The prescribed Ricci curvature problem for naturally reductive metrics on non-compact simple Lie groups
Abstract
We investigate the prescribed Ricci curvature problem in the class of left-invariant naturally reductive Riemannian metrics on a non-compact simple Lie group. We obtain a number of conditions for the solvability of the underlying equations and discuss several examples.
1 Introduction
The study of the prescribed Ricci curvature problem is an important part of modern geometry with ties to flows, relativity and other subjects. The first wave of interest in this problem came in the 1980s; see [Bes87, Chapter 5] and [Aub98, Section 6.5]. Particularly extensive contributions were made at that time by DeTurck and his collaborators. For a discussion of the subsequent advances, including the recent progress in the framework on homogeneous spaces, see the survey [BP19].
Let be a smooth manifold. In its original interpretation, the prescribed Ricci curvature problem comes down to the equation
(1.1) |
where the Riemannian metric on is the unknown and the (0,2)-tensor field is given. The paper [DeT81] proved, for nondegenerate , the existence of satisfying this equation in a neighbourhood of a point on ; see also [DG99, Pul13, Pul16a]. However, subsequent research into the solvability of (1.1) on all of revealed the need for a more nuanced interpretation of the prescribed Ricci curvature problem. Specifically, suppose is closed and is positive-definite. The results of [Ham84, DeT85, Delan03] and other papers suggest that there exists at most one such that the equation
(1.2) |
can be solved for on all of . This is certainly the case if is the 2-dimensional sphere and is positive-definite; see [WW70, DeT85, Ham84]. Thus, on a closed manifold, one customarily interprets the prescribed Ricci curvature problem as the question of finding and such that (1.2) holds. This paradigm was originally proposed by DeTurck and Hamilton in [Ham84, DeT85]. As it turns out, (1.2) arises in applications, such as the construction of the Ricci iteration; see [PR19, BPRZ19] and also [BP19, Section 3.10]. On the other hand, if is open, it may be possible to obtain compelling existence theorems for (1.1) without the additional constant . We refer to [Delay02, Delay18] for examples of such theorems.
In recent years, the third-named author and his collaborators produced a series of results, surveyed in [BP19], on the prescribed Ricci curvature problem in the class of homogeneous metrics. More precisely, suppose is a connected Lie group. Let be a homogeneous space with respect to . Assume that the metric and the tensor field are -invariant. Then (1.1) reduces to an overdetermined system of algebraic equations, whereas (1.2) reduces to a determined one. For compact and positive-semidefinite , the third-named author showed in [Pul16b] that homogeneous metrics satisfying (1.2) are, up to scaling, critical points of the scalar curvature functional subject to the constraint . This observation led to the discovery of several sufficient conditions for the solvability of (1.2) in [Pul16b, GP17, Pul20]. It parallels the well-known variational approach to the Einstein equation; see, e.g., [WZ86, §1]. In the case where is compact but is not positive-semidefinite, the question of solvability of (1.2) remains largely open. We hope that the present paper will stimulate its investigation; see Remark 5.1.
As for the prescribed Ricci curvature problem for homogeneous metrics on non-compact spaces, progress has been scarce so far. Buttsworth conducted in [But19] a comprehensive study of this problem on unimodular three-dimensional Lie groups. In most of the cases he considered, there is at most one constant such that a metric satisfying (1.2) exists. Several questions related to, but distinct from, the solvability of (1.1) and (1.2) on non-compact Lie groups have been studied by Milnor, Kowalski–Nikcevic, Eberlein, Kremlev–Nikonorov, Ha–Lee, Pina–dos Santos, the first-named author in collaboration with Lafuente (forthcoming work), and others. For a discussion of those results and a collection of references, see [BP19, Sections 2 and 4.1].
Left-invariant naturally reductive metrics on a Lie group form an important family nested between the set of all left-invariant metrics and the set of bi-invariant ones. The investigation of this family has led to several significant advances in geometry. For instance, it yielded new solutions to the Einstein equation and new insights into the spectral theory of the Laplacian on manifolds; see [DZ79, GS10, Lau19]. In the recent paper [APZ20], Ziller and two of the authors studied (1.2) for naturally reductive metrics on a compact Lie group using variational methods. That work exposed several interesting and previously unseen patterns of behaviour of the scalar curvature functional. For instance, one of the main theorems of [APZ20] is a necessary condition for the existence of a critical point subject to the constraint . No results of this kind had appeared in the literature before.
The present paper studies the prescribed Ricci curvature problem for naturally reductive metrics on a non-compact Lie group . We assume that is simple. The more general case of semisimple seems to be much more difficult analytically—we intend to consider it elsewhere. Since is an open manifold, it is reasonable for us to view the prescribed Ricci curvature problem as the question of finding solutions to (1.1). On the other hand, the fact that naturally reductive metrics are homogeneous suggests that a “better” interpretation of this problem may be given by (1.2). The present paper studies both equations. We show that (1.1) reduces to an overdetermined algebraic system. Even so, we are able to obtain a comprehensive existence and uniqueness theorem. Equation (1.2) reduces to a determined system. In order to find conditions for solvability, we characterise metrics satisfying (1.2) as critical points of the scalar curvature functional subject to one of three -dependent constraints. While this characterisation is similar in spirit to the one obtained for compact Lie groups in [APZ20], it bears some conceptual distinctions and requires a different proof. We obtain existence theorems for global maxima and classify some of the other critical points. The development and application of our variational methods presents many interesting analytical challenges and provides a wealth of insight for the investigation of (1.2) on compact homogeneous spaces in the case of mixed-signature (see, e.g., Remark 5.1).
The paper is organised as follows. In Section 2, we recall the characterisation, originally obtained by Gordon, of naturally reductive metrics on a non-compact simple Lie group. This characterisation underpins all of our results. In Section 3, we compute the Ricci curvature of a naturally reductive metric on . We believe that the formulas we obtain are of independent interest. Section 4 is devoted to equation (1.1). We produce a necessary and sufficient condition for the existence of a solution. We also establish uniqueness up to scaling. Section 5 focuses on (1.2). We develop the variational approach to this equation and describe several types of critical points of the scalar curvature functional. At the end, we summarise the implications for the existence and the number of solutions. Section 6 examines the case where the metrics we consider are naturally reductive with respect to for a simple subgroup . Here, we find conditions for the solvability of (1.2) that are both necessary and sufficient. We also determine the precise number of solutions. Finally, Section 7 offers a series of examples.
2 Naturally reductive metrics on non-compact simple Lie groups
Consider a connected non-compact simple Lie group with Lie algebra . The results of [Gor85, Section 5] yield a convenient characterisation of left-invariant naturally reductive metrics on . We present this characterisation in Theorem 2.1 below. For the basic theory of naturally reductive metrics, see [DZ79, Section 1] and [Gor85, Section 2]. In what follows, we identify every left-invariant (0,2)-tensor field on with the bilinear form it induces on .
Let be a maximal compact subgroup of with Lie algebra . Suppose is the Killing form of . Denote by the -orthogonal complement of in . Then
is a Cartan decomposition. We have the inclusions
The Killing form is positive-definite on and negative-definite on . Thus,
is an inner product on . Clearly, is -invariant, and
(2.1) |
The quotient is a symmetric space. Because is simple, this space is irreducible. Consequently, the pair must appear in Table 3 or 4 of [Bes87, Section 7.H]. Let be the simple ideals of . Denote by the centre of . Then
(2.2) |
where if is trivial and otherwise. Analysing the tables in [Bes87, Section 7.H], we conclude that is at most 1-dimensional.
The direct product acts on in accordance with the formula
The isotropy subgroup at the identity element of is
Denote by the set of left-invariant metrics on that are naturally reductive with respect to and some decomposition of the Lie algebra of . The main purpose of this paper is to study the prescribed Ricci curvature problem in . Gordon showed in [Gor85, Section 5] that every left-invariant naturally reductive metric on must lie in for some choice of . Moreover, she obtained the following characterisation result.
Theorem 2.1 (Gordon).
A left-invariant metric on the simple group lies in if and only if
(2.3) |
for some .
3 The Ricci curvature
Our main objective in this section is to produce formulas for the Ricci curvature and the scalar curvature of a metric given by (2.3). To do so, we need to introduce an array of constants, , associated with the pair . Throughout the paper,
As we explained above, if the centre of is non-trivial.
Suppose is the Killing form of for . There exists such that
Using the assumption that is simple, one can easily check that for . If the centre of is non-trivial, then .
Next, we state a proposition that provides a way of computing for a specific pair and . In what follows, superscript means complexification. Clearly, the algebra is semisimple. We preserve the notation for the adjoint representation of . Choose Cartan subalgebras in and in . It is easy to verify that is a completely reducible -module under the action given by . This observation implies that every element of must be semisimple in . Consequently, we may assume that contains . Let and be sets of positive roots of and . The notation stands for the trace of a linear operator.
Proposition 3.1.
Given and , the constant satisfies
Proof.
We preserve the notation and for the Killing forms of and . Because is compact, is a simple subalgebra of . Therefore,
Using basic properties of root systems, we find
∎
Remark 3.2.
The assertion of Proposition 3.1 holds even if is not simple but merely semisimple.
Remark 3.3.
One can use properties of Casimir elements to produce another formula for . More precisely, given , there exists a decomposition
such that every is a non-trivial irreducible -module under the action defined by . Let be the highest weight of . Denote by the half-sum of positive roots of . Using classical results on eigenvalues of Casimir elements, one can show that
where we preserve the notation for the bilinear form induced on by the Killing form of . Related formulas can be found in Dynkin’s work; see [Dyn57].
Example 3.4.
Assume and with . Then and in the decomposition (2.2). Clearly,
Denote by the matrix of size that has 1 in the th slot and 0 elsewhere. Suppose
where is the linear functional on such that is the difference of Kronecker deltas . Choosing , we find
Proposition 3.1 implies . A similar argument with yields . Since is abelian, .
Remark 3.5.
We are now ready to state the main result of this section.
Theorem 3.6.
Suppose is a naturally reductive metric on the simple group satisfying (2.3). The Ricci curvature of is given by the formulas
To prove Theorem 3.6, we apply the strategy developed in [DZ79, Section 5]. In what follows, stands for the trace of a bilinear form with respect to an inner product . The notation is used for the -orthogonal projection onto . For , define a bilinear form on by setting
Fix -orthonormal bases of and of . We need the following auxiliary result; cf. [Jen73, pages 609–610] and [DZ79, pages 32–34].
Lemma 3.7.
Given ,
Proof.
Invoking (2.1), we compute
Since is an ideal of ,
We conclude that
This proves the first equality in the statement of the lemma.
For ,
It is easy to see that the forms are symmetric. Consequently,
We conclude that
which implies the second inequality. ∎
Proof of Theorem 3.6.
Let be the Levi-Civita connection of the metric . The Koszul formula yields
(3.1) |
see [Gor85, page 485]. The Ricci curvature of satisfies
(3.2) |
This fact goes back to [Sag70]; a simpler proof appeared in [DZ79, Section 5]. Substituting (3.1) into (3.2), we easily obtain the required identities for , and with ; cf. [DZ79, page 33].
Because lies in , it is -invariant. Consequently, there exists such that . Taking the trace with respect to on both sides and exploiting Lemma 3.7, we find
Therefore,
which yields the required identity for . ∎
Denote by the scalar curvature functional on . Our next goal is to produce a formula for . Taking the trace of the second equality in Lemma 3.7 and using the first one, we obtain
(3.3) |
Corollary 3.8.
Suppose is a naturally reductive metric on the simple group satisfying (2.3). The scalar curvature of is given by the formula
4 Metrics with prescribed Ricci curvature
Consider a (0,2)-tensor field on the simple group . In this section, we state a necessary and sufficient condition for the solvability of the equation
(4.1) |
in the class . If a metric satisfying (4.1) exists, then must be left-invariant. Moreover, by Theorem 3.6, the formula
(4.2) |
holds with .
Theorem 4.1.
Proof.
It is tempting to use Theorem 4.1 to study the solvability of the equation
(4.5) |
in the class . Indeed, suppose is the set of left-invariant tensor fields on satisfying (4.2), (4.3) and (4.4). Theorem 4.1 states that (4.1) has a solution if and only if lies in . Using this result, we can easily obtain a description of the set of tensor fields that coincide with Ricci curvatures of metrics in up to scaling. Namely, a pair satisfying (4.5) exists if and only if
However, in general, it is difficult to determine whether a specific given by (4.2) lies in . To do so, one has to check whether (4.3) and (4.4) hold with replaced by for some . Already when , this involves the tricky task of understanding if a polynomial of degree has roots that obey several constraints; when , the question seems to be substantially harder. In the present paper, we take a different approach to the analysis of (4.5). We are able to show that the existence of a pair satisfying (4.5) follows from simple inequalities for the components of . Moreover, we draw interesting conclusions regarding the non-uniqueness of such a pair.
5 Metrics with Ricci curvature prescribed up to scaling
Suppose is a left-invariant symmetric (0,2)-tensor field on . Our next goal is to study the solvability of equation (4.5) in the class . As above, we assume the group is simple. This implies, in particular, that for all .
First, we re-state the problem in variational terms. More precisely, define
(5.1) |
Each of these three spaces carries a manifold structure induced from . In Section 5.1, we show that satisfies (4.5) if and only if it is (up to scaling) a critical point of , or . This resembles the variational interpretation of (4.5) on compact Lie groups for positive-semidefinite (see [APZ20, Proposition 3.1]); however, in that case, only needs to be considered. In Sections 5.2 and 5.3, we obtain sufficient conditions for the existence of global maxima of and , respectively. This requires complex estimates on the scalar curvature obtained in Lemmas 5.4 and 5.6. In Section 5.4, we classify completely the critical points of assuming . The analysis here can be reduced, as Lemma 5.9 demonstrates, to the study of a cubic polynomial in one variable. Finally, in Section 5.5, we summarise the implications of our results for the prescribed Ricci curvature problem.
Remark 5.1.
It appears that (4.5) admits a similar variational interpretation on compact homogeneous spaces when has mixed signature. Thus, our arguments yield new insight into the study of (4.5) in that setting. For instance, Buttsworth showed in [But19] through methods of elementary polynomial analysis that, for certain on , a left-invariant metric satisfying (4.5) exists for precisely two distinct constants . This was somewhat surprising at the time, as nothing similar had occurred in previously understood examples. We observe an analogous phenomenon in Theorem 5.10 below. In our arguments, the two constants arise naturally as Lagrange multipliers for and .
5.1 The variational approach
The following result underpins our approach to the study of (4.5).
Proposition 5.2.
Let be a left-invariant symmetric (0,2)-tensor field on . A metric satisfies (4.5) for some if and only if it is (up to scaling) a critical point of , or .
Proof.
Denote by the space of left-invariant bilinear form fields
with . Theorem 3.6 shows that lies in . We identify with the space tangent to at in the natural way. The left-invariant bilinear form fields
where denotes pullback, make a basis of .
5.2 Global maxima on
Our goal in this subsection is to show that simple inequalities for guarantee the existence of a critical point of .
Theorem 5.3.
Suppose is a left-invariant (0,2)-tensor field on satisfying (4.2) for some . Choose an index such that
(5.3) |
If
(5.4) |
and
(5.5) |
then the functional attains its global maximum.
The proof of Theorem 5.3 requires the following estimate for .
Lemma 5.4.
Proof.
Consider a metric satisfying (2.3). The definition of implies
(5.7) |
For , denote
In view of Corollary 3.8 and formula (5.7), if , then
Thus, in this case, estimate (5.6) holds.
Denote
Formula (5.7) implies
Invoking Corollary 3.8 again, we find
In view of (5.5), if , then
In this case, again, estimate (5.6) holds.
With Lemma 5.4 at hand, we can prove Theorem 5.3 using the approach from [APZ20, Proof of Theorem 3.3]. The main idea behind this approach goes back to [GP17].
Proof of Theorem 5.3.
Denote . For , consider the metric satisfying
Straightforward verification shows that lies in . By Corollary 3.8,
Furthermore, in light of (5.4),
(5.10) |
We conclude that for sufficiently large , which implies the existence of such that
Using Lemma 5.4 with
yields
(5.11) |
Since is compact, the functional attains its global maximum at some . Obviously, lies in . Therefore, by (5.11), for all . ∎
5.3 Global maxima on
Now we focus on the space .
Theorem 5.5.
The proof relies on the following estimate.
Lemma 5.6.
Assume that (5.5) holds. Given , there exists a compact set such that for every .
Proof.
Let be a metric satisfying (2.3). Then
(5.12) |
which implies . Moreover,
Given , denote
Corollary 3.8 implies
By (5.5), if , then .
Choose as in (5.9). For , denote
As shown above, if , then . Recalling that and assuming that and , we obtain
Thus, the inequality implies .
Let be the set of those that satisfy (2.3) with
This set is compact. By the arguments above, whenever lies in . ∎
5.4 Critical points on
If in formula (2.2), then
(5.13) |
for some . In this case, straightforward analysis shows that has no critical points unless
(5.14) |
On the other hand, when (5.14) holds, the scalar curvature of every metric in equals 0. If , we are able to obtain a complete classification of the critical points of . We present this classification in Theorem 5.8 below. While its statement is quite bulky, its conditions are easy to verify once the tensor field and the geometric parameters of and are given. According to Table 3 in [Bes87, Section 7.H], the sum can be greater than 2 only if is one of the pairs
It seems difficult to classify the critical points of in these cases without using software, such as Maple, to solve the Euler–Lagrange equations numerically. Nevertheless, for all values of , the following result holds.
Proposition 5.7.
If is a critical point of , then .
Proof.
Assume that in formula (2.2). For the list of satisfying this assumption, see Table 3 in [Bes87, Section 7.H]. Equality (4.2) becomes
(5.15) |
It will be convenient for us to denote
The proof of Theorem 5.8 below shows that the variational properties of are largely determined by those of the polynomial
The discriminant of this polynomial is
Denote
According to the classical theory of cubic equations (see, e.g., [Jan10] for a modern interpretation), if and , then is a triple root of . It is also a saddle point. If and , then and are a double root and a simple root of , respectively. Both are local extremum points.
Theorem 5.8.
Assume that in formula (2.2). The scalar curvature functional does not have a global minimum. Critical points of other types exist under the following conditions:
-
1.
A saddle if and only if
(5.16) -
2.
A global maximum if and only if
(5.17) -
3.
A local maximum that is not a global maximum if and only if
(5.18) -
4.
A local minimum if and only if
(5.19)
When it exists, the critical point of is unique up to scaling.
Let us make a few remarks in preparation for the proof. Consider a metric . There are such that
(5.20) |
The equality implies
By Corollary 3.8,
(5.21) |
Note that the factor in front of is necessarily negative. Our arguments will involve two curves, and , in the space given by the formulas
Both these curves pass through at .
Lemma 5.9.
The metric given by (5.20) is a critical point of if and only if is a multiple root of .
Proof.
Assume is a critical point of . Proposition 5.7 implies . In light of (5.4), this means must be a root of . Furthermore, because is a critical point of ,
Thus, the derivative of at vanishes. This proves the “only if” part of the claim.
Assume that is a multiple root of . We need to show that is a critical point of . Clearly, the vectors tangent to the curves and at are linearly independent. Therefore, it suffices to prove that
Computing as above, we find
Formula (5.4) implies
∎
Proof of Theorem 5.8.
Recalling that is a cubic polynomial, we find
Consequently, never attains its global minimum. This proves the first statement.
Suppose is a saddle point of . Proposition 5.7 implies . Moreover, every neighbourhood of in contains a metric with negative scalar curvature and one with positive scalar curvature. By Lemma 5.9, is a multiple root of . Formula (5.4) shows that every interval around contains a point where is positive and one where is negative. This is only possible if is a triple root. By the classical theory of cubic equations, conditions (5.16) hold. Conversely, these conditions ensure that has a triple root at . Consider a metric defined by
Lemma 5.9 implies that is a critical point of . Using (5.4), one easily shows that every neighbourhood of contains a metric with negative scalar curvature and one with positive scalar curvature. In light of Proposition 5.7, this means is a saddle point.
The functional attains its global maximum if and only if for some and for all . Formula (5.4) implies that this happens if and only if has a double root in the interval and is nonnegative on this interval. Conditions (5.17) are necessary and sufficient for to have such properties.
Next, has a local maximum that is not a global maximum if and only if for some , for all in a neighbourhood of , and the scalar curvature of at least one metric in is positive. This is equivalent to having a simple root in the interval and a double root in . Conditions (5.18) are necessary and sufficient for to have such properties.
Analogously, has a local minimum if and only if has a double root in and is nonpositive in a neighbourhood of this root. Conditions (5.19) are necessary and sufficient for this.
Finally, in view of Lemma 5.9, can have at most one critical point up to scaling since a cubic polynomial can have at most one multiple root. ∎
5.5 Summary
The results of Sections 5.1–5.4 enable us to make several conclusions about the solvability of (4.5). We summarise these conclusions in Theorem 5.10 below. The constant in (4.5) must be positive if satisfies (4.2). This is an immediate consequence of the formulas for the Ricci curvature obtained in Section 3.
Theorem 5.10.
Suppose is a left-invariant (0,2)-tensor field on given by (4.2).
- 1.
- 2.
- 3.
Proof.
Statements 1 and 3 follow from Theorems 5.5 and 5.8 combined with Proposition 5.2. Next, assume that (5.4) and (5.5) hold. According to Theorems 5.3 and 5.5, the functionals and attain their global maxima at some and . Proposition 5.2 implies that both and have Ricci curvature equal to up to scaling. These metrics cannot be homothetic because and are not of the same sign. ∎
When , Theorem 5.10 is essentially optimal. We explain this in detail in Remark 6.2. At the same time, when , it seems that (5.4) and (5.5) may fail to hold even if and attain their global maxima. Indeed, on compact Lie groups, inequalities that are similar in spirit to these provide merely a “linear approximation” to the necessary and sufficient conditions for the existence of a critical point; see [APZ20, Section 5].
6 The case where is simple
As above, let be a left-invariant (0,2)-tensor field on . Assume that is simple. Our next result settles the question of solvability of (4.5) under this assumption. We do not use the variational approach developed in Section 5.1; however, see Remarks 6.2 and 6.3 below.
Since is simple, the numbers and in (2.2) equal 1 and 0, respectively. By Theorem 3.6, if (5.13) holds for some , the constant in (4.5) must be positive.
Proposition 6.1.
Proof.
Choose a metric . There exist such that
Theorem 3.6 and formula (3.3) imply that satisfies (4.5) if and only if
where . Clearing from the second line and substituting into the first, we obtain
This is a quadratic equation with discriminant
It has no solutions if , precisely one positive solution if or , and precisely two positive solutions if . Together with (3.3), this implies the result. ∎
Remark 6.2.
Proposition 6.1 shows that Theorem 5.10 is essentially optimal in our current setting. Indeed, since is simple, (5.5) becomes
In view of (3.3), this is equivalent to
Theorem 5.10 and Remark 5.11 assert that (5.5) and (5.14) are sufficient conditions for the solvability of (4.5). Conversely, as Proposition 6.1 shows, the existence of a pair satisfying (4.5) implies that either (5.5) or (5.14) must hold. Theorem 5.10 provides lower bounds on the number of solutions to (4.5). Using Proposition 6.1, one can easily demonstrate that these bounds are sharp.
7 Examples
Example 7.1.
Assume and . Then
Formula (3.3) yields . Suppose is given by (5.13) with . Since is simple, Proposition 6.1 applies. Formula (6.1) becomes
Equivalently,
If this holds, then (4.5) has no solutions. Similarly, if
then there is one pair , up to scaling of , that satisfies (4.5). If
there are two such pairs.
Example 7.2.
Assume and with . Then
Using (3.3), we find
Suppose is given by (5.15) with . Inequality (5.5) becomes
(7.1) |
According to Theorem 5.5, if (7.1) holds, then attains its global maximum. In this case, there exists a metric with Ricci curvature for some . Inequality (5.4) takes the form
(7.2) |
The set of triples for which both (7.1) and (7.2) are satisfied is non-empty and open in . We depict it in Figure 1 for . We also indicate where (7.1) holds without (7.2). Because these inequalities are invariant under scaling of , we make our sketch assuming . By Theorem 5.3, if both (7.1) and (7.2) are satisfied, then attains its global maximum. In this case, there exists a metric with Ricci curvature for some .

Theorem 5.8 enables us to classify the critical points of . For instance, suppose and . Then the discriminant of the polynomial satisfies
By Theorem 5.8, has no critical points. To give another example, suppose and , where is the unique positive root of the polynomial
Then . Moreover, using an approximate value of calculated in Maple, we find and
This means attains its global maximum.
One can use Maple to produce the graphs of , and ; cf. [APZ20, Section 5].
Example 7.3.
Assume and with . Then
As we showed in Example 3.4,
Suppose
for some . Inequality (5.5) becomes
(7.3) |
By Theorem 5.5, if (7.3) holds, then attains its global maximum. This is the case, e.g., when
Assume for simplicity that . Then (5.4) takes the form
(7.4) |
By Theorem 5.5, if both (7.3) and this inequality hold, then attains its global maximum. This happens, e.g., for and .
Acknowledgements
The authors are grateful to Jorge Lauret and Cynthia Will for their careful reading of the paper and useful comments.
References
- [APZ20] R. Arroyo, A. Pulemotov, W. Ziller, The prescribed Ricci curvature problem for naturally reductive metrics on compact Lie groups, arXiv:2001.09441 [math.DG], 2020, submitted.
- [Aub98] T. Aubin, Some nonlinear problems in Riemannian geometry, Springer-Verlag, Berlin, 1998.
- [Bes87] A.L. Besse, Einstein manifolds, Springer-Verlag, Berlin, 1987.
- [But19] T. Buttsworth, The prescribed Ricci curvature problem on three-dimensional unimodular Lie groups, Math. Nachr. 292 (2019) 747–759.
- [BP19] T. Buttsworth, A. Pulemotov, The prescribed Ricci curvature problem for homogeneous metrics, arXiv:1911.08214 [math.DG], 2019, to appear in: O. Dearricott, W. Tuschmann, Y. Nikolayevsky, T. Leistner, D. Crowley (Eds), Differential geometry in the large, Cambridge University Press.
- [BPRZ19] T. Buttsworth, A. Pulemotov, Y.A. Rubinstein, W. Ziller, On the Ricci iteration for homogeneous metrics on spheres and projective spaces, arXiv:1811.01724 [math.DG], 2018, to appear in Transf. Groups.
- [DZ79] J. D’Atri and W. Ziller, Naturally reductive metrics and Einstein metrics on compact Lie groups, Mem. Amer. Math. Soc. 215 (1979).
- [Delan03] Ph. Delanoë, Local solvability of elliptic, and curvature, equations on compact manifolds, J. Reine Angew. Math. 558 (2003) 23–45.
- [Delay02] E. Delay, Studies of some curvature operators in a neighborhood of an asymptotically hyperbolic Einstein manifold, Adv. Math. 168 (2002) 213–224.
- [Delay18] E. Delay, Inversion of some curvature operators near a parallel Ricci metric II: Non-compact manifold with bounded geometry, Ark. Mat. 56 (2018) 285–297.
- [DeT81] D.M. DeTurck, Existence of metrics with prescribed Ricci curvature: local theory, Invent. Math. 65 (1981/82) 179–207.
- [DeT85] D.M. DeTurck, Prescribing positive Ricci curvature on compact manifolds, Rend. Sem. Mat. Univ. Politec. Torino 43 (1985) 357–369.
- [DG99] D. DeTurck, H. Goldschmidt, Metrics with prescribed Ricci curvature of constant rank. I. The integrable case, Adv. Math. 145 (1999) 1–97.
- [Dyn57] E.B. Dynkin, Semisimple subalgebras of semisimple Lie algebras, Amer. Math. Soc. Transl. (2) 6 (1957) 111–244.
- [Gor85] C.S. Gordon, Naturally reductive homogeneous Riemannian manifolds, Can. J. Math. 37 (1985) 467–487.
- [GS10] C.S. Gordon, C.J. Sutton, Spectral isolation of naturally reductive metrics on simple Lie groups, Math. Z. 266 (2010) 979–995.
- [GP17] M. Gould, A. Pulemotov, The prescribed Ricci curvature problem on homogeneous spaces with intermediate subgroups, arXiv:1710.03024 [math.DG], 2017, to appear in Comm. Anal. Geom.
- [Ham84] R.S. Hamilton, The Ricci curvature equation, in: S.-S. Chern (Ed.), Seminar on nonlinear partial differential equations, Springer-Verlag, New York, 1984, pages 47–72.
- [Jan10] S. Janson, Roots of polynomials of degrees 3 and 4, arXiv:1009.2373 [math.HO], 2010.
- [Jen73] G.R. Jensen, Einstein metrics on principal fibre bundles, J. Differential Geom. 8 (1973) 599–614.
- [Lau19] E.A. Lauret, On the smallest Laplace eigenvalue for naturally reductive metrics on compact simple Lie groups, Proc. Amer. Math. Soc. 148 (2020) 3375–3380.
- [Pul13] A. Pulemotov, Metrics with prescribed Ricci curvature near the boundary of a manifold, Math. Ann. 357 (2013) 969–986.
- [Pul16a] A. Pulemotov, The Dirichlet problem for the prescribed Ricci curvature equation on cohomogeneity one manifolds, Ann. Mat. Pura Appl. 195 (2016) 1269–1286.
- [Pul16b] A. Pulemotov, Metrics with prescribed Ricci curvature on homogeneous spaces, J. Geom. Phys. 106 (2016) 275–283.
- [Pul20] A. Pulemotov, Maxima of curvature functionals and the prescribed Ricci curvature problem on homogeneous spaces, J. Geom. Anal. 30 (2020) 987–1010.
- [PR19] A. Pulemotov, Y.A. Rubinstein, Ricci iteration on homogeneous spaces, Trans. Amer. Math. Soc. 371 (2019) 6257–6287.
- [Sag70] A.A. Sagle, Some homogeneous Einstein manifolds, Nagoya Math. J. 39 (1970) 81–106.
- [WW70] N.R. Wallach, F.W. Warner, Curvature forms for 2-manifolds, Proc. Amer. Math. Soc. 25 (1970) 712–713.
- [WZ86] M.Y. Wang, W. Ziller, Existence and nonexistence of homogeneous Einstein metrics, Invent. Math. 84 (1986) 177–194.