AN UNKNOTTEDNESS RESULT FOR SELF SHRINKERS WITH MULTIPLE ENDS

ALEXANDER MRAMOR

Abstract. In this article we prove a new unknottedness result for self shrinkers in R3 with multiple asymptotically conical ends satisfying some apriori topological assumptions, obtaining a stronger result when the number of ends is 2.

1. Introduction In this article we prove the following isotopy rigidity fact about self shrinkers with multiple asymptotically conical ends, self shrinkers being fundamental singularity models for the mean curvature flow: Theorem 1.1. Let M ⊂ R3 be a 2-sided self shrinker with k ≥ 2 asymptotically conical ends. Supposing for a choice of normal that M bounds a region R diffoemor- phic to the interior of a compact closed surface Σ with k solid half cylinders attached and that Σ is itself topologically standard, then M is ambiently isotopic to a stan- dardly embedded genus g surface with k straight round half cylinders attached along its convex hull. We say a closed genus g surface Σ is topologically standard when it is ambiently isotopic to a surface given by k standardly embedded tori arranged along a curve with adjaent ones glued together by unknotted bridges. We’ll also often refer to surfaces M satisfying the conclusion of the theorem as standardly embedded. Obviously that Σ is topologically standard is a nontrivial assumption. Which might be less immediately clear, as discussed in more detail below that the ends bound solid cylinders is a nontrivial assumption when k > 2 as well. For k > 2 its uncertain that the set of such shrinkers is nonempty, but the bounded solid cylinder part of the assumption is always met (see lemma 2.4) for shrinkers with 2 ends, and arXiv:2011.09373v4 [math.DG] 29 Nov 2020 there is numerical evidence of many shrinkers with 2 ends besides the cylinder so the set in the following result is almost certainly nontrivial: Corollary 1.2. Let M ⊂ R3 be a 2-sided complete self shrinker with two asymp- totically conical ends, and suppose that capping M off gives a standardly embedded genus g surface. Then M is ambiently isotopic to a standardly embedded genus g surface with two straight round half cylinders attached along its convex hull. 1 MULTIPLE ENDS 2

Naturally here and throughout “capping” a noncompact surface whose ends bound solid cylinders refers to cutting these cylinders and gluing in discs in the obvious way. That the ends are asymptotically conical is a very natural assumption with L. Wang’s work and Ilmanen’s no cylinder conjecture in mind – these are described more in section 2. In particular (since cylinders are unknotted) our result implies all self shrinkers with finite topology and two ends that are given by adding ends to a topologically standard surface are unknotted in the conjectural picture. Previously analogous versions of this result were established in the compact case by the author with S. Wang in [40] and in the noncompact case with one asymptotically conical end by the author in the preprint [39] (the results in these cases don’t need any other apriori control on the topology along the lines of the statements above). In a result allowing even for multiple ends, Brendle showed in [4] that noncompact self shrinkers in R3 with no topology (roughly speaking) must coincide with either the plane or round cylinder, but the result in this work is an improvement in some ways in that it allows for nontrivial topology. There are rigourously constructed examples of noncompact self shrinkers with one end besides the plane via desingularizations of the sphere and plane by Kapouleas, Kleene, and Møller [31]. To the author’s knowledge though there are no rigorous constructions of self shrinkers with more than one end besides the cylinder: the ex- amples constructed in [26] might possibly be noncompact and hence could furnish examples of shrinkers with more than one end, and the very recent work of A. Sun, Z. Wang and Zhou [42] gives many examples of self shrinkers (even of arbitrarily high entropy), which from their techniques it seems plausible at least some of which would be noncompact. On the other hand as already mentioned there is good numer- ical evidence of self shrinkers with two asymptotically conical ends with interesting topology by Chopp [7] and also Ilmanen [29]. Of course, its clear upon inspection these numerical examples are unknotted in the sense of the theorem above.

There are related unknottedness results for minimal surfaces in R3, the most relevent one being the result of Meeks and Yau [37] for minimal surfaces of finite topology. However, there additional topological facts of classical minimal surfaces are used, most apparently to the author the topologically parallel ends theorem which says that a classical minimal surface with k ends and finite topology is isotopic to a surface asymptotic to k parallel planes – this is intuitively true because the blow- down of a minimal surface (with some growth assumptions) yields a minimal cone (possibly with multiplicity) which in the surface case must be a plane. To the au- thor’s knowledge though the situation for shrinkers is unclear and there seems to be partial evidence both for and against the ends of a self shrinker to be parallel though, although in the case where apriori the ends are parallel it seems plausible that their MULTIPLE ENDS 3 arguments can be applied (resolving this apparent uncertainty one way or the other seems to be a natural question for future work). This and more is discussed in the concluding remarks. At any rate our argument is significantly different from theirs, even when k = 2 (when k = 2, the ends are automatically parallel so an adapatation (if possible) of their work would apply to give corollary 1.2, in fact without assuming apriori that the ends are attached to a topologically standard surface). One facet of our argument that might be of interest is that we do not employ Waldhausen’s theorem in our scheme (note the argument of Meeks and Yau needed some results to the effect saying certain parts of their minimal surfaces are unknotted without appealing to the theorem as well). Granted by assumption the compact “soul” Σ of M is toplogically standard (which is the type of information one would expect to obtain from Waldhausen), but as figure 1 below shows the problem is far from solved with just this information in short because topological standardness of a collection of surfaces does not imply their combinations will be. We feel the general shape of the argument could possibly be built on to give stronger unknottedness reults while still bypassing the theorem (to the author’s knowledge, this general type of question originates with Yau), although there are complications as discussed more in the concluding remarks. To deal with situations like that posed in figure 1, recall that an unstable minimal surface at least in principle can be perturbed to be mean convex in their metric using the first eigenfunction of their Jacobi operator – for shrinkers with asymptotically conical ends this was established by Bernstein and L. Wang in [2] (indeed they are unstable). The mean convexity then lets one construct a (shrinker) mean convex level set flow starting from it using a limit of surgery flows as carried out in the author’s previous article [39] – this will alow us to show in many undesirable cases the flow must “snag” on itself so that one finds a stable self shrinker giving a contradiction. Using this principle we give two arguments, one particular to the cases k = 1, 2 (although since the case k = 1 is already covered by the stronger result in the author’s previous work [39] we ignore it as we did in the theorem statement) and another which applies to any number of ends inspired by an observation of Bernstein and L. Wang. In the first argument, particular to when k = 2, the renormalized mean curvature flow is used to reduce the problem to showing there are no outwardly shrinker convex knotted cylinders coming for which our construction of the flow applies: to get this fact one sees for such a knotted cylinder one would find a nonempty stable shrinker by the flow and gain a subsequent contradiction as described. To make this reduction its easy to see handles of Σ must break under the flow, and they must not link with the ends of M, so that eventually the flow of M must be an embedded cylinder which MULTIPLE ENDS 4 is isotopic to M after adding some handles to the cylinder. Hence if this cylinder is unknotted, then M must be topologically standard. Using ideas from the first argument, in the second argument one analogously reduces to the question of showing an outwardly mean convex self shrinker diffeomo- prhic to a sphere connect sum k half cylinders is standardly embedded. The simple but key observation, more or less gleaned from the aforementioned paper of Bern- stein and L. Wang [2] on conical shrinkers of low entropy (see theorem 4.6 therein) with some slight twists, is that the flow of such a surface preserves must eventually become starshaped (modulo in our setting a possible singular set which can be un- derstood using stratifcation results for the mean curvature flow due to Colding and Minicozzi [11]), finishing the proof of theorem 1.1. Acknowledgments: The author is supported by an AMS–Simons travel grant and is grateful for their assistance. He also thanks Jacob Bernstein for discussing the state of the art in numerical examples of self shrinkers with him.

2. Preliminaries Let X : M → N n+1 be an embedding of M realizing it as a smooth closed hy- persurface of N, which by abuse of notation we also refer to as M. Then the mean n+1 curvature flow Mt of M is given by (the image of) X : M × [0,T ) → N satisfying the following, where ν is the outward normal: dX = H~ = −Hν, X(M, 0) = X(M) (2.1) dt

By the comprison principle singularties occur often and makes their study impor- tant. To study these singularities, one may parabolically rescale about the developing high curvature region to obtain an ancient flow defined for times (−∞,T ]; when the base point is fixed this is called a tangent flow blowup which, as described by Ilmanen in his preprint [27] for flows of 2 dimensional surfaces, will be modeled on smooth self shrinkers: these are surfaces satisfying the following equivalent definitions: n n+1 X⊥ (1) M ⊂ R which satisfy H − 2 = 0, where X is the position vector −|x|2 (2) minimal surfaces in the Gaussian metric Gij = e 2n δij (3) surfaces M which√ give rise to ancient flows Mt that move by dilations by setting Mt = −tM Of course, as the degenerate neckpinch of Angenent and Velasquez [1] illustrates tangent flows do not capture quite all the information about a developing singularity but they are a natural starting point. The Gaussian metric is a poorly behaved MULTIPLE ENDS 5 metric in many regards; it is incomplete and by the calculations in [10] its scalar curvature at a point x is given by:

2   |x| n − 1 2 R = e 2n n + 1 − |x| (2.2) 4n We see that as |x| → ∞ the scalar curvature diverges, so there is no way to complete the metric. Also since R is positive for |x| small and negative for large |x|, there is no sign on sectional or Ricci curvatures. On the other hand it is f-Ricci positive, in 1 2 the sense of Bakry and Emery with f = − 2n |x| , suggesting it should satisfy many of the same properties of true Ricci positive metrics (see [46]). Indeed, this provides some idea as to why one might expect an unknottedness result for self shrinkers, because analogous unknottedness results hold in Ricci positive metrics on S3. As is well known, the second variation for formula for area shows there are no stable minimal surfaces in Ricci positive manifolds, see for instance chapter 1 of [9]. Crucial for our arguments in the proof of the theorem is that this turns out to also be true for minimal surfaces of polynomial volume growth in Rn endowed with the Gaussian metric as discussed in [10]. To see why this is so, the Jacobi operator for the Gaussian metric is given by: 1 1 L = ∆ + |A|2 − hX, ∇(·)i + (2.3) 2 2 1 The extra 2 term is essentially the reason such self shrinkers unstable in the Gaussian metric: for example owing to the constant term its clear in the compact case from this that one could simply plug in the function “1” to get a variation with Lu > 0 which doesn’t change sign implying the first eigenvalue is negative. In fact, every properly embedded shrinker has polynomial volume growth by Q. Ding and Y.L. Xin: Theorem 2.1 (Theorem 1.1 of [14]). Any complete non-compact properly immersed self-shrinker M n in Rn+m has Euclidean volume growth at most. We will combine these facts below to conclude the self shrinker we could find in some cases in the argument below must in fact be unstable. Now we discuss some terminology describing possible behavior of the ends of self shrinkers. 3 A regular cone in R is a surface of the form Cγ = {rγ}r∈(0,∞) where γ is smooth simple closed curve in S2. An end of a surface M 2 ,→ R3 is asymptotically conical 2 with asymptotic cross section γ if ρM → Cγ in the Cloc sense of graphs as ρ & 0 restricted to that end. Similarly we define asymptotically cylindrical ends to be ends which are asymptot- ically graphs over cylinders (with some precsribed axis and diameter) which converge 2 to that cylinder in Cloc on that end. MULTIPLE ENDS 6

The reason we focus on such ends is the following important result of L. Wang, which says that these are the only possible types of ends which may arise in the case of finite topology: Theorem 2.2 (Theorem 1.1 of [44]). If M is an end of a noncompact self-shrinker in R3 of finite topology, then either of the following holds: −1 ∞ 3 3 (1) limτ→∞ τ M = C(M) in Cloc(R \ 0) for C(M) a regular cone in R −1 1 ∞ 3 3 (2) limτ→∞ τ (M − τv(M)) = Rv(M) × S in Cloc(R ) for a v(M) ∈ R \{0} In particular, theorem 2.2 applies to self shrinkers which arises as the tangent flow to compact mean curvature flows. Ilmanen in [28] conjectured that there would be no cylidrical ends unless the self shrinker was itself a cylinder: this is a natural working assumption from an analytical perspective because analysis on cones is in some sense easier than on cylinders. Later, L. Wang [45] gave evidence that it was true by verifying it for shrinkers very quickly asymptotic to cylinders. Note that since ∞ the convergence is in Cloc, higher multiplicity convergence is ruled out and hence the number of links of the cone must be the same as the number of ends. The mean curvature flow is best understood in the mean convex case, especially so for 2-convex surfaces (λ1 +λ2 > 0) and a surgery theory with this convexity condition similar to the Ricci flow with surgery. For the mean curvature flow with surgery one finds for a mean convex surface M (in higher dimensions, 2-convex) curvature scales Hth < Hneck < Htrig so that when H = Htrig at some point p and time t, the flow is stopped and suitable points where H ∼ Hneck are found to do surgery where “necks” (at these points the surface will be approximately cylindrical) are cut and caps are glued in. The high curvature regions are then topologically identified as Sn or Sn−1 × S1 and discarded and the low curvature regions will have curvature bounded on the order of Hth. The flow is then restarted and the process repeated. It was initially eastablished for compact 2-convex hypersurfaces in Rn+1 where n ≥ 3 by Huisken and Sinestrari in [25], and their approach was later extended to the case n = 2 by Brendle and Huisken in [6]. A somewhat different approach covering all dimensions simultaneously was given later by Haslhofer and Kleiner in [19] shortly afterwards. Haslhofer and Ketover then showed several years later in section 8 of their paper [20] enroute to proving their main result that the mean curvature flow with surgery can be applied to compact mean convex hypersurfaces in general ambient manifolds. An important advantage of the mean curvature flow with surgery is that the topo- logical change across discontinous times, when necks are cut and high curvature regions discarded, is easy to understand. A disadvantage is that it isn’t quite a Brakke flow (a geometric measure theory formulation of the mean curvature flow) so does not immediately inherit some of the consequences thereof, but at least it is MULTIPLE ENDS 7 closely related to the level set flow by results of Laurer [32] and Head [21, 22] which is at least in their common domain of defintion. In their work they show that surgery converges to the level set flow in Hausdorff distance (and in fact in varifold sense as Head shows) as the surgery parameters degenerate (i.e. as one lets Hth → ∞). This connection is useful for us because deep results of White [47] show that a mean convex LSF will converge to a (possibly empty) stable minimal surface long term – this will be used in conjunction with the instability results mentioned above to garner a contradiction. As mentioned above mean curvature flow with surgery in the round metric has been already accomplished by Haslhofer and Ketover but some extra care is needed for the Gaussian metric especially in the noncompact case. This is because the metric is poorly behaved at infinity (as one sees from the calculation of its scalar curvature) which introduces some analytic difficulites for using the flow, so instead we consider the renormalized mean curvature flow (which we’ll abreviate RMCF) defined by the following equation:

dX X = H~ + (2.4) dt 2 Where here as before X is the position vector on M. It is related to the regular mean curvature flow by the following reparameterization; this allows one to transfer many deep theorems on the MCF to the RMCF. Supposing that Mt is a mean curvature flow on [−1,T ), −1 < T ≤ 0 (T = 0 is the case for a self shrinker). Then ˆ the renormalized flow Mτ of Mt defined on [0, − log(−T )) is given by ˆ τ/2 Xτ = e X−e−τ , τ = − log (−t) (2.5) (Note up to any finite time the reparameterization is bounded and preserves many properties of the regular MCF, like the avoidance principle). This is a natural flow for us to consider because it is up to a multiplicative term the gradient flow of the Gaussian area and fixed points with respect to it are precisely self shrinkers. More precisely, writing HG for the mean curvature of a surface with respect to the Gaussian metric: ⊥ |x|2 X H = e 4 (H − ) (2.6) G 2

With this in mind, the author showed in his previous article [39] that one can then construct a flow with surgery using the RMCF on suitable perturbations of noncompact self shrinkers, and that as one let’s the surgery parameters degenerate indeed the surgery converges to the level set flow: MULTIPLE ENDS 8

3 X⊥ Theorem 2.3. Let M ⊂ R be an asymptotically conical surface such that H − 2 ≥ c(1 + |X|2)−α for some c, α > 0, and so that as R → ∞ |A(p)|2 → 0 for any c p ∈ M ∩ B(p, R) . Then there is a level set flow Lt out of M with respect to the renormalized mean curvature flow which is

(1) inward in that Lt1 ⊂ Lt2 for any t1 > t2. k (2) the Hausdorff limit of surgery flows St with initial data M. In [39] the author then uses this result to show asymptotically conical self shrinkers with one end by finding a stable self shrinker if the original surface was not a Heegaard spliting (in the appropriate sense) using a covering space characterization of Heegaard surfaces observed by Lawson, giving a contradiction. The order of decay assumption on shrinker mean convexity in the statement above is to ensure the flow indeed stays shrinker mean convex, using section 3 of Bernstein and L. Wang [2]. While in [39] it was only used to study single ended self shrinkers, theorem 2.3 can be applied to self shrinkers with multiple asymptotically conical ends as well since the convergence to the asymptotic cone is with multiplicity 1. We end this preliminaries section with the following easy but important topological lemma for two sided surfaces with two ends; in the following note the distinction between diffeomorphic and isotopic: Lemma 2.4. Suppose M is a 2-sided asymptotically conical self shrinker with two ends. Then it bounds a region R with which is diffeomorphic to the interior of a region bounded by a closed compact surface Σ with two solid half cylinders attached.

Proof: Writing the ends of M as E1 and E2, it suffices to show that there is a choice of unit normal on M such that the links of E1 and E2 bound disjoint discs (using as outward normal the normal inherited from M), because then a cross section of each end will bound a disk by the asymptotically conical assumption which will give the decomposition. To see that this is so, consider the associated components of the link, L1 and L2 (there are two different links by the work of L. Wang as already mentioned). By the two sided assumption we may color the sphere with two colors which alternates across the link of M. Since the link of M when there are 2 ends splits the sphere into 3 components the links must bound disjoint discs of the same color, implying the conclusion. 

3. Proof of theorem 1.1 Although we will ultimately not use this fact, arguing as in [39] by perturbing M inwards and outwards and running the flow one sees that M must be a Heegaard spliting, in that in a very large ball it bounds regions which are both handlebodies. MULTIPLE ENDS 9

Equivalently as observed in [33] it’s a 2-sided surface for which the inclusion map into the regions it bounds induces a surjection on fundamental groups. At this step one would wish to invoke a Waldhausen type theorem, the most well known one being: Theorem 3.1. ([43]) Any two Heegaard splitings of genus g of S3 are ambiently isotopic and, in particular, are standardly embedded. Waldhausen’s theorem also applies to Heegaard splitings of the ball with one boundary component, as utilized in [39], and the same statement also is true for those with two boundary components which are diffeomorphic to annuli (see corollary at the bottom of page 408 in [36]). Unfortunately though, generally speaking Wald- hausen’s theorem doesn’t apply immediately when there is more than one boundary component as a consequence of the knotted minimal examples P. Hall [18], where he produces “knotted” minimal surfaces (which are Heegaard splitings) in the ball with two boundary components and genus 1, showing that Meeks’ claim is sharp. An upshot of this discussion as observed in [37] is that one should expect an unknottedness theorem for noncompact complete minimal surfaces (in their case classical minimal surfaces) with multiple ends and nontrivial topology to employ the completeness assumption. The proof we give below does because, for instance, the surgery flow we use is for surfaces without boundary and, because our surfaces are without boundary, we may employ the maximum principle without worrying about behavior along a boundary. Of course, we also make some admittedly strong apriori assumptions in the arrangement of the surfaces (that the ends bound cylinders and are attached to a surface which is already known to be topologically standard) which will be used quickly below. In particular we point out that the assumption of Σ being topologically standard does indeed preclude self shrinkers analogous to that of Hall’s examples. Despite Hall’s examples it is tempting to try to at least extract some amount of information from Waldhausen’s theorem by compactifying the ends with caps to get a closed surface under the pretense such capping might “destroy” the knottedness, because the result itself one might reasonably expect to be a Heegaard splitting of S3 (after additionally compactifying R3 at infinity as in [40]) so would be topologically standard by Waldhausen’s theorem. If this were true, this would give that we may write our surface as a connect sum of Σ, a standardly embedded genus g surface, with ends E1,...EK solid half cylinders attached (although perhaps in a complicated way), in particular not assuming firstoff that Σ is standardly embedded surface. Unfortunately though, capping off preserving the quality of being a Heegaard split- ing even in our simple case as above seems to be false in general. For instance, capping the boundaries in Hall’s example (specifically, see figure 4 in [18]) and extending the interior of the ball to all of R3 gives a trefoil knot. In our context, this implies there MULTIPLE ENDS 10 could be a self shrinker with two ends attached along a trefoil knot (without further observations to rule this out). This explains the unfortunate hypothesis that Σ in the theorem statement is assumed to be topologically standard. Perhaps, if the capping could be done in a shrinker mean convex way (which is plausible for complete self shrinkers because one might be able to take advantage of the hx, νi term) one would be able to sidestep this issue. At any rate even assuming that Σ is topologically standard actually doesn’t yield much information about M on the whole, because there can be synergy of sorts in terms of isotopy class between each of the components when glued together. As an extreme example since the ends of M are asymptotically conical we can always take the Ei to be isotopic to standardly embedded round cylinders, “hiding” much of the knottedness in Σ, while Σ could itself be standardly embedded because one could push along where an end would be connected as the figure below shows:

Figure 1. This figure shows an asymptotically conical knotted cylin- der M can be decomposed into three regions, two ends colored blue and a tubular neighborhood of a “knotted interval” diffeomorphic to a sphere colored red, which are each individually standardly embedded but combined produce a surface which isn’t. Of course the blue dots represent where the ends would attach to the sphere.

From this we see more careful arguing is needed. As an aside (since we do know from above M is a Heegaard splitting) one idea and perhaps the most intuitive one might be to attach the ends to a large ball “at infinity” and appeal to Waldhausen’s theorem for splittings of S3, although because of the suprisingly delicate nature of Heegaard splittings under such manipulations we take a different tack using the flow more strongly. Below we discuss two arugments. The first argument, as sketched in MULTIPLE ENDS 11 the introduction, is specifc to 2 ends and is superceded by our second argument, but elements of its proof will be used in the proof of the general case and the argument might of independent interest.

3.1. An argument specific to 2 ends. In our first arugment, specific to the case of 2 ends, we will start by arguing the ends are unknotted in an appropriate sense and. As a consequence, one sees it would be helpful if the knottedness is shared “more” with the ends, in other words reducing the topological synergy between Σ and E1,E2. To do this at least in the case encapsulated in figure 1 (which we will reduce down to later in the course of the proof) we see that if the boundaries of the ends in the diagram were taken to be the same (eliminating Σ completely), then the knottedness of M would be detected by the Ei in that there wouldn’t be an isotopy of M which takes then both simultaneuously to straight round half cylinders. In fact, by taking the common boundary sufficiently far to the left or right, we would only have to contend with one of the ends being “knotted” although in intermediate steps it’s helpful to make the distinction. We’ll return to this point below.

In the following we will consider the flow of M “restricted” to the Ei, so slightly more precision is needed. We write Ei as the diffeomorphic image of the standard half cylinders on M, and by “flow of the end E” we mean the spacetime track of the points Ei under the flow of M be it a/the flow with surgery or level set flow (we will specify which we refer to in the course of the argument). As a first order approximation to the problem of incompatibility of the isotopies of the Ei to standard cylinders in relation to the other regions of M (in particular refering to the discussion above if the boundary curves of the Ei aren’t taken very far away from the origin), it is possible that (there is no claim this list necessarily exhausts all errant possibilities, but ruling these out will suffice in our proof):

(1) some of the Ei link nontrivially with Σ, in that the principal axis of an Ei links with a homotopically nontrivial curve on Σ. (2) some of the Ei are self-knotted in that there is no isotopy of M which brings an individual Ei to a round half cylinder. (3) some the Ei are knotted amongst each other, in that for some Ei, Ej there is no isotopy of M fixing Σ for which Ei and Ej are isotopic to round parallel half cylinders. Strictly speaking in the proof we may reduce to only considering cases (1) and (2) and below we will actually be interested in just a cylinder but we thought the reader might find it independently interesting to consider all three cases; in fact in the sequel, we will actually only use case (1). Indeed, the argument we will give in fact MULTIPLE ENDS 12 works for shrinkers with any number of conical ends which bound solid cylinders (which is automatically true for 2 ends by lemma 2.4) and will be used later in the argument for the case of more ends. To rule these cases out, we consider the flow of M, perturbing our asymptoti- cally conical shrinker M outward i.e expanding out of the region R defined in the statement of the theorem. Roughly speaking this is the right choice because ends can’t be broken off so information about their knottedness won’t be destroyed: such information would be destroyed if this were possible essentially from observing that a capped, asymptotically straight cylinder is always isotopic to a capped standard cylinder by “pushing” along the cap far enough. Relabeling this perturbed surface M (it is isotopic to the original M), let’s first discuss ruling out case (1). Suppose on the contrary that an end E linked nontrvially with Σ. Writing Σ as the connect sum of k tori, E passes through the central axis A of at least one of them preventing a neckpinch from occuring along A.

Figure 2. Here is a diagram representing a nontrivial linking when Σ is a torus and there is just one end E1. The comparison principle keeps the “hole” of Σ from collapsing, forcing some points of the level set flow to always stay near the initial data by the mean convexity of the flow. In this diagram, the boundary of E1 is the blue circle and the red arrows indicate roughly the direction of the flow.

Since the flow is “outward” though, this forces the level set flow Mt of M to contain points intersecting the convex hull of (the region diffeomorphic to) Σ, forcing a nonempty asymptotic limit because the flow still must stay embedded (note since the flow remains asymptotic to the cone for any given time, separate sheets of the flow can only approach each other at bounded points so the maximum principle can be applied). This limit must be a stable self shrinker, giving a contradiction. MULTIPLE ENDS 13

The central observation we will employ to deal with case (2) is the following:

Lemma 3.2. Let Mt be either a mean curvature flow with surgery or the level set flow of M. Then for each end E, Et, the flow restricted to E in the sense described above, will consist of precisely one noncompact component with nonempty boundary isotopic to E and possibly a number of compact components each diffeomorphic to S2.

Proof: We will discuss this for just surgery flows; passing to a limit in surgery pa- rameters gives the level set flow version. This clearly holds during times the flow is smooth so consider the first surgery time if any. Since the end is diffeomorphic to a cylinder and since the perturbation is outwards (so one side of disconnecting surgery can’t be a boundaryless noncompact piece), if cutting along a neck increases the number of connected components one of the components must be compact and spherical. If each surgery neck is of this type then the noncompact part must be isotopic to the E and we are done. Suppose then that there is a neck which doesn’t disconnect. Then this implies there is a handle attached to E, which can’t be since E is diffeomorphic to a half cylinder. 

With this in mind we now discuss ruling out case (2): consider an end E and sup- pose its is knotted. Choosing a parameterization of the solid half cylinder bounded by E consider the corresponding central curve γ. Since E is knotted there must be, considering a projection to a generic plane P , points ci ∈ P over which γ self crosses. Corresponding to these self crossings we may locally parameterize over these ci the level set flow of the end E for some short forward time δi as two disjoint continu- ously embedded cylinders Ci,1(t),Ci,2(t) (indexing the crossings by i) in coordinates (x, θ) ∈ [−ri,j(t), ri,j(t)]×[0, 2π). The important point being since the flow is outward from R for all i the sets Ci,1(t) and Ci,2(t) approach each other. MULTIPLE ENDS 14

Figure 3. When an end E is self knotted it must cross over itself essentially as the diagram suggests, and there one can find many points p, q whose normals point approximately towards each other and so, since the flow is outward, move closer.

The catch is that δi could be finite for some (or potentially all) i: in this scenario as t → δi, ri,j(t) → 0 (for both i and j – if one is greater than 0 is is the other) and so that when t > δi there is no crossing over the point ci. For example, the cylinder could locally untwist along the flow as in the first Reidmeister move. But by lemma 3.2 and since the flow is monotone E(t) crosses over some of the ci for all time (i.e. so for some i, δi = ∞) which implies like in case (1) that the level set flow Mt of M is nonempty as t → ∞, so that it converges to a stable self shrinker gaining a contradiction.

To discuss case (3) we introduce some more notation. Parameterizing the Ei as cylinders with parameters di denoting position along the principal axis, consder for two ends say E1, E2 a curve c(s) on M connnecting a point on d1 = 0 with a point on d2 = 0. Parameterizing the mean curvature flow on the surface so it has no tangential components, we will consder the images c(t) = c(s, t) of c under the flow. Note since the flow is outward the level set flow of M will remain connected under the flow so the curve c is defined for all times. The notation is explained in the diagram below: MULTIPLE ENDS 15

Figure 4. A schematic explaining the notation of this section. Here the ends E1 and E2 are drawn to be “knotted” together; the curve c(t) would have to be “snagged” forcing a nonempty asymptotic limit.

If min |c(s, t)| → ∞ as t → ∞, then E1 and E2 span a strip (given by the path s∈[0,1] of c(s, t) in t) so are unknotted as claimed. If not, then as t → ∞ the level set flow of M converges to what must be a nonempty stable self shrinker like the first two cases, giving a contradiction. Now we discuss how to complete the proof of the main theorem when the number of ends is 2. The idea is to reduce to the picture encapsulated in the knotted cylinder figure above, because there it is clear one could take the boundary of E1 to be the boundary of E2 so, having ruled out cases (2) and (3) we would be done. In fact, moving their common boundary component very far onto the asymptotically conical portion of either end, only ruling out case (2) suffices. The obvious obstruction is that Σ need not be a sphere (and of course if this were so we would at this point already fall into the jursidiction of Brendle’s theorem). However after some finite amount of time T ,ΣT must eventually be diffeomorphic to a sphere (considering the flow of Σ in the same way we considered the flow of the ends) or else we would find a nonempty stable limit using the argument we used to rule out case (1) above. Since Σ was already known to be standardly embedded it is isotopic to ΣT after attaching g handles in a standard way, where g is the genus of Σ. Here by attaching a handle in a standard way we mean to an unknotted cylinder along two discs of M so that its convex hull is disjoint from other handles.

In fact, M is isotopic to MT with g handles attached. To see this in a very explicit way, first we describe an isotopy of M to a presentation Mf of M with Σ “standard” as follows. Up to isotopy in a large ball B(0,D) containing the isotopy path of Σ MULTIPLE ENDS 16 the ends E1 and E2 can be taken to be small tubular neighborhoods of intervals γ1, γ2 with endpoints p1, p2 on Σ and their other endpoints on S(0,D) and as such we may conflate the ends Ei with their representative curves γi. From the definition of ambient isotopy (in that its an isotopy of all of R3) one can then extend the isotopy of Σ to that of M by following the induced isotopies of the γi attached points pi on Σ as its isotoped to a round sphere S with g handles attached to give an isotopy class presentation of M as a round sphere with g handles attached in a standard way along the north pole hemisphere and, by the argument ruling out case (1) above, arranged so the ends are attached on the south hemisphere and so that they do not link with the handles. Because the smooth mean curvature flow is an isotopy itself we can undo it so up to isotopy M and hence Mf are left unchanged by the flow untill the first surgery time is reached. Since the flow is outward from the region R from lemma 2.4 the ends remain connected to S across surgery times and, from lemma 3.2 above, the ends do not change isotopy class under the flow (modulo some spherical components surgered away). At the surgery time some number of handles, say ` handles, may be either filled, broken, or disconnected from S ⊂ Mf by the surgery but since Σ was standardly embedded the presurgery domain containing the ends is isotopic to the post surgery domain by adding back ` handles say along the equator, using again the fact that these handles don’t link with the ends. After the first surgery time the process is repeated until the second surgery time, et cetera. Since no new handles may form under the flow, the claim that M is isotopic to MT with g handles attached in a standard way follows.

So, if MT is a standardly embedded cylinder then M will be isotopic to a standardly embedded cylinder with g handles atached in a standard way, completing the proof. This we already know from the discussion above though, because MT is a cylinder so we may choose the boundaries of E1 and E2 to agree so theorem 1.1 for two ends/corollary 1.2 follows from having ruled out cases (2) and (3) above.

3.2. An argument for any number of ends. Using the elements of the argument above and the aforementioned observation of Bernstein and L. Wang we are now ready to describe how to proceed when there are more than two ends (since we use some of the arguments above, this argument is relatively short). Following the argument above, we see there is a time T for which i) MT is diffeomorphic to a sphere with k half cylinders attached and ii) that M is isotopic to MT with some number of handles attached in a standard way.

Proceeding exactly as in the proof of lemma 3.2 we see that for t > T , Mt is isotopic to MT . Undoing the reparameterization in the RMCF back to the “regular” MULTIPLE ENDS 17 mean curvature flow and writing it as Mft, this implies there is a time −1 < s < 0 for which Mfs is shrinker mean convex in the sense that equation (4.44) of [2] holds at smooth points (for us what matters in the following is that −2tH + x · ν ≥ 0), which by abuse of notation we will also refer to as shrinker mean convexity when the context is clear, that Mfs is isotopic to MT refering to the last paragraph, and that for s < t < 1 Mft is isotopic to Mfs.

As observed by Bernstein and L. Wang at time t = 0 if Mft is smooth its must be starshaped with respect to the origin in that x · ν ≥ 0; note that the reparam- eterization in the relation of RMCF to MCF only has time domain [−1, 0) but in the unparameterized picture the corresponding convexity condition makes sense and for smooth flows is (in our setting) preserved for nonnegative times as well. On the other hand, since our flow is not smooth and since we worked in the RMCF setting in [39] in our construction of the surgery/level set flow we only really know the level set flow Mft is shrinker mean convex at smooth points for s ∈ [−1, 0) from what’s been shown thus far.

By standard compactness results, we may take a varifold limit to define Mf0 (it will be nonempty by using appropriately placed large spheres as barriers). Since for t < 0 Mf0 is shrinker mean convex the singular set (if any, in fact from [2] there most likely is none although our set of assumptions is slightly different) of Mf0 must consist only of cylindrical singularities or round points (from what was shown in [39] in the unrescaled picture the noncollapsing constant can be taken to be the same in any fixed ball from t = −1 through to t = 0, and in fact from [35] this holds for the RMCF as well), and we also see the multiplicity of convergence is 1. In fact, by Colding and Minicozzi’s stratification theory for the mean curvature flow [11] (and specialized to n = 2), in any ball B the singular set SB of Mf0 ∩ B must consist of only finitely many Lipschitz curves along with some number of round points – the latter we henceforth ignore since these must be disconnected from the noncompact component of Mf. Note although their results are stated for MCF’s with compact initial condition, we may apply their work in our setting because entropy is bounded throughout (by the entropy of the initial self shrinker), the flow constructed is an integral varifold a.e. time, and is unit density.

With this in mind we next claim then that at any smooth point of Mf0 we must have x · ν ≥ 0 (or locally starshaped with respect to the origin). To see this, from above (that the singular set is very tame) we know for any such point p there is a small ball B(p, ) where the surface is smooth and has bounded curvature. Taking  potentially even smaller, we have Mf0 ∩ B(p, ) is approximately a disk. Applying one-sided minimization to the flow in B(p, ) to get local area bounds (using the adaptation of MULTIPLE ENDS 18

L. Lin in [34]) along with Brakke regularity then gives smooth convergence in this ball for some small time. From this we get Mf0 ∩ B(p, ) is shrinker mean convex and hence, since t = 0, locally starshaped with respect to the origin.

Now none of the curves in the singular set S may be along one of the ends, because the flow is outward. As an upshot they are embedded intervals connected to a “smooth” part Mfbulk gotten by simply deleting S and replacing endpoints in the result as the following diagram illustrates:

Figure 5. Mf0 consists of a starshaped surface with possibly some singular “filligree” attached colored in red above, the blue dot repre- senting the origin.

The singular set holds no important topological information or else there would ∗ be times s < t < 1 for which Mft∗ would not be isotopic to Mfs, and so its clear that bulk bulk Mf is isotopic to Mfs and hence MT . Since Mf is smooth and satisfies x · ν ≥ 0 except at a discrete set of points, it must be unknotted (i.e. isotopic to a round sphere with some straight half cylinders attached) finishing the argument.

Relating to their argument, note that in their paper Bernstein and L. Wang addi- tionally show that the link is connected (and hence there is just one end). The extra ingredient in their argument that allows them to conclude this is their low entropy assumption which lets them conclude number of ends is greater than 1 one could find a stable self shrinker. MULTIPLE ENDS 19

4. Concluding remarks In this section we mainly discuss the result of Meeks and Yau on classical minimal surfaces with finite topology and the theorems from classical minimal surface theory their argument uses, along the way comparing with the situation for self shrinkers. As an aside there are also results for classical minimal surfaces with infinite topology by Frohman and Meeks [16, 17]; analogous results for self shrinkers would be interesting as well, but these problems seem further on the horizon so we don’t discuss them. Specifically, Meeks and Yau show that minimal surfaces in R3 are isotopic to a standardly embedded genus g surface with some number k parallel planes, each one attached to the adjacent ones by an unknotted arc (see fig. 1 in their paper).

In their approach on the unknottedness problem for classical minimal surfaces in R3 with finite topology, Meeks and Yau in [37] instead argue by reducing to the unknottedness question of minimal surfaces with one end (which is a consequence of Waldhausen’s theorem) by cutting along carefully chosen disks. An important ingredient in carrying this out is the topologically parallel ends theorem, which says that a classical minimal surface with k ends and finite topology is isotopic to a surface asymptotic to k parallel planes.

The topologically parallel ends theorem is intuitively true for classical minimal because the blowdown of a minimal surface often yields a smooth cone with minimal link (possibly with multiplicity) which in the surface case must be a plane – if there were multiple planes in the limit it seems clear that there must be infinite topology so there is just one plane giving the statement. Unfortunately though this isn’t a generally applicable argument for classical minimal surfaces, an issue being if some ends have infinite total curvature. To deal with this, Meeks and Yau use that all but two ends have finite total curvature by work of Hoffman and Meeks [24] (the details of which use some deep results on classical minimal surfaces in turn e.g. the paragraph after claim 2 in their paper, which clearly fails for the cylinder, although maybe these could be somehow worked around) and that the ones with finite total curvature must have great spheres as links as shown by Jorge and Meeks in [30] (this is not specific to minimal surfaces), from which the result easily follows.

As mentioned in the introduction at a superficial level there deosn’t seem to be any reason to expect the ends of a self shrinker to be parallel, and there are well known differences two cases. For example, even when the link of the blowdown cone is smooth, the link need not be minimal as the examples of Kapouleas, Kleene, and Møller show, and it follows from L. Wang’s work that the blowdown of a self shrinker with multiple ends could never be a single great circle, giving already rigourously justified relevent differences between the asymptotics of self shrinkers and classical MULTIPLE ENDS 20 minimal surfaces. In fact it seems there could be examples of nonparallel examples by somehow desingularizing two transverse cylinders. On the other hand, there are some plausible candidates for self shrinkers with parallel ends as well. For example, a desingularization of the plane and cylinder would yield a self shrinker with 3 parallel ends (and, if the desingularization process preserves the quality of being asymptotically cylindrical, would also be a counterex- ample to the no-cylinder conjecture as observed by Angenent – see [41]) but it seems it remains for even any convincing numerical evidence for shrinkers with more than 2 ends to be given. We note it’s clear from their figures that it’s plausible one could desingularize any of Chopp’s and Ilmanen’s proposed shrinkers with a plane to pro- duce examples with 3 parallel ends as well. With Kapouleas, Kleene, and Møller’s construction in mind it seems likely that the plane under desingularization becomes “wobbly,” implying all three ends would have infinite total curvature in these cases. Note that in these proposals all three ends seem likely to not have flat link and hence have infinite total curvature, in contrast to whats possible in the classical minimal case [24]. Note that when the number of ends is greater than 3 in the statement of theo- rem 1.1 they cannot be parallel, so, if a parallel ends theorem for shrinkers is true this would render this paper useless for k > 2 (although the result is still new for k = 2). The upshot of the previous paragraphs is that the status of this assertion seems unclear, with some evidence of varying degrees of certainty both against it and suggestive of it. Continuing along with our discussion of the argument of Meeks and Yau, the next step after knowing the ends are parallel is to argue, using an application of Meeks- Simon-Yau [38], to argue that adjacent parallel ends are up to isotopy connected by unknotted arcs. Apriori, [38] applies only in ambient manifolds of bounded geometry. But since one eventually works in a large ball anyway (in which the Gaussian metric is well controlled –its incompleteness is relegated to spatial infinity), the argument that adjacent ends are connected by unknotted arcs seems to carry though without ad- justment. After this step, Meeks and Yau find appropriate loops to fill in with minimal discs and in so doing reduce the question of unknottedness to minimal surfaces in the ball with one boundary component. This seems to be an issue in the case of shrinkers, because the proof in [39] that shrinkers with one end are topologically standard used that the shrinkers in question were complete. On the other hand, it seems plausible (even though as discussed above this type of issue can be subtle) that the surgeries done in Meeks and Yau preserve the quality of being a Heegaard splitting which MULTIPLE ENDS 21 would circumvent this issue because we know from the flow argument the original self shrinker will be a Heegaard splitting (even though Waldhausen’s theorem doesn’t apply with just this knowledge) so that, if a parallel ends theorem for shrinkers were true, it seems plausible that then one could conclude self shrinkers of finite topology were topologically standard exactly in the sense of Meeks and Yau. It could be possible to show this under the assumption of parallel ends and would seem to furnish a different proof of corollary 1.2 as well (and giving a stronger statement even).

Along the same lines, in the conclusion of the first argument above (for two ends) it seems possible one argue that after every surgery the surface remains a Heegaard splitting under some additional assumptions. Then after all the handles are surgered away one could use that a Heegaard spliting which is an annulus is topologically standard to conclude. Of course, even surgeries which seem “uncomplicated” could ruin the property of being a Heegaard spliting as Hall’s examples show, so as above it seems some care would be needed to justify this.

Without a parallel ends theorem in hand, the most general configuration for asymp- totically conical ends seems to be that they come in families of parallel ends, or in other words a mix of the phenomena our article considers with what Meeks and Yau contend with in theirs. Hastily jumping ahead and applying the argument (if possible) of Meeks-Yau to each family of parallel ends, it seems one could still possibly reduce down the general question of knottedness to the case considered in theorem 1.1. Then any self shrinker with asymptotically conical ends attached along a standardly embedded surface Σ could be represented as a standardly embedded compact surface with a number of “standard parallel families” of ends attached, as the diagram below exhibits. MULTIPLE ENDS 22

Figure 6. An example of a conjectural “standard configuration” for a self shrinker M with 4 families of parallel ends. Here the red “legless Y’s” are a 1-dimensional representation of the conical ends, and the blue curves represent bridges between parallel ends and from parallel families to the compact “soul” of M.

Of course and as already mentioned above, it would be preferable to remove the assumption that Σ is standardly embedded. Intuitively this seems plausible because the flows of knotted handles (in the suitable sense) should snag when flowing in the appropriate direction, but their configurations with each other could concievably be quite complicated and so its unclear presently to the author what types of changes might take place across surgery times. For example, it seems concievable that a num- ber of surgeries could occur that take a nonstandardly embedded surface eventually to one which is standardly embedded (at least without further input from topology, perhaps using more advanced insights from topology but still weaker than Wald- hausen) so that the flow never detects that the original surface was nonstandard.

References [1] Sigurd Angenent and J.J.L. Vel´azquez. Degenerate neckpinches in mean curvature flow. J. reine angew. Math. 482 (1997), 15-66. [2] Jacob Bernstein and Lu Wang. A Topological Property of Asymptotically Conical Self-Shrinkers of Small Entropy. Duke Math. J. 166, no. 3 (2017), 403-435 [3] Kenneth Brakke. The Motion of a Surface by its Mean Curvature. Press, 1978. [4] Simon Brendle. Embedded self-similar shrinkers of genus 0. Annals of Mathemat- ics, vol. 183, no. 2, 2016, pp. 715–728. [5] Simon Brendle. Minimal surfaces in S3: a survey of recent results. Bull. Math Sci 3 (2013), 133-171. MULTIPLE ENDS 23

[6] Simon Brendle and . Mean curvature flow with surgery of mean convex surfaces in R3. Invent. Math 203, 615-654 [7] David Chopp. Computation of self-similar solutions for mean curvature flow. Ex- periment. Math. 3 (1994), no. 1, 1–15. [8] Otis Chodosh, Kyeongsu Choi, Christos Mantoulidis, and Felix Schulze. Mean curvature flow with generic initial data. arXiv:2003.14344 [9] and William Minicozzi II. Generic mean curvature flow I; generic singularities. Annals of Mathematics. (2) 175 (2012), 755-833. [10] Tobias Colding and William Minicozzi II. Smooth compactness of self-shrinkers. Comment. Math. Helv. 87 (2012), 463-475. [11] Tobias Colding and William Minicozzi II. The singular set of mean curva- ture flow with generic singularities. Inventiones mathematicae volume 204, pages 443–471(2016) [12] Tobias Colding, William Minicozzi II and Erik Pedersen. Mean Curvature Flow. Bull. Amer. Math. Soc. 52 (2015), 297-333 [13] Tobias Colding, Tom Illmanen, William Minicozzi II, and Brian White. The round sphere minimizes entropy among closed self-shrinkers. J. Differential Geom. 95 (2013), 53-69 [14] Qi Ding and Y.L. Xin. Volume growth eigenvalue and compactness for self- shrinkers. Asian J. Math. Volume 17, Number 3 (2013), 443-456. [15] Theodore Frankel. On the fundamental group of a compact minimal submanifold. Annals of Mathematics 83 (1964), 68-73. [16] Charles Frohman and W. H. Meeks III. The topological classification of minimal surfaces in R3. Annals of Math, 167 (2008), 681–700 [17] Charles Frohman and W. H. Meeks III. The topological uniqueness of complete one-ended minimal surfaces and Heegaard surfaces in R3. J. Amer. Math. Soc. 10 (1997), 495–512. [18] Peter Hall. Two topological examples in minimal surface theory. J. Differential Geom. Volume 19, Number 2 (1984), 475-481. [19] Robert Haslhofer and Bruce Kleiner. Mean curvature flow with surgery. Duke Math. J., Volume 166, Number 9 (2017), 1591-1626. [20] Robert Haslhofer and Daniel Ketover. Minimal 2-spheres in 3-spheres. Duke Math. J., Volume 168, Number 10 (2019), 1929-1975. [21] John Head. The Surgery and Level-Set Approaches to Mean Curvature Flow. Thesis. [22] John Head. On the mean curvature evolution of two-convex hypersurfaces. J. differential geometry 94 (2013) 241-266 [23] Gerhard Huisken. Asymptotic behavior for singularities of the mean curvature flow. J. Differential Geom. 31 (1990), no. 1, 285–299. MULTIPLE ENDS 24

[24] David Hoffman and William Meeks III. The asymptotic behavior of properly embedded minimal surfaces of finite topology. J. AMS 2(4) (1989), 667-681. [25] Gerhard Huisken and Carlo Sinestrari. Mean curvature flow with surgeries of two-convex hypersurfaces. Invent. Math 175, 137-221 (2009) [26] Dan Ketover. Self-shrinking Platonic solids. Preprint, arxiv:1602.07271 [27] Tom Ilmanen. Singularities of mean curvature flow of surfaces. Preprint, 1995. [28] Tom Ilmanen. Problems in mean curvature flow. Available at http://people.math.ethz.ch/ ilmanen/classes/eil03/problems03.ps [29] Tom Ilmanen. Lectures on Mean Curvature Flow and Related Equations. Online. [30] Luquesio Jorge and William Meeks. The topology ofcomplete minimal surfaces of finite total Gaussian curvature. Topology 22 (1983). 203-221. [31] Nikolaos Kapouleas, Stephen Kleene, and Niels Martin Møller. Mean curvature self-shrinkers of high genus: Non-compact examples. Journal f¨urdie reine und angewandte Mathematik, 2018(739), 1-39. [32] Joseph Laurer. Convergence of mean curvature flows with surgery. Comm. Anal. Geom., Volume 21, Number 2, 355-363, 2013. [33] Blaine Lawson. The unknottedness of minimal embeddings. Invent. Math. 11 (1970), 41–46. [34] Longzhi Lin. Mean curvature flow of star-shaped hypersurfaces. To appear in Comm Anal Geom. [35] Zhengjiang Lin and Ao Sun. Bifurcation of perturbations of non-generic closed self-shrinkers. To appear in journal of topology and analysis, arXiv:2004.07787 [36] William Meeks III. The topological uniqueness of minimal surfaces in three di- mensional Euclidean space. Topology 20:389-410, 1981. [37] William Meeks III and Shing-Tung Yau. The topological uniqueness of complete minimal surfaces of finite topological type. Topology 31:305-315, 1992. [38] William Meeks III, Leon Simon, and Shing-Tung Yau. Embedded minimal sur- faces, exotic spheres, and manifolds with positive Ricci curvature. Annals of Math- ematics (1982) 621-659. [39] Alexander Mramor. An unknottedness result for noncompact self shrinkers. Preprint, arXiv:2005.01688 [40] Alexander Mramor and Shengwen Wang. On the topological rigidity of compact self shrinkers in R3. Int. Mat. Res. Not., rny050. [41] Xuan Hien Nguyen. Construction of complete embedded self-similar surfaces un- der mean curvature flow. Part III. Duke Math J. Volume 163, number 11 (2014), 2023-2056 [42] Ao Sun, Zhichao Wang, and Xin Zhou. Multiplicity one for min-max theory in compact manifolds with boundary and its applications. Preprint, arXiv:2011.04136 [43] Friedhelm Waldhausen. Heegaard-zerlegungen der 3-sphere. Topology 7:195-203, 1968. MULTIPLE ENDS 25

[44] Lu Wang. Asymptotic structure of self-shrinkers. Preprint, arXiv: 1610.04904 [45] Lu Wang. Uniqueness of self-similar shrinkers with asymptotically cylindrical ends. J. Reine Angew. Math. 715 (2016), 207-230. [46] Guofang Wei, Will Wylie. Comparison geometry for the Bakry-Emery Ricci ten- sor. J. Differential Geom. Volume 83, Number 2 (2009), 337-405. [47] Brian White. The size of the singular set in mean curvature flow of mean-convex sets. J. Amer. Math. Soc. 13 (2000), 665-695. [48] Brian White. A local regularity theorem for mean curvature flow. Annals of Mathematics. (2) 161 (2005), 1487-1519.

Department of Mathematics, , Baltimore, MD, 21231 Email address: [email protected]