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

Existence and Uniqueness of the Motion by Curvature of Regular Networks

Michael Gößwein 111Fakultät für Mathematik, Universität Duisburg–Essen, Thea-Leymann-Straße 9, 45127 Essen, Germany    Julia Menzel 222Fakultät für Mathematik, Universität Regensburg, Universitätsstrasse 31, 93053 Regensburg, Germany    Alessandra Pluda 333Dipartimento di Matematica, Università di Pisa, Largo Bruno Pontecorvo 5, 56127 Pisa, Italy
Abstract

We prove existence and uniqueness of the motion by curvature of networks with triple junctions in n\mathbb{R}^{n} when the initial datum is of class Wp22/pW^{2-\nicefrac{{2}}{{p}}}_{p} and the unit tangent vectors to the concurring curves form angles of 120120 degrees. Moreover we investigate the regularisation effect due to the parabolic nature of the system. An application of this wellposedness result is a new proof of [23, Theorem 3.18] where the possible behaviours of the solutions at the maximal time of existence are described.
Our study is motivated by an open question proposed in [22]: does there exist a unique solution of the motion by curvature of networks with initial datum being a regular network of class C2C^{2}? We give a positive answer.

MSC (2010): 53C44, 35K51 (primary); 35K59, 35D35 (secondary).

Keywords: Networks, motion by curvature, local existence and uniqueness, parabolic regularisation, nonlinear boundary conditions.

1 Introduction

The mean curvature flow of surfaces in n\mathbb{R}^{n}, and in Riemannian manifolds in general, is one of the most significant examples of geometric evolution equations. This evolution can be understood as the gradient flow of the area functional: a time–dependent surface evolves with normal velocity equal to its mean curvature at any point and time.

From the 80s the curve shortening flow (mean curvature flow of one–dimensional objects) was widely studied by many authors both for closed curves [9, 10, 11, 15] and for curves with fixed end–points [16, 28, 29]. Also two concurring curves forming an angle or a cusp can be regarded as a single curve with a singular point which will vanish immediately under the flow [2, 3, 4]. When more than two curves meet at a junction, the description of the motion cannot be reduced to the case of a single curve and the problem presents new interesting features. The simplest example of motion by mean curvature of a set which is essentially singular is indeed the motion by curvature of networks that are finite unions of curves that meet at junctions.

Although after the seminal work by Brakke [5] several weak definitions of the motion by curvature of singular surfaces have been proposed, the first attempt to find strong solutions to the network flow was by Bronsard and Reitich [6] providing a well posedness result for initial data of class C2+αC^{2+\alpha}. The analysis of the long time behavior of the evolving networks was undertaken in [23], completed in [21] for trees composed of three curves and extended to more general cases in [17, 22, 24].

In this paper we restrict to regular networks that possess only triple junctions where the unit tangent vectors of the concurring curves form angles of 120120 degrees. The motion by curvature of networks can be expressed as a boundary value problem where the evolution of each curve is described by a second order quasilinear PDE as given in Definition 2.22.

Our main result is the following.

Theorem 1.1 (Existence, uniqueness and smoothness of the motion by curvature).

Let p(3,)p\in(3,\infty) and 𝒩0\mathcal{N}_{0} be a regular network in n\mathbb{R}^{n} of class Wp22/pW^{2-\nicefrac{{2}}{{p}}}_{p}. Then there exists a maximal solution (𝒩(t))t[0,Tmax)\left(\mathcal{N}(t)\right)_{t\in[0,T_{\max})} to the motion by curvature with initial datum 𝒩0\mathcal{N}_{0} in the maximal time interval [0,Tmax)[0,T_{\max}) which is geometrically unique and locally of regularity

𝑬T:=Wp1((an,bn);Lp((0,1);(n)3))Lp((0,T);Wp2((an,bn);(n)3))\boldsymbol{E}_{T}:=W_{p}^{1}\left((a_{n},b_{n});L_{p}((0,1);(\mathbb{R}^{n})^{3})\right)\cap L_{p}\left((0,T);W_{p}^{2}\left((a_{n},b_{n});(\mathbb{R}^{n})^{3}\right)\right)

for intervals [an,bn][0,Tmax)[a_{n},b_{n}]\subset[0,T_{max}), nn\in\mathbb{N}, with n[an,bn]=[0,Tmax)\bigcup_{n\in\mathbb{N}}[a_{n},b_{n}]=[0,T_{max}). Furthermore, the parametrisation γ:[0,Tmax)×[0,1](n)3\gamma:[0,T_{max})\times[0,1]\to(\mathbb{R}^{n})^{3} that parametrises the curves of the networks with constant speed equal to their length is smooth for positive times.

As several solutions to the motion by curvature can be obtained by parametrising the same set with different maps, the uniqueness has to be understood in a purely geometric sense namely, up to reparametrisations.

Local existence and uniqueness was proved in [6] for admissible initial networks of class C2+αC^{2+\alpha} with the sum of the curvature at the junctions equal to zero. When the initial datum is a regular network of class C2C^{2} without any restriction on the curvature at the junctions, existence (but not uniqueness) has been established in [22, Theorem 6.8]. Theorem 1.1 improves the uniqueness result by Bronsard and Reitich passing from initial data in C2+αC^{2+\alpha} to Wp22/pW^{2-\nicefrac{{2}}{{p}}}_{p} which gives a fortiori uniqueness for regular networks of class C2C^{2} and even of class H2H^{2} (take any p(3,6]p\in(3,6]).

We discuss now in more details Theorem 1.1. The motion by curvature of networks is described by a parabolic system of degenerate PDEs where only the normal movements of the curves are prescribed. We specify a suitable tangential component of the velocity to turn the problem into a system of non–degenerate second order quasilinear PDEs, the so–called Special Flow (Definition 2.26). Then we linearise the Special Flow around the initial datum and prove existence and uniqueness for the linearised problem in Section 3.1. Wellposedness of the linear system follows by Solonnikov’s theory [27] provided that the system is parabolic and that the complementary conditions hold. Both properties were already shown in [6], nevertheless we present a new and shorter proof of the complementary conditions. Solutions to the Special Flow are obtained by a contraction argument in Section 3.2. The solution to the Special Flow induces a solution to the motion by curvature of networks. To conclude the uniqueness result it is then enough to prove that two different solutions differ only by a reparametrisation but they are actually the same set as shown in Section 3.3. Existence and uniqueness of maximal solutions can then be deduced with standard arguments. Given p(3,)p\in(3,\infty) our solution space 𝑬T\boldsymbol{E}_{T} embeds into C([0,T];C1+α([0,1];(n)3))C\left([0,T];C^{1+\alpha}\left([0,1];(\mathbb{R}^{n})^{3}\right)\right). This choice allows us to define the boundary conditions pointwise and to use the theory of [27] for the associated linear system. Moreover the above regularity is needed in the contraction estimates because of the quasilinear nature of the equations.

Because of the parabolicity of the problem it is natural to ask whether the regularity of the evolving network increases during the flow. We give a positive answer to this question in Section 4 proving that the flow is smooth for all positive times (see Theorem 4.8). The idea of the proof is based on the so called parameter trick which is due to Angenent [2]. Although this strategy has been generalized to several situations [19, 20, 25], it should be pointed out that our system is not among the cases treated above because of the fully non-linear boundary condition

i=13γxi|γxi|=0.\sum_{i=1}^{3}\frac{\gamma^{i}_{x}}{|\gamma^{i}_{x}|}=0\,.

In [12] a strategy has been developed to prove smoothness for positive times of the surface diffusion flow for triple junction clusters with the same non–linear boundary condition. We follow this approach and modify the arguments to our setting to complete the proof of Theorem 1.1.

Finally a description of the possible different behaviours of the solutions as time tends to the maximal time of existence is desirable. Thanks to Theorem 1.1 and the quantification of the existence time of solutions to the Special Flow in terms of the initial values as given in Theorem 3.13 we are also able to prove the following:

Theorem 1.2 (Long time behaviour).

Let p(3,6]p\in(3,6], 𝒩0\mathcal{N}_{0} an admissible network of class Wp22/pW_{p}^{2-\nicefrac{{2}}{{p}}} and (𝒩(t))t[0,Tmax)\left(\mathcal{N}(t)\right)_{t\in[0,T_{\max})} be a maximal solution to the motion by curvature with initial datum 𝒩0\mathcal{N}_{0} in [0,Tmax)[0,T_{\max}) where Tmax(0,){}T_{\max}\in(0,\infty)\cup\{\infty\}. Then

Tmax=T_{\max}=\infty

or as tTmaxt\nearrow T_{\max} at least one of the following happens:

  • i)i)

    the inferior limit of the length of at least one curve of the network 𝒩(t)\mathcal{N}(t) is zero;

  • ii)ii)

    the superior limit of the L2L^{2}–norm of the curvature of the network is ++\infty.

This result was first shown for planar networks in [23, Theorem 3.18]. The benefit of our proof is that energy estimates are completely avoided.

We describe here the structure of the paper. In Section 2 we define the motion by curvature of networks and we introduce the solution space together with useful properties. Section 3 is devoted to prove existence of solutions to the motion by curvature and their geometric uniqueness. Then in Section 4 we explore the regularisation effect of the flow resulting in the proof of Theorem 1.1. We conclude with the proof of Theorem 1.2 in Section 5 giving a description of the behaviour of solutions at their maximal time of existence.

Acknowledgements

The authors gratefully acknowledge the support by the Deutsche Forschungsgemeinschaft (DFG) via the GRK 1692 “Curvature, Cycles, and Cohomology”.

2 Solutions to the Motion by Curvature of Networks

2.1 Preliminaries on function spaces

This paper is devoted to show well-posedness of a second order evolution equation. One natural solution space is given by

Wp1,2((0,T)×(0,1);d):=Wp1((0,T);Lp((0,1);d))Lp((0,T);Wp2((0,1);d))W_{p}^{1,2}\left((0,T)\times(0,1);\mathbb{R}^{d}\right):=W_{p}^{1}\left((0,T);L_{p}((0,1);\mathbb{R}^{d})\right)\cap L_{p}\left((0,T);W_{p}^{2}\left((0,1);\mathbb{R}^{d}\right)\right)

where TT is positive representing the time of existence and dd\in\mathbb{N} is any natural number. This space should be understood as the intersection of two Bochner spaces that are Sobolev spaces defined on a measure space with values in a Banach space. We give a brief summary in the case that the measure space is an interval. A detailed introduction on Bochner spaces can be found in [31].

Definition 2.1.

Let II\subset\mathbb{R} be an open interval and XX be a Banach space. A function f:IXf:I\to X is called strongly measurable if there exists a family of simple functions fn:IXf_{n}:I\to X, nn\in\mathbb{N}, such that for almost every xIx\in I,

limnfn(x)f(x)=0.\lim\limits_{n\to\infty}\left\lVert f_{n}(x)-f(x)\right\rVert=0\,.

Here, a function g:IXg:I\to X is called simple if

g=k=1Nakχ(bk,ck)g=\sum_{k=1}^{N}a_{k}\chi_{(b_{k},c_{k})}

for NN\in\mathbb{N}, akXa_{k}\in X, bk,ckIb_{k},c_{k}\in I and bk<ckb_{k}<c_{k} for k{1,,N}k\in\{1,\dots,N\}.

If f:IXf:I\to X is strongly measurable, then fX:I\left\lVert f\right\rVert_{X}:I\to\mathbb{R} is Lebesgue measurable. This justifies the following definition.

Definition 2.2 (LpL_{p}–spaces).

Let II\subset\mathbb{R} be an open interval and XX be a Banach space. For 1p1\leq p\leq\infty, we define the LpL_{p}–space

Lp(I;X):={f:IX strongly measurable :fLp(I;X)<},L_{p}\left(I;X\right):=\left\{f:I\to X\text{ strongly measurable }:\left\lVert f\right\rVert_{L_{p}(I;X)}<\infty\right\}\,,

where fLp(I;X):=f()XLp(I;)\left\lVert f\right\rVert_{L_{p}\left(I;X\right)}:=\left\lVert\left\lVert f(\cdot)\right\rVert_{X}\right\rVert_{L_{p}\left(I;\mathbb{R}\right)}. Furthermore, we let

L1,loc(I;X):={f:IX strongly measurable: for all KI compact, f|KL1(K;X)}.L_{1,loc}\left(I;X\right):=\left\{f:I\to X\text{ strongly measurable}:\text{ for all }K\subset I\text{ compact, }f_{|K}\in L_{1}\left(K;X\right)\right\}\,.
Definition 2.3.

Let II\subset\mathbb{R} be an open interval, XX be a Banach space, fL1,loc(I;X)f\in L_{1,loc}(I;X) and k0k\in\mathbb{N}_{0}. The kk-th distributional derivative xkf\partial_{x}^{k}f of ff is the functional on C0(I;)C_{0}^{\infty}(I;\mathbb{R}) given by

ϕ,xkf:=(1)|α|If(x)xkϕ(x)dx.\langle\phi\,,\,\partial_{x}^{k}f\rangle:=(-1)^{\lvert\alpha\rvert}\int_{I}f(x)\partial_{x}^{k}\phi(x)\mathrm{d}x\,.

The distribution xkf\partial^{k}_{x}f is called regular if there exists vL1,loc(I;X)v\in L_{1,loc}(I;X) such that

ϕ,xkf:=Iv(x)ϕ(x)dx.\langle\phi\,,\,\partial^{k}_{x}f\rangle:=\int_{I}v(x)\phi(x)\mathrm{d}x\,.

In this case we write xkf=vL1,loc(I;X)\partial^{k}_{x}f=v\in L_{1,loc}(I;X).

Definition 2.4 (Sobolev spaces).

Let mm\in\mathbb{N}, II\subset\mathbb{R} be an open interval and XX be a Banach space. For 1p1\leq p\leq\infty the Sobolev space of order mm\in\mathbb{N} is defined as

Wpm(I;X):={fLp(I;X):xkfLp(I;X) for all 1km},W_{p}^{m}(I;X):=\left\{f\in L_{p}(I;X):\partial^{k}_{x}f\in L_{p}(I;X)\text{ for all }1\leq k\leq m\right\}\,,

where xkf\partial^{k}_{x}f is the distributional derivative defined in Definition 2.3.

It is well-known that the space Wpm(I;X)W^{m}_{p}\left(I;X\right) is a Banach space in the norm

fWpm(I;X):={(0kmxkfLp(I;X)p)1/p,1p<,max0kmxkfL(I;X),p=.\lVert f\rVert_{W^{m}_{p}(I;X)}:=\left\{\begin{array}[]{cl}\left(\sum_{0\leq k\leq m}\lVert\partial^{k}_{x}f\rVert_{L_{p}(I;X)}^{p}\right)^{\nicefrac{{1}}{{p}}}\,,&1\leq p<\infty\,,\\ \max_{0\leq k\leq m}\lVert\partial^{k}_{x}f\rVert_{L_{\infty}(I;X)}\,,&p=\infty\,.\end{array}\right. (2.1)

Elements in the solution space

𝑬T:=Wp1((0,T);Lp((0,1);(n)3))Lp((0,T);Wp2((0,1);(n)3))\boldsymbol{E}_{T}:=W_{p}^{1}\left((0,T);L_{p}((0,1);(\mathbb{R}^{n})^{3})\right)\cap L_{p}\left((0,T);W_{p}^{2}\left((0,1);(\mathbb{R}^{n})^{3}\right)\right)

are thus functions fLp((0,T);Lp((0,1);(n)3)))f\in L_{p}\left((0,T);L_{p}\left((0,1);(\mathbb{R}^{n})^{3})\right)\right) possessing one distributional derivative with respect to time tfLp((0,T);Lp((0,1);(n)3)))\partial_{t}f\in L_{p}\left((0,T);L_{p}\left((0,1);(\mathbb{R}^{n})^{3})\right)\right) in the sense of Definition 2.3. Furthermore, for almost every t(0,T)t\in(0,T), the function f(t)f(t) lies in Wp2((0,1);(n)3))W_{p}^{2}\left((0,1);(\mathbb{R}^{n})^{3})\right) and thus has two spacial derivatives x(f(t))\partial_{x}(f(t)), x2(f(t))Lp((0,1);(n)3))\partial_{x}^{2}\left(f(t)\right)\in L_{p}\left((0,1);(\mathbb{R}^{n})^{3})\right). One easily sees that the functions txk(f(t))t\mapsto\partial_{x}^{k}(f(t)) for k{1,2}k\in\{1,2\} lie in Lp((0,T);Lp((0,1);(n)3)))L_{p}\left((0,T);L_{p}\left((0,1);(\mathbb{R}^{n})^{3})\right)\right). The space 𝑬T\boldsymbol{E}_{T} is often denoted by Wp1,2((0,T)×(0,1);(n)3))W_{p}^{1,2}\left((0,T)\times(0,1);(\mathbb{R}^{n})^{3})\right). We also use the notation Wp1,2:=𝑬T\left\lVert\cdot\right\rVert_{W_{p}^{1,2}}:=\left\lVert\cdot\right\rVert_{\boldsymbol{E}_{T}} where 𝑬T\left\lVert\cdot\right\rVert_{\boldsymbol{E}_{T}} is the corresponding norm on 𝑬T\boldsymbol{E}_{T} as defined in (2.1).

Definition 2.5 (Sobolev–Slobodeckij spaces).

Given dd\in\mathbb{N}, p[1,)p\in[1,\infty) and θ(0,1)\theta\in(0,1) the Slobodeckij semi-norm of an element fLp((0,1);d)f\in L_{p}\left((0,1);\mathbb{R}^{d}\right) is defined as

[f]θ,p:=(0101|f(x)f(y)|p|xy|θp+1dxdy)1/p.\left[f\right]_{\theta,p}:=\left(\int_{0}^{1}\int_{0}^{1}\frac{\left\lvert f(x)-f(y)\right\rvert^{p}}{\lvert x-y\rvert^{\theta p+1}}\,\mathrm{d}x\,\mathrm{d}y\right)^{\nicefrac{{1}}{{p}}}\,.

Let s(0,)s\in(0,\infty) be non–integer. The Sobolev–Slobodeckij space Wps((0,1);d)W_{p}^{s}\left((0,1);\mathbb{R}^{d}\right) is defined by

Wps((0,1);d):={fWps((0,1);d):[xsf]ss,p<}.W_{p}^{s}\left((0,1);\mathbb{R}^{d}\right):=\left\{f\in W_{p}^{\lfloor s\rfloor}\left((0,1);\mathbb{R}^{d}\right):\left[\partial_{x}^{\lfloor s\rfloor}f\right]_{s-\lfloor s\rfloor,p}<\infty\right\}\,.

The following result characterises the regularity of the initial values.

Theorem 2.6.

Let TT be positive, p(3,)p\in(3,\infty) and α(0,13/p]\alpha\in\left(0,1-\nicefrac{{3}}{{p}}\right]. We have continuous embeddings

Wp1,2((0,T)×(0,1))C([0,T];Wp22/p((0,1)))C([0,T];C1+α([0,1])).W_{p}^{1,2}\left((0,T)\times(0,1)\right)\hookrightarrow C\left([0,T];W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1)\right)\right)\hookrightarrow C\left([0,T];C^{1+\alpha}\left([0,1]\right)\right)\,.
Proof.

The first embedding follows from [7, Lemma 4.4], the second is an immediate consequence of the Sobolev Embedding Theorem [30, Theorem 4.6.1.(e)]. ∎

Similarly, we can specify the spaces of the boundary values.

Lemma 2.7.

Let TT be positive, dd\in\mathbb{N} and p[1,)p\in[1,\infty). Then the operator

Wp1,2((0,T)×(0,1);)\displaystyle W_{p}^{1,2}\left((0,T)\times(0,1);\mathbb{R}\right) Wp1/21/2p((0,T);Lp({0};))Lp((0,T);Wp11/p({0};)),\displaystyle\to W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left((0,T);L_{p}\left(\{0\};\mathbb{R}\right)\right)\cap L_{p}\left((0,T);W_{p}^{1-\nicefrac{{1}}{{p}}}\left(\{0\};\mathbb{R}\right)\right)\,,
f\displaystyle f (fx)|x=0\displaystyle\mapsto\left(f_{x}\right)_{|x=0}

is linear and continuous.

Proof.

This follows from [27, Theorem 5.1]. ∎

In this work we will use the following identification.

Proposition 2.8.

Let TT be positive, dd\in\mathbb{N} and p[1,)p\in[1,\infty). There is an isometric isomorphism

Wp1/21/2p((0,T);Lp({0};d))Lp((0,T);Wp11/p({0};d))Wp1/21/2p((0,T);d)W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left((0,T);L_{p}\left(\{0\};\mathbb{R}^{d}\right)\right)\cap L_{p}\left((0,T);W_{p}^{1-\nicefrac{{1}}{{p}}}\left(\{0\};\mathbb{R}^{d}\right)\right)\simeq W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left((0,T);\mathbb{R}^{d}\right)

via the map f(tf(t,0))f\mapsto\left(t\mapsto f(t,0)\right).

Proof.

It is shown in [18, page 406] that integration with respect to the volume element on the 0–dimensional manifold {0}\{0\} is given by integration with respect to the counting measure. That allows us to identify the space

Lp({0};d):={f:{0}d:{0}|f|pdσ=|f(0)|p<}L^{p}\left(\{0\};\mathbb{R}^{d}\right):=\left\{f:\{0\}\to\mathbb{R}^{d}:\int_{\{0\}}\left\lvert f\right\rvert^{p}\,\mathrm{d}\sigma=\left\lvert f(0)\right\rvert^{p}<\infty\right\}

with d\mathbb{R}^{d} via the isometric isomorphism I:Lp({0};d)dI:L_{p}\left(\{0\};\mathbb{R}^{d}\right)\to\mathbb{R}^{d}, ff(0)f\mapsto f(0). One easily sees that this operator restricts to I:Wps({0};d)dI:W_{p}^{s}\left(\{0\};\mathbb{R}^{d}\right)\to\mathbb{R}^{d} for every s>0s>0. ∎

Another important feature of Sobolev Slobodeckij spaces is their Banach algebra property.

Proposition 2.9.

Let II\subset\mathbb{R} be a bounded open interval, p[1,)p\in[1,\infty) and s(0,1)s\in(0,1) with s1p>0s-\frac{1}{p}>0. Then for f,gWps(I;)f,g\in W_{p}^{s}\left(I;\mathbb{R}\right) the product fgfg lies in Wps(I;)W_{p}^{s}\left(I;\mathbb{R}\right) and satisfies

fgWps(I;)C(s,p)(fC(I¯)gWps(I;)+gC(I¯)fWps(I;)).\left\lVert fg\right\rVert_{W_{p}^{s}(I;\mathbb{R})}\leq C(s,p)\left(\left\lVert f\right\rVert_{C(\overline{I})}\left\lVert g\right\rVert_{W_{p}^{s}(I;\mathbb{R})}+\left\lVert g\right\rVert_{C(\overline{I})}\left\lVert f\right\rVert_{W_{p}^{s}(I;\mathbb{R})}\right)\,.

Furthermore, given a smooth function F:dF:\mathbb{R}^{d}\to\mathbb{R}, dd\in\mathbb{N}, and a function fWps(I;d)f\in W_{p}^{s}\left(I;\mathbb{R}^{d}\right), the function tF(f(t))t\mapsto F(f(t)) lies in Wps(I;)W_{p}^{s}\left(I;\mathbb{R}\right).

Proof.

As Wps((0,1);)C(I¯;)W_{p}^{s}\left((0,1);\mathbb{R}\right)\hookrightarrow C(\overline{I};\mathbb{R}) due to the Sobolev Embedding Theorem [30, Theorem 4.6.1.(e)], we obtain for f,gWps(I;)f,g\in W_{p}^{s}\left(I;\mathbb{R}\right) the estimate

fgLp(I;)fC(I¯)gLp(I;)C(s,p)fWps(I;)gWps(I;)\left\lVert fg\right\rVert_{L_{p}(I;\mathbb{R})}\leq\left\lVert f\right\rVert_{C(\overline{I})}\left\lVert g\right\rVert_{L_{p}(I;\mathbb{R})}\leq C(s,p)\lVert f\rVert_{W_{p}^{s}(I;\mathbb{R})}\left\lVert g\right\rVert_{W_{p}^{s}(I;\mathbb{R})}

and

[fg]s,pp\displaystyle[fg]^{p}_{s,p} =II|(fg)(x)(fg)(y)|p|xy|sp+1dxdy\displaystyle=\int_{I}\int_{I}\frac{|(fg)(x)-(fg)(y)|^{p}}{|x-y|^{sp+1}}\mathrm{d}x\,\mathrm{d}y
II|g(x)|p|f(x)f(y)|p+|f(y)|p|g(x)g(y)|p|xy|sp+1dxdy\displaystyle\leq\int_{I}\int_{I}\frac{|g(x)|^{p}|f(x)-f(y)|^{p}+|f(y)|^{p}|g(x)-g(y)|^{p}}{|x-y|^{sp+1}}\mathrm{d}x\,\mathrm{d}y
gC(I¯)p[f]s,pp+fC(I¯)[g]s,ppC(s,p)fWps(I;)gWps(I;).\displaystyle\leq\left\lVert g\right\rVert_{C(\overline{I})}^{p}[f]^{p}_{s,p}+\left\lVert f\right\rVert_{C(\overline{I})}[g]_{s,p}^{p}\leq C(s,p)\lVert f\rVert_{W_{p}^{s}(I;\mathbb{R})}\left\lVert g\right\rVert_{W_{p}^{s}(I;\mathbb{R})}\,.

Let F:dF:\mathbb{R}^{d}\to\mathbb{R} be smooth and fWps(I;d)f\in W_{p}^{s}\left(I;\mathbb{R}^{d}\right). As ff lies in C(I¯;d)C(\overline{I};\mathbb{R}^{d}), there exists R>0R>0 such that f(I¯)BR(0)¯f(\overline{I})\subset\overline{B_{R}(0)}. Thus we obtain

F(f)Lp(I;)p=I|F(f(x))|pdxmaxzBR(0)¯|F(z)|p|I|\left\lVert F(f)\right\rVert^{p}_{L_{p}(I;\mathbb{R})}=\int_{I}|F(f(x))|^{p}\mathrm{d}x\leq\max_{z\in\overline{B_{R}(0)}}|F(z)|^{p}|I|

where |I||I| denotes the length of the interval II. Using

|F(f(x))F(f(y))|\displaystyle|F(f(x))-F(f(y))| =|01(DF)(ξf(x)+(1ξ)f(y))dξ(f(x)f(y))|\displaystyle=\left|\int_{0}^{1}(DF)\left(\xi f(x)+(1-\xi)f(y)\right)\mathrm{d}\xi\,(f(x)-f(y))\right|
maxzBR(0)¯|DF(z)||f(x)f(y)|\displaystyle\leq\max_{z\in\overline{B_{R}(0)}}|DF(z)||f(x)-f(y)|

we obtain

[F(f)]s,pp=II|F(f(x))F(f(y))|p|xy|sp+1dxdy[f]s,ppmaxzBR(0)¯|DF(z)|p.\displaystyle[F(f)]^{p}_{s,p}=\int_{I}\int_{I}\frac{|F(f(x))-F(f(y))|^{p}}{|x-y|^{sp+1}}\mathrm{d}x\,\mathrm{d}y\leq[f]_{s,p}^{p}\max_{z\in\overline{B_{R}(0)}}|DF(z)|^{p}\,.

To show well-posedness of evolution equations it is important to have embeddings with constants independent of the time interval one is working with. To this end one needs to change the norm on the solution space. In the following, we collect the results that are needed in our specific case.

Corollary 2.10.

Let p(3,)p\in(3,\infty). For every T>0T>0,

|g|Wp1,2((0,T)×(0,1)):=gWp1,2((0,T)×(0,1))+g(0)Wp22/p((0,1)){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|g\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{W_{p}^{1,2}\left((0,T)\times(0,1)\right)}:=\left\lVert g\right\rVert_{W_{p}^{1,2}\left((0,T)\times(0,1)\right)}+\left\lVert g(0)\right\rVert_{W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1)\right)}

defines a norm on Wp1,2((0,T)×(0,1))W_{p}^{1,2}\left((0,T)\times(0,1)\right) that is equivalent to the usual one.

Proof.

This is a consequence of Theorem 2.6. ∎

Lemma 2.11.

Let T0T_{0} be positive, T(0,T0)T\in(0,T_{0}) and p(3,)p\in(3,\infty). There exists a linear operator

𝑬:Wp1,2((0,T)×(0,1))Wp1,2((0,T0)×(0,1))\boldsymbol{E}:W_{p}^{1,2}\left((0,T)\times(0,1)\right)\to W_{p}^{1,2}\left((0,T_{0})\times(0,1)\right)

such that for all gWp1,2((0,T)×(0,1))g\in W_{p}^{1,2}\left((0,T)\times(0,1)\right), (𝐄g)|(0,T)=g\left(\boldsymbol{E}g\right)_{|(0,T)}=g and

𝑬gWp1,2((0,T0)×(0,1))C(gWp1,2((0,T)×(0,1))+g(0)Wp22/p(0,1))=C|g|Wp1,2((0,T)×(0,1))\left\lVert\boldsymbol{E}g\right\rVert_{W_{p}^{1,2}\left((0,T_{0})\times(0,1)\right)}\leq C\left(\left\lVert g\right\rVert_{W_{p}^{1,2}\left((0,T)\times(0,1)\right)}+\left\lVert g(0)\right\rVert_{W_{p}^{2-\nicefrac{{2}}{{p}}}(0,1)}\right)=C{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|g\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{W_{p}^{1,2}\left((0,T)\times(0,1)\right)}

with a constant C=C(p,T0)C=C(p,T_{0}) depending only on pp and T0T_{0}.

Proof.

In the case that g(0)=0g(0)=0, the function gg can be extended to (0,)(0,\infty) by reflecting it with respect to the axis t=Tt=T. The general statement can be deduced from this case by solving a linear parabolic equation of second order and using results on maximal regularity as given in [25, Proposition 3.4.3]. ∎

Given dd\in\mathbb{N} we obtain an extension operator on the space Wp1,2((0,T)×(0,1);d)W_{p}^{1,2}\left((0,T)\times(0,1);\mathbb{R}^{d}\right) by applying 𝑬\boldsymbol{E} to every component.

Lemma 2.12.

Let p(1,)p\in(1,\infty) and α>1p\alpha>\frac{1}{p}. For every positive TT,

|b|Wpα((0,T);):=bWpα((0,T);)+|b(0)|{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|b\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{W_{p}^{\alpha}\left((0,T);\mathbb{R}\right)}:=\left\lVert b\right\rVert_{W_{p}^{\alpha}\left((0,T);\mathbb{R}\right)}+|b(0)|

defines a norm on Wpα((0,T);)W_{p}^{\alpha}\left((0,T);\mathbb{R}\right) that is equivalent to the usual one.

Proof.

This is an immediate consequence of the Sobolev Embedding Theorem [30, Theorem 4.6.1.(e)]. ∎

Lemma 2.13.

Let TT be positive, p(1,)p\in(1,\infty) and α>1p\alpha>\frac{1}{p}. There exists a linear operator

E:Wpα((0,T);)Wpα((0,);)E:W_{p}^{\alpha}\left((0,T);\mathbb{R}\right)\to W_{p}^{\alpha}\left((0,\infty);\mathbb{R}\right)

such that for all bWpα((0,T);)b\in W_{p}^{\alpha}\left((0,T);\mathbb{R}\right), (Eb)|(0,T)=b\left(Eb\right)_{|(0,T)}=b and

EbWpα((0,);)Cp(bWpα((0,T);)+|b(0)|)=Cp|b|Wpα((0,T);)\left\lVert Eb\right\rVert_{W_{p}^{\alpha}\left((0,\infty);\mathbb{R}\right)}\leq C_{p}\left(\left\lVert b\right\rVert_{W_{p}^{\alpha}\left((0,T);\mathbb{R}\right)}+\lvert b(0)\rvert\right)=C_{p}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|b\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{W_{p}^{\alpha}\left((0,T);\mathbb{R}\right)}

with a constant CpC_{p} depending only on pp.

Proof.

In the case b(0)=0b(0)=0 the operator obtained by reflecting the function with respect to the axis t=Tt=T has the desired properties. The general statement can be deduced from this case using surjectivity of the temporal trace :|t=0Wpα((0,);).{}_{|t=0}:W_{p}^{\alpha}\left((0,\infty);\mathbb{R}\right)\to\mathbb{R}\,.

Theorem 2.14 (Uniform embedding I).

Let p(3,)p\in(3,\infty) and T0T_{0} be positive. There exist constants C(p)C(p) and C(T0,p)C\left(T_{0},p\right) such that for all T(0,T0]T\in(0,T_{0}] and all gWp1,2((0,T)×(0,1))g\in W_{p}^{1,2}\left((0,T)\times(0,1)\right),

gC([0,T];C1([0,1]))C(p)gC([0,T];Wp22/p((0,1)))C(T0,p)|g|Wp1,2((0,T)×(0,1)).\left\lVert g\right\rVert_{C\left([0,T];C^{1}\left([0,1]\right)\right)}\leq C(p)\left\lVert g\right\rVert_{C\left([0,T];W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1)\right)\right)}\leq C\left(T_{0},p\right){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|g\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{W_{p}^{1,2}\left((0,T)\times(0,1)\right)}\,.
Proof.

Let T(0,T0]T\in(0,T_{0}] be arbitrary, gWp1,2((0,T)×(0,1))g\in W_{p}^{1,2}\left((0,T)\times(0,1)\right) and 𝑬g\boldsymbol{E}g the extension according to Lemma 2.11. Then 𝑬g\boldsymbol{E}g lies in Wp1,2((0,T0)×(0,1))W_{p}^{1,2}\left(\left(0,T_{0}\right)\times(0,1)\right) and Theorem 2.6 and Lemma 2.11 imply

gC([0,T];Wp22/p((0,1)))\displaystyle\left\lVert g\right\rVert_{C\left([0,T];W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1)\right)\right)} 𝑬gC([0,T0];Wp22/p((0,1)))C(T0,p)𝑬gWp1,2((0,T0)×(0,1))\displaystyle\leq\left\lVert\boldsymbol{E}g\right\rVert_{C\left(\left[0,T_{0}\right];W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1)\right)\right)}\leq C\left(T_{0},p\right)\left\lVert\boldsymbol{E}g\right\rVert_{W_{p}^{1,2}\left(\left(0,T_{0}\right)\times(0,1)\right)}
C(T0,p)|g|Wp1,2((0,T)×(0,1)).\displaystyle\leq C\left(T_{0},p\right){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|g\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{W_{p}^{1,2}\left((0,T)\times(0,1)\right)}\,.

Theorem 2.15 (Uniform embedding II).

Let p(3,)p\in(3,\infty), θ(1+1/p22/p,1)\theta\in\left(\frac{1+\nicefrac{{1}}{{p}}}{2-\nicefrac{{2}}{{p}}},1\right), δ(0,11/p)\delta\in\left(0,1-\nicefrac{{1}}{{p}}\right) and T0T_{0} be positive. There exists a constant C(T0,p,θ,δ)>0C\left(T_{0},p,\theta,\delta\right)>0 such that for all T(0,T0]T\in(0,T_{0}] there holds the embedding

Wp1,2((0,T)×(0,1))C(1θ)(11/pδ)([0,T];C1([0,1]))W_{p}^{1,2}\left((0,T)\times(0,1)\right)\hookrightarrow C^{(1-\theta)\left(1-\nicefrac{{1}}{{p}}-\delta\right)}\left([0,T];C^{1}\left([0,1]\right)\right)

and all gWp1,2((0,T)×(0,1))g\in W_{p}^{1,2}\left((0,T)\times(0,1)\right) satisfy the uniform estimate

gC(1θ)(11/pδ)([0,T];C1([0,1]))C(T0,p,θ,δ)|g|Wp1,2((0,T)×(0,1)).\left\lVert g\right\rVert_{C^{(1-\theta)\left(1-\nicefrac{{1}}{{p}}-\delta\right)}\left([0,T];C^{1}\left([0,1]\right)\right)}\leq C\left(T_{0},p,\theta,\delta\right){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|g\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{W_{p}^{1,2}\left((0,T)\times(0,1)\right)}\,.
Proof.

By [26, Corollary 26] there holds for any δ(0,11/p)\delta\in\left(0,1-\nicefrac{{1}}{{p}}\right) the continuous embedding

Wp1,2((0,T0)×(0,1))C11/pδ([0,T0];Lp((0,1)))W_{p}^{1,2}\left((0,T_{0})\times(0,1)\right)\hookrightarrow C^{1-\nicefrac{{1}}{{p}}-\delta}\left([0,T_{0}];L_{p}((0,1))\right)

with operator norm depending on T0T_{0}. Furthermore, Theorem 2.6 gives

Wp1,2((0,T0)×(0,1))C([0,T0];Wp22/p((0,1))).W_{p}^{1,2}\left((0,T_{0})\times(0,1)\right)\hookrightarrow C\left([0,T_{0}];W_{p}^{2-\nicefrac{{2}}{{p}}}((0,1))\right)\,.

The results in [30] yield that the real interpolation space satisfies

Wpθ(22/p)((0,1))=(Lp((0,1)),Wp22/p((0,1)))θ,pW_{p}^{\theta(2-\nicefrac{{2}}{{p}})}((0,1))=\left(L_{p}((0,1)),W_{p}^{2-\nicefrac{{2}}{{p}}}((0,1))\right)_{\theta,p}

with equivalent norms. In particular, for all fWpθ(22/p)((0,1))f\in W_{p}^{\theta(2-\nicefrac{{2}}{{p}})}((0,1)) there holds the estimate

fWpθ(22/p)((0,1))CfLp((0,1))1θfWp22/p((0,1))θ.\left\lVert f\right\rVert_{W_{p}^{\theta(2-\nicefrac{{2}}{{p}})}((0,1))}\leq C\left\lVert f\right\rVert^{1-\theta}_{L_{p}((0,1))}\left\lVert f\right\rVert^{\theta}_{W_{p}^{2-\nicefrac{{2}}{{p}}}((0,1))}\,.

A direct computation using the above estimate shows that for all α(0,1)\alpha\in(0,1),

C([0,T0];Wp22/p((0,1)))Cα([0,T0];Lp((0,1)))C(1θ)α([0,T0];Wpθ(22/p)((0,1)))C\left([0,T_{0}];W_{p}^{2-\nicefrac{{2}}{{p}}}((0,1))\right)\cap C^{\alpha}\left([0,T_{0}];L_{p}((0,1))\right)\hookrightarrow C^{(1-\theta)\alpha}\left([0,T_{0}];W_{p}^{\theta(2-\nicefrac{{2}}{{p}})}((0,1))\right)

which yields for all δ(0,11/p)\delta\in\left(0,1-\nicefrac{{1}}{{p}}\right) the continuous embedding

Wp1,2((0,T0)×(0,1))C(1θ)(11/pδ)([0,T0];Wpθ(22/p)((0,1))).W_{p}^{1,2}\left((0,T_{0})\times(0,1)\right)\hookrightarrow C^{(1-\theta)(1-\nicefrac{{1}}{{p}}-\delta)}\left([0,T_{0}];W_{p}^{\theta(2-\nicefrac{{2}}{{p}})}((0,1))\right)\,.

Due to θ(22/p)1p>1\theta\left(2-\nicefrac{{2}}{{p}}\right)-\frac{1}{p}>1 the Sobolev Embedding Theorem yields

Wp1,2((0,T0)×(0,1))C(1θ)(11/pδ)([0,T0];C1([0,1])).W_{p}^{1,2}\left((0,T_{0})\times(0,1)\right)\hookrightarrow C^{(1-\theta)(1-\nicefrac{{1}}{{p}}-\delta)}\left([0,T_{0}];C^{1}([0,1])\right)\,.

The claim now follows using the extension operator 𝑬\boldsymbol{E} constructed in Lemma 2.11 with similar arguments as in the proof of Theorem 2.14. ∎

2.2 Motion by curvature of networks

Let n,n2n\in\mathbb{N},n\geq 2. Consider a curve σ:[0,1]n\sigma:[0,1]\to\mathbb{R}^{n} of class C1C^{1}. A curve is said to be regular if |σx(x)|0|\sigma_{x}(x)|\neq 0 for every x[0,1]x\in[0,1]. Let us denote with ss the arclength parameter. We remind that s=x|σx|\partial_{s}=\frac{\partial_{x}}{|\sigma_{x}|}. If a curve σ\sigma is of class C1C^{1} and regular, its unit tangent vector is given by τ=σs=σx|σx|{\tau}=\sigma_{s}=\frac{\sigma_{x}}{\left|\sigma_{x}\right|}. The curvature vector of a regular C2C^{2}–curve σ\sigma is defined by

𝜿:=σss=τs=σxx|σx|2σxx,σxσx|σx|4.\boldsymbol{\kappa}:=\sigma_{ss}=\tau_{s}=\frac{\sigma_{xx}}{\left|\sigma_{x}\right|^{2}}-\frac{\left\langle\sigma_{xx},\sigma_{x}\right\rangle\sigma_{x}}{\left|\sigma_{x}\right|^{4}}\,.

The curvature is given by κ=|τs|\kappa=|\tau_{s}|.

Definition 2.16.

A network 𝒩\mathcal{N} is a connected set in n\mathbb{R}^{n} consisting of a finite union of regular curves 𝒩i\mathcal{N}^{i} that meet at their endpoints in junctions. Each curve 𝒩i\mathcal{N}^{i} admits a regular C1C^{1}–parametrisation, namely a map γi:[0,1]n\gamma^{i}:[0,1]\to\mathbb{R}^{n} of class C1C^{1} with |γxi|0|\gamma^{i}_{x}|\neq 0 on [0,1][0,1] and γi([0,1])=𝒩i\gamma^{i}\left([0,1]\right)=\mathcal{N}^{i}.

Although a network is a set by definition, we will mainly deal with its parametrisations. It is then natural to speak about the regularity of these maps.

Definition 2.17.

Let kk\in\mathbb{N}, k2k\geq 2, and 1p1\leq p\leq\infty with p>1k1p>\frac{1}{k-1}. A network 𝒩\mathcal{N} is of class CkC^{k} (or WpkW_{p}^{k}, respectively) if it admits a regular parametrisation of class CkC^{k} (or WpkW_{p}^{k}, respectively).

In this paper we restrict to the class of regular networks.

Definition 2.18.

A network is called regular if its curves meet at triple junctions forming equal angles.

Notice that this notion is geometric in the sense that it does not depend on the choice of the parametrisations σi\sigma^{i} of the curves of the network 𝒩\mathcal{N}.

We define now the motion by curvature of regular networks: a time dependent family of regular networks evolves with normal velocity equal to the curvature vector at any point and any time, namely

Vi=𝜿i.V^{i}=\boldsymbol{\kappa}^{i}\,.

To be more precise, given a time dependent family of curves γi\gamma^{i}, we denote by 𝑷i:nn\boldsymbol{P}^{i}:\mathbb{R}^{n}\to\mathbb{R}^{n} the projection onto the normal space to γi\gamma^{i}, namely 𝑷i:=Idγsγs\boldsymbol{P}^{i}:=\mathrm{Id}-\gamma_{s}\otimes\gamma_{s}. The motion equation reads as

𝑷iγi=𝜿i.\boldsymbol{P}^{i}\gamma^{i}=\boldsymbol{\kappa}^{i}\,.

For the sake of presentation we restrict to the motion by curvature of a Triod.

Definition 2.19.

A Triod 𝕋=i=13σ([0,1])i\mathbb{T}=\bigcup_{i=1}^{3}\sigma{}^{i}([0,1]) is a network composed of three regular C1C^{1}–curves σi:[0,1]n\sigma^{i}:\left[0,1\right]\to\mathbb{R}^{n} that intersect each other at the triple junction O:=σ1(0)=σ2(0)=σ3(0)O:=\sigma^{1}(0)=\sigma^{2}(0)=\sigma^{3}(0). The other three endpoints of the curves σi(1)\sigma^{i}(1) with i{1,2,3}i\in\{1,2,3\} coincide with three points Pi:=σi(1)nP^{i}:=\sigma^{i}\left(1\right)\in\mathbb{R}^{n}. The Triod is called regular if it is a regular network.

P1P^{1}σ1\sigma^{1}σ3\sigma^{3}σ2\sigma^{2}OOP3\,\,\,\,\,\,P^{3}P2\,\,\,\,P^{2}
Figure 1: A regular Triod in 2\mathbb{R}^{2}.
Definition 2.20 (Geometrically admissible initial Triod).

A Triod 𝕋0\mathbb{T}_{0} is a geometrically admissible initial datum for the motion by curvature if it is regular and each of its curves can be parametrised by a regular curve σiWp22/p([0,1],n)\sigma^{i}\in W^{2-2/p}_{p}([0,1],\mathbb{R}^{n}) with p(3,)p\in(3,\infty).

Remark 2.21.

For p(3,)p\in(3,\infty) the Sobolev Embedding Theorem [30, Theorem 4.6.1.(e)] implies

Wp22/p((0,1);n)C1+α([0,1];n)W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1);\mathbb{R}^{n}\right)\hookrightarrow C^{1+\alpha}\left([0,1];\mathbb{R}^{n}\right)

for α(0,13/p)\alpha\in\left(0,1-\nicefrac{{3}}{{p}}\right). In particular, any admissible initial network is of class C1C^{1} and the angle condition at the boundary is well-defined.

Definition 2.22 (Solutions to the motion by curvature).

Let p(3,)p\in(3,\infty) and T>0T>0. Let 𝕋0\mathbb{T}_{0} be a geometrically admissible initial Triod with endpoints P1P^{1}, P2P^{2}, P3P^{3}. A time dependent family of Triods (𝕋(t))\left(\mathbb{T}(t)\right) is a solution to the motion by curvature in [0,T][0,T] with initial datum 𝕋0\mathbb{T}_{0} if there exists a collection of time dependent parametrisations

γniWp1(In;Lp((0,1);n))Lp(In;Wp2((0,1);n)),\gamma^{i}_{n}\in W^{1}_{p}(I_{n};L_{p}((0,1);\mathbb{R}^{n}))\cap L_{p}(I_{n};W^{2}_{p}((0,1);\mathbb{R}^{n}))\,,

with n{0,,N}n\in\{0,\dots,N\} for some NN\in\mathbb{N}, In:=(an,bn)I_{n}:=(a_{n},b_{n})\subset\mathbb{R}, anan+1a_{n}\leq a_{n+1}, bnbn+1b_{n}\leq b_{n+1}, an<bna_{n}<b_{n} and n(an,bn)=(0,T)\bigcup_{n}(a_{n},b_{n})=(0,T) such that for all n{0,,N}n\in\{0,\dots,N\} and tInt\in I_{n}, γn(t)=(γ1(t),γ2(t),γ3(t))\gamma_{n}(t)=\left(\gamma^{1}(t),\gamma^{2}(t),\gamma^{3}(t)\right) is a regular parametrisation of 𝕋(t)\mathbb{T}(t). Moreover, each γn\gamma_{n} needs to satisfy the following system:

{𝑷iγti(t,x)=𝜿i(t,x) motion by curvature,γi(t,1)=Pi fixed endpoints,γ1(t,0)=γ2(t,0)=γ3(t,0) concurrency condition,i=13τi(t,0)=0 angle condition,\begin{cases}\begin{array}[]{lll}\boldsymbol{P}^{i}\gamma^{i}_{t}(t,x)=\boldsymbol{\kappa}^{i}(t,x)&\text{ motion by curvature,}\\ \gamma^{i}(t,1)=P^{i}&\text{ fixed endpoints,}\\ \gamma^{1}(t,0)=\gamma^{2}(t,0)=\gamma^{3}(t,0)&\text{ concurrency condition,}\\ \sum_{i=1}^{3}\tau^{i}(t,0)=0&\text{ angle condition,}\end{array}\end{cases} (2.2)

for almost every tIn,x(0,1)t\in I_{n},x\in\left(0,1\right) and for i{1,2,3}i\in\{1,2,3\}. Finally, we ask that γn(0,[0,1])=𝕋0\gamma_{n}(0,[0,1])=\mathbb{T}_{0} whenever an=0a_{n}=0.

Remark 2.23.

In the motion by curvature equation only the normal component of the velocity γti\gamma^{i}_{t} is prescribed. This does not mean that there is no tangential motion. Indeed, a non–trivial tangential velocity is generally needed to allow for motion of the triple junction.

Remark 2.24.

We are interested in finding a time–dependent family of networks (𝒩(t))\left(\mathcal{N}(t)\right) solving the motion by curvature. Our notion of solution allows the network to be parametrised by different sets of functions in different (but overlapping) time intervals. Namely a solution can be parametrised by γ=(γ1,γ2,γ3)\gamma=(\gamma^{1},\gamma^{2},\gamma^{3}) with γi:(a0,b0)×[0,1]n\gamma^{i}:(a_{0},b_{0})\times[0,1]\to\mathbb{R}^{n} and η=(η1,η2,η3)\eta=(\eta^{1},\eta^{2},\eta^{3}) with ηi:(a1,b1)×[0,1]n\eta^{i}:(a_{1},b_{1})\times[0,1]\to\mathbb{R}^{n} if a0a1<b0b1a_{0}\leq a_{1}<b_{0}\leq b_{1} and γi((a1,b0)×[0,1])=ηi((a1,b0)×[0,1])\gamma^{i}((a_{1},b_{0})\times[0,1])=\eta^{i}((a_{1},b_{0})\times[0,1]). Requiring that the family of networks (𝒩(t))\left(\mathcal{N}(t)\right) is parametrised by one map γ(t)=(γ1(t),γ2(t),γ3(t))\gamma(t)=(\gamma^{1}(t),\gamma^{2}(t),\gamma^{3}(t)) in the whole time interval of existence [0,T][0,T] as in [22] gives a slightly stronger definition of the motion by curvature in comparison to Definition 2.22. This difference does not affect the proof of the short time existence result, but in principle using our definition the maximal time interval of existence could be longer.

The first step to find solutions to the motion by curvature is to turn system (2.2) into a system of quasilinear parabolic PDEs by choosing a suitable tangential velocity TT. We choose TT such that

γti(t,x)=𝑷iγti(t,x)+γti(t,x),τi(t,x)τi(t,x)=𝜿i(t,x)+Ti(t,x)τi(t,x)=γxxi(t,x)|γxi(t,x)|2.\gamma^{i}_{t}(t,x)=\boldsymbol{P}^{i}\gamma^{i}_{t}(t,x)+\left\langle\gamma^{i}_{t}(t,x)\,,\tau^{i}(t,x)\right\rangle\tau^{i}(t,x)=\boldsymbol{\kappa}^{i}(t,x)+T^{i}(t,x)\tau^{i}(t,x)=\frac{\gamma^{i}_{xx}(t,x)}{|\gamma^{i}_{x}(t,x)|^{2}}\,. (2.3)

Since the expression of the curvature reads as

𝜿i(t,x)=γxxi(t,x)|γxi(t,x)|2γxxi(t,x)|γxi(t,x)|2,τi(t,x)τi(t,x)\boldsymbol{\kappa}^{i}(t,x)=\frac{\gamma^{i}_{xx}(t,x)}{|\gamma^{i}_{x}(t,x)|^{2}}-\left\langle\frac{\gamma^{i}_{xx}(t,x)}{|\gamma^{i}_{x}(t,x)|^{2}}\,,\tau^{i}(t,x)\right\rangle\tau^{i}(t,x)

we choose

Ti(t,x)=γxxi(t,x)|γxi(t,x)|2,τi(t,x).T^{i}(t,x)=\left\langle\frac{\gamma^{i}_{xx}(t,x)}{|\gamma^{i}_{x}(t,x)|^{2}}\,,\tau^{i}(t,x)\right\rangle\,.

The equation γti=γxxi|γxi|2\gamma^{i}_{t}=\frac{\gamma^{i}_{xx}}{|\gamma^{i}_{x}|^{2}} is called Special Flow.

Definition 2.25 (Admissible initial parametrisation).

Let p(3,)p\in(3,\infty). An admissible initial parametrisation for a Triod 𝕋0\mathbb{T}_{0} is a triple σ=(σ1,σ2,σ3)\sigma=(\sigma^{1},\sigma^{2},\sigma^{3}) where iσi([0,1])=𝕋0\bigcup_{i}\sigma^{i}([0,1])=\mathbb{T}_{0}, σ1(0)=σ2(0)=σ3(0)\sigma^{1}(0)=\sigma^{2}(0)=\sigma^{3}(0) and i=13σxi(0)|σxi(0)|=0\sum_{i=1}^{3}\frac{\sigma^{i}_{x}(0)}{|\sigma^{i}_{x}(0)|}=0 with σi\sigma^{i} regular and of class Wp22/p((0,1),n)W^{2-2/p}_{p}((0,1),\mathbb{R}^{n}).

Notice that it follows by the very definition that a geometrically admissible Triod admits an admissible parametrisation.

Definition 2.26 (Solution of the Special Flow).

Let T>0T>0 and p(3,)p\in(3,\infty). Consider an admissible initial parametrisation σ=(σ1,σ2,σ3)\sigma=(\sigma^{1},\sigma^{2},\sigma^{3}) for a Triod 𝕋0\mathbb{T}_{0} in n\mathbb{R}^{n} with σi(1)=Pin\sigma^{i}(1)=P^{i}\in\mathbb{R}^{n}. Then we say that γ=(γ1,γ2,γ3)\gamma=(\gamma^{1},\gamma^{2},\gamma^{3}) is a solution of the Special Flow in the time interval [0,T][0,T] with initial datum σ\sigma if

γ=(γ1,γ2,γ3)𝑬T=Wp1((0,T);Lp((0,1);(n)3))Lp((0,T);Wp2((0,1);(n)3)),\gamma=\left(\gamma^{1},\gamma^{2},\gamma^{3}\right)\in\boldsymbol{E}_{T}=W^{1}_{p}((0,T);L_{p}((0,1);(\mathbb{R}^{n})^{3}))\cap L_{p}((0,T);W^{2}_{p}((0,1);(\mathbb{R}^{n})^{3}))\,,

|γxi(t,x)|0|\gamma^{i}_{x}(t,x)|\neq 0 for all (t,x)[0,T]×[0,1](t,x)\in[0,T]\times[0,1] and the following system is satisfied for i{1,2,3}i\in\{1,2,3\} and for almost every x(0,1)x\in(0,1), t(0,T)t\in(0,T):

{γti(t,x)=γxxi(t,x)|γxi(t,x)|2Special Flow,γi(t,1)=Pifixed endpoints,γ1(t,0)=γ2(t,0)=γ3(t,0)concurrency condition,i=13γxi(t,0)|γxi(t,0)|=0angle condition,γi(0,x)=σi(x)initial datum.\begin{cases}\begin{array}[]{lll}\gamma^{i}_{t}(t,x)=\frac{\gamma_{xx}^{i}\left(t,x\right)}{\left|\gamma_{x}^{i}\left(t,x\right)\right|^{2}}&\text{Special Flow,}\\ \gamma^{i}(t,1)=P^{i}&\text{fixed endpoints,}\\ \gamma^{1}(t,0)=\gamma^{2}(t,0)=\gamma^{3}(t,0)&\text{concurrency condition,}\\ \sum_{i=1}^{3}\frac{\gamma_{x}^{i}\left(t,0\right)}{\left|\gamma_{x}^{i}\left(t,0\right)\right|}=0&\text{angle condition,}\\ \gamma^{i}(0,x)=\sigma^{i}(x)&\text{initial datum.}\\ \end{array}\end{cases} (2.4)
Remark 2.27.

Both in [6] and in [23] the authors define the motion by curvature introducing directly the Special Flow. This is not restrictive to get a short time existence result because a solution of the Special Flow as defined in Definition 2.26 induces a solution of the motion by curvature in the sense of Definition 2.22 which is shown in Theorem 3.14 below. However, we will see that it is not trivial to deduce geometric uniqueness of solutions to the motion by curvature from uniqueness of solutions to the Special Flow.

3 Existence and Uniqueness of the Motion by Curvature

3.1 Existence and uniqueness of the linearised Special Flow

We fix an admissible initial parametrisation σ=(σ1,σ2,σ3)\sigma=(\sigma^{1},\sigma^{2},\sigma^{3}). Linearising the main equation of system (2.4) around the initial datum we obtain:

γti(t,x)1|σxi(x)|2γxxi(t,x)=(1|γxi(t,x)|21|σxi(x)|2)γxxi(t,x).\gamma^{i}_{t}(t,x)-\frac{1}{\left|\sigma^{i}_{x}(x)\right|^{2}}\,\gamma^{i}_{xx}(t,x)=\left(\frac{1}{\left|\gamma^{i}_{x}(t,x)\right|^{2}}-\frac{1}{\left|\sigma^{i}_{x}(x)\right|^{2}}\right)\gamma^{i}_{xx}(t,x)\,. (3.1)

The linearisation of the angle condition in x=0x=0 is given by

i=13(γxi|σxi|σxiγxi,σxi|σxi|3)=i=13((1|γxi|1|σxi|)γxi+σxiγxi,σxi|σxi|3),-\sum_{i=1}^{3}\left(\frac{\gamma^{i}_{x}}{|\sigma^{i}_{x}|}-\frac{\sigma^{i}_{x}\left\langle\gamma^{i}_{x},\sigma^{i}_{x}\right\rangle}{|\sigma^{i}_{x}|^{3}}\right)=\sum_{i=1}^{3}\left(\left(\frac{1}{|\gamma^{i}_{x}|}-\frac{1}{|\sigma^{i}_{x}|}\right)\gamma^{i}_{x}+\frac{\sigma^{i}_{x}\left\langle\gamma^{i}_{x},\sigma^{i}_{x}\right\rangle}{|\sigma^{i}_{x}|^{3}}\right)\,, (3.2)

where we have omitted the dependence of the left-hand side on (t,0)(t,0). The concurrency and the fixed endpoints conditions are already linear and affine. We obtain the following linearised system for a general right hand side (f,η,b,ψ)(f,\eta,b,\psi).

{γti(t,x)1|σxi(x)|2γxxi(t,x)=fi(t,x),t(0,T),x(0,1),i{1,2,3},γ(t,1)=η(t),t[0,T],γ1(t,0)γ2(t,0)=0,t[0,T],γ2(t,0)γ3(t,0)=0,t[0,T],i=13(γxi(t,0)|σxi(0)|σxi(0)γxi(t,0),σxi(0)|σxi(0)|3)=b(t),t[0,T],γ(0,x)=ψ(x),x[0,1].\begin{cases}\begin{array}[]{rll}\gamma^{i}_{t}(t,x)-\frac{1}{\left|\sigma^{i}_{x}(x)\right|^{2}}\,\gamma^{i}_{xx}(t,x)&=f^{i}(t,x)\,,&t\in(0,T)\,,x\in(0,1)\,,i\in\{1,2,3\}\,,\\ \gamma(t,1)&=\eta(t)\,,&t\in[0,T]\,,\\ \gamma^{1}\left(t,0\right)-\gamma^{2}\left(t,0\right)&=0\,,&t\in[0,T]\,,\\ \gamma^{2}(t,0)-\gamma^{3}\left(t,0\right)&=0\,,&t\in[0,T]\,,\\ -\sum_{i=1}^{3}\left(\frac{\gamma^{i}_{x}(t,0)}{|\sigma^{i}_{x}(0)|}-\frac{\sigma^{i}_{x}(0)\left\langle\gamma^{i}_{x}(t,0),\sigma^{i}_{x}(0)\right\rangle}{|\sigma^{i}_{x}(0)|^{3}}\right)&=b(t)\,,&t\in[0,T]\,,\\ \gamma\left(0,x\right)&=\psi\left(x\right)\,,&x\in[0,1]\,.\end{array}\end{cases} (3.3)
Definition 3.1 (Linear compatibility conditions).

Let p(3,)p\in(3,\infty). A function ψ=(ψ1,ψ2,ψ3)\psi=(\psi^{1},\psi^{2},\psi^{3}) of class Wp22/p((0,1);(n)3)W^{2-\nicefrac{{2}}{{p}}}_{p}\left((0,1);(\mathbb{R}^{n})^{3}\right) satisfies the linear compatibility conditions for system (3.3) with respect to given functions ηWp11/2p((0,T);(n)3)\eta\in W_{p}^{1-\nicefrac{{1}}{{2p}}}((0,T);(\mathbb{R}^{n})^{3}), bWp1/21/2p((0,T);n)b\in W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}((0,T);\mathbb{R}^{n}) if for i,j{1,2,3}i,j\in\{1,2,3\} it holds ψi(0)=ψj(0)\psi^{i}(0)=\psi^{j}(0), ψi(1)=ηi(0)\psi^{i}(1)=\eta^{i}(0) and

i=13(ψxi(0)|σxi(0)|σxi(0)ψxi(0),σxi(0)|σxi(0)|3)=b(0).-\sum_{i=1}^{3}\left(\frac{\psi^{i}_{x}(0)}{|\sigma^{i}_{x}(0)|}-\frac{\sigma^{i}_{x}(0)\left\langle\psi^{i}_{x}(0),\sigma^{i}_{x}(0)\right\rangle}{|\sigma^{i}_{x}(0)|^{3}}\right)=b(0)\,.

We want to show that system (3.3) admits a unique solution γ=(γ1,γ2,γ3)\gamma=(\gamma^{1},\gamma^{2},\gamma^{3}) in 𝑬T\boldsymbol{E}_{T}. The result follows from the classical theory for linear parabolic systems by Solonnikov [27] provided that the system is parabolic and that the complementary conditions hold (see [27, p. 11]). Both the parabolicity and the complementary (initial and boundary) conditions have been proven in [6]. We remark that the complementary conditions at the boundary follow from the Lopatinskii–Shapiro condition (see for instance [8, pages 11–15]).

Definition 3.2 (Lopatinskii–Shapiro condition).

Let λ\lambda\in\mathbb{C} with (λ)>0\Re(\lambda)>0 be arbitrary. The Lopatinskii–Shapiro condition for system (3.3) is satisfied at the triple junction if every solution ϱ=(ϱ1,ϱ2,ϱ3)C2([0,),(2)3)\varrho=(\varrho^{1},\varrho^{2},\varrho^{3})\in C^{2}([0,\infty),(\mathbb{C}^{2})^{3}) to

{λϱi(x)1|σxi(0)|2ϱxxi(x)=0,x[0,),i{1,2,3},ϱ1(0)ϱ2(0)=0,ϱ2(0)ϱ3(0)=0,i=13ϱxi(x)|σxi(0)|σxi(0)ϱxi(x),σxi(0)|σxi(0)|3=0\begin{cases}\begin{array}[]{rll}\lambda\varrho^{i}(x)-\frac{1}{|\sigma^{i}_{x}(0)|^{2}}\varrho^{i}_{xx}(x)&=0\,,&\;\;x\in[0,\infty)\,,\;i\in\{1,2,3\}\,,\\ \varrho^{1}(0)-\varrho^{2}(0)&=0\,,&\\ \varrho^{2}(0)-\varrho^{3}(0)&=0\,,&\\ \sum_{i=1}^{3}\frac{\varrho_{x}^{i}(x)}{\left|\sigma_{x}^{i}(0)\right|}-\frac{\sigma^{i}_{x}(0)\left\langle\varrho^{i}_{x}(x),\sigma^{i}_{x}(0)\right\rangle}{|\sigma^{i}_{x}(0)|^{3}}&=0&\end{array}\end{cases} (3.4)

which satisfies limx|ϱi(x)|=0\lim_{x\to\infty}\lvert\varrho^{i}(x)\rvert=0 is the trivial solution.

Similarly, the Lopatinskii–Shapiro condition for system (3.3) is satisfied at the fixed endpoints if every solution ϱ=(ϱ1,ϱ2,ϱ3)C2([0,),(2)3)\varrho=(\varrho^{1},\varrho^{2},\varrho^{3})\in C^{2}([0,\infty),(\mathbb{C}^{2})^{3}) to

{λϱi(x)1|σxi(0)|2ϱxxi(x)=0,x[0,),i{1,2,3},ϱi(0)=0,i{1,2,3}\begin{cases}\begin{array}[]{rll}\lambda\varrho^{i}(x)-\frac{1}{|\sigma^{i}_{x}(0)|^{2}}\varrho^{i}_{xx}(x)&=0\,,&\;x\in[0,\infty)\,,\;i\in\{1,2,3\}\,,\\ \varrho^{i}(0)&=0\,,&\;i\in\{1,2,3\}\end{array}\end{cases} (3.5)

which satisfies limx|ϱi(x)|=0\lim_{x\to\infty}\lvert\varrho^{i}(x)\rvert=0 is the trivial solution.

Lemma 3.3.

The Lopatinskii–Shapiro condition is satisfied.

Proof.

We first check the condition at the triple junction. Let ϱ\varrho be a solution to (3.4) satisfying limx|ϱi(x)|=0\lim_{x\to\infty}\lvert\varrho^{i}(x)\rvert=0. We multiply λϱi(x)1|σxi(0)|2ϱxxi(x)=0\lambda\varrho^{i}(x)-\frac{1}{|\sigma^{i}_{x}(0)|^{2}}\varrho^{i}_{xx}(x)=0 by |σxi(0)|𝑷iϱ¯i(x)|\sigma^{i}_{x}(0)|\boldsymbol{P}^{i}\overline{\varrho}^{i}(x) (where with 𝑷i\boldsymbol{P}^{i} we mean here Idσsi(0)σsi(0)\mathrm{Id}-\sigma^{i}_{s}(0)\otimes\sigma^{i}_{s}(0)), then we integrate and sum. Using the two conditions at the boundary we get

0\displaystyle 0 =i=130λ|σxi(0)||𝑷i(ϱi(x))|21|σxi(0)|ϱxxi(x),𝑷iϱ¯i(x)dx\displaystyle=\sum_{i=1}^{3}\int_{0}^{\infty}\lambda|\sigma^{i}_{x}(0)||\boldsymbol{P}^{i}(\varrho^{i}(x))|^{2}-\frac{1}{|\sigma^{i}_{x}(0)|}\left\langle\varrho^{i}_{xx}(x),\boldsymbol{P}^{i}\overline{\varrho}^{i}(x)\right\rangle\,\mathrm{d}x (3.6)
=i=130λ|σxi(0)||𝑷i(ϱi(x))|2+|𝑷i(ϱxi(x))|2|σxi(0)|dxi=131|σxi(0)|𝑷iϱxi(0),𝑷iϱ¯i(0)\displaystyle=\sum_{i=1}^{3}\int_{0}^{\infty}\lambda|\sigma^{i}_{x}(0)||\boldsymbol{P}^{i}(\varrho^{i}(x))|^{2}+\frac{|\boldsymbol{P}^{i}(\varrho^{i}_{x}(x))|^{2}}{|\sigma^{i}_{x}(0)|}\,\mathrm{d}x-\sum_{i=1}^{3}\frac{1}{|\sigma^{i}_{x}(0)|}\left\langle\boldsymbol{P}^{i}\varrho^{i}_{x}(0),\boldsymbol{P}^{i}\overline{\varrho}^{i}(0)\right\rangle (3.7)
=i=130λ|σxi(0)||𝑷i(ϱi(x))|2+|𝑷i(ϱxi(x))|2|σxi(0)|dxϱ¯1(0),i=13𝑷i(ϱxi(0)|σxi(0)|)\displaystyle=\sum_{i=1}^{3}\int_{0}^{\infty}\lambda|\sigma^{i}_{x}(0)||\boldsymbol{P}^{i}(\varrho^{i}(x))|^{2}+\frac{|\boldsymbol{P}^{i}(\varrho^{i}_{x}(x))|^{2}}{|\sigma^{i}_{x}(0)|}\,\mathrm{d}x-\left\langle\overline{\varrho}^{1}(0),\sum_{i=1}^{3}\boldsymbol{P}^{i}\left(\frac{\varrho^{i}_{x}(0)}{|\sigma^{i}_{x}(0)|}\right)\right\rangle (3.8)
=i=130λ|σxi(0)||𝑷i(ϱi(x))|2+|𝑷i(ϱxi(x))|2|σxi(0)|dx.\displaystyle=\sum_{i=1}^{3}\int_{0}^{\infty}\lambda|\sigma^{i}_{x}(0)||\boldsymbol{P}^{i}(\varrho^{i}(x))|^{2}+\frac{|\boldsymbol{P}^{i}(\varrho^{i}_{x}(x))|^{2}}{|\sigma^{i}_{x}(0)|}\,\mathrm{d}x\,. (3.9)

As a consequence we get that 𝑷i(ϱi(x))=0\boldsymbol{P}^{i}(\varrho^{i}(x))=0 for all x[0,)x\in[0,\infty) and i{1,2,3}i\in\{1,2,3\} and in particular 𝑷i(ϱ1(0))=0\boldsymbol{P}^{i}(\varrho^{1}(0))=0 for all i{1,2,3}i\in\{1,2,3\}. As the orthogonal complements of σxi(0)\sigma_{x}^{i}(0) with i{1,2,3}i\in\{1,2,3\} span all n\mathbb{R}^{n}, we conclude that ϱi(0)=0\varrho^{i}(0)=0 for all i{1,2,3}i\in\{1,2,3\}. Repeating the argument and testing the motion equation by |σxi(0)|ϱ¯i(x),σsi(0)σsi(0)|\sigma^{i}_{x}(0)|\langle\overline{\varrho}^{i}(x),\sigma^{i}_{s}(0)\rangle\sigma^{i}_{s}(0) we can conclude that ϱi(x)=0\varrho^{i}(x)=0 for every x[0,)x\in[0,\infty). Indeed, we obtain

i=13λ|σxi(0)|0|ϱi(x),σsi(0)|2dx+i=131|σxi(0)|0|ϱxi(x),σsi(0)|2dx\displaystyle\sum_{i=1}^{3}\lambda|\sigma^{i}_{x}(0)|\int_{0}^{\infty}|\left\langle\varrho^{i}(x),\sigma^{i}_{s}(0)\right\rangle|^{2}\,\mathrm{d}x+\sum_{i=1}^{3}\frac{1}{|\sigma^{i}_{x}(0)|}\int_{0}^{\infty}|\left\langle\varrho^{i}_{x}(x),\sigma^{i}_{s}(0)\right\rangle|^{2}\,\mathrm{d}x
+\displaystyle+ i=131|σxi(0)|ϱ¯i(0),σsi(0)ϱxi(0),σsi(0)=0.\displaystyle\sum_{i=1}^{3}\frac{1}{|\sigma^{i}_{x}(0)|}\left\langle\overline{\varrho}^{i}(0),\sigma^{i}_{s}(0)\right\rangle\left\langle\varrho^{i}_{x}(0),\sigma^{i}_{s}(0)\right\rangle=0\,. (3.10)

This time the boundary condition vanishes since we get ϱi(0)=0\varrho^{i}(0)=0 from the previous step. Taking again the real part of (3.1) we can conclude that ϱi(x),σsi(0)=0\left\langle\varrho^{i}(x),\sigma^{i}_{s}(0)\right\rangle=0 for all x[0,)x\in[0,\infty). Hence ϱi(x)=0\varrho^{i}(x)=0 for every x[0,)x\in[0,\infty) as desired.

The condition at the fixed endpoints follows in exactly the same way using the boundary condition ϱi(0)=0\varrho^{i}(0)=0. ∎

Given T>0T>0 we introduce the spaces

  • 𝔼T:={γ𝑬T,γ1(t,0)=γ2(t,0)=γ3(t,0) fori{1,2,3},t[0,T]}\mathbb{E}_{T}:=\left\{\gamma\in\boldsymbol{E}_{T}\,,\gamma^{1}(t,0)=\gamma^{2}(t,0)=\gamma^{3}(t,0)\text{ for}\;i\in\{1,2,3\},t\in[0,T]\,\right\},

  • 𝔽T:={(f,η,0,b,ψ)withfLp((0,T);Lp((0,1);(n)3)),ηWp11/2p((0,T);(n)3),  0Wp11/2p((0,T);2n),bWp1/21/2p((0,T);n),ψWp22/p((0,1);(n)3) such that the linear compatibility conditions in Definition 3.1 hold}\mathbb{F}_{T}:=\left\{(f,\eta,0,b,\psi)\;\text{with}\,f\in L_{p}((0,T);L_{p}((0,1);(\mathbb{R}^{n})^{3})),\,\eta\in W^{1-\nicefrac{{1}}{{2p}}}_{p}((0,T);(\mathbb{R}^{n})^{3})\,,\right.\\ \left.\qquad\quad\;\,0\in W^{1-\nicefrac{{1}}{{2p}}}_{p}((0,T);\mathbb{R}^{2n})\,,b\in W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}((0,T);\mathbb{R}^{n})\,,\,\psi\in W_{p}^{2-2/p}((0,1);(\mathbb{R}^{n})^{3})\right.\\ \left.\qquad\quad\;\text{ such that the linear compatibility conditions in Definition~\ref{linearcompcond} hold}\right\}.

Theorem 3.4.

Let p(3,)p\in(3,\infty). For every T>0T>0 system (3.3) has a unique solution γ𝔼T\gamma\in\mathbb{E}_{T} provided that fLp((0,T);Lp((0,1);(n)3))f\in L_{p}((0,T);L_{p}((0,1);(\mathbb{R}^{n})^{3})), ηWp11/2p((0,T);(n)3)\eta\in W^{1-\nicefrac{{1}}{{2p}}}_{p}((0,T);(\mathbb{R}^{n})^{3}) bWp1/21/2p((0,T);n)b\in W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}((0,T);\mathbb{R}^{n}) and ψWp22/p((0,1);(n)3)\psi\in W_{p}^{2-\nicefrac{{2}}{{p}}}((0,1);(\mathbb{R}^{n})^{3}) fulfil the linear compatibility conditions given in Definition 3.1. Moreover, there exists a constant C=C(T)>0C=C(T)>0 such that the following estimate holds:

γ𝑬TC(fLp((0,T);Lp((0,1)))+ηWp11/2p((0,T))+bWp1/21/2p((0,T))+ψWp22/p((0,1))).\|\gamma\|_{\boldsymbol{E}_{T}}\leq C\left(\|f\|_{L_{p}((0,T);L_{p}((0,1)))}+\|\eta\|_{W^{1-\nicefrac{{1}}{{2p}}}_{p}((0,T))}+\|b\|_{W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}((0,T))}+\|\psi\|_{W_{p}^{2-\nicefrac{{2}}{{p}}}((0,1))}\right)\,. (3.11)
Proof.

This follows from [27, Theorem 5.4]. ∎

Theorem 3.4 implies in particular that the linear operator LT:𝔼T𝔽TL_{T}:\mathbb{E}_{T}\to\mathbb{F}_{T} defined by

LT(γ)=((γtiγxxi|σxi|2)i{1,2,3}γ|x=1(γ|x=01γ|x=02,γ|x=02γ|x=03)i=13(γxi|σxi|σxiγxi,σxi|σxi|3)|x=0γ|t=0)L_{T}(\gamma)=\begin{pmatrix}\left(\gamma^{i}_{t}-\frac{\gamma^{i}_{xx}}{|\sigma^{i}_{x}|^{2}}\right)_{i\in\{1,2,3\}}\\ \gamma_{|x=1}\\ \left(\gamma^{1}_{|x=0}-\gamma^{2}_{|x=0},\gamma^{2}_{|x=0}-\gamma^{3}_{|x=0}\right)\\ -\sum_{i=1}^{3}\left(\frac{\gamma^{i}_{x}}{|\sigma^{i}_{x}|}-\frac{\sigma^{i}_{x}\left\langle\gamma^{i}_{x},\sigma^{i}_{x}\right\rangle}{|\sigma^{i}_{x}|^{3}}\right)_{|x=0}\\ \gamma_{|t=0}\end{pmatrix}

is a continuous isomorphism.

Corollary 2.10 and Lemma 2.12 imply that for every positive TT the spaces 𝔼T\mathbb{E}_{T} and 𝔽T\mathbb{F}_{T} endowed with the norms

|γ|𝑬T:=|γ|Wp1,2((0,T)×(0,1);(n)3)=γWp1,2((0,T)×(0,1);(n)3)+γ(0)Wp22/p((0,1);(n)3){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\gamma\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}:={\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\gamma\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{W_{p}^{1,2}\left((0,T)\times(0,1);(\mathbb{R}^{n})^{3}\right)}=\left\lVert\gamma\right\rVert_{W_{p}^{1,2}\left((0,T)\times(0,1);(\mathbb{R}^{n})^{3}\right)}+\left\lVert\gamma(0)\right\rVert_{W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1);(\mathbb{R}^{n})^{3}\right)}

and

|(f,η,0,b,ψ)|𝔽T:=\displaystyle{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|(f,\eta,0,b,\psi)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathbb{F}_{T}}:= fLp((0,T);Lp((0,1);(n)3))+|η|Wp11/2p((0,T);(n)3)\displaystyle\left\lVert f\right\rVert_{L_{p}\left((0,T);L_{p}((0,1);(\mathbb{R}^{n})^{3})\right)}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\eta\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{W_{p}^{1-\nicefrac{{1}}{{2p}}}\left((0,T);(\mathbb{R}^{n})^{3}\right)}
+|b|Wp1/21/2p((0,T);n)+ψWp22/p((0,1);(n)3),\displaystyle+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|b\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left((0,T);\mathbb{R}^{n}\right)}+\left\lVert\psi\right\rVert_{W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1);(\mathbb{R}^{n})^{3}\right)}\,,

respectively, are Banach spaces. Given a linear operator A:𝔽T𝔼TA:\mathbb{F}_{T}\to\mathbb{E}_{T} we let

|||A|||(𝔽T,𝔼T):=sup{|||A(f,η,0,b,ψ)|||𝑬T:(f,η,0,b,ψ)𝔽T,|||(f,η,0,b,ψ)|||𝔽T1}.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|A\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathcal{L}\left(\mathbb{F}_{T},\mathbb{E}_{T}\right)}:=\sup\{{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|A(f,\eta,0,b,\psi)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}:(f,\eta,0,b,\psi)\in\mathbb{F}_{T},{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|(f,\eta,0,b,\psi)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathbb{F}_{T}}\leq 1\}\,.
Lemma 3.5.

Let p(3,)p\in(3,\infty). For all T0>0T_{0}>0 there exists a constant c(T0,p)c(T_{0},p) such that

supT(0,T0]|LT1|(𝔽T,𝔼T)c(T0,p).\sup_{T\in(0,T_{0}]}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|L_{T}^{-1}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathcal{L}(\mathbb{F}_{T},\mathbb{E}_{T})}\leq c(T_{0},p)\,.
Proof.

Let T(0,T0]T\in(0,T_{0}] be arbitrary, (f,η,0,b,ψ)𝔽T\left(f,\eta,0,b,\psi\right)\in\mathbb{F}_{T} and ET0b:=(Eb)|(0,T0)E_{T_{0}}b:=\left(Eb\right)_{|(0,T_{0})}, ET0η:=(Eη)|(0,T0)E_{T_{0}}\eta:=\left(E\eta\right)_{|(0,T_{0})} where EE is the extension operator defined in Lemma 2.13. Extending ff by 0 to ET0fLp((0,T0);Lp((0,1);(n)3))E_{T_{0}}f\in L_{p}\left((0,T_{0});L_{p}\left((0,1);(\mathbb{R}^{n})^{3}\right)\right) we observe that (ET0f,ET0η,0,ET0b,ψ)\left(E_{T_{0}}f,E_{T_{0}}\eta,0,E_{T_{0}}b,\psi\right) lies in 𝔽T0\mathbb{F}_{T_{0}}. As LTL_{T} and LT0L_{T_{0}} are isomorphisms, there exist unique γ𝔼T\gamma\in\mathbb{E}_{T} and γ~𝔼T0\widetilde{\gamma}\in\mathbb{E}_{T_{0}} such that LTγ=(f,η,0,b,ψ)L_{T}\gamma=(f,\eta,0,b,\psi) and LT0γ~=(ET0f,ET0η,0,ET0b,ψ)L_{T_{0}}\widetilde{\gamma}=\left(E_{T_{0}}f,E_{T_{0}}\eta,0,E_{T_{0}}b,\psi\right) satisfying

LTγ=(f,η,0,b,ψ)=(ET0f,ET0η,0,ET0b,ψ)|(0,T)=(LT0γ~)|(0,T)=LT(γ~|(0,T))L_{T}\gamma=(f,\eta,0,b,\psi)=\left(E_{T_{0}}f,E_{T_{0}}\eta,0,E_{T_{0}}b,\psi\right)_{|(0,T)}=\left(L_{T_{0}}\widetilde{\gamma}\right)_{|(0,T)}=L_{T}\left(\widetilde{\gamma}_{|(0,T)}\right)

and thus γ=γ~|(0,T)\gamma=\widetilde{\gamma}_{|(0,T)}. Using Theorem 3.4, Lemma 2.13 and the equivalence of norms on 𝑬T0\boldsymbol{E}_{T_{0}} this implies

|LT1(f,η,0,b,ψ)|𝑬T=|(LT01(ET0f,ET0η,0,ET0b,ψ))|(0,T)|𝑬T\displaystyle{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|L_{T}^{-1}\left(f,\eta,0,b,\psi\right)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}={\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\left(L_{T_{0}}^{-1}\left(E_{T_{0}}f,E_{T_{0}}\eta,0,E_{T_{0}}b,\psi\right)\right)_{|(0,T)}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}
|LT01(ET0f,ET0η,0,ET0b,ψ)|𝑬T0c(T0,p)LT01(ET0f,ET0η,0,ET0b,ψ)𝑬T0\displaystyle\leq{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|L_{T_{0}}^{-1}\left(E_{T_{0}}f,E_{T_{0}}\eta,0,E_{T_{0}}b,\psi\right)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T_{0}}}\leq c\left(T_{0},p\right)\left\lVert L_{T_{0}}^{-1}\left(E_{T_{0}}f,E_{T_{0}}\eta,0,E_{T_{0}}b,\psi\right)\right\rVert_{\boldsymbol{E}_{T_{0}}}
c(T0,p)(ET0f,ET0η,0,ET0b,ψ)𝔽T0c(T0,p)|(f,η,0,b,ψ)|𝔽T.\displaystyle\leq c\left(T_{0},p\right)\left\lVert\left(E_{T_{0}}f,E_{T_{0}}\eta,0,E_{T_{0}}b,\psi\right)\right\rVert_{\mathbb{F}_{T_{0}}}\leq c\left(T_{0},p\right){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|(f,\eta,0,b,\psi)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathbb{F}_{T}}\,.

3.2 Existence and uniqueness of the Special Flow

Given MM positive we introduce the notation

BM¯:={γ𝑬T:|γ|𝑬TM}.\overline{B_{M}}:=\left\{\gamma\in\boldsymbol{E}_{T}:{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\gamma\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}\leq M\right\}\,.

This section is devoted to the proof of the following:

Theorem 3.6.

Let p(3,)p\in(3,\infty) and let σ=(σ1,σ2,σ3)\sigma=(\sigma^{1},\sigma^{2},\sigma^{3}) be an admissible initial parametrisation. There exists a positive radius MM and a positive time TT such that the system (2.4) has a unique solution σ\mathcal{E}\sigma in 𝐄TBM¯.\boldsymbol{E}_{T}\cap\overline{B_{M}}\,.

Given an admissible initial parametrisation σ\sigma and T>0T>0 we consider the complete metric spaces

𝔼Tσ\displaystyle\mathbb{E}^{\sigma}_{T} :={γ𝔼Tsuch thatγ|t=0=σ and γ|x=1=σ(1)},\displaystyle:=\{\gamma\in\mathbb{E}_{T}\;\text{such that}\;\gamma_{|t=0}=\sigma\,\text{ and }\gamma_{|x=1}=\sigma(1)\}\,,
𝔽Tσ\displaystyle\mathbb{F}^{\sigma}_{T} :=𝔽T(Lp((0,T);Lp((0,1);(n)3))×{σ(1)}×{0}×Wp1/21/2p((0,T);n)×{σ}).\displaystyle:=\mathbb{F}_{T}\cap\left(L_{p}\left((0,T);L_{p}\left((0,1);(\mathbb{R}^{n})^{3}\right)\right)\times\{\sigma(1)\}\times\{0\}\times W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left((0,T);\mathbb{R}^{n}\right)\times\{\sigma\}\right)\,.
Lemma 3.7.

Let p(3,)p\in(3,\infty), T>0T>0 and σ=(σ1,σ2,σ3)\sigma=(\sigma^{1},\sigma^{2},\sigma^{3}) be an admissible initial parametrisation. Then the space 𝔼Tσ\mathbb{E}_{T}^{\sigma} is non-empty.

Proof.

As σ\sigma is an admissible initial parametrisation, one easily checks that f0f\equiv 0, ησ(1)\eta\equiv\sigma(1), b0b\equiv 0 and ψσ\psi\equiv\sigma is an admissible right hand side for system (3.3). In other words, (0,σ(1),0,0,σ)𝔽T\left(0,\sigma(1),0,0,\sigma\right)\in\mathbb{F}_{T} and hence Theorem 3.4 yields the existence of ϱ𝔼T\varrho\in\mathbb{E}_{T} with LTϱ=(0,σ(1),0,0,σ)L_{T}\varrho=\left(0,\sigma(1),0,0,\sigma\right). In particular, ϱ|t=0=σ\varrho_{|t=0}=\sigma and ϱ|x=1=σ(1)\varrho_{|x=1}=\sigma(1) which gives ϱ𝔼Tσ\varrho\in\mathbb{E}_{T}^{\sigma}. ∎

Lemma 3.8.

Let p(3,)p\in(3,\infty) and

𝒄:=12mini{1,2,3},x[0,1]|σxi(x)|.\boldsymbol{c}:=\frac{1}{2}\min_{i\in\{1,2,3\},x\in[0,1]}|\sigma^{i}_{x}(x)|\,.

Given T0>0T_{0}>0 and M>0M>0 there exists a time T~(𝐜,M)(0,T0]\widetilde{T}(\boldsymbol{c},M)\in(0,T_{0}] such that for all γ𝔼TσBM¯\gamma\in\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}} with T[0,T~(𝐜,M)]T\in[0,\widetilde{T}(\boldsymbol{c},M)] it holds

infx[0,1],t[0,T],i{1,2,3}|γxi(t,x)|𝒄.\inf_{x\in[0,1],t\in[0,T],i\in\{1,2,3\}}\left\lvert\gamma^{i}_{x}(t,x)\right\rvert\geq\boldsymbol{c}\,.

In particular, the curves γi(t)\gamma^{i}(t) are regular for all t[0,T]t\in[0,T].

Proof.

Let p(3,)p\in(3,\infty), θ(1+1/p22/p,1)\theta\in\left(\frac{1+\nicefrac{{1}}{{p}}}{2-\nicefrac{{2}}{{p}}},1\right) and δ(0,11/p)\delta\in\left(0,1-\nicefrac{{1}}{{p}}\right). By Theorem 2.15 there exists a constant C(T0,p,θ,δ)>0C(T_{0},p,\theta,\delta)>0 such that for all T(0,T0]T\in(0,T_{0}] and all γ𝔼TσBM¯\gamma\in\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}} with α:=(1θ)(11/pδ)\alpha:=(1-\theta)(1-\nicefrac{{1}}{{p}}-\delta) it holds

γCα([0,T];C1([0,1];(n)3))C(T0,p,θ,δ)|γ|𝑬TC(T0,p,θ,δ)M,\lVert\gamma\rVert_{C^{\alpha}\left([0,T];C^{1}([0,1];(\mathbb{R}^{n})^{3})\right)}\leq C(T_{0},p,\theta,\delta){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\gamma\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}\leq C(T_{0},p,\theta,\delta)M\,,

which implies in particular for all t[0,T]t\in[0,T],

γ(t)σC1([0,1];(n)3)TαC(T0,p,θ,δ)M.\lVert\gamma(t)-\sigma\rVert_{C^{1}([0,1];(\mathbb{R}^{n})^{3})}\leq T^{\alpha}C(T_{0},p,\theta,\delta)M\,.

We let T~(𝒄,M)\widetilde{T}(\boldsymbol{c},M) be so small that T~(𝒄,M)αC(T0,p,θ,δ)M𝒄\widetilde{T}(\boldsymbol{c},M)^{\alpha}C(T_{0},p,\theta,\delta)M\leq\boldsymbol{c}. Then it follows for all γ𝔼Tσ\gamma\in\mathbb{E}_{T}^{\sigma} with T(0,T~(𝒄,M))T\in(0,\widetilde{T}(\boldsymbol{c},M)),

inft[0,T],x[0,1]|γxi(t,x)|infx[0,1]|σxi(x)|supt[0,T],x[0,1]|γxi(t,x)γxi(0,x)|𝒄.\inf_{t\in[0,T],x\in[0,1]}|\gamma^{i}_{x}(t,x)|\geq\inf_{x\in[0,1]}|\sigma^{i}_{x}(x)|-\sup_{t\in[0,T],x\in[0,1]}|\gamma^{i}_{x}(t,x)-\gamma^{i}_{x}(0,x)|\geq\boldsymbol{c}\,.

Let us now define the operator NTN_{T} that encodes the non–linearity of our problem. The map NT:𝔼Tσ𝔽TσN_{T}:\mathbb{E}^{\sigma}_{T}\to\mathbb{F}_{T}^{\sigma} is given by γ(NT1(γ),γ|x=1,0,NT2(γ),γ|t=0)\gamma\mapsto\left(N_{T}^{1}(\gamma),\gamma_{|x=1},0,N_{T}^{2}(\gamma),\gamma_{|t=0}\right) where the two components NT1,NT2N^{1}_{T},N^{2}_{T} are defined as

NT1:\displaystyle N^{1}_{T}: {𝔼TσLp((0,T);Lp((0,1);(n)3)),γf(γ),\displaystyle\begin{cases}\mathbb{E}^{\sigma}_{T}&\to L_{p}((0,T);L_{p}((0,1);(\mathbb{R}^{n})^{3}))\,,\\ \gamma&\mapsto f(\gamma)\,,\end{cases}
NT2:\displaystyle N^{2}_{T}: {𝔼TσWp1/21/2p((0,T);n),γb(γ)\displaystyle\begin{cases}\mathbb{E}^{\sigma}_{T}&\to W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}((0,T);\mathbb{R}^{n})\,,\\ \gamma&\mapsto b(\gamma)\end{cases}

with

f(γ)i(t,x)\displaystyle f(\gamma)^{i}(t,x) :=(1|γxi(t,x)|21|σxi(x)|2)γxxi(t,x),\displaystyle:=\left(\frac{1}{\left|\gamma^{i}_{x}(t,x)\right|^{2}}-\frac{1}{\left|\sigma^{i}_{x}(x)\right|^{2}}\right)\gamma^{i}_{xx}(t,x)\,,
b(γ)(t)\displaystyle b(\gamma)(t) :=i=13((1|γxi(t,0)|1|σxi(0)|)γxi(t,0)+σxi(0)γxi(t,0),σxi(0)|σxi(0)|3)\displaystyle:=\sum_{i=1}^{3}\left(\left(\frac{1}{|\gamma^{i}_{x}(t,0)|}-\frac{1}{|\sigma^{i}_{x}(0)|}\right)\gamma^{i}_{x}(t,0)+\frac{\sigma^{i}_{x}(0)\left\langle\gamma^{i}_{x}(t,0),\sigma^{i}_{x}(0)\right\rangle}{|\sigma^{i}_{x}(0)|^{3}}\right)

defined by the right hand side of (3.1) and (3.2), respectively.

Proposition 3.9.

Let p(3,)p\in(3,\infty) and MM be positive. Then for all T(0,T~(𝐜,M)]T\in(0,\widetilde{T}(\boldsymbol{c},M)] the map

NT:𝔼TσBM¯𝔽Tσ,NT(γ):=(NT1(γ),γ|x=1,0,NT2(γ),γ|t=0)N_{T}:\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}}\to\mathbb{F}_{T}^{\sigma}\,,\quad N_{T}(\gamma):=\left(N_{T}^{1}(\gamma),\gamma_{|x=1},0,N_{T}^{2}(\gamma),\gamma_{|t=0}\right)

is well-defined.

Proof.

Let T(0,T~(𝒄,M)]T\in(0,\widetilde{T}(\boldsymbol{c},M)] and γ𝔼TσBM¯\gamma\in\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}} be given. Lemma 3.8 implies

(1|γxi|21|σxi|2)γxxiLp((0,T);Lp((0,1);n))=0T01|1|γxi|21|σxi|2|p|γxxi|pdxdt\displaystyle\Big{\|}\left(\frac{1}{|\gamma^{i}_{x}|^{2}}-\frac{1}{|\sigma^{i}_{x}|^{2}}\right)\gamma^{i}_{xx}\Big{\|}_{L_{p}((0,T);L_{p}((0,1);\mathbb{R}^{n}))}=\int_{0}^{T}\int_{0}^{1}\left\lvert\frac{1}{|\gamma^{i}_{x}|^{2}}-\frac{1}{|\sigma^{i}_{x}|^{2}}\right\rvert^{p}|\gamma^{i}_{xx}|^{p}\,\mathrm{d}x\,\mathrm{d}t
C(supx[0,1],t[0,T]1|γxi|2p+supx[0,1]1|σxi|2p)0T01|γxxi|pdxdt\displaystyle\leq C\left(\sup_{x\in[0,1],t\in[0,T]}\frac{1}{|\gamma^{i}_{x}|^{2p}}+\sup_{x\in[0,1]}\frac{1}{|\sigma^{i}_{x}|^{2p}}\right)\int_{0}^{T}\int_{0}^{1}|\gamma^{i}_{xx}|^{p}\,\mathrm{d}x\,\mathrm{d}t
C(𝒄)γxxiLp((0,T);Lp((0,1);n))pC(𝒄,M)<.\displaystyle\leq C(\boldsymbol{c})\|\gamma^{i}_{xx}\|^{p}_{L_{p}((0,T);L_{p}((0,1);\mathbb{R}^{n}))}\leq C(\boldsymbol{c},M)<\infty\,.

We now show that NT2(γ)N_{T}^{2}(\gamma) lies in Wp1/21/2p((0,T);n)W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left((0,T);\mathbb{R}^{n}\right). Let h:nnh:\mathbb{R}^{n}\to\mathbb{R}^{n} be a smooth function such that h(p)=p|p|h(p)=\frac{p}{|p|} for all pnB𝒄/2(0)p\in\mathbb{R}^{n}\setminus B_{\nicefrac{{\boldsymbol{c}}}{{2}}}(0). Then one observes that for all t[0,T]t\in[0,T]

b(γ)(t)=i=13h(γxi(t))(Dh)(σxi)γxi(t)b(\gamma)(t)=\sum_{i=1}^{3}h(\gamma^{i}_{x}(t))-\left(Dh\right)(\sigma^{i}_{x})\gamma^{i}_{x}(t) (3.12)

where we omitted the evaluation in x=0x=0 to ease notation. Each term in the sum can be expressed as

h(γxi(t))(Dh)(σxi)γxi(t)\displaystyle h(\gamma^{i}_{x}(t))-\left(Dh\right)(\sigma^{i}_{x})\gamma^{i}_{x}(t) =01(Dh)(ξγxi(t)+(1ξ)σxi)dξ(γxi(t)σxi)\displaystyle=\int_{0}^{1}(Dh)(\xi\gamma^{i}_{x}(t)+(1-\xi)\sigma^{i}_{x})\mathrm{d}\xi\,(\gamma^{i}_{x}(t)-\sigma^{i}_{x})
(Dh)(σxi)(γxi(t)σxi)+h(σxi)Dh(σxi)σxi.\displaystyle\phantom{=}-(Dh)(\sigma^{i}_{x})\left(\gamma^{i}_{x}(t)-\sigma^{i}_{x}\right)+h(\sigma^{i}_{x})-Dh(\sigma^{i}_{x})\sigma^{i}_{x}\,.

All terms that are constant in tt are smooth in tt and by Lemma 2.7 we have

tγxi(t,0)Wp1/21/2p((0,T);n).t\mapsto\gamma^{i}_{x}(t,0)\in W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left((0,T);\mathbb{R}^{n}\right)\,.

As Wp1/21/2p((0,T);)W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left((0,T);\mathbb{R}\right) is a Banach algebra according to Proposition 2.9, it only remains to show

t01(Dh)(ξγxi(t,0)+(1ξ)σxi(0))dξWp1/21/2p((0,T);n×n)t\mapsto\int_{0}^{1}(Dh)(\xi\gamma^{i}_{x}(t,0)+(1-\xi)\sigma^{i}_{x}(0))\mathrm{d}\xi\in W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left((0,T);\mathbb{R}^{n\times n}\right)

which follows from the second assertion in Proposition 2.9. Observe that γ|x=1=σ(1)\gamma_{|x=1}=\sigma(1) and γ|t=0=σ\gamma_{|t=0}=\sigma by definition of 𝔼Tσ\mathbb{E}_{T}^{\sigma}. As

NT2(γ)|t=0=i=13σxi(0)|σxi(0)|=0=i=13(σxi(0)|σxi(0)|σxi(0)σxi(0),σxi(0)|σxi(0)|3)N_{T}^{2}(\gamma)_{|t=0}=\sum_{i=1}^{3}\frac{\sigma^{i}_{x}(0)}{|\sigma^{i}_{x}(0)|}=0=-\sum_{i=1}^{3}\left(\frac{\sigma^{i}_{x}(0)}{|\sigma^{i}_{x}(0)|}-\frac{\sigma^{i}_{x}(0)\left\langle\sigma^{i}_{x}(0),\sigma^{i}_{x}(0)\right\rangle}{|\sigma^{i}_{x}(0)|^{3}}\right)

and as σi(0)=σj(0)\sigma^{i}(0)=\sigma^{j}(0), σi(1)=γi(0,1)\sigma^{i}(1)=\gamma^{i}(0,1), we may conclude that

(NT1(γ),γ|x=1,0,NT2(γ),γ|t=0)=(NT1(γ),σ(1),0,NT2(γ),σ)𝔽Tσ.\left(N_{T}^{1}(\gamma),\gamma_{|x=1},0,N_{T}^{2}(\gamma),\gamma_{|t=0}\right)=\left(N_{T}^{1}(\gamma),\sigma(1),0,N_{T}^{2}(\gamma),\sigma\right)\in\mathbb{F}_{T}^{\sigma}.

Corollary 3.10.

Let p(3,)p\in(3,\infty) and MM be positive. Then for all T(0,T~(𝐜,M)]T\in(0,\widetilde{T}(\boldsymbol{c},M)] the map

KT:𝔼TσBM¯𝔼Tσ,KT:=LT1NTK_{T}:\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}}\to\mathbb{E}_{T}^{\sigma}\,,\quad K_{T}:=L_{T}^{-1}N_{T}

is well-defined.

Proof.

Let T(0,T~(𝒄,M)]T\in(0,\widetilde{T}(\boldsymbol{c},M)] and γ𝔼TσBM¯\gamma\in\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}}. By the previous proof we have

NT(γ)=(NT1(γ),γ|x=1,0,NT2(γ),γ|t=0)𝔽Tσ𝔽TN_{T}(\gamma)=\left(N_{T}^{1}(\gamma),\gamma_{|x=1},0,N_{T}^{2}(\gamma),\gamma_{|t=0}\right)\in\mathbb{F}_{T}^{\sigma}\subset\mathbb{F}_{T}

and thus in particular

KT(γ)=LT1(NT(γ))𝔼T.K_{T}(\gamma)=L_{T}^{-1}(N_{T}(\gamma))\in\mathbb{E}_{T}\,.

To verify that KT(γ)K_{T}(\gamma) lies in 𝔼Tσ\mathbb{E}_{T}^{\sigma} we observe that

KT(γ)|t=0\displaystyle K_{T}(\gamma)_{|t=0} =(LT(KT(γ)))5=NT(γ)5=γ|t=0=σ,\displaystyle=\left(L_{T}\left(K_{T}(\gamma)\right)\right)_{5}=N_{T}(\gamma)_{5}=\gamma_{|t=0}=\sigma\,,
KT(γ)|x=1\displaystyle K_{T}(\gamma)_{|x=1} =(LT(KT(γ)))2=NT(γ)2=γ|x=1=σ(1).\displaystyle=\left(L_{T}\left(K_{T}(\gamma)\right)\right)_{2}=N_{T}(\gamma)_{2}=\gamma_{|x=1}=\sigma(1)\,.

Proposition 3.11.

Let p(3,)p\in(3,\infty) and MM be positive. There exists T(𝐜,M)(0,T~(𝐜,M)]T(\boldsymbol{c},M)\in(0,\widetilde{T}(\boldsymbol{c},M)] such that for all T(0,T(𝐜,M)]T\in(0,T(\boldsymbol{c},M)] the map KT:𝔼TσBM¯𝔼TσK_{T}:\mathbb{E}^{\sigma}_{T}\cap\overline{B_{M}}\to\mathbb{E}^{\sigma}_{T} is a contraction.

Proof.

Let T(0,T~(𝒄,M)]T\in(0,\widetilde{T}(\boldsymbol{c},M)] and γ,γ~𝔼TσBM¯\gamma,\widetilde{\gamma}\in\mathbb{E}^{\sigma}_{T}\cap\overline{B_{M}} be fixed. We begin by estimating

NT1(γ)NT1(γ~)Lp((0,T);Lp((0,1);(n)3))=f(γ)f(γ~)Lp((0,T);Lp((0,1);(n)3)).\|N^{1}_{T}(\gamma)-N^{1}_{T}(\widetilde{\gamma})\|_{L_{p}\left((0,T);L_{p}\left((0,1);(\mathbb{R}^{n})^{3}\right)\right)}=\|f(\gamma)-f(\widetilde{\gamma})\|_{L_{p}\left((0,T);L_{p}\left((0,1);(\mathbb{R}^{n})^{3}\right)\right)}\,.

The ii-th component of f(γ)f(γ~)f(\gamma)-f(\widetilde{\gamma}) is given by

(1|γxi|21|σxi|2)(γxxiγ~xxi)+(1|γxi|21|γ~xi|2)γ~xxi\displaystyle\left(\frac{1}{|\gamma_{x}^{i}|^{2}}-\frac{1}{|\sigma_{x}^{i}|^{2}}\right)(\gamma_{xx}^{i}-\widetilde{\gamma}_{xx}^{i})+\left(\frac{1}{|\gamma_{x}^{i}|^{2}}-\frac{1}{|\widetilde{\gamma}_{x}^{i}|^{2}}\right)\widetilde{\gamma}^{i}_{xx}
=(1|γxi|2|σxi|+1|γxi||σxi|2)(|σxi||γxi|)(γxxiγ~xxi)\displaystyle=\left(\frac{1}{|\gamma^{i}_{x}|^{2}|\sigma^{i}_{x}|}+\frac{1}{|\gamma^{i}_{x}||\sigma^{i}_{x}|^{2}}\right)\left(|\sigma^{i}_{x}|-|\gamma^{i}_{x}|\right)(\gamma_{xx}^{i}-\widetilde{\gamma}_{xx}^{i})
+(1|γxi|2|γ~xi|+1|γxi||γ~xi|2)(|γ~xi||γxi|)γ~xxi.\displaystyle\quad+\left(\frac{1}{|\gamma^{i}_{x}|^{2}|\widetilde{\gamma}^{i}_{x}|}+\frac{1}{|\gamma^{i}_{x}||\widetilde{\gamma}^{i}_{x}|^{2}}\right)\left(|\widetilde{\gamma}^{i}_{x}|-|\gamma^{i}_{x}|\right)\widetilde{\gamma}^{i}_{xx}\,.

Lemma 3.8 implies

supt[0,T],x[0,1]|1|γxi|2|σxi|+1|γxi||σxi|2|C(𝒄)<,\displaystyle\sup_{t\in[0,T],x\in[0,1]}\left|\frac{1}{|\gamma^{i}_{x}|^{2}|\sigma^{i}_{x}|}+\frac{1}{|\gamma^{i}_{x}||\sigma^{i}_{x}|^{2}}\right|\leq C(\boldsymbol{c})<\infty\,,

and

supt[0,T],x[0,1]|1|γxi|2|γ~xi|+1|γxi||γ~xi|2|C(𝒄)<.\displaystyle\sup_{t\in[0,T],x\in[0,1]}\left\lvert\frac{1}{|\gamma^{i}_{x}|^{2}|\widetilde{\gamma}^{i}_{x}|}+\frac{1}{|\gamma^{i}_{x}||\widetilde{\gamma}^{i}_{x}|^{2}}\right\rvert\leq C(\boldsymbol{c})<\infty\,.

Hence we obtain

f(γ)if(γ~)iLp(0,T;Lp((0,1);(n)3))\displaystyle\left\lVert f(\gamma)^{i}-f(\widetilde{\gamma})^{i}\right\rVert_{L_{p}(0,T;L_{p}\left((0,1);(\mathbb{R}^{n})^{3}\right))}
C(𝒄)((|σxi||γxi|)(γxxiγ~xxi)Lp((0,T);Lp(0,1);n)+(|γ~xi||γxi|)γ~xxiLp((0,T);Lp(0,1);n))\displaystyle\leq C(\boldsymbol{c})\left(\left\lVert\left(|\sigma^{i}_{x}|-|\gamma^{i}_{x}|\right)(\gamma_{xx}^{i}-\widetilde{\gamma}_{xx}^{i})\right\rVert_{L_{p}\left((0,T);L_{p}(0,1);\mathbb{R}^{n}\right)}+\left\lVert\left(|\widetilde{\gamma}^{i}_{x}|-|\gamma^{i}_{x}|\right)\widetilde{\gamma}^{i}_{xx}\right\rVert_{L_{p}\left((0,T);L_{p}(0,1);\mathbb{R}^{n}\right)}\right)
C(𝒄)(supt[0,T],x[0,1]||σxi(x)||γxi(t,x)||γxxiγ~xxiLp((0,T);Lp(0,1);n))\displaystyle\leq C(\boldsymbol{c})\left(\sup_{t\in[0,T],x\in[0,1]}\left||\sigma^{i}_{x}(x)|-|\gamma^{i}_{x}(t,x)|\right|\left\lVert\gamma^{i}_{xx}-\widetilde{\gamma}^{i}_{xx}\right\rVert_{L_{p}\left((0,T);L_{p}(0,1);\mathbb{R}^{n})\right)}\right.
+supt[0,T],x[0,1]||γ~xi(t,x)||γxi(t,x)||γ~xxiLp((0,T);Lp((0,1);n)))\displaystyle\left.\phantom{C(\boldsymbol{c})}\,\,\,\,\,\,\,\,+\sup_{t\in[0,T],x\in[0,1]}\left||\widetilde{\gamma}^{i}_{x}(t,x)|-|\gamma^{i}_{x}(t,x)|\right|\left\lVert\widetilde{\gamma}^{i}_{xx}\right\rVert_{L_{p}\left((0,T);L_{p}((0,1);\mathbb{R}^{n})\right)}\right)
C(𝒄)supt[0,T],x[0,1]|σxi(x)γxi(t,x)||γγ~|𝑬T\displaystyle\leq C(\boldsymbol{c})\sup_{t\in[0,T],x\in[0,1]}\left\lvert\sigma^{i}_{x}(x)-\gamma^{i}_{x}(t,x)\right\rvert{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\gamma-\widetilde{\gamma}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}
+C(𝒄)supt[0,T],x[0,1]|γ~xi(t,x)γxi(t,x)||γ~|𝑬T.\displaystyle\quad+C(\boldsymbol{c})\sup_{t\in[0,T],x\in[0,1]}\left\lvert\widetilde{\gamma}^{i}_{x}(t,x)-\gamma^{i}_{x}(t,x)\right\rvert{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\widetilde{\gamma}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}\,.

Let θ(1+1/p22/p,1)\theta\in\left(\frac{1+\nicefrac{{1}}{{p}}}{2-\nicefrac{{2}}{{p}}},1\right), δ(0,11/p)\delta\in\left(0,1-\nicefrac{{1}}{{p}}\right) be fixed and define α:=(1θ)(11/pδ)\alpha:=(1-\theta)(1-\nicefrac{{1}}{{p}}-\delta). Theorem 2.15 implies

supt[0,T],x[0,1]|σxi(x)γxi(t,x)|=supt[0,T]γxi(0)γxi(t)C([0,1];n)\displaystyle\sup_{t\in[0,T],x\in[0,1]}\left\lvert\sigma^{i}_{x}(x)-\gamma^{i}_{x}(t,x)\right\rvert=\sup_{t\in[0,T]}\left\lVert\gamma^{i}_{x}(0)-\gamma^{i}_{x}(t)\right\rVert_{C([0,1];\mathbb{R}^{n})}
supt[0,T]γi(t)γi(0)C1([0,1];n)supt[0,T]tαγiCα([0,T];C1([0,1];n))\displaystyle\leq\sup_{t\in[0,T]}\left\lVert\gamma^{i}(t)-\gamma^{i}(0)\right\rVert_{C^{1}([0,1];\mathbb{R}^{n})}\leq\sup_{t\in[0,T]}t^{\alpha}\left\lVert\gamma^{i}\right\rVert_{C^{\alpha}\left([0,T];C^{1}([0,1];\mathbb{R}^{n})\right)}
TαγiCα([0,T];C1([0,1];n))TαC(T0,p,θ,δ)|γ|𝑬TC(M)Tα.\displaystyle\leq T^{\alpha}\left\lVert\gamma^{i}\right\rVert_{C^{\alpha}\left([0,T];C^{1}([0,1];\mathbb{R}^{n})\right)}\leq T^{\alpha}C(T_{0},p,\theta,\delta){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\gamma\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}\leq C(M)T^{\alpha}\,.

Similarly we obtain

supt[0,T],x[0,1]|γ~xi(t,x)γxi(t,x)|\displaystyle\sup_{t\in[0,T],x\in[0,1]}\left\lvert\widetilde{\gamma}^{i}_{x}(t,x)-\gamma^{i}_{x}(t,x)\right\rvert =supt[0,T],x[0,1]|(γ~xiγxi)(t,x)(γ~xiγxi)(0,x)|\displaystyle=\sup_{t\in[0,T],x\in[0,1]}\left\lvert\left(\widetilde{\gamma}^{i}_{x}-\gamma^{i}_{x}\right)(t,x)-\left(\widetilde{\gamma}^{i}_{x}-\gamma^{i}_{x}\right)(0,x)\right\rvert
supt[0,T](γ~iγi)(t)(γ~iγi)(0)C1([0,1];n)\displaystyle\leq\sup_{t\in[0,T]}\left\lVert\left(\widetilde{\gamma}^{i}-\gamma^{i}\right)(t)-\left(\widetilde{\gamma}^{i}-\gamma^{i}\right)(0)\right\rVert_{C^{1}([0,1];\mathbb{R}^{n})}
supt[0,T]tαγ~iγiCα([0,T];C1([0,1];n))CTα|γ~γ|𝑬T.\displaystyle\leq\sup_{t\in[0,T]}t^{\alpha}\left\lVert\widetilde{\gamma}^{i}-\gamma^{i}\right\rVert_{C^{\alpha}\left([0,T];C^{1}([0,1];\mathbb{R}^{n})\right)}\leq CT^{\alpha}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\widetilde{\gamma}-\gamma\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}\,.

This allows us to conclude

f(γ)f(γ~)Lp((0,T);Lp((0,1);(n)3))C(𝒄,M)Tα|γγ~|𝑬T.\left\lVert f(\gamma)-f(\widetilde{\gamma})\right\rVert_{L_{p}\left((0,T);L_{p}((0,1);(\mathbb{R}^{n})^{3})\right)}\leq C(\boldsymbol{c},M)T^{\alpha}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\gamma-\widetilde{\gamma}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}\,.

We proceed by estimating

NT2(γ)NT2(γ~)Wp1/21/2p((0,T);n)=b(γ)b(γ~)Wp1/21/2p((0,T);n).\left\lVert N_{T}^{2}(\gamma)-N_{T}^{2}(\widetilde{\gamma})\right\rVert_{W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}((0,T);\mathbb{R}^{n})}=\left\lVert b(\gamma)-b(\widetilde{\gamma})\right\rVert_{W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}((0,T);\mathbb{R}^{n})}\,.

Let T(0,T~(𝒄,M)]T\in(0,\widetilde{T}(\boldsymbol{c},M)] be fixed and h:nnh:\mathbb{R}^{n}\to\mathbb{R}^{n} be a smooth function such that h(p)=p|p|h(p)=\frac{p}{|p|} on nB𝒄/2(0)\mathbb{R}^{n}\setminus B_{\nicefrac{{\boldsymbol{c}}}{{2}}}(0). As for all t[0,T]t\in[0,T] and η𝔼TσBM¯\eta\in\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}},

|ηxi(t,0)|𝒄,|\eta^{i}_{x}(t,0)|\geq\boldsymbol{c}\,,

we may conclude that for all γ,γ~𝔼TσBM¯\gamma,\widetilde{\gamma}\in\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}}, the function

tgi(t):=01(Dh)(ξγxi(t,0)+(1ξ)γ~xi(t,0))dξt\mapsto g^{i}(t):=\int_{0}^{1}\left(Dh\right)(\xi\gamma^{i}_{x}(t,0)+(1-\xi)\widetilde{\gamma}^{i}_{x}(t,0))\mathrm{d}\xi

lies in Wp1/21/2p(0,T;n×n)W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}(0,T;\mathbb{R}^{n\times n}). To ease notation we let s:=1/21/2ps:=\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}. Observe that gi(0)=(Dh)(σxi(0))g^{i}(0)=(Dh)(\sigma^{i}_{x}(0)) and thus using identity (3.12) we obtain

b(γ)(t)b(γ~)(t)=i=13(gi(t)gi(0))(γxi(t,0)γ~xi(t,0)).b(\gamma)(t)-b(\widetilde{\gamma})(t)=\sum_{i=1}^{3}\left(g^{i}(t)-g^{i}(0)\right)\left(\gamma^{i}_{x}(t,0)-\widetilde{\gamma}^{i}_{x}(t,0)\right)\,.

Using the product estimate in Proposition 2.9 we obtain

b(γ)b(γ~)Wps((0,T);n)\displaystyle\left\lVert b(\gamma)-b(\widetilde{\gamma})\right\rVert_{W_{p}^{s}\left((0,T);\mathbb{R}^{n}\right)} i=13(gigi(0))(γxi(,0)γ~xi(,0))Wps((0,T);n)\displaystyle\leq\sum_{i=1}^{3}\left\lVert\left(g^{i}-g^{i}(0)\right)\left(\gamma^{i}_{x}(\cdot,0)-\widetilde{\gamma}^{i}_{x}(\cdot,0)\right)\right\rVert_{W_{p}^{s}\left((0,T);\mathbb{R}^{n}\right)}
i=13gigi(0)C([0,T];n×n)γxi(,0)γ~xi(,0)Wps(0,T;n)\displaystyle\leq\sum_{i=1}^{3}\left\lVert g^{i}-g^{i}(0)\right\rVert_{C([0,T];\mathbb{R}^{n\times n})}\left\lVert\gamma^{i}_{x}(\cdot,0)-\widetilde{\gamma}^{i}_{x}(\cdot,0)\right\rVert_{W_{p}^{s}\left(0,T;\mathbb{R}^{n}\right)}
+gigi(0)Wps(0,T;n×n)γxi(,0)γ~xi(,0)C([0,T];n).\displaystyle\;\;\;\;\quad+\left\lVert g^{i}-g^{i}(0)\right\rVert_{W_{p}^{s}\left(0,T;\mathbb{R}^{n\times n}\right)}\left\lVert\gamma^{i}_{x}(\cdot,0)-\widetilde{\gamma}^{i}_{x}(\cdot,0)\right\rVert_{C([0,T];\mathbb{R}^{n})}\,.

As s1p>0s-\frac{1}{p}>0 due to p(3,)p\in(3,\infty) there exists β(0,1)\beta\in(0,1) such that

Wps(0,T;n)Cβ([0,T];n)\displaystyle W_{p}^{s}\left(0,T;\mathbb{R}^{n}\right)\hookrightarrow C^{\beta}\left([0,T];\mathbb{R}^{n}\right)

with embedding constant C(s,p)C(s,p). This implies in particular

supt[0,T]|gi(t)gi(0)|TβgiCβ([0,T];n×n)TβC(s,p)giWps((0,T);n×n).\sup_{t\in[0,T]}|g^{i}(t)-g^{i}(0)|\leq T^{\beta}\left\lVert g^{i}\right\rVert_{C^{\beta}\left([0,T];\mathbb{R}^{n\times n}\right)}\leq T^{\beta}C(s,p)\left\lVert g^{i}\right\rVert_{W_{p}^{s}\left((0,T);\mathbb{R}^{n\times n}\right)}\,.

Reading carefully through the estimates in Proposition 2.9 we observe that

giWps((0,T);n×n)C(T0,𝒄,M).\left\lVert g^{i}\right\rVert_{W_{p}^{s}\left((0,T);\mathbb{R}^{n\times n}\right)}\leq C(T_{0},\boldsymbol{c},M)\,.

Furthermore, given θ(1+1/p22/p,1)\theta\in\left(\frac{1+\nicefrac{{1}}{{p}}}{2-\nicefrac{{2}}{{p}}},1\right) and δ(0,11/p)\delta\in\left(0,1-\nicefrac{{1}}{{p}}\right), Theorem 2.15 implies with α:=(1θ)(11/pδ)>0\alpha:=(1-\theta)\left(1-\nicefrac{{1}}{{p}}-\delta\right)>0 the estimate

supt[0,T]|γxi(t,0)γ~xi(t,0)|\displaystyle\sup_{t\in[0,T]}\left\lvert\gamma^{i}_{x}(t,0)-\widetilde{\gamma}^{i}_{x}(t,0)\right\rvert =supt[0,T]|(γxiγ~xi)(t,0)(γxiγ~xi)(0,0)|\displaystyle=\sup_{t\in[0,T]}\left\lvert\left(\gamma^{i}_{x}-\widetilde{\gamma}^{i}_{x}\right)(t,0)-(\gamma^{i}_{x}-\widetilde{\gamma}^{i}_{x})(0,0)\right\rvert
supt[0,T](γiγ~i)(t)(γiγ~i)(0)C1([0,1];n)\displaystyle\leq\sup_{t\in[0,T]}\left\lVert\left(\gamma^{i}-\widetilde{\gamma}^{i}\right)(t)-\left(\gamma^{i}-\widetilde{\gamma}^{i}\right)(0)\right\rVert_{C^{1}\left([0,1];\mathbb{R}^{n}\right)}
Tαγiγ~iCα([0,T];C1([0,1],n))Tα|γγ~|𝑬T.\displaystyle\leq T^{\alpha}\left\lVert\gamma^{i}-\widetilde{\gamma}^{i}\right\rVert_{C^{\alpha}\left([0,T];C^{1}\left([0,1],\mathbb{R}^{n}\right)\right)}\leq T^{\alpha}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\gamma-\widetilde{\gamma}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}\,.

This allows us to conclude

b(γ)b(γ~)Wp1/21/2p(0,T;n)C(s,p,T0,𝒄,M)Tα|γγ~|𝑬T.\displaystyle\left\lVert b(\gamma)-b(\widetilde{\gamma})\right\rVert_{W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left(0,T;\mathbb{R}^{n}\right)}\leq C(s,p,T_{0},\boldsymbol{c},M)T^{\alpha}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\gamma-\widetilde{\gamma}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}\,.

Finally, Lemma 3.5 implies for all T(0,T~(𝒄,M)]T\in(0,\widetilde{T}(\boldsymbol{c},M)],

|KT(γ)KT(γ~)|𝑬T=|LT1(NT(γ)NT(γ~))|𝑬Tc(T0,p)|NT(γ)NT(γ~)|𝔽T\displaystyle{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|K_{T}(\gamma)-K_{T}(\widetilde{\gamma})\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}={\left|\kern-1.07639pt\left|\kern-1.07639pt\left|L_{T}^{-1}\left(N_{T}(\gamma)-N_{T}(\widetilde{\gamma})\right)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}\leq c(T_{0},p){\left|\kern-1.07639pt\left|\kern-1.07639pt\left|N_{T}(\gamma)-N_{T}(\widetilde{\gamma})\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathbb{F}_{T}}
=c(T0,p)(f(γ)f(γ~)Lp((0,T);Lp((0,1);(n)3))+b(γ)b(γ~)Wp1/21/2p(0,T;n))\displaystyle=c(T_{0},p)\left(\|f(\gamma)-f(\widetilde{\gamma})\|_{L_{p}\left((0,T);L_{p}\left((0,1);(\mathbb{R}^{n})^{3}\right)\right)}+\left\lVert b(\gamma)-b(\widetilde{\gamma})\right\rVert_{W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left(0,T;\mathbb{R}^{n}\right)}\right)
C(T0,p,𝒄,M)Tmin{α,β}|γγ~|𝑬T.\displaystyle\leq C(T_{0},p,\boldsymbol{c},M)T^{\min\{\alpha,\beta\}}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\gamma-\widetilde{\gamma}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}\,.

This completes the proof. ∎

To conclude the existence of a solution with the Banach Fixed Point Theorem we have to show that there exists a radius M>0M>0 such that KTK_{T} is a self-mapping of 𝔼TσBM¯\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}}.

Proposition 3.12.

Let p(3,)p\in(3,\infty). There exists a positive radius MM depending on 𝐜\boldsymbol{c} and the norm of σ\sigma in Wp22/p((0,1);(n)3)W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1);(\mathbb{R}^{n})^{3}\right) and a positive time T^(𝐜,M)\widehat{T}(\boldsymbol{c},M) such that for all T(0,T^(𝐜,M)]T\in(0,\widehat{T}(\boldsymbol{c},M)] the map

KT:𝔼TσBM¯𝔼TσBM¯K_{T}:\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}}\to\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}}

is well-defined.

Proof.

We let T0=1T_{0}=1 and define

M:=2max{supT(0,1]|LT1|(𝔽T,𝔼T),1}max{|σ|𝑬1,|(N11(σ),σ(1),0,N12(σ),σ)|𝔽1}M:=2\max\left\{\sup_{T\in(0,1]}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|L_{T}^{-1}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathcal{L}\left(\mathbb{F}_{T},\mathbb{E}_{T}\right)},1\right\}\max\left\{{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\mathcal{L}\sigma\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{1}},{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\left(N_{1}^{1}(\mathcal{L}\sigma),\sigma(1),0,N_{1}^{2}(\mathcal{L}\sigma),\sigma\right)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathbb{F}_{1}}\right\}

where σ:=L11(0,σ(1),0,0,σ)\mathcal{L}\sigma:=L_{1}^{-1}\left(0,\sigma(1),0,0,\sigma\right) denotes the extension defined in Lemma 3.7 with T=1T=1. In particular, σ\mathcal{L}\sigma lies in 𝔼TσBM¯\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}} for all T(0,1]T\in(0,1]. Moreover, for all T(0,1]T\in(0,1] we have

|KT(σ)|𝑬TsupT(0,1]|LT1|(𝔽T,𝔼T)|(N11(σ),σ(1),0,N12(σ),σ)|𝔽TM/2.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|K_{T}\left(\mathcal{L}\sigma\right)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}\leq\sup_{T\in(0,1]}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|L_{T}^{-1}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathcal{L}\left(\mathbb{F}_{T},\mathbb{E}_{T}\right)}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\left(N_{1}^{1}(\mathcal{L}\sigma),\sigma(1),0,N_{1}^{2}(\mathcal{L}\sigma),\sigma\right)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathbb{F}_{T}}\leq\nicefrac{{M}}{{2}}\,.

Let T(𝒄,M)T\left(\boldsymbol{c},M\right) be the time as in Proposition 3.11. Given T(0,T(𝒄,M)]T\in(0,T\left(\boldsymbol{c},M\right)] and γ𝔼TσBM¯\gamma\in\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}} we observe that for some β(0,1)\beta\in(0,1),

|KT(γ)KT(σ)|𝑬TC(𝒄,M)Tβ|γσ|𝑬TC(𝒄,M)Tβ2M.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|K_{T}\left(\gamma\right)-K_{T}\left(\mathcal{L}\sigma\right)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}\leq C\left(\boldsymbol{c},M\right)T^{\beta}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\gamma-\mathcal{L}\sigma\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{T}}\leq C\left(\boldsymbol{c},M\right)T^{\beta}2M\,.

We choose a time T^(𝒄,M)(0,T(𝒄,M)]\widehat{T}(\boldsymbol{c},M)\in(0,T\left(\boldsymbol{c},M\right)] so small that for all T(0,T^(𝒄,M)]T\in(0,\widehat{T}(\boldsymbol{c},M)] it holds C(𝒄,M)Tβ2MM/2C\left(\boldsymbol{c},M\right)T^{\beta}2M\leq\nicefrac{{M}}{{2}}. Finally, we conclude for all T(0,T^(𝒄,M)]T\in(0,\widehat{T}(\boldsymbol{c},M)] and γ𝔼TσBM¯\gamma\in\mathbb{E}_{T}^{\sigma}\cap\overline{B_{M}},

|KT(γ)|𝔼T|KT(γ)KT(σ)|𝔼T+|KT(σ)|𝔼TM/2+M/2=M.{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|K_{T}(\gamma)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathbb{E}_{T}}\leq{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|K_{T}(\gamma)-K_{T}(\mathcal{L}\sigma)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathbb{E}_{T}}+{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|K_{T}(\mathcal{L}\sigma)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathbb{E}_{T}}\leq\nicefrac{{M}}{{2}}+\nicefrac{{M}}{{2}}=M\,.

Theorem 3.13.

Let p(3,)p\in(3,\infty) and σ\sigma be an admissible initial parametrisation. There exists a positive time T~(σ)\widetilde{T}\left(\sigma\right) depending on mini{1,2,3},x[0,1]|σxi(x)|\min_{i\in\{1,2,3\},x\in[0,1]}|\sigma^{i}_{x}(x)| and σWp22/p((0,1);(n)3)\left\lVert\sigma\right\rVert_{W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1);(\mathbb{R}^{n})^{3}\right)} such that for all 𝐓(0,T~(σ)]\boldsymbol{T}\in(0,\widetilde{T}(\sigma)] the system (2.4) has a solution σ\mathcal{E}\sigma in

𝑬𝑻=Wp1((0,𝑻);Lp((0,1);(n)3))Lp((0,𝑻);Wp2((0,1);(n)3))\boldsymbol{E}_{\boldsymbol{T}}=W_{p}^{1}\left((0,\boldsymbol{T});L_{p}\left((0,1);(\mathbb{R}^{n})^{3}\right)\right)\cap L_{p}\left((0,\boldsymbol{T});W_{p}^{2}\left((0,1);(\mathbb{R}^{n})^{3}\right)\right)

which is unique in 𝐄𝐓BM¯\boldsymbol{E}_{\boldsymbol{T}}\cap\overline{B_{M}} with

M:=2max{supT(0,1]|LT1|(𝔽T,𝔼T),1}max{|σ|𝑬1,|(N11(σ),σ(1),0,N12(σ),σ)|𝔽1}M:=2\max\left\{\sup_{T\in(0,1]}{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|L_{T}^{-1}\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathcal{L}\left(\mathbb{F}_{T},\mathbb{E}_{T}\right)},1\right\}\max\left\{{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\mathcal{L}\sigma\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\boldsymbol{E}_{1}},{\left|\kern-1.07639pt\left|\kern-1.07639pt\left|\left(N_{1}^{1}(\mathcal{L}\sigma),\sigma(1),0,N_{1}^{2}(\mathcal{L}\sigma),\sigma\right)\right|\kern-1.07639pt\right|\kern-1.07639pt\right|}_{\mathbb{F}_{1}}\right\}

where σ:=L11(0,σ(1),0,0,σ)\mathcal{L}\sigma:=L_{1}^{-1}\left(0,\sigma(1),0,0,\sigma\right) denotes the extension defined in Lemma 3.7 with T=1T=1.

Proof.

Let MM and T^(𝒄,M)\widehat{T}\left(\boldsymbol{c},M\right) be as in Proposition 3.12 and let 𝑻(0,T^(𝒄,M)]\boldsymbol{T}\in(0,\widehat{T}\left(\boldsymbol{c},M\right)]. The fixed points of the mapping K𝑻K_{\boldsymbol{T}} in 𝔼𝑻σBM¯\mathbb{E}_{\boldsymbol{T}}^{\sigma}\cap\overline{B_{M}} are precisely the solutions of the system (2.4) in the space 𝑬𝑻BM¯\boldsymbol{E}_{\boldsymbol{T}}\cap\overline{B_{M}}. As K𝑻K_{\boldsymbol{T}} is a contraction of the complete metric space 𝔼𝑻σBM¯\mathbb{E}_{\boldsymbol{T}}^{\sigma}\cap\overline{B_{M}}, existence and uniqueness of a solution follow from the Contraction Mapping Principle. ∎

Proof of Theorem 3.6.

This follows from Theorem 3.13 where the appropriate time 𝑻\boldsymbol{T} and radius MM are specified. ∎

3.3 Existence and uniqueness of solutions to the motion by curvature

Now that we obtained existence and uniqueness of solutions to the Special Flow (2.4) we can come back to our geometric problem.

Theorem 3.14 (Local existence of the motion by curvature).

Let p(3,)p\in(3,\infty) and 𝕋0\mathbb{T}_{0} be a geometrically admissible initial Triod. Then there exists T>0T>0 such that there exists a solution to the motion by curvature in [0,T][0,T] with initial datum 𝕋0\mathbb{T}_{0} as defined in Definition 2.22 which can be described by one parametrisation in the whole time interval [0,T][0,T].

Proof.

By Definition 2.20 the geometrically admissible initial Triod 𝕋0\mathbb{T}_{0} admits a parametrisation σ=(σ1,σ2,σ3)\sigma=(\sigma^{1},\sigma^{2},\sigma^{3}) that is an admissible initial parametrisation for the Special Flow. Theorem 3.6 implies that there exists 𝑻>0\boldsymbol{T}>0 and a solution σ𝑬𝑻\mathcal{E}\sigma\in\boldsymbol{E}_{\boldsymbol{T}} to the Special Flow (2.4) in [0,𝑻][0,\boldsymbol{T}] with (σ)i(0)=σi(\mathcal{E}\sigma)^{i}(0)=\sigma^{i}. Then by Definition 2.22 𝕋=i=13(σ)i([0,𝑻]×[0,1])\mathbb{T}=\bigcup_{i=1}^{3}(\mathcal{E}\sigma)^{i}([0,\boldsymbol{T}]\times[0,1]) is a solution to the motion by curvature in [0,𝑻][0,\boldsymbol{T}] with initial datum 𝕋0\mathbb{T}_{0}. ∎

Lemma 3.15 (A composition property).

Let p(3,)p\in(3,\infty), TT be positive and

f,gLp((0,T);Wp2((0,1)))Wp1((0,T);Lp((0,1)))f,g\in L_{p}\left((0,T);W_{p}^{2}\left((0,1)\right)\right)\cap W_{p}^{1}\left((0,T);L_{p}((0,1))\right)

be such that for every t[0,T]t\in[0,T] the map g(t,):[0,1][0,1]g(t,\cdot):[0,1]\to[0,1] is a C1C^{1}–diffeomorphism. Then the map h(t,x):=f(t,g(t,x))h(t,x):=f(t,g(t,x)) lies in Lp((0,T);Wp2((0,1)))Wp1((0,T);Lp((0,1)))L_{p}\left((0,T);W_{p}^{2}\left((0,1)\right)\right)\cap W_{p}^{1}\left((0,T);L_{p}((0,1))\right) and all derivatives can be calculated by the chain rule.

Proof.

This can be shown with similar arguments as in [14, Lemma 5.3] using the embedding in Theorem 2.6. ∎

Theorem 3.16 (Local uniqueness of the motion by curvature).

Let p(3,)p\in(3,\infty), T,T~>0T,\widetilde{T}>0, 𝕋0\mathbb{T}_{0} be a geometrically admissible initial Triod and (𝕋(t))(\mathbb{T}(t)), (𝕋~(t))(\widetilde{\mathbb{T}}(t)) be two solutions to the motion by curvature with initial datum 𝕋0\mathbb{T}_{0} in [0,T][0,T] and [0,T~][0,\widetilde{T}], respectively, as defined in Definition 2.22. Then there exists a positive time T^min{T,T~}\widehat{T}\leq\min\{T,\widetilde{T}\} such that 𝕋(t)=𝕋~(t)\mathbb{T}(t)=\widetilde{\mathbb{T}}(t) for all t[0,T^]t\in[0,\widehat{T}].

Proof.

Let 𝕋0\mathbb{T}_{0} be a geometrically admissible initial Triod with regular parametrisation σWp22/p((0,1);(n)3)\sigma\in W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1);(\mathbb{R}^{n})^{3}\right). Then σ\sigma is an admissible initial value for the Special Flow (2.4) and Theorem 3.6 yields that there exists 𝑻>0\boldsymbol{T}>0 and a solution σ=((σ)1,(σ)2,(σ)3)𝑬𝑻\mathcal{E}\sigma=((\mathcal{E}\sigma)^{1},(\mathcal{E}\sigma)^{2},(\mathcal{E}\sigma)^{3})\in\boldsymbol{E}_{\boldsymbol{T}} of (2.4) with initial datum σ\sigma which is unique in 𝑬𝑻BM¯\boldsymbol{E}_{\boldsymbol{T}}\cap\overline{B_{M}} with MM as in Theorem 3.13. In particular, 𝕋(t):=(σ)(t,[0,1])\mathbb{T}(t):=(\mathcal{E}\sigma)\left(t,[0,1]\right) defines a solution to the motion by curvature (2.2) in [0,𝑻][0,\boldsymbol{T}] with initial datum 𝕋0\mathbb{T}_{0}. Suppose that there is another solution (𝕋~(t))(\widetilde{\mathbb{T}}(t)) to the motion by curvature in the sense of Definition 2.22 with initial datum 𝕋0\mathbb{T}_{0} in a time interval [0,T~][0,\widetilde{T}] for some positive T~\widetilde{T}. By possibly decreasing the time of existence T~\widetilde{T} we may assume that there exists one parametrisation γ~𝑬T~\widetilde{\gamma}\in\boldsymbol{E}_{\widetilde{T}} for the evolution (𝕋~(t))(\widetilde{\mathbb{T}}(t)) in the whole time interval [0,T~][0,\widetilde{T}].

We show that there exists a family of time dependent diffeomorphisms ψi(t):[0,1][0,1]\psi^{i}(t):[0,1]\to[0,1] with t[0,T^]t\in[0,\widehat{T}] for some T^min{T~,𝑻}\widehat{T}\leq\min\{\widetilde{T},\boldsymbol{T}\} such that the equality

γ~i(t,ψi(t,x))=(σ)i(t,x)\widetilde{\gamma}^{i}(t,\psi^{i}(t,x))=(\mathcal{E}\sigma)^{i}(t,x)

holds in the space 𝑬T^\boldsymbol{E}_{\widehat{T}}. In order to make use of the uniqueness assertion in Theorem 3.6 we construct the reparametrisations ψ=(ψ1,ψ2,ψ3)\psi=\left(\psi^{1},\psi^{2},\psi^{3}\right) in such a way that the functions (t,x)γ~i(t,ψi(t,x))(t,x)\mapsto\widetilde{\gamma}^{i}(t,\psi^{i}(t,x)) are a solution to the Special Flow in 𝑬T^\boldsymbol{E}_{\widehat{T}} with initial datum σ\sigma.

One easily shows that there exist unique diffeomorphisms ψ0i:[0,1][0,1]\psi^{i}_{0}:[0,1]\to[0,1], i{1,2,3}i\in\{1,2,3\}, of regularity ψ0iWp22/p((0,1);)\psi^{i}_{0}\in W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1);\mathbb{R}\right) such that ψ0i(0)=0\psi^{i}_{0}(0)=0, ψ0i(1)=1\psi^{i}_{0}(1)=1 and γ~i(0,ψ0i(x))=σi(x)\widetilde{\gamma}^{i}(0,\psi^{i}_{0}(x))=\sigma^{i}(x). Taking into account the special tangential velocity in (2.4) (formal) differentiation shows that the reparametrisations ψi\psi^{i} need to satisfy the following boundary value problem:

{ψti(t,x)=ψxxi(t,x)|γ~xi(t,ψi(t,x))|2ψxi(t,x)2γ~ti(t,ψi(t,x)),γ~xi(t,ψi(t,x))|γ~xi(t,ψi(t,x))|2+1|γ~xi(t,ψi(t,x))|γ~xxi(t,ψi(t,x))|γ~xi(t,ψi(t,x))|2,γ~xi(t,ψi(t,x))|γ~xi(t,ψi(t,x))|,ψi(t,0)=0,ψi(t,1)=1,ψi(0,x)=ψ0i(x).\begin{cases}\begin{array}[]{ll}\psi^{i}_{t}(t,x)&=\frac{\psi_{xx}^{i}\left(t,x\right)}{\left|\widetilde{\gamma}_{x}^{i}\left(t,\psi^{i}(t,x)\right)\right|^{2}\psi^{i}_{x}(t,x)^{2}}-\frac{\left\langle\widetilde{\gamma}_{t}^{i}(t,\psi^{i}(t,x)),\widetilde{\gamma}^{i}_{x}(t,\psi^{i}(t,x))\right\rangle}{|\widetilde{\gamma}^{i}_{x}(t,\psi^{i}(t,x))|^{2}}\\ &\,\,\,\,\,+\frac{1}{\left|\widetilde{\gamma}_{x}^{i}\left(t,\psi^{i}(t,x)\right)\right|}\left\langle\frac{\widetilde{\gamma}_{xx}^{i}\left(t,\psi^{i}(t,x)\right)}{\left|\widetilde{\gamma}_{x}^{i}\left(t,\psi^{i}(t,x)\right)\right|^{2}}\,,\,\frac{\widetilde{\gamma}_{x}^{i}(t,\psi^{i}(t,x))}{\left|\widetilde{\gamma}_{x}^{i}\left(t,\psi^{i}(t,x)\right)\right|}\right\rangle\,,\\ \psi^{i}(t,0)&=0\,,\\ \psi^{i}(t,1)&=1\,,\\ \psi^{i}(0,x)&=\psi_{0}^{i}(x)\,.\\ \end{array}\end{cases} (3.13)

Lemma 3.17 yields that there exists a solution

ψ=(ψ1,ψ2,ψ3)Wp1((0,T^);Lp((0,1);3))Lp((0,T^);Wp2((0,1);3))\psi=\left(\psi^{1},\psi^{2},\psi^{3}\right)\in W_{p}^{1}((0,\widehat{T});L_{p}((0,1);\mathbb{R}^{3}))\cap L_{p}((0,\widehat{T});W_{p}^{2}((0,1);\mathbb{R}^{3}))

to system (3.13) for some T^min{T~,𝑻}\widehat{T}\leq\min\{\widetilde{T},\boldsymbol{T}\} such that ψi(t):[0,1][0,1]\psi^{i}(t):[0,1]\to[0,1] is a diffeomorphism for every t[0,T^]t\in[0,\widehat{T}]. Then Lemma 3.15 implies that the composition (t,x)γ~i(t,ψi(t,x))(t,x)\mapsto\widetilde{\gamma}^{i}(t,\psi^{i}(t,x)) lies in 𝑬T^\boldsymbol{E}_{\widehat{T}} and by construction, it is a solution to the Special Flow. We may now argue as in the proof of [14, Theorem 5.4] to obtain that (t,x)(σ)i(t,x)(t,x)\mapsto(\mathcal{E}\sigma)^{i}(t,x) and (t,x)γ~i(t,ψi(t,x))(t,x)\mapsto\widetilde{\gamma}^{i}(t,\psi^{i}(t,x)) coincide in 𝑬T^\boldsymbol{E}_{\widehat{T}}. In particular, the networks 𝕋(t)\mathbb{T}(t) and 𝕋~(t)\widetilde{\mathbb{T}}(t) coincide for all t[0,T^]t\in[0,\widehat{T}]. ∎

Lemma 3.17.

Let p(3,)p\in(3,\infty), ψ0=(ψ01,ψ02,ψ03)Wp22/p((0,1);3)\psi_{0}=(\psi_{0}^{1},\psi_{0}^{2},\psi_{0}^{3})\in W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1);\mathbb{R}^{3}\right) with ψ0i:[0,1][0,1]\psi_{0}^{i}:[0,1]\to[0,1] a diffeomorphism with ψ0i(0)=0\psi_{0}^{i}(0)=0, ψ0i(1)=1\psi_{0}^{i}(1)=1, T~>0\widetilde{T}>0 and γ~𝔼T~\widetilde{\gamma}\in\mathbb{E}_{\widetilde{T}} be such that γ~xi(x)0\widetilde{\gamma}^{i}_{x}(x)\neq 0 for all x[0,1]x\in[0,1]. Then there exists a time T^(0,T~]\widehat{T}\in(0,\widetilde{T}] and a solution

ψ=(ψ1,ψ2,ψ3)Wp1((0,T^);Lp((0,1);3))Lp((0,T^);Wp2((0,1);3))\psi=\left(\psi^{1},\psi^{2},\psi^{3}\right)\in W_{p}^{1}((0,\widehat{T});L_{p}((0,1);\mathbb{R}^{3}))\cap L_{p}((0,\widehat{T});W_{p}^{2}((0,1);\mathbb{R}^{3}))

to system (3.13) such that ψi(t):[0,1][0,1]\psi^{i}(t):[0,1]\to[0,1] is a diffeomorphism for every t[0,T^]t\in[0,\widehat{T}].

Proof.

We observe that the right hand side of the motion equation in system (3.13) contains terms of the form fi(t,ψi(t,x))f^{i}(t,\psi^{i}(t,x)) with fiLp((0,T);Lp((0,1)))f^{i}\in L_{p}\left((0,T);L_{p}((0,1))\right). To remove this dependence it is convenient to consider the associated problem for the inverse diffeomorphisms ξ=(ξ1,ξ2,ξ3)\xi=(\xi^{1},\xi^{2},\xi^{3}) given by ξi(t):=ψi(t)1\xi^{i}(t):=\psi^{i}(t)^{-1}. Indeed suppose that ψWp1,2((0,T~)×(0,1);3)\psi\in W^{1,2}_{p}((0,\widetilde{T})\times(0,1);\mathbb{R}^{3}) is a solution to (3.13) with ψi(t):[0,1][0,1]\psi^{i}(t):[0,1]\to[0,1] a C1C^{1}–diffeomorphism. Similar arguments as in [14, Lemma 5.3] show that also ξ\xi is of class Wp1,2((0,T~)×(0,1);3)W^{1,2}_{p}((0,\widetilde{T})\times(0,1);\mathbb{R}^{3}). Moreover, the differentiation rules

ξyi(t,y)\displaystyle\xi^{i}_{y}(t,y) =ψxi(t,ξi(t,y))1,\displaystyle=\psi^{i}_{x}(t,\xi^{i}(t,y))^{-1}\,,
ξyyi(t,y)\displaystyle\xi^{i}_{yy}(t,y) =ξyi(t,y)3ψxxi(t,ξi(t,y))\displaystyle=-\xi^{i}_{y}(t,y)^{3}\psi^{i}_{xx}(t,\xi^{i}(t,y))

yield the evolution equation

ξti(t,y)=\displaystyle\xi^{i}_{t}(t,y)= ψti(t,ξi(t,y))ξyi(t,y)\displaystyle-\psi^{i}_{t}(t,\xi^{i}(t,y))\xi_{y}^{i}(t,y)
=\displaystyle= ψxxi(t,ξi(t,y))|γ~xi(t,y)|2ξyi(t,y)3+γ~ti(t,y),γ~xi(t,y)|γ~xi(t,y)|2ξyi(t,y)\displaystyle-\frac{\psi_{xx}^{i}\left(t,\xi^{i}(t,y)\right)}{\left|\widetilde{\gamma}_{x}^{i}\left(t,y\right)\right|^{2}}\xi^{i}_{y}(t,y)^{3}+\frac{\left\langle\widetilde{\gamma}_{t}^{i}(t,y),\widetilde{\gamma}^{i}_{x}(t,y)\right\rangle}{|\widetilde{\gamma}^{i}_{x}(t,y)|^{2}}\xi^{i}_{y}(t,y)
ξyi(t,y)|γ~xi(t,y)|γ~xxi(t,y)|γ~xi(t,y)|2,γ~xi(t,y)|γ~xi(t,y)|,\displaystyle-\frac{\xi^{i}_{y}(t,y)}{\left|\widetilde{\gamma}_{x}^{i}\left(t,y\right)\right|}\left\langle\frac{\widetilde{\gamma}_{xx}^{i}\left(t,y\right)}{\left|\widetilde{\gamma}_{x}^{i}\left(t,y\right)\right|^{2}}\,,\,\frac{\widetilde{\gamma}_{x}^{i}(t,y)}{\left|\widetilde{\gamma}_{x}^{i}\left(t,y\right)\right|}\right\rangle\,,

and in conclusion the following system for ξ\xi:

{ξti(t,y)=ξyyi(t,y)|γ~xi(t,y)|2+γ~ti(t,y),γ~xi(t,y)|γ~xi(t,y)|2ξyi(t,y)ξyi(t,y)|γ~xi(t,y)|γ~xxi(t,y)|γ~xi(t,y)|2,γ~xi(t,y)|γ~xi(t,y)|,ξi(t,0)=0,ξi(t,1)=1,ξi(0,y)=(ψ0i)1(y)\begin{cases}\begin{array}[]{ll}\xi^{i}_{t}(t,y)&=\frac{\xi^{i}_{yy}(t,y)}{\left|\widetilde{\gamma}_{x}^{i}\left(t,y\right)\right|^{2}}+\frac{\left\langle\widetilde{\gamma}_{t}^{i}(t,y),\widetilde{\gamma}^{i}_{x}(t,y)\right\rangle}{|\widetilde{\gamma}^{i}_{x}(t,y)|^{2}}\xi^{i}_{y}(t,y)-\frac{\xi^{i}_{y}(t,y)}{\left|\widetilde{\gamma}_{x}^{i}\left(t,y\right)\right|}\left\langle\frac{\widetilde{\gamma}_{xx}^{i}\left(t,y\right)}{\left|\widetilde{\gamma}_{x}^{i}\left(t,y\right)\right|^{2}}\,,\,\frac{\widetilde{\gamma}_{x}^{i}(t,y)}{\left|\widetilde{\gamma}_{x}^{i}\left(t,y\right)\right|}\right\rangle\,,\\ \xi^{i}(t,0)&=0\,,\\ \xi^{i}(t,1)&=1\,,\\ \xi^{i}(0,y)&=(\psi_{0}^{i})^{-1}(y)\\ \end{array}\end{cases} (3.14)

for all t[0,T~]t\in[0,\widetilde{T}], y[0,1]y\in[0,1]. We observe that the boundary value problem (3.14) has a very similar structure as the Special Flow. Analogous arguments as in the proof of Theorem 3.6 allow us to conclude that there exists a solution ξWp1,2((0,T^)×(0,1);(2)3)\xi\in W^{1,2}_{p}((0,\widehat{T})\times(0,1);(\mathbb{R}^{2})^{3}) to (3.14) with T^(0,T~]\widehat{T}\in(0,\widetilde{T}] such that for t[0,T^]t\in[0,\widehat{T}] the map ξi(t):[0,1][0,1]\xi^{i}(t):[0,1]\to[0,1] is a C1C^{1}–diffeomorphism. Reversing the above argumentation yields that the inverse functions ψi(t):=ξi(t)1\psi^{i}(t):=\xi^{i}(t)^{-1} solve (3.13) and possess the desired properties. ∎

Theorem 3.18 (Geometric uniqueness of the motion by curvature).

Let p(3,)p\in(3,\infty), 𝕋0\mathbb{T}_{0} be a geometrically admissible initial Triod and TT be positive. Solutions to the motion by curvature in [0,T][0,T] with initial datum 𝕋0\mathbb{T}_{0} are geometrically unique in the sense that given any two solutions (𝕋(t))(\mathbb{T}(t)) and (𝕋~(t))(\widetilde{\mathbb{T}}(t)) to the motion by curvature in the time interval [0,T][0,T] with initial datum 𝕋0\mathbb{T}_{0} the networks 𝕋(t)\mathbb{T}(t) and 𝕋~(t)\widetilde{\mathbb{T}}(t) coincide for all t[0,T]t\in[0,T].

Proof.

Let (𝕋(t))(\mathbb{T}(t)) and (𝕋~(t))(\widetilde{\mathbb{T}}(t)) be two solutions to the motion by curvature in [0,T][0,T] with initial datum 𝕋0\mathbb{T}_{0}. Suppose by contradiction that the set

𝒮:={t[0,T]:𝕋(t)𝕋~(t)}\mathcal{S}:=\left\{t\in[0,T]:\mathbb{T}(t)\neq\widetilde{\mathbb{T}}(t)\right\}

is non-empty and let t:=inf𝒮t^{*}:=\inf\mathcal{S}. As 𝒮\mathcal{S} is an open subset of [0,T][0,T], we have t[0,T)t^{*}\in[0,T) and 𝕋(t)=𝕋~(t)\mathbb{T}(t^{*})=\widetilde{\mathbb{T}}(t^{*}). The Triod 𝕋(t)\mathbb{T}(t^{*}) is geometrically admissible and both t𝕋(t+t)t\mapsto\mathbb{T}(t^{*}+t) and t𝕋~(t+t)t\mapsto\widetilde{\mathbb{T}}(t^{*}+t) are solutions to the motion by curvature in the time interval [0,Tt][0,T-t^{*}] with initial datum 𝕋(t)\mathbb{T}(t^{*}). Theorem 3.16 yields that there exists a time T^(0,Tt]\widehat{T}\in(0,T-t^{*}] such that for all t[0,T^]t\in[0,\widehat{T}], 𝕋(t+t)=𝕋~(t+t)\mathbb{T}(t^{*}+t)=\widetilde{\mathbb{T}}(t^{*}+t) which contradicts the definition of tt^{*}. ∎

Definition 3.19 (Maximal solutions to the motion by curvature).

Let p(3,)p\in(3,\infty) and 𝕋0\mathbb{T}_{0} be a geometrically admissible initial network. A time–dependent family of Triods (𝕋(t))t[0,T)\left(\mathbb{T}(t)\right)_{t\in[0,T)} with T(0,){}T\in(0,\infty)\cup\{\infty\} is a maximal solution to the motion by curvature in [0,T)[0,T) with initial datum 𝕋0\mathbb{T}_{0} if it is a solution (in the sense of Definition 2.22) in [0,T^][0,\hat{T}] for all T^<T\hat{T}<T and if there does not exist a solution (𝕋~(τ))(\widetilde{\mathbb{T}}(\tau)) to the motion by curvature in the sense of Definition 2.22 in [0,T~][0,\widetilde{T}] with T~T\widetilde{T}\geq T and such that 𝕋=𝕋~\mathbb{T}=\widetilde{\mathbb{T}} in [0,T)[0,T). In this case the time TT is called maximal time of existence and is denoted by TmaxT_{max}.

Proposition 3.20 (Existence and uniqueness of maximal solutions).

Let p(3,)p\in(3,\infty) and 𝕋0\mathbb{T}_{0} be a geometrically admissible initial network. There exists a maximal solution to the motion by curvature with initial datum 𝕋0\mathbb{T}_{0} which is geometrically unique.

Proof.

Given an admissible Triod 𝕋0\mathbb{T}_{0} we let

Tmax:=\displaystyle T_{max}:= sup{T>0: there exists a solution (𝕋T(t)) to the motion by curvature in [0,T]\displaystyle\sup\left\{T>0:\text{ there exists a solution }(\mathbb{T}^{T}(t))\text{ to the motion by curvature in }[0,T]\right.
 with initial datum 𝕋0}.\displaystyle\qquad\quad\,\,\,\,\qquad\left.\text{ with initial datum }\mathbb{T}_{0}\right\}\,.

Theorem 3.14 yields Tmax(0,){}T_{max}\in(0,\infty)\cup\{\infty\}. Given any t[0,Tmax)t\in[0,T_{max}) we may consider a solution 𝕋T\mathbb{T}^{T} with T(t,Tmax)T\in(t,T_{max}) to the motion by curvature in [0,T][0,T] with initial datum 𝕋0\mathbb{T}_{0} and set

𝕋(t):=𝕋T(t).\mathbb{T}(t):=\mathbb{T}^{T}(t)\,.

We note that 𝕋\mathbb{T} is well-defined on [0,Tmax)[0,T_{max}) as any two solutions 𝕋T1\mathbb{T}^{T_{1}} and 𝕋T2\mathbb{T}^{T_{2}} with T1T_{1}, T2[0,Tmax)T_{2}\in[0,T_{max}) to the motion by curvature with initial datum 𝕋0\mathbb{T}_{0} coincide on their common interval of existence by Theorem 3.18. One easily verifies that (𝕋(t))t[0,Tmax)(\mathbb{T}(t))_{t\in[0,T_{max})} satisfies the properties of a maximal solution stated in Definition 3.19. Indeed, if there existed a solution 𝕋~(τ)\widetilde{\mathbb{T}}(\tau) to the motion by curvature in [0,T~][0,\widetilde{T}] for T~Tmax\widetilde{T}\geq T_{max}, Theorem 3.14 would imply the existence of a solution with initial datum 𝕋~(T~)\widetilde{\mathbb{T}}(\widetilde{T}) in a time interval [0,δ][0,\delta], δ>0\delta>0. This would yield the existence of a solution in the time interval [0,T~+δ][0,\widetilde{T}+\delta] with initial datum 𝕋0\mathbb{T}_{0} contradicting the definition of TmaxT_{max}. The uniqueness assertion follows from Theorem 3.18. ∎

4 Smoothness of the Special Flow

This section is devoted to prove that solutions to the Special Flow are smooth for positive times. Heuristically, this regularisation effect is due to the parabolic nature of the problem. The basic idea of the proof is based on the so called parameter trick which is due to Angenent [2] and has been generalized to several situations [19, 20, 25]. However, due to the fully non-linear boundary condition

i=13γxi(t,0)|γxi(t,0)|=0\sum_{i=1}^{3}\frac{\gamma^{i}_{x}(t,0)}{|\gamma^{i}_{x}(t,0)|}=0

the Special Flow is not treated in the above mentioned results. An adaptation of the parameter trick that allows to treat fully non-linear boundary terms is presented in [12]. We follow [12, Section 4] modifying the arguments for the application in our Sobolev setting.

In the following we let σ𝑬T\mathcal{E}\sigma\in\boldsymbol{E}_{T} be a solution to the Special Flow on [0,T][0,T], T>0T>0, with initial datum σWp22/p((0,1);(n)3)\sigma\in W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1);(\mathbb{R}^{n})^{3}\right).

The key idea to apply Angenent’s parameter trick lies in an implicit function type argument which itself relies on the invertibility of the linearisation of the Special Flow in the solution σ\mathcal{E}\sigma. Thus, the linear analysis from Subsection 3.1 will not be enough to apply this method. So before we can actually start we have to generalise Theorem 3.4.

Definition 4.1.

We consider the full linearisation of system (2.4) around σ\mathcal{E}\sigma which gives

{γti(t,x)1|(σ)xi(t,x)|2γxxi(t,x)2(σ)xxi(t,x)γxi(t,x),(σ)xi(t,x)|(σ)xi(t,x)|4=fi(t,x),γ(t,1)=η(t),γ1(t,0)γ2(t,0)=0,γ2(t,0)γ3(t,0)=0,i=13(γxi(t,0)|(σ)xi(t,0)|(σ)xi(t,0)γxi(t,0),(σ)xi(t,0)|(σ)xi(t,0)|3)=b(t),γ(0,x)=ψ(x).\displaystyle\begin{cases}\begin{array}[]{rl}\gamma^{i}_{t}(t,x)-\frac{1}{\left|(\mathcal{E}\sigma)^{i}_{x}(t,x)\right|^{2}}\,\gamma^{i}_{xx}(t,x)-2\frac{(\mathcal{E}\sigma)^{i}_{xx}(t,x)\left\langle\gamma^{i}_{x}(t,x),(\mathcal{E}\sigma)^{i}_{x}(t,x)\right\rangle}{|(\mathcal{E}\sigma)^{i}_{x}(t,x)|^{4}}&=f^{i}(t,x)\,,\\ \gamma(t,1)&=\eta(t)\,,\\ \gamma^{1}\left(t,0\right)-\gamma^{2}\left(t,0\right)&=0\,,\\ \gamma^{2}(t,0)-\gamma^{3}\left(t,0\right)&=0\,,\\ -\sum_{i=1}^{3}\left(\frac{\gamma^{i}_{x}(t,0)}{|(\mathcal{E}\sigma)^{i}_{x}(t,0)|}-\frac{(\mathcal{E}\sigma)^{i}_{x}(t,0)\left\langle\gamma^{i}_{x}(t,0),(\mathcal{E}\sigma)^{i}_{x}(t,0)\right\rangle}{|(\mathcal{E}\sigma)^{i}_{x}(t,0)|^{3}}\right)&=b(t)\,,\\ \gamma\left(0,x\right)&=\psi\left(x\right)\,.\\ \end{array}\end{cases} (4.1)

Here ψ\psi is an admissible initial value with respect to the given right hand side η\eta and bb. For γ𝑬T\gamma\in\boldsymbol{E}_{T} we define 𝒜T,(γ)Lp((0,T);Lp((0,1);(n)3))\mathcal{A}_{T,\mathcal{E}}(\gamma)\in L_{p}\left((0,T);L_{p}((0,1);(\mathbb{R}^{n})^{3})\right) by

(𝒜T,(γ))i:=1|(σ)xi(t,x)|2γxxi(t,x)+2(σ)xxi(t,x)γxi(t,x),(σ)xi(t,x)|(σ)xi(t,x)|4.\left(\mathcal{A}_{T,\mathcal{E}}(\gamma)\right)^{i}:=\frac{1}{\left|(\mathcal{E}\sigma)^{i}_{x}(t,x)\right|^{2}}\,\gamma^{i}_{xx}(t,x)+2\frac{(\mathcal{E}\sigma)^{i}_{xx}(t,x)\left\langle\gamma^{i}_{x}(t,x),(\mathcal{E}\sigma)^{i}_{x}(t,x)\right\rangle}{|(\mathcal{E}\sigma)^{i}_{x}(t,x)|^{4}}\,.
Definition 4.2 (The linearised boundary operator).

Let T>0T>0 and

T,:𝑬T=Wp1,2((0,T)×(0,1);(n)3)Wp11/2p((0,T);(n)5)×Wp1/21/2p((0,T);n)\mathcal{B}_{T,\mathcal{E}}:\boldsymbol{E}_{T}=W_{p}^{1,2}\left((0,T)\times(0,1);(\mathbb{R}^{n})^{3}\right)\to W_{p}^{1-\nicefrac{{1}}{{2p}}}\left((0,T);(\mathbb{R}^{n})^{5}\right)\times W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left((0,T);\mathbb{R}^{n}\right)

be the linearised boundary operator induced by the linearisation in σ\mathcal{E}\sigma, i.e.,

T,(γ)=(γ(,1)γ1(,0)γ2(,0)γ2(,0)γ3(,0)i=13γxi(,0)|(σ)xi(,0)|(σ)xi(,0)γxi(,0),(σ)xi(,0)|(σ)xi(,0)|3).\mathcal{B}_{T,\mathcal{E}}(\gamma)=\begin{pmatrix}\gamma(\cdot,1)\\ \gamma^{1}(\cdot,0)-\gamma^{2}(\cdot,0)\\ \gamma^{2}(\cdot,0)-\gamma^{3}(\cdot,0)\\ -\sum_{i=1}^{3}\frac{\gamma^{i}_{x}(\cdot,0)}{|(\mathcal{E}\sigma)^{i}_{x}(\cdot,0)|}-\frac{(\mathcal{E}\sigma)^{i}_{x}(\cdot,0)\left\langle\gamma^{i}_{x}(\cdot,0),(\mathcal{E}\sigma)^{i}_{x}(\cdot,0)\right\rangle}{|(\mathcal{E}\sigma)^{i}_{x}(\cdot,0)|^{3}}\end{pmatrix}.

Moreover we let

𝑿T\displaystyle\boldsymbol{X}_{T} :=ker(T,).\displaystyle:=\ker(\mathcal{B}_{T,\mathcal{E}})\,.

As T,\mathcal{B}_{T,\mathcal{E}} is continuous, the space 𝑿T\boldsymbol{X}_{T} is a closed subspace of 𝑬T\boldsymbol{E}_{T} and thus a Banach space.

Remark 4.3 (Existence analysis for (4.1)).

Note that the compatibility conditions in Definition 3.1 for system (3.3) are precisely the same as the ones for (4.1) due to the fact that T,|t=0\mathcal{B}_{T,\mathcal{E}}\big{|}_{t=0} equals the original linearisation. Also, with the same arguments as in the proof of Lemma 3.3 we can derive the Lopatinskii-Shapiro conditions for T,\mathcal{B}_{T,\mathcal{E}}. Therefore, the result from Theorem 3.4 holds also for problem (4.1). For γ𝑬T\gamma\in\boldsymbol{E}_{T} we write

LT,(γ):=(γtAT,(γ)T,(γ)γ|t=0).L_{T,\mathcal{E}}(\gamma):=\begin{pmatrix}\gamma_{t}-A_{T,\mathcal{E}}(\gamma)\\ \mathcal{B}_{T,\mathcal{E}}(\gamma)\\ \gamma_{|t=0}\end{pmatrix}\,.

With the previous considerations we have the basics to start the work on the parameter trick. As a first step we have to construct a parametrisation of the non-linear boundary conditions over the linear boundary conditions. We need to do this as we cannot have the non-linear boundary operator to be part of the operator used in the parameter trick due to technical reasons with the compatibility conditions.

In the following lemma we construct a partition of the solution space 𝑬T=𝑿T𝒁T\boldsymbol{E}_{T}=\boldsymbol{X}_{T}\oplus\boldsymbol{Z}_{T}.

Lemma 4.4.

Let T>0T>0. There exists a closed subspace 𝐙T\boldsymbol{Z}_{T} of 𝐄T\boldsymbol{E}_{T} such that 𝐄T=𝐗T𝐙T\boldsymbol{E}_{T}=\boldsymbol{X}_{T}\oplus\boldsymbol{Z}_{T}.

Proof.

Firstly, we consider the space

Z¯T1:={𝔟Wp11/2p((0,T);(n)5)×Wp1/21/2p((0,T);n):𝔟|t=0=0}.\overline{Z}^{1}_{T}:=\left\{\mathfrak{b}\in W_{p}^{1-\nicefrac{{1}}{{2p}}}\left((0,T);(\mathbb{R}^{n})^{5}\right)\times W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left((0,T);\mathbb{R}^{n}\right):\mathfrak{b}_{|t=0}=0\right\}.

We notice that f=0f=0, 𝔟Z¯T1\mathfrak{b}\in\overline{Z}^{1}_{T}, ψ=0\psi=0 is a suitable right hand side for problem (4.1). Hence for every 𝔟Z¯T1\mathfrak{b}\in\overline{Z}_{T}^{1} there exists a unique solution LT,1(0,𝔟,0)𝑬TL_{T,\mathcal{E}}^{-1}\left(0,\mathfrak{b},0\right)\in\boldsymbol{E}_{T} to (4.1) and the space ZT1:=LT,1((0,Z¯T1,0))Z_{T}^{1}:=L_{T,\mathcal{E}}^{-1}\left((0,\overline{Z}_{T}^{1},0)\right) is a closed subspace of 𝑬T\boldsymbol{E}_{T}.

Next we define the space

Z¯2:=(n)5×n.\overline{Z}^{2}:=(\mathbb{R}^{n})^{5}\times\mathbb{R}^{n}\,.

Given b~Z¯2\tilde{b}\in\overline{Z}^{2} the elliptic system L~η=(0,b~)\tilde{L}\eta=(0,\tilde{b}) defined by

{1|σxi(x)|2ηxxi(x)=0,x(0,1),i{1,2,3},η1(1)=b~1,η2(1)=b~2,η3(1)=b~3,η1(0)η2(0)=b~4,η2(0)η3(0)=b~5,i=13(ηxi(0)|σxi(0)|σxi(0)ηxi(0),σxi(0)|σxi(0)|3)=b~6,\begin{cases}\begin{array}[]{rl}-\frac{1}{\left|\sigma^{i}_{x}(x)\right|^{2}}\,\eta^{i}_{xx}(x)&=0\,,\qquad\quad x\in(0,1)\,,i\in\{1,2,3\}\,,\\ \eta^{1}(1)&=\tilde{b}^{1}\,,\\ \eta^{2}(1)&=\tilde{b}^{2}\,,\\ \eta^{3}(1)&=\tilde{b}^{3}\,,\\ \eta^{1}(0)-\eta^{2}(0)&=\tilde{b}^{4}\,,\\ \eta^{2}(0)-\eta^{3}(0)&=\tilde{b}^{5}\,,\\ -\sum_{i=1}^{3}\left(\frac{\eta^{i}_{x}(0)}{|\sigma^{i}_{x}(0)|}-\frac{\sigma^{i}_{x}(0)\left\langle\eta^{i}_{x}(0),\sigma^{i}_{x}(0)\right\rangle}{|\sigma^{i}_{x}(0)|^{3}}\right)&=\tilde{b}^{6}\,,\end{array}\end{cases} (4.2)

has a unique solution ηWp2((0,1);(n)3)\eta\in W_{p}^{2}\left((0,1);(\mathbb{R}^{n})^{3}\right) which we denote by L~1(0,b~)\tilde{L}^{-1}(0,\tilde{b}). This is guaranteed due to the results in [1] and the fact that the boundary operator fulfils the Lopatinskii-Shapiro conditions according to Lemma 3.3. The space L~1(0,Z¯2)\tilde{L}^{-1}(0,\overline{Z}^{2}) is a closed subspace of Wp2((0,1);(n)3)W_{p}^{2}\left((0,1);(\mathbb{R}^{n})^{3}\right) due to continuity of the solution operator which is guaranteed by the energy estimates in [1]. Extending every function in L~1(0,Z¯2)\tilde{L}^{-1}(0,\overline{Z}^{2}) constantly in time we can view L~1(0,Z¯2)\tilde{L}^{-1}(0,\overline{Z}^{2}) as a closed subspace of 𝑬T\boldsymbol{E}_{T}. This space will be denoted by ZT2Z^{2}_{T}. It is straightforward to check that ZT1ZT2={0}Z_{T}^{1}\cap Z_{T}^{2}=\{0\} which allows us to define 𝒁T\boldsymbol{Z}_{T} as the subspace of 𝑬T\boldsymbol{E}_{T} given by

𝒁T:=ZT1ZT2.\boldsymbol{Z}_{T}:=Z_{T}^{1}\oplus Z_{T}^{2}\,.

Note that 𝒁T\boldsymbol{Z}_{T} is indeed a closed subspace which one sees as follows. Suppose that

(zn)n=(zn1+zn2)n𝒁T(z_{n})_{n\in\mathbb{N}}=(z_{n}^{1}+z_{n}^{2})_{n\in\mathbb{N}}\subset\boldsymbol{Z}_{T}

is a convergent sequence in 𝑬T\boldsymbol{E}_{T}.

Due to 𝑬TC([0,T];C1+α([0,1];(n)3))\boldsymbol{E}_{T}\hookrightarrow C([0,T];C^{1+\alpha}([0,1];({{\mathbb{R}}}^{n})^{3})) for α(0,13/p]\alpha\in\left(0,1-\nicefrac{{3}}{{p}}\right] according to Theorem 2.6 we may conclude that the sequence (zn|t=0)n=(zn2|t=0)n(z_{n}\big{|}_{t=0})_{n\in\mathbb{N}}=(z_{n}^{2}\big{|}_{t=0})_{n\in\mathbb{N}} converges in C1+α([0,1];(n)3)C^{1+\alpha}([0,1];({{\mathbb{R}}}^{n})^{3}). In particular, this yields the convergence of the boundary data needed for the elliptic system we used to construct zn2z^{2}_{n}. Continuity of the elliptic solution operator then implies that (zn2|t=0)n(z_{n}^{2}\big{|}_{t=0})_{n\in\mathbb{N}} converges in Wp2((0,1);(n)3)W^{2}_{p}((0,1);({{\mathbb{R}}}^{n})^{3}). Due to its constant extension in time we see that (zn2)n(z_{n}^{2})_{n\in\mathbb{N}} converges in 𝑬T\boldsymbol{E}_{T} to a limit z2z^{2} which is also in ZT2Z^{2}_{T} being a closed subspace of 𝑬T\boldsymbol{E}_{T}. Then (zn1)n=(zn)n(zn2)n(z_{n}^{1})_{n\in\mathbb{N}}=(z_{n})_{n\in\mathbb{N}}-(z_{n}^{2})_{n\in\mathbb{N}} converges in 𝑬T\boldsymbol{E}_{T} as sum of two convergent sequences to an element z1z^{1} of the closed space ZT1Z_{T}^{1}. We conclude that (zn)n(z_{n})_{n\in\mathbb{N}} converges to z1+z2𝒁Tz^{1}+z^{2}\in\boldsymbol{Z}_{T} which shows that 𝒁T\boldsymbol{Z}_{T} is closed.
It remains to prove that 𝑿T𝒁T={0}\boldsymbol{X}_{T}\cap\boldsymbol{Z}_{T}=\{0\} and 𝑬T=𝑿T+𝒁T\boldsymbol{E}_{T}=\boldsymbol{X}_{T}+\boldsymbol{Z}_{T}. To this end let γ𝑿T𝒁T\gamma\in\boldsymbol{X}_{T}\cap\boldsymbol{Z}_{T}. By definition of 𝑿T\boldsymbol{X}_{T} we have T,(γ)=0\mathcal{B}_{T,\mathcal{E}}(\gamma)=0 which implies in particular T,(γ)|t=0=0\mathcal{B}_{T,\mathcal{E}}(\gamma)_{|t=0}=0. As γ\gamma lies in 𝒁T\boldsymbol{Z}_{T}, there exist z1ZT1z_{1}\in Z_{T}^{1}, z2ZT2z_{2}\in Z_{T}^{2} with γ=z1+z2\gamma=z_{1}+z_{2}. Using that T,(z1)\mathcal{B}_{T,\mathcal{E}}(z_{1}) lies in Z¯T1\overline{Z}_{T}^{1}, we observe

0=T,(z1+z2)|t=0=T,(z1)|t=0+T,(z2)|t=0=T,(z2)|t=0.0=\mathcal{B}_{T,\mathcal{E}}(z_{1}+z_{2})_{|t=0}=\mathcal{B}_{T,\mathcal{E}}(z_{1})_{|t=0}+\mathcal{B}_{T,\mathcal{E}}(z_{2})_{|t=0}=\mathcal{B}_{T,\mathcal{E}}(z_{2})_{|t=0}\,.

Due to the uniqueness of the elliptic system (4.2) this shows (z2)|t=0=0(z_{2})_{|t=0}=0. By definition of ZT2Z_{T}^{2} we obtain z2=0z_{2}=0. This implies 0=T,(γ)=T,(z1)0=\mathcal{B}_{T,\mathcal{E}}(\gamma)=\mathcal{B}_{T,\mathcal{E}}(z_{1}) which gives z1=LT,1(0,0,0)=0z_{1}=L_{T,\mathcal{E}}^{-1}(0,0,0)=0.

To prove that 𝑬T=𝑿T+𝒁T\boldsymbol{E}_{T}=\boldsymbol{X}_{T}+\boldsymbol{Z}_{T} we let γ𝑬T\gamma\in\boldsymbol{E}_{T}. We define

z2:=L~1(0,T,(γ)|t=0)ZT2z_{2}:=\tilde{L}^{-1}(0,\mathcal{B}_{T,\mathcal{E}}(\gamma)_{|t=0})\in Z_{T}^{2}

viewing z2z_{2} as an element of 𝑬T\boldsymbol{E}_{T} by extending it constantly in time. By definition of the boundary operator in the elliptic system (4.2) and due to (σ)|t=0=σ(\mathcal{E}\sigma)_{|t=0}=\sigma we have

T,(z2)|t=0=T,(γ)|t=0.\mathcal{B}_{T,\mathcal{E}}(z_{2})_{|t=0}=\mathcal{B}_{T,\mathcal{E}}(\gamma)_{|t=0}\,.

In particular, T,(γ)T,(z2)\mathcal{B}_{T,\mathcal{E}}(\gamma)-\mathcal{B}_{T,\mathcal{E}}(z_{2}) lies in Z¯T1\overline{Z}^{1}_{T} and we may define

z1:=LT,1(0,T,(γ)T,(z2),0)ZT1.z_{1}:=L_{T,\mathcal{E}}^{-1}\left(0,\mathcal{B}_{T,\mathcal{E}}(\gamma)-\mathcal{B}_{T,\mathcal{E}}(z_{2}),0\right)\in Z_{T}^{1}\,.

Now it remains to show that γz1z2\gamma-z_{1}-z_{2} lies in 𝑿T\boldsymbol{X}_{T} which is equivalent to T,(γz1z2)=0\mathcal{B}_{T,\mathcal{E}}(\gamma-z_{1}-z_{2})=0 which follows by construction. ∎

Lemma 4.5 (Parametrisation of the nonlinear boundary conditions).

Let T>0T>0. There exists a neighbourhood UU of 0 in 𝐗T\boldsymbol{X}_{T}, a function ϱ:U𝐙T\varrho:U\to\boldsymbol{Z}_{T} and a neighbourhood VV of σ\mathcal{E}\sigma in 𝐄T\boldsymbol{E}_{T} such that

{σ+𝒖+ϱ(𝒖):𝒖U}={γV:𝒢(γ)=0}\left\{\mathcal{E}\sigma+\boldsymbol{u}+\varrho(\boldsymbol{u})\,:\,\boldsymbol{u}\in U\right\}=\left\{\gamma\in V:\mathcal{G}(\gamma)=0\right\}

where 𝒢\mathcal{G} denotes the operator

γ𝒢(γ):=(γ1(,1)σ1(1)γ2(,1)σ2(1)γ3(,1)σ3(1)γ1(,0)γ2(,0)γ2(,0)γ3(,0)i=13γxi(,0)|γxi(,0)|).\gamma\mapsto\mathcal{G}(\gamma):=\begin{pmatrix}\gamma^{1}(\cdot,1)-\sigma^{1}(1)\,\\ \gamma^{2}(\cdot,1)-\sigma^{2}(1)\,\\ \gamma^{3}(\cdot,1)-\sigma^{3}(1)\,\\ \gamma^{1}(\cdot,0)-\gamma^{2}(\cdot,0)\,\\ \gamma^{2}(\cdot,0)-\gamma^{3}(\cdot,0)\,\\ \sum_{i=1}^{3}\frac{\gamma^{i}_{x}(\cdot,0)}{|\gamma^{i}_{x}(\cdot,0)|}\,\end{pmatrix}\,.

Furthermore, it holds that (Dϱ)|00(D\varrho)_{|0}\equiv 0.

Proof.

We let

𝒀T:=Wp11/2p((0,T);(n)5)×Wp1/21/2p((0,T);n)\boldsymbol{Y}_{T}:=W_{p}^{1-\nicefrac{{1}}{{2p}}}\left((0,T);(\mathbb{R}^{n})^{5}\right)\times W_{p}^{\nicefrac{{1}}{{2}}-\nicefrac{{1}}{{2p}}}\left((0,T);\mathbb{R}^{n}\right)

and consider the operator

F:𝑿T𝒁T\displaystyle F:\boldsymbol{X}_{T}\oplus\boldsymbol{Z}_{T} 𝒀T,\displaystyle\to\boldsymbol{Y}_{T}\,,
(𝒙,𝒛)\displaystyle(\boldsymbol{x},\boldsymbol{z}) 𝒢(σ+𝒙+𝒛).\displaystyle\mapsto\mathcal{G}\left(\mathcal{E}\sigma+\boldsymbol{x}+\boldsymbol{z}\right)\,.

By definition of σ\mathcal{E}\sigma we have that F(0,0)=0F(0,0)=0. We observe that DF|(0,0)(0,𝒛)=T,(𝒛)DF_{|(0,0)}(0,\boldsymbol{z})=\mathcal{B}_{T,\mathcal{E}}(\boldsymbol{z}). To apply the implicit function theorem we have to show that

T,:𝒁T𝒀T\mathcal{B}_{T,\mathcal{E}}:\boldsymbol{Z}_{T}\to\boldsymbol{Y}_{T}

is an isomorphism. The map is injective as kerT,𝒁T=𝑿T𝒁T={0}\ker{\mathcal{B}_{T,\mathcal{E}}}\cap\boldsymbol{Z}_{T}=\boldsymbol{X}_{T}\cap\boldsymbol{Z}_{T}=\{0\}. Given 𝒃𝒀T\boldsymbol{b}\in\boldsymbol{Y}_{T} we let z2:=L~1(0,𝒃|t=0)ZT2z_{2}:=\tilde{L}^{-1}(0,\boldsymbol{b}_{|t=0})\in Z_{T}^{2} and z1:=LT,1(0,𝒃T,(z2))ZT1z_{1}:=L_{T,\mathcal{E}}^{-1}(0,\boldsymbol{b}-\mathcal{B}_{T,\mathcal{E}}(z_{2}))\in Z^{1}_{T} and observe that z1+z2𝒁Tz_{1}+z_{2}\in\boldsymbol{Z}_{T} satisfies

T,(z1+z2)=T,(z1)+T,(z2)=𝒃T,(z2)+T,(z2)=𝒃.\mathcal{B}_{T,\mathcal{E}}(z_{1}+z_{2})=\mathcal{B}_{T,\mathcal{E}}(z_{1})+\mathcal{B}_{T,\mathcal{E}}(z_{2})=\boldsymbol{b}-\mathcal{B}_{T,\mathcal{E}}(z_{2})+\mathcal{B}_{T,\mathcal{E}}(z_{2})=\boldsymbol{b}\,.

The implicit function theorem implies that there exist neighbourhoods UU and WW of 0 in 𝑿T\boldsymbol{X}_{T} and 𝒁T\boldsymbol{Z}_{T}, respectively, and a function ϱ:UW\varrho:U\to W with ϱ(0)=0\varrho(0)=0 such that for a neighbourhood V~\tilde{V} of 0 in 𝑬T\boldsymbol{E}_{T}, it holds

{𝒖+ϱ(𝒖):𝒖U}={𝒙+𝒛𝑬T:F(𝒙,𝒛)=0}V~.\{\boldsymbol{u}+\varrho(\boldsymbol{u}):\boldsymbol{u}\in U\}=\{\boldsymbol{x}+\boldsymbol{z}\in\boldsymbol{E}_{T}:F(\boldsymbol{x},\boldsymbol{z})=0\}\cap\tilde{V}\,.

To show that (Dϱ)|0=0(D\varrho)_{|0}=0 we let 𝒖𝑿T\boldsymbol{u}\in\boldsymbol{X}_{T} be arbitrary. Due to (Dϱ)|0:𝑿T𝒁T(D\varrho)_{|0}:\boldsymbol{X}_{T}\to\boldsymbol{Z}_{T} we obtain (Dϱ)|0𝒖𝒁T(D\varrho)_{|0}\boldsymbol{u}\in\boldsymbol{Z}_{T}. Hence it is enough to show that (Dϱ)|0𝒖(D\varrho)_{|0}\boldsymbol{u} lies also in 𝑿T\boldsymbol{X}_{T}. To this end we differentiate the identity

0=F(δ𝒖,ϱ(δ𝒖))=𝒢(σ+δ𝒖+ϱ(δ𝒖))0=F(\delta\boldsymbol{u},\varrho(\delta\boldsymbol{u}))=\mathcal{G}\left(\mathcal{E}\sigma+\delta\boldsymbol{u}+\varrho(\delta\boldsymbol{u})\right)

with respect to δ\delta and obtain

0=ddδ𝒢(σ+δ𝒖+ϱ(δ𝒖))|δ=0=(D𝒢)(σ)(𝒖+(Dϱ)|0𝒖)=T,(𝒖+(Dϱ)|0𝒖).0=\frac{\mathrm{d}}{\mathrm{d}\delta}\mathcal{G}\left(\mathcal{E}\sigma+\delta\boldsymbol{u}+\varrho(\delta\boldsymbol{u})\right)_{|\delta=0}=\left(D\mathcal{G}\right)(\mathcal{E}\sigma)(\boldsymbol{u}+(D\varrho)_{|0}\boldsymbol{u})=\mathcal{B}_{T,\mathcal{E}}(\boldsymbol{u}+(D\varrho)_{|0}\boldsymbol{u})\,.

This implies 𝒖+(Dϱ)|0𝒖kerT,=𝑿T\boldsymbol{u}+(D\varrho)_{|0}\boldsymbol{u}\in\ker\mathcal{B}_{T,\mathcal{E}}=\boldsymbol{X}_{T} and thus (Dϱ)|0𝒖𝑿T(D\varrho)_{|0}\boldsymbol{u}\in\boldsymbol{X}_{T}. ∎

With this result we can finally start the proof of the parabolic smoothing. We will first derive higher time regularity of the solution (this is actually the classical parameter trick argument by Angenent), and will then get from this higher regularity in space using the parabolic problem and finally start a bootstrap procedure.

Proposition 4.6 (Higher time regularity of solutions to the Special Flow).

Let σ𝐄T\mathcal{E}\sigma\in\boldsymbol{E}_{T} be a solution to the Special Flow in [0,T][0,T] with T>0T>0 and initial value σWp22/p((0,1);(n)3)\sigma\in W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1);(\mathbb{R}^{n})^{3}\right). Then we have for all t~(0,T]\tilde{t}\in(0,T] the increased time regularity

t(σ)𝑬T|[t~,T].\partial_{t}(\mathcal{E}{\sigma})\in\boldsymbol{E}_{T}\big{|}_{[\tilde{t},T]}\,. (4.3)
Proof.

We consider the space

𝑰:=\displaystyle\boldsymbol{I}:= {ψWp22/p((0,1);(n)3):ψ(1)=0,ψ1(0)=ψ2(0)=ψ3(0),\displaystyle\left\{\psi\in W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1);(\mathbb{R}^{n})^{3}\right):\psi(1)=0\,,\psi^{1}(0)=\psi^{2}(0)=\psi^{3}(0)\,,\right.
i=13ψxi(0)|σxi(0)|σxi(0)ψxi(0),σxi(0)|σxi(0)|3=0}.\displaystyle\left.\quad\sum_{i=1}^{3}\frac{\psi^{i}_{x}(0)}{|\sigma^{i}_{x}(0)|}-\frac{\sigma^{i}_{x}(0)\left\langle\psi^{i}_{x}(0),\sigma^{i}_{x}(0)\right\rangle}{|\sigma^{i}_{x}(0)|^{3}}=0\right\}\,.

We let UU, VV and ϱ\varrho be as in the previous Lemma and define ϱ¯(𝒖):=σ+𝒖+ϱ(𝒖)\overline{\varrho}(\boldsymbol{u}):=\mathcal{E}\sigma+\boldsymbol{u}+\varrho(\boldsymbol{u}). For some small ε(0,1)\varepsilon\in(0,1) we consider the map

G:(1ε,1+ε)×𝑰×𝑿T\displaystyle G:(1-\varepsilon,1+\varepsilon)\times\boldsymbol{I}\times\boldsymbol{X}_{T} 𝑰×Lp((0,T)×(0,1);(n)3),\displaystyle\to\boldsymbol{I}\times L_{p}\left((0,T)\times(0,1);(\mathbb{R}^{n})^{3}\right)\,,
(λ,ψ,𝒖)\displaystyle\left(\lambda,\psi,\boldsymbol{u}\right) (𝒖|t=0ψ,tϱ¯(𝒖)λϱ¯(𝒖)xx|ϱ¯(𝒖)x|2).\displaystyle\mapsto\left(\boldsymbol{u}_{|t=0}-\psi,\partial_{t}\overline{\varrho}(\boldsymbol{u})-\lambda\frac{\overline{\varrho}(\boldsymbol{u})_{xx}}{|\overline{\varrho}(\boldsymbol{u})_{x}|^{2}}\right)\,.

Notice that G(1,0,0)=0G(1,0,0)=0. Due to (Dϱ)|0=0(D\varrho)_{|0}=0 the Fréchet derivative

(DG)|(1,0,0)(0,0,):𝑿T𝑰×Lp((0,T)×(0,1);(n)3)(DG)_{|(1,0,0)}(0,0,\cdot):\boldsymbol{X}_{T}\to\boldsymbol{I}\times L_{p}\left((0,T)\times(0,1);(\mathbb{R}^{n})^{3}\right)

is given by

(DG)|(1,0,0)(0,0,𝒖)=(𝒖|t=0,t𝒖𝒜T,(𝒖)).(DG)_{|(1,0,0)}(0,0,\boldsymbol{u})=\left(\boldsymbol{u}_{|t=0},\partial_{t}\boldsymbol{u}-\mathcal{A}_{T,\mathcal{E}}(\boldsymbol{u})\right)\,.

As explained in Remark 4.3 we have that (DG)|(1,0,0)(0,0,)(DG)_{|(1,0,0)}(0,0,\cdot) is an isomorphism. Hence the implicit function theorem implies the existence of neighbourhoods 𝒰\mathcal{U} of (1,0)(1,0) in (1ε,1+ε)×𝑰(1-\varepsilon,1+\varepsilon)\times\boldsymbol{I} and 𝒱\mathcal{V} of 0 in 𝑿T\boldsymbol{X}_{T} and a function ζ:𝒰𝒱\zeta:\mathcal{U}\to\mathcal{V} with ζ((1,0))=0\zeta((1,0))=0 and

{(λ,ψ,𝒖)𝒰×𝒱:G(λ,ψ,𝒖)=0}={(λ,ψ,ζ(λ,ψ)):(λ,ψ)𝒰}.\{(\lambda,\psi,\boldsymbol{u})\in\mathcal{U}\times\mathcal{V}:G(\lambda,\psi,\boldsymbol{u})=0\}=\{(\lambda,\psi,\zeta(\lambda,\psi)):(\lambda,\psi)\in\mathcal{U}\}\,.

Consider now the map P:𝑬T𝑿TP:\boldsymbol{E}_{T}\to\boldsymbol{X}_{T} given by P(γ):=P𝑿T(γσ)P(\gamma):=P_{\boldsymbol{X}_{T}}(\gamma-\mathcal{E}\sigma) with P𝑿T(η)=𝒖P_{\boldsymbol{X}_{T}}(\eta)=\boldsymbol{u} for the unique partition η=𝒖+𝒖¯𝑿T𝒁T\eta=\boldsymbol{u}+\overline{\boldsymbol{u}}\in\boldsymbol{X}_{T}\oplus\boldsymbol{Z}_{T}. Clearly, we have that ϱ¯(P(γ))=γ\overline{\varrho}(P(\gamma))=\gamma for all γ\gamma in the neighbourhood VV constructed in Lemma 4.5. Given λ\lambda close to 11 we consider the time-scaled function

(σ)λ(t,x):=(σ)(λt,x).(\mathcal{E}{\sigma})_{\lambda}(t,x):=(\mathcal{E}{\sigma})(\lambda t,x)\,.

By definition this satisfies for ψ:=P((σ)λ)|t=0\psi:=P((\mathcal{E}{\sigma})_{\lambda})\big{|}_{t=0}

G(λ,ψ,P((σ)λ))=0.G(\lambda,\psi,P((\mathcal{E}{\sigma})_{\lambda}))=0\,.

By uniqueness we conclude that

P((σ)λ)=ζ(λ,ψ)P((\mathcal{E}{\sigma})_{\lambda})=\zeta(\lambda,\psi)

and therefore

(σ)λ=ϱ¯(ζ(λ,ψ)).(\mathcal{E}{\sigma})_{\lambda}=\bar{\varrho}(\zeta(\lambda,\psi))\,.

As both ζ\zeta and ϱ¯\bar{\varrho} are smooth, this shows that (σ)λ(\mathcal{E}{\sigma})_{\lambda} is a smooth function in λ\lambda with values in 𝑬T\boldsymbol{E}_{T}. This implies now

tt(σ)=λ((σ)λ)|λ=1𝑬T\displaystyle t\partial_{t}(\mathcal{E}{\sigma})=\partial_{\lambda}((\mathcal{E}{\sigma})_{\lambda})\big{|}_{\lambda=1}\in\boldsymbol{E}_{T} (4.4)

from which we directly conclude (4.3). ∎

Next, we want to derive higher regularity in space for our solution. But this follows almost immediately from the associated ODE we have at a fixed time.

Corollary 4.7 (Higher space regularity of solutions to the Special Flow).

Let σ𝐄T\mathcal{E}\sigma\in\boldsymbol{E}_{T} be a solution to the Special Flow in [0,T][0,T] with T>0T>0 and initial value σWp22/p((0,1);(n)3)\sigma\in W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1);(\mathbb{R}^{n})^{3}\right). Given t~(0,T]\tilde{t}\in(0,T] we have for almost all t[t~,T]t\in[\tilde{t},T] that

(σ)(t)Wp/23((0,1);(n)3).(\mathcal{E}\sigma)(t)\in W^{3}_{\nicefrac{{p}}{{2}}}((0,1);({{\mathbb{R}}}^{n})^{3})\,.

In particular, there is an α>0\alpha>0 such that (σ)(t)C2+α((0,1);(n)3)(\mathcal{E}\sigma)(t)\in C^{2+\alpha}((0,1);({{\mathbb{R}}}^{n})^{3}) for almost all t[t~,T]t\in[\tilde{t},T].

Proof.

Considering t((σ)i)(t)\partial_{t}((\mathcal{E}\sigma)^{i})(t) as given functions 𝔣iWp1((0,1);n)\mathfrak{f}^{i}\in W^{1}_{p}((0,1);{{\mathbb{R}}}^{n}) we see that (σ)i(t,)(\mathcal{E}\sigma)^{i}(t,\cdot) solves

(σ)xxi(t,)|(σ)xi(t,)|2=𝔣i\displaystyle\frac{(\mathcal{E}\sigma)^{i}_{xx}(t,\cdot)}{\left|(\mathcal{E}\sigma)^{i}_{x}(t,\cdot)\right|^{2}}=\mathfrak{f}^{i}

in Wp1((0,1);(2)3)W^{1}_{p}((0,1);({{\mathbb{R}}}^{2})^{3}) for almost every t[t~,T]t\in[\tilde{t},T]. As for almost every t[t~,T]t\in[\tilde{t},T] the function (σ)i(t,)(\mathcal{E}\sigma)^{i}(t,\cdot) is in Wp2((0,1);n)W^{2}_{p}((0,1);{{\mathbb{R}}}^{n}), we know that |(σ)xi(t,)|2W1,p/2((0,1);)\left|(\mathcal{E}\sigma)^{i}_{x}(t,\cdot)\right|^{2}\in W^{1,\nicefrac{{p}}{{2}}}((0,1);{{\mathbb{R}}}) for almost every t[t~,T]t\in[\tilde{t},T] and thus we can multiply the above equations to see that

(σ)xxi(t,)=𝔣~iWp/21((0,1);n)\displaystyle(\mathcal{E}\sigma)^{i}_{xx}(t,\cdot)=\tilde{\mathfrak{f}}^{i}\in W^{1}_{\nicefrac{{p}}{{2}}}((0,1);{{\mathbb{R}}}^{n})

for almost every t[t~,T]t\in[\tilde{t},T] with new given inhomogeneities 𝔣~i\tilde{\mathfrak{f}}^{i}. Note that we used here that with our choice of pp the Sobolev space Wp/21W^{1}_{\nicefrac{{p}}{{2}}} is indeed a Banach algebra on one dimensional domains. But from the last equation we directly conclude (σ)iWp/23((0,1);(n)3)(\mathcal{E}\sigma)^{i}\in W^{3}_{\nicefrac{{p}}{{2}}}((0,1);({{\mathbb{R}}}^{n})^{3}). The second claim is just a direct consequence of the Sobolev embeddings. ∎

With the two previous results we are now able to start a bootstrap procedure.

Theorem 4.8 (Smoothness of solutions to the Special Flow).

Let σ𝐄T\mathcal{E}\sigma\in\boldsymbol{E}_{T} be a solution to the Special Flow in [0,T][0,T] with T>0T>0 and initial value σWp22/p((0,1);(n)3)\sigma\in W_{p}^{2-\nicefrac{{2}}{{p}}}\left((0,1);(\mathbb{R}^{n})^{3}\right). Then σ\mathcal{E}\sigma is smooth on [t~,T]×[0,1][\tilde{t},T]\times[0,1] for all t~(0,T)\tilde{t}\in(0,T).

Proof.

Due to Corollary 4.7 we can use (σ)(t)(\mathcal{E}\sigma)(t) for almost all t>0t>0 as initial data for a regularity result in parabolic Hölder space, cf. [13] for such a result for the Willmore flow. As we checked that the Lopatinskii-Shapiro conditions are still valid in higher co-dimensions, the analysis works as in the planar case. Additionally, the needed compatibility conditions due to the zero order boundary conditions are guaranteed by the fact that t(σ)\partial_{t}(\mathcal{E}\sigma) lies in C([t~,T];C([0,1];(n)3)C([\tilde{t},T];C([0,1];({{\mathbb{R}}}^{n})^{3}). With this new maximal regularity result, which is the key argument in the proof of Proposition 4.6, we can repeat the whole procedure to derive C3+α,(3+α)/2C^{3+\alpha,(3+\alpha)/2}-regularity. This starts now the bootstrapping yielding the desired smoothness result. Note that in every step the needed compatibility conditions are guaranteed by the fact that our flow already has the regularity related to these compatibility conditions (see for instance [23, Theorem 3.1]). ∎

In analogy to [14] we may now use smoothness of the Special Flow to prove Theorem 1.1.

Proof of Theorem 1.1.

The existence of maximal solutions and their geometric uniqueness are shown in Proposition 3.20. Using smoothness of the Special Flow shown in Theorem 4.8 one may argue analogously to [14, Section 5.2, Section 7.2] to show that parametrising each curve 𝕋i(t)\mathbb{T}^{i}(t) with constant speed equal to the length of 𝕋i(t)\mathbb{T}^{i}(t) yields a global parametrisation γ:[0,Tmax)×[0,1](n)3\gamma:[0,T_{max})\times[0,1]\to(\mathbb{R}^{n})^{3} of the evolution that is smooth for positive times. ∎

5 Long Time Behaviour of the Motion by Curvature

Proof of Theorem 1.2.

Let ε(0,Tmax/1000)\varepsilon\in(0,\nicefrac{{T_{\max}}}{{1000}}) be fixed. Suppose that TmaxT_{\max} is finite and that the two assertions i)i) and ii)ii) are not fulfilled. Let γ=(γ1,γ2,γ3):[0,Tmax)×[0,1](n)3\gamma=(\gamma^{1},\gamma^{2},\gamma^{3}):[0,T_{max})\times[0,1]\to(\mathbb{R}^{n})^{3} be the parametrisation of the evolution such that each curve 𝕋i(t)\mathbb{T}^{i}(t) is parametrised with constant speed equal to its length L(𝕋i(t))L(\mathbb{T}^{i}(t)). As γ\gamma is smooth on [ε,T][\varepsilon,T] for all positive ε\varepsilon and all T(ε,Tmax)T\in(\varepsilon,T_{\max}), hypothesis ii)ii) yields

𝜿iL((ε,Tmax);L2((0,1);n)).\boldsymbol{\kappa}^{i}\in L^{\infty}\left((\varepsilon,T_{\max});L^{2}((0,1);\mathbb{R}^{n})\right)\,.

As 𝑬T\boldsymbol{E}_{T} embeds continuously into C([0,T];C1([0,1];(n)3))C\left([0,T];C^{1}([0,1];(\mathbb{R}^{n})^{3})\right), hypothesis i)i) implies that the lengths L(𝕋i)L(\mathbb{T}^{i}) of all three curves composing the network are uniformly bounded away from zero in [0,Tmax)[0,T_{\max}). Moreover, thanks to the gradient flow structure of the motion by curvature the single lengths of the networks satisfy L(𝕋i(t))L(𝕋0)L(\mathbb{T}^{i}(t))\leq L(\mathbb{T}_{0}) for all t[0,Tmax)t\in[0,T_{\max}). In particular, we obtain for all t[0,Tmax)t\in[0,T_{max}), x[0,1]x\in[0,1],

0<𝔠|γxi(t,x)|=L(𝕋i(t))C<.0<\mathfrak{c}\leq|\gamma^{i}_{x}(t,x)|=L(\mathbb{T}^{i}(t))\leq{C}<\infty\,. (5.1)

With our choice of parametrisation the curvature can be expressed as 𝜿i=γxxi/L(𝕋i)2\boldsymbol{\kappa}^{i}=\nicefrac{{\gamma^{i}_{xx}}}{{L(\mathbb{T}^{i})^{2}}} from which we can infer for all t[0,Tmax)t\in[0,T_{max}),

01|γxxi|2dx=(L(𝕋i))3𝕋|𝜿i|2dsC<.\int_{0}^{1}|\gamma^{i}_{xx}|^{2}\,\mathrm{d}x=\left(L(\mathbb{T}^{i})\right)^{3}\int_{\mathbb{T}}|\boldsymbol{\kappa}^{i}|^{2}\,\mathrm{d}s\leq C<\infty\,.

As the endpoints P1P^{1}, P2P^{2}, P3P^{3} are fixed and as the single lengths L(𝕋i(t))L(\mathbb{T}^{i}(t)) are uniformly bounded from above in [0,Tmax)[0,T_{max}), there exists a constant R>0R>0 such that for every t[0,Tmax)t\in[0,T_{\max}) it holds 𝕋(t)BR(0)\mathbb{T}(t)\subset B_{R}(0). With the above arguments we conclude

γiL((ε,Tmax);W22((0,1);n)).\gamma^{i}\in L^{\infty}((\varepsilon,T_{\max});W^{2}_{2}((0,1);\mathbb{R}^{n}))\,.

The Sobolev Embedding Theorem implies for all p(3,6]p\in(3,6] the estimate

supt(ε,Tmax)γi(t)Wp22/p((0,1);n)𝑪\sup_{t\in(\varepsilon,T_{max})}\left\lVert\gamma^{i}(t)\right\rVert_{W^{2-\nicefrac{{2}}{{p}}}_{p}\left((0,1);\mathbb{R}^{n}\right)}\leq\boldsymbol{C} (5.2)

for a uniform constant 𝑪>0\boldsymbol{C}>0. We note further that for all δ(0,Tmax/4)\delta\in\left(0,\nicefrac{{T_{max}}}{{4}}\right) the parametrisation γ(Tmaxδ)\gamma(T_{max}-\delta) is an admissible initial value for the Special Flow (2.4). Due to (5.1) and (5.2) Theorem 3.13 yields that there exists a uniform time 𝑻\boldsymbol{T} of existence of solutions to the Special Flow (2.4) for all initial values γ(Tmaxδ)\gamma(T_{max}-\delta) depending on 𝑪\boldsymbol{C} and 𝔠\mathfrak{c}. Let δ:=min{𝑻/2,Tmax/4}\delta:=\min\left\{\nicefrac{{\boldsymbol{T}}}{{2}},\nicefrac{{T_{max}}}{{4}}\right\}. Then Theorem 3.13 implies the existence of a solution η=(η1,η2,η3)\eta=(\eta^{1},\eta^{2},\eta^{3}) with ηi\eta^{i} regular and

ηiWp1((Tmaxδ,Tmax+δ);Lp((0,1);n))Lp((Tmaxδ,Tmax+δ);Wp2((0,1);n))\eta^{i}\in W_{p}^{1}\left((T_{max}-\delta,T_{max}+\delta);L_{p}\left((0,1);\mathbb{R}^{n}\right)\right)\cap L_{p}\left((T_{max}-\delta,T_{max}+\delta);W_{p}^{2}\left((0,1);\mathbb{R}^{n}\right)\right)

to system (2.4) with η(Tmaxδ)=γ(Tmaxδ)\eta\left(T_{max}-\delta\right)=\gamma\left(T_{max}-\delta\right). The two parametrisations γ\gamma and η\eta defined on (0,Tmaxδ3)(0,T_{max}-\frac{\delta}{3}) and (Tmaxδ2,Tmax+δ)\left(T_{max}-\frac{\delta}{2},T_{max}+\delta\right), respectively, define a solution (𝕋~(t))(\widetilde{\mathbb{T}}(t)) to the motion by curvature on the time interval (0,Tmax+δ](0,T_{max}+\delta] with initial datum 𝕋0\mathbb{T}_{0} coinciding with 𝕋\mathbb{T} on (0,Tmax)(0,T_{max}). This contradicts the maximality of TmaxT_{max}. ∎

References

  • [1] S. Agmon, A. Douglis, and L. Nirenberg, Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions. I, Comm. Pure Appl. Math. 12 (1959), 623–727. MR 0125307
  • [2] S. Angenent, Parabolic equations for curves on surfaces. I. Curves with pp-integrable curvature, Ann. of Math. (2) 132 (1990), no. 3, 451–483.
  • [3]   , On the formation of singularities in the curve shortening flow, J. Diff. Geom. 33 (1991), 601–633.
  • [4]   , Parabolic equations for curves on surfaces. II. Intersections, blow–up and generalized solutions, Ann. of Math. (2) 133 (1991), no. 1, 171–215.
  • [5] K. A. Brakke, The motion of a surface by its mean curvature, Princeton University Press, NJ, 1978.
  • [6] L. Bronsard and F. Reitich, On three-phase boundary motion and the singular limit of a vector-valued Ginzburg-Landau equation, Archive for Rational Mechanics and Analysis 124 (1993), no. 4, 355–379.
  • [7] R. Denk, J. Saal, and J. Seiler, Inhomogeneous symbols, the Newton polygon, and maximal LpL^{p}-regularity, Russ. J. Math. Phys. 15 (2008), no. 2, 171–191. MR 2410829
  • [8] S. D. Eidelman and N. V. Zhitarashu, Parabolic boundary value problems, Operator Theory: Advances and Applications, vol. 101, Birkhäuser Verlag, Basel, 1998, Translated from the Russian original by Gennady Pasechnik and Andrei Iacob. MR 1632789
  • [9] M. Gage, An isoperimetric inequality with applications to curve shortening, Duke Math. J. 50 (1983), no. 4, 1225–1229.
  • [10]   , Curve shortening makes convex curves circular, Invent. Math. 76 (1984), 357–364.
  • [11] M. Gage and R. S. Hamilton, The heat equation shrinking convex plane curves, J. Diff. Geom. 23 (1986), 69–95.
  • [12] H. Garcke and M. Gößwein, Non-linear Stability of Double Bubbles under Surface Diffusion, preprint 2019, arXiv:1910.01041.
  • [13] H. Garcke, J. Menzel, and A. Pluda, Willmore flow of planar networks, J. Differential Equations 266 (2019), no. 4, 2019–2051. MR 3906239
  • [14]   , Long time existence of solutions to an elastic flow of networks, Comm. Partial Differential Equations, doi:10.1080/03605302.2020.1771364 (2020).
  • [15] M. A. Grayson, The heat equation shrinks embedded plane curves to round points, J. Diff. Geom. 26 (1987), 285–314.
  • [16] G. Huisken, A distance comparison principle for evolving curves, Asian J. Math. 2 (1998), 127–133.
  • [17] T. Ilmanen, A. Neves, and F. Schulze, On short time existence for the planar network flow, J. Differential Geom. 111 (2019), no. 1, 39–89. MR 3909904
  • [18] J. M. Lee, Introduction to smooth manifolds, second ed., Graduate Texts in Mathematics, vol. 218, Springer, New York, 2013. MR 2954043
  • [19] A. Lunardi, Analytic semigroups and optimal regularity in parabolic problems, Birkhäuser, Basel, 1995.
  • [20]   , Nonlinear parabolic equations and systems, Evolutionary equations. Vol. I, Handb. Differ. Equ., North-Holland, Amsterdam, 2004, pp. 385–436.
  • [21] A. Magni, C. Mantegazza, and M. Novaga, Motion by curvature of planar networks II, Ann. Sc. Norm. Sup. Pisa 15 (2016), 117–144.
  • [22] C. Mantegazza, M. Novaga, A. Pluda, and F. Schulze, Evolution of networks with multiple junctions, preprint 2016.
  • [23] C. Mantegazza, M. Novaga, and V. M. Tortorelli, Motion by curvature of planar networks, Ann. Sc. Norm. Sup. Pisa 3 (5) (2004), 235–324.
  • [24] A. Pluda, Evolution of spoon–shaped networks, Networks and Heterogeneus Media 11 (2016), no. 3, 509–526.
  • [25] J. Prüss and G. Simonett, Moving interfaces and quasilinear parabolic evolution equations, Monographs in Mathematics, vol. 105, Birkhäuser/Springer, [Cham], 2016. MR 3524106
  • [26] J. Simon, Sobolev, Besov and Nikolskii fractional spaces: imbeddings and comparisons for vector valued spaces on an interval, Ann. Mat. Pura Appl. (4) 157 (1990), 117–148. MR 1108473
  • [27] V. A. Solonnikov, Boundary value problems of mathematical physics. III, Proceedings of the Steklov Institute of Mathematics, no. 83 (1965), Amer. Math. Soc., Providence, R.I., 1967.
  • [28] A. Stahl, Convergence of solutions to the mean curvature flow with a Neumann boundary condition, Calc. Var. Partial Differential Equations 4 (1996), no. 5, 421–441.
  • [29]   , Regularity estimates for solutions to the mean curvature flow with a Neumann boundary condition, Calc. Var. Partial Differential Equations 4 (1996), no. 4, 385–407.
  • [30] H. Triebel, Interpolation Theory, Function Spaces, Differential Operators, North-Holland Mathematical Library, vol. 18, North-Holland Publishing Co., Amsterdam-New York, 1978.
  • [31] K. Yosida, Functional Analysis, Classics in Mathematics, Springer-Verlag, Berlin, 1995, Reprint of the sixth (1980) edition. MR 1336382