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

The blowdown of ancient noncollapsed mean curvature flows

Wenkui Du, Robert Haslhofer
Abstract.

In this paper, we consider ancient noncollapsed mean curvature flows Mt=Ktn+1M_{t}=\partial K_{t}\subset\mathbb{R}^{n+1} that do not split off a line. It follows from general theory that the blowdown of any time-slice, limλ0λKt0\lim_{\lambda\to 0}\lambda K_{t_{0}}, is at most n1n-1 dimensional. Here, we show that the blowdown is in fact at most n2n-2 dimensional. Our proof is based on fine cylindrical analysis, which generalizes the fine neck analysis that played a key role in many recent papers. Moreover, we show that in the uniformly kk-convex case, the blowdown is at most k2k-2 dimensional. This generalizes the recent results from Choi-Haslhofer-Hershkovits [CHH21a] to higher dimensions, and also has some applications towards the classification problem for singularities in 3-convex mean curvature flow.

1. Introduction

In the analysis of mean curvature flow it is crucial to understand ancient noncollapsed flows. We recall that a mean curvature flow MtM_{t} is called ancient if it is defined for all t0t\ll 0, and noncollapsed if it is mean-convex and there is an α>0\alpha>0 so that every point pMtp\in M_{t} admits interior and exterior balls of radius at least α/H(p)\alpha/H(p), c.f. [SW09, And12, HK17] (in fact, by [Bre15, HK15] one can always take α=1\alpha=1). By the work of White [Whi00, Whi03, Whi15] and by [HH18] it is known that all blowup limits of mean-convex mean curvature flow are ancient noncollapsed flows. More generally, by Ilmanen’s mean-convex neighborhood conjecture [Ilm03], which has been proved recently in the case of neck-singularities in [CHH18, CHHW19], it is expected even without mean-convexity assumption that all blowup limits near any cylindrical singularity are ancient noncollapsed flows.

In this paper, we consider the asymptotic structure of ancient noncollapsed flows, specifically the blowdown of any time-slice. To describe this, denote by Mt=Ktn+1M_{t}=\partial K_{t}\subset\mathbb{R}^{n+1} any ancient noncollapsed mean curvature flow. Recall that such flows are always convex thanks to [HK17, Theorem 1.10]. We can assume without essential loss of generality that the solution does not split off a line (otherwise we can reduce the problem to a lower dimensional one). Equivalently, this means that MtM_{t} is strictly convex.

Definition 1.1 (blowdown).

Given any time t0t_{0}, the blowdown of Kt0K_{t_{0}} is defined by

(1.1) Kˇt0:=limλ0λKt0.\check{K}_{t_{0}}:=\lim_{\lambda\to 0}\lambda\cdot K_{t_{0}}.

It is easy to see that the blowdown exists (in fact it is given as a decreasing intersection of convex sets) and is a cone, i.e. satsifies λKˇt0=Kˇt0\lambda\check{K}_{t_{0}}=\check{K}_{t_{0}} for all λ>0\lambda>0. The blowdown captures important information about the asymptotic structure of the time slices. By general theory the blowdown is at most n1n-1 dimensional. Indeed, the blowdown cannot be bigger than the singular set in a mean convex flow, which is at most n1n-1 dimensional by the deep theory of White [Whi00, Whi03] (see also [HK17]). Here, we improve this to:

Theorem 1.2 (blowdown of ancient noncollapsed flows).

Let Mt=Ktn+1M_{t}=\partial K_{t}\subset\mathbb{R}^{n+1}, where n3n\geq 3, be an ancient noncollapsed mean curvature flow that does not split off a line. Then the blowdown of any time slice is at most n2n-2 dimensional, namely

(1.2) dimKˇt0n2.\dim\check{K}_{t_{0}}\leq n-2.

This generalizes the main result from [CHH21a], where Choi, Hershkovits and the second author ruled out the potential scenario ancient wing-like noncollapsed flows in 4\mathbb{R}^{4}. Our theorem shows that the blowdown is always at least 11 dimension smaller than what one gets from the general bound for the dimension of the singular set. In particular, it implies:

Corollary 1.3 (nonexistence of ancient asymptotically conical flows).

There does not exist any ancient noncollapsed solution of the mean curvature flow in n+1\mathbb{R}^{n+1} that is asymptotic to an (n1)(n-1)-dimensional cone.

For comparison, we recall the basic result that there does not exist any nontrivial ancient noncollapsed flow in n+1\mathbb{R}^{n+1} that is asymptotic to an nn-dimensional cone, as follows easily from the fact that the asymptotic area ratio vanishes [HK17]. Our corollary rules out solutions that are asymptotic to n1n-1 dimension cones, i.e. solutions asymptotic to cones whose dimension is smaller by 11 than the dimension of the evolving hypersurfaces, which is much more subtle. It even rules out flows asymptotic to singular cones.

The dimension bound can be improved further in the uniformly kk-convex case. Recall that an ancient noncollapsed flow is uniformly kk-convex if there is a constant β>0\beta>0 such that λ1++λkβH\lambda_{1}+\ldots+\lambda_{k}\geq\beta H, where λ1λn\lambda_{1}\leq\ldots\leq\lambda_{n} denote the principal curvatures. We prove:

Theorem 1.4 (blowdown of ancient kk-convex flows).

If we assume in addition the flow is uniformly kk-convex, where k3k\geq 3, then the blowdown of any time slice is at most k2k-2 dimensional.

Setting k=3k=3 and relaxing the strict convexity assumption we obtain:

Corollary 1.5 (blowdown of ancient 3-convex flows).

Let Mt=KtM_{t}=\partial K_{t} be an ancient noncollapsed uniformly 3-convex mean curvature flow in n+1\mathbb{R}^{n+1}. Then its blowdown is

  • either a point (which only happens if the solution is compact),

  • or a halfline,

  • or a line (which only happens for a round shrinking Sn1×S^{n-1}\times\mathbb{R} or an (n1)(n-1)-dimensional rotationally symmetric oval times \mathbb{R}),

  • or a 22-dimensional halfplane (which only happens if the solution is a (n1)(n-1)-dimensional rotationally symmetric bowl soliton times \mathbb{R}),

  • or a 22-dimensional plane (which only happens for a round shrinking Sn2×2S^{n-2}\times\mathbb{R}^{2}),

  • or an nn-dimensional hyperplane (which only happens for flat n\mathbb{R}^{n}).

Moreover, in the halfline case, the neutral eigenfunctions are dominant and the diameter of the level sets grows slower than |t|12+δ|t|^{\tfrac{1}{2}+\delta} for any δ>0\delta>0.

Let us explain how the corollary follows: If the flow is strictly convex, then its blowdown is a halfline by our theorem (unless the solution is compact in which case the blowdown is simply a point). If the flow is not strictly convex, then after splitting off a line we get an ancient uniformly 22-convex noncollapsed flow, and such flows have been classified in the recent breakthrough by Angenent-Daskalopoulos-Sesum [ADS19, ADS20] and Brendle-Choi [BC19, BC]. Finally, the ‘moreover statement’ follows as consequence of our proofs as explained in Section 5 and Section 7.

As another corollary, for self-similarly translating solutions we obtain:

Corollary 1.6 (3-convex translators).

Any noncollapsed uniformly 3-convex translator in n+1\mathbb{R}^{n+1} is SO(n1)\mathrm{SO}(n-1)-symmetric.

Indeed, any such translator is either an (n1)(n-1)-dimensional rotationally symmetric bowl soliton times \mathbb{R}, or its blowdown is a halfline. The latter in particular implies that the translator has a unique tip point, and hence is SO(n1)\mathrm{SO}(n-1)-symmetric by a recent result of Zhu [Zhu21].

Remark 1.7 (classification problem for 3-convex translators).

Having established the above corollaries it seems likely that the arguments from [CHH21b, Section 3-5] generalize to higher dimensions to yield that every noncollapsed uniformly 3-convex translator in n+1\mathbb{R}^{n+1} is either ×Bowln1\mathbb{R}\times\mathrm{Bowl}_{n-1}, or Bowln\mathrm{Bowl}_{n}, or belongs to the one-parameter family of O(n1)×2\mathrm{O}(n-1)\times\mathbb{Z}_{2}-symmetric oval-bowls constructed by Hoffman-Ilmanen-Martin-White [HIMW19, Theorem 8.1].

Finally, inspecting all the possibilities in Corollary 1.6 (blowdown of ancient 3-convex flows), as yet another corollary we obtain:

Corollary 1.8 (classification in unstable mode).

The only noncompact ancient noncollpased uniformly 33-convex flow, with a bubble-sheet tangent flow at -\infty, whose unstable mode is dominant is ×Bowln1\mathbb{R}\times\mathrm{Bowl}_{n-1}.

To prove our main theorems, we generalize the arguments from [CHH21a], by Choi, Hershkovits and the second author, to higher dimensions as follows:

In Section 2, we collect some basic properties of ancient noncollapsed flows in n+1\mathbb{R}^{n+1}. In particular, we show that it is enough to analyze the case where the tangent flow at -\infty is given by

(1.3) limλ0λMλ2t=×Sn(2(n)t)\lim_{\lambda\rightarrow 0}\lambda M_{\lambda^{-2}t}=\mathbb{R}^{\ell}\times S^{n-\ell}(\sqrt{2(n-\ell)t})

with =n1\ell=n-1 (respectively k1k-1).

In Section 3, to facilitate barrier and calibration arguments in later sections, we construct an O()×O(n+1)\mathrm{O}(\ell)\times\mathrm{O}(n+1-\ell) symmetric foliation by rotating and shifting the d=n+1d=n+1-\ell dimensional shrinker foliation from [ADS19].

In Section 4, we set up a fine cylindrical analysis, which generalizes the fine neck analysis from [ADS19, ADS20, BC19, BC, CHH18, CHHW19] and the fine bubble sheet analysis from [CHH21a]. Given any space-time point X0=(x0,t0)X_{0}=(x_{0},t_{0}), we consider the renormalized flow

(1.4) M¯τ=eτ2(Mt0eτx0).\bar{M}_{\tau}=e^{\frac{\tau}{2}}\,\left(M_{t_{0}-e^{-\tau}}-x_{0}\right).

Then, the hypersurfaces M¯τ\bar{M}_{\tau} converge for τ\tau\to-\infty to the cylinder

(1.5) Γ=×Sn(2(n)).\Gamma=\mathbb{R}^{\ell}\times S^{n-\ell}(\sqrt{2(n-\ell)}).

The analysis over Γ\Gamma is governed by the Ornstein-Uhlenbeck operator

(1.6) =+12(n)Sn12i=1xixi+1.\mathcal{L}=\triangle_{\mathbb{R}^{\ell}}+\frac{1}{2(n-\ell)}\triangle_{S^{n-\ell}}-\frac{1}{2}\sum_{i=1}^{\ell}x_{i}\frac{\partial}{\partial x_{i}}+1.

Applying the Merle-Zaag lemma, we see that either the unstable eigenfunctions, namely

(1.7) 1,x1,,xn+1,1,x_{1},\ldots,x_{n+1},

are dominant, or the neutral eigenfunctions, namely

(1.8) {xi22}i=1,,{xixj}1i<j{xixJ}1i<Jn+1\{x_{i}^{2}-2\}_{i=1,\ldots,\ell}\cup\{x_{i}x_{j}\}_{1\leq i<j\leq\ell}\cup\{x_{i}x_{J}\}_{1\leq i\leq\ell<J\leq n+1}

are dominant. Further, considering M~τ=S(τ)M¯τ\tilde{M}_{\tau}=S(\tau)\bar{M}_{\tau}, where S(τ)SO(n+1)S(\tau)\in\mathrm{SO}(n+1) is a suitable rotation, we can kill the contribution from {xixJ}\{x_{i}x_{J}\}.

In Section 5, we analyze the case where the neutral eigenfunctions are dominant. We show that along a suitable sequence τi\tau_{i}\to-\infty, the truncated graphical function u^\hat{u} of M~τ\tilde{M}_{\tau} satisfies

(1.9) limτiu^(,τi)u^(,τi)=xQxT2trQ,\lim_{\tau_{i}\to-\infty}\frac{\hat{u}(\cdot,\tau_{i})}{\|\hat{u}(\cdot,\tau_{i})\|}={x^{\prime}}Qx^{\prime T}-2\mathrm{tr}\,Q,

where x=(x1,,x)x^{\prime}=(x_{1},\ldots,x_{\ell}), and Q={qij}Q=\{q_{ij}\} is a nontrivial semi-negative definite ×\ell\times\ell-matrix. We then prove the crucial inclusion

(1.10) Kˇt0kerQ.\check{K}_{t_{0}}\subseteq\mathrm{ker}\,Q.

Roughly speaking, directions that are not in the kernel of the quadratic form are directions of inwards quadratic bending, and thus must be compact directions. We make this precise using the Brunn-Minkowski inequality. Once (1.10) is established, we conclude that dimKˇt01k2\dim\check{K}_{t_{0}}\leq\ell-1\leq k-2 for uniformly kk-convex flows.

In Section 6, we consider ancient noncollapsed flows whose unstable mode is dominant. We prove the fine cylindrical theorem, which says that there exists a nonvanishing vector (a1,,a)(a_{1},\ldots,a_{\ell}) with the following significance: For every space-time center XX, after suitable recentering in the n+1\mathbb{R}^{n+1-\ell}-plane, the truncated graph function uˇX(,τ)\check{u}^{X}(\cdot,\tau) of the renormalized flow M¯τX\bar{M}_{\tau}^{X} satisfies

(1.11) uˇX=eτ2(a1x1+a2x2+ax)+o(eτ2)\displaystyle\check{u}^{X}=e^{\tfrac{\tau}{2}}\left(a_{1}x_{1}+a_{2}x_{2}+\ldots a_{\ell}x_{\ell}\right)+o(e^{\tfrac{\tau}{2}})

for τ0\tau\ll 0 depending only on an upper bound on the cylindrical scale of XX. To show this, we generalize the proofs of the fine neck theorem from [CHH18, CHHW19] and the fine bubble sheet theorem from [CHH21a].

In Section 7, we rule out the case where the unstable mode is dominant. We suppose towards a contradiction that there is an ancient noncollapsed uniformly kk-convex flow Mt=KtM_{t}=\partial K_{t} that does not split off a line, which has a (k1)(k-1)-dimensional blowdown Kˇt0\check{K}_{t_{0}}. We select two distinct rays of differentiability R±Kˇt0R^{\pm}\subset\partial\check{K}_{t_{0}}. Considering suitable corresponding sequences of points in Kt\partial K_{t}, we can pass to limits Kt±K^{\pm}_{t} that split off k2k-2 lines. Using the classification from Brendle-Choi [BC], we argue that Kt±K^{\pm}_{t} must be (n+2k)(n+2-k)-dimensional bowls times (k2)(k-2)-dimensional planes. However, since these bowls translate in different directions, we obtain a contradiction with the fact that the fine cylindrical vector (a1,,a)(a_{1},\ldots,a_{\ell}) is independent of the space-time center point. This concludes the outline of the proof.

Acknowledgments. This research was supported by the NSERC Discovery Grant and the Sloan Research Fellowship of the second author.


2. Basic properties of ancient noncollapsed flows

Let Mt=KtM_{t}=\partial K_{t} be a noncollapsed mean curvature flow in n+1\mathbb{R}^{n+1}, where n3n\geq 3, that is noncompact and strictly convex. Assume further that it is uniformly kk-convex, where k3k\geq 3 (in case k=nk=n, there is no further assumption). It is not hard to see that the blowdown

(2.1) Kˇ=limλ0λKt0\check{K}=\lim_{\lambda\to 0}\lambda K_{t_{0}}

is in fact independent of the choice of t0t_{0} prior to the extinction time (indeed, as one increases t0t_{0}, by mean-convexity the blowdown cannot increase, and using suitable spheres as barriers, similarly as in [CHH21a, Section 2], one sees that it cannot decrease either). By [HK17, CM15] the tangent flow of {Kt}\{K_{t}\} at -\infty in suitable coordinates, is

(2.2) limλ0λKλ2t=×Dn+1(2(n)t),\lim_{\lambda\rightarrow 0}\lambda K_{\lambda^{-2}t}=\mathbb{R}^{\ell}\times D^{n+1-\ell}(\sqrt{2(n-\ell)t}),

for some {1,,k1}\ell\in\{1,\ldots,k-1\}. Together with mean-convexity it follows that

(2.3) Kˇ×{0}.\check{K}\subset\mathbb{R}^{\ell}\times\{0\}.

If k2\ell\leq k-2, then dimKˇk2\dim\check{K}\leq k-2 and we are done. Hence, we can assume from now on that =k1\ell=k-1. Moreover, since our solution is strictly convex, Kˇ\check{K} does not contain any line.

Given any ancient noncollapsed mean curvature flow \mathcal{M}, that is not a round shrinking cylinder, any point X=(x,t)X=(x,t)\in\mathcal{M} and scale r>0r>0, we consider the flow

(2.4) X,r=𝒟1/r(X),\mathcal{M}_{X,r}=\mathcal{D}_{1/r}(\mathcal{M}-X),

that is obtained from \mathcal{M} by translating XX to the space-time origin and parabolically rescaling by 1/r1/r.

Definition 2.1 (ε\varepsilon-cylindrical).

We say that \mathcal{M} is ε\varepsilon-cylindrical around XX at scale rr, if X,r\mathcal{M}_{X,r} is ε\varepsilon-close in C1/εC^{\lfloor 1/\varepsilon\rfloor} in B(0,1/ε)×[1,2]B(0,1/\varepsilon)\times[-1,-2] to the evolution of a round shrinking cylinder ×Sn(2(n)t)\mathbb{R}^{\ell}\times S^{n-\ell}(\sqrt{2(n-\ell)t}) with axis through the origin.

We fix a small enough parameter ε>0\varepsilon>0 quantifying the quality of the cylinders for the rest of the paper. Given X=(x,t)X=(x,t)\in\mathcal{M}, we analyze the solution around XX at the diadic scales rj=2jr_{j}=2^{j}, where jj\in\mathbb{Z}. Using Huisken’s monotonicity formula [Hui90] and quantitative differentiation (see e.g. [CHN13]), for every XX\in\mathcal{M}, we can find an integer J(X)J(X)\in\mathbb{Z} such that

(2.5)  is not ε-cylindrical around X at scale rj for all j<J(X),\textrm{$\mathcal{M}$ is not $\varepsilon$-cylindrical around $X$ at scale $r_{j}$ for all $j<J(X)$},

and

(2.6)  is ε2-cylindrical around X at scale rj for all jJ(X)+N.\textrm{$\mathcal{M}$ is $\tfrac{\varepsilon}{2}$-cylindrical around $X$ at scale $r_{j}$ for all $j\geq J(X)+N$}.
Definition 2.2 (cylindrical scale).

The cylindrical scale of XX\in\mathcal{M} is defined by

(2.7) Z(X)=2J(X).Z(X)=2^{J(X)}.

3. Foliations and barriers

We recall from Angenent-Daskalopoulos-Sesum [ADS19] that there is some L0>1L_{0}>1 such that for every aL0a\geq L_{0} and b>0b>0, there are dd-dimensional shrinkers in d+1\mathbb{R}^{d+1},

(3.1) Σa\displaystyle{\Sigma}_{a} ={surface of revolution with profile r=ua(x1),0x1a},\displaystyle=\{\textrm{surface of revolution with profile }r=u_{a}(x_{1}),0\leq x_{1}\leq a\},
Σ~b\displaystyle\tilde{{\Sigma}}_{b} ={surface of revolution with profile r=u~b(x1),0x1<},\displaystyle=\{\textrm{surface of revolution with profile }r=\tilde{u}_{b}(x_{1}),0\leq x_{1}<\infty\},

as illustrated in [ADS19, Fig. 1]. Here, the parameter aa captures where the concave functions uau_{a} meet the x1x_{1}-axis, namely ua(a)=0u_{a}(a)=0, and the parameter bb is the asymptotic slope of the convex functions u~b\tilde{u}_{b}, namely limx1u~b(x1)=b\lim_{x_{1}\to\infty}\tilde{u}_{b}^{\prime}(x_{1})=b. We recall:

Lemma 3.1 (Lemmas 4.9 and 4.10 in [ADS19]).

There exists some δ>0\delta>0 such that the shrinkers Σa,Σ~b{\Sigma}_{a},\tilde{{\Sigma}}_{b} and Σ:={x22++xd+12=2(d1)}d+1{\Sigma}:=\{x_{2}^{2}+\ldots+x_{d+1}^{2}=2(d-1)\}\subset\mathbb{R}^{d+1} foliate the region

(3.2) {(x1,,xd+1)d+1|x1L0 and x22++xd+122(d1)+δ}.\{(x_{1},\ldots,x_{d+1})\in\mathbb{R}^{d+1}\;|\;x_{1}\geq L_{0}\textrm{ and }x_{2}^{2}+\ldots+x_{d+1}^{2}\leq 2(d-1)+\delta\;\}.

Moreover, denoting by νfol\nu_{\textrm{fol}} the outward unit normal of this family, we have

(3.3) div(ex2/4νfol)=0.\mathrm{div}(e^{-x^{2}/4}\nu_{\mathrm{fol}})=0.

We now write points in n+1\mathbb{R}^{n+1} in the form x=(x,x′′)x=(x^{\prime},x^{\prime\prime}), where x=(x1,,x)x^{\prime}=(x_{1},\dots,x_{\ell}) and x′′=(x+1,,xn+1)x^{\prime\prime}=(x_{\ell+1},\dots,x_{n+1}). We choose d=n+1d=n+1-\ell and shift and rotate the above foliation to construct a suitable foliation in n+1\mathbb{R}^{n+1}:

Definition 3.2 (cylindrical foliation).

For every aL0a\geq L_{0}, we denote by Γa\Gamma_{a} the O()×O(n+1)\mathrm{O}(\ell)\times\mathrm{O}(n+1-\ell) symmetric hypersurface in n+1\mathbb{R}^{n+1} given by

(3.4) Γa={(x,x′′):(|x|1,x′′)Σa}.\Gamma_{a}=\{(x^{\prime},x^{\prime\prime}):(|x^{\prime}|-1,x^{\prime\prime})\in\Sigma_{a}\}.

Similarly, for every bL0b\geq L_{0}, we denote by Γ~b\tilde{\Gamma}_{b} the O()×O(n+1)\mathrm{O}(\ell)\times\mathrm{O}(n+1-\ell) symmetric hypersurface in n+1\mathbb{R}^{n+1} given by

(3.5) Γ~b={(x,x′′):(|x|1,x′′)Σ~b}.\tilde{\Gamma}_{b}=\{(x^{\prime},x^{\prime\prime}):(|x^{\prime}|-1,x^{\prime\prime})\in\tilde{\Sigma}_{b}\}.
Lemma 3.3 (Foliation lemma).

There exist δ>0\delta>0 and L1>2L_{1}>2 such that the hypersurfaces Γa{\Gamma}_{a}, Γ~b{\tilde{\Gamma}}_{b}, and the cylinder Γ:=×Sn(2(n)){\Gamma}:=\mathbb{R}^{\ell}\times S^{n-\ell}(\sqrt{2(n-\ell)}) foliate the domain

Ω:={(x1,,xn+1)|x12++x2L12,x+12++xn+122(n)+δ}.{\Omega}:=\left\{(x_{1},\ldots,x_{n+1})|x_{1}^{2}+\ldots+x_{\ell}^{2}\geq L_{1}^{2},\,x_{\ell+1}^{2}+\ldots+x_{n+1}^{2}\leq 2(n-\ell)+\delta\right\}.

Moreover, denoting by νfol\nu_{\mathrm{fol}} the outward unit normal of this foliation, we have

(3.6) div(νfole|x|2/4)0inside the cylinder,\mathrm{div}(\nu_{\mathrm{fol}}e^{-|x|^{2}/4})\leq 0\;\;\;\textrm{inside the cylinder},

and

(3.7) div(νfole|x|2/4)0outside the cylinder.\mathrm{div}(\nu_{\mathrm{fol}}e^{-|x|^{2}/4})\geq 0\;\;\;\textrm{outside the cylinder}.
Proof.

Let δ>0\delta>0 be as in Lemma 3.1, and set L1=L0+2(1)L_{1}=L_{0}+2(\ell-1). The fact that Γa{\Gamma}_{a}, Γ~b{\tilde{\Gamma}}_{b} and Γ\Gamma foliate Ω\Omega is implied by Lemma 3.1 and Definition 3.2.
Now, observe that for every element in Γ\Gamma_{*} in the foliation of Ω\Omega, we have

(3.8) div(νfole|x|2/4)=(HΓ12x,νfol)e|x|2/4.\mathrm{div}(\nu_{\mathrm{fol}}e^{-|x|^{2}/4})=(H_{\Gamma_{*}}-\frac{1}{2}\left\langle x,\nu_{\mathrm{fol}}\right\rangle)e^{-|x|^{2}/4}\,.

By symmetry, it suffices to compute the curvatures HΓH_{\Gamma_{*}} of Γ\Gamma_{*} in the region {x1>0,x2==x=0}\{x_{1}>0,x_{2}=\dots=x_{\ell}=0\}, where we can identify points and unit normals in Γ\Gamma_{*} with the corresponding ones in Σ\Sigma_{*}, by disregarding the x2,,xx_{2},\dots,x_{\ell} components. The relation between the mean curvature of a surface Σ\Sigma_{*} and its (unshifted) rotation Γn+1\Gamma_{*}\subset\mathbb{R}^{n+1} on points with x1>0,x2==x=0x_{1}>0,x_{2}=\dots=x_{\ell}=0 is given by

(3.9) HΓ=HΣ+1x1e1,νfol,H_{\Gamma_{*}}=H_{\Sigma_{*}}+\frac{\ell-1}{x_{1}}\left\langle e_{1},\nu_{\mathrm{fol}}\right\rangle,

where e1=(1,0,,0)n+2e_{1}=(1,0,\dots,0)\in\mathbb{R}^{n-\ell+2}. For the shrinkers Σa\Sigma_{a}, the concavity of uau_{a} implies e1,νfol0\left\langle e_{1},\nu_{\mathrm{fol}}\right\rangle\geq 0, so we infer that

(3.10) HΓa=12xe1,νfol+1x1e1,νfol12x,νfol,H_{\Gamma_{a}}=\frac{1}{2}\left\langle x-e_{1},\nu_{\mathrm{fol}}\right\rangle+\frac{\ell-1}{x_{1}}\left\langle e_{1},\nu_{\mathrm{fol}}\right\rangle\leq\frac{1}{2}\left\langle x,\nu_{\mathrm{fol}}\right\rangle,

since in Ω{x1>0,x2==x=0}\Omega\cap\{x_{1}>0,x_{2}=\dots=x_{\ell}=0\}, we have x1L12(1)x_{1}\geq L_{1}\geq 2(\ell-1).
For the shrinkers Σ~b\tilde{\Sigma}_{b}, the convexity of u~b\tilde{u}_{b} implies that e1,νfol0\left\langle e_{1},\nu_{\mathrm{fol}}\right\rangle\leq 0, so similarly we have

(3.11) HΓ~b12x,νfol.H_{\tilde{\Gamma}_{b}}\geq\frac{1}{2}\left\langle x,\nu_{\mathrm{fol}}\right\rangle.

This proves the lemma. ∎

Corollary 3.4 (Inner Barriers).

Let {Kτ}τ[τ1,τ2]\{K_{\tau}\}_{\tau\in[\tau_{1},\tau_{2}]} be compact domains, whose boundary evolves by renormalized mean curvature flow. If Γa{\Gamma}_{a} is contained in KτK_{\tau} for every τ[τ1,τ2]\tau\in[\tau_{1},\tau_{2}], and KτΓa=\partial K_{\tau}\cap{\Gamma}_{a}=\emptyset for all τ<τ2\tau<\tau_{2}, then

(3.12) Kτ2ΓaΓa.\partial K_{\tau_{2}}\cap{\Gamma}_{a}\subseteq\partial{\Gamma}_{a}.
Proof.

Lemma 3.3 implies that the vector H+x2\vec{H}+\frac{x^{\perp}}{2} points outwards of Γa{\Gamma}_{a}. The result now follows from the maximum principle. ∎


4. Setting up the fine cylindrical analysis

In this section, we set up the fine cylindrical analysis. This is similar to setting up fine bubble sheet analysis in [CHH21a, Section 4], so we will discuss this briefly. Throughout this section, let \mathcal{M} be an ancient noncollapsed flow in n+1\mathbb{R}^{n+1} whose tangent flow at -\infty is a round shrinking cylinder ×Sn(2(n)t)\mathbb{R}^{\ell}\times S^{n-\ell}(\sqrt{2(n-\ell)t}). Assume further that \mathcal{M} is not self-similarly shrinking. Given any space-time point X0=(x0,t0)X_{0}=(x_{0},t_{0}), we consider the renormalized flow

(4.1) M¯τX0=eτ2(Mt0eτx0).\bar{M}^{X_{0}}_{\tau}=e^{\frac{\tau}{2}}\,\left(M_{t_{0}-e^{-\tau}}-x_{0}\right).

Then, as τ\tau\rightarrow-\infty, the hypersurfaces M¯τX0\bar{M}^{X_{0}}_{\tau} converges to the cylinder

(4.2) Γ=×Sn(2(n)).\Gamma=\mathbb{R}^{\ell}\times S^{n-\ell}({\sqrt{2(n-\ell)}}).

After rescaling, we can assume that Z(X0)1Z(X_{0})\leq 1, and we can find a universal function ρ(τ)>0\rho(\tau)>0 with

(4.3) limτρ(τ)=,andρ(τ)ρ(τ)0,\lim_{\tau\to-\infty}\rho(\tau)=\infty,\quad\textrm{and}-\rho(\tau)\leq\rho^{\prime}(\tau)\leq 0,

so that for every SSO(n+1)S\in\mathrm{SO}(n+1) with (S(Γ),Γ)<ρ(τ)3\sphericalangle(S(\Gamma),\Gamma)<\rho(\tau)^{-3}, the rotated surface S(MτX0)S(M^{X_{0}}_{\tau}) is the graph of a function u=uS(,τ)u=u_{S}(\cdot,\tau) over ΓB2ρ(τ)(0)\Gamma\cap B_{2\rho(\tau)}(0) with

(4.4) u(,τ)C4(ΓB2ρ(τ)(0))ρ(τ)2.\|u(\cdot,\tau)\|_{C^{4}(\Gamma\cap B_{2\rho(\tau)}(0))}\leq\rho(\tau)^{-2}.

We fix a nonnegative smooth cutoff function χ\chi satisfying χ(s)=1\chi(s)=1 for |s|12|s|\leq\frac{1}{2} and χ(s)=0\chi(s)=0 for |s|1|s|\geq 1. We consider the truncated function

(4.5) u^(x,τ):=u(x,τ)χ(rρ(τ)),\hat{u}(x,\tau):=u(x,\tau)\chi\left(\frac{r}{\rho(\tau)}\right),

where

(4.6) r(x):=x12++x2.r(x):=\sqrt{x_{1}^{2}+\ldots+x_{\ell}^{2}}.

Using the implicit function theorem, similarly as in [CHH21a, Proposition 4.1], we can find a differentiable function SX0(τ)SO(n+1)S^{X_{0}}(\tau)\in\mathrm{SO}(n+1), defined for τ\tau sufficiently negative, such that for u:=uSX0(τ)u:=u_{S^{X_{0}}(\tau)} we have

(4.7) ΓBρ(τ)(0)Ax,νΓu^(x,τ)e|x|24=0 for all Ao(n+1).\int_{\Gamma\cap B_{\rho(\tau)}(0)}\langle Ax,\nu_{\Gamma}\rangle\hat{u}(x,\tau)e^{-\frac{|x|^{2}}{4}}=0\qquad\textrm{ for all }A\in o(n+1).

Moreover, we can arrange that for all τ0\tau\ll 0, the matrix

(4.8) A(τ)=S(τ)S(τ)1o(n+1)A(\tau)=S^{\prime}(\tau)S(\tau)^{-1}\in o(n+1)

satisfies Aij(τ)=0A_{ij}(\tau)=0 for 1i,j1\leq i,j\leq\ell, and AIJ(τ)=0A_{IJ}(\tau)=0 for +1I,Jn+1\ell+1\leq I,J\leq n+1.

We now set

(4.9) M~τX0=SX0(τ)M¯τX0,\tilde{M}_{\tau}^{X_{0}}=S^{X_{0}}(\tau)\bar{M}_{\tau}^{X_{0}},

where SX0(τ)SO(n+1)S^{X_{0}}(\tau)\in SO(n+1) is the fine tuning rotation from above, and let

(4.10) u:=uSX0(τ),u:=u_{S^{X_{0}}(\tau)},

so uu describes M~τX0\tilde{M}_{\tau}^{X_{0}} as a graph over the cylinder Γ\Gamma.

Proposition 4.1 (Inverse Poincare inequality).

The graph function uu satisfies the integral estimates

(4.11) Γ{|r|L}e|x|24|u(x,τ)|2CΓ{|r|L2}e|x|24u(x,τ)2\int_{\Gamma\cap\{|r|\leq L\}}e^{-\frac{|x|^{2}}{4}}\,|\nabla u(x,\tau)|^{2}\leq C\int_{\Gamma\cap\{|r|\leq\frac{L}{2}\}}e^{-\frac{|x|^{2}}{4}}\,u(x,\tau)^{2}

and

(4.12) Γ{L2|r|L}e|x|24u(x,τ)2CL2Γ{|r|L2}e|x|24u(x,τ)2\int_{\Gamma\cap\{\frac{L}{2}\leq|r|\leq L\}}e^{-\frac{|x|^{2}}{4}}\,u(x,\tau)^{2}\leq\frac{C}{L^{2}}\int_{\Gamma\cap\{|r|\leq\frac{L}{2}\}}e^{-\frac{|x|^{2}}{4}}\,u(x,\tau)^{2}

for all L[L1,ρ(τ)]L\in[L_{1},\rho(\tau)] and τ0\tau\ll 0, where C<C<\infty is a numerical constant.

Proof.

The proof is similar to the ones of [BC19, Proposition 2.3] and Proposition [CHH21a, Proposition 4.4], with the only change that we now use the higher dimensional foliation from Definition 3.2 and Lemma 3.3. ∎

Let \mathcal{H} be the Hilbert space of all functions ff on Γ\Gamma such that

(4.13) f2=Γ1(4π)n/2e|x|24f2<.\|f\|_{\mathcal{H}}^{2}=\int_{\Gamma}\frac{1}{(4\pi)^{n/2}}e^{-\frac{|x|^{2}}{4}}\,f^{2}<\infty.
Proposition 4.2 (evolution equation and error estimate).

The truncated graph function u^(x,τ)=u(x,τ)χ(rρ(τ))\hat{u}(x,\tau)=u(x,\tau)\,\chi\big{(}\frac{r}{\rho(\tau)}\big{)} satisfies the evolution equation

(4.14) τu^=u^+E^+A(τ)x,νΓχ(rρ(τ)),\partial_{\tau}\hat{u}=\mathcal{L}\hat{u}+\hat{E}+\langle A(\tau)x,\nu_{\Gamma}\rangle\chi\!\left(\frac{r}{\rho(\tau)}\right),

where =ΔΓ12xtanΓ+1\mathcal{L}=\Delta_{\Gamma}-\tfrac{1}{2}\,x^{\text{\rm tan}}\cdot\nabla^{\Gamma}+1, and where the error term E^\hat{E} and the rotation A=SS1A=S^{\prime}S^{-1} satisfy the estimates

(4.15) E^Cρ1(u^+|A(τ)|)and|A(τ)|Cρ1u^\|\hat{E}\|_{\mathcal{H}}\leq C\rho^{-1}\big{(}\|\hat{u}\|_{\mathcal{H}}+|A(\tau)|\big{)}\quad\textrm{and}\quad|A(\tau)|\leq C\rho^{-1}\|\hat{u}\|_{\mathcal{H}}

for τ0\tau\ll 0.

Proof.

The proof is similar as in [CHH21a, Section 4], where we now use Proposition 4.1 to control the error terms. ∎

The cylindrical representation of the operator \mathcal{L} is explicitly given by

(4.16) =+12(n)Sn12i=1xixi+1\mathcal{L}=\triangle_{\mathbb{R}^{\ell}}+\frac{1}{2(n-\ell)}\triangle_{S^{n-\ell}}-\frac{1}{2}\sum_{i=1}^{\ell}x_{i}\frac{\partial}{\partial x_{i}}+1

Analysing the spectrum of \mathcal{L}, the Hilbert space \mathcal{H} from (4.13) can be decomposed as

(4.17) =+0,\mathcal{H}=\mathcal{H}_{+}\oplus\mathcal{H}_{0}\oplus\mathcal{H}_{-},

where

(4.18) +=span{1,x1,,xn+1},\displaystyle\mathcal{H}_{+}=\text{span}\{1,x_{1},\ldots,x_{n+1}\},

and

(4.19) 0=span({xi22}1i{xixj}1i<j{xixJ}1i<Jn+1),\displaystyle\mathcal{H}_{0}=\text{span}\left(\{x_{i}^{2}-2\}_{1\leq i\leq\ell}\cup\{x_{i}x_{j}\}_{1\leq i<j\leq\ell}\cup\{x_{i}x_{J}\}_{1\leq i\leq\ell<J\leq n+1}\right),

and where x1,,xn+1x_{1},\dots,x_{n+1} denotes the restriction of the Euclidean coordinate functions to the cylinder Γ\Gamma. Moreover, we have

f,f12f2\displaystyle\langle\mathcal{L}f,f\rangle_{\mathcal{H}}\geq\frac{1}{2}\,\|f\|_{\mathcal{H}}^{2} for f+f\in\mathcal{H}_{+},
(4.20) f,f=0\displaystyle\langle\mathcal{L}f,f\rangle_{\mathcal{H}}=0 for f0f\in\mathcal{H}_{0},
f,f1nf2\displaystyle\langle\mathcal{L}f,f\rangle_{\mathcal{H}}\leq-\frac{1}{n-\ell}\,\|f\|_{\mathcal{H}}^{2} for ff\in\mathcal{H}_{-}.

Consider the functions

U+(τ):=P+u^(,τ)2,\displaystyle U_{+}(\tau):=\|P_{+}\hat{u}(\cdot,\tau)\|_{\mathcal{H}}^{2},
(4.21) U0(τ):=P0u^(,τ)2,\displaystyle U_{0}(\tau):=\|P_{0}\hat{u}(\cdot,\tau)\|_{\mathcal{H}}^{2},
U(τ):=Pu^(,τ)2,\displaystyle U_{-}(\tau):=\|P_{-}\hat{u}(\cdot,\tau)\|_{\mathcal{H}}^{2},

where P+,P0,PP_{+},P_{0},P_{-} denote the orthogonal projections to +,0,\mathcal{H}_{+},\mathcal{H}_{0},\mathcal{H}_{-}, respectively.

Proposition 4.3 (Merle-Zaag alternative).

For τ\tau\to-\infty, either the neutral mode is dominant, i.e.

(4.22) U+U+=o(U0),U_{-}+U_{+}=o(U_{0}),

or the unstable mode is dominant, i.e.

(4.23) U+U0Cρ1U+.U_{-}+U_{0}\leq C\rho^{-1}U_{+}.
Proof.

Using Proposition 4.2 (evolution equation and error estimate), we obtain

ddτU+(τ)U+(τ)Cρ1(U+(τ)+U0(τ)+U(τ)),\displaystyle\frac{d}{d\tau}U_{+}(\tau)\geq U_{+}(\tau)-C\rho^{-1}\,(U_{+}(\tau)+U_{0}(\tau)+U_{-}(\tau)),
(4.24) |ddτU0(τ)|Cρ1(U+(τ)+U0(τ)+U(τ)),\displaystyle\Big{|}\frac{d}{d\tau}U_{0}(\tau)\Big{|}\leq C\rho^{-1}\,(U_{+}(\tau)+U_{0}(\tau)+U_{-}(\tau)),
ddτU(τ)2nU(τ)+Cρ1(U+(τ)+U0(τ)+U(τ)).\displaystyle\frac{d}{d\tau}U_{-}(\tau)\leq-\frac{2}{n-\ell}U_{-}(\tau)+C\rho^{-1}\,(U_{+}(\tau)+U_{0}(\tau)+U_{-}(\tau)).

Hence, the Merle-Zaag ODE lemma [MZ98, CM19] implies the assertion. ∎


5. Fine cylindrical analysis in the neutral mode

In this section, we prove our main theorems in the case where the neutral mode is dominant. Namely, throughout this section we assume that our truncated graph function u^(,τ)\hat{u}(\cdot,\tau) satisfies

(5.1) U+U+=o(U0).U_{-}+U_{+}=o(U_{0}).

Using this together with the evolution inequality (4.24), given any δ>0\delta>0, we see that |ddτlogU0|12δ|\tfrac{d}{d\tau}\log U_{0}|\leq\tfrac{1}{2}\delta for sufficiently large τ-\tau, and thus

(5.2) U0(τ)eδτU_{0}(\tau)\geq e^{\delta\tau}

for sufficiently large τ-\tau.

Proposition 5.1.

Every sequence {τi}\{\tau_{i}\} converging to -\infty has a subsequence {τim}\{\tau_{i_{m}}\} such that

(5.3) limτimu^(,τim)u^(,τim)=xQxT2trQ,\lim_{\tau_{i_{m}}\to-\infty}\frac{\hat{u}(\cdot,\tau_{i_{m}})}{\|\hat{u}(\cdot,\tau_{i_{m}})\|_{\mathcal{H}}}={x^{\prime}}Qx^{\prime T}-2\mathrm{tr}\,Q,

in \mathcal{H}-norm, where Q={qij}Q=\{q_{ij}\} is a nontrivial semi-negative definite ×\ell\times\ell-matrix, and x=(x1,,x)x^{\prime}=(x_{1},\dots,x_{\ell}).

Proof.

Recalling (4.19) and using the orthogonality condition (4.7), we see that

(5.4) P0u^span({xi22}1i{xixj}1i<j)0.P_{0}\hat{u}\in\text{span}\left(\{x_{i}^{2}-2\}_{1\leq i\leq\ell}\cup\{x_{i}x_{j}\}_{1\leq i<j\leq\ell}\right)\subset\mathcal{H}_{0}.

Therefore, every sequence τi\tau_{i}\to-\infty has a subsequence {τim}\{\tau_{i_{m}}\} such that

(5.5) limτimu^(,τim)u^(,τim)=𝒬(x)\lim_{\tau_{i_{m}}\to-\infty}\frac{\hat{u}(\cdot,\tau_{i_{m}})}{\|\hat{u}(\cdot,\tau_{i_{m}})\|_{\mathcal{H}}}=\mathcal{Q}(x^{\prime})

with respect to the {\mathcal{H}}-norm, where x=(x1,,x)x^{\prime}=(x_{1},\dots,x_{\ell}), and

(5.6) 𝒬(x):=xQxT2trQ\mathcal{Q}(x^{\prime}):={x^{\prime}}Qx^{\prime T}-2\mathrm{tr}\,Q

for some nontrivial matrix Q={qij}Q=\{q_{ij}\}. After an orthogonal change of coordinates in the xx^{\prime}-plane, we can assume that qij=0q_{ij}=0 for iji\not=j. Let us show that q110q_{11}\leq 0 (the same argument yields qii0q_{ii}\leq 0 for 1i1\leq i\leq\ell).
We denote by K~τ\tilde{K}_{\tau} the region enclosed by M~τ\tilde{M}_{\tau} and denote by 𝒜(x,τ)\mathcal{A}(x^{\prime},\tau) the area of the cross section of K~τ\tilde{K}_{\tau} with fixed xx^{\prime} coordinates. Explicitly,

(5.7) 𝒜(x,τ)=1n+1Sn(2(n)+u(x,ω,τ))n+1𝑑S(ω),\displaystyle\mathcal{A}(x^{\prime},\tau)=\frac{1}{n+1-\ell}\int_{S^{n-\ell}}\big{(}\sqrt{2(n-\ell)}+u(x^{\prime},\omega,\tau)\big{)}^{n+1-\ell}dS(\omega),

which can be expanded as

𝒜(x,τ)=\displaystyle\mathcal{A}(x^{\prime},\tau)= |Bn+1(2(n))|+(2(n))n2Snu(x,ω,τ)𝑑S(ω)\displaystyle|B^{n+1-\ell}(\sqrt{2(n-\ell)})|+\left(2(n-\ell)\right)^{\frac{n-\ell}{2}}\int_{S^{n-\ell}}u(x^{\prime},\omega,\tau)dS(\omega)
(5.8) +k=2n+1ck(n,)Snuk(x,ω,τ)𝑑S(ω).\displaystyle+\sum_{k=2}^{n-\ell+1}c_{k}(n,\ell)\int_{S^{n-\ell}}u^{k}(x^{\prime},\omega,\tau)dS(\omega).

By Brunn’s concavity principle, the function x𝒜1n+1(x,τ)x^{\prime}\mapsto{\mathcal{A}^{\frac{1}{n-\ell+1}}(x^{\prime},\tau)} is concave. In particular, we have

(5.9) 𝒜1n+1(x12,x2,,x,τ)+𝒜1n+1(x1+2,x2,,x,τ)2𝒜1n+1(x1,x2,,x,τ).{\mathcal{A}^{\frac{1}{n-\ell+1}}(x_{1}-2,x_{2},\dots,x_{\ell},\tau)}+{\mathcal{A}^{\frac{1}{n-\ell+1}}(x_{1}+2,x_{2},\dots,x_{\ell},\tau)}\\ \leq 2{\mathcal{A}^{\frac{1}{n-\ell+1}}(x_{1},x_{2},\dots,x_{\ell},\tau)}.

This implies

(5.10) [1,1]𝒜1n+1(x,τ)𝑑x13331111𝒜1n+1(x,τ)𝑑x1𝑑x.\int_{[-1,1]^{\ell}}\!\!\!\!\!\!\mathcal{A}^{\frac{1}{n-\ell+1}}(x^{\prime},\tau)\,dx^{\prime}\geq\frac{1}{3}\int_{-3}^{3}\int_{-1}^{1}\ldots\int_{-1}^{1}\mathcal{A}^{\frac{1}{n-\ell+1}}(x^{\prime},\tau)\,dx_{1}\dots dx_{\ell}.

Combining this with (5.5) and (5) yields

(5.11) [1,1]𝒬(x)𝑑x13331111𝒬(x)𝑑x1𝑑x.\int_{[-1,1]^{\ell}}\mathcal{Q}(x^{\prime})\,dx^{\prime}\geq\frac{1}{3}\int_{-3}^{3}\int_{-1}^{1}\ldots\int_{-1}^{1}\mathcal{Q}(x^{\prime})\,dx_{1}\dots dx_{\ell}.

This implies q110q_{11}\leq 0, and thus concludes the proof. ∎

Theorem 5.2.

If the neutral mode is dominant, then the blowdown Kˇ\check{K} satisfies

(5.12) KˇkerQ.\check{K}\subseteq\mathrm{ker}\,Q.

for any QQ from Proposition 5.1. In particular, dimKˇ1k2\dim\check{K}\leq\ell-1\leq k-2.

Proof.

Let

(5.13) 𝒬(x):=xQxT2trQ.\mathcal{Q}(x^{\prime}):={x^{\prime}}Qx^{\prime T}-2\mathrm{tr}\,Q.

Given a unit vector vv in the xx^{\prime}-plane, let {v,w1,,w1}\{v,w_{1},\dots,w_{\ell-1}\} be an orthonormal basis of the xx^{\prime}-plane. Set c=1/nc=1/n. Since 𝒬(x)>0\mathcal{Q}(x^{\prime})>0 for |x|c|x^{\prime}|\leq c, we have

(5.14) 0c[0,c]1𝒬(rv+s1w1++s1w1)𝑑r𝑑s>0.\int_{0}^{c}\int_{[0,c]^{\ell-1}}\mathcal{Q}(rv+s_{1}w_{1}+\dots+s_{\ell-1}w_{\ell-1})\,drds>0.

On the other hand, if vkerQv\notin\ker Q, then for d=d((v,kerQ))d=d(\sphericalangle(v,\ker Q)) sufficiently large, we have

(5.15) dd+c[0,c]1𝒬(rv+s1w1++s1w1)𝑑r𝑑s<0.\int_{d}^{d+c}\int_{[0,c]^{\ell-1}}\mathcal{Q}(rv+s_{1}w_{1}+\dots+s_{\ell-1}w_{\ell-1})\,drds<0.

Defining 𝒜\mathcal{A} as in (5.7), similarly as in the previous proof we have

ab[0,c]1𝒜1n+1(rv+s1w1++s1w1,τim)|Bn+1(2(n))|1n+1u^(,τim)𝑑r𝑑s\displaystyle\int_{a}^{b}\int_{[0,c]^{\ell-1}}\!\!\!\!\frac{{\scriptstyle\mathcal{A}^{\frac{1}{n-\ell+1}}(rv+s_{1}w_{1}+\dots+s_{\ell-1}w_{\ell-1},\tau_{i_{m}})-|B^{n-\ell+1}(\sqrt{2(n-\ell)})|^{\frac{1}{n-\ell+1}}}}{||\hat{u}(\cdot,\tau_{i_{m}})||_{\mathcal{H}}}drds
(5.16) c(n,)ab[0,c]1𝒬(rv+s1w1++s1w1)𝑑r𝑑s\displaystyle\rightarrow c(n,\ell)\int_{a}^{b}\int_{[0,c]^{\ell-1}}\mathcal{Q}(rv+s_{1}w_{1}+\dots+s_{\ell-1}w_{\ell-1})drds

as mm\rightarrow\infty. Combining (5.14), (5.15) and (5) we see that for every mm sufficiently large, there exist rm,s1,m,,s1,m[0,1]r_{m},s_{1,m},\dots,s_{\ell-1,m}\in[0,1] such that

(5.17) 𝒜(rmv+s1,mw1++s1,,mw1,τim)>𝒜((rm+d)v+s1,mw1++s1,,mw1,τim).\quad\mathcal{A}(r_{m}v+s_{1,m}w_{1}+\dots+s_{\ell-1,,m}w_{\ell-1},\tau_{i_{m}})\\ >\mathcal{A}((r_{m}+d)v+s_{1,m}w_{1}+\dots+s_{\ell-1,,m}w_{\ell-1},\tau_{i_{m}}).

Now, suppose towards a contradiction there is some ωKˇkerQ\omega\in\check{K}\setminus\ker Q. Since S(τ)IS(\tau)\rightarrow I as τ\tau\rightarrow-\infty, for all τ-\tau sufficiently large, we have

(5.18) (P(S(τ)ω),kerQ)>12(ω,kerQ),\sphericalangle(P(S(\tau)\omega),\ker Q)>\tfrac{1}{2}\sphericalangle(\omega,\ker Q),

where PP denotes the projection to the xx^{\prime}-plane. Set

(5.19) vm:=P(S(τim)ω)|P(S(τim)ω)|.v_{m}:=\frac{P(S(\tau_{i_{m}})\omega)}{|P(S(\tau_{i_{m}})\omega)|}.

Take v=vmv=v_{m} in the previous discussion, and let {vm,w1,m,,w1,m}\{v_{m},w_{1,m},\dots,w_{\ell-1,m}\} be an orthonromal basis of xx^{\prime}-plane. Note that

(5.20) rmvm+s1,mw1,m++s1,mw1,m+λS(τim)ωK~τmλ[0,).r_{m}v_{m}+s_{1,m}w_{1,m}+\dots+s_{\ell-1,m}w_{\ell-1,m}+\lambda S(\tau_{i_{m}})\omega\in\tilde{K}_{\tau_{m}}\,\forall\lambda\in[0,\infty).

On the other hand, by Brunn’s concavity principle, the function

(5.21) r𝒜1nl+1(rvm+s1,mw1,m++s1,mw1,m)r\mapsto\mathcal{A}^{\frac{1}{n-l+1}}(rv_{m}+s_{1,m}w_{1,m}+\dots+s_{\ell-1,m}w_{\ell-1,m})

is concave, as long as it does not vanish. Together with (5.17), this implies that for all mm sufficiently large, the area of the cross sections is decreasing for r>rmr>r_{m}, and vanishes at some finite rr_{\ast}. This contradicts (5.20), as the ray would have nowhere to go. This proves the theorem. ∎

Corollary 5.3.

Assume the solution is noncompact and uniformly 33-convex. Choosing coordinates such that Kˇ={λe1|λ0}\check{K}=\{\lambda e_{1}|\lambda\geq 0\}, we have

(5.22) limτu^(,τ)u^(,τ)=x222x222.\lim_{\tau\to-\infty}\frac{\hat{u}(\cdot,\tau)}{\|\hat{u}(\cdot,\tau)\|_{\mathcal{H}}}=-\frac{x_{2}^{2}-2}{\|x_{2}^{2}-2\|_{\mathcal{H}}}.

Furthermore, given any δ>0\delta>0 for τ0\tau\ll 0 the level-sets satisfy

(5.23) M¯τ{x1=0}Bexp(δ|τ|)(0).\bar{M}_{\tau}\cap\{x_{1}=0\}\subseteq B_{\exp{(\delta|\tau|)}}(0).
Proof.

Since a normalized semi-negative definite rank one 2×22\times 2 quadratic form is uniquely determined by its kernel, we have convergence without the need of passing to a subsequence. Furthermore, arguing as in [CHH21a, Corollary 5.4] we get the bound for the diameter of the level sets. ∎


6. Fine cylindrical analysis in the unstable mode

In this section, we prove the fine cylindrical theorem for ancient noncollapsed flows whose tangent flow at -\infty is ×Dn+1(2(n)t)\mathbb{R}^{\ell}\times D^{n+1-\ell}(\sqrt{2(n-\ell)t}). This is similar to the fine neck theorem in [CHH18, CHHW19] and the fine bubble sheet theorem in [CHH21a], so we will discuss this rather briefly.

Throughout this section, we assume the tilted renormalized flow M~τX0\tilde{M}^{X_{0}}_{\tau} around some point X0X_{0}, has a dominant unstable mode, i.e.

(6.1) U0+UCρ1U+.U_{0}+U_{-}\leq C\rho^{-1}U_{+}.

Together with (4.24), this implies ddτU+(τ)(1Cρ1)U+\frac{d}{d\tau}U_{+}(\tau)\geq(1-C\rho^{-1})U_{+}, hence

(6.2) U=(1+o(1))U+Ce910τ.U=(1+o(1))U_{+}\leq Ce^{\frac{9}{10}\tau}.

for sufficiently large τ-\tau. Therefore, Proposition 4.2 yields

(6.3) |SX0(τ)I|Ce920τ.|S^{X_{0}}(\tau)-I|\leq Ce^{\frac{9}{20}\tau}.

We now consider the untilted renormalized flow M¯τX\bar{M}^{X}_{\tau} around any center XX. For τ𝒯(Z(X))\tau\leq\mathcal{T}(Z(X)). we can write M¯τX\bar{M}^{X}_{\tau} as a graph of a function u¯(,τ)\bar{u}(\cdot,\tau) over ΓB2ρ(τ)(0)\Gamma\cap B_{2\rho(\tau)}(0), where ρ=ρX(,τ)\rho=\rho^{X}(\cdot,\tau) satisfies (4.3) and (4.4). We will work with the truncated function

(6.4) uˇ=u¯χ(r/ρ(τ)).\check{u}=\bar{u}\chi(r/\rho(\tau)).

Moreover, we set

Uˇ+(τ):=P+uˇ(,τ)2,\displaystyle\check{U}_{+}(\tau):=\|P_{+}\check{u}(\cdot,\tau)\|_{\mathcal{H}}^{2},
(6.5) Uˇ0(τ):=P0uˇ(,τ)2,\displaystyle\check{U}_{0}(\tau):=\|P_{0}\check{u}(\cdot,\tau)\|_{\mathcal{H}}^{2},
Uˇ(τ):=Puˇ(,τ)2.\displaystyle\check{U}_{-}(\tau):=\|P_{-}\check{u}(\cdot,\tau)\|_{\mathcal{H}}^{2}.

Applying the Merle-Zaag ODE lemma [MZ98, CM19], we infer that there are universal constants C,R<C,R<\infty, such that for every XX either

(6.6) Uˇ++Uˇ=o(Uˇ0),\check{U}_{+}+\check{U}_{-}=o(\check{U}_{0}),

or

(6.7) Uˇ0+UˇCρ1Uˇ+\check{U}_{0}+\check{U}_{-}\leq C\rho^{-1}\check{U}_{+}

whenever ρR\rho\geq R.

Lemma 6.1 (dominant mode).

The unstable mode is dominant for the untilted renormalized flow M¯τX\bar{M}^{X}_{\tau} with any center XX, i.e. (6.7) holds for all XX.

Proof.

Let us first show that the statement holds for X=X0X=X_{0}. Combining (6.2) and (6.3), we infer that the untilted flow satisfies

(6.8) UˇCe920τ.\check{U}\leq Ce^{\frac{9}{20}\tau}.

If we had Uˇ++Uˇ=o(Uˇ0)\check{U}_{+}+\check{U}_{-}=o(\check{U}_{0}), then arguing as at the beginning of Section 5 we would see that Uˇeδτ\check{U}\geq e^{\delta\tau} for every δ>0\delta>0 and τ-\tau sufficiently large, contradicting (6.8). Thus, for X=X0X=X_{0}, we indeed get (6.7). Finally, since any neck centered at a general point XX merges with the neck centered at X0X_{0} as τ\tau\to-\infty, the inequality (6.7) holds for every XX. ∎

Proposition 6.2 (improved graphical radius).

There exists some 𝒯=𝒯(Z(X))>\mathcal{T}=\mathcal{T}(Z(X))>-\infty, such that

(6.9) ρ¯(τ)=e19τ\bar{\rho}(\tau)=e^{-\frac{1}{9}\tau}

is a graphical radius function satisfying (4.3) and (4.4) for τ𝒯\tau\leq\mathcal{T}.

Proof.

The proof is similar to the one of [CHH21a, Proposition 6.4] (see also [CHH18, CHHW19]), with the only difference that we now use the higher dimensional barriers from Definition 3.2 and Corollary 3.4. ∎

From now on, we work with ρ=ρ¯\rho=\bar{\rho} from Proposition 6.2 (improved graphical radius), and in particular, we define uˇ,Uˇ0,Uˇ±,\check{u},\check{U}_{0},\check{U}_{\pm},\ldots with respect to this improved graphical radius. Note that the unstable mode is still dominant.

Proposition 6.3 (sharp decay estimate).

There exist constants C<C<\infty and 𝒯>\mathcal{T}>-\infty, depending only on an upper bound for Z(X)Z(X), such that for τ𝒯\tau\leq\mathcal{T}:

(6.10) uˇ(,τ)Ceτ/2,\|\check{u}(\cdot,\tau)\|_{\mathcal{H}}\leq Ce^{\tau/2},

and

(6.11) uˇ(,τ)Cn({r100})Ceτ/2.\|\check{u}(\cdot,\tau)\|_{C^{n}(\{r\leq 100\})}\leq Ce^{\tau/2}.
Proof.

Since M¯𝒯X\bar{M}^{X}_{\mathcal{T}} is an ε\varepsilon-graph over ΓB2ρX(0)\Gamma\cap B_{2\rho^{X}}(0), we get uˇ(,𝒯)1\|\check{u}(\cdot,\mathcal{T})\|_{\mathcal{H}}\leq 1. Together with the evolution inequality ddτ(e910τUˇ+)0\tfrac{d}{d\tau}(e^{-\frac{9}{10}\tau}\check{U}_{+})\geq 0, this implies

(6.12) uˇ(,τ)Ce920τ\|\check{u}(\cdot,\tau)\|_{\mathcal{H}}\leq Ce^{\frac{9}{20}\tau}

for τ𝒯\tau\leq\mathcal{T}, where C<C<\infty and 𝒯>\mathcal{T}>-\infty only depend on Z(X)Z(X).

Now, combining the evolution inequality ddτUˇ+(1Cρ1)Uˇ+\tfrac{d}{d\tau}\check{U}_{+}\geq(1-C\rho^{-1})\check{U}_{+} with the rough estimate (6.12) and Proposition 6.2 (improved graphical radius) yields

(6.13) ddτ(eτUˇ+)Ceτ+19τ+910τ=Ce190τ\frac{d}{d\tau}\left(e^{-\tau}\check{U}_{+}\right)\geq-Ce^{-\tau+\frac{1}{9}\tau+\frac{9}{10}\tau}=-Ce^{\frac{1}{90}\tau}

for all τ𝒯\tau\leq\mathcal{T}. Thus, Uˇ+Ceτ\check{U}_{+}\leq Ce^{\tau}. This proves (6.10). Finally, (6.11) follows from the parabolic estimates in [CHH21a, Appendix A]. ∎

Theorem 6.4 (fine cylindrical theorem).

If {Mt}\{M_{t}\} is an ancient noncollapsed flow in n+1\mathbb{R}^{n+1}, whose tangent flow at -\infty is ×Sn(2(n)t)\mathbb{R}^{\ell}\times S^{n-\ell}(\sqrt{2(n-\ell)t}), and whose unstable mode is dominant, then there exist constants a1,,an+1a_{1},\ldots,a_{n+1} with a12++a2>0a_{1}^{2}+\ldots+a_{\ell}^{2}>0 with the following significance. For every center point X=(x01,x0n+1,t0)X=(x_{0}^{1},\ldots x_{0}^{n+1},t_{0}), the truncated graphical function uˇX(,τ)\check{u}^{X}(\cdot,\tau) of the renormalized flow M¯τX\bar{M}_{\tau}^{X} satisfies the fine cylindrical expansion estimates

(6.14) uˇXeτ2(i=1aixi+J=+1n+1aJXxJ)Ce59τ,\Big{\|}\check{u}^{X}-e^{\frac{\tau}{2}}\Big{(}\sum_{i=1}^{\ell}a_{i}x_{i}+\sum_{J=\ell+1}^{n+1}a_{J}^{X}x_{J}\Big{)}\Big{\|}_{\mathcal{H}}\leq Ce^{\frac{5}{9}\tau},

and

(6.15) uˇXeτ2(i=1aixi+J=+1n+1aJXxJ)L({r100})Ce1936τ\Big{\|}\check{u}^{X}-e^{\frac{\tau}{2}}\Big{(}\sum_{i=1}^{\ell}a_{i}x_{i}+\sum_{J=\ell+1}^{n+1}a_{J}^{X}x_{J}\Big{)}\Big{\|}_{L^{\infty}(\{r\leq 100\})}\leq Ce^{\frac{19}{36}\tau}

for τ𝒯\tau\leq\mathcal{T}, where aJX=aJ12(n)x0Ja_{J}^{X}=a_{J}-\frac{1}{\sqrt{2(n-\ell)}}x^{J}_{0} for J+1J\geq\ell+1, and where C<C<\infty and 𝒯>\mathcal{T}>-\infty only depend on an upper bound for the cylindrical scale Z(X)Z(X).

Proof.

In the following C<C<\infty and 𝒯>\mathcal{T}>-\infty denote constants that might change from line to line, but only depend on an upper bound for the cylindrical scale Z(X)Z(X). Recalling the basis of +\mathcal{H}_{+} from (4.18), we can write

(6.16) P+uˇX=a0X(τ)+i=1n+1aiX(τ)xi,\displaystyle P_{+}\check{u}^{X}=a_{0}^{X}(\tau)+\sum_{i=1}^{n+1}a_{i}^{X}(\tau)x_{i},

where the superscript indicates (a priori) dependence on the center XX.

Letting Eˇ:=(τ)uˇ\check{E}:=(\partial_{\tau}-\mathcal{L})\check{u}, and using  1=1\mathcal{L}\,1=1, we compute

(6.17) ddτa0X(τ)=a0X(τ)+en2(4π)2|Sn(2(n))|1ΓEˇe|x|24.\displaystyle\frac{d}{d\tau}a_{0}^{X}(\tau)=a_{0}^{X}(\tau)+{e^{\frac{n-\ell}{2}}(4\pi)^{\frac{-\ell}{2}}}|S^{n-\ell}(\sqrt{2(n-\ell)})|^{-1}\int_{\Gamma}\check{E}e^{-\frac{|x|^{2}}{4}}\,.

Hence, using Proposition 4.2, Proposition 6.2 and Proposition 6.3, we obtain

(6.18) |ddτ(eτa0X(τ))|CeτEˇCeτ+τ2+τ9Ce718τ.\displaystyle\Big{|}\frac{d}{d\tau}\left(e^{-\tau}a_{0}^{X}(\tau)\right)\Big{|}\leq Ce^{-\tau}\|\check{E}\|_{\mathcal{H}}\leq Ce^{-\tau+\frac{\tau}{2}+\frac{\tau}{9}}\leq Ce^{-\frac{7}{18}\tau}.

Integrating this from τ\tau to 𝒯\mathcal{T} implies

(6.19) |a0X(τ)|Ce1118τ(τ𝒯).|a_{0}^{X}(\tau)|\leq Ce^{\tfrac{11}{18}\tau}\qquad(\tau\leq\mathcal{T}).

In a similar manner, using xi=12xi\mathcal{L}x_{i}=\frac{1}{2}x_{i} for i=1,,n+1i=1,\ldots,n+1, we get

(6.20) |ddτ(eτ2aiX(τ))|Ceτ2EˇCeτ9,\displaystyle\Big{|}\frac{d}{d\tau}\left(e^{-\frac{\tau}{2}}a_{i}^{X}(\tau)\right)\Big{|}\leq Ce^{-\frac{\tau}{2}}\|\check{E}\|_{\mathcal{H}}\leq Ce^{\frac{\tau}{9}},

so integrating from -\infty to τ\tau yields

(6.21) i=1n+1|eτ2aiX(τ)a¯iX|Ce19τ(τ𝒯),\sum_{i=1}^{n+1}\left|e^{-\tfrac{\tau}{2}}a_{i}^{X}(\tau)-\bar{a}_{i}^{X}\right|\leq Ce^{\frac{1}{9}\tau}\qquad(\tau\leq\mathcal{T}),

where

(6.22) a¯iX=limτeτ/2aiX(τ)(i=1,,n+1).\bar{a}_{i}^{X}=\lim_{\tau\to-\infty}e^{-\tau/2}a_{i}^{X}(\tau)\qquad(i=1,\ldots,n+1).

Now, consider the difference

(6.23) DX:=uˇXeτ2i=1n+1a¯iXxi.D^{X}:=\check{u}^{X}-e^{\frac{\tau}{2}}\sum_{i=1}^{n+1}\bar{a}_{i}^{X}x_{i}.

Using the above, we see that

(6.24) |DX||uˇXP+uˇX|+C(|x|+1)e1118τ.|D^{X}|\leq|\check{u}^{X}-P_{+}\check{u}^{X}|+C(|x|+1)e^{\tfrac{11}{18}\tau}.

Together with the inequality Uˇ0+UˇCe109τ\check{U}_{0}+\check{U}_{-}\leq Ce^{\frac{10}{9}\tau}, which follows from Lemma 6.1, Proposition 6.2 and Proposition 6.3, this yields

(6.25) DXCe59τ,\|D^{X}\|_{\mathcal{H}}\leq Ce^{\frac{5}{9}\tau},

which proves (6.14) modulo the claim about the coefficients.

Next, we observe that

(6.26) DXL2({r100})CDXCe59τ,\displaystyle\|D^{X}\|_{L^{2}(\{r\leq 100\})}\leq C\|D^{X}\|_{\mathcal{H}}\leq Ce^{\frac{5}{9}\tau},

and, using Proposition 6.3, that

(6.27) nDXL2({r100})Cu^XCn({r100})+Ceτ2Ceτ2.\displaystyle\|\nabla^{n}D^{X}\|_{L^{2}(\{r\leq 100\})}\leq C\|\hat{u}^{X}\|_{C^{n}(\{r\leq 100\})}+Ce^{\frac{\tau}{2}}\leq Ce^{\frac{\tau}{2}}.

Applying Agmon’s inequality, this yields

(6.28) DXL({r100})CDXL2({r100})12DXHn({r100})12Ce1936τ.\|D^{X}\|_{L^{\infty}(\{r\leq 100\})}\leq C\|D^{X}\|_{L^{2}(\{r\leq 100\})}^{\frac{1}{2}}\|D^{X}\|_{H^{n}(\{r\leq 100\})}^{\frac{1}{2}}\leq Ce^{\tfrac{19}{36}\tau}.

Now, comparing the renormalized flows with center X=(x01,,x0n+1,t0)X=(x_{0}^{1},\ldots,x_{0}^{n+1},t_{0}) and center 0, we need to relate the parameters of the functions uˇX\check{u}^{X} and uˇ0\check{u}^{0}. Since it is easy to see that the parameters do not depend on t0t_{0}, we may choose t0=0t_{0}=0. We then have

(6.29) 2(n)+uˇX(xx0eτ/2,x′′+O(eτ/2),τ)=dist((2(n)+uˇ0(x,x′′,τ))x′′|x′′|,eτ/2x0′′),\sqrt{2(n-\ell)}+\check{u}^{X}(x^{\prime}-x_{0}^{\prime}e^{\tau/2},x^{\prime\prime}+O(e^{\tau/2}),\tau)\\ =\mathrm{dist}\left(\left(\sqrt{2(n-\ell)}+\check{u}^{0}(x^{\prime},x^{\prime\prime},\tau)\right)\frac{x^{\prime\prime}}{|x^{\prime\prime}|},e^{\tau/2}x_{0}^{\prime\prime}\right),

where we use the notation x=(x1,,x)x^{\prime}=(x_{1},\ldots,x_{\ell}) and x′′=(x+1,,xn+1)x^{\prime\prime}=(x_{\ell+1},\ldots,x_{n+1}) as before. By Taylor expansion and Proposition 6.3, we have

(6.30) dist((2(n)+uˇ0(x,x′′,τ))x′′|x′′|,eτ/2x0′′)=2(n)+uˇ0(x,x′′,τ)eτ/2x′′x0′′|x′′|+o(eτ/2).\mathrm{dist}\left(\left(\sqrt{2(n-\ell)}+\check{u}^{0}(x^{\prime},x^{\prime\prime},\tau)\right)\frac{x^{\prime\prime}}{|x^{\prime\prime}|},e^{\tau/2}x_{0}^{\prime\prime}\right)\\ =\sqrt{2(n-\ell)}+\check{u}^{0}(x^{\prime},x^{\prime\prime},\tau)-e^{\tau/2}\frac{x^{\prime\prime}\cdot x^{\prime\prime}_{0}}{|x^{\prime\prime}|}+o(e^{\tau/2}).

Together with (6.28), the above formulas imply that a¯iX=a¯i0\bar{a}^{X}_{i}=\bar{a}^{0}_{i} for 1i1\leq i\leq\ell, and a¯JX=a¯J012(n)x0J\bar{a}^{X}_{J}=\bar{a}^{0}_{J}-\frac{1}{\sqrt{2(n-\ell)}}x^{J}_{0} for +1Jn+1\ell+1\leq J\leq n+1.

Finally, by the same contradiction argument as in [CHH21a, Section 6.4] (see also [ADS19, Lemma 5.11]) we obtain

(6.31) a12++a2>0.a_{1}^{2}+\ldots+a_{\ell}^{2}>0.

This finishes the proof of the fine cylindrical theorem. ∎


7. Conclusion of the proof

In this final section, we conclude the proof of our main theorem:

Theorem 7.1.

Let Mt=Ktn+1M_{t}=\partial K_{t}\subset\mathbb{R}^{n+1} be an ancient noncollapsed mean curvature flow that does not split off a line. Asume that the flow is uniformly kk-convex for some k3k\geq 3. Then the blowdown of any time slice is at most k2k-2 dimensional, namely

(7.1) dimKˇt0k2.\dim\check{K}_{t_{0}}\leq k-2.
Proof.

Let Mt=Ktn+1M_{t}=\partial K_{t}\subset\mathbb{R}^{n+1} be an ancient noncollapsed uniformly kk-convex mean curvature flow that does not split off a line. By the reduction from Section 2, we can assume that

(7.2) limλ0λMλ2t=×Sn(2(n)t),\lim_{\lambda\rightarrow 0}\lambda M_{\lambda^{-2}t}=\mathbb{R}^{\ell}\times S^{n-\ell}(\sqrt{2(n-\ell)t}),

where =k1\ell=k-1. Considering the expansion of the renormalized flow over the cylinder Γ=×Sn(2(n))\Gamma=\mathbb{R}^{\ell}\times S^{n-\ell}(\sqrt{2(n-\ell)}), by Proposition 4.3 (Merle-Zaag alternative) and Lemma 6.1 (dominant mode), for τ\tau\to-\infty, either the neutral mode dominates for the tilted flow, or the unstable mode dominates for the untilted flow with respect to any center. In the dominant neutral mode case, we have shown in Theorem 5.2 that the blowdown is at most k2k-2 dimensional. We can thus assume that we are in the dominant unstable mode case. Since we already know that dimKˇ\dim\check{K}\leq\ell, and since the dimension of convex sets is always an integer, it suffices to rule out the scenario that dimKˇ==k1\dim\check{K}=\ell=k-1.

Suppose towards a contradiction that dimKˇ==k1\dim\check{K}=\ell=k-1. Recall that Kˇ\check{K} does not contain any line and also recall that k3k\geq 3. Choosing suitable coordinates, we can assume that Kˇ{0}\check{K}\setminus\{0\} is contained in (×{0}){x1>0}(\mathbb{R}^{\ell}\times\{0\})\cap\{x_{1}>0\} and contains the positive x1x_{1}-axis in its interior.

Claim 7.2 (points of differentiability).

There exists a two-dimensional plane PP\subseteq\mathbb{R}^{\ell} containing the x1x_{1}-axis, such that all points 0pKˇP0\neq p\in\partial\check{K}\cap P are twice differentiable points of Kˇ\partial\check{K}\subseteq\mathbb{R}^{\ell}.

Proof of Claim 7.2.

Observe first that since Kˇ\check{K} is a cone, whenever there is some point 0pKˇP0\neq p\in\partial\check{K}\cap P that is not twice differentiable, then KˇP\partial\check{K}\cap P in fact contains a 11-dimensional ray of points that are not twice differentiable.

Now suppose towards a contradiction that for every two-plane PP\subseteq\mathbb{R}^{\ell} containing the x1x_{1}-axis, there is some point 0pKˇP0\neq p\in\partial\check{K}\cap P that is not twice differentiable. Since the space of two-planes containing the x1x_{1}-axis is (2)(\ell-2)-dimensional, together with the above it follows that the set of points in Kˇ\partial\check{K} that are not twice differentiable has positive (1)(\ell-1)-dimensional Hausdorff measure. This contradicts Alexandrov’s theorem. ∎

Rotating coordinates, we can assume that PP is the x1x2x_{1}x_{2}-plane. Observe that KˇP\partial\check{K}\cap P consists of two rays R±R^{\pm} with directional vectors e±e_{\pm} satisfying e+e2>0e_{+}\cdot e_{2}>0 and ee2<0e_{-}\cdot e_{2}<0, respectively. By a space-time translation, we may also assume that t0=0t_{0}=0 and that 0M00\in M_{0} is the point in M0M_{0} with smallest x1x_{1}-value. Now, for every h>0h>0, let xh±M0{x1=h}Px_{h}^{\pm}\in M_{0}\cap\{x_{1}=h\}\cap P be a point which maximizes/minimizes the value of x2x_{2} in K0{x1=h}PK_{0}\cap\{x_{1}=h\}\cap P.

Claim 7.3 (cylindrical scale).

There exists some constant C<C<\infty such that

(7.3) suphZ(xh±)C.\sup_{h}Z(x^{\pm}_{h})\leq C.
Proof of the Claim 7.3.

We will argue similarly as in the proofs of [CHH21a, Claim 7.2], [CHH18, Proposition 5.8] and [CHHW19, Proposition 6.2].
Suppose towards a contradiction that Z(xhi±)Z(x^{\pm}_{h_{i}})\to\infty for some sequence hih_{i}\to\infty. Let i\mathcal{M}^{i} be the sequence of flows obtained by shifting xhi±x^{\pm}_{h_{i}} to the origin, and parabolically rescaling by Z(xhi±)1Z(x^{\pm}_{h_{i}})^{-1}. By [HK17, Thm. 1.14] we can pass to a subsequential limit \mathcal{M}^{\infty}, which is an ancient noncollapsed flow that is weakly convex and smooth until it becomes extinct. Note also that \mathcal{M}^{\infty} has an ×Sn\mathbb{R}^{\ell}\times S^{n-\ell} tangent flow at -\infty.

Note that \mathcal{M}^{\infty} cannot be a round shrinking ×Sn\mathbb{R}^{\ell}\times S^{n-\ell}. Indeed, if such a cylinder became extinct at time 0, this would contradict the definition of the cylindrical scale, and if it became extinct at some later time, this would contradict the fact that K0PK^{\infty}_{0}\cap P is a strict subset of PP by construction.

Thus, by Proposition 4.3 (Merle-Zaag alternative), for the flow \mathcal{M}^{\infty}, either the neutral mode is dominant or the unstable mode is dominant. If the neutral mode is dominant, then for large ii we obtain a contradiction with the fact that i\mathcal{M}^{i} has dominant unstable mode, using in particular equation (5.2). If the unstable mode is dominant, then by the fine cylindrical theorem (Theorem 6.4), the limit \mathcal{M}^{\infty} has some nonvanishing fine cylindrical vector (a1,,a)(a_{1}^{\infty},\ldots,a_{\ell}^{\infty}). However, this contradicts the fact that the fine cylindrical vector of i\mathcal{M}^{i} is obtained from the fine cylindrical vector (a1,,a)(a_{1},\ldots,a_{\ell}) of \mathcal{M} by scaling by Z(xhi+)10Z(x^{+}_{h_{i}})^{-1}\to 0. This concludes the proof of the claim. ∎

Take hih_{i}\to\infty and consider the sequence i,±:=(xhi±,0)\mathcal{M}^{i,\pm}:=\mathcal{M}-(x_{h_{i}}^{\pm},0). By Claim 7.3 (cylindrical scale), any subsequential limit ,±\mathcal{M}^{\infty,\pm} is an ancient noncollapsed flow with an ×Sn\mathbb{R}^{\ell}\times S^{n-\ell} tangent flow at -\infty. Moreover, arguing as in the proof of the claim, we see that ,±\mathcal{M}^{\infty,\pm} has a dominant unstable mode, with the same fine cylindrical vector (a1,,a)(a_{1},\ldots,a_{\ell}) as our original flow \mathcal{M}.

Claim 7.4 (splitting off lines).

The hypersurfaces M0,+M_{0}^{\infty,+} and M0,M_{0}^{\infty,-} contain (1)(\ell-1)-dimensional planes P+PP^{+}\neq P^{-}.

Proof of Claim 7.4.

First observe that since xhi±/xhi±e±R±{x_{h_{i}}^{\pm}}/{\|x_{h_{i}}^{\pm}\|}\to e_{\pm}\in R^{\pm}, the hypersurfaces M0,+M_{0}^{\infty,+} and M0,M_{0}^{\infty,-} contain a line in direction e+ee_{+}\neq e_{-}. Thus, it remains to find 2\ell-2 lines in nonradial directions.

By the definition of the blowdown, we have the Hausdorff convergence

(7.4) hi1(K{x1=hi})Kˇ{x1=1}.h_{i}^{-1}(K\cap\{x_{1}=h_{i}\})\rightarrow\check{K}\cap\{x_{1}=1\}.

This implies

(7.5) hi1xhi±q±KˇP{x1=1},h_{i}^{-1}x^{\pm}_{h_{i}}\rightarrow q^{\pm}\in\partial\check{K}\cap P\cap\{x_{1}=1\},

where q±q^{\pm} is a twice differentiable point of Kˇ\partial\check{K} by Claim 7.2.

Now, let ν\nu be the inwards unit normal of Kˇ{x1=1}1\check{K}\cap\{x_{1}=1\}\subset\mathbb{R}^{\ell-1} at q=q±q=q^{\pm}, and consider any unit tangent vector wTq(Kˇ{x1=1})w\in T_{q}(\partial\check{K}\cap\{x_{1}=1\}). Since Kˇ\check{K} is twice differentiable at qq, we have

(7.6) qλ:=q+λw+|λ|3/2νInt(Kˇ)q_{\lambda}:=q+\lambda w+|\lambda|^{3/2}\nu\in\textrm{Int}(\check{K})

for 0<|λ|<λ00<|\lambda|<\lambda_{0} small enough. By convexity, the segments from qq to qλq_{\lambda} and qλq_{-\lambda} are contained in Kˇ\check{K}. Note also that these segments have length at least λ\lambda and meet at angle πε(λ)\pi-\varepsilon(\lambda), where ε0\varepsilon\to 0 as λ0\lambda\to 0. Together with the Hausdorff convergence from above, we see that for ii large enough the segments connecting xhix_{h_{i}} to hiqλh_{i}q_{\lambda} and hiqλh_{i}q_{-\lambda} are contained in KK, have length at least hiλh_{i}\lambda and meet at angle πε(λ)\pi-\varepsilon(\lambda). Taking hih_{i}\to\infty, and sending λ=λi0\lambda=\lambda_{i}\to 0 slowly enough, we conclude that M0,±M_{0}^{\infty,\pm} contains a line in direction ww. Since ww was arbitrary, this completes the proof of the claim. ∎

By Claim 7.4 (splitting off lines), we get Mt,±=P±×Nt±M_{t}^{\infty,\pm}=P^{\pm}\times N_{t}^{\pm}, where Nt±N_{t}^{\pm} is an ancient noncollapsed uniformly 22-convex flow. Using the classification by Brendle-Choi [BC], we infer that Nt±N_{t}^{\pm} must be a translating bowl soliton. By inspection, we see that the fine cylindrical vector of P±×Nt±P^{\pm}\times N_{t}^{\pm} points in the direction of translation. However, since Nt+N_{t}^{+} and NtN_{t}^{-} translate in different directions, this contradicts the fact that ,+\mathcal{M}^{\infty,+} and ,\mathcal{M}^{\infty,-} have the same fine cylindrical vector. This concludes the proof of the theorem. ∎


References

  • [ADS19] S. Angenent, P. Daskalopoulos, and N. Sesum. Unique asymptotics of ancient convex mean curvature flow solutions. J. Differential Geom., 111(3):381–455, 2019.
  • [ADS20] S. Angenent, P. Daskalopoulos, and N. Sesum. Uniqueness of two-convex closed ancient solutions to the mean curvature flow. Ann. of Math. (2), 192(2):353–436, 2020.
  • [And12] B. Andrews. Noncollapsing in mean-convex mean curvature flow. Geom. Topol., 16(3):1413–1418, 2012.
  • [BC] S. Brendle and K. Choi. Uniqueness of convex ancient solutions to mean curvature flow in higher dimensions. Geom. Top. (to appear).
  • [BC19] S. Brendle and K. Choi. Uniqueness of convex ancient solutions to mean curvature flow in 3\mathbb{R}^{3}. Invent. Math., 217(1):35–76, 2019.
  • [Bre15] S. Brendle. A sharp bound for the inscribed radius under mean curvature flow. Invent. Math., 202(1):217–237, 2015.
  • [CHH18] K. Choi, R. Haslhofer, and O. Hershkovits. Ancient low entropy flows, mean convex neighborhoods, and uniqueness. arXiv:1810.08467, 2018.
  • [CHH21a] K. Choi, R. Haslhofer, and O. Hershkovits. A nonexistence result for wing-like mean curvature flows in 4\mathbb{R}^{4}. arXiv:2105.13100, 2021.
  • [CHH21b] K. Choi, R. Haslhofer, and O. Hershkovits. Classification of noncollapsed translators in 4\mathbb{R}^{4}. preprint, 2021.
  • [CHHW19] K. Choi, R. Haslhofer, O. Hershkovits, and B. White. Ancient asymptotically cylindrical flows and applications. arXiv:1910.00639, 2019.
  • [CHN13] J. Cheeger, R. Haslhofer, and A. Naber. Quantitative stratification and the regularity of mean curvature flow. Geom. Funct. Anal., 23(3):828–847, 2013.
  • [CM15] T. Colding and W. Minicozzi. Uniqueness of blowups and Lojasiewicz inequalities. Ann. of Math. (2), 182(1):221–285, 2015.
  • [CM19] K. Choi and C. Mantoulidis. Ancient gradient flows of elliptic functionals and Morse index. arXiv:1902.07697, 2019.
  • [HH18] R. Haslhofer and O. Hershkovits. Singularities of mean convex level set flow in general ambient manifolds. Adv. Math., 329:1137–1155, 2018.
  • [HIMW19] D. Hoffman, T. Ilmanen, F. Martín, and B. White. Graphical translators for mean curvature flow. Calc. Var. Partial Differential Equations, 58(4):Art. 117, 29, 2019.
  • [HK15] R. Haslhofer and B. Kleiner. On Brendle’s estimate for the inscribed radius under mean curvature flow. Int. Math. Res. Not. IMRN, (15):6558–6561, 2015.
  • [HK17] R. Haslhofer and B. Kleiner. Mean curvature flow of mean convex hypersurfaces. Comm. Pure Appl. Math., 70(3):511–546, 2017.
  • [Hui90] G. Huisken. Asymptotic behavior for singularities of the mean curvature flow. J. Differential Geom., 31(1):285–299, 1990.
  • [Ilm03] T. Ilmanen. Problems in mean curvature flow. https://people.math.ethz.ch/ ilmanen/classes/eil03/problems03.ps, 2003.
  • [MZ98] F. Merle and H. Zaag. Optimal estimates for blowup rate and behavior for nonlinear heat equations. Comm. Pure Appl. Math., 51(2):139–196, 1998.
  • [SW09] W. Sheng and X. Wang. Singularity profile in the mean curvature flow. Methods Appl. Anal., 16(2):139–155, 2009.
  • [Whi00] B. White. The size of the singular set in mean curvature flow of mean convex sets. J. Amer. Math. Soc., 13(3):665–695, 2000.
  • [Whi03] B. White. The nature of singularities in mean curvature flow of mean-convex sets. J. Amer. Math. Soc., 16(1):123–138, 2003.
  • [Whi15] B. White. Subsequent singularities in mean-convex mean curvature flow. Calc. Var. Partial Differential Equations, 54(2):1457–1468, 2015.
  • [Zhu21] J. Zhu. Rotational symmetry of uniformly 3-convex translating solitons of mean curvature flow in higher dimensions. arXiv:2103.16382, 2021.

Wenkui du, Department of Mathematics, University of Toronto, 40 St George Street, Toronto, ON M5S 2E4, Canada

Robert Haslhofer, Department of Mathematics, University of Toronto, 40 St George Street, Toronto, ON M5S 2E4, Canada