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

Smooth convergence to the enveloping cylinder for mean curvature flow of complete graphical hypersurfaces

Wolfgang Maurer funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) Project number 336454636
Abstract

For a mean curvature flow of complete graphical hypersurfaces Mt=graphu(,t)M_{t}=\mathop{\mathrm{graph}}u(\cdot,t) defined over domains Ωt\Omega_{t}, the enveloping cylinder is Ωt×\partial\Omega_{t}\times\mathbb{R}. We prove the smooth convergence of Mthen+1M_{t}-h\,e_{n+1} to the enveloping cylinder under certain circumstances. Moreover, we give examples demonstrating that there is no uniform curvature bound in terms of the inital curvature and the geometry of Ωt\Omega_{t}. Furthermore, we provide an example where the hypersurface increasingly oscillates towards infinity in both space and time. It has unbounded curvature at all times and is not smoothly asymptotic to the enveloping cylinder. We also prove a relation between the initial spatial asymptotics at the boundary and the temporal asymptotics of how the surface vanishes to infinity for certain rates in the case Ωt\Omega_{t} are balls.

1 Introduction

Mean curvature flow is the evolution of a hypersurface that moves with normal velocity equal to the mean curvature. It is described by a particularly appealing geometric evolution equation for the embedding (ddtX=ΔtX\frac{\mathrm{d}}{\mathrm{d}t}X=\Delta_{t}X), while it is also the negative L2L^{2}-gradient flow of the area functional.

For hypersurfaces MtM_{t} that are the graphs of function u(,t)u(\cdot,t), they solve mean curvature flow if and only if uu solves the quasilinear parabolic partial differential equation

u˙=(δijuiuj1+|u|2)uij.\dot{u}=\left(\delta^{ij}-\frac{u^{i}\,u^{j}}{1+|\nabla u|^{2}}\right)u_{ij}\,.

(We make use of Einstein summation convention.) This equation has been extensively studied. To name a few of these works, that are in line with the present article, the author would like to mention [10], where the case of boundary values on a bounded domain is considered, [5], where the long-time existence of solutions for entire graphs is proven, and [19], where complete graphical hypersurfaces are considered. In the case of complete graphs, the representing functions u(,t)u(\cdot,t) are defined on open subsets Ωtn\Omega_{t}\subset\mathbb{R}^{n} and they diverge to infinity toward the boundary Ωt\partial\Omega_{t} such that no boundary values occur and the graphs of u(,t)u(\cdot,t) are in fact complete hypersurfaces. The domains Ωt\Omega_{t} must be time dependent in this setting and it turns out that they, or respectively their boundaries Ωt\partial\Omega_{t}, form a weak solution of mean curvature flow. This can be understood by the following heuristic. The hypersurfaces graphu(,t)\mathop{\mathrm{graph}}u(\cdot,t) are in some sense asymptotic to the cylinder Ωt×\partial\Omega_{t}\times\mathbb{R}. Therefore, mean curvature flow is also expected for Ωt×\partial\Omega_{t}\times\mathbb{R}, and hence for Ωt\partial\Omega_{t} because the \mathbb{R}-factor does not contribute to the mean curvature or the evolution by mean curvature flow.

The main aim of this article is to prove that graphu(,t)\mathop{\mathrm{graph}}u(\cdot,t) is smoothly asymptotic to the cylinder Ωt×\partial\Omega_{t}\times\mathbb{R} under certain assumptions. For initially bounded curvature and bounded Ω0\Omega_{0} the asymptotic is smooth for positive times t>0t>0 as long as Ωt\partial\Omega_{t} is not singular. It seems to be difficult to apply the method of proof, which relies on local graphical representations and pseudolocality, past singularities. Therefore, we investigate the case of noncollapsed mean curvature flow with a different method (the curvature bound of [9]) that allows going beyond singularities. So for the noncollapsed mean curvature flow of complete graphs the smooth asymptotics holds generally (still assuming initially bounded curvature).

If one drops the assumption of an initial curvature bound, the asymptotic may not be smooth anymore. We provide an interesting example where the asymptotic is non-smooth at any time and instead the graphical hypersurfaces infinitely sheets towards the enveloping cylinder. For the purpose of the construction we explicitely construct a barrier function which is defined over a shrinking annulus. This barrier allows for an estimate of uu in terms of the value of uu on a surrounding annulus earlier in time. Using this estimate we can prove that there are solutions uu over a shrinking ball Ωt\Omega_{t} for which u˙(0,t)\dot{u}(0,t) oscillates infinitely such that there are sequences tkTt_{k}\to T and τkT\tau_{k}\to T with u˙(0,tk)+\dot{u}(0,t_{k})\to+\infty and u˙(0,τk)\dot{u}(0,\tau_{k})\to-\infty. The example shows that complicated kinds of singularities can appear at infinity for complete graphical hypersurfaces.

The barrier can also be used to prove a relationship between the spatial asymptotic of u(x,0)u(x,0) as xΩ0x\to\partial\Omega_{0} and the temporal asymptotic u(0,t)u(0,t) as tTt\to T for rotationally symmetric graphs. In this context it is mandatory to mention [14, 15] where examples of rotationally symmetric, complete, graphical surfaces with a continuous spectrum of blow-up rates are constructed. The blow-up rates in their examples were intimately connected to the spatial asymptotics, too.

The article is organized as follows. Firstly, we will review the mean curvature flow of complete graphical hypersurfaces, because it is central to this paper and to set up the notation. In Section 3, we prove the smooth asymptotics to the enveloping cylinder up to the first singularity for initially bounded curvature. A curvature bound above some height is central to the proof. Despite this curvature bound above a height, we include in this section a sequence of examples demonstrating that there is no uniform curvature bound (independent of height) depending only on the initial curvature bound and the geometry of Ωt\Omega_{t}. In Section 4, we prove the smooth asymptotics for noncollapsed mean curvature flows of complete graphs. The section also provides a construction establishing existence for such flows. Section 5 is devoted to the barrier over an annulus and its applications. The appendix contains invaluable information about normal graphs and noncollapsed mean curvature flow as well as weak mean curvature flows.

The author thanks O. Schnürer for posing the question discussed here and for his supervision of the authors PhD-thesis, where this work originates from. The author is also very grateful for the conversations, stimulations and for O. Schnürer’s patience.

2 Recapitulation of mean curvature flow without singularities

Since this paper builds on the ideas and results of “Mean curvature flow without singularities” ([19]), it is worthwhile to summarize the main points of [19]. This section does not contain new results. (We follow [19], or [17] when we diverge from [19].) Instead, it helps setting the notation and allows us to shorten our exposition lateron when similar steps as here are needed to be taken.

Mean curvature flow without singularities is mainly about the mean curvature flow of complete graphical hypersurfaces.

  1. 1.

    Initial Data: Let Ω0n\Omega_{0}\subset\mathbb{R}^{n} be open and let u0:Ω0u_{0}\colon\Omega_{0}\to\mathbb{R} be a locally Lipschitz-continuous function. We assume that there is a continuous extension u¯0:n¯=[,]\overline{u}_{0}\colon\mathbb{R}^{n}\to\overline{\mathbb{R}}=[-\infty,\infty] of u0u_{0} such that {x:<u¯0(x)<}=Ω0\{x:-\infty<\overline{u}_{0}(x)<\infty\}=\Omega_{0} and u¯0|Ω0=u0\overline{u}_{0}|_{\Omega_{0}}=u_{0} hold.

  2. 2.

    Solution Data: A mean curvature flow without singularities is a pair (u,Ω)(u,\Omega) of an relatively open subset Ωn×[0,)\Omega\subset\mathbb{R}^{n}\times[0,\infty) and a continuous function u:Ωu\colon\Omega\to\mathbb{R}. The zero time slice of Ω\Omega is supposed to be Ω0\Omega_{0}, in line with a consistent notation Ωt\Omega_{t} for the time slices (Ω=t0Ωt×{t}\Omega=\bigcup_{t\geq 0}\Omega_{t}\times\{t\}). Moreover, we suppose that u(,0)=u0u(\cdot,0)=u_{0} holds. For this reason, we call (u0,Ω0)(u_{0},\Omega_{0}) the initial data for (u,Ω)(u,\Omega).

    Maximality condition: We suppose that there exists a continuous function u¯:n×[0,)¯\overline{u}\colon\mathbb{R}^{n}\times[0,\infty)\to\overline{\mathbb{R}} such that {(x,t):<u¯(x,t)<}=Ω\{(x,t):-\infty<\overline{u}(x,t)<\infty\}=\Omega and u|Ω=uu|_{\Omega}=u hold.

    Equation: The function uu is supposed to be smooth and to satisfy the equation of graphical mean curvature flow on Ω(Ω0×{0})\Omega\setminus(\Omega_{0}\times\{0\}), i.e.,

    (1) tu=(δijuiuj1+|Du|2)uij.\partial_{t}u=\left(\delta^{ij}-\frac{u^{i}\,u^{j}}{1+|Du|^{2}}\right)u_{ij}\;.
  3. 3.

    Hypersurfaces: We denote by Mtgraphu(,t)M_{t}\coloneqq\mathop{\mathrm{graph}}u(\cdot,t) the graphical hypersurfaces that move by their mean curvature (locally in a classical sense).

  4. 4.

    Shadow flow: The family (Ωt)t0(\Omega_{t})_{t\geq 0} is called the shadow flow.

Theorem 1.

For any such initial data (u0,Ω0)(u_{0},\Omega_{0}), there exists a corresponding mean curvature flow without singularities (u,Ω)(u,\Omega). The shadow flow is a weak solution of mean curvature flow in dimension n1n-1 in a sense explained below (Remark 2 (3)).

Remark 2.
  1. 1.

    The maximality condition implies |u(x,t)||u(x,t)|\to\infty for (x,t)Ω(x,t)\to\partial\Omega. (Ω\partial\Omega denotes the relative boundary of Ω\Omega in n×[0,)\mathbb{R}^{n}\times[0,\infty).) In particular, the hypersurfaces MtM_{t} are complete. Moreover, the maximality condition implies that the solution is maximal in tt; stopping the flow at an arbitrary time may prevent the maximality condition to hold.

    The maximality condition is defined slightly differently in [19]. Only positive and proper functions uu are considered there. We follow [17] here, where these assumptions are dropped to some extent and the maximality condition is adapted accordingly.

  2. 2.

    Although MtM_{t} is smooth, the formalism allows for changes of the topology of MtM_{t}. Singularities of Ωt\partial\Omega_{t} may be interpreted as singularities of MtM_{t} at infinity.

  3. 3.

    In [19], the shadow flow is advertised as a weak solution. They underpin this by showing that (for their solution) (Ωt)t[0,)(\Omega_{t})_{t\in[0,\infty)} coincides with the level-set flow starting from Ω0\Omega_{0} n\mathcal{H}^{n}-almost everywhere if the level-set flow is non-fattening.

    In [17], for an arbitrary solution (u,Ω)(u,\Omega) the shadow flow is interpreted as a weak solution in the sense of a domain flow (cf. Definition 25).

Discussion of the proof of Theorem 1.

One constructs an approximating sequence of functions vk:n×[0,)¯v_{k}\colon\mathbb{R}^{n}\times[0,\infty)\to\overline{\mathbb{R}}. It needs to satisfy the following form of local equicontinuity: For any aa\in\mathbb{R}, any RR\in\mathbb{R}, and any ε>0\varepsilon>0 there is δ=δ(a,R,ε)>0\delta=\delta(a,R,\varepsilon)>0 and an index K=K(a,R,ε)K=K(a,R,\varepsilon)\in\mathbb{N} such that for any kKk\geq K and any (x,s),(y,t)n×[0,)(x,s),(y,t)\in\mathbb{R}^{n}\times[0,\infty) with |x|<R|x|<R, |vk(x,s)|<a|v_{k}(x,s)|<a, and |xy|+|st|<δ|x-y|+|s-t|<\delta we have |vk(x,s)vk(y,t)|<ε|v_{k}(x,s)-v_{k}(y,t)|<\varepsilon. A variation on the Arzelà-Ascoli theorem then shows that a subsequence of (vk)k(v_{k})_{k\in\mathbb{N}} converges pointwise to a continuous function u¯:n×[0,)¯\overline{u}\colon\mathbb{R}^{n}\times[0,\infty)\to\overline{\mathbb{R}}. We set Ω{(x,t):<u¯(x,t)<}\Omega\coloneqq\{(x,t):-\infty<\overline{u}(x,t)<\infty\} and uu¯|Ωu\coloneqq\overline{u}|_{\Omega}. The convergence is locally uniform on Ω\Omega.

To be approximating, the sequence needs to satisfy vk(,0)u¯0v_{k}(\cdot,0)\to\overline{u}_{0} pointwise, where u¯0\overline{u}_{0} is the above extension of the initial function such that u¯(,0)=u¯0\overline{u}(\cdot,0)=\overline{u}_{0} holds. Furthermore, for any a,Ra,R\in\mathbb{R} and any t0>0t_{0}>0, we suppose that there is an index K=K(a,R,t0)K=K(a,R,t_{0}) such that vkv_{k} is a smooth solution of (1) on the set {(x,t):|vk(x,t)|<a,|x|<R,t0<|t|<R}\{(x,t):|v_{k}(x,t)|<a,\,|x|<R,\,t_{0}<|t|<R\}. Moreover, we assume uniform estimates of the form |αvk|C(α,a,R,t0)|\partial^{\alpha}v_{k}|\leq C(\alpha,a,R,t_{0}) for kKk\geq K and for all multi-indices α\alpha on this set. The subsequential convergence vkuv_{k}\to u is then locally smooth on Ω(Ω0×{0})\Omega\setminus(\Omega_{0}\times\{0\}). As a consequence, uu solves (1) on Ω(Ω0×{0})\Omega\setminus(\Omega_{0}\times\{0\}) and is as asserted.

We have summarized how one obtains a solution from an approximating sequence and what are sufficient conditions on this sequence. One still has to find the approximating sequence and prove the local estimates on the functions and its derivatives. For the approximations one could either solve initial boundary value problems or use the flow of closed hypersurfaces which have graphical parts. These are just two options and one cannot say in general how to approximate; it depends on the given problem.

For example, we can find an approximating sequence in the following way. One considers for a+a\in\mathbb{R}_{+} the functions φa:¯\varphi_{a}\colon\overline{\mathbb{R}}\to\mathbb{R} with

(2) φa(x){x|x|<a,axa,axa.\varphi_{a}(x)\coloneqq\begin{cases}x&|x|<a,\\ a&x\geq a,\\ -a&x\leq-a.\end{cases}

Then one mollifies φau¯\varphi_{a}\circ\overline{u} and restricts to a ball. Solving graphical mean curvature flow on this ball with this initial function and holding the boundary values fixed over time, we find an approximator. (It can be extended to all of n\mathbb{R}^{n} by an arbitrary value.) An approximating sequence is obtained by taking increasingly larger aa, finer mollification parameter, and larger balls.

For the local estimates it is often possible to use the height function to construct a cut-off function.

We don’t say anything about the shadow flow here because the details of that part of the proof are not important to us. ∎

3 Curvature bounds for mean curvature flow of complete graphs

In this section we give hypothesis that render the heuristics rigorous that for a mean curvature flow without singularities (u,Ω)(u,\Omega), the graph of u(,t)u(\cdot,t) and Ωt×\partial\Omega_{t}\times\mathbb{R} behave alike in sufficient heights. The first section, however, lowers our expectations towards a curvature bound for mean curvature flow of complete graphs.

3.1 Impossibility of a controlled curvature bound

In the title of this paragraph, one would need to explain what is meant by “controlled”. The issue is, what quantities is the curvature bound allowed to depend on? We will only allow the geometry of Ωt\Omega_{t} and the initial curvature here. We cannot exclude curvature bounds that depend on more elaborate quantities.

We are going to construct a sequence of examples with uniformly bounded initial curvature over a domain which is a shrinking ball. Nevertheless, the curvature in these examples becomes arbitrarily large at some fixed time before the ball disappears. This demonstrates that there is no uniform curvature bound for the mean curvature flow without singularities that depends only on the initial curvature and the geometry of Ωt\Omega_{t}.

We start with Ω0=B1(0)2\Omega_{0}=B_{1}(0)\subset\mathbb{R}^{2}. Every solution (u,Ω)(u,\Omega) of the mean curvature without singularities has Ω\Omega resemble the shrinking ball solution, which exists up to the time t=12t=\frac{1}{2} when uu vanishes to infinity. We start with an initial u0u_{0} that is sketched in Figure 1. The torus shown in that figure is assumed to close at time t=14t=\frac{1}{4} when moved by its mean curvature. The mean curvature flow without singularities MtM_{t} stays a smooth graph. So by the avoidance principle MtM_{t} must have already passed through the torus by the time the torus closes. Now consider a sequence of initial surfaces M0M_{0} that look essentially the same but have their tip further and further downstairs. These surfaces can be constructed such that their curvature is uniformly bounded among all of them. But at time t=14t=\frac{1}{4}, the tips must have passed through the torus. For this reason, the speeds of the tips must become arbitrarily large because they have to travel increasing distances in the same time. That means the mean curvature becomes arbitrarily large. We conclude that controlling the curvature of M0M_{0} and the geometry of Ωt\Omega_{t} is not sufficient to deduce a curvature bound for MtM_{t} at later times.

Refer to caption
Figure 1: The torus closes and the solution must have passed through before.

In this counterexample sequence the problem does not arise from infinity. In fact, there everything is well-behaved. Furthermore, none of the members of the sequence actually has unbounded curvature. Therefore, we will keep the hypothesis of initially bounded curvature and of controlled geometry of Ωt\Omega_{t}, but we will weaken the assertion. We will only obtain an uncontrolled curvature bound. In accordance with the fact that in our counterexample sequence the problem does not come from infinity, we obtain that under those hypothesis the solution MtM_{t} is smoothly asymptotic to the cylinder Ωt×\Omega_{t}\times\mathbb{R}.

3.2 Smooth asymptotics to the cylinder

For interior curvature estimates we rely on the pseudolocality of mean curvature flow. This remarkable property of mean curvature flow means that one has control on solutions by means of local data; the behavior of the solution far out is irrelevant for that control. This is not to be confused with locality because a change far out will have effects everywhere, but these are mild enough such that estimates can still be derived from local data. This should be contrasted to the heat equation, where one can realize any given temperature after any given time at any given point in a room by making the walls of the room sufficiently hot.

We use the interior curvature estimates from [2], but one could also use [16], which is based on the pseudolocality theorem in [13], in conjunction with [5].

Theorem 3.

There exist 0<ε10<\varepsilon\leq 1 and L>0L>0 with the following property. Let r>0r>0. Let (Mt)0tT(M_{t})_{0\leq t\leq T} be a smooth solution of mean curvature flow which is properly embedded in Br(x0)B_{r}(x_{0}) with x0M0x_{0}\in M_{0} and such that M0M_{0} is graphical in Br(x0)B_{r}(x_{0}) with gradient bounded by LL. Then the second fundamental form is estimated by

(3) |A(x,t)|21t+(εr)2|A(x,t)|^{2}\leq\frac{1}{t}+(\varepsilon\,r)^{-2}

for t(0,min{(εr)2,T}]t\in(0,\min\{(\varepsilon r)^{2},T\}] and xBεr(x0)Mtx\in B_{\varepsilon r}(x_{0})\cap M_{t}.

If, in addition, the second fundamental form of M0Br(x0)M_{0}\cap B_{r}(x_{0}) is bounded by r1r^{-1}, then

(4) |A(x,t)|2(εr)2|A(x,t)|^{2}\leq(\varepsilon\,r)^{-2}

holds for t[0,min{(εr)2,T}]t\in[0,\min\{(\varepsilon r)^{2},T\}] and xBεr(x0)Mtx\in B_{\varepsilon r}(x_{0})\cap M_{t}.

Theorem 4.

Let (u,Ω)(u,\Omega) be a solution of the mean curvature flow without singularities. Suppose that (Ωt)0tT({\partial}{\Omega}_{t})_{0\leq t\leq T} is a smooth and compact solution of the mean curvature flow in dimension n1n-1 for some T>0T>0. Moreover, suppose M0M_{0} has bounded curvature.

Then Mthen+1M_{t}-h\,e_{n+1} converges for 0<tT0<t\leq T in ClocC_{\mathrm{loc}}^{\infty} to the cylinder Ωt×\partial\Omega_{t}\times\mathbb{R} for hh\to\infty. In particular, supxMt|A|2(x,t)\sup\limits_{x\in M_{t}}|A|^{2}(x,t) is bounded for 0tT0\leq t\leq T.

Remark.

Instead of bounded curvature, one can assume that M0M_{0} admits local graph representations of some radius rr with small Lipschitz constant (cf. Definition 17). Then the curvature of MtM_{t} is bounded for small positive times by Theorem 3.

Proof of Theorem 4.

Firstly, we are going to prove a curvature estimate above some height. The asymptotic behavior is then treated afterwards. The idea for the curvature estimate is to alternately use Theorem 3, which provides a curvature estimate from a local graph representation, and that a curvature estimate ensures graph representations, as we are going to explain below. These two complement each other perfectly if done carefully. Essential is that the curvature is below a certain threshold CAC_{A} so that we can ensure graph representations. Theorem 3 then supplies us with a curvature bound a small, but fixed, time later. If this bound is below CAC_{A}, we can then repeat the argument. The rough idea is depicted in Figure 2.

Refer to caption
Figure 2: The rough idea of the proof.

We set ZtΩt×Z_{t}\coloneqq\partial\Omega_{t}\times\mathbb{R}.

Let 0<ε0<\varepsilon, L1L\leq 1 be as in Theorem 3. By Proposition 18 there is r>0r>0 dependent on

L,supt[0,T]supxZt|AZt(x,t)|,andsupt[0,T]supx,yZtdistZt(x,y)|xy|L,\quad\sup_{t\in[0,T]}\sup_{x\in Z_{t}}|A_{Z_{t}}(x,t)|\,,\quad\text{and}\quad\sup_{t\in[0,T]}\sup_{x,y\in Z_{t}}\frac{\mathrm{dist}_{Z_{t}}(x,y)}{|x-y|}

with the following property. If MM is any normal graph over ZtZ_{t} with gradient bounded by L/6L/6 and which is in a tubular neighborhood of thickness at most half of the maximal tubular neighborhood thickness, then MM admits local graph representations of radius rr with gradient bounded by LL. By decreasing rr if necessary, we may assume r<1r<1 and supM0|A|r1\sup_{M_{0}}|A|\leq r^{-1}, where |A||A| corresponds to M0M_{0}.

We set CA2(εr)1C_{A}\coloneqq 2(\varepsilon\,r)^{-1}. This will function as a threshold for the curvature. By Corollary 20 there is δ>0\delta>0 such that for any two hypersurfaces MMM^{\prime}\subset M, where MM^{\prime} has buffer 11 in MM, with curvature bounded by CAC_{A} and which lie in a (tubular) neighborhood of ZtZ_{t} of size δ\delta, can be locally written as a normal graph over ZtZ_{t} with gradient bounded by L/8L/8.

We set

asup{u(x,t):t[0,T],xΩt,dist(x,Ωt)δ},a\coloneqq\sup\{u(x,t)\colon t\in[0,T],\,x\in\Omega_{t},\,\mathrm{dist}(x,\partial\Omega_{t})\geq\delta\}\;,

such that MtaMt{xn+1>a}M_{t}^{a}\coloneqq M_{t}\cap\{x^{n+1}>a\} lies in a tubular neighborhood of ZtZ_{t} of size δ\delta. By Lemma 21, Mta+1M_{t}^{a+1} is a normal graph over ZtZ_{t} (following Definition 14, where it can be a graph over a subset) if the curvature of MtaM_{t}^{a} is bounded by CAC_{A}. By our choice of rr in the beginning of the proof, this shows that Mta+1M_{t}^{a+1} admits local graph representations of radius rr with gradient bound LL if MtaM_{t}^{a}’s curvature is bounded by CAC_{A}.

Because of supM0|A|r1<CA\sup_{M_{0}}|A|\leq r^{-1}<C_{A}, M0a+1M_{0}^{a+1} admits local graph representations of radius rr with gradient bounded by LL. The interior estimate (4) yields

(5) |A(x,t)|2(εr)2<(CA)2|A(x,t)|^{2}\leq(\varepsilon\,r)^{-2}<(C_{A})^{2}

for t[0,min{(εr)2,T}]t\in[0,\min\{(\varepsilon r)^{2},T\}] and xMta+1+rMta+2x\in M_{t}^{a+1+r}\subset M_{t}^{a+2}.

From (3) we infer that, in the case that the curvature of MtaM_{t}^{a^{\prime}} (aaa^{\prime}\geq a) is smaller than CAC_{A} and, therefore, Mta+1M_{t}^{a^{\prime}+1} admits local graph representations of radius rr, like it is demonstrated above, the estimate

(6) |A(x,t+(εr)2)|22(εr)2<(CA)2\big{|}A\big{(}x,t+(\varepsilon\,r)^{2}\big{)}\big{|}^{2}\leq 2\,(\varepsilon\,r)^{-2}<(C_{A})^{2}

holds for xMt+(εr)2a+2x\in M_{t+(\varepsilon\,r)^{2}}^{a^{\prime}+2}.

Starting from the bound (5) on the interval [0,ε2r2][0,T][0,\varepsilon^{2}\,r^{2}]\cap[0,T], the estimate (6) can be applied iteratively to show that the curvature of Mta+2NM_{t}^{a+2N} (NN is some controlled number of steps) cannot reach the threshold CAC_{A} for any t[0,T]t\in[0,T]. The procedure is demonstrated in Figure 3. Starting with the curvature at the brown dot, (5) establishes the curvature bound depicted by the brown line. Since the curvature stays below the threshold CAC_{A}, we can apply the curvature bound (6) to obtain the curvature bound in red. In fact, the picture even shows how to obtain (6) from (3). Starting from the red curvature bound, which is below the threshold CAC_{A}, one can again apply (6) to obtain the orange curvature bound. In this way, we proceed through all the colors of the rainbow (and beyond). We will reach the time TT in finitely many steps, providing us with curvature estimates on the whole time interval. However, the region where the estimates hold go up in height by 2 units per step.

Refer to captionsup|AM0|\sup|A_{M_{0}}|(εr)1(\varepsilon r)^{-1}(εr)2(\varepsilon r)^{2}2(εr)1\sqrt{2}(\varepsilon r)^{-1}CA=2(εr)1C_{A}=2(\varepsilon r)^{-1}|A||A|tt
Figure 3: The iterative procedure to obtain the curvature estimate.

Below the height a+2Na+2N the curvature is bounded (in a non-controlled fashion) because it is a continuous function on a compact set.

Now that we have curvature estimates, we can work on improving them: Because of the curvature bound, in great heights the solution MtM_{t} is C1C^{1}-close to the cylinder ZtZ_{t}. This follows from Lemma 21. We translate the solutions downwards, so we consider Mthen+1M_{t}-h\,e_{n+1}, and by the Arzelà-Ascoli-Theorem we can extract a limit for hh\to\infty which must be the enveloping cylinder ZtZ_{t}. For t>0t>0, the convergence is not only uniform (C0C^{0}) but by interior estimates for |kA||\nabla^{k}A|, that hold for mean curvature flows of bounded curvature [5], and interpolation inequalities the uniform convergence expands to smooth convergence. In other words, the solution MtM_{t} is smoothly asymptotic to the cylinder ZtZ_{t}. ∎

Remark.

In the dimensions n=1,2n=1,2, the result holds beyond singularities: One can obtain the smooth convergence of Mthen+1M_{t}-h\,e_{n+1} to the cylinder Ωt×\partial\Omega_{t}\times\mathbb{R} for all positive times t>0t>0 provided Ωt\partial\Omega_{t} is not singular. We are assuming that Ω0\Omega_{0} is smooth and bounded.

In the case n=1n=1, Ω0\Omega_{0} is a union of bounded intervals with fixed distances between them and Ω0=Ωt\Omega_{0}=\Omega_{t} holds for all times. Thus, there are no singular times and MtM_{t} is smoothly asymptotic to Ωt×\partial\Omega_{t}\times\mathbb{R} for all t>0t>0.

In the case n=2n=2, the Gage-Hamilton-Grayson theorem [7, 8] ensures that any singularity of (Ωt)t(\partial\Omega_{t})_{t} looks like a shrinking circle when magnified. Therefore, the connected component of Ωt\partial\Omega_{t} on which the singularity appears becomes extinct in that singularity. The proof of Theorem 4 can be carried out on each connected component of Ωt\partial\Omega_{t} separately. And none of the connected components have seen singularities in their past. Loosely speaking, when a singularity arises, all problems related to it disappear in that singularity.

4 α\alpha-noncollapsed mean curvature flow without singularities

4.1 Asymptotics to the cylinder

Let (u,Ω)(u,\Omega) be a mean curvature flow without singularities with u0u\geq 0. We note that u(x,t)u(x,t)\to\infty as (x,t)Ω(x,t)\to\partial\Omega. We assume that Ω\Omega is bounded. In particular, any Ωtn\Omega_{t}\subset\mathbb{R}^{n} is bounded and Ωt=\Omega_{t}=\emptyset for all tTt\geq T with some T>0T>0. Furthermore, we suppose that Ω0\Omega_{0} is smooth, mean convex, and α\alpha-noncollapsed for some α>0\alpha>0 (Definition 32). We also assume that u0=u(,0)u_{0}=u(\cdot,0) is smooth, that M0M_{0} has bounded curvature, and that Mt=graphu|ΩtM_{t}=\mathop{\mathrm{graph}}u|_{\Omega_{t}} is mean convex and α\alpha-noncollapsed for all t0t\geq 0 (if MtM_{t}\neq\emptyset).

Our aim in this section is to prove

Theorem 5.

With the above assumptions, if Ωtn\partial\Omega_{t}\subset\mathbb{R}^{n} is a smooth hypersurface for some t>0t>0 then Mthen+1M_{t}-h\,e_{n+1} converges in ClocC_{\textrm{loc}}^{\infty} to the cylinder Ωt×\partial\Omega_{t}\times\mathbb{R} for hh\to\infty.

Like in section 3, establishing a curvature bound is crucial. We exploit the HH-dependent curvature bound from Corollary 35. Firstly, we will use it to demonstrate that the mean curvature HH behaves continuously at infinity and converges to the mean curvature of Ωt\partial\Omega_{t} (Theorem 6). Once we have control on HH, we can directly use the HH-dependent estimates to bound all the terms |kA||\nabla^{k}A|, kk\in\mathbb{N}, for MtM_{t}.

Proof of Theorem 5.

Let t>0t>0 be a time when Ωt\partial\Omega_{t} is smooth. Theorem 6 asserts that H[u](,t)H[u](\cdot,t) is continuous with boundary values H[Ωt]H[\partial\Omega_{t}] (see Theorem 6 for the terminology). The continuity of H[u](,t)H[u](\cdot,t), the compactness of Ω¯t\overline{\Omega}_{t}, and the boundedness of H[Ωt]H[\partial\Omega_{t}] together imply the boundedness of H[u](,t)H[u](\cdot,t) and therefore the boundedness of the mean curvature HH of MtM_{t}.

Because MtM_{t} belongs to an α\alpha-noncollapsed mean curvature flow, Corollary 35 implies that the whole geometry of MtM_{t} is bounded, i.e., all terms |lA||\nabla^{l}A| (l=0,1,l=0,1,\ldots) are bounded. This applies even for times slightly before. The smooth convergence now follows from Lemma 21, the theorem of Arzelà-Ascoli, and interpolation inequalities. ∎

Theorem 6.

Let H[u](,t)H[u](\cdot,t) be the mean curvature of the graph of u(,t)u(\cdot,t) defined as a function on Ωt\Omega_{t}. Then H[u]H[u] is continuously extendable to Ω¯\overline{\Omega} with values in (0,](0,\infty], and there holds H[u](x,t)=H[Ωt](x)H[u](x,t)=H[\partial\Omega_{t}](x) for xΩtx\in\partial\Omega_{t} with H[Ωt]H[\partial\Omega_{t}] the mean curvature of the boundary Ωtn\partial\Omega_{t}\subset\mathbb{R}^{n} (may be infinite).

Proof.

We denote the spacetime track by

(7) t[0,)Mt×{t}n+1×[0,).\mathcal{M}\coloneqq\bigcup_{t\in[0,\infty)}M_{t}\times\{t\}\subset\mathbb{R}^{n+1}\times[0,\infty)\;.

In what follows, unspecified geometric quantities refer to MtM_{t}. Points on MtM_{t} are tracked in time along the normal direction.

We have

(8) |H|=|gijkhij||gij||khij|=n|A||\nabla H|=|g^{ij}\,\nabla_{k}h_{ij}|\leq|g^{ij}|\cdot|\nabla_{k}h_{ij}|=\sqrt{n}\,|\nabla A|

and

(9) |tH|=|ΔH+|A|2H||gij||gkl||ij2hkl|+|A|2|gkl||hkl|=n|2A|+n|A|3.|\partial_{t}H|=\big{|}\Delta H+|A|^{2}\,H\big{|}\leq|g^{ij}|\cdot|g^{kl}|\cdot|\nabla_{ij}^{2}h_{kl}|+|A|^{2}\cdot|g^{kl}|\cdot|h_{kl}|=n\,|\nabla^{2}A|+\sqrt{n}\,|A|^{3}\;.

Thus, by Lemma 7 and these inequalities, |A|2|A_{\mathcal{M}}|^{2} is bounded by terms of the form |lA|2(1+H2)l+1\frac{|\nabla^{l}A|^{2}}{(1+H^{2})^{l+1}} with l=0,1,2l=0,1,2. By virtue of Corollary 35, the curvature of the spacetime track \mathcal{M} is uniformly bounded for tt0>0t\geq t_{0}>0. For times 0<t<t00<t<t_{0}, t0t_{0} sufficiently small, we may use Theorem 4 to see that \mathcal{M} has bounded curvature for these times, too.

By assumption every MtM_{t} is α\alpha-noncollapsed and because MtM_{t} is not a hyperplane we have H>0H>0 everywhere. Because HH is the normal speed, this implies that the spacetime track \mathcal{M} is graphical over n+1\mathbb{R}^{n+1} in the time direction. More precisely, the domain is D{(x,y)n+1:yu0(x)}D\coloneqq\{(x,y)\in\mathbb{R}^{n+1}\colon y\geq u_{0}(x)\}. The representation function is the time of arrival function τ:D\tau\colon D\to\mathbb{R}. By the α\alpha-noncollapsedness of the MtM_{t} and because of the boundedness of interior balls, the mean curvature is uniformly bounded from below, Hc>0H\geq c>0. This implies that |τ||\nabla\tau| is uniformly bounded. Together with the curvature bound for =graphτ\mathcal{M}=\mathop{\mathrm{graph}}\tau, this implies a finite bound for τC2\|\tau\|_{C^{2}}.

The sequence of functions τ(x,y+h)\tau(x,y+h) is monotonically increasing and, by the C2C^{2}-bound, converges for hh\to\infty in Cloc1C_{\mathrm{loc}}^{1} to a C1,1C^{1,1} function σ\sigma defined on Ω0×\Omega_{0}\times\mathbb{R}. It is independent of the yy-variable and represents Ω×\partial\Omega\times\mathbb{R} as a graph.

The mean curvature depends on the gradient of the time of arrival function:

(10) H[u](x,t)=|τ(x,u(x,t))|1.H[u](x,t)=\left|\nabla\tau(x,u(x,t))\right|^{-1}\,.

For (x,t)Ω(x,t)\to\partial\Omega we have u(x,t)u(x,t)\to\infty and therefore

(11) H[u](x,t)=|τ(x,u(x,t))|1|σ(x,0)|1=H[Ωt](x).H[u](x,t)=\left|\nabla\tau(x,u(x,t))\right|^{-1}\to\left|\nabla\sigma(x,0)\right|^{-1}=H[\partial\Omega_{t}](x)\;.\qed
Remark.

From the proof one can easily see that ΩC1,1\partial\Omega\in C^{1,1} holds.

Remark.

It is worth mentioning that, in contrast to |A||A_{\mathcal{M}}|, |A||\nabla A_{\mathcal{M}}| can not be estimated by terms of the form |lA|2(1+H2)l+1\frac{|\nabla^{l}A|^{2}}{(1+H^{2})^{l+1}}. Therefore, we do not get higher estimates for the spacetime track from Corollary 35.

Lemma 7.

The squared norm of the second fundamental form of the spacetime track \mathcal{M} is given by

(12) |A|2=|A|21+H2+2|H|2(1+H2)2+|tH|2(1+H2)3|A_{\mathcal{M}}|^{2}=\frac{|A|^{2}}{1+H^{2}}+2\,\frac{|\nabla H|^{2}}{(1+H^{2})^{2}}+\frac{|\partial_{t}H|^{2}}{(1+H^{2})^{3}}

where the geometric quantities on the right hand side correspond to those of MtM_{t}.

Proof.

For a point (x0,y0,t0)(x_{0},y_{0},t_{0})\in\mathcal{M} (x0nx_{0}\in\mathbb{R}^{n}, y0y_{0}\in\mathbb{R}) we represent \mathcal{M} locally as the graph of a function v(x,t)v(x,t). By a rotation in (x,y)(x,y)-space (n+1\mathbb{R}^{n+1}) we may assume that xv(x0,t0)=0\nabla_{x}v(x_{0},t_{0})=0 holds. Because (Mt)t(M_{t})_{t} is a mean curvature flow, vv solves graphical mean curvature flow, i.e., tv=1+|xv|2H\partial_{t}v=\sqrt{1+|\nabla_{x}v|^{2}}\,H holds. At the point (x0,t0)(x_{0},t_{0}) we obtain tv=H\partial_{t}v=H, xtv=xH\nabla_{x}\partial_{t}v=\nabla_{x}H, and t2v=tH\partial_{t}^{2}v=\partial_{t}H. Before stating computations with these identities at (x0,y0,t0)(x_{0},y_{0},t_{0}), let us briefly fix some notation: Dxf\mathrm{D}_{x}f is a linear form and xf\nabla_{x}f is the gradient of ff, which is given by xf=g1(Dxf)𝖳\nabla_{x}f=g^{-1}\cdot(\mathrm{D}_{x}f)^{\mathsf{T}}.

h\displaystyle h_{\mathcal{M}} =D(x,t)2v1+|(x,t)v|2=11+|(0,H)|2(Dx2v(Dxtv)𝖳Dxtvt2v)\displaystyle=\frac{\mathrm{D}_{(x,t)}^{2}v}{\sqrt{1+|\nabla_{(x,t)}v|^{2}}}=\frac{1}{\sqrt{1+|(0,H)|^{2}}}\begin{pmatrix}\mathrm{D}_{x}^{2}v&(\mathrm{D}_{x}\partial_{t}v)^{\mathsf{T}}\\ \mathrm{D}_{x}\partial_{t}v&\partial_{t}^{2}v\end{pmatrix}
=11+H2(h(DxH)𝖳DxHtH)\displaystyle=\frac{1}{\sqrt{1+H^{2}}}\begin{pmatrix}h&(\mathrm{D}_{x}H)^{\mathsf{T}}\\ \mathrm{D}_{x}H&\partial_{t}H\end{pmatrix}
g1\displaystyle g_{\mathcal{M}}^{-1} =((δαβ)(x,t)v(x,t)v1+|(x,t)v|2)=((δij)001H21+H2)=(g10011+H2)\displaystyle=\left(\left(\delta^{\alpha\beta}\right)-\frac{\nabla_{(x,t)}v\otimes\nabla_{(x,t)}v}{1+|\nabla_{(x,t)}v|^{2}}\right)=\begin{pmatrix}\left(\delta^{ij}\right)&0\\ 0&1-\frac{H^{2}}{1+H^{2}}\end{pmatrix}=\begin{pmatrix}g^{-1}&0\\ 0&\frac{1}{1+H^{2}}\end{pmatrix}
A\displaystyle A_{\mathcal{M}} =g1h=11+H2(g1hg1(DxH)𝖳DxH1+H2tH1+H2)\displaystyle=g_{\mathcal{M}}^{-1}\cdot h_{\mathcal{M}}=\frac{1}{\sqrt{1+H^{2}}}\begin{pmatrix}g^{-1}\cdot h&g^{-1}\cdot(\mathrm{D}_{x}H)^{\mathsf{T}}\\ \frac{\mathrm{D}_{x}H}{1+H^{2}}&\frac{\partial_{t}H}{1+H^{2}}\end{pmatrix}
=11+H2(AxHDxH1+H2tH1+H2)\displaystyle=\frac{1}{\sqrt{1+H^{2}}}\begin{pmatrix}A&\nabla_{x}H\\ \frac{\mathrm{D}_{x}H}{1+H^{2}}&\frac{\partial_{t}H}{1+H^{2}}\end{pmatrix}
|A|2\displaystyle|A_{\mathcal{M}}|^{2} =tr(A2)=11+H2tr(A2+xHDxH1+H2AxH+xHtH1+H2DxH1+H2A+tH1+H2DxH1+H2DxH1+H2xH+(tH1+H2)2)\displaystyle=\textrm{tr}(A_{\mathcal{M}}^{2})=\frac{1}{1+H^{2}}\mathrm{tr}\begin{pmatrix}A^{2}+\nabla_{x}H\cdot\frac{\mathrm{D}_{x}H}{1+H^{2}}&A\cdot\nabla_{x}H+\nabla_{x}H\cdot\frac{\partial_{t}H}{1+H^{2}}\\ \frac{\mathrm{D}_{x}H}{1+H^{2}}\cdot A+\frac{\partial_{t}H}{1+H^{2}}\,\frac{\mathrm{D}_{x}H}{1+H^{2}}&\frac{\mathrm{D}_{x}H}{1+H^{2}}\cdot\nabla_{x}H+\left(\frac{\partial_{t}H}{1+H^{2}}\right)^{2}\end{pmatrix}
=11+H2(|A|2+2|xH|21+H2+(tH)2(1+H2)2).\displaystyle=\frac{1}{1+H^{2}}\left(|A|^{2}+2\,\frac{|\nabla_{x}H|^{2}}{1+H^{2}}+\frac{(\partial_{t}H)^{2}}{(1+H^{2})^{2}}\right)\;.\qed

4.2 Construction of an α\alpha-noncollapsed mean curvature flow without singularities

Under the presumption that Ω0\Omega_{0} is bounded and smooth, H[Ω0]>0H[\partial\Omega_{0}]>0, |AM0|C|A_{M_{0}}|\leq C, and HM0c>0H_{M_{0}}\geq c>0, we will construct a mean curvature flow without singularities that is α\alpha-noncollapsed for some α>0\alpha>0 and matches these data. Theorem 5 from the last section tells us that this flow is smoothly asymptotic to the cylinder Ωt×\partial\Omega_{t}\times\mathbb{R} for all times tt when Ωt\partial\Omega_{t} is smooth.

At first, we describe our approach. We approximate M0M_{0} by closed hypersurfaces. This is done in a way such that these approximating hypersurfaces are α\alpha-noncollapsed. We let flow these by the level-set flow. By the results of [9], the α\alpha-noncollapsedness persists along this flow with the same α\alpha. The level-set flow will in fact be a smooth graphical flow below a certain height. This height tends in the limit of the approximation to infinity. We can then employ a limit process to obtain a mean curvature flow without singularities which is α\alpha-noncollapsed.

Approximation of M0M_{0} by compact M0δM_{0}^{\delta}.

Let δ>0\delta>0 (to be thought of being small). The initial surface M0M_{0} is above a height aδ=max{u0(x):xΩ0,dist(x,Ω0)δ}a_{\delta}=\max\{u_{0}(x)\colon x\in\Omega_{0},\,\mathrm{dist}(x,\partial\Omega_{0})\geq\delta\} in a δ\delta-neighborhood of NΩ0×N\coloneqq\partial\Omega_{0}\times\mathbb{R}. If δ\delta is sufficiently small, then Lemma 21 yields that M0M_{0} is a normal graph over NN above the height aδa_{\delta} (according to our Definition 14 we also speak of a normal graph over NN if we have a graph over a subset of NN). Let vv be the representation function for this normal graph. We note that we always choose the outward pointing normal. But because we define the tubular diffeomorphism in the form (p,d)pdν(p,d)\mapsto p-d\,\nu, the function vv is positive (cf. Definitions 13 and 14). We have v=𝒪(δ)v=\mathcal{O}(\delta), |v|=𝒪(δ)|\nabla v|=\mathcal{O}(\sqrt{\delta}), and |2v|C|\nabla^{2}v|\leq C. By Proposition 15 there hold

(13) gM0\displaystyle g_{M_{0}} =gN+𝒪(δ)\displaystyle=g_{N}+\mathcal{O}(\delta)
(14) hM0\displaystyle h_{M_{0}} =hN+2v+𝒪(δ).\displaystyle=h_{N}+\nabla^{2}v+\mathcal{O}(\delta)\;.

We choose a smooth function λλδ:[0,1]\lambda\equiv\lambda_{\delta}\colon\mathbb{R}\to[0,1] with λ(x)=0\lambda(x)=0 for xaδx\leq a_{\delta} and λ(x)=1\lambda(x)=1 for xaδ+1δx\geq a_{\delta}+\frac{1}{\delta}. We can engineer this function in such a way that on (aδ,aδ+1δ)(a_{\delta},a_{\delta}+\frac{1}{\delta}) we have λ>0\lambda^{\prime}>0 and that |λ|+|λ′′|1/2=𝒪(δ)|\lambda^{\prime}|+|\lambda^{\prime\prime}|^{1/2}=\mathcal{O}(\delta) holds. By slight abuse of notation, λ\lambda becomes a function on NN as λ(xn+1)\lambda(x^{n+1}).

On N{aδ<xn+1aδ+2δ}N\cap\{a_{\delta}<x^{n+1}\leq a_{\delta}+\frac{2}{\delta}\} we define for 0<ε<δ30<\varepsilon<\delta^{3}

(15) w(1λ)v+λε2(aδ+2δxn+1)2.w\coloneqq(1-\lambda)\,v+\lambda\,\frac{\varepsilon}{2}\left(a_{\delta}+\frac{2}{\delta}-x^{n+1}\right)^{2}.

Then we have

(16) en+1w=(1λ)en+1v+λε(xn+1(aδ+2δ))+λ(ε2(aδ+2δxn+1)2v),\displaystyle\begin{split}\nabla_{e_{n+1}}w&=(1-\lambda)\,\nabla_{e_{n+1}}v+\lambda\,\varepsilon\left(x^{n+1}-\left(a_{\delta}+\frac{2}{\delta}\right)\right)\\ &\quad+\lambda^{\prime}\left(\frac{\varepsilon}{2}\left(a_{\delta}+\frac{2}{\delta}-x^{n+1}\right)^{2}-v\right),\end{split}
(17) Ω0w\displaystyle\nabla_{\partial\Omega_{0}}w =(1λ)Ω0v,\displaystyle=(1-\lambda)\,\nabla_{\partial\Omega_{0}}v\;,
(18) Ω02w\displaystyle\nabla_{\partial\Omega_{0}}^{2}w =(1λ)Ω02v,\displaystyle=(1-\lambda)\,\nabla_{\partial\Omega_{0}}^{2}v\;,
(19) Ω0en+1w\displaystyle\nabla_{\partial\Omega_{0}}\nabla_{e_{n+1}}w =en+1Ω0w=(1λ)Ω0en+1vλΩ0v,\displaystyle=\nabla_{e_{n+1}}\nabla_{\partial\Omega_{0}}w=(1-\lambda)\,\nabla_{\partial\Omega_{0}}\nabla_{e_{n+1}}v-\lambda^{\prime}\,\nabla_{\partial\Omega_{0}}v\;,
(20) en+12w=(1λ)en+12v+λε+2λ(ε(xn+1(aδ+2δ))en+1v)+λ′′(ε2(aδ+2δxn+1)2v).\displaystyle\begin{split}\nabla_{e_{n+1}}^{2}w&=(1-\lambda)\,\nabla_{e_{n+1}}^{2}v+\lambda\,\varepsilon+2\,\lambda^{\prime}\left(\varepsilon\left(x^{n+1}-\left(a_{\delta}+\frac{2}{\delta}\right)\right)-\nabla_{e_{n+1}}v\right)\\ &\quad+\lambda^{\prime\prime}\left(\frac{\varepsilon}{2}\left(a_{\delta}+\frac{2}{\delta}-x^{n+1}\right)^{2}-v\right)\;.\end{split}

If 0<ε<δ30<\varepsilon<\delta^{3} is sufficiently small (depending on vv), we have 0<w<v0<w<v, w=(1λ)v+𝒪(δ)\nabla w=(1-\lambda)\nabla v+\mathcal{O}(\delta), and 2w=(1λ)2v+𝒪(δ)\nabla^{2}w=(1-\lambda)\,\nabla^{2}v+\mathcal{O}(\delta) on N{aδ<xn+1aδ+2δ}N\cap\{a_{\delta}<x^{n+1}\leq a_{\delta}+\frac{2}{\delta}\}.

Let WW be the hypersurface that is given as the normal graph over NN with representation function ww. Then, again by Proposition 15

(21) gW\displaystyle g_{W} =gN+𝒪(δ)\displaystyle=g_{N}+\mathcal{O}(\delta)
(22) hW=hN+2w+𝒪(δ)=hN+(1λ)2v+𝒪(δ)=(1λ)hM0+λhN+𝒪(δ).\displaystyle\begin{split}h_{W}&=h_{N}+\nabla^{2}w+\mathcal{O}(\delta)=h_{N}+(1-\lambda)\,\nabla^{2}v+\mathcal{O}(\delta)\\ &=(1-\lambda)\,h_{M_{0}}+\lambda\,h_{N}+\mathcal{O}(\delta)\;.\end{split}

Therefore, if we adjust the constants a little bit and choose δ\delta sufficiently small, there hold |AW|2C|A_{W}|^{2}\leq C and HWc>0H_{W}\geq c>0.

The compact approximation M0δM_{0}^{\delta} to M0M_{0} is now constructed as follows: In the region {xn+1aδ}\{x^{n+1}\leq a_{\delta}\}, it coincides with M0M_{0}. In the region {aδ<xn+1aδ+2δ}\{a_{\delta}<x^{n+1}\leq a_{\delta}+\frac{2}{\delta}\}, we set it equal to WW. From this part we obtain the part above the height aδ+2δa_{\delta}+\frac{2}{\delta} by reflection at the hyperplane {xn+1=aδ+2δ}\{x^{n+1}=a_{\delta}+\frac{2}{\delta}\}. By construction, M0δM_{0}^{\delta} is mirror-symmetrical. In fact, its interior is a vain set if ε\varepsilon is chosen sufficiently small. This stems from the fact that then en+1w<0\nabla_{e_{n+1}}w<0 holds, which one can check from (16). Of course, M0δM_{0}^{\delta} is smooth and we have |AM0δ|2C|A_{M_{0}^{\delta}}|^{2}\leq C and HM0δc>0H_{M_{0}^{\delta}}\geq c>0.

Uniform α\alpha-noncollapsedness of the M0δM_{0}^{\delta}.

Lemma 8.

For C>0C>0, c>0c>0, and r>0r>0 there is α>0\alpha>0 such that: If a hypersurface MM has bounded second fundamental form |A|C|A|\leq C, mean curvature HcH\geq c, and admits local graph representations of radius rr, then MM is α\alpha-noncollapsed.

Proof.

Let xMx\in M. Then, inside Br(x)B_{r}(x), MM is a graph over TxM\textrm{T}_{x}M. Let ρmin{r2,1C}\rho\coloneqq\min\{\frac{r}{2},\,\frac{1}{C}\}. Then, the two closed balls B¯ρ(x±ρν(x))\overline{B}_{\rho}(x\pm\rho\,\nu(x)) intersect MM only in xx. Thus, if we choose αcρ\alpha\coloneqq c\,\rho, xx admits interior and exterior balls of radius αH(x)αc=ρ\frac{\alpha}{H(x)}\leq\frac{\alpha}{c}=\rho. Because α\alpha does not depend on xx, this demonstrates that MM is α\alpha-noncollapsed. ∎

To apply the lemma it remains to show that the M0δM_{0}^{\delta} admit local graph representations of a radius rr that is independent of δ\delta. To this end, we realize that the M0δM_{0}^{\delta} are normal graphs over Ω0×\partial\Omega_{0}\times\mathbb{R} for the most part, with two “M0M_{0}-caps.” By Proposition 18 the normal graph part of each M0δM_{0}^{\delta} admits local graph representations of a controlled radius independent of δ\delta. Therefore, Lemma 8 is applicable and yields an α>0\alpha>0 such that the M0δM_{0}^{\delta} are α\alpha-noncollapsed.

Evolution of the M0δM_{0}^{\delta} and passing to a limit.

So far, we have constructed smooth closed hypersurfaces M0δM_{0}^{\delta} such that

  • M0δM_{0}^{\delta} coincides with M0M_{0} below the height aδa_{\delta},

  • the M0δM_{0}^{\delta} are mean convex and α\alpha-noncollapsed with α>0\alpha>0 independent of δ\delta,

  • The M0δM_{0}^{\delta} are symmetric double graphs with respect to the hyperplanes {x:xn+1=aδ+2δ}\{x:x^{n+1}=a_{\delta}+\frac{2}{\delta}\} in the sense of [18].

We can flow M0δM_{0}^{\delta} by the mean curvature flow as in [18]: Singularities only occur on the hyperplane {x:xn+1=aδ+2δ}\{x:x^{n+1}=a_{\delta}+\frac{2}{\delta}\} and the surfaces stay symmetric double graphs. In particular, the part below the hyperplane is smooth and graphical.

By Proposition 29, the weak solution is unique. In particular, (Mtδ)t(M_{t}^{\delta})_{t} coincides with the level-set flow. By Theorem 33, the level-set flow is α\alpha-noncollapsed for all t0t\geq 0 with the same α\alpha as for M0δM_{0}^{\delta}. Hence, (Mtδ)t[0,)\left(M_{t}^{\delta}\right)_{t\in[0,\infty)} is α\alpha-noncollapsed.

We now may use the a priori estimates and the Arzelà-Ascoli argument as employed before in Section 2 to pass to a limit as we let δ0\delta\to 0. The limit is a smooth mean curvature flow (Mt)t(M_{t})_{t} starting from M0M_{0} that is graphical and provides us with a mean curvature flow without singularities (u,Ω)(u,\Omega) with initial value (u0,Ω0)(u_{0},\Omega_{0}). Clearly, the approximation outlined in Section 2 would have given us this result as well. But importantly, the uniform α\alpha-noncollapsedness of the approximators carries over to the limit because MtδMtM_{t}^{\delta}\to M_{t} locally smoothly, proving that MtM_{t} is α\alpha-noncollapsed.

Thus, we have proven the following theorem.

Theorem 9.

Let Ω0n\Omega_{0}\subset\mathbb{R}^{n} be a bounded, smooth, and mean convex domain. Let u0:Ω0u_{0}\colon\Omega_{0}\to\mathbb{R} be a positive and smooth function such that u0(x)u_{0}(x)\to\infty for xΩ0x\to\partial\Omega_{0}. Let M0graphu0M_{0}\coloneqq\mathop{\mathrm{graph}}u_{0}. We assume that there are constants C,c>0C,c>0 such that |AM0|C|A_{M_{0}}|\leq C and HM0cH_{M_{0}}\geq c. Then there exists a mean curvature flow without singularities (u,Ω)(u,\Omega) with initial values (u0,Ω0)(u_{0},\Omega_{0}) such that (Mt)t[0,)(M_{t})_{t\in[0,\infty)} is α\alpha-noncollapsed for some α>0\alpha>0.

5 Barrier over annuli

In the present section we assume n2n\geq 2.

Central to this section is a barrier which is defined over annuli. This barrier enables us to prove estimates for a mean curvature flow in terms of the height over an annulus earlier in time. This will be exploited to obtain various results.

Refer to caption
Figure 4: Sketch of the barrier.

The upper barrier has the following geometry (Fig. 4.1). Initially start with an annulus in nn-dimensional Euclidean space. Over this, we consider a special function which tends to infinity at the boundary of the annulus. As time goes by, the initial annulus shrinks by mean curvature flow and the function is adjusted accordingly to have the shrinking annulus as its domain; while at the same time, the function is shifted upwards.

In the first section, we will write down the barrier and prove that it really functions as a barrier. The upshot is Theorem 11. Afterwards, in the second section, we will utilize the barrier to construct an “ugly” example of mean curvature flow without singularities with wild oscillations which persist up to the vanishing time. As another application of the barrier, we will prove in the third section a certain relationship between the spatial asymptotics of a mean curvature flow without singularities and its temporal asymptotics at the vanishing time.

5.1 Construction of the barrier

The barrier construction starts with a hypersurface that is obtained from a grim reaper curve which is rotated around the xn+1x^{n+1}-axis. The grim reaper is a well-known special solution of curve-shortening flow (one-dimensional mean curvature flow). It has the explicit graphical representation (λ>0\lambda>0)

(23) ugr(x,t)=1λlogcos(λx)+λtfor x(π2λ,π2λ),t.u_{\mathrm{gr}}(x,t)=-\frac{1}{\lambda}\log\cos(\lambda\,x)+\lambda\,t\qquad\text{for }x\in\left(-\frac{\pi}{2\lambda},\frac{\pi}{2\lambda}\right),\;t\in\mathbb{R}\;.

In this form graphugr\mathop{\mathrm{graph}}u_{\mathrm{gr}} translates with speed λ\lambda upwards.

Our barrier construction is motivated by this solution and initially we start with a grim reaper curve rotated around a vertical axis. Let us have a look at the barrier function now:

(24) w(x,t)=1λlogcos[λ(|x|2+2(n1)tR)]+λt+f(t),w(x,t)=-\frac{1}{\lambda}\log\cos\left[\lambda\left(\sqrt{|x|^{2}+2(n-1)t}-R\right)\right]+\lambda\,t+f(t)\;,

where R,λ>0R,\lambda>0. As can be seen, ww is similar to ugru_{\mathrm{gr}}. But xx is replaced by the term |x|2+2(n1)t\sqrt{|x|^{2}+2(n-1)t} which is motivated by the radius of an (n1)(n-1)-dimensional sphere flowing by its mean curvature. The function f(t)f(t) will be adequately chosen to make ww satisfy the correct differential inequality. We will give two possible choices for ff, one rather simple, the other more elaborate but much smaller as λ\lambda becomes large. However, we include the second one only for the sake of completeness because in our applications λt\lambda\,t dominates f(t)f(t) for large λ\lambda in either case.

Before we start proving the differential inequality, we must talk about the domain of definition of ww. We set

(25) tI(R)[0,T(R)][0,R22(n1)]\displaystyle t\in I(R)\coloneqq[0,T(R)]\coloneqq\left[0,\frac{R^{2}}{2(n-1)}\right]
(26) xAt(R,λ){xn:(Rπ2λ)<|x|2+2(n1)t<(R+π2λ)}.\displaystyle x\in A_{t}(R,\lambda)\coloneqq\left\{x\in\mathbb{R}^{n}\colon\left(R-\frac{\pi}{2\lambda}\right)<\sqrt{|x|^{2}+2(n-1)t}<\left(R+\frac{\pi}{2\lambda}\right)\right\}.

At(R,λ)A_{t}(R,\lambda) is an annulus or a ball, depending on tt, RR, and λ\lambda.

Differential inequality.

To keep the computations accessible, we will make the abbreviations \sqrt{\rule{0.0pt}{4.30554pt}\quad} for |x|2+2(n1)t\sqrt{|x|^{2}+2(n-1)t} and [][\;\cdot\;] for λ(R)\lambda\left(\sqrt{\rule{0.0pt}{4.30554pt}\quad}-R\right).

To show that ww is an upper barrier, we must prove

(27) w˙+(δijwiwj1+|w|2)wij!0.-\dot{w}+\left(\delta^{ij}-\frac{w^{i}w^{j}}{1+|\nabla w|^{2}}\right)w_{ij}\stackrel{{\scriptstyle!}}{{\leq}}0\;.

First, we determine the derivatives.

(28) w˙\displaystyle\dot{w} =tan[]n1+λ+f(t),\displaystyle=\tan[\;\cdot\;]\,\frac{n-1}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}}+\lambda+f^{\prime}(t)\;,
(29) wi\displaystyle w_{i} =tan[]xi,\displaystyle=\tan[\;\cdot\;]\,\frac{x_{i}}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}}\;,
(30) wij\displaystyle w_{ij} =λ(1+tan2[])xixj2+tan[](δijxixj2).\displaystyle=\lambda\left(1+\tan^{2}[\;\cdot\;]\right)\frac{x_{i}\,x_{j}}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}}+\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}}\left(\delta_{ij}-\frac{x_{i}\,x_{j}}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}}\right).

Substituting into (27) yields

(31) w˙+(δijwiwj1+|w|2)wij=w˙+11+|w|2((1+|w|2)δijwiwj)wij=λf(t)tan[]n1+λ(1+tan2[])1+tan2[]|x|22((1+tan2[]|x|22)|x|22tan2[]|x|44)+tan[]1+tan2[]|x|22((1+tan2[]|x|22)(n|x|22)tan2[](|x|22|x|44))=λf(t)tan[](n1)+λ(1+tan2[])1+tan2[]|x|22(|x|22)+tan[]1+tan2[]|x|22((1+tan2[]|x|22)(n1)+(1|x|22))=f(t)+(1|x|22)1+tan2[]|x|22(tan[]λ)!0.\begin{split}&-\dot{w}+\left(\delta^{ij}-\frac{w^{i}w^{j}}{1+|\nabla w|^{2}}\right)w_{ij}\\ &\quad=-\dot{w}+\frac{1}{1+|\nabla w|^{2}}\bigg{(}(1+|\nabla w|^{2})\delta^{ij}-w^{i}w^{j}\bigg{)}w_{ij}\\ &\quad=-\lambda-f^{\prime}(t)-\tan[\;\cdot\;]\,\frac{n-1}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}}\\ &\qquad+\frac{\lambda\,(1+\tan^{2}[\;\cdot\;])}{1+\tan^{2}[\;\cdot\;]\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}^{2}}}\left(\left(1+\tan^{2}[\;\cdot\;]\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}}\right)\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}}-\tan^{2}[\;\cdot\;]\frac{|x|^{4}}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{4}}\right)\\ &\qquad+\frac{\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}}}{1+\tan^{2}[\;\cdot\;]\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}^{2}}}\left(\left(1+\tan^{2}[\;\cdot\;]\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}}\right)\left(n-\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}}\right)\right.\\ &\hskip 128.0374pt\left.-\tan^{2}[\;\cdot\;]\left(\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}}-\frac{|x|^{4}}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{4}}\right)\right)\\ &\quad=-\lambda-f^{\prime}(t)-\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}}(n-1)+\frac{\lambda\,(1+\tan^{2}[\;\cdot\;])}{1+\tan^{2}[\;\cdot\;]\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}^{2}}}\left(\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}}\right)\\ &\qquad+\frac{\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}}}{1+\tan^{2}[\;\cdot\;]\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}^{2}}}\left(\left(1+\tan^{2}[\;\cdot\;]\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}}\right)(n-1)+\left(1-\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}}\right)\right)\\ &\quad=-f^{\prime}(t)+\frac{\left(1-\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}^{2}}\right)}{1+\tan^{2}[\;\cdot\;]\frac{|x|^{2}}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}^{2}}}\left(\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}}-\lambda\right)\stackrel{{\scriptstyle!}}{{\leq}}0\;.\end{split}

In the case tan[]λ\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}}\leq\lambda, the inequality clearly holds. Thus, let us from now on assume tan[]>λ\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}}>\lambda.

Neglecting terms with the appropriate signs, we aim to show

(32) tan[]1+tan2[]2|x|2!f(t).\frac{\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}}}{1+\frac{\tan^{2}[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}^{2}}|x|^{2}}\stackrel{{\scriptstyle!}}{{\leq}}f^{\prime}(t)\;.

If we can choose ff in such a way then the differential inequality (27) follows.

We are going to utilize the following lemma.

Lemma 10.

Let J>0J\subset\mathbb{R}_{>0} be an interval and y:J>0y\colon J\to\mathbb{R}_{>0} a monotonically increasing function, which is continuous and surjective. Then

(33) y(r)1+y2(r)r21a\frac{y(r)}{1+y^{2}(r)\,r^{2}}\leq\frac{1}{a}

holds for all rJr\in J, where aa is the unique solution of y(a)=1ay(a)=\frac{1}{a}.

Proof.

It is not hard to see that there is a unique solution aa of y(a)=1ay(a)=\frac{1}{a} given the hypothesis on yy.

We distinguish two cases. If rar\leq a holds, we find

(34) y(r)1+y2(r)r2y(r)y(a)=1a.\frac{y(r)}{1+y^{2}(r)\,r^{2}}\leq y(r)\leq y(a)=\frac{1}{a}\;.

In the case that r>ar>a holds, we have

(35) y(r)1+y2(r)r2y(r)1+y2(r)a2.\frac{y(r)}{1+y^{2}(r)\,r^{2}}\leq\frac{y(r)}{1+y^{2}(r)\,a^{2}}\;.

The expression y1+y2a2\frac{y}{1+y^{2}a^{2}} tends to 0 for both y0y\to 0 and yy\to\infty. Hence, it attains its maximum at an interior point bb. We can determine bb from the extremality condition

(36) 0=ddy(y1+y2a2)|y=b=1+b2a22b2a2(1+b2a2)2=1b2a2(1+b2a2)2.0=\left.\frac{\mathrm{d}}{\mathrm{d}y}\left(\frac{y}{1+y^{2}a^{2}}\right)\right|_{y=b}=\frac{1+b^{2}a^{2}-2b^{2}a^{2}}{(1+b^{2}a^{2})^{2}}=\frac{1-b^{2}a^{2}}{(1+b^{2}a^{2})^{2}}\;.

We infer that b=1ab=\frac{1}{a} is the maximum point. That leads to

(37) y1+y2a2b1+b2a2=1/a1+1<1a\frac{y}{1+y^{2}a^{2}}\leq\frac{b}{1+b^{2}a^{2}}=\frac{1/a}{1+1}<\frac{1}{a}

for any y>0y\in\mathbb{R}_{>0}. Bringing together (35) and (37), the assertion follows. ∎

Application of the lemma.

We shall derive (32) with the help of Lemma 10. Clearly, tan[]\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}}, seen as a function of r|x|r\coloneqq|x|, will take the role of y(r)y(r). Notice that we fix tI(R)t\in I(R) here. The interval JJ is appropriately chosen to be

(38) J=(R22(n1)t,(R+π2λ)22(n1)t),J=\left(\sqrt{R^{2}-2(n-1)t},\sqrt{\left(R+\frac{\pi}{2\lambda}\right)^{2}-2(n-1)t}\right),

such that []λ(R)(0,π2)[\;\cdot\;]\coloneqq\lambda(\sqrt{\rule{0.0pt}{4.30554pt}\quad}-R)\in(0,\frac{\pi}{2}). This makes y(r)=tan[]y(r)=\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}} well defined on JJ and surjective onto >0\mathbb{R}_{>0}.

Now we show, that our choice of yy is monotonically increasing. The derivative with respect to \sqrt{\rule{0.0pt}{4.30554pt}\quad} is

(39) ddtan[λ(R)]=λcos2[]tan[]2=λsin[]cos[]2cos2[]λ[]2cos2[]=λR2cos2[]>0.\begin{split}\frac{\mathrm{d}}{\mathrm{d}\sqrt{\rule{0.0pt}{4.30554pt}\quad}}\frac{\tan[\lambda(\sqrt{\rule{0.0pt}{4.30554pt}\quad}-R)]}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}}&=\frac{\lambda}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}\,\cos^{2}[\;\cdot\;]}-\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}}=\frac{\lambda\sqrt{\rule{0.0pt}{4.30554pt}\quad}-\sin[\;\cdot\;]\cos[\;\cdot\;]}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}\cos^{2}[\;\cdot\;]}\\ &\geq\frac{\lambda\sqrt{\rule{0.0pt}{4.30554pt}\quad}-[\;\cdot\;]}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}\cos^{2}[\;\cdot\;]}=\frac{\lambda\,R}{\sqrt{\rule{0.0pt}{4.30554pt}\quad}^{2}\cos^{2}[\;\cdot\;]}>0\;.\end{split}

So tan[]\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}} is increasing as a function of \sqrt{\rule{0.0pt}{4.30554pt}\quad}. And because \sqrt{\rule{0.0pt}{4.30554pt}\quad} is increasing in rr, y(r)=tan[]y(r)=\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}} is increasing in rr as well.

Finally, Lemma 10 is applicable and yields

(40) tan[]1+tan2[]2|x|21at,\frac{\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}}}{1+\frac{\tan^{2}[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}^{2}}\,|x|^{2}}\leq\frac{1}{a_{t}}\;,

where ata_{t} is the unique solution in JJ of the equation

(41) tan[λ(at2+2(n1)tR)]at2+2(n1)t=1at.\frac{\tan\left[\lambda\left(\sqrt{a_{t}^{2}+2(n-1)t}-R\right)\right]}{\sqrt{a_{t}^{2}+2(n-1)t}}=\frac{1}{a_{t}}\;.

To further estimate 1at\frac{1}{a_{t}} in (40), we must extract information from (41). We will now demonstrate two possible ways to do so.

First choice for ff.

We notice that

(42) tan(λ(at2+2(n1)tR))=at2+2(n1)tat1=tan(π4).\tan\left(\lambda\left(\sqrt{a_{t}^{2}+2(n-1)t}-R\right)\right)=\frac{\sqrt{a_{t}^{2}+2(n-1)t}}{a_{t}}\geq 1=\tan\left(\frac{\pi}{4}\right)\,.

Consequently, and by 2(n1)tR22(n-1)t\leq R^{2},

(43) at2(π4λ+R)22(n1)tπR2λ.a_{t}^{2}\geq\left(\frac{\pi}{4\lambda}+R\right)^{2}-2(n-1)t\geq\frac{\pi R}{2\lambda}\;.

Choosing

(44) f(t)2λπRt\boxed{f(t)\coloneqq\sqrt{\frac{2\lambda}{\pi R}}\;t}

we infer from (40) and (43)

(45) tan[]1+tan2[]2|x|21at2λπR=f(t).\frac{\frac{\tan[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}}}{1+\frac{\tan^{2}[\;\cdot\;]}{\sqrt{\rule{0.0pt}{3.01389pt}\quad}^{2}}\,|x|^{2}}\leq\frac{1}{a_{t}}\leq\sqrt{\frac{2\lambda}{\pi R}}=f^{\prime}(t)\;.

I.e., (32), and hence (27) hold.

Second choice for ff.

We write

(46) sat2+2(n1)t.s\coloneqq\sqrt{a_{t}^{2}+2(n-1)t}\;.

We note that ss depends on tt and lies in the range [R+π4λ,R+π2λ)\left[R+\tfrac{\pi}{4\lambda},R+\tfrac{\pi}{2\lambda}\right), which can be seen from (42). With ss at hand we can rewrite (41) as

(47) 1at=1s22(n1)t=tan(λ(sR))s.\frac{1}{a_{t}}=\frac{1}{\sqrt{s^{2}-2(n-1)t}}=\frac{\tan(\lambda\,(s-R))}{s}\;.

This can easily be solved for 2(n1)t2(n-1)t. We write τ(s)\tau(s) for 2(n1)t2(n-1)t viewed as a function of ss:

(48) τ(s)s2(11tan2(λ(sR))),s[R+π4λ,R+π2λ).\tau(s)\coloneqq s^{2}\left(1-\frac{1}{\tan^{2}(\lambda(s-R))}\right),\qquad s\in\left[R+\tfrac{\pi}{4\lambda},\,R+\tfrac{\pi}{2\lambda}\right)\;.

We continuously extend τ\tau to the closed interval through τ(R+π2λ)=(R+π2λ)2\tau(R+\frac{\pi}{2\lambda})=(R+\frac{\pi}{2\lambda})^{2}. We will estimate τ(s)\tau(s) from below by

(49) σ(s)(14λ(R+π2λ))s2+4λs3;\sigma(s)\coloneqq\left(1-4\lambda\left(R+\frac{\pi}{2\lambda}\right)\right)s^{2}+4\lambda s^{3}\;;

σ\sigma is the solution of σ(s)=2sσ(s)+4λs2\sigma^{\prime}(s)=\frac{2}{s}\sigma(s)+4\lambda s^{2} with σ(R+π2λ)=τ(R+π2λ)=(R+π2λ)2\sigma\left(R+\tfrac{\pi}{2\lambda}\right)=\tau\left(R+\tfrac{\pi}{2\lambda}\right)=\left(R+\tfrac{\pi}{2\lambda}\right)^{2}. The function τ\tau fulfills the differential inequality

(50) τ(s)=2s(11tan2(λ(sR)))+2λs21+tan2(λ(sR))tan3(λ(sR))2sτ(s)+4λs2.\begin{split}\tau^{\prime}(s)&=2s\left(1-\frac{1}{\tan^{2}(\lambda\,(s-R))}\right)+2\lambda\,s^{2}\frac{1+\tan^{2}(\lambda\,(s-R))}{\tan^{3}(\lambda\,(s-R))}\\ &\leq\frac{2}{s}\tau(s)+4\lambda\,s^{2}\;.\end{split}

It follows that

(51) dds(τ(s)σ(s))2sτ(s)+4λs2(2sσ(s)+4λs2)=2s(τ(s)σ(s)).\frac{\mathrm{d}}{\mathrm{d}s}(\tau(s)-\sigma(s))\leq\frac{2}{s}\tau(s)+4\lambda\,s^{2}-\left(\frac{2}{s}\sigma(s)+4\lambda\,s^{2}\right)=\frac{2}{s}(\tau(s)-\sigma(s))\;.

In fact the inequality is strict except for s=R+π4λs=R+\frac{\pi}{4\lambda}. Thus, whenever τ(s)=σ(s)\tau(s)=\sigma(s) for some s(R+π4λ,R+π2λ)s\in(R+\frac{\pi}{4\lambda},R+\frac{\pi}{2\lambda}), we have τ(s)<σ(s)\tau^{\prime}(s)<\sigma^{\prime}(s) and there is an ε>0\varepsilon>0 such that τ>σ\tau>\sigma on (sε,s)(s-\varepsilon,s). Considering sup{s:τ(s)<σ(s)}\sup\{s\colon\tau(s)<\sigma(s)\}, it is now easy to show that τσ\tau\geq\sigma on the whole interval [R+π4λ,R+π2λ]\left[R+\frac{\pi}{4\lambda},\,R+\frac{\pi}{2\lambda}\right]. So we have shown for s[R+π4λ,R+π2λ]s\in\left[R+\frac{\pi}{4\lambda},\,R+\frac{\pi}{2\lambda}\right]

(52) τ(s)\displaystyle\tau(s) (14λ(R+π2λ))s2+4λs3,\displaystyle\geq\left(1-4\lambda\left(R+\frac{\pi}{2\lambda}\right)\right)s^{2}+4\lambda s^{3}\;,
(53) τ(s)(R+π4λ)2\displaystyle\frac{\tau(s)}{\left(R+\frac{\pi}{4\lambda}\right)^{2}} τ(s)s214λ(R+π2λ)+4λs,\displaystyle\geq\frac{\tau(s)}{s^{2}}\geq 1-4\lambda\left(R+\frac{\pi}{2\lambda}\right)+4\lambda\,s\;,
(54) sR+π2λ14λ(1τ(s)(R+π4λ)2)=R+1λ(π214(1τ(s)(R+π4λ)2)).\displaystyle\begin{split}s&\leq R+\frac{\pi}{2\lambda}-\frac{1}{4\lambda}\left(1-\frac{\tau(s)}{\left(R+\frac{\pi}{4\lambda}\right)^{2}}\right)\\ &=R+\frac{1}{\lambda}\left(\frac{\pi}{2}-\frac{1}{4}\left(1-\frac{\tau(s)}{\left(R+\frac{\pi}{4\lambda}\right)^{2}}\right)\right).\end{split}

Let us return to (46), the value for ss we are interested in. Let us also remember that τ(s)=2(n1)t\tau(s)=2(n-1)t holds in this context. With sR+π4λs\geq R+\frac{\pi}{4\lambda} and (54) we can continue from (47)

(55) 1attan(λ(sR))s1R+π4λtan(π214(12(n1)t(R+π4λ)2)).\frac{1}{a_{t}}\leq\frac{\tan(\lambda(s-R))}{s}\leq\frac{1}{R+\frac{\pi}{4\lambda}}\,\tan\left(\frac{\pi}{2}-\frac{1}{4}\left(1-\frac{2(n-1)t}{\left(R+\frac{\pi}{4\lambda}\right)^{2}}\right)\right).

We integrate the last term:

(56) f(t)2(R+π4λ)n1[logsin(14(12(n1)t(R+π4λ)2))+logsin(14)].\boxed{f(t)\coloneqq\frac{2\left(R+\frac{\pi}{4\lambda}\right)}{n-1}\left[-\log\sin\left(\frac{1}{4}\left(1-\frac{2(n-1)t}{\left(R+\frac{\pi}{4\lambda}\right)^{2}}\right)\right)+\log\sin\left(\frac{1}{4}\right)\right].}

Then we obtain (32) from (55) and (40). So the differential inequality (27) for ww follows with this choice for ff.

Asymptotics of ff for large λ\lambda.

The merit of the second choice for ff is the milder growth in λ\lambda. While (44) clearly grows like λ\sqrt{\lambda}, (56) only grows like log(λ)\log(\lambda). To see this let us evaluate ff at the final time T(R)T(R) of the time interval under consideration (T(R)R22(n1)T(R)\coloneqq\frac{R^{2}}{2(n-1)}). Let f1f_{1} be the first choice, (44), and let f2f_{2} be the second, (56). Then we have

(57) f1(T(R))=2λπRR22(n1)=R4(n1)8Rλπ.f_{1}(T(R))=\sqrt{\frac{2\lambda}{\pi\,R}}\;\frac{R^{2}}{2(n-1)}=\frac{R}{4(n-1)}\sqrt{\frac{8R\,\lambda}{\pi}}\;.

On the other hand, for λ\lambda\to\infty:

(58) f2(T(R))=2(R+π4λ)n1[logsin(14(R+π4λ)2R2(R+π4λ)2)+logsin(14)]2Rn1logsin(14π2λRR2)2Rn1log(π8Rλ)=2Rn1log(8Rλπ).\begin{split}f_{2}(T(R))&=\frac{2\left(R+\frac{\pi}{4\lambda}\right)}{n-1}\left[-\log\sin\left(\frac{1}{4}\frac{\left(R+\frac{\pi}{4\lambda}\right)^{2}-R^{2}}{\left(R+\frac{\pi}{4\lambda}\right)^{2}}\right)+\log\sin(\frac{1}{4})\right]\\ &\simeq-\frac{2R}{n-1}\;\log\sin\left(\frac{1}{4}\frac{\frac{\pi}{2\lambda}R}{R^{2}}\right)\\ &\simeq-\frac{2R}{n-1}\;\log\left(\frac{\pi}{8R\,\lambda}\right)\\ &=\frac{2R}{n-1}\;\log\left(\frac{8R\,\lambda}{\pi}\right)\;.\end{split}

Barrier property and annulus dependent estimate.

So far we have proven that for any choice of ff ((44) or (56)) ww fulfills the differential inequality (27) for all tI(R)t\in I(R) and xAt(R,λ){0}x\in A_{t}(R,\lambda)\setminus\{0\}. For this reason, a mean curvature flow cannot touch graphw(,t)\mathop{\mathrm{graph}}w(\cdot,t) from below at any interior point if the flow is disjoint initially. A special case, however, is the point (0,w(0,t))(0,w(0,t)) (if defined) because ww does not solve (27) at x=0x=0. In fact, ww is not even differentiable at x=0x=0. But still, no smooth surface can touch graphw\mathop{\mathrm{graph}}w from below at (0,w(0,t))(0,w(0,t)): All of the directional derivatives of ww at zero are defined and they all are strictly negative, and hence there is not even a smooth curve, let alone a hypersurface, that can touch ww from below at that point.

It remains to exclude that graphw(,t)\mathop{\mathrm{graph}}w(\cdot,t) is crossed at infinity. Then graphw(,t)\mathop{\mathrm{graph}}w(\cdot,t) is a barrier and we obtain the following annulus dependent estimate.

Theorem 11.

Let R,λ>0R,\lambda>0. Recall the definition of At(R,λ)A_{t}(R,\lambda) in (26). Let (u,Ω)(u,\Omega) be a solution of mean curvature flow without singularities in n+1\mathbb{R}^{n+1} (n2n\geq 2). Suppose A0(R,λ)Ω0A_{0}(R,\lambda)\Subset\Omega_{0} and

(59) u(x,0)1λlogcos[λ(|x|R)] for all xA0(R,λ).u(x,0)\leq-\frac{1}{\lambda}\log\cos\big{[}\lambda(|x|-R)\big{]}\qquad\text{ for all }x\in A_{0}(R,\lambda)\;.

Then for tI(R)=[0,R22(n1)]t\in I(R)=[0,\frac{R^{2}}{2(n-1)}] there holds At(R,λ)ΩtA_{t}(R,\lambda)\Subset\Omega_{t} and

(60) u(x,t)w(x,t)=1λlogcos[λ(|x|2+2(n1)tR)]+λt+f(t)u(x,t)\leq w(x,t)=-\frac{1}{\lambda}\log\cos\left[\lambda\left(\sqrt{|x|^{2}+2(n-1)t}-R\right)\right]+\lambda\,t+f(t)

holds for all xAt(R,λ).x\in A_{t}(R,\lambda). The expression f(t)f(t) can be given by (44) or (56) depending on your preferences.

Proof.

From [17] we know that At(R,λ)ΩtA_{t}(R,\lambda)\Subset\Omega_{t} for all t>0t>0, because the boundary of At(R,λ)A_{t}(R,\lambda) moves by mean curvature flow. In particular, graphu(,t)\mathop{\mathrm{graph}}u(\cdot,t) will not touch graphw(,t)\mathop{\mathrm{graph}}w(\cdot,t) at infinity. As we have demonstrated above, there will be no touching in the interior, either. Thus, (60) holds and the theorem is proven. ∎

Clearly w-w resembles a subsolution, and consequently, graphw(,t)\mathop{\mathrm{graph}}-w(\cdot,t) is a lower barrier which cannot be touched from above by any mean curvature flow.

5.2 Example with wild oscillations

We can use Theorem 11 to construct an example of mean curvature flow without singularities which behaves quite badly. The associated domain Ω\Omega simply describes a shrinking ball and the associated function uu converges to ++\infty at Ω\partial\Omega. So the graphical surface Mt=graphu(,t)M_{t}=\mathop{\mathrm{graph}}u(\cdot,t) vanishes at ++\infty the moment when Ωt\Omega_{t} shrinks to a point. Special about the surface is that it does not vanish monotonically to infinity. Instead, u(0,t)u(0,t) increasingly oscillates as tt approaches the final time. The surface MtM_{t} has unbounded curvature at any time, a behavior that is not mirrored by the domain Ωt\Omega_{t}. In fact, as we go upwards at a fixed time tt, MtM_{t} has to get closer and closer to Ωt×\partial\Omega_{t}\times\mathbb{R}. However, MtM_{t} is not smoothly asymptotic to Ωt×\partial\Omega_{t}\times\mathbb{R} but MtM_{t} approaches Ωt×\Omega_{t}\times\mathbb{R} with more and more sheets. Figure 5 is an attempt to depict MtM_{t}. For simplicity we will only consider the case n=2n=2, although it is straightforward to generalize to higher nn.

Refer to caption
Figure 5: Wildly Oscillating Mean Curvature Flow Without Singularities.

For all kk\in\mathbb{N} with k2k\geq 2 we define the Intervals Ik(k2k1,k1k)I_{k}\coloneqq\left(\frac{k-2}{k-1},\frac{k-1}{k}\right). The length of IkI_{k} is given by 1(k1)k\frac{1}{(k-1)k}. Let λk(k1)kπ\lambda_{k}\coloneqq(k-1)k\,\pi and let RkR_{k} be the centers of the IkI_{k}. Then we define

(61) w+(x)\displaystyle w_{+}(x) {1λklogcos[λk(|x|Rk)]+kfor |x|Ik with even k,+else,\displaystyle\coloneqq\begin{cases}-\frac{1}{\lambda_{k}}\log\cos\big{[}\lambda_{k}\,(|x|-R_{k})\big{]}+k&\text{for }|x|\in I_{k}\text{ with even }k,\\ +\infty&\text{else},\end{cases}
(62) w(x)\displaystyle w_{-}(x) {+1λklogcos[λk(|x|Rk)]+k3for |x|Ik with odd k,else.\displaystyle\coloneqq\begin{cases}+\frac{1}{\lambda_{k}}\log\cos\big{[}\lambda_{k}\,(|x|-R_{k})\big{]}+k^{3}&\text{for }|x|\in I_{k}\text{ with odd }k,\\ -\infty&\text{else.}\end{cases}

The functions w±w_{\pm} are continuous and w<w+w_{-}<w_{+} holds, in fact, even w+w=w_{+}-w_{-}=\infty is true. Figure 6 sketches w+w_{+} and ww_{-}. The different added constants kk and k3k^{3} are there to ensure that the barriers corresponding to w+w_{+} and ww_{-} stay interlocked for all times such that a mean curvature flow between those is bound to oscillate infinitely towards Ωt×\partial\Omega_{t}\times\mathbb{R}. But we are going to have a closer look at this now.

Refer to caption
Figure 6: Sketch of a cross-section for w±w_{\pm}.

Let u0u_{0} be any smooth function on B1(0)2B_{1}(0)\subset\mathbb{R}^{2} with w<u0<w+w_{-}<u_{0}<w_{+} and with u0(x)u_{0}(x)\to\infty for xB1(0)x\to\partial B_{1}(0). Let u:Ωu\colon\Omega\to\mathbb{R} be a mean curvature flow without singularities that starts from u0u_{0}. The solution is defined up to the time T=12T=\frac{1}{2}, at which the balls Ωt\Omega_{t} shrink to a point.

Let TkRk22T_{k}\coloneqq\frac{R_{k}^{2}}{2}. Theorem 11 yields for even kk

(63) u(0,Tk)k1λklogcos[λk(2TkRk)]+(λk+2λkπRk)Tk=k+((k1)kπ+2(k1)kRk)Tkπ2k2(k).\begin{split}u(0,T_{k})&\leq k-\frac{1}{\lambda_{k}}\log\cos\left[\lambda_{k}\left(\sqrt{2T_{k}}-R_{k}\right)\right]+\left(\lambda_{k}+\sqrt{\frac{2\lambda_{k}}{\pi\,R_{k}}}\right)T_{k}\\ &=k+\left((k-1)\,k\,\pi+\sqrt{\frac{2(k-1)\,k}{R_{k}}}\right)T_{k}\\ &\simeq\frac{\pi}{2}\,k^{2}\qquad(k\to\infty)\;.\end{split}

Analogously, we obtain for odd kk

(64) u(0,Tk)k3((k1)kπ+2(k1)kRk)Tkk3(k).\begin{split}u(0,T_{k})&\geq k^{3}-\left((k-1)\,k\,\pi+\sqrt{\frac{2(k-1)\,k}{R_{k}}}\right)T_{k}\\ &\simeq k^{3}\qquad(k\to\infty)\;.\end{split}

This says that u(0,T2l+1)u(0,T_{2l+1}) is larger than 12(2l+1)3\frac{1}{2}(2l+1)^{3}, while u(0,T2l)u(0,T_{2l}) is smaller than 2(2l)22(2l)^{2} if the number ll\in\mathbb{N} is large enough. This shows that u(0,t)u(0,t) has no clear rate with which it tends to infinity. Since Tk12T_{k}\to\frac{1}{2} it also shows that u(0,t)u(0,t) oscillates more and more on smaller and smaller time intervals and that there must be a sequence (tk)k(t_{k})_{k\in\mathbb{N}} of times with tkT=12t_{k}\to T=\frac{1}{2} and u˙(0,tk)\dot{u}(0,t_{k})\to-\infty, while it is clear that there are also sequences (t~k)k(\tilde{t}_{k})_{k\in\mathbb{N}} with t~kT\tilde{t}_{k}\to T and u˙(0,t~k)+\dot{u}(0,\tilde{t}_{k})\to+\infty. In summary, one can say that much of the behavior of u0u_{0} for |x|R=1|x|\to R=1 is transmitted by virtue of Theorem 11 to u(0,t)u(0,t) for tTt\to T. In the next paragraph we will see more of this.

Remark.

An example of a mean curvature flow without singularities akin to the one we have discussed is defined on a half-space. It also has unbounded curvature for all t0t\geq 0 and it sheets towards a plane. It can be constructed in a similar fashion but instead of the annulus barrier one uses grim reapers cross \mathbb{R} in that case.

5.3 Relation between spatial and temporal asymptotics

Theorem 11 shows that the height of a solution can be estimated by the height of the solution on an annulus at a prior time. This can be used to derive a relationship between temporal and spatial asymptotics.

Theorem 12.

Let n2n\geq 2 and let u0:nBρ(0)u_{0}\colon\mathbb{R}^{n}\supset B_{\rho}(0)\to\mathbb{R} be a smooth function and suppose that there are α>1\alpha>1 and C>0C>0 such that

(65) u0(x)C(ρ|x|)α(|x|ρ).u_{0}(x)\simeq C\,(\rho-|x|)^{-\alpha}\qquad(|x|\to\rho)\;.

Let u:Ωu\colon\Omega\to\mathbb{R} be a mean curvature flow without singularities starting from u0u_{0}. Then

(66) u(0,t)C(ρ2(n1)t)α(tT)C(n1ρ(Tt))α(tT)\begin{split}u(0,t)&\simeq C\left(\rho-\sqrt{2(n-1)t}\right)^{-\alpha}\qquad(t\to T)\\ &\simeq C\left(\frac{n-1}{\rho}\,(T-t)\right)^{-\alpha}\qquad(t\to T)\end{split}

holds with Tρ22(n1)T\coloneqq\frac{\rho^{2}}{2(n-1)}.

Proof.

The idea is to use the barriers over annuli as depicted in Figure 7.

Refer to caption
Figure 7: Use of the barriers for relation of spatial and temporal asymptotics.

Let 0<t<T0<t<T, rt2(n1)tr_{t}\coloneqq\sqrt{2(n-1)t}, and λtπ2(ρrt)1+α2\lambda_{t}\coloneqq\frac{\pi}{2}\,(\rho-r_{t})^{-\frac{1+\alpha}{2}}. We set

(67) A(t)A0(rt,λt)={x:rtπ2λt<|x|<rt+π2λt}.A(t)\coloneqq A_{0}(r_{t},\lambda_{t})=\left\{x\colon r_{t}-\frac{\pi}{2\lambda_{t}}<|x|<r_{t}+\frac{\pi}{2\lambda_{t}}\right\}.

We will assume that tt is sufficiently close to TT such that ρrt>(ρrt)1+α2\rho-r_{t}>(\rho-r_{t})^{\frac{1+\alpha}{2}} and hence A(t)Bρ(0)A(t)\Subset B_{\rho}(0) hold.

Theorem 11 yields

(68) u(0,t)supxA(t)u0(x)+(λt+2λtπrt)t.u(0,t)\leq\sup_{x\in A(t)}u_{0}(x)+\left(\lambda_{t}+\sqrt{\frac{2\lambda_{t}}{\pi\,r_{t}}}\right)t\;.

We determine the asymptotic behavior: Let xtA(t)¯x_{t}\in\overline{A(t)} with supxA(t)u0(x)=u0(xt)\sup_{x\in A(t)}u_{0}(x)=u_{0}(x_{t}). We observe that

(69) ρrt±π2λtρrt=ρrt±(ρrt)(1+α)/2ρrt=1±(ρrt)α121(tT).\frac{\rho-r_{t}\pm\frac{\pi}{2\lambda_{t}}}{\rho-r_{t}}=\frac{\rho-r_{t}\pm(\rho-r_{t})^{(1+\alpha)/2}}{\rho-r_{t}}=1\pm(\rho-r_{t})^{\frac{\alpha-1}{2}}\simeq 1\qquad(t\to T)\;.

From this we infer that ρ|xt|ρrt\rho-|x_{t}|\simeq\rho-r_{t} holds and we deduce

(70) supxA(t)u0(x)=u0(xt)C(ρ|xt|)αC(ρrt)α(tT).\sup_{x\in A(t)}u_{0}(x)=u_{0}(x_{t})\simeq C\,(\rho-|x_{t}|)^{-\alpha}\simeq C\,(\rho-r_{t})^{-\alpha}\qquad(t\to T)\;.

Similar to (68) and (70), we find

(71) u(0,t)infxA(t)u0(x)(λt+2λtπrt)tu(0,t)\geq\inf_{x\in A(t)}u_{0}(x)-\left(\lambda_{t}+\sqrt{\frac{2\lambda_{t}}{\pi\,r_{t}}}\right)t

and

(72) infxA(t)u0(x)C(ρrt)α(tT).\inf_{x\in A(t)}u_{0}(x)\simeq C\,(\rho-r_{t})^{-\alpha}\qquad(t\to T)\;.

Lastly, we note that

(73) (λt+2λtπrt)tπ2(ρrt)(1+α)/2T(tT).\left(\lambda_{t}+\sqrt{\frac{2\lambda_{t}}{\pi\,r_{t}}}\right)t\simeq\frac{\pi}{2}\,(\rho-r_{t})^{-(1+\alpha)/2}\,T\qquad(t\to T)\;.

Taking together (68), (70), (71), (72), and (73), we find (note α>1+α2\alpha>\frac{1+\alpha}{2})

(74) u(0,t)C(ρrt)α=C(ρ2(n1)t)α(tT).u(0,t)\simeq C\,(\rho-r_{t})^{-\alpha}=C\left(\rho-\sqrt{2(n-1)t}\right)^{-\alpha}\qquad(t\to T)\;.

With a Taylor expansion at t=T=ρ22(n1)t=T=\frac{\rho^{2}}{2(n-1)}, we can also write this as

(75) u(0,t)C(n1ρ(Tt))α(tT).u(0,t)\simeq C\left(\frac{n-1}{\rho}(T-t)\right)^{-\alpha}\qquad(t\to T)\;.\qed
Remark.

The assumption on the rate of u0u_{0} is not expected to be optimal: In [14] and [15] rotational symmetric solutions of the mean curvature flow without singularities that fit into our setting have been constructed. Their solutions have the asymptotics Theorem 12 is concerned with, while Isenberg and Wu only state the asymptotics for blow-ups. Their results concentrate more on the blow-up rate of the curvature and they prove that any blow-up rate (Tt)α(T-t)^{-\alpha} with α1\alpha\geq 1 for the curvature is possible. The case α=1\alpha=1 is dealt with in [15] and is beyond the scope of our Theorem 12.

Appendix A Normal graphs

Definition 13 (Tubular Neighborhood).

Let NN be a C2C^{2}-hypersurface of n+1\mathbb{R}^{n+1} and let ν\nu be a continuous normal to NN. If for δ>0\delta>0 (δ=\delta=\infty is possible) the map Φ:(p,d)pdν\Phi\colon(p,d)\mapsto p-d\,\nu, defined on N×(δ,δ)N\times(-\delta,\delta), is a diffeomorphism onto its image, then this image is called a tubular neighborhood of NN of thickness δ\delta and we denote it by NδN_{\delta}. The diffeomorphism Φ\Phi is called tubular diffeomorphism.

Definition 14 (Normal Graph).

Let NN be a C2C^{2}-hypersurface of n+1\mathbb{R}^{n+1} with tubular diffeomorphism Φ:N×(δ,δ)Nδ\Phi\colon N\times(-\delta,\delta)\to N_{\delta}. A hypersurface MM of n+1\mathbb{R}^{n+1} is called a normal graph over NN (or has a normal graph representation over NN) if M=Φ(graphu)M=\Phi(\mathop{\mathrm{graph}}u) for some function u:NN(δ,δ)u\colon N\supset N^{\prime}\to(-\delta,\delta). This function is called the representation function.

A.1 Geometry of normal graphs

Proposition 15.

Let MM be a normal graph over a C3C^{3}-hypersurface NnN^{n} of n+1\mathbb{R}^{n+1} with representation function uu. We denote the metric and second fundamental form of NN by gijg_{ij} and hijh_{ij} and the Weingarten map with AA. On MM we choose the normal νM\nu^{M} which satisfies νM,ν0\left\langle\nu^{M},\nu\right\rangle\geq 0, where ν\nu is the normal of NN. Then, with Ba=0(uA)aB\coloneqq\sum_{a=0}^{\infty}(u\,A)^{a},

(76) gijM\displaystyle g_{ij}^{M} =gij+uiuj2uhij+u2hikhjk,\displaystyle=g_{ij}+u_{i}\,u_{j}-2\,u\,h_{ij}+u^{2}\,h_{ik}\,h_{j}^{k}\;,
(77) hijM\displaystyle h_{ij}^{M} =11+|DuB|2[hij+uijuhjkhkj+(uihjk+ujhik+ujhik)ulBkl].\displaystyle=\frac{1}{\sqrt{1+\left|\mathrm{D}u\cdot B\right|^{2}}}\big{[}h_{ij}+u_{ij}-u\,h_{j}^{k}\,h_{kj}+\big{(}u_{i}\,h_{j}^{k}+u_{j}\,h_{i}^{k}+u\,\nabla_{j}h_{i}^{k}\big{)}\,u_{l}\,B_{k}^{l}\big{]}\;.
Proof.

We will denote the embedding of NN into n+1\mathbb{R}^{n+1} with NN, too. A parametrization of MM is given by X(x)=N(x)u(x)ν(x)X(x)=N(x)-u(x)\,\nu(x) with xNNx\in N^{\prime}\subset N. We compute derivatives up to the second order while noting Nij=hijνN_{ij}=-h_{ij}\,\nu and νi=hikNk\nu_{i}=h_{i}^{k}\,N_{k}:

(78) Xi\displaystyle X_{i} =Niuiνuνi=NiuiνuhikNk,\displaystyle=N_{i}-u_{i}\,\nu-u\,\nu_{i}=N_{i}-u_{i}\,\nu-u\,h_{i}^{k}\,N_{k}\;,
Xij\displaystyle X_{ij} =NijuijνuiνjujhikNku(jhik)NkuhikNkj\displaystyle=N_{ij}-u_{ij}\,\nu-u_{i}\,\nu_{j}-u_{j}\,h_{i}^{k}\,N_{k}-u\,(\nabla_{j}h_{i}^{k})\,N_{k}-u\,h_{i}^{k}\,N_{kj}
(79) =(hijuij+uhikhkj)ν(uihjk+ujhik+ujhik)Nk.\displaystyle=(-h_{ij}-u_{ij}+u\,h_{i}^{k}\,h_{kj})\,\nu-(u_{i}\,h_{j}^{k}+u_{j}\,h_{i}^{k}+u\,\nabla_{j}h_{i}^{k})\,N_{k}\;.

The asserted identity (76) for gijM=Xi,Xjg_{ij}^{M}=\left\langle X_{i},X_{j}\right\rangle follows.

For the second identity we compute hijM=Xij,νM.h_{ij}^{M}=-\left\langle X_{ij},\nu^{M}\right\rangle. To this end, we observe that the submanifold MM is given by du~=0d-\tilde{u}=0, where dd is the distance function to NN and u~=uπ\tilde{u}=u\circ\pi is the extension of uu to the tubular neighborhood of NN which is constant in normal direction; π\pi is the projection to the closest point in NN. (We actually have Φ1=(π,d)\Phi^{-1}=(\pi,d), where Φ\Phi is the tubular diffeomorphism.) So for vTpMv\in\mathrm{T}_{p}M we have (DdDu~)|pv=0(\mathrm{D}d-\mathrm{D}\tilde{u})|_{p}\cdot v=0. From here we see that (du~)X(\nabla d-\nabla\tilde{u})\circ X is proportional to the normal νM\nu^{M} of MM. It actually points in the direction of νM-\nu^{M}. Moreover, |du~|2=1+|u~|2|\nabla d-\nabla\tilde{u}|^{2}=1+|\nabla\tilde{u}|^{2} since d\nabla d and u~\nabla\tilde{u} are orthogonal and |d|=1|\nabla d|=1. Using d=νπ\nabla d=-\nu\circ\pi and Du~|Xν=0\mathrm{D}\tilde{u}|_{X}\cdot\nu=0, we obtain

(80) hijM=Xij,νM=(DdDu~1+|Du~|2X)Xij=11+|u~|X|2[hij+uijuhikhkj+(uihjk+ujhik+ujhik)Du~|XNk].\begin{split}h_{ij}^{M}&=-\left\langle X_{ij},\nu^{M}\right\rangle=\left(\frac{\mathrm{D}d-\mathrm{D}\tilde{u}}{\sqrt{1+|\mathrm{D}\tilde{u}|^{2}}}\circ X\right)\cdot X_{ij}\\ &=\frac{1}{\sqrt{1+\left|\nabla\tilde{u}|_{X}\right|^{2}}}\left[h_{ij}+u_{ij}-u\,h_{i}^{k}\,h_{kj}+\left(u_{i}\,h_{j}^{k}+u_{j}\,h_{i}^{k}+u\,\nabla_{j}h_{i}^{k}\right)\,\mathrm{D}\tilde{u}|_{X}\cdot N_{k}\right]\;.\end{split}

Now we turn our attention to Du~|X=(Du|πDπ)|X=DuDπ|X\mathrm{D}\tilde{u}|_{X}=(\mathrm{D}u|_{\pi}\cdot\mathrm{D}\pi)|_{X}=\mathrm{D}u\cdot\mathrm{D}\pi|_{X} and focus on Dπ|X\mathrm{D}\pi|_{X}. Because id=Φ(π,d)=N(π)dν(π)\textrm{id}=\Phi(\pi,d)=N(\pi)-d\,\nu(\pi) holds, we compute, using Dν=DNA\mathrm{D}\nu=\mathrm{D}N\cdot A and d|X=ν\nabla d|_{X}=-\nu,

(81) idTNδ|X=DNDπ|Xd|XDνDπ|XνDd|X=DN(idTNd|XA)Dπ|X+νν.\begin{split}\mathrm{id}_{\mathrm{T}N_{\delta}}|_{X}&=\mathrm{D}N\cdot\mathrm{D}\pi|_{X}-d|_{X}\,\mathrm{D}\nu\cdot\mathrm{D}\pi|_{X}-\nu\otimes\mathrm{D}d|_{X}\\ &=\mathrm{D}N\cdot\left(\mathrm{id}_{\mathrm{T}N}-d|_{X}\,A\right)\cdot\mathrm{D}\pi|_{X}+\nu\otimes\nu^{\flat}\;.\end{split}

We observe that d|X=ud|_{X}=u holds. Because νN=ν,N=0\nu^{\flat}\cdot\nabla N=\left\langle\nu,\nabla N\right\rangle=0, a multiplication of N\nabla N from the right and a subsequent cancellation of N\nabla N on the left yields

(82) idTN=(idTNuA)Dπ|XDN.\mathrm{id}_{\mathrm{T}N}=\left(\mathrm{id}_{\mathrm{T}N}-u\,A\right)\cdot\mathrm{D}\pi|_{X}\cdot\mathrm{D}N\,.

We only mention here that uA<1\|u\,A\|<1 holds because the thickness of the maximal tubular neighborhood is bounded by the curvature in that way, though there might be global effects further cutting down the maximal thickness. By the Neumann series, (idTNuA)1=a=0(uA)aB(\mathrm{id}_{\mathrm{T}N}-u\,A)^{-1}=\sum_{a=0}^{\infty}(u\,A)^{a}\eqqcolon B. Hence, we obtain the identity

(83) Dπ|XDN=B.\mathrm{D}\pi|_{X}\cdot\mathrm{D}N=B\;.

From this equation we deduce

(84) Du~|XDN=DuB.\mathrm{D}\tilde{u}|_{X}\cdot\mathrm{D}N=\mathrm{D}u\cdot B\;.

Because the metric on NN is the metric on n+1\mathbb{R}^{n+1} pulled back via the embedding NN, we have

(85) |Du~|X|n+1=|Du~|XDN|g=|DuB|g.|\mathrm{D}\tilde{u}_{|X}|_{\mathbb{R}^{n+1}}=|\mathrm{D}\tilde{u}_{|X}\cdot\mathrm{D}N|_{g}=|\mathrm{D}u\cdot B|_{g}\;.

Substituting (84) and (85) into (80) yields (77). ∎

A.2 Normal graphs and local graph representations

Definition 16.

Let MM be a Riemannian manifold. We say a point xMx\in M has buffer r>0r>0 in MM if any curve γ:[0,a)M\gamma\colon[0,a)\to M of length at most rr and with γ(0)=x\gamma(0)=x is extendable to a curve γ:[0,r]M\gamma\colon[0,r]\to M.

A subset MMM^{\prime}\subset M has buffer rr in MM if every point of MM^{\prime} has buffer rr in MM.

Definition 17.

Let MM be a hypersurface. We say MM admits local graph representations of radius r>0r>0 if Br(x)MB_{r}(x)\cap M is graphical over the tangential hyperplane at xx for any point xMx\in M.

The following lemma says that normal graphs over a given hypersurface admit local graph representations of a controlled radius.

Proposition 18.

Let NN be a complete hypersurface and let δ\delta be the thickness of a tubular neighborhood of NN. Let MM be a normal graph over NN with representation function u:N(δ2,δ2)u\colon N\to(-\frac{\delta}{2},\frac{\delta}{2}). We assume that |Du|L<16|\mathrm{D}u|\leq L<\frac{1}{6} holds. Let NNN^{\prime}\subset N have buffer ρ>0\rho>0 and suppose that the distance functions distn\mathrm{dist}_{\mathbb{R}^{n}} and distN\mathrm{dist}_{N} are equivalent on NN: distNC0distn\mathrm{dist}_{N}\leq C_{0}\,\mathrm{dist}_{\mathbb{R}^{n}}.

Then there is r>0r>0 such that MM admits local graph representations of radius rr. The radius rr depends only on LL, ρ\rho, C0C_{0}, and sup|AN|\sup|A_{N}|. Moreover, the gradients in the local graph representations are bounded by 8L8L.

Proof.

From the proof of Proposition 15 we know that

(86) d,νM=11+|DuB|2\left\langle\nabla d,-\nu_{M}\right\rangle=\frac{1}{\sqrt{1+|\mathrm{D}u\cdot B|^{2}}}

holds with B=(IuA)1=k=0(uA)kB=(I-uA)^{-1}=\sum_{k=0}^{\infty}(uA)^{k}. Because of |u|<δ2|u|<\frac{\delta}{2} the eigenvalues of uAuA are at most 12\frac{1}{2} in modulus. So the eigenvalues of BB are bounded by 22. Thus, from the gradient estimate |Du|L|\mathrm{D}u|\leq L we obtain

(87) d,νM11+(2|Du|)211+(2L)2,\left\langle\nabla d,-\nu_{M}\right\rangle\geq\frac{1}{\sqrt{1+(2|\mathrm{D}u|)^{2}}}\geq\frac{1}{\sqrt{1+(2L)^{2}}}\;,

where dd is the distance function to NN.

Let x0,xMx_{0},x\in M. Then

(88) d(x0),νM(x)=d(x),νM(x)+d(x0)d(x),νM(x).\left\langle\nabla d(x_{0}),-\nu_{M}(x)\right\rangle=\left\langle\nabla d(x),-\nu_{M}(x)\right\rangle+\left\langle\nabla d(x_{0})-\nabla d(x),-\nu_{M}(x)\right\rangle\;.

For a curve γ:[0,1]M\gamma\colon[0,1]\to M with γ(0)=x0\gamma(0)=x_{0} and γ(1)=x\gamma(1)=x, we have

(89) d(x)d(x0)=01ddtd(γ)dt=012d(γ)γdt,\nabla d(x)-\nabla d(x_{0})=\int_{0}^{1}\frac{\mathrm{d}}{\mathrm{d}t}\nabla d(\gamma)\,\mathrm{d}t=\int_{0}^{1}\nabla^{2}d(\gamma)\cdot\gamma^{\prime}\,\mathrm{d}t\;,

and therefore

(90) |d(x0)d(x)|2dL(Nδ/2)distM(x,x0)Cdistn(x,x0).|\nabla d(x_{0})-\nabla d(x)|\leq\|\nabla^{2}d\|_{L^{\infty}(N_{{\delta}/2})}\,\mathrm{dist}_{M}(x,x_{0})\leq C\,\mathrm{dist}_{\mathbb{R}^{n}}(x,x_{0})\;.

If distn(x,x0)r\mathrm{dist}_{\mathbb{R}^{n}}(x,x_{0})\leq r for sufficiently small r>0r>0, we obtain from (88), (87), and (90)

(91) d(x0),νM(x)11+(2L)2Cdistn(x,x0)11+(3L)2.\left\langle\nabla d(x_{0}),-\nu_{M}(x)\right\rangle\geq\frac{1}{\sqrt{1+(2L)^{2}}}-C\,\mathrm{dist}_{\mathbb{R}^{n}}(x,x_{0})\geq\frac{1}{\sqrt{1+(3L)^{2}}}\;.

That means that in the ball Br(x0)B_{r}(x_{0}), MM is locally graphical over the hyperplane determined by d(x0)\nabla d(x_{0}). The gradient is bounded by 3L3L because for a hyperplane the corresponding endomorphism BB is the identity. Thus, MM is also locally graphical over the hyperplane Tx0M\textrm{T}_{x_{0}}M with gradient bounded by 8L8L. The calculation goes like this: Let ξ,ηn+1\xi,\eta\in\mathbb{R}^{n+1} be two unit-vectors with ξ1,η1(1+(3L)2)1/2\xi^{1},\eta^{1}\geq(1+(3L)^{2})^{-1/2}. Then the remaining components ξ^=(ξ2,,ξn+1)\hat{\xi}=(\xi^{2},\ldots,\xi^{n+1}) satisfy

|ξ^|2=1(ξ1)2111+(3L)2=(3L)21+(3L)2,\left|\hat{\xi}\right|^{2}=1-(\xi^{1})^{2}\leq 1-\frac{1}{1+(3L)^{2}}=\frac{(3L)^{2}}{1+(3L)^{2}}\;,

and analogous for η\eta. Hence,

ξ,η=ξ1η1+ξ^,η^11+(3L)2(3L)21+(3L)2=1(3L)21+(3L)2.\left\langle\xi,\eta\right\rangle=\xi^{1}\,\eta^{1}+\left\langle\hat{\xi},\hat{\eta}\right\rangle\geq\frac{1}{1+(3L)^{2}}-\frac{(3L)^{2}}{1+(3L)^{2}}=\frac{1-(3L)^{2}}{1+(3L)^{2}}\;.

We want to write this in the form (compare to (86))

ξ,η1(3L)21+(3L)2=!11+L2.\left\langle\xi,\eta\right\rangle\geq\frac{1-(3L)^{2}}{1+(3L)^{2}}\stackrel{{\scriptstyle!}}{{=}}\frac{1}{\sqrt{1+L^{\prime 2}}}.

We obtain

L=(1+(3L)2)2(1(3L)2)21=4(3L)21(3L)2<L<168L.L^{\prime}=\sqrt{\frac{(1+(3L)^{2})^{2}}{(1-(3L)^{2})^{2}}-1}=\frac{\sqrt{4(3L)^{2}}}{1-(3L)^{2}}\stackrel{{\scriptstyle L<\frac{1}{6}}}{{<}}8L\;.

Now, we need to prove that not only the normals point in the right directions and we have local (on MM) graph representations, but we also need to prove that the projection to the hyperplane Tx0M\mathrm{T}_{x_{0}}M is one-to-one. Assume that this is not the case and there exist two points whose projections on Tx0M\mathrm{T}_{x_{0}}M are the same. Let a>0a>0 be the Euclidean distance of those points. Because MM is a normal graph over NN with gradient bounded by 16\frac{1}{6} and MM is in a tubular neighborhood of half the maximal thickness (we are sufficently far away from focal points), we can make the radius rr small depending on sup|AN|\sup|A_{N}|, that in this case the projection of the two points to NN have NN-distance bounded by 12a\frac{1}{2}a. Because MM is a normal graph over NN with gradient bound 16\frac{1}{6}, the difference in distance to NN of the two points must be bounded by 112a\frac{1}{12}a. But if rr is sufficiently small compared sup|AN|\sup|A_{N}|, then the Euclidean squared distance |pq|2|p-q|^{2} of two points is comparable to |d(p)d(q)|2+|πN(p)πN(q)|2|d(p)-d(q)|^{2}+|\pi_{N}(p)-\pi_{N}(q)|^{2}. As we can see, this is not the case for the above points. Therefore, the assumption that there exist two points in MBr(x0)M\cap B_{r}(x_{0}) whose projections to Tx0M\mathrm{T}_{x_{0}}M agree is false. We conclude that MBr(x0)M\cap B_{r}(x_{0}) is graphical over Tx0M\mathrm{T}_{x_{0}}M. ∎

A.3 Hypersurfaces close to each other

In this paragraph, we show that a hypersurface MM with bounded curvature which lies sufficiently close to a hypersurface NN can be written as a normal graph over NN. In fact, the C1C^{1}-norm of the graphical representation is small if MM is only close enough to NN; very much in the spirit of the interpolation inequality uC12CuC0uC2\|u\|_{C^{1}}^{2}\leq C\,\|u\|_{C^{0}}\,\|u\|_{C^{2}}.

Proposition 19.

Let Nn+1N\subset\mathbb{R}^{n+1} be a hypersurface with bounded curvature |AN|C<|A_{N}|\leq C<\infty. Let 0<δ<(2C)10<\delta<(2\,C)^{-1} be chosen such that δ\delta is the thickness of a tubular neighborhood of NN, denoted by NδN_{\delta}. Let MNδM\subset N_{\delta} be a hypersurface, also of bounded curvature |AM|C|A_{M}|\leq C. Let x0Mx_{0}\in M be a point with buffer r>0r>0 in MM. Then holds

(92) |Md|2(x0)δmax{6C,(πr)2δ}.|\nabla^{M}d|^{2}(x_{0})\leq\delta\cdot\max\left\{6\,C,\,\left(\frac{\pi}{r}\right)^{2}\delta\right\}\;.

Herein, d:Nδ(δ,δ)d\colon N_{\delta}\to(-\delta,\delta) denotes the signed distance to NN.

Proof.

Without loss of generality, we assume d(x0)0d(x_{0})\geq 0. Locally around x0x_{0}, we choose a continuous normal νM\nu_{M} of MM such that w0w(x0)d,νM|x00w_{0}\coloneqq w(x_{0})\coloneqq\left\langle\nabla d,-\nu_{M}\right\rangle|_{x_{0}}\geq 0. A continuous choice of a normal is always possible in MM-balls with radius bounded by π2C\frac{\pi}{2C}. This is because we can join any two points of the ball by a curve through x0x_{0} of length less than πC\frac{\pi}{C}. Noting that CC bounds the curvature, the direction of a normal is changing along such a curve by an angle less than π\pi. But any point admits only two directions for a normal, which are separated by an angle of π\pi. Thus, a normal field which is continuously extended along “short” curves starting from x0x_{0} is continuous.

In what follows, we note that |d|=1|\nabla d|=1, 2d(,d)=0\nabla^{2}d(\cdot,\nabla d)=0 and |2d|op2C|\nabla^{2}d|_{\mathrm{op}}\leq 2\,C\, hold, where ||op|\cdot|_{\mathrm{op}} denotes the operator norm (largest eigenvalue in modulus). This inequality follows from |d|<δ<(2C)1|d|<\delta<(2\,C)^{-1} and that the eigenvalues of 2d\nabla^{2}d which correspond to directions perpendicular to d\nabla d are given by κi1dκi-\frac{\kappa_{i}}{1-d\,\kappa_{i}}, where κi\kappa_{i} denote the principal curvatures of NN at the closest point on NN.

The assertion is trivial if |Md(x0)|=0|\nabla^{M}d(x_{0})|=0 holds. Therefore, we assume |Md(x0)|0|\nabla^{M}d(x_{0})|\neq 0 from now on. Let x(t)x(t) be the maximal solution of the initial value problem

(93) x˙(t)=Md(x(t))|Md|(x(t))=d(x(t))+w(x(t))νM(x(t))1w2(x(t))\dot{x}(t)=\frac{\nabla^{M}d(x(t))}{|\nabla^{M}d|(x(t))}=\frac{\nabla d(x(t))+w(x(t))\,\nu_{M}(x(t))}{\sqrt{1-w^{2}(x(t))}}

with x(0)=x0x(0)=x_{0} and wd,νMw\coloneqq\left\langle\nabla d,-\nu_{M}\right\rangle. With the second fundamental form hh of MM and noting that |Md|=1w2=|νM+wd||\nabla^{M}d|=\sqrt{1-w^{2}}=|\nu_{M}+w\,\nabla d|, along this curve holds

(94) |ddtw(x(t))|=|x˙d,νM+d,x˙νM|=|2d(x˙,νM+wd)h(x˙,Md)|3C1w2.\begin{split}\left|\frac{\mathrm{d}}{\mathrm{d}t}w(x(t))\right|&=\left|\left\langle\nabla_{\dot{x}}\nabla d,-\nu_{M}\right\rangle+\left\langle\nabla d,-\nabla_{\dot{x}}\nu_{M}\right\rangle\right|\\ &=\left|-\nabla^{2}d(\dot{x},\nu_{M}+w\,\nabla d)-h(\dot{x},\nabla^{M}d)\right|\\ &\leq 3\,C\sqrt{1-w^{2}}\;.\end{split}

It follows for the change in angle |ddtarccos(w(x(t)))|3C\left|\frac{\mathrm{d}}{\mathrm{d}t}\arccos(w(x(t)))\right|\leq 3\,C, and therefore

(95) arccos(w(x0))3Ctarccos(w(x(t))arccos(w(x0))+3Ct.\arccos(w(x_{0}))-3\,C\,t\leq\arccos(w(x(t))\leq\arccos(w(x_{0}))+3\,C\,t\;.

Thus, the solution x(t)x(t) exists for times 0t<min{r,T}0\leq t<\min\{r,T\}, where Tarccos(w(x0))3CT\coloneqq\frac{\arccos(w(x_{0}))}{3\,C}. It is important here that αarccos(w(x0))π2\alpha\coloneqq\arccos(w(x_{0}))\leq\frac{\pi}{2} which follows from the choice of the normal such that w0w(x0)0w_{0}\coloneqq w(x_{0})\geq 0. This is also crucial for the inequality in the third line of the following estimate.

(96) δd(x(t))d(x(0))=0tddtd(x(s))ds=0td,x˙ds=0t1w2(x)ds=0tsinarccosw(x)ds0tsin(arccos(w0)3Cs)ds=13C(cos(α3Ct)w0)=13C\begin{split}\delta&\geq d(x(t))-d(x(0))=\int_{0}^{t}\frac{\mathrm{d}}{\mathrm{d}t}d(x(s))\,\mathrm{d}s=\int_{0}^{t}\left\langle\nabla d,\dot{x}\right\rangle\,\mathrm{d}s\\ &=\int_{0}^{t}\sqrt{1-w^{2}(x)}\,\textrm{d}s=\int_{0}^{t}\sin\arccos w(x)\,\textrm{d}s\\ &\geq\int_{0}^{t}\sin(\arccos(w_{0})-3\,C\,s)\,\textrm{d}s=\frac{1}{3\,C}(\cos(\alpha-3\,C\,t)-w_{0})\\ &=\frac{1}{3\,C}\end{split}

From this we infer

(97) |Md|2(x0)=1w02(1+w0)αtδ2arccos(w0)1tδ.|\nabla^{M}d|^{2}(x_{0})=1-w_{0}^{2}\leq(1+w_{0})\,\frac{\alpha}{t}\,\delta\leq 2\arccos(w_{0})\,\frac{1}{t}\,\delta\;.

In the case that the solution x(t)x(t) exists up to time TT, we obtain |Md|2(x0)6Cδ|\nabla^{M}d|^{2}(x_{0})\leq 6\,C\,\delta from the last inequality by tTt\to T. Otherwise, from trt\to r follows (note arcsinxπ2x\arcsin x\leq\frac{\pi}{2}\,x)

(98) 2π|Md|(x0)|Md|2(x0)arcsin(|Md|(x0))=|Md|2(x0)arccos(w0)21rδ.\frac{2}{\pi}\,|\nabla^{M}d|(x_{0})\leq\frac{|\nabla^{M}d|^{2}(x_{0})}{\arcsin(|\nabla^{M}d|(x_{0}))}=\frac{|\nabla^{M}d|^{2}(x_{0})}{\arccos(w_{0})}\leq 2\,\frac{1}{r}\,\delta\;.

The assertion follows. ∎

Corollary 20.

In the situation of Proposition 19, but with δ<(24C)1\delta<(24\,C)^{-1} and δ<r2π\delta<\frac{r}{2\pi}, we denote the projection from MM onto the closest point in NN with p:MNp\colon M\to N. Let MM^{\prime} be a subset of MM with buffer r>0r>0. Then p|Mp|_{M^{\prime}} is a local diffeomorphism and MM^{\prime} is locally graphical over NN with gradient bounded by max{12Cδ,2πrδ}\max\left\{\sqrt{12\,C\,\delta},\,\sqrt{2}\frac{\pi}{r}\,\delta\right\}.

Proof.

From Proposition 19 we have

(99) |Md|2δmax{6C,(πr)2δ}<14,|\nabla^{M}d|^{2}\leq\delta\,\max\left\{6\,C,\,\left(\frac{\pi}{r}\right)^{2}\,\delta\right\}<\frac{1}{4}\;,

and it follows that in points of MM^{\prime}, Dp\mathrm{D}p is an isomorphism of the corresponding tangential spaces. By the inverse function theorem, pp is a local diffeomorphism. The representation function of the local graphical representation of MM^{\prime} over NN as a normal graph is denoted by uu. For the gradient bound we make use of (notice (86)):

(100) |Md|2=|u(IuA)1|21+|u(IuA)1|2(67)2|u|21+(67)2|u|2.|\nabla^{M}d|^{2}=\frac{|\nabla u\cdot(I-uA)^{-1}|^{2}}{1+|\nabla u\cdot(I-uA)^{-1}|^{2}}\geq\frac{(\frac{6}{7})^{2}|\nabla u|^{2}}{1+(\frac{6}{7})^{2}|\nabla u|^{2}}\,.

This yields

(101) |u|2(76)2|Md|21|Md|2(76)243|Md|22|Md|2max{12Cδ,2(πr)2δ2}.\begin{split}|\nabla u|^{2}&\leq\left(\frac{7}{6}\right)^{2}\,\frac{|\nabla^{M}d|^{2}}{1-|\nabla^{M}d|^{2}}\leq\left(\frac{7}{6}\right)^{2}\,\frac{4}{3}\,|\nabla^{M}d|^{2}\leq 2\,|\nabla^{M}d|^{2}\\ &\leq\max\left\{12\,C\,\delta,2\left(\frac{\pi}{r}\right)^{2}\,\delta^{2}\right\}\>.\qed\end{split}

In the special situation we face in our applications, we can overcome that the representation as a normal graph is only local. We formulate this as

Lemma 21.

In the situation of Corollary 20, let now be MM specifically given as M=graphuM=\mathop{\mathrm{graph}}u for a smooth function u:Ωu\colon\Omega\to\mathbb{R}. The set Ωn\Omega\subset\mathbb{R}^{n} is assumed to be open, smooth, and bounded and we assume u(x^)u(\hat{x})\to\infty for x^x^0Ω\hat{x}\to\hat{x}_{0}\in\partial\Omega. Moreover, let NN be specifically given as N=Ω×N=\partial\Omega\times\mathbb{R}.

For δ>0\delta>0, chosen like in Corollary 20, there exists a>0a>0 such that M{xn+1>a}M\cap\{x^{n+1}>a\} lies in the tubular neighborhood NδN_{\delta} and M{xn+1>a+1}M\cap\{x^{n+1}>a+1\} is a normal graph over NN with gradient bounded by max{12Cδ,2πδ}\max\left\{\sqrt{12\,C\,\delta},\,\sqrt{2}\,\pi\,\delta\right\}.

Proof.

We set a=min{u(x):xΩ,dist(x,Ω)δ}a=\min\{u(x)\colon x\in\Omega,\,\mathrm{dist}(x,\partial\Omega)\geq\delta\}. Then M{xn+1>a}M\cap\{x^{n+1}>a\} lies in the tubular neighborhood NδN_{\delta}.

We rename the hypersurfaces such that M{xn+1>a}M\cap\{x^{n+1}>a\} becomes MM and M{xn+1>a+1}M\cap\{x^{n+1}>a+1\} becomes MM^{\prime}. On MM we consider the projection pp onto NN, which by Corollary 20 is a local diffeomorphism on MM^{\prime}. Let X(Dp)1(en+1)X\coloneqq(\mathrm{D}p)^{-1}(e_{n+1}) and let α:M×[0,)M\alpha\colon M^{\prime}\times[0,\infty)\to M^{\prime} be the flow of the vector field XX. Then pα(x,s)=p(x)+sen+1p\circ\alpha(x,s)=p(x)+s\,e_{n+1} holds because s(pα)=Dpsα=DpDp1(en+1)=en+1\partial_{s}(p\circ\alpha)=\mathrm{D}p\cdot\partial_{s}\alpha=\mathrm{D}p\cdot\mathrm{D}p^{-1}(e_{n+1})=e_{n+1} and α(x,0)=x\alpha(x,0)=x.

Let x=(x^,xn+1)Mx=(\hat{x},x^{n+1})\in M^{\prime} be arbitrary. We consider x(s)α(x,s)x(s)\coloneqq\alpha(x,s). Without loss of generality, we may assume that e1e_{1} is the normal vector to NN at p(x)p(x) and that p(x)=(0,,0,h)p(x)=(0,\ldots,0,h) holds. Then x(s)x(s) is of the form

(102) x(s)=(x1(s),0,,0,h+s)=(x1(s),0,,0,u(x^(s))).x(s)=\big{(}x^{1}(s),0,\ldots,0,h+s\big{)}=\big{(}x^{1}(s),0,\ldots,0,u(\hat{x}(s))\big{)}\;.

It follows x1(s)0x^{1}(s)\to 0 for ss\to\infty. Suppose there is another point yMy\in M^{\prime} with p(y)=p(x)p(y)=p(x). Then y=(y1,0,,0,h)y=(y^{1},0,\ldots,0,h) holds, and we may assume without loss of generality that y1y^{1} is between x1=x1(0)x^{1}=x^{1}(0) and 0. Because x1(s)0x^{1}(s)\to 0, there is s>0s^{\prime}>0 such that x1(s)=y1x^{1}(s^{\prime})=y^{1}. But then h=u(y^)=u(x^)=h+sh=u(\hat{y})=u(\hat{x})=h+s^{\prime}, a contradiction.

The argument shows that pp is injective on MM^{\prime}. So pp is an injective local diffeomorphism. As a consequence, MM^{\prime} is a normal graph over NN.

The gradient bound follows from Corollary 20. ∎

Appendix B Set flow, domain flow, and α\alpha-noncollapsed mean curvature flow

We follow [9], which is based on [11, 12, 6, 3]. We also refer to [17].

Definition 22 (Set flow).

Let II\subset\mathbb{R} be an interval and let (Kt)tI(K_{t})_{t\in I} be a family of closed subsets of n+1\mathbb{R}^{n+1}. We say that (Kt)tI(K_{t})_{t\in I} is a set flow if for any smooth mean curvature flow (Mt)t[t0,t1](M_{t})_{t\in[t_{0},t_{1}]} of closed hypersurfaces and with [t0,t1]I[t_{0},t_{1}]\subset I, we have

(103) Kt0Mt0=t[t0,t1]:KtMt=.K_{t_{0}}\cap M_{t_{0}}=\emptyset\quad\Longrightarrow\quad\forall t\in[t_{0},t_{1}]:K_{t}\cap M_{t}=\emptyset\;.
Definition 23 (Level-set flow).

The level-set flow is the maximal set flow, where “maximal” is understood with respect to inclusion of sets.

Proposition 24.

For any compact subset K0n+1K_{0}\subset\mathbb{R}^{n+1} there exists a unique level-set flow (Kt)t[0,)(K_{t})_{t\in[0,\infty)}. It coincides with the level-set flow from the definition of Evans-Spruck and Chen-Giga-Goto.

Definition 25 (Domain flow).

Let II\subset\mathbb{R} be an interval and let (Ωt)tI(\Omega_{t})_{t\in I} be a family of open subsets of n+1\mathbb{R}^{n+1}. Then (Ωt)tI(\Omega_{t})_{t\in I} is a domain flow if for any family (Kt)t[a,b](K_{t})_{t\in[a,b]} of compact subsets of n\mathbb{R}^{n} whose boundaries (Kt)t[a,b](\partial K_{t})_{t\in[a,b]} form a classical mean curvature flow, we have KaΩat[a,b]:KtΩtK_{a}\subset\Omega_{a}\Longrightarrow\forall t\in[a,b]\colon K_{t}\subset\Omega_{t}; and the same holds with Ω¯c\overline{\Omega}^{c} instead of Ω\Omega.

Remark 26.

It can be shown with the help of the Jordan-Brouwer separation theorem that for any domain flow (Ωt)tI(\Omega_{t})_{t\in I}, (Ωt¯)t0(\overline{\Omega_{t}})_{t\geq 0} and (Ωt)t0(\partial\Omega_{t})_{t\geq 0} are set flows. However, we cannot have sudden vanishing, a problem which set flows may exhibit.

Proposition 27.

Let Kn+1K\subset\mathbb{R}^{n+1} have a mean convex C2C^{2}-boundary K\partial K. Then the level-set flow (Kt)t[0,)(K_{t})_{t\in[0,\infty)} starting from K0=KK_{0}=K satisfies Kt1Kt2K_{t_{1}}\supset K_{t_{2}} for any t1t2t_{1}\leq t_{2}.

Proposition 28.

If (Kt)t[0,)(K_{t})_{t\in[0,\infty)} and (Lt)t[0,)(L_{t})_{t\in[0,\infty)} are two set flows which are initially disjoint, then they stay disjoint.

Proposition 29.

If Ω0\Omega_{0} is bounded and has a strictly mean convex C2C^{2}-boundary, then there is a unique domain flow (Ωt)t[0,)(\Omega_{t})_{t\in[0,\infty)} starting from Ω0\Omega_{0}. It satisfies Ωt1Ωt2\Omega_{t_{1}}\Supset\Omega_{t_{2}} for t1<t2t_{1}<t_{2}.

Proof.

For some time, the flow (Ωt)t(\partial\Omega_{t})_{t} is smooth before singularities form. For this time, any weak flow is unique. By Proposition 28, for any ε>0\varepsilon>0 smaller then the first singular time, ΩtεΩtΩt+ε\Omega_{t-\varepsilon}\Supset\Omega_{t}\Supset\Omega_{t+\varepsilon} holds for tεt\geq\varepsilon. Any other weak solution is strictly contained between Ωt±ε\Omega_{t\pm\varepsilon} for tεt\geq\varepsilon as well. Note that ε>0\varepsilon>0 is arbitrary. Let (Ψt)t(\Psi_{t})_{t} be another weak solution. Let t>0t>0. If xΨtx\in\Psi_{t} then, by openness, there is a closed ball around xx that is completely contained inside Ψt\Psi_{t}. It takes some time τ\tau to shrink that ball to half its radius. Hence, xΨt+τΩtx\in\Psi_{t+\tau}\Subset\Omega_{t}. Since xΨtx\in\Psi_{t} was arbitrary, ΨtΩt\Psi_{t}\subset\Omega_{t} follows. The same argument can be made to show the reverse inclusion. We have just shown Ψt=Ωt\Psi_{t}=\Omega_{t} for arbitrary tt. ∎

Definition 30.

A closed subset K0n+1K_{0}\subset\mathbb{R}^{n+1} is said to be mean convex if Kt1Kt2K_{t_{1}}\supset K_{t_{2}} for t1t2t_{1}\leq t_{2}, where (Kt)t(K_{t})_{t} denotes the level-set flow starting from K0K_{0}.

Remark.

Adopting this definition, it is obvious that mean convexity is preserved along the level-set flow.

Definition 31.

Let Kn+1K\subset\mathbb{R}^{n+1} be a closed subset. We define the mean curvature in the viscosity sense in points xKx\in\partial K by

(104) H(x)=inf{HA(x):AK is a smooth domain and xA}.H(x)=\inf\big{\{}H_{\partial A}(x)\colon A\subset K\text{ is a smooth domain and }x\in\partial A\big{\}}\;.

The mean curvature in the viscosity sense may be infinite.

Definition 32 (α\alpha-Noncollapsedness).

Let α>0\alpha>0. A closed subset of n+1\mathbb{R}^{n+1} is called α\alpha-noncollapsed if for any point xKx\in\partial K the (viscosity) mean curvature satisfies H(x)[0,]H(x)\in[0,\infty] and there exist closed balls B¯int\overline{B}_{\mathrm{int}} and B¯ext\overline{B}_{\mathrm{ext}} of radius r(x)=αH(x)r(x)=\frac{\alpha}{H(x)} that contain xx and such that B¯intK\overline{B}_{\mathrm{int}}\subset K and B¯extn+1Int(K)\overline{B}_{\mathrm{ext}}\subset\mathbb{R}^{n+1}\setminus\mathrm{Int}(K) (see Fig. 8).

We also say that a mean convex hypersurface MtM_{t} is α\alpha-noncollapsed if the bounded closed region it bounds is α\alpha-noncollapsed in the above sense.

A family (Kt)tI(K_{t})_{t\in I} is called α{\alpha}-noncollapsed if KtK_{t} is α{\alpha}-noncollapsed for all tIt\in I.

Refer to caption
Figure 8: Exterior and interior touching balls.

Smooth closed hypersurfaces with positive mean curvature H>0H>0 are α\alpha-noncollapsed for some α>0\alpha>0. B. Andrews has shown in [1] that the hypersurface stays α\alpha-noncollapsed with the same α\alpha if one lets flow the hypersurface by its mean curvature. This result holds even in a weak setting:

Theorem 33.

Let K0n+1K_{0}\subset\mathbb{R}^{n+1} be a compact, smooth, and mean convex domain that is α\alpha-noncollapsed for α>0\alpha>0. Then the level-set flow that starts from K0K_{0} is α\alpha-noncollapsed.

Theorem 34.

For any α>0\alpha>0, there are ρ=ρ(α)>0\rho=\rho(\alpha)>0 and Cl=Cl(α)C_{l}=C_{l}(\alpha) (l=0,1,2,l=0,1,2,\ldots) with the following property. If (Mt)tI(M_{t})_{t\in I} is an α\alpha-noncollapsed mean curvature flow in a parabolic ball P(x,t,r)n+1×P(x,t,r)\subset\mathbb{R}^{n+1}\times\mathbb{R} with xMtx\in M_{t} and H(x,t)r1H(x,t)\leq r^{-1}, then

(105) supP(x,t,ρr)|lA|Clr(l+1).\sup_{P(x,t,\rho\,r)}|\nabla^{l}A|\leq C_{l}\,r^{-(l+1)}\;.
Idea of the proof.

The theorem is proven with a blow-up argument. One considers a rescaled sequence of counterexamples with sup|A|1\sup|A|\geq 1 and H(0)0H(0)\to 0. Using the α\alpha-noncollapsedness, one can show that the sequence converges locally smoothly to a hyperplane. This is the halfspace convergence result of [9]. Contradiction with sup|A|1\sup|A|\geq 1. ∎

Corollary 35.

For an α\alpha-noncollapsed mean curvature flow (Mt)t(0,T)(M_{t})_{t\in(0,T)}, there are constants Cl=Cl(α)C_{l}=C_{l}(\alpha) (l=0,1,2,l=0,1,2,\ldots) such that

(106) |lA|2(x,t)Cl(t1+H2(x,t))l+1.|\nabla^{l}A|^{2}(x,t)\leq C_{l}\,\big{(}t^{-1}+H^{2}(x,t)\big{)}^{l+1}\;.

In particular, for t0>0t_{0}>0 there are constants Cl=Cl(α,t0)C_{l}=C_{l}(\alpha,t_{0}) such that

(107) |lA|2(1+H2)l+1Cl(α,t0)holds for tt0.\frac{|\nabla^{l}A|^{2}}{(1+H^{2})^{l+1}}\leq C_{l}(\alpha,t_{0})\qquad\text{holds for }t\geq t_{0}\;.
Proof.

For xMtx\in M_{t} we set r(t1+H2(x,t))1/2r\coloneqq\big{(}t^{-1}+H^{2}(x,t)\big{)}^{-1/2}. Then there hold rH(x,t)1r\leq H(x,t)^{-1} and rt1/2r\leq t^{1/2}, which implies P(x,t,r)n+1×(0,T)P(x,t,r)\subset\mathbb{R}^{n+1}\times(0,T). By Theorem 34 there are constants Cl(α)C_{l}(\alpha) such that |lA|(x,t)Clr(l+1)=Cl(t1+H2(x,t))l+12|\nabla^{l}A|(x,t)\leq C_{l}\,r^{-(l+1)}=C_{l}\,\big{(}t^{-1}+H^{2}(x,t)\big{)}^{\frac{l+1}{2}} (l=0,1,l=0,1,\ldots) holds. The assertion follows with Cl2C_{l}^{2} instead of ClC_{l}, which is just as fine. ∎

References

  • [1] Ben Andrews, Noncollapsing in mean -convex mean curvature flow, Geom. Topol. 16 (2012), 1413–1418.
  • [2] Bing-Long Chen and Le Yin, Uniqeuness and pseudolocality theorems of the mean curvature flow, Comm. Anal. Geom. 15.3 (2007), 435–490.
  • [3] Yun Gang Chen, Yoshikazu Giga, and Shun’ichi Goto, Uniqueness and existence of viscosity solutions of generalized mean curvature flow equations, J. Differential Geom. 33.3 (1991), 749–786.
  • [4] Tobias H. Colding and William P. Minicozzi, II., Differentiability of the arrival time, arXiv e-prints (2015), arXiv:1501.07899.
  • [5] Klaus Ecker and Gerhard Huisken, Interior estimates for hypersurfaces moving by mean curvature, Invent. Math. 105.3 (1991), 547–569.
  • [6] Lawrence C. Evans and Joel Spruck, Motion of level sets by mean curvature. I, J. Differential Geom. 33.3 (1991), 635–681.
  • [7] Michael Gage and Richard S. Hamilton, The heat equations shrinking convex plane curves, J. Differential Geom.  23.1 (1986), 69–96.
  • [8] Matthew A. Grayson, The heat equation shrinks embedded plane curves to round points, J. Differential Geom. 26.2 (1987), 285–314.
  • [9] Robert Haslhofer and Bruce Kleiner, Mean Curvature Flow of Mean Convex Hypersurfaces, Comm Pure Appl. Math. 70.3 (2016), 511–546.
  • [10] Gerhard Huisken, Nonparametric mean curvature evolution with boundary conditions, J. Differential Equat.  77.2 (1989), 369–378.
  • [11] Tom Ilmanen, The Level-set Flow on a Manifold, Proc. Symp. Pure Math. 54.1 (1993), 193–204.
  • [12] Tom Ilmanen, Elliptic regularization and partial regularity for motion by mean curvature, Mem. Amer. Math. Soc. 108.520 (1994).
  • [13] Tom Ilmanen, André Neves, and Felix Schulze, On short time existence for the planar network flow, J. Differential Geom. 111.1 (2019), 39–89.
  • [14] James Isenberg and Haotian Wu, Mean curvature flow of noncompact hypersurfaces with Type-II curvature blow-up, J. Reine Angew. Math. 754 (2019), 225–251.
  • [15] James Isenberg, Haotian Wu, and Zhou Zhang, Mean curvature flow of non-compact hypersurfaces with Type-II curvature blow-up. II, arXiv e-Prints (2019), arXiv:1911.07282.
  • [16] Ananda Lahiri, Almost graphical hypersurfaces become graphical under mean curvature flow, arXiv e-Prints (2015), arXiv:1505.00543.
  • [17] Wolfgang Maurer, Shadows of graphical mean curvature flow, Comm. Anal. Geom. 29.1 (2021), 183–206
  • [18] Wolfgang Maurer, Mean curvature flow of symmetric double graphs only develops singularities on the hyperplane of symmetry, arXiv e-Prints (2021), arXiv:2103.06072.
  • [19] Mariel Sáez Trumper and Oliver Schnürer, Mean curvature flow without singularities, J. Differential Geom. 97.3 (2014), 545–570.

Wolfgang A. Maurer, Fachbereich Mathematik und Statistik, Universität Konstanz, 78457 Konstanz, Germany

e-mail: wolfgang.maurer@uni-konstanz.de