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

Convergence of Sobolev gradient trajectories to elastica

Shinya Okabe Mathematical Institute, Tohoku University, Aoba, Sendai 980-8578, Japan. [email protected]  and  Philip Schrader Mathematical Institute, Tohoku University, Aoba, Sendai 980-8578, Japan. [email protected]
Abstract.

In this paper we study the H2(ds)H^{2}(ds)-gradient flow for the modified elastic energy defined on closed immersed curves in n\mathbb{R}^{n}. We prove the existence of a unique global-in-time solution to the flow and establish full convergence to elastica by way of a Łojasiewicz–Simon gradient inequality.

Mathematics subject classification (2020): 53E99, 58E99, 58B20

Key words and phrases:
H2(ds)H^{2}(ds) Sobolev gradient flow, elastic energy, Łojasiewicz–Simon gradient inequality, full convergence.
2020 Mathematics Subject Classification:
Primary: 53E99, Secondary: 58E99, 58B20
The first author is supported in part by JSPS KAKENHI Grant Numbers JP19H05599, JP20KK0057 and JP21H00990.
The second author is supported by an International Postdoctoral Research Fellowship of the Japan Society for the Promotion of Science and JSPS KAKENHI Grant Number JP19F19710.

1. Introduction

This paper is concerned with a Sobolev gradient flow for the modified elastic energy defined on closed immersed curves γ:S1n\gamma:S^{1}\to\mathbb{R}^{n}:

(1) (γ):=γk2𝑑s+λ2(γ),\operatorname{\mathcal{E}}(\gamma):=\int_{\gamma}k^{2}\,ds+\lambda^{2}\mathcal{L}(\gamma),

where n2n\in\mathbb{N}_{\geq 2}; λ\lambda is a nonzero constant; (γ)\mathcal{L}(\gamma), kk and ss denote the length, the curvature and the arc length parameter of γ\gamma, respectively; and throughout the paper S1S^{1} is identified with [0,1][0,1] modulo endpoints. In the case n=2n=2 the critical points of \operatorname{\mathcal{E}} are the classical Euler–Bernoulli elastica.

A variety of gradient flows toward elastica have been studied by other authors, beginning with the curve-straightening flow studied by Langer and Singer for closed curves in 3\mathbb{R}^{3} [16] and Riemannian manifolds [17]. Langer and Singer consider the restriction of the total squared curvature to constant speed curves. Under this restriction the total squared curvature of a curve is equivalent to the Dirichlet energy of its tangent indicatrix. Langer and Singer show that the Dirichlet energy satisfies the Palais–Smale condition on a Hilbert manifold of tangent indicatrices of Sobolev class H1H^{1}. Consequences of this are existence for all positive time as well as sub-convergence (convergence of a subsequence) of the associated gradient flow, where the gradient is defined by the H1H^{1}-metric on indicatrices.

The work of Langer and Singer was extended in several directions by Linnér [19, 20] and also Wen who studied the L2L^{2}-gradient flow of Dirichlet energy on indicatrices [42] and then the L2(ds)L^{2}(ds)-gradient flow of the total squared curvature on constant speed planar curves [43]. In both cases Wen proved that in the case of smooth initial data with non-zero winding number, solutions exist for all positive time and converge to circles. Around the same time Koiso [14] obtained similar results for unit speed space curves, and Polden [34] considered the L2(ds)L^{2}(ds)-flow of the modified elastic energy \operatorname{\mathcal{E}} for planar curves without constraint. Polden found that given smooth initial data the flow exists globally and subconverges modulo translations to critical points of the energy. Dziuk et. al. [9] extended the results of Wen and Polden to closed curves in n\mathbb{R}^{n}.

More recently the L2(ds)L^{2}(ds)-gradient flow for \operatorname{\mathcal{E}} has become known as the elastic flow and has been studied on spaces of closed curves with area constraint [29, 30], open curves with boundary conditions [18, 6, 41], non-compact curves [27] and planar networks [7]. For a more complete list of references we refer to the recent survey [22]. Standard results are solvability of the elastic flow for smooth initial curve and subconvergence of solutions to elastica. There is room for improvement on both fronts. With initial data in the energy class H2H^{2}, well-posedness of the elastic flow with boundary conditions was proved in [37] by way of analytic semigroup theory; and for the more general pp-elastic flow (which is the elastic flow when p=2p=2) [2, 28, 31, 32] show existence of weak solutions by the minimizing movements method. Stronger convergence results are obtained in [8, 23] using Łojasiewicz–Simon gradient inequalities. We note however that except for some special cases of weak solutions to the pp-elastic flow, the existence results do not include uniqueness. Moreover, the convergence results are always up to reparametrisation and sometimes translation111[14, 29, 30] proved full convergence without any translation corrections due to the invariance of center of gravity of curves under the elastic flow with the inextensibility condition, but the inextensibility condition fixes parametrisation. In [23] translation assumptions are neatly avoided without constraints, but reparametrisation is still required. .

In this paper we introduce a new gradient flow for \operatorname{\mathcal{E}}. We prove the existence of unique global-in-time strong solutions starting from initial curves in the energy space H2H^{2} and full convergence of gradient trajectories to elastica without any corrections to translation or parametrisation. Namely, we consider the Cauchy problem for the H2(ds)H^{2}(ds)-gradient flow for \operatorname{\mathcal{E}} defined on closed curves in n\mathbb{R}^{n}:

(GF) {tγ=gradγ,γ(,0)=γ0().\displaystyle\begin{cases}&\partial_{t}\gamma=-\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma},\\ &\gamma(\cdot,0)=\gamma_{0}(\cdot).\end{cases}

Here, gradγ\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma} denotes the H2(ds)H^{2}(ds)-gradient for \operatorname{\mathcal{E}} at γ\gamma (for the precise definition and its expression, see Section 2), i.e. the gradient of \operatorname{\mathcal{E}} with respect to the inner product

(2) v,wH2(ds),γ:=γv,w𝑑s+γvs,ws𝑑s+γvss,wss𝑑s\left\langle{v,w}\right\rangle_{H^{2}(ds),\gamma}:=\int_{\gamma}\left\langle{v,w}\right\rangle\,ds+\int_{\gamma}\left\langle{v_{s},w_{s}}\right\rangle\,ds+\int_{\gamma}\left\langle{v_{ss},w_{ss}}\right\rangle ds

for variations v,wH2(S1,n)v,w\in H^{2}(S^{1},\mathbb{R}^{n}) along γ\gamma. We consider initial data γ0\gamma_{0} in the space of H2H^{2} immersions

2(S1,n):={γH2(S1,n):|γ(u)|>0}\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}):=\{\gamma\in H^{2}(S^{1},\mathbb{R}^{n}):|\gamma^{\prime}(u)|>0\}

which is an open subset of H2(S1,n)H^{2}(S^{1},\mathbb{R}^{n}) (see Lemma 3.1 (i)). The main result of this paper is as follows:

Theorem 1.1.

Let γ02(S1,n)\gamma_{0}\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}). Then problem (GF) possesses a unique global-in-time solution γ\gamma in the class C1([0,),2(S1,n))C^{1}([0,\infty),\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})). Moreover, the solution γ\gamma converges to an elastica as tt\to\infty in the H2H^{2}-topology.

Since it is defined by the H2(ds)H^{2}(ds)-metric, gradγ\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma} is also of class H2H^{2} and so (GF) is an ODE in the Sobolev space H2(S1,n)H^{2}(S^{1},\mathbb{R}^{n}). Thus we prove the existence of local-in-time solutions to problem (GF) by the use of the generalized (to Banach space) Picard–Lindelöf theorem (Proposition 3.2). Moreover, thanks to the metric completeness of the space 2(S1,n)\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) with respect to the H2(ds)H^{2}(ds)-Riemannian metric (see [4, Theorem 4.3]), the proof of the existence of global-in-time solutions follows from a simple adaptation of the method used in [33, Theorem 9.1.6].

The metric completeness is also a key ingredient in the proof of convergence: we use a Łojasiewicz–Simon gradient inequality to show that the H2(ds)H^{2}(ds)-length of a gradient trajectory is finite, and therefore converges by completeness. Proving that such an inequality holds for \operatorname{\mathcal{E}} is complicated by the fact that, due to reparametrisation invariance, the second derivative of \operatorname{\mathcal{E}} has an infinite dimensional nullspace and therefore cannot be Fredholm. This can be overcome by restricting the energy to a submanifold consisting of curves which form a cross-section of the reparametrisation symmetry, proving that a Łojasiewicz–Simon inequality holds for this restriction, and then extending the inequality by symmetry. This is the method used in e.g. [5],[8] and [23] where the chosen submanifold consists of normal graphs over the critical point. Here we instead restrict to the submanifold of arc length proportionally parametrised curves. Even in finite codimension, verifying a Łojasiewicz-Simon gradient inequality on a submanifold is far from trivial (see [36]), but here we are helped by the fact that the restriction of \operatorname{\mathcal{E}} to arc length proportionally parametrised curves takes a simple form.

We remark that the H2(ds)H^{2}(ds)-gradient flow for \operatorname{\mathcal{E}} is different from the flow considered by Langer and Singer [16, 17] and Linnér [19, 21]. These authors carry out their analysis using an H1H^{1}-metric on a space of indicatrices, which is effectively H2H^{2} on the space of curves but without measuring the zeroth order product of variations – the first summand in (2). As observed by Wen [43] and then clearly demonstrated by Linnér [21], different representations of the space of curves and different choices for the metric on these spaces result in geometrically distinct flows.

Finally, we mention that Sobolev gradient flows for other geometric functionals have been studied, e.g. the H1(ds)H^{1}(ds)-curve shortening flow [40], the H2H^{2}-elastic flow for graphs with an obstacle [25], and fractional Sobolev gradient flows for knot energies [35, 13]. We also mention the comprehensive book by Neuberger [26] which discusses many applications as well as the numerical advantages of Sobolev gradient flows.

The rest of this paper is organized as follows: In Section 2 we give the precise formulation of the H2(ds)H^{2}(ds)-gradient flow for \operatorname{\mathcal{E}} and introduce some notation. In Section 3 we prove the existence of a unique global-in-time solution of (GF). We prove the full convergence of solutions to (GF) as follows: in Section 4.2 we verify the analyticity of the functional \operatorname{\mathcal{E}}; in Section 4.3 we introduce the submanifold of arc length proportionally parametrised curves, and in Section 4.4 we prove a Łojasiewicz–Simon gradient inequality on this submanifold which we then extend to a Łojasiewicz–Simon gradient inequality on all of 2(S1,n)\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}); finally we prove full convergence of solutions to an elastica by way of the Łojasiewicz–Simon gradient inequality in Section 5.

Acknowledgements. The authors are very grateful to the anonymous referees for their careful reading and for many valuable suggestions, and the second author would like to thank Glen Wheeler for helpful discussions. Most of the work in this paper was completed while the second author was a JSPS postdoctoral fellow at Tohoku University, and he wishes to express his gratitude for the kind hospitality of the staff at Tohoku University and in the overseas fellowship division of the JSPS.

2. Formulation

2.1. The H2(ds)H^{2}(ds)-gradient

We derive the H2(ds)H^{2}(ds)-gradient for the modified elastic energy \operatorname{\mathcal{E}} (1) defined on closed immersed curves γ2(S1,n)\gamma\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}). For v,wH2(S1,n)v,w\in H^{2}(S^{1},\mathbb{R}^{n}) define the H2(ds)H^{2}(ds)-inner product by

(3) v,wH2(ds),γ:=v,wL2(ds),γ+vs,wsL2(ds),γ+vss,wssL2(ds),γ\left\langle{v,w}\right\rangle_{H^{2}(ds),\gamma}:=\left\langle{v,w}\right\rangle_{L^{2}(ds),\gamma}+\left\langle{v_{s},w_{s}}\right\rangle_{L^{2}(ds),\gamma}+\left\langle{v_{ss},w_{ss}}\right\rangle_{L^{2}(ds),\gamma}

with

(4) v,wL2(ds),γ:=γv,w𝑑s,\left\langle{v,w}\right\rangle_{L^{2}(ds),\gamma}:=\int_{\gamma}\left\langle{v,w}\right\rangle\,ds,

where ,\left\langle{\cdot,\cdot}\right\rangle denotes the Euclidean product. From now on we will omit the subscript γ\gamma from the H2(ds),L2(ds)H^{2}(ds),L^{2}(ds) products unless it is needed. Because they depend on the base curve γ\gamma the L2(ds)L^{2}(ds) and H2(ds)H^{2}(ds) products (3), (4) are Riemannian metrics on 2(S1,n)\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}). The H2(ds)H^{2}(ds)-norm is equivalent to the usual H2H^{2}-norm but the constants c1,c2c_{1},c_{2} in

c1vH2vH2(ds)c2vH2c_{1}{\left\|{v}\right\|}_{H^{2}}\leq{\left\|{v}\right\|}_{H^{2}(ds)}\leq c_{2}{\left\|{v}\right\|}_{H^{2}}

will depend on γ\gamma. Indeed from vs=v|γ|v_{s}=\frac{v^{\prime}}{{\left|{\gamma^{\prime}}\right|}} and vss=1|γ|2v′′γ′′,γ|γ|4vv_{ss}=\frac{1}{{\left|{\gamma^{\prime}}\right|}^{2}}v^{\prime\prime}-\frac{\left\langle{\gamma^{\prime\prime},\gamma^{\prime}}\right\rangle}{{\left|{\gamma^{\prime}}\right|}^{4}}v^{\prime}, setting c0=(min|γ|)1c_{0}=(\min{\left|{\gamma^{\prime}}\right|})^{-1}

vH2(ds)2\displaystyle{\left\|{v}\right\|}_{H^{2}(ds)}^{2} γLvL22+c0vL22+c03v′′L22+c07γL2γ′′L22vL2\displaystyle\leq{\left\|{\gamma^{\prime}}\right\|}_{L^{\infty}}{\left\|{v}\right\|}_{L^{2}}^{2}+c_{0}{\left\|{v^{\prime}}\right\|}^{2}_{L^{2}}+c_{0}^{3}{\left\|{v^{\prime\prime}}\right\|}^{2}_{L^{2}}+c_{0}^{7}{\left\|{\gamma^{\prime}}\right\|}_{L^{\infty}}^{2}{\left\|{\gamma^{\prime\prime}}\right\|}_{L^{2}}^{2}{\left\|{v^{\prime}}\right\|}_{L^{\infty}}^{2}
c22vH22.\displaystyle\leq c_{2}^{2}{\left\|{v}\right\|}_{H^{2}}^{2}.

The left inequality in the norm equivalence is similar.

It then follows from the Lax-Milgram theorem that H2(ds)H^{2}(ds) is a strong Riemannian metric, meaning that at each γ\gamma it gives an continuous linear isomorphism between H2(S1,n)H^{2}(S^{1},\mathbb{R}^{n}) and its dual. For contrast, the L2(ds)L^{2}(ds)-product does not - it is a weak Riemannian metric.

For the Gateaux derivative of \operatorname{\mathcal{E}} at γ2(S1,n)\gamma\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) we find

(5) dγv=ddε(γ+εv)|ε=0=0(γ)[2γss,vss3k2γs,vsλ2γss,v]dsd\operatorname{\mathcal{E}}_{\gamma}v=\dfrac{d}{d\varepsilon}\mathcal{E}(\gamma+\varepsilon v)\Bigl{|}_{\varepsilon=0}=\int^{\mathcal{L}(\gamma)}_{0}[2\left\langle{\gamma_{ss},v_{ss}}\right\rangle-3k^{2}\left\langle{\gamma_{s},v_{s}}\right\rangle-\lambda^{2}\left\langle{\gamma_{ss},v}\right\rangle]\,ds

for all vH2(S1;n)v\in H^{2}(S^{1};\mathbb{R}^{n}). Using Lemma 3.1 it is not too difficult to show that dγd\operatorname{\mathcal{E}}_{\gamma} is bounded in (H2)(H^{2})^{*}. Moreover, it depends continuously on γ\gamma and is therefore a Frechét derivative. Then since H2(ds)H^{2}(ds) is a strong metric there exists an H2(ds)H^{2}(ds)-gradient of \mathcal{E} at γ\gamma, meaning

(6) gradγ,vH2(ds)=0(γ)[2γss,vss3k2γs,vsλ2γss,v]𝑑s\left\langle{\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma},v}\right\rangle_{H^{2}(ds)}=\int^{\mathcal{L}(\gamma)}_{0}[2\left\langle{\gamma_{ss},v_{ss}}\right\rangle-3k^{2}\left\langle{\gamma_{s},v_{s}}\right\rangle-\lambda^{2}\left\langle{\gamma_{ss},v}\right\rangle]\,ds

for all vH2(S1;n)v\in H^{2}(S^{1};\mathbb{R}^{n}). To derive the explicit form of gradγ\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}, observe that (6) is the weak formulation of

(7) (gradγ)ssss(gradγ)ss+gradγ=γ(\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma})_{ssss}-(\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma})_{ss}+\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}=\nabla\operatorname{\mathcal{E}}_{\gamma}

where γ=2γssss+3(k2γs)sλ2γss\nabla\operatorname{\mathcal{E}}_{\gamma}=2\gamma_{ssss}+3(k^{2}\gamma_{s})_{s}-\lambda^{2}\gamma_{ss} is the classical L2(ds)L^{2}(ds)-gradient.

To solve (7) we derive the Green’s function, i.e. the solution to

(8) Gxxxx(x,y)Gxx(x,y)+G(x,y)=δ(xy)G_{xxxx}(x,y)-G_{xx}(x,y)+G(x,y)=\delta(x-y)

which is C2C^{2} periodic. Using the general solution to the homogeneous equation we set

G(x,y)=\displaystyle G(x,y)=
{b1e32xcosx2+b2e32xcosx2+b3e32xsinx2+b4e32xsinx2ifx<y,c1e32xcosx2+c2e32xcosx2+c3e32xsinx2+c4e32xsinx2ifx>y.\displaystyle\begin{cases}b_{1}e^{\frac{\sqrt{3}}{2}x}\cos\tfrac{x}{2}+b_{2}e^{\frac{-\sqrt{3}}{2}x}\cos\tfrac{x}{2}+b_{3}e^{\frac{\sqrt{3}}{2}x}\sin\tfrac{x}{2}+b_{4}e^{\frac{-\sqrt{3}}{2}x}\sin\tfrac{x}{2}&\text{if}\quad x<y,\\ c_{1}e^{\frac{\sqrt{3}}{2}x}\cos\frac{x}{2}+c_{2}e^{\frac{-\sqrt{3}}{2}x}\cos\frac{x}{2}+c_{3}e^{\frac{\sqrt{3}}{2}x}\sin\frac{x}{2}+c_{4}e^{\frac{-\sqrt{3}}{2}x}\sin\frac{x}{2}&\text{if}\quad x>y.\end{cases}

In order to solve for the unknown parameters bib_{i} and cic_{i} we enforce periodic boundary conditions 1iG(0,y)=1iG((γ),y)\partial_{1}^{i}G(0,y)=\partial_{1}^{i}G(\mathcal{L}(\gamma),y) for i=0,1,2,3i=0,1,2,3, continuity at x=yx=y of 1iG(x,y)\partial_{1}^{i}G(x,y) for i=0,1,2i=0,1,2, and the third order discontinuity

limxy+13G(x,y)limxy13G(x,y)=1.\lim_{x\to y^{+}}\partial_{1}^{3}G(x,y)-\lim_{x\to y^{-}}\partial_{1}^{3}G(x,y)=1.

We find:

(9) G(x,y;γ)=A((γ)|xy|,|xy|)β((γ)),0x,y(γ),G(x,y;\gamma)=\frac{A(\mathcal{L}(\gamma)-{\left|{x-y}\right|},|x-y|)}{\beta(\mathcal{L}(\gamma))},\qquad 0\leq x,y\leq\mathcal{L}(\gamma),

where

A(x1,x2)\displaystyle A(x_{1},x_{2}) =sinh3x12cosx22+sinh3x22cosx12\displaystyle=\sinh\frac{\sqrt{3}x_{1}}{2}\cos\frac{x_{2}}{2}+\sinh\frac{\sqrt{3}x_{2}}{2}\cos\frac{x_{1}}{2}
(10) +3cosh3x12sinx22+3cosh3x22sinx12,\displaystyle\qquad+\sqrt{3}\cosh\frac{\sqrt{3}x_{1}}{2}\sin\frac{x_{2}}{2}+\sqrt{3}\cosh\frac{\sqrt{3}x_{2}}{2}\sin\frac{x_{1}}{2},
β()\displaystyle\beta(\ell) =23(cosh32cos2).\displaystyle=2\sqrt{3}\Bigl{(}\cosh\frac{\sqrt{3}\ell}{2}-\cos\frac{\ell}{2}\Bigr{)}.

Now, setting

s:=sγ(u)=0u|γ(ξ)|𝑑ξ,s~:=sγ(u~)=0u~|γ(ξ)|𝑑ξ,foru,u~S1,\displaystyle s:=s_{\gamma}(u)=\int^{u}_{0}|\gamma^{\prime}(\xi)|\,d\xi,\quad\tilde{s}:=s_{\gamma}(\tilde{u})=\int^{\tilde{u}}_{0}|\gamma^{\prime}(\xi)|\,d\xi,\quad\text{for}\quad u,\tilde{u}\in S^{1},

the solution to (6) is

gradγ(s)\displaystyle\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}(s) =0(γ)G(s,s~;γ)(γ)(s~)𝑑s~\displaystyle=\int_{0}^{\mathcal{L}(\gamma)}G(s,\tilde{s};\gamma)\nabla\mathcal{E}(\gamma)(\tilde{s})d\tilde{s}
=0(γ)G[2γs~s~s~s~+3(k2γs~)s~λ2γs~s~]𝑑s~\displaystyle=\int_{0}^{\mathcal{L}(\gamma)}G[2\gamma_{\tilde{s}\tilde{s}\tilde{s}\tilde{s}}+3(k^{2}\gamma_{\tilde{s}})_{\tilde{s}}-\lambda^{2}\gamma_{\tilde{s}\tilde{s}}]\,d\tilde{s}
=0(γ)[2Gs~s~s~s~γGs~(3k2λ2)γs~]𝑑s~\displaystyle=\int_{0}^{\mathcal{L}(\gamma)}[2G_{\tilde{s}\tilde{s}\tilde{s}\tilde{s}}\gamma-G_{\tilde{s}}(3k^{2}-\lambda^{2})\gamma_{\tilde{s}}]\,d\tilde{s}
=2γ(s)+0(γ)[2(Gs~s~G)γGs~(3k2λ2)γs~]𝑑s~\displaystyle=2\gamma(s)+\int_{0}^{\mathcal{L}(\gamma)}[2(G_{\tilde{s}\tilde{s}}-G)\gamma-G_{\tilde{s}}(3k^{2}-\lambda^{2})\gamma_{\tilde{s}}]\,d\tilde{s}
=2γ(s)0(γ)[2Gγ+Gs~γs~(3k2+2λ2)]𝑑s~,\displaystyle=2\gamma(s)-\int_{0}^{\mathcal{L}(\gamma)}[2G\gamma+G_{\tilde{s}}\gamma_{\tilde{s}}(3k^{2}+2-\lambda^{2})]\,d\tilde{s},

where we have used (8) and the relation G(s,s~;γ)=G(s~,s;γ)G(s,\tilde{s};\gamma)=G(\tilde{s},s;\gamma). Thus we obtain

(11) gradγ(u)=2γ(u)01[\displaystyle\operatorname{grad}\mathcal{E}_{\gamma}(u)=2\gamma(u)-\int_{0}^{1}\Bigl{[} 2G(s,s~;γ)γ(u~)|γ(u~)|\displaystyle 2G(s,\tilde{s};\gamma)\gamma(\tilde{u})|\gamma^{\prime}(\tilde{u})|
+1|γ(u~)|u~G(s,s~;γ)γ(u~)(3k(u~)2+2λ2)]du~.\displaystyle+\frac{1}{|\gamma^{\prime}(\tilde{u})|}\partial_{\tilde{u}}G(s,\tilde{s};\gamma)\gamma^{\prime}(\tilde{u})(3k(\tilde{u})^{2}+2-\lambda^{2})\Bigr{]}\,d\tilde{u}.

By definition gradγH2(S1,n)\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}\in H^{2}(S^{1},\mathbb{R}^{n}) (this can also be checked directly from (11), as in Lemma 3.1 (iv)) and so problem (GF) is an ODE in 2(S1,n)\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}).

2.2. Metric completeness

The H2(ds)H^{2}(ds)-Riemannian distance between γ,β2(S1,n)\gamma,\beta\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) is

dist(γ,β):=infp01p(t)H2(ds)𝑑t\operatorname{dist}(\gamma,\beta):=\inf_{p}\int_{0}^{1}{\left\|{p^{\prime}(t)}\right\|}_{H^{2}(ds)}\,dt

where the infimum is taken over all piecewise C1C^{1} paths p:[0,1]2(S1,n)p:[0,1]\to\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) with p(0)=γp(0)=\gamma and p(1)=βp(1)=\beta. By Theorem 1.9.5 in [12], since H2(ds)H^{2}(ds) is a strong Riemannian metric222Note that the definition of a Riemannian metric used in [12] includes the assumption that it is strong. the distance function defines a metric on 2(S1,n)\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) whose topology coincides with the H2H^{2}-topology.

Lemma 2.1 ([4], Lemma 4.2).

Write Brdist(γ0)B^{\operatorname{dist}}_{r}(\gamma_{0}) for the open ball with radius rr with respect to the H2(ds)H^{2}(ds)-Riemannian distance.

  1. (i)

    Given γ02(S1,n)\gamma_{0}\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) there exist r>0r>0 and C>0C>0 such that

    dist(γ1,γ2)Cγ1γ2H2\operatorname{dist}(\gamma_{1},\gamma_{2})\leq C{\left\|{\gamma_{1}-\gamma_{2}}\right\|}_{H^{2}}

    for all γ1,γ2Brdist(γ0)\gamma_{1},\gamma_{2}\in B^{\operatorname{dist}}_{r}(\gamma_{0}).

  2. (ii)

    Given Brdist(γ0)2(S1,n)B^{\operatorname{dist}}_{r}(\gamma_{0})\subset\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) there exists C>0C>0 such that

    γ1γ2H2Cdist(γ1,γ2){\left\|{\gamma_{1}-\gamma_{2}}\right\|}_{H^{2}}\leq C\operatorname{dist}(\gamma_{1},\gamma_{2})

    for all γ1,γ2Brdist(γ0)\gamma_{1},\gamma_{2}\in B^{\operatorname{dist}}_{r}(\gamma_{0}).

Theorem 2.2 ([4], Theorem 4.3).

(2(S1,n),dist)(\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}),{\rm dist}) is a complete metric space.

For contrast once again, the distance obtained from the L2(ds)L^{2}(ds)-metric does not give a complete metric space and in fact vanishes on any path component. See [24] and the discussion in [40].

2.3. Symmetries

Given a diffeomorphism ϕDiff(S1)\phi\in\text{Diff}(S^{1}) we have the action by reparametrisation Φ:2(S1,n)2(S1,n)\Phi:\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})\to\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}), Φγ:=γϕ\Phi\gamma:=\gamma\circ\phi. Since Φ\Phi is linear we have dΦ=Φd\Phi=\Phi. Along the same lines as in [40] Section 3, this action is an isometry of the H2(ds)H^{2}(ds)-metric. That is,

Φv,ΦwH2(ds),Φγ=v,wH2(ds),γ.\left\langle{\Phi v,\Phi w}\right\rangle_{H^{2}(ds),\Phi\gamma}=\left\langle{v,w}\right\rangle_{H^{2}(ds),\gamma}.

Moreover, the energy \operatorname{\mathcal{E}} is invariant under reparametrisation: (γ)=(Φγ)\operatorname{\mathcal{E}}(\gamma)=\operatorname{\mathcal{E}}(\Phi\gamma) and therefore dγ=dΦγΦd\operatorname{\mathcal{E}}_{\gamma}=d\operatorname{\mathcal{E}}_{\Phi\gamma}\Phi. Applying the definition of the gradient and the isometry property, we have

dγv=gradγ,vH2(ds),γ=Φgradγ,ΦvH2(ds),Φγ.d\operatorname{\mathcal{E}}_{\gamma}v=\left\langle{\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma},v}\right\rangle_{H^{2}(ds),\gamma}=\left\langle{\Phi\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma},\Phi v}\right\rangle_{H^{2}(ds),\Phi\gamma}.

On the other side dΦγΦv=gradΦγ,ΦvH2(ds),Φγd\operatorname{\mathcal{E}}_{\Phi\gamma}\Phi v=\left\langle{\operatorname{grad}\operatorname{\mathcal{E}}_{\Phi\gamma},\Phi v}\right\rangle_{H^{2}(ds),\Phi\gamma} and equating the two we observe that Φgradγ=gradΦγ\Phi\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}=\operatorname{grad}\operatorname{\mathcal{E}}_{\Phi\gamma}. Using the isometry property again we have

(12) gradΦγH2(ds)=gradγH2(ds).{\left\|{\operatorname{grad}\operatorname{\mathcal{E}}_{\Phi\gamma}}\right\|}_{H^{2}(ds)}={\left\|{\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}}\right\|}_{H^{2}(ds)}.

By a similar argument, the above also holds when Φ\Phi is the map induced by a fixed translation in n\mathbb{R}^{n}.

3. Existence and uniqueness

From now on, we set

(13) F(γ):=grad(γ)forγ2(S1,n).F(\gamma):=-\operatorname{grad}\mathcal{E}(\gamma)\quad\text{for}\quad\gamma\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}).

Moreover, we denote by CSC_{S} the Sobolev constant of the imbedding H1(S1)C12(S1)H^{1}(S^{1})\subset C^{\frac{1}{2}}(S^{1}):

fC12(S1)CSfH1(S1).\|f\|_{C^{\frac{1}{2}}(S^{1})}\leq C_{S}\|f\|_{H^{1}(S^{1})}.
Lemma 3.1.

Let γ02(S1,n)\gamma_{0}\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) and b=12minuS1|γ0(u)|.b=\tfrac{1}{2}\min_{u\in S^{1}}{\left|{\gamma^{\prime}_{0}(u)}\right|}. Then there exist positive constants c1,c2,c3c_{1},c_{2},c_{3} depending on γ0\gamma_{0} such that for all γ\gamma in the open H2H^{2}-ball U=Bb/CSH2(γ0):={γH2(S1,n):γγ0H2<b/Cs}U=B^{H^{2}}_{b/C_{S}}(\gamma_{0}):=\{\gamma\in H^{2}(S^{1},\mathbb{R}^{n}):{\left\|{\gamma-\gamma_{0}}\right\|}_{H^{2}}<b/C_{s}\}:

  1. (i)

    for all uS1u\in S^{1}: 0<c1<|γ(u)|<c20<c_{1}<{\left|{\gamma^{\prime}(u)}\right|}<c_{2}, and therefore c1<(γ)<c2c_{1}<\mathcal{L}(\gamma)<c_{2},

  2. (ii)

    T:UH1(S1,n),γγsT:U\to H^{1}(S^{1},\mathbb{R}^{n}),\gamma\mapsto\gamma_{s} is locally Lipschitz,

  3. (iii)

    κ:UL2(S1,n),γγss\kappa:U\to L^{2}(S^{1},\mathbb{R}^{n}),\gamma\mapsto\gamma_{ss} is locally Lipschitz, and kL2<c3{\left\|{k}\right\|}_{L^{2}}<c_{3},

  4. (iv)

    F(γ)H2c3{\left\|{F(\gamma)}\right\|}_{H^{2}}\leq c_{3},

  5. (v)

    DFγopc3{\left\|{DF_{\gamma}}\right\|}_{\operatorname{\rm{op}}}\leq c_{3}.

where op{\left\|{\cdot}\right\|}_{\operatorname{\rm{op}}} is the operator norm.

Proof.

By the Sobolev imbedding we have

||γ(u)||γ0(u)||γγ0C0CSγγ0H2<b\bigl{|}|\gamma^{\prime}(u)|-|\gamma_{0}^{\prime}(u)|\bigr{|}\leq\|\gamma^{\prime}-\gamma_{0}^{\prime}\|_{C^{0}}\leq C_{S}\|\gamma-\gamma_{0}\|_{H^{2}}<b

for all uS1u\in S^{1}, and then

0<|γ0(u)|b<|γ(u)|<b+γ0C00<|\gamma_{0}^{\prime}(u)|-b<|\gamma^{\prime}(u)|<b+\|\gamma_{0}^{\prime}\|_{C^{0}}

which proves (i).

Suppose γ,μ\gamma,\mu are in the open H2H^{2}-ball centered at γ0\gamma_{0} with radius b/CSb/C_{S}, then by (i)

(14) |T(γ)T(μ)|=|γ|γ|μ|μ||\displaystyle|T(\gamma)-T(\mu)|={\left|{\frac{\gamma^{\prime}}{{\left|{\gamma^{\prime}}\right|}}-\frac{\mu^{\prime}}{{\left|{\mu^{\prime}}\right|}}}\right|} =1|γ||μ||γ|μ|γ|γ|+γ|γ|μ|γ||\displaystyle=\frac{1}{|\gamma^{\prime}||\mu^{\prime}|}{\left|{\gamma^{\prime}|\mu^{\prime}|-\gamma^{\prime}|\gamma^{\prime}|+\gamma^{\prime}|\gamma^{\prime}|-\mu^{\prime}|\gamma^{\prime}|}\right|}
(15) 2c1|γμ|\displaystyle\leq\frac{2}{c_{1}}|\gamma^{\prime}-\mu^{\prime}|

and therefore T(γ)T(μ)L22c1γμL2\|T(\gamma)-T(\mu)\|_{L^{2}}\leq\frac{2}{c_{1}}\|\gamma^{\prime}-\mu^{\prime}\|_{L^{2}}. Showing that

T(γ)T(μ)L2cγμH1\|T(\gamma)^{\prime}-T(\mu)^{\prime}\|_{L^{2}}\leq c\|\gamma^{\prime}-\mu^{\prime}\|_{H^{1}}

uses basically the same method but with more terms. Thus (ii) follows.

From γss=γ′′|γ|2γ′′,γγ|γ|4\gamma_{ss}=\gamma^{\prime\prime}|\gamma^{\prime}|^{-2}-\left\langle{\gamma^{\prime\prime},\gamma^{\prime}}\right\rangle\gamma^{\prime}|\gamma^{\prime}|^{-4} we get

|κ|2=|γ′′|2|γ|4γ′′,γ2|γ|62|γ′′|2|γ|42c14|γ′′|2|\kappa|^{2}=\frac{|\gamma^{\prime\prime}|^{2}}{|\gamma^{\prime}|^{4}}-\frac{\left\langle{\gamma^{\prime\prime},\gamma^{\prime}}\right\rangle^{2}}{|\gamma^{\prime}|^{6}}\leq 2\frac{|\gamma^{\prime\prime}|^{2}}{|\gamma^{\prime}|^{4}}\leq\frac{2}{c_{1}^{4}}|\gamma^{\prime\prime}|^{2}

and then since |κ|=k|\kappa|=k:

k2𝑑u=|κ|2𝑑u2c14γ′′L222c14(bCS+γ0H2)2.\int k^{2}du=\int|\kappa|^{2}du\leq\frac{2}{c_{1}^{4}}\|\gamma^{\prime\prime}\|^{2}_{L^{2}}\leq\frac{2}{c_{1}^{4}}\Bigl{(}\frac{b}{C_{S}}+\|\gamma_{0}\|_{H^{2}}\Bigr{)}^{2}.

The proof that

κ(γ)κ(μ)L2cγμH2\|\kappa(\gamma)-\kappa(\mu)\|_{L^{2}}\leq c\|\gamma-\mu\|_{H^{2}}

is similar to (ii). Thus (iii) follows.

From now on the arguments of GG and its derivatives, when ommitted, should be taken to be s,s~;γs,\tilde{s};\gamma. By (11) and (13) we have

(16) F(γ)(u)\displaystyle F(\gamma)(u) =2γ(u)+01[2Gγ(u~)|γ(u~)|\displaystyle=-2\gamma(u)+\int_{0}^{1}\Bigl{[}2G\gamma(\tilde{u})|\gamma^{\prime}(\tilde{u})|
+1|γ(u~)|u~Gγ(u~)(3k(u~)2+2λ2)]du~,\displaystyle\qquad\qquad\qquad+\frac{1}{|\gamma^{\prime}(\tilde{u})|}\partial_{\tilde{u}}G\gamma^{\prime}(\tilde{u})\left(3k(\tilde{u})^{2}+2-\lambda^{2}\right)\Bigr{]}\,d\tilde{u},
(17) F(γ)(u)\displaystyle F(\gamma)^{\prime}(u) =2γ(u)+01[2uGγ(u~)|γ(u~)|\displaystyle=-2\gamma^{\prime}(u)+\int_{0}^{1}\Bigl{[}2\partial_{u}G\gamma(\tilde{u})|\gamma^{\prime}(\tilde{u})|
(18) +1|γ(u~)|uu~Gγ(u~)(3k(u~)2+2λ2)]du~,\displaystyle\qquad\qquad\qquad+\frac{1}{|\gamma^{\prime}(\tilde{u})|}\partial_{u}\partial_{\tilde{u}}G\gamma^{\prime}(\tilde{u})\left(3k(\tilde{u})^{2}+2-\lambda^{2}\right)\Bigr{]}\,d\tilde{u},
(19) F(γ)′′(u)\displaystyle F(\gamma)^{\prime\prime}(u) =2γ′′(u)+01[2u2Gγ(u~)|γ(u~)|\displaystyle=-2\gamma^{\prime\prime}(u)+\int_{0}^{1}\Bigl{[}2\partial^{2}_{u}G\gamma(\tilde{u})|\gamma^{\prime}(\tilde{u})|
(20) +1|γ(u~)|u2u~Gγ(u~)(3k(u~)2+2λ2)]du~.\displaystyle\qquad\qquad\qquad+\frac{1}{|\gamma^{\prime}(\tilde{u})|}\partial^{2}_{u}\partial_{\tilde{u}}G\gamma^{\prime}(\tilde{u})\left(3k(\tilde{u})^{2}+2-\lambda^{2}\right)\Bigr{]}\,d\tilde{u}.

Then, suppressing arguments in A((γ)|ss~|,|ss~|)A(\mathcal{L}(\gamma)-{\left|{s-{\tilde{s}}}\right|},{\left|{s-{\tilde{s}}}\right|}) and β((γ))\beta(\mathcal{L}(\gamma)), the derivatives of GG are

(21) uG\displaystyle\partial_{u}G =(21)Aβsgn(ss~)|γ(u)|,\displaystyle=\frac{(\partial_{2}-\partial_{1})A}{\beta}\operatorname{sgn}(s-\tilde{s})|\gamma^{\prime}(u)|,
(22) u~G\displaystyle\partial_{\tilde{u}}G =uG,\displaystyle=-\partial_{u}G,
(23) uu~G\displaystyle\partial_{u}\partial_{\tilde{u}}G =(21)2Aβ|γ(u~)||γ(u)|,\displaystyle=-\frac{(\partial_{2}-\partial_{1})^{2}A}{\beta}|\gamma^{\prime}(\tilde{u})||\gamma^{\prime}(u)|,
(24) u2G\displaystyle\partial^{2}_{u}G =(21)2Aβ|γ(u)|2+uGγ(u),γ′′(u)|γ(u)|2,\displaystyle=\frac{(\partial_{2}-\partial_{1})^{2}A}{\beta}|\gamma^{\prime}(u)|^{2}+\partial_{u}G\frac{\left\langle{\gamma^{\prime}(u),\gamma^{\prime\prime}(u)}\right\rangle}{|\gamma^{\prime}(u)|^{2}},
(25) u2u~G\displaystyle\partial^{2}_{u}\partial_{\tilde{u}}G =(21)3Aβsgn(ss~)|γ(u~)||γ(u)|2\displaystyle=-\frac{(\partial_{2}-\partial_{1})^{3}A}{\beta}\operatorname{sgn}(s-\tilde{s})|\gamma^{\prime}(\tilde{u})||\gamma^{\prime}(u)|^{2}
(21)2Aβ|γ(u~)|γ(u),γ′′(u)|γ(u)|,\displaystyle\qquad-\frac{(\partial_{2}-\partial_{1})^{2}A}{\beta}|\gamma^{\prime}(\tilde{u})|\frac{\left\langle{\gamma^{\prime}(u),\gamma^{\prime\prime}(u)}\right\rangle}{|\gamma^{\prime}(u)|},

where in the expressions for uu~G\partial_{u}\partial_{\tilde{u}}G and u2G\partial^{2}_{u}G we have used the fact, computed from (10), that (21)A((γ),0)(\partial_{2}-\partial_{1})A(\mathcal{L}(\gamma),0) vanishes. Now AA and β\beta are both smooth functions, and each of s,s~,(γ)s,\tilde{s},\mathcal{L}(\gamma) is LL^{\infty}-bounded by (i). Moreover, β((γ))\beta(\mathcal{L}(\gamma)) is bounded away from zero by (i), and so the quotients in the above expressions are all bounded. It follows that uG,u~G,\partial_{u}G,\partial_{\tilde{u}}G, and uu~G\partial_{u}\partial_{\tilde{u}}G are bounded and then from (16),(18), (21), (22) and (23), using (iii), we confirm that F(γ)H1{\left\|{F(\gamma)}\right\|}_{H^{1}} is bounded. From (24),(25) there exist positive constants C1,C2C_{1},C_{2} such that

|u2G(s,s~;γ)|\displaystyle|\partial_{u}^{2}G(s,\tilde{s};\gamma)| C1(|γ(u)|2+|γ′′(u)|),\displaystyle\leq C_{1}(|\gamma^{\prime}(u)|^{2}+|\gamma^{\prime\prime}(u)|),
|u2u~G(s,s~;γ)|\displaystyle|\partial_{u}^{2}\partial_{\tilde{u}}G(s,\tilde{s};\gamma)| C2|γ(u~)|(|γ(u)|2+|γ′′(u)|).\displaystyle\leq C_{2}|\gamma^{\prime}(\tilde{u})|(|\gamma^{\prime}(u)|^{2}+|\gamma^{\prime\prime}(u)|).

Using these inequalities with (20), we have

|F(γ)′′(u)|\displaystyle|F(\gamma)^{\prime\prime}(u)| 2|γ′′(u)|\displaystyle\leq 2|\gamma^{\prime\prime}(u)|
+01[2C1(|γ(u)|2+|γ′′(u)|)|γ(u~)||γ(u~)|+C2|γ(u~)|(|γ(u)|2+|γ′′(u)|)(3k(u~)2+2+λ2)]du~.\displaystyle\quad+\int_{0}^{1}\begin{multlined}\Bigl{[}2C_{1}\bigl{(}|\gamma^{\prime}(u)|^{2}+|\gamma^{\prime\prime}(u)|\bigr{)}|\gamma(\tilde{u})||\gamma^{\prime}(\tilde{u})|\\ +C_{2}|\gamma^{\prime}(\tilde{u})|\bigl{(}|\gamma^{\prime}(u)|^{2}+|\gamma^{\prime\prime}(u)|\bigr{)}\bigl{(}3k(\tilde{u})^{2}+2+\lambda^{2}\bigr{)}\Bigr{]}\,d\tilde{u}.\end{multlined}\Bigl{[}2C_{1}\bigl{(}|\gamma^{\prime}(u)|^{2}+|\gamma^{\prime\prime}(u)|\bigr{)}|\gamma(\tilde{u})||\gamma^{\prime}(\tilde{u})|\\ +C_{2}|\gamma^{\prime}(\tilde{u})|\bigl{(}|\gamma^{\prime}(u)|^{2}+|\gamma^{\prime\prime}(u)|\bigr{)}\bigl{(}3k(\tilde{u})^{2}+2+\lambda^{2}\bigr{)}\Bigr{]}\,d\tilde{u}.

Using (iii) again we have that F(γ)′′L2{\left\|{F(\gamma)^{\prime\prime}}\right\|}_{L^{2}} is bounded. Thus (iv) follows.

Let α:(a,b)2(S1,n)\alpha:(a,b)\to\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) be a variation through γ\gamma in the direction vH2(S1,n)v\in H^{2}(S^{1},\mathbb{R}^{n}), i.e α(0)=γ\alpha(0)=\gamma and α(0)=v\alpha^{\prime}(0)=v. Treating α\alpha as a function of two variables α(ε,u)\alpha(\varepsilon,u), we calculate

ε|αu||ε=0=αuε,αu|αu|1|ε=0=αεu,αs|ε=0=vu,γs.\partial_{\varepsilon}|\alpha_{u}|\Bigm{|}_{\varepsilon=0}=\left\langle{\alpha_{u\varepsilon},\alpha_{u}}\right\rangle|\alpha_{u}|^{-1}\Bigm{|}_{\varepsilon=0}=\left\langle{\alpha_{\varepsilon u},\alpha_{s}}\right\rangle\Bigm{|}_{\varepsilon=0}=\left\langle{v_{u},\gamma_{s}}\right\rangle.

Hence also at ε=0\varepsilon=0 we have εds=vs,γsds\partial_{\varepsilon}ds=\langle v_{s},\gamma_{s}\rangle ds. Similarly for any function ϕ\phi depending on α\alpha we find the following relation for commutation with the arc length derivative

ϕsε|ε=0=ϕεsαεs,αsϕs|ε=0=ϕεsvs,γsϕs.\phi_{s\varepsilon}\Bigm{|}_{\varepsilon=0}=\phi_{\varepsilon s}-\left\langle{\alpha_{\varepsilon s},\alpha_{s}}\right\rangle\phi_{s}\Bigm{|}_{\varepsilon=0}=\phi_{\varepsilon s}-\langle v_{s},\gamma_{s}\rangle\phi_{s}.

Using these formulae we calculate

(26) DFγv\displaystyle DF_{\gamma}v =εF(α)|ε=0\displaystyle=\partial_{\varepsilon}F(\alpha)\bigm{|}_{\varepsilon=0}
=2v+0(γ)[2DGγv+2Gv+2Gγvs~,γs~\displaystyle=-2v+\int_{0}^{\mathcal{L}(\gamma)}\Bigl{[}2DG_{\gamma}v+2Gv+2G\gamma\left\langle{v_{\tilde{s}},\gamma_{\tilde{s}}}\right\rangle
+((DGγv)s~γs~+Gs~vs~)(3k2+2λ2)\displaystyle\qquad\qquad\qquad\quad\,\,+\bigl{(}(DG_{\gamma}v)_{\tilde{s}}\gamma_{\tilde{s}}+G_{{\tilde{s}}}v_{\tilde{s}}\bigr{)}(3k^{2}+2-\lambda^{2})
+Gs~γs~(6vs~s~,γs~s~(15k2+2λ2)vs~,γs~)]ds~.\displaystyle\qquad\qquad\qquad\quad\,\,+G_{\tilde{s}}\gamma_{\tilde{s}}\bigl{(}6\left\langle{v_{{\tilde{s}}{\tilde{s}}},\gamma_{{\tilde{s}}{\tilde{s}}}}\right\rangle-(15k^{2}+2-\lambda^{2})\left\langle{v_{\tilde{s}},\gamma_{\tilde{s}}}\right\rangle\bigr{)}\Bigr{]}\,d{\tilde{s}}.

To see that DFγop{\left\|{DF_{\gamma}}\right\|}_{\operatorname{\rm{op}}} is bounded, note from (9) that G(x,y)G(x,y) and 2G(x,y)\partial_{2}G(x,y) are both continuous functions, and so by (i) we have LL^{\infty}-bounds for G(s,s~)G(s,\tilde{s}) and Gs~(s,s~)G_{{\tilde{s}}}(s,\tilde{s}). Since (iii) gives an L2L^{2}-bound for kk, it remains to show that DGγvDG_{\gamma}v and (DGγv)s(DG_{\gamma}v)_{s} are bounded. From sγ=0u|γ(τ)|𝑑τs_{\gamma}=\int_{0}^{u}|\gamma^{\prime}(\tau)|\,d\tau and (γ)=sγ(1)\mathcal{L}(\gamma)=s_{\gamma}(1) we calculate

Dsγv\displaystyle Ds_{\gamma}v =0uv,γ/|γ|𝑑τ,\displaystyle=\int_{0}^{u}\left\langle{v^{\prime},\gamma^{\prime}/|\gamma^{\prime}|}\right\rangle\,d\tau,
Dγv\displaystyle D\mathcal{L}_{\gamma}v =01v,γ/|γ|𝑑τ,\displaystyle=\int_{0}^{1}\left\langle{v^{\prime},\gamma^{\prime}/|\gamma^{\prime}|}\right\rangle\,d\tau,
D(|ss~|)γv\displaystyle D({\left|{s-\tilde{s}}\right|})_{\gamma}v =sgn(ss~)u~uv,γ/|γ|𝑑τ,\displaystyle=\operatorname{sgn}(s-{\tilde{s}})\int_{\tilde{u}}^{u}\left\langle{v^{\prime},\gamma^{\prime}/|\gamma^{\prime}|}\right\rangle\,d\tau,

and, assuming vH2=1{\left\|{v}\right\|}_{H^{2}}=1, each of these is LL^{\infty}-bounded. Then, suppressing arguments in AA and β\beta again we have

DGγv=1β1A(DγvD(|ss~|)γv)+1β2AD(|ss~|)γvββ2ADγv.DG_{\gamma}v=\frac{1}{\beta}\partial_{1}A\left(D\mathcal{L}_{\gamma}v-D({\left|{s-{\tilde{s}}}\right|})_{\gamma}v\right)+\frac{1}{\beta}\partial_{2}AD({\left|{s-\tilde{s}}\right|})_{\gamma}v-\frac{\beta^{\prime}}{\beta^{2}}AD\mathcal{L}_{\gamma}v.

Since AA and β\beta are smooth functions, and (γ)\mathcal{L}(\gamma) is bounded away from zero by (i), we now have that |DGγ||DG_{\gamma}| is bounded. Next we calculate

(DGγv)s\displaystyle(DG_{\gamma}v)_{s} =1β(21A12A)sgn(ss~)(DγvD(|ss~|)γv))+1β(2A1A)(D(|ss~|)γv)s+1β(22A12A)sgn(ss~)D(|ss~|)γvββ2(2A1A)sgn(ss~)Dγv,\displaystyle=\begin{multlined}\frac{1}{\beta}(\partial_{2}\partial_{1}A-\partial_{1}^{2}A)\operatorname{sgn}(s-{\tilde{s}})(D\mathcal{L}_{\gamma}v-D({\left|{s-{\tilde{s}}}\right|})_{\gamma}v))\\ +\frac{1}{\beta}(\partial_{2}A-\partial_{1}A)(D({\left|{s-{\tilde{s}}}\right|})_{\gamma}v)_{s}\\ +\frac{1}{\beta}(\partial_{2}^{2}A-\partial_{1}\partial_{2}A)\operatorname{sgn}(s-{\tilde{s}})D({\left|{s-{\tilde{s}}}\right|})_{\gamma}v\\ -\frac{\beta^{\prime}}{\beta^{2}}(\partial_{2}A-\partial_{1}A)\operatorname{sgn}(s-{\tilde{s}})D\mathcal{L}_{\gamma}v,\end{multlined}\frac{1}{\beta}(\partial_{2}\partial_{1}A-\partial_{1}^{2}A)\operatorname{sgn}(s-{\tilde{s}})(D\mathcal{L}_{\gamma}v-D({\left|{s-{\tilde{s}}}\right|})_{\gamma}v))\\ +\frac{1}{\beta}(\partial_{2}A-\partial_{1}A)(D({\left|{s-{\tilde{s}}}\right|})_{\gamma}v)_{s}\\ +\frac{1}{\beta}(\partial_{2}^{2}A-\partial_{1}\partial_{2}A)\operatorname{sgn}(s-{\tilde{s}})D({\left|{s-{\tilde{s}}}\right|})_{\gamma}v\\ -\frac{\beta^{\prime}}{\beta^{2}}(\partial_{2}A-\partial_{1}A)\operatorname{sgn}(s-{\tilde{s}})D\mathcal{L}_{\gamma}v,
(D(|ss~|)γv)s\displaystyle(D({\left|{s-\tilde{s}}\right|})_{\gamma}v)_{s} =sgn(ss~)vsγs.\displaystyle=\operatorname{sgn}(s-{\tilde{s}})v_{s}\gamma_{s}.

This implies that (DGγv)s(DG_{\gamma}v)_{s} is also LL^{\infty}-bounded when vH2=1{\left\|{v}\right\|}_{H^{2}}=1, and then we observe from (26) that DFγop\|DF_{\gamma}\|_{\operatorname{\rm{op}}} is bounded. Thus (v) follows, and the proof of Lemma 3.1 is complete. ∎

Proposition 3.2.

Let γ02(S1,n)\gamma_{0}\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}). Then there exists T>0T>0 such that problem (GF) possesses a unique solution in C1([0,T),2(S1,n))C^{1}([0,T),\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})).

Proof.

According to the Banach space Picard–Lindelöf theorem in [45, Theorem 3.A] it is sufficient to show that there is an H2H^{2}-ball containing γ0\gamma_{0} on which FF is bounded and Lipschitz. Using the ball from Lemma 3.1, FF is bounded by (iv) and is Lipschitz by (v) and the mean value inequality (e.g. see [15, Corollary. 4.2]). ∎

For long-time existence we will imitate the method used in e.g. [33, Theorem 9.1.6].

Lemma 3.3.

Assume that γ:(a,b)2(S1,n)\gamma:(a,b)\to\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) is a C1C^{1} curve with finite H2(ds)H^{2}(ds)-length. Then limtbγ(t)\lim_{t\to b}\gamma(t) exists in (2(S1,n),dist)(\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}),{\rm dist}).

Proof.

For any increasing sequence of times tibt_{i}\to b we claim that γi:=γ(ti)\gamma_{i}:=\gamma(t_{i}) is Cauchy in (2(S1,n),dist)(\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}),{\rm dist}). Suppose not, then there exists ε>0\varepsilon>0 such that for all N>0N>0 there exist j,k>Nj,k>N such that dist(γj,γk)>ε\operatorname{dist}(\gamma_{j},\gamma_{k})>\varepsilon. Since dist(γj,γk)<tjtkγtH2(ds)dt\operatorname{dist}(\gamma_{j},\gamma_{k})<\int_{t_{j}}^{t_{k}}{\left\|{\gamma_{t}}\right\|}_{H^{2}(ds)}\,dt, this contradicts the assumption that the length abγtH2(ds)𝑑t\int_{a}^{b}{\left\|{\gamma_{t}}\right\|}_{H^{2}(ds)}\,dt should be finite. By Theorem 2.2 we see that γi\gamma_{i} converges to some γb\gamma_{b}. Moreover, γb\gamma_{b} is unique: if we have another sequence γ~i:=γ(t~i)γ~b\tilde{\gamma}_{i}:=\gamma(\tilde{t}_{i})\to\tilde{\gamma}_{b} and we take t¯i\bar{t}_{i} to be the ordered union of tit_{i} and t~i\tilde{t}_{i} then γ(t¯i)\gamma(\bar{t}_{i}) must also converge, which requires γ~b=γb\tilde{\gamma}_{b}=\gamma_{b}. Finally, we have limtbγ(t)=γb\lim_{t\to b}\gamma(t)=\gamma_{b}, because if not then there exists an increasing sequence tjbt_{j}\to b such that γ(tj)\gamma(t_{j}) does not converge to γb\gamma_{b}. ∎

Proposition 3.4.

Assume that γ02(S1,n)\gamma_{0}\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}). Then problem (GF) possesses a unique global-in-time solution γ\gamma.

Proof.

Suppose γ:[0,T)2(S1,n)\gamma:[0,T)\to\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) is a maximal solution curve to γt=gradγ\gamma_{t}=-\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}. Abbreviating (t):=(γ(t))\operatorname{\mathcal{E}}(t):=\operatorname{\mathcal{E}}(\gamma(t)), since

(t)(0)=0t(τ)𝑑τ=0tgradγH2(ds)2𝑑τ\operatorname{\mathcal{E}}(t)-\operatorname{\mathcal{E}}(0)=\int_{0}^{t}\operatorname{\mathcal{E}}^{\prime}(\tau)\,d\tau=-\int_{0}^{t}\|\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}\|_{H^{2}(ds)}^{2}\,d\tau

we have

(27) 0TgradγH2(ds)2𝑑t(0),\int_{0}^{T}\|\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}\|_{H^{2}(ds)}^{2}\,dt\leq\operatorname{\mathcal{E}}(0),

and then using the Hölder inequality we obtain

0TgradγH2(ds)𝑑tT(0TgradγH2(ds)2𝑑t)1/2T(0).\int_{0}^{T}\|\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}\|_{H^{2}(ds)}\,dt\leq\sqrt{T}\Bigl{(}\int_{0}^{T}\|\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}\|_{H^{2}(ds)}^{2}\,dt\Bigr{)}^{1/2}\leq\sqrt{T\operatorname{\mathcal{E}}(0)}.

Observe that if TT is finite then the length of γ\gamma is finite, and then it follows from Lemma 3.3 that limtTγ(t)\lim_{t\to T}\gamma(t) exists in (2(S1,n),dist)(\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}),{\rm dist}). This contradicts the maximality of TT. Therefore Proposition 3.4 follows. ∎

4. Łojasiewicz–Simon gradient inequality

Our general strategy for proving convergence of the flow to stationary points is well known: upgrade the L2L^{2}-in-time estimate (27) to an L1L^{1} estimate using a Łojasiewicz–Simon gradient inequality, and then the length of any trajectory is finite (see e.g. Chapter 7 of [10] for abstract results). Typically the Łojasiewicz–Simon inequality requires that \operatorname{\mathcal{E}} should be analytic, which we verify in Section 4.2. The other standard requisite is a Fredholm second derivative, and so as mentioned in the introduction, an immediate obstruction is that the reparametrisation invariance of \operatorname{\mathcal{E}} implies that its second derivative has infinite dimensional nullspace, and is therefore not Fredholm. To get around this, we restrict \operatorname{\mathcal{E}} to the submanifold of arc length proportionally parametrized curves. We show that on this submanifold a gradient inequality is satisfied, and then use the reparametrisation symmetry to extend the inequality to the full space.

4.1. Regularity of critical points and the second variation

First we show that stationary points of \operatorname{\mathcal{E}} have higher regularity, and then we calculate the second derivative of \operatorname{\mathcal{E}}. The higher regularity of critical points will be required in order to show in Proposition 4.12 that the second derivative is Fredholm.

Proposition 4.1.

γH2(S1,n)\gamma\in H^{2}(S^{1},\mathbb{R}^{n}) is a stationary point of \operatorname{\mathcal{E}} iff s4γL2\partial_{s}^{4}\gamma\in L^{2} and

(28) 2s4γ+s(3k2γsλ2γs)=0a.e.,\displaystyle 2\partial^{4}_{s}\gamma+\partial_{s}(3k^{2}\gamma_{s}-\lambda^{2}\gamma_{s})=0\quad\text{a.e.},
(29) s2γ(0)=s2γ((γ)),s3γ(0)=s3γ((γ)).\displaystyle\partial^{2}_{s}\gamma(0)=\partial^{2}_{s}\gamma(\mathcal{L}(\gamma)),\quad\partial^{3}_{s}\gamma(0)=\partial^{3}_{s}\gamma(\mathcal{L}(\gamma)).
Proof.

Recalling (5) we have

(30) dγV=0[2γss,Vss(3k2λ2)γs,Vs]𝑑s,d\operatorname{\mathcal{E}}_{\gamma}V=\int^{\mathcal{L}}_{0}\bigl{[}2\left\langle{\gamma_{ss},V_{ss}}\right\rangle-(3k^{2}-\lambda^{2})\left\langle{\gamma_{s},V_{s}}\right\rangle\bigr{]}\,ds,

where :=(γ)\mathcal{L}:=\mathcal{L}(\gamma). Assume γ\gamma is a stationary point of \operatorname{\mathcal{E}}, i.e. dγV=0d\operatorname{\mathcal{E}}_{\gamma}V=0 for all VH2(S1,n)V\in H^{2}(S^{1},\mathbb{R}^{n}). Let yy and ww be the solutions to

(31) ys=3k2γsλ2γs,y(0)=0,y_{s}=3k^{2}\gamma_{s}-\lambda^{2}\gamma_{s},\qquad y(0)=0,
wss=2γss+y,w(0)=0,ws(0)=0,w_{ss}=2\gamma_{ss}+y,\qquad w(0)=0,\quad w_{s}(0)=0,

and define

w~(s):=w(s)s22w()s22(s)(ws()2w()),\tilde{w}(s):=w(s)-\frac{s^{2}}{\mathcal{L}^{2}}w(\mathcal{L})-\frac{s^{2}}{\mathcal{L}^{2}}(s-\mathcal{L})\Bigl{(}w_{s}(\mathcal{L})-\frac{2}{\mathcal{L}}w(\mathcal{L})\Bigr{)},

so that w~(0),w~(),w~s(0)\tilde{w}(0),\tilde{w}(\mathcal{L}),\tilde{w}_{s}(0) and w~s()\tilde{w}_{s}(\mathcal{L}) all vanish and w~H2(S1,n)\tilde{w}\in H^{2}(S^{1},\mathbb{R}^{n}). Moreover, w~ss=wssP\tilde{w}_{ss}=w_{ss}-P where Pss=0P_{ss}=0 and so 0P,w~ss𝑑s=0\int_{0}^{\mathcal{L}}\left\langle{P,\tilde{w}_{ss}}\right\rangle ds=0. Since γ\gamma is stationary dγw~=0d\operatorname{\mathcal{E}}_{\gamma}\tilde{w}=0, from which it follows that

0=02γss+y,w~ss𝑑s\displaystyle 0=\int_{0}^{\mathcal{L}}\left\langle{2\gamma_{ss}+y,\tilde{w}_{ss}}\right\rangle\,ds =02γss+yP,w~ss𝑑s\displaystyle=\int^{\mathcal{L}}_{0}\left\langle{2\gamma_{ss}+y-P,\tilde{w}_{ss}}\right\rangle\,ds
=0|2γss+yP|2𝑑s\displaystyle=\int_{0}^{\mathcal{L}}|2\gamma_{ss}+y-P|^{2}\,ds

and therefore

(32) 2γss+yP=0a.e. inS1.2\gamma_{ss}+y-P=0\quad\text{a.e. in}\quad S^{1}.

Considering only the normal component, since PP is smooth and from (31) yy is H1H^{1}, we observe that kk is also in H1H^{1}. But then yuu=6kkuγu+(3k2λ2)γuuy_{uu}=6kk_{u}\gamma_{u}+(3k^{2}-\lambda^{2})\gamma_{uu} is in L2L^{2}, so yy is H2H^{2} and so is kk. Note that since (32) gives no information about the tangential component of γuu\gamma_{uu} we cannot bootstrap any further333yuuu=(6ku2+6kkuu)γu+12kkuγuu+(3k2λ2)γuuuy_{uuu}=(6k_{u}^{2}+6kk_{uu})\gamma_{u}+12kk_{u}\gamma_{uu}+(3k^{2}-\lambda^{2})\gamma_{uuu}, but γuuu\gamma_{uuu} is not necessarily L2L^{2}.. However, we already have enough L2L^{2}-derivatives to differentiate (32) twice with respect to ss and for γ\gamma to satisfy the Euler-Lagrange equation (28) (a.e.). As for the periodicity conditions (29), integration by parts in (30) gives

dγV=20γs4+((3k2λ2)γs)s,V𝑑s+[2γss,Vs2γs3,V+(3k2λ)γs,V]0.\displaystyle d\operatorname{\mathcal{E}}_{\gamma}V=\begin{multlined}2\int^{\mathcal{L}}_{0}\left\langle{\gamma_{s^{4}}+((3k^{2}-\lambda^{2})\gamma_{s})_{s},V}\right\rangle\,ds\\ +\Bigl{[}2\left\langle{\gamma_{ss},V_{s}}\right\rangle-2\left\langle{\gamma_{s^{3}},V}\right\rangle+\left\langle{(3k^{2}-\lambda)\gamma_{s},V}\right\rangle\Bigr{]}_{0}^{\mathcal{L}}.\end{multlined}2\int^{\mathcal{L}}_{0}\left\langle{\gamma_{s^{4}}+((3k^{2}-\lambda^{2})\gamma_{s})_{s},V}\right\rangle\,ds\\ +\Bigl{[}2\left\langle{\gamma_{ss},V_{s}}\right\rangle-2\left\langle{\gamma_{s^{3}},V}\right\rangle+\left\langle{(3k^{2}-\lambda)\gamma_{s},V}\right\rangle\Bigr{]}_{0}^{\mathcal{L}}.

For γ\gamma to be stationary the boundary terms must vanish for all VH2(S1,n)V\in H^{2}(S^{1},\mathbb{R}^{n}). Choosing VV such that V(0)=0=V()V(0)=0=V(\mathcal{L}) we see that γss\gamma_{ss} (and therefore kk) must be periodic. Then, setting Vs(0)=Vs()V_{s}(0)=V_{s}(\mathcal{L}) (but V(0)0V(0)\neq 0) and using the periodicity of kk and γs\gamma_{s} we confirm that γs3\gamma_{s^{3}} is periodic. ∎

Next we calculate the second variation of \operatorname{\mathcal{E}}:

Lemma 4.2.

Let γ2(S1,n)\gamma\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}). Then for V,WH2(S1,n)V,W\in H^{2}(S^{1},\mathbb{R}^{n}) the second variation formula at γ\gamma is given by

(33) d2γ(V,W)=0(γ)[\displaystyle d^{2}\operatorname{\mathcal{E}}_{\gamma}(V,W)=\int_{0}^{\mathcal{L}(\gamma)}\bigl{[} Wss,2Vss2Vss,TT2Vs,κT6Vs,Tκ\displaystyle\left\langle{W_{ss},2V_{ss}-2\left\langle{V_{ss},T}\right\rangle T-2\left\langle{V_{s},\kappa}\right\rangle T-6\left\langle{V_{s},T}\right\rangle\kappa}\right\rangle
+Ws,2Vs,κκ6Vss,κT2Vss,Tκ\displaystyle+\left\langle{W_{s},-2\left\langle{V_{s},\kappa}\right\rangle\kappa-6\left\langle{V_{ss},\kappa}\right\rangle T-2\left\langle{V_{ss},T}\right\rangle\kappa}\right\rangle
+Ws,(3k2λ2)Vs+(15k2λ2)Vs,TT]ds.\displaystyle+\left\langle{W_{s},-(3k^{2}-\lambda^{2})V_{s}+(15k^{2}-\lambda^{2})\left\langle{V_{s},T}\right\rangle T}\right\rangle\bigr{]}\,ds.
Proof.

Consider d:UH2(S1,n)d\operatorname{\mathcal{E}}:U\to H^{2}(S^{1},\mathbb{R}^{n})^{*}, UU an open subset of 2(S1,n)\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}), and let α:=γ+εW\alpha:=\gamma+\varepsilon W be a variation of γU\gamma\in U in the direction of WH2(S1,n)W\in H^{2}(S^{1},\mathbb{R}^{n}). Then using εds=αεs,αsds\partial_{\varepsilon}ds=\left\langle{\alpha_{\varepsilon s},\alpha_{s}}\right\rangle ds, from (5) we calculate

εdαV=0[2Vssε,αss+2Vss,αssε6αssε,αssVs,αs(3k2λ2)(Vsε,αs+Vs,αsε)+(2Vss,αss(3k2λ2)Vs,αs)αεs,αs]ds.\displaystyle\partial_{\varepsilon}d\operatorname{\mathcal{E}}_{\alpha}V=\begin{multlined}\int_{0}^{\mathcal{L}}\bigl{[}2\left\langle{V_{ss\varepsilon},\alpha_{ss}}\right\rangle+2\left\langle{V_{ss},\alpha_{ss\varepsilon}}\right\rangle-6\left\langle{\alpha_{ss\varepsilon},\alpha_{ss}}\right\rangle\left\langle{V_{s},\alpha_{s}}\right\rangle\\ -(3k^{2}-\lambda^{2})\bigl{(}\left\langle{V_{s\varepsilon},\alpha_{s}}\right\rangle+\left\langle{V_{s},\alpha_{s\varepsilon}}\right\rangle\bigr{)}\\ +\bigl{(}2\left\langle{V_{ss},\alpha_{ss}}\right\rangle-(3k^{2}-\lambda^{2})\left\langle{V_{s},\alpha_{s}}\right\rangle\bigr{)}\left\langle{\alpha_{\varepsilon s},\alpha_{s}}\right\rangle\bigr{]}\,ds.\end{multlined}\int_{0}^{\mathcal{L}}\bigl{[}2\left\langle{V_{ss\varepsilon},\alpha_{ss}}\right\rangle+2\left\langle{V_{ss},\alpha_{ss\varepsilon}}\right\rangle-6\left\langle{\alpha_{ss\varepsilon},\alpha_{ss}}\right\rangle\left\langle{V_{s},\alpha_{s}}\right\rangle\\ -(3k^{2}-\lambda^{2})\bigl{(}\left\langle{V_{s\varepsilon},\alpha_{s}}\right\rangle+\left\langle{V_{s},\alpha_{s\varepsilon}}\right\rangle\bigr{)}\\ +\bigl{(}2\left\langle{V_{ss},\alpha_{ss}}\right\rangle-(3k^{2}-\lambda^{2})\left\langle{V_{s},\alpha_{s}}\right\rangle\bigr{)}\left\langle{\alpha_{\varepsilon s},\alpha_{s}}\right\rangle\bigr{]}\,ds.

Using the formula for commutation with the arc length derivative:

ϕsε=ϕεsαεs,αsϕs,\phi_{s\varepsilon}=\phi_{\varepsilon s}-\left\langle{\alpha_{\varepsilon s},\alpha_{s}}\right\rangle\phi_{s},

and writing T=γsT=\gamma_{s}, we calculate

αsε|ε=0\displaystyle\alpha_{s\varepsilon}\bigm{|}_{\varepsilon=0} =αεsαεs,αsαs|ε=0=WsWs,TT,\displaystyle=\alpha_{\varepsilon s}-\left\langle{\alpha_{\varepsilon s},\alpha_{s}}\right\rangle\alpha_{s}\bigm{|}_{\varepsilon=0}=W_{s}-\left\langle{W_{s},T}\right\rangle T,
αssε|ε=0\displaystyle\alpha_{ss\varepsilon}\bigm{|}_{\varepsilon=0} =αεssαεss,αsαsαεs,αssαs2αεs,αsαss|ε=0\displaystyle=\alpha_{\varepsilon ss}-\left\langle{\alpha_{\varepsilon ss},\alpha_{s}}\right\rangle\alpha_{s}-\left\langle{\alpha_{\varepsilon s},\alpha_{ss}}\right\rangle\alpha_{s}-2\left\langle{\alpha_{\varepsilon s},\alpha_{s}}\right\rangle\alpha_{ss}\bigm{|}_{\varepsilon=0}
=WssWss,TTWs,κT2Ws,Tκ.\displaystyle=W_{ss}-\left\langle{W_{ss},T}\right\rangle T-\left\langle{W_{s},\kappa}\right\rangle T-2\left\langle{W_{s},T}\right\rangle\kappa.

Moreover, noting that VV is independent of ε\varepsilon, we have

Vsε|ε=0\displaystyle V_{s\varepsilon}\bigm{|}_{\varepsilon=0} =αεs,αsVs|ε=0=Ws,TVs,\displaystyle=-\left\langle{\alpha_{\varepsilon s},\alpha_{s}}\right\rangle V_{s}\bigm{|}_{\varepsilon=0}=-\left\langle{W_{s},T}\right\rangle V_{s},
Vssε|ε=0\displaystyle V_{ss\varepsilon}\bigm{|}_{\varepsilon=0} =αεss,αsVsαεs,αssVs2αεs,αsVss|ε=0\displaystyle=-\left\langle{\alpha_{\varepsilon ss},\alpha_{s}}\right\rangle V_{s}-\left\langle{\alpha_{\varepsilon s},\alpha_{ss}}\right\rangle V_{s}-2\left\langle{\alpha_{\varepsilon s},\alpha_{s}}\right\rangle V_{ss}\bigm{|}_{\varepsilon=0}
=Wss,TVsWs,κVs2Ws,TVss.\displaystyle=-\left\langle{W_{ss},T}\right\rangle V_{s}-\left\langle{W_{s},\kappa}\right\rangle V_{s}-2\left\langle{W_{s},T}\right\rangle V_{ss}.

Therefore, evaluating εdαV\partial_{\varepsilon}d\operatorname{\mathcal{E}}_{\alpha}V at ε=0\varepsilon=0 and using αss,αs=0\left\langle{\alpha_{ss},\alpha_{s}=0}\right\rangle:

d2γ(V,W)\displaystyle d^{2}\operatorname{\mathcal{E}}_{\gamma}(V,W)
=0(γ)[2Wss,TVs,κ2Ws,κVs,κ4Ws,TVss,κ\displaystyle=\int_{0}^{\mathcal{L}(\gamma)}\bigl{[}-2\left\langle{W_{ss},T}\right\rangle\left\langle{V_{s},\kappa}\right\rangle-2\left\langle{W_{s},\kappa}\right\rangle\left\langle{V_{s},\kappa}\right\rangle-4\left\langle{W_{s},T}\right\rangle\left\langle{V_{ss},\kappa}\right\rangle
+2Vss,Wss2Wss,TT,Vss2Ws,κT,Vss\displaystyle\qquad\qquad\quad+2\left\langle{V_{ss},W_{ss}}\right\rangle-2\left\langle{W_{ss},T}\right\rangle\left\langle{T,V_{ss}}\right\rangle-2\left\langle{W_{s},\kappa}\right\rangle\left\langle{T,V_{ss}}\right\rangle
4Ws,Tκ,Vss6Wss,κVs,T+12k2Ws,TVs,T\displaystyle\qquad\qquad\quad-4\left\langle{W_{s},T}\right\rangle\left\langle{\kappa,V_{ss}}\right\rangle-6\left\langle{W_{ss},\kappa}\right\rangle\left\langle{V_{s},T}\right\rangle+12k^{2}\left\langle{W_{s},T}\right\rangle\left\langle{V_{s},T}\right\rangle
+(3k2λ2)(2Ws,TVs,TVs,Ws)\displaystyle\qquad\qquad\quad+(3k^{2}-\lambda^{2})\left(2\left\langle{W_{s},T}\right\rangle\left\langle{V_{s},T}\right\rangle-\left\langle{V_{s},W_{s}}\right\rangle\right)
+(2Vss,κ(3k2λ2)Vs,T)Ws,T]ds\displaystyle\qquad\qquad\quad+(2\left\langle{V_{ss},\kappa}\right\rangle-(3k^{2}-\lambda^{2})\left\langle{V_{s},T}\right\rangle)\left\langle{W_{s},T}\right\rangle\bigr{]}\,ds
=0(γ)[Wss,2Vss2Vss,TT2Vs,κT6Vs,Tκ\displaystyle=\int_{0}^{\mathcal{L}(\gamma)}\bigl{[}\left\langle{W_{ss},2V_{ss}-2\left\langle{V_{ss},T}\right\rangle T-2\left\langle{V_{s},\kappa}\right\rangle T-6\left\langle{V_{s},T}\right\rangle\kappa}\right\rangle
+Ws,2Vs,κκ6Vss,κT2Vss,Tκ\displaystyle\qquad\qquad\quad+\left\langle{W_{s},-2\left\langle{V_{s},\kappa}\right\rangle\kappa-6\left\langle{V_{ss},\kappa}\right\rangle T-2\left\langle{V_{ss},T}\right\rangle\kappa}\right\rangle
+Ws,(3k2λ2)Vs+(15k2λ2)Vs,TT]ds\displaystyle\qquad\qquad\quad+\left\langle{W_{s},-(3k^{2}-\lambda^{2})V_{s}+(15k^{2}-\lambda^{2})\left\langle{V_{s},T}\right\rangle T}\right\rangle\bigr{]}\,ds

which proves the lemma. ∎

4.2. Analyticity

We are assuming less differentiability than [8, Section 3.1] and so the proof is a little different.

Definition (cf. [8], Section 3.1).

Let X,YX,Y be Banach spaces and DXD\subset X an open subset. A map f:DYf:D\to Y is analytic at x0Dx_{0}\in D if there exist continuous symmetric multilinear functions aj:Xj=X××XY,ja_{j}:X^{j}=X\times\ldots\times X\to Y,\,j\in\mathbb{N} and a0Ya_{0}\in Y such that in a neighbourhood of x0x_{0}

j=0ajopxx0j<,andf(x)=j=0aj(xx0)j,\sum_{j=0}^{\infty}{\left\|{a_{j}}\right\|}_{\operatorname{\rm{op}}}{\left\|{x-x_{0}}\right\|}^{j}<\infty,\quad\text{and}\quad f(x)=\sum_{j=0}^{\infty}a_{j}(x-x_{0})^{j},

where ajop{\left\|{a_{j}}\right\|}_{\operatorname{\rm{op}}} is the operator norm and aj(xx0)j:=aj(xx0,,xx0)a_{j}(x-x_{0})^{j}:=a_{j}(x-x_{0},\ldots,x-x_{0}).

Our goal is to prove that :H2(S1,n)\operatorname{\mathcal{E}}:H^{2}(S^{1},\mathbb{R}^{n})\to\mathbb{R} is analytic. The proof will require a series of lemmas showing that the constituent functions such as γ|γu|\gamma\mapsto|\gamma_{u}| are analytic. Let H+1(S1,):={αH1(S1,):α(u)>0}H^{1}_{+}(S^{1},\mathbb{R}):=\{\alpha\in H^{1}(S^{1},\mathbb{R}):\alpha(u)>0\} and define

f:H+1(S1,)H+1(S1,),f(α)(u):=α(u).f:H^{1}_{+}(S^{1},\mathbb{R})\to H^{1}_{+}(S^{1},\mathbb{R}),\quad f(\alpha)(u):=\sqrt{\alpha(u)}.

Note that by the Sobolev imbedding α\alpha is continuous and so

f(α)H12=S1α(u)𝑑u+S114ααu2𝑑uαL+14(min|α|)1αuL22<{\left\|{f(\alpha)}\right\|}_{H^{1}}^{2}=\int_{S^{1}}\alpha(u)\,du+\int_{S^{1}}\frac{1}{4\alpha}\alpha_{u}^{2}\,du\leq{\left\|{\alpha}\right\|}_{L^{\infty}}+\frac{1}{4}(\min|\alpha|)^{-1}{\left\|{\alpha_{u}}\right\|}^{2}_{L^{2}}<\infty

which shows that f(α)f(\alpha) is indeed in H1H^{1}.

Lemma 4.3.

The function f:H+1(S1,)H+1(S1,)f:H^{1}_{+}(S^{1},\mathbb{R})\to H^{1}_{+}(S^{1},\mathbb{R}) is analytic.

Proof.

Let α0H+1(S1,)\alpha_{0}\in H^{1}_{+}(S^{1},\mathbb{R}) and consider the open H1H^{1}-ball Bε(α0)B_{\varepsilon}(\alpha_{0}) with radius ε\varepsilon centred at α0\alpha_{0}. Assume εCS<minuS1|α0(u)|\varepsilon C_{S}<\min_{u\in S^{1}}|\alpha_{0}(u)|, where CSC_{S} is the Sobolev imbedding constant, so that for all αBε(α0)\alpha\in B_{\varepsilon}(\alpha_{0}) we have

(34) 0<α0(u)εCS<α(u)<εCS+α0(u)0<\alpha_{0}(u)-\varepsilon C_{S}<\alpha(u)<\varepsilon C_{S}+\alpha_{0}(u)

for all uS1u\in S^{1}, and therefore Bε(α0)H+1(S1,)B_{\varepsilon}(\alpha_{0})\subset H^{1}_{+}(S^{1},\mathbb{R}). Considered as a function from ++\mathbb{R}^{+}\to\mathbb{R}^{+}, the square root is analytic. Indeed, the Taylor series about x0x_{0} is

j=0bj(x0)(xx0)j,bj(x0):=cjx012j,cj:=(1)j1(2j3)!22j2j!(j2)!,\sum_{j=0}^{\infty}b_{j}(x_{0})(x-x_{0})^{j},\quad b_{j}(x_{0}):=c_{j}x_{0}^{\frac{1}{2}-j},\quad c_{j}:=\frac{(-1)^{j-1}(2j-3)!}{2^{2j-2}j!(j-2)!},

the ratio of successive terms is (12j)2(j+1)(xx0)x01\frac{(1-2j)}{2(j+1)}(x-x_{0})x_{0}^{-1}, and so by the ratio test the series converges when |xx0|<x0|x-x_{0}|<x_{0}. It follows from (34) that if αBε(α0)\alpha\in B_{\varepsilon}(\alpha_{0}) then for each uu the series

j=0bj(α0(u))(α(u)α0(u))j\sum_{j=0}^{\infty}b_{j}(\alpha_{0}(u))(\alpha(u)-\alpha_{0}(u))^{j}

converges to f(α)(u)f(\alpha)(u). Moreover, since α0(u)>0\alpha_{0}(u)>0 we have

bj(α0)H12cj2(minuS1α0)12j+(12j)2cj2(minuS1α0)12juα0L22,{\left\|{b_{j}(\alpha_{0})}\right\|}_{H^{1}}^{2}\leq c_{j}^{2}\Bigl{(}\min_{u\in S^{1}}\alpha_{0}\Bigr{)}^{1-2j}+\Bigl{(}\frac{1}{2}-j\Bigr{)}^{2}c_{j}^{2}\Bigl{(}\min_{u\in S^{1}}\alpha_{0}\Bigr{)}^{-1-2j}{\left\|{\partial_{u}\alpha_{0}}\right\|}_{L^{2}}^{2},

and so bj(α0)H1b_{j}(\alpha_{0})\in H^{1}. Since H1(S1,)H^{1}(S^{1},\mathbb{R}) is a Banach algebra (see e.g. [1, Theorem 4.39]) we may identify bj(α0)b_{j}(\alpha_{0}) with the linear map from H1(S1,)H^{1}(S^{1},\mathbb{R}) to itself given by α(u)bj(α0(u))α(u)\alpha(u)\mapsto b_{j}(\alpha_{0}(u))\alpha(u). Then to conclude the proof we need to show that j=0bj(α0)opαα0H1j\sum_{j=0}^{\infty}{\left\|{b_{j}(\alpha_{0})}\right\|}_{\operatorname{\rm{op}}}{\left\|{\alpha-\alpha_{0}}\right\|}_{H^{1}}^{j} converges, where the norm on bj(α0)b_{j}(\alpha_{0}) is the operator norm. Note that

bj+1(α0)=12j2j+2α01bj(α0).b_{j+1}(\alpha_{0})=\frac{1-2j}{2j+2}\alpha_{0}^{-1}b_{j}(\alpha_{0}).

Then using the Banach algebra property ([1, Theorem 4.39]) there is a constant KK such that

bj+1(α0)op=supvH1=112j2j+2α01bj(α0)vH1K12j2j+2α01H1bj(α0)op\|b_{j+1}(\alpha_{0})\|_{\operatorname{\rm{op}}}=\sup_{\|v\|_{H^{1}}=1}\Bigl{\|}\frac{1-2j}{2j+2}\alpha_{0}^{-1}b_{j}(\alpha_{0})v\Bigr{\|}_{H^{1}}\leq K\Bigl{\|}\frac{1-2j}{2j+2}\alpha_{0}^{-1}\Bigr{\|}_{H^{1}}\|b_{j}(\alpha_{0})\|_{\operatorname{\rm{op}}}

and so the ratio of successive terms is

bj+1(α0)opbj(α0)opαα0H1K12j2j+2α01H1αα0H1Kα01H1αα0H1\frac{\|b_{j+1}(\alpha_{0})\|_{\operatorname{\rm{op}}}}{\|b_{j}(\alpha_{0})\|_{\operatorname{\rm{op}}}}\|\alpha-\alpha_{0}\|_{H^{1}}\leq K\Bigl{\|}\frac{1-2j}{2j+2}\alpha_{0}^{-1}\Bigr{\|}_{H^{1}}\|\alpha-\alpha_{0}\|_{H^{1}}\\ \to K\|\alpha_{0}^{-1}\|_{H^{1}}\|\alpha-\alpha_{0}\|_{H^{1}}

as jj\to\infty. By the ratio test the series converges if αα0H1<K1α01H11\|\alpha-\alpha_{0}\|_{H^{1}}<K^{-1}\|\alpha_{0}^{-1}\|_{H^{1}}^{-1} which can be arranged by choosing ε\varepsilon sufficiently small. ∎

Lemma 4.4.

The function 2(S1,n)H+1(S1,),γ|γu|\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})\to H^{1}_{+}(S^{1},\mathbb{R}),\gamma\mapsto|\gamma_{u}| is analytic.

Proof.

First note that u:2(S1,n)H1(S1,n)\partial_{u}:\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})\to H^{1}(S^{1},\mathbb{R}^{n}) is linear and continuous, therefore analytic. The map induced by the Euclidean inner product ,:H1(S1,n)×H1(S1,n)H1(S1,)\left\langle{\cdot,\cdot}\right\rangle:H^{1}(S^{1},\mathbb{R}^{n})\times H^{1}(S^{1},\mathbb{R}^{n})\to H^{1}(S^{1},\mathbb{R}) satisfies

v,wH12=S1[|v,w|2+|v,w+v,w|2]𝑑ucvH12wH12.\|\left\langle{v,w}\right\rangle\|_{H^{1}}^{2}=\int_{S^{1}}\bigl{[}|\left\langle{v,w}\right\rangle|^{2}+|\left\langle{v^{\prime},w}\right\rangle+\left\langle{v,w^{\prime}}\right\rangle|^{2}\bigr{]}\,du\leq c\|v\|_{H^{1}}^{2}\|w\|_{H^{1}}^{2}.

So it is continuous bilinear and therefore the map q(γ):=γ,γq(\gamma):=\left\langle{\gamma,\gamma}\right\rangle is analytic. Now by Lemma 4.3 we have that γ|γu|\gamma\mapsto|\gamma_{u}| is a composition of analytic functions fquf\circ q\circ\partial_{u} and is therefore analytic by e.g. [44, p. 1079]. ∎

Lemma 4.5.

The function rec:H+1(S1,)H+1(S1,),ψ(u)ψ(u)1\operatorname{\rm{rec}}:H^{1}_{+}(S^{1},\mathbb{R})\to H^{1}_{+}(S^{1},\mathbb{R}),\,\psi(u)\mapsto\psi(u)^{-1} is analytic.

Proof.

This is proved for H3H^{3} functions in [8, Appendix B.1], and the same method of proof works for H1H^{1} functions. ∎

We also require the following property of analytic functions: If F:DY1,G:DY2F:D\to Y_{1},\,G:D\to Y_{2} are analytic and there exists a bilinear continuous mapping :Y1×Y2Z*:Y_{1}\times Y_{2}\to Z into another Banach space, then the product FG:DZ,xF(x)G(x)F*G:D\to Z,x\mapsto F(x)*G(x) is also analytic. According to [8, p. 2175] this can be proved using similar ideas as for the Cauchy product of series.

Lemma 4.6.

The function s:2(S1,n)H1(S1,n)\partial_{s}:\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})\to H^{1}(S^{1},\mathbb{R}^{n}), γγs=γu/|γu|\gamma\mapsto\gamma_{s}=\gamma_{u}/|\gamma_{u}| is analytic.

Proof.

We observed in the two previous lemmas that u\partial_{u} and γ|γu|1\gamma\mapsto|\gamma_{u}|^{-1} are analytic on 2(S1,n)\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}), with respective codomains H1(S1,n)H^{1}(S^{1},\mathbb{R}^{n}) and H+1(S1,)H^{1}_{+}(S^{1},\mathbb{R}). Moreover the pointwise product :H1(S1,)×H1(S1,n)H1(S1,n)*:H^{1}(S^{1},\mathbb{R})\times H^{1}(S^{1},\mathbb{R}^{n})\to H^{1}(S^{1},\mathbb{R}^{n}) is bilinear and continuous. Indeed if αH1(S1,)\alpha\in H^{1}(S^{1},\mathbb{R}) and γH1(S1,n)\gamma\in H^{1}(S^{1},\mathbb{R}^{n}) then

αγH1αLγL2+αLγuL2+αuL2γL.\|\alpha*\gamma\|_{H^{1}}\leq\|\alpha\|_{L^{\infty}}\|\gamma\|_{L^{2}}+\|\alpha\|_{L^{\infty}}\|\gamma_{u}\|_{L^{2}}+\|\alpha_{u}\|_{L^{2}}\|\gamma\|_{L^{\infty}}.

Since sγ\partial_{s}\gamma is the pointwise product of γu\gamma_{u} and |γu|1|\gamma_{u}|^{-1}, the result follows. ∎

Lemma 4.7.

The function s2:2(S1,n)L2(S1,n)\partial_{s}^{2}:\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})\to L^{2}(S^{1},\mathbb{R}^{n}) is analytic.

Proof.

Since u:H1(S1,n)L2(S1,n)\partial_{u}:H^{1}(S^{1},\mathbb{R}^{n})\to L^{2}(S^{1},\mathbb{R}^{n}) is also linear and continuous, by Lemma 4.6 we have that us:2(S1,n)L2(S1,n)\partial_{u}\partial_{s}:\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})\to L^{2}(S^{1},\mathbb{R}^{n}) is a composition of analytic functions, therefore analytic. The product :H1(S1,)×L2(S1,n)L2(S1,n)*:H^{1}(S^{1},\mathbb{R})\times L^{2}(S^{1},\mathbb{R}^{n})\to L^{2}(S^{1},\mathbb{R}^{n}) is also bilinear and continuous, and γss=|γu|1γsu\gamma_{ss}=|\gamma_{u}|^{-1}\gamma_{su}, so s2\partial_{s}^{2} is analytic. ∎

Proposition 4.8.

The energy :2(S1,n)\operatorname{\mathcal{E}}:\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})\to\mathbb{R} is analytic.

Proof.

Recall that (γ)=S1[γss,γss|γu|+λ|γu|]𝑑u\operatorname{\mathcal{E}}(\gamma)=\int_{S^{1}}[\left\langle{\gamma_{ss},\gamma_{ss}}\right\rangle|\gamma_{u}|+\lambda|\gamma_{u}|]\,du. Lemma 4.7 implies that s2\partial_{s}^{2} is analytic. Moreover, the Euclidean inner product is continuous bilinear on L2(S1,n)×L2(S1,n)L1(S1,)L^{2}(S^{1},\mathbb{R}^{n})\times L^{2}(S^{1},\mathbb{R}^{n})\to L^{1}(S^{1},\mathbb{R}), pointwise multiplication :L1(S1,)×H1(S1,)L1(S1,)*:L^{1}(S^{1},\mathbb{R})\times H^{1}(S^{1},\mathbb{R})\to L^{1}(S^{1},\mathbb{R}) is continuous bilinear, and the sum of analytic functions is analytic. Thus the integrand is analytic as a function 2(S1,n)L1(S1,)\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})\to L^{1}(S^{1},\mathbb{R}). Integration is of course linear and bounded on L1L^{1}, so the energy is a composition of analytic functions. ∎

4.3. The submanifold of arc length proportional parametrized curves

Denote Hzm1(S1,):={αH1(S1,):S1α𝑑u=0}H^{1}_{zm}(S^{1},\mathbb{R}):=\{\alpha\in H^{1}(S^{1},\mathbb{R}):\int_{S^{1}}\alpha\,du=0\} and define

Φ:2(S1,n)Hzm1(S1,),Φ(γ):=|γu|(γ).\Phi:\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})\to H^{1}_{zm}(S^{1},\mathbb{R}),\quad\Phi(\gamma):=|\gamma_{u}|-\mathcal{L}(\gamma).

Then Ω:=Φ1(0)\Omega:=\Phi^{-1}(0) is the subset of 2(S1,n)\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) consisting of curves which are parametrized proportional to arc length.

Proposition 4.9.

The set Ω\Omega of arc length proportional parametrized curves is an analytic submanifold of H2(S1,n)H^{2}(S^{1},\mathbb{R}^{n}).

Proof.

We will show that 𝟎H1(S1,)\mathbf{0}\in H^{1}(S^{1},\mathbb{R}) is a regular value of Φ\Phi. For the derivative we get

(35) dΦγv=1|γ|v,γS11|γ|v,γ𝑑u,d\Phi_{\gamma}v=\frac{1}{|\gamma^{\prime}|}\left\langle{v^{\prime},\gamma^{\prime}}\right\rangle-\int_{S^{1}}\frac{1}{|\gamma^{\prime}|}\left\langle{v^{\prime},\gamma^{\prime}}\right\rangle du,

and so 𝟎\mathbf{0} is a regular value if for all wHzm1(S1,)w\in H^{1}_{zm}(S^{1},\mathbb{R}) there exists vH2(S1,n)v\in H^{2}(S^{1},\mathbb{R}^{n}) such that

(36) 1|γ|v,γS11|γ|v,γ𝑑u=w.\displaystyle\frac{1}{|\gamma^{\prime}|}\left\langle{v^{\prime},\gamma^{\prime}}\right\rangle-\int_{S^{1}}\frac{1}{|\gamma^{\prime}|}\left\langle{v^{\prime},\gamma^{\prime}}\right\rangle du=w.

To show that such a vv exists we will need an orthonormal frame {T,νi}\{T,\nu_{i}\} along γ\gamma which we construct as follows. First let {T(0),νi0}\{T(0),\nu_{i}^{0}\} be an orthonormal basis at γ(0)\gamma(0), and then let νi(u),i=1,n1\nu_{i}(u),\,i=1,\ldots n-1 be the solutions to

(37) νi=1|γ|2νi,γ′′γ,νi(0)=νi0.\nu_{i}^{\prime}=-\frac{1}{{\left|{\gamma^{\prime}}\right|}^{2}}\left\langle{\nu_{i},\gamma^{\prime\prime}}\right\rangle\gamma^{\prime},\quad\nu_{i}(0)=\nu_{i}^{0}\,.

Then

dduγ,νi\displaystyle\frac{d}{du}\left\langle{\gamma^{\prime},\nu_{i}}\right\rangle =γ′′,νi1|γ|2γ,γνi,γ′′=0\displaystyle=\left\langle{\gamma^{\prime\prime},\nu_{i}}\right\rangle-\frac{1}{{\left|{\gamma^{\prime}}\right|}^{2}}\left\langle{\gamma^{\prime},\gamma^{\prime}}\right\rangle\left\langle{\nu_{i},\gamma^{\prime\prime}}\right\rangle=0

and since T(0),νi0=0\left\langle{T(0),\nu_{i}^{0}}\right\rangle=0 we have γ,νi=0\left\langle{\gamma^{\prime},\nu_{i}}\right\rangle=0 identically. Moreover

dduνi,νj\displaystyle\frac{d}{du}\left\langle{\nu_{i},\nu_{j}}\right\rangle =1|γ|2νi,γ′′γ,νj1|γ|2νj,γ′′νi,γ=0\displaystyle=-\frac{1}{{\left|{\gamma^{\prime}}\right|}^{2}}\left\langle{\nu_{i},\gamma^{\prime\prime}}\right\rangle\left\langle{\gamma^{\prime},\nu_{j}}\right\rangle-\frac{1}{{\left|{\gamma^{\prime}}\right|}^{2}}\left\langle{\nu_{j},\gamma^{\prime\prime}}\right\rangle\left\langle{\nu_{i},\gamma^{\prime}}\right\rangle=0

from which it follows that the orthonormality of {T(0),νi0}\{T(0),\nu_{i}^{0}\} is indeed preserved and {T,νi}\{T,\nu_{i}\} is an orthonormal frame.

Now given wHzm1(S1,)w\in H^{1}_{zm}(S^{1},\mathbb{R}) we let vv be a solution of

(38) v=wT+βi=1n1νiξi,v^{\prime}=wT+\beta\sum_{i=1}^{n-1}\nu_{i}\xi_{i},

where β:S1\beta:S^{1}\to\mathbb{R} is any smooth function satisfying β(0)=0=β(1)\beta(0)=0=\beta(1) and thereby ensuring v(0)=v(1)v^{\prime}(0)=v^{\prime}(1), and ξH1(S1,n1)\xi\in H^{1}(S^{1},\mathbb{R}^{n-1}) is a control function which we are free to choose in order to ensure vv satisfies the zeroth order periodicity condition v(0)=v(1)v(0)=v(1). If this is possible then the solution vv satisfies (36) and dΦγd\Phi_{\gamma} is surjective.

In fact, it will be sufficient to consider the system

(39) x=βi=1n1νiξix^{\prime}=\beta\sum_{i=1}^{n-1}\nu_{i}\xi_{i}

because if yy is any solution to y=wTy^{\prime}=wT and we can control xx from e.g. x(0)=y(0)x(0)=y(0) to x(1)=y(1)x(1)=y(1), then we let v=yxv=y-x and v(0)=0=v(1)v(0)=0=v(1). According to [3, p. 76], a sufficient condition for the existence of such a control is that the matrix

(40) W:=01BBT𝑑u,where B:=β[ν1νn1]W:=\int_{0}^{1}BB^{T}du,\quad\text{where }B:=\beta[\nu_{1}\ldots\nu_{n-1}]

should be non-singular (here BB is an n×(n1)n\times(n-1) matrix with columns νi\nu_{i}, and BTB^{T} its transpose), in which case a particular control which drives the solution from x(0)x(0) to x(1)x(1) is ξ=BTW1(x(1)x(0))\xi=B^{T}W^{-1}(x(1)-x(0)).

Suppose WW is singular. Then there is a constant non-zero vector ana\in\mathbb{R}^{n} such that

0=aTWa=01(aTB)(aTB)T𝑑u0=a^{T}Wa=\int_{0}^{1}(a^{T}B)(a^{T}B)^{T}\,du

which implies that aTB=β[a,ν1,,a,νn1]=0a^{T}B=\beta[\left\langle{a,\nu_{1}}\right\rangle,\ldots,\left\langle{a,\nu_{n-1}}\right\rangle]=0 on (0,1)(0,1), i.e. aa is in the direction of ±T\pm T. But this is impossible because γ\gamma is closed and not constant, and TT is continuous.

We therefore have that dΦγd\Phi_{\gamma} is surjective, and its kernel splits because it is a closed subspace of a Hilbert space. Therefore 𝟎\mathbf{0} is a regular value of Φ\Phi and Ω\Omega is a submanifold (see e.g. [38, Theorem  2.2.2], with [15, Proposition  2.3]). To see that it is an analytic submanifold we note that (e.g. in the proof of [15, Proposition  2.3]) the charts for Ω\Omega are constructed by applying the inverse function theorem to Φ\Phi. Since Φ\Phi is analytic as a consequence of Lemma 4.4, local inverses of Φ\Phi will also be analytic by a theorem of Whittlesey ([44, p. 1081]). So the charts are analytic, i.e. Ω\Omega is analytic. ∎

Combining Proposition 4.8 with Proposition 4.9, we have:

Corollary 4.10.

The restriction |Ω\operatorname{\mathcal{E}}|\Omega is analytic.

Here we give a characterization of the tangent space TγΩT_{\gamma}\Omega.

Corollary 4.11.

The tangent space TγΩT_{\gamma}\Omega is equal to kerdΦγ\ker d\Phi_{\gamma} and it consists of all VH2(S1,n)V\in H^{2}(S^{1},\mathbb{R}^{n}) satisfying

(41) Vs,T+1(γ)0(γ)V,κ𝑑s=0.\left\langle{V_{s},T}\right\rangle+\frac{1}{\mathcal{L}(\gamma)}\int^{\mathcal{L}(\gamma)}_{0}\left\langle{V,\kappa}\right\rangle ds=0.
Proof.

The relation TγΩ=kerdΦγT_{\gamma}\Omega=\ker d\Phi_{\gamma} is a consequence of the regular values theorem (see e.g. [38, Theorem 2.2.2]). From (35) we see that

Vu,TS1Vu,T𝑑u=0\left\langle{V_{u},T}\right\rangle-\int_{S^{1}}\left\langle{V_{u},T}\right\rangle du=0

for all VTγΩV\in T_{\gamma}\Omega. Since |γu|=(γ)|\gamma_{u}|=\mathcal{L}(\gamma) for γΩ\gamma\in\Omega, multiplication by 1/(γ)1/\mathcal{L}(\gamma) and integration by parts gives the desired expression. ∎

4.4. The gradient inequality

Proposition 4.12.

Let γ\gamma be a stationary point of |Ω\operatorname{\mathcal{E}}|\Omega, then d2(|Ω)γd^{2}(\operatorname{\mathcal{E}}|\Omega)_{\gamma} is a Fredholm operator with index zero.

Proof.

Let γΩ\gamma\in\Omega, and fix V,WTγΩV,W\in T_{\gamma}\Omega arbitrarily. Taking a derivative of (41), we have

(42) Vss,T=Vs,κ.\left\langle{V_{ss},T}\right\rangle=-\left\langle{V_{s},\kappa}\right\rangle.

Combining Lemma 4.2 with (42) and Corollary 4.11, we reduce the second variation formula to

d2γ(V,W)\displaystyle d^{2}\operatorname{\mathcal{E}}_{\gamma}(V,W) =0(γ)[Wss,2Vss6T,Vsκ+Ws,6κ,VssT+(15k2λ2)Vs,TT(3k2λ2)Ws,Vs]ds.\displaystyle=\int^{\mathcal{L}(\gamma)}_{0}\begin{multlined}\bigl{[}\left\langle{W_{ss},2V_{ss}-6\left\langle{T,V_{s}}\right\rangle\kappa}\right\rangle\\ +\left\langle{W_{s},-6\left\langle{\kappa,V_{ss}}\right\rangle T+(15k^{2}-\lambda^{2})\left\langle{V_{s},T}\right\rangle T}\right\rangle\\ -(3k^{2}-\lambda^{2})\left\langle{W_{s},V_{s}}\right\rangle\bigr{]}\,ds.\end{multlined}\bigl{[}\left\langle{W_{ss},2V_{ss}-6\left\langle{T,V_{s}}\right\rangle\kappa}\right\rangle\\ +\left\langle{W_{s},-6\left\langle{\kappa,V_{ss}}\right\rangle T+(15k^{2}-\lambda^{2})\left\langle{V_{s},T}\right\rangle T}\right\rangle\\ -(3k^{2}-\lambda^{2})\left\langle{W_{s},V_{s}}\right\rangle\bigr{]}\,ds.

Assuming furthermore that γ\gamma is a critical point of \operatorname{\mathcal{E}} and therefore admits higher derivatives, we can differentiate

s(κ,VsT)=γs3,VsT+κ,VssT+κ,Vsκ\partial_{s}(\left\langle{\kappa,V_{s}}\right\rangle T)=\left\langle{\gamma_{s^{3}},V_{s}}\right\rangle T+\left\langle{\kappa,V_{ss}}\right\rangle T+\left\langle{\kappa,V_{s}}\right\rangle\kappa

and use this to eliminate the κ,VssT\left\langle{\kappa,V_{ss}}\right\rangle T term from the expression for d2d^{2}\operatorname{\mathcal{E}}. Then after further simplifications we find

d2γ(V,W)=0(γ)[2Wss,Vss+Ws,6γs3,VsT+6T,Vsγs3+Ws,(15k2λ2)Vs,TT(3k2λ2)Vs]ds.\displaystyle d^{2}\operatorname{\mathcal{E}}_{\gamma}(V,W)=\begin{multlined}\int^{\mathcal{L}(\gamma)}_{0}\bigl{[}2\left\langle{W_{ss},V_{ss}}\right\rangle+\left\langle{W_{s},6\left\langle{\gamma_{s^{3}},V_{s}}\right\rangle T+6\left\langle{T,V_{s}}\right\rangle\gamma_{s^{3}}}\right\rangle\\ +\left\langle{W_{s},(15k^{2}-\lambda^{2})\left\langle{V_{s},T}\right\rangle T-(3k^{2}-\lambda^{2})V_{s}}\right\rangle\bigr{]}\,ds.\end{multlined}\int^{\mathcal{L}(\gamma)}_{0}\bigl{[}2\left\langle{W_{ss},V_{ss}}\right\rangle+\left\langle{W_{s},6\left\langle{\gamma_{s^{3}},V_{s}}\right\rangle T+6\left\langle{T,V_{s}}\right\rangle\gamma_{s^{3}}}\right\rangle\\ +\left\langle{W_{s},(15k^{2}-\lambda^{2})\left\langle{V_{s},T}\right\rangle T-(3k^{2}-\lambda^{2})V_{s}}\right\rangle\bigr{]}\,ds.

Since γΩ\gamma\in\Omega we have

V,WH2=3(γ)Vss,WssL2(ds)+(γ)Vs,WsL2(ds)+V,WL2\left\langle{V,W}\right\rangle_{H^{2}}=\mathcal{L}^{3}(\gamma)\left\langle{V_{ss},W_{ss}}\right\rangle_{L^{2}(ds)}+\mathcal{L}(\gamma)\left\langle{V_{s},W_{s}}\right\rangle_{L^{2}(ds)}+\left\langle{V,W}\right\rangle_{L^{2}}

and observe that the second derivative of \operatorname{\mathcal{E}} has the form

d2γ(V,W)=2(γ)3V,WH2+0(γ)τ(V),W𝑑s,d^{2}\operatorname{\mathcal{E}}_{\gamma}(V,W)=\frac{2}{\mathcal{L}(\gamma)^{3}}\left\langle{V,W}\right\rangle_{H^{2}}+\int^{\mathcal{L}(\gamma)}_{0}\left\langle{\tau(V),W}\right\rangle\,ds,

where τ\tau is a continuous linear map from TγΩT_{\gamma}\Omega into L2(S1,n)L^{2}(S^{1},\mathbb{R}^{n}). We define the associated operator B:TγΩH2(S1,n)TγΩB:T_{\gamma}\Omega\subset H^{2}(S^{1},\mathbb{R}^{n})\to T_{\gamma}\Omega^{*} by

B(V)=d2(|Ω)γ(V,).B(V)=d^{2}(\operatorname{\mathcal{E}}|\Omega)_{\gamma}(V,\cdot).

Since TγΩT_{\gamma}\Omega is a closed subspace in H2(S1,n)H^{2}(S^{1},\mathbb{R}^{n}), we see that (TγΩ,,H2)(T_{\gamma}\Omega,\langle\cdot,\cdot\rangle_{H^{2}}) is a Hilbert space. Moreover, by the form of d2(|Ω)γd^{2}(\operatorname{\mathcal{E}}|\Omega)_{\gamma} we observe that

B(V)=2(γ)3I(V)+T(V),B(V)=\dfrac{2}{\mathcal{L}(\gamma)^{3}}I(V)+T(V),

where II is the Riesz map I(V)=V,H2I(V)=\left\langle{V,\cdot}\right\rangle_{H^{2}} and T(V):=0(γ)V,τ()𝑑sT(V):=\int^{\mathcal{L(\gamma)}}_{0}\left\langle{V,\tau(\cdot)}\right\rangle\,ds. Employing the Riesz representation theorem in (TγΩ,,H2)(T_{\gamma}\Omega,\langle\cdot,\cdot\rangle_{H^{2}}), we see that the operator II is an isomorphism, and then it is Fredholm with index zero. As for TT, recalling that τ:TγΩL2(S1,n)\tau:T_{\gamma}\Omega\to L^{2}(S^{1},\mathbb{R}^{n}) is bounded, we have

T(V)(TγΩ)=supWTγΩ,WH2=1|0(γ)V,τ(W)𝑑s|CVL2\|T(V)\|_{(T_{\gamma}\Omega)^{*}}=\sup_{W\in T_{\gamma}\Omega,\,\|W\|_{H^{2}}=1}\Bigl{|}\int^{\mathcal{L(\gamma)}}_{0}\left\langle{V,\tau(W)}\right\rangle\,ds\Bigr{|}\leq C\|V\|_{L^{2}}

for all VTγΩV\in T_{\gamma}\Omega. Now it follows from the compactness of the imbedding H2L2H^{2}\subset L^{2} that a bounded sequence (Vi)TγΩ(V_{i})\subset T_{\gamma}\Omega has a subsequence converging in L2(S1,n)L^{2}(S^{1},\mathbb{R}^{n}), and then (TVi)(TV_{i}) has a convergent subsequence because the space ((TγΩ),(TγΩ))((T_{\gamma}\Omega)^{*},\|\cdot\|_{(T_{\gamma}\Omega)^{*}}) is a Banach space. Thus T:TγΩ(TγΩ)T:T_{\gamma}\Omega\to(T_{\gamma}\Omega)^{*} is compact. Then since BB is the sum of a Fredholm index zero operator and a compact operator, it is also Fredholm with index zero (e.g. [45, Example  8.16]). ∎

Proposition 4.13.

(Łojasiewicz–Simon gradient inequality on Ω\Omega). Let ςΩ\varsigma\in\Omega be a stationary point of \operatorname{\mathcal{E}}. There are constants Z(0,),δ(0,1]Z\in(0,\infty),\delta\in(0,1] and θ[12,1)\theta\in\left[\tfrac{1}{2},1\right) such that if αΩ\alpha\in\Omega with αςH2<δ{\left\|{\alpha-\varsigma}\right\|}_{H^{2}}<\delta then

d(|Ω)αTαΩZ|(α)(ς)|θ.\|d(\operatorname{\mathcal{E}}|\Omega)_{\alpha}\|_{T_{\alpha}\Omega^{*}}\geq Z|\operatorname{\mathcal{E}}(\alpha)-\operatorname{\mathcal{E}}(\varsigma)|^{\theta}.
Proof.

Let ϕ:UΩϕ(U)B\phi:U\subset\Omega\to\phi(U)\subset B be a local chart for Ω\Omega with ςU\varsigma\in U, where BB is a subspace of H2(S1,n)H^{2}(S^{1},\mathbb{R}^{n}), and choose δ\delta such that αΩ\alpha\in\Omega is also in UU when αςH2<δ\|\alpha-\varsigma\|_{H^{2}}<\delta. Define E:=ϕ1:ϕ(U)E:=\operatorname{\mathcal{E}}\circ\phi^{-1}:\phi(U)\to\mathbb{R}, then dE=ddϕ1dE=d\operatorname{\mathcal{E}}\circ d\phi^{-1}, and since ς\varsigma is stationary we have

d2Eϕ(ς)=d2ς(dϕϕ(ς)1,dϕϕ(ς)1).d^{2}E_{\phi(\varsigma)}=d^{2}\operatorname{\mathcal{E}}_{\varsigma}(d\phi^{-1}_{\phi(\varsigma)}\cdot,d\phi^{-1}_{\phi(\varsigma)}\cdot).

Moreover, ϕ\phi is analytic and dϕς:TςΩBd\phi_{\varsigma}:T_{\varsigma}\Omega\to B is an isomorphism so it follows from Corollary 4.10 and Proposition 4.12 that EE is analytic and d2Eϕ(ς)d^{2}E_{\phi(\varsigma)} is Fredholm. Therefore by [11, Theorem 1] there exist constants Z~(0,),δ~(0,1],θ[12,1)\tilde{Z}\in(0,\infty),\tilde{\delta}\in(0,1],\theta\in\left[\tfrac{1}{2},1\right) such that if ϕ(α)ϕ(ς)B<δ~\|\phi(\alpha)-\phi(\varsigma)\|_{B}<\tilde{\delta} then

dEϕ(α)BZ~|E(ϕ(α))E(ϕ(ς))|θ=Z~|(α)(ς)|θ.\|dE_{\phi(\alpha)}\|_{B^{*}}\geq\tilde{Z}\left|E(\phi(\alpha))-E(\phi(\varsigma))\right|^{\theta}=\tilde{Z}\left|\operatorname{\mathcal{E}}(\alpha)-\operatorname{\mathcal{E}}(\varsigma)\right|^{\theta}.

Since dϕ1d\phi^{-1} is continuous we have that there is a constant cc such that for any αΩ\alpha\in\Omega with αςH2<δ{\left\|{\alpha-\varsigma}\right\|}_{H^{2}}<\delta and any VTαΩV\in T_{\alpha}\Omega

VH2cdϕα(V)B\|V\|_{H^{2}}\leq c\|d\phi_{\alpha}(V)\|_{B}

and therefore

dEϕ(α)B=supdϕα(V)B|dEϕ(α)dϕα(V)|dϕα(V)BsupVTαΩ|dαV|1cVH2=cd(|Ω)αTαΩ.\|dE_{\phi(\alpha)}\|_{B^{*}}=\sup_{d\phi_{\alpha}(V)\in B}\frac{|dE_{\phi(\alpha)}d\phi_{\alpha}(V)|}{\|d\phi_{\alpha}(V)\|_{B}}\leq\sup_{V\in T_{\alpha}\Omega}\frac{|d\operatorname{\mathcal{E}}_{\alpha}V|}{\frac{1}{c}\|V\|_{H^{2}}}=c\|d(\operatorname{\mathcal{E}}|\Omega)_{\alpha}\|_{{T_{\alpha}\Omega}^{*}}.

The existence of a δ\delta such that if αςH2<δ\|\alpha-\varsigma\|_{H^{2}}<\delta then ϕ(α)ϕ(ς)B<δ~\|\phi(\alpha)-\phi(\varsigma)\|_{B}<\tilde{\delta} follows from the continuity of ϕ\phi and the fact that Ω\Omega is a submanifold of H2(S1,n)H^{2}(S^{1},\mathbb{R}^{n}). ∎

Lemma 4.14.

The projection P:2(S1,n)ΩP:\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})\to\Omega which takes each γ\gamma to its arc length proportional reparametrisation is continuous with respect to H2H^{2} at any σ2(S1,n)\sigma\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) which is stationary for \operatorname{\mathcal{E}}.

Proof.

Write wγ(u)=1(γ)0u|γ(τ)|𝑑τw_{\gamma}(u)=\frac{1}{\mathcal{L}(\gamma)}\int_{0}^{u}|\gamma^{\prime}(\tau)|\,d\tau so α(w):=P(γ)(w)=γwγ1(w)=γ(u)\alpha(w):=P(\gamma)(w)=\gamma\circ w_{\gamma}^{-1}(w)=\gamma(u) and

α(w)\displaystyle\alpha^{\prime}(w) =(γ)γ(u)|γ(u)|=:(γ)Tγ(u),\displaystyle=\mathcal{L}(\gamma)\frac{\gamma^{\prime}(u)}{|\gamma^{\prime}(u)|}=:\mathcal{L}(\gamma)T_{\gamma}(u),
α′′(w)\displaystyle\alpha^{\prime\prime}(w) =γ′′(u)(γ)2|γ(u)|2γ(u)(γ)2|γ(u)|4γ′′(u),γ(u)=(γ)2κγ(u).\displaystyle=\gamma^{\prime\prime}(u)\frac{\mathcal{L}(\gamma)^{2}}{|\gamma^{\prime}(u)|^{2}}-\gamma^{\prime}(u)\frac{\mathcal{L}(\gamma)^{2}}{|\gamma^{\prime}(u)|^{4}}\left\langle{\gamma^{\prime\prime}(u),\gamma^{\prime}(u)}\right\rangle=\mathcal{L}(\gamma)^{2}\kappa_{\gamma}(u).

Let σ2(S1,n)\sigma\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) be a stationary point of \operatorname{\mathcal{E}} and γ2(S1,n)\gamma\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) such that γσH2<b/CS{\left\|{\gamma-\sigma}\right\|}_{H^{2}}<b/C_{S}, where bb and CsC_{s} are as in Lemma 3.1. Then we have constants c1,c2,c3c_{1},c_{2},c_{3} depending only on σ\sigma such that Lemma 3.1(i)-(iii) hold. Using

(43) dwγ1dw=(γ)|γ(w)|,\dfrac{dw^{-1}_{\gamma}}{dw}=\dfrac{\mathcal{{L}(\gamma)}}{|\gamma^{\prime}(w)|},

we obtain the following estimate on parameters

(44) |uωγ1ωσ(u)|\displaystyle|u-\omega_{\gamma}^{-1}\circ\omega_{\sigma}(u)| =|ωγ1ωγ(u)ωγ1ωσ(u)|\displaystyle=|\omega_{\gamma}^{-1}\circ\omega_{\gamma}(u)-\omega_{\gamma}^{-1}\circ\omega_{\sigma}(u)|
=|ωσ(u)ωγ(u)(ωγ1)(τ)𝑑τ|\displaystyle=\Bigl{|}\int_{\omega_{\sigma}(u)}^{\omega_{\gamma}(u)}(\omega_{\gamma}^{-1})^{\prime}(\tau)\,d\tau\Bigr{|}
(γ)|γ|1L0u||γ(τ)|(γ)|σ(τ)|(σ)|𝑑τ\displaystyle\leq\mathcal{L}(\gamma)\||\gamma^{\prime}|^{-1}\|_{L^{\infty}}\int_{0}^{u}\left|\frac{|\gamma^{\prime}(\tau)|}{\mathcal{L}(\gamma)}-\frac{|\sigma^{\prime}(\tau)|}{\mathcal{L}(\sigma)}\right|d\tau
1(σ)|γ|1L(γL+(γ))γσL1\displaystyle\leq\frac{1}{\mathcal{L}(\sigma)}\||\gamma^{\prime}|^{-1}\|_{L^{\infty}}\left({\left\|{\gamma^{\prime}}\right\|}_{L^{\infty}}+\mathcal{L}(\gamma)\right)\|\gamma^{\prime}-\sigma^{\prime}\|_{L^{1}}
2c2c12γσL1\displaystyle\leq\frac{2c_{2}}{c_{1}^{2}}\|\gamma^{\prime}-\sigma^{\prime}\|_{L^{1}}

where we have also used |(γ)(σ)|γσL1|\mathcal{L}(\gamma)-\mathcal{L}(\sigma)|\leq{\left\|{\gamma-\sigma}\right\|}_{L^{1}}. Then

P(σ)P(γ)L22\displaystyle\|P(\sigma)-P(\gamma)\|^{2}_{L^{2}} =01|σωσ1(w)γωγ1(w)|2𝑑w\displaystyle=\int_{0}^{1}|\sigma\circ\omega_{\sigma}^{-1}(w)-\gamma\circ\omega_{\gamma}^{-1}(w)|^{2}\,dw
201|σ(u)σωγ1ωσ(u)|2|σ(u)|(σ)𝑑u+201|σωγ1(w)γωγ1(w)|2𝑑w\displaystyle\leq\begin{multlined}2\int_{0}^{1}|\sigma(u)-\sigma\circ\omega_{\gamma}^{-1}\circ\omega_{\sigma}(u)|^{2}\frac{|\sigma^{\prime}(u)|}{\mathcal{L}(\sigma)}\,du\\ +2\int_{0}^{1}|\sigma\circ\omega_{\gamma}^{-1}(w)-\gamma\circ\omega_{\gamma}^{-1}(w)|^{2}\,dw\end{multlined}2\int_{0}^{1}|\sigma(u)-\sigma\circ\omega_{\gamma}^{-1}\circ\omega_{\sigma}(u)|^{2}\frac{|\sigma^{\prime}(u)|}{\mathcal{L}(\sigma)}\,du\\ +2\int_{0}^{1}|\sigma\circ\omega_{\gamma}^{-1}(w)-\gamma\circ\omega_{\gamma}^{-1}(w)|^{2}\,dw
2σL3(σ)01|uωγ1ωσ(u)|2𝑑u+2γL(γ)σγL22\displaystyle\leq\frac{2\|\sigma^{\prime}\|_{L^{\infty}}^{3}}{\mathcal{L}(\sigma)}\int_{0}^{1}|u-\omega_{\gamma}^{-1}\circ\omega_{\sigma}(u)|^{2}\,du+\frac{2\|\gamma^{\prime}\|_{L^{\infty}}}{\mathcal{L}(\gamma)}\|\sigma-\gamma\|_{L^{2}}^{2}
8c5γσL12+2cσγL22,\displaystyle\leq 8c^{5}\|\gamma^{\prime}-\sigma^{\prime}\|_{L^{1}}^{2}+2c\|\sigma-\gamma\|_{L^{2}}^{2},

where c:=c2/c1c:=c_{2}/c_{1}. For the difference of first derivatives we have

P(σ)P(γ)L22\displaystyle\|P(\sigma)^{\prime}-P(\gamma)^{\prime}\|_{L^{2}}^{2} =01|(σ)Tσωσ1(w)(γ)Tγωγ1(w)|2𝑑w\displaystyle=\int_{0}^{1}|\mathcal{L}(\sigma)T_{\sigma}\circ\omega_{\sigma}^{-1}(w)-\mathcal{L}(\gamma)T_{\gamma}\circ\omega_{\gamma}^{-1}(w)|^{2}\,dw
2(σ)201|Tσωσ1(w)Tσωγ1(w)|2𝑑w\displaystyle\leq 2\mathcal{L}(\sigma)^{2}\int_{0}^{1}|T_{\sigma}\circ\omega_{\sigma}^{-1}(w)-T_{\sigma}\circ\omega_{\gamma}^{-1}(w)|^{2}\,dw
+201|(σ)Tσωγ1(w)(γ)Tγωγ1(w)|2𝑑w.\displaystyle\qquad+2\int_{0}^{1}|\mathcal{L}(\sigma)T_{\sigma}\circ\omega_{\gamma}^{-1}(w)-\mathcal{L}(\gamma)T_{\gamma}\circ\omega_{\gamma}^{-1}(w)|^{2}\,dw.

For the first term on the right hand side of the inequality, using a change of variable, the fundamental theorem of calculus and (44) we have

(σ)201\displaystyle\mathcal{L}(\sigma)^{2}\int_{0}^{1} |Tσωσ1(w)Tσωγ1(w)|2dw\displaystyle|T_{\sigma}\circ\omega_{\sigma}^{-1}(w)-T_{\sigma}\circ\omega_{\gamma}^{-1}(w)|^{2}\,dw
(σ)σLTσL2201|uωγ1ωσ(u)|2𝑑u\displaystyle\leq\mathcal{L}(\sigma)\|\sigma^{\prime}\|_{L^{\infty}}\|T_{\sigma}^{\prime}\|^{2}_{L^{2}}\int_{0}^{1}|u-\omega_{\gamma}^{-1}\circ\omega_{\sigma}(u)|^{2}\,du
4c4c22c32γσL22,\displaystyle\leq 4c^{4}c^{2}_{2}c_{3}^{2}\|\gamma^{\prime}-\sigma^{\prime}\|_{L^{2}}^{2},

where we have used Tσ=|σ|κσT_{\sigma}^{\prime}={\left|{\sigma^{\prime}}\right|}\kappa_{\sigma}. For the second term, recalling (14)

01|(σ)Tσωγ1(w)(γ)Tγωγ1(w)|2𝑑w\displaystyle\int_{0}^{1}|\mathcal{L}(\sigma)T_{\sigma}\circ\omega_{\gamma}^{-1}(w)-\mathcal{L}(\gamma)T_{\gamma}\circ\omega_{\gamma}^{-1}(w)|^{2}\,dw
2γL(γ)01[|(σ)(γ)|2+(γ)2|Tσ(u)Tγ(u)|2]𝑑u\displaystyle\qquad\leq\frac{2\|\gamma^{\prime}\|_{L^{\infty}}}{\mathcal{L}(\gamma)}\int_{0}^{1}\bigl{[}|\mathcal{L}(\sigma)-\mathcal{L}(\gamma)|^{2}+\mathcal{L}(\gamma)^{2}|T_{\sigma}(u)-T_{\gamma}(u)|^{2}\bigr{]}\,du
2(c+2c3)σγL22.\displaystyle\qquad\leq 2(c+2c^{3})\|\sigma^{\prime}-\gamma^{\prime}\|_{L^{2}}^{2}.

Similarly for the difference of second derivatives

P(σ)′′P(γ)′′L22\displaystyle\|P(\sigma)^{\prime\prime}-P(\gamma)^{\prime\prime}\|_{L^{2}}^{2} =01|(σ)2κσωσ1(w)(γ)2κγωγ1(w)|2𝑑w\displaystyle=\int_{0}^{1}|\mathcal{L}(\sigma)^{2}\kappa_{\sigma}\circ\omega_{\sigma}^{-1}(w)-\mathcal{L}(\gamma)^{2}\kappa_{\gamma}\circ\omega_{\gamma}^{-1}(w)|^{2}\,dw
2(σ)3σLκσL2201|uωγ1ωσ(u)|2𝑑u+4γL(γ)κσL22((σ)+(γ))2σγL22+4γL(γ)3κσκγL22\displaystyle\leq\begin{multlined}2\mathcal{L}(\sigma)^{3}\|\sigma^{\prime}\|_{L^{\infty}}\|\kappa_{\sigma}^{\prime}\|_{L^{2}}^{2}\int_{0}^{1}|u-\omega_{\gamma}^{-1}\circ\omega_{\sigma}(u)|^{2}\,du\\ +4\frac{\|\gamma^{\prime}\|_{L^{\infty}}}{\mathcal{L}(\gamma)}\|\kappa_{\sigma}\|_{L^{2}}^{2}(\mathcal{L}(\sigma)+\mathcal{L}(\gamma))^{2}\|\sigma^{\prime}-\gamma^{\prime}\|_{L^{2}}^{2}\\ +4\|\gamma^{\prime}\|_{L^{\infty}}\mathcal{L}(\gamma)^{3}\|\kappa_{\sigma}-\kappa_{\gamma}\|^{2}_{L^{2}}\end{multlined}2\mathcal{L}(\sigma)^{3}\|\sigma^{\prime}\|_{L^{\infty}}\|\kappa_{\sigma}^{\prime}\|_{L^{2}}^{2}\int_{0}^{1}|u-\omega_{\gamma}^{-1}\circ\omega_{\sigma}(u)|^{2}\,du\\ +4\frac{\|\gamma^{\prime}\|_{L^{\infty}}}{\mathcal{L}(\gamma)}\|\kappa_{\sigma}\|_{L^{2}}^{2}(\mathcal{L}(\sigma)+\mathcal{L}(\gamma))^{2}\|\sigma^{\prime}-\gamma^{\prime}\|_{L^{2}}^{2}\\ +4\|\gamma^{\prime}\|_{L^{\infty}}\mathcal{L}(\gamma)^{3}\|\kappa_{\sigma}-\kappa_{\gamma}\|^{2}_{L^{2}}
4cc22(κσL22+2c32)σγL22+4c24κσκγL22.\displaystyle\leq\begin{multlined}4cc_{2}^{2}\left({\left\|{\kappa_{\sigma}^{\prime}}\right\|}_{L^{2}}^{2}+2c_{3}^{2}\right){\left\|{\sigma^{\prime}-\gamma^{\prime}}\right\|}_{L^{2}}^{2}+4c_{2}^{4}{\left\|{\kappa_{\sigma}-\kappa_{\gamma}}\right\|}_{L^{2}}^{2}.\end{multlined}4cc_{2}^{2}\left({\left\|{\kappa_{\sigma}^{\prime}}\right\|}_{L^{2}}^{2}+2c_{3}^{2}\right){\left\|{\sigma^{\prime}-\gamma^{\prime}}\right\|}_{L^{2}}^{2}+4c_{2}^{4}{\left\|{\kappa_{\sigma}-\kappa_{\gamma}}\right\|}_{L^{2}}^{2}.

We have assumed σ\sigma is stationary so that, by the proof of Lemma 4.1, κσL2{\left\|{\kappa_{\sigma}^{\prime}}\right\|}_{L^{2}} is bounded. Finally, we recall that by Lemma 3.1 (iii) κ\kappa is Lipschitz, and this completes the proof. ∎

Theorem 4.15.

(Gradient inequality on 2(S1,n)\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})) Let σ2(S1,n)\sigma\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) be a stationary point of \operatorname{\mathcal{E}}. Then there are constants Z(0,),δ(0,1]Z\in(0,\infty),\delta\in(0,1] and θ[12,1)\theta\in\left[\tfrac{1}{2},1\right) such that if γ2(S1,n)\gamma\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) with γσH2<δ\|\gamma-\sigma\|_{H^{2}}<\delta then

gradγH2(ds),γZ|(γ)(σ)|θ.\|\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}\|_{H^{2}(ds),\gamma}\geq Z|\operatorname{\mathcal{E}}(\gamma)-\operatorname{\mathcal{E}}(\sigma)|^{\theta}.
Proof.

Let α,ςΩ\alpha,\varsigma\in\Omega be the respective arc length proportional reparametrisations of γ,σ\gamma,\sigma. Then since \operatorname{\mathcal{E}} and dd\operatorname{\mathcal{E}} are parametrisation invariant and d(|Ω)αTγΩdαH2\|d(\operatorname{\mathcal{E}}|\Omega)_{\alpha}\|_{T_{\gamma}\Omega^{*}}\leq\|d\operatorname{\mathcal{E}}_{\alpha}\|_{{H^{2}}^{*}} we have by Proposition 4.13

dαH2d(|Ω)αTγΩZ|(α)(ς)|θ=Z|(γ)(σ)|θ\|d\operatorname{\mathcal{E}}_{\alpha}\|_{{H^{2}}^{*}}\geq\|d(\operatorname{\mathcal{E}}|\Omega)_{\alpha}\|_{T_{\gamma}\Omega^{*}}\geq Z|\operatorname{\mathcal{E}}(\alpha)-\operatorname{\mathcal{E}}(\varsigma)|^{\theta}=Z|\operatorname{\mathcal{E}}(\gamma)-\operatorname{\mathcal{E}}(\sigma)|^{\theta}

provided αςH2\|\alpha-\varsigma\|_{H^{2}} is sufficiently small, which can be arranged according to Lemma 4.14 because σ\sigma is stationary. Since reparametrisation is a linear map on H2H^{2} we have dα(V)=dγ(Vωγ)d\operatorname{\mathcal{E}}_{\alpha}(V)=d\operatorname{\mathcal{E}}_{\gamma}(V\circ\omega_{\gamma}) and then

(45) dαH2\displaystyle\|d\operatorname{\mathcal{E}}_{\alpha}\|_{{H^{2}}^{*}} =supVH2=1|dγ(Vωγ)|\displaystyle=\sup_{\|V\|_{H^{2}}=1}|d\operatorname{\mathcal{E}}_{\gamma}(V\circ\omega_{\gamma})|
=supVH2=1gradγ,VωγH2(ds),γ.\displaystyle=\sup_{\|V\|_{H^{2}}=1}\left\langle{\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma},V\circ\omega_{\gamma}}\right\rangle_{H^{2}(ds),\gamma}.

From (43) we calculate

(46) VωγH2(ds),γ2=(γ)VL22+1(γ)VL22+1(γ)3V′′L22.\|V\circ\omega_{\gamma}\|^{2}_{H^{2}(ds),\gamma}=\mathcal{L}(\gamma)\|V\|_{L^{2}}^{2}+\dfrac{1}{\mathcal{L}(\gamma)}\|V^{\prime}\|_{L^{2}}^{2}+\frac{1}{\mathcal{L}(\gamma)^{3}}\|V^{\prime\prime}\|_{L^{2}}^{2}.

By Lemma 3.1 we have upper and lower bounds for (γ)\mathcal{L}(\gamma) and therefore a constant c1c_{1} such that

dαH2c1gradγH2(ds).\|d\operatorname{\mathcal{E}}_{\alpha}\|_{{H^{2}}^{*}}\leq c_{1}\|\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}\|_{H^{2}(ds)}.

5. Convergence

Since the Łojasiewicz–Simon gradient inequality proved above only holds in an H2H^{2}-neighbourhood of a critical point we will need subconvergence in H2H^{2} of minimizing sequences in order to use it. To this end we will prove a Palais–Smale type condition for |Ω\operatorname{\mathcal{E}}|\Omega by adapting the method used in [39].

First let us define an auxiliary functional J:2(S1,n)J:\mathcal{I}^{2}(S^{1},\mathbb{R}^{n})\to\mathbb{R} by

(47) J(γ):=1(γ)301|γ′′|2𝑑u+λ2(γ).J(\gamma):=\frac{1}{\mathcal{L}(\gamma)^{3}}\int_{0}^{1}{\left|{\gamma^{\prime\prime}}\right|}^{2}\,du+\lambda^{2}\mathcal{L}(\gamma)\,.

This function has the property J|Ω=|ΩJ|\Omega=\operatorname{\mathcal{E}}|\Omega, and also the following.

Lemma 5.1.

JJ is locally coercive modulo C1C^{1} in the following sense. Let UU be an open neighbourhood in 2(S1,n)\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) for which there are constants c1,c2c_{1},c_{2} such that for all γU\gamma\in U: γH2<c1\|\gamma\|_{H^{2}}<c_{1} and 0<c2<|γ(u)|0<c_{2}<{\left|{\gamma^{\prime}(u)}\right|}. Then there exist positive constants c3c_{3} and c4c_{4} depending on UU such that for any VH2(S1,n)V\in H^{2}(S^{1},\mathbb{R}^{n})

(48) d2Jγ(V,V)c3VH22c4VC12.\displaystyle d^{2}J_{\gamma}(V,V)\geq c_{3}\|V\|_{H^{2}}^{2}-c_{4}\|V\|_{C^{1}}^{2}.
Proof.

To calculate the second derivative of JJ first calculate

dγV=01V,T𝑑ud\mathcal{L}_{\gamma}V=\int_{0}^{1}\left\langle{V^{\prime},T}\right\rangle\,du

and then

d2γ(V,V)=011|γ||V|21|γ|3V,γ2du.d^{2}\mathcal{L}_{\gamma}(V,V)=\int_{0}^{1}\frac{1}{{\left|{\gamma^{\prime}}\right|}}{\left|{V^{\prime}}\right|}^{2}-\frac{1}{{\left|{\gamma^{\prime}}\right|}^{3}}\left\langle{V^{\prime},\gamma^{\prime}}\right\rangle^{2}\,du.

Using the assumptions on UU there are constants c¯,c~\bar{c},\tilde{c} such that

(49) |dγV|\displaystyle{\left|{d\mathcal{L}_{\gamma}V}\right|} c¯VL1,\displaystyle\leq\bar{c}{\left\|{V^{\prime}}\right\|}_{L^{1}},
|d2γ(V,V)|\displaystyle{\left|{d^{2}\mathcal{L}_{\gamma}(V,V)}\right|} c~VL22.\displaystyle\leq\tilde{c}{\left\|{V^{\prime}}\right\|}_{L^{2}}^{2}.

The derivatives of JJ are

(50) dJγV\displaystyle dJ_{\gamma}V =3(γ)4dγ(V)01|γ′′|2𝑑u+2(γ)301V′′,γ′′𝑑u+λ2dγV,\displaystyle=\frac{-3}{\mathcal{L}(\gamma)^{4}}d\mathcal{L}_{\gamma}(V)\int_{0}^{1}{\left|{\gamma^{\prime\prime}}\right|}^{2}\,du+\frac{2}{\mathcal{L}(\gamma)^{3}}\int_{0}^{1}\left\langle{V^{\prime\prime},\gamma^{\prime\prime}}\right\rangle\,du+\lambda^{2}d\mathcal{L}_{\gamma}V,

and

d2Jγ\displaystyle d^{2}J_{\gamma} (V,V)\displaystyle(V,V)
=12(γ)5(dγV)2γ′′L223(γ)4d2γ(V,V)γ′′L22\displaystyle=-\frac{12}{\mathcal{L}(\gamma)^{5}}(d\mathcal{L}_{\gamma}V)^{2}{\left\|{\gamma^{\prime\prime}}\right\|}_{L^{2}}^{2}-\frac{3}{\mathcal{L}(\gamma)^{4}}d^{2}\mathcal{L}_{\gamma}(V,V){\left\|{\gamma^{\prime\prime}}\right\|}_{L^{2}}^{2}
12(γ)4dγV01V′′,γ′′𝑑u+2(γ)3V′′L22+λ2d2γ(V,V),\displaystyle\qquad-\frac{12}{\mathcal{L}(\gamma)^{4}}d\mathcal{L}_{\gamma}V\int_{0}^{1}\left\langle{V^{\prime\prime},\gamma^{\prime\prime}}\right\rangle\,du+\frac{2}{\mathcal{L}(\gamma)^{3}}{\left\|{V^{\prime\prime}}\right\|}_{L^{2}}^{2}+\lambda^{2}d^{2}\mathcal{L}_{\gamma}(V,V),

and then using (49) and the assumptions on UU again, there are positive constants a1,a2,a3a_{1},a_{2},a_{3} such that

d2Jγ(V,V)\displaystyle d^{2}J_{\gamma}(V,V) a1V′′L22a2VL22a3VL2V′′L2\displaystyle\geq a_{1}{\left\|{V^{\prime\prime}}\right\|}_{L^{2}}^{2}-a_{2}{\left\|{V^{\prime}}\right\|}_{L^{2}}^{2}-a_{3}{\left\|{V^{\prime}}\right\|}_{L^{2}}{\left\|{V^{\prime\prime}}\right\|}_{L^{2}}
a1VH22(a1+a2)VH12a3VL2V′′L2.\displaystyle\geq a_{1}{\left\|{V}\right\|}_{H^{2}}^{2}-(a_{1}+a_{2}){\left\|{V}\right\|}^{2}_{H^{1}}-a_{3}{\left\|{V^{\prime}}\right\|}_{L^{2}}{\left\|{V^{\prime\prime}}\right\|}_{L^{2}}.

If we apply the inequality 2abεa2+1εb2,ε>02ab\leq\varepsilon a^{2}+\frac{1}{\varepsilon}b^{2},\varepsilon>0 to the last term, choosing ε\varepsilon sufficiently large and using a Sobolev imbedding we obtain (48). ∎

Corollary 5.2.

If U2(S1,n)U\subset\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}) satisfies the same conditions as in Lemma 5.1 and is convex, then there exist constants c1,c2c_{1},c_{2} such that for any γ,βU\gamma,\beta\in U

(51) (dJβdJγ)(βγ)c1βγH22c2βγC12.\bigl{(}dJ_{\beta}-dJ_{\gamma}\bigr{)}(\beta-\gamma)\geq c_{1}\|\beta-\gamma\|_{H^{2}}^{2}-c_{2}\|\beta-\gamma\|_{C^{1}}^{2}.
Proof.

Since γ+t(βγ)U\gamma+t(\beta-\gamma)\in U for all t[0,1]t\in[0,1], (48) holds and

(dJβdJγ)(βγ)\displaystyle\bigl{(}dJ_{\beta}-dJ_{\gamma}\bigr{)}(\beta-\gamma) =01ddt𝑑Jγ+t(βγ)(βγ)𝑑t\displaystyle=\int_{0}^{1}\frac{d}{dt}dJ_{\gamma+t(\beta-\gamma)}(\beta-\gamma)\,dt
=01d2Jγ+t(βγ)(βγ,βγ)𝑑t\displaystyle=\int_{0}^{1}d^{2}J_{\gamma+t(\beta-\gamma)}(\beta-\gamma,\beta-\gamma)\,dt
c1βγH22c2βγC1.\displaystyle\geq c_{1}\|\beta-\gamma\|_{H^{2}}^{2}-c_{2}\|\beta-\gamma\|_{C^{1}}.

Next we construct a continuous projection onto TαΩT_{\alpha}\Omega using a right inverse for the map dΦd\Phi from Section 4.3. For this we just need to choose initial conditions for the construction described in the proof of Lemma 4.9. We define rα:Hzm2(S1,)H2(S1,n)r_{\alpha}:H^{2}_{zm}(S^{1},\mathbb{R})\to H^{2}(S^{1},\mathbb{R}^{n}) by

(52) rαw\displaystyle r_{\alpha}w :=yx,where\displaystyle:=y-x,\quad\text{where}
y\displaystyle y^{\prime} =wα|α|,y(0)=0,\displaystyle=w\frac{\alpha^{\prime}}{{\left|{\alpha^{\prime}}\right|}},\quad y(0)=0,
x\displaystyle x^{\prime} =Bξ,x(0)=0,ξ=BTW1y(1),\displaystyle=B\xi,\quad x(0)=0,\quad\xi=B^{T}W^{-1}y(1),

and BB and WW are the matrices from (40). Let us confirm that rr has the desired properties. It follows from (35) that

(53) dΦαrαw=(rαw),α|α|01(rαw)α|α|𝑑u=w01w𝑑u=w.d\Phi_{\alpha}r_{\alpha}w=\left\langle{\left(r_{\alpha}w\right)^{\prime},\frac{\alpha^{\prime}}{{\left|{\alpha^{\prime}}\right|}}}\right\rangle-\int_{0}^{1}(r_{\alpha}w)^{\prime}\frac{\alpha^{\prime}}{{\left|{\alpha^{\prime}}\right|}}\,du=w-\int_{0}^{1}w\,du=w.

Moreover, rαw(0)=0r_{\alpha}w(0)=0 and

rαw(1)=y(1)01x𝑑u\displaystyle r_{\alpha}w(1)=y(1)-\int_{0}^{1}x^{\prime}\,du =y(1)01BBTW1y(1)𝑑u\displaystyle=y(1)-\int_{0}^{1}BB^{T}W^{-1}y(1)\,du
=y(1)WW1y(1)\displaystyle=y(1)-WW^{-1}y(1)
=0.\displaystyle=0.

Thus rαwr_{\alpha}w is periodic. Periodicity of (rαw)(r_{\alpha}w)^{\prime} follows from that of α,w\alpha^{\prime},w and β\beta.

Now for each αΩ\alpha\in\Omega we define prTαΩ:H2(S1,n)TαΩ\operatorname{\rm{pr}}_{T_{\alpha}\Omega}:H^{2}(S^{1},\mathbb{R}^{n})\to T_{\alpha}\Omega by

(54) prTαΩV:=(1rαdΦα)V.\operatorname{\rm{pr}}_{T_{\alpha}\Omega}V:=(1-r_{\alpha}d\Phi_{\alpha})V.

Indeed dΦαprTαΩV=dΦαVdΦαrαdΦαV=0d\Phi_{\alpha}\operatorname{\rm{pr}}_{T_{\alpha}\Omega}V=d\Phi_{\alpha}V-d\Phi_{\alpha}r_{\alpha}d\Phi_{\alpha}V=0 by (53) and so prTαΩkerdΦα=TαΩ\operatorname{\rm{pr}}_{T_{\alpha}\Omega}\in\ker d\Phi_{\alpha}=T_{\alpha}\Omega.

The following lemma will be used in Proposition 5.4 to estimate terms involving the projection onto TαΩT_{\alpha}\Omega. Its proof requires some estimates on the matrix WW, which we have included in Appendix A.

Lemma 5.3.

Let UU be an H2H^{2}-bounded subset of Ω\Omega for which there exists c1c_{1} such that 0<c1<|α(u)|0<c_{1}<{\left|{\alpha^{\prime}(u)}\right|} for all αU\alpha\in U. Then there exists a constant cc such that for all αU\alpha\in U and vH2v\in H^{2}

(55) |dJαrαdΦαv|cvC1.{\left|{dJ_{\alpha}r_{\alpha}d\Phi_{\alpha}v}\right|}\leq c{\left\|{v}\right\|}_{C^{1}}.
Proof.

For this first we note from (52) that |y(1)|wL{\left|{y(1)}\right|}\leq{\left\|{w}\right\|}_{L^{\infty}}, and since the νi\nu_{i} are orthonormal, we deduce from Lemma A.2 that

|(rαw)||w|+W1|y(1)|cwL.{\left|{(r_{\alpha}w)^{\prime}}\right|}\leq{\left|{w}\right|}+{\left\|{W^{-1}}\right\|}{\left|{y(1)}\right|}\leq c{\left\|{w}\right\|}_{L^{\infty}}.

Moreover, it follows from (35) that |dΦαv|2vL{\left|{d\Phi_{\alpha}v}\right|}\leq 2{\left\|{v^{\prime}}\right\|}_{L^{\infty}}, and then

(56) |(rαdΦαv)|cdΦαvL2cvL.{\left|{(r_{\alpha}d\Phi_{\alpha}v)^{\prime}}\right|}\leq c{\left\|{d\Phi_{\alpha}v}\right\|}_{L^{\infty}}\leq 2c{\left\|{v^{\prime}}\right\|}_{L^{\infty}}.

From (52) again, recalling that |α|=(α){\left|{\alpha^{\prime}}\right|}=\mathcal{L}(\alpha), calculate the second derivative

(57) (rαdΦαv)′′=(dΦαv)T+dΦαvα′′(α)+(Bξ).(r_{\alpha}d\Phi_{\alpha}v)^{\prime\prime}=(d\Phi_{\alpha}v)^{\prime}T+d\Phi_{\alpha}v\frac{\alpha^{\prime\prime}}{\mathcal{L}(\alpha)}+(B\xi)^{\prime}\,.

Note that because |α|{\left|{\alpha^{\prime}}\right|} is constant T,α′′=0\left\langle{T,\alpha^{\prime\prime}}\right\rangle=0 and then from (50),

dJα\displaystyle dJ_{\alpha} rαdΦαv\displaystyle r_{\alpha}d\Phi_{\alpha}v
=3(α)4dα(rαdΦαv)01|α′′|2𝑑u+2(α)301(rαdΦαv)′′,α′′𝑑u\displaystyle=\frac{-3}{\mathcal{L}(\alpha)^{4}}d\mathcal{L}_{\alpha}(r_{\alpha}d\Phi_{\alpha}v)\int_{0}^{1}{\left|{\alpha^{\prime\prime}}\right|}^{2}\,du+\frac{2}{\mathcal{L}(\alpha)^{3}}\int_{0}^{1}\left\langle{(r_{\alpha}d\Phi_{\alpha}v)^{\prime\prime},\alpha^{\prime\prime}}\right\rangle\,du
+λ2dαrαdΦαv\displaystyle\qquad\qquad+\lambda^{2}d\mathcal{L}_{\alpha}r_{\alpha}d\Phi_{\alpha}v
(58) =3(α)4dα(rαdΦαv)α′′L22+2(α)301dΦαv(α)|α′′|2+(Bξ),α′′du\displaystyle=\frac{-3}{\mathcal{L}(\alpha)^{4}}d\mathcal{L}_{\alpha}(r_{\alpha}d\Phi_{\alpha}v){\left\|{\alpha^{\prime\prime}}\right\|}^{2}_{L^{2}}+\frac{2}{\mathcal{L}(\alpha)^{3}}\int_{0}^{1}\frac{d\Phi_{\alpha}v}{\mathcal{L}(\alpha)}{\left|{\alpha^{\prime\prime}}\right|}^{2}+\left\langle{(B\xi)^{\prime},\alpha^{\prime\prime}}\right\rangle\,du
+λ2dαrαdΦα.\displaystyle\qquad\qquad+\lambda^{2}d\mathcal{L}_{\alpha}r_{\alpha}d\Phi_{\alpha}.

From the definition (40) of BB we have

(Bξ)=βiνiξi+βiνiξi+Bξ,(B\xi)^{\prime}=\beta^{\prime}\sum_{i}\nu_{i}\xi_{i}+\beta\sum_{i}\nu_{i}^{\prime}\xi_{i}+B\xi^{\prime},

and from (37), νi,α′′=0\left\langle{\nu_{i}^{\prime},\alpha^{\prime\prime}}\right\rangle=0, so recalling that β\beta is smooth and the νi\nu_{i} are normalised, we have

|(Bξ),α′′|c1|ξ|+c2|ξ|.{\left|{\left\langle{(B\xi)^{\prime},\alpha^{\prime\prime}}\right\rangle}\right|}\leq c_{1}{\left|{\xi}\right|}+c_{2}{\left|{\xi^{\prime}}\right|}.

Now as in (56), we have |ξ|W1|y(1)|2cvL{\left|{\xi}\right|}\leq{\left\|{W^{-1}}\right\|}{\left|{y(1)}\right|}\leq 2c{\left\|{v^{\prime}}\right\|}_{L^{\infty}}. Since

ξ=(β[ν1νn1]T+β[ν1νn1]T)W1y(1),\xi^{\prime}=\left(\beta^{\prime}[\nu_{1}\ldots\nu_{n-1}]^{T}+\beta[\nu_{1}^{\prime}\ldots\nu_{n-1}^{\prime}]^{T}\right)W^{-1}y(1),

using (37) to estimate νi\nu_{i}^{\prime}, we obtain

|ξ|(c1+c2|α′′|)vL.{\left|{\xi^{\prime}}\right|}\leq\left(c_{1}+c_{2}{\left|{\alpha^{\prime\prime}}\right|}\right){\left\|{v^{\prime}}\right\|}_{L^{\infty}}.

Thus it follows that |(Bξ),α′′|(c1+c2|α′′|)|α′′|vL{\left|{\left\langle{(B\xi)^{\prime},\alpha^{\prime\prime}}\right\rangle}\right|}\leq\left(c_{1}+c_{2}{\left|{\alpha^{\prime\prime}}\right|}\right){\left|{\alpha^{\prime\prime}}\right|}{\left\|{v^{\prime}}\right\|}_{L^{\infty}}. Combining this into (5) with the estimates (49), (56), and |dΦαv|vL{\left|{d\Phi_{\alpha}v}\right|}\leq{\left\|{v^{\prime}}\right\|}_{L^{\infty}}, we see that

|dJαrαdΦαv|cvL.\displaystyle{\left|{dJ_{\alpha}r_{\alpha}d\Phi_{\alpha}v}\right|}\leq c{\left\|{v^{\prime}}\right\|}_{L^{\infty}}.

Therefore Lemma 5.3 follows. ∎

Now we can prove the following Palais–Smale type condition for |Ω\operatorname{\mathcal{E}}|\Omega (it is not quite the standard condition because we need to assume an L2L^{2}-bound on the sequence).

Proposition 5.4.

Let (αi)(\alpha_{i}) be a sequence of curves in Ω\Omega such that (αi)\operatorname{\mathcal{E}}(\alpha_{i}) and αiL2\|\alpha_{i}\|_{L^{2}} are bounded, and d(αi)0{\left\|{d(\operatorname{\mathcal{E}}_{\alpha_{i}})}\right\|}\to 0. Then (αi)(\alpha_{i}) has a subsequence converging in H2H^{2}.

Proof.

The assumed bounds on (αi)\operatorname{\mathcal{E}}(\alpha_{i}) and αiL2\|\alpha_{i}\|_{L^{2}} together imply an H2H^{2}-bound and then by compactness of the Sobolev imbedding H2C1H^{2}\subset C^{1} we have a subsequence, still denoted (αi)(\alpha_{i}), which converges in C1C^{1} to some α\alpha_{\infty}. Let ε>0\varepsilon>0 be sufficiently small so that the open C1C^{1}-ball BεC1(α)B_{\varepsilon}^{C^{1}}(\alpha_{\infty}) contains only immersions (c.f. Lemma 3.1) and choose NN sufficiently large that for all i>Ni>N, αiBεC1(α)\alpha_{i}\in B_{\varepsilon}^{C^{1}}(\alpha_{\infty}). Then for i>Ni>N, which we assume from now on, αi\alpha_{i} is contained in a neighbourhood UU satisfying the assumptions of Corollary 5.2. Hence from (51) we get

(dJαjdJαi)(αjαi)c1αiαjH22c2αiαjC12\bigl{(}dJ_{\alpha_{j}}-dJ_{\alpha_{i}}\bigr{)}(\alpha_{j}-\alpha_{i})\geq c_{1}\|\alpha_{i}-\alpha_{j}\|_{H^{2}}^{2}-c_{2}\|\alpha_{i}-\alpha_{j}\|_{C^{1}}^{2}\,

with c1>0c_{1}>0. From (54) and J|Ω=|ΩJ|\Omega=\operatorname{\mathcal{E}}|\Omega

(dJαjdJαi)(αjαi)\displaystyle\bigl{(}dJ_{\alpha_{j}}-dJ_{\alpha_{i}}\bigr{)}(\alpha_{j}-\alpha_{i}) =(dαjprTαjΩdαiprTαiΩ)(αjαi)\displaystyle=\bigl{(}d\operatorname{\mathcal{E}}_{\alpha_{j}}\operatorname{\rm{pr}}_{T_{\alpha_{j}}\Omega}-d\operatorname{\mathcal{E}}_{\alpha_{i}}\operatorname{\rm{pr}}_{T_{\alpha_{i}}\Omega}\bigr{)}(\alpha_{j}-\alpha_{i})
+(dJαjrαjdΦαjdJαirαidΦαi)(αjαi).\displaystyle\qquad+(dJ_{\alpha_{j}}r_{\alpha_{j}}d\Phi_{\alpha_{j}}-dJ_{\alpha_{i}}r_{\alpha_{i}}d\Phi_{\alpha_{i}})(\alpha_{j}-\alpha_{i}).

Then rearranging the inequality

c1αiαjH22\displaystyle c_{1}\|\alpha_{i}-\alpha_{j}\|_{H^{2}}^{2} c2αiαjC12+(dαjprTαjΩdαiprTαiΩ)(αjαi)\displaystyle\leq c_{2}\|\alpha_{i}-\alpha_{j}\|_{C^{1}}^{2}+\bigl{(}d\operatorname{\mathcal{E}}_{\alpha_{j}}\operatorname{\rm{pr}}_{T_{\alpha_{j}}\Omega}-d\operatorname{\mathcal{E}}_{\alpha_{i}}\operatorname{\rm{pr}}_{T_{\alpha_{i}}\Omega}\bigr{)}(\alpha_{j}-\alpha_{i})
+(dJαjrαjdΦαjdJαirαidΦαi)(αjαi).\displaystyle\qquad+(dJ_{\alpha_{j}}r_{\alpha_{j}}d\Phi_{\alpha_{j}}-dJ_{\alpha_{i}}r_{\alpha_{i}}d\Phi_{\alpha_{i}})(\alpha_{j}-\alpha_{i}).

Now by Lemma 5.3, since d(αi)0{\left\|{d(\operatorname{\mathcal{E}}_{\alpha_{i}})}\right\|}\to 0 by assumption and αi\alpha_{i} converges in C1C^{1}, the right hand side of the above inequality converges to zero and αi\alpha_{i} converges in H2H^{2}. ∎

Theorem 5.5.

Let γ\gamma be a solution to the H2(ds)H^{2}(ds)-gradient flow of the modified elastic energy \operatorname{\mathcal{E}}. Then there is a stationary point γH2(S1,n)\gamma_{\infty}\in H^{2}(S^{1},\mathbb{R}^{n}) such that γ(t)γ\gamma(t)\to\gamma_{\infty} in H2H^{2} as tt\to\infty.

Proof.

Consider the projected and translated flow

α(t):=P(γ(t))1(γ(t))0(γ(t))γ(t)𝑑s,\alpha(t):=P(\gamma(t))-\frac{1}{\mathcal{L}(\gamma(t))}\int^{\mathcal{L}(\gamma(t))}_{0}\gamma(t)\,ds,

where as in Lemma 4.14, P(γ(t))P(\gamma(t)) is the arc length proportional reparametrisation of γ(t)\gamma(t). From parametrisation and translation invariance of the energy we have λ2(α)<(α)=(γ)(γ(0))\lambda^{2}\mathcal{L}(\alpha)<\operatorname{\mathcal{E}}(\alpha)=\operatorname{\mathcal{E}}(\gamma)\leq\operatorname{\mathcal{E}}(\gamma(0)). Moreover, using the estimates in Lemma 4.14 and the Poincaré–Wirtinger inequality, we see that α(t)L2\|\alpha(t)\|_{L^{2}} is also bounded. From (27) (with TT\to\infty) there exists a monotone, divergent sequence (ti)(t_{i}) such that

grad(γ(ti))H2(ds)0.{\left\|{\operatorname{grad}\operatorname{\mathcal{E}}(\gamma(t_{i}))}\right\|}_{H^{2}(ds)}\to 0.

Then (45) and (46) imply that

dαH2((γ)+1(γ)+1(γ)3)1/2gradγH2(ds).\|d\operatorname{\mathcal{E}}_{\alpha}\|_{{H^{2}}^{*}}\leq\Bigl{(}\mathcal{L}(\gamma)+\frac{1}{\mathcal{L}(\gamma)}+\frac{1}{\mathcal{L}(\gamma)^{3}}\Bigr{)}^{1/2}\|\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}\|_{H^{2}(ds)}.

Since (γ)<(γ(0))/λ2\mathcal{L}(\gamma)<\operatorname{\mathcal{E}}(\gamma(0))/\lambda^{2}, and applying the Hölder inequality to Fenchel’s theorem 2π|k|𝑑s2\pi\leq\int|k|ds gives

1(γ)14π2k2𝑑s<14π2(γ(0)),\frac{1}{\mathcal{L}(\gamma)}\leq\frac{1}{4\pi^{2}}\int k^{2}ds<\frac{1}{4\pi^{2}}\operatorname{\mathcal{E}}(\gamma(0)),

it follows that dα(ti)H20\|d\operatorname{\mathcal{E}}_{\alpha(t_{i})}\|_{{H^{2}}^{*}}\to 0 too. Hence α(ti)\alpha(t_{i}) which we abbreviate to αi\alpha_{i} satisfies the assumptions of Proposition 5.4 and there exists a subsequence, still denoted αi\alpha_{i}, converging in H2H^{2} to a stationary point α\alpha_{\infty}. Now by Theorem 4.15 there are constants Z>0Z>0, δ(0,1]\delta\in(0,1], and θ[1/2,1)\theta\in[1/2,1) such that for any x2x\in\mathcal{I}^{2} with xαH2<δ\|x-\alpha_{\infty}\|_{H^{2}}<\delta:

(59) gradxH2(ds)Z|(x)(α)|θ.\|\operatorname{grad}\operatorname{\mathcal{E}}_{x}\|_{H^{2}(ds)}\geq Z{\left|{\operatorname{\mathcal{E}}(x)-\operatorname{\mathcal{E}}(\alpha_{\infty})}\right|}^{\theta}.

Since the H2(ds)H^{2}(ds)-Riemannian distance and the standard H2H^{2} metric are equivalent (Section 2.2), there exist δ~>0,r>0\tilde{\delta}>0,r>0 such that Bδ~H2(α)Brdist(α)BδH2(α)B^{H^{2}}_{\tilde{\delta}}(\alpha_{\infty})\subset B^{\operatorname{dist}}_{r}(\alpha_{\infty})\subset B^{H^{2}}_{\delta}(\alpha_{\infty}). For any ii such that αiBδ~H2(α)\alpha_{i}\in B^{H^{2}}_{\tilde{\delta}}(\alpha_{\infty}) we let βi(t)\beta_{i}(t) be the H2(ds)H^{2}(ds)-gradient flow with intial data βi(ti)=αi\beta_{i}(t_{i})=\alpha_{i}. Then due to the uniqueness of the flow, for all t>tit>t_{i}, βi(t)\beta_{i}(t) is a fixed (i.e. time independent) reparametrisation and translation of γ(t)\gamma(t) , namely βi(t)=γ(t)ωγ(ti)11(γ(ti))0(γ(ti))γ(ti)𝑑s\beta_{i}(t)=\gamma(t)\circ\omega_{\gamma(t_{i})}^{-1}-\frac{1}{\mathcal{L}(\gamma(t_{i}))}\int^{\mathcal{L}(\gamma(t_{i}))}_{0}\gamma(t_{i})\,ds, and therefore by (12) we have

(60) gradβi(t)H2(ds)=gradγ(t)H2(ds).\|\operatorname{grad}\operatorname{\mathcal{E}}_{\beta_{i}(t)}\|_{H^{2}(ds)}=\|\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma(t)}\|_{H^{2}(ds)}.

It follows that the trajectories βi(t)\beta_{i}(t) and γ(t)\gamma(t) have the same H2(ds)H^{2}(ds)-length. Let TiT_{i} be the maximum time such that βi(t)αH2<δ~\|\beta_{i}(t)-\alpha_{\infty}\|_{H^{2}}<\tilde{\delta} for all t[ti,Ti)t\in[t_{i},T_{i}). Define

H(t):=((γ(t))(α))1θ.H(t):=(\operatorname{\mathcal{E}}(\gamma(t))-\operatorname{\mathcal{E}}(\alpha_{\infty}))^{1-\theta}.

Then H>0H>0 and is monotonically decreasing because (α)=(γ)\operatorname{\mathcal{E}}(\alpha)=\operatorname{\mathcal{E}}(\gamma). Since (59) holds for βi(t)\beta_{i}(t) with t[ti,Ti)t\in[t_{i},T_{i}), we observe from (βi(t))=(γ(t))\operatorname{\mathcal{E}}(\beta_{i}(t))=\operatorname{\mathcal{E}}(\gamma(t)) and (60) that

H(t)\displaystyle-H^{\prime}(t) =(1θ)((γ(t))(α))θd(γ(t))dt\displaystyle=-(1-\theta)\left(\operatorname{\mathcal{E}}(\gamma(t))-\operatorname{\mathcal{E}}(\alpha_{\infty})\right)^{-\theta}\frac{d\operatorname{\mathcal{E}}(\gamma(t))}{dt}
=(1θ)((γ(t))(α))θgradγH2(ds)2\displaystyle=(1-\theta)\left(\operatorname{\mathcal{E}}(\gamma(t))-\operatorname{\mathcal{E}}(\alpha_{\infty})\right)^{-\theta}\|\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}\|^{2}_{H^{2}(ds)}
(1θ)ZgradγH2(ds).\displaystyle\geq(1-\theta)Z\|\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}\|_{H^{2}(ds)}.

Integrating over [ti,Ti)[t_{i},T_{i}) we get

(1θ)ZtiTigradγH2(ds)𝑑tH(ti)H(Ti).(1-\theta)Z\int_{t_{i}}^{T_{i}}\|\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}\|_{H^{2}(ds)}\,dt\leq H(t_{i})-H(T_{i}).

Now if we fix a jj such that αjαH2<δ~\|\alpha_{j}-\alpha_{\infty}\|_{H^{2}}<\tilde{\delta} and let W:=ij[ti,Ti)W:=\cup_{i\geq j}[t_{i},T_{i}) we have that

(61) WgradγH2(ds)𝑑tH(ti)(1θ)Z.\int_{W}\|\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}\|_{H^{2}(ds)}\,dt\leq\dfrac{H(t_{i})}{(1-\theta)Z}.

In fact, there exists NN\in\mathbb{N} such that βN(t)αH2<δ~\|\beta_{N}(t)-\alpha_{\infty}\|_{H^{2}}<\tilde{\delta} for all t>tNt>t_{N}. If not, then for each ii\in\mathbb{N} there exists TiT_{i} such that βi(Ti)\beta_{i}(T_{i}) is on the boundary of the ball Bδ~H2(α)B^{H^{2}}_{\tilde{\delta}}(\alpha_{\infty}), and there exists a subsequence, still denoted (ti)(t_{i}), such that the intersection ij[ti,Ti)\cap_{i\geq j}[t_{i},T_{i}) is empty. By the choice of δ~\tilde{\delta}, Lemma 2.1 applies and there is a C>0C>0, depending only on α\alpha_{\infty} and rr, such that

δ~=βi(Ti)αH2\displaystyle\tilde{\delta}=\|\beta_{i}(T_{i})-\alpha_{\infty}\|_{H^{2}} βi(ti)αH2+βi(ti)βi(Ti)H2\displaystyle\leq\|\beta_{i}(t_{i})-\alpha_{\infty}\|_{H^{2}}+\|\beta_{i}(t_{i})-\beta_{i}(T_{i})\|_{H^{2}}
(62) α(ti)αH2+Cdist(βi(ti),βi(Ti))\displaystyle\leq\|\alpha(t_{i})-\alpha_{\infty}\|_{H^{2}}+C\operatorname{dist}(\beta_{i}(t_{i}),\beta_{i}(T_{i}))
α(ti)αH2+CtiTiγtH2(ds)𝑑t,\displaystyle\leq\|\alpha(t_{i})-\alpha_{\infty}\|_{H^{2}}+C\int_{t_{i}}^{T_{i}}\|\gamma_{t}\|_{H^{2}(ds)}\,dt,

where we have used (60). But then the integral WgradγH2(ds)𝑑t\int_{W}\|\operatorname{grad}\operatorname{\mathcal{E}}_{\gamma}\|_{H^{2}(ds)}\,dt cannot be finite, contradicting (61). So there exists NN\in\mathbb{N} such that βN(t)Bδ~H2(α)\beta_{N}(t)\in B_{\tilde{\delta}}^{H^{2}}(\alpha_{\infty}) for all t>tNt>t_{N} and therefore

tNγtH2(ds)𝑑t<,\int_{t_{N}}^{\infty}\|\gamma_{t}\|_{H^{2}(ds)}\,dt<\infty,

that is, the H2(ds)H^{2}(ds)-length of γ(t)\gamma(t) is finite. Hence by Lemma 3.3 the flow converges in the H2(ds)H^{2}(ds)-distance, and therefore also in H2H^{2} (Lemma 2.1). ∎

We are now in a position to prove Theorem 1.1.

Proof of Theorem 1.1.

By Proposition 3.2 we see that problem (GF) possesses a unique local-in-time solution γC1([0,T),(S1,n))\gamma\in C^{1}([0,T),\mathcal{I}(S^{1},\mathbb{R}^{n})) for some T>0T>0. Proposition 3.4 extends the local-in-time solution into a global-in-time solution. Finally, we observe from Theorem 5.5 that the global-in-time solution converges to an elastica as tt\to\infty in the H2H^{2}-topology. ∎

Remark 5.6.

As in [16] the classification of closed elastica in 2\mathbb{R}^{2} and 3\mathbb{R}^{3} allows us to determine the shape of the limit in these cases. In 2\mathbb{R}^{2} the only closed elastica are the (geometric) circle, the figure eight elastica and their multiple covers. Since the flow must remain in a path component of 2(S1,2)\mathcal{I}^{2}(S^{1},\mathbb{R}^{2}), for rotation index p>0p>0 the limit is a pp-fold circle, and an initial curve with rotation index 0 will converge to a (possible multiply covered) figure eight. In 3\mathbb{R}^{3} there are more closed elastica, however it was proved in [16] that circles are the only stable closed elastica. In both 2\mathbb{R}^{2} and 3\mathbb{R}^{3} it follows from (4.1) that the limiting circle has radius |λ|1{\left|{\lambda}\right|}^{-1}.

Remark 5.7.

It was mentioned in the introduction that in [8] and [23] the convergence results for the L2(ds)L^{2}(ds)-gradient flow of elastic energy are modulo reparametrisation. They are obtained by proving a Łojasiewicz–Simon gradient inequality for the L2L^{2}-norm of the gradient in a space of graphs over the critical point, which is different to the approach taken here. However, it seems that the key difference is not in the method, but in the fact that the L2(ds)L^{2}(ds)-metric does not dominate L2L^{2} without extra assumptions on parametrisation, and therefore it would not be possible to verify an L2L^{2} version of (62).

Appendix A Auxiliary lemmata

In this appendix, we prove some estimates on the matrix WW defined by (40) that are used in the proof of Lemma 5.3.

Lemma A.1.

Let γ2(S1,n)\gamma\in\mathcal{I}^{2}(S^{1},\mathbb{R}^{n}). Then there exists a constant C>0C>0 such that

W1<C,\|W^{-1}\|<C,

where WW is defined by (40).

Proof.

Let {T(u),ν1(u),,νn1(u)}\{T(u),\nu_{1}(u),\ldots,\nu_{n-1}(u)\} be an orthonormal basis at γ(u)\gamma(u) for uS1u\in S^{1} (e.g. as in (37)). Let μ\mu\in\mathbb{R} be the smallest eigenvalue of WW and η\eta a corresponding eigenvector with |η|=1|\eta|=1. Since

ηTWη=μηTη=μ,\eta^{T}W\eta=\mu\eta^{T}\eta=\mu,

we have

μ=01ηTBBTη𝑑u=01|BTη|2𝑑u=j=1n101β2η,νj2𝑑u,\mu=\int^{1}_{0}\eta^{T}BB^{T}\eta\,du=\int^{1}_{0}|B^{T}\eta|^{2}\,du=\sum^{n-1}_{j=1}\int^{1}_{0}\beta^{2}\left\langle{\eta,\nu_{j}}\right\rangle^{2}\,du,

where we used the definition of BB (40). Recalling that

1=|η|2=η,T2+j=1n1η,νj2,1=|\eta|^{2}=\left\langle{\eta,T}\right\rangle^{2}+\sum^{n-1}_{j=1}\left\langle{\eta,\nu_{j}}\right\rangle^{2},

we see that

(63) μ=01β2(1η,T2)𝑑u.\mu=\int^{1}_{0}\beta^{2}\bigl{(}1-\left\langle{\eta,T}\right\rangle^{2}\bigr{)}\,du.

If μ=0\mu=0, then it follows from (63) that η,T(u)1\left\langle{\eta,T(u)}\right\rangle\equiv 1 and T(u)=±ηT(u)=\pm\eta for all uu. But this is impossible because γ\gamma is closed and TT is continuous. Hence

(64) W1:=supvn|W1v||v|=supxn|x||Wx|=1μ<.\|W^{-1}\|:=\sup_{v\in\mathbb{R}^{n}}\frac{|W^{-1}v|}{{\left|{v}\right|}}=\sup_{x\in\mathbb{R}^{n}}\frac{{\left|{x}\right|}}{{\left|{Wx}\right|}}=\dfrac{1}{\mu}<\infty.

Therefore Lemma A.1 follows. ∎

Lemma A.2.

Let UΩU\subset\Omega such that αH2<{\left\|{\alpha}\right\|}_{H^{2}}<\infty for all αU\alpha\in U. Then there exists a constant C>0C>0 such that

Wα1<C\|W_{\alpha}^{-1}\|<C

for all αU\alpha\in U, where WαW_{\alpha} denotes the matrix defined by (40) with γ=α\gamma=\alpha.

Proof.

Suppose there is no such CC, and therefore there exists a sequence (αi)U(\alpha_{i})\subset U such that Wαi1{\left\|{W_{\alpha_{i}}^{-1}}\right\|}\to\infty. Equivalently, abbreviating to WiW_{i} and letting μi\mu_{i} be the smallest eigenvalue of WiW_{i}, by (64) μi0\mu_{i}\to 0. We will show that this leads to a contradiction.

Since the sequence (αi)(\alpha_{i}) is bounded in H2(S1,n)H^{2}(S^{1},\mathbb{R}^{n}), there exists αH2(S1,n)C1+δ(S1,n)\alpha_{\infty}\in H^{2}(S^{1},\mathbb{R}^{n})\cap C^{1+\delta}(S^{1},\mathbb{R}^{n}) such that

(65) αiαweakly inH2(S1,n),\displaystyle\alpha_{i}\rightharpoonup\alpha_{\infty}\quad\text{weakly in}\quad H^{2}(S^{1},\mathbb{R}^{n}),
αiαinC1+δ(S1,n),\displaystyle\alpha_{i}\to\alpha_{\infty}\quad\text{in}\quad C^{1+\delta}(S^{1},\mathbb{R}^{n}),

up to a subsequence, where δ(0,1/2)\delta\in(0,1/2). This implies that

Ti:=αi|αi|T:=α|α|inCδ(S1,n).T_{i}:=\dfrac{\alpha_{i}^{\prime}}{|\alpha_{i}^{\prime}|}\to T_{\infty}:=\dfrac{\alpha_{\infty}^{\prime}}{|\alpha_{\infty}^{\prime}|}\quad\text{in}\quad C^{\delta}(S^{1},\mathbb{R}^{n}).

As in (37) we construct an orthonormal frame {Ti,νi1,,νin1}\{T_{i},\nu_{i}^{1},\ldots,\nu_{i}^{n-1}\} along αi\alpha_{i} by starting with an orthonormal basis {Ti(0),νi,01,,νi,0n1}\{T_{i}(0),\nu_{i,0}^{1},\ldots,\nu_{i,0}^{n-1}\} and letting {νi1,,νin1}\{\nu_{i}^{1},\ldots,\nu_{i}^{n-1}\} be the solutions of

(66) (νij)=1|αi|2(νij),αi′′αi,νij(0)=νi,0j(\nu_{i}^{j})^{\prime}=-\dfrac{1}{|\alpha_{i}^{\prime}|^{2}}\langle(\nu_{i}^{j}),\alpha_{i}^{\prime\prime}\rangle\alpha_{i}^{\prime},\quad\nu_{i}^{j}(0)=\nu_{i,0}^{j}

for j{1,,n1}j\in\{1,\ldots,n-1\}.

Since |νij|1|\nu_{i}^{j}|\equiv 1, choosing NN sufficiently large so that for i>Ni>N, αi\alpha_{i} is contained in a C1C^{1}-ball centred at α\alpha_{\infty} and therefore |αi(u)||\alpha_{i}^{\prime}(u)| is bounded below, we have that

(νij)L2Cαi′′L2(S1)<C\|(\nu_{i}^{j})^{\prime}\|_{L^{2}}\leq C\|\alpha_{i}^{\prime\prime}\|_{L^{2}(S^{1})}<C

for all i>Ni>N and j{1,,n1}j\in\{1,\ldots,n-1\}. Thus (νij)(\nu^{j}_{i}) is bounded in H1(S1,n)H^{1}(S^{1},\mathbb{R}^{n}), and then we find {ν1,,νn1}\{\nu_{\infty}^{1},\ldots,\nu_{\infty}^{n-1}\} such that

(67) νijνjweakly inH1(S1,n),\displaystyle\nu_{i}^{j}\rightharpoonup\nu_{\infty}^{j}\quad\text{weakly in}\quad H^{1}(S^{1},\mathbb{R}^{n}),
νijνjinCδ(S1),\displaystyle\nu_{i}^{j}\to\nu_{\infty}^{j}\quad\text{in}\quad C^{\delta}(S^{1}),

up to a subsequence. We note here that |νj|1|\nu_{\infty}^{j}|\equiv 1. By (66) we have

(68) νij(u)=νij(0)0u1|αi(u~)|2νij(u~),αi′′(u~)αi(u~)𝑑u~.\nu_{i}^{j}(u)=\nu_{i}^{j}(0)-\int^{u}_{0}\dfrac{1}{|\alpha_{i}^{\prime}(\tilde{u})|^{2}}\langle\nu_{i}^{j}(\tilde{u}),\alpha_{i}^{\prime\prime}(\tilde{u})\rangle\alpha_{i}^{\prime}(\tilde{u})\,d\tilde{u}.

Letting ii\to\infty in (68), we observe from (65) and (67) that

νj(u)=νj(0)0u1|α(u~)|2νj(u~),α′′(u~)α(u~)𝑑u~\nu_{\infty}^{j}(u)=\nu_{\infty}^{j}(0)-\int^{u}_{0}\dfrac{1}{|\alpha_{\infty}^{\prime}(\tilde{u})|^{2}}\langle\nu_{\infty}^{j}(\tilde{u}),\alpha_{\infty}^{\prime\prime}(\tilde{u})\rangle\alpha_{\infty}^{\prime}(\tilde{u})\,d\tilde{u}

for j{1,,n1}j\in\{1,\ldots,n-1\}, where {T(0),ν1(0),,νn1(0)}\{T_{\infty}(0),\nu_{\infty}^{1}(0),\ldots,\nu_{\infty}^{n-1}(0)\} is an orthonormal basis at α(0)\alpha_{\infty}(0) because of the strong convergence in (65) and (67). Then we can check as in (37) that {T(u),ν1(u),,νn1(u)}\{T_{\infty}(u),\nu_{\infty}^{1}(u),\ldots,\nu_{\infty}^{n-1}(u)\} is an orthonormal frame at α(u)\alpha_{\infty}(u) for all uS1u\in S^{1}.

Now we have

Bi:=β[νi1νi2νin1],Wi:=01BiBiT𝑑u,B_{i}:=\beta\bigl{[}\nu_{i}^{1}\,\nu_{i}^{2}\ldots\nu_{i}^{n-1}\bigr{]},\qquad W_{i}:=\int^{1}_{0}B_{i}B_{i}^{T}\,du,

and let ηi\eta_{i} be a unit eigenvector corresponding to the smallest eigenvalue μi\mu_{i} of WiW_{i}. By (67) we see that

WiW:=01BBT𝑑uW_{i}\to W_{\infty}:=\int^{1}_{0}B_{\infty}B_{\infty}^{T}\,du

with

B:=β[ν1ν2νn1].B_{\infty}:=\beta\bigl{[}\nu_{\infty}^{1}\nu_{\infty}^{2}\ldots\nu_{\infty}^{n-1}\bigr{]}.

Thus we find μ\mu_{\infty}\in\mathbb{R} such that

μiμasi,\mu_{i}\to\mu_{\infty}\quad\text{as}\quad i\to\infty,

and μ\mu_{\infty} is the smallest eigenvalue of WW_{\infty}. Let η\eta_{\infty} be the unit eigenvector corresponding to μ\mu_{\infty}, then as in (63), we have

(69) μ=01β2(1η,T2)𝑑u.\mu_{\infty}=\int^{1}_{0}\beta^{2}\Bigl{(}1-\left\langle{\eta_{\infty},T_{\infty}}\right\rangle^{2}\Bigr{)}\,du.

Repeating the argument following (63), μ\mu_{\infty} cannot be zero, and we have our contradiction to the assumption that there is no CC bounding Wα1{\left\|{W_{\alpha}^{-1}}\right\|}.

References

  • [1] Adams, R. A., and Fournier, J. J. F. Sobolev spaces, second ed., vol. 140 of Pure and Applied Mathematics (Amsterdam). Elsevier/Academic Press, Amsterdam, 2003.
  • [2] Blatt, S., Hopper, C. P., and Vorderobermeier, N. A minimising movement scheme for the p-elastic energy of curves. Journal of Evolution Equations 22, 2 (2022), 1–25.
  • [3] Brockett, R. Finite dimensional linear systems. Series in decision and control. Wiley, 1970.
  • [4] Bruveris, M. Completeness properties of Sobolev metrics on the space of curves. J. Geom. Mech. 7, 2 (2015), 125–150.
  • [5] Chill, R., Fašangová, E., and Schätzle, R. Willmore blowups are never compact. Duke Mathematical Journal 147, 2 (2009), 345 – 376.
  • [6] Dall’Acqua, A., Lin, C.-C., and Pozzi, P. A gradient flow for open elastic curves with fixed length and clamped ends. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 17, 3 (2017), 1031–1066.
  • [7] Dall’Acqua, A., Lin, C.-C., and Pozzi, P. Elastic flow of networks: short-time existence result. Journal of Evolution Equations (2020), 1–46.
  • [8] Dall’Acqua, A., Pozzi, P., and Spener, A. The Łojasiewicz-Simon gradient inequality for open elastic curves. J. Differential Equations 261, 3 (2016), 2168–2209.
  • [9] Dziuk, G., Kuwert, E., and Schätzle, R. Evolution of elastic curves in n\mathbb{R}^{n}: existence and computation. SIAM J. Math. Anal. 33, 5 (2002), 1228–1245.
  • [10] Feehan, P. M. N. Global existence and convergence of solutions to gradient systems and applications to Yang-Mills gradient flow, 2016.
  • [11] Feehan, P. M. N., and Maridakis, M. Łojasiewicz-Simon gradient inequalities for analytic and Morse-Bott functions on Banach spaces. J. Reine Angew. Math. 765 (2020), 35–67.
  • [12] Klingenberg, W. P. A. Riemannian geometry, second ed., vol. 1 of De Gruyter Studies in Mathematics. Walter de Gruyter & Co., Berlin, 1995.
  • [13] Knappmann, J., Schumacher, H., Steenebrügge, D., and von der Mosel, H. A speed preserving Hilbert gradient flow for generalized integral menger curvature, 2021.
  • [14] Koiso, N. On the motion of a curve towards elastica. In Actes de la Table Ronde de Géométrie Différentielle (Luminy, 1992), vol. 1 of Sémin. Congr. Soc. Math. France, Paris, 1996, pp. 403–436.
  • [15] Lang, S. Fundamentals of differential geometry, vol. 191 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1999.
  • [16] Langer, J., and Singer, D. A. Curve straightening and a minimax argument for closed elastic curves. Topology 24, 1 (1985), 75–88.
  • [17] Langer, J., and Singer, D. A. Curve-straightening in Riemannian manifolds. Ann. Global Anal. Geom. 5, 2 (1987), 133–150.
  • [18] Lin, C.-C. L2L^{2}-flow of elastic curves with clamped boundary conditions. J. Differential Equations 252, 12 (2012), 6414–6428.
  • [19] Linnér, A. Some properties of the curve straightening flow in the plane. Trans. Amer. Math. Soc. 314, 2 (1989), 605–618.
  • [20] Linnér, A. Curve-straightening in closed Euclidean submanifolds. Comm. Math. Phys. 138, 1 (1991), 33–49.
  • [21] Linnér, A. Symmetrized curve-straightening. Differential Geom. Appl. 18, 2 (2003), 119–146.
  • [22] Mantegazza, C., Pluda, A., and Pozzetta, M. A survey of the elastic flow of curves and networks. Milan J. Math. 89, 1 (2021), 59–121.
  • [23] Mantegazza, C., and Pozzetta, M. The Łojasiewicz-Simon inequality for the elastic flow. Calc. Var. Partial Differential Equations 60, 1 (2021), Paper No. 56, 17.
  • [24] Michor, P. W., and Mumford, D. Riemannian geometries on spaces of plane curves. J. Eur. Math. Soc. (JEMS) 8, 1 (2006), 1–48.
  • [25] Müller, M. On gradient flows with obstacles and Euler’s elastica. Nonlinear Anal. 192 (2020), 111676, 48.
  • [26] Neuberger, J. W. Sobolev gradients and differential equations, second ed., vol. 1670 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 2010.
  • [27] Novaga, M., and Okabe, S. Curve shortening-straightening flow for non-closed planar curves with infinite length. J. Differential Equations 256, 3 (2014), 1093–1132.
  • [28] Novaga, M., and Pozzi, P. A second order gradient flow of pp-elastic planar networks. SIAM J. Math. Anal. 52, 1 (2020), 682–708.
  • [29] Okabe, S. The motion of elastic planar closed curves under the area-preserving condition. Indiana Univ. Math. J. 56, 4 (2007), 1871–1912.
  • [30] Okabe, S. The dynamics of elastic closed curves under uniform high pressure. Calc. Var. Partial Differential Equations 33, 4 (2008), 493–521.
  • [31] Okabe, S., Pozzi, P., and Wheeler, G. A gradient flow for the pp-elastic energy defined on closed planar curves. Math. Ann. 378, 1-2 (2020), 777–828.
  • [32] Okabe, S., and Wheeler, G. The pp-elastic flow for planar closed curves with constant parametrization. arXiv:2104.03570 (2021).
  • [33] Palais, R. S., and Terng, C.-L. Critical point theory and submanifold geometry, vol. 1353 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1988.
  • [34] Polden, A. Curves and surfaces of least total curvature and fourth-order flows. PhD Thesis, Universität Tübingen (1996).
  • [35] Reiter, P., and Schumacher, H. Sobolev gradients for the Möbius energy. Archive for Rational Mechanics and Analysis 242, 2 (2021), 701–746.
  • [36] Rupp, F. On the Łojasiewicz–Simon gradient inequality on submanifolds. Journal of Functional Analysis 279, 8 (2020), 108708.
  • [37] Rupp, F., and Spener, A. Existence and convergence of the length-preserving elastic flow of clamped curves. arXiv:2009.06991 (2020).
  • [38] Schrader, P. Global analysis of one-dimensional variational problems. PhD thesis, The University of Western Australia, 2016.
  • [39] Schrader, P. Morse theory for elastica. J. Geom. Mech. 8, 2 (2016), 235–256.
  • [40] Schrader, P., Wheeler, G., and Wheeler, V.-M. On the H1(ds){H}^{1}(ds)-gradient flow for the length functional. arXiv:2102.07305 (2021).
  • [41] Spener, A. Short time existence for the elastic flow of clamped curves. Math. Nachr. 290, 13 (2017), 2052–2077.
  • [42] Wen, Y. L2L^{2} flow of curve straightening in the plane. Duke Math. J. 70, 3 (1993), 683–698.
  • [43] Wen, Y. Curve straightening flow deforms closed plane curves with nonzero rotation number to circles. J. Differential Equations 120, 1 (1995), 89–107.
  • [44] Whittlesey, E. F. Analytic functions in Banach spaces. Proc. Amer. Math. Soc. 16 (1965), 1077–1083.
  • [45] Zeidler, E. Nonlinear functional analysis and its applications. I. Springer-Verlag, New York, 1986. Fixed-point theorems, Translated from the German by Peter R. Wadsack.