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

Tangent Flows of Kähler Metric Flows

Max Hallgren and Wangjian Jian
Abstract.

We improve the description of 𝔽\mathbb{F}-limits of noncollapsed Ricci flows in the Kähler setting. In particular, the singular strata 𝒮k\mathcal{S}^{k} of such metric flows satisfy 𝒮2j=𝒮2j+1\mathcal{S}^{2j}=\mathcal{S}^{2j+1}. We also prove an analogous result for quantitative strata, and show that any tangent flow admits a nontrivial one-parameter action by isometries, which is locally free on the cone link in the static case. The main results are established using parabolic regularizations of conjugate heat kernel potential functions based at almost-selfsimilar points, which may be of independent interest.

1. Introduction

Suppose (Mi2n,gi,pi)(M_{i}^{2n},g_{i},p_{i}) is a sequence of pointed, complete Riemannian manifolds satisfying

(1.1) Rc(gi)\displaystyle Rc(g_{i}) (n1)gi,\displaystyle\geq-(n-1)g_{i},
(1.2) Vol(B(pi,1))\displaystyle\text{Vol}(B(p_{i},1)) ν,\displaystyle\geq\nu,

where ν>0\nu>0. Assume moreover that the sequence (Mi,dgi,pi)(M_{i},d_{g_{i}},p_{i}) converges in the pointed Gromov-Hausdorff sense to a metric space (X,d,p)(X,d,p_{\infty}). It was shown in [CC97] that any tangent cone based at a point in XX is a metric cone. The singular strata 𝒮k\mathcal{S}^{k} were defined to be set of xXx\in X such that no tangent cone based at xx is isometric to C(Z)×k+1C(Z)\times\mathbb{R}^{k+1}, where ZZ is any compact metric space with diameter at most π\pi. Moreover, the Hausdorff dimension estimates dim(𝒮k)k\text{dim}_{\mathcal{H}}(\mathcal{S}^{k})\leq k were established.

It is natural to ask what additional properties XX satisfies if (Mi,gi)(M_{i},g_{i}) are assumed to be Kähler. Theorem 9.1 of [CCT02] showed that in this case 𝒮2j=𝒮2j+1\mathcal{S}^{2j}=\mathcal{S}^{2j+1} for j=0,,n1j=0,...,n-1; roughly speaking, if a tangent cone C(Y)C(Y) of some xXx\in X splits a factor of 2j+1\mathbb{R}^{2j+1}, then it actually splits a factor of 2j+2\mathbb{R}^{2j+2}. It was also shown in [Liu18] that any tangent cone C(Y)C(Y) of XX admits a 1-parameter action by isometries, which extends to an effective isometric action by a torus. This action is used in [LS21] to obtain an embedding C(Y)NC(Y)\hookrightarrow\mathbb{C}^{N} whose image is a normal affine algebraic variety, such that the action on C(Y)C(Y) is the restriction of a linear torus action on N\mathbb{C}^{N}.

The goal of this paper is to prove analogous results in the setting of Ricci flow. A Ricci flow analogue of the singular stratification was first introduced for Ricci flows satisfying a Type-I curvature assumption in [Gia17]. A version defined for general Ricci flows was later defined and studied in [Bam20b]. To state our results more precisely, we let (Mi2n,(gi,t)t(Ti,0])(M_{i}^{2n},(g_{i,t})_{t\in(-T_{i},0]}) be any sequence of Ricci flows equipped with conjugate heat kernel measures (νxi,0;t)t(Ti,0](\nu_{x_{i},0;t})_{t\in(-T_{i},0]} based at (xi,0)(x_{i},0), where T:=limiTi(0,]T_{\infty}:=\lim_{i\to\infty}T_{i}\in(0,\infty] and 𝒩xi,0(1)Y\mathcal{N}_{x_{i},0}(1)\geq-Y for some Y<Y<\infty. In [Bam20a, Bam20, Bam20b], Bamler establishes a Ricci flow version of Gromov’s compactness theorem and Cheeger-Colding theory, where 𝔽\mathbb{F}-convergence takes the role of pointed Gromov-Hausdorff, and the volume noncollapsing assumption is replaced by the assumed lower bound for Nash entropy. We will review related definitions in Section 2. After passing to a subsequence, we can suppose that

(Mi2n,(gi,t)t(Ti,0],(νxi,0;t)t(Ti,0])i𝔽(𝒳,(νx;t)t(T,0])(M_{i}^{2n},(g_{i,t})_{t\in(-T_{i},0]},(\nu_{x_{i},0;t})_{t\in(-T_{i},0]})\xrightarrow[i\to\infty]{\mathbb{F}}(\mathcal{X},(\nu_{x_{\infty};t})_{t\in(-T_{\infty},0]})

uniformly on compact time intervals, for some future-continuous metric flow 𝒳\mathcal{X} of full support over (T,0](T_{\infty},0]. There is stratification 𝒮0𝒮2n2=𝒮\mathcal{S}^{0}\subseteq\cdots\subseteq\mathcal{S}^{2n-2}=\mathcal{S} of the singular set 𝒮\mathcal{S} of 𝒳\mathcal{X} analogous to that of Ricci limit spaces (see Section 2 for details). Our first main theorem can then be stated as follows.

Theorem 1.1.

If (Mi,(gi,t)t(Ti,0])(M_{i},(g_{i,t})_{t\in(-T_{i},0]}) are Kähler-Ricci flows, then 𝒮2j+1=𝒮2j\mathcal{S}^{2j+1}=\mathcal{S}^{2j} for j=0,,n1j=0,...,n-1.

For applications to smooth Ricci flows, it is often useful to study the quantitative stratification. A similar stratification was studied for Riemannian manifolds satisfying (1.1),(1.2) in [CN13], where it was used to prove LpL^{p} estimates for the Riemannian curvature tensor of Einstein manifolds. The Ricci flow version was again first studied for Type-I Ricci flows [Gia19], while two different but related definitions for general Ricci flows were used in [Bam20b]. These are the quantitative strata 𝒮r1,r2ϵ,k\mathcal{S}_{r_{1},r_{2}}^{\epsilon,k} and the weak quantitative strata 𝒮^r1,r2ϵ,k\widehat{\mathcal{S}}_{r_{1},r_{2}}^{\epsilon,k}. Our next result is a quantitative form of Theorem 1.1, from which Theorem 1.1 is easily derived. We note that our definition of quantitative strata 𝒮r1,r2ϵ,k\mathcal{S}_{r_{1},r_{2}}^{\epsilon,k} is slightly more restrictive than the definition given in [Bam20b] (see Section 2).

Theorem 1.2.

For any ϵ>0\epsilon>0 and Y,A<Y,A<\infty, there exists δ=δ(ϵ,Y,A)>0\delta=\delta(\epsilon,Y,A)>0 such that for all r2>r10r_{2}>r_{1}\geq 0 and j{0,,n1}j\in\{0,...,n-1\}, we have

𝒮r1,r2ϵ,2j+1P(x;A,A2)𝒮r1,r2δ,2jP(x;A,A2),\mathcal{S}_{r_{1},r_{2}}^{\epsilon,2j+1}\cap P^{\ast}(x_{\infty};A,-A^{2})\subseteq\mathcal{S}_{r_{1},r_{2}}^{\delta,2j}\cap P^{\ast}(x_{\infty};A,-A^{2}),
𝒮^r1,r2ϵ,2j+1P(x;A,A2)𝒮^r1,r2δ,2jP(x;A,A2).\widehat{\mathcal{S}}_{r_{1},r_{2}}^{\epsilon,2j+1}\cap P^{\ast}(x_{\infty};A,-A^{2})\subseteq\widehat{\mathcal{S}}_{r_{1},r_{2}}^{\delta,2j}\cap P^{\ast}(x_{\infty};A,-A^{2}).

We observe that unlike Theorem 1.1, Theorem 1.2 applies directly to smooth Kähler-Ricci flows, since the quantitative strata are generally nonempty even for smooth flows.

Next, we consider a fixed point x0𝒳t0x_{0}\in\mathcal{X}_{t_{0}} with t0<0t_{0}<0, and consider a sequence of parabolic rescalings

(𝒳t0,λk,(νx0;tt0,λk)t[λk2(t0T),0])),(\mathcal{X}^{-t_{0},\lambda_{k}},(\nu_{x_{0};t}^{-t_{0},\lambda_{k}})_{t\in[-\lambda_{k}^{2}(t_{0}-T_{\infty}),0])}),

where λk\lambda_{k}\nearrow\infty. After passing to a subsequence, we can assume 𝔽\mathbb{F}-convergence

(𝒳t0,λk,(νx0;tt0,λk)t[λk2(t0T),0]))i𝔽(𝒴,(νy;t)t(,0]),(\mathcal{X}^{-t_{0},\lambda_{k}},(\nu_{x_{0};t}^{-t_{0},\lambda_{k}})_{t\in[-\lambda_{k}^{2}(t_{0}-T_{\infty}),0])})\xrightarrow[i\to\infty]{\mathbb{F}}(\mathcal{Y},(\nu_{y_{\infty};t})_{t\in(-\infty,0]}),

where 𝒳t0,λk\mathcal{X}^{-t_{0},\lambda_{k}} is the metric flow 𝒳\mathcal{X} with a time translation by t0-t_{0} and a parabolic rescaling by λk\lambda_{k}, and 𝒴\mathcal{Y} is a metric soliton modeled on a singular shrinking Kähler-Ricci soliton (Y,dY,Y,gY,f)(Y,d_{Y},\mathcal{R}_{Y},g_{Y},f) with singularities of codimension four (see Section 2). We now give an analogue of Liu’s construction [Liu18] of a 1-parameter isometric action in the setting of these singular solitons.

Theorem 1.3.

(Y,dY)(Y,d_{Y}) admits a nontrivial 1-parameter action by isometries (σs)s(\sigma_{s})_{s\in\mathbb{R}} which preserves Y\mathcal{R}_{Y}. Moreover, the infinitesimal generator of the restriction to Y\mathcal{R}_{Y} is JfJ\nabla f. If in addition Rc(gY)=0Rc(g_{Y})=0, then YY is a metric cone (by [Bam20b]), and the action restricted to the cone link is locally free.

Taking the closure of this 1-parameter subgroup would induce a faithful action of a torus on YY, if we knew that the isometry group of (Y,dY)(Y,d_{Y}) is a Lie group – this holds for Ricci limit spaces by [CC00, CN12], and it is likely the arguments can be extended to certain 𝔽\mathbb{F}-limits of Ricci flows (c.f. Remark 2.7 of [Bam20b]). However, it is currently uncertain whether the arguments of [LS21] can be adapted to show that every tangent cone of a Kähler-Ricci flow is an affine variety. This is because the proof of Lemma 2.2 in [LS21] relied on sharp estimates (see [JN21]) for the size of singular sets of Gromov-Hausdorff limits of manifolds satisfying (1.2) and a two-sided Ricci curvature bound. The analogous estimates for 𝔽\mathbb{F}-limits of noncollapsed Ricci flows are so far unavailable.

The basic idea for proving Theorems 1.1 and 1.2 is similar to that in the case of Ricci-limit spaces [CCT02]. We first consider the model case where (M2n,g,J,f)(M^{2n},g,J,f) is a smooth, complete gradient Kähler-Ricci soliton which isometrically splits a factor of \mathbb{R}: (M,g,f)=(M×,g+dy2,f+y24)(M,g,f)=(M^{\prime}\times\mathbb{R},g^{\prime}+dy^{2},f^{\prime}+\frac{y^{2}}{4}), where (M,g,f)(M^{\prime},g^{\prime},f^{\prime}) is a gradient Ricci soliton of dimension 2n12n-1. Then z:=2f,Jyz:=2\langle\nabla f,J\nabla y\rangle satisfies

z=Jy2Rc(Jy)=Jy,\nabla z=J\nabla y-2Rc(J\nabla y)=J\nabla y,

since Rc(Jy)=JRc(y)=0Rc(J\nabla y)=JRc(\nabla y)=0. In particular, z\nabla z is parallel and pointwise orthogonal to y\nabla y. It follows that zz is a coordinate for another factor of \mathbb{R} split by (M,g)(M,g).

We now outline the steps we will take to implement this idea in the singular setting:

  • We show that any point y𝒳y\in\mathcal{X} close in 𝔽\mathbb{F}-distance to a metric soliton which either splits 2k+1\mathbb{R}^{2k+1} or is static and splits a factor of 2k1\mathbb{R}^{2k-1} is well-approximated by sequences of points (yi,ti)Mi×(Ti,0](y_{i},t_{i})\in M_{i}\times(-T_{i},0] which are (ϵ(ϵ),ri)(\epsilon^{\prime}(\epsilon),r_{i})-selfsimilar for some ri[r1,r2]r_{i}\in[r_{1},r_{2}], and either (2k+1,ϵ(ϵ),ri)(2k+1,\epsilon^{\prime}(\epsilon),r_{i})-split, or else (ϵ(ϵ),ri)(\epsilon^{\prime}(\epsilon),r_{i})-static and (2k1,ϵ(ϵ),ri)(2k-1,\epsilon^{\prime}(\epsilon),r_{i})-split, where limϵ0ϵ(ϵ)=0\lim_{\epsilon\to 0}\epsilon^{\prime}(\epsilon)=0. The converse was shown in [Bam20b], so it suffices to consider only the weak quantitative strata.

  • We construct a parabolic regularization qq associated to the conjugate heat kernel based at any almost-selfsimilar point (x0,t0)(x_{0},t_{0}), which satisfies estimates similar to 4τ(fW)4\tau(f-W), where ff is the potential function for a shrinking GRS.

  • Given a strong (2k+1,ϵ,r)(2k+1,\epsilon,r)-splitting map y=(y1,,y2k+1)y=(y_{1},...,y_{2k+1}) based at an almost-selfsimilar point (x0,t0)(x_{0},t_{0}), we show that the functions

    zi:=12q,Jyiz_{i}:=\frac{1}{2}\langle\nabla q,J\nabla y_{i}\rangle

    are almost-splitting functions whose gradients are almost-orthogonal to those of yiy_{i}, and use appropriate linear combinations of these functions and yiy_{i} to conclude that (x0,t0)(x_{0},t_{0}) is (2k+2,ϵ(ϵ),r)(2k+2,\epsilon^{\prime}(\epsilon),r)-split.

  • The previous step gives 𝒮^r1,r2ϵ(ϵ),2k+1𝒮^r1,r2ϵ,2k\widehat{\mathcal{S}}_{r_{1},r_{2}}^{\epsilon^{\prime}(\epsilon),2k+1}\subseteq\widehat{\mathcal{S}}_{r_{1},r_{2}}^{\epsilon,2k}, hence

    𝒮2k+1=ϵ(0,1)𝒮^0,ϵϵ(ϵ),2k+1ϵ(0,1)𝒮^0,ϵϵ,2k=𝒮2k.\mathcal{S}^{2k+1}=\cup_{\epsilon\in(0,1)}\widehat{\mathcal{S}}_{0,\epsilon}^{\epsilon^{\prime}(\epsilon),2k+1}\subseteq\cup_{\epsilon\in(0,1)}\widehat{\mathcal{S}}_{0,\epsilon}^{\epsilon,2k}=\mathcal{S}^{2k}.

In Section 2, we review definitions from [Bam20, Bam20b] relevant to our methods and results. In Section 3, we show that if a limiting metric soliton isometrically splits a factor of k\mathbb{R}^{k}, this can be used to find almost-split points in the approximating Ricci flows. In Section 4, we construct a parabolic regularization of approximate Ricci soliton potentials. In Section 5, we use these regularizations to construct almost-splitting maps on Kähler-Ricci flows, and finish the proof of Theorems 1.1 and 1.2. In Section 6, we construct isometric actions on tangent flows, and prove Theorem 1.3.

1.1. Acknowledgements

The authors would like to thank Xiaodong Cao and Jian Song for helpful discussions.

2. Preliminaries and Notation

Throughout this paper, we use the following notation convention of Cheeger-Colding theory (c.f. [CC96]): we let Ψ(a1,,ak|b1,,b)\Psi(a_{1},...,a_{k}|b_{1},...,b_{\ell}) denote a quantity depending on parameters a1,,ak,b1,,ba_{1},...,a_{k},b_{1},...,b_{\ell}, which satisfies

lim(a1,,ak)(0,,0)Ψ(a1,,ak|b1,,b)=0\lim_{(a_{1},...,a_{k})\to(0,...,0)}\Psi(a_{1},...,a_{k}|b_{1},...,b_{\ell})=0

for any fixed b1,,bb_{1},...,b_{\ell}. We also adhere to the following convention in [BK18]: if we say that a proposition P(ϵ)P(\epsilon) depending on a parameter ϵ\epsilon holds if ϵϵ¯(b1,,b)\epsilon\leq\overline{\epsilon}(b_{1},...,b_{\ell}), this means there exists a constant ϵ¯\overline{\epsilon} depending on parameters b1,,bb_{1},...,b_{\ell} such that P(ϵ)P(\epsilon) holds whenever ϵ(0,ϵ¯]\epsilon\in(0,\overline{\epsilon}]. The notation EE¯(b1,,b)E\geq\underline{E}(b_{1},...,b_{\ell}) is defined analogously. We also let 𝒫(X)\mathcal{P}(X) denote the space of Borel probability measures on a metric space XX.

The parabolic analogue of a metric space, defined in [Bam20], is that of a metric flow; metric flow pairs play the role of pointed metric spaces. The definition of a metric flow relies on an auxiliary function Φ(x):=x14πey24𝑑y\Phi(x):=\int_{-\infty}^{x}\frac{1}{\sqrt{4\pi}}e^{-\frac{y^{2}}{4}}dy.

Definition 2.1 (Metric Flow Pairs, Definitions 3.2, 5.1 in [Bam20]).

A metric flow over II\subseteq\mathbb{R} is a tuple (𝒳,𝔱,(dt)tI,(νx;s)x𝒳,sI(,𝔱(x)])(\mathcal{X},\mathfrak{t},(d_{t})_{t\in I},(\nu_{x;s})_{x\in\mathcal{X},s\in I\cap(-\infty,\mathfrak{t}(x)]}), where 𝒳\mathcal{X} is a set, 𝔱:𝒳I\mathfrak{t}:\mathcal{X}\to I is a function, dtd_{t} are metrics on the level sets 𝒳t:=𝔱1(t)\mathcal{X}_{t}:=\mathfrak{t}^{-1}(t), and νx;s𝒫(𝒳s)\nu_{x;s}\in\mathcal{P}(\mathcal{X}_{s}), s𝔱(x)s\leq\mathfrak{t}(x) are such that νx;𝔱(x)=δx\nu_{x;\mathfrak{t}(x)}=\delta_{x} and the following hold:
(i)(i) (Gradient estimate for heat flows) For s,tIs,t\in I, s<ts<t, T0T\geq 0, if us:𝒳s[0,1]u_{s}:\mathcal{X}_{s}\to[0,1] is such that Φ1us\Phi^{-1}\circ u_{s} is T12T^{-\frac{1}{2}}-Lipschitz (or just measurable if T=0T=0), then either ut:𝒳t[0,1]u_{t}:\mathcal{X}_{t}\to[0,1], x𝒳sus𝑑νx;sx\mapsto\int_{\mathcal{X}_{s}}u_{s}d\nu_{x;s}, is constant or Φ1ut\Phi^{-1}\circ u_{t} is (T+ts)12(T+t-s)^{-\frac{1}{2}}-Lipschitz,
(ii)(ii) (Reproduction formula) For t1t2t3t_{1}\leq t_{2}\leq t_{3} in II, νx;t1(E)=𝒳t2νy;t1(E)𝑑νx;t2(y)\nu_{x;t_{1}}(E)=\int_{\mathcal{X}_{t_{2}}}\nu_{y;t_{1}}(E)d\nu_{x;t_{2}}(y) for x𝒳t3x\in\mathcal{X}_{t_{3}} and all Borel sets E𝒳t1E\subseteq\mathcal{X}_{t_{1}}.
A conjugate heat flow on 𝒳\mathcal{X} is a family μt𝒫(𝒳t)\mu_{t}\in\mathcal{P}(\mathcal{X}_{t}), tIt\in I^{\prime}, such that for sts\leq t in II^{\prime}, we have μs(E)=𝒳tνx;s(E)𝑑μt(x)\mu_{s}(E)=\int_{\mathcal{X}_{t}}\nu_{x;s}(E)d\mu_{t}(x) for any Borel subset E𝒳sE\subseteq\mathcal{X}_{s}. A metric flow pair (𝒳,(μt)tI)(\mathcal{X},(\mu_{t})_{t\in I^{\prime}}) consists of a metric flow 𝒳\mathcal{X}, along with a conjugate heat flow (μt)tI(\mu_{t})_{t\in I^{\prime}} such that supp(μt)=𝒳t\text{supp}(\mu_{t})=\mathcal{X}_{t} and |II|=0|I\setminus I^{\prime}|=0.

The parabolic analogue of pointed Gromov-Hausdorff convergence is replaced with 𝔽\mathbb{F}-convergence, also introduced in [Bam20].

Definition 2.2 (Correspondences and 𝔽\mathbb{F}-Distance, Definitions 5.4, 5.6 in [Bam20]).

Given metric flows (𝒳i)i(\mathcal{X}^{i})_{i\in\mathcal{I}} defined over I,iI^{\prime,i}, a correspondence over I′′I^{\prime\prime}\subseteq\mathbb{R} is a pair

=((Zt,dt)tI′′,(φti)tI′′,i,i)\mathfrak{C}=\left((Z_{t},d_{t})_{t\in I^{\prime\prime}},(\varphi_{t}^{i})_{t\in I^{\prime\prime,i},i\in\mathcal{I}}\right)

where (Zt,dtZ)(Z_{t},d_{t}^{Z}) are metric spaces, I′′,iI,iI′′I^{\prime\prime,i}\subseteq I^{\prime,i}\cap I^{\prime\prime}, and φti:(𝒳ti,dti)(Zt,dtZ)\varphi_{t}^{i}:(\mathcal{X}_{t}^{i},d_{t}^{i})\to(Z_{t},d_{t}^{Z}) are isometric embeddings. The 𝔽\mathbb{F}-distance between metric flow pairs (𝒳j,(μtj)tI,j)(\mathcal{X}^{j},(\mu_{t}^{j})_{t\in I^{\prime,j}}), j=1,2j=1,2, within \mathfrak{C} is the infimum of r>0r>0 such that there exists a measurable set EI′′E\subseteq I^{\prime\prime} such that I′′EI′′,1I′′,2I^{\prime\prime}\setminus E\subseteq I^{\prime\prime,1}\cap I^{\prime\prime,2}, |E|r2|E|\leq r^{2}, and there exist couplings qtq_{t} of (μt1,μt2)(\mu_{t}^{1},\mu_{t}^{2}), tI′′Et\in I^{\prime\prime}\setminus E, such that for all s,tI′′Es,t\in I^{\prime\prime}\setminus E with sts\leq t, we have

𝒳t1×𝒳t2dW1Zs((φs1)νx1;s1,(φs2)νx2;s2)𝑑qt(x1,x2)r.\int_{\mathcal{X}_{t}^{1}\times\mathcal{X}_{t}^{2}}d_{W_{1}}^{Z_{s}}\left((\varphi_{s}^{1})_{\ast}\nu_{x^{1};s}^{1},(\varphi_{s}^{2})_{\ast}\nu_{x^{2};s}^{2}\right)dq_{t}(x^{1},x^{2})\leq r.

The 𝔽\mathbb{F}-distance between metric flow pairs is the infimum of 𝔽\mathbb{F}-distances within a correspondence \mathfrak{C}, where \mathfrak{C} is varied among all correspondences.

For the next definition, we suppose (𝒳i,(μti)tI,i)(\mathcal{X}^{i},(\mu_{t}^{i})_{t\in I^{\prime,i}}) 𝔽\mathbb{F}-converge to (𝒳,(μt)tI)(\mathcal{X}^{\infty},(\mu_{t}^{\infty})_{t\in I^{\prime\infty}}) within the correspondence \mathfrak{C}.

Definition 2.3 (Convergence within a correspondence, Definition 6.18 in [Bam20]).

Given μi𝒫(𝒳tii)\mu^{i}\in\mathcal{P}(\mathcal{X}_{t_{i}}^{i}) and μ𝒫(𝒳t)\mu^{\infty}\in\mathcal{P}(\mathcal{X}_{t_{\infty}}^{\infty}), we write μiiμ\mu^{i}\xrightarrow[i\to\infty]{\mathfrak{C}}\mu^{\infty} if titt_{i}\to t_{\infty} and there exist EiI′′E_{i}\subseteq I^{\prime\prime} such that |I′′Ei|0|I^{\prime\prime}\setminus E_{i}|\to 0, EiI′′E_{i}\subseteq I^{\prime\prime} and

limisuptI′′EdWZt((φti)μti,(φt)μt)=0,\lim_{i\to\infty}\sup_{t\in I^{\prime\prime}\setminus E}d_{W}^{Z_{t}}\left((\varphi_{t}^{i})_{\ast}\mu_{t}^{i},(\varphi_{t}^{\infty})_{\ast}\mu_{t}^{\infty}\right)=0,

where μti\mu_{t}^{i} is the conjugate heat flow on 𝒳i\mathcal{X}^{i} with μtii=μi\mu_{t_{i}}^{i}=\mu^{i}, for i{}i\in\mathbb{N}\cup\{\infty\}. We write xiixx_{i}\xrightarrow[i\to\infty]{\mathfrak{C}}x_{\infty} if δxiiδx\delta_{x_{i}}\xrightarrow[i\to\infty]{\mathfrak{C}}\delta_{x_{\infty}}.

Next, we recall Kleiner-Lott’s notion of a Ricci flow spacetime; it was shown in [Bam20, Bam20b] that the regular part of a metric flow obtained as a limit of Ricci flows possesses this structure.

Definition 2.4 (Ricci Flow Spacetime, Definition 1.2 in [KL17]).

A Ricci flow spacetime is a tuple (,𝔱,𝔱,g)(\mathcal{M},\mathfrak{t},\partial_{\mathfrak{t}},g) consisting of a manifold \mathcal{M}, a time function 𝔱:\mathfrak{t}:\mathcal{M}\to\mathbb{R}, a "time-like" vector field 𝔱𝔛()\partial_{\mathfrak{t}}\in\mathfrak{X}(\mathcal{M}) with 𝔱𝔱=1\partial_{\mathfrak{t}}\mathfrak{t}=1, and a bundle metric gg on the subbundle ker(d𝔱)T\ker(d\mathfrak{t})\subseteq T\mathcal{M} satisfying 𝔱g=2Rc(g)\mathcal{L}_{\partial_{\mathfrak{t}}}g=-2Rc(g), where Rc(g)|𝔱1(t)Rc(g)|\mathfrak{t}^{-1}(t) is defined to be the Ricci curvature of g|𝔱1(t)g|\mathfrak{t}^{-1}(t). We write t:=𝔱1(t)\mathcal{M}_{t}:=\mathfrak{t}^{-1}(t) and gt:=g|tg_{t}:=g|\mathcal{M}_{t}.

Given a Ricci flow (M,(gt)tI)(M,(g_{t})_{t\in I}) and some (x,t)M×I(x,t)\in M\times I, we let K(x,t;,):M×(I(,t))(0,)K(x,t;\cdot,\cdot):M\times(I\cap(-\infty,t))\to(0,\infty) denote the conjugate heat kernel based at (x,t)(x,t), and define dνx,t;s:=K(x,t;,s)dgs𝒫(M)d\nu_{x,t;s}:=K(x,t;\cdot,s)dg_{s}\in\mathcal{P}(M). We now summarize some of the main points of Bamler’s weak compactness and partial regularity theory. The notation in this statement will be used throughout the remainder of the paper.

Theorem 2.5 (c.f. Theorems 7.6, 9.12, 9.31 in [Bam20]).

Suppose (Min,(gi,t)t(Ti,0],(xi,0))(M_{i}^{n},(g_{i,t})_{t\in(-T_{i},0]},(x_{i},0)) is a sequence of pointed Ricci flows satisfying 𝒩xi,0(1)Y\mathcal{N}_{x_{i},0}(1)\geq-Y for some Y<Y<\infty. Then we can pass to a subsequence to obtain a future-continuous metric flow pair (𝒳,(νx,t)t(T,0])(\mathcal{X},(\nu_{x_{\infty},t})_{t\in(-T,0]}) along with a correspondence \mathfrak{C} such that we have the following 𝔽\mathbb{F}-convergence within the correspondence on compact time intervals:

(Min,(gi,t)t(Ti,0],(νxi,0;t)t(Ti,0])i𝔽,(𝒳,(νx,t)t(T,0]).(M_{i}^{n},(g_{i,t})_{t\in(-T_{i},0]},(\nu_{x_{i},0;t})_{t\in(-T_{i},0]})\xrightarrow[i\to\infty]{\mathbb{F},\mathfrak{C}}(\mathcal{X},(\nu_{x_{\infty},t})_{t\in(-T,0]}).

Moreover, there is an open, dense subset 𝒳\mathcal{R}\subseteq\mathcal{X} (with respect to the natural topology defined in Section 3 of [Bam20]) which admits the structure of a Ricci flow spacetime (,𝔱,𝔱,g)(\mathcal{R},\mathfrak{t},\partial_{\mathfrak{t}},g), where 𝔱\mathfrak{t} is the restricted function from the metric flow structure, and each (𝒳t,dt)(\mathcal{X}_{t},d_{t}) is the completion of the Riemannian length metric on (t,dgt)(\mathcal{R}_{t},d_{g_{t}}). In addition, the subbundle ker(d𝔱)T\ker(d\mathfrak{t})\subseteq T\mathcal{R} admits an endomorphism JJ satisfying 𝔱J=0\mathcal{L}_{\partial_{\mathfrak{t}}}J=0 and restricting to an almost-complex structure JtJ_{t} on each t\mathcal{R}_{t} such that each (t,gt,Jt)(\mathcal{R}_{t},g_{t},J_{t}) a Kähler manifold, and there is an increasing exhaustion (Ui)(U_{i}) of \mathcal{R} by precompact open sets along with time-preserving diffeomorphisms ψi:UiMi\psi_{i}:U_{i}\to M_{i} such that the following hold:
(i)(i) ψigig\psi_{i}^{\ast}g_{i}\to g in Cloc()C_{loc}^{\infty}(\mathcal{R}),
(ii)(ii) (ψi1)t𝔱(\psi_{i}^{-1})_{\ast}\partial_{t}\to\partial_{\mathfrak{t}} in Cloc()C_{loc}^{\infty}(\mathcal{R}),
(iii)(iii) If JiEnd(TMi)J_{i}\in\text{End}(TM_{i}) denote the given complex structures, then ψiJiJ\psi_{i}^{\ast}J_{i}\to J in Cloc()C_{loc}^{\infty}(\mathcal{R}),
(iv)(iv) If we write dνx,t=vtdgtd\nu_{x_{\infty},t}=v_{t}dg_{t} on \mathcal{R}, then ψiK(xi,0;,)v\psi_{i}^{\ast}K(x_{i},0;\cdot,\cdot)\to v in Cloc()C_{loc}^{\infty}(\mathcal{R}).

Proof.

By the mentioned theorems in [Bam20, Bam20b], it suffices to verify the claims concerning the complex structures. Because |Ji|gi,t=2n|J_{i}|_{g_{i,t}}=\sqrt{2n} and Ji=0\nabla J_{i}=0, the Arzela-Ascoli theorem lets us pass to a subsequence so that ψiJiJ\psi_{i}^{\ast}J_{i}\to J, where JJ restricts to an almost-complex structure on TtT\mathcal{R}_{t} for each t(T,0]t\in(-T,0]. Moreover, if ωi,tΩ2(Mi)\omega_{i,t}\in\Omega^{2}(M_{i}) denote the Kähler forms of (Mi,gi,t)(M_{i},g_{i,t}), then ψiωi,tωt\psi_{i}^{\ast}\omega_{i,t}\to\omega_{t}, where ωt(,):=gt(J,)\omega_{t}(\cdot,\cdot):=g_{t}(J\cdot,\cdot). Then dωi,t=0d\omega_{i,t}=0 and tJi=0\partial_{t}J_{i}=0 pass to the limit to give dωt=0d\omega_{t}=0 and 𝔱J=0\mathcal{L}_{\partial_{\mathfrak{t}}}J=0, so (t,gt,Jt)(\mathcal{R}_{t},g_{t},J_{t}) is Kähler, where Jt:=J|TtJ_{t}:=J|T\mathcal{R}_{t}. ∎

Next, we will recall Bamler’s description of the infinitesimal structure of the metric flows 𝒳\mathcal{X} obtained as 𝔽\mathbb{F}-limits of closed Ricci flows as in Theorem 2.5.

Definition 2.6 (Singular Solitons, Definition 2.15 in [Bam20b]).

A singular space is a tuple (X,d,X,gX)(X,d,\mathcal{R}_{X},g_{X}), where (X,d)(X,d) is a complete, locally compact metric length space, XX\mathcal{R}_{X}\subseteq X is a dense open subset admiting a smooth structure so that (X,gX)(\mathcal{R}_{X},g_{X}) is a smooth Riemannian manifold whose length metric is d|(X×X)d|(\mathcal{R}_{X}\times\mathcal{R}_{X}), and such that for any compact subset KXK\subseteq X and D<D<\infty, there exist 0<κ1(K,D)<κ2(K,D)<0<\kappa_{1}(K,D)<\kappa_{2}(K,D)<\infty such that for all xKx\in K and r(0,D)r\in(0,D), we have

κ1rn<|B(x,r)X|<κ2rn.\kappa_{1}r^{n}<|B(x,r)\cap\mathcal{R}_{X}|<\kappa_{2}r^{n}.

A singular shrinking gradient Ricci soliton (GRS) consists of a singular space along with a function fC()f\in C^{\infty}(\mathcal{R}) satisfying the Ricci soliton equation on X\mathcal{R}_{X}:

Rc(gX)+2f=12gX.Rc(g_{X})+\nabla^{2}f=\frac{1}{2}g_{X}.

In the following, we will not directly use the precise definition of a metric soliton or static flow modeled on a metric space, but the interested reader may find these definitions in Section 3.8 of [Bam20].

Theorem 2.7 (Theorems 2.6, 2.16, 2.18 in [Bam20b]).

If 𝒳\mathcal{X} is a metric flow obtained as in Theorem 2.5, and y𝒳y\in\mathcal{X}, t0:=𝔱(x)t_{0}:=\mathfrak{t}(x), then for any sequence λk\lambda_{k}\nearrow\infty, we can pass to a further subsequence so that the time shifted and parabolically rescaled metric flows (𝒳t0,λk,(νy;tt0,λk)t(λk2(t0T),0])(\mathcal{X}^{-t_{0},\lambda_{k}},(\nu_{y;t}^{-t_{0},\lambda_{k}})_{t\in(-\lambda_{k}^{2}(t_{0}-T),0]}) 𝔽\mathbb{F}-converge to a metric flow pair (𝒴,(νy;t)t(,0])(\mathcal{Y},(\nu_{y_{\infty};t})_{t\in(-\infty,0]}), where (𝒴<0,(νy;t)t(,0))(\mathcal{Y}_{<0},(\nu_{y_{\infty};t_{\infty}})_{t\in(-\infty,0)}) is a metric soliton modeled on a singular shrinking Kähler GRS (Y,d,Y,gY,JY,fY)(Y,d,\mathcal{R}_{Y},g_{Y},J_{Y},f_{Y}). Also, there are diffeomorphisms as in Theorem 2.5 realizing smooth convergence on the regular part of 𝒴\mathcal{Y}. There is an identification 𝒴<0Y×(,0)\mathcal{Y}_{<0}\cong Y\times(-\infty,0) restricting to isometries (𝒴t,dt)(Y,|t|d)(\mathcal{Y}_{t},d_{t})\cong(Y,\sqrt{|t|}d), and also identifying the spacetime 𝒴\mathcal{R}\subseteq\mathcal{Y} with (Y×(,0),t,tfY,|t|gY)(\mathcal{R}_{Y}\times(-\infty,0),t,\partial_{t}-\nabla f_{Y},|t|g_{Y}). Writing dνy;t=(4πτ)n2efdgtd\nu_{y_{\infty};t}=(4\pi\tau)^{-\frac{n}{2}}e^{-f}dg_{t}, we have that f(,t)f(\cdot,t) corresponds to fYf_{Y} for all t<0t<0 with respect to this identification.

If Rc(gY)=0Rc(g_{Y})=0, then 𝒴<0\mathcal{Y}_{<0} is a static metric flow modeled on the metric cone (Y,d)(Y,d) with vertex oo. Moreover, in this case, there is an identification 𝒴<0Y×(,0)\mathcal{Y}_{<0}\cong Y\times(-\infty,0) restricting to isometries (𝒴t,dt)(Y,d)(\mathcal{Y}_{t},d_{t})\cong(Y,d), and identifying the spacetime \mathcal{R} with (Y×(,0),t,t,gY)(\mathcal{R}_{Y}\times(-\infty,0),t,\partial_{t},g_{Y}); f(,t)f(\cdot,t) then corresponds to 14|t|d2(o,)+W\frac{1}{4|t|}d^{2}(o,\cdot)+W_{\infty} for each t<0t<0. Moreover, YB(o,1)\mathcal{R}_{Y}\cap\partial B(o,1) equipped with the restricted Riemannian metric is a Sasaki-Einstein manifold.

Proof.

By the mentioned theorems in [Bam20b], and by Theorem 2.5, it suffices to recall that a gradient Ricci soliton structure on a Kähler manifold is automatically a Kähler-Ricci soliton (see section 2.2 of [FIK03]). ∎

In order to define Bamler’s stratification of the singular set, we must review the notions of almost-split, almost-static, and almost-selfsimilar.

Definition 2.8 (Definitions 5.1, 5.5, 5.6, 5.7 in [Bam20b]).

Suppose (Mn,(gt)tI)(M^{n},(g_{t})_{t\in I}) is a closed Ricci flow, (x0,t0)M×I(x_{0},t_{0})\in M\times I, r>0r>0, ϵ(0,1)\epsilon\in(0,1), and write dνx0,t0=(4πτ)n2efdgd\nu_{x_{0},t_{0}}=(4\pi\tau)^{-\frac{n}{2}}e^{-f}dg.
(i)(i) (x0,t0)(x_{0},t_{0}) is (ϵ,r)(\epsilon,r)-selfsimilar if [t0ϵ1r2,t0]I[t_{0}-\epsilon^{-1}r^{2},t_{0}]\subseteq I and the following hold for W:=𝒩x,t(r2)W:=\mathcal{N}_{x,t}(r^{2}):

t0ϵ1r2t0ϵr2Mτ|Rc+2f12τg|2𝑑νx0,t0;t𝑑tϵ,\int_{t_{0}-\epsilon^{-1}r^{2}}^{t_{0}-\epsilon r^{2}}\int_{M}\tau\left|Rc+\nabla^{2}f-\frac{1}{2\tau}g\right|^{2}d\nu_{x_{0},t_{0};t}dt\leq\epsilon,
supt[t0ϵ1r2,t0ϵr2]M|τ(R+2Δf|f|2)+fnW|𝑑νx0,t0;tϵ,\sup_{t\in[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}]}\int_{M}\left|\tau(R+2\Delta f-|\nabla f|^{2})+f-n-W\right|d\nu_{x_{0},t_{0};t}\leq\epsilon,
infM×[t0ϵ1r2,t0ϵr2]r2Rϵ.\inf_{M\times[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}]}r^{2}R\geq-\epsilon.

(ii)(ii) (x0,t0)(x_{0},t_{0}) is (ϵ,r)(\epsilon,r)-static if [t0ϵ1r2,t0]I[t_{0}-\epsilon^{-1}r^{2},t_{0}]\subseteq I and the following hold:

r2t0ϵ1r2t0ϵr2M|Rc|2𝑑νx0,t0;t𝑑tϵ,r^{2}\int_{t_{0}-\epsilon^{-1}r^{2}}^{t_{0}-\epsilon r^{2}}\int_{M}|Rc|^{2}d\nu_{x_{0},t_{0};t}dt\leq\epsilon,
supt[t0ϵ1r2,t0ϵr2]MR𝑑νx0,t0;tϵ,\sup_{t\in[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}]}\int_{M}Rd\nu_{x_{0},t_{0};t}\leq\epsilon,
infM×[t0ϵ1r2,t0ϵr2]r2Rϵ.\inf_{M\times[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}]}r^{2}R\geq-\epsilon.

(iii)(iii) (x0,t0)(x_{0},t_{0}) is weakly (k,ϵ,r)(k,\epsilon,r)-split if [t0ϵ1r2,t0]I[t_{0}-\epsilon^{-1}r^{2},t_{0}]\subseteq I and there is a map y=(y1,,yk):M×[t0ϵ1r2,t0ϵr2]ky=(y_{1},...,y_{k}):M\times[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}]\to\mathbb{R}^{k}, called a weak (k,ϵ,r)(k,\epsilon,r)-splitting map, satisfying the following:

r1t0ϵ1r2t0ϵr2M|yi|𝑑νx0,t0;t𝑑tϵ,r^{-1}\int_{t_{0}-\epsilon^{-1}r^{2}}^{t_{0}-\epsilon r^{2}}\int_{M}|\square y_{i}|d\nu_{x_{0},t_{0};t}dt\leq\epsilon,
r2t0ϵ1r2t0ϵr2M|yi,yjδij|𝑑νx0,t0;t𝑑tϵ.r^{-2}\int_{t_{0}-\epsilon^{-1}r^{2}}^{t_{0}-\epsilon r^{2}}\int_{M}|\langle\nabla y_{i},\nabla y_{j}\rangle-\delta_{ij}|d\nu_{x_{0},t_{0};t}dt\leq\epsilon.

(iv)(iv) (x0,t0)(x_{0},t_{0}) is strongly (k,ϵ,r)(k,\epsilon,r)-split if [t0ϵ1r2,t0]I[t_{0}-\epsilon^{-1}r^{2},t_{0}]\subseteq I and there is a map y:M×[t0ϵ1r2,t0ϵr2]ky:M\times[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}]\to\mathbb{R}^{k}, called a strong (k,ϵ,r)(k,\epsilon,r)-splitting map, satisfying the following:

yi=0 on M×[t0ϵ1r2,t0ϵr2],\square y_{i}=0\text{ on }M\times[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}],
r2t0ϵ1r2t0ϵr2M|yi,yjδij|𝑑νx0,t0;t𝑑tϵ,r^{-2}\int_{t_{0}-\epsilon^{-1}r^{2}}^{t_{0}-\epsilon r^{2}}\int_{M}|\langle\nabla y_{i},\nabla y_{j}\rangle-\delta_{ij}|d\nu_{x_{0},t_{0};t}dt\leq\epsilon,
Myi𝑑νx0,t0;t=0 for all t[t0ϵ1r2,t0ϵr2].\int_{M}y_{i}d\nu_{x_{0},t_{0};t}=0\text{ for all }t\in[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}].

The following estimate gives rough LpL^{p} bounds on various geometric quantities in almost-selfsimilar regions, which we will use frequently.

Proposition 2.9 (c.f. Proposition 6.2 in [Bam20b]).

Given ϵ>0\epsilon>0, if αα¯\alpha\leq\overline{\alpha} and δδ¯(ϵ)\delta\leq\overline{\delta}(\epsilon), then the following holds. Suppose (Mn,(gt)tI)(M^{n},(g_{t})_{t\in I}) is a Ricci flow, r>0r>0, (x0,t0)M×I(x_{0},t_{0})\in M\times I is (δ,r)(\delta,r)-selfsimilar, W:=𝒩x0,t0(1)YW:=\mathcal{N}_{x_{0},t_{0}}(1)\geq-Y, and write (4πτ)n2efdg:=dν:=dνx0,t0(4\pi\tau)^{-\frac{n}{2}}e^{-f}dg:=d\nu:=d\nu_{x_{0},t_{0}}. Then

(2.1) t0ϵ1r2t0ϵr2M(τ|Rc|2+τ|2f|2+|f|2+τ|f|4+τ1eαf)e2αf𝑑νt𝑑tC(Y,ϵ),\int_{t_{0}-\epsilon^{-1}r^{2}}^{t_{0}-\epsilon r^{2}}\int_{M}\left(\tau|Rc|^{2}+\tau|\nabla^{2}f|^{2}+|\nabla f|^{2}+\tau|\nabla f|^{4}+\tau^{-1}e^{\alpha f}\right)e^{2\alpha f}d\nu_{t}dt\leq C(Y,\epsilon),
(2.2) supt[t0ϵ1r2,t0ϵr2]M(τ|R|+τ|Δf|+τ|f|2+eαf)e2αf𝑑νtC(Y,ϵ).\sup_{t\in[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}]}\int_{M}\left(\tau|R|+\tau|\Delta f|+\tau|\nabla f|^{2}+e^{\alpha f}\right)e^{2\alpha f}d\nu_{t}\leq C(Y,\epsilon).
Proof.

By Proposition 7.1 of [Bam20b], we have 𝒩x0,t0(τ)WΨ(δ|Y,ϵ)\mathcal{N}_{x_{0},t_{0}}(\tau)\geq W-\Psi(\delta|Y,\epsilon). We can therefore apply Proposition 6.2 of [Bam20b] for any r[ϵ12,ϵ12]r\in[\epsilon^{\frac{1}{2}},\epsilon^{-\frac{1}{2}}], taking θ:=ϵ2\theta:=\epsilon^{2}, to obtain (2.1), (2.2). ∎

We now review the quantitative stratification of metric flows introduced by Bamler. We will use a slightly stricter definition of (k,ϵ,r)(k,\epsilon,r)-symmetric points than that in [Bam20b]. Let (μtk)t<0(\mu_{t}^{\mathbb{R}^{k}})_{t<0} be the Euclidean backwards heat kernel based at 0kk0^{k}\in\mathbb{R}^{k}.

Definition 2.10 ((k,ϵ,r)(k,\epsilon,r)-symmetric points, c.f. Definition 2.21 in [Bam20]).

Given a metric flow 𝒳\mathcal{X} over II, a point x0𝒳t0x_{0}\in\mathcal{X}_{t_{0}} is (k,ϵ,r)(k,\epsilon,r)-symmetric if [t0ϵ1r2,t0]I[t_{0}-\epsilon^{-1}r^{2},t_{0}]\subseteq I and there is a metric flow pair (𝒳,(μt)t0)(\mathcal{X}^{\prime},(\mu_{t}^{\prime})_{t\leq 0}) over (,0](-\infty,0] which is an 𝔽\mathbb{F}-limit of noncollapsed Ricci flows as in Theorem 2.5, and which satisfies one of the following:

  • (b1)(b1)

    (𝒳<0,(μt)t(,0))(𝒳′′×k,(μt′′μk)t(,0))(\mathcal{X}^{\prime}_{<0},(\mu_{t}^{\prime})_{t\in(-\infty,0)})\cong(\mathcal{X}^{\prime\prime}\times\mathbb{R}^{k},(\mu_{t}^{\prime\prime}\otimes\mu_{\mathbb{R}^{k}})_{t\in(-\infty,0)}) as metric flow pairs for some metric soliton (𝒳′′,(μt′′)t(,0))(\mathcal{X}^{\prime\prime},(\mu_{t}^{\prime\prime})_{t\in(-\infty,0)}), and this identification restricts to an isometry of Ricci flow spacetimes ′′×k\mathcal{R}^{\prime}\cong\mathcal{R}^{\prime\prime}\times\mathbb{R}^{k},

  • (b2)(b2)

    (𝒳<0,(μt)t(,0))(𝒳′′×k2,(νx;tμk2)t(,0))(\mathcal{X}^{\prime}_{<0},(\mu_{t}^{\prime})_{t\in(-\infty,0)})\cong(\mathcal{X}^{\prime\prime}\times\mathbb{R}^{k-2},(\nu_{x^{\prime};t}^{\prime}\otimes\mu_{\mathbb{R}^{k-2}})_{t\in(-\infty,0)}) as metric flow pairs for some static cone 𝒳′′\mathcal{X}^{\prime\prime} with vertex xx^{\prime}, and this identification restricts to an isometry of Ricci flow spacetimes ′′×k2\mathcal{R}^{\prime}\cong\mathcal{R}^{\prime\prime}\times\mathbb{R}^{k-2}.

In addition, 𝒳,𝒳′′\mathcal{X}^{\prime},\mathcal{X}^{\prime\prime} must satisfy the following:

  • (c)(c)

    Writing dμt=(4πτ)n2efdgd\mu_{t}^{\prime}=(4\pi\tau)^{-\frac{n}{2}}e^{-f^{\prime}}dg on \mathcal{R}^{\prime} and dμt′′=(4πτ)n2ef′′dg′′d\mu_{t}^{\prime\prime}=(4\pi\tau)^{-\frac{n}{2}}e^{-f^{\prime\prime}}dg^{\prime\prime} on \mathcal{R}^{\prime}, we have Rc(g)+2f=12τgRc(g^{\prime})+\nabla^{2}f^{\prime}=\frac{1}{2\tau}g^{\prime} on \mathcal{R}^{\prime}, Rc(g′′)+2f′′=12τg′′Rc(g^{\prime\prime})+\nabla^{2}f^{\prime\prime}=\frac{1}{2\tau}g^{\prime\prime} on ′′\mathcal{R}^{\prime\prime}, and f=f′′+14τ|x|2f^{\prime}=f^{\prime\prime}+\frac{1}{4\tau}|x|^{2} on \mathcal{R}^{\prime}. In case (b2)(b2), we also have Rc(g)=0Rc(g^{\prime})=0 and Rc(g′′)=0Rc(g^{\prime\prime})=0,

  • (d)(d)

    𝒩(μt)(τ)=W\mathcal{N}_{(\mu_{t}^{\prime})}(\tau)=W for all τ>0\tau>0, where W[Y,0]W\in[-Y,0].

Finally, we require that |𝒩x0(r2)W|<ϵ|\mathcal{N}_{x_{0}}(r^{2})-W|<\epsilon and

d𝔽((𝒳[ϵ1,0]t0,r1,(νx0;tt0,r1)t[ϵ1,0]),(𝒳[ϵ1,0],(μt)t[ϵ1,0]))<ϵ.d_{\mathbb{F}}\left((\mathcal{X}_{[-\epsilon^{-1},0]}^{-t_{0},r^{-1}},(\nu_{x_{0};t}^{-t_{0},r^{-1}})_{t\in[-\epsilon^{-1},0]}),(\mathcal{X}_{[-\epsilon^{-1},0]}^{\prime},(\mu_{t}^{\prime})_{t\in[-\epsilon^{-1},0]})\right)<\epsilon.

The main difference between Definition 2.21 in [Bam20b] and Definition 2.10 is the added assumptions on the Nash entropy.

There is another notion of almost-symmetric points of a metric flow, defined in terms of smooth Ricci flow approximants.

Definition 2.11 (Weakly (k,ϵ,r)(k,\epsilon,r)-symmetric points, Definition 20.1 in [Bam20b]).

Given a metric flow 𝒳\mathcal{X} over II, a point x𝒳t0x\in\mathcal{X}_{t_{0}} is weakly (k,ϵ,r)(k,\epsilon,r)-symmetric if [t0ϵ1r2,t0]I[t_{0}-\epsilon^{-1}r^{2},t_{0}]\subseteq I and there is a closed, pointed Ricci flow (M,(gt)t[ϵ1,0],x)(M^{\prime},(g_{t}^{\prime})_{t\in[-\epsilon^{-1},0]},x^{\prime}) such that (x,0)(x^{\prime},0) is (ϵ,1)(\epsilon,1)-selfsimilar and either strongly (k,ϵ,1)(k,\epsilon,1)-split or both (ϵ,1)(\epsilon,1)-static and strongly (k2,ϵ,1)(k-2,\epsilon,1)-split, which satisfies |𝒩x,0(1)𝒩x0(r2)|ϵ|\mathcal{N}_{x^{\prime},0}(1)-\mathcal{N}_{x_{0}}(r^{2})|\leq\epsilon and

d𝔽((𝒳t0,r,(νx0;tt0,r)t[ϵ1,0]),(M,(gt)t[ϵ1,0],(νx,0;t)t[ϵ1,0]))<ϵ.d_{\mathbb{F}}\left((\mathcal{X}^{-t_{0},r},(\nu_{x_{0};t}^{-t_{0},r})_{t\in[-\epsilon^{-1},0]}),(M^{\prime},(g_{t}^{\prime})_{t\in[-\epsilon^{-1},0]},(\nu_{x^{\prime},0;t})_{t\in[-\epsilon^{-1},0]})\right)<\epsilon.
Definition 2.12 (Strata and Quantitative strata of a metric flow, Definitions 2.21, 20.2 in [Bam20b]).

If 𝒳\mathcal{X} is a metric flow over II and 0r1<r2<0\leq r_{1}<r_{2}<\infty, ϵ>0\epsilon>0, then 𝒮r1,r2ϵ,k\mathcal{S}_{r_{1},r_{2}}^{\epsilon,k} consists of the points x𝒳tx\in\mathcal{X}_{t} such that [tϵ1r22,t]I[t-\epsilon^{-1}r_{2}^{2},t]\subseteq I and xx is not (k+1,ϵ,r)(k+1,\epsilon,r^{\prime})-symmetric for any r(r1,r2)r^{\prime}\in(r_{1},r_{2}). Analogously, x𝒮^r1,r2ϵ,kx\in\widehat{\mathcal{S}}_{r_{1},r_{2}}^{\epsilon,k} if [tϵ1r22,t]I[t-\epsilon^{-1}r_{2}^{2},t]\subseteq I and xx is not weakly (k+1,ϵ,r)(k+1,\epsilon,r^{\prime})-symmetric for any r(r1,r2)r^{\prime}\in(r_{1},r_{2}). Define

𝒮k:=ϵ(0,1)𝒮^0,ϵϵ,k.\mathcal{S}^{k}:=\cup_{\epsilon\in(0,1)}\widehat{\mathcal{S}}_{0,\epsilon}^{\epsilon,k}.

Note that, because ϵ𝒮^0,rϵ,k\epsilon\mapsto\widehat{\mathcal{S}}_{0,r}^{\epsilon,k} and r𝒮^0,rϵ,kr\mapsto\widehat{\mathcal{S}}_{0,r}^{\epsilon,k} are both decreasing, we can write

𝒮k=ϵ(0,1)r(0,1)𝒮^0,rϵ,k.\mathcal{S}^{k}=\cup_{\epsilon\in(0,1)}\cup_{r\in(0,1)}\widehat{\mathcal{S}}_{0,r}^{\epsilon,k}.

Bamler showed that, roughly speaking, the weak quantitative strata are qualitatively at least as large as the quantitative strata.

Lemma 2.13 (c.f. Lemma 20.3 in [Bam20b]).

Suppose 𝒳\mathcal{X} is an 𝔽\mathbb{F}-limit of closed noncollapsed Ricci flows as in Theorem 2.5. Given Y<Y<\infty, ϵ>0\epsilon>0 there exists ϵ(Y,ϵ)>0\epsilon^{\prime}(Y,\epsilon)>0 such that for all 0r1<r2<0\leq r_{1}<r_{2}<\infty, we have

{x𝒳;𝒩x(r22)Y}𝒮r1,r2ϵ,k𝒮^r1,r2ϵ(Y,ϵ),k.\{x\in\mathcal{X};\mathcal{N}_{x}(r_{2}^{2})\geq-Y\}\cap\mathcal{S}_{r_{1},r_{2}}^{\epsilon,k}\subseteq\widehat{\mathcal{S}}_{r_{1},r_{2}}^{\epsilon^{\prime}(Y,\epsilon),k}.

As a consequence, we have

ϵ(0,1)𝒮0,ϵϵ,k𝒮k.\cup_{\epsilon\in(0,1)}\mathcal{S}_{0,\epsilon}^{\epsilon,k}\subseteq\mathcal{S}^{k}.
Proof.

Suppose x𝒳𝒮^r1,r2ϵ(Y,ϵ),kx\in\mathcal{X}\setminus\widehat{\mathcal{S}}_{r_{1},r_{2}}^{\epsilon^{\prime}(Y,\epsilon),k} and 𝒩x(r22)Y\mathcal{N}_{x}(r_{2}^{2})\geq-Y. By definition, there exists r(r1,r2)r\in(r_{1},r_{2}) such that xx is weakly (k,ϵ(Y,ϵ),r)(k,\epsilon^{\prime}(Y,\epsilon),r)-symmetric. Lemma 20.3 of [Bam20b] then states that xx is (k,ϵ,r)(k,\epsilon,r)-symmetric, hence x𝒮r1,r2ϵ,kx\notin\mathcal{S}_{r_{1},r_{2}}^{\epsilon,k}. We note that this Lemma holds even with our stricter definition of (k,ϵ,r)(k,\epsilon^{\prime},r)-symmetric points, by Nash entropy convergence (Theorem 15.45 of [Bam20b]) and Proposition 7.1 of [Bam20b].

Taking r1=0r_{1}=0, r2=ϵr_{2}=\epsilon, and taking the union over ϵ(0,1)\epsilon\in(0,1) gives

{x𝒳;𝒩x(1)Y}(ϵ(0,1)𝒮0,ϵϵ,k)𝒮k,\{x\in\mathcal{X};\mathcal{N}_{x}(1)\geq-Y\}\cap\left(\cup_{\epsilon\in(0,1)}\mathcal{S}_{0,\epsilon}^{\epsilon,k}\right)\subseteq\mathcal{S}^{k},

so the remaining claim follows by taking YY\nearrow\infty. ∎

Remark 2.14.

We will later show that ϵ(0,1)𝒮0,ϵϵ,k=𝒮k\cup_{\epsilon\in(0,1)}\mathcal{S}_{0,\epsilon}^{\epsilon,k}=\mathcal{S}^{k}.

We now recall a result from [Bam20b] asserting the existence of good cutoff functions vanishing near the singular set of an 𝔽\mathbb{F}-limit of Ricci flows, which will be useful in Section 6. Assume 𝒳\mathcal{X} is a metric flow as in Theorem 2.5.

Lemma 2.15 (Lemma 15.27 in [Bam20b]).

There is a family of smooth functions ηrC(,[0,1])\eta_{r}\in C^{\infty}(\mathcal{R},[0,1]) satisfying the following:
(i)(i) rRmrr_{Rm}\geq r on {ηr>0}\{\eta_{r}>0\},
(ii)(ii) ηr|{rRm2r}1\eta_{r}|\{r_{Rm}\geq 2r\}\equiv 1,
(iii)(iii) |ηr|C0r1|\nabla\eta_{r}|\leq C_{0}r^{-1}, |𝔱ηr|C0r2|\partial_{\mathfrak{t}}\eta_{r}|\leq C_{0}r^{-2}, where C0<C_{0}<\infty is universal
(iv)(iv) for any x𝒳tx\in\mathcal{X}_{t}, A<A<\infty, and r>0r>0, the set {ηr>0}P(x;A,A2)t\{\eta_{r}>0\}\cap P^{\ast}(x_{\infty};A,-A^{2})\cap\mathcal{R}_{t} is relatively compact in t\mathcal{R}_{t},
(v)(v) for any maximal integral curve γ:I\gamma:I\to\mathcal{R} of 𝔱\mathfrak{t} (assume 𝔱(γ(t))=t\mathfrak{t}(\gamma(t))=t by a constant reparametrization), we have either ηr(γ(t))=0\eta_{r}(\gamma(t))=0 for tt near tmin:=inf(I)t_{\min}:=\inf(I), or else tmin=Tt_{\min}=-T.

3. Sequences of Noncollapsed Ricci Flows whose 𝔽\mathbb{F}-Limits are Split Solitons

In this section, we show the qualitative equivalence of Bamler’s notions of quantitative strata and weak quantitative strata, with one direction already established in [Bam20b]. We begin with an elementary lemma concerning the 1-Wasserstein distance between product measures.

Lemma 3.1.

Suppose (X,dX),(Y,dY)(X,d^{X}),(Y,d^{Y}) are metric spaces and μ1,μ2𝒫(X)\mu_{1},\mu_{2}\in\mathcal{P}(X), ν1,ν2𝒫(Y)\nu_{1},\nu_{2}\in\mathcal{P}(Y). Then

dW1X×Y(μ1ν1,μ2ν2)dW1X(μ1,μ2)+dW1Y(ν1,ν2).d_{W_{1}}^{X\times Y}\left(\mu_{1}\otimes\nu_{1},\mu_{2}\otimes\nu_{2}\right)\leq d_{W_{1}}^{X}(\mu_{1},\mu_{2})+d_{W_{1}}^{Y}(\nu_{1},\nu_{2}).
Proof.

Suppose qXq_{X} is a coupling of (μ1,μ2)(\mu_{1},\mu_{2}) and qYq_{Y} is a coupling of (ν1,ν2)(\nu_{1},\nu_{2}). Define σ:X×X×Y×YX×Y×X×Y\sigma:X\times X\times Y\times Y\to X\times Y\times X\times Y, (x1,x2,y1,y2)(x1,y1,x2,y2)(x_{1},x_{2},y_{1},y_{2})\mapsto(x_{1},y_{1},x_{2},y_{2}), and q:=σ(qXqY)q:=\sigma_{\ast}(q_{X}\otimes q_{Y}) . Then qq is a coupling of (μ1ν1,μ2ν2)(\mu_{1}\otimes\nu_{1},\mu_{2}\otimes\nu_{2}), so we can estimate

dW1X×Y(μ1ν1,μ2ν2)\displaystyle d_{W_{1}}^{X\times Y}(\mu_{1}\otimes\nu_{1},\mu_{2}\otimes\nu_{2})\leq X×Y×X×YdX×Y((x1,y1),(x2,y2))𝑑q(x1,y1,x2,y2)\displaystyle\int_{X\times Y\times X\times Y}d^{X\times Y}\left((x_{1},y_{1}),(x_{2},y_{2})\right)dq(x_{1},y_{1},x_{2},y_{2})
\displaystyle\leq X×X×Y×Y(dX(x1,x2)+dY(y1,y2))𝑑qX(x1,x2)𝑑qY(y1,y2)\displaystyle\int_{X\times X\times Y\times Y}\left(d^{X}(x_{1},x_{2})+d^{Y}(y_{1},y_{2})\right)dq_{X}(x_{1},x_{2})dq_{Y}(y_{1},y_{2})
=\displaystyle= X×XdX(x1,x2)𝑑qX(x1,x2)+Y×YdY(y1,y2)𝑑qY(y1,y2).\displaystyle\int_{X\times X}d^{X}(x_{1},x_{2})dq_{X}(x_{1},x_{2})+\int_{Y\times Y}d^{Y}(y_{1},y_{2})dq_{Y}(y_{1},y_{2}).

Taking the infimum over all such couplings qX,qYq_{X},q_{Y} gives the remaining claim. ∎

We now show that if a metric soliton splits a factor of k\mathbb{R}^{k}, this can be used to extract a sequence of approximating points in smooth Ricci flows which are almost-selfsimilar and almost-split. This is analogous to the existence of almost-splitting maps in section 2.6 of [CC96] given Gromov-Hausdorff closeness to a metric product.

Proposition 3.2.

Suppose (𝒳,(μt)t(,0])(\mathcal{X},(\mu_{t})_{t\in(-\infty,0]}) is a future continuous metric soliton satisfying the following:

  • (a)(a)

    (𝒳,(μt)t(,0))(\mathcal{X},(\mu_{t})_{t\in(-\infty,0)}) is an 𝔽\mathbb{F}-limit of nn-dimensional closed Ricci flows (Mi,(gi,t)t(δi1,0],(νxi,0;t)t(δ1,0])(M_{i},(g_{i,t})_{t\in(-\delta_{i}^{-1},0]},(\nu_{x_{i},0;t})_{t\in(-\delta^{-1},0]}) as in Theorem 2.5, with 𝒩xi,0(1)Y\mathcal{N}_{x_{i},0}(1)\geq-Y,

  • (b)(b)

    (𝒳,(μt)t(,0))(𝒳×k,(μtμk)t(,0))(\mathcal{X},(\mu_{t})_{t\in(-\infty,0)})\cong(\mathcal{X}^{\prime}\times\mathbb{R}^{k},(\mu_{t}^{\prime}\otimes\mu_{\mathbb{R}^{k}})_{t\in(-\infty,0)}) as metric flow pairs for some metric soliton (𝒳,(μt)t(,0))(\mathcal{X}^{\prime},(\mu_{t}^{\prime})_{t\in(-\infty,0)}), and this identification restricts to an isometry of Ricci flow spacetimes ×k\mathcal{R}\cong\mathcal{R}^{\prime}\times\mathbb{R}^{k},

  • (c)(c)

    Writing dμt=(4πτ)n2efdgd\mu_{t}=(4\pi\tau)^{-\frac{n}{2}}e^{-f}dg on \mathcal{R} and dμt=(4πτ)n2efdgd\mu_{t}^{\prime}=(4\pi\tau)^{-\frac{n}{2}}e^{-f^{\prime}}dg^{\prime} on \mathcal{R}^{\prime}, we have Rc(g)+2f=12τgRc(g)+\nabla^{2}f=\frac{1}{2\tau}g on \mathcal{R}, Rc(g)+2f=12τgRc(g^{\prime})+\nabla^{2}f^{\prime}=\frac{1}{2\tau}g^{\prime} on \mathcal{R}^{\prime}, and f=f+14τ|x|2f=f^{\prime}+\frac{1}{4\tau}|x|^{2} on \mathcal{R}.

  • (d)(d)

    𝒩(μt)(τ)=W\mathcal{N}_{(\mu_{t})}(\tau)=W for all τ>0\tau>0, where W[Y,0]W\in[-Y,0].

Then for any ϵ>0\epsilon>0, (xi,0)(x_{i},0) is (ϵ,1)(\epsilon,1)-selfsimilar and (k,ϵ,1)(k,\epsilon,1)-split for sufficiently large i=i(ϵ)i=i(\epsilon)\in\mathbb{N}. If in addition, Rc(g)=0Rc(g)=0, then (xi,0)(x_{i},0) is also (ϵ,1)(\epsilon,1)-static for sufficiently large ii\in\mathbb{N}.

Proof.

By Nash entropy convergence (Theorem 15.45 of [Bam20b]), assumption (d)(d) and Proposition 7.1 of [Bam20b] imply that (xi,0)(x_{i},0) is (ϵ,1)(\epsilon,1)-selfsimilar for sufficiently large ii\in\mathbb{N}. If Rc(g)=0Rc(g)=0, then (xi,0)(x_{i},0) is (ϵ,1)(\epsilon,1)-static for large ii\in\mathbb{N} by Claim 22.7 of [Bam20b]. Suppose the remaining claim fails, so that there exists ϵ>0\epsilon>0 such that, after passing to a subsequence, (xi,0)(x_{i},0) is not (k,ϵ,1)(k,\epsilon,1)-split. Fix a correspondence \mathfrak{C} such that

(3.1) (Min,(gi,t)t(δi1,0],(νxi,0;t)t(δi1,0])i𝔽,(𝒳,(νx;t)t(,0])(M_{i}^{n},(g_{i,t})_{t\in(-\delta_{i}^{-1},0]},(\nu_{x_{i},0;t})_{t\in(\delta_{i}^{-1},0]})\xrightarrow[i\to\infty]{\mathbb{F},\mathfrak{C}}(\mathcal{X},(\nu_{x_{\infty};t})_{t\in(-\infty,0]})

on compact time intervals. By passing to a subsequence, we can assume the convergence is time-wise for all times in some subset I(,0)I^{\prime}\subseteq(-\infty,0), where |I|=0|\mathbb{R}\setminus I^{\prime}|=0.

Choose a sequence tj0t_{j}\nearrow 0, and recall that

Hn|t|Var(μt)=Var(μt)+Var(μtk)Var(μt),H_{n}|t|\geq\text{Var}(\mu_{t})=\text{Var}(\mu_{t}^{\prime})+\text{Var}(\mu_{t}^{\mathbb{R}^{k}})\geq\text{Var}(\mu_{t}^{\prime}),

so we can find zj𝒳tjz_{j}\in\mathcal{X}_{t_{j}}^{\prime} such that Var(δzj,μtj)<Hn|tj|\text{Var}(\delta_{z_{j}},\mu_{t_{j}}^{\prime})<H_{n}|t_{j}|. Then

dW1𝒳tj(δ(zj,0k),μtj)=\displaystyle d_{W_{1}}^{\mathcal{X}_{t_{j}}}(\delta_{(z_{j},0^{k})},\mu_{t_{j}})= dW1𝒳tj×k(δzjδ0k,μtjμtjk)dW1𝒳tj(δzj,μtj)+dW1k(δ0k,μtjk)2Hn|tj|.\displaystyle d_{W_{1}}^{\mathcal{X}_{t_{j}}^{\prime}\times\mathbb{R}^{k}}(\delta_{z_{j}}\otimes\delta_{0^{k}},\mu_{t_{j}}^{\prime}\otimes\mu_{t_{j}}^{\mathbb{R}^{k}})\leq d_{W_{1}}^{\mathcal{X}_{t_{j}}^{\prime}}(\delta_{z_{j}},\mu_{t_{j}}^{\prime})+d_{W_{1}}^{\mathbb{R}^{k}}(\delta_{0^{k}},\mu_{t_{j}}^{\mathbb{R}^{k}})\leq 2H_{n}|t_{j}|.

Fix τ>0\tau>0. Letting eαke^{\alpha}\in\mathbb{R}^{k} be the standard basis vectors, we can find zj,iα,zj,iMiz_{j,i}^{\alpha},z_{j,i}\in M_{i} such that (zj,iα,t0)i(zj,eα)(z_{j,i}^{\alpha},t_{0})\xrightarrow[i\to\infty]{\mathfrak{C}}(z_{j},e^{\alpha}) and (zj,i,tj)i(zj,0k)(z_{j,i},t_{j})\xrightarrow[i\to\infty]{\mathfrak{C}}(z_{j},0^{k}), so that there are subsets Ej,i[τ1,0]E_{j,i}\subseteq[-\tau^{-1},0] such that limi|Ej,i|=0\lim_{i\to\infty}|E_{j,i}|=0, [τ1,tj]Ej,iI′′,j[-\tau^{-1},t_{j}]\setminus E_{j,i}\subseteq I^{\prime\prime,j} and

limisupt[τ1,tj]Ej,idW1Zt((φti)ν(zj,iα,tj);ti,(φt)ν(zj,eα);ti)=0\lim_{i\to\infty}\sup_{t\in[-\tau^{-1},t_{j}]\setminus E_{j,i}}d_{W_{1}}^{Z_{t}}\left((\varphi_{t}^{i})_{\ast}\nu_{(z_{j,i}^{\alpha},t_{j});t}^{i},(\varphi_{t}^{\infty})_{\ast}\nu_{(z_{j},e^{\alpha});t}^{i}\right)=0
limisupt[τ1,tj]Ej,idW1Zt((φti)ν(zj,i,tj);ti,(φt)ν(zj,0k);ti)=0.\lim_{i\to\infty}\sup_{t\in[-\tau^{-1},t_{j}]\setminus E_{j,i}}d_{W_{1}}^{Z_{t}}\left((\varphi_{t}^{i})_{\ast}\nu_{(z_{j,i},t_{j});t}^{i},(\varphi_{t}^{\infty})_{\ast}\nu_{(z_{j},0^{k});t}^{i}\right)=0.

We can pass to further subsequences and use a diagonal argument to obtain subsets Ej[j,0]E_{j}\subseteq[-j,0] with [tj,0]Ej[t_{j},0]\subseteq E_{j} such that |Ej|2j|E_{j}|\leq 2^{-j}, [τ1,0]EjI′′,j[-\tau^{-1},0]\setminus E_{j}\subseteq I^{\prime\prime,j}, and zj,zj,αMjz_{j}^{\prime},z_{j}^{\prime,\alpha}\in M_{j} such that

supt[j,0]EjdW1Zt((φtj)ν(zj,tj);tj,(φt)ν(zj,0k);t)2j,\sup_{t\in[-j,0]\setminus E_{j}}d_{W_{1}}^{Z_{t}}\left((\varphi_{t}^{j})_{\ast}\nu_{(z_{j}^{\prime},t_{j});t}^{j},(\varphi_{t}^{\infty})_{\ast}\nu_{(z_{j},0^{k});t}\right)\leq 2^{-j},
supt[j,0]EjdW1Zt((φtj)ν(zj,α,tj);tj,(φt)ν(zj,eα);t)2j\sup_{t\in[-j,0]\setminus E_{j}}d_{W_{1}}^{Z_{t}}\left((\varphi_{t}^{j})_{\ast}\nu_{(z_{j}^{\prime,\alpha},t_{j});t}^{j},(\varphi_{t}^{\infty})_{\ast}\nu_{(z_{j},e^{\alpha});t}\right)\leq 2^{-j}

We now show (ν(zj,tj);tj)t[j,tj]i(μt)t(,0)(\nu_{(z_{j}^{\prime},t_{j});t}^{j})_{t\in[-j,t_{j}]}\xrightarrow[i\to\infty]{\mathfrak{C}}(\mu_{t})_{t\in(-\infty,0)} uniformly on compact time intervals. In fact

dW1Zt((φtj)ν(zj,tj);tj,(φt)μt)\displaystyle d_{W_{1}}^{Z_{t}}\left((\varphi_{t}^{j})_{\ast}\nu_{(z_{j}^{\prime},t_{j});t}^{j},(\varphi_{t}^{\infty})_{\ast}\mu_{t}\right)\leq dW1Zt((φtj)ν(zj,tj);tj,(φt)ν(zj,0k);t)+dW1𝒳t(ν(zj,0k);t,μt)\displaystyle d_{W_{1}}^{Z_{t}}\left((\varphi_{t}^{j})_{\ast}\nu_{(z_{j}^{\prime},t_{j});t}^{j},(\varphi_{t}^{\infty})_{\ast}\nu_{(z_{j},0^{k});t}\right)+d_{W_{1}}^{\mathcal{X}_{t}}(\nu_{(z_{j},0^{k});t},\mu_{t})
\displaystyle\leq 2j+2Hn|tj|\displaystyle 2^{-j}+2H_{n}|t_{j}|

for all t[j,0]Ejt\in[-j,0]\setminus E_{j}. Similarly, for any τ>0\tau>0, for sufficiently large jj\in\mathbb{N}, we have

dW1𝒳tj(δ(zj,eα),μtjνeα;tjk)dW1𝒳tj(δzj,μtj)+dW1𝒳tj(δeα,νeα;tjk)2Hn|tj|,d_{W_{1}}^{\mathcal{X}_{t_{j}}}(\delta_{(z_{j},e^{\alpha})},\mu_{t_{j}}^{\prime}\otimes\nu_{e^{\alpha};t_{j}}^{\mathbb{R}^{k}})\leq d_{W_{1}}^{\mathcal{X}_{t_{j}}}(\delta_{z_{j}},\mu_{t_{j}}^{\prime})+d_{W_{1}}^{\mathcal{X}_{t_{j}}}(\delta_{e^{\alpha}},\nu_{e^{\alpha};t_{j}}^{\mathbb{R}^{k}})\leq 2H_{n}|t_{j}|,

so by repeating the above reasoning with (zj,tj)(z_{j}^{\prime},t_{j}) replaced by (zj,α,tj)(z_{j}^{\prime,\alpha},t_{j}), we can also assume that

(ν(zj,α,tj);t)t[Tj,0]j(μtνeα;tk)t(,0)(\nu_{(z_{j}^{\prime,\alpha},t_{j});t})_{t\in[-T_{j},0]}\xrightarrow[j\to\infty]{\mathfrak{C}}(\mu_{t}^{\prime}\otimes\nu_{e^{\alpha};t}^{\mathbb{R}^{k}})_{t\in(-\infty,0)}

on compact time intervals. Because (νxj,0;t)t[Tj,0]j(μt)t(,0)(\nu_{x_{j},0;t})_{t\in[-T_{j},0]}\xrightarrow[j\to\infty]{\mathfrak{C}}(\mu_{t})_{t\in(-\infty,0)}, we can pass to a further subsequence to find subsets EjE_{j}^{\prime} with |Ej|2j|E_{j}^{\prime}|\leq 2^{-j} such that

dW1Zt((φt)μt,(φtj)ν(xj,0);tj)2jd_{W_{1}}^{Z_{t}}\left((\varphi_{t}^{\infty})_{\ast}\mu_{t},(\varphi_{t}^{j})_{\ast}\nu_{(x_{j},0);t}^{j}\right)\leq 2^{-j}

for all t[j,0]Ejt\in[-j,0]\setminus E_{j}^{\prime}. It follows that

(3.2) dW1gj,t(ν(zj,tj);tj,ν(xj,0);tj)dW1Zt((φtj)ν(zj,tj);tj,(φt)μt)+dW1Zt((φt)μt,(φtj)ν(xj,0);tj)2j+1d_{W_{1}}^{g_{j,t}}(\nu_{(z_{j}^{\prime},t_{j});t}^{j},\nu_{(x_{j},0);t}^{j})\leq d_{W_{1}}^{Z_{t}}\left((\varphi_{t}^{j})_{\ast}\nu_{(z_{j}^{\prime},t_{j});t}^{j},(\varphi_{t}^{\infty})_{\ast}\mu_{t}\right)+d_{W_{1}}^{Z_{t}}\left((\varphi_{t}^{\infty})_{\ast}\mu_{t},(\varphi_{t}^{j})_{\ast}\nu_{(x_{j},0);t}^{j}\right)\leq 2^{-j+1}

for all t[j,0](EjEj)t\in[-j,0]\setminus(E_{j}\cup E_{j}^{\prime}), where |EjEj|2j+1|E_{j}\cup E_{j}^{\prime}|\leq 2^{-j+1}.

Let ψj:UjVjMj×[j,0]\psi_{j}:U_{j}\to V_{j}\subseteq M_{j}\times[-j,0] be time-preserving diffeomorphisms from a precompact exhaustion (Uj)(U_{j}) of \mathcal{R} realizing smooth convergence as in Theorem 2.5. Then ψjK(zj,tj;,)(4πτ)n2ef\psi_{j}^{\ast}K(z_{j}^{\prime},t_{j};\cdot,\cdot)\to(4\pi\tau)^{-\frac{n}{2}}e^{-f} in Cloc()C_{loc}^{\infty}(\mathcal{R}) and f=f+14τ|x|2f=f^{\prime}+\frac{1}{4\tau}|x|^{2}, so we have the following convergence in Cloc()C_{loc}^{\infty}(\mathcal{R}):

ψjK(zj,tj;,)(4πτ)n2ef|x|24τ,\psi_{j}^{\ast}K(z_{j}^{\prime},t_{j};\cdot,\cdot)\to(4\pi\tau)^{-\frac{n}{2}}e^{-f^{\prime}-\frac{|x|^{2}}{4\tau}},
ψjK(zj,α,tj;,)(4πτ)n2ef|xeα|24τ.\psi_{j}^{\ast}K(z_{j}^{\prime,\alpha},t_{j};\cdot,\cdot)\to(4\pi\tau)^{-\frac{n}{2}}e^{-f^{\prime}-\frac{|x-e^{\alpha}|^{2}}{4\tau}}.

Writing K(zj,tj;,)=(4πτj)n2efjK(z_{j}^{\prime},t_{j};\cdot,\cdot)=(4\pi\tau_{j})^{-\frac{n}{2}}e^{-f_{j}} and K(zj,α,tj;,)=(4πτj)n2efjαK(z_{j}^{\prime,\alpha},t_{j};\cdot,\cdot)=(4\pi\tau_{j})^{-\frac{n}{2}}e^{-f_{j}^{\alpha}}, this means

ψjfjf+|x|24τ,ψjfjαf+|xeα|24τ\psi_{j}^{\ast}f_{j}\to f^{\prime}+\frac{|x|^{2}}{4\tau},\qquad\psi_{j}^{\ast}f_{j}^{\alpha}\to f^{\prime}+\frac{|x-e^{\alpha}|^{2}}{4\tau}

in Cloc()C_{loc}^{\infty}(\mathcal{R}) .

Claim 1: For any δ>0\delta^{\prime}>0 when j=j(δ)j=j(\delta^{\prime}) is sufficiently large, (zj,tj)(z_{j}^{\prime},t_{j}) and (zj,α,tj)(z_{j}^{\prime,\alpha},t_{j}) are (δ,1)(\delta^{\prime},1)-selfsimilar.

Because τf𝑑μt(nk)2=W\int_{\mathcal{R}_{-\tau}^{\prime}}f^{\prime}d\mu_{t}^{\prime}-\frac{(n-k)}{2}=W for all t(,0)t\in(-\infty,0), the proof of Corollary 15.47 in [Bam20b] gives the following for any τ(0,)\tau\in(0,\infty):

limj𝒩zj;tj(τ)=\displaystyle\lim_{j\to\infty}\mathcal{N}_{z_{j}^{\prime};t_{j}}(\tau)= τf𝑑μtn2=τ×k(f+|x|24τ)d(μtμtk)n2\displaystyle\int_{\mathcal{R}_{-\tau}}fd\mu_{t}-\frac{n}{2}=\int_{\mathcal{R}_{-\tau}^{\prime}\times\mathbb{R}^{k}}\left(f^{\prime}+\frac{|x|^{2}}{4\tau}\right)d(\mu_{t}^{\prime}\otimes\mu_{t}^{\mathbb{R}^{k}})-\frac{n}{2}
=\displaystyle= τfdμt(nk)2=:W\displaystyle\int_{\mathcal{R}_{-\tau}^{\prime}}f^{\prime}d\mu_{t}^{\prime}-\frac{(n-k)}{2}=:W

and similarly,

limj𝒩zj,α;tj(τ)=τ×k(f+|xeα|24τ)d(μtμtk)n2=W.\lim_{j\to\infty}\mathcal{N}_{z_{j}^{\prime,\alpha};t_{j}}(\tau)=\int_{\mathcal{R}_{-\tau}^{\prime}\times\mathbb{R}^{k}}\left(f^{\prime}+\frac{|x-e^{\alpha}|^{2}}{4\tau}\right)d(\mu_{t}^{\prime}\otimes\mu_{t}^{\mathbb{R}^{k}})-\frac{n}{2}=W.

The claim follows by choosing τ\tau sufficiently large and appealing to Proposition 7.1 of [Bam20b]. \square

Choose t[1,12]It\in[-1,-\frac{1}{2}]\cap I^{\prime}. Then

dW1gj,1(νzj,α,tj;1j,νzj,tj;1j)\displaystyle d_{W_{1}}^{g_{j,-1}}\left(\nu_{z_{j}^{\prime,\alpha},t_{j};-1}^{j},\nu_{z_{j}^{\prime},t_{j};-1}^{j}\right)\leq dW1Zt((φtj)νzj,α,tj;tj,(φt)(μtν(eα,0k);tk))\displaystyle d_{W_{1}}^{Z_{t}}\left((\varphi_{t}^{j})_{\ast}\nu_{z_{j}^{\prime,\alpha},t_{j};t}^{j},(\varphi_{t}^{\infty})_{\ast}(\mu_{t}^{\prime}\otimes\nu_{(e^{\alpha},0^{k});t}^{\mathbb{R}^{k}})\right)
+dW1Zt((φtj)νzj,tj;tj,(φt)(μtμtk))+dW1k(νeα,0;tk,μtk).\displaystyle+d_{W_{1}}^{Z_{t}}\left((\varphi_{t}^{j})_{\ast}\nu_{z_{j}^{\prime},t_{j};t}^{j},(\varphi_{t}^{\infty})_{\ast}(\mu_{t}^{\prime}\otimes\mu_{t}^{\mathbb{R}^{k}})\right)+d_{W_{1}}^{\mathbb{R}^{k}}\left(\nu_{e^{\alpha},0;t}^{\mathbb{R}^{k}},\mu_{t}^{\mathbb{R}^{k}}\right).

Because dW1k(νeα,0;tk,μtk)1d_{W_{1}}^{\mathbb{R}^{k}}\left(\nu_{e^{\alpha},0;t}^{\mathbb{R}^{k}},\mu_{t}^{\mathbb{R}^{k}}\right)\leq 1 and the 𝔽\mathbb{F}-convergence (3.1) is timewise at time tt, we have

dW1gj,1(νzj,α,tj;1j,νzj,tj;1j)2d_{W_{1}}^{g_{j,-1}}\left(\nu_{z_{j}^{\prime,\alpha},t_{j};-1}^{j},\nu_{z_{j}^{\prime},t_{j};-1}^{j}\right)\leq 2

for sufficiently large jj\in\mathbb{N}. Now fix δ>0\delta>0 and β>0\beta>0 small. Taking δδ¯(δ,n,Y)\delta^{\prime}\leq\overline{\delta}^{\prime}(\delta,n,Y) in Claim 1, we may therefore appeal to the proof of Proposition 10.8 in [Bam20b] (up to Claim 10.31) to conclude that, setting Wj:=𝒩zj,tj(1)W_{j}^{\prime}:=\mathcal{N}_{z_{j}^{\prime},t_{j}}(1), Wjα:=𝒩zj,α,tj(1)W_{j}^{\alpha}:=\mathcal{N}_{z_{j}^{\prime,\alpha},t_{j}}(1) , the functions

ujα:=2τj(fjαfj)2τj(WjαWj)u_{j}^{\alpha}:=2\tau_{j}(f_{j}^{\alpha}-f_{j}^{\prime})-2\tau_{j}(W_{j}^{\alpha}-W_{j}^{\prime})

satisfy the following properties for all j=j(δ)j=j(\delta)\in\mathbb{N} sufficiently large, assuming γγ¯\gamma\leq\overline{\gamma}:

(i)(i) δ1δMj(τj12|tujα|+|2ujα|2)eγfj𝑑νzj,tj;tj𝑑tδ\int_{-\delta^{-1}}^{-\delta}\int_{M_{j}}\left(\tau_{j}^{-\frac{1}{2}}|\partial_{t}u_{j}^{\alpha}|+|\nabla^{2}u_{j}^{\alpha}|^{2}\right)e^{\gamma f_{j}^{\prime}}d\nu_{z_{j}^{\prime},t_{j};t}^{j}dt\leq\delta,

(ii)(ii) δ1δMjτj1|ujα|4e2γfj𝑑νzj,tj;tj𝑑tC(Y,γ),\int_{-\delta^{-1}}^{-\delta}\int_{M_{j}}\tau_{j}^{-1}|\nabla u_{j}^{\alpha}|^{4}e^{2\gamma f_{j}^{\prime}}d\nu_{z_{j}^{\prime},t_{j};t}^{j}dt\leq C(Y,\gamma),

(iii)(iii) δ1δMj|ujα,ujβqαβj|𝑑νzj,tj;tj𝑑tδ\int_{-\delta^{-1}}^{-\delta}\int_{M_{j}}|\langle\nabla u_{j}^{\alpha},\nabla u_{j}^{\beta}\rangle-q_{\alpha\beta}^{j}|d\nu_{z_{j}^{\prime},t_{j};t}^{j}dt\leq\delta for some qαβjq_{\alpha\beta}^{j}\in\mathbb{R}.

Moreover, for any b>0b>0, we can combine (ii),(iii)(ii),(iii) to estimate

2ϵ112ϵMj|ujα,ujβqαβj|eγfj𝑑νt𝑑t\displaystyle\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M_{j}}|\langle\nabla u_{j}^{\alpha},\nabla u_{j}^{\beta}\rangle-q_{\alpha\beta}^{j}|e^{\gamma f_{j}^{\prime}}d\nu_{t}dt\leq 1b2ϵ112ϵMj|ujα,ujβqαβj|𝑑νt𝑑t\displaystyle\frac{1}{b}\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M_{j}}|\langle\nabla u_{j}^{\alpha},\nabla u_{j}^{\beta}\rangle-q_{\alpha\beta}^{j}|d\nu_{t}dt
+b2ϵ112ϵMj|ujα,ujβqαβj|e2γfj𝑑νzj,tj;tj\displaystyle+b\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M_{j}}|\langle\nabla u_{j}^{\alpha},\nabla u_{j}^{\beta}\rangle-q_{\alpha\beta}^{j}|e^{2\gamma f_{j}^{\prime}}d\nu_{z_{j}^{\prime},t_{j};t}^{j}
\displaystyle\leq b1δ+C(ϵ,Y,γ)b.\displaystyle b^{-1}\delta+C(\epsilon,Y,\gamma)b.

For any δ>0\delta^{\prime}>0, we can therefore choose bb¯(δ,Y,γ)b\leq\overline{b}(\delta^{\prime},Y,\gamma) and then δδ¯(δ,Y,γ)\delta\leq\overline{\delta}(\delta^{\prime},Y,\gamma) so that

(iii)(iii^{\prime}) 2ϵ112ϵMj|ujα,ujβqαβj|eγfj𝑑νt𝑑tδ\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M_{j}}|\langle\nabla u_{j}^{\alpha},\nabla u_{j}^{\beta}\rangle-q_{\alpha\beta}^{j}|e^{\gamma f_{j}^{\prime}}d\nu_{t}dt\leq\delta^{\prime} for some qαβjq_{\alpha\beta}^{j}\in\mathbb{R}
whenever j=j(ϵ,δ)j=j(\epsilon,\delta^{\prime})\in\mathbb{N} is sufficiently large.
Claim 2: For any δ′′>0\delta^{\prime\prime}>0, twe have |qαβjδαβ|δ′′|q_{\alpha\beta}^{j}-\delta_{\alpha\beta}|\leq\delta^{\prime\prime} for sufficiently large j=j(δ′′)j=j(\delta^{\prime\prime})\in\mathbb{N}.

Fix any compact set K[2,1]K\subseteq\mathcal{R}_{[-2,-1]}, and set Kt:=KtK_{t}:=K\cap\mathcal{R}_{t} or t[2,1]t\in[-2,-1]. For sufficiently large jj, we can estimate

21Mj|ujα,ujβδαβ|𝑑νzj,tj;tj𝑑t\displaystyle\int_{-2}^{-1}\int_{M_{j}}|\langle\nabla u_{j}^{\alpha},\nabla u_{j}^{\beta}\rangle-\delta_{\alpha\beta}|d\nu_{z_{j}^{\prime},t_{j};t}^{j}dt
=\displaystyle= 21ψj,t(Kt)|ujα,ujβδαβ|𝑑νzj,tj;tj𝑑t+21Mjψj,t(Kt)|ujα,ujβδαβ|𝑑νzj,tj;tj𝑑t\displaystyle\int_{-2}^{-1}\int_{\psi_{j,t}(K_{t})}|\langle\nabla u_{j}^{\alpha},\nabla u_{j}^{\beta}\rangle-\delta_{\alpha\beta}|d\nu_{z_{j}^{\prime},t_{j};t}^{j}dt+\int_{-2}^{-1}\int_{M_{j}\setminus\psi_{j,t}(K_{t})}|\langle\nabla u_{j}^{\alpha},\nabla u_{j}^{\beta}\rangle-\delta_{\alpha\beta}|d\nu_{z_{j}^{\prime},t_{j};t}^{j}dt
\displaystyle\leq 21Kt2τj|(fjαfj),(fjβfj)ψjδαβ|(K(zj,tj;,)ψj)d(ψj,tgj,t)𝑑t\displaystyle\int_{-2}^{-1}\int_{K_{t}}2\tau_{j}\left|\langle\nabla(f_{j}^{\alpha}-f_{j}),\nabla(f_{j}^{\beta}-f_{j})\rangle\circ\psi_{j}-\delta_{\alpha\beta}\right|\left(K(z_{j}^{\prime},t_{j};\cdot,\cdot)\circ\psi_{j}\right)d(\psi_{j,t}^{\ast}g_{j,t})dt
+(21Mjψj(K)|ujα,ujβδαβ|2𝑑νzj,tj;tj𝑑t)12(21νzj,tj;tj(Mψj(K))𝑑t)12\displaystyle+\left(\int_{-2}^{-1}\int_{M_{j}\setminus\psi_{j}(K)}|\langle\nabla u_{j}^{\alpha},\nabla u_{j}^{\beta}\rangle-\delta_{\alpha\beta}|^{2}d\nu_{z_{j}^{\prime},t_{j};t}^{j}dt\right)^{\frac{1}{2}}\left(\int_{-2}^{-1}\nu_{z_{j}^{\prime},t_{j};t}^{j}(M\setminus\psi_{j}(K))dt\right)^{\frac{1}{2}}
\displaystyle\leq supK|2τj(fjαfj),(fjβfj)ψjδαβ|\displaystyle\sup_{K}\left|2\tau_{j}\langle\nabla(f_{j}^{\alpha}-f_{j}),\nabla(f_{j}^{\beta}-f_{j})\rangle\circ\psi_{j}-\delta_{\alpha\beta}\right|
+C(Y,γ)(121νzj,tj;tj(ψj,t(Kt))𝑑t)12,\displaystyle+C(Y,\gamma)\left(1-\int_{-2}^{-1}\nu_{z_{j}^{\prime},t_{j};t}^{j}(\psi_{j,t}(K_{t}))dt\right)^{\frac{1}{2}},

where we used estimate (ii)(ii) to obtain the last inequality. Next, we observe that as jj\to\infty, 2τj(fjαfj),(fjβfj)ψj2\tau_{j}\langle\nabla(f_{j}^{\alpha}-f_{j}),\nabla(f_{j}^{\beta}-f_{j})\rangle\circ\psi_{j} converges in Cloc()C_{loc}^{\infty}(\mathcal{R}) to

2τ(f+|xeα|24τ)(f+|x|24τ),(f+|xeβ|24τ)(f+|x|24τ)\displaystyle 2\tau\left\langle\nabla\left(f^{\prime}+\frac{|x-e^{\alpha}|^{2}}{4\tau}\right)-\nabla\left(f^{\prime}+\frac{|x|^{2}}{4\tau}\right),\nabla\left(f^{\prime}+\frac{|x-e^{\beta}|^{2}}{4\tau}\right)-\nabla\left(f^{\prime}+\frac{|x|^{2}}{4\tau}\right)\right\rangle
=(xeα)x,(xeβ)x=δαβ\displaystyle\hskip 34.1433pt=\left\langle(x-e^{\alpha})-x,(x-e^{\beta})-x\right\rangle=\delta_{\alpha\beta}

Moreover, we have

21νzj,tj;tj(ψj,t(Kt))𝑑t=21Kt(K(zj,tj;,t)ψj,t)d(ψj,tgj,t)𝑑t21μt(K)𝑑t.\int_{-2}^{-1}\nu_{z_{j}^{\prime},t_{j};t}^{j}(\psi_{j,t}(K_{t}))dt=\int_{-2}^{-1}\int_{K_{t}}\left(K(z_{j}^{\prime},t_{j};\cdot,t)\circ\psi_{j,t}\right)d(\psi_{j,t}^{\ast}g_{j,t})dt\to\int_{-2}^{-1}\mu_{t}(K)dt.

Combining expressions, we get

lim supj|δαβqαβj|\displaystyle\limsup_{j\to\infty}|\delta_{\alpha\beta}-q_{\alpha\beta}^{j}|\leq lim supj21Mj|ujα,ujβδαβ|𝑑νzj,tj;tj𝑑t\displaystyle\limsup_{j\to\infty}\int_{-2}^{-1}\int_{M_{j}}|\langle\nabla u_{j}^{\alpha},\nabla u_{j}^{\beta}\rangle-\delta_{\alpha\beta}|d\nu_{z_{j}^{\prime},t_{j};t}^{j}dt
+lim supj21Mj|ujα,ujβqαβj|𝑑νzj,tj;tj𝑑t\displaystyle+\limsup_{j\to\infty}\int_{-2}^{-1}\int_{M_{j}}|\langle\nabla u_{j}^{\alpha},\nabla u_{j}^{\beta}\rangle-q_{\alpha\beta}^{j}|d\nu_{z_{j}^{\prime},t_{j};t}^{j}dt
\displaystyle\leq C(γ,n)(121μt(K)𝑑t)12+δ\displaystyle C(\gamma,n)\left(1-\int_{-2}^{-1}\mu_{t}(K)dt\right)^{\frac{1}{2}}+\delta^{\prime}

for any compact set K[2,1]K\subseteq\mathcal{R}_{[-2,-1]}. Choosing a compact exhaustion of [2,1]\mathcal{R}_{[-2,-1]} and δδ¯(δ′′)\delta^{\prime}\leq\overline{\delta}^{\prime}(\delta^{\prime\prime}) then gives the claim. \square

By the W1W_{1}-distance estimate (3.2), we can apply Proposition 8.1 of [Bam20b] with the following choice of parameters: α0=γ\alpha_{0}=\gamma, α1=0\alpha_{1}=0, t0=tjt_{0}=t_{j}, t1=0t_{1}=0, D=1D=1, s[ϵ1,ϵ]s\in[-\epsilon^{-1},-\epsilon], t=12θ(1)γϵt^{\ast}=\frac{1}{2}\theta(1)\gamma\epsilon, where θ\theta is defined in the aforementioned proposition; then for jj\in\mathbb{N} sufficiently large, we have

dνxj,0;sjC(Y,γ)eγfjdνzj,tj;sjd\nu_{x_{j},0;s}^{j}\leq C(Y,\gamma)e^{\gamma f_{j}^{\prime}}d\nu_{z_{j}^{\prime},t_{j};s}^{j}

for all s[ϵ1,ϵ]s\in[-\epsilon^{-1},-\epsilon]. Combining this with (i),(iii)(i),(iii^{\prime}), the (δ,1)(\delta,1)-selfsimilarity of (zi,ti)(z_{i}^{\prime},t_{i}), and taking δ\delta sufficiently small gives the desired contradiction. ∎

Next, we verify that an 𝔽\mathbb{F}-limit of metric solitons splitting k\mathbb{R}^{k} is also a metric soliton splitting k\mathbb{R}^{k}.

Lemma 3.3.

Suppose (𝒳i,(μti)t(,0))(\mathcal{X}_{i},(\mu_{t}^{i})_{t\in(-\infty,0)}) is a sequence of metric solitons satisfying the following:

  • (a)i(a)_{i}

    Each pair (𝒳i,(μti)t(,0))(\mathcal{X}_{i},(\mu_{t}^{i})_{t\in(-\infty,0)}) is an 𝔽\mathbb{F}-limit of nn-dimensional closed Ricci flows as in Theorem 2.5, with Nash entropy lower bound Y-Y,

  • (b)i(b)_{i}

    (𝒳i,(μti)t(,0))(𝒳i×k,(μt,iμk)t(,0))(\mathcal{X}_{i},(\mu_{t}^{i})_{t\in(-\infty,0)})\cong(\mathcal{X}_{i}^{\prime}\times\mathbb{R}^{k},(\mu_{t}^{\prime,i}\otimes\mu_{\mathbb{R}^{k}})_{t\in(-\infty,0)}) as metric flow pairs for some metric solitons (𝒳i,(μt,i)t(,0))(\mathcal{X}_{i}^{\prime},(\mu_{t}^{\prime,i})_{t\in(-\infty,0)}), and this identification restricts to an isometry of Ricci flow spacetimes ii×k\mathcal{R}_{i}\cong\mathcal{R}_{i}^{\prime}\times\mathbb{R}^{k},

  • (c)i(c)_{i}

    Writing dμti=(4πτ)n2efidgid\mu_{t}^{i}=(4\pi\tau)^{-\frac{n}{2}}e^{-f_{i}}dg_{i} on i\mathcal{R}_{i} and dμt,i=(4πτ)n2efidgid\mu_{t}^{\prime,i}=(4\pi\tau)^{-\frac{n}{2}}e^{-f_{i}^{\prime}}dg_{i}^{\prime} on i\mathcal{R}_{i}^{\prime}, we have Rc(gi)+2fi=12τgiRc(g_{i})+\nabla^{2}f_{i}=\frac{1}{2\tau}g_{i} on i\mathcal{R}_{i}, Rc(gi)+2fi=12τgiRc(g_{i}^{\prime})+\nabla^{2}f_{i}^{\prime}=\frac{1}{2\tau}g_{i}^{\prime} on i\mathcal{R}_{i}^{\prime}, and fi=fi+14τ|x|2f_{i}=f_{i}^{\prime}+\frac{1}{4\tau}|x|^{2} on i\mathcal{R}_{i}.

  • (d)i(d)_{i}

    𝒩(μi,t)(τ)=Wi\mathcal{N}_{(\mu_{i,t})}(\tau)=W_{i} for all τ>0\tau>0, where Wi[Y,0]W_{i}\in[-Y,0].

Assume that (𝒳i,(μti)t(,0))(\mathcal{X}_{i},(\mu_{t}^{i})_{t\in(-\infty,0)}) 𝔽\mathbb{F}-converge to another metric flow pair (𝒳,(μt)t(,0))(\mathcal{X}_{\infty},(\mu_{t}^{\infty})_{t\in(-\infty,0)}). Then (𝒳,(μt)t(,0))(\mathcal{X},(\mu_{t}^{\infty})_{t\in(-\infty,0)}) is a metric soliton satisfying (a)(d)(a)-(d) of Proposition 3.2.

If in addition Rc(gi)=0Rc(g_{i}^{\prime})=0, then we have (𝒳,(μt))(𝒳′′×k,(νx′′;t)t(,0))(\mathcal{X}_{\infty},(\mu_{t}^{\infty}))\cong(\mathcal{X}^{\prime\prime}\times\mathbb{R}^{k},(\nu_{x^{\prime\prime};t})_{t\in(-\infty,0)}) as metric flow pairs for some static metric cone 𝒳′′\mathcal{X}^{\prime\prime} with vertex x′′x^{\prime\prime}. Moreover, this identification restricts to an isometry of Ricci flow spacetimes ′′×k\mathcal{R}^{\infty}\cong\mathcal{R}^{\prime\prime}\times\mathbb{R}^{k}, and Rc(g)=0Rc(g^{\infty})=0, Rc(g′′)=0Rc(g^{\prime\prime})=0.

Proof.

By passing to subsequences and using a diagonal argument, we may assume that

(Min,(gi,t)t(ϵi1,0],(xi,0))(M_{i}^{n},(g_{i,t})_{t\in(-\epsilon_{i}^{-1},0]},(x_{i},0))

is a sequence of closed, pointed Ricci flows such that 𝒩xi,0(1)Y\mathcal{N}_{x_{i},0}(1)\geq-Y,

d𝔽((Mi,(gi,t)t(ϵi1,0],(νxi,0;t)t(ϵi1,0]),(𝒳i,(μti)t(ϵi1,0]))ϵi,d_{\mathbb{F}}\left((M_{i},(g_{i,t})_{t\in(-\epsilon_{i}^{-1},0]},(\nu_{x_{i},0;t})_{t\in(-\epsilon_{i}^{-1},0]}),(\mathcal{X}_{i},(\mu_{t}^{i})_{t\in(-\epsilon_{i}^{-1},0]})\right)\leq\epsilon_{i},

where limiϵi=0\lim_{i\to\infty}\epsilon_{i}=0. By Proposition 3.2, we may moreover assume that (xi,0)(x_{i},0) are (ϵi,1)(\epsilon_{i},1)-selfsimilar and strongly (k,ϵi,1)(k,\epsilon_{i},1)-split. In particular,

(Mi,(gi,t)t(Ti,0],(νxi,0;t)t(Ti,0])i𝔽(𝒳,(μt)t(,0]).(M_{i},(g_{i,t})_{t\in(-T_{i},0]},(\nu_{x_{i}^{\prime},0;t})_{t\in(-T_{i},0]})\xrightarrow[i\to\infty]{\mathbb{F}}(\mathcal{X}_{\infty},(\mu_{t}^{\infty})_{t\in(-\infty,0]}).

Then 𝒳\mathcal{X} satisfies (a)(a) by construction, (b)(b) by Theorem 15.50 in [Bam20b], and (d)(d) by Proposition 7.1 in [Bam20b] and the Nash entropy convergence Theorem 15.45 of [Bam20b]. Moreover, we have Rc(g)+2f=12τgRc(g)+\nabla^{2}f=\frac{1}{2\tau}g on \mathcal{R} by Theorem 15.69 of [Bam20b]. By (b)(b), we have

(4πτ)n2efdg=((4πτ)nk2efdg)((4πτ)k2e|x|24τ)=(4πτ)n2ef|x|24τdg(4\pi\tau)^{-\frac{n}{2}}e^{-f}dg=\left((4\pi\tau)^{-\frac{n-k}{2}}e^{-f^{\prime}}dg^{\prime}\right)\otimes\left((4\pi\tau)^{-\frac{k}{2}}e^{-\frac{|x|^{2}}{4\tau}}\right)=(4\pi\tau)^{-\frac{n}{2}}e^{-f^{\prime}-\frac{|x|^{2}}{4\tau}}dg

on \mathcal{R}, so that f=f+|x|24τf=f^{\prime}+\frac{|x|^{2}}{4\tau}. Thus Rc(g),g,2fRc(g),g,\nabla^{2}f all split with respect to the decomposition T=TTkT\mathcal{R}=T\mathcal{R}^{\prime}\oplus T\mathbb{R}^{k}, so restricting Rc(g)+2f=12τgRc(g)+\nabla^{2}f=\frac{1}{2\tau}g to TT\mathcal{R}^{\prime} gives Rc(g)+2f=12τgRc(g^{\prime})+\nabla^{2}f^{\prime}=\frac{1}{2\tau}g^{\prime} on \mathcal{R}^{\prime}.

Finally, suppose that Rc(gi)=0Rc(g_{i}^{\prime})=0 for all ii\in\mathbb{N}. By the proof of Claim 22.7 in [Bam20b], we can also assume (xi,0)(x_{i},0) are (ϵi,1)(\epsilon_{i},1)-static. Then Theorem 15.80 of [Bam20b] guarantees the remaining claims. ∎

We may now apply the previous results to prove the reverse qualitative inclusion of quantitative singular strata as that proved in [Bam20b]

Proposition 3.4.

For any ϵ>0\epsilon>0, there exists δ=δ(ϵ,Y,A)>0\delta=\delta(\epsilon,Y,A)>0 such that the following holds. Suppose (𝒳,(νx;t)t(T,0])(\mathcal{X},(\nu_{x_{\infty};t})_{t\in(-T,0]}) is a metric flow pair obtained as an 𝔽\mathbb{F}-limit of noncollapsed Ricci flows as in Theorem 2.5 (with 𝒩xi,0(1)Y\mathcal{N}_{x_{i},0}(1)\geq-Y). Assume y𝒳<0P(x,A,A2)y_{\infty}\in\mathcal{X}_{<0}\cap P^{\ast}(x_{\infty},A,-A^{2}) is (k,δ,r)(k,\delta,r)-symmetric and that (M,(gt)t[δ1,0],(y0,0))(M,(g_{t})_{t\in[-\delta^{-1},0]},(y_{0},0)) is a closed pointed Ricci flow such that |𝒩y0,0(1)𝒩y(1)|<δ|\mathcal{N}_{y_{0},0}(1)-\mathcal{N}_{y_{\infty}}(1)|<\delta and

d𝔽((M,(gt)t[δ1,0],(νy0,0;t)t[δ1,0]),(𝒳t0,r1,(νy;tt0,r1)t[δ1,0]))<δ,d_{\mathbb{F}}\left((M,(g_{t})_{t\in[-\delta^{-1},0]},(\nu_{y_{0},0;t})_{t\in[-\delta^{-1},0]}),(\mathcal{X}^{-t_{0},r^{-1}},(\nu_{y_{\infty};t}^{-t_{0},r^{-1}})_{t\in[-\delta^{-1},0]})\right)<\delta,

where t0:=𝔱(y)t_{0}:=\mathfrak{t}(y_{\infty}). Then one of the following holds:

(i)(i) (y0,t0)(y_{0},t_{0}) is (k,ϵ,1)(k,\epsilon,1)-split and (ϵ,1)(\epsilon,1)-selfsimilar,

(ii)(ii) (y0,t0)(y_{0},t_{0}) is (k2,ϵ,1)(k-2,\epsilon,1)-split, (ϵ,1)(\epsilon,1)-static, and (ϵ,1)(\epsilon,1)-selfsimilar.
In particular, yy_{\infty} is weakly (k,ϵ,1)(k,\epsilon,1)-symmetric, and 𝒮^r1,r2ϵ,k𝒮r1,r2δ(ϵ,Y,A),k\widehat{\mathcal{S}}_{r_{1},r_{2}}^{\epsilon,k}\subseteq\mathcal{S}_{r_{1},r_{2}}^{\delta(\epsilon,Y,A),k}. Moreover, we have

𝒮=ϵ(0,1)𝒮0,ϵϵ,k.\mathcal{S}=\cup_{\epsilon\in(0,1)}\mathcal{S}_{0,\epsilon}^{\epsilon,k}.
Proof.

The hypotheses ensure that 𝒩y(1)Y(Y,A)\mathcal{N}_{y_{\infty}}(1)\geq-Y^{\prime}(Y,A). By time translation and parabolic rescaling, we can assume r=1r=1 and 𝔱(y)=0\mathfrak{t}(y_{\infty})=0. Suppose by way of contradiction there is a sequence δi0\delta_{i}\searrow 0 along with metric flows 𝒳i\mathcal{X}^{i} each obtained as 𝔽\mathbb{F}-limits of closed noncollapsed nn-dimensional Ricci flows, (k,δi,1)(k,\delta_{i},1)-symmetric points y,i𝒳0iy_{\infty,i}\in\mathcal{X}_{0}^{i} with 𝒩y,i(1)Y\mathcal{N}_{y_{\infty,i}}(1)\geq-Y^{\prime}, closed pointed Ricci flows (Mi,(gi,t)t[δi1,0],(yi,0))(M_{i},(g_{i,t})_{t\in[-\delta_{i}^{-1},0]},(y_{i},0)) satifying |𝒩yi,0(1)𝒩y,i(1)|<δi|\mathcal{N}_{y_{i},0}(1)-\mathcal{N}_{y_{\infty,i}}(1)|<\delta_{i} and

d𝔽((Mi,(gi,t)t[δi1,0],(νyi,0;t)t[δi1,0]),(𝒳i,(νy,i;t)t[δi1,0]))<δi,d_{\mathbb{F}}\left((M_{i},(g_{i,t})_{t\in[-\delta_{i}^{-1},0]},(\nu_{y_{i},0;t})_{t\in[-\delta_{i}^{-1},0]}),(\mathcal{X}_{i},(\nu_{y_{\infty,i};t})_{t\in[-\delta_{i}^{-1},0]})\right)<\delta_{i},

such that (i),(ii)(i),(ii) both fail for (yi,0)(y_{i},0). Because y,iy_{\infty,i} are (k,δi,1)(k,\delta_{i},1)-symmetric, we can find metric flow pairs (𝒳i,(μt,i)t[δi1,0])(\mathcal{X}_{i}^{\prime},(\mu_{t}^{\prime,i})_{t\in[-\delta_{i}^{-1},0]}) satisfying properties (b)(d)(b)-(d) in Definition 2.10, along with |𝒩μt,i(1)𝒩y,i(1)|<δi|\mathcal{N}_{\mu_{t}^{\prime,i}}(1)-\mathcal{N}_{y_{\infty,i}}(1)|<\delta_{i} and

d𝔽((𝒳i,(νy,i;t)t[δi1,0]),(𝒳i,(μt,i)t[δi1,0]))<δi.d_{\mathbb{F}}\left((\mathcal{X}_{i},(\nu_{y_{\infty},i};t)_{t\in[-\delta_{i}^{-1},0]}),(\mathcal{X}_{i}^{\prime},(\mu_{t}^{\prime,i})_{t\in[-\delta_{i}^{-1},0]})\right)<\delta_{i}.

Because 𝒩yi,0(1)Yδi2Y\mathcal{N}_{y_{i},0}(1)\geq-Y^{\prime}-\delta_{i}\geq-2Y^{\prime}, Theorem 2.5 lets us pass to a subsequence to obtain a future-continuous metric flow pair (𝒳,(μt)t(,0])(\mathcal{X}^{\prime},(\mu_{t}^{\prime})_{t\in(-\infty,0]}) such that

(Mi,(gi,t)t[δi1,0],(νyi,0;t)t[δi1,0])i𝔽(𝒳,(μt)t(,0])(M_{i},(g_{i,t})_{t\in[-\delta_{i}^{-1},0]},(\nu_{y_{i},0;t})_{t\in[-\delta_{i}^{-1},0]})\xrightarrow[i\to\infty]{\mathbb{F}}(\mathcal{X}^{\prime},(\mu_{t}^{\prime})_{t\in(-\infty,0]})

uniformly on compact time intervals. By construction we also have

(𝒳i,(μt,i)t[δi1,0])i𝔽(𝒳,(μt)t(,0])(\mathcal{X}_{i}^{\prime},(\mu_{t}^{\prime,i})_{t\in[-\delta_{i}^{-1},0]})\xrightarrow[i\to\infty]{\mathbb{F}}(\mathcal{X}^{\prime},(\mu_{t}^{\prime})_{t\in(-\infty,0]})

on compact time intervals, so by Lemma 3.3, 𝒳\mathcal{X}^{\prime} also satisfies properties (b)(d)(b)-(d) of Definition 2.10. By Proposition 3.2, we conclude that (yi,0)(y_{i},0) satisfies one of (i)(i) or (ii)(ii) when ii\in\mathbb{N} is sufficiently large, a contradiction. ∎

4. Strong Almost-GRS Potentials

In this section, we construct parabolic regularizations of potential functions associated to conjugate heat kernels based at almost-selfsimilar points of a Ricci flow. These functions still satisfy the almost-soliton identities, but also satisfy additional estimates which will be useful in Section 5.

Definition 4.1.

A strong (ϵ,r)(\epsilon,r)-soliton potential based at (x0,t0)M×I(x_{0},t_{0})\in M\times I is a function hC(M×[t0ϵ1r2,t0ϵr2])h\in C^{\infty}(M\times[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}]) such that if W:=𝒩x0,t0(r2)W:=\mathcal{N}_{x_{0},t_{0}}(r^{2}), then

(i)(4τ(hW))=2n,\displaystyle\hskip 14.22636pt(i)\hskip 8.53581pt\square\left(4\tau(h-W)\right)=-2n,
(ii)r2t0ϵ1r2t0ϵr2M|τ(R+|h|2)(hW)|𝑑νt𝑑tϵ,,\displaystyle\hskip 14.22636pt(ii)\hskip 8.53581ptr^{-2}\int_{t_{0}-\epsilon^{-1}r^{2}}^{t_{0}-\epsilon r^{2}}\int_{M}\left|\tau(R+|\nabla h|^{2})-(h-W)\right|d\nu_{t}dt\leq\epsilon,,
(iii)t0ϵ1r2t0ϵr2Mτ|Rc+2h12τg|2𝑑νt𝑑tϵ,\displaystyle\hskip 14.22636pt(iii)\hskip 8.53581pt\int_{t_{0}-\epsilon^{-1}r^{2}}^{t_{0}-\epsilon r^{2}}\int_{M}\tau\left|Rc+\nabla^{2}h-\frac{1}{2\tau}g\right|^{2}d\nu_{t}dt\leq\epsilon,
(iv)supt[t0ϵ1r2,t0ϵr2]M|τ(R+2Δh|h|2)+hnW|𝑑νtϵ,\displaystyle\hskip 14.22636pt(iv)\hskip 8.53581pt\sup_{t\in[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}]}\int_{M}\left|\tau(R+2\Delta h-|\nabla h|^{2})+h-n-W\right|d\nu_{t}\leq\epsilon,
(v)M(hn2)𝑑νt=W for all t[t0ϵ1r2,t0ϵr2].\displaystyle\hskip 14.22636pt(v)\hskip 8.53581pt\int_{M}\left(h-\frac{n}{2}\right)d\nu_{t}=W\text{ for all }t\in[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}].

The following proposition is an analogue of Bamler’s construction (Theorem 12.1 of [Bam20b]) of strong almost-splitting maps which approximate weak almost-splitting maps.

Proposition 4.2.

For any ϵ>0\epsilon>0, Y<Y<\infty, the following holds whenever δδ¯(ϵ,Y)\delta\leq\overline{\delta}(\epsilon,Y). Suppose (Mn,(gt)t[T,0])(M^{n},(g_{t})_{t\in[-T,0]}) is a closed Ricci flow with 𝒩x0,t0(r2)Y\mathcal{N}_{x_{0},t_{0}}(r^{2})\geq-Y. Assume (x0,t0)M×I(x_{0},t_{0})\in M\times I is (δ,r)(\delta,r)-selfsimilar, and set q:=4τ(fW)q:=4\tau(f-W), where dν=dνx0,t0=(4πτ)n2efdgd\nu=d\nu_{x_{0},t_{0}}=(4\pi\tau)^{-\frac{n}{2}}e^{-f}dg and W:=𝒩x0,t0(r2)W:=\mathcal{N}_{x_{0},t_{0}}(r^{2}). Then there exists a function qC(M×[t0ϵ1r2,t0ϵr2])q^{\prime}\in C^{\infty}(M\times[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}]) such that f:=14τq+Wf^{\prime}:=\frac{1}{4\tau}q^{\prime}+W is a strong (ϵ,r)(\epsilon,r)-soliton potential based at (x0,t0)(x_{0},t_{0}), and

(4.1) t0ϵ1r2t0ϵr2M|(ff)|2𝑑νt𝑑t+supt[t0ϵ1r2,t0ϵr2]M(ff)2𝑑νtϵ.\int_{t_{0}-\epsilon^{-1}r^{2}}^{t_{0}-\epsilon r^{2}}\int_{M}|\nabla(f-f^{\prime})|^{2}d\nu_{t}dt+\sup_{t\in[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}]}\int_{M}(f-f^{\prime})^{2}d\nu_{t}\leq\epsilon.
Proof.

Without loss of generality, we can assume r=1r=1 and t0=0t_{0}=0. Bamler’s on-diagonal upper bounds for the heat kernel (Theorem 7.1 in [Bam20a]) imply fΛ(Y)f\geq-\Lambda(Y) on M×[ϵ1,ϵ]M\times[-\epsilon^{-1},-\epsilon] if δδ¯\delta\leq\overline{\delta}. Fix Z(Λ,)Z\in(\Lambda,\infty) to be determined, and let χC()\chi\in C^{\infty}(\mathbb{R}) be such that χ(s)=s\chi(s)=s for s(,12Z]s\in(-\infty,\frac{1}{2}Z], |χ|2|\chi^{\prime}|\leq 2, |χ′′|10Z1|\chi^{\prime\prime}|\leq 10Z^{-1}, and χ(s)=Z\chi(s)=Z for all s[Z,]s\in[Z,\infty]. Set q~:=χq\widetilde{q}:=\chi\circ q, which satisfies q~4τ(Λ+Y)\widetilde{q}\geq-4\tau(\Lambda+Y).

Step 1: (Bound the truncation errors) We first apply Proposition 6.5 of [Bam20b] to obtain

νt({qZ})eZ8τ{qZ}eq8τ𝑑νtC(Y)eZ8τ{qZ}e12f𝑑νtC(Y)eZ8τ\nu_{t}(\{q\geq Z\})\leq e^{-\frac{Z}{8\tau}}\int_{\{q\geq Z\}}e^{\frac{q}{8\tau}}d\nu_{t}\leq C(Y)e^{-\frac{Z}{8\tau}}\int_{\{q\geq Z\}}e^{\frac{1}{2}f}d\nu_{t}\leq C(Y)e^{-\frac{Z}{8\tau}}

for all t[12δ1,δ]t\in[-\frac{1}{2}\delta^{-1},-\delta]. Thus, for any p[1,)p\in[1,\infty),

{qZ}qp𝑑νt=\displaystyle\int_{\{q\geq Z\}}q^{p}d\nu_{t}= {qZ}(p0q(x)rp1𝑑r)𝑑νt(x)=p0{qZ}1{rq(x)}rp1𝑑νt(x)𝑑r\displaystyle\int_{\{q\geq Z\}}\left(p\int_{0}^{q(x)}r^{p-1}dr\right)d\nu_{t}(x)=p\int_{0}^{\infty}\int_{\{q\geq Z\}}1_{\{r\leq q(x)\}}r^{p-1}d\nu_{t}(x)dr
=\displaystyle= pZrp1νt({qr})𝑑r+pZpνt({qZ})\displaystyle p\int_{Z}^{\infty}r^{p-1}\nu_{t}(\{q\geq r\})dr+pZ^{p}\nu_{t}(\{q\geq Z\})
\displaystyle\leq C(Y,p)Zrp1er8τ𝑑r+C(Y,p)ZpeZ8τ\displaystyle C(Y,p)\int_{Z}^{\infty}r^{p-1}e^{-\frac{r}{8\tau}}dr+C(Y,p)Z^{p}e^{-\frac{Z}{8\tau}}

In particular, we have

{qZ}qp𝑑νtC(Y,p,ϵ)eϵZ10.\int_{\{q\geq Z\}}q^{p}d\nu_{t}\leq C(Y,p,\epsilon)e^{-\frac{\epsilon Z}{10}}.

for all t[10ϵ1,110ϵ]t\in[-10\epsilon^{-1},-\frac{1}{10}\epsilon]. Using (2.1), we can estimate

10ϵ1110ϵM|(qq~)|2𝑑νt𝑑t\displaystyle\int_{-10\epsilon^{-1}}^{-\frac{1}{10}\epsilon}\int_{M}|\nabla(q-\widetilde{q})|^{2}d\nu_{t}dt\leq 210ϵ1110ϵ{qZ/2}|q|2𝑑νt𝑑t\displaystyle 2\int_{-10\epsilon^{-1}}^{-\frac{1}{10}\epsilon}\int_{\{q\geq Z/2\}}|\nabla q|^{2}d\nu_{t}dt
(10ϵ1110ϵM|q|4𝑑νt𝑑t)12(10ϵ1110ϵνt({qZ/2})𝑑t)12\displaystyle\leq\left(\int_{-10\epsilon^{-1}}^{-\frac{1}{10}\epsilon}\int_{M}|\nabla q|^{4}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-10\epsilon^{-1}}^{-\frac{1}{10}\epsilon}\nu_{t}(\{q\geq Z/2\})dt\right)^{\frac{1}{2}}
C(Y,ϵ)eϵZ20.\displaystyle\leq C(Y,\epsilon)e^{-\frac{\epsilon Z}{20}}.

Next, we estimate

M(qq~)2𝑑νt{qZ/2}2q2𝑑νtC(Y,ϵ)eϵZ10\int_{M}(q-\widetilde{q})^{2}d\nu_{t}\leq\int_{\{q\geq Z/2\}}2q^{2}d\nu_{t}\leq C(Y,\epsilon)e^{-\frac{\epsilon Z}{10}}

for any t[10ϵ1,110ϵ]t\in[-10\epsilon^{-1},-\frac{1}{10}\epsilon], and apply (2.1) to obtain

10ϵ1110ϵM|Δ(qq~)|𝑑νt𝑑t\displaystyle\int_{-10\epsilon^{-1}}^{-\frac{1}{10}\epsilon}\int_{M}|\Delta(q-\widetilde{q})|d\nu_{t}dt 10ϵ1110ϵM|1(χq)||Δq|𝑑νt𝑑t+10ϵ1110ϵM|χ′′q||q|2𝑑νt𝑑t\displaystyle\leq\int_{-10\epsilon^{-1}}^{-\frac{1}{10}\epsilon}\int_{M}\left|1-(\chi^{\prime}\circ q)\right|\cdot|\Delta q|d\nu_{t}dt+\int_{-10\epsilon^{-1}}^{-\frac{1}{10}\epsilon}\int_{M}|\chi^{\prime\prime}\circ q|\cdot|\nabla q|^{2}d\nu_{t}dt
(10ϵ1110ϵM(4|Δq|2+|q|4)𝑑νt𝑑t)12(10ϵ1110ϵνt({qZ/2})𝑑t)12\displaystyle\leq\left(\int_{-10\epsilon^{-1}}^{-\frac{1}{10}\epsilon}\int_{M}\left(4|\Delta q|^{2}+|\nabla q|^{4}\right)d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-10\epsilon^{-1}}^{-\frac{1}{10}\epsilon}\nu_{t}(\{q\geq Z/2\})dt\right)^{\frac{1}{2}}
C(Y,ϵ)eϵZ20.\displaystyle\leq C(Y,\epsilon)e^{-\frac{\epsilon Z}{20}}.

Step 2: (Estimate the error by parabolic regularization) By Step 1, we can choose t[2ϵ1,2ϵ1+1]t^{\ast}\in[-2\epsilon^{-1},-2\epsilon^{-1}+1] such that

M|(qq~)|2𝑑νt+M|qq~|2𝑑νt+M|Δ(qq~)|𝑑νtC(Y,ϵ)eϵZ20.\int_{M}|\nabla(q-\widetilde{q})|^{2}d\nu_{t^{\ast}}+\int_{M}|q-\widetilde{q}|^{2}d\nu_{t^{\ast}}+\int_{M}|\Delta(q-\widetilde{q})|d\nu_{t^{\ast}}\leq C(Y,\epsilon)e^{-\frac{\epsilon Z}{20}}.

Let qC(M×[t,0])q^{\prime}\in C^{\infty}(M\times[t^{\ast},0]) solve q=2n\square q^{\prime}=-2n, with q(,t):=q~q^{\prime}(\cdot,t^{\ast}):=\widetilde{q}. Because Λq~Z-\Lambda\leq\widetilde{q}\leq Z, the maximum principle gives 8ϵ1(Λ+Y)4nϵ1qZ-8\epsilon^{-1}(\Lambda+Y)-4n\epsilon^{-1}\leq q^{\prime}\leq Z on M×[t,0]M\times[t^{\ast},0], so if Z8ϵ1(Λ+Y)Z\geq 8\epsilon^{-1}(\Lambda+Y), then (2.2) and the almost-selfsimilar inequalities (Definition 2.8(i)(i)) imply

ddtM(q~q)2𝑑νt=\displaystyle\frac{d}{dt}\int_{M}(\widetilde{q}-q^{\prime})^{2}d\nu_{t}= 2M(q~q)(q~+2n)𝑑νt2M|(q~q)|2𝑑νt\displaystyle 2\int_{M}(\widetilde{q}-q^{\prime})(\square\widetilde{q}+2n)d\nu_{t}-2\int_{M}|\nabla(\widetilde{q}-q^{\prime})|^{2}d\nu_{t}
\displaystyle\leq C(ϵ,Y)ZM(|q+2n|+|1χq|+(χ′′q)|q|2)𝑑νt\displaystyle C(\epsilon,Y)Z\int_{M}\left(|\square q+2n|+|1-\chi^{\prime}\circ q|+(\chi^{\prime\prime}\circ q)|\nabla q|^{2}\right)d\nu_{t}
2M|(q~q)|2𝑑νt\displaystyle-2\int_{M}|\nabla(\widetilde{q}-q^{\prime})|^{2}d\nu_{t}
\displaystyle\leq C(ϵ,Y)Zνt({qZ/2})+(νt({qZ/2}))12(M|q|4𝑑νt)12\displaystyle C(\epsilon,Y)Z\nu_{t}(\{q\geq Z/2\})+\left(\nu_{t}(\{q\geq Z/2\})\right)^{\frac{1}{2}}\left(\int_{M}|\nabla q|^{4}d\nu_{t}\right)^{\frac{1}{2}}
+Ψ(δ|ϵ,Y,Z)2M|(q~q)|2𝑑νt\displaystyle+\Psi(\delta|\epsilon,Y,Z)-2\int_{M}|\nabla(\widetilde{q}-q^{\prime})|^{2}d\nu_{t}

so we may integrate in time, using Step 1, Hölder’s inequality, and (2.1) to obtain

supt[t,110ϵ]M(q~q)2𝑑νt+t110ϵM|(q~q)|2𝑑νt𝑑t\displaystyle\sup_{t\in[t^{\ast},-\frac{1}{10}\epsilon]}\int_{M}(\widetilde{q}-q^{\prime})^{2}d\nu_{t}+\int_{t^{\ast}}^{-\frac{1}{10}\epsilon}\int_{M}|\nabla(\widetilde{q}-q^{\prime})|^{2}d\nu_{t}dt\leq Ψ(δ|Y,ϵ,Z)+C(Y,ϵ)ZeϵZ20.\displaystyle\Psi(\delta|Y,\epsilon,Z)+C(Y,\epsilon)Ze^{-\frac{\epsilon Z}{20}}.

Combining this with the estimates of Step 1 gives

supt[t,110ϵ]M(qq)2𝑑νt+t110ϵM|(qq)|2𝑑νt𝑑tΨ(δ|Y,ϵ,Z)+C(Y,ϵ)ZeϵZ20,\sup_{t\in[t^{\ast},-\frac{1}{10}\epsilon]}\int_{M}(q-q^{\prime})^{2}d\nu_{t}+\int_{t^{\ast}}^{-\frac{1}{10}\epsilon}\int_{M}|\nabla(q-q^{\prime})|^{2}d\nu_{t}dt\leq\Psi(\delta|Y,\epsilon,Z)+C(Y,\epsilon)Ze^{-\frac{\epsilon Z}{20}},

hence (4.1) holds if we choose ZZ¯(ϵ,Y)Z\geq\underline{Z}(\epsilon,Y) and then δδ¯(ϵ,Y,Z)\delta\leq\overline{\delta}(\epsilon,Y,Z).

Step 3: (Show ff^{\prime} is an almost-soliton potential function) We now consider a quantity analogous to Perelman’s differential Harnack quantity:

w:=τ(R+12τΔq116τ2|q|2)+14τqn+W,w^{\prime}:=\tau\left(R+\frac{1}{2\tau}\Delta q^{\prime}-\frac{1}{16\tau^{2}}|\nabla q^{\prime}|^{2}\right)+\frac{1}{4\tau}q^{\prime}-n+W,

so that

(16τ(wW))\displaystyle\square\left(16\tau(w^{\prime}-W)\right) =(16τ2+8τΔq|q|2+4q16τn)\displaystyle=\square\left(16\tau^{2}+8\tau\Delta q^{\prime}-|\nabla q^{\prime}|^{2}+4q^{\prime}-16\tau n\right)
(4.2) =32τ2|Rc+14τ2q12τg|2.\displaystyle=32\tau^{2}\left|Rc+\frac{1}{4\tau}\nabla^{2}q^{\prime}-\frac{1}{2\tau}g\right|^{2}.

We can write

wW\displaystyle w^{\prime}-W =(ww)+(wW)\displaystyle=(w^{\prime}-w)+(w-W)
=12Δ(qq)116τ(|q|2|q|2)+14τ(qq)+(wW).\displaystyle=\frac{1}{2}\Delta(q^{\prime}-q)-\frac{1}{16\tau}\left(|\nabla q^{\prime}|^{2}-|\nabla q|^{2}\right)+\frac{1}{4\tau}(q^{\prime}-q)+(w-W).

By Step 1, we can estimate

M||q|2|q|2|𝑑νt\displaystyle\int_{M}\left||\nabla q^{\prime}|^{2}-|\nabla q|^{2}\right|d\nu_{t^{\ast}} C{qZ/2}|q|2𝑑νtC(Y,ϵ)eϵZ20,\displaystyle\leq C\int_{\{q\geq Z/2\}}|\nabla q|^{2}d\nu_{t^{\ast}}\leq C(Y,\epsilon)e^{-\frac{\epsilon Z}{20}},

and so (also using the almost-selfsimilar identities)

(4.3) M16τ|wW|𝑑νtC(Y,ϵ)eϵZ20+C(ϵ)δ.\int_{M}16\tau|w^{\prime}-W|d\nu_{t^{\ast}}\leq C(Y,\epsilon)e^{-\frac{\epsilon Z}{20}}+C(\epsilon)\delta.

Using (2.1) and Step 2, we have

(4.4) tϵ2M||q|2|q|2|𝑑νt𝑑t\displaystyle\int_{t^{\ast}}^{-\frac{\epsilon}{2}}\int_{M}\left||\nabla q^{\prime}|^{2}-|\nabla q|^{2}\right|d\nu_{t}dt\leq tϵ2M|(qq),(qq)+2q|𝑑νt𝑑t\displaystyle\int_{t^{\ast}}^{-\frac{\epsilon}{2}}\int_{M}\left|\langle\nabla(q^{\prime}-q),\nabla(q^{\prime}-q)+2\nabla q\rangle\right|d\nu_{t}dt
\displaystyle\leq tϵ2M|(qq)|2𝑑νt𝑑t+C(Y,ϵ)(tϵ2M|(qq)|2𝑑νt𝑑t)12\displaystyle\int_{t^{\ast}}^{-\frac{\epsilon}{2}}\int_{M}\left|\nabla(q^{\prime}-q)\right|^{2}d\nu_{t}dt+C(Y,\epsilon)\left(\int_{t^{\ast}}^{-\frac{\epsilon}{2}}\int_{M}|\nabla(q^{\prime}-q)|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}
\displaystyle\leq Ψ(δ|ϵ,Y,Z)+C(Y,ϵ)ZeϵZ40,\displaystyle\Psi(\delta|\epsilon,Y,Z)+C(Y,\epsilon)Ze^{-\frac{\epsilon Z}{40}},
(4.5) tϵ2|MΔ(qq)𝑑νt|𝑑t\displaystyle\int_{t^{\ast}}^{-\frac{\epsilon}{2}}\left|\int_{M}\Delta(q^{\prime}-q)d\nu_{t}\right|dt =tϵ2|Mf,(qq)𝑑νt|𝑑t\displaystyle=\int_{t^{\ast}}^{-\frac{\epsilon}{2}}\left|\int_{M}\langle\nabla f,\nabla(q^{\prime}-q)\rangle d\nu_{t}\right|dt
(tϵ2M|f|2𝑑νt𝑑t)12(tϵ2M|(qq)|2𝑑νt𝑑t)12\displaystyle\leq\left(\int_{t^{\ast}}^{-\frac{\epsilon}{2}}\int_{M}|\nabla f|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{t^{\ast}}^{-\frac{\epsilon}{2}}\int_{M}|\nabla(q^{\prime}-q)|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}
Ψ(δ|ϵ,Y,Z)+C(Y,ϵ)ZeϵZ40.\displaystyle\leq\Psi(\delta|\epsilon,Y,Z)+C(Y,\epsilon)Ze^{-\frac{\epsilon Z}{40}}.

Choose a cutoff function ζC([t,12ϵ])\zeta\in C^{\infty}([t^{\ast},-\frac{1}{2}\epsilon]) satisfying ζ(t)|[t,ϵ]1\zeta(t)|[t^{\ast},-\epsilon]\equiv 1, |ζ|4ϵ1|\zeta^{\prime}|\leq 4\epsilon^{-1}, and ζ(ϵ2)=0\zeta(-\frac{\epsilon}{2})=0. Then (4) implies

(4.6) tϵ2ζddtMτ(wW)𝑑νt𝑑t=2tϵ2Mζτ2|Rc+14τ2q12τg|2𝑑νt𝑑t.\int_{t^{\ast}}^{-\frac{\epsilon}{2}}\zeta\frac{d}{dt}\int_{M}\tau(w^{\prime}-W)d\nu_{t}dt=2\int_{t^{\ast}}^{-\frac{\epsilon}{2}}\int_{M}\zeta\tau^{2}\left|Rc+\frac{1}{4\tau}\nabla^{2}q^{\prime}-\frac{1}{2\tau}g\right|^{2}d\nu_{t}dt.

On the other hand, integration by parts gives

(4.7) tϵ2ζddtMτ(wW)𝑑νt𝑑t=tϵ2ζMτ(wW)𝑑νt𝑑tMτ(wW)𝑑νt.\int_{t^{\ast}}^{-\frac{\epsilon}{2}}\zeta\frac{d}{dt}\int_{M}\tau(w^{\prime}-W)d\nu_{t}dt=-\int_{t^{\ast}}^{-\frac{\epsilon}{2}}\zeta^{\prime}\int_{M}\tau(w^{\prime}-W)d\nu_{t}dt-\int_{M}\tau(w^{\prime}-W)d\nu_{t^{\ast}}.

By combining (4.6),(4.7) with estimates (4.3),(4.4),(4.5), and Steps 1,2, we obtain

(4.8) tϵMτ2|Rc+14τ2q12τg|2𝑑νt𝑑tΨ(δ|ϵ,Y,Z)+C(Y,ϵ)ZeϵZ40.\int_{t^{\ast}}^{-\epsilon}\int_{M}\tau^{2}\left|Rc+\frac{1}{4\tau}\nabla^{2}q^{\prime}-\frac{1}{2\tau}g\right|^{2}d\nu_{t}dt\leq\Psi(\delta|\epsilon,Y,Z)+C(Y,\epsilon)Ze^{-\frac{\epsilon Z}{40}}.

Thus, property (iii)(iii) of strong almost-soliton potential functions holds if we choose ZZ¯(ϵ,Y)Z\geq\underline{Z}(\epsilon,Y) and δδ¯(ϵ,Y,Z)\delta\leq\overline{\delta}(\epsilon,Y,Z). In the sense of distributions,

|τ(wW)|2τ2|Rc+14τ2q12τg|2,\square|\tau(w^{\prime}-W)|\leq 2\tau^{2}\left|Rc+\frac{1}{4\tau}\nabla^{2}q^{\prime}-\frac{1}{2\tau}g\right|^{2},

so we can use (4.8) to estimate

M|τ(wW)|𝑑νtM|τ(wW)|𝑑νt\displaystyle\int_{M}|\tau(w^{\prime}-W)|d\nu_{t}-\int_{M}|\tau(w^{\prime}-W)|d\nu_{t^{\ast}} =ttM|τ(wW)|𝑑νs𝑑s\displaystyle=\int_{t^{\ast}}^{t}\int_{M}\square|\tau(w^{\prime}-W)|d\nu_{s}ds
Ψ(δ|Y,ϵ,Z)+C(Y,ϵ)ZeϵZ40\displaystyle\leq\Psi(\delta|Y,\epsilon,Z)+C(Y,\epsilon)Ze^{-\frac{\epsilon Z}{40}}

for all t[t,ϵ]t\in[t^{\ast},-\epsilon]. Combining this with (4.3) gives property (iv)(iv) of strong almost-soliton functions if we choose ZZ¯(ϵ,Y)Z\geq\underline{Z}(\epsilon,Y) and δδ¯(ϵ,Y,Z)\delta\leq\overline{\delta}(\epsilon,Y,Z), while (ii)(ii) follows by combining (iii),(iv)(iii),(iv). To verify (v)(v), we note that

ddt(τM(fn2)𝑑νtτW)=14M(q+2n)𝑑νt=0,\frac{d}{dt}\left(\tau\int_{M}\left(f^{\prime}-\frac{n}{2}\right)d\nu_{t}-\tau W\right)=\frac{1}{4}\int_{M}(\square q+2n)d\nu_{t}=0,

and moreover

|M(fn2)𝑑ν1W|M|qq|𝑑ν1Ψ(δ|Y,ϵ)+C(Y,ϵ)ZeϵZ40,\left|\int_{M}\left(f^{\prime}-\frac{n}{2}\right)d\nu_{-1}-W\right|\leq\int_{M}|q^{\prime}-q|d\nu_{-1}\leq\Psi(\delta|Y,\epsilon)+C(Y,\epsilon)Ze^{-\frac{\epsilon Z}{40}},

so we can add a small constant to qq to obtain ff^{\prime} satisfying (v)(v), without affecting properties (i)(iv)(i)-(iv) or (4.1) ∎

In the case where (x0,t0)(x_{0},t_{0}) is also almost-static, the scalar and Ricci curvature terms are small, and 4τ(hW)4\tau(h-W) is a regularization of Bamler’s almost-radial function (see Proposition 13.1 of [Bam20b] when k=0k=0).

Definition 4.3.

A strong (ϵ,r)(\epsilon,r)-radial function based at (x0,t0)M×I(x_{0},t_{0})\in M\times I is a function hC(M×[t0ϵ1r2,t0ϵr2])h\in C^{\infty}(M\times[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}]) such that if W:=𝒩x0,t0(r2)W:=\mathcal{N}_{x_{0},t_{0}}(r^{2}), then

(i)q=2n,\displaystyle\hskip 14.22636pt(i)\hskip 8.53581pt\square q=-2n,
(ii)r4t0ϵ1r2t0ϵr2M||q|24q|𝑑νt𝑑tϵ,\displaystyle\hskip 14.22636pt(ii)\hskip 8.53581ptr^{-4}\int_{t_{0}-\epsilon^{-1}r^{2}}^{t_{0}-\epsilon r^{2}}\int_{M}\left||\nabla q|^{2}-4q\right|d\nu_{t}dt\leq\epsilon,
(iii)r2t0ϵ1r2t0ϵr2M|2q2g|2𝑑νt𝑑tϵ,\displaystyle\hskip 14.22636pt(iii)\hskip 8.53581ptr^{-2}\int_{t_{0}-\epsilon^{-1}r^{2}}^{t_{0}-\epsilon r^{2}}\int_{M}|\nabla^{2}q-2g|^{2}d\nu_{t}dt\leq\epsilon,
(iv)Mq𝑑νt=2nτ for all t[t0ϵ1r2,t0ϵr2].\displaystyle\hskip 14.22636pt(iv)\hskip 8.53581pt\int_{M}qd\nu_{t}=2n\tau\text{ for all }t\in[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}].

Given these definitions, we can rephrase Proposition 4.2, and give a criterion for the existence of strong (ϵ,r)(\epsilon,r)-radial functions. Moreover, we will establish slightly improved estimates, which will be useful for the proof of Theorem 1.3.

Proposition 4.4.

For any ϵ>0\epsilon>0, Y<Y<\infty and p[1,)p\in[1,\infty), the following holds if δδ¯(ϵ,Y,p)\delta\leq\overline{\delta}(\epsilon,Y,p) and αα¯(ϵ,Y)\alpha\leq\overline{\alpha}(\epsilon,Y). Suppose (Mn,(gt)tI,(x0,t0))(M^{n},(g_{t})_{t\in I},(x_{0},t_{0})) is a closed, pointed Ricci flow satisfying 𝒩x0,t0(r2)Y\mathcal{N}_{x_{0},t_{0}}(r^{2})\geq-Y. Assume (x0,t0)(x_{0},t_{0}) is (δ,r)(\delta,r)-selfsimilar and hh is a strong (δ,r)(\delta,r)-soliton potential. Then

(4.9) supt[t0ϵ1r2,t0ϵr2]M|h|p𝑑νtC(Y,ϵ,p),\sup_{t\in[t_{0}-\epsilon^{-1}r^{2},t_{0}-\epsilon r^{2}]}\int_{M}|\nabla h|^{p}d\nu_{t}\leq C(Y,\epsilon,p),
(4.10) t0ϵ1r2t0ϵr2M(τ|Rc+2h12τg|2+r2|τ(R+|h|2)(hW)|)eαf𝑑νt𝑑tϵ.\int_{t_{0}-\epsilon^{-1}r^{2}}^{t_{0}-\epsilon r^{2}}\int_{M}\left(\tau\left|Rc+\nabla^{2}h-\frac{1}{2\tau}g\right|^{2}+r^{-2}\left|\tau(R+|\nabla h|^{2})-(h-W)\right|\right)e^{\alpha f}d\nu_{t}dt\leq\epsilon.

If in addition (x0,t0)(x_{0},t_{0}) is (δ,r)(\delta,r)-static, then q:=4τ(hW)q:=4\tau(h-W) is a strong (ϵ,r)(\epsilon,r)-radial function satisfying

t0ϵ1r2t0ϵr2M(r2|2q2g|2+r4||q|24q|)eαf𝑑νt𝑑tϵ.\int_{t_{0}-\epsilon^{-1}r^{2}}^{t_{0}-\epsilon r^{2}}\int_{M}\left(r^{-2}|\nabla^{2}q-2g|^{2}+r^{-4}\left||\nabla q|^{2}-4q\right|\right)e^{\alpha f}d\nu_{t}dt\leq\epsilon.
Proof.

By time translation and parabolic rescaling, we can assume r=1r=1 and t0=0t_{0}=0. Fixing T[ϵ1,δ1]T\in[\epsilon^{-1},\delta^{-1}], we use properties (ii),(v)(ii),(v) of strong almost-soliton potentials to get

TϵM|q|2𝑑νt𝑑t\displaystyle\int_{-T}^{-\epsilon}\int_{M}|\nabla q|^{2}d\nu_{t}dt TϵM16τ(τ(R+|h|2)(hW))𝑑νt𝑑t+16T3δ1+TϵM4q𝑑νt𝑑t\displaystyle\leq\int_{-T}^{-\epsilon}\int_{M}16\tau\left(\tau(R+|\nabla h|^{2})-(h-W)\right)d\nu_{t}dt+16T^{3}\delta^{-1}+\int_{-T}^{-\epsilon}\int_{M}4qd\nu_{t}dt
16Tδ+16T3δ+8nT210nT2\displaystyle\leq 16T\delta+16T^{3}\delta+8nT^{2}\leq 10nT^{2}

assuming ϵϵ¯\epsilon\leq\overline{\epsilon}. If we choose T=2pϵ1T=2p\epsilon^{-1}, then we can therefore find t^[2pϵ11,2pϵ1]\widehat{t}\in[2p\epsilon^{-1}-1,2p\epsilon^{-1}] such that

M|q|2𝑑νt^10nT2,\int_{M}|\nabla q|^{2}d\nu_{\widehat{t}}\leq 10nT^{2},

so the hypercontractivity of the heat kernel (Theorem 12.1 in [Bam20a]) gives

supt[2ϵ,ϵ2]M|q|p𝑑νtCT2=C(p,ϵ).\sup_{t\in[-2\epsilon,-\frac{\epsilon}{2}]}\int_{M}|\nabla q|^{p}d\nu_{t}\leq CT^{2}=C(p,\epsilon).

By Cauchy’s inequality, (4.10) will follow from

ϵ1ϵM(τ|Rc+2h12τg|2+|τ(R+|h|2)(hW)|)eαf𝑑νt𝑑tC(Y,ϵ)\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left(\tau\left|Rc+\nabla^{2}h-\frac{1}{2\tau}g\right|^{2}+|\tau(R+|\nabla h|^{2})-(h-W)|\right)e^{\alpha f}d\nu_{t}dt\leq C(Y,\epsilon)

if αα¯(ϵ,Y)\alpha\leq\overline{\alpha}(\epsilon,Y). Fix a cutoff function ζC([2ϵ1,12ϵ])\zeta\in C^{\infty}([-2\epsilon^{-1},-\frac{1}{2}\epsilon]) such that ζ(2ϵ1)=ζ(12ϵ)=0\zeta(-2\epsilon^{-1})=\zeta(-\frac{1}{2}\epsilon)=0, ζ|[ϵ1,ϵ]1\zeta|[-\epsilon^{-1},-\epsilon]\equiv 1, and |ζ|4ϵ1|\zeta^{\prime}|\leq 4\epsilon^{-1}. We compute (recalling the definition of ww^{\prime} from the proof of Proposition 4.2)

ddtMτ(wW)eαf\displaystyle\frac{d}{dt}\int_{M}\tau(w^{\prime}-W)e^{\alpha f} =M(τ(wW))eαf𝑑νtτM(wW)((4πτ)n2e(1α)f)𝑑gt\displaystyle=\int_{M}\square\left(\tau(w^{\prime}-W)\right)e^{\alpha f}d\nu_{t}-\tau\int_{M}(w^{\prime}-W)\square^{\ast}\left((4\pi\tau)^{-\frac{n}{2}}e^{-(1-\alpha)f}\right)dg_{t}
=2τ2M|Rc+2h12τg|2eαf𝑑νt\displaystyle=2\tau^{2}\int_{M}\left|Rc+\nabla^{2}h-\frac{1}{2\tau}g\right|^{2}e^{\alpha f}d\nu_{t}
+ατ2M(wW)(R+(1α)|f|2n2τ)eαf𝑑νt.\displaystyle\hskip 17.07164pt+\alpha\tau^{2}\int_{M}(w^{\prime}-W)\left(R+(1-\alpha)|\nabla f|^{2}-\frac{n}{2\tau}\right)e^{\alpha f}d\nu_{t}.

Multiplying both sides by ζ\zeta and integrating, then rearranging gives (assuming αα¯(Y,ϵ)\alpha\leq\overline{\alpha}(Y,\epsilon))

ϵ1ϵMτ2|Rc+2h12τg|2eαf𝑑νt\displaystyle\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\tau^{2}\left|Rc+\nabla^{2}h-\frac{1}{2\tau}g\right|^{2}e^{\alpha f}d\nu_{t}\leq C(Y,ϵ)(2ϵ112ϵM|wW|2𝑑νt𝑑t)12\displaystyle C(Y,\epsilon)\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|w^{\prime}-W|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}
+C(Y,ϵ)(2ϵ112ϵM|wW|2𝑑νt𝑑t)12(2ϵ112ϵM(R2+|f|4+1)e2αf𝑑νt𝑑t)12.\displaystyle+C(Y,\epsilon)\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|w^{\prime}-W|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}(R^{2}+|\nabla f|^{4}+1)e^{2\alpha f}d\nu_{t}dt\right)^{\frac{1}{2}}.

The L2L^{2} Poincare inequality and property (iv)(iv) of Definition 4.3 give

2ϵ112ϵM|wW|2𝑑νt𝑑t\displaystyle\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|w^{\prime}-W|^{2}d\nu_{t}dt C(Y,ϵ)2ϵ112ϵM(R2+|h|4+2|2h|2+h2+1)𝑑νt𝑑t\displaystyle\leq C(Y,\epsilon)\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}\left(R^{2}+|\nabla h|^{4}+2|\nabla^{2}h|^{2}+h^{2}+1\right)d\nu_{t}dt
C(Y,ϵ)2ϵ112ϵM(|Rc|2+|h|4+1)𝑑νt𝑑t.\displaystyle\leq C(Y,\epsilon)\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}\left(|Rc|^{2}+|\nabla h|^{4}+1\right)d\nu_{t}dt.

The L4L^{4} estimate for |h||\nabla h| is a consequence of (4.9), while the Ricci curvature is bounded using (2.1).

Now suppose (x0,0)(x_{0},0) is also (δ,1)(\delta,1)-static. Then qq clearly satisfies properties (i),(iv)(i),(iv), while (iii),(iv)(iii),(iv) follow from combining properties (iii),(ii)(iii),(ii), respectively, of strong almost-soliton potentials with the almost-static inequalities. The remaining inequality follows from the improved estimate for strong almost-soliton potentials and the estimate

2ϵ112ϵM|Rc|2eαf𝑑νt+supt[2ϵ1,12ϵ]MReαf𝑑νtΨ(δ|Y,ϵ),\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|Rc|^{2}e^{\alpha f}d\nu_{t}+\sup_{t\in[-2\epsilon^{-1},-\frac{1}{2}\epsilon]}\int_{M}Re^{\alpha f}d\nu_{t}\leq\Psi(\delta|Y,\epsilon),

which itself follows from Cauchy’s inequality, the almost-static inequalities, and (2.1),(2.2). ∎

Remark 4.5.

It is also possible to construct regularized versions of Bamler’s almost radial functions (c.f. Section 13 of [Bam20b]) when k>0k>0 near points which are almost-selfsimilar, almost-static, and almost-split, but we will not need this.

5. Improved Splitting for Noncollapsed Kähler-Ricci Flows

Near a point which is almost-selfsimilar and almost-split, we obtain estimates on the Ricci curvature in the direction of the almost-splitting.

Lemma 5.1.

Suppose (Mn,(gt)tI,(x,0))(M^{n},(g_{t})_{t\in I},(x,0)) is a closed, pointed Ricci flow, 𝒩x,0(1)Y\mathcal{N}_{x,0}(1)\geq-Y, yC(M×[δ1,δ])y\in C^{\infty}(M\times[-\delta^{-1},-\delta]) is a strong (1,δ,1)(1,\delta,1)-splitting map, and (x,0)(x,0) is (δ,1)(\delta,1)-selfsimilar. Then

ϵ1ϵM|Rc(y)|𝑑νt𝑑tΨ(δ|ϵ,Y).\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|Rc(\nabla y)|d\nu_{t}dt\leq\Psi(\delta|\epsilon,Y).
Proof.

Suppose by way of contradiction there exist ϵ>0\epsilon>0, Y<Y<\infty, a sequence δi0\delta_{i}\searrow 0 and closed Ricci flows (Mi,(gi,t)t(δi1,0])(M_{i},(g_{i,t})_{t\in(-\delta_{i}^{-1},0]}) along with (δi,1)(\delta_{i},1)-selfsimilar points (xi,0)(x_{i},0) and strong (1,δi,1)(1,\delta_{i},1)-splitting maps yiC(Mi×[δi1,δi])y_{i}\in C^{\infty}(M_{i}\times[-\delta_{i}^{-1},-\delta_{i}]) based at (xi,0)(x_{i},0) such that

lim infiϵ1ϵM|Rcgi(yi)|𝑑νti𝑑t>0,\liminf_{i\to\infty}\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|Rc_{g_{i}}(\nabla y_{i})|d\nu_{t}^{i}dt>0,

where νi\nu^{i} is the conjugate heat kernel of (Mi,(gi,t)t[δi1,0])(M_{i},(g_{i,t})_{t\in[-\delta_{i}^{-1},0]}) based at (xi,0)(x_{i},0). By passing to a subsequence, we can assume 𝔽\mathbb{F}-convergence

(Mi,(gi,t)t[δi1,0],(νti)t[δi1,0])i𝔽(𝒳,(μt)t(,0])(M_{i},(g_{i,t})_{t\in[-\delta_{i}^{-1},0]},(\nu_{t}^{i})_{t\in[-\delta_{i}^{-1},0]})\xrightarrow[i\to\infty]{\mathbb{F}}(\mathcal{X},(\mu_{t})_{t\in(-\infty,0]})

on compact time intervals, where 𝒳\mathcal{X} is a future-continuous metric soliton; moreover, dμt=(4πτ)n2efdgtd\mu_{t}=(4\pi\tau)^{-\frac{n}{2}}e^{-f}dg_{t}, on \mathcal{R}, where fC()f\in C^{\infty}(\mathcal{R}) satisfies Rc+2f=12τgRc+\nabla^{2}f=\frac{1}{2\tau}g on \mathcal{R}. Let (Ui)(U_{i}) be a precompact exhaustion of \mathcal{R}, and let ψi:UiMi\psi_{i}:U_{i}\to M_{i} be time-preserving diffeomorphisms such that ψigig\psi_{i}^{\ast}g_{i}\to g and ψiKi(xi,0;,)(4πτ)n2ef\psi_{i}^{\ast}K^{i}(x_{i},0;\cdot,\cdot)\to(4\pi\tau)^{-\frac{n}{2}}e^{-f} in Cloc()C_{loc}^{\infty}(\mathcal{R}), where KiK^{i} is the conjugate heat kernel of (Mi,(gi,t)t(δi1,0])(M_{i},(g_{i,t})_{t\in(-\delta_{i}^{-1},0]}). By Theorem 15.50 of [Bam3], we have a splitting of Ricci flow spacetimes ×\mathcal{R}\cong\mathcal{R}^{\prime}\times\mathbb{R} and of metric flows 𝒳𝒳×\mathcal{X}\cong\mathcal{X}^{\prime}\times\mathbb{R}, and ψiyiy\psi_{i}^{\ast}y_{i}\to y_{\infty} in Cloc()C_{loc}^{\infty}(\mathcal{R}), where y:𝒳y_{\infty}:\mathcal{X}\to\mathbb{R} is the projection onto the \mathbb{R}-factor. In particular, we have Rcg(y)=0Rc_{g_{\infty}}(\nabla y_{\infty})=0, hence Rcψigi((ψiyi))0Rc_{\psi_{i}^{\ast}g_{i}}(\nabla(\psi_{i}^{\ast}y_{i}))\to 0 in Cloc()C_{loc}^{\infty}(\mathcal{R}) as ii\to\infty. Let K[ϵ1,ϵ]K\subseteq\mathcal{R}_{[-\epsilon^{-1},-\epsilon]} be an arbitrary compact subset, and set Kt:=𝒳tKK_{t}:=\mathcal{X}_{t}\cap K, so that

lim supiϵ1ϵM|Rcgi(yi)|𝑑νti𝑑t\displaystyle\limsup_{i\to\infty}\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|Rc_{g_{i}}(\nabla y_{i})|d\nu_{t}^{i}dt
\displaystyle\leq lim supi(ϵ1ϵψi,t(Kt)|Rcgi(yi)|𝑑νti𝑑t+ϵ1ϵMiψi,t(Kt)|Rcgi(yi)|𝑑νti𝑑t)\displaystyle\limsup_{i\to\infty}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{\psi_{i,t}(K_{t})}|Rc_{g_{i}}(\nabla y_{i})|d\nu_{t}^{i}dt+\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M_{i}\setminus\psi_{i,t}(K_{t})}|Rc_{g_{i}}(\nabla y_{i})|d\nu_{t}^{i}dt\right)
\displaystyle\leq lim supi(ϵ1ϵKt|Rcψigi((ψiyi))|ψiKi(xi,0;,)d(ψigi,t)𝑑t)\displaystyle\limsup_{i\to\infty}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{K_{t}}|Rc_{\psi_{i}^{\ast}g_{i}}(\nabla(\psi_{i}^{\ast}y_{i}))|\psi_{i}^{\ast}K^{i}(x_{i},0;\cdot,\cdot)d(\psi_{i}^{\ast}g_{i,t})dt\right)
+lim supi(ϵ1ϵMiψi,t(Kt)|Rcgi(yi)|32𝑑νti𝑑t)23(ϵ1ϵνti(Miψi,t(Kt))𝑑t)13\displaystyle+\limsup_{i\to\infty}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M_{i}\setminus\psi_{i,t}(K_{t})}|Rc_{g_{i}}(\nabla y_{i})|^{\frac{3}{2}}d\nu_{t}^{i}dt\right)^{\frac{2}{3}}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\nu_{t}^{i}\left(M_{i}\setminus\psi_{i,t}(K_{t})\right)dt\right)^{\frac{1}{3}}
\displaystyle\leq lim supi(ϵ1ϵMi|Rcgi(yi)|32𝑑νti𝑑t)23(ϵ1ϵ(1νti(ψi,t(Kt)))𝑑t)13\displaystyle\limsup_{i\to\infty}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M_{i}}|Rc_{g_{i}}(\nabla y_{i})|^{\frac{3}{2}}d\nu_{t}^{i}dt\right)^{\frac{2}{3}}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\left(1-\nu_{t}^{i}(\psi_{i,t}(K_{t}))\right)dt\right)^{\frac{1}{3}}
\displaystyle\leq (ϵ1ϵ(1μt(Kt))𝑑t)12lim supi(ϵ1ϵMi|Rcgi(yi)|32𝑑νti𝑑t)23.\displaystyle\left(\int_{-\epsilon^{-1}}^{-\epsilon}\left(1-\mu_{t}(K_{t})\right)dt\right)^{\frac{1}{2}}\limsup_{i\to\infty}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M_{i}}|Rc_{g_{i}}(\nabla y_{i})|^{\frac{3}{2}}d\nu_{t}^{i}dt\right)^{\frac{2}{3}}.

We then estimate

ϵ1ϵMi|Rcgi(yi)|32𝑑νti𝑑t(ϵ1ϵM|Rcgi|2𝑑νti𝑑t)34(ϵ1ϵM|yi|4𝑑νti𝑑t)14C(Y,ϵ)\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M_{i}}|Rc_{g_{i}}(\nabla y_{i})|^{\frac{3}{2}}d\nu_{t}^{i}dt\leq\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|Rc_{g_{i}}|^{2}d\nu_{t}^{i}dt\right)^{\frac{3}{4}}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla y_{i}|^{4}d\nu_{t}^{i}dt\right)^{\frac{1}{4}}\leq C(Y,\epsilon)

for sufficiently large ii\in\mathbb{N}. Since K[ϵ1,ϵ]K\subseteq\mathcal{R}_{[-\epsilon^{-1},-\epsilon]} was arbitrary, we obtain

lim supiϵ1ϵM|Rcgi(yi)|𝑑νti𝑑t=0,\limsup_{i\to\infty}\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|Rc_{g_{i}}(\nabla y_{i})|d\nu_{t}^{i}dt=0,

a contradiction. ∎

Remark 5.2.

This proof is easily adapted to show LpL^{p} bounds on Rc(y)Rc(\nabla y) for any p[1,2)p\in[1,2), but fails for p=2p=2. This creates additional technical difficulties in the proof of Proposition 5.3.

Next, we use the estimates for strong almost-soliton potential functions and strong almost-splitting maps to construct new almost splitting maps on a Kähler-Ricci flows. We observe that many of these estimates would fail without the use of the parabolic approximation hh.

Proposition 5.3.

Suppose hC(M×[δ1,δ])h\in C^{\infty}(M\times[-\delta^{-1},-\delta]) is a strong (δ,1)(\delta,1)-soliton potential function at (x0,0)(x_{0},0) satisfying

δ1δM|(hf)|2𝑑νt𝑑tδ,\int_{-\delta^{-1}}^{-\delta}\int_{M}|\nabla(h-f)|^{2}d\nu_{t}dt\leq\delta,

and assume yC(M×[δ1,δ])y\in C^{\infty}(M\times[-\delta^{-1},-\delta]) is a strong (1,δ,1)(1,\delta,1)-splitting map, where (x0,0)(x_{0},0) is (δ,1)(\delta,1)-selfsimilar. If δδ¯(Y,ϵ,p)\delta\leq\overline{\delta}(Y,\epsilon,p), then the following hold, where q:=4τ(hW)q:=4\tau(h-W) and z:=12q,Jyz:=\frac{1}{2}\langle\nabla q,J\nabla y\rangle:

(i)ϵ1ϵM|2q|2(|y|2p+|q|2p)𝑑νt𝑑tC(Y,ϵ,p) for each p,\displaystyle\hskip 14.22636pt(i)\hskip 8.53581pt\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}q|^{2}\left(|\nabla y|^{2p}+|\nabla q|^{2p}\right)d\nu_{t}dt\leq C(Y,\epsilon,p)\text{ for each }p\in\mathbb{N},
(ii)ϵ1ϵM||z|21|𝑑νt𝑑tΨ(δ|Y,ϵ),\displaystyle\hskip 14.22636pt(ii)\hskip 8.53581pt\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left||\nabla z|^{2}-1\right|d\nu_{t}dt\leq\Psi(\delta|Y,\epsilon),
(iii)ϵ1ϵM|z,y|𝑑νt𝑑tΨ(δ|Y,ϵ),\displaystyle\hskip 14.22636pt(iii)\hskip 8.53581pt\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left|\langle\nabla z,\nabla y\rangle\right|d\nu_{t}dt\leq\Psi(\delta|Y,\epsilon),
(iv)ϵ1ϵM|z|𝑑νt𝑑tΨ(δ|Y,ϵ),\displaystyle\hskip 14.22636pt(iv)\hskip 8.53581pt\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\square z|d\nu_{t}dt\leq\Psi(\delta|Y,\epsilon),
(v)ϵ1ϵM|zJy|2𝑑νt𝑑tΨ(δ|Y,ϵ).\displaystyle\hskip 14.22636pt(v)\hskip 8.53581pt\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla z-J\nabla y|^{2}d\nu_{t}dt\leq\Psi(\delta|Y,\epsilon).

In particular, for δδ¯(Y,ϵ)\delta\leq\overline{\delta}(Y,\epsilon), (y,z)(y,z) is a weak (2,ϵ,1)(2,\epsilon,1)-splitting map.

Proof.

(i)(i) Observe that

|q|2(p+1)(p+1)|q|2p|q|2=2(p+1)|q|2p|2q|2.\square|\nabla q|^{2(p+1)}\leq(p+1)|\nabla q|^{2p}\square|\nabla q|^{2}=-2(p+1)|\nabla q|^{2p}|\nabla^{2}q|^{2}.

Upon integration, (4.9) lets us estimate

2(p+1)ϵ1ϵM|2q|2|q|2p𝑑νt𝑑tM|q|2(p+1)𝑑νt|t=2ϵ1t=12ϵC(Y,ϵ,p)2(p+1)\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}q|^{2}|\nabla q|^{2p}d\nu_{t}dt\leq-\left.\int_{M}|\nabla q|^{2(p+1)}d\nu_{t}\right|_{t=-2\epsilon^{-1}}^{t=-\frac{1}{2}\epsilon}\leq C(Y,\epsilon,p)

assuming δδ¯(Y,ϵ,p)\delta\leq\overline{\delta}(Y,\epsilon,p). Next, we estimate

(|y|2(p+1)|q|2)\displaystyle\square\left(|\nabla y|^{2(p+1)}|\nabla q|^{2}\right) 2(p+1)|2y|2|y|2p|q|22|y|2(p+1)|2q|22|y|2(p+1),|q|2\displaystyle\leq-2(p+1)|\nabla^{2}y|^{2}|\nabla y|^{2p}|\nabla q|^{2}-2|\nabla y|^{2(p+1)}|\nabla^{2}q|^{2}-2\langle\nabla|\nabla y|^{2(p+1)},\nabla|\nabla q|^{2}\rangle
2|y|2(p+1)|2q|2+8(p+1)|2y||y|2p+1|2q||q|\displaystyle\leq-2|\nabla y|^{2(p+1)}|\nabla^{2}q|^{2}+8(p+1)|\nabla^{2}y|\cdot|\nabla y|^{2p+1}|\nabla^{2}q|\cdot|\nabla q|

Integration on M×[2ϵ1,12ϵ]M\times[-2\epsilon^{-1},-\frac{1}{2}\epsilon] then gives

2ϵ112ϵM|2q|2|y|2p𝑑νt𝑑t\displaystyle\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla^{2}q|^{2}|\nabla y|^{2p}d\nu_{t}dt
\displaystyle\leq M|y|2(p+1)|q|2𝑑νt|t=2ϵ1t=12ϵ\displaystyle\left.\int_{M}|\nabla y|^{2(p+1)}|\nabla q|^{2}d\nu_{t}\right|_{t=-2\epsilon^{-1}}^{t=-\frac{1}{2}\epsilon}
+8(p+1)(2ϵ112ϵM|2y|2|y|2(2p+1)𝑑νt𝑑t)12(2ϵ112ϵM|2q|2|q|2𝑑νt𝑑t)12\displaystyle+8(p+1)\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla^{2}y|^{2}|\nabla y|^{2(2p+1)}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla^{2}q|^{2}|\nabla q|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}
\displaystyle\leq C(Y,ϵ,p)\displaystyle C(Y,\epsilon,p)

assuming δδ¯(Y,ϵ,p)\delta\leq\overline{\delta}(Y,\epsilon,p), where we used Bamler’s estimates for strong almost-splitting maps (Proposition 12.21 of [Bam3]).

(ii)(ii) For any t[2ϵ1,12ϵ]t\in[-2\epsilon^{-1},-\frac{1}{2}\epsilon],

M(q,y2y)𝑑νt=\displaystyle\int_{M}\left(\langle\nabla q,\nabla y\rangle-2y\right)d\nu_{t}= (4πτ)n24τMef,y𝑑gt+4τM(hf),y𝑑νt\displaystyle-(4\pi\tau)^{-\frac{n}{2}}4\tau\int_{M}\langle\nabla e^{-f},\nabla y\rangle dg_{t}+4\tau\int_{M}\langle\nabla(h-f),\nabla y\rangle d\nu_{t}
=\displaystyle= 4τM(Δy)𝑑νt+4τM(hf),y𝑑νt,\displaystyle 4\tau\int_{M}(\Delta y)d\nu_{t}+4\tau\int_{M}\langle\nabla(h-f),\nabla y\rangle d\nu_{t},

so we can estimate

2ϵ112ϵ|M(q,y2y)𝑑νt|𝑑t\displaystyle\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\left|\int_{M}\left(\langle\nabla q,\nabla y\rangle-2y\right)d\nu_{t}\right|dt\leq 4τn2ϵ112ϵM|2y|2𝑑νt𝑑t\displaystyle 4\tau n\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla^{2}y|^{2}d\nu_{t}dt
+4τ(2ϵ112ϵM|(hf)|2𝑑νt𝑑t)12(2ϵ112ϵM|y|2𝑑νt𝑑t)12\displaystyle+4\tau\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla(h-f)|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla y|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}
\displaystyle\leq Ψ(δ|Y,ϵ).\displaystyle\Psi(\delta|Y,\epsilon).

Using the L1L^{1}-Poincare inequality and Lemma 5.1,

2ϵ112ϵM|q,y2y|𝑑νt𝑑t\displaystyle\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}\left|\langle\nabla q,\nabla y\rangle-2y\right|d\nu_{t}dt
\displaystyle\leq 2ϵ112ϵ|M(q,y2y)𝑑νt|𝑑t+C2ϵ112ϵτM|(q,y2y)|𝑑νt𝑑t\displaystyle\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\left|\int_{M}\left(\langle\nabla q,\nabla y\rangle-2y\right)d\nu_{t}\right|dt+C\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\tau\int_{M}\left|\nabla\left(\langle\nabla q,\nabla y\rangle-2y\right)\right|d\nu_{t}dt
\displaystyle\leq Ψ(δ|Y,ϵ)+C(ϵ)2ϵ112ϵM|(4τRc+2q2g)(y)|𝑑νt𝑑t+C(ϵ)2ϵ112ϵM|Rc(y)|𝑑νt𝑑t\displaystyle\Psi(\delta|Y,\epsilon)+C(\epsilon)\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}\left|(4\tau Rc+\nabla^{2}q-2g)(\nabla y)\right|d\nu_{t}dt+C(\epsilon)\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|Rc(\nabla y)|d\nu_{t}dt
+C(ϵ)(2ϵ112ϵM|2y|2𝑑νt𝑑t)12(2ϵ112ϵM|q|2𝑑νt𝑑t)12\displaystyle+C(\epsilon)\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla^{2}y|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla q|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}
\displaystyle\leq Ψ(δ|Y,ϵ)+C(ϵ)(2ϵ112ϵM|4τRc+2q2g|2𝑑νt𝑑t)12(2ϵ112ϵM|y|2𝑑νt𝑑t)12\displaystyle\Psi(\delta|Y,\epsilon)+C(\epsilon)\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|4\tau Rc+\nabla^{2}q-2g|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla y|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}
\displaystyle\leq Ψ(δ|Y,ϵ).\displaystyle\Psi(\delta|Y,\epsilon).

Then Hölder’s inequality and (4.9) give

2ϵ112ϵM(q,y2y)2𝑑νt𝑑t\displaystyle\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}\left(\langle\nabla q,\nabla y\rangle-2y\right)^{2}d\nu_{t}dt
\displaystyle\leq (2ϵ112ϵM|q,y2y|𝑑νt𝑑t)12(2ϵ112ϵM|q,y2y|3𝑑νt𝑑t)12\displaystyle\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}\left|\langle\nabla q,\nabla y\rangle-2y\right|d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}\left|\langle\nabla q,\nabla y\rangle-2y\right|^{3}d\nu_{t}dt\right)^{\frac{1}{2}}
\displaystyle\leq Ψ(δ|Y,ϵ)(2ϵ112ϵM(|q|3|y|3+|y|3)𝑑νt𝑑t)12\displaystyle\Psi(\delta|Y,\epsilon)\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}(|\nabla q|^{3}|\nabla y|^{3}+|y|^{3})d\nu_{t}dt\right)^{\frac{1}{2}}
\displaystyle\leq Ψ(δ|Y,ϵ).\displaystyle\Psi(\delta|Y,\epsilon).

Next, we compute

(q,y2y)2=2|(q,y2y)|24(q,y2y)2q,2y.\square\left(\langle\nabla q,\nabla y\rangle-2y\right)^{2}=-2\left|\nabla\left(\langle\nabla q,\nabla y\rangle-2y\right)\right|^{2}-4\left(\langle\nabla q,\nabla y\rangle-2y\right)\langle\nabla^{2}q,\nabla^{2}y\rangle.

Fix a cutoff function ζC([2ϵ1,12ϵ])\zeta\in C^{\infty}([-2\epsilon^{-1},-\frac{1}{2}\epsilon]) such that ζ(2ϵ1)=ζ(12ϵ)=0\zeta(-2\epsilon^{-1})=\zeta(-\frac{1}{2}\epsilon)=0, ζ|[ϵ1,ϵ]1\zeta|[-\epsilon^{-1},-\epsilon]\equiv 1, and |ζ|4ϵ1|\zeta^{\prime}|\leq 4\epsilon^{-1}. Then

2ϵ112ϵMζ(q,y2y)2𝑑νt𝑑t=\displaystyle\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}\zeta^{\prime}\left(\langle\nabla q,\nabla y\rangle-2y\right)^{2}d\nu_{t}dt= 2ϵ112ϵζ(ddtM(q,y2y)2𝑑νt)𝑑t\displaystyle-\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\zeta\left(\frac{d}{dt}\int_{M}\left(\langle\nabla q,\nabla y\rangle-2y\right)^{2}d\nu_{t}\right)dt
=\displaystyle= 22ϵ112ϵMζ|(q,y2y)|2𝑑νt𝑑t\displaystyle 2\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}\zeta\left|\nabla\left(\langle\nabla q,\nabla y\rangle-2y\right)\right|^{2}d\nu_{t}dt
+42ϵ112ϵMζ(q,y2y)2q,2y𝑑νt𝑑t,\displaystyle+4\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}\zeta\left(\langle\nabla q,\nabla y\rangle-2y\right)\langle\nabla^{2}q,\nabla^{2}y\rangle d\nu_{t}dt,

so that (using part (i)(i) and Proposition 12.21 of [Bam20b])

ϵ1ϵM|(q,y2y)|2𝑑νt𝑑t\displaystyle\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left|\nabla\left(\langle\nabla q,\nabla y\rangle-2y\right)\right|^{2}d\nu_{t}dt
\displaystyle\leq 4ϵ12ϵ112ϵM(q,y2y)2𝑑νt𝑑t+42ϵ112ϵM|q||2q||y||2y|𝑑νt𝑑t\displaystyle 4\epsilon^{-1}\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}\left(\langle\nabla q,\nabla y\rangle-2y\right)^{2}d\nu_{t}dt+4\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla q||\nabla^{2}q|\cdot|\nabla y||\nabla^{2}y|d\nu_{t}dt
+42ϵ112ϵM|2q||2y||y|𝑑νt𝑑t\displaystyle+4\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla^{2}q|\cdot|\nabla^{2}y||y|d\nu_{t}dt
\displaystyle\leq Ψ(δ|Y,ϵ)+4(2ϵ112ϵM|2q|2|q|2𝑑νt𝑑t)12(2ϵ112ϵM|2y|2|y|2𝑑νt𝑑t)12\displaystyle\Psi(\delta|Y,\epsilon)+4\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla^{2}q|^{2}|\nabla q|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla^{2}y|^{2}|\nabla y|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}
+4(2ϵ112ϵM|2q|2𝑑νt𝑑t)12(2ϵ112ϵM|2y|2|y|2𝑑νt𝑑t)12\displaystyle+4\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla^{2}q|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-2\epsilon^{-1}}^{-\frac{1}{2}\epsilon}\int_{M}|\nabla^{2}y|^{2}|y|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}
\displaystyle\leq Ψ(δ|Y,ϵ).\displaystyle\Psi(\delta|Y,\epsilon).

Integrating

(|y|2|q|4)2|2y|2|q|4162y(y),2q(q)|q|2\square\left(|\nabla y|^{2}|\nabla q|^{4}\right)\leq-2|\nabla^{2}y|^{2}|\nabla q|^{4}-16\langle\nabla^{2}y(\nabla y),\nabla^{2}q(\nabla q)\rangle|\nabla q|^{2}

against the conjugate heat kernel, and applying part (i)(i) and (4.9) gives

ϵ1ϵM|2y|2|q|4𝑑νt𝑑t\displaystyle\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}y|^{2}|\nabla q|^{4}d\nu_{t}dt\leq |M|y|2|q|2dνt|t=ϵ1t=ϵ|\displaystyle\left|\left.\int_{M}|\nabla y|^{2}|\nabla q|^{2}d\nu_{t}\right|_{t=-\epsilon^{-1}}^{t=-\epsilon}\right|
+8(ϵ1ϵM|2y|2|y|2𝑑νt𝑑t)12(ϵ1ϵM|2q|2|q|6𝑑νt𝑑t)12\displaystyle+8\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}y|^{2}|\nabla y|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}q|^{2}|\nabla q|^{6}d\nu_{t}dt\right)^{\frac{1}{2}}
\displaystyle\leq C(Y,ϵ)\displaystyle C(Y,\epsilon)

assuming δδ¯(Y,ϵ)\delta\leq\overline{\delta}(Y,\epsilon). We can use Hölder’s inequality to estimate

ϵ1ϵM|(2q2g)(y)|2𝑑νt𝑑t\displaystyle\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left|(\nabla^{2}q-2g)(\nabla y)\right|^{2}d\nu_{t}dt
2ϵ1ϵM|(q,y2y)|2𝑑νt𝑑t+2ϵ1ϵM|2y|2|q|2𝑑νt𝑑t\displaystyle\leq 2\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left|\nabla(\langle\nabla q,\nabla y\rangle-2y)\right|^{2}d\nu_{t}dt+2\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}y|^{2}|\nabla q|^{2}d\nu_{t}dt
Ψ(δ|Y,ϵ)+(ϵ1ϵ|2y|2|q|4𝑑νt𝑑t)12(ϵ1ϵM|2y|2𝑑νt𝑑t)12\displaystyle\leq\Psi(\delta|Y,\epsilon)+\left(\int_{-\epsilon^{-1}}^{-\epsilon}|\nabla^{2}y|^{2}|\nabla q|^{4}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}y|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}
Ψ(δ|Y,ϵ)\displaystyle\leq\Psi(\delta|Y,\epsilon)

and we also obtain the rough estimate

ϵ1ϵM|(2q+2g)(y)|2𝑑νt𝑑t2ϵ1ϵM(|2q|2|y|2+4|y|2)𝑑νt𝑑tC(Y,ϵ).\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left|(\nabla^{2}q+2g)(\nabla y)\right|^{2}d\nu_{t}dt\leq 2\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left(|\nabla^{2}q|^{2}|\nabla y|^{2}+4|\nabla y|^{2}\right)d\nu_{t}dt\leq C(Y,\epsilon).

Now use Hölder’s inequality, and combine estimates:

ϵ1ϵM||2q(y)|24|𝑑νt𝑑t\displaystyle\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left||\nabla^{2}q(\nabla y)|^{2}-4\right|d\nu_{t}dt
\displaystyle\leq ϵ1ϵM(2q+2g)(y),(2q2g)(y)𝑑νt𝑑t+4ϵ1ϵM|1|y|2|𝑑νt𝑑t\displaystyle\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left\langle(\nabla^{2}q+2g)(\nabla y),(\nabla^{2}q-2g)(\nabla y)\right\rangle d\nu_{t}dt+4\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left|1-|\nabla y|^{2}\right|d\nu_{t}dt
\displaystyle\leq Ψ(δ|Y,ϵ).\displaystyle\Psi(\delta|Y,\epsilon).

On the other hand, we can estimate

ϵ1ϵM||2q(y)|2|2q(Jy)|2|𝑑νt𝑑t\displaystyle\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left||\nabla^{2}q(\nabla y)|^{2}-|\nabla^{2}q(J\nabla y)|^{2}\right|d\nu_{t}dt
\displaystyle\leq (ϵ1ϵM||2q(y1+|y|2)|2|2q(Jy1+|y|2)|2|𝑑νt𝑑t)12\displaystyle\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left|\left|\nabla^{2}q\left(\frac{\nabla y}{\sqrt{1+|\nabla y|^{2}}}\right)\right|^{2}-\left|\nabla^{2}q\left(J\frac{\nabla y}{\sqrt{1+|\nabla y|^{2}}}\right)\right|^{2}\right|d\nu_{t}dt\right)^{\frac{1}{2}}
×(ϵ1ϵM||2q(y)|2|2q(Jy)|2|(1+|y|2)𝑑νt𝑑t)12.\displaystyle\times\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left||\nabla^{2}q(\nabla y)|^{2}-|\nabla^{2}q(J\nabla y)|^{2}\right|\left(1+|\nabla y|^{2}\right)d\nu_{t}dt\right)^{\frac{1}{2}}.

To estimate the first integral, we observe that

||2q(y)|2|2q(Jy)|2|\displaystyle\left|\left|\nabla^{2}q\left(\nabla y\right)\right|^{2}-\left|\nabla^{2}q\left(J\nabla y\right)\right|^{2}\right|
=\displaystyle= |(|2q(y)|2|(4τRc2g)(y)|2)(|2q(Jy)|2|(4τRc2g)(Jy)|2)|\displaystyle\left|\left(\left|\nabla^{2}q\left(\nabla y\right)\right|^{2}-|(4\tau Rc-2g)(\nabla y)|^{2}\right)-\left(\left|\nabla^{2}q\left(J\nabla y\right)\right|^{2}-|(4\tau Rc-2g)(J\nabla y)|^{2}\right)\right|
\displaystyle\leq |(4τRc+2q+2g)(y),(4τRc+2q2g)(y)|\displaystyle\left|\left\langle\left(-4\tau Rc+\nabla^{2}q+2g\right)(\nabla y),\left(4\tau Rc+\nabla^{2}q-2g\right)(\nabla y)\right\rangle\right|
+|(4τRc+2q+2g)(Jy),(4τRc+2q2g)(Jy)|,\displaystyle+\left|\left\langle\left(-4\tau Rc+\nabla^{2}q+2g\right)(J\nabla y),\left(4\tau Rc+\nabla^{2}q-2g\right)(J\nabla y)\right\rangle\right|,

so that

ϵ1ϵM||2q(y1+|y|2)|2|2q(Jy1+|y|2)|2|𝑑νt𝑑t\displaystyle\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left|\left|\nabla^{2}q\left(\frac{\nabla y}{\sqrt{1+|\nabla y|^{2}}}\right)\right|^{2}-\left|\nabla^{2}q\left(J\frac{\nabla y}{\sqrt{1+|\nabla y|^{2}}}\right)\right|^{2}\right|d\nu_{t}dt
\displaystyle\leq 2(ϵ1ϵM|4τRc+2q2g|2𝑑νt𝑑t)12(ϵ1ϵM|4τRc2q2g|2𝑑νt𝑑t)12\displaystyle 2\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|4\tau Rc+\nabla^{2}q-2g|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|4\tau Rc-\nabla^{2}q-2g|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}
\displaystyle\leq Ψ(δ|Y,ϵ).\displaystyle\Psi(\delta|Y,\epsilon).

For the second integral, we only need the course upper bound

ϵ1ϵM||2q(y)|2|2q(Jy)|2|(1+|y|2)𝑑νt𝑑t\displaystyle\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left||\nabla^{2}q(\nabla y)|^{2}-|\nabla^{2}q(J\nabla y)|^{2}\right|\left(1+|\nabla y|^{2}\right)d\nu_{t}dt ϵ1ϵM4|2q|2(|y|4+|y|2)𝑑νt𝑑t\displaystyle\leq\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}4|\nabla^{2}q|^{2}\left(|\nabla y|^{4}+|\nabla y|^{2}\right)d\nu_{t}dt
C(Y,ϵ)\displaystyle\leq C(Y,\epsilon)

by part (i)(i). Combining expressions, we finally obtain

ϵ1ϵM||z|21|𝑑νt𝑑t\displaystyle\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left||\nabla z|^{2}-1\right|d\nu_{t}dt\leq 14ϵ1ϵM||2q(Jy)|24|dνtdt+14ϵ1ϵM|2y|2|q|2𝑑νt𝑑t\displaystyle\frac{1}{4}\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left||\nabla^{2}q(J\nabla y)|^{2}-4\right|d\nu_{t}dt+\frac{1}{4}\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}y|^{2}|\nabla q|^{2}d\nu_{t}dt
+12(ϵ1ϵM|2q|2|q|2𝑑νt𝑑t)12(ϵ1ϵM|2y|2|q|2𝑑νt𝑑t)12\displaystyle+\frac{1}{2}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}q|^{2}|\nabla q|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}y|^{2}|\nabla q|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}
\displaystyle\leq Ψ(δ|Y,ϵ).\displaystyle\Psi(\delta|Y,\epsilon).

(iii)(iii) Because XRc(JX,X)X\mapsto Rc(JX,X) and Xg(JX,X)X\mapsto g(JX,X) are skew-symmetric, we have

z,y=\displaystyle\langle\nabla z,\nabla y\rangle= 122q(Jy,y)+12yJy,q\displaystyle\frac{1}{2}\nabla^{2}q(J\nabla y,\nabla y)+\frac{1}{2}\langle\nabla_{\nabla y}J\nabla y,\nabla q\rangle
=\displaystyle= 12(4τRc+2q2g)(Jy,y)122y(y,Jq),\displaystyle\frac{1}{2}\left(4\tau Rc+\nabla^{2}q-2g\right)(J\nabla y,\nabla y)-\frac{1}{2}\nabla^{2}y(\nabla y,J\nabla q),

which allows us to estimate (again using Proposition 12.21 of [Bam20b])

ϵ1ϵM|z,y|𝑑νt𝑑t\displaystyle\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left|\langle\nabla z,\nabla y\rangle\right|d\nu_{t}dt\leq ϵ1ϵM|4τRc+2q2g,yJy|𝑑νt𝑑t\displaystyle\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left|\langle 4\tau Rc+\nabla^{2}q-2g,\nabla y\otimes J\nabla y\rangle\right|d\nu_{t}dt
+ϵ1ϵM|2y||y||q|𝑑νt𝑑t\displaystyle+\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}y|\cdot|\nabla y|\cdot|\nabla q|d\nu_{t}dt
\displaystyle\leq (ϵ1ϵM|4τRc+2q2g|2𝑑νt𝑑t)12(ϵ1ϵM|y|4𝑑νt𝑑t)12\displaystyle\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left|4\tau Rc+\nabla^{2}q-2g\right|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla y|^{4}d\nu_{t}dt\right)^{\frac{1}{2}}
+(ϵ1ϵM|2y|2|y|2𝑑νt𝑑t)12(ϵ1ϵM|q|2𝑑νt𝑑t)12\displaystyle+\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}y|^{2}|\nabla y|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla q|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}
\displaystyle\leq Ψ(δ|Y,ϵ).\displaystyle\Psi(\delta|Y,\epsilon).

(iv)(iv) We compute

q,Jy=\displaystyle\square\langle\nabla q,J\nabla y\rangle= 2Rc(q,Jy)+q,Jy+q,Jy2q,(Jy)\displaystyle 2Rc(\nabla q,J\nabla y)+\langle\square\nabla q,J\nabla y\rangle+\langle\nabla q,J\square\nabla y\rangle-\langle\nabla^{2}q,\nabla(J\nabla y)\rangle
=\displaystyle= 2Rc(q,Jy)Rc(q),Jyq,JRc(y)2q,(Jy)\displaystyle 2Rc(\nabla q,J\nabla y)-\langle Rc(\nabla q),J\nabla y\rangle-\langle\nabla q,JRc(\nabla y)\rangle-\langle\nabla^{2}q,\nabla(J\nabla y)\rangle
=\displaystyle= 2q,(Jy),\displaystyle-\langle\nabla^{2}q,\nabla(J\nabla y)\rangle,

so that

ϵ1ϵM|z|𝑑νt𝑑t(ϵ1ϵM|2q|2𝑑νt𝑑t)12(ϵ1ϵM|2y|2𝑑νt𝑑t)12Ψ(δ|Y,ϵ).\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\square z|d\nu_{t}dt\leq\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}q|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\left(\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}y|^{2}d\nu_{t}dt\right)^{\frac{1}{2}}\leq\Psi(\delta|Y,\epsilon).

(v)(v) Using part (ii)(ii) and Lemma 5.1, we estimate

ϵ1ϵM|zJy|2𝑑νt𝑑t\displaystyle\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left|\nabla z-J\nabla y\right|^{2}d\nu_{t}dt
=ϵ1ϵM(|z|2+|y|22z,Jy)𝑑νt𝑑t\displaystyle=\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left(|\nabla z|^{2}+|\nabla y|^{2}-2\langle\nabla z,J\nabla y\rangle\right)d\nu_{t}dt
ϵ1ϵM(22q(Jy,Jy))𝑑νt𝑑t+ϵ1ϵM|2y||y||q|𝑑νt𝑑t+Ψ(δ|Y,ϵ)\displaystyle\leq\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left(2-\nabla^{2}q(J\nabla y,J\nabla y)\right)d\nu_{t}dt+\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla^{2}y|\cdot|\nabla y|\cdot|\nabla q|d\nu_{t}dt+\Psi(\delta|Y,\epsilon)
ϵ1ϵM|4τRc+2q2g||y|2𝑑νt𝑑t\displaystyle\leq\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\left|4\tau Rc+\nabla^{2}q-2g\right|\cdot|\nabla y|^{2}d\nu_{t}dt
+ϵ1ϵM|Rc(y)||y|𝑑νt𝑑t+Ψ(δ|Y,ϵ)\displaystyle\hskip 11.38109pt+\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|Rc(\nabla y)|\cdot|\nabla y|d\nu_{t}dt+\Psi(\delta|Y,\epsilon)
ϵ1ϵM|Rc(y)|||y|1|dνtdt+ϵ1ϵM|Rc(y)|𝑑νt𝑑t+Ψ(δ|Y,ϵ)\displaystyle\leq\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|Rc(\nabla y)|\cdot\left||\nabla y|-1\right|d\nu_{t}dt+\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|Rc(\nabla y)|d\nu_{t}dt+\Psi(\delta|Y,\epsilon)
Ψ(δ|Y,ϵ).\displaystyle\leq\Psi(\delta|Y,\epsilon).

Next, we prove an elementary lemma which will allow us to form almost splitting maps using a linear combination of almost splitting maps along with the new almost-splitting maps constructed in Lemma 5.3.

Lemma 5.4.

Given N,kN,k\in\mathbb{N}, there exists C=C(N)<C=C(N)<\infty such that the following holds. Suppose (𝒱,,)(\mathcal{V},\langle\cdot,\cdot\rangle) is a real inner product space of dimension at most NN, and let JJ be a complex structure on 𝒱\mathcal{V} compatible with the inner product: Jv,Jw=v,w\langle Jv,Jw\rangle=\langle v,w\rangle for all v,w𝒱v,w\in\mathcal{V}. If v1,,v2k+1𝒱v_{1},...,v_{2k+1}\in\mathcal{V} are orthonormal, then there are (aij)1i2k+21j2k+1(a_{ij})_{1\leq i\leq 2k+2}^{1\leq j\leq 2k+1} and (bij)1i2k+21j2k+1(b_{ij})_{1\leq i\leq 2k+2}^{1\leq j\leq 2k+1} with |aij|+|bij|C|a_{ij}|+|b_{ij}|\leq C such that

v~i:=j=12k+1(aijvj+bijJvj),1i2k+2,\widetilde{v}_{i}:=\sum_{j=1}^{2k+1}\left(a_{ij}v_{j}+b_{ij}Jv_{j}\right),\hskip 17.07164pt1\leq i\leq 2k+2,

are orthonormal.

Proof.

It suffices to show the existence of c(n)>0c(n)>0 such that for any orthonormal tuple (v1,,v2k+1)(v_{1},...,v_{2k+1}) in 𝒱\mathcal{V}, there exists i{1,,2k+1}i\in\{1,...,2k+1\} such that

wi:=Jvij=12k+1Jvi,vjvjw_{i}:=Jv_{i}-\sum_{j=1}^{2k+1}\langle Jv_{i},v_{j}\rangle v_{j}

satisfies |wi|c(n)|w_{i}|\geq c(n). Suppose by way of contradiction there exist orthonormal tuples (v1α,,v2k+1α)α(v_{1}^{\alpha},...,v_{2k+1}^{\alpha})_{\alpha\in\mathbb{N}} such that max1i2k+1|wiα|0\max_{1\leq i\leq 2k+1}|w_{i}^{\alpha}|\to 0 as α\alpha\to\infty. We can pass to subsequences so that limαviα=vi\lim_{\alpha\to\infty}v_{i}^{\alpha}=v_{i}^{\infty}, where (v1,,v2k+1)(v_{1}^{\infty},...,v_{2k+1}^{\infty}) is an orthonormal tuple. Then, for each i{1,,2k+1}i\in\{1,...,2k+1\}, we have

0=limαwiα=Jvij=12k+1Jvi,vjvj.0=\lim_{\alpha\to\infty}w_{i}^{\alpha}=Jv_{i}^{\infty}-\sum_{j=1}^{2k+1}\langle Jv_{i}^{\infty},v_{j}^{\infty}\rangle v_{j}^{\infty}.

That is, Jv1,,Jv2k+1Jv_{1},...,Jv_{2k+1} are in the \mathbb{R}-linear span of v1,,v2k+1v_{1},...,v_{2k+1}. This means that the \mathbb{R}-linear span of v1,,v2k+1v_{1},...,v_{2k+1}, equipped with the restriction of the complex structure, is a complex vector space of real dimension 2k+12k+1, a contradiction. ∎

We finally establish the improved splitting for Kähler-Ricci flows.

Proposition 5.5.

For any ϵ>0\epsilon>0, k{0,,n}k\in\{0,...,n\}, Y<Y<\infty, the following holds whenever δδ¯(ϵ,Y)\delta\leq\overline{\delta}(\epsilon,Y). Suppose (M2n,(gt)tI)(M^{2n},(g_{t})_{t\in I}) is a Kähler-Ricci flow with 𝒩x0,t0(r2)Y\mathcal{N}_{x_{0},t_{0}}(r^{2})\geq-Y for some r>0r>0. If (x0,t0)M×I(x_{0},t_{0})\in M\times I is (δ,r)(\delta,r)-selfsimilar and strongly (2k+1,δ,r)(2k+1,\delta,r)-split, then (x0,t0)(x_{0},t_{0}) is weakly (2k+2,ϵ,r)(2k+2,\epsilon,r)-split.

Proof.

By parabolic rescaling and time translation, we can assume t0=0t_{0}=0 and r=1r=1. For ease of notation, write νt:=νx0,0;t\nu_{t}:=\nu_{x_{0},0;t}. Let y:M×[δ1,δ]2k+1y:M\times[-\delta^{-1},\delta]\to\mathbb{R}^{2k+1} be a strong (2k+1,δ,1)(2k+1,\delta,1)-splitting map, and let qC(M×[δ1,δ])q\in C^{\infty}(M\times[-\delta^{-1},-\delta]) be a strong (δ,1)(\delta,1)-soliton potential, both based at (x0,0)(x_{0},0). For each i{1,,2k+1}i\in\{1,...,2k+1\}, Proposition 5.3 states that the functions zi:=12q,Jyiz_{i}:=\frac{1}{2}\langle\nabla q,J\nabla y_{i}\rangle are weak (1,Ψ(δ|Y),1)(1,\Psi(\delta|Y),1)-splitting maps based at (x0,0)(x_{0},0) which satisfy

ϵ1ϵM|ziJyi|2𝑑νt𝑑tΨ(δ|Y,ϵ).\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}|\nabla z_{i}-J\nabla y_{i}|^{2}d\nu_{t}dt\leq\Psi(\delta|Y,\epsilon).

By replacing yy with Ay+bA\circ y+b for some A(2k+1)×(2k+1)A\in\mathbb{R}^{(2k+1)\times(2k+1)} with |AI2k+1|Ψ(δ|Y,ϵ)|A-I_{2k+1}|\leq\Psi(\delta|Y,\epsilon) and |b|Ψ(δ|Y,ϵ)|b|\leq\Psi(\delta|Y,\epsilon), we can assume that

ϵ1ϵMyi,yj𝑑νt𝑑t=δij.\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\langle\nabla y_{i},\nabla y_{j}\rangle d\nu_{t}dt=\delta_{ij}.

Similar to the proof of Proposition 10.8 in [Bam20b], we consider the finite-dimensional real vector space 𝒱\mathcal{V} spanned by yi,Jyi𝔛(M×[ϵ1,ϵ])\nabla y_{i},J\nabla y_{i}\in\mathfrak{X}(M\times[-\epsilon^{-1},-\epsilon]), equipped with the restricted L2L^{2} inner product

X1,X2L2=ϵ1ϵMX1,X2𝑑νt𝑑t\langle X_{1},X_{2}\rangle_{L^{2}}=\int_{-\epsilon^{-1}}^{-\epsilon}\int_{M}\langle X_{1},X_{2}\rangle d\nu_{t}dt

and the obvious complex structure. Lemma 5.4 then provides aij,bija_{ij},b_{ij}, where 1i2k+21\leq i\leq 2k+2, 1j2k+11\leq j\leq 2k+1, such that

V~i:=j=12k+1aijyj+bijJyj\widetilde{V}_{i}:=\sum_{j=1}^{2k+1}a_{ij}\nabla y_{j}+b_{ij}J\nabla y_{j}

are orthonormal in 𝒱\mathcal{V}, and |aij|,|bij|C(k)|a_{ij}|,|b_{ij}|\leq C(k). It follows that

y~i:=j=12k+1aijyj+bijzj\widetilde{y}_{i}:=\sum_{j=1}^{2k+1}a_{ij}y_{j}+b_{ij}z_{j}

define a weak (2k+2,Ψ(δ|Y,ϵ),1)(2k+2,\Psi(\delta|Y,\epsilon),1)-splitting map y~=(y~1,,y~2k+2)\widetilde{y}=(\widetilde{y}_{1},...,\widetilde{y}_{2k+2}) based at (x0,0)(x_{0},0). ∎

Proof of Theorem 1.2.

We first verify the existence of a lower bound of the Nash entropy on all of P(x,A,A2)P^{\ast}(x_{\infty},A,-A^{2}). Given yP(x,A,A2)y\in P^{\ast}(x_{\infty},A,-A^{2}), we can find a sequence (yi,ti)Mi×(Ti,0](y_{i},t_{i})\in M_{i}\times(-T_{i},0] such that (yi,ti)iy(y_{i},t_{i})\xrightarrow[i\to\infty]{\mathfrak{C}}y_{\infty}. Lemma 15.8 of [Bam20b] implies that (yi,ti)P(xi,2A,(2A)2)(y_{i},t_{i})\in P^{\ast}(x_{i},2A,-(2A)^{2}), and Bamler’s Nash entropy oscillation estimate (Corollary 5.11 in [Bam20a]) then gives 𝒩yi,ti(1)Y(Y,A)\mathcal{N}_{y_{i},t_{i}}(1)\geq-Y^{\prime}(Y,A). Taking the limit as ii\to\infty, we obtain (via Theorem 15.45 in [Bam20b]) 𝒩y(1)Y(Y,A)\mathcal{N}_{y}(1)\geq-Y^{\prime}(Y,A). The inclusion

𝒮^r1,r2ϵ,2j+1P(x;A,A2)𝒮^r1,r2δ(ϵ,Y,A),2jP(x;A,A2)\widehat{\mathcal{S}}_{r_{1},r_{2}}^{\epsilon,2j+1}\cap P^{\ast}(x_{\infty};A,-A^{2})\subseteq\widehat{\mathcal{S}}_{r_{1},r_{2}}^{\delta(\epsilon,Y,A),2j}\cap P^{\ast}(x_{\infty};A,-A^{2})

thus follows from Proposition 5.5. The remaining claim is then a consequence of the inclusions between quantitative strata and weak quantitative strata (Lemma 20.3 of [Bam20b] and Proposition 3.4). ∎

Proof of Theorem 1.1.

Taking r1=0r_{1}=0 and r2=ϵr_{2}=\epsilon in Theorem 1.2 gives

𝒮^0,ϵϵ,2j+1P(x,A,A2)𝒮^0,ϵδ(ϵ,Y,A),2jP(x;A,A2).\widehat{\mathcal{S}}_{0,\epsilon}^{\epsilon,2j+1}\cap P^{\ast}(x_{\infty},A,-A^{2})\subseteq\widehat{\mathcal{S}}_{0,\epsilon}^{\delta(\epsilon,Y,A),2j}\cap P^{\ast}(x_{\infty};A,-A^{2}).

Then taking the union over ϵ>0\epsilon>0 gives

𝒮2j+1P(x,A,A2)𝒮2j+2P(x,A,A2).\mathcal{S}^{2j+1}\cap P^{\ast}(x_{\infty},A,-A^{2})\subseteq\mathcal{S}^{2j+2}\cap P^{\ast}(x_{\infty},A,-A^{2}).

Finally, taking AA\nearrow\infty gives the claim. ∎

6. An Isometric Action on Tangent Flows

Suppose (M,g,J,f)(M,g,J,f) is a (not necessarily complete) shrinking gradient Kähler-Ricci soliton, and let ω:=g(J,)\omega:=g(J\cdot,\cdot) be the corresponding Kähler form. Then the Ricci soliton equation gives

(JfJ)(W)\displaystyle(\mathcal{L}_{J\nabla f}J)(W) =Jf(JW)J([Jf,W])=Jf(JW)JW(Jf)J(JfWW(Jf))\displaystyle=\mathcal{L}_{J\nabla f}(JW)-J([J\nabla f,W])=\nabla_{J\nabla f}(JW)-\nabla_{JW}(J\nabla f)-J\left(\nabla_{J\nabla f}W-\nabla_{W}(J\nabla f)\right)
=JJWfWf=JRc(JW)+Rc(W)12(J2W+W)=0\displaystyle=-J\nabla_{JW}\nabla f-\nabla_{W}\nabla f=JRc(JW)+Rc(W)-\frac{1}{2}(J^{2}W+W)=0

for any vector field W𝔛()W\in\mathfrak{X}(\mathcal{R}), and

Jfω=diJfω=d(ω(Jf,))=d(g(f,))=d(df)=0,\mathcal{L}_{J\nabla f}\omega=di_{J\nabla f}\omega=d\left(\omega(J\nabla f,\cdot)\right)=-d\left(g(\nabla f,\cdot)\right)=-d(df)=0,

so that JfJ\nabla f is a real holomorphic Killing vector field on MM. We now prove the completeness of the flow of this vector field for tangent flows.

Proposition 6.1.

Suppose (Mi2n,(gi,t)t[ϵi1,0])(M_{i}^{2n},(g_{i,t})_{t\in[-\epsilon_{i}^{-1},0]}) are closed Kähler-Ricci flows, and that (xi,0)(x_{i},0) are (ϵi,1)(\epsilon_{i},1)-selfsimilar, where ϵi0\epsilon_{i}\searrow 0. Assume

(M,(gi,t)t(ϵi1,0),(νxi,0;t)t[ϵi1,0))i𝔽(𝒳,(νx;t)t(,0))(M,(g_{i,t})_{t\in(-\epsilon_{i}^{-1},0)},(\nu_{x_{i},0;t})_{t\in[-\epsilon_{i}^{-1},0)})\xrightarrow[i\to\infty]{\mathbb{F}}(\mathcal{X},(\nu_{x_{\infty};t})_{t\in(-\infty,0)})

on compact time intervals where 𝒳\mathcal{X} is a metric soliton modeled on a singular shrinking Kähler-Ricci soliton (X,d,X,gX,fX)(X,d,\mathcal{R}_{X},g_{X},f_{X}) as in Theorem 2.7. Set q:=4τ(fXW)C()q:=4\tau(f_{X}-W)\in C^{\infty}(\mathcal{R}), where 𝒳\mathcal{R}\subseteq\mathcal{X} is the regular part of the metric flow. Then JqJ\nabla q is complete, and the heat kernel satisfies the following infinitesimal symmetry for all (x1,x0)×(x_{1},x_{0})\in\mathcal{R}\times\mathcal{R} with 𝔱(x0)<𝔱(x1)\mathfrak{t}(x_{0})<\mathfrak{t}(x_{1}):

x1K(x1;x0),Jq(x1)+x0K(x1;x0),Jq(x0)=0.\langle\nabla_{x_{1}}K(x_{1};x_{0}),J\nabla q(x_{1})\rangle+\langle\nabla_{x_{0}}K(x_{1};x_{0}),J\nabla q(x_{0})\rangle=0.

Moreover, the flow of JqJ\nabla q extends to a 1-parameter action by isometries on all of XX.

Remark 6.2.

This proof is modeled on Theorem 15.50 in [Bam20b].

Proof.

Let (Ui)(U_{i}) be a precompact exhaustion of ,\mathcal{R},with open embeddings ψi:UiMi×(ϵi1,0)\psi_{i}:U_{i}\to M_{i}\times(-\epsilon_{i}^{-1},0) realizing the 𝔽\mathbb{F}-convergence on the regular part as in Theorem 2.5. By Proposition 4.2 and the proof of Theorem 15.69 in [Bam20b], we can find almost-GRS potential functions hiC(Mi×[ϵi1,ϵi])h_{i}\in C^{\infty}(M_{i}\times[-\epsilon_{i}^{-1},-\epsilon_{i}]) such that, hiψifXh_{i}\circ\psi_{i}\to f_{X} in Cloc()C_{loc}^{\infty}(\mathcal{R}), where we identify X×(,0)\mathcal{R}\cong\mathcal{R}_{X}\times(-\infty,0) as in Theorem 2.7. Now fix t0(,0)t_{0}\in(-\infty,0) and uCc(t0)u^{\infty}\in C_{c}^{\infty}(\mathcal{R}_{t_{0}}), and let uiC(Mi×{t0})u^{i}\in C^{\infty}(M_{i}\times\{t_{0}\}) be an approximating sequence: uiψiuu^{i}\circ\psi_{i}\to u^{\infty} in Cloc()C_{loc}^{\infty}(\mathcal{R}). Let (ζhi)h(α,α)(\zeta_{h}^{i})_{h\in(-\alpha,\alpha)} be the flow of JiqiJ_{i}\nabla q_{i}, and define u,iC(Mi×{t0}×(α,α))u^{\prime,i}\in C^{\infty}(M_{i}\times\{t_{0}\}\times(-\alpha,\alpha)) by u,i(x,t0,h):=ui(ζhi(x),t0)u^{\prime,i}(x,t_{0},h):=u^{i}(\zeta_{h}^{i}(x),t_{0}), so that u,i(,t0,0)=ui(,t0)u^{\prime,i}(\cdot,t_{0},0)=u^{i}(\cdot,t_{0}) and hu,i(x,t0,h)=ui,Jiqi(ζhi(x),t0)\partial_{h}u^{\prime,i}(x,t_{0},h)=\langle\nabla u^{i},J_{i}\nabla q_{i}\rangle(\zeta_{h}^{i}(x),t_{0}). Next, let u′′,iC(Mi×[t0,0]×(α,α))u^{\prime\prime,i}\in C^{\infty}(M_{i}\times[t_{0},0]\times(-\alpha,\alpha)) be given by u′′,i(,t0,)=u,iu^{\prime\prime,i}(\cdot,t_{0},\cdot)=u^{\prime,i} and u′′,i(,,h)=0\square u^{\prime\prime,i}(\cdot,\cdot,h)=0 for all h(α,α)h\in(-\alpha,\alpha). Letting ωi,ρiΩ2(Mi)\omega_{i},\rho_{i}\in\Omega^{2}(M_{i}) denote the Kähler and Ricci 2-forms, respectively, we have

2u′′,i,gi(Ji,)=2u′′,i,ωi=0,\langle\nabla^{2}u^{\prime\prime,i},g_{i}(J_{i}\cdot,\cdot)\rangle=\langle\nabla^{2}u^{\prime\prime,i},\omega_{i}\rangle=0,
2u′′,i,Rcgi(Ji,)=2u′′,i,ρi=0,\langle\nabla^{2}u^{\prime\prime,i},Rc_{g_{i}}(J_{i}\cdot,\cdot)\rangle=\langle\nabla^{2}u^{\prime\prime,i},\rho_{i}\rangle=0,

so we can estimate

|hu′′,iu′′,i,Jiqi|\displaystyle\square\left|\partial_{h}u^{\prime\prime,i}-\langle\nabla u^{\prime\prime,i},J_{i}\nabla q_{i}\rangle\right|\leq 2|2u′′,i,(Jiqi)|=2|2u′′,i,(4τRc(gi)+2qi2gi)(Ji,)|\displaystyle 2\left|\langle\nabla^{2}u^{\prime\prime,i},\nabla(J_{i}\nabla q_{i})\rangle\right|=2\left|\left\langle\nabla^{2}u^{\prime\prime,i},\left(4\tau Rc(g_{i})+\nabla^{2}q_{i}-2g_{i}\right)(J_{i}\cdot,\cdot)\right\rangle\right|
\displaystyle\leq 2|2u′′,i||4τRc(gi)+2qi2gi|.\displaystyle 2|\nabla^{2}u^{\prime\prime,i}|\cdot|4\tau Rc(g_{i})+\nabla^{2}q_{i}-2g_{i}|.

Let νi:=νxi,0\nu^{i}:=\nu_{x_{i},0}, and integrate |u′′,i|2=2|2u′′,i|2\square|\nabla u^{\prime\prime,i}|^{2}=-2|\nabla^{2}u^{\prime\prime,i}|^{2} against νi\nu^{i} to obtain

2t0t1Mi|2u′′,i|2𝑑νti𝑑tMi|u,i|2𝑑νt0i.2\int_{t_{0}}^{t_{1}}\int_{M_{i}}|\nabla^{2}u^{\prime\prime,i}|^{2}d\nu_{t}^{i}dt\leq\int_{M_{i}}|\nabla u^{\prime,i}|^{2}d\nu_{t_{0}}^{i}.

However, we know that u,iψiu,u^{\prime,i}\circ\psi_{i}\to u^{\prime,\infty} in Cloc(t0)C_{loc}^{\infty}(\mathcal{R}_{t_{0}}), where u,(x)=u(ζh(x))u^{\prime,\infty}(x)=u^{\infty}(\zeta_{h}^{\infty}(x)), and (ζh)(\zeta_{h}^{\infty}) is the (partially defined) flow of JqJ\nabla q. In particular, we can estimate, for any t1(t0,0)t_{1}\in(t_{0},0),

supt[t0,t1]Mi|hu′′,iu′′,i,Jiqi|𝑑νti\displaystyle\sup_{t\in[t_{0},t_{1}]}\int_{M_{i}}\left|\partial_{h}u^{\prime\prime,i}-\langle\nabla u^{\prime\prime,i},J_{i}\nabla q_{i}\rangle\right|d\nu_{t}^{i}\leq (t0t1Mi|4τRc(gi)+2qi2gi|2𝑑νti𝑑t)12(Mi|u,i|2𝑑νt0i)12\displaystyle\left(\int_{t_{0}}^{t_{1}}\int_{M_{i}}|4\tau Rc(g_{i})+\nabla^{2}q_{i}-2g_{i}|^{2}d\nu_{t}^{i}dt\right)^{\frac{1}{2}}\left(\int_{M_{i}}|\nabla u^{\prime,i}|^{2}d\nu_{t_{0}}^{i}\right)^{\frac{1}{2}}
\displaystyle\leq Ψ(ϵi|t0,t1).\displaystyle\Psi(\epsilon_{i}|t_{0},t_{1}).

Because u′′,iψiu′′,u^{\prime\prime,i}\circ\psi_{i}\to u^{\prime\prime,\infty} in Cloc([t0,0))C_{loc}^{\infty}(\mathcal{R}_{[t_{0},0)}), where u′′,(x,h)=t0K(x;y)u(y,h)𝑑gt0(y)u^{\prime\prime,\infty}(x,h)=\int_{\mathcal{R}_{t_{0}}}K(x;y)u^{\prime}(y,h)dg_{t_{0}}(y), we obtain hu′′,=u′′,,Jq\partial_{h}u^{\prime\prime,\infty}=\langle\nabla u^{\prime\prime,\infty},J\nabla q\rangle. We can therefore compute, for all xt1x\in\mathcal{R}_{t_{1}},

h|h=0u′′,(x,0)=\displaystyle\partial_{h}|_{h=0}u^{\prime\prime,\infty}(x,0)= h|h=0t0K(x;y)u(y,h)𝑑gt0(y)=t0K(x;y)u,Jq(y)𝑑gt0(y)\displaystyle\left.\frac{\partial}{\partial h}\right|_{h=0}\int_{\mathcal{R}_{t_{0}}}K(x;y)u^{\prime}(y,h)dg_{t_{0}}(y)=\int_{\mathcal{R}_{t_{0}}}K(x;y)\langle\nabla u,J\nabla q\rangle(y)dg_{t_{0}}(y)
=\displaystyle= t0u(y)(yK(x;y),Jq(y)+K(x;y)div(Jq)(y))𝑑gt0(y)\displaystyle-\int_{\mathcal{R}_{t_{0}}}u(y)\left(\langle\nabla_{y}K(x;y),J\nabla q(y)\rangle+K(x;y)\text{div}(J\nabla q)(y)\right)dg_{t_{0}}(y)
=\displaystyle= t0u(y)yK(x;y),Jq(y)𝑑gt0(y)\displaystyle-\int_{\mathcal{R}_{t_{0}}}u(y)\langle\nabla_{y}K(x;y),J\nabla q(y)\rangle dg_{t_{0}}(y)

since

div(Jq)=(Jq),g=2ω4τρ,g=0.\text{div}(J\nabla q)=\langle\nabla(J\nabla q),g\rangle=-\langle 2\omega-4\tau\rho,g\rangle=0.

On the other hand, we have

h|h=0u′′,(x,0)=u′′,(x,0),Jq(x)=t0xK(x;y),Jq(x)u(y)𝑑gt0(y),\partial_{h}|_{h=0}u^{\prime\prime,\infty}(x,0)=\langle\nabla u^{\prime\prime,\infty}(x,0),J\nabla q(x)\rangle=\int_{\mathcal{R}_{t_{0}}}\langle\nabla_{x}K(x;y),J\nabla q(x)\rangle u(y)dg_{t_{0}}(y),

and the infinitesimal symmetry follows.

By Theorem 15.45(c)(c) in [Bam20b], any xt1x\in\mathcal{R}_{t_{1}} satisfies limτ02ττ2τ𝒩x(τ′′)𝑑τ′′=0\lim_{\tau^{\prime}\searrow 0}\frac{2}{\tau^{\prime}}\int_{\frac{\tau^{\prime}}{2}}^{\tau^{\prime}}\mathcal{N}_{x}(\tau^{\prime\prime})d\tau^{\prime\prime}=0. Suppose γ:It1\gamma:I^{\ast}\to\mathcal{R}_{t_{1}} is an integral curve of JqJ\nabla q, and fix τ>0\tau^{\prime}>0 sufficiently small so that if t0:=t1τt_{0}:=t_{1}-\tau^{\prime} and t0:=t112τt_{0}^{\prime}:=t_{1}-\frac{1}{2}\tau^{\prime}, then there exists x0t0x_{0}\in\mathcal{R}_{t_{0}} which exists until time t0t_{0}^{\prime}. Write K(γ(s);)=(4πτ)n2efsK(\gamma(s);\cdot)=(4\pi\tau)^{-\frac{n}{2}}e^{-f_{s}}, where fsC([t0,t0])f_{s}\in C^{\infty}(\mathcal{R}_{[t_{0},t_{0}^{\prime}]}), sIs\in I^{\ast}. By Theorem 14.54(b)(b) of [Bam20b], the completeness of JqJ\nabla q will follow from showing the following identity:

ddst0t0tfsefs𝑑gt𝑑t=0.\frac{d}{ds}\int_{t_{0}}^{t_{0}^{\prime}}\int_{\mathcal{R}_{t}}f_{s}e^{-f_{s}}dg_{t}dt=0.

For r>0r>0, let ηrC()\eta_{r}\in C^{\infty}(\mathcal{R}) be the cutoff functions from Lemma 2.15. Fix δ>0\delta>0 and a cutoff η¯δC([0,))\overline{\eta}_{\delta}\in C^{\infty}([0,\infty)) with η¯δ|[0,δ]0\overline{\eta}_{\delta}|[0,\delta]\equiv 0 and η¯δ(a)=a\overline{\eta}_{\delta}(a)=a for all a[2δ,)a\in[2\delta,\infty).

Claim: There exists A=A(δ)<A=A(\delta)<\infty such that supp(η¯δefs)P(x,A,A2)\text{supp}(\overline{\eta}_{\delta}\circ e^{-f_{s}})\subseteq P^{\ast}(x_{\infty},A,-A^{2}) for all sI[σ1,σ1]s\in I^{\ast}\cap[-\sigma^{-1},\sigma^{-1}].

We recall the following Gaussian estimate for the conjugate heat kernel on 𝒳\mathcal{X} (Lemma 15.9 of [Bam20b]):

ef(y)C(T)exp((dW1𝒳𝔱(y)(νx;t0,δy))210𝔱(y))e^{-f(y)}\leq C(T)\exp\left(-\frac{\left(d_{W_{1}}^{\mathcal{X}_{\mathfrak{t}(y)}}(\nu_{x_{\infty};t_{0}},\delta_{y})\right)^{2}}{10\mathfrak{t}(y)}\right)

for all y[T,0)y\in\mathcal{R}_{[-T,0)}. We let y=γ(s)y=\gamma(s), and observe that sf(γ(s))s\mapsto f(\gamma(s)) is constant, so there exists Λ(|t0|12,)\Lambda\in(|t_{0}|^{\frac{1}{2}},\infty) such that

dW1𝒳t1(νx;t0,δγ(s))Λd_{W_{1}}^{\mathcal{X}_{t_{1}}}(\nu_{x_{\infty};t_{0}},\delta_{\gamma(s)})\leq\Lambda

for all sIs\in I^{\ast}. This implies γ(I)P(x,Λ,Λ2)\gamma(I^{\ast})\subseteq P^{\ast}(x_{\infty},\Lambda,-\Lambda^{2}). We may therefore apply Lemma 15.9 of [Bam20b] to conclude

fs(y)C+110τ(dW1𝒳t(νγ(s);t,δy))2f_{s}(y)\geq-C+\frac{1}{10\tau^{\prime}}\left(d_{W_{1}}^{\mathcal{X}_{t}}(\nu_{\gamma(s);t},\delta_{y})\right)^{2}

for any t[t0,t0]t\in[t_{0},t_{0}^{\prime}] and yty\in\mathcal{R}_{t}. Thus, there exists A=A(δ)<A^{\prime}=A^{\prime}(\delta)<\infty such that

supp(η¯δefs)P(γ(s),A,(A)2)\text{supp}(\overline{\eta}_{\delta}\circ e^{-f_{s}})\subseteq P^{\ast}(\gamma(s),A^{\prime},-(A^{\prime})^{2})

for all sI[σ1,σ1]s\in I^{\ast}\cap[-\sigma^{-1},\sigma^{-1}]. The Claim then follows from Proposition 3.40 of [Bam20], which describes inclusion properties of PP^{\ast}-parabolic neighborhoods. \square

By the Claim and Lemma 2.11(iv)(iv), we see that

s[σ1,σ1]supp((η¯δefs)ηr)[t0,t0]\bigcup_{s\in[-\sigma^{-1},\sigma^{-1}]}\text{supp}((\overline{\eta}_{\delta}\circ e^{-f_{s}})\eta_{r})\cap\mathcal{R}_{[t_{0},t_{0}^{\prime}]}

is relatively compact in [t0,t0]\mathcal{R}_{[t_{0},t_{0}^{\prime}]} for any fixed δ,r>0\delta,r>0. Thus, for any s1,s2Is_{1},s_{2}\in I^{\ast}, the infinitesimal symmetry of KK gives

t0t0tfs(η¯δefs)ηr𝑑gt𝑑t|s=s1s=s2=\displaystyle\left.\int_{t_{0}}^{t_{0}^{\prime}}\int_{\mathcal{R}_{t}}f_{s}(\overline{\eta}_{\delta}\circ e^{-f_{s}})\eta_{r}dg_{t}dt\right|_{s=s_{1}}^{s=s_{2}}= s1s2t0t0tJq,(fs(η¯δefs))ηr𝑑gt𝑑t𝑑s\displaystyle-\int_{s_{1}}^{s_{2}}\int_{t_{0}}^{t_{0}^{\prime}}\int_{\mathcal{R}_{t}}\left\langle J\nabla q,\nabla\left(f_{s}(\overline{\eta}_{\delta}\circ e^{-f_{s}})\right)\right\rangle\eta_{r}dg_{t}dtds
=\displaystyle= s1s2t0t0t(Jq,ηr+div(Jq)ηr)fs(η¯δefs)𝑑gt𝑑t𝑑s\displaystyle\int_{s_{1}}^{s_{2}}\int_{t_{0}}^{t_{0}^{\prime}}\int_{\mathcal{R}_{t}}\left(\left\langle J\nabla q,\nabla\eta_{r}\right\rangle+\text{div}(J\nabla q)\eta_{r}\right)f_{s}(\overline{\eta}_{\delta}\circ e^{-f_{s}})dg_{t}dtds
(6.1) =\displaystyle= s1s2t0t0tJq,ηrfs(η¯δefs)𝑑gt𝑑t𝑑s.\displaystyle\int_{s_{1}}^{s_{2}}\int_{t_{0}}^{t_{0}^{\prime}}\int_{\mathcal{R}_{t}}\left\langle J\nabla q,\nabla\eta_{r}\right\rangle f_{s}(\overline{\eta}_{\delta}\circ e^{-f_{s}})dg_{t}dtds.

We recall that fsf_{s} is bounded uniformly (in ss) from below on [t0,t0]\mathcal{R}_{[t_{0},t_{0}^{\prime}]}, so fs(η¯δefs)f_{s}(\overline{\eta}_{\delta}\circ e^{-f_{s}}) is uniformly bounded on [t0,t0]\mathcal{R}_{[t_{0},t_{0}^{\prime}]}. We note that

s[σ1,σ1]supp(η¯δefs)[t0,t0]L\bigcup_{s\in[-\sigma^{-1},\sigma^{-1}]}\text{supp}\left(\overline{\eta}_{\delta}\circ e^{-f_{s}}\right)\cap\mathcal{R}_{[t_{0},t_{0}^{\prime}]}\subseteq L

for any fixed δ>0\delta>0, where L𝒳[t0,t0]L\subseteq\mathcal{X}_{[t_{0},t_{0}^{\prime}]} is compact. Integrating the estimate |fW|12τ|\nabla\sqrt{f-W}|\leq\frac{1}{2\sqrt{\tau}} along almost-minimizing curves in t\mathcal{R}_{t} we obtain supL|f|<\sup_{L}|f|<\infty, and so supL|f|<\sup_{L}|\nabla f|<\infty. Thus, we can bound the right hand side of (6.1) by

Ct0t0Lt|f||ηr|𝑑gt𝑑tCt0t0Lt|ηr|𝑑gt𝑑tCr2,C\int_{t_{0}}^{t_{0}^{\prime}}\int_{L\cap\mathcal{R}_{t}}|\nabla f|\cdot|\nabla\eta_{r}|dg_{t}dt\leq C\int_{t_{0}}^{t_{0}^{\prime}}\int_{L\cap\mathcal{R}_{t}}|\nabla\eta_{r}|dg_{t}dt\leq Cr^{2},

where C<C<\infty is independent of r>0r>0, and the last inequality follows from the estimate for {rRm<r}\{r_{Rm}<r\} in Lemma 15.27 of [Bam20b]. We can therefore take r0r\to 0 to obtain

t0t0tfs1(η¯δefs1)𝑑gt𝑑t=t0t0tfs2(η¯δefs2)𝑑gt𝑑t.\int_{t_{0}}^{t_{0}^{\prime}}\int_{\mathcal{R}_{t}}f_{s_{1}}(\overline{\eta}_{\delta}\circ e^{-f_{s_{1}}})dg_{t}dt=\int_{t_{0}}^{t_{0}^{\prime}}\int_{\mathcal{R}_{t}}f_{s_{2}}(\overline{\eta}_{\delta}\circ e^{-f_{s_{2}}})dg_{t}dt.

Finally, we take δ0\delta\searrow 0 and appeal to the dominated convergence theorem to get the desired identity.

Now let (ϕs)s(\phi_{s})_{s\in\mathbb{R}} be the flow on X\mathcal{R}_{X} generated by JqJ\nabla q (restricted to a time slice of \mathcal{R}). For any x1,x2Xx_{1},x_{2}\in X, if ϵ>0\epsilon>0 and γ:[0,1]X\gamma:[0,1]\to X is a curve with image in the regular set X\mathcal{R}_{X} of XX and length(γ)<d(x1,x2)+ϵ\text{length}(\gamma)<d(x_{1},x_{2})+\epsilon, then ϕsγ\phi_{s}\circ\gamma is a curve in X\mathcal{R}_{X} from ϕs(x1)\phi_{s}(x_{1}) to ϕs(x2)\phi_{s}(x_{2}) with length(γ)\text{length}(\gamma), so d(ϕs(x1),ϕs(x2))<d(x1,x2)+ϵd(\phi_{s}(x_{1}),\phi_{s}(x_{2}))<d(x_{1},x_{2})+\epsilon; taking ϵ0\epsilon\searrow 0, and replacing x1,x2x_{1},x_{2} with ϕs(x1),ϕs(x2)\phi_{-s}(x_{1}),\phi_{-s}(x_{2}) implies that ϕs:(X,d)(X,d)\phi_{s}:(\mathcal{R}_{X},d)\to(\mathcal{R}_{X},d) is an isometry for all ss\in\mathbb{R}. We can therefore extend to a unique isometry ϕs:(X,d)(X,d)\phi_{s}:(X,d)\to(X,d), whose image is closed and contains X\mathcal{R}_{X}, hence is bijective.

The proof strategy for the following proposition is roughly similar to that of Theorem 2 in [Liu18].

Proposition 6.3.

Let 𝒳\mathcal{X} be as in Proposition 6.1, and assume Rc(gX)=0Rc(g_{X})=0, so that X=C(Y)X=C(Y) is a metric cone with vertex {o}\{o\}. Then the 1-parameter group of isometries (ϕs)s(\phi_{s})_{s\in\mathbb{R}} of C(Y)C(Y) acts locally freely on the link YY.

Remark 6.4.

The rough idea to assume by way of contradiction that a point zC(Y){o}z\in C(Y)\setminus\{o\} is fixed by the action (ϕs)(\phi_{s}), so that ϕs\phi_{s} preserves the distance to zz. Let qiq_{i} be a sequence of almost-radial functions based at (xi,0)(x_{i},0), and let (zi,1)Mi×[ϵi1,0](z_{i},-1)\in M_{i}\times[\epsilon_{i}^{-1},0] converge to (z,1)(z,-1). At sufficiently small scales near (zi,1)(z_{i},-1), appropriate rescalings of qiq_{i} look like almost-splitting functions, so Proposition 5.3 gives almost-splitting functions yiy_{i} with yiJqi\nabla y_{i}\approx J\nabla q_{i}. By a diagonal argument, after parabolic rescaling of flows, we get convergence of yiy_{i} to a function yy_{\infty} on the tangent cone C(Z)C(Z) at zz which induces a metric splitting. On the other hand, yiJqi\nabla y_{i}\approx J\nabla q_{i} implies that the flow of y\nabla y_{\infty} preserves the distance to the vertex of C(Z)C(Z), a contradiction.

Proof.

Fix a correspondence \mathfrak{C} realized the 𝔽\mathbb{F}-convergence to 𝒳\mathcal{X}. It suffices to show that there is no point zB(o,1)z\in\partial B(o,1) satisfying ϕs(z)=z\phi_{s}(z)=z for all ss\in\mathbb{R}. Suppose by way of contradiction such a point exists. For any xC(Y)x\in\mathcal{R}_{C(Y)}, we then have d(ϕs(x),z)=d(x,z)d(\phi_{s}(x),z)=d(x,z) for all ss\in\mathbb{R}. Choose a sequence ziMiz_{i}\in M_{i} such that

(zi,1)i(z,1)C(Y)×(,0)=𝒳<0.(z_{i},-1)\xrightarrow[i\to\infty]{\mathfrak{C}}(z,-1)\in C(Y)\times(-\infty,0)=\mathcal{X}_{<0}.

By Proposition 4.4, there is a sequence δi0\delta_{i}\searrow 0 such that if Wi:=𝒩xi,0(1)W_{i}:=\mathcal{N}_{x_{i},0}(1), then qi:=4τ(hiWi)q_{i}:=4\tau(h_{i}-W_{i}) are strong (δi,1)(\delta_{i},1)-conical functions based at (xi,0)(x_{i},0) which satisfy

(6.2) δi1δiMi(|2qi2gi|2+||qi|24qi|)eαfi𝑑νxi,0;ti𝑑tδi\int_{-\delta_{i}^{-1}}^{-\delta_{i}}\int_{M_{i}}\left(\left|\nabla^{2}q_{i}-2g_{i}\right|^{2}+\left||\nabla q_{i}|^{2}-4q_{i}\right|\right)e^{\alpha f_{i}}d\nu_{x_{i},0;t}^{i}dt\leq\delta_{i}

for some α>0\alpha>0, where we have written νxi,0;ti=(4πτ)n2efidgi,t\nu_{x_{i},0;t}^{i}=(4\pi\tau)^{-\frac{n}{2}}e^{-f_{i}}dg_{i,t}. By the proof of Theorem 15.80 in [Bam20b], we can therefore pass to a subsequence so that qiψiq:=d2(,o)q_{i}\circ\psi_{i}\to q_{\infty}:=d^{2}(\cdot,o) in Cloc(<0)C_{loc}^{\infty}(\mathcal{R}_{<0}) as ii\to\infty, where ψi\psi_{i} are as in Theorem 2.5. This implies

lim infi1τ12τ1τMi(qi)+𝑑νzi,1;ti𝑑t1τ12τ1τC(Y)q𝑑νz,1;t𝑑t.\liminf_{i\to\infty}\frac{1}{\tau^{\ast}}\int_{-1-2\tau^{\ast}}^{-1-\tau^{\ast}}\int_{M_{i}}(q_{i})_{+}d\nu_{z_{i},-1;t}^{i}dt\geq\frac{1}{\tau^{\ast}}\int_{-1-2\tau^{\ast}}^{-1-\tau^{\ast}}\int_{\mathcal{R}_{C(Y)}}q_{\infty}d\nu_{z,-1;t}dt.

Claim: limt1tq𝑑νz,1;t=q(z)=1\lim_{t\nearrow-1}\int_{\mathcal{R}_{t}}q_{\infty}d\nu_{z,-1;t}=q_{\infty}(z)=1.

Now choose sequences tj1t_{j}\nearrow-1, yjXy_{j}\in X such that (yj,tj)(y_{j},t_{j}) are HnH_{n}-centers of (z,1)(z,-1). Because min(q,4)\min(q_{\infty},4) is 2-Lipschitz, then have

|tmin(q,4)𝑑νz,1;tjmin(q(yj),4)|2Hn(1+tj)0\left|\int_{\mathcal{R}_{t}}\min(q_{\infty},4)d\nu_{z,-1;t_{j}}-\min(q_{\infty}(y_{j}),4)\right|\leq 2\sqrt{H_{n}(1+t_{j})}\to 0

as jj\to\infty. However, Claim 22.9(d) of [Bam20b] implies that the natural topology agrees on 𝒳\mathcal{X} agrees with the product topology on C(Y)×(,0)C(Y)\times(-\infty,0); because (yj,tj)(z,1)(y_{j},t_{j})\to(z,-1) in the natural topology, we have yjzy_{j}\to z in C(Y)C(Y), hence

limjq(yj)=q(z)=1.\lim_{j\to\infty}q_{\infty}(y_{j})=q_{\infty}(z)=1.\hskip 56.9055pt\square

We can therefore find γ>0\gamma>0 such that, for any τ(0,1)\tau^{\ast}\in(0,1), we have

1τ12τ1τMi(qi)+𝑑νzi,1;ti𝑑tγ2\frac{1}{\tau^{\ast}}\int_{-1-2\tau^{\ast}}^{-1-\tau^{\ast}}\int_{M_{i}}(q_{i})_{+}d\nu_{z_{i},-1;t}^{i}dt\geq\gamma^{2}

for i=i(τ,γ)i=i(\tau^{\ast},\gamma)\in\mathbb{N} sufficiently large. Because (zi,1)i(z,1)(z_{i},-1)\xrightarrow[i\to\infty]{\mathfrak{C}}(z,-1), there exists A<A<\infty such that (zi,1)P(x;A,A2)(z_{i},-1)\in P^{\ast}(x_{\infty};A,-A^{2}) for all ii\in\mathbb{N}. We can therefore use Bamler’s conjugate heat kernel comparison theorem (Proposition 8.1 in [Bam20b]) and (6.2) to obtain

1(δi)11δiMi(|2qi2gi|2+||qi|24qi|)𝑑νzi,1;ti𝑑tδi\int_{-1-(\delta_{i}^{\prime})^{-1}}^{-1-\delta_{i}^{\prime}}\int_{M_{i}}\left(|\nabla^{2}q_{i}-2g_{i}|^{2}+\left||\nabla q_{i}|^{2}-4q_{i}\right|\right)d\nu_{z_{i},-1;t}^{i}dt\leq\delta_{i}^{\prime}

for some sequence δi0\delta_{i}^{\prime}\searrow 0. We may then proceed as in the proof of Proposition 13.19 of [Bam20b] to conclude that, for any ϵ>0\epsilon>0,

12aiqi:Mi×[1(γβ)2ϵ1,1(γβ)2ϵ]\frac{1}{2\sqrt{a_{i}}}q_{i}:M_{i}\times[-1-(\gamma\beta)^{2}\epsilon^{-1},-1-(\gamma\beta)^{2}\epsilon]\to\mathbb{R}

are (1,ϵ,γβ)(1,\epsilon,\gamma\beta)-splitting maps for ββ¯(ϵ)\beta\leq\overline{\beta}(\epsilon), where

ai:=Miqi𝑑νzi,1;t1(βγ)2ia_{i}:=\int_{M_{i}}q_{i}d\nu_{z_{i},-1;t_{1}-(\beta\gamma)^{2}}^{i}

satisfies

12γ2aiCMiqieαfi𝑑νtiC(Miqi2𝑑νti)12(Mie2αfi𝑑νti)12C(Y,ϵ,β,γ).\frac{1}{2}\gamma^{2}\leq a_{i}\leq C\int_{M_{i}}q_{i}e^{\alpha f_{i}}d\nu_{t}^{i}\leq C\left(\int_{M_{i}}q_{i}^{2}d\nu_{t}^{i}\right)^{\frac{1}{2}}\left(\int_{M_{i}}e^{2\alpha f_{i}}d\nu_{t}^{i}\right)^{\frac{1}{2}}\leq C(Y,\epsilon,\beta,\gamma).

In fact, the lower bound follows from the estimate for |qi||\square q_{i}|, (c.f. the proof of Proposition 12.1 of [Bam20b]), while the upper bound follows from the L2L^{2}-Poincare inequality and and property (iv)(iv) of strong almost-radial functions.

We now apply Proposition 12.1 of [Bam20b] to obtain strong (1,ϵ′′,γβ)(1,\epsilon^{\prime\prime},\gamma\beta)-splitting maps yi′′y_{i}^{\prime\prime} with

β21(βγ)2ϵ′′11(βγ)2ϵ′′Mi|(qi2aiyi′′)|2𝑑νzi,1;ti𝑑tΨ(β|Y,ϵ′′)\beta^{-2}\int_{-1-(\beta\gamma)^{2}\epsilon^{\prime\prime-1}}^{-1-(\beta\gamma)^{2}\epsilon^{\prime\prime}}\int_{M_{i}}\left|\nabla\left(\frac{q_{i}}{2\sqrt{a_{i}}}-y_{i}^{\prime\prime}\right)\right|^{2}d\nu_{z_{i},-1;t}^{i}dt\leq\Psi(\beta|Y,\epsilon^{\prime\prime})

for sufficiently large ii\in\mathbb{N}. Next, apply Proposition 5.3 to obtain a weak (1,ϵ,γβ)(1,\epsilon^{\prime},\gamma\beta)-splitting map yiy_{i}^{\prime} satisfying

β21(βγ)2ϵ11(βγ)2ϵMi|Jqi2aiyi|2𝑑νzi,1;ti𝑑tΨ(β|Y,ϵ)\beta^{-2}\int_{-1-(\beta\gamma)^{2}\epsilon^{\prime-1}}^{-1-(\beta\gamma)^{2}\epsilon^{\prime}}\int_{M_{i}}\left|\frac{J\nabla q_{i}}{2\sqrt{a_{i}}}-\nabla y_{i}^{\prime}\right|^{2}d\nu_{z_{i},-1;t}^{i}dt\leq\Psi(\beta|Y,\epsilon^{\prime})

for sufficiently large ii\in\mathbb{N}, assuming ϵ′′ϵ¯(ϵ′′)\epsilon^{\prime\prime}\leq\overline{\epsilon}^{\prime}(\epsilon^{\prime\prime}). Another application of Proposition 12.1 of [Bam20b] yields strong (1,ϵ,γβ)(1,\epsilon,\gamma\beta)-splitting maps yiβ:Mi×[1β2ϵ1,1β2ϵ]y_{i}^{\beta}:M_{i}\times[-1-\beta^{2}\epsilon^{-1},-1-\beta^{2}\epsilon]\to\mathbb{R} satisfying

β21(βγ)2ϵ11(βγ)2ϵMi|Jqi2aiyiβ|2𝑑νzi,1;ti𝑑tΨ(β|Y,ϵ)\beta^{-2}\int_{-1-(\beta\gamma)^{2}\epsilon^{-1}}^{-1-(\beta\gamma)^{2}\epsilon}\int_{M_{i}}\left|\frac{J\nabla q_{i}}{2\sqrt{a_{i}}}-\nabla y_{i}^{\beta}\right|^{2}d\nu_{z_{i},-1;t}^{i}dt\leq\Psi(\beta|Y,\epsilon)

for large i=i(β)i=i(\beta)\in\mathbb{N}, assuming ϵϵ¯(ϵ)\epsilon^{\prime}\leq\overline{\epsilon}^{\prime}(\epsilon). We also pass to a subsequence so that aia(0,)a_{i}\to a\in(0,\infty). Then

(ψi1)(Jiqi2ai)V(\psi_{i}^{-1})_{\ast}\left(J_{i}\frac{\nabla q_{i}}{2\sqrt{a_{i}}}\right)\to V

in Cloc(C(Y))C_{loc}^{\infty}(\mathcal{R}_{C(Y)}), where V:=12aJqV:=\frac{1}{2\sqrt{a}}J\nabla q_{\infty}.

Using Theorem 2.7, choose a sequence βj0\beta_{j}\searrow 0 such that we have 𝔽\mathbb{F}-convergence of the corresponding parabolic rescalings to a tangent flow of 𝒳\mathcal{X} based at (z,1)(z,-1):

(𝒳1,βj1,(ν(z,0);t1,βj1)t[2,0])i𝔽(𝒴,(νy;t)t[2,0]),\left(\mathcal{X}^{-1,\beta_{j}^{-1}},(\nu_{(z,0);t}^{-1,\beta_{j}^{-1}})_{t\in[-2,0]}\right)\xrightarrow[i\to\infty]{\mathbb{F}}\left(\mathcal{Y},(\nu_{y_{\infty};t})_{t\in[-2,0]}\right),

where 𝒴\mathcal{Y} is a static metric flow modeled on a a Ricci flat cone C(Z)C(Z). By Theorem 2.16 of [Bam20b], there is a precompact exhaustion (Wj)(W_{j}) of C(Z)\mathcal{R}_{C(Z)} along with diffeomorphisms ηj:WjC(Y)\eta_{j}:W_{j}\to\mathcal{R}_{C(Y)} such that ηj(βj2gC(Y))gC(Z)\eta_{j}^{\ast}(\beta_{j}^{-2}g_{C(Y)})\to g_{C(Z)} in Cloc(C(Z))C_{loc}^{\infty}(\mathcal{R}_{C(Z)}), and so that for any ϵ>0\epsilon>0 and D<D<\infty,

ηj:(WjB(oZ,D),dC(Z))(C(Y),βj1dC(Y))\eta_{j}:\left(W_{j}\cap B(o_{Z},D),d_{C(Z)}\right)\to\left(C(Y),\beta_{j}^{-1}d_{C(Y)}\right)

are ϵ\epsilon-Gromov-Hausdorff maps to B(z,βjD)B(z,\beta_{j}D) for sufficiently large j=j(ϵ,D)j=j(\epsilon,D)\in\mathbb{N}. Define g~j:=βj2gC(Y)\widetilde{g}_{j}:=\beta_{j}^{-2}g_{C(Y)} and Vj:=βjVV_{j}:=\beta_{j}V, so that |Vj|g~j10|V_{j}|_{\widetilde{g}_{j}}\leq 10 on C(Y)B(o,10)\mathcal{R}_{C(Y)}\cap B(o,10); by ΔVj=0\Delta V_{j}=0 and elliptic regularity, we can pass to a subsequence so that (ηj1)VjV(\eta_{j}^{-1})_{\ast}V_{j}\to V_{\infty} in Cloc(C(Z))C_{loc}^{\infty}(\mathcal{R}_{C(Z)}). Moreover, if we define rj:=βj1d(z,)r_{j}:=\beta_{j}^{-1}d(z,\cdot), then rjηjr:=d(oZ,)r_{j}\circ\eta_{j}\to r_{\infty}:=d(o_{Z},\cdot) locally uniformly on C(Z)\mathcal{R}_{C(Z)}; this follows from the Gromov-Hausdorff convergence (X,βj1d,z)(C(Z),dC(Z),oZ)(X,\beta_{j}^{-1}d,z)\to(C(Z),d_{C(Z)},o_{Z}) (Theorem 2.16 of [Bam20b]). Because rjr_{j} are Lipschitz, we know VjrjV_{j}r_{j} is well-defined almost everywhere on C(Y)\mathcal{R}_{C(Y)}. Because rjϕs=rjr_{j}\circ\phi_{s}=r_{j}, we may then conclude Vjrj=0V_{j}r_{j}=0 almost everywhere. Given χCc(C(Z))\chi\in C_{c}^{\infty}(\mathcal{R}_{C(Z)}), we have

C(Z)ϕVr𝑑gC(Z)\displaystyle\int_{\mathcal{R}_{C(Z)}}\phi V_{\infty}r_{\infty}dg_{C(Z)} =C(Z)rdiv(ϕV)𝑑gC(Z)=limjC(Y)rjdiv((ϕηj1)Vj)𝑑gC(Y)\displaystyle=\int_{\mathcal{R}_{C(Z)}}r_{\infty}\text{div}(\phi V_{\infty})dg_{C(Z)}=\lim_{j\to\infty}\int_{\mathcal{R}_{C(Y)}}r_{j}\text{div}\left((\phi\circ\eta_{j}^{-1})V_{j}\right)dg_{C(Y)}
=limjC(Y)(ϕηj1)Vjrj𝑑gC(Y)=0\displaystyle=\lim_{j\to\infty}\int_{\mathcal{R}_{C(Y)}}(\phi\circ\eta_{j}^{-1})V_{j}r_{j}dg_{C(Y)}=0

since rjηjrr_{j}\circ\eta_{j}\to r_{\infty} uniformly and (ηj1)VjV(\eta_{j}^{-1})_{\ast}V_{j}\to V_{\infty} in ClocC_{loc}^{\infty}. Thus Vr=0V_{\infty}r_{\infty}=0 almost everywhere in C(Z)\mathcal{R}_{C(Z)}, so the flow of VV_{\infty} preserves rr_{\infty}.

By Bamler’s change of basepoint theorem (Theorem 6.40 in [Bam20]), we have

(Mi,(gi,t)t[δi1,1],(νzi,1;t)t[δi1,1])i𝔽,(𝒳,(νz,1;t)t(,1])(M_{i},(g_{i,t})_{t\in[-\delta_{i}^{-1},-1]},(\nu_{z_{i},-1;t})_{t\in[-\delta_{i}^{-1},-1]})\xrightarrow[i\to\infty]{\mathbb{F},\mathfrak{C}}(\mathcal{X},(\nu_{z,-1;t})_{t\in(-\infty,-1]})

on compact time intervals. For each jj\in\mathbb{N}, we can therefore choose i(j)i(j)\in\mathbb{N} such that ηj(Wj)Ui(j)\eta_{j}(W_{j})\subseteq U_{i(j)},

ψi(j)gi(j)gC(Y)Cj(Ui(j)×[2,1βj4],gC(Y))+ψi(j)qi(j)qCj(Ui(j)×[2,1βj4],gC(Y))βj4,||\psi_{i(j)}^{\ast}g_{i(j)}-g_{C(Y)}||_{C^{j}(U_{i(j)}\times[-2,-1-\beta_{j}^{4}],g_{C(Y)})}+||\psi_{i(j)}^{\ast}q_{i(j)}-q_{\infty}||_{C^{j}(U_{i(j)}\times[-2,-1-\beta_{j}^{4}],g_{C(Y)})}\leq\beta_{j}^{4},
d𝔽((Mi(j),(gi(j),t)t[βj4,0],(νzi(j),1;t)t[βj4,1]),(𝒳[βj4,1],(νz,1;t)t[βj4,1]))<βj4,d_{\mathbb{F}}\left((M_{i(j)},(g_{i(j),t})_{t\in[-\beta_{j}^{-4},0]},(\nu_{z_{i(j)},-1;t})_{t\in[-\beta_{j}^{-4},-1]}),(\mathcal{X}_{[\beta_{j}^{-4},-1]},(\nu_{z,-1;t})_{t\in[-\beta_{j}^{-4},-1]})\right)<\beta_{j}^{4},
1ϵj1βj21ϵjβj2Mi(j)|Jqi(j)2ai(j)yi(j)βj|2𝑑νzi(j),1;t𝑑tϵjβj2\int_{-1-\epsilon_{j}^{-1}\beta_{j}^{2}}^{-1-\epsilon_{j}\beta_{j}^{2}}\int_{M_{i(j)}}\left|\frac{J\nabla q_{i(j)}}{2\sqrt{a_{i(j)}}}-\nabla y_{i(j)}^{\beta_{j}}\right|^{2}d\nu_{z_{i(j)},-1;t}dt\leq\epsilon_{j}\beta_{j}^{2}

for some sequence ϵj0\epsilon_{j}\searrow 0, where we now view ηj\eta_{j} as maps Wj×[2,1]C(Y)×[2,1]W_{j}\times[-2,-1]\to\mathcal{R}_{C(Y)}\times[-2,-1] which are constant in time. . Define parabolic rescalings g^j,t:=βj2gi(j),1+βj2t\widehat{g}_{j,t}:=\beta_{j}^{-2}g_{i(j),-1+\beta_{j}^{2}t}, ν^tj:=νzi(j),1;1+βj2t\widehat{\nu}_{t}^{j}:=\nu_{z_{i(j)},-1;-1+\beta_{j}^{2}t}, (1,ϵj,1)(1,\epsilon_{j},1)-splitting maps

y^j(,t):=βj1yi(j)βj(,1+βj2t)\widehat{y}_{j}(\cdot,t):=\beta_{j}^{-1}y_{i(j)}^{\beta_{j}}(\cdot,-1+\beta_{j}^{2}t)

based at (zi(j),0)(z_{i(j)},0) in the rescaled flow, a^j:=βj2ai(j)\widehat{a}_{j}:=\beta_{j}^{-2}a_{i(j)}, and

q^j(,t):=βj2qi(j)(,1+βj2t),\widehat{q}_{j}(\cdot,t):=\beta_{j}^{-2}q_{i(j)}(\cdot,-1+\beta_{j}^{2}t),

Then

(Mi(j),(g^j,t)t[2,0],(ν^tj)t[2,0])i𝔽(𝒴,(νoZ,0;t)t[2,0]),(M_{i(j)},(\widehat{g}_{j,t})_{t\in[-2,0]},(\widehat{\nu}_{t}^{j})_{t\in[-2,0]})\xrightarrow[i\to\infty]{\mathbb{F}}(\mathcal{Y},(\nu_{o_{Z},0;t})_{t\in[-2,0]}),

and ψi(j)ηj\psi_{i(j)}\circ\eta_{j} realizes smooth convergence on C(Z)\mathcal{R}_{C(Z)}; for example, (ψi(j)ηj)g^jgC(Z)(\psi_{i(j)}\circ\eta_{j})^{\ast}\widehat{g}_{j}\to g_{C(Z)} in Cloc(C(Z))C_{loc}^{\infty}(\mathcal{R}_{C(Z)}). so the proof of Theorem 15.50 in [Bam20b] shows that 𝒴=𝒴×\mathcal{Y}=\mathcal{Y}^{\prime}\times\mathbb{R} splits as a metric flow, and (ψi(j)ηj)y^jy(\psi_{i(j)}\circ\eta_{j})^{\ast}\widehat{y}_{j}\to y_{\infty}, where y:𝒴y_{\infty}:\mathcal{Y}\to\mathbb{R} denotes the projection onto the \mathbb{R}-factor. On the other hand, our assumptions guarantee that

12a^j(ψi(j)1)g^jq^j(,t)VjCj1(Ui(j),g~j)Cβj2,\left\|\frac{1}{2\sqrt{\widehat{a}_{j}}}(\psi_{i(j)}^{-1})_{\ast}\nabla^{\widehat{g}_{j}}\widehat{q}_{j}(\cdot,t)-V_{j}\right\|_{C^{j-1}(U_{i(j)},\widetilde{g}_{j})}\leq C\beta_{j}^{2},

which implies

12a^j((ψi(j)ηj)1)g^jq^j(,t)V\frac{1}{2\sqrt{\widehat{a}_{j}}}((\psi_{i(j)}\circ\eta_{j})^{-1})_{\ast}\nabla^{\widehat{g}_{j}}\widehat{q}_{j}(\cdot,t)\to V_{\infty}

in Cloc(C(Z))C_{loc}^{\infty}(\mathcal{R}_{C(Z)}). Here, we again view ηj\eta_{j} as a map WjUi(j)W_{j}\to U_{i(j)} which is constant in time. Let K^j\widehat{K}_{j} denote the heat kernel of the rescaled flows. For any compact subset KC(Z)K\subseteq\mathcal{R}_{C(Z)}, we then have

21K|Vy|2𝑑νoZ,1;t𝑑t\displaystyle\int_{-2}^{-1}\int_{K}|V_{\infty}-\nabla y_{\infty}|^{2}d\nu_{o_{Z},-1;t}dt
=limj21K|((ψi(j)ηj)1)(g^jq^j2a^jg^jy^j)|2(ψi(j)ηj)K^j(zi(j),0;,t)d((ψi(j)ηj)g^j)𝑑t\displaystyle=\lim_{j\to\infty}\int_{-2}^{-1}\int_{K}\left|\left((\psi_{i(j)}\circ\eta_{j})^{-1}\right)_{\ast}\left(\frac{\nabla^{\widehat{g}_{j}}\widehat{q}_{j}}{2\sqrt{\widehat{a}_{j}}}-\nabla^{\widehat{g}_{j}}\widehat{y}_{j}\right)\right|^{2}(\psi_{i(j)}\circ\eta_{j})^{\ast}\widehat{K}_{j}(z_{i(j)},0;\cdot,t)d\left((\psi_{i(j)}\circ\eta_{j})^{\ast}\widehat{g}_{j}\right)dt
=limj21(ψi(j)ηj)(K)|g^jq^j2a^jg^jy^j|2𝑑ν^tj𝑑t\displaystyle=\lim_{j\to\infty}\int_{-2}^{-1}\int_{(\psi_{i(j)}\circ\eta_{j})(K)}\left|\frac{\nabla^{\widehat{g}_{j}}\widehat{q}_{j}}{2\sqrt{\widehat{a}_{j}}}-\nabla^{\widehat{g}_{j}}\widehat{y}_{j}\right|^{2}d\widehat{\nu}_{t}^{j}dt
lim infjβj212βj21βj2Mi(j)|gi(j)qj2ai(j)gi(j)yi(j)|2𝑑νzi(j),1;t𝑑t=0,\displaystyle\leq\liminf_{j\to\infty}\beta_{j}^{-2}\int_{-1-2\beta_{j}^{2}}^{-1-\beta_{j}^{2}}\int_{M_{i(j)}}\left|\frac{\nabla^{g_{i(j)}}q_{j}}{2\sqrt{a_{i(j)}}}-\nabla^{g_{i(j)}}y_{i(j)}\right|^{2}d\nu_{z_{i(j)},-1;t}dt=0,

where we used that

((ψi(j)ηj)1)g^jy^j=(ψi(j)ηj)g^j(ψi(j)ηj)y^jgC(Z)y\left((\psi_{i(j)}\circ\eta_{j})^{-1}\right)_{\ast}\nabla^{\widehat{g}_{j}}\widehat{y}_{j}=\nabla^{(\psi_{i(j)}\circ\eta_{j})^{\ast}\widehat{g}_{j}}(\psi_{i(j)}\circ\eta_{j})^{\ast}\widehat{y}_{j}\to\nabla^{g_{C(Z)}}y_{\infty}

in Cloc(C(Z))C_{loc}^{\infty}(\mathcal{R}_{C(Z)}). Thus V=yV_{\infty}=\nabla y_{\infty}. However, y\nabla y_{\infty} is a complete vector field on C(Z)\mathcal{R}_{C(Z)} which leaves any compact set in finite time, whereas the flow of VV_{\infty} preserves any geodesic ball centered at oZo_{Z}, a contradiction. ∎

Proof of Theorem 1.3.

This is a consequence of Propositions 6.1 and 6.3. ∎

References

  • [Bam20] Richard H Bamler “Compactness theory of the space of Super Ricci flows”, 2020 arXiv:2008.09298 [math.DG]
  • [Bam20a] Richard H Bamler “Entropy and heat kernel bounds on a Ricci flow background”, 2020 arXiv:2008.07093 [math.DG]
  • [Bam20b] Richard H Bamler “Structure theory of non-collapsed limits of Ricci flows”, 2020 arXiv:2009.03243 [math.DG]
  • [BK18] Richard H. Bamler and Bruce Kleiner “Uniqueness and stability of Ricci flow through singularities”, 2018 arXiv:1709.04122 [math.DG]
  • [CC00] Jeff Cheeger and Tobias H. Colding “On the structure of spaces with Ricci curvature bounded below. II” In J. Differential Geom. 54.1, 2000, pp. 13–35 URL: http://projecteuclid.org.proxy.library.cornell.edu/euclid.jdg/1214342145
  • [CC96] Jeff Cheeger and Tobias H. Colding “Lower bounds on Ricci curvature and the almost rigidity of warped products” In Ann. of Math. (2) 144.1, 1996, pp. 189–237 DOI: 10.2307/2118589
  • [CC97] Jeff Cheeger and Tobias H. Colding “On the structure of spaces with Ricci curvature bounded below. I” In J. Differential Geom. 46.3, 1997, pp. 406–480 URL: http://projecteuclid.org.proxy.library.cornell.edu/euclid.jdg/1214459974
  • [CCT02] J. Cheeger, T.. Colding and G. Tian “On the singularities of spaces with bounded Ricci curvature” In Geom. Funct. Anal. 12.5, 2002, pp. 873–914 DOI: 10.1007/PL00012649
  • [CN12] Tobias Holck Colding and Aaron Naber “Sharp Hölder continuity of tangent cones for spaces with a lower Ricci curvature bound and applications” In Ann. of Math. (2) 176.2, 2012, pp. 1173–1229 DOI: 10.4007/annals.2012.176.2.10
  • [CN13] Jeff Cheeger and Aaron Naber “Lower bounds on Ricci curvature and quantitative behavior of singular sets” In Invent. Math. 191.2, 2013, pp. 321–339 DOI: 10.1007/s00222-012-0394-3
  • [FIK03] Mikhail Feldman, Tom Ilmanen and Dan Knopf “Rotationally symmetric shrinking and expanding gradient Kähler-Ricci solitons” In J. Differential Geom. 65.2, 2003, pp. 169–209 URL: http://projecteuclid.org.proxy.library.cornell.edu/euclid.jdg/1090511686
  • [Gia17] Panagiotis Gianniotis “The size of the singular set of a type I Ricci flow” In J. Geom. Anal. 27.4, 2017, pp. 3099–3119 DOI: 10.1007/s12220-017-9797-0
  • [Gia19] Panagiotis Gianniotis “Regularity theory for type I Ricci flows” In Calc. Var. Partial Differential Equations 58.6, 2019, pp. Paper No. 200\bibrangessep24 DOI: 10.1007/s00526-019-1644-7
  • [JN21] Wenshuai Jiang and Aaron Naber L2L^{2} curvature bounds on manifolds with bounded Ricci curvature” In Ann. of Math. (2) 193.1, 2021, pp. 107–222 DOI: 10.4007/annals.2021.193.1.2
  • [KL17] Bruce Kleiner and John Lott “Singular Ricci flows I” In Acta Math. 219.1, 2017, pp. 65–134 DOI: 10.4310/ACTA.2017.v219.n1.a4
  • [Liu18] Gang Liu “On the tangent cone of Kähler manifolds with Ricci curvature lower bound” In Math. Ann. 370.1-2, 2018, pp. 649–667 DOI: 10.1007/s00208-017-1536-0
  • [LS21] Gang Liu and Gábor Székelyhidi “Gromov-Hausdorff limits of Kähler manifolds with Ricci curvature bounded below II” In Comm. Pure Appl. Math. 74.5, 2021, pp. 909–931 DOI: 10.1002/cpa.21900

M. Hallgren, Department of Mathematics, Cornell University, Ithaca NY 14850
Email: meh249@cornell.edu

W. Jian, Institute of Mathematics, Academy of Mathematics and Systems Science, Chinese Academy of Sciences, Beijing, 100190, China.
Email: wangjian@amss.ac.cn