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

Classification of convex ancient free boundary curve shortening flows in the disc.

Theodora Bourni  and  Mat Langford Department of Mathematics, University of Tennessee Knoxville, Knoxville TN, 37996-1320 [email protected], [email protected] School of Mathematical and Physical Sciences, The University of Newcastle, Newcastle, NSW, Australia, 2308 [email protected]
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 B2B^{2} is convex if it bounds a convex region in B2B^{2} and locally uniformly convex if it is of class C2C^{2} 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 (0,1)(0,1) as t0t\to 0 and smoothly to the segment [1,1]×{0}[-1,1]\times\{0\} as tt\to-\infty. It is invariant under reflection across the yy-axis. As a graph over the xx-axis, it satisfies

eλ2ty(x,t)Acosh(λx)uniformly in x ast\mathrm{e}^{\lambda^{2}t}y(x,t)\to A\cosh(\lambda x)\;\;\text{uniformly in $x$ as}\;\;t\to-\infty

for some A>0A>0, where λ\lambda is the solution to λtanhλ=1\lambda\tanh\lambda=1.

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 (0,1)(0,1). 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 θ(0,π2)\theta\in(0,\frac{\pi}{2}), denote by Cθ\mathrm{C}_{\theta} the circle centred on the yy-axis which meets B2\partial B^{2} orthogonally at (cosθ,sinθ)(\cos\theta,\sin\theta). That is,

(1) Cθ{(x,y)2:x2+(cscθy)2=cot2θ}.\mathrm{C}_{\theta}\doteqdot\left\{(x,y)\in\mathbb{R}^{2}:x^{2}+\left(\csc\theta-y\right)^{2}=\cot^{2}\theta\right\}\,.

If we set

θ(t)arcsinetandθ+(t)arcsine2t,\theta^{-}(t)\doteqdot\arcsin\mathrm{e}^{t}\;\;\text{and}\;\;\theta^{+}(t)\doteqdot\arcsin\mathrm{e}^{2t}\,,

then we find that the inward normal speed of Cθ(t)\mathrm{C}_{\theta^{-}(t)} is no greater than its curvature κ\kappa^{-} and the inward normal speed of Cθ+(t)\mathrm{C}_{\theta^{+}(t)} is no less than its curvature κ+\kappa^{+}. The maximum principle and the Hopf boundary point lemma then imply that

Proposition 2.1.

a solution to free boundary curve shortening flow in B2B^{2} which lies below (resp. above) the circle Cθ0\mathrm{C}_{\theta_{0}} at time t0t_{0} lies below Cθ+(t0++tt0)\mathrm{C}_{\theta^{+}(t^{+}_{0}+t-t_{0})} (resp. above Cθ(t0+tt0)\mathrm{C}_{\theta^{-}(t^{-}_{0}+t-t_{0})}) for all t>t0t>t_{0}, where 2t0+=logsinθ02t^{+}_{0}=\log\sin\theta_{0} (resp. t0=logsinθ0t^{-}_{0}=\log\sin\theta_{0}).

Consider now the shifted scaled Angenent oval {Atλ}t(,0)\{\mathrm{A}^{\lambda}_{t}\}_{t\in(-\infty,0)}, where

Atλ{(x,y)×(0,π2λ):sin(λy)=eλ2tcosh(λx)}.\mathrm{A}^{\lambda}_{t}\doteqdot\left\{(x,y)\in\mathbb{R}\times(0,\tfrac{\pi}{2\lambda}):\sin(\lambda y)=\mathrm{e}^{\lambda^{2}t}\cosh\big{(}\lambda x\big{)}\right\}\,.

This evolves by curve shortening flow and satisfies

νλ(cosθ,sinθ)(cosθ,sinθ)=cosθtanh(λcosθ)sinθcot(λsinθ))tanh2(λcosθ)+cot2(λsinθ))\nu_{\lambda}(\cos\theta,\sin\theta)\cdot(\cos\theta,\sin\theta)=\frac{\cos\theta\tanh(\lambda\cos\theta)-\sin\theta\cot(\lambda\sin\theta))}{\sqrt{\tanh^{2}(\lambda\cos\theta)+\cot^{2}(\lambda\sin\theta))}}

at a point (cosθ,sinθ)B2(\cos\theta,\sin\theta)\in\partial B^{2}.

Lemma 2.2.

For each θ(0,π2)\theta\in(0,\frac{\pi}{2}), there is a unique λ(θ)(0,π2sinθ)\lambda(\theta)\in(0,\frac{\pi}{2\sin\theta}) such that

νλ(θ)(cosθ,sinθ)(cosθ,sinθ)=0.\nu_{\lambda(\theta)}(\cos\theta,\sin\theta)\cdot(\cos\theta,\sin\theta)=0\,.

Given θ,θ0(0,π2)\theta,\theta_{0}\in(0,\frac{\pi}{2}) with θ>θ0\theta>\theta_{0},

νλ(θ0)(cosθ,sinθ)(cosθ,sinθ)<0.\nu_{\lambda(\theta_{0})}(\cos\theta,\sin\theta)\cdot(\cos\theta,\sin\theta)<0\,.
Proof.

Define

f(λ,θ)cosθtanh(λcosθ)sinθcot(λsinθ)).f(\lambda,\theta)\doteqdot\cos\theta\tanh(\lambda\cos\theta)-\sin\theta\cot(\lambda\sin\theta))\,.

Observe that

limλ0f(λ,θ)=,limλπ2sinθf(λ,θ)=cosθtanh(π2cotθ)>0,\lim_{\lambda\searrow 0}f(\lambda,\theta)=-\infty\,,\;\;\lim_{\lambda\nearrow\frac{\pi}{2\sin\theta}}f(\lambda,\theta)=\cos\theta\tanh(\tfrac{\pi}{2}\cot\theta)>0\,,

and

(2) fλ=cos2θ(1tanh2(λcosθ))+sin2θ(1+cot2(λsinθ))>0.\frac{\partial f}{\partial\lambda}=\cos^{2}\theta(1-\tanh^{2}(\lambda\cos\theta))+\sin^{2}\theta(1+\cot^{2}(\lambda\sin\theta))>0\,.

The first claim follows.

Next observe that

fθ=\displaystyle\frac{\partial f}{\partial\theta}={} sinθtanh(λcosθ)λcosθsinθsech2(λcosθ)\displaystyle-\sin\theta\tanh(\lambda\cos\theta)-\lambda\cos\theta\sin\theta\operatorname{sech}^{2}(\lambda\cos\theta)
cosθcot(λsinθ)+λsinθcosθcsc2(λsinθ).\displaystyle-\cos\theta\cot(\lambda\sin\theta)+\lambda\sin\theta\cos\theta\csc^{2}(\lambda\sin\theta)\,.

Given θ(0,π2)\theta\in(0,\frac{\pi}{2}), we obtain, at the unique zero λ(0,π2sinθ)\lambda\in(0,\frac{\pi}{2\sin\theta}) of f(,θ)f(\cdot,\theta),

fθ=\displaystyle\frac{\partial f}{\partial\theta}={} sinθtanθcot(λsinθ)λcosθsinθ(1tan2θcot2(λsinθ))\displaystyle-\sin\theta\tan\theta\cot(\lambda\sin\theta)-\lambda\cos\theta\sin\theta(1-\tan^{2}\theta\cot^{2}(\lambda\sin\theta))
cosθcot(λsinθ)+λsinθcosθcsc2(λsinθ)\displaystyle-\cos\theta\cot(\lambda\sin\theta)+\lambda\sin\theta\cos\theta\csc^{2}(\lambda\sin\theta)
=\displaystyle={} secθcot(λsinθ)(1λsinθcot(λsinθ)).\displaystyle-\sec\theta\cot(\lambda\sin\theta)\left(1-\lambda\sin\theta\cot(\lambda\sin\theta)\right).

Since YcotY<1Y\cot Y<1 for Y(0,π2)Y\in(0,\frac{\pi}{2}), 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 {Γt}t[α,ω)\{\Gamma_{t}\}_{t\in[\alpha,\omega)} be a solution to free boundary curve shortening flow in B2B^{2}. Suppose that λλ(θα)\lambda\leq\lambda(\theta_{\alpha}), where θα\theta_{\alpha} denotes the smaller, in absolute value, of the two turning angles to Γα\Gamma_{\alpha} at its boundary. If Γα\Gamma_{\alpha} lies above Asλ\mathrm{A}^{\lambda}_{s}, then Γt\Gamma_{t} lies above As+tαλ\mathrm{A}^{\lambda}_{s+t-\alpha} for all t(α,ω)(,αs)t\in(\alpha,\omega)\cap(-\infty,\alpha-s).

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 (cosθ,sinθ)(\cos\theta,\sin\theta) with |θ|θα|\theta|\leq\theta_{\alpha}. On the other hand, since λλ(θα)\lambda\leq\lambda(\theta_{\alpha}), (2) implies that

f(λ,θα)f(λα,θα)=0,f(\lambda,\theta_{\alpha})\leq f(\lambda_{\alpha},\theta_{\alpha})=0\,,

and hence, by the argument of Lemma 2.2,

f(λ,θ)0forθθα.f(\lambda,\theta)\leq 0\;\;\text{for}\;\;\theta\geq\theta_{\alpha}\,.

So the Hopf boundary point lemma implies that no boundary contact can develop for θθα\theta\geq\theta_{\alpha} either. ∎

Remark 2.4.

Since scots1s\cot s\to 1 as s0s\to 0, f(λ,θ)f(\lambda,\theta) is non-negative at θ=0\theta=0 so long as λλ0\lambda\geq\lambda_{0}, where λ0tanhλ0=1\lambda_{0}\tanh\lambda_{0}=1.

2.2. Old-but-not-ancient solutions

For each ρ>0\rho>0, choose a curve Γρ\Gamma^{\rho} in B¯2\overline{B}{}^{2} with the following properties:

  • Γρ\Gamma^{\rho} meets B2\partial B^{2} orthogonally at (cosρ,sinρ)(\cos\rho,\sin\rho),

  • Γρ\Gamma^{\rho} is reflection symmetric about the yy-axis,

  • ΓρB2\Gamma^{\rho}\cap B^{2} is the relative boundary of a convex region ΩρB2\Omega^{\rho}\subset B^{2}, and

  • κsρ>0\kappa^{\rho}_{s}>0 in B2{x>0}B^{2}\cap\{x>0\}.

For example, we could take ΓρAtρλρB2\Gamma^{\rho}\doteqdot\mathrm{A}^{\lambda_{\rho}}_{t_{\rho}}\cap B^{2}, where λρ>λ0\lambda_{\rho}>\lambda_{0} and tρt_{\rho} are (uniquely) chosen so that

cosρtanh(λρcosρ)sinρcot(λρsinρ))=0\cos\rho\tanh(\lambda_{\rho}\cos\rho)-\sin\rho\cot(\lambda_{\rho}\sin\rho))=0

and

tρ=λρ2log(cosh(λρcosρ)sin(λρsinρ)).-t_{\rho}=\lambda_{\rho}^{-2}\log\left(\frac{\cosh(\lambda_{\rho}\cos\rho)}{\sin(\lambda_{\rho}\sin\rho)}\right)\,.

Observe that the circle Cθρ\mathrm{C}_{\theta_{\rho}} defined by

sinθρ=2sinρ1+sin2ρ\sin\theta_{\rho}=\frac{2\sin\rho}{1+\sin^{2}\rho}

is tangent to the line y=sinρy=\sin\rho, and hence lies above Γρ\Gamma^{\rho}.

Work of Stahl [22, 21] now yields the following old-but-not-ancient solutions.

Lemma 2.5.

For each ρ(0,π2)\rho\in(0,\frac{\pi}{2}), there exists a smooth solution222Given by a one parameter family of immersions X:[1,1]×[αρ,0)B¯2X:[-1,1]\times[\alpha_{\rho},0)\to\overline{B}{}^{2} satisfying XC([1,1]×(αρ,0))C2+β,1+β2([1,1]×[αρ,0))X\in C^{\infty}([-1,1]\times(\alpha_{\rho},0))\cap C^{2+\beta,1+\frac{\beta}{2}}([-1,1]\times[\alpha_{\rho},0)) for some β(0,1)\beta\in(0,1). {Γtρ}t[αρ,0)\{\Gamma^{\rho}_{t}\}_{t\in[\alpha_{\rho},0)} to curve shortening flow with Γαρρ=Γρ\Gamma^{\rho}_{\alpha_{\rho}}=\Gamma^{\rho} which satisfies the following properties:

  • Γtρ\Gamma^{\rho}_{t} meets B2\partial B^{2} orthogonally for each t(αρ,0)t\in(\alpha_{\rho},0),

  • Γtρ\Gamma^{\rho}_{t} is convex and locally uniformly convex for each t(αρ,0)t\in(\alpha_{\rho},0),

  • Γtρ\Gamma^{\rho}_{t} is reflection symmetric about the yy-axis for each t(αρ,0)t\in(\alpha_{\rho},0),

  • Γtρ(0,1)\Gamma_{t}^{\rho}\to(0,1) uniformly as t0t\to 0,

  • κsρ>0\kappa^{\rho}_{s}>0 in B2{x>0}B^{2}\cap\{x>0\}, and

  • αρ<12log(2sinρ1+sin2ρ)\alpha_{\rho}<\frac{1}{2}\log\left(\frac{2\sin\rho}{1+\sin^{2}\rho}\right)\to-\infty as ρ0\rho\to 0.

Proof.

Existence of a maximal solution to curve shortening flow out of Γρ\Gamma^{\rho} which meets B2\partial B^{2} 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 B2B^{2} at the final time (which is finite) [21, Proposition 1.4]. We obtain {Γtρ}t[αρ,0)\{\Gamma_{t}^{\rho}\}_{t\in[\alpha_{\rho},0)} by time-translating Stahl’s solution.

By uniqueness of solutions Γtρ\Gamma^{\rho}_{t} remains reflection symmetric about the yy-axis for t(αρ,0)t\in(\alpha_{\rho},0), so the final point is (0,1)(0,1).

The reflection symmetry also implies that κsρ=0\kappa^{\rho}_{s}=0 at the point ptΓtρ{x=0}p_{t}\doteqdot\Gamma^{\rho}_{t}\cap\{x=0\} for all t[αρ,0)t\in[\alpha_{\rho},0). By [21, Proposition 2.1], κsρ=κρ>0\kappa^{\rho}_{s}=\kappa^{\rho}>0 at the boundary point qtΓtρ{x>0}q_{t}\doteqdot\partial\Gamma^{\rho}_{t}\cap\{x>0\} for all t(αρ,0)t\in(\alpha_{\rho},0). Applying Sturm’s theorem [4] to κsρ\kappa^{\rho}_{s}, we thus find that κsρ>0\kappa^{\rho}_{s}>0 on ΓtρB2{x>0}\Gamma^{\rho}_{t}\cap B^{2}\cap\{x>0\} for all t(αρ,0)t\in(\alpha_{\rho},0).

Since CθρΩρ\mathrm{C}_{\theta_{\rho}}\subset\Omega^{\rho}, the final property follows from Proposition 2.1. ∎

We now fix ρ>0\rho>0 and drop the super/subscript ρ\rho. Set

κ¯(t)minΓtκ=κ(pt)andκ¯(t)maxΓtκ=κ(qt),\underline{\kappa}(t)\doteqdot\min_{\Gamma_{t}}\kappa=\kappa(p_{t})\;\;\text{and}\;\;\overline{\kappa}(t)\doteqdot\max_{\Gamma_{t}}\kappa=\kappa(q_{t})\,,

and define y¯(t)\underline{y}(t), y¯(t)\overline{y}(t) and θ¯(t)\overline{\theta}(t) by

pt=(0,y¯(t)),qt=(cosθ¯(t),sinθ¯(t)),andy¯(t)=sinθ¯(t).p_{t}=(0,\underline{y}(t))\,,\;\;q_{t}=(\cos\overline{\theta}(t),\sin\overline{\theta}(t))\,,\;\;\text{and}\;\;\overline{y}(t)=\sin\overline{\theta}(t)\,.
Lemma 2.6.

Each old-but-not-ancient solution satisfies

(3) κ¯tanθ¯κ¯,\underline{\kappa}\leq\tan\overline{\theta}\leq\overline{\kappa}\,,
(4) sinθ¯et,\sin\overline{\theta}\leq e^{t}\,,

and

(5) sinθ¯1+cosθ¯y¯sinθ¯.\frac{\sin\overline{\theta}}{1+\cos\overline{\theta}}\leq\underline{y}\leq\sin\overline{\theta}\,.
Proof.

To prove the lower bound for κ¯\overline{\kappa}, it suffices to show that the circle Cθ¯(t)\mathrm{C}_{\overline{\theta}(t)} (see (1)) lies locally below Γt\Gamma_{t} near qtq_{t}. If this is not the case, then, locally around qtq_{t}, Γt\Gamma_{t} lies below Cθ¯(t)\mathrm{C}_{\overline{\theta}(t)} and hence κ(qt)tanθ¯(t)\kappa(q_{t})\leq\tan\overline{\theta}(t). But then we can translate Cθ¯(t)\mathrm{C}_{\overline{\theta}(t)} downwards until it touches Γt\Gamma_{t} from below in an interior point at which the curvature must satisfy κtanθ¯(t)\kappa\geq\tan\overline{\theta}(t). This contradicts the unique maximization of the curvature at qtq_{t}.

The estimate (4) now follows by integrating the inequality

ddtsinθ¯=cosθ¯κ¯sinθ¯\frac{d}{dt}\sin\overline{\theta}=\cos\overline{\theta}\,\overline{\kappa}\geq\sin\overline{\theta}

between any initial time tt and the final time 0 (at which θ¯=π2\overline{\theta}=\frac{\pi}{2} since the solution contracts to the point (0,1)(0,1)).

The upper bound for y¯\underline{y} follows from convexity and the boundary condition y¯=sinθ¯\overline{y}=\sin\overline{\theta}. To prove the lower bound, we will show that the circle Cθ¯(t)\mathrm{C}_{\overline{\theta}(t)} lies nowhere above Γt\Gamma_{t}. Suppose that this is not the case. Then, since Cθ¯(t)\mathrm{C}_{\overline{\theta}(t)} lies locally below Γt\Gamma_{t} near qtq_{t}, we can move Cθ¯(t)\mathrm{C}_{\overline{\theta}(t)} downwards until it is tangent from below to a point ptp^{\prime}_{t} on Γt{x0}\Gamma_{t}\cap\{x\geq 0\}, at which we must have κtanθ¯(t)\kappa\geq\tan\overline{\theta}(t). But then, since κs0\kappa_{s}\geq 0 in {x>0}\{x>0\}, we find that κtanθ¯(t)\kappa\geq\tan\overline{\theta}(t) for all points between ptp^{\prime}_{t} and qtq_{t}. But this implies that this whole arc (including ptp^{\prime}_{t}) lies above Cθ¯(t)\mathrm{C}_{\overline{\theta}(t)}, a contradiction.

To prove the upper bound for κ¯\underline{\kappa}, fix tt and consider the circle CC centred on the yy-axis through the points ptp_{t} and qtq_{t}. Its radius is r(t)r(t), where

rcos2θ¯+(sinθ¯y¯)22(sinθ¯y¯).r\doteqdot\frac{\cos^{2}\overline{\theta}+(\sin\overline{\theta}-\underline{y})^{2}}{2(\sin\overline{\theta}-\underline{y})}\,.

We claim that Γt\Gamma_{t} lies locally below CC near ptp_{t}. Suppose that this is not the case. Then, by the symmetry of Γt\Gamma_{t} and CC across the yy-axis, Γt\Gamma_{t} lies locally above CC near ptp_{t}. This implies two things: first, that

κ(pt)r1,\kappa(p_{t})\geq r^{-1},

and second, that, by moving CC vertically upwards, we can find a point ptp^{\prime}_{t} (the final point of contact) which satisfies

κ(pt)r1.\kappa(p^{\prime}_{t})\leq r^{-1}\,.

These two inequalities contradict the (unique) minimization of κ\kappa at ptp_{t}. We conclude that

κ¯2(sinθ¯y¯)cos2θ¯+(sinθ¯y¯)2tanθ¯\underline{\kappa}\leq\frac{2(\sin\overline{\theta}-\underline{y})}{\cos^{2}\overline{\theta}+(\sin\overline{\theta}-\underline{y})^{2}}\leq\tan\overline{\theta}

due to the lower bund for y¯\underline{y}. ∎

Remark 2.7.

If we parametrize by turning angle θ[θ¯,θ¯]\theta\in[-\overline{\theta},\overline{\theta}], so that

τ=(cosθ,sinθ),\tau=(\cos\theta,\sin\theta)\,,

then the estimates (3) are also easily obtained from the monotonicity of κ\kappa and the formulae

(6) x(θ)=x0+0θcosuκ(u)𝑑uandy(θ)=y0+0θsinuκ(u)𝑑u.x(\theta)=x_{0}+\int_{0}^{\theta}\frac{\cos u}{\kappa(u)}du\;\;\text{and}\;\;y(\theta)=y_{0}+\int_{0}^{\theta}\frac{\sin u}{\kappa(u)}du\,.

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 ρ>0\rho>0, consider the old-but-not-ancient solution {Γtρ}t[αρ,0)\{\Gamma^{\rho}_{t}\}_{t\in[\alpha_{\rho},0)}, Γtρ=Ωtρ\Gamma^{\rho}_{t}=\partial\Omega^{\rho}_{t}, constructed in Lemma 2.5. By (4), Ωtρ\Omega^{\rho}_{t} contains Cω(t)B2\mathrm{C}_{\omega(t)}\cap B^{2}, where ω(t)(0,π2)\omega(t)\in(0,\frac{\pi}{2}) is uniquely defined by

1cosω(t)sinω(t)=et.\frac{1-\cos\omega(t)}{\sin\omega(t)}=\mathrm{e}^{t}\,.

If we represent Γtρ\Gamma^{\rho}_{t} as a graph xyρ(x,t)x\mapsto y^{\rho}(x,t) over the xx-axis, then convexity and the boundary condition imply that |yxρ|tanω|y^{\rho}_{x}|\leq\tan\omega. Since ω(t)\omega(t) is independent of ρ\rho, Stahl’s (global in space, interior in time) Ecker–Huisken type estimates [22] imply uniform-in-ρ\rho bounds for the curvature and its derivatives. So the limit

{Γtρ}t[αρ,0){Γt}t(,0)\{\Gamma^{\rho}_{t}\}_{t\in[\alpha_{\rho},0)}\to\{\Gamma_{t}\}_{t\in(-\infty,0)}

exists in CC^{\infty} (globally in space on compact subsets of time) and the limit {Γt}t(,0)\{\Gamma_{t}\}_{t\in(-\infty,0)} satisfies curve shortening flow with free boundary in B2B^{2}. On the other hand, since {Γtρ}t(αρ,0)\{\Gamma^{\rho}_{t}\}_{t\in(\alpha_{\rho},0)} contracts to (0,1)(0,1) as t0t\to 0, (the contrapositive of) Proposition 2.1 implies that Γtρ\Gamma^{\rho}_{t} must intersect the closed region enclosed by Cθ+(t)\mathrm{C}_{\theta^{+}(t)} for all t<0t<0. It follows that Γt\Gamma_{t} must intersect the closed region enclosed by Cθ+(t)\mathrm{C}_{\theta^{+}(t)} for all t<0t<0. Since each Γt\Gamma_{t} is the limit of convex boundaries, each is convex. It follows that Γt\Gamma_{t} converges to (0,1)(0,1) as t0t\to 0 and, by [22, Corollary 4.5], that Γt\Gamma_{t} is locally uniformly convex for each tt. ∎

2.4. Asymptotics for the height

For the purposes of this section, we fix an ancient solution {Γt}(,0)\{\Gamma_{t}\}_{(-\infty,0)} obtained as in Proposition 2.8 by taking a sublimit as λλ0\lambda\searrow\lambda_{0} of the specific old-but-not ancient solutions {Γtλ}t[αλ,0)\{\Gamma^{\lambda}_{t}\}_{t\in[\alpha_{\lambda},0)} corresponding to Γαλλ=AtλλB2\Gamma^{\lambda}_{\alpha_{\lambda}}=\mathrm{A}^{\lambda}_{t_{\lambda}}\cap B^{2}, tλt_{\lambda} being the time at which {Atλ}t(,0)\{\mathrm{A}^{\lambda}_{t}\}_{t\in(-\infty,0)} meets B2\partial B^{2} orthogonally. The asymptotics we obtain for this solution will be used to prove its uniqueness.

We will need to prove that the limit limteλ02ty¯(t)\lim_{t\to-\infty}\mathrm{e}^{-\lambda_{0}^{2}t}\underline{y}(t) exists in (0,)(0,\infty). The following speed bound will imply that it exists in [0,)[0,\infty).

Lemma 2.9.

The ancient solution {Γt}(,0)\{\Gamma_{t}\}_{(-\infty,0)} satisfies

(7) κcosθλ0tan(λ0y).\frac{\kappa}{\cos\theta}\geq\lambda_{0}\tan(\lambda_{0}y)\,.
Proof.

It suffices to prove that κcosθλtan(λy)\frac{\kappa}{\cos\theta}\geq\lambda\tan(\lambda y) on each of the old-but-not-ancient solutions {Γtλ}t[αλ,0)\{\Gamma^{\lambda}_{t}\}_{t\in[\alpha_{\lambda},0)}. Note that equality holds on the initial timeslice Γαλλ=Atλλ\Gamma^{\lambda}_{\alpha_{\lambda}}=\mathrm{A}^{\lambda}_{t_{\lambda}}.

Given any μ<λ\mu<\lambda, set uμtan(μy)u\doteqdot\mu\tan(\mu y) and vxs=cosθ=ν,e2v\doteqdot x_{s}=\cos\theta=\left\langle\nu,e_{2}\right\rangle. Observe that

us=μ2sec2(μy)sinθ,(tΔ)u=2μ2sec2(μy)sin2θu,u_{s}=\mu^{2}\sec^{2}(\mu y)\sin\theta\,,\;\;(\partial_{t}-\Delta)u=-2\mu^{2}\sec^{2}(\mu y)\sin^{2}\theta u\,,
vs=κsinθand(tΔ)v=κ2v.v_{s}=-\kappa\sin\theta\;\;\text{and}\;\;(\partial_{t}-\Delta)v=\kappa^{2}v\,.

At an interior maximum of uvκ\frac{uv}{\kappa} we observe that

κκ=uu+vv\frac{\nabla\kappa}{\kappa}=\frac{\nabla u}{u}+\frac{\nabla v}{v}

and hence

0(tΔ)uvκ=\displaystyle 0\leq(\partial_{t}-\Delta)\frac{uv}{\kappa}={} uvκ((tΔ)uu2uu,vv)\displaystyle\frac{uv}{\kappa}\left(\frac{(\partial_{t}-\Delta)u}{u}-2\left\langle\frac{\nabla u}{u},\frac{\nabla v}{v}\right\rangle\right)
(8) =\displaystyle={} 2μ2sec2(μy)sin2θ(1uvκ).\displaystyle 2\mu^{2}\sec^{2}(\mu y)\sin^{2}\theta\left(1-\frac{uv}{\kappa}\right).

At a (without loss of generality right) boundary maximum of uvκ\frac{uv}{\kappa}, we have ys=yy_{s}=y and κs=κ\kappa_{s}=\kappa, and hence

(uvκ)s=\displaystyle\left(\frac{uv}{\kappa}\right)_{s}={} uvκ(usu+vsvκsκ)\displaystyle\frac{uv}{\kappa}\left(\frac{u_{s}}{u}+\frac{v_{s}}{v}-\frac{\kappa_{s}}{\kappa}\right)
=\displaystyle={} uvκ(sec2(μy)μytanμyκyv1)\displaystyle\frac{uv}{\kappa}\left(\frac{\sec^{2}(\mu y)\mu y}{\tan{\mu y}}-\kappa\frac{y}{v}-1\right)
(9) \displaystyle\leq{} (uvκ1)tan(μy)μy.\displaystyle\left(\frac{uv}{\kappa}-1\right)\tan(\mu y)\mu y\,.

We may now conclude that maxΓ¯tλuvκ\max_{\overline{\Gamma}^{\lambda}_{t}}\frac{uv}{\kappa} remains less than one. Indeed, if uvκ\frac{uv}{\kappa} ever reaches 11, then there must be a first time t0>0t_{0}>0 and a point x0Γ¯tx_{0}\in\overline{\Gamma}_{t} at which this occurs (note that uv/κuv/\kappa is continuous on Γ¯t\overline{\Gamma}_{t} up to the initial time). The point x0x_{0} 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

κcosθμtan(μy)\frac{\kappa}{\cos\theta}\geq\mu\tan(\mu y)

on {Γtλ}t[αλ,0)\{\Gamma^{\lambda}_{t}\}_{t\in[\alpha_{\lambda},0)} for all μ<λ\mu<\lambda. Now take μλ\mu\to\lambda. ∎

If we parametrize Γt\Gamma_{t} as a graph xy(x,t)x\mapsto y(x,t) over the xx-axis, then (7) yields

(sin(λ0y))t=\displaystyle\big{(}\sin(\lambda_{0}y)\big{)}_{t}={} λ0cos(λ0y)κ1+|yx|2=λ0cos(λ0y)κcosθλ0sin(λ0y)\displaystyle\lambda_{0}\cos(\lambda_{0}y)\kappa\sqrt{1+|y_{x}|^{2}}=\lambda_{0}\cos(\lambda_{0}y)\frac{\kappa}{\cos\theta}\geq\lambda_{0}\sin(\lambda_{0}y)

and hence

(10) (eλ02tsin(λ0y(x,t)))t0.\left(\mathrm{e}^{-\lambda_{0}^{2}t}\sin(\lambda_{0}y(x,t))\right)_{t}\geq 0\,.

In particular, the limit

A(x)limteλ02ty(x,t)A(x)\doteqdot\lim_{t\to-\infty}\mathrm{e}^{-\lambda_{0}^{2}t}y(x,t)

exists in [0,)[0,\infty) for each x(1,1)x\in(-1,1), 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 T>T>-\infty and C<C<\infty such that

(11) κ¯Cetfor allt<T.\overline{\kappa}\leq Ce^{t}\;\;\text{for all}\;\;t<T\,.
Proof.

We will prove the estimate for each old-but-not-ancient solution {Γtλ}t(αλ,0)\{\Gamma^{\lambda}_{t}\}_{t\in(\alpha_{\lambda},0)}. We first prove a crude gradient estimate of the form

(12) |κs|2κ|\kappa_{s}|\leq 2\kappa

for tt sufficiently negative. It will suffice to prove that

(13) |κs|κ+γ,ν0.|\kappa_{s}|-\kappa+\left\langle\gamma,\nu\right\rangle\leq 0\,.

Indeed, since γ,νs=κγ,τ\left\langle\gamma,\nu\right\rangle_{s}=\kappa\left\langle\gamma,\tau\right\rangle has the same sign as the xx-coordinate, we may estimate, as in (7),

(14) |γ,ν||γ,ν|x=0λ2κ|x=0=λ2minΓtκκ.|\!\left\langle\gamma,\nu\right\rangle\!|\leq|\!\left\langle\gamma,\nu\right\rangle\!|_{x=0}\leq\lambda^{-2}\kappa|_{x=0}=\lambda^{-2}\min_{\Gamma_{t}}\kappa\leq\kappa\,.

For λ\lambda sufficiently close to λ0\lambda_{0}, we have κ|t=αλ<1/2\kappa|_{t=\alpha_{\lambda}}<1/2. Denote by TλT^{\lambda} the first time at which κ\kappa reaches 1/21/2. Since κ\kappa is continuous up to the initial time αλ\alpha_{\lambda}, we have Tλ>αλT^{\lambda}>\alpha_{\lambda}. We claim that (13) holds for t<Tλt<T^{\lambda}. Indeed, it is satisfied on the initial timeslice Γαλλ=Atλλ\Gamma^{\lambda}_{\alpha_{\lambda}}=\mathrm{A}^{\lambda}_{t_{\lambda}} since

κs2κ2=λ2(cos2θsin2θsin2θaλ2)=λ2(sin4θ+aλ2)0,\kappa_{s}^{2}-\kappa^{2}=\lambda^{2}\left(\cos^{2}\theta\sin^{2}\theta-\sin^{2}\theta-a^{2}_{\lambda}\right)=-\lambda^{2}(\sin^{4}\theta+a_{\lambda}^{2})\leq 0\,,

whereas γ,ν0\left\langle\gamma,\nu\right\rangle\leq 0. We will show that

fε|κs|κ+γ,νεetαλf_{\varepsilon}\doteqdot|\kappa_{s}|-\kappa+\left\langle\gamma,\nu\right\rangle-\varepsilon\mathrm{e}^{t-\alpha_{\lambda}}

remains negative up to time TλT^{\lambda}. Suppose, to the contrary, that fεf_{\varepsilon} reaches zero at some time t<Tλt<T^{\lambda} at some point pΓ¯tp\in\overline{\Gamma}_{t}. Since |κs|κ+γ,ν|\kappa_{s}|-\kappa+\left\langle\gamma,\nu\right\rangle vanishes at the boundary, pp must be an interior point. Since κs\kappa_{s} vanishes at the yy-axis, and the curve is symmetric, we may assume that x(p)>0x(p)>0. At such a point,

0(tΔ)fε=\displaystyle 0\leq(\partial_{t}-\Delta)f_{\varepsilon}={} κ2(4κsκ+γ,ν)2κεetαλ\displaystyle\kappa^{2}(4\kappa_{s}-\kappa+\left\langle\gamma,\nu\right\rangle)-2\kappa-\varepsilon\mathrm{e}^{t-\alpha_{\lambda}}
=\displaystyle={} κ2(3[κγ,ν]+4εetαλ)2κεetαλ.\displaystyle\kappa^{2}(3[\kappa-\left\langle\gamma,\nu\right\rangle]+4\varepsilon\mathrm{e}^{t-\alpha_{\lambda}})-2\kappa-\varepsilon\mathrm{e}^{t-\alpha_{\lambda}}\,.

Recalling (14) and estimating κ12\kappa\leq\frac{1}{2} yields

06κ32κ+(4κ21)εetαλ<0,0\leq 6\kappa^{3}-2\kappa+(4\kappa^{2}-1)\varepsilon\mathrm{e}^{t-\alpha_{\lambda}}<0\,,

which is absurd. So fεf_{\varepsilon} does indeed remain negative, and taking ε0\varepsilon\to 0 yields (12) for t<Tλt<T^{\lambda}.

Since Length(Γt{x0})1\mathrm{Length}(\Gamma_{t}\cap\{x\geq 0\})\leq 1, integrating (12) yields

κ¯e2κ¯fort<Tλ.\overline{\kappa}\leq\mathrm{e}^{2}\underline{\kappa}\;\;\text{for}\;\;t<T^{\lambda}\,.

Recalling (3) and (4), this implies that

κ¯e2et1e2tfort<Tλ.\overline{\kappa}\leq\mathrm{e}^{2}\frac{\mathrm{e}^{t}}{\sqrt{1-\mathrm{e}^{2t}}}\;\;\text{for}\;\;t<T^{\lambda}\,.

Taking t=Tλt=T^{\lambda} we find that TλTT^{\lambda}\geq T, where TT is independent of λ\lambda, so we conclude that

κ¯Cetfort<T,\overline{\kappa}\leq C\mathrm{e}^{t}\;\;\text{for}\;\;t<T\,,

where CC and TT do not depend on λ\lambda. ∎

Lemma 2.11.

There exist C<C<\infty and T>T>-\infty such that

κyλ02+Ce2tfort<T.\frac{\kappa}{y}\leq\lambda_{0}^{2}+C\mathrm{e}^{2t}\;\;\text{for}\;\;t<T\,.
Proof.

Consider the old-but-not-ancient solution {Γtλ}t(,0)\{\Gamma^{\lambda}_{t}\}_{t\in(-\infty,0)}. By (11), we can find C<C<\infty and T>T>-\infty such that

(tΔ)κy=\displaystyle(\partial_{t}-\Delta)\frac{\kappa}{y}={} κ2κy+2κy,yy\displaystyle\kappa^{2}\frac{\kappa}{y}+2\left\langle\nabla\frac{\kappa}{y},\frac{\nabla y}{y}\right\rangle
\displaystyle\leq{} Ce2tκy+2κy,yyfort<T.\displaystyle C\mathrm{e}^{2t}\frac{\kappa}{y}+2\left\langle\nabla\frac{\kappa}{y},\frac{\nabla y}{y}\right\rangle\;\;\text{for}\;\;t<T\,.

Since, at a boundary point,

(κy)s=κsyκyysy=0,\left(\frac{\kappa}{y}\right)_{s}=\frac{\kappa_{s}}{y}-\frac{\kappa}{y}\frac{y_{s}}{y}=0\,,

the Hopf boundary point lemma and the ode comparison principle yield

maxΓtλκyCmaxΓαλλκyfort(αλ,T).\max_{\Gamma^{\lambda}_{t}}\frac{\kappa}{y}\leq C\max_{\Gamma^{\lambda}_{\alpha_{\lambda}}}\frac{\kappa}{y}\;\;\text{for}\;\;t\in(\alpha_{\lambda},T)\,.

But now

(tΔ)κy\displaystyle(\partial_{t}-\Delta)\frac{\kappa}{y}\leq{} Ce2tmaxΓαλλκy+2κy,yyfort<T,\displaystyle C\mathrm{e}^{2t}\max_{\Gamma^{\lambda}_{\alpha_{\lambda}}}\frac{\kappa}{y}+2\left\langle\nabla\frac{\kappa}{y},\frac{\nabla y}{y}\right\rangle\;\;\text{for}\;\;t<T\,,

and hence, by ode comparison,

maxΓtλκymaxΓαλλκy(1+Ce2t)fort(αλ,T).\displaystyle\max_{\Gamma^{\lambda}_{t}}\frac{\kappa}{y}\leq\max_{\Gamma^{\lambda}_{\alpha_{\lambda}}}\frac{\kappa}{y}\left(1+C\mathrm{e}^{2t}\right)\;\;\text{for}\;\;t\in(\alpha_{\lambda},T)\,.

Since, on the initial timeslice Γαλλ=Atλλ\Gamma^{\lambda}_{\alpha_{\lambda}}=\mathrm{A}_{t_{\lambda}}^{\lambda},

κy=λtan(λy)ycosθ,\frac{\kappa}{y}=\frac{\lambda\tan(\lambda y)}{y}\cos\theta\,,

the claim follows upon taking λλ0\lambda\to\lambda_{0}. ∎

It follows that

(logy¯(t)λ02t)tCe2tfort<T\left(\log\underline{y}(t)-\lambda_{0}^{2}t\right)_{t}\leq C\mathrm{e}^{2t}\;\;\text{for}\;\;t<T

and hence, integrating from time tt up to time TT,

logy¯(t)λ02tlogy¯(T)λ02TCfort<T.\log\underline{y}(t)-\lambda_{0}^{2}t\geq\log\underline{y}(T)-\lambda_{0}^{2}T-C\;\;\text{for}\;\;t<T\,.

So we indeed find that

Lemma 2.12.

the limit

(15) Alimteλ02ty¯(t)A\doteqdot\lim_{t\to-\infty}\mathrm{e}^{-\lambda_{0}^{2}t}\underline{y}(t)

exists in (0,)(0,\infty) on the particular ancient solution {Γt}(,0)\{\Gamma_{t}\}_{(-\infty,0)}.

3. Uniqueness

Now let {Γt}t(,0)\{\Gamma_{t}\}_{t\in(-\infty,0)}, Γt=relΩt\Gamma_{t}=\partial_{\mathrm{rel}}\Omega_{t}, be any convex, locally uniformly convex ancient free boundary curve shortening flow in the disc. By Stahl’s theorem [21], we may assume that Γt\Gamma_{t} contracts to a point on the boundary as t0t\to 0.

3.1. Backwards convergence

We first show that Γ¯t\overline{\Gamma}_{t} converges to a bisector as tt\to-\infty.

Lemma 3.1.

Up to a rotation of the plane,

Γ¯tC[1,1]×{0}ast.\overline{\Gamma}_{t}\underset{C^{\infty}}{\longrightarrow}[-1,1]\times\{0\}\,\,\text{as}\;\;t\to-\infty\,.
Proof.

Set A(t)area(Ωt)A(t)\doteqdot\operatorname{area}(\Omega_{t}). Integrating the variational formula for area yields

A(t)=t0Γt𝑑θ,A(t)=\int_{t}^{0}\!\!\!\int_{\Gamma_{t}}d\theta\,,

where θ\theta is the turning angle. Since convexity ensures that the total turning angle Γt𝑑θ\int_{\Gamma_{t}}d\theta is increasing and A(t)πA(t)\leq\pi for all tt, we find that

Γt𝑑θ0ast.\int_{\Gamma_{t}}d\theta\to 0\,\,\text{as}\;\;t\to-\infty\,.

Monotonicity of the flow, the free boundary condition and convexity now imply that the enclosed regions Ωt\Omega_{t} satisfy

Ω¯tB2{y0}ast\overline{\Omega}_{t}\to B^{2}\cap\{y\geq 0\}\,\,\text{as}\;\;t\to-\infty

in the Hausdorff topology.

If we now represent Γt\Gamma_{t} graphically over the xx-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 κ\kappa 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 tjt_{j}\to-\infty. 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.

Γt\Gamma_{t} is reflection symmetric about the yy-axis for all tt.

Proof.

Given any ω(0,π)\omega\in(0,\pi), we define the halfspace

Hω={(x,y):(x,y)(sinω,cosω)>0}H_{\omega}=\{(x,y):(x,y)\cdot(\sin\omega,-\cos\omega)>0\}

and denote by RωR_{\omega} the reflection about Hω\partial H_{\omega}. We first claim that, for every ω\omega, there exists t=tωt=t_{\omega} such that

(RωΓt)(ΓtHω)=for allt<tω.(R_{\omega}\cdot\Gamma_{t})\cap(\Gamma_{t}\cap H_{\omega})=\emptyset\;\;\text{for all}\;\;t<t_{\omega}\,.

Assume that the claim is not true. Then there exists ω(0,π)\omega\in(0,\pi), a sequence of times tit_{i}\to-\infty, and a sequence of pairs of points pi,qiΓtip_{i},q_{i}\in\Gamma_{t_{i}} such that Rω(pi)=qiR_{\omega}(p_{i})=q_{i}. This implies that the line passing through pip_{i} and qiq_{i} is parallel to the vector (sinω,cosω)(\sin\omega,-\cos\omega), so the mean value theorem yields for each ii a point rir_{i} on Γti\Gamma_{t_{i}} where the normal is parallel to (cosω,sinω)(\cos\omega,\sin\omega). This contradicts Lemma 3.1. ∎

3.3. Asymptotics for the height

We begin with a lemma.

Lemma 3.3.

For all t<0t<0,

κs>0in{x>0}Γt\kappa_{s}>0\;\;\text{in}\;\;\{x>0\}\cap\Gamma_{t}

and hence

(16) sinθ¯1+cosθ¯y¯.\frac{\sin\overline{\theta}}{1+\cos\overline{\theta}}\leq\underline{y}\,.
Proof.

Choose T>T>-\infty so that κ<27\kappa<\frac{2}{7} for t<Tt<T and, given ε>0\varepsilon>0, set

vεκs+ε(1γ,ν).v_{\varepsilon}\doteqdot\kappa_{s}+\varepsilon(1-\langle\gamma,\nu\rangle)\,.

We claim that vε0v_{\varepsilon}\geq 0 in {x0}(,T)\{x\geq 0\}\cap(-\infty,T). Suppose that this is not the case. Since at the spatial boundary vε>εv_{\varepsilon}>\varepsilon, and vεεv_{\varepsilon}\to\varepsilon as tt\to-\infty, there must exist a first time in (,T)(-\infty,T) and an interior point at which vε=0v_{\varepsilon}=0. But, at such a point,

0(tΔ)vε=κ2(κsεx,ν)+3κ2κs+2εκ=εκ23εκ2(1x,ν)+2εκε(27κ)κ>0,\begin{split}0\geq\left(\partial_{t}-\Delta\right)v_{\varepsilon}&=\kappa^{2}(\kappa_{s}-\varepsilon\langle x,\nu\rangle)+3\kappa^{2}\kappa_{s}+2\varepsilon\kappa\\ &=-\varepsilon\kappa^{2}-3\varepsilon\kappa^{2}(1-\langle x,\nu\rangle)+2\varepsilon\kappa\\ &\geq\varepsilon(2-7\kappa)\kappa\\ &>0\,,\end{split}

which is absurd. Now take ε0\varepsilon\to 0 to obtain κs0\kappa_{s}\geq 0 in {x0}Γt\{x\geq 0\}\cap\Gamma_{t} for t(,T]t\in(-\infty,T]. Since κs=0\kappa_{s}=0 at the yy-axis and κs=κ>0\kappa_{s}=\kappa>0 at the right boundary point, the strong maximum principle and the Hopf boundary point lemma imply that κs>0\kappa_{s}>0 in {x>0}Γt\{x>0\}\cap\Gamma_{t} for t(,T]t\in(-\infty,T]. But then Sturm’s theorem implies that κs\kappa_{s} does not develop additional zeroes up to time 0.

Having established the first claim, the second follows as in Lemma 2.6. ∎

Proposition 3.4.

If we define A(0,)A\in(0,\infty) as in (15), then

eλ02ty(x,t)Acosh(λ0x)uniformly ast.\mathrm{e}^{\lambda_{0}^{2}t}y(x,t)\to A\cosh(\lambda_{0}x)\;\;\text{uniformly as}\;\;t\to-\infty\,.
Proof.

Given τ<0\tau<0, consider the rescaled height function

yτ(x,t)eλ02τy(x,t+τ),y^{\tau}(x,t)\doteqdot\mathrm{e}^{-\lambda_{0}^{2}\tau}y(x,t+\tau)\,,

which is defined on the time-translated flow {Γtτ}t(,τ)\{\Gamma^{\tau}_{t}\}_{t\in(-\infty,-\tau)}, where ΓtτΓt+τ\Gamma^{\tau}_{t}\doteqdot\Gamma_{t+\tau}. Note that

(17) {(tΔτ)yτ=0in{Γtτ}t(,τ)τyτ,N=yon{Γtτ}t(,τ),\left\{\begin{aligned} (\partial_{t}-\Delta^{\tau})y^{\tau}={}&0\;\;\text{in}\;\;\{\Gamma_{t}^{\tau}\}_{t\in(-\infty,-\tau)}\\ \left\langle\nabla^{\tau}y^{\tau},N\right\rangle={}&y\;\;\text{on}\;\;\{\partial\Gamma_{t}^{\tau}\}_{t\in(-\infty,-\tau)}\,,\end{aligned}\right.

where τ\nabla^{\tau} and Δτ\Delta^{\tau} are the gradient and Laplacian on {Γtτ}t(,τ)\{\Gamma^{\tau}_{t}\}_{t\in(-\infty,-\tau)}, respectively, and NN is the outward unit normal to B2\partial B^{2}.

Since {Γt}t(,0)\{\Gamma_{t}\}_{t\in(-\infty,0)} reaches the origin at time zero, it must intersect the constructed solution for all t<0t<0. In particular, the value of y¯\underline{y} on the former can at no time exceed the value of y¯\overline{y} on the latter. But then (15) and (16) yield

(18) lim supteλ02ty¯<.\limsup_{t\to-\infty}\mathrm{e}^{-\lambda_{0}^{2}t}\overline{y}<\infty\,.

This implies a uniform bound for yτy^{\tau} on {Γtτ}t(,T]\{\Gamma^{\tau}_{t}\}_{t\in(-\infty,T]} for any TT\in\mathbb{R}. So Alaoglu’s theorem yields a sequence of times τj\tau_{j}\to-\infty such that yτjy^{\tau_{j}} converges in the weak topology as jj\to\infty to some yLloc2([1,1]×(,))y^{\infty}\in L^{2}_{\mathrm{loc}}([-1,1]\times(-\infty,\infty)). Since convexity and the boundary condition imply a uniform bound for τyτ\nabla^{\tau}y^{\tau} on any time interval of the form (,T](-\infty,T], we may also arrange that the convergence is uniform in space at time zero, say.

Weak convergence ensures that yy^{\infty} satisfies the problem

(19) {yt=yxxin[1,1]×(,)yx(±1)=±y(±1).\left\{\begin{aligned} y_{t}={}&y_{xx}\;\;\text{in}\;\;[-1,1]\times(-\infty,\infty)\\ y_{x}(\pm 1)={}&\pm y(\pm 1)\,.\end{aligned}\right.

Indeed, a smooth function yτy^{\tau} satisfies the boundary value problem (17) (and analogously for (19)) if and only if

τΓtτyτ(tΔτ)η=0\int_{-\infty}^{-\tau}\!\int_{\Gamma_{t}^{\tau}}y^{\tau}(\partial_{t}-\Delta^{\tau})^{\ast}\eta=0

for all smooth η\eta which are compactly supported in time and satisfy

τηN=ηonΓtτ,\nabla^{\tau}\eta\cdot N=\eta\;\;\text{on}\;\;\partial\Gamma^{\tau}_{t}\,,

where (tΔτ)(t+Δτ)(\partial_{t}-\Delta^{\tau})^{\ast}\doteqdot-(\partial_{t}+\Delta^{\tau}) is the formal L2L^{2}-adjoint of the heat operator. Since {Γtτ}t(,τ)\{\Gamma_{t}^{\tau}\}_{t\in(-\infty,-\tau)} converges uniformly in the smooth topology to the stationary interval {[1,1]×{0}}t(,)\{[-1,1]\times\{0\}\}_{t\in(-\infty,\infty)} as τ\tau\to-\infty, we conclude that the limit yy^{\infty} must satisfy (19) in the L2L^{2} sense (and hence in the classical sense due to the L2L^{2} theory for the heat equation). Indeed, by the definition of smooth convergence, we may (after possibly applying a diffeomorphism) parametrize each flow {Γ¯}tτjt(,τj)\{\overline{\Gamma}{}_{t}^{\tau_{j}}\}_{t\in(-\infty,-\tau_{j})} over I[1,1]I\doteqdot[-1,1] by a family of embeddings γtj:I×(,τj)B¯2\gamma^{j}_{t}:I\times(-\infty,-\tau_{j})\to\overline{B}{}^{2} which converge in Cloc(I×(,))C^{\infty}_{\mathrm{loc}}(I\times(-\infty,\infty)) as jj\to\infty to the stationary embedding (x,t)xe1(x,t)\mapsto xe_{1}. Given ηC0(I×(,))\eta\in C^{\infty}_{0}(I\times(-\infty,\infty)) satisfying ηζ(±1)=±η\eta_{\zeta}(\pm 1)=\pm\eta, set ηjφjη\eta^{j}\doteqdot\varphi^{j}\eta, where φj:[1,1]×(,τj)\varphi^{j}:[-1,1]\times(-\infty,-\tau^{j})\to\mathbb{R} is defined by

φζj+(1|γζj|)φj=0,φj(0,t)=1.\varphi^{j}_{\zeta}+(1-|\gamma^{j}_{\zeta}|)\varphi^{j}=0\,,\;\;\varphi^{j}(0,t)=1\,.

That is, φj(ζ,t)=esj(ζ,t)ζ\varphi^{j}(\zeta,t)=\mathrm{e}^{s^{j}(\zeta,t)-\zeta}, where sj(ζ,t)0ζ|γζj(ξ,t)|𝑑ξs^{j}(\zeta,t)\doteqdot\int_{0}^{\zeta}|\gamma^{j}_{\zeta}(\xi,t)|\,d\xi. This ensures that τjηjN=ηj\nabla^{\tau^{j}}\eta^{j}\cdot N=\eta^{j} at the boundary, and hence

0=\displaystyle 0={} Iyτj(tΔτj)ηj𝑑sj𝑑t.\displaystyle\int_{-\infty}^{\infty}\int_{I}y^{\tau_{j}}(\partial_{t}-\Delta^{\tau_{j}})^{\ast}\eta^{j}ds^{j}dt\,.

Since φj1\varphi^{j}\to 1 in Cloc(I×(,))C^{\infty}_{\mathrm{loc}}(I\times(-\infty,\infty)), a short computation reveals that

0=Iy(tΔ)η𝑑ζ𝑑t.0=\int_{-\infty}^{\infty}\int_{I}y^{\infty}(\partial_{t}-\Delta)^{\ast}\eta\,d\zeta dt\,.

Finally, we characterize the limit (uniqueness of which implies full convergence, completing the proof). Separation of variables leads us to consider the problem

{ϕxx=μϕin[1,1]ϕx(±1)=±ϕ(±1).\left\{\begin{aligned} -\phi_{xx}={}&\mu\phi\;\;\text{in}\;\;[-1,1]\\ \phi_{x}(\pm 1)={}&\pm\phi(\pm 1)\,.\end{aligned}\right.

There is only one negative eigenspace, and its frequency turns out to be k1=λ0k_{-1}=\lambda_{0}, with the corresponding mode given by

ϕ1(x)cosh(k1x).\phi_{-1}(x)\doteqdot\cosh(k_{-1}x)\,.

Thus, recalling (18), we are able to conclude that

y(x,t)=Aeλ02tcosh(λ0x)y^{\infty}(x,t)=A\mathrm{e}^{\lambda_{0}^{2}t}\cosh(\lambda_{0}x)

for some A0A\geq 0. In particular,

eλ02τjy(x,τj)=yτj(x,0)Acosh(λ0x)uniformly asj.\mathrm{e}^{-\lambda_{0}^{2}\tau_{j}}y(x,\tau_{j})=y^{\tau_{j}}(x,0)\to A\cosh(\lambda_{0}x)\;\;\text{uniformly as}\;\;j\to\infty\,.

Now, if AA 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 τj\tau_{j} for jj 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 {Γt}t(,0)\{\Gamma_{t}\}_{t\in(-\infty,0)} the constructed ancient solution and let {Γt}t(,0)\{\Gamma^{\prime}_{t}\}_{t\in(-\infty,0)} be a second ancient solution which, without loss of generality, contracts to the point (0,1)(0,1) at time 0. Given any τ>0\tau>0, consider the time-translated solution {Γtτ}t(,τ)\{\Gamma_{t}^{\tau}\}_{t\in(-\infty,-\tau)} defined by Γtτ=Γt+τ\Gamma_{t}^{\tau}=\Gamma^{\prime}_{t+\tau}. By Proposition 3.4,

eλ02tyτ(x,t)Aeλ02τcosh(λ0x)ast\mathrm{e}^{-\lambda_{0}^{2}t}y^{\tau}(x,t)\to A\mathrm{e}^{\lambda_{0}^{2}\tau}\cosh(\lambda_{0}x)\;\;\text{as}\;\;t\to-\infty

uniformly in xx. So Γtτ\Gamma^{\tau}_{t} lies above Γt\Gamma_{t} for t-t sufficiently large. The avoidance principle then ensures that Γtτ\Gamma^{\tau}_{t} lies above Γt\Gamma_{t} for all t(,0)t\in(-\infty,0). Taking τ0\tau\to 0, we find that Γt\Gamma^{\prime}_{t} lies above Γt\Gamma_{t} for all t<0t<0. Since the two curves reach the point (0,1)(0,1) at time zero, they intersect for all t<0t<0 by the avoidance principle. The strong maximum principle then implies that the two solutions coincide for all tt. ∎

4. Supplement: nonexistence of rotators

Free boundary curve shortening flow in B2B^{2} is invariant under rotations about the origin, so it is natural to seek solutions which move by rotation; that is, solutions γ:(L2,L2)×(,)B¯2\gamma:(-\frac{L}{2},\frac{L}{2})\times(-\infty,\infty)\to\overline{B}{}^{2} satisfying

γ(,t)=eiBtγ(,0)\gamma(\cdot,t)=\mathrm{e}^{iBt}\gamma(\cdot,0)

for some B>0B>0. Differentiating yields the rotator equation

(20) κ=Bγ,τ.\kappa=-B\left\langle\gamma,\tau\right\rangle.

It turns out, however, that there are no solutions to (20) in B2B^{2} 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) x=B+xyandy=x2,x^{\prime}=B+xy\;\;\text{and}\;\;y^{\prime}=-x^{2}\,,

where

xBγ,τandyBγ,ν.x\doteqdot B\left\langle\gamma,\tau\right\rangle\;\;\text{and}\;\;y\doteqdot B\left\langle\gamma,\nu\right\rangle\,.

Arc-length parametrized solutions γ\gamma to the rotator equation (20) can be recovered from solutions to the system (21) via

γB1(x+iy)eiθ,θ(s)0sx(σ)𝑑σ,\gamma\doteqdot B^{-1}(x+iy)\mathrm{e}^{i\theta}\,,\;\;\theta(s)\doteqdot-\int_{0}^{s}x(\sigma)d\sigma\,,

and this parametrization is unique up to an ambient rotation and a unit linear reparametrization, i.e. (θ,s)(±θ+θ0,±s+s0)(\theta,s)\mapsto(\pm\theta+\theta_{0},\pm s+s_{0}) .

Note that

|γ|=B1x2+y2.|\gamma|=B^{-1}\sqrt{x^{2}+y^{2}}\,.

So we seek solutions (x,y):(L2,L2)B2(x,y):(-\frac{L}{2},\frac{L}{2})\to B{}^{2} to (21) satisfying the free boundary condition (x(±L2),y(±L2))=(±B,0)(x(\pm\frac{L}{2}),y(\pm\frac{L}{2}))=(\pm B,0).

Let γ\gamma be such a solution. Since (21) can be uniquely solved with initial condition (x(s0),y(s0))=(B,0)(x(s_{0}),y(s_{0}))=(B,0) (which corresponds to γ(s0)B2\gamma(s_{0})\in\partial B^{2} with γ,τ|s0=1\left\langle\gamma,\tau\right\rangle|_{s_{0}}=1), we find that γ\gamma must be invariant under rotation by π\pi about the origin. In particular, the points γ(L2)\gamma(-\frac{L}{2}) and γ(L2)\gamma(\frac{L}{2}) are diametrically opposite. It follows that γ(0)\gamma(0) is the origin. Indeed, for topological reasons, γ\gamma 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

0=y(L2)=0L2y=0L2x2𝑑s,0=y(\tfrac{L}{2})=\int_{0}^{\frac{L}{2}}y^{\prime}=-\int_{0}^{\frac{L}{2}}x^{2}ds\,,

which is impossible since x(L2)=B>0x(\frac{L}{2})=B>0. 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.