The blowdown of ancient noncollapsed mean curvature flows
Abstract.
In this paper, we consider ancient noncollapsed mean curvature flows that do not split off a line. It follows from general theory that the blowdown of any time-slice, , is at most dimensional. Here, we show that the blowdown is in fact at most dimensional. Our proof is based on fine cylindrical analysis, which generalizes the fine neck analysis that played a key role in many recent papers. Moreover, we show that in the uniformly -convex case, the blowdown is at most dimensional. This generalizes the recent results from Choi-Haslhofer-Hershkovits [CHH21a] to higher dimensions, and also has some applications towards the classification problem for singularities in 3-convex mean curvature flow.
1. Introduction
In the analysis of mean curvature flow it is crucial to understand ancient noncollapsed flows. We recall that a mean curvature flow is called ancient if it is defined for all , and noncollapsed if it is mean-convex and there is an so that every point admits interior and exterior balls of radius at least , c.f. [SW09, And12, HK17] (in fact, by [Bre15, HK15] one can always take ). By the work of White [Whi00, Whi03, Whi15] and by [HH18] it is known that all blowup limits of mean-convex mean curvature flow are ancient noncollapsed flows. More generally, by Ilmanen’s mean-convex neighborhood conjecture [Ilm03], which has been proved recently in the case of neck-singularities in [CHH18, CHHW19], it is expected even without mean-convexity assumption that all blowup limits near any cylindrical singularity are ancient noncollapsed flows.
In this paper, we consider the asymptotic structure of ancient noncollapsed flows, specifically the blowdown of any time-slice. To describe this, denote by any ancient noncollapsed mean curvature flow. Recall that such flows are always convex thanks to [HK17, Theorem 1.10]. We can assume without essential loss of generality that the solution does not split off a line (otherwise we can reduce the problem to a lower dimensional one). Equivalently, this means that is strictly convex.
Definition 1.1 (blowdown).
Given any time , the blowdown of is defined by
(1.1) |
It is easy to see that the blowdown exists (in fact it is given as a decreasing intersection of convex sets) and is a cone, i.e. satsifies for all . The blowdown captures important information about the asymptotic structure of the time slices. By general theory the blowdown is at most dimensional. Indeed, the blowdown cannot be bigger than the singular set in a mean convex flow, which is at most dimensional by the deep theory of White [Whi00, Whi03] (see also [HK17]). Here, we improve this to:
Theorem 1.2 (blowdown of ancient noncollapsed flows).
Let , where , be an ancient noncollapsed mean curvature flow that does not split off a line. Then the blowdown of any time slice is at most dimensional, namely
(1.2) |
This generalizes the main result from [CHH21a], where Choi, Hershkovits and the second author ruled out the potential scenario ancient wing-like noncollapsed flows in . Our theorem shows that the blowdown is always at least dimension smaller than what one gets from the general bound for the dimension of the singular set. In particular, it implies:
Corollary 1.3 (nonexistence of ancient asymptotically conical flows).
There does not exist any ancient noncollapsed solution of the mean curvature flow in that is asymptotic to an -dimensional cone.
For comparison, we recall the basic result that there does not exist any nontrivial ancient noncollapsed flow in that is asymptotic to an -dimensional cone, as follows easily from the fact that the asymptotic area ratio vanishes [HK17]. Our corollary rules out solutions that are asymptotic to dimension cones, i.e. solutions asymptotic to cones whose dimension is smaller by than the dimension of the evolving hypersurfaces, which is much more subtle. It even rules out flows asymptotic to singular cones.
The dimension bound can be improved further in the uniformly -convex case. Recall that an ancient noncollapsed flow is uniformly -convex if there is a constant such that , where denote the principal curvatures. We prove:
Theorem 1.4 (blowdown of ancient -convex flows).
If we assume in addition the flow is uniformly -convex, where , then the blowdown of any time slice is at most dimensional.
Setting and relaxing the strict convexity assumption we obtain:
Corollary 1.5 (blowdown of ancient 3-convex flows).
Let be an ancient noncollapsed uniformly 3-convex mean curvature flow in . Then its blowdown is
-
•
either a point (which only happens if the solution is compact),
-
•
or a halfline,
-
•
or a line (which only happens for a round shrinking or an -dimensional rotationally symmetric oval times ),
-
•
or a -dimensional halfplane (which only happens if the solution is a -dimensional rotationally symmetric bowl soliton times ),
-
•
or a -dimensional plane (which only happens for a round shrinking ),
-
•
or an -dimensional hyperplane (which only happens for flat ).
Moreover, in the halfline case, the neutral eigenfunctions are dominant and the diameter of the level sets grows slower than for any .
Let us explain how the corollary follows: If the flow is strictly convex, then its blowdown is a halfline by our theorem (unless the solution is compact in which case the blowdown is simply a point). If the flow is not strictly convex, then after splitting off a line we get an ancient uniformly -convex noncollapsed flow, and such flows have been classified in the recent breakthrough by Angenent-Daskalopoulos-Sesum [ADS19, ADS20] and Brendle-Choi [BC19, BC]. Finally, the ‘moreover statement’ follows as consequence of our proofs as explained in Section 5 and Section 7.
As another corollary, for self-similarly translating solutions we obtain:
Corollary 1.6 (3-convex translators).
Any noncollapsed uniformly 3-convex translator in is -symmetric.
Indeed, any such translator is either an -dimensional rotationally symmetric bowl soliton times , or its blowdown is a halfline. The latter in particular implies that the translator has a unique tip point, and hence is -symmetric by a recent result of Zhu [Zhu21].
Remark 1.7 (classification problem for 3-convex translators).
Having established the above corollaries it seems likely that the arguments from [CHH21b, Section 3-5] generalize to higher dimensions to yield that every noncollapsed uniformly 3-convex translator in is either , or , or belongs to the one-parameter family of -symmetric oval-bowls constructed by Hoffman-Ilmanen-Martin-White [HIMW19, Theorem 8.1].
Finally, inspecting all the possibilities in Corollary 1.6 (blowdown of ancient 3-convex flows), as yet another corollary we obtain:
Corollary 1.8 (classification in unstable mode).
The only noncompact ancient noncollpased uniformly -convex flow, with a bubble-sheet tangent flow at , whose unstable mode is dominant is .
To prove our main theorems, we generalize the arguments from [CHH21a], by Choi, Hershkovits and the second author, to higher dimensions as follows:
In Section 2, we collect some basic properties of ancient noncollapsed flows in . In particular, we show that it is enough to analyze the case where the tangent flow at is given by
(1.3) |
with (respectively ).
In Section 3, to facilitate barrier and calibration arguments in later sections, we construct an symmetric foliation by rotating and shifting the dimensional shrinker foliation from [ADS19].
In Section 4, we set up a fine cylindrical analysis, which generalizes the fine neck analysis from [ADS19, ADS20, BC19, BC, CHH18, CHHW19] and the fine bubble sheet analysis from [CHH21a]. Given any space-time point , we consider the renormalized flow
(1.4) |
Then, the hypersurfaces converge for to the cylinder
(1.5) |
The analysis over is governed by the Ornstein-Uhlenbeck operator
(1.6) |
Applying the Merle-Zaag lemma, we see that either the unstable eigenfunctions, namely
(1.7) |
are dominant, or the neutral eigenfunctions, namely
(1.8) |
are dominant. Further, considering , where is a suitable rotation, we can kill the contribution from .
In Section 5, we analyze the case where the neutral eigenfunctions are dominant. We show that along a suitable sequence , the truncated graphical function of satisfies
(1.9) |
where , and is a nontrivial semi-negative definite -matrix. We then prove the crucial inclusion
(1.10) |
Roughly speaking, directions that are not in the kernel of the quadratic form are directions of inwards quadratic bending, and thus must be compact directions. We make this precise using the Brunn-Minkowski inequality. Once (1.10) is established, we conclude that for uniformly -convex flows.
In Section 6, we consider ancient noncollapsed flows whose unstable mode is dominant. We prove the fine cylindrical theorem, which says that there exists a nonvanishing vector with the following significance: For every space-time center , after suitable recentering in the -plane, the truncated graph function of the renormalized flow satisfies
(1.11) |
for depending only on an upper bound on the cylindrical scale of . To show this, we generalize the proofs of the fine neck theorem from [CHH18, CHHW19] and the fine bubble sheet theorem from [CHH21a].
In Section 7, we rule out the case where the unstable mode is dominant. We suppose towards a contradiction that there is an ancient noncollapsed uniformly -convex flow that does not split off a line, which has a -dimensional blowdown . We select two distinct rays of differentiability . Considering suitable corresponding sequences of points in , we can pass to limits that split off lines. Using the classification from Brendle-Choi [BC], we argue that must be -dimensional bowls times -dimensional planes. However, since these bowls translate in different directions, we obtain a contradiction with the fact that the fine cylindrical vector is independent of the space-time center point. This concludes the outline of the proof.
Acknowledgments. This research was supported by the NSERC Discovery Grant and the Sloan Research Fellowship of the second author.
2. Basic properties of ancient noncollapsed flows
Let be a noncollapsed mean curvature flow in , where , that is noncompact and strictly convex. Assume further that it is uniformly -convex, where (in case , there is no further assumption). It is not hard to see that the blowdown
(2.1) |
is in fact independent of the choice of prior to the extinction time (indeed, as one increases , by mean-convexity the blowdown cannot increase, and using suitable spheres as barriers, similarly as in [CHH21a, Section 2], one sees that it cannot decrease either). By [HK17, CM15] the tangent flow of at in suitable coordinates, is
(2.2) |
for some . Together with mean-convexity it follows that
(2.3) |
If , then and we are done. Hence, we can assume from now on that . Moreover, since our solution is strictly convex, does not contain any line.
Given any ancient noncollapsed mean curvature flow , that is not a round shrinking cylinder, any point and scale , we consider the flow
(2.4) |
that is obtained from by translating to the space-time origin and parabolically rescaling by .
Definition 2.1 (-cylindrical).
We say that is -cylindrical around at scale , if is -close in in to the evolution of a round shrinking cylinder with axis through the origin.
We fix a small enough parameter quantifying the quality of the cylinders for the rest of the paper. Given , we analyze the solution around at the diadic scales , where . Using Huisken’s monotonicity formula [Hui90] and quantitative differentiation (see e.g. [CHN13]), for every , we can find an integer such that
(2.5) |
and
(2.6) |
Definition 2.2 (cylindrical scale).
The cylindrical scale of is defined by
(2.7) |
3. Foliations and barriers
We recall from Angenent-Daskalopoulos-Sesum [ADS19] that there is some such that for every and , there are -dimensional shrinkers in ,
(3.1) | ||||
as illustrated in [ADS19, Fig. 1]. Here, the parameter captures where the concave functions meet the -axis, namely , and the parameter is the asymptotic slope of the convex functions , namely . We recall:
Lemma 3.1 (Lemmas 4.9 and 4.10 in [ADS19]).
There exists some such that the shrinkers and foliate the region
(3.2) |
Moreover, denoting by the outward unit normal of this family, we have
(3.3) |
We now write points in in the form , where and . We choose and shift and rotate the above foliation to construct a suitable foliation in :
Definition 3.2 (cylindrical foliation).
For every , we denote by the symmetric hypersurface in given by
(3.4) |
Similarly, for every , we denote by the symmetric hypersurface in given by
(3.5) |
Lemma 3.3 (Foliation lemma).
There exist and such that the hypersurfaces , , and the cylinder foliate the domain
Moreover, denoting by the outward unit normal of this foliation, we have
(3.6) |
and
(3.7) |
Proof.
Let be as in Lemma 3.1, and set . The fact that , and foliate is implied by Lemma 3.1 and Definition 3.2.
Now, observe that for every element in in the foliation of , we have
(3.8) |
By symmetry, it suffices to compute the curvatures of in the region , where we can identify points and unit normals in with the corresponding ones in , by disregarding the components. The relation between the mean curvature of a surface and its (unshifted) rotation on points with is given by
(3.9) |
where . For the shrinkers , the concavity of implies , so we infer that
(3.10) |
since in , we have .
For the shrinkers , the convexity of implies that , so similarly we have
(3.11) |
This proves the lemma. ∎
Corollary 3.4 (Inner Barriers).
Let be compact domains, whose boundary evolves by renormalized mean curvature flow. If is contained in for every , and for all , then
(3.12) |
Proof.
Lemma 3.3 implies that the vector points outwards of . The result now follows from the maximum principle. ∎
4. Setting up the fine cylindrical analysis
In this section, we set up the fine cylindrical analysis. This is similar to setting up fine bubble sheet analysis in [CHH21a, Section 4], so we will discuss this briefly. Throughout this section, let be an ancient noncollapsed flow in whose tangent flow at is a round shrinking cylinder . Assume further that is not self-similarly shrinking. Given any space-time point , we consider the renormalized flow
(4.1) |
Then, as , the hypersurfaces converges to the cylinder
(4.2) |
After rescaling, we can assume that , and we can find a universal function with
(4.3) |
so that for every with , the rotated surface is the graph of a function over with
(4.4) |
We fix a nonnegative smooth cutoff function satisfying for and for . We consider the truncated function
(4.5) |
where
(4.6) |
Using the implicit function theorem, similarly as in [CHH21a, Proposition 4.1], we can find a differentiable function , defined for sufficiently negative, such that for we have
(4.7) |
Moreover, we can arrange that for all , the matrix
(4.8) |
satisfies for , and for .
We now set
(4.9) |
where is the fine tuning rotation from above, and let
(4.10) |
so describes as a graph over the cylinder .
Proposition 4.1 (Inverse Poincare inequality).
The graph function satisfies the integral estimates
(4.11) |
and
(4.12) |
for all and , where is a numerical constant.
Proof.
Let be the Hilbert space of all functions on such that
(4.13) |
Proposition 4.2 (evolution equation and error estimate).
The truncated graph function satisfies the evolution equation
(4.14) |
where , and where the error term and the rotation satisfy the estimates
(4.15) |
for .
Proof.
The cylindrical representation of the operator is explicitly given by
(4.16) |
Analysing the spectrum of , the Hilbert space from (4.13) can be decomposed as
(4.17) |
where
(4.18) |
and
(4.19) |
and where denotes the restriction of the Euclidean coordinate functions to the cylinder . Moreover, we have
for , | ||||
(4.20) | for , | |||
for . |
Consider the functions
(4.21) | |||
where denote the orthogonal projections to , respectively.
Proposition 4.3 (Merle-Zaag alternative).
For , either the neutral mode is dominant, i.e.
(4.22) |
or the unstable mode is dominant, i.e.
(4.23) |
Proof.
5. Fine cylindrical analysis in the neutral mode
In this section, we prove our main theorems in the case where the neutral mode is dominant. Namely, throughout this section we assume that our truncated graph function satisfies
(5.1) |
Using this together with the evolution inequality (4.24), given any , we see that for sufficiently large , and thus
(5.2) |
for sufficiently large .
Proposition 5.1.
Every sequence converging to has a subsequence such that
(5.3) |
in -norm, where is a nontrivial semi-negative definite -matrix, and .
Proof.
Recalling (4.19) and using the orthogonality condition (4.7), we see that
(5.4) |
Therefore, every sequence has a subsequence such that
(5.5) |
with respect to the -norm, where , and
(5.6) |
for some nontrivial matrix . After an orthogonal change of coordinates in the -plane, we can assume that for . Let us show that (the same argument yields for ).
We denote by the region enclosed by and denote by the area of the cross section of with fixed coordinates. Explicitly,
(5.7) |
which can be expanded as
(5.8) |
By Brunn’s concavity principle, the function is concave. In particular, we have
(5.9) |
This implies
(5.10) |
Combining this with (5.5) and (5) yields
(5.11) |
This implies , and thus concludes the proof. ∎
Theorem 5.2.
If the neutral mode is dominant, then the blowdown satisfies
(5.12) |
for any from Proposition 5.1. In particular, .
Proof.
Let
(5.13) |
Given a unit vector in the -plane, let be an orthonormal basis of the -plane. Set . Since for , we have
(5.14) |
On the other hand, if , then for sufficiently large, we have
(5.15) |
Defining as in (5.7), similarly as in the previous proof we have
(5.16) |
as . Combining (5.14), (5.15) and (5) we see that for every sufficiently large, there exist such that
(5.17) |
Now, suppose towards a contradiction there is some . Since as , for all sufficiently large, we have
(5.18) |
where denotes the projection to the -plane. Set
(5.19) |
Take in the previous discussion, and let be an orthonromal basis of -plane. Note that
(5.20) |
On the other hand, by Brunn’s concavity principle, the function
(5.21) |
is concave, as long as it does not vanish. Together with (5.17), this implies that for all sufficiently large, the area of the cross sections is decreasing for , and vanishes at some finite . This contradicts (5.20), as the ray would have nowhere to go. This proves the theorem. ∎
Corollary 5.3.
Assume the solution is noncompact and uniformly -convex. Choosing coordinates such that , we have
(5.22) |
Furthermore, given any for the level-sets satisfy
(5.23) |
Proof.
Since a normalized semi-negative definite rank one quadratic form is uniquely determined by its kernel, we have convergence without the need of passing to a subsequence. Furthermore, arguing as in [CHH21a, Corollary 5.4] we get the bound for the diameter of the level sets. ∎
6. Fine cylindrical analysis in the unstable mode
In this section, we prove the fine cylindrical theorem for ancient noncollapsed flows whose tangent flow at is . This is similar to the fine neck theorem in [CHH18, CHHW19] and the fine bubble sheet theorem in [CHH21a], so we will discuss this rather briefly.
Throughout this section, we assume the tilted renormalized flow around some point , has a dominant unstable mode, i.e.
(6.1) |
Together with (4.24), this implies , hence
(6.2) |
for sufficiently large . Therefore, Proposition 4.2 yields
(6.3) |
We now consider the untilted renormalized flow around any center . For . we can write as a graph of a function over , where satisfies (4.3) and (4.4). We will work with the truncated function
(6.4) |
Moreover, we set
(6.5) | |||
Applying the Merle-Zaag ODE lemma [MZ98, CM19], we infer that there are universal constants , such that for every either
(6.6) |
or
(6.7) |
whenever .
Lemma 6.1 (dominant mode).
The unstable mode is dominant for the untilted renormalized flow with any center , i.e. (6.7) holds for all .
Proof.
Let us first show that the statement holds for . Combining (6.2) and (6.3), we infer that the untilted flow satisfies
(6.8) |
If we had , then arguing as at the beginning of Section 5 we would see that for every and sufficiently large, contradicting (6.8). Thus, for , we indeed get (6.7). Finally, since any neck centered at a general point merges with the neck centered at as , the inequality (6.7) holds for every . ∎
Proposition 6.2 (improved graphical radius).
Proof.
From now on, we work with from Proposition 6.2 (improved graphical radius), and in particular, we define with respect to this improved graphical radius. Note that the unstable mode is still dominant.
Proposition 6.3 (sharp decay estimate).
There exist constants and , depending only on an upper bound for , such that for :
(6.10) |
and
(6.11) |
Proof.
Since is an -graph over , we get . Together with the evolution inequality , this implies
(6.12) |
for , where and only depend on .
Theorem 6.4 (fine cylindrical theorem).
If is an ancient noncollapsed flow in , whose tangent flow at is , and whose unstable mode is dominant, then there exist constants with with the following significance. For every center point , the truncated graphical function of the renormalized flow satisfies the fine cylindrical expansion estimates
(6.14) |
and
(6.15) |
for , where for , and where and only depend on an upper bound for the cylindrical scale .
Proof.
In the following and denote constants that might change from line to line, but only depend on an upper bound for the cylindrical scale . Recalling the basis of from (4.18), we can write
(6.16) |
where the superscript indicates (a priori) dependence on the center .
Letting , and using , we compute
(6.17) |
Hence, using Proposition 4.2, Proposition 6.2 and Proposition 6.3, we obtain
(6.18) |
Integrating this from to implies
(6.19) |
In a similar manner, using for , we get
(6.20) |
so integrating from to yields
(6.21) |
where
(6.22) |
Now, consider the difference
(6.23) |
Using the above, we see that
(6.24) |
Together with the inequality , which follows from Lemma 6.1, Proposition 6.2 and Proposition 6.3, this yields
(6.25) |
which proves (6.14) modulo the claim about the coefficients.
Next, we observe that
(6.26) |
and, using Proposition 6.3, that
(6.27) |
Applying Agmon’s inequality, this yields
(6.28) |
Now, comparing the renormalized flows with center and center , we need to relate the parameters of the functions and . Since it is easy to see that the parameters do not depend on , we may choose . We then have
(6.29) |
where we use the notation and as before. By Taylor expansion and Proposition 6.3, we have
(6.30) |
Together with (6.28), the above formulas imply that for , and for .
7. Conclusion of the proof
In this final section, we conclude the proof of our main theorem:
Theorem 7.1.
Let be an ancient noncollapsed mean curvature flow that does not split off a line. Asume that the flow is uniformly -convex for some . Then the blowdown of any time slice is at most dimensional, namely
(7.1) |
Proof.
Let be an ancient noncollapsed uniformly -convex mean curvature flow that does not split off a line. By the reduction from Section 2, we can assume that
(7.2) |
where . Considering the expansion of the renormalized flow over the cylinder , by Proposition 4.3 (Merle-Zaag alternative) and Lemma 6.1 (dominant mode), for , either the neutral mode dominates for the tilted flow, or the unstable mode dominates for the untilted flow with respect to any center. In the dominant neutral mode case, we have shown in Theorem 5.2 that the blowdown is at most dimensional. We can thus assume that we are in the dominant unstable mode case. Since we already know that , and since the dimension of convex sets is always an integer, it suffices to rule out the scenario that .
Suppose towards a contradiction that . Recall that does not contain any line and also recall that . Choosing suitable coordinates, we can assume that is contained in and contains the positive -axis in its interior.
Claim 7.2 (points of differentiability).
There exists a two-dimensional plane containing the -axis, such that all points are twice differentiable points of .
Proof of Claim 7.2.
Observe first that since is a cone, whenever there is some point that is not twice differentiable, then in fact contains a -dimensional ray of points that are not twice differentiable.
Now suppose towards a contradiction that for every two-plane containing the -axis, there is some point that is not twice differentiable. Since the space of two-planes containing the -axis is -dimensional, together with the above it follows that the set of points in that are not twice differentiable has positive -dimensional Hausdorff measure. This contradicts Alexandrov’s theorem. ∎
Rotating coordinates, we can assume that is the -plane. Observe that consists of two rays with directional vectors satisfying and , respectively. By a space-time translation, we may also assume that and that is the point in with smallest -value. Now, for every , let be a point which maximizes/minimizes the value of in .
Claim 7.3 (cylindrical scale).
There exists some constant such that
(7.3) |
Proof of the Claim 7.3.
We will argue similarly as in the proofs of [CHH21a, Claim 7.2], [CHH18, Proposition 5.8] and [CHHW19, Proposition 6.2].
Suppose towards a contradiction that for some sequence . Let be the sequence of flows obtained by shifting to the origin, and parabolically rescaling by . By [HK17, Thm. 1.14] we can pass to a subsequential limit , which is an ancient noncollapsed flow that is weakly convex and smooth until it becomes extinct. Note also that has an tangent flow at .
Note that cannot be a round shrinking . Indeed, if such a cylinder became extinct at time , this would contradict the definition of the cylindrical scale, and if it became extinct at some later time, this would contradict the fact that is a strict subset of by construction.
Thus, by Proposition 4.3 (Merle-Zaag alternative), for the flow , either the neutral mode is dominant or the unstable mode is dominant. If the neutral mode is dominant, then for large we obtain a contradiction with the fact that has dominant unstable mode, using in particular equation (5.2). If the unstable mode is dominant, then by the fine cylindrical theorem (Theorem 6.4), the limit has some nonvanishing fine cylindrical vector . However, this contradicts the fact that the fine cylindrical vector of is obtained from the fine cylindrical vector of by scaling by . This concludes the proof of the claim. ∎
Take and consider the sequence . By Claim 7.3 (cylindrical scale), any subsequential limit is an ancient noncollapsed flow with an tangent flow at . Moreover, arguing as in the proof of the claim, we see that has a dominant unstable mode, with the same fine cylindrical vector as our original flow .
Claim 7.4 (splitting off lines).
The hypersurfaces and contain -dimensional planes .
Proof of Claim 7.4.
First observe that since , the hypersurfaces and contain a line in direction . Thus, it remains to find lines in nonradial directions.
By the definition of the blowdown, we have the Hausdorff convergence
(7.4) |
This implies
(7.5) |
where is a twice differentiable point of by Claim 7.2.
Now, let be the inwards unit normal of at , and consider any unit tangent vector . Since is twice differentiable at , we have
(7.6) |
for small enough. By convexity, the segments from to and are contained in . Note also that these segments have length at least and meet at angle , where as . Together with the Hausdorff convergence from above, we see that for large enough the segments connecting to and are contained in , have length at least and meet at angle . Taking , and sending slowly enough, we conclude that contains a line in direction . Since was arbitrary, this completes the proof of the claim. ∎
By Claim 7.4 (splitting off lines), we get , where is an ancient noncollapsed uniformly -convex flow. Using the classification by Brendle-Choi [BC], we infer that must be a translating bowl soliton. By inspection, we see that the fine cylindrical vector of points in the direction of translation. However, since and translate in different directions, this contradicts the fact that and have the same fine cylindrical vector. This concludes the proof of the theorem. ∎
References
- [ADS19] S. Angenent, P. Daskalopoulos, and N. Sesum. Unique asymptotics of ancient convex mean curvature flow solutions. J. Differential Geom., 111(3):381–455, 2019.
- [ADS20] S. Angenent, P. Daskalopoulos, and N. Sesum. Uniqueness of two-convex closed ancient solutions to the mean curvature flow. Ann. of Math. (2), 192(2):353–436, 2020.
- [And12] B. Andrews. Noncollapsing in mean-convex mean curvature flow. Geom. Topol., 16(3):1413–1418, 2012.
- [BC] S. Brendle and K. Choi. Uniqueness of convex ancient solutions to mean curvature flow in higher dimensions. Geom. Top. (to appear).
- [BC19] S. Brendle and K. Choi. Uniqueness of convex ancient solutions to mean curvature flow in . Invent. Math., 217(1):35–76, 2019.
- [Bre15] S. Brendle. A sharp bound for the inscribed radius under mean curvature flow. Invent. Math., 202(1):217–237, 2015.
- [CHH18] K. Choi, R. Haslhofer, and O. Hershkovits. Ancient low entropy flows, mean convex neighborhoods, and uniqueness. arXiv:1810.08467, 2018.
- [CHH21a] K. Choi, R. Haslhofer, and O. Hershkovits. A nonexistence result for wing-like mean curvature flows in . arXiv:2105.13100, 2021.
- [CHH21b] K. Choi, R. Haslhofer, and O. Hershkovits. Classification of noncollapsed translators in . preprint, 2021.
- [CHHW19] K. Choi, R. Haslhofer, O. Hershkovits, and B. White. Ancient asymptotically cylindrical flows and applications. arXiv:1910.00639, 2019.
- [CHN13] J. Cheeger, R. Haslhofer, and A. Naber. Quantitative stratification and the regularity of mean curvature flow. Geom. Funct. Anal., 23(3):828–847, 2013.
- [CM15] T. Colding and W. Minicozzi. Uniqueness of blowups and Lojasiewicz inequalities. Ann. of Math. (2), 182(1):221–285, 2015.
- [CM19] K. Choi and C. Mantoulidis. Ancient gradient flows of elliptic functionals and Morse index. arXiv:1902.07697, 2019.
- [HH18] R. Haslhofer and O. Hershkovits. Singularities of mean convex level set flow in general ambient manifolds. Adv. Math., 329:1137–1155, 2018.
- [HIMW19] D. Hoffman, T. Ilmanen, F. Martín, and B. White. Graphical translators for mean curvature flow. Calc. Var. Partial Differential Equations, 58(4):Art. 117, 29, 2019.
- [HK15] R. Haslhofer and B. Kleiner. On Brendle’s estimate for the inscribed radius under mean curvature flow. Int. Math. Res. Not. IMRN, (15):6558–6561, 2015.
- [HK17] R. Haslhofer and B. Kleiner. Mean curvature flow of mean convex hypersurfaces. Comm. Pure Appl. Math., 70(3):511–546, 2017.
- [Hui90] G. Huisken. Asymptotic behavior for singularities of the mean curvature flow. J. Differential Geom., 31(1):285–299, 1990.
- [Ilm03] T. Ilmanen. Problems in mean curvature flow. https://people.math.ethz.ch/ ilmanen/classes/eil03/problems03.ps, 2003.
- [MZ98] F. Merle and H. Zaag. Optimal estimates for blowup rate and behavior for nonlinear heat equations. Comm. Pure Appl. Math., 51(2):139–196, 1998.
- [SW09] W. Sheng and X. Wang. Singularity profile in the mean curvature flow. Methods Appl. Anal., 16(2):139–155, 2009.
- [Whi00] B. White. The size of the singular set in mean curvature flow of mean convex sets. J. Amer. Math. Soc., 13(3):665–695, 2000.
- [Whi03] B. White. The nature of singularities in mean curvature flow of mean-convex sets. J. Amer. Math. Soc., 16(1):123–138, 2003.
- [Whi15] B. White. Subsequent singularities in mean-convex mean curvature flow. Calc. Var. Partial Differential Equations, 54(2):1457–1468, 2015.
- [Zhu21] J. Zhu. Rotational symmetry of uniformly 3-convex translating solitons of mean curvature flow in higher dimensions. arXiv:2103.16382, 2021.
Wenkui du, Department of Mathematics, University of Toronto, 40 St George Street, Toronto, ON M5S 2E4, Canada
Robert Haslhofer, Department of Mathematics, University of Toronto, 40 St George Street, Toronto, ON M5S 2E4, Canada
E-mail: [email protected], [email protected]