Classification of convex ancient free boundary curve shortening flows in the disc.
Abstract.
We classify convex ancient curve shortening flows in the disc with free boundary on the circle.
1. Introduction
Curve shortening flow is the gradient flow of length for regular curves. It models the evolution of grain boundaries [19, 23] and the shapes of worn stones [11] in two dimensions, and has been exploited in a multitude of further applications (see, for example, [20]).
The evolution of closed planar curves by curve shortening was initiated by Mullins [19] and was later taken up by Gage [13] and Gage–Hamilton [12], who proved that closed convex curves remain convex and shrink to “round” points in finite time. Soon after, Grayson showed that closed embedded planar curves become convex in finite time under the flow, thereafter shrinking to round points according to the Gage–Hamilton theorem. Different proofs of these results were discovered later by others [1, 2, 3, 16, 18].
Ancient solutions to geometric flows (that is, solutions defined on backwards-infinite time-intervals) are important from an analytical standpoint as they model singularity formation [15]. They also arise in quantum field theory, where they model the ultraviolet regime in certain Dirichlet sigma models [5]. They have generated a great deal of interest from a purely geometric standpoint due to their symmetry and rigidity properties. Indeed, ancient solutions to curve shortening flow of convex planar curves have been classified through the work of Daskalopoulos–Hamilton–Šešum [10] and the authors in collaboration with Tinaglia [6]. Bryan and Louie [7] proved that the shrinking parallel is the only convex ancient solution to curve shortening flow on the two-sphere, and Choi and Mantoulidis showed that it is the only embedded ancient solution on the two-sphere with uniformly bounded length [8].
The natural Neumann boundary value problem for curve shortening flow, called the free boundary problem, asks for a family of curves whose endpoints lie on (but are free to move on) a fixed barrier curve which is met by the solution curve orthogonally. Study of the free boundary problem was initiated by Huisken [17] and further developed by Stahl [21, 22]. In particular, Stahl proved that convex curves with free boundary on a smooth, convex, locally uniformly convex barrier remain convex and shrink to a point on the barrier curve.
The analysis of ancient solutions to free boundary curve shortening flow remains in its infancy. Indeed, to our knowledge, the only examples previously known seem to be those inherited from closed or complete examples (one may restrict the shrinking circle, for example, to the upper halfplane).
We provide here a classification of convex111A free boundary curve in the open disc is convex if it bounds a convex region in and locally uniformly convex if it is of class and its curvature is positive. ancient free boundary curve shortening flows in the disc.
Theorem 1.1.
Modulo rotation about the origin and translation in time, there exists exactly one convex, locally uniformly convex ancient solution to free boundary curve shortening flow in the disc. It converges to the point as and smoothly to the segment as . It is invariant under reflection across the -axis. As a graph over the -axis, it satisfies
for some , where is the solution to .
Theorem 1.1 is a consequence of Propositions 2.8, 3.4, and 3.5 proved below. Note that it is actually a classification of all convex ancient solutions, since the strong maximum principle and the Hopf boundary point lemma imply that any convex solution to the flow is either a stationary segment (and hence a bisector of the disc by the free boundary condition) or is locally uniformly convex at interior times.
A higher dimensional counterpart of Theorem 1.1 will be treated in a forthcoming paper.
Another natural setting in which to seek ancient solutions is within the class of soliton solutions. Since free boundary curve shortening flow in the disc is invariant under ambient rotations, one might expect to find rotating solutions. In §4, we provide a short proof that none exist.
Theorem 1.2.
There exist no proper rotating solutions to free boundary curve shortening flow in the disc.
Acknowledgements
We wish to thank Jonathan Zhu for sharing his thoughts on the problem.
TB was supported through grant 707699 of the Simons Foundation and grant DMS-2105026 of the National Science Foundation. ML was supported through an Australian Research Council DECRA fellowship (grant DE200101834).
2. Existence
Our first goal is the explicit construction of a non-trivial ancient free boundary curve shortening flow in the disc. It will be clear from the construction that the solution is reflection symmetric about the vertical axis, emerges at time negative infinity from the horizontal bisector, and converges at time zero to the point . We shall also prove an estimate for the height of the constructed solution (which will be needed to prove its uniqueness).
2.1. Barriers
Given , denote by the circle centred on the -axis which meets orthogonally at . That is,
(1) |
If we set
then we find that the inward normal speed of is no greater than its curvature and the inward normal speed of is no less than its curvature . The maximum principle and the Hopf boundary point lemma then imply that
Proposition 2.1.
a solution to free boundary curve shortening flow in which lies below (resp. above) the circle at time lies below (resp. above ) for all , where (resp. ).
Consider now the shifted scaled Angenent oval , where
This evolves by curve shortening flow and satisfies
at a point .
Lemma 2.2.
For each , there is a unique such that
Given with ,
Proof.
Define
Observe that
and
(2) |
The first claim follows.
Next observe that
Given , we obtain, at the unique zero of ,
Since for , this is less than zero. The second claim follows. ∎
The maximum principle and the Hopf boundary point lemma now imply the following.
Proposition 2.3.
Let be a solution to free boundary curve shortening flow in . Suppose that , where denotes the smaller, in absolute value, of the two turning angles to at its boundary. If lies above , then lies above for all .
Proof.
By the strong maximum principle, the two families of curves can never develop contact at an interior point. Since the families are monotonic, they cannot develop boundary contact at a boundary point with . On the other hand, since , (2) implies that
and hence, by the argument of Lemma 2.2,
So the Hopf boundary point lemma implies that no boundary contact can develop for either. ∎
Remark 2.4.
Since as , is non-negative at so long as , where .
2.2. Old-but-not-ancient solutions
For each , choose a curve in with the following properties:
-
•
meets orthogonally at ,
-
•
is reflection symmetric about the -axis,
-
•
is the relative boundary of a convex region , and
-
•
in .
For example, we could take , where and are (uniquely) chosen so that
and
Observe that the circle defined by
is tangent to the line , and hence lies above .
Lemma 2.5.
For each , there exists a smooth solution222Given by a one parameter family of immersions satisfying for some . to curve shortening flow with which satisfies the following properties:
-
•
meets orthogonally for each ,
-
•
is convex and locally uniformly convex for each ,
-
•
is reflection symmetric about the -axis for each ,
-
•
uniformly as ,
-
•
in , and
-
•
as .
Proof.
Existence of a maximal solution to curve shortening flow out of which meets orthogonally was proved by Stahl [22, Theorem 2.1]. Stahl also proved that this solution remains convex and locally uniformly convex and shrinks to a point on the boundary of at the final time (which is finite) [21, Proposition 1.4]. We obtain by time-translating Stahl’s solution.
By uniqueness of solutions remains reflection symmetric about the -axis for , so the final point is .
The reflection symmetry also implies that at the point for all . By [21, Proposition 2.1], at the boundary point for all . Applying Sturm’s theorem [4] to , we thus find that on for all .
Since , the final property follows from Proposition 2.1. ∎
We now fix and drop the super/subscript . Set
and define , and by
Lemma 2.6.
Each old-but-not-ancient solution satisfies
(3) |
(4) |
and
(5) |
Proof.
To prove the lower bound for , it suffices to show that the circle (see (1)) lies locally below near . If this is not the case, then, locally around , lies below and hence . But then we can translate downwards until it touches from below in an interior point at which the curvature must satisfy . This contradicts the unique maximization of the curvature at .
The estimate (4) now follows by integrating the inequality
between any initial time and the final time (at which since the solution contracts to the point ).
The upper bound for follows from convexity and the boundary condition . To prove the lower bound, we will show that the circle lies nowhere above . Suppose that this is not the case. Then, since lies locally below near , we can move downwards until it is tangent from below to a point on , at which we must have . But then, since in , we find that for all points between and . But this implies that this whole arc (including ) lies above , a contradiction.
To prove the upper bound for , fix and consider the circle centred on the -axis through the points and . Its radius is , where
We claim that lies locally below near . Suppose that this is not the case. Then, by the symmetry of and across the -axis, lies locally above near . This implies two things: first, that
and second, that, by moving vertically upwards, we can find a point (the final point of contact) which satisfies
These two inequalities contradict the (unique) minimization of at . We conclude that
due to the lower bund for . ∎
Remark 2.7.
If we parametrize by turning angle , so that
then the estimates (3) are also easily obtained from the monotonicity of and the formulae
(6) |
2.3. Taking the limit
Proposition 2.8.
There exists a convex, locally uniformly convex ancient curve shortening flow in the disc with free boundary on the circle.
Proof.
For each , consider the old-but-not-ancient solution , , constructed in Lemma 2.5. By (4), contains , where is uniquely defined by
If we represent as a graph over the -axis, then convexity and the boundary condition imply that . Since is independent of , Stahl’s (global in space, interior in time) Ecker–Huisken type estimates [22] imply uniform-in- bounds for the curvature and its derivatives. So the limit
exists in (globally in space on compact subsets of time) and the limit satisfies curve shortening flow with free boundary in . On the other hand, since contracts to as , (the contrapositive of) Proposition 2.1 implies that must intersect the closed region enclosed by for all . It follows that must intersect the closed region enclosed by for all . Since each is the limit of convex boundaries, each is convex. It follows that converges to as and, by [22, Corollary 4.5], that is locally uniformly convex for each . ∎
2.4. Asymptotics for the height
For the purposes of this section, we fix an ancient solution obtained as in Proposition 2.8 by taking a sublimit as of the specific old-but-not ancient solutions corresponding to , being the time at which meets orthogonally. The asymptotics we obtain for this solution will be used to prove its uniqueness.
We will need to prove that the limit exists in . The following speed bound will imply that it exists in .
Lemma 2.9.
The ancient solution satisfies
(7) |
Proof.
It suffices to prove that on each of the old-but-not-ancient solutions . Note that equality holds on the initial timeslice .
Given any , set and . Observe that
At an interior maximum of we observe that
and hence
(8) |
At a (without loss of generality right) boundary maximum of , we have and , and hence
(9) |
We may now conclude that remains less than one. Indeed, if ever reaches , then there must be a first time and a point at which this occurs (note that is continuous on up to the initial time). The point cannot be an interior point, due to (8), and it cannot be a boundary point, due to (9) and the Hopf boundary point lemma. We conclude that
on for all . Now take . ∎
In particular, the limit
exists in for each , as claimed.
We next prove that the limit is positive. The following lemma will be used to prove the requisite speed bound.
Lemma 2.10.
There exist and such that
(11) |
Proof.
We will prove the estimate for each old-but-not-ancient solution . We first prove a crude gradient estimate of the form
(12) |
for sufficiently negative. It will suffice to prove that
(13) |
Indeed, since has the same sign as the -coordinate, we may estimate, as in (7),
(14) |
For sufficiently close to , we have . Denote by the first time at which reaches . Since is continuous up to the initial time , we have . We claim that (13) holds for . Indeed, it is satisfied on the initial timeslice since
whereas . We will show that
remains negative up to time . Suppose, to the contrary, that reaches zero at some time at some point . Since vanishes at the boundary, must be an interior point. Since vanishes at the -axis, and the curve is symmetric, we may assume that . At such a point,
Recalling (14) and estimating yields
which is absurd. So does indeed remain negative, and taking yields (12) for .
Lemma 2.11.
There exist and such that
Proof.
Consider the old-but-not-ancient solution . By (11), we can find and such that
Since, at a boundary point,
the Hopf boundary point lemma and the ode comparison principle yield
But now
and hence, by ode comparison,
Since, on the initial timeslice ,
the claim follows upon taking . ∎
It follows that
and hence, integrating from time up to time ,
So we indeed find that
Lemma 2.12.
the limit
(15) |
exists in on the particular ancient solution .
3. Uniqueness
Now let , , be any convex, locally uniformly convex ancient free boundary curve shortening flow in the disc. By Stahl’s theorem [21], we may assume that contracts to a point on the boundary as .
3.1. Backwards convergence
We first show that converges to a bisector as .
Lemma 3.1.
Up to a rotation of the plane,
Proof.
Set . Integrating the variational formula for area yields
where is the turning angle. Since convexity ensures that the total turning angle is increasing and for all , we find that
Monotonicity of the flow, the free boundary condition and convexity now imply that the enclosed regions satisfy
in the Hausdorff topology.
If we now represent graphically over the -axis, then convexity and the boundary condition ensure that the height and gradient are bounded by the height at the boundary. Stahl’s estimates [22] now give bounds for and its derivatives up to the boundary depending only on the height at the boundary. We then get smooth subsequential convergence along any sequence of times . The claim follows since any sublimit is the horizontal segment. ∎
We henceforth assume, without loss of generality, that the backwards limit is the horizontal bisector.
3.2. Reflection symmetry
We can now prove that the solution is reflection symmetric using Alexandrov reflection across lines through the origin (see Chow and Gulliver [9]).
Lemma 3.2.
is reflection symmetric about the -axis for all .
Proof.
Given any , we define the halfspace
and denote by the reflection about . We first claim that, for every , there exists such that
Assume that the claim is not true. Then there exists , a sequence of times , and a sequence of pairs of points such that . This implies that the line passing through and is parallel to the vector , so the mean value theorem yields for each a point on where the normal is parallel to . This contradicts Lemma 3.1. ∎
3.3. Asymptotics for the height
We begin with a lemma.
Lemma 3.3.
For all ,
and hence
(16) |
Proof.
Choose so that for and, given , set
We claim that in . Suppose that this is not the case. Since at the spatial boundary , and as , there must exist a first time in and an interior point at which . But, at such a point,
which is absurd. Now take to obtain in for . Since at the -axis and at the right boundary point, the strong maximum principle and the Hopf boundary point lemma imply that in for . But then Sturm’s theorem implies that does not develop additional zeroes up to time .
Having established the first claim, the second follows as in Lemma 2.6. ∎
Proposition 3.4.
If we define as in (15), then
Proof.
Given , consider the rescaled height function
which is defined on the time-translated flow , where . Note that
(17) |
where and are the gradient and Laplacian on , respectively, and is the outward unit normal to .
Since reaches the origin at time zero, it must intersect the constructed solution for all . In particular, the value of on the former can at no time exceed the value of on the latter. But then (15) and (16) yield
(18) |
This implies a uniform bound for on for any . So Alaoglu’s theorem yields a sequence of times such that converges in the weak∗ topology as to some . Since convexity and the boundary condition imply a uniform bound for on any time interval of the form , we may also arrange that the convergence is uniform in space at time zero, say.
Weak∗ convergence ensures that satisfies the problem
(19) |
Indeed, a smooth function satisfies the boundary value problem (17) (and analogously for (19)) if and only if
for all smooth which are compactly supported in time and satisfy
where is the formal -adjoint of the heat operator. Since converges uniformly in the smooth topology to the stationary interval as , we conclude that the limit must satisfy (19) in the sense (and hence in the classical sense due to the theory for the heat equation). Indeed, by the definition of smooth convergence, we may (after possibly applying a diffeomorphism) parametrize each flow over by a family of embeddings which converge in as to the stationary embedding . Given satisfying , set , where is defined by
That is, , where . This ensures that at the boundary, and hence
Since in , a short computation reveals that
Finally, we characterize the limit (uniqueness of which implies full convergence, completing the proof). Separation of variables leads us to consider the problem
There is only one negative eigenspace, and its frequency turns out to be , with the corresponding mode given by
Thus, recalling (18), we are able to conclude that
for some . In particular,
Now, if is not equal to the corresponding value on the constructed solution (note that the full limit exists for the latter), then one of the two solutions must lie above the other at time for sufficiently large. But this violates the avoidance principle. ∎
3.4. Uniqueness
Uniqueness of the constructed ancient solution now follows directly from the avoidance principle.
Proposition 3.5.
Modulo time translation and rotation about the origin, there is only one convex, locally uniformly convex ancient solution to free boundary curve shortening flow in the disc.
Proof.
Denote by the constructed ancient solution and let be a second ancient solution which, without loss of generality, contracts to the point at time . Given any , consider the time-translated solution defined by . By Proposition 3.4,
uniformly in . So lies above for sufficiently large. The avoidance principle then ensures that lies above for all . Taking , we find that lies above for all . Since the two curves reach the point at time zero, they intersect for all by the avoidance principle. The strong maximum principle then implies that the two solutions coincide for all . ∎
4. Supplement: nonexistence of rotators
Free boundary curve shortening flow in is invariant under rotations about the origin, so it is natural to seek solutions which move by rotation; that is, solutions satisfying
for some . Differentiating yields the rotator equation
(20) |
It turns out, however, that there are no solutions to (20) in satisfying the free boundary condition.
Proof of Theorem 1.2.
Following Halldorsson [14], we rewrite the rotator equation as the pair of ordinary differential equations
(21) |
where
Arc-length parametrized solutions to the rotator equation (20) can be recovered from solutions to the system (21) via
and this parametrization is unique up to an ambient rotation and a unit linear reparametrization, i.e. .
Let be such a solution. Since (21) can be uniquely solved with initial condition (which corresponds to with ), we find that must be invariant under rotation by about the origin. In particular, the points and are diametrically opposite. It follows that is the origin. Indeed, for topological reasons, must cross the line orthogonally bisecting the segment joining its endpoints an odd number of times (with multiplicity). But since the rotational invariance pairs each crossing above the origin with one below, we are forced to include the origin in the set of crossings. We conclude that
which is impossible since . This completes the proof. ∎
References
- [1] Andrews, B. Noncollapsing in mean-convex mean curvature flow. Geom. Topol. 16, 3 (2012), 1413–1418.
- [2] Andrews, B., and Bryan, P. A comparison theorem for the isoperimetric profile under curve-shortening flow. Comm. Anal. Geom. 19, 3 (2011), 503–539.
- [3] Andrews, B., and Bryan, P. Curvature bound for curve shortening flow via distance comparison and a direct proof of Grayson’s theorem. J. Reine Angew. Math. 653 (2011), 179–187.
- [4] Angenent, S. The zero set of a solution of a parabolic equation. J. Reine Angew. Math. 390 (1988), 79–96.
- [5] Bakas, I., and Sourdis, C. Dirichlet sigma models and mean curvature flow. Journal of High Energy Physics 2007, 06 (2007), 057.
- [6] Bourni, T., Langford, M., and Tinaglia, G. Convex ancient solutions to curve shortening flow. Calc. Var. Partial Differential Equations 59, 4 (2020), 133.
- [7] Bryan, P., and Louie, J. Classification of convex ancient solutions to curve shortening flow on the sphere. J. Geom. Anal. 26, 2 (2016), 858–872.
- [8] Choi, K., and Mantoulidis, C. Ancient gradient flows of elliptic functionals and Morse index. Preprint available at arXiv:1902.07697.
- [9] Chow, B., and Gulliver, R. Aleksandrov reflection and geometric evolution of hypersurfaces. Comm. Anal. Geom. 9, 2 (2001), 261–280.
- [10] Daskalopoulos, P., Hamilton, R., and Sesum, N. Classification of compact ancient solutions to the curve shortening flow. J. Differential Geom. 84, 3 (2010), 455–464.
- [11] Firey, W. J. Shapes of worn stones. Mathematika 21 (1974), 1–11.
- [12] Gage, M., and Hamilton, R. The heat equation shrinking convex plane curves. J. Differ. Geom. 23 (1986), 69–96.
- [13] Gage, M. E. Curve shortening makes convex curves circular. Invent. Math. 76, 2 (1984), 357–364.
- [14] Halldorsson, H. P. Self-similar solutions to the curve shortening flow. Trans. Amer. Math. Soc. 364, 10 (2012), 5285–5309.
- [15] Hamilton, R. S. The formation of singularities in the Ricci flow. In Surveys in differential geometry, Vol. II (Cambridge, MA, 1993). Int. Press, Cambridge, MA, 1995, pp. 7–136.
- [16] Hamilton, R. S. Isoperimetric estimates for the curve shrinking flow in the plane. In Modern methods in complex analysis (Princeton, NJ, 1992), vol. 137 of Ann. of Math. Stud. Princeton Univ. Press, Princeton, NJ, 1995, pp. 201–222.
- [17] Huisken, G. Nonparametric mean curvature evolution with boundary conditions. J. Differential Equations 77, 2 (1989), 369–378.
- [18] Huisken, G. A distance comparison principle for evolving curves. Asian J. Math. 2, 1 (1998), 127–133.
- [19] Mullins, W. W. Two-dimensional motion of idealized grain boundaries. J. Appl. Phys. 27 (1956), 900–904.
- [20] Sapiro, G. Geometric partial differential equations and image analysis. Cambridge University Press, Cambridge, 2001.
- [21] Stahl, A. Convergence of solutions to the mean curvature flow with a Neumann boundary condition. Calc. Var. Partial Differential Equations 4, 5 (1996), 421–441.
- [22] Stahl, A. Regularity estimates for solutions to the mean curvature flow with a Neumann boundary condition. Calc. Var. Partial Differential Equations 4, 4 (1996), 385–407.
- [23] von Neumann, J. In Metal Interfaces. American Society for Testing Materials, Cleveland, 1952, p. 108.