On the variational properties of the prescribed Ricci curvature functional
Abstract.
We study the prescribed Ricci curvature problem for homogeneous metrics. Given a (0,2)-tensor field , this problem asks for solutions to the equation for some constant . 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.
The prescribed Ricci curvature problem consists in finding a Riemannian metric on a manifold such that for a given (0,2)-tensor field . As was suggested by DeTurck [11] and Hamilton [17], it is more natural to ask instead whether one can solve the equation
for some constant . In fact, on a compact manifold, such a constant appears to be necessary. For example, if 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 and the tensor are invariant under a Lie group acting on . In the case where the quotient is one-dimensional, the problem was addressed by Hamilton [17], Cao–DeTurck [10], Pulemotov [23, 24] and Buttsworth–Krishnan [7]. The case where is a homogeneous space 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 is positive-definite. Without this assumption, the behavior of is very different. General existence theorems in the homogeneous setting rely on the fact, proven in [25], that -invariant metrics on with Ricci curvature are precisely (up to scaling) the critical points of the scalar curvature functional on the set of -invariant metrics on subject to the constraint . We will assume that and 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 . 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 is maximal in , the supremum of is always attained. Otherwise, the global behavior of depends on the set of intermediate subgroups, i.e., Lie groups with . For every such , one has the homogeneous fibration
If we fix homogeneous metrics and on the fiber and the base , it is natural to study the two-parameter variation
of . Notice that we define this variation using the reciprocals of and . In fact, we often deal with inverses of metrics rather than metrics themselves, which enables us to view as a pre-compact space. Solving the constraint for , we obtain a one-parameter family , called a canonical variation, with scalar curvature given by
Here and are the traces of on the fiber and the base, and are the scalar curvatures of and , and is the O’Neil tensor of the Riemannian submersion. As a consequence,
which means that and control the behavior of at infinity. It is easy to see that is bounded from above if is positive-definite, and it is therefore natural to search for a global maximum of . The above formulas motivate two invariants of the intermediate subgroup with Lie algebra :
These quantities maximize the limits and the “derivatives” of at infinity. We furthermore introduce an invariant of the homogeneous space :
where the supremum is taken over all Lie algebras of intermediate subgroups. We can now state our main result.
Theorem.
Let be a compact homogeneous space and a positive-definite -invariant (0,2)-tensor field on . If is not maximal in , then the set of intermediate subgroups with is non-empty. If is such a subgroup of the lowest possible dimension and , then achieves its supremum at some metric and hence for some .
In fact, we show that there exists such that the set is non-empty and compact, which implies the result. This requires careful estimates of the behavior of at infinity.
We illustrate this result by examining as examples the Stiefel manifold 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 are complicated and cannot be solved explicitly. Our theorem implies the existence of a large set of tensors such that has a global maximum, although in many cases, has no critical points at all. These examples also have interesting features for critical points besides global maxima. We find:
-
A tensor such that admits a two-dimensional submanifold of critical points. This submanifold is non-degenerate and a local maximum.
-
A large set of critical points that are global maxima among diagonal metrics but saddles in the space of all invariant metrics.
-
A set of tensors for which has both a circle of saddles and a strict local maximum, although for some of these , 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 -invariant metrics on which the Ricci map is a local diffeomorphism (up to scaling).
Finally, we observe that the functional has a large set of critical points on flag manifolds, e.g., . Let be a complex structure on a compact generalized flag manifold , where is the centralizer of a torus in . It is well known that there exists a unique -invariant Hermitian bi-linear form such that for every Kähler metric compatible with ; see [20]. Since the set of such metrics has the same dimension as the torus , we obtain a smooth critical submanifold in of dimension . Furthermore, this critical submanifold contains the unique Kähler–Einstein metric associated to . For instance, if , there is an -dimensional smooth submanifold of Kähler metrics consisting of critical points of for equal to (up to scaling). In general, there exist several inequivalent complex structures on and hence non-isometric critical submanifolds.
We now outline the strategy of our proofs. To describe the set of homogeneous metrics on , one decomposes the tangent space into a sum of irreducible modules and considers so-called diagonal metrics
where is a fixed bi-invariant metric on the Lie algebra of . Not every homogeneous metric is of this form when some of the summands are equivalent. However, it can always be reduced to this form by choosing another decomposition of as above. With the substitution , the constraint becomes simply , where and . Thus the set of diagonal metrics in is parameterized by a simplex with stratified boundary . In this construction, the strata are indexed by the sets of those variables that vanish. Some of them are marked by subalgebras with and denoted . If is a divergent sequence of metrics of bounded scalar curvature, then there exist such an intermediate subalgebra and a subsequence of that converges to a point in . At the remaining strata, denoted , the scalar curvature goes to . We will in fact show that is the largest possible limit of the scalar curvature achieved as one approaches . Since is bounded from above, it is natural to look at subalgebras with and determine conditions under which there are metrics with scalar curvature larger than . This is achieved by using the “derivatives” , and if such metrics exists, we show that is compact for some . One of the main difficulties is that, generally speaking, the scalar curvature does not extend continuously to and may not be achieved by a metric on . In addition, may not be an actual derivative. This also explains, in part, why we cannot arbitrarily choose a subalgebra with in our main theorem. For the proof, we need to produce careful estimates for the scalar curvature near . The final difficulty is to extend our arguments to the set of all metrics in when some of the isotropy summands are equivalent and 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 and the stratification of by subalgebras, as well as the behavior of the scalar curvature near the strata in . In Section 3 we use Riemannian submersions defined by intermediate subgroups to understand when there exist metrics with near a stratum . 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 be two compact Lie groups with Lie algebras such that is an almost effective homogeneous space. We fix a bi-invariant metric on , which defines a -orthogonal -invariant splitting , i.e., . The tangent space is identified with , and acts on via the adjoint representation . A -invariant metric on is determined by an -invariant inner product on . We denote by , or sometimes simply by , the space of -invariant metrics on . We will assume throughout the paper that is not a torus since in this case all -invariant metrics are flat, in particular, for all .
We describe metrics in in terms of -invariant splittings. Under the action of on , we decompose
where acts irreducibly on . Some of these summands may need to be one-dimensional if there exists a subspace of on which acts as the identity. We denote by the space of all such decompositions and use the letter for a particular choice of decomposition. The space has a natural topology induced from the embedding into the products of Grassmannians of -planes in . Clearly, is compact.
If is a -invariant symmetric bi-linear form field on , it is determined by its value on . We are interested when such a bi-linear form field is (up to scaling) the Ricci curvature of a metric , i.e., when
(1.1) |
Throughout the paper we will assume that is positive-definite. We may also assume that in (1.1) since a compact homogeneous space does not admit any metrics with unless it is a torus (see [4, Theorem 1.84]), which we excluded above.
Define the hypersurface
where is the trace with respect to . We denote it simply by 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
where is the scalar curvature of . More precisely, the following result holds.
Proposition 1.2.
The Ricci curvature of a metric equals for some if and only if is a critical point of .
Here is the Lagrange multiplier of the variational problem. Our main interest in this paper is to describe the geometry of the functional 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 , we have for some constant . In general, and do not have to be orthogonal if some of these summands are equivalent. But we can diagonalize and simultaneously, and hence there exists a decomposition such that the metric has the form
(1.3) |
We call such metrics diagonal with respect to our choice of and denote their set by , or simply . Thus, . We also denote . The scalars are simply the eigenvalues of with respect to . When these eigenvalues have multiplicity, and some of the modules in the corresponding eigenspace are equivalent under the action of , we note that for a compact infinite family of decompositions . In order to describe all homogeneous metrics on , 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 -invariant Einstein metrics.
We define the structure constants
where , and are -orthonormal bases of , and . Clearly, , and if and only if . We will denote by the Killing form of . By the irreducibility of , there exist constants such that
with if and only if lies in the center of . Furthermore, not all of vanish since otherwise is in the center and is a torus. Using this notation, and , the scalar curvature is given by
(1.4) |
(see [28]). For the tensor , we introduce the constants such that
Varying over all decompositions, these constants determine uniquely.
We now recall some formulas for Riemannian submersions which will be useful for us. Let be a compact subgroup with Lie algebra lying between and , i.e., . We then have a homogeneous fibration
We will often consider metrics with respect to which the projection is a Riemannian submersion. Assume that the decomposition is orthogonal with respect to both and . Then is a Riemannian submersion metric if and only if is -invariant. In this case, can be thought of as a homogeneous metric on the fiber , and a homogeneous metric on the base . We can introduce a new submersion metric on by scaling the fiber and the base, i.e.,
One has the following formula for the scalar curvature (see, e.g., [4, Proposition 9.70]):
(1.5) |
where is the O’Neill tensor of the submersion, while and are the scalar curvatures of and . The proof of this formula is a pointwise calculation and hence extends to the situation where is not compact (i.e., is not a manifold). In our case it can indeed happen that for some Lie algebra the connected subgroup with Lie algebra is not compact. However, this will not affect our discussions.
Let denote the normalizer of . Every element acts on as a diffeomorphism via right translation: . Combining this with the action by left translation, we obtain an action on the space of metrics via pullback: . This induces an isometry between and , and hence for the scalar curvature. This also holds for the possibly larger group of automorphisms of that preserve . Note though that is invariant under , or one of its subgroups, only if 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 is invariant under a subgroup , and if is the set metrics in invariant under , then critical points of the restriction are also critical points of . Indeed, given , the Ricci curvature and hence the gradient is invariant under . Consequently, must be tangent to .
In the remainder of the paper we will assume for simplicity that , 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 or is not connected, we consider only intermediate subalgebras that are Lie algebras of intermediates subgroups . This is easily seen to be equivalent to saying that must be invariant under . 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 and will reduce and may easily imply the existence of some critical points. This happens, for instance, when is isotropy irreducible; see [29] for many examples.
2. The simplicial complex
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 is given by (1.3), we set and obtain the following formulas for the scalar curvature and its constraint:
(2.1) |
We need to study the behavior of at infinity, which means that at least one of the variables goes to . It is natural to introduce a simplicial complex and its stratification. Specifically, let
Notice that the numbers , and thus the simplex , depend on the choice of . This simplex is a natural parametrization of the set . We identify a metric with .
The boundary of consists of lower-dimensional simplices. For every nonempty proper subset of the index set , let
Thus is a -dimensional simplex, which we call a stratum of . The closure of satisfies
and we call a stratum adjacent to if is a nonempty proper subset of . It will also be useful for us to consider tubular -neighborhoods of strata for :
Finally, we associate to each stratum an -invariant subspace of :
We can fill out the closure with geodesics starting at the center. To this end, consider the unit sphere
of dimension . Define a geodesic by setting , where
The stratification of induces one of the sphere:
Our first observation is that we can mark the strata with subalgebras.
Proposition 2.2.
The functional is bounded from above. Furthermore, for any either as or for some such that is a subalgebra of .
Proof.
Let , which is well-defined since by assumption. We also have since is not a torus. Then (2.1) implies that .
For the second claim, let be the index set with . Obviously, and . If is not a subalgebra, then there exist and such that . Then in formula (2.1), we have a contribution of the form
with . Since and , we know that stays bounded away from and as . This implies that as . ∎
We also need to control how fast the scalar curvature goes to . For this purpose we prove the following result.
Proposition 2.3.
Consider a stratum such that is not a subalgebra. Then for every and , there exist an open neighborhood in and a positive number such that whenever and .
Proof.
Let as before. There exist and such that . Moreover, and . Define
where . Evidently, this quantity is always positive. Choose a neighborhood of in such that and
for all . This implies
provided and . ∎
If the space has pairwise inequivalent isotropy summands, Proposition 2.3 implies the following result, originally proved in [25].
Corollary 2.4.
If is maximal in , then attains its global maximum at a metric , and hence for some .
We can add a marking to the strata in . If is a subalgebra, we denote the stratum by or simply ; if it is not, we denote the stratum by or .
Next, we need an estimate for near . Let be the connected subgroup of with Lie algebra . Define
where the letter is preserved for the decomposition of the tangent space to induced by . Since a normal homogeneous metric has non-negative scalar curvature, . Also, if and only if is flat. It is important for us to note that, since is not closed in general, the supremum in the definition of may not be achieved in , which will complicate our discussion.
Consider a metric in identified with . If for some , then
We may regard as a metric on . Its scalar curvature is given by
where the Killing form of restricted to equals . One easily shows that
Proposition 2.5.
Consider a stratum . If satisfies , then
Proof.
We break up the formula for the scalar curvature in (2.1) as follows, using the assumption that for and :
where in the last step we used the estimate . Now observe that
since and the trace of with respect to equals 1. Consequently,
When for all , we get the desired result. ∎
We can reformulate Proposition 2.5 as follows.
Corollary 2.6.
Let be a subalgebra stratum. Then for every there exists a constant such that the set does not intersect .
Corollary 2.7.
Suppose for every subalgebra stratum . Then is a (possibly empty) compact subset of .
Proposition 2.5 shows that is an upper bound for the possible values of the scalar curvature as we approach points in . However, it is important to keep in mind that does not, in general, extend continuously to the closure of .
We end this section with the following observation. Recall that even if and are connected and compact, and if is an intermediate subalgebra, then the connected (intermediate) subgroup with Lie algebra is not necessarily compact.
Lemma 2.8.
If is an intermediate subgroup with Lie algebra , then the closure is compact and , where is the Lie algebra of .
Proof.
Let with semisimple and the center of . Suppose that and are the connected Lie subgroups of with Lie algebras and . Then (i.e., is the quotient of by a discrete subgroup of the center of ) and is compact. The closure is compact abelian, and hence a torus. Denote its Lie algebra by . Thus . If , then any extension satisfies since . ∎
Thus it is sufficient to compute the invariants only for intermediate subalgebras for which is compact.
3. Riemannian submersions
In this section we study the behavior of near the subalgebra stratum geometrically. It will be more convenient to choose the path below instead of since we can then use formula (1.5) for Riemannian submersions. The goal is to see if there are metrics near whose scalar curvature is larger than . As before, suppose is an intermediate connected subgroup with Lie algebra and associated stratum . Thus for some . Define
We will show that controls the desired behavior.
We have the homogeneous fibration
(3.1) |
Let us consider metrics on for which this fibration is a Riemannian submersion. We start with a metric of the form
Assume that lies in , i.e., , where
We also require the metric to be -invariant so that the projection in (3.1) is a Riemannian submersion with and the metrics on the fiber and the base.
Consider the two-parameter family
Substituting , we obtain a one-parameter family of metrics
(3.2) |
lying in . We call this the canonical variation associated to . By (1.5), the scalar curvature of is
(3.3) |
Since , every point in is a limit of such a path . Thus we have
(3.4) |
Notice that
We now use these formulas to understand the relationship between the numbers corresponding to different strata.
Proposition 3.5.
If is a stratum adjacent to with , then .
Proof.
Let be the subgroup of with Lie algebra . Thus . Given with , define a one-parameter family of metrics
where the sum is taken over all with . Applying (3.4) to the homogeneous fibration
we conclude that
This means that, for every with , there exists a metric with and scalar curvature arbitrarily close to . ∎
As we noted in Section 3, it is possible that the supremum in the definition of is not attained by a metric in . In this case, we have the following result.
Proposition 3.6.
Assume that is not attained. Then there exists an adjacent stratum such that and for some with .
Proof.
Since is not attained, it is possible to find a sequence with converging to some , where the stratum is adjacent to . Applying Proposition 2.3 to the homogeneous space and using the nonnegativity of , we conclude that for some . Similarly, applying Proposition 2.5 to shows that
In light of Proposition 3.5, this means . If the supremum is attained, then we are done. Otherwise, we repeat the argument until we reach a subalgebra for which is achieved. By Corollary 2.4, this will be the case at the latest for a subalgebra in which is maximal. ∎
Finally, we show how the difference controls the behavior of the scalar curvature functional.
Proposition 3.7.
Consider a subalgebra stratum such that is attained. If , then there exists a metric , arbitrarily close to , with .
Proof.
Combining Propositions 3.6 and 3.7 implies our main theorem if there exists only one decomposition (up to order of summands).
Proposition 3.8.
Assume that is a compact homogeneous space such that the modules are inequivalent. Let be an intermediate subalgebra of the lowest possible dimension such that , where the supremum is taken over all intermediate subalgebras . If , then achieves its maximum at some metric , and hence for some .
It is natural to add an additional marking to the strata of by labeling a subalgebra stratum , which encodes the behavior of in a neighborhood of .
As remarked before, if is an intermediate subalgebra, then the connected subgroup with Lie algebra may not be compact. Thus, 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, is well-defined for any intermediate subalgebra.
4. Global maxima
From now on, we allow the decomposition to vary. Clearly, the numbers and the structure constants depend continuously on . Recall also that the space of all decompositions is compact.
Consider an intermediate subgroup with Lie algebra . The numbers and introduced above depend on the choice of the decomposition . Removing this dependence, we define
and introduce an invariant for given by
where the supremum is taken over all intermediate subalgebras .
Our first goal is to extend Propositions 2.3 and 2.5 to all of . We will use the following parametrization of the space , convenient in our context. Specifically, consider the map defined by
where and is the decomposition with irreducible modules . 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 are equivalent. However, given a metric , the preimage is compact.
To state our next result, we fix a point in the boundary of . Let be the simplex associated with the decomposition . Clearly, lies in a stratum for some . The following result generalises Proposition 2.5 to all of .
Proposition 4.1.
Assume that lies in a subalgebra stratum . Then for every there exists an open neighborhood of in such that
whenever .
Proof.
We use the notation , and (respectively, , and ) for the constants associated with the decomposition (respectively, ). Given , there exists a neighborhood of in such that
for all whenever . Let us choose small enough to ensure that in this set. The trace constraint implies that . Notice also that there exist common lower and upper bounds for independent of .
If , we find, as in the proof of Proposition 2.5, that
Consequently, for small enough , since when , we have
Shrinking further if necessary and using the continuity of the scalar curvature, we conclude that
which implies the result. ∎
Using a similar (but simpler) proof, we can generalize Proposition 2.3.
Proposition 4.2.
Assume that lies in a stratum . Given , there exists a neighborhood of in such that whenever .
The precompactness of and Propositions 4.1 and 4.2 yield the following extension of Corollary 2.7 to all of .
Corollary 4.3.
If , then is a compact subset of .
Our next result generalises Proposition 3.6.
Proposition 4.4.
Assume that is not maximal in . Then there exists an intermediate subgroup with Lie algebra such that . If has the least possible dimension of all such subgroups, then there exists with
Proof.
For each , choose a subgroup with Lie algebra such that . Assume that has the least possible dimension among all subgroups with this property. Let be a metric on with
We may assume that converge to an intermediate subalgebra and that for all . Our goal is to show that a subsequence of converges to a metric .
Consider a decomposition with modules such that
for some and all with . Passing to a subsequence if necessary, we may assume that these decompositions converge to some with modules and that does not depend on (thus we can omit the index from the notation ). Let be the connected subgroup of with Lie algebra . There exist positive numbers such that
Note that , where the supremum is taken over all intermediate subalgebras between and . Indeed, if not, there exists a subalgebra with , contradicting the assumption that is chosen to be of smallest possible dimension. Therefore,
for large . We now claim that the constants all lie in some compact subset of . To see this, we can argue as in Propositions 4.1 and 4.2 and Corollary 4.3 replacing with the sequence and using the fact that the structure constants of converge to those of .
Thus, passing to a subsequence if necessary, we may assume that
for . The metric satisfies and . ∎
We are now ready to prove our main theorem.
Proof of the main theorem.
The existence of the subgroup follows from Proposition 4.4. By Proposition 3.7, there exists such that the superlevel set is nonempty. Corollary 4.3 implies that this set is also compact. Consequently, assumes its maximum at a metric . As a critical point, such a metric satisfies for some constant . ∎
5. Examples
5.1. The Stiefel manifold .
This example is interesting since some of the modules 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, is two-dimensional. In addition, there exists a circle of intermediate subgroups and hence we need to maximize . This is the first example of this type where the existence of global maxima was studied.
Consider the homogeneous space where the circle group is embedded diagonally into . The bi-invariant metric we choose on is such that . Since the two-fold cover sends the diagonal embedding to , the space coincides with the Stiefel manifold .
We can identify and with the group of unit quaternions and the Lie algebra of purely imaginary quaternions. Then
Consider the following -invariant decomposition of :
(5.1) |
Under , the element takes and to and , respectively. This implies that the restrictions of to and are equivalent complex representations. Using the -orthonormal bases of , of , and of , we can thus represent the metric and the tensor field by the matrices
(5.2) |
which must be positive-definite. This is the case if and and similarly for . The outer automorphism of that switches the two factors preserves and hence induces an isometry of . It acts on by taking to , and similarly on . Conjugation by preserves as well. It acts on by taking to . Thus the composition takes to , which shows that we may assume . It will be convenient for us to denote
As we will see shortly, the maximal subgroups of containing are the subgroups and and a one-parameter family of three-dimensional subgroups , , isomorphic to . Our main theorem leads to the following result. We distinguish the case where the supremum 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 and . Then we have:
-
(a)
and if and only if
-
(b)
and for some if and only if
In both cases, attains its global maximum.
Proof.
We start by describing the space of all decompositions. Since the representations of on and are complex equivalent representations, is the two-parameter family of decompositions with modules
(5.4) |
where . Let us choose a -orthonormal basis in consisting of the vectors
(5.5) |
and one in consisting of
(5.6) |
Using these bases, for the decomposition , one easily computes
(5.7) |
The structure constants unrelated to these by permutation are 0.
The decomposition corresponding to gives rise to the natural set of diagonal metrics with intermediate subgroups
with Lie algebras
If , we have, for each , the intermediate subgroup
The remaining decompositions do not produce any subgroups. Thus, there are three isolated intermediate subgroups, , and , as well as a one-parameter family of subgroups .
The identity component of the normalizer of is given by and hence , represented by elements of the form . These elements act via right translation on and via conjugation on . Thus they also act on and, via pullback, on . It is easy to see that takes to . This implies, in particular, that the subalgebras are all conjugate to each other by elements of . Since acts on by isometries in , it follows that , as well as , are all isometric to each other.
We now compute the constants and for the intermediate subalgebras. For the maximal subgroups and we have
where is a permutation of . Indeed, are isotropy irreducible, which easily determines . The space has two irreducible summands, and the scalar curvature, under the trace constraint, is
The above value is its maximum. The assumption implies that . Furthermore, if and only if
(5.8) |
For the three-dimensional subalgebras , we use the bases (5.5) and (5.6). Now, the parameters and both equal . The tensor field satisfies
(5.9) |
and . The action of the quotient establishes an isometry between and the space , which one easily sees has scalar curvature . The trace constraint means that , implying
The action of the normalizer shows that are all isometric to the symmetric space . Thus, they have scalar curvature 12. The trace constraint for -invariant metrics on takes the form
Thus
We now choose an angle such that is maximal, i.e., . Observe that
for some phase shift . The largest possible value of this quantity is , which means
Clearly, if and only if
(5.10) |
Remark 5.11.
For the set of diagonal metrics one easily solves the equation directly. This equation reduces to the system
(5.12) |
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 fixes the diagonal metrics and acts by isometries. Thus the Ricci curvature of a diagonal metric must be diagonal as well. In particular, if , then every solution to the system (5.12) is a critical point of on all of . In Figure 1 these metrics lie in the union of the 3 grey regions. On the other hand, given a critical point of with and , one obtains a circle of further critical points by applying the normalizer.
For a metric as in (5.2), the constraint becomes
and the scalar curvature satisfies
where ; see, e.g., [5]. We may assume that . If , we show in Figure 1 points in the -plane that correspond to various behaviors of . 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 yields a global maximum, and in the middle-grey region . 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 , the picture is the same but with the line and the curve shifted to the right and the reflection of shifted up.

Next, we give several examples that illustrate various behaviors.
Example 1.
Suppose that , and . A straightforward computation shows that the diagonal metrics with
(5.13) |
solve the system (5.12) and hence are critical points of the functional 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 . Maple shows that all four eigenvalues of the Hessian are negative for , 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 , 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 can, in fact, be saddles on all of .
Example 2.
From the previous example we see that the metric with must be special. Thus we choose , and . This choice of is marked by the red dot in Figure 1. Direct verification shows that the metrics with
(5.14) |
are critical points of for and with scalar curvature . They form a surface diffeomorphic to , described in the coordinates . The normalizer acts on this surface via . Thus metrics with the same value of are isometric. On the other hand, the squared volume of the metric (5.14) equals , and hence metrics with different values of are not isometric. To determine the critical point type of (5.14), we compute the eigenvalues of the Hessian of . 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 near would have to have scalar curvature larger than .
Example 3.
Assume that . As explained below, has a non-diagonal critical point if and only if lies in the pink region in Figure 2. Since the action of leaves unchanged, we obtain a circle of non-diagonal critical points for each such . It is always a non-degenerate critical submanifold of index 2 and co-index 1. In addition, we have a diagonal critical point for between the thick curves. By Proposition 5.3, this diagonal critical point must be a global maximum when 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 in the rest of the dotted region and a saddle with index and co-index 2 for in the yellow region. The transition from the pink to the yellow region is achieved across the curve
(5.15) |
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 having a non-diagonal critical point for in the pink region, we first observe that is diagonal if and only if or . This is because the off-diagonal entries in the Ricci curvature are
Since the action of leaves diagonal Ricci curvature unchanged, it suffices to consider only metrics with . We take two million non-diagonal satisfying and and calculate . After checking that is positive-definite and normalizing so that its component becomes 1, we mark the and components in the -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 with . 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 , and , the green dot in Figure 2. Direct verification shows that the gradient of vanishes at the circle of metrics with , , and . These metrics are all saddles with index 2 and co-index 1. At the same time, the diagonal metric with
is a strict local maximum. It cannot be a global maximum because its scalar curvature is less than 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.

As far as non-diagonal Ricci candidates are concerned, we make the following remarks. Recall that supports two circles of (non-diagonal) Einstein metrics isometric to the canonical product Einstein metric on ; 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 , each one of which admits two arcs of critical points.
The Stiefel manifold is also a generalized Wallach space under the decomposition with and any . These spaces we will study in detail in [27] in order to produce saddle critical points. Using the methods of [27], one shows that admits a critical point of co-index 0 or 1 if
When , this should be interpreted as , 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 is a non-degenerate critical point of if and only if . Assuming and , a Maple computation shows that this only fails for three families of metrics. The first one satisfies , and these are the metrics with diagonal Ricci tensor in Example 3. The second one is given by with , and for these the Ricci tensor is not diagonal. The third family consists of the metrics for which
5.2. The Ledger–Obata space and
Our second example is motivated by the following observations. Let the homogeneous space be such that with both summands irreducible. When and are inequivalent, one can solve the equation directly; see [25]. There are only two cases where and are equivalent representations. If is simple, then must equal ; see the classification in [13, 18]. If it is not, one easily checks that the only possibility is that is the Ledger–Obata space , where is simple and embedded diagonally.
We discuss here the latter case, the former one being quite similar; see Remark 5.22. In what follows, is a basis of orthonormal with respect to , the negative of the Killing form of .
Let . There are exactly three intermediate subgroups, namely,
The outer automorphism given by interchanges these subgroups. Choose the bi-invariant metric on to be . Then is an isometry of .
Fix an -invariant, -orthogonal decomposition of by setting
(5.16) |
Then . Furthermore, is an irreducible representation of real type on each since is simple. The collections and constitute -orthonormal bases of and . With respect to these bases, the metric and have the form
(5.17) |
with and . (Each entry in these matrices represents a scalar matrix, where .) Applying our main theorem, we obtain the following result.
Proposition 5.18.
Suppose that and . The functional attains its global maximum if
(5.19) |
Proof.
We need to compute and . A well-known formula for the sectional curvature of a normal homogeneous metric (see [4, Proposition 7.87b]) implies
where is the norm corresponding to . Similarly,
In addition, is an isotropy irreducible symmetric space, and hence . It follows from (1.4) that
with . The trace constraints on and have the form and , respectively. We conclude that
which means if and only if .
As observed above, the automorphism takes to . Since is the identity, the matrix of the pullback of by with respect to our fixed bases of and is
The maps from to and from to induced by are isometries. They preserve the scalar curvature, so
Consequently, if and only if .
To obtain the third subgroup, we need to apply to , or to . The matrix of the pullback of by with respect to (5.16) is
This implies
Thus if and only if .
There are two scenarios for the condition of our main theorem to be satisfied. One is that (equivalently, ), in which case it suffices to demand that . The other is that . Then we need . Combining these conditions, we arrive at (5.19). ∎
For and of the form (5.17), the constraint is
and a computation shows that the scalar curvature is given by
(5.20) |
Thus we have the same critical points, up to scaling, no matter what group we choose. The formulas
define coordinates in the space of Ricci candidates normalized so that . Figure 3 shows points in the -plane that correspond to different behaviors of . The tensor is positive-definite if and only if . The order-three automorphism yields a natural symmetry on the space of Ricci candidates. Its action rotates the picture in Figure 3 by . The inside of the large triangle is the set of that satisfy the condition of Proposition 5.18. The equalities , and hold in the dark-grey, the middle grey and the light-grey triangle, respectively. A computer-assisted experiment with one million metrics shows that also has a critical point for 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 in the white region has no critical points.
Maple is able to solve the Euler–Lagrange equations for for any specific choice of . It suggests that the solution representing a metric, when it exists, is unique and is always a local maximum (global if lies inside the triangle).

Example 4.
Assume that , and . One easily checks that the formulas
(5.21) |
define a metric with positive Ricci curvature for , and this metric is a critical point of . Computing the Hessian, we conclude that it is a strict local maximum unless . In Figure 3 the Ricci candidates with , and occupy the segment of the -axis with . The orbit of this segment under , indicated by the green lines, gives two further segments with the same behavior.
The case of , respectively , is special. The orbit of the corresponding tensor under is depicted by the red dots in Figure 3. The three tensors in this orbit are, in fact, Einstein metrics. The -equivariant diffeomorphisms from to given by
are isometries between these Einstein metrics and the canonical Einstein product metric on . 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 and hence consist of local maxima.
The only other Einstein metric on is the normal homogeneous metric induced by ; see [16]. It corresponds to the origin in Figure 3 and is a strict local maximum of the associated functional .
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 are Morse or Morse–Bott functions.
Remark 5.22.
The remaining homogeneous space with two equivalent summands is . It is, in fact, quite similar since the scalar curvature is, remarkably, given by the same formula as (5.20) with ; see [19]. There are three intermediate subgroups isomorphic to , which are permuted by the triality automorphism of . One easily sees that the conditions for a global maximum on 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 , 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.