-invariant Ricci solitons and ancient flows on
Abstract
Consider the standard action of on . We establish the existence of a uniform constant so that any -invariant Ricci soliton on with Einstein constant must have Riemann curvature and volume bounded by , and injectivity radius bounded below by . This observation, coupled with basic numerics, gives strong evidence to suggest that the only -invariant Ricci solitons on are round. We also encounter the so-called ‘pancake’ ancient solution of the Ricci flow.
1 Introduction
Let be a smooth manifold. A gradient Ricci soliton on is a Riemannian metric for which there exists a constant and a smooth function so that
(1) |
Solutions of (1) arise as self-similar solutions to the well-known Ricci flow
(2) |
In the quest for new solutions to (1), one often assumes that and are invariant under a certain group action of . It is of no use to assume that acts transitively on , because then is homogeneous and is constant, so must be Einstein. Finding homogeneous Einstein metrics is its own well-studied problem; perhaps the crowning achievement in this area of study is the work in [3], where the authors show that the mountain pass theorem is quite generally applicable in the construction of compact homogeneous Einstein metrics. The classification of non-compact homogeneous Einstein metrics is the subject of the long-standing Alekseevskii conjecture (see Conjecture 7.57 of [1]).
After assuming that acts transitively, the next natural step is assuming that acts with cohomogeneity one, which means that the generic orbits of the action of in have dimension one less than that of the manifold. Several examples of gradient Ricci solitons have been constructed using cohomogeneity one invariance (perhaps the most notable examples are [2] and [9]). In this paper, we consider the problem of solving (1) on with for a pair which is invariant under the usual cohomogeneity one action of . We first show a compactness result for solutions of this problem.
Theorem 1.
There exists a so that any -invariant solution of (1) (with ) on has , , and Riemann curvature bounded pointwise by .
Theorem 1 forms part of a well-established area of study which seeks to produce various types of compactness results for spaces of gradient shrinking Ricci solitons, especially on four-dimensional manifolds. The strongest general result to date appears to be Theorem 1.1 in [12], which shows compactness in the orbifold sense (with the pointed Cheeger-Gromov topology), but only once a uniform lower bound on the Perelman entropy is known. The proof of this orbifold compactness result essentially boils down to establishing a uniform norm on the Riemann curvature, rather than the stronger uniform bounds we establish with Theorem 1.
We hope that Theorem 1 brings us a step closer to actually determining uniqueness of invariant Ricci solitons.
Conjecture 1.
Any -invariant solution of (1) on with is the round sphere, up to diffeomorphism.
The prospects of verifying this conjecture numerically are discussed in Section 7.
In the course of proving Theorem 1, we discover that our notion of compactness is not very rigid, in the sense that we find a sequence of ‘almost’ Ricci solitons with unbounded Riemann curvature. These ‘almost’ solitons are an interpolation between the Gaussian shrinker on and a rescaled product of the Bryant soliton on with a flat metric on . The only reason we can conclude that these are not Ricci solitons is that these metrics have non-negative Riemann curvature; if these metrics were solitons, they would be round by Hamilton’s pinching results in [11]. However, it turns out that this ‘pancake’ shape is a -noncollapsed ancient Ricci flow on with positive Riemann curvature.
Theorem 2.
There exists a and a -noncollapsed, non-round -invariant ancient solution to the Ricci flow on with positive Riemann curvature operator.
Acknowledgements
I would like to thank David Buttsworth for guidance in implementing some of the numerical aspects of this project. I would also like to thank Max Hallgren, Mat Langford, Jason Lotay and Yongjia Zhang for several useful conversations regarding the ‘pancake’ ancient solution of the Ricci flow.
2 Preliminaries
In case the metric and the function are invariant under a certain cohomogeneity one action, the Ricci soliton equation (1) reduces to a system of ordinary differential equations. In this section, we discuss these equations in the case that our cohomogeneity one action is , and we provide some initial results on their solutions using the maximum principle. It turns out that all of the material in this section applies to solitons on both and which are invariant under the obvious action of , so for this section only, we consider both of these four-dimensional manifolds.
2.1 The boundary value problem
Under the action of , the principal orbits of or are product spheres . In the case of , the two singular orbits are one copy of and one copy of , whereas for , both singular orbits are copies of . Up to diffeomorphism, any -invariant Riemannian metric on or has the form
(3) |
where are smooth and positive functions, is the standard one-form on , and is the round metric of unit Ricci curvature on . In order for the Riemannian metric to close up smoothly at the singular orbits, the functions and must be smoothly extendable to functions on so that
(4) |
in the case of , and so that
(5) |
in the case of . At any point in a principal orbit, let be an orthonormal basis adapted to the standard product metric on . Then in the basis , we compute (cf. [10]) the Riemann curvature operator:
(6) |
Suppose we have an -invariant metric of the form (3) which is a gradient shrinking Ricci soliton with an -invariant potential function . Then depends only on the parameter, and can be smoothly extended to a function on so that
(7) |
and the functions satisfy (4) or (5), alongside the Ricci soliton equation (1) on which becomes
(8) |
We find it useful to set and introduce the new functions , , , so the Ricci soliton equations in these variables are
(9) |
Using the boundary conditions, a solution of (9) on uniquely determines a solution of (8). Note that in these co-ordinates, the Riemann curvature eigenvalues for a Ricci soliton are:
2.2 An initial step towards compactness: the maximum principle
If we were lucky enough to know a priori that any -invariant shrinking soliton on had to have non-negative Riemann curvature operator, then Hamilton’s pinching results for -manifolds under the Ricci flow in [11] would prove Conjecture 1 in the affirmative immediately (and would therefore give us the proof of Theorem 1 as well). Although there does not seem to be a way to cheaply show that our Ricci solitons must have non-negative Riemann curvature, a relatively simple application of the maximum principle does show that at least two of the six Riemann curvature eigenvalues must be non-negative everywhere. Since this observation is straightforward, and is frequently used in the remainder of the paper, we include the proofs in this preliminary section.
Proposition 1.
For any -invariant gradient shrinking Ricci soliton on or of the form (3), the quantity is non-negative everywhere.
Proof.
For a given point on a principal orbit, consider the selfdual/anti-selfdual basis
In this basis, the Riemann curvature operator is given by
where
After applying the Uhlenbeck trick, Hamilton shows in [11] that, under the Ricci flow, the curvatures satisfy the evolution equations
Since the second and third diagonal entries of both and are identical, the first entries of and are non-negative. We therefore find that the first eigenvalue of the matrix must be increasing under the Ricci flow. However, the Riemann curvatures of gradient shrinking Ricci solitons evolve only by diffeomorphisms and scalings under the flow, so the first eigenvalue of the matrix , which is , must be non-negative everywhere. ∎
Proposition 2.
For any -invariant gradient shrinking Ricci soliton on or of the form (3), the quantity is non-negative everywhere. More generally, suppose we have an -invariant Riemannian metric of the form (3), as well as two points so that:
-
•
at ; and
-
•
the metric satisfies the Ricci soliton equations on .
Then on .
Proof.
Clearly it suffices to show that on . Differentiating and using the equations of (8), we find
Therefore, if somewhere on , then there must be a point in with , and . At this point, we find
a contradiction. We obtain a similar contradiction if there were a point with . A metric which is a Ricci soliton everywhere has at and for both and by (4) and (5), so the claim follows. ∎
3 The shooting problem for : a curve and a surface
We turn to the problem of establishing Theorem 1. It is convenient to cast the problem of finding solutions of (4), (7) and (9) as a shooting problem: the idea is to study the initial value problem for (9) around the two singular orbits at , and examine how the solutions meet at a specified principal orbit. In particular, we will examine how the solutions meet at an orbit where , since (9) implies that there will be exactly one such orbit for a shrinking soliton.
To begin, we examine more carefully the initial value problem at the orbit (). Note that, by (4) and (7), there must be smooth functions so that
(10) |
and for all . It turns out that whenever solving (9) subject to (10) close to , solutions are uniquely determined by the number
(11) |
Proposition 3.
The proof of Proposition 3 essentially follows from the techniques discussed in [5]. We can similarly examine the initial value problem at the orbit. This time, (4) and (7) imply that there must be smooth functions so that
(12) |
where again for . This time, solutions are uniquely determined by
(13) |
Proposition 4.
The main idea behind the proof of Theorem 1 is to provide bounds on the possible values that can achieve. Fortunately, we can find some of these bounds immediately.
Proposition 5.
Proof.
Let be the smooth curve in consisting of all values of evaluated at found from Proposition 3 for , and let be the smooth surface in consisting of all values of evaluated at found from Proposition 4 for and . Proposition 5 implies that any -invariant Ricci soliton on must correspond to a point in where the curve intersects the surface . The following images give various views of an approximation to (in red) and (in blue) which were found using Matlab’s ODE solver.
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/f546cc90-ade0-49f1-88a5-ff8192b72756/InitialView.png)
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/f546cc90-ade0-49f1-88a5-ff8192b72756/L1L2.png)
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/f546cc90-ade0-49f1-88a5-ff8192b72756/L1R.png)
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/f546cc90-ade0-49f1-88a5-ff8192b72756/L2R.png)
Some important observations:
-
•
The intersection we see corresponds to the round sphere with Einstein constant , and is found by setting , and .
-
•
As increases, it appears the corresponding point on the curve approaches the point on the surface corresponding to and . This behaviour appears to resemble that of the ancient ‘pancake’ Ricci flow solution discussed in Section 8.
4 The surface close to ,
The biggest difficulty in proving Theorem 1 is establishing an a priori upper bound for . This part of the proof is achieved by showing that any Ricci soliton that occurs with too large must have non-negative Riemann curvature everywhere. This is a two-step process: the first step is showing that if is too large, then the Riemann curvature must be non-negative between the singular orbit and the unique orbit with , and the data at this point resembles the Gaussian soliton on ; the second shows that if the data at is close to Gaussian, then the Riemann curvature must also be non-negative between the orbit and the singular orbit. This section is devoted to the proof of Theorem 3 below, which achieves the second step. Indeed, Theorem 3 explicitly describes just how Gaussian we need to be at the principal orbit to guarantee curvature non-negativity between this orbit and the orbit. The proof essentially involves an analysis of the behaviour of the soliton equations for values of and close to and respectively.
Theorem 3.
Remark 1.
The value of in the statement of the above theorem arises in relation to the Gaussian soliton on . Indeed, this soliton is the special solution coming out of the orbit with , , , ; when , we have .
The proof of Theorem 3 follows from the three lemmas presented below. The first lemma sets the goal posts, in that it tells us that curvature positivity between the orbit and the orbit is guaranteed if curvature is positive at and is close to .
Lemma 1.
Suppose we have a soliton so that , where . Then the corresponding solutions of the IVP in Proposition 4 are such that the sectional curvatures and do not change sign on .
The curvature positivity condition of the previous lemma is assumed in the hypothesis of Theorem 3, so we turn attention to the task of ensuring that by having the soliton at the orbit quite close to the Gaussian soliton.
Lemma 2.
A Ricci soliton with at the orbit must have for all . In particular, .
Lemma 3.
A Ricci soliton with at time and for all must have .
Proof of Theorem 3.
Using the hypothesis of Theorem 3, Lemma 2 implies that and
for all . Combining with the hypothesis of Theorem 3, Lemma 3 implies that . Lemma 1 then implies that two of the Riemann curvature eigenvalues do not change sign. Since they are non-negative at time , we obtain that these curvatures are non-negative on . We already know from Propositions 1 and 2 that the other two Riemann curvature eigenvalues are non-negative. ∎
We conclude this section with the proof of the three lemmas.
Proof of Lemma 1.
The strategy behind the proof of this lemma is simple enough: check the signs of and using a Taylor series approximation, and show that the error of such an approximation is small enough. The result is obvious if , because then and uniformly. Otherwise, consider the new functions , , , so that the vector satisfies
(14) |
where
and we have
Let be the second-order Taylor series approximation to the solution, so that
Therefore, the function satisfies
where
provided , and
Therefore,
provided , and . We then find that, for ,
(15) |
since . The estimate (15) gives on , so does not change sign for . We also find that, at the principal orbit ,
which is sufficiently close to for by (15). Therefore, these two sectional curvatures do not change sign on . Since , we can also use these estimates to conclude that is close to , so , as required.
∎
Proof of Lemma 2.
A key quantity to consider is because , so that
(16) |
Using (16) and the fact that (since ), we find that
(17) |
Now let be the last time so that
(18) |
Such a must exist because everywhere by Proposition 1, so the estimate holds at time . Then using (16) and (18), we obtain that ; integrating from to gives
(19) |
Now, using the fact that and , where , we find that . Combining this with (19) and the definition of , we obtain that for all . This implies that, for , we have
Since is negative at time , and is negative everywhere, we find that for , so .
5 The curve C for large values of
In our quest to prove that normalised -invariant solitons on have uniformly bounded Riemann curvature and volume (as well as a uniform lower bound on the injectivity radius), it is necessary to find uniform bounds for the numbers discussed in Propositions 3 and 4. This section is dedicated to the proof of Theorem 4 below, which combines with Theorem 3 above to produce the required bound for (recall that a Ricci soliton on with non-negatve Riemann curvature operator must be round).
Theorem 4.
Suppose we have an -invariant gradient shrinking Ricci soliton on of the form (3). Consider the quantities associated to the soliton which satisfy (9), and suppose and . Then there is a principal orbit with , , , , and so that the Riemann curvature is non-negative between this principal orbit and the orbit.
When proving this theorem, it is useful to again change variables with , , , so that
(24) |
Therefore our study essentially reduces to the analysis of the integral curves of the following system of equations:
(25) |
In these co-ordinates, the non-negativity of and is implied by the non-negativity of and (since we already know that ). The solutions we are interested in are part of the two-dimensional unstable manifold of the critical point of (25) at ; the parameter now describes the initial direction of travel through this two-dimensional unstable manifold, and is given precisely by for the solutions of (24).
In this section, we are only interested in evolving the system (24) up until the time , so we can assume that . It is handy to keep track of the evolution of the quantities , and . Indeed, the closeness of to measures the Gaussian structure, positivity of ensures positivity of the last sectional curvature, and is an analogue of a quantity that arises in the construction of the Bryant soliton. We find that
(26) |
where whenever , and whenever and . Therefore, to prove Theorem 4, it suffices to prove the following:
Theorem 5.
Choose . If we have a Ricci soliton with , then the corresponding trajectory of (25) which -limits to includes a point at which , , , and , and so that on the trajectory up until that point.
To proof of this theorem essentially follows by noting the behaviour of the unstable trajectory coming out of as becomes large:
-
•
the trajectory travels from , and gets close to the critical point ;
-
•
the trajectory then travels close to the line of critical points for ;
-
•
the trajectory then breaks away from this line of critical points as begins to grow while stays close to , and we get a point satisfying the conclusion of Theorem 5.
To prove Theorem 5, we proceed working backwards, and start by determining how close the unstable trajectory needs to be to touching the point in order to get the desired conclusion.
Lemma 4.
Choose and suppose our trajectory of (25) includes a point with , , and . Then the trajectory includes a later point at which , , , and . Furthermore, on the trajectory between these two points.
Proof.
We can assume that we have a solution of (25) so that is the time of the first point.
Step One: construct an interval on which we expect to find the required terminal point; find some basic estimates. Define so that
(27) |
On such an interval, (26) implies that
so since , we find that for all . The equations for and in (25) can then be written
(28) |
Now we claim that . Indeed, (28) and the fact that with implies that
(29) |
Estimate (29), coupled with the equation for in (25) and the definition of implies that
(30) |
The estimates (29) and (30) combine with the definition of in (27) to show that .
Now, on , (28) gives us and . Therefore, and , where is some function of . Combining these estimates gives
(31) |
We conclude this step with the assertion that . We show this by demonstrating that all of the other inequalities defining in (27) are strict at time itself. On the interval , , so on since . For any with , it is clear that . On the other hand, if , then by (31) as well. It is clear from (28) that , so it must be the case that .
Step Two: show that and are non-negative on . We have on by (25), since on and . The estimates on provided by the definition of in (27) then imply that on ; this observation coupled with (26) and implies that here as well.
Step Three: find the required point in . Let be the connected interval of times ending at so that . We find from (31) that in . The definition of then implies that in as well. The intermediate value theorem also gives us that there is some at which . Finally, we need to demonstrate that at . At time (which is less than by the above estimate for on ), (31) tells us that . The fact that on implies that the distance between and is greater than . This large amount of time tells us that must get quite close to by the time we land in . Indeed, from (26) and (31) we find that
(32) |
on , while
(33) |
holds on . Estimate (32) tells us that whenever and , we have . The estimate (33) coupled with implies that at time , so it takes no more than time to get . Therefore, is achieved for all times in (including ). ∎
Now we discuss how making large can force our trajectory to be close to the point, in the sense of Lemma 4. This is achieved with the two lemmas below.
Lemma 5.
Proof.
This time, we find it convenient to have our solution of (25) so that the initial point described in the hypothesis of the lemma corresponds to time . This is so we can easily compare our solution to the special solution. Strictly speaking, these functions solve (53), but we now reparametrise them so that they solve (54) instead, with a parametrisation that preserves .
Step One: estimates on the limiting solution. As discussed in Appendix A, we have
(34) |
Combining (34) with (54) gives
(35) |
Also note that (34), combined with Theorem 10 and Proposition 7 in Appendix A implies that
(36) |
It is handy to note that
with
By (35), we find that on so that . Also, on so that . Finally, , because for .
Step Two: closeness to the limiting solution. Define the positive number so that
(37) |
We will now obtain the required estimates by showing that the quantities and are small on the interval . By examining (25) and (54), we compute
(38) |
where provided , and provided . We now use (38) to obtain smallness of on . Let ; (38) implies that
on , as long as , and
on , provided . Therefore, since , we find that
(39) |
Step Three: concluding estimates. We need to check that are all smaller than at time , that , and that on . We claim that . To see this, we use (25) to estimate
on . Therefore, for all , so . On , we estimate that , so for all and . On the other hand, (35) gives on , so that
for . We consequently find
so that for . Therefore, as required.
The smallness of and (39) imply the required estimates for . The estimate for follows from the definition of . The estimates for and follow immediately from (39) and the equation for in (25). Using Proposition 8, we also have
since , while . It is clear that is non-negative on , since and . To show that is non-negative on , note that is non-negative on since , , and (by (25) and the estimates on and ). The non-negativity of and the fact that implies that on by (26).
∎
Lemma 6.
For each , choose an aribtrary . Then the solution of (24) satisfies , , and , where is the Bryant soliton solution discussed in the Appendix. Furthermore, on .
Proof.
Let , , and consider the rescaled functions for . Then these new functions satisfy
with the ‘new’ equal to , i.e., . The discussion of the Bryant soliton on discussed in the Appendix implies the existence of smooth functions of the form (10) so that
Of course, uniformly here. Recall that by using Theorem 10 and Propositon 7, we find that the corresponding and functions satisfy (36). It is also well-known that and are monotonically decreasing functions.
We prove this lemma by comparing to on the interval . We use the notation and to mean the functions from to consisting of components formed by breaking the corresponding sets of functions according to (10). Letting , we find
(40) |
where
and
provided and . One can easily check using the results in the Appendix that on .
Now we let so that
where
This almost-diagonal form of the equations makes it clear that
for all , since is small.
The rest of the proof involves using the smallness of v to obtain the estimates discussed in the statement of the lemma. First note that
In this previous computation, we used the estimates on and from Theorem 10 in the Appendix. Similarly,
To check the closeness of and to , first note that
(41) |
since by Proposition 2. Also, the above estimates for v and imply that , so the equation for implies immediately that for each , so we find
To conclude the proof, we need show that on this interval. It is clear that . To show that , it suffices to show that for all because of the inequality , the equation for in (26) and the fact that , so that is initially positive. From (41), we have
but we also have
by Theorem 10, so we obtain the required estimates. ∎
6 Bounds on curvature at the singular orbit
By Theorems 3 and 4, we find that any Ricci soliton must satisfy . We now construct bounds on and . Fortunately, these bounds are simpler to construct, and can be found without using the (already large) bound on . The bound on is easier to construct once we have a bound on , so we treat first.
Theorem 6.
Suppose the metric of the form (3) is a gradient shrinking Ricci soliton with Einstein constant . Then .
Proof.
We already know from Proposition 5 that , so suppose for the sake of contradiction that . In this case, we claim that
(42) |
To verify (42), note that if , then Proposition 2 implies that , so that on . Then the estimate violates the boundary conditions and , . Since , Proposition 2 again implies that on , so the boundary condtions , and the inequality together imply that .
With (42) in hand, we now claim that
(43) |
The second inequality in (43) is an immediate consequence of the fact that on (follows from Proposition 1), and . The first inequality is a consequence of the fact that
and the observation that . In fact, we know that everywhere (Proposition 1 again), so since everywhere, we rearrange to find so we can estimate further on :
since (follows from (42)) and (Proposition 2). Let , so that and for all . But we can estimate the evolution of :
since for all . If , we integrate to get the following estimate:
which is a contradiction. ∎
Theorem 7.
Suppose the metric of the form (3) is a gradient shrinking Ricci soliton with Einstein constant . Then .
Proof.
Once again consider the non-negative quantity defined by , and note that
(44) |
Also note that
(45) |
because of the estimate , where and . Now Theorem 6 tells us that , so (44) combined with (45) implies that
(46) |
The equation then implies that for all . Now consider the quantities and . Since everywhere, we find that and . We compute
(47) |
and
Since and , we find from (46) and (47) that for all . Therefore
so that , which coincides with , can be no more than .
∎
7 Compactness and uniqueness
We summarise the results for the we have seen so far.
Theorem 8.
An -invariant gradient shrinking Ricci soliton on of the form (3) has , and .
We are now ready to prove the main result of this paper.
Proof of Theorem 1.
We assume for the sake of contradiction that there is no such value of . Then there is a sequence of -invariant solutions to (1) with unbounded Riemann curvature, unbounded volume, or an injectivity radius shrinking to . Theorem 8 implies that are all bounded uniformly, so we can assume that there numbers are all convergent. Propositions 3 and 4 imply that our sequence of solutions converge to another solution. It is clear that the Riemann curvature, volume and injectivity radii all depend continuously on the values of , so we obtain a contradiction. ∎
With Theorem 1 in hand, we discuss how one could prove Conjecture 1. First note that by Propositions 3 and 4, there is a smooth function whose zeroes are precisely the Ricci solitons we aim to classify. By Theorem 8, there is a compact domain that contains all the zeroes of . In fact, we have explicitly described this domain. Therefore, Conjecture 1 would follow with the following steps.
-
1.
Find an explicit open neighbourhood of the canonical metric on on which no other zeroes of occur. This is essentially a quantitative use of the inverse function theorem.
-
2.
Show numerically that there are no zeroes of in . This could be achieved by finding an upper bound for on , discretising the set accordingly, and showing that is sufficiently far away from at each of these finitely-many points using an appropriate numerical ODE solver.
We do not pursue these ideas in this paper, primarily because the bounds we have found are far too large for the numerics described here to provide an answer in a reasonable amount of time. However, we do emphasise that these techniques could be used to resolve Conjecture 1 in the affirmative in ‘finite time’.
8 An -invariant ancient solution on
The sequence of ‘almost Ricci solitons’ on we found by making large appears to have a pancake shape. In this section, we describe a ‘pancake’ -noncollapsed ancient solution to the Ricci flow; it is likely that this is the precise geometric structure that our ‘almost Ricci solitons’ are detecting. It is worth noting that by the recent classification result in [4], this -noncollapsed ancient Ricci flow on cannot be uniformly PIC and weakly PIC2.
We restate Theorem 2 for convenience.
Theorem 9.
There exists a and a -noncollapsed -invariant ancient Ricci flow on with positive Riemann curvature operator which is not isometric to the round shrinking sphere. The group acts on in the obvious way.
The proof of this result is broken up into several steps. Apart from the first step, the construction of this ancient Ricci flow solution is almost identical to that of the Perelman ancient ‘sausage’ solution; the details are available in Chapter 19 of [8].
Proof.
Step One: a sequence of initial Riemannian metrics. For each large , choose an -invariant Riemannian metric on of the form (3) with and
This metric clearly satisfies the smoothness conditions at the singular orbits, but it does fail to be smooth at . However, we can mollify the two functions on so that the resulting functions are smooth. Recall that the Riemann curvatures of this Riemannian metric are
it is clear that after mollification, these curvatures are all non-negative. Since we mollify on , the quantity is uniformly bounded from below, independently of because does not depend on in this region. Also the supremum of all four eigenvalues is uniformly bounded from above ( does not affect the terms, and only makes large so the corresponding curvatures can only get smaller). Therefore,
(48) |
Now we claim that there is a so that is -noncollapsed on all scales , and for all large . To see this, we must show that any geodesic ball on which has volume at least . By the lower bound on the scalar curvature (48), it suffices to find a so that the volume of any geodesic ball is at least whenever . To this end, take an arbitrary geodesic ball and use the manifold decomposition:
where and . Now for each , is isometrically contained in equipped with the standard metric, so we have
(49) |
On the other hand, let be an -invariant Riemannian metric on found by extending on smoothly to a function on with the appropriate smoothness conditions at . Also let be a smooth extension of to a function with domain and appropriate smoothness conditions at . Note that is independent of , and is isometrically embedded in the warped product manifold with metric , where is the standard one-form on . Therefore,
(50) |
Now it is well-known that the right hand side of (49), which is independent of , is strictly positive. Also, we can arrange it so that for all and , so it is also clear that the right hand side of (50) is strictly positive for all , and can be bounded from below by a positive number, independently of .
Now any other geodesic ball of radius contains a geodesic ball of radius which is entirely contained in at least one of or , so the existence of a independent of follows from (49) and (50).
Step Two: Ricci flows from the initial metrics. Since the Riemann curvature of is non-negative, for each there exists a so that the Ricci flow starting at becomes singular for the first time at , and with the round sphere as a singularity model. By (48), there exists a uniform so that for all . Furthermore, by (48) and the fact that each is -noncollapsed on all scales, Theorem 19.52 (no local collapsing) of [8] implies the existence of a so that all of the Ricci flows on are uniformly -noncollapsed on all scales . Note that we can apply this theorem because we can uniformly bound on the time interval independently of ; this is a result of the evolution equations for discussed in the proof of Proposition 1.
For a given small , let be the unique time at which the ratio of largest to smallest Riemann curvature eigenvalues is , and so that the ratio is less than or equal to on .
Step Three: rescaled flows. Consider the rescaled Ricci flows
which start at time , have sectional curvature ratio at time , become singular at time , and are uniformly -noncollapsed on for scales .
We now show that . To see this, note that for small enough , we have
(51) |
thanks to the evolution equation
and the pinching of that we have already established. The estimate (51) coupled with Hamilton’s trace Harnack inequality (Corollary 15.3 in [7]) then implies that for . Therefore,
Note that in the last computation, we have used the following facts:
-
•
For any , , which follows from the Ricci curvature upper bound.
-
•
By making small, we can force to be arbitrarily close to a round sphere, and the scalar curvature of this particular sphere must be because the remaining time until blow up is exactly . Therefore, Myer’s theorem implies that .
The estimate implies that as well, so , because of our uniform estimates on itself.
Step Four: convergence. We are now in a position to take limits. Indeed, the uniform -noncollapsing and curvature bounds are enough to get smooth convergence to an ancient Ricci flow by Theorem 3.10 in [6]; the ancient Ricci flow must have non-negative (hence positive) Riemann curvature. Since was eventually fixed, we can examine the sequence at time to find that the ancient Ricci flow solution is on , but it is not the round sphere. By following the arguments in Chapter 19 of [6], we find that acts via isometries on this ancient solution on in the obvious way. ∎
We conclude this section by observing that, not only is this -invariant ancient solution non-round, but it is not isometric to Perelman’s rotationally-invariant solution either.
Proposition 6.
Let act on in the obvious way. Any -invariant continuous function which is also invariant under an rotation group action must be constant.
Proof.
Let be a continuous function which parametrises the action, and let be a continuous function which parametrises the action. Note that, almost by construction, we have the following:
-
•
and are single points;
-
•
is a copy of ;
-
•
is a copy of ;
-
•
is a copy of for each ;
-
•
is a copy of for each ;
-
•
the function is constant on the level sets for each and .
We claim that there is an so that is constant on and . To see this, note that must be constant on the two submanifolds and . Since these submanifolds are compact, connected and at least one-dimensional, their images under the continuous function must be non-trivial closed subintervals of containing or , respectively. However, since is constant on the level sets of , must actually be constant on the pre-images of these closed subintervals.
Choose so that is the maximal connected subinterval of containing so that is constant on . Similarly, choose so that is the maximal connected subinterval of containing so that is constant on . The argument in the last paragraph shows that and both exist. If , the proof will be complete, so we assume that . Define the pairwise-disjoint connected sets , , . It is clear that . Also note that both and are -invariant because any orbit containing points both in and out of either of the sets would violate the maximality of or . This implies that is -invariant as well. Now and are connected subintervals which must cover ; they must be pairwise disjoint because orbits stay in exactly one of , or . Since and are compact, we find that and are closed, so must be of the form for some . Therefore, must be homeomorphic to both and , which is a contradiction (these two topological spaces have non-isomorphic fundamental groups).
∎
Appendix A The Bryant soliton revisited
The Bryant steady soliton is a rotationally-invariant metric on of the form , where is the standard metric on with Ricci curvature , and is smooth, and can be extended to a smooth and odd function on with . A detailed construction of this metric is given in [6], where the authors also show that the Bryant soliton has positive Riemann curvature everywhere. If we let , and , where is the potential function, then
(52) |
For any , solutions of (52) are invariant under the transformation that sends a function to , so to uniquely specify the Bryant soliton, we need to prescribe the value
Let , and , then
(53) |
The following facts about the Bryant soliton curve are well-known (they are also discussed in [6]):
-
•
;
-
•
;
-
•
for all ;
-
•
.
It is therefore clear that there is a function so that along the Bryant soliton, and this function does not depend on the choice of (this parameter only affects how quickly one travels through the curve ). The following propositions tell us some valuable information about this function.
Proposition 7.
The function satisfies for all . In fact, for all .
Proof.
Consider the system
(54) |
the critical point has a one-dimensional unstable manifold. The part of the unstable manifold lying in consists of exactly those points of the form for . It therefore suffices to show that any point in the unstable manifold with satisfies
(55) |
Our first step in verifying (55) for is to find an so that (55) holds on . To find such an , note that must be smooth close to because it describes an unstable manifold of a hyperbolic critical point of a smooth vector field. By looking at the linearisation of (54) at , we see that the unstable manifold points in the direction of , so . Now the function also satisfies
(56) |
for close to . Using the equalities and , we can write . Then (56) becomes
so . The existence of the required follows.
Consider the quantities and . We find that that whenever , so that whenever and . On the other hand, we have
which is non-positive whenever and . The required estimates follow. ∎
Proposition 8.
We have for . In fact, for all
Proof.
As before, it suffices to consider the unstable trajectory of the system (54) which travels from to . The estimate follows from the fact that , and the observation that whenever the quantity vanishes and , we have
so that is preserved. On the other hand, consider the quantity , so that
Whenever and , we find that . It therefore suffices to show that when . To show this point, we consider the quantity . Then
so that whenever and , we have . Since at the point , we have that when , i.e., our solution curve has when . ∎
We conclude this appendix with some short-time estimates on the original functions solving (53) with .
Theorem 10.
If , then up until time , we have
Proof.
For the estimate, note that follows from curvature positivity of the Bryant soliton. For the lower bound, we use Proposition 7 to conclude that , so that
Now since , we find that , and with . We therefore claim that
(57) |
provided . To see this, note that the solution for corresponds to a curve in space starting at , with increasing and decreasing to the point . We write this curve as for each . The equality then implies that , as long as . The solution of is precisely ; estimate (57) follows for all so that .
We now move on to the estimates. Using and Proposition 7, we find that
provided . Using (52) and the fact that , we find that . One can then verify that . It is clear that , where with , . We can easily show that for all , so that ; solving this implies that . For the lower bound, we estimate
which implies that . ∎
References
- [1] A. Besse. Einstein Manifolds, Springer-Verlag (2007), Berlin.
- [2] C. Böhm, Inhomogeneous Einstein metrics on low-dimensional spheres and other low-dimensional spaces, Invent. Math. 134 (1998), no. 1, 145–176.
- [3] C. Böhm, M. Wang and W. Ziller, A variational approach for compact homogeneous Einstein manifolds, Geom. Funct. Anal. 14 (2004), no. 4, 681–733.
- [4] S. Brendle, P. Daskalopoulos, K. Naff and N. Sesum, Uniqueness of compact ancient solutions to the higher dimensional Ricci flow, arXiv:2102.07180.
- [5] M. Buzano, Initial value problem for cohomogeneity one gradient Ricci solitons, J. Geom. Phys. 61 (2011), no. 6, 1033–1044.
- [6] B. Chow, S. Chu, D. Glickenstein, C. Guenther, J. Isenberg, T. Ivey, D. Knopf, P. Lu, F. Luo and L. Ni, The Ricci flow: techniques and applications. Part I: Geometric aspects, volume 163, Mathematical Surveys and Monographs, American Mathematical Society. Providence, RI (2010).
- [7] B. Chow, S. Chu, D. Glickenstein, C. Guenther, J. Isenberg, T. Ivey, D. Knopf, P. Lu, F. Luo and L. Ni, The Ricci flow: techniques and applications. Part II: Analytic aspects, volume 163, Mathematical Surveys and Monographs, American Mathematical Society. Providence, RI (2010).
- [8] B. Chow, S. Chu, D. Glickenstein, C. Guenther, J. Isenberg, T. Ivey, D. Knopf, P. Lu, F. Luo and L. Ni, The Ricci flow: techniques and applications. Part III: Geometric-analytic aspects, volume 163, Mathematical Surveys and Monographs, American Mathematical Society. Providence, RI (2010).
- [9] A. Dancer and M. Wang, On Ricci solitons of cohomogeneity one, Ann. Global Anal. Geom. 39 (2011), no. 3, 259–292.
- [10] K. Grove and W. Ziller, Cohomogeneity one manifolds with positive Ricci curvature, Invent. Math. 149 (2002), no. 3., 619–646.
- [11] R. Hamilton, Four-manifolds with positive curvature operator, J. Differential Geom. 24 (1986), no. 2, 153–179.
- [12] R. Haslhoffer and R. Müller, A note on the compactness theorem for 4d Ricci shrinkers, Proc. Amer. Math. Soc. 143 (2015), no. 10, 4433–4437.