\(\newcommand{\footnotename}{footnote}\) \(\def \LWRfootnote {1}\) \(\newcommand {\footnote }[2][\LWRfootnote ]{{}^{\mathrm {#1}}}\) \(\newcommand {\footnotemark }[1][\LWRfootnote ]{{}^{\mathrm {#1}}}\) \(\let \LWRorighspace \hspace \) \(\renewcommand {\hspace }{\ifstar \LWRorighspace \LWRorighspace }\) \(\newcommand {\TextOrMath }[2]{#2}\) \(\newcommand {\mathnormal }[1]{{#1}}\) \(\newcommand \ensuremath [1]{#1}\) \(\newcommand {\LWRframebox }[2][]{\fbox {#2}} \newcommand {\framebox }[1][]{\LWRframebox } \) \(\newcommand {\setlength }[2]{}\) \(\newcommand {\addtolength }[2]{}\) \(\newcommand {\setcounter }[2]{}\) \(\newcommand {\addtocounter }[2]{}\) \(\newcommand {\arabic }[1]{}\) \(\newcommand {\number }[1]{}\) \(\newcommand {\noalign }[1]{\text {#1}\notag \\}\) \(\newcommand {\cline }[1]{}\) \(\newcommand {\directlua }[1]{\text {(directlua)}}\) \(\newcommand {\luatexdirectlua }[1]{\text {(directlua)}}\) \(\newcommand {\protect }{}\) \(\def \LWRabsorbnumber #1 {}\) \(\def \LWRabsorbquotenumber "#1 {}\) \(\newcommand {\LWRabsorboption }[1][]{}\) \(\newcommand {\LWRabsorbtwooptions }[1][]{\LWRabsorboption }\) \(\def \mathchar {\ifnextchar "\LWRabsorbquotenumber \LWRabsorbnumber }\) \(\def \mathcode #1={\mathchar }\) \(\let \delcode \mathcode \) \(\let \delimiter \mathchar \) \(\def \oe {\unicode {x0153}}\) \(\def \OE {\unicode {x0152}}\) \(\def \ae {\unicode {x00E6}}\) \(\def \AE {\unicode {x00C6}}\) \(\def \aa {\unicode {x00E5}}\) \(\def \AA {\unicode {x00C5}}\) \(\def \o {\unicode {x00F8}}\) \(\def \O {\unicode {x00D8}}\) \(\def \l {\unicode {x0142}}\) \(\def \L {\unicode {x0141}}\) \(\def \ss {\unicode {x00DF}}\) \(\def \SS {\unicode {x1E9E}}\) \(\def \dag {\unicode {x2020}}\) \(\def \ddag {\unicode {x2021}}\) \(\def \P {\unicode {x00B6}}\) \(\def \copyright {\unicode {x00A9}}\) \(\def \pounds {\unicode {x00A3}}\) \(\let \LWRref \ref \) \(\renewcommand {\ref }{\ifstar \LWRref \LWRref }\) \( \newcommand {\multicolumn }[3]{#3}\) \(\require {textcomp}\) \(\newcommand {\intertext }[1]{\text {#1}\notag \\}\) \(\let \Hat \hat \) \(\let \Check \check \) \(\let \Tilde \tilde \) \(\let \Acute \acute \) \(\let \Grave \grave \) \(\let \Dot \dot \) \(\let \Ddot \ddot \) \(\let \Breve \breve \) \(\let \Bar \bar \) \(\let \Vec \vec \) \(\require {mathtools}\) \(\newcommand {\vcentcolon }{\mathrel {\unicode {x2236}}}\) \(\newcommand {\approxcolon }{\approx \vcentcolon }\) \(\newcommand {\Approxcolon }{\approx \dblcolon }\) \(\newcommand {\simcolon }{\sim \vcentcolon }\) \(\newcommand {\Simcolon }{\sim \dblcolon }\) \(\newcommand {\dashcolon }{\mathrel {-}\vcentcolon }\) \(\newcommand {\Dashcolon }{\mathrel {-}\dblcolon }\) \(\newcommand {\colondash }{\vcentcolon \mathrel {-}}\) \(\newcommand {\Colondash }{\dblcolon \mathrel {-}}\) \(\newenvironment {crampedsubarray}[1]{}{}\) \(\newcommand {\smashoperator }[2][]{#2\limits }\) \(\newcommand {\SwapAboveDisplaySkip }{}\) \(\newcommand {\LaTeXunderbrace }[1]{\underbrace {#1}}\) \(\newcommand {\LaTeXoverbrace }[1]{\overbrace {#1}}\) \(\Newextarrow \xLongleftarrow {10,10}{0x21D0}\) \(\Newextarrow \xLongrightarrow {10,10}{0x21D2}\) \(\let \xlongleftarrow \xleftarrow \) \(\let \xlongrightarrow \xrightarrow \) \(\newcommand {\LWRmultlined }[1][]{\begin {multline*}}\) \(\newenvironment {multlined}[1][]{\LWRmultlined }{\end {multline*}}\) \(\let \LWRorigshoveleft \shoveleft \) \(\renewcommand {\shoveleft }[1][]{\LWRorigshoveleft }\) \(\let \LWRorigshoveright \shoveright \) \(\renewcommand {\shoveright }[1][]{\LWRorigshoveright }\) \(\newcommand {\shortintertext }[1]{\text {#1}\notag \\}\) \(\newcommand {\R }{\ensuremath {\mathbb {R}}}\) \(\DeclareMathOperator {\sech }{\mathrm {sech}}\) \(\DeclareMathOperator {\area }{\mathrm {area}}\) \(\DeclareMathOperator {\dvg }{\mathrm {div}}\) \(\newcommand {\inner }[2]{\left \langle #1 \, , \, #2\right \rangle }\) \(\newcommand {\norm }[1]{\left \Vert #1\right \Vert }\) \(\newcommand {\pushright }[1]{\ifmeasuring@ #1\else \omit \hfill $\displaystyle #1$\fi \ignorespaces }\)


DOI:10.56994/ARMJ.011.002.006


Received: 4 September 2024; Accepted: 10 March 2025.


Ancient curve shortening flow in the disc with mixed boundary condition

Mat Langford Mathematical Sciences Institute, Australian National University,
Canberra, ACT, Australia
email mathew.langford@anu.edu.au×
, Yuxing Liu Mathematical Sciences Institute, Australian National University,
Canberra, ACT, Australia
email yuxing.liu@anu.edu.au×
, George McNamara Mathematical Sciences Institute, Australian National University,
Canberra, ACT, Australia
email george.mcnamara@anu.edu.au×


Abstract

Given any non-central interior point \(o\) of the unit disc \(D\), the diameter \(L\) through \(o\) is the union of two linear arcs emanating from \(o\) which meet \(\partial D\) orthogonally, the shorter of them stable and the longer unstable (under these boundary conditions). In each of the two half discs bounded by \(L\), we construct a convex eternal solution to curve shortening flow which fixes \(o\) and meets \(\partial D\) orthogonally, and evolves out of the unstable critical arc at \(t=-\infty \) and into the stable one at \(t=+\infty \). We then prove that these two (congruent) solutions are the only non-flat convex ancient solutions to the curve shortening flow satisfying the specified boundary conditions. We obtain analogous conclusions in the “degenerate” case \(o\in \partial D\) as well, although in this case the solution contracts to the point \(o\) at a finite time with asymptotic shape that of a half Grim Reaper, thus providing an interesting example for which an embedded flow develops a collapsing singularity.

1. Introduction

Variational problems subject to boundary constraints are ubiquitous in pure and applied mathematics and physics. One of the simplest such problems is to find and study paths of critical (e.g. minimal) length amongst those joining a given point \(o\) in some domain \(\Omega \) to its boundary \(\partial \Omega \). When \(\Omega \) is a Euclidean domain, such paths are, of course, straight linear arcs from \(o\) to \(\partial \Omega \) which meet \(\partial \Omega \) orthogonally.

While characterizing all such curves is a non-trivial problem in general (even for convex Euclidean domains, say), the “Dirichlet–Neumann geodesics” in the unit disc in \(\R ^2\) are easily found: when \(o\) is the origin, they are the radii; when \(o\) is not the origin, there are exactly two, and their union is the diameter through \(o\).

One useful tool for analyzing such variational problems is the (formal) gradient flow (a.k.a. steepest descent flow), which in this case is the “Dirichlet–Neuman curve shortening flow”; this equation evolves each point of a given curve with velocity equal to the curvature vector at that point, subject to holding one endpoint fixed at \(o\) with the other constrained to \(\partial \Omega \), which is met orthogonally.

While curve shortening flow is now well-studied under other boundary conditions — particularly the “periodic” (i.e. no-boundary) [2] , [3] , [4] , [7] , [9] , [11] , [12] , [13] , [15] , “Neumann–Neumann” (a.k.a. free boundary) [5] , [6] , [8] , [10] , [14] , [16] , [17] , [19] , [20] and “Dirichlet–Dirichlet” [1] , [15] conditions — we are aware of no literature considering the mixed “Dirichlet–Neumann” condition.

Our main result (inspired by [6] ) is the following classification of the convex ancient solutions which arise in the simple setting of the unit disc.

Theorem 1.1.
Given any \(d\in (0,1]\), there exists a convex, locally uniformly convex ancient solution \(\{\Gamma _t^d\}_{t\in (-\infty ,\omega _d)}\) to curve shortening flow in the unit disc \(D\) with one endpoint fixed at \(o:=(-d,0)\) and the other meeting \(\partial D\) orthogonally. The timeslices \(\Gamma ^d_t\) each lie in the upper half-disc, and converge uniformly in the smooth topology as \(t\to -\infty \) to the unstable critical arc \(\{(x,0):x\in [-d,1]\}\); as a graph over the \(x\)-axis,

\[ \mathrm {e}^{\lambda ^2t}y(x,t)\to A\sinh (\lambda (x+d))\;\;\text {uniformly in $x$ as}\;\; t\to -\infty \]

for some \(A>0\), where \(\lambda \) is the positive solution to \(\tanh (\lambda (1+d))=\lambda \).

When \(d<1\), \(\omega _d=+\infty \) and the timeslices converge uniformly in the smooth topology as \(t\to +\infty \) to the minimizing arc \(\{(x,0):x\in [-1,-d]\}\). When \(d=1\), \(\omega _d<\infty \) and the timeslices contract uniformly as \(t\to \omega _d\) to the point \(o\) and, after performing a standard type-II blow-up, converge locally uniformly in the smooth topology to the right half of the downward translating Grim Reaper.

Modulo time translations and reflection about the \(x\)-axis, \(\{\Gamma _t^d\}_{t\in (-\infty ,\omega _d)}\) is the only non-flat convex ancient curve shortening flow subject to the same boundary conditions.

En route to proving Theorem 1.1 , we establish the following convergence result (cf. [1] , [11] , [12] , [13] , [17] ), which is of independent interest (see the proof of Lemma 3.1 ).

Theorem 1.2.
Let \(\Gamma \) be an oriented smooth convex arc in the upper unit half-disc \(D_+\) with left endpoint \(o=(-d,0)\), \(d\in (0,1]\), where its curvature vanishes, and right endpoint on \(\partial D\), which is met orthogonally. Suppose that the curvature of \(\Gamma \) increases monotonically with arclength from \(o\). If \(d<1\), then the Dirichlet–Neumann curve shortening flow starting from \(\Gamma \) exists for all positive time \(t\) and converges uniformly in the smooth topology as \(t\to \infty \) to the minimizing arc joining \(o\) to \(\partial D\). If \(d=1\), then the Dirichlet–Neumann curve shortening flow starting from \(\Gamma \) converges uniformly to the point \(o\) as \(t\to \omega <\infty \) and, after performing a standard type-II blow-up, converges locally uniformly in the smooth topology to a half Grim Reaper.

Though the curvature monotonicity hypothesis appears unnaturally restrictive in Theorem 1.2 , we note that some such additional condition is required to prevent the development of self-intersections at the Dirichlet endpoint (resulting in subsequent cusplike singularities). Moreover, as Theorem 1.2 demonstrates in case the Dirichlet endpoint lies on the boundary, collapsing singularities may form at the Dirichlet endpoint even when the flow remains embedded. It is not hard to see that this can also occur when the Dirichlet endpoint lies to the interior (as a limiting case of the flow forming a cusp singularity just after losing embeddedness, say).

Acknowledgements.

This work originated as an undergraduate summer research project undertaken at the ANU by YL and GM under the supervision of ML. We are grateful to the MSI for funding the project. ML thanks Julie Clutterbuck for bringing the mixed boundary value problem to his attention.

2. Preliminaries

Fix a point \(o=(-d,0)\in D\) in the unit disc \(D\subset \R ^2\), with \(d\in (0,1]\). Denote by \(C_\theta \subset D\) the circular arc which passes through \(o\) and meets the boundary of \(D\) orthogonally at \((\sin \theta ,\cos \theta )\); that is,

\[ C_\theta :=\{(x,y)\in D:(x-\xi )^2+(y-\eta )^2=r^2\}\,, \]

where, defining \(a:=\frac {1}{2}(d^{-1}+d)\),

\[ (\xi ,\eta ):=(\cos \theta ,\sin \theta )+r(-\sin \theta ,\cos \theta )\;\; \text { and}\;\; r:=\frac {1+d^2+2d\cos \theta }{2d\sin \theta }=\frac {a+\cos \theta }{\sin \theta }\,. \]

Consider also the circular arc \(\check C_{\theta }\subset D\) which is symmetric about the \(y\)-axis and meets \(\partial D\) orthogonally at \((\cos \theta ,\sin \theta )\). That is,

\[ \check C_\theta :=\{x^2+(y-\check \eta )^2=\check r^2\}\,, \]

where

\[ \check \eta :=\csc \theta \;\;\text {and}\;\;\check r:=\cot \theta \,. \]

Proposition 2.1.
The family \(\{\check C_{\theta ^+(t)}\}_{t\in (-\infty ,0)}\), where \(\theta ^+(t):=\arcsin \mathrm {e}^{2t}\), is a supersolution to curve shortening flow. The family \(\{C_{\theta ^-(t)}\}_{t\in (-\infty ,\omega _d)}\), where \(\theta ^-\) is the solution to

\begin{equation} \label {eq:characteristic ODE} \left \{\begin{aligned}\frac {d\theta }{dt}={}&\frac {\sin \theta }{a+\cos \theta }\\ \theta (0)={}&\tfrac {\pi }{2}\,,\end {aligned}\right . \end{equation}

is a subsolution to curve shortening flow.

Remark 1.
Separating variables, the problem ( 1 ) becomes

\[ \int _t^0dt=\int _{\theta }^{\frac {\pi }{2}}\frac {a+\cos \omega }{\sin \omega }d\omega =\int _{\omega =\theta }^{\frac {\pi }{2}}d\log \left (2\sin ^{1+a}\left (\tfrac {\omega }{2}\right )\cos ^{1-a}\left (\tfrac {\omega }{2}\right )\right )\,, \]

and hence

\[ \mathrm {e}^{t}=2\sin ^{1+a}\left (\tfrac {\theta ^-(t)}{2}\right )\cos ^{1-a}\left (\tfrac {\theta ^-(t)}{2}\right )\,. \]

In particular, for all \(d\in (0,1]\), the solution certainly exists for all \(t<0\), with \(\theta ^-(t)\sim 2^{\frac {a}{1+a}}\mathrm {e}^{\frac {t}{a+1}}\) as \(t\to -\infty \). When \(d\in (0,1)\), the solution exists up to time \(\omega _d=+\infty \), and \(\lim _{t\to +\infty }\theta ^-(t)=\pi \). When \(d=1\), the solution exists up to time \(\omega _d=\log 2\), and \(\lim _{t\to \omega _d}\theta ^-(t)=\pi \).

Proof of Proposition 2.1 .

The first claim is proved in [6], Proposition 2.1 .

To prove the second claim, consider any monotone increasing function \(\theta \) of \(t\), and let \(\gamma (u,t)=(x(u,t),y(u,t))\) be a general parametrization of \(C_{\theta (t)}\). Differentiation of the equation

\[ (x-\xi )^2+(y-\eta )^2=r^2 \]

with respect to \(t\) along \(\gamma \) and \(\theta \) yields

\[ (x-\xi )(x_t-\xi _\theta \theta _t)+(y-\eta )(y_t-\eta _\theta \theta _t)=rr_\theta \theta _t\,. \]

Since the outward unit normal to \(C_\theta \) at \((x,y)\) is \(\nu =\frac {1}{r}(x-\xi ,y-\eta )\), this becomes

\[ -\gamma _t\cdot \nu =-\left (\frac {x-\xi }{r}\xi _\theta +\frac {y-\eta }{r}\eta _\theta +rr_\theta \right )\theta _t\,. \]

We claim that

\[ \frac {1}{r}(x-\xi ,y-\eta )\cdot (\xi _\theta ,\eta _\theta )+r_\theta =-\frac {y}{\sin \theta }\,. \]

Indeed,

\begin{align*} \frac {1}{r}({}&x-\xi ,y-\eta )\cdot (\xi _\theta ,\eta _\theta )\\ ={}&\frac {1}{r}\Big ((x,y)-(\cos \theta ,\sin \theta )-r(-\sin \theta ,\cos \theta ))\Big )\cdot \Big ((1+r_\theta )(-\sin \theta ,\cos \theta )-r(\cos \theta ,\sin \theta )\Big )\\ ={}&-\left ((x,y)-(\cos \theta ,\sin \theta )-r(-\sin \theta ,\cos \theta ))\right )\cdot \left (\cot \theta (-\sin \theta ,\cos \theta )+(\cos \theta ,\sin \theta )\right )\\ ={}&-(x,y)\cdot \left (\cot \theta (-\sin \theta ,\cos \theta )+(\cos \theta ,\sin \theta )\right )+1+r\cot \theta \\ ={}&-(x,y)\cdot (0,\csc \theta )-r_\theta \,, \end{align*} from which the claim follows. Since, \(y\le \sin \theta \) along \(C_\theta \), taking \(\theta \) to be the solution to the specified initial value problem yields

\[ -\gamma _t\cdot \nu =\frac {y}{\sin \theta }\theta _t=\frac {y}{\sin \theta }\frac {1}{r}\le \frac {1}{r}=\kappa \,, \]

as claimed.

Next consider \(\{\mathrm {H}_t\}_{t\in (-\infty ,\infty )}\), the fundamental domain of the horizontally oriented hairclip solution to curve shortening flow centred at \(o\); that is,

\[ \mathrm {H}_t:=\{(x,y)\in [0,\infty )\times [0,\tfrac {\pi }{2}]:\sin (y)=\mathrm {e}^{t}\sinh (x+d)\}\,. \]

(image)
Figure 1. Some timeslices of (one period of) the “hairclip” solution.

Given any \(\lambda >0\), define \(\{\mathrm {H}^\lambda _t\}_{t\in (-\infty ,\infty )}\) by parabolically rescaling the hairclip by \(\lambda \). That is,

\[ \mathrm {H}^\lambda _t:=\lambda ^{-1}\mathrm {H}_{\lambda ^{2}t}=\{(x,y)\in [0,\infty )\times [0,\tfrac {\pi }{2\lambda }]:\sin (\lambda y)=\mathrm {e}^{\lambda ^2t}\sinh (\lambda (x+d))\}\,. \]

Observe that \(\{\mathrm {H}^\lambda _t\}_{t\in (-\infty ,\infty )}\) satisfies

\[ \frac {\kappa }{\cos \theta }=\lambda \tan (\lambda y)\;\;\text {and}\;\;\frac {\kappa }{\sin \theta }=\lambda \tanh (\lambda (x+d))\,, \]

where \(\theta \in [0,\frac {\pi }{2}]\) is the angle the tangent vector makes with the \(x\)-axis. From this we see, in particular, that \(\kappa \) is positive and monotone increasing with respect to arclength from \(o\).

Proposition 2.2.
For each \(\theta \in (0,\frac {\pi }{2})\) there exists a unique pair \((\lambda ,t)\) such that \(\mathrm {H}^\lambda _t\) intersects \(\partial D\) orthogonally at \((\cos \theta ,\sin \theta )\).

Proof.

Given any \(\theta \in (0,\frac {\pi }{2})\), substituting the point \((\cos \theta ,\sin \theta )\) for \((x,y)\) in the defining equation \(\sin (\lambda y)=\mathrm {e}^{\lambda ^2t}\sinh (\lambda (x+d))\) and solving for \(t\) yields for each \(\lambda \in (0,\frac {\pi }{2\sin \theta })\) the unique timeslice of the (fundamental domain of the) Hairclip solution which intersects \(\partial D\) at \((\cos \theta ,\sin \theta )\); namely,

\[ t=\lambda ^{-2}\ln \left (\frac {\sin (\lambda \sin \theta )}{\sinh (\lambda (\cos \theta +d))}\right )\,. \]

At that point, the normal satisfies

\begin{align*} \nu _\lambda (\cos \theta ,\sin \theta )\cdot (\cos \theta ,\sin \theta )={}&\frac {\sin (\lambda \sin \theta )\cos \theta -\tanh (\lambda (\cos \theta +d))\cos (\lambda \sin \theta )\sin \theta }{\tanh (\lambda (\cos \theta +d))\cos (\lambda \sin \theta )}\\ ={}&-\frac {\tan (\lambda \sin \theta )\cos \theta }{\tanh (\lambda (\cos \theta +d))}g(\lambda ,\theta )\,, \end{align*} where

\[ g(\lambda ,\theta ):=\tanh (\lambda (\cos \theta +d))\cot (\lambda \sin \theta )\tan \theta -1\,. \]

Observe that

\[ \lim _{\lambda \searrow 0}g(\lambda ,\theta )=d\cdot \sec \theta >0, \ \lim _{\lambda \nearrow \frac {\pi }{2\sin \theta }}g(\lambda ,\theta )=-1<0 \]

and

(2–1) \{begin}{align*} \frac {dg}{d\lambda }={}&\tan \theta \Big [(\cos \theta +d)\cot (\lambda \sin \theta )\sech ^2(\lambda (\cos \theta +d))\\ {}&\pushright {-\sin \theta \csc ^2(\lambda \sin \theta )\tanh (\lambda
(\cos \theta +d))\Big ]}\\ ={}&\frac {\tan \theta \tanh (\lambda (\cos \theta +d))}{\lambda \sin (\lambda \sin \theta )}\left [\frac {\lambda (\cos \theta +d)}{\sinh (\lambda (\cos \theta +d))}\cos (\lambda \sin \theta )\sech (\lambda
(\cos \theta +d))\right .\\ {}&\pushright {\left .-\frac {\lambda \sin \theta }{\sin (\lambda \sin \theta )}\right ]}\\ \le {}&\frac {\tan \theta \tanh (\lambda (\cos \theta +d))}{\lambda \sin (\lambda \sin \theta )}\left [\cos
(\lambda \sin \theta )\sech (\lambda (\cos \theta +d))-1\right ]\\ <{}&0 \{end}{align*}

for \(\lambda \ \in \ (0,\frac {\pi }{2\sin \theta })\). It follows that there exists a unique \(\lambda \) such that

\[ \nu _\lambda (\cos \theta ,\sin \theta )\cdot (\cos \theta ,\sin \theta )=0\,. \]

The claim follows.

Remark 2.
Note that, since \(\lim _{\theta \to 0}g(\lambda ,\theta )=\frac {\tanh (\lambda (d+1)))}{\lambda }-1\), the function \(g(\lambda ,\theta )\) is non-negative at \(\theta =0\) so long as \(\lambda \geq \lambda _0\), where \(\lambda _0\) is the unique positive solution to the equation

\[ \lambda =\tanh (\lambda (d+1))\,. \]

A priori estimates">
A priori estimates" class="Theorem TheoremMain">   
Proposition 2.3 ( A priori estimates) . A priori estimates'); rotateIcon(this);">
Let \(\Gamma \subset D_+\) be a smooth, convex embedding of a closed interval with left endpoint \(o=(-d,0)\), \(d\in (0,1]\), and right endpoint meeting \(\partial D\) orthogonally, and suppose that the curvature \(\kappa \) of \(\Gamma \) increases monotonically with respect to arclength from \(o\). Denote by \(\underline \theta \) resp. \(\overline \theta \) the least resp. greatest value taken by the turning angle along \(\Gamma \) and by \(\overline \kappa =\kappa (\overline \theta )\) the greatest value taken by \(\kappa \).

The circle \(C_{\overline \theta }\) lies below \(\Gamma \). Thus,

\begin{equation} \label {eq:curvature lower bound} \overline \kappa \ge \frac {\sin \overline \theta }{a+\cos \overline \theta } \end{equation}

and

\begin{equation} \label {eq:gradient lower bound} \underline \theta \ge \mathrm {arccot}\left (\frac {1+a\cos \overline \theta }{b\sin \overline \theta }\right )\,, \end{equation}

where \(b:=\frac {1}{2}(d^{-1}-d)\) and we recall that \(a:=\frac {1}{2}(d^{-1}+d)\), with the right hand side taken to be zero in case \(d=1\).

Proof.

Suppose, to the contrary, that \(C_{\overline \theta }\) does not lie below \(\Gamma \). Then some point of \(\Gamma \) must lie strictly below \(C_{\overline \theta }\), and hence (since the endpoints of the two curves agree) upon translating \(C_{\overline \theta }\) downwards, the two curves will continue to intersect until some final moment, at which they must make first order contact at some interior point \(q\in \Gamma \). At this point, the curvature \(\kappa \) of \(\Gamma \) must be no less than \(1/r(\overline \theta )\) (the curvature of \(C_{\overline \theta }\)). But then, by the monotonicity of \(\kappa \), \(\kappa \) must exceed \(1/r(\overline \theta )\) on the whole segment of \(\Gamma \) joining \(q\) to \(\partial D\), in which case (since \(\Gamma \) and \(C_{\overline \theta }\) make first order contact at \(\partial D\)) the point \(q\) must lie strictly above \(C_{\overline \theta }\), which is absurd. So \(C_{\overline \theta }\) must indeed lie below \(\Gamma \). The first inequality is then immediate and the second is straightforward.


Figure 2. Scaled hairclip timeslice and circular arcs through the prescribed boundary points \(o\) and \((\cos \theta ,\sin \theta )\).

3. Existence

For each \(d\in (0,1]\) and \(\rho \in (0,\frac {\pi }{2})\), let \(\Gamma ^\rho \subset D_+\) be a smooth oriented arc satisfying the following properties.


  • The left endpoint of \(\Gamma ^\rho \) is \(o=(-d,0)\), where its curvature vanishes, and its right endpoint meets \(\partial D\) orthogonally at \((\cos \rho ,\sin \rho )\).
  • \(\Gamma ^\rho \) is convex.
  • The curvature of \(\Gamma ^\rho \) is monotone increasing with respect to arclength from \(o\).

For example, we could take \(\Gamma ^\rho :=\mathrm {H}^{\lambda _\rho }_{t_\rho }\cap D\), where \((\lambda _\rho ,t_\rho )\) are the unique choice of \((\lambda ,t)\) which ensure that \(\mathrm {H}_t^\lambda \) meets \(\partial D\) orthogonally at \((\cos \rho ,\sin \rho )\).

  
Lemma 3.1 (Very old (but not ancient) solutions) .
For each \(d\in (0,1]\) and \(\rho \in (0,\frac {\pi }{2})\) there exist \(\alpha _\rho <0\) such that \(\alpha _\rho \to -\infty \) as \(\rho \to 0\) and a smooth 1 curve shortening flow \(\{\Gamma ^\rho _t\}_{t\in [\alpha _\rho ,\omega _d)}\) in \(D\) exhibiting the following properties.

  • \(\Gamma ^\rho _{\alpha _\rho }=\Gamma ^\rho \).
  • For each \(t\in [\alpha _\rho ,\omega _d)\), \(\Gamma ^\rho _t\) is an oriented embedding of a closed interval, with left endpoint \(o=(-d,0)\) and right endpoint meeting \(\partial D\) orthogonally.
  • For each \(t\in [\alpha _\rho ,\omega _d)\), \(\Gamma ^\rho _t\) is convex.
  • For each \(t\in [\alpha _\rho ,\omega _d)\), the curvature of \(\Gamma ^\rho _t\) is monotone increasing with respect to arclength from \(o\).
  • If \(d<1\), then \(\omega _d=\infty \) and \(\Gamma ^\rho _t\) converges uniformly in the smooth topology as \(t\to \infty \) to the minimizing arc \(\{(x,0):x\in [-1,-d]\}\).
  • If \(d=1\), then \(\omega _d\in (0,\infty )\) and \(\Gamma ^\rho _t\) converges uniformly as \(t\to \omega _d\) to the point \(o\), and there are sequences of times \(t_j\nearrow \omega _d\), points \(p_j\in \Gamma ^\rho _{t_j}\), and scales \(\lambda _j\nearrow \infty \) such that the sequence \(\{\lambda _j(\Gamma ^\rho _{\lambda _j^{-2}t+t_j}-p_j)\}_{t\in [\lambda _j^2(\alpha _\rho -t_j),\lambda _j^2(\omega _\rho -j^{-1}-t_j))}\) converges locally uniformly in the smooth topology as \(j\to \infty \) to the right half of the downwards translating Grim Reaper.

Proof.

Form the “odd doubling” \(\check \Gamma ^\rho \) of \(\Gamma ^\rho \) by taking the union of \(\Gamma ^\rho \) with its rotation through angle \(\pi \) about \(o\). Since \(\check \Gamma ^\rho \) is a regular curve of class \(C^2\) and there exists a ball \(B\) about \((\cos \rho ,\sin \rho )\) (of radius \(1/10\), say) which is disjoint from the rotation of \(\Gamma ^\rho \) through angle \(\pi \) about \(o\), where \(\Gamma ^\rho \) meets \(\partial D\) orthogonally, Stahl's short-time existence theorem for free boundary mean curvature flow [20], Theorem 2.1 yields a solution \(\{\check \Gamma ^\rho _t\}_{t\in [0,\delta )}\) to Neumann–Neumann curve shortening flow with boundary on the odd doubling of \(\partial D\cap B\) for a short time \(\delta >0\). Since this solution is uniquely determined by its initial condition, it must be invariant under rotation through angle \(\pi \) about \(o\), and hence descend to a solution \(\{\hat \Gamma ^\rho _t\}_{t\in [0,\delta )}\) to Dirichlet–Neumann curve shortening flow in \(D\) with Dirichlet condition \(o\) and initial condition \(\Gamma ^\rho \). Denote by \(T\) the maximal time of existence of the latter.

Since the curvature of \(\{\hat \Gamma ^\rho _t\}_{t\in [0,T)}\) satisfies

\[ \left \{\begin {aligned}(\partial _t-\Delta )\kappa ={}&\kappa ^3\\ \kappa ={}&0\;\;\text {at $o$, and}\\ \kappa _s={}&\kappa \;\;\text {at}\;\;\partial D\,, \end {aligned}\right . \]

where \(s\) denotes arclength from \(o\), the maximum principle (and Hopf boundary point lemma) ensure that \(\kappa \) remains positive on \(\hat \Gamma ^\rho _t\setminus \{o\}\) for \(t>0\).

For similar reasons, positivity of \(\kappa _s\) is also preserved. Indeed, using the commutator relation

\[ [\partial _t,\partial _s]=\kappa ^2\partial _s\,, \]

the identity \(0=\kappa _t=\Delta \kappa \) at \(o\), and the positivity of \(\kappa \) away from \(o\), we find that

\[ \left \{\begin {aligned}(\partial _t-\Delta )\kappa _s={}&4\kappa ^2\kappa _s\\ (\kappa _s)_s={}&0\;\;\text {at $o$, and}\\ \kappa _s>{}&0\;\;\text {at}\;\;\partial D\,, \end {aligned}\right . \]

so the claim once again follows from the maximum principle.

Since \(\overline \theta _t=\overline \kappa >0\) and \(\overline \theta <\pi \) (when \(d<1\), the maximum principle prevents \(\hat \Gamma ^\rho _t\) from ever reaching the minimizing arc — a stationary solution to the flow) we find that \(\overline \theta \) must attain a limit as \(t\to T\). We claim that this limit is \(\pi \). Indeed, if \(\overline \theta \le \theta _0<\pi \) for all \(t\in [0,T)\), then, representing the solution as a graph over the line \(\{(-d,y):y\in \R \}\), the “gradient estimate” ( 3 ) yields a uniform bound for the gradient, at least when \(d<1\). But then, by applying parabolic regularity theory (see, for instance, [18] ) to the graphical Dirichlet–Neumann curve shortening flow equation

\[ \left \{\begin {aligned}x_t={}&\frac {x_{yy}}{1+x_y^2}\;\;\text {in}\;\; [0,\overline y(t)]\\ x(0,t)={}&0\\ x_y(\overline y(t),t)={}&\cot \overline \theta (t)\,, \end {aligned}\right . \]

where \(\overline y(t):=\sin \overline \theta (t)\), we obtain uniform estimates for all derivatives of the graph functions \(x(\cdot ,t)\) (cf. [20] ). To obtain corresponding estimates when \(d=1\), we instead represent the solution as a graph over the “tilted” line through \((-1,0)\) and \((\cos (\overline \theta (T)),\sin (\overline \theta (T)))\) and use the “gradient estimate” \(\underline \theta \ge 0\). The Arzelà–Ascoli theorem and monotonicity of the flow now ensure that \(x(\cdot ,t)\) takes a smooth limit as \(t\to T\), at which point the flow can be smoothly continued by the above short time existence argument, violating the maximality of \(T\). We conclude that \(\overline \theta (t)\to \pi \) as \(t\to T\).

It now follows from ( 3 ) that \(\underline \theta (t)\to \pi \) as \(t\to T\). When \(d=1\), we conclude that \(\hat \Gamma ^\rho _t\) contracts to \(o\) as \(t\to T\). Note that in this case \(T<\infty \) since the lower barriers \(C_{\theta ^-(t)}\) contract to \(o\) in finite time. A more or less standard “type-I vs type-II" blow-up argument (cf. [17] ) then guarantees convergence to the half Grim Reaper after performing a standard type-II blow-up. (The flow must be type-II because the limit of a standard type-I blow-up — a shrinking semi-circle — violates the Dirichlet boundary condition.)

When \(d<1\), we conclude that \(\hat \Gamma ^\rho _t\) converges uniformly to the minimizing arc \(\{(x,0):x\in [-1,-d]\}\) as \(t\to T\). But then, for large enough \(t\), \(\hat \Gamma ^\rho _t\) may be represented as a graph over the \(x\)-axis with small gradient, at which point parabolic regularity, short-time existence and the Arzelà–Ascoli theorem guarantee that \(T=\infty \) and \(\hat \Gamma ^\rho _t\) converges uniformly in the smooth topology to the minimizing arc.

Finally, since \(\overline \theta \) is monotone, there is a unique time \(-\alpha _\rho >0\) such that \(\overline \theta (-\alpha _\rho )=\frac {\pi }{2}\); since the Neumann–Neumann circle \(\check C_{\theta _\rho }\), where \(\sin \theta _\rho =\frac {2\sin \rho }{1+\sin ^2\rho }\), lies above \(\Gamma ^\rho \), we find (by suitably time translating the upper barrier \(\{\check C_{\theta ^+(t)}\}_{t\in (-\infty ,0)}\), as in [6] ) that

\[ \alpha _\rho <\frac {1}{2}\log \left (\frac {2\sin \rho }{1+\sin ^2\rho }\right )\,. \]

Time-translating the solution \(\{\hat \Gamma ^\rho \}_{t\in [0,\infty )}\) by \(\alpha _\rho \) now yields the desired very old (but not ancient) solution \(\{\Gamma ^\rho _t\}_{t\in [\alpha _\rho ,\infty )}\).

Taking the limit as \(\rho \to 0\) of these very old (but not ancient) solutions yields our desired ancient solution.

Theorem 3.2.
Given any \(d\in (0,1]\), there exists a convex ancient Dirichlet–Neumann curve shortening flow \(\{\Gamma _t\}_{t\in (-\infty ,\omega _d)}\) in the upper half disc \(D_+\) which converges uniformly in the smooth topology to the unstable critical arc \([-d,1]\times \{0\}\) as \(t\to -\infty \). When \(d<1\), \(\omega _d=\infty \) and \(\{\Gamma _t\}_{t\in (-\infty ,\omega _d)}\) converges uniformly in the smooth topology as \(t\to +\infty \) to the minimizing arc \([-1,-d]\times \{0\}\). When \(d=1\), \(\omega _d<\infty \) and \(\Gamma ^\rho _t\) converges uniformly as \(t\to \omega _d\) to the point \(o\), and there are a sequence of times \(t_j\nearrow \omega _d\), right endpoints \(p_j\in \Gamma ^\rho _{t_j}\), and scales \(\lambda _j\nearrow \infty \) such that the sequence \(\{\lambda _j(\Gamma ^\rho _{\lambda _j^{-2}t+t_j}-p_j)\}_{t\in [\lambda _j^2(\alpha _\rho -t_j),\lambda _j^2(\omega _\rho -j^{-1}-t_j))}\) converges locally uniformly in the smooth topology as \(j\to \infty \) to the right half of the downwards translating Grim Reaper.

Proof.

Given any sequence of angles \(\rho _j\searrow 0\), consider the sequence of corresponding very old (but not ancient) solutions \(\{\Gamma ^j_t\}_{t\in [\alpha _j,\infty )}\) constructed in Lemma 3.1 . Differentiating the Neumann boundary condition and applying the estimate ( 2 ) yields the inequality

\[ \overline \theta _t=\overline \kappa \ge \frac {\sin \overline \theta }{a+\cos \overline \theta } \]

on each of these solutions. It follows, by the ODE comparison principle, that each \(\{\Gamma ^j_t\}_{t\in [\alpha _j,\infty )}\) satisfies

\begin{equation} \label {eq:gradient upper bound} \overline \theta \le \theta ^- \end{equation}

for \(t\in [\alpha _j,0]\), where we recall that \(\theta ^-\) is the solution to ( 1 ). Since \(\theta ^-\) is independent of \(j\), this implies uniform estimates for the gradient on any time interval of the form \([-\infty ,-T]\), \(T>0\), when we represent \(\{\Gamma ^j_t\}_{t\in [\alpha _j,-T]}\) graphically over the \(x\)-axis. By parabolic regularity theory and the Arzelà–Ascoli theorem, we may then extract a smooth limit of the very old solutions \(\{\Gamma ^{j}_t\}_{t\in (-\infty ,0)}\) after passing to a subsequence. This limit is ancient, since \(\alpha _\rho \to -\infty \) as \(\rho \to 0\), reaches the point \((0,1)\) at time zero (since each \(\Gamma ^{j}_t\) intersects the convex domain bounded by \(\check C_{\theta ^+(t)}\) for each \(t\in (-\infty ,0)\)), and converges uniformly in the smooth topology as \(t\to -\infty \) to the unstable Dirichlet–Neumann geodesic from \(o\) due to the estimate ( 4 ) (and parabolic regularity theory). The longtime behaviour follows from the argument presented above.

3.1. Asymptotics.

We now prove precise asymptotics for the height of the ancient solution constructed in Theorem 3.2 , assuming the initial conditions for the old-but-not-ancient solutions \(\{\Gamma ^\rho _t\}_{t\in [\alpha _\rho ,\omega _d)}\) are given by the hairclip timeslices \(\Gamma ^\rho =\mathrm {H}^{\lambda _\rho }_{t_\rho }\cap D\).

Lemma 3.3.
On each old-but-not-ancient solution \(\{\Gamma ^\rho _t\}_{t\in [\alpha _\rho ,\omega _d)}\),

\begin{equation*} \frac {\kappa }{\cos \theta }\ge \lambda _\rho \tan (\lambda _\rho y)\,. \end{equation*}

Proof.

Note that equality holds on the initial curve \(\Gamma ^\rho =\mathrm {H}^{\lambda _\rho }_{t_\rho }\cap D\). Thus, given any \(\mu <\lambda _\rho \), the function

\[ w:=\frac {\kappa }{\cos \theta }-\mu \tan (\mu y) \]

is strictly positive on the initial curve \(\Gamma ^\rho \), except at the left endpoint, where it vanishes. Observe that

\[ w_s= \frac {\kappa _s}{\cos \theta }+\sin \theta \left (\frac {\kappa ^2}{\cos ^2\theta }-\mu ^2\sec ^2(\mu y)\right )\,. \]

In particular, at the left endpoint on the initial curve,

\begin{align*} w_s={}&\frac {\kappa _s}{\cos \theta }-\mu ^2\sin \theta =(\lambda _\rho ^2-\mu ^2)\sin \theta >0\,. \end{align*} Thus (since \(w_s\) is continuous at \(o\) at time zero), if \(w\) fails to remain non-negative at positive times, then this failure must occur immediately following some interior time \(t_\ast >0\). There are three possibilities: 1. \(w_s(\cdot ,t_\ast )=0\) at the left endpoint, 2. \(w(\cdot ,t_\ast )=0\) at the right endpoint; or 3. \(w(\cdot ,t_\ast )=0\) at some interior point, \(p_\ast \).

The first of the three possibilities is immediately ruled out by the Hopf boundary point lemma.

In the second case, the Hopf boundary point lemma and the Neumann boundary condition yield, at the right endpoint,

\begin{align*} 0>w_s ={}&\frac {\kappa }{\cos \theta }+\sin \theta \left (\frac {\kappa ^2}{\cos ^2\theta }-\mu ^2(1+\tan ^2(\mu y))\right )=\mu \tan (\mu y)-\mu ^2y\ge 0\,, \end{align*} which is absurd.

In the final case (having ruled out the first two), \(w\) must attain a negative interior minumum just following time \(t_\ast \). But at such a point, \(w<0\), \(w_s=0\) and

\begin{align*} 0\ge {}& (\partial _t-\Delta )w\\ ={}&\frac {(\partial _t-\Delta )\kappa }{\cos \theta }-\frac {\kappa (\partial _t-\Delta )\cos \theta }{\cos ^2\theta }+2\left (\frac {\kappa }{\cos \theta }\right )_s\frac {(\cos \theta )_s}{\cos \theta }-(\partial _t-\Delta )(\mu \tan (\mu y))\,. \end{align*} Since

\[ (\partial _t-\Delta )\kappa =\kappa ^3\,,\;\;(\partial _t-\Delta )\cos \theta =\kappa ^2\cos \theta \;\;\text {and}\;\;(\partial _t-\Delta )y=0\,, \]

we conclude that

\begin{align*} 0\ge {}&-2\left (\frac {\kappa }{\cos \theta }\right )_s\kappa \tan \theta +2\mu \tan (\mu y)(\mu \tan (\mu y))_s\sin \theta \\ ={}&2(\mu \tan (\mu y))_s\sin \theta \left (\mu \tan (\mu y)-\frac {\kappa }{\cos \theta }\right )\\ ={}&2\mu ^2\sec ^2(\mu y)\left (\mu \tan (\mu y)-\frac {\kappa }{\cos \theta }\right )\\ >{}&0\,, \end{align*} which is absurd.

Having ruled out each of the three possibilities, we conclude that \(w\ge 0\) for any \(\mu <\lambda _\rho \). The claim follows.

In the limit as \(\rho \to 0\), we then obtain

\begin{equation} \label {eq:sharp speed lower bound} \frac {\kappa }{\cos \theta }\ge \lambda _0\tan (\lambda _0y) \end{equation}

on the ancient solution, where \(\lambda _0=\lim _{\rho \to 0}\lambda _\rho \). We now find that, as a graph over the \(x\)-axis (for \(t\) sufficiently negative),

\[ (\sin (\lambda _0y))_t=\lambda \cos (\lambda _0 y)y_t=\lambda _0\cos (\lambda _0 y)\sqrt {1+y_x^2}\kappa =\lambda _0\cos (\lambda _0 y)\frac {\kappa }{\cos \theta }\ge \lambda _0^2\sin (\lambda y)\,, \]

and hence

\[ \left (\mathrm {e}^{-\lambda _0^2t}\sin (\lambda _0 y)\right )_t\ge 0\,, \]

which implies that the limit

\[ A(x):=\lim _{t\to -\infty }\mathrm {e}^{-\lambda _0^2t}y(x,t) \]

exists in \([0,\infty )\) for each \(x\in (-d,1)\).

Recall that \(\theta ^-(t)\sim \mathrm {e}^{\frac {t}{a+1}}\) for \(t\sim -\infty \). In particular, \(\overline \theta (t)\le \theta ^-(t)\) is integrable. We will exploit this fact to show that the limit \(A(x)\) is positive (at least near \(x=1\)). First, we shall show that \(\overline \kappa \) is integrable.

Lemma 3.4.
There exist \(\rho _0>0\), \(T>-\infty \), \(C<\infty \) and \(\delta >0\) such that

\[ \kappa \le C\mathrm {e}^{\delta t}\;\;\text {for}\;\; t\le T \]

on each old-but-not-ancient solution \(\{\Gamma ^\rho _t\}_{t\in [\alpha _\rho ,\omega _d)}\) with \(\rho <\rho _0\).

Proof.

Since

\[ (\partial _t-\Delta )\sin \theta =\kappa ^2\sin \theta \,, \]

we find that

\[ (\partial _t-\Delta )\frac {\kappa }{\sin \theta }=2\nabla \frac {\kappa }{\sin \theta }\cdot \frac {\nabla \sin \theta }{\sin \theta }\,. \]

So the maximum principle guarantees that the maximum of \(\frac {\kappa }{\sin \theta }\) occurs at the parabolic boundary. Now, at the left boundary point, \(\frac {\kappa }{\sin \theta }=0\), while at the right,

\begin{align*} \left (\frac {\kappa }{\sin \theta }\right )_s={}&\frac {\kappa }{\sin \theta }\left (\frac {\kappa _s}{\kappa }-\frac {\cos \theta \kappa }{\sin \theta }\right )=\frac {\overline \kappa }{\sin \overline \theta }\left (1-\frac {\overline \kappa }{\tan \overline \theta }\right )\,. \end{align*} By ( 4 ), we can find \(T>-\infty \) (independent of \(\rho \)) so that \(\cos \overline \theta (t)\ge \frac {1}{2}\) for all \(t\le T\). We thereby conclude that

\[ \frac {\kappa }{\sin \theta }\le \max \left \{2,\max _{t=\alpha _\rho }\frac {\kappa }{\sin \theta }\right \} \]

for all \(t\le T\). Since \(\max _{t=\alpha _\rho }\frac {\kappa }{\sin \theta }\le \lambda _\rho \tanh (\lambda _\rho (1+d))\to \lambda _0^2<1\) as \(\rho \to 0\), we find that

\begin{equation} \label {eq:C2 estimate constructed solution} \overline \kappa \le 2\sin \overline \theta \le 2\sin \theta ^- \end{equation}

for all \(\rho \) sufficiently small. The claim follows since \(\theta ^-\) is comparable to \(2^{1+\frac {a}{1+a}}\mathrm {e}^{\frac {t}{1+a}}\) as \(t\to -\infty \).

Corollary 3.5.
There exist \(T>-\infty \), \(C<\infty \) and \(\delta >0\) such that

\[ \frac {\kappa }{y}\le \lambda _0^2+C\mathrm {e}^{\delta t}\;\;\text {for}\;\; t<T \]

on the ancient solution.

Proof.

Given \(\rho <\rho _0\), set

\[ \eta _\rho (t):=\lambda _\rho ^2\left (\exp \big (\tfrac {C^2}{2\delta }\mathrm {e}^{2\delta t}\big )-1\right )\,, \]

where \(\rho _0\), \(C\) and \(\delta \) are the constants from Lemma 3.4 , so that

\[ \frac {\eta _\rho '}{\lambda _\rho ^2+\eta _\rho }=C^2\mathrm {e}^{2\delta t} \]

and hence, for \(t<T\),

\begin{align*} (\partial _t-\Delta )\big (\kappa -(\lambda _\rho ^2+\eta _\rho )y\big )={}&\kappa ^3-\eta _\rho 'y\\ \le {}&C^2\mathrm {e}^{2\delta t}\kappa -\frac {\eta _\rho '}{\lambda _\rho ^2+\eta _\rho }(\lambda _\rho ^2+\eta _\rho )y\\ ={}&C^2\mathrm {e}^{2\delta t}\big (\kappa -(\lambda _\rho ^2+\eta _\rho )y\big )\,. \end{align*} Since \(\kappa -(\lambda _\rho ^2+\eta _\rho )y=0\) at the left endpoint and \((\kappa -(\lambda _\rho ^2+\eta _\rho )y)_s=\kappa -(\lambda _\rho ^2+\eta _\rho )y\) at the right endpoint, we find that

\[ \kappa -(\lambda _\rho ^2+\eta _\rho )y\le \exp \left (\frac {C^2}{2\delta }\mathrm {e}^{2\delta t}\right )\big (\kappa -(\lambda _\rho ^2+\eta _\rho )y\big )\Big |_{t=\alpha _\rho } \]

on each of the old-but-not-ancient solutions with \(\rho <\rho _0\), and hence, taking \(\rho \to 0\),

\[ \kappa \le (\lambda _0^2+\eta _0)y \]

on the ancient solution. The claim follows since, by the mean value theorem, we may estimate \(\eta _0\le \frac {\lambda _0^2C^4}{4\delta ^2}\mathrm {e}^{2\delta t}\) for \(t<0\).

By the estimate ( 4 ) and Corollary 3.5 , we can find \(T>-\infty \), \(C<\infty \) and \(\delta >0\) such that our ancient solution satisfies

\[ (\log \overline y)_t=\frac {\overline \kappa }{\overline y\cos \overline \theta }\le \frac {1}{\sqrt {1-4C^2\mathrm {e}^{2\delta t}}}\frac {\overline \kappa }{\overline y}\le (1+8C^2\mathrm {e}^{2\delta t})\frac {\overline \kappa }{\overline y}\le \lambda _0^2+C^4\mathrm {e}^{2\delta t} \]

for \(t<T\). Integrating from time \(t<T\) to time \(T\) and rearranging then yields

\[ \overline y\ge B\mathrm {e}^{\lambda _0^2t}\,,\;\; B>0\,. \]

Since the gradient of the solution is bounded by \(\tan \overline \theta \le C\mathrm {e}^{\lambda _0^2t}\) for \(t\le T\), this guarantees that the limit \(A(x):=\mathrm {e}^{-\lambda _0^2t}y(x,t)\) is positive for all \(x>x_0\) where \(x_0<1\).

4. Uniqueness

4.1. Unique asymptotics.

Consider now any convex ancient Dirichlet–Neumann curve shortening flow \(\{\Gamma _t\}_{t\in (-\infty ,\omega )}\) with Dirichlet endpoint \(o\in \overline D\setminus \{0\}\).

Lemma 4.1.
Up to a time-translation, a rotation about the origin, and a reflection about the \(x\)-axis, we may arrange that

  • \(o=(-d,0)\) for some \(d\in (0,1]\),
  • \((0,1)\in \Gamma _0\),
  • \(\Gamma _t\) lies in the upper half disc for all \(t\), and
  • \(\Gamma _t\to \{(x,0):x\in [-d,1]\}\) uniformly in the smooth topology as \(t\to -\infty \).

Proof.

Up to a time translation, we may arrange that \(\omega >0\). Up to rotation and a reflection, we may then arrange that \(o=(-d,0)\), \(d\in (0,1]\), and \((\cos \overline \theta (0),\sin \overline \theta (0))\) lies in the upper half-disc. This ensures that \((\cos \overline \theta (t),\sin \overline \theta (t))\) lies in the upper half-disc for all \(t<0\). Indeed, if \(\overline \theta (t_\ast )=0\) for some \(t_\ast <0\), then convexity and the boundary conditions guarantee that \(\Gamma _{t_\ast }=\{(x,0):x\in [-d,1]\}\); so \(\{\Gamma _t\}_{t\in (-\infty ,\omega )}\) is the stationary unstable critical arc.

Denote by \(\Omega _t\) the region lying above \(\Gamma _t\cup \{(0,x):x\in [-1,-d]\}\) and set \(\Omega :=\cup _{t<\omega }\Omega _t\). The first variation formula for enclosed area yields

\[ \frac {d}{dt}\area (\Omega _t)=-\int _{\Gamma _t}\kappa \,ds=-(\overline \theta (t)-\underline \theta (t)) \]

and hence

\[ \area (\Omega _t)=\area (\Omega _0)+\int _t^0(\overline \theta (\tau )-\underline \theta (\tau ))\,d\tau \,. \]

Since \(\area (\Omega )\) is finite, \(\overline \theta -\underline \theta \) must converge to zero along some sequence of times \(t_j\to -\infty \). Since \(\overline \theta >0\), this ensures that \(\Omega \) is the upper half-disc, and hence \(\Gamma _t\) converges uniformly to the unstable critical arc as \(t\to -\infty \). Parabolic regularity theory then guarantees smooth convergence.

Since the flow is monotone, \(\Gamma _t\) must then lie in the upper half disc for all \(t\). We have thus shown, when \(d<1\), that \(\omega =\infty \) and \(\Gamma _t\) converges smoothly to the minimizing arc as \(t\to \infty \) and, when \(d=1\), that \(\omega <\infty \) and \(\Gamma _t\) converges uniformly to \(o\) as \(t\to \omega \). Up to a further time-translation, we may therefore arrange that the point \((0,1)\) lies in \(\Gamma _0\).

Lemma 4.2.
For every \(t\in (-\infty ,\omega )\), \(\kappa _s>0\).

Proof.

Since \(\kappa _s\ge 0\) at both endpoints, the claim may be obtained by applying the maximum principle exactly as in [6], Lemma 3.3 .

Proposition 4.3.
There exists \(A\in [0,\infty )\) such that

\begin{equation} \label {eq:asymptotics} y(x,t)=A\mathrm {e}^{\lambda _0^2t}(\sinh (\lambda (x+d))+o(1)) \end{equation}

uniformly as \(t\to -\infty \).

Proof.

Denote by \(\{\Gamma ^\ast _t\}_{t\in (-\infty ,\omega )}\) the constructed solution. Since \(\Gamma _0\) and \(\Gamma _0^\ast \) both contain the point \((0,1)\), the contrapositive of the avoidance principle guarantees that \(\Gamma _t\) must intersect \(\Gamma ^\ast _t\) away from \(o\) at every time \(t<0\). It follows that the value of \(\underline \theta \) on the second solution must at no time exceed the value of \(\overline \theta {}^\ast \) on the constructed solution. But then, applying the gradient bound ( 3 ) and estimating \(\sin \overline \theta {}^\ast \le A\mathrm {e}^{\lambda _0^2t}+o(\mathrm {e}^{\lambda _0^2t})\) yields

\[ \frac {b}{1+a}\overline y\le \frac {b\sin \overline \theta }{1+a\cos \overline \theta }\le \tan \underline \theta \le 2\sin \underline \theta \le 2\sin \overline \theta {}^\ast \le 2(A\mathrm {e}^{\lambda _0^2t}+o(\mathrm {e}^{\lambda _0^2t})) \]

as \(t\to -\infty \), and hence, when \(d<1\),

\begin{equation} \label {eq:unique height asymptotics} \limsup _{t\to -\infty }\mathrm {e}^{\lambda _0^2t}\overline y(t)<\infty . \end{equation}

Since the height function \(y\) satisfies the (intrinsic) Dirichlet–Robin heat equation

\[ \left \{\begin {aligned}(\partial _t-\Delta )y={}&0\\ y=0\;\;\text {at}\;\; o\,, {}&\; y_s=y\;\;\text {at}\;\;(\cos \overline \theta ,\sin \overline \theta )\,,\end {aligned}\right . \]

we may apply Alaoglu's theorem and elementary Fourier analysis as in [6], Proposition 3.4 to obtain ( 7 ).

When \(d=1\), we need to work a little harder to obtain ( 8 ): at any time \(t<0\), either \(\overline y(t)\le \overline y{}^\ast (t)\), as desired, or \(\overline y(t)>\overline y{}^\ast (t)\). In the latter case, the avoidance principle and the Dirichlet condition ensure that \(y{}^\ast (\cdot ,t)-y(\cdot ,t)\) attains a positive maximum at an interior point. Since the Dirichlet–Neumann circular arc \(C_{\overline \theta (t)}\) lies below \(\Gamma _t\) (with common boundary), we can find some \(t_0>t\) and \(x_0\in (-1,\cos \overline \theta (t_0))\) such that the advanced arc \(C_{\overline \theta (t_0)}\) touches \(\Gamma _t^\ast \) from above at the interior point \((x_0,y^\ast (x_0,t))\), and hence

\begin{equation*} y_{\overline \theta (t_0)}(x_0)=y^\ast (x_0,t)=:A_0\;\;\text {and}\;\; (y_{\overline \theta (t_0)})_x(x_0)= y^\ast _x(x_0,t)=:B_0\,, \end{equation*}

where

\[ y_{\theta }(x)=r(\theta )-\sqrt {r^2(\theta )-(x+1)^2}\,,\;\; r(\theta ):=\frac {1+\cos \theta }{\sin \theta }\,. \]

That is,

\[ r_0-\sqrt {r^2_0-(x_0+1)^2}=A_0\;\;\text {and}\;\;\frac {x_0+1}{\sqrt {r^2_0-(x_0+1)^2}}=B_0\,, \]

where \(r_0:= r(\overline \theta (t_0))\). Rearranging, these become

\[ x_0+1=\frac {B_0r_0}{\sqrt {1+B_0^2}} \;\;\text {and}\;\; r_0=A_0+\frac {x_0+1}{B_0}=A_0+\frac {r_0}{\sqrt {1+B_0^2}}\,, \]

which together imply that

\[ \left (\sqrt {1+B_0^2}-1\right )r_0=A_0\sqrt {1+B_0^2}=\frac {A_0}{x_0+1}B_0r_0\,. \]

Eliminating \(r_0\) and rearranging, we conclude that

\begin{equation} \label {eq:intersection points for unique asymptotics when d=1} \frac {A_0}{(x_0+1)B_0}=\frac {1}{1+\sqrt {1+B_0^2}}\,. \end{equation}

We claim that this is only possible (when \(-t\) is sufficiently large) if \(x_0\) is close to one. Indeed, the asymptotic linear analysis yields, for some \(A\in (0,\infty )\),

\[ \left \{\begin {aligned} A_0={}&y^\ast (x_0,t)=A\mathrm {e}^{\lambda _0^2t}\left (\sinh (\lambda _0(x_0+1))+o(1)\right )\\ B_0={}&y^\ast _x(x_0,t)=A\lambda _0\mathrm {e}^{\lambda _0^2t}\left (\cosh (\lambda _0(x_0+1))+o(1)\right ) \end {aligned}\right .\;\;\text {as}\;\; t\to -\infty \,. \]

(Note that, recalling ( 6 ), we may estimate \(y^\ast _{xx}\lesssim \overline \kappa {}^\ast \le 2\sin \overline \theta {}^\ast \le C\mathrm {e}^{\lambda _0^2t}\), which justifies the uniform \(C^1\) convergence of \(\mathrm {e}^{-\lambda _0^2t}y^\ast (\cdot ,t)\).) Recalling ( 9 ), we conclude that

\[ \frac {\tanh (\lambda _0(x_0+1))}{\lambda _0(x_0+1)}\to \frac {1}{2}\;\;\text {as}\;\; t\to -\infty \,. \]

This implies that \(x_0=1-o(1)\) as \(t\to -\infty \) and hence, as \(t\to -\infty \),

\[ \sin \overline \theta (t)\le \sin \overline \theta (t_0)=(1+\overline x(t_0))r_0^{-1}\sim (1+x_0)r_0^{-1}=\frac {B_0}{\sqrt {1+B_0^2}}\le B_0\sim \mathrm {e}^{\lambda _0^2t} \]

as desired.

4.2. Uniqueness.

Uniqueness may now be established using the avoidance principle, as in [6], Proposition 3.5 . Combined with Theorem 3.2 and the asymptotics ( 7 ), this completes the proof of Theorem 1.1 .

References

  • [1]  Paul T. Allen, Adam Layne, and Katharine Tsukahara. The Dirichlet problem for curve shortening flow. Preprint, arXiv:1208.3510v2.

  • [2]  Ben Andrews. Noncollapsing in mean-convex mean curvature flow. Geom. Topol., 16(3):1413–1418, 2012.

  • [3]  Ben Andrews and Paul Bryan. A comparison theorem for the isoperimetric profile under curve-shortening flow. Comm. Anal. Geom., 19(3):503–539, 2011.

  • [4]  Ben Andrews and Paul Bryan. Curvature bound for curve shortening flow via distance comparison and a direct proof of Grayson's theorem. J. Reine Angew. Math., 653:179–187, 2011.

  • [5]  Theodora Bourni, Nathan Burns, and Spencer Catron. Classification of convex ancient solutions to free boundary curve shortening flow in convex domains. Preprint, arXiv:2404.08824.

  • [6]  Theodora Bourni and Mat Langford. Classification of convex ancient free-boundary curve-shortening flows in the disc. Anal. PDE, 16(9):2225–2240, 2023.

  • [7]  Theodora Bourni, Mat Langford, and Giuseppe Tinaglia. Convex ancient solutions to curve shortening flow. Calc. Var. Partial Differential Equations, 59(4):133, 2020.

  • [8]  John A. Buckland. Mean curvature flow with free boundary on smooth hypersurfaces. J. Reine Angew. Math., 586:71–90, 2005.

  • [9]  Panagiota Daskalopoulos, Richard Hamilton, and Natasa Sesum. Classification of compact ancient solutions to the curve shortening flow. J. Differential Geom., 84(3):455–464, 2010.

  • [10]  Nick Edelen. The free-boundary Brakke flow. J. Reine Angew. Math., 758:95–137, 2020.

  • [11]  M. Gage and R.S. Hamilton. The heat equation shrinking convex plane curves. J. Differ. Geom., 23:69–96, 1986.

  • [12]  M. E. Gage. Curve shortening makes convex curves circular. Invent. Math., 76(2):357–364, 1984.

  • [13]  Matthew A. Grayson. The heat equation shrinks embedded plane curves to round points. J. Differential Geom., 26(2):285–314, 1987.

  • [14]  Gerhard Huisken. Nonparametric mean curvature evolution with boundary conditions. J. Differential Equations, 77(2):369–378, 1989.

  • [15]  Gerhard Huisken. A distance comparison principle for evolving curves. Asian J. Math., 2(1):127–133, 1998.

  • [16]  Dongyeong Ko. Existence and Morse index of two free boundary embedded geodesics on Riemannian 2-disks with convex boundary. Preprint, arXiv:2309.09896.

  • [17]  Mat Langford and Jonathan J. Zhu. A distance comparison principle for curve shortening flow with free boundary. Preprint, arXiv:2302.14258.

  • [18]  Gary M. Lieberman. Second order parabolic differential equations. World Scientific Publishing Co., Inc., River Edge, NJ, 1996.

  • [19]  Axel Stahl. Convergence of solutions to the mean curvature flow with a Neumann boundary condition. Calculus of Variations and Partial Differential Equations, 4(5):421–441, 1996.

  • [20]  Axel Stahl. Regularity estimates for solutions to the mean curvature flow with a Neumann boundary condition. Calculus of Variations and Partial Differential Equations, 4(4):385–407, 1996.