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

Genus one singularities in mean curvature flow

Adrian Chun-Pong Chu  and  Ao Sun The University of Chicago, Department of Mathematics, Eckhart Hall, 5734 S University Ave, Chicago, IL, 60637
Current Address: Cornell University, Malott Hall, 212 Garden Ave, Ithaca, NY 14853
[email protected] Lehigh University, Department of Mathematics, Chandler-Ullmann Hall, Bethlehem, PA 18015 [email protected]
Abstract.

We show that for certain one-parameter families of initial conditions in 3\mathbb{R}^{3}, when we run mean curvature flow, a genus one singularity must appear in one of the flows. Moreover, such a singularity is robust under perturbation of the family of initial conditions. This contrasts sharply with the case of just a single flow. As an application, we construct an embedded, genus one self-shrinker with entropy lower than a shrinking doughnut.

1. Introduction

Mean curvature flow (MCF) is the most rapid process to decrease the area of a surface. With an initial motivation from applied science, this geometric evolution equation has gained much interest recently due to its potential for studying the geometry and topology of surfaces embedded in three-manifolds. As a nonlinear geometric heat flow, MCF may have singularities, which may lead to changes in the geometry and topology of the surfaces.

The blow-up method, pioneered by Huisken [Hui90], Ilmanen [Ilm95], and White [Whi97], shows that the singularities are modeled by a special class of surfaces called self-shrinkers. They satisfy the equation H+x/2=0\vec{H}+\vec{x}^{\perp}/2=0. Determining the possible singularity models that can arise in an arbitrary MCF is a challenging problem. With the convexity assumption, Huisken [Hui84] proved that the singularities must be modeled by spheres. With the mean convexity assumption, White [Whi97, Whi00, Whi03] proved that the singularities must be modeled by spheres and cylinders. However, in the absence of curvature assumptions, the question of which type of singularities must arise in MCF remains widely open. In this paper, we find a condition that guarantees the appearance of a singularity modeled by a genus one self-shrinker. To the best of our knowledge, this is the first resultthat produces a singularity, that appears in a non-self-shrinking flow and is modeled by a self-shrinker of non-zero genus.

Refer to caption

Figure 1.

Let us first explain the heuristics, which involves an interpolation argument. In Figure 1, we have a one-parameter family {Ms}s[0,1]\{M^{s}\}_{s\in[0,1]} of tori in the top row. Suppose that the initial torus M0M^{0} has a thin “inward neck,” which will eventually pinch under the MCF. On the other hand, the final torus M1M^{1} has a thin “outward neck” in the middle, which will also pinch under MCF. Then, there should exist a critical value s0[0,1]s_{0}\in[0,1] such that for the torus Ms0M^{s_{0}}, both the inward and outward necks pinch under MCF, giving rise to a genus one singularity.

The following is our main theorem. We will provide a precise definition of “inward (or outward) torus neck will pinch” later in Definition 1.8.

Theorem 1.1.

Let {Ms}s[0,1]\{M^{s}\}_{s\in[0,1]} be a smooth family of tori in 3\mathbb{R}^{3} such that for the MCF starting from M0M^{0} (resp. M1M^{1}), the inward (resp. outward) torus neck will pinch. Then there exists s0[0,1]s_{0}\in[0,1] such that the MCF starting from Ms0M^{s_{0}} would develop a singularity that is not multiplicity one cylindrical or multiplicity one spherical.

Note that, in precise terms, by MCF we actually refer the level set flow (see §2). In fact, before the flow encounter a genus one singularity, it is possible that it passed through some cylindrical singularities or spherical singularities. We also remark that Brendle [Bre16] proved that the only genus 0 self-shrinkers are the spheres and the cylinders. In contrast, there are many higher genus self-shrinkers, as constructed in [Ang92, Ngu14, KKM18, Møl11, SWZ20], among others.

Now, immediately, we can exclude the possibility of multiplicity if the entropy of each torus MsM^{s} is less than 22. The entropy of a surface Σ\Sigma was defined by Colding-Minicozzi [CM12]:

Ent(Σ):=supx03,t0>0(4πt0)1Σe|xx0|24t0.\mathrm{Ent}(\Sigma)\mathrel{\mathop{\mathchar 58\relax}}=\sup_{x_{0}\in\mathbb{R}^{3},t_{0}>0}(4\pi t_{0})^{-1}\int_{\Sigma}e^{-\frac{|x-x_{0}|^{2}}{4t_{0}}}.
Corollary 1.2.

In the setting of Theorem 1.1, if each initial torus MsM^{s} has entropy less than 22, then at the singularity concerned, every tangent flow is given by a multiplicity one, embedded, genus one self-shrinker.

Recall that the tangent flow represents a specific blow-up limit of a MCF at a singularity, as discussed in §2.2. By employing Huisken’s monotonicity formula [Hui90], Ilmanen [Ilm95], and White [Whi97] proved that the tangent flow must be a self-shrinker with multiplicity.

Let us now explicitly provide a family of tori that satisfies the assumption of Corollary 1.2. Consider the rotationally symmetric, compact, genus one self-shrinker in 3\mathbb{R}^{3} constructed by Drugan-Nguyen [DN18], which we will denote by 𝕋\mathbb{T}. It is worth noting that both 𝕋\mathbb{T} and the Angenent torus [Ang92] are referred to as shrinking doughnuts, and they may be the same. It was shown that 𝕋\mathbb{T} has entropy strictly less than 22 [DN18], while Berchenko-Kogan [BK21] provided numerical evidence that the Angenent torus has an entropy of approximately 1.851.85.

Theorem 1.3.

Let {Ms}s[0,1]\{M^{s}\}_{s\in[0,1]} be a smooth family of tori in 3\mathbb{R}^{3} that are sufficiently close in CC^{\infty} to the shrinking doughnut 𝕋\mathbb{T}, with M0M^{0} strictly inside 𝕋\mathbb{T} while M1M^{1} strictly outside. Then there exists s0[0,1]s_{0}\in[0,1] such that the MCF starting from Ms0M^{s_{0}} would develop a singularity at which every tangent flow is given by a multiplicity one, embedded, genus one self-shrinker.

The idea of Theorem 1.3 can be traced back to the work of Lin and the second author in [LS22]. In earlier work, Colding-Ilmanen-Minicozzi-White [CIMIW13] observed that one can perturb a closed embedded self-shrinker in 3\mathbb{R}^{3} such that the MCF has only neck and spherical singularities. Lin and the second author observed a bifurcation phenomenon: Inward (resp. outward) perturbations cause the MCF pinch from inside (resp. outside). After we completed this manuscript, we were notified by the anonymous referee that the idea of Theorem 1.1 has been discussed and explained orally by Edelen and White.

It is also interesting to compare our results with the recent developments in generic MCF [CM12, CCMS20, CCMS21, SX21a, SX21b, CCS23, Sun23]: One can perturb a single MCF to avoid a singularity that is not spherical or cylindrical. In contrast, our results imply that for a certain one-parameter family of MCFs, a singularity that is modeled by a genus one shrinker remains robust under perturbations.

It is natural to ask whether Theorem 1.1 extends to surfaces with genus two or above. Actually, it would not: See a counterexample in Remark 5.2. Nevertheless, a similar theory might be established for a multi-parameter family of higher genus surfaces (see Question 1.10).

Let us now present several applications of the above theorems.

Theorem 1.4.

An embedded, genus one self-shrinker in 3\mathbb{R}^{3} of the least entropy either is non-compact or has index 55.

Note that the existence of an entropy minimizer among all embedded, genus gg self-shrinkers in 3\mathbb{R}^{3}, with a fixed gg, was proved by Sun-Wang [SW20].

Theorem 1.5.

There exists an ancient MCF through cylindrical and spherical singularities {M(t)}t<0\{M(t)\}_{t<0} in 3\mathbb{R}^{3} such that:

  • As tt\to-\infty, 1tM(t)𝕋\frac{1}{\sqrt{-t}}M(t)\to\mathbb{T} smoothly.

  • As t0t\to 0, M(t)M(t) hits a singularity at which every tangent flow is given by a multiplicity one, embedded, genus one self-shrinker of lower entropy than 𝕋\mathbb{T}.

In fact, Theorem 1.5 remains valid even with 𝕋\mathbb{T} replaced by any other closed, embedded, rotationally symmetric, genus one shrinker (if they indeed exist), and the same proof will hold.

Recalling that the rotationally symmetric shrinker 𝕋\mathbb{T} must have index of at least 77, as shown by Liu [Liu16], we can deduce the following corollary from Theorem 1.4 and 1.5.

Corollary 1.6.

There exists an embedded, genus one self-shrinker in 3\mathbb{R}^{3} with entropy lower than 𝕋\mathbb{T}.

Finally, the three self-shrinkers in 3\mathbb{R}^{3} with the lowest entropy are the plane, the sphere, and the cylinder ([CIMIW13, BW17]). Notably, all three of them are rotationally symmetric. Kleene-Møller [KMl14] proved that all other rotationally symmetric smooth embedded self-shrinkers are closed with genus 11.

Now, the space of smooth embedded self-shrinkers in 3\mathbb{R}^{3} with entropy less than some constant δ<2\delta<2 is known to be compact in the ClocC_{\text{loc}}^{\infty} topology (see [Lee21]). Together with the rigidity of the cylinder as a self-shrinker by [CIM15], there exists a smooth embedded self-shrinker that minimizes entropy among all smooth embedded self-shrinkers with entropy larger than that of the cylinder.

Corollary 1.7.

A smooth embedded self-shrinker in 3\mathbb{R}^{3} with the fourth lowest entropy is not rotationally symmetric.

1.1. Main ideas: Change in homology under MCF

The major challenge of this paper is to introduce some new concepts to rigorously state and prove the interpolation argument we outlined on page 1 and Figure 1. Particularly, it is crucial to describe the topological change of the surfaces more precisely. Let ={M(t)}t0\mathcal{M}=\{M(t)\}_{t\geq 0} be a MCF in 3\mathbb{R}^{3}, where the initial condition M(0)M(0) is a closed, smooth, embedded surface. Since we would allow M(t)M(t) to have singularities and thus change its topology, \mathcal{M} is, more precisely, a level set flow. In this paper, we often use the phrases MCF and level set flow interchangeably.

It is known that the topology of M(t)M(t) simplifies over time. In [Whi95], White focused on describing the complement 3\M(t)\mathbb{R}^{3}\backslash M(t) (instead of M(t)M(t) itself), and how it changes over time. For example, he showed rank(H1(3\M(t)))\mathrm{rank}(H_{1}(\mathbb{R}^{3}\backslash M(t))) is non-increasing in tt, where H1H_{1} denotes the first homology group in \mathbb{Z}-coefficients. Therefore, heuristically, the topology can only be destroyed but not created during the evolution of the surface.

In this paper, we will further describe this phenomenon by keeping track of which elements of the initial homology group H1(3\M(0))H_{1}(\mathbb{R}^{3}\backslash M(0)) are destroyed, and how they are destroyed. To illustrate, let us use the flow depicted in Figure 2 as an example.

Refer to caption

Figure 2.
1.1.1. Heuristic observation

Let us begin by providing some heuristic observations regarding Figure 2. We will elaborate on them more precisely shortly. We fix four elements of H1(3\M(0))H_{1}(\mathbb{R}^{3}\backslash M(0)) at time t=0t=0, as shown in the figure. Note that a0a_{0} and a1a_{1} are in the bounded region inside the genus two surface M(0)M(0), whereas b0b_{0} and b1b_{1} are in the region outside M(0)M(0).

  1. (1)

    At time t=T1t=T_{1}, a0a_{0} is “broken” by the cylindrical singularity xx of the flow. As a result, for later time t>T1t>T_{1}, a0a_{0} no longer exists. Apparently, it “terminates” at time T1T_{1}.

  2. (2)

    On the other hand, a1a_{1}, b0b_{0}, and b1b_{1} all can survive through time T1T_{1}. For example, for b0b_{0}, we can clearly have a continuous family of loops, {βt}t0\{\beta_{t}\}_{t\geq 0}, where [β0]=b0[\beta_{0}]=b_{0} and each βt\beta_{t} is a loop outside the surface M(t)M(t). In this sense, b0b_{0} will survive for all time, although it becomes trivial after time T1T_{1}.

  3. (3)

    As for b1b_{1}, although it survives through t=T1t=T_{1}, it will terminate at t=T2t=T_{2}, when it is broken by the cylindrical singularity yy.

Let us now provide precise descriptions of these observations.

1.1.2. Three new concepts

To our knowledge, these concepts are new, but they seem natural in the context of geometric flows. We believe these concepts may hold independent interest as well.

To set up, for any two times t1<t2t_{1}<t_{2}, let us consider the complement of the spacetime track of the flow within the time interval [t1,t2][t_{1},t_{2}]:

W[t1,t2]:=t[t1,t2](3\M(t))×{t}3×[t1,t2].W[t_{1},t_{2}]\mathrel{\mathop{\mathchar 58\relax}}=\bigcup_{t\in[t_{1},t_{2}]}(\mathbb{R}^{3}\backslash M(t))\times\{t\}\subset\mathbb{R}^{3}\times[t_{1},t_{2}].

In order to discuss the “termination” of an element c0H1(3\M(0))c_{0}\in H_{1}(\mathbb{R}^{3}\backslash M(0)) under the flow, we first need to relate elements of H1(3\M(0))H_{1}(\mathbb{R}^{3}\backslash M(0)) and elements of H1(3\M(t))H_{1}(\mathbb{R}^{3}\backslash M(t)) at some later time t>0t>0.

Homology descent. (Definition 3.1.) Given two elements c0H1(3\M(0))c_{0}\in H_{1}(\mathbb{R}^{3}\backslash M(0)) and cH1(3\M(t))c\in H_{1}(\mathbb{R}^{3}\backslash M(t)) with t>0t>0, we say that cc descends from c0c_{0}, and denote

c0c,c_{0}\succ c,

if the following holds: For every representative γ0c0\gamma_{0}\in c_{0} and γc\gamma\in c, if we view them as subsets

γ0(3\M(0))×{0},γ(3\M(t))×{t},\gamma_{0}\subset(\mathbb{R}^{3}\backslash M(0))\times\{0\},\;\;\gamma\subset(\mathbb{R}^{3}\backslash M(t))\times\{t\},

then they bound some singular 2-chain ΓW[0,t]\Gamma\subset W[0,t], i.e. γ0γ=Γ.\gamma_{0}-\gamma=\partial\Gamma. (See Figure 3.)

Refer to caption

Figure 3.

As we will prove, the above notion satisfies some desirable properties. For example, given a c0H1(3\M(0))c_{0}\in H_{1}(\mathbb{R}^{3}\backslash M(0)), the element cH1(3\M(t))c\in H_{1}(\mathbb{R}^{3}\backslash M(t)) described above, if exists, turns out to be unique. Consequently, we denote this unique element as c0(t)c_{0}(t).

This enables us to further define:

Homology termination. (Definition 3.8.) Let c0H1(3\M(0))c_{0}\in H_{1}(\mathbb{R}^{3}\backslash M(0)). If

𝔱(c0):=sup{t0:c0c for some cH1(3\M(t))}\mathfrak{t}(c_{0})\mathrel{\mathop{\mathchar 58\relax}}=\sup\{t\geq 0\mathrel{\mathop{\mathchar 58\relax}}c_{0}\succ c\textrm{ for some }c\in H_{1}(\mathbb{R}^{3}\backslash M(t))\}

is finite, then we say that c0c_{0} terminates at time 𝔱(c0)\mathfrak{t}(c_{0}).

For instance, in Figure 2, we observe that a0a_{0} terminates at time T1T_{1}, and b1b_{1} terminates at time T2T_{2}. However, b0b_{0} never terminates, despite the fact that b0(t)b_{0}(t) becomes trivial for t>T1t>T_{1}. Similarly, a1a_{1} also never terminates, even though a1(t)a_{1}(t) becomes trivial for t>T2t>T_{2}. Note that a1a_{1} would not terminate at time T3T_{3}: For any t>T3t>T_{3}, any loop in 3\M(t)=3\mathbb{R}^{3}\backslash M(t)=\mathbb{R}^{3} would bound a disc in 3\mathbb{R}^{3}, so it follows easily that for any loop γ0a1\gamma_{0}\in a_{1} and loop γ3×{t}\gamma\subset\mathbb{R}^{3}\times\{t\}, γ0γ\gamma_{0}-\gamma would bound some 2-dimensional chain in the complement of the spacetime track.

Finally, we can describe what “a0a_{0} breaks at a cylindrical singularity xx” means.

Homology breakage. (Definition 3.12.) Let c0H1(3\M(0))c_{0}\in H_{1}(\mathbb{R}^{3}\backslash M(0)), T>0T>0, and xM(T)x\in M(T). Suppose the following holds:

  • For each t[0,T)t\in[0,T), the element c0(t)H1(3\M(t))c_{0}(t)\in H_{1}(\mathbb{R}^{3}\backslash M(t)) (such that c0c0(t)c_{0}\succ c_{0}(t)) exists.

  • For every neighborhood U3U\subset\mathbb{R}^{3} of xx, for each t<Tt<T sufficiently close to TT, every element of c0(t)c_{0}(t) intersects UU.

Then we say that c0c_{0} breaks at (x,T)(x,T). (See Figure 4.)

Refer to caption

Figure 4. The picture at time tt, for all t<Tt<T sufficiently close to TT.

For example, in Figure 2, a0a_{0} breaks at (x,T1)(x,T_{1}), while b1b_{1} breaks at (y,T2)(y,T_{2}).

As we will see, these three new concepts are quite useful and satisfy several nice properties. Here are a few examples:

  • A homology class cannot break at a regular point, nor a spherical singularity of the flow (Proposition 3.14 and 3.15).

  • If the initial condition M(0)M(0) is a closed surface of non-zero genus, then some initial homology class must terminate at finite time (Remark 4.10).

  • Suppose {M(t)}t0\{M(t)\}_{t\geq 0} is a MCF with only spherical and cylindrical singularities. If a homology class terminates at some time TT, then it must break at (x,T)(x,T) for some cylindrical singularity xM(T)x\in M(T) (Theorem 4.5).

These properties are all crucial in proving the main theorems.

Finally, let us provide a precise definition of “inward (or outward) torus neck will pinch” in Theorem 1.1.

Definition 1.8.

Given a torus MM in 3\mathbb{R}^{3}, let a0a_{0} (resp. b0b_{0}) be a generator of the first homology group of the interior (resp. exterior) region of MM, which is isomorphic to \mathbb{Z} (see Figure 5). We say that the inward (resp. outward) torus neck of MM will pinch if a0a_{0} (resp. b0b_{0}) will terminate under MCF.

Refer to caption

Figure 5.

Clearly, a0a_{0} (and b0b_{0}) is unique up to a sign, and the above notion is independent of which sign we choose.

1.2. Structure of cylindrical singularities

Once we establish the topological concepts to keep track of the homology classes under the MCF, another challenge arises: We need to understand what happens to these homology classes as the MCF encounters the cylindrical singularities.

Intuitively, a cylindrical singularity is just like a neck, and as we approach the singular time, the neck pinches as in Figure 1. However, the actual situation can be much more complicated. For example, consider the MCF of the boundary of a tubular neighborhood of a rotationally symmetric S1S^{1} in 3\mathbb{R}^{3}. It will shrink to a singular set that is a rotationally symmetric S1S^{1}, where each singular point is cylindrical, but it does not look like a neck pinching.

First, one has the partial regularity of the singular set of cylindrical singularities, studied by White [Whi97] and Colding-Minicozzi [CM15, CM16]. This allows us to control the singular set. We can establish the compactness of the singular set of cylindrical singularities that are inward (or outward), and know that they only appear for a zero-measured set of time.

Another important theory is the mean convex neighborhood theory of cylindrical singularities by Choi-Haslhofer-Hershkovits [CHH22], and a generalized version by Choi-Haslhofer-Hershkovits-White [CHHW22]. In these works, they classified the possible limit flows at a cylindrical singularity. As a consequence, they derived a canonical neighborhood theorem at a cylindrical singularity, which describes the local behavior of the MCF.

We will study the local behavior of MCF at cylindrical singularities based on these two theories. Nevertheless, the particular local behavior we need to understand does not directly come from [CHH22, CHHW22]. We present these relevant results in §2.3.

1.3. Outline of proofs

1.3.1. Theorem 1.1

We will prove by contradiction. For each s[0,1]s\in[0,1], let s={Ms(t)}t0\mathcal{M}^{s}=\{M^{s}(t)\}_{t\geq 0} be the MCF (more precisely, a level set flow) with Ms(0)=MsM^{s}(0)=M^{s} as its initial condition. Let a0a_{0} (resp. b0)b_{0}) be a generator of the first homology group of the inside (resp. outside) region of each torus MsM^{s} (recall Definition 1.8). Assuming that Theorem 1.1 were false, s\mathcal{M}^{s} would be a MCF through cylindrical and spherical singularities for each ss. This flow is unique and well-defined by Choi-Haslhofer-Hershkovits [CHH22]. Next, we show that for each ss, either a0a_{0} or b0b_{0} will terminate, but not both. This claim relies on the fact, mentioned above, that if a homology class will terminate, it must break at a neck singularity. This crucial fact is established based on the mean convex neighborhood theorem and the canonical neighborhood theorem by Choi-Haslhofer-Hershkovits-White [CHH22, CHHW22].

Thus, we can partition [0,1][0,1] into a disjoint union ABA\sqcup B, where AA is the set of ss for which a0a_{0} will terminate, and BB is the set of ss for which b0b_{0} will terminate. Furthermore, we will show that AA and BB are both closed sets. Recall that we are given 0A0\in A and 1B1\in B. Since [0,1][0,1] is a connected interval, this leads to a contradiction.

1.3.2. Theorem 1.3

We can apply Theorem 1.1 to prove Theorem 1.3, provided that we can show the inward torus neck will pinch (i.e., a0a_{0} will terminate) for the starting flow (s=0s=0), and the outward torus neck will pinch (i.e., b0b_{0} will terminate) for the ending flow (s=1s=1). To prove, for instance, that a0a_{0} will terminate for the starting flow, we recall that M0(0)M^{0}(0) lies strictly inside the shrinker Σ\Sigma. Then we will run MCF to these two surfaces and use the avoidance principle, which states that the distance between the two surfaces will increase, to conclude that a0a_{0} must terminate.

1.3.3. Theorem 1.4

Let Σ\Sigma be an embedded, genus one shrinker with the least entropy. Suppose by contradiction that it is compact with index at least 66. Disregarding the four (orthogonal) deformations induced by translation and scaling, there are still two other deformations that decrease the entropy, one of which is the one-sided deformation given by the first eigenfunction of the Jacobi operator. Thus, we can construct a one-parameter family of tori with entropy less than Σ\Sigma, such that the starting torus is inside Σ\Sigma, and the ending torus is outside Σ\Sigma. Then, as in the proof of Theorem 1.3, we apply Theorem 1.1 to obtain another genus one shrinker with less entropy than Σ\Sigma. This contradicts the definition of Σ\Sigma.

1.3.4. Theorem 1.5

According to Liu [Liu16], the shrinking doughnut 𝕋\mathbb{T} has an index of at least 77. Consequently, based on the result of Choi-Mantoulidis [CM22], there exists a one-parameter family of ancient rescaled MCF originating from 𝕋\mathbb{T} that decreases the entropy. As before, we can apply Theorem 1.1 to immediately obtain the desired genus one, self-shrinking tangent flow with lower entropy.

1.4. Open questions

We propose several open problems. The first one is motivated by generic MCF and min-max theory.

Conjecture 1.9.

There exists an embedded, genus one, index 55 self-shrinker in 3\mathbb{R}^{3} that is the “second most generic” one.

We say a self-shrinker Σ\Sigma is the “second most generic”, after the generic ones (the cylinder and the sphere), in the following sense: Suppose we have a one-parameter family of embedded surfaces {Ms}s[0,1]\{M^{s}\}_{s\in[0,1]} in 3\mathbb{R}^{3}. Then, we can perturb this family such that when we run MCF for every MsM^{s}, every singularity is either cylindrical, spherical, or modeled by Σ\Sigma.

Note that Theorem 1.4 and its proof can be seen as evidence of a very “local” version of this conjecture: They say that any closed, embedded, genus one self-shrinker with an index of at least 66 is not the second most generic.

Now, we note that Theorem 1.1 does not hold for initial conditions with genus greater than one, see Remark 5.2.

Question 1.10.

Can Theorem 1.1 be generalized to the higher genus case, possibly by considering higher parameter families of initial conditions?

Finally, notice that many concepts that we introduce in this paper heavily rely on the extrinsic structure of mean curvature flow.

Question 1.11.

Can the concepts of homology descent, homology termination, and homology breakage be adapted to the setting of Ricci flow?

1.5. Organizations.

In §2, we will introduce the preliminary materials, including a refined canonical neighborhood theorem. In §3, we will define the concepts of homology descent, homology termination, and homology breakage, and prove some relevant basic propositions. In §4, we focus on the case of MCF through cylindrical and spherical singularities, with torus as the initial condition. In §5, we prove the main theorems.

Acknowledgement

We would like to thank Professor André Neves for all the fruitful discussions and his constant support. We are grateful to Zhihan Wang for the valuable conversations. And the first author would also like to thank Chi Cheuk Tsang for the helpful discussions. We are also grateful to anonymous referees for many helpful comments and suggestions, especially the work by Edelen and White.

2. Preliminaries

In §2 we will set up the language and provide the necessary background to define MCF through cylindrical and spherical singularities.

The classical mean curvature flow is a family of hypersurfaces {M(t)}t[0,T)\{M(t)\}_{t\in[0,T)} in n+1\mathbb{R}^{n+1} satisfying the equation

(1) tx=H(x),\partial_{t}x=\vec{H}(x),

where xx is the position vector and H\vec{H} is the mean curvature vector. When the hypersurface is not C2C^{2}, we can not define the mean curvature flow using this PDE, and we need to use some weak notions to define the flow.

2.1. Weak solutions of MCF

Throughout this paper, we will focus on two different types of weak solutions of MCF. One is a set-theoretic weak solution defined by the level set flow, and another one is a geometric measure theoretic weak solution called Brakke flow. Readers interested in detailed discussions of level set flows can refer to [ES91, Ilm92], while those interested in Brakke flow can refer to [Bra78, Ilm94].

The level set flow equation is a degenerate parabolic equation

(2) tu=Δu(D2u(Du,Du)|Du|2).\partial_{t}u=\Delta u-\left(\frac{D^{2}u(Du,Du)}{|Du|^{2}}\right).

Suppose M(0)M(0) is a closed hypersurface in n+1\mathbb{R}^{n+1}, then if u(,t)u(\cdot,t) solves (2) with M(0)={xn+1:u(,0)=0}M(0)=\{x\in\mathbb{R}^{n+1}\mathrel{\mathop{\mathchar 58\relax}}u(\cdot,0)=0\}, then M(t):={xn+1:u(,0)=0}M(t)\mathrel{\mathop{\mathchar 58\relax}}=\{x\in\mathbb{R}^{n+1}\mathrel{\mathop{\mathchar 58\relax}}u(\cdot,0)=0\} can be viewed as a weak solution to MCF. In particular, when M(t)M(t) is smooth, this weak solution coincides with the classical solution of MCF.

The level set flow was introduced by Osher-Sethian in [OS88]. Chen-Giga-Goto [CGG91] and Evans-Spruck [ES91] introduced the viscosity solutions to equation (2), and these solutions are Lipschitz. Throughout this paper, when we refer to a level set function or a solution to the level set flow equation, we mean a viscosity solution to equation (2).

The set-theoretic solution of a MCF will be called the level set flow or biggest flow. These notions are used by Ilmanen [Ilm92] and White [Whi95, Whi00, Whi03]. The term “biggest flow” is used to avoid ambiguity when dealing with weak solutions for noncompact flows. Such a weak solution may have a nonempty interior. In this case, we say the level set flow fattens.


Brakke flow is defined using geometric measure theory. Let XX be a complete manifold without boundary. The Brakke flow is a family of Radon measures {μt}t0\{\mu_{t}\}_{t\geq 0}, such that for any test function ϕCc2(X)\phi\in C_{c}^{2}(X) with ϕ0\phi\geq 0,

lim supstμs(ϕ)μt(ϕ)st(ϕH2+H)𝑑μt,\limsup_{s\to t}\frac{\mu_{s}(\phi)-\mu_{t}(\phi)}{s-t}\leq\int(-\phi H^{2}+\nabla^{\perp}\cdot\vec{H})d\mu_{t},

where H\vec{H} is the mean curvature vector of μt\mu_{t} whenever μt\mu_{t} is rectifiable and has L2L^{2}-mean curvature in the varifold sense. Otherwise, the right-hand side is defined to be -\infty.

In general, the Brakke flow starting from a given initial data is not unique. We will be interested in unit regular cyclic integral Brakke flows. For detailed discussions on these notions, we refer the readers to [Whi09]. The existence of such a flow starting from a smooth surface is guaranteed by Ilmanen’s elliptic regularization, see [Ilm94]. These flows have a well-established compactness theory.

2.2. Setting and notations

Let M(0)M(0) be a closed smooth nn-dimensional hypersurface in n+1\mathbb{R}^{n+1} that bounds a compact set Kin(0)K_{\mathrm{in}}(0). Let Kout(0)=n+1\Kin(0)¯K_{\mathrm{out}}(0)=\overline{\mathbb{R}^{n+1}\backslash K_{\mathrm{in}}(0)}. Now, denote by

{M(t)}t0,{Kin(t)}t0, and {Kout(t)}t0\{M(t)\}_{t\geq 0},\{K_{\mathrm{in}}(t)\}_{t\geq 0},\textrm{ and }\{K_{\mathrm{out}}(t)\}_{t\geq 0}

respectively the level set flow (i.e. the biggest flow) with initial condition M(0),Kin(0)M(0),K_{\mathrm{in}}(0), and Kout(0)K_{\mathrm{out}}(0). Then we define their spacetime tracks

\displaystyle\mathcal{M} ={(x,t):xM(t),t0},\displaystyle=\{(x,t)\mathrel{\mathop{\mathchar 58\relax}}x\in M(t),t\geq 0\},
𝒦in\displaystyle\mathcal{K}_{\mathrm{in}} ={(x,t):xKin(t),t0},\displaystyle=\{(x,t)\mathrel{\mathop{\mathchar 58\relax}}x\in K_{\mathrm{in}}(t),t\geq 0\},
𝒦out\displaystyle\mathcal{K}_{\mathrm{out}} ={(x,t):xKout(t),t0}.\displaystyle=\{(x,t)\mathrel{\mathop{\mathchar 58\relax}}x\in K_{\mathrm{out}}(t),t\geq 0\}.

We then define the inner flow of M(0)M(0),

Min(t)={x:(x,t)𝒦in}M_{\mathrm{in}}(t)=\{x\mathrel{\mathop{\mathchar 58\relax}}(x,t)\in\partial\mathcal{K}_{\mathrm{in}}\}

and the outer flow of M(0)M(0),

Mout(t)={x:(x,t)𝒦out}.M_{\mathrm{out}}(t)=\{x\mathrel{\mathop{\mathchar 58\relax}}(x,t)\in\partial\mathcal{K}_{\mathrm{out}}\}.
Lemma 2.1.

Let u:n+1×[0,)u\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}^{n+1}\times[0,\infty)\to\mathbb{R} be a level set function of \mathcal{M}, with u(,0)0u(\cdot,0)\leq 0 on Kin(0)K_{\mathrm{in}}(0). Then

n+1\Kin(t)={x:u(x,t)>0},n+1\Kout(t)={x:u(x,t)<0}.\mathbb{R}^{n+1}\backslash K_{\mathrm{in}}(t)=\{x\mathrel{\mathop{\mathchar 58\relax}}u(x,t)>0\},\;\;\mathbb{R}^{n+1}\backslash K_{\mathrm{out}}(t)=\{x\mathrel{\mathop{\mathchar 58\relax}}u(x,t)<0\}.
Proof.

For the first claim, we let Φ:\Phi\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}\to\mathbb{R} by Φ(x)=x\Phi(x)=x if x>0x>0 and Φ(x)=0\Phi(x)=0 otherwise. By the relabelling lemma ([Ilm92, Lemma 3.2]), v:=Φuv\mathrel{\mathop{\mathchar 58\relax}}=\Phi\circ u also satisfies the level set equation. Noting v(,0)=0v(\cdot,0)=0 precisely on Kin(0)K_{\mathrm{in}}(0), which is compact, we know by the uniqueness of level set flow that vv is a level set function of 𝒦in\mathcal{K}_{\mathrm{in}}. Hence,

n+1\Kin(t)={x:u(x,t)>0}.\mathbb{R}^{n+1}\backslash K_{\mathrm{in}}(t)=\{x\mathrel{\mathop{\mathchar 58\relax}}u(x,t)>0\}.

The second claim is similar. We let Ψ:\Psi\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}\to\mathbb{R} by Ψ(x)=x\Psi(x)=x if x<0x<0 and Ψ(x)=0\Psi(x)=0 otherwise. Then v=Ψuv=\Psi\circ u satisfies the level set equation by the relabelling lemma, and {x:u(x,t)0}={x:v(x,t)=0}\{x\mathrel{\mathop{\mathchar 58\relax}}u(x,t)\geq 0\}=\{x\mathrel{\mathop{\mathchar 58\relax}}v(x,t)=0\}, which is non-compact. Nevertheless, by Ilmanen [Ilm92], because any level sets other than KoutK_{\mathrm{out}} are compact, {x:v(x,t)=0}\{x\mathrel{\mathop{\mathchar 58\relax}}v(x,t)=0\} is the biggest flow, which is unique. Then the second claim will follow. ∎

Finally, we denote

Win(t)=n+1\Kout(t),Wout(t)=n+1\Kin(t),W(t)=Win(t)Wout(t).W_{\mathrm{in}}(t)=\mathbb{R}^{n+1}\backslash K_{\mathrm{out}}(t),\;\;W_{\mathrm{out}}(t)=\mathbb{R}^{n+1}\backslash K_{\mathrm{in}}(t),\;\;W(t)=W_{\mathrm{in}}(t)\cup W_{\mathrm{out}}(t).

In fact, we will further define the spacetime track

Win[t0,t1]=t[t0,t1]Win(t)×{t},W_{\mathrm{in}}[t_{0},t_{1}]=\bigcup_{t\in[t_{0},t_{1}]}W_{\mathrm{in}}(t)\times\{t\},

and we can similarly define Wout[t0,t1]W_{\mathrm{out}}[t_{0},t_{1}] and W[t0,t1]W[t_{0},t_{1}]. The reason we care about these sets is that their topological changes are described by White [Whi95], which will be crucial for us later. We remark that, when we need to specify the flow \mathcal{M}, we will add a superscript M to the symbols: e.g. we will write Win(t)W^{\mathcal{M}}_{\mathrm{in}}(t) in place of Win(t)W_{\mathrm{in}}(t).

Let (x,T)(x,T) be a singularity of \mathcal{M}, and λj\lambda_{j}\to\infty. Then any subsequential limit, in the sense of Brakke flow (see [Ilm94, Section 7]), of the rescaled flows

{λj(M(λj2t+T)x)}λj2T<t<0\{\lambda_{j}(M(\lambda^{-2}_{j}t+T)-x)\}_{-\lambda^{2}_{j}T<t<0}

is called a tangent flow at (x,T)(x,T). The tangent flow is unique if it is the shrinking cylinder or has only conical ends, by Colding-Minicozzi [CM15] and Chodosh-Choi-Schulze [CS21] respectively. Moreover, the convergence is in ClocC_{\text{loc}}^{\infty} by Brakke’s regularity theorem (see [Whi05]).

Now, following [CHHW22], we call (x,T)(x,T) an inward neck singularity of \mathcal{M} if, as λ\lambda\to\infty, the rescaled flows

{λ(Kin(λ2t+T)x)}λ2T<t<0\{\lambda(K_{\mathrm{in}}(\lambda^{-2}t+T)-x)\}_{-\lambda^{2}T<t<0}

converge locally smoothly with multiplicity one to the solid shrinking cylinder

{Bn(2(n1)t)×}t<0\{B^{n}(\sqrt{-2(n-1)t})\times\mathbb{R}\}_{t<0}

up to rotation and translation. Similarly, we can define an outward neck singularity. If, instead, those rescaled flows converge with multiplicity one to the solid shrinking ball

{Bn+1(2nt)}t<0\{B^{n+1}(\sqrt{-2nt})\}_{t<0}

up to translation, then we call (x,T)(x,T) an inward spherical singularity. We can again similarly define an outward spherical singularity.

2.3. MCF through cylindrical and spherical singularities

If every singularity of \mathcal{M} is a neck or a spherical singularity, then we call \mathcal{M} a MCF through cylindrical and spherical singularities. In this case, building on Hershkovits-White [HW20], Choi-Haslhofer-Hershkovits-White showed M(t),Min(t)M(t),M_{\mathrm{in}}(t), and Mout(t)M_{\mathrm{out}}(t) are all the same [CHHW22, Theorem 1.19], i.e. fattening does not occur.

Neck singularities are well-understood after the work of many researchers [HS99a, HS99b, Whi00, Whi03, SW09, Wan11, And12, Bre15, CM15, HK17, ADS19, ADS20, CHH22, CHHW22], among others. In Theorem 2.4, we will state the canonical neighborhood theorem of Choi-Haslhofer-Hershkovits-White [CHHW22]. Using that, we obtain a more detailed topological description of neck singularities in Theorem 2.5.

Definition 2.2.

Let X=(x,T)X=(x,T) be a regular point in a level-set flow \mathcal{M}. Let λ:=|𝐇(x)|\lambda\mathrel{\mathop{\mathchar 58\relax}}=|{\bf H}(x)|. Suppose there exists an ancient MCF {Σ(t)}\{\Sigma(t)\} that is, up to spacetime translation and parabolic rescaling, one of the following:

  • the shrinking sphere,

  • the shrinking cylinder with axis \ell,

  • the translating bowl with axis \ell,

  • the ancient oval with axis \ell,

such that: For each t(1/ϵ2,0]t\in(-1/\epsilon^{2},0] and inside B1/ϵ(0)n+1B_{1/\epsilon}(0)\subset\mathbb{R}^{n+1},

λ(M(λ2t+T)x) and Σ(t)\lambda(M(\lambda^{-2}t+T)-x)\textrm{ and }\Sigma(t)

are ϵ\epsilon-close in C1/ϵC^{\lfloor{1/\epsilon}\rfloor}. Then, we call

(T1λ2ϵ2,T]×B1λϵ(x)\left(T-\frac{1}{\lambda^{2}\epsilon^{2}},T\right]\times B_{\frac{1}{\lambda\epsilon}}(x)

an ϵ\epsilon-canonical neighborhood of XX with axis \ell.

We will also have a weaker definition, for situations when we focus on a time slice:

Definition 2.3.

Let xx be a regular point in a subset MM. Let λ:=|𝐇(x)|\lambda\mathrel{\mathop{\mathchar 58\relax}}=|{\bf H}(x)|. Suppose there exists a hypersurface Σ\Sigma that is, up to translation and rescaling, a time slice of one of the following:

  • the shrinking sphere,

  • the shrinking cylinder with axis \ell,

  • the translating bowl with axis \ell,

  • the ancient oval with axis \ell,

and such that: Inside B1/ϵ(0)n+1B_{1/\epsilon}(0)\subset\mathbb{R}^{n+1}, λ(Mx) and Σ\lambda(M-x)\textrm{ and }\Sigma are ϵ\epsilon-close in C1/ϵC^{\lfloor{1/\epsilon}\rfloor}. Then, we call B1λϵ(x)B_{\frac{1}{\lambda\epsilon}}(x) an ϵ\epsilon-canonical neighborhood of xx with axis \ell.

One can compare the above with the notion of ϵ\epsilon-canonical neighborhoods in 3-dimensional Ricci flow [MF10, Lecture 2].

Theorem 2.4 (Canonical neighborhood).

Let (x,T)(x,T) be a neck singularity of a MCF through cylindrical and spherical singularities \mathcal{M}, and \ell be the axis of the cylindrical tangent flow at (x,T)(x,T). Then for every ϵ>0\epsilon>0, there exists δ,δ¯>0\delta,\bar{\delta}>0 such that every regular point of \mathcal{M} in B2δ(x)×(Tδ¯,T+δ¯)B_{2\delta}(x)\times(T-\bar{\delta},T+\bar{\delta}) has an ϵ\epsilon-canonical neighborhood with axis \ell in the sense of Definition 2.2.

We used balls of radius 2δ2\delta (instead of δ\delta): This is solely for the sake of notational convenience, so that it can be directly quoted in Theorem 2.5.

Proof.

This is from [CHHW22, Corollary 1.18]. Note that all limit flows at (x,T)(x,T) have the same axis [CHHW22, p.163]. ∎

2.4. Consequence of almost all time regularity

Recall that throughout this paper, a cylindrical singularity has tangent flow given by the cylinder Sn1×S^{n-1}\times\mathbb{R}. By White’s stratification of singular set of MCF ([Whi97, Whi03]), at almost every time, the time-slice of a MCF through cylindrical and spherical singularities is smooth. Based on this, in items (3) - (6) of the following theorem, we will obtain a topologically more refined picture of neck-pinches. The shapes of the surfaces described in items (3) - (6) are illustrated in Figure 6.

Refer to caption

Figure 6.
Theorem 2.5.

There exists a universal constant R0=R0(n)R_{0}=R_{0}(n) with the following significance. Let (x,T)(x,T) be an inward neck singularity of a MCF through cylindrical and spherical singularities \mathcal{M} in n+1\mathbb{R}^{n+1}, and \ell be the axis of the cylindrical tangent flow at (x,T)(x,T). For every δ0>0\delta_{0}>0 and every R>R0R>R_{0}, there exists δ(0,δ0)\delta\in(0,\delta_{0}) and δ¯>0\bar{\delta}>0 such that:

  1. (1)

    Let B=Bδ(x)B=B_{\delta}(x). Then the set M(Tδ¯)BM(T-\bar{\delta})\cap B

    • is up to scaling and translation 1R\frac{1}{R}-close in CC^{\infty} to the cylinder (Sn1×\cong S^{n-1}\times\mathbb{R}) in BR(0)B_{R}(0) with axis \ell and radius 11,

    • and as a topological cylinder has Kin(Tδ¯)BK_{\mathrm{in}}(T-\bar{\delta})\cap B on its inside.

    • δ¯0\bar{\delta}\to 0 as RR\to\infty.

  2. (2)

    (Mean convex neighborhood) For every Tδ¯<t1<t2<T+δ¯T-\bar{\delta}<t_{1}<t_{2}<T+\bar{\delta},

    Kin(t2)BKin(t1)\M(t1).K_{\mathrm{in}}(t_{2})\cap B\subset K_{\mathrm{in}}(t_{1})\backslash M(t_{1}).

Moreover, there exists some countable dense set J[Tδ¯,T+δ¯]J\subset[T-\bar{\delta},T+\bar{\delta}] with Tδ¯JT-\bar{\delta}\in J such that we have for every tJt\in J:

  1. (3)

    M(t)M(t) is smooth, and intersects B\partial B transversely.

  2. (4)

    Each connected component of Kin(t)BK_{\mathrm{in}}(t)\cap\partial B is a convex nn-ball in B\partial B.

  3. (5)

    Denote the two connected components of Kin(Tδ¯)BK_{\mathrm{in}}(T-\bar{\delta})\cap\partial B by D1D_{1} and D2D_{2}. Then M(t)DiM(t)\cap D_{i} has at most one connected component for i=1,2i=1,2.

  4. (6)

    Let KK be a connected component of Kin(t)BK_{\mathrm{in}}(t)\cap B. Then KK satisfies one of the following:

    • K\partial K is a connected component of M(t)BM(t)\cap B that is a sphere.

    • K\partial K consists of a connected component of M(t)BM(t)\cap B that is an nn-ball and another ball on B\partial B.

    • K\partial K consists of a connected component of M(t)BM(t)\cap B that is a cylinder Sn1×(0,1)\cong S^{n-1}\times(0,1) and two balls on B\partial B.

And the case for outward neck singularities is analogous.

Proof.

We will just do the case of inward neck singularity.

To obtain (1) and (2).

Let us first arbitrarily pick some ϵ,R>0\epsilon,R>0, which we will further specify later. Let δ,δ¯>0\delta,\bar{\delta}>0 be obtained from applying the canonical neighborhood theorem (Theorem 2.4) to (x,T)(x,T) and ϵ\epsilon. We can decrease δ¯\bar{\delta} such that it lies in the range (0,δ0)(0,\delta_{0}).

By possibly further decreasing δ,δ¯\delta,\bar{\delta}, we can guarantee (2) by the mean convex neighborhood theorem of Choi-Haslhofer-Hershkovits-White [CHHW22, Theorem 1.17]. In fact, further decreasing δ,δ¯\delta,\bar{\delta}, we can by the definition of neck singularity assume that M(Tδ¯)B2δ(x)M(T-\bar{\delta})\cap B_{2\delta}(x)

  • is, up to scaling and translation, 1R\frac{1}{R}-close in CC^{\infty} to the cylinder (Sn1×\cong S^{n-1}\times\mathbb{R}) in B2R(0)B_{2R}(0) with axis \ell and radius 11,

  • and as a topological cylinder has Kin(Tδ¯)B2δ(x)K_{\mathrm{in}}(T-\bar{\delta})\cap B_{2\delta}(x) on its inside.

In particular, (1) is fulfilled.

To define JJ and obtain (3).

Note that using [CM16, Corollary 0.6], for some set I1[Tδ¯,T+δ¯]I_{1}\subset[T-\bar{\delta},T+\bar{\delta}] of full measure, M(t)M(t) is smooth for all tI1t\in I_{1}. Then (3) just follows from a standard transversality argument. Namely, for each tI1t\in I_{1}, via the transversality theorem, Br(x)B_{r}(x) intersects M(t)M(t) transversely for a.e. r(δ/2,δ)r\in(\delta/2,\delta). Hence, for some countable dense subset JI1J\subset I_{1} and some set I2(δ/2,δ)I_{2}\subset(\delta/2,\delta) of full measure, for all (t,r)J×I2(t,r)\in J\times I_{2}, Br(x)B_{r}(x) intersects M(t)M(t) transversely. Hence, by slightly decreasing δ\delta, (3) can be fulfilled.

To obtain (4).

Let us first state a lemma, which gives us the constant R0R_{0} we need.

Lemma 2.6.

There exist constants R0>2R_{0}>2, and ϵ0,ϵ1>0\epsilon_{0},\epsilon_{1}>0, all depending only on nn, with the following significance.

  • Consider some ball B2R0(x)B_{2R_{0}}(x), and fix a diameter line \ell. Let 𝒞B2R0(x)\mathcal{C}\subset B_{2R_{0}}(x) be the solid cylinder with radius 22 and axis \ell.

  • Let xx^{\prime} be a regular point of some time-slice M(t)M(t) of a level set flow in n+1\mathbb{R}^{n+1}, and xx^{\prime} has an ϵ0\epsilon_{0}-canonical neighborhood with axis \ell.

  • Assume xBR0(x)x^{\prime}\in B_{R_{0}}(x), M(t)B2R0(x)𝒞M(t)\cap B_{2R_{0}}(x)\subset\mathcal{C}.

  • Let SS be a smooth nn-disc properly embedded in 𝒞\mathcal{C}, with S\partial S lying on and transversely intersecting the cylindrical part of 𝒞\partial\mathcal{C}, and xSx^{\prime}\in S, such that:

  • SS is ϵ1\epsilon_{1}-close in CC^{\infty} to some planar nn-disc perpendicular to \ell. (See Figure 7.)

Then we have:

  • If M(t)M(t) intersects SS transversely at xx^{\prime}, then the connected component DD of Kin(t)SK_{\mathrm{in}}(t)\cap S that contains xx^{\prime} is a convex nn-disc in SS, and M(t)D=DM(t)\cap D=\partial D with the intersection being transverse.

  • If M(t)M(t) does not intersect SS transversely at xx^{\prime}, then DD is just the point xx^{\prime}.

Refer to caption

Figure 7.
Proof.

By an inspection of the geometry of the sphere, cylinder, bowl, and ancient oval, for all sufficiently large R0R_{0} and small ϵ0\epsilon_{0}, if M(t)B2R0(x)𝒞M(t)\cap B_{2R_{0}}(x)\subset\mathcal{C} then

M(t)B2R0(x)(ϵ0-canonical neighborhood of x)M(t)\cap B_{2R_{0}}(x)\cap(\epsilon_{0}\textrm{-canonical neighborhood of }x^{\prime})

has curvature |A|>1/2|A|>1/2. Thus, if the smooth nn-disc SS is sufficiently planar, the desired claim follows easily. ∎

Now, we begin proving (4). Let us assume the R,ϵR,\epsilon we chose satisfy R>R0R>R_{0} and ϵ<ϵ0\epsilon<\epsilon_{0}, with R0,ϵ0R_{0},\epsilon_{0} from the above lemma. By how we chose RR in the proof of (1) above, we can rescale M(Tδ¯)M(T-\bar{\delta}) by some factor λ\lambda such that

λ(M(Tδ¯)x)B2R(0)\lambda(M(T-\bar{\delta})-x)\cap B_{2R}(0)

lies in the solid cylinder CB2R(0)C\subset B_{2R}(0) with axis \ell and radius 22. Thus, by the mean convex neighborhood property (2), for all t(Tδ¯,T+δ¯)t\in(T-\bar{\delta},T+\bar{\delta}),

λ(M(t)x)B2R(0)C.\lambda(M(t)-x)\cap B_{2R}(0)\subset C.

Now, remember that we should focus on those tJ(Tδ¯,T+δ¯)t\in J\subset(T-\bar{\delta},T+\bar{\delta}). By Theorem 2.4 and ϵ<ϵ0\epsilon<\epsilon_{0}, M(t)M(t) has an ϵ0\epsilon_{0}-canonical neighborhood with \ell, and so does λ(M(t)x)\lambda(M(t)-x) since the property is independent of scaling and translation. Let SS be a connected component of BR(0)C\partial B_{R}(0)\cap C. By increasing RR, we can make SS arbitrarily close to being planar. Hence, we can apply Lemma 2.6. Then (4) follows immediately.

To obtain (5).

We will just do the case for D1D_{1}. Let

T1:=sup{tJ:M(t)D1 has only one connected component}.T_{1}\mathrel{\mathop{\mathchar 58\relax}}=\sup\{t\in J\mathrel{\mathop{\mathchar 58\relax}}M(t)\cap D_{1}\textrm{ has only one connected component}\}.

Note that T1>Tδ¯T_{1}>T-\bar{\delta} by (1) and Tδ¯JT-\bar{\delta}\in J. To prove that M(t)D1M(t)\cap D_{1} has at most one connected component for each tJt\in J, it suffices to prove that T1=T+δ¯T_{1}=T+\bar{\delta}. Suppose the otherwise, i.e. T1<T+δ¯T_{1}<T+\bar{\delta} so that there exists a sequence in JJ, t1,t2,T1t_{1},t_{2},...\downarrow T_{1}, such that M(ti)D1M(t_{i})\cap D_{1} contains at least two components.

Now, let

K1=Tδ¯<t<T1Kin(t)D1,K2=Kin(T1)D1,K3=iKin(ti)D1.K_{1}=\bigcap_{T-\bar{\delta}<t<T_{1}}K_{\mathrm{in}}(t)\cap D_{1},\;\;K_{2}=K_{\mathrm{in}}(T_{1})\cap D_{1},\;\;K_{3}=\bigcup_{i}K_{\mathrm{in}}(t_{i})\cap D_{1}.

Note that K1K2K3K_{1}\supset K_{2}\supset K_{3} by the mean convex neighborhood property (2).

Proposition 2.7.

K1K_{1} is a convex nn-ball in B\partial B, K1=K2K_{1}=K_{2}, and K3K_{3} is dense in K1K_{1}.

Proof.

By the mean convex property,

K1=tJ,t<T1Kin(t)D1.K_{1}=\bigcap_{t\in J,t<T_{1}}K_{\mathrm{in}}(t)\cap D_{1}.

Then by (4), K1K_{1} is a convex nn-ball.

To prove K1=K2K_{1}=K_{2}, it suffices to prove K1K2K_{1}\subset K_{2}. Note that by Lemma 2.1, for every xK1x\in K_{1} and t(Tδ¯,T1)t\in(T-\bar{\delta},T_{1}) we have u(x,t)0u(x,t)\leq 0, where uu is a level set function for \mathcal{M}. Since uu is continuous, u(x,T1)0u(x,T_{1})\leq 0, implying xK2x\in K_{2} by Lemma 2.1.

Finally, to prove K3K_{3} is dense in K1K_{1}, it suffices to prove K1\K3K_{1}\backslash K_{3} has empty interior (as a subset of B\partial B) since K1K_{1} is a convex nn-ball. We claim that K2\K3Min(T1)K_{2}\backslash K_{3}\subset M_{\mathrm{in}}(T_{1}). Indeed, if xK2\K3x\in K_{2}\backslash K_{3}, then for every spacetime neighborhood UU of (x,T1)(x,T_{1}) in n+1×\mathbb{R}^{n+1}\times\mathbb{R}, for each ii, UU contains the point

(x,ti)(n+1×)\𝒦in.(x,t_{i})\in(\mathbb{R}^{n+1}\times\mathbb{R})\backslash\mathcal{K}_{\mathrm{in}}.

Thus, (x,T1)𝒦in(x,T_{1})\in\partial\mathcal{K}_{\mathrm{in}}, and so xMin(T1)x\in M_{\mathrm{in}}(T_{1}).

As a result,

K1\K3=K2\K3Min(T1)D1=M(T1)D1,K_{1}\backslash K_{3}=K_{2}\backslash K_{3}\subset M_{\mathrm{in}}(T_{1})\cap D_{1}=M(T_{1})\cap D_{1},

where the last equality is by the non-fattening of \mathcal{M} [CHHW22, Theorem 1.19]. We will prove that M(T1)D1M(T_{1})\cap D_{1} consists entirely of singularities (of \mathcal{M}), and then immediately we would know M(T1)D1M(T_{1})\cap D_{1} has empty interior using [CM16, Theorem 0.1], which says that the singular set of \mathcal{M} is contained in finitely many compact embedded Lipschitz submanifolds each of dimension at most n1n-1 together with a set of dimension n2n-2.

Suppose by contradiction that M(T1)D1M(T_{1})\cap D_{1} contains some regular point pp. So around some neighborhood of pp in n+1\mathbb{R}^{n+1}, M(T1)M(T_{1}) is a smooth surface, with Kin(T1)K_{\mathrm{in}}(T_{1}) on one side. Thus, we have pK2p\in\partial K_{2}, with K2K_{2} a convex nn-ball. Then we repeat the argument in the above proof of (4) to apply Lemma 2.6 around the point pp, and conclude:

  • K2\partial K_{2} is a smooth (n1)(n-1)-sphere and consists entirely of regular points.

  • The interior of K2K_{2} does not intersects M(T1)M(T_{1}).

  • M(T1)M(T_{1}) intersects D1D_{1} transversely along K2\partial K_{2}.

So, for some short amount of time after T1T_{1}, M(T1)D1M(T_{1})\cap D_{1} would still have only one connected component by pseudolocality of (locally) smooth MCF (see [INS19, Theorem 1.5]). This contradicts the definition of T1T_{1}. ∎

Let us continue the proof of (5). Now, for each ii, Kin(ti)D1K_{\mathrm{in}}(t_{i})\cap D_{1} has finitely many connected components by transversality (3). Let EiE_{i} be the one with the maximal diameter (measured inside B\partial B), denoted did_{i}. Then by the canonical neighborhood property Theorem 2.4, assuming ϵ\epsilon small, for some geodesic ball E~iB\widetilde{E}_{i}\subset\partial B of diameter 3di3d_{i}, E~iKin(ti)=Ei\widetilde{E}_{i}\cap K_{\mathrm{in}}(t_{i})=E_{i}.

Now, note did_{i} is increasing in ii by the mean convex neighborhood property (2). Let d=limidid=\lim_{i}d_{i}. There are two cases: (a) ddiam(K1)/2d\geq\mathrm{diam}(K_{1})/2, and (b) d<diam(K1)/2d<\mathrm{diam}(K_{1})/2. For case (a), by the definition of tit_{i}, we know for sufficiently large ii, the neighborhood E~i\widetilde{E}_{i} would then need to contain a connected component of Kin(ti)D1K_{\mathrm{in}}(t_{i})\cap D_{1} other than EiE_{i}, contradicting the definition of E~i\widetilde{E}_{i}. So case (a) is impossible. Case (b) is also impossible since it, together with the existence of E~i\widetilde{E}_{i}, violates Proposition 2.7 which says K3K_{3} is dense in K1K_{1}. This finishes the proof of (5).

To obtain (6).

Choose a connected component KK of Kin(t)Bδ(x)K_{\mathrm{in}}(t)\cap B_{\delta}(x). Let us foliate B2δ(x)B_{2\delta}(x) with planar nn-discs that are perpendicular to the axis \ell. Then as in the proof of (4), we apply Lemma 2.6 to characterize the intersection of KK with every such planar nn-discs. Namely, every such set of intersections consists of convex nn-discs and isolated points. Viewing these sets of intersection as level sets of some function defined on KK, Morse theory then immediately implies (6).

This finishes the proof of Theorem 2.5. ∎

Finally, we discuss some convergence theorems of MCF through cylindrical and spherical singularities.

Proposition 2.8.

Let i={Mi(t)}t0\mathcal{M}^{i}=\{M^{i}(t)\}_{t\geq 0}, with i=1,2,i=1,2,..., and ={M(t)}t0\mathcal{M}=\{M(t)\}_{t\geq 0} be MCF through neck and spherical singularities in n+1\mathbb{R}^{n+1}. Assume that each Mi(0)M^{i}(0) and M(0)M(0) are smooth, closed hypersurfaces, with Mi(0)M(0)M^{i}(0)\to M(0) in CC^{\infty}. Then

  1. (1)

    For a.e. tt, Mi(t)M(t)M^{i}(t)\to M(t) in CC^{\infty}.

  2. (2)

    The spacetime tracks i\mathcal{M}^{i}\to\mathcal{M} in the Hausdorff sense.

Proof.

By Ilmanen’s elliptic regularization (see [Ilm94, Whi09]), for any closed smooth hypersurface Mi(0)M^{i}(0), there exists a unit regular cyclic Brakke flow {μti}t0\{\mu_{t}^{i}\}_{t\geq 0} such that μ0i=Mi(0)n\mu_{0}^{i}=M^{i}(0)\lfloor\mathcal{H}^{n}, where n\mathcal{H}^{n} is the nn-dimensional Hausdorff measure. By the mean convex neighborhood theorem [CHH22] and the nonfattening of level set flow with singularities that have mean convex neighborhood [HW20], {μti}t0\{\mu^{i}_{t}\}_{t\geq 0} is supported on i\mathcal{M}^{i}. Then the compactness of Brakke flows ([Ilm94, Whi09]) implies that {μti}t0\{\mu^{i}_{t}\}_{t\geq 0} subsequentially converges to a limit unit regular cyclic Brakke flow {μt}t0\{\mu^{\infty}_{t}\}_{t\geq 0}.

Because Mi(0)M(0)M^{i}(0)\to M(0) smoothly, μ0=μ0\mu^{\infty}_{0}=\mu_{0}, and by the uniqueness of unit regular cyclic Brakke flow, μt=μt\mu^{\infty}_{t}=\mu_{t} a.e. for all t0t\geq 0. In particular, the regular part of μt\mu^{\infty}_{t} equals the regular part of μt\mu_{t}. Then by Brakke’s regularity theorem and a.e. time regularity of i\mathcal{M}^{i} with neck and spherical singularities, we have for a.e. tt, Mi(t)M(t)M^{i}(t)\to M(t).

The compactness of weak set flow shows that i\mathcal{M}^{i} subsequentially converges to a limit weak set flow \mathcal{M}^{\infty} in Hausdorff distance. Because {μt}t0\{\mu_{t}\}_{t\geq 0} is supported on \mathcal{M}^{\infty}, we have \mathcal{M}\subset\mathcal{M}^{\infty}. Meanwhile, \mathcal{M} is the biggest flow, therefore \mathcal{M}^{\infty}\subset\mathcal{M}. Thus, =\mathcal{M}^{\infty}=\mathcal{M}. This also shows the uniqueness of the limit. Therefore, i\mathcal{M}^{i} converges to \mathcal{M} in Hausdorff distance.

3. Homology descent, homology termination, and homology breakage

In this section, we consider general level set flows ={M(t)}t0\mathcal{M}=\{M(t)\}_{t\geq 0} in n+1\mathbb{R}^{n+1}, where M(0)M(0) is not necessarily a closed hypersurface. We will introduce three new concepts. For a heuristic explanation of them, see §1.1.

Let Hk()H_{k}(\cdot) denotes the kk-th homology group in \mathbb{Z}-coefficients.

Definition 3.1 (Homology descent).

We define a relation \succ on the disjoint union

t0Hn1(W(t))\bigsqcup_{t\geq 0}H_{n-1}(W(t))

as follows. Given two times T0T1T_{0}\leq T_{1}, and two homology classes c0Hn1(W(T0))c_{0}\in H_{n-1}(W(T_{0})) and c1Hn1(W(T1))c_{1}\in H_{n-1}(W(T_{1})), we say that c1c_{1} descends from c0c_{0}, and denote

c0c1,c_{0}\succ c_{1},

if every representative γ0c0\gamma_{0}\in c_{0} and γ1c1\gamma_{1}\in c_{1} together bound some nn-chain ΓW[T0,T1]\Gamma\subset W[T_{0},T_{1}], i.e. γ0γ1=Γ.\gamma_{0}-\gamma_{1}=\partial\Gamma. (See Figure 3.)

Clearly, in the above definition, we can interchangeably replace “every representative” with “some representative”. Note that we are using singular homology, which means that γ0,γ1,\gamma_{0},\gamma_{1}, and Γ\Gamma are just singular chains.

Remark 3.2.

The relation \succ is a partial order. Indeed, let ciHn1(W(Ti))c_{i}\in H_{n-1}(W(T_{i})) for i=0,1,2i=0,1,2. Clearly c0c0c_{0}\succ c_{0}. If c0c1c_{0}\succ c_{1} and c1c0c_{1}\succ c_{0}, then T0=T1T_{0}=T_{1}, implying c0=c1c_{0}=c_{1}. Moreover, if c0c1c_{0}\succ c_{1} and c1c2c_{1}\succ c_{2}, then T0T2T_{0}\leq T_{2} and it readily follows from definition that c0c2c_{0}\succ c_{2}.

This relation has certain favorable properties.

Proposition 3.3.

Let c0Hn1(W(T0))c_{0}\in H_{n-1}(W(T_{0})) and T0T1T_{0}\leq T_{1}. Then there exists at most one c1Hn1(W(T1))c_{1}\in H_{n-1}(W(T_{1})) such that c0c1c_{0}\succ c_{1}.

Proof.

Suppose c1,c2Hn1(W(T1))c_{1},c_{2}\in H_{n-1}(W(T_{1})) satisfy c0c1c_{0}\succ c_{1} and c0c2c_{0}\succ c_{2}. Our aim is to show c1=c2c_{1}=c_{2}. Choose γici\gamma_{i}\in c_{i} for i=0,1,2i=0,1,2. Then by definition, γ0γ1=A\gamma_{0}-\gamma_{1}=\partial A for some AW[T0,T1]A\subset W[T_{0},T_{1}], and similarly γ0γ2=B\gamma_{0}-\gamma_{2}=\partial B for some BW[T0,T1]B\subset W[T_{0},T_{1}]. Thus, γ1\gamma_{1} and γ2\gamma_{2} bound ABW[T0,T1]A-B\subset W[T_{0},T_{1}]. Since the map

Hn1(W(T1))Hn1(W[T0,T1])H_{n-1}(W(T_{1}))\to H_{n-1}(W[T_{0},T_{1}])

induced by the inclusion W(T1)W[T0,T1]W(T_{1})\to W[T_{0},T_{1}] is injective by White [Whi95, Theorem 1 (iii)], we deduce that γ1\gamma_{1} and γ2\gamma_{2} are homologous within W(T1)W(T_{1}). Consequently, c1=c2c_{1}=c_{2}. ∎

Remark 3.4.

Note that in the above it is possible that there does not exist any c1Hn1(W(T1))c_{1}\in H_{n-1}(W(T_{1})) for which c0c1c_{0}\succ c_{1}. As illustrated in Figure 8, after time TT, no homology class c1c_{1} satisfies a0c1a_{0}\succ c_{1}.

Refer to caption
Figure 8.
Remark 3.5.

On the other hand, there may be multiple homology classes c0H1(W(T0))c_{0}\in H_{1}(W(T_{0})) satisfying the relation c0c1c_{0}\succ c_{1}. As an example, consider the flow shown in Figure 8, where both b0H1(Wout(0))b_{0}\in H_{1}(W_{\mathrm{out}}(0)) and the trivial element of H1(Wout(0))H_{1}(W_{\mathrm{out}}(0)) descend to the trivial element of H1(Wout(T1))H_{1}(W_{\mathrm{out}}(T_{1})).

In fact, precisely because of Proposition 3.3 and Remark 3.5, we chose the symbol \succ (instead of \prec) to pictographically reflect that more than one homology class may descend into one, but not the other way around.

Proposition 3.6.

We focus on the case n=2n=2. Let c1H1(W(T1))c_{1}\in H_{1}(W(T_{1})) and T0T1T_{0}\leq T_{1}. Then there exists at least one c0H1(W(T0))c_{0}\in H_{1}(W(T_{0})) such that c0c1c_{0}\succ c_{1}.

Proof.

Choose some γc1\gamma\in c_{1}. By White [Whi95, Theorem 1 (ii)], γ\gamma can be homotoped through W[T0,T1]W[T_{0},T_{1}] to some loop γ\gamma^{\prime} in W(T0)W(T_{0}). So c0:=[γ]c1c_{0}\mathrel{\mathop{\mathchar 58\relax}}=[\gamma^{\prime}]\succ c_{1}. ∎

The following proposition says that a homology class cannot disappear and then reappear later.

Proposition 3.7.

Let T0<T1T_{0}<T_{1}, c0Hn1(W(T0))c_{0}\in H_{n-1}(W(T_{0})), and c1Hn1(W(T1))c_{1}\in H_{n-1}(W(T_{1})) with c0c1c_{0}\succ c_{1}. Then for every t[T0,T1]t\in[T_{0},T_{1}] there exists a unique cH1(W(t))c\in H_{1}(W(t)) such that c0cc1c_{0}\succ c\succ c_{1}.

Proof.

We only need to prove existence, as then uniqueness would follow from Proposition 3.3.

Under our assumption, we have γ0c0\gamma_{0}\in c_{0} and γ1c1\gamma_{1}\in c_{1} such that they together bound some nn-chain CC in W[T0,T1]W[T_{0},T_{1}]. Since W[T0,T1]W[T_{0},T_{1}] is an open subset of Euclidean space, we can choose a representative of the nn-chain CC as a polyhedron chain. By tilting the faces appropriately, we can ensure that they do not lie entirely within any specific slice n+1×{t}\mathbb{R}^{n+1}\times\{t\}. This enables us to find βt={x:(x,t)C}\beta_{t}=\{x\mathrel{\mathop{\mathchar 58\relax}}(x,t)\in C\} as an (n1)(n-1)-chain without a boundary for each t[T0,T1]t\in[T_{0},T_{1}]. Consequently, we have [βt]Hn1(W(t))[\beta_{t}]\in H_{n-1}(W(t)), and c0[βt]c1c_{0}\succ[\beta_{t}]\succ c_{1}. ∎

Based on Proposition 3.7, the following definition is well-defined.

Definition 3.8 (Homology termination).

Let c0Hn1(W(T0))c_{0}\in H_{n-1}(W(T_{0})).

  • If

    𝔱(c0):=sup{tT0:c0c for some cHn1(W(t))}\mathfrak{t}(c_{0})\mathrel{\mathop{\mathchar 58\relax}}=\sup\{t\geq T_{0}\mathrel{\mathop{\mathchar 58\relax}}c_{0}\succ c\textrm{ for some }c\in H_{n-1}(W(t))\}

    is finite, then we say that c0c_{0} terminates at time 𝔱(c0)\mathfrak{t}(c_{0}), otherwise we say c0c_{0} never terminates.

  • And for each tT0t\geq T_{0}, the unique cHn1(W(t))c\in H_{n-1}(W(t)) such that c0cc_{0}\succ c, if exists, is denoted c0(t)c_{0}(t).

If needed, we use 𝔱\mathfrak{t}^{\mathcal{M}} in place of 𝔱\mathfrak{t} to specify the flow.

Note that since WW is open, if c0c_{0} terminates at time 𝔱(c0)\mathfrak{t}(c_{0}) then there is no cHn1(W(𝔱(c0)))c\in H_{n-1}(W(\mathfrak{t}(c_{0}))) such that c0cc_{0}\succ c. So c0(𝔱(c0))c_{0}(\mathfrak{t}(c_{0})) is not well-defined, and every c0Hn1(W(T0))c_{0}\in H_{n-1}(W(T_{0})) does not terminate at time T0T_{0}. Therefore, one can interpret the time interval [T0,𝔱(c0))[T_{0},\mathfrak{t}(c_{0})) as the “maximal interval of existence” for c0c_{0}.

Remark 3.9 (Trivial homology classes).

Let us also elaborate on trivial homology classes. At each time tt, Hn1(W(t))H_{n-1}(W(t)) has a unique trivial homology class 0t0_{t}. This is true even for situations like Figure 8 when the surfaces have inside and outside regions: The trivial elements of H1(Win(t))H_{1}(W_{\mathrm{in}}(t)) and H1(Wout(t))H_{1}(W_{\mathrm{out}}(t)) are viewed as the same.

However, 0t0_{t} is considered distinct for different tt, because we used disjoint union in Definition 3.1. Nonetheless, for any t1<t2t_{1}<t_{2}, it is vacuously true that 0t10t20_{t_{1}}\succ 0_{t_{2}}. Thus, we can denote each 0t0_{t} as 0(t)0(t), following the notation in Definition 3.8. In addition, clearly, the trivial homology class never terminates.

Example 3.10.

Let us revisit Figure 8. It is clear that a0a_{0} terminates at time TT, whereas b0b_{0} does not. In fact, b0b_{0} will never terminate: b0(t)b_{0}(t) would just become trivial for each t>Tt>T.

Example 3.11.

Let us now instead consider the flow in Figure 9. At time TT, b0b_{0} terminates while a0a_{0} does not. In fact, a0(t)a_{0}(t) becomes trivial after time TT, and thus it will never terminate.

Refer to caption
Figure 9.

Now, we introduce another concept. In Figure 8, a0a_{0} terminates at time TT because, intuitively, it “breaks” at the cylindrical singularity xx. Similarly, in Figure 9, b0b_{0} terminates at time TT because it “breaks” at the outward cylindrical singularity. The following definition provides a precise characterization of this breakage phenomenon.

Definition 3.12 (Homology breakage).

Let c0Hn1(W(T0))c_{0}\in H_{n-1}(W(T_{0})), T1>T0T_{1}>T_{0}, and KM(T1)K\subset M(T_{1}) be a compact set. Suppose the following holds:

  • For each T0t<T1T_{0}\leq t<T_{1}, there exists c0(t)Hn1(W(t))c_{0}(t)\in H_{n-1}(W(t)) such that c0c0(t)c_{0}\succ c_{0}(t).

  • For every neighborhood Un+1U\subset\mathbb{R}^{n+1} of KK, for each t<T1t<T_{1} sufficiently close to T1T_{1}, every element of c0(t)c_{0}(t) intersects UU. (Recall Figure 4.)

Then we say that c0c_{0} breaks in (K,T1)(K,T_{1}). We will often concern the case when KK is just a point xM(T)x\in M(T), for which we say that c0c_{0} breaks at (x,T1)(x,T_{1}).

One might wonder why Definition 3.12 does not require c0c_{0} to terminate at time T1T_{1}. This is because it is not necessary:

Proposition 3.13.

If a homology class c0Hn1(W(T0))c_{0}\in H_{n-1}(W(T_{0})) breaks in some (K,T1)(K,T_{1}), then c0c_{0} terminates at time T1T_{1}.

Proof.

Suppose the otherwise: There exists T2>T1T_{2}>T_{1} and c2Hn1(W(T2))c_{2}\in H_{n-1}(W(T_{2})) such that c0c2c_{0}\succ c_{2}. Then there exists γ0c0\gamma_{0}\in c_{0} and γ2c2\gamma_{2}\in c_{2} that together in W[T0,T2]W[T_{0},T_{2}] bound some nn-chain CC. Without loss of generality we can assume that βt:={x:(x,t)C}\beta_{t}\mathrel{\mathop{\mathchar 58\relax}}=\{x\mathrel{\mathop{\mathchar 58\relax}}(x,t)\in C\} is an (n1)(n-1)-chain without boundary for each t[T0,T2]t\in[T_{0},T_{2}], as in the proof of Proposition 3.7. Then c0(t)=[βt]Hn1(W[t])c_{0}(t)=[\beta_{t}]\in H_{n-1}(W[t]) satisfies c0c0(t)c_{0}\succ c_{0}(t).

By assumption c0c_{0} breaks in some (K,T1)(K,T_{1}) with KM(T1)K\subset M(T_{1}). Therefore, KC=K\cap C=\emptyset. Since KK is compact and CC is closed, there exists a neighborhood of KK in n+1×\mathbb{R}^{n+1}\times\mathbb{R} of the form Br(K)×[T1δ,T1+δ]B_{r}(K)\times[T_{1}-\delta,T_{1}+\delta] that does not intersect CC. Consequently, for all t[T1δ,T1+δ]t\in[T_{1}-\delta,T_{1}+\delta], βt\beta_{t} avoids Br(K)B_{r}(K). This contradicts the assumption that c0c_{0} breaks at (K,T1)(K,T_{1}). ∎

Note that, vacuously, the trivial homology class does not break in any (K,T)(K,T). Moreover, if a homology class breaks in (K1,T)(K_{1},T) and K1K2M(T)K_{1}\subset K_{2}\subset M(T), then it also breaks in (K2,T)(K_{2},T).

One might wonder whether the converse of the above proposition is true. Actually, in the case of 2-dimensional MCF through cylindrical and spherical singularities, if a homology class terminates at some time TT, then it actually breaks at some cylindrical singularity (x,T)(x,T). This is the statement of Theorem 4.5, which is one of the main results in §4. However, we are unsure whether the converse is true in general.

Proposition 3.14.

No homology class breaks at a regular point.

Proof.

Suppose (x,T)(x,T) is a regular point. Then there exists a small ball BB around xx such that for all tt close to TT, MtBM_{t}\cap B is a smooth nn-disk. It is clear that every nn-chain can be homotoped to avoid BB. Therefore, no homology class breaks at (x,T)(x,T). ∎

Proposition 3.15.

No homology class breaks at a spherical singularity.

Proof.

Suppose otherwise. Without loss of generality, suppose some c0Hn1(W(T0))c_{0}\in H_{n-1}(W(T_{0})) breaks at some spherical singularity (x,T)(x,T). Then there exists a small ball BB around xx such that for all t<Tt<T close to TT, M(t)BM(t)\cap B is a smooth sphere. For each such tt, let γ\gamma be a representative of c0(t)c_{0}(t). By removing the components of γ\gamma inside the sphere M(t)BM(t)\cap B, we can assume that γ\gamma lies outside the sphere. Thus clearly γ\gamma can be homotoped within W(t)W(t) to avoid BB. This again contradicts the assumption that c0c_{0} breaks at (x,T)(x,T). ∎

Lastly, we conclude this section with the following proposition, which provides us with a scenario where we know the inside homology classes must terminate. Namely, if we take a compact shrinker and push it inward, then all non-trivial inside homology classes will terminate, while the outward ones will not. This proposition will be crucial for us when we use Theorem 1.1 to prove other main theorems.

Proposition 3.16.

The setting is as follows.

  • Let Σ\Sigma be a smooth, embedded, compact shrinker in 3\mathbb{R}^{3}.

  • Let S0(1)S^{0}(-1) be a surface, lying strictly inside Σ\Sigma, given by deforming Σ\Sigma within the inside region of Σ\Sigma.

  • Let S1(1)S^{1}(-1) be a surface, lying strictly outside Σ\Sigma, given by deforming Σ\Sigma within the outside region of Σ\Sigma.

  • Note that the first homology groups of

    3\Σ,3\S0(1), and 3\S1(1)\mathbb{R}^{3}\backslash\Sigma,\;\;\mathbb{R}^{3}\backslash S^{0}(-1),\textrm{ and }\mathbb{R}^{3}\backslash S^{1}(-1)

    can be canonically identified.

  • Let

    𝒮={tΣ}1t0,𝒮0={S0(t)}t1, and 𝒮1={S1(t)}t1\mathcal{S}=\{\sqrt{-t}\Sigma\}_{-1\leq t\leq 0},\;\;\mathcal{S}^{0}=\{S^{0}(t)\}_{t\geq-1},\textrm{ and }\mathcal{S}^{1}=\{S^{1}(t)\}_{t\geq-1}

    be the associated level set flows.

Then there exist times T,T~(1,0)T,\widetilde{T}\in(-1,0) such that

  1. (1)

    For each non-trivial element a0H1(Win𝒮0(1))a_{0}\in H_{1}(W_{\mathrm{in}}^{\mathcal{S}^{0}}(-1)), 𝔱(a0)T~\mathfrak{t}(a_{0})\leq\widetilde{T}.

  2. (2)

    For each element b0H1(Wout𝒮0(1))b_{0}\in H_{1}(W_{\mathrm{out}}^{\mathcal{S}^{0}}(-1)), b0(T~)b_{0}(\widetilde{T}) exists and is trivial.

  3. (3)

    For each element a1H1(Win𝒮1(1))a_{1}\in H_{1}(W_{\mathrm{in}}^{\mathcal{S}^{1}}(-1)), a1(T)a_{1}(T) exists and is trivial.

  4. (4)

    For each non-trivial element b1H1(Wout𝒮1(1))b_{1}\in H_{1}(W_{\mathrm{out}}^{\mathcal{S}^{1}}(-1)), 𝔱(b1)T\mathfrak{t}(b_{1})\leq T.

Proof.

For the first claim, note that:

  • S0(1)S^{0}(-1) is inside Σ\Sigma.

  • dist(tΣ,S0(t))\mathrm{dist}(\sqrt{-t}\Sigma,S^{0}(t)) is non-decreasing in tt by [ES91, Theorem 7.3].

  • Σ\Sigma shrinks self-similarly under the flow.

Thus, we can deduce the existence of T~<0\widetilde{T}<0 such that for every tT~t\geq\widetilde{T}, S0(t)S^{0}(t) is empty. Consequently, for any non-trivial element a0H1(Win𝒮0(1))a_{0}\in H_{1}(W^{\mathcal{S}^{0}}_{\mathrm{in}}(-1)), either 𝔱(a0)T~\mathfrak{t}(a_{0})\leq\widetilde{T}, or a0(T~)a_{0}(\widetilde{T}) still exists but is trivial. Suppose by contradiction that the latter holds. Then we can pick some α0a0\alpha_{0}\in a_{0} such that α0=A\alpha_{0}=\partial A for some

AWin𝒮0([1,T~])Win𝒮([1,T~]).A\subset W^{\mathcal{S}^{0}}_{\mathrm{in}}([-1,\widetilde{T}])\subset W^{\mathcal{S}}_{\mathrm{in}}([-1,\widetilde{T}]).

By rescaling each time slice of AA, we can ensure that α0\alpha_{0} bounds some

A~(interior region of Σ)×[1,T~].\widetilde{A}\subset(\textrm{interior region of }\Sigma)\times[-1,\widetilde{T}].

Projecting A~\widetilde{A} into the interior region of Σ\Sigma, we have that α0\alpha_{0} is homologically trivial, which contradicts the definition of α0\alpha_{0}. This concludes the proof of the first claim.

For the second claim, since Σ\Sigma shrinks self-similarly under the flow, we know that b0b_{0} has not terminated by the time T~(<0)\widetilde{T}(<0) for the flow tΣ{\sqrt{-t}\Sigma}. Then by the fact that S0(t)S^{0}(t) lies inside tΣ\sqrt{-t}\Sigma for each t[1,T~]t\in[-1,\widetilde{T}], which is a result of the avoidance principle, we can deduce that b0(T~)b_{0}(\widetilde{T}) still exists for the flow 𝒮0\mathcal{S}^{0}. However, as S0(T~)S^{0}(\widetilde{T}) is empty, it follows that b0(T~)b_{0}(\widetilde{T}) must be trivial.

Let us define

ϵ=dist(Σ,S1(1))}.\epsilon=\mathrm{dist}(\Sigma,S^{1}(-1))\}.

Pick a loop β1b1\beta_{1}\in b_{1}. Define Bϵ(tΣ)B_{\epsilon}(\sqrt{-t}\Sigma) as the ϵ\epsilon-neighborhood of tΣ\sqrt{-t}\Sigma, and denote

Y(t)\displaystyle Y(t) :=3\Bϵ(tΣ)\displaystyle\mathrel{\mathop{\mathchar 58\relax}}=\mathbb{R}^{3}\backslash B_{\epsilon}(\sqrt{-t}\Sigma)
Y[t1,t2]\displaystyle Y[t_{1},t_{2}] :=t[t1,t2](3\Bϵ(tΣ))×{t}.\displaystyle\mathrel{\mathop{\mathchar 58\relax}}=\bigcup_{t\in[t_{1},t_{2}]}(\mathbb{R}^{3}\backslash B_{\epsilon}(\sqrt{-t}\Sigma))\times\{t\}.

We prove the fourth claim before the third. In order to prove the fourth claim, it suffices to show for some 1<T<0-1<T<0, there exists no 22-chain CWout𝒮1[0,T]C\subset W^{\mathcal{S}^{1}}_{\mathrm{out}}[0,T] such that C=β1β2\partial C=\beta_{1}-\beta_{2}, where β2\beta_{2} is a closed 11-chain outside S1(T)S^{1}(T). Since S1(1)S^{1}(-1) lies outside Σ\Sigma, by the avoidance principle it suffices to prove that:

Lemma 3.17.

For some 1<T<0-1<T<0, there does not exists a 22-chain CY[1,T]C\subset Y[-1,T] such that C=β1β2\partial C=\beta_{1}-\beta_{2} for some closed 11-chain β2Y(T)\beta_{2}\subset Y(T).

Proof.

Choose a value of TT that is sufficiently close to 0 such that diam(TΣ)<ϵ\mathrm{diam}(\sqrt{-T}\Sigma)<\epsilon. With this choice, the set Bϵ(TΣ)B_{\epsilon}(\sqrt{-T}\Sigma) is star-shaped with respect to any point on TΣ\sqrt{-T}\Sigma. Thus, the boundary Bϵ(TΣ)\partial B_{\epsilon}(\sqrt{-T}\Sigma) has genus 0.

Suppose by contradiction that there exists a 22-chain CY[1,T]C\subset Y[-1,T] such that C=β1β2\partial C=\beta_{1}-\beta_{2} for some closed 11-chain β2Y(T)\beta_{2}\subset Y(T). By rescaling CC at each time slice tt, we can construct another 22-chain C~\widetilde{C} outside Σ\Sigma such that C~=β1Tβ2\partial\widetilde{C}=\beta_{1}-\sqrt{-T}\beta_{2}.

Since β1\beta_{1}, which lies outside Σ\Sigma, is homologically non-trivial, we can pick a non-trivial loop α\alpha inside Σ\Sigma such that [β1]H1(3\α)[\beta_{1}]\in H_{1}(\mathbb{R}^{3}\backslash\alpha) is non-trivial. Then by the existence of C~\widetilde{C}, we have [β2]0[\beta_{2}]\neq 0 in H1(3\α)H_{1}(\mathbb{R}^{3}\backslash\alpha) too. However, this is impossible because Tβ2\sqrt{-T}\beta_{2} lies outside Bϵ/T(Σ)B_{\epsilon/\sqrt{-T}}(\Sigma) while α\alpha lies inside, and Bϵ/T(Σ)\partial B_{\epsilon/{\sqrt{-T}}}(\Sigma) has genus 0 by the first paragraph of this proof. ∎

This finishes proving the fourth claim of Proposition 3.16. Finally, for the third claim, since a1(T)a_{1}(T) exists for the flow {tΣ}t0\{\sqrt{-t}\Sigma\}_{t\leq 0}, it follows from the avoidance principle that a1(T)a_{1}(T) exists for 𝒮1\mathcal{S}^{1}. Moreover, since the inside of S1(T)S^{1}(T) contains Bϵ(TΣ)B_{\epsilon}(\sqrt{-T}\Sigma), which is star-shaped, we know a1(T)=0a_{1}(T)=0 in H1(W𝒮1(T))H_{1}(W^{\mathcal{S}^{1}}(T)). ∎

4. Homology breakage of MCF through cylindrical and spherical singularities

4.1. MCF through cylindrical and spherical singularities

In this section, we focus on 2-dimensional MCF ={M(t)}t0\mathcal{M}=\{M(t)\}_{t\geq 0} through cylindrical and spherical singularities in 3\mathbb{R}^{3}, where the initial condition M(0)M(0) is a smooth, closed surface.

Proposition 4.1.

For any T00T_{0}\geq 0, no element of H1(Wout(T0))H_{1}(W_{\mathrm{out}}(T_{0})) can break at an inward neck singularity, and no element of H1(Win(T0))H_{1}(W_{\mathrm{in}}(T_{0})) can break at an outward neck singularity.

Proof.

Let us just prove the first claim. Suppose by contradiction some c0H1(Wout(T0))c_{0}\in H_{1}(W_{\mathrm{out}}(T_{0})) breaks at an inward neck singularity (x,T)(x,T), with T>T0T>T_{0}. Applying Theorem 2.5 to (x,T)(x,T) with δ0=1\delta_{0}=1 and any R>R0R>R_{0}, we obtain constants δ,δ¯>0\delta,\bar{\delta}>0 and a dense subset J[Tδ¯,T+δ¯]J\subset[T-\bar{\delta},T+\bar{\delta}] satisfying the properties in Theorem 2.5. Let B=Bδ(x)B=B_{\delta}(x).

Pick a time tJ[Tδ¯,T)t\in J\cap[T-\bar{\delta},T). Since c0c_{0} breaks at TT, c0(t)c_{0}(t) still exists. Pick a loop γc0(t)\gamma\in c_{0}(t). By Theorem 2.5 (6) (and recall Figure 6), we can homotope γ\gamma within Wout(t)W_{\mathrm{out}}(t) to avoid BB. This can be done for all tt in J[Tδ¯,T)J\cap[T-\bar{\delta},T), which is dense in [Tδ¯,T)[T-\bar{\delta},T). So we obtain a contradiction to the fact that c0c_{0} breaks at (x,T)(x,T). ∎

In the following proposition, we provide a more detailed description of the shape around a neck pinch at which homology class breaks. Namely, in this case, prior to the singular time, only the last bullet point of Theorem 2.5 (6) can occur, i.e. M(t)BM(t)\cap B is a cylinder.

Proposition 4.2.

There exists a universal constant R0>0R_{0}>0 with the following significance. Suppose c0H1(Win(T0))c_{0}\in H_{1}(W_{\mathrm{in}}(T_{0})) breaks at some inward neck singularity (x,T)(x,T). Let δ0>0\delta_{0}>0. Then for each R>R0R>R_{0}, there exist constants δ(0,δ0)\delta\in(0,\delta_{0}), δ¯>0\bar{\delta}>0, and a dense subset J(Tδ¯,T+δ¯)J\subset(T-\bar{\delta},T+\bar{\delta}) with Tδ¯JT-\bar{\delta}\in J, such that:

  1. (1)

    The first five items of Theorem 2.5 hold.

  2. (2)

    For each tJ[Tδ¯,T)t\in J\cap[T-\bar{\delta},T), Kin(t)Bδ(x)K_{\mathrm{in}}(t)\cap B_{\delta}(x) is a solid cylinder such that its boundary consists of a connected component of M(t)Bδ(x)M(t)\cap B_{\delta}(x) that is a cylinder and two disks D1,D2D_{1},D_{2} on Bδ(x)\partial B_{\delta}(x).

  3. (3)

    Moreover, for such tt, every element γc0(t)\gamma\in c_{0}(t) has a non-zero intersection number (in \mathbb{Z}-coefficients) with each DiD_{i}.

The outward case is analogous.

Proof.

We will just prove the inward case. Let us apply Theorem 2.5 to (x,T)(x,T) to obtain the constants δ,δ¯\delta,\bar{\delta} and the subset J[Tδ¯,T+δ¯]J\subset[T-\bar{\delta},T+\bar{\delta}]. Let B=Bδ(x)B=B_{\delta}(x). In addition the first five items of Theorem 2.5 will hold.

We need to show that for each tJ(T0,T)t\in J\cap(T_{0},T) sufficiently close to TT, Kin(t)Bδ(x)K_{\mathrm{in}}(t)\cap B_{\delta}(x) satisfies the description in (2): After that we could just shrink δ¯\bar{\delta} and the set JJ to guarantee (2). Suppose by contradiction that there exists a sequence in JJ, t1,t2,Tt_{1},t_{2},...\uparrow T, such that Kin(ti)Bδ(x)K_{\mathrm{in}}(t_{i})\cap B_{\delta}(x) violates the description in (2). Fix one tit_{i}. Note that Theorem 2.5 (5) and (6) together imply that Kin(ti)BK_{\mathrm{in}}(t_{i})\cap B can have at most one cylindrical component. Thus, in our case, Kin(ti)BK_{\mathrm{in}}(t_{i})\cap B actually has no cylindrical component. As a result, any connected component KK of Kin(ti)BK_{\mathrm{in}}(t_{i})\cap B satisfies either one of the following by Theorem 2.5 (6):

  • K\partial K is a connected component of M(t)BM(t)\cap B that is a sphere.

  • K\partial K consists of a connected component of M(t)BM(t)\cap B that is an disc and another disc on B\partial B.

In either situation, any element of c0(ti)c_{0}(t_{i}) can be perturbed to avoid BB. Applying this argument to each tit_{i}, we obtain a contradiction to the fact that c0c_{0} breaks at (x,T)(x,T).

Finally, to prove (3), it suffices to show that for each tJ(T0,T)t\in J\cap(T_{0},T) sufficiently close to TT, c0(t)c_{0}(t) satisfies the description of (3): Then we could just shrink JJ, and we would be done. Suppose otherwise, so that there exists a sequence in JJ, t1,t2,Tt_{1},t_{2},...\uparrow T, such that c0(ti)c_{0}(t_{i}) violates the description of (3). Then for each tit_{i}, we can find a loop γc0(ti)\gamma\in c_{0}(t_{i}) with intersection number zero with some connected component of Kin(ti)BK_{\mathrm{in}}(t_{i})\cap\partial B. In fact, since Kin(ti)BK_{\mathrm{in}}(t_{i})\cap B is a cylinder by (2), γ\gamma has intersection number zero with both connected components D1,D2D_{1},D_{2} of Kin(ti)BK_{\mathrm{in}}(t_{i})\cap\partial B (which are discs). To contradict the fact that c0c_{0} breaks at (x,T)(x,T), it suffices to find another element of c0(ti)c_{0}(t_{i}) that avoids BB.

Indeed, this can be proved as follows. We can assume γ\gamma intersects B\partial B transversely. Since γ\gamma has intersection number zero with D1D_{1}, we can pair up each positive intersection point of γD1\gamma\cap D_{1} with a negative one. Now fix a pair, and draw a line segment LL on D1D_{1} to connect the pair of points. Adding LL and L-L to γ\gamma, and slightly pushing the resulting curve away from D1D_{1} around L,LL,-L, we can obtain another representative of c0(ti)c_{0}(t_{i}) that avoids this pair of intersection points. And we do this for each pair. Then at the end, we get a curve belonging to c0(ti)c_{0}(t_{i}) that avoids D1D_{1} completely. Then, we repeat this process with D2D_{2}, to get a curve that avoids D2D_{2} too. Lastly, we discard all connected components of the curve that are in KK, which are all trivial as KK is a solid cylinder, to obtain an element of c0(ti)c_{0}(t_{i}) that avoids BB, as desired. ∎

Denote by 𝒮spherein\mathcal{S}^{\mathrm{in}}_{\mathrm{sphere}} the set of inward spherical singularities of \mathcal{M}, and by 𝒮neckin\mathcal{S}^{\mathrm{in}}_{\mathrm{neck}} the set of inward neck singularities of \mathcal{M}. Similarly, we define 𝒮sphereout\mathcal{S}^{\mathrm{out}}_{\mathrm{sphere}} and 𝒮neckout\mathcal{S}^{\mathrm{out}}_{\mathrm{neck}}. Then, we denote by Sspherein(t)3S^{\mathrm{in}}_{\mathrm{sphere}}(t)\subset\mathbb{R}^{3} the slice of 𝒮spherein\mathcal{S}^{\mathrm{in}}_{\mathrm{sphere}} at time tt, and proceed similarly for the other three sets.

Lemma 4.3.

Sneckin(T)S_{\mathrm{neck}}^{\mathrm{in}}(T) and Sneckout(T)S_{\mathrm{neck}}^{\mathrm{out}}(T) are compact sets.

Proof.

We only show Sneckin(T)S_{\mathrm{neck}}^{\mathrm{in}}(T) is compact and the proof for Sneckout(T)S_{\mathrm{neck}}^{\mathrm{out}}(T) is the same. It suffices to show Sneckin(T)¯=Sneckin(T)\overline{S_{\mathrm{neck}}^{\mathrm{in}}(T)}=S_{\mathrm{neck}}^{\mathrm{in}}(T). By the semi-continuity of the Gaussian density, a limit point pp of Sneckin(T)S_{\mathrm{neck}}^{\mathrm{in}}(T) must be a neck singularity. Hence it suffices to show pSneckin(T)p\in S_{\mathrm{neck}}^{\mathrm{in}}(T). We prove by contradiction: Suppose not, then pSneckout(T)p\in S_{\mathrm{neck}}^{\mathrm{out}}(T), and by mean convex neighborhood theorem, there is a neighborhood UU of pp and δ>0\delta>0 such that the MCF {Mt}t[Tδ,T+δ]\{M_{t}\}_{t\in[T-\delta,T+\delta]} in UU moves outward. This contradicts the assumption that pp is a limit point of Sneckin(T)S_{\mathrm{neck}}^{\mathrm{in}}(T). ∎

Proposition 4.4.

Suppose c0H1(Win(T0))c_{0}\in H_{1}(W_{\mathrm{in}}(T_{0})) terminates at some time T>T0T>T_{0}. Then c0c_{0} breaks in (Sneckin(T),T)(S^{\mathrm{in}}_{\mathrm{neck}}(T),T). The outward case is analogous.

Proof.

We will only prove the inward case, as the outward case follows analogously. Suppose the otherwise: There exist a neighborhood UU of Sneckin(T)S^{\mathrm{in}}_{\mathrm{neck}}(T) in 3\mathbb{R}^{3}, an increasing sequence of times t1,t2,Tt_{1},t_{2},...\uparrow T, and elements γic0(ti)\gamma_{i}\in c_{0}(t_{i}) such that each γi\gamma_{i} is disjoint from UU.

By the mean convex neighborhood theorem and the compactness of Sneckin(T)S^{\mathrm{in}}_{\mathrm{neck}}(T) and Sneckout(T)S^{\mathrm{out}}_{\mathrm{neck}}(T) from Lemma 4.3, we can further pick open neighborhoods Uin,U~inU_{\mathrm{in}},\widetilde{U}_{\mathrm{in}} with

Sneckin(T)UinU~inU,S^{\mathrm{in}}_{\mathrm{neck}}(T)\subset U_{\mathrm{in}}\subset\subset\widetilde{U}_{\mathrm{in}}\subset\subset U,

an open neighborhood UoutU_{\mathrm{out}} of Sneckout(T)S^{\mathrm{out}}_{\mathrm{neck}}(T), and two times T1<T<T2T_{1}<T<T_{2} such that:

  • U~in\widetilde{U}_{\mathrm{in}} and UoutU_{\mathrm{out}} are disjoint.

  • In the time interval (T1,T2)(T_{1},T_{2}), M(t)U~inM(t)\cap\widetilde{U}_{\mathrm{in}} evolves inward (i.e.

    Kin(t2)U~inKin(t1)\M(t)K_{\mathrm{in}}(t_{2})\cap\widetilde{U}_{\mathrm{in}}\subset K_{\mathrm{in}}(t_{1})\backslash M(t)

    for every T1<t1<t2<T2T_{1}<t_{1}<t_{2}<T_{2}) while M(t)UoutM(t)\cap U_{\mathrm{out}} evolves outward.

By Huisken’s analysis of spherical singularities (see also the special case of [CM16, Theorem 4.6]), each spherical singularity is isolated in spacetime. Therefore, the limit points of spherical singularities can only be cylindrical singularities.

We claim that after appropriately shrinking the time interval [T1,T2][T_{1},T_{2}],

(3\(UinUout))×[T1,T2](\mathbb{R}^{3}\backslash(U_{\mathrm{in}}\cup U_{\mathrm{out}}))\times[T_{1},T_{2}]

has only finitely many singular points, and we can thus assume such singular points all are spherical singularities at time TT. In fact, suppose not, there exists a sequence of distinct singular points {pi}i=1\{p_{i}\}_{i=1}^{\infty} outside UinUoutU_{\mathrm{in}}\cup U_{\mathrm{out}}, with singular time tiTt_{i}\to T. Then by the compactness of the singular set of \mathcal{M} and the previous paragraph, there is a subsequence converging to a cylindrical singularity in (Sneckin(T)Sneckout(T))×{T}(S^{\mathrm{in}}_{\mathrm{neck}}(T)\cup S^{\mathrm{out}}_{\mathrm{neck}}(T))\times\{T\}. This contradicts our choice of the pip_{i}’s.

As a consequence of the claim, by shrinking [T1,T2][T_{1},T_{2}] and the neighborhoods U~in\widetilde{U}_{\mathrm{in}} and UoutU_{\mathrm{out}}, we can assume

U~in\Uin¯×[T1,T2]\overline{\widetilde{U}_{\mathrm{in}}\backslash U_{\mathrm{in}}}\times[T_{1},T_{2}]

consists only of smooth points. Furthermore, we can choose a neighborhood VinV_{\mathrm{in}} of Sspherein(T)\U~in¯S^{\mathrm{in}}_{\mathrm{sphere}}(T)\backslash\overline{\widetilde{U}_{\mathrm{in}}} such that M(t)VinM(t)\cap V_{\mathrm{in}} is a finite union of convex smooth spheres for each t[T1,T2]t\in[T_{1},T_{2}], using what we proved in the previous paragraph. Similarly, we can find a neighborhood VoutV_{\mathrm{out}} for Ssphereout(T)\U~out¯S^{\mathrm{out}}_{\mathrm{sphere}}(T)\backslash\overline{\widetilde{U}_{\mathrm{out}}} with analogous properties. We can assume the closures of U~in,Uout,Vin,Vout\widetilde{U}_{\mathrm{in}},U_{\mathrm{out}},V_{\mathrm{in}},V_{\mathrm{out}} are all disjoint. Moreover, M(t)\(UinUoutVinVout)M(t)\backslash(U_{\mathrm{in}}\cup U_{\mathrm{out}}\cup V_{\mathrm{in}}\cup V_{\mathrm{out}}) evolves smoothly for t[T1,T2]t\in[T_{1},T_{2}].

To derive a contradiction to 𝔱(c0)=T\mathfrak{t}(c_{0})=T, we are going to prove that for some tit_{i}, there exists a smooth deformation of γi\gamma_{i}, {γtWin(t)}t[ti,T]\{\gamma^{t}\subset W_{\mathrm{in}}(t)\}_{t\in[t_{i},T]}, with γti=γi\gamma^{t_{i}}=\gamma_{i}, thereby letting γi\gamma_{i} “survive” up to time TT. Note that:

  • By the smoothness of M(t)M(t) in U~in\Uin¯\overline{\widetilde{U}_{\mathrm{in}}\backslash U_{\mathrm{in}}} for t[T1,T2]t\in[T_{1},T_{2}],

    C:=supt[T1,T2],xM(t)U~in\Uin¯|A|<.C\mathrel{\mathop{\mathchar 58\relax}}=\sup_{t\in[T_{1},T_{2}],\;x\in M(t)\cap\overline{\widetilde{U}_{\mathrm{in}}\backslash U_{\mathrm{in}}}}|A|<\infty.

    Thus, the velocity of the flow in this spacetime region is bounded by CC. Thus, since γi\gamma_{i} avoids U~in\widetilde{U}_{\mathrm{in}}, we can take a ti(T1,T)t_{i}\in(T_{1},T) sufficiently close to TT such that there is not enough time for any point of M(ti)\U~inM(t_{i})\backslash\widetilde{U}_{\mathrm{in}} to be pushed into UinU_{\mathrm{in}} by time TT.

  • Note that M(t)M(t) evolves outward in U~out\widetilde{U}_{\mathrm{out}} for t[T1,T2]t\in[T_{1},T_{2}].

  • Since VinV_{\mathrm{in}} and VoutV_{\mathrm{out}} consists of spheres, we can remove the components of γi\gamma_{i} inside the spheres, so we may assume γi\gamma_{i} avoids VinV_{\mathrm{in}} and VoutV_{\mathrm{out}}.

Combining the above observations, we can construct a smooth deformation of γi\gamma_{i}, {γtWin(t)}t[ti,T]\{\gamma^{t}\subset W_{\mathrm{in}}(t)\}_{t\in[t_{i},T]}, using the evolution of MCF, with γti=γi\gamma^{t_{i}}=\gamma_{i}. This contradicts that 𝔱(c0)=T\mathfrak{t}(c_{0})=T. ∎

Here comes a key theorem which supports that our definition of homology termination and breakage would accurately describe the heuristic phenomenon shown in Figure 8.

Theorem 4.5.

Suppose c0H1(Win(T0))c_{0}\in H_{1}(W_{\mathrm{in}}(T_{0})) terminates at some time T>T0T>T_{0}. Then c0c_{0} breaks at some inward neck singularity (x,T)(x,T).

The outward case is analogous.

Note that such xx may be non-unique: Consider a flow that is a thin torus collapsing into a closed curve consisting entirely of neck singularities.

Proof.

We prove the inward case as the outward case is analogous. We will prove by contradiction. Suppose that the theorem is false, meaning:

Assumption (\star): For every inward neck singularity (x,T)(x,T), there is a neighborhood UxU_{x} of xx such that it is not true that “for every time t<Tt<T close enough to TT, every element of c0(t)c_{0}(t) intersects UxU_{x}”.

Applying Theorem 2.5 to each inward neck singularity (x,T)(x,T), with a constant δ0(x)>0\delta_{0}(x)>0 such that Bδ0(x)(x)UxB_{\delta_{0}(x)}(x)\subset U_{x} and an R>max{R0,100}R>\max\{R_{0},100\}, we obtain constants δ(x),δ¯(x)>0\delta(x),\bar{\delta}(x)>0 and a set of full measure J(x)[Tδ¯(x),T+δ¯(x)]J(x)\subset[T-\bar{\delta}(x),T+\bar{\delta}(x)] satisfying the properties of Theorem 2.5.

Since Sneckin(T)S^{\mathrm{in}}_{\mathrm{neck}}(T) is compact by Lemma 4.3, there exist x1,,xnSneckin(T)x_{1},...,x_{n}\in S^{\mathrm{in}}_{\mathrm{neck}}(T) such that

Bδ(x1)/2(x1),,Bδ(xn)/2(xn)B_{\delta(x_{1})/2}(x_{1}),...,B_{\delta(x_{n})/2}(x_{n})

cover Sneckin(T)S^{\mathrm{in}}_{\mathrm{neck}}(T). For simplicity, we denote those balls by 12B1,,12Bn\frac{1}{2}B_{1},...,\frac{1}{2}B_{n}, while

B1:=Bδ(x1)(x1),,Bn:=Bδ(xn)(xn)B_{1}\mathrel{\mathop{\mathchar 58\relax}}=B_{\delta(x_{1})}(x_{1}),...,B_{n}\mathrel{\mathop{\mathchar 58\relax}}=B_{\delta(x_{n})}(x_{n})

Since c0c_{0} terminates at time TT, we know that c0c_{0} breaks in (Sneckin(T),T)(S^{\mathrm{in}}_{\mathrm{neck}}(T),T) by Proposition 4.4. Thus, by definition, there exists a time T1T_{1} with maxiTδ¯(xi)<T1<T\max_{i}T-\bar{\delta}(x_{i})<T_{1}<T such that for each t[T1,T)t\in[T_{1},T), every element of c0(t)c_{0}(t) intersects i12Bi\cup_{i}\frac{1}{2}B_{i}. We can assume T1iJ(xi)T_{1}\in\cap_{i}J(x_{i}) so that M(T1)M(T_{1}) is smooth and intersects each Bi\partial B_{i} transversely by Theorem 2.5 (3).

Lemma 4.6.

Let DD be a connected component of Kin(T1)BiK_{\mathrm{in}}(T_{1})\cap\partial B_{i} (of which there are at most two according to Theorem 2.5 (5)), and γc0(T1)\gamma\in c_{0}(T_{1}). Then, it follows that the linking number link(γ,D)=0\mathrm{link}(\gamma,\partial D)=0.

Proof.

Suppose the otherwise, that there exists some DD as above and γc0(t0)\gamma\in c_{0}(t_{0}) such that link(γ,D)0\mathrm{link}(\gamma,\partial D)\neq 0. Now, pick any t1[T1,T)t_{1}\in[T_{1},T) and γ1c0(t1)\gamma_{1}\in c_{0}(t_{1}). By definition, γ1\gamma_{1} is homologous to γ\gamma within Win[T1,t1]W_{\mathrm{in}}[T_{1},t_{1}]. Thus, γ1\gamma_{1} is homologous to γ\gamma within 3\D\mathbb{R}^{3}\backslash\partial D, as the mean convex neighborhood property (Theorem 2.5 (2)) implies that D3\Win(t)\partial D\subset\mathbb{R}^{3}\backslash W_{\mathrm{in}}(t) for all t[T1,t1]t\in[T_{1},t_{1}]. Therefore, link(γ1,D)0\mathrm{link}(\gamma_{1},\partial D)\neq 0, which implies that γ1\gamma_{1} must intersect DD. However, since DB¯iUxiD\subset\bar{B}_{i}\subset U_{x_{i}}, this implies that for all t1[T1,T)t_{1}\in[T_{1},T), any element of c0(t1)c_{0}(t_{1}) must intersect UxiU_{x_{i}}. This contradicts the assumption (\star). ∎

Let ϵ1:=miniδ(xi)/2\epsilon_{1}\mathrel{\mathop{\mathchar 58\relax}}=\min_{i}\delta(x_{i})/2. Let γc0(T1)\gamma\in c_{0}(T_{1}) be such that

(3) length(γ)<infγc0(T1)length(γ)+ϵ1/100\mathrm{length}(\gamma)<\inf_{\gamma^{\prime}\in c_{0}(T_{1})}\mathrm{length}(\gamma^{\prime})+\epsilon_{1}/100

Without loss of generality, we can assume γ\gamma intersects all Bi\partial B_{i} transversely. To finish the proof, it suffices to show that γ\gamma avoids i12Bi\cup_{i}\frac{1}{2}B_{i}: This would contradict the definition of T1T_{1}.

Lemma 4.7.

γ\gamma does not intersect i12Bi\cup_{i}\frac{1}{2}B_{i}.

Proof.

We prove by contradiction. Suppose that γ\gamma intersects some 12Bi\frac{1}{2}B_{i}. We will produce an element of c0(T1)c_{0}(T_{1}) whose length is too small.

Without loss of generality, we can assume that no connected component of γBi\gamma\cap B_{i} is a closed loop. This is because we could just remove all such loops from γ\gamma, and the resulting curve is still in c0(T1)c_{0}(T_{1}) by Theorem 2.5 (6). Hence, letting β\beta be a connected component of γBi\gamma\cap B_{i}, we can assume that β\beta is a line segment.

Now, by Theorem 2.5 (5) and our choice that T1iJ(xi)T_{1}\in\cap_{i}J(x_{i}), Win(T1)BiW_{\mathrm{in}}(T_{1})\cap\partial B_{i} consists of at most two disks. There are two cases: Either (1) β\beta starts and ends on the same disk, say D1D_{1}, or (2) β\beta starts and ends on different disks, D1D_{1} and D2D_{2}. We will show that both are impossible.

For case (1), since β\beta intersects 12Bi\frac{1}{2}B_{i}, whose distance to Bi\partial B_{i} is δ(xi)/2\delta(x_{i})/2, we know that length(β)\mathrm{length}(\beta) is at least δ(xi)\delta(x_{i}). On the other hand, note that by Theorem 2.5 (1), (2), and (4), D1D_{1} is a convex disc on Bi\partial B_{i} with diameter less than δ(xi)/50\delta(x_{i})/50 (recall R>100R>100). Thus, we can join the end points of β\beta, from β(1)\beta(1) to β(0)\beta(0), by a segment β1\beta_{1} on D1D_{1} of length less than δ(xi)/50\delta(x_{i})/50: See Figure 10. Then, we consider the new loop γββ\gamma-\beta-\beta^{\prime}, which replaces βγ\beta\subset\gamma with β\beta^{\prime}. This loop lies in c0(T1)c_{0}(T_{1}), because β+β\beta+\beta^{\prime} bounds a disc in Win(T1)B¯iW_{\mathrm{in}}(T_{1})\cap\bar{B}_{i} by Theorem 2.5 (6).

Refer to caption

Figure 10.

Moreover, this new loop is impossibly short:

length(γββ)\displaystyle\mathrm{length}(\gamma-\beta-\beta^{\prime}) length(γ)δ(xi)+δ(xi)/50\displaystyle\leq\mathrm{length}(\gamma)-\delta(x_{i})+\delta(x_{i})/50
<length(γ)δ(xi)/2\displaystyle<\mathrm{length}(\gamma)-\delta(x_{i})/2
length(γ)ϵ1\displaystyle\leq\mathrm{length}(\gamma)-\epsilon_{1}
<infγc0(T1)length(γ),\displaystyle<\inf_{\gamma^{\prime}\in c_{0}(T_{1})}\mathrm{length}(\gamma^{\prime}),

in which the last inequality is from the definition of γ\gamma. Thus, a contradiction arises, and case (1) is impossible.

For case (2), suppose the starting point β(0)\beta(0) is in D1D_{1} and the ending point β(1)\beta(1) is in D2D_{2}. We claim that there is another connected component β^\hat{\beta} of γBi\gamma\cap B_{i} such that starting point β^(0)\hat{\beta}(0) is in D2D_{2} and ending point β^(1)\hat{\beta}(1) is in D1\in D_{1}. This claim follows immediately from:

  • By Theorem 2.5 (6), M(T1)BiM(T_{1})\cap\partial B_{i} is a cylinder.

  • By Lemma 4.6, link(γ,D1)=link(γ,D2)=0\mathrm{link}(\gamma,\partial D_{1})=\mathrm{link}(\gamma,\partial D_{2})=0.

  • Case (1) was proven impossible.

Finally, let β1\beta_{1} be a segment on D1D_{1} connecting β^(1)\hat{\beta}(1) to β(0)\beta(0), and β2\beta_{2} be a segment on D2D_{2} connecting β^(0)\hat{\beta}(0) to β(1)\beta(1) (see Figure 11). As in case (1), we can guarantee length(β1),length(β2)<δ(xi)/50\mathrm{length}(\beta_{1}),\mathrm{length}(\beta_{2})<\delta(x_{i})/50. Hence, we consider the new loop γββ^β1β2\gamma-\beta-\hat{\beta}-\beta_{1}-\beta_{2}, which replaces β+β^γ\beta+\hat{\beta}\subset\gamma with β1β2-\beta_{1}-\beta_{2}. This new loop lies in c0(T1)c_{0}(T_{1}), because β+β^+β1+β2\beta+\hat{\beta}+\beta_{1}+\beta_{2} bounds a disc in Win(T1)B¯iW_{\mathrm{in}}(T_{1})\cap\bar{B}_{i} by Theorem 2.5 (6). Moreover, as in case (1), we can show that

length(γββ^β1β2)<infγc0(T1)length(γ),\mathrm{length}(\gamma-\beta-\hat{\beta}-\beta_{1}-\beta_{2})<\inf_{\gamma^{\prime}\in c_{0}(T_{1})}\mathrm{length}(\gamma^{\prime}),

which is a contradiction. Therefore, case (2) is also impossible. This leads to a contradiction.

Refer to caption

Figure 11.

This finishes the proof of Theorem 4.5. ∎

4.2. MCF through cylindrical and spherical singularities from torus

In §4.2 , we will focus on 2-dimensional MCF ={M(t)}t0\mathcal{M}=\{M(t)\}_{t\geq 0} through cylindrical and spherical singularities in 3\mathbb{R}^{3}, where M(0)M(0) is a smooth torus. The main goal of §4.2 is to prove the following.

Theorem 4.8.

The setting is as follows.

  • Let {M(t)}t0\{M(t)\}_{t\geq 0} be a MCF through cylindrical and spherical singularities with M(0)M(0) a smooth torus in 3\mathbb{R}^{3}.

  • Let a0a_{0} be a generator of H1(Win(0))H_{1}(W_{\mathrm{in}}(0))\cong\mathbb{Z}, and b0b_{0} be a generator of H1(Wout(0))H_{1}(W_{\mathrm{out}}(0))\cong\mathbb{Z}.

  • Let T=min{𝔱(a0),𝔱(b0)}.T=\min\{\mathfrak{t}(a_{0}),\mathfrak{t}(b_{0})\}.

Then T<T<\infty, and genus(M(t))=1\mathrm{genus}(M(t))=1 for a.e. t<Tt<T, while genus(M(t))=0\mathrm{genus}(M(t))=0 or M(t)M(t) is empty for a.e. t>Tt>T.

Throughout §4.2, we will retain the notations in this theorem.

Let us first sketch the proof. By [CM16], M(t)M(t) is smooth for a.e. time. And by [Whi95], genus(M(t))\mathrm{genus}(M(t)), when well-defined, is non-increasing in tt. Thus, there exists some time TgT_{g} such that genus(M(t))=1\mathrm{genus}(M(t))=1 for a.e. t<Tgt<T_{g}, while genus(M(t))=0\mathrm{genus}(M(t))=0 or M(t)M(t) is empty for a.e. t>Tgt>T_{g}. Our goal is to show T=TgT=T_{g}.

The proof consists of proving the following six claims one-by-one:

  • T<T<\infty.

  • Let t0t\geq 0. If M(t)M(t) is a smooth torus and a0(t)a_{0}(t) exists, then a0(t)a_{0}(t) generates H1(Win(t))H_{1}(W_{\mathrm{in}}(t)). And the case for b0b_{0} is analogous.

  • TgTT_{g}\geq T.

  • 𝔱(a0)𝔱(b0)\mathfrak{t}(a_{0})\neq\mathfrak{t}(b_{0}).

  • If 𝔱(a0)<𝔱(b0)\mathfrak{t}(a_{0})<\mathfrak{t}(b_{0}), then b0(t)b_{0}(t) is trivial for each t>𝔱(a0)t>\mathfrak{t}(a_{0}). And if 𝔱(b0)<𝔱(a0)\mathfrak{t}(b_{0})<\mathfrak{t}(a_{0}), then a0(t)a_{0}(t) is trivial for each t>𝔱(b0)t>\mathfrak{t}(b_{0}).

  • TgTT_{g}\leq T.

We now begin the proof.

Proposition 4.9.

T<T<\infty.

Proof.

Suppose otherwise, i.e. a0a_{0} and b0b_{0} both never terminate. Since M(0)M(0) is compact, eventually Kout(t)=3K_{\mathrm{out}}(t)=\mathbb{R}^{3}. So a0(T)a_{0}(T) and b0(T)b_{0}(T) both become trivial for some large T>0T>0. As a result, if we pick some loops α0a0\alpha_{0}\in a_{0} and β0b0\beta_{0}\in b_{0}, then there exist 2-chain AWin[0,T]A\subset W_{\mathrm{in}}[0,T] and BWout[0,T]B\subset W_{\mathrm{out}}[0,T] such that A=α0\partial A=\alpha_{0} and B=β0\partial B=\beta_{0}.

Now, denote by B^3×[T,0]\hat{B}\subset\mathbb{R}^{3}\times[-T,0] the reflection of BB across 3×{0}\mathbb{R}^{3}\times\{0\}. Let B~=BB^\widetilde{B}=B\cup\hat{B}, which can be viewed as a closed 2-chain in 4\mathbb{R}^{4}. Then we view A4\B~A\subset\mathbb{R}^{4}\backslash\widetilde{B}. Thus, to derive a contradiction, it suffices to show that α0\alpha_{0} is homologically non-trivial in 4\B~\mathbb{R}^{4}\backslash\widetilde{B}.

Without loss of generality, we can assume B~\widetilde{B} is connected by discarding all those connected components that do not contain β0\beta_{0}. By Alexander duality,

H1(4\B~)H2(B~).H_{1}(\mathbb{R}^{4}\backslash\widetilde{B})\cong H^{2}(\widetilde{B})\cong\mathbb{Z}.

One can check that α04\B~\alpha_{0}\subset\mathbb{R}^{4}\backslash\widetilde{B} actually generates \mathbb{Z} as the linking number link(a0,b0)=1\mathrm{link}(a_{0},b_{0})=1. This shows α0\alpha_{0} is homologically non-trivial in 4\B~\mathbb{R}^{4}\backslash\widetilde{B}, contradicting the existence of AA. ∎

Remark 4.10.

The above proof works also in the case when M(0)M(0) is a closed surface of any genus with a0H1(Win(0))a_{0}\in H_{1}(W_{\mathrm{in}}(0)) and b0H1(Wout(0))b_{0}\in H_{1}(W_{\mathrm{out}}(0)) linked, and the flow {M(t)}t0\{M(t)\}_{t\geq 0} is a general level set flow (whose singularities are not necessarily cylindrical or spherical).

Proposition 4.11.

Let t0t\geq 0. If M(t)M(t) is a smooth torus and a0(t)a_{0}(t) exists, then a0(t)a_{0}(t) generates H1(Win(t))H_{1}(W_{\mathrm{in}}(t)). And the case for b0b_{0} is analogous.

Proof.

We will just prove the case for a0a_{0}. Let a¯\bar{a} be a generator of H1(Win(t))H_{1}(W_{\mathrm{in}}(t))\cong\mathbb{Z}. It suffices to show a¯=a0(t)\bar{a}=a_{0}(t) up to a sign.

By definition, there exists α0a0\alpha_{0}\in a_{0}, α1a0(t)\alpha_{1}\in a_{0}(t) such that α0α1=A\alpha_{0}-\alpha_{1}=\partial A for some AW[0,t]A\subset W[0,t]. On the other hand, pick a loop α¯1a¯\bar{\alpha}_{1}\in\bar{a}, then by [Whi95, Theorem 1 (ii)], there exists a homotopy HH in W[0,T]W[0,T] joining α¯1\bar{\alpha}_{1} back to some loop α¯0W(0)\bar{\alpha}_{0}\subset W(0) (which means H=α¯1α¯0\partial H=\bar{\alpha}_{1}-\bar{\alpha}_{0}). So [α¯0]=ka0[\bar{\alpha}_{0}]=ka_{0} for some integer kk, and so α¯0kα0=A0\bar{\alpha}_{0}-k\alpha_{0}=\partial A_{0} for some A0W(0)A_{0}\subset W(0). If we manage to show a0=[α¯0]a_{0}=[\bar{\alpha}_{0}] or [α¯0]-[\bar{\alpha}_{0}], then by the fact that a0a_{0} can only descend into one class at time tt (Proposition 3.3), we would know a0(t)=a¯a_{0}(t)=\bar{a} or a¯-\bar{a}, as desired. Hence, it suffices to show that k=±1k=\pm 1.

Let us glue H,A0,H,A_{0}, and kAkA together, so that we have

α¯1kα1=(H+A0+kA).\bar{\alpha}_{1}-k\alpha_{1}=\partial(H+A_{0}+kA).

Thus, since the inclusion H1(Win(t))H1(Win[0,t])H_{1}(W_{\mathrm{in}}(t))\to H_{1}(W_{\mathrm{in}}[0,t]) is injective by [Whi95, Theorem 1 (iii)], a¯=kα0(t)\bar{a}=k\alpha_{0}(t) in H1(Win(t))H_{1}(W_{\mathrm{in}}(t)). Since a¯\bar{a} is a generator by definition, k=±1k=\pm 1, as desired. ∎

Proposition 4.12.

TgTT_{g}\geq T.

Proof.

Let us assume T=𝔱(a0)T=\mathfrak{t}(a_{0}), as the other case T=𝔱(b0)T=\mathfrak{t}(b_{0}) is analogous. Recall that we have shown T<T<\infty. Since genus(M(t))\mathrm{genus}(M(t)), if well-defined, is non-increasing in tt, it suffices to prove that there exists T1<TT_{1}<T such that for a dense set of t(T1,T)t\in(T_{1},T), genus(M(t))=1\mathrm{genus}(M(t))=1.

By Theorem 4.5, T=𝔱(a0)T=\mathfrak{t}(a_{0}) implies a0a_{0} breaks at some inward neck singularity (x,T)(x,T). Then, applying Proposition 4.2 to (x,T)(x,T) with δ0=1\delta_{0}=1 and an R>R0R>R_{0}, we obtain constants δ,δ¯\delta,\bar{\delta} and a dense set J[Tδ¯,T+δ¯]J\subset[T-\bar{\delta},T+\bar{\delta}] with Tδ¯JT-\bar{\delta}\in J. We let T1=Tδ¯T_{1}=T-\bar{\delta}, and B=Bδ(x)B=B_{\delta}(x).

Now, fix any t(T1,T)t\in(T_{1},T), and DD let be one of the two connected component of Kin(t)BK_{\mathrm{in}}(t)\cap\partial B: Recall that Kin(t)BK_{\mathrm{in}}(t)\cap B is a solid cylinder by Proposition 4.2. By Proposition 4.2, some element αa0(t)\alpha\in a_{0}(t) has a non-zero intersection number with DD. Now, we push D\partial D slightly into Kout(t)BK_{\mathrm{out}}(t)\cap B and call that loop β\beta. Then the linking number link(β,α)\mathrm{link}(\beta,\alpha) is non-zero, with α\alpha inside M(t)M(t) and β\beta outside M(t)M(t). Hence, genus(M(t))\mathrm{genus}(M(t)) is non-zero, and thus has to be one, as desired. ∎

Proposition 4.13.

𝔱(a0)𝔱(b0)\mathfrak{t}(a_{0})\neq\mathfrak{t}(b_{0}).

Proof.

If 𝔱(b0)<𝔱(a0)\mathfrak{t}(b_{0})<\mathfrak{t}(a_{0}), we are done. So let us assume 𝔱(a0)𝔱(b0)\mathfrak{t}(a_{0})\leq\mathfrak{t}(b_{0}) and aim to show 𝔱(b0)>𝔱(a0)\mathfrak{t}(b_{0})>\mathfrak{t}(a_{0}).

Let us focus on the time t=T1t=T_{1}, with T1:=Tδ¯T_{1}\mathrel{\mathop{\mathchar 58\relax}}=T-\bar{\delta}, as defined in the proof of Proposition 4.12. We know genus(M(T1))=1\mathrm{genus}(M(T_{1}))=1 from before. Now, consider the loops αa0(T1)\alpha\in a_{0}(T_{1}) and βWout(T1)B\beta\subset W_{\mathrm{out}}(T_{1})\cap B defined in the previous proof. Then by Proposition 4.11, α\alpha is a generator of H1(Win(T1))H_{1}(W_{\mathrm{in}}(T_{1})), and from the construction of β\beta it is clear link(β,α)=±1\mathrm{link}(\beta,\alpha)=\pm 1. So β\beta actually generates H1(Wout(T1))H_{1}(W_{\mathrm{out}}(T_{1})). Then by Proposition 4.11 again and the assumption 𝔱(b0)𝔱(a0)\mathfrak{t}(b_{0})\geq\mathfrak{t}(a_{0}), we have [β]=b0(T1)[\beta]=b_{0}(T_{1}), possibly after changing the orientation of β\beta.

Finally, by the mean convex neighborhood property, βWout(T1)B\beta\subset W_{\mathrm{out}}(T_{1})\cap B will survive after time TT. So 𝔱(b0)>𝔱(a0)\mathfrak{t}(b_{0})>\mathfrak{t}(a_{0}). ∎

Proposition 4.14.

If 𝔱(a0)<𝔱(b0)\mathfrak{t}(a_{0})<\mathfrak{t}(b_{0}), then b0(t)b_{0}(t) exists and is trivial for each t>𝔱(a0)t>\mathfrak{t}(a_{0}). And if 𝔱(b0)<𝔱(a0)\mathfrak{t}(b_{0})<\mathfrak{t}(a_{0}), then a0(t)a_{0}(t) exists and is trivial for each t>𝔱(b0)t>\mathfrak{t}(b_{0}).

Proof.

We prove the first statement and the second statement is similar. Let us retain the notation from the previous proof. By Proposition 4.2, M(T1)BM(T_{1})\cap B (recall T1=Tδ¯T_{1}=T-\bar{\delta}) is close to a round cylinder. Now, enclose this cylinder by an Angenent torus, and run the MCF. Note that:

  • Since the time interval around TT given by the mean convex neighborhood property is independent of RR (in Proposition 4.2), we can, by making RR very large and thus the Angenent torus very small, assume that the mean convex neighborhood property still holds at the moment the Angenent torus vanishes.

  • By the avoidance principle, the distance between the Angenent torus and M(t)M(t) is non-decreasing.

Hence, when the Angenent torus vanishes, the neck M(t)BM(t)\cap B has already been “cut into disconnected pieces.” As a result, the loop β\beta, which remains disjoint from the evolving surface, would have become trivial at the moment the Angenent torus disappears.

Finally, note that as RR\to\infty, δ¯0\bar{\delta}\to 0 (see Theorem 2.5, item 1). By the definition of cylindrical singularity, we know T1=Tδ¯TT_{1}=T-\bar{\delta}\to T and M(Tδ¯)BM(T-\bar{\delta})\cap B tends to be an actual round cylinder after rescaling by the factor RR. This shows that the moment when Angenent torus vanishes will tend to TT. Therefore, b0(t)b_{0}(t) is trivial for each t>Tt>T. ∎

Finally, since we have already proven TgTT_{g}\geq T, to complete the proof of Theorem 4.8, it remains to show:

Proposition 4.15.

TgTT_{g}\leq T.

Proof.

Suppose by contradiction Tg>TT_{g}>T. Again, we can just consider the case 𝔱(a0)<𝔱(b0)\mathfrak{t}(a_{0})<\mathfrak{t}(b_{0}). By our Proposition 4.14, we can pick a time T2(T,Tg)T_{2}\in(T,T_{g}) when M(T2)M(T_{2}) is a smooth torus and b0(T2)b_{0}(T_{2}) exists and is trivial. This contradicts Proposition 4.11, which says that b0(T2)b_{0}(T_{2}) generates H1(Wout(T2))H_{1}(W_{\mathrm{out}}(T_{2})). ∎

This completes the proof of Theorem 4.8.

4.3. Termination time of limit of MCF

Finally, in §4.3, let us mention a proposition that describes a relationship between the termination time and a convergent sequence of initial conditions.

Proposition 4.16.

The setting is as follows.

  • Let i={Mi(t)}t0\mathcal{M}^{i}=\{M^{i}(t)\}_{t\geq 0}, i=1,2,i=1,2,..., and ={M(t)}t0\mathcal{M}=\{M(t)\}_{t\geq 0} all be MCF through cylindrical and spherical singularities, such that each Mi(0)M^{i}(0) and M(0)M(0) are smooth, close hypersurfaces.

  • For each ii, assume Mi(0)M^{i}(0) is sufficiently close in CC^{\infty} to M(0)M(0) such that each H1(Wi(0))H_{1}(W^{\mathcal{M}^{i}}(0)) can be canonically identified with H1(W(0))H_{1}(W^{\mathcal{M}}(0)). Moreover, Mi(0)M(0)M^{i}(0)\to M(0) in CC^{\infty}.

  • Let c0H1(W(0))c_{0}\in H_{1}(W^{\mathcal{M}}(0)). Note that c0c_{0} can be viewed as an element of H1(Wi(0))H_{1}(W^{\mathcal{M}^{i}}(0)) for each ii too.

Then

lim infi𝔱i(c0)𝔱(c0).\liminf_{i}\mathfrak{t}^{\mathcal{M}^{i}}(c_{0})\geq\mathfrak{t}^{\mathcal{M}}(c_{0}).
Proof.

Let T=𝔱(c0)T=\mathfrak{t}^{\mathcal{M}}(c_{0}).

We first consider the case T<T<\infty. Suppose by contradiction that there exists a subsequence {ik}k=1\{i_{k}\}_{k=1}^{\infty} and some T1<TT_{1}<T such that 𝔱ik(c0)T1\mathfrak{t}^{\mathcal{M}^{i_{k}}}(c_{0})\leq T_{1} for each kk. Pick some element γ0W(0)\gamma_{0}\subset W^{\mathcal{M}}(0) with [γ0]=c0[\gamma_{0}]=c_{0}, and γ1W(T1+T2)\gamma_{1}\subset W^{\mathcal{M}}(\frac{T_{1}+T}{2}) with [γ1]=c0(T1+T2)[\gamma_{1}]=c_{0}(\frac{T_{1}+T}{2}). By definition, γ0\gamma_{0} and γ1\gamma_{1} together bound some ΓW[0,T1+T2]\Gamma\subset W^{\mathcal{M}}[0,\frac{T_{1}+T}{2}].

Now, recall that i\mathcal{M}^{i}\to\mathcal{M} in the Hausdorff sense by Proposition 2.8. Thus, since Γ\Gamma is compact, for all sufficiently large ii, ΓWi[0,T1+T2]\Gamma\subset W^{\mathcal{M}^{i}}[0,\frac{T_{1}+T}{2}]. Moreover, γ0\gamma_{0} represents c0H1(Wi(0))c_{0}\in H_{1}(W^{\mathcal{M}^{i}}(0)) for such large ii. This contradicts that 𝔱ik(c0)T1\mathfrak{t}^{\mathcal{M}^{i_{k}}}(c_{0})\leq T_{1} for each kk.

Lastly, the case T=T=\infty can be done similarly using the fact that the flow \mathcal{M} vanishes in finite time. ∎

5. Proof of main theorems

5.1. Proof of Theorem 1.1

Suppose by contradiction that for each s[0,1]s\in[0,1], {Ms(t)}t0\{M^{s}(t)\}_{t\geq 0} is a MCF through cylindrical and spherical singularities. For each s[0,1]s\in[0,1], let

Ts=min{𝔱s(a0),𝔱s(b0)}.T^{s}=\min\{\mathfrak{t}^{\mathcal{M}^{s}}(a_{0}),\mathfrak{t}^{\mathcal{M}^{s}}(b_{0})\}.

Furthermore, Propositions 4.13 and 4.14 show that either a0a_{0} or b0b_{0} will terminate, but not both. Consequently, we can represent [0,1][0,1] as a disjoint union ABA\sqcup B, where AA contains those ss for which Ts=𝔱s(a0)T^{s}=\mathfrak{t}^{\mathcal{M}^{s}}(a_{0}), and BB contains those ss for which Ts=𝔱s(b0)T^{s}=\mathfrak{t}^{\mathcal{M}^{s}}(b_{0}). Note that 0A0\in A and 1B1\in B by the assumption. Thus, the following lemma leads us directly to a contradiction.

Lemma 5.1.

The sets AA and BB are both closed.

Proof.

We will just prove that AA is closed. Let s[0,1]s\in[0,1] be an accumulation point of AA, and pick a sequence sis_{i} in AA with siss_{i}\to s. Note that:

  • For each ii, by Theorem 4.8, genus(Msi(t))=1\mathrm{genus}(M^{s_{i}}(t))=1 for a.e. t<Tsit<T^{s_{i}} and genus(Msi(t))=0\mathrm{genus}(M^{s_{i}}(t))=0 for a.e. t>Tsit>T^{s_{i}}.

  • Similarly, genus(Ms(t))=1\mathrm{genus}(M^{s}(t))=1 for a.e. t<Tst<T^{s} and genus(Ms(t))=0\mathrm{genus}(M^{s}(t))=0 for a.e. t>Tst>T^{s}.

Thus, together with Proposition 2.8, which says Mis(t)Ms(t)M^{s}_{i}(t)\to M^{s}(t) in CC^{\infty} for a.e. t0t\geq 0, we know TsiTsT^{s_{i}}\to T^{s}. Hence,

Ts=lim infiTsi=lim infi𝔱si(a0)𝔱s(a0).T^{s}=\liminf_{i}T^{s_{i}}=\liminf_{i}\mathfrak{t}^{\mathcal{M}^{s_{i}}}(a_{0})\geq\mathfrak{t}^{\mathcal{M}^{s}}(a_{0}).

Note that the second equality holds because siAs_{i}\in A, and the inequality holds by Proposition 4.16. Thus, we know Ts=𝔱s(a0)T^{s}=\mathfrak{t}^{\mathcal{M}^{s}}(a_{0}), which means for the flow s\mathcal{M}^{s}, a0a_{0} will terminate but b0b_{0} will not. So sAs\in A. This shows that AA is closed. ∎

This finishes the proof of Theorem 1.1.

Remark 5.2.

Let us explain why Theorem 1.1 would not hold if the initial conditions have genus greater than one. For example, consider the genus 22 surface depicted in Figure 13, where a0a_{0} and b0b_{0} are linked as shown. Then, the MCF actually could develop both inward and outward cylindrical singularities simultaneously, with a0a_{0} breaking at the inward one and b0b_{0} breaking at the outward one. This phenomenon may prevent a genus one singularity from showing up in any intermediate flow between {M0(t)}t0\{M^{0}(t)\}_{t\geq 0} and {M1(t)}t0\{M^{1}(t)\}_{t\geq 0}, in the setting of Theorem 1.1.

One might think if we choose a0a_{0} and b0b_{0} better, like in Figure 13, then the conclusion of Theorem 1.1 may hold. However, Figure 13 and 13 are actually homotopic to each other. In conclusion, in a genus two surface, we cannot force a genus one singularity to appear just by topology: The geometry of the initial conditions must play a role.

Refer to caption
Figure 12.
Refer to caption
Figure 13.

5.2. Proof of Corollary 1.2

Let s:={Ms(t)}t0\mathcal{M}^{s}\mathrel{\mathop{\mathchar 58\relax}}=\{M^{s}(t)\}_{t\geq 0} be the level set flow starting from Ms(0):=MsM^{s}(0)\mathrel{\mathop{\mathchar 58\relax}}=M^{s}. We can apply Theorem 1.1 to the flows s\mathcal{M}^{s}, s[0,1]s\in[0,1], which shows there exists s0[0,1]s_{0}\in[0,1] such that s0\mathcal{M}^{s_{0}} has a singularity (x,T)(x,T) that is not (multiplicity one) cylindrical or spherical. In other words, every tangent flow \mathcal{M}^{\prime} at (x,T)(x,T) is not the shrinking cylinder or sphere of multiplicity one. Recall that by [Ilm95], \mathcal{M}^{\prime} is a smooth, embedded, self-shrinking flow {tmΣ}t<0\{\sqrt{-t}m\Sigma^{\prime}\}_{t<0} with genus at most one and has multiplicity mm. But the multiplicity can only be 11 by the entropy bound Ent(Ms0)<2\mathrm{Ent}(M^{s_{0}})<2 and the monotonicity formula. Thus, Σ\Sigma^{\prime} has genus 11.

5.3. Proof of Theorem 1.3

Note that we have Ent(Ms)<2\mathrm{Ent}(M^{s})<2 for each ss as MsM^{s} is close to 𝕋\mathbb{T}, which has entropy less than 22. To apply Corollary 1.2, it suffices to show that for the MCF starting from M0M^{0} (resp. M1M^{1}), the inward (resp. outward) torus neck will pinch. But this is given by Proposition 3.16.

5.4. Proof of Theorem 1.4

Let Σ1\Sigma_{1} be a genus one embedded shrinker in 3\mathbb{R}^{3} with the least entropy. Recall that by [CM12] index(Σ1)5\mathrm{index}(\Sigma_{1})\geq 5. Therefore, in order to prove Theorem 1.4, let us suppose by contradiction that Σ1\Sigma_{1} is compact with index at least 66.

We first need a family of initial conditions to run MCF. That will be provided by the following lemma.

Lemma 5.3.

Let Σn\Sigma^{n} by any smooth, embedded, compact, nn-dimensional shrinker in n+1\mathbb{R}^{n+1} with index at least 66. Let ϵ>0\epsilon>0 be sufficiently small. Then there exists a one-parameter family of smooth, compact, embedded surfaces {Ms(0)}s[0,1]\{M^{s}(0)\}_{s\in[0,1]} such that:

  1. (1)

    The family varies continuous in the CC^{\infty}-topology, and each Ms(0)M^{s}(0) is ϵ\epsilon-close to CC^{\infty} to Σ\Sigma.

  2. (2)

    Each Ms(0)M^{s}(0) has entropy less than that of Σ\Sigma.

  3. (3)

    M0(0),M1(0),M^{0}(0),M^{1}(0), and Σ\Sigma are all disjoint, with M0(0)M^{0}(0) inside Σ\Sigma and M1(0)M^{1}(0) outside.

Proof.

Fix an outward unit normal vector field 𝐧\bf n to Σ\Sigma. Since index(Σ)6\mathrm{index}(\Sigma)\geq 6, the eigenfunctions of its Jacobi operator, with respect to the Gaussian metric, that have negative eigenvalues include:

  • three induced by translation in 3\mathbb{R}^{3},

  • one by scaling,

  • the unique one-sided one which has the lowest eigenvalue, denoted ϕ0\phi_{0},

  • and at least one more, denoted ϕ1\phi_{1},

all of which are orthonormal under the L2L^{2}-inner product. We will choose ϕ0>0\phi_{0}>0.

Let ϵ>0\epsilon>0, and define Ms(0)M^{s}(0) to be the following perturbation of Σ\Sigma:

Ms(0):=Σ+ϵ(cos(sπ)ϕ0+sin(sπ)ϕ1)𝐧.M^{s}(0)\mathrel{\mathop{\mathchar 58\relax}}=\Sigma+\epsilon(-\cos(s\pi)\phi_{0}+\sin(s\pi)\phi_{1})\bf n.

Thus, if ϵ>0\epsilon>0 is sufficiently small, clearly the family {Ms(0)}s[0,1]\{M^{s}(0)\}_{s\in[0,1]} is smooth. Item (3) holds because ϕ0>0\phi_{0}>0. Finally, (2) holds because ϕ0,ϕ1\phi_{0},\phi_{1} are not induced by translation or scaling (see Theorem 0.15 in [CM12]). ∎

Applying the above lemma to Σ1\Sigma_{1}, we obtain a one-parameter family {Ms(0)}s[0,1]\{M^{s}(0)\}_{s\in[0,1]} of tori. Then

Ent(Ms(0))<Ent(Σ1)Ent(𝕋)<2.\mathrm{Ent}(M^{s}(0))<\mathrm{Ent}(\Sigma_{1})\leq\mathrm{Ent}(\mathbb{T})<2.

Thus, applying Corollary 1.2, and by the monotonicity formula, we obtain another embedded genus one shrinker with entropy less than Σ1\Sigma_{1}, which contradicts the definition of Σ1\Sigma_{1}.

5.5. Proof of Theorem 1.5

Since 𝕋\mathbb{T} is rotationally symmetric, by [Liu16], it has index at least 77. Again, we need a family of MCF. We will apply [CM22, Theorem 1.6] of Choi-Mantoulidis. Namely, since 𝕋\mathbb{T} is a minimal surface with index at least 66 under the Gaussian metric, it has, as we saw in the proof of Lemma 5.3, two orthonormal eigenfunctions ϕ0,ϕ1\phi_{0},\phi_{1} to the Jacobi operator that

  • have negative eigenvalues,

  • and are both orthogonal to the other 4 eigenfunctions induced by translation and scaling.

Now, pick an ϵ>0\epsilon>0. Applying [CM22, Theorem 1.6] to the 2-dimensional function space spanned by ϕ0\phi_{0} and ϕ1\phi_{1}, we obtain a one-parameter family of smooth ancient rescaled MCF (i.e. MCF under the Gaussian metric) ~s={M~s(τ)}τ0\widetilde{\mathcal{M}}^{s}=\{\widetilde{M}^{s}(\tau)\}_{\tau\leq 0}, s[0,1]s\in[0,1], such that:

  • For each ss, M~s(t)𝕋\widetilde{M}^{s}(t)\to\mathbb{T} in CC^{\infty} as tt\to-\infty.

  • M~0(0)\widetilde{M}^{0}(0) lies inside 𝕋\mathbb{T} while M~1(0)\widetilde{M}^{1}(0) lies outside.

  • {M~s(0)}s[0,1]\{\widetilde{M}^{s}(0)\}_{s\in[0,1]} is a smooth family of tori, each being ϵ\epsilon-close to 𝕋\mathbb{T} in CC^{\infty} (see [CM22, Corollary 3.4]).

If ϵ\epsilon is small enough, we can apply Theorem 1.3 to the family {M~s(0)}s[0,1]\{\widetilde{M}^{s}(0)\}_{s\in[0,1]} to obtain an s0[0,1]s_{0}\in[0,1] such that the level set flow {M(t)}t0\{M(t)\}_{t\geq 0} with initial condition M(0)=M~s(0)M(0)=\widetilde{M}^{s}(0) would develop a singularity at which every tangent flow is given by a multiplicity one, embedded, genus one self-shrinker.

Finally, we define an ancient smooth MCF {M^(t)}t1\{\hat{M}(t)\}_{t\leq-1} by rescaling the rescaled MCF {M~s0(τ)}τ0\{\widetilde{M}^{s_{0}}(\tau)\}_{\tau\leq 0}:

M^(t)=tM~(log(t)),t1.\hat{M}(t)=\sqrt{-t}\widetilde{M}(-\log(-t)),\;\;t\leq-1.

Note that M^(1)=M~(0)=M(0)\hat{M}(-1)=\widetilde{M}(0)=M(0). Hence, combining the two flows {M^(t)}t1\{\hat{M}(t)\}_{t\leq-1} and {M(t)}t0\{M(t)\}_{t\geq 0}, we obtain an ancient MCF satisfying Theorem 1.5.

5.6. Proof of Corollary 1.7

Let Σ\Sigma be an embedded shrinker with the fourth least entropy in 3\mathbb{R}^{3}, whose existence was established in §1 already. Suppose by contradiction that Σ\Sigma is rotationally symmetric. Then by Kleene-Møller [KMl14], Σ\Sigma is closed with genus one. Moreover, Σ\Sigma has entropy less than 22 since the shrinking doughnut 𝕋\mathbb{T} in [DN18] does, and by [Liu16], Σ\Sigma has index at least 77. Therefore, Theorem 1.5 still holds with 𝕋\mathbb{T} replaced by Σ\Sigma: The exact same proof will work. As a result, we obtain a genus one shrinker with entropy strictly lower than Σ\Sigma. However, the self-shrinkers with the three lowest entropy are the plane, the sphere, and the cylinder ([CIMIW13, BW17]). Contradiction arises.

References

  • [ADS19] Sigurd Angenent, Panagiota Daskalopoulos, and Natasa Sesum. Unique asymptotics of ancient convex mean curvature flow solutions. J. Differential Geom., 111(3):381–455, 2019.
  • [ADS20] Sigurd Angenent, Panagiota Daskalopoulos, and Natasa Sesum. Uniqueness of two-convex closed ancient solutions to the mean curvature flow. Ann. of Math. (2), 192(2):353–436, 2020.
  • [And12] Ben Andrews. Noncollapsing in mean-convex mean curvature flow. Geom. Topol., 16(3):1413–1418, 2012.
  • [Ang92] Sigurd B Angenent. Shrinking doughnuts. In Nonlinear diffusion equations and their equilibrium states, 3, pages 21–38. Springer, 1992.
  • [BK21] Yakov Berchenko-Kogan. The entropy of the Angenent torus is approximately 1.85122. Exp. Math., 30(4):587–594, 2021.
  • [Bra78] Kenneth A. Brakke. The Motion of a Surface by Its Mean Curvature. (MN-20). Princeton University Press, Princeton, 1978.
  • [Bre15] Simon Brendle. A sharp bound for the inscribed radius under mean curvature flow. Invent. Math., 202(1):217–237, 2015.
  • [Bre16] Simon Brendle. Embedded self-similar shrinkers of genus 0. Ann. of Math. (2), 183(2):715–728, 2016.
  • [BW17] Jacob Bernstein and Lu Wang. A topological property of asymptotically conical self-shrinkers of small entropy. Duke Math. J., 166(3):403–435, 2017.
  • [CCMS20] Otis Chodosh, Kyeongsu Choi, Christos Mantoulidis, and Felix Schulze. Mean curvature flow with generic initial data. arXiv preprint arXiv:2003.14344, 2020.
  • [CCMS21] Otis Chodosh, Kyeongsu Choi, Christos Mantoulidis, and Felix Schulze. Mean curvature flow with generic low-entropy initial data. arXiv preprint arXiv:2102.11978, 2021.
  • [CCS23] Otis Chodosh, Kyeongsu Choi, and Felix Schulze. Mean curvature flow with generic initial data ii. arXiv preprint arXiv:2302.08409, 2023.
  • [CGG91] Yun Gang Chen, Yoshikazu Giga, and Shun’ichi Goto. Uniqueness and existence of viscosity solutions of generalized mean curvature flow equations. Journal of differential geometry, 33(3):749–786, 1991.
  • [CHH22] Kyeongsu Choi, Robert Haslhofer, and Or Hershkovits. Ancient low-entropy flows, mean-convex neighborhoods, and uniqueness. Acta Math., 228(2):217–301, 2022.
  • [CHHW22] Kyeongsu Choi, Robert Haslhofer, Or Hershkovits, and Brian White. Ancient asymptotically cylindrical flows and applications. Invent. Math., 229(1):139–241, 2022.
  • [CIM15] Tobias Holck Colding, Tom Ilmanen, and William P Minicozzi. Rigidity of generic singularities of mean curvature flow. Publications mathématiques de l’IHÉS, 121(1):363–382, 2015.
  • [CIMIW13] Tobias Holck Colding, Tom Ilmanen, William P Minicozzi II, and Brian White. The round sphere minimizes entropy among closed self-shrinkers. Journal of Differential Geometry, 95(1):53–69, 2013.
  • [CM12] Tobias H Colding and William P Minicozzi. Generic mean curvature flow i; generic singularities. Annals of mathematics, pages 755–833, 2012.
  • [CM15] Tobias Holck Colding and William P. Minicozzi, II. Uniqueness of blowups and łojasiewicz inequalities. Ann. of Math. (2), 182(1):221–285, 2015.
  • [CM16] Tobias Holck Colding and William P Minicozzi. The singular set of mean curvature flow with generic singularities. Inventiones mathematicae, 204(2):443–471, 2016.
  • [CM22] Kyeongsu Choi and Christos Mantoulidis. Ancient gradient flows of elliptic functionals and morse index. American Journal of Mathematics, 144(2):541–573, 2022.
  • [CS21] Otis Chodosh and Felix Schulze. Uniqueness of asymptotically conical tangent flows. Duke Math. J., 170(16):3601–3657, 2021.
  • [DN18] Gregory Drugan and Xuan Hien Nguyen. Shrinking doughnuts via variational methods. The Journal of Geometric Analysis, 28:3725–3746, 2018.
  • [ES91] Lawrence C Evans and Joel Spruck. Motion of level sets by mean curvature. i. Journal of Differential Geometry, 33(3):635–681, 1991.
  • [HK17] Robert Haslhofer and Bruce Kleiner. Mean curvature flow of mean convex hypersurfaces. Communications on Pure and Applied Mathematics, 70(3):511–546, 2017.
  • [HS99a] Gerhard Huisken and Carlo Sinestrari. Convexity estimates for mean curvature flow and singularities of mean convex surfaces. Acta mathematica, 183(1):45–70, 1999.
  • [HS99b] Gerhard Huisken and Carlo Sinestrari. Mean curvature flow singularities for mean convex surfaces. Calculus of Variations and Partial Differential Equations, 8(1):1–14, 1999.
  • [Hui84] Gerhard Huisken. Flow by mean curvature of convex surfaces into spheres. J. Differential Geom., 20(1):237–266, 1984.
  • [Hui90] Gerhard Huisken. Asymptotic behavior for singularities of the mean curvature flow. Journal of Differential Geometry, 31(1):285–299, 1990.
  • [HW20] Or Hershkovits and Brian White. Nonfattening of mean curvature flow at singularities of mean convex type. Communications on Pure and Applied Mathematics, 73(3):558–580, 2020.
  • [Ilm92] Tom Ilmanen. Generalized flow of sets by mean curvature on a manifold. Indiana University mathematics journal, pages 671–705, 1992.
  • [Ilm94] Tom Ilmanen. Elliptic regularization and partial regularity for motion by mean curvature, volume 520. American Mathematical Soc., 1994.
  • [Ilm95] Tom Ilmanen. Singularities of mean curvature flow of surfaces. preprint, 1995.
  • [INS19] Tom Ilmanen, André Neves, and Felix Schulze. On short time existence for the planar network flow. J. Differential Geom., 111(1):39–89, 2019.
  • [KKM18] Nikolaos Kapouleas, Stephen James Kleene, and Niels Martin Møller. Mean curvature self-shrinkers of high genus: non-compact examples. Journal für die reine und angewandte Mathematik (Crelles Journal), 2018(739):1–39, 2018.
  • [KMl14] Stephen Kleene and Niels Martin Mø ller. Self-shrinkers with a rotational symmetry. Trans. Amer. Math. Soc., 366(8):3943–3963, 2014.
  • [Lee21] Tang-Kai Lee. Compactness and rigidity of self-shrinking surfaces. arXiv preprint arXiv:2108.03919, 2021.
  • [Liu16] Zihan Hans Liu. The index of shrinkers of the mean curvature flow. arXiv preprint arXiv:1603.06539, 2016.
  • [LS22] Zhengjiang Lin and Ao Sun. Bifurcation of perturbations of non-generic closed self-shrinkers. Journal of Topology and Analysis, 14(04):979–999, 2022.
  • [MF10] John W Morgan and Frederick Tsz-Ho Fong. Ricci flow and geometrization of 3-manifolds, volume 53. American Mathematical Soc., 2010.
  • [Møl11] Niels Martin Møller. Closed self-shrinking surfaces in 3\mathbb{R}^{3} via the torus. arXiv preprint arXiv:1111.7318, 2011.
  • [Ngu14] Xuan Hien Nguyen. Construction of complete embedded self-similar surfaces under mean curvature flow, part iii. Duke Mathematical Journal, 163(11):2023–2056, 2014.
  • [OS88] Stanley Osher and James A. Sethian. Fronts propagating with curvature-dependent speed: algorithms based on Hamilton-Jacobi formulations. J. Comput. Phys., 79(1):12–49, 1988.
  • [Sun23] Ao Sun. Local entropy and generic multiplicity one singularities of mean curvature flow of surfaces. J. Differential Geom., 124(1):169–198, 2023.
  • [SW09] Weimin Sheng and Xu-Jia Wang. Singularity profile in the mean curvature flow. Methods Appl. Anal., 16(2):139–155, 2009.
  • [SW20] Ao Sun and Zhichao Wang. Compactness of self-shrinkers in r3 with fixed genus. Advances in Mathematics, 367:107110, 2020.
  • [SWZ20] Ao Sun, Zhichao Wang, and Xin Zhou. Multiplicity one for min-max theory in compact manifolds with boundary and its applications. arXiv preprint arXiv:2011.04136, 2020.
  • [SX21a] Ao Sun and Jinxin Xue. Initial perturbation of the mean curvature flow for asymptotical conical limit shrinker. arXiv preprint arXiv:2107.05066, 2021.
  • [SX21b] Ao Sun and Jinxin Xue. Initial perturbation of the mean curvature flow for closed limit shrinker. arXiv preprint arXiv:2104.03101, 2021.
  • [Wan11] Xu-Jia Wang. Convex solutions to the mean curvature flow. Annals of mathematics, pages 1185–1239, 2011.
  • [Whi95] Brian White. The topology of hypersurfaces moving by mean curvature. Communications in analysis and geometry, 3(2):317–333, 1995.
  • [Whi97] Brian White. Stratification of minimal surfaces, mean curvature flows, and harmonic maps. Journal für die reine und angewandte Mathematik, 488:1–36, 1997.
  • [Whi00] Brian White. The size of the singular set in mean curvature flow of mean-convex sets. Journal of the American Mathematical Society, 13(3):665–695, 2000.
  • [Whi03] Brian White. The nature of singularities in mean curvature flow of mean-convex sets. Journal of the American Mathematical Society, 16(1):123–138, 2003.
  • [Whi05] Brian White. A local regularity theorem for mean curvature flow. Annals of mathematics, pages 1487–1519, 2005.
  • [Whi09] Brian White. Currents and flat chains associated to varifolds, with an application to mean curvature flow. Duke Mathematical Journal, 148(1):41–62, 2009.