vi],vmj)Q([vmj,vi],vl)+Q([[vmj,vi],vi],vmj))\displaystyle=-\tfrac{1}{2}\sum_{i=1}^{n}\sum_{j=1}^{r+s}\Big{(}\frac{\alpha_{j}}{2\beta}+1\Big{)}\sum_{m=1}^{d_{j}}\bigg{(}\sum_{l=1}^{n}Q\big{(}[v_{l},v_{i}],v_{m}^{j}\big{)}Q\big{(}[v_{m}^{j},v_{i}],v_{l}\big{)}+Q\big{(}[[v_{m}^{j},v_{i}],v_{i}],v_{m}^{j}\big{)}\bigg{)}
=i=1nj=1r+s(αj2β+1)m=1djQ([vi,[vi,vmj]],vmj)\displaystyle=-\sum_{i=1}^{n}\sum_{j=1}^{r+s}\Big{(}\frac{\alpha_{j}}{2\beta}+1\Big{)}\sum_{m=1}^{d_{j}}Q\big{(}[v_{i},[v_{i},v_{m}^{j}]],v_{m}^{j}\big{)}
=j=1r+s(αj2β+1)trQ|𝔭Aj=j=1r+s(αj2β+1)dj(1κj).\displaystyle=-\sum_{j=1}^{r+s}\Big{(}\frac{\alpha_{j}}{2\beta}+1\Big{)}\operatorname{tr}_{Q|_{\mathfrak{p}}}A_{j}=-\sum_{j=1}^{r+s}\Big{(}\frac{\alpha_{j}}{2\beta}+1\Big{)}d_{j}(1-\kappa_{j}).

Therefore,

τ=j=1r+s(αj2β+1)dj(1κj)n,\displaystyle\tau=-\sum_{j=1}^{r+s}\Big{(}\frac{\alpha_{j}}{2\beta}+1\Big{)}\frac{d_{j}(1-\kappa_{j})}{n},

which yields the required identity for Ric(g)|𝔭\operatorname{Ric}(g)|_{\mathfrak{p}}. ∎

Denote by SS the scalar curvature functional on K\mathcal{M}_{K}. Our next goal is to produce a formula for SS. Taking the trace of the second equality in Lemma 3.7 and using the first one, we obtain

2i=1r+sdi(1κi)=n.\displaystyle 2\sum_{i=1}^{r+s}d_{i}(1-\kappa_{i})=n. (3.3)

Theorem 3.6 and (3.3) imply the following result.

Corollary 3.8.

Suppose gKg\in\mathcal{M}_{K} is a naturally reductive metric on the simple group GG satisfying (2.3). The scalar curvature of gg is given by the formula

S(g)=14i=1r+sαiβ2di(1κi)n2β+14i=1rdiκiαi.S(g)=-\tfrac{1}{4}\sum_{i=1}^{r+s}\frac{\alpha_{i}}{\beta^{2}}d_{i}(1-\kappa_{i})-\frac{n}{2\beta}+\tfrac{1}{4}\sum_{i=1}^{r}\frac{d_{i}\kappa_{i}}{\alpha_{i}}.

4 Metrics with prescribed Ricci curvature

Consider a (0,2)-tensor field TT on the simple group GG. In this section, we state a necessary and sufficient condition for the solvability of the equation

Ric(g)=T\displaystyle\operatorname{Ric}(g)=T (4.1)

in the class K\mathcal{M}_{K}. If a metric gKg\in\mathcal{M}_{K} satisfying (4.1) exists, then TT must be left-invariant. Moreover, by Theorem 3.6, the formula

T=T𝔭Q|𝔭+T1Q|𝔨1++Tr+sQ|𝔨r+sT=-T_{\mathfrak{p}}Q|_{\mathfrak{p}}+T_{1}Q|_{\mathfrak{k}_{1}}+\ldots+T_{r+s}Q|_{\mathfrak{k}_{r+s}} (4.2)

holds with T𝔭,T1,,Tr+s>0T_{\mathfrak{p}},T_{1},\ldots,T_{r+s}>0.

Theorem 4.1.

Suppose TT is a left-invariant (0,2)-tensor field on GG given by (4.2). A naturally reductive metric gKg\in\mathcal{M}_{K} satisfying (4.1) exists if and only if

4Tiκi>0\displaystyle 4T_{i}-\kappa_{i}>0 (4.3)

for all i=1,,ri=1,\ldots,r and

T𝔭=i=1r+s2di(1κi)+di(4Tiκi)(1κi)2n.\displaystyle T_{\mathfrak{p}}=\sum_{i=1}^{r+s}\frac{2d_{i}(1-\kappa_{i})+d_{i}\sqrt{(4T_{i}-\kappa_{i})(1-\kappa_{i})}}{2n}. (4.4)

There is at most one such gg, up to scaling.

Proof.

Consider a naturally reductive metric gKg\in\mathcal{M}_{K}. It satisfies (2.3) for some β,α1,,αr+s>0\beta,\alpha_{1},\ldots,\alpha_{r+s}>0. By Theorem 3.6, the Ricci curvature of gg equals TT if and only if

αiβ\displaystyle\frac{\alpha_{i}}{\beta} =4Tiκi1κi,1ir+s,\displaystyle=\sqrt{\frac{4T_{i}-\kappa_{i}}{1-\kappa_{i}}},\qquad 1\leq i\leq r+s,
T𝔭\displaystyle T_{\mathfrak{p}} =i=1r+s(1+αi2β)(1κi)din=i=1r+s2di(1κi)+di(4Tiκi)(1κi)2n.\displaystyle=\sum_{i=1}^{r+s}\Big{(}1+\frac{\alpha_{i}}{2\beta}\Big{)}\frac{(1-\kappa_{i})d_{i}}{n}=\sum_{i=1}^{r+s}\frac{2d_{i}(1-\kappa_{i})+d_{i}\sqrt{(4T_{i}-\kappa_{i})(1-\kappa_{i})}}{2n}.

This observation proves the result. ∎

It is tempting to use Theorem 4.1 to study the solvability of the equation

Ric(g)=cT\operatorname{Ric}(g)=cT (4.5)

in the class K\mathcal{M}_{K}. Indeed, suppose Ξ\Xi is the set of left-invariant tensor fields on GG satisfying (4.2), (4.3) and (4.4). Theorem 4.1 states that (4.1) has a solution if and only if TT lies in Ξ\Xi. Using this result, we can easily obtain a description of the set of tensor fields that coincide with Ricci curvatures of metrics in K\mathcal{M}_{K} up to scaling. Namely, a pair (g,c)K×(0,)(g,c)\in\mathcal{M}_{K}\times(0,\infty) satisfying (4.5) exists if and only if

TΞ={τh|τ>0andhΞ}.\displaystyle T\in\Xi^{\prime}=\{\tau h\,|\,\tau>0~{}\mbox{and}~{}h\in\Xi\}.

However, in general, it is difficult to determine whether a specific TT given by (4.2) lies in Ξ\Xi^{\prime}. To do so, one has to check whether (4.3) and (4.4) hold with T𝔭,T1,,Tr+sT_{\mathfrak{p}},T_{1},\ldots,T_{r+s} replaced by cT𝔭,cT1,,cTr+scT_{\mathfrak{p}},cT_{1},\ldots,cT_{r+s} for some c>0c>0. Already when r+s=2r+s=2, this involves the tricky task of understanding if a polynomial of degree 44 has roots that obey several constraints; when r+s3r+s\geq 3, the question seems to be substantially harder. In the present paper, we take a different approach to the analysis of (4.5). We are able to show that the existence of a pair (g,c)K×(0,)(g,c)\in\mathcal{M}_{K}\times(0,\infty) satisfying (4.5) follows from simple inequalities for the components of TT. Moreover, we draw interesting conclusions regarding the non-uniqueness of such a pair.

5 Metrics with Ricci curvature prescribed up to scaling

Suppose TT is a left-invariant symmetric (0,2)-tensor field on GG. Our next goal is to study the solvability of equation (4.5) in the class K\mathcal{M}_{K}. As above, we assume the group GG is simple. This implies, in particular, that κi<1\kappa_{i}<1 for all ii.

First, we re-state the problem in variational terms. More precisely, define

T+\displaystyle\mathcal{M}^{+}_{T} ={gK|trgT=1},\displaystyle=\{g\in\mathcal{M}_{K}\,|\,\operatorname{tr}_{g}T=1\},
T\displaystyle\mathcal{M}^{-}_{T} ={gK|trgT=1},\displaystyle=\{g\in\mathcal{M}_{K}\,|\,\operatorname{tr}_{g}T=-1\},
T0\displaystyle\mathcal{M}^{0}_{T} ={gK|trgT=0}.\displaystyle=\{g\in\mathcal{M}_{K}\,|\,\operatorname{tr}_{g}T=0\}. (5.1)

Each of these three spaces carries a manifold structure induced from K\mathcal{M}_{K}. In Section 5.1, we show that gg satisfies (4.5) if and only if it is (up to scaling) a critical point of S|T+S|_{\mathcal{M}^{+}_{T}}, S|TS|_{\mathcal{M}^{-}_{T}} or S|T0S|_{\mathcal{M}^{0}_{T}}. This resembles the variational interpretation of (4.5) on compact Lie groups for positive-semidefinite TT (see [APZ20, Proposition 3.1]); however, in that case, only S|T+S|_{\mathcal{M}^{+}_{T}} needs to be considered. In Sections 5.2 and 5.3, we obtain sufficient conditions for the existence of global maxima of S|T+S|_{\mathcal{M}^{+}_{T}} and S|TS|_{\mathcal{M}^{-}_{T}}, respectively. This requires complex estimates on the scalar curvature obtained in Lemmas 5.4 and 5.6. In Section 5.4, we classify completely the critical points of S|T0S|_{\mathcal{M}^{0}_{T}} assuming r+s2r+s\leq 2. The analysis here can be reduced, as Lemma 5.9 demonstrates, to the study of a cubic polynomial in one variable. Finally, in Section 5.5, we summarise the implications of our results for the prescribed Ricci curvature problem.

Remark 5.1.

It appears that (4.5) admits a similar variational interpretation on compact homogeneous spaces when TT has mixed signature. Thus, our arguments yield new insight into the study of (4.5) in that setting. For instance, Buttsworth showed in [But19] through methods of elementary polynomial analysis that, for certain TT on 𝖲𝖴(2)\mathsf{SU}(2), a left-invariant metric gg satisfying (4.5) exists for precisely two distinct constants cc\in\mathbb{R}. This was somewhat surprising at the time, as nothing similar had occurred in previously understood examples. We observe an analogous phenomenon in Theorem 5.10 below. In our arguments, the two constants arise naturally as Lagrange multipliers for S|T+S|_{\mathcal{M}^{+}_{T}} and S|TS|_{\mathcal{M}^{-}_{T}}.

5.1 The variational approach

The following result underpins our approach to the study of (4.5).

Proposition 5.2.

Let TT be a left-invariant symmetric (0,2)-tensor field on GG. A metric gKg\in\mathcal{M}_{K} satisfies (4.5) for some cc\in{\mathbb{R}} if and only if it is (up to scaling) a critical point of S|T+S|_{\mathcal{M}^{+}_{T}}, S|TS|_{\mathcal{M}^{-}_{T}} or S|T0S|_{\mathcal{M}^{0}_{T}}.

Proof.

Denote by 𝒯\mathcal{T} the space of left-invariant bilinear form fields

γQ|𝔭+γ1Q|𝔨1++γr+sQ|𝔨r+s\gamma Q|_{\mathfrak{p}}+\gamma_{1}Q|_{\mathfrak{k}_{1}}+\ldots+\gamma_{r+s}Q|_{\mathfrak{k}_{r+s}}

with γ,γ1,,γr+s\gamma,\gamma_{1},\ldots,\gamma_{r+s}\in\mathbb{R}. Theorem 3.6 shows that Ric(g)\operatorname{Ric}(g) lies in 𝒯\mathcal{T}. We identify 𝒯\mathcal{T} with the space tangent to K\mathcal{M}_{K} at gg in the natural way. The left-invariant bilinear form fields

Q𝔭=π𝔭Q,Qi=π𝔨iQ,i=1,,r+s,\displaystyle Q_{\mathfrak{p}}=\pi_{\mathfrak{p}}^{*}Q,\qquad Q_{i}=\pi_{\mathfrak{k}_{i}}^{*}Q,\qquad i=1,\ldots,r+s,

where * denotes pullback, make a basis of 𝒯\mathcal{T}.

Suppose gg is given by (2.3). Using Corollary 3.8, (3.3) and Theorem 3.6, we find that the differential of the scalar curvature functional SS satisfies

dSg(Q𝔭)\displaystyle dS_{g}(Q_{\mathfrak{p}}) =β(14i=1r+sαiβ2di(1κi)n2β)\displaystyle=\frac{\partial}{\partial\beta}\bigg{(}-\tfrac{1}{4}\sum_{i=1}^{r+s}\frac{\alpha_{i}}{\beta^{2}}d_{i}(1-\kappa_{i})-\frac{n}{2\beta}\bigg{)}
=1β2i=1r+s(αi2β+1)di(1κi)=g(Ric(g),Q𝔭),\displaystyle=\frac{1}{\beta^{2}}\sum_{i=1}^{r+s}\Big{(}\frac{\alpha_{i}}{2\beta}+1\Big{)}d_{i}(1-\kappa_{i})=-g(\operatorname{Ric}(g),Q_{\mathfrak{p}}),
dSg(Qj)\displaystyle dS_{g}(Q_{j}) =αj(14i=1r+sαiβ2di(1κi)+14i=1rdiκiαi)\displaystyle=\frac{\partial}{\partial\alpha_{j}}\bigg{(}-\tfrac{1}{4}\sum_{i=1}^{r+s}\frac{\alpha_{i}}{\beta^{2}}d_{i}(1-\kappa_{i})+\tfrac{1}{4}\sum_{i=1}^{r}\frac{d_{i}\kappa_{i}}{\alpha_{i}}\bigg{)}
=dj4αj2(αj24β2(1κj)+κj)=g(Ric(g),Qj),j=1,,r+s.\displaystyle=-\frac{d_{j}}{4\alpha_{j}^{2}}\Big{(}\frac{\alpha_{j}^{2}}{4\beta^{2}}(1-\kappa_{j})+\kappa_{j}\Big{)}=-g(\operatorname{Ric}(g),Q_{j}),\qquad j=1,\ldots,r+s.

(We preserve the notation gg for the inner product induced by gg on the tensor bundle over GG.) Consequently,

dSg(h)=g(Ric(g),h)dS_{g}(h)=-g(\operatorname{Ric}(g),h) (5.2)

for all h𝒯h\in\mathcal{T}.

Let us scale gg by the factor

τ={|trgT|iftrgT0,1iftrgT=0.\displaystyle\tau=\begin{cases}|\operatorname{tr}_{g}T|&\mbox{if}~{}\operatorname{tr}_{g}T\neq 0,\\ 1&\mbox{if}~{}\operatorname{tr}_{g}T=0.\end{cases}

Clearly, τg\tau g lies in Tσ\mathcal{M}_{T}^{\sigma} for some σ{+,,0}\sigma\in\{+,-,0\}. The space tangent to Tσ\mathcal{M}^{\sigma}_{T} at τg\tau g consists of those h𝒯h\in\mathcal{T} that satisfy g(T,h)=0g(T,h)=0. Formula (5.2) implies that dSτgdS_{\tau g} vanishes on this space if and only if gg satisfies (4.5). ∎

By Theorem 3.6, the constant cc in Proposition 5.2 must be positive if TT is given by (4.2) with T𝔭,T1,,Tr+s>0T_{\mathfrak{p}},T_{1},\ldots,T_{r+s}>0. The above proof shows that one may think of cc as a Lagrange multiplier.

5.2 Global maxima on T+\mathcal{M}_{T}^{+}

Our goal in this subsection is to show that simple inequalities for TT guarantee the existence of a critical point of S|T+S|_{\mathcal{M}_{T}^{+}}.

Theorem 5.3.

Suppose TT is a left-invariant (0,2)-tensor field on GG satisfying (4.2) for some T𝔭,T1,,Tr+s>0T_{\mathfrak{p}},T_{1},\ldots,T_{r+s}>0. Choose an index mm such that

κmTm=maxi=1,,rκiTi.\displaystyle\frac{\kappa_{m}}{T_{m}}=\max_{i=1,\ldots,r}\frac{\kappa_{i}}{T_{i}}. (5.3)

If

κmtrQTTm<dim𝔨+dm(1κm)3n\frac{\kappa_{m}\operatorname{tr}_{Q}T}{T_{m}}<\dim\mathfrak{k}+d_{m}(1-\kappa_{m})-3n (5.4)

and

i=1r+sn2κiT𝔭2di2(1κi)Ti2nT𝔭Ti2n<0,\sum_{i=1}^{r+s}\frac{n^{2}\kappa_{i}T_{\mathfrak{p}}^{2}-d_{i}^{2}(1-\kappa_{i})T_{i}^{2}}{nT_{\mathfrak{p}}T_{i}}-2n<0, (5.5)

then the functional S|T+S|_{\mathcal{M}^{+}_{T}} attains its global maximum.

The proof of Theorem 5.3 requires the following estimate for S|T+S|_{\mathcal{M}^{+}_{T}}.

Lemma 5.4.

Let mm be as in (5.3). Assume that (5.5) holds. Given ϵ>0\epsilon>0, there exists a compact set 𝒞ϵ+T+\mathcal{C}_{\epsilon}^{+}\subset\mathcal{M}^{+}_{T} such that

S(g)<κm4Tm+ϵS(g)<\frac{\kappa_{m}}{4T_{m}}+\epsilon (5.6)

for every gT+𝒞ϵ+g\in\mathcal{M}^{+}_{T}\setminus\mathcal{C}_{\epsilon}^{+}.

Proof.

Consider a metric gT+g\in\mathcal{M}^{+}_{T} satisfying (2.3). The definition of T+\mathcal{M}^{+}_{T} implies

trgT=nT𝔭β+i=1r+sdiTiαi=1.\displaystyle\operatorname{tr}_{g}T=-\frac{nT_{\mathfrak{p}}}{\beta}+\sum_{i=1}^{r+s}\frac{d_{i}T_{i}}{\alpha_{i}}=1. (5.7)

For ϵ>0\epsilon>0, denote

Λ+(ϵ)=nκmT𝔭4Tmϵ.\displaystyle\Lambda_{\infty}^{+}(\epsilon)=\frac{n\kappa_{m}T_{\mathfrak{p}}}{4T_{m}\epsilon}.

In view of Corollary 3.8 and formula (5.7), if β>Λ+(ϵ)\beta>\Lambda_{\infty}^{+}(\epsilon), then

S(g)<14i=1rdiκiαiκm4Tmi=1r+sdiTiαi=κm4Tm(1+nT𝔭β)<κm4Tm+ϵ.S(g)<\tfrac{1}{4}\sum_{i=1}^{r}\frac{d_{i}\kappa_{i}}{\alpha_{i}}\leq\frac{\kappa_{m}}{4T_{m}}\sum_{i=1}^{r+s}\frac{d_{i}T_{i}}{\alpha_{i}}=\frac{\kappa_{m}}{4T_{m}}\bigg{(}1+\frac{nT_{\mathfrak{p}}}{\beta}\bigg{)}<\frac{\kappa_{m}}{4T_{m}}+\epsilon.

Thus, in this case, estimate (5.6) holds.

Denote

Λ0+=12(i=1r+sn2κiT𝔭2+di2(1κi)Ti2n2T𝔭2Ti)1|i=1r+sn2κiT𝔭2di2(1κi)Ti2nT𝔭Ti2n|.\displaystyle\Lambda_{0}^{+}=\tfrac{1}{2}\bigg{(}\sum_{i=1}^{r+s}\frac{n^{2}\kappa_{i}T_{\mathfrak{p}}^{2}+d_{i}^{2}(1-\kappa_{i})T_{i}^{2}}{n^{2}T_{\mathfrak{p}}^{2}T_{i}}\bigg{)}^{-1}\bigg{|}\sum_{i=1}^{r+s}\frac{n^{2}\kappa_{i}T_{\mathfrak{p}}^{2}-d_{i}^{2}(1-\kappa_{i})T_{i}^{2}}{nT_{\mathfrak{p}}T_{i}}-2n\bigg{|}.

Formula (5.7) implies

βαj<βdjTji=1r+sdiTiαi=β+nT𝔭djTj,j=1,,r+s.\displaystyle\frac{\beta}{\alpha_{j}}<\frac{\beta}{d_{j}T_{j}}\sum_{i=1}^{r+s}\frac{d_{i}T_{i}}{\alpha_{i}}=\frac{\beta+nT_{\mathfrak{p}}}{d_{j}T_{j}},\qquad j=1,\ldots,r+s.

Invoking Corollary 3.8 again, we find

S(g)\displaystyle S(g) =14β(i=1r+s(diκiβαidi(1κi)αiβ)2n)\displaystyle=\frac{1}{4\beta}\bigg{(}\sum_{i=1}^{r+s}\Big{(}d_{i}\kappa_{i}\frac{\beta}{\alpha_{i}}-d_{i}(1-\kappa_{i})\frac{\alpha_{i}}{\beta}\Big{)}-2n\bigg{)}
<14β(i=1r+s(κiβ+nT𝔭Tidi2(1κi)Tiβ+nT𝔭)2n)\displaystyle<\frac{1}{4\beta}\bigg{(}\sum_{i=1}^{r+s}\bigg{(}\kappa_{i}\frac{\beta+nT_{\mathfrak{p}}}{T_{i}}-\frac{d_{i}^{2}(1-\kappa_{i})T_{i}}{\beta+nT_{\mathfrak{p}}}\bigg{)}-2n\bigg{)}
=14β(i=1r+s(nκiT𝔭Tidi2(1κi)TinT𝔭)2n+βi=1r+s(κiTi+di2(1κi)TinT𝔭(β+nT𝔭)))\displaystyle=\frac{1}{4\beta}\bigg{(}\sum_{i=1}^{r+s}\bigg{(}\frac{n\kappa_{i}T_{\mathfrak{p}}}{T_{i}}-\frac{d_{i}^{2}(1-\kappa_{i})T_{i}}{nT_{\mathfrak{p}}}\bigg{)}-2n+\beta\sum_{i=1}^{r+s}\bigg{(}\frac{\kappa_{i}}{T_{i}}+\frac{d_{i}^{2}(1-\kappa_{i})T_{i}}{nT_{\mathfrak{p}}(\beta+nT_{\mathfrak{p}})}\bigg{)}\bigg{)}
<14β(i=1r+sn2κiT𝔭2di2(1κi)Ti2nT𝔭Ti2n+βi=1r+sn2κiT𝔭2+di2(1κi)Ti2n2T𝔭2Ti).\displaystyle<\frac{1}{4\beta}\bigg{(}\sum_{i=1}^{r+s}\frac{n^{2}\kappa_{i}T_{\mathfrak{p}}^{2}-d_{i}^{2}(1-\kappa_{i})T_{i}^{2}}{nT_{\mathfrak{p}}T_{i}}-2n+\beta\sum_{i=1}^{r+s}\frac{n^{2}\kappa_{i}T_{\mathfrak{p}}^{2}+d_{i}^{2}(1-\kappa_{i})T_{i}^{2}}{n^{2}T_{\mathfrak{p}}^{2}T_{i}}\bigg{)}.

In view of (5.5), if β<Λ0+\beta<\Lambda_{0}^{+}, then

S(g)<18β(i=1r+sn2κiT𝔭2di2(1κi)Ti2nT𝔭Ti2n)<0.\displaystyle S(g)<\frac{1}{8\beta}\bigg{(}\sum_{i=1}^{r+s}\frac{n^{2}\kappa_{i}T_{\mathfrak{p}}^{2}-d_{i}^{2}(1-\kappa_{i})T_{i}^{2}}{nT_{\mathfrak{p}}T_{i}}-2n\bigg{)}<0.

In this case, again, estimate (5.6) holds.

Choose pp and qq such that

dpTp=mini=1,,r+sdiTi,αq=mini=1,,r+sαi.\displaystyle d_{p}T_{p}=\min_{i=1,\ldots,r+s}d_{i}T_{i},\qquad\alpha_{q}=\min_{i=1,\ldots,r+s}\alpha_{i}. (5.8)

Denote

Γ0+=12dpTpmin{1,Λ0+nT𝔭}.\displaystyle\Gamma_{0}^{+}=\tfrac{1}{2}d_{p}T_{p}\min\Big{\{}1,\frac{\Lambda_{0}^{+}}{nT_{\mathfrak{p}}}\Big{\}}.

By (5.7), if αq<Γ0+\alpha_{q}<\Gamma_{0}^{+}, then

β=nT𝔭(i=1r+sdiTiαi1)1<nT𝔭(dqTqαq1)1nT𝔭αqdpTpαq2nT𝔭αqdpTpΛ0+,\displaystyle\beta=nT_{\mathfrak{p}}\bigg{(}\sum_{i=1}^{r+s}\frac{d_{i}T_{i}}{\alpha_{i}}-1\bigg{)}^{-1}<nT_{\mathfrak{p}}\Big{(}\frac{d_{q}T_{q}}{\alpha_{q}}-1\Big{)}^{-1}\leq\frac{nT_{\mathfrak{p}}\alpha_{q}}{d_{p}T_{p}-\alpha_{q}}\leq\frac{2nT_{\mathfrak{p}}\alpha_{q}}{d_{p}T_{p}}\leq\Lambda_{0}^{+},

which means (5.6) holds.

Choose ll such that

αl=maxi=1,,r+sαi.\displaystyle\alpha_{l}=\max_{i=1,\ldots,r+s}\alpha_{i}. (5.9)

For ϵ>0\epsilon>0, denote

Γ+(ϵ)=2Λ+(ϵ)2mini=1,,r+sdi(1κi)i=1rdiκiΓ0+.\displaystyle\Gamma_{\infty}^{+}(\epsilon)=\frac{2\Lambda_{\infty}^{+}(\epsilon)^{2}}{\min_{i=1,\ldots,r+s}d_{i}(1-\kappa_{i})}\sum_{i=1}^{r}\frac{d_{i}\kappa_{i}}{\Gamma_{0}^{+}}.

As we showed above, if β>Λ+(ϵ)\beta>\Lambda_{\infty}^{+}(\epsilon) or αq<Γ0+\alpha_{q}<\Gamma_{0}^{+}, then (5.6) holds. Assuming αqΓ0+\alpha_{q}\geq\Gamma_{0}^{+} and αl>Γ+(ϵ)\alpha_{l}>\Gamma_{\infty}^{+}(\epsilon), we find

S(g)\displaystyle S(g) <14i=1r+sαiβ2di(1κi)+14i=1rκidiαi<αl4β2dl(1κl)+14i=1rκidiαq\displaystyle<-\tfrac{1}{4}\sum_{i=1}^{r+s}\frac{\alpha_{i}}{\beta^{2}}d_{i}(1-\kappa_{i})+\tfrac{1}{4}\sum_{i=1}^{r}\frac{\kappa_{i}d_{i}}{\alpha_{i}}<-\frac{\alpha_{l}}{4\beta^{2}}d_{l}(1-\kappa_{l})+\tfrac{1}{4}\sum_{i=1}^{r}\frac{\kappa_{i}d_{i}}{\alpha_{q}}
Γ+(ϵ)4Λ+(ϵ)2mini=1,,r+sdi(1κi)+14i=1rκidiΓ0+<14i=1rκidiΓ0+<0.\displaystyle\leq-\frac{\Gamma_{\infty}^{+}(\epsilon)}{4\Lambda_{\infty}^{+}(\epsilon)^{2}}\min_{i=1,\ldots,r+s}d_{i}(1-\kappa_{i})+\tfrac{1}{4}\sum_{i=1}^{r}\frac{\kappa_{i}d_{i}}{\Gamma_{0}^{+}}<-\tfrac{1}{4}\sum_{i=1}^{r}\frac{\kappa_{i}d_{i}}{\Gamma_{0}^{+}}<0.

Thus, the inequality αl>Γ+(ϵ)\alpha_{l}>\Gamma_{\infty}^{+}(\epsilon) implies (5.6).

Let 𝒞ϵ+\mathcal{C}_{\epsilon}^{+} be the set of metrics gT+g\in\mathcal{M}_{T}^{+} satisfying (2.3) with

min{Λ0+,Γ0+}min{β,α1,,αr+s}max{β,α1,,αr+s}max{Λ+(ϵ),Γ+(ϵ)}.\displaystyle\min\{\Lambda_{0}^{+},\Gamma_{0}^{+}\}\leq\min\{\beta,\alpha_{1},\ldots,\alpha_{r+s}\}\leq\max\{\beta,\alpha_{1},\ldots,\alpha_{r+s}\}\leq\max\{\Lambda_{\infty}^{+}(\epsilon),\Gamma_{\infty}^{+}(\epsilon)\}.

Clearly, this set is compact. Summarising the arguments above, we conclude that (5.6) holds for all gT+𝒞ϵ+g\in\mathcal{M}_{T}^{+}\setminus\mathcal{C}_{\epsilon}^{+}. ∎

With Lemma 5.4 at hand, we can prove Theorem 5.3 using the approach from [APZ20, Proof of Theorem 3.3]. The main idea behind this approach goes back to [GP17].

Proof of Theorem 5.3.

Denote U=trQTdmTmU=\operatorname{tr}_{Q}T-d_{m}T_{m}. For t>Ut>U, consider the metric gtKg_{t}\in\mathcal{M}_{K} satisfying

gt=tQ|𝔭\displaystyle g_{t}=tQ|_{\mathfrak{p}} +tQ|𝔨1++tQ|𝔨m1\displaystyle+tQ|_{\mathfrak{k}_{1}}+\cdots+tQ|_{\mathfrak{k}_{m-1}}
+ϕ(t)Q|𝔨m+tQ|𝔨m+1++tQ|𝔨r+s,ϕ(t)=dmTmttU.\displaystyle+\phi(t)Q|_{\mathfrak{k}_{m}}+tQ|_{\mathfrak{k}_{m+1}}+\cdots+tQ|_{\mathfrak{k}_{r+s}},\qquad\phi(t)=\frac{d_{m}T_{m}t}{t-U}.

Straightforward verification shows that gtg_{t} lies in T+\mathcal{M}_{T}^{+}. By Corollary 3.8,

S(gt)\displaystyle S(g_{t}) =14tdm(1κm)ϕ(t)4t2dm(1κm)14ti=1r+sdi(1κi)\displaystyle=\frac{1}{4t}d_{m}(1-\kappa_{m})-\frac{\phi(t)}{4t^{2}}d_{m}(1-\kappa_{m})-\frac{1}{4t}\sum_{i=1}^{r+s}d_{i}(1-\kappa_{i})
n2t+14ti=1rκidiκmdm4t+κmdm4ϕ(t)\displaystyle\hphantom{=}~{}-\frac{n}{2t}+\frac{1}{4t}\sum_{i=1}^{r}\kappa_{i}d_{i}-\frac{\kappa_{m}d_{m}}{4t}+\frac{\kappa_{m}d_{m}}{4\phi(t)}
=14tdm(1κm)ϕ(t)4t2dm(1κm)3n4t+dim𝔨4tκmdm4t+κmdm4ϕ(t).\displaystyle=\frac{1}{4t}d_{m}(1-\kappa_{m})-\frac{\phi(t)}{4t^{2}}d_{m}(1-\kappa_{m})-\frac{3n}{4t}+\frac{\dim\mathfrak{k}}{4t}-\frac{\kappa_{m}d_{m}}{4t}+\frac{\kappa_{m}d_{m}}{4\phi(t)}.

Furthermore, in light of (5.4),

4limtt2ddtS(gt)\displaystyle 4\lim_{t\to\infty}t^{2}\frac{d}{dt}S(g_{t}) =dm(1κm)dim𝔨+3n+κmdm+κmUTm\displaystyle=-d_{m}(1-\kappa_{m})-\dim\mathfrak{k}+3n+\kappa_{m}d_{m}+\frac{\kappa_{m}U}{T_{m}}
=dm(1κm)dim𝔨+3n+κmtrQTTm<0.\displaystyle=-d_{m}(1-\kappa_{m})-\dim\mathfrak{k}+3n+\frac{\kappa_{m}\operatorname{tr}_{Q}T}{T_{m}}<0. (5.10)

We conclude that ddtS(gt)<0\frac{d}{dt}S(g_{t})<0 for sufficiently large tt, which implies the existence of t0(U,)t_{0}\in(U,\infty) such that

S(gt0)>limtS(gt)=κm4Tm.S(g_{t_{0}})>\lim_{t\to\infty}S(g_{t})=\frac{\kappa_{m}}{4T_{m}}.

Using Lemma 5.4 with

ϵ=12(S(gt0)κm4Tm)>0\epsilon=\tfrac{1}{2}\Big{(}S(g_{t_{0}})-\frac{\kappa_{m}}{4T_{m}}\Big{)}>0

yields

S(h)<κm4Tm+ϵ=12S(gt0)+κm8Tm<S(gt0),hT+𝒞ϵ+.S(h)<\frac{\kappa_{m}}{4T_{m}}+\epsilon=\tfrac{1}{2}S(g_{t_{0}})+\frac{\kappa_{m}}{8T_{m}}<S(g_{t_{0}}),\qquad h\in\mathcal{M}^{+}_{T}\setminus\mathcal{C}_{\epsilon}^{+}. (5.11)

Since 𝒞ϵ+\mathcal{C}_{\epsilon}^{+} is compact, the functional S|𝒞ϵ+S|_{\mathcal{C}^{+}_{\epsilon}} attains its global maximum at some gmx𝒞ϵ+{g_{\mathrm{mx}}\in\mathcal{C}_{\epsilon}^{+}}. Obviously, gt0g_{t_{0}} lies in 𝒞ϵ+\mathcal{C}_{\epsilon}^{+}. Therefore, by (5.11), S(h)S(gmx)S(h)\leq S(g_{\mathrm{mx}}) for all hT+h\in\mathcal{M}^{+}_{T}. ∎

5.3 Global maxima on T\mathcal{M}_{T}^{-}

Now we focus on the space T\mathcal{M}_{T}^{-}.

Theorem 5.5.

Suppose TT is a left-invariant (0,2)-tensor field on GG satisfying (4.2) for some T𝔭,T1,,Tr+s>0T_{\mathfrak{p}},T_{1},\ldots,T_{r+s}>0. If condition (5.5) holds, then the functional S|TS|_{\mathcal{M}^{-}_{T}} attains its global maximum.

The proof relies on the following estimate.

Lemma 5.6.

Assume that (5.5) holds. Given θ>0\theta>0, there exists a compact set 𝒞θT\mathcal{C}_{\theta}^{-}\subset\mathcal{M}^{-}_{T} such that S(g)<θS(g)<-\theta for every gT𝒞θg\in\mathcal{M}^{-}_{T}\setminus\mathcal{C}_{\theta}^{-}.

Proof.

Let gT+g\in\mathcal{M}^{+}_{T} be a metric satisfying (2.3). Then

trgT=nT𝔭β+i=1r+sdiTiαi=1,\displaystyle\operatorname{tr}_{g}T=-\frac{nT_{\mathfrak{p}}}{\beta}+\sum_{i=1}^{r+s}\frac{d_{i}T_{i}}{\alpha_{i}}=-1, (5.12)

which implies β<nT𝔭\beta<nT_{\mathfrak{p}}. Moreover,

βαj<βdjTji=1r+sdiTiαi=nT𝔭βdjTj,j=1,,r+s.\displaystyle\frac{\beta}{\alpha_{j}}<\frac{\beta}{d_{j}T_{j}}\sum_{i=1}^{r+s}\frac{d_{i}T_{i}}{\alpha_{i}}=\frac{nT_{\mathfrak{p}}-\beta}{d_{j}T_{j}},\qquad j=1,\ldots,r+s.

Given θ>0\theta>0, denote

Λ0(θ)=14θ|i=1r+sn2κiT𝔭2di2(1κi)Ti2nT𝔭Ti2n|.\displaystyle\Lambda_{0}^{-}(\theta)=\frac{1}{4\theta}\bigg{|}\sum_{i=1}^{r+s}\frac{n^{2}\kappa_{i}T_{\mathfrak{p}}^{2}-d_{i}^{2}(1-\kappa_{i})T_{i}^{2}}{nT_{\mathfrak{p}}T_{i}}-2n\bigg{|}.

Corollary 3.8 implies

S(g)\displaystyle S(g) <14β(i=1r+s(κinT𝔭βTidi2(1κi)TinT𝔭β)2n)\displaystyle<\frac{1}{4\beta}\bigg{(}\sum_{i=1}^{r+s}\bigg{(}\kappa_{i}\frac{nT_{\mathfrak{p}}-\beta}{T_{i}}-\frac{d_{i}^{2}(1-\kappa_{i})T_{i}}{nT_{\mathfrak{p}}-\beta}\bigg{)}-2n\bigg{)}
<14β(i=1r+s(κinT𝔭Tidi2(1κi)TinT𝔭)2n)\displaystyle<\frac{1}{4\beta}\bigg{(}\sum_{i=1}^{r+s}\bigg{(}\kappa_{i}\frac{nT_{\mathfrak{p}}}{T_{i}}-\frac{d_{i}^{2}(1-\kappa_{i})T_{i}}{nT_{\mathfrak{p}}}\bigg{)}-2n\bigg{)}
=14β(i=1r+sn2κiT𝔭2di2(1κi)Ti2nT𝔭Ti2n).\displaystyle=\frac{1}{4\beta}\bigg{(}\sum_{i=1}^{r+s}\frac{n^{2}\kappa_{i}T_{\mathfrak{p}}^{2}-d_{i}^{2}(1-\kappa_{i})T_{i}^{2}}{nT_{\mathfrak{p}}T_{i}}-2n\bigg{)}.

By (5.5), if β<Λ0(θ)\beta<\Lambda_{0}^{-}(\theta), then S(g)<θS(g)<-\theta.

Choose pp and qq as in (5.8). For θ>0\theta>0, denote

Γ0(θ)=dpTpΛ0(θ)nT𝔭.\displaystyle\Gamma_{0}^{-}(\theta)=\frac{d_{p}T_{p}\Lambda_{0}^{-}(\theta)}{nT_{\mathfrak{p}}}.

If αq<Γ0(θ)\alpha_{q}<\Gamma_{0}^{-}(\theta), then

β=nT𝔭(i=1r+sdiTiαi+1)1<nT𝔭αqdqTqnT𝔭αqdpTp<Λ0(θ),\displaystyle\beta=nT_{\mathfrak{p}}\bigg{(}\sum_{i=1}^{r+s}\frac{d_{i}T_{i}}{\alpha_{i}}+1\bigg{)}^{-1}<\frac{nT_{\mathfrak{p}}\alpha_{q}}{d_{q}T_{q}}\leq\frac{nT_{\mathfrak{p}}\alpha_{q}}{d_{p}T_{p}}<\Lambda_{0}^{-}(\theta),

which means S(g)<θS(g)<-\theta.

Choose ll as in (5.9). For θ>0\theta>0, denote

Γ(θ)=n2T𝔭2mini=1,r+sdi(1κi)(4θ+i=1rκidiΓ0(θ)).\displaystyle\Gamma_{\infty}^{-}(\theta)=\frac{n^{2}T_{\mathfrak{p}}^{2}}{\min_{i=1,\ldots r+s}d_{i}(1-\kappa_{i})}\bigg{(}4\theta+\sum_{i=1}^{r}\frac{\kappa_{i}d_{i}}{\Gamma_{0}^{-}(\theta)}\bigg{)}.

As shown above, if αq<Γ0(θ)\alpha_{q}<\Gamma_{0}^{-}(\theta), then S(g)<θS(g)<-\theta. Recalling that β<nT𝔭\beta<nT_{\mathfrak{p}} and assuming that αqΓ0(θ)\alpha_{q}\geq\Gamma_{0}^{-}(\theta) and αl>Γ(θ)\alpha_{l}>\Gamma_{\infty}^{-}(\theta), we obtain

S(g)\displaystyle S(g) <αl4β2dl(1κl)+14i=1rκidiαq\displaystyle<-\frac{\alpha_{l}}{4\beta^{2}}d_{l}(1-\kappa_{l})+\tfrac{1}{4}\sum_{i=1}^{r}\frac{\kappa_{i}d_{i}}{\alpha_{q}}
Γ(θ)4n2T𝔭2mini=1,,r+sdi(1κi)+14i=1rκidiΓ0(θ)<θ.\displaystyle\leq-\frac{\Gamma_{\infty}^{-}(\theta)}{4n^{2}T_{\mathfrak{p}}^{2}}\min_{i=1,\ldots,r+s}d_{i}(1-\kappa_{i})+\tfrac{1}{4}\sum_{i=1}^{r}\frac{\kappa_{i}d_{i}}{\Gamma_{0}^{-}(\theta)}<-\theta.

Thus, the inequality αl>Γ(θ)\alpha_{l}>\Gamma_{\infty}^{-}(\theta) implies S(g)<θS(g)<-\theta.

Let 𝒞θ\mathcal{C}_{\theta}^{-} be the set of those gT+g\in\mathcal{M}_{T}^{+} that satisfy (2.3) with

min{Λ0(θ),Γ0(θ)}\displaystyle\min\{\Lambda_{0}^{-}(\theta),\Gamma_{0}^{-}(\theta)\} min{β,α1,,αr+s}\displaystyle\leq\min\{\beta,\alpha_{1},\ldots,\alpha_{r+s}\}
max{β,α1,,αr+s}max{nT𝔭,Γ(θ)}.\displaystyle\leq\max\{\beta,\alpha_{1},\ldots,\alpha_{r+s}\}\leq\max\{nT_{\mathfrak{p}},\Gamma_{\infty}^{-}(\theta)\}.

This set is compact. By the arguments above, S(g)<θS(g)<-\theta whenever gg lies in T+𝒞θ\mathcal{M}_{T}^{+}\setminus\mathcal{C}_{\theta}^{-}. ∎

Proof of Theorem 5.5.

Fix a metric hTh\in\mathcal{M}_{T}^{-}. Applying Lemma 5.6 with θ=|S(h)|+1\theta=|S(h)|+1, we conclude that

S(g)<|S(h)|1<S(h)\displaystyle S(g)<-|S(h)|-1<S(h)

for all gTg\in\mathcal{M}_{T}^{-} outside a compact set 𝒞θT\mathcal{C}_{\theta}^{-}\subset\mathcal{M}_{T}^{-}. Clearly, hh lies in 𝒞θ\mathcal{C}_{\theta}^{-}, and the functional S|𝒞θS|_{\mathcal{C}_{\theta}^{-}} attains its global maximum at some hmx𝒞θh_{\mathrm{mx}}\in\mathcal{C}_{\theta}^{-}. This implies S(g)S(hmx)S(g)\leq S(h_{\mathrm{mx}}) for all gTg\in\mathcal{M}_{T}^{-}. ∎

5.4 Critical points on T0\mathcal{M}_{T}^{0}

If r+s=1r+s=1 in formula (2.2), then

T=T𝔭Q|𝔭+T1Q|𝔨1\displaystyle T=-T_{\mathfrak{p}}Q|_{\mathfrak{p}}+T_{1}Q|_{\mathfrak{k}_{1}} (5.13)

for some T𝔭,T1>0T_{\mathfrak{p}},T_{1}>0. In this case, straightforward analysis shows that S|T0S|_{\mathcal{M}_{T}^{0}} has no critical points unless

d1T12+2nT𝔭(2T1κ1T𝔭)=0.\displaystyle d_{1}T_{1}^{2}+2nT_{\mathfrak{p}}(2T_{1}-\kappa_{1}T_{\mathfrak{p}})=0. (5.14)

On the other hand, when (5.14) holds, the scalar curvature of every metric in T0\mathcal{M}_{T}^{0} equals 0. If r+s=2r+s=2, we are able to obtain a complete classification of the critical points of S|T0S|_{\mathcal{M}_{T}^{0}}. We present this classification in Theorem 5.8 below. While its statement is quite bulky, its conditions are easy to verify once the tensor field TT and the geometric parameters of GG and KK are given. According to Table 3 in [Bes87, Section 7.H], the sum r+sr+s can be greater than 2 only if (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}) is one of the pairs

(𝔰𝔲(p,q),𝔰𝔲(p)𝔰𝔲(q)),\displaystyle(\mathfrak{su}(p,q),\mathfrak{su}(p)\oplus\mathfrak{su}(q)\oplus{\mathbb{R}}), 1<pq,\displaystyle 1<p\leq q,
(𝔰𝔬(4,m),𝔰𝔬(4)𝔰𝔬(m)),\displaystyle(\mathfrak{so}(4,m),\mathfrak{so}(4)\oplus\mathfrak{so}(m)), m4,\displaystyle m\geq 4,
(𝔰𝔬(3,4),𝔰𝔬(3)𝔰𝔬(4)).\displaystyle(\mathfrak{so}(3,4),\mathfrak{so}(3)\oplus\mathfrak{so}(4)).

It seems difficult to classify the critical points of S|T0S|_{\mathcal{M}_{T}^{0}} in these cases without using software, such as Maple, to solve the Euler–Lagrange equations numerically. Nevertheless, for all values of r+sr+s, the following result holds.

Proposition 5.7.

If gg is a critical point of S|T0S|_{\mathcal{M}_{T}^{0}}, then S(g)=0S(g)=0.

Proof.

According to Proposition 5.2, the fact that gg is a critical point of S|T0S|_{\mathcal{M}_{T}^{0}} implies equality (4.5). Taking the trace on both sides of this equality, we obtain

S(g)=trgRic(g)=ctrgT=0.\displaystyle S(g)=\operatorname{tr}_{g}\operatorname{Ric}(g)=c\operatorname{tr}_{g}T=0.

Assume that r+s=2r+s=2 in formula (2.2). For the list of 𝔨\mathfrak{k} satisfying this assumption, see Table 3 in [Bes87, Section 7.H]. Equality (4.2) becomes

T=T𝔭Q|𝔭+T1Q|𝔨1+T2Q|𝔨2.T=-T_{\mathfrak{p}}Q|_{\mathfrak{p}}+T_{1}Q|_{\mathfrak{k}_{1}}+T_{2}Q|_{\mathfrak{k}_{2}}. (5.15)

It will be convenient for us to denote

a=nd2(1κ2)T𝔭,b\displaystyle a=nd_{2}(1-\kappa_{2})T_{\mathfrak{p}},\qquad b =d12(1κ1)T1d22(1κ2)T2+2n2T𝔭n2κ1T𝔭2T1,\displaystyle=d_{1}^{2}(1-\kappa_{1})T_{1}-d_{2}^{2}(1-\kappa_{2})T_{2}+2n^{2}T_{\mathfrak{p}}-\frac{n^{2}\kappa_{1}T_{\mathfrak{p}}^{2}}{T_{1}},
c\displaystyle c =2nd2T2+2nd2κ1T𝔭T2T1nd2κ2T𝔭,\displaystyle=-2nd_{2}T_{2}+\frac{2nd_{2}\kappa_{1}T_{\mathfrak{p}}T_{2}}{T_{1}}-nd_{2}\kappa_{2}T_{\mathfrak{p}},
d\displaystyle d =d22κ1T22T1+κ2d22T2.\displaystyle=-\frac{d_{2}^{2}\kappa_{1}T_{2}^{2}}{T_{1}}+\kappa_{2}d_{2}^{2}T_{2}.

The proof of Theorem 5.8 below shows that the variational properties of S|T0S|_{\mathcal{M}_{T}^{0}} are largely determined by those of the polynomial

P(x)=ax3+bx2+cx+d.\displaystyle P(x)=ax^{3}+bx^{2}+cx+d.

The discriminant of this polynomial is

D=18abcd4b3d+b2c24ac327a2d2.\displaystyle D=18abcd-4b^{3}d+b^{2}c^{2}-4ac^{3}-27a^{2}d^{2}.

Denote

Rt=b3a,Rd=9adbc2(b23ac),Rs=4abc9a2db3a(b23ac).\displaystyle R_{t}=-\frac{b}{3a},\qquad R_{d}=\frac{9ad-bc}{2(b^{2}-3ac)},\qquad R_{s}=\frac{4abc-9a^{2}d-b^{3}}{a(b^{2}-3ac)}.

According to the classical theory of cubic equations (see, e.g., [Jan10] for a modern interpretation), if D=0D=0 and b2=3acb^{2}=3ac, then x=Rtx=R_{t} is a triple root of P(x)P(x). It is also a saddle point. If D=0D=0 and b23acb^{2}\neq 3ac, then x=Rdx=R_{d} and x=Rsx=R_{s} are a double root and a simple root of P(x)P(x), respectively. Both are local extremum points.

Theorem 5.8.

Assume that r+s=2r+s=2 in formula (2.2). The scalar curvature functional S|T0S|_{\mathcal{M}_{T}^{0}} does not have a global minimum. Critical points of other types exist under the following conditions:

  1. 1.

    A saddle if and only if

    D=0,b2=3ac,d2T2nT𝔭<Rt.\displaystyle D=0,\qquad b^{2}=3ac,\qquad\frac{d_{2}T_{2}}{nT_{\mathfrak{p}}}<R_{t}. (5.16)
  2. 2.

    A global maximum if and only if

    D=0,b23ac,Rsd2T2nT𝔭<Rd.\displaystyle D=0,\qquad b^{2}\neq 3ac,\qquad R_{s}\leq\frac{d_{2}T_{2}}{nT_{\mathfrak{p}}}<R_{d}. (5.17)
  3. 3.

    A local maximum that is not a global maximum if and only if

    D=0,b23ac,d2T2nT𝔭<Rs<Rd.\displaystyle D=0,\qquad b^{2}\neq 3ac,\qquad\frac{d_{2}T_{2}}{nT_{\mathfrak{p}}}<R_{s}<R_{d}. (5.18)
  4. 4.

    A local minimum if and only if

    D=0,b23ac,d2T2nT𝔭<Rd<Rs.\displaystyle D=0,\qquad b^{2}\neq 3ac,\qquad\frac{d_{2}T_{2}}{nT_{\mathfrak{p}}}<R_{d}<R_{s}. (5.19)

When it exists, the critical point of S|T0S|_{\mathcal{M}_{T}^{0}} is unique up to scaling.

Let us make a few remarks in preparation for the proof. Consider a metric gT0g\in\mathcal{M}_{T}^{0}. There are β,α1,α2>0\beta,\alpha_{1},\alpha_{2}>0 such that

g=βQ|𝔭+α1Q|𝔨1+α2Q|𝔨2.g=\beta Q|_{\mathfrak{p}}+\alpha_{1}Q|_{\mathfrak{k}_{1}}+\alpha_{2}Q|_{\mathfrak{k}_{2}}. (5.20)

The equality trgT=0\operatorname{tr}_{g}T=0 implies

nT𝔭β+d1T1α1+d2T2α2=0,βα1=1d1T1(nT𝔭d2T2βα2),α2β>d2T2nT𝔭.\displaystyle-\frac{nT_{\mathfrak{p}}}{\beta}+\frac{d_{1}T_{1}}{\alpha_{1}}+\frac{d_{2}T_{2}}{\alpha_{2}}=0,\qquad\frac{\beta}{\alpha_{1}}=\frac{1}{d_{1}T_{1}}\bigg{(}nT_{\mathfrak{p}}-\frac{d_{2}T_{2}\beta}{\alpha_{2}}\bigg{)},\qquad\frac{\alpha_{2}}{\beta}>\frac{d_{2}T_{2}}{nT_{\mathfrak{p}}}.

By Corollary 3.8,

S(g)\displaystyle S(g) =14β(α1βd1(1κ1)α2βd2(1κ2)2n+κ1d1βα1+κ2d2βα2)\displaystyle=\frac{1}{4\beta}\Big{(}-\frac{\alpha_{1}}{\beta}d_{1}(1-\kappa_{1})-\frac{\alpha_{2}}{\beta}d_{2}(1-\kappa_{2})-2n+\frac{\kappa_{1}d_{1}\beta}{\alpha_{1}}+\frac{\kappa_{2}d_{2}\beta}{\alpha_{2}}\Big{)}
=β4α2(d2T2βnT𝔭α2)P(α2β).\displaystyle=\frac{\beta}{4\alpha_{2}(d_{2}T_{2}\beta-nT_{\mathfrak{p}}\alpha_{2})}P\Big{(}\frac{\alpha_{2}}{\beta}\Big{)}. (5.21)

Note that the factor in front of P(α2β)P(\frac{\alpha_{2}}{\beta}) is necessarily negative. Our arguments will involve two curves, γ1\gamma_{1} and γ2\gamma_{2}, in the space T0\mathcal{M}_{T}^{0} given by the formulas

γ1(t)\displaystyle\gamma_{1}(t) =βQ|𝔭d1T1etα2βd2T2βnT𝔭etα2Q|𝔨1+etα2Q|𝔨2,\displaystyle=\beta Q|_{\mathfrak{p}}-\frac{d_{1}T_{1}e^{t}\alpha_{2}\beta}{d_{2}T_{2}\beta-nT_{\mathfrak{p}}e^{t}\alpha_{2}}Q|_{\mathfrak{k}_{1}}+e^{t}\alpha_{2}Q|_{\mathfrak{k}_{2}}, t>lnd2T2βnT𝔭α2,\displaystyle t>\ln\frac{d_{2}T_{2}\beta}{nT_{\mathfrak{p}}\alpha_{2}},
γ2(t)\displaystyle\gamma_{2}(t) =etβQ|𝔭+etα1Q|𝔨1+etα2Q|𝔨2,\displaystyle=e^{t}\beta Q|_{\mathfrak{p}}+e^{t}\alpha_{1}Q|_{\mathfrak{k}_{1}}+e^{t}\alpha_{2}Q|_{\mathfrak{k}_{2}}, t.\displaystyle t\in\mathbb{R}.

Both these curves pass through gg at t=0t=0.

Lemma 5.9.

The metric gg given by (5.20) is a critical point of S|T0S|_{\mathcal{M}_{T}^{0}} if and only if x=α2βx=\frac{\alpha_{2}}{\beta} is a multiple root of P(x)P(x).

Proof.

Assume gg is a critical point of S|T0S|_{\mathcal{M}_{T}^{0}}. Proposition 5.7 implies S(g)=0S(g)=0. In light of (5.4), this means x=α2βx=\frac{\alpha_{2}}{\beta} must be a root of P(x)P(x). Furthermore, because gg is a critical point of S|T0S|_{\mathcal{M}_{T}^{0}},

0\displaystyle 0 =ddtS(γ1(t))|t=0=ddt(β4etα2(d2T2βnT𝔭etα2)P(etα2β))|t=0\displaystyle=\frac{d}{dt}S(\gamma_{1}(t))|_{t=0}=\frac{d}{dt}\Big{(}\frac{\beta}{4e^{t}\alpha_{2}(d_{2}T_{2}\beta-nT_{\mathfrak{p}}e^{t}\alpha_{2})}P\Big{(}\frac{e^{t}\alpha_{2}}{\beta}\Big{)}\Big{)}\Big{|}_{t=0}
=ddtβ4etα2(d2T2βnT𝔭etα2)|t=0P(α2β)+β4α2(d2T2βnT𝔭α2)ddtP(etα2β)|t=0\displaystyle=\frac{d}{dt}\frac{\beta}{4e^{t}\alpha_{2}(d_{2}T_{2}\beta-nT_{\mathfrak{p}}e^{t}\alpha_{2})}\Big{|}_{t=0}P\Big{(}\frac{\alpha_{2}}{\beta}\Big{)}+\frac{\beta}{4\alpha_{2}(d_{2}T_{2}\beta-nT_{\mathfrak{p}}\alpha_{2})}\frac{d}{dt}P\Big{(}\frac{e^{t}\alpha_{2}}{\beta}\Big{)}\Big{|}_{t=0}
=14(d2T2βnT𝔭α2)ddxP(x)|x=α2β.\displaystyle=\frac{1}{4(d_{2}T_{2}\beta-nT_{\mathfrak{p}}\alpha_{2})}\frac{d}{dx}P(x)|_{x=\frac{\alpha_{2}}{\beta}}.

Thus, the derivative of P(x)P(x) at x=α2βx=\frac{\alpha_{2}}{\beta} vanishes. This proves the “only if” part of the claim.

Assume that x=α2βx=\frac{\alpha_{2}}{\beta} is a multiple root of P(x)P(x). We need to show that gg is a critical point of S|T0S|_{\mathcal{M}_{T}^{0}}. Clearly, the vectors tangent to the curves γ1\gamma_{1} and γ2\gamma_{2} at gg are linearly independent. Therefore, it suffices to prove that

ddtS(γ1(t))|t=0=ddtS(γ2(t))|t=0=0.\displaystyle\frac{d}{dt}S(\gamma_{1}(t))|_{t=0}=\frac{d}{dt}S(\gamma_{2}(t))|_{t=0}=0.

Computing as above, we find

ddt\displaystyle\frac{d}{dt} S(γ1(t))|t=0\displaystyle S(\gamma_{1}(t))|_{t=0}
=ddtβ4etα2(d2T2βnT𝔭etα2)|t=0P(α2β)+14(d2T2βnT𝔭α2)ddxP(x)|x=α2β=0.\displaystyle=\frac{d}{dt}\frac{\beta}{4e^{t}\alpha_{2}(d_{2}T_{2}\beta-nT_{\mathfrak{p}}e^{t}\alpha_{2})}\Big{|}_{t=0}P\Big{(}\frac{\alpha_{2}}{\beta}\Big{)}+\frac{1}{4(d_{2}T_{2}\beta-nT_{\mathfrak{p}}\alpha_{2})}\frac{d}{dx}P(x)|_{x=\frac{\alpha_{2}}{\beta}}=0.

Formula (5.4) implies

ddtS(γ2(t))|t=0=ddtS(etg)|t=0=ddtetS(g)|t=0=β4α2(d2T2βnT𝔭α2)P(α2β)=0.\displaystyle\frac{d}{dt}S(\gamma_{2}(t))|_{t=0}=\frac{d}{dt}S(e^{t}g)|_{t=0}=\frac{d}{dt}e^{-t}S(g)|_{t=0}=-\frac{\beta}{4\alpha_{2}(d_{2}T_{2}\beta-nT_{\mathfrak{p}}\alpha_{2})}P\Big{(}\frac{\alpha_{2}}{\beta}\Big{)}=0.

Proof of Theorem 5.8.

Recalling that P(x)P(x) is a cubic polynomial, we find

limtS(γ1(t))\displaystyle\lim_{t\to\infty}S(\gamma_{1}(t)) =limtβ4etα2(d2T2βnT𝔭etα2)P(etα2β)=.\displaystyle=\lim_{t\to\infty}\frac{\beta}{4e^{t}\alpha_{2}(d_{2}T_{2}\beta-nT_{\mathfrak{p}}e^{t}\alpha_{2})}P\Big{(}\frac{e^{t}\alpha_{2}}{\beta}\Big{)}=-\infty.

Consequently, S|T0S|_{\mathcal{M}_{T}^{0}} never attains its global minimum. This proves the first statement.

Suppose gg is a saddle point of S|T0S|_{\mathcal{M}_{T}^{0}}. Proposition 5.7 implies S(g)=0S(g)=0. Moreover, every neighbourhood of gg in T0\mathcal{M}_{T}^{0} contains a metric with negative scalar curvature and one with positive scalar curvature. By Lemma 5.9, x=α2βx=\frac{\alpha_{2}}{\beta} is a multiple root of P(x)P(x). Formula (5.4) shows that every interval around x=α2βx=\frac{\alpha_{2}}{\beta} contains a point where P(x)P(x) is positive and one where P(x)P(x) is negative. This is only possible if x=α2βx=\frac{\alpha_{2}}{\beta} is a triple root. By the classical theory of cubic equations, conditions (5.16) hold. Conversely, these conditions ensure that P(x)P(x) has a triple root at x=Rtx=R_{t}. Consider a metric gsdlT0g_{\mathrm{sdl}}\in\mathcal{M}_{T}^{0} defined by

gsdl=Q|𝔭d1T1Rtd2T2nT𝔭RtQ|𝔨1+RtQ|𝔨2.g_{\mathrm{sdl}}=Q|_{\mathfrak{p}}-\frac{d_{1}T_{1}R_{t}}{d_{2}T_{2}-nT_{\mathfrak{p}}R_{t}}Q|_{\mathfrak{k}_{1}}+R_{t}Q|_{\mathfrak{k}_{2}}.

Lemma 5.9 implies that gsdlg_{\mathrm{sdl}} is a critical point of S|T0S|_{\mathcal{M}_{T}^{0}}. Using (5.4), one easily shows that every neighbourhood of gsdlg_{\mathrm{sdl}} contains a metric with negative scalar curvature and one with positive scalar curvature. In light of Proposition 5.7, this means gsdlg_{\mathrm{sdl}} is a saddle point.

The functional S|T0S|_{\mathcal{M}_{T}^{0}} attains its global maximum if and only if S(ggmx)=0S(g_{\mathrm{gmx}})=0 for some ggmxT0g_{\mathrm{gmx}}\in\mathcal{M}_{T}^{0} and S(h)0S(h)\leq 0 for all hT0h\in\mathcal{M}_{T}^{0}. Formula (5.4) implies that this happens if and only if P(x)P(x) has a double root in the interval (d2T2nT𝔭,)\big{(}\frac{d_{2}T_{2}}{nT_{\mathfrak{p}}},\infty\big{)} and is nonnegative on this interval. Conditions (5.17) are necessary and sufficient for P(x)P(x) to have such properties.

Next, S|T0S|_{\mathcal{M}_{T}^{0}} has a local maximum that is not a global maximum if and only if S(glmx)=0S(g_{\mathrm{lmx}})=0 for some glmxT0g_{\mathrm{lmx}}\in\mathcal{M}_{T}^{0}, S(h)0S(h)\leq 0 for all hh in a neighbourhood of glmxg_{\mathrm{lmx}}, and the scalar curvature of at least one metric in T0\mathcal{M}_{T}^{0} is positive. This is equivalent to P(x)P(x) having a simple root in the interval (d2T2nT𝔭,)\big{(}\frac{d_{2}T_{2}}{nT_{\mathfrak{p}}},\infty\big{)} and a double root in (Rs,)(R_{s},\infty). Conditions (5.18) are necessary and sufficient for P(x)P(x) to have such properties.

Analogously, S|T0S|_{\mathcal{M}_{T}^{0}} has a local minimum if and only if P(x)P(x) has a double root in (d2T2nT𝔭,)\big{(}\frac{d_{2}T_{2}}{nT_{\mathfrak{p}}},\infty\big{)} and is nonpositive in a neighbourhood of this root. Conditions (5.19) are necessary and sufficient for this.

Finally, in view of Lemma 5.9, S|T0S|_{\mathcal{M}_{T}^{0}} can have at most one critical point up to scaling since a cubic polynomial can have at most one multiple root. ∎

5.5 Summary

The results of Sections 5.15.4 enable us to make several conclusions about the solvability of (4.5). We summarise these conclusions in Theorem 5.10 below. The constant cc in (4.5) must be positive if TT satisfies (4.2). This is an immediate consequence of the formulas for the Ricci curvature obtained in Section 3.

Theorem 5.10.

Suppose TT is a left-invariant (0,2)-tensor field on GG given by (4.2).

  1. 1.

    If (5.5) holds, then there exists at least one pair (g,c)K×(0,)(g,c)\in\mathcal{M}_{K}\times(0,\infty) satisfying (4.5).

  2. 2.

    If both (5.4) and (5.5) hold, then there are at least two pairs (g,c)K×(0,)(g,c)\in\mathcal{M}_{K}\times(0,\infty) that satisfy (4.5) and have non-homothetic metrics gg.

  3. 3.

    If r+s=2r+s=2 and conditions (5.16), (5.17), (5.18) or (5.19) hold, then there exists at least one pair (g,c)K×(0,)(g,c)\in\mathcal{M}_{K}\times(0,\infty) satisfying (4.5).

Proof.

Statements 1 and 3 follow from Theorems 5.5 and 5.8 combined with Proposition 5.2. Next, assume that (5.4) and (5.5) hold. According to Theorems 5.3 and 5.5, the functionals S|T+S|_{\mathcal{M}_{T}^{+}} and S|TS|_{\mathcal{M}_{T}^{-}} attain their global maxima at some g1T+g_{1}\in\mathcal{M}_{T}^{+} and g2Tg_{2}\in\mathcal{M}_{T}^{-}. Proposition 5.2 implies that both g1g_{1} and g2g_{2} have Ricci curvature equal to TT up to scaling. These metrics cannot be homothetic because trg1T\operatorname{tr}_{g_{1}}T and trg2T\operatorname{tr}_{g_{2}}T are not of the same sign. ∎

When r+s=1r+s=1, Theorem 5.10 is essentially optimal. We explain this in detail in Remark 6.2. At the same time, when r+s2r+s\geq 2, it seems that (5.4) and (5.5) may fail to hold even if S|T+S|_{\mathcal{M}_{T}^{+}} and S|TS|_{\mathcal{M}_{T}^{-}} attain their global maxima. Indeed, on compact Lie groups, inequalities that are similar in spirit to these provide merely a “linear approximation” to the necessary and sufficient conditions for the existence of a critical point; see [APZ20, Section 5].

Different pairs (g,c)K×(0,)(g,c)\in\mathcal{M}_{K}\times(0,\infty) satisfying (4.5) must have distinct cc. More precisely, by Theorem 4.1, if

Ric(g1)=Ric(g2)=cT,g1,g2K,\displaystyle\operatorname{Ric}(g_{1})=\operatorname{Ric}(g_{2})=cT,\qquad g_{1},g_{2}\in\mathcal{M}_{K},

then g1g_{1} and g2g_{2} are equal up to scaling.

Remark 5.11.

The discussion at the beginning of Section 5.4 shows that (5.14) is a sufficient condition for the solvability of (4.5) if r+s=1r+s=1.

6 The case where KK is simple

As above, let TT be a left-invariant (0,2)-tensor field on GG. Assume that KK is simple. Our next result settles the question of solvability of (4.5) under this assumption. We do not use the variational approach developed in Section 5.1; however, see Remarks 6.2 and 6.3 below.

Since KK is simple, the numbers rr and ss in (2.2) equal 1 and 0, respectively. By Theorem 3.6, if (5.13) holds for some T𝔭,T1>0T_{\mathfrak{p}},T_{1}>0, the constant cc in (4.5) must be positive.

Proposition 6.1.

Assume KK is simple. Let the tensor field TT satisfy (5.13) for some T𝔭,T1>0T_{\mathfrak{p}},T_{1}>0.

  1. 1.

    If

    2nT𝔭(2T1κ1T𝔭)<d1T12\displaystyle 2nT_{\mathfrak{p}}(2T_{1}-\kappa_{1}T_{\mathfrak{p}})<-d_{1}T_{1}^{2} (6.1)

    then there exists no metric gKg\in\mathcal{M}_{K} such that (4.5) holds.

  2. 2.

    If

    2nT𝔭(2T1κ1T𝔭)=d1T12or2T1κ1T𝔭0,\displaystyle 2nT_{\mathfrak{p}}(2T_{1}-\kappa_{1}T_{\mathfrak{p}})=-d_{1}T_{1}^{2}\qquad or\qquad 2T_{1}-\kappa_{1}T_{\mathfrak{p}}\geq 0,

    then there exists precisely one pair (g,c)K×(0,)(g,c)\in\mathcal{M}_{K}\times(0,\infty), up to scaling of gg, such that (4.5) holds.

  3. 3.

    If

    d1T12<2nT𝔭(2T1κ1T𝔭)<0,\displaystyle-d_{1}T_{1}^{2}<2nT_{\mathfrak{p}}(2T_{1}-\kappa_{1}T_{\mathfrak{p}})<0,

    then there are precisely two pairs (g,c)K×(0,)(g,c)\in\mathcal{M}_{K}\times(0,\infty), up to scaling of gg, such that (4.5) holds.

Proof.

Choose a metric gKg\in\mathcal{M}_{K}. There exist β,α1>0\beta,\alpha_{1}>0 such that

g=βQ|𝔭+α1Q|𝔨1.g=\beta Q|_{\mathfrak{p}}+\alpha_{1}Q|_{\mathfrak{k}_{1}}.

Theorem 3.6 and formula (3.3) imply that gg satisfies (4.5) if and only if

14(κ1(1x2)+x2)\displaystyle\tfrac{1}{4}(\kappa_{1}(1-x^{2})+x^{2}) =cT1,\displaystyle=cT_{1},
14(2+x)\displaystyle\tfrac{1}{4}(2+x) =cT𝔭,\displaystyle=cT_{\mathfrak{p}},

where x=α1βx=\frac{\alpha_{1}}{\beta}. Clearing cc from the second line and substituting into the first, we obtain

(1κ1)T𝔭x2T1x+κ1T𝔭2T1=0.\displaystyle(1-\kappa_{1})T_{\mathfrak{p}}x^{2}-T_{1}x+\kappa_{1}T_{\mathfrak{p}}-2T_{1}=0.

This is a quadratic equation with discriminant

E=T12+4(1κ1)T𝔭(2T1κ1T𝔭).\displaystyle E=T_{1}^{2}+4(1-\kappa_{1})T_{\mathfrak{p}}(2T_{1}-\kappa_{1}T_{\mathfrak{p}}).

It has no solutions if E<0E<0, precisely one positive solution if E=0E=0 or ET12E\geq T_{1}^{2}, and precisely two positive solutions if 0<E<T120<E<T_{1}^{2}. Together with (3.3), this implies the result. ∎

Remark 6.2.

Proposition 6.1 shows that Theorem 5.10 is essentially optimal in our current setting. Indeed, since KK is simple, (5.5) becomes

κ1n2T𝔭2(1κ1)d12T12nT𝔭T12n<0.\displaystyle\frac{\kappa_{1}n^{2}T_{\mathfrak{p}}^{2}-(1-\kappa_{1})d_{1}^{2}T_{1}^{2}}{nT_{\mathfrak{p}}T_{1}}-2n<0.

In view of (3.3), this is equivalent to

2nT𝔭(2T1κ1T𝔭)>d1T12.\displaystyle 2nT_{\mathfrak{p}}(2T_{1}-\kappa_{1}T_{\mathfrak{p}})>-d_{1}T_{1}^{2}.

Theorem 5.10 and Remark 5.11 assert that (5.5) and (5.14) are sufficient conditions for the solvability of (4.5). Conversely, as Proposition 6.1 shows, the existence of a pair (g,c)K×(0,)(g,c)\in\mathcal{M}_{K}\times(0,\infty) satisfying (4.5) implies that either (5.5) or (5.14) must hold. Theorem 5.10 provides lower bounds on the number of solutions to (4.5). Using Proposition 6.1, one can easily demonstrate that these bounds are sharp.

Remark 6.3.

In our current setting, every metric satisfying (4.5) is, up to scaling, a global maximum point of S|T+S|_{\mathcal{M}_{T}^{+}}, S|TS|_{\mathcal{M}_{T}^{-}} or S|T0S|_{\mathcal{M}_{T}^{0}}. This observation follows from the results of Section 5 and Proposition 6.1.

7 Examples

Let us illustrate how the results of Sections 56 apply to specific groups.

Example 7.1.

Assume G=𝖦2G=\mathsf{G}_{2}^{\mathbb{C}} and K=𝖦2K=\mathsf{G}_{2}. Then

r=1,s=0,n=d1=14.\displaystyle r=1,\qquad s=0,\qquad n=d_{1}=14.

Formula (3.3) yields κ1=12\kappa_{1}=\tfrac{1}{2}. Suppose TT is given by (5.13) with T𝔭,T1>0T_{\mathfrak{p}},T_{1}>0. Since KK is simple, Proposition 6.1 applies. Formula (6.1) becomes

4T𝔭T1T𝔭2<T12.\displaystyle 4T_{\mathfrak{p}}T_{1}-T_{\mathfrak{p}}^{2}<-T_{1}^{2}.

Equivalently,

T1T𝔭<(52).\displaystyle\frac{T_{1}}{T_{\mathfrak{p}}}<(\sqrt{5}-2).

If this holds, then (4.5) has no solutions. Similarly, if

T1T𝔭=(52)orT1T𝔭14,\displaystyle\frac{T_{1}}{T_{\mathfrak{p}}}=(\sqrt{5}-2)\qquad\mbox{or}\qquad\frac{T_{1}}{T_{\mathfrak{p}}}\geq\tfrac{1}{4},

then there is one pair (g,c)K×(0,)(g,c)\in\mathcal{M}_{K}\times(0,\infty), up to scaling of gg, that satisfies (4.5). If

(52)<T1T𝔭<14,\displaystyle(\sqrt{5}-2)<\frac{T_{1}}{T_{\mathfrak{p}}}<\tfrac{1}{4},

there are two such pairs.

Example 7.2.

Assume G=𝖲𝖮+(2,q)G=\mathsf{SO}^{+}(2,q) and K=𝖲𝖮(2)×𝖲𝖮(q)K=\mathsf{SO}(2)\times\mathsf{SO}(q) with q5q\geq 5. Then

r=1,s=1,n=2q,d1=q(q1)2,d2=1.\displaystyle r=1,\qquad s=1,\qquad n=2q,\qquad d_{1}=\frac{q(q-1)}{2},\qquad d_{2}=1.

Using (3.3), we find

κ1=q2q,κ2=0.\displaystyle\kappa_{1}=\frac{q-2}{q},\qquad\kappa_{2}=0.

Suppose TT is given by (5.15) with T𝔭,T1,T2>0T_{\mathfrak{p}},T_{1},T_{2}>0. Inequality (5.5) becomes

8q(q2)T𝔭2q(q1)2T122T1T216q2T𝔭T1<0.\displaystyle 8q(q-2)T_{\mathfrak{p}}^{2}-q(q-1)^{2}T_{1}^{2}-2T_{1}T_{2}-16q^{2}T_{\mathfrak{p}}T_{1}<0. (7.1)

According to Theorem 5.5, if (7.1) holds, then S|TS|_{\mathcal{M}_{T}^{-}} attains its global maximum. In this case, there exists a metric gTg\in\mathcal{M}_{T}^{-} with Ricci curvature cTcT for some c>0c>0. Inequality (5.4) takes the form

2q(q2)T𝔭+q(4q+1)T1+(q2)T2<0.\displaystyle-2q(q-2)T_{\mathfrak{p}}+q(4q+1)T_{1}+(q-2)T_{2}<0. (7.2)

The set of triples (T𝔭,T1,T2)(T_{\mathfrak{p}},T_{1},T_{2}) for which both (7.1) and (7.2) are satisfied is non-empty and open in 3\mathbb{R}^{3}. We depict it in Figure 1 for q=10q=10. We also indicate where (7.1) holds without (7.2). Because these inequalities are invariant under scaling of (T𝔭,T1,T2)(T_{\mathfrak{p}},T_{1},T_{2}), we make our sketch assuming T𝔭=1T_{\mathfrak{p}}=1. By Theorem 5.3, if both (7.1) and (7.2) are satisfied, then S|T+S|_{\mathcal{M}_{T}^{+}} attains its global maximum. In this case, there exists a metric gT+g\in\mathcal{M}_{T}^{+} with Ricci curvature cTcT for some c>0c>0.

Refer to caption
Figure 1: Existence of global maxima of S|TS|_{\mathcal{M}_{T}^{-}} and S|T+S|_{\mathcal{M}_{T}^{+}} on 𝖲𝖮+(2,10)\mathsf{SO}^{+}(2,10)

Theorem 5.8 enables us to classify the critical points of S|T0S|_{\mathcal{M}_{T}^{0}}. For instance, suppose q=5q=5 and (T𝔭,T1,T2)=(1,1,1)(T_{\mathfrak{p}},T_{1},T_{2})=(1,1,1). Then the discriminant DD of the polynomial P(x)P(x) satisfies

D=7994800850.\displaystyle D=\tfrac{79948008}{5}\neq 0.

By Theorem 5.8, S|T0S|_{\mathcal{M}_{T}^{0}} has no critical points. To give another example, suppose q=5q=5 and (T𝔭,T1,T2)=(1,15,η)(T_{\mathfrak{p}},T_{1},T_{2})=\big{(}1,\tfrac{1}{5},\eta\big{)}, where η\eta is the unique positive root of the polynomial

R(x)=12x34412x2583104x+4198144.\displaystyle R(x)=-12x^{3}-4412x^{2}-583104x+4198144.

Then D=η2R(η)=0D=\eta^{2}R(\eta)=0. Moreover, using an approximate value of η\eta calculated in Maple, we find b23acb^{2}\neq 3ac and

Rs=η3+1376η2121808η+77868810η210160η+84640<η10<115η(16η)η21016η+8464=Rd.\displaystyle R_{s}=\frac{\eta^{3}+1376\eta^{2}-121808\eta+778688}{10\eta^{2}-10160\eta+84640}<\frac{\eta}{10}<\frac{115\eta(16-\eta)}{\eta^{2}-1016\eta+8464}=R_{d}.

This means S|T0S|_{\mathcal{M}_{T}^{0}} attains its global maximum.

One can use Maple to produce the graphs of S|TS|_{\mathcal{M}_{T}^{-}}, S|T+S|_{\mathcal{M}_{T}^{+}} and S|T0S|_{\mathcal{M}_{T}^{0}}; cf. [APZ20, Section 5].

Example 7.3.

Assume G=𝖲𝖴(p,q)G=\mathsf{SU}(p,q) and K=𝖲𝖴(p)×𝖴(q)K=\mathsf{SU}(p)\times\mathsf{U}(q) with 2pq2\leq p\leq q. Then

r=2,s=1,n=2pq,d1=p21,d2=q21,d3=1.\displaystyle r=2,\qquad s=1,\qquad n=2pq,\qquad d_{1}=p^{2}-1,\qquad d_{2}=q^{2}-1,\qquad d_{3}=1.

As we showed in Example 3.4,

κ1=pp+q,κ2=qp+q,κ3=0.\displaystyle\kappa_{1}=\frac{p}{p+q},\qquad\kappa_{2}=\frac{q}{p+q},\qquad\kappa_{3}=0.

Suppose

T=T𝔭Q|𝔭+T1Q|𝔨1+T1Q|𝔨3+T3Q|𝔨3\displaystyle T=-T_{\mathfrak{p}}Q|_{\mathfrak{p}}+T_{1}Q|_{\mathfrak{k}_{1}}+T_{1}Q|_{\mathfrak{k}_{3}}+T_{3}Q|_{\mathfrak{k}_{3}}

for some T𝔭,T1,T2,T3>0T_{\mathfrak{p}},T_{1},T_{2},T_{3}>0. Inequality (5.5) becomes

4p3q2T𝔭2T2\displaystyle 4p^{3}q^{2}T_{\mathfrak{p}}^{2}T_{2} (p21)2qT12T2+4p2q3T𝔭2T1\displaystyle-(p^{2}-1)^{2}qT_{1}^{2}T_{2}+4p^{2}q^{3}T_{\mathfrak{p}}^{2}T_{1}
(q21)2pT1T22(p+q)T1T2T38p2q2(p+q)T𝔭T1T2<0.\displaystyle-(q^{2}-1)^{2}pT_{1}T_{2}^{2}-(p+q)T_{1}T_{2}T_{3}-8p^{2}q^{2}(p+q)T_{\mathfrak{p}}T_{1}T_{2}<0. (7.3)

By Theorem 5.5, if (7.3) holds, then S|TS|_{\mathcal{M}_{T}^{-}} attains its global maximum. This is the case, e.g., when

(T𝔭,T1,T2,T3)=(1,1,1,4p2q2).\displaystyle(T_{\mathfrak{p}},T_{1},T_{2},T_{3})=(1,1,1,4p^{2}q^{2}).

Assume for simplicity that T1T2T_{1}\geq T_{2}. Then (5.4) takes the form

2pq2T𝔭+q(p21)T1+qT3\displaystyle-2pq^{2}T_{\mathfrak{p}}+q(p^{2}-1)T_{1}+qT_{3} <(p35p2q4pq22p)T2.\displaystyle<(p^{3}-5p^{2}q-4pq^{2}-2p)T_{2}. (7.4)

By Theorem 5.5, if both (7.3) and this inequality hold, then S|T+S|_{\mathcal{M}_{T}^{+}} attains its global maximum. This happens, e.g., for (p,q)=(2,12)(p,q)=(2,12) and (T𝔭,T1,T2,T3)=(1,1,38,1)(T_{\mathfrak{p}},T_{1},T_{2},T_{3})=\big{(}1,1,\tfrac{3}{8},1\big{)}.

Acknowledgements

The authors are grateful to Jorge Lauret and Cynthia Will for their careful reading of the paper and useful comments.

References

  • [APZ20] R. Arroyo, A. Pulemotov, W. Ziller, The prescribed Ricci curvature problem for naturally reductive metrics on compact Lie groups, arXiv:2001.09441 [math.DG], 2020, submitted.
  • [Aub98] T. Aubin, Some nonlinear problems in Riemannian geometry, Springer-Verlag, Berlin, 1998.
  • [Bes87] A.L. Besse, Einstein manifolds, Springer-Verlag, Berlin, 1987.
  • [But19] T. Buttsworth, The prescribed Ricci curvature problem on three-dimensional unimodular Lie groups, Math. Nachr. 292 (2019) 747–759.
  • [BP19] T. Buttsworth, A. Pulemotov, The prescribed Ricci curvature problem for homogeneous metrics, arXiv:1911.08214 [math.DG], 2019, to appear in: O. Dearricott, W. Tuschmann, Y. Nikolayevsky, T. Leistner, D. Crowley (Eds), Differential geometry in the large, Cambridge University Press.
  • [BPRZ19] T. Buttsworth, A. Pulemotov, Y.A. Rubinstein, W. Ziller, On the Ricci iteration for homogeneous metrics on spheres and projective spaces, arXiv:1811.01724 [math.DG], 2018, to appear in Transf. Groups.
  • [DZ79] J. D’Atri and W. Ziller, Naturally reductive metrics and Einstein metrics on compact Lie groups, Mem. Amer. Math. Soc. 215 (1979).
  • [Delan03] Ph. Delanoë, Local solvability of elliptic, and curvature, equations on compact manifolds, J. Reine Angew. Math. 558 (2003) 23–45.
  • [Delay02] E. Delay, Studies of some curvature operators in a neighborhood of an asymptotically hyperbolic Einstein manifold, Adv. Math. 168 (2002) 213–224.
  • [Delay18] E. Delay, Inversion of some curvature operators near a parallel Ricci metric II: Non-compact manifold with bounded geometry, Ark. Mat. 56 (2018) 285–297.
  • [DeT81] D.M. DeTurck, Existence of metrics with prescribed Ricci curvature: local theory, Invent. Math. 65 (1981/82) 179–207.
  • [DeT85] D.M. DeTurck, Prescribing positive Ricci curvature on compact manifolds, Rend. Sem. Mat. Univ. Politec. Torino 43 (1985) 357–369.
  • [DG99] D. DeTurck, H. Goldschmidt, Metrics with prescribed Ricci curvature of constant rank. I. The integrable case, Adv. Math. 145 (1999) 1–97.
  • [Dyn57] E.B. Dynkin, Semisimple subalgebras of semisimple Lie algebras, Amer. Math. Soc. Transl. (2) 6 (1957) 111–244.
  • [Gor85] C.S. Gordon, Naturally reductive homogeneous Riemannian manifolds, Can. J. Math. 37 (1985) 467–487.
  • [GS10] C.S. Gordon, C.J. Sutton, Spectral isolation of naturally reductive metrics on simple Lie groups, Math. Z. 266 (2010) 979–995.
  • [GP17] M. Gould, A. Pulemotov, The prescribed Ricci curvature problem on homogeneous spaces with intermediate subgroups, arXiv:1710.03024 [math.DG], 2017, to appear in Comm. Anal. Geom.
  • [Ham84] R.S. Hamilton, The Ricci curvature equation, in: S.-S. Chern (Ed.), Seminar on nonlinear partial differential equations, Springer-Verlag, New York, 1984, pages 47–72.
  • [Jan10] S. Janson, Roots of polynomials of degrees 3 and 4, arXiv:1009.2373 [math.HO], 2010.
  • [Jen73] G.R. Jensen, Einstein metrics on principal fibre bundles, J. Differential Geom. 8 (1973) 599–614.
  • [Lau19] E.A. Lauret, On the smallest Laplace eigenvalue for naturally reductive metrics on compact simple Lie groups, Proc. Amer. Math. Soc. 148 (2020) 3375–3380.
  • [Pul13] A. Pulemotov, Metrics with prescribed Ricci curvature near the boundary of a manifold, Math. Ann. 357 (2013) 969–986.
  • [Pul16a] A. Pulemotov, The Dirichlet problem for the prescribed Ricci curvature equation on cohomogeneity one manifolds, Ann. Mat. Pura Appl. 195 (2016) 1269–1286.
  • [Pul16b] A. Pulemotov, Metrics with prescribed Ricci curvature on homogeneous spaces, J. Geom. Phys. 106 (2016) 275–283.
  • [Pul20] A. Pulemotov, Maxima of curvature functionals and the prescribed Ricci curvature problem on homogeneous spaces, J. Geom. Anal. 30 (2020) 987–1010.
  • [PR19] A. Pulemotov, Y.A. Rubinstein, Ricci iteration on homogeneous spaces, Trans. Amer. Math. Soc. 371 (2019) 6257–6287.
  • [Sag70] A.A. Sagle, Some homogeneous Einstein manifolds, Nagoya Math. J. 39 (1970) 81–106.
  • [WW70] N.R. Wallach, F.W. Warner, Curvature forms for 2-manifolds, Proc. Amer. Math. Soc. 25 (1970) 712–713.
  • [WZ86] M.Y. Wang, W. Ziller, Existence and nonexistence of homogeneous Einstein metrics, Invent. Math. 84 (1986) 177–194.