Rigidity of spherical product Ricci solitons
Abstract.
We show that is isolated as a shrinking Ricci soliton in the space of metrics, up to scaling and diffeomorphism. We also prove the same rigidity for , where belongs to a certain class of closed Einstein manifolds. These results are the Ricci flow analogues of our results for Clifford-type shrinking solitons for the mean curvature flow.
1. Introduction
Ricci solitons are metrics that flow by rescaling (up to diffeomorphism) under the Ricci flow. They were first introduced by Hamilton [Ham82] to study the singular behaviour of the Ricci flow. Shrinking solitons are those that contract under Ricci flow, and relate to the formation of singularities. A gradient shrinking Ricci soliton is a Riemannian manifold that satisfies the equation
(1.1) |
for some smooth function on and some . Note that equation (1.1) has two natural symmetries: diffeomorphisms of , and rescaling (which results in a rescaled ). In the compact setting, every closed shrinking Ricci soliton is a gradient shrinking soliton. Shrinking solitons often arise as Type I singularity models of Ricci flow, and also as tangent flows, which were recently introduced by Bamler in [Bam20].
In this paper, we study the rigidity of certain gradient shrinking Ricci solitons , that is, whether is isolated in the space of shrinking Ricci solitons (modulo diffeomorphism and rescaling). Our first main theorem is the following local rigidity theorem for .
Theorem 1.1.
Let be with the round unit metric on each factor. There is a -neighbourhood of such that any other gradient shrinking Ricci soliton in , is equivalent to up to diffeomorphism and rescaling.
The question of local rigidity of was previously posed by Kröncke [Krö16, Example 6.4]. Our approach here is the Ricci flow analogue of our methods in [SZ20] for as mean curvature flow solitons (see also [ELS20], [Zhu20]). These methods are needed to handle non-integrable infinitesimal deformations. In contrast to the MCF setting, as Ricci solitons such deformations only arise when at least one spherical factor has dimension .
Our methods also can be used to show local rigidity for a class of product manifolds , where is a 1-Einstein manifold satisfying
-
()
, and .
Here we use the sign convention for the Laplacian on functions, is the first nonzero eigenvalue and is the Lichnerowicz Laplacian; refers to the kernel when acting on transverse traceless 2-tensors. We prove the following:
Theorem 1.2.
Let be , where has the round unit metric and is a closed -Einstein manifold satisfying . There is a -neighbourhood of such that any other gradient shrinking Ricci soliton in , is equivalent to up to diffeomorphism and rescaling.
There are many which satisfy the assumptions in Theorem 1.2. For example, any listed in [CH15, Table 2], that is not listed as “i.d.” and has (in their notation) .
Both theorems above are direct consequences of the following quantitative rigidity theorem. For this we introduce the shrinker quantity
(1.2) |
where are the pair realising the Perelman shrinker entropy ; an important property is that is equivariant under pullback by diffeomorphisms, and under rescalings (see Section 2). Any compact gradient shrinking Ricci soliton must satisfy (1.1) with , that is, (see [Krö15c, Remark 3.4]).
Theorem 1.3.
Let be either or , where has the round unit metric and is a closed -Einstein manifold satisfying . There exist so that if is a metric on with , then there exists and a diffeomorphism of so that,
(1.3) |
As may be regarded as the -gradient of the shrinker entropy , Theorem 1.3 may be regarded as the Hölder version of a Łojasiewicz inequality for about the critical points , . The version is somewhat more natural and would yield a gradient Łojasiewicz inequality as well, but the theory is significantly more technical so we leave it for future work; the Hölder version is more than sufficient for rigidity.
Background
The question of rigidity - whether a geometric object is isolated in its class, modulo the symmetries of that class - has been studied for various geometric structures in many contexts. For Ricci solitons, low dimensional solitons have been essentially classified, so their rigidity or non-rigidity is clear (the dimension case is due to Hamilton [Ham88] and the dimension case is due to Ivey [Ive93]). In higher dimensions, the classification is far from well-understood. From a geometric PDE perspective, Ricci solitons are natural generalizations of Einstein manifolds. Kröncke studied the local rigidity of Ricci solitons in [Krö16], using similar techniques as for Einstein manifolds in [Krö15b].
For non-rigidity results, similar to the Einstein case, Inoue [Ino19] studied the moduli space of Fano manifolds with Kähler-Ricci solitons. As a consequence, Inoue proved that some Kähler-Ricci solitons are non-rigid. It seems that the local rigidity of a Ricci soliton actually has deep connections to the topological/algebraic structure of the manifold, and we look forward to such interpretations for the class of Ricci solitons we study in this paper.
The study of rigidity of Einstein manifolds has a more developed history. The first rigidity result for Einstein manifolds was proved by Berger in [Ber66], where Berger proved that all Einstein metrics on whose sectional curvature is -pinched are isometric to the standard sphere metric. We consider this to be a ‘global’ result in that it gives a large explicit neighbourhood in which the special structure is unique. Later, Berger-Ebin [BE69] proposed a general strategy to prove local rigidity of Einstein metrics. In particular, they introduced the notion of “infinitesimal deformations”. The local rigidity of symmetric spaces as Einstein manifolds was first studied by [Koi80], and Kröncke extended Koiso’s result in [Krö15b].
Meanwhile, it is known that certain moduli spaces of Einstein manifolds, even with an extra structure (e.g. Kähler), can be nontrivial. Then any metric in such a moduli space is an example of non-rigid Einstein manifold. For example, even in dimension 2, the moduli space of surfaces with genus of constant curvature has dimension . Hence when , such hyperbolic metrics are not rigid. We refer the readers to [Bes87, Chapter 12.J] for further discussion.
The Ricci soliton equation is both more general and more complicated than the Einstein equation, making it more challenging to obtain a rigidity result. It is also interesting to compare Ricci flow with other geometric flows, especially the most natural extrinsic flow - the mean curvature flow. There are a number of results on the rigidity of shrinking solitons to the mean curvature flow - see [CIM15], [ELS20], [SZ20], [Zhu20]. All these results are about product self-shrinkers. During the completion of this work, we learnt that Colding-Minicozzi are developing techniques which could be applied to study certain noncompact Ricci solitons [CMa, CMb].
Deformation theory for solitons
The starting point in the study of local rigidity is to consider infinitesimal deformations which satisfy the equation to first order. That is, one looks for -tensors in the kernel of the linearized operator . By analogy with the mean curvature flow setting, we call elements of this kernel Jacobi deformations; in the Ricci flow literature these are sometimes referred to as infinitesimal solitonic deformations [Krö16]. A Jacobi deformation is called integrable if it induces a continuous family of gradient shrinking Ricci solitons. Any diffeomorphism or rescaling generates an integrable Jacobi deformation.
In [Krö16], Kröncke found conditions under which certain Einstein manifolds are rigid as gradient shrinking Ricci solitons. His idea was to expand to higher and higher order and examine whether there are solutions to each order. If there are no solutions at a given order, we call this an obstruction of that order. Kröncke calculates only up to order 2: has a second order obstruction, whilst for all Jacobi deformations are integrable to second order. (For , almost all Jacobi deformations are non-integrable to second order, but there is a set of positive codimension which is indeed integrable.)
The manifolds we study in this paper, also have Jacobi deformations which are integrable to order 2, but we will show they have an obstruction at order 3. In particular, we resolve the case of . We also hope that our method may be adapted to determine the rigidity of . As described below, expanding at orders higher than 2 result in inhomogenous elliptic PDEs, which makes the analysis significantly more complicated. For instance, we use the special structure of eigenfunctions on to solve (1.6) explicitly.
In [SZ20], we studied the rigidity of products of spheres as mean curvature flow self-shrinkers, and found that despite the existence of nontrivial Jacobi deformations, there is an obstruction at third order. (Note that [ELS20], by a similar method, had previously covered the case.) As mentioned above, when , as a Ricci shrinker does not have Jacobi deformations other than those from symmetry (a first order obstruction). However, when at least one of the factors is , nontrivial Jacobi deformations appear once more. We remark that when considered in the class of Einstein manifolds, there are again no nontrivial deformations, even for products.
Proof strategy
Our proof strategy is a deformation-obstruction theory for the shrinker quantity , mirroring that we used in [SZ20] (also developed by the second named author in [Zhu20]). One may consult those papers for an overview of the method, including in toy cases, but we reproduce a sketch below for the readers’ convenience. The basic idea is similar to that of Kröncke [Krö16], but we study the expansion of more directly to find a ‘uniformly’ obstructed term, which gives the quantitative rigidity.
Consider an elliptic functional defined on a Banach space . Let be the Euler-Lagrange operator of and suppose is a critical point of , so that . In the Ricci soliton setting, will be the shrinker entropy and will be the Ricci shrinker quantity .
For , we have the formal Taylor expansion
(1.4) |
Here , where is the linearised operator at 0. Its kernel is the space of Jacobi deformations.
If , ellipticity gives that is invertible, hence is obstructed at order 1. (As mentioned above, modulo symmetries, this is the case for , .) Otherwise , and the first order term suggests the decomposition , where and is in a complement of . This gives the further expansion
(1.5) |
Invertibility of on implies that is (at least) second order compared to and . The second order term is then . By elliptic theory, there exists for which this term is 0, if and only if . In [Krö16], Kröncke essentially finds Einstein manifolds for which for all nonzero . This would imply , which we refer to as an order 2 obstruction, and in turn implies .
For the cases considered in this paper, it turns out that . That is, there is always a solution of
(1.6) |
for some , which suggests the further decomposition . The expansion then becomes
(1.7) |
where again invertibility of shows that the residual is of third order compared to . Repeating the process, we may attempt to solve the third order condition
(1.8) |
which depends on the projection Note that the third order condition involves some lower order cross terms.
If the projection vanishes identically, we proceed to the next higher order. If at the th order, the corresponding projection is nonzero for any nonzero , then we have an th order obstruction, which would imply the quantitative rigidity . Again, for the cases considered here, we find a third order obstruction
(1.9) |
Organisation of the paper
We collect some preliminary results in Section 2 and record descriptions of the Jacobi deformations on our products in Section 3. In Section 4 we compute the variation of the shrinker quantities , with the variation formulae for more basic geometric quantities recorded in Appendix A. Using the explicit description of Jacobi deformations, we specialise these variations to our products in Section 5 and establish a formal obstruction at order 3. Finally, in Section 6, we use Taylor expansion and the formal obstruction to prove the main quantitative rigidity Theorem 1.3. The a priori bounds on are deferred to Appendix B.
Acknowledgements.
The authors want to thank Professors Bill Minicozzi and Toby Colding for sharing in enlightening discussions about their work and ours. JZ would also like to thank Prof. Richard Bamler for helpful conversations. We thank Prof. Klaus Kröncke for rectifying our understanding of his work as well as Jiangtao Yu for pointing out an inaccuracy in the prior computation.
JZ was supported in part by the Australian Research Council under grant FL150100126.
2. Preliminaries
Throughout, we consider a closed Riemannian manifold .
Our sign convention for the (rough) Laplacian is . Here, and henceforth, we suppress dependence on the metric when clear from context. When we wish to emphasise the domain, we will use for the Laplacian on functions, for the rough Laplacian on 1-forms and for the rough Laplacian on symmetric 2-tensors. When not otherwise specified or clear from context, should be understood as the Laplacian on functions. With our sign convention, we say is an eigenfunction of with eigenvalue if , and similarly for other operators. We write for the first nonzero eigenvalue.
A metric is said to be -) Einstein if for some . Let denote the space of (smooth) symmetric -tensors on , and the space of (smooth) 1-forms on . Throughout the paper, we will use several different norms like the Sobolev norm and the Hölder norm. We can also study the tensors in the completion under these norms, but approximation by smooth sections easily yields the same estimates.
An important subspace is the space of transverse traceless (TT) -tensors, which we denote by . Then we have the well-known (-) orthogonal decomposition
(2.1) |
Here is the Lie derivative, . Note that is actually independent of the metric, and is the -adjoint of for any .
We define the Lichnerowicz Laplacian as acting by
Note that this definition also makes sense on all of but we choose to restrict the domain so that and so forth will refer to the TT kernel.
Define
The shrinker entropy is defined by
(2.2) |
As in [Krö15c, Lemma 4.1], given any gradient shrinking Ricci soliton, there is a neighbourhood in the space of metrics on which the minimisers are unique and depend analytically on the metric; consequently also depends analytically on the metric on this neighbourhood. Moreover, the first variation of is given by
(2.3) |
where the Ricci shrinker quantity is defined by
(2.4) |
(Compact) gradient shrinking Ricci solitons are those metrics which satisfy . For convenience we denote . By definition, we have the normalisation . Henceforth, for readability, we will suppress dependence on the metric where clear from context.
Any Einstein manifold is a gradient shrinking Ricci soliton with constant. Henceforth, will always refer to the Lebesgue space weighted by ; if is Einstein this is equivalent to up to the constant weight .
Through this paper, will denote constants which may change from line to line but retain their stated dependencies.
We record the following observation:
Lemma 2.1.
(2.7) |
Proof.
Using the Bianchi identity gives . By commuting derivatives we have . On the other hand, differentiating the defining equation for in space gives
This implies the result. ∎
2.1. Variations of shrinker quantities
Consider a soliton metric . We consider the formal expansion of given by
where is a symmetric -linear map . For instance, is the linearisation of at . We will henceforth suppress dependence on the initial metric . We will use the corresponding notation for variations of other quantities such as .
2.2. Deformations and Jacobi deformations
Given a 1-parameter family of metrics with , the corresponding infinitesimal deformation is . At a shrinking soliton , we define to be the kernel of . We call elements of Jacobi deformations; they correspond to deformations which preserve to first order and have previously been referred to as infinitesimal solitonic deformations.
The Ricci shrinker quantity has two natural symmetries: under scaling by we have , and under pullback by a diffeomorphism we have . The corresponding spaces of infinitesimal deformations are and , and if these are integrable subspaces of Jacobi deformations. We define the subspace of Jacobi deformations from symmetry to be .
If is a -Einstein metric, Kröncke [Krö16, Section 6] checked that
Note that . A theorem of Lichnerowicz ([Lic58], see also [IV15, Theorem 2.1]) implies that , and in particular that is invertible.
Let , so that the conformal deformation space satisfies
(2.8) |
Without loss of generality, we will assume . The action of is well-known and also computed in Section 4.1: It acts on the conformal part , where , , by
and on the TT part by
Recall that we defined as an operator on . The above implies the -orthogonal decomposition of the kernel
(2.9) |
where
(2.10) |
We denote by the -projection to , and similarly for other subspaces.
To study deformations of a gradient shrinking Ricci soliton, we need to account for the symmetry action at an integrated level, not just infinitesimal. More precisely, given a gradient shrinking Ricci soliton and a metric close to , we want to compare to the symmetry orbit of (or vice versa).
Lemma 2.2.
Suppose is a -Einstein manifold. For any there exists such that if , then there exists and such that
and moreover .
The proof is standard, using implicit function theorem for Banach spaces. We refer the readers to the proof of Theorem 3.6 in [Via14].
3. Jacobi deformations on Einstein manifolds
In this section we discuss Jacobi deformations on products of Einstein manifolds, with the goal of giving an explicit description of the Jacobi deformations on and certain products . We also demonstrate that there are no Jacobi deformations on , , other than those from scaling and diffeomorphism.
3.1. Spectrum on product Einstein manifolds
By a standard separation of variables, spectra on a product manifold can be derived from the spectra of each component. For the Laplacian on functions, we have
(3.1) |
For the spectrum of , it is convenient to first work on all of . To match with previous literature, we define the Einstein operator by
Note that this differs from the expression for by exactly the Ricci terms; in the case of a -Einstein manifold, for we have .
We have the following theorem (see [AM11] and [Krö15a]; note that our sign convention is opposite to Kröncke’s).
Theorem 3.1.
Suppose are -Einstein manifolds. Then
(3.2) |
Moreover, the eigentensors on the product manifold may be given by (symmetric tensor) products of the corresponding eigensections.
3.2. Spectra on round spheres
We will always consider with the metric induced as the sphere of radius in , which is -Einstein. The spectra of the Laplace operators has been computed (see for instance [Bou99, Bou09]):
For define , and .
-
•
The spectrum of the Laplacian on functions consists of eigenfunctions with eigenvalue , for .
-
•
The Hodge Laplacian acts on by and has spectrum consisting of , , together with divergence-free eigenforms with eigenvalue , for .
-
•
The (generally defined) Lichnerowicz Laplacian acts on by , and its spectrum consists of: , ; , ; , for ; and TT eigenforms with eigenvalue , for .
That is,
In particular, since for any , there are no TT eigentensors of eigenvalue 0, that is, .
Later we focus on the case (which is in fact the unit sphere), so we record separately
On , the -eigenspace consists of the coordinate functions , which satisfy , , and .
3.3. Jacobi deformations on product manifolds
If the product manifold is for , there is no Jacobi deformation other than those induced by diffeomorphisms from the spectrum on product manifolds. This is a consequence of the following lemma.
Lemma 3.2.
Let , with the round 1-Einstein metric . If then
Proof.
For the Laplacian on functions, note that for we have , so the sum of any two nonzero eigenvalues is strictly greater than 2. But , so .
For the Lichnerowicz Laplacian, note that is strictly positive, so clearly . Also note that the least eigenvalue of is , and the next eigenvalue is . Since and we conclude using Theorem 3.1 that . ∎
In the following we only focus on the product manifolds with components.
We give precise descriptions of Jacobi deformations on and . Recall that we decomposed the kernel as
(3.3) |
where is the space of Jacobi deformations from symmetries, and
(3.4) |
Let be the projections onto each factor.
Lemma 3.3.
Let with the round 1-Einstein metric , we have
Proof.
For the Laplacian on functions, since , the only product eigenfunctions with eigenvalue 2 are those which are constant on one factor and have eigenvalue 2 on the other.
For the Lichnerowicz Laplacian, is strictly positive, so clearly . Also the least eigenvalue of is , and the next eigenvalue is . Moreover the eigenspaces are and . So the only product eigentensors in are of the forms , , or . Here is the (pulled-back) metric on the th factor, , and , . But , and . The TT kernel is therefore . ∎
We will consider 1-Einstein manifolds , which additionally satisfy:
-
()
, and .
Lemma 3.4.
Let , where has the round 1-Einstein metric and is a -Einstein manifold satisfying (), we have
Proof.
For the Laplacian on functions, since , the only product eigenfunctions with eigenvalue 2 are those which are constant on and have eigenvalue 2 on the factor.
For the Lichnerowicz Laplacian, is strictly positive and is nonnegative for any Einstein manifold (cf. [Krö15a, Lemma 4.4]), so . Then by , the only product eigentensors in are of the forms or , where and , . But , and again . The TT kernel is therefore . ∎
Later, it will be convenient to collect the two previous cases as . We will write any as , where and is the Laplacian on the th factor. It will be understood that each is independent of any other factors; explicitly, in the above.
4. Variations
In this section we compute the variation of the quantities , by considering certain families of metrics with . The initial soliton will be an Einstein manifold , with Einstein constant , hence and . In particular the initial weight is also a constant.
Note that the variations of more familiar differential geometric quantities are given in Appendix A. For the more complicated quantities here, the procedure will be as follows:
We always denote where , and similarly for .
Note that any variation of or will act as 0 on any constant function on . In particular, at the initial metric is constant so , .
For convenience, we define
where understood as the volume ratio using the initial metric as the reference metric.
Step (2) is a general procedure, but here as our initial soliton is Einstein, it will be expedient for us to actually consider , where . Note that , so variations of match those of . Moreover, , while all variations of match those of .
4.1. First variation
We begin with the first variation as it is distinguished in our method. So we consider the 1-parameter family of metrics , where to start is a general element of satisfying . Here we use primes to denote the derivative in , evaluated at 0.
As at , we immediately have
Note that . Then
Taking , since we have Therefore
Differentiating (2.5) and using gives
(4.1) |
4.1.1. TT direction
Here we compute , where is transverse traceless. To match the notation used for higher variations, we use subscripts to denote derivatives in .
As , the variation formulae in Appendix A.1 give and therefore
They also give
where is the Lichnerowicz Laplacian.
Then
4.1.2. Conformal direction
Here we compute , where , . For use in higher variations, it will be convenient to change the parameter to . That is, here we consider the family .
Since commutes with itself, (4.1) gives that
Here we have used the variation formulae in Appendix A.1 (or Appendix A.2). It will be useful for later to record
Then we have
This identifies the conformal kernel as .
4.2. Higher variations in the conformal kernel direction
We now proceed to find variations in the directions , where , by consider the 1-parameter family of metrics .
4.2.1. Second variation
First, we have . Since at we have , this immediately implies
We calculate
By the above, since , we have and
We also compute
Then once again taking , we have
This gives
Now differentiating (2.5) twice, we have
Using the conformal variation formulae in Appendix A.2 gives
(4.2) |
At we then calculate , so using we have
(4.3) |
Note that, by differentiating (2.7) twice, at (using ) we have
(4.4) |
4.2.2. Third variation
Here we compute , where .
For brevity, we will omit the solution of (and ) - it is constant on , and therefore will not contribute after integration against anyway.
The third derivative of (2.5) gives
Using the variation formulae in Appendix A.2 then gives the following defining equation for ,
(4.5) |
We then have
Since at , we then have
(4.6) |
Note that the term will vanish upon integration against .
Taking the inner product we have
(4.7) |
4.3. Second variation cross terms
Finally, we calculate certain cross terms for .
4.3.1. Kernel-TT direction
Here we compute , where , and is any smooth function on , by considering the 2-parameter variation .
By the calculations in Section 4.1 we have and , .
First, using at , we have
as the Lichnerowicz Laplacian preserves the TT decomposition.
Calculate
Now as we have , and from the calculations in Section 4.1 we have , . Taking again, we have and . So since we have
and hence
Differentiating (2.5) in then gives
Note that acts as . Using the formulae in Section A.3 then gives hence
As we then have Since is TT, the first variation Lemma A.1 gives . Therefore .
Again using Appendix A.3 we conclude that
(4.8) |
4.3.2. Kernel-conformal direction
Here we compute , where , and , by considering the metric variation . We furthermore assume (hence ) is orthogonal to . By calculations in Section 4.1 we have and , .
Again since at , we have
As before we have
Now as we have , and as in Section 4.1 we have , . Taking again, since we again have
It then follows from being orthogonal to that
Differentiating (2.5) in then gives . This yields the defining equation for ,
(4.9) |
Then, using the conformal cross-variations in Appendix A.2.1, we have
Taking the inner product with and collecting terms, it follows that
(4.10) |
where is defined as in (4.9).
5. Specialising to products
In this section we consider a product of 1-Einstein manifolds , where has the (unit) round metric. The goal is to derive a formal obstruction for at order 3. To do so, we will use explicit knowledge of the Jacobi deformations on together with the formulae in the previous section.
5.0.1. Explicit form
We assume that either is , or satisfies assumption (), so that the results of Section 3.3 apply. It will actually be convenient to unify the cases by collecting factors: We write , where if , and otherwise . Note here the subscript serves only to distinguish the isometric factors.
Throughout this section, we consider . As in Section 3.3, we have , where we may write and for some and some vector . Then as in Section 3.2, we have and .
We introduce the notation and . Then , , and . Finally note .
5.1. Spherical integrals
Proof.
Note that , and . Also note that . The first four equations are then routine computations. For the last four, note that spherical polynomials of odd degree integrate to 0. ∎
5.2. Explicit solutions
In this subsection, we use the explicit form of to explicitly solve for certain quantities in Section 4.
We will frequently use the ansatz that a function is a linear combination
(5.1) |
and formally solve the resulting equations as linear systems. Under this ansatz, we represent such a function by the column vector , and the Laplacian acts as
The Laplacian on a particular factor acts as
5.2.1. Second variation of
Here we find an explicit solution for the function in Section 4.2.1. Recall that the defining equation (4.2) is
where we have used
We use the ansatz , and solve the linear system as
One may verify that this solution indeed satisfies the defining equation, which proves
5.2.2. Second variation of
Proposition 5.3.
Let and be as in Section 5.0.1. Then
Proof.
By the results of Section 3.3, we have , and . Since , we immediately have .
5.2.3. Second variation of the metric
Since , there is now a unique , and such that
(5.3) |
where . We will give an explicit solution for . For , it will later suffice to give an explicit solution for modulo .
First, tracing (5.3) using the first variation formulae in Section 4.1 and the second variation formulae (4.3) gives the defining equation for :
(5.4) |
Using the same ansatz , and the explicit solution of , we can solve the linear system as
One may verify that the corresponding solution
indeed satisfies equation (5.4). For convenience we also record
(5.5) |
With in hand, we consider the pointwise inner products . We have
so pairing (5.3) with we find the defining equation
(5.6) | |||||
We will find so that equals the right hand side of (5.6); then . Here we use the ansatz . Using the solution of (hence ), we may solve the linear system as
One may verify that the corresponding solution indeed solves (5.6). In conclusion, we have proven
Lemma 5.4.
Let and be as in Section 5.0.1. There is a unique solution of , where , , , . Moreover they satisfy
(5.7) |
(5.8) |
for some .
5.3. Integration by parts
We will use the following integration by parts tricks. The first equation below, in particular, is used for terms involving derivatives of .
5.4. Second variation cross terms
Proposition 5.6.
Proof.
Pair the second variation formula (4.10) with . Using Lemma 5.5 and the explicit forms (5.2, 5.7) of and , every term in may be expanded as a product of the functions . All such products have integral computed in Lemma 5.1, and collecting terms gives the first equation.
By formula (4.8), we have . Now note that for any , the product is a spherical polynomial of total degree 3, and hence . Therefore we need only integrate against the explicit portion of the solution (5.8) for - that is, against in the proof of Lemma 5.4.
Consider the basis functions . Multiplying by and summing over respectively give . Only the last two do not have integral listed in Lemma 5.1. For these we again note that spherical polynomials of odd degree integrate to 0, which implies and . Collecting terms then gives the second equation. ∎
5.5. Third variation
Proposition 5.7.
Let and . Then
(5.11) |
5.6. Formal obstruction
Combining the above formulae, we have
Theorem 5.8.
Let and . Let be the unique solution of . Then we have
(5.12) |
where
Corollary 5.9.
Let and suppose . Let be the unique solution of . If , or if , then there exists such that
(5.13) |
Proof.
Recall that .
First assume . Then have explicitly and . By the trivial inequality we then have
Now assume . Then , and
Note that the numerator has no integer roots.
Therefore, in either case,
which implies the result. ∎
Note that one may replace in (5.13) with any norm since is finite-dimensional.
6. Hölder theory
In this section, we use Taylor expansion to turn the formal obstruction into the quantitative rigidity Theorem 1.3. Throughout this section will denote the 1-Einstein manifold , where either or satisfies ().
Recall that is the space of Jacobi deformations, and is the subspace of Jacobi deformations modulo symmetries. We use to denote an unspecified contraction of the tensors . We will freely use that Hölder norms behave well under contractions, i.e. .
6.1. Taylor expansion
In the sequel, we would like to apply Taylor expansion to . To do so, we need a priori bounds for in terms of a finite number of derivatives of . However, this is not as simple as in [SZ20], since the explicit form of depends not only on the 2-jet of , but also on the implicitly defined and (the 2-jet of) .
The goal is to show that , , and are uniformly bounded as -linear operators on with values in , , and respectively. Explicitly, let
Then we have
Lemma 6.1.
Fix a reference soliton metric . There exists such that if , then for any the variations at satisfy
(6.1) |
The proof of Lemma 6.1 is deferred to Appendix B, and uses an inductive process similar to Section 4. We now proceed with our Taylor expansion:
Lemma 6.2.
Fix a metric . There exist so that if , then
(6.2) |
Proof.
Set , so that . We then apply Taylor’s theorem with remainder (in integral form) pointwise to and to get the result. Note that Lemma 6.1 gives uniform bounds on in a neighbourhood of , which controls the remainder. ∎
6.2. Hölder estimate
Here we first prove a quantitative rigidity for deformations modulo symmetries as follows:
Theorem 6.3.
Let , where or satisfies (). There exists such that if and , then
(6.3) |
The proof of Theorem 6.3 is somewhat formal and proceeds as in [SZ20]. We reproduce the strategy here in detail for those unfamiliar with the setting there.
We will repeatedly use the following estimate, which takes advantage of the fact that is finite-dimensional.
Lemma 6.4.
For any , we have
(6.4) |
Proof.
The first two inequalities are the equivalence of norms on a finite-dimensional space. Pythagoras theorem in space implies that The last inequality is just integration as has finite volume.
∎
We also need the following elliptic Schauder estimate:
Lemma 6.5.
There exists so that if , then
In what follows we consider a soliton metric , and recall that . Also, we assume for small enough so that . We will frequently use this to absorb terms of higher degree, and also freely use Young’s inequality.
6.2.1. 1st order analysis
By Lemma 6.2 with we have
We consider , and decompose , where and so . Note that indeed by the assumption on . The above then becomes
(6.5) |
This allows us to show that the residual is of higher order again:
Lemma 6.6.
There exist so that if and is as above, then
(6.6) |
(6.7) |
6.2.2. 2nd order analysis
By Lemma 6.2 with we have
Expanding by our decomposition of and using the uniform estimates on ,
As discussed in Section 5.2.3, there is a unique solution , where , , and . Then we have the further decomposition , and the above becomes
(6.10) |
This allows us to show that the residual is of higher order:
Lemma 6.7.
There exist so that if and is as above, then
(6.11) |
Proof.
We also record estimates on :
Lemma 6.8.
There exist so that if and are as above, then
(6.12) |
6.2.3. 3rd order analysis
By Lemma 6.2 with , we have
Using our decomposition , this becomes
By the uniform estimates on , after expanding further we get
(6.13) |
Proof of Theorem 6.3.
Decompose as above. By the formal obstruction Corollary 5.9 and equivalence of norms on the finite-dimensional space , we have
(6.14) |
In the second line we have used Lemma 6.4 and that (by self-adjointness). Then by the triangle inequality and (6.13), we have
(6.15) |
Absorbing higher order terms yields
(6.16) |
Absorbing higher order terms again gives the desired estimate
(6.18) |
∎
Appendix A Variation of basic geometric quantities
In this section we compute various variation formulae for the geometric quantities and their traces .
A.1. First variation
Here we list the well-known first variations of the basic geometric quantities.
Lemma A.1.
Consider a manifold with a 1-parameter family of metrics . Then at , if is an orthonormal frame, we have
(A.1) |
(A.2) |
(A.3) |
(A.4) |
A.2. Conformal direction
Here we consider higher order variations under conformal change.
Lemma A.2.
Let be a manifold with the metric family , where (and ) is a smooth function on . Then we have the variation formulae
Proof.
The relevant quantities are given explicitly under conformal variation by
The above variation formulae follow by setting and expanding in to third order. Note that .
∎
A.2.1. Second variation cross term
Lemma A.3.
Let be a manifold with the 2-parameter family of metrics , where are smooth functions on . Then we have
Proof.
Polarise the second variation formulae in the previous lemma. ∎
A.3. Second variation in kernel-TT direction
Here we compute the basic variations needed in Section 4.3.1, by considering the 2-parameter variation . This will allow us to use the variation formulae under conformal change, although now so for any quantity we will have .
Lemma A.4.
Let be a closed 1-Einstein manifold, and and . Then for any smooth function on , we have
Appendix B Uniform bounds for shrinker quantities
Here we prove Lemma 6.1, which is reproduced below. Recall that denotes an unspecified contraction of the tensors , and .
Lemma B.1 (= Lemma 6.1).
Fix a reference soliton metric . There exists such that if , then for any the variations at satisfy
(B.1) |
Proof.
We proceed by induction on . For brevity we abbreviate and . The main idea is that we can estimate derivatives of by differentiating their defining equations in the direction and focussing on the highest order terms; all other terms will be estimated by the inductive hypothesis.
The case is clear as each quantity depends smoothly on . So suppose that
for all . For a tensor quantity we denote an unspecified linear combination of contractions of copies of with total order by
In particular, if , then by the inductive hypothesis
Recall that , where . Differentiating this further, we find that
where . It follows immediately that .
By differentiating the normalisation , we have
Similarly by differentiating , we have
Eliminating , we then have
It follows that .
Recall the defining equation (2.5) for , which states
All the basic geometric quantities involve at most the 2-jet of . Let denote the drift Laplacian. By differentiating the defining equation, we will have
Here , denote the 2-jets of and (with respect to the reference metric on ), and the sum is over all partitions of .
By the previous estimates, we then have
For a shrinking soliton, it is known that (see for instance [CZ12]). Therefore, by continuity, for small enough we will have . In particular, is coercive (uniformly in ), so by elliptic Schauder theory
Finally, note that again all the basic geometric quantities in the definition
depend on at most the 2-jet of . Then
By the previous estimates we conclude that indeed , which completes the induction. ∎
References
- [AM11] Lars Andersson and Vincent Moncrief, Einstein spaces as attractors for the Einstein flow, J. Differential Geom. 89 (2011), no. 1, 1–47. MR 2863911
- [Bam20] Richard Bamler, Compactness theory of the space of super ricci flows, arXiv:2008.09298 (2020).
- [BE69] M. Berger and D. Ebin, Some decompositions of the space of symmetric tensors on a Riemannian manifold, J. Differential Geometry 3 (1969), 379–392. MR 266084
- [Ber66] M. Berger, Sur les variétés d’Einstein compactes, Comptes Rendus de la IIIe Réunion du Groupement des Mathématiciens d’Expression Latine (Namur, 1965), Librairie Universitaire, Louvain, 1966, pp. 35–55. MR 0238226
- [Bes87] Arthur L. Besse, Einstein manifolds, Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], vol. 10, Springer-Verlag, Berlin, 1987. MR 867684
- [Bou99] Mohamed Boucetta, Spectre des laplaciens de Lichnerowicz sur les sphères et les projectifs réels, Publ. Mat. 43 (1999), no. 2, 451–483. MR 1744618
- [Bou09] by same author, Spectra and symmetric eigentensors of the Lichnerowicz Laplacian on , Osaka J. Math. 46 (2009), no. 1, 235–254. MR 2531148
- [CCG+07] Bennett Chow, Sun-Chin Chu, David Glickenstein, Christine Guenther, James Isenberg, Tom Ivey, Dan Knopf, Peng Lu, Feng Luo, and Lei Ni, The Ricci flow: techniques and applications. Part I, Mathematical Surveys and Monographs, vol. 135, American Mathematical Society, Providence, RI, 2007, Geometric aspects. MR 2302600
- [CH15] Huai-Dong Cao and Chenxu He, Linear stability of Perelman’s -entropy on symmetric spaces of compact type, J. Reine Angew. Math. 709 (2015), 229–246. MR 3430881
- [CIM15] Tobias Holck Colding, Tom Ilmanen, and William P. Minicozzi, II, Rigidity of generic singularities of mean curvature flow, Publ. Math. Inst. Hautes Études Sci. 121 (2015), 363–382. MR 3349836
- [CMa] Tobias Holck Colding and William P. Minicozzi, II, Optimal growth bounds for eigenfunctions, in preparation.
- [CMb] by same author, Singularities of ricci flow and diffeomorphisms, in preparation.
- [CZ12] Huai-Dong Cao and Meng Zhu, On second variation of Perelman’s Ricci shrinker entropy, Math. Ann. 353 (2012), no. 3, 747–763. MR 2923948
- [ELS20] Christopher G. Evans, Jason D. Lotay, and Felix Schulze, Remarks on the self-shrinking Clifford torus, J. Reine Angew. Math. 765 (2020), 139–170. MR 4129358
- [Ham82] Richard S. Hamilton, Three-manifolds with positive Ricci curvature, J. Differential Geometry 17 (1982), no. 2, 255–306. MR 664497
- [Ham88] by same author, The Ricci flow on surfaces, Mathematics and general relativity (Santa Cruz, CA, 1986), Contemp. Math., vol. 71, Amer. Math. Soc., Providence, RI, 1988, pp. 237–262. MR 954419
- [Ino19] Eiji Inoue, The moduli space of Fano manifolds with Kähler-Ricci solitons, Adv. Math. 357 (2019), 106841, 65. MR 4017922
- [IV15] Stefan Ivanov and Dimiter Vassilev, The lichnerowicz and obata first eigenvalue theorems and the obata uniqueness result in the yamabe problem on cr and quaternionic contact manifolds, arXiv:1504.03259 (2015).
- [Ive93] Thomas Ivey, Ricci solitons on compact three-manifolds, Differential Geom. Appl. 3 (1993), no. 4, 301–307. MR 1249376
- [Koi80] Norihito Koiso, Rigidity and stability of Einstein metrics—the case of compact symmetric spaces, Osaka Math. J. 17 (1980), no. 1, 51–73. MR 558319
- [Krö15a] Klaus Kröncke, On infinitesimal Einstein deformations, Differential Geom. Appl. 38 (2015), 41–57. MR 3304669
- [Krö15b] by same author, On the stability of Einstein manifolds, Ann. Global Anal. Geom. 47 (2015), no. 1, 81–98. MR 3302177
- [Krö15c] by same author, Stability and instability of Ricci solitons, Calc. Var. Partial Differential Equations 53 (2015), no. 1-2, 265–287. MR 3336320
- [Krö16] by same author, Rigidity and infinitesimal deformability of Ricci solitons, J. Geom. Anal. 26 (2016), no. 3, 1795–1807. MR 3511458
- [Lic58] André Lichnerowicz, Géométrie des groupes de transformations, Travaux et Recherches Mathématiques, III. Dunod, Paris, 1958. MR 0124009
- [SZ20] Ao Sun and Jonathan J. Zhu, Rigidity and łojasiewicz inequalities for clifford self-shrinkers, arXiv:2011.01636 (2020).
- [Via14] Jeff A. Viaclovsky, Critical metrics for riemannian curvature functionals, arXiv:1405.6080 (2014).
- [Zhu20] Jonathan J. Zhu, Łojasiewicz inequalities for cylindrical self-shrinkers by perturbation, In preparation (2020).