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

Sobolev metrics on spaces of manifold valued curves

Martin Bauer Martin Bauer: Department of Mathematics, Florida State University bauer@math.fsu.edu Cy Maor Cy Maor: Einstein Institute of Mathematics, The Hebrew University of Jerusalem cy.maor@mail.huji.ac.il  and  Peter W. Michor Peter W. Michor: Faculty for Mathematics, University of Vienna peter.michor@univie.ac.at
Abstract.

We study completeness properties of reparametrization invariant Sobolev metrics of order n2n\geq 2 on the space of manifold valued open and closed immersed curves. In particular, for several important classes of metrics, we show that Sobolev immersions are metrically and geodesically complete (thus the geodesic equation is globally well-posed). These results were previously known only for closed curves with values in Euclidean space. For the class of constant coefficient Sobolev metrics on open curves, we show that they are metrically incomplete, and that this incompleteness only arises from curves that vanish completely (unlike “local” failures that occur in lower order metrics).

2010 Mathematics Subject Classification:
58B20, 58D10, 35G55, 35G60
M. Bauer was partially supported by NSF-grant 1912037 (collaborative research in connection with NSF-grant 1912030). C. Maor was partially supported by ISF-grant 1269/19.

1. Introduction and main results

1.1. Background

In recent years Riemannian geometry on the space of curves has been an area of active research. The motivation for these investigations can be found in the area of shape analysis, where the space of geometric curves plays an important role: closed planar curves are used to encode the outlines (shapes) of planar objects, and elastic (reparametrization invariant) Riemannian metrics have been successfully used to compare these objects in a variety of different applications [32, 33, 38, 39]. More recently, curves with values in a manifold have emerged as a topic of interest in shape analysis as well. Examples include the study of trajectories on the earth [34, 35], of computer animations [18], or of brain connectivity data [19]. Here the brain connectivity of a patient over time is represented as a path in the space of positive, symmetric matrices. Motivated by these applications several of the numerical algorithms, as originally developed for planar curves, have been generalized to this more complicated situation.

In this article we are interested in the mathematical properties of these Riemannian metrics and in particular in questions related to completeness of the corresponding geodesic equations. These investigations build up on classical questions related to diffeomorphism groups, as reparametrization invariant metrics on spaces of immersions can be viewed as generalizations of right-invariant metrics on diffeomorphism groups. These have been in the focus of intense research due to their relations to many prominent PDEs via Arnold’s approach to hydrodynamics [1, 2, 36]. Local well-posedness in this setup was established for a wide variety of invariant metrics, typically using an Ebin–Marsden-type analysis [20, 29, 23, 28, 7]. The focus of this article is geodesic and metric completeness, which is well understood for strong enough metrics in the case of diffeomorphism groups [39, 30, 29, 16, 5], but is mostly open for spaces of immersions. For closed, regular curves with values in Euclidean space, a series of completeness results both on the space of parametrized and unparametrized curves has been obtained, beginning with Bruveris, Michor and Mumford [14], see also [12, 15, 8]. The goal of this article is to generalize these results to the case of open and closed, regular curves with values in a Riemannian manifold. While the manifold structure of the target space is of little relevance for the local results mentioned before, it significantly complicates the analysis for the global results studied in the present article. We will comment on the differences with the Euclidean situation in Section 1.4 below; first we describe the main contributions of the present article.

1.2. Main Result

To formulate our main result we first introduce the manifold of regular curves and the class of Riemannian metrics that we will consider in this article. For n2n\geq 2, we consider the space of Sobolev immersions from a one-dimensional parameter space DD with values in a complete Riemannian manifold with bounded geometry (𝒩,g)(\mathcal{N},g):

(1.1) n(D,𝒩)={cHn(D,𝒩):c(θ)0,θD}.\mathcal{I}^{n}(D,\mathcal{N})=\left\{c\in H^{n}(D,\mathcal{N}):c^{\prime}(\theta)\neq 0,\;\forall\theta\in D\right\}.

Here D=[0,2π]D=[0,2\pi] for open curves and D=S1D=S^{1} for closed curves. The Sobolev space Hn(D,𝒩)H^{n}(D,\mathcal{N}) is defined in more detail in Section 2; note that Hn(D,𝒩)C1(D,𝒩)H^{n}(D,\mathcal{N})\subset C^{1}(D,\mathcal{N}), hence the condition c(θ)0c^{\prime}(\theta)\neq 0 is well defined. On this space we can consider reparametrization invariant (elastic) Sobolev metrics of order nn. The class we focus on in this paper is given by

(1.2) Gc(h,k)=i=0nai(c)Dg(sih,sik)ds,\begin{split}&G_{c}(h,k)=\sum_{i=0}^{n}a_{i}(\ell_{c})\int_{D}g(\nabla_{\partial_{s}}^{i}h,\nabla_{\partial_{s}}^{i}k)\,\mathrm{d}s,\end{split}

where aiC((0,),[0,))a_{i}\in C^{\infty}((0,\infty),[0,\infty)), \nabla is the covariant derivative in 𝒩\mathcal{N}, and s=|c|s=|c^{\prime}| is the norm of cc^{\prime} with respect to the Riemannian metric gg. Furthermore, ds=|c|dθ\,\mathrm{d}s=|c^{\prime}|\,\mathrm{d}\theta is the arc length one form, s=1|c|θ\partial_{s}=\frac{1}{|c^{\prime}|}\partial_{\theta} is the arc length vector field along the curve, and c=Dds\ell_{c}=\int_{D}\,\mathrm{d}s is the length of the curve. The two most important sub-families of these type are:

  1. (1)

    The constant coefficient Sobolev metrics, where ai(c)=Ci0a_{i}(\ell_{c})=C_{i}\geq 0 are constants and do not depend on the length c\ell_{c};

  2. (2)

    The family of scale invariant Sobolev metrics where ai(c)=Cic2n3a_{i}(\ell_{c})=C_{i}\ell_{c}^{2n-3} with Ci0C_{i}\geq 0 being again constants. In this case, when the target manifold 𝒩\mathcal{N} is the Euclidean space, composition with rescaling xαxx\mapsto\alpha x of the target manifold is an isometry of (n(D,𝒩),G)(\mathcal{I}^{n}(D,\mathcal{N}),G), for each α>0\alpha>0.

In both cases we assume that C0C_{0} and CnC_{n} are strictly positive, to avoid degeneracy. The main focus of the present article lies on completeness properties of these Riemannian metrics. In a slightly simplified version our main results can be summarized as follows:

Theorem (Main Theorem).

Let D=[0,2π]D=[0,2\pi] or D=S1D=S^{1}, and let GG be the scale invariant Sobolev metric (1.2) of order n2n\geq 2. The following completeness properties hold:

  1. (1)

    (n(D,𝒩),distG)(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}) is a complete metric space.

  2. (2)

    (n(D,𝒩),G)(\mathcal{I}^{n}(D,\mathcal{N}),G) is geodesically complete

  3. (3)

    Any two immersions in the same connected component of (n(D,𝒩),G)(\mathcal{I}^{n}(D,\mathcal{N}),G) can be joined by a minimizing geodesic.

For D=S1D=S^{1} the results continue to hold for the family of constant coefficient Sobolev metrics.

Previously this result was only known for closed curves in Euclidean space (see [12] for constant coefficients and [16] for a wider class that includes scale invariant ones), and thus the results of the present article generalize these previous works in two important directions (open curves and curves with values in a manifold). In fact, we will prove these statements for a wider class of metrics, see Theorems 5.15.3. Note that in this infinite dimensional situation the theorem of Hopf-Rinow is not valid [3] and thus item (3) does not follow directly from the metric completeness, but has to be proven separately.

1.3. Further contributions of the article

In the following we describe several further key contributions of the current article:

  • Completeness in the smooth setting: In the main theorem above, we have formulated the results only in the Sobolev category. Using an Ebin–Marsden-type no-loss-no-gain result [20], we show that geodesic completeness (i.e., global existence of geodesics) extends to the space of smooth, closed curves (Corollary 5.13). For open curves, we only obtain regularity in the interior of the curve, as explained in Section 5.6.

  • Metric incompleteness of constant coefficient metrics on open curves: In [4] it was observed that the space of open curves, with respect to constant coefficient Sobolev metrics, is metrically incomplete; indeed, in the same paper the authors constructed a path of immersed curves, whose lengths tend to zero after finite time. In Section 6 we elaborate on this example, and show that vanishing of the entire curve is the only way a path (or a sequence) of immersed curves can leave the space of immersions n([0,2π],𝒩)\mathcal{I}^{n}([0,2\pi],\mathcal{N}) in finite time (Theorem 6.3). That is, a path cannot leave the space by some ”local” failure, say, by losing the immersion property at a point (such a failure of completeness can occur in lower-order metrics, e.g., in shockwaves in the inviscid Burgers equation). We give some evidence that the completion of the space of open curves in this case is a one-point completion, where the additional point represents all the Cauchy sequences converging to vanishing length curves.

  • Existence of minimizing geodesics for constant coefficient metrics on open curves: We show that if the distance between two open curves is lower than some explicit threshold depending only on their lengths, then they can be connected by a minimizing geodesic (Theorem 6.7). We note, however, that this threshold is not necessarily sharp; in fact, in view of the rather rigid way in which curves can leave the space, we cannot rule out that a minimizing geodesic exists between any two immersions. We also do not know whether geodesics (unlike general paths of finite length) may cease to exist after finite time, that is, we do not know if the space is geodesically incomplete (only that it is metrically incomplete). These questions will be considered in future works.

  • Local well-posedness: Our completeness results are only valid for metrics of order n2n\geq 2, and it can be shown that metrics of lower order can never have these properties. Nevertheless, using an Ebin–Marsden-type approach, we show local well-posedness for all smooth metrics of the type (1.2) of order n1n\geq 1, see Theorem 3.8. This result was previously known for closed curves and the case of open curves requires some additional considerations for dealing with the boundary terms that appear in the geodesic equation.

  • Completeness of the intrinsic metric on Hn(D,𝒩)H^{n}(D,\mathcal{N}): It is well known that Hn(D,𝒩)H^{n}(D,\mathcal{N}), for n>dimD/2n>\dim D/2, is a Hilbert manifold, and that its topology coincides with the one induced via the inclusion Hn(D,𝒩)Hn(D,m)H^{n}(D,\mathcal{N})\subset H^{n}(D,\mathbb{R}^{m}) that is defined by a closed isometric embedding ι:𝒩m\iota:\mathcal{N}\to\mathbb{R}^{m}. This inclusion also induces a complete metric space structure on Hn(D,𝒩)H^{n}(D,\mathcal{N}). As part of the proof of the main theorem, we show that the natural Riemannian metric on Hn(D,𝒩)H^{n}(D,\mathcal{N}),

    (1.3) c(h,k):=Dgc(h,k)+gc(θnh,θnk)dθ,\mathcal{H}_{c}(h,k):=\int_{D}g_{c}(h,k)+g_{c}(\nabla_{\partial_{\theta}}^{n}h,\nabla_{\partial_{\theta}}^{n}k)\,\mathrm{d}\theta,

    is also metrically complete (Proposition 2.2), thus defining a complete metric space structure that is intrinsic (independent of an isometric embedding). We study these different definitions and their equivalence in Section 2.

1.4. Main ideas in the proof and structure of the article

The techniques used in the proof of our main theorem, Theorems 5.15.3, expand upon the ones used to study completeness of Euclidean curves [12]. The main difficulties arise from taking into account the more complicated structure of the space n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) and the effects of the curvature of 𝒩\mathcal{N} on various estimates (in particular, on the behavior of some Sobolev interpolation inequalities). To give the reader a first glimpse, we will outline the strategy and main steps below.

Local well-posedness

As a basis to the rest of the analysis, we first study the metric GG (as in (1.2)) in Section 3, and prove that it is a smooth, strong metric on n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}).111A Riemannian metric GG on a manifold \mathcal{M} is a section of non-degenerate bilinear forms on the tangent bundle. A strong Riemmanian metric also satisfies that for each xx\in\mathcal{M}, the topology induced by GxG_{x} on TxT_{x}\mathcal{M} coincides with the original topology (induced by the manifold structure) on TxT_{x}\mathcal{M}. If dim<\dim\mathcal{M}<\infty, every metric is a strong one, but in infinite dimensions this is not the case. In this section we also give some details on the associated geodesic equation and formulate the local well-posedness result (as this theorem is not the focus of the present article, we postpone its proof to Appendix A.2).

Metric and geodesic completeness

The space of Sobolev immersions n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) is an open subset of Hn(D,𝒩)H^{n}(D,\mathcal{N}), which is metrically complete with respect to the metric \mathcal{H}, defined in (1.3); this is established in Section 2. Note that for 𝒩=d\mathcal{N}=\mathbb{R}^{d} this is trivial, as Hn(D,𝒩)H^{n}(D,\mathcal{N}) is a Hilbert space in this case.

Since (Hn(D,𝒩),dist)(H^{n}(D,\mathcal{N}),\operatorname{dist}^{\mathcal{H}}) is a complete metric space, showing metric completeness of (n(D,𝒩),distG)(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}) can be reduced to showing that GG and \mathcal{H} are equivalent metrics, uniformly on every distG\operatorname{dist}^{G}-ball in n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}), and that the speed |c||c^{\prime}| of an immersion cn(D,𝒩)c\in\mathcal{I}^{n}(D,\mathcal{N}) is bounded away from zero on every distG\operatorname{dist}^{G}-ball. This reduction is done in detail in Section 5.1.

In order to obtain the uniform equivalence of GG and \mathcal{H} on metric balls, one needs to obtain bounds on the length c\ell_{c} of the curve, and on certain norms of the velocity cc^{\prime}, uniformly for all immersions cc in a metric ball. This is done in Section 5.2, and the proof of metric completeness is then concluded in Sections 5.35.4. As metric completeness implies geodesic completeness of strong Riemannian metrics also in infinite dimensions, see [24, VIII, Proposition 6.5], this also concludes the proof of geodesic completeness.

The main technical tool for establishing the bounds on c\ell_{c} and cc^{\prime} are Sobolev interpolation inequalities on the tangent space Tcn(D,𝒩)T_{c}\mathcal{I}^{n}(D,\mathcal{N}), with explicit dependence of the inequalities constants on the length of the base curve cc. In the case of closed curves, there is non-trivial holonomy along the curves, hence we need to control the holonomy along a curve in terms of its length, and apply these estimates to the interpolation inequalities (this is one of the main technical differences from the Euclidean case). These are done in Section 4, though some of the geometric estimates are postponed to Appendix B.

Existence of minimal geodesics

To prove existence of minimal geodesics between two immersions c0c_{0} and c1c_{1}, we consider the energy of paths ct:[0,1]n(D,𝒩)c_{t}:[0,1]\to\mathcal{I}^{n}(D,\mathcal{N}) between c0c_{0} and c1c_{1} (defined by the metric GG), and use the direct methods of the calculus of variations to prove that a minimizing sequence of paths converges, in an appropriate sense, to an energy minimizer (which is, by definition, a geodesic). This is done in Section 5.5. Since this approach relies heavily on weak convergence of paths, and the weak topology is not readily available on the Hilbert manifold n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}), we first embed it into the Hilbert space Hn(D,m)H^{n}(D,\mathbb{R}^{m}) via a closed isometric embedding ι:𝒩m\iota:\mathcal{N}\to\mathbb{R}^{m}. The analysis then combines the same type of bounds that are used to prove metric completeness, with bounds that relate the metric on Hn(D,m)H^{n}(D,\mathbb{R}^{m}) to the metric GG on n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) (similar bounds are also used in proving the completeness of Hn(D,𝒩)H^{n}(D,\mathcal{N}) with respect to the dist\operatorname{dist}^{\mathcal{H}} metric in Section 2).

Acknowledgements

We would like to thank to Martins Bruveris, FX Vialard and Amitai Yuval for various discussions during the work on this paper.

2. Spaces of manifold valued functions and immersions

Let (𝒩,g)(\mathcal{N},g) be a (possibly non-compact) complete Riemannian manifold with bounded geometry, where the induced norm of the Riemannian metric will be denoted by ||=g(,)|\cdot|=\sqrt{g(\cdot,\cdot)}. We will denote its covariant derivative by \nabla, or, where ambiguity might arise, by 𝒩\nabla^{\mathcal{N}}. With a slight abuse of notation, we will also use it as the covariant derivative on pullbacks of T𝒩T\mathcal{N}.

We consider the space of (closed or open) regular curves with values in 𝒩\mathcal{N}, which we denote by

(2.1) Imm(D,𝒩)={cC(D,𝒩):c(θ)0,θD}.\operatorname{Imm}(D,\mathcal{N})=\left\{c\in C^{\infty}(D,\mathcal{N}):c^{\prime}(\theta)\neq 0,\;\forall\theta\in D\right\}.

Here D=S1D=S^{1} for closed curves and D=[0,2π]D=[0,2\pi] for open curves. This space is an infinite dimensional manifold, whose tangent space at a curve cc is the space of vector fields along cc:

(2.2) TcImm(D,𝒩)={hC(D,T𝒩):π(h)=c},T_{c}\operatorname{Imm}(D,\mathcal{N})=\left\{h\in C^{\infty}(D,T\mathcal{N}):\pi(h)=c\right\},

where π\pi denotes the foot point projection from T𝒩T\mathcal{N} to 𝒩\mathcal{N}.

To obtain the desired completeness and well-posedness results we need to consider a larger space of Sobolev immersions n(D,𝒩)Imm(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N})\supset\operatorname{Imm}(D,\mathcal{N}), for n2n\geq 2, which we define below.

Definition 2.1.

Let 𝒩\mathcal{N} be a Riemannian manifold as above, and fix a proper, smooth, isometric embedding ι:𝒩m\iota:\mathcal{N}\to\mathbb{R}^{m}, for large enough mm\in\mathbb{N}. For n2n\geq 2, we define the Sobolev space Hn(D,𝒩)H^{n}(D,\mathcal{N}) and the space of Sobolev immersions n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) as follows:

  1. (1)

    Hn(D,𝒩)H^{n}(D,\mathcal{N}) consists of all maps c:D𝒩c:D\to\mathcal{N} such that ιcHn(D;m)\iota\circ c\in H^{n}(D;\mathbb{R}^{m}).

  2. (2)

    n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) consists of all cHn(D,𝒩)c\in H^{n}(D,\mathcal{N}) such that c(θ)0,θDc^{\prime}(\theta)\neq 0,\,\forall\theta\in D.

With this (extrinsic) definition of Hn(D,𝒩)H^{n}(D,\mathcal{N}), it inherits the metric structure of Hn(D;m)H^{n}(D;\mathbb{R}^{m}), which we denote by distext\operatorname{dist}^{\text{ext}}; since convergence in the space Hn(D;m)H^{n}(D;\mathbb{R}^{m}) implies uniform convergence, we have that Hn(D,𝒩)H^{n}(D,\mathcal{N}) is a closed subset of Hn(D;m)H^{n}(D;\mathbb{R}^{m}), hence a complete metric space with respect to distext\operatorname{dist}^{\text{ext}}. We are interested in characterizing Hn(D,𝒩)H^{n}(D,\mathcal{N}) as an infinite dimensional Riemannian manifold. The main goal of this section is to prove the following:

Proposition 2.2.

The space Hn(D,𝒩)H^{n}(D,\mathcal{N}), 2n2\leq n\in\mathbb{N} is a Hilbert manifold whose tangent space at cc is Hn(D;cT𝒩)H^{n}(D;c^{*}T\mathcal{N}). Moreover, it is a complete metric space with respect to the distance function dist\operatorname{dist}^{\mathcal{H}} induced by the smooth Riemannian metric (1.3):

c(h,k):=Dgc(h,k)+gc(θnh,θnk)dθ.\mathcal{H}_{c}(h,k):=\int_{D}g_{c}(h,k)+g_{c}(\nabla_{\partial_{\theta}}^{n}h,\nabla_{\partial_{\theta}}^{n}k)\,\mathrm{d}\theta.

Finally, the space of Sobolev immersions n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) is an open subset of Hn(D,𝒩)H^{n}(D,\mathcal{N}) and in particular is a Hilbert manifold with the same tangent space.

Henceforth, we will always endow Hn(D,𝒩)H^{n}(D,\mathcal{N}) with the metric dist\operatorname{dist}^{\mathcal{H}} (rather than distext\operatorname{dist}^{\text{ext}}). Note that, in general, dist\operatorname{dist}^{\mathcal{H}} and distext\operatorname{dist}^{\text{ext}} need not to be equivalent metrics. Note also that for hTcn(D,𝒩)h\in T_{c}\mathcal{I}^{n}(D,\mathcal{N}) there are two natural L2L^{2} metrics: in one we integrate with respect to dθ\,\mathrm{d}\theta, and in the other with respect to arc length ds=|c|dθ\,\mathrm{d}s=|c^{\prime}|\,\mathrm{d}\theta; we denote the first one by L2(dθ)L^{2}(\,\mathrm{d}\theta) and the second by L2(ds)L^{2}(\,\mathrm{d}s).

Proposition 2.2 holds for much more general manifold domain DD: namely, the Hilbert manifold structure exists whenever 2n>dimD2n>\dim D, and the openness of n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) in Hn(D,𝒩)H^{n}(D,\mathcal{N}) holds whenever 2(n1)>dimD2(n-1)>\dim D. These are known results and we describe their proofs below for completeness. To the best of our knowledge, the completeness of (Hn(D,𝒩),dist)(H^{n}(D,\mathcal{N}),\operatorname{dist}^{\mathcal{H}}) has not been considered before; we expect it to hold, again, whenever 2n>dimD2n>\dim D, virtually with the same proof as the one below (using Hölder inequalities instead of uniform bounds).

We start by proving a technical lemma that shows the local equivalence of the c\mathcal{H}_{c} norm and the restriction of the standard Hn(D;m)H^{n}(D;\mathbb{R}^{m}) norm. This lemma will be used both in the proof of Proposition 2.2, and also later, when we prove existence of minimizing geodesics between immersions in Section 5.5.

Lemma 2.3.

Let ι:𝒩m\iota:\mathcal{N}\to\mathbb{R}^{m} be an isometric embedding, and let n2n\geq 2. Let K𝒩K\subset\mathcal{N} be a compact set, and let cHn(D,𝒩)c\in H^{n}(D,\mathcal{N}) be a curve whose image lies in KK. Let C>0C>0 be such that

θkcL2(dθ)<C,k=0,,n1.\|\nabla_{\partial_{\theta}}^{k}c^{\prime}\|_{L^{2}(\,\mathrm{d}\theta)}<C,\qquad k=0,\ldots,n-1.

For hHn(D;cT𝒩)h\in H^{n}(D;c^{*}T\mathcal{N}), denote by ιhHn(D;m)\iota_{*}h\in H^{n}(D;\mathbb{R}^{m}) the image of hh under the embedding. The extrinsic norm of hh is then defined by

hHn(ι)2:=ιhHn(D;m)2=02π|ιh|2+|θnιh|2dθ,\|h\|_{H^{n}(\iota)}^{2}:=\|\iota_{*}h\|_{H^{n}(D;\mathbb{R}^{m})}^{2}=\int_{0}^{2\pi}|\iota_{*}h|^{2}+|\partial_{\theta}^{n}\iota_{*}h|^{2}\,\,\mathrm{d}\theta,

where |||\cdot| is the norm in m\mathbb{R}^{m}, and θ=cN\partial_{\theta}=\nabla^{\mathbb{R}^{N}}_{c^{\prime}} is the standard derivative on m\mathbb{R}^{m}. Then, there exists a constant β>0\beta>0, depending only on ι\iota, KK and CC such that for every hHn(D;cT𝒩)h\in H^{n}(D;c^{*}T\mathcal{N}),

β1hHn(ι)hcβhHn(ι).\beta^{-1}\|h\|_{H^{n}(\iota)}\leq\|h\|_{\mathcal{H}_{c}}\leq\beta\|h\|_{H^{n}(\iota)}.

Throughout the proof we will use standard Sobolev embedding results of the space Hn1(D,cT𝒩)H^{n-1}(D,c^{*}T\mathcal{N}); these estimates can be found in any standard book on Sobolev spaces, e.g., [25], and the adaptation from real-valued functions to vector-bundle-valued functions is straightforward using parallel transport along the curve to a single tangent space. For completion, the estimates and their reduction to the real-valued case appear in Lemma 4.1 below.

Proof.

First, note that cHn1(D,cT𝒩)c^{\prime}\in H^{n-1}(D,c^{*}T\mathcal{N}), so the fact that there exists a bound on the L2L^{2} norms of θkc\nabla_{\partial_{\theta}}^{k}c^{\prime} for k=0,,n1k=0,\ldots,n-1 is not an additional assumption. Also, by standard Sobolev estimates, Hk1(D,cT𝒩)H^{k-1}(D,c^{*}T\mathcal{N}) continuously embeds into Ck2(D,cT𝒩)C^{k-2}(D,c^{*}T\mathcal{N}), that is

Ck2(D,cT𝒩)Ck,n,dim𝒩Hk1(D,cT𝒩),\|\cdot\|_{C^{k-2}(D,c^{*}T\mathcal{N})}\leq C_{k,n,\dim\mathcal{N}}\|\cdot\|_{H^{k-1}(D,c^{*}T\mathcal{N})},

where the constant Ck,n,dim𝒩C_{k,n,\dim\mathcal{N}} depends only on k,nk,n and the dimension of 𝒩\mathcal{N}. Therefore, we have that our bounds on θkcL2(dθ)\|\nabla_{\partial_{\theta}}^{k}c^{\prime}\|_{L^{2}(\,\mathrm{d}\theta)} imply that

θkc<C,k=0,,n2,\|\nabla_{\partial_{\theta}}^{k}c^{\prime}\|_{\infty}<C,\qquad k=0,\ldots,n-2,

by possibly enlarging the constant CC.

Next, note that hL2(ι)=hL2(dθ)\|h\|_{L^{2}(\iota)}=\|h\|_{L^{2}(\,\mathrm{d}\theta)}, since ι\iota is an isometric embedding |ιh|=|h||\iota_{*}h|=|h| pointwise for every θ\theta (here, the m\mathbb{R}^{m}-norm appears on the left-hand side, the T𝒩T\mathcal{N}-norm on the right-hand side).

Denote by II the second fundamental form of 𝒩\mathcal{N} in m\mathbb{R}^{m}, that is, for v,wTx𝒩v,w\in T_{x}\mathcal{N}, we have

II(v,w)=vmwv𝒩w.\textup{II}(v,w)=\nabla^{\mathbb{R}^{m}}_{v}w-\nabla^{\mathcal{N}}_{v}w.

In a coordinate patch on a tubular neighborhood of 𝒩\mathcal{N}, with coordinates (xi)i=1m(x_{i})_{i=1}^{m} such that (xa)a=1d(x_{a})_{a=1}^{d}, where d=dim𝒩d=\dim\mathcal{N} are coordinates on 𝒩\mathcal{N} and xαxa\partial_{x_{\alpha}}\perp\partial_{x_{a}} for a=1,,da=1,\ldots,d and α=d+1,,m\alpha=d+1,\ldots,m, we have

II(v,w)=Γabα(x)vawbα,\textup{II}(v,w)=\Gamma_{ab}^{\alpha}(x)v^{a}w^{b}\partial_{\alpha},

where Γijk\Gamma_{ij}^{k} are the Christoffel symbols of m\nabla^{\mathbb{R}^{m}} in these coordinates. Since v𝒩wII(v,w)\nabla^{\mathcal{N}}_{v}w\perp\textup{II}(v,w), we have

(2.3) |θιh|2=|θ𝒩h|2+|II(c,h)|2|θ𝒩h|2+C2|II|2|h|2|θ𝒩h|2+C|h|2,\begin{split}|\partial_{\theta}\iota_{*}h|^{2}&=|\nabla^{\mathcal{N}}_{\partial_{\theta}}h|^{2}+|\textup{II}(c^{\prime},h)|^{2}\leq|\nabla^{\mathcal{N}}_{\partial_{\theta}}h|^{2}+C^{2}|\textup{II}|^{2}|h|^{2}\\ &\leq|\nabla^{\mathcal{N}}_{\partial_{\theta}}h|^{2}+C^{\prime}|h|^{2},\end{split}

where C=C2supxK|II|2C^{\prime}=C^{2}\sup_{x\in K}|\textup{II}|^{2}. Integrating, we obtain

θ𝒩hL2(dθ)2θιhL2(dθ)2θ𝒩hL2(dθ)2+ChL2(dθ)2hH1(dθ)2.\|\nabla^{\mathcal{N}}_{\partial_{\theta}}h\|_{L^{2}(\,\mathrm{d}\theta)}^{2}\leq\|\partial_{\theta}\iota_{*}h\|_{L^{2}(\,\mathrm{d}\theta)}^{2}\leq\|\nabla^{\mathcal{N}}_{\partial_{\theta}}h\|_{L^{2}(\,\mathrm{d}\theta)}^{2}+C^{\prime}\|h\|_{L^{2}(\,\mathrm{d}\theta)}^{2}\lesssim\|h\|_{H^{1}(\,\mathrm{d}\theta)}^{2}.

Here and in the following we use the notation \lesssim to indicate that there exists a constant, which does not depend on hh, such that the inequality holds. For the second order terms we calculate

(2.4) θ2ιh=θθ𝒩h+θ(II(c,h))=(θ𝒩)2h+II(c,θ𝒩h)+θ(II(c,h)),\begin{split}\partial_{\theta}^{2}\iota_{*}h&=\partial_{\theta}\nabla^{\mathcal{N}}_{\partial_{\theta}}h+\partial_{\theta}(\textup{II}(c^{\prime},h))=(\nabla^{\mathcal{N}}_{\partial_{\theta}})^{2}h+\textup{II}(c^{\prime},\nabla^{\mathcal{N}}_{\partial_{\theta}}h)+\partial_{\theta}(\textup{II}(c^{\prime},h)),\end{split}

Since II and its derivatives are bounded on the compact set KK, we have

|θ2ιh||(θ𝒩)2h|+|c||θ𝒩h|+|c||h|+|θc||h|+|c||θh||(θ𝒩)2h|+|c||θ𝒩h|+|c|(1+|c|)|h|+|θ𝒩c||h|.\begin{split}|\partial_{\theta}^{2}\iota_{*}h|&\lesssim|(\nabla^{\mathcal{N}}_{\partial_{\theta}})^{2}h|+|c^{\prime}||\nabla^{\mathcal{N}}_{\partial_{\theta}}h|+|c^{\prime}||h|+|\partial_{\theta}c^{\prime}||h|+|c^{\prime}||\partial_{\theta}h|\\ &\lesssim|(\nabla^{\mathcal{N}}_{\partial_{\theta}})^{2}h|+|c^{\prime}||\nabla^{\mathcal{N}}_{\partial_{\theta}}h|+|c^{\prime}|(1+|c^{\prime}|)|h|+|\nabla^{\mathcal{N}}_{\partial_{\theta}}c^{\prime}||h|.\end{split}

where we used (2.3) when changing θ\partial_{\theta} to θ𝒩\nabla^{\mathcal{N}}_{\partial_{\theta}} (applied to cc^{\prime} and hh). Since n2n\geq 2, we use again the Sobolev embedding Hn1(D,cT𝒩)Cn2(D,cT𝒩)C0(D,cT𝒩)H^{n-1}(D,c^{*}T\mathcal{N})\subset C^{n-2}(D,c^{*}T\mathcal{N})\subset C^{0}(D,c^{*}T\mathcal{N}) to obtain that |c|<C|c^{\prime}|<C and hLC2hc\|h\|_{L^{\infty}}\leq C_{2}\|h\|_{\mathcal{H}_{c}} for some C2>0C_{2}>0 depending only on the dimension. We therefore have

|θ2ιh||(θ𝒩)2h|+|θ𝒩h|+|h|+hc|θ𝒩c|.|\partial_{\theta}^{2}\iota_{*}h|\lesssim|(\nabla^{\mathcal{N}}_{\partial_{\theta}})^{2}h|+|\nabla^{\mathcal{N}}_{\partial_{\theta}}h|+|h|+\|h\|_{\mathcal{H}_{c}}|\nabla^{\mathcal{N}}_{\partial_{\theta}}c^{\prime}|.

Squaring and integrating, and using that θ𝒩cL2<C\|\nabla^{\mathcal{N}}_{\partial_{\theta}}c^{\prime}\|_{L^{2}}<C, we obtain that,

θ2ιhL2(θ𝒩)2hL2+θ𝒩hL2+hL2+hchc,\|\partial_{\theta}^{2}\iota_{*}h\|_{L^{2}}\lesssim\|(\nabla^{\mathcal{N}}_{\partial_{\theta}})^{2}h\|_{L^{2}}+\|\nabla^{\mathcal{N}}_{\partial_{\theta}}h\|_{L^{2}}+\|h\|_{L^{2}}+\|h\|_{\mathcal{H}_{c}}\lesssim\|h\|_{\mathcal{H}_{c}},

and therefore

hH2(ι)hc.\|h\|_{H^{2}(\iota)}\lesssim\|h\|_{\mathcal{H}_{c}}.

The converse inequality follows in a similar manner, by using (2.4), to bound |(θ𝒩)2h||(\nabla^{\mathcal{N}}_{\partial_{\theta}})^{2}h| with |θ2ιh||\partial_{\theta}^{2}\iota_{*}h| and lower order terms.

For n>2n>2 the proof proceeds inductively in the same way — writing θnιh\partial_{\theta}^{n}\iota_{*}h in terms of (θ𝒩)nh(\nabla^{\mathcal{N}}_{\partial_{\theta}})^{n}h and lower order terms that involve the second fundamental form and its derivatives (as in (2.4)), and bounding the lower order terms in a similar manner. ∎

Proof of Proposition 2.2. Part I: Smooth structure and topology. An alternative characterization of Hn(D,𝒩)H^{n}(D,\mathcal{N}) is

Hn(D,𝒩)={cC(D,𝒩):c=exps(V) for some sC(D,𝒩),VHn(D;sT𝒩)},\begin{split}H^{n}(D,\mathcal{N})&=\Big{\{}c\in C(D,\mathcal{N})\,:\,c=\exp_{s}(V)\\ &\qquad\qquad\text{ for some }s\in C^{\infty}(D,\mathcal{N}),\,V\in H^{n}(D;s^{*}T\mathcal{N})\Big{\}},\end{split}

where exp\exp is the exponential map with respect to the Riemannian metric gg on 𝒩\mathcal{N} (see, e.g., [37, Lemma B.5]). This characterization induces a smooth structure on Hn(D,𝒩)H^{n}(D,\mathcal{N}), where the charts, modeled on Hn(D;sT𝒩)H^{n}(D;s^{*}T\mathcal{N}), are given by exps\exp_{s} for sC(D,𝒩)s\in C^{\infty}(D,\mathcal{N}). The tangent space at cc is Hn(D;cT𝒩)H^{n}(D;c^{*}T\mathcal{N}). See [27, 5.3–5.8] for details. This smooth structure is described in detail in [22, Section 3] (it is denoted there by 𝒜gs\mathcal{A}^{s}_{g}). In [22, Proposition 3.7] it is shown that this smooth structure coincides with the one induced by considering local charts on DD and 𝒩\mathcal{N} (which provides yet another characterization to Hn(D,𝒩)H^{n}(D,\mathcal{N})).

Next, note that the topology induced by this smooth structure is equivalent to the topology induced on Hn(D,𝒩)H^{n}(D,\mathcal{N}) by distext\operatorname{dist}^{\text{ext}} [37, Lemma B.7]. The inner product c\mathcal{H}_{c} describes the Hilbert space topology on the tangent space TcHn(D,𝒩)=Hn(D;cT𝒩)T_{c}H^{n}(D,\mathcal{N})=H^{n}(D;c^{*}T\mathcal{N}). Since these are also the modeling spaces for the natural chart construction, \mathcal{H} is a strong Riemannian metric. Thus the distance function dist\operatorname{dist}^{\mathcal{H}} induced by \mathcal{H} induces the topology of Hn(D,𝒩)H^{n}(D,\mathcal{N}).

Part II: Openness of n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) in Hn(D,𝒩)H^{n}(D,\mathcal{N}). Taking again the extrinsic point of view n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) is the intersection of Hn(D,𝒩)H^{n}(D,\mathcal{N}) with the set of maps cHn(D;m)c\in H^{n}(D;\mathbb{R}^{m}) such that c0c^{\prime}\neq 0. By the Sobolev embedding cL(D;m)CcHn(D;m)\|c^{\prime}\|_{L^{\infty}(D;\mathbb{R}^{m})}\leq C\|c\|_{H^{n}(D;\mathbb{R}^{m})}, which holds since n2n\geq 2, it is immediate that c0c^{\prime}\neq 0 is an open condition in Hn(D;m)H^{n}(D;\mathbb{R}^{m}), and hence n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) is open in Hn(D,𝒩)H^{n}(D,\mathcal{N}).

Part III: Completeness of (Hn(D,𝒩),dist)(H^{n}(D,\mathcal{N}),\operatorname{dist}^{\mathcal{H}}). Let cjHn(D,𝒩)c_{j}\in H^{n}(D,\mathcal{N}) be a Cauchy sequence with respect to dist\operatorname{dist}^{\mathcal{H}}. We aim to show that cjc_{j} is also a Cauchy sequence with respect to distext\operatorname{dist}^{\text{ext}}. Then, since (Hn(D,𝒩),distext)(H^{n}(D,\mathcal{N}),\operatorname{dist}^{\text{ext}}) is complete, we will obtain that the sequence converges to some cHn(D,𝒩)c_{\infty}\in H^{n}(D,\mathcal{N}); since the topologies induced by distext\operatorname{dist}^{\text{ext}} and dist\operatorname{dist}^{\mathcal{H}} coincide, we will obtain that (Hn(D,𝒩),dist)(H^{n}(D,\mathcal{N}),\operatorname{dist}^{\mathcal{H}}) is complete as well.

Since (cj)j(c_{j})_{j\in\mathbb{N}} is a dist\operatorname{dist}^{\mathcal{H}}-Cauchy sequence, it lies inside some dist\operatorname{dist}^{\mathcal{H}}-ball BB of radius r>0r>0 centered at some c0Hn(D,𝒩)c_{0}\in H^{n}(D,\mathcal{N}). By taking a slightly larger rr we can also assume that for every jk𝒩j\leq k\in\mathcal{N} there exists a path cjk:[0,1]Hn(D,𝒩)c_{jk}:[0,1]\to H^{n}(D,\mathcal{N}) connecting cjc_{j} and ckc_{k} (that is, cjk(0)=cjc_{jk}(0)=c_{j} and cjk(1)=ckc_{jk}(1)=c_{k}), such that cjk(t)Bc_{jk}(t)\in B for every t[0,1]t\in[0,1] and L(cjk)<dist(cj,ck)+1jL^{\mathcal{H}}(c_{jk})<\operatorname{dist}^{\mathcal{H}}(c_{j},c_{k})+\frac{1}{j}, where LL^{\mathcal{H}} is the length of cjkc_{jk} with respect to the metric \mathcal{H}.

We now show that all the curves in BB lie inside a compact subset of 𝒩\mathcal{N}; moreover, we show that for some C>0C>0, all curves cBc\in B satisfy

θkcL2(dθ)<C,k=0,,n1.\|\nabla_{\partial_{\theta}}^{k}c^{\prime}\|_{L^{2}(d\theta)}<C,\qquad k=0,\ldots,n-1.

It then follows by Lemma 2.3 that there exists a constant β>0\beta>0, such that for every cBc\in B and every hHn(D;cT𝒩)h\in H^{n}(D;c^{*}T\mathcal{N}),

ιhHn(D;m)βhc,\|\iota_{*}h\|_{H^{n}(D;\mathbb{R}^{m})}\leq\beta\|h\|_{\mathcal{H}_{c}},

where ιhHn(D;m)\iota_{*}h\in H^{n}(D;\mathbb{R}^{m}) is the image of hh under the embedding, and where Hn(D;m)\|\cdot\|_{H^{n}(D;\mathbb{R}^{m})} is the standard norm in Hn(D;m)H^{n}(D;\mathbb{R}^{m}) (see Lemma 2.3). Therefore, for every jk𝒩j\leq k\in\mathcal{N},

distext(cj,ck)Lext(cjk)βL(cjk)<β(dist(cj,ck)+1j),\operatorname{dist}^{\text{ext}}(c_{j},c_{k})\leq L^{\text{ext}}(c_{jk})\leq\beta L^{\mathcal{H}}(c_{jk})<\beta\left(\operatorname{dist}^{\mathcal{H}}(c_{j},c_{k})+\frac{1}{j}\right),

where LextL^{\text{ext}} is the length with respect to the external structure. Thus (cj)j(c_{j})_{j\in\mathbb{N}} is a distext\operatorname{dist}^{\text{ext}}-Cauchy sequence and the proof is complete.

It remains to verify the assumptions of Lemma 2.3. Let now c¯B=B(c0,r)\bar{c}\in B=B(c_{0},r). By definition, there exists a path c:[0,1]Hn(D,𝒩)c:[0,1]\to H^{n}(D,\mathcal{N}), with c(0)=c0c(0)=c_{0} and c(1)=c¯c(1)=\bar{c} such that L(c)<rL^{\mathcal{H}}(c)<r. Now, for every θ0D\theta_{0}\in D, we have

dist𝒩(c0(θ0),c¯(θ0))01|tc(t,θ0)|dt01tcLC01tctc=CL(c)<Cr,\begin{split}\operatorname{dist}_{\mathcal{N}}(c_{0}(\theta_{0}),\bar{c}(\theta_{0}))&\leq\int_{0}^{1}|\partial_{t}c(t,\theta_{0})|\,\,\mathrm{d}t\leq\int_{0}^{1}\|\partial_{t}c\|_{L^{\infty}}\leq C\int_{0}^{1}\|\partial_{t}c_{t}\|_{\mathcal{H}_{c}}\\ &=CL^{\mathcal{H}}(c)<Cr,\end{split}

where we used the Sobolev embedding on vector bundles Hn1(D,cT𝒩)Cn2(D,cT𝒩)C0(D,cT𝒩)H^{n-1}(D,c^{*}T\mathcal{N})\subset C^{n-2}(D,c^{*}T\mathcal{N})\subset C^{0}(D,c^{*}T\mathcal{N}) as in the proof of Lemma 2.3. It follows that the images of all the curves in BB lie in a compact subset of 𝒩\mathcal{N} (namely a neighborhood of radius CrCr around the image of c0c_{0}).

Now, let k=0,,n1k=0,\ldots,n-1, then

θkc¯L2(dθ)θkc0L2(dθ)=01t(D|θkc|2dθ)1/2dt=01Dg(θkc,θktc)dθ(D|θkc|2dθ)1/2dt01(D|θktc|2dθ)1/2dt01tccdt=L(c)<r,\begin{split}\|\nabla_{\partial_{\theta}}^{k}\bar{c}^{\prime}\|_{L^{2}(\,\mathrm{d}\theta)}-\|\nabla_{\partial_{\theta}}^{k}c_{0}^{\prime}\|_{L^{2}(\,\mathrm{d}\theta)}&=\int_{0}^{1}\partial_{t}\left(\int_{D}|\nabla_{\partial_{\theta}}^{k}c^{\prime}|^{2}\,\,\mathrm{d}\theta\right)^{1/2}\,\,\mathrm{d}t\\ &=\int_{0}^{1}\frac{\int_{D}g(\nabla_{\partial_{\theta}}^{k}c^{\prime},\nabla_{\partial_{\theta}}^{k}\partial_{t}c^{\prime})\,\,\mathrm{d}\theta}{\left(\int_{D}|\nabla_{\partial_{\theta}}^{k}c^{\prime}|^{2}\,\,\mathrm{d}\theta\right)^{1/2}}\,\,\mathrm{d}t\\ &\leq\int_{0}^{1}\left(\int_{D}|\nabla_{\partial_{\theta}}^{k}\partial_{t}c^{\prime}|^{2}\,\,\mathrm{d}\theta\right)^{1/2}\,\,\mathrm{d}t\\ &\leq\int_{0}^{1}\|\partial_{t}c\|_{\mathcal{H}_{c}}\,\,\mathrm{d}t=L^{\mathcal{H}}(c)<r,\end{split}

where we used again that the L2L^{2} norms of θkh\nabla_{\partial_{\theta}}^{k}h for k=0,,nk=0,\ldots,n are controlled by hc\|h\|_{\mathcal{H}_{c}} (again, we refer to Lemma 4.1 for an exact statement). The uniform bound on θkc¯L2(dθ)\|\nabla_{\partial_{\theta}}^{k}\bar{c}^{\prime}\|_{L^{2}(\,\mathrm{d}\theta)} immediately follows, and thus the assumptions of Lemma 2.3 are fulfilled, uniformly on BB. ∎

3. Reparametrization invariant Sobolev metrics on spaces of curves

3.1. The metric and geodesic equation in the smooth category

As detailed in the introduction, we are interested in reparametrization invariant Sobolev metrics on the spaces Imm(D,𝒩)\operatorname{Imm}(D,\mathcal{N}) and n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) defined above, and, more accurately, in metrics of the type (1.2):

Gc(h,k)=i=0nai(c)Dg(sih,sik)ds,aiC((0,),[0,)), for i=0,,nanda0,an>0.\begin{split}&G_{c}(h,k)=\sum_{i=0}^{n}a_{i}(\ell_{c})\int_{D}g(\nabla_{\partial_{s}}^{i}h,\nabla_{\partial_{s}}^{i}k)\,\mathrm{d}s,\\ &a_{i}\in C^{\infty}((0,\infty),[0,\infty)),\text{ for }i=0,\ldots,n\quad\text{and}\quad a_{0},a_{n}>0.\end{split}

We now calculate the geodesic equation associated with GcG_{c} in smooth settings; in the next subsection we extend the treatment to Sobolev settings. To derive the geodesic equation it will be more convenient to write the metric using the so-called inertia operator, i.e., use integration by parts to write GG as

(3.1) Gc(h,k)=Dg(Ach,k)ds+Bc(h,k).G_{c}(h,k)=\int_{D}g(A_{c}h,k)\,\mathrm{d}s+B_{c}(h,k).

Here

(3.2) Ac:TcImm(D,𝒩)TcImm(D,𝒩),A_{c}:T_{c}\operatorname{Imm}(D,\mathcal{N})\to T_{c}\operatorname{Imm}(D,\mathcal{N}),

is called the inertia operator of the metric GG and Bc(h,k)B_{c}(h,k) depends solely on the boundary of DD and stems from the integration by parts process. Thus for closed curves the operator BB is not present.

Lemma 3.1.

The inertia operator of the metric (1.2) takes the form:

(3.3) Ac(h)=i=0n(1)iai(c)s2ih,A_{c}(h)=\sum_{i=0}^{n}(-1)^{i}a_{i}(\ell_{c})\nabla_{\partial_{s}}^{2i}h,

For open curves, i.e. D=[0,2π]D=[0,2\pi], the boundary operator BB is given by:

(3.4) Bc(h,k)=i=1nai(c)j=0i1(1)i+j1g(si+jh,sij1k)|02π.B_{c}(h,k)=\sum_{i=1}^{n}a_{i}(\ell_{c})\sum_{j=0}^{i-1}(-1)^{i+j-1}g(\nabla_{\partial_{s}}^{i+j}h,\nabla_{\partial_{s}}^{i-j-1}k)\Big{|}^{2\pi}_{0}\;.
Proof.

These formulas follow directly from the integration by parts formula

(3.5) Dg(h,sk)ds=g(h,k)|DDg(sh,k)ds.\int_{D}g(h,\nabla_{\partial_{s}}k)\,\mathrm{d}s=g(h,k)|_{\partial D}-\int_{D}g(\nabla_{\partial_{s}}h,k)\,\mathrm{d}s\;.

Note that for closed curves we have D=S1D=S^{1} and thus D=\partial D=\emptyset. ∎

Before we calculate the geodesic equation we will collect variational formulas of several quantities that appear in the metric. In the following we will denote the variation of a quantity in direction hTcImm(D,𝒩)h\in T_{c}\operatorname{Imm}(D,\mathcal{N}) by Dc,hD_{c,h}.

Lemma 3.2.

Let cImm(D,𝒩)c\in\operatorname{Imm}(D,\mathcal{N}) and hTcImm(D,𝒩)h\in T_{c}\operatorname{Imm}(D,\mathcal{N}). Then

(3.6) Dc,h|c|\displaystyle D_{c,h}|c^{\prime}| =g(v,sh)|c|\displaystyle=g(v,\nabla_{\partial_{s}}h)|c^{\prime}|
(3.7) Dc,hds\displaystyle D_{c,h}\,\mathrm{d}s =g(v,sh)ds\displaystyle=g(v,\nabla_{\partial_{s}}h)\,\mathrm{d}s
(3.8) Dc,hc\displaystyle D_{c,h}\ell_{c} =Dg(v,sh)ds\displaystyle=\int_{D}g(v,\nabla_{\partial_{s}}h)\,\mathrm{d}s

where v=c/|c|v=c^{\prime}/{|c^{\prime}|} denotes the unit length tangent vector to the curve cc. Extending the connection, as described in [9, Section 3], we can also calculate the variation of the covariant derivtive s\nabla_{\partial_{s}} applied to a tangent vector kTcImm(D,𝒩)k\in T_{c}\operatorname{Imm}(D,\mathcal{N}):

(3.9) hsk\displaystyle\nabla_{h}\nabla_{\partial_{s}}k =g(v,sh)sk+shk+(v,h)k;\displaystyle=-g(v,\nabla_{\partial_{s}}h)\nabla_{\partial_{s}}k+\nabla_{\partial_{s}}\nabla_{h}k+\mathcal{R}(v,h)k;

where \mathcal{R} denotes the (Riemannian) curvature of (𝒩,g)(\mathcal{N},g).

Proof.

The first three formulas follow by straight-forward calculations, similar as for curves with values in Euclidean spaces, see, e.g., [28, 12]. For the last formula we follow the more general presentation in [9], where the variation of the Laplacian for DD being a compact manifold of arbitrary dimension has been derived. Using the formula

(3.10) hθk=θhk+(h,c)k\displaystyle\nabla_{h}\nabla_{\partial_{\theta}}k=\nabla_{\partial_{\theta}}\nabla_{h}k+\mathcal{R}(h,c^{\prime})k

for swapping covariant derivatives, see e.g. [9, Section 3.8], we obtain

hsk\displaystyle\nabla_{h}\nabla_{\partial_{s}}k =Df,h(|c|1)θk+|c|1hθk\displaystyle=D_{f,h}\left(|c^{\prime}|^{-1}\right)\nabla_{\partial_{\theta}}k+|c^{\prime}|^{-1}\nabla_{h}\nabla_{\partial_{\theta}}k
=g(sh,v)sk+|c|1θhk+|c|1(h,c)k\displaystyle=-g(\nabla_{\partial_{s}}h,v)\nabla_{\partial_{s}}k+|c^{\prime}|^{-1}\nabla_{\partial_{\theta}}\nabla_{h}k+|c^{\prime}|^{-1}\mathcal{R}(h,c^{\prime})k

which concludes the proof since v=|c|1cv=|c^{\prime}|^{-1}c^{\prime}. ∎

We are now able to calculate the geodesic equation. In the following calculation we will restrict to first order metrics, for which the exact form of the geodesic spray will be of importance in the proof of the local well-posedness result. For higher order metrics the existence and well-posedness of the geodesic equation will follow from general principles on strong metrics and we will thus not include these cumbersome calculations. The interested reader can consult the related calculations in [9], where the geodesic equations are derived for general higher order metrics (under the assumption that DD has no boundary). The geodesic equation for constant coefficient metrics on closed curves in Euclidean space also appears in [14, Theorem 1.1].

Lemma 3.3.

The geodesic equation of the first-order Sobolev-type metric, as defined in (1.2) for n=1n=1, is given by the set of equations:

t(Acct)=g(v,sct)Acct12Ψc(ct,ct)svg(sct,Acct)v+a1(c)(ct,sct)v,\nabla_{\partial_{t}}(A_{c}c_{t})=-g(v,\nabla_{\partial_{s}}c_{t})A_{c}c_{t}-\frac{1}{2}\Psi_{c}(c_{t},c_{t})\nabla_{\partial_{s}}v-g(\nabla_{\partial_{s}}c_{t},A_{c}c_{t})v\\ +a_{1}(\ell_{c})\mathcal{R}(c_{t},\nabla_{\partial_{s}}c_{t})v,

where the quadratic form Ψc(ct,ct)\Psi_{c}(c_{t},c_{t}) is given by

Ψc(ct,ct)\displaystyle\Psi_{c}(c_{t},c_{t}) =a0(c)g(ct,ct)+a0(c)Dg(ct,ct)ds\displaystyle=a_{0}(\ell_{c})g(c_{t},c_{t})+a_{0}^{\prime}(\ell_{c})\int_{D}g(c_{t},c_{t})\,\mathrm{d}s
a1(c)g(sct,sct)+a1(c)Dg(sct,sct)ds\displaystyle\quad-a_{1}(\ell_{c})g(\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})+a_{1}^{\prime}(\ell_{c})\int_{D}g(\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s

For open curves, D=[0,2π]D=[0,2\pi], we get the following boundary conditions:

(2t(a1(c)sct)+Ψc(ct,ct)v)|θ=0,2π=0.\displaystyle\Big{(}-2\nabla_{\partial_{t}}\left(a_{1}(\ell_{c})\nabla_{\partial_{s}}c_{t}\right)+\Psi_{c}(c_{t},c_{t})v\Big{)}\bigg{|}_{\theta=0,2\pi}=0\;.

The proof of this result is postponed to Appendix A.1.

3.2. The induced metric on Sobolev immersions

To obtain the desired completeness and well-posedness results we consider the extension of the metric GG (of order nn) on the Banach manifolds q(D,𝒩)Imm(D,𝒩)\mathcal{I}^{q}(D,\mathcal{N})\supset\operatorname{Imm}(D,\mathcal{N}), for qmax{n,2}q\geq\max\{n,2\}, as defined in Definition 2.1 above.

Our aim in the rest of the section is to show the smoothness of the metrics GG on q(D,𝒩)\mathcal{I}^{q}(D,\mathcal{N}) (assuming qnq\geq n). First, we need to introduce some mixed order spaces:

Definition 3.4.

Let q2q\geq 2 and qk0q\geq k\geq 0. We define the function space:

Hqk(D,T𝒩)={hHk(D,T𝒩):πhq(D,𝒩)}.H^{k}_{\mathcal{I}^{q}}(D,T\mathcal{N})=\left\{h\in H^{k}(D,T\mathcal{N}):\pi\circ h\in\mathcal{I}^{q}(D,\mathcal{N})\right\}\;.

We have the following result concerning their manifold structure and the operator θ\nabla_{\partial_{\theta}}:

Lemma 3.5.

The spaces Hqk(D,T𝒩)H^{k}_{\mathcal{I}^{q}}(D,T\mathcal{N}) are smooth Hilbert manifolds for any q2q\geq 2 and qk0q\geq k\geq 0. The mapping

(3.11) θ:Hqk(D,T𝒩)Hqk1(D,T𝒩)\nabla_{\partial_{\theta}}:H^{k}_{\mathcal{I}^{q}}(D,T\mathcal{N})\to H^{k-1}_{\mathcal{I}^{q}}(D,T\mathcal{N})

is a bounded linear mapping for 1kq1\leq k\leq q.

Proof.

The first part of this result can be found in [10, Theorem 2.4], while the second part follows directly from the definition of the space Hqk(D,T𝒩)H^{k}_{\mathcal{I}^{q}}(D,T\mathcal{N}). ∎

Note that Hqq(D,T𝒩)=Tq(D,𝒩)H^{q}_{\mathcal{I}^{q}}(D,T\mathcal{N})=T\mathcal{I}^{q}(D,\mathcal{N}). If k<qk<q then Hqk(D,T𝒩)H^{k}_{\mathcal{I}^{q}}(D,T\mathcal{N}) is the the robust fiber completion of the weak Riemannian manifold (q(D,T𝒩),Gk)(\mathcal{I}^{q}(D,T\mathcal{N}),G^{k}) with the Sobolev metric GkG^{k} from (1.2) in the sense described in [26]. These spaces will appear, when we repeatedly apply s\nabla_{\partial_{s}} to a vector field hh along an HnH^{n}-immersion (s\nabla_{\partial_{s}} will reduce the order of the vector field, but not of its foot point). To show the smoothness of the metric we need the following result:

Lemma 3.6.

Let q2q\geq 2. Then the mapping

(3.12) Hqk+1(D,T𝒩)\displaystyle H^{k+1}_{\mathcal{I}^{q}}(D,T\mathcal{N}) Hqk(D,T𝒩)\displaystyle\rightarrow H^{k}_{\mathcal{I}^{q}}(D,T\mathcal{N})
(3.13) h\displaystyle h sh=1|π(h)|θh\displaystyle\mapsto\nabla_{\partial_{s}}h=\frac{1}{|\pi(h)|}\nabla_{\partial_{\theta}}h

is smooth for any k0k\geq 0.

Proof.

The mapping

(3.14) Hqk+1(D,T𝒩)\displaystyle H^{k+1}_{\mathcal{I}^{q}}(D,T\mathcal{N}) Hqk(D,T𝒩)\displaystyle\rightarrow H^{k}_{\mathcal{I}^{q}}(D,T\mathcal{N})
(3.15) h\displaystyle h θh\displaystyle\mapsto\nabla_{\partial_{\theta}}h

is smooth by Lemma 3.5. By the module properties of Sobolev spaces, multiplication Hq(D,)×Hqk(D,T𝒩)Hqk(D,T𝒩)H^{q}(D,\mathbb{R})\times H^{k}_{\mathcal{I}^{q}}(D,T\mathcal{N})\rightarrow H^{k}_{\mathcal{I}^{q}}(D,T\mathcal{N}) is smooth for q2q\geq 2 and k0k\geq 0. Thus the result follows since |π(h)|Hq(D,)|\pi(h)|\in H^{q}(D,\mathbb{R}). ∎

Using this lemma we immediately obtain the smoothness of the metric:

Theorem 3.7.

Let q2q\geq 2. Consider the Sobolev metric GG on Imm(D,𝒩)\operatorname{Imm}(D,\mathcal{N}) of order nqn\leq q of the form (1.2). Then GG extends to a smooth Riemannian metric on q(D,𝒩)\mathcal{I}^{q}(D,\mathcal{N}). For q=nq=n the metric GG is a strong Riemannian metric on n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}).

Proof.

Iterating Lemma 3.6 we have that

(3.16) si:Tq(D,𝒩)=Hqq(D,T𝒩)Hqqi(D,T𝒩)Lq2(D,T𝒩)\nabla_{\partial_{s}}^{i}:T\mathcal{I}^{q}(D,\mathcal{N})=H^{q}_{\mathcal{I}^{q}}(D,T\mathcal{N})\rightarrow H^{q-i}_{\mathcal{I}^{q}}(D,T\mathcal{N})\subset L^{2}_{\mathcal{I}^{q}}(D,T\mathcal{N})

is smooth for 0in0\leq i\leq n. Thus the mapping

Tq(D,𝒩)×qTq(D,𝒩)\displaystyle T\mathcal{I}^{q}(D,\mathcal{N})\times_{\mathcal{I}^{q}}T\mathcal{I}^{q}(D,\mathcal{N}) L1(D,)\displaystyle\to L^{1}(D,\mathbb{R})
(h,k)\displaystyle(h,k) gc(sih,sik)|c|\displaystyle\mapsto g_{c}(\nabla_{\partial_{s}}^{i}h,\nabla_{\partial_{s}}^{i}k)|c^{\prime}|

is smooth as well. Here we used again the module properties of Sobolev spaces. It remains to show the smoothness of ccc\mapsto\ell_{c}. Therefore we use the fact that

L1(D,)\displaystyle L^{1}(D,\mathbb{R}) \displaystyle\to\mathbb{R}
f\displaystyle f fdθ\displaystyle\mapsto\int f\,\mathrm{d}\theta

is a bounded linear operator, hence it immediately follows that the length function cc=|c|dθc\mapsto\ell_{c}=\int|c^{\prime}|\,\mathrm{d}\theta is smooth. For n2n\geq 2 the metric GG is a strong Riemannian metric on n(D,N)\mathcal{I}^{n}(D,N) since for each cn(D,𝒩)c\in\mathcal{I}^{n}(D,\mathcal{N}) the inner product Gc(h,k)G_{c}(h,k) describes the Hilbert space structure on Tcn(D,N)T_{c}\mathcal{I}^{n}(D,N) (This is best seen in a local chart, whose base is, by definition, around a smooth cc, otherwise one has to deal with ΓHn(cTN)\Gamma_{H^{n}}(c^{*}TN) for cc a Sobolev HnH^{n}-immersion). ∎

3.3. Local well-posedness of the geodesic equation

The local well-posedness results as summarized in the following theorem are based on the seminal method of Ebin and Marsden [20]. They are known in the case of closed curves, see [10, 28, 6], but to the best of our knowledge they are new for the case of open curves. However, as local well-posedness is not the focus of the current article, we postpone the proof of this result to Appendix A.2.

Theorem 3.8.

Let D=[0,2π]D=[0,2\pi] or D=S1D=S^{1}. Let GG be a Sobolev metric of order n1n\geq 1 of the form (1.2) on q(D,𝒩)\mathcal{I}^{q}(D,\mathcal{N}), with either q2nq\geq 2n or q=n2q=n\geq 2. We have:

  1. (1)

    The initial value problem for the geodesic equation has unique local solutions in the Banach manifold q(D,𝒩)\mathcal{I}^{q}(D,\mathcal{N}). The solutions depend smoothly on tt and on the initial conditions c(0,)c(0,\cdot) and ct(0,)c_{t}(0,\cdot). Moreover, the Riemannian exponential mapping exp\operatorname{exp} exists and is smooth on a neighborhood of the zero section of the tangent bundle, and (π,exp)(\pi,\operatorname{exp}) is a diffeomorphism from a (possibly smaller) neighborhood of the zero section of Tq(D,𝒩)T\mathcal{I}^{q}(D,\mathcal{N}) to a neighborhood of the diagonal in q(D,𝒩)×q(D,𝒩)\mathcal{I}^{q}(D,\mathcal{N})\times\mathcal{I}^{q}(D,\mathcal{N}).

  2. (2)

    The results of part 1 (local well-posedness of the geodesic equation and properties of the exponential map) continue to hold on q(D,𝒩)C(Do,𝒩)\mathcal{I}^{q}(D,\mathcal{N})\cap C^{\infty}(D^{o},\mathcal{N}), where DoD^{o} is the interior of DD.

Note that for D=S1D=S^{1} we have Imm(S1,𝒩)=q(S1,𝒩)C(S1,𝒩)\operatorname{Imm}(S^{1},\mathcal{N})=\mathcal{I}^{q}(S^{1},\mathcal{N})\cap C^{\infty}(S^{1},\mathcal{N}), i.e., the local well-posedness continues to hold in the smooth category.

4. Estimates

In this section we prove some interpolation inequalities for Sobolev sections of the tangent bundle, that will be needed for proving metric completeness of (n(D,𝒩),G)(\mathcal{I}^{n}(D,\mathcal{N}),G) in various cases. For vector-space-valued functions, these inequalities are rather simple adaptations of standard inequalities; this is the case when 𝒩=d\mathcal{N}=\mathbb{R}^{d}, as sections of cTdc^{*}T\mathbb{R}^{d} can be regarded as vector-space-valued functions (see [14, Lemmas 2.14–2.15], [12, Lemma 2.4] for the case D=S1D=S^{1}).

For a general target manifold, two things change: first, instead of working with a section hHk(D,cT𝒩)h\in H^{k}(D,c^{*}T\mathcal{N}) directly, we need to parallel transport hh to a single base point, that is, to work with

H(θ)=Πθθ0h(θ)Hk(D,Tc(θ0)𝒩)Hk(D,dim𝒩),H(\theta)=\Pi_{\theta}^{\theta_{0}}h(\theta)\in H^{k}(D,T_{c(\theta_{0})}\mathcal{N})\simeq H^{k}(D,\mathbb{R}^{\dim\mathcal{N}}),

where θ0D\theta_{0}\in D is a base point, and Πθθ0\Pi_{\theta}^{\theta_{0}} is the parallel transport, in 𝒩\mathcal{N}, from Tc(θ)𝒩T_{c(\theta)}\mathcal{N} to Tc(θ0)𝒩T_{c(\theta_{0})}\mathcal{N}, along cc. The reason for using HH is that it is a vector-space-valued function, and so we can take regular derivatives of HH and use the fundamental theorem of calculus. The derivatives of HH relate to covariant derivatives of hh via

(4.1) H(θ)=ddθΠθθ0h(θ)=Πθθ0θh(θ).H^{\prime}(\theta)=\frac{d}{\,\mathrm{d}\theta}\Pi_{\theta}^{\theta_{0}}h(\theta)=\Pi_{\theta}^{\theta_{0}}\nabla_{\partial_{\theta}}h(\theta).

See, e.g., [17, Chapter 2, exercise 2]. Note that, since the parallel transport operator is an isometry, we have |H(θ)|=|h(θ)||H(\theta)|=|h(\theta)|, |H(θ)|=|θh(θ)||H^{\prime}(\theta)|=|\nabla_{\partial_{\theta}}h(\theta)|, and so on for higher order derivatives.

The second difference from the Euclidean case arises when D=S1D=S^{1}. In the Euclidean case we obtain inequalities for periodic functions, that are generally better than the ones for general functions (and this fact is essential for completeness of constant coefficients metrics). However, when 𝒩d\mathcal{N}\neq\mathbb{R}^{d}, even though h(0)=h(2π)h(0)=h(2\pi), it is not true that H(0)=H(2π)H(0)=H(2\pi), because the holonomy along the curve cc is in general non-trivial (that is, Π2π0idTc(0)𝒩\Pi_{2\pi}^{0}\neq\mathrm{id}_{T_{c(0)}\mathcal{N}}). Therefore, we need to bound the amount by which HH fails to be periodic, and to prove estimates for such “almost periodic” functions.

We now state the estimates; first the inequalities that hold for both D=S1D=S^{1} or D=[0,2π]D=[0,2\pi], and then inequalities that hold only in the periodic case. As the proof of the periodic case is long and somewhat different from the rest of the analysis in this paper, we postpone it to Appendix B. This is done solely for the sake of readability — these estimates are at the core of proving the metric completeness of (n(S1;𝒩);G)(\mathcal{I}^{n}(S^{1};\mathcal{N});G) for GG with constant coefficients, and are one of the main differences between the analysis of manifold-valued curves and of d\mathbb{R}^{d}-valued curves.

Lemma 4.1 (General estimates).

If n2n\geq 2, cn(D,𝒩)c\in\mathcal{I}^{n}(D,\mathcal{N}) and hHn(D,cT𝒩)h\in H^{n}(D,c^{*}T\mathcal{N}), then for 0k<n0\leq k<n, there exists C=C(k,n,dim𝒩)>0C=C(k,n,\dim\mathcal{N})>0 such that

(4.2) a2kskhL2(ds)2C(hL2(ds)2+a2nsnhL2(ds)2),a^{2k}\|\nabla_{\partial_{s}}^{k}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\leq C\left(\|h\|^{2}_{L^{2}(\,\mathrm{d}s)}+a^{2n}\|\nabla_{\partial_{s}}^{n}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\right)\,,

and

(4.3) a2kskhL2C(a1hL2(ds)2+a2n1snhL2(ds)2),a^{2k}\|\nabla_{\partial_{s}}^{k}h\|^{2}_{L^{\infty}}\leq C\left(a^{-1}\|h\|^{2}_{L^{2}(\,\mathrm{d}s)}+a^{2n-1}\|\nabla_{\partial_{s}}^{n}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\right)\,,

for every a(0,c]a\in(0,\ell_{c}]. The same holds when we replace s\nabla_{\partial_{s}} with θ\nabla_{\partial_{\theta}} and ds\,\mathrm{d}s with dθ\,\mathrm{d}\theta, with a(0,2π]a\in(0,2\pi].

Proof.

Since all the norms involved (in the ds\,\mathrm{d}s case) are reparametrization-invariant, we can assume that cc is arc-length parametrized. In this case, we have s=θ\nabla_{\partial_{s}}=\nabla_{\partial_{\theta}}, ds=dθ\,\mathrm{d}s=\,\mathrm{d}\theta, where θ[0,c]\theta\in[0,\ell_{c}] (and in the case D=S1D=S^{1}, we identify the points θ=0\theta=0 and θ=c\theta=\ell_{c}). Define

H:[0,c]Tc(0)𝒩dim𝒩H(θ)=Πθ0h(θ).H:[0,\ell_{c}]\to T_{c(0)}\mathcal{N}\equiv\mathbb{R}^{\dim\mathcal{N}}\qquad H(\theta)=\Pi_{\theta}^{0}h(\theta).

From (4.1) we have

|θkh(θ)|=|Πθ0θkh(θ)|=|dkdθkH(θ)|.\left|\nabla_{\partial_{\theta}}^{k}h(\theta)\right|=\left|\Pi_{\theta}^{0}\nabla_{\partial_{\theta}}^{k}h(\theta)\right|=\left|\frac{d^{k}}{\,\mathrm{d}\theta^{k}}H(\theta)\right|.

In order to prove (4.2), we therefore need to prove that

a2k0c|θkH|2dθC(0c|H|2dθ+a2n0c|θnH|2dθ),a^{2k}\int_{0}^{\ell_{c}}|\partial_{\theta}^{k}H|^{2}\,\,\mathrm{d}\theta\leq C\left(\int_{0}^{\ell_{c}}|H|^{2}\,\,\mathrm{d}\theta+a^{2n}\int_{0}^{\ell_{c}}|\partial_{\theta}^{n}H|^{2}\,\,\mathrm{d}\theta\right),

for every a(0,c]a\in(0,\ell_{c}], and similarly for (4.3). Since HH is valued in dim𝒩\mathbb{R}^{\dim\mathcal{N}}, this is a standard Sobolev inequality, see, e.g., [25, Theorem 7.40].

The dθ\,\mathrm{d}\theta case is similar, but simpler (no need to reparametrize cc first). ∎

Lemma 4.2 (Estimates for S1S^{1}).

If n2n\geq 2, cn(S1,𝒩)c\in\mathcal{I}^{n}(S^{1},\mathcal{N}) and hHn(S1,cT𝒩)h\in H^{n}(S^{1},c^{*}T\mathcal{N}), then for 0<k<n0<k<n, there exists C>0C>0, depending on k,n,dim𝒩k,n,\dim\mathcal{N}, the injectivity radius and the upper and lower bounds on the sectional curvature of 𝒩\mathcal{N}, such that

(4.4) skhL2(ds)2Cmin{1,c2}(hL2(ds)2+snhL2(ds)2).\|\nabla_{\partial_{s}}^{k}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\leq C\min\left\{1,\ell_{c}^{2}\right\}\left(\|h\|^{2}_{L^{2}(\,\mathrm{d}s)}+\|\nabla_{\partial_{s}}^{n}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\right).
Proof.

See Appendix B. ∎

Remark 4.3.

It is interesting to compare inequality (4.4) to the equivalent one in the Euclidean settings [14, Lemma 2.14], that is, when 𝒩=d\mathcal{N}=\mathbb{R}^{d}. There we have

shL2(ds)2c24s2hL2(ds)2,\|\nabla_{\partial_{s}}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\leq\frac{\ell_{c}^{2}}{4}\|\nabla_{\partial_{s}}^{2}h\|^{2}_{L^{2}(\,\mathrm{d}s)},

from which higher order inequalities readily follow. The zeroth order term that appears in the right-hand side of (4.4) is a curvature term, and, as the proof in Appendix B shows, arise from the non-trivial holonomy along the closed curve cc.

5. Metric and geodesic completeness

We now want to prove the main result of this article, i.e., extend the completeness results, obtained for planar curves, to the situation studied in this article. The exact statement of the main results is now detailed in Theorems 5.15.3 below (the main result as presented in the introduction is a slightly simplified form of them).

Theorem 5.1.

Let n2n\geq 2, let D=[0,2π]D=[0,2\pi] or D=S1D=S^{1}, and let GG be a smooth Riemannian metric on n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}). Assume that for every metric ball B(c0,r)(n(D,𝒩),distG)B(c_{0},r)\in(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}), there exists a constant C=C(c0,r)>0C=C(c_{0},r)>0, such that for any cB(c0,r)c\in B(c_{0},r) and hTcn(D,𝒩)h\in T_{c}\mathcal{I}^{n}(D,\mathcal{N}) we have

(5.1) hGc\displaystyle\|h\|_{G_{c}} Cc1/2shL2(ds),\displaystyle\geq C\ell_{c}^{-1/2}\|\nabla_{\partial_{s}}h\|_{L^{2}(\,\mathrm{d}s)},
(5.2) hGc\displaystyle\|h\|_{G_{c}} CskhL\displaystyle\geq C\|\nabla_{\partial_{s}}^{k}h\|_{L^{\infty}} k=0,,n1,\displaystyle k=0,\ldots,n-1,
(5.3) hGc\displaystyle\|h\|_{G_{c}} CsnhL2(ds).\displaystyle\geq C\|\nabla_{\partial_{s}}^{n}h\|_{L^{2}(\,\mathrm{d}s)}.

Then GG is a strong metric, and we have:

  1. (1)

    (n(D,𝒩),distG)(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}) is a complete metric space.

  2. (2)

    (n(D,𝒩),G)(\mathcal{I}^{n}(D,\mathcal{N}),G) is geodesically complete

For Sobolev metrics of the type (1.2) we also obtain geodesic convexity:

Theorem 5.2.

Let D=[0,2π]D=[0,2\pi] or D=S1D=S^{1}, and let GG be a smooth Sobolev metric of the type (1.2) on n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}), that satisfies assumptions (5.1)–(5.3). Then any two immersions in the same connected component can be joined by a minimizing geodesic.

The reason that in Theorem 5.2 we further assume, unlike in Theorem 5.1, that GG is of the type (1.2) is merely a technical one; both theorems are first proved for metrics of this type, and in Theorem 5.1 the extension to the general case is immediate. Theorem 5.2, with the same method of proof, definitely holds for metrics that are not of type (1.2), but this needs to be checked on a case-by-case basis, and thus we present this theorem only for this type of metrics. The assumptions (5.1)–(5.3) are satisfied in the following cases:

Theorem 5.3.

Let D=[0,2π]D=[0,2\pi] or D=S1D=S^{1}, and let GG be a Sobolev metric of order n2n\geq 2 of the type (1.2) on n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}). Assume that one of the following holds:

  1. (1)

    Length weighted case: There exists α>0\alpha>0 such that either a1(x)αx1a_{1}(x)\geq\alpha x^{-1} or both a0(x)αx3a_{0}(x)\geq\alpha x^{-3} and ak(x)αx2k3a_{k}(x)\geq\alpha x^{2k-3} for some k>1k>1.

  2. (2)

    Constant coefficient case: D=S1D=S^{1} and both a0a_{0} and ana_{n} are positive constants.

Then assumptions (5.1)–(5.3) hold, and the completeness results of Theorem 5.1 hold for (n(D,𝒩),G)(\mathcal{I}^{n}(D,\mathcal{N}),G).

Remark 5.4.

Note that the family of scale-invariant Sobolev metric, as introduced in Section 1.2, satisfies conditions (1) of Theorem 5.3. In the article [16], where the authors study completeness properties for length weighted metrics on curves with values in Euclidean space, more general conditions on the coefficient functions that still ensure completeness have been derived. While such an analysis should be also possible in our situation, the resulting conditions would be much more complicated. The reason for this essentially lies in the fact that the manifold valued Sobolev estimates are more complicated (and involve lower-order terms), compared to the d\mathbb{R}^{d}-valued one, as described in Remark 4.3. Thus, for the sake of clarity, we discuss here only conditions of the type (1).

The remaining part of this section will contain the proof of these theorems. To prove Theorem 5.1 we will first show the metric completeness, which then implies the geodesic completeness, see [24, VIII, Proposition 6.5]. Since the theorem of Hopf-Rinow is not valid in infinite dimensions222Atkin constructed in [3] an example of a geodesically complete Riemannian manifold where the exponential map is not surjective, see also [21]. we cannot conclude the existence of geodesics by abstract arguments. Instead we show this statement by hand using the direct methods of the calculus of variations, in Section 5.5. Finally, in Section 5.6, we deduce geodesic completeness in the smooth category.

5.1. Reduction from metric completeness to equivalence of strong Riemannian metrics

In this section we reduce the question of metric completeness (n(D,𝒩),distG)(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}) to a question on uniform equivalence of the Riemannian metrics GG and \mathcal{H} on metric balls. This is done in two steps.

First reduction — distance equivalence on balls. The space n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) is an open subset of Hn(D,𝒩)H^{n}(D,\mathcal{N}) (see Proposition 2.2). In addition to the metric distG\operatorname{dist}^{G} induced by GG, it therefore inherits also the distance function dist\operatorname{dist}^{\mathcal{H}} induced from Hn(D,𝒩)H^{n}(D,\mathcal{N}). In general, distG\operatorname{dist}^{G} and dist\operatorname{dist}^{\mathcal{H}} are not equivalent.333This follows by the fact that \mathcal{H} and GG are no equivalent: For GG of the type (1.2), the highest order derivative it involves is sn\nabla_{\partial_{s}}^{n}, which equals to |c|nθn|c^{\prime}|^{-n}\nabla_{\partial_{\theta}}^{n} plus lower order terms. The highest order derivative in \mathcal{H} is, on the other hand, θn\nabla_{\partial_{\theta}}^{n}. In particular, if we take a curve cc on which |c||c^{\prime}| is very close to being zero in some interval, it follows that c\mathcal{H}_{c} and GcG_{c} can have extremely large ratio. However, we do have the following:

Proposition 5.5.

Assume that GG is a strong Riemannian metric on n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) and that the following holds:

  1. (1)

    For every metric ball B(c0,r)(n(D,𝒩),distG)B(c_{0},r)\subset\left(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}\right), there exists a constant C>0C>0 such that distCdistG\operatorname{dist}^{\mathcal{H}}\leq C\operatorname{dist}^{G} on B(c0,r)B(c_{0},r).

  2. (2)

    For every metric ball B(c0,r)(n(D,𝒩),distG)B(c_{0},r)\subset\left(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}\right), |c|1L\|{|c^{\prime}|}^{-1}\|_{L^{\infty}} is bounded.

Then (n(D,𝒩),distG)(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}) is metrically complete.

Proof.

The proof below is similar to the proof of [12, Theorem 4.3]. For the convenience of the reader we repeat the arguments here. Given a Cauchy sequence (cn)(c_{n}) in (n(D,𝒩),distG)\left(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}\right), the sequence remains in a bounded metric ball in (n(D,𝒩),distG)\left(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}\right), hence by (1) the sequence is also a Cauchy sequence in Hn(D,𝒩)H^{n}(D,\mathcal{N}), hence cncHn(D,𝒩)c_{n}\to c\in H^{n}(D,\mathcal{N}) (modulo a subsequence). Moreover, since the sequence cnc_{n} lies in a metric ball, |cn|1<C<|c_{n}^{\prime}|^{-1}<C<\infty for all nn by (2), and since HnH^{n}-convergence implies C1C^{1}-convergence, we obtain that |c|1C|{c^{\prime}}|^{-1}\leq C, and thus cn(D,𝒩)c\in\mathcal{I}^{n}(D,\mathcal{N}). Since both \mathcal{H} and GG are strong metrics on n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}), they induce the same topology (the manifold topology) [24, VII, Proposition 6.1], and thus dist(cn,c)0\operatorname{dist}^{\mathcal{H}}(c_{n},c)\to 0 implies that distG(cn,c)0\operatorname{dist}^{G}(c_{n},c)\to 0, hence (n(D,𝒩),distG)\left(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}\right) is metrically complete. ∎

Second reduction — metric equivalence implies distance equivalence.

Next, we show that distance-equivalence on metric balls follows from metric-equivalence on metric balls. The following proposition is the content of Proposition 3.5 and Remark 3.6 in [12], adapted to our setting.

Proposition 5.6.

Assume that for each metric ball

B(c0,r)(n(D,𝒩),distG),B(c_{0},r)\subset\left(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}\right),

there exists C=C(c0,r)>0C=C(c_{0},r)>0 such that for every cB(c0,r)c\in B(c_{0},r) and hHn(D;cT𝒩)h\in H^{n}(D;c^{*}T\mathcal{N}), we have

(5.4) hcChGc.\|h\|_{\mathcal{H}_{c}}\leq C\|h\|_{G_{c}}.

Then, property (1) in Proposition 5.5 holds.

Proof.

The following proof is an adaptation of the proof of [12, Lemma 4.2]. Let c1,c2B(c0,r)c_{1},c_{2}\in B(c_{0},r) and ε>0\varepsilon>0, and let γ\gamma be a piecewise smooth curve between c1c_{1} and c2c_{2} with LG(γ)distG(c1,c2)+εL^{G}(\gamma)\leq\operatorname{dist}^{G}(c_{1},c_{2})+\varepsilon. Since distG(c1,c2)<2r\operatorname{dist}^{G}(c_{1},c_{2})<2r, by the triangle inequality, we have that γB(c0,3r)\gamma\subset B(c_{0},3r). We then have, using assumption (5.4) for B(c0,3r)B(c_{0},3r), that

dist(c1,c2)L(γ)CLG(γ)C(distG(c1,c2)+ε).\operatorname{dist}^{\mathcal{H}}(c_{1},c_{2})\leq L^{\mathcal{H}}(\gamma)\leq CL^{G}(\gamma)\leq C(\operatorname{dist}^{G}(c_{1},c_{2})+\varepsilon).

Since ε\varepsilon is arbitrary, completes the proof. ∎

5.2. Estimates on c\ell_{c} and |c||c^{\prime}| in metric balls

In this section we bound various quantities that depend on the curve cc uniformly on metric balls in n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}). These will enable us to prove the assumption of Proposition 5.6, as well as assumption (2) of Proposition 5.5.

To this end, we will repeatedly use the following result (see [12, Lemma 3.2] for a proof444In fact, for the case in which CC is independent of the metric ball, the statement in [12, Lemma 3.2] is inaccurate; the statement of Lemma 5.7 is the corrected one, and the proof follows exactly as in [12, Lemma 3.2], by checking carefully which constants appear when using Gronwall’s inequality [12, Corollary 2.7].):

Lemma 5.7.

Let (,𝔤)(\mathcal{M},\mathfrak{g}) be a Riemannian manifold, possibly of infinite dimension, and let FF be a normed space. Let f:Ff:\mathcal{M}\to F be a C1C^{1}-function, such that for each metric ball B(y,r)B(y,r) in \mathcal{M} there exists a constant CC, such that

Txf.vFC(1+f(x)F)vxfor all xB(y,r),vTx.\|T_{x}f.v\|_{F}\leq C(1+\|f(x)\|_{F})\|v\|_{x}\qquad\text{for all }\,\,x\in B(y,r),\,\,v\in T_{x}\mathcal{M}.

Then ff is Lipschitz continuous on every metric ball, and in particular bounded on every metric ball. Moreover, if the constant CC is independent of the metric ball B(y,r)B(y,r), then the Lipschitz constant in B(y,r)B(y,r) can be bounded by a function L:[0,)3(0,)L:[0,\infty)^{3}\to(0,\infty), increasing in all variables, as follows:

f(x1)f(x2)FL(C,f(y)F,r)dist(x1,x2) for every x1,x2B(y,r).\|f(x_{1})-f(x_{2})\|_{F}\leq L(C,\|f(y)\|_{F},r)\operatorname{dist}(x_{1},x_{2})\quad\text{ for every }x_{1},x_{2}\in B(y,r).

In particular the Lipschitz constant in B(y,r)B(y,r) depends on yy only through f(y)F\|f(y)\|_{F}.

Remark 5.8.

Tracking the constants in Lemma 5.7 carefully, one can obtain the bound

(5.5) L(C,t,r)=C2(1+r)(1+2r)e2Cr(1+t).L(C,t,r)=C^{2}(1+r)(1+2r)e^{2Cr}(1+t).

Note that this is not sharp, it is simply what is obtained by the method of the proof (using Gronwall’s inequality).

Lemma 5.9 (Bounds on length).

Assume that assumption (5.1) holds. Then the length function ccc\mapsto\ell_{c} is bounded from above and away from zero on every metric ball.

Proof.

From (3.8) we have that

|Dc,hc|D|g(v,sh)|dsc1/2(D|g(v,sh)|2ds)1/2c1/2shL2(ds).|D_{c,h}\ell_{c}|\leq\int_{D}|g(v,\nabla_{\partial_{s}}h)|\,\,\mathrm{d}s\leq\ell_{c}^{1/2}\left(\int_{D}|g(v,\nabla_{\partial_{s}}h)|^{2}\,\,\mathrm{d}s\right)^{1/2}\leq\\ \leq\ell_{c}^{1/2}\|\nabla_{\partial_{s}}h\|_{L^{2}(\,\mathrm{d}s)}.

Therefore, under the assumption (5.1), we have

|Dc,hc|hGc|D_{c,h}\ell_{c}|\lesssim\|h\|_{G_{c}}

and we obtain from Lemma 5.7 that ccc\mapsto\ell_{c} is bounded on every metric ball.

Similarly, for the map cc1c\mapsto\ell_{c}^{-1}, we have, under the assumption (5.1), that

|Dc,hc1|=c2|Dc,hc|c3/2shL2(ds)c1hGc,|D_{c,h}\ell_{c}^{-1}|=\ell_{c}^{-2}|D_{c,h}\ell_{c}|\leq\ell_{c}^{-3/2}\|\nabla_{\partial_{s}}h\|_{L^{2}(\,\mathrm{d}s)}\lesssim\ell_{c}^{-1}\|h\|_{G_{c}},

which concludes the proof using again Lemma 5.7. ∎

Lemma 5.10 (Bounds on speed).

Assume that assumption (5.2) holds for k=1k=1. Then, there exists a constant α=α(c0,r)>0\alpha=\alpha(c_{0},r)>0 such that

α1|c(θ)|α\alpha^{-1}\leq|c^{\prime}(\theta)|\leq\alpha

for every cB(c0,r)c\in B(c_{0},r) and θD\theta\in D.

Proof.

Consider the function

log|c|:(n(D,𝒩),G)L(D;).\log|c^{\prime}|\,:\,(\mathcal{I}^{n}(D,\mathcal{N}),G)\to L^{\infty}(D;\mathbb{R}).

By (3.6) and assumption (5.2) we have

Dc,hlog|c|Lg(v,sh)LshLhGc.\|D_{c,h}\log|c^{\prime}|\|_{L^{\infty}}\leq\|g(v,\nabla_{\partial_{s}}h)\|_{L^{\infty}}\leq\|\nabla_{\partial_{s}}h\|_{L^{\infty}}\lesssim\|h\|_{G_{c}}.

By Lemma 5.7 we thus have that log|c|\log|c^{\prime}| is bounded on metric balls, from which the claim follows. ∎

Lemma 5.11.

Assume that assumption (5.2) holds for k=0k=0. Then the image in 𝒩\mathcal{N} of every metric ball B(c0,r)B(c_{0},r) is bounded. That is, there exists R=R(c0,r)>0R=R(c_{0},r)>0 such that for every cB(c0,r)c\in B(c_{0},r) and every θD\theta\in D,

dist𝒩(c(θ),c0(0))<R.\operatorname{dist}_{\mathcal{N}}(c(\theta),c_{0}(0))<R.
Proof.

Let cB(c0,r)c\in B(c_{0},r), and let c(t,θ):[0,1]B(c0,r)c(t,\theta):[0,1]\to B(c_{0},r) be a path from c0=c(0,)c_{0}=c(0,\cdot) to c=c(1,)c=c(1,\cdot), whose length is smaller than rr. Using (5.2), we have

dist𝒩(c(θ),c0(θ))01|tc(t,θ)|dtC01tc(t,θ)Gcdt<Cr.\operatorname{dist}_{\mathcal{N}}(c(\theta),c_{0}(\theta))\leq\int_{0}^{1}|\partial_{t}c(t,\theta)|\,\,\mathrm{d}t\leq C\int_{0}^{1}\|\partial_{t}c(t,\theta)\|_{G_{c}}\,\,\mathrm{d}t<Cr.

This completes the proof, as the length of c0c_{0} is finite. ∎

Lemma 5.12.

Assume that assumptions (5.1)–(5.3) hold. Then the following quantities are bounded on every metric ball

(5.6) sk|c|L\displaystyle\|\nabla_{\partial_{s}}^{k}|c^{\prime}|\|_{L^{\infty}} k=0,,n2,\displaystyle k=0,\ldots,n-2,
(5.7) sk|c|L2\displaystyle\|\nabla_{\partial_{s}}^{k}|c^{\prime}|\|_{L^{2}} k=0,,n1,\displaystyle k=0,\ldots,n-1,

where L2L^{2} is with respect to either ds\,\mathrm{d}s or dθ\,\mathrm{d}\theta.

Proof.

The proof of this is result follows by an induction on kk using iteratively Lemma 5.9 and 5.10. It is mainly an adaptation of Lemma 3.3 and Proposition 3.4 in [12], though the calculations in our situation are more involved due to the appearance of curvature terms of the manifold 𝒩\mathcal{N}. To keep the presentation simple we postpone it to the Appendix C. ∎

5.3. Proof of Theorem 5.1: metric and geodesic completeness

We are now able to prove Theorem 5.1, that is, that (n(D,𝒩),G)(\mathcal{I}^{n}(D,\mathcal{N}),G) is metrically and geodesically complete. We first prove it for a metric GG of the type (1.2) of order nn that satisfies assumptions (5.1)–(5.3). Afterwards the assumption that GG is of the type (1.2) will be removed.

In particular, GG satisfies (5.2), and therefore Lemma 5.10 implies that assumption (2) in Proposition 5.5 holds. Therefore, in order to prove that (n(D,𝒩),distG)(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}) is metrically complete, we need to show that GG is a strong metric, and prove property (1), which by Proposition 5.6 follows from (5.4). In fact, we will show a stronger result and prove that GG and \mathcal{H} are equivalent uniformly on metric balls. This will also imply that GG is a strong metric.

From Lemma 4.1, we have

i=0nθihL2(dθ)2Ch2\sum_{i=0}^{n}\|\nabla_{\partial_{\theta}}^{i}h\|_{L^{2}(\,\mathrm{d}\theta)}^{2}\leq C\|h\|_{\mathcal{H}}^{2}

for some universal constant C>0C>0. Similarly, we have

(5.8) θhLCh2\|\nabla_{\partial_{\theta}}h\|_{L^{\infty}}\leq C^{\prime}\|h\|_{\mathcal{H}}^{2}

for some universal constant C>0C^{\prime}>0.

From the definition of s\nabla_{\partial_{s}} we have, by using the Leibniz rule,

skh=1|c|kθkh+i=1k1Pi,kθih,\nabla_{\partial_{s}}^{k}h=\frac{1}{|c^{\prime}|^{k}}\nabla_{\partial_{\theta}}^{k}h+\sum_{i=1}^{k-1}P_{i,k}\nabla_{\partial_{\theta}}^{i}h,

where Pi,kP_{i,k} are polynomials in |c|,s|c|,,ski|c||c^{\prime}|,\nabla_{\partial_{s}}|c^{\prime}|,\ldots,\nabla_{\partial_{s}}^{k-i}|c^{\prime}| and |c|1,,|c|k|c^{\prime}|^{-1},\ldots,|c^{\prime}|^{-k}, which are linear in ski|c|\nabla_{\partial_{s}}^{k-i}|c^{\prime}|. Similarly,

θkh=|c|kskh+i=1k1Qi,ksih\nabla_{\partial_{\theta}}^{k}h=|c^{\prime}|^{k}\nabla_{\partial_{s}}^{k}h+\sum_{i=1}^{k-1}Q_{i,k}\nabla_{\partial_{s}}^{i}h

where Qi,kQ_{i,k} is a polynomial in the variables |c|,s|c|,,ski|c||c^{\prime}|,\nabla_{\partial_{s}}|c^{\prime}|,\ldots,\nabla_{\partial_{s}}^{k-i}|c^{\prime}| and the variables |c|,,|c|k1|c^{\prime}|,\ldots,|c^{\prime}|^{k-1}, which are linear in ski|c|\nabla_{\partial_{s}}^{k-i}|c^{\prime}|. Using Lemma 5.10 and Lemma 5.12, we therefore have that for k<nk<n, Pi,kP_{i,k} and Qi,kQ_{i,k} are uniformly bounded on any metric ball, and so are |c|±1|c^{\prime}|^{\pm 1}, hence

|skh|i=1k|θih|,|θkh|i=1k|kih|,|\nabla_{\partial_{s}}^{k}h|\lesssim\sum_{i=1}^{k}|\nabla_{\partial\theta}^{i}h|,\quad|\nabla_{\partial\theta}^{k}h|\lesssim\sum_{i=1}^{k}|\nabla_{k}^{i}h|,

uniformly on every metric ball. The bound on |c|±1|c^{\prime}|^{\pm 1} also implies that integration with respect to ds\,\mathrm{d}s or dθ\,\mathrm{d}\theta are equivalent, hence

(5.9) skhL2(ds)i=1kθihL2(dθ),θkhL2(dθ)i=1kkihL2(ds),\|\nabla_{\partial_{s}}^{k}h\|_{L^{2}(\,\mathrm{d}s)}\lesssim\sum_{i=1}^{k}\|\nabla_{\partial\theta}^{i}h\|_{L^{2}(\,\mathrm{d}\theta)},\quad\|\nabla_{\partial\theta}^{k}h\|_{L^{2}(\,\mathrm{d}\theta)}\lesssim\sum_{i=1}^{k}\|\nabla_{k}^{i}h\|_{L^{2}(\,\mathrm{d}s)},

uniformly on every metric ball.

For k=nk=n, we have, uniformly on every metric ball,

|snh||sn1|c|||θh|+i=2n|θih||\nabla_{\partial_{s}}^{n}h|\lesssim\left|\nabla_{\partial_{s}}^{n-1}|c^{\prime}|\right||\nabla_{\partial_{\theta}}h|+\sum_{i=2}^{n}|\nabla_{\partial_{\theta}}^{i}h|

and

|θnh||sn1|c|||sh|+i=2n|sih|,|\nabla_{\partial\theta}^{n}h|\lesssim\left|\nabla_{\partial_{s}}^{n-1}|c^{\prime}|\right||\nabla_{\partial_{s}}h|+\sum_{i=2}^{n}|\nabla_{\partial_{s}}^{i}h|,

and therefore, invoking Lemma 5.12 again and using (5.8), we have,

(5.10) snhL2(ds)θhL+i=2nθihL2(dθ)Ch\|\nabla_{\partial_{s}}^{n}h\|_{L^{2}(\,\mathrm{d}s)}\lesssim\|\nabla_{\partial_{\theta}}h\|_{L^{\infty}}+\sum_{i=2}^{n}\|\nabla_{\partial_{\theta}}^{i}h\|_{L^{2}(\,\mathrm{d}\theta)}\lesssim C\|h\|_{\mathcal{H}}

and, using (5.2) again,

(5.11) θnhL2(dθ)shL+i=2nsihL2(ds)hGc+i=2nsihL2(ds).\|\nabla_{\partial\theta}^{n}h\|_{L^{2}(\,\mathrm{d}\theta)}\lesssim\|\nabla_{\partial_{s}}h\|_{L^{\infty}}+\sum_{i=2}^{n}\|\nabla_{\partial_{s}}^{i}h\|_{L^{2}(\,\mathrm{d}s)}\lesssim\|h\|_{G_{c}}+\sum_{i=2}^{n}\|\nabla_{\partial_{s}}^{i}h\|_{L^{2}(\,\mathrm{d}s)}.

Since (5.1) holds, we have by Lemma 5.9 that c\ell_{c} is uniformly bounded from above and below on metric balls, hence all the coefficient functions ai(c)0a_{i}(\ell_{c})\geq 0 are bounded from above on metric balls, and a0,ana_{0},a_{n} are also bounded away from zero. We therefore have that, on each metric ball

hL2(ds)+snhL2(ds)hGci=0nsihL2(ds).\|h\|_{L^{2}(\,\mathrm{d}s)}+\|\nabla_{\partial_{s}}^{n}h\|_{L^{2}(\,\mathrm{d}s)}\lesssim\|h\|_{G_{c}}\lesssim\sum_{i=0}^{n}\|\nabla_{\partial_{s}}^{i}h\|_{L^{2}(\,\mathrm{d}s)}.

Since c\ell_{c} is bounded from below and above uniformly on metric balls, Lemma 4.2 enables us to improve that to

i=0nsihL2(ds)hGci=0nsihL2(ds)\sum_{i=0}^{n}\|\nabla_{\partial_{s}}^{i}h\|_{L^{2}(\,\mathrm{d}s)}\lesssim\|h\|_{G_{c}}\lesssim\sum_{i=0}^{n}\|\nabla_{\partial_{s}}^{i}h\|_{L^{2}(\,\mathrm{d}s)}

Combining this with the estimate (5.9), (5.10) and (5.11) immediately imply

hchGhc,\|h\|_{\mathcal{H}_{c}}\lesssim\|h\|_{G}\lesssim\|h\|_{\mathcal{H}_{c}},

uniformly on metric balls. In particular, this implies (5.4) and show that GG is a strong metric, thus all the assumptions of Propositions 5.55.6 are satisfied, which completes the proof of metric completeness. As stated before, geodesic completeness follows directly as for strong Riemannian metrics (in infinite dimensions) metric completeness still implies geodesic completeness, see, e.g., [24, VIII, Proposition 6.5].

We now remove the assumption that GG is of the type (1.2), and only assume that it is a smooth metric that satisfies (5.1)–(5.3). Denote by G~\tilde{G} the metric

hG~c2:=hL2(ds)2+c1shL2(ds)2+snhL2(ds).\|h\|_{\tilde{G}_{c}}^{2}:=\|h\|_{L^{2}(\,\mathrm{d}s)}^{2}+\ell_{c}^{-1}\|\nabla_{\partial_{s}}h\|_{L^{2}(\,\mathrm{d}s)}^{2}+\|\nabla_{\partial_{s}}^{n}h\|_{L^{2}(\,\mathrm{d}s)}.

This metric is of the type (1.2), and in Section 5.4 below we show that this metric indeed satisfies (5.1)–(5.3). Therefore, it is metrically complete.

Now assume that GG is another metric that satisfies (5.1)–(5.3). We claim that on every metric ball BG(c0,r)B^{G}(c_{0},r), there exists a constant C=C(c0,r)C=C(c_{0},r) such that G~cCGc\|\cdot\|_{\tilde{G}_{c}}\leq C\|\cdot\|_{G_{c}}. Indeed, assumptions (5.1) and (5.3) imply that GG controls the second and third addends in the definition on G~\tilde{G}; since hLc1/2hL2(ds)\|h\|_{L^{\infty}}\geq\ell_{c}^{-1/2}\|h\|_{L^{2}(\,\mathrm{d}s)}, assumption (5.2) for k=0k=0 and Lemma 5.9 imply that GG controls the second addend in G~\tilde{G} as well (uniformly on every metric ball). This implies, in particular, that GG is a strong metric (since G~\tilde{G} is).

The proof is now concluded by similar arguments as Section 5.1 (with G~\tilde{G} instead of \mathcal{H}): Let ck(n(D,𝒩),distG)c_{k}\in(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}) be a Cauchy sequence. It follows that ckc_{k} is also a Cauchy sequence in (n(D,𝒩),distG~)(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{\tilde{G}}). Since (n(D,𝒩),distG~)(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{\tilde{G}}) is metrically complete, ckc_{k} converges in (n(D,𝒩),distG~)(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{\tilde{G}}) to some limit cn(D,𝒩)c\in\mathcal{I}^{n}(D,\mathcal{N}). Since both GG and G~\tilde{G} are strong metrics on n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}), they induce the same topology. Therefore, ckcc_{k}\to c in (n(D,𝒩),distG)(\mathcal{I}^{n}(D,\mathcal{N}),\operatorname{dist}^{G}) as well, thus proving metric completeness, from which geodesic completeness follows as before.

5.4. Proof of Theorem 5.3

Length weighted case. If both a0(x)αx3a_{0}(x)\geq\alpha x^{-3} and ak(x)αx2k3a_{k}(x)\geq\alpha x^{2k-3} for some k>1k>1, then by (4.2) we have that

c1shL2(ds)2ChG2\ell_{c}^{-1}\|\nabla_{\partial_{s}}h\|_{L^{2}(\,\mathrm{d}s)}^{2}\leq C\|h\|_{G}^{2}

for some C>0C>0. This is also obviously true if a1(x)αx1a_{1}(x)\geq\alpha x^{-1}. Thus (5.1) holds, and from Lemma 5.9 we obtain that the length function ccc\mapsto\ell_{c} is bounded from above and away from zero on any metric ball. Since GG is of the type (1.2), we have that

hGc2a0(c)hL2(ds)2+an(c)snhL2(ds)2,\|h\|_{G_{c}}^{2}\geq a_{0}(\ell_{c})\|h\|_{L^{2}(\,\mathrm{d}s)}^{2}+a_{n}(\ell_{c})\|\nabla_{\partial_{s}}^{n}h\|_{L^{2}(\,\mathrm{d}s)}^{2},

and the bound on the length implies that on each metric ball, the constants a0(c)a_{0}(\ell_{c}) and an(c)a_{n}(\ell_{c}) are bounded away from zero. This immediately implies (5.3), and also that on every metric ball

hGc2C(hL2(ds)2+snhL2(ds)2),\|h\|_{G_{c}}^{2}\geq C(\|h\|_{L^{2}(\,\mathrm{d}s)}^{2}+\|\nabla_{\partial_{s}}^{n}h\|_{L^{2}(\,\mathrm{d}s)}^{2}),

for some C>0C>0. On the other hand, using (4.3) with a=ca=\ell_{c} we have, for every k=0,,n1k=0,\ldots,n-1,

skhL2C(c2k1hL2(ds)2+c2(nk)1snhL2(ds)2),\|\nabla_{\partial_{s}}^{k}h\|^{2}_{L^{\infty}}\leq C\left(\ell_{c}^{-2k-1}\|h\|^{2}_{L^{2}(\,\mathrm{d}s)}+\ell_{c}^{2(n-k)-1}\|\nabla_{\partial_{s}}^{n}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\right),

hence on each metric ball, we have

skhL2C(hL2(ds)2+snhL2(ds)2)C′′hGc2,\|\nabla_{\partial_{s}}^{k}h\|^{2}_{L^{\infty}}\leq C^{\prime}\left(\|h\|^{2}_{L^{2}(\,\mathrm{d}s)}+\|\nabla_{\partial_{s}}^{n}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\right)\leq C^{\prime\prime}\|h\|_{G_{c}}^{2},

which implies (5.2).

Constant coefficient case. Assume that a0a_{0} and ana_{n} are positive constants. We then immediately have (5.3). Furthermore, using (4.4) for k=1k=1, we have

shL2(ds)2Cc2hGc2\|\nabla_{\partial_{s}}h\|_{L^{2}(\,\mathrm{d}s)}^{2}\leq C\ell_{c}^{2}\|h\|_{G_{c}}^{2}

for some constant CC that is independent of the curve cc.555This is the crucial point in which the improved estimates for closed curves in Lemma 4.2 are needed. This implies (5.1), and hence the boundedness of ccc\mapsto\ell_{c} by Lemma 5.9. The proof of (5.2) now follows in the same manner as the length weighted case.

5.5. Proof of Theorem 5.2: existence of minimizing geodesics

We now prove that any two immersions in the same connected component can be joined by a minimizing geodesic. The approach is a variational one: we consider the energy

E(c):=01Gc(c˙,c˙)dt,E(c):=\int_{0}^{1}G_{c}(\dot{c},\dot{c})\,\,\mathrm{d}t,

defined on the set

Ac0,c1:={c:[0,1]n(D,𝒩):c˙L2((0,1);Hn(D;cTN)),c(0)=c0,c(1)=c1},A_{c_{0},c_{1}}:=\Big{\{}c:[0,1]\to\mathcal{I}^{n}(D,\mathcal{N}):\dot{c}\in L^{2}((0,1);H^{n}(D;c^{*}TN)),\\ \qquad c(0)=c_{0},c(1)=c_{1}\Big{\}},

where c0,c1n(D,𝒩)c_{0},c_{1}\in\mathcal{I}^{n}(D,\mathcal{N}) are two immersions in the same connected component (thus Ac0,c1A_{c_{0},c_{1}} is a non-empty set). We aim to show that there exists a minimizer to EE over Ac0,c1A_{c_{0},c_{1}}, which is, by definition, a minimizing geodesic.

We prove the existence of minimizers using the direct methods in the calculus of variations; namely, we take a minimizing sequence cjc^{j}, prove that it is weakly sequentially precompact, and that any limit point must be a minimizer. In order to use weak convergence, we embed the curves in a Hilbert space, which neither n(D,𝒩)\mathcal{I}^{n}(D,\mathcal{N}) or Hn(D,𝒩)H^{n}(D,\mathcal{N}) are (this is the point where 𝒩\mathcal{N}-valued curves differ from d\mathbb{R}^{d}-valued curves treated in [12, Theorem 5.2]). To this end, we again isometrically embed 𝒩\mathcal{N} into m\mathbb{R}^{m} for some large enough m𝒩m\in\mathcal{N}, as in the definition of Hn(D,𝒩)H^{n}(D,\mathcal{N}) that we started with (Definition 2.1). This will require us, as in Section 2, to use Lemma 2.3 to relate the metric \mathcal{H} on Hn(D,𝒩)H^{n}(D,\mathcal{N}) with the standard Sobolev norm on Hn(D;m)H^{n}(D;\mathbb{R}^{m}).

Let now cjAc0,c1c^{j}\in A_{c_{0},c_{1}} be a minimizing sequence of EE, that is,

E(cj)infAc0,c1E.E(c^{j})\to\inf_{A_{c_{0},c_{1}}}E.

In particular, E(cj)E(c^{j}) is a bounded sequence. Denote by R2R^{2} its supremum. We also fix an isometric embedding ι:𝒩m\iota:\mathcal{N}\to\mathbb{R}^{m}, and, using this embedding, we consider cjc^{j} as elements of the Hilbert space H1([0,1];Hn(D;m))H^{1}([0,1];H^{n}(D;\mathbb{R}^{m})).

Step I: The family (cj(t))j,t[0,1](c^{j}(t))_{j\in\mathbb{N},t\in[0,1]} lies in a bounded ball around c0c_{0}. Fix t0[0,1]t_{0}\in[0,1] and jj\in\mathbb{N}. Since cj:[0,t0]n(D,𝒩)c^{j}:[0,t_{0}]\to\mathcal{I}^{n}(D,\mathcal{N}) is a path from c0c_{0} to cj(t0)c^{j}(t_{0}), we have

distG2(cj(t0),c0)(0t0c˙j(t)Gcj(t)dt)201c˙j(t)Gcj(t)2dt=E(cj)R2.\operatorname{dist}_{G}^{2}(c^{j}(t_{0}),c_{0})\leq\left(\int_{0}^{t_{0}}\|\dot{c}^{j}(t)\|_{G_{c^{j}(t)}}\,\mathrm{d}t\right)^{2}\leq\int_{0}^{1}\|\dot{c}^{j}(t)\|_{G_{c^{j}(t)}}^{2}\,\mathrm{d}t=E(c^{j})\leq R^{2}.

Therefore, (cj(t))j,t[0,1]B(c0,R)(c^{j}(t))_{j\in\mathbb{N},t\in[0,1]}\subset B(c_{0},R), where the ball is with respect to the metric GG.

Step II: The family (cj)j(c^{j})_{j\in\mathbb{N}} is a bounded set in H1([0,1];Hn(D;m))H^{1}([0,1];H^{n}(D;\mathbb{R}^{m})). Since GG satisfies (5.1)–(5.3), we have that (5.4) hold uniformly on B(c0,R)B(c_{0},R), that is, there exists C>0C>0 such that

C1hchGcChc, for all cB(c0,R),hHn(D;cT𝒩).C^{-1}\|h\|_{\mathcal{H}_{c}}\leq\|h\|_{G_{c}}\leq C\|h\|_{\mathcal{H}_{c}},\quad\text{ for all }c\in B(c_{0},R),\,\,h\in H^{n}(D;c^{*}T\mathcal{N}).

This was proved in Section 5.3. Moreover, from Lemmata 5.115.12, we have that the assumptions of Lemma 2.3 hold uniformly on B(c0,R)B(c_{0},R), hence, combining with the above inequality, we obtain that there exists C>0C>0 such that

C1hHn(ι)hGcChHn(ι), for all cB(c0,R),hHn(D;cT𝒩).C^{-1}\|h\|_{H^{n}(\iota)}\leq\|h\|_{G_{c}}\leq C\|h\|_{H^{n}(\iota)},\quad\text{ for all }c\in B(c_{0},R),\,\,h\in H^{n}(D;c^{*}T\mathcal{N}).

Since (cj(t))j,t[0,1]B(c0,R)(c^{j}(t))_{j\in\mathbb{N},t\in[0,1]}\subset B(c_{0},R), we obtain that for any fixed t0t_{0} and jj,

c0cj(t0)Hn(ι)2=D|c0cj(t0)|2+|θn(c0cj(t0))|2dθ=D|0t0c˙j(t)dt|2+|0t0θnc˙j(t0)dt|2dθD01|c˙j(t)|2+|θnc˙j(t0)|2dtdθ=01c˙j(t)Hn(ι)2dt=C01c˙j(t)Gcj(t)2dtCR2.\begin{split}\|c_{0}-c^{j}(t_{0})\|_{H^{n}(\iota)}^{2}&=\int_{D}|c_{0}-c^{j}(t_{0})|^{2}+|\partial_{\theta}^{n}(c_{0}-c^{j}(t_{0}))|^{2}\,\,\mathrm{d}\theta\\ &=\int_{D}\left|\int_{0}^{t_{0}}\dot{c}^{j}(t)\,\,\mathrm{d}t\right|^{2}+\left|\int_{0}^{t_{0}}\partial_{\theta}^{n}\dot{c}^{j}(t_{0})\,\,\mathrm{d}t\right|^{2}\,\,\mathrm{d}\theta\\ &\leq\int_{D}\int_{0}^{1}|\dot{c}^{j}(t)|^{2}+|\partial_{\theta}^{n}\dot{c}^{j}(t_{0})|^{2}\,\,\mathrm{d}t\,\,\mathrm{d}\theta\\ &=\int_{0}^{1}\|\dot{c}^{j}(t)\|_{H^{n}(\iota)}^{2}\,\,\mathrm{d}t\\ &=C\int_{0}^{1}\|\dot{c}^{j}(t)\|_{G_{c^{j}(t)}}^{2}\,\,\mathrm{d}t\leq CR^{2}.\end{split}

Therefore

cjH1([0,1];Hn(D;m))2=01cj(t)Hn(ι)2+c˙j(t)Hn(ι)2dt012c0Hn(ι)2+2c0cj(t0)Hn(ι)2dt+01c˙j(t)Hn(ι)2dt012c0Hn(ι)2+2CR2dt+C01c˙j(t)Gcj(t)2dt3CR2+2c0Hn(ι)2\begin{split}\|c^{j}&\|_{H^{1}([0,1];H^{n}(D;\mathbb{R}^{m}))}^{2}=\int_{0}^{1}\|c^{j}(t)\|_{H^{n}(\iota)}^{2}+\|\dot{c}^{j}(t)\|_{H^{n}(\iota)}^{2}\,\,\mathrm{d}t\\ &\leq\int_{0}^{1}2\|c_{0}\|_{H^{n}(\iota)}^{2}+2\|c_{0}-c^{j}(t_{0})\|_{H^{n}(\iota)}^{2}\,\,\mathrm{d}t+\int_{0}^{1}\|\dot{c}^{j}(t)\|_{H^{n}(\iota)}^{2}\,\,\mathrm{d}t\\ &\leq\int_{0}^{1}2\|c_{0}\|_{H^{n}(\iota)}^{2}+2CR^{2}\,\,\mathrm{d}t+C\int_{0}^{1}\|\dot{c}^{j}(t)\|_{G_{c^{j}(t)}}^{2}\,\,\mathrm{d}t\\ &\leq 3CR^{2}+2\|c_{0}\|_{H^{n}(\iota)}^{2}\end{split}

Hence, the sequence cjc^{j} is bounded in the Hilbert space H1([0,1];Hn(D;m))H^{1}([0,1];H^{n}(D;\mathbb{R}^{m})). Therefore, it has a subsequence (not relabeled) that weakly converges to some cH1([0,1];Hn(D;m))c^{*}\in H^{1}([0,1];H^{n}(D;\mathbb{R}^{m})).

Step III: The limit point cc^{*} belongs to Ac0,c1A_{c_{0},c_{1}}. Let ε(0,1/2)\varepsilon\in(0,1/2). We then have that the embedding H1([0,1];Hn(D;m))C([0,1];Hnε(D;m))H^{1}([0,1];H^{n}(D;\mathbb{R}^{m}))\subset C([0,1];H^{n-\varepsilon}(D;\mathbb{R}^{m})) is compact (due to the Aubin–Lions–Simon lemma666See, e.g., [11, Theorem II.5.16]. With respect to the notation there we use the lemma for p=p=\infty, r=2r=2, B0=HnB_{0}=H^{n}, B1=HnεB_{1}=H^{n-\varepsilon} and B2=Hn1B_{2}=H^{n-1}. We can use p=p=\infty because H1H^{1} embeds in LL^{\infty}.) and Hnε(D;m)H^{n-\varepsilon}(D;\mathbb{R}^{m}) is compactly embedded in Cn1(D;m)C^{n-1}(D;\mathbb{R}^{m}). In particular, we thus have that cjcc^{j}\to c^{*} in the strong topology of C([0,1];Cn1(D;m))C([0,1];C^{n-1}(D;\mathbb{R}^{m})). Since cj(θ)𝒩c^{j}(\theta)\in\mathcal{N} for all jj and θ\theta, the uniform convergence implies that c(θ)𝒩c^{*}(\theta)\in\mathcal{N} for all θ\theta as well. Since cj(0)=c0c^{j}(0)=c_{0} and cj(1)=c1c^{j}(1)=c_{1} for all jj, the same holds for cc^{*}. Finally, since cj(t)B(c0,R)c^{j}(t)\in B(c_{0},R) for every jj and tt, Lemma 5.10 implies that

|θcj(t,θ)|>α|\partial_{\theta}c^{j}(t,\theta)|>\alpha

for some α>0\alpha>0. Since cjcc^{j}\to c^{*} in C([0,1];Cn1(D;m))C([0,1];C^{n-1}(D;\mathbb{R}^{m})), the same holds for cc^{*}, hence cn(D,𝒩)c^{*}\in\mathcal{I}^{n}(D,\mathcal{N}). This shows that indeed cAc0,c1c^{*}\in A_{c_{0},c_{1}}.

Step IV: Weak convergence of derivatives. It will be helpful now to emphasize the particular curve that is used to define the s\nabla_{\partial_{s}} derivative. Therefore, for the rest of this proof, denote Dcj:=|cj|1θ𝒩D_{c^{j}}:=|c^{j}|^{-1}\nabla^{\mathcal{N}}_{\partial\theta}. We now show that, for k=0,,nk=0,\ldots,n, we have

(5.12) Dcjnc˙jDcnc˙ in L2([0,1];L2(D;m)).D_{c^{j}}^{n}\dot{c}^{j}\rightharpoonup D_{c^{*}}^{n}\dot{c}^{*}\quad\text{ in }\,L^{2}([0,1];L^{2}(D;\mathbb{R}^{m})).

By the definition of cc^{*}, we have that

c˙jc˙ in L2([0,1];Hn(D;m)),\dot{c}^{j}\rightharpoonup\dot{c}^{*}\text{ in }L^{2}([0,1];H^{n}(D;\mathbb{R}^{m})),

hence the case k=0k=0 is immediate. We will show that for k=1,,nk=1,\ldots,n,

(5.13) hjh in L2([0,1];Hk(D;m))h^{j}\rightharpoonup h\text{ in }L^{2}([0,1];H^{k}(D;\mathbb{R}^{m}))

implies

(5.14) DcjhjDch in L2([0,1];Hk1(D;m)),D_{c^{j}}h^{j}\rightharpoonup D_{c^{*}}h\quad\text{ in }\,L^{2}([0,1];H^{k-1}(D;\mathbb{R}^{m})),

from which (5.12) follows by induction. First, considering all the vector fields as sections of D×mD\times\mathbb{R}^{m}, we have that

Dcjhj=1|θcj|(θhjIIcj(θcj,hj)),Dch=1|θc|(θhIIc(θc,h)),D_{c^{j}}h^{j}=\frac{1}{|\partial_{\theta}c^{j}|}\left(\partial_{\theta}h^{j}-\textup{II}_{c^{j}}(\partial_{\theta}c^{j},h^{j})\right),\quad D_{c^{*}}h=\frac{1}{|\partial_{\theta}c^{*}|}\left(\partial_{\theta}h-\textup{II}_{c^{*}}(\partial_{\theta}c^{*},h)\right),

where the subscript of II denotes the point where it is evaluated (recall that II is the second fundamental form of 𝒩\mathcal{N} in m\mathbb{R}^{m}).

Since cjcc^{j}\to c^{*} in C([0,1];Cn1(D;m))C([0,1];C^{n-1}(D;\mathbb{R}^{m})) and |θcj||\partial_{\theta}c^{j}| is uniformly bounded from below, we have that |θcj|1|θc|1|\partial_{\theta}c^{j}|^{-1}\to|\partial_{\theta}c^{*}|^{-1} uniformly (in tt and θ\theta). In particular, since IIcj\textup{II}_{c^{j}} are uniformly bounded bilinear forms (this follows again from Lemma 5.11), it follows that DcjhjD_{c^{j}}h^{j} is a bounded sequence in L2([0,1];Hk1(D;m))L^{2}([0,1];H^{k-1}(D;\mathbb{R}^{m})). Therefore, in order to prove (5.14), it is enough to check it with respect to smooth test functions. Let uC([0,1];C(D;m))u\in C([0,1];C^{\infty}(D;\mathbb{R}^{m})), and denote w=u+(1)k1θ2k2uw=u+(-1)^{k-1}\partial_{\theta}^{2k-2}u; we then have

DcjhjDch,uL2([0,1];Hk1(D;m))=DcjhjDch,wL2([0,1];L2(D;m)).\left\langle D_{c^{j}}h^{j}-D_{c^{*}}h,u\right\rangle_{L^{2}([0,1];H^{k-1}(D;\mathbb{R}^{m}))}=\left\langle D_{c^{j}}h^{j}-D_{c^{*}}h,w\right\rangle_{L^{2}([0,1];L^{2}(D;\mathbb{R}^{m}))}.

Since |θcj|1|θc|1|\partial_{\theta}c^{j}|^{-1}\to|\partial_{\theta}c^{*}|^{-1} uniformly, the right-hand side converges to zero if

θhjθh,wL2([0,1];L2(D;m))0,IIcj(θcj,hj)IIc(θc,h),wL2([0,1];L2(D;m))0.\begin{split}\left\langle\partial_{\theta}h^{j}-\partial_{\theta}h,w\right\rangle_{L^{2}([0,1];L^{2}(D;\mathbb{R}^{m}))}&\to 0,\\ \left\langle\textup{II}_{c^{j}}(\partial_{\theta}c^{j},h^{j})-\textup{II}_{c^{*}}(\partial_{\theta}c^{*},h),w\right\rangle_{L^{2}([0,1];L^{2}(D;\mathbb{R}^{m}))}&\to 0.\end{split}

The first one follows from (5.13). The second one follows also from (5.13), using in additon the fact that cjcc^{j}\to c^{*} in C([0,1];Cn1(D;m))C([0,1];C^{n-1}(D;\mathbb{R}^{m})) implies that IIcjIIc\textup{II}_{c^{j}}\to\textup{II}_{c^{*}} uniformly, and θcjθc\partial_{\theta}c^{j}\to\partial_{\theta}c^{*} uniformly. This completes the proof of (5.14), and hence also of (5.12).

Step V: cc^{*} is a minimizer. Using the embedding ι\iota, and considering all curves as curves in m\mathbb{R}^{m}, we can write the energy as

E(c)=k=0n0102πak(c)|Dckc˙|2|θc|dθdt=k=0nak(c)|θc|Dckc˙L2([0,1];L2(D;m))2,\begin{split}E(c)&=\sum_{k=0}^{n}\int_{0}^{1}\int_{0}^{2\pi}a_{k}(\ell_{c})|D_{c}^{k}\dot{c}|^{2}|\partial_{\theta}c|\,\,\mathrm{d}\theta\,\,\mathrm{d}t\\ &=\sum_{k=0}^{n}\|\sqrt{a_{k}(\ell_{c})}\sqrt{|\partial_{\theta}c|}D_{c}^{k}\dot{c}\|_{L^{2}([0,1];L^{2}(D;\mathbb{R}^{m}))}^{2},\end{split}

where the transition to the second line uses the fact that ι\iota is an isometric embedding. Since cjcc^{j}\to c^{*} in C([0,1];Cn1(D;m))C([0,1];C^{n-1}(D;\mathbb{R}^{m})), we have that ak(cj)ak(c)\sqrt{a_{k}(\ell_{c^{j}})}\to\sqrt{a_{k}(\ell_{c^{*}})} uniformly (for k=0,,nk=0,\ldots,n), and that |θcj||θc|\sqrt{|\partial_{\theta}c^{j}|}\to\sqrt{|\partial_{\theta}c^{*}|} uniformly. Therefore, (5.12) implies that for all k=0,,nk=0,\ldots,n,

ak(cj)|θcj|Dcjkc˙jak(c)|θc|Dckc˙ in L2([0,1];L2(D;m)).\sqrt{a_{k}(\ell_{c^{j}})}\sqrt{|\partial_{\theta}c^{j}|}D_{c^{j}}^{k}\dot{c}^{j}\rightharpoonup\sqrt{a_{k}(\ell_{c^{*}})}\sqrt{|\partial_{\theta}c^{*}|}D_{c^{*}}^{k}\dot{c}^{*}\quad\text{ in }\,L^{2}([0,1];L^{2}(D;\mathbb{R}^{m})).

Since the map xx2x\mapsto\|x\|^{2} in a Hilbert space is weakly sequentially lower semicontinuous, we obtain that

infAc0,c1EE(c)lim infE(cj)infAc0,c1E,\inf_{A_{c_{0},c_{1}}}E\leq E(c^{*})\leq\liminf E(c^{j})\to\inf_{A_{c_{0},c_{1}}}E,

hence cc^{*} is a minimizer.

5.6. Geodesic completeness in the smooth category

For closed curves, i.e., D=S1D=S^{1} we obtain also completeness in the smooth category using the no-loss-no-gain result.

Corollary 5.13.

Let n2n\geq 2 and let GG be a smooth Riemannian metric on n(S1,𝒩)\mathcal{I}^{n}(S^{1},\mathcal{N}). Assume that for every metric ball B(c0,r)(n(S1,𝒩),distG)B(c_{0},r)\in(\mathcal{I}^{n}(S^{1},\mathcal{N}),\operatorname{dist}^{G}), there exists a constant C=C(c0,r)>0C=C(c_{0},r)>0 such that conditions from theorem 5.1 hold, i.e.,

(5.1) hGc\displaystyle\|h\|_{G_{c}} Cc1/2shL2(ds),\displaystyle\geq C\ell_{c}^{-1/2}\|\nabla_{\partial_{s}}h\|_{L^{2}(\,\mathrm{d}s)},
(5.2) hGc\displaystyle\|h\|_{G_{c}} CskhL\displaystyle\geq C\|\nabla_{\partial_{s}}^{k}h\|_{L^{\infty}} k=0,,n1,\displaystyle k=0,\ldots,n-1,
(5.3) hGc\displaystyle\|h\|_{G_{c}} CsnhL2.\displaystyle\geq C\|\nabla_{\partial_{s}}^{n}h\|_{L^{2}}.

Then the space (Imm(S1,𝒩),G|Imm(S1,𝒩))(\operatorname{Imm}(S^{1},\mathcal{N}),G|_{\operatorname{Imm}(S^{1},\,\mathcal{N})}) is geodesically complete, where G|Imm(S1,𝒩)G|_{\operatorname{Imm}(S^{1},\,\mathcal{N})} is the restriction of the metric GG to the space of smooth immersions.

Proof.

The proof of this result follows directly by applying Lemma A.1, for V=Tn(D,𝒩)V=T\mathcal{I}^{n}(D,\mathcal{N}), an open subset of Hn(D,T𝒩)H^{n}(D,T\mathcal{N}), and FF the exponential map of GG. ∎

For open curves D=[0,2π]D=[0,2\pi] one has to be slightly more careful, due to the potential loss of smoothness at the boundary; in this case Lemma A.1 only yields that solutions to the geodesic equation with smooth initial data remain at all times in n([0,2π],𝒩)C((0,2π),𝒩)\mathcal{I}^{n}([0,2\pi],\mathcal{N})\cap C^{\infty}((0,2\pi),\mathcal{N}).

6. Incompleteness of constant coefficient metrics on open curves

In our main result we have seen a significant difference between open and closed curves: while we prove that the constant coefficient metrics of order n2n\geq 2 are geodesically and metrically complete on spaces of closed curves, for open curves we had to assume certain non-trivial length-weighted coefficients. In fact, for open curves with values in d\mathbb{R}^{d} it has been observed in [4, Remark 2.7] that constant coefficient Sobolev metrics are in fact metrically incomplete, by constructing an explicit example of a path that leaves the space in finite time. Essentially, they showed that one can shrink a straight line to a point using finite energy. This behavior does not appear for closed curves as blow-up of curvature is an obstruction and thus ensures the completeness of the space.777This is true for metrics of order n2n\geq 2 that are discussed in this paper. Metrics of order n<2n<2 are not strong enough to detect this curvature blowup, which results in metric- and geodesic-incompleteness, as seen in [28, Section 6.1]. The goals of this section are twofold:

  1. (1)

    to extend the example of metric incompleteness from [4] (Example 6.1);

  2. (2)

    to show that shrinking to a point is the only possibility to leave the space with finite energy (Theorem 6.3), and deduce from it a condition that ensures the existence of geodesics between given curves (Theorem 6.7).

The following example of metric incompleteness is a generalization of the example given in [4, Remark 2.7]. We only present it for 2\mathbb{R}^{2}-valued curves for the sake of clarity; it can be adapted easily to arbitrary target manifolds (disappearing along a geodesic instead of a straight line).

Example 6.1.

Consider n([0,2π];2)\mathcal{I}^{n}([0,2\pi];\mathbb{R}^{2}) with the metric

hGc2=hL2(ds)2+snhL2(ds)2.\|h\|_{G_{c}}^{2}=\|h\|_{L^{2}(\,\mathrm{d}s)}^{2}+\|\nabla_{\partial_{s}}^{n}h\|_{L^{2}(\,\mathrm{d}s)}^{2}.

Consider the path c:[0,1)nc:[0,1)\to\mathcal{I}^{n}, defined by

c(t,θ)=((1t)(θπ)+f(t),g(t))c(t,\theta)=((1-t)(\theta-\pi)+f(t),g(t))

for some smooth functions f,g:[0,1)f,g:[0,1)\to\mathbb{R} to be determined. Note that

cθ\displaystyle c_{\theta} =(1t,0),\displaystyle=(1-t,0), ct\displaystyle\qquad c_{t} =((θπ)+f(t),g(t)),\displaystyle=(-(\theta-\pi)+f^{\prime}(t),g^{\prime}(t)),
sct\displaystyle\nabla_{\partial_{s}}c_{t} =(11t,0),\displaystyle=\left(\frac{-1}{1-t},0\right), skct\displaystyle\nabla_{\partial_{s}}^{k}c_{t} =0 for k>1.\displaystyle=0\text{ for }k>1.

Hence

ctGc2\displaystyle\|c_{t}\|_{G_{c}}^{2} =02π((f(t)(θπ))2+g(t)2)(1t)dθ\displaystyle=\int_{0}^{2\pi}\left((f^{\prime}(t)-(\theta-\pi))^{2}+g^{\prime}(t)^{2}\right)(1-t)\,\,\mathrm{d}\theta
=2π(1t)(π23+f(t)2+g(t)2),\displaystyle=2\pi(1-t)\left(\frac{\pi^{2}}{3}+f^{\prime}(t)^{2}+g^{\prime}(t)^{2}\right),

and therefore

length(c)=01ctGc=2π01(1t)1/2(π23+f(t)2+g(t)2)1/2dt2π01(1t)1/2(π3+|f(t)|+|g(t)|)dt,\begin{split}\operatorname{length}(c)&=\int_{0}^{1}\|c_{t}\|_{G_{c}}=\sqrt{2\pi}\int_{0}^{1}(1-t)^{1/2}\left(\frac{\pi^{2}}{3}+f^{\prime}(t)^{2}+g^{\prime}(t)^{2}\right)^{1/2}\,\mathrm{d}t\\ &\leq\sqrt{2\pi}\int_{0}^{1}(1-t)^{1/2}\left(\frac{\pi}{\sqrt{3}}+|f^{\prime}(t)|+|g^{\prime}(t)|\right)\,\mathrm{d}t,\end{split}

hence length(c)<\operatorname{length}(c)<\infty if 01|f(t)|(1t)1/2dt<\int_{0}^{1}|f^{\prime}(t)|(1-t)^{1/2}\,\,\mathrm{d}t<\infty and similarly for gg. Under these restrictions on ff and gg many things can happen, for example:

  1. (1)

    For f=g=0f=g=0 we obtain that cc converges, as t1t\to 1, to the constant curve at the origin;

  2. (2)

    For f(t)=tx0f(t)=tx_{0} and g(t)=ty0g(t)=ty_{0}, cc converges to the constant curve at (x0,y0)(x_{0},y_{0}).

  3. (3)

    For f(t)=log(1t)f(t)=-\log(1-t) and g=0g=0, cc converges to a point at infinity at the positive end of the xx axis.

  4. (4)

    For f(t)=sin(log(1t))f(t)=\sin(-\log(1-t)) and g=0g=0, cc does not converge pointwise to anything in 2\mathbb{R}^{2}.

Note that this analysis does not change if we replace GG with another constant coefficient metric. This shows that (n([0,2π];2),distG)(\mathcal{I}^{n}([0,2\pi];\mathbb{R}^{2}),\operatorname{dist}^{G}) is not metrically complete. However, from the point of view of the metric completion, all these different choices of ff and gg are the same point in the completion — indeed, let

ci(t,θ)=((1t)(θπ)+fi(t),gi(t)),i=1,2,c^{i}(t,\theta)=((1-t)(\theta-\pi)+f_{i}(t),g_{i}(t)),\quad i=1,2,

and define, for a fixed t[0,1)t\in[0,1), the path γt(τ,θ)\gamma^{t}(\tau,\theta) as the affine homotopy between c1(t,)=γt(0,)c^{1}(t,\cdot)=\gamma^{t}(0,\cdot) and c2(t,)=γt(1,)c^{2}(t,\cdot)=\gamma^{t}(1,\cdot), that is,

γt(τ,θ)=((1t)(θπ)+τf1(t)+(1τ)f2(t),τg1(t)+(1τ)g2(t)).\gamma^{t}(\tau,\theta)=((1-t)(\theta-\pi)+\tau f_{1}(t)+(1-\tau)f_{2}(t),\tau g_{1}(t)+(1-\tau)g_{2}(t)).

Since |γθt|=1t|\gamma^{t}_{\theta}|=1-t and γτt=(f1(t)f2(t),g1(t)g2(t))\gamma^{t}_{\tau}=(f_{1}(t)-f_{2}(t),g_{1}(t)-g_{2}(t)) is independent of θ\theta and τ\tau, it follows immediately that

length(γt)1t.\operatorname{length}(\gamma^{t})\propto 1-t.

Therefore,

distG(c1(t,),c2(t,))length(γt)1t0\operatorname{dist}^{G}(c^{1}(t,\cdot),c^{2}(t,\cdot))\leq\operatorname{length}(\gamma^{t})\propto 1-t\to 0

as t1t\to 1. This means, that in the metric completion, all the Cauchy sequences obtained by choosing different ffs and ggs are equivalent, hence converge to a single point.

This example leads to the following open question:

Question 6.2.

Let GG be a constant coefficient Sobolev metric of order n2n\geq 2 of the type (1.2) on n([0,2π],𝒩)\mathcal{I}^{n}([0,2\pi],\mathcal{N}). For i=1,2i=1,2, let cnin([0,2π],𝒩)c^{i}_{n}\in\mathcal{I}^{n}([0,2\pi],\mathcal{N}) be two Cauchy sequences with cni0\ell_{c^{i}_{n}}\to 0. Does it hold that

limndistG(cn1,cn2)=0?\lim_{n\to\infty}\operatorname{dist}^{G}(c^{1}_{n},c^{2}_{n})=0?

We now show that if a Cauchy sequence of curves does not converge, its lengths must tend to zero:

Theorem 6.3.

Let GG be a constant coefficient Sobolev metric of order n2n\geq 2 of the type (1.2) on n([0,2π];𝒩)\mathcal{I}^{n}([0,2\pi];\mathcal{N}), where both a0a_{0} and ana_{n} are strictly positive constants. Assume that (cn)nn([0,2π];𝒩)(c_{n})_{n\in\mathbb{N}}\subset\mathcal{I}^{n}([0,2\pi];\mathcal{N}) is a Cauchy sequence with respect to distG\operatorname{dist}^{G}, whose lengths are bounded from below, that is cn>δ>0\ell_{c_{n}}>\delta>0 for all nn. Then cnc_{n} converges to some cn([0,2π];𝒩)c_{\infty}\in\mathcal{I}^{n}([0,2\pi];\mathcal{N}).

Before proving this result we note a consequence of it: if the answer to Question 6.2 is positive, then, together with Theorem 6.3, it would give a positive answer to the following conjecture on the metric completion of (n([0,2π],𝒩),distG)(\mathcal{I}^{n}([0,2\pi],\mathcal{N}),\operatorname{dist}^{G}):

Question 6.4.

Let GG be a constant coefficient Sobolev metric of order n2n\geq 2 of the type (1.2) on n([0,2π],𝒩)\mathcal{I}^{n}([0,2\pi],\mathcal{N}). Is the metric completion of (n([0,2π],𝒩),distG)(\mathcal{I}^{n}([0,2\pi],\mathcal{N}),\operatorname{dist}^{G}) given by n([0,2π],𝒩){0}\mathcal{I}^{n}([0,2\pi],\mathcal{N})\cup\{0\}, where {0}\{0\} represents the limit of all vanishing-length Cauchy sequences?

In our infinite dimensional situation metric incompleteness does not imply geodesic incompleteness. Furthermore the paths constructed in Example 6.1 are not geodesics (a direct calculations shows that the boundary equations in the geodesic equations are not satisfied). This leads to the following question:

Question 6.5.

Let GG be a constant coefficient Sobolev metric of order n2n\geq 2 of the type (1.2) on n([0,2π],𝒩)\mathcal{I}^{n}([0,2\pi],\mathcal{N}). Is n([0,2π],𝒩)\mathcal{I}^{n}([0,2\pi],\mathcal{N}) geodesically complete?

We proceed with the proof of Theorem 6.3. We will need the following lemma, which is similar to Lemma 5.9:

Lemma 6.6.

Let GG be a Sobolev metric of order n2n\geq 2 on n([0,2π],𝒩)\mathcal{I}^{n}([0,2\pi],\mathcal{N}), such that, for every hHn([0,2π],cT𝒩)h\in H^{n}([0,2\pi],c^{*}T\mathcal{N}),

shL2(ds)Cmax{1,c1}hGc\|\nabla_{\partial_{s}}h\|_{L^{2}(\,\mathrm{d}s)}\leq C\max\left\{1,\ell_{c}^{-1}\right\}\|h\|_{G_{c}}

for some uniform constant C>0C>0. Then, the function cc3/2c\mapsto\ell_{c}^{3/2} is Lipschitz continuous on every metric ball in (n([0,2π],𝒩),distG)(\mathcal{I}^{n}([0,2\pi],\mathcal{N}),\operatorname{dist}^{G}). Moreover, the Lipschitz constant of in B(c0,r)B(c_{0},r) depends only on c0\ell_{c_{0}} and rr, and is an increasing function of both, that is, there exists a function L(C,,r)\textup{L}(C,\ell,r), increasing in all variables, such that

|c3/2c~3/2|L(C,c0,r)distG(c,c~) for every c,c~B(c0,r).|\ell_{c}^{3/2}-\ell_{\tilde{c}}^{3/2}|\leq\textup{L}(C,\ell_{c_{0}},r)\operatorname{dist}^{G}(c,\tilde{c})\quad\text{ for every }c,\tilde{c}\in B(c_{0},r).
Proof of Lemma 6.6.

As in the proof of Lemma 5.9, we have

|Dc,hc3/2|32cshL2(ds)32Cmax{c,1}hGc32C(1+c3/2)hGc|D_{c,h}\ell_{c}^{3/2}|\leq\frac{3}{2}\ell_{c}\|\nabla_{\partial_{s}}h\|_{L^{2}(\,\mathrm{d}s)}\leq\frac{3}{2}C\max\left\{\ell_{c},1\right\}\|h\|_{G_{c}}\leq\frac{3}{2}C(1+\ell_{c}^{3/2})\|h\|_{G_{c}}

from which the claim follows by Lemma 5.7, with L(C,,r):=L(C,3/2,r)\textup{L}(C,\ell,r):=L(C,\ell^{3/2},r). ∎

Proof of Theorem 6.3.

Assume that cnc_{n} is a Cauchy sequence with cn>δ\ell_{c_{n}}>\delta for some δ>0\delta>0.

Since GG has constant coefficients (with a0,an>0a_{0},a_{n}>0), we have, using (4.2) for k=1k=1, that

shL2(ds)2Cmax{1,c2}(hL2(ds)2+snhL2(ds))Cmax{1,c2}hGc2,\begin{split}\|\nabla_{\partial_{s}}h\|_{L^{2}(\,\mathrm{d}s)}^{2}&\leq C\max\left\{1,\ell_{c}^{-2}\right\}\left(\|h\|_{L^{2}(\,\mathrm{d}s)}^{2}+\|\nabla_{\partial_{s}}^{n}h\|_{L^{2}(\,\mathrm{d}s)}\right)\\ &\leq C^{\prime}\max\left\{1,\ell_{c}^{-2}\right\}\|h\|_{G_{c}}^{2},\end{split}

where the constants C,CC,C^{\prime} depend only on nn, a0a_{0} and ana_{n}, hence we can apply Lemma 6.6.

There exists N1N_{1} large enough such that cnB(cN1,1/2)c_{n}\in B(c_{N_{1}},1/2) for all nN1n\geq N_{1}. Applying Lemma 6.6, for B(cN1,1)B(c_{N_{1}},1) we obtain that there exists a constant ¯\bar{\ell}, depending on cN1c_{N_{1}} such that

c¯ for all cB(cN1,1).\ell_{c}\leq\bar{\ell}\qquad\text{ for all }c\in B(c_{N_{1}},1).

In particular, this applies to all cnc_{n} for nN1n\geq N_{1}.

Let L(C,,r)\textup{L}(C^{\prime},\ell,r) be the Lipschitz constant bound as in Lemma 6.6, and denote L¯:=L(C,¯,1)\bar{L}:=\textup{L}(C^{\prime},\bar{\ell},1). Denote r0:=min{δ3/22L¯,1/2}r_{0}:=\min\left\{\frac{\delta^{3/2}}{2\bar{L}},1/2\right\}. There exists an index N2>N1N_{2}>N_{1} such that for nN2n\geq N_{2} we have that cnB(cN2,r0/3)c_{n}\in B(c_{N_{2}},r_{0}/3), that is distG(cn,cN)<r0/3\operatorname{dist}^{G}(c_{n},c_{N})<r_{0}/3. Applying Lemma 6.6 to B(cN2,r0)B(c_{N_{2}},r_{0}) and the bound cN2¯\ell_{c_{N_{2}}}\leq\bar{\ell}, we have that

|c3/2c~3/2|L¯distG(c,c~) for every c,c~B(cN2,r0).\left|\ell_{c}^{3/2}-\ell_{\tilde{c}}^{3/2}\right|\leq\bar{L}\operatorname{dist}^{G}(c,\tilde{c})\text{ for every }c,\tilde{c}\in B(c_{N_{2}},r_{0}).

Since cN2>δ\ell_{c_{N_{2}}}>\delta, and B(cN2,r0)B(cN1,1)B(c_{N_{2}},r_{0})\subset B(c_{N_{1}},1), we obtain

(6.1) c[δ22/3,¯] for every cB(cN2,r0).\ell_{c}\in\left[\frac{\delta}{2^{2/3}},\bar{\ell}\right]\qquad\text{ for every }c\in B(c_{N_{2}},r_{0}).

Denote by GG^{\prime} the standard scale-invariant metric of order nn on n([0,2π],𝒩)\mathcal{I}^{n}([0,2\pi],\mathcal{N}); that is,

hGc2=c3hL2(ds)2+c2n3snhL2(ds)2.\|h\|_{G^{\prime}_{c}}^{2}=\ell_{c}^{-3}\|h\|_{L^{2}(\,\mathrm{d}s)}^{2}+\ell_{c}^{2n-3}\|\nabla_{\partial_{s}}^{n}h\|_{L^{2}(\,\mathrm{d}s)}^{2}.

Recall that (n,distG)(\mathcal{I}^{n},\operatorname{dist}^{G^{\prime}}) is metrically complete by Theorem 5.3. Using (6.1) and Lemma 4.1, it follows that GG^{\prime} and GG are equivalent in B(cN2,r0)B(c_{N_{2}},r_{0}). From here we continue in a similar way as in Propositions 5.55.6: Let c,c~B(cN2,r0/3)c,\tilde{c}\in B(c_{N_{2}},r_{0}/3), and 0<ε<r0/3distG(c,c~)0<\varepsilon<r_{0}/3-\operatorname{dist}^{G}(c,\tilde{c}). Let γ\gamma be a curve between cc and c~\tilde{c} such that lengthG(γ)<distG(c,c~)+ε\operatorname{length}^{G}(\gamma)<\operatorname{dist}^{G}(c,\tilde{c})+\varepsilon. By triangle inequality, we have that γB(cN2,r0)\gamma\subset B(c_{N_{2}},r_{0}), and since GG^{\prime} and GG are equivalent there, we have that for some constant C>0C>0 (independent of γ\gamma),

distG(c,c~)lengthG(γ)ClengthG(γ)<C(distG(c,c~)+ε),\operatorname{dist}^{G^{\prime}}(c,\tilde{c})\leq\operatorname{length}^{G^{\prime}}(\gamma)\leq C\operatorname{length}^{G}(\gamma)<C(\operatorname{dist}^{G}(c,\tilde{c})+\varepsilon),

and since ε\varepsilon is arbitrarily small, we conclude that

distG(c,c~)CdistG(c,c~),for every c,c~B(cN2,r0/3).\operatorname{dist}^{G^{\prime}}(c,\tilde{c})\leq C\operatorname{dist}^{G}(c,\tilde{c}),\qquad\text{for every }c,\tilde{c}\in B(c_{N_{2}},r_{0}/3).

Since for every nN2n\geq N_{2}, cnB(cN2,r0/3)c_{n}\in B(c_{N_{2}},r_{0}/3), it follows that cnc_{n} is a Cauchy sequence with respect to GG^{\prime} as well. Since (n,distG)(\mathcal{I}^{n},\operatorname{dist}^{G^{\prime}}) is metrically complete, we have that there exists cnc_{\infty}\in\mathcal{I}^{n} such that distG(cn,c)0\operatorname{dist}^{G^{\prime}}(c_{n},c_{\infty})\to 0. Since both GG and GG^{\prime} are strong metrics on n\mathcal{I}^{n}, they induce the same topology [24, VII, Proposition 6.1], and thus cncnc_{n}\to c_{\infty}\in\mathcal{I}^{n} with respect to GG as well, which completes the proof. ∎

From the arguments in the proof of Theorem 6.3, we also obtain that for close enough immersions c0,c1n([0,2π];𝒩)c_{0},c_{1}\in\mathcal{I}^{n}([0,2\pi];\mathcal{N}), there exists a connecting minimizing geodesic:

Theorem 6.7.

Let GG be a constant coefficient Sobolev metric of order n2n\geq 2 of the type (1.2) on n([0,2π];𝒩)\mathcal{I}^{n}([0,2\pi];\mathcal{N}), where both a0a_{0} and ana_{n} are strictly positive constants. Let c0n([0,2π];𝒩)c_{0}\in\mathcal{I}^{n}([0,2\pi];\mathcal{N}). Then, there exists a constant r0r_{0}, depending only on the coefficients aka_{k} and on c0\ell_{c_{0}}, such that for every c1B(c0,r0)c_{1}\in B(c_{0},r_{0}), there exists a minimizing geodesic between c0c_{0} and c1c_{1}.

Remark 6.8.

The proof below, together with the bound (5.5), imply that r0r_{0} can be chosen such that

r0=Cc03/21+c03/2Cmin(c03/2,12),r_{0}=C\frac{\ell_{c_{0}}^{3/2}}{1+\ell_{c_{0}}^{3/2}}\geq C\min\left(\ell_{c_{0}}^{3/2},\frac{1}{2}\right),

where CC depends only on the coefficients aka_{k}, k=0,,nk=0,\ldots,n. Note that we do not know whether the existence of minimizing geodesics fails in general; it might be that although the space in metrically incomplete, a minimizing geodesic between any two curves c0,c1n([0,2π],𝒩)c_{0},c_{1}\in\mathcal{I}^{n}([0,2\pi],\mathcal{N}) exists.

Proof.

As in Theorem 6.3, there exists a constant CC, depending only on nn, a0a_{0} and ana_{n} (or alternatively, on aka_{k}, k=0,,nk=0,\ldots,n) such that the assumption of Lemma 6.6 holds. Fix L~:=L(C,c0,1)\tilde{L}:=\textup{L}(C,\ell_{c_{0}},1), where L(C,,r)\textup{L}(C,\ell,r) is the Lipschitz constant function from Lemma 6.6. Let

r0=min(c03/22L~,1).r_{0}=\min\left(\frac{\ell_{c_{0}}^{3/2}}{2\tilde{L}},1\right).

It follows that

c[122/3c0,32/322/3c0] for every cB(c0,r0).\ell_{c}\in\left[\frac{1}{2^{2/3}}\ell_{c_{0}},\frac{3^{2/3}}{2^{2/3}}\ell_{c_{0}}\right]\quad\text{ for every }c\in B(c_{0},r_{0}).

As in Theorem 6.3, it follows that in this ball GG is uniformly equivalent to a scale-invariant Sobolev metric of order nn on n([0,2π];𝒩)\mathcal{I}^{n}([0,2\pi];\mathcal{N}), hence Lemmata 5.105.12 hold uniformly on B(c0,r0)B(c_{0},r_{0}) (rather than on every metric ball).

Let c1B(c0,r0)c_{1}\in B(c_{0},r_{0}). Define the energy E(c)E(c) and the set of paths Ac0,c1A_{c_{0},c_{1}} as in Section 5.5. Let cAc0,c1c\in A_{c_{0},c_{1}}, with length(c)<r0\operatorname{length}(c)<r_{0}. Assume that cc has constant speed; we then have

E(c)=length(c)2<r02.E(c)=\operatorname{length}(c)^{2}<r_{0}^{2}.

Therefore,

infAc0,c1E<r02.\inf_{A_{c_{0},c_{1}}}E<r_{0}^{2}.

We can now take a minimizing sequence cjAc0,c1c^{j}\in A_{c_{0},c_{1}}, and assume without loss of generality that E(cj)<r02E(c^{j})<r_{0}^{2} for all jj. The proof now follows in the same way as in Theorem 6.3. ∎

Appendix A The geodesic equation

A.1. Proof of Lemma 3.3: the geodesic equation

Proof of Lemma 3.3.

To prove the formula for the geodesic equation we consider the energy of a path of immersions c(t,θ)c(t,\theta). Furthermore, we will treat the zeroth and first order terms separately. Varying c(t,θ)c(t,\theta) in direction h(t,θ)h(t,\theta) with h(0,θ)=h(1,θ)=0h(0,\theta)=h(1,\theta)=0 we obtain for the zeroth-order term:

d(01a0(c)Dg(ct,ct)|c|dθdt)(h)\displaystyle d\left(\int_{0}^{1}a_{0}(\ell_{c})\int_{D}g(c_{t},c_{t})|c^{\prime}|\,\mathrm{d}\theta\,\mathrm{d}t\right)(h)
=01a0(c)Dc,hcDg(ct,ct)|c|dθdt\displaystyle\qquad=\int_{0}^{1}a_{0}^{\prime}(\ell_{c})D_{c,h}\ell_{c}\int_{D}g(c_{t},c_{t})|c^{\prime}|\,\mathrm{d}\theta\,\mathrm{d}t
+01a0(c)D2g(hct,ct)+g(ct,ct)g(v,sh)dsdt\displaystyle\qquad\qquad\qquad+\int_{0}^{1}a_{0}(\ell_{c})\int_{D}2g(\nabla_{h}c_{t},c_{t})+g(c_{t},c_{t})g(v,\nabla_{\partial_{s}}h)\,\mathrm{d}s\,\mathrm{d}t
=01(a0(c)Dg(v,sh)dsDg(ct,ct)ds)dt\displaystyle\qquad=\int_{0}^{1}\left(a_{0}^{\prime}(\ell_{c})\int_{D}g(v,\nabla_{\partial_{s}}h)\,\mathrm{d}s\int_{D}g(c_{t},c_{t})\,\mathrm{d}s\right)\,\mathrm{d}t
+01(a0(c)D2g(th,ct)+g(sh,vg(ct,ct))ds)dt.\displaystyle\qquad\qquad\qquad+\int_{0}^{1}\left(a_{0}(\ell_{c})\int_{D}2g(\nabla_{\partial_{t}}h,c_{t})+g(\nabla_{\partial_{s}}h,vg(c_{t},c_{t}))\,\mathrm{d}s\right)\,\mathrm{d}t.

where we used in the last step that

(A.1) hct=th.\nabla_{h}c_{t}=\nabla_{\partial_{t}}h\;.

and the variation formula for the length c\ell_{c} from Lemma 3.2. Here, as before, v=c/|c|v=c^{\prime}/|c^{\prime}| is the unit length tangent vector to the curve cc.

Similarly we calculate for the first-order terms:

d(01a1(c)Dg(sct,sct)|c|dθdt)(h)\displaystyle d\left(\int_{0}^{1}a_{1}(\ell_{c})\int_{D}g(\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})|c^{\prime}|\,\mathrm{d}\theta\,\mathrm{d}t\right)(h)
=01a1(c)(Dg(v,sh)dsDg(sct,sct)ds)dt\displaystyle=\int_{0}^{1}a_{1}^{\prime}(\ell_{c})\left(\int_{D}g(v,\nabla_{\partial_{s}}h)\,\mathrm{d}s\int_{D}g(\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\right)\,\mathrm{d}t
+01a1(c)(D2g(hsct,sct)+g(sct,sct)g(v,sh)ds)dt\displaystyle\;+\int_{0}^{1}a_{1}(\ell_{c})\left(\int_{D}2g(\nabla_{h}\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})+g(\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})g(v,\nabla_{\partial_{s}}h)\,\mathrm{d}s\right)\,\mathrm{d}t
=01a1(c)(Dg(v,sh)dsDg(sct,sct)ds)dt\displaystyle=\int_{0}^{1}a_{1}^{\prime}(\ell_{c})\left(\int_{D}g(v,\nabla_{\partial_{s}}h)\,\mathrm{d}s\int_{D}g(\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\right)\,\mathrm{d}t
+01a1(c)(D2g(g(v,sh)sct+shct+(v,h)ct,sct)ds)dt\displaystyle\;+\int_{0}^{1}a_{1}(\ell_{c})\left(\int_{D}2g(-g(v,\nabla_{\partial_{s}}h)\nabla_{\partial_{s}}c_{t}+\nabla_{\partial_{s}}\nabla_{h}c_{t}+\mathcal{R}(v,h)c_{t},\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\right)\,\mathrm{d}t
+01a1(c)(Dg(sct,sct)g(v,sh)ds)dt\displaystyle\;+\int_{0}^{1}a_{1}(\ell_{c})\left(\int_{D}g(\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})g(v,\nabla_{\partial_{s}}h)\,\mathrm{d}s\right)\,\mathrm{d}t
=01a1(c)(Dg(v,sh)dsDg(sct,sct)ds)dt\displaystyle=\int_{0}^{1}a_{1}^{\prime}(\ell_{c})\left(\int_{D}g(v,\nabla_{\partial_{s}}h)\,\mathrm{d}s\int_{D}g(\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\right)\,\mathrm{d}t
+01a1(c)(Dg(g(v,sh)sct+2sth+2(v,h)ct,sct)ds)dt.\displaystyle\;+\int_{0}^{1}a_{1}(\ell_{c})\left(\int_{D}g(-g(v,\nabla_{\partial_{s}}h)\nabla_{\partial_{s}}c_{t}+2\nabla_{\partial_{s}}\nabla_{\partial_{t}}h+2\mathcal{R}(v,h)c_{t},\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\right)\,\mathrm{d}t.

Sorting this by derivatives of hh we obtain

d(01a1(c)Dg(sct,sct)|c|dθdt)(h)\displaystyle d\left(\int_{0}^{1}a_{1}(\ell_{c})\int_{D}g(\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})|c^{\prime}|\,\mathrm{d}\theta\,\mathrm{d}t\right)(h)
=01a1(c)D2g(sth,sct)+2g((v,h)ct,sct)\displaystyle=\int_{0}^{1}a_{1}(\ell_{c})\int_{D}2g(\nabla_{\partial_{s}}\nabla_{\partial_{t}}h,\nabla_{\partial_{s}}c_{t})+2g(\mathcal{R}(v,h)c_{t},\nabla_{\partial_{s}}c_{t})
+g(sh,a1(c)a1(c)Dg(sct,sct)dsvg(sct,sct)v)dsdt.\displaystyle\qquad+g\left(\nabla_{\partial_{s}}h,\frac{a_{1}^{\prime}(\ell_{c})}{a_{1}(\ell_{c})}\int_{D}g(\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\;v-g(\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})v\right)\,\mathrm{d}s\,\mathrm{d}t.

Putting both together we obtain:

dE(c).h=01D2a0(c)g(th,ct)+2a1(c)g(sth,sct)+2a1(c)g((v,h)ct,sct)+g(sh,Ψc(ct,ct)v)dsdt,dE(c).h=\int_{0}^{1}\int_{D}2a_{0}(\ell_{c})g(\nabla_{\partial_{t}}h,c_{t})+2a_{1}(\ell_{c})g(\nabla_{\partial_{s}}\nabla_{\partial_{t}}h,\nabla_{\partial_{s}}c_{t})\\ +2a_{1}(\ell_{c})g(\mathcal{R}(v,h)c_{t},\nabla_{\partial_{s}}c_{t})+g(\nabla_{\partial_{s}}h,\Psi_{c}(c_{t},c_{t})v)\,\mathrm{d}s\,\mathrm{d}t,

where

Ψc(ct,ct)\displaystyle\Psi_{c}(c_{t},c_{t}) =a0(c)g(ct,ct)+a0(c)Dg(ct,ct)ds\displaystyle=a_{0}(\ell_{c})g(c_{t},c_{t})+a_{0}^{\prime}(\ell_{c})\int_{D}g(c_{t},c_{t})\,\mathrm{d}s
a1(c)g(sct,sct)+a1(c)Dg(sct,sct)ds.\displaystyle\quad-a_{1}(\ell_{c})g(\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})+a_{1}^{\prime}(\ell_{c})\int_{D}g(\nabla_{\partial_{s}}c_{t},\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s.

To obtain the geodesic equation, we have to integrate by parts to free hh from all derivatives. We will treat the four terms separately. For the first two terms we recall that ds\,\mathrm{d}s depends on the curve cc (and thus on time tt), i.e., for time dependent vector fields hh and kk, with h(0,θ)=h(1,θ)=0h(0,\theta)=h(1,\theta)=0, we have

(A.2) 01Dg(th,k)dsdt=01Dg(h,t(|c|k))dθdt=01Dg(h,tk+g(v,sct)k)dsdt.\begin{split}\int_{0}^{1}\int_{D}g(\nabla_{\partial_{t}}h,k)\,\mathrm{d}s\,\mathrm{d}t&=-\int_{0}^{1}\int_{D}g(h,\nabla_{\partial_{t}}(|c^{\prime}|k))\,\mathrm{d}\theta\,\mathrm{d}t\\ &=-\int_{0}^{1}\int_{D}g(h,\nabla_{\partial_{t}}k+g(v,\nabla_{\partial_{s}}c_{t})k)\,\mathrm{d}s\,\mathrm{d}t\;.\end{split}

Applying this formula to the first term yields:

201Dg(th,a0(c)ct)dsdt=2Da0(c)g(h,tct+g(v,sct)ct+a0(c)a0(c)Dg(v,sct)dsct)dsdt.2\int_{0}^{1}\int_{D}g(\nabla_{\partial_{t}}h,a_{0}(\ell_{c})c_{t})\,\mathrm{d}s\,\mathrm{d}t\\ =-2\int_{D}a_{0}(\ell_{c})\,g\Big{(}h,\nabla_{t}c_{t}+g(v,\nabla_{\partial_{s}}c_{t})c_{t}\\ +\tfrac{a_{0}^{\prime}(\ell_{c})}{a_{0}(\ell_{c})}\int_{D}g(v,\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\;c_{t}\Big{)}\,\mathrm{d}s\,\mathrm{d}t\;.

For the second term we need to apply integration by parts in space first:

201Dg(sth,a1(c)sct)dsdt\displaystyle 2\int_{0}^{1}\int_{D}g(\nabla_{\partial_{s}}\nabla_{\partial_{t}}h,a_{1}(\ell_{c})\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\,\mathrm{d}t
=201g(th,a1(c)sct)|02πdt201Dg(th,a1(c)ssct)dsdt\displaystyle=2\int_{0}^{1}g(\nabla_{\partial_{t}}h,a_{1}(\ell_{c})\nabla_{\partial_{s}}c_{t})|^{2\pi}_{0}\,\mathrm{d}t-2\int_{0}^{1}\int_{D}g(\nabla_{\partial_{t}}h,a_{1}(\ell_{c})\nabla_{\partial_{s}}\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\,\mathrm{d}t
=201g(h,t(a1(c)sct))|02πdt+201Da1(c)g(h,ts2ct)dsdt\displaystyle=-2\int_{0}^{1}g(h,\nabla_{\partial_{t}}\left(a_{1}(\ell_{c})\nabla_{\partial_{s}}c_{t}\right))|^{2\pi}_{0}\,\mathrm{d}t+2\int_{0}^{1}\int_{D}a_{1}(\ell_{c})g(h,\nabla_{\partial_{t}}\nabla^{2}_{s}c_{t})\,\mathrm{d}s\,\mathrm{d}t
+201Da1(c)g(h,g(v,sct)s2ct+a1(c)a1(c)Dg(v,sct)dss2ct)dsdt\displaystyle\;+2\int_{0}^{1}\!\!\int_{D}\!a_{1}(\ell_{c})g\!\left(h,g(v,\nabla_{\partial_{s}}c_{t})\nabla^{2}_{s}c_{t}+\tfrac{a_{1}^{\prime}(\ell_{c})}{a_{1}(\ell_{c})}\int_{D}g(v,\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\;\nabla^{2}_{s}c_{t}\right)\!\,\mathrm{d}s\,\mathrm{d}t

For the third term we use the symmetries of the curvature tensor to obtain

201Da1(c)g((v,h)ct,sct)dsdt=201Da1(c)g((ct,sct)v,h)dsdt.\displaystyle 2\!\!\int_{0}^{1}\!\!\int_{D}\!\!a_{1}(\ell_{c})g(\mathcal{R}(v,h)c_{t},\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\,\mathrm{d}t=2\!\!\int_{0}^{1}\!\!\int_{D}a_{1}(\ell_{c})g(\mathcal{R}(c_{t},\nabla_{\partial_{s}}c_{t})v,h)\,\mathrm{d}s\,\mathrm{d}t.

Finally for the last term we need to integrate in parts in space again, taking into account the boundary terms:

01Dg(sh,Ψc(ct,ct)v)dsdt\displaystyle\int_{0}^{1}\int_{D}g(\nabla_{\partial_{s}}h,\Psi_{c}(c_{t},c_{t})v)\,\mathrm{d}s\,\mathrm{d}t
=01g(h,Ψc(ct,ct)v)|02πdt01Dg(h,s(Ψc(ct,ct)v))dsdt.\displaystyle\qquad=\int_{0}^{1}g\big{(}h,\Psi_{c}(c_{t},c_{t})v)\Big{|}^{2\pi}_{0}\,\mathrm{d}t-\int_{0}^{1}\int_{D}g\big{(}h,\nabla_{\partial_{s}}(\Psi_{c}(c_{t},c_{t})v)\big{)}\,\mathrm{d}s\,\mathrm{d}t\,.

We can now read off the geodesic equation. We will fist start by collecting the terms on the interior of DD:

a0(c)tcta1(c)ts2ct\displaystyle a_{0}(\ell_{c})\nabla_{\partial_{t}}c_{t}-a_{1}(\ell_{c})\nabla_{\partial_{t}}\nabla^{2}_{s}c_{t}
=a0(c)g(v,sct)cta0(c)(Dg(v,sct)ds)ct\displaystyle\qquad=-a_{0}(\ell_{c})g(v,\nabla_{\partial_{s}}c_{t})c_{t}-a_{0}^{\prime}(\ell_{c})\left(\int_{D}g(v,\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\right)c_{t}
+a1(c)g(v,sct)s2ct+a1(c)(Dg(v,sct)ds)s2ct\displaystyle\qquad\qquad\qquad+a_{1}(\ell_{c})g(v,\nabla_{\partial_{s}}c_{t})\nabla_{\partial_{s}}^{2}c_{t}+a_{1}^{\prime}(\ell_{c})\left(\int_{D}g(v,\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\right)\nabla^{2}_{s}c_{t}
+a1(c)(ct,sct)v12s(Ψc(ct,ct)v).\displaystyle\qquad\qquad\qquad+a_{1}(\ell_{c})\mathcal{R}(c_{t},\nabla_{\partial_{s}}c_{t})v-\frac{1}{2}\nabla_{\partial_{s}}(\Psi_{c}(c_{t},c_{t})v).

From here the result follows using the definition of the inertia operator AcA_{c}, the product rule for the term s(Ψc(ct,ct)v)\nabla_{\partial_{s}}(\Psi_{c}(c_{t},c_{t})v), by using the formula

t(Acct)\displaystyle\nabla_{t}(A_{c}c_{t}) =(tAc)ct+Ac(tct)=t(a0(c)cta1(c)s2ct)\displaystyle=(\nabla_{t}A_{c})c_{t}+A_{c}(\nabla_{t}c_{t})=\nabla_{t}(a_{0}(\ell_{c})c_{t}-a_{1}(\ell_{c})\nabla_{\partial_{s}}^{2}c_{t})
=(Dg(v,sct)ds)a0(c)ct+a0(c)tct\displaystyle=\left(\int_{D}g(v,\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\right)\;a_{0}^{\prime}(\ell_{c})c_{t}+a_{0}(\ell_{c})\nabla_{\partial_{t}}c_{t}
(Dg(v,sct)ds)a1(c)s2cta1(c)ts2ct,\displaystyle\qquad-\left(\int_{D}g(v,\nabla_{\partial_{s}}c_{t})\,\mathrm{d}s\right)\;a_{1}^{\prime}(\ell_{c})\nabla^{2}_{s}c_{t}-a_{1}(\ell_{c})\nabla_{\partial_{t}}\nabla^{2}_{s}c_{t}\;,

and by collecting the boundary terms if D=[0,2π]D=[0,2\pi]. ∎

A.2. Proof of Theorem 3.8: local well-posedness

In this section we will use the method of Ebin–Marsden to obtain local well-posedness and uniqueness of the geodesic equations. Before we prove the local well-posedness we formulate a variant of the no-loss-no-gain result, which is also used in Section 5.6.

Lemma A.1.

Let q2q\geq 2, VHq(D,T𝒩)V\subset H^{q}(D,T\mathcal{N}) an open subset and let F:VHq(D,T𝒩)F:V\to H^{q}(D,T\mathcal{N}) be a smooth and 𝒟q(D)\mathcal{D}^{q}(D) equivariant map, i.e., F(hφ)=F(h)φF(h\circ\varphi)=F(h)\circ\varphi for all hHq(D,T𝒩)h\in H^{q}(D,T\mathcal{N}) and φ𝒟q(D)\varphi\in\mathcal{D}^{q}(D). Then FF is a smooth map from VHlocq+l(Do,T𝒩)V\cap H_{\operatorname{loc}}^{q+l}(D^{o},T\mathcal{N}) to itself for any ll\in\mathbb{N}, where DoD^{o} is the interior of DD.

Proof.

For D=S1D=S^{1} this result is shown in [13, Corollary 4.1]. For the case D=[0,2π]D=[0,2\pi] the proof is essentially the same, see also the arguments of Ebin and Marsden [20, Theorem 12.1, Lemma 12.2] who proved the original no-loss-no-gain results for manifolds with boundary. ∎

Proof of Theorem 3.8.

For closed curves, i.e., D=S1D=S^{1}, this result can be found in [10, Theorem 4.4], see also [28, 6]. In the following we will focus on the case of open curves, where the proof will be slightly more involved due to the existence of a boundary. For a strong Riemannian metric (q=nq=n) the existence of the geodesic equation and its local well-posedness is always guaranteed, see, e.g., [24, VIII, Theorem 4.2]. Thus we obtain the first part of the theorem for the Sobolev metric of order n2n\geq 2 on n([0,2π],𝒩)\mathcal{I}^{n}([0,2\pi],\mathcal{N}) by Theorem 3.7. For qnq\neq n we have to prove the well-posedness by hand. In the follwoing we will assume that n=1n=1; the proof for n>1n>1 follows similarly. Following the seminal method of Ebin and Marsden [20] we will show that the geodesic spray, as derived in Lemma 3.3, extends to a smooth vector field on the Sobolev completion, which will allow us to apply Cauchy’s theorem to conclude the local well-posedness of the equation. To this end, we need the following statement regarding the invertibility of the operator AcA_{c} under Neumann boundary conditions:

Claim: Let fHqr([0,2π],T𝒩)f\in H^{r}_{\mathcal{I}^{q}}([0,2\pi],T\mathcal{N}) and cq([0,2π],𝒩)c\in\mathcal{I}^{q}([0,2\pi],\mathcal{N}) with q2r0q-2\geq r\geq 0 and πf=c\pi\circ f=c. Then the boundary value problem

(A.3) Acu(θ)=f(θ),θu(0)=u0,θu(2π)=u1A_{c}u(\theta)=f(\theta)\;,\qquad\nabla_{\partial_{\theta}}u(0)=u_{0},\;\nabla_{\partial_{\theta}}u(2\pi)=u_{1}

has a unique solution uHqr+2([0,2π],T𝒩)u\in H^{r+2}_{\mathcal{I}^{q}}([0,2\pi],T\mathcal{N}), with πu=c\pi\circ u=c.

Note that by subtracting any HqH^{q} section that satisfy the boundary conditions, we can assume that the boundary conditions are homogeneous. Then, a weak form of this equation is simply Gc(u,w)=02πg(f,w)dθG_{c}(u,w)=\int_{0}^{2\pi}g(f,w)\,\,\mathrm{d}\theta for every wH1([0,2π],cT𝒩)w\in H^{1}([0,2\pi],c^{*}T\mathcal{N}). Since cc is fixed, GcG_{c} is equivalent to the standard H1H^{1} norm on H1([0,2π],cT𝒩)H^{1}([0,2\pi],c^{*}T\mathcal{N}). By the Lax-Milgram theorem, there exists a unique solution uH1u_{*}\in H^{1}. We can then consider the equation Acu(θ)=f(θ)A_{c}u(\theta)=f(\theta) with initial conditions u(0)=u(0)u(0)=u_{*}(0), θu(0)=0\nabla_{\partial_{\theta}}u(0)=0. By moving to the weak form again, it follows that the solution for this initial value problem must be uu_{*}, and so its regularity follows from standard initial-value ODE theory. This completes the proof of this claim.

To apply this theorem to the geodesic equation we need to observe that for any fixed time tt the boundary terms of the geodesic equation can be rewritten to yield exactly Neumann conditions for the system

Ac(tct)=((tAc)ctg(v,sct)Acct12Ψc(ct,ct)svg(sct,Acct)v+a1(c)(ct,sct)v),A_{c}(\nabla_{\partial_{t}}c_{t})=\bigg{(}-(\nabla_{\partial_{t}}A_{c})c_{t}-g(v,\nabla_{\partial_{s}}c_{t})A_{c}c_{t}-\frac{1}{2}\Psi_{c}(c_{t},c_{t})\nabla_{\partial_{s}}v\\ -g(\nabla_{\partial_{s}}c_{t},A_{c}c_{t})v+a_{1}(\ell_{c})\mathcal{R}(c_{t},\nabla_{\partial_{s}}c_{t})v\bigg{)},\;\\

where (tAc)=tAcAct(\nabla_{\partial_{t}}A_{c})=\nabla_{\partial_{t}}\circ A_{c}-A_{c}\circ\nabla_{\partial_{t}}, which is an operator of order 2. In addition we have the boundary conditions

θtct|θ=0\displaystyle\nabla_{\theta}\nabla_{\partial_{t}}c_{t}\bigg{|}_{\theta=0} =F0(c,ct)\displaystyle=F_{0}(c,c_{t})\in\mathbb{R}
θtct|θ=2π\displaystyle\nabla_{\theta}\nabla_{\partial_{t}}c_{t}\bigg{|}_{\theta=2\pi} =F1(c,ct).\displaystyle=F_{1}(c,c_{t})\in\mathbb{R}\;.

where F0F_{0} and F1F_{1} can be calculated by applying the product formula for differentiation and the formula for swapping covariant derivatives to the boundary conditions in Lemma 3.3.

Thus by the claim above we can invert AcA_{c} to rewrite the geodesic equation as

tct=Ac1((tAc)ctg(v,sct)Acct12Ψc(ct,ct)svg(sct,Acct)v+a1(c)(ct,sct)v).\nabla_{\partial_{t}}c_{t}=A_{c}^{-1}\bigg{(}-(\nabla_{\partial_{t}}A_{c})c_{t}-g(v,\nabla_{\partial_{s}}c_{t})A_{c}c_{t}-\frac{1}{2}\Psi_{c}(c_{t},c_{t})\nabla_{\partial_{s}}v\\ -g(\nabla_{\partial_{s}}c_{t},A_{c}c_{t})v+a_{1}(\ell_{c})\mathcal{R}(c_{t},\nabla_{\partial_{s}}c_{t})v\bigg{)}\;.

The right hand side of this equation defines a smooth mapping

Φ:Tq(D,𝒩)Tq(D,𝒩),\Phi:T\mathcal{I}^{q}(D,\mathcal{N})\to T\mathcal{I}^{q}(D,\mathcal{N}),

where the smoothness of Φ\Phi follows directly by counting derivatives, using the Sobolev embedding theorem and the result that AcA_{c} and thus also (tAc)(\nabla_{\partial_{t}}A_{c}) and Ac1A_{c}^{-1} are smooth. Thus we have interpreted the geodesic equation as an ODE (in tt) on a Banach space of functions. From here the proof of item 1 of Theorem 3.8 follows directly as in [10, Theorem 4.4] and reduces to an application of the Cauchy theorem and the equivalence of fiber-wise quadratic smooth mappings Φ:Tq(D,𝒩)Tq(D,𝒩)\Phi\colon T\mathcal{I}^{q}(D,\mathcal{N})\to T\mathcal{I}^{q}(D,\mathcal{N}) and smooth sprays S:Tq(D,𝒩)TTq(D,𝒩)S\colon T\mathcal{I}^{q}(D,\mathcal{N})\to TT\mathcal{I}^{q}(D,\mathcal{N}).

To prove item 2 of Theorem 3.8, we use Lemma A.1, for FF the exponential map GG on q(D,𝒩)\mathcal{I}^{q}(D,\mathcal{N}), and VHqq(D,T𝒩)V\subset H^{q}_{\mathcal{I}^{q}}(D,T\mathcal{N}) a neighborhood of the zero section on which the exponential map is defined. It follows that the domain of existence of the geodesic equation (in tt) and the neighborhoods for the exponential mapping are uniform in the Sobolev exponential l𝒩l\in\mathcal{N} and thus the result continues to hold on locq+l(D,𝒩)\mathcal{I}^{q+l}_{\textup{loc}}(D,\mathcal{N}) and therefore also locally in the smooth category. ∎

Appendix B Holonomy estimates: proof of Lemma 4.2

We now prove the Sobolev estimates for manifolds-valued curves as stated in Lemma 4.2. We start be proving some geometric estimates, culminating in bounds on the holonomy along a closed curve (Proposition B.3). The settings for the geometric estimates is as follows:

Let (𝒩,g)(\mathcal{N},g) be a complete Riemannian manifold of finite dimension, with bounded sectional curvature, |K|K𝒩|K|\leq K_{\mathcal{N}} and positive injectivity radius inj𝒩>0\operatorname{inj}_{\mathcal{N}}>0. We denote by \mathcal{R} the Riemann curvature of gg.

Let c:[0,a]𝒩c:[0,a]\to\mathcal{N} be a curve, and let VV be a vector field along cc. Let Πθ1θ2:Tc(θ1)𝒩Tc(θ2)𝒩\Pi_{\theta_{1}}^{\theta_{2}}:T_{c(\theta_{1})}\mathcal{N}\to T_{c(\theta_{2})}\mathcal{N} be the parallel transport operator along cc, and Ddθ\frac{D}{d\theta} the covariant derivative along cc.

Lemma B.1.
|V(a)Π0aV(0)|0a|DdθV(θ)|dθ.\left|V(a)-\Pi_{0}^{a}V(0)\right|\leq\int_{0}^{a}\left|\frac{D}{d\theta}V(\theta)\right|\,\,\mathrm{d}\theta.
Proof.

Define f(θ)=ΠθaV(θ)Π0aV(0)f(\theta)=\Pi_{\theta}^{a}V(\theta)-\Pi_{0}^{a}V(0). Our goal is to bound |f(a)||f(a)|. Note that f(0)=0f(0)=0, and that

θf(θ)=ΠθaDdθV(θ).\frac{\partial}{\partial\theta}f(\theta)=\Pi_{\theta}^{a}\frac{D}{d\theta}V(\theta).

Therefore, using the fact that the parallel transport is an isometry, we have

|f(a)|=|0aθf(θ)dθ|0a|DdθV(θ)|dθ.|f(a)|=\left|\int_{0}^{a}\frac{\partial}{\partial\theta}f(\theta)\,\,\mathrm{d}\theta\right|\leq\int_{0}^{a}\left|\frac{D}{d\theta}V(\theta)\right|\,\,\mathrm{d}\theta.

Let c:[0,a]𝒩c:[0,a]\to\mathcal{N} be a closed curve, c(0)=c(a)=pc(0)=c(a)=p, with c<2inj𝒩\ell_{c}<2\operatorname{inj}_{\mathcal{N}}. Define a map c(θ,t):[0,a]×[0,1]𝒩c(\theta,t):[0,a]\times[0,1]\to\mathcal{N}, such that c(θ,)c(\theta,\cdot) is the unique geodesic connecting pp and c(θ)c(\theta). This is well defined since c<2inj𝒩\ell_{c}<2\operatorname{inj}_{\mathcal{N}} implies that dist(p,c(θ))<c/2<inj𝒩\operatorname{dist}(p,c(\theta))<\ell_{c}/2<\operatorname{inj}_{\mathcal{N}} for any θ\theta. In other words, if we define γ(θ)=expp1(c(θ))\gamma(\theta)=\exp_{p}^{-1}(c(\theta)), then c(θ,t):=expp(tγ(θ))c(\theta,t):=\exp_{p}(t\gamma(\theta)). For every t0[0,1]t_{0}\in[0,1], ct0:=c(,t0):[0,a]𝒩c^{t_{0}}:=c(\cdot,t_{0}):[0,a]\to\mathcal{N} is a closed curve based in pp, and for t0=0t_{0}=0 it is the constant curve.

Lemma B.2.

There exists a constant C1C_{1}, depending only on inj𝒩\operatorname{inj}_{\mathcal{N}} and the upper bound for the sectional curvature of 𝒩\mathcal{N}, such that if the curve cc satisfies c<C1\ell_{c}<C_{1}, then ctc\ell_{c^{t}}\leq\ell_{c} for every t[0,1]t\in[0,1].

Proof.

In the following we will assume that c<2inj𝒩\ell_{c}<2\operatorname{inj}_{\mathcal{N}}, otherwise the family ctc^{t} is not well-defined.

It is obviously sufficient to prove that |θc(θ,t)||θc(θ,1)||\partial_{\theta}c(\theta,t)|\leq|\partial_{\theta}c(\theta,1)| for every θ\theta and tt. Note that for a fixed θ0\theta_{0}, J(t):=θc(θ0,t)J(t):=\partial_{\theta}c(\theta_{0},t) is a Jacobi field, hence it satisfies the Jacobi equation

D2dt2J+(J,tc(θ0,t))tc(θ0,t)=0,\frac{D^{2}}{dt^{2}}J+\mathcal{R}\left(J,\partial_{t}c(\theta_{0},t)\right)\partial_{t}c(\theta_{0},t)=0,

with the initial conditions

J(0)=0,DdtJ(0)=θ|θ=θ0expp1c(θ)=:γ(θ0).J(0)=0,\qquad\frac{D}{dt}J(0)=\left.\frac{\partial}{\partial\theta}\right|_{\theta=\theta_{0}}\exp_{p}^{-1}c(\theta)=:\gamma^{\prime}(\theta_{0}).

These initial conditions follow from the fact that

J(0)\displaystyle J(0) =θ|θ=θ0c(θ,0)=θ|θ=θ0p=0,\displaystyle=\left.\frac{\partial}{\partial\theta}\right|_{\theta=\theta_{0}}c(\theta,0)=\left.\frac{\partial}{\partial\theta}\right|_{\theta=\theta_{0}}p=0,
DdtJ(0)\displaystyle\frac{D}{dt}J(0) =Dtθc|(θ,t)=(θ0,0)=Dθtc|(θ,t)=(θ0,0)\displaystyle=\left.\frac{D}{\partial t}\frac{\partial}{\partial\theta}c\right|_{(\theta,t)=(\theta_{0},0)}=\left.\frac{D}{\partial\theta}\frac{\partial}{\partial t}c\right|_{(\theta,t)=(\theta_{0},0)}
=Dθdtγ(θ)expp[γ(θ)]|(θ,t)=(θ0,0)=Dθd0expp[γ(θ)]|θ=θ0\displaystyle=\left.\frac{D}{\partial\theta}d_{t\gamma(\theta)}\exp_{p}[\gamma(\theta)]\right|_{(\theta,t)=(\theta_{0},0)}=\left.\frac{D}{\partial\theta}d_{0}\exp_{p}[\gamma(\theta)]\right|_{\theta=\theta_{0}}
=Dθγ(θ)|θ=θ0=γ(θ0),\displaystyle=\left.\frac{D}{\partial\theta}\gamma(\theta)\right|_{\theta=\theta_{0}}=\gamma^{\prime}(\theta_{0}),

where we used the fact that d0expp=idTp𝒩d_{0}\exp_{p}=\mathrm{id}_{T_{p}\mathcal{N}}, and that when t=0t=0, c(θ,0)=pc(\theta,0)=p for all θ\theta, hence covariant derivative along θ\theta is the same as the regular derivative in idTp𝒩\mathrm{id}_{T_{p}\mathcal{N}}.

Note that we can always reparametrize θ\theta such that |γ(θ)|=1|\gamma^{\prime}(\theta)|=1 for any θ\theta, hence |DdtJ(0)|=1\left|\frac{D}{dt}J(0)\right|=1.

Our aim is to prove that |J(t)||J(1)||J(t)|\leq|J(1)|. The proof mimics the proof of Rauch’s comparison theorem. Define f(t):=|J(t)|f(t):=|J(t)|; we want to prove that f˙(t)0\dot{f}(t)\geq 0 for t(0,1)t\in(0,1). For brevity, write J˙:=DdtJ\dot{J}:=\frac{D}{dt}J, J¨:=D2dt2J\ddot{J}:=\frac{D^{2}}{dt^{2}}J. We then have

f˙=g(J,J˙)|J|2.\dot{f}=\frac{g(J,\dot{J})}{|J|^{2}}.

We have J(0)=0J(0)=0 and therefore, by the Jacobi equations, also J¨(0)=0\ddot{J}(0)=0. We therefore obtain that

J˙(t)=J˙(0)+O(t2),J(t)=tJ˙(0)+O(t3),\dot{J}(t)=\dot{J}(0)+O(t^{2}),\qquad J(t)=t\dot{J}(0)+O(t^{3}),

hence

f˙(t)=1t+O(t).\dot{f}(t)=\frac{1}{t}+O(t).

Using the Jacobi equations and the upper bound KK on the sectional curvature of 𝒩\mathcal{N}, we obtain

f¨=(|J˙|2+g(J,J¨))|J|22g(J,J˙)2|J|4g(J,J¨)|J|2g(J,J˙)2|J|4=g(J,J¨)|J|2f˙2=g(J,(J,tc)tc)|J|2f˙2K|tc|2|J|2)|J|2f˙2=K|tc|2f˙2Kdist𝒩2(p,c(θ0))f˙2Kc24f˙2,\begin{split}\ddot{f}&=\frac{\left(|\dot{J}|^{2}+g(J,\ddot{J})\right)|J|^{2}-2g(J,\dot{J})^{2}}{|J|^{4}}\\ &\geq\frac{g(J,\ddot{J})}{|J|^{2}}-\frac{g(J,\dot{J})^{2}}{|J|^{4}}=\frac{g(J,\ddot{J})}{|J|^{2}}-\dot{f}^{2}\\ &=-\frac{g(J,\mathcal{R}(J,\partial_{t}c)\partial_{t}c)}{|J|^{2}}-\dot{f}^{2}\\ &\geq-\frac{K|\partial_{t}c|^{2}|J|^{2})}{|J|^{2}}-\dot{f}^{2}=-K|\partial_{t}c|^{2}-\dot{f}^{2}\\ &\geq-K\operatorname{dist}_{\mathcal{N}}^{2}(p,c(\theta_{0}))-\dot{f}^{2}\geq-\frac{K\ell_{c}^{2}}{4}-\dot{f}^{2},\end{split}

where we used the fact that |tc(θ0,t)|=dist𝒩2(p,c(θ0))|\partial_{t}c(\theta_{0},t)|=\operatorname{dist}_{\mathcal{N}}^{2}(p,c(\theta_{0})) since c(θ0,t)c(\theta_{0},t) is a constant speed geodesic from pp to c(θ0)c(\theta_{0}). We obtain that

f¨+f˙2Kc24,f˙(t)=1t+O(t).\ddot{f}+\dot{f}^{2}\geq-\frac{K\ell_{c}^{2}}{4},\qquad\dot{f}(t)=\frac{1}{t}+O(t).

From the Riccati comparison estimate [31, Corollary 6.4.2], it follows that for t>0t>0 we have

f˙(t){Kc2cot(Kc2t)K>0,t2πKctK=0Kc2coth(Kc2t)K<0.\dot{f}(t)\geq\begin{cases}\frac{\sqrt{K}\ell_{c}}{2}\cot\left(\frac{\sqrt{K}\ell_{c}}{2}t\right)&K>0,\,t\leq\frac{2\pi}{\sqrt{K}\ell_{c}}\\ t&K=0\\ \frac{\sqrt{-K}\ell_{c}}{2}\coth\left(\frac{\sqrt{-K}\ell_{c}}{2}t\right)&K<0.\end{cases}

If K0K\leq 0, it follows that f˙(t)>0\dot{f}(t)>0 for any t>0t>0, and we are done. If K>0K>0, then by choosing c<π/K\ell_{c}<\pi/\sqrt{K}, we obtain that f˙(t)\dot{f}(t) is larger than a function that is positive in (0,1](0,1]. ∎

We now state the main geometric estimate we need. Recall that, in two dimensions, the holonomy of a small closed curve is roughly the area enclosed by the curve times the curvature inside it, and that by the isoperimetric inequality, the area grows at most like the length of the curve squared. The following proposition combines these statements (in any dimension) into a quantitative estimate on the holonomy:

Proposition B.3.

There exists a constant C=C(K𝒩,inj𝒩,dim𝒩)>0C=C(K_{\mathcal{N}},\operatorname{inj}_{\mathcal{N}},\dim\mathcal{N})>0, such that for every closed curve c𝒩c\subset\mathcal{N} based in Tp𝒩T_{p}\mathcal{N},

|HolcidTp𝒩|min{Cc2, 2dim𝒩}\left|\operatorname{Hol}_{c}-\mathrm{id}_{T_{p}\mathcal{N}}\right|\leq\min\left\{C\ell_{c}^{2}\,,\,2\sqrt{\dim\mathcal{N}}\right\}

where Holc\operatorname{Hol}_{c} is the holonomy along cc and c\ell_{c} is the length of the curve cc.

Proof.

Since Holc\operatorname{Hol}_{c} is an isometry of Tp𝒩T_{p}\mathcal{N}, |Holc|=|idTp𝒩|=dim𝒩|\operatorname{Hol}_{c}|=|\mathrm{id}_{T_{p}\mathcal{N}}|=\sqrt{\dim\mathcal{N}}. Therefore, by triangle inequality, we have |HolcidTp𝒩|2dim𝒩\left|\operatorname{Hol}_{c}-\mathrm{id}_{T_{p}\mathcal{N}}\right|\leq 2\sqrt{\dim\mathcal{N}}.

In the following, we assume that C2dim𝒩/C12C\geq 2\sqrt{\dim\mathcal{N}}/C_{1}^{2}, where C1C_{1} is defined in Lemma B.2, and therefore it is sufficient to prove that |HolcidTp𝒩|Cc2\left|\operatorname{Hol}_{c}-\mathrm{id}_{T_{p}\mathcal{N}}\right|\leq C\ell_{c}^{2} under the assumption that cC1\ell_{c}\leq C_{1}.

Fix a unit vector vTp𝒩v\in T_{p}\mathcal{N}. Our goal is to prove that |Holcvv|Cc2\left|\operatorname{Hol}_{c}v-v\right|\leq C\ell_{c}^{2}. Define the family of curves c(θ,t)=ct(θ):[0,a]×[0,1]𝒩c(\theta,t)=c^{t}(\theta):[0,a]\times[0,1]\to\mathcal{N} as in Lemma B.2. Define a vector field XΓ(cT𝒩)X\in\Gamma(c^{*}T\mathcal{N}) by

X(θ,t):=Πpct(θ)v,X(\theta,t):=\Pi_{p}^{c^{t}(\theta)}v,

where Πpct(θ):Tp𝒩Tct(θ)𝒩\Pi_{p}^{c^{t}(\theta)}:T_{p}\mathcal{N}\to T_{c^{t}(\theta)}\mathcal{N} is the parallel transport along the curve ctc^{t}. We have

X(θ,0)=v,X(0,t)=v,X(a,t)=Holctv.X(\theta,0)=v,\quad X(0,t)=v,\quad X(a,t)=\operatorname{Hol}_{c^{t}}v.

Since c(a,t)=pc(a,t)=p, the parallel transport along the curve c(a,)c(a,\cdot) is the identity, and so, by Lemma B.1 we have that

|Holcvv|=|X(a,1)X(a,0)|01|DtX(a,t)|dt.\left|\operatorname{Hol}_{c}v-v\right|=\left|X(a,1)-X(a,0)\right|\leq\int_{0}^{1}\left|\frac{D}{\partial t}X(a,t)\right|\,\,\mathrm{d}t.

Since c(0,t)=pc(0,t)=p for all tt, the covariant derivative Dt\frac{D}{\partial t} along (0,t)(0,t) is simply the standard derivative t\frac{\partial}{\partial t}. Therefore, since X(0,t)=vX(0,t)=v does not depend on tt, we have DtX(0,t)=0\frac{D}{\partial t}X(0,t)=0. Hence, using Lemma B.1 again, we have

|DtX(a,t)|0a|DθDtX(θ,t)|dθ.\left|\frac{D}{\partial t}X(a,t)\right|\leq\int_{0}^{a}\left|\frac{D}{\partial\theta}\frac{D}{\partial t}X(\theta,t)\right|\,\,\mathrm{d}\theta.

Since X(θ,t)X(\theta,t) is the parallel transport of X(0,t)=vX(0,t)=v along the constant tt curve, we have DθX(θ,t)=0\frac{D}{\partial\theta}X(\theta,t)=0, and therefore DtDθX(θ,t)=0\frac{D}{\partial t}\frac{D}{\partial\theta}X(\theta,t)=0. Combining this with

DθDtXDtDθX=(θt)X\frac{D}{\partial\theta}\frac{D}{\partial t}X-\frac{D}{\partial t}\frac{D}{\partial\theta}X=\mathcal{R}\left(\frac{\partial}{\partial\theta}\frac{\partial}{\partial t}\right)X

(see, e.g., [17, Chapter 4, Lemma 4.1]), we have

|DθDtX|=|(cθct)X|K𝒩|θ||t|,\left|\frac{D}{\partial\theta}\frac{D}{\partial t}X\right|=\left|\mathcal{R}\left(\frac{\partial c}{\partial\theta}\frac{\partial c}{\partial t}\right)X\right|\leq K_{\mathcal{N}}\left|\frac{\partial}{\partial\theta}\right|\,\left|\frac{\partial}{\partial t}\right|,

where we used the fact that |X|=|v|=1|X|=|v|=1 since the parallel transport is an isometry. Since c(θ,)c(\theta,\cdot) is a constant speed geodesic from pp to c(θ)=c(θ,1)c(\theta)=c(\theta,1), and that dist(p,c(θ))c/2\operatorname{dist}(p,c(\theta))\leq\ell_{c}/2, we have that

|ct|c/2.\left|\frac{\partial c}{\partial t}\right|\leq\ell_{c}/2.

Combining these estimates, we obtain

|Holcvv|01|DtX(a,t)|dt010a|DθDtX|dθdtK𝒩c2010a|cθ|dθdt=K𝒩c201ctdtK𝒩2c2,\begin{split}\left|\operatorname{Hol}_{c}v-v\right|&\leq\int_{0}^{1}\left|\frac{D}{\partial t}X(a,t)\right|\,\,\mathrm{d}t\leq\int_{0}^{1}\int_{0}^{a}\left|\frac{D}{\partial\theta}\frac{D}{\partial t}X\right|\,\,\mathrm{d}\theta\,\,\mathrm{d}t\\ &\leq K_{\mathcal{N}}\frac{\ell_{c}}{2}\int_{0}^{1}\int_{0}^{a}\left|\frac{\partial c}{\partial\theta}\right|\,\,\mathrm{d}\theta\,\,\mathrm{d}t=K_{\mathcal{N}}\frac{\ell_{c}}{2}\int_{0}^{1}\ell_{c^{t}}\,\,\mathrm{d}t\leq\frac{K_{\mathcal{N}}}{2}\ell_{c}^{2},\end{split}

where in the last inequality we used Lemma B.2 to estimate ct\ell_{c^{t}}. ∎

Using these holonomy estimates, we can now prove Lemma 4.2:

Proof of Lemma 4.2. As mentioned at the beginning of Section 4, although h(0)=h(2π)h(0)=h(2\pi) when D=S1D=S^{1}, it is not true that H(0)=H(2π)H(0)=H(2\pi), where

H(θ)=Πθ0h(θ),H(\theta)=\Pi_{\theta}^{0}h(\theta),

because of holonomy effects. Therefore, in order to prove (4.4) we cannot use Sobolev inequalities for periodic functions verbatim, but rather use the result of Proposition B.3, which implies that for short curves HH is “almost” periodic since the holonomy is small. We will do so by induction over kk and nn.

Base step: the case k=1k=1, n=2n=2. Assume that k=1k=1 and n=2n=2. When c1\ell_{c}\geq 1, the inequality (4.2) implies (4.4) by taking a=1a=1. We are left with the case c<1\ell_{c}<1.

Recall that we denote by Πθ1θ2\Pi_{\theta_{1}}^{\theta_{2}} the parallel transport from Tc(θ1)𝒩T_{c(\theta_{1})}\mathcal{N} to Tc(θ2)𝒩T_{c(\theta_{2})}\mathcal{N} along cc (in the direction dictated by the parameter θ\theta). Now, by applying (4.1) for sh\nabla_{\partial_{s}}h and using the fundamental theorem of calculus, we have:

Πθ0sh(θ)sh(0)=0θddσΠσ0sh(σ)dσ=0θΠσ0(θsh(σ))dσ.\Pi_{\theta}^{0}\nabla_{\partial_{s}}h(\theta)-\nabla_{\partial_{s}}h(0)=\int_{0}^{\theta}\frac{d}{d\sigma}\Pi_{\sigma}^{0}\nabla_{\partial_{s}}h(\sigma)\,\,\mathrm{d}\sigma=\int_{0}^{\theta}\Pi_{\sigma}^{0}(\nabla_{\partial_{\theta}}\nabla_{\partial_{s}}h(\sigma))\,\,\mathrm{d}\sigma.

Integrating over θ\theta with respect to ds\,\mathrm{d}s, we obtain

sh(0)1cS1Πθ0sh(θ)ds(θ)=1cS10θΠσ0(θsh(σ))dσds(θ).\nabla_{\partial_{s}}h(0)-\frac{1}{\ell_{c}}\int_{S^{1}}\Pi_{\theta}^{0}\nabla_{\partial_{s}}h(\theta)\,\,\mathrm{d}s(\theta)=-\frac{1}{\ell_{c}}\int_{S^{1}}\int_{0}^{\theta}\Pi_{\sigma}^{0}(\nabla_{\partial_{\theta}}\nabla_{\partial_{s}}h(\sigma))\,\,\mathrm{d}\sigma\,\,\mathrm{d}s(\theta).

Using again (4.1), we have

S1Πθ0sh(θ)ds(θ)=02πΠθ0θh(θ)dθ=02πddθΠθ0h(θ)dθ=Π2π0h(0)h(0),\int_{S^{1}}\Pi_{\theta}^{0}\nabla_{\partial_{s}}h(\theta)\,\,\mathrm{d}s(\theta)=\int_{0}^{2\pi}\Pi_{\theta}^{0}\nabla_{\partial_{\theta}}h(\theta)\,\,\mathrm{d}\theta\\ =\int_{0}^{2\pi}\frac{d}{d\theta}\Pi_{\theta}^{0}h(\theta)\,\,\mathrm{d}\theta=\Pi_{2\pi}^{0}h(0)-h(0),

which is not necessarily zero since there the holonomy along cc might be non-trivial. We therefore obtain

(B.1) sh(0)1c(Π2π0h(0)h(0))=1cS10θΠσ0(θsh(σ))dσds(θ).\nabla_{\partial_{s}}h(0)-\frac{1}{\ell_{c}}\left(\Pi_{2\pi}^{0}h(0)-h(0)\right)\\ =-\frac{1}{\ell_{c}}\int_{S^{1}}\int_{0}^{\theta}\Pi_{\sigma}^{0}(\nabla_{\partial_{\theta}}\nabla_{\partial_{s}}h(\sigma))\,\,\mathrm{d}\sigma\,\,\mathrm{d}s(\theta).

Similarly,

sh(2π)Πθ2πsh(θ)=θ2πddσΠσ2πsh(σ)dσ=θ2πΠσ2π(θsh(σ))dσ,\nabla_{\partial_{s}}h(2\pi)-\Pi_{\theta}^{2\pi}\nabla_{\partial_{s}}h(\theta)=\int_{\theta}^{2\pi}\frac{d}{d\sigma}\Pi_{\sigma}^{2\pi}\nabla_{\partial_{s}}h(\sigma)\,\,\mathrm{d}\sigma\\ =\int_{\theta}^{2\pi}\Pi_{\sigma}^{2\pi}(\nabla_{\partial_{\theta}}\nabla_{\partial_{s}}h(\sigma))\,\,\mathrm{d}\sigma,

and

S1Πθ2πsh(θ)ds(θ)=02πddθΠθ2πh(θ)dθ=h(2π)Π02πh(2π).\int_{S^{1}}\Pi_{\theta}^{2\pi}\nabla_{\partial_{s}}h(\theta)\,\,\mathrm{d}s(\theta)=\int_{0}^{2\pi}\frac{d}{d\theta}\Pi_{\theta}^{2\pi}h(\theta)\,\,\mathrm{d}\theta=h(2\pi)-\Pi_{0}^{2\pi}h(2\pi).

Thus, using the fact that h(0)=h(2π)h(0)=h(2\pi) and sh(0)=sh(2π)\nabla_{\partial_{s}}h(0)=\nabla_{\partial_{s}}h(2\pi), we have

(B.2) sh(0)1c(h(0)Π02πh(0))=1cS1θ2πΠσ2π(θsh(σ))dσds(θ).\nabla_{\partial_{s}}h(0)-\frac{1}{\ell_{c}}\left(h(0)-\Pi_{0}^{2\pi}h(0)\right)\\ =\frac{1}{\ell_{c}}\int_{S^{1}}\int_{\theta}^{2\pi}\Pi_{\sigma}^{2\pi}(\nabla_{\partial_{\theta}}\nabla_{\partial_{s}}h(\sigma))\,\,\mathrm{d}\sigma\,\,\mathrm{d}s(\theta).

Adding (B.1) and (B.2), we obtain

sh(0)1c(Π2π0h(0)Π02πh(0))=12cS1(θ2πΠσ2π(θsh(σ))dσ0θΠσ0(θsh(σ))dσ)ds(θ).\begin{split}&\nabla_{\partial_{s}}h(0)-\frac{1}{\ell_{c}}\left(\Pi_{2\pi}^{0}h(0)-\Pi_{0}^{2\pi}h(0)\right)\\ &=\frac{1}{2\ell_{c}}\int_{S^{1}}\left(\int_{\theta}^{2\pi}\Pi_{\sigma}^{2\pi}(\nabla_{\partial_{\theta}}\nabla_{\partial_{s}}h(\sigma))\,\,\mathrm{d}\sigma-\int_{0}^{\theta}\Pi_{\sigma}^{0}(\nabla_{\partial_{\theta}}\nabla_{\partial_{s}}h(\sigma))\,\mathrm{d}\sigma\right)\,\,\mathrm{d}s(\theta).\end{split}

Therefore, using the fact that Πθ1θ2\Pi_{\theta_{1}}^{\theta_{2}} is an isometry, we obtain that

|sh(0)||Π2π0Π02π|c|h(0)|+12cS102π|θsh(σ)|dσds(θ)=|Π2π0Π02π|c|h(0)|+12S1|s2h(σ)|ds(σ).\begin{split}|\nabla_{\partial_{s}}h(0)|&\leq\frac{\left|\Pi_{2\pi}^{0}-\Pi_{0}^{2\pi}\right|}{\ell_{c}}|h(0)|+\frac{1}{2\ell_{c}}\int_{S^{1}}\int_{0}^{2\pi}\left|\nabla_{\partial_{\theta}}\nabla_{\partial_{s}}h(\sigma)\right|\,\mathrm{d}\sigma\,\,\mathrm{d}s(\theta)\\ &=\frac{\left|\Pi_{2\pi}^{0}-\Pi_{0}^{2\pi}\right|}{\ell_{c}}|h(0)|+\frac{1}{2}\int_{S^{1}}\left|\nabla_{\partial_{s}}^{2}h(\sigma)\right|\,\mathrm{d}s(\sigma).\end{split}

Using the estimate on the magnitude of the holonomy in Proposition B.3, we have

|sh(0)|min{Cc,2dim𝒩c}|h(0)|+12S1|s2h(σ)|ds,|\nabla_{\partial_{s}}h(0)|\leq\min\left\{C\ell_{c},\frac{2\sqrt{\dim\mathcal{N}}}{\ell_{c}}\right\}|h(0)|+\frac{1}{2}\int_{S^{1}}\left|\nabla_{\partial_{s}}^{2}h(\sigma)\right|\,\mathrm{d}s,

for some C>0C>0 that depends only on the injectivity radius and on the bounds on the sectional curvature of 𝒩\mathcal{N}. In this inequality the point 0 is arbitrary, hence the above holds for h(θ),sh(θ)h(\theta),\nabla_{\partial_{s}}h(\theta) instead of h(0),sh(0)h(0),\nabla_{\partial_{s}}h(0). Squaring this inequality, and using the inequality (a+b)22(a2+b2)(a+b)^{2}\leq 2(a^{2}+b^{2}) and Cauchy-Schwartz (or Jensen’s) inequality, we obtain, for every θ\theta,

(B.3) |sh(θ)|2min{2C2c2,8dim𝒩c2}|h(θ)|2+c2s2hL2(ds)2Cmin{c2,1c2}|h(θ)|2+c2s2hL2(ds)2.\begin{split}|\nabla_{\partial_{s}}h(\theta)|^{2}&\leq\min\left\{2C^{2}\ell_{c}^{2},\frac{8\dim\mathcal{N}}{\ell_{c}^{2}}\right\}|h(\theta)|^{2}+\frac{\ell_{c}}{2}\|\nabla_{\partial_{s}}^{2}h\|_{L^{2}(\,\mathrm{d}s)}^{2}\\ &\leq C^{\prime}\min\left\{\ell_{c}^{2},\frac{1}{\ell_{c}^{2}}\right\}|h(\theta)|^{2}+\frac{\ell_{c}}{2}\|\nabla_{\partial_{s}}^{2}h\|_{L^{2}(\,\mathrm{d}s)}^{2}.\end{split}

Since we assumed c1\ell_{c}\leq 1, Inequality (B.3) implies (4.4) by integrating with respect to ds\,\mathrm{d}s.

Induction step. Now assume we have (4.4) for n=2,,mn=2,\ldots,m and k=1,,n1k=1,\ldots,n-1; we will now prove it for n=m+1n=m+1, k=1,,mk=1,\ldots,m. Denote the constant in (4.4) by Ck,nC_{k,n}. Besides kk and nn, Ck,nC_{k,n} will depend also on the properties of the manifold 𝒩\mathcal{N} as stated in the formulation of the lemma, but we omit this dependence as it is fixed throughout the induction.

First, assume k=1k=1. If cmin{1,(2Cm1,mC1,m)1/2}\ell_{c}\geq\min\left\{1,\left(2C_{m-1,m}C_{1,m}\right)^{-1/2}\right\}, then (4.2) implies (4.4) for k=1,n=m+1k=1,n=m+1, by letting a=min{1,(2Cm1,mC1,m)1/2}a=\min\left\{1,\left(2C_{m-1,m}C_{1,m}\right)^{-1/2}\right\}. If cmin{1,(2Cm1,mC1,m)1/2}\ell_{c}\leq\min\left\{1,\left(2C_{m-1,m}C_{1,m}\right)^{-1/2}\right\}, we have

shL2(ds)2C1,mc2(hL2(ds)2+smhL2(ds)2)C1,mc2(hL2(ds)2+Cm1,m(shL2(ds)2+sm+1hL2(ds)2)),\|\nabla_{\partial_{s}}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\leq C_{1,m}\ell_{c}^{2}\left(\|h\|^{2}_{L^{2}(\,\mathrm{d}s)}+\|\nabla_{\partial_{s}}^{m}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\right)\\ \leq C_{1,m}\ell_{c}^{2}\left(\|h\|^{2}_{L^{2}(\,\mathrm{d}s)}+C_{m-1,m}\left(\|\nabla_{\partial_{s}}h\|^{2}_{L^{2}(\,\mathrm{d}s)}+\|\nabla_{\partial_{s}}^{m+1}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\right)\right),

where in the second line we applied the induction hypothesis to sh\nabla_{\partial_{s}}h. Moving the C1,mCm1,mc2shL2(ds)2C_{1,m}C_{m-1,m}\ell_{c}^{2}\|\nabla_{\partial_{s}}h\|^{2}_{L^{2}(\,\mathrm{d}s)} to the other side, and noting that C1,mCm1,mc21/2C_{1,m}C_{m-1,m}\ell_{c}^{2}\leq 1/2 by assumption, we obtain that

shL2(ds)22C1,mc2(hL2(ds)2+Cm1,msm+1hL2(ds)2),\|\nabla_{\partial_{s}}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\leq 2C_{1,m}\ell_{c}^{2}\left(\|h\|^{2}_{L^{2}(\,\mathrm{d}s)}+C_{m-1,m}\|\nabla_{\partial_{s}}^{m+1}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\right),

which completes the proof for k=1k=1.

We now assume k>1k>1. If cmin{1,(2Ck1,mC1,k)1/2}\ell_{c}\geq\min\left\{1,\left(2C_{k-1,m}C_{1,k}\right)^{-1/2}\right\}, then (4.2) implies (4.4) for k=1,n=m+1k=1,n=m+1, by letting a=min{1,(2Ck1,mC1,k)1/2}a=\min\left\{1,\left(2C_{k-1,m}C_{1,k}\right)^{-1/2}\right\}. If cmin{1,(2Ck1,mC1,k)1/2}\ell_{c}\leq\min\left\{1,\left(2C_{k-1,m}C_{1,k}\right)^{-1/2}\right\}, we have (by applying the induction hypothesis for sh\nabla_{\partial_{s}}h),

skhL2(ds)2Ck1,mc2(shL2(ds)2+sm+1hL2(ds)2)Ck1,mc2(C1,k(hL2(ds)2+skhL2(ds)2)+sm+1hL2(ds)2).\|\nabla_{\partial_{s}}^{k}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\leq C_{k-1,m}\ell_{c}^{2}\left(\|\nabla_{\partial_{s}}h\|^{2}_{L^{2}(\,\mathrm{d}s)}+\|\nabla_{\partial_{s}}^{m+1}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\right)\\ \leq C_{k-1,m}\ell_{c}^{2}\left(C_{1,k}\left(\|h\|^{2}_{L^{2}(\,\mathrm{d}s)}+\|\nabla_{\partial_{s}}^{k}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\right)+\|\nabla_{\partial_{s}}^{m+1}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\right).

Moving the Ck1,mC1,kc2skhL2(ds)2C_{k-1,m}C_{1,k}\ell_{c}^{2}\|\nabla_{\partial_{s}}^{k}h\|^{2}_{L^{2}(\,\mathrm{d}s)} to the other side, and noting that Ck1,mC1,kc21/2C_{k-1,m}C_{1,k}\ell_{c}^{2}\leq 1/2 by assumption, we obtain that

skhL2(ds)22Ck1,mc2(C1,khL2(ds)2+sm+1hL2(ds)2),\|\nabla_{\partial_{s}}^{k}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\leq 2C_{k-1,m}\ell_{c}^{2}\left(C_{1,k}\|h\|^{2}_{L^{2}(\,\mathrm{d}s)}+\|\nabla_{\partial_{s}}^{m+1}h\|^{2}_{L^{2}(\,\mathrm{d}s)}\right),

which completes the proof for k>1k>1.

Appendix C Proof of Lemma 5.12

First, we note that for a function fL2(D)f\in L^{2}(D) we have, for every cn(D,𝒩)c\in\mathcal{I}^{n}(D,\mathcal{N}),

fL2(dθ)|D|1/2fL\|f\|_{L^{2}(\,\mathrm{d}\theta)}\leq|D|^{1/2}\|f\|_{L^{\infty}}

hence boundedness on metric balls of sk|c|L\|\nabla_{\partial_{s}}^{k}|c^{\prime}|\|_{L^{\infty}} implies boundedness of sk|c|L2(dθ)\|\nabla_{\partial_{s}}^{k}|c^{\prime}|\|_{L^{2}(\,\mathrm{d}\theta)}. Lemma 5.9 implies that under the assumption (5.1), the L2(dθ)L^{2}(\,\mathrm{d}\theta) and L2(ds)L^{2}(\,\mathrm{d}s) norms are equivalent on metric balls, hence boundedness on metric balls of sk|c|L\|\nabla_{\partial_{s}}^{k}|c^{\prime}|\|_{L^{\infty}} also implies boundedness of sk|c|L2(ds)\|\nabla_{\partial_{s}}^{k}|c^{\prime}|\|_{L^{2}(\,\mathrm{d}s)}. Therefore, by Lemma 5.7, our goal is to show that

Dc,h(sk|c|)LpC(1+sk|c|Lp)hGc,\|D_{c,h}(\nabla_{\partial_{s}}^{k}|c^{\prime}|)\|_{L^{p}}\leq C(1+\|\nabla_{\partial_{s}}^{k}|c^{\prime}|\|_{L^{p}})\|h\|_{G_{c}},

where p=p=\infty for k=0,,n2k=0,\ldots,n-2 and p=2p=2 for k=n1k=n-1. We will first prove the case p=p=\infty by induction on kk, and then treat the case p=2p=2, k=n1k=n-1 (in which the cases L2(dθ)L^{2}(\,\mathrm{d}\theta) and L2(ds)L^{2}(\,\mathrm{d}s) are similar, so for brevity, we simply write L2L^{2}).

The claim for k=0k=0 was proven in Lemma 5.10. We now assume the claim is true up to k1k-1 and prove it for kk. First, note that

Dc,h(sk|c|)=sk(g(v,sh)|c|)i=0k1(ki+1)sig(v,sh)ski|c|=i=0k((ki)(ki+1))sig(v,sh)ski|c|,\begin{split}D_{c,h}(\nabla_{\partial_{s}}^{k}|c^{\prime}|)&=\nabla_{\partial_{s}}^{k}(g(v,\nabla_{\partial_{s}}h)|c^{\prime}|)-\sum_{i=0}^{k-1}\binom{k}{i+1}\nabla_{\partial_{s}}^{i}g(v,\nabla_{\partial_{s}}h)\nabla_{\partial_{s}}^{k-i}|c^{\prime}|\\ &=\sum_{i=0}^{k}\left(\binom{k}{i}-\binom{k}{i+1}\right)\nabla_{\partial_{s}}^{i}g(v,\nabla_{\partial_{s}}h)\nabla_{\partial_{s}}^{k-i}|c^{\prime}|,\end{split}

where we use the convention (kk+1)=0\binom{k}{k+1}=0. This can be easily proved by induction using (3.6). From this it follows that

(C.1) |Dc,h(sk|c|)|i=0k|sig(v,sh)||ski|c||i=0kj=0i|sjvsij+1hski|c||,\left|D_{c,h}(\nabla_{\partial_{s}}^{k}|c^{\prime}|)\right|\lesssim\sum_{i=0}^{k}\left|\nabla_{\partial_{s}}^{i}g(v,\nabla_{\partial_{s}}h)\right|\left|\nabla_{\partial_{s}}^{k-i}|c^{\prime}|\right|\\ \lesssim\sum_{i=0}^{k}\sum_{j=0}^{i}\left|\nabla_{\partial_{s}}^{j}v\right|\left|\nabla_{\partial_{s}}^{i-j+1}h\right|\left|\nabla_{\partial_{s}}^{k-i}|c^{\prime}|\right|,

where the constant depends only on the indices i,j,ki,j,k. Using the induction hypothesis, we obtain (using the fact that |v|=1|v|=1),

|Dc,h(sk|c|)||sh||sk|c||+i=0kj=0i|sjv||sij+1h|\left|D_{c,h}(\nabla_{\partial_{s}}^{k}|c^{\prime}|)\right|\lesssim\left|\nabla_{\partial_{s}}h\right|\left|\nabla_{\partial_{s}}^{k}|c^{\prime}|\right|+\sum_{i=0}^{k}\sum_{j=0}^{i}\left|\nabla_{\partial_{s}}^{j}v\right|\left|\nabla_{\partial_{s}}^{i-j+1}h\right|

on every metric ball. Our assumption (5.2) implies that for i=1,,n1i=1,\ldots,n-1, we have sihLChGc\|\nabla_{\partial_{s}}^{i}h\|_{L^{\infty}}\leq C\|h\|_{G_{c}} on every metric ball. Therefore, we obtain, as long as kn2k\leq n-2

|Dc,h(sk|c|)|(|sk|c||+j=0k|sjv|)hGc\left|D_{c,h}(\nabla_{\partial_{s}}^{k}|c^{\prime}|)\right|\lesssim\left(\left|\nabla_{\partial_{s}}^{k}|c^{\prime}|\right|+\sum_{j=0}^{k}\left|\nabla_{\partial_{s}}^{j}v\right|\right)\|h\|_{G_{c}}

on every metric ball.

In order to complete the proof (for the LL^{\infty} case), we need to show that

(C.2) skvLk=0,,n2\displaystyle\|\nabla_{\partial_{s}}^{k}v\|_{L^{\infty}}\qquad k=0,\ldots,n-2

is bounded on every metric ball. The case k=0k=0 is trivial, since |v|=1|v|=1 by definition. Note that

Dc,h|skv|=g(hskv,skv|skv|)|hskv|.D_{c,h}|\nabla_{\partial_{s}}^{k}v|=g\left(\nabla_{h}\nabla_{\partial_{s}}^{k}v,\frac{\nabla_{\partial_{s}}^{k}v}{|\nabla_{\partial_{s}}^{k}v|}\right)\leq|\nabla_{h}\nabla_{\partial_{s}}^{k}v|.

Therefore, in order to use Lemma 5.7 for the function |skv||\nabla_{\partial_{s}}^{k}v|, we need to show that

|hskv|C(1+skv)hGc|\nabla_{h}\nabla_{\partial_{s}}^{k}v|\leq C(1+\|\nabla_{\partial_{s}}^{k}v\|_{\infty})\|h\|_{G_{c}}

on every metric ball. Using (3.9), we obtain

hskv=shsk1vg(v,sh)skv+(v,h)sk1v=skhvi=0k1si(g(v,sh)skiv)+i=0k1si((v,h)sk1iv)=sk+1hi=0ksi(g(v,sh)skiv)+i=0k1si((v,h)sk1iv),\begin{split}\nabla_{h}\nabla_{\partial_{s}}^{k}v&=\nabla_{\partial_{s}}\nabla_{h}\nabla_{\partial_{s}}^{k-1}v-g(v,\nabla_{\partial_{s}}h)\nabla_{\partial_{s}}^{k}v+\mathcal{R}(v,h)\nabla_{\partial_{s}}^{k-1}v\\ &=\nabla_{\partial_{s}}^{k}\nabla_{h}v-\sum_{i=0}^{k-1}\nabla_{\partial_{s}}^{i}(g(v,\nabla_{\partial_{s}}h)\nabla_{\partial_{s}}^{k-i}v)+\sum_{i=0}^{k-1}\nabla_{\partial_{s}}^{i}(\mathcal{R}(v,h)\nabla_{\partial_{s}}^{k-1-i}v)\\ &=\nabla_{\partial_{s}}^{k+1}h-\sum_{i=0}^{k}\nabla_{\partial_{s}}^{i}(g(v,\nabla_{\partial_{s}}h)\nabla_{\partial_{s}}^{k-i}v)+\sum_{i=0}^{k-1}\nabla_{\partial_{s}}^{i}(\mathcal{R}(v,h)\nabla_{\partial_{s}}^{k-1-i}v),\end{split}

where in the last line we used the fact that

hv=shg(v,sh)v,\nabla_{h}v=\nabla_{\partial_{s}}h-g(v,\nabla_{\partial_{s}}h)v,

which follows immediately from (3.6). We therefore have,

hskv=sk+1hi=0kj=0il=0j(ij)(jl)g(slv,sjl+1h)skjv+i=0k1j=0il=0jm=0l(ij)(jl)(lm)sjl(smv,slmh)sk1jv,\begin{split}\nabla_{h}\nabla_{\partial_{s}}^{k}v&=\nabla_{\partial_{s}}^{k+1}h-\sum_{i=0}^{k}\sum_{j=0}^{i}\sum_{l=0}^{j}\binom{i}{j}\binom{j}{l}g(\nabla_{\partial_{s}}^{l}v,\nabla_{\partial_{s}}^{j-l+1}h)\nabla_{\partial_{s}}^{k-j}v\\ &\quad+\sum_{i=0}^{k-1}\sum_{j=0}^{i}\sum_{l=0}^{j}\sum_{m=0}^{l}\binom{i}{j}\binom{j}{l}\binom{l}{m}\nabla_{\partial_{s}}^{j-l}\mathcal{R}(\nabla_{\partial_{s}}^{m}v,\nabla_{\partial_{s}}^{l-m}h)\nabla_{\partial_{s}}^{k-1-j}v,\end{split}

where we repeatedly used

s((X,Y)Z)=(s)(X,Y)Z+(sX,Y)Z+(X,sY)Z+(X,Y)sZ.\nabla_{\partial_{s}}\left(\mathcal{R}(X,Y)Z\right)=(\nabla_{\partial_{s}}\mathcal{R})(X,Y)Z+\mathcal{R}(\nabla_{\partial_{s}}X,Y)Z\\ +\mathcal{R}(X,\nabla_{\partial_{s}}Y)Z+\mathcal{R}(X,Y)\nabla_{\partial_{s}}Z.

Using the fact that sr\nabla_{\partial_{s}}^{r}\mathcal{R} is bounded for every rr,888Note that by Lemma 5.11, the whole analysis here is done on a compact subset of 𝒩\mathcal{N} (the closure of the image of B(c0,r)B(c_{0},r)). Hence the boundedness of \mathcal{R} and its covariant derivatives follows from the smoothness of 𝒩\mathcal{N}, and does not require any global bounded geometry assumption on 𝒩\mathcal{N} (except from completeness). we obtain the bound

(C.3) |hskv|\displaystyle|\nabla_{h}\nabla_{\partial_{s}}^{k}v| |sk+1h|+j=0kl=0j|slv||sjl+1h||skjv|\displaystyle\lesssim|\nabla_{\partial_{s}}^{k+1}h|+\sum_{j=0}^{k}\sum_{l=0}^{j}|\nabla_{\partial_{s}}^{l}v|\,|\nabla_{\partial_{s}}^{j-l+1}h|\,|\nabla_{\partial_{s}}^{k-j}v|
(C.4) +j=0k1|sk1jv|l=0jm=0l|smv||slmh|\displaystyle\qquad\qquad+\sum_{j=0}^{k-1}|\nabla_{\partial_{s}}^{k-1-j}v|\sum_{l=0}^{j}\sum_{m=0}^{l}|\nabla_{\partial_{s}}^{m}v|\,|\nabla_{\partial_{s}}^{l-m}h|
(C.5) |sk+1h|+|skv||sh|+i=0kPi|sih|,\displaystyle\lesssim|\nabla_{\partial_{s}}^{k+1}h|+|\nabla_{\partial_{s}}^{k}v||\nabla_{\partial_{s}}h|+\sum_{i=0}^{k}P_{i}|\nabla_{\partial_{s}}^{i}h|,

where PiP_{i} are polynomials in |sv|,,|sk1v||\nabla_{\partial_{s}}v|,\ldots,|\nabla_{\partial_{s}}^{k-1}v|. By the induction hypothesis sjv\|\nabla_{\partial_{s}}^{j}v\|_{\infty} is bounded on metric balls for j=0,,k1j=0,\ldots,k-1, hence PiP_{i} is bounded on metric balls. Using, this, and assumption (5.2), we obtain that, as long as kn2k\leq n-2,

|hskv|(1+|skv|)hGc,|\nabla_{h}\nabla_{\partial_{s}}^{k}v|\lesssim(1+|\nabla_{\partial_{s}}^{k}v|)\|h\|_{G_{c}},

which completes the proof of (C.2) and hence of (5.6).

It remains to prove (5.7) for k=n1k=n-1, that is, to prove that

Dc,h(sn1|c|)L2C(1+sn1|c|L2)hGc.\|D_{c,h}(\nabla_{\partial_{s}}^{n-1}|c^{\prime}|)\|_{L^{2}}\leq C(1+\|\nabla_{\partial_{s}}^{n-1}|c^{\prime}|\|_{L^{2}})\|h\|_{G_{c}}.

Using (C.1) we have

|Dc,h(sn1|c|)|i=0n1j=0i|sjvsij+1hsn1i|c|||snh||c|+hGci=0n1j=0i|sjv||sn1i|c|||snh||c|+hGc(|sn1|c||+|c||sn1v|+i,j=0n2|sjv||si|c||)|snh|+hGc(|sn1|c||+|sn1v|+1)\begin{split}&\left|D_{c,h}(\nabla_{\partial_{s}}^{n-1}|c^{\prime}|)\right|\lesssim\sum_{i=0}^{n-1}\sum_{j=0}^{i}\left|\nabla_{\partial_{s}}^{j}v\right|\left|\nabla_{\partial_{s}}^{i-j+1}h\right|\left|\nabla_{\partial_{s}}^{n-1-i}|c^{\prime}|\right|\\ &\quad\lesssim\left|\nabla_{\partial_{s}}^{n}h\right||c^{\prime}|+\|h\|_{G_{c}}\sum_{i=0}^{n-1}\sum_{j=0}^{i}\left|\nabla_{\partial_{s}}^{j}v\right|\left|\nabla_{\partial_{s}}^{n-1-i}|c^{\prime}|\right|\\ &\quad\lesssim\left|\nabla_{\partial_{s}}^{n}h\right||c^{\prime}|+\|h\|_{G_{c}}\left(\left|\nabla_{\partial_{s}}^{n-1}|c^{\prime}|\right|+|c^{\prime}|\left|\nabla_{\partial_{s}}^{n-1}v\right|+\sum_{i,j=0}^{n-2}\left|\nabla_{\partial_{s}}^{j}v\right|\left|\nabla_{\partial_{s}}^{i}|c^{\prime}|\right|\right)\\ &\quad\lesssim\left|\nabla_{\partial_{s}}^{n}h\right|+\|h\|_{G_{c}}\left(\left|\nabla_{\partial_{s}}^{n-1}|c^{\prime}|\right|+\left|\nabla_{\partial_{s}}^{n-1}v\right|+1\right)\end{split}

where in the second inequality we used (5.2), and in the bounds (5.6) and (C.2) on metric balls. Squaring this and integrating, we obtain, using (5.3) for the first term,

Dc,h(sn1|c|)L2C(1+|sn1v|L2+sn1|c|L2)hGc.\|D_{c,h}(\nabla_{\partial_{s}}^{n-1}|c^{\prime}|)\|_{L^{2}}\leq C(1+\||\nabla_{\partial_{s}}^{n-1}v|\|_{L^{2}}+\|\nabla_{\partial_{s}}^{n-1}|c^{\prime}|\|_{L^{2}})\|h\|_{G_{c}}.

Therefore, we are left to show that |sn1v|L2\||\nabla_{\partial_{s}}^{n-1}v|\|_{L^{2}} is bounded on metric balls.

As before, we need to show that

(C.6) hsn1vL2C(1+sn1vL2)hGc,\displaystyle\|\nabla_{h}\nabla_{\partial_{s}}^{n-1}v\|_{L^{2}}\leq C(1+\|\nabla_{\partial_{s}}^{n-1}v\|_{L^{2}})\|h\|_{G_{c}},

and we have shown that

|hsn1v||snh|+|sn1v||sh|+i=0n1Pi|sih|\begin{split}|\nabla_{h}\nabla_{\partial_{s}}^{n-1}v|&\lesssim|\nabla_{\partial_{s}}^{n}h|+|\nabla_{\partial_{s}}^{n-1}v||\nabla_{\partial_{s}}h|+\sum_{i=0}^{n-1}P_{i}|\nabla_{\partial_{s}}^{i}h|\end{split}

where PiP_{i} are polynomials in |sv|,,|sn2v||\nabla_{\partial_{s}}v|,\ldots,|\nabla_{\partial_{s}}^{n-2}v|, which are bounded on metric balls. We therefore have, using (5.2) that

|hsn1v||snh|+hGc(1+|sn1v|).|\nabla_{h}\nabla_{\partial_{s}}^{n-1}v|\lesssim|\nabla_{\partial_{s}}^{n}h|+\|h\|_{G_{c}}\left(1+|\nabla_{\partial_{s}}^{n-1}v|\right).

Squaring, integrating and using (5.3), we obtain (C.6), which completes the proof.

References

  • [1] V. Arnold. Sur la géométrie différentielle des groupes de lie de dimension infinie et ses applications à l’hydrodynamique des fluides parfaits. In Annales de l’institut Fourier, volume 16, pages 319–361, 1966.
  • [2] V. I. Arnold and B. A. Khesin. Topological methods in hydrodynamics, volume 125. Springer Science & Business Media, 1999.
  • [3] C. Atkin. The Hopf–Rinow theorem is false in infinite dimensions. Bulletin of the London Mathematical Society, 7(3):261–266, 1975.
  • [4] M. Bauer, M. Bruveris, N. Charon, and J. Møller-Andersen. A relaxed approach for curve matching with elastic metrics. ESAIM: Control, Optimisation and Calculus of Variations, 25:72, 2019.
  • [5] M. Bauer, M. Bruveris, E. Cismas, J. Escher, and B. Kolev. Well-posedness of the epdiff equation with a pseudo-differential inertia operator. Journal of Differential Equations, 269(1):288–325, 2020.
  • [6] M. Bauer, M. Bruveris, and B. Kolev. Fractional Sobolev metrics on spaces of immersed curves. Calculus of Variations and Partial Differential Equations, 57(1):27, 2018.
  • [7] M. Bauer, M. Bruveris, and P. W. Michor. Overview of the geometries of shape spaces and diffeomorphism groups. Journal of Mathematical Imaging and Vision, 50(1-2):60–97, 2014.
  • [8] M. Bauer and P. Harms. Metrics on spaces of immersions where horizontality equals normality. Differential Geometry and its Applications, 39:166–183, 2015.
  • [9] M. Bauer, P. Harms, and P. W. Michor. Sobolev metrics on shape space of surfaces. Journal of Geometric Mechanics, 3(4):389–438, 2011.
  • [10] M. Bauer, P. Harms, and P. W. Michor. Fractional sobolev metrics on spaces of immersions. Calculus of Variations and Partial Differential Equations, 59(2):1–27, 2020.
  • [11] F. Boyer and P. Fabrie. Mathematical Tools for the Study of the Incompressible Navier–Stokes Equations and Related Models. Springer-Verlag New York, 2013.
  • [12] M. Bruveris. Completeness properties of Sobolev metrics on the space of curves. Journal of Geometric Mechanics, 7(2):125–150, 2015.
  • [13] M. Bruveris. Regularity of maps between Sobolev spaces. Annals of Global Analysis and Geometry, 52(1):11–24, 2017.
  • [14] M. Bruveris, P. W. Michor, and D. Mumford. Geodesic completeness for Sobolev metrics on the space of immersed plane curves. Forum of Mathematics, Sigma, 2, 2014.
  • [15] M. Bruveris and J. Møller-Andersen. Completeness of length-weighted Sobolev metrics on the space of curves. arXiv preprint arXiv:1705.07976, 2017.
  • [16] M. Bruveris and F.-X. Vialard. On completeness of groups of diffeomorphisms. Journal of the European Mathematical Society, 19(5):1507–1544, 2017.
  • [17] M. P. d. Carmo. Riemannian geometry. Birkhäuser, 1992.
  • [18] E. Celledoni, M. Eslitzbichler, and A. Schmeding. Shape analysis on lie groups with applications in computer animation. Journal of Geometric Mechanics, 8(3):273, 2016.
  • [19] M. Dai, Z. Zhang, and A. Srivastava. Analyzing dynamical brain functional connectivity as trajectories on space of covariance matrices. IEEE transactions on medical imaging, 39(3):611–620, 2019.
  • [20] D. G. Ebin and J. Marsden. Groups of diffeomorphisms and the motion of an incompressible fluid. Annals of Mathematics, pages 102–163, 1970.
  • [21] I. Ekeland et al. The Hopf–Rinow theorem in infinite dimension. Journal of Differential Geometry, 13(2):287–301, 1978.
  • [22] H. Inci, T. Kappeler, and P. Topalov. On the regularity of the composition of diffeomorphisms. Mem. Amer. Math. Soc., 226(1062):vi+60, 2013.
  • [23] B. Kolev. Local well-posedness of the epdiff equation: A survey. Journal of Geometric Mechanics, 9(2):167, 2017.
  • [24] S. Lang. Fundamentals of differential geometry, volume 191. Springer Science & Business Media, 1999.
  • [25] G. Leoni. A first course in Sobolev spaces. American Mathematical Soc., 2nd edition, 2017.
  • [26] M. Micheli, P. W. Michor, and D. Mumford. Sobolev metrics on diffeomorphism groups and the derived geometry of spaces of submanifolds. Izvestiya: Mathematics, 77:3:541–570, 2013.
  • [27] P. W. Michor. Manifolds of mappings for continuum mechanics. In Geometric Continuum Mechanics, volume 42 of Advances in Continuum Mechanics, pages 3–75. Birkhäuser Basel, 2020.
  • [28] P. W. Michor and D. Mumford. An overview of the Riemannian metrics on spaces of curves using the Hamiltonian approach. Applied and Computational Harmonic Analysis, 23(1):74–113, 2007.
  • [29] G. Misiołek and S. C. Preston. Fredholm properties of riemannian exponential maps on diffeomorphism groups. Inventiones mathematicae, 179(1):191, 2010.
  • [30] D. Mumford and P. W. Michor. On Euler’s equation and ’EPDiff’. Journal of Geometric Mechanics, 5(3):319, 2013.
  • [31] P. Petersen. Riemannian Geometry. Springer, 3rd edition, 2016.
  • [32] A. Srivastava, E. Klassen, S. H. Joshi, and I. H. Jermyn. Shape analysis of elastic curves in euclidean spaces. IEEE Transactions on Pattern Analysis and Machine Intelligence, 33(7):1415–1428, 2010.
  • [33] A. Srivastava and E. P. Klassen. Functional and shape data analysis, volume 1. Springer, 2016.
  • [34] J. Su, S. Kurtek, E. Klassen, A. Srivastava, et al. Statistical analysis of trajectories on riemannian manifolds: bird migration, hurricane tracking and video surveillance. The Annals of Applied Statistics, 8(1):530–552, 2014.
  • [35] Z. Su, E. Klassen, and M. Bauer. Comparing curves in homogeneous spaces. Differential Geometry and its Applications, 60:9–32, 2018.
  • [36] C. Vizman et al. Geodesic equations on diffeomorphism groups. SIGMA. Symmetry, Integrability and Geometry: Methods and Applications, 4:030, 2008.
  • [37] K. Wehrheim. Uhlenbeck Compactness. European Mathematical Society, 2004.
  • [38] L. Younes. Computable elastic distances between shapes. SIAM Journal on Applied Mathematics, 58(2):565–586, 1998.
  • [39] L. Younes. Shapes and diffeomorphisms, volume 171. Springer, 2010.