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

On the variational properties of the prescribed Ricci curvature functional

Artem Pulemotov University of Queensland [email protected]  and  Wolfgang Ziller University of Pennsylvania [email protected]
Abstract.

We study the prescribed Ricci curvature problem for homogeneous metrics. Given a (0,2)-tensor field TT, this problem asks for solutions to the equation Ric(g)=cT\operatorname{Ric}(g)=cT for some constant cc. Our approach is based on examining global properties of the scalar curvature functional whose critical points are solutions to this equation. We produce conditions for a general homogeneous space under which it has a global maximum. Finally, we study the behavior of the functional in specific examples to illustrate our result.

This research was supported by the Australian Government through the Australian Research Council’s Discovery Projects funding scheme (project DP180102185). The second author was also supported by the NSF grant 1506148 and a Reybould Fellowship.

The prescribed Ricci curvature problem consists in finding a Riemannian metric gg on a manifold MM such that Ric(g)=T\operatorname{Ric}(g)=T for a given (0,2)-tensor field TT. As was suggested by DeTurck [11] and Hamilton [17], it is more natural to ask instead whether one can solve the equation

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

for some constant cc. In fact, on a compact manifold, such a constant appears to be necessary. For example, if MM is a surface, this follows from the Gauss–Bonnet theorem.

The prescribed Ricci curvature problem has been studied by many authors since the 1980s; see, e.g., [8] for an overview of the literature. Considering the difficulty of the equation involved, it is natural to make symmetry assumptions. More precisely, suppose that the metric gg and the tensor TT are invariant under a Lie group GG acting on MM. In the case where the quotient M/GM/G is one-dimensional, the problem was addressed by Hamilton [17], Cao–DeTurck [10], Pulemotov [23, 24] and Buttsworth–Krishnan [7]. The case where MM is a homogeneous space G/HG/H has been studied extensively; see the survey [8] and the more recent references [9, 21, 22, 3, 2]. In some simple situations, the equation can be solved explicitly, as shown, e.g., in [25, 6].

Throughout the paper we assume that TT is positive-definite. Without this assumption, the behavior of S|TS_{|\mathcal{M}_{T}} is very different. General existence theorems in the homogeneous setting rely on the fact, proven in [25], that GG-invariant metrics on G/HG/H with Ricci curvature cTcT are precisely (up to scaling) the critical points of the scalar curvature functional SS on the set T=T(G/H)\mathcal{M}_{T}=\mathcal{M}_{T}(G/H) of GG-invariant metrics on G/HG/H subject to the constraint trgT=1\operatorname{tr}_{g}T=1. We will assume that GG and HH are compact Lie groups. The goal of this paper is to study the global behavior of the scalar curvature functional. In [14, 26] one finds a general theorem for the existence of a global maximum, under certain assumptions on the isotropy representation of G/HG/H. Our main result removes these assumptions, thereby expanding greatly the class of spaces to which the theorem applies. Moreover, we improve the conditions for the existence of the global maximum and interpret them geometrically.

If HH is maximal in GG, the supremum of SS is always attained. Otherwise, the global behavior of SS depends on the set of intermediate subgroups, i.e., Lie groups KK with HKGH\subset K\subset G. For every such KK, one has the homogeneous fibration

K/HG/HG/K.K/H\to G/H\to G/K.

If we fix homogeneous metrics gFg_{F} and gBg_{B} on the fiber K/HK/H and the base G/KG/K, it is natural to study the two-parameter variation

gs,t=1sgF+1tgBg_{s,t}=\frac{1}{s}\,g_{F}+\frac{1}{t}\,g_{B}

of g=gF+gBg=g_{F}+g_{B}. Notice that we define this variation using the reciprocals of ss and tt. In fact, we often deal with inverses of metrics rather than metrics themselves, which enables us to view T\mathcal{M}_{T} as a pre-compact space. Solving the constraint trgT=1\operatorname{tr}_{g}T=1 for ss, we obtain a one-parameter family gtS|Tg_{t}\in S_{|\mathcal{M}_{T}}, called a canonical variation, with scalar curvature given by

S(gt)=SFT1+T2(SBT2SFT1)tt2T11tT2|A|g.S(g_{t})=\frac{S_{F}}{T_{1}^{*}}+T_{2}^{*}\Big{(}\frac{S_{B}}{T_{2}^{*}}-\frac{S_{F}}{T_{1}^{*}}\Big{)}t-\frac{t^{2}T_{1}^{*}}{1-tT_{2}^{*}}|A|_{g}.

Here T1=trgFT|FT_{1}^{*}=\operatorname{tr}_{g_{F}}T_{|F} and T2=trgBT|BT_{2}^{*}=\operatorname{tr}_{g_{B}}T_{|B} are the traces of TT on the fiber and the base, SFS_{F} and SBS_{B} are the scalar curvatures of gFg_{F} and gBg_{B}, and AA is the O’Neil tensor of the Riemannian submersion. As a consequence,

limt0S(gt)=SFT1andlimt0dS(gt)dt=T2(SBT2SFT1),\lim_{t\to 0}S(g_{t})=\frac{S_{F}}{T_{1}^{*}}\qquad\text{and}\qquad\lim_{t\to 0}\frac{dS(g_{t})}{dt}=T_{2}^{*}\Big{(}\frac{S_{B}}{T_{2}^{*}}-\frac{S_{F}}{T_{1}^{*}}\Big{)},

which means that SFS_{F} and SBSFS_{B}-S_{F} control the behavior of SS at infinity. It is easy to see that SS is bounded from above if TT is positive-definite, and it is therefore natural to search for a global maximum of SS. The above formulas motivate two invariants of the intermediate subgroup KK with Lie algebra 𝔨{\mathfrak{k}}:

α𝔨=sup{S(h)trhT|F|hT(K/H)}andβ𝔨=sup{S(h)trhT|B|hT(G/K)}.\displaystyle\alpha_{\mathfrak{k}}=\sup\Big{\{}\frac{S(h)}{\operatorname{tr}_{h}T_{|F}}\,\Big{|}\,h\in\mathcal{M}_{T}(K/H)\Big{\}}\qquad\text{and}\qquad\beta_{\mathfrak{k}}=\sup\Big{\{}\frac{S(h)}{\operatorname{tr}_{h}T_{|B}}\,\Big{|}\,h\in\mathcal{M}_{T}(G/K)\Big{\}}.

These quantities maximize the limits and the “derivatives” of SS at infinity. We furthermore introduce an invariant of the homogeneous space G/HG/H:

αG/H=sup𝔨α𝔨,\alpha_{G/H}=\textstyle\sup_{{\mathfrak{k}}}\alpha_{\mathfrak{k}},

where the supremum is taken over all Lie algebras of intermediate subgroups. We can now state our main result.

Theorem.

Let G/HG/H be a compact homogeneous space and TT a positive-definite GG-invariant (0,2)-tensor field on G/HG/H. If HH is not maximal in GG, then the set of intermediate subgroups KK with α𝔨=αG/H\alpha_{\mathfrak{k}}=\alpha_{G/H} is non-empty. If KK is such a subgroup of the lowest possible dimension and β𝔨α𝔨>0\beta_{{\mathfrak{k}}}-\alpha_{{\mathfrak{k}}}>0, then S|TS_{|\mathcal{M}_{T}} achieves its supremum at some metric gTg\in\mathcal{M}_{T} and hence Ric(g)=cT\operatorname{Ric}(g)=cT for some c>0c>0.

In fact, we show that there exists ϵ>0\epsilon>0 such that the set {gT(G/H)S(g)>αG/H+ϵ}\{g\in\mathcal{M}_{T}(G/H)\mid S(g)>\alpha_{G/H}+\epsilon\} is non-empty and compact, which implies the result. This requires careful estimates of the behavior of S|TS_{|\mathcal{M}_{T}} at infinity.

We illustrate this result by examining as examples the Stiefel manifold V2(5)V_{2}({\mathbb{R}}^{5}) and Ledger–Obata spaces. These manifolds are especially interesting since the isotropy representation has equivalent modules and there are families of intermediate subgroups. Hence not every metric is “diagonal” (see discussion below) as was assumed in all previously considered examples. The Euler–Lagrange equations for S|TS_{|\mathcal{M}_{T}} are complicated and cannot be solved explicitly. Our theorem implies the existence of a large set of tensors TT such that S|TS_{|\mathcal{M}_{T}} has a global maximum, although in many cases, S|TS_{|\mathcal{M}_{T}} has no critical points at all. These examples also have interesting features for critical points besides global maxima. We find:

  • \blacktriangleright

    A tensor TT such that S|TS_{|\mathcal{M}_{T}} admits a two-dimensional submanifold of critical points. This submanifold is non-degenerate and a local maximum.

  • \blacktriangleright

    A large set of critical points that are global maxima among diagonal metrics but saddles in the space of all invariant metrics.

  • \blacktriangleright

    A set of tensors TT for which S|TS_{|\mathcal{M}_{T}} has both a circle of saddles and a strict local maximum, although for some of these TT, the functional does not achieve its global maximum.

We note that generic critical points are isolated. More precisely, as shown in [21], there exists an open and dense subset in the space of GG-invariant metrics on which the Ricci map is a local diffeomorphism (up to scaling).

Finally, we observe that the functional S|TS_{|\mathcal{M}_{T}} has a large set of critical points on flag manifolds, e.g., SU(3)/T2SU(3)/T^{2}. Let JJ be a complex structure on a compact generalized flag manifold G/H=G/C(τ)G/H=G/C(\tau), where C(τ)C(\tau) is the centralizer of a torus τ\tau in GG. It is well known that there exists a unique GG-invariant Hermitian bi-linear form hJh_{J} such that Ric(g)=hJ\operatorname{Ric}(g)=-h_{J} for every Kähler metric gg compatible with JJ; see [20]. Since the set of such metrics has the same dimension as the torus τ\tau, we obtain a smooth critical submanifold in T\mathcal{M}_{T} of dimension dimτ1\dim\tau-1. Furthermore, this critical submanifold contains the unique Kähler–Einstein metric associated to JJ. For instance, if G/H=SU(n+1)/TnG/H=SU(n+1)/T^{n}, there is an (n1)(n-1)-dimensional smooth submanifold of Kähler metrics consisting of critical points of S|TS_{|\mathcal{M}_{T}} for TT equal to hJ-h_{J} (up to scaling). In general, there exist several inequivalent complex structures on G/HG/H and hence non-isometric critical submanifolds.

We now outline the strategy of our proofs. To describe the set of homogeneous metrics on G/HG/H, one decomposes the tangent space 𝔪𝔤/𝔥{\mathfrak{m}}\simeq{\mathfrak{g}}/{\mathfrak{h}} into a sum of irreducible modules 𝔪1𝔪r{\mathfrak{m}}_{1}\oplus\cdots\oplus{\mathfrak{m}}_{r} and considers so-called diagonal metrics

g=x1Q|𝔪1+x2Q|𝔪2++xrQ|𝔪r,g=x_{1}Q_{|{\mathfrak{m}}_{1}}+x_{2}Q_{|{\mathfrak{m}}_{2}}+\cdots+x_{r}Q_{|{\mathfrak{m}}_{r}},

where QQ is a fixed bi-invariant metric on the Lie algebra of GG. Not every homogeneous metric is of this form when some of the summands 𝔪i{\mathfrak{m}}_{i} are equivalent. However, it can always be reduced to this form by choosing another decomposition of 𝔪{\mathfrak{m}} as above. With the substitution yi=1xiy_{i}=\frac{1}{x_{i}}, the constraint trgT=1\operatorname{tr}_{g}T=1 becomes simply diTiyi=1\sum d_{i}T_{i}y_{i}=1, where di=dim𝔪id_{i}=\dim{\mathfrak{m}}_{i} and T|𝔪i=TiQ|𝔪iT_{|{\mathfrak{m}}_{i}}=T_{i}\,Q_{|{\mathfrak{m}}_{i}}. Thus the set of diagonal metrics in T\mathcal{M}_{T} is parameterized by a simplex Δ\Delta with stratified boundary Δ\partial\Delta. In this construction, the strata are indexed by the sets of those variables yiy_{i} that vanish. Some of them are marked by subalgebras 𝔨{\mathfrak{k}} with 𝔥𝔨𝔤{\mathfrak{h}}\subset{\mathfrak{k}}\subset{\mathfrak{g}} and denoted Δ𝔨\Delta_{\mathfrak{k}}. If (gi)T(g_{i})\subset\mathcal{M}_{T} is a divergent sequence of metrics of bounded scalar curvature, then there exist such an intermediate subalgebra and a subsequence of (gi)(g_{i}) that converges to a point in Δ𝔨\Delta_{\mathfrak{k}}. At the remaining strata, denoted Δ\Delta_{\infty}, the scalar curvature goes to -\infty. We will in fact show that α𝔨\alpha_{\mathfrak{k}} is the largest possible limit of the scalar curvature achieved as one approaches Δ𝔨\Delta_{\mathfrak{k}}. Since S|TS_{|\mathcal{M}_{T}} is bounded from above, it is natural to look at subalgebras 𝔨{\mathfrak{k}} with α𝔨=αG/H\alpha_{\mathfrak{k}}=\alpha_{G/H} and determine conditions under which there are metrics with scalar curvature larger than αG/H\alpha_{G/H}. This is achieved by using the “derivatives” β𝔨α𝔨\beta_{\mathfrak{k}}-\alpha_{\mathfrak{k}}, and if such metrics exists, we show that {gT(G/H)S(g)>αG/H+ϵ}\{g\in\mathcal{M}_{T}(G/H)\mid S(g)>\alpha_{G/H}+\epsilon\} is compact for some ϵ>0\epsilon>0. One of the main difficulties is that, generally speaking, the scalar curvature does not extend continuously to Δ\partial\Delta and α𝔨\alpha_{\mathfrak{k}} may not be achieved by a metric on K/HK/H. In addition, β𝔨α𝔨\beta_{\mathfrak{k}}-\alpha_{\mathfrak{k}} may not be an actual derivative. This also explains, in part, why we cannot arbitrarily choose a subalgebra 𝔨{\mathfrak{k}} with α𝔨=αG/H\alpha_{\mathfrak{k}}=\alpha_{G/H} in our main theorem. For the proof, we need to produce careful estimates for the scalar curvature near Δ\partial\Delta. The final difficulty is to extend our arguments to the set of all metrics in T\mathcal{M}_{T} when some of the isotropy summands are equivalent and 𝔪{\mathfrak{m}} can be decomposed into irreducible modules in many substantially different ways. This requires further careful estimates.

The paper is organized as follows. In Section 1 we recall properties of homogeneous spaces and Riemannian submersion that we will need in our proofs. In Section 2 we describe the simplicial complex Δ\Delta and the stratification of Δ\partial\Delta by subalgebras, as well as the behavior of the scalar curvature near the strata in Δ\partial\Delta. In Section 3 we use Riemannian submersions defined by intermediate subgroups to understand when there exist metrics gg with S(g)>α𝔨S(g)>\alpha_{\mathfrak{k}} near a stratum Δ𝔨\Delta_{\mathfrak{k}}. In Section 4 we vary the decompositions and prove our main theorem. Finally, in Section 5 we study the Stiefel manifold and the Ledger–Obata space as examples.


1. Preliminaries

We first recall some basics of the geometry of a homogeneous space. Let HGH\subset G be two compact Lie groups with Lie algebras 𝔥𝔤{\mathfrak{h}}\subset{\mathfrak{g}} such that G/HG/H is an almost effective homogeneous space. We fix a bi-invariant metric QQ on 𝔤{\mathfrak{g}}, which defines a QQ-orthogonal AdH\operatorname{Ad}_{H}-invariant splitting 𝔤=𝔥𝔪{\mathfrak{g}}={\mathfrak{h}}\oplus{\mathfrak{m}}, i.e., 𝔪=𝔥{\mathfrak{m}}={\mathfrak{h}}^{\perp}. The tangent space TeH(G/H)T_{eH}(G/H) is identified with 𝔪{\mathfrak{m}}, and HH acts on 𝔪{\mathfrak{m}} via the adjoint representation AdH\operatorname{Ad}_{H}. A GG-invariant metric on G/HG/H is determined by an AdH\operatorname{Ad}_{H}-invariant inner product on 𝔪{\mathfrak{m}}. We denote by (G/H)\mathcal{M}(G/H), or sometimes simply by \mathcal{M}, the space of GG-invariant metrics on G/HG/H. We will assume throughout the paper that G/HG/H is not a torus since in this case all GG-invariant metrics are flat, in particular, Ric(g)=0\operatorname{Ric}(g)=0 for all gg\in\mathcal{M}.

We describe metrics in \mathcal{M} in terms of AdH\operatorname{Ad}_{H}-invariant splittings. Under the action of AdH\operatorname{Ad}_{H} on 𝔪{\mathfrak{m}}, we decompose

𝔪=𝔪1𝔪r,{\mathfrak{m}}={\mathfrak{m}}_{1}\oplus\ldots\oplus{\mathfrak{m}}_{r},

where AdH\operatorname{Ad}_{H} acts irreducibly on 𝔪i{\mathfrak{m}}_{i}. Some of these summands may need to be one-dimensional if there exists a subspace of 𝔪{\mathfrak{m}} on which AdH\operatorname{Ad}_{H} acts as the identity. We denote by 𝒟\mathcal{D} the space of all such decompositions and use the letter D𝒟D\in\mathcal{D} for a particular choice of decomposition. The space 𝒟\mathcal{D} has a natural topology induced from the embedding into the products of Grassmannians Gk(𝔤)G_{k}({\mathfrak{g}}) of kk-planes in 𝔤{\mathfrak{g}}. Clearly, 𝒟\mathcal{D} is compact.

If TT is a GG-invariant symmetric bi-linear form field on G/HG/H, it is determined by its value on 𝔪{\mathfrak{m}}. We are interested when such a bi-linear form field is (up to scaling) the Ricci curvature of a metric gg\in\mathcal{M}, i.e., when

(1.1) Ric(g)=cTfor some constant c.\operatorname{Ric}(g)=cT\qquad\text{for some constant $c$.}

Throughout the paper we will assume that TT is positive-definite. We may also assume that c>0c>0 in (1.1) since a compact homogeneous space does not admit any metrics with Ric0\operatorname{Ric}\leq 0 unless it is a torus (see [4, Theorem 1.84]), which we excluded above.

Define the hypersurface

T(G/H)={gtrgT=1},\mathcal{M}_{T}(G/H)=\{g\in\mathcal{M}\mid\operatorname{tr}_{g}T=1\}\subset\mathcal{M},

where trg\operatorname{tr}_{g} is the trace with respect to gg. We denote it simply by T\mathcal{M}_{T} when the homogeneous space is clear from context. As shown in [25], a solution to (1.1) can we viewed as a critical point of the functional

S:T,S\colon\mathcal{M}_{T}\to{\mathbb{R}},

where S(g)S(g) is the scalar curvature of gg. More precisely, the following result holds.

Proposition 1.2.

The Ricci curvature of a metric gTg\in\mathcal{M}_{T} equals cTcT for some cc\in\mathbb{R} if and only if gg is a critical point of S|TS_{|\mathcal{M}_{T}}.

Here cc is the Lagrange multiplier of the variational problem. Our main interest in this paper is to describe the geometry of the functional SS and its implications for when (1.1) has a solution.

We now recall the formulas for the scalar curvature and the Ricci curvature of a homogeneous metric. Given gg\in\mathcal{M}, we have Q|𝔪i=xiQ|𝔪iQ_{|{\mathfrak{m}}_{i}}=x_{i}\,Q_{|{\mathfrak{m}}_{i}} for some constant xi>0x_{i}>0. In general, 𝔪i{\mathfrak{m}}_{i} and 𝔪j{\mathfrak{m}}_{j} do not have to be orthogonal if some of these summands are equivalent. But we can diagonalize gg and QQ simultaneously, and hence there exists a decomposition D𝒟D\in\mathcal{D} such that the metric has the form

(1.3) g=x1Q|𝔪1+x2Q|𝔪2++xrQ|𝔪r.g=x_{1}Q_{|{\mathfrak{m}}_{1}}+x_{2}Q_{|{\mathfrak{m}}_{2}}+\cdots+x_{r}Q_{|{\mathfrak{m}}_{r}}.

We call such metrics diagonal with respect to our choice of DD and denote their set by D(G/H)\mathcal{M}^{D}(G/H), or simply D\mathcal{M}^{D}. Thus, =D𝒟D\mathcal{M}=\cup_{D\in\mathcal{D}}\mathcal{M}^{D}. We also denote TD=TD\mathcal{M}_{T}^{D}=\mathcal{M}_{T}\cap\mathcal{M}^{D}. The scalars xix_{i} are simply the eigenvalues of gg with respect to QQ. When these eigenvalues have multiplicity, and some of the modules in the corresponding eigenspace are equivalent under the action of AdH\operatorname{Ad}_{H}, we note that gDg\in\mathcal{M}^{D} for a compact infinite family of decompositions DD. In order to describe all homogeneous metrics on G/HG/H, we can thus restrict ourselves to diagonal metrics but allow the decomposition to change. The advantage is that, while the scalar curvature of a homogeneous metric for a fixed decomposition is quite complicated and hence somewhat intractable, for a diagonal metric it has a much simpler form. This idea was first used in [28] to study GG-invariant Einstein metrics.

We define the structure constants

[ijk]=α,β,γQ([eα,eβ],eγ)2,i,j,k=1,,r,[ijk]=\sum_{\alpha,\beta,\gamma}Q([e_{\alpha},e_{\beta}],e_{\gamma})^{2},\qquad i,j,k=1,\cdots,r,

where (eα)(e_{\alpha}), (eβ)(e_{\beta}) and (eγ)(e_{\gamma}) are QQ-orthonormal bases of 𝔪i{\mathfrak{m}}_{i}, 𝔪j{\mathfrak{m}}_{j} and 𝔪k{\mathfrak{m}}_{k}. Clearly, [ijk]0[ijk]\geq 0, and [ijk]=0[ijk]=0 if and only if Q([𝔪i,𝔪j],𝔪k)=0Q([{\mathfrak{m}}_{i},{\mathfrak{m}}_{j}],{\mathfrak{m}}_{k})=0. We will denote by BB the Killing form of GG. By the irreducibility of 𝔪i{\mathfrak{m}}_{i}, there exist constants bi0b_{i}\geq 0 such that

B|𝔪i=biQ|𝔪iB_{|{\mathfrak{m}}_{i}}=-b_{i}Q_{|{\mathfrak{m}}_{i}}

with bi=0b_{i}=0 if and only if 𝔪i{\mathfrak{m}}_{i} lies in the center of 𝔤{\mathfrak{g}}. Furthermore, not all of bib_{i} vanish since otherwise 𝔪{\mathfrak{m}} is in the center 𝔷(𝔤){\mathfrak{z}}({\mathfrak{g}}) and G/HG/H is a torus. Using this notation, and di=dim𝔪id_{i}=\dim{\mathfrak{m}}_{i}, the scalar curvature is given by

(1.4) S(g)\displaystyle S(g) =12idibixi14i,j,k[ijk]xkxixj\displaystyle=\frac{1}{2}\sum_{i}\frac{d_{i}b_{i}}{x_{i}}-\frac{1}{4}\sum_{i,j,k}[ijk]\frac{x_{k}}{x_{i}x_{j}}

(see [28]). For the tensor TT, we introduce the constants TiT_{i} such that

T|𝔪i=TiQ|𝔪i.T_{|{\mathfrak{m}}_{i}}=T_{i}\,Q_{|{\mathfrak{m}}_{i}}.

Varying over all decompositions, these constants determine TT uniquely.

We now recall some formulas for Riemannian submersions which will be useful for us. Let KK be a compact subgroup with Lie algebra 𝔨{\mathfrak{k}} lying between HH and GG, i.e., HKGH\subset K\subset G. We then have a homogeneous fibration

K/HG/HG/K.K/H\to G/H\to G/K.

We will often consider metrics with respect to which the projection G/HG/KG/H\to G/K is a Riemannian submersion. Assume that the decomposition 𝔪=(𝔨𝔪)𝔨{\mathfrak{m}}=({\mathfrak{k}}\cap{\mathfrak{m}})\oplus{\mathfrak{k}}^{\perp} is orthogonal with respect to both QQ and gg. Then gg\in\mathcal{M} is a Riemannian submersion metric if and only if g|𝔨g_{|{\mathfrak{k}}^{\perp}} is AdK\operatorname{Ad}_{K}-invariant. In this case, g|𝔨𝔪g_{|{\mathfrak{k}}\cap{\mathfrak{m}}} can be thought of as a homogeneous metric on the fiber F=K/HF=K/H, and g|𝔨g_{|{\mathfrak{k}}^{\perp}} a homogeneous metric on the base B=G/KB=G/K. We can introduce a new submersion metric gs,tg_{s,t} on G/HG/H by scaling the fiber and the base, i.e.,

gs,t=1sgF+1tgB, where gF=g|𝔨𝔪andgB=g|𝔨.g_{s,t}=\frac{1}{s}\,g_{F}+\frac{1}{t}\,g_{B},\qquad\text{ where }\qquad g_{F}=g_{|{\mathfrak{k}}\cap{\mathfrak{m}}}\qquad\text{and}\qquad g_{B}=g_{|{\mathfrak{k}}^{\perp}}.

One has the following formula for the scalar curvature (see, e.g., [4, Proposition 9.70]):

(1.5) S(gs,t)=sSF+tSBt2s|A|g,S(g_{s,t})=sS_{F}+tS_{B}-\frac{t^{2}}{s}|A|_{g},

where AA is the O’Neill tensor of the submersion, while SFS_{F} and SBS_{B} are the scalar curvatures of gFg_{F} and gBg_{B}. The proof of this formula is a pointwise calculation and hence extends to the situation where KK is not compact (i.e., G/KG/K is not a manifold). In our case it can indeed happen that for some Lie algebra 𝔨𝔤{\mathfrak{k}}\subset{\mathfrak{g}} the connected subgroup KGK\subset G with Lie algebra 𝔨{\mathfrak{k}} is not compact. However, this will not affect our discussions.

Let N(H)N(H) denote the normalizer of HH. Every element nN(H)n\in N(H) acts on G/HG/H as a diffeomorphism via right translation: Rn(gH)=gn1HR_{n}(gH)=gn^{-1}H. Combining this with the action by left translation, we obtain an action on the space of metrics via pullback: g(Adn)(g)\mathcal{M}\ni g\mapsto(\operatorname{Ad}_{n})^{*}(g)\in\mathcal{M}. This induces an isometry between gg and (Adn)(g)(\operatorname{Ad}_{n})^{*}(g), and hence S((Adn)(g))=S(g)S((\operatorname{Ad}_{n})^{*}(g))=S(g) for the scalar curvature. This also holds for the possibly larger group Aut(G,H)\operatorname{Aut}(G,H) of automorphisms of GG that preserve HH. Note though that S|TS_{|\mathcal{M}_{T}} is invariant under Aut(G,H)\operatorname{Aut}(G,H), or one of its subgroups, only if TT is as well, in which case the orbit of a critical point consists of further critical points.

The following remark, based on Palais’s principle of symmetric criticality, will be useful for us. If TT is invariant under a subgroup LAut(G,H)L\subset\operatorname{Aut}(G,H), and if TL\mathcal{M}_{T}^{L} is the set metrics in T\mathcal{M}_{T} invariant under LL, then critical points of the restriction S|TLS_{|\mathcal{M}_{T}^{L}} are also critical points of S|TS_{|\mathcal{M}_{T}}. Indeed, given gTLg\in\mathcal{M}_{T}^{L}, the Ricci curvature Ric(g)\operatorname{Ric}(g) and hence the gradient gradS|T(g)\operatorname{grad}S_{|\mathcal{M}_{T}}(g) is invariant under LL. Consequently, gradS|T(g)\operatorname{grad}S_{|\mathcal{M}_{T}}(g) must be tangent to TL\mathcal{M}_{T}^{L}.

In the remainder of the paper we will assume for simplicity that GG, HH and all the intermediate subgroups are connected. Let us explain why this is in fact not necessary. One only needs to make the following modification. If GG or HH is not connected, we consider only intermediate subalgebras 𝔨{\mathfrak{k}} that are Lie algebras of intermediates subgroups HKGH\subset K\subset G. This is easily seen to be equivalent to saying that 𝔨{\mathfrak{k}} must be invariant under AdH\operatorname{Ad}_{H}. The proofs of all of our results apply without any changes, and the conclusions are also the same. This is useful when the space of invariants has large dimension or there are many intermediate subgroups. Adding components to GG and HH will reduce dimT\dim\mathcal{M}_{T} and may easily imply the existence of some critical points. This happens, for instance, when G/HG/H is isotropy irreducible; see [29] for many examples.


2. The simplicial complex

In Sections 23 we fix the decomposition D𝒟D\in\mathcal{D} and study the set of metrics D\mathcal{M}^{D} diagonal with respect to DD.

It will be convenient for us to describe a homogeneous metric in terms of its inverse since this makes the space of metrics precompact. If gTDg\in\mathcal{M}_{T}^{D} is given by (1.3), we set yi=1xiy_{i}=\frac{1}{x_{i}} and obtain the following formulas for the scalar curvature and its constraint:

(2.1) S(g)=12idibiyi14i,j,kyiyjyk[ijk],trgT=idiTiyi=1.S(g)=\frac{1}{2}\sum_{i}d_{i}b_{i}y_{i}-\frac{1}{4}\sum_{i,j,k}\frac{y_{i}y_{j}}{y_{k}}[ijk],\qquad\operatorname{tr}_{g}T=\sum_{i}d_{i}T_{i}y_{i}=1.

We need to study the behavior of S|TDS_{|\mathcal{M}_{T}^{D}} at infinity, which means that at least one of the variables yiy_{i} goes to 0. It is natural to introduce a simplicial complex and its stratification. Specifically, let

Δ=ΔD={(y1,,yr)r|idiTiyi=1 and yi>0}.\Delta=\Delta^{D}=\big{\{}(y_{1},\cdots,y_{r})\in{\mathbb{R}}^{r}\,\big{|}\,\textstyle\sum_{i}d_{i}T_{i}y_{i}=1\text{ and }y_{i}>0\big{\}}.

Notice that the numbers TiT_{i}, and thus the simplex Δ\Delta, depend on the choice of DD. This simplex is a natural parametrization of the set TD\mathcal{M}_{T}^{D}. We identify a metric gTDg\in\mathcal{M}_{T}^{D} with y=(y1,,yr)Δy=(y_{1},\cdots,y_{r})\in\Delta.

The boundary of Δ\Delta consists of lower-dimensional simplices. For every nonempty proper subset JJ of the index set I={1,,r}I=\{1,\cdots,r\}, let

ΔJ={yΔyi>0 for iJ,yi=0 for iJc}.\Delta_{J}=\{y\in\partial\Delta\mid y_{i}>0\text{ for }i\in J,\ y_{i}=0\text{ for }i\in J^{c}\}.

Thus ΔJ\Delta_{J} is a |J||J|-dimensional simplex, which we call a stratum of Δ\partial\Delta. The closure of ΔJ\Delta_{J} satisfies

Δ¯J=JJΔJ,\bar{\Delta}_{J}=\bigcup_{J^{\prime}\subset J}\Delta_{J^{\prime}},

and we call ΔJ\Delta_{J^{\prime}} a stratum adjacent to ΔJ\Delta_{J} if JJ^{\prime} is a nonempty proper subset of JJ. It will also be useful for us to consider tubular ϵ\epsilon-neighborhoods of strata for ϵ>0\epsilon>0:

Tϵ(ΔJ)={yΔyiϵ for iJc}.T_{\epsilon}(\Delta_{J})=\{y\in\Delta\mid y_{i}\leq\epsilon\text{ for }i\in J^{c}\}.

Finally, we associate to each stratum an AdH\operatorname{Ad}_{H}-invariant subspace of 𝔪{\mathfrak{m}}:

𝔪J=iJ𝔪i.{\mathfrak{m}}_{J}=\bigoplus_{i\in J}{\mathfrak{m}}_{i}.

We can fill out the closure Δ¯\bar{\Delta} with geodesics starting at the center. To this end, consider the unit sphere

𝕊={vr|idiTivi=0,vi2=1}\mathbb{S}=\big{\{}v\in{\mathbb{R}}^{r}\,\big{|}\,\textstyle\sum_{i}d_{i}T_{i}v_{i}=0,~{}\sum v_{i}^{2}=1\big{\}}

of dimension r2r-2. Define a geodesic γv:[0,tv]Δ¯\gamma_{v}\colon[0,t_{v}]\to\bar{\Delta} by setting γv(t)=v0tv\gamma_{v}(t)=v_{0}-tv, where

v𝕊,tv=1rmaxidiTivi, and v0=(v01,,v0r)=(1rd1T1,,1rdrTr).v\in\mathbb{S},\qquad t_{v}=\frac{1}{r\max_{i}d_{i}T_{i}v_{i}},\qquad\text{ and }\qquad v_{0}=(v_{01},\ldots,v_{0r})=\Big{(}\frac{1}{rd_{1}T_{1}},\ldots,\frac{1}{rd_{r}T_{r}}\Big{)}.

The stratification of Δ\partial\Delta induces one of the sphere:

𝕊J={v𝕊γv(tv)ΔJ}.\mathbb{S}_{J}=\{v\in\mathbb{S}\mid\gamma_{v}(t_{v})\in\Delta_{J}\}.

Our first observation is that we can mark the strata with subalgebras.

Proposition 2.2.

The functional S|TDS_{|\mathcal{M}_{T}^{D}} is bounded from above. Furthermore, for any v𝕊v\in\mathbb{S} either S(γv(t))S(\gamma_{v}(t))\to-\infty as ttvt\to t_{v} or v𝕊Jv\in\mathbb{S}_{J} for some JJ such that 𝔥𝔪J{\mathfrak{h}}\oplus{\mathfrak{m}}_{J} is a subalgebra of 𝔤{\mathfrak{g}}.

Proof.

Let A=maxibiminiTiA=\frac{\max_{i}b_{i}}{\min_{i}T_{i}}, which is well-defined since Ti>0T_{i}>0 by assumption. We also have A>0A>0 since G/HG/H is not a torus. Then (2.1) implies that S(g)AS(g)\leq A.

For the second claim, let JJ be the index set with Jc={iv0tvvi=0}J^{c}=\{i\mid v_{0}-t_{v}v_{i}=0\}. Obviously, JIJ\neq I and tv>0t_{v}>0. If 𝔥𝔪J{\mathfrak{h}}\oplus{\mathfrak{m}}_{J} is not a subalgebra, then there exist i,jJi,j\in J and kJc\ k\in J^{c} such that [ijk]0[ijk]\neq 0. Then in formula (2.1), we have a contribution of the form

[ijk]yiyjyk=[ijk](v0itvi)(v0jtvj)v0ktvk-[ijk]\,\frac{y_{i}y_{j}}{y_{k}}=-[ijk]\,\frac{(v_{0i}-tv_{i})(v_{0j}-tv_{j})}{v_{0k}-tv_{k}}

with 0t<tv0\leq t<t_{v}. Since i,jJi,j\in J and kJck\in J^{c}, we know that (v0itvi)(v0jtvj)(v_{0i}-tv_{i})(v_{0j}-tv_{j}) stays bounded away from 0 and v0ktvk0v_{0k}-tv_{k}\to 0 as ttvt\to t_{v}. This implies that S(γv(t))S(\gamma_{v}(t))\to-\infty as ttvt\to t_{v}. ∎

We also need to control how fast the scalar curvature goes to -\infty. For this purpose we prove the following result.

Proposition 2.3.

Consider a stratum ΔJ\Delta_{J} such that 𝔥𝔪J{\mathfrak{h}}\oplus{\mathfrak{m}}_{J} is not a subalgebra. Then for every v𝕊Jv\in\mathbb{S}_{J} and a>0a>0, there exist an open neighborhood U(v)U(v) in 𝕊\mathbb{S} and a positive number ϵ(v)\epsilon(v) such that S(γu(t))<aS(\gamma_{u}(t))<-a whenever uU(v)u\in U(v) and (1ϵ(v))tu<t<tu(1-\epsilon(v))t_{u}<t<t_{u}.

Proof.

Let A=maxibiminiTiA=\frac{\max_{i}b_{i}}{\min_{i}T_{i}} as before. There exist i,jJi,j\in J and kJck\in J^{c} such that [ijk]0[ijk]\neq 0. Moreover, (v0itvvi)(v0jtvvj)>0(v_{0i}-t_{v}v_{i})(v_{0j}-t_{v}v_{j})>0 and (v0ktvvk)=0(v_{0k}-t_{v}v_{k})=0. Define

ϵ(v)=min{[ijk](v0itvvi+)(v0jtvvj+)4(A+a)tvvk,12},\epsilon(v)=\min\Big{\{}[ijk]\frac{(v_{0i}-t_{v}v_{i}^{+})(v_{0j}-t_{v}v_{j}^{+})}{4(A+a)t_{v}v_{k}},\frac{1}{2}\Big{\}},

where vi+=max{vi,0}v_{i}^{+}=\max\{v_{i},0\}. Evidently, this quantity is always positive. Choose a neighborhood U(v)U(v) of vv in 𝕊\mathbb{S} such that uk>0u_{k}>0 and

(v0i\displaystyle(v_{0i} tuui+)(v0jtuuj+)>12(v0itvvi+)(v0jtvvj+)and\displaystyle-t_{u}u_{i}^{+})(v_{0j}-t_{u}u_{j}^{+})>\tfrac{1}{2}(v_{0i}-t_{v}v_{i}^{+})(v_{0j}-t_{v}v_{j}^{+})\qquad\mbox{and}
v0k\displaystyle v_{0k} (1ϵ(v))tuuk<2(v0k(1ϵ(v))tvvk)=2(v0ktvvk+ϵ(v)tvvk)=2ϵ(v)tvvk\displaystyle-(1-\epsilon(v))t_{u}u_{k}<2(v_{0k}-(1-\epsilon(v))t_{v}v_{k})=2(v_{0k}-t_{v}v_{k}+\epsilon(v)t_{v}v_{k})=2\epsilon(v)t_{v}v_{k}

for all u=(u1,,ur)U(v)u=(u_{1},\ldots,u_{r})\in U(v). This implies

S(γu(t))\displaystyle S(\gamma_{u}(t)) A[ijk](v0itui)(v0jtuj)v0ktuk\displaystyle\leq A-[ijk]\frac{(v_{0i}-tu_{i})(v_{0j}-tu_{j})}{v_{0k}-tu_{k}}
<A[ijk](v0ituui+)(v0jtuuj+)v0k(1ϵ(v))tuuka,\displaystyle<A-[ijk]\frac{(v_{0i}-t_{u}u_{i}^{+})(v_{0j}-t_{u}u_{j}^{+})}{v_{0k}-(1-\epsilon(v))t_{u}u_{k}}\leq-a,

provided uU(v)u\in U(v) and (1ϵ(v))tu<t<tu(1-\epsilon(v))t_{u}<t<t_{u}. ∎

If the space G/HG/H has pairwise inequivalent isotropy summands, Proposition 2.3 implies the following result, originally proved in [25].

Corollary 2.4.

If 𝔥{\mathfrak{h}} is maximal in 𝔤{\mathfrak{g}}, then S|TS_{|\mathcal{M}_{T}} attains its global maximum at a metric gT{g\in\mathcal{M}_{T}}, and hence Ric(g)=cT\operatorname{Ric}(g)=cT for some c>0c>0.

We can add a marking to the strata in Δ\partial\Delta. If 𝔥𝔪J=𝔨{\mathfrak{h}}\oplus{\mathfrak{m}}_{J}={\mathfrak{k}} is a subalgebra, we denote the stratum ΔJ\Delta_{J} by (ΔJ,𝔨)(\Delta_{J},{\mathfrak{k}}) or simply Δ𝔨\Delta_{\mathfrak{k}}; if it is not, we denote the stratum by (ΔJ,)(\Delta_{J},\infty) or Δ\Delta_{\infty}.

Next, we need an estimate for SS near (ΔJ,𝔨)(\Delta_{J},{\mathfrak{k}}). Let KK be the connected subgroup of GG with Lie algebra 𝔨{\mathfrak{k}}. Define

α𝔨D=sup{S(h)hD(K/H) with trhT|𝔨𝔪=1},\displaystyle\alpha_{\mathfrak{k}}^{D}=\sup\{S(h)\mid h\in\mathcal{M}^{D}(K/H)\text{ with }\operatorname{tr}_{h}T_{|{\mathfrak{k}}\cap{\mathfrak{m}}}=1\},

where the letter DD is preserved for the decomposition of the tangent space to K/HK/H induced by DD. Since a normal homogeneous metric has non-negative scalar curvature, α𝔨D0\alpha_{\mathfrak{k}}^{D}\geq 0. Also, α𝔨D=0\alpha_{\mathfrak{k}}^{D}=0 if and only if K/HK/H is flat. It is important for us to note that, since Δ𝔨\Delta_{\mathfrak{k}} is not closed in general, the supremum in the definition of α𝔨D\alpha_{\mathfrak{k}}^{D} may not be achieved in Δ𝔨\Delta_{\mathfrak{k}}, which will complicate our discussion.

Consider a metric in TD\mathcal{M}_{T}^{D} identified with y=(y1,,yr)Δy=(y_{1},\ldots,y_{r})\in\Delta. If 𝔨=𝔥𝔪J{\mathfrak{k}}={\mathfrak{h}}\oplus{\mathfrak{m}}_{J} for some JIJ\subset I, then

y=y|𝔨𝔪+y|𝔨,y|𝔨𝔪=iJ1yiQ|𝔪i,y|𝔨=iJc1yiQ|𝔪i.\displaystyle y=y_{|{\mathfrak{k}}\cap{\mathfrak{m}}}+y_{|{\mathfrak{k}}^{\perp}},\qquad y_{|{\mathfrak{k}}\cap{\mathfrak{m}}}=\sum_{i\in J}\tfrac{1}{y_{i}}Q_{|{\mathfrak{m}}_{i}},\qquad y_{|{\mathfrak{k}}^{\perp}}=\sum_{i\in J^{c}}\tfrac{1}{y_{i}}Q_{|{\mathfrak{m}}_{i}}.

We may regard y|𝔨𝔪y_{|{\mathfrak{k}}\cap{\mathfrak{m}}} as a metric on K/HK/H. Its scalar curvature is given by

S(y|𝔨𝔪)=12iJdib¯iyi14i,j,kJ[ijk]yiyjyk,\displaystyle S(y_{|{\mathfrak{k}}\cap{\mathfrak{m}}})=\frac{1}{2}\sum_{i\in J}d_{i}\bar{b}_{i}y_{i}-\frac{1}{4}\sum_{i,j,k\in J}[ijk]\frac{y_{i}y_{j}}{y_{k}},

where the Killing form of KK restricted to 𝔪i{\mathfrak{m}}_{i} equals b¯iQ|𝔪i-\bar{b}_{i}Q_{|{\mathfrak{m}}_{i}}. One easily shows that

b¯i=bij,kJc[ijk]di.\bar{b}_{i}=b_{i}-\sum_{j,k\in J^{c}}\frac{[ijk]}{d_{i}}.
Proposition 2.5.

Consider a stratum (ΔJ,𝔨)(\Delta_{J},{\mathfrak{k}}). If yTDy\in\mathcal{M}_{T}^{D} satisfies maxiJcyiϵ\displaystyle\max_{i\in J^{c}}y_{i}\leq\epsilon, then

S(y)α𝔨D+ϵiJcdibi2.S(y)\leq\alpha_{\mathfrak{k}}^{D}+\epsilon\sum_{i\in J^{c}}\frac{d_{i}b_{i}}{2}.
Proof.

We break up the formula for the scalar curvature in (2.1) as follows, using the assumption that [ijk]=0[ijk]=0 for i,jJi,j\in J and kJck\in J^{c}:

S(y)\displaystyle S(y) =12iJdibiyi+12iJcdibiyi14i,j,kJ[ijk]yiyjyk\displaystyle=\frac{1}{2}\sum_{i\in J}d_{i}b_{i}y_{i}+\frac{1}{2}\sum_{i\in J^{c}}d_{i}b_{i}y_{i}-\frac{1}{4}\sum_{i,j,k\in J}[ijk]\frac{y_{i}y_{j}}{y_{k}}
12iJj,kJc[ijk]yiyjyk14iJj,kJc[ijk]yjykyi14i,j,kJc[ijk]yiyjyk\displaystyle\hphantom{=}~{}-\frac{1}{2}\sum_{i\in J}\sum_{j,k\in J^{c}}[ijk]\frac{y_{i}y_{j}}{y_{k}}-\frac{1}{4}\sum_{i\in J}\sum_{j,k\in J^{c}}[ijk]\frac{y_{j}y_{k}}{y_{i}}-\frac{1}{4}\sum_{i,j,k\in J^{c}}[ijk]\frac{y_{i}y_{j}}{y_{k}}
12iJdib¯iyi+12iJj,kJc[ijk]yi+12iJcdibiyi14i,j,kJ[ijk]yiyjyk\displaystyle\leq\frac{1}{2}\sum_{i\in J}d_{i}{\bar{b}}_{i}y_{i}+\frac{1}{2}\sum_{i\in J}\sum_{j,k\in J^{c}}[ijk]y_{i}+\frac{1}{2}\sum_{i\in J^{c}}d_{i}b_{i}y_{i}-\frac{1}{4}\sum_{i,j,k\in J}[ijk]\frac{y_{i}y_{j}}{y_{k}}
14iJj,kJc[ijk]yi(yjyk+ykyj)\displaystyle\hphantom{=}~{}-\frac{1}{4}\sum_{i\in J}\sum_{j,k\in J^{c}}[ijk]y_{i}\Big{(}\frac{y_{j}}{y_{k}}+\frac{y_{k}}{y_{j}}\Big{)}
S(y|𝔨𝔪)+12iJcdibiyi,\displaystyle\leq S(y_{|{\mathfrak{k}}\cap{\mathfrak{m}}})+\frac{1}{2}\sum_{i\in J^{c}}d_{i}b_{i}y_{i},

where in the last step we used the estimate yjyk+ykyj2\frac{y_{j}}{y_{k}}+\frac{y_{k}}{y_{j}}\geq 2. Now observe that

S(y|𝔨𝔪)=(try|𝔨𝔪T|𝔨𝔪)S((try|𝔨𝔪T|𝔨𝔪)y|𝔨𝔪)<S((try|𝔨𝔪T|𝔨𝔪)y|𝔨𝔪)α𝔨D\displaystyle S(y_{|{\mathfrak{k}}\cap{\mathfrak{m}}})=(\operatorname{tr}_{y_{|{\mathfrak{k}}\cap{\mathfrak{m}}}}T_{|{\mathfrak{k}}\cap{\mathfrak{m}}})S((\operatorname{tr}_{y_{|{\mathfrak{k}}\cap{\mathfrak{m}}}}T_{|{\mathfrak{k}}\cap{\mathfrak{m}}})y_{|{\mathfrak{k}}\cap{\mathfrak{m}}})<S((\operatorname{tr}_{y_{|{\mathfrak{k}}\cap{\mathfrak{m}}}}T_{|{\mathfrak{k}}\cap{\mathfrak{m}}})y_{|{\mathfrak{k}}\cap{\mathfrak{m}}})\leq\alpha_{\mathfrak{k}}^{D}

since try|𝔨𝔪T|𝔨𝔪<tryT=1\operatorname{tr}_{y_{|{\mathfrak{k}}\cap{\mathfrak{m}}}}T_{|{\mathfrak{k}}\cap{\mathfrak{m}}}<\operatorname{tr}_{y}T=1 and the trace of T|𝔨𝔪T_{|{\mathfrak{k}}\cap{\mathfrak{m}}} with respect to (try|𝔨𝔪T|𝔨𝔪)y|𝔨𝔪(\operatorname{tr}_{y_{|{\mathfrak{k}}\cap{\mathfrak{m}}}}T_{|{\mathfrak{k}}\cap{\mathfrak{m}}})y_{|{\mathfrak{k}}\cap{\mathfrak{m}}} equals 1. Consequently,

S(y)<α𝔨D+12iJcdibiyi.\displaystyle S(y)<\alpha_{\mathfrak{k}}^{D}+\frac{1}{2}\sum_{i\in J^{c}}d_{i}b_{i}y_{i}.

When yi<ϵy_{i}<\epsilon for all iJci\in J^{c}, we get the desired result. ∎

We can reformulate Proposition 2.5 as follows.

Corollary 2.6.

Let Δ𝔨D\Delta_{\mathfrak{k}}^{D} be a subalgebra stratum. Then for every a>α𝔨Da>\alpha_{\mathfrak{k}}^{D} there exists a constant ϵ>0\epsilon>0 such that the set {gTDS(g)a}\{g\in\mathcal{M}_{T}^{D}\mid S(g)\geq a\} does not intersect Tϵ(Δ𝔨)T_{\epsilon}(\Delta_{\mathfrak{k}}).

Combining Propositions 2.3 and 2.5, we arrive at the following conclusion.

Corollary 2.7.

Suppose a>α𝔨Da>\alpha_{\mathfrak{k}}^{D} for every subalgebra stratum Δ𝔨\Delta_{\mathfrak{k}}. Then {gTDS(g)a}\{g\in\mathcal{M}_{T}^{D}\mid S(g)\geq a\} is a (possibly empty) compact subset of TD\mathcal{M}_{T}^{D}.

Proposition 2.5 shows that α𝔨D\alpha_{\mathfrak{k}}^{D} is an upper bound for the possible values of the scalar curvature as we approach points in Δ𝔨\Delta_{\mathfrak{k}}. However, it is important to keep in mind that SS does not, in general, extend continuously to the closure of Δ\Delta.

We end this section with the following observation. Recall that even if GG and HH are connected and compact, and if 𝔨{\mathfrak{k}} is an intermediate subalgebra, then the connected (intermediate) subgroup with Lie algebra 𝔨{\mathfrak{k}} is not necessarily compact.

Lemma 2.8.

If KK is an intermediate subgroup with Lie algebra 𝔨{\mathfrak{k}}, then the closure K¯\bar{K} is compact and α𝔨¯D=α𝔨D\alpha_{\bar{\mathfrak{k}}}^{D}=\alpha_{\mathfrak{k}}^{D}, where 𝔨¯\bar{\mathfrak{k}} is the Lie algebra of K¯\bar{K}.

Proof.

Let 𝔨=𝔨s𝔷{\mathfrak{k}}={\mathfrak{k}}_{\mathrm{s}}\oplus{\mathfrak{z}} with 𝔨s{\mathfrak{k}}_{\mathrm{s}} semisimple and 𝔷{\mathfrak{z}} the center of 𝔨{\mathfrak{k}}. Suppose that KsK_{\mathrm{s}} and ZZ are the connected Lie subgroups of GG with Lie algebras 𝔨s{\mathfrak{k}}_{\mathrm{s}} and 𝔷{\mathfrak{z}}. Then K=KsZK=K_{\mathrm{s}}\cdot Z (i.e., KK is the quotient of Ks×ZK_{\mathrm{s}}\times Z by a discrete subgroup of the center of Ks×ZK_{\mathrm{s}}\times Z) and KsK_{\mathrm{s}} is compact. The closure Z¯\bar{Z} is compact abelian, and hence a torus. Denote its Lie algebra by 𝔷¯\bar{\mathfrak{z}}. Thus K¯=KsZ¯\bar{K}=K_{\mathrm{s}}\cdot\bar{Z}. If g(K/H)g\in\mathcal{M}(K/H), then any extension g¯(K¯/H)\bar{g}\in\mathcal{M}({\bar{K}}/H) satisfies S(g¯)=S(g)S(\bar{g})=S(g) since [𝔷¯,𝔷¯]=[𝔷¯,𝔨]=0[\bar{\mathfrak{z}},\bar{\mathfrak{z}}]=[\bar{\mathfrak{z}},{\mathfrak{k}}]=0. ∎

Thus it is sufficient to compute the invariants α𝔨D\alpha_{\mathfrak{k}}^{D} only for intermediate subalgebras for which KK is compact.


3. Riemannian submersions

In this section we study the behavior of SS near the subalgebra stratum Δ𝔨\Delta_{\mathfrak{k}} geometrically. It will be more convenient to choose the path gtg_{t} below instead of γv(t)\gamma_{v}(t) since we can then use formula (1.5) for Riemannian submersions. The goal is to see if there are metrics near Δ𝔨\Delta_{\mathfrak{k}} whose scalar curvature is larger than α𝔨D\alpha_{\mathfrak{k}}^{D}. As before, suppose KK is an intermediate connected subgroup with Lie algebra 𝔨{\mathfrak{k}} and associated stratum Δ𝔨\Delta_{\mathfrak{k}}. Thus 𝔨=𝔥𝔪J\mathfrak{k}={\mathfrak{h}}\oplus{\mathfrak{m}}_{J} for some JIJ\subset I. Define

β𝔨D=sup{S(h)hD(G/K) with trhT|𝔨=1}.\beta_{\mathfrak{k}}^{D}=\sup\{S(h)\mid h\in\mathcal{M}^{D}(G/K)\text{ with }\operatorname{tr}_{h}T_{|{\mathfrak{k}}^{\perp}}=1\}.

We will show that β𝔨α𝔨\beta_{\mathfrak{k}}-\alpha_{\mathfrak{k}} controls the desired behavior.

We have the homogeneous fibration

(3.1) F=K/HG/HG/K=B.F=K/H\to G/H\to G/K=B.

Let us consider metrics on G/HG/H for which this fibration is a Riemannian submersion. We start with a metric of the form

g=gF+gB=iJ1yiQ|𝔪i+iJc1yiQ|𝔪i.g=g_{F}+g_{B}=\sum_{i\in J}\tfrac{1}{y_{i}}Q_{|{\mathfrak{m}}_{i}}+\sum_{i\in J^{c}}\tfrac{1}{y_{i}}Q_{|{\mathfrak{m}}_{i}}.

Assume that gg lies in T\mathcal{M}_{T}, i.e., T1+T2=1T_{1}^{*}+T_{2}^{*}=1, where

T1=trgFT|𝔨𝔪=iJdiyiTi,T2=trgBT|𝔨=iJcdiyiTi.\displaystyle T_{1}^{*}=\operatorname{tr}_{g_{F}}T_{|{\mathfrak{k}}\cap{\mathfrak{m}}}=\sum_{i\in J}d_{i}y_{i}T_{i},\qquad T_{2}^{*}=\operatorname{tr}_{g_{B}}T_{|{\mathfrak{k}}^{\perp}}=\sum_{i\in J^{c}}d_{i}y_{i}T_{i}.

We also require the metric gBg_{B} to be AdK\operatorname{Ad}_{K}-invariant so that the projection in (3.1) is a Riemannian submersion with gFg_{F} and gBg_{B} the metrics on the fiber and the base.

Consider the two-parameter family

gs,t=1sgF+1tgBwithsT1+tT2=1.g_{s,t}=\frac{1}{s}g_{F}+\frac{1}{t}g_{B}\qquad\text{with}\qquad sT_{1}^{*}+tT_{2}^{*}=1.

Substituting s=1tT2T1s=\frac{1-tT_{2}^{*}}{T_{1}^{*}}, we obtain a one-parameter family of metrics

(3.2) gt=T11tT2gF+1tgB\displaystyle g_{t}=\frac{T_{1}^{*}}{1-tT_{2}^{*}}\,g_{F}+\frac{1}{t}\,g_{B}

lying in TD\mathcal{M}_{T}^{D}. We call this the canonical variation associated to KK. By (1.5), the scalar curvature of gtg_{t} is

(3.3) S(gt)=sSF+tSBt2s|A|g=SFT1+T2(SBT2SFT1)tt2T11tT2|A|g.S(g_{t})=s\,S_{F}+t\,S_{B}-\frac{t^{2}}{s}|A|_{g}=\frac{S_{F}}{T_{1}^{*}}+T_{2}^{*}\Big{(}\frac{S_{B}}{T_{2}^{*}}-\frac{S_{F}}{T_{1}^{*}}\Big{)}t-\frac{t^{2}T_{1}^{*}}{1-tT_{2}^{*}}|A|_{g}.

Since limt0gt=gFT1Δ𝔨\lim_{t\to 0}g_{t}=\frac{g_{F}}{T_{1}^{*}}\in\Delta_{\mathfrak{k}}, every point in Δ𝔨\Delta_{\mathfrak{k}} is a limit of such a path gtg_{t}. Thus we have

(3.4) limt0S(gt)=SFT1andlimt0dS(gt)dt=T2(SBT2SFT1).\lim_{t\to 0}S(g_{t})=\frac{S_{F}}{T_{1}^{*}}\qquad\text{and}\qquad\lim_{t\to 0}\frac{dS(g_{t})}{dt}=T_{2}^{*}\bigg{(}\frac{S_{B}}{T_{2}^{*}}-\frac{S_{F}}{T_{1}^{*}}\bigg{)}.

Notice that

SFT1=S(T1gF)α𝔨DandSBT2=S(T2gB)β𝔨D.\frac{S_{F}}{T_{1}^{*}}=S(T_{1}^{*}g_{F})\leq\alpha_{\mathfrak{k}}^{D}\qquad\text{and}\qquad\frac{S_{B}}{T_{2}^{*}}=S(T_{2}^{*}g_{B})\leq\beta_{\mathfrak{k}}^{D}.

We now use these formulas to understand the relationship between the numbers α𝔨D\alpha_{\mathfrak{k}}^{D} corresponding to different strata.

Proposition 3.5.

If Δ𝔨\Delta_{{\mathfrak{k}}^{\prime}} is a stratum adjacent to Δ𝔨\Delta_{\mathfrak{k}} with 𝔨𝔨{\mathfrak{k}}^{\prime}\subset{\mathfrak{k}}, then α𝔨Dα𝔨D\alpha_{{\mathfrak{k}}^{\prime}}^{D}\leq\alpha_{\mathfrak{k}}^{D}.

Proof.

Let KK^{\prime} be the subgroup of GG with Lie algebra 𝔨{\mathfrak{k}}^{\prime}. Thus HKKGH\subset K^{\prime}\subset K\subset G. Given hD(K/H)h\in\mathcal{M}^{D}(K^{\prime}/H) with trhT|𝔨𝔪=1\operatorname{tr}_{h}T_{|{\mathfrak{k}}^{\prime}\cap{\mathfrak{m}}}=1, define a one-parameter family of metrics

ht=11tdiTih+1tQ𝔨𝔨D(K/H),\displaystyle h_{t}=\frac{1}{1-t\sum d_{i}T_{i}}\,h+\frac{1}{t}\,Q_{{\mathfrak{k}}^{\prime\perp}\cap{\mathfrak{k}}}\in\mathcal{M}^{D}(K/H),

where the sum is taken over all ii with 𝔪i𝔨𝔨{\mathfrak{m}}_{i}\subset{\mathfrak{k}}^{\prime\perp}\cap{\mathfrak{k}}. Applying (3.4) to the homogeneous fibration

K/HK/HK/K,K^{\prime}/H\to K/H\to K/K^{\prime},

we conclude that

limt0S(ht)=S(h).\displaystyle\lim_{t\to 0}S(h_{t})=S(h).

This means that, for every hD(K/H)h\in\mathcal{M}^{D}(K^{\prime}/H) with trhT|𝔨𝔪=1\operatorname{tr}_{h}T_{|{\mathfrak{k}}^{\prime}\cap{\mathfrak{m}}}=1, there exists a metric gD(K/H)g\in\mathcal{M}^{D}(K/H) with trgT|𝔨𝔪=1\operatorname{tr}_{g}T_{|{\mathfrak{k}}\cap{\mathfrak{m}}}=1 and scalar curvature arbitrarily close to S(h)S(h). ∎

As we noted in Section 3, it is possible that the supremum in the definition of α𝔨D\alpha_{\mathfrak{k}}^{D} is not attained by a metric in Δ𝔨\Delta_{\mathfrak{k}}. In this case, we have the following result.

Proposition 3.6.

Assume that α𝔨D\alpha_{{\mathfrak{k}}}^{D} is not attained. Then there exists an adjacent stratum Δ𝔨\Delta_{{\mathfrak{k}}^{\prime}} such that α𝔨D=α𝔨D\alpha_{{\mathfrak{k}}^{\prime}}^{D}=\alpha_{\mathfrak{k}}^{D} and S(h)=α𝔨DS(h)=\alpha_{{\mathfrak{k}}^{\prime}}^{D} for some hD(K/H)h\in\mathcal{M}^{D}(K^{\prime}/H) with trhT|𝔨𝔪=1\operatorname{tr}_{h}T_{|{\mathfrak{k}}^{\prime}\cap{\mathfrak{m}}}=1.

Proof.

Since α𝔨D\alpha_{{\mathfrak{k}}}^{D} is not attained, it is possible to find a sequence hi(K/H)h_{i}\in\mathcal{M}(K/H) with limiS(hi)=α𝔨D\lim_{i\to\infty}S(h_{i})=\alpha_{\mathfrak{k}}^{D} converging to some hΔJh\in\Delta_{J^{\prime}}, where the stratum ΔJ\Delta_{J^{\prime}} is adjacent to (ΔJ,𝔨)(\Delta_{J},{\mathfrak{k}}). Applying Proposition 2.3 to the homogeneous space K/HK/H and using the nonnegativity of α𝔨D\alpha_{\mathfrak{k}}^{D}, we conclude that ΔJ=Δ𝔨\Delta_{J^{\prime}}=\Delta_{{\mathfrak{k}}^{\prime}} for some 𝔨𝔨{\mathfrak{k}}^{\prime}\subset{\mathfrak{k}}. Similarly, applying Proposition 2.5 to K/HK/H shows that

α𝔨D=limiS(hi)α𝔨D.\displaystyle\alpha_{\mathfrak{k}}^{D}=\lim_{i\to\infty}S(h_{i})\leq\alpha_{{\mathfrak{k}}^{\prime}}^{D}.

In light of Proposition 3.5, this means α𝔨D=α𝔨D\alpha_{{\mathfrak{k}}^{\prime}}^{D}=\alpha_{{\mathfrak{k}}}^{D}. If the supremum α𝔨D\alpha_{{\mathfrak{k}}^{\prime}}^{D} is attained, then we are done. Otherwise, we repeat the argument until we reach a subalgebra 𝔨′′{\mathfrak{k}}^{\prime\prime} for which α𝔨′′D\alpha_{{\mathfrak{k}}^{\prime\prime}}^{D} is achieved. By Corollary 2.4, this will be the case at the latest for a subalgebra 𝔨′′{\mathfrak{k}}^{\prime\prime} in which 𝔥{\mathfrak{h}} is maximal. ∎

Finally, we show how the difference β𝔨Dα𝔨D\beta_{\mathfrak{k}}^{D}-\alpha_{\mathfrak{k}}^{D} controls the behavior of the scalar curvature functional.

Proposition 3.7.

Consider a subalgebra stratum Δ𝔨\Delta_{\mathfrak{k}} such that α𝔨D\alpha_{{\mathfrak{k}}}^{D} is attained. If β𝔨Dα𝔨D>0\beta_{\mathfrak{k}}^{D}-\alpha_{\mathfrak{k}}^{D}>0, then there exists a metric gΔg\in\Delta, arbitrarily close to Δ𝔨\Delta_{\mathfrak{k}}, with S(g)>α𝔨DS(g)>\alpha_{\mathfrak{k}}^{D}.

Proof.

Choose g¯FD(K/H)\bar{g}_{F}\in\mathcal{M}^{D}(K/H) such that trg¯FT|𝔨𝔪=1\operatorname{tr}_{\bar{g}_{F}}T_{|{\mathfrak{k}}\cap{\mathfrak{m}}}=1 and S(g¯F)=α𝔨DS(\bar{g}_{F})=\alpha_{\mathfrak{k}}^{D}. If β𝔨Dα𝔨D>0\beta_{\mathfrak{k}}^{D}-\alpha_{\mathfrak{k}}^{D}>0, it is possible to find g¯BD(G/K)\bar{g}_{B}\in\mathcal{M}^{D}(G/K) with trg¯BT|𝔨=1\operatorname{tr}_{\bar{g}_{B}}T_{|{\mathfrak{k}}^{\perp}}=1 and S(g¯B)α𝔨D>0S(\bar{g}_{B})-\alpha_{\mathfrak{k}}^{D}>0. Consider the metric gTD(G/H)g\in\mathcal{M}_{T}^{D}(G/H) given by

g=2(g¯F+g¯B).\displaystyle g=2(\bar{g}_{F}+\bar{g}_{B}).

If we let gtg_{t} be the canonical variation as in (3.2), we conclude from (3.4) that

limt0S(gt)=S(g¯F)=α𝔨Dandlimt0dS(gt)dt=12(S(g¯B)α𝔨D)>0.\displaystyle\lim_{t\to 0}S(g_{t})=S(\bar{g}_{F})=\alpha_{\mathfrak{k}}^{D}\qquad\text{and}\qquad\lim_{t\to 0}\frac{dS(g_{t})}{dt}=\tfrac{1}{2}(S(\bar{g}_{B})-\alpha_{\mathfrak{k}}^{D})>0.

Clearly, S(gt)>α𝔨S(g_{t})>\alpha_{\mathfrak{k}} for small tt. ∎

Combining Propositions 3.6 and 3.7 implies our main theorem if there exists only one decomposition DD (up to order of summands).

Proposition 3.8.

Assume that G/HG/H is a compact homogeneous space such that the modules 𝔪i{\mathfrak{m}}_{i} are inequivalent. Let 𝔨{\mathfrak{k}} be an intermediate subalgebra of the lowest possible dimension such that α𝔨D=supαD\alpha_{{\mathfrak{k}}}^{D}=\sup_{\ell}\alpha_{\ell}^{D}, where the supremum is taken over all intermediate subalgebras \ell. If β𝔨Dα𝔨D>0\beta_{\mathfrak{k}}^{D}-\alpha_{{\mathfrak{k}}}^{D}>0, then S|TS_{|\mathcal{M}_{T}} achieves its maximum at some metric gTg\in\mathcal{M}_{T}, and hence Ric(g)=cT\operatorname{Ric}(g)=cT for some c>0c>0.

It is natural to add an additional marking to the strata of Δ\partial\Delta by labeling a subalgebra stratum (Δ𝔨,α𝔨,β𝔨)(\Delta_{\mathfrak{k}},\alpha_{\mathfrak{k}},\beta_{\mathfrak{k}}), which encodes the behavior of SS in a neighborhood of Δ𝔨\Delta_{\mathfrak{k}}.

As remarked before, if 𝔨{\mathfrak{k}} is an intermediate subalgebra, then the connected subgroup KK with Lie algebra 𝔨{\mathfrak{k}} may not be compact. Thus, G/KG/K is not necessarily a manifold, and hence (3.1) is not an actual Riemannian submersion. Nevertheless, (3.3) still holds since, by homogeneity, this is a local formula and (3.1) is still a Riemannian submersion locally. Thus, β𝔨D\beta_{{\mathfrak{k}}}^{D} is well-defined for any intermediate subalgebra.


4. Global maxima

From now on, we allow the decomposition DD to vary. Clearly, the numbers bib_{i} and the structure constants [ijk][ijk] depend continuously on DD. Recall also that the space 𝒟\mathcal{D} of all decompositions is compact.

Consider an intermediate subgroup KK with Lie algebra 𝔨{\mathfrak{k}}. The numbers α𝔨D\alpha_{\mathfrak{k}}^{D} and β𝔨D\beta_{\mathfrak{k}}^{D} introduced above depend on the choice of the decomposition DD. Removing this dependence, we define

α𝔨\displaystyle\alpha_{\mathfrak{k}} =sup{S(g)g(K/H) with trg(T|𝔨𝔪)=1}and\displaystyle=\sup\{S(g)\mid g\in\mathcal{M}(K/H)\text{ with }\operatorname{tr}_{g}(T_{|{\mathfrak{k}}\cap{\mathfrak{m}}})=1\}\qquad\mbox{and}
β𝔨\displaystyle\beta_{\mathfrak{k}} =sup{S(g)g(G/K) with trg(T|𝔨)=1},\displaystyle=\sup\{S(g)\mid g\in\mathcal{M}(G/K)\text{ with }\operatorname{tr}_{g}(T_{|{\mathfrak{k}}^{\perp}})=1\},

and introduce an invariant for G/HG/H given by

αG/H=sup𝔨α𝔨,\alpha_{G/H}=\textstyle\sup_{{\mathfrak{k}}}\alpha_{\mathfrak{k}},

where the supremum is taken over all intermediate subalgebras 𝔨{\mathfrak{k}}.

Our first goal is to extend Propositions 2.3 and 2.5 to all of T\mathcal{M}_{T}. We will use the following parametrization of the space \mathcal{M}, convenient in our context. Specifically, consider the map σ:+r×𝒟{\sigma:\mathbb{R}_{+}^{r}\times\mathcal{D}\to\mathcal{M}} defined by

σ(y,D)=i1yiQ|𝔪i,\sigma(y,D)=\sum_{i}\tfrac{1}{y_{i}}Q_{|\mathfrak{m}_{i}},

where y=(y1,,yr)y=(y_{1},\ldots,y_{r}) and DD is the decomposition with irreducible modules 𝔪1,,𝔪r\mathfrak{m}_{1},\ldots,\mathfrak{m}_{r}. This map is clearly continuous. While it is surjective, it may not be injective. The preimages of some metrics are infinite when some of the isotropy summands of G/HG/H are equivalent. However, given a metric gg\in\mathcal{M}, the preimage σ1(g)\sigma^{-1}(g) is compact.

To state our next result, we fix a point (y,D)(y,D) in the boundary of σ1(T)\sigma^{-1}(\mathcal{M}_{T}). Let Δ\Delta be the simplex associated with the decomposition DD. Clearly, yy lies in a stratum ΔJ\Delta_{J} for some JIJ\subset I. The following result generalises Proposition 2.5 to all of T\mathcal{M}_{T}.

Proposition 4.1.

Assume that yy lies in a subalgebra stratum (ΔJ,𝔨)(\Delta_{J},{\mathfrak{k}}). Then for every ϵ>0\epsilon>0 there exists an open neighborhood UU of (y,D)(y,D) in r×𝒟{\mathbb{R}}^{r}\times\mathcal{D} such that

S(σ(y,D))α𝔨+ϵ\displaystyle S(\sigma(y^{\prime},D^{\prime}))\leq\alpha_{\mathfrak{k}}+\epsilon

whenever (y,D)Uσ1(T)(y^{\prime},D^{\prime})\in U\cap\sigma^{-1}(\mathcal{M}_{T}).

Proof.

We use the notation bib_{i}, TiT_{i} and [ijk][ijk] (respectively, bib_{i}^{\prime}, TiT_{i}^{\prime} and [ijk][ijk]^{\prime}) for the constants associated with the decomposition DD (respectively, DD^{\prime}). Given δ>0\delta>0, there exists a neighborhood UδU_{\delta} of (y,D)(y,D) in r×𝒟{\mathbb{R}}^{r}\times\mathcal{D} such that

|yiyi|+|bibi|+|TiTi|+|[ijk][ijk]|<δ\displaystyle|y_{i}-y_{i}^{\prime}|+|b_{i}-b_{i}^{\prime}|+|T_{i}-T_{i}^{\prime}|+|[ijk]-[ijk]^{\prime}|<\delta

for all i,j,ki,j,k whenever (y,D)Uδσ1(T)(y^{\prime},D^{\prime})\in U_{\delta}\cap\sigma^{-1}(\mathcal{M}_{T}). Let us choose δ\delta small enough to ensure that yi>yi2y_{i}^{\prime}>\frac{y_{i}}{2} in this set. The trace constraint implies that yi<1diTiy_{i}^{\prime}<\frac{1}{d_{i}T_{i}^{\prime}}. Notice also that there exist common lower and upper bounds for TiT_{i} independent of DD.

If (y,D)Uδσ1(T)(y^{\prime},D^{\prime})\in U_{\delta}\cap\sigma^{-1}(\mathcal{M}_{T}), we find, as in the proof of Proposition 2.5, that

S(σ(y,D))\displaystyle S(\sigma(y^{\prime},D^{\prime})) =12idibiyi14i,j,kyiyjyk[ijk]\displaystyle=\frac{1}{2}\sum_{i}d_{i}b_{i}^{\prime}y_{i}^{\prime}-\frac{1}{4}\sum_{i,j,k}\frac{y_{i}^{\prime}y_{j}^{\prime}}{y_{k}^{\prime}}[ijk]^{\prime}
12iJ(dibij,kJc[ijk])yi14i,j,kJyiyjyk[ijk]+12iJcdibiyi\displaystyle\leq\frac{1}{2}\sum_{i\in J}\bigg{(}d_{i}b_{i}^{\prime}-\sum_{j,k\in J^{c}}[ijk]^{\prime}\bigg{)}y_{i}^{\prime}-\frac{1}{4}\sum_{i,j,k\in J}\frac{y_{i}^{\prime}y_{j}^{\prime}}{y_{k}^{\prime}}[ijk]^{\prime}+\frac{1}{2}\sum_{i\in J^{c}}d_{i}b_{i}^{\prime}y_{i}^{\prime}
12iJdib¯iyi14i,j,kJyiyjyk[ijk]+12iJdi(bibi)yi\displaystyle\leq\frac{1}{2}\sum_{i\in J}d_{i}\bar{b}_{i}y_{i}^{\prime}-\frac{1}{4}\sum_{i,j,k\in J}\frac{y_{i}^{\prime}y_{j}^{\prime}}{y_{k}^{\prime}}[ijk]+\frac{1}{2}\sum_{i\in J}d_{i}(b_{i}^{\prime}-b_{i})y_{i}^{\prime}
12iJj,kJc([ijk][ijk])yi14i,j,kJyiyjyk([ijk][ijk])+12iJcdibiyi.\displaystyle\hphantom{=}~{}-\frac{1}{2}\sum_{i\in J}\sum_{j,k\in J^{c}}([ijk]^{\prime}-[ijk])y_{i}^{\prime}-\frac{1}{4}\sum_{i,j,k\in J}\frac{y_{i}^{\prime}y_{j}^{\prime}}{y_{k}^{\prime}}([ijk]^{\prime}-[ijk])+\frac{1}{2}\sum_{i\in J^{c}}d_{i}b_{i}^{\prime}y_{i}^{\prime}.

Consequently, for small enough δ\delta, since yi=0y_{i}=0 when iJci\in J^{c}, we have

S(σ(y,D))\displaystyle S(\sigma(y^{\prime},D^{\prime})) S(σ(y,D)|𝔨𝔪)+12iJ|bibi|Ti\displaystyle\leq S(\sigma(y^{\prime},D)_{|{\mathfrak{k}}\cap{\mathfrak{m}}})+\frac{1}{2}\sum_{i\in J}\frac{|b_{i}^{\prime}-b_{i}|}{T_{i}^{\prime}}
+12iJj,kJc|[ijk][ijk]|diTi+12i,j,kJ|[ijk][ijk]|didjTiTjyk+12iJcdibi|yiyi|\displaystyle\hphantom{=}~{}+\frac{1}{2}\sum_{i\in J}\sum_{j,k\in J^{c}}\frac{|[ijk]^{\prime}-[ijk]|}{d_{i}T_{i}^{\prime}}+\frac{1}{2}\sum_{i,j,k\in J}\frac{|[ijk]^{\prime}-[ijk]|}{d_{i}d_{j}T_{i}^{\prime}T_{j}^{\prime}y_{k}}+\frac{1}{2}\sum_{i\in J^{c}}d_{i}b_{i}^{\prime}|y_{i}^{\prime}-y_{i}|
S(σ(y,D)|𝔨𝔪)+ϵ2.\displaystyle\leq S(\sigma(y^{\prime},D)_{|{\mathfrak{k}}\cap{\mathfrak{m}}})+\frac{\epsilon}{2}.

Shrinking δ\delta further if necessary and using the continuity of the scalar curvature, we conclude that

S(σ(y,D)|𝔨𝔪)<S(σ(y,D))+ϵ2α𝔨+ϵ2,S(\sigma(y^{\prime},D)_{|{\mathfrak{k}}\cap{\mathfrak{m}}})<S(\sigma(y,D))+\frac{\epsilon}{2}\leq\alpha_{\mathfrak{k}}+\frac{\epsilon}{2},

which implies the result. ∎

Using a similar (but simpler) proof, we can generalize Proposition 2.3.

Proposition 4.2.

Assume that (y,D)(y,D) lies in a stratum (ΔJ,)(\Delta_{J},\infty). Given a>0a>0, there exists a neighborhood UU of (y,D)(y,D) in r×𝒟{\mathbb{R}}^{r}\times\mathcal{D} such that S(σ(y,D))<aS(\sigma(y^{\prime},D^{\prime}))<-a whenever (y,D)Uσ1(T)(y^{\prime},D^{\prime})\in U\cap\sigma^{-1}(\mathcal{M}_{T}).

The precompactness of σ1(T)r×𝒟\sigma^{-1}(\mathcal{M}_{T})\subset{\mathbb{R}}^{r}\times\mathcal{D} and Propositions 4.1 and 4.2 yield the following extension of Corollary 2.7 to all of T\mathcal{M}_{T}.

Corollary 4.3.

If a>αG/Ha>\alpha_{G/H}, then {gTS(g)a}\{g\in\mathcal{M}_{T}\mid S(g)\geq a\} is a compact subset of T\mathcal{M}_{T}.

Our next result generalises Proposition 3.6.

Proposition 4.4.

Assume that 𝔥{\mathfrak{h}} is not maximal in 𝔤{\mathfrak{g}}. Then there exists an intermediate subgroup KK with Lie algebra 𝔨{\mathfrak{k}} such that α𝔨=αG/H\alpha_{{\mathfrak{k}}}=\alpha_{G/H}. If KK has the least possible dimension of all such subgroups, then there exists h(K/H)h\in\mathcal{M}(K/H) with

S(h)=α𝔨andtrhT|𝔨𝔪=1.\displaystyle S(h)=\alpha_{{\mathfrak{k}}}\qquad\text{and}\qquad\operatorname{tr}_{h}T_{|{\mathfrak{k}}\cap{\mathfrak{m}}}=1.
Proof.

For each ii\in{\mathbb{N}}, choose a subgroup KiGK_{i}\subset G with Lie algebra 𝔨i{\mathfrak{k}}_{i} such that α𝔨iαG/H12i\alpha_{{\mathfrak{k}}_{i}}\geq\alpha_{G/H}-\frac{1}{2i}. Assume that KiK_{i} has the least possible dimension among all subgroups with this property. Let hih_{i} be a metric on Ki/HK_{i}/H with

S(hi)>αG/H1iandtrhiT|𝔨i𝔪=1.S(h_{i})>\alpha_{G/H}-\tfrac{1}{i}\qquad\text{and}\qquad\operatorname{tr}_{h_{i}}T_{|{\mathfrak{k}}_{i}\cap{\mathfrak{m}}}=1.

We may assume that 𝔨i{\mathfrak{k}}_{i} converge to an intermediate subalgebra 𝔨{\mathfrak{k}} and that dim𝔨i=dim𝔨\dim{\mathfrak{k}}_{i}=\dim{\mathfrak{k}} for all ii. Our goal is to show that a subsequence of (hi)(h_{i}) converges to a metric h(K/H)h\in\mathcal{M}(K/H).

Consider a decomposition Di𝒟D_{i}\in\mathcal{D} with modules 𝔪1i,,𝔪ri\mathfrak{m}_{1}^{i},\ldots,\mathfrak{m}_{r}^{i} such that

𝔨i=𝔥jJi𝔪jiandhi(𝔪ki,𝔪li)=0{\mathfrak{k}}_{i}={\mathfrak{h}}\oplus\bigoplus_{j\in J_{i}}\mathfrak{m}_{j}^{i}\qquad\text{and}\qquad h_{i}(\mathfrak{m}_{k}^{i},\mathfrak{m}_{l}^{i})=0

for some JiIJ_{i}\subset I and all k,lJik,l\in J_{i} with klk\neq l. Passing to a subsequence if necessary, we may assume that these decompositions converge to some D𝒟D\in\mathcal{D} with modules 𝔪1,,𝔪r\mathfrak{m}_{1},\ldots,\mathfrak{m}_{r} and that JiJ_{i} does not depend on ii (thus we can omit the index ii from the notation JiJ_{i}). Let KK be the connected subgroup of GG with Lie algebra 𝔨=𝔥𝔪J{\mathfrak{k}}={\mathfrak{h}}\oplus\mathfrak{m}_{J}. There exist positive numbers xjix_{ji} such that

hi=jJxjiQ|𝔪ji.\displaystyle h_{i}=\sum_{j\in J}x_{ji}Q_{|\mathfrak{m}_{j}^{i}}.

Note that αG/H>supα\alpha_{G/H}>\sup_{\ell}\alpha_{\ell}, where the supremum is taken over all intermediate subalgebras \ell between 𝔥{\mathfrak{h}} and 𝔨{\mathfrak{k}}. Indeed, if not, there exists a subalgebra \ell with α>αG/H12i\alpha_{\ell}>\alpha_{G/H}-\frac{1}{2i}, contradicting the assumption that KiK_{i} is chosen to be of smallest possible dimension. Therefore,

S(hi)>αG/H1i>supαS(h_{i})>\alpha_{G/H}-\tfrac{1}{i}>\sup\nolimits_{\ell}\alpha_{\ell}

for large ii. We now claim that the constants xjix_{ji} all lie in some compact subset of +{\mathbb{R}}_{+}. To see this, we can argue as in Propositions 4.1 and 4.2 and Corollary 4.3 replacing G/HG/H with the sequence (Ki/H)(K_{i}/H) and using the fact that the structure constants of 𝔨i{\mathfrak{k}}_{i} converge to those of 𝔨{\mathfrak{k}}.

Thus, passing to a subsequence if necessary, we may assume that

limixji=xi+\lim_{i\to\infty}x_{ji}=x_{i}\in{\mathbb{R}}_{+}

for jJj\in J. The metric h=iJxiQ|𝔪ih=\sum_{i\in J}x_{i}Q_{|\mathfrak{m}_{i}} satisfies S(h)=α𝔨=αG/HS(h)=\alpha_{{\mathfrak{k}}}=\alpha_{G/H} and trhT|𝔨𝔪=1\operatorname{tr}_{h}T_{|{\mathfrak{k}}\cap{\mathfrak{m}}}=1. ∎

We are now ready to prove our main theorem.

Proof of the main theorem.

The existence of the subgroup KK follows from Proposition 4.4. By Proposition 3.7, there exists ϵ>0\epsilon>0 such that the superlevel set {gTS(g)α𝔨+ϵ}\{g\in\mathcal{M}_{T}\mid S(g)\geq\alpha_{{\mathfrak{k}}}+\epsilon\} is nonempty. Corollary 4.3 implies that this set is also compact. Consequently, S|TS_{|\mathcal{M}_{T}} assumes its maximum at a metric gTg\in\mathcal{M}_{T}. As a critical point, such a metric satisfies Ric(g)=cT\operatorname{Ric}(g)=cT for some constant c>0c>0. ∎


5. Examples

5.1. The Stiefel manifold V𝟐(𝟒)\mathbf{\emph{V}_{2}({\mathbb{R}}^{4})}.

This example is interesting since some of the modules 𝔪i{\mathfrak{m}}_{i} are equivalent and hence we need to consider non-diagonal metrics as well. Furthermore, the set of decompositions is not discrete; in fact, as we will see below, 𝒟\mathcal{D} is two-dimensional. In addition, there exists a circle of intermediate subgroups and hence we need to maximize α𝔨\alpha_{\mathfrak{k}}. This is the first example of this type where the existence of global maxima was studied.

Consider the homogeneous space G/H=(SU(2)×SU(2))/S1G/H=(SU(2)\times SU(2))/S^{1} where the circle group is embedded diagonally into SU(2)×SU(2)SU(2)\times SU(2). The bi-invariant metric QQ we choose on GG is such that |(X,Y)|Q=12(tr(X2)+tr(Y2))|(X,Y)|_{Q}=-\tfrac{1}{2}(\operatorname{tr}(X^{2})+\operatorname{tr}(Y^{2})). Since the two-fold cover SU(2)×SU(2)SO(4)SU(2)\times SU(2)\to SO(4) sends the diagonal embedding diag(SU(2))SU(2)×SU(2)\operatorname{diag}(SU(2))\subset SU(2)\times SU(2) to SO(3)SO(4)SO(3)\subset SO(4), the space G/HG/H coincides with the Stiefel manifold V2(4)=SO(4)/SO(2)V_{2}(\mathbb{R}^{4})=SO(4)/SO(2).

We can identify SU(2)SU(2) and 𝔰𝔲(2){\mathfrak{su}}(2) with the group of unit quaternions and the Lie algebra of purely imaginary quaternions. Then

H={(eη𝐢,eη𝐢)|η[0,2π)}and𝔥=span{(𝐢,𝐢)}.H=\{(e^{\eta\bf i},e^{\eta\bf i})\,|\,\eta\in[0,2\pi)\}\qquad\mbox{and}\qquad{\mathfrak{h}}=\operatorname{span}\{(\,\mathbf{i},\,\mathbf{i})\}.

Consider the following AdH\operatorname{Ad}_{H}-invariant decomposition of 𝔪{\mathfrak{m}}:

(5.1) 𝔪0=span{(𝐢,𝐢)},𝔪1={(z𝐣,0)z}and𝔪2={(0,z𝐣)z}.\displaystyle{\mathfrak{m}}_{0}=\operatorname{span}\{(\,\mathbf{i},-\mathbf{i})\},\qquad{\mathfrak{m}}_{1}=\{(z\,\mathbf{j},0)\mid z\in{\mathbb{C}}\}\qquad\mbox{and}\qquad{\mathfrak{m}}_{2}=\{(0,z\,\mathbf{j})\mid z\in{\mathbb{C}}\}.

Under AdH\operatorname{Ad}_{H}, the element (eη𝐢,eη𝐢)H(e^{\eta\bf i},e^{\eta\bf i})\in H takes (z𝐣,0)(z{\bf j},0) and (0,z𝐣)(0,z\bf j) to (e2η𝐢z𝐣,0)(e^{2\eta\bf i}z{\bf j},0) and (0,e2η𝐢z𝐣)(0,e^{2\eta\bf i}z\bf j), respectively. This implies that the restrictions of AdH\operatorname{Ad}_{H} to 𝔪1{\mathfrak{m}}_{1} and 𝔪2{\mathfrak{m}}_{2} are equivalent complex representations. Using the QQ-orthonormal bases {(𝐢,𝐢)2}\big{\{}\frac{({\bf i},-{\bf i})}{\sqrt{2}}\big{\}} of 𝔪0{\mathfrak{m}}_{0}, {(𝐣,0),(𝐤,0)}\{({\bf j},0),({\bf k},0)\} of 𝔪1{\mathfrak{m}}_{1}, and {(0,𝐣),(0,𝐤)}\{(0,{\bf j}),(0,{\bf k})\} of 𝔪2{\mathfrak{m}}_{2}, we can thus represent the metric gg\in\mathcal{M} and the tensor field TT by the matrices

(5.2) g=(x000000x10x3x400x1x4x30x3x4x200x4x30x2)andT=(T000000T10T3T400T1T4T30T3T4T200T4T30T2),g=\left(\begin{matrix}x_{0}&0&0&0&0\\ 0&x_{1}&0&x_{3}&x_{4}\\ 0&0&x_{1}&-x_{4}&x_{3}\\ 0&x_{3}&-x_{4}&x_{2}&0\\ 0&x_{4}&x_{3}&0&x_{2}\end{matrix}\right)\qquad\mbox{and}\qquad T=\left(\begin{matrix}T_{0}&0&0&0&0\\ 0&T_{1}&0&T_{3}&T_{4}\\ 0&0&T_{1}&-T_{4}&T_{3}\\ 0&T_{3}&-T_{4}&T_{2}&0\\ 0&T_{4}&T_{3}&0&T_{2}\end{matrix}\right),

which must be positive-definite. This is the case if x0,x1,x2>0x_{0},x_{1},x_{2}>0 and x1x2x32x42>0x_{1}x_{2}-x_{3}^{2}-x_{4}^{2}>0 and similarly for TT. The outer automorphism of G=SU(2)×SU(2)G=SU(2)\times SU(2) that switches the two factors preserves HH and hence induces an isometry of (V2(4),Q)(V_{2}({\mathbb{R}}^{4}),Q). It acts on gg by taking (x1,x2,x3,x4)(x_{1},x_{2},x_{3},x_{4}) to (x2,x1,x3,x4)(x_{2},x_{1},x_{3},-x_{4}), and similarly on TT. Conjugation by (𝐣,𝐣)(\bf{j},\bf{j}) preserves HH as well. It acts on gg by taking (x1,x2,x3,x4)(x_{1},x_{2},x_{3},x_{4}) to (x1,x2,x3,x4)(x_{1},x_{2},x_{3},-x_{4}). Thus the composition takes (x1,x2,x3,x4)(x_{1},x_{2},x_{3},x_{4}) to (x2,x1,x3,x4)(x_{2},x_{1},x_{3},x_{4}), which shows that we may assume T1T2T_{1}\geq T_{2}. It will be convenient for us to denote

γ(T0,T2,T3,T4)=T0+T02+16T0T2+16T32+T428.\displaystyle\gamma(T_{0},T_{2},T_{3},T_{4})=\frac{T_{0}+\sqrt{T_{0}^{2}+16T_{0}T_{2}}+16\sqrt{T_{3}^{2}+T_{4}^{2}}}{8}.

As we will see shortly, the maximal subgroups of GG containing HH are the subgroups K1=SU(2)×S1K_{1}=SU(2)\times S^{1} and K2=S1×SU(2)K_{2}=S^{1}\times SU(2) and a one-parameter family of three-dimensional subgroups KθK^{\theta}, θ[0,2π)\theta\in[0,2\pi), isomorphic to SU(2)SU(2). Our main theorem leads to the following result. We distinguish the case where the supremum αG/H\alpha_{G/H} is attained by one of the four-dimensional subgroups from the case where it is attained by a three-dimensional subgroup.

Proposition 5.3.

Suppose that G/H=V2(4)G/H=V_{2}({\mathbb{R}}^{4}) and T1T2T_{1}\geq T_{2}. Then we have:

  • (a)

    αG/H=α𝔨2\alpha_{G/H}=\alpha_{{\mathfrak{k}}_{2}} and β𝔨2α𝔨2>0\beta_{{\mathfrak{k}}_{2}}-\alpha_{{\mathfrak{k}}_{2}}>0 if and only if

    γ(T0,T2,T3,T4)T1<T2+T0+T02+16T0T28.\gamma(T_{0},T_{2},T_{3},T_{4})\leq T_{1}<T_{2}+\frac{T_{0}+\sqrt{T_{0}^{2}+16T_{0}T_{2}}}{8}.
  • (b)

    αG/H=α𝔨θ\alpha_{G/H}=\alpha_{{\mathfrak{k}}^{\theta}} and β𝔨θα𝔨θ>0\beta_{{\mathfrak{k}}^{\theta}}-\alpha_{{\mathfrak{k}}^{\theta}}>0 for some θ[0,2π)\theta\in[0,2\pi) if and only if

    T02T2+4T32+T42<T1γ(T0,T2,T3,T4).\frac{T_{0}}{2}-T_{2}+4\sqrt{T_{3}^{2}+T_{4}^{2}}<T_{1}\leq\gamma(T_{0},T_{2},T_{3},T_{4}).

In both cases, S|TS_{|\mathcal{M}_{T}} attains its global maximum.

Proof.

We start by describing the space of all decompositions. Since the representations of HH on 𝔪1{\mathfrak{m}}_{1} and 𝔪2{\mathfrak{m}}_{2} are complex equivalent representations, 𝒟\mathcal{D} is the two-parameter family of decompositions Ds,θD_{s,\theta} with modules

(5.4) 𝔪0=span{(𝐢,𝐢)},𝔪1s,θ={(seθ𝐢z𝐣,tz𝐣)z}and𝔪2s,θ={(teθ𝐢z𝐣,sz𝐣)z},\displaystyle{\mathfrak{m}}_{0}=\operatorname{span}\{(\,\mathbf{i},-\,\mathbf{i})\},\quad{\mathfrak{m}}_{1}^{s,\theta}=\{(se^{\theta\mathbf{i}}z\mathbf{j},-tz\,\mathbf{j})\mid z\in{\mathbb{C}}\}\quad\mbox{and}\quad{\mathfrak{m}}_{2}^{s,\theta}=\{(te^{\theta\mathbf{i}}z\,\mathbf{j},sz\,\mathbf{j})\mid z\in{\mathbb{C}}\},

where s2+t2=1s^{2}+t^{2}=1. Let us choose a QQ-orthonormal basis in 𝔪1s,θ{\mathfrak{m}}_{1}^{s,\theta} consisting of the vectors

(5.5) v1=(scosθ𝐣+ssinθ𝐤,t𝐣)andv2=(ssinθ𝐣+scosθ𝐤,t𝐤)\displaystyle v_{1}=(s\cos\theta\,\mathbf{j}+s\sin\theta\,\mathbf{k},-t\,\mathbf{j})\qquad\mbox{and}\qquad v_{2}=(-s\sin\theta\,\mathbf{j}+s\cos\theta\,\mathbf{k},-t\,\mathbf{k})

and one in 𝔪2s,θ{\mathfrak{m}}_{2}^{s,\theta} consisting of

(5.6) w1=(tcosθ𝐣+tsinθ𝐤,s𝐣)andw2=(tsinθ𝐣+tcosθ𝐤,s𝐤).\displaystyle w_{1}=(t\cos\theta\,\mathbf{j}+t\sin\theta\,\mathbf{k},s\mathbf{j})\qquad\mbox{and}\qquad w_{2}=(-t\sin\theta\,\mathbf{j}+t\cos\theta\,\mathbf{k},s\mathbf{k}).

Using these bases, for the decomposition Ds,θD_{s,\theta}, one easily computes

(5.7) [011]=[022]=4(s2t2)2and[012]=8s2t2.[011]=[022]=4(s^{2}-t^{2})^{2}\qquad\mbox{and}\qquad[012]=8s^{2}t^{2}.

The structure constants unrelated to these by permutation are 0.

The decomposition corresponding to s=1,t=0s=1,t=0 gives rise to the natural set of diagonal metrics x3=x4=0x_{3}=x_{4}=0 with intermediate subgroups

K0=T2,K1=SU(2)×Srt1andK2=Slt1×SU(2)K_{0}=T^{2},\qquad K_{1}=SU(2)\times S^{1}_{\textrm{rt}}\qquad\mbox{and}\qquad K_{2}=S^{1}_{\textrm{lt}}\times SU(2)

with Lie algebras

𝔨0=𝔥𝔪0,𝔨1=𝔥𝔪0𝔪1and𝔨2=𝔥𝔪0𝔪2.\displaystyle{\mathfrak{k}}_{0}={\mathfrak{h}}\oplus{\mathfrak{m}}_{0},\qquad{\mathfrak{k}}_{1}={\mathfrak{h}}\oplus{\mathfrak{m}}_{0}\oplus{\mathfrak{m}}_{1}\qquad\mbox{and}\qquad{\mathfrak{k}}_{2}={\mathfrak{h}}\oplus{\mathfrak{m}}_{0}\oplus{\mathfrak{m}}_{2}.

If s=t=22s=t=\frac{\sqrt{2}}{2}, we have, for each θ\theta, the intermediate subgroup

KθSU(2)with Lie algebra𝔨θ=𝔥𝔪1s,θ.K^{\theta}\simeq SU(2)\qquad\text{with Lie algebra}\qquad{\mathfrak{k}}^{\theta}={\mathfrak{h}}\oplus{\mathfrak{m}}_{1}^{s,\theta}.

The remaining decompositions do not produce any subgroups. Thus, there are three isolated intermediate subgroups, K0K_{0}, K1K_{1} and K2K_{2}, as well as a one-parameter family of subgroups KθK^{\theta}.

The identity component of the normalizer of HH is given by N0(H)=T2SU(2)×SU(2),N_{0}(H)=T^{2}\subset SU(2)\times SU(2), and hence N0(H)/HS1N_{0}(H)/H\simeq S^{1}, represented by elements of the form (eη𝐢,1)N0(H)(e^{\eta\bf i},1)\in N_{0}(H). These elements act via right translation on G/HG/H and via conjugation on 𝔪{\mathfrak{m}}. Thus they also act on 𝒟\mathcal{D} and, via pullback, on \mathcal{M}. It is easy to see that (eη𝐢,1)(e^{\eta\bf i},1) takes Ds,θD_{s,\theta} to Ds,θ+2ηD_{s,\theta+2\eta}. This implies, in particular, that the subalgebras 𝔨θ{\mathfrak{k}}^{\theta} are all conjugate to each other by elements of N0(H)N_{0}(H). Since N0(H)N_{0}(H) acts on G/HG/H by isometries in QQ, it follows that (G/Kθ,Q)(G/K^{\theta},Q), as well as (Kθ/H,Q)(K^{\theta}/H,Q), are all isometric to each other.

We now compute the constants α𝔨\alpha_{\mathfrak{k}} and β𝔨\beta_{\mathfrak{k}} for the intermediate subalgebras. For the maximal subgroups K1K_{1} and K2K_{2} we have

α𝔨i=8Ti+T0T02+16T0Ti2Ti2andβ𝔨i=4Tj,\displaystyle\alpha_{{\mathfrak{k}}_{i}}=\frac{8T_{i}+T_{0}-\sqrt{T_{0}^{2}+16T_{0}T_{i}}}{2T_{i}^{2}}\qquad\mbox{and}\qquad\beta_{{\mathfrak{k}}_{i}}=\frac{4}{T_{j}},

where (i,j)(i,j) is a permutation of {1,2}\{1,2\}. Indeed, G/KiG/K_{i} are isotropy irreducible, which easily determines β𝔨i\beta_{{\mathfrak{k}}_{i}}. The space Ki/HK_{i}/H has two irreducible summands, and the scalar curvature, under the trace constraint, is

S=8y1y12y0.S=8y_{1}-\frac{y_{1}^{2}}{y_{0}}.

The above value α𝔨i\alpha_{{\mathfrak{k}}_{i}} is its maximum. The assumption T1T2T_{1}\geq T_{2} implies that α𝔨2α𝔨1\alpha_{{\mathfrak{k}}_{2}}\geq\alpha_{{\mathfrak{k}}_{1}}. Furthermore, β𝔨2α𝔨2>0\beta_{{\mathfrak{k}}_{2}}-\alpha_{{\mathfrak{k}}_{2}}>0 if and only if

(5.8) T1<T2+T0+T02+16T0T28.\displaystyle T_{1}<T_{2}+\frac{T_{0}+\sqrt{T_{0}^{2}+16T_{0}T_{2}}}{8}.

For the three-dimensional subalgebras 𝔨θ{\mathfrak{k}}^{\theta}, we use the bases (5.5) and (5.6). Now, the parameters ss and tt both equal 12\frac{1}{\sqrt{2}}. The tensor field TT satisfies

2T(v1,v1)\displaystyle 2T(v_{1},v_{1}) =2T(v2,v2)=T1+T22(cosθT3sinθT4),\displaystyle=2T(v_{2},v_{2})=T_{1}+T_{2}-2(\cos\theta\,T_{3}-\sin\theta\,T_{4}),
2T(w1,w1)\displaystyle 2T(w_{1},w_{1}) =2T(w2,w2)=T1+T2+2(cosθT3sinθT4),\displaystyle=2T(w_{2},w_{2})=T_{1}+T_{2}+2(\cos\theta\,T_{3}-\sin\theta\,T_{4}),
2T(v1,w1)\displaystyle 2T(v_{1},w_{1}) =2T(v2,w2)=T1T2,\displaystyle=2T(v_{2},w_{2})=T_{1}-T_{2},
(5.9) T(v1,w2)\displaystyle T(v_{1},w_{2}) =T(v2,w1)=cosθT4+sinθT3,\displaystyle=-T(v_{2},w_{1})=\cos\theta\,T_{4}+\sin\theta\,T_{3},

and T(v1,v2)=T(w1,w2)=0T(v_{1},v_{2})=T(w_{1},w_{2})=0. The action of the quotient N0(H)/HN_{0}(H)/H establishes an isometry between (Kθ/H,Q)(K^{\theta}/H,Q) and the space (diag(SU(2))/diag(S1),Q)(\operatorname{diag}(SU(2))/\operatorname{diag}(S^{1}),Q), which one easily sees has scalar curvature 44. The trace constraint means that g(v1,v1)=2T(v1,v1)g(v_{1},v_{1})=2T(v_{1},v_{1}), implying

α𝔨θ=4T1+T22(cosθT3sinθT4).\displaystyle\alpha_{{\mathfrak{k}}^{\theta}}=\frac{4}{T_{1}+T_{2}-2(\cos\theta\,T_{3}-\sin\theta\,T_{4})}.

The action of the normalizer shows that (G/Kθ,Q)(G/K^{\theta},Q) are all isometric to the symmetric space ((SU(2)×SU(2))/diag(SU(2)),Q)((SU(2)\times SU(2))/\operatorname{diag}(SU(2)),Q). Thus, they have scalar curvature 12. The trace constraint for KθK^{\theta}-invariant metrics on G/KθG/K^{\theta} takes the form

g((𝐢,𝐢)2,(𝐢,𝐢)2)=g(w1,w1)\displaystyle g\bigg{(}\frac{({\bf i},-{\bf i})}{\sqrt{2}},\frac{(\bf i,-\bf i)}{\sqrt{2}}\bigg{)}=g(w_{1},w_{1}) =g(w2,w2)\displaystyle=g(w_{2},w_{2})
=T((𝐢,𝐢)2,(𝐢,𝐢)2)+T(w1,w1)+T(w2,w2)\displaystyle=T\bigg{(}\frac{({\bf i},-{\bf i})}{\sqrt{2}},\frac{(\bf i,-\bf i)}{\sqrt{2}}\bigg{)}+T(w_{1},w_{1})+T(w_{2},w_{2})
=T0+T1+T2+2(cosθT3sinθT4).\displaystyle=T_{0}+T_{1}+T_{2}+2(\cos\theta\,T_{3}-\sin\theta\,T_{4}).

Thus

β𝔨θ=12T0+T1+T2+2(cosθT3sinθT4).\beta_{{\mathfrak{k}}^{\theta}}=\frac{12}{T_{0}+T_{1}+T_{2}+2(\cos\theta\,T_{3}-\sin\theta\,T_{4})}.

We now choose an angle θ0\theta_{0} such that α𝔨θ0\alpha_{{\mathfrak{k}}^{\theta_{0}}} is maximal, i.e., α𝔨θ0=supθα𝔨θ\alpha_{{\mathfrak{k}}^{\theta_{0}}}=\sup_{\theta}\alpha_{{\mathfrak{k}}^{\theta}}. Observe that

cosθT3sinθT4=T32+T42sin(θη)\cos\theta\,T_{3}-\sin\theta\,T_{4}=\sqrt{T_{3}^{2}+T_{4}^{2}}\,\sin(\theta-\eta)

for some phase shift η\eta. The largest possible value of this quantity is T32+T42\sqrt{T_{3}^{2}+T_{4}^{2}}, which means

α𝔨θ0=4T1+T22T32+T42andβ𝔨θ0=12T0+T1+T2+2T32+T42.\alpha_{{\mathfrak{k}}^{\theta_{0}}}=\frac{4}{T_{1}+T_{2}-2\sqrt{T_{3}^{2}+T_{4}^{2}}}\qquad\mbox{and}\qquad\beta_{{\mathfrak{k}}^{\theta_{0}}}=\frac{12}{T_{0}+T_{1}+T_{2}+2\sqrt{T_{3}^{2}+T_{4}^{2}}}.

Clearly, β𝔨θ0α𝔨θ0>0\beta_{{\mathfrak{k}}^{\theta_{0}}}-\alpha_{{\mathfrak{k}}^{\theta_{0}}}>0 if and only if

(5.10) T0<2T1+2T28T32+T42.T_{0}<2T_{1}+2T_{2}-8\sqrt{T_{3}^{2}+T_{4}^{2}}.

Let us determine αG/H\alpha_{G/H}. As noted above, the assumption T1T2T_{1}\geq T_{2} implies that α𝔨2α𝔨1\alpha_{{\mathfrak{k}}_{2}}\geq\alpha_{{\mathfrak{k}}_{1}}. Elementary analysis shows that α𝔨2α𝔨θ0\alpha_{{\mathfrak{k}}_{2}}\geq\alpha_{{\mathfrak{k}}^{\theta_{0}}} if and only if T1γ(T0,T2,T3,T4)T_{1}\geq\gamma(T_{0},T_{2},T_{3},T_{4}). Combining this condition with (5.8), we arrive at statement (a) of the proposition. Reversing the inequality and taking note of (5.10), we obtain statement (b). ∎

Remark 5.11.

For the set of diagonal metrics x3=x4=0x_{3}=x_{4}=0 one easily solves the equation Ric(g)=cT\operatorname{Ric}(g)=cT directly. This equation reduces to the system

(5.12) x0=(4cT1)x1,x0=(4cT2)x2,cT0\displaystyle x_{0}=(4-cT_{1})x_{1},\qquad x_{0}=(4-cT_{2})x_{2},\qquad cT_{0} =(4cT1)2+(4cT2)2.\displaystyle=(4-cT_{1})^{2}+(4-cT_{2})^{2}.

One easily shows that the solution is unique. In fact, our main theorem, applied to such diagonal metrics, shows that the critical point is always a global maximum. The action of N0(H)/HN_{0}(H)/H fixes the diagonal metrics and acts by isometries. Thus the Ricci curvature of a diagonal metric must be diagonal as well. In particular, if T3=T4=0T_{3}=T_{4}=0, then every solution to the system (5.12) is a critical point of S|TS_{|\mathcal{M}_{T}} on all of T\mathcal{M}_{T}. In Figure 1 these metrics lie in the union of the 3 grey regions. On the other hand, given a critical point of S|TS_{|\mathcal{M}_{T}} with T3=T4=0T_{3}=T_{4}=0 and x30x_{3}\neq 0, one obtains a circle of further critical points by applying the normalizer.

For a metric gg as in (5.2), the constraint trgT=1\operatorname{tr}_{g}T=1 becomes

T0x0+2x1T2+2x2T14x3T34x4T4Λ=1,\frac{T_{0}}{x_{0}}+\frac{2x_{1}T_{2}+2x_{2}T_{1}-4x_{3}T_{3}-4x_{4}T_{4}}{\Lambda}=1,

and the scalar curvature satisfies

S(g)=8x0+8(x1+x2)Λ8x1x2x0Λx0(x1x2)2Λ22x0Λ,S(g)=\frac{8}{x_{0}}+\frac{8(x_{1}+x_{2})}{\Lambda}-\frac{8x_{1}x_{2}}{x_{0}\Lambda}-\frac{x_{0}(x_{1}-x_{2})^{2}}{\Lambda^{2}}-\frac{2x_{0}}{\Lambda},

where Λ=x1x2x32x42\Lambda=x_{1}x_{2}-x_{3}^{2}-x_{4}^{2}; see, e.g., [5]. We may assume that T0=1T_{0}=1. If T32+T42=0T_{3}^{2}+T_{4}^{2}=0, we show in Figure 1 points in the (T1,T2)(T_{1},T_{2})-plane that correspond to various behaviors of S|TS_{|\mathcal{M}_{T}}. In the union of the three grey regions the solutions to (5.12) guarantee the existence of a (diagonal) critical point. In the dark-grey region 𝔨2{\mathfrak{k}}_{2} yields a global maximum, and in the middle-grey region 𝔨θ{{\mathfrak{k}}_{\theta}} . A computer-assisted experiment shows that this global maximum is always a diagonal metric; see Example 3. However, in the light grey region the diagonal critical point might not be a global maximum, as we demonstrate below.

If T32+T420T_{3}^{2}+T_{4}^{2}\neq 0, the picture is the same but with the line 2T1=12T22T_{1}=1-2T_{2} and the curve 8T1=1+1+16T28T_{1}=1+\sqrt{1+16T_{2}} shifted to the right and the reflection of 8T1=1+1+16T28T_{1}=1+\sqrt{1+16T_{2}} shifted up.

Refer to caption
Figure 1. Prescribed Ricci curvature on the Stiefel manifold V2(4)V_{2}({\mathbb{R}}^{4})

Next, we give several examples that illustrate various behaviors.

Example 1.

Suppose that T0=1T_{0}=1, T1=T2=t>0T_{1}=T_{2}=t>0 and T32+T42=0T_{3}^{2}+T_{4}^{2}=0. A straightforward computation shows that the diagonal metrics with

(5.13) x0=32t+1,x1=x2=32t+1+32t+18andx3=x4=0\displaystyle x_{0}=\sqrt{32t+1},\qquad x_{1}=x_{2}=\frac{32t+1+\sqrt{32t+1}}{8}\qquad\mbox{and}\qquad x_{3}=x_{4}=0

solve the system (5.12) and hence are critical points of the functional S|TS_{|\mathcal{M}_{T}} when restricted to the set of diagonal metrics. By Remark 5.11, they are also critical points of the scalar curvature functional on all of T\mathcal{M}_{T}. Maple shows that all four eigenvalues of the Hessian are negative for t>14t>\frac{1}{4}, i.e., (5.13) defines a local maximum. By the Maple experiment in Example 3, it is a global maximum. On the other hand, if t<14t<\frac{1}{4}, we have two negative and two positive eigenvalues, which means (5.13) is a saddle. Thus global maxima among the set of diagonal metrics in T\mathcal{M}_{T} can, in fact, be saddles on all of T\mathcal{M}_{T}.

We point out that (5.13) is the Jensen Einstein metric when t=34t=\frac{3}{4}, and a global maximum. It is marked with the blue dot in Figure 1.

Example 2.

From the previous example we see that the metric with t=14t=\frac{1}{4} must be special. Thus we choose T0=1T_{0}=1, T1=T2=14T_{1}=T_{2}=\frac{1}{4} and T32+T42=0T_{3}^{2}+T_{4}^{2}=0. This choice of TT is marked by the red dot in Figure 1. Direct verification shows that the metrics gTg\in\mathcal{M}_{T} with

(5.14) x0=2t,x1=x2=t,x3=t2t32t1cosψandx4=t2t32t1sinψx_{0}=2t,\qquad x_{1}=x_{2}=t,\qquad x_{3}=t\sqrt{\frac{2t-3}{2t-1}}\cos\psi\qquad\mbox{and}\qquad x_{4}=t\sqrt{\frac{2t-3}{2t-1}}\sin\psi

are critical points of S|TS_{|\mathcal{M}_{T}} for t[32,)t\in\big{[}\frac{3}{2},\infty\big{)} and ψ[0,2π)\psi\in[0,2\pi) with scalar curvature 88. They form a surface diffeomorphic to 2{\mathbb{R}}^{2}, described in the coordinates (t,ψ)(t,\psi). The normalizer N0(H)/HN_{0}(H)/H acts on this surface via (t,ψ)(t,ψ+2η)(t,\psi)\to(t,\psi+2\eta). Thus metrics with the same value of ψ\psi are isometric. On the other hand, the squared volume of the metric (5.14) equals 8t5(2t1)2\frac{8t^{5}}{(2t-1)^{2}}, and hence metrics with different values of tt are not isometric. To determine the critical point type of (5.14), we compute the eigenvalues of the Hessian of S|TS_{|\mathcal{M}_{T}}. Two of them are always negative, and the other two vanish. The 0-eigenspace is tangent to the surface of critical points. Consequently, this surface is a non-degenerate critical submanifold with index 2. Using the Morse–Bott lemma, we conclude that it is isolated and a local maximum. Using the numerical discussion in Example 3, we conclude that it must in fact be a global maximum since otherwise the metrics in Example 1 with tt near 1/41/4 would have to have scalar curvature larger than 88.

Example 3.

Assume that T32+T42=0T_{3}^{2}+T_{4}^{2}=0. As explained below, S|TS_{|\mathcal{M}_{T}} has a non-diagonal critical point if and only if TT lies in the pink region in Figure 2. Since the action of N0(H)/HN_{0}(H)/H leaves TT unchanged, we obtain a circle of non-diagonal critical points for each such TT. It is always a non-degenerate critical submanifold of index 2 and co-index 1. In addition, we have a diagonal critical point for TT between the thick curves. By Proposition 5.3, this diagonal critical point must be a global maximum when TT lies in the dark-grey or the middle-grey region in Figure 1. Computation of the eigenvalues of the Hessian indicates that it is a local maximum for TT in the rest of the dotted region and a saddle with index and co-index 2 for TT in the yellow region. The transition from the pink to the yellow region is achieved across the curve

(5.15) T1(t)=4t2(1t)16t4+1,T2(t)=t(4t1)16t4+1,t(14,1).T_{1}(t)=\frac{4t^{2}(1-t)}{16t^{4}+1},\qquad T_{2}(t)=\frac{t(4t-1)}{16t^{4}+1},\qquad t\in\big{(}\tfrac{1}{4},1\big{)}.

The Hessian at the critical points corresponding to this curve has two zero eigenvalues. However, we do not know whether these critical points lie on critical submanifolds as in Example 2.

To verify the above claim about S|TS_{|\mathcal{M}_{T}} having a non-diagonal critical point for TT in the pink region, we first observe that Ric(g)\operatorname{Ric}(g) is diagonal if and only if x3=x4=0x_{3}=x_{4}=0 or x02=4x1x2x_{0}^{2}=4x_{1}x_{2}. This is because the off-diagonal entries in the Ricci curvature are

xi(4x1x2x02)x0(x1x2x32x42),i=1,2.\frac{x_{i}(4x_{1}x_{2}-x_{0}^{2})}{x_{0}(x_{1}x_{2}-x_{3}^{2}-x_{4}^{2})},\qquad i=1,2.

Since the action of N0(H)/HN_{0}(H)/H leaves diagonal Ricci curvature unchanged, it suffices to consider only metrics gg with x4=0x_{4}=0. We take two million non-diagonal gg satisfying x02=4x1x2x_{0}^{2}=4x_{1}x_{2} and x4=0x_{4}=0 and calculate Ric(g)\operatorname{Ric}(g). After checking that Ric(g)\operatorname{Ric}(g) is positive-definite and normalizing so that its 𝔪0{\mathfrak{m}}_{0} component becomes 1, we mark the 𝔪1{\mathfrak{m}}_{1} and 𝔪2{\mathfrak{m}}_{2} components in the (T1,T2)(T_{1},T_{2})-plane as in Figure 2. The obtained values fill up the pink region. Note that the critical circles, which are obtained by applying the normalizer, have radius x3x_{3} with 0<x3<120<x_{3}<\frac{1}{2}. As this radius becomes zero, we obtain the curve in (5.15). Hence here the critical circles merge with the diagonal critical point. In the boundary on the right-hand side of the pink region, we necessarily have critical circles with at least two zero eigenvalues in the Hessian since, as one moves to the right, they must disappear.

For a specific example, choose T0=1T_{0}=1, T1=135472T_{1}=\frac{135}{472} and T2=15118T_{2}=\frac{15}{118}, the green dot in Figure 2. Direct verification shows that the gradient of S|TS_{|\mathcal{M}_{T}} vanishes at the circle of metrics with x0=18159x_{0}=\frac{181}{59}, x1=905472x_{1}=\frac{905}{472}, x2=362295x_{2}=\frac{362}{295} and x32+x42=3276155696x_{3}^{2}+x_{4}^{2}=\frac{32761}{55696}. These metrics are all saddles with index 2 and co-index 1. At the same time, the diagonal metric with

x0=2086659,x1=62598+2772086654044andx2=834644392086622892x_{0}=\frac{\sqrt{20866}}{59},\qquad x_{1}=\frac{62598+277\sqrt{20866}}{54044}\qquad\mbox{and}\qquad x_{2}=\frac{83464-439\sqrt{20866}}{22892}

is a strict local maximum. It cannot be a global maximum because its scalar curvature is less than α𝔨0\alpha_{{\mathfrak{k}}^{0}} and there are no further critical points.

It follows that, outside the pink region, the functional has only a diagonal critical point. This must be the unique critical point. Thus in the dark-grey or middle-grey region in Figure 1 the global maximum must be a diagonal metric.

Refer to caption
Figure 2. Ricci curvature of non-diagonal metrics on the Stiefel manifold V2(4)V_{2}({\mathbb{R}}^{4})

As far as non-diagonal Ricci candidates TT are concerned, we make the following remarks. Recall that V2(4)V_{2}({\mathbb{R}}^{4}) supports two circles of (non-diagonal) Einstein metrics isometric to the canonical product Einstein metric on V2(4)𝕊3×𝕊2V_{2}(\mathbb{R}^{4})\simeq\mathbb{S}^{3}\times\mathbb{S}^{2}; see [5]. Scaling each factor does not change the Ricci curvature and hence yields an arc of critical points. These arcs turn out to be non-degenerate critical submanifolds with index 3 and local maxima. Applying the normalizer, we obtain a circle of candidates TT, each one of which admits two arcs of critical points.

The Stiefel manifold is also a generalized Wallach space under the decomposition Ds,θD_{s,\theta} with s=12s=\frac{1}{\sqrt{2}} and any θ\theta. These spaces we will study in detail in [27] in order to produce saddle critical points. Using the methods of [27], one shows that S|TS_{|\mathcal{M}_{T}} admits a critical point of co-index 0 or 1 if

T1=T2<142|T32T42|T32+T42.T_{1}=T_{2}<\frac{1}{4}-\frac{2|T_{3}^{2}-T_{4}^{2}|}{\sqrt{T_{3}^{2}+T_{4}^{2}}}.

When T3=T4=0T_{3}=T_{4}=0, this should be interpreted as T1=T2<14T_{1}=T_{2}<\frac{1}{4}, corresponding to the saddles discussed above in Example 1.

Finally, we remark that one can determine the metrics that are degenerate critical points. In [27, Proposition 2.30] we show that gg is a non-degenerate critical point of S|TS_{|\mathcal{M}_{T}} if and only if rkdRicg=dim1\operatorname{rk}d\operatorname{Ric}_{g}=\dim\mathcal{M}-1. Assuming x0=1x_{0}=1 and x4=0x_{4}=0, a Maple computation shows that this only fails for three families of metrics. The first one satisfies 4x1x2=14x_{1}x_{2}=1, and these are the metrics with diagonal Ricci tensor in Example 3. The second one is given by x3=12x_{3}=\frac{1}{2} with x1x2x_{1}\neq x_{2}, and for these the Ricci tensor is not diagonal. The third family consists of the metrics for which

(16x12x32+4x128x32)x22+(16x1x34+8x1x32x1)x2+x328x12x32+4x34=0.(16x_{1}^{2}x_{3}^{2}+4x_{1}^{2}-8x_{3}^{2})x_{2}^{2}+(-16x_{1}x_{3}^{4}+8x_{1}x_{3}^{2}-x_{1})x_{2}+x_{3}^{2}-8x_{1}^{2}x_{3}^{2}+4x_{3}^{4}=0.

5.2. The Ledger–Obata space H𝟑/diag(H)\mathbf{\emph{H}^{3}/\text{diag}(\!\emph{H})} and Spin(𝟖)/G𝟐\mathbf{\emph{Spin}(8)/\emph{G}_{2}}

Our second example is motivated by the following observations. Let the homogeneous space G/HG/H be such that 𝔪=𝔪1𝔪2{\mathfrak{m}}={\mathfrak{m}}_{1}\oplus{\mathfrak{m}}_{2} with both summands irreducible. When 𝔪1{\mathfrak{m}}_{1} and 𝔪2{\mathfrak{m}}_{2} are inequivalent, one can solve the equation Ric(g)=cT\operatorname{Ric}(g)=cT directly; see [25]. There are only two cases where 𝔪1{\mathfrak{m}}_{1} and 𝔪2{\mathfrak{m}}_{2} are equivalent representations. If GG is simple, then G/HG/H must equal Spin(8)/G2=𝕊7×𝕊7Spin(8)/G_{2}=\mathbb{S}^{7}\times\mathbb{S}^{7}; see the classification in [13, 18]. If it is not, one easily checks that the only possibility is that G/HG/H is the Ledger–Obata space H3/diag(H)H^{3}/\operatorname{diag}(H), where HH is simple and embedded diagonally.

We discuss here the latter case, the former one being quite similar; see Remark 5.22. In what follows, (ei)(e_{i}) is a basis of 𝔥{\mathfrak{h}} orthonormal with respect to BH-B_{H}, the negative of the Killing form of HH.

Let G/H=H3/diag(H)G/H=H^{3}/\operatorname{diag}(H). There are exactly three intermediate subgroups, namely,

K1\displaystyle K_{1} ={(a,b,b)Ga,bH},\displaystyle=\{(a,b,b)\in G\mid a,b\in H\},
K2\displaystyle K_{2} ={(a,b,a)Ga,bH}andK3={(a,a,b)Ga,bH}.\displaystyle=\{(a,b,a)\in G\mid a,b\in H\}\qquad\mbox{and}\qquad K_{3}=\{(a,a,b)\in G\mid a,b\in H\}.

The outer automorphism R:GGR:G\to G given by R(a,b,c)=(c,a,b)R(a,b,c)=(c,a,b) interchanges these subgroups. Choose the bi-invariant metric QQ on GG to be Q=BHBHBHQ=-B_{H}-B_{H}-B_{H}. Then RR is an isometry of (G,Q)(G,Q).

Fix an AdH\operatorname{Ad}_{H}-invariant, QQ-orthogonal decomposition of 𝔪{\mathfrak{m}} by setting

(5.16) 𝔪1={(2X,X,X)𝔤X𝔥}and𝔪2={(0,X,X)𝔤X𝔥}.{\mathfrak{m}}_{1}=\{(-2X,X,X)\in{\mathfrak{g}}\mid X\in{\mathfrak{h}}\}\qquad\text{and}\qquad{\mathfrak{m}}_{2}=\{(0,X,-X)\in{\mathfrak{g}}\mid X\in{\mathfrak{h}}\}.

Then 𝔨1=𝔥𝔪1{\mathfrak{k}}_{1}={\mathfrak{h}}\oplus{\mathfrak{m}}_{1}. Furthermore, AdH\operatorname{Ad}_{H} is an irreducible representation of real type on each 𝔪i{\mathfrak{m}}_{i} since HH is simple. The collections (16(2ej,ej,ej))\Big{(}\frac{1}{\sqrt{6}}(-2e_{j},e_{j},e_{j})\Big{)} and (12(0,ej,ej))\Big{(}\frac{1}{\sqrt{2}}(0,e_{j},-e_{j})\Big{)} constitute QQ-orthonormal bases of 𝔪1{\mathfrak{m}}_{1} and 𝔪2{\mathfrak{m}}_{2}. With respect to these bases, the metric and TT have the form

(5.17) g=(x1x3x3x2)andT=(T1T3T3T2)\centering g=\left(\begin{matrix}x_{1}&x_{3}\\ x_{3}&x_{2}\end{matrix}\right)\qquad\text{and}\qquad T=\left(\begin{matrix}T_{1}&T_{3}\\ T_{3}&T_{2}\end{matrix}\right)\@add@centering

with detg>0\det g>0 and detT>0\det T>0. (Each entry in these matrices represents a scalar a×aa\times a matrix, where a=dimHa=\dim H.) Applying our main theorem, we obtain the following result.

Proposition 5.18.

Suppose that G/H=H3/diag(H)G/H=H^{3}/\operatorname{diag}(H) and T1T2>T32T_{1}T_{2}>T_{3}^{2}. The functional S|TS_{|\mathcal{M}_{T}} attains its global maximum if

(5.19) 34T2<T1<95T21435|T3|.\displaystyle\frac{3}{4}T_{2}<T_{1}<\frac{9}{5}T_{2}-\frac{14\sqrt{3}}{5}|T_{3}|.
Proof.

We need to compute α𝔨i\alpha_{{\mathfrak{k}}_{i}} and β𝔨i\beta_{{\mathfrak{k}}_{i}}. A well-known formula for the sectional curvature of a normal homogeneous metric (see [4, Proposition 7.87b]) implies

S(Q|G/K1)=j,ksec(Q|G/K1)((0,ej,ej)2,(0,ek,ek)2)=12j,k|[ej,ek]|2,\displaystyle S(Q_{|G/K_{1}})=\sum_{j,k}\operatorname{sec}(Q_{|G/K_{1}})\bigg{(}\frac{(0,e_{j},-e_{j})}{\sqrt{2}},\frac{(0,e_{k},-e_{k})}{\sqrt{2}}\bigg{)}=\frac{1}{2}\sum_{j,k}|[e_{j},e_{k}]|^{2},

where |||\cdot| is the norm corresponding to BH-B_{H}. Similarly,

S(Q|K1/H)=j,ksec(Q|K1/H)((2ej,ej,ej)6,(2ek,ek,ek)6)=38j,k|[ej,ek]|2.\displaystyle S(Q_{|K_{1}/H})=\sum_{j,k}\operatorname{sec}(Q_{|K_{1}/H})\bigg{(}\frac{(-2e_{j},e_{j},e_{j})}{\sqrt{6}},\frac{(-2e_{k},e_{k},e_{k})}{\sqrt{6}}\bigg{)}=\frac{3}{8}\sum_{j,k}|[e_{j},e_{k}]|^{2}.

In addition, (K1/H,Q)(K_{1}/H,Q) is an isotropy irreducible symmetric space, and hence [𝔪1,𝔪1]𝔥[{\mathfrak{m}}_{1},{\mathfrak{m}}_{1}]\subset{\mathfrak{h}}. It follows from (1.4) that

S(Q|G/K1)=a2and, by the above,S(Q|K1/H)=3a8S(Q_{|G/K_{1}})=\frac{a}{2}\qquad\text{and, by the above,}\qquad S(Q_{|K_{1}/H})=\frac{3a}{8}

with a=dimHa=\dim H. The trace constraints on G/K1G/K_{1} and K1/HK_{1}/H have the form x2=aT2x_{2}=aT_{2} and x1=aT1x_{1}=aT_{1}, respectively. We conclude that

α𝔨1=38T1andβ𝔨1=12T2,\alpha_{{\mathfrak{k}}_{1}}=\frac{3}{8T_{1}}\qquad\text{and}\qquad\beta_{{\mathfrak{k}}_{1}}=\frac{1}{2T_{2}},

which means β𝔨1α𝔨1>0\beta_{{\mathfrak{k}}_{1}}-\alpha_{{\mathfrak{k}}_{1}}>0 if and only if 4T1>3T24T_{1}>3T_{2}.

As observed above, the automorphism RR takes K1K_{1} to K2K_{2}. Since R3R^{3} is the identity, the matrix of the pullback of TT by RR with respect to our fixed bases of 𝔪1{\mathfrak{m}}_{1} and 𝔪2{\mathfrak{m}}_{2} is

(12323212)(T1T3T3T2)(12323212)=(T1+3T223T343(T1T2)2T343(T1T2)2T343T1+T2+23T34).\left(\begin{matrix}-\frac{1}{2}&\frac{\sqrt{3}}{2}\\ -\frac{\sqrt{3}}{2}&-\frac{1}{2}\end{matrix}\right)\left(\begin{matrix}T_{1}&T_{3}\\ T_{3}&T_{2}\end{matrix}\right)\left(\begin{matrix}-\frac{1}{2}&-\frac{\sqrt{3}}{2}\\ \frac{\sqrt{3}}{2}&-\frac{1}{2}\end{matrix}\right)=\left(\begin{matrix}\frac{T_{1}+3T_{2}-2\sqrt{3}T_{3}}{4}&\frac{\sqrt{3}(T_{1}-T_{2})-2T_{3}}{4}\\ \frac{\sqrt{3}(T_{1}-T_{2})-2T_{3}}{4}&\frac{3T_{1}+T_{2}+2\sqrt{3}T_{3}}{4}\end{matrix}\right).

The maps from (G/K1,Q)(G/K_{1},Q) to (G/K2,Q)(G/K_{2},Q) and from (K1/H,Q)(K_{1}/H,Q) to (K2/H,Q)(K_{2}/H,Q) induced by RR are isometries. They preserve the scalar curvature, so

α𝔨2=32(T1+3T223T3)andβ𝔨2=23T1+T2+23T3.\alpha_{{\mathfrak{k}}_{2}}=\frac{3}{2(T_{1}+3T_{2}-2\sqrt{3}T_{3})}\qquad\text{and}\qquad\beta_{{\mathfrak{k}}_{2}}=\frac{2}{3T_{1}+T_{2}+2\sqrt{3}T_{3}}.

Consequently, β𝔨2α𝔨2>0\beta_{{\mathfrak{k}}_{2}}-\alpha_{{\mathfrak{k}}_{2}}>0 if and only if 5T1<9T2143T35T_{1}<9T_{2}-14\sqrt{3}T_{3}.

To obtain the third subgroup, we need to apply RR to K2K_{2}, or R2R^{2} to K1K_{1}. The matrix of the pullback of TT by R2R^{2} with respect to (5.16) is

(T1+3T2+23T343(T2T1)2T343(T2T1)2T343T1+T223T34).\left(\begin{matrix}\frac{T_{1}+3T_{2}+2\sqrt{3}T_{3}}{4}&\frac{\sqrt{3}(T_{2}-T_{1})-2T_{3}}{4}\\ \frac{\sqrt{3}(T_{2}-T_{1})-2T_{3}}{4}&\frac{3T_{1}+T_{2}-2\sqrt{3}T_{3}}{4}\end{matrix}\right).

This implies

α𝔨3=32(T1+3T2+23T3)andβ𝔨3=23T1+T223T3.\alpha_{{\mathfrak{k}}_{3}}=\frac{3}{2(T_{1}+3T_{2}+2\sqrt{3}T_{3})}\qquad\text{and}\qquad\beta_{{\mathfrak{k}}_{3}}=\frac{2}{3T_{1}+T_{2}-2\sqrt{3}T_{3}}.

Thus β𝔨3α𝔨3>0\beta_{{\mathfrak{k}}_{3}}-\alpha_{{\mathfrak{k}}_{3}}>0 if and only if 5T1<9T2+143T35T_{1}<9T_{2}+14\sqrt{3}T_{3}.

There are two scenarios for the condition of our main theorem to be satisfied. One is that α𝔨1max{α𝔨2,α𝔨3}\alpha_{{\mathfrak{k}}_{1}}\geq\max\{\alpha_{{\mathfrak{k}}_{2}},\alpha_{{\mathfrak{k}}_{3}}\} (equivalently, 3T13T223|T3|3T_{1}\leq 3T_{2}-2\sqrt{3}|T_{3}|), in which case it suffices to demand that 4T1>3T2{4T_{1}>3T_{2}}. The other is that α𝔨1max{α𝔨2,α𝔨3}\alpha_{{\mathfrak{k}}_{1}}\leq\max\{\alpha_{{\mathfrak{k}}_{2}},\alpha_{{\mathfrak{k}}_{3}}\}. Then we need 5T1<9T2143|T3|5T_{1}<9T_{2}-14\sqrt{3}|T_{3}|. Combining these conditions, we arrive at (5.19). ∎

For gg and TT of the form (5.17), the constraint trgT=1\operatorname{tr}_{g}T=1 is

a(x1T2+x2T12x3T3)x1x2x32=1,\frac{a(x_{1}T_{2}+x_{2}T_{1}-2x_{3}T_{3})}{x_{1}x_{2}-x_{3}^{2}}=1,

and a computation shows that the scalar curvature is given by

(5.20) S(g)=a(9x1x22+12x12x26x1x3218x2x32x13)24(x1x2x32)2.S(g)=\frac{a(9x_{1}x_{2}^{2}+12x_{1}^{2}x_{2}-6x_{1}x_{3}^{2}-18x_{2}x_{3}^{2}-x_{1}^{3})}{24(x_{1}x_{2}-x_{3}^{2})^{2}}.

Thus we have the same critical points, up to scaling, no matter what group HH we choose. The formulas

x=12(T1T2)andy=T3x=\frac{1}{2}(T_{1}-T_{2})\qquad\mbox{and}\qquad y=T_{3}

define coordinates in the space of Ricci candidates TT normalized so that T1+T2=1T_{1}+T_{2}=1. Figure 3 shows points in the (x,y)(x,y)-plane that correspond to different behaviors of S|TS_{|\mathcal{M}_{T}}. The tensor TT is positive-definite if and only if x2+y2<14x^{2}+y^{2}<\frac{1}{4}. The order-three automorphism RR yields a natural symmetry on the space of Ricci candidates. Its action rotates the picture in Figure 3 by 2π3\frac{2\pi}{3}. The inside of the large triangle is the set of (x,y)(x,y) that satisfy the condition of Proposition 5.18. The equalities α𝔨1=αG/H\alpha_{{\mathfrak{k}}_{1}}=\alpha_{G/H}, α𝔨2=αG/H\alpha_{{\mathfrak{k}}_{2}}=\alpha_{G/H} and α𝔨3=αG/H\alpha_{{\mathfrak{k}}_{3}}=\alpha_{G/H} hold in the dark-grey, the middle grey and the light-grey triangle, respectively. A computer-assisted experiment with one million metrics shows that S|TS_{|\mathcal{M}_{T}} also has a critical point for (x,y)(x,y) in the pink regions. Indefinite tensors that are Ricci curvatures of metrics up to scaling fill up the blue regions. Thus the image of the Ricci map is the union of the grey, pink and blue areas. In particular, for any TT in the white region S|TS_{|\mathcal{M}_{T}} has no critical points.

Maple is able to solve the Euler–Lagrange equations for S|TS_{|\mathcal{M}_{T}} for any specific choice of TT. It suggests that the solution representing a metric, when it exists, is unique and is always a local maximum (global if TT lies inside the triangle).

Refer to caption
Figure 3. Prescribed Ricci curvature on the Ledger–Obata space H3/diag(H)H^{3}/\operatorname{diag}(H)
Example 4.

Assume that T1=tT_{1}=t, T2=1tT_{2}=1-t and T3=0T_{3}=0. One easily checks that the formulas

(5.21) x1=a20t2+30t9,x2=a20t2+30t9+t20t2+30t93(7t3)andx3=0\displaystyle x_{1}=a\sqrt{-20t^{2}+30t-9},\quad x_{2}=a\frac{-20t^{2}+30t-9+t\sqrt{-20t^{2}+30t-9}}{3(7t-3)}\quad\text{and}\qquad x_{3}=0

define a metric with positive Ricci curvature for 37<t<1\frac{3}{7}<t<1, and this metric is a critical point of S|TS_{|\mathcal{M}_{T}}. Computing the Hessian, we conclude that it is a strict local maximum unless t=34t=\frac{3}{4}. In Figure 3 the Ricci candidates with T1=t(37,1)T_{1}=t\in\big{(}\frac{3}{7},1\big{)}, T2=1tT_{2}=1-t and T3=0T_{3}=0 occupy the segment of the xx-axis with x(114,12)x\in\big{(}\!-\frac{1}{14},\frac{1}{2}\big{)}. The orbit of this segment under RR, indicated by the green lines, gives two further segments with the same behavior.

The case of t=34t=\frac{3}{4}, respectively x=14x=\frac{1}{4}, is special. The orbit of the corresponding tensor TT under RR is depicted by the red dots in Figure 3. The three tensors in this orbit are, in fact, Einstein metrics. The H3H^{3}-equivariant diffeomorphisms from H3/diag(H)H^{3}/\operatorname{diag}(H) to H×HH\times H given by

(a,b,c)diag(H)\displaystyle(a,b,c)\operatorname{diag}(H) (ac1,bc1),\displaystyle\mapsto(ac^{-1},bc^{-1}),
(a,b,c)diag(H)\displaystyle(a,b,c)\operatorname{diag}(H) (ab1,cb1)and(a,b,c)diag(H)(ba1,ca1)\displaystyle\mapsto(ab^{-1},cb^{-1})\qquad\mbox{and}\qquad(a,b,c)\operatorname{diag}(H)\mapsto(ba^{-1},ca^{-1})

are isometries between these Einstein metrics and the canonical Einstein product metric on H×HH\times H. Scaling the factors in the product metric, we obtain arcs of critical points. As it turns out, these arcs are non-degenerate critical submanifolds with index 1 and co-index 0 and hence consist of local maxima.

The only other Einstein metric on H3/diag(H)H^{3}/\operatorname{diag}(H) is the normal homogeneous metric induced by QQ; see [16]. It corresponds to the origin in Figure 3 and is a strict local maximum of the associated functional S|TS_{|\mathcal{M}_{T}}.

As in the case of the Stiefel manifold, one easily determines which critical points are degenerate. It turns out that this only happens for the “red dot” Einstein metrics discussed above. Thus for the Ledger–Obata space all the functionals S|TS_{|\mathcal{M}_{T}} are Morse or Morse–Bott functions.

Remark 5.22.

The remaining homogeneous space with two equivalent summands is Spin(8)/G2Spin(8)/G_{2}. It is, in fact, quite similar since the scalar curvature SS is, remarkably, given by the same formula as (5.20) with a=84a=84; see [19]. There are three intermediate subgroups isomorphic to Spin(7)Spin(7), which are permuted by the triality automorphism of Spin(8)Spin(8). One easily sees that the conditions for a global maximum on Spin(8)/G2Spin(8)/G_{2} are the same as on the Ledger–Obata space. The set of Einstein metrics consists of three metrics isometric to the canonical product Einstein metric on 𝕊7×𝕊7\mathbb{S}^{7}\times\mathbb{S}^{7}, and the normal homogeneous metric.


References

  • [1]
  • [2] R.M. Arroyo, M.D. Gould, A. Pulemotov, The prescribed Ricci curvature problem for naturally reductive metrics on non-compact simple Lie groups, submitted, arXiv:2006.15765, 2020.
  • [3] R.M. Arroyo, A. Pulemotov, W. Ziller, The prescribed Ricci curvature problem for naturally reductive metrics on compact Lie groups, Differential Geom. Appl. 78 (2021), article 101794.
  • [4] A. Besse, Einstein manifolds, Springer-Verlag, Berlin, 1987.
  • [5] C. Böhm, M. Wang, W. Ziller, A variational approach for compact homogeneous Einstein manifolds, Geom. Funct. Anal. 14 (2004) 681–733.
  • [6] T. Buttsworth, The prescribed Ricci curvature problem on three-dimensional unimodular Lie groups, Math. Nachr. 292 (2019) 747–759.
  • [7] T. Buttsworth, A.M. Krishnan, Prescribing Ricci curvature on a product of spheres, to appear in Ann. Mat. Pura Appl, arXiv:2010.05118, 2020.
  • [8] T. Buttsworth, A. Pulemotov, The prescribed Ricci curvature problem for homogeneous metrics, in: Differential geometry in the large (O. Dearricott et al., eds), Cambridge University Press, 2021, 169–192.
  • [9] T. Buttsworth, A. Pulemotov, Y.A. Rubinstein, W. Ziller, On the Ricci iteration for homogeneous metrics on spheres and projective spaces, Transform. Groups 26 (2021) 145–164.
  • [10] J. Cao, D.M. DeTurck, The Ricci curvature equation with rotational symmetry, Amer. J. Math. 116 (1994) 219–241.
  • [11] D.M. DeTurck, Prescribing positive Ricci curvature on compact manifolds, Rend. Sem. Mat. Univ. Politec. Torino 43 (1985) 357–369.
  • [12] D.M. DeTurck, N. Koiso, Uniqueness and nonexistence of metrics with prescribed Ricci curvature, Ann. Inst. H. Poincaré Anal. Non Linéaire 1 (1984) 351–359.
  • [13] W. Dickinson, M. Kerr, The geometry of compact homogeneous spaces with two isotropy summands, Ann. Global Anal. Geom. 34 (2008) 329–350.
  • [14] M.D. Gould, A. Pulemotov, Homogeneous metrics with prescribed Ricci curvature on spaces with non-maximal isotropy, to appear in Commun. Anal. Geom., arXiv:1710.03024, 2017.
  • [15] K. Grove, W. Ziller, Cohomogeneity one manifolds with positive Ricci curvature, Invent. Math. 149 (2002) 619–646.
  • [16] Z. Chen, Y.G. Nikonorov, Y.V. Nikonorova, Invariant Einstein metrics on Ledger–Obata spaces, Differential Geom. Appl. 50 (2017) 71–87.
  • [17] R.S. Hamilton, The Ricci curvature equation, in: Seminar on nonlinear partial differential equations (S.-S. Chern, ed.), Springer-Verlag, New York, 1984, 47–72.
  • [18] C. He, Cohomogeneity one manifolds with a small family of invariant metrics, Geom. Dedicata 157 (2012) 41–90.
  • [19] M. Kerr, New examples of homogeneous Einstein metrics, Michigan J. Math. 45 (1998) 115–134.
  • [20] J.L. Koszul, Sur la forme hermitienne canonique des espaces homogènes complexes, Canadian J. Math. 7 (1955) 562–576.
  • [21] J. Lauret, C.E. Will, Prescribing Ricci curvature on homogeneous spaces, arXiv:2010.03643, 2020.
  • [22] J. Lauret, C.E. Will, On the stability of homogeneous Einstein manifolds II, arXiv:2107.00354, 2021.
  • [23] A. Pulemotov, Metrics with prescribed Ricci curvature near the boundary of a manifold, Math. Ann. 357 (2013) 969–986.
  • [24] A. Pulemotov, The Dirichlet problem for the prescribed Ricci curvature equation on cohomogeneity one manifolds, Ann. Mat. Pura Appl. 195 (2016) 1269–1286.
  • [25] A. Pulemotov, Metrics with prescribed Ricci curvature on homogeneous spaces, J. Geom. Phys. 106 (2016) 275–283.
  • [26] A. Pulemotov, Maxima of curvature functionals and the prescribed Ricci curvature problem on homogeneous spaces, J. Geom. Anal. 30 (2020) 987–1010.
  • [27] A. Pulemotov, W. Ziller, Palais–Smale sequences for the prescribed Ricci curvature functional, in preparation, 2023.
  • [28] M.Y. Wang, W. Ziller, Existence and nonexistence of homogeneous Einstein metrics, Invent. Math. 84 (1986) 177–194.
  • [29] M.Y. Wang, W. Ziller, On isotropy irreducible Riemannian manifolds, Acta Math. 166 (1991) 223–261.