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

SO(2)×SO(3)SO(2)\times SO(3)-invariant Ricci solitons and ancient flows on 𝕊4\mathbb{S}^{4}

Timothy Buttsworth
Abstract

Consider the standard action of SO(2)×SO(3)SO(2)\times SO(3) on 5=23\mathbb{R}^{5}=\mathbb{R}^{2}\oplus\mathbb{R}^{3}. We establish the existence of a uniform constant 𝒞>0\mathcal{C}>0 so that any SO(2)×SO(3)SO(2)\times SO(3)-invariant Ricci soliton on 𝕊45\mathbb{S}^{4}\subset\mathbb{R}^{5} with Einstein constant 11 must have Riemann curvature and volume bounded by 𝒞\mathcal{C}, and injectivity radius bounded below by 1𝒞\frac{1}{\mathcal{C}}. This observation, coupled with basic numerics, gives strong evidence to suggest that the only SO(2)×SO(3)SO(2)\times SO(3)-invariant Ricci solitons on 𝕊4\mathbb{S}^{4} are round. We also encounter the so-called ‘pancake’ ancient solution of the Ricci flow.

1 Introduction

Let MM be a smooth manifold. A gradient Ricci soliton on MM is a Riemannian metric gg for which there exists a constant λ\lambda and a smooth function u:Mu:M\to\mathbb{R} so that

Ric(g)+Hessg(u)=λg.\text{Ric}(g)+Hess_{g}(u)=\lambda g. (1)

Solutions of (1) arise as self-similar solutions to the well-known Ricci flow

gt=2Ric(g).\displaystyle\frac{\partial g}{\partial t}=-2\text{Ric}(g). (2)

In the quest for new solutions to (1), one often assumes that gg and uu are invariant under a certain group action GG of MM. It is of no use to assume that GG acts transitively on MM, because then gg is homogeneous and uu is constant, so gg 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 GG acts transitively, the next natural step is assuming that GG acts with cohomogeneity one, which means that the generic orbits of the action of GG in MM 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 𝕊4\mathbb{S}^{4} with λ=1\lambda=1 for a pair (g,u)(g,u) which is invariant under the usual cohomogeneity one action of SO(2)×SO(3)SO(2)\times SO(3). We first show a compactness result for solutions of this problem.

Theorem 1.

There exists a 𝒞>0\mathcal{C}>0 so that any SO(2)×SO(3)SO(2)\times SO(3)-invariant solution (g,u)(g,u) of (1) (with λ=1\lambda=1) on 𝕊4\mathbb{S}^{4} has injg1𝒞\text{inj}_{g}\geq\frac{1}{\mathcal{C}}, vol(g)𝒞\text{vol}(g)\leq\mathcal{C}, and Riemann curvature bounded pointwise by 𝒞\mathcal{C}.

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 L2L^{2} norm on the Riemann curvature, rather than the stronger uniform LL^{\infty} 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 SO(2)×SO(3)SO(2)\times SO(3)-invariant solution of (1) on 𝕊4\mathbb{S}^{4} with λ=1\lambda=1 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 2×𝕊2\mathbb{R}^{2}\times\mathbb{S}^{2} and a rescaled product of the Bryant soliton on 3\mathbb{R}^{3} with a flat metric on 𝕊1\mathbb{S}^{1}. 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 κ\kappa-noncollapsed ancient Ricci flow on 𝕊4\mathbb{S}^{4} with positive Riemann curvature.

Theorem 2.

There exists a κ>0\kappa>0 and a κ\kappa-noncollapsed, non-round SO(2)×SO(3)SO(2)\times SO(3)-invariant ancient solution to the Ricci flow on 𝕊4\mathbb{S}^{4} 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 gg and the function uu 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 SO(2)×SO(3)SO(2)\times SO(3), 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 𝕊4\mathbb{S}^{4} and 𝕊2×𝕊2\mathbb{S}^{2}\times\mathbb{S}^{2} which are invariant under the obvious action of SO(2)×SO(3)SO(2)\times SO(3), so for this section only, we consider both of these four-dimensional manifolds.

2.1 The boundary value problem

Under the action of SO(2)×SO(3)SO(2)\times SO(3), the principal orbits of 𝕊4\mathbb{S}^{4} or 𝕊2×𝕊2\mathbb{S}^{2}\times\mathbb{S}^{2} are product spheres 𝕊1×𝕊2\mathbb{S}^{1}\times\mathbb{S}^{2}. In the case of 𝕊4\mathbb{S}^{4}, the two singular orbits are one copy of 𝕊1\mathbb{S}^{1} and one copy of 𝕊2\mathbb{S}^{2}, whereas for 𝕊2×𝕊2\mathbb{S}^{2}\times\mathbb{S}^{2}, both singular orbits are copies of 𝕊2\mathbb{S}^{2}. Up to diffeomorphism, any SO(2)×SO(3)SO(2)\times SO(3)-invariant Riemannian metric on 𝕊4\mathbb{S}^{4} or 𝕊2×𝕊2\mathbb{S}^{2}\times\mathbb{S}^{2} has the form

dt2+f12dθ2+f22Q,dt^{2}+f_{1}^{2}d\theta^{2}+f_{2}^{2}Q, (3)

where fi:(0,T)f_{i}:(0,T)\to\mathbb{R} are smooth and positive functions, dθd\theta is the standard one-form on 𝕊1\mathbb{S}^{1}, and QQ is the round metric of unit Ricci curvature on 𝕊2\mathbb{S}^{2}. In order for the Riemannian metric gg to close up smoothly at the singular orbits, the functions f1f_{1} and f2f_{2} must be smoothly extendable to functions on (T,2T)(-T,2T) so that

f1(0)>0and f1(t) is even about t=0,f2(0)=1and f2(t) is odd about t=0,f1(T)=1and f1(t) is odd about t=T,f2(T)>0and f2(t) is even about t=T,\displaystyle\begin{split}f_{1}(0)>0\ \text{and $f_{1}(t)$ is even about $t=0$},\qquad f_{2}^{\prime}(0)=1\ \text{and $f_{2}(t)$ is odd about $t=0$},\\ f_{1}^{\prime}(T)=-1\ \text{and $f_{1}(t)$ is odd about $t=T$},\qquad f_{2}(T)>0\ \text{and $f_{2}(t)$ is even about $t=T$},\\ \end{split} (4)

in the case of 𝕊4\mathbb{S}^{4}, and so that

f1(0)=1and f1(t) is odd about t=0,f2(0)>0and f2(t) is even about t=0,f1(T)=1and f1(t) is odd about t=T,f2(T)>0and f2(t) is even about t=T,\displaystyle\begin{split}f_{1}^{\prime}(0)=1\ \text{and $f_{1}(t)$ is odd about $t=0$},\qquad f_{2}(0)>0\ \text{and $f_{2}(t)$ is even about $t=0$},\\ f_{1}^{\prime}(T)=-1\ \text{and $f_{1}(t)$ is odd about $t=T$},\qquad f_{2}(T)>0\ \text{and $f_{2}(t)$ is even about $t=T$},\\ \end{split} (5)

in the case of 𝕊2×𝕊2\mathbb{S}^{2}\times\mathbb{S}^{2}. At any point in a principal orbit, let {e1,e2,e3}\{e_{1},e_{2},e_{3}\} be an orthonormal basis adapted to the standard product metric dθ2+Qd\theta^{2}+Q on 𝕊1×𝕊2\mathbb{S}^{1}\times\mathbb{S}^{2}. Then in the basis {te1,te2,te3,e2e3,e1e2,e1e3}\{\partial_{t}\wedge e_{1},\partial_{t}\wedge e_{2},\partial_{t}\wedge e_{3},e_{2}\wedge e_{3},e_{1}\wedge e_{2},e_{1}\wedge e_{3}\}, we compute (cf. [10]) the Riemann curvature operator:

=(f1′′f1000000f2′′f2000000f2′′f20000001f22f22000000f1f2f1f2000000f1f2f1f2).\displaystyle\mathcal{R}=\begin{pmatrix}-\frac{f_{1}^{\prime\prime}}{f_{1}}&0&0&0&0&0\\ 0&-\frac{f_{2}^{\prime\prime}}{f_{2}}&0&0&0&0\\ 0&0&-\frac{f_{2}^{\prime\prime}}{f_{2}}&0&0&0\\ 0&0&0&\frac{1-f_{2}^{\prime 2}}{f_{2}^{2}}&0&0\\ 0&0&0&0&-\frac{f_{1}^{\prime}f_{2}^{\prime}}{f_{1}f_{2}}&0\\ 0&0&0&0&0&-\frac{f_{1}^{\prime}f_{2}^{\prime}}{f_{1}f_{2}}\end{pmatrix}. (6)

Suppose we have an SO(2)×SO(3)SO(2)\times SO(3)-invariant metric of the form (3) which is a gradient shrinking Ricci soliton with an SO(2)×SO(3)SO(2)\times SO(3)-invariant potential function uu. Then uu depends only on the tt parameter, and can be smoothly extended to a function on (T,2T)(-T,2T) so that

u(t)is even about both t=0 and t=T,\displaystyle u(t)\ \text{is even about both $t=0$ and $t=T$}, (7)

and the functions (f1,f2,u)(f_{1},f_{2},u) satisfy (4) or (5), alongside the Ricci soliton equation (1) on (0,T)(0,T) which becomes

f1′′f12f2′′f2+u′′=λ,f1′′f12f1f2f1f2+uf1f1=λ,f2′′f2f1f2f1f2+1f22f22+uf2f2=λ.\displaystyle\begin{split}-\frac{f_{1}^{\prime\prime}}{f_{1}}-2\frac{f_{2}^{\prime\prime}}{f_{2}}+u^{\prime\prime}&=\lambda,\\ -\frac{f_{1}^{\prime\prime}}{f_{1}}-2\frac{f_{1}^{\prime}f_{2}^{\prime}}{f_{1}f_{2}}+u^{\prime}\frac{f_{1}^{\prime}}{f_{1}}&=\lambda,\\ -\frac{f_{2}^{\prime\prime}}{f_{2}}-\frac{f_{1}^{\prime}f_{2}^{\prime}}{f_{1}f_{2}}+\frac{1-f_{2}^{\prime 2}}{f_{2}^{2}}+u^{\prime}\frac{f_{2}^{\prime}}{f_{2}}&=\lambda.\end{split} (8)

We find it useful to set λ=1\lambda=1 and introduce the new functions Li=fifiL_{i}=\frac{f_{i}^{\prime}}{f_{i}}, ξ=L1+2L2u\xi=L_{1}+2L_{2}-u^{\prime}, R=1f2R=\frac{1}{f_{2}}, so the Ricci soliton equations in these variables are

ξ=L122L221,L1=ξL11,L2=ξL2+R21,R=L2R.\displaystyle\begin{split}\xi^{\prime}=-L_{1}^{2}-2L_{2}^{2}-1,\\ L_{1}^{\prime}=-\xi L_{1}-1,\\ L_{2}^{\prime}=-\xi L_{2}+R^{2}-1,\\ R^{\prime}=-L_{2}R.\end{split} (9)

Using the boundary conditions, a solution of (9) on (0,T)(0,T) uniquely determines a solution of (8). Note that in these co-ordinates, the Riemann curvature eigenvalues for a Ricci soliton are:

  • ξL1+1L12\xi L_{1}+1-L_{1}^{2} (we know this must be non-negative for a soliton by Proposition 1 below);

  • R2L22R^{2}-L_{2}^{2} (we know this must be non-negative for a soliton by Proposition 2 below);

  • L1L2-L_{1}L_{2} (occurs twice); and

  • ξL2+1R2L22\xi L_{2}+1-R^{2}-L_{2}^{2} (occurs twice).

2.2 An initial step towards compactness: the maximum principle

If we were lucky enough to know a priori that any SO(2)×SO(3)SO(2)\times SO(3)-invariant shrinking soliton on 𝕊4\mathbb{S}^{4} had to have non-negative Riemann curvature operator, then Hamilton’s pinching results for 44-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 SO(2)×SO(3)SO(2)\times SO(3)-invariant gradient shrinking Ricci soliton (g,u)(g,u) on 𝕊4\mathbb{S}^{4} or 𝕊2×𝕊2\mathbb{S}^{2}\times\mathbb{S}^{2} of the form (3), the quantity f1′′f1\frac{-f_{1}^{\prime\prime}}{f_{1}} is non-negative everywhere.

Proof.

For a given point on a principal orbit, consider the selfdual/anti-selfdual basis

{te1+e2e3,te2+e1e3,e3t+e1e2,te1e2e3,te2e1e3,e3te1e2}.\displaystyle\{\partial_{t}\wedge e_{1}+e_{2}\wedge e_{3},\partial_{t}\wedge e_{2}+e_{1}\wedge e_{3},e_{3}\wedge\partial_{t}+e_{1}\wedge e_{2},\partial_{t}\wedge e_{1}-e_{2}\wedge e_{3},\partial_{t}\wedge e_{2}-e_{1}\wedge e_{3},e_{3}\wedge\partial_{t}-e_{1}\wedge e_{2}\}.

In this basis, the Riemann curvature operator is given by

=(ABBTA)\displaystyle\mathcal{R}=\begin{pmatrix}A&B\\ B^{T}&A\end{pmatrix}

where

2A=(f1′′f1+1(f2)2f22000f2′′f2f1f2f1f2000f2′′f2f1f2f1f2), 2B=(f1′′f11(f2)2f22000f2′′f2+f1f2f1f2000f2′′f2+f1f2f1f2).\displaystyle 2A=\begin{pmatrix}-\frac{f_{1}^{\prime\prime}}{f_{1}}+\frac{1-(f_{2}^{\prime})^{2}}{f_{2}^{2}}&0&0\\ 0&-\frac{f_{2}^{\prime\prime}}{f_{2}}-\frac{f_{1}^{\prime}f_{2}^{\prime}}{f_{1}f_{2}}&0\\ 0&0&-\frac{f_{2}^{\prime\prime}}{f_{2}}-\frac{f_{1}^{\prime}f_{2}^{\prime}}{f_{1}f_{2}}\end{pmatrix},\ 2B=\begin{pmatrix}-\frac{f_{1}^{\prime\prime}}{f_{1}}-\frac{1-(f_{2}^{\prime})^{2}}{f_{2}^{2}}&0&0\\ 0&-\frac{f_{2}^{\prime\prime}}{f_{2}}+\frac{f_{1}^{\prime}f_{2}^{\prime}}{f_{1}f_{2}}&0\\ 0&0&-\frac{f_{2}^{\prime\prime}}{f_{2}}+\frac{f_{1}^{\prime}f_{2}^{\prime}}{f_{1}f_{2}}\end{pmatrix}.

After applying the Uhlenbeck trick, Hamilton shows in [11] that, under the Ricci flow, the curvatures satisfy the evolution equations

Aτ=Δg(τ)A+A2+BTB+2A#,Bτ=Δg(τ)B+AB+BA+2B#.\displaystyle\frac{\partial A}{\partial\tau}=\Delta_{g(\tau)}A+A^{2}+B^{T}B+2A^{\#},\qquad\frac{\partial B}{\partial\tau}=\Delta_{g(\tau)}B+AB+BA+2B^{\#}.

Since the second and third diagonal entries of both AA and BB are identical, the first entries of A#A^{\#} and B#B^{\#} are non-negative. We therefore find that the first eigenvalue of the matrix A+BA+B 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 A+BA+B, which is f1′′f1-\frac{f_{1}^{\prime\prime}}{f_{1}}, must be non-negative everywhere. ∎

Proposition 2.

For any SO(2)×SO(3)SO(2)\times SO(3)-invariant gradient shrinking Ricci soliton on 𝕊4\mathbb{S}^{4} or 𝕊2×𝕊2\mathbb{S}^{2}\times\mathbb{S}^{2} of the form (3), the quantity 1f22f22\frac{1-f_{2}^{\prime 2}}{f_{2}^{2}} is non-negative everywhere. More generally, suppose we have an SO(2)×SO(3)SO(2)\times SO(3)-invariant Riemannian metric of the form (3), as well as two points r1,r2[0,T]r_{1},r_{2}\in[0,T] so that:

  • 1f22f220\frac{1-f_{2}^{\prime 2}}{f_{2}^{2}}\geq 0 at r1,r2r_{1},r_{2}; and

  • the metric satisfies the Ricci soliton equations on (r1,r2)(r_{1},r_{2}).

Then 1f22f220\frac{1-f_{2}^{\prime 2}}{f_{2}^{2}}\geq 0 on (r1,r2)(r_{1},r_{2}).

Proof.

Clearly it suffices to show that f2[1,1]f_{2}^{\prime}\in[-1,1] on [r1,r2][r_{1},r_{2}]. Differentiating and using the equations of (8), we find

0\displaystyle 0 =f2′′′+(f1f2f1+1f22f2)+(u′′λ)f2+uf2′′\displaystyle=-f_{2}^{\prime\prime\prime}+(-\frac{f_{1}^{\prime}f_{2}^{\prime}}{f_{1}}+\frac{1-f_{2}^{\prime 2}}{f_{2}})^{\prime}+(u^{\prime\prime}-\lambda)f_{2}^{\prime}+u^{\prime}f_{2}^{\prime\prime}
=f2′′′+f12f2f12f1f2′′f1+2f2f2′′f2+f2(f221)f22+2f2′′f2f2+uf2′′.\displaystyle=-f_{2}^{\prime\prime\prime}+\frac{f_{1}^{\prime 2}f_{2}^{\prime}}{f_{1}^{2}}-\frac{f_{1}^{\prime}f_{2}^{\prime\prime}}{f_{1}}+\frac{-2f_{2}^{\prime}f_{2}^{\prime\prime}f_{2}+f_{2}^{\prime}(f_{2}^{\prime 2}-1)}{f_{2}^{2}}+2\frac{f_{2}^{\prime\prime}f_{2}^{\prime}}{f_{2}}+u^{\prime}f_{2}^{\prime\prime}.

Therefore, if f2>1f_{2}^{\prime}>1 somewhere on (r1,r2)(r_{1},r_{2}), then there must be a point in (r1,r2)(r_{1},r_{2}) with f2>1f_{2}^{\prime}>1, f2′′=0f_{2}^{\prime\prime}=0 and f2′′′0f_{2}^{\prime\prime\prime}\leq 0. At this point, we find

0\displaystyle 0 =f2′′′+f12f2f12+f2(f221)f22>0,\displaystyle=-f_{2}^{\prime\prime\prime}+\frac{f_{1}^{\prime 2}f_{2}^{\prime}}{f_{1}^{2}}+\frac{f_{2}^{\prime}(f_{2}^{\prime 2}-1)}{f_{2}^{2}}>0,

a contradiction. We obtain a similar contradiction if there were a point with f2<1f_{2}^{\prime}<-1. A metric which is a Ricci soliton everywhere has f2[1,1]f_{2}^{\prime}\in[-1,1] at 0 and TT for both 𝕊4\mathbb{S}^{4} and 𝕊2×𝕊2\mathbb{S}^{2}\times\mathbb{S}^{2} by (4) and (5), so the claim follows. ∎

3 The shooting problem for 𝕊4\mathbb{S}^{4}: 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 t=0,Tt=0,T, and examine how the solutions meet at a specified principal orbit. In particular, we will examine how the solutions meet at an orbit where ξ=0\xi=0, 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 𝕊1\mathbb{S}^{1} orbit (t=0t=0). Note that, by (4) and (7), there must be smooth functions η0,η1,η2,η3\eta_{0},\eta_{1},\eta_{2},\eta_{3} so that

ξ(t)=2t+η0(t),L1(t)=η1(t),L2(t)=1t+η2(t),R(t)=1t+η3(t),\displaystyle\xi(t)=\frac{2}{t}+\eta_{0}(t),\qquad L_{1}(t)=\eta_{1}(t),\qquad L_{2}(t)=\frac{1}{t}+\eta_{2}(t),\qquad R(t)=\frac{1}{t}+\eta_{3}(t), (10)

and ηi(0)=0\eta_{i}(0)=0 for all i=0,1,2,3i=0,1,2,3. It turns out that whenever solving (9) subject to (10) close to t=0t=0, solutions are uniquely determined by the number

δ1=η3(0)=limt0(R(t)1tt).\displaystyle\delta_{1}=\eta_{3}^{\prime}(0)=\lim_{t\to 0}\left(\frac{R(t)-\frac{1}{t}}{t}\right). (11)
Proposition 3.

For each value of δ1\delta_{1}\in\mathbb{R}, there exists a unique solution to the initial value problem of (9), subject to (10) and (11). The solution can be extended smoothly to a point with ξ=0\xi=0, and ξ1(0)\xi^{-1}(0) depends smoothly on δ1\delta_{1}. For i=0,1,2,3i=0,1,2,3, the quantity ηi(t)t\frac{\eta_{i}(t)}{t} depends continuously on δ1\delta_{1} and t[0,ξ1(0)]t\in[0,\xi^{-1}(0)].

The proof of Proposition 3 essentially follows from the techniques discussed in [5]. We can similarly examine the initial value problem at the 𝕊2\mathbb{S}^{2} orbit. This time, (4) and (7) imply that there must be smooth functions η0,η1,η2,η3\eta_{0},\eta_{1},\eta_{2},\eta_{3} so that

ξ(Tt)=1t+η0(t),L1(Tt)=1t+η1(t),L2(Tt)=η2(t),R(Tt)=R(T)+η3(t),\displaystyle-\xi(T-t)=\frac{1}{t}+\eta_{0}(t),\qquad-L_{1}(T-t)=\frac{1}{t}+\eta_{1}(t),\qquad-L_{2}(T-t)=\eta_{2}(t),\qquad R(T-t)=R(T)+\eta_{3}(t), (12)

where again ηi(0)=0\eta_{i}(0)=0 for i=0,1,2,3i=0,1,2,3. This time, solutions are uniquely determined by

δ2=η0(0),δ3=R(T).\displaystyle\delta_{2}=\eta_{0}^{\prime}(0),\qquad\delta_{3}=R(T). (13)
Proposition 4.

For each value of (δ2,δ3)2(\delta_{2},\delta_{3})\in\mathbb{R}^{2}, there exists a unique solution to the initial value problem of (9), subject to (12) and (13). The solution can be extended smoothly to a point with ξ=0\xi=0, and ξ1(0)\xi^{-1}(0) depends smoothly on (δ2,δ3)(\delta_{2},\delta_{3}). For i=0,1,2,3i=0,1,2,3, the quantity ηi(t)t\frac{\eta_{i}(t)}{t} depends continuously on δ2,δ3\delta_{2},\delta_{3} and t[ξ1(0),T]t\in[\xi^{-1}(0),T].

The main idea behind the proof of Theorem 1 is to provide bounds on the possible values that δ1,δ2,δ3\delta_{1},\delta_{2},\delta_{3} can achieve. Fortunately, we can find some of these bounds immediately.

Proposition 5.

If the solutions to the IVPs found in Propositions 3 and 4 coincide at time ξ1(0)\xi^{-1}(0), then δ10\delta_{1}\geq 0, δ21\delta_{2}\geq-1 and δ30\delta_{3}\geq 0.

Proof.

The δ3\delta_{3} bound is obvious because limt0R(t)=\lim_{t\to 0}R(t)=\infty, and R=L2RR^{\prime}=-L_{2}R. For δ1\delta_{1}, note that R2L220R^{2}-L_{2}^{2}\geq 0 by Proposition 2, so using (10) and (11), we find 0limt0(R(t)2L2(t)2)=2η3(0)2η2(0)0\leq\lim_{t\to 0}(R(t)^{2}-L_{2}(t)^{2})=2\eta_{3}^{\prime}(0)-2\eta_{2}^{\prime}(0). But by using (9) we see that η2(0)=2η3(0)\eta_{2}^{\prime}(0)=-2\eta_{3}^{\prime}(0), so 06δ10\leq 6\delta_{1}. For δ2\delta_{2}, we can use Proposition 1 to find that ξL1+1L120\xi L_{1}+1-L_{1}^{2}\geq 0 everywhere; the result then follows similarly to the analogous result for δ1\delta_{1}. ∎

Let CC be the smooth curve in 3\mathbb{R}^{3} consisting of all values of (L1,L2,R)(L_{1},L_{2},R) evaluated at ξ1(0)\xi^{-1}(0) found from Proposition 3 for δ10\delta_{1}\geq 0, and let SS be the smooth surface in 3\mathbb{R}^{3} consisting of all values of (L1,L2,R)(L_{1},L_{2},R) evaluated at ξ1(0)\xi^{-1}(0) found from Proposition 4 for δ21\delta_{2}\geq-1 and δ30\delta_{3}\geq 0. Proposition 5 implies that any SO(2)×SO(3)SO(2)\times SO(3)-invariant Ricci soliton on 𝕊4\mathbb{S}^{4} must correspond to a point in 3\mathbb{R}^{3} where the curve CC intersects the surface SS. The following images give various views of an approximation to CC (in red) and SS (in blue) which were found using Matlab’s ODE solver.

[Uncaptioned image][Uncaptioned image][Uncaptioned image][Uncaptioned image]

Some important observations:

  • The intersection we see corresponds to the round sphere with Einstein constant 11, and is found by setting δ1=118\delta_{1}=\frac{1}{18}, δ2=79\delta_{2}=-\frac{7}{9} and δ3=13\delta_{3}=\frac{1}{\sqrt{3}}.

  • As δ1\delta_{1} increases, it appears the corresponding point on the curve CC approaches the point on the surface SS corresponding to δ2=1\delta_{2}=-1 and δ3=1\delta_{3}=1. This behaviour appears to resemble that of the ancient ‘pancake’ Ricci flow solution discussed in Section 8.

4 The surface SS close to δ2=1\delta_{2}=-1, δ3=1\delta_{3}=1

The biggest difficulty in proving Theorem 1 is establishing an a priori upper bound for δ1\delta_{1}. This part of the proof is achieved by showing that any Ricci soliton that occurs with δ1\delta_{1} too large must have non-negative Riemann curvature everywhere. This is a two-step process: the first step is showing that if δ1\delta_{1} is too large, then the Riemann curvature must be non-negative between the 𝕊1\mathbb{S}^{1} singular orbit and the unique orbit with ξ=10\xi=10, and the data at this point resembles the Gaussian soliton on 2×𝕊2\mathbb{R}^{2}\times\mathbb{S}^{2}; the second shows that if the data at ξ=10\xi=10 is close to Gaussian, then the Riemann curvature must also be non-negative between the ξ=10\xi=10 orbit and the 𝕊2\mathbb{S}^{2} 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 ξ=10\xi=10 principal orbit to guarantee curvature non-negativity between this orbit and the 𝕊2\mathbb{S}^{2} orbit. The proof essentially involves an analysis of the behaviour of the soliton equations for values of δ2\delta_{2} and δ3\delta_{3} close to 1-1 and 11 respectively.

Theorem 3.

Suppose we have an SO(2)×SO(3)SO(2)\times SO(3)-invariant gradient shrinking Ricci soliton on 𝕊4\mathbb{S}^{4} of the form (3). Let B2=10146B_{2}=10^{-146}, and consider the quantities ξ,L1,L2,R\xi,L_{1},L_{2},R associated to the soliton which satisfy (9). If there is a principal orbit with ξ=10\xi=10, |L1+15+26|B2\left|L_{1}+\frac{1}{5+\sqrt{26}}\right|\leq B_{2}, |R1|B2\left|R-1\right|\leq B_{2}, |L2|B2\left|L_{2}\right|\leq B_{2}, and with 0\mathcal{R}\geq 0, then 0\mathcal{R}\geq 0 between this principal orbit and the 𝕊2\mathbb{S}^{2} orbit (\mathcal{R} is described in (6)).

Remark 1.

The value of 15+26-\frac{1}{5+\sqrt{26}} in the statement of the above theorem arises in relation to the Gaussian soliton on 2×𝕊2\mathbb{R}^{2}\times\mathbb{S}^{2}. Indeed, this soliton is the special solution coming out of the 𝕊2\mathbb{S}^{2} orbit with ξ(Tt)=1tt-\xi(T-t)=\frac{1}{t}-t, L1(Tt)=1t-L_{1}(T-t)=\frac{1}{t}, L2(Tt)=0L_{2}(T-t)=0, R(Tt)=1R(T-t)=1; when ξ=10\xi=10, we have L1=15+26L_{1}=-\frac{1}{5+\sqrt{26}}.

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 ξ=10\xi=10 orbit and the 𝕊2\mathbb{S}^{2} orbit is guaranteed if curvature is positive at ξ=10\xi=10 and (δ2,δ3)(\delta_{2},\delta_{3}) is close to (1,1)(-1,1).

Lemma 1.

Suppose we have a soliton so that (δ2,δ3)[1,1+B1]×[1B1,1+B1](\delta_{2},\delta_{3})\in[-1,-1+B_{1}]\times[1-B_{1},1+B_{1}], where B1=1070B_{1}=10^{-70}. Then the corresponding solutions of the IVP in Proposition 4 are such that the sectional curvatures L1L2-L_{1}L_{2} and ξL2+1R2L22\xi L_{2}+1-R^{2}-L_{2}^{2} do not change sign on [ξ1(10),T)[\xi^{-1}(10),T).

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 max{|δ31|,δ2+1}B1\max\{\left|\delta_{3}-1\right|,\delta_{2}+1\}\leq B_{1} by having the soliton at the ξ=10\xi=10 orbit quite close to the Gaussian soliton.

Lemma 2.

A Ricci soliton with |R1|,|L2|B2\left|R-1\right|,\left|L_{2}\right|\leq B_{2} at the ξ1(10)\xi^{-1}(10) orbit must have max{|R(t)1|,|L2(t)|}B1\max\{\left|R(t)-1\right|,\left|L_{2}(t)\right|\}\leq B_{1} for all t(ξ1(10),T)t\in(\xi^{-1}(10),T). In particular, |δ31|B1\left|\delta_{3}-1\right|\leq B_{1}.

Lemma 3.

A Ricci soliton with |L1+15+26|B2\left|L_{1}+\frac{1}{5+\sqrt{26}}\right|\leq B_{2} at time ξ1(10)\xi^{-1}(10) and max{|R(t)1|,|L2(t)|}B1\max\{\left|R(t)-1\right|,\left|L_{2}(t)\right|\}\leq B_{1} for all t(ξ1(10),T)t\in(\xi^{-1}(10),T) must have 0δ2+1B10\leq\delta_{2}+1\leq B_{1}.

Proof of Theorem 3.

Using the hypothesis of Theorem 3, Lemma 2 implies that |δ31|B1\left|\delta_{3}-1\right|\leq B_{1} and

max{|R(t)1|,|L2(t)|}B1\max\{\left|R(t)-1\right|,\left|L_{2}(t)\right|\}\leq B_{1}

for all t(ξ1(10),T)t\in(\xi^{-1}(10),T). Combining with the hypothesis of Theorem 3, Lemma 3 implies that 0δ2+1B10\leq\delta_{2}+1\leq B_{1}. Lemma 1 then implies that two of the Riemann curvature eigenvalues do not change sign. Since they are non-negative at time ξ1(10)\xi^{-1}(10), we obtain that these curvatures are non-negative on (ξ1(10),T)(\xi^{-1}(10),T). 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 L1L2-L_{1}L_{2} and ξL2+1R2L22\xi L_{2}+1-R^{2}-L_{2}^{2} using a Taylor series approximation, and show that the error of such an approximation is small enough. The result is obvious if δ3=1\delta_{3}=1, because then L2=0L_{2}=0 and R=1R=1 uniformly. Otherwise, consider the new functions w(t)=ξ(Tt)1tw(t)=-\xi(T-t)-\frac{1}{t}, x(t)=L1(Tt)+ξ(Tt)x(t)=-L_{1}(T-t)+\xi(T-t), y(t)=L2(Tt)δ31y(t)=\frac{-L_{2}(T-t)}{\delta_{3}-1}, z(t)=R(Tt)1δ31z(t)=\frac{R(T-t)-1}{\delta_{3}-1} so that the vector u=(w,x,y,z)u=(w,x,y,z) satisfies

u+Atu=Bu+C(u)+D\displaystyle u^{\prime}+\frac{A}{t}u=Bu+C(u)+D (14)

where

A=(2200010000100000),B=(0000000000020010),C(u)=((x+w)22(δ31)2y2x2+xw+2(δ31)2y2wy+(δ31)z2(δ31)yz),D=(1000),\displaystyle A=\begin{pmatrix}2&2&0&0\\ 0&-1&0&0\\ 0&0&1&0\\ 0&0&0&0\end{pmatrix},\ B=\begin{pmatrix}0&0&0&0\\ 0&0&0&0\\ 0&0&0&2\\ 0&0&-1&0\end{pmatrix},\ C(u)=\begin{pmatrix}-(x+w)^{2}-2(\delta_{3}-1)^{2}y^{2}\\ x^{2}+xw+2(\delta_{3}-1)^{2}y^{2}\\ -wy+(\delta_{3}-1)z^{2}\\ -(\delta_{3}-1)yz\end{pmatrix},\ D=\begin{pmatrix}-1\\ 0\\ 0\\ 0\end{pmatrix},

and we have

u(0)=(0,0,0,1),u(0)=(δ2,3δ2+12,δ3+12,0),u′′(0)=(0,0,0,δ3(δ3+1)2).\displaystyle u(0)=(0,0,0,1),\qquad u^{\prime}(0)=(\delta_{2},-\frac{3\delta_{2}+1}{2},\frac{\delta_{3}+1}{2},0),\qquad u^{\prime\prime}(0)=(0,0,0,-\frac{\delta_{3}(\delta_{3}+1)}{2}).

Let u2=u(0)+u(0)t+u′′(0)t22u_{2}=u(0)+u^{\prime}(0)t+\frac{u^{\prime\prime}(0)t^{2}}{2} be the second-order Taylor series approximation to the solution, so that

u2+Atu2Bu2C(u2)\displaystyle u_{2}^{\prime}+\frac{A}{t}u_{2}-Bu_{2}-C(u_{2}) =D+((δ2+1)24+(δ31)2(δ3+1)22(3δ2+1)(δ2+1)4(δ31)2(δ3+1)22(δ3+1)(δ3+δ2)2+(δ31)δ3(δ3+1)20)t2+(000(δ31)(δ3+1)2δ38)t3+(00(1δ3)δ32(1+δ3)2160)t4\displaystyle=D+\begin{pmatrix}\frac{(\delta_{2}+1)^{2}}{4}+\frac{(\delta_{3}-1)^{2}(\delta_{3}+1)^{2}}{2}\\ -\frac{(3\delta_{2}+1)(\delta_{2}+1)}{4}-\frac{(\delta_{3}-1)^{2}(\delta_{3}+1)^{2}}{2}\\ \frac{(\delta_{3}+1)(\delta_{3}+\delta_{2})}{2}+\frac{(\delta_{3}-1)\delta_{3}(\delta_{3}+1)}{2}\\ 0\end{pmatrix}t^{2}+\begin{pmatrix}0\\ 0\\ 0\\ \frac{(\delta_{3}-1)(\delta_{3}+1)^{2}\delta_{3}}{8}\end{pmatrix}t^{3}+\begin{pmatrix}0\\ 0\\ \frac{(1-\delta_{3})\delta_{3}^{2}(1+\delta_{3})^{2}}{16}\\ 0\end{pmatrix}t^{4}

Therefore, the function v=uu2v=u-u_{2} satisfies

v+Atv=Bv+F(t)v+C(v)+E2t2+E3t3+E4t4,\displaystyle v^{\prime}+\frac{A}{t}v=Bv+F(t)v+C(v)+E_{2}t^{2}+E_{3}t^{3}+E_{4}t^{4},
v(0)=0,v(0)=0,v′′(0)=0,\displaystyle v(0)=0,\qquad v^{\prime}(0)=0,\qquad v^{\prime\prime}(0)=0,

where

|E2|4max{|δ2+1|,|δ31|},|E3|max{|δ2+1|,|δ31|},|E4|13max{|δ2+1|,|δ31|},\displaystyle\left|E_{2}\right|_{\infty}\leq 4\max\{\left|\delta_{2}+1\right|,\left|\delta_{3}-1\right|\},\qquad\left|E_{3}\right|_{\infty}\leq\max\{\left|\delta_{2}+1\right|,\left|\delta_{3}-1\right|\},\qquad\left|E_{4}\right|_{\infty}\leq\frac{1}{3}\max\{\left|\delta_{2}+1\right|,\left|\delta_{3}-1\right|\},

provided max{|δ2+1|,|δ31|}106\max\{\left|\delta_{2}+1\right|,\left|\delta_{3}-1\right|\}\leq 10^{-6}, and

F(t)\displaystyle F(t) =(000000000002(δ31)00(δ31)0)\displaystyle=\begin{pmatrix}0&0&0&0\\ 0&0&0&0\\ 0&0&0&2(\delta_{3}-1)\\ 0&0&-(\delta_{3}-1)&0\end{pmatrix}
+(1+δ21+δ22(δ31)2(δ3+1)03δ2+122δ212(δ31)2(δ3+1)0δ3+120δ20000(δ31)(δ3+1)2)t+(00000000000(δ31)(δ3+12)δ300(δ31)(δ3+14)δ30)t2.\displaystyle+\begin{pmatrix}1+\delta_{2}&1+\delta_{2}&-2(\delta_{3}-1)^{2}(\delta_{3}+1)&0\\ -\frac{3\delta_{2}+1}{2}&-2\delta_{2}-1&2(\delta_{3}-1)^{2}(\delta_{3}+1)&0\\ -\frac{\delta_{3}+1}{2}&0&-\delta_{2}&0\\ 0&0&0&\frac{-(\delta_{3}-1)(\delta_{3}+1)}{2}\end{pmatrix}t+\begin{pmatrix}0&0&0&0\\ 0&0&0&0\\ 0&0&0&-(\delta_{3}-1)(\frac{\delta_{3}+1}{2})\delta_{3}\\ 0&0&(\delta_{3}-1)(\frac{\delta_{3}+1}{4})\delta_{3}&0\end{pmatrix}t^{2}.

Therefore,

|v(t)||v(t)|(2.1t+1t+2.1)+max{|δ2+1|,|δ31|}(4t2+t3+13t4),\displaystyle\left|v^{\prime}(t)\right|_{\infty}\leq\left|v(t)\right|_{\infty}\left(2.1t+\frac{1}{t}+2.1\right)+\max\{\left|\delta_{2}+1\right|,\left|\delta_{3}-1\right|\}\left(4t^{2}+t^{3}+\frac{1}{3}t^{4}\right),

provided |v|1100\left|v\right|_{\infty}\leq\frac{1}{100}, and max{|δ2+1|,|δ31|}106\max\{\left|\delta_{2}+1\right|,\left|\delta_{3}-1\right|\}\leq 10^{-6}. We then find that, for t[0,11]t\in[0,11],

|v(t)|t1065×max{|δ2+1|,|δ31|}105,\displaystyle\frac{\left|v(t)\right|_{\infty}}{t}\leq 10^{65}\times\max\{\left|\delta_{2}+1\right|,\left|\delta_{3}-1\right|\}\leq 10^{-5}, (15)

since max{|δ2+1|,|δ31|}B1\max\{\left|\delta_{2}+1\right|,\left|\delta_{3}-1\right|\}\leq B_{1}. The estimate (15) gives y(t)δ3+12t|v(t)|0y(t)\geq\frac{\delta_{3}+1}{2}t-\left|v(t)\right|\geq 0 on [0,11][0,11], so L2(Tt)L_{2}(T-t) does not change sign for t[0,11]t\in[0,11]. We also find that, at the principal orbit TtT-t,

ξL2+1R2L22δ31=(1t+w(t))y(t)z(t)(2+(δ31)z(t))(δ31)y(t)2,\displaystyle\frac{\xi L_{2}+1-R^{2}-L_{2}^{2}}{\delta_{3}-1}=-\left(\frac{1}{t}+w(t)\right)y(t)-z(t)(2+(\delta_{3}-1)z(t))-(\delta_{3}-1)y(t)^{2},

which is sufficiently close to 2-2 for t[0,11]t\in[0,11] by (15). Therefore, these two sectional curvatures do not change sign on [T11,T)[T-11,T). Since max{|δ2+1|,|δ31|}B1\max\{\left|\delta_{2}+1\right|,\left|\delta_{3}-1\right|\}\leq B_{1}, we can also use these estimates to conclude that ξ(Tt)-\xi(T-t) is close to 1tt\frac{1}{t}-t, so ξ1(10)[T11,T)\xi^{-1}(10)\in[T-11,T), as required.

Proof of Lemma 2.

A key quantity to consider is K(t)2=L2(t)2+(R(t)1)2K(t)^{2}=L_{2}(t)^{2}+(R(t)-1)^{2} because K(t)=ξL22+L2(R1)K(t)K^{\prime}(t)=\frac{-\xi L_{2}^{2}+L_{2}(R-1)}{K(t)}, so that

K(t)(12max{0,ξ})K(t)(12+max{0,ξ})K(t),K(ξ1(10))2δ.\displaystyle\begin{split}K(t)\left(-\frac{1}{2}-\max\{0,\xi\}\right)\leq K^{\prime}(t)\leq\left(\frac{1}{2}+\max\{0,-\xi\}\right)K(t),\\ K(\xi^{-1}(10))\leq\sqrt{2}\delta.\end{split} (16)

Using (16) and the fact that ξ1(0)ξ1(10)10\xi^{-1}(0)-\xi^{-1}(10)\leq 10 (since ξ1\xi^{\prime}\leq-1), we find that

K(t)2e5B2,t(ξ1(10),ξ1(0)].\displaystyle K(t)\leq\sqrt{2}e^{5}B_{2},\ t\in(\xi^{-1}(10),\xi^{-1}(0)]. (17)

Now let t(ξ1(0),TB12]t^{*}\in(\xi^{-1}(0),T-\frac{B_{1}}{2}] be the last time so that

ξ(t)L1(t)+121Tt+12for allt[ξ1(0),t).\displaystyle-\xi(t)\leq-L_{1}(t)+\frac{1}{2}\leq\frac{1}{T-t}+\frac{1}{2}\ \text{for all}\ t\in[\xi^{-1}(0),t^{*}). (18)

Such a tt^{*} must exist because 0L11Tt0\leq-L_{1}\leq\frac{1}{T-t} everywhere by Proposition 1, so the estimate holds at time ξ1(0)\xi^{-1}(0). Then using (16) and (18), we obtain that K(t)(1+1Tt)K(t)K^{\prime}(t)\leq\left(1+\frac{1}{T-t}\right)K(t); integrating from ξ1(0)\xi^{-1}(0) to tt^{*} gives

K(t)8B2eTξ1(0)+5(Tξ1(0))B1,t(ξ1(0),t).\displaystyle K(t)\leq\frac{\sqrt{8}B_{2}e^{T-\xi^{-1}(0)+5}(T-\xi^{-1}(0))}{B_{1}},\ t\in(\xi^{-1}(0),t^{*}). (19)

Now, using the fact that yy2+1y^{\prime}\geq y^{2}+1 and y(ξ1(0))0y(\xi^{-1}(0))\geq 0, where y=min{ξ,L1}y=\min\{-\xi,-L_{1}\}, we find that Tξ1(0)π2T-\xi^{-1}(0)\leq\frac{\pi}{2}. Combining this with (19) and the definition of B1B_{1}, we obtain that K(t)B110K(t)\leq\frac{B_{1}}{10} for all t(ξ1(0),t)t\in(\xi^{-1}(0),t^{*}). This implies that, for t(ξ1(0),t)t\in(\xi^{-1}(0),t^{*}), we have

(ξ+L1)\displaystyle(-\xi+L_{1})^{\prime} =L1(ξ+L1)+2L22\displaystyle=L_{1}(-\xi+L_{1})+2L_{2}^{2}
L1(ξ+L1)+104.\displaystyle\leq L_{1}(-\xi+L_{1})+10^{-4}.

Since ξ+L1-\xi+L_{1} is negative at time ξ1(0)\xi^{-1}(0), and L1L_{1} is negative everywhere, we find that ξ(t)<L1(t)+12-\xi(t)<-L_{1}(t)+\frac{1}{2} for t(ξ1(0),t)t\in(\xi^{-1}(0),t^{*}), so t=TB12t^{*}=T-\frac{B_{1}}{2}.

Combining K(TB12)B110K(T-\frac{B_{1}}{2})\leq\frac{B_{1}}{10} with |R|R2\left|R^{\prime}\right|\leq R^{2} then gives

|R(t)1|B1fort(TB12,T).\displaystyle\left|R(t)-1\right|\leq B_{1}\ \text{for}\ t\in(T-\frac{B_{1}}{2},T). (20)

Equation (20) combined with (19) and (17) then gives

|R(t)1|B1fort(ξ1(10),T).\displaystyle\left|R(t)-1\right|\leq B_{1}\ \text{for}\ t\in(\xi^{-1}(10),T). (21)

In particular, we find that |δ31|B1\left|\delta_{3}-1\right|\leq B_{1}.

To conclude the proof, it therefore suffices to check the smallness of L2L_{2} on [ξ1(10),T)[\xi^{-1}(10),T). The required bound for L2L_{2} on [ξ1(10),TB12][\xi^{-1}(10),T-\frac{B_{1}}{2}] follows from (17), (19) and the fact that t=TB12t^{*}=T-\frac{B_{1}}{2}. The required L2L_{2} bound on [TB12,T)[T-\frac{B_{1}}{2},T) follows from the estimate |L2(t)||R(t)21|B1(R(t)+1)52B1\left|L_{2}(t)\right|^{\prime}\geq-\left|R(t)^{2}-1\right|\geq-B_{1}(R(t)+1)\geq-\frac{5}{2}B_{1} for t(ξ1(0),T)t\in(\xi^{-1}(0),T) coupled with L2(T)=0L_{2}(T)=0. ∎

Proof of Lemma 3.

The estimate |R(t)1|B1\left|R(t)-1\right|\leq B_{1} implies that |L2(t)|B1(R(t)+1)52B1\left|L_{2}(t)\right|^{\prime}\geq-B_{1}(R(t)+1)\geq-\frac{5}{2}B_{1} for t(ξ1(0),T)t\in(\xi^{-1}(0),T). Therefore, |L2(t)|5B1(Tt)2\left|L_{2}(t)\right|\leq\frac{5B_{1}(T-t)}{2} for all t(ξ1(0),T)t\in(\xi^{-1}(0),T), because L2(T)=0L_{2}(T)=0. In fact, we actually obtain

|L2(t)|5B1(Tt)2for allt(ξ1(10),T)\displaystyle\left|L_{2}(t)\right|\leq\frac{5B_{1}(T-t)}{2}\ \text{for all}\ t\in(\xi^{-1}(10),T) (22)

because |L2(t)|B1\left|L_{2}(t)\right|\leq B_{1} on (ξ1(10),ξ1(0))(\xi^{-1}(10),\xi^{-1}(0)) by Lemma 2, and Tξ1(0)>25T-\xi^{-1}(0)>\frac{2}{5}. This estimate on Tξ1(0)T-\xi^{-1}(0) follows from the fact that L22+(R1)22B12L_{2}^{2}+(R-1)^{2}\leq 2B_{1}^{2} is small at time ξ1(0)\xi^{-1}(0), and L1(ξ1(0))[1,0]L_{1}(\xi^{-1}(0))\in[-1,0] (since ξL1+1L120\xi L_{1}+1-L_{1}^{2}\geq 0 everywhere).

Consider the function x(t)=ξ(Tt)+L1(Tt)+tx(t)=-\xi(T-t)+L_{1}(T-t)+t. In light of (22), the function xx satisfies

x(t)25B12t2+xt,x(0)=32(δ2+1),\displaystyle x^{\prime}(t)\geq-25B_{1}^{2}t^{2}+\frac{x}{t},\qquad x^{\prime}(0)=\frac{3}{2}(\delta_{2}+1),

for all tt with x(t)tx(t)\leq t. If δ2+1>B1\delta_{2}+1>B_{1}, we have that x(t)>B1t2x(t)>\frac{B_{1}t}{2} for all t(0,12)t\in(0,12). Since 0L1(Tt)1t0\leq-L_{1}(T-t)\leq\frac{1}{t}, we then find that

ξ(Tt)+L1(Tt)1L1(Tt)x(t)>B1t2.\displaystyle-\xi(T-t)+L_{1}(T-t)-\frac{1}{L_{1}(T-t)}\geq x(t)>\frac{B_{1}t}{2}. (23)

Now, the smallness of (R(ξ1(0))1)2+L2(ξ1(0))2(R(\xi^{-1}(0))-1)^{2}+L_{2}(\xi^{-1}(0))^{2} and the closeness of L1(ξ1(0))L_{1}(\xi^{-1}(0)) to 15+26-\frac{1}{5+\sqrt{26}} implies that Tξ1(10)12T-\xi^{-1}(10)\geq\frac{1}{2}. Therefore, (23) is in contradiction with |L1(Tt)+15+26|B2\left|L_{1}(T-t)+\frac{1}{5+\sqrt{26}}\right|\leq B_{2} at time t=Tξ1(10)t=T-\xi^{-1}(10). ∎

5 The curve C for large values of δ1\delta_{1}

In our quest to prove that normalised SO(2)×SO(3)SO(2)\times SO(3)-invariant solitons on 4\mathbb{R}^{4} 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 δ1,δ2,δ3\delta_{1},\delta_{2},\delta_{3} 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 δ1\delta_{1} (recall that a Ricci soliton on 𝕊4\mathbb{S}^{4} with non-negatve Riemann curvature operator must be round).

Theorem 4.

Suppose we have an SO(2)×SO(3)SO(2)\times SO(3)-invariant gradient shrinking Ricci soliton on 𝕊4\mathbb{S}^{4} of the form (3). Consider the quantities ξ,L1,L2,R\xi,L_{1},L_{2},R associated to the soliton which satisfy (9), and suppose δ(0,1025)\delta\in(0,10^{-25}) and δ1>1δ120\delta_{1}>\frac{1}{\delta^{120}}. Then there is a principal orbit with ξ=10\xi=10, |L1+15+26|δ\left|L_{1}+\frac{1}{5+\sqrt{26}}\right|\leq\delta, |R1|δ\left|R-1\right|\leq\delta, |L2|δ\left|L_{2}\right|\leq\delta, and so that the Riemann curvature is non-negative between this principal orbit and the 𝕊1\mathbb{S}^{1} orbit.

When proving this theorem, it is useful to again change variables with w=L1w=L_{1}, x=L2Rx=\frac{L_{2}}{R}, y=Rξy=\frac{R}{\xi}, z=1Rz=\frac{1}{R} so that

w=wyzyz,x=x+yyz2+x2yyz,y=xy2+y3(w2z2+2x2+z2)yz,z=xyzyz.\displaystyle\begin{split}w^{\prime}&=\frac{-w-yz}{yz},\\ x^{\prime}&=\frac{-x+y-yz^{2}+x^{2}y}{yz},\\ y^{\prime}&=\frac{-xy^{2}+y^{3}(w^{2}z^{2}+2x^{2}+z^{2})}{yz},\\ z^{\prime}&=\frac{xyz}{yz}.\end{split} (24)

Therefore our study essentially reduces to the analysis of the integral curves of the following system of equations:

w=wyz,x=x+yyz2+x2y,y=xy2+y3(w2z2+2x2+z2),z=xyz.\displaystyle\begin{split}w^{\prime}&=-w-yz,\\ x^{\prime}&=-x+y-yz^{2}+x^{2}y,\\ y^{\prime}&=-xy^{2}+y^{3}(w^{2}z^{2}+2x^{2}+z^{2}),\\ z^{\prime}&=xyz.\end{split} (25)

In these co-ordinates, the non-negativity of L1L2-L_{1}L_{2} and ξL2+1R2L22\xi L_{2}+1-R^{2}-L_{2}^{2} is implied by the non-negativity of xx and xy+z21x2\frac{x}{y}+z^{2}-1-x^{2} (since we already know that L10L_{1}\leq 0). The solutions we are interested in are part of the two-dimensional unstable manifold of the critical point of (25) at (0,1,12,0)(0,1,\frac{1}{2},0); the δ1\delta_{1} parameter now describes the initial direction of travel through this two-dimensional unstable manifold, and is given precisely by δ1=limt01z(t)1tt\delta_{1}=\lim_{t\to 0}\frac{\frac{1}{z(t)}-\frac{1}{t}}{t} for the solutions of (24).

In this section, we are only interested in evolving the system (24) up until the time ξ1(10)\xi^{-1}(10), so we can assume that y(0,)y\in(0,\infty). It is handy to keep track of the evolution of the quantities C=xy(1z)C=\frac{x}{y(1-z)}, D=wyw2D=\frac{w}{y}-w^{2} and E=xy+z21x2E=\frac{x}{y}+z^{2}-1-x^{2}. Indeed, the closeness of DD to 1-1 measures the Gaussian structure, positivity of EE ensures positivity of the last sectional curvature, and CC is an analogue of a quantity that arises in the construction of the Bryant soliton. We find that

C=C+(1+z)+Cxy+C(xyy2(w2z2+2x2+z2))+C2y2zD=(D1)(1yw)+D~E=xy(1w2z2x2)+E~,\displaystyle\begin{split}C^{\prime}&=-C+(1+z)+Cxy+C(xy-y^{2}(w^{2}z^{2}+2x^{2}+z^{2}))+C^{2}y^{2}z\\ D^{\prime}&=(-D-1)(1-yw)+\tilde{D}\\ E^{\prime}&=xy\left(1-w^{2}z^{2}-x^{2}\right)+\tilde{E},\end{split} (26)

where |D~|3|z1|+3|x|\left|\tilde{D}\right|\leq 3\left|z-1\right|+3\left|x\right| whenever 0x,y,w,z10\leq x,y,-w,z\leq 1, and E~=0\tilde{E}=0 whenever E=0E=0 and y(0,)y\in(0,\infty). Therefore, to prove Theorem 4, it suffices to prove the following:

Theorem 5.

Choose δ(0,1025)\delta\in(0,10^{-25}). If we have a Ricci soliton with δ11δ120\delta_{1}\geq\frac{1}{\delta^{120}}, then the corresponding trajectory of (25) which α\alpha-limits to (0,1,12,0)(0,1,\frac{1}{2},0) includes a point at which yz=110yz=\frac{1}{10}, 1δ3z11-\frac{\delta}{3}\leq z\leq 1, 0w10\leq-w\leq 1, |D+1|δ3\left|D+1\right|\leq\frac{\delta}{3} and xδ3x\leq\frac{\delta}{3}, and so that min{x,E}0\min\{x,E\}\geq 0 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 (0,1,12,0)(0,1,\frac{1}{2},0) as δ1\delta_{1} becomes large:

  • the trajectory travels from (0,1,12,0)(0,1,\frac{1}{2},0), and gets close to the critical point (0,0,0,0)(0,0,0,0);

  • the trajectory then travels close to the line of critical points (0,0,0,z)(0,0,0,z) for z[0,1]z\in[0,1];

  • the trajectory then breaks away from this line of critical points as y>0y>0 begins to grow while zz stays close to 11, 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 (w,x,y,z)=(0,0,0,0)(w,x,y,z)=(0,0,0,0) in order to get the desired conclusion.

Lemma 4.

Choose δ(0,1025)\delta\in(0,10^{-25}) and suppose our trajectory of (25) includes a point with 0w,x,y,zδ60\leq-w,x,y,z\leq\delta^{6}, |yx1|δ6\left|\frac{y}{x}-1\right|\leq\delta^{6}, |D|1\left|D\right|\leq 1 and E0E\geq 0. Then the trajectory includes a later point at which yz=110yz=\frac{1}{10}, 1δ3z11-\frac{\delta}{3}\leq z\leq 1, 0w10\leq-w\leq 1, |D+1|δ3\left|D+1\right|\leq\frac{\delta}{3} and xδ3x\leq\frac{\delta}{3}. Furthermore, min{E,x}0\min\{E,x\}\geq 0 on the trajectory between these two points.

Proof.

We can assume that we have a solution of (25) so that 0 is the time of the first point.

Step One: construct an interval (0,t)(0,t^{*}) on which we expect to find the required terminal point; find some basic estimates. Define t>0t^{*}>0 so that

(0,t)is the maximal time interval on which|x|110,|y|19,|w|12andz(0,1).\displaystyle(0,t^{*})\ \textit{is the maximal time interval on which}\ \left|x\right|\leq\frac{1}{10},\left|y\right|\leq\frac{1}{9},\left|w\right|\leq\frac{1}{2}\ \textit{and}\ z\in(0,1). (27)

On such an interval, (26) implies that

C+1C10CC+2+C250+C10,\displaystyle-C+1-\frac{C}{10}\leq C^{\prime}\leq-C+2+\frac{C^{2}}{50}+\frac{C}{10},

so since |C(0)1|<103\left|C(0)-1\right|<10^{-3}, we find that C(t)[0.9,2.4]C(t)\in[0.9,2.4] for all t(0,t)t\in(0,t^{*}). The equations for yy and zz in (25) can then be written

yy2=y(C(1z)+w2z2+2C2y2(1z)2+z2),zy2=Cz(1z).\displaystyle\begin{split}\frac{y^{\prime}}{y^{2}}&=y\left(-C(1-z)+w^{2}z^{2}+2C^{2}y^{2}(1-z)^{2}+z^{2}\right),\\ \frac{z^{\prime}}{y^{2}}&=Cz(1-z).\end{split} (28)

Now we claim that t>z1(12)t^{*}>z^{-1}(\frac{1}{2}). Indeed, (28) and the fact that y(0)δ6y(0)\leq\delta^{6} with C[0.9,2.4]C\in[0.9,2.4] implies that

y(t)δ6fort(0,min{t,z1(12)}).\displaystyle y(t)\leq\delta^{6}\ \text{for}\ t\in(0,\min\{t^{*},z^{-1}(\frac{1}{2})\}). (29)

Estimate (29), coupled with the equation for ww^{\prime} in (25) and the definition of CC implies that

|x|=Cy|1z|5δ6and|w|<14fort(0,min{t,z1(12)}).\displaystyle\left|x\right|=\frac{Cy}{\left|1-z\right|}\leq 5\delta^{6}\ \text{and}\ \left|w\right|<\frac{1}{4}\ \text{for}\ t\in(0,\min\{t^{*},z^{-1}(\frac{1}{2})\}). (30)

The estimates (29) and (30) combine with the definition of tt^{*} in (27) to show that t>z1(12)t^{*}>z^{-1}(\frac{1}{2}).

Now, on [z1(12),t][z^{-1}(\frac{1}{2}),t^{*}], (28) gives us yy21.5y\frac{y^{\prime}}{y^{2}}\leq 1.5y and (1z)y20.4(1z)\frac{(1-z)^{\prime}}{y^{2}}\leq-0.4(1-z). Therefore, y(t)δ6e1.5(sz1(12))y(t)\leq\delta^{6}e^{1.5(s-z^{-1}(\frac{1}{2}))} and (1z(t))12e0.4(sz1(12))(1-z(t))\leq\frac{1}{2}e^{-0.4(s-z^{-1}(\frac{1}{2}))}, where ss is some function of tt. Combining these estimates gives

y(t)(1z(t))3.75δ610for allt[z1(12),t].\displaystyle y(t)(1-z(t))^{3.75}\leq\frac{\delta^{6}}{10}\ \text{for all}\ t\in[z^{-1}(\frac{1}{2}),t^{*}]. (31)

We conclude this step with the assertion that y(t)=19y(t^{*})=\frac{1}{9}. We show this by demonstrating that all of the other inequalities defining tt^{*} in (27) are strict at time tt^{*} itself. On the interval (0,t)(0,t^{*}), yz<14yz<\frac{1}{4}, so w<14-w<\frac{1}{4} on (0,t)(0,t^{*}) since w(0)<14-w(0)<\frac{1}{4}. For any t[0,t]t\in[0,t^{*}] with (1z)<92.4δ(1-z)<\frac{9}{2.4}\delta, it is clear that xCy(1z)<δx\leq Cy(1-z)<\delta. On the other hand, if (1z)92.4δ(1-z)\geq\frac{9}{2.4}\delta, then x2.4y(1z)3.75(1z)2.75<δx\leq 2.4\frac{y(1-z)^{3.75}}{(1-z)^{2.75}}<\delta by (31) as well. It is clear from (28) that z(t)<1z(t^{*})<1, so it must be the case that y(t)=19y(t^{*})=\frac{1}{9}.

Step Two: show that xx and EE are non-negative on (0,t)(0,t^{*}). We have x0x\geq 0 on (0,t)(0,t^{*}) by (25), since z(0,1)z\in(0,1) on (0,t)(0,t^{*}) and x(0)0x(0)\geq 0. The estimates on x,z,wx,z,w provided by the definition of tt^{*} in (27) then imply that 1w2z2x201-w^{2}z^{2}-x^{2}\geq 0 on (0,t)(0,t^{*}); this observation coupled with (26) and E(0)0E(0)\geq 0 implies that E0E\geq 0 here as well.

Step Three: find the required point in (0,t)(0,t^{*}). Let II be the connected interval of times ending at tt^{*} so that y(t)[110,19]y(t)\in[\frac{1}{10},\frac{1}{9}]. We find from (31) that (1z)δ63.75<δ200(1-z)\leq\delta^{\frac{6}{3.75}}<\frac{\delta}{200} in II. The definition of CC then implies that 0xδ30\leq x\leq\frac{\delta}{3} in II as well. The intermediate value theorem also gives us that there is some t1It_{1}\in I at which yz=110yz=\frac{1}{10}. Finally, we need to demonstrate that |D+1|δ3\left|D+1\right|\leq\frac{\delta}{3} at t1t_{1}. At time (1z)1(δ100)(1-z)^{-1}(\frac{\delta}{100}) (which is less than t1t_{1} by the above estimate for (1z)(1-z) on II), (31) tells us that yδ61003.7510δ3.75δ1.25y\leq\frac{\delta^{6}100^{3.75}}{10\delta^{3.75}}\leq\delta^{1.25}. The fact that y1.5y3y^{\prime}\leq 1.5y^{3} on [z1(12),t][z^{-1}(\frac{1}{2}),t^{*}] implies that the distance between II and (1z)1(δ100)(1-z)^{-1}(\frac{\delta}{100}) is greater than 1100δ2\frac{1}{100\delta^{2}}. This large amount of time tells us that DD must get quite close to 1-1 by the time we land in II. Indeed, from (26) and (31) we find that

|D+1|34|D1|+δ10\displaystyle\left|D+1\right|^{\prime}\leq-\frac{3}{4}\left|D-1\right|+\frac{\delta}{10} (32)

on ((1z)1(δ100),t)((1-z)^{-1}(\frac{\delta}{100}),t^{*}), while

|D+1|34|D1|+10\displaystyle\left|D+1\right|^{\prime}\leq-\frac{3}{4}\left|D-1\right|+10 (33)

holds on (0,t)(0,t^{*}). Estimate (32) tells us that whenever |D+1|δ3\left|D+1\right|\geq\frac{\delta}{3} and t((1z)1(δ100),t)t\in((1-z)^{-1}(\frac{\delta}{100}),t^{*}) , we have |D+1|δ10\left|D+1\right|^{\prime}\leq-\frac{\delta}{10}. The estimate (33) coupled with |D(0)|1\left|D(0)\right|\leq 1 implies that |D|15\left|D\right|\leq 15 at time (1z)1(δ100)(1-z)^{-1}(\frac{\delta}{100}), so it takes no more than time 105δ>1100δ2\frac{10^{5}}{\delta}>\frac{1}{100\delta^{2}} to get |D+1|δ3\left|D+1\right|\leq\frac{\delta}{3}. Therefore, |D+1|δ3\left|D+1\right|\leq\frac{\delta}{3} is achieved for all times in II (including t1t_{1}). ∎

Now we discuss how making δ1\delta_{1} large can force our trajectory to be close to the (0,0,0,0)(0,0,0,0) point, in the sense of Lemma 4. This is achieved with the two lemmas below.

Lemma 5.

Let B3(0,10100)B_{3}\in(0,10^{-100}). Suppose our trajectory solution of (25) has a point so that 0zB3120\leq z\leq B_{3}^{12}, 0wB30\leq-w\leq B_{3}, E0E\geq 0 and

{|xx()(19)|,|yy()(19)|}B36,\{\left|x-x^{(\infty)}(\frac{1}{9})\right|,\left|y-y^{(\infty)}(\frac{1}{9})\right|\}\leq B_{3}^{6},

where (x(),y())(x^{(\infty)},y^{(\infty)}) is the special Bryant soliton solution discussed in Theorem 10 of the Appendix. Then there is a later point on the trajectory at which 0w,x,y,zB30\leq-w,x,y,z\leq B_{3}, |yx1|B3\left|\frac{y}{x}-1\right|\leq B_{3}, |D|1\left|D\right|\leq 1, and so that min{E,x}0\min\{E,x\}\geq 0 on the trajectory between these two points.

Proof.

This time, we find it convenient to have our (w,x,y,z)(w,x,y,z) solution of (25) so that the initial point described in the hypothesis of the lemma corresponds to time 19\frac{1}{9}. This is so we can easily compare our solution to the special x(),y()x^{(\infty)},y^{(\infty)} solution. Strictly speaking, these functions x,yx^{\infty},y^{\infty} solve (53), but we now reparametrise them so that they solve (54) instead, with a parametrisation that preserves 19\frac{1}{9}.

Step One: estimates on the limiting solution. As discussed in Appendix A, we have

(x())(t)0,(y())(t)0,y()(t)x()(t)[12,1]for allt[19,).\displaystyle(x^{(\infty)})^{\prime}(t)\leq 0,\qquad(y^{(\infty)})^{\prime}(t)\leq 0,\qquad\frac{y^{(\infty)}(t)}{x^{(\infty)}(t)}\in[\frac{1}{2},1]\ \text{for all}\ t\in[\frac{1}{9},\infty). (34)

Combining (34) with (54) gives

(y())(y())3(14(y())2).\displaystyle(y^{(\infty)})^{\prime}\leq-(y^{(\infty)})^{3}(1-4(y^{(\infty)})^{2}). (35)

Also note that (34), combined with Theorem 10 and Proposition 7 in Appendix A implies that

10.0375x()(19)10.0331120.01875y()(19)120.01545.\displaystyle\begin{split}1-0.0375\leq x^{(\infty)}(\frac{1}{9})\leq 1-0.0331\\ \frac{1}{2}-0.01875\leq y^{(\infty)}(\frac{1}{9})\leq\frac{1}{2}-0.01545.\end{split} (36)

It is handy to note that

[19,)ABCD¯\displaystyle[\frac{1}{9},\infty)\subseteq\overline{A^{*}\cup B^{*}\cup C^{*}\cup D^{*}}

with

A=(y())1(13,120.01545),B=(y())1(18,13),C=(y())1(B32,18),D=(y())1(0,B32).\displaystyle A^{*}=(y^{(\infty)})^{-1}(\frac{1}{3},\frac{1}{2}-0.01545),B^{*}=(y^{(\infty)})^{-1}(\frac{1}{8},\frac{1}{3}),\ C^{*}=(y^{(\infty)})^{-1}(\frac{B_{3}}{2},\frac{1}{8}),\ D^{*}=(y^{(\infty)})^{-1}(0,\frac{B_{3}}{2}).

By (35), we find that (y())(t)127(14(y()(t))2)(y^{(\infty)})^{\prime}(t)\leq-\frac{1}{27}(1-4(y^{(\infty)}(t))^{2}) on AA^{*} so that |A|18\left|A^{*}\right|\leq 18. Also, (y())(t)1283(y^{(\infty)})^{\prime}(t)\leq-\frac{1}{2\cdot 8^{3}} on BB^{*} so that |B|250\left|B^{*}\right|\leq 250. Finally, |C|3B32\left|C^{*}\right|\leq\frac{3}{B_{3}^{2}}, because (y())(t)15(y()(t))316(y^{(\infty)})^{\prime}(t)\leq-\frac{15(y^{(\infty)}(t))^{3}}{16} for tCt\in C^{*}.

Step Two: closeness to the limiting solution. Define the positive number t0supC268+3B32t_{0}\leq\sup C\leq 268+\frac{3}{B_{3}^{2}} so that

[19,t0]is the maximal time interval on whichz(t)B36.\displaystyle[\frac{1}{9},t_{0}]\ \text{is the maximal time interval on which}\ z(t)\leq B_{3}^{6}. (37)

We will now obtain the required estimates by showing that the quantities X=xx()X=x-x^{(\infty)} and Y=yy()Y=y-y^{(\infty)} are small on the interval [19,t0][\frac{1}{9},t_{0}]. By examining (25) and (54), we compute

(XY)\displaystyle\begin{pmatrix}X\\ Y\end{pmatrix}^{\prime} =(1+2x()y()1+(x())2(y())2(4x()y()1)x()y()(6x()y()2))(XY)+O(X2+Y2)+Z\displaystyle=\begin{pmatrix}-1+2x^{(\infty)}y^{(\infty)}&1+(x^{(\infty)})^{2}\\ (y^{(\infty)})^{2}(4x^{(\infty)}y^{(\infty)}-1)&x^{(\infty)}y^{(\infty)}(6x^{(\infty)}y^{(\infty)}-2)\end{pmatrix}\begin{pmatrix}X\\ Y\end{pmatrix}+O(X^{2}+Y^{2})+Z (38)

where |O(X2+Y2)|100(|X|2+|Y|2)\left|O(X^{2}+Y^{2})\right|_{\infty}\leq 100(\left|X\right|^{2}+\left|Y\right|^{2}) provided |X|2+|Y|21\left|X\right|^{2}+\left|Y\right|^{2}\leq 1, and |Z|4|z|2\left|Z\right|_{\infty}\leq 4\left|z\right|^{2} provided x,y,w,z[0,1.1]x,y,-w,z\in[0,1.1]. We now use (38) to obtain smallness of (X,Y)(X,Y) on (19,t0)(\frac{1}{9},t_{0}). Let V=max{|Y|,2|XY|}V=\max\{\left|Y\right|,2\left|X-Y\right|\}; (38) implies that

V\displaystyle V^{\prime} B328V+16B36\displaystyle\leq-\frac{B_{3}^{2}}{8}V+16B_{3}^{6}

on CC^{*}, as long as |V|B33\left|V\right|\leq B_{3}^{3}, and

V\displaystyle V^{\prime} 2V+16B36\displaystyle\leq 2V+16B_{3}^{6}

on ABA^{*}\cup B^{*}, provided |V|B33\left|V\right|\leq B_{3}^{3}. Therefore, since V(19)4B36V(\frac{1}{9})\leq 4B_{3}^{6}, we find that

V(t)B33for alltABC¯[19,t0].\displaystyle V(t)\leq B_{3}^{3}\ \text{for all}\ t\in\overline{A^{*}\cup B^{*}\cup C^{*}}\cap[\frac{1}{9},t_{0}]. (39)

Step Three: concluding estimates. We need to check that x,y,z,yx1,wx,y,z,\frac{y}{x}-1,w are all smaller than B3B_{3} at time t0t_{0}, that |D|1\left|D\right|\leq 1, and that E,x0E,x\geq 0 on [19,t0][\frac{1}{9},t_{0}]. We claim that y()(t0)=B32y^{(\infty)}(t_{0})=\frac{B_{3}}{2}. To see this, we use (25) to estimate

zz\displaystyle\frac{z^{\prime}}{z} =xy\displaystyle=xy
(x()+3B33)(y()+B33)\displaystyle\leq(x^{(\infty)}+3B_{3}^{3})(y^{(\infty)}+B_{3}^{3})
2(y()+3B332)22.5(y())2\displaystyle\leq 2(y^{(\infty)}+\frac{3B_{3}^{3}}{2})^{2}\leq 2.5(y^{(\infty)})^{2}

on [19,t0][\frac{1}{9},t_{0}]. Therefore, z(t)B312e28z(t)\leq B_{3}^{12}e^{28} for all tAt\in A^{*}, so t0A¯t_{0}\notin\overline{A^{*}}. On BB^{*}, we estimate that zz2z^{\prime}\leq\frac{z}{2}, so z(t)B312e153B311z(t)\leq B_{3}^{12}e^{153}\leq B_{3}^{11} for all tBt\in B^{*} and t0B¯t_{0}\notin\overline{B^{*}}. On the other hand, (35) gives (y)15(y())316(y^{\infty})^{\prime}\leq-\frac{15(y^{(\infty)})^{3}}{16} on CC^{*}, so that

(y()(t))2815(t(y())1(18))+512\displaystyle(y^{(\infty)}(t))^{2}\leq\frac{8}{15(t-(y^{(\infty)})^{-1}(\frac{1}{8}))+512}

for tC¯t\in\overline{C^{*}}. We consequently find

z(t)z(t)2015(t(y())1(18))+512,\displaystyle\frac{z^{\prime}(t)}{z(t)}\leq\frac{20}{15(t-(y^{(\infty)})^{-1}(\frac{1}{8}))+512},

so that z(t)<B36z(t)<B_{3}^{6} for tCt\in C^{*}. Therefore, y()(t0)=B32y^{(\infty)}(t_{0})=\frac{B_{3}}{2} as required.

The smallness of y()(t0)y^{(\infty)}(t_{0}) and (39) imply the required estimates for x(t0),y(t0)x(t_{0}),y(t_{0}). The estimate for z(t0)z(t_{0}) follows from the definition of t0t_{0}. The estimates for ww and DD follow immediately from (39) and the equation for ww^{\prime} in (25). Using Proposition 8, we also have

|yx1|\displaystyle\left|\frac{y}{x}-1\right| |yxy()x()|+|y()x()1|\displaystyle\leq\left|\frac{y}{x}-\frac{y^{(\infty)}}{x^{(\infty)}}\right|+\left|\frac{y^{(\infty)}}{x^{(\infty)}}-1\right|
|yx()xy()|x()(x()2V)+3B34\displaystyle\leq\frac{\left|yx^{(\infty)}-xy^{(\infty)}\right|}{x^{(\infty)}(x^{(\infty)}-2V)}+\frac{3B_{3}}{4}
B3,\displaystyle\leq B_{3},

since VB33V\leq B_{3}^{3}, while x()y()=B22x^{(\infty)}\geq y^{(\infty)}=\frac{B_{2}}{2}. It is clear that xx is non-negative on [19,t0][\frac{1}{9},t_{0}], since x()B32x^{(\infty)}\geq\frac{B_{3}}{2} and VB33V\leq B_{3}^{3}. To show that EE is non-negative on [19,t0][\frac{1}{9},t_{0}], note that 1w2z2x21-w^{2}z^{2}-x^{2} is non-negative on [19,t0][\frac{1}{9},t_{0}] since x()[B32,10.331]x^{(\infty)}\in[\frac{B_{3}}{2},1-0.331], VB33V\leq B_{3}^{3}, zB3z\leq B_{3} and wB3w\leq B_{3} (by (25) and the estimates on yy and ww). The non-negativity of 1w2z2x21-w^{2}z^{2}-x^{2} and the fact that E(19)0E(\frac{1}{9})\geq 0 implies that E0E\geq 0 on [19,t0][\frac{1}{9},t_{0}] by (26).

Lemma 6.

For each B3(0,10100)B_{3}\in(0,10^{-100}), choose an aribtrary δ1B320\delta_{1}\geq B_{3}^{-20}. Then the solution of (24) satisfies 0z(19δ1)B3120\leq z(\frac{1}{9\sqrt{\delta_{1}}})\leq B_{3}^{12}, 0w(19δ1)B30\leq-w(\frac{1}{9\sqrt{\delta_{1}}})\leq B_{3}, and {|x(19δ1)x()(19)|,|y(19δ1)y()(19)|}B64\{\left|x(\frac{1}{9\sqrt{\delta_{1}}})-x^{(\infty)}(\frac{1}{9})\right|,\left|y(\frac{1}{9\sqrt{\delta_{1}}})-y^{(\infty)}(\frac{1}{9})\right|\}\leq B_{6}^{4}, where (x(),y())(x^{(\infty)},y^{(\infty)}) is the Bryant soliton solution discussed in the Appendix. Furthermore, min{E,x}0\min\{E,x\}\geq 0 on (0,19δ1)(0,\frac{1}{9\sqrt{\delta_{1}}}).

Proof.

Let p=1δ1p=\frac{1}{\sqrt{\delta_{1}}}, t=19t^{*}=\frac{1}{9}, and consider the rescaled functions f~(t)=pf(pt)\tilde{f}(t)=pf(pt) for f=R,L1,L2,ξf=R,L_{1},L_{2},\xi. Then these new functions satisfy

ξ~\displaystyle\tilde{\xi}^{\prime} =L1~22L2~2p2,\displaystyle=-\tilde{L_{1}}^{2}-2\tilde{L_{2}}^{2}-p^{2},
L1~\displaystyle\tilde{L_{1}}^{\prime} =ξ~L1~p2,\displaystyle=-\tilde{\xi}\tilde{L_{1}}-p^{2},
L2~\displaystyle\tilde{L_{2}} =ξ~L2~+R~2p2,\displaystyle=-\tilde{\xi}\tilde{L_{2}}+\tilde{R}^{2}-p^{2},
R~\displaystyle\tilde{R}^{\prime} =L2~R~,\displaystyle=-\tilde{L_{2}}\tilde{R},

with the ‘new’ δ1\delta_{1} equal to 11, i.e., limt0R~(t)1tt=1\lim_{t\to 0}\frac{\tilde{R}(t)-\frac{1}{t}}{t}=1. The discussion of the Bryant soliton on 3\mathbb{R}^{3} discussed in the Appendix implies the existence of smooth functions ξ(),L1(),L2(),R():(0,)\xi^{(\infty)},L_{1}^{(\infty)},L_{2}^{(\infty)},R^{(\infty)}:(0,\infty)\to\mathbb{R} of the form (10) so that

(ξ())\displaystyle(\xi^{(\infty)})^{\prime} =(L1())22(L2())2,\displaystyle=-(L_{1}^{(\infty)})^{2}-2(L_{2}^{(\infty)})^{2},
(L1())\displaystyle(L_{1}^{(\infty)})^{\prime} =ξ()L1(),\displaystyle=-\xi^{(\infty)}L_{1}^{(\infty)},
(L2())\displaystyle(L_{2}^{(\infty)})^{\prime} =ξ()L2()+(R())2,\displaystyle=-\xi^{(\infty)}L_{2}^{(\infty)}+(R^{(\infty)})^{2},
(R())\displaystyle(R^{(\infty)})^{\prime} =L2()R(),\displaystyle=-L_{2}^{(\infty)}R^{(\infty)},
1\displaystyle 1 =limt0R()(t)1tt.\displaystyle=\lim_{t\to 0}\frac{R^{(\infty)}(t)-\frac{1}{t}}{t}.

Of course, L1()=0L_{1}^{(\infty)}=0 uniformly here. Recall that by using Theorem 10 and Propositon 7, we find that the corresponding x()x^{(\infty)} and y()y^{(\infty)} functions satisfy (36). It is also well-known that x()x^{(\infty)} and y()y^{(\infty)} are monotonically decreasing functions.

We prove this lemma by comparing (ξ~,L1~,L2~,R~)(\tilde{\xi},\tilde{L_{1}},\tilde{L_{2}},\tilde{R}) to (ξ(),L1(),L2(),R())(\xi^{(\infty)},L_{1}^{(\infty)},L_{2}^{(\infty)},R^{(\infty)}) on the interval [0,19][0,\frac{1}{9}]. We use the notation η~\tilde{\eta} and η()\eta^{(\infty)} to mean the functions from [0,)[0,\infty) to 4\mathbb{R}^{4} consisting of components formed by breaking the corresponding sets of functions according to (10). Letting v=η~η()\textbf{v}=\tilde{\eta}-\eta^{(\infty)}, we find

v(t)=Avt+B(v)P,v(0)=0,v(0)=(p2,p23,0,0),\displaystyle\textbf{v}^{\prime}(t)=\frac{A\textbf{v}}{t}+B(\textbf{v})-P,\qquad\textbf{v}(0)=0,\qquad\textbf{v}^{\prime}(0)=(-p^{2},-\frac{p^{2}}{3},0,0), (40)

where

A=(0040020010220011)P=(p2p2p20)\displaystyle A=\begin{pmatrix}0&0&-4&0\\ 0&-2&0&0\\ -1&0&-2&2\\ 0&0&-1&-1\end{pmatrix}\qquad P=\begin{pmatrix}p^{2}\\ p^{2}\\ p^{2}\\ 0\end{pmatrix}

and

|B(v)|5|v|,\displaystyle\left|B(\textbf{v})\right|_{\infty}\leq 5\left|\textbf{v}\right|_{\infty},

provided |η~|1\left|\tilde{\eta}\right|_{\infty}\leq 1 and |v|1100\left|\textbf{v}\right|_{\infty}\leq\frac{1}{100}. One can easily check using the results in the Appendix that |η~|1\left|\tilde{\eta}\right|_{\infty}\leq 1 on [0,t][0,t^{*}].

Now we let s=S1v\textbf{s}=S^{-1}\textbf{v} so that

s=S1ASst+S1B(Ss)S1P,s(0)=0,s(0)=S1v(0),\displaystyle\textbf{s}^{\prime}=\frac{S^{-1}AS\textbf{s}}{t}+S^{-1}B(S\textbf{s})-S^{-1}P,\qquad\textbf{s}(0)=0,\qquad\textbf{s}^{\prime}(0)=S^{-1}\textbf{v}^{\prime}(0),

where

S=(2108001011021001),S1=(19019109130234301001901919),S1AS=(2100020000200001).\displaystyle S=\begin{pmatrix}2&-1&0&8\\ 0&0&1&0\\ 1&-1&0&-2\\ 1&0&0&1\end{pmatrix},\qquad S^{-1}=\begin{pmatrix}-\frac{1}{9}&0&\frac{1}{9}&\frac{10}{9}\\ -\frac{1}{3}&0&-\frac{2}{3}&\frac{4}{3}\\ 0&1&0&0\\ \frac{1}{9}&0&-\frac{1}{9}&-\frac{1}{9}\end{pmatrix},\qquad S^{-1}AS=\begin{pmatrix}-2&1&0&0\\ 0&-2&0&0\\ 0&0&-2&0\\ 0&0&0&1\end{pmatrix}.

This almost-diagonal form of the equations makes it clear that

|v(t)|11|s(t)|106p2t1100\displaystyle\left|\textbf{v}(t)\right|_{\infty}\leq 11\left|\textbf{s}(t)\right|_{\infty}\leq 10^{6}p^{2}t\leq\frac{1}{100}

for all t[0,t]t\in[0,t^{*}], since pp 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

|x(pt)x(t)|\displaystyle\left|x(pt)-x^{\infty}(t)\right| |L2~(t)R~(t)L2()(t)R()(t)|\displaystyle\leq\left|\frac{\tilde{L_{2}}(t)}{\tilde{R}(t)}-\frac{L_{2}^{(\infty)}(t)}{R^{(\infty)}(t)}\right|
t2|(R()(t)L2~(t)L2()(t)R~(t))|\displaystyle\leq t^{2}\left|\left(R^{(\infty)}(t)\tilde{L_{2}}(t)-L_{2}^{(\infty)}(t)\tilde{R}(t)\right)\right|
t2(R()(t)|L2~(t)L2()(t)|+|L2()(t)||R()(t)R~(t)|)\displaystyle\leq t^{2}\left(R^{(\infty)}(t)\left|\tilde{L_{2}}(t)-L_{2}^{(\infty)}(t)\right|+\left|L_{2}^{(\infty)}(t)\right|\left|R^{(\infty)}(t)-\tilde{R}(t)\right|\right)
2106p2t2,t[0,19].\displaystyle\leq 2\cdot 10^{6}p^{2}t^{2},\qquad t\in[0,\frac{1}{9}].

In this previous computation, we used the estimates on R()R^{(\infty)} and L2()L_{2}^{(\infty)} from Theorem 10 in the Appendix. Similarly,

|y(pt)y(t)|\displaystyle\left|y(pt)-y^{\infty}(t)\right| |R~(t)ξ~(t)R()(t)ξ()(t)|\displaystyle\leq\left|\frac{\tilde{R}(t)}{\tilde{\xi}(t)}-\frac{R^{(\infty)}(t)}{\xi^{(\infty)}(t)}\right|
1ξ~(t)ξ()(t)|(R()(t)ξ~(t)ξ()(t)R~(t))|\displaystyle\leq\frac{1}{\tilde{\xi}(t)\xi^{(\infty)}(t)}\left|\left(R^{(\infty)}(t)\tilde{\xi}(t)-\xi^{(\infty)}(t)\tilde{R}(t)\right)\right|
t2(R()(t)|ξ~(t)ξ()(t)|+ξ()(t)|R()(t)R~(t)|)\displaystyle\leq t^{2}\left(R^{(\infty)}(t)\left|\tilde{\xi}(t)-\xi^{(\infty)}(t)\right|+\xi^{(\infty)}(t)\left|R^{(\infty)}(t)-\tilde{R}(t)\right|\right)
4106p2t2,t[0,19].\displaystyle\leq 4\cdot 10^{6}p^{2}t^{2},\qquad t\in[0,\frac{1}{9}].

To check the closeness of ww and zz to 0, first note that

0z(pt)pt\displaystyle 0\leq z(pt)\leq pt (41)

since |R|R2\left|R^{\prime}\right|\leq R^{2} by Proposition 2. Also, the above estimates for v and η()\eta^{(\infty)} imply that ξ~(19)0\tilde{\xi}(\frac{1}{9})\geq 0, so the equation for L~1\tilde{L}_{1} implies immediately that 0L~1(t)p20\leq-\tilde{L}_{1}^{\prime}(t)\leq p^{2} for each t[0,t]t\in[0,t^{*}], so we find

0w(pt)=L1(pt)=L~1(t)ppt,t[0,19].\displaystyle 0\leq-w(pt)=-L_{1}(pt)=\frac{-\tilde{L}_{1}(t)}{p}\leq pt,\qquad t\in[0,\frac{1}{9}].

To conclude the proof, we need show that min{x,E}0\min\{x,E\}\geq 0 on this interval. It is clear that x0x\geq 0. To show that E0E\geq 0, it suffices to show that |z(pt)|2|1x(pt)|10\left|z(pt)\right|^{2}\leq\frac{\left|1-x(pt)\right|}{10} for all t[0,19]t\in[0,\frac{1}{9}] because of the inequality 0w10\leq-w\leq 1, the equation for EE^{\prime} in (26) and the fact that limt0ξL2+1R2L22=η0(0)+12η3(0)=3η2(0)=6δ1\lim_{t\to 0}\xi L_{2}+1-R^{2}-L_{2}^{2}=\eta_{0}^{\prime}(0)+1-2\eta_{3}^{\prime}(0)=-3\eta_{2}^{\prime}(0)=6\delta_{1}, so that ξL2+1R2L22\xi L_{2}+1-R^{2}-L_{2}^{2} is initially positive. From (41), we have

0z(pt)pt1,\displaystyle 0\leq\frac{z(pt)}{pt}\leq 1,

but we also have

1x(pt)p2t2\displaystyle\frac{1-x(pt)}{p^{2}t^{2}} =1x~(t)p2t2\displaystyle=\frac{1-\tilde{x}(t)}{p^{2}t^{2}}
=1x()(t)p2t2+x()(t)x~(pt)p2t2\displaystyle=\frac{1-x^{(\infty)}(t)}{p^{2}t^{2}}+\frac{x^{(\infty)}(t)-\tilde{x}(pt)}{p^{2}t^{2}}
2tan2(32t)t2p24106,\displaystyle\geq\frac{2\tan^{2}\left(\sqrt{\frac{3}{2}}t\right)}{t^{2}p^{2}}-4\cdot 10^{6},

by Theorem 10, so we obtain the required estimates. ∎

6 Bounds on curvature at the 𝕊2\mathbb{S}^{2} singular orbit

By Theorems 3 and 4, we find that any Ricci soliton must satisfy δ1[0,1020,000]\delta_{1}\in[0,10^{20,000}]. We now construct bounds on δ2\delta_{2} and δ3\delta_{3}. Fortunately, these bounds are simpler to construct, and can be found without using the (already large) bound on δ1\delta_{1}. The bound on δ2\delta_{2} is easier to construct once we have a bound on δ3\delta_{3}, so we treat δ3\delta_{3} first.

Theorem 6.

Suppose the metric of the form (3) is a gradient shrinking Ricci soliton with Einstein constant λ=1\lambda=1. Then δ3[0,40]\delta_{3}\in[0,40].

Proof.

We already know from Proposition 5 that δ30\delta_{3}\geq 0, so suppose for the sake of contradiction that δ3>40\delta_{3}>40. In this case, we claim that

T>12andL2(t)<0for allt(T12,T).\displaystyle T>\frac{1}{2}\ \text{and}\ L_{2}(t)<0\ \text{for all}\ t\in(T-\frac{1}{2},T). (42)

To verify (42), note that if T12T\leq\frac{1}{2}, then Proposition 2 implies that |R|R2\left|R^{\prime}\right|\leq R^{2}, so that R>1R>1 on (0,T)(0,T). Then the estimate L2=ξL2+R21>ξL2L_{2}^{\prime}=-\xi L_{2}+R^{2}-1>-\xi L_{2} violates the boundary conditions limt0L2(t)=limt0ξ(t)=+\lim_{t\to 0}L_{2}(t)=\lim_{t\to 0}\xi(t)=+\infty and limtTL2(t)=0\lim_{t\to T}L_{2}(t)=0, limtTξ(t)=\lim_{t\to T}\xi(t)=-\infty. Since T>12T>\frac{1}{2}, Proposition 2 again implies that R21>0R^{2}-1>0 on (T12,T)(T-\frac{1}{2},T), so the boundary condtions L2(T)=0L_{2}(T)=0, limtTξ(t)=\lim_{t\to T}\xi(t)=-\infty and the inequality L2>ξL2L_{2}^{\prime}>-\xi L_{2} together imply that L2<0L_{2}<0.

With (42) in hand, we now claim that

L2(t)L1(t)1tTfor allt(0,T).\displaystyle L_{2}(t)\geq L_{1}(t)\geq\frac{1}{t-T}\ \text{for all}\ t\in(0,T). (43)

The second inequality in (43) is an immediate consequence of the fact that L1L12L_{1}^{\prime}\leq-L_{1}^{2} on (0,T)(0,T) (follows from Proposition 1), and limtTL1(t)=\lim_{t\to T}L_{1}(t)=-\infty. The first inequality is a consequence of the fact that

(L2L1)\displaystyle(L_{2}-L_{1})^{\prime} =ξ(L2L1)+R2,\displaystyle=-\xi(L_{2}-L_{1})+R^{2},

and the observation that limt0(L2L1)(t)=+\lim_{t\to 0}(L_{2}-L_{1})(t)=+\infty. In fact, we know that ξL1+1L120\xi L_{1}+1-L_{1}^{2}\geq 0 everywhere (Proposition 1 again), so since L10L_{1}\leq 0 everywhere, we rearrange to find ξ1L1L1-\xi\geq\frac{1}{L_{1}}-L_{1} so we can estimate further on (T12,T)(T-\frac{1}{2},T):

(L2L1)\displaystyle(L_{2}-L_{1})^{\prime} (1L1L1)(L2L1)+R2\displaystyle\geq(\frac{1}{L_{1}}-L_{1})(L_{2}-L_{1})+R^{2}
=L1(L2L1)+L2L1L1+R2\displaystyle=-L_{1}(L_{2}-L_{1})+\frac{L_{2}-L_{1}}{L_{1}}+R^{2}
(L2L1)2+L2L1L1+R2\displaystyle\geq(L_{2}-L_{1})^{2}+\frac{L_{2}-L_{1}}{L_{1}}+R^{2}
(L2L1)2+R21\displaystyle\geq(L_{2}-L_{1})^{2}+R^{2}-1
(L2L1)2+1(1δ3+Tt)21\displaystyle\geq(L_{2}-L_{1})^{2}+\frac{1}{(\frac{1}{\delta_{3}}+T-t)^{2}}-1

since L2<0L_{2}<0 (follows from (42)) and RR2R^{\prime}\leq R^{2} (Proposition 2). Let y=(L2L1)(Tt)y=(L_{2}-L_{1})(T-t), so that y(T12)0y(T-\frac{1}{2})\geq 0 and y(t)1y(t)\leq 1 for all t(T12,T)t\in(T-\frac{1}{2},T). But we can estimate the evolution of yy:

y\displaystyle y^{\prime} ((L2L1)2+1(1δ3+Tt)21)(Tt)(L2L1)\displaystyle\geq\left((L_{2}-L_{1})^{2}+\frac{1}{(\frac{1}{\delta_{3}}+T-t)^{2}}-1\right)(T-t)-(L_{2}-L_{1})
=y2y(Tt)+(1(1δ3+Tt)21)(Tt)\displaystyle=\frac{y^{2}-y}{(T-t)}+\left(\frac{1}{(\frac{1}{\delta_{3}}+T-t)^{2}}-1\right)(T-t)
14(Tt)+(Tt)(1δ3+Tt)2(Tt),\displaystyle\geq-\frac{1}{4(T-t)}+\frac{(T-t)}{(\frac{1}{\delta_{3}}+T-t)^{2}}-(T-t),

since y2y14y^{2}-y\geq-\frac{1}{4} for all yy\in\mathbb{R}. If δ3>40\delta_{3}>40, we integrate to get the following estimate:

1\displaystyle 1 y(T140)y(T12)\displaystyle\geq y(T-\frac{1}{40})-y(T-\frac{1}{2})
=T12T140y(t)𝑑t\displaystyle=\int_{T-\frac{1}{2}}^{T-\frac{1}{40}}y^{\prime}(t)dt
T12T140((Tt)(1δ3+Tt)2(Tt)14(Tt))𝑑t\displaystyle\geq\int_{T-\frac{1}{2}}^{T-\frac{1}{40}}\left(\frac{(T-t)}{(\frac{1}{\delta_{3}}+T-t)^{2}}-(T-t)-\frac{1}{4(T-t)}\right)dt
1.8918ln(20)4>1,\displaystyle\geq 1.89-\frac{1}{8}-\frac{\ln(20)}{4}>1,

which is a contradiction. ∎

Theorem 7.

Suppose the metric of the form (3) is a gradient shrinking Ricci soliton with Einstein constant λ=1\lambda=1. Then δ2[1,1020,000]\delta_{2}\in[-1,10^{20,000}].

Proof.

Once again consider the non-negative quantity defined by K(t)2=L2(t)2+(R(t)1)2K(t)^{2}=L_{2}(t)^{2}+(R(t)-1)^{2}, and note that

K(t)K(t)(12max{0,ξ}).\displaystyle K^{\prime}(t)\geq K(t)\left(-\frac{1}{2}-\max\{0,\xi\}\right). (44)

Also note that

Tξ1(0)2\displaystyle T-\xi^{-1}(0)\leq 2 (45)

because of the estimate yy2+1y^{\prime}\geq y^{2}+1, where y=min{ξ,L1}y=\min\{-\xi,-L_{1}\} and y(ξ1(0))0y(\xi^{-1}(0))\geq 0. Now Theorem 6 tells us that limtTK(t)39\lim_{t\to T}K(t)\leq 39, so (44) combined with (45) implies that

K(t)39efor allt(ξ1(0),T).\displaystyle K(t)\leq 39e\ \text{for all}\ t\in(\xi^{-1}(0),T). (46)

The equation L2=ξL2+R21L_{2}^{\prime}=-\xi L_{2}+R^{2}-1 then implies that |L2(t)|e(392e+239)(Tt)\left|L_{2}(t)\right|\leq e(39^{2}e+2\cdot 39)(T-t) for all t(ξ1(0),T)t\in(\xi^{-1}(0),T). Now consider the quantities X=ξL1X=\xi-L_{1} and Y=ξL1+1L12Y=\xi L_{1}+1-L_{1}^{2}. Since Y0Y\geq 0 everywhere, we find that L1(ξ1(0))[1,0]L_{1}(\xi^{-1}(0))\in[-1,0] and X(ξ1(0))[0,1]X(\xi^{-1}(0))\in[0,1]. We compute

X=L1X2L22,\displaystyle X^{\prime}=L_{1}X-2L_{2}^{2}, (47)

and

Y\displaystyle Y^{\prime} =L1(ξL1)+L1(ξL1)\displaystyle=L_{1}^{\prime}(\xi-L_{1})+L_{1}(\xi^{\prime}-L_{1}^{\prime})
=(ξL11)(ξL1)+L1(L122L221+ξL1+1)\displaystyle=(-\xi L_{1}-1)(\xi-L_{1})+L_{1}(-L_{1}^{2}-2L_{2}^{2}-1+\xi L_{1}+1)
=(ξL11+L12)(ξL1)2L1L22\displaystyle=(-\xi L_{1}-1+L_{1}^{2})(\xi-L_{1})-2L_{1}L_{2}^{2}
=XY2L1L22.\displaystyle=-XY-2L_{1}L_{2}^{2}.

Since L10L_{1}\leq 0 and Tξ1(0)2T-\xi^{-1}(0)\leq 2, we find from (46) and (47) that X(t)4(39e)2-X(t)\leq 4\cdot(39e)^{2} for all t(ξ1(0),T)t\in(\xi^{-1}(0),T). Therefore

Y4(39e)2Y+2(392e2+239e)2(Tt),\displaystyle Y^{\prime}\leq 4\cdot(39e)^{2}Y+2(39^{2}e^{2}+2\cdot 39e)^{2}(T-t),

so that Y(T)Y(T), which coincides with 32(δ2+1)\frac{3}{2}(\delta_{2}+1), can be no more than 1020,00010^{20,000}.

7 Compactness and uniqueness

We summarise the results for the 𝕊4\mathbb{S}^{4} we have seen so far.

Theorem 8.

An SO(2)×SO(3)SO(2)\times SO(3)-invariant gradient shrinking Ricci soliton on 𝕊4\mathbb{S}^{4} of the form (3) has δ1[0,1020,000]\delta_{1}\in[0,10^{20,000}], δ2[1,1020,000]\delta_{2}\in[-1,10^{20,000}] and δ3[0,40]\delta_{3}\in[0,40].

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 𝒞>0\mathcal{C}>0. Then there is a sequence of SO(2)×SO(3)SO(2)\times SO(3)-invariant solutions (g,u)(g,u) to (1) with unbounded Riemann curvature, unbounded volume, or an injectivity radius shrinking to 0. Theorem 8 implies that δ1,δ2,δ3\delta_{1},\delta_{2},\delta_{3} 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 (δ1,δ2,δ3)(\delta_{1},\delta_{2},\delta_{3}), 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 F:33F:\mathbb{R}^{3}\to\mathbb{R}^{3} whose zeroes are precisely the Ricci solitons we aim to classify. By Theorem 8, there is a compact domain Ω3\Omega\subset\mathbb{R}^{3} that contains all the zeroes of FF. In fact, we have explicitly described this domain. Therefore, Conjecture 1 would follow with the following steps.

  1. 1.

    Find an explicit open neighbourhood NN of the canonical metric on 𝕊4\mathbb{S}^{4} on which no other zeroes of FF occur. This is essentially a quantitative use of the inverse function theorem.

  2. 2.

    Show numerically that there are no zeroes of FF in ΩN\Omega\setminus N. This could be achieved by finding an upper bound for |dF|\left|dF\right| on Ω\Omega, discretising the set ΩN\Omega\setminus N accordingly, and showing that FF is sufficiently far away from 0 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 δ1,δ2\delta_{1},\delta_{2} 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 SO(2)×SO(3)SO(2)\times SO(3)-invariant ancient solution on 𝕊4\mathbb{S}^{4}

The sequence of ‘almost Ricci solitons’ on 𝕊4\mathbb{S}^{4} we found by making δ1\delta_{1} large appears to have a pancake shape. In this section, we describe a ‘pancake’ κ\kappa-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 κ\kappa-noncollapsed ancient Ricci flow on 𝕊4\mathbb{S}^{4} cannot be uniformly PIC and weakly PIC2.

We restate Theorem 2 for convenience.

Theorem 9.

There exists a κ>0\kappa>0 and a κ\kappa-noncollapsed SO(2)×SO(3)SO(2)\times SO(3)-invariant ancient Ricci flow on 𝕊4\mathbb{S}^{4} with positive Riemann curvature operator which is not isometric to the round shrinking sphere. The group SO(2)×SO(3)SO(2)\times SO(3) acts on 𝕊45=23\mathbb{S}^{4}\subseteq\mathbb{R}^{5}=\mathbb{R}^{2}\oplus\mathbb{R}^{3} 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 L>10L>10, choose an SO(2)×SO(3)SO(2)\times SO(3)-invariant Riemannian metric g0g_{0} on 𝕊4\mathbb{S}^{4} of the form (3) with T=L+1T=L+1 and

f~1(r)={Lif 0<r<1(L+1r)if 1r<L+1,f~2(r)={sin(πr2)if 0<r<11if 1r<L+1.\displaystyle\tilde{f}_{1}(r)=\begin{cases}L\ &\text{if}\ 0<r<1\\ (L+1-r)\ &\text{if}\ 1\leq r<L+1\end{cases},\qquad\tilde{f}_{2}(r)=\begin{cases}\sin(\frac{\pi r}{2})\ &\text{if}\ 0<r<1\\ 1\ &\text{if}\ 1\leq r<L+1.\end{cases}

This metric clearly satisfies the smoothness conditions at the singular orbits, but it does fail to be smooth at r=1r=1. However, we can mollify the two functions f~1,f~2\tilde{f}_{1},\tilde{f}_{2} on (12,32)(-\frac{1}{2},\frac{3}{2}) so that the resulting functions f1,f2f_{1},f_{2} are smooth. Recall that the Riemann curvatures of this Riemannian metric are

f1′′f1,f2′′f2,1(f2)2f22,f1f2f1f2;\displaystyle-\frac{f_{1}^{\prime\prime}}{f_{1}},\qquad-\frac{f_{2}^{\prime\prime}}{f_{2}},\qquad\frac{1-(f_{2}^{\prime})^{2}}{f_{2}^{2}},\qquad\frac{-f_{1}^{\prime}f_{2}^{\prime}}{f_{1}f_{2}};

it is clear that after mollification, these curvatures are all non-negative. Since we mollify on (12,32)(\frac{1}{2},\frac{3}{2}), the quantity f2′′f2+1(f2)2f22-\frac{f_{2}^{\prime\prime}}{f_{2}}+\frac{1-(f_{2}^{\prime})^{2}}{f_{2}^{2}} is uniformly bounded from below, independently of LL because f2f_{2} does not depend on LL in this region. Also the supremum of all four eigenvalues is uniformly bounded from above (LL does not affect the f2f_{2} terms, and only makes f1f_{1} large so the corresponding curvatures can only get smaller). Therefore,

there exists aC>0so that1CS(g0)=|(g0)|Cfor allL>10.\displaystyle\textit{there exists a}\ C>0\ \textit{so that}\ \frac{1}{C}\leq S(g_{0})=\left|\mathcal{R}(g_{0})\right|\leq C\ \textit{for all}\ L>10. (48)

Now we claim that there is a κ0>0\kappa_{0}>0 so that (M,g0)(M,g_{0}) is κ0\kappa_{0}-noncollapsed on all scales r0>0r_{0}>0, and for all large L>10L>10. To see this, we must show that any geodesic ball Br0(x0)B_{r_{0}}(x_{0}) on which S(g0)1r02S(g_{0})\leq\frac{1}{r_{0}^{2}} has volume at least κ0r04\kappa_{0}r_{0}^{4}. By the lower bound on the scalar curvature (48), it suffices to find a κ0>0\kappa_{0}>0 so that the volume of any geodesic ball Br0(x0)B_{r_{0}}(x_{0}) is at least κ0r04\kappa_{0}r_{0}^{4} whenever 0<r02C0<r_{0}^{2}\leq C. To this end, take an arbitrary geodesic ball Br0(x0)𝕊4B_{r_{0}}(x_{0})\subseteq\mathbb{S}^{4} and use the manifold decomposition:

𝕊4=AB¯,\displaystyle\mathbb{S}^{4}=\overline{A\cup B},

where A=(0,32]×𝕊1×𝕊2A=(0,\frac{3}{2}]\times\mathbb{S}^{1}\times\mathbb{S}^{2} and B=[32,L+1)×𝕊1×𝕊2B=[\frac{3}{2},L+1)\times\mathbb{S}^{1}\times\mathbb{S}^{2}. Now for each L>10L>10, BB is isometrically contained in 2×𝕊2\mathbb{R}^{2}\times\mathbb{S}^{2} equipped with the standard metric, so we have

inf{Vol(Br0(x0))r04|Br0(x0)B,r0(0,C]}inf{Vol(Br0(x0))r04|Br0(x0)2×𝕊2,r0(0,C]}.\displaystyle\inf\Bigg{\{}\ \frac{\text{Vol}(B_{r_{0}}(x_{0}))}{r_{0}^{4}}\ |\ B_{r_{0}}(x_{0})\subseteq B,r_{0}\in(0,C]\ \Bigg{\}}\geq\inf\Bigg{\{}\ \frac{\text{Vol}(B_{r_{0}}(x_{0}))}{r_{0}^{4}}\ |\ B_{r_{0}}(x_{0})\subseteq\mathbb{R}^{2}\times\mathbb{S}^{2},r_{0}\in(0,C]\ \Bigg{\}}. (49)

On the other hand, let g¯\overline{g} be an SO(3)SO(3)-invariant Riemannian metric on 𝕊3\mathbb{S}^{3} found by extending f2f_{2} on (0,32)(0,\frac{3}{2}) smoothly to a function on (0,2)(0,2) with the appropriate smoothness conditions at 22. Also let f1¯(r)\overline{f_{1}}(r) be a smooth extension of f1:(0,32)f_{1}:(0,\frac{3}{2})\to\mathbb{R} to a function with domain (0,2)(0,2) and appropriate smoothness conditions at r=2r=2. Note that g¯\overline{g} is independent of LL, and AA is isometrically embedded in the warped product manifold 𝕊1×𝕊3\mathbb{S}^{1}\times\mathbb{S}^{3} with metric (f1¯(r)2dθ2,g¯)(\overline{f_{1}}(r)^{2}d\theta^{2},\overline{g}), where dθd\theta is the standard one-form on 𝕊1\mathbb{S}^{1}. Therefore,

inf{Vol(Br0(x0))r04|Br0(x0)A,r0(0,C]}inf{Vol(Br0(x0))r04|Br0(x0)𝕊1×𝕊3,r0(0,C]}.\displaystyle\inf\Bigg{\{}\ \frac{\text{Vol}(B_{r_{0}}(x_{0}))}{r_{0}^{4}}\ |\ B_{r_{0}}(x_{0})\subseteq A,r_{0}\in(0,C]\ \Bigg{\}}\geq\inf\Bigg{\{}\ \frac{\text{Vol}(B_{r_{0}}(x_{0}))}{r_{0}^{4}}\ |\ B_{r_{0}}(x_{0})\subseteq\mathbb{S}^{1}\times\mathbb{S}^{3},r_{0}\in(0,C]\ \Bigg{\}}. (50)

Now it is well-known that the right hand side of (49), which is independent of LL, is strictly positive. Also, we can arrange it so that f1¯(r)1\overline{f_{1}}(r)\geq 1 for all r(0,2)r\in(0,2) and L>10L>10, so it is also clear that the right hand side of (50) is strictly positive for all L>10L>10, and can be bounded from below by a positive number, independently of L>10L>10.

Now any other geodesic ball of radius r0r_{0} contains a geodesic ball of radius r02\frac{r_{0}}{2} which is entirely contained in at least one of AA or BB, so the existence of a κ0>0\kappa_{0}>0 independent of LL follows from (49) and (50).

Step Two: Ricci flows from the initial metrics. Since the Riemann curvature of g0g_{0} is non-negative, for each L>10L>10 there exists a TLT_{L} so that the Ricci flow starting at g0g_{0} becomes singular for the first time at TLT_{L}, and with the round sphere as a singularity model. By (48), there exists a uniform C>0C^{\prime}>0 so that TL(1C,C)T_{L}\in(\frac{1}{C^{\prime}},C^{\prime}) for all L>10L>10. Furthermore, by (48) and the fact that each g0g_{0} is κ0\kappa_{0}-noncollapsed on all scales, Theorem 19.52 (no local collapsing) of [8] implies the existence of a κ>0\kappa>0 so that all of the Ricci flows on (2TL3,TL)(\frac{2T_{L}}{3},T_{L}) are uniformly κ\kappa-noncollapsed on all scales r2<TL3r^{2}<\frac{T_{L}}{3}. Note that we can apply this theorem because we can uniformly bound ||\left|\mathcal{R}\right| on the time interval [0,116𝒞][0,\frac{1}{16\mathcal{C}}] independently of LL; this is a result of the evolution equations for \mathcal{R} discussed in the proof of Proposition 1.

For a given small δ>0\delta>0, let tLt_{L} be the unique time at which the ratio of largest to smallest Riemann curvature eigenvalues is 1+δ1+\delta, and so that the ratio is less than or equal to 1+δ1+\delta on (tL,TL)(t_{L},T_{L}).

Step Three: rescaled flows. Consider the rescaled Ricci flows

g^L(t)=1TLtLgL(TL+(TLtL)t)\displaystyle\hat{g}_{L}(t)=\frac{1}{T_{L}-t_{L}}g_{L}(T_{L}+(T_{L}-t_{L})t)

which start at time t0(L)=TLTLtlt_{0}(L)=-\frac{T_{L}}{T_{L}-t_{l}}, have sectional curvature ratio 1+δ1+\delta at time 1-1, become singular at time 0, and are uniformly κ\kappa-noncollapsed on (TL3(TLtL),0)(-\frac{T_{L}}{3(T_{L}-t_{L})},0) for scales r2t0(L)3r^{2}\leq\frac{-t_{0}(L)}{3}.

We now show that limLt0(L)=\lim_{L\to\infty}t_{0}(L)=-\infty. To see this, note that for small enough δ>0\delta>0, we have

S^(1)3,\displaystyle\hat{S}(-1)\leq 3, (51)

thanks to the evolution equation

S^tΔg^(t)S^+S^22\displaystyle\frac{\partial\hat{S}}{\partial t}\geq\Delta_{\hat{g}(t)}\hat{S}+\frac{\hat{S}^{2}}{2}

and the pinching of 1+δ1+\delta that we have already established. The estimate (51) coupled with Hamilton’s trace Harnack inequality (Corollary 15.3 in [7]) then implies that S^(t)31t0(L)tt0(L)\hat{S}(t)\leq 3\frac{-1-t_{0}(L)}{t-t_{0}(L)} for t(t0(L),1]t\in(t_{0}(L),-1]. Therefore,

LTLtl\displaystyle\frac{L}{\sqrt{T_{L}-t_{l}}} diam(g^L(t0(L)))\displaystyle\leq\text{diam}(\hat{g}_{L}(t_{0}(L)))
=diam(g^L(1))t0(L)1diam(g^L(t))t𝑑t\displaystyle=\text{diam}(\hat{g}_{L}(-1))-\int_{t_{0}(L)}^{-1}\frac{\partial\text{diam}(\hat{g}_{L}(t))}{\partial t}dt
diam(g^L(1))+12t0(L)11t0(L)tt0(L)𝑑t\displaystyle\leq\text{diam}(\hat{g}_{L}(-1))+12\int_{t_{0}(L)}^{-1}\sqrt{\frac{-1-t_{0}(L)}{t-t_{0}(L)}}dt
=diam(g^L(1))+24(1t0(L))\displaystyle=\text{diam}(\hat{g}_{L}(-1))+24(-1-t_{0}(L))
24TLTLtl.\displaystyle\leq 24\frac{T_{L}}{T_{L}-t_{l}}.

Note that in the last computation, we have used the following facts:

  • For any x,y𝕊4x,y\in\mathbb{S}^{4}, d(x,y)t121t0(L)tt0(L)\frac{\partial d(x,y)}{\partial t}\geq-12\sqrt{\frac{-1-t_{0}(L)}{t-t_{0}(L)}}, which follows from the Ricci curvature upper bound.

  • By making δ\delta small, we can force g^L(1)\hat{g}_{L}(-1) to be arbitrarily close to a round sphere, and the scalar curvature of this particular sphere must be 22 because the remaining time until blow up is exactly 11. Therefore, Myer’s theorem implies that diam(g^L(1))23π\text{diam}(\hat{g}_{L}(-1))\leq 2\sqrt{3}\pi.

The estimate LTLtl24TLTLtl\frac{L}{\sqrt{T_{L}-t_{l}}}\leq 24\frac{T_{L}}{T_{L}-t_{l}} implies that limLTLtL=0\lim_{L\to\infty}T_{L}-t_{L}=0 as well, so limLt0(L)=\lim_{L\to\infty}t_{0}(L)=-\infty, because of our uniform estimates on TLT_{L} itself.

Step Four: convergence. We are now in a position to take limits. Indeed, the uniform κ\kappa-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 δ>0\delta>0 was eventually fixed, we can examine the sequence at time 1-1 to find that the ancient Ricci flow solution is on 𝕊4\mathbb{S}^{4}, but it is not the round sphere. By following the arguments in Chapter 19 of [6], we find that SO(2)×SO(3)SO(2)\times SO(3) acts via isometries on this ancient solution on 𝕊45=23\mathbb{S}^{4}\subset\mathbb{R}^{5}=\mathbb{R}^{2}\oplus\mathbb{R}^{3} in the obvious way. ∎

We conclude this section by observing that, not only is this SO(2)×SO(3)SO(2)\times SO(3)-invariant ancient solution non-round, but it is not isometric to Perelman’s rotationally-invariant κ\kappa solution either.

Proposition 6.

Let SO(2)×SO(3)SO(2)\times SO(3) act on 𝕊45=23\mathbb{S}^{4}\subset\mathbb{R}^{5}=\mathbb{R}^{2}\oplus\mathbb{R}^{3} in the obvious way. Any SO(2)×SO(3)SO(2)\times SO(3)-invariant continuous function f:𝕊4f:\mathbb{S}^{4}\to\mathbb{R} which is also invariant under an SO(4)SO(4) rotation group action must be constant.

Proof.

Let r1:𝕊4[0,1]r_{1}:\mathbb{S}^{4}\to[0,1] be a continuous function which parametrises the SO(4)SO(4) action, and let r2:𝕊4[0,1]r_{2}:\mathbb{S}^{4}\to[0,1] be a continuous function which parametrises the SO(2)×SO(3)SO(2)\times SO(3) action. Note that, almost by construction, we have the following:

  • r11(0)r_{1}^{-1}(0) and r11(1)r_{1}^{-1}(1) are single points;

  • r21(0)r_{2}^{-1}(0) is a copy of 𝕊1\mathbb{S}^{1};

  • r21(1)r_{2}^{-1}(1) is a copy of 𝕊2\mathbb{S}^{2};

  • r11(k)r_{1}^{-1}(k) is a copy of 𝕊3\mathbb{S}^{3} for each k(0,1)k\in(0,1);

  • r21(k)r_{2}^{-1}(k) is a copy of 𝕊1×𝕊2\mathbb{S}^{1}\times\mathbb{S}^{2} for each k(0,1)k\in(0,1);

  • the function ff is constant on the level sets ri1(k)r_{i}^{-1}(k) for each i=1,2i=1,2 and k[0,1]k\in[0,1].

We claim that there is an ϵ>0\epsilon>0 so that ff is constant on r11([0,ϵ])r_{1}^{-1}([0,\epsilon]) and r11([ϵ,1])r_{1}^{-1}([\epsilon,1]). To see this, note that ff must be constant on the two submanifolds (SO(2)×SO(3))r11(0)(SO(2)\times SO(3))\cdot r_{1}^{-1}(0) and (SO(2)×SO(3))r11(1)(SO(2)\times SO(3))\cdot r_{1}^{-1}(1). Since these submanifolds are compact, connected and at least one-dimensional, their images under the continuous function r1r_{1} must be non-trivial closed subintervals of [0,1][0,1] containing 0 or 11, respectively. However, since ff is constant on the level sets of r1r_{1}, ff must actually be constant on the r1r_{1} pre-images of these closed subintervals.

Choose ϵ~0(0,1]\tilde{\epsilon}_{0}\in(0,1] so that [0,ϵ~0][0,\tilde{\epsilon}_{0}] is the maximal connected subinterval of [0,1][0,1] containing 0 so that ff is constant on [0,ϵ~0][0,\tilde{\epsilon}_{0}]. Similarly, choose ϵ~1[0,1)\tilde{\epsilon}_{1}\in[0,1) so that [ϵ1,1][\epsilon_{1},1] is the maximal connected subinterval of [0,1][0,1] containing 11 so that ff is constant on [ϵ1,1][\epsilon_{1},1]. The argument in the last paragraph shows that ϵ~0\tilde{\epsilon}_{0} and ϵ~1\tilde{\epsilon}_{1} both exist. If ϵ~1ϵ~0\tilde{\epsilon}_{1}\leq\tilde{\epsilon}_{0}, the proof will be complete, so we assume that 0<ϵ~0<ϵ~1<10<\tilde{\epsilon}_{0}<\tilde{\epsilon}_{1}<1. Define the pairwise-disjoint connected sets A=r11([0,ϵ0~])A=r_{1}^{-1}([0,\tilde{\epsilon_{0}}]), B=r11((ϵ~0,ϵ~1))B=r_{1}^{-1}((\tilde{\epsilon}_{0},\tilde{\epsilon}_{1})), C=r11([ϵ~1,1])C=r_{1}^{-1}([\tilde{\epsilon}_{1},1]). It is clear that 𝕊4=ABC\mathbb{S}^{4}=A\cup B\cup C. Also note that both AA and CC are SO(2)×SO(3)SO(2)\times SO(3)-invariant because any orbit containing points both in and out of either of the sets would violate the maximality of [0,ϵ0~][0,\tilde{\epsilon_{0}}] or [ϵ~1,1][\tilde{\epsilon}_{1},1]. This implies that BB is SO(2)×SO(3)SO(2)\times SO(3)-invariant as well. Now r2(A),r2(B)r_{2}(A),r_{2}(B) and r2(C)r_{2}(C) are connected subintervals which must cover [0,1][0,1]; they must be pairwise disjoint because SO(2)×SO(3)SO(2)\times SO(3) orbits stay in exactly one of AA, BB or CC. Since AA and CC are compact, we find that r2(A)r_{2}(A) and r2(C)r_{2}(C) are closed, so r2(B)r_{2}(B) must be of the form (a,c)(a,c) for some 0<a<c<10<a<c<1. Therefore, BB must be homeomorphic to both (a,c)×𝕊1×𝕊2(a,c)\times\mathbb{S}^{1}\times\mathbb{S}^{2} and (ϵ~0,ϵ~1)×𝕊3(\tilde{\epsilon}_{0},\tilde{\epsilon}_{1})\times\mathbb{S}^{3}, 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 3\mathbb{R}^{3} of the form g=dtdt+f(t)2Qg=dt\otimes dt+f(t)^{2}Q, where QQ is the standard metric on 𝕊2\mathbb{S}^{2} with Ricci curvature 11, and f:(0,)(0,)f:(0,\infty)\to(0,\infty) is smooth, and can be extended to a smooth and odd function on (,)(-\infty,\infty) with f(0)=1f^{\prime}(0)=1. 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 R=1fR=\frac{1}{f}, L2=ffL_{2}=\frac{f^{\prime}}{f} and ξ=2L2u\xi=2L_{2}-u^{\prime}, where uu is the potential function, then

ξ=2L22,L2=ξL2+R2,R=L2R.\displaystyle\begin{split}\xi^{\prime}&=-2L_{2}^{2},\\ L_{2}^{\prime}&=-\xi L_{2}+R^{2},\\ R^{\prime}&=-L_{2}R.\end{split} (52)

For any p>0p>0, solutions of (52) are invariant under the transformation that sends a function f(t)f(t) to pf(pt)pf(pt), so to uniquely specify the Bryant soliton, we need to prescribe the value δ1:=limt0(R(t)1tt).\delta_{1}:=\lim_{t\to 0}\left(\frac{R(t)-\frac{1}{t}}{t}\right).

Let x=L2Rx=\frac{L_{2}}{R}, y=Rξy=\frac{R}{\xi} and z=1Rz=\frac{1}{R}, then

x=1yz(x+y+yx2),y=xy2+2x2y3yz,z=x.\displaystyle\begin{split}x^{\prime}&=\frac{1}{yz}\left(-x+y+yx^{2}\right),\\ y^{\prime}&=\frac{-xy^{2}+2x^{2}y^{3}}{yz},\\ z^{\prime}&=x.\end{split} (53)

The following facts about the Bryant soliton curve (x(t),y(t),z(t))(x(t),y(t),z(t)) are well-known (they are also discussed in [6]):

  • (x(0),y(0))=(1,12)(x(0),y(0))=(1,\frac{1}{2});

  • limt(x(t),y(t))=(0,0)\lim_{t\to\infty}(x(t),y(t))=(0,0);

  • x(t),y(t)<0x^{\prime}(t),y^{\prime}(t)<0 for all t(0,)t\in(0,\infty);

  • limty(t)x(t)=1\lim_{t\to\infty}\frac{y(t)}{x(t)}=1.

It is therefore clear that there is a function f:[0,1][0,12]f:[0,1]\to[0,\frac{1}{2}] so that y(t)=f(x(t))y(t)=f(x(t)) along the Bryant soliton, and this function does not depend on the choice of δ1>0\delta_{1}>0 (this parameter only affects how quickly one travels through the curve y=f(x)y=f(x)). The following propositions tell us some valuable information about this function.

Proposition 7.

The function ff satisfies x2f(x)x2+(1x)2\frac{x}{2}\leq f(x)\leq\frac{x}{2}+(1-x)^{2} for all x[34,1]x\in[\frac{3}{4},1]. In fact, x2f(x)\frac{x}{2}\leq f(x) for all x[0,1]x\in[0,1].

Proof.

Consider the system

x=x+y+yx2,y=xy2+2x2y3;\displaystyle\begin{split}x^{\prime}&=-x+y+yx^{2},\\ y^{\prime}&=-xy^{2}+2x^{2}y^{3};\end{split} (54)

the critical point (1,12)(1,\frac{1}{2}) has a one-dimensional unstable manifold. The part of the unstable manifold lying in [0,1]×[0,12][0,1]\times[0,\frac{1}{2}] consists of exactly those points of the form (x,f(x))(x,f(x)) for x[0,1]x\in[0,1]. It therefore suffices to show that any point in the unstable manifold with x[34,1]x\in[\frac{3}{4},1] satisfies

x2yx2+(1x)2.\displaystyle\frac{x}{2}\leq y\leq\frac{x}{2}+(1-x)^{2}. (55)

Our first step in verifying (55) for x[34,1]x\in[\frac{3}{4},1] is to find an ϵ>0\epsilon>0 so that (55) holds on [1ϵ,1][1-\epsilon,1]. To find such an ϵ>0\epsilon>0, note that ff must be smooth close to x=1x=1 because it describes an unstable manifold of a hyperbolic critical point of a smooth vector field. By looking at the linearisation of (54) at (1,12)(1,\frac{1}{2}), we see that the unstable manifold points in the direction of (2,1)(2,1), so f(1)=12f^{\prime}(1)=\frac{1}{2}. Now the function ff also satisfies

(x+f(x)+f(x)x2)f(x)=xf(x)2+2x2f(x)3\displaystyle(-x+f(x)+f(x)x^{2})f^{\prime}(x)=-xf(x)^{2}+2x^{2}f(x)^{3} (56)

for xx close to 11. Using the equalities f(1)=12f(1)=\frac{1}{2} and f(1)=12f^{\prime}(1)=\frac{1}{2}, we can write f(x)=x2+a(x1)2+O(x1)3f(x)=\frac{x}{2}+a(x-1)^{2}+O(x-1)^{3}. Then (56) becomes

x12+(x1)2(34+3a)+O(x1)3=x12+(x1)2(74+a2)+O(x1)3,\displaystyle\frac{x-1}{2}+(x-1)^{2}\left(\frac{3}{4}+3a\right)+O(x-1)^{3}=\frac{x-1}{2}+(x-1)^{2}\left(\frac{7}{4}+\frac{a}{2}\right)+O(x-1)^{3},

so a=25(0,1)a=\frac{2}{5}\in(0,1). The existence of the required ϵ>0\epsilon>0 follows.

Consider the quantities M1=yx2M_{1}=y-\frac{x}{2} and M2=yx2(x1)2M_{2}=y-\frac{x}{2}-(x-1)^{2}. We find that that y=0y^{\prime}=0 whenever M1=0M_{1}=0, so that M10M_{1}^{\prime}\geq 0 whenever M1=0M_{1}=0 and x[0,1]x\in[0,1]. On the other hand, we have

M2=xy2+2x2y3(12+2(x1))(x+y+yx2),\displaystyle M_{2}^{\prime}=-xy^{2}+2x^{2}y^{3}-(\frac{1}{2}+2(x-1))(-x+y+yx^{2}),

which is non-positive whenever M2=0M_{2}=0 and x[34,1]x\in[\frac{3}{4},1]. The required estimates follow. ∎

Proposition 8.

We have xx2f(x)xx-x^{2}\leq f(x)\leq x for x[0,14]x\in[0,\frac{1}{4}]. In fact, f(x)xf(x)\leq x for all x[0,1]x\in[0,1]

Proof.

As before, it suffices to consider the unstable trajectory of the system (54) which travels from (1,12)(1,\frac{1}{2}) to (0,0)(0,0). The estimate f(x)xf(x)\leq x follows from the fact that f(1)<1f(1)<1, and the observation that whenever the quantity M1=yxM_{1}=y-x vanishes and x(0,1)x\in(0,1), we have

M1\displaystyle M_{1}^{\prime} =xy2(2xy1)Myx2\displaystyle=xy^{2}(2xy-1)-M-yx^{2}
=2x3(x21)<0\displaystyle=2x^{3}\left(x^{2}-1\right)<0

so that M10M_{1}\leq 0 is preserved. On the other hand, consider the quantity M2=yx1+xM_{2}=\frac{y}{x}-1+x, so that

M2\displaystyle M_{2}^{\prime} =yxxyx2+x\displaystyle=\frac{y^{\prime}x-x^{\prime}y}{x^{2}}+x^{\prime}
=(xy2+2x2y3)x(x+y+yx2)yx2x+y+yx2\displaystyle=\frac{(-xy^{2}+2x^{2}y^{3})x-(-x+y+yx^{2})y}{x^{2}}-x+y+yx^{2}
=2y2+2xy3+(1yxyx)yxx+y+yx2\displaystyle=-2y^{2}+2xy^{3}+(1-\frac{y}{x}-yx)\frac{y}{x}-x+y+yx^{2}
=2y2+2xy3+(xM2yx)(M2+1x)x+y+yx2.\displaystyle=-2y^{2}+2xy^{3}+(x-M_{2}-yx)(M_{2}+1-x)-x+y+yx^{2}.

Whenever y=xx2y=x-x^{2} and x(0,310]x\in(0,\frac{3}{10}], we find that M2>0M_{2}^{\prime}>0. It therefore suffices to show that y>0.21y>0.21 when x=0.3x=0.3. To show this point, we consider the quantity M3=(y0.21)+0.3x5M_{3}=(y-0.21)+\frac{0.3-x}{5}. Then

M3=xy2+2x2y3+xyyx25\displaystyle M_{3}^{\prime}=-xy^{2}+2x^{2}y^{3}+\frac{x-y-yx^{2}}{5}

so that whenever M3=0M_{3}=0 and x[0.3,1]x\in[0.3,1], we have M3>0M_{3}^{\prime}>0. Since M3>0M_{3}>0 at the point (1,12)(1,\frac{1}{2}), we have that M3>0M_{3}>0 when x=0.3x=0.3, i.e., our solution curve has y>0.21y>0.21 when x=0.3x=0.3. ∎

We conclude this appendix with some short-time estimates on the original functions solving (53) with δ1=1\delta_{1}=1.

Theorem 10.

If δ1=1\delta_{1}=1, then up until time 19\frac{1}{9}, we have

sin(6t)6\displaystyle\frac{\sin(\sqrt{6}t)}{\sqrt{6}} z(t)t,\displaystyle\leq z(t)\leq t,
12tan2(32t)\displaystyle 1-2\tan^{2}\left(\sqrt{\frac{3}{2}}t\right) x(t)13t2e9t2.\displaystyle\leq x(t)\leq 1-3t^{2}e^{-9t^{2}}.
Proof.

For the zz estimate, note that z(t)tz(t)\leq t follows from curvature positivity of the Bryant soliton. For the lower bound, we use Proposition 7 to conclude that yx2y\geq\frac{x}{2}, so that

z′′\displaystyle z^{\prime\prime} =1yz(x+y+yx2)\displaystyle=\frac{1}{yz}\left(-x+y+yx^{2}\right)
xyz(1+(x2+1)2)\displaystyle\geq\frac{x}{yz}\left(-1+\frac{(x^{2}+1)}{2}\right)
2z(1+(x2+1)2)\displaystyle\geq\frac{2}{z}\left(-1+\frac{(x^{2}+1)}{2}\right)
=(z)21z.\displaystyle=\frac{(z^{\prime})^{2}-1}{z}.

Now since R(t)=1t+t+O(t3)R(t)=\frac{1}{t}+t+O(t^{3}), we find that z(0)=0=z′′(0)z(0)=0=z^{\prime\prime}(0), and z(0)=1z^{\prime}(0)=1 with z′′′(0)=6z^{\prime\prime\prime}(0)=-6. We therefore claim that

z(t)sin(6t)6,\displaystyle z(t)\geq\frac{\sin(\sqrt{6}t)}{\sqrt{6}}, (57)

provided t[0,19]t\in[0,\frac{1}{9}]. To see this, note that the solution sin(pt)p\frac{\sin(\sqrt{p}t)}{\sqrt{p}} for pt[0,π2]\sqrt{p}t\in[0,\frac{\pi}{2}] corresponds to a curve in (z,z)(z,z^{\prime}) space starting at (0,1)(0,1), with zz increasing and zz^{\prime} decreasing to the point (1p,0)(\frac{1}{\sqrt{p}},0). We write this curve as z=fp(z)z^{\prime}=f_{p}(z) for each p>0p>0. The equality z′′′(0)=6z^{\prime\prime\prime}(0)=-6 then implies that zf6(z)z^{\prime}\geq f_{6}(z), as long as z[0,16]z\in[0,\frac{1}{\sqrt{6}}]. The solution of z=f6(z)z^{\prime}=f_{6}(z) is precisely sin(6t)6\frac{\sin(\sqrt{6}t)}{\sqrt{6}}; estimate (57) follows for all tt so that 6t[0,π2]\sqrt{6}t\in[0,\frac{\pi}{2}].

We now move on to the xx estimates. Using z(t)tz(t)\leq t and Proposition 7, we find that

x\displaystyle x^{\prime} 1t((x1)(2x3x2+3x2)2(x2+(1x)2))\displaystyle\leq\frac{1}{t}\left(\frac{(x-1)(2x^{3}-x^{2}+3x-2)}{2\left(\frac{x}{2}+(1-x)^{2}\right)}\right)
2(x1)t(13(1x)),\displaystyle\leq\frac{2(x-1)}{t}\left(1-3(1-x)\right),

provided x[34,1]x\in[\frac{3}{4},1]. Using (52) and the fact that R(t)=1t+t+O(t3)R(t)=\frac{1}{t}+t+O(t^{3}), we find that L2(t)=1t2t+O(t3)L_{2}(t)=\frac{1}{t}-2t+O(t^{3}). One can then verify that x(0)=1,x(0)=0,x′′(0)=6x(0)=1,x^{\prime}(0)=0,x^{\prime\prime}(0)=-6. It is clear that (1x)x(t)(1-x)\geq x_{*}(t), where x(t)=2x(t)t(13x(t))x_{*}(t)^{\prime}=\frac{2x_{*}(t)}{t}(1-3x_{*}(t)) with x(0)=x(0)=0x_{*}(0)=x_{*}^{\prime}(0)=0, x′′(0)=6x_{*}^{\prime\prime}(0)=6. We can easily show that x(t)3t2x^{*}(t)\leq 3t^{2} for all t0t\geq 0, so that x(t)2x(t)t(19t2)x_{*}(t)^{\prime}\geq\frac{2x_{*}(t)}{t}(1-9t^{2}); solving this implies that (1x)(t)3t2e9t2(1-x)(t)\geq 3t^{2}e^{-9t^{2}}. For the xx lower bound, we estimate

x\displaystyle x^{\prime} x+x2(1+x2)yz\displaystyle\geq\frac{-x+\frac{x}{2}(1+x^{2})}{yz}
x+x2(1+x2)xz2\displaystyle\geq\frac{-x+\frac{x}{2}(1+x^{2})}{\frac{xz}{2}}
=x21z\displaystyle=\frac{x^{2}-1}{z}
6(x21)sin(6t)\displaystyle\geq\frac{\sqrt{6}(x^{2}-1)}{\sin(\sqrt{6}t)}
26(x1)sin(6t),\displaystyle\geq\frac{2\sqrt{6}(x-1)}{\sin(\sqrt{6}t)},

which implies that (1x)(t)2tan2(32t)(1-x)(t)\leq 2\tan^{2}(\sqrt{\frac{3}{2}}t). ∎

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.