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

Completeness and geodesic distance properties for fractional Sobolev metrics on spaces of immersed curves

Martin Bauer Department of Mathematics, Florida State University, Tallahassee, United States. bauer@math.fsu.edu Patrick Heslin Department of Mathematics and Statistics, National University of Ireland, Maynooth, Kildare, Ireland. patrick.heslin@mu.ie  and  Cy Maor Einstein Institute of Mathematics, Hebrew University of Jerusalem, Israel. cy.maor@mail.huji.ac.il
(Date: December 2023)
Abstract.

We investigate the geometry of the space of immersed closed curves equipped with reparametrization-invariant Riemannian metrics; the metrics we consider are Sobolev metrics of possible fractional order q[0,)q\in[0,\infty). We establish the critical Sobolev index on the metric for several key geometric properties. Our first main result shows that the Riemannian metric induces a metric space structure if and only if q>1/2q>1/2. Our second main result shows that the metric is geodesically-complete (i.e., the geodesic equation is globally well-posed) if q>3/2q>3/2, whereas if q<3/2q<3/2 then finite-time blowup may occur. The geodesic-completeness for q>3/2q>3/2 is obtained by proving metric-completeness of the space of HqH^{q}-immersed curves with the distance induced by the Riemannian metric.

Key words and phrases:
infinite-dimensional Riemannian geometry, immersions, geodesic distance, completeness, global well-posedness, fractional Sobolev space
2020 Mathematics Subject Classification:
58B20, 58D10, 35G55, 35A01

1. Introduction

Background and Motivation.

Reparametrization-invariant Sobolev metrics on the space of immersed curves have been of central interest in recent years: from an application point of view, they take a central role in the area of mathematical shape analysis, see e.g. [41, 46, 6] and the references therein. These metrics also arise in higher-order gradient flows for various functionals [40, 37]. From a theoretical point of view, they are the natural generalization of right-invariant Sobolev metrics on the diffeomorphism group of 𝕊1\mathbb{S}^{1}; whose geodesic equations reduce to many important PDEs from hydrodynamics, including the Burgers, Camassa-Holm and Hunter-Saxton equations. For a comprehensive list of examples, see the book of Arnold and Khesin [1].

More recently, there has been an interest to extend the study of reparamerization-invariant Sobolev metrics to those of fractional order. This can be motivated, e.g., by applications in shape optimization in geometric knot theory [39, 28]; there, a main tool is using gradient-based approach for H3/2H^{3/2} and H3/2+ϵH^{3/2+\epsilon}-type metrics, an exponent we show in this work to be critical for the completeness of the metric. Fractional order metrics have already been investigated in the context of the aforementioned geometric approach to hydrodynamics. Well-known PDEs, including the Surface Quasi-Geostrophic equations [44] and the modified Constantin-Lax-Majda equation [45], arise as reduced geodesic equations for right-invariant Sobolev metrics of fractional order on diffeomorphism groups.

The geometry of infinite dimensional Riemannian manifolds is subtle and susceptible to pathologies and many elementary facts from finite dimensional geometry do not necessarily carry over. Indeed a smooth exponential map may not exist [19], the geodesic distance can be degenerate or even vanish identically [22, 35, 32, 27], and almost all statements of the classical Hopf-Rinow theorem fail to hold [26, 34, 2].

In the context of reparametrization-invariant Sobolev metrics on spaces of immersions, generally speaking, the higher the order of the Sobolev metric, the better behaved the Riemannian structure. The goal of this current paper is to identify the exact thresholds in which transitions between “bad” and “good” behaviors occur, as described below, and are summarized in Table 1.

Main results.

We will now describe the main results of the present article. Our central object of interest is the space Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) of smooth immersions of closed curves in d\mathbb{R}^{d}, endowed with a reparametrization-invariant Sobolev metric of order q[0,)q\in[0,\infty), denoted by GqG^{q}. Each result is restated later in greater detail and generality, and includes also immersions of Sobolev (rather than smooth) regularity. Exact definitions of the spaces and the metrics considered here are given in Section 2.2.

Our first main result concerns the induced geodesic distance: Since the space of immersed curves Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) is an infinite dimensional manifold, the induced geodesic distance of a Riemannian metric is a-priori only guaranteed to be a semi-metric, i.e., distinct elements can be of zero distance [35, 22]. The following result characterizes precisely for which metrics this occurs:

Theorem (Geodesic distance).

The geodesic distance of the reparametrization-invariant Sobolev metric of order q[0,)q\in[0,\infty), on the space of smooth immersed closed curves Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}), induced a metric space structure if and only if q>1/2q>1/2.

A more detailed version of this Theorem is given in Theorem 3.1. The fact that for q=0q=0 the geodesic distance collapses was obtained by Michor–Mumford [36], who also showed that the geodesic distance is not degenerate on the quotient shape space for q1q\geq 1. Using different methods we extend their results in both directions.

Our second main result concerns the well-posedness of the corresponding geodesic equation: Bauer–Bruveris–Kolev [5] showed that these equations are locally well-posed when the order of the metric is at least 11. Here we determine the critical index for global existence, i.e, geodesic completeness of the metric:

Theorem (Geodesic completeness).

The reparametrization-invariant Sobolev metric GqG^{q} on the space of smooth immersed closed curves Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) is geodesically-complete if q>3/2q>3/2 and is not if q<3/2q<3/2.

A more detailed version of this theorem, including for results on geodesic convexity and metric completeness for curves of Sobolev regularity, is given in Theorem 4.1. Together with the local well-posedness result for the geodesic equation [5], our result implies that the corresponding geodesic equation is globally well-posed for q>3/2q>3/2. This was previously known only for integer order metrics q2q\geq 2 [12].

The proof of this result extends a method previously used for proving completeness of integer-order metrics [12, 11]: first proving that for q>3/2q>3/2 the space of HqH^{q}-Sobolev immersions, endowed with our reparametrization-invariant Sobolev metric of order qq, is metrically-complete. Since this metric is a strong metric on this space, the geodesic completeness then follows by the only part of the Hopf-Rinow theorem that holds in infinite dimensions. The geodesic completeness in the smooth category is shown by an Ebin-Marsden-type no-loss-no-gain argument [21]. The main challenges here are the more complicated estimates that arise due to the fractional order norms. The proof for the geodesic incompleteness for q<3/2q<3/2 follows from a simple example of shrinking circles.

qq (order of the metric) 0 (0,12)(0,\frac{1}{2}) 12\frac{1}{2} (12,1)(\frac{1}{2},1) 11 (1,32)(1,\frac{3}{2}) 32\frac{3}{2} >32>\frac{3}{2}
smoothness111For the results concerning smoothness of the extended spray for Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}) we refer to [19, 24] and for Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) to [5] (this is not an extensive list, and also the ones below are not).
metric space222For the results concerning the metric space structure of Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}) we refer to [4, 27] and for the previously known cases on Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) to [36].
geodesic completeness333For the results concerning the geodesic completeness on Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}) we refer to [23, 17, 18, 9, 38, 10] and for the previously known cases on Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) to [14, 12]. ? ? ? ?
Table 1. Geometric properties of right-invariant HqH^{q} metrics on Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}) ( blue) and Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) ( red). Smoothness refers to whether the geodesic spray extends to a smooth vector field on the space of Sobolev diffeomorphisms (curves, resp.), and implies local well-posedness of the geodesic equation in both the Sobolev and smooth categories. Together with geodesic completeness this implies global well-posedness. The contribution of this paper is virtually to all the results in red in the last two lines.

Comparisons with results for right-invariant metrics on diffeomorphism groups.

The results of this paper are analogous to those on Sobolev metrics on the group of diffeomorphism of the circle Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}); in particular, the critical exponents for completeness and vanishing-distance turn out to be the same, see Table 1.

While the eventual results are mostly similar for Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}) and Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}), obtaining them for Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) is generally significantly harder. The reason for this is that Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) is a much richer space — one can think of Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}) as a subspace of Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) of curves with a fixed image. From another perspective, results on Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}) can be reduced to arguments over the Lie algebra 𝔛(𝕊1)\mathfrak{X}(\mathbb{S}^{1}) of vector fields, whereas there is no equivalent for this on Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}). This can be seen, for example, in the proof of non-vanishing distance of the geodesic distance for q>1/2q>1/2 (Theorem 3.1): In the case of Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}), this is a simple application of the embedding LHqL^{\infty}\subset H^{q} [4]. However, using the same embedding for paths in Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}), we obtain weights, depending on the length of the curves in the path (whereas in the case of Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}) the length is fixed); these lengths are not controlled from below when q(1/2,3/2]q\in(1/2,3/2], and thus a more convoluted argument is needed.

Similarly the global well-posedness for q>32q>\frac{3}{2} on Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}) follows directly by abstract arguments [16], using the fact that the GqG^{q}-metric extends to a strong, right invariant metric on the space of Sobolev diffeomorphisms of regularity qq. Similar arguments can be used to show that the GqG^{q}-metric induces a strong, reparametrization-invariant metric on the space of curves of Sobolev regularity qq. Due to the more intricate nature of Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) this result cannot be used to directly conclude global well-posedness. Instead one has to carefully bound the dependence of several geometric quantities on GqG^{q}-metric balls and use this to prove metric completeness by direct estimates.

Future directions

The results of this paper also work for scale-invariant versions of the metrics GqG^{q}, and other length weights; it would be interesting to find optimal (or nearly optimal) conditions on the length-weights for which completeness holds for q>3/2q>3/2, in the spirit of [15]. We expect our results to also extend to the case of manifold-valued curves, by combining the techniques of this paper with those of [11]. Regarding vanishing geodesic distance, it is still open whether, for q<1/2q<1/2, the geodesic distance collapses completely, and whether it collapses also on shape space, which both hold in the case q=0q=0 [36]. Finally, we note that the geodesic completeness for the critical index, i.e., for the Gc3/2G_{c}^{3/2} metric on Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}), is still open. The completeness for the corresponding metric on Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}) was recently established [10].

Structure of the paper

In Section 2 we provide the necessary background, including definitions of the fractional Sobolev norms we use in this paper, the space of Sobolev immersions, and the reparametrization-invariant metrics. We also prove some useful inequalities, both on the Sobolev norms and on the Riemannian metrics. In Section 3 we restate and prove the results regarding the geodesic distance. In Section 4 we restate and prove the results regarding completeness properties. The appendix contains proofs of some results for fractional Sobolev spaces used throughout the paper.

Acknowledgements

CM and MB were partially funded by BSF grant #2022076. MB was partially funded by NSF grant DMS-1953244 and by the Austrian Science Fund grant P 35813-N. CM was partially funded by ISF grant 1269/19. PH was supported by the National University of Ireland’s Dr. Éamon de Valera Postdoctoral Fellowship.

2. Preliminaries

2.1. Fractional Sobolev Spaces

Here we record some definitions and estimates pertaining to fractional Sobolev spaces. Our presentation follows closely that of Escher and Kolev [23].

Throughout this paper, we identify 𝕊1=/\mathbb{S}^{1}=\mathbb{R}/\mathbb{Z}, and let θ[0,1]\theta\in[0,1] be a parametrization of 𝕊1\mathbb{S}^{1} (with 010\sim 1). The fractional Sobolev space Hq(𝕊1,d)H^{q}(\mathbb{S}^{1},\mathbb{R}^{d}) for qq\in\mathbb{R} is acquired by completing the space of smooth functions C(𝕊1,d)C^{\infty}(\mathbb{S}^{1},\mathbb{R}^{d}) under the norm

(2.1) fHq2=nd(1+n2)q|f^(n)|2,\left\lVert f\right\rVert_{H^{q}}^{2}=\sum_{n\in\mathbb{Z}^{d}}(1+n^{2})^{q}\lvert\hat{f}(n)\rvert^{2},

where f^(n)\displaystyle\hat{f}(n) denotes the Fourier transform of ff. We recall the Sobolev embedding theorem for fractional spaces:

Proposition 2.1 (Sobolev Embedding Theorem).

For q>12+kq>\frac{1}{2}+k the space Hq(𝕊1,d)H^{q}(\mathbb{S}^{1},\mathbb{R}^{d}) continuously embeds into the classical space Ck(𝕊1,d)C^{k}(\mathbb{S}^{1},\mathbb{R}^{d}) of kk-times continuously differentiable functions.

Proofs of this statement can be found in many standard references, e.g., [43, Section 2.7.1]. We further define the space Hq(𝕊1,𝕊1)H^{q}(\mathbb{S}^{1},\mathbb{S}^{1}) to consist of all self-maps of the circle which, when composed with any chart, are in Hq(𝕊1,)H^{q}(\mathbb{S}^{1},\mathbb{R}). If we require that q>32q>\frac{3}{2}, it follows from the Sobolev Embedding Theorem that Hq(𝕊1,𝕊1)C1(𝕊1,𝕊1)H^{q}(\mathbb{S}^{1},\mathbb{S}^{1})\hookrightarrow C^{1}(\mathbb{S}^{1},\mathbb{S}^{1}). Hence, by the Inverse Function Theorem, we may define the space of HqH^{q}-diffeomorphisms of the circle as

𝒟q(𝕊1)=Hq(𝕊1,𝕊1){C1diffeomorphisms of 𝕊1}.\operatorname{\mathcal{D}}^{q}(\mathbb{S}^{1})=H^{q}(\mathbb{S}^{1},\mathbb{S}^{1})\cap\{C^{1}-\text{diffeomorphisms of }\mathbb{S}^{1}\}.

This space is an infinite-dimensional Hilbert manifold modeled on Hq(𝕊1,)H^{q}(\mathbb{S}^{1},\mathbb{R}) c.f., [21]. It is in addition a half-Lie group [33, 8], i.e., a topological group under composition, where for any φ𝒟q(𝕊1)\varphi\in\operatorname{\mathcal{D}}^{q}(\mathbb{S}^{1}) right translation

𝒟q(𝕊1)𝒟q(𝕊1);ηηφ\displaystyle\operatorname{\mathcal{D}}^{q}(\mathbb{S}^{1})\rightarrow\operatorname{\mathcal{D}}^{q}(\mathbb{S}^{1})\ ;\ \eta\mapsto\eta\circ\varphi

is smooth, but left translation

𝒟q(𝕊1)𝒟q(𝕊1);ηφη\displaystyle\operatorname{\mathcal{D}}^{q}(\mathbb{S}^{1})\rightarrow\operatorname{\mathcal{D}}^{q}(\mathbb{S}^{1})\ ;\ \eta\mapsto\varphi\circ\eta

is only continuous. It further acts on Hq(𝕊1,)H^{q}(\mathbb{S}^{1},\mathbb{R}) where, again for fixed φ𝒟q(𝕊1)\varphi\in\operatorname{\mathcal{D}}^{q}(\mathbb{S}^{1}), the following map is smooth

Hq(𝕊1,d)Hq(𝕊1,d);ffφ.\displaystyle H^{q}(\mathbb{S}^{1},\mathbb{R}^{d})\rightarrow H^{q}(\mathbb{S}^{1},\mathbb{R}^{d})\ ;\ f\mapsto f\circ\varphi.

Throughout our arguments we will require estimates on products and compositions with respect to the homogeneous Sobolev seminorm defined by

(2.2) fH˙q2=nn2q|f^(n)|2.\displaystyle\left\lVert f\right\rVert_{\dot{H}^{q}}^{2}=\sum_{n\in\mathbb{Z}}n^{2q}\lvert\hat{f}(n)\rvert^{2}.

Finally, we note that this seminorm can be rewritten as

fH˙q2=𝕊1Λ2qh,h𝑑θ,\left\lVert f\right\rVert^{2}_{\dot{H}^{q}}=\int_{\mathbb{S}^{1}}\left\langle\Lambda^{2q}h,h\right\rangle\,d\theta,

where Λ:=Hθ\Lambda:=H\partial_{\theta} is the pseudo-differential operator with symbol Λ^(m)=|m|\widehat{\Lambda}(m)=\left\lvert m\right\rvert.

Our central estimates are as follows. From a notational standpoint, we will write 1R2\left\lVert\cdot\right\rVert_{1}\simeq_{R}\left\lVert\cdot\right\rVert_{2}, 1R2\left\lVert\cdot\right\rVert_{1}\lesssim_{R}\left\lVert\cdot\right\rVert_{2}, etc. to indicate an equivalence or inequality is valid up to a constant depending continuously on RR.

Lemma 2.2.

Consider the Sobolev spaces and norms as defined above.

  1. (i)

    For 0<ab0<a\leq b we have

    (2.3) fH˙afH˙b,\left\lVert f\right\rVert_{\dot{H}^{a}}\leq\left\lVert f\right\rVert_{\dot{H}^{b}},

    for all fHb(𝕊1,d)f\in H^{b}(\mathbb{S}^{1},\mathbb{R}^{d}).

  2. (ii)

    For b>12b>\frac{1}{2} and 0ab0\leq a\leq b, we have the following estimate on products

    (2.4) fgHa(a,b)fHagHb,\left\lVert f\cdot g\right\rVert_{{H}^{a}}\lesssim_{(a,b)}\left\lVert f\right\rVert_{{H}^{a}}\left\lVert g\right\rVert_{{H}^{b}},

    for all fHa(𝕊1,d)f\in H^{a}(\mathbb{S}^{1},\mathbb{R}^{d}) and gHb(𝕊1,d)g\in H^{b}(\mathbb{S}^{1},\mathbb{R}^{d}).

  3. (iii)

    For b>12b>\frac{1}{2} and 0ab0\leq a\leq b, we have the following estimate on products for the homogeneous norm

    (2.5) fgH˙a(a,b)|f^(0)|gH˙a+|g^(0)|fH˙a+fH˙agH˙b,\left\lVert f\cdot g\right\rVert_{\dot{H}^{a}}\lesssim_{(a,b)}|\hat{f}(0)|\left\lVert g\right\rVert_{\dot{H}^{a}}+\left\lvert\hat{g}(0)\right\rvert\left\lVert f\right\rVert_{\dot{H}^{a}}+\left\lVert f\right\rVert_{\dot{H}^{a}}\left\lVert g\right\rVert_{\dot{H}^{b}},

    for all fHa(𝕊1,d)f\in H^{a}(\mathbb{S}^{1},\mathbb{R}^{d}) and gHb(𝕊1,d)g\in H^{b}(\mathbb{S}^{1},\mathbb{R}^{d}).

  4. (iv)

    For 0a10\leq a\leq 1 we have the following estimate on products for the homogeneous norm

    (2.6) fgH˙aafH˙agL+fLgH˙a,\left\lVert f\cdot g\right\rVert_{\dot{H}^{a}}\lesssim_{a}\left\lVert f\right\rVert_{\dot{H}^{a}}\left\lVert g\right\rVert_{L^{\infty}}+\left\lVert f\right\rVert_{L^{\infty}}\left\lVert g\right\rVert_{\dot{H}^{a}},

    for all f,gHa(𝕊1,d)f,g\in H^{a}(\mathbb{S}^{1},\mathbb{R}^{d}).

  5. (v)

    For b>32b>\frac{3}{2} and 0a10\leq a\leq 1, we have the following estimate on compositions

    (2.7) fφH˙a(φ1)θL1a2φθLa2fH˙a,\left\lVert f\circ\varphi\right\rVert_{\dot{H}^{a}}\leq\left\lVert(\varphi^{-1})_{\theta}\right\rVert_{L^{\infty}}^{\tfrac{1-a}{2}}\left\lVert\varphi_{\theta}\right\rVert_{L^{\infty}}^{\tfrac{a}{2}}\left\lVert f\right\rVert_{\dot{H}^{a}},

    for all fHa(𝕊1,d)f\in H^{a}(\mathbb{S}^{1},\mathbb{R}^{d}) and φ𝒟b(𝕊1)\varphi\in\operatorname{\mathcal{D}}^{b}(\mathbb{S}^{1}).

Details of the proofs for these estimates can be found in Appendix A.

2.2. Riemannian Geometry of Immersed Curves

Here we introduce the setting for the results contained in this paper. Further details of the constructions can be found in Michor–Mumford [36].

We consider the space of smooth immersions of 𝕊1\mathbb{S}^{1} into Euclidean space

(2.8) Imm(𝕊1,d):={cC(𝕊1,d)||cθ|0}.\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}):=\{c\in C^{\infty}(\mathbb{S}^{1},\mathbb{R}^{d})\ |\ \left\lvert c_{\theta}\right\rvert\neq 0\}.

It is an open subset of C(𝕊1,d)C^{\infty}(\mathbb{S}^{1},\mathbb{R}^{d}) and hence inherits the structure of an infinite-dimensional Frechét manifold with tangent space at the point cImm(𝕊1,d)c\in\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) given by

(2.9) TcImm(𝕊1,d):={hC(𝕊1,Td)|πTdh=c}.T_{c}\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}):=\{h\in C^{\infty}(\mathbb{S}^{1},T\mathbb{R}^{d})\ |\ \pi_{T\mathbb{R}^{d}}\circ h=c\}.

Next, we define the space of smooth, orientation preserving diffeomorphisms of 𝕊1\mathbb{S}^{1}

(2.10) Diff(𝕊1):={φC(𝕊1,𝕊1)|φ is a bijection}.\operatorname{\operatorname{Diff}}(\mathbb{S}^{1}):=\{\varphi\in C^{\infty}(\mathbb{S}^{1},\mathbb{S}^{1})\ |\ \varphi\text{ is a bijection}\}.

This is an infinite-dimensional Frechét Lie group which acts on Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) on the right by composition

(2.11) Imm(𝕊1,d)×Diff(𝕊1)Imm(𝕊1,d);(c,φ)cφ.\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d})\times\operatorname{\operatorname{Diff}}(\mathbb{S}^{1})\rightarrow\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d})\ ;\ (c,\varphi)\mapsto c\circ\varphi.

We are interested in equipping Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) with reparametrization-invariant Riemannian metrics GG, i.e., for all cImm(𝕊1,d),h,kTfImm(𝕊1,d)c\in\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}),h,k\in T_{f}\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) and φDiff(𝕊1)\varphi\in\operatorname{\operatorname{Diff}}(\mathbb{S}^{1})

(2.12) Gcφ(hφ,kφ)=Gc(h,k).G_{c\circ\varphi}\left(h\circ\varphi,k\circ\varphi\right)=G_{c}(h,k).

The importance of these metrics stems from the fact that they descend to Riemannian metrics on the shape space

(2.13) Bi(𝕊1,d):=Imm(𝕊1,d)/Diff(𝕊1),\text{B}_{i}(\mathbb{S}^{1},\mathbb{R}^{d}):=\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d})/\operatorname{\operatorname{Diff}}(\mathbb{S}^{1}),

which carries the structure of an infinite-dimensional orbifold.

A subclass of reparametrization-invariant metrics are Sobolev-type metrics of order qq\in\mathbb{R}:

(2.14) Gcq(h,k)=𝕊1Lcqh,k𝑑sG_{c}^{q}(h,k)=\int_{\mathbb{S}^{1}}\left\langle L_{c}^{q}h,k\right\rangle\ ds

where LcqL^{q}_{c} is a self-adjoint, invertible, pseudo-differential operator of order 2q2q depending on cc in such a way as to keep GG reparametrization-invariant and ds=|cθ|dθds=\left\lvert c_{\theta}\right\rvert d\theta denotes integration with respect to arc length. It has been observed in [5], that the requirement that the metric (2.14) be invariant under Diff(𝕊1)\operatorname{\operatorname{Diff}}(\mathbb{S}^{1}) tells us that ,Gcq\left\langle\cdot,\cdot\right\rangle_{G_{c}^{q}} is completely determined by its behavior on constant speed curves. In particular, if we define the constant speed reparametrization for cc:

(2.15) ψc(θ)=1lc0θ|cθ(σ)|𝑑σ,\psi_{c}(\theta)=\frac{1}{l_{c}}\int_{0}^{\theta}|c_{\theta}(\sigma)|\ d\sigma,

where lcl_{c} denotes the length of cc, we have |(cψc1)θ|=lc|(c\circ\psi^{-1}_{c})_{\theta}|=l_{c}; which in turn gives us that Lcq=RψcLcψc1qRψc1L^{q}_{c}=R_{\psi_{c}}L^{q}_{c\circ\psi_{c}^{-1}}R_{\psi_{c}^{-1}}. We now assume LcqL^{q}_{c} has the form:

(2.16) Lcq=Rψc(1+(1lc)2qΛ2q)Rψc1,L^{q}_{c}=R_{\psi_{c}}\left(1+\left(\frac{1}{l_{c}}\right)^{2q}\Lambda^{2q}\right)R_{\psi_{c}^{-1}},

where, as before, Λ:=Hθ\Lambda:=H\partial_{\theta} is the pseudo-differential operator with symbol Λ^(m)=|m|\widehat{\Lambda}(m)=\left\lvert m\right\rvert. In summary, our full and our homogeneous reparametrization-invariant metrics of interest are, respectively, given by:

(2.17) Gcq(h,k):=𝕊1Rψc(1+(1lc)2qΛ2q)Rψc1h,k𝑑sG_{c}^{q}(h,k):=\int_{\mathbb{S}^{1}}\left\langle R_{\psi_{c}}\left(1+\left(\frac{1}{l_{c}}\right)^{2q}\Lambda^{2q}\right)R_{\psi_{c}^{-1}}h,k\right\rangle\ ds

and

(2.18) G˙cq(h,k):=𝕊1Rψc(1lc)2qΛ2qRψc1h,k𝑑s.\dot{G}_{c}^{q}(h,k):=\int_{\mathbb{S}^{1}}\left\langle R_{\psi_{c}}\left(\frac{1}{l_{c}}\right)^{2q}\Lambda^{2q}R_{\psi_{c}^{-1}}h,k\right\rangle\ ds.

We further denote differentiation with respect to arc length by Ds=1|cθ|θD_{s}=\frac{1}{\left\lvert c_{\theta}\right\rvert}\partial_{\theta}. It is not difficult to show that DshG˙cq=hG˙cq+1\left\lVert D_{s}h\right\rVert_{\dot{G}_{c}^{q}}=\left\lVert h\right\rVert_{\dot{G}_{c}^{q+1}}. For constant speed curves this follows from integration by parts. The result for non-constant speed curves then follows from the reparametrization-invariance of G˙cq\dot{G}_{c}^{q}.

We also consider alongside the above space Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}), its finite smoothness counterparts. For r>32{r}>\frac{3}{2} we have the Sobolev completions

(2.19) r(𝕊1,d):={cHr(𝕊1,d)||cθ|0},\displaystyle\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}):=\{c\in H^{r}(\mathbb{S}^{1},\mathbb{R}^{d})\ |\ \left\lvert c_{\theta}\right\rvert\neq 0\}\ ,
(2.20) Tcr(𝕊1,d):={hHr(𝕊1,Td)|πTdh=c}.\displaystyle T_{c}\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}):=\{h\in H^{r}(\mathbb{S}^{1},T\mathbb{R}^{d})\ |\ \pi_{T\mathbb{R}^{d}}\circ h=c\}.

The following result is due to Bauer et al. [5, Theorems 6.4 and 7.1].

Proposition 2.3.

For r>32r>\frac{3}{2} and q[12,r2]q\in[\frac{1}{2},\frac{r}{2}] the metric GcqG_{c}^{q} as in (2.17) is a smooth metric on r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}). If q1q\geq 1, then it induces a smooth exponential map on r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}), which is a local diffeomorphism. If q=rq=r, then the metric extends to a strong metric on r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}), i.e., it induces on each tangent space the original HrH^{r}-topology.

We now express the induced homogeneous reparametrization-invariant norm in terms of the usual homogeneous Sobolev norm.

Lemma 2.4.

For r>32r>\frac{3}{2}, crc\in\mathcal{I}^{r} and hTcrh\in T_{c}\mathcal{I}^{r} and qrq\leq r we have

(2.21) hG˙cq2=lc12qhψc1H˙q2.\left\lVert h\right\rVert_{\dot{G}_{c}^{q}}^{2}=l_{c}^{1-2q}\left\lVert h\circ\psi_{c}^{-1}\right\rVert_{\dot{H}^{q}}^{2}.

Moreover, the subsequent inequality holds for any q1q2rq_{1}\leq q_{2}\leq r

(2.22) hG˙cq1lcq2q1hG˙cq2.\left\lVert h\right\rVert_{\dot{G}_{c}^{{q_{1}}}}\leq l_{c}^{{q_{2}-q_{1}}}\left\lVert h\right\rVert_{\dot{G}_{c}^{q_{2}}}.

Lastly, when 1qr1\leq q\leq r we have

(2.23) hG˙c1hGcq.\left\lVert h\right\rVert_{\dot{G}_{c}^{1}}\leq\left\lVert h\right\rVert_{{G}_{c}^{q}}.
Proof.

Applying a change of coordinates we compute directly

hG˙cq2\displaystyle\left\lVert h\right\rVert_{\dot{G}_{c}^{q}}^{2} =𝕊1(1lc)2qΛ2qRψc1h,Rψc1hlc𝑑θ\displaystyle=\int_{\mathbb{S}^{1}}\left\langle\left(\frac{1}{l_{c}}\right)^{2q}\Lambda^{2q}R_{\psi_{c}^{-1}}h,R_{\psi_{c}^{-1}}h\right\rangle\ l_{c}\ d\theta
=lc12q𝕊1Λ2qRψc1h,Rψc1h𝑑θ\displaystyle=l_{c}^{1-2q}\int_{\mathbb{S}^{1}}\left\langle\Lambda^{2q}R_{\psi_{c}^{-1}}h,R_{\psi_{c}^{-1}}h\right\rangle\ d\theta
=lc12qhψc1H˙q2.\displaystyle=l_{c}^{1-2q}\left\lVert h\circ\psi_{c}^{-1}\right\rVert_{\dot{H}^{q}}^{2}.

The inequality (2.22) follows immediately from the above and (2.3).

Finally, note that, as the metrics (2.17) and (2.18) are invariant under reparametrization, it will suffice to establish the inequality (2.23) on constant speed curves, i.e., where ψc(θ)=θ\psi_{c}(\theta)=\theta. By (2.21) we acquire that

hG˙c12\displaystyle\left\lVert h\right\rVert_{\dot{G}_{c}^{1}}^{2} =lc1hH˙12=lc1j2|j|2|h^(j)|2\displaystyle=l_{c}^{-1}\left\lVert h\right\rVert_{\dot{H}^{1}}^{2}=l_{c}^{-1}\sum_{j\in\mathbb{Z}^{2}}\left\lvert j\right\rvert^{2}\left\lvert\hat{h}(j)\right\rvert^{2}

and

hGcq2\displaystyle\left\lVert h\right\rVert_{G_{c}^{q}}^{2} =hGc02+hG˙cq2=lchL22+lc12qhH˙q2\displaystyle=\left\lVert h\right\rVert_{G_{c}^{0}}^{2}+\left\lVert h\right\rVert_{\dot{G}_{c}^{q}}^{2}=l_{c}\left\lVert h\right\rVert_{L^{2}}^{2}+l_{c}^{1-2q}\left\lVert h\right\rVert_{\dot{H}^{q}}^{2}
=lcj2|h^(j)|2+lc12qj2|j|2q|h^(j)|2\displaystyle=l_{c}\sum_{j\in\mathbb{Z}^{2}}\left\lvert\hat{h}(j)\right\rvert^{2}+l_{c}^{1-2q}\sum_{j\in\mathbb{Z}^{2}}\left\lvert j\right\rvert^{2q}\left\lvert\hat{h}(j)\right\rvert^{2}
=j2(lc+lc12q|j|2q)|h^(j)|2.\displaystyle=\sum_{j\in\mathbb{Z}^{2}}\left(l_{c}+l_{c}^{1-2q}\left\lvert j\right\rvert^{2q}\right)\left\lvert\hat{h}(j)\right\rvert^{2}.

Hence, (2.23) will hold if, for any lc>0l_{c}>0 and |j|2\left\lvert j\right\rvert\in\mathbb{Z}^{2}, we have

lc1|j|2lc+lc12q|j|2q.\displaystyle l_{c}^{-1}\left\lvert j\right\rvert^{2}\leq l_{c}+l_{c}^{1-2q}\left\lvert j\right\rvert^{2q}.

For |j|=0\left\lvert j\right\rvert=0, this is immediate. For |j|0\left\lvert j\right\rvert\neq 0 we divide both sides by lc1|j|2l_{c}^{-1}\left\lvert j\right\rvert^{2} and obtain

1(lc|j|)2+(|j|lc)2(q1),\displaystyle 1\leq\left(\frac{l_{c}}{\left\lvert j\right\rvert}\right)^{2}+\left(\frac{\left\lvert j\right\rvert}{l_{c}}\right)^{2(q-1)},

which holds for any q1q\geq 1. ∎

The next lemma, which establishes a Sobolev Embedding-type theorem for our reparametrisation-invariant norms, extends a result for first order metrics contained in [14, Lemma 2.14] to fractional orders. The important point here is the explicit dependence of the embedding constant on the length of the underlying curve.

Lemma 2.5.

If r>32r>\frac{3}{2} and 12<qr\frac{1}{2}<q\leq r, then there exists a constant C=C(q,d)>0C=C(q,d)>0 such that, for all hTcrh\in T_{c}\mathcal{I}^{r} and all (0,lc]\ell\in(0,l_{c}], we have

(2.24) hL(𝕊1)C1(hGc02+2qhG˙cq2)Cmax{12,q12}hGcq.\left\lVert h\right\rVert_{L^{\infty}(\mathbb{S}^{1})}\leq C\sqrt{\frac{1}{\ell}\left(\left\lVert h\right\rVert_{G_{c}^{0}}^{2}+\ell^{2q}\left\lVert h\right\rVert_{\dot{G}_{c}^{q}}^{2}\right)}{\leq C\max\{\ell^{-\frac{1}{2}},\ell^{q-\frac{1}{2}}\}\left\lVert h\right\rVert_{G_{c}^{q}}}.
Proof.

This inequality, for standard Sobolev spaces, i.e., with HqH^{q} instead of GqG^{q}, is well-known: For integer order q1q\geq 1 this result appears in [30, Theorem 7.40], and for q(1/2,1)q\in(1/2,1) it appears in [31, Theorem 2.8]. The general case follows by a combination of these results. Our case reduces to this case by reparametrization, as shown below for the case =lc\ell=l_{c}.

By the usual Sobolev Embedding Theorem, there exists C=C(q,d)>0C=C(q,d)>0 such that hL(𝕊1)ChHq(dθ)\left\lVert h\right\rVert_{L^{\infty}(\mathbb{S}^{1})}\leq C\left\lVert h\right\rVert_{H^{q}(d\theta)}. Hence, we have

hL(𝕊1)\displaystyle\left\lVert h\right\rVert_{L^{\infty}(\mathbb{S}^{1})} =hψc1L(𝕊1)\displaystyle=\left\lVert h\circ\psi_{c}^{-1}\right\rVert_{L^{\infty}(\mathbb{S}^{1})}
Chψc1Hq(dθ)\displaystyle\leq C\left\lVert h\circ\psi_{c}^{-1}\right\rVert_{H^{q}(d\theta)}
=C1lc(hGc02+(1lc)2qhG˙cq2).\displaystyle=C\sqrt{\frac{1}{l_{c}}\left(\left\lVert h\right\rVert_{G_{c}^{0}}^{2}+\left(\frac{1}{l_{c}}\right)^{-2q}\left\lVert h\right\rVert_{\dot{G}_{c}^{q}}^{2}\right)}.\qed

3. Geodesic Distance

In this section we study the induced geodesic distance of our class of metrics. Recall that any Riemannian metric induces a geodesic distance defined as the infimum over the length of all differentiable paths with fixed end points. As mentioned in the introduction, in finite dimensions this will always induce a metric space structure, however, in infinite dimensions this is not necessarily the case. We say that an induced geodesic distance is degenerate if there exists a pair of points for which we can find an arbitrarily short path connecting them, i.e., the geodesic distance between the points is zero.

Initial investigations into geodesic distance in the context of spaces of immersed curves are due to Michor and Mumford [36, 3]. Their results show that the geodesic distance is degenerate if q=0q=0, but that it is a true distance on the quotient shape space Bi(𝕊1,d)=Imm(𝕊1,d)/Diff(𝕊1)B_{i}(\mathbb{S}^{1},\mathbb{R}^{d})=\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d})/\operatorname{Diff}(\mathbb{S}^{1}) if q1q\geq 1. Similar non-degeneracy results can be obtained on the whole space for q1q\geq 1 using the square-root transform [42]. These results naturally raise the question, for which q(0,1)q\in(0,1) this change of behavior occurs. Our main result of this section provides an answer to this question.

Theorem 3.1.

The geodesic distance of the reparametrization invariant Sobolev metric GcqG_{c}^{q}, denoted distGcq\operatorname{dist}_{G_{c}^{q}}, is non-degenerate if and only if q>12q>\frac{1}{2}. More precisely,

  1. (i)

    For any q12q\leq\frac{1}{2} there exists distinct curves c0c1Imm(𝕊1,d)c_{0}\neq c_{1}\in\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) such that distGcq(c0,c1)=0\operatorname{dist}_{G_{c}^{q}}(c_{0},c_{1})=0.

  2. (ii)

    For any q>12q>\frac{1}{2} and any c0c1Imm(𝕊1,d)c_{0}\neq c_{1}\in\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) the geodesic distance distGcq(c0,c1)\operatorname{dist}_{G_{c}^{q}}(c_{0},c_{1}) is non-zero.

Furthermore, if q>12q>\frac{1}{2} we obtain the bound

distGcq(c0,c1)Cmin{c0c1,diammax}min{diammax1/2,1},\operatorname{dist}_{G_{c}^{q}}(c_{0},c_{1})\geq C\min\{\|c_{0}-c_{1}\|_{\infty},\operatorname{diam}_{\text{max}}\}\min\{\operatorname{diam}_{\text{max}}^{1/2},1\},

for some constant C=C(q,n)>0C=C(q,n)>0. Here diam(ci)\operatorname{diam}(c_{i}) is the diameter of the image of the curve cic_{i} and diammax=max{diam(c0),diam(c1)}\operatorname{diam}_{\text{max}}=\max\{\operatorname{diam}(c_{0}),\operatorname{diam}(c_{1})\}.

Lastly, if q1q\geq 1 we obtain the additional estimate

distGcq(c0,c1)𝕊1|θc0|θc0|12θc1|θc1|12|2𝑑θ=lc0+lc1𝕊1θc0,θc1|θc0|12|θc1|12dθ.\operatorname{dist}_{G_{c}^{q}}(c_{0},c_{1})\geq\sqrt{\int_{\mathbb{S}^{1}}\left|\frac{\partial_{\theta}c_{0}}{|\partial_{\theta}c_{0}|^{\tfrac{1}{2}}}-\frac{\partial_{\theta}c_{1}}{|\partial_{\theta}c_{1}|^{\tfrac{1}{2}}}\right|^{2}d\theta}=\sqrt{l_{c_{0}}+l_{c_{1}}-\int_{\mathbb{S}^{1}}\frac{\langle\partial_{\theta}c_{0},\partial_{\theta}c_{1}\rangle}{|\partial_{\theta}c_{0}|^{\tfrac{1}{2}}|\partial_{\theta}c_{1}|^{\tfrac{1}{2}}}d\theta.}

These results continue to hold on the space of Sobolev immersions r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}), as long as the metric GcqG_{c}^{q} is defined on it.

Remark 3.2.

For q=0q=0 it has been shown in [36], that the geodesic distance vanishes identically on all of Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}). This implies, in particular, also the degeneracy of the induced geodesic distance on the shape space Bi(𝕊1,d)B_{i}(\mathbb{S}^{1},\mathbb{R}^{d}). Our result regarding the degeneracy for q12q\leq\frac{1}{2} is significantly weaker: we only show the existence of distinct immersions, such that their geodesic distance is zero and we do not prove that this holds for arbitrary immersions. Furthermore, our examples are of the type c1=c0φ1c_{1}=c_{0}\circ\varphi_{1} with φ1Diff(𝕊1)\varphi_{1}\in\operatorname{Diff}(\mathbb{S}^{1}). These elements are, however, identified in shape space Bi(𝕊1,d)B_{i}(\mathbb{S}^{1},\mathbb{R}^{d}). Thus this result does not resolve the degeneracy on the quotient space, but only for the space of immersions. We believe that the index q=12q=\frac{1}{2} is also critical for the geodesic distance on Bi(𝕊1,d)B_{i}(\mathbb{S}^{1},\mathbb{R}^{d}), but the necessary estimates seem quite challenging and we leave this question open for future research.

To prove Theorem 3.1 we will first collect a useful estimate pertaining to the diameter of the initial curve c0c_{0}.

Lemma 3.3.

Let c0,c1Imm(𝕊1,d)c_{0},c_{1}\in\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}). If lc1diam(c0)l_{c_{1}}\leq\operatorname{diam}(c_{0}), then

14diam(c0)c0c1.\frac{1}{4}\operatorname{diam}(c_{0})\leq\|c_{0}-c_{1}\|_{\infty}.
Proof.

The diameter of any closed curve is at most half its length. Thus, by assumption,

diam(c1)12diam(c0).\operatorname{diam}(c_{1})\leq\frac{1}{2}\operatorname{diam}(c_{0}).

Now, let t,s𝕊1t,s\in\mathbb{S}^{1} such that |c0(t)c0(s)|=diam(c0)|c_{0}(t)-c_{0}(s)|=\operatorname{diam}(c_{0}), then

diam(c0)|c0(t)c1(t)|+|c1(t)c1(s)|+|c1(s)c0(s)|2c0c1+diam(c1)2c0c1+12diam(c0),\begin{split}\operatorname{diam}(c_{0})&\leq|c_{0}(t)-c_{1}(t)|+|c_{1}(t)-c_{1}(s)|+|c_{1}(s)-c_{0}(s)|\\ &\leq 2\|c_{0}-c_{1}\|_{\infty}+\operatorname{diam}(c_{1})\leq 2\|c_{0}-c_{1}\|_{\infty}+\frac{1}{2}\operatorname{diam}(c_{0}),\end{split}

from which the claim follows. ∎

With this at hand, we now proceed to the proof of the main theorem for this section.

Proof of Theorem 3.1.

We start with showing the non-degeneracy for q>12q>\frac{1}{2}. Let c(t)=c(t,)c(t)=c(t,\cdot) with t[0,1]t\in[0,1] be any path between c0c_{0} and c1c_{1}. Denote

t0=max{t[0,1]:diam(c0)lc(s) for all st}.t_{0}=\max\{t\in[0,1]~{}:~{}\operatorname{diam}(c_{0})\leq l_{c(s)}\text{ for all }s\leq t\}.

Note that t0>0t_{0}>0 since lc02diam(c0)l_{c_{0}}\geq 2\operatorname{diam}(c_{0}). By definition, lc(t)diam(c0)l_{c(t)}\geq\operatorname{diam}(c_{0}) for all t[0,t0]t\in[0,t_{0}], and we can use (2.24) with =min{1,diam(c0)}\ell=\min\{1,\operatorname{diam}(c_{0})\} to obtain

c0c(t0)0t0tcdtCmax{diam(c0)1/2,1}0t0tcGcqdt.\|c_{0}-c(t_{0})\|_{\infty}\leq\int_{0}^{t_{0}}\|\partial_{t}c\|_{\infty}\,dt\leq C\max\{\operatorname{diam}(c_{0})^{-1/2},1\}\int_{0}^{t_{0}}\|\partial_{t}c\|_{G_{c}^{q}}\,dt.

If t0=1t_{0}=1 we are done. Otherwise, t0<1t_{0}<1, and thus c(t0)=diam(c0)\ell_{c(t_{0})}=\operatorname{diam}(c_{0}). Therefore, by Lemma 3.3, we have

(3.1) 14diam(c0)\displaystyle\frac{1}{4}\operatorname{diam}(c_{0}) c0c(t0)\displaystyle\leq\|c_{0}-c(t_{0})\|_{\infty}
(3.2) Cmax{diam(c0)1/2,1}0t0tcGcqdt\displaystyle\leq C\max\{\operatorname{diam}(c_{0})^{-1/2},1\}\int_{0}^{t_{0}}\|\partial_{t}c\|_{G_{c}^{q}}\,dt
(3.3) Cmax{diam(c0)1/2,1}01tcGcqdt.\displaystyle\leq C\max\{\operatorname{diam}(c_{0})^{-1/2},1\}\int_{0}^{1}\|\partial_{t}c\|_{G_{c}^{q}}\,dt.

As this estimate holds for any path connecting c0c_{0} to c1c_{1}, it also holds for the infimum and thus we have obtained the desired bound for the geodesic distance.

Next we prove the additional bound for q1q\geq 1. Therefore we first introduce the so-called SRV transform [42], which is given as the mapping

(3.4) ccθ|cθ|12.c\mapsto\frac{c_{\theta}}{|c_{\theta}|^{\tfrac{1}{2}}}.

It has been shown that the SRV transform [42] is a Riemannian isometry from Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) equipped with the H1H^{1}-type metric

(3.5) GcSRV(h,h)=𝕊1Dsh,Dsk+14Dsh,Dskds\displaystyle G^{\operatorname{SRV}}_{c}(h,h)=\int_{\mathbb{S}^{1}}\langle D_{s}h^{\top},D_{s}k^{\top}\rangle+\frac{1}{4}\langle D_{s}h^{\bot},D_{s}k^{\bot}\rangle ds
(3.6) with Dsh=Dsh,DscDsh, and Dsh=DshDsh\displaystyle\text{with }D_{s}h^{\top}=\langle D_{s}h,D_{s}c\rangle D_{s}h,\text{ and }D_{s}h^{\bot}=D_{s}h-D_{s}h^{\top}

to a submanifold of the space of all smooth functions C(𝕊1,d)C^{\infty}(\mathbb{S}^{1},\mathbb{R}^{d}) equipped with the flat L2L^{2} (Riemannian) metric. It is easy to see that the H1H^{1}-metric Gc1G_{c}^{1} is lower bounded by the SRV metric GSRVG^{\operatorname{SRV}} and thus the same is true for their geodesic distances. Finally we note, that the geodesic distance of a submanifold is bounded by the geodesic distance on the surrounding space and that the geodesic distance of the flat L2L^{2}-metric is simply given by the L2L^{2}-difference of the functions. Combining these observations leads to the desired lower bound.

It remains to prove the degeneracy for q12q\leq\frac{1}{2}. To this end, consider any c0Imm(𝕊1,d)c_{0}\in\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) and let c1=c0φ1c_{1}=c_{0}\circ\varphi_{1} for some fixed reparametrization idφ1Diff(𝕊1)\operatorname{id}\neq\varphi_{1}\in\operatorname{Diff}(\mathbb{S}^{1}). We aim to show that the geodesic distance between c0c_{0} and c1c_{1} is zero. For the sake of simplicity we assume that c0c_{0} is parametrized by arc length. We now consider any path φ:[0,1]Diff(𝕊1)\varphi:[0,1]\mapsto\operatorname{\operatorname{Diff}}(\mathbb{S}^{1}) connecting id\operatorname{id} to φ1\varphi_{1}. Then c(t)=c0(φ(t))c(t)=c_{0}(\varphi(t)) is a path in the manifold of immersions that connects c0c_{0} to c1c_{1}. From this we have

(3.7) distGcq(c0,c1)\displaystyle\operatorname{dist}_{G_{c}^{q}}(c_{0},c_{1}) 01tcGcq𝑑t\displaystyle\leq\int_{0}^{1}\|\partial_{t}c\|_{G_{c}^{q}}\,dt
(3.8) 01(θc0φ)tφGcq𝑑t\displaystyle\leq\int_{0}^{1}\|\left(\partial_{\theta}c_{0}\circ\varphi\right)\partial_{t}\varphi\|_{G_{c}^{q}}\,dt
(3.9) =01θc0(tφφ1)Gcq𝑑t\displaystyle=\int_{0}^{1}\|\partial_{\theta}c_{0}\left(\partial_{t}\varphi\circ\varphi^{-1}\right)\|_{G_{c}^{q}}\,dt
(3.10) 01C(lc)θc0(tφφ1)Hq𝑑t,\displaystyle\leq\int_{0}^{1}C(l_{c})\|\partial_{\theta}c_{0}\left(\partial_{t}\varphi\circ\varphi^{-1}\right)\|_{H^{q}}\,dt,

where we used Lemma 2.4 for the expression of the GqG^{q} metric in the last step. Since the length of c(t)c(t) is constant in time, i.e., lc0=lc(t)l_{c_{0}}=l_{c(t)} we can bound this via

(3.11) distGcq(c0,c1)\displaystyle\operatorname{dist}_{G_{c}^{q}}(c_{0},c_{1}) C(lc0)01θc0(tφφ1)Hq𝑑t\displaystyle\leq C(l_{c_{0}})\int_{0}^{1}\|\partial_{\theta}c_{0}\left(\partial_{t}\varphi\circ\varphi^{-1}\right)\|_{H^{q}}\,dt
(3.12) C(lc0)01θc0H1tφtφt1Hq𝑑t\displaystyle\leq C(l_{c_{0}})\int_{0}^{1}\|\partial_{\theta}c_{0}\|_{H^{1}}\|\partial_{t}\varphi_{t}\circ\varphi_{t}^{-1}\|_{H^{q}}\,dt
(3.13) =C~(lc0,c0H2)01tφtφt1Hq𝑑t,\displaystyle=\tilde{C}(l_{c_{0}},\|c_{0}\|_{H^{2}})\int_{0}^{1}\|\partial_{t}\varphi_{t}\circ\varphi_{t}^{-1}\|_{H^{q}}\,dt\,,

where we used (2.4) with a=qa=q and b=1b=1 in the last inequality. Note that the norm on the right hand side is exactly the right invariant HqH^{q}-norm on Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}). Since this inequality holds for any path connecting id\operatorname{id} to φ1\varphi_{1} in Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}) this implies

(3.14) distGcq(c0,c1)\displaystyle\operatorname{dist}_{G_{c}^{q}}(c_{0},c_{1}) C(c0)distHqDiff(id,φ1)\displaystyle\leq C(c_{0})\operatorname{dist}^{\operatorname{Diff}}_{H^{q}}(\operatorname{id},\varphi_{1})

and thus we obtain the desired result since the geodesic distance of the right invariant HqH^{q}-metric on Diff(𝕊1)\operatorname{Diff}(\mathbb{S}^{1}) vanishes for every q12q\leq\frac{1}{2}, cf. [4]. ∎

4. Completeness Properties

This section concerns the second central goal of this paper: the extension of the results of Bruveris, Michor and Mumford [14, 12] on completeness of integer order Sobolev metrics to fractional orders. The main result of this section is the following:

Theorem 4.1.

If r>32r>\frac{3}{2} then

  1. (1)

    The space r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}) equipped with the geodesic distance distGcr\operatorname{dist}_{G_{c}^{r}} induced by the metric GcrG_{c}^{r} given in (2.17) is metrically complete.

  2. (2)

    The Riemannian manifold (r(𝕊1,d),Gcr)\left(\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}),G_{c}^{r}\right) is geodesically convex, i.e., for any pair of curves c1c_{1} and c2r(𝕊1,d)c_{2}\in\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}) there exists a minimizing geodesic connecting them.

  3. (3)

    If 32<qr\frac{3}{2}<q\leq r, then the Riemannian manifold (r(𝕊1,d),Gcq)\left(\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}),G_{c}^{q}\right) is geodesically complete. This result continues to holds for r=r=\infty, i.e., the Fréchet manifold Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) equipped with the GcqG_{c}^{q}-metric is geodesically complete.

  4. (4)

    If 12<q<r\frac{1}{2}<q<r, then the space r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}) equipped with the geodesic distance distGcq\operatorname{dist}_{G_{c}^{q}} induced by the metric GcqG_{c}^{q} given in (2.17) is metrically incomplete. If q>32q>\frac{3}{2} then the corresponding metric completion is exactly q(𝕊1,d)\mathcal{I}^{q}(\mathbb{S}^{1},\mathbb{R}^{d}).

  5. (5)

    If q<32q<\frac{3}{2} then the Riemannian manifold (r(𝕊1,d),Gcq)\left(\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}),G_{c}^{q}\right) is geodesically incomplete.

Note that fourth claim only discusses the case q>1/2q>1/2 since, by Theorem 3.1, for q1/2q\leq 1/2, the geodesic distance does not induce a metric space structure at all.

The bulk of the work in establishing this theorem lies in proving the first result. The second result then follows from an analogous argument to the integer order case. The proof of the third claim follows from the first by an Ebin-Marsden-type no-loss no-gain result. The fourth part is shown mainly via some rather soft arguments and finally the fifth is proven using an explicit example.

The proof of metric completeness hinges on the following Lemma, which establishes an equivalence of the rrth-order invariant and non-invariant norms on GcrG_{c}^{r}-metric balls.

Lemma 4.2.

If r>32r>\frac{3}{2}, then, for any GcrG_{c}^{r}-metric ball BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r}, there exists a constant α(r,c0,ρ)>0\alpha(r,c_{0},\rho)>0 such that, for all cBGcr(c0,ρ)c\in B_{G_{c}^{r}}(c_{0},\rho) and all hTcrh\in T_{c}\mathcal{I}^{r}, we have

(4.1) α1hHrhGcrαhHr.\alpha^{-1}\left\lVert h\right\rVert_{H^{r}}\leq\left\lVert h\right\rVert_{G_{c}^{r}}\leq\alpha\left\lVert h\right\rVert_{H^{r}}.

To establish this result we first recall a useful lemma for establishing boundedness on GcrG_{c}^{r}-metric balls, cf. [12, Lemma 3.2].

Lemma 4.3.

Let (X,X)(X,\left\lVert\cdot\right\rVert_{X}) be a normed space with f:(r,distGcr)(X,X)f:(\mathcal{I}^{r},\operatorname{dist}_{G_{c}^{r}})\rightarrow(X,\left\lVert\cdot\right\rVert_{X}) a C1C^{1} function. For any crc\in\mathcal{I}^{r} and hTcrh\in T_{c}\mathcal{I}^{r} we denote the derivative of ff at cc in the direction hh by

(4.2) Dc,h(f)=ddt|t=0f(σ(t)),D_{c,h}(f)=\frac{d}{dt}\bigg{|}_{t=0}f(\sigma(t)),

where σ:(ε,ε)r\sigma:(-\varepsilon,\varepsilon)\rightarrow\mathcal{I}^{r} is a C1C^{1} path with σ(0)=c\sigma(0)=c and σ˙(0)=h\dot{\sigma}(0)=h.

Assume that for every metric ball BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) there exists a constant β(r,c0,ρ)>0\beta(r,c_{0},\rho)>0 such that

(4.3) Dc,h(f)Xβ(1+f(c)X)hGcr,\left\lVert D_{c,h}(f)\right\rVert_{X}\leq\beta(1+\left\lVert f(c)\right\rVert_{X})\left\lVert h\right\rVert_{G_{c}^{r}},

for all cBGcr(c0,ρ)c\in B_{G_{c}^{r}}(c_{0},\rho) and hTcrh\in T_{c}\mathcal{I}^{r}. Then ff is Lipschitz continuous, and, in particular, bounded on every GcrG_{c}^{r}-metric ball BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r}.

The above lemma also holds for vector spaces equipped with semi-norms. In particular we may consider mappings such as f:(r,distGcr)(Hp,H˙p)f:(\mathcal{I}^{r},\operatorname{dist}_{G_{c}^{r}})\rightarrow(H^{p},\left\lVert\cdot\right\rVert_{\dot{H}^{p}}).

Throughout this section, for notational simplicity, we will denote ABA\lesssim B when there exists a constant c>0c>0, depending on rr, c0c_{0} and ρ\rho, such that AcBA\leq cB on a GcrG_{c}^{r}-metric ball BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r}. Similarly, we will write ABA\simeq B if ABA\lesssim B and BAB\lesssim A.

Lemma 4.4.

If r>32r>\frac{3}{2}, then the following functions are bounded on every GcrG_{c}^{r}-metric ball BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r} by a constant depending only on rr, c0c_{0} and ρ\rho

(r,distGcr)(,||);clc,\displaystyle(\mathcal{I}^{r},\operatorname{dist}_{G_{c}^{r}})\rightarrow(\mathbb{R},\left\lvert\cdot\right\rvert)\ ;\ c\mapsto l_{c}\ ,
(r,distGcr)(,||);clc1,\displaystyle(\mathcal{I}^{r},\operatorname{dist}_{G_{c}^{r}})\rightarrow(\mathbb{R},\left\lvert\cdot\right\rvert)\ ;\ c\mapsto l_{c}^{-1}\ ,
(r,distGcr)(L(𝕊1,),L);c|cθ|,\displaystyle(\mathcal{I}^{r},\operatorname{dist}_{G_{c}^{r}})\rightarrow(L^{\infty}(\mathbb{S}^{1},\mathbb{R}),\left\lVert\cdot\right\rVert_{L^{\infty}})\ ;\ c\mapsto\left\lvert c_{\theta}\right\rvert\ ,
(r,distGcr)(L(𝕊1,),L);c|cθ|1.\displaystyle(\mathcal{I}^{r},\operatorname{dist}_{G_{c}^{r}})\rightarrow(L^{\infty}(\mathbb{S}^{1},\mathbb{R})\ ,\left\lVert\cdot\right\rVert_{L^{\infty}})\ ;\ c\mapsto\left\lvert c_{\theta}\right\rvert^{-1}.

It should be noted that one can actually show that, for r>32r>\frac{3}{2}, the function clcc\mapsto l_{c} is bounded on GcqG_{c}^{q}-metric balls in r\mathcal{I}^{r} for 1qr1\leq q\leq r. This can be achieved by carefully following the proof below and suitably adapting Lemma 4.3.

Proof.

Calculating the derivative of clcc\mapsto l_{c} in the direction of hTcrh\in T_{c}\mathcal{I}^{r} we acquire

(4.4) Dc,h(lc)=ddt|t=0lσ(t)=ddt|t=0(01|σθ(t)|𝑑θ)=𝕊1Dsh,Dsc𝑑s,D_{c,h}(l_{c})=\frac{d}{dt}\bigg{|}_{t=0}l_{\sigma(t)}=\frac{d}{dt}\bigg{|}_{t=0}\left(\int_{0}^{1}\left\lvert\sigma_{\theta}(t)\right\rvert\ d\theta\right)=\int_{\mathbb{S}^{1}}\left\langle D_{s}h,D_{s}c\right\rangle\ ds,

where Dsc=cθ|cθ|D_{s}c=\frac{c_{\theta}}{\left\lvert c_{\theta}\right\rvert} is the unit tangent vector to cc. From this we estimate

|Dc,h(lc)|\displaystyle\left\lvert D_{c,h}(l_{c})\right\rvert =|𝕊1Dsh,Dsc𝑑s|𝕊1|Dsh|𝑑slc12DshGc0=lc12hG˙c1,\displaystyle=\left|\int_{\mathbb{S}^{1}}\left\langle D_{s}h,D_{s}c\right\rangle\ ds\right|\leq\int_{\mathbb{S}^{1}}\left\lvert D_{s}h\right\rvert\ ds\leq l_{c}^{\frac{1}{2}}\left\lVert D_{s}h\right\rVert_{G_{c}^{0}}=l_{c}^{\frac{1}{2}}\left\lVert h\right\rVert_{\dot{G}_{c}^{1}},

where the second inequality follows from Hölder inequality. If lc>1l_{c}>1 we have

|Dc,h(lc)|\displaystyle\left\lvert D_{c,h}(l_{c})\right\rvert lc12hG˙c1lchG˙c1lchGcr.\displaystyle\leq l_{c}^{\frac{1}{2}}\left\lVert h\right\rVert_{\dot{G}_{c}^{1}}\leq l_{c}\left\lVert h\right\rVert_{\dot{G}_{c}^{1}}\leq l_{c}\left\lVert h\right\rVert_{{G}_{c}^{r}}.

where in the last inequality we have used (2.23). On the other hand, if lc1l_{c}\leq 1 we have

|Dc,h(lc)|\displaystyle\left\lvert D_{c,h}(l_{c})\right\rvert lc12hG˙c1hG˙c1hGcr,\displaystyle\leq l_{c}^{\frac{1}{2}}\left\lVert h\right\rVert_{\dot{G}_{c}^{1}}\leq{\left\lVert h\right\rVert_{\dot{G}_{c}^{1}}\leq\left\lVert h\right\rVert_{G_{c}^{r}}},

where again in the last inequality we have used (2.23). In either scenario the first result follows from Lemma 4.3.

For the second estimate we compute

Dc,h(lc1)\displaystyle D_{c,h}(l_{c}^{-1}) =lc2Dc,h(lc).\displaystyle=l_{c}^{-2}D_{c,h}(l_{c}).

If lc>1l_{c}>1, using the above we have

|Dc,h(lc1)|\displaystyle\left\lvert D_{c,h}(l_{c}^{-1})\right\rvert =lc2|Dc,h(lc)|lc2lchGcr=lc1hGcr.\displaystyle=l_{c}^{-2}\left\lvert D_{c,h}(l_{c})\right\rvert\leq l_{c}^{-2}l_{c}\left\lVert h\right\rVert_{{G}_{c}^{r}}=l_{c}^{-1}\left\lVert h\right\rVert_{{G}_{c}^{r}}.

On the other hand, if lc1l_{c}\leq 1 we estimate

|Dc,h(lc1)|\displaystyle\left\lvert D_{c,h}(l_{c}^{-1})\right\rvert =lc2|Dc,h(lc)|lc2lc12hG˙c1=lc2hψc1H˙1\displaystyle=l_{c}^{-2}\left\lvert D_{c,h}(l_{c})\right\rvert\leq l_{c}^{-2}l_{c}^{\frac{1}{2}}\left\lVert h\right\rVert_{\dot{G}_{c}^{1}}=l_{c}^{-2}\left\lVert h\circ\psi_{c}^{-1}\right\rVert_{\dot{H}^{1}}
lc2hψc1H˙r=lc2lcr12hG˙cr=lc1lcr32hG˙cr\displaystyle\leq l_{c}^{-2}\left\lVert h\circ\psi_{c}^{-1}\right\rVert_{\dot{H}^{r}}=l_{c}^{-2}l_{c}^{r-\frac{1}{2}}\left\lVert h\right\rVert_{\dot{G}_{c}^{r}}=l_{c}^{-1}l_{c}^{r-\frac{3}{2}}\left\lVert h\right\rVert_{\dot{G}_{c}^{r}}
lc1hG˙cr,\displaystyle\leq l_{c}^{-1}\left\lVert h\right\rVert_{\dot{G}_{c}^{r}},

where in the second and third equalities we have used (2.21), in the second inequality we have used (2.3) and in the final inequality we have used the fact that r>32r>\frac{3}{2}. The result once again follows from Lemma 4.3.

To establish the last two results simultaneously, we prove the boundedness of the map clog|cθ|c\mapsto\log\left\lvert c_{\theta}\right\rvert. Calculating the derivative in the direction of hTcrh\in T_{c}\mathcal{I}^{r} we have

Dc,h(log|cθ|)=1|cθ|Dc,h(|cθ|)=Dsh,Dsc.\displaystyle D_{c,h}(\log\left\lvert c_{\theta}\right\rvert)=\frac{1}{\left\lvert c_{\theta}\right\rvert}D_{c,h}(\left\lvert c_{\theta}\right\rvert)=\left\langle D_{s}h,D_{s}c\right\rangle.

From this we estimate

Dc,h(log|cθ|)L\displaystyle\left\lVert D_{c,h}(\log\left\lvert c_{\theta}\right\rvert)\right\rVert_{L^{\infty}} =Dsh,DscLDshLDshGcr1hGcr,\displaystyle=\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle\right\rVert_{L^{\infty}}\leq\left\lVert D_{s}h\right\rVert_{L^{\infty}}\lesssim\left\lVert D_{s}h\right\rVert_{G_{c}^{r-1}}\lesssim\left\lVert h\right\rVert_{G_{c}^{r}},

where the last two inequalities follow from Lemma 2.24, (2.22) and the boundedness of lcl_{c} and lc1l_{c}^{-1} on GcrG_{c}^{r}-metric balls. The result then follows from Lemma 4.3. ∎

We now prove a weaker version of Lemma 4.2. We show that, for r>32r>\frac{3}{2} and 0p10\leq p\leq 1, on GcrG_{c}^{r}-metric balls in r\mathcal{I}^{r} we have the equivalence G˙cpH˙p\left\lVert\cdot\right\rVert_{\dot{G}_{c}^{p}}\simeq\left\lVert\cdot\right\rVert_{\dot{H}^{p}}; whereas Lemma 4.2 is concerned with the equivalence GcrHr\left\lVert\cdot\right\rVert_{{G}_{c}^{r}}\simeq\left\lVert\cdot\right\rVert_{{H}^{r}}.

Lemma 4.5.

Let r>32r>\frac{3}{2} and 0p10\leq p\leq 1. Then, for any GcrG_{c}^{r}-metric ball BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r}, there exists an β(c0,ρ,r,p)>0\beta(c_{0},\rho,{r},p)>0 such that, for all cBGcr(c0,ρ)c\in B_{G_{c}^{r}}(c_{0},\rho) and all hTcrh\in T_{c}\mathcal{I}^{r}, we have

(4.5) β1hH˙phG˙cpβhH˙p.\beta^{-1}\left\lVert h\right\rVert_{\dot{H}^{p}}\leq\left\lVert h\right\rVert_{\dot{G}_{c}^{p}}\leq\beta\left\lVert h\right\rVert_{\dot{H}^{p}}.
Proof.

From Lemma 2.4 and (2.7) we estimate

hG˙cp\displaystyle\left\lVert h\right\rVert_{\dot{G}_{c}^{p}} =lc12phψc1H˙p\displaystyle=l_{c}^{\frac{1}{2}-p}\left\lVert h\circ\psi_{c}^{-1}\right\rVert_{\dot{H}^{p}}
lc12p|cθ|lcL1p2lc|cθ|Lp2hH˙p\displaystyle\leq l_{c}^{\frac{1}{2}-p}\left\lVert\frac{\left\lvert c_{\theta}\right\rvert}{l_{c}}\right\rVert_{L^{\infty}}^{\frac{1-p}{2}}\left\lVert\frac{l_{c}}{\left\lvert c_{\theta}\right\rvert}\right\rVert_{L^{\infty}}^{\frac{p}{2}}\left\lVert h\right\rVert_{\dot{H}^{p}}
=|cθ|L1p2|cθ|1Lp2hH˙p\displaystyle=\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{L^{\infty}}^{\frac{1-p}{2}}\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{L^{\infty}}^{\frac{p}{2}}\left\lVert h\right\rVert_{\dot{H}^{p}}

and

hG˙cp\displaystyle\left\lVert h\right\rVert_{\dot{G}_{c}^{p}} =lc12phψc1H˙p\displaystyle=l_{c}^{\frac{1}{2}-p}\left\lVert h\circ\psi_{c}^{-1}\right\rVert_{\dot{H}^{p}}
lc12plc|cθ|L1+p2|cθ|lcLp2hH˙p\displaystyle\geq l_{c}^{\frac{1}{2}-p}\left\lVert\frac{l_{c}}{\left\lvert c_{\theta}\right\rvert}\right\rVert_{L^{\infty}}^{\frac{-1+p}{2}}\left\lVert\frac{\left\lvert c_{\theta}\right\rvert}{l_{c}}\right\rVert_{L^{\infty}}^{-\frac{p}{2}}\left\lVert h\right\rVert_{\dot{H}^{p}}
=|cθ|1L1+p2|cθ|Lp2hH˙p.\displaystyle=\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{L^{\infty}}^{\frac{-1+p}{2}}\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{L^{\infty}}^{-\frac{p}{2}}\left\lVert h\right\rVert_{\dot{H}^{p}}.

The result then follows from Lemma 4.4. ∎

The next two lemmas will play key technical roles in the proof of Lemma 4.2.

Lemma 4.6.

For r>3/2r>3/2 the following functions are bounded on every GcrG_{c}^{r}-metric ball BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r} by a constant depending only on r,c0r,c_{0} and ρ\rho:

(r,distGcr)(H˙r~(𝕊1,d),H˙r~);cDsc,\displaystyle(\mathcal{I}^{r},\operatorname{dist}_{G_{c}^{r}})\rightarrow(\dot{H}^{{\tilde{r}}}(\mathbb{S}^{1},\mathbb{R}^{d}),\left\lVert\cdot\right\rVert_{\dot{H}^{{\tilde{r}}}})\ ;\ c\mapsto D_{s}c\ ,
(r,distGcr)(H˙r~(𝕊1,d),H˙r~);c|cθ|±1,\displaystyle(\mathcal{I}^{r},\operatorname{dist}_{G_{c}^{r}})\rightarrow(\dot{H}^{{\tilde{r}}}(\mathbb{S}^{1},\mathbb{R}^{d}),\left\lVert\cdot\right\rVert_{\dot{H}^{{\tilde{r}}}})\ ;\ c\mapsto\left\lvert c_{\theta}\right\rvert^{\pm 1},

where r~=min{r1,1}\tilde{r}=\min\{r-1,1\}.

Proof.

Note that, for r>32r>\frac{3}{2}, we have both 12<r~1\frac{1}{2}<\tilde{r}\leq 1 and r~+1r\tilde{r}+1\leq r. Computing the derivatives of both functions in the direction hTcrh\in T_{c}\mathcal{I}^{r} we have

(4.6) Dc,h(Dsc)=DshDsh,DscDscD_{c,h}(D_{s}c)=D_{s}h-\left\langle D_{s}h,D_{s}c\right\rangle D_{s}c

and

(4.7) Dc,h(|cθ|)=Dsh,Dsc|cθ|.D_{c,h}(\left\lvert c_{\theta}\right\rvert)=\left\langle D_{s}h,D_{s}c\right\rangle\left\lvert c_{\theta}\right\rvert.

Applying the triangle inequality to (4.6) we have

Dc,h(Dsc)H˙r~\displaystyle\left\lVert D_{c,h}(D_{s}c)\right\rVert_{\dot{H}^{\tilde{r}}} DshH˙r~+Dsh,DscDscH˙r~.\displaystyle\leq\left\lVert D_{s}h\right\rVert_{\dot{H}^{\tilde{r}}}+\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle D_{s}c\right\rVert_{\dot{H}^{\tilde{r}}}.

For the first term above we apply Lemma 4.5 and obtain

DshH˙r~\displaystyle\left\lVert D_{s}h\right\rVert_{\dot{H}^{\tilde{r}}} DshG˙cr~=hG˙crhGcr.\displaystyle\simeq\left\lVert D_{s}h\right\rVert_{\dot{G}_{c}^{\tilde{r}}}=\left\lVert h\right\rVert_{\dot{G}_{c}^{r}}\leq\left\lVert h\right\rVert_{G_{c}^{r}}.

While, for the second term, noting that DscL=1\left\lVert D_{s}c\right\rVert_{L^{\infty}}=1, we apply (2.6) and obtain

(4.8) Dsh,DscDscH˙r~Dsh,DscH˙r~+Dsh,DscLDscH˙r~Dsh,DscH˙r~+DshLDscLDscH˙r~Dsh,DscH˙r~+DshLDscH˙r~.\begin{split}\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle D_{s}c\right\rVert_{\dot{H}^{\tilde{r}}}&\lesssim\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle\right\rVert_{\dot{H}^{\tilde{r}}}+\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle\right\rVert_{L^{\infty}}\left\lVert D_{s}c\right\rVert_{\dot{H}^{\tilde{r}}}\\ &\lesssim\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle\right\rVert_{\dot{H}^{\tilde{r}}}+\left\lVert D_{s}h\right\rVert_{L^{\infty}}\left\lVert D_{s}c\right\rVert_{L^{\infty}}\left\lVert D_{s}c\right\rVert_{\dot{H}^{\tilde{r}}}\\ &\lesssim\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle\right\rVert_{\dot{H}^{\tilde{r}}}+\left\lVert D_{s}h\right\rVert_{L^{\infty}}\left\lVert D_{s}c\right\rVert_{\dot{H}^{\tilde{r}}}.\end{split}

We approach the term Dsh,DscH˙r~\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle\right\rVert_{\dot{H}^{\tilde{r}}} in an identical fashion, applying (2.6)

(4.9) Dsh,DscH˙r~DshH˙r~+DshLDscH˙r~.\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle\right\rVert_{\dot{H}^{\tilde{r}}}\lesssim\left\lVert D_{s}h\right\rVert_{\dot{H}^{\tilde{r}}}+\left\lVert D_{s}h\right\rVert_{L^{\infty}}\left\lVert D_{s}c\right\rVert_{\dot{H}^{\tilde{r}}}.

By Lemmas 2.5 and 4.4 and (2.22) we have DshLDshGcr~hGcr\left\lVert D_{s}h\right\rVert_{L^{\infty}}\lesssim\left\lVert D_{s}h\right\rVert_{G_{c}^{\tilde{r}}}\leq\left\lVert h\right\rVert_{G_{c}^{r}} on GcrG_{c}^{r}-metric balls. Combining all of this, (4.8) takes the form

Dsh,DscDscH˙r~\displaystyle\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle D_{s}c\right\rVert_{\dot{H}^{\tilde{r}}} hGcr+DscH˙r~hGcr+DscH˙r~hGcr\displaystyle\lesssim\left\lVert h\right\rVert_{G_{c}^{r}}+\left\lVert D_{s}c\right\rVert_{\dot{H}^{\tilde{r}}}\left\lVert h\right\rVert_{G_{c}^{r}}+\left\lVert D_{s}c\right\rVert_{\dot{H}^{\tilde{r}}}\left\lVert h\right\rVert_{G_{c}^{r}}
(1+DscH˙r~)hGcr.\displaystyle\lesssim\left(1+\left\lVert D_{s}c\right\rVert_{\dot{H}^{\tilde{r}}}\right)\left\lVert h\right\rVert_{G_{c}^{r}}.

Hence, by Lemma 4.3 we have that DscH˙r~\left\lVert D_{s}c\right\rVert_{\dot{H}^{\tilde{r}}} is bounded on GcrG_{c}^{r}-metric balls BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r} by a constant depending only on r,c0r,c_{0} and ρ\rho.

We now turn our attention to the map

(r,distGcr)(H˙r~(𝕊1,d),H˙r~);c|cθ|.\displaystyle(\mathcal{I}^{r},\operatorname{dist}_{G_{c}^{r}})\rightarrow(\dot{H}^{{\tilde{r}}}(\mathbb{S}^{1},\mathbb{R}^{d}),\left\lVert\cdot\right\rVert_{\dot{H}^{{\tilde{r}}}})\ ;\ c\mapsto\left\lvert c_{\theta}\right\rvert.

Using (4.7) and applying (2.5) with f=Dsh,Dsc,g=|cθ|,a=b=r~>12f=\left\langle D_{s}h,D_{s}c\right\rangle,g=\left\lvert c_{\theta}\right\rvert,a=b=\tilde{r}>\frac{1}{2} we have

(4.10) Dc,h(|cθ|)H˙r~=Dsh,Dsc|cθ|H˙r~|Dsh,Dsc^(0)||cθ|H˙r~+||cθ|^(0)|Dsh,DscH˙r~+Dsh,DscH˙r~|cθ|H˙r~.\begin{split}\left\lVert D_{c,h}(\left\lvert c_{\theta}\right\rvert)\right\rVert_{\dot{H}^{\tilde{r}}}&=\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{\tilde{r}}}\\ &\lesssim\left\lvert\widehat{\left\langle D_{s}h,D_{s}c\right\rangle}(0)\right\rvert\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{\tilde{r}}}+\left\lvert\widehat{\left\lvert c_{\theta}\right\rvert}(0)\right\rvert\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle\right\rVert_{\dot{H}^{\tilde{r}}}\\ &\quad+\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle\right\rVert_{\dot{H}^{\tilde{r}}}\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{\tilde{r}}}.\end{split}

For the first term of (4.10) we have, using the above, that

|Dsh,Dsc^(0)|\displaystyle\left\lvert\widehat{\left\langle D_{s}h,D_{s}c\right\rangle}(0)\right\rvert Dsh,DscLDshLDscLhGcr.\displaystyle\leq\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle\right\rVert_{L^{\infty}}\lesssim\left\lVert D_{s}h\right\rVert_{L^{\infty}}\left\lVert D_{s}c\right\rVert_{L^{\infty}}\leq\left\lVert h\right\rVert_{G_{c}^{r}}.

For the second term we estimate ||cθ|^(0)||cθ|L\left\lvert\widehat{\left\lvert c_{\theta}\right\rvert}(0)\right\rvert\leq\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{L^{\infty}} and recall, by Lemma 4.4, that this is bounded on GcrG_{c}^{r}-metric balls BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) by a constant depending only on r,c0r,c_{0} and ρ\rho. Recycling (4.9) and remarking that, from the above argument, we now have that DscH˙r~\left\lVert D_{s}c\right\rVert_{\dot{H}^{\tilde{r}}} is bounded on GcrG_{c}^{r}-metric balls BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r} by a constant depending only on r,c0r,c_{0} and ρ\rho, we acquire

Dsh,DscH˙r~\displaystyle\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle\right\rVert_{\dot{H}^{\tilde{r}}} DshH˙r~+DshLDscH˙r~hGcr.\displaystyle\lesssim\left\lVert D_{s}h\right\rVert_{\dot{H}^{\tilde{r}}}+\left\lVert D_{s}h\right\rVert_{L^{\infty}}\left\lVert D_{s}c\right\rVert_{\dot{H}^{\tilde{r}}}\lesssim\left\lVert h\right\rVert_{G_{c}^{r}}.

Combining all of this, (4.10) becomes

(4.11) Dc,h(|cθ|)H˙r~|cθ|H˙r~hGcr+hGcr+|cθ|H˙r~hGcr(1+|cθ|H˙r~)hGcr.\begin{split}\left\lVert D_{c,h}(\left\lvert c_{\theta}\right\rvert)\right\rVert_{\dot{H}^{\tilde{r}}}&\lesssim\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{\tilde{r}}}\left\lVert h\right\rVert_{G_{c}^{r}}+\left\lVert h\right\rVert_{G_{c}^{r}}+\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{\tilde{r}}}\left\lVert h\right\rVert_{G_{c}^{r}}\\ &\lesssim\left(1+\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{\tilde{r}}}\right)\left\lVert h\right\rVert_{G_{c}^{r}}.\end{split}

Hence, by Lemma 4.3 we have that |cθ|H˙r~\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{\tilde{r}}} is bounded on GcrG_{c}^{r}-metric balls BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r} by a constant depending only on r,c0r,c_{0} and ρ\rho. Similarly, |cθ|1H˙r~\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{\dot{H}^{\tilde{r}}} is bounded on GcrG_{c}^{r}-metric balls BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r} by a constant depending only on r,c0r,c_{0} and ρ\rho. ∎

Lemma 4.7.

If r>2r>2 with decomposition r=p+nr=p+n for some 0<p10<p\leq 1 and n2n\geq 2 an integer, then, for 1kn1\leq k\leq n, the following functions are bounded on every GcrG_{c}^{r}-metric ball BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r} by a constant depending only on r,p,n,k,c0r,p,n,k,c_{0} and ρ\rho.

(r,distGcr)(H˙p(𝕊1,),H˙p);cDskc,\displaystyle(\mathcal{I}^{r},\operatorname{dist}_{G_{c}^{r}})\rightarrow(\dot{H}^{p}(\mathbb{S}^{1},\mathbb{R}),\left\lVert\cdot\right\rVert_{\dot{H}^{p}})\ ;\ c\mapsto D_{s}^{k}c\ ,
(r,distGcr)(H˙p(𝕊1,),H˙p);cDsk1|cθ|,\displaystyle(\mathcal{I}^{r},\operatorname{dist}_{G_{c}^{r}})\rightarrow(\dot{H}^{p}(\mathbb{S}^{1},\mathbb{R}),\left\lVert\cdot\right\rVert_{\dot{H}^{p}})\ ;\ c\mapsto D_{s}^{k-1}\left\lvert c_{\theta}\right\rvert\ ,
(r,distGcr)(H˙p+k1(𝕊1,),H˙p+k1);c|cθ|±1.\displaystyle(\mathcal{I}^{r},\operatorname{dist}_{G_{c}^{r}})\rightarrow(\dot{H}^{p+{k-1}}(\mathbb{S}^{1},\mathbb{R}),\left\lVert\cdot\right\rVert_{\dot{H}^{p+{k-1}}})\ ;\ c\mapsto\left\lvert c_{\theta}\right\rvert^{\pm 1}.
Proof.

Notice that the k=1k=1 case for each function follows immediately from Lemma 4.6. We proceed by induction on kk. For k2k\geq 2 recall the following formula from [12, Lemma 3.3].

(4.12) Dc,h(Dskc)=DskhDskh,DscDsckDsh,DscDskcDsh,DskcDsc+lower order terms,\begin{split}D_{c,h}\left(D_{s}^{k}c\right)&=D_{s}^{k}h-\left\langle D_{s}^{k}h,D_{s}c\right\rangle D_{s}c-k\left\langle D_{s}h,D_{s}c\right\rangle D_{s}^{k}c\\ &\qquad-\left\langle D_{s}h,D_{s}^{k}c\right\rangle D_{s}c+\text{lower order terms,}\end{split}

where the lower order terms include only products of terms with less than kk derivative. Applying the triangle inequality and ignoring the contributions of the lower order terms (one can readily show that these terms are bounded by hGcr\left\lVert h\right\rVert_{G_{c}^{r}} on GcrG_{c}^{r}-metric balls up to constants depending only on r,p,k,c0r,p,k,c_{0} and ρ\rho) we have

(4.13) Dc,h(Dskc)H˙pDskhH˙p+Dskh,DscDscH˙p+kDsh,DscDskcH˙p+Dsh,DskcDscH˙p+hGcr.\begin{split}\left\lVert D_{c,h}\left(D_{s}^{k}c\right)\right\rVert_{\dot{H}^{p}}&\lesssim\left\lVert D_{s}^{k}h\right\rVert_{\dot{H}^{p}}+\left\lVert\left\langle D_{s}^{k}h,D_{s}c\right\rangle D_{s}c\right\rVert_{\dot{H}^{p}}+k\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle D_{s}^{k}c\right\rVert_{\dot{H}^{p}}\\ &\quad+\left\lVert\left\langle D_{s}h,D_{s}^{k}c\right\rangle D_{s}c\right\rVert_{\dot{H}^{p}}+{\left\lVert h\right\rVert_{G_{c}^{r}}}.\end{split}

For the first term we apply Lemma 4.5, (2.22) and Lemma 4.4

DskhH˙p\displaystyle\left\lVert D_{s}^{k}h\right\rVert_{\dot{H}^{p}} DskhG˙cp=hG˙cp+klcrpkhG˙crhG˙crhGcr.\displaystyle\simeq\left\lVert D_{s}^{k}h\right\rVert_{\dot{G}_{c}^{p}}=\left\lVert h\right\rVert_{\dot{G}_{c}^{p+k}}\leq l_{c}^{r-p-k}\left\lVert h\right\rVert_{\dot{G}_{c}^{r}}\lesssim\left\lVert h\right\rVert_{\dot{G}_{c}^{r}}\leq\left\lVert h\right\rVert_{G_{c}^{r}}.

For the second term in (4.13) we first apply (2.5) with f=Dskh,Dscf=\left\langle D_{s}^{k}h,D_{s}c\right\rangle, g=Dscg=D_{s}c, a=pa=p and b=1b=1

(4.14) Dskh,DscDscH˙p|Dskh,Dsc^(0)|DscH˙p+|Dsc^(0)|Dskh,DscH˙p+Dskh,DscH˙pDscH˙1.\begin{split}\left\lVert\left\langle D_{s}^{k}h,D_{s}c\right\rangle D_{s}c\right\rVert_{\dot{H}^{p}}&\lesssim\left\lvert\widehat{\left\langle D_{s}^{k}h,D_{s}c\right\rangle}(0)\right\rvert\left\lVert D_{s}c\right\rVert_{\dot{H}^{p}}\\ &\quad+\left\lvert\widehat{D_{s}c}(0)\right\rvert\left\lVert\left\langle D_{s}^{k}h,D_{s}c\right\rangle\right\rVert_{\dot{H}^{p}}\\ &\quad+\left\lVert\left\langle D_{s}^{k}h,D_{s}c\right\rangle\right\rVert_{\dot{H}^{p}}\left\lVert D_{s}c\right\rVert_{\dot{H}^{1}}.\end{split}

For the first term of (4.14) we apply the Cauchy-Schwartz inequality for the L2L^{2}-inner product

|Dskh,Dsc^(0)|\displaystyle\left\lvert\widehat{\left\langle D_{s}^{k}h,D_{s}c\right\rangle}(0)\right\rvert =|𝕊1Dskh,Dsc𝑑θ|DskhL2DscL2.\displaystyle=\left\lvert\int_{\mathbb{S}^{1}}\left\langle D_{s}^{k}h,D_{s}c\right\rangle\,d\theta\right\rvert\lesssim\left\lVert D_{s}^{k}h\right\rVert_{L^{2}}\left\lVert D_{s}c\right\rVert_{L^{2}}.

By Lemma 4.5, (2.22) and Lemma 4.4 we have

DskhL2\displaystyle\left\lVert D_{s}^{k}h\right\rVert_{L^{2}} DskhGc0=hG˙cklcrkhG˙crhG˙crhGcr.\displaystyle\simeq\left\lVert D_{s}^{k}h\right\rVert_{G_{c}^{0}}=\left\lVert h\right\rVert_{\dot{G}_{c}^{k}}\leq l_{c}^{r-k}\left\lVert h\right\rVert_{\dot{G}_{c}^{r}}\lesssim\left\lVert h\right\rVert_{\dot{G}_{c}^{r}}\leq\left\lVert h\right\rVert_{G_{c}^{r}}.

Recalling now that DscD_{s}c is a unit vector we have DscL2=1\left\lVert D_{s}c\right\rVert_{L^{2}}=1 and |Dsc^(0)|DscL=1\left\lvert\widehat{D_{s}c}(0)\right\rvert\leq\left\lVert D_{s}c\right\rVert_{L^{\infty}}=1. By Lemma 4.6, DscH˙1\left\lVert D_{s}c\right\rVert_{\dot{H}^{1}} is bounded on GcrG_{c}^{r}-metric balls in r\mathcal{I}^{r} by constants depending only on r,p,c0r,p,c_{0} and ρ\rho, and thus, by (2.22), also DscH˙p\left\lVert D_{s}c\right\rVert_{\dot{H}^{p}}. Combining all this gives us |Dskh,Dsc^(0)|hGcr\left\lvert\widehat{\left\langle D_{s}^{k}h,D_{s}c\right\rangle}(0)\right\rvert\lesssim\left\lVert h\right\rVert_{G_{c}^{r}} and (4.14) becomes

(4.15) Dskh,DscDscH˙pDskh,DscH˙p+hGcr.\left\lVert\left\langle D_{s}^{k}h,D_{s}c\right\rangle D_{s}c\right\rVert_{\dot{H}^{p}}\lesssim\left\lVert\left\langle D_{s}^{k}h,D_{s}c\right\rangle\right\rVert_{\dot{H}^{p}}+\left\lVert h\right\rVert_{G_{c}^{r}}.

For the first term in (4.15) we apply (2.5) with f=Dskh,g=Dsc,a=pf=D_{s}^{k}h,g=D_{s}c,a=p and b=1b=1

Dskh,DscH˙p\displaystyle\left\lVert\left\langle D_{s}^{k}h,D_{s}c\right\rangle\right\rVert_{\dot{H}^{p}} |Dskh^(0)|DscH˙p+|Dsc^(0)|DskhH˙p+DskhH˙pDscH˙1.\displaystyle\lesssim\left\lvert\widehat{D_{s}^{k}h}(0)\right\rvert\left\lVert D_{s}c\right\rVert_{\dot{H}^{p}}+\left\lvert\widehat{D_{s}c}(0)\right\rvert\left\lVert D_{s}^{k}h\right\rVert_{\dot{H}^{p}}+\left\lVert D_{s}^{k}h\right\rVert_{\dot{H}^{p}}\left\lVert D_{s}c\right\rVert_{\dot{H}^{1}}.

Similar to before we estimate using Hölder’s inequality

|Dskh^(0)|=|𝕊1Dskh𝑑θ|DskhL1DskhL2.\displaystyle\left\lvert\widehat{D_{s}^{k}h}(0)\right\rvert=\left\lvert\int_{\mathbb{S}^{1}}D_{s}^{k}h\ d\theta\right\rvert\leq\left\lVert D_{s}^{k}h\right\rVert_{L^{1}}\lesssim\left\lVert D_{s}^{k}h\right\rVert_{L^{2}}.

Recalling from above that DskhL2hGcr\left\lVert D_{s}^{k}h\right\rVert_{L^{2}}\lesssim\left\lVert h\right\rVert_{G_{c}^{r}} and DskhH˙phGcr\left\lVert D_{s}^{k}h\right\rVert_{\dot{H}^{p}}\lesssim\left\lVert h\right\rVert_{G_{c}^{r}} and the boundedness of DscL2\left\lVert D_{s}c\right\rVert_{L^{2}}, |Dsc^(0)|\left\lvert\widehat{D_{s}c}(0)\right\rvert, DscH˙p\left\lVert D_{s}c\right\rVert_{\dot{H}^{p}} and DscH˙1\left\lVert D_{s}c\right\rVert_{\dot{H}^{1}} on GcrG_{c}^{r}-metric balls in r\mathcal{I}^{r}, (4.15) then becomes

Dskh,DscDscH˙phGcr.\displaystyle\left\lVert\left\langle D_{s}^{k}h,D_{s}c\right\rangle D_{s}c\right\rVert_{\dot{H}^{p}}\lesssim\left\lVert h\right\rVert_{G_{c}^{r}}.

The remaining terms in (4.13) are bounded in an almost identical fashion as

Dsh,DscDskcDskcH˙phGcr\displaystyle\left\lVert\left\langle D_{s}h,D_{s}c\right\rangle D_{s}^{k}c\right\rVert\lesssim\left\lVert D_{s}^{k}c\right\rVert_{\dot{H}^{p}}\left\lVert h\right\rVert_{G_{c}^{r}}

and

Dsh,DskcDscH˙pDskcH˙phGcr.\displaystyle\left\lVert\left\langle D_{s}h,D_{s}^{k}c\right\rangle D_{s}c\right\rVert_{\dot{H}^{p}}\lesssim\left\lVert D_{s}^{k}c\right\rVert_{\dot{H}^{p}}\left\lVert h\right\rVert_{G_{c}^{r}}.

Hence Dc,h(Dskc)H˙p(1+DskcH˙p)hGcr\left\lVert D_{c,h}\left(D_{s}^{k}c\right)\right\rVert_{\dot{H}^{p}}\lesssim(1+\left\lVert D_{s}^{k}c\right\rVert_{\dot{H}^{p}})\left\lVert h\right\rVert_{G_{c}^{r}} on BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) and the first result follows from Lemma 4.3.

The boundedness of the second function on GcrG_{c}^{r}-metric balls can be argued exactly as above using the formula from [12, Lemma 3.3]

(4.16) Dc,h(Dsk1|cθ|)\displaystyle D_{c,h}\left(D_{s}^{k-1}\left\lvert c_{\theta}\right\rvert\right) =Dskh,Dsc|cθ|(k2)Dsh,DscDsk1|cθ|\displaystyle=\left\langle D_{s}^{k}h,D_{s}c\right\rangle\left\lvert c_{\theta}\right\rvert-(k-2)\left\langle D_{s}h,D_{s}c\right\rangle D_{s}^{k-1}\left\lvert c_{\theta}\right\rvert
+Dsh,Dskc|cθ|+lower order terms.\displaystyle\qquad+\left\langle D_{s}h,D_{s}^{k}c\right\rangle\left\lvert c_{\theta}\right\rvert+\text{lower order terms.}

Finally, for bounding the third function we use the boundedness of the second one, together with the expansion

(4.17) θk1|cθ|=j=1k2αAjcj,αi=0k2(θi|cθ|)αiDsk1|cθ|,\partial_{\theta}^{{k-1}}\left\lvert c_{\theta}\right\rvert=\sum_{j=1}^{k-2}\sum_{\alpha\in A_{j}}c_{j,\alpha}\prod_{i=0}^{k-2}\big{(}\partial_{\theta}^{i}\left\lvert c_{\theta}\right\rvert\big{)}^{\alpha_{i}}D_{s}^{{k-1}}\left\lvert c_{\theta}\right\rvert,

where cj,αc_{j,\alpha} are constants and α=(α0,,αk2)\alpha=(\alpha_{0},...,\alpha_{k-2}) are multi-indices with index sets

Aj={α|i=0k2iαi=k1jandi=0k2αi=j}.\displaystyle A_{j}=\left\{\alpha\ \left|\ \sum_{i=0}^{k-2}i\alpha_{i}=k-1-j\ \text{and}\ \sum_{i=0}^{k-2}\alpha_{i}=j\right.\right\}.

Applying the triangle inequality and (2.5) to (4.17) and using an induction argument, we acquire |cθ|H˙p+k11\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{p+k-1}}\lesssim 1 on BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) for all 2kn2\leq k\leq n. To establish the result for c|cθ|1c\mapsto\left\lvert c_{\theta}\right\rvert^{-1}, we simply apply the chain rule to express θk1|cθ|1\partial_{\theta}^{k-1}\left\lvert c_{\theta}\right\rvert^{-1} as a linear combination of powers of |cθ|1\left\lvert c_{\theta}\right\rvert^{-1} and derivatives up to order k1{k-1} of |cθ|\left\lvert c_{\theta}\right\rvert. ∎

Armed with the above, we are now ready to prove the central estimate.

Proof of Lemma  4.2.

Firstly, note that GcrGc0+G˙cr\left\lVert\cdot\right\rVert_{G_{c}^{r}}\simeq\left\lVert\cdot\right\rVert_{G_{c}^{0}}+\left\lVert\cdot\right\rVert_{\dot{G}_{c}^{r}}. The equivalence Gc0L2\left\lVert\cdot\right\rVert_{G_{c}^{0}}\simeq\left\lVert\cdot\right\rVert_{L^{2}} on GcrG_{c}^{r}-metric balls in r\mathcal{I}^{r} follows directly from Lemma 4.5. Hence, to establish the estimate, we need to show the equivalence of the homogeneous norms G˙crH˙r\left\lVert\cdot\right\rVert_{\dot{G}_{c}^{r}}\simeq\left\lVert\cdot\right\rVert_{\dot{H}^{r}} on GcrG_{c}^{r}-metric balls in r\mathcal{I}^{r}.

We begin with the case 32<r2\frac{3}{2}<r\leq 2. As 12<r11\frac{1}{2}<{r-1}\leq 1 we have, by Lemma 4.5, that G˙cr1H˙r1\left\lVert\cdot\right\rVert_{\dot{G}_{c}^{{r-1}}}\simeq\left\lVert\cdot\right\rVert_{\dot{H}^{{r-1}}} on GcrG_{c}^{r}-metric balls in r\mathcal{I}^{r}. From this we have

hG˙cr\displaystyle\left\lVert h\right\rVert_{\dot{G}_{c}^{r}} =DshG˙cr1DshH˙r1=|cθ|1hθH˙r1.\displaystyle=\left\lVert D_{s}h\right\rVert_{\dot{G}_{c}^{{r-1}}}\simeq\left\lVert D_{s}h\right\rVert_{\dot{H}^{{r-1}}}=\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}h_{\theta}\right\rVert_{\dot{H}^{{r-1}}}.

Applying (2.5) with f=|cθ|1f=\left\lvert c_{\theta}\right\rvert^{-1}, g=hθg=h_{\theta} and a=b=r1>12a=b=r-1>\frac{1}{2} we acquire

hG˙cr\displaystyle\left\lVert h\right\rVert_{\dot{G}_{c}^{r}} |cθ|1hθH˙r1\displaystyle\simeq\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}h_{\theta}\right\rVert_{\dot{H}^{{r-1}}}
||cθ|1^(0)|hθH˙r1+|hθ^(0)||cθ|1H˙r1+|cθ|1H˙r1hθH˙r1.\displaystyle\lesssim\left\lvert\widehat{\left\lvert c_{\theta}\right\rvert^{-1}}(0)\right\rvert\left\lVert h_{\theta}\right\rVert_{\dot{H}^{{r-1}}}+\left\lvert\widehat{h_{\theta}}(0)\right\rvert\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{\dot{H}^{{r-1}}}+\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{\dot{H}^{{r-1}}}\left\lVert h_{\theta}\right\rVert_{\dot{H}^{{r-1}}}.

Note now that ||cθ|1^(0)||cθ|1L\left\lvert\widehat{\left\lvert c_{\theta}\right\rvert^{-1}}(0)\right\rvert\leq\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{L^{\infty}} and hθ^(0)=0\widehat{h_{\theta}}(0)=0, which gives us

hG˙cr\displaystyle\left\lVert h\right\rVert_{\dot{G}_{c}^{r}} |cθ|1LhθH˙r1+|cθ|1H˙r1hθH˙r1\displaystyle\lesssim\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{L^{\infty}}\left\lVert h_{\theta}\right\rVert_{\dot{H}^{{r-1}}}+\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{\dot{H}^{{r-1}}}\left\lVert h_{\theta}\right\rVert_{\dot{H}^{{r-1}}}
=(|cθ|1L+|cθ|1H˙r1)hH˙r.\displaystyle=\left(\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{L^{\infty}}+\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{\dot{H}^{{r-1}}}\right)\left\lVert h\right\rVert_{\dot{H}^{r}}.

By Lemmas 4.4 and 4.6, we have that |cθ|1L\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{L^{\infty}} and |cθ|1H˙r1\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{\dot{H}^{{r-1}}} are bounded on GcrG_{c}^{r}-metric balls BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r} by constants depending only on r,c0r,c_{0} and ρ\rho. Hence we have

(4.18) hG˙cr\displaystyle\left\lVert h\right\rVert_{\dot{G}_{c}^{r}} hH˙r\displaystyle\lesssim\left\lVert h\right\rVert_{\dot{H}^{r}}

on GcrG_{c}^{r}-metric balls BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r}.

For the other direction, note that

hH˙r\displaystyle\left\lVert h\right\rVert_{\dot{H}^{r}} =hθH˙r1=|cθ|DshH˙r1.\displaystyle=\left\lVert h_{\theta}\right\rVert_{\dot{H}^{r-1}}=\left\lVert\left\lvert c_{\theta}\right\rvert D_{s}h\right\rVert_{\dot{H}^{r-1}}.

Mirroring the above, we apply (2.5) with f=|cθ|f=\left\lvert c_{\theta}\right\rvert, g=Dshg=D_{s}h, a=b=r1>12a=b=r-1>\frac{1}{2} and acquire

hH˙r\displaystyle\left\lVert h\right\rVert_{\dot{H}^{r}} =|cθ|DshH˙r1\displaystyle=\left\lVert\left\lvert c_{\theta}\right\rvert D_{s}h\right\rVert_{\dot{H}^{{r-1}}}
||cθ|^(0)|DshH˙r1+|Dsh^(0)||cθ|H˙r1+|cθ|H˙r1DshH˙r1.\displaystyle\lesssim\left\lvert\widehat{\left\lvert c_{\theta}\right\rvert}(0)\right\rvert\left\lVert D_{s}h\right\rVert_{\dot{H}^{{r-1}}}+\left\lvert\widehat{D_{s}h}(0)\right\rvert\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{{r-1}}}+\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{{r-1}}}\left\lVert D_{s}h\right\rVert_{\dot{H}^{{r-1}}}.

For the first term we again bound ||cθ|^(0)||cθ|L\left\lvert\widehat{\left\lvert c_{\theta}\right\rvert}(0)\right\rvert\leq\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{L^{\infty}}. For the term |Dsh^(0)|\left\lvert\widehat{D_{s}h}(0)\right\rvert we estimate

|Dsh^(0)|\displaystyle\left\lvert\widehat{D_{s}h}(0)\right\rvert =|𝕊1|cθ|1hθ𝑑θ||cθ|1L2hH˙1|cθ|1LhH˙1,\displaystyle=\left\lvert\int_{\mathbb{S}^{1}}\left\lvert c_{\theta}\right\rvert^{-1}h_{\theta}\ d\theta\right\rvert\leq\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{L^{2}}\left\lVert h\right\rVert_{\dot{H}^{1}}\leq\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{L^{\infty}}\left\lVert h\right\rVert_{\dot{H}^{1}},

where, in the first inequality, we have used Cauchy-Schwartz for the L2L^{2} inner product. This gives us

hH˙r\displaystyle\left\lVert h\right\rVert_{\dot{H}^{r}} |cθ|LDshH˙r1+|cθ|1L|cθ|H˙r1hH˙1+|cθ|H˙r1DshH˙r1.\displaystyle\lesssim\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{L^{\infty}}\left\lVert D_{s}h\right\rVert_{\dot{H}^{{r-1}}}+\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{L^{\infty}}\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{{r-1}}}\left\lVert h\right\rVert_{\dot{H}^{1}}+\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{{r-1}}}\left\lVert D_{s}h\right\rVert_{\dot{H}^{{r-1}}}.

Using Lemma 4.5 we have DshH˙r1DshG˙cr1=hG˙cr\left\lVert D_{s}h\right\rVert_{\dot{H}^{{r-1}}}\simeq\left\lVert D_{s}h\right\rVert_{\dot{G}_{c}^{r-1}}=\left\lVert h\right\rVert_{\dot{G}_{c}^{r}} and hH˙1hG˙c1lcr1hG˙cr\left\lVert h\right\rVert_{\dot{H}^{1}}\simeq\left\lVert h\right\rVert_{\dot{G}_{c}^{1}}\leq l_{c}^{r-1}\left\lVert h\right\rVert_{\dot{G}_{c}^{r}} where, in the final inequality, we have used (2.22). Hence we have

hH˙r\displaystyle\left\lVert h\right\rVert_{\dot{H}^{r}} (|cθ|L+lcr1|cθ|1L|cθ|H˙r1+|cθ|H˙r1)hG˙cr.\displaystyle\lesssim\left(\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{L^{\infty}}+l_{c}^{r-1}\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{L^{\infty}}\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{{r-1}}}+\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{{r-1}}}\right)\left\lVert h\right\rVert_{\dot{G}_{c}^{r}}.

Finally, by Lemmas 4.4 and 4.6, |cθ|L,lc\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{L^{\infty}},l_{c} and |cθ|H˙r1\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{{r-1}}} are bounded on GcrG_{c}^{r}-metric balls BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r} by constants depending only on r,c0r,c_{0} and ρ\rho. Hence we have

(4.19) hH˙rhG˙cr\displaystyle\left\lVert h\right\rVert_{\dot{H}^{r}}\lesssim\left\lVert h\right\rVert_{\dot{G}_{c}^{r}}

on GcrG_{c}^{r}-metric balls BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r}. This delivers the lemma for the cases 32<r2\frac{3}{2}<r\leq 2.

Next consider 2<r2<r with decomposition r=p+nr=p+n for some 0<p10<p\leq 1 and n2n\geq 2 an integer. As 0<p10<p\leq 1 we have, by Lemma 4.5, that G˙cpH˙p\left\lVert\cdot\right\rVert_{\dot{G}_{c}^{p}}\simeq\left\lVert\cdot\right\rVert_{\dot{H}^{p}} on GcrG_{c}^{r}-metric balls in r\mathcal{I}^{r}. From this we have

hG˙cp+1\displaystyle\left\lVert h\right\rVert_{\dot{G}_{c}^{p+1}} =DshG˙cpDshH˙p=|cθ|1hθH˙p.\displaystyle=\left\lVert D_{s}h\right\rVert_{\dot{G}_{c}^{p}}\simeq\left\lVert D_{s}h\right\rVert_{\dot{H}^{p}}=\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}h_{\theta}\right\rVert_{\dot{H}^{p}}.

Repeating the same argument as for (4.18) (with a slight change of using a=pa=p and b=1b=1 instead of a=b=r1a=b=r-1), we obtain

hG˙cp+1\displaystyle\left\lVert h\right\rVert_{\dot{G}_{c}^{p+1}} hH˙p+1\displaystyle\lesssim\left\lVert h\right\rVert_{\dot{H}^{p+1}}

on GcrG_{c}^{r}-metric balls BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r}.

For the other direction, note that

hH˙p+1\displaystyle\left\lVert h\right\rVert_{\dot{H}^{p+1}} =hθH˙p=|cθ|DshH˙p.\displaystyle=\left\lVert h_{\theta}\right\rVert_{\dot{H}^{p}}=\left\lVert\left\lvert c_{\theta}\right\rvert D_{s}h\right\rVert_{\dot{H}^{p}}.

We now repeat the same argument as for (4.19) (again with a=pa=p and b=1b=1 instead of a=b=r1a=b=r-1), and obtain

hH˙p+1hG˙cp+1\displaystyle\left\lVert h\right\rVert_{\dot{H}^{p+1}}\lesssim\left\lVert h\right\rVert_{\dot{G}_{c}^{p+1}}

on GcrG_{c}^{r}-metric balls BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r}.

We now establish an inductive step. Assume that, for some kk with 2kn2\leq k\leq n we have hG˙cp+k1hH˙p+k1\left\lVert h\right\rVert_{\dot{G}_{c}^{p+k-1}}\simeq\left\lVert h\right\rVert_{\dot{H}^{p+k-1}} on GcrG_{c}^{r}-metric balls in r\mathcal{I}^{r}. From this we obtain

hG˙cp+k\displaystyle\left\lVert h\right\rVert_{\dot{G}_{c}^{p+k}} =DshG˙cp+k1DshH˙p+k1=|cθ|1hθH˙p+k1,\displaystyle=\left\lVert D_{s}h\right\rVert_{\dot{G}_{c}^{p+k-1}}\simeq\left\lVert D_{s}h\right\rVert_{\dot{H}^{p+k-1}}=\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}h_{\theta}\right\rVert_{\dot{H}^{p+k-1}},

and

hH˙p+k\displaystyle\left\lVert h\right\rVert_{\dot{H}^{p+k}} =hθH˙p+k1=|cθ|DshH˙p+k1.\displaystyle=\left\lVert h_{\theta}\right\rVert_{\dot{H}^{p+k-1}}=\left\lVert\left\lvert c_{\theta}\right\rvert D_{s}h\right\rVert_{\dot{H}^{p+k-1}}.

By the same argument as above, with a=b=p+k1>1/2a=b=p+k-1>1/2, and using Lemma 4.7 instead of Lemma 4.6 to bound |cθ|H˙p+k1\left\lVert\left\lvert c_{\theta}\right\rvert\right\rVert_{\dot{H}^{p+k-1}} and |cθ|1H˙p+k1\left\lVert\left\lvert c_{\theta}\right\rvert^{-1}\right\rVert_{\dot{H}^{p+k-1}}, we obtain

hG˙cp+k\displaystyle\left\lVert h\right\rVert_{\dot{G}_{c}^{p+k}} hH˙p+k\displaystyle\simeq\left\lVert h\right\rVert_{\dot{H}^{p+k}}

on GcrG_{c}^{r}-metric balls BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) in r\mathcal{I}^{r}. The result for 2<r2<r now follows by induction on kk. ∎

Using Lemma 4.2 we can now relate the induced geodesic distance on (r,Gcr)(\mathcal{I}^{r},G_{c}^{r}) to the standard norm distance on the ambient linear space (Hr,Hr)(H^{r},\left\lVert\cdot\right\rVert_{H^{r}}).

Lemma 4.8.

Let r>32r>\frac{3}{2}. Then, for every GcrG_{c}^{r}-metric ball BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho) and every c1,c2BGcr(c0,ρ)c_{1},c_{2}\in B_{G_{c}^{r}}(c_{0},\rho), we have

(4.20) c2c1Hrα(r,c0,4ρ)distGcr(c1,c2),\left\lVert c_{2}-c_{1}\right\rVert_{H^{r}}\leq\alpha(r,c_{0},4\rho)\operatorname{dist}_{G_{c}^{r}}(c_{1},c_{2}),

where α\alpha is as in Lemma 4.2.

Proof.

Let c1,c2BGcr(c0,ρ)c_{1},c_{2}\in B_{G_{c}^{r}}(c_{0},\rho) and σ:[0,1]r\sigma:[0,1]\rightarrow\mathcal{I}^{r} a piecewise smooth curve connecting them with GcrG_{c}^{r}-length LGcr(σ)<distGcr(c1,c2)+εL_{G_{c}^{r}}(\sigma)<\operatorname{dist}_{G_{c}^{r}}(c_{1},c_{2})+\varepsilon. Then, for ε<ρ\varepsilon<\rho, we have:

LGcr(σ)<distGcr(c1,c2)+ε<3ρ.L_{G_{c}^{r}}(\sigma)<\operatorname{dist}_{G_{c}^{r}}(c_{1},c_{2})+\varepsilon<3\rho.

Hence, as

distGcr(σ(t),c0)distGcr(σ(t),c1)+distGcr(c1,c0)LGcr(σ)+ρ<3ρ+ε<4ρ,\operatorname{dist}_{G_{c}^{r}}(\sigma(t),c_{0})\leq\operatorname{dist}_{G_{c}^{r}}(\sigma(t),c_{1})+\operatorname{dist}_{G_{c}^{r}}(c_{1},c_{0})\leq L_{G_{c}^{r}}(\sigma)+\rho<3\rho+\varepsilon<4\rho,

we have that σ([0,1])BGcr(c0,4ρ)\sigma([0,1])\subset B_{G_{c}^{r}}(c_{0},4\rho). Finally, applying Lemma 4.2, we have that

c1c2HrLHr(σ)=01σ˙Hr𝑑tα(r,c0,4ρ)01σ˙Gcr𝑑t=αLGcr(σ).\displaystyle\left\lVert c_{1}-c_{2}\right\rVert_{H^{r}}\leq L_{H^{r}}(\sigma)=\int_{0}^{1}\left\lVert\dot{\sigma}\right\rVert_{H^{r}}\ dt\leq\alpha(r,c_{0},4\rho)\int_{0}^{1}\left\lVert\dot{\sigma}\right\rVert_{G_{c}^{r}}\ dt=\alpha L_{G_{c}^{r}}(\sigma).

Taking an infimum over all such σ\sigma yields (4.20). ∎

We are now ready to present the proof of the main theorem.

Proof of Theorem 4.1.

(1) Let {cn}r\{c_{n}\}\subset\mathcal{I}^{r} be a GcrG_{c}^{r}-Cauchy sequence. Then there exists ρ>0\rho>0 such that {cn}\{c_{n}\} is contained in some BGcr(c0,ρ)B_{G_{c}^{r}}(c_{0},\rho). Hence, by Lemma 4.8 there exists α>0\alpha>0 such that

(4.21) cNcMHrαdistGcr(cN,cM),\left\lVert c_{N}-c_{M}\right\rVert_{H^{r}}\leq\alpha\operatorname{dist}_{G_{c}^{r}}(c_{N},c_{M}),

for all N,MN,M\in\mathbb{N}. So {cn}\{c_{n}\} is Cauchy in (Hr,Hr)(H^{r},\left\lVert\cdot\right\rVert_{H^{r}}) and converges to some cHrc_{\infty}\in H^{r}. From Lemma 4.3 we have that {|(cn)θ|1}\{\left\lvert(c_{n})_{\theta}\right\rvert^{-1}\} is bounded away from 0. As r>32r>\frac{3}{2}, HrH^{r} convergence implies C1C^{1} convergence and hence crc_{\infty}\in\mathcal{I}^{r}. Finally, as GcrG_{c}^{r} is a strong metric, distGcr\operatorname{dist}_{G_{c}^{r}} induces the same topology as the manifold topology [29], which is in our case the Hilbert space topology of HrH^{r}. Thus, cncHr0\left\lVert c_{n}-c_{\infty}\right\rVert_{H^{r}}\rightarrow 0 implies that distGcr(cn,c)0\operatorname{dist}_{G_{c}^{r}}(c_{n},c_{\infty})\rightarrow 0.

(2) Proving geodesic convexity follows exactly as in the integer-order results, using the direct methods in the calculus of variations and utilizing the estimates of Lemma 4.7. See [12, Section 5] or [11, Section 5.5].

(3) Next, for geodesic completeness, note that although the Hopf-Rinow theorem does not hold in infinite dimensions [26, 34, 2] one still has that metric completeness implies geodesic completeness for strong metrics [29]. Hence we immediately have that, for q>32q>\frac{3}{2}, (q,Gq)(\mathcal{I}^{q},G^{q}) is geodesically complete.

To extend geodesic completeness to the case r>qr>q we will apply a no-loss-no-gain argument in spatial regularity, as originally developed by Ebin and Marsden to prove local well-posedness of the incompressible Euler equation [21]. In the context of the situation of the present article, the same argument has already been used in [5] to prove the local well-posedness for the GqG^{q}-metric on the whole scale of Sobolev immersions r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}) with r>32r>\frac{3}{2} and rq0r-q\geq 0. The exact same argument yields the desired global existence for all initial conditions if q>32q>\frac{3}{2}. Since the metric GqG^{q} is invariant by reparametrization, the same also holds for the corresponding geodesic spray. This allows one to apply the no loss-no gain result as formulated in [13]. Using this we obtain that solving the geodesic equation for initial conditions in r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}) with rqr\geq q the corresponding solution (geodesic) in r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}) exists for the same maximal time interval as in q(𝕊1,d)\mathcal{I}^{q}(\mathbb{S}^{1},\mathbb{R}^{d}). As (q(𝕊1,d),Gq)\left(\mathcal{I}^{q}(\mathbb{S}^{1},\mathbb{R}^{d}),G^{q}\right) is geodesically complete by the results of the previous section, this implies that all geodesics exist for all time in q(𝕊1,d)\mathcal{I}^{q}(\mathbb{S}^{1},\mathbb{R}^{d}) and thus also in r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}). This concludes the proof of geodesic completeness for r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}) with r>qr>q and consequently also for r=r=\infty, i.e., for the Fréchet manifold Imm(𝕊1,d)\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}).

(4) Let η0Cc()\eta_{0}\in C_{c}^{\infty}(\mathbb{R}) be a standard mollifier, and let η(t,x)=1tη0(x/t)\eta(t,x)=\frac{1}{t}\eta_{0}(x/t). Fix r(3/2,r)r^{\prime}\in(3/2,r) with rqr^{\prime}\geq q. Let c0r(𝕊1,d)r(𝕊1,d)c_{0}\in\mathcal{I}^{r^{\prime}}(\mathbb{S}^{1},\mathbb{R}^{d})\setminus\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}), and define c:[0,ϵ]r(𝕊1,d)c:[0,\epsilon]\to\mathcal{I}^{r^{\prime}}(\mathbb{S}^{1},\mathbb{R}^{d}) by

c(t)={c0t=0c0η(t,)t>0,c(t)=\begin{cases}c_{0}&t=0\\ c_{0}*\eta(t,\cdot)&t>0,\end{cases}

where * denotes the convolution operator.

By standard theory of mollifiers, c(t)c(0)c(t)\to c(0) as t0t\to 0 in Hr(𝕊1,d)H^{r^{\prime}}(\mathbb{S}^{1},\mathbb{R}^{d}); since r>3/2r^{\prime}>3/2, it is true also in C1(𝕊1,d)C^{1}(\mathbb{S}^{1},\mathbb{R}^{d}). Therefore, for ϵ>0\epsilon>0 small enough, it follows that indeed c(t)c(t) is an immersion for all t[0,ϵ]t\in[0,\epsilon]. In particular, c(t)Imm(𝕊1,d)c(t)\in\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{d}) for any t(0,ϵ]t\in(0,\epsilon]. Now, since cc is a smooth curve on the complete metric space (r(𝕊1,d),distGr)(\mathcal{I}^{r^{\prime}}(\mathbb{S}^{1},\mathbb{\mathbb{R}}^{d}),\operatorname{dist}_{G^{r^{\prime}}}), it has a finite GrG^{r^{\prime}}-length, and thus also a finite GqG^{q}-length (e.g., using (2.22) and the fact that length is uniformly bounded along the curve). Now, consider the path c|t(0,1]c|_{t\in(0,1]} in (r(𝕊1,d),distGq)(\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{\mathbb{R}}^{d}),\operatorname{dist}_{G^{q}}). By what we proved, it is a finite length path that leaves the space (at t=0t=0), since c0r(𝕊1,d)c_{0}\notin\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{\mathbb{R}}^{d}). Thus the space is incomplete, as long it was a metric space to begin with (i.e., if q>1/2q>1/2). This completes the metric incompleteness proof. Now, if q>3/2q>3/2, we can repeat the same argument for r=qr^{\prime}=q; this shows that r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{\mathbb{R}}^{d}) is dense in q(𝕊1,d)\mathcal{I}^{q}(\mathbb{S}^{1},\mathbb{\mathbb{R}}^{d}) with respect to distGq\operatorname{dist}_{G^{q}}. Since (q(𝕊1,d),distGq)(\mathcal{I}^{q}(\mathbb{S}^{1},\mathbb{\mathbb{R}}^{d}),\operatorname{dist}_{G^{q}}) is complete, we obtain it is the completion of r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{\mathbb{R}}^{d}).

(5) It remains to prove the statement on geodesic incompleteness for q<32q<\frac{3}{2} and r>32r>\frac{3}{2}. Therefore we will follow a similar argument as in [7], where geodesic incompleteness for integer order metrics on the space of immersions has been studied. Namely, we consider the space 𝒞\mathcal{C} of all concentric circles as a subset of r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}), i.e.,

(4.22) 𝒞:={(rcos(θ),rsin(θ)):r>0}r(𝕊1,d).\mathcal{C}:=\left\{(r\operatorname{cos}(\theta),r\operatorname{sin}(\theta)):r\in\mathbb{R}_{>0}\right\}\subset\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}).

A straight forward calculation shows that the space 𝒞\mathcal{C} equipped with the restriction of the GqG^{q}-metric is in fact a totally geodesic subset of r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}). Consequently, if we can show that 𝒞\mathcal{C} is geodesically incomplete for q<32q<\frac{3}{2} this also implies that r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}) is geodesically incomplete. Furthermore, since 𝒞\mathcal{C} is finite dimensional, by the theorem of Hopf-Rinow this can be reduced to proving metric incompleteness. This allows us to conclude the proof by showing that one can scale down a circle to zero with finite GqG^{q}-length. To this end, let c:[0,1)Imm(𝕊1,2)c:[0,1)\to\operatorname{Imm}(\mathbb{S}^{1},\mathbb{R}^{2}), c(t)=(1t)(cos(θ),sin(θ))c(t)=(1-t)(\operatorname{cos}(\theta),\operatorname{sin}(\theta)). If q<32q<\frac{3}{2}, a straightforward calculation yields the following inequality

(4.23) 01Gcq(tc,tc)1/2𝑑t132q,\displaystyle\int_{0}^{1}G_{c}^{q}(\partial_{t}c,\partial_{t}c)^{1/2}dt\lesssim\frac{1}{\tfrac{3}{2}-q},

from which it follows that length of cc is finite. Hence we have constructed a path of finite length that leaves the space 𝒞\mathcal{C}. This yields the desired metric and geodesic incompleteness result of 𝒞\mathcal{C} and consequently geodesic incompleteness of r(𝕊1,d)\mathcal{I}^{r}(\mathbb{S}^{1},\mathbb{R}^{d}). ∎

Appendix A Products and compositions in fractional Sobolev spaces

Here we provide details for the proof of the estimates given in Lemma 2.2. As mentioned in Section 2, our approach closely follows that of [23].

Proof of Lemma 2.2.
  1. (i)

    The inequality (2.3) is immediate from the definition (2.2).

  2. (ii)

    The proof of the inequality (2.4) can be found in [23, Lemma B.1].

  3. (iii)

    The proof of the (2.5) is essentially identical to the proof of (2.4). However, we record it here for completeness. We will deal explicitly with the case d=1d=1; the extension to d>1d>1 is straightforward.

    For 0<ab0<a\leq b, note that from (2.2) we have

    fgH˙a2\displaystyle\left\lVert fg\right\rVert_{\dot{H}^{a}}^{2} =n{0}|n|2a|fg^(n)|2,\displaystyle=\sum_{n\in\mathbb{Z}\setminus\{0\}}\left\lvert n\right\rvert^{2a}\lvert\widehat{fg}(n)\rvert^{2},

    where

    fg^(n)=f^g^(n)=j+k=nf^(j)g^(k).\displaystyle\widehat{fg}(n)=\hat{f}*\hat{g}(n)=\sum_{j+k=n}\hat{f}(j)\hat{g}(k).

    This gives the inequality

    |n|a|fg^(n)|\displaystyle\left\lvert n\right\rvert^{a}\lvert\widehat{fg}(n)\rvert j+k=n|j+k|a|f^(j)||g^(k)|\displaystyle\leq\sum_{j+k=n}\left\lvert j+k\right\rvert^{a}\lvert\hat{f}(j)\rvert\left\lvert\hat{g}(k)\right\rvert
    j+k=n|j||k||j+k|a|f^(j)||g^(k)|+j+k=n|j|>|k||j+k|a|f^(j)||g^(k)|,\displaystyle\leq\sum_{\begin{subarray}{c}j+k=n\\ \left\lvert j\right\rvert\leq\left\lvert k\right\rvert\end{subarray}}\left\lvert j+k\right\rvert^{a}\lvert\hat{f}(j)\rvert\left\lvert\hat{g}(k)\right\rvert+\sum_{\begin{subarray}{c}j+k=n\\ \left\lvert j\right\rvert>\left\lvert k\right\rvert\end{subarray}}\left\lvert j+k\right\rvert^{a}\lvert\hat{f}(j)\rvert\left\lvert\hat{g}(k)\right\rvert,

    which, up to a multiplication by 2a2^{a}, gives us

    |n|a|fg^(n)|\displaystyle\left\lvert n\right\rvert^{a}\lvert\widehat{fg}(n)\rvert j+k=n|j||k||k|a|f^(j)||g^(k)|+j+k=n|k|<|j||j|a|f^(j)||g^(k)|.\displaystyle\lesssim\sum_{\begin{subarray}{c}j+k=n\\ \left\lvert j\right\rvert\leq\left\lvert k\right\rvert\end{subarray}}\left\lvert k\right\rvert^{a}\lvert\hat{f}(j)\rvert\left\lvert\hat{g}(k)\right\rvert+\sum_{\begin{subarray}{c}j+k=n\\ \left\lvert k\right\rvert<\left\lvert j\right\rvert\end{subarray}}\left\lvert j\right\rvert^{a}\lvert\hat{f}(j)\rvert\left\lvert\hat{g}(k)\right\rvert.

    Separating the terms with j,k=0j,k=0 we acquire

    (A.1) |n|a|fg^(n)||f^(0)||n|a|g^(n)|+|g^(0)||n|a|f^(n)|+j+k=n0<|j||k||k|a|f^(j)||g^(k)|+j+k=n0<|k|<|j||j|a|f^(j)||g^(k)|.\begin{split}\left\lvert n\right\rvert^{a}\lvert\widehat{fg}(n)\rvert&\leq\lvert\hat{f}(0)\rvert\left\lvert n\right\rvert^{a}\left\lvert\hat{g}(n)\right\rvert+\left\lvert\hat{g}(0)\right\rvert\left\lvert n\right\rvert^{a}\lvert\hat{f}(n)\rvert\\ &\qquad+\sum_{\begin{subarray}{c}j+k=n\\ 0<\left\lvert j\right\rvert\leq\left\lvert k\right\rvert\end{subarray}}\left\lvert k\right\rvert^{a}\lvert\hat{f}(j)\rvert\left\lvert\hat{g}(k)\right\rvert+\sum_{\begin{subarray}{c}j+k=n\\ 0<\left\lvert k\right\rvert<\left\lvert j\right\rvert\end{subarray}}\left\lvert j\right\rvert^{a}\lvert\hat{f}(j)\rvert\left\lvert\hat{g}(k)\right\rvert.\end{split}

    Focusing on the third term above, we have

    j+k=n0<|j||k||k|a|f^(j)||g^(k)|\displaystyle\sum_{\begin{subarray}{c}j+k=n\\ 0<\left\lvert j\right\rvert\leq\left\lvert k\right\rvert\end{subarray}}\left\lvert k\right\rvert^{a}\lvert\hat{f}(j)\rvert\left\lvert\hat{g}(k)\right\rvert j+k=n0<|j||k||kj|ba|k|a|f^(j)||g^(k)|\displaystyle\leq\sum_{\begin{subarray}{c}j+k=n\\ 0<\left\lvert j\right\rvert\leq\left\lvert k\right\rvert\end{subarray}}\left\lvert\frac{k}{j}\right\rvert^{b-a}\left\lvert k\right\rvert^{a}\lvert\hat{f}(j)\rvert\left\lvert\hat{g}(k)\right\rvert
    j+k=n0<|j||k|1|j|b|j|a|f^(j)||k|b|g^(k)|\displaystyle\leq\sum_{\begin{subarray}{c}j+k=n\\ 0<\left\lvert j\right\rvert\leq\left\lvert k\right\rvert\end{subarray}}\frac{1}{\left\lvert j\right\rvert^{b}}\left\lvert j\right\rvert^{a}\lvert\hat{f}(j)\rvert\left\lvert k\right\rvert^{b}\left\lvert\hat{g}(k)\right\rvert
    j+k=nj01|j|b|j|a|f^(j)||k|b|g^(k)|\displaystyle\leq\sum_{\begin{subarray}{c}j+k=n\\ j\neq 0\end{subarray}}\frac{1}{\left\lvert j\right\rvert^{b}}\left\lvert j\right\rvert^{a}\lvert\hat{f}(j)\rvert\left\lvert k\right\rvert^{b}\left\lvert\hat{g}(k)\right\rvert
    =(λbf~ag~b)(n),\displaystyle=(\lambda_{b}\tilde{f}_{a}*\tilde{g}_{b})(n),

    where, for n{0}n\in\mathbb{Z}\setminus\{0\}, we define λb(n)=1|n|b\lambda_{b}(n)=\frac{1}{\left\lvert n\right\rvert^{b}}, f~a(n)=|n|a|f^(n)|\tilde{f}_{a}(n)=\left\lvert n\right\rvert^{a}\lvert\hat{f}(n)\rvert and g~b(n)=|n|b|g^(n)|\tilde{g}_{b}(n)=\left\lvert n\right\rvert^{b}\left\lvert\hat{g}(n)\right\rvert along with λb(0)=f~a(0)=g~b(0)=0\lambda_{b}(0)=\tilde{f}_{a}(0)=\tilde{g}_{b}(0)=0. We note here that, for a>0a>0, we have the equality fH˙a=f~a2\left\lVert f\right\rVert_{\dot{H}^{a}}=\lVert\tilde{f}_{a}\rVert_{\ell^{2}}.

    Next, for the fourth term of (A.1), we have

    j+k=n0<|k|<|j||j|a|f^(j)||g^(k)|\displaystyle\sum_{\begin{subarray}{c}j+k=n\\ 0<\left\lvert k\right\rvert<\left\lvert j\right\rvert\end{subarray}}\left\lvert j\right\rvert^{a}\lvert\hat{f}(j)\rvert\left\lvert\hat{g}(k)\right\rvert =j+k=n0<|k|<|j||kk|b|j|a|f^(j)||g^(k)|\displaystyle=\sum_{\begin{subarray}{c}j+k=n\\ 0<\left\lvert k\right\rvert<\left\lvert j\right\rvert\end{subarray}}\left\lvert\frac{k}{k}\right\rvert^{b}\left\lvert j\right\rvert^{a}\lvert\hat{f}(j)\rvert\left\lvert\hat{g}(k)\right\rvert
    =j+k=n0<|k|<|j||j|a|f^(j)|1|k|b|k|b|g^(k)|\displaystyle=\sum_{\begin{subarray}{c}j+k=n\\ 0<\left\lvert k\right\rvert<\left\lvert j\right\rvert\end{subarray}}\left\lvert j\right\rvert^{a}\lvert\hat{f}(j)\rvert\frac{1}{\left\lvert k\right\rvert^{b}}\left\lvert k\right\rvert^{b}\left\lvert\hat{g}(k)\right\rvert
    j+k=nk0|j|a|f^(j)|1|k|b|k|b|g^(k)|\displaystyle\leq\sum_{\begin{subarray}{c}j+k=n\\ k\neq 0\end{subarray}}\left\lvert j\right\rvert^{a}\lvert\hat{f}(j)\rvert\frac{1}{\left\lvert k\right\rvert^{b}}\left\lvert k\right\rvert^{b}\left\lvert\hat{g}(k)\right\rvert
    =(f~aλbg~b)(n).\displaystyle=\left(\tilde{f}_{a}*\lambda_{b}\tilde{g}_{b}\right)(n).

    Combining all this, (A.1) becomes

    (A.2) |n|a|fg^(n)||f^(0)|g~a(n)+|g^(0)|f~a(n)+(λbf~ag~b)(n)+(f~aλbg~b)(n),\left\lvert n\right\rvert^{a}\lvert\widehat{fg}(n)\rvert\leq\lvert\hat{f}(0)\rvert\tilde{g}_{a}(n)+\left\lvert\hat{g}(0)\right\rvert\tilde{f}_{a}(n)+\left(\lambda_{b}\tilde{f}_{a}*\tilde{g}_{b}\right)(n)+\left(\tilde{f}_{a}*\lambda_{b}\tilde{g}_{b}\right)(n),

    which gives

    (A.3) fgH˙a|f^(0)|g~a2+|g^(0)|f~a2+λbf~ag~b2+f~aλbg~b2.\left\lVert fg\right\rVert_{\dot{H}^{a}}\leq\lvert\hat{f}(0)\rvert\left\lVert\tilde{g}_{a}\right\rVert_{\ell^{2}}+\left\lvert\hat{g}(0)\right\rvert\lVert\tilde{f}_{a}\rVert_{\ell^{2}}+\lVert\lambda_{b}\tilde{f}_{a}*\tilde{g}_{b}\rVert_{\ell^{2}}+\lVert\tilde{f}_{a}*\lambda_{b}\tilde{g}_{b}\rVert_{\ell^{2}}.

    By Young’s inequality and the Cauchy-Schwartz inequality for the 2\ell^{2}-inner product we have

    λbf~ag~b2\displaystyle\lVert\lambda_{b}\tilde{f}_{a}*\tilde{g}_{b}\rVert_{\ell^{2}} λbf~a1g~b2λb2f~a2g~b2fH˙agH˙b\displaystyle\lesssim\lVert\lambda_{b}\tilde{f}_{a}\rVert_{\ell^{1}}\lVert\tilde{g}_{b}\rVert_{\ell^{2}}\lesssim\lVert\lambda_{b}\rVert_{\ell^{2}}\lVert\tilde{f}_{a}\rVert_{\ell^{2}}\lVert\tilde{g}_{b}\rVert_{\ell^{2}}\lesssim\lVert f\rVert_{\dot{H}^{a}}\lVert g\rVert_{\dot{H}^{b}}

    and

    f~aλbg~b2\displaystyle\lVert\tilde{f}_{a}*\lambda_{b}\tilde{g}_{b}\rVert_{\ell^{2}} f~a2λbg~b1f~a2λb2g~b2fH˙agH˙b.\displaystyle\lesssim\lVert\tilde{f}_{a}\rVert_{\ell^{2}}\lVert\lambda_{b}\tilde{g}_{b}\rVert_{\ell^{1}}\lesssim\lVert\tilde{f}_{a}\rVert_{\ell^{2}}\lVert\lambda_{b}\rVert_{\ell^{2}}\lVert\tilde{g}_{b}\rVert_{\ell^{2}}\lesssim\lVert f\rVert_{\dot{H}^{a}}\lVert g\rVert_{\dot{H}^{b}}.

    Hence, from (A.3) we have

    (A.4) fgH˙a(a,b)|f^(0)|gH˙a+|g^(0)|fH˙a+fH˙agH˙b.\displaystyle\left\lVert fg\right\rVert_{\dot{H}^{a}}\lesssim_{(a,b)}|\hat{f}(0)|\left\lVert g\right\rVert_{\dot{H}^{a}}+\left\lvert\hat{g}(0)\right\rvert\left\lVert f\right\rVert_{\dot{H}^{a}}+\left\lVert f\right\rVert_{\dot{H}^{a}}\left\lVert g\right\rVert_{\dot{H}^{b}}.

  4. (iv)

    Recalling the Gagliardo semi-norm, cf. [20], we have

    (A.5) fgH˙a𝕊1𝕊1|f(θ)g(θ)f(α)g(α)||θα|1+2a𝑑θ𝑑α.\left\lVert fg\right\rVert_{\dot{H}^{a}}\simeq\int_{\mathbb{S}^{1}}\int_{\mathbb{S}^{1}}\frac{\left\lvert f(\theta)g(\theta)-f(\alpha)g(\alpha)\right\rvert}{\left\lvert\theta-\alpha\right\rvert^{1+2a}}\ d\theta d\alpha.

    Applying the triangle inequality, we have

    |f(θ)g(θ)f(α)g(α)|\displaystyle\left\lvert f(\theta)g(\theta)-f(\alpha)g(\alpha)\right\rvert =|f(θ)g(θ)f(θ)g(α)+f(θ)g(α)f(α)g(α)|\displaystyle=\left\lvert f(\theta)g(\theta)-f(\theta)g(\alpha)+f(\theta)g(\alpha)-f(\alpha)g(\alpha)\right\rvert
    |f(θ)g(θ)f(θ)g(α)|+|f(θ)g(α)f(α)g(α)|\displaystyle\leq\left\lvert f(\theta)g(\theta)-f(\theta)g(\alpha)\right\rvert+\left\lvert f(\theta)g(\alpha)-f(\alpha)g(\alpha)\right\rvert
    fL|g(θ)g(α)|+gL|f(θ)f(α)|.\displaystyle\leq\left\lVert f\right\rVert_{L^{\infty}}\left\lvert g(\theta)-g(\alpha)\right\rvert+\left\lVert g\right\rVert_{L^{\infty}}\left\lvert f(\theta)-f(\alpha)\right\rvert.

    Substituting this into (A.5) immediately yields (2.6).

  5. (v)

    Applying a change of variables immediately gives

    fϕL2\displaystyle\left\lVert f\circ\phi\right\rVert_{L^{2}} (ϕ1)θ12fL2\displaystyle\leq\left\lVert(\phi^{-1})_{\theta}\right\rVert^{\frac{1}{2}}\left\lVert f\right\rVert_{L^{2}}

    and

    fϕH˙1\displaystyle\left\lVert f\circ\phi\right\rVert_{\dot{H}^{1}} ϕθ12fH˙1.\displaystyle\leq\left\lVert\phi_{\theta}\right\rVert^{\frac{1}{2}}\left\lVert f\right\rVert_{\dot{H}^{1}}.

    The inequality (2.7) then follows by interpolation, cf. [25, Corollary 8.3] or [43].

References

  • [1] V. Arnold and B. Khesin. Topological Methods in Hydrodynamics. Springer, Cham, Second edition, 2021.
  • [2] C. J. Atkin. The Hopf-Rinow theorem is false in infinite dimensions. Bull. London Math. Soc., 7(3):261–266, 1975.
  • [3] M. Bauer, M. Bruveris, P. Harms, and P. Michor. Vanishing geodesic distance for the Riemannian metric with geodesic equation the KdV-equation. Annals of Global Analysis and Geometry, 41:461–472, 2012.
  • [4] M. Bauer, M. Bruveris, P. Harms, and P. Michor. Geodesic distance for right invariant Sobolev metrics of fractional order on the diffeomorphism group. Annals of Global Analysis and Geometry, 44(1):5–21, 2013.
  • [5] M. Bauer, M. Bruveris, and B. Kolev. Fractional Sobolev metrics on spaces of immersed curves. Calc. Var. Partial Differential Equations, 57(1):Paper No. 27, 24, 2018.
  • [6] M. Bauer, M. Bruveris, and P. Michor. Overview of the geometries of shape spaces and diffeomorphism groups. Journal of Mathematical Imaging and Vision, 50:60–97, 2014.
  • [7] M. Bauer, P. Harms, and P. Michor. Sobolev metrics on shape space of surfaces. Journal of Geometric Mechanics, 3(4):389–438, 2012.
  • [8] M. Bauer, P. Harms, and P. Michor. Regularity and completeness of half-lie groups. Journal of the European Mathematical Society, forthcoming.
  • [9] M. Bauer, B. Kolev, and S. Preston. Geometric investigations of a vorticity model equation. Journal of Differential Equations, 260(1):478–516, 2016.
  • [10] M. Bauer, B. Kolev, and S. Preston. Geodesic completeness of the H3/2H^{3/2} metric on Diff(S1)\operatorname{Diff}(S^{1}). Monatshefte für Mathematik, 193(2):233–245, 2020.
  • [11] M. Bauer, C. Maor, and P. Michor. Sobolev metrics on spaces of manifold valued curves. Annali della Scuola Normale Superiore di Pisa, Vol. XXIV:1895–1948, 2023.
  • [12] M. Bruveris. Completeness properties of Sobolev metrics on the space of curves. J. Geom. Mech., 7(2):125–150, 2015.
  • [13] M. Bruveris. Regularity of maps between Sobolev spaces. Annals of Global Analysis and Geometry, 52:11–24, 2017.
  • [14] M. Bruveris, P. Michor, and D. Mumford. Geodesic completeness for Sobolev metrics on the space of immersed plane curves. In Forum of Mathematics, Sigma, volume 2, page e19. Cambridge University Press, 2014.
  • [15] M. Bruveris and J. Møller-Andersen. Completeness of length-weighted Sobolev metrics on the space of curves. arXiv preprint arXiv:1705.07976, pages 1–15, 2017.
  • [16] M. Bruveris and F. Vialard. On completeness of groups of diffeomorphisms. Journal of the European Mathematical Society, 19(5):1507–1544, 2017.
  • [17] R. Camassa and D. Holm. An integrable shallow water equation with peaked solitons. Physical review letters, 71(11):1661, 1993.
  • [18] A. Constantin and J. Escher. Wave breaking for nonlinear nonlocal shallow water equations. Acta Mathematica, 181(2):229–243, 1998.
  • [19] A. Constantin and B. Kolev. On the geometric approach to the motion of inertial mechanical systems. Journal of Physics A: Mathematical and General, 35(32):R51, 2002.
  • [20] E. Di Nezza, G. Palatucci, and E. Valdinoci. Hitchhiker’s guide to the fractional Sobolev spaces. Bull. Sci. Math., 136(5):521–573, 2012.
  • [21] D. Ebin and J. Marsden. Groups of diffeomorphisms and the motion of an incompressible fluid. Ann. of Math. (2), 92:102–163, 1970.
  • [22] Y. Eliashberg and L. Polterovich. Bi-invariant metrics on the group of Hamiltonian diffeomorphisms. Internat. J. Math, 4(5):727–738, 1993.
  • [23] J. Escher and B. Kolev. Right-invariant Sobolev metrics of fractional order on the diffeomorphism group of the circle. J. Geom. Mech., 6(3):335–372, 2014.
  • [24] J. Escher, B. Kolev, and M. Wunsch. The geometry of a vorticity model equation. Communications on Pure and Applied Analysis, 11(4):1407–1419, 2011.
  • [25] M. Frazier and B. Jawerth. A discrete transform and decompositions of distribution spaces. J. Funct. Anal., 93(1):34–170, 1990.
  • [26] N. Grossman. Hilbert manifolds without epiconjugate points. Proc. Amer. Math. Soc., 16:1365–1371, 1965.
  • [27] R. Jerrard and C. Maor. Geodesic distance for right-invariant metrics on diffeomorphism groups: critical Sobolev exponents. Annals of Global Analysis and Geometry, 56:351–360, 2019.
  • [28] J. Knappmann, H. Schumacher, D. Steenebrügge, and H. von der Mosel. A speed preserving Hilbert gradient flow for generalized integral Menger curvature. Advances in Calculus of Variations, 16(3):597–635, 2023.
  • [29] S. Lang. Fundamentals of differential geometry, volume 191 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1999.
  • [30] G. Leoni. A first course in Sobolev spaces. American Mathematical Soc., 2017.
  • [31] G. Leoni. A first course in fractional Sobolev spaces, volume 229. American Mathematical Society, 2023.
  • [32] V. Magnani and D. Tiberio. A remark on vanishing geodesic distances in infinite dimensions. Proceedings of the American Mathematical Society, 148(8):3653–3656, 2020.
  • [33] T. Marquis and K.-H. Neeb. Half-lie groups. Transformation Groups, 23(3):801–840, 2018.
  • [34] J. McAlpin. Infinite dimensional manifolds and Morse theory. ProQuest LLC, Ann Arbor, MI, 1965. Thesis (Ph.D.)–Columbia University.
  • [35] P. Michor and D. Mumford. Vanishing geodesic distance on spaces of submanifolds and diffeomorphisms. Documenta Mathematica, 10:217–245, 2005.
  • [36] P. Michor and D. Mumford. Riemannian geometries on spaces of plane curves. Journal of the European Mathematical Society, 8(1):1–48, 2006.
  • [37] S. Okabe and P. Schrader. Convergence of Sobolev gradient trajectories to elastica. arXiv preprint arXiv:2107.06504, pages 1–29, 2021.
  • [38] S. Preston and P. Washabaugh. Euler–Arnold equations and Teichmüller theory. Differential Geometry and its Applications, 59:1–11, 2018.
  • [39] P. Reiter and H. Schumacher. Sobolev gradients for the Möbius energy. Archive for Rational Mechanics and Analysis, 242(2):701–746, 2021.
  • [40] P. Schrader, G. Wheeler, and V. Wheeler. On the H1(dsγ)H^{1}(ds^{\gamma})-gradient flow for the length functional. The Journal of Geometric Analysis, 33(9):297, 2023.
  • [41] A. Srivastava and E. Klassen. Functional and shape data analysis, volume 1. Springer, 2016.
  • [42] A. Srivastava, E. Klassen, S. Joshi, and I. Jermyn. Shape analysis of elastic curves in euclidean spaces. IEEE Transactions on Pattern Analysis and Machine Intelligence, 33(7):1415–1428, 2010.
  • [43] H. Triebel. Theory of function spaces. Birkhäuser, 1983.
  • [44] P. Washabaugh. The SQG equation as a geodesic equation. Archive for Rational Mechanics and Analysis, 222(3):1269–1284, 2016.
  • [45] M. Wunsch. On the geodesic flow on the group of diffeomorphisms of the circle with a fractional Sobolev right-invariant metric. Journal of Nonlinear Mathematical Physics, 17(1):7–11, 2010.
  • [46] L. Younes. Shapes and diffeomorphisms, volume 171. Springer, 2010.