1: \errorcontextlines10
2: %\long\def\meta#1{\texttt{#1}}
3: \documentclass[reqno]{amsart}
4: %\usepackage[notref]{showkeys}
5: %
6: \usepackage{geompsfi}
7: %\input supp-pdf.tex
8: %\DeclareGraphicsExtensions{.ps,.0}
9: %\DeclareGraphicsRule{*}{mps}{*}{}
10: \usepackage{latexsym}
11: \usepackage{amstext}
12: \usepackage {amsmath}
13: \usepackage {amsfonts}
14: \usepackage {amssymb}
15: \usepackage {amsthm}
16: \usepackage {bbm}
17: \usepackage{enumerate}
18: \usepackage{array}
19: %\usepackage{pictex}
20: %\usepackage[metapost]{mfpic}
21: \usepackage{comment}
22: \usepackage{xspace}
23: \sloppy
24: %\usepackage{showkeys}\renewcommand{\baselinestretch}{1.5}
25: %==================
26: % new commands
27: %==================
28: \def\cringle{\mathaccent"7017 }
29: \newcommand{\GammaD}{\ensuremath{\Gamma_{\!\text{\tiny D}}}}
30: \newcommand{\uD}{\ensuremath{u_{\text{\tiny D}}}}
31: \DeclareMathOperator{\dist}{dist}
32: \DeclareMathOperator{\sgn}{sgn}
33: \DeclareMathOperator{\spt}{supp}
34: \DeclareMathOperator\supp{supp}
35: \DeclareMathOperator{\aplim}{ap-lim}
36: \DeclareMathOperator{\dive}{div}
37: %\DeclareMathOperator{\loc}{loc}
38: \def\loc{{\mathrm{loc}}}
39: \DeclareMathOperator{\Lp}{L}
40: \DeclareMathOperator{\BV}{BV}
41: \DeclareMathOperator{\Lip}{Lip}
42: \DeclareMathOperator{\HS}{H}
43: \DeclareMathOperator{\p}{p}
44: \DeclareMathOperator{\diam}{diam}
45: \DeclareMathOperator{\modulo}{\, mod}
46: \DeclareMathOperator{\Int}{Int}
47: \DeclareMathOperator{\length}{length}
48: \newcommand{\HO}{\ensuremath{\cringle{\HS}\mbox{}}}
49: \newcommand{\eps}{\varepsilon}
50: \newcommand{\Wasserstein}{Monge-Kantorovich\xspace}
51: \newcommand{\multiplicity}{multiplicity\xspace}
52: \newcommand{\Chi}{\mathcal{X}}
53: \newcommand{\R}{\ensuremath{\mathbb{R}}}
54: \newcommand{\Rn}{\ensuremath{\mathbb{R}^2}}
55: \newcommand{\N}{\ensuremath{\mathbb{N}}\ }
56: \newcommand{\LL}{\ensuremath{\mathcal{L}}}
57: \newcommand{\K}{\ensuremath{\mathcal K}}
58: \newcommand{\sphere}{\ensuremath{\mathcal S}}
59: \newcommand{\F}{\ensuremath{\mathcal F}}
60: \newcommand{\G}{\ensuremath{\mathcal G}}
61: \newcommand{\Fone}{\ensuremath{\mathcal F_1}}
62: \newcommand{\A}{\ensuremath{\mathcal{A}}}
63: \newcommand{\Ha}{\ensuremath{\mathcal{H}}}
64: \newcommand{\ma}{\ensuremath{\mathfrak{m}}}
65: \newcommand{\te}{\ensuremath{\mathfrak{t}}}
66: \newcommand{\alphaprime}{\ensuremath{\alpha^\prime}}
67: \newcommand{\schwsternto}{\ensuremath{\overset{*}{\rightharpoonup}}}
68: \newcommand{\eing}{\ensuremath{\lfloor}}
69: \newcommand{\schwto}{\ensuremath{\to}}
70: \newcommand{\schwstto}{\ensuremath{\overset{*}{\rightharpoonup}}}
71: \let\ds\displaystyle
72: \def\Mod#1{\left\|#1\right\|}
73: \def\mod#1{\left|#1\right|}
74: \def\vc#1{\ensuremath{\vcenter{\hbox{#1}}}}
75: %
76: % environments
77: %
78: \newcounter{counter-liste}
79: \newcounter{counter-liste2}
80: \newenvironment{liste2}{\
81: \begin{list}{{\arabic{counter-liste2}.}\hfill}{\usecounter{counter-liste2}
82: \setlength{\topsep}{0bp}
83: \setlength{\labelwidth}{10bp}
84: \setlength{\leftmargin}{12bp}%0.04\linewidth}
85: \setlength{\labelsep}{2bp}
86: \setlength{\itemindent}{0bp}%-0.2\leftmargin}
87: %\setlength{\itemsep}{3bp}%{3bp}
88: \setlength{\parsep}{0bp}}
89: }{\end{list}}
90: \newenvironment{liste}%
91: {\ \begin{list}{{(\arabic{counter-liste})}\hfill}%
92: {\topsep2mm\itemindent1ex\leftmargin0cm\usecounter{counter-liste}}
93: }%
94: {\end{list}}
95: \newenvironment{liste(a)}%
96: {\ \begin{list}{{(\alph{counter-liste})}\hfill}%
97: {\topsep2mm\itemindent1ex\leftmargin0cm\usecounter{counter-liste}}
98: }%
99: {\end{list}}
100: %
101: \def\R{\mathbb R}
102: \def\H{\mathcal H}
103: \def\W{\mathcal W}
104: \let\e\varepsilon
105: \def\mod#1{\left|#1\right|}
106: \def\BV{\mathrm{BV}}
107: \let\ds\displaystyle
108: \def\pref#1{(\ref{#1})}
109: \theoremstyle{plain}
110: \numberwithin{equation}{section}
111: \newtheorem{lemma}{Lemma}[section]
112: \newtheorem{theorem}[lemma]{Theorem}
113: \newtheorem{proposition}[lemma]{Proposition}
114: \newtheorem{definition}[lemma]{Definition}
115: \newtheorem{assumption}[lemma]{Assumption}
116: \newtheorem{corollary}[lemma]{Corollary}
117: \newtheorem*{prob-mass-transport}{Optimal mass transport}
118: \newtheorem*{prob-Kantorovich}{Optimal Kantorovich potential}
119: \theoremstyle{definition}
120: \newtheorem{remark}[lemma]{Remark}
121: %
122: %\opengraphsfile{helf-figs}
123: %
124: % From Blom-Peletier
125: %
126: \def\nw{{N_w}}
127: \def\nl{{N_\ell}}
128: \def\X{{\mathcal{X}}}
129: \def\E{{\mathcal{E}}}
130: \def\Fid{F^{\textrm{id}}}
131: \def\Fnid{F^{\textrm{nid}}}
132: \def\Hid{H^{\textrm{id}}}
133:
134: %=====================================================================================================
135: % main
136: %=====================================================================================================
137: \begin{document}
138: \title{Partial Localization, Lipid Bilayers, and the Elastica Functional}
139: \author{Mark A. Peletier \and Matthias R\"oger}
140: \begin{abstract}
141: \emph{Partial localization} is the phenomenon of self-aggregation of
142: mass into high-density structures that are thin in one direction and
143: extended in the others.
144: We give a detailed study of an energy functional that arises in a
145: simplified model for lipid bilayer membranes. We demonstrate that this
146: functional, defined on a class of two-dimensional spatial mass
147: densities, exhibits partial localization and displays the `solid-like'
148: behavior of cell membranes.
149:
150: Specifically, we show that density fields of moderate energy are
151: partially localized, \emph{i.e.} resemble thin structures. Deviation
152: from a specific uniform thickness, creation of `ends', and the bending of such
153: structures all carry an energy penalty, of different orders in terms
154: of the thickness of the structure.
155:
156: These findings are made precise in a Gamma-convergence result. We prove
157: that a rescaled version of
158: the energy functional converges in the zero-thickness limit to a
159: functional that is defined on
160: a class of planar curves. Finiteness of the limit enforces both
161: optimal thickness and non-fracture; if these conditions are met, then
162: the limit value is given by the classical Elastica
163: (bending) energy of the curve.
164: \begin{comment}
165: \vspace*{2cm}
166: \emph{Partial localization} is the phenomenon of self-aggregation of
167: mass into high-density structures that are thin in one direction and
168: extended in the others.
169: We give a detailed study of a simplified energy functional that is
170: inspired by well-known block copolymer energies, and show that this
171: functional exhibits such partial localization. The energy functional is
172: definied on a class of two-dimensional spatial mass densities, and we
173: characterize the properties of this functional by identifying its
174: behaviour in a particular limit $\e\to0$.
175:
176: Specifically, we show that (a) density fields of moderate energy are
177: partially localized, \emph{i.e.} are thin (order $\e$) in one direction
178: and extended in the other; (b) these thin structures prefer a thickness
179: of \emph{exactly} $2\e$, and deviations from this optimal thickness are
180: penalized at order $\e^0$; (c) they resist fracture, in the sense that
181: an `end' carries an energy penalty that is larger than $O(\e^2)$; (d)
182: they resist bending, and that bending is penalized at order $\e^2$.
183:
184: These findings are formulated as a Gamma-convergence result, in which we
185: show that an $\e^{-2}$-rescaled version of the functional converges to a
186: limit functional that is defined on a class of planar curves. Finiteness
187: of the limit value enforces both optimal thickness and non-fracture; if
188: these conditions are met, then the value of this functional is given by
189: the classical Elastica (bending) energy of the curve.
190: \end{comment}
191: \end{abstract}
192: \keywords{Partial localization, concentration on curves, lipid bilayers,
193: Elastica functional, Willmore functional, Monge-Kantorovich distance,
194: Geometric Measure Theory, Gamma-convergence, block copolymers}
195: \maketitle
196:
197: \section{Introduction}
198:
199: In this paper we study the asymptotic expansion as $\eps\to 0$ of the
200: functional
201: \begin{equation}
202: \label{def:Fe}
203: \F_\eps(u,v) := \begin{cases}
204: \;\ds\e\! \int \mod{\nabla u} + \frac1\e d_1(u,v) \qquad\qquad &
205: \text{if }(u,v)\in \K_\eps, \\
206: \;\infty & \text{otherwise.}
207: \end{cases}
208: \end{equation}
209: Here $d_1(\cdot,\cdot)$ is the \Wasserstein distance
210: (Definition~\ref{def:wasserstein}), and
211: \begin{multline}
212: \label{def:Ke}
213: \K_\eps := \biggl\{ (u,v)\in
214: \BV\bigl(\R^2;\{0,\e^{-1}\}\bigr)\times L^1\bigl(\R^2;\{0,\eps^{-1}\}\bigr): \\
215: \left.\int u = \int v =M , \ uv = 0\text{ a.e.}\right\}.
216: \end{multline}
217: How this singular-perturbation problem is related to the terms in the title---partial
218: localization, lipid bilayers, and the Elastica functional---we explain in the rest
219: of this introductory section.
220: %Starting with Section~\ref{sec:heuristics}
221: %we then discuss and prove our main results, and we conclude with
222: %an interpretation of our mathematical findings in both mathematical and biological context.
223:
224:
225: \subsection{Lipid Bilayers}
226: Lipid bilayers, biological membranes,
227: are the living cell's main separating structure.
228: They shield the interior of the cell from the outside, and their
229: mechanical properties determine a large part of the interior
230: organization of living cells.
231: The main component is a lipid molecule (Fig.~\ref{fig:bl}) which consists
232: of a head and two tails. The head is usually charged, and therefore
233: hydrophilic, while the tails are hydrophobic. This difference in water
234: affinity causes lipids to aggregate, and in the biological setting
235: the lipids are typically found in a bilayer
236: structure as shown in Fig.~\ref{fig:bl}. In such a structure the
237: energetically unfavorable tail-water interactions are avoided by
238: grouping the tails together in a water-free zone, shielded from the water
239: by the heads. Note that, despite the structured appearance of the bilayer,
240: there is no covalent bonding between any two lipids; the bilayer structure is
241: entirely the result of the hydrophobic effect.
242: %While in the context of living cells lipids are exclusively found
243: %in structures that are planar, it is known that by adding or removing
244: %tails a preference can be created for structures with other symmetries.
245: %
246: %
247: %While for other amphiphilic molecules spherical or
248: %linear arrangements are also possible, most lipids have a preference
249: %for planar structures.
250:
251: \begin{figure}[t]
252: \centering
253: \includegraphics[width=\textwidth]{membr-1azw}
254: \caption{Lipid molecules aggregate into macroscopically surface-like structures}
255: \label{fig:bl}
256: \end{figure}
257: As planar structures, lipid bilayers have a remarkable combination of solid-like and
258: liquid-like properties. On one hand they resist various types
259: of deformation, such as extension, bending, and fracture, much in the way
260: a sheet of rubber or another elastic material does:
261: for instance, in order to stretch the bilayer, a tensile
262: force has to be applied, and when this force is removed the structure will again relax to
263: its original length.
264: With respect to in-plane rearrangements of the lipid molecules,
265: however, the material behaviour is viscous, and there is no penalty to large
266: in-plane deformations.
267:
268: While this interesting combination of properties is clearly related to the chemical
269: makeup of the lipids---most importantly, the hydrophobic character
270: of the tails---quantitative and detailed understanding of the phenomenon is
271: still lacking. Here we focus on a simple question that has already been alluded to above:
272: how can we understand the stability of these planar structures,
273: and their pseudo-solid behaviour, if they are constructed
274: from independent, non-bound molecules? This is the main question behind
275: the analysis of this paper.
276:
277: \subsection{Partial localization}
278:
279: The self-aggregation of lipids into bilayers is fundamentally different
280: from the self-aggregation of `simple' hydrophobic compounds in
281: water.
282: Lipid bilayers are \emph{thin structures}, in the sense that
283: there is a separation of length scales: the thickness of a lipid bilayer is fixed
284: to approximately two lipid lengths, while the in-plane spatial
285: extent is only limited by the surroundings. By contrast,
286: `simple' hydrophobic compounds aggregate in water to form
287: drops that lack the small intrinsic length scale of lipid bilayers.
288:
289: We use the term \emph{partial localization} for the self-aggregation
290: into structures that are thin in one or more directions and
291: `large' in others. The word \emph{localization} is taken from
292: the literature of reaction-diffusion equations, in which localized
293: solutions are those that are concentrated, in a well-defined way, in a
294: small neighbourhood of a point. By \emph{partial} localization
295: we refer to localization to the neighbourhood of a set that
296: has intermediate dimension, \emph{i.e.} dimension
297: larger than zero (a point) but
298: smaller than that of the ambient space.
299:
300: \subsection{Energy on a mesoscale: The functional \boldmath $\F_\eps$}
301: The functional $\F_\eps$ that is defined above is
302: the result of an attempt to capture enough of the essence
303: of lipid bilayers to address the issues of stability
304: and pseudo-solid behaviour while
305: keeping the description as simple as possible.
306: %The
307: %goal is to find an energy functional for which functions of
308: %near-minimal energy display partial localization;
309: %existing models for lipid bilayers and, more generally,
310: %block co-polymers
311: %provide a promising starting-point.
312: %
313: The derivation of $\F_\eps$, which is given in Appendix~\ref{app:derivation},
314: starts with a simple two-bead model of lipid molecules, inspired by
315: well-known models of block copolymers~\cite{Leibler80,BatesFredrickson90,FredricksonBates96}, and
316: proceeds to reduce complexity through a number of sometimes
317: radical simplifications. Despite these simplifications the physical
318: origin of the various elements of $\F_\eps$ remains identifiable:
319: \begin{itemize}
320: \item The functions $u$ and $v$ represent densities of the (hydrophobic) tail and (hydrophilic) headb eads;
321: \item The term $\int \mod{\nabla u}$, coupled with the restriction to
322: functions $u$ and $v$ that take only two values, and have disjoint
323: support, represents an interfacial energy that arises from the hydrophobic effect;
324: \item The \Wasserstein distance $d_1(u,v)$ between $u$ and
325: $v$ is a weak remnant of the covalent bonding between head and tail
326: particles.
327: \end{itemize}
328: The parameter $\eps$ appears in \eqref{def:Ke} as a density scaling.
329: We prove that $\eps$ in fact is related
330: to the \emph{thickness} of structures with moderate energy.
331:
332: We show in this paper that
333: the functional $\F_\eps$ favours partial localization, \emph{i.e.}
334: that pairs $(u,v)$ for which $\F_\eps(u,v)$ is not too large
335: are necessarily partially localized. We will also show that $\F_\eps$
336: displays three additional
337: properties that were already mentioned: resistance to stretching,
338: to fracture, and to bending. To describe the last of these in more
339: detail we briefly return to biophysics.
340:
341: \subsection{Energy on the macroscale: The Helfrich Hamiltonian and the Elastica functional}
342:
343: Intrigued by the shape of red blood cells
344: Canham, Helfrich, and Evans pioneered the modelling
345: of lipid bilayer vesicles by energy
346: methods~\cite{Canham70,Helfrich73,Evans74}.
347: The name of Helfrich is now associated with
348: a surface energy for closed vesicles,
349: %On the basis of symmetry arguments the biophysicist Helfrich
350: %postulated a surface energy density for closed
351: %vesicles,
352: represented by a smooth boundaryless surface $S$, of the form
353: \begin{equation}
354: \label{def:EHelfrich}
355: E_{\mathrm{Helfrich}}(S) =
356: \int_S \bigl[ k(H - H_0)^2 + \overline k K\bigr]\, d\H^2.
357: \end{equation}
358: Here $k,\ \overline k$, and $H_0$ are constants, $H$ and $K$
359: are the (scalar) total and Gaussian curvature, and $\H^2$ is the
360: two-dimensional Hausdorff measure. This energy functional,
361: and many generalizations in the same vein, have been
362: remarkably successful in describing the
363: wide variety of vesicle shapes that are observed in laboratory
364: experiments~\cite{Seifert97}.
365: If we add an assumption of symmetry, in which
366: both `sides' of the surface have the same properties, the `spontaneous
367: curvature' $H_0$ vanishes.
368: For closed surfaces in the same homotopy class, by the Gau\ss-Bonnet
369: Theorem, $E_{\mathrm{Helfrich}}$
370: is essentially equal to the
371: \emph{Willmore functional}~\cite{Willmore65}
372: \begin{gather}
373: W(S) \,=\, \frac{1}{2}\int_S H^2 \, d\H^2,\label{eq:willmore}
374: \end{gather}
375: which arises in a wide variety of
376: situations other than that mentioned here.
377:
378: %Since the original derivation of~\pref{def:EHelfrich}
379: %is phenomenological in nature,
380: %many authors have presented alternative derivations
381: %from more elementary descriptions~\cite{SzleiferKramerBenShaulGelbartSafran90,%
382: %ChaconSomozaTarazona98,%
383: %OversteegenBlokhuis00,LaradjiMouritsen00,Wuerger00}.
384:
385: \medskip
386:
387: In this paper we discuss the above mentioned bilayer
388: models in two space dimensions,
389: \emph{i.e.}\ we consider functions
390: whose support resembles fattened curves in $\R^2$;
391: in the limit $\e\to0$ the thickness of these curves tends to zero.
392: If we consider the limit curves to be two-dimensional restrictions
393: of cylindrical surfaces, we observe that the
394: Gaussian curvature $K$ vanishes.
395: If we also assume that the bilayer is symmetric
396: then the Helfrich energy reduces to the Elastica functional $\W$, the
397: classical bending energy of the curve,
398: \begin{gather}
399: \W(\gamma) \,=\, \frac{1}{2}\int_\gamma \kappa^2 \,
400: d\H^1,\label{eq:def-elastica}
401: \end{gather}
402: where $\kappa$ is the curvature of the curve. This functional has a
403: long history going back at least to Jakob Bernoulli; critical points of
404: this energy are known as \emph{Euler elastica}, see \cite{truesdell} for
405: a historical review.
406:
407: \subsection{The main result: the singular limit $\e\to0$}
408:
409: In the main result of this paper we study the limit $\e\to0$, and
410: connect all the aspects that were discussed above. The result revolves
411: around a rescaled functional
412: \[
413: \G_\e = \frac{\F_\eps - 2M}{\e^2},
414: \]
415: where $M$ is the mass of $u$ and $v$ (see~\pref{def:Ke}).
416:
417: Take any sequence \mbox{$(u_\e,v_\e)_{\eps>0}\subset \K_\eps$}.
418: \emph{If the rescaled energy $\G_\e(u_\e,v_\e)$ is
419: bounded}, then
420: \begin{enumerate}
421: \item the sequence $u_\e$ converges as measures;
422: \item the limit measure can be represented as a collection of curves (which may overlap);
423: \item the support of $u_\e$ is of `thickness' approximately $2\e$;
424: \item in particular the \emph{total length} of the curves is $M/2$;
425: \item each curve in the collection is \emph{closed}.
426: \end{enumerate}
427: Boundedness of $\G_\e$ along the sequence therefore implies partial
428: localization: the support of $u_\e$ resembles a tubular
429: $\e$-neighbourhood of a curve.
430: In addition, \emph{stretching} and \emph{fracture} are also represented
431: in this result: stretching corresponds to deviation in total length
432: %\footnote{Actually the result is stronger: it is \emph{local} deviation
433: %from optimal length that is penalized; deviation from optimal
434: %\emph{total} length follows as a consequence}
435: from the optimal value $M/2$, and fracture to creation of non-closed
436: curves. Neither is possible for a sequence of bounded $\G_\e$; both are
437: penalized in $\G_\e$ at an order larger than $O(1)$, or equivalently, at
438: an order larger than $O(\e^2)$ in $\F_\e$.
439:
440: The resistance to \emph{bending} arises as the Gamma-limit of the
441: functional $\G_\e$ itself:
442: \begin{enumerate}
443: \item For any sequence \mbox{$(u_\e,v_\e)_{\eps>0}\subset \K_\eps$},
444: \[
445: \W(\Gamma) := \sum_{\gamma\in\Gamma} \W(\gamma) \leq
446: \liminf_{\e\to0} \G_\e(u_\e,v_\e),
447: \]
448: where $\Gamma $ is the collection of limit curves introduced above, and $\W(\gamma)$ is the curve bending energy defined in
449: \eqref{eq:def-elastica};
450: \item For any given collection of curves $\Gamma$, there exists a
451: sequence \mbox{$(u_\e,v_\e)_{\eps>0}\subset \K_\eps$} such that
452: $\W(\Gamma) = \lim_{\e\to0} \G_\e(u_\e,v_\e)$.
453: \end{enumerate}
454: The details of this result are given as Theorem~\ref{the:main}.
455:
456: Summarizing, only partially localized structures can have moderate
457: energy, and
458: such thin structures also display resistance to bending, stretching, and
459: fracture.
460: %Summarizing, only partially localized structures can have moderate
461: %energy, and
462: %such thin structures also display the three properties mentioned above:
463: %the resistance to fracture and stretching
464: %is encoded in the requirement that the limit curves are
465: %closed and of fixed length, and the resistance to bending
466: %is represented by the limit functional $\W$.
467:
468: \medskip
469: So far we have described the result that we prove in this paper. The
470: proof, however, also suggests a more refined result, that we present
471: here as a conjecture: if a sequence \mbox{$(u_\e,v_\e)_{\eps>0}\subset
472: \K_\eps$} converges
473: to a collection of curves~$\Gamma$,
474: then
475: \begin{align}\label{eq:asym-Fe}
476: \F_\eps(u_\eps,v_\eps)\,\geq\, 2M + \mathfrak L_M(\Gamma)
477: + \eps \mathfrak{B}(\Gamma)
478: + \eps^2 \sum_{\gamma\in\Gamma}\W(\gamma) + O(\e^3).
479: \end{align}
480: In this development,
481: \begin{itemize}
482: \item the function $\mathfrak L_M(\Gamma)$ is a measure of
483: the deviation of $\sum_{\gamma\in\Gamma}\length(\gamma)$ from length
484: $M/2$: $\mathfrak L_M(\Gamma)\geq0$,
485: and $\mathfrak L_M(\Gamma)=0 $ only if this sum equals $M/2$;
486: \item $\mathfrak B(\Gamma)$ is zero if all curves $\gamma\in\Gamma$ are
487: closed, and strictly positive (a positive constant times the
488: number of endpoints) if at least one $\gamma\in\Gamma$ is open.
489: \end{itemize}
490: As for the Gamma-convergence above,
491: the inequality in this statement should be interpreted as
492: a `best-possible' behaviour; while for given limit curves $\gamma$
493: sequences exist for which this inequality is an equality,
494: it is also possible to construct sequences along which the difference
495: between the left- and right-hand sides is unbounded.
496: \subsection{Strategy of the analysis and overview of the paper}
497: At first glance it is not obvious that the functionals $\F_\eps$
498: encode bending stiffness effects. Curvature terms in fact appear at
499: order $\eps^2$ of an asymptotic development, which asks for a
500: careful analysis. Most information is hidden in the
501: nonlocal \Wasserstein
502: distance term. The special property that makes our analysis work is that
503: this distance decouples into one-dimensional problems. This reduction is
504: the major step in our analysis and the connection to the \emph{optimal mass
505: transport problem} is the most important tool here. Whereas the idea
506: becomes very clear in simple situations which resemble the lim-sup
507: construction in the Gamma-convergence proof, some effort is required to
508: prove the lim-inf estimate in the general case, where certain `defects'
509: such as high curvature and vanishing thickness of approximating structures
510: have to be controlled simultaneously.
511:
512: In the next section we motivate our choice of scaling in the functional
513: $\F_\eps$ and illustrate the behaviour of $\F_\eps$ and $\G_\e$ in two
514: particular examples. In Section \ref{sec:wass} we give definitions of
515: the
516: \Wasserstein distance and `systems of curves' which describe the class of
517: limit structures.
518: A precise statement of our results is given in Section
519: \ref{sec:main}. We briefly review the optimal mass transport problem
520: in Section \ref{sec:mtp}. In Section~\ref{sec:overview} we provide a short guide to the structure of the proof of our
521: main result, Theorem~\ref{the:main}, and the proof itself is given
522: in Sections \ref{sec:liminf} and \ref{sec:limsup}.
523: We conclude in Section \ref{sec:discussion} with some discussions of generalisations and the wider implications of this result.
524:
525: \bigskip
526: \textbf{Acknowledgement.} The authors are grateful for several inspiring
527: discussions with Prof.\ Felix Otto and Prof.\ Reiner Sch\"atzle.
528: \\[5mm]
529: %==================
530: % notation
531: %==================
532: \begin{minipage}{12cm}
533: \centering{\bf Summary of notation}
534: \ \\[3mm]
535: \begin{small}
536: \begin{tabular}{lll}
537: $\F_\eps(\cdot,\cdot)$ & functionals describing the mesoscale energy & \eqref{def:Fe}\\
538: $\G_\eps(\cdot,\cdot)$ & rescaled functionals & \eqref{eq:def-Ge-uv}\\
539: $\K_\eps$ & domain of $\F_\eps$, $\G_\eps$ & \eqref{def:Ke}\\
540: $\W(\gamma)$ & bending energy of a curve $\gamma$ & \eqref{eq:def-elastica}\\
541: $\W(\Gamma)$ & generalized bending energy of a system of
542: curves $\Gamma$ & \eqref{eq:def-W-Gamma}\\
543: $d_1(\cdot,\cdot)$ & \Wasserstein distance &\eqref{eq:def-d-1}\\
544: $d_p(\cdot,\cdot)$ & $p$-Wasserstein distance & \eqref{eq:def-d-p}\\
545: $\spt(\Gamma)$
546: & support of a system of
547: curves $\Gamma$ & Def.~\ref{def:sys-curves}\\
548: $\theta(\Gamma,\cdot)$
549: & \multiplicity of a system of
550: curves $\Gamma$ & Def.~\ref{def:sys-curves}\\
551: $|\Gamma|$
552: & total mass of a system of
553: curves $\Gamma$ & Def.~\ref{def:sys-curves}\\
554: $\Lip_1(\R^2)$ & Lipschitz continuous functions with Lipschitz constant 1
555: & \pref{eq:phi_eps}\\
556: $\mathcal{T},\mathcal{E}$
557: & transport set and set of endpoints of rays &
558: Def.~\ref{def:rays}\\
559: $E$ & $\{s: \gamma(s) \text{ lies inside a transport ray}\}$ &
560: Def.~\ref{def:para}\\
561: $\theta(s)$ & ray direction in $\gamma(s)$ & Def.~\ref{def:para}\\
562: $L^+(s),L^-(s),l^+(s)$ & positive, negative and effective ray
563: length in $\gamma(s)$&
564: Def.~\ref{def:para}\\
565: $\psi(\cdot,\cdot)$ & parametrization & Def.~
566: \ref{def:para}\\
567: $E_\delta$ & boundary points with uniformly bounded ray lengths
568: & \eqref{eq:def-E-delta}\\
569: $\alpha(s),\beta(s)$ & direction of ray and difference to tangent
570: at $\gamma(s)$& \eqref{eq:def-beta}\\
571: $\ma(s,\cdot)$ & mass coordinates & \eqref{eq:def-mass}\\
572: $\te(s,\cdot)$ & length coordinates & \eqref{eq:def-t_s}\\
573: $M(s)$ & mass over $\gamma(s)$& \eqref{eq:def-Ms}\\
574: $E_i,\theta_i,\psi_i$,
575: \\
576: $L^+_i,L^-_i,l_i^+$,& corresponding quantities for a collection
577: $\{\gamma_i\}$ & Rem.~\ref{rem:corr-lengths}\\
578: $\alpha_i,\beta_i,\ma_i,\te_i,M_i$\\
579: $E_{\eps,i},\theta_{\eps,i},\psi_{\eps,i}$,\\
580: $L^+_{\eps,i},L^-_{\eps,i},l_{\eps,i}^+$
581: & corresponding quantities for a collection
582: $\{\gamma_{\eps,i}\}$ & Rem.~
583: \ref{rem:corr-lengths}\\
584: $\alpha_{\eps,i},\beta_{\eps,i},\ma_{\eps,i},\te_{\eps,i},M_{\eps,i}$
585: \end{tabular}
586: \end{small}
587: \end{minipage}
588:
589: \section{Heuristics: the functional $\F_\eps$}
590: \label{sec:heuristics}
591:
592: In the preceding section we postulated that the energy
593: functional $\F_\eps$, defined in~\pref{def:Fe}, favours
594: %as a candidate for stable
595: partial localization. More precisely,
596: we claim that for
597: small but fixed $\e>0$ functions $(u,v)$ with moderate $\G_\eps(u,v)$
598: will be partially localized, in the sense that
599: their support resembles one or more fattened curves.
600:
601: In this section we first provide some heuristic arguments to support
602: this claim, mostly to develop some intuitive
603: understanding of the functional~$\F_\eps$ and its properties.
604:
605: For the discussion below it is
606: useful to remark that the distance $d_1(u,v)$ as defined in
607: Definition \ref{def:wasserstein}
608: scales as the mass
609: $\int u = \int v$ times a length scale, and can be interpreted as
610: a $u$-weighted spatial distance between the mass distributions
611: $u$ and~$v$.
612:
613: \subsection{Fixed thickness (\boldmath $\e=1$), part 1:
614: comparing disc and strip structures}
615: To gain some insight into the properties of $\F_\eps$ we start with the case of
616: fixed $\e=1$, and study the limit of large mass
617: \[
618: M = \int u = \int v,\qquad M \longrightarrow \infty.
619: \]
620: We now compare two different specific realizations, a disc and a strip.
621: \begin{figure}[ht]
622: \centerline{\includegraphics[width=\textwidth]{disc}}%width=\textwidth
623: \caption{Disc and strip structures.}
624: \label{fig:disc}
625: \end{figure}
626:
627: \begin{enumerate}
628: \item The \textbf{disc: } we
629: concentrate all the mass of $u$ into a disc
630: in $\R^2$,
631: of radius $R\sim M^{1/2}$, surrounded by the mass of $v$ in an
632: annulus (Figure~\ref{fig:disc}). In this setup mass is transported (in
633: the sense of the definition of $d_1(u,v)$) over
634: a distance $O(R) = O(M^{1/2})$, so that
635: \begin{equation}
636: \label{sim:d1}
637: d_1(u,v)\sim MR\sim M^{3/2} \qquad \text{as }M\to\infty.
638: \end{equation}
639: On the other hand, the interfacial length
640: $\int\mod{\nabla u}$ equals $2\pi R\sim M^{1/2}$,
641: and the functional $\Fone$ picks up the larger of the two:
642: \[
643: \Fone(u,v) = d_1(u,v) + \int\mod{\nabla u} \sim M^{3/2}.
644: \]
645:
646: \item For the \textbf{strip} we take the support of $u$ to be
647: a rectangle of width 2 and length $M/2$, flanked by two strips
648: of half this width for $\supp(v)$ (Figure~\ref{fig:disc}).
649: The transport distance can now be taken constant and equal to 1,
650: so that we find
651: \[
652: \Fone(u,v) = M + M + 4 = 2M+4.
653: \]
654: For this choice of geometry $\Fone$ scales linearly with $M$ in
655: the limit $M\to\infty$.
656: \end{enumerate}
657:
658: Comparing the two we observe that the linear structure has lower energy,
659: for large mass, than a spherical one. This is a first
660: indication that $\Fone$ may favour partial localization.
661:
662: \medskip
663:
664: With this example we can also illustrate an additional property.
665: Let us take for $\supp(u)$ a strip of thickness $t$ and length $M/t$, and recalculate
666: the value of $\Fone$:
667: \[
668: \Fone(u,v) = \left(\frac t2 + \frac 2t\right) M + 2t.
669: \]
670: For large $M$ this expression is dominated by the value of
671: the prefactor $t/2+ 2/t$, suggesting two additional features:
672: \begin{enumerate}
673: \item There is a preferred thickness, which is that value of $t$ for which
674: $t/2+2/t$ is minimal (i.e. $t=2$);
675: \item For the preferred thickness, the functional $\Fone$ equals $2M$
676: plus `other terms'.
677: \end{enumerate}
678: Although these statements only give suggestions, not proofs, for the case of general geometry,
679: we shall see below that they both are true, and that these examples
680: do demonstrate the general behaviour. The `other terms' in the
681: example above are the single term~$2t$, a term which is associated
682: with the ends of the strip; we now turn to a different case,
683: in which there are no loose ends, but instead the curvature of the structure
684: creates an energy penalty.
685:
686: \subsection{Fixed thickness (\boldmath $\e=1$), part 2: comparing line with
687: ring structures}
688:
689: In the previous section we argued that the energy $\Fone$ favours
690: objects of thickness~$2$ and penalizes loose ends. We now consider
691: \emph{ring} structures, and we will see that also the \emph{curvature} carries
692: an energy penalty.
693:
694: \medskip
695:
696: Let the supports of $u$ and $v$ be the ring structures of
697: Fig.~\ref{fig:rings}: the support of $u$ is a single ring between
698: circles of radii $r_2$ and $r_3$, and the support of $v$ consists
699: of two rings flanking $\supp(u)$.
700:
701: \begin{figure}[ht]
702: \centerline{\psfig{width=\textwidth,figure=rings}}
703: \caption{Ring structures; the pairs $(u,v)$ on the right-hand figure
704: indicate the values of $u$ and $v$ in that region in the plane.}
705: \label{fig:rings}
706: \end{figure}
707:
708: We first do some direct optimization to reduce the number of
709: degrees of freedom in this geometry.
710: For fixed outer radii $r_1$ and $r_4$ one may optimize $\Fone$
711: over variations of $r_2$ and $r_3$ that respect the mass
712: constraint $\int u = \int v$, and this optimization results
713: in the values of $r_2$ and $r_3$ given by
714: \begin{equation}
715: \label{def:r2r3}
716: r_2^2 = \tfrac12(R^2 + r_1^2), \qquad \text{and}\qquad
717: r_3^2 = \tfrac12(R^2 + r_4^2)
718: \end{equation}
719: in terms of the mean radius $R = \tfrac12(r_1+r_4)$. The functional $\Fone$
720: can then be computed explicitly in terms of $r_1$ and $r_4$ (see
721: Appendix~\ref{app:ring_solutions}). The interesting quantity is actually
722: the energy per unit mass,
723: $\Fone/M$, as a function of the mean radius $R$ and the thickness $t$ of the
724: structure,
725: \begin{gather*}
726: M\,:=\, \int u, \qquad t\,:=\, \frac{r_4-r_1}{2}.
727: \end{gather*}
728: Expanding $\Fone/M$ around $R=\infty$ and $t=2$, we find
729: \begin{align}
730: \label{asymp:F1}
731: \frac{\Fone}{M}(R,t)\,=\,& 2 +\frac{1}{4}(t-2)^2
732: +\frac{1}{4}R^{-2} + O(|t-2|^3 + R^{-3}),
733: \end{align}
734: see Appendix \ref{app:ring_solutions}.
735:
736: Again we recognize a preference for structures of thickness $t=2$;
737: we now also observe a penalization of the curvature in the term $R^{-2}/4$.
738: The main result of this paper indeed is to identify this
739: curvature penalization for structures of arbitrary geometry, in the form of an elastica limit energy.
740:
741:
742:
743:
744: \subsection{Rescaling and renormalization}
745:
746: In the previous sections we have seen that the functional $\F_\eps$
747: at $\e=1$ has a preference for structures of thickness $2$. It is easy
748: to see, by repeating the arguments above, that this becomes a
749: preference for thickness $2\e$ in the general case. For instance,
750: the development~\pref{asymp:F1} generalizes to
751: \[
752: \frac{\F_\eps}{M}(R,t) = 2 + \frac1{4}(t/\e-2)^2 + \frac14 \e^2R^{-2}
753: + O\bigl(|t/\e-2|^3+ \e^3R^{-3}\bigr).
754: \]
755:
756: This expression provides us with a recipe for identifying the bending
757: energy in the limit $\e\to0$. The curvature, which is equal to $1/R$ for
758: this
759: spherical geometry, enters the expression above as the term $\e^2R^{-2}/4$.
760: This suggests that the alternative functional
761: \[
762: \G_\e(u,v) := \frac1{\e^2}\bigl[\F_\eps(u,v) - 2M\big],
763: \qquad \text{where}\qquad M = \int u = \int v,
764: \]
765: may have a (Gamma-)limit similar to the Elastica functional.
766: This suggestion is proved in Theorem~\ref{the:main}.
767: \begin{comment}
768: To relate this to the asymptotic devellopment \eqref{eq:asym-Fe} we
769: remark that $M=4\pi R$ so that the curvature term becomes
770: $2\cdot\eps^2\frac{1}{4}2\pi R \frac{1}{R^2}$.
771: \end{comment}
772: %==========================================================
773: % Wasserstein etc
774: %==========================================================
775: \section{\Wasserstein distance, systems of curves and a generalized curve
776: bending energy}
777: \label{sec:wass}
778: In this section we introduce some basic definitions and concepts.
779: \begin{definition}
780: \label{def:wasserstein}
781: Consider $u,v\in L^1(\Rn)$ with compact support in $\Rn$ satisfying the mass balance
782: \begin{gather}\label{eq:fun-eq-mass}
783: \int_{\Rn} u\,d\LL^2 \,=\, \int_{\Rn} v\,d\LL^2\,=\, 1.
784: \end{gather}
785: The \Wasserstein distance $d_1(u,v)$ is defined as
786: \begin{eqnarray}\label{eq:def-d-1}
787: d_1(u,v) &:=& \min \Big( \int_{\Rn\times\Rn} |x-y| \,d\gamma(x,y)\Big)
788: \end{eqnarray}
789: where the minimum is taken over all Radon measures $\gamma$ on $\Rn\times\Rn$
790: with \emph{marginals}
791: $u\LL^2$ and $v\LL^2$, that means $\gamma$ satisfies
792: \begin{eqnarray}
793: \int_{\Rn\times\Rn} \varphi(x) \,d\gamma(x,y) &=& \int_{\Rn}\varphi u\,d\LL^2,\label{eq:marg-u}\\
794: \int_{\Rn\times\Rn} \psi(y) \,d\gamma(x,y) &=& \int_{\Rn}\psi v\,d\LL^2\label{eq:marg-v}
795: \end{eqnarray}
796: for all $\varphi,\psi\in C^0_c(\Rn)$.
797: \end{definition}
798: The \Wasserstein distance
799: is characterized by the optimal mass transport problem described in
800: \mbox{section \ref{sec:mtp}}.
801: \begin{remark}\label{rem:wass}
802: For $\mu,\nu\in \mathcal{P}_1$, where
803: \begin{gather*}
804: \mathcal{P}_1(\Rn)\,:=\, \big\{ \mu\text{ Radon measure on }\Rn \,:\, \int_{\Rn}d\mu\,=\,1,
805: \,\,\int_{\Rn} |x|\,d\mu(x)\,<\,\infty\big\},
806: \end{gather*}
807: we can define the \Wasserstein distance $d_1(\mu,\nu)$ analogously to \eqref{eq:def-d-1},
808: substituting $u\LL^2$ and $v\LL^2$
809: in \eqref{eq:marg-u}, \eqref{eq:marg-v} by $\mu$ and $\nu$.
810: The minimum in \eqref{eq:def-d-1} is always attained and $d_1$ is a distance function on
811: $\mathcal{P}_1$~\cite{Amb04}. Moreover $d_1$ is continuous in the sense that for
812: $(\mu_k)_{k\in\N}\subset \mathcal{P}_1(\Rn), \mu\in \mathcal{P}_1(\Rn)$
813: \begin{gather*}
814: d_1(\mu_k,\mu) \,\to\, 0\qquad \Longleftrightarrow\qquad
815: \begin{cases}
816: \mu_k\to\mu\text{ as Radon measures and}\\
817: \int_{\Rn} |x|\,d\mu_k\,\to\,\int_{\Rn} |x|\,d\mu,
818: \end{cases}
819: \end{gather*}
820: see \cite[Theorem 3.2]{Amb04}.
821:
822: A family of related distances is given by
823: the $p$-Wasserstein distances
824: \begin{eqnarray}\label{eq:def-d-p}
825: d_p(\mu,\nu) &:=& \min \Big( \int_{\Rn\times\Rn} |x-y|^p \,d\gamma(x,y)\Big)^{\frac{1}{p}},
826: \end{eqnarray}
827: where the minimum is taken over all Radon measures $\gamma$ on $\Rn\times\Rn$
828: with \emph{marginals} $\mu$ and $\nu$. The \Wasserstein distance
829: coincides with the 1-Wasserstein distance.
830: \end{remark}
831: We will describe the limit structures in terms of \emph{systems of closed $W^{2,2}$-curves}.
832: The following definitions are mainly taken from \cite{BeM}.
833: \begin{definition}\label{def:sys-curves}
834: We call $\gamma:\R\to\R^2$ a closed $W^{2,2}$-curve if
835: \begin{align*}
836: &\gamma\in W^{2,2}_{\loc}(\R,\R^2),\\
837: &\gamma^\prime(r)\neq 0\text{ for all }r\in \R,\\
838: &\gamma\text{ is $L$-periodic for some }0<L<\infty.
839: \end{align*}
840: If $\gamma$ is $L$-periodic and injective on $[0,L)$ the
841: curve
842: $\gamma$ is called \emph{simple}. The length $L(\gamma)$ of a closed curve
843: $\gamma$ is defined by
844: \[
845: L(\gamma) := \int_0^L |\gamma'(r)|\, dr,
846: \]
847: where $L$ is the the minimal period of $\gamma$.
848:
849: A $W^{2,2}$-\emph{system of closed curves} is a finite collection of
850: closed $W^{2,2}$-curves. We represent a system of curves by a
851: \emph{multiset} $\{\gamma_i\}_{i=1,...,N}$, $N\in\N$, \emph{i.e.}
852: a set in which repeated elements are counted with multiplicity.
853:
854: A system of curves $\Gamma=\{\gamma_i\}_{i=1,...,N}$ is called \emph{disjoint}
855: if each $\gamma_i$ is simple and $\gamma_i(\R)\cap\gamma_j(\R)=\emptyset$ for $i\not=j$.
856: $\Gamma$
857: has \emph{no transversal crossings} if for any $1\leq i,j\leq m$, $s_i,s_j\in\R$,
858: \begin{gather*}
859: \gamma_i(s_i)\,=\,\gamma_j(s_j)\text{ implies that }\gamma_i^\prime(s_i)\text{ and }
860: \gamma_j^\prime(s_j)\text{ are parallel}.
861: \end{gather*}
862: We define the \emph{support}, \emph{multiplicity} and \emph{total mass}
863: of a system of curves $\Gamma=\{\gamma_i\}_{i=1,...,N}$ by
864: \begin{align*}
865: \spt(\Gamma) &:=\, \bigcup_{i=1}^N \gamma_i(\R),\\
866: \theta(\Gamma,x) &:=\, \#\big\{(i,s) : \gamma_i(s) =x,\ 1\leq i\leq N,\
867: 0\leq s< L(\gamma_i)\big\},\\
868: |\Gamma| &:=\, \sum_{i=1}^N L(\gamma_i)
869: \end{align*}
870: and define a corresponding
871: Radon measure $\mu_\Gamma$ on $\Rn$ to be the measure that
872: satisfies
873: \begin{eqnarray}\label{eq:def-gamma-mu}
874: \int_{\Rn} \varphi\,d\mu_\Gamma &=& \sum_{i=1}^N
875: \int_0^{L(\gamma_i)}\varphi(\gamma_i(s))\,|\gamma_i^\prime(s)|\,ds
876: \end{eqnarray}
877: for all $\varphi\in C^0_c(\Rn)$.
878:
879: Two systems of curves are identified if the corresponding Radon measures
880: coincide.
881: \end{definition}
882: \begin{remark}
883: We can represent a given system of closed curves $\Gamma$ by a multiset
884: $\Gamma\,=\,\{\gamma_i\}_{i=1,...,N},\,\text{ where for all }i=1,...,N$
885: \begin{gather}
886: \gamma_i\text{ is one-periodic, with $1$ being the smallest
887: possible period},\label{eq:stand-sys}\\
888: \gamma_i\text{ is parametrized proportional to arclength.}
889: \label{eq:curve-arclength}
890: \end{gather}
891: \end{remark}
892: We generalize the classical curve bending energy %\emph{Willmore functional}
893: defined in \eqref{eq:def-elastica} to $W^{2,2}$-systems of closed curves.
894: \begin{definition}\label{def:W-Gamma}
895: Let $\Gamma$ be a $W^{2,2}$-system of closed
896: curves represented by $\Gamma=\{\gamma_i\}_{i=1,...,N}$ as in
897: \eqref{eq:stand-sys}-\eqref{eq:curve-arclength} and set $L_i = L(\gamma_i)$.
898: Then we define
899: \begin{eqnarray}\label{eq:def-W-Gamma}
900: \W(\Gamma) &:=& \frac{1}{2}\sum_{i=1}^N L_i^{-3}\int_0^1
901: \gamma_i^{\prime\prime}(s)^2 \,ds.
902: \end{eqnarray}
903: \end{definition}
904: \begin{remark}
905: The support and total mass of a system of curves coincide with
906: the support and total mass of the corresponding Radon measure~\cite{Sim}.
907: The multiplicity functions of $\Gamma$ and of $\mu_\Gamma$ coincide
908: $\Ha^{1}$-almost everywhere on the support of $\Gamma$.\\
909: By Remark 3.8, Proposition 4.5, and Corollary 4.8 of \cite{BeM05},
910: we obtain that
911: $\Gamma$ is a $W^{2,2}$-system of closed curves without transversal crossings if and only if
912: $\mu_\Gamma$ is a \emph{Hutchinson varifold} with weak mean curvature
913: $\vec{H}\in L^2(\mu_\Gamma)$
914: such that in every point of $\supp(\Gamma)$ a unique tangent line exists. Moreover
915: \begin{eqnarray}
916: \W(\Gamma) &=& \frac{1}{2}\int |\vec{H}|^2\,d\mu_\Gamma\label{eq:W-hutch}
917: \end{eqnarray}
918: holds.
919: \end{remark}
920: %==================
921: % Main results
922: %==================
923: \section{Main results}\label{sec:main}
924: To investigate the limit behaviour of the functionals
925: $ \F_\eps$ as $\eps\to 0$
926: we fix $M>0$ and study the rescaled functionals
927: \begin{eqnarray}\label{eq:def-Ge-uv}
928: \G_\eps(u,v) &:=& \frac{1}{\eps^2} \Big( \F_\eps(u,v) - 2M\Big).
929: \end{eqnarray}
930:
931: \begin{theorem}
932: \label{the:main}
933: The curve bending energy $\W$ as defined in \eqref{eq:def-W-Gamma} is the Gamma-limit
934: of the functionals $\G_\eps$ in the following sense.
935: \begin{liste2}
936: \item
937: Let $(u_\eps, v_\eps)_{\eps>0}\subset
938: L^1(\Rn)\times L^1(\Rn)$, $R>0$ and a Radon measure $\mu$ on $\Rn$ be given with
939: \begin{eqnarray}
940: \spt(u_\eps)&\subset& B_R(0)\quad\text{ for all }\eps>0,\label{eq:spt-ue}\\
941: u_\eps\LL^2 &\to& \mu\quad\text{ as Radon measures on
942: }\R^2,\label{eq:conv-ue}
943: \end{eqnarray}
944: and
945: \begin{eqnarray}
946: \liminf_{\eps\to 0} \G_\eps(u_\eps,v_\eps)&<&\infty.\label{eq:liminf-bounded}
947: \end{eqnarray}
948: Then there is a $W^{2,2}$-system of closed curves $\Gamma$
949: such that
950: \begin{gather}
951: 2\mu_\Gamma = \mu,\label{eq:rep-Gamma}\\
952: \spt(\Gamma)\text{ is bounded},\label{eq:Gamma-0}\\
953: \Gamma\text{ has no transversal crossings},\label{eq:Gamma-1}\\
954: 2|\Gamma|\,=\, M\label{eq:Gamma-3},
955: \end{gather}
956: and such that the \emph{liminf-estimate}
957: \begin{eqnarray}\label{eq:liminf-esti}
958: \W(\Gamma) &\leq& \liminf_{\eps\to 0} \G_\eps(u_\eps)
959: \end{eqnarray}
960: holds.
961: \item
962: Let $\Gamma$ be a $W^{2,2}$-system of closed curves such that
963: \eqref{eq:Gamma-0}-\eqref{eq:Gamma-3} holds. Then there exists a
964: sequence $(u_\eps,v_\eps)_{\eps>0}\subset \K_\eps$
965: such that \eqref{eq:spt-ue} holds for some $R>0$,
966: \begin{eqnarray}\label{eq:conv-ue-Gamma}
967: u_\eps\LL^2 &\to& 2\mu_\Gamma\quad\text{ as Radon measures on }\R^2,
968: \end{eqnarray}
969: and such that the \emph{limsup-estimate}
970: \begin{eqnarray}\label{eq:limsup-esti}
971: \W(\Gamma) &\geq& \limsup_{\eps\to 0} \G_\eps(u_\eps)
972: \end{eqnarray}
973: is satisfied.
974: \end{liste2}
975: \end{theorem}
976: \begin{remark}
977: \label{rem:convergence}
978: Let $X$ be the space of nonnegative Radon measures on $\R^2$. Define a concept of convergence on $X\times X$ by which
979: \begin{gather*}
980: (\mu_i,\nu_i) \,\to\, (\mu,\nu)\quad\text{ in }X\times X\quad\text{ iff }\quad
981: \begin{cases}
982: \quad\mu_i\,\to\,\mu\quad\text{ as Radon measures on }\R^2\text{ and }\\
983: \quad\bigcup_{i\geq i_0}\spt(\mu_i) \text{ is bounded for some }i_0\in\N.
984: \end{cases}
985: \end{gather*}
986: Note that no condition is placed on $\nu_i$.
987: We now define
988: \begin{eqnarray*}
989: \G_\eps(\mu,\nu) &:=&
990: \begin{cases}
991: \G_\eps(u,v) \quad &\text{ if }\mu\,=\,u\LL^2,\,\nu\,=\,v\LL^2
992: \text{ for some }(u,v)\in \K_\eps,\\
993: \infty &\text{ otherwise}
994: \end{cases}
995: \end{eqnarray*}
996: and
997: \begin{eqnarray*}%\label{eq:def-W-mu}
998: \G_0(\mu,\nu) &:=&
999: \begin{cases}
1000: \W(\Gamma) &\begin{cases}
1001: \text{if $\mu=\nu$ is given by a $W^{2,2}$-system of closed curves}\\
1002: \text{$\Gamma=\{\gamma_i\}_{i=1,...,N}$ satisfying \eqref{eq:rep-Gamma}-\eqref{eq:Gamma-3}}
1003: \end{cases} \\[5\jot]
1004: \infty & \text{otherwise.}
1005: \end{cases}
1006: \end{eqnarray*}
1007: Then Theorem \ref{the:main} can be rephrased as
1008: \begin{eqnarray}\label{eq:gamma-conv}
1009: \lim_{\eps\to 0} \G_\eps &=& \G_0
1010: \end{eqnarray}
1011: in the sense of Gamma-convergence with respect to the topology of
1012: $X\times X$ defined above.
1013: \end{remark}
1014:
1015: \begin{remark}
1016: Implicit in the formulation of Theorem~\ref{the:main} is the following
1017: (contrapositive) statement: if a sequence $(u_\e,v_\e)_{\e>0}\subset
1018: \K_\e$ converges to a measure that \emph{cannot} be represented
1019: by a system of curves satisfying~\eqref{eq:rep-Gamma}--\eqref{eq:Gamma-3},
1020: then $\G_\e$ is unbounded along this sequence. Put differently,
1021: each of
1022: the following is penalized in $\F_\e$ at an order lower (\emph{i.e.} larger) than $\eps^2$:
1023: \begin{enumerate}
1024: \item a fracture in the limit structure, occurring as a nonclosed curve;
1025: \item a total mass $|\Gamma|$ that deviates from~$M/2$;
1026: \item a non-even \multiplicity.
1027: \end{enumerate}
1028:
1029: In the Introduction we conjectured the asymptotic
1030: development~\eqref{eq:asym-Fe}.
1031: Theorem~\ref{the:main} justifies the zeroth and second order of this
1032: development in the following way: if $F_\e-2M$ is of order $O(\e^2)$,
1033: then the three types of degeneracy above can not occur, and the dominant term
1034: is given in the limit by $\e^2 \W$.
1035: \end{remark}
1036:
1037: \begin{remark}
1038: By the definition of the set $\K_\eps$ and the functionals $\F_\eps$ in
1039: \eqref{def:Fe}, \eqref{def:Ke} it is clear that for a
1040: sequence $(u_\eps, v_\eps)_{\eps>0}$ along which $\G_\e$ is bounded the supports of $u_\eps$ and $v_\eps$ necessarily concentrate as $\eps \to
1041: 0$. The crucial conclusions are (a) that the concentration is on curves
1042: rather than on points or other sets and (b) that the limit measure can
1043: be constructed as a sum of curves, each with density two. The latter
1044: property
1045: implies a uniform thickness of structures with bounded energy and
1046: paraphrases the stretching resistance of lipid bilayers.
1047: \end{remark}
1048:
1049: \begin{remark}
1050: We also obtain a compactness result for sequences with bounded $\G_\e$: Let
1051: $(u_\eps, v_\eps)_{\eps>0}\subset L^1(\R^2)\times
1052: L^1(\R^2)$ satisfy \eqref{eq:spt-ue}, \eqref{eq:liminf-bounded}.
1053: Then there exists a Radon measure $\mu$ on $\R^2$ such that
1054: \eqref{eq:conv-ue} holds for a subsequence $\eps\to 0$. Theorem
1055: \ref{the:main} then implies that there
1056: is a $W^{2,2}$-system $\Gamma$ of closed
1057: curves such
1058: that the conclusions \eqref{eq:rep-Gamma}-\eqref{eq:liminf-esti} hold.
1059: \end{remark}
1060: %====================================================
1061: % Mass transport
1062: %====================================================
1063: \section{The Mass Transport Problem}
1064: \label{sec:mtp}
1065: In this section we recall the classical Monge problem of optimal mass transport and review
1066: some results that we will use later.
1067: \begin{definition}\label{def:trans-maps}
1068: Fix two non-negative Borel functions $u,v\in L^1(\R^2)$ with compact support
1069: satisfying the
1070: mass balance \eqref{eq:fun-eq-mass}.
1071: By $\mathcal{A}(u,v)$ we denote the set of all Borel maps $S:\R^2\to\Rn$
1072: \emph{pushing $u$ forward to $v$}, that is
1073: \begin{eqnarray}\label{eq:trans-map}
1074: \int\eta(S(x)) u(x)\,dx &=& \int_{\Rn} \eta(y)v(y)\,dy
1075: \end{eqnarray}
1076: holds for all $\eta\in C^0(\Rn)$.
1077: \end{definition}
1078: \begin{prob-mass-transport}
1079: For $u,v$ as in Definition \ref{def:trans-maps} minimize the \emph{transport cost}
1080: $I:\mathcal{A}(u,v)\to \R$,
1081: \begin{eqnarray}\label{eq:transportcost}
1082: I(S) &:=& \int_\Omega |S(x)-x| u(x)\,dx.
1083: \end{eqnarray}
1084: \end{prob-mass-transport}
1085: The following dual formulation is due to Kantorovich \cite{Kan}.
1086: \begin{prob-Kantorovich}
1087: For $u,v$ as in Definition \ref{def:trans-maps} maximize $K: \Lip_1(\Rn)\to\R$,
1088: \begin{eqnarray}\label{eq:phi_eps}
1089: K(\phi) &:=& \int_\Omega \phi(x)(u-v)(x)dx,
1090: \end{eqnarray}
1091: where $\Lip_1(\Rn)$ denotes the space of Lipschitz functions on $\R^2$
1092: with Lipschitz constant not larger than $1$.
1093: \end{prob-Kantorovich}
1094: There is a vast literature on the optimal mass transportation problem
1095: and an impressive number of applications, see for example \cite{EGan,TW,CFC,Amb,Vil,JKF,Ott}.
1096: We only list a few results which we will use later.
1097: \begin{theorem}[\cite{CFC},\cite{FeC}]%[Caffarelli, Feldman, McCann], [Feldman, McCann]
1098: Let $u,v$ be given as in Definition \ref{def:trans-maps}.
1099: \begin{liste2}
1100: \item There exists an optimal transport map $S\in \mathcal{A}(u,v)$.
1101: \item There exists an optimal Kantorovich potential $\phi\in \Lip_1(\Rn)$.
1102: \item The identities
1103: \begin{gather*}
1104: d_1(u,v)\,=\, I(S)\,=\,K(\phi)
1105: \end{gather*}
1106: hold.
1107: \item Every optimal transport map $S$ and every optimal Kantorovich potential $\phi$
1108: satisfy
1109: \begin{eqnarray}\label{eq:dual-crit}
1110: \phi(x) -\phi(S(x)) &=& |x-S(x)|\quad\text{ for almost all }x\in\spt(u).
1111: \end{eqnarray}
1112: \end{liste2}
1113: \end{theorem}
1114: The optimal transport map and the optimal Kantorovich potential are in general not
1115: unique. We can choose $S$ and $\phi$ enjoying some additional properties.
1116: \begin{proposition}[\cite{CFC,FeC}]\label{prop:add-prop}
1117: There exists an optimal transport map $S\in \mathcal{A}(u,v)$ and an optimal
1118: Kantorovich potential $\phi$ such that
1119: \begin{eqnarray}
1120: \phi(x) &=& \min_{y\in\spt(v)}\big(\phi(y)+|x-y|\big)\quad\text{ for
1121: any }x\in\spt(u),\label{eq:add-prop-phi1}\\
1122: \phi(y) &=& \max_{x\in\spt(u)}\big(\phi(x)-|x-y|\big)\quad\text{ for
1123: any }y\in\spt(v),\label{eq:add-prop-phi2}
1124: \end{eqnarray}
1125: and such that $S$ is the unique monotone transport map in the sense of
1126: \cite{FeC},
1127: \begin{gather*}
1128: \frac{x_1-x_2}{|x_1-x_2|}
1129: +\frac{S(x_1)-S(x_2)}{|S(x_1)-S(x_2)|}\,\neq\, 0\quad\text{ for all
1130: }x_1\neq x_2\in\Rn\text{ with }S(x_1)\neq S(x_2).
1131: \end{gather*}
1132:
1133: \begin{comment}
1134: We can choose $S$ to be essentially one-to-one
1135: in the sense that there exists an (optimal) transport map $\tilde{S}\in\mathcal{A}(v,u)$ pushing
1136: $v$ forward to $u$ such that
1137: \begin{eqnarray}
1138: \tilde{S}\circ S &=& Id\quad\text{ almost everywhere on }\spt(u),\label{eq:1-1-u}\\
1139: S\circ \tilde{S} &=& Id\quad\text{ almost everywhere on }\spt(v).\label{eq:1-1-v}
1140: \end{eqnarray}
1141: \end{comment}
1142: \end{proposition}
1143: \begin{comment}
1144: \begin{proof}
1145: We let $\phi,S$ be the optimal Kantorovich potential and optimal transport map constructed in
1146: \cite{CFC}. Then \eqref{eq:add-prop-phi1} and \eqref{eq:add-prop-phi2} hold and $S$ is the unique
1147: monotone transport map in the sense of \cite{FeC}. Moreover the pushforward
1148: $(Id\times S)_\#\, u\LL^2$ gives the unique solution of the secondary variational problem
1149: described in \cite{AP}. Choose $\tilde{S}\in \A(v,u)$ to be the optimal transport pushing
1150: $v$ forward to $u$. Then $(\tilde{S}\times Id)_\#\, v\LL^2$ coincides with $(Id\times S)_\#\, u\LL^2$
1151: and we deduce \eqref{eq:1-1-u}, \eqref{eq:1-1-v}.
1152: \end{proof}
1153: \end{comment}
1154: We will extensively use the fact that, by~\eqref{eq:dual-crit} the optimal transport is organized along
1155: {\itshape transport rays} which are defined as follows.
1156: \begin{definition}[\cite{CFC}]
1157: \label{def:rays}
1158: Let $u,v$ be as in Definition \ref{def:trans-maps} and let $\phi\in \Lip_1(\Rn)$
1159: be the optimal transport map as in Proposition \ref{prop:add-prop}.
1160: A {\itshape transport ray} is a line segment in $\Rn$ with endpoints $a,b\in\Rn$ such that
1161: $\phi$ has slope one on that segment and $a,b$ are maximal, that is
1162: \begin{gather*}
1163: a\in\spt(u),\,b\in\spt(v),\quad a\neq b,\\
1164: \phi(a)-\phi(b)\,=\,|a-b|\\
1165: |\phi(a+t(a-b))-\phi(b)| \,<\, |a+t(a-b)-b|\quad\text{ for all }t>0,\\
1166: |\phi(b+t(b-a))-\phi(a)| \,<\, |b+t(b-a)-a|\quad\text{ for all }t>0.
1167: \end{gather*}
1168: We define the {\itshape transport set }$\mathcal{T}$ to consist of all points which
1169: lie in the (relative) interior of some transport ray and define $\mathcal{E}$ to be the
1170: set of all endpoints of rays.
1171: \end{definition}
1172:
1173: Some important properties of
1174: transport rays are given in the next proposition.
1175: \begin{proposition}[\cite{CFC}]\label{prop:cfm-rays}
1176: \begin{liste2}
1177: \item Two rays can only intersect in a common endpoint.
1178: %\item The support of $u,v$ is almost covered by the interior of rays:\\
1179: % $\spt(u_\eps)\cup\spt(v_\eps)\,\subset\, \mathcal{T}\cup \mathcal{B}$ where $\mathcal{B}$ is
1180: % an $\LL^2$-nullset.
1181: \item
1182: The endpoints $\mathcal{E}$ form a Borel set of measure zero.
1183: \item If $z$ lies in the interior of a ray with endpoints $a\in\spt(u),b\in\spt(v)$
1184: then $\phi$ is differentiable in $z$ with $\nabla\phi(z)\,=\,(a-b)/|a-b|$.
1185: \end{liste2}
1186: \end{proposition}
1187: In Section \ref{subsec:para} we will use the transport rays to parametrize the support of $u$ and
1188: to compute the \Wasserstein distance between $u$ and $v$.
1189: %==================
1190: % Overview of proof
1191: %==================
1192:
1193: \section{Overview of the proof of Theorem~\ref{the:main}}
1194: \label{sec:overview}
1195:
1196: \subsection{The lower bound}
1197: The heart of the proof of the lower bound~\pref{eq:liminf-esti} is a
1198: parametrization of $\supp(u)\cup\supp(v)$ that allows us to rewrite the
1199: functionals $\F_\e$ and $\G_\e$ in terms of the geometry of $\supp(u)$ and
1200: $\supp(v)$ and the structure of the transport rays.
1201:
1202: The first step in the proof of~\pref{eq:liminf-esti} is to pass to
1203: functions $u$ and $v$ with smooth boundaries (Proposition~\pref{prop:red})
1204: so that $\partial \supp(u)$ can be represented by a collection $\{\gamma_i\}$
1205: of smooth curves. For the discussion in this section we will pretend
1206: that there is only one curve $\gamma:[0,L)\to\R^2$, which is closed and
1207: which we parametrize by arclength.
1208:
1209: The properties of the \Wasserstein distance $d_1$ give that almost
1210: every point $x\in \supp(u)\cup\supp(v)$ lies in the interior of exactly
1211: one transport
1212: ray; the set $\supp(u)\cup \supp(v)$ can therefore be written, up to a
1213: null set, as the disjoint union of transport rays.
1214:
1215: In addition, along each ray mass is transported `from $u$ to $v$', implying
1216: that each ray intersects $\partial\supp(u)$. We therefore can parametrize
1217: the collection of rays by their intersections with $\partial \supp(u)$.
1218:
1219:
1220: The ray direction of the ray $\mathcal R(s)$ that passes through $\gamma(s)$
1221: defines a unit length vector $\theta(s)$;
1222: we now introduce a parametrization $\psi$ by
1223: \[
1224: \psi(s,t) := \gamma(s) + t\theta(s).
1225: \]
1226: In terms of such a parametrization the first term of $\F_\e$ has a simple
1227: description as the length of the curve $\gamma$,
1228: \[
1229: \e\int|\nabla u| = \int_0^L 1\,ds = L.
1230: \]
1231: To rewrite $d_1(u,v)$ we
1232: use the fact that transport takes place along transport rays; when restricted to a single ray the transport problem becomes one-dimensional, and even explicitly solvable.
1233:
1234: The central estimate in this part of the proof is the following lower bound:
1235: \begin{align}
1236: \label{est:central}
1237: d_1(u,v) %&\geq \int_0^L \frac{\sin^3\beta(s)}{3\alphaprime(s)^2\eps}
1238: % \Bigg[\Big(
1239: % 1+\frac{2\alphaprime(s)\eps}{\sin^2\beta(s)}M(s)\Big)^{\frac{3}{2}}
1240: % +\Big(
1241: % 1-\frac{2\alphaprime(s)\eps}{\sin^2\beta(s)}M(s)\Big)^{\frac{3}{2}}
1242: % -2\,\Bigg]ds\\
1243: &\geq \int_0^L \Biggl[\frac{\eps}{\sin\beta(s)}M(s)^2+
1244: \frac{\eps^3}{4\sin^5\beta(s)}\alphaprime(s)^2M(s)^4\Biggr]ds.
1245: \end{align}
1246: In this inequality the geometry of $\supp(u)$ is characterized by
1247: functions $\alpha$, $\beta$, and~$M$, which are illustrated in Figure~\ref{fig:explanation}.
1248: \begin{figure}[ht]
1249: \centering
1250: \includegraphics[height=5.5cm]{geometry2}
1251: \caption{The parametrization of $\supp(u)\cup\supp(v)$. On the left six transport rays are drawn over the two supports; on the right the functions $\alpha$, $\beta$, and $M$ are indicated. At a boundary point $\gamma(s)$, $\alpha(s)$ is the angle of the transport ray through that point, and $\beta(s)$ is the angle between $\gamma$ and the ray. Finally, $M(s)ds$ is the amount of mass contained between the rays at positions $s$ and $s+ds$.}
1252: \label{fig:explanation}
1253: \end{figure}
1254: The functions $\alpha$ and $\beta$ are angles: $\alpha(s)$ is the angle
1255: between the ray direction $\theta(s)$ and a reference direction and $\beta(s)$ is the angle between $\theta(s)$ and the tangent $\gamma'(s)$ to $\gamma$. The function $M(\cdot)$, finally, measures the amount of mass supported on a ray. The relation between the function $M(\cdot)$ and the scalar $M$ in~\pref{def:Ke} is
1256: \[
1257: M = \int_0^L M(s)\, ds.
1258: \]
1259:
1260:
1261: With the estimate~\pref{est:central}, the main result can readily be appreciated. Let a sequence $(u_\e,v_\e)$ be such that $\G_\e(u_\e,v_\e)$
1262: is bounded as $\e\to0$, and let $\alpha_\e$, $\beta_\e$, and $M_\e$ be the
1263: associated geometric quantities. With the inequality above,
1264: \begin{align}
1265: \G_\e(u_\e,v_\e) &= \frac{\F_\e(u_\e,v_\e)-2M_\e}{\e^2} \notag \\
1266: &\geq \int_0^L \Bigl[\frac1{\e^2} - \frac{2M_\e(s)}{\e^2}+\frac{M_\e(s)^2}{\e^2\sin\beta_\e(s)}+
1267: \frac{\alphaprime_\e(s)^2M_\e(s)^4}{4\sin^5\beta_\e(s)}\Bigr]ds\notag\\
1268: &= \int_0^L \Biggl[\frac1{\e^2}\bigl(1-M_\e(s)\bigr)^2\
1269: +\frac1{\e^2}\left(\frac1{\sin\beta_\e(s)}-1\right) M_\e(s)^2+
1270: \frac{\alphaprime_\e(s)^2M_\e(s)^4}{4\sin^5\beta_\e(s)}\Biggr]ds
1271: \label{est:informal}
1272: \end{align}
1273: Note that each of the three terms above is non-negative.
1274: If $\G_\e(u,v)$ is bounded, then one concludes that as $\e\to0$ the mass $M_\eps(s)$ tends to one for almost all $s$. Similarly, $\beta_\e$ converges to $\pi/2$, implying that
1275: the ray angle $\alpha_\eps$ converges to the angle
1276: of the normal to $\gamma_\eps$, and in a weak sense therefore $\alpha^\prime_\eps$ converges to the curvature $\kappa_\eps=\gamma_\eps^{\prime\prime}$.
1277: The inequality above suggests
1278: a $W^{2,2}_{\mathrm{loc}}$ bound for curves $\gamma_\eps$ with the last term approximating
1279: \begin{gather*}
1280: \frac{1}{4}\int_0^L \kappa_\eps(s)^2\,ds
1281: \end{gather*}
1282: This integral is the bending energy that we expect in the limit.
1283:
1284: Based on the inequality~\pref{est:informal} we show that the boundary curves of $\spt(u_\eps)$ are compact, that the limit is given
1285: by a $W^{2,2}$ system of closed curves as in \eqref{eq:rep-Gamma}-\eqref{eq:Gamma-3}
1286: and that the lim-inf estimate is satisfied. We finally conclude
1287: that the limit of the mass distributions $u_\eps$ is identical to the limit of the boundary curves.
1288: \begin{comment}
1289: \[
1290: M_\e(s) \to 1\qquad\text{and} \qquad \sin\beta_\e(s) \to 1,
1291: \]
1292: implying that in the limit the ray angle $\alpha$ coincides with
1293: the angle of the normal to $\gamma$; therefore $\alpha'$ coincides
1294: in the limit with the curvature $\kappa=\gamma''$ of $\gamma$, and the last term in the integral above converges to
1295: \[
1296: \frac14\int_0^L \kappa(s)^2\,ds.
1297: \]
1298: \end{comment}
1299:
1300: The main reasons why the proof in Section~\ref{sec:liminf} is more involved than this brief explanation are related to the gaps in the reasoning above:
1301: \begin{itemize}
1302: \item A ray can intersect $\partial\supp(u)$ multiple times, and care must be taken to ensure that the parametrization is a bijection.
1303: \item The ray direction $\theta$ is not necessarily smooth; to be precise, when the ray length tends to zero (which is equivalent to $M(\cdot)$ vanishing), $\theta$ may vary wildly.
1304: \item Similarly, when $M(\cdot)$ vanishes, the $L^2$-estimate on $\alpha'$ in~\pref{est:informal} degenerates,
1305: resulting in a compactness problem for the boundary curves.
1306: \end{itemize}
1307:
1308: \subsection{The upper bound}
1309: With the machinery of the lower bound in place, and with the insight that is provided, the upper bound becomes a relatively simple construction. Around a given limit curve tubular neighbourhoods are constructed for the supports of $u$ and $v$. The thickness of these neigbourhoods can be chosen just right, and the calculation that leads to~\pref{est:central} now gives an exact value for $\G_\e$.
1310:
1311: The main remaining difficulty of the proof is the fact that a $W^{2,2}$-curve need not have a non-self-intersecting tubular neighbourhood of any thickness, and therefore an approximation argument is needed.
1312:
1313: %==================
1314: % liminf
1315: %==================
1316: \section{Proof of the lim-inf estimate}
1317: \label{sec:liminf}
1318: In this section we prove the first part of Theorem~\ref{the:main}.
1319: The main idea is to use the optimal Kantorovich potential
1320: of the mass transport problem of $u_\eps$ to $v_\eps$ to construct a
1321: parametrization of the support of $u_\eps$
1322: and $v_\eps$ and to derive a lower bound for the functionals
1323: $\G_\eps(u_\eps,v_\eps)$. This lower bound yields the compactness of the
1324: systems of curves
1325: which describe the boundary of the support of $u_\eps$ as well as the
1326: desired lim-inf estimate.
1327:
1328: We first show that we can restrict ourselves to a class of `generic
1329: data'.
1330: \begin{proposition}\label{prop:red}
1331: It is sufficient to prove the \emph{liminf} part of Theorem
1332: \ref{the:main} under the additional assumptions that
1333: \begin{align}
1334: &M=1,\quad\text{ that is } \int u_\e \,=\, \int v_\e \,=\,1\quad\text{ for
1335: }(u_\e,v_\e)\in\K_\eps, \label{eq:M=1}\\[2mm]
1336: &\sup_{\eps>0}\G_\eps(u_\eps,v_\eps)\,\leq
1337: \,\Lambda\quad\text{ for some }\quad 0<\Lambda<\infty,\label{eq:bound-G-eps}\\[2mm]
1338: &(u_\eps, v_\eps)\,\in\, \K_\eps\quad\text{ for all
1339: }\eps>0,\label{eq:ass-K}\\[2mm]
1340: &\lim_{\eps\to 0} \G_\eps(u_\eps,v_\eps)\,=\,\liminf_{\eps\to 0}
1341: \G_\eps(u_\eps,v_\eps)\quad\text{ exists,} \label{eq:ex-lim-G}\\[2mm]
1342: &\partial\spt(u_\eps)\text{ is given by a disjoint system of
1343: smooth, simple curves } \Gamma_{\eps}. \label{eq:reg-u-eps}
1344: \end{align}
1345: \end{proposition}
1346: \begin{proof}
1347: To prove that we can restrict ourselves to $M=1$ we do a
1348: spatial rescaling and set
1349: \begin{gather*}
1350: \tilde{\eps}\,:=\,\frac{\eps}{M},\quad
1351: \tilde{u}_{\tilde{\eps}}(x)\,:=\, M u_\eps(Mx),\quad
1352: \tilde{v}_{\tilde{\eps}}(x)\,:=\, M v_\eps(Mx).
1353: \end{gather*}
1354: We then obtain
1355: \begin{gather}
1356: \int_{\R^2} \tilde{u}_{\tilde{\eps}}\,=\,\int_{\R^2} \tilde{v}_{\tilde{\eps}}\,=\,1,\qquad\qquad
1357: {\tilde{\eps}}\int |\nabla \tilde{u}_\eps|\,=\,
1358: \frac{\eps}{M}\int|\nabla\label{eq:tilde-u-bdry}
1359: u_\eps|.
1360: \end{gather}
1361: Moreover, if $S$ is the optimal transport map from $u_\eps$ to
1362: $v_\eps$, then $\tilde{S}(x):=
1363: M^{-1}S(Mx)$ is the optimal transport map from $\tilde{u}_{\tilde{\eps}}$
1364: to $\tilde{v}_{\tilde{\eps}}$ and
1365: \begin{gather}
1366: d_1(\tilde{u}_{\tilde{\eps}},\tilde{v}_{\tilde{\eps}})\,=\, M^{-2}
1367: d_1(u_\eps,v_\eps). \label{eq:tilde-d1}
1368: \end{gather}
1369: We obtain from \eqref{eq:tilde-u-bdry},
1370: \eqref{eq:tilde-d1} that
1371: \begin{align}
1372: {\G}_{\tilde{\eps}}(\tilde{u}_{\tilde{\eps}},\tilde{v}_{\tilde{\eps}})
1373: \,=\,&\frac{1}{\tilde{\eps}^2}
1374: \Big(\frac{1}{\tilde{\eps}}d_1(\tilde{u}_{\tilde{\eps}},\tilde{v}_{\tilde{\eps}})
1375: +\tilde{\eps}\int|\nabla\tilde{u}_{\tilde{\eps}}| -2\Big)\notag\\
1376: \,=\,& \frac{M^2}{\eps^2}\Big( \frac{1}{M\eps}d_1(u_\eps,v_\eps) +
1377: \frac{\eps}{M}\int|\nabla u_\eps| - \frac{2}{M}\int u_\eps\Big)\notag\\
1378: =&\, M \G_\eps(u_\eps,v_\eps).\label{eq:tilde-Ge}
1379: \end{align}
1380: We observe that $(\tilde{u}_{\tilde{\eps}},\tilde{v}_{\tilde{\eps}})\in
1381: \K_{\tilde{\eps}}$ iff
1382: $(u_\eps,v_\eps)\in \K_\eps$ and that
1383: $(\tilde{u}_{\tilde{\e}})_{\tilde{\e}>0}$ satisfies
1384: \eqref{eq:spt-ue}, \eqref{eq:conv-ue} for \mbox{$\tilde{R}:= R/M$} and
1385: $\tilde{\mu}$ defined by
1386: \begin{gather}
1387: \int \eta(x)\,d\tilde{\mu}(x)\,=\, \int
1388: \frac{1}{M}\eta\big(\frac{x}{M}\big)\,d\mu(x)\quad\text{ for }\eta\in
1389: C^0_c(\R^2). \label{eq:tilde-mu}
1390: \end{gather}
1391: If Theorem \ref{the:main} holds for $M=1$ then there exists a
1392: $W^{2,2}$-system of closed curves $\tilde{\Gamma}$ with
1393: \begin{gather}
1394: 2\mu_{\tilde{\Gamma}}\,=\,\tilde{\mu},\qquad
1395: 2|\tilde{\Gamma}|\,=\,1,\label{eq:the-mu-tilde}
1396: \end{gather}
1397: such that \eqref{eq:Gamma-0}, \eqref{eq:Gamma-1} holds and
1398: \begin{gather}
1399: \W(\tilde{\Gamma}) \,\leq\, \liminf_{\tilde{\eps}\to
1400: 0}\G_{\tilde{\eps}}(\tilde{u}_{\tilde{\eps}},\tilde{v}_{\tilde{\eps}}). \label{eq:liminf-tilde}
1401: \end{gather}
1402: We deduce from \eqref{eq:tilde-mu} and \eqref{eq:the-mu-tilde}
1403: that $\mu$ is
1404: given as system of closed curves:
1405: Let $\tilde{\Gamma}\,=\,\{\tilde{\gamma}_i\}_{i=1,...,N}$ with
1406: \begin{gather*}
1407: \tilde{\gamma}_i : [0,\tilde{L}_i)\,\to\,\R^2, \qquad
1408: |\tilde{\gamma}_i^\prime|\,=\,1
1409: \end{gather*}
1410: and define a
1411: system of curves $\Gamma=\{\gamma_i\}_{i=1,...,N}$ by
1412: \begin{gather*}
1413: \gamma_i : [0,L_i)\,\to\,\R^2,\quad L_i:= M\tilde{L}_i,\\
1414: \gamma_i(s):= M\tilde{\gamma}\big(\frac{s}{M}\big)\quad\text{ for }s\in
1415: [0,L_i).
1416: \end{gather*}
1417: Then $|\gamma_i^\prime|=1$ and $\Gamma$ satisfies \eqref{eq:Gamma-0},
1418: \eqref{eq:Gamma-1}.
1419: We obtain from \eqref{eq:tilde-mu} that
1420: \begin{align*}
1421: \int \eta \,d\mu\,&=\, M\int
1422: \eta(Mx)\,d\tilde{\mu}(x)\,=\,2M\sum_{i=1}^m\int_0^{\tilde{L}_i}
1423: \eta(M\tilde{\gamma}_i(s))\,ds \\
1424: \,&=\, 2\sum_{i=1}^m\int_0^{L_i}\eta(\gamma_i(s))\,ds
1425: \,=\, 2\int \eta\,d\mu_{\Gamma},
1426: \end{align*}
1427: which yields $\mu=2\mu_\Gamma$ and, by taking $\eta=1$ inside $B_R(0)$,
1428: that \eqref{eq:Gamma-3} holds.
1429: Finally we deduce from \eqref{eq:tilde-mu}
1430: \begin{align*}
1431: \W(\Gamma)\,&=\,\frac{1}{2}\sum_{i=1}^m \int_0^{L_i}
1432: |\gamma_i^{\prime\prime}(s)|^2\,ds \\
1433: \,&=\,
1434: \frac{1}{2}\sum_{i=1}^m \int_0^{L_i}
1435: \frac{1}{M^2}\Big|\tilde{\gamma}_i^{\prime\prime}\Big(\frac{s}{M}\Big)\Big|^2\,ds
1436: \,=\,
1437: \frac{1}{2}\sum_{i=1}^m \int_0^{\tilde{L}_i}
1438: \frac{1}{M}|\tilde{\gamma}_i^{\prime\prime}(s)|^2\,ds \\
1439: \,&=\,
1440: \frac{1}{M}\W(\tilde{\Gamma})
1441: \end{align*}
1442: and by \eqref{eq:tilde-Ge}, \eqref{eq:liminf-tilde} this yields that
1443: \eqref{eq:liminf-esti}
1444: holds. Therefore we have shown that it is sufficient to prove the
1445: lim-inf part of Theorem \ref{the:main} for $M=1$.
1446:
1447: Eventually restricting ourselves to a subsequence $\eps\to 0$ we can
1448: assume \eqref{eq:ex-lim-G}. The existence of a $0<\Lambda<\infty$
1449: such that \eqref{eq:bound-G-eps} holds follows from
1450: \eqref{eq:liminf-bounded}. By the definition of $\G_\eps$ this in
1451: particular implies \eqref{eq:ass-K}.
1452:
1453: It remains to show that it is sufficient to consider functions $u_\eps$
1454: such that $\partial\spt(u_\eps)$ is smooth. We already have seen that we
1455: can assume that $(u_\eps,v_\eps)\in\K_\eps$. This implies that
1456: $\spt(u_\eps)$ is a set of finite perimeter. Moreover
1457: $\spt(u_\eps)\subset B_R(0)$ and by
1458: \cite[Theorem~3.42]{AFP} there exists a sequence of
1459: open sets $(E_k)_{k\in\N}$ with smooth boundary such that
1460: \begin{eqnarray}
1461: \Chi_{E_k} &\to& \eps u_\eps\quad\text{ in }L^1(\Rn),\label{eq:E-k-1}\\
1462: |\nabla\Chi_{E_k}|(\Rn) &\to& \eps|\nabla u_\eps|(\Rn)\label{eq:E-k-2}\\
1463: E_k &\subset& B_{2R}(0).\label{eq:E-k-3}
1464: \end{eqnarray}
1465: We set
1466: \begin{eqnarray}\label{eq:def-r-k}
1467: r_k &:=& \eps^{\frac{1}{2}}|E_k|^{-\frac{1}{2}}
1468: \end{eqnarray}
1469: and define functions $\hat{u}_k$,
1470: \begin{eqnarray}
1471: \hat{u}_k(x) &:=& \frac{1}{\eps}\Chi_{E_k}(r_k^{-1}x).\label{eq:def-u-hat}
1472: \end{eqnarray}
1473: This yields
1474: \begin{gather}
1475: \int_{\R^2} \hat{u}_k \,=\, \frac{r_k^2}{\eps}|E_k|\,=\,1\label{eq:mass-hat-u}
1476: \end{gather}
1477: and
1478: \begin{gather}
1479: \int_{\R^2} |\nabla\hat{u}_k| \,=\, \frac{r_k}{\eps}\int_{\Rn}
1480: |\nabla E_k|.\label{eq:surf-hat-u}
1481: \end{gather}
1482: Moreover, from \eqref{eq:E-k-1}, \eqref{eq:def-r-k} and
1483: $|\spt(u_\eps)|=\eps$ we
1484: observe that
1485: \begin{gather}
1486: r_k\,\to\, 1\quad\text{ as }k\to\infty.\label{eq:conv-k}
1487: \end{gather}
1488: We compute that
1489: \begin{eqnarray*}
1490: \int_{\R^2} |\hat{u}_k -u_\eps| &=&
1491: \frac{1}{\eps}\int_{\R^2} \Big|\Chi_{E_k}(r_k^{-1}x)-\eps
1492: u_\eps(x)\Big|\,dx\\
1493: &\leq& \frac{1}{\eps}\int_{\R^2} \big|\Chi_{E_k}(r_k^{-1}x)
1494: -\eps u_\eps(r_k^{-1}x)\big|\,dx\\
1495: &&+\frac{1}{\eps}\int_{\R^2} \big|\eps u_\eps(r_k^{-1}x)-\eps
1496: u_\eps(x)|\,dx\\
1497: &=& \frac{r_k^2}{\eps}\int_{\R^2} \big|\Chi_{E_k}-\eps
1498: u_\eps(x)\big|\,dx
1499: +\int_{\R^2} \big|u_\eps(r_k^{-1}x)-u_\eps(x)|\,dx.
1500: \end{eqnarray*}
1501: Since the right-hand side converges to zero as $k\to\infty$ by
1502: \eqref{eq:E-k-1}, \eqref{eq:conv-k} we deduce that
1503: \begin{eqnarray}
1504: \hat{u}_k &\to& u_\eps\quad\text{ in }L^1({\R^2})\label{eq:conv-u-hat-L1}
1505: \end{eqnarray}
1506: as $k\to\infty$. Equations \eqref{eq:E-k-2}, \eqref{eq:surf-hat-u} and
1507: \eqref{eq:conv-k} yield that
1508: \begin{eqnarray}
1509: \int_{\R^2} |\nabla\hat{u}_k| &\to& \int_{\R^2} |\nabla
1510: u_\eps|\quad\text{ as }k\to\infty.\label{eq:conv-u-hat-area}
1511: \end{eqnarray}
1512: We construct approximations $\hat{v}_k$ of $v_\eps$ by setting
1513: \begin{gather}
1514: \hat{v}_k(x) \,:=\, v_\eps(x)\big(1-\eps \hat{u}_k(x)\big) +w_k(x),
1515: \label{eq:def-hat-v}
1516: \end{gather}
1517: for a suitable $w_k\in L^1\big(B_{2R}(0),\{0,\eps^{-1}\}\big)$ with
1518: \begin{gather}
1519: \int_{\R^2} ( w_k -\eps v_\eps\hat{u}_k)\,=\, 0. \label{eq:def-wk}
1520: \end{gather}
1521: We then deduce from \eqref{eq:def-hat-v}, \eqref{eq:def-wk} that
1522: $(\hat{u}_k,\hat{v}_k )\in\K_\eps$. It follows from
1523: \eqref{eq:conv-u-hat-L1}, \eqref{eq:def-wk} and $u_\eps v_\eps=0$ that
1524: $w_k\to 0$ as $k\to \infty$. Together with \eqref{eq:def-hat-v} we
1525: obtain that
1526: \begin{gather}
1527: \lim_{k\to\infty}\hat{v}_k\,=\, v_\eps\quad\text{ in
1528: }L^1(\R^2). \label{eq:conv-hat-vk}
1529: \end{gather}
1530:
1531:
1532: By \eqref{eq:E-k-3}, \eqref{eq:conv-u-hat-L1} and \eqref{eq:conv-hat-vk}
1533: we deduce from Remark \ref{rem:wass} the continuity of the \Wasserstein
1534: distance term in $\F_\eps(\hat{u}_k,\hat{v}_k)$. Together with
1535: \eqref{eq:conv-u-hat-area} this yields
1536: \begin{eqnarray*}
1537: \F_\eps(\hat{u}_k,\hat{v}_k) &\to& \F_\eps(u_\eps,v_\eps)\quad\text{ as }k\to\infty.
1538: \end{eqnarray*}
1539: Therefore we can choose $k(\eps)\in\N$ such that for
1540: $\tilde{u}_\eps:=\hat{u}_{k(\eps)}$, $\tilde{v}_\eps:=\hat{v}_{k(\eps)}$
1541: satisfy the estimate
1542: \begin{eqnarray*}
1543: \|u_\eps -\tilde{u}_\eps\|_{L^1({\R^2})} +
1544: |\G_\eps(u_\eps,v_\eps)-\G_\eps(\tilde{u}_\eps,\tilde{v}_\eps)| &\leq&
1545: \eps.
1546: \end{eqnarray*}
1547: This yields a sequence $(\tilde{u}_\eps,\tilde{v}_\eps)_{\eps>0}$ that satisfies
1548: \eqref{eq:reg-u-eps} and has the same limit as
1549: $(u_\eps,v_\eps)_{\eps>0}$ in \eqref{eq:conv-ue}, \eqref{eq:liminf-esti}.
1550: \end{proof}
1551:
1552:
1553: From now on, for the rest of this section, we assume that
1554: \eqref{eq:M=1}-\eqref{eq:reg-u-eps} hold.
1555: %============================================================
1556: % Parametrization by rays
1557: %============================================================
1558: \subsection{Parametrization by rays}
1559: \label{subsec:para}
1560: Due to \eqref{eq:reg-u-eps} the boundary $\partial\spt(u_\eps)$ is smooth.
1561: We are free to choose a suitable representation.
1562: \begin{remark}\label{rem:arclength-curves}
1563: We can represent $\partial\spt(u_\eps)$ by a disjoint system
1564: $\Gamma_\eps$ of closed curves such that for all $\gamma\in \Gamma_\eps$
1565: \begin{gather}
1566: \gamma \text{ is parametrized by arclength}, |\gamma^\prime|=1,\notag
1567: \\
1568: \spt(u_\eps)\text{ is `on the left hand side' of
1569: }\gamma. \label{eq:spt-ue-lhs}
1570: \end{gather}
1571: In particular $\gamma$ is $L(\gamma)$-periodic and
1572: \begin{gather*}
1573: \det (\gamma^\prime(s),\nu(s)) \,\leq\, 0,
1574: \end{gather*}
1575: where $\nu(s)$ denotes the outward unit normal of $\spt(u_\eps)$ at
1576: $\gamma(s)$.
1577: \end{remark}
1578: Let $\phi\in \Lip_1(\R^2)$ be an optimal Kantorovich potential for
1579: the mass transport from $u_\eps$ to $v_\eps$ as in Proposition
1580: \ref{prop:add-prop}, with
1581: $\mathcal{T}$ being the set of transport rays as in Definition
1582: \ref{def:rays}.
1583: Recall that $\phi$ is differentiable, with $|\nabla \phi| = 1$, in the
1584: relative interior of any ray.
1585:
1586: \begin{definition}[Parametrization by rays]\label{def:para}
1587: For $\gamma\in \Gamma_\eps$ we define
1588: \begin{liste}
1589: \item
1590: a set $E$ of `inner points' with respect to the transport set
1591: $\mathcal{T}$,
1592: \begin{eqnarray*}
1593: E &:=& \{s\in \R :\gamma(s) \in \mathcal{T}\},
1594: \end{eqnarray*}
1595: \item
1596: a direction field
1597: \begin{eqnarray*}
1598: \theta: E\to \sphere^1, \qquad \theta(s) := \nabla\phi(\gamma(s)),
1599: \end{eqnarray*}
1600: \item
1601: the positive and negative total ray length $L^+,L^-:E\to \R$,
1602: \begin{eqnarray}
1603: L^+(s) &:=& \sup\{t>0 : \phi(\gamma(s)+t\theta
1604: (s))-\phi(\gamma(s))\,=\,t\},\label{eq:def-L+}\\
1605: L^-(s) &:=& \inf\{t<0 : \phi(\gamma(s)+t\theta
1606: (s))-\phi(\gamma(s))\,=\,t\},\label{eq:def-L-}
1607: \end{eqnarray}
1608: \item the \emph{effective} positive ray length $l^+: E\to\R$,
1609: \begin{gather}
1610: l^+(s) := \sup\ \{t\in [0,L^+(s)] : \gamma(s) + \tau\theta(s)\in
1611: \Int\bigl(\spt(u_\eps)\bigr)\text{ for all }0<\tau<
1612: t\}.\label{eq:def-l+}
1613: \end{gather}
1614: \end{liste}
1615: Finally we define a map $\psi$ which will serve as a parametrization of
1616: $\spt(u_\eps)\cup\spt(v_\eps)$ by
1617: \begin{gather}
1618: \psi: \big[0,L(\gamma)\big)\times\R\,\to\,\R^2,\quad
1619: \psi(s,t) \,:=\, \gamma(s) + t\theta(s).\label{eq:def-psi}
1620: \end{gather}
1621: \end{definition}
1622: \begin{remark}\label{rem:corr-lengths}
1623: All objects defined above are properties of $\gamma$ even if we
1624: do not denote this dependence explicitely. When dealing with a
1625: collection of curves $\{\gamma_i :i=1,...,N\}$ or $\{\gamma_{\eps,i} : \eps>0,
1626: i=1,...,N(\eps)\}$ then $E_i$, $l_{\eps,i}^+$ etc. refer to the
1627: objects defined for the corresponding curves.
1628: \end{remark}
1629:
1630: \begin{figure}[ht]
1631: \centerline{\psfig{width=\textwidth,figure=para}}%width=\textwidth
1632: \caption{The parametrization in a `simple' situation. Here
1633: $l^+(s)=L^+(s)$ coincide.}
1634: \label{fig:para}
1635: \end{figure}
1636:
1637:
1638: See the Figures
1639: \ref{fig:para} and \ref{fig:para2} for an illustration of the
1640: parametrization defined above. The {\itshape effective ray length} is
1641: introduced to obtain the injectivity of the parametrization
1642: in the case that a ray crosses several times the boundary
1643: of $\spt(u_\eps)$ (as it is the case in the situation depicted in Figure
1644: \ref{fig:para2}).
1645: The set $\{\gamma(s): l^+(s)>0\}$ represents the points of the boundary
1646: where mass is transported in the `right
1647: direction'.
1648:
1649: We will first prove some results which are analogous to those in \cite{CFC}.
1650: \begin{lemma}\label{lem:usc}
1651: The ray direction $\theta: E\to\R^2$ is continuous.
1652: The positive, negative, and effective positive ray lengths $L^+, L^-,l^+ : E\to\R$
1653: are measurable.
1654: \end{lemma}
1655:
1656: \begin{proof}
1657: Consider a sequence $(s_k)_{k\in\N}\subset E$, $s\in E$, with $s_k\to s$.
1658: Choose $\tilde{t}_k\leq t_k$ such that
1659: \begin{align}\label{eq:t_n}
1660: L^+(s_k)-\frac{1}{k} \,&<\, t_k \, <\, L^+(s_k),\\
1661: l^+(s_k)-\frac{1}{k} \,&<\, \tilde{t}_k\, < \, l^+(s_k).\label{eq:tilde-t_n}
1662: \end{align}
1663: It follows from \eqref{eq:def-L+} and \eqref{eq:t_n} that
1664: \begin{eqnarray}
1665: \phi(\gamma(s_k)+t_k\theta(s_k)) &=& \phi(\gamma(s_k))+t_k.\label{eq:theta_n}
1666: \end{eqnarray}
1667: Since $(t_k)_{k\in\N},(\tilde{t}_k)_{k\in\N}$ are uniformly bounded by
1668: $2R$ and $|\theta(s_k)|=1$ there exists a subsequence $k\to\infty$ and
1669: $\tilde{t}$,
1670: $t\in\R,\theta\in\R^2$
1671: with $\tilde{t}\leq t$, $|\theta|=1$ such that
1672: \begin{gather*}
1673: t_k \,\to\, t,\quad \tilde{t}_k\,\to\, \tilde{t},\quad\text{and} \quad
1674: \theta(s_k)\,\to\, \theta.
1675: \end{gather*}
1676: We deduce from \eqref{eq:t_n}-\eqref{eq:theta_n} that
1677: \begin{eqnarray}\label{eq:t}
1678: \limsup_{k\to\infty} L^+(s_k) &\leq& t,\\\label{eq:tilde-t}
1679: \limsup_{k\to\infty}\, l^+(s_k) &\leq& \tilde{t},\\\label{eq:theta}
1680: \phi(\gamma(s)+t\theta) &=& \phi(\gamma(s)) + t.
1681: \end{eqnarray}
1682: Since $\phi$ has Lipschitz constant one it follows from
1683: \eqref{eq:theta} that
1684: \begin{gather}
1685: \theta \,=\, \nabla\phi(\gamma(s))\,=\,\theta(s).\label{eq:equal-theta}
1686: \end{gather}
1687: We therefore deduce by \eqref{eq:theta} that $t\leq
1688: L^+(s)$ and by \eqref{eq:t} we find that
1689: \begin{eqnarray*}
1690: \limsup_{k\to\infty} L^+(s_k) &\leq& L^+(s),
1691: \end{eqnarray*}
1692: which proves the upper-semicontinuity and therefore the measurability
1693: of $L^+$. By analogous arguments one proves that $L^-$
1694: is lower-semicontinuous and measurable.\\
1695: Let now $0<\tau<\tilde{\tau}$. Since $\tilde{t}_k\to\tilde{t}$ we
1696: obtain from \eqref{eq:tilde-t_n} that
1697: \begin{gather*}
1698: \phi\big(\gamma(s_{k_i}) +\tau\theta(s_{k_i})\big) \,\in\, \spt
1699: (u_\eps)\quad\text{ for all sufficiently large }k\in\N.
1700: \end{gather*}
1701: Since $\spt(u_\eps)$ is closed and since $\phi$ is continuous this yields
1702: \begin{eqnarray*}
1703: \phi\big(\gamma(s)+\tau\theta(s)\big) &\in& \spt(u).
1704: \end{eqnarray*}
1705: But $0<\tau<\tilde{\tau}$ was arbitrary and we deduce that
1706: $\tilde{t}\leq l^+(s)$.
1707: By \eqref{eq:tilde-t} this shows the upper-semicontinuity
1708: and measurability of $l^+$.
1709: \end{proof}
1710:
1711: \begin{figure}[ht]
1712: \centerline{\psfig{width=\textwidth,figure=para2}}%width=\textwidth
1713: \caption{A situation where a ray crosses several times the boundary of
1714: $\spt(u_\eps)$, given by curves $\gamma_1,\gamma_2,\gamma_3$. On the
1715: right for each of the curves $\gamma_i$ the positive, negative
1716: and effective ray lengths $L^+_i,L^-_i,l^+_i$ is depicted. Observe that
1717: $l^+_2(s)=0$.}
1718: \label{fig:para2}
1719: \end{figure}
1720:
1721: The next result is similar to Lemma~16 of \cite{CFC}:
1722: since rays can only intersect in a common endpoint, the ray directions vary
1723: Lipschitz continuously.
1724:
1725: \begin{proposition}\label{prop:lip-l>c}
1726: Let $\gamma\in \Gamma_\eps$ be given as in Remark
1727: \ref{rem:arclength-curves} and set for
1728: $\delta>0$
1729: \begin{gather}
1730: E_\delta \,:=\, \{s\in E :
1731: L^+(s),|L^-(s)|\geq\delta\}. \label{eq:def-E-delta}
1732: \end{gather}
1733: Then $\theta$ is Lipschitz continuous on $E_\delta$ with
1734: \begin{eqnarray}
1735: \big|\theta(s_1)-\theta(s_2)\big|
1736: &\leq& \frac{2}{\delta}\, |s_1-s_2|\quad\text{ for all }s_1,s_2\in
1737: E_\delta. \label{eq:lipconst-theta}
1738: \end{eqnarray}
1739: \end{proposition}
1740: \begin{proof}
1741: To prove \eqref{eq:lipconst-theta} it is sufficient
1742: to consider the case $\theta(s_1)\neq\theta(s_2)$. Then there exist
1743: $t_1,t_2\in\R$ such that
1744: \begin{eqnarray}
1745: \gamma(s_1)+t_1 \theta(s_1) &=& \gamma(s_2) +t_2\theta(s_2).
1746: \label{eq:triangle}
1747: \end{eqnarray}
1748: By the definition of $L^+,L^-$ and since rays can only intersect in
1749: common endpoints we obtain
1750: \begin{gather}
1751: t_i\,\in\, \R\setminus\big(L^-(s_i),L^+(s_i)\big),\quad
1752: i=1,2. \label{eq:bounds-triangle}
1753: \end{gather}
1754: By taking the determinant with $\theta(s_2)$ in
1755: \eqref{eq:triangle} we compute
1756: \begin{gather}
1757: \det\big(\gamma(s_1)-\gamma(s_2),\theta(s_2)\big)\,=\,
1758: -t_1\det\big(\theta(s_1),\theta(s_2)\big).\label{eq:eq-det}
1759: \end{gather}
1760: From \eqref{eq:bounds-triangle}
1761: and since $s_1,s_2\in E_\delta$ we deduce that
1762: %\begin{gather*}
1763: $t_1^2\,\geq\,\delta^2$.
1764: %\end{gather*}
1765: Taking squares in \eqref{eq:eq-det} we therefore obtain
1766: \begin{align}
1767: (s_1-s_2)^2 \,&\geq\,
1768: \delta^2\det\big(\theta(s_1),\theta(s_2)\big)^2\notag\\
1769: &=\, \delta^2\Big[ 1-(\theta(s_1)\cdot\theta(s_2))^2\Big] \notag\\
1770: &=\, \frac{\delta^2}{4}\big|\theta(s_1)-\theta(s_2)\big|^2
1771: \big|\theta(s_1)+\theta(s_2)\big|^2. \label{eq:est-rays-1}
1772: \end{align}
1773: Next we observe that
1774: \begin{gather}
1775: \big|\theta(s_1)+\theta(s_2)\big|\,\geq\, 1 \quad\text{ if }
1776: |s_1-s_2|\leq \delta.\label{eq:small-sdiff}
1777: \end{gather}
1778: Indeed we compute
1779: \begin{align*}
1780: 4\delta\,&=\, \phi(\gamma(s_1)+\delta \theta(s_1)) -
1781: \phi(\gamma(s_1)-\delta\theta(s_1)) + \phi(\gamma(s_2)+\delta
1782: \theta(s_2)) - \phi(\gamma(s_2)-\delta\theta(s_2))\\
1783: &=\, \phi(\gamma(s_1)+\delta \theta(s_1))-
1784: \phi(\gamma(s_2)-\delta\theta(s_2))+ \phi(\gamma(s_2)+\delta
1785: \theta(s_2))-\phi(\gamma(s_1)-\delta\theta(s_1))\\
1786: &\leq\, 2|\gamma(s_1)-\gamma(s_2)| +2\delta|\theta(s_1)+\theta(s_2)|\\
1787: &\leq\, 2|s_1-s_2| +2\delta|\theta(s_1)+\theta(s_2)|
1788: \end{align*}
1789: which proves \eqref{eq:small-sdiff}.\\
1790: By \eqref{eq:est-rays-1}, \eqref{eq:small-sdiff} we deduce that
1791: \eqref{eq:lipconst-theta} holds for
1792: $s_1,s_2\in E_\delta$
1793: with $|s_1-s_2|\leq \delta$. In the case $|s_1-s_2|>\delta$
1794: we obtain \eqref{eq:lipconst-theta} since $|\theta(s_1)-\theta(s_2)|\leq
1795: 2$.
1796: \end{proof}
1797: \begin{definition}\label{def:alpha}
1798: Define two functions $\alpha,\beta: E\,\to\, \R\modulo 2\pi$ by
1799: requiring that
1800: \begin{gather}
1801: \theta(s) \,=\,
1802: \begin{pmatrix}
1803: \cos \alpha(s)\\
1804: \sin \alpha(s)
1805: \end{pmatrix},
1806: \qquad\det\big(\gamma^\prime(s),\theta(s)\big) \,=\, \sin\beta(s).\label{eq:def-beta}
1807: \end{gather}
1808: \end{definition}
1809: We note that
1810: \begin{gather}
1811: \sin\beta(s)\,\geq\, 0\quad\text{ for }s\in E\text{ with }l^+(s)>0
1812: \label{eq:sin-beta}
1813: \end{gather}
1814: by \eqref{eq:spt-ue-lhs} and the definition of $l^+$.
1815:
1816: Proposition \ref{prop:lip-l>c} yields the Lipschitz continuity of
1817: $\theta,\alpha$ on sets where the positive and negative ray lengths are
1818: strictly positive. By Kirszbraun's Theorem and Rademacher's Theorem we
1819: obtain an extension which is differentiable almost everywhere. To
1820: identify these derivatives as a property of $\alpha$ we prove
1821: the existence of $\alpha^\prime$ in the sense of \emph{approximate
1822: derivatives}~\cite[section 6.1.3]{EG}.
1823:
1824: \begin{lemma}\label{lem:alpha-prime}
1825: The function $\alpha: E\to \R\modulo 2\pi$ as defined in Definition
1826: \ref{def:alpha} is almost everywhere approximately differentiable and
1827: its approximate differential
1828: satisfies
1829: \begin{eqnarray}
1830: \sin \beta(s) - t \alpha^\prime(s) &\geq& 0\quad\text{ for all }\quad
1831: L^-(s)\,\leq t\, \leq\, L^+(s)
1832: \label{eq:det>0}
1833: \end{eqnarray}
1834: for almost all $s\in E$ with $l^+(s)>0$.
1835: \end{lemma}
1836: \begin{proof}
1837: Consider for $\delta>0$ the set $E_\delta$ defined in \eqref{eq:def-E-delta}.
1838: By Proposition \ref{prop:lip-l>c} the restriction of $\theta$ to
1839: $E_\delta$ is Lipschitz continuous.
1840: From Kirszbraun's Theorem we deduce the existence of a Lipschitz
1841: continuous extension $\tilde{\alpha}_\delta:\R\to\R\modulo 2\pi$ of $\alpha$, see for example the construction in \cite{Gan}.
1842: This extension is almost everywhere differentiable by Rademacher's Theorem.
1843: By \cite[Corollary 1.7.3]{EG} almost all points of $E_\delta$ have
1844: density one in $E_\delta$ and
1845: we deduce that $\alpha$ is approximately differentiable in these points
1846: and that the approximate differential coincides with
1847: $\tilde{\alpha}_\delta^\prime$.
1848: Since \mbox{$E=\cup_{k\in\N}E_{1/k}$} this proves that $\alpha$ is
1849: approximately differentiable almost everywhere in $E$.
1850:
1851: Since a measurable function is almost everywhere approximately continuous
1852: by \cite[Theorem 1.7.3]{EG}, it is sufficient to prove \eqref{eq:det>0}
1853: for $s\in E$ with $l^+(s)>0$ such that
1854: \begin{itemize}
1855: \item $\alpha$ is approximately differentiable in $s$, and
1856: \item there exists a sequence
1857: $(s_k)_{k\in\N}\subset E$ with $s_k\to s$ as $k\to\infty$ and
1858: \begin{gather*}
1859: L^+(s_k)\to L^+(s),\quad L^-(s_k)\to L^-(s),\quad l^+(s_k)\to l^+(s),\\
1860: L^+(s_k),\, -L^-(s_k),\, l^+(s_k)\,>\,0\quad\text{ for all }k\in \N.
1861: \end{gather*}
1862: \end{itemize}
1863: Since \eqref{eq:det>0} is satisfied for $\alpha'(s)=0$ we can
1864: also assume that
1865: \begin{itemize}
1866: \item $\alpha'(s)>0$ (the other case is analogous);
1867: \item $(\alpha(s_k)-\alpha(s))/(s_k-s)\to \alpha'(s)$, or equivalently, that
1868: $(\theta(s_k)-\theta(s))/(s_k-s)\to\theta'(s)$;
1869: \item for all $k\in\N$ there exist $t_k,t^*_k$ such that
1870: \begin{gather}
1871: \label{eq:gamma-theta}
1872: \gamma(s_k)+t_k\theta(s_k) \,=\, \gamma(s)+t^*_k\theta(s).
1873: \end{gather}
1874: \end{itemize}
1875: Analogously to \eqref{eq:bounds-triangle}, \eqref{eq:eq-det} we find
1876: from~\pref{eq:gamma-theta}
1877: \begin{gather}
1878: t_k\,\in\, \R\setminus\big(L^-(s_k),L^+(s_k)\big),\quad
1879: t^*_k\,\in\, \R\setminus\big(L^-(s),L^+(s)\big),
1880: \label{eq:length-t-k}\\
1881: \det\big(\gamma(s_k)-\gamma(s),\theta(s)\big)\,=\,
1882: -t_k\det\big(\theta(s_k),\theta(s)\big).\label{eq:eq-det-k}
1883: \end{gather}
1884: Writing equation~\pref{eq:eq-det-k} as
1885: \[
1886: \det\Big(\frac{\gamma(s_k)-\gamma(s)}{s_k-s},\theta(s)\Big)\,=\,
1887: -t_k\det\Big(\frac{\theta(s_k)-\theta(s)}{s_k-s},\theta(s)\Big),
1888: \]
1889: we observe that the left-hand side converges to
1890: $\det(\gamma'(s),\theta(s)) = \sin \beta(s)\geq0$
1891: and the determinant on the right-hand side to $\det(\theta'(s),\theta(s)) = -\alpha'(s)<0$.
1892: For sufficiently large $k$ we
1893: can therefore eliminate the negative range for $t_k$ in~\pref{eq:length-t-k}
1894: and obtain
1895: \[
1896: t_k\,\geq\, L^+(s_k).
1897: \]
1898: Again using the negative sign of $\det\bigl((\theta(s_k)-\theta(s))/(s_k-s),\theta(s)\bigr)$ we then find
1899: \begin{gather*}
1900: \det\Big(\frac{\gamma(s_k)-\gamma(s)}{s_k-s},\theta(s)\Big)\,\geq\,
1901: -L^+(s_k)\det\Big(\frac{\theta(s_k)-\theta(s)}{s_k-s},\theta(s)\Big)
1902: \end{gather*}
1903: Taking the limit $k\to\infty$ we
1904: deduce
1905: \begin{gather*}
1906: \sin\beta(s) \,\geq\,
1907: L^+(s)\,\alpha'(s).
1908: \end{gather*}
1909: This proves the Lemma.
1910: \end{proof}
1911:
1912: For each $\gamma\in\Gamma_\eps$ we define analogously to
1913: \eqref{eq:def-psi} a map $\psi$. Restricting these maps
1914: suitably we obtain a parametrization of $\spt(u_\eps)$ which is
1915: essentially
1916: one-to-one.
1917: \begin{proposition}\label{prop:parametrization}
1918: Let $\Gamma_\eps=\{\gamma_i:i=1,...,N\}$ be chosen as in Remark
1919: \ref{rem:arclength-curves}. For all $i=1,...,N$ let $L_i=L(\gamma_i)$.
1920: Then, with
1921: \begin{eqnarray}\label{eq:def-E^i}
1922: D_i &:=& \big\{(s,t): s\in E_i\cap [0,L_i), 0\leq t<l_i^+(s)\},
1923: \end{eqnarray}
1924: the restrictions $\psi_i: D_i\to\R^2$ give, up to a Lebesgue
1925: nullset, an injective map onto $\spt(u_\eps)$:
1926: for almost all $x\in \spt(u_\eps)$ there exists a unique
1927: $i\in\{1,...,N\}$ and a unique
1928: $(s,t)\in D_i$ such that $\psi_i(s,t)=x$.
1929: \end{proposition}
1930: \begin{proof}
1931: For both the injectivity and the surjectivity it
1932: is sufficient to consider only interior points of
1933: $\spt(u_\eps)$ that are not ray ends, since the boundary of
1934: $\spt(u_\eps)$ and the sets
1935: of ray ends form an $\LL^2$-nullset (Proposition \ref{prop:cfm-rays}).
1936:
1937: For the surjectivity, let $x$ be such a point in $\spt(u_\eps)$. By
1938: \eqref{eq:add-prop-phi1} there exists $y\in \spt(v)$ such
1939: that
1940: \begin{eqnarray*}
1941: \phi(x)-\phi(y) &=& |x-y|
1942: \end{eqnarray*}
1943: and it follows that $x$ is on the interior of a ray with direction
1944: $\nabla\phi(x)=\frac{x-y}{|x-y|}$.
1945: Define
1946: \begin{eqnarray*}
1947: t &:=& \min\ \{t>0: x-t\nabla\phi(x)\in\partial\spt(u_\eps)\}.
1948: \end{eqnarray*}
1949: Then there exists $i\in\{1,...,N\}$ and $s\in E_i$ such that $l^+(s)\geq
1950: t>0$ and
1951: \[
1952: x \,=\,\gamma_i(s)+t\theta_i(s) \,=\, \psi_i(s,t).
1953: \]
1954:
1955: To prove the injectivity part, let $x$ be the same point again, and
1956: assume that there exists $j\in\{1,...,N\}$,
1957: $(\sigma,\tau)\in D_j$ such that
1958: \begin{gather*}
1959: x \,=\, \psi_i(s,t)\,=\,\psi_j(\sigma,\tau).
1960: \end{gather*}
1961: Since different rays cannot intersect in interior points such as $x$,
1962: the three points $x$, $\gamma_i(s)$, and
1963: $\gamma_j(\sigma)$ are on the same ray, and since $t>0$ and $\tau>0$
1964: we have
1965: \begin{gather*}
1966: \phi(\gamma_i(s))\, <\, \phi(x) \qquad\text{and}\qquad\phi(\gamma_j(\sigma)) \,<\, \phi(x).
1967: \end{gather*}
1968: This implies that $\gamma_i(s)$ and $\gamma_j(\sigma)$ are on the same side of the ray with respect to $x$.
1969: On the other hand by the definition of $l^+$ and since $t<l^+(s)$,
1970: $\tau<l^+(\sigma)$,
1971: neither the part of the ray between $x$ and $\gamma_i(s)$
1972: nor the part between $x$ and $\gamma_j(\sigma)$ contains a point of
1973: $\partial\spt(u_\eps)$. This implies that $\gamma_i(s)=\gamma_j(\sigma)$
1974: which proves the injectivity part of the proposition.
1975: \end{proof}
1976:
1977: Using Proposition \ref{prop:lip-l>c} and Proposition \ref{prop:parametrization}
1978: we can justify the following transformation formula for integrals.
1979:
1980: \begin{lemma}\label{lem:trafo-formula}
1981: Let $\Gamma_\eps=\{\gamma_i\}_{i=1,...,N}$ be as in Remark
1982: \ref{rem:arclength-curves} and $D_i$ as defined in \eqref{eq:def-E^i}.
1983: Then for all \mbox{$g\in L^1(\R^2)$}
1984: \begin{gather}\label{eq:trafo}
1985: \int g(x)u_\eps(x)\,dx =
1986: \sum_{i=1}^N \int_{D_i} g(\psi_i(s,t))\frac{1}{\eps}
1987: \big(\sin\beta_i(s)-t\alphaprime_i(s)\big)\,dt\,ds
1988: \end{gather}
1989: holds.
1990: \end{lemma}
1991: \begin{proof}
1992: We deduce from Proposition \ref{prop:lip-l>c} that for all $i\in
1993: \{1,...,N\}$ the functions $\theta_i$ and $\psi_i$ are approximately
1994: differentiable on $D_i$ with approximate differentials
1995: \begin{align*}
1996: \theta_i^\prime(s)\,&=\, \alphaprime_i(s)
1997: \begin{pmatrix}
1998: -\sin \alpha_i(s)\\ \cos \alpha_i(s)
1999: \end{pmatrix},\\
2000: \partial_s\psi_i(s,t)\,&=\,
2001: \gamma_i^{\prime}(s)+t\theta_i^\prime(s),\qquad
2002: \partial_t\psi_i(s,t)\,=\, \theta_i^\prime(s).
2003: \end{align*}
2004: This yields that
2005: \begin{align*}
2006: |\det D\psi_i(s,t)|
2007: \,=\,& \big|\det \big(\gamma_i^\prime(s),\theta_i(s)\big)+
2008: t \det\big(\theta_i^\prime(s),\theta_i(s)\big)\big|\\
2009: =\,& | \sin\beta_i(s) - t\alphaprime_i(s)|\\
2010: =\,& \sin\beta_i(s) -
2011: t\alphaprime_i(s)
2012: \end{align*}
2013: where we have used \eqref{eq:det>0} in the last equality.\\
2014: We then deduce from the generalized transformation formula~\cite[Remark 5.5.2]{AGS} that
2015: \begin{gather*}
2016: \int_{D_i} g(\psi_i(s,t))\frac{1}{\eps}
2017: \big(\sin\beta_i(s)-t\alphaprime_i(s)\big)\,dt\,ds\,=\,
2018: \int_{\psi_i(D_i)} g(x)u_\eps(x)\,dx.
2019: \end{gather*}
2020: Summing these equalities over $i=1,...,N$ we deduce by Proposition
2021: \ref{prop:parametrization} that \eqref{eq:trafo} holds.
2022: \end{proof}
2023: Often it is more convenient to work not in {\itshape length coordinates} but rather in
2024: {\itshape mass coordinates} which are defined as follows.
2025:
2026: \begin{definition}\label{def:mass-coordinates}
2027: For $\gamma\in\Gamma_\eps$ and $s\in \R$ we define a map
2028: $\mathfrak{m}_s:\R\to \R$ and a map $M: \R\to\R$ by
2029: \begin{eqnarray}\label{eq:def-mass}
2030: \mathfrak{m}(s,t) &:=&
2031: \begin{cases}
2032: \frac{t}{\eps}\sin\beta(s) -\frac{t^2}{2\eps}\alphaprime(s)&\text{ if }l^+(s)>0,\\
2033: 0 &\text{ otherwise.}
2034: \end{cases}\\
2035: \label{eq:def-Ms}
2036: M(s) &:=& \mathfrak{m}(s,l^+(s)).
2037: \end{eqnarray}
2038: \end{definition}
2039:
2040: \begin{figure}[ht]
2041: \centerline{\includegraphics[height=6cm]{masscoord2}}
2042: \caption{The function $\mathfrak m$. Two rays delimit a section of $u$-mass (below the curve $\gamma$, in this figure) that is transported to the associated section of $v$-mass. For given $t>0$, $\mathfrak m(s,t)\,ds$ is the amount of $u$-mass contained between the two rays and stretching from $\gamma(s)$ to the point $\gamma(s)+t\theta(s)$. For $t<0$, $\mathfrak m(t,s)$ is the corresponding amount of $v$-mass, with a minus sign.}
2043: \end{figure}
2044:
2045: \begin{lemma}\label{lem-mass}
2046: The map $\mathfrak{m}(s,\cdot)$ is strictly monotone on $\big(L^-(s),L^+(s)\big)$ with
2047: inverse
2048: \begin{eqnarray}\label{eq:def-t_s}
2049: \te(s,m) &=& \frac{\sin\beta(s)}{\alphaprime(s)}
2050: \Big[1- \Big(1-\frac{2\alphaprime(s)\eps}
2051: {\sin^2\beta(s)}m\Big)^\frac{1}{2}\Big].
2052: \end{eqnarray}
2053: Let $\Gamma_\eps=\{\gamma_i\}_{i=1,...,N}$ be given as in Remark
2054: \ref{rem:arclength-curves}, set $L_i=L(\gamma_i)$, and let $M_i$ be the
2055: mass function~\pref{eq:def-Ms} for curve $\gamma_i$. We obtain for
2056: $g\in L^1(\R^2)$
2057: that
2058: \begin{gather}\label{eq:trafo2}
2059: \int g(x)u_\eps(x)\,dx =
2060: \sum_{i=1}^N \int_0^{L_i}\int_0^{M_i(s)}
2061: g(\psi_i(s,\te_i(s,m)))\,dm\,ds.
2062: \end{gather}
2063: In particular, the total mass of $u_\eps$ is given by
2064: \begin{eqnarray}
2065: \int u_\eps(x)\,dx &=&
2066: \sum_{i=1}^N \int_0^{L_i} M_i(s)\,ds.\label{eq:tot-mass}
2067: \end{eqnarray}
2068: \end{lemma}
2069:
2070: \begin{proof}
2071: The calculation of the inverse is straightforward. For the transformation formula~\pref{eq:trafo2} we combine~\pref{eq:trafo} with the remark that
2072: \[
2073: \frac{\partial}{\partial t}\,\mathfrak m(s,t)= \frac{1}{\eps}
2074: \big(\sin\beta_i(s)-t\alphaprime_i(s)\big).
2075: \]
2076: Note that the integration interval $0<t<l^+_i(s)$ in the definition~\pref{eq:def-E^i} of $D_i$ transforms to $0<m<M_i(s)$.
2077: \end{proof}
2078:
2079: Typically, the optimal transport map is built by gluing solutions of
2080: one-dimensional transport problems on single rays together.
2081: We observe a similar structure here.
2082:
2083: \begin{proposition}\label{prop:mass-balance}
2084: Let $\gamma\in \Gamma_\eps$ be as in Remark \ref{rem:arclength-curves}
2085: and $S$ be
2086: the optimal transport map from $u_\eps$ to $v_\eps$, as in Proposition
2087: \ref{prop:add-prop}. Define
2088: for $s\in E$ with $l^+(s)>0$
2089: an interval $I(s)\subset\R$ and
2090: measures $f_s^+,f_s^-$ on $I(s)$,
2091: \begin{align*}
2092: I(s) &:=
2093: \Big(\mathfrak{m}_s\big(L^-(s)\big),
2094: \mathfrak{m}_s\big(L^+(s)\big)\Big),\\
2095: df_s^+ &:= u_\eps\big(\psi(s,\te(s,m))\big)\,dm,\qquad
2096: df_s^- := v_\eps\big(\psi(s,\te(s,m))\big)\,dm.
2097: \end{align*}
2098: Moreover, define a map
2099: \begin{eqnarray*}
2100: \hat{S}:I(s) \cap\spt(f_s^+) &\to& I(s)\cap\spt(f_s^-),
2101: \end{eqnarray*}
2102: by the equation
2103: %(see Section~\ref{sec:mtp} for the optimal transport map $S$),
2104: \begin{gather}\label{eq:def-s-hat}
2105: S\big(\psi(s,\te(s,m))\big) \,=\,
2106: \psi\big(s,\te(s,\hat{S}(m))\big)\quad\text{ for }m\in I\cap\spt(f_s^+).
2107: \end{gather}
2108: Then for almost all $s\in E$ with $l^+(s)>0$
2109: the map $\hat{S}$
2110: is the unique monotone transport map pushing $f_s^+$ forward to $f_s^-$.
2111: \end{proposition}
2112:
2113: \begin{proof}
2114: It is sufficient to prove the claim for almost all $s\in \{l^+>k^{-1}\}$,
2115: $k\in\N$ arbitrary.
2116: Let $A\subset E$ be any set such that
2117: \begin{gather}
2118: A\,\subset\, \{s\in E: l^+(s)>k^{-1}\},\quad \diam(A)\,<\,k^{-1}\label{eq:setA}
2119: \end{gather}
2120: We define rays and a set of rays
2121: \begin{gather*}
2122: \mathcal{R}(s)\,:=\,\{\gamma(s)+ \tau \theta(s) : \tau \in (L^-(s),
2123: L^+(s))\},\\
2124: \mathcal{R}(A)\,:=\, \bigcup_{s\in A} \mathcal{R}(s).
2125: \end{gather*}
2126: Since different rays can only intersect in a common endpoint, two rays
2127: $\mathcal{R}(s), \mathcal{R}(\sigma)$ with $s,\sigma\in A$, $s\neq
2128: \sigma$ can only be disjoint or identical. The latter case is excluded
2129: by \eqref{eq:setA}: since $l^+(s),l^+(\sigma)>k^{-1}$ the distance between
2130: $\gamma(s)$ and $\gamma(\sigma)$ has to be at least $k^{-1}$ in
2131: contradiction to the assumption on the diameter of $A$ in \eqref{eq:setA}.
2132: Therefore the union $\{ \mathcal{R}(s):s\in A\}$ is
2133: pairwise disjoint.
2134:
2135: Let now $(g_j)_{j\in\N}$ be a dense subset of $C^0(\R)$. Since $S$ pushes
2136: $u_\eps|_{\mathcal{R}(A)}$
2137: forward to $v_\eps|_{\mathcal{R}(A)}$ we obtain for all $j\in\N$
2138: \begin{eqnarray}\label{eq:loc-trans}
2139: \int_{\mathcal{R}(A)}g_j(\phi(S(x)))u_\eps(x)\,dx &=&
2140: \int_{\mathcal{R}(A)}g_j(\phi(x)) v_\eps(x)\, dx.
2141: \end{eqnarray}
2142: Repeating the changes of variables of Lemmas~\ref{lem:trafo-formula} and~\ref{lem-mass}, and using the fact that $\{
2143: \mathcal{R}(s):s\in A\}$ is
2144: pairwise disjoint we obtain
2145: \begin{align*}
2146: &\int_A \int_{I(s)}
2147: g_j\Big(\phi\bigl(S(\psi(s,\te(s,m)))\bigr)\Bigr)u_\eps\big(\psi(s,\te(s,m))\big)\,dm\,ds\\
2148: &\qquad =\,
2149: \int_A \int_{I(s)} g_j\Bigl(\phi\bigl(\psi(s,\te(s,m))\bigr)\Bigr)v_\eps\big(\psi(s,\te(s,m))\big)\,dm\,ds,
2150: \end{align*}
2151: and by using the definition of $\hat S$~\pref{eq:def-s-hat} and the linearity of
2152: $\phi$ along rays we write this
2153: as
2154: \begin{align*}
2155: &\int_A \int_{I(s)}
2156: g_j\big(\phi(\gamma(s)) + \te(s,\hat S(m))\bigr)u_\eps\big(\psi(s,\te(s,m))\big)\,dm\,ds\\
2157: &\qquad =\,
2158: \int_A \int_{I(s)} g_j\big(\phi(\gamma(s))
2159: +\te(s,m)\big)v_\eps\big(\psi(s,\te(s,m))\big)\,dm\,ds,
2160: \end{align*}
2161: Since $A$ as above was arbitrary we deduce that for almost all $s\in \{l^+>0\}$
2162: and all $j\in\N$
2163: \begin{eqnarray*}
2164: \int_{I(s)} g_j\big(\phi(\gamma(s))+\te(s,\hat{S}(m))\big)\,df_s^+(m) &=&
2165: \int_{I(s)} g_j\big(\phi(\gamma(s))+\te(s,m)\big)\,df_s^-(m).
2166: \end{eqnarray*}
2167: Since $(g_j)_{j\in\N}$ is dense in $C^0(\R)$ and since for fixed $s$ the
2168: function $m\mapsto
2169: \phi(\gamma(s)) + \te(s,m)$ is a homeomorphism it follows that
2170: \begin{eqnarray*}
2171: \int_{I(s)} g(\hat{S}(m))\,df_s^+(m) &=&
2172: \int_{I(s)} g(m)\,df_s^-(m)
2173: \end{eqnarray*}
2174: for all $g\in C^0(\R)$. Therefore $\hat{S}$ pushes $f_s^+$ forward to $f_s^-$.
2175: The monotonicity of $\hat{S}$ follows from the monotonicity of $S$. By \cite[Theorem 3.1]{Amb}
2176: the map $\hat{S}$ is the unique monotone transport map from $f_s^+$ to
2177: $f_s^-$.
2178: \end{proof}
2179:
2180:
2181: \begin{lemma}\label{lem:eff-mass-balance}
2182: Let $\gamma\in \Gamma_\eps$ as in Remark \ref{rem:arclength-curves}.
2183: Then for almost all $s\in E$ with $l^+(s)>0$ we
2184: obtain that
2185: \begin{gather}
2186: \big|\psi\big(s,\te(s,m)\big)-
2187: S\big(\psi\big(s,\te(s,m)\big)\big)\big|
2188: \,\geq\, \te(s,m)-\te(s,m-M(s))
2189: \label{eq:esti-t-diffs}
2190: \end{gather}
2191: for all $0<m<M(s)$.
2192: \end{lemma}
2193: \begin{proof}
2194: We let $\hat{S}$ be as in \eqref{eq:def-s-hat}.
2195: Using the monotonicity of $\te(s,.)$ we obtain
2196: \begin{align}
2197: \big|\psi\big(s,\te(s,m)\big)-
2198: S\big(\psi\big(s,\te(s,m)\big)\big)\big|
2199: &= \big|\psi\big(s,\te(s,m)\big)-
2200: \psi\big(s,\te(s,\hat{S}(m))\big)\big|\notag\\
2201: &= \te(s,m)-\te(s,\hat{S}(m)).\label{eq:t-diff}
2202: \end{align}
2203: By Proposition \ref{prop:mass-balance} the map $\hat{S}$ is the unique monotone
2204: transport map from $f_s^+$ to $f^-_s$. In particular
2205: \begin{gather}
2206: f^-_s\big(\big[\hat{S}(m),\hat{S}(M(s))\big]\big)
2207: \,=\, f^+_s\big([m,M(s)]\big)\label{eq:eq-opt-fs}
2208: \end{gather}
2209: holds. Since $f^+_s$ restricted to the set $[0,M(s)]$ coincides
2210: with the Lebesgue measure and since $0\leq f^-_s,f^+_s\leq \LL^1$ we deduce from
2211: \eqref{eq:eq-opt-fs} that
2212: \begin{gather}
2213: \hat{S}(M(s))-\hat{S}(m)\,\geq M(s)-m\label{eq:eq-opt-fs2}
2214: \end{gather}
2215: for any $0\leq m\leq M(s)$. The value $\hat{S}(M(s))$ is
2216: less or equal than zero since $\hat{S}$ is decreasing and maps onto the
2217: support of
2218: $f^-_s$. This implies by \eqref{eq:eq-opt-fs2} that \mbox{$\hat{S}(m)\leq
2219: m-M(s)$}. By the montonicity of $\te_s$ we deduce
2220: \begin{gather*}
2221: \te(s,m)-\te(s,\hat{S}(m))\,\geq\,
2222: \te(s,m)-\te(s,m-M(s)).
2223: \end{gather*}
2224: Together with \eqref{eq:t-diff} this proves \eqref{eq:esti-t-diffs}.
2225: \end{proof}
2226:
2227: By Lemma \ref{prop:lip-l>c} the ray directions vary Lipschitz
2228: continuously on sets where the positive and negative ray lengths are
2229: bounded from below. We obtain a Lipschitz bound also on sets
2230: where $M(\cdot)$ is bounded from below.
2231: \begin{lemma}\label{lem:lip-mass>c}
2232: Let $\gamma\in \Gamma_\eps$ be as in Remark \ref{rem:arclength-curves}.
2233: Then the inequality
2234: \begin{eqnarray}\label{eq:est-l-m}
2235: \sin\beta(s) \min\big( L^+(s),|L^-(s)|\big) &\geq& \frac{\eps}{2} M(s)
2236: \end{eqnarray}
2237: holds. In particular, for all $0<\kappa<1$ the function $\alpha$ is
2238: Lipschitz continuous on the set
2239: $\{s: M(s)\geq (1-\kappa)\}$ and
2240: \begin{eqnarray}\label{eq:lip-bound-m>0}
2241: |\alphaprime(s)| &\leq& \frac{2}{\eps(1-\kappa)}
2242: \end{eqnarray}
2243: holds for almost all $s\in \R$ with $M(s)\geq (1-\kappa)$.
2244: \end{lemma}
2245: \begin{proof}
2246: By \eqref{eq:det>0} we get that $\sin\beta(s)\geq 0$ for $s\in E$ with
2247: $l^+(s)>0$ and that $\alphaprime(s)=0$ if $\sin\beta(s)=0$. Since in the
2248: latter case \eqref{eq:est-l-m} holds it is sufficient to assume that
2249: $\sin\beta(s)>0$.
2250: We consider first the case $\alphaprime(s)\leq 0$.
2251: From \eqref{eq:def-mass} and \eqref{eq:def-Ms} we obtain
2252: \begin{eqnarray}
2253: M(s) &=& \frac{1}{\eps}\Big(l^+(s)\sin\beta(s)
2254: -\frac{1}{2}l^+(s)^2\alphaprime(s)\Big)\notag\\
2255: &\leq&
2256: \frac{\sin\beta(s)}{\eps}l^+(s)-\frac{\sin\beta(s)}{\eps}
2257: \cdot\frac{l^+(s)^2}{2L^-(s)}.\label{eq:eff-mass-est1}
2258: \end{eqnarray}
2259: By Proposition \ref{prop:mass-balance} the estimate
2260: \begin{gather*}
2261: M(s)\,=\, f_s^+((0,M(s))\,\leq\, f^-_s\big(I(s)\cap (L^-(s),0]\big)
2262: \end{gather*}
2263: holds and we deduce that
2264: \begin{eqnarray}
2265: M(s) &\leq&
2266: \int_{L^-(s)}^0\frac{1}{\eps}\big(\sin\beta(s)-t\alphaprime(s)\big)\,dt\notag\\
2267: &\leq& \frac{|L^-(s)|}{\eps}\sin\beta(s)\notag\\
2268: &=&
2269: -\frac{\sin\beta(s)}{\eps}L^-(s).\label{eq:eff-mass-est5}
2270: \end{eqnarray}
2271: Together with \eqref{eq:eff-mass-est1} we deduce
2272: \begin{eqnarray*}
2273: M(s) &\leq& \frac{\sin\beta(s)}{\eps}l^+(s) + \frac{1}{2M(s)}\Big(\frac{\sin\beta(s)}{\eps}l^+(s)\Big)^2.
2274: \end{eqnarray*}
2275: Therefore $\xi=\frac{\sin\beta(s)}{\eps}l^+(s)$ satisfies the inequality
2276: \begin{eqnarray*}
2277: \xi^2 +2M(s)\xi-2M(s)^2 &\geq & 0
2278: \end{eqnarray*}
2279: and we obtain that
2280: \begin{eqnarray}\label{eq:eff-mass-est2}
2281: \frac{\sin\beta(s)}{\eps}l^+(s)&\geq&(\sqrt{3}-1)M(s).
2282: \end{eqnarray}
2283: Together with \eqref{eq:eff-mass-est5} and $l^+(s)\leq L^+(s)$
2284: this proves \eqref{eq:est-l-m} in the case that $\alphaprime(s)\leq0$.\\
2285: Next we assume that $\alphaprime(s)>0$. From
2286: \eqref{eq:def-mass}, \eqref{eq:def-Ms} we observe that
2287: \begin{eqnarray}\label{eq:eff-mass-est4}
2288: M(s) &\leq& \frac{1}{\eps}L^+(s)\sin\beta(s).
2289: \end{eqnarray}
2290: To prove \eqref{eq:est-l-m} also for $L^-(s)$ let us assume---without loss
2291: of generality---that
2292: $|L^-(s)|\leq \frac{\eps}{\sin\beta(s)}M(s)$. We then deduce that
2293: \begin{gather}
2294: |L^-(s)|\,\leq\, L^+(s)\label{eq:ass-l+-}
2295: \end{gather}
2296: and compute
2297: \begin{eqnarray*}
2298: M(s) &\leq& \int_{L^-(s)}^0 v_\eps(\psi(s,t))\big(\sin\beta(s)
2299: -t\alphaprime(s)\big)\,dt\notag\\
2300: &\leq& \int_{L^-(s)}^0 \frac{1}{\eps}\big(\sin\beta(s)
2301: -L^-(s)\alphaprime(s)\big)\,dt\,=\,\frac{|L^-(s)|}{\eps}\big(\sin\beta(s)
2302: -L^-(s)\alphaprime(s)\big).
2303: \end{eqnarray*}
2304: By \eqref{eq:ass-l+-} and \eqref{eq:det>0} we deduce that
2305: \begin{gather*}
2306: M(s) \,\leq\, \frac{|L^-(s)|}{\eps}\big(\sin\beta(s)
2307: +L^+(s)\alphaprime(s)\big)\,\leq\,2|L^-(s)|\frac{\sin\beta(s)}{\eps}.
2308: \end{gather*}
2309: Together with \eqref{eq:eff-mass-est4} this proves \eqref{eq:est-l-m} in
2310: the case $\alphaprime(s)>0$.
2311:
2312: The estimate \eqref{eq:lip-bound-m>0} follows from \eqref{eq:est-l-m}
2313: and \eqref{eq:det>0}.
2314: \end{proof}
2315: %===============================================================
2316: % estimate from below
2317: %===============================================================
2318: \subsection{Estimate from below}
2319: \label{subsec:est-below}
2320: Let $S$ be the optimal transport map and $\phi\in \Lip_1(\R^2)$
2321: be an
2322: optimal Kantorovich potential for the mass transport from $u_\eps$ to
2323: $v_\eps$ as in Proposition
2324: \ref{prop:add-prop}.
2325:
2326: Using Lemma \ref{lem-mass} and Lemma \ref{lem:eff-mass-balance}
2327: we obtain for the
2328: \Wasserstein distance between $u$ to $v$ that
2329: \begin{eqnarray}
2330: d_1(u,v) &=&\int_{\R^2} |x-S(x)|u_\eps(x)\,dx\notag\\
2331: &=& \sum_{i=1}^N\int_0^{L_i}\int_0^{M_i(s)}
2332: \big|\psi_i(s,\te_i(s,m))-S
2333: \big(\psi_i(s,\te_i(s,m))\big|\,dm\,ds\notag\\
2334: &\geq& \sum_{i=1}^N\int_0^{L_i}\int_0^{M_i(s)}\bigl[
2335: \te_i(s,m) -
2336: \te_i(s,m-M_i(s))\bigr]\,dm.\label{eq:cost-1}
2337: \end{eqnarray}
2338: We further estimate the right-hand side of this inequality.
2339:
2340: \begin{lemma}\label{lem:stand-trans-est}
2341: Let $\Gamma_\eps$ be a disjoint system of curves parametrized by
2342: arclength as in Remark \ref{rem:arclength-curves} and let $\gamma\in
2343: \Gamma_\eps$.
2344: Then we obtain for all $s\in E$ with $l^+(s)>0$ and
2345: $\sin\beta(s)>0$ that
2346: \begin{align}
2347: &\int_0^{M(s)} \bigl[\te(s,m) -
2348: \te(s,m-M(s))\bigr]\,dm\notag\\
2349: =\,& \frac{\eps}{\sin\beta(s)}M(s)^2+
2350: \frac{\eps^3}{4\sin^5\beta(s)}\alphaprime(s)^2M(s)^4 +
2351: R(s)\eps^5,\label{eq:stand-trans-est}
2352: \end{align}
2353: where
2354: \begin{gather}
2355: 0\,\leq\, R(s)\,\leq\,
2356: \frac{7}{9}\frac{M(s)^6}{\sin\beta(s)^9}\alphaprime(s)^4.\label{eq:est-R}
2357: \end{gather}
2358: \end{lemma}
2359:
2360: \begin{proof}
2361: We compute for $r\in\R$
2362: \begin{eqnarray}
2363: \int_0^r \te(s,m)\,dm &=& \int_0^r
2364: \frac{\sin\beta(s)}{\alphaprime(s)}\Big[1- \Big(1-\frac{2\alphaprime(s)\eps}
2365: {\sin^2\beta(s)}m\Big)^\frac{1}{2}\Big]\, dm\notag\\
2366: &=& \frac{\sin^3\beta(s)}{3\alphaprime(s)^2\eps}\Big[\Big(
2367: 1-\frac{2\alphaprime(s)\eps}{\sin^2\beta(s)}r\Big)^{\frac{3}{2}}
2368: +\frac{3\eps\alphaprime(s)}{\sin^2\beta(s)}r -1\Big]\label{eq:cost-4}
2369: \end{eqnarray}
2370: and thus
2371: \begin{align}
2372: &\int_0^{M(s)} \bigl[\te(s,m)-
2373: \te(s,m-M(s))\bigr]\,dm\notag\\
2374: =\,&\int_0^{M(s)} \te(s,m)\, dm+ \int_0^{-M(s)}
2375: \te(s,m)\,dm\notag\\
2376: =\,&\frac{\sin^3\beta(s)}{3\alphaprime(s)^2\eps}
2377: \Bigg[\Big(
2378: 1+\frac{2\alphaprime(s)\eps}{\sin^2\beta(s)}M(s)\Big)^{\frac{3}{2}}
2379: +\Big(
2380: 1-\frac{2\alphaprime(s)\eps}{\sin^2\beta(s)}M(s)\Big)^{\frac{3}{2}}
2381: -2\,\Bigg]\label{eq:cost-5}
2382: \end{align}
2383: By a Taylor expansion we observe that for all $\xi>0$
2384: \begin{eqnarray}
2385: \label{eq:Taylor_xi}
2386: (1+\xi)^{\frac{3}{2}} + (1-\xi)^{\frac{3}{2}} -2 &=& \frac{3}{4}\xi^2
2387: +\frac{3}{64}\xi^4 + \frac{7}{9}2^{-6}\zeta(\xi)^6
2388: \end{eqnarray}
2389: with $0\leq \zeta(\xi)\leq \xi$.
2390: Using this expansion in \eqref{eq:cost-5} we obtain
2391: \eqref{eq:stand-trans-est}, \eqref{eq:est-R}.
2392: \end{proof}
2393:
2394: Putting all information together we derive the following
2395: estimate.
2396:
2397: \begin{proposition}\label{prop:lower-bound}
2398: Let $\partial\spt(u_\eps)$ be given by a disjoint system $\Gamma_\eps$
2399: of closed curves parametrized by arclength as in Remark
2400: \ref{rem:arclength-curves}. Then the lower bound
2401: \begin{eqnarray}\label{eq:lower-bound}
2402: \G_\eps(u_\eps,v_\eps) &\geq& \sum_{i=1}^N\int_0^{L_i}
2403: \left[\frac{1}{\eps^2}\Big(\frac{1}{\sin\beta_i(s)}-1\Big)M_i(s)^2
2404: +\frac{1}{\eps^2}\big(M_i(s)-1\big)^2\right]\,ds \notag\\
2405: &&+\sum_{i=1}^N\int_0^{L_i}\frac{1}{4\sin\beta_i(s)}
2406: \Big(\frac{M_i(s)}{\sin\beta_i(s)}\Big)^4\alphaprime_i(s)^2\,ds
2407: \end{eqnarray}
2408: holds.
2409: \end{proposition}
2410:
2411: \begin{proof}
2412: Since $L_i=L(\gamma_i)$ and by \eqref{eq:tot-mass} we obtain that
2413: \begin{gather*}
2414: \eps\int_{\R^2}|\nabla u_\eps| \,=\, \sum_{i=1}^N L_i\qquad\text{and}\qquad
2415: 1\,=\, \int_{\R^2} u_\eps \,=\, \sum_{i=1}^N\int_0^{L_i} M_i(s)\,ds.
2416: \end{gather*}
2417: From \eqref{eq:cost-1} and Lemma \ref{lem:stand-trans-est} we therefore obtain
2418: \begin{align*}
2419: \G_\eps(u_\eps,v_\eps)
2420: \;\geq\;& \sum_{i=1}^N\int_0^{L_i}\Big(\frac{1}{\eps^2\sin\beta_i(s)}M_i(s)^2
2421: -\frac{2}{\eps^2}M_i(s)
2422: +\frac{1}{\eps^2}\Big) \\
2423: & +\sum_{i=1}^N\int_0^{L_i}
2424: \frac{1}{4\sin^5\beta_i(s)}
2425: \alphaprime_i(s)^2M_i(s)^4,
2426: \end{align*}
2427: and observing that
2428: \begin{align*}
2429: \frac{1}{\sin\beta_i(s)}M_i(s)^2 -2M_i(s) + 1 \,=\,
2430: \Big(\frac{1}{\sin\beta_i(s)}-1\Big)M_i(s)^2
2431: +\big(M_i(s)-1\big)^2
2432: \end{align*}
2433: the estimate \eqref{eq:lower-bound} follows.
2434: \end{proof}
2435: %===============================================================
2436: % Convergence and lower-semicontinuity
2437: %===============================================================
2438: \subsection{Compactness of the boundary curves and proof of the lim-inf
2439: estimate}
2440: \label{subsec:conv}
2441: In this subsection we prove that the boundaries of $\spt(u_\eps)$
2442: converge
2443: to a system of closed curves which satisfies the
2444: lim-inf estimate. We later identify this limit with the limit of
2445: $u_\eps\LL^2$.
2446:
2447: Let $\partial\spt(u_\eps)$ be given by a system of closed curves
2448: $\Gamma_{\eps}=\{\gamma_{\eps,i}\}_{i=1,...,N(\eps)}$
2449: parametrized by arclength as in Remark \ref{rem:arclength-curves}. In
2450: particular we recall that \mbox{$L_{\eps,i}=L(\gamma_{\eps,i})$}.
2451:
2452: The total length of the boundary of $\spt(u_\eps)$ is given by
2453: $|\Gamma_{\eps}|$ and we obtain from
2454: \eqref{eq:bound-G-eps} and \eqref{eq:lower-bound}
2455: \begin{gather}
2456: |\Gamma_{\eps}|\,=\,\sum_{i=1}^{N(\eps)} L_{\eps,i}\,=\, \eps\int_{\R^2}|\nabla
2457: u_\eps|\,\leq\, 2+\eps^2 \G_\eps(u_\eps,v_\eps)\,\leq\, 2+\eps^2\Lambda.
2458: \label{eq:bound-boundaries}
2459: \end{gather}
2460: This implies the convergence of the measures $\eps |\nabla u_\eps|
2461: =\mu_{\Gamma_\eps}$ as Radon measures on $\R^2$ for a subsequence
2462: $\eps\to 0$.
2463:
2464: We first define modified curves $\tilde{\gamma}_{\eps,i}$ which are
2465: uniformly bounded in
2466: $W^{2,2}\big(0,L_{\eps,i}\big)$.
2467: % as we will prove in Lemma\ref{lem:mod-curves}.
2468:
2469: \begin{definition}\label{def:mod-curves}
2470: Let $\gamma\in \Gamma_\eps$ and $0<\lambda<1$.
2471: Choose a periodic Lipschitz continuous function
2472: $\tilde{\alpha}_{\eps,i}:\R\to\R\modulo 2\pi$ with period $L_{\eps,i}$ such that
2473: \begin{align}
2474: \tilde{\alpha}_{\eps,i}\,&=\,\alpha_{\eps,i}\quad\text{ on }\quad
2475: \{s\in \R\,:\, M_{\eps,i}(s)\geq 1-\lambda\},\label{eq:ext-equal}\\
2476: |\tilde{\alpha}_{\eps,i}^\prime| \,&\leq\,
2477: \frac{2}{\eps(1-\lambda)}\label{eq:ext-curv-bound}
2478: \end{align}
2479: and set
2480: \begin{gather*}
2481: \tilde{\theta}_{\eps,i} \,=\, \begin{pmatrix}
2482: \cos \alpha_{\eps,i}\\
2483: \sin \alpha_{\eps,i}
2484: \end{pmatrix},
2485: \qquad
2486: \tilde{\theta}_{\eps,i}^\perp \,=\, \begin{pmatrix}
2487: \sin \alpha_{\eps,i}\\
2488: -\cos \alpha_{\eps,i}
2489: \end{pmatrix}.
2490: \end{gather*}
2491: We then define $\tilde{\gamma}_{\eps,i}:\R\to\R^2$ to be the curve which satisfies
2492: \begin{eqnarray}
2493: \tilde{\gamma}_{\eps,i}(0) &=&
2494: \gamma_{\eps,i}(0),\label{eq:new-curve-0}\\
2495: \tilde{\gamma}_{\eps,i}^\prime(s) &=& \tilde{\theta}_{\eps,i}^\perp(s)
2496: \quad\text{ for all }s\in\R.\label{eq:new-curve-tan}
2497: \end{eqnarray}
2498: \end{definition}
2499: \begin{remark}
2500: By Proposition \ref{prop:lip-l>c} the
2501: function $\alpha_{\eps,i}$ is Lipschitz continuous on the set
2502: $\{M_{\eps,i}>1-\lambda\}$. By Kirszbraun's Theorem and
2503: \eqref{eq:lip-bound-m>0}
2504: an extension $\tilde{\alpha}_{\eps,i}$ as in Definition
2505: \ref{def:mod-curves} exists.
2506:
2507: The curves $\tilde{\gamma}_{\eps,i}$ are not necessarily closed (or, equivalently, the functions $\tilde{\gamma}_{\eps,i}$ are not necessarily periodic);
2508: the restriction $\tilde{\gamma}_{\eps,i}|_{[0,L_{\eps,i})}$ may also
2509: have self-intersections. However the \emph{tangents} of these
2510: curves, the functions
2511: $\tilde{\gamma}_{\eps,i}^\prime$, are periodic with period~$L_{\eps,i}$.
2512: \end{remark}
2513:
2514: \begin{lemma}\label{lem:mod-curves}
2515: Let $\eps>0$, $0<\lambda<1$, $1\leq p\leq 2$, and consider for $i\in \{1,...,N(\eps)\}$
2516: functions
2517: $\tilde{\alpha}_{\eps,i}$ and modified curves $\tilde{\gamma}_{\eps,i}$
2518: as in Definition \ref{def:mod-curves}. Then there exists a constant
2519: $C(\lambda,\Lambda)$, which is independent of $\eps$, such that
2520: \begin{align}
2521: \sum_{i=1}^{N(\eps)} \int_0^{L_{\eps,i}}
2522: |\gamma_{\eps,i}^\prime(s)-\tilde{\gamma}_{\eps,i}^\prime(s)|\, ds &\leq
2523: \, \eps C(\lambda,\Lambda)
2524: %2 \eps (1-\lambda)^{-1} (1+\eps^2\Lambda)^{1/2}\Lambda^{1/2} +\eps^2\lambda^{-2}\Lambda
2525: \label{eq:mod-curves-1}\\
2526: \sum_{i=1}^{N(\eps)} \int_0^{L_{\eps,i}}
2527: |\tilde{\gamma}_{\eps,i}^{\prime\prime}(s)|^2\,ds
2528: &\leq\, C(\lambda,\Lambda),
2529: \label{eq:mod-curves-2}\\
2530: \sum_{i=1}^{N(\eps)} \int_0^{L_{\eps,i}}
2531: |\tilde{\gamma}_{\eps,i}^{\prime\prime}(s)|^p\,ds
2532: &\leq\, \Big[\sum_{i=1}^{N(\eps)}L_{\eps,i}\Big]^{1-\frac{p}{2}}
2533: \Big[4(1-\lambda)^{-4} \G_\eps(u_\eps,v_\eps)\Big]^{\frac{p}{2}}\notag\\
2534: &\qquad{}+ 4\lambda^{-2}(1-\lambda)^{-2}\eps^{2-p}\G_\eps(u_\eps,v_\eps).
2535: \label{eq:mod-curves-3}
2536: \end{align}
2537: \end{lemma}
2538: \begin{proof}
2539: Dropping the indexes $\eps,i$ for a moment and letting
2540: $\theta(s)^\perp = (\sin \alpha(s),-\cos \alpha(s))^T$ we compute
2541: that
2542: \begin{eqnarray}
2543: \int_0^L |\gamma^\prime(s)-\tilde{\gamma}^\prime(s)|\,ds &\leq&
2544: \int_0^L \Chi_{\{M\geq 1-\lambda\}}(s) \Big|\gamma^\prime(s) -
2545: \theta(s)^\perp \Big|\, ds\notag\\
2546: &&+ 2\int_0^L \Chi_{\{M< 1-\lambda\}}(s)\,ds\label{eq:diff-gamma-prime}
2547: \end{eqnarray}
2548: and
2549: \begin{align*}
2550: & |\gamma^\prime(s) - \theta(s)^\perp|^2
2551: \,=\, 2-2\gamma^\prime(s)\cdot \theta(s)^\perp\\
2552: =\,& 2- 2\det \big(\gamma^\prime(s),\theta(s)\big)\,=\,
2553: 2\sin\beta(s)\Big(\frac{1}{\sin\beta(s)} -1\Big),
2554: \end{align*}
2555: which implies
2556: \begin{gather}
2557: \int_0^L \Chi_{\{M\geq 1-\lambda\}}(s) \Big|\gamma^\prime(s) -
2558: \theta(s)^\perp \Big|^2\, ds \leq\,
2559: 2(1-\lambda)^{-2}\int_0^L \Big(\frac{1}{\sin\beta(s)}
2560: -1\Big)M(s)^2\,ds.\label{eq:diff-g-p1}
2561: \end{gather}
2562: Moreover we calculate
2563: \begin{eqnarray}\label{eq:diff-g-p2}
2564: \int_0^L \Chi_{\{M< 1-\lambda\}}(s)\,ds &\leq&
2565: \eps^2\lambda^{-2}\int_0^L \frac{1}{\eps^2}\big(1-M(s)\big)^2\,ds.
2566: \end{eqnarray}
2567: By \eqref{eq:diff-g-p1}, \eqref{eq:diff-g-p2} we obtain from
2568: \eqref{eq:diff-gamma-prime} that
2569: \begin{eqnarray*}
2570: \int_0^L |\gamma^\prime(s)-\tilde{\gamma}^\prime(s)|\,ds &\leq&
2571: \sqrt{2L}\eps(1-\lambda)^{-1}\Big( \int_0^L \frac{1}{\eps^2}\Big(\frac{1}{\sin\beta(s)}
2572: -1\Big)M(s)^2\,ds\Big)^{1/2}\\
2573: &&+ \eps^2\lambda^{-2}\int_0^L \frac{1}{\eps^2}\big(1-M(s)\big)^2\,ds.
2574: \end{eqnarray*}
2575: Summing the corresponding inequalities for $i=1,...,N(\eps)$, and using
2576: H\"older's inequality, we deduce
2577: that
2578: \begin{align}
2579: \sum_{i=1}^{N(\eps)}&\int_0^{L_{\eps,i}}
2580: |\gamma_{\eps,i}^\prime(s)-\tilde{\gamma}_{\eps,i}^\prime(s)|\,ds\notag\\
2581: &\leq
2582: \sqrt{2}\eps(1-\lambda)^{-1}\Big(\sum_{i=1}^{N(\eps)}L_i\Big)^{1/2}
2583: \Big( \sum_{i=1}^{N(\eps)}\int_0^{L_{\eps,i}} \frac{1}{\eps^2}\Big(\frac{1}{\sin\beta_{\eps,i}(s)}
2584: -1\Big)M_{\eps,i}(s)^2\,ds\Big)^{1/2}\notag\\
2585: &\qquad+ \eps^2\lambda^{-2}\sum_{i=1}^{N(\eps)}\int_0^{L_{\eps,i}}
2586: \frac{1}{\eps^2}\big(1-M_{\eps,i}(s)\big)^2\,ds.\\
2587: &\leq
2588: \sqrt{2}\eps(1-\lambda)^{-1}\Big(\sum_{i=1}^{N(\eps)}L_i\Big)^{1/2}
2589: \Lambda^{\frac{1}{2}} + \eps^2\lambda^{-2}\Lambda,\label{eq:aux-mod-curves}
2590: \end{align}
2591: where we have used \eqref{eq:bound-G-eps} and \eqref{eq:lower-bound}. By \eqref{eq:bound-boundaries}
2592: the inequality \eqref{eq:mod-curves-1} follows from \eqref{eq:aux-mod-curves}.\\
2593: Dropping indexes $\eps,i$ again we obtain that
2594: \begin{eqnarray}
2595: \int_0^L |\tilde{\gamma}^{\prime\prime}(s)|^p\,ds
2596: &=& \int_0^L \Chi_{\{M\geq 1-\lambda\}}(s) |\alphaprime(s)|^p\, ds\notag\\
2597: &&\qquad+ \int_0^L \Chi_{\{M<
2598: 1-\lambda\}}(s)|\tilde{\alpha}^\prime(s)|^p\,ds\notag\\
2599: &\leq& L^{1-\frac{p}{2}}\Big[4(1-\lambda)^{-4}\int_0^L
2600: \frac{1}{4\sin\beta(s)}\Big(\frac{M(s)}{\sin\beta(s)}\Big)^4|\alphaprime(s)|^2\,ds\Big]^{\frac{p}{2}}
2601: \notag\\
2602: && \qquad{}+4(1-\lambda)^{-2}\frac{1}{\eps^p}\int_0^L \Chi_{\{M<
2603: 1-\lambda\}}(s)\,ds,\notag
2604: \end{eqnarray}
2605: where we have used \eqref{eq:ext-curv-bound} in the last inequality. We
2606: can further estimate the second term on the right-hand side by
2607: \eqref{eq:diff-g-p2},
2608: \begin{gather*}
2609: 4(1-\lambda)^{-2}\frac{1}{\eps^p}\int_0^L \Chi_{\{M<
2610: 1-\lambda\}}(s)\,ds\,\leq\, 4\lambda^{-2}(1-\lambda)^{-2}\eps^{2-p}\int_0^L
2611: \frac{1}{\eps^2}(1-M(s))^2\,ds.
2612: \end{gather*}
2613: Putting everything together and summing over $i=1,...,N(\eps)$ we obtain
2614: \begin{eqnarray*}
2615: \sum_{i=1}^{N(\eps)}\int_0^{L_{\eps,i}} |\tilde{\gamma}_{\eps,i}^{\prime\prime}(s)|^p\,ds
2616: &\leq& \Big[\sum_{i=1}^{N(\eps)}L_{\eps,i}\Big]^{1-\frac{p}{2}}
2617: \Big[4(1-\lambda)^{-4} \G_\eps(u_\eps,v_\eps)\Big]^{\frac{p}{2}}\\
2618: &&\qquad {}+ 4\lambda^{-2}(1-\lambda)^{-2}\eps^{2-p}\G_\eps(u_\eps,v_\eps)
2619: \end{eqnarray*}
2620: which is \eqref{eq:mod-curves-3}. Setting $p=2$
2621: and using \eqref{eq:bound-G-eps} the estimate \eqref{eq:mod-curves-2} follows.
2622: \end{proof}
2623:
2624: The next result shows that those curves $\gamma_{\eps,i}$ that
2625: are `too short' do not contribute to the limit.
2626:
2627: \begin{lemma}\label{lem:density}
2628: There exists $\delta=\delta(\Lambda)$ and $C=C(\Lambda)$ independent of
2629: $\eps$ such that
2630: for the index set
2631: \begin{eqnarray*}
2632: J(\eps) &:=& \big\{i\in \{1,...,N(\eps)\} : L_{\eps,i} \leq\delta\big\}
2633: \end{eqnarray*}
2634: the estimate
2635: \begin{eqnarray}\label{eq:density}
2636: \sum_{i\in J(\eps)} L_{i,\eps} &\leq& \eps C(\Lambda)
2637: \end{eqnarray}
2638: holds.
2639: \end{lemma}
2640:
2641: \begin{proof}
2642: Let $\eps>0$ and $i\in J(\eps)$ be fixed and consider
2643: $\tilde{\alpha}_{\eps,i}$ and $\tilde{\gamma}_{\eps,i}$ as in Definition
2644: \ref{def:mod-curves} with $\lambda=1/2$. We drop for a moment the indexes $\eps,i$.
2645: We then have
2646: \begin{eqnarray}
2647: L &=& \int_0^L |\tilde{\gamma}^\prime(s)|^2\,ds\notag\\
2648: &=& \int_0^L \tilde{\gamma}^\prime(s)\cdot \big(\tilde{\gamma}(s)-\tilde{\gamma}(0)\big)^\prime\,ds\notag\\
2649: &=& -\int_0^L \tilde{\gamma}^{\prime\prime}(s)\big(\tilde{\gamma}(s)-\tilde{\gamma}(0)\big)\,ds+
2650: \tilde{\gamma}^\prime(L)\cdot \big(\tilde{\gamma}(L)-\tilde{\gamma}(0)\big)\notag\\
2651: &\leq& \int_0^L s |\tilde{\gamma}^{\prime\prime}(s)|\,ds+
2652: \Big|\int_0^L \bigl(\tilde{\gamma}^\prime(s)-\gamma^\prime(s)\bigr)\,ds \Big|\notag\\
2653: &\leq& \frac{1}{\sqrt{3}}L^{3/2}\Big(\int_0^L |\tilde{\gamma}^{\prime\prime}(s)|^2\,ds\Big)^{1/2}
2654: + \int_0^L \big| \tilde{\gamma}^\prime(s)-\gamma^\prime(s)\big|\,ds. \label{eq:proof-dens-1}
2655: \end{eqnarray}
2656: Summing the corresponding inequalities over $i\in J(\eps)$ and using
2657: \eqref{eq:mod-curves-1} and \eqref{eq:mod-curves-2} we obtain with the
2658: H\"older inequality
2659: \begin{eqnarray*}
2660: \sum_{i\in J(\eps)} L_{\eps,i} &\leq&
2661: C(\Lambda)\Big(\sum_{i\in J(\eps)} L_{\eps,i}^3\Big)^{1/2}\Lambda +
2662: \eps C(\Lambda)\\
2663: &\leq& {\delta} \Big(\sum_{i\in J(\eps)} L_{\eps,i}\Big)C(\Lambda) +
2664: \eps C(\Lambda)\\
2665: &\leq& \frac{1}{2}\Big(\sum_{i\in J(\eps)} L_{\eps,i}\Big) +
2666: \eps C(\Lambda)
2667: \end{eqnarray*}
2668: for $\delta=\delta(\Lambda)$ small enough. This proves \eqref{eq:density}.
2669: \end{proof}
2670:
2671: \begin{proposition}\label{prop:conv-interface}
2672: There exists a
2673: subsequence $\eps\to 0$ and a $W^{2,2}$-system
2674: $\Gamma=\{c_i, i=1,...,N_1\}$ of closed curves without transversal crossings
2675: such that
2676: \begin{eqnarray}
2677: \eps|\nabla u_\eps| &\to& \mu_{\Gamma}\label{eq:conv-nabla-ue}
2678: \end{eqnarray}
2679: as measures on $\R^2$. Moreover $\Gamma$ has even \multiplicity and satisfies
2680: \begin{eqnarray}\label{eq:willmore-1}
2681: \W(\Gamma) &\leq& 2\liminf_{\eps\to 0} \G_\eps(u_\eps,v_\eps).
2682: \end{eqnarray}
2683: \end{proposition}
2684:
2685: \begin{proof}
2686: By Lemma \ref{lem:density} we can, without
2687: loss of generality, assume that $L_{\eps,i}> \delta(\Lambda)$ for
2688: all $\eps,i$. We then deduce that the number of curves $N(\eps)$ is
2689: bounded uniformly in $\eps$.
2690: Next we consider the reparametrized one-periodic functions
2691: $c_{\eps,i}:\R\to\R^2$ defined by
2692: \begin{gather*}
2693: c_{\eps,i}(r)\,:=\, \gamma_{\eps,i}(L_{\eps,i}r)\quad\text{ for
2694: }r\in\R.
2695: \end{gather*}
2696: Since the functions $c_{\eps,i}$ are Lipschitz continuous with uniformly
2697: bounded Lipschitz constant, and since $L_{\eps,i}$ are bounded from above
2698: and away from zero, there exists a subsequence
2699: $\eps\to 0$, a number $N_1\in\N$ and
2700: \begin{gather*}
2701: L_i>0,\,
2702: c_i:\R\to\R^2,\quad i=1,...,N_1,
2703: \end{gather*}
2704: such that for all $i\in\{1,...,N_1\}$ the functions $c_i$ are Lipschitz
2705: continuous and
2706: \begin{eqnarray}
2707: N(\eps) &=& N_1\quad\text{ for all }\eps>0,\notag\\
2708: L_{\eps,i} &\to& L_i\quad\text{ as }\eps\to 0,\label{eq:conv-curve-org-L}\\
2709: c_{\eps,i} &\to& c_i\quad\text{ in } C^{0,\sigma}(\R;\R^2)\quad\text{ as
2710: }\eps\to 0\label{eq:conv-curve-org}
2711: \end{eqnarray}
2712: holds for all $0<\sigma<1$.
2713: In particular, the functions $c_i$ are
2714: one-periodic and $\eps|\nabla u_\eps|$ converge as measures to
2715: $\mu_{\Gamma}$.
2716:
2717: We next turn to the $W^{2,2}$-bound on the limit curves $c_i$.
2718: We fix
2719: $0<\lambda<1$ and consider for $\gamma_{\eps,i}$ modified functions
2720: $\tilde{\gamma}_{\eps,i}$ as in Definition \ref{def:mod-curves} and the
2721: reparametrizations $\tilde{c}_{\eps,i}$,
2722: \begin{eqnarray*}
2723: \tilde{c}_{\eps,i}(r) &=&
2724: \tilde{\gamma}_{\eps,i}(L_{\eps,i}r)\quad\text{ for }r\in\R.
2725: \end{eqnarray*}
2726: By Lemma \ref{lem:mod-curves} these curves are uniformly bounded in
2727: $W^{2,2}_{\loc}(\R;\R^2)$ and we deduce that there exists a subsequence
2728: $\eps\to
2729: 0$ of the sequence in \eqref{eq:conv-curve-org} such that the curves
2730: $\tilde{c}_{\eps,i}$ converge weakly in $W^{2,2}_{\loc}(\R;\R^2)$ and
2731: strongly in
2732: $C^{1,\sigma}([-K,K];\R^2)$ for all $0 <\sigma<1/2$, $K>0$. Since
2733: $c_{\eps,i}(0)=\tilde{c}_{\eps,i}(0)$ we obtain for any $r\in \R$
2734: \begin{eqnarray*}
2735: \lim_{\eps\to 0} |c_{\eps,i}(r) -\tilde{c}_{\eps,i}(r)| &\leq&
2736: \lim_{\eps\to 0} \Big|\int_0^r |c_{\eps,i}^\prime(\varrho)-
2737: \tilde{c}_{\eps,i}^\prime(\varrho)|\,d\varrho\Big|\\
2738: &\leq& \lim_{\eps\to 0}\Big(1+\frac{|r|}{L_{\eps,i}}\Big) \int_0^{L_{\eps,i}}
2739: |\gamma_{\eps,i}^\prime(s)-
2740: \tilde{\gamma}_{\eps,i}^\prime(s)|\,ds\\
2741: &=& 0,
2742: \end{eqnarray*}
2743: where we have used \eqref{eq:mod-curves-1}, the periodicity of
2744: $c_{\eps,i}$, $\tilde{c}^\prime_{\eps,i}$ and
2745: $L_{\eps,i}\geq\delta(\Lambda)$. Therefore we can identify
2746: the limits of $c_{\eps,i},\tilde{c}_{\eps,i}$ and deduce that
2747: \begin{alignat}{3}
2748: \tilde{c}_{\eps,i} \,&\schwto\, c_i\quad&&\text{ weakly in }
2749: W^{2,2}_{\loc}(\R;\R^2)\quad&&\text{ as
2750: }\eps\to 0,\label{eq:conv-curve-mod-W22}\\
2751: \tilde{c}_{\eps,i} \,&\to\, c_i\quad&&\text{ in }
2752: C^{1,\sigma}([-K,K];\R^2)\quad&&\text{ as
2753: }\eps\to 0,\label{eq:conv-curve-mod-C1}
2754: \end{alignat}
2755: for all $0\leq\sigma<1/2$, $K>0$. In particular $c_i\in W^{2,2}_{\loc}(\R;\R^2)$.
2756: The lower-semicontinuity of the norm under weak
2757: convergence and \eqref{eq:mod-curves-3} yield that
2758: \begin{eqnarray*}
2759: \sum_{i=1}^{N_1} L_i^{1-2p}\int_0^1 c_i^{\prime\prime}(r)^p \,dr &\leq&
2760: \liminf_{\eps\to
2761: 0}\sum_{i=1}^{N_1} L_{\eps,i}^{1-2p}\int_0^1
2762: \tilde{c}_{\eps,i}^{\prime\prime}(r)^p \,dr\\
2763: &=& \liminf_{\eps\to
2764: 0}\sum_{i=1}^{N_1} \int_0^{L_{\eps,i}}
2765: \tilde{\gamma}_{\eps,i}^{\prime\prime}(s)^p \,ds\\
2766: &\leq& \Big[\sum_{i=1}^{{N_1}}L_{i}\Big]^{1-\frac{p}{2}}
2767: \Big[4(1-\lambda)^{-4} \liminf_{\eps\to
2768: 0}\G_\eps(u_\eps,v_\eps)\Big]^{\frac{p}{2}}
2769: \end{eqnarray*}
2770: holds for all
2771: $1\leq p<2$. Since $0<\lambda<1$ and $1\leq p<2$ are arbitrary we
2772: obtain \eqref{eq:willmore-1}.
2773:
2774: Since the curves $c_{\eps,i}$
2775: are pairwise disjoint and have no self-intersections, by
2776: \eqref{eq:conv-curve-org} we deduce that the curves $c_i$ have no
2777: transversal crossings.
2778: Finally we obtain for any $\eta\in C^1_c(\R^2)$ by the divergence theorem
2779: that
2780: \begin{eqnarray*}
2781: 0 &=& \lim_{\eps\to 0} \int_{\spt(u_\eps)} \nabla\cdot\eta(x)\,dx\\
2782: &=& \lim_{\eps\to 0} \sum_{i=1}^{N_1} \int_0^{1}
2783: \eta(c_{\eps,i}(s))\cdot c_{\eps,i}^\prime(s)^\perp\,ds\\
2784: &=& \sum_{i=1}^{N_1} \int_0^1 \eta(c_{i}(s))\cdot
2785: c_i^\prime(s)^\perp\,ds\\
2786: &=& \int_{\spt(\Gamma)} \eta(x)\cdot \Big[ \sum_{\{(i,s): c_i(s)=x\}}
2787: \frac{1}{L_i}c_i^\prime(s)^\perp\Big]\,d\Ha^{1}(x).
2788: \end{eqnarray*}
2789: We deduce that
2790: \begin{eqnarray*}
2791: \sum_{(i,s): c_i(s)=x} \frac{1}{L_i}c_i^\prime(s)^\perp &=& 0\quad\text{ for
2792: }\Ha^1-\text{almost all }x\in \spt(\Gamma)
2793: \end{eqnarray*}
2794: and since $|c_i^\prime|=L_i$ and the vectors
2795: $\{c_i^\prime(s)\,:\,c_i(s)=x\}$ are collinear, this implies that
2796: $\theta_\Gamma(x)=\#\{(i,s): c_i(s)=x\}$ is even.
2797: \end{proof}
2798:
2799: Using results from \cite{BeM05} we deduce that $\mu_\Gamma$ as in
2800: Proposition \ref{prop:conv-interface} is given by an alternative system of curves,
2801: where each curve is passed twice.
2802: \begin{lemma}\label{lem:alt-sys}
2803: Let $\Gamma=\{c_i, i=1,...,N\}$ be a $W^{2,2}$-system of closed curves without
2804: transversal crossings and with an even \multiplicity function $\theta_{\Gamma}$. Then
2805: there exists a system of curves $\overline{\Gamma}=\{\overline\gamma_i :
2806: i=1,...,\overline{N}\}$ such that
2807: \begin{align}
2808: \mu_\Gamma\,&=\, 2\mu_{\overline{\Gamma}},\label{eq:alt-sys-1}\\
2809: \W(\Gamma) \,&=\, 2\W(\overline{\Gamma}),\label{eq:alt-sys-2}\\
2810: \theta_{\Gamma} \,&=\, 2\theta_{\overline{\Gamma}}.\label{eq:alt-sys-3}
2811: \end{align}
2812: \end{lemma}
2813:
2814: \begin{proof}
2815: This Lemma requires some machinery from geometric measure
2816: theory. For the relevant definitions of Sobolev type submanifolds,
2817: Hutchinson varifolds and the various definitions for tangential lines
2818: and \multiplicity functions see \cite{BeM05}.
2819:
2820: By \cite{BeM05} Remark 4.9 we obtain that
2821: \begin{gather*}
2822: f\,:=\, \theta(\Gamma,\cdot)\Chi_{\spt(\Gamma)}
2823: \end{gather*}
2824: belongs to $\mathcal{S}^2_{tg}(\R^2)$, that is the set of Sobolev-type
2825: submanifold with
2826: the additional property that a unique tangential line exists in all points of
2827: the support.
2828: By \cite[Proposition 4.5]{BeM05} this implies that $f\in HV^2(\R^2)$ is a
2829: Hutchinson varifold such that a unique tangential line exists in all points of
2830: the support. Since the \multiplicity is even we conclude that
2831: also $\frac{1}{2}f$ belongs to $HV^2(\R^2)$ and that a unique tangential
2832: line exists in all points of the support. Using again \cite[Proposition 4.5]{BeM05}
2833: we obtain that $\frac{1}{2}f\in \mathcal{S}^2_{tg}(\R^2)$. Now
2834: \cite[Corollary 4.8]{BeM05} gives that $\frac{1}{2}f$ is given by a
2835: $W^{2,2}$-system $\overline{\Gamma}=\{\overline\gamma_i :
2836: i=1,...,\overline{N}\}$ of closed curves without transversal crossings. This implies
2837: \eqref{eq:alt-sys-1}. Equations \eqref{eq:alt-sys-2} and
2838: \eqref{eq:alt-sys-3} are an immediate consequence.
2839: \end{proof}
2840:
2841: We next show that the limit of the boundaries of the support of
2842: $u_\eps$ and the limit of the mass distributions $u_\eps$ are identical.
2843:
2844: \begin{proposition}\label{prop:lim-mass}
2845: For a system of curves $\Gamma$ as obtained in Proposition
2846: \ref{prop:conv-interface} and the measure $\mu$ in \eqref{eq:conv-ue} we have
2847: \begin{gather*}
2848: \mu_\Gamma\,=\, \mu.
2849: \end{gather*}
2850: In particular, \eqref{eq:conv-nabla-ue} holds for the whole sequence
2851: $\eps\to 0$.
2852: \end{proposition}
2853:
2854: \begin{proof}
2855: Let $\partial\spt(u_\eps)$ be given by a system of closed curves
2856: $\Gamma_\eps=\{\gamma_{\eps,i}\}_{i=1,...,N(\eps)}$ as in Remark
2857: \ref{rem:arclength-curves}.
2858: defined in Definition \ref{def:para} we obtain from Lemma \ref{lem:trafo-formula} that
2859: for all $\eta\in C^1_c(\R^{2};R^+_0)$
2860: \begin{align}
2861: &\Big|\int_{\R^2} \eta u_\eps -\int_{\R^2} \eta \eps|\nabla u_\eps|\,\Big|\notag\\
2862: =\,& \Bigg|
2863: \sum_{i=1}^{N(\eps)}\int_0^{L_{\eps,i}}\Big( \int_0^{M_{\eps,i}(s)}
2864: \eta\big(\gamma_{\eps,i}(s)+\te_{\eps,i}(s,m)\theta_{\eps,i}(s)\big)\,dm
2865: -\eta\big(\gamma_{\eps,i}(s)\big)\Big)\,ds\;\Bigg|\notag\\
2866: \leq\,& \sum_{i=1}^{N(\eps)}\int_0^{L_{\eps,i}}
2867: \big|M_{\eps,i}(s)-1\big|\eta\big(\gamma_{\eps,i}(s)\big)\,ds\notag\\
2868: &+ \sum_{i=1}^{N(\eps)}\int_0^{L_{\eps,i}}\int_0^{M_{\eps,i}(s)}
2869: \Big|\eta\big(\gamma_{\eps,i}(s)+\te_{\eps,i}(s,m)\theta_{\eps,i}(s)\big)-
2870: \eta\big(\gamma_{\eps,i}(s)\big)\Big|\,dm\,ds\notag\\
2871: \leq\,& \Big(\sum_{i=1}^{N(\eps)}\int_0^{L_{\eps,i}}
2872: \big(M_{\eps,i}(s)-1\big)^2\,ds\Big)^{\frac{1}{2}}\Big(\sum_{i=1}^{N(\eps)}\int_0^{L_{\eps,i}}
2873: \eta\big(\gamma_{\eps,i}(s)\big)^2\,ds\Big)^{\frac{1}{2}}\notag\\
2874: &+\|\eta\|_{C^1(\R^{2})}\sum_{i=1}^{N(\eps)}\int_0^{L_{\eps,i}}\int_0^{M_{\eps,i}(s)}
2875: \te_{\eps,i}(s,m)\,dm\,ds\notag\\
2876: \leq\,& \eps \G_\eps(u_\eps,v_\eps)^{\frac{1}{2}}2\|\eta\|_{C^0(\R^{2})}+
2877: \|\eta\|_{C^1(\R^{2})}d_1(u_\eps,v_\eps),
2878: \label{eq:proof-conv-20}
2879: \end{align}
2880: where we have used \eqref{eq:cost-1}. We observe that by
2881: \eqref{eq:bound-G-eps}
2882: \begin{gather*}
2883: \frac{1}{\eps} d_1(u_\eps,v_\eps)\,\leq\, 2 +\eps^2 \Lambda
2884: \end{gather*}
2885: and we deduce from \eqref{eq:proof-conv-20} that
2886: \begin{gather*}
2887: \Big|\int_{\R^{2}} \eta u_\eps -\int_{\R^2} \eta \eps|\nabla u_\eps|\,\Big|
2888: \,\leq\, \eps \Big(2\Lambda^{\frac{1}{2}}\|\eta\|_{C^0(\R^{2})}+
2889: (2+\eps^2\Lambda)\|\eta\|_{C^1(\R^{2})}\Big)
2890: \end{gather*}
2891: This shows that
2892: \begin{gather*}
2893: \mu\,=\, \lim_{\eps\to 0}u_\eps\LL^2\,=\,\lim_{\eps\to 0}\eps|\nabla
2894: u_\eps| \,=\, \mu_\Gamma
2895: \end{gather*}
2896: in the sense of convergence as measures.
2897: \end{proof}
2898: The proof of the lim-inf estimate now follows:
2899: by Proposition \ref{prop:conv-interface} there exists a system
2900: of curves $\Gamma$ of even \multiplicity such that
2901: \[
2902: \frac{1}{2}\W(\Gamma) \leq \liminf_{\eps\to 0} \G_\eps(u_\eps,v_\eps).
2903: \]
2904: By Lemma \ref{lem:alt-sys} there is an alternative system of curves
2905: $\tilde \Gamma$, of integer \multiplicity, such that $\mu_\Gamma =
2906: 2\mu_{\tilde\Gamma}$
2907: and $\W(\Gamma) = 2\W(\tilde \Gamma)$. This second system $\tilde \Gamma$
2908: is the system referred to in Theorem~\ref{the:main} as $\Gamma$. Finally,
2909: Proposition~\ref{prop:lim-mass} guarantees that
2910: $\mu_\Gamma = 2\mu_{\tilde\Gamma}$ is the limit of the sequence
2911: $(u_\e,v_\e)$ in the appropriate sense (see \eqref{eq:conv-ue}).
2912:
2913: %========================================================================
2914: % limsup estimate
2915: %========================================================================
2916: \section{The limsup estimate}
2917: \label{sec:limsup}
2918: %
2919: \subsection{The heart of the construction}
2920: Let us first perform the lim-sup construction in the case that $\mu=2\mu_{\Gamma}$,
2921: where $\Gamma$ consists of one simple curve.
2922: \begin{lemma}\label{lem:single-curve}
2923: Let $\gamma: [0,L)\to\R^2$ be a closed $C^2$-curve without self-intersections
2924: and with $|\gamma^\prime|=1$ and let $\Gamma=\{\gamma\}$.
2925: Then there exists $\eps_0>0$,
2926: $\eps_0=\eps_0(\|\gamma^{\prime\prime}\|_{C^0([0,L))})$ and sequences
2927: $(u_\eps, v_\eps)_{0<\eps<\eps_0}$ with
2928: \begin{gather}
2929: u_\eps,v_\eps\,\in\, BV(\R^2;\{0,1/\eps\}),\quad u_\eps v_\eps\,=\,0,
2930: \label{eq:ls-reg}\\
2931: \int_{\R^2} u_\eps \,=\, \int_{\R^2} v_\eps \,=\,
2932: 2L,\label{eq:ls-mass}\\
2933: \spt(u_\eps)\cup\spt(v_\eps)\,\subset\, \{x: \dist(x,\Gamma)\leq 3\eps\},
2934: \label{eq:ls-spt-uv}
2935: \end{gather}
2936: such that
2937: \begin{align}
2938: 2\mu_{\Gamma} &\,=\, \lim_{\eps\to 0}u_\eps\,\LL^2,\label{eq:ls-conv}\\
2939: \frac{1}{2}\int_0^{L} \gamma^{\prime\prime}(s)^2\,ds &\,\geq\, \limsup_{\eps\to
2940: 0} \frac{1}{\eps^2}\Big(\frac{1}{\eps} d_1(u_\eps,v_\eps)
2941: +\eps\int_{\R^2}|\nabla u_\eps| -4L\Big).\label{eq:ls-ls}
2942: \end{align}
2943: \end{lemma}
2944: \begin{proof}
2945: There exists a $\eps_0>0$ such that
2946: \begin{gather}
2947: \sup_{s\in [0,L)} |\gamma^{\prime\prime}(s)| \,\leq\,
2948: \frac{1}{4\eps_0},\label{eq:ls-eps0-curv}\\
2949: x \mapsto \dist(x,\Gamma)\text{ is of class }C^2\text{ in }\{x:\dist(x,\Gamma)< 4\eps_0\}.
2950: \label{eq:ls-eps0-dist}
2951: \end{gather}
2952: Let $\nu:[0,L)\to S^1$ be the $C^1$ unit normal field of $\gamma$ that satisfies
2953: \begin{gather*}
2954: \det (\gamma^\prime,\nu)\,=\,-1
2955: \end{gather*}
2956: and let $\kappa(s)$ be the curvature
2957: in direction of $\nu(s)$,
2958: \begin{gather*}
2959: \kappa(s) \,=\, -\nu^\prime(s)\cdot\gamma^\prime(s) \,=\,
2960: \nu(s)\cdot\gamma^{\prime\prime}(s).
2961: \end{gather*}
2962: For $\eps<\eps_0$ we set
2963: \begin{eqnarray*}
2964: u_\eps(x) &:=&
2965: \begin{cases}
2966: \frac{1}{\eps} &\text{ if }\dist(x,\Gamma)<\eps,\\
2967: 0 &\text{ elsewise,}
2968: \end{cases}
2969: \end{eqnarray*}
2970: and define two curves $\hat{\gamma}_+,\hat{\gamma}_- :[0,L)\to\R^2$ by
2971: \begin{eqnarray*}
2972: \hat{\gamma}_+ (s)&:=& \gamma(s) +\eps\nu(s),\\
2973: \hat{\gamma}_- (s)&:=& \gamma(s) -\eps\nu(s),
2974: \end{eqnarray*}
2975: see Figure \ref{fig:limsup}.
2976: Using the parametrization $\Phi(s,r):=\gamma(s)+r\nu(s)$ we calculate
2977: that
2978: \begin{gather*}
2979: |\det D\Phi(s,r)|\,=\, 1-r\kappa(s)\quad\text{ for }0<
2980: s<L,\,-\eps<r<\eps.
2981: \end{gather*}
2982: Therefore for any $\eta\in C^0(\R^2)$
2983: \begin{gather*}
2984: \lim_{\eps\to 0}\int_{\R^2} \eta(x)u_\eps(x)\,ds\,=\, \lim_{\eps\to 0}
2985: \int_0^L\int_{-\eps}^\eps
2986: \eta(\Phi(s,r))\frac{1}{\eps}(1-r\kappa(s))\,dr\,ds\\
2987: =\, \int_0^L
2988: 2\eta(\Phi(s,0))\,ds\,=\, 2\int_0^L \eta(\gamma(s))\,ds
2989: \end{gather*}
2990: holds and we deduce \eqref{eq:ls-conv}.
2991:
2992: The arc length functions corresponding to $\hat{\gamma}_+, \hat{\gamma}_-$ are given by
2993: \begin{eqnarray*}
2994: \vartheta_{+}(s) &=& \int_0^s|\hat{\gamma}_+^\prime(\sigma)|\,
2995: d\sigma\,=\,\int_0^s (1-\eps\kappa(\sigma))\,d\sigma,\\
2996: \vartheta_{-}(s) &=& \int_0^s|\hat{\gamma}_-^\prime(\sigma)|\,
2997: d\sigma\,=\,\int_0^s (1+\eps\kappa(\sigma))\,d\sigma.
2998: \end{eqnarray*}
2999: Let $\tilde{\gamma}_+,\tilde{\gamma}_-$ be the corresponding
3000: parametrizations with respect to arclength,
3001: \begin{gather*}
3002: \tilde{\gamma}_{+} \,=\, \hat{\gamma}_{+}\circ \vartheta_{+}^{-1},\qquad
3003: \tilde{\gamma}_+: [0,L_+)\to\R^2,\quad L_+\,=\, L
3004: -\eps\int_0^L\kappa(\sigma)\,d\sigma,\\
3005: \tilde{\gamma}_{-} \,=\, \hat{\gamma}_{-}\circ
3006: \vartheta_{-}^{-1},\qquad
3007: \tilde{\gamma}_-: [0,L_-)\to\R^2,\quad L_-\,=\, L
3008: +\eps\int_0^L\kappa(\sigma)\,d\sigma.\\
3009: \end{gather*}
3010: The curves $\tilde{\gamma}_+,\tilde{\gamma}_-$ parametrize the boundary of
3011: $\{u_\eps=1/\eps\}$ and we obtain that
3012: \begin{eqnarray}\label{eq:boundary-limsup}
3013: \int_{\R^2} \eps|\nabla u_\eps| &=& L_+ + L_- \,=\, 2L.
3014: \end{eqnarray}
3015: Corresponding to $\tilde{\gamma}_+,\tilde{\gamma}_-$ we define unit
3016: normal fields
3017: \begin{eqnarray*}
3018: \tilde{\nu}_+ &:=& \nu\circ\vartheta_+^{-1},\\
3019: \tilde{\nu}_- &:=& \nu\circ\vartheta_-^{-1}
3020: \end{eqnarray*}
3021: and we compute the curvature of $\tilde{\gamma}_+,\tilde{\gamma}_-$ in
3022: direction of $\nu_{\pm}$
3023: \begin{eqnarray*}
3024: \tilde{\kappa}_+ &=&
3025: \frac{\kappa\circ\vartheta_+^{-1}}{1-\eps\kappa\circ\vartheta_+^{-1}},\\
3026: \tilde{\kappa}_- &=&
3027: \frac{\kappa\circ\vartheta_-^{-1}}{1+\eps\kappa\circ\vartheta_-^{-1}}.
3028: \end{eqnarray*}
3029: Observe that by \eqref{eq:ls-eps0-curv} for $\eps<\eps_0$
3030: \begin{gather}
3031: |\tilde{\kappa}_+|,|\tilde{\kappa}_-|\,\leq\,
3032: \frac{1}{3\eps_0}.\label{eq:ls-est-kappa}
3033: \end{gather}
3034: As in subsection \ref{subsec:para} we define \emph{mass coordinates} $\ma_+,\ma_-$ by
3035: \begin{eqnarray*}
3036: \ma_+(r,t) &:=& \frac{1}{\eps}\big( t-\frac{t^2}{2}\tilde{\kappa}_+(r)\big),\\
3037: \ma_-(r,t) &:=& \frac{1}{\eps}\big( t-\frac{t^2}{2}\tilde{\kappa}_-(r)\big)
3038: \end{eqnarray*}
3039: and the inverse mappings $\te_+(r,\cdot)=\ma_+(r,\cdot)^{-1}$, $\te_-(r,\cdot)=\ma_-(r,\cdot)^{-1}$,
3040: \begin{eqnarray*}
3041: \te_+(r,m) &=& \frac{1}{\tilde{\kappa}_+(r)}\Big(1-\sqrt{1-2\eps\tilde{\kappa}_+m}\Big),\\
3042: \te_-(r,m) &=& \frac{1}{\tilde{\kappa}_-(r)}\Big(1-\sqrt{1-2\eps\tilde{\kappa}_-m}\Big).
3043: \end{eqnarray*}
3044: By \eqref{eq:ls-est-kappa} the expressions
3045: $1-2\eps\tilde{\kappa}_\pm m$ are positive for $|m|\leq 1$. Using that
3046: \begin{alignat*}{2}
3047: \sqrt{1+2z}\,&\leq\, 1+z,\quad&&\text{ for }z\geq -\frac{1}{2},\\
3048: \sqrt{1+2z}\,&\geq\, 1+2z,\quad&&\text{ for } -\frac{1}{2}\leq z\leq 0
3049: \end{alignat*}
3050: we deduce that
3051: \begin{gather}
3052: 0\,\leq\, \frac{1}{m}\te_+(r,m)\,\leq\,
3053: 2\eps,\qquad 0\,\leq\, \frac{1}{m}\te_-(r,m)\,\leq\,
3054: 2\eps.\label{eq:range-l}
3055: \end{gather}
3056: We construct parametrizations $\psi_+,\psi_-$ in the same way as
3057: we did in subsection \ref{subsec:para}. Considering
3058: $\tilde{\gamma}_+$, the unit
3059: normal field $\tilde{\nu}_+$ takes the role of the direction field
3060: $\theta$ in subsection \ref{subsec:para}. Therefore $\tilde{\kappa}_+$ and
3061: $1$ appear in place of $\alphaprime$ and $\sin\beta$. Note that,
3062: with respect to $\tilde{\gamma}_+$, $\spt(u_\eps)$ is in the direction of
3063: $-\tilde{\nu}_+$ and, with respect to $\tilde{\gamma}_-$, in direction of
3064: $\tilde{\nu}_-$.\\
3065: \begin{figure}[ht]
3066: \centerline{\psfig{width=\textwidth,figure=limsup}}
3067: \caption{The limsup construction.}
3068: \label{fig:limsup}
3069: \end{figure}
3070: For $\tilde{\gamma}_+, \tilde{\gamma}_-$ respectively we define `ray
3071: lengths' $l_+(r),l_-(r)$ such that their values in mass
3072: coordinates become $-1$ and $1$,
3073: \begin{align}
3074: l_+(r) &:= \te_+(r,-1)\,=\,
3075: \frac{1}{\tilde{\kappa}_+(r)}\Big(1-\sqrt{1+2\eps\tilde{\kappa}_+}\Big),
3076: \label{eq:ls-def-l+}\\
3077: l_-(r) &:= \te_-(r,1)\,=\,
3078: \frac{1}{\tilde{\kappa}_-(r)}\Big(1-\sqrt{1-2\eps\tilde{\kappa}_-}\Big).
3079: \label{eq:ls-def-l-}
3080: \end{align}
3081: To parametrize $\spt(u_\eps)$ we define sets
3082: \begin{alignat*}{1}
3083: D_+ &:=\{ (r,t) : 0\leq r< L_+, l_+(r)<t<0\}\\
3084: D_- &:=\{ (r,t) : 0\leq r< L_-, 0<t<l_-(r)\}.
3085: \end{alignat*}
3086: and the maps
3087: \begin{xalignat*}{2}
3088: \psi_+(r,t) &:=
3089: \tilde{\gamma}_+(r)+ t\tilde{\nu}_+(r)&&\text{ for }0<r<L_+, -2\eps<t<2\eps\\
3090: \psi_-(r,t) &:=
3091: \tilde{\gamma}_-(r)+ t\tilde{\nu}_-(r)&&\text{ for }0<r<L_-, -2\eps<t<2\eps,
3092: \end{xalignat*}
3093: which are, for
3094: $\eps<\eps_0$, by \eqref{eq:ls-eps0-dist} injective on their
3095: domains. Analogous to Lemma~\ref{lem-mass} we observe that for any
3096: $\eta\in C^0_c(\R^2)$
3097: \begin{align}
3098: \int_{\psi_+(D_+)} \eta(x)u_\eps(x)\,dx \,&=\,
3099: \int_0^{L_+}\int_{-1}^0 \eta\big(\psi_+(r,\te_+(r,m))\big)\,dm\,dr,
3100: \label{eq:ls-trafo} \\
3101: \int_{\psi_-(D_-)} \eta(x)u_\eps(x)\,dx \,&=\,
3102: \int_0^{L_-}\int_0^1
3103: \eta\big(\psi_-(r,\te_-(r,m))\big)\,dm\,dr. \label{eq:ls-trafo2}
3104: \end{align}
3105: Moreover we deduce from \eqref{eq:ls-def-l+}, \eqref{eq:ls-def-l-} that
3106: \begin{align*}
3107: l_+(\vartheta_+(s)) &= \frac{1}{\kappa(s)}\Big(1-\eps\kappa(s)
3108: -\sqrt{1-(\eps\kappa(s))^2} \Big),\\
3109: l_-(\vartheta_-(s)) &= \frac{1}{\kappa(s)}\Big(1+\eps\kappa(s)
3110: -\sqrt{1-(\eps\kappa(s))^2} \Big).
3111: \end{align*}
3112: which implies that
3113: \begin{gather*}
3114: l_-(\vartheta_-(s)) - l_+(\vartheta_+(s))\,=\, 2\eps.
3115: \end{gather*}
3116: We conclude that the sets $\psi_+(D_+)$ and $\psi_-(D_-)$
3117: are pairwise disjoint and cover almost all of $\spt(u_\eps)$. We
3118: therefore obtain by \eqref{eq:ls-trafo}, \eqref{eq:ls-trafo2} that
3119: \begin{align*}
3120: &\int \eta(x)u_\eps(x)\,dx \\
3121: &\,=
3122: \int_0^{L_+}\int_{-1}^0 \eta\big(\psi_+(r,\te_+(r,m))\big)\,dm\,dr +
3123: \int_0^{L_-}\int_0^1 \eta\big(\psi_-(r,\te_-(r,m))\big)\,dm\,dr
3124: \end{align*}
3125: holds.
3126: In particular we obtain
3127: \begin{gather}
3128: \int_{\R^2} u_\eps \,=\, L_+ +L_- \,=\, 2L. \label{eq:mass-u-eps}
3129: \end{gather}
3130: We now define an injective transport map $S:\spt(u_\eps)\to\R^2$
3131: by
3132: \begin{alignat*}{2}
3133: S(\psi_+(r,t))\,:=&\,
3134: \psi_+\big(r,\te_+\big(r,\ma_+(r,t)+1\big)\big)
3135: &&\text{ for }(r,t)\in D_+,\\
3136: S(\psi_-(r,t))\,:=&\,
3137: \psi_-\big(r,\te_-\big(r,\ma_-(r,t)-1\big)\big)
3138: &&\text{ for }(r,t)\in D_-
3139: \end{alignat*}
3140: and set
3141: \begin{gather*}
3142: v_\eps \,:=\, \frac{1}{\eps}\Chi_{S(\spt(u_\eps))}.
3143: \end{gather*}
3144: This gives
3145: \begin{gather*}
3146: \spt(v_\eps)\,=\, \{\psi_+(r,t): 0\leq t\leq \te_+(r,1)\} \cup
3147: \{\psi_-(r,t): \te_-(r,-1)\leq t\leq 0\}.
3148: \end{gather*}
3149: By \eqref{eq:range-l} we obtain that $\te_+(r,1),|\te_-(r,-1)|\leq
3150: 2\eps$ and
3151: since the distance of $\hat{\gamma}_+$ and $\hat{\gamma}_-$ to $\gamma$
3152: equals $\eps$ the equation \eqref{eq:ls-spt-uv} follows.
3153: We compute for an arbitrary $\eta\in C^0_c(\R^2)$ that
3154: \begin{align*}
3155: &\int\eta(S(x))u_\eps(x)\,dx\\
3156: =\,& \int_0^{L_+}\int_{-1}^0
3157: \eta\big(S(\psi_+(r,\te_+(r,m)))\big)\,dm\,dr
3158: + \int_0^{L_-}\int_0^1
3159: \eta\big(S(\psi_-(r,\te_-(r,m)))\big)\,dm\,dr\\
3160: =\,& \int_0^{L_+}\int_{-1}^0
3161: \eta\big(\psi_+(r,\te_+(r,m+1))\big)\,dm\,dr
3162: + \int_0^{L_-}\int_0^1
3163: \eta\big(\psi_-(r,\te_-(r,m-1))\big)\,dm\,dr\\
3164: =\,& \int_0^{L_+}\int_0^1
3165: \eta\big(\psi_+(r,\te_+(r,m))\big)\,dm\,dr
3166: + \int_0^{L_-}\int_{-1}^0
3167: \eta\big(\psi_-(r,\te_-(r,m))\big)\,dm\,dr\\
3168: =\,& \int \eta(x)v_\eps(x)\,dx,
3169: \end{align*}
3170: where we have used the change-of-variables formula in the last equality.
3171: This shows that $S$ indeed is a transport map. In particular it follows
3172: from \eqref{eq:mass-u-eps} that \mbox{$\int_{\R^2} v_\eps =2L$} holds,
3173: which
3174: proves \eqref{eq:ls-mass}. Moreover
3175: we obtain
3176: \begin{eqnarray}
3177: d_1(u_\eps,v_\eps) &\leq& \int_{\R^2} |x-S(x)| u_\eps(x)\,dx\notag\\
3178: &=& \int_0^{L_+}\int_{-1}^0 \big(
3179: \te_+(r,m+1) -\te_+(r,m)\big)\,dm\,dr \notag\\
3180: && +\int_0^{L_-}\int_0^1 \big( \te_-(r,m)
3181: -\te_-(r,m-1))\big)\,dm\,dr \label{eq:ls-d1}
3182: \end{eqnarray}
3183: By \eqref{eq:ls-def-l+} we have $\ma_+(r,l_+(r))=-1$ and we deduce for
3184: the inner integral of the first term on the right-hand side of
3185: \eqref{eq:ls-d1} that
3186: \begin{align*}
3187: & \int_{-1}^0 \big(
3188: \te_+(r,m+1) -\te_+(r,m)\big)\,dm\notag\\
3189: =\,& \int_0^{\ma_+(r,l_+(r))}\te_+(r,m)
3190: -\te_+\bigl(r,m-\ma_+(r,l_+(r))\bigr)\,dm.
3191: \end{align*}
3192: Therefore Lemma \ref{lem:stand-trans-est} yields that
3193: \begin{gather}
3194: \int_{-1}^0 \big(
3195: \te_+(r,m+1) -\te_+(r,m)\big)\,dm
3196: \,=\, \eps +\frac{1}{4}\eps^3
3197: \tilde{\kappa}_+(r)^2 +R_+(r)\eps^5,\label{eq:ls-d1+}
3198: \end{gather}
3199: where
3200: \begin{gather}
3201: 0\,\leq\, R_+(r)\,\leq\, \tilde{\kappa}_+(r)^4\,\leq\, C(\eps_0)
3202: \label{eq:ls-R+}
3203: \end{gather}
3204: by \eqref{eq:est-R} and \eqref{eq:ls-eps0-curv}.
3205: For the inner integral of the second term in \eqref{eq:ls-d1} we deduce
3206: by similar arguments
3207: \begin{align}
3208: \int_0^1 \te_-(r,m)-\te_-(r,m-1)\,dm
3209: =\,& \eps +\frac{1}{4}\eps^3
3210: \tilde{\kappa}_-(r)^2 +R_-(r)\eps^5,\label{eq:ls-d1-}
3211: \end{align}
3212: with
3213: \begin{gather}
3214: 0\,\leq\, R_-(r)\,\leq\, \tilde{\kappa}_-(r)^4\,<\, C(\eps_0).
3215: \label{eq:ls-R-}
3216: \end{gather}
3217: By \eqref{eq:ls-d1} and \eqref{eq:ls-d1+}-\eqref{eq:ls-R-} we obtain
3218: that
3219: \begin{align}
3220: &\frac{1}{\eps} d_1(u_\eps,v_\eps)\notag\\
3221: \leq\,& L_+ +L_- +\frac{1}{4}\eps^2
3222: \int_0^{L_+} \tilde{\kappa}_+(r)^2\,dr +\frac{1}{4}\eps^2
3223: \int_0^{L_-} \tilde{\kappa}_-(r)^2\,dr +\eps^4 L
3224: C(\eps_0).\label{eq:ls-d1-sum}
3225: \end{align}
3226: We compute that
3227: \begin{align*}
3228: \int_0^{L_+} \tilde{\kappa}_+(r)^2\,dr +
3229: \int_0^{L_-} \tilde{\kappa}_-(r)^2\,dr \,&=\, \int_0^L\Big(
3230: \frac{\kappa(s)^2}{1-\eps\kappa(s)} +
3231: \frac{\kappa(s)^2}{1+\eps\kappa(s)}\Big)\,ds\\
3232: &=\, 2\int_0^L \kappa(s)^2\,ds + 2\eps^2\int_0^L
3233: \frac{\kappa(s)^2}{1-\eps^2\kappa(s)^2}\,ds\\
3234: &\leq\, 2\int_0^L \kappa(s)^2\,ds + 2\eps^2 LC(\eps_0)
3235: \end{align*}
3236: which gives with \eqref{eq:ls-d1-sum} and $2L=L_++L_-$ that
3237: \begin{gather}
3238: \frac{1}{\eps}d_1(u_\eps,v_\eps) \,\leq\, 2L
3239: +\frac{\eps^2}{2}\int_0^L\kappa(s)^2\,ds +\eps^4LC(\eps_0).
3240: \label{eq:ls-d1-sum2}
3241: \end{gather}
3242: Since $|\kappa|=|\gamma^{\prime\prime}|$ we obtain from
3243: \eqref{eq:boundary-limsup} and \eqref{eq:ls-d1-sum2} that
3244: \begin{gather*}
3245: \frac{1}{\eps^2}\Big(\frac{1}{\eps} d_1(u_\eps,v_\eps)
3246: +\eps\int_{\R^2}|\nabla u_\eps| -4L\Big) \,\leq\,
3247: \frac{1}{2}\int_0^L\gamma^{\prime\prime}(s)^2\,ds +\eps^2 LC(\eps_0),
3248: \end{gather*}
3249: which proves \eqref{eq:ls-ls}.
3250: \end{proof}
3251: %
3252: %
3253: %
3254: \subsection{The general case}
3255: \label{subsec:ls-gen}
3256: We prove now the limsup estimate stated in Theorem \ref{the:main}.
3257: First we need a technical Lemma which follows from
3258: results in \cite{BMP} and \cite{BeM} and approximates a given system of
3259: curves as in Theorem \ref{the:main} strongly in $W^{2,2}$ by a sequence
3260: of more regular systems of curves.
3261: \begin{lemma}\label{lem:detach}
3262: Let $\Gamma$ be a $W^{2,2}$-system of closed curves without
3263: transversal crossings. Then there exists a sequence $(\Gamma^j)_{j\in \N}$
3264: of $W^{2,2}$-systems of closed curves and a number
3265: $m\in\N$ such that the following holds:
3266: \begin{liste(a)}
3267: \item
3268: For all $j\in\N$ the system of curves
3269: $
3270: \Gamma^j = \{\gamma^j_k\}_{k=1,...,m}
3271: $
3272: is given by a pairwise disjoint family of simple closed curves and satisfies
3273: \begin{gather}
3274: |\Gamma^j| = |\Gamma|.\label{eq:equal-length}
3275: \end{gather}
3276: \item
3277: For all $1\leq k\leq m$ holds
3278: \begin{align}\label{eq:conv-strong-gamma}
3279: & \gamma^j_k \,\to\, \gamma_k\quad\text{ in }W^{2,2}(0,1)\quad\text{ as }
3280: j\to\infty.
3281: \end{align}
3282: \item $\tilde{\Gamma}:=\{\gamma_k\}_{k=1,...,m}$ is a
3283: $W^{2,2}$-system of closed curves without transversal crossings.
3284: \item
3285: $\tilde{\Gamma}$ is equivalent to $\Gamma$ in the sense that
3286: \begin{gather*}
3287: \mu_{\Gamma}\,=\, \mu_{\tilde{\Gamma}}.
3288: \end{gather*}
3289: \end{liste(a)}
3290: In particular we have
3291: \begin{align}
3292: \W(\Gamma^j)\,&\to \W(\Gamma)\quad\text{ as }
3293: j\to\infty.\label{eq:conv-G}
3294: \end{align}
3295: \end{lemma}
3296:
3297: \begin{proof}
3298: By \cite[Corollary 5.2]{BeM} there exists a sequence
3299: $(\tilde{\Gamma}^j)_{j\in\N}$ and a $W^{2,2}$-system of closed curves
3300: $\tilde{\Gamma}$ such that
3301: \begin{liste}
3302: \item
3303: for all $j\in\N$ $\tilde{\Gamma}^j$ is an oriented parametrization of a bounded
3304: open smooth set $E_j\subset\R^2$,\label{enum:i}
3305: \item
3306: $\tilde{\Gamma}^j$ converge weakly in $W^{2,2}$ to $\tilde{\Gamma}$ as
3307: $j\to\infty$,\label{enum:ii}
3308: \item
3309: $\tilde{\Gamma}^j$ converge `in energy' to $\tilde{\Gamma}$ as
3310: $j\to\infty$.\label{enum:iii}
3311: \item
3312: $\tilde{\Gamma}$ and $\Gamma$ are equivalent.
3313: \end{liste}
3314: Property \eqref{enum:i} implies that $\tilde{\Gamma}^j$ is a disjoint system of
3315: simple closed curves parametrized on $(0,1)$ proportional to arclength.
3316:
3317: By property \eqref{enum:ii} there exists a number
3318: $m\in\N$ such that
3319: \begin{gather*}
3320: \tilde{\Gamma}^j\,=\, \{\tilde{\gamma}^j_k\}_{k=1,...,m},\quad
3321: \tilde{\Gamma}\,=\, \{\tilde{\gamma}_k\}_{k=1,...,m},\\
3322: \tilde{\gamma}^j_k,\,\tilde{\gamma}_k : (0,1)\,\to\,\R^2,\\
3323: |(\tilde{\gamma}^j_k)^\prime|\,=\, L^j_k,\quad
3324: |(\tilde{\gamma}_k)^\prime|\,=\, L_k\text{ on }(0,1),\notag\\
3325: \end{gather*}
3326: with
3327: \begin{gather}
3328: \tilde{\gamma}^j_k\,\to\, \tilde{\gamma}_k\quad\text{ weakly in
3329: }W^{2,2}(0,1)\text{ as }j\to\infty.\label{eq:conv-gamma}
3330: \end{gather}
3331: In particular we deduce that $L^j_k\to\ L_k$ as $j\to\infty$ for all
3332: $k=1,...,m$
3333: and
3334: \begin{gather}
3335: |\tilde{\Gamma}^j|\,\to\, |\tilde{\Gamma}|\label{eq:conv-length}
3336: \end{gather}
3337: holds.
3338: \\
3339: The `convergence in energy' stated in \eqref{enum:iii} gives that, as
3340: $j\to\infty$,
3341: \begin{gather}
3342: |\tilde{\Gamma}^j| + \sum_{k=1}^m \int_0^1 (L^j_k)^{-3}
3343: |(\tilde{\gamma}^j_k)^{\prime\prime}|^2\,\to\,
3344: |\tilde{\Gamma}|+\sum_{k=1}^m \int_0^1 (L_k)^{-3}
3345: |(\tilde{\gamma}_k)^{\prime\prime}|^2.\label{eq:conv-energy}
3346: \end{gather}
3347: Together with \eqref{eq:conv-length} and the weak convergence
3348: \eqref{eq:conv-gamma} this
3349: yields
3350: \begin{gather}
3351: \tilde{\gamma}^j_k\,\to\, \tilde{\gamma}_k\quad\text{ strongly in
3352: }W^{2,2}(0,1)\text{ as }j\to\infty,\label{eq:conv-tilde-gamma}\\
3353: \W(\tilde{\Gamma}^j)\,\to
3354: \W(\Gamma)\text{ as }j\to\infty.\label{eq:conv-tilde-G}
3355: \end{gather}
3356: Finally we define for $j\in\N$ a modified system of curves
3357: $\Gamma^j_k=\{\gamma^j_k\}_{k=1,...,m}$ by
3358: \begin{gather*}
3359: \tilde{\gamma}^j_k(s)\,:=\,
3360: \frac{|\Gamma|}{|\tilde{\Gamma}^j|}\tilde{\gamma}^j_k(s).
3361: \end{gather*}
3362: Then \eqref{eq:equal-length} is satisfied. By \eqref{eq:conv-length}
3363: we see that $\tilde{\gamma}^j_k$ and $\gamma^j_k$ have for all $1\leq
3364: k\leq m$ the same limits as $j\to\infty$. We therefore obtain from
3365: \eqref{eq:conv-tilde-gamma}, \eqref{eq:conv-tilde-G} that
3366: \eqref{eq:conv-strong-gamma}, \eqref{eq:conv-G}
3367: hold.
3368: \end{proof}
3369: \ \\
3370: {\bf Proof of Theorem \ref{the:main} (lim-sup part):}\\
3371: Let $\Gamma$ be given as in Theorem \ref{the:main} and
3372: $(\Gamma^j)_{j\in\N}$ be a sequence of systems of closed curves as in
3373: Lemma \ref{lem:detach}, given by $\Gamma^j=\{\gamma^j_k\}_{k=1,...,m}$.
3374: We now parametrize the curves $\gamma^j_k$ by arclength,
3375: \begin{gather*}
3376: \gamma^j_k: (0,L^j_k)\to \R^2,\quad |(\gamma^j_k)^\prime|=1.
3377: \end{gather*}
3378: Choose $\eps^j_0$ such that the distance functions $\dist(\cdot,\Gamma^j)$
3379: is smooth in the set $\{x:\dist(x,\Gamma^j)<4\eps^j_0\}$.\\
3380: For $0<\eps<\eps^j_0$ we let $u^{j,k}_\eps$, $v^{j,k}_\eps$ be the
3381: approximations constructed in Lemma \ref{lem:single-curve} and define
3382: \begin{gather*}
3383: u^j_{\eps}\,:=\, \sum_{k=1}^m u^{j,k}_\eps,\qquad
3384: v^j_{\eps}\,:=\, \sum_{k=1}^m v^{j,k}_\eps.
3385: \end{gather*}
3386: By \eqref{eq:ls-spt-uv} all the functions $u^{j,k}_\eps,
3387: v^{j,k}_\eps$ have pairwise disjoint supports.
3388: We compute that
3389: \begin{align*}
3390: \int_{\R^2} u^j_\eps\,d\LL^2\,=\, \sum_{k=1}^m \int_{\R^2}
3391: u^{j,k}_\eps\,d\LL^2\,=\,
3392: \sum_{k=1}^m 2L^k_j\,=\, 2|\Gamma^j|\,=\, 2|\Gamma|\,=\, M
3393: \end{align*}
3394: and, by similar calculations, $\int v^j_\eps =M$.
3395: Moreover, by \eqref{eq:ls-conv}, \eqref{eq:conv-strong-gamma},
3396: \begin{align*}
3397: \lim_{j\to\infty}\lim_{\eps\to 0} u^j_\eps\LL^2\,=\, \lim_{j\to\infty}
3398: \sum_{k=1}^m \lim_{\eps\to 0} u^{j,k}_\eps\LL^2 \,=\, 2\lim_{j\to\infty}
3399: \mu_{\Gamma^j}\,=\, 2\mu_{\Gamma}
3400: \end{align*}
3401: and, by \eqref{eq:conv-G}, \eqref{eq:ls-ls},
3402: \begin{align*}
3403: \W(\Gamma)\,&=\,\lim_{j\to\infty} \W(\Gamma^j)\\
3404: &=\, \lim_{j\to\infty}\frac{1}{2}\sum_{k=1}^m \int_0^{L^j_k}
3405: \big|(\gamma^j_k)^{\prime\prime}\big|^2 \,ds\\
3406: &\geq\, \limsup_{j\to\infty}\sum_{k=1}^m \limsup_{\eps\to 0} \frac{1}{\eps^2}\Big(
3407: \frac{1}{\eps}d_1(u^{j,k}_\eps,v^{j,k}_\eps) + \eps\int |\nabla u^{j,k}_\eps| -
3408: 2\int u^{j,k}_\eps\,d\LL^2\Big)\\
3409: &\geq\, \limsup_{j\to\infty}\limsup_{\eps\to 0} \frac{1}{\eps^2}\Big(
3410: \frac{1}{\eps}d_1(u^j_\eps,v^{j}_\eps) + \eps\int |\nabla u^j_\eps| -
3411: 2\int u^j_\eps\,d\LL^2\Big).
3412: \end{align*}
3413: The final inequality follows from constructing an admissible joint transport map for $(u_\eps^j,v_\eps^j)$ from the transport maps of $(u_\eps^{j,k},v_\eps^{j,k})$.
3414:
3415:
3416: Therefore there exist subsequences $(j_l)_{l\in\N}, (\eps_l)_{l\in\N}$
3417: such that
3418: \begin{gather*}
3419: \lim_{l\to\infty} u^{j_l}_{\eps_l}\LL^2\,=\,2\mu_\Gamma
3420: \end{gather*}
3421: as Radon measures on $\R^2$ and
3422: \begin{gather*}
3423: \W(\Gamma) \,\geq\, \limsup_{l\to\infty} \G_{\eps_l}(u_{\eps_l},v_{\eps_l}).
3424: \end{gather*}
3425: This proves the lim-sup estimate in Theorem \ref{the:main}.\hfill $\Box$
3426: %========================================================================================
3427: % discussion
3428: %========================================================================================
3429: \section{Discussion}
3430: \label{sec:discussion}
3431:
3432: \subsection{General remarks}
3433: In the Introduction we formulated two basic questions about
3434: self-aggregating, partially localized structures: first, why do they
3435: exist at all (and why are they stable), and secondly, how can we
3436: understand their resistance to stretching, bending, and fracture, that
3437: is observed experimentally.
3438:
3439: The fact that the functional $\F_\eps$ favours structures in which $u$
3440: and $v$ alternate on a length scale $\e$ is well illustrated by the
3441: special cases of Section~\ref{sec:heuristics}. Simply put, for
3442: $d_1(u,v)$ to be moderate, the masses represented by $u$ and $v$ have to
3443: be close; the fact that $\Mod u_{L^\infty}$ and $\Mod v_{L^\infty}$ are
3444: limited implies that these masses occupy a certain amount of volume; and
3445: finally the interface penalization prevents the masses of $u$ and $v$
3446: from being too finely interspersed. Together these restrictions impose a
3447: length scale on structures with moderate $\F_\eps$.
3448:
3449: This argument does not determine, however, the \emph{geometry} of
3450: structures with moderate (or even lowest) energy. The fact that the
3451: straight lamellar geometry has lowest energy is not obvious, as is
3452: demonstrated by the `wriggled lamellar' patterns of Ren and Wei in a
3453: strongly related system~\cite{RenWei05}. The reason for the penalization
3454: of curvature in the current context can be traced back to a property of
3455: strict convexity of the function $\mathfrak t(s,m)$ defined in
3456: \eqref{eq:def-t_s}. Since this convexity property lies at the
3457: heart of the stability of these partially localized structures, we take
3458: some time to investigate it further.
3459:
3460: \subsection{The geometric basis for stability}
3461:
3462: An essential observation in the proof is that it is more convenient to
3463: use
3464: \emph{mass} coordinates along rays instead of \emph{length} coordinates,
3465: since the \Wasserstein distance measures spatial distances between pairs
3466: of infinitesimal portions of mass. The function $\mathfrak t(s,m)$
3467: (see~\eqref{eq:def-t_s}) connects the two descriptions by giving the
3468: length coordinate $t$ as a function of mass coordinate $m$.
3469:
3470: For fixed $m$, the value of $\te(s,m)$ depends on the value of
3471: $\alpha'(s)$, as is obvious both from the definition and from a
3472: geometric point of view (see Fig.~\ref{fig:effectcurvature} (left)).
3473: For the purpose of this discussion we write the dependence on
3474: $\alpha'(s)$ explicitly, as
3475: $\mathfrak t(s,m;\alpha'(s))$. The pertinent observation is that for
3476: fixed $m\not=0$, the function $\alpha'(s) \mapsto \mathfrak
3477: t(s,m;\alpha'(s))$ is strictly convex.
3478: This convexity can be seen in a simple plot of the function $\te$, as in
3479: Fig.~\ref{fig:effectcurvature} (right), but can also be recognized in
3480: the geometry of rays that are rotated with respect to each other
3481: (Fig.~\ref{fig:effectcurvature} (left)).
3482:
3483:
3484: \begin{figure}[ht]
3485: \centering
3486: \centerline{\vc{\psfig{figure=effectcurvature2,height=5cm}}
3487: \hskip5mm \vc{\psfig{figure=ts2,height=5cm}}}
3488: \caption{Left: the function $\te$ maps the mass coordinate $m$ to a
3489: length coordinate $t$, in the following way. Fixing an infinitesimal
3490: section $ds$ of interface, a section of mass $mds$, delimited by $ds$
3491: and by two rays, extends away from the interface by a distance of
3492: $\te(s,m;\alpha'(s))$.
3493: Right:
3494: the function $\alpha'(s)\mapsto\mathfrak t(s,m;\alpha'(s))$
3495: (see~\eqref{eq:def-t_s}), plotted for $m\in\{1/3,2/3,1\}$ (curves
3496: (1--3)). Here we have taken $\e=1$ and $\sin \beta(s)=1$.}
3497: \label{fig:effectcurvature}
3498: \end{figure}
3499:
3500: With this convexity we can understand why curvature is penalized.
3501: In~\eqref{eq:cost-5} the relevant expression is
3502: \[
3503: \int_0^{M(s)} \te(s,m;\alpha'(s))\, dm+ \int_0^{-M(s)} \te(s,m;\alpha'(s))\,dm
3504: \]
3505: which can be written as
3506: \[
3507: \int_0^{M(s)} \bigl[\te(s,m;\alpha'(s)) - \te(s,-m;\alpha'(s))\bigr]\, dm.
3508: \]
3509: This integral describes the transport cost for the transport along an
3510: individual ray at position $s$.
3511: Using a symmetry property of $\te$ and the strict convexity mentioned above,
3512: \begin{align*}
3513: \mathfrak t(s,m;\alpha'(s)) - \mathfrak t(s,-m;\alpha'(s))
3514: &= \mathfrak t(s,m;\alpha'(s)) + \mathfrak t(s,m;-\alpha'(s)) \\
3515: &> 2\mathfrak t(s,m;0)\\
3516: &= \mathfrak t(s,m;0) - \mathfrak t(s,-m;0).
3517: \end{align*}
3518: It is in this inequality that we see how curving the interface (or, more
3519: precisely, rotating the rays) creates a higher transport cost and
3520: therefore a larger value of the energy.
3521:
3522:
3523:
3524: \subsection{Connections to other work: continuum models and phase separation}
3525:
3526: Beyond the direct context of lipid bilayers, this work is at the
3527: intersection of a variety of different lines of research. There is of
3528: course a long-running tradition of slender-body limits in solid
3529: mechanics, resulting in a wide variety of simplified models (see e.g.\
3530: \cite{FrieseckeJamesMueller05TR} for an overview of the case of elastic
3531: plates and shells). Although our final result is rather similar---the
3532: Gamma-limiting energy $\W$ is the classical elastica bending
3533: energy---the current work is fundamentally different, in that the
3534: material at the $\e>0$ level is closer to a fluid than a solid:
3535: particles may be arbitrarily rearranged without incurring an energy
3536: penalty. Indeed, the penalization of curvature that we find in the limit
3537: is an effect of \emph{global} geometry rather than of (local)
3538: deformation. One might compare this with the difference between an
3539: elastic ball and a drop of water; both resist deformation, but for the
3540: ball this arises from local intermolecular forces, while the form of the
3541: drop results from the global effect of surface area minimization.
3542:
3543: Turning to fluid-like systems, the simple fact that the mass represented
3544: by $u$ and $v$ concentrates onto a curve is in itself remarkable. Phase
3545: separation naturally leads to penalization of interfacial length, as
3546: reflected by the term $\int \mod{\nabla u}$ in $\F_\eps$, and the distance
3547: penalization by the term $d_1(u,v)$ is necessary to prevent bulk-scale
3548: separation. To our knowledge the current work is the first example of a
3549: phase-separating system that concentrates on low-dimensional sets.
3550:
3551: Concentration onto low-dimensional sets is a commonly observed feature
3552: in many (often non-conservative, non-variational) reaction-diffusion
3553: systems. For domains with sufficient symmetry a soliton- or pulse-like
3554: solution in one dimension can be interpreted as a solution in higher
3555: dimensions that concentrates on a hyperplane. In recent years less
3556: trivial examples of low-dimensional concentration have been
3557: uncovered~\cite{DAprile00,BadialeDAprile02,AmbrosettiMalchiodiNi02,AmbrosettiMalchiodiNi03,AmbrosettiMalchiodiNi04,DoelmanVanderPloeg02,MalchiodiMontenegro02,MalchiodiMontenegro03,Malchiodi05}.
3558: Those concentrated solutions are often unstable, however, and sometimes
3559: even highly unstable
3560: (e.g.\cite{MalchiodiMontenegro02,MalchiodiMontenegro03,Malchiodi05}),
3561: thus strongly contrasting with the stable nature of the structures of
3562: this paper.
3563:
3564: %Although some of the results above are on systems with a variational
3565: %structure, for those systems the associated solutions are unstable;. To
3566: %our knowledge this work is the first example of \emph{stable}
3567: %concentration on lower-dimensional sets in the variational context.
3568:
3569: \subsection{Connections to other work: The elastica functional}
3570:
3571: Originally the elastica functional was introduced by Daniel Bernoulli and
3572: Euler as the bending energy of an elastic rod, but it has many other
3573: applications in different fields. In variational methods for image
3574: reconstruction it is widely used~\cite{Mum}. In
3575: the theory of phase separations the elastica functional appears
3576: as the two dimensional reduction of the Willmore functional
3577: \eqref{eq:willmore}.
3578:
3579: It is often natural to consider the elastica functional as acting on
3580: boundaries of subsets of the ambient space. In \cite{BMP}
3581: the lower semicontinuity of the elastica functional under
3582: $L^1$-convergence of sets is investigated; see
3583: also \cite{Sch}
3584: for a generalization to arbitrary dimensions.
3585: The Gamma-convergence of a diffuse interface approximation
3586: of the elastica, and
3587: more generally the Willmore functional, was conjectured by De Giorgi
3588: \cite{DG} and proved in a modified form in \cite{RS} for space
3589: dimensions two and three.
3590:
3591: Our approach is not restricted to curves which are
3592: boundaries of sets, and therefore open curves can be represented in this framework.
3593: A related approach can be found in \cite{BrM}, where
3594: functionals which act on phase-field approximations of `thickened
3595: curves' are considered and the Gamma-convergence of these functionals
3596: is proved in a topology based on the Haussdorf metric. The
3597: elastica functional is part of the limit, however the approximation is
3598: completely different.
3599:
3600: \subsection{Relevance to the understanding of lipid bilayers}
3601:
3602: The derivation that leads from a self-consistent mean-field theory for
3603: two-bead copolymers to the energy $\F_\eps$ (Appendix~\ref{app:derivation})
3604: contains a number of highly suspicious assumptions. The most glaring one
3605: is the severing of the head-tail bond in individual polymers: under this
3606: assumption a given head does not keep track of to \emph{which} tail it
3607: is connected, as long as it is connected to exactly one tail. The
3608: \Wasserstein distance $d_1(u,v)$ is the mathematical implementation
3609: of this assumption, and this term is all that keeps heads and tails from
3610: separating to arbitrarily large distances.
3611:
3612: The most interesting conclusion, from the point of view of applications,
3613: might be that this very weak remnant of the covalent bond
3614: suffices; that apparently \emph{very little} is needed to convert a
3615: system exhibiting bulk-scale phase separation (e.g.\ an oil-water
3616: mixture) into a system which separates at a smaller scale.
3617:
3618:
3619: \subsection{Other approaches to the upscaling limit}
3620: In the physico-chemical literature many authors have established
3621: connections between the Helfrich Hamiltonian~\pref{def:EHelfrich} on one
3622: hand and various smaller-scale models of lipid bilayers on the
3623: other~\cite{SzleiferKramerBenShaulGelbartSafran90,OversteegenBlokhuis00,
3624: LaradjiMouritsen00}. Besides being an interesting scientific issue,
3625: such a connection should also give insight in the dependence of the
3626: coefficients~$k$ and~$\overline k$ on molecular parameters; indeed the
3627: derivation given by Helfrich and others is phenomenological, based on
3628: scaling and dimensionality arguments, and therefore the coefficients are
3629: unknown.
3630:
3631:
3632: %SzleiferKramerBenShaulGelbartSafran90: berekenen de PDF (maar zeggen niet hoe)
3633: %OversteegenBlokhuis00: lattice gas model
3634: %MarkvoortVanSantenHilbers06TR: MD, relation between headgroup charge vesicle shape
3635: %LaradjiMouritsen00: direct MD simulations
3636: %ChaconSomozaTarazona98: from a density+direction model
3637:
3638: An approach that comes close to ours is that
3639: of~\cite{ChaconSomozaTarazona98}, in which the authors connect the
3640: Helfrich description with a model in which lipids are represented by a
3641: density and a vectorial quantity that may be interpreted as an alignment
3642: order parameter. As in the examples above, however, the final
3643: coefficients are again found by a numerical method
3644: (minimization of a grand potential energy). Our result here appears to
3645: be unique in giving a fully analytic formula for the bending modulus,
3646: which is not so surprising since the model contains no unscalable
3647: parameters other than $\e$.
3648:
3649: Another way in which our contribution differs from existing work is the
3650: possibility to assess the energy penalties associated with non-optimal
3651: thickness and with curve `ends'.
3652: In the current paper we only show that these penalities are larger than
3653: $O(\eps^2)$, but determining the actual scale of penalization is an
3654: interesting open question.
3655:
3656: \subsection{Generalization to $d_p(u,v)$ for $p>1$}
3657:
3658: The derivation of the energies $\F_\eps$ presented in
3659: Appendix \ref{app:derivation} suggests substituting the
3660: \Wasserstein distance by the 2-Wasserstein distance. Some of the
3661: results of this paper carry over to this situation, and in fact to the
3662: $d_p$-metric for all $1<p<\infty$. This follows from the remark that
3663: \begin{align*}
3664: d_1(u,v) &= \min \Big( \int_{\Rn\times\Rn} |x-y| \,d\gamma(x,y)\Big) \\
3665: &\leq \min \Big( \int_{\Rn\times\Rn} |x-y|^p \,d\gamma(x,y)\Big)^{1/p}
3666: \Big( \int_{\Rn\times\Rn} d\gamma(x,y)\Big)^{(p-1)/p}\\
3667: &= d_p(u,v),
3668: \end{align*}
3669: where we take total mass $M$ equal to one for simplicity. Those statements of Theorem~\ref{the:main} that only depend on the upper
3670: bound~\pref{eq:liminf-bounded} therefore carry over without change for functionals such as
3671: \begin{align*}
3672: \F^p_\e(u,v) &= \e\int\mod{\nabla u} + \frac1{\e}d_p(u,v) \\
3673: \tilde \F^p_\e(u,v) &= \e\int\mod{\nabla u} + \frac1{p\e^p}d_p(u,v)^p
3674: \end{align*}
3675: To be concrete, sequences $(u_\e,v_\e)$ along which the associated functionals
3676: $\G^p_\e$ or $\tilde\G^p_\e$ remain bounded will converge along subsequences to a system of closed $W^{2,2}$-curves.
3677:
3678: The Gamma-convergence result of this paper, however, can not be proved in this way: there is a gap between this lower bound and the upper bound that follows from a generalization of Section~\ref{sec:limsup}; this gap results from the strict convexity of the function $z\mapsto z^p$ for $p>1$. Nonetheless, we believe a similar result to be true for all $1<p<\infty$.
3679:
3680:
3681: \subsection{Generalizations}
3682: There are several other interesting directions in which further research could
3683: continue, among them
3684: \begin{itemize}
3685: \item generalizations to higher dimension and higher codimension,
3686: %\item different $p$-Wasserstein metrics and more realistic cost functions,
3687: \item characterisation of local minimizers of the energies $\F_\eps$,
3688: \item formulation of consistent evolution equations on the mesoscale,
3689: \emph{i.e.} for the densities $(u_\e,v_\e)$.
3690: \end{itemize}
3691: In higher dimensions, and for limit structures of dimension larger than
3692: one, a concept similar to the `systems of curves' will probably not be
3693: appropriate, and we expect that a
3694: varifold approach will be better suited. Some results of the present
3695: paper are
3696: easily generalized but there are also fundamental differences.
3697:
3698: %The derivation of the energies $\F_\eps$ presented in
3699: %Appendix \ref{app:derivation} suggests substituting the
3700: %\Wasserstein distance by the 2-Wasserstein distance. It would be interesting
3701: %to allow a distance which is built on a general cost
3702: %function (see the remark in Appendix \ref{app:derivation}).
3703:
3704: To investigate how close the proposed mesoscale approximation is
3705: to the Helfrich energy or to the elastica functional it is interesting to
3706: compare (local) minimizers of the energy $\F_\eps$
3707: to the variety of shapes found as local minimizers for the Helfrich
3708: energy.
3709:
3710: A natural idea for extending the mesoscale description of lipid bilayers to
3711: a time-dependent evolution is to set up a gradient flow for the energies
3712: $\F_\eps$, such as in~\cite{Fraaije93} or~\cite{BlomPeletier04}. The
3713: latter corresponds to a gradient flow based on the 2-Wasserstein
3714: distance.
3715:
3716:
3717: %===============================================
3718: % appendix
3719: %===============================================
3720: \newpage
3721: \begin{appendix}
3722: \section{Derivation of $\F_\eps$}
3723: \label{app:derivation}
3724:
3725: The functional $\F_\eps$ that is the basis of this paper arises in a
3726: simple model for a water-lipid system. In this model a lipid is
3727: represented by a two-bead chain: a head bead and a tail bead connected
3728: by a spring. The water molecules are represented by a third type of
3729: bead.
3730:
3731: The state space at the microscopic level is given by the positions
3732: $X^i_t$, $X^i_h$, and $X^j_w$ of the lipid tail, lipid head, and water
3733: beads; the lipids are numbered by $i=1,\ldots,\nl$ and the water beads
3734: by $j=1,\ldots,\nw$.
3735: Assuming that the beads are confined to a space $\Omega\subset\R^d$, the
3736: full microscopic state space for the system is then
3737: \[
3738: \X = \Omega^{2\nl+\nw}.
3739: \]
3740: Elements $X=(X_t^1,\ldots,X_t^\nl,X_h^1,\ldots,X_h^\nl,
3741: X_w^1,\ldots,X_w^\nw)\in \X$ are called microstates.
3742:
3743: We describe the system in terms of probabilities on $\X$, i.e. the
3744: (probabilistic) state $\psi$ is a probability measure on $\X$:
3745: \[
3746: \psi\in \E, \qquad \text{where} \qquad
3747: \E = \left\{\,\psi : \X \to \R^+, \quad \int_\X \psi = 1 \,\right\}.
3748: \]
3749: We assume that neither the microstates themselves nor the
3750: measure $\psi$ can be observed at the continuum level;
3751: the observables are three derived quantities, the volume fractions
3752: of tails, of heads, and of water.
3753:
3754: For a given probability measure $\psi\in \E$, the water volume fraction
3755: $r_w(\psi):\Omega\to\R^+$ is defined by
3756: \[
3757: r_w(\psi)(x) = v \sum_{j=1}^\nw \int_{\X} \psi\, \delta(X_w^j-x) \, dX
3758: \qquad \text{for all }x\in\Omega.
3759: \]
3760: Here $v$ is the volume fraction of a single bead. The tail and head
3761: volume fractions $r_t(\psi)$ and $r_t(\psi)$ are defined similarly.
3762:
3763:
3764: To specify the behaviour of the system we introduce two free energy
3765: functionals. The `ideal' free energy $\Fid:\E\to\R$ models the effects
3766: of entropy and the interactions between beads of the same molecule;
3767: the `non-ideal' free energy $\Fnid:\E\to\R$ represents the interactions
3768: between the beads of different molecules. The total free energy is the
3769: sum of the two,
3770: \[
3771: F(\psi) = \Fid(\psi) + \Fnid(\psi).
3772: \]
3773: For $\Fid$ we assume zero temperature and postulate
3774: \[
3775: \Fid(\psi) = \int_\X \psi \Hid,
3776: \]
3777: where the function $\Hid:\X\to\R$ is the internal energy of a microstate,
3778: and the superscript `id' again refers to a restriction to interaction
3779: within a single lipid molecule. While remarking that many different
3780: choices are found in the literature, here we choose simply to implement
3781: a spring by penalizing head-tail distance,
3782: \[
3783: \Hid(X) = \frac k2\sum_{i=1}^\nl \mod{X_t^i-X_h^i}^p.
3784: \]
3785: A natural choice is $p=2$, which gives the energy of a linear
3786: spring. Even more realistic seems a general distance term that becomes
3787: infinite if $\mod{X_t^i-X_h^i}$ exceeds a certain maximal value.
3788:
3789: For the non-ideal free energy $\Fnid$ we make an important
3790: simplifying assumption: $\Fnid$ can be written as a function
3791: of only the observables,
3792: \[
3793: \Fnid(\psi)= \Fnid(r_w(\psi),r_h(\psi),r_t(\psi)).
3794: \]
3795: Typical terms in the non-ideal energy
3796: $\Fnid$ are a convolution integral, in which
3797: proximity of hydrophilic (heads and water) beads and hydrophobic
3798: tail beads is penalized:
3799: \[
3800: \mu\int_\Omega \int_\Omega
3801: \bigl(r_w(\psi)(x)+r_h(\psi)(x)\bigr)\,r_t(\psi)(y)\, \kappa(x-y) \, dxdy,
3802: \]
3803: and a compressibility term that penalizes
3804: deviation from unit total volume:
3805: \[
3806: \frac K2 \int_\Omega \bigl(r_w(\psi)(x)+r_t(\psi)(x) + r_h(\psi)(x)-1\bigr)^2\, dx.
3807: \]
3808: In these expressions we take two limits:
3809: \begin{itemize}
3810: \item In the limit $K\to\infty$, we find that the observables satisfy an
3811: incompressibility condition
3812: \begin{equation}
3813: \label{eq:incompressibility}
3814: r_t(\psi)+r_h(\psi)+ r_w(\psi) \equiv 1.
3815: \end{equation}
3816: With this condition the convolution integral above can be written as
3817: \begin{equation}
3818: \label{int:conv}
3819: \mu\int_\Omega \int_\Omega
3820: \bigl(1-r_t(\psi)(x)\bigr)\,r_t(\psi)(y)\, \kappa(x-y) \, dxdy.
3821: \end{equation}
3822: \item Replacing the fixed function $\kappa$ by a rescaled version
3823: $\kappa_\delta$, defined by
3824: \[
3825: \kappa_\delta(x) := \delta^{-d}\kappa(x/\delta),
3826: \]
3827: the integral~\pref{int:conv} Gamma-converges in the following way,
3828: \[
3829: \frac 2\delta \int_\Omega \int_\Omega
3830: \bigl(1-r_t(\psi)(x)\bigr)\,r_t(\psi)(y)\, \kappa_\delta(x-y) \, dxdy
3831: \stackrel{\delta\to0}\longrightarrow
3832: K_{1,d} \left(\int \kappa\right) F^{\mathrm{int}}\bigl(r_t(\psi)\bigr)
3833: \]
3834: where the limit interface functional $F^{\mathrm{int}}$ is given by
3835: \begin{equation}
3836: F^{\mathrm{int}}(u) = \begin{cases}
3837: \int_\Omega \mod{\nabla u}\, dx & \text{if $u\in \BV\bigl(\Omega;\{0,1\}\bigr)$}
3838: \\
3839: \infty & \text{otherwise}.
3840: \end{cases}
3841: \label{result:Davila}
3842: \end{equation}
3843: Here $K_{1,1} = 1$, $K_{1,2} = 2/\pi$, and $K_{1,3} = 1/2$
3844: \cite{AlbertiBellettiniCassandroPresutti96,Davila02}.
3845: \end{itemize}
3846:
3847: Putting the ideal and non-ideal energies together, and applying the
3848: simplifications above, we find (up to rescaling)
3849: \[
3850: F(\psi) := \Fid(\psi) + \Fnid(\psi)
3851: = \int_\X \sum_{i=1}^\nl\mod{X_t^i-X_h^i}^p\psi(X)\, dX +
3852: F^{\mathrm{int}}\bigl(r_t(\psi)\bigr).
3853: \]
3854: We now assume that we minimize $F$ under all variations of $\psi$ that conserve the observables $r_{t,h,w}(\psi)$. By a convexity argument, using the fact that the functions $r_{t,h,w}$ do not distinguish between different lipids or water molecules, it follows that all lipids are identically distributed, as are all water molecules:
3855: %If we assume that any external forcing does not distinguish between
3856: %different lipids or between different water beads (\emph{i.e.} that the
3857: %external forcing acts on the observables $r_{t,h,w}(\psi)$ rather than
3858: %directly on $\psi$), then a standard argument gives that $\psi$ is
3859: %identically distributed for all lipids and for all water molecules,
3860: %i.e.
3861: \[
3862: \psi(X) = \prod_{i=1}^\nl \psi_\ell(X^i_t,X_h^i)
3863: \prod_{j=1}^\nw \psi_w(X^j_w),
3864: \]
3865: and that the total energy becomes a function of the observables alone,
3866: by minimization over all other degrees of freedom (again we disregard
3867: rescaling)
3868: \begin{align*}
3869: &F(\rho_t,\rho_h,\rho_w) := F^{\mathrm{int}}(\rho_t) +{}\\
3870: &\qquad {}+\inf_{\psi_\ell}\Bigl\{\int_{\Omega\times\Omega}
3871: \mod{X_t-X_h}^p\psi_\ell(X_t,X_h)\, dX_tdX_h:
3872: r_t(\psi_\ell) = \rho_t, \ r_h(\psi_\ell) = \rho_h\Bigr\}
3873: \end{align*}
3874: with the remaining constraint $\rho_t+\rho_h+\rho_w \equiv 1.$
3875: Remark that the final infimum is exactly the Wasserstein distance of
3876: degree $p$, $d_p(\rho_t,\rho_h)^p$.
3877:
3878: \medskip
3879: We now recover the functional $\F_\eps$ in~\eqref{def:Fe} for $\eps=1$
3880: by defining $u:=\rho_t$, $v:=\rho_h$ and by choosing $p=1$.
3881: Combining the pure-phase condition $u(x) = \rho_t(x)\in\{0,1\}$ for all $x$~\pref{result:Davila} with the incompressbility condition~\pref{eq:incompressibility} gives $v(x)=\rho_h(x)\in[0,1]$ and $uv=0$. It is easy to see that there is nothing to be gained in letting $v$ take values in the interior $(0,1)$, and we can therefore consider $\F_1$ to be given on the set of admissible functions
3882: \begin{multline*}
3883: \K_1 := \biggl\{ (u,v)\in
3884: \BV\bigl(\R^2;\{0,1\}\bigr)\times L^1\bigl(\R^2;\{0,1\}\bigr): \\
3885: \left.\int u = \int v =M , \ uv = 0\text{ a.e.}\right\}.
3886: \end{multline*}
3887: The remaining dependence on $\eps$ is a consequence of simple rescaling.
3888:
3889: \section{Ring solutions}
3890: \label{app:ring_solutions}
3891: We consider $r_1<r_2<R<r_3<r_4$ with
3892: \begin{align}
3893: r_2^2 \,&=\, \tfrac12(R^2 + r_1^2), \qquad
3894: r_3^2 \,=\, \tfrac12(R^2 + r_4^2),\label{eq:a-r2r3}\\
3895: 2R \,&=\, r_1+r_4\label{eq:a-R}
3896: \end{align}
3897: and $u,v$ given by
3898: \begin{gather*}
3899: u\,=\, \Chi_{B_{r_3}(0)\setminus B_{r_2}(0)},\qquad v\,=\,
3900: \Chi_{B_{r_2}(0)\setminus B_{r_1}(0)} +\Chi_{B_{r_4}(0)\setminus
3901: B_{r_3}}.
3902: \end{gather*}
3903: In this setting the unique monotone optimal transport map $S$ from $u$
3904: to $v$ is radially symmetric, $S(x)=S(|x|)$ and determined
3905: by the requirement that
3906: \begin{alignat}2
3907: \int_{r_2}^s 2\pi r\,dr \,&=\, \int_{r_1}^{S(s)}2\pi
3908: r\,dr&\qquad&\text{ for }
3909: r_2<s<R,\label{eq:a-cond-S-1}\\
3910: \int_{s}^{r_3} 2\pi r\,dr \,&=\, \int_{S(s)}^{r_4}2\pi r\,dr
3911: &\qquad&\text{ for }
3912: R<s<r_3.\label{eq:a-cond-S-2}
3913: \end{alignat}
3914: Then
3915: \begin{align}
3916: d_1(u,v)\,&=\, \int |x-S(x)|u(x)\,dx\notag\\
3917: &=\, \int_{r_2}^R 2\pi r(r-S(r))\,dr + \int_R^{r_3} 2\pi
3918: r(S(r)-r)\,dr.\label{eq:a-d1-1}
3919: \end{align}
3920: To compute the first integral we introduce new coordinates
3921: \begin{gather}
3922: m(r)\,=\,\pi r^2-\pi r_2^2\quad\text{ with inverse }\quad
3923: r(m)\,=\, \big(r_2^2+\frac{1}{\pi}m\big)^{1/2}.\label{eq:a-def-m1}
3924: \end{gather}
3925: In this coordinates \eqref{eq:a-cond-S-1} implies
3926: \begin{gather}
3927: m(r)\,=\, m(S(r)) -m(r_1),\\
3928: S(r(m))\,=\, r\big(m+m(r_1)\big)\,=\,r\big(m-m(R)\big),
3929: \end{gather}
3930: where in the last equality we have used that
3931: \begin{gather*}
3932: m(r_1) + m(R)\,=\,\pi\big(r_1^2-2r_2^2+R^2\big)\,=\, 0
3933: \end{gather*}
3934: by \eqref{eq:a-r2r3}.
3935: The first integral in \eqref{eq:a-d1-1} therefore becomes
3936: \begin{align}
3937: \int_{r_2}^R 2\pi r(r-S(r))\,dr \,&=\, \int_0^{m(R)}
3938: \big(r(m)-r(m-m(R))\big)\,dm\notag\\
3939: &=\, \int_0^{m(R)} r(m)-r(-m)\,dm\notag\\
3940: &=\, \frac{2\pi}{3}\Big[\Big(r_2^2 +\frac{1}{\pi}m\Big)^{3/2}
3941: +\Big(r_2^2 -\frac{1}{\pi}m\Big)^{3/2}\Big]_0^{m(R)}\notag\\
3942: &=\, \frac{2\pi}{3}\Big[\Big(r_2^2 +\frac{1}{\pi}m(R)\Big)^{3/2}
3943: +\Big(r_2^2 -\frac{1}{\pi}m(R)\Big)^{3/2}-2r_2^3\Big].\label{eq:a-d1-2}
3944: \end{align}
3945: If we now introduce $t=(r_4-r_1)/2$ then
3946: \begin{gather*}
3947: R-r_1\,=\,r_4-R\,=\, t,\qquad m(R)\,=\, \frac{\pi}{2}(2Rt-t^2),
3948: \end{gather*}
3949: and \eqref{eq:a-d1-2} yields
3950: \begin{align}
3951: \int_{r_2}^R 2\pi r(r-S(r))\,dr \,&=
3952: \frac{2\pi}{3}
3953: \Big[R^3+(R-t)^3 -2R^3\Big(1-\frac{t}{R}+\frac{t^2}{2R^2}\Big)^{\frac{3}{2}}\Big].
3954: \label{eq:a-d1-3}
3955: \end{align}
3956: If we do the analogous calculations for the second integral in
3957: \eqref{eq:a-d1-1} we obtain that
3958: \begin{align}
3959: \int_R^{r_3} 2\pi r(S(r)-r)\,dr \,&=
3960: \frac{2\pi}{3}
3961: \Big[R^3+(R+t)^3 -2R^3\Big(1+\frac{t}{R}+\frac{t^2}{2R^2}\Big)^{\frac{3}{2}}\Big].
3962: \label{eq:a-d1-4}
3963: \end{align}
3964: which is the expression on the right-hand side of \eqref{eq:a-d1-3} with
3965: $t$ substituted by $-t$.\\
3966: We consider $R\to\infty$ with $t$ of order one and obtain by a Taylor expansion
3967: in $1/R$
3968: \begin{align*}
3969: & d_1(u,v) +\int |\nabla u|\\
3970: =\,& \frac{2\pi}{3}
3971: \Big[2R^3+(R+t)^3 +(R-t)^3
3972: -2R^3\Big(1+\frac{t}{R}+\frac{t^2}{2R^2}\Big)^{\frac{3}{2}}
3973: -2R^3\Big(1-\frac{t}{R}+\frac{t^2}{2R^2}\Big)^{\frac{3}{2}}\Big]\\
3974: & +2\pi R\Big[\Big(1+\frac{t}{R}+\frac{t^2}{2R^2}\Big)^{\frac{1}{2}}
3975: +\Big(1-\frac{t}{R}+\frac{t^2}{2R^2}\Big)^{\frac{1}{2}}\Big]\\
3976: =\,& \pi Rt^2 + \frac{\pi}{16}\frac{t^4}{R}+ 4\pi R
3977: +\frac{\pi}{2}\frac{t^2}{R} + O(R^{-3}) \\
3978: =\,& 2\pi Rt \Big[\Big(\frac{t}{2} +\frac{2}{t}\Big) +
3979: \frac{1}{4R^2} + \frac{(2-t)}{32R^2}\big(t^2+2t-4\big) + O(R^{-2})O(|t-2|) \Big]\\
3980: =\,& 2\pi Rt\Big[ 2 +\frac{1}{4}(t-2)^2 +\frac{1}{4}R^{-2}+ O(R^{-2})O(|t-2|)\Bigr],
3981: \end{align*}
3982: which is the desired asymptotic development.
3983: \end{appendix}
3984: %%%%%%%%%%%%%%%%%%%%%%%%
3985: % bibliography
3986: %%%%%%%%%%%%%%%%%%%%%%%%
3987:
3988: \bibliography{helf-lib}
3989: \bibliographystyle{plain}
3990: %
3991: \end{document}
3992: