Skip to main content
Log in

Subsequent singularities in mean-convex mean curvature flow

  • Published:
Calculus of Variations and Partial Differential Equations Aims and scope Submit manuscript

Abstract

We use Ilmanen’s elliptic regularization to prove that for an initially smooth mean convex hypersurface in \(\mathbf {R}^n\) moving by mean curvature flow, the surface is very nearly convex in a spacetime neighborhood of every singularity. In particular, the tangent flows are all shrinking spheres or cylinders. Previously this was known only (1) for \(n\le 7\), and (2) for arbitrary \(n\) up to the first singular time.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. Part (2) of Theorem 1 was also proved by Huisken and Sinestrari [3] using different methods.

  2. See Appendix B for more details.

  3. Spacetime singular sets and parabolic Hausdorff dimension are not needed in this paper. The spacetime singular set is \(\{(x,u(x)): x\in X\}\). The parabolic Hausdorff dimension of a subset of \(\mathbf {R}^n\times \mathbf {R}\) is its Hausdorff dimension with respect to the metric \(d((x,t),(x',t'))=\max \{ |x-x'|, |t-t'|^{1/2} \}\). See [7, p. 666] or [6, §7].

  4. At first, one only knows subsequential convergence to some flow \(t\in [0,\infty )\mapsto M'_t\times \mathbf {R}\) with \(M'_0=M_0\). But then the uniqueness (or nonfattening) for mean curvature flow of mean convex surfaces implies that \(M'_t\equiv M_t\) for \(t\ge 0\). Since the limit is independent of the subsequence, we in fact have convergence and not just subsequential convergence.

References

  1. Brakke, K.A.: The motion of a surface by its mean curvature. In: Mathematical Notes, vol. 20. Princeton University Press, Princeton (1978)

  2. Hardt, R., Simon, L.: Boundary regularity and embedded solutions for the oriented Plateau problem. Ann. Math. (2) 110(3), 439–486 (1979). doi:10.2307/1971233

    Article  MathSciNet  MATH  Google Scholar 

  3. Huisken, G., Sinestrari, C.: Convexity estimates for mean curvature flow and singularities of mean convex surfaces. Acta Math. 183(1), 45–70 (1999). doi:10.1007/BF02392946

    Article  MathSciNet  MATH  Google Scholar 

  4. Ilmanen, T.: Elliptic regularization and partial regularity for motion by mean curvature. Mem. Am. Math. Soc. 108(520), x+90 (1994)

    MathSciNet  Google Scholar 

  5. White, B.: The topology of hypersurfaces moving by mean curvature. Commun. Anal. Geom. 3(1–2), 317–333 (1995)

    MATH  Google Scholar 

  6. White, B.: Stratification of minimal surfaces, mean curvature flows, and harmonic maps. J. Reine Angew. Math. 488, 1–35 (1997). doi:10.1515/crll.1997.488.1

    MathSciNet  MATH  Google Scholar 

  7. White, B.: The size of the singular set in mean curvature flow of mean-convex sets. J. Am. Math. Soc. 13(3), 665–695 (2000). doi:10.1090/S0894-0347-00-00338-6. (electronic)

    Article  MATH  Google Scholar 

  8. White, B.: The nature of singularities in mean curvature flow of mean-convex sets. J. Am. Math. Soc. 16(1), 123–138 (2003). doi:10.1090/S0894-0347-02-00406-X. (electronic)

    Article  MATH  Google Scholar 

  9. White, B.: A local regularity theorem for mean curvature flow. Ann. Math. (2) 161(3), 1487–1519 (2005). doi:10.4007/annals.2005.161.1487

    Article  MATH  Google Scholar 

  10. White, B.: Boundary regularity for mean curvature flow. In: Preparation (2011)

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Brian White.

Additional information

Communicated by G. Huisken.

This research was supported by the National Science Foundation under Grants DMS-0406209 and DMS 1105330.

Appendices

Appendix A: translating solitions

Here we prove existence and regularity of the translating solitons \(N_\lambda \) that were used in the proofs of Theorems 3 and 6.

Lemma 8

Let \(M\) be a connected, oriented hypersurface such that the mean curvature \(H\) is everywhere a nonnegative multiple of the unit normal \(\nu \) and such that \(M\) translates with constant velocity \(\mathbf {v}\ne 0\) under mean curvature flow, i.e., such that

$$\begin{aligned} H\cdot \nu = \mathbf {v}\cdot \nu . \end{aligned}$$

Then \(H\cdot \nu \) has no local interior minimum unless \(H\equiv 0\) (in which case, \(M\) lies in \(\Sigma \times L\) where \(L\) is a line parallel to \(\mathbf {v}\) and where \(\Sigma \) is a minimal hypersurface in \(L^\perp \)).

Proof

This is because under mean curvature flow, \(H\cdot \nu \) satisfies a nice second order parabolic equation. The result then follows immediately from the strong maximum principle. See for example [8, Theorem 2]. \(\square \)

Lemma 9

Let \(M_1\) and \(M_2\) be two disjoint, compact, connected smooth hypersurfaces in \(\mathbf {R}^{n+1}\) such that \(M_i\) translates by \(\mathbf {v}\ne 0\) under mean curvature flow. If the function

$$\begin{aligned} (p_1,p_2)\in M_1\times M_2 \mapsto |p-q| \end{aligned}$$

achieves its minimum value \(\mu \) at \((p,q)\) for some interior points \(p\in M_1\) and \(q\in M_2\), then \(M_1\) and \(M_2\) lie in parallel hyperplanes, and the vector \(\mathbf {v}\) is parallel to those planes.

Proof

Since the result is essentially local, we may assume that \(M_1\) and \(M_2\) are graphs of functions \(f_1\) and \(f_2\) over the same domain \(D\) in a hyperplane perpendicular to the line \(p_1p_2\), with \(f_2>f_1\) at all points. Note that \(f_2-f_1\) has a local minimum. Thus by the maximum principle, it is constant: \(f_2-f_1\equiv \mu \). Consequently, for each point \(x\) in the domain of the \(f_i\), we have \(f_2(x)-f_1(x)=\mu \), the minimum distance between the two graphs. But that implies that \(Df_2(x)=Df_1(x)=0\). In other words, \(f_1\) and \(f_2\) must be constant. Thus \(M_1\) and \(M_2\) are planar. The fact that \(\mathbf {v}\) is parallel to the those planes follows immediately (by Lemma 8, for example). \(\square \)

Theorem 10

Let \(n\ge 2\). Let \(W\) be a bounded region in an \(\mathbf {R}^n\) with piecewise smooth, mean convex boundary and let \(\lambda \ge 0\). Let \(S\) be a smooth, closed \((n-1)\)-manifold in \((\partial W)\times \mathbf {R}\) that is graph-like in the following sense: each line \(\{x\}\times \mathbf {R}\) with \(x\in \partial W\) intersects \(S\) either in a point or in a line segment.

Then there is a smooth, compact \(n\)-manifold \(N\) in \(\overline{W}\times \mathbf {R}\) such that

  1. (1)

    \(\partial N=S\),

  2. (2)

    \(N{\setminus } S\) is the graph of a smooth function \(f: W\rightarrow \mathbf {R}\),

  3. (3)

    \(N\) translates with velocity \(-\lambda \mathbf {e}_{n+1}\).

Proof

First we recall that there is an entire function \(\phi : \mathbf {R}^n\rightarrow \mathbf {R}\) such that \(\phi \) is rotationally symmetric (i.e., such that \(\phi (x)\) depends only on \(|x|\)) and such that the graph of \(\phi \) translates with velocity \(-\lambda \mathbf {e}_{n+1}\). By adding a constant to \(\phi \), we can assume that \(S\) lies below the graph of \(\phi \) Let \(b= \min \{y: (x,y)\in S\}\). Let \(R\) be the region

$$\begin{aligned} \{ (x,y) \in \overline{W}\times \mathbf {R}: b \le y \le \phi (x)\}. \end{aligned}$$

Now let \(N\) be a variety (e.g. an integral current) such that \(N\) minimizes the \(\lambda \)-area among all varieties that have boundary \(S\) and that lie in the region \(R\).

By the maximum principle, \(N{\setminus } S\) lies in the interior of \(R\). By [2], \(N\) is a smooth manifold with boundary in a neighborhood of \(S\). Thus the singular set \(X\) is a compact subset of the interior of \(R\). Also, the singular set has Hausdorff dimension at most \(n-7\).

Let \(N(s)\) be the result of translating \(N\) upward by distance \(s\).

Claim 1: for \(s>0\), the interiors of \(N\) and \(N(s)\) are disjoint.

Proof of Claim 1: Suppose not. Let \(s\) be the largest number such that \(N\) and \(N(s)\) intersect in the interior or are tangent to each other at a common boundary point. By the boundary maximum principle, they cannot be tangent to each other at a common boundary point. Thus they touch at an one or more interior points. Let \(Z\) be the set of interior points where they touch. Then \(Z\) is a compact subset \(W\times \mathbf {R}\). Now it follows that for sufficiently small \(\epsilon >0\), the function

$$\begin{aligned} (p,q) \in N\times N(s+\epsilon ) \mapsto |p-q| \end{aligned}$$

will have an interior local minimum \((p,q)\). But Lemma 9 then implies that \(N\) and \(N(s+\epsilon )\) lie in parallel vertical planes, which is clearly incompatible with the boundary conditions. This completes the Proof of Claim 2.

Claim 1 implies that \(N{\setminus } S\) is the graph of a continuous function \(f: W\rightarrow \mathbf {R}\).

It remains to show that the singular set \(X\) is empty.

Let \(K\subset W\) be a compact set with smooth boundary such that the singular set \(X\) is contained in the interior of \(K\times \mathbf {R}\).

By Lemma 8, \(\hbox {Tan}_pN\) cannot be vertical at any regular point. In particular, this means that the function \(f\) is smooth on except on the compact set \(\pi (X)\) (where \(\pi :\mathbf {R}^n\times \mathbf {R}\rightarrow \mathbf {R}^n\) is projection onto the first factor).

Claim 2: \(\sup \{ |Df(x)| : x\in K{\setminus } \pi (X)\} \le \sup \{|Df(x)|: x\in \partial K\}\).

For suppose not. Then for every small \(\epsilon >0\), the minimum distance between the graph of \(f|K\) and the graph of \((f+\epsilon )|K\) is attained at a pair of interior points \(p\in \text {graph}(f|K)\) and \(q\in \text {graph}((f+\epsilon )|K)\). Note that the tangent cone to \(N\) at \(p\) is contained in the halfspace \(\{v: v\times (q-p)\le 0\}\), which implies that the tangent cone is a plane, which by the standard regularity theory for area minimizing hypersurfaces implies that \(p\) is a regular point. Likewise \(q\) is a regular point. But now \(p\) and \(q\) violate Lemma 9. This completes the Proof of Claim 2.

Now Claim 2 implies that \(f\) is locally Lipschitz. But that implies (by standard PDE or by standard GMT) that \(f\) is smooth on \(W\). (For the GMT argument, consider the tangent cone at any point. By Claim 2, the tangent cone is the graph of a lipschitz function. Thus it suffices to show that there is no nonplanar minimal cone that is the graph of lipshitz function \(u\). By dimension reducing, we may assume that the cone is smooth away from the origin. Let \(\mathbf {v}\) be the unit vector for which \(f(\mathbf {v})\) is a maximum. Near \(\mathbf {v}\), the graph of \(f\) lies below the tangent plane at \((\mathbf {v},f(\mathbf {v}))\). Hence by the strong maximum principle, \(f\) is linear in a neighborhood of \(\mathbf {v}\). By analytic continuation, \(f\) is linear everywhere). \(\square \)

Appendix B: \(\kappa _1/h\) bounded below implies convex type

Theorem 11

Suppose \(t \in [0,\infty )\mapsto M(t)\) is a (possibly singular) mean curvature flow of mean convex \((n-1)\)-dimensional surfaces in a smooth \(n\)-dimensional Riemannian manifold. Let \(Z=(z,t)\) be a spacetime singular point of the flow with \(t>0\), and suppose that

$$\begin{aligned} \kappa _1/h \ge c > -\infty \end{aligned}$$
(*)

in a space-time neighborhood of \(Z\).

Then the singular point \(z\in M_t\) has convex type.

Proof

In [8], this (and all the results of that paper) are proved under the restriction:

  • \((\dag )\) For \(n>7\), we work in \(\mathbf {R}^n\) and we consider the flow only up to and including the first singular time.

In fact, the only place that restriction is explicitly used is in the proof of [8, Theorem 4]. In the proof of that theorem, the restriction \((\dag )\) is used only to show that \(\kappa _1/h\) is bounded below in a (space-time) neighborhood of the singularity in question up to and including the time \(t\). Thus, if we assume (as here) the hypothesis (*) that \(\kappa _1/h\) is bounded below in a neighborhood of \(Z\), then Theorem 4, and therefore all the other theorems of [8], hold at that singularity, without assuming the restriction \((\dag )\). In particular, the singularity is of convex type by [8, Theorem 1] and its corollary.

Incidentally, the assumption (*) that \(\kappa _1/h\) is bounded below in a spacetime neighborhood of the singularity (both before and after the singularity) significantly simplifies the proofs in [8] in various places. In particular, one can (throughout that paper) work with the class \(\mathcal {F}\) of all blowup flows (limit flows) at \(Z\) rather than the class of what are are called there “special limit flows”. For example, [8, Theorem 9] asserts that any each flow in \(\mathcal {F}\) is convex for all times \(\tau \le 0\). But the class of blowup flows at a space-time point is trivially closed under time translation, so it follows immediately that the restriction \(\tau \le 0\) is not necessary. (In [8], an additional argument involving an unpublished reference [W6] was given to remove the restriction \(\tau \le 0\)). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

White, B. Subsequent singularities in mean-convex mean curvature flow. Calc. Var. 54, 1457–1468 (2015). https://doi.org/10.1007/s00526-015-0831-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00526-015-0831-4

Mathematics Subject Classification

Navigation