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

Self-similar curve shortening flow in hyperbolic 2-space

Eric Woolgar Dept of Mathematical and Statistical Sciences, and Theoretical Physics Institute, University of Alberta, Edmonton, AB, Canada T6G 2G1. ewoolgar(at)ualberta.ca  and  Ran Xie School of Mathematics and Statistics, Xi’an Jiaotong University, Xi’an, PR China dhsieh29(at)gmail.com
Abstract.

We find and classify self-similar solutions of the curve shortening flow in standard hyperbolic 2-space. Together with earlier work of Halldórsson on curve shortening flow in the plane and Santos dos Reis and Tenenblat in the 2-sphere, this completes the classification of self-similar curve shortening flows in the constant curvature model spaces in 2-dimensions.

1. Introduction

Consider a smooth map X:I×[0,T)M:(x,t)X(x,t)=:Xt(x)X:I\times[0,T)\to M:(x,t)\mapsto X(x,t)=:X_{t}(x), for II a connected interval of the real line and T(0,]T\in(0,\infty]. We will take M3M\subset{\mathbb{R}}^{3} to be a surface, though the flow can also be contemplated for MM a manifold with arbitrary dimension. Then a curve shortening flow is a solution of the differential equation

(1.1) Xt=κgN,\frac{\partial X}{\partial t}=\kappa_{g}N\ ,

where κg\kappa_{g} is the geodesic curvature of the curve Xt(x)X_{t}(x) (for fixed tt) and NN is the principal normal to the curve, with the signs chosen so that κgN\kappa_{g}N points to the concave side of the curve.

The theory of the curve shortening flow, often denoted simply as CSF, was developed in a number of papers in the 1980s, among them Gage [2, 3], Gage and Hamilton [5], and Grayson [6]. There is now an extensive literature, including at least one book [1].

Geodesics are trivial solutions of this flow since they have κg=0\kappa_{g}=0. The next simplest solutions are curves that evolve self-similarly, sometimes called soliton solutions. This class includes geodesics but also includes nontrivial examples. The well-known grim reaper or paperclip, whose trace at fixed tt is the graph y(x)=C+logsecxy(x)=C+\log\sec x, is an example of a non-geodesic self-similar solution of CSF in 2{\mathbb{R}}^{2}. In a thesis in 2013, Halldórsson [7, 8] classified the self-similar solutions in 2{\mathbb{R}}^{2}. Self-similar solutions on the 2-sphere 𝕊2{\mathbb{S}}^{2} were classified by Santos dos Reis and Tenenblat [12] in 2019. In the present paper, we study self-similar curve shortening flows in the remaining complete constant curvature surface, standard hyperbolic 2-space 2{\mathbb{H}}^{2}.

Definition 1.1.

Two unit speed curves γ,γ~:I(M,g)\gamma,{\tilde{\gamma}}:I\to(M,g) are congruent if there is an isometry φ\varphi of (M,g)(M,g) such that φγ=γ~\varphi\circ\gamma={\tilde{\gamma}}. A curve shortening flow evolves by isometries if γt(s)\gamma_{t}(s) is a unit speed curve for each tJt\in J, 0J0\in J\subset{\mathbb{R}}, and there is a one-parameter family of isometries φt\varphi_{t} such that φtγ=γt\varphi_{t}\circ\gamma=\gamma_{t}; i.e., each γt\gamma_{t} is congruent to every other one. A curve that evolves by isometries is also said to be self-similar or a soliton.

We will further take solitons to be inextendible, meaning that the soliton curve admits a unit speed parametrization X(s)X(s) on an open connected interval II\in{\mathbb{R}} such that X(s)X(s) cannot be defined as a unit speed curve on any open connected interval JIJ\supset I containing the closure of II as a proper subset.

Note that self-similarity of curves only makes sense on surfaces that have families of sufficiently smooth isometries. By the fundamental theorem of curves in surfaces, when MM is a surface we can determine whether the curves γ\gamma and γ~{\tilde{\gamma}} are congruent by comparing their curvatures as functions of arclength. Using this fact, a consequence of [4, Proposition 4.2] is that closed curves on surfaces cannot evolve by isometries under CSF. However they can evolve by the composition of isometries and time-dependent rescalings; often the terms“self-similar” and “scaling soliton” are employed to describe curves that evolve in this more general sense, but this paper is concerned with solitons that do not rescale in time.

This brings us to our main theorem.

Theorem 1.2.

Let 𝕄3{\mathbb{M}}^{3} be 3{\mathbb{R}}^{3} equipped with the quadratic form (Minkowski metric) η=diag(1,1,1)\eta=\operatorname{diag}(1,1,-1) and denote the Minkowski inner product by U,Wη=η(U,W)\langle U,W\rangle_{\eta}=\eta(U,W) for U,W𝕄3U,W\in{\mathbb{M}}^{3}. For each v~𝕄3{0}{\tilde{v}}\in{\mathbb{M}}^{3}\setminus\{0\} there is a 22-parameter family of nontrivial solutions XX of CSF evolving by isometries in standard hyperbolic 22-space 2{\mathbb{H}}^{2}. These soliton curves are complete, unbounded, and properly embedded, and asymptote either to a horocycle or to a geodesic. The soliton cannot asymptote to a geodesic

  • (i)

    at both ends, or

  • (ii)

    if v~{\tilde{v}} is timelike, or

  • (iii)

    if μ(s)=X(s),v~η\mu(s)=\left\langle X(s),{\tilde{v}}\right\rangle_{\eta} has a critical point (and it can have at most one critical point), or

  • (iv)

    if μ(s)\mu(s) has a zero (and it can have at most one zero).

While the results of this paper can be viewed as a natural outgrowth of the work reported in [7, 8] and [12], there are important independent reasons to study the curve shortening flow and its solitons in hyperbolic space. There are two generalization of this problem to higher dimensions which are important for physics. The first generalization is the mean curvature flow of codimension one objects (hypersurfaces) flowing in standard hyperbolic nn-space. The fixed points are minimal surfaces bounded by a curve on the boundary at conformal infinity. These surfaces play an important role in the AdS/CFT correspondence. The surface areas of the minimal surfaces are proportional to the entanglement entropy of the region RR enclosed by the curve on the conformal boundary. This entropy is a property of quantum states of a conformal field theory defined on the conformal boundary, and encodes uncertainty in measurements of those states within RR due to correlations which extend beyond RR. One way to construct these minimal surfaces is as limits of convergent mean curvature flows. But these mean curvature flows could instead approach “generalized fixed points”, the self-similar solutions, as limits. The problem of self-similar curve shortening flows in 2{\mathbb{H}}^{2} is the obvious first step to complete before addressing the higher dimensional mean curvature flow version that arises in AdS/CFT physics. Our result that self-similar curves asymptote to geodesics only in limited cases, and never at both ends, may suggest that solitons of the higher dimensional problem would not meet the boundary at infinity orthogonally, and perhaps not even transversally.

The second application arises by considering dimension one flows in a spacetime of dimension n+1n+1 with metric g=dt2+a2(t)hg=-dt^{2}+a^{2}(t)h, for hh a metric on a Riemannian nn-manifold. This situation is of interest in cosmology, particularly when hh is any constant curvature metric. So-called cosmic strings in this spacetime can be modelled as solutions of the wavemap equation for an embedded timelike 2-surface X:(u,w)(X0,Xi)X:(u,w)\mapsto(X^{0},X^{i}), i=1,,ni=1,\dots,n, in spacetime. In this setting this becomes

(1.2) ηXi+2H(t)ηabaXibX0=0,\Box_{\eta}X^{i}+2H(t)\eta^{ab}\partial_{a}X^{i}\partial_{b}X^{0}=0\ ,

where H(t):=a(t)a(t)H(t):=\frac{a^{\prime}(t)}{a(t)}, η\eta is the induced Lorentzian metric on the 2-surface, and η\Box_{\eta} is the trace of the Hessian of η\eta (regarded as operating on an nn-tuple of functions XiX^{i}). If one takes the ww parameter to be the tt-coordinate and the uu parameter to be an arclength ss along the curves Xt(s)=X(s,t)X_{t}(s)=X(s,t), then X0=tX^{0}=t and one obtains [9, Equation 15 with Δ=2s2\Delta=\frac{\partial^{2}}{\partial s^{2}} and assuming the second time derivative term is small enough to ignore]

(1.3) 2Xis2=2H(t)Xit.\frac{\partial^{2}X^{i}}{\partial s^{2}}=2H(t)\frac{\partial X^{i}}{\partial t}\ .

For inflationary cosmology models, H(t)H(t) is constant, and then (1.3) is a curve shortening flow. Therefore, for this problem we increase the dimension of the ambient manifold, which should now be any constant curvature space, and study curve shortening flows in it.

The manuscript is organized as follows. In Section 2, we first review the model of 2{\mathbb{H}}^{2} which represents it as the z1z\geq 1 sheet of the hyperboloid x2+y2z2=1x^{2}+y^{2}-z^{2}=-1 in the Minkowski spacetime 𝕄3{\mathbb{M}}^{3}. This model permits us straightforwardly to adapt ideas of [8, 12] to our setting, which we do in Sections 2.2–2.4. Specifically, we formulate the problem as an autonomous system of ordinary differential equations. In Section 3, we discuss certain special solutions such as geodesics and horocycles. Horocycles were discussed by Grayson [6] and arguably are hyperbolic space analogues of the grim reaper self-similar flow in 2{\mathbb{R}}^{2}. For completeness, we briefly discuss hypercycles, which are scaling solitons (so they evolve by a composition of isometries and rescalings). In Section 4 we analyze the autonomous system from Section 2 and prove a series of lemmata leading to the proof of Theorem 1.2.

After this paper appeared in preprint form, we learned of the beautiful thesis [10] which independently derived similar results in great detail. This work is now described in [11].

1.1. Acknowledgements

We are indebted to K Tenenblat for comments on the preprint version of this paper, which helped us to improve our presentation (especially, to correct an error in Proposition 2.3), and for bringing reference [10] to our attention. EW is grateful to the organizers and audience of the Clark University Geometric Analysis seminar of 13 Nov 2020, where these results were presented, for their interest and insightful comments. We thank to Michael Yi Li and Yingfei Yi for organizing the 2019 International Undergraduate Summer Enrichment Programme (IUSEP) at the University of Alberta, during which this work was begun. EW is supported by NSERC Discovery Grant RGPIN–2017–04896.

2. The hyperboloidal model of hyperbolic space

2.1. Elementary properties

We review some basic facts of the hyperboloidal model, which represents hyperbolic 2-space 2{\mathbb{H}}^{2} as the z1z\geq 1 sheet of the hyperboloid x2+y2z2=1x^{2}+y^{2}-z^{2}=-1 in Minkowski 33-space 𝕄3:=(3,η){\mathbb{M}}^{3}:=\left({\mathbb{R}}^{3},\eta\right) where η\eta is the quadratic form diag(1,1,1)\operatorname{diag}(1,1,-1). We will adopt the convention that our coordinates are enumerated as xi=(x1,x2,x3)=(x,y,z)x^{i}=(x^{1},x^{2},x^{3})=(x,y,z) with η(x,x)=η(y,y)=+1\eta(\partial_{x},\partial_{x})=\eta(\partial_{y},\partial_{y})=+1 so that x\partial_{x} and y\partial_{y} are spacelike and η(z,z)=1\eta(\partial_{z},\partial_{z})=-1 so that z\partial_{z} is timelike in the terminology common in physics.

The hyperboloid modelling 2{\mathbb{H}}^{2} is a spacelike hypersurface in 𝕄3{\mathbb{M}}^{3}; i.e., any vector tangent to this surface is spacelike. However, any vector XX from the origin to a point on 2{\mathbb{H}}^{2} is timelike and future-directed (η(X,z)<0\eta(X,\partial_{z})<0), and has η(X,X)=1\eta(X,X)=-1. In fact, any such vector is a future-directed timelike unit normal field for the surface.

The group GG that preserves the quadratic form η\eta is the orthogonal group O(2,1)\operatorname{O}(2,1). The proper orthochronous subgroup GG that preserves spatial orientation and time orientation will preserve the hyperboloid sheet z=1+x2+y2z=\sqrt{1+x^{2}+y^{2}} and the orientation of bases for its tangent spaces (as well as preserving the choice of future and past). This subgroup lies in the connected component of the identity in the isometry group of 2{\mathbb{H}}^{2}. The one-parameter subgroups of GG can be classified as compositions of boosts in the xx-direction

(2.1) A1(ζ)=[coshζ0sinhζ010sinhζ0coshζ],A_{1}(\zeta)=\left[\begin{array}[]{ccc}\cosh\zeta&0&-\sinh\zeta\\ 0&1&0\\ -\sinh\zeta&0&\cosh\zeta\end{array}\right]\ ,

boosts in the yy-direction

(2.2) A2(ξ)=[1000coshξsinhξ0sinhξcoshξ],A_{2}(\xi)=\left[\begin{array}[]{ccc}1&0&0\\ 0&\cosh\xi&-\sinh\xi\\ 0&-\sinh\xi&\cosh\xi\end{array}\right]\ ,

and rotations about the zz-axis

(2.3) A3(θ)=[cosθsinθ0sinθcosθ0001].A_{3}(\theta)=\left[\begin{array}[]{ccc}\cos\theta&-\sin\theta&0\\ \sin\theta&\cos\theta&0\\ 0&0&1\end{array}\right]\ .

The corresponding Lie algebra is spanned by

(2.4) A1(0)=[001000100],A2(0)=[000001010],A3(0)=[010100000].A_{1}^{\prime}(0)=\left[\begin{array}[]{ccc}0&0&-1\\ 0&0&0\\ -1&0&0\end{array}\right]\ ,\ A_{2}^{\prime}(0)=\left[\begin{array}[]{ccc}0&0&0\\ 0&0&-1\\ 0&-1&0\end{array}\right]\ ,\ A_{3}^{\prime}(0)=\left[\begin{array}[]{ccc}0&-1&0\\ 1&0&0\\ 0&0&0\end{array}\right]\ .

The boosts in the xx-direction preserve y\partial_{y} and each of the two null planes in which it lies. Likewise, the boosts in the yy-direction preserve x\partial_{x} and the two null planes in which it lies. Using this, one can see that a general GG-transformation mapping any orthonormal basis of 𝕄3{\mathbb{M}}^{3} to any other (preserving orientation and time orientation) can be constructed from a product A1(ζ)A2(ξ)A3(θ)A_{1}(\zeta)A_{2}(\xi)A_{3}(\theta), whose parameters ζ\zeta, ξ\xi, and θ\theta play the role of “Euler angles”. We will sketch the argument. Consider two η\eta-orthonormal basis (ONB) sets {ei}\{e_{i}\} and {e~i}\{{\tilde{e}}_{i}\}, where {ei}\{e_{i}\} is the coordinate basis defined by the xix^{i} coordinates, with e3=ze_{3}=\partial_{z} future-timelike and {e1,e2}\{e_{1},e_{2}\} right-handed (from here on, we simply say an oriented basis). Likewise, e~3{\tilde{e}}_{3} will be future-timelike as well. The span of {e~1,e~3}\{{\tilde{e}}_{1},{\tilde{e}}_{3}\} is a timelike plane Π\Pi. It’s a simple matter to find a (normalized spacelike) vector, call it e1e_{1}^{\prime}, that lies in Π\Pi and is orthogonal to e3e_{3}. It’s also a simple exercise in linear algebra to find a rotation A3(θ)A_{3}(\theta) about e3e_{3} such that A3(θ)e1=e1A_{3}(\theta)e_{1}=e_{1}^{\prime}. This obviously leaves e3e_{3} invariant, but maps e2e_{2} to e2=A3(θ)e2e_{2}^{\prime}=A_{3}(\theta)e_{2}. Now apply a boost in the plane spanned by e2e_{2}^{\prime} and e3e_{3}. Such a boost A2(ξ)A_{2}(\xi) will leave e1e_{1}^{\prime} invariant, but we can choose ξ\xi such that e3:=A2(ξ)e3e_{3}^{\prime}:=A_{2}(\xi)e_{3}, which remains timelike under a boost, lies in Π\Pi. This boost, incidentally, acts on e2e_{2}^{\prime} to produce e2′′:=A2(ξ)e2e_{2}^{\prime\prime}:=A_{2}(\xi)e_{2}^{\prime}. The plane Π\Pi is now a coordinate plane for the ONB {e1,e2′′,e3}\{e_{1}^{\prime},e_{2}^{\prime\prime},e_{3}^{\prime}\}, with Π=Span{e1,e3}=Span{e~1,e~3}\Pi=\operatorname{Span}\{e_{1}^{\prime},e_{3}^{\prime}\}=\operatorname{Span}\{{\tilde{e}}_{1},{\tilde{e}}_{3}\}, and e2′′e_{2}^{\prime\prime} is normal (i.e., η\eta-normal) to this plane. A final boost A1(ζ)A_{1}(\zeta) in the plane Π\Pi preserves e2′′e_{2}^{\prime\prime} and can be chosen such that e~1=e1′′:=A1(ζ)e1{\tilde{e}}_{1}=e_{1}^{\prime\prime}:=A_{1}(\zeta)e_{1}^{\prime}. Since e3′′:=A1(ζ)e3e_{3}^{\prime\prime}:=A_{1}(\zeta)e_{3}^{\prime} must be orthogonal to e1′′e_{1}^{\prime\prime}, it follows that {e1′′,e2′′,e3′′}={e~1,e~2,e~3}\{e_{1}^{\prime\prime},e_{2}^{\prime\prime},e_{3}^{\prime\prime}\}=\{{\tilde{e}}_{1},{\tilde{e}}_{2},{\tilde{e}}_{3}\}.

Then a curve of isometries may be written as

(2.5) A(t)=A1(ζ(t))A2(ξ(t))A3(θ(t)).A(t)=A_{1}(\zeta(t))A_{2}(\xi(t))A_{3}(\theta(t))\ .

If this curve passes through the identity isometry at t=0t=0, we may write its tangent there as

(2.6) A˙(0)=A˙1(0)ζ˙(0)+A˙2(0)ξ˙(0)+A˙3(0)θ˙(0)=[001000100]ζ˙(0)+[000001010]ξ˙(0)+[010100000]θ˙(0)=[0θ˙(0)ξ˙(0)θ˙(0)0ζ˙(0)ξ˙(0)ζ˙(0)0].\begin{split}{\dot{A}}(0)=&\,{\dot{A}}_{1}(0){\dot{\zeta}}(0)+{\dot{A}}_{2}(0){\dot{\xi}}(0)+{\dot{A}}_{3}(0){\dot{\theta}}(0)\\ =&\,\left[\begin{array}[]{ccc}0&0&-1\\ 0&0&0\\ -1&0&0\end{array}\right]{\dot{\zeta}}(0)+\left[\begin{array}[]{ccc}0&0&0\\ 0&0&-1\\ 0&-1&0\end{array}\right]{\dot{\xi}}(0)+\left[\begin{array}[]{ccc}0&-1&0\\ 1&0&0\\ 0&0&0\end{array}\right]{\dot{\theta}}(0)\\ =&\,\left[\begin{array}[]{ccc}0&-{\dot{\theta}}(0)&-{\dot{\xi}}(0)\\ {\dot{\theta}}(0)&0&-{\dot{\zeta}}(0)\\ -{\dot{\xi}}(0)&-{\dot{\zeta}}(0)&0\end{array}\right]\ .\end{split}

For any curve of isometries containing the identity at t=0t=0, differentiation at t=t0t=t_{0} can be accomplished using that A(t0+ϵ)=A(t0+ϵ)A1(t0)A(t0)=:B(ϵ)A(t0)A(t_{0}+\epsilon)=A(t_{0}+\epsilon)A^{-1}(t_{0})A(t_{0})=:B(\epsilon)A(t_{0}) where B(ϵ):=A(t0+ϵ)A1(t0)B(\epsilon):=A(t_{0}+\epsilon)A^{-1}(t_{0}) so B(0)=idB(0)=\operatorname{id}. But B(ϵ)B(\epsilon) can be written as a product B(ϵ)=A1(ζ(ϵ))A2(ξ(ϵ))A3(θ(ϵ))B(\epsilon)=A_{1}(\zeta(\epsilon))A_{2}(\xi(\epsilon))A_{3}(\theta(\epsilon)), and so

(2.7) A˙(t0)=ddϵ|ϵ=0B(ϵ)A(t0)=B˙(0)A(t0),{\dot{A}}(t_{0})=\frac{d}{d\epsilon}\bigg{|}_{\epsilon=0}B(\epsilon)A(t_{0})={\dot{B}}(0)A(t_{0})\ ,

with B˙{\dot{B}} given by the right-hand side of (2.6).

2.2. Curves in the hyperboloid model

Given a hypersurface SS in a Riemannian manifold and vectors UU and VV tangent to it, a standard formula in Riemannian geometry gives

(2.8) UV=DUV+K(U,V),\nabla_{U}V=D_{U}V+K(U,V)\ ,

where \nabla is the connection on the ambient manifold, DD is the connection induced on the hypersurface, and KK is the vector-valued second fundamental form of the hypersurface. In the case of present interest, \nabla is the Levi-Civita connection of the Minkowski metric η=,η\eta=\langle\cdot,\cdot\rangle_{\eta}. Since the vector XX from the origin to the unit hyperboloid is a unit future-timelike vector orthogonal to the hyperboloid, we may write a general vector VV as

(2.9) V=P(V)X,VηX,V=P(V)-\langle X,V\rangle_{\eta}X\ ,

where PP is orthogonal projection to the tangent space of the hyperboloid. The negative sign in the second term arises because XX is timelike. The second fundamental form can then be written as

(2.10) K=P(X)X,K=P(\nabla X_{\flat})X\ ,

where X:=,XηX_{\flat}:=\langle\cdot,X\rangle_{\eta} is the 1-form metric-dual to XX. If we write the first fundamental form of the hyperboloid as hh then the hyperboloid is umbilic in 𝕄3{\mathbb{M}}^{3} such that K=XhK=Xh (so that X,K(V,V)η=1\langle X,K(V,V)\rangle_{\eta}=-1 for any unit vector VV tangent to the hyperboloid).

Now consider a (smooth) unit-speed curve X(s)X(s) on the unit hyperboloid. The unit tangent vector is T(s):=dXdsT(s):=\frac{dX}{ds}. It follows from the above formulas that

(2.11) TT=DTT+K(T,T)=DTT+X.\nabla_{T}T=D_{T}T+K(T,T)=D_{T}T+X\ .

Since TT is a unit vector, TT\nabla_{T}T lies in its orthogonal complement, from which we see that DTTD_{T}T does as well. But DTTD_{T}T must be tangent to the hyperboloid and so orthogonal to XX, so we write

(2.12) DTT=κgN,D_{T}T=\kappa_{g}N\ ,

where NN is called the principal normal vector to the curve sX(s)s\mapsto X(s) and κg\kappa_{g} is the geodesic curvature. The sign choices are such that {T,N,X}\{T,N,X\} is an orthonormal oriented basis oriented such that XX is future-pointing and the vector κgN\kappa_{g}N points to the concave side of the curve, viewed as a curve in the hyperboloid, whenever κg\kappa_{g} is not zero. Now denote η(TT,TT)=:ϵκ2\eta(\nabla_{T}T,\nabla_{T}T)=:\epsilon\kappa^{2} where ϵ=1\epsilon=1 if TT\nabla_{T}T is spacelike, 0 if it’s null, and 1-1 if it’s timelike. Then from (2.11) we can relate the curvature κ\kappa of XX as a space curve in (𝕄3,η)({\mathbb{M}}^{3},\eta) to its geodesic curvature κg\kappa_{g} in the hyperboloid by

(2.13) κg2=ϵκ2+1.\kappa_{g}^{2}=\epsilon\kappa^{2}+1\ .

Note that T,TNη=N,TTη=N,DTTη=κg\langle T,\nabla_{T}N\rangle_{\eta}=-\langle N,\nabla_{T}T\rangle_{\eta}=-\langle N,D_{T}T\rangle_{\eta}=-\kappa_{g}. By similar reasoning, we see that X,TNη=0\langle X,\nabla_{T}N\rangle_{\eta}=0, and of course since NN is a unit vector then N,TNη=0\langle N,\nabla_{T}N\rangle_{\eta}=0. Thus TN=κgT\nabla_{T}N=-\kappa_{g}T. Collecting our results, we have that the {T,N,X}\{T,N,X\} basis evolves along the curve according to the Frenet-Serret equations in 2{\mathbb{H}}^{2}, which are

(2.14) TT=κgN+X,TN=κgT,TX=T.\begin{split}\nabla_{T}T=&\,\kappa_{g}N+X\ ,\\ \nabla_{T}N=&\,-\kappa_{g}T\ ,\\ \nabla_{T}X=&\,T\ .\end{split}

2.3. Curves and axes

Here we follow closely the work of [12] for the 𝕊2{\mathbb{S}}^{2} case, making changes as necessary. Choose an arbitrary vector v~𝕄3\{0}{\tilde{v}}\in{\mathbb{M}}^{3}\backslash\{0\}, which can be either timelike, spacelike, or null. It is convenient to define v~=av{\tilde{v}}=av where a=|η(v~,v~)|a=\sqrt{|\eta({\tilde{v}},{\tilde{v}})|} when vv is not null. For presentation purposes, we will introduce the scale factor even when v~{\tilde{v}} is null, but in that case a0a\neq 0 will not be determined and vv, being null, cannot be normailzed. Then η(v~,v~)=ϵ=0,±1\eta({\tilde{v}},{\tilde{v}})=\epsilon=0,\pm 1 depending on whether vv is null (0), spacelike (+1+1), or timelike (1-1).

Fix a unit speed curve X:I𝕄3X:I\to{\mathbb{M}}^{3} on the unit hyperboloid, and define the functions

(2.15) τ(s):=T(s),vη,ν(s):=N(s),vη,μ(s):=X(s),vη.\begin{split}\tau(s):=&\,\langle T(s),v\rangle_{\eta}\ ,\\ \nu(s):=&\,\langle N(s),v\rangle_{\eta}\ ,\\ \mu(s):=&\,\langle X(s),v\rangle_{\eta}\ .\end{split}

Then we can write

(2.16) v=τ(s)T(s)+ν(s)N(s)μ(s)X(s)v=\tau(s)T(s)+\nu(s)N(s)-\mu(s)X(s)

and

(2.17) ϵ=η(v,v)=τ2(s)+ν2(s)μ2(s).\epsilon=\eta(v,v)=\tau^{2}(s)+\nu^{2}(s)-\mu^{2}(s)\ .

The next result shows that if the geodesic curvature of the curve X(s)X(s) takes the form κg(s)=aτ(s)\kappa_{g}(s)=a\tau(s), a choice corresponding to a flow by isometries under CSF as we will see in the next subsection, then τ\tau, ν\nu, and μ\mu obey a certain autonomous system of differential equations along X(s)X(s). The analogous result for curves in 𝕊2{\mathbb{S}}^{2} is found in [12, Proposition 3.1].

Proposition 2.1.

Along any smooth unit speed curve X(s)X(s) on the hyperboloid we have that κg(s)=aτ(s)\kappa_{g}(s)=a\tau(s) if and only if

(2.18) {τ(s)=aτ(s)ν(s)+μ(s),ν(s)=aτ2(s),μ(s)=τ(s).\begin{cases}\tau^{\prime}(s)=a\tau(s)\nu(s)+\mu(s),\\ \nu^{\prime}(s)=-a\tau^{2}(s),\\ \mu^{\prime}(s)=\tau(s).\end{cases}
Proof.

Equations (2.14) imply that

(2.19) τ(s)=κg(s)ν(s)+μ(s),ν(s)=κg(s)τ(s),μ(s)=τ(s).\begin{split}\tau^{\prime}(s)=&\,\kappa_{g}(s)\nu(s)+\mu(s)\ ,\\ \nu^{\prime}(s)=&\,-\kappa_{g}(s)\tau(s)\ ,\\ \mu^{\prime}(s)=&\,\tau(s)\ .\end{split}

To prove the forward implication, simply substitute κg(s)=aτ(s)\kappa_{g}(s)=a\tau(s) into (2.19).

To prove the reverse implication, note that we can combine equations (2.18) and (2.19) to write that (aτκg)ν=0\left(a\tau-\kappa_{g}\right)\nu=0 and (aτκg)τ=0\left(a\tau-\kappa_{g}\right)\tau=0. But if τ\tau and ν\nu both vanish, then vv must be orthogonal to both TT and NN and hence parallel to XX, so vv is normal to the hyperboloid. But vv is a constant vector (i.e., it is parallel in (𝕄3,η)({\mathbb{M}}^{3},\eta)), so this can only happen at isolated points along a nontrivial curve XX, so it must instead be that aτκg=0a\tau-\kappa_{g}=0. ∎

Next we show that solutions of the system (2.18) are always realized by actual curves.

Proposition 2.2.

Given a solution (τ(s),ν(s),μ(s))(\tau(s),\nu(s),\mu(s)) of (2.18) obeying initial conditions (τ0,ν0,μ0)=(τ(0),ν(0),μ(0))(\tau_{0},\nu_{0},\mu_{0})=(\tau(0),\nu(0),\mu(0)), there is a smooth unit speed curve X:I𝕄3X:I\to{\mathbb{M}}^{3} on the unit hyperboloid satisfying (2.15).

Proof.

Choose a point X0X_{0} on the hyperboloid and two vectors T0T_{0} and N0N_{0} such that {T0,N0,X0}\{T_{0},N_{0},X_{0}\} is an oriented orthonormal frame. Then define vv, normalized as above (if non-null), by v=τ0T0+ν0N0μ0X0v=\tau_{0}T_{0}+\nu_{0}N_{0}-\mu_{0}X_{0}. Now, given τ(s)\tau(s) for sIs\in I and given a>0a>0, define a function k:I:saτ(s)k:I\to{\mathbb{R}}:s\mapsto a\tau(s). By the fundamental theorem for curves in 2{\mathbb{H}}^{2}, there is a unique curve X:I𝕄3X:I\to{\mathbb{M}}^{3} such that X(0)=X0X(0)=X_{0}, X(0)=:T(0)=T0X^{\prime}(0)=:T(0)=T_{0}, and N(0)=N0N(0)=N_{0}, whose curvature is k(s)k(s). This curve must satisfy equations (2.14) with κg(s)=k(s)\kappa_{g}(s)=k(s). Then equations (2.15) hold for any constant vector vv, and so hold for v=τ0T0+ν0N0μ0X0v=\tau_{0}T_{0}+\nu_{0}N_{0}-\mu_{0}X_{0}. ∎

2.4. CSF and flow by isometries

Having defined a moving basis along an arbitrary curve in the hyperboloid, we can now define the curve shortening flow in the hyperboloid to be a flow Xt(s)=X(t,s)X_{t}(s)=X(t,s) of smooth curves Xt(s)=(xt(s),yt(s),zt(s))X_{t}(s)=\left(x_{t}(s),y_{t}(s),z_{t}(s)\right) with principal normal NtN_{t}, for tJt\in J\subset{\mathbb{R}} some connected interval about t=0t=0, such that

(2.20) Xtt,Ntη=κg,Xtt,Xtη= 0,|X0|η2=x02+y02z02=1,z01.\begin{split}\left\langle\frac{\partial X_{t}}{\partial t},N_{t}\right\rangle_{\eta}=&\,\kappa_{g}\ ,\\ \left\langle\frac{\partial X_{t}}{\partial t},X_{t}\right\rangle_{\eta}=&\,0\ ,\\ \left|X_{0}\right|_{\eta}^{2}=x_{0}^{2}+y_{0}^{2}-z_{0}^{2}=&\,-1\ ,\ z_{0}\geq 1\ .\end{split}

The last two equations are equivalent to the conditions Xt,Xtη=1\left\langle X_{t},X_{t}\right\rangle_{\eta}=-1 and X0=(x0,y0,z0)X_{0}=\left(x_{0},y_{0},z_{0}\right).

We are interested in those solutions of (2.20) that are of the form

(2.21) Xt=A(t)X0,tJ,A(0)=id,X_{t}=A(t)X_{0}\ ,\ t\in J\ ,\ A(0)=\operatorname{id}\ ,

where A(t):=A1(ζ(t))A2(ξ(t))A3(θ(t))A(t):=A_{1}(\zeta(t))A_{2}(\xi(t))A_{3}(\theta(t)) with ζ(0)=ξ(0)=θ(0)=0\zeta(0)=\xi(0)=\theta(0)=0, with the AiA_{i} defined in equations (2.1)–(2.3). Then tXtt\mapsto X_{t} is called a flow by isometries. Motion by an isometry preserves the curvature, so κg(s,t)=κ)g(s)\kappa_{g}(s,t)=\kappa)g(s) (i.e., it is independent of tt).

As we saw in Proposition 2.1, if there is a (normalized or null) vector vv such that aT,vη=κga\langle T,v\rangle_{\eta}=\kappa_{g}, then a system of three scalar equations governs the moving frame. The next result shows that if a curve shortening flow is also a flow by isometries, there is always such a vv which is constant with respect to the flow parameter tt. For convenience, we work with the unrescaled vector v~=av{\tilde{v}}=av (a(s)a(s) is used for another purpose in the proof). We follow the proof of the analogous result for curves in 𝕊2{\mathbb{S}}^{2} [12, Theorem 2.2].

Proposition 2.3.

Let X(s,t)=:Xt(s)X(s,t)=:X_{t}(s) be a smooth function of two variables such that Xt(s)X_{t}(s) is a regular unit speed curve for each tJt\in J, lying in the hyperboloid z=1+x2+y2z=\sqrt{1+x^{2}+y^{2}} in 𝕄3{\mathbb{M}}^{3} and evolving by isometries as in (2.21). Then (2.20) holds if and only if there is a v~𝕄3\{0}{\tilde{v}}\in{\mathbb{M}}^{3}\backslash\{0\} such that

(2.22) T(s),v~η=κg(s)\left\langle T(s),{\tilde{v}}\right\rangle_{\eta}=\kappa_{g}(s)

for all sIs\in I, where T(s)=dXt0ds=:Xt0(s)T(s)=\frac{dX_{t_{0}}}{ds}=:X_{t_{0}}^{\prime}(s) and κg(s)\kappa_{g}(s) is the geodesic curvature of Xt0(s)=:X(s)X_{t_{0}}(s)=:X(s).

Proof.

Let Xt(s)=A(t)X0X_{t}(s)=A(t)X_{0}, so that XX flows by isometries. Then at any t=t0Jt=t_{0}\in J, we compute

(2.23) t|t0Xt(s)=[0θ˙ξ˙θ˙0ζ˙ξ˙ζ˙0][xt0(s)yt0(s)zt0(s)]=[θ˙yt0ξ˙zt0θ˙xt0ζ˙zt0ξ˙xt0ζ˙yt0],\frac{\partial}{\partial t}\bigg{|}_{t_{0}}X_{t}(s)=\left[\begin{array}[]{ccc}0&-{\dot{\theta}}&-{\dot{\xi}}\\ {\dot{\theta}}&0&-{\dot{\zeta}}\\ -{\dot{\xi}}&-{\dot{\zeta}}&0\end{array}\right]\left[\begin{array}[]{c}x_{t_{0}}(s)\\ y_{t_{0}}(s)\\ z_{t_{0}}(s)\end{array}\right]=\left[\begin{array}[]{c}-{\dot{\theta}}y_{t_{0}}-{\dot{\xi}}z_{t_{0}}\\ {\dot{\theta}}x_{t_{0}}-{\dot{\zeta}}z_{t_{0}}\\ -{\dot{\xi}}x_{t_{0}}-{\dot{\zeta}}y_{t_{0}}\end{array}\right]\ ,

using (2.7) and (2.6). Let Tt0(s):=Xt(s)|t=t0T_{t_{0}}(s):=X_{t}^{\prime}(s)\big{|}_{t=t_{0}} be the unit tangent vector to Xt0(s)X_{t_{0}}(s) and let Nt0(s)N_{t_{0}}(s) be the unit normal vector to Xt0(s)X_{t_{0}}(s) in the tangent space to the hyperboloid, defined so that {Tt0,Nt0,Xt0}\{T_{t_{0}},N_{t_{0}},-X_{t_{0}}\} is an oriented orthonormal basis. Writing Nt0(s)=(a(s),b(s),c(s))N_{t_{0}}(s)=(a(s),b(s),c(s)), we have

(2.24) Nt0,Xt0η=a(s)xt0(s)+b(s)yt0(s)c(s)zt0(s)=0Nt0,Tt0η=a(s)xt0(s)+b(s)yt0(s)c(s)zt0(s)=0,\begin{split}\left\langle N_{t_{0}},X_{t_{0}}\right\rangle_{\eta}=&\,a(s)x_{t_{0}}(s)+b(s)y_{t_{0}}(s)-c(s)z_{t_{0}}(s)=0\,\\ \left\langle N_{t_{0}},T_{t_{0}}\right\rangle_{\eta}=&\,a(s)x_{t_{0}}^{\prime}(s)+b(s)y_{t_{0}}^{\prime}(s)-c(s)z_{t_{0}}^{\prime}(s)=0\ ,\end{split}

so that at each ss along Xt0X_{t_{0}} we have

(2.25) Nt0=(a(s),b(s),c(s))=(yt0zt0yt0zt0,zt0xt0zt0xt0,xt0yt0+xt0yt0).N_{t_{0}}=(a(s),b(s),c(s))=\left(y_{t_{0}}^{\prime}z_{t_{0}}-y_{t_{0}}z_{t_{0}}^{\prime},z_{t_{0}}^{\prime}x_{t_{0}}-z_{t_{0}}x_{t_{0}}^{\prime},-x_{t_{0}}^{\prime}y_{t_{0}}+x_{t_{0}}y_{t_{0}}^{\prime}\right)\ .

It’s not difficult to check that this vector has norm 11 and has the correct sign. Then

(2.26) Nt0,Xtt|t=t0η=[a(s)b(s)c(s)][100010001][θ˙yξ˙zθ˙xζ˙zξ˙xζ˙y]=ζ˙(xyyxzzxy2+xz2)+ξ˙(xxy+zzy+x2yz2y)+θ˙(xxzyyz+x2z+y2z)\begin{split}\left\langle N_{t_{0}},\frac{\partial X_{t}}{\partial t}\bigg{|}_{t=t_{0}}\right\rangle_{\eta}=&\,\left[\begin{array}[]{ccc}a(s)&b(s)&c(s)\end{array}\right]\left[\begin{array}[]{ccc}1&0&0\\ 0&1&0\\ 0&0&-1\end{array}\right]\left[\begin{array}[]{c}-{\dot{\theta}}y-{\dot{\xi}}z\\ {\dot{\theta}}x-{\dot{\zeta}}z\\ -{\dot{\xi}}x-{\dot{\zeta}}y\end{array}\right]\\ =&\,{\dot{\zeta}}\left(xyy^{\prime}-xzz^{\prime}-x^{\prime}y^{2}+x^{\prime}z^{2}\right)+{\dot{\xi}}\left(-xx^{\prime}y+zz^{\prime}y+x^{2}y^{\prime}-z^{2}y^{\prime}\right)\\ &\,+{\dot{\theta}}\left(-xx^{\prime}z-yy^{\prime}z+x^{2}z^{\prime}+y^{2}z^{\prime}\right)\end{split}

where we’ve removed the t0t_{0} subscripts on the right to lessen the clutter. Using that x2+y2z2=1x^{2}+y^{2}-z^{2}=-1 and, therefore, that xx+yy=zzxx^{\prime}+yy^{\prime}=zz^{\prime}, this last line simplifies and implies that, for any curve moving by isometries described by A(ζ(t),ξ(t),θ(t))A(\zeta(t),\xi(t),\theta(t)), we have

(2.27) Nt0,Xtt|t=t0η=ζ˙xξ˙yθ˙z=Tt0,v~η,\left\langle N_{t_{0}},\frac{\partial X_{t}}{\partial t}\bigg{|}_{t=t_{0}}\right\rangle_{\eta}={\dot{\zeta}}x^{\prime}-{\dot{\xi}}y^{\prime}-{\dot{\theta}}z^{\prime}=\langle T_{t_{0}},{\tilde{v}}\rangle_{\eta}\ ,

where v~:=(ζ˙,ξ˙,θ˙){\tilde{v}}:=\left({\dot{\zeta}},-{\dot{\xi}},{\dot{\theta}}\right). (Note that v~{\tilde{v}} can be either timelike, spacelike, or null.)

To prove necessity (“only if”), by (2.20) we have Xt0t,Nt0η=κg(s)\left\langle\frac{\partial X_{t_{0}}}{\partial t},N_{t_{0}}\right\rangle_{\eta}=\kappa_{g}(s), so

(2.28) Tt0(s),v~ηT(s),v~η=κg(s).\left\langle T_{t_{0}}(s),{\tilde{v}}\right\rangle_{\eta}\equiv\left\langle T(s),{\tilde{v}}\right\rangle_{\eta}=\kappa_{g}(s)\ .

To prove sufficiency, we note that one can by direct computation prove that if Ai(t)A_{i}(t) is any of the matrices (2.1)–(2.3) then in matrix notation (Ai(t))TηAi(t)=(Ai(0))TηA0(0)(A_{i}^{\prime}(t))^{T}\eta A_{i}(t)=(A_{i}^{\prime}(0))^{T}\eta A_{0}(0) where MTM^{T} denotes the transpose of MM (note that Ai(0)=idA_{i}(0)=\operatorname{id}). For a flow of the form X(s,t)=Ai(t)X(s)X(s,t)=A_{i}(t)X(s), we have N(s,t)=Ai(t)N(s)N(s,t)=A_{i}(t)N(s), so

(2.29) κg(s,t)=Xt(s,t),N(s,t)=XT(s)(Ai(t))TηAi(t)N(s)=XT(s)(Ai(0))TηA0(0)N(s)=κg(s,0).\begin{split}\kappa_{g}(s,t)=&\,\left\langle\frac{\partial X}{\partial t}(s,t),N(s,t)\right\rangle=X^{T}(s)\left(A_{i}^{\prime}(t)\right)^{T}\eta A_{i}(t)N(s)=X^{T}(s)(A_{i}^{\prime}(0))^{T}\eta A_{0}(0)N(s)\\ =&\,\kappa_{g}(s,0)\ .\end{split}

Thus, any of the families of isometries listed in (2.1)–(2.3) produce curve shortening flows such that κg(s)\kappa_{g}(s) is independent of tt. So if there is a vector vv and a unit speed curve X(s)X(s) with curvature function κg(s)\kappa_{g}(s) such that X(s),v=κg(s)\langle X^{\prime}(s),v\rangle=\kappa_{g}(s), then a flow by any of these families will be a curve shortening flow with κg(s,t)=κg(s,0)κg(s)\kappa_{g}(s,t)=\kappa_{g}(s,0)\equiv\kappa_{g}(s). ∎

Remark 2.4.

It is always possible to reduce (2.29) to the form Xtt=κgNt\frac{\partial X_{t}}{\partial t}=\kappa_{g}N_{t} by a reparametrization uφ(u,t)u\mapsto\varphi(u,t) of the curve X(u,t)X(u,t) (e.g., [1, Proposition 1.1]).

3. Examples

3.1. Geodesics

These are solutions of (2.18) with τ0\tau\equiv 0, so μ0\mu\equiv 0 as well, and ν\nu is constant. They are fixed points of (3.1). Since μ=0\mu=0, geodesics are plane curves lying in the plane orthogonal to vv.

We may view the system (2.18) as an autonomous third-order system

(3.1) Φ=Fa(Φ),\Phi^{\prime}=F_{a}(\Phi)\ ,

where Φ=(τ,ν,μ)\Phi=(\tau,\nu,\mu), a>0a>0 is a constant, and Fa:33:(τ,ν,μ)(aτν+μ,aτ2,τ)F_{a}:{\mathbb{R}}^{3}\to{\mathbb{R}}^{3}:(\tau,\nu,\mu)\mapsto(a\tau\nu+\mu,-a\tau^{2},\tau). We now show that at fixed points, the differential of FaF_{a} has real eigenvalues of all three signs (positive, negative, and zero), so geodesics are unstable fixed points but attract in one direction.

Lemma 3.1.

For FaF_{a} as above, then Fa=0(τ,ν,μ)=(0,±1,0)=±e2F_{a}=0\Leftrightarrow(\tau,\nu,\mu)=(0,\pm 1,0)=\pm e_{2}, and the corresponding curves are geodesics. The eigenvalues of dFa|e2dF_{a}|_{e_{2}} are

(3.2) λ= 0,a+a2+42>0,aa2+42<0.\lambda=\ 0,\ \frac{a+\sqrt{a^{2}+4}}{2}>0,\ \frac{a-\sqrt{a^{2}+4}}{2}<0\ .

The eigenvalues of dFa|e2dF_{a}|_{-e_{2}} are the negatives of the above eigenvalues.

Specifically, when a=1a=1 and (τ,ν,μ)=(0,1,0)(\tau,\nu,\mu)=(0,1,0) (respectively, (τ,ν,μ)=(0,1,0)(\tau,\nu,\mu)=(0,1,0)), then λ=0,1+52,152\lambda=0,\frac{1+\sqrt{5}}{2},\frac{1-\sqrt{5}}{2} (respectively, λ=0,1+52,152\lambda=0,\frac{-1+\sqrt{5}}{2},\frac{-1-\sqrt{5}}{2}).

Proof.

The fixed points at ±e2\pm e_{2} are obvious. Then since μ=0\mu=0 and τ=0\tau=0, we have from (2.15) that vv is normal to a timelike plane containing XX and TT. The curve X(s)X(s) is then the intersection curve of this timelike plane, which contains the origin, and the hyperboloid. It is therefore a hyperbolic great circle, and thus a geodesic (this can also be seen from (2.22)).

The differential of FaF_{a} is given by

(3.3) dFa=(aνaτ12aτ00100).dF_{a}=\begin{pmatrix}a\nu&a\tau&1\\ -2a\tau&0&0\\ 1&0&0\end{pmatrix}\ .

Hence, at ±e2=(0,±1,0)\pm e_{2}=(0,\pm 1,0) we have

(3.4) dFa|±e2=(±a01000100).dF_{a}|_{\pm e_{2}}=\begin{pmatrix}\pm a&0&1\\ 0&0&0\\ 1&0&0\end{pmatrix}\ .

Therefore, λ\lambda is an eigenvalue of dFa|±e2dF_{a}|_{\pm e_{2}} iff λ2(±aλ)+λ=λ(λ2aλ1)=0\lambda^{2}(\pm a-\lambda)+\lambda=-\lambda\left(\lambda^{2}\mp a\lambda-1\right)=0. ∎

3.2. Horocycles

In the Poincaré disk model, horocycles are Euclidean circles tangent to the boundary of the Poincaré disk at one point and otherwise lying within the disk. If the point of tangency is (1,0)(1,0) then the horocycle with centre (ϖ,0)(\varpi,0) can be written in parametrized form as

(3.5) u(ρ)=ϖ+(1ϖ)cosρ(1ϖ),v(ρ)=(1ϖ)sinρ(1ϖ),u(\rho)=\varpi+(1-\varpi)\cos\frac{\rho}{(1-\varpi)}\quad,\quad v(\rho)=(1-\varpi)\sin\frac{\rho}{(1-\varpi)}\ ,

where ρ\rho is a parameter. By inverting the stereographic projection (u,v)=(x1+z,y1+z)(u,v)=\left(\frac{x}{1+z},\frac{y}{1+z}\right) where z=1+x2+y2z=\sqrt{1+x^{2}+y^{2}}, we obtain a parametric description of this horocycle in the hyperboloidal model

(3.6) x=ϖ2s2+2ϖ12ϖ(1ϖ),y=s,z=ϖ2s2+12ϖ(1ϖ)1.x=\frac{\varpi^{2}s^{2}+2\varpi-1}{2\varpi(1-\varpi)}\quad,\quad y=s\quad,\quad z=\frac{\varpi^{2}s^{2}+1}{2\varpi(1-\varpi)}-1\ .

Here ss is an arclength parameter. It is straightforward to compute the Frenet frame

(3.7) T=(ϖs(1ϖ),1,ϖs(1ϖ))N=(ϖ2s22ϖ2+2ϖ12ϖ(1ϖ),s,ϖ2s22ϖ+12ϖ(1ϖ)),X=(ϖ2s2+2ϖ12ϖ(1ϖ),s,ϖ2s2+12ϖ(1ϖ)1).\begin{split}T=&\,\left(\frac{\varpi s}{(1-\varpi)},1,\frac{\varpi s}{(1-\varpi)}\right)\quad\,\quad N=\left(\frac{\varpi^{2}s^{2}-2\varpi^{2}+2\varpi-1}{2\varpi(1-\varpi)},s,\frac{\varpi^{2}s^{2}-2\varpi+1}{2\varpi(1-\varpi)}\right)\ ,\\ X=&\,\left(\frac{\varpi^{2}s^{2}+2\varpi-1}{2\varpi(1-\varpi)},s,\frac{\varpi^{2}s^{2}+1}{2\varpi(1-\varpi)}-1\right)\ .\end{split}

Choosing v=(0,1,0)v=(0,1,0) we obtain

(3.8) τ=1,ν=μ=s.\tau=1\ ,\ \nu=-\mu=-s\ .

Entering these values into the system (2.18), we see that the system is satisfied if a=1a=1. Then the relation κg=aτ\kappa_{g}=a\tau yields κg=1\kappa_{g}=1, as it must for horocycles.

As discussed by Grayson [6], under the curve shortening flow horocycles remain horocycles, and so evolve self-similarly, with curvature κg=1\kappa_{g}=1 throughout the flow. In the next section, we will seek all solutions of the system (2.18) corresponding to self-similar evolutions with v=(0,1,0)v=(0,1,0). This will be the case of curves which evolve purely by a boost in 𝕄3{\mathbb{M}}^{3}.

3.3. Hypercycles

In the Poincaré disk model, hypercycles are arcs of Euclidean circles that end when they meet the boundary of the Poincaré disk transversally. Under CSF they do not evolve by isometries, but do evolve under the composition of isometries and rescalings.

Applying a rotation if necessary, we can place the centre of the Euclidean circle along one of the axes, say at (ϖ,0)(\varpi,0) where ϖ>0\varpi>0. Let the Euclidean radius of the circle be c<1+ϖc<1+\varpi. We can parametrize the hypercycle as we would any such Euclidean circle, say as

(3.9) u(t)=ϖ+ccostc,v(t)=csintc,u(t)=\varpi+c\cos\frac{t}{c}\quad,\quad v(t)=c\sin\frac{t}{c}\ ,

with the domain of tt chosen so that u2+v2<1u^{2}+v^{2}<1.

In the hyperboloidal model, by using stereographic projection to lft the above parametrized curve to the unit hyperboloid we obtain

(3.10) x(t)=2(ϖ+ccostc)1c2ϖ22cϖcostc,y(t)=2csintc1c2ϖ22cϖcostc,z(t)=1+c2+ϖ2+2cϖcostc1c2ϖ22cϖcostc.\begin{split}x(t)=&\,\frac{2\left(\varpi+c\cos\frac{t}{c}\right)}{1-c^{2}-\varpi^{2}-2c\varpi\cos\frac{t}{c}}\quad,\quad y(t)=\frac{2c\sin\frac{t}{c}}{1-c^{2}-\varpi^{2}-2c\varpi\cos\frac{t}{c}}\ ,\\ z(t)=&\,\frac{1+c^{2}+\varpi^{2}+2c\varpi\cos\frac{t}{c}}{1-c^{2}-\varpi^{2}-2c\varpi\cos\frac{t}{c}}\ .\end{split}

These curves are curves of intersection of the hyperboloid z=1+x2+y2z=\sqrt{1+x^{2}+y^{2}} in 𝕄3{\mathbb{M}}^{3} with the planes

(3.11) 2ϖ(x+1ϖ)+(c2ϖ21)(z+1)=0,2\varpi\left(x+\frac{1}{\varpi}\right)+\left(c^{2}-\varpi^{2}-1\right)(z+1)=0\ ,

for c+ϖ>1c+\varpi>1. These planes are timelike.

In his seminal paper, Grayson computes the evolution of hypercycles in 2{\mathbb{H}}^{2} under CSF [6, p 76]. The curve remains a hypercycle, but the curvature evolves as

(3.12) κg(s,t)=κg(t)=11Ae2t.\kappa_{g}(s,t)=\kappa_{g}(t)=\frac{1}{\sqrt{1-Ae^{2t}}}\ .

4. Curve shortening flows evolving by isometries

4.1. Properties of solutions

From the system of equations (2.18), we can deduce the following lemmata. We exclude the case of μ\mu identically zero. In this case, the curve X(s)X(s) and the constant vector vv are orthogonal, and then XX is a geodesic. We also take X(s)X(s) to be defined for all ss\in{\mathbb{R}}. Note that by equations (2.15) and (2.22) tautau, ν\nu, μ\mu, and κ\kappa will be bounded on the intersection of XX with any compact subset of 2{\mathbb{H}}^{2}, so inextendible curves trapped in a bounded set have infinite arclength. For those that escape any bounded set, the distance from any chosen p2p\in{\mathbb{H}}^{2} increases without bound, so the arclength parameter is unbounded in this case as well.

Lemma 4.1 (Behaviour of μ\mu).

Let (τ(s),ν(s),μ(s))(\tau(s),\nu(s),\mu(s)) be a solution of the system (2.18) and let s0s_{0} be a critical point of μ\mu. If μ\mu is not identically zero then s0s_{0} is the unique critical point of μ\mu. If μ(s0)>0\mu(s_{0})>0 it is a global minimum. If μ(s0)<0\mu(s_{0})<0 it is a global maximum. Furthermore, μ\mu has at most one zero. If μ\mu converges as ss\to\infty, then in fact limsμ(s)=0\lim_{s\to\infty}\mu(s)=0. Likewise, if μ\mu converges as ss\to-\infty, then limsμ(s)=0\lim_{s\to-\infty}\mu(s)=0. Otherwise, μ\mu diverges to ±\pm\infty.

Proof.

We have from (2.18) that

(4.1) μ′′(s)=τ(s)=aτ(s)ν(s)+μ(s)=aμ(s)ν(s)+μ(s).\mu^{\prime\prime}(s)=\tau^{\prime}(s)=a\tau(s)\nu(s)+\mu(s)=a\mu^{\prime}(s)\nu(s)+\mu(s)\ .

Now at a critical point we have μ(s0)=τ(s0)=0\mu^{\prime}(s_{0})=\tau(s_{0})=0. Since by (2.17) ν\nu cannot diverge at s0s_{0} then from (4.1) we see that μ′′(s0)=μ(s0)\mu^{\prime\prime}(s_{0})=\mu(s_{0}). If μ(s0)>0\mu(s_{0})>0 then the second derivative test implies that it must be a local minimum. Now say there is another critical point at s1s_{1}, and that no critical point lies between s0s_{0} and s1s_{1}. Then necessarily μ(s1)>μ(s0)>0\mu(s_{1})>\mu(s_{0})>0, so μ(s1)\mu(s_{1}) is also a local minimum and so there must be a local maximum between these critical points, which is a contradiction. Hence s0s_{0} is the unique critical point, and therefore a global minimum. The dual statement for the case of μ(s0)<0\mu(s_{0})<0 now follows in precisely similar fashion. In either of these cases, μ\mu has no zero.

If μ(s0)=0\mu(s_{0})=0 but μ(s)\mu(s) is not identically zero, then either (i) μ(s)<0\mu(s)<0 for s0<s<s0+δs_{0}<s<s_{0}+\delta (for some δ>0\delta>0) or (ii) μ(s)>0\mu(s)>0 for s0<s<s0+δs_{0}<s<s_{0}+\delta. In case (i), then by the above argument the next critical point at some s1>s0s_{1}>s_{0} must have μ′′(s1)<0\mu^{\prime\prime}(s_{1})<0 and would therefore be a local maximum, a contradiction; in case (ii) it must have μ′′(s1)>0\mu^{\prime\prime}(s_{1})>0, again a contradiction. So there is no critical point at s1>s0s_{1}>s_{0}. A similar argument shows that no critical point at s1<s0s_{1}<s_{0} either. Therefore, if μ(s0)=0\mu(s_{0})=0 then either s0s_{0} is the unique critical point, or μ\mu has no critical points at all in this case, and in either situation s0s_{0} is then the unique zero of μ\mu.

Since μ\mu has at most one critical point s0s_{0}, on the domain s<s0s<s_{0} we may take μ\mu to be monotonic and τ\tau to have a sign. The same statements are true on the domain s>s0s>s_{0}, and for all ss if μ\mu has no critical point. Then if |μ||\mu| does not diverge to \infty, μ\mu will converge. By way of contradiction, assume that μ\mu converges to a constant c0c\neq 0. Then from the third equation in (2.18) and recalling the τ\tau has a sign, we have τ0\tau\to 0. But then η(v,v)=τ2+ν2μ2\eta(v,v)=\tau^{2}+\nu^{2}-\mu^{2}, so τ2+ν2\tau^{2}+\nu^{2} is bounded, which implies that ν2\nu^{2} cannot diverge to \infty. Then the first equation in (2.18) yields τc\tau^{\prime}\to c. But then τ\tau would diverge, a contradiction, unless c=0c=0. ∎

Lemma 4.2 (Behaviour of ν\nu).

The value s0s_{0} is a critical point of ν\nu if and only if s0s_{0} is a critical point of μ\mu, and then s0s_{0} is a point of inflection for ν\nu, which is therefore monotonic. Either ν\nu is bounded and thus converges or it diverges to ++\infty as ss\to-\infty or to -\infty as s+s\to+\infty.

Proof.

Since by (2.18) we have ν=aτ2=a(μ)2\nu^{\prime}=-a\tau^{2}=-a(\mu^{\prime})^{2}, then it is obvious that s0s_{0} is a critical point of ν\nu if and only if s0s_{0} is a critical point of μ\mu, and that ν(s)\nu(s) is monotonic so any critical point is an inflection point. It also follows from monotonicity that if ν\nu is bounded it converges as s±s\to\pm\infty, and since it decreases monotonically, if it is not bounded it diverges as claimed. ∎

Notice that we have both a monotone quantity ν\nu and a conserved quantity η(v,v)=τ2+ν2μ2\eta(v,v)=\tau^{2}+\nu^{2}-\mu^{2}.

Lemma 4.3 (Linear growth).

Let (τ(s),ν(s),μ(s))(\tau(s),\nu(s),\mu(s)) be a solution of the system (2.18) such that μ\mu diverges as ss\to\infty. Then τ(s),ν(s),μ(s)𝒪(s)\tau(s),\nu(s),\mu(s)\in{\mathcal{O}}(s); i.e., they are each bounded above in magnitude by CsCs for some constant C>0C>0. This is also true as ss\to-\infty.

Proof.

Since xx2x\mapsto x^{2} is a convex function, Jensen’s inequality yields for s>s0s>s_{0} that

(4.2) (1(ss0)s0sτ(r)𝑑r)21(ss0)s0sτ2(r)𝑑r.\left(\frac{1}{(s-s_{0})}\int\limits_{s_{0}}^{s}\tau(r)dr\right)^{2}\leq\frac{1}{(s-s_{0})}\int\limits_{s_{0}}^{s}\tau^{2}(r)dr\ .

Since μ=τ\mu^{\prime}=\tau and ν=aτ2\nu^{\prime}=-a\tau^{2}, the above inequality implies that

(4.3) (μ(s)μ(s0))2(ss0)2(ν(s)ν(s0))a(ss0)μ(s)μ(s0)+1a(ss0)(ν(s0)ν(s)).\begin{split}&\,\frac{\left(\mu(s)-\mu(s_{0})\right)^{2}}{\left(s-s_{0}\right)^{2}}\leq-\frac{\left(\nu(s)-\nu(s_{0})\right)}{a\left(s-s_{0}\right)}\\ \implies&\,\mu(s)\leq\mu(s_{0})+\sqrt{\frac{1}{a}\left(s-s_{0}\right)\left(\nu(s_{0})-\nu(s)\right)}\ .\end{split}

In the last line, we recall that ν\nu is a decreasing function. Writing μ0:=μ(s0)\mu_{0}:=\mu(s_{0}) and ν0:=ν(s0)\nu_{0}:=\nu(s_{0}), then

(4.4) η(v,v)τ2+ν[ν+(ss0)a]2μ01a(ss0)(ν0ν)(ss0)aν0μ02,\eta(v,v)\geq\tau^{2}+\nu\left[\nu+\frac{(s-s_{0})}{a}\right]-2\mu_{0}\sqrt{\frac{1}{a}\left(s-s_{0}\right)\left(\nu_{0}-\nu\right)}-\frac{(s-s_{0})}{a}\nu_{0}-\mu_{0}^{2}\ ,

so that

(4.5) η(v,v)+μ02(ss0)2+ν0a(ss0)+2μ0(ν0ν)a(ss0)3τ2(ss0)2+ν(ss0)[ν(ss0)+1a].\frac{\eta(v,v)+\mu_{0}^{2}}{(s-s_{0})^{2}}+\frac{\nu_{0}}{a(s-s_{0})}+2\mu_{0}\sqrt{\frac{\left(\nu_{0}-\nu\right)}{a(s-s_{0})^{3}}}\geq\frac{\tau^{2}}{(s-s_{0})^{2}}+\frac{\nu}{(s-s_{0})}\left[\frac{\nu}{(s-s_{0})}+\frac{1}{a}\right]\ .

If there is a sequence of values si±s_{i}\to\pm\infty such that either |τ(si)||\tau(s_{i})| or |ν(si)||\nu(s_{i})| (or both) grows faster than linearly, then this inequality cannot hold. But then by (4.3), |μ||\mu| also cannot grow faster than linearly. ∎

Lemma 4.4 (Bounded curvature).

The curvature and τ\tau are bounded.

Proof.

Since κg=aτ\kappa_{g}=a\tau, then κg\kappa_{g} is bounded iff τ\tau is.

Consider first ss\to\infty. If ν\nu diverges, then necessarily ν\nu\to-\infty as ss\to\infty, so take s1s_{1} such that aν(s1)+2<0a\nu(s_{1})+2<0. Since ν\nu is monotonic, then aν(s)+2<0a\nu(s)+2<0 for all ss1.s\geq s_{1}. Since μ\mu has at most one critical point and therefore τ\tau has at most one zero, by increasing s1s_{1} if necessary we can take τ(s)>0\tau(s)>0 and μ(s)>0\mu(s)>0 for ss1s\geq s_{1}. (We will deal with other cases at the end.)

Under these circumstances, we have from equations (2.18) and (2.17) that

(4.6) τ=aντ+μ=avτ+τ2+ν2η(v,v)aντ+τ2+ν2+|η(v,v)|aντ+2τ2+ν2,\begin{split}\tau^{\prime}=&\,a\nu\tau+\mu=av\tau+\sqrt{\tau^{2}+\nu^{2}-\eta(v,v)}\leq a\nu\tau+\sqrt{\tau^{2}+\nu^{2}+\left|\eta(v,v)\right|}\\ \leq&\,a\nu\tau+2\sqrt{\tau^{2}+\nu^{2}}\ ,\end{split}

where we may have to increase s1s_{1} so that the last inequality holds (indeed, we have τaντ+τ2+ν2\tau^{\prime}\leq a\nu\tau+\sqrt{\tau^{2}+\nu^{2}} if vv is null or spacelike). Furthermore, since ν<0\nu<0 and τ>0\tau>0, then τ2+ν2<τ2+ν22τν=τν\sqrt{\tau^{2}+\nu^{2}}<\sqrt{\tau^{2}+\nu^{2}-2\tau\nu}=\tau-\nu, so we get from (4.6) that

(4.7) τ<aντ+2τ2ν=(aν+2)τ2ν.\tau^{\prime}<a\nu\tau+2\tau-2\nu=(a\nu+2)\tau-2\nu\ .

But aν+2<0a\nu+2<0, and so whenever τ>2νaν+2\tau>\frac{2\nu}{a\nu+2} equation (4.7) will yield τ<2ν2ν=0\tau^{\prime}<2\nu-2\nu=0. Rewriting

(4.8) 2νaν+2=2a4a(aν(s)+2)2a4a(aν(s1)+2)=2ν(s1)aν(s1)+2\frac{2\nu}{a\nu+2}=\frac{2}{a}-\frac{4}{a(a\nu(s)+2)}\leq\frac{2}{a}-\frac{4}{a(a\nu(s_{1})+2)}=\frac{2\nu(s_{1})}{a\nu(s_{1})+2}

using the monotonicity of ν\nu, we therefore obtain for ss1s\geq s_{1} that

(4.9) 0<τ(s)max{τ(s1),2ν(s1)aν(s1)+2}.0<\tau(s)\leq\max\left\{\tau(s_{1}),\frac{2\nu(s_{1})}{a\nu(s_{1})+2}\right\}\ .

Next we deal briefly with the case where τ(s1)<0\tau(s_{1})<0 and μ(s1)<0\mu(s_{1})<0. Then subsequently τ(s)<0\tau(s)<0 and μ(s)<0\mu(s)<0 for all s>s1s>s_{1}. As above, ν(s)<0\nu(s)<0 and ν\nu\to-\infty. Repeating the calculations of equations (4.6) and (4.7) with the appropriate sign changes, we now have

(4.10) τ=aνττ2+ν2η(v,v)aνττ2+ν2+|η(v,v)|aντ2τ2+ν2aντ2(τ+ν)(aν2)τ,\begin{split}\tau^{\prime}=&\,a\nu\tau-\sqrt{\tau^{2}+\nu^{2}-\eta(v,v)}\geq a\nu\tau-\sqrt{\tau^{2}+\nu^{2}+\left|\eta(v,v)\right|}\\ \geq&\,a\nu\tau-2\sqrt{\tau^{2}+\nu^{2}}\geq a\nu\tau-2(\tau+\nu)\\ \geq&\,(a\nu-2)\tau\ ,\end{split}

where in the last line we dropped the term 2ν-2\nu since ν(s)<0\nu(s)<0. Using aν(s)2<0a\nu(s)-2<0 for ss1s\geq s_{1} when s1s_{1} large enough, then we have that τ>0\tau^{\prime}>0 whenever τ<1aν2<0\tau<\frac{1}{a\nu-2}<0. Alternatively, we can simply observe that the logarithmic derivative of τ\tau is negative, so the magnitude of τ\tau decreases. Hence τ\tau is bounded.

To treat the limit ss\to-\infty, note that we can replace ss by r:=sr:=-s in the derivatives in (2.18) and then treat the limit rr\to\infty. If we also replace μμ\mu\mapsto-\mu and νν\nu\mapsto-\nu, we recover the same system as (2.18). We will again obtain that τ\tau is bounded as rr\to\infty and therefore as ss\to-\infty. ∎

Lemma 4.5 (Convergence).

If either μ\mu or ν\nu is bounded, then all three functions τ\tau, ν\nu, and μ\mu converge, with μ0\mu\to 0, τ0\tau\to 0, and ν±η(v,v)\nu\to\pm\sqrt{\eta(v,v)}. This can occur only when vv is achronal (i.e., spacelike or null). If μ\mu and ν\nu are not bounded, then τ±1a\tau\to\pm\frac{1}{a}.

Proof.

If μ\mu is bounded (and thus converges to zero by Lemma 4.1), then from (2.17) it is clear that τ\tau and ν\nu are bounded. Since ν\nu is monotonic, it must converge, and then by (2.17) so must τ\tau.

If instead ν\nu is bounded, since it’s monotonic it must converge, and then as argued in the proof in the previous lemma, we have τ0\tau\to 0. Then by (2.17) μ\mu converges as well. By Lemma 4.1, then μ0\mu\to 0. Hence (2.17) yields ν2η(v,v)\nu^{2}\to\eta(v,v), which requires that η(v,v)0\eta(v,v)\geq 0 so vv must be achronal.

If instead neither μ\mu nor ν\nu are bounded, then write (2.17) as

(4.11) (ν+μ)(νμ)=τ2+η(v,v).(\nu+\mu)(\nu-\mu)=\tau^{2}+\eta(v,v)\ .

Since τ\tau is bounded, so is the product (ν+μ)(νμ)(\nu+\mu)(\nu-\mu). But neither μ\mu nor ν\nu is bounded so one factor in this product diverges and so the other factor must converge to zero. Then limsμν=±1\lim_{s\to\infty}\frac{\mu}{\nu}=\pm 1.

Because τ\tau is bounded, lim supsτ(s)\limsup_{s\to\infty}\tau(s) and lim infsτ(s)\liminf_{s\to\infty}\tau(s) exist. Now lim supsτ(s)=limατ(sα)\limsup_{s\to\infty}\tau(s)=\lim_{\alpha\to\infty}\tau(s_{\alpha}) where each τ(sα)\tau(s_{\alpha}) is a local maximum of τ\tau. At each local maximum, we have by the first equation in (2.18) that τ(sα)=μ(sα)aν(sα)\tau(s_{\alpha})=-\frac{\mu(s_{\alpha})}{a\nu(s_{\alpha})}, so lim supsτ=lim supsμ(sα)aν(sα)=1a\limsup_{s\to\infty}\tau=-\limsup_{s\to\infty}\frac{\mu(s_{\alpha})}{a\nu(s_{\alpha})}=\mp\frac{1}{a} using the result of the previous paragraph. But we can replace the local maxima τ(sα)\tau(s_{\alpha}) by local minima, say τ(sβ)\tau(s_{\beta}) and obtain lim infsτ=1a\liminf_{s\to\infty}\tau=\mp\frac{1}{a} (with the same choice of sign). Hence limsτ=1a\lim_{s\to\infty}\tau=\mp\frac{1}{a}. ∎

4.2. Spacelike v~{\tilde{v}}

Each spacelike vector v~𝕄3{\tilde{v}}\in{\mathbb{M}}^{3} is contained in a 11-parameter family of planes, exactly two of which are null. Boosts in the spacelike directions orthogonal to each such plane preserve the planes and preserve ±v~\pm{\tilde{v}}, but do not preserve any spacelike vector in these planes.

We can apply a constant boost to the standard basis of 𝕄3{\mathbb{M}}^{3} so that an arbitrary spacelike v~{\tilde{v}} has the form v~=(v~1,v~2,0){\tilde{v}}=({\tilde{v}}_{1},{\tilde{v}}_{2},0) in the transformed basis. A subsequent constant rotation of the basis brings vv to the form v~=a(0,1,0){\tilde{v}}=a(0,1,0) for some a>0a>0. As before, we will normalize and write v=v~/av={\tilde{v}}/a. Then vv lies in the intersection of the null plane Span{(0,1,0),(1,0,1)}\operatorname{Span}\{(0,1,0),(1,0,1)\} with the vertical x=0x=0 plane, while μ(s)\mu(s) as defined by (2.15) is simply the yy-coordinate of X(s)X(s). The curve of intersection of the x=0x=0 plane and the unit hyperboloid is the geodesic in 𝕄3{\mathbb{M}}^{3} whose trace is the graph of z=1+y2z=\sqrt{1+y^{2}}, x=0x=0. The intersection of the unit hyperboloid and the plane y=0y=0 orthogonal to vv is also a geodesic, which we denote by Γv\Gamma_{v}. We will label the y>0y>0 half-space (into which vv points) in 𝕄3{\mathbb{M}}^{3} by H+H_{+} and the y<0y<0 half-space by HH_{-}.

Lemma 4.6.

Let X(s)X(s) be a non-geodesic curve evolving by isometries under CSF with spacelike v~{\tilde{v}} and let v=v~/η(v~,v~)v={\tilde{v}}/\sqrt{\eta({\tilde{v}},{\tilde{v}})}. Then exactly one of the following holds:

  • (i)

    μ(s)\mu(s) has exactly one critical point s0s_{0}, a minimum, X(s)X(s) lies entirely in H+H_{+}, and X(s)X(s) extends to infinity at both ends. See Figure 1.

  • (ii)

    μ(s)\mu(s) has exactly one critical point s0s_{0}, a maximum, X(s)X(s) lies entirely in HH_{-}, and X(s)X(s) extends to infinity at both ends. See Figure 2.

  • (iii)

    μ(s)\mu(s) has no critical point, and either converges to Γv\Gamma_{v} at one end while extending to infinity at the other or intersects Γv\Gamma_{v} at one point and extends to infinity at both ends. See Figures 3 and 4.

Proof.

If μ\mu has a critical point in H+H_{+}, by Lemma 4.1 it is a global minimum. Then μ(s)μ(s0)\mu(s)\geq\mu(s_{0}) for all ss, so the curve must remain in H+H_{+}. Likewise if μ\mu has a critical point in HH_{-}, is it a global maximum, and so the curve must remain in HH_{-}. In either case, μ\mu is bounded away from zero so by Lemma 4.1 μ\mu\to\infty at both ends s±s\to\pm\infty if the curve is in H+H_{+}, and μ\mu\to-\infty at both ends s±s\to\pm\infty if the curve is in HH_{-}.

If a critical point s0s_{0} of μ\mu were to lie on Γv\Gamma_{v}, then μ(s0)=μ(s0)=0\mu(s_{0})=\mu^{\prime}(s_{0})=0. Then τ(s0)=0\tau(s_{0})=0 and ν(s0)=±1\nu(s_{0})=\pm 1. (We can further deduce that τ(s0)=0\tau^{\prime}(s_{0})=0 and μ′′(s0)=0\mu^{\prime\prime}(s_{0})=0.) For these initial data, the unique maximal solution of the system (2.18) has τ(s)=0\tau(s)=0 for all ss, and thus κg(s)=aτ(s)=0\kappa_{g}(s)=a\tau(s)=0 for all ss, so the corresponding curve is a geodesic, contrary to assumption. This possibility therefore does not occur.

If a curve has no critical point of μ\mu then μ\mu is monotonic so μ\mu has at most one zero. Then the curve meets Γv\Gamma_{v} in at most one point. If it does not meet Γv\Gamma_{v}, then it must approach Γv\Gamma_{v} in the limit, since it cannot have a limit for μ\mu other than μ0\mu\to 0. ∎

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 1. Curve in the Poincaré disk model, evolving by isometry with spacelike v~{\tilde{v}}, as in Lemma 4.6.(i). The valley in the graph of τ\tau indicates the maximum of |κg||\kappa_{g}|.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 2. Curve in the Poincaré disk model evolving by isometries with spacelike v~{\tilde{v}}, as in Lemma 4.6.(ii).
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 3. Curve in the Poincaré disk model evolving by isometries with spacelike v~{\tilde{v}} and converging to a geodesic at one end, as in Lemma 4.6.(iii).
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4. Curve in the Poincaré disk model of hyperbolic space that evolves by isometry with spacelike vv corresponding to Lemma 4.6.(iii), such that μ\mu has a zero and diverges at both ends.

4.3. Timelike v~{\tilde{v}}

For a timelike v~{\tilde{v}}, one can apply a constant boost to the axes of the standard basis of 𝕄3{\mathbb{M}}^{3} so that in the boosted basis v~{\tilde{v}} takes the form v~=(0,0,θ˙)=θ˙(0,0,1)=θ˙v{\tilde{v}}=(0,0,{\dot{\theta}})={\dot{\theta}}(0,0,1)={\dot{\theta}}v. In this case, μ\mu is the negative of the zz-coordinate of the curve. We’ve chosen the sign so that vv is future-timelike, though v~{\tilde{v}} can be either future- or past-timelike depending on the sign of θ˙{\dot{\theta}}.

Lemma 4.7.

Let X(s)X(s) be a non-geodesic curve evolving by isometries under CSF with future-timelike vv. Then there is exactly one critical point of μ(s)\mu(s) along X(s)X(s), a global maximum, and μ\mu\to-\infty as s±s\to\pm\infty along X(s)X(s). See Figures 5 and 6.

Proof.

Since both XX and vv are future-timelike and nonzero, then μ(s)c<0\mu(s)\leq c<0 for some c>0c>0 and all ss, so by Lemma 4.1 μ\mu has a negative global maximum and no other critical point. Since μ(s)\mu(s) cannot converge to 0 along X(s)X(s), it cannot converge at all and so must diverge to -\infty as s±s\to\pm\infty. ∎

Since μ\mu is proportional to z-z along a soliton with future-timelike vv, then μ\mu\to\infty implies that the curve X(s)X(s) extends to infinity along the unit hyperboloid in 𝕄3{\mathbb{M}}^{3}.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 5. Curve in the Poincaré disk model evolving by isometries with timelike v~{\tilde{v}}.
Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6. Curve in the Poincaré disk model evolving by isometries with timelike v~{\tilde{v}}, which passes (nearly) through the origin, where μ\mu is maximized.

4.4. Null v~{\tilde{v}}

The remaining possibility is that v~{\tilde{v}} is null. Applying a constant rotation to the axes if necessary, we can take v~=a(0,1,1)=:av{\tilde{v}}=a(0,1,1)=:av, so μ=y(s)z(s)<0\mu=y(s)-z(s)<0 along X(s)=(x(s),y(s),z(s))X(s)=(x(s),y(s),z(s)), since y<zy<z at any point on the unit hyperboloid z=x2+y2z=\sqrt{x^{2}+y^{2}} in 𝕄3{\mathbb{M}}^{3}.

Lemma 4.8.

Let X(s)X(s) be a non-geodesic curve evolving by isometries under CSF with future-null vv. Then there is at most one critical point of μ(s)\mu(s) along the curve X(s)=(x(s),y(s),z(s))X(s)=(x(s),y(s),z(s)), a global maximum, and z(s)z(s)\to\infty at both ends of its domain. See Figure 7.

Proof.

Since y<zy<z at each point of the hyperboloid x2+y2z2=1x^{2}+y^{2}-z^{2}=-1, z>0z>0, then μ(s)=y(s)z(s)<0\mu(s)=y(s)-z(s)<0 so by Lemma 4.1 there is at most one critical point s0s_{0} of μ\mu, and it must be a global maximum. When this occurs, then μ(s)μ(s0)<0\mu(s)\leq\mu(s_{0})<0 so μ(s)\mu(s) cannot converge to zero, and so μ\mu\to-\infty and zz\to\infty as s±s\to\pm\infty.

On the other hand, if there is no such maximum then μ\mu converges to zero from below as ss\to\infty (or as ss\to-\infty, but not both). But for future-null vv, v,Xμ0\left\langle v,X\right\rangle\equiv\mu\to 0 implies that yz0y-z\to 0 along the curve. Since x2+(yz)(y+z)=1x^{2}+(y-z)(y+z)=-1, this can only happen if y,z±y,z\to\pm\infty. ∎

For past-null vv, we take vvv\mapsto-v so μμ\mu\mapsto-\mu and the global maximum in the lemma becomes a global minimum.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 7. Curve in the Poincaré disk model evolving by isometries with null v~{\tilde{v}}.

4.5. Proof of the main theorem

First we give a brief proof of the following simple property.

Lemma 4.9.

Curves evolving by isometries in 2{\mathbb{H}}^{2} have no self-intersections, and are properly embedded.

Proof.

By way of contradiction, say that a curve X(s)X(s) self-intersects at two parameter values s2>s1s_{2}>s_{1} with no self-intersections in between. At the self-intersection, the ingoing and outgoing tangents make angle α\alpha. The region bounded by X(s)X(s) for s1ss2s_{1}\leq s\leq s_{2} has area A>0A>0. Since the Gauss curvature of 2{\mathbb{H}}^{2} is 1-1, Gauss-Bonnet yields

(4.12) s1s2κg𝑑s=A+π+α>π.\oint\limits_{s_{1}}^{s_{2}}\kappa_{g}ds=A+\pi+\alpha>\pi\ .

But at a point of self-intersection we must have μ(s1)=μ(s2)\mu(s_{1})=\mu(s_{2}), and so

(4.13) 0=μ(s2)μ(s1)=s1s2μ(s)𝑑s=s1s2τ𝑑s=1as1s2κg𝑑s.0=\mu(s_{2})-\mu(s_{1})=\oint\limits_{s_{1}}^{s_{2}}\mu^{\prime}(s)ds=\oint\limits_{s_{1}}^{s_{2}}\tau ds=\frac{1}{a}\oint\limits_{s_{1}}^{s_{2}}\kappa_{g}ds\ .

Comparing (4.12) and (4.13), we arrive at a contradiction, so there are no self-intersections. Then since every curve in Lemmata 4.64.8 escapes any bounded set, there are no accumulation points either, so the curves are properly embedded. ∎

The main theorem now follows easily from the above results.

Proof of Theorem 1.2.

Choose a vector v~{\tilde{v}}. If it is not null, write v~=av{\tilde{v}}=av such that vv is normalized and define the variables τ\tau, ν\nu, and μ\mu along smooth curves X(s)X(s) using (2.15). If we require that the system (2.18) holds then from Proposition 2.2, each solution of (2.18) determines a curve X(s)X(s). Solutions of (2.18) are parametrized by the initial data (τ0,ν0,μ0)(\tau_{0},\nu_{0},\mu_{0}) for (2.18), subject to the constraint (2.20) at s=0s=0. Hence we obtain 22-parameter families of solutions of (2.18) and of the associated curves X(s)X(s). The solution curves have domain ss\in{\mathbb{R}}, for ss a unit speed parameter, and μ=X,v\mu=\langle X,v\rangle is divergent at least at one end, so the solutions are complete and noncompact.

If μ\mu converges at one end, then by Lemma 4.1 we have μ0\mu\to 0 while by Lemma 4.5 we also have τ0\tau\to 0 and thus by Proposition 2.3 then κg0\kappa_{g}\to 0. In this case X(s)X(s) approaches a geodesic at this end. Lemmata 4.64.8 give conditions under which this situation does not arise.

If μ\mu diverges, invoking Lemma 4.5 for ss\to\infty, or for ss\to-\infty as the case may be, we have that the curvature obeys κg±1a0\kappa_{g}\to\pm\frac{1}{a}\neq 0, and so XX approaches a horosphere.

Lemma 4.9 establishes that XX is properly embedded. ∎

References

  • [1] K-S Chou and X-P Zhu, The curve shortening problem, (CRC Press, Boca Raton, 2001).
  • [2] M Gage, An isoperimetric inequality with application to curve shortening, Duke Math J 50 (1983, 1225–1229.
  • [3] M Gage, Curve shortening makes convex curves circular, Invent Math 76 (1984) 357–364.
  • [4] ME Gage, Curve shortening on surfaces, Ann Sci l’ÉNS 4e4^{e} série, 23 (1990) 229–256.
  • [5] M Gage and RS Hamilton, The heat equation shrinking convex plane curves, J Diff Geom 23 (1986) 69–96.
  • [6] MA Grayson, Shortening embedded curves, Ann Math 129 (1989) 71–111.
  • [7] HP Halldórsson, Self-similar solutions to the mean curvature flow in Euclidean and Minkowski space, PhD thesis, Massachusetts Institute of Technology, 2013 (unpublished).
  • [8] HP Halldórsson, Self-similar solutions to the curve shortening flow, Trans Amer Math Soc 364 (2012) 5285–5309.
  • [9] TWB Kibble, Topology of cosmic domains and strings, J Phys A: Math Gen 9 (1976) 1387–1398.
  • [10] FN da Silva, Fluxos de curvas no espaco hiperbólico e no cone de luz, University of Brasilia PhD thesis, (2020, unpublished).
  • [11] FN da Silva and K Tenenblat, Soliton solutions to the curve shortening flow on the 2-dimensional hyperbolic plane, preprint [arxiv:2102.07916v2].
  • [12] HF Santos Dos Reis and K Tenenblat, Soliton solutions to the curve shortening flow on the sphere, Proc Amer Math Soc 147 (2019) 11, 4955-4967.