0803.0937/as.tex
1: %-----------------%
2: % arXived version %
3: % 6 March 2008    %
4: %-----------------%
5: %
6: \documentclass{article}
7: %
8: \usepackage{amssymb,amsmath,amsthm}
9: \usepackage{epsfig}
10: %
11: \newcommand{\ie}{\emph{i.e.}}
12: \newcommand{\eg}{\emph{e.g.}}
13: \newcommand{\cf}{\emph{cf}}
14: \newcommand{\etc}{\emph{etc}}
15: \newcommand{\Real}{\mathbb{R}}
16: \newcommand{\Nat}{\mathbb{N}}
17: \newcommand{\Int}{\mathbb{Z}}
18: \newcommand{\sii}{L^2}
19: \newcommand{\sobi}{\mathop{W_0^{1,2}}\nolimits}
20: \newcommand{\Sobi}{\mathop{W^{1,2}}\nolimits}
21: \newcommand{\sobii}{\mathop{W_0^{2,2}}\nolimits}
22: \newcommand{\Sobii}{\mathop{W^{2,2}}\nolimits}
23: \newcommand{\eps}{\varepsilon}
24: \newcommand{\Hilbert}{\mathcal{H}}
25: %
26: \numberwithin{equation}{section}
27: %
28: \theoremstyle{plain}
29: \newtheorem{Theorem}{Theorem}[section]
30: \newtheorem{Lemma}[Theorem]{Lemma}
31: \theoremstyle{remark}
32: \newtheorem{Remark}[Theorem]{Remark}
33: %
34: \begin{document}
35: %
36: %%%%%%%%%%%%%%%
37: %%%% TITLE %%%%
38: %%%%%%%%%%%%%%%
39: %
40: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
41: \title{\Large\textbf{%
42: Spectrum of the Laplacian in a narrow curved strip
43: with combined Dirichlet and Neumann boundary conditions%
44: }}
45: \author{David Krej\v{c}i\v{r}\'{\i}k}
46: \date{
47: \footnotesize
48: %
49: \begin{center}
50: \emph{
51: Department of Theoretical Physics, Nuclear Physics Institute,
52: \\
53: Academy of Sciences, 250\,68 \v{R}e\v{z} near Prague, Czech Republic
54: \smallskip \\
55: \emph{E-mail:} krejcirik@ujf.cas.cz
56: }
57: \end{center}
58: %
59: 6 March 2008}
60: \maketitle
61: %
62: \begin{abstract}
63: \noindent
64: We consider the Laplacian in a domain squeezed
65: between two parallel curves in the plane,
66: subject to Dirichlet boundary conditions on one of the curves
67: and Neumann boundary conditions on the other.
68: We derive two-term asymptotics for eigenvalues
69: in the limit when the distance between the curves tends to zero.
70: The asymptotics are uniform and local in the sense that
71: the coefficients depend only on the extremal points where
72: the ratio of the curvature radii of the Neumann boundary
73: to the Dirichlet one is the biggest.
74: We also show that the asymptotics can be obtained
75: from a form of norm-resolvent convergence
76: which takes into account the width-dependence
77: of the domain of definition of the operators involved.
78: %
79: \bigskip
80: \begin{itemize}
81: %
82: \item[\textbf{MSC\,2000:}]
83: 35P15; 49R50; 58J50; 81Q15.
84: %
85: \item[\textbf{Keywords:}]
86: Laplacian in tubes;
87: Dirichlet and Neumann boundary conditions;
88: dimension reduction; norm-resolvent convergence;
89: binding effect of curvature;
90: waveguides.
91: %
92: \bigskip
93: \item[\textbf{To appear in:}]
94: ESAIM: Control, Optimisation and Calculus of Variations 
95: \\
96: \verb|http://www.esaim-cocv.org|
97: \end{itemize}
98: %
99: \end{abstract}
100: %
101: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
102: %
103: \newpage
104: %%%%%%%%%%%%%%
105: %%%% BODY %%%%
106: %%%%%%%%%%%%%%
107: %
108: %---------------------%
109: \section{Introduction}
110: %---------------------%
111: %
112: Given an open interval $I\subseteq\Real$ (bounded or unbounded),
113: let $\gamma\in C^2(\overline{I};\Real^2)$
114: be a unit-speed planar curve.
115: The derivative~$\dot{\gamma}\equiv(\dot{\gamma}^1,\dot{\gamma}^2)$
116: and $n:=(-\dot{\gamma}^2,\dot{\gamma}^1)$
117: define unit tangent and normal vector fields along~$\gamma$, respectively.
118: The curvature is defined through the Frenet-Serret formulae by
119: $\kappa:=\det(\dot{\gamma},\ddot{\gamma})$;
120: it is a bounded and uniformly continuous function on~$I$.
121: 
122: For any positive~$\eps$, we introduce a mapping~$\mathcal{L}_\eps$
123: from $\overline{I} \times [0,1]$ to~$\Real^2$ by
124: %
125: \begin{equation}\label{StripMap}
126:   \mathcal{L}_\eps(s,t) := \gamma(s) + \eps \, t \, n(s) \,.
127: \end{equation}
128: %
129: Assuming that~$\mathcal{L}_\eps$ is injective
130: and that~$\eps$ is so small that the supremum norm of~$\kappa$
131: is less than~$\eps^{-1}$,
132: $\mathcal{L}_\eps$~induces a diffeomorphism
133: and the image
134: %
135: \begin{equation}\label{strip}
136:   \Omega_\eps :=
137:   \mathcal{L}_\eps\big(I \times (0,1)\big)
138: \end{equation}
139: %
140: has a geometrical meaning
141: of an open non-self-intersecting strip,
142: contained between the parallel curves~$\gamma(I)$ and
143: $
144:   \gamma_\eps(I) := \mathcal{L}_\eps(I\times\{1\})
145: $,
146: and, if~$\partial I$ is not empty, the straight lines
147: $\mathcal{L}_\eps\big(\{\inf I\}\times(0,1)\big)$
148: and $\mathcal{L}_\eps\big(\{\sup I\}\times(0,1)\big)$.
149: The geometry is set in such a way that $\kappa>0$
150: implies that the parallel curve~$\gamma_\eps$
151: is ``locally shorter'' than~$\gamma$, and vice versa,
152: \cf~Figure~\ref{figure}.
153: 
154: Let~$-\Delta_{DN}^{\Omega_\eps}$ be the Laplacian in $\sii(\Omega_\eps)$
155: with Dirichlet and Neumann boundary conditions
156: on~$\gamma$ and~$\gamma_\eps$, respectively.
157: If~$\partial I$ is not empty, we impose Dirichlet
158: boundary conditions on the remaining parts of~$\partial\Omega_\eps$.
159: 
160: For any self-adjoint operator~$H$ which is bounded from below,
161: we denote by $\{\lambda_j(H)\}_{j=1}^\infty$
162: the non-decreasing sequence of numbers corresponding to
163: the spectral problem of~$H$ according to
164: the Rayleigh-Ritz variational formula~\cite[Sec.~4.5]{Davies}.
165: Each~$\lambda_j(H)$ represents either a (discrete) eigenvalue
166: (repeated according to multiplicity) below the essential spectrum
167: or the threshold of the essential spectrum of~$H$.
168: All the eigenvalues below the essential spectrum
169: may be characterized by this variational/minimax principle.
170: 
171: Under the above assumptions,
172: our main result reads as follows:
173: %
174: \begin{Theorem}\label{Thm.mine}
175: For all $j \geq 1$,
176: %
177: \begin{equation}\label{expansion}
178:   \lambda_j(-\Delta_{DN}^{\Omega_\eps})
179:   \ = \ \left(\frac{\pi}{2\eps}\right)^2
180:   + \frac{\inf\kappa}{\eps}
181:   + o(\eps^{-1})
182:   \qquad\mbox{as}\qquad
183:   \eps \to 0
184:   \,.
185: \end{equation}
186: %
187: \end{Theorem}
188: %
189: 
190: %
191: \begin{figure}[t]
192: %
193: \begin{center}
194: \epsfig{file=figure.eps,width=0.8\textwidth}
195: \end{center}
196: %
197: \caption{
198: The geometry of the strip $\Omega_\eps$ in a case of bounded~$I$.
199: The thick and thin lines correspond to Dirichlet
200: and Neumann boundary conditions, respectively.
201: }
202: \label{figure}
203: \end{figure}
204: %
205: 
206: Theorem~\ref{Thm.mine} has important consequences
207: for the spectral properties of the operator $-\Delta_{DN}^{\Omega_\eps}$,
208: especially in the physically interesting situation $I=\Real$.
209: In this case, assuming that the curvature~$\kappa$ vanishes at infinity,
210: the leading term $\pi^2/(2\eps)^2$ of~\eqref{expansion}
211: coincides with the threshold of the essential spectrum of
212: $-\Delta_{DN}^{\Omega_\eps}$.
213: The next term in the expansion then tells us that
214: %
215: \begin{enumerate}
216: \item[(a)]
217: the discrete spectrum exists
218: whenever~$\kappa$ assumes a negative value
219: and~$\eps$ is sufficiently small;
220: \item[(b)]
221: the number of the eigenvalues increases to infinity
222: as $\eps \to 0$.
223: \end{enumerate}
224: %
225: This provides an insight into the mechanism
226: which is behind the qualitative results
227: obtained by Dittrich and K\v{r}{\'\i}\v{z}
228: in their 2002 letter~\cite{DKriz2}.
229: Using $-\Delta_{DN}^{\Omega_\eps}$ as a model
230: for the Hamiltonian of a quantum waveguide,
231: they show that the discrete eigenvalues exist if, and only if,
232: the reference curve~$\gamma$ of sign-definite~$\kappa$
233: is curved ``in the right direction'',
234: namely if the Neumann boundary condition is imposed
235: on the ``locally longer'' boundary
236: (\ie~$\kappa<0$ in our setting),
237: and that~(b) holds.
238: The results were
239: further generalized in \cite{KKriz,FK3},
240: numerically tested in~\cite{OM},
241: and established in a different physical model in~\cite{JLP}.
242: 
243: The purely Dirichlet or Neumann strips
244: differ from the present situation in many respects
245: (see~\cite{KKriz} for a comparison).
246: The case of the Neumann Laplacian~$-\Delta_{N}^{\Omega_\eps}$
247: is trivial in the sense that
248: $$
249:   \lambda_1(-\Delta_{N}^{\Omega_\eps})
250:   = \lambda_1(-\Delta_N^I)
251:   = 0
252:   \,,
253: $$
254: independently of the geometry and smallness of~$\eps$,
255: where $-\Delta_{N}^I$ denotes the Neumann Laplacian in $\sii(I)$.
256: For $j \geq 2$, one has
257: $$
258:   \lambda_j(-\Delta_{N}^{\Omega_\eps})
259:   = \lambda_j(-\Delta_N^I) + o(1)
260:   \qquad\mbox{as}\qquad
261:   \eps \to 0
262:   \,,
263: $$
264: independently of the geometry.
265: More generally, it is well known that the spectrum
266: of the Neumann Laplacian on an $\eps$-tubular neighbourhood
267: of a Riemannian manifold converges when $\eps \to 0$
268: to the spectrum of the Laplace-Beltrami operator
269: on the manifold~\cite{Schatzman_1996}.
270: 
271: As for the Dirichlet Laplacian~$-\Delta_{D}^{\Omega_\eps}$,
272: it is well known~\cite{ES,GJ,DE,KKriz}
273: that the existence of discrete spectrum
274: in unbounded strips is robust,
275: \ie\ independent of the sign of~$\kappa$.
276: This is also reflected in the asymptotic formula, $j \geq 1$,
277: %
278: \begin{equation}\label{Dirichlet}
279:   \lambda_j(-\Delta_{D}^{\Omega_\eps})
280:   = \left(\frac{\pi}{\eps}\right)^2
281:   + \lambda_j\big(-\Delta_D^I - \frac{\kappa^2}{4}\big)
282:   + o(1)
283:   \qquad\mbox{as}\qquad
284:   \eps \to 0
285:   \,,
286: \end{equation}
287: %
288: known for many years \cite{Karp-Pinsky_1988,DE},
289: where $-\Delta_{D}^I$ denotes the Dirichlet Laplacian in $\sii(I)$.
290: That is, contrary to Theorem~\ref{Thm.mine},
291: in the purely Dirichlet case
292: the second term in the asymptotic expansion is independent of~$\eps$,
293: always negative unless~$\gamma$ is a straight line,
294: and its value is determined by the \emph{global} geometry of~$\gamma$.
295: 
296: The local character of~\eqref{expansion} rather resembles
297: the problem of a straight narrow strip of \emph{variable} width
298: studied recently by Friedlander and Solomyak
299: \cite{Friedlander-Solomyak_2007a,Friedlander-Solomyak_2007b},
300: and also by Borisov and Freitas~\cite{BF}
301: -- see also~\cite{F} for related work.
302: In view of their asymptotics,
303: the spectrum of the Dirichlet Laplacian
304: is basically determined by the points
305: where the strip is the widest.
306: In our model the cross-section is uniform
307: but the curvature \emph{and} boundary conditions
308: are not homogeneous.
309: 
310: The purely Dirichlet case with uniform cross-section
311: differs from the present situation
312: also in the direct method of the proof of~\eqref{Dirichlet}.
313: Using the parametrization~\eqref{StripMap},
314: the spectral problem for the Laplacian $-\Delta_{D}^{\Omega_\eps}$
315: in the ``curved'' and $\eps$-dependent Hilbert space $\sii(\Omega_\eps)$
316: is transferred to the spectral problem
317: for a more complicated operator~$H_\eps^D$
318: in $\sii\big(I\times(0,1)\big)$.
319: Inspecting the dependence of the coefficients of~$H_\eps^D$ on~$\eps$,
320: it turns out that the operator is in the limit $\eps \to 0$
321: decoupled into a sum of the ``transverse'' Laplacian
322: multiplied by~$\eps^{-2}$
323: and of the $\eps$-independent Schr\"odinger operator on~$\gamma$.
324: At this stage, the minimax principle is sufficient
325: to establish~\eqref{Dirichlet}.
326: Furthermore, since the ``straightened'' Hilbert space is independent of~$\eps$,
327: it is also possible to show that~\eqref{Dirichlet} is
328: obtained as a consequence of
329: some sort of norm-resolvent convergence \cite{DE,FK4}.
330: An alternative approach is based on the $\Gamma$-convergence method~\cite{BMT}.
331: See also~\cite{Grieser} for a recent survey of 
332: the thin-limit problem in a wider context.
333: 
334: The above procedure does not work in the present situation
335: because the transformed operator~$H_\eps^{DN}$
336: does not decouple as $\eps \to 0$,
337: at least at the stage of the elementary usage of the minimax principle.
338: Moreover, the operator domain of~$H_\eps^{DN}$
339: becomes dependent on~$\eps$;
340: contrary to the Dirichlet boundary condition,
341: the Neumann one is transferred
342: to an $\eps$-dependent and variable Robin-type boundary condition
343: (\cf~Remark~\ref{Rem.domain} below).
344: In this paper we propose an alternative approach,
345: which enables us to treat the case of combined boundary conditions.
346: Our method of proof is based on refined applications
347: of the minimax principle.
348: 
349: In the following Section~\ref{Sec.proof},
350: we prove Theorem~\ref{Thm.mine} as a consequence
351: of upper and lower bounds to $\lambda_j(-\Delta_{DN}^{\Omega_\eps})$.
352: More specifically, these estimates imply
353: %
354: \begin{Theorem}\label{Thm.stronger}
355: For all $j \geq 1$,
356: %
357: \begin{equation}\label{stronger}
358:   \lambda_j(-\Delta_{DN}^{\Omega_\eps})
359:   = \left(\frac{\pi}{2\eps}\right)^2
360:   + \lambda_j\big(-\Delta_D^I + \frac{\kappa}{\eps}\big)
361:   + \mathcal{O}(1)
362:   \qquad\mbox{as}\qquad
363:   \eps \to 0
364:   \,.
365: \end{equation}
366: %
367: \end{Theorem}
368: %
369: \noindent
370: Then Theorem~\ref{Thm.mine} follows at once as
371: a weaker version of Theorem~\ref{Thm.stronger},
372: by using known results about the strong-coupling/semiclassical asymptotics
373: of eigenvalues of the one-dimensional Schr\"odinger operator.
374: Indeed, for all $j \geq 1$, one has
375: %
376: \begin{equation}\label{strong}
377:   \lambda_j\big(-\Delta_D^I + \frac{\kappa}{\eps}\big)
378:   = \frac{\inf\kappa}{\eps} + o(\eps^{-1})
379:   \qquad\mbox{as}\qquad
380:   \eps \to 0
381:   \,.
382: \end{equation}
383: %
384: This result seems to be well known;
385: we refer to~\cite[App.~A]{FK1} for a proof in any dimension.
386: 
387: Another goal of the present paper is to show
388: that the eigenvalue convergence of Theorem~\ref{Thm.mine}
389: can be obtained as a consequence of the norm-resolvent ``convergence''
390: of~$-\Delta_{DN}^{\Omega_\eps}$ to $-\Delta_D^I+\kappa/\eps$
391: as $\eps \to 0$.
392: We use the quotation marks because the latter operator
393: is in fact $\eps$-dependent and the norm-resolvent convergence
394: should be rather interpreted as the convergence
395: of the difference of corresponding resolvent operators in norm.
396: However, the operators act in different Hilbert spaces
397: and the norm-resolvent convergence still requires
398: a meaningful reinterpretation.
399: Because of the technical complexity,
400: we postpone the statement of this convergence result
401: until Section~\ref{Sec.norm}.
402: 
403: The paper is concluded by Section~\ref{Sec.end}
404: in which we discuss possible extensions
405: of our main results.
406: 
407: %-----------------------------%
408: \section{Spectral convergence}\label{Sec.proof}
409: %-----------------------------%
410: %
411: In this section we give a simple proof of Theorem~\ref{Thm.stronger}
412: by establishing upper and lower bounds to $\lambda_j(-\Delta_{DN}^{\Omega_\eps})$.
413: We begin with necessary geometric preliminaries.
414: 
415: %-----------------------------------%
416: \subsection{Curvilinear coordinates}\label{Sec.coord}
417: %-----------------------------------%
418: %
419: As usual, the Laplacian $-\Delta_{DN}^{\Omega_\eps}$
420: is introduced as the self-adjoint operator in $\sii(\Omega_\eps)$
421: associated with the quadratic form~$Q_{DN}^{\Omega_\eps}$
422: defined by
423: %
424: \begin{align*}
425:   Q_{DN}^{\Omega_\eps}[\Psi]
426:   &:= \int_{\Omega_\eps} |\nabla\Psi(x)|^2 \, dx
427:   \,,
428:   \\
429:   \Psi \in D(Q_{DN}^{\Omega_\eps}) &:=
430:   \left\{
431:   \Psi \in W^{1,2}(\Omega_\eps) \ | \quad
432:   \Psi = 0 \quad \mbox{on} \quad
433:   \partial\Omega_\eps \setminus \gamma_\eps(I)
434:   \right\}
435:   \,.
436: \end{align*}
437: %
438: Here $\Psi$ on~$\partial\Omega_\eps$
439: is understood in the sense of traces.
440: It is natural to express the Laplacian
441: in the ``coordinates'' $(s,t)$ determined
442: by the inverse of~$\mathcal{L}_\eps$.
443: 
444: As stated in Introduction,
445: under the hypotheses that~$\mathcal{L}_\eps$ is injective and
446: %
447: \begin{equation}\label{Ass.basic}
448:   \eps \,\sup|\kappa| < 1 \,,
449: \end{equation}
450: %
451: the mapping~\eqref{StripMap} induces a global diffeomorphism
452: between $I\times(0,1)$ and~$\Omega_\eps$.
453: This is readily seen by the inverse function theorem
454: and the expression for the Jacobian
455: $
456:   \det(\partial_1\mathcal{L}_\eps,\partial_2\mathcal{L}_\eps)
457:   =\eps h_\eps
458: $
459: of~$\mathcal{L}_\eps$, where
460: %
461: \begin{equation}\label{Jacobian}
462:   h_\eps(s,t) := 1 - \kappa(s) \, \eps \, t
463:   \,.
464: \end{equation}
465: %
466: In fact, \eqref{Ass.basic}~yields the uniform estimates
467: %
468: \begin{equation}\label{uniform}
469:   0 <
470:   1 - \eps \sup\kappa
471:   \leq h_\eps \leq
472:   1 - \eps \inf\kappa
473:   < \infty
474:   \,,
475: \end{equation}
476: %
477: where the lower bound ensures that
478: the Jacobian never vanishes in $\overline{I}\times[0,1]$.
479: 
480: The passage to the natural coordinates
481: (together with a simple scaling)
482: is then performed via the unitary transformation
483: %
484: \begin{equation*}
485:   U_\eps : \sii(\Omega_\eps) \to
486:   \Hilbert_\eps := \sii\big(I\times(0,1),h_\eps(s,t)\,ds\,dt\big):
487:   \left\{\Psi\mapsto \sqrt{\eps} \ \Psi\circ\mathcal{L}_\eps\right\}
488:   \,.
489: \end{equation*}
490: %
491: This leads to a unitarily equivalent operator
492: $H_\eps:=U_\eps(-\Delta_{DN}^{\Omega_\eps})U_\eps^{-1}$ in~$\Hilbert_\eps$,
493: which is associated with the quadratic form~$Q_\eps$ defined by
494: %
495: \begin{align*}
496:   Q_\eps[\psi] &:=
497:   \int_{I\times(0,1)} \frac{|\partial_1\psi(s,t)|^2}{h_\eps(s,t)} \, ds\,dt
498:   + \int_{I\times(0,1)} \frac{|\partial_2\psi(s,t)|^2}{\eps^2}
499:   \, h_\eps(s,t) \, ds\,dt
500:   \,,
501:   \\
502:   \psi \in D(Q_\eps) &:=
503:   \left\{
504:   \psi \in W^{1,2}\big(I\times(0,1)\big) \ | \
505:   \psi = 0 \quad \mbox{on} \quad
506:   \partial\big(I\times(0,1)\big) \setminus \big(I\times\{1\}\big)
507:   \right\}
508:   .
509: \end{align*}
510: %
511: 
512: As a consequence of~\eqref{uniform},
513: $\Hilbert_\eps$ and $\sii\big(I\times(0,1)\big)$
514: can be identified as vector spaces
515: due to the equivalence of norms,
516: denoted respectively by $\|\cdot\|_\eps$ and $\|\cdot\|$
517: in the following.
518: More precisely, we have
519: %
520: \begin{equation}\label{norms}
521:   1 - \eps \sup\kappa
522:   \leq \frac{\|\psi\|_{\eps}^2}{\|\psi\|^2} \leq
523:   1 - \eps \inf\kappa
524:   \,.
525: \end{equation}
526: %
527: That is, the fraction of norms
528: behaves as $1+\mathcal{O}(\eps)$ as $\eps \to 0$.
529: 
530: %-----------------------%
531: \subsection{Upper bound}
532: %-----------------------%
533: %
534: Let~$\psi$ be a test function from the domain $D(Q_\eps)$
535: of the form
536: %
537: \begin{equation}\label{chi}
538:   \psi(s,t) := \varphi(s) \chi_1(t) \,,
539:   \qquad\mbox{where}\qquad
540:   \chi_1(t):=\sqrt{2} \sin\left(\pi t/2\right)
541: \end{equation}
542: %
543: and~$\varphi \in \sobi(I)$ is arbitrary.
544: Note that~$\chi_1$ is a normalized eigenfunction
545: corresponding to the lowest eigenvalue of $-\Delta_{DN}^{(0,1)}$,
546: \ie\ the Laplacian in $\sii((0,1))$,
547: subject to the Dirichlet and Neumann boundary condition
548: at~$0$ and~$1$, respectively.
549: A straightforward calculation yields
550: $$
551:   Q_\eps[\psi] - \left(\frac{\pi}{2\eps}\right)^2 \|\psi\|_\eps^2
552:   = \int_I
553:   \left(
554:   a_\eps(s)\,|\varphi'(s)|^2 + \frac{\kappa(s)}{\eps}\,|\varphi(s)|^2
555:   \right)
556:   ds
557:   \,,
558: $$
559: where
560: $$
561:   a_\eps(s) := \int_0^1 \frac{|\chi_1(t)|^2}{h_\eps(s,t)} \, dt
562:   \,.
563: $$
564: Note that $\sup a_\eps = 1+\mathcal{O}(\eps)$
565: due to~\eqref{uniform} and the normalization of~$\chi_1$.
566: Using in addition the boundedness of~$\kappa$
567: and
568: $
569:   \|\varphi\|_{\sii(I)} = \|\psi\|
570: $
571: together with~\eqref{norms}, we can therefore write
572: $$
573: \displaystyle
574:   \frac{Q_\eps[\psi]}{\,\|\psi\|_\eps^2}
575:   - \left(\frac{\pi}{2\eps}\right)^2
576:   - \mathcal{O}(1)
577:   \, \leq \,
578:   \big[1+\mathcal{O}(\eps)\big] \,
579:   \frac{
580:   \int_I
581:   \left(
582:   |\varphi'(s)|^2 + \frac{\kappa(s)}{\eps}\,|\varphi(s)|^2
583:   \right)
584:   ds
585:   }
586:   {\int_I |\varphi(s)|^2 \, ds}
587:   \,.
588: $$
589: From this inequality,
590: the minimax principle gives the upper bound
591: %
592: \begin{align}\label{upper}
593:   \lambda_j(H_\eps)
594:   - \left(\frac{\pi}{2\eps}\right)^2
595:   & \, \leq \, \big[1+\mathcal{O}(\eps)\big] \,
596:   \lambda_j\big(-\Delta_D^I + \frac{\kappa}{\eps}\big)
597:   + \mathcal{O}(1)
598:   \nonumber \\
599:   & \, = \, \lambda_j\big(-\Delta_D^I + \frac{\kappa}{\eps}\big)
600:   + \mathcal{O}(1)
601:   \qquad\mbox{as}\qquad
602:   \eps \to 0 \
603: \end{align}
604: %
605: for all $j \geq 1$.
606: Here the equality follows by~\eqref{strong}.
607: 
608: %-----------------------%
609: \subsection{Lower bound}\label{Sec.lower}
610: %-----------------------%
611: %
612: For all $\psi \in D(Q_\eps)$,
613: we have
614: $$
615:   Q_\eps[\psi] \geq
616:   \int_{I\times(0,1)} \frac{|\partial_1\psi(s,t)|^2}{h_\eps(s,t)} \, ds\,dt
617:   + \int_{I\times(0,1)} \frac{\nu\big(\eps\kappa(s)\big)}{\eps^2}
618:   \ |\psi(s,t)|^2
619:   \, h_\eps(s,t) \, ds\,dt
620:   \,,
621: $$
622: where $\nu(\epsilon) \equiv \lambda_1(T_\epsilon)$
623: denotes the lowest eigenvalue
624: of the operator~$T_\epsilon$ in the Hilbert space
625: $\sii\big((0,1),(1-\epsilon t)dt\big)$
626: defined by
627: %
628: \begin{align*}
629:   (T_\epsilon\chi)(t)
630:   &:= -\chi''(t) + \frac{\epsilon}{1-\epsilon t} \, \chi'(t)
631:   \,,
632:   \\
633:   \chi \in D(T_\epsilon)
634:   &:= \left\{
635:   \chi \in W^{2,2}\big((0,1)\big) \ | \quad
636:   \chi(0) = \chi'(1) = 0
637:   \right\}
638:   .
639: \end{align*}
640: %
641: Note that $\nu(0)=(\pi/2)^2$ and that the corresponding
642: eigenfunction for $\epsilon=0$ can be identified with~$\chi_1$.
643: The analytic perturbation theory yields
644: %
645: \begin{equation}\label{analytic}
646:   \nu(\epsilon) = \left(\frac{\pi}{2}\right)^2 + \epsilon
647:   + \mathcal{O}\big(\epsilon^2\big)
648:   \qquad\mbox{as}\qquad
649:   \epsilon \to 0
650:   \,.
651: \end{equation}
652: %
653: Using this expansion and the boundedness of~$\kappa$,
654: we can estimate
655: $$
656:   Q_\eps[\psi] - \left(\frac{\pi}{2\eps}\right)^2 \|\psi\|_\eps^2
657:   \geq \int_{I\times(0,1)}
658:   \left(
659:   \frac{|\partial_1\psi(s,t)|^2}{1-\eps \inf\kappa}
660:   + \frac{\kappa}{\eps} \, |\psi(s,t)|^2
661:   - C \, |\psi(s,t)|^2
662:   \right)
663:   ds \, dt
664:   \,,
665: $$
666: where~$C$ is a positive constant depending uniquely
667: on $\|\kappa\|_{L^\infty(I)}$.
668: Using in addition~\eqref{norms}, we therefore get
669: $$
670: \displaystyle
671:   \frac{Q_\eps[\psi]}{\,\|\psi\|_\eps^2}
672:   - \left(\frac{\pi}{2\eps}\right)^2
673:   - \mathcal{O}(1)
674:   \, \geq \,
675:   \big[1+\mathcal{O}(\eps)\big] \,
676:   \frac{
677:   \int_{I\times(0,1)}
678:   \left(
679:   |\partial_1\psi(s,t)|^2 + \frac{\kappa(s)}{\eps}\,|\psi(s,t)|^2
680:   \right)
681:   ds
682:   }
683:   {\int_{I\times(0,1)} |\psi(s,t)|^2 \, ds}
684:   \,.
685: $$
686: Consequently, the minimax principle gives
687: %
688: \begin{align}\label{lower}
689:   \lambda_j(H_\eps)
690:   - \left(\frac{\pi}{2\eps}\right)^2
691:   & \, \geq \, \big[1+\mathcal{O}(\eps)\big] \,
692:   \lambda_j\big(-\Delta_D^I + \frac{\kappa}{\eps}\big)
693:   + \mathcal{O}(1)
694:   \nonumber \\
695:   & \, = \, \lambda_j\big(-\Delta_D^I + \frac{\kappa}{\eps}\big)
696:   + \mathcal{O}(1)
697:   \qquad\mbox{as}\qquad
698:   \eps \to 0 \
699: \end{align}
700: %
701: for all $j \geq 1$.
702: Again, here the equality follows by~\eqref{strong}.
703: 
704: In view of the unitary equivalence of~$H_\eps$
705: with $-\Delta_{DN}^{\Omega_\eps}$,
706: the estimates~\eqref{upper} and~\eqref{lower}
707: prove Theorem~\ref{Thm.stronger}.
708: 
709: %----------------------------------%
710: \section{Norm-resolvent convergence}\label{Sec.norm}
711: %----------------------------------%
712: %
713: In this section we study the mechanism which is behind
714: the eigenvalue convergence of Theorem~\ref{Thm.mine} in more details.
715: First we explain what we mean by the norm-resolvent convergence
716: of the family of operators $\{-\Delta_{DN}^{\Omega_\eps}\}_{\eps>0}$.
717: 
718: %------------------------------------------------------%
719: \subsection{The reference Hilbert space and the result}
720: %------------------------------------------------------%
721: %
722: In Section~\ref{Sec.coord},
723: we identified the Laplacian $-\Delta_{DN}^{\Omega_\eps}$
724: with a Laplace-Beltrami-type operator~$H_\eps$ in~$\Hilbert_\eps$.
725: Now it is more convenient to pass to another
726: unitarily equivalent operator~$\hat{H}_\eps$
727: which acts in the ``fixed'' (\ie~$\eps$-independent) Hilbert space
728: $$
729:   \Hilbert_0 := \sii\big(I\times(0,1)\big)
730:   \,.
731: $$
732: This is enabled by means of the unitary mapping
733: %
734: \begin{equation*}
735:   \hat{U}_\eps : \Hilbert_\eps \to \Hilbert_0:
736:   \big\{\psi\mapsto \sqrt{h_\eps} \ \psi\big\}
737:   \,,
738: \end{equation*}
739: %
740: provided that the curvature~$\kappa$
741: is differentiable in a weak sense;
742: henceforth we assume that
743: %
744: \begin{equation}\label{Ass.derivative}
745:   \kappa' \in L^\infty(I)
746:   \,.
747: \end{equation}
748: %
749: We set $\hat{H}_\eps:=\hat{U}_\eps H_\eps \hat{U}_\eps^{-1}$.
750: As a comparison operator to $\hat{H}_\eps$ for small~$\eps$,
751: we consider the decoupled operator
752: $$
753:   \hat{H}_0 :=
754:   \left(-\Delta_D^I + \frac{\kappa}{\eps}\right) \otimes 1
755:   + 1 \otimes \Big(-\frac{1}{\eps^{2}}\,\Delta_{DN}^{(0,1)}\Big)
756:   \qquad\mbox{in}\qquad
757:   \sii(I)\otimes\sii\big((0,1)\big)
758:   \,.
759: $$
760: Here the subscript~$0$ is just a notational convention, of course,
761: since~$\hat{H}_0$ still depends on~$\eps$.
762: Using natural isomorphisms, we may reconsider~$\hat{H}_0$
763: as an operator in~$\Hilbert_0$.
764: 
765: We clearly have
766: %
767: \begin{equation}\label{pre.lb1}
768:   \hat{H}_0 \geq \left(\frac{\pi}{2\eps}\right)^2 + \frac{\inf\kappa}{\eps}
769:   \,.
770: \end{equation}
771: %
772: At the same time,
773: %
774: \begin{equation}\label{pre.lb2}
775:   \hat{H}_\eps
776:   \geq \frac{\nu(\eps\kappa)}{\eps^2}
777:   \geq \frac{\nu(\eps\inf\kappa)}{\eps^2}
778:   = \left(\frac{\pi}{2\eps}\right)^2 + \frac{\inf\kappa}{\eps}
779:   + \mathcal{O}(1)
780:   \,,
781: \end{equation}
782: %
783: where the first inequality was established
784: (for the unitarily equivalent operator~$H_\eps$)
785: in the beginning of Section~\ref{Sec.lower},
786: the second inequality holds due to the monotonicity of
787: $\eps\mapsto\nu(\eps)$ proved in \cite[Thm.~2]{FK3}
788: and the equality follows from~\eqref{analytic}.
789: (Alternatively, we could use Theorem~\ref{Thm.mine} to get~\eqref{pre.lb2},
790: however, one motivation of the present section is to show
791: that the former can be obtained as a consequence of Theorem~\ref{Thm.norm} below.)
792: Fix any number
793: %
794: \begin{equation}\label{k}
795:   k > -\inf\kappa \,.
796: \end{equation}
797: %
798: It follows that $\hat{H}_\eps-\pi^2/(2\eps)^2+k/\eps$
799: and $\hat{H}_0-\pi^2/(2\eps)^2+k/\eps$ are positive operators
800: for all sufficiently small~$\eps$.
801: 
802: Now we are in a position to state the main result of this section.
803: %
804: \begin{Theorem}\label{Thm.norm}
805: In addition to the injectivity of~$\mathcal{L}_\eps$
806: and the boundedness of~$\kappa$, let us assume~\eqref{Ass.derivative}.
807: Then there exist positive constants~$\eps_0$ and~$C_0$,
808: depending uniquely on~$k$
809: and the supremum norms of~$\kappa$ and~$\kappa'$,
810: such that for all $\eps\in(0,\eps_0)$:
811: %
812: \begin{equation*}
813:   \left\|
814:   \left[\hat{H}_\eps-\left(\frac{\pi}{2\eps}\right)^2+\frac{k}{\eps}\right]^{-1}
815:   - \left[\hat{H}_0-\left(\frac{\pi}{2\eps}\right)^2+\frac{k}{\eps}\right]^{-1}
816:   \right\|
817:   \ \leq \
818:   C_0 \, \eps^{3/2}
819:   \,.
820: \end{equation*}
821: %
822: \end{Theorem}
823: %
824: 
825: The theorem is proved in several steps
826: divided into the following subsections.
827: In particular, it follows as a direct consequence
828: of Lemmata~\ref{Lem.inter} and~\ref{Lem.complement} below.
829: In the final subsection we show how it implies
830: the convergence of eigenvalues of Theorem~\ref{Thm.mine}.
831: 
832: %------------------------------------%
833: \subsection{The transformed Laplacian}
834: %------------------------------------%
835: %
836: Let us now find an explicit expression for the quadratic form~$\hat{Q}_\eps$
837: associated with the operator~$\hat{H}_\eps$.
838: By definition, it is given by
839: %
840: \begin{equation*}
841:   \hat{Q}_\eps[\psi] := Q_\eps[\hat{U}_\eps^{-1}\psi] \,,
842:   \qquad
843:   \psi \in D(\hat{Q}_\eps) := \hat{U}_\eps D(Q_\eps)
844:   \,.
845: \end{equation*}
846: %
847: One easily verifies that
848: %
849: \begin{equation}\label{form.domain}
850:   D(\hat{Q}_\eps) = D(Q_\eps) =: \mathcal{Q}
851:   \,,
852: \end{equation}
853: %
854: which is actually independent of~$\eps$.
855: Furthermore, for any $\psi \in \mathcal{Q}$,
856: we have
857: $$
858:   \hat{Q}_\eps[\psi] = \hat{Q}_\eps^1[\psi]+\hat{Q}_\eps^2[\psi]
859:   \,,
860: $$
861: where
862: %
863: \begin{align*}
864:   \hat{Q}_\eps^1[\psi]
865:   &:=
866:   \int
867:   \frac{\big|\partial_1 (h_\eps^{-1/2}\psi)\big|^2}{h_\eps}
868:   &=&
869:   \int \left\{
870:   \frac{|\partial_1\psi|^2}{h_\eps^2}
871:   + V_\eps^1 \, |\psi|^2
872:   + V_\eps^2 \, \Re\big(\overline{\psi}\partial_1\psi\big)
873:   \right\} ,
874:   \\
875:   \hat{Q}_\eps^2[\psi]
876:   &:=
877:   \int \frac{\big|\partial_2(h_\eps^{-1/2}\psi)\big|^2}{\eps^2} \ h_\eps
878:   &=& \int \left\{
879:   \frac{|\partial_2\psi|^2}{\eps^2}
880:   + V_\eps^3 \, |\psi|^2
881:   + V_\eps^4  \, \Re\big(\overline{\psi}\partial_2\psi\big)
882:   \right\} ,
883: \end{align*}
884: %
885: with
886: %
887: \begin{align*}
888:  V_\eps^1(s,t) &:=  \frac{1}{4} \frac{\kappa'(s)^2 \eps^2 t^2}{h_\eps(s,t)^4} \,,
889:  &
890:  V_\eps^2(s,t) &:= \frac{\kappa'(s) \eps t}{h_\eps(s,t)^3} \,,
891:  \\
892:  V_\eps^3(s,t) &:= \frac{1}{4} \frac{\kappa(s)^2}{h_\eps(s,t)^2} \,,
893:  &
894:  V_\eps^4(s,t) &:= \frac{\kappa(s)}{\eps h_\eps(s,t)} \,.
895: \end{align*}
896: %
897: Here and in the sequel the integral sign~$\int$ refers
898: to an integration over $I\times(0,1)$.
899: Integrating by parts in the expression for~$\hat{Q}_\eps^2[\psi]$,
900: we finally arrive at
901: $$
902:   \hat{Q}_\eps[\psi] =
903:   \int \left\{
904:   \frac{|\partial_1\psi|^2}{h_\eps^2}
905:   + \frac{|\partial_2\psi|^2}{\eps^2}
906:   + (V_\eps^1-V_\eps^3)  |\psi|^2
907:   + V_\eps^2 \, \Re\big(\overline{\psi}\partial_1\psi\big)
908:   \right\}
909:   + \int_\partial v_\eps \, |\psi|^2
910:   \,,
911: $$
912: where
913: $$
914:   v_\eps(s,t) := \frac{1}{2} \frac{\kappa(s)}{\eps \big(1-\eps\kappa(s)\big)}
915:   \,.
916: $$
917: Here and in the sequel the integral sign~$\int_\partial$ refers
918: to an integration over the boundary $I\times\{1\}$.
919: %
920: \begin{Remark}\label{Rem.domain}
921: $\hat{H}_\eps$ is exactly
922: the operator $H_\eps^{DN}$ mentioned briefly in Introduction.
923: Let us remark in this context that,
924: contrary to the form domains~\eqref{form.domain},
925: the operator domains of~$H_\eps$ and~$\hat{H}_\eps$ do differ
926: (unless the curvature~$\kappa$ vanishes identically).
927: Indeed, under additional regularity conditions about~$\gamma$,
928: it can be shown that while functions from $D(H_\eps)$
929: satisfy Neumann boundary conditions on $I\times\{1\}$,
930: the functions~$\psi$ from $D(\hat{H}_\eps)$ satisfy
931: non-homogeneous Robin-type boundary conditions
932: $
933:   \partial_2\psi + \eps^2 v_\eps \psi = 0
934: $
935: on $I\times\{1\}$.
936: This is the reason why the decoupling of~$\hat{H}_\eps$
937: for small~$\eps$ is not obvious in this situation.
938: At the same time, we see that the operator domain of~$\hat{H}_\eps$
939: heavily depends on the geometry of~$\gamma$.
940: For our purposes, however, it will be enough to work
941: with the associated quadratic form~$\hat{Q}_\eps$
942: whose domain is independent of~$\eps$ and~$\kappa$.
943: \end{Remark}
944: %
945: 
946: %-------------------------------------------------------%
947: \subsection{Renormalized operators and resolvent bounds}
948: %-------------------------------------------------------%
949: %
950: It will be more convenient to work with the shifted operators
951: $$
952:   L_\eps := \hat{H}_\eps-\left(\frac{\pi}{2\eps}\right)^2+\frac{k}{\eps}
953:   \qquad\mbox{and}\qquad
954:   L_0 := \hat{H}_0-\left(\frac{\pi}{2\eps}\right)^2+\frac{k}{\eps}
955:   \,.
956: $$
957: Let~$l_\eps$ and~$l_0$ denote the associated quadratic forms.
958: It is important that they have the same domain~$\mathcal{Q}$.
959: More precisely, $\hat{H}_0$~was initially defined as a direct sum,
960: however, using natural isomorphisms, it is clear that
961: we can identify the form domain of~$L_0$ with~$\mathcal{Q}$ and
962: $$
963:   l_0[\psi] =
964:   \int \left\{
965:   |\partial_1\psi|^2
966:   + \frac{1}{\eps^2}
967:   \Big[
968:   |\partial_2\psi|^2-\left(\frac{\pi}{2}\right)^2|\psi|^2
969:   \Big]
970:   + \frac{k+\kappa}{\eps} \, |\psi|^2
971:   \right\}
972: $$
973: for all $\psi \in \mathcal{Q}$.
974: 
975: It will be also useful to have an intermediate operator~$L$,
976: obtained from~$L_\eps$ after neglecting its non-singular dependence on~$\eps$
977: but keeping the boundary term.
978: For simplicity, henceforth we assume that~$\eps$ is less than one
979: and that it is in fact so small that~\eqref{Ass.basic}
980: holds with a number less than one on the right hand side.
981: Consequently,
982: %
983: \begin{equation}\label{V-estimates}
984:   |h_\eps-1| \leq C \eps
985:   \,, \
986:   |V_\eps^1| \leq C \eps^2
987:   \,, \
988:   |V_\eps^2| \leq C \eps
989:   \,, \
990:   |V_\eps^3| \leq C
991:   \,, \
992:   |V_\eps^4| \leq C \eps^{-1}
993:   \,, \
994:   |v_\eps| \leq C \eps^{-1}
995:   \,.
996: \end{equation}
997: %
998: Here and in the sequel,
999: we use the convention that~$C$ and~$c$ are positive constants
1000: which possibly depend on~$k$ and the supremum norms of~$\kappa$ and~$\kappa'$,
1001: and which may vary from line to line.
1002: In view of these estimates, it is reasonable to introduce~$L$
1003: as the operator associated with the quadratic form~$l$
1004: defined by $D(l):=\mathcal{Q}$ and
1005: $$
1006:   l[\psi] :=
1007:   \int \left\{
1008:   |\partial_1\psi|^2
1009:   + \frac{1}{\eps^2}
1010:   \Big[
1011:   |\partial_2\psi|^2-\left(\frac{\pi}{2}\right)^2|\psi|^2
1012:   \Big]
1013:   + \frac{k}{\eps} \, |\psi|^2
1014:   \right\}
1015:   + \int_\partial v_\eps \, |\psi|^2
1016: $$
1017: for all $\psi \in \mathcal{Q}$.
1018: Indeed, it follows from~\eqref{V-estimates} that
1019: %
1020: \begin{align}\label{diff.inter}
1021:   \big|l_\eps[\psi] - l[\psi]\big|
1022:   &\leq \int \Big\{
1023:   |h_\eps^{-2}-1| |\partial_1\psi|^2
1024:   + |V_\eps^1-V_\eps^3| |\psi|^2
1025:   + |V_\eps^2| \, |\psi||\partial_1\psi|
1026:   \Big\}
1027:   \nonumber \\
1028:   &\leq C \big(
1029:   \eps \|\partial_1\psi\|^2 + \|\psi\|^2
1030:   \big)
1031: \end{align}
1032: %
1033: for all $\psi\in\mathcal{Q}$.
1034: 
1035: Let us now argue that, for every $\psi \in \mathcal{Q}$
1036: and for all sufficiently small~$\eps$
1037: (which precisely means that~$\eps$ has to be less than an explicit constant
1038: depending on~$k$ and the supremum norms of~$\kappa$ and~$\kappa'$),
1039: we have
1040: %
1041: \begin{equation}\label{lbs}
1042:   \min\left\{
1043:   l_\eps[\psi], l_0[\psi], l[\psi]
1044:   \right\}
1045:   \geq
1046:   c \,
1047:   \left(
1048:   \|\partial_1\psi\|^2 + \eps^{-1} \|\psi\|^2
1049:   \right)
1050:   \,.
1051: \end{equation}
1052: %
1053: Here the bound for~$l_0$ follows at once by improving
1054: the crude bound~\eqref{pre.lb1} and recalling~\eqref{k}.
1055: The bound for~$l$ follows from that for~$l_\eps$ and from~\eqref{diff.inter}.
1056: As for the bound for~$l_\eps$,
1057: we first remark that the estimates~\eqref{pre.lb2}
1058: actually hold for the part of~$\hat{H}_\eps$
1059: associated with~$\hat{Q}_\eps^2$.
1060: Second, using~\eqref{V-estimates} and some elementary estimates,
1061: we have
1062: $
1063:   \hat{Q}_\eps^1[\psi]
1064:   \geq (c-C\eps) \|\partial_1\psi\|^2
1065:   - C \|\psi\|^2
1066: $.
1067: Hence, for~$\eps$ small enough,
1068: we indeed conclude with the bound for~$l_\eps$.
1069: 
1070: The estimates~\eqref{lbs} imply that,
1071: for all sufficiently small~$\eps$,
1072: %
1073: \begin{equation}\label{res.bounds}
1074:   \|L_\eps^{-1}\| \leq C \eps
1075:   \,, \qquad
1076:   \|L_0^{-1}\| \leq C \eps
1077:   \,, \qquad
1078:   \|L^{-1}\| \leq C \eps
1079:   \,.
1080: \end{equation}
1081: %
1082: 
1083: %----------------------------------------------%
1084: \subsection{An intermediate convergence result}
1085: %----------------------------------------------%
1086: %
1087: As the first step in the proof of Theorem~\ref{Thm.norm},
1088: we show that it is actually enough to establish
1089: the norm-resolvent convergence for a simpler operator~$L$ instead of~$L_\eps$.
1090: %
1091: \begin{Lemma}\label{Lem.inter}
1092: Under the assumptions of Theorem~\ref{Thm.norm},
1093: there exist positive constants~$\eps_0$ and~$C_0$,
1094: depending uniquely on~$k$
1095: and the supremum norms of~$\kappa$ and~$\kappa'$,
1096: such that for all $\eps\in(0,\eps_0)$:
1097: %
1098: \begin{equation*}
1099:   \left\|
1100:   L_\eps^{-1} - L^{-1}
1101:   \right\|
1102:   \ \leq \
1103:   C_0 \, \eps^{2}
1104:   \,.
1105: \end{equation*}
1106: %
1107: \end{Lemma}
1108: %
1109: \begin{proof}
1110: We are inspired by~\cite[Sec.~3]{Friedlander-Solomyak_2007a}.
1111: Adapting the estimate~\eqref{diff.inter}
1112: for the sesquilinear form generated by $l_\eps-l$
1113: and using~\eqref{lbs}, we get
1114: %
1115: \begin{align*}
1116:   \big|l_\eps(\phi,\psi) - l(\phi,\psi)\big|
1117:   &\leq C
1118:   \sqrt{\eps \|\partial_1\phi\|^2 + \|\phi\|^2}
1119:   \sqrt{\eps \|\partial_1\psi\|^2 + \|\psi\|^2}
1120:   \\
1121:   &\leq (C/c) \, \eps \sqrt{l[\phi]\,l_\eps[\psi]}
1122: \end{align*}
1123: %
1124: for every $\phi,\psi \in \mathcal{Q}$.
1125: Choosing $\phi:=L^{-1} f$ and $\psi:=L_\eps^{-1} g$,
1126: where $f,g\in\Hilbert_0$ are arbitrary,
1127: we arrive at
1128: $$
1129:   \big|(f,L^{-1}g) - (f,L_\eps^{-1}g)\big|
1130:   \leq (C/c) \, \eps \sqrt{(f,L^{-1}f)(g,L_\eps^{-1}g)}
1131:   \leq (C^2/c) \, \eps^2 \, \|f\| \, \|g\|
1132:   \,.
1133: $$
1134: Here $(\cdot,\cdot)$ denotes the inner product in~$\Hilbert_0$
1135: and the second inequality follows from~\eqref{res.bounds}.
1136: This completes the proof with $C_0:=C^2/c$.
1137: \end{proof}
1138: %
1139: 
1140: %------------------------------------------------------------%
1141: \subsection{An orthogonal decomposition of the Hilbert space}
1142: %------------------------------------------------------------%
1143: %
1144: Contrary to Lemma~\ref{Lem.inter},
1145: the convergence of $\|L^{-1}-L_0^{-1}\|$ is less obvious.
1146: We follow the idea of~\cite{Friedlander-Solomyak_2007a}
1147: and decompose the Hilbert space~$\Hilbert_0$
1148: into an orthogonal sum
1149: $$
1150:   \Hilbert_0 = \mathfrak{H}_1 \oplus \mathfrak{H}_1^\bot
1151:   \,,
1152: $$
1153: where the subspace~$\mathfrak{H}_1$ consists of functions~$\psi_1$
1154: such that
1155: %
1156: \begin{equation}\label{psi1}
1157:   \psi_1(s,t) = \varphi_1(s) \chi_1(t)
1158:   \,.
1159: \end{equation}
1160: %
1161: Recall that~$\chi_1$ has been introduced in~\eqref{chi}.
1162: Since~$\chi_1$ is normalized, we clearly have
1163: $
1164:   \|\psi_1\|=\|\varphi_1\|_{\sii(I)}
1165: $.
1166: Given any $\psi\in\Hilbert_0$, we have the decomposition
1167: %
1168: \begin{equation}\label{psi.decomposition}
1169:   \psi = \psi_1 + \psi_\bot
1170:   \qquad\mbox{with}\qquad
1171:   \psi_1 \in \mathfrak{H}_1, \ \psi_\bot\in \mathfrak{H}_1^\bot
1172:   \,,
1173: \end{equation}
1174: %
1175: where~$\psi_1$ has the form~\eqref{psi1}
1176: with $\varphi_1(s):=\int_0^1 \psi(s,t) \chi_1(t) dt$.
1177: Note that $\psi_1\in\mathcal{Q}$ if $\psi\in\mathcal{Q}$.
1178: The inclusion $\psi_\bot\in\mathfrak{H}_1^\bot$ means that
1179: %
1180: \begin{equation}\label{orth.identity1}
1181:   \int_0^1 \psi_\bot(s,t) \, \chi_1(t) \, dt = 0
1182:   \qquad\mbox{for a.e.}\quad s \in I
1183:   \,.
1184: \end{equation}
1185: %
1186: If in addition $\psi_\bot \in \mathcal{Q}$,
1187: then one can differentiate the last identity to get
1188: %
1189: \begin{equation}\label{orth.identity2}
1190:   \int_0^1 \partial_1\psi_\bot(s,t) \, \chi_1(t) \, dt = 0
1191:   \qquad\mbox{for a.e.}\quad s \in I
1192:   \,.
1193: \end{equation}
1194: %
1195: 
1196: %----------------------------------------------%
1197: \subsection{A complementary convergence result}
1198: %----------------------------------------------%
1199: %
1200: Now we are in a position to prove the following result,
1201: which together with Lemma~\ref{Lem.inter}
1202: establishes Theorem~\ref{Thm.norm}.
1203: %
1204: \begin{Lemma}\label{Lem.complement}
1205: Under the assumptions of Theorem~\ref{Thm.norm},
1206: there exist positive constants~$\eps_0$ and~$C_0$,
1207: depending uniquely on~$k$
1208: and the supremum norms of~$\kappa$ and~$\kappa'$,
1209: such that for all $\eps\in(0,\eps_0)$:
1210: %
1211: \begin{equation*}
1212:   \left\|
1213:   L^{-1} - L_0^{-1}
1214:   \right\|
1215:   \ \leq \
1216:   C_0 \, \eps^{3/2}
1217:   \,.
1218: \end{equation*}
1219: %
1220: \end{Lemma}
1221: %
1222: \begin{proof}
1223: Again, we use some of the ideas of~\cite[Sec.~3]{Friedlander-Solomyak_2007a}.
1224: As a consequence of~\eqref{orth.identity1} and~\eqref{orth.identity2},
1225: we get that $l_0(\psi_1,\psi_\bot)=0$; therefore
1226: %
1227: \begin{equation}\label{nomixed}
1228:   l_0[\psi] = l_0[\psi_1]+l_0[\psi_\bot]
1229: \end{equation}
1230: %
1231: for every $\psi \in \mathcal{Q}$.
1232: At the same time, for sufficiently small~$\eps$,
1233: %
1234: \begin{equation}\label{lbs.bis}
1235: \begin{aligned}
1236:   l_0[\psi_1]
1237:   &\geq c
1238:   \left(
1239:   \|\varphi_1'\|_{\sii(I)}^2 + \eps^{-1} \, \|\varphi_1\|_{\sii(I)}^2
1240:   \right)
1241:   \,,
1242:   \\
1243:   l_0[\psi_\bot]
1244:   &\geq c
1245:   \Big(
1246:   \|\partial_1\psi_\bot\|^2 + \eps^{-2} \, \|\partial_2\psi_\bot\|^2
1247:   + \eps^{-2} \, \|\partial\psi_\bot\|^2
1248:   \Big)
1249:   \,,
1250: \end{aligned}
1251: \end{equation}
1252: %
1253: where the second inequality is based on
1254: $\int |\partial_2\psi_\bot|^2 \geq \pi^2 \int |\psi_\bot|^2$.
1255: 
1256: Let us now compare~$l_0$ with~$l$.
1257: For every $\psi \in \mathcal{Q}$, we define
1258: $$
1259:   m[\psi] := l[\psi]-l_0[\psi]
1260:   = \int_\partial v_\eps \, |\psi|^2
1261:   - \int \frac{\kappa}{\eps} \, |\psi|^2
1262:   \,.
1263: $$
1264: Using the estimates~\eqref{V-estimates} and~\eqref{lbs.bis},
1265: we get, for any $\psi \in \mathcal{Q}$
1266: decomposed as in~\eqref{psi.decomposition},
1267: %
1268: \begin{align*}
1269:   m[\psi_1]
1270:   &=
1271:   \int_I \frac{\kappa(s)^2}{1-\eps \kappa(s)} \, |\varphi_1(s)|^2 ds
1272:   \leq C \, \|\varphi_1\|_{\sii(\Real)}^2
1273:   \leq (C/c) \, \eps \, l_0[\psi_1]
1274:   \,,
1275:   \\
1276:   |m[\psi_\bot]|
1277:   &\leq C \eps^{-1} \Big(
1278:   \|\psi_\bot\| \|\partial_2\psi_\bot|| + \|\psi_\bot\|^2
1279:   \Big)
1280:   \leq 2 (C/c) \, \eps \, l_0[\psi_\bot]
1281:   \,,
1282:   \\
1283:   |m(\psi_1,\psi_\bot)|
1284:   &= \left| \int_\partial
1285:   v_\eps \, \overline{\psi_1} \psi_\bot
1286:   \right|
1287:   \leq C \eps^{-1} \|\psi_1\| \sqrt{\|\psi_\bot\| \|\partial_2\psi_\bot||}
1288:   \\
1289:   &\leq (C/c) \, \eps^{1/2} \sqrt{l_0[\psi_1]l_0[\psi_\bot]}
1290:   \,.
1291: \end{align*}
1292: %
1293: Except for~$m[\psi_1]$,
1294: here the boundary integral was estimated via
1295: $$
1296:   \int_\partial |\psi|^2
1297:   = \int \partial_2|\psi|^2
1298:   = \int 2\,\Re\big(\overline{\psi}\partial_2\psi\big)
1299:   \leq 2 \, \|\psi\| \, \|\partial_2\psi\|
1300:   \,.
1301: $$
1302: Taking~\eqref{nomixed} into account,
1303: we conclude with the estimate
1304: (which can be again adapted for the corresponding sesquilinear form)
1305: $$
1306:   |m[\psi]| \leq C \, \eps^{1/2} \ l_0[\psi]
1307: $$
1308: valid for every $\psi \in \mathcal{Q}$
1309: and all sufficiently small~$\eps$.
1310: In particular, this implies the crude estimates
1311: $$
1312:   c \, l_0[\psi] \leq l[\psi] \leq C \, l_0[\psi]
1313:   \,.
1314: $$
1315: 
1316: Summing up, we have the crucial bound
1317: $$
1318:   \big| l(\phi,\psi)-l_0(\phi,\psi) \big|
1319:   \leq C \, \eps^{1/2} \, \sqrt{l_0[\phi] \, l[\psi]}
1320:   \,,
1321: $$
1322: valid for arbitrary $\phi,\psi \in \mathcal{Q}$.
1323: The rest of the proof then follows the lines
1324: of the proof of Lemma~\ref{Lem.inter}.
1325: \end{proof}
1326: %
1327: 
1328: %--------------------------------------%
1329: \subsection{Convergence of eigenvalues}
1330: %--------------------------------------%
1331: %
1332: As an application of Theorem~\ref{Thm.norm},
1333: we shall show now how it implies the eigenvalue asymptotics
1334: of Theorem~\ref{Thm.mine}.
1335: Recall that the numbers $\lambda_j(H)$ represent
1336: either eigenvalues below the essential spectrum
1337: or the threshold of the essential spectrum of~$H$.
1338: In particular, they provide a complete information
1339: about the spectrum of~$H$ if it is an operator with compact resolvent.
1340: In our situation, this will be the case if~$I$ is bounded,
1341: but let us stress that we allow infinite or semi-infinite intervals, too.
1342: 
1343: We begin with analysing the spectrum of the comparison operator.
1344: %
1345: \begin{Lemma}\label{Lem.comparison}
1346: Let~$\kappa$ be bounded.
1347: One has
1348: $$
1349:   \lambda_1(\hat{H}_0)
1350:   = \lambda_1\big(-\Delta_D^I+\frac{\kappa}{\eps}\big)
1351:   + \left(\frac{\pi}{2\eps}\right)^2
1352:   \,.
1353: $$
1354: Moreover, for any integer $N \geq 2$,
1355: there exists a positive constant~$\eps_0$
1356: depending on~$N$, $\kappa$ and~$I$
1357: such that for all $\eps < \eps_0$:
1358: $$
1359:   \forall j \in\{1,\dots,N\}, \qquad
1360:   \lambda_j(\hat{H}_0)
1361:   = \lambda_j\big(-\Delta_D^I+\frac{\kappa}{\eps}\big)
1362:   + \left(\frac{\pi}{2\eps}\right)^2
1363:   \,.
1364: $$
1365: \end{Lemma}
1366: %
1367: \begin{proof}
1368: Since~$\hat{H}_0$ is decoupled,
1369: we know that (\cf~\cite[Corol.\ of Thm.~VIII.33]{RS1})
1370: %
1371: \begin{equation*}
1372:   \left\{\lambda_j(\hat{H}_0)\right\}_{j=1}^\infty
1373:   = \left\{\lambda_j\big(-\Delta_D^I+\frac{\kappa}{\eps}\big)\right\}_{j=1}^\infty
1374:   + \left\{ \left(\frac{j\pi}{2\eps}\right)^2 \right\}_{j=1}^\infty
1375:   \,,
1376: \end{equation*}
1377: %
1378: and it only remains to arrange the sum of the numbers
1379: on the right hand side into a non-decreasing sequence.
1380: The assertion for $N=1$ is therefore trivial.
1381: Let $j \geq 2$ and assume by induction that
1382: $
1383:   \lambda_{j-1}(\hat{H}_0) - \pi^2/(2\eps)^2
1384:   = \lambda_{j-1}(-\Delta_D^I+\kappa/\eps)
1385: $.
1386: Then
1387: $$
1388:   \lambda_{j}(\hat{H}_0) - \pi^2/(2\eps)^2
1389:   = \min\left\{
1390:   \lambda_{j-1}(-\Delta_D^I+\kappa/\eps),
1391:   3\pi^2/(2\eps)^2
1392:   \right\}
1393: $$
1394: and the assertion of Lemma follows at once
1395: due to the asymptotics~\eqref{strong}.
1396: \end{proof}
1397: %
1398: 
1399: Now, fix $j \geq 1$ and assume that~$\eps$ is so small
1400: that the conclusions of Theorem~\ref{Thm.norm}
1401: and Lemma~\ref{Lem.comparison} hold.
1402: By virtue of Theorem~\ref{Thm.norm}, we have
1403: $$
1404:   \left|
1405:   \left[\lambda_j(\hat{H}_\eps)
1406:   -\left(\frac{\pi}{2\eps}\right)^2+\frac{k}{\eps}\right]^{-1}
1407:   - \left[\lambda_j(\hat{H}_0)
1408:   -\left(\frac{\pi}{2\eps}\right)^2+\frac{k}{\eps}\right]^{-1}
1409:   \right|
1410:   \ \leq \
1411:   C_0 \, \eps^{3/2}
1412:   \,,
1413: $$
1414: since the left hand side is estimated
1415: by the norm of the resolvent difference.
1416: Using now Lemma~\ref{Lem.comparison},
1417: the above estimate is equivalent to
1418: %
1419: \begin{equation}\label{1/2}
1420:   \left|
1421:   \frac{1}{\eps \big[\lambda_j(\hat{H}_\eps)-\pi^2/(2\eps)^2\big]+k}
1422:   -\frac{1}{\eps \, \lambda_j(-\Delta_D^I+\kappa/\eps)+k}
1423:   \right|
1424:   \ \leq \
1425:   C_0 \, \eps^{1/2}
1426:   \,.
1427: \end{equation}
1428: %
1429: Consequently, recalling~\eqref{strong}, we conclude with
1430: $$
1431:   \lim_{\eps \to 0} \eps \left[\lambda_j(\hat{H}_\eps)
1432:   -\left(\frac{\pi}{2\eps}\right)^2\right]
1433:   = \lim_{\eps \to 0} \eps \, \lambda_j\big(-\Delta_D^I+\frac{\kappa}{\eps}\big)
1434:   = \inf\kappa
1435:   \,.
1436: $$
1437: This is indeed equivalent to Theorem~\ref{Thm.mine}
1438: because~$\hat{H}_\eps$ and~$-\Delta_{DN}^{\Omega_\eps}$
1439: are unitarily equivalent (therefore isospectral).
1440: 
1441: %
1442: \begin{Remark}
1443: Because of the lack of one half in the power of~$\eps$
1444: in Theorem~\ref{Thm.norm}, it turns out that~\eqref{1/2}
1445: yields a slightly worse result than Theorem~\ref{Thm.stronger}.
1446: Let us also emphasize that Theorem~\ref{Thm.mine} and~\ref{Thm.stronger}
1447: have been proved in Section~\ref{Sec.proof}
1448: without the need to assume the extra condition~\eqref{Ass.derivative}.
1449: On the other hand, Theorem~\ref{Thm.norm} contains an operator-type
1450: convergence result of independent interest.
1451: \end{Remark}
1452: %
1453: 
1454: %----------------------------%
1455: \section{Possible extensions}\label{Sec.end}
1456: %----------------------------%
1457: %
1458: \subsection{Different boundary conditions}
1459: %
1460: The result of Theorem~\ref{Thm.stronger}
1461: readily extends to the case of other boundary conditions
1462: imposed on the sides
1463: $\mathcal{L}_\eps\big(\{\inf I\}\times(0,1)\big)$
1464: and $\mathcal{L}_\eps\big(\{\sup I\}\times(0,1)\big)$,
1465: provided that the boundary conditions for the one-dimensional
1466: Schr\"odinger operator are changed accordingly.
1467: 
1468: It is more interesting to impose different boundary
1469: conditions on the approaching parallel curves as $\eps \to 0$.
1470: As an example, let us keep the Dirichlet boundary conditions
1471: but replace the Neumann boundary condition
1472: by the Robin condition of the type
1473: $$
1474:   \frac{\partial \Psi}{\partial n}
1475:   + (\alpha \circ \ell_\eps^{-1}) \, \Psi = 0
1476:   \qquad\mbox{on}\qquad
1477:   \gamma_\eps(I)
1478:   \,.
1479: $$
1480: Here $\ell_\eps:=\mathcal{L}_\eps(\cdot,1)$
1481: and~$\alpha:I\to\Real$
1482: is assumed to be bounded and uniformly continuous.
1483: Let us denote the corresponding Laplacian by
1484: $-\Delta_{DR_\alpha}^{\Omega_\eps}$.
1485: Then the method of the present paper gives
1486: %
1487: \begin{Theorem}\label{Thm.Robin}
1488: For all $j \geq 1$,
1489: %
1490: \begin{align*}
1491:   \lambda_j(-\Delta_{DR_\alpha}^{\Omega_\eps})
1492:   &= \left(\frac{\pi}{2\eps}\right)^2
1493:   + \lambda_j\big(-\Delta_D^I + \frac{\kappa+2\alpha}{\eps}\big)
1494:   + \mathcal{O}(1)
1495:   \\
1496:   &= \left(\frac{\pi}{2\eps}\right)^2
1497:   + \frac{\inf(\kappa+2\alpha)}{\eps}
1498:   + o(\eps^{-1})
1499:   \qquad\mbox{as}\qquad
1500:   \eps \to 0
1501:   \,.
1502: \end{align*}
1503: %
1504: \end{Theorem}
1505: %
1506: Let us mention that strips with this combination
1507: of Dirichlet and Robin boundary conditions
1508: were studied in~\cite{FK3}.
1509: 
1510: %
1511: \subsection{Higher-dimensional generalization}
1512: %
1513: Let~$\Omega_\eps$ be a three-dimensional layer instead of the planar strip.
1514: That is, we keep the definition~\eqref{strip} with~\eqref{StripMap},
1515: but now $\gamma: I \subseteq \Real^2 \to \Real^3$ is a parametrization of
1516: a two-dimensional surface embedded in~$\Real^3$
1517: and $n:=(\partial_1\gamma)\times(\partial_2\gamma)$,
1518: where the cross denotes the vector product in $\Real^3$.
1519: Let~$M$ be the corresponding mean curvature.
1520: Proceeding in the same way as in Section~\ref{Sec.proof}
1521: (we omit the details but refer to~\cite{DEK2}
1522: for a necessary geometric background), we get
1523: %
1524: \begin{Theorem}\label{Thm.layer}
1525: For all $j \geq 1$,
1526: %
1527: \begin{align*}
1528:   \lambda_j(-\Delta_{DN}^{\Omega_\eps})
1529:   &= \left(\frac{\pi}{2\eps}\right)^2
1530:   + \lambda_j\big(-\Delta_D^\gamma + \frac{2M}{\eps}\big)
1531:   + \mathcal{O}(1)
1532:   \\
1533:   &= \left(\frac{\pi}{2\eps}\right)^2
1534:   + \frac{2\inf M}{\eps}
1535:   + o(\eps^{-1})
1536:   \qquad\mbox{as}\qquad
1537:   \eps \to 0
1538:   \,,
1539: \end{align*}
1540: %
1541: where $-\Delta_D^\gamma$ denotes the Laplace-Beltrami operator
1542: in $\sii\big(\gamma(I)\big)$, subject to Di\-richlet boundary conditions.
1543: \end{Theorem}
1544: %
1545: 
1546: Notice that the leading geometric term in the asymptotic expansions
1547: depends on the extrinsic curvature only.
1548: This suggests that the spectral properties of the Dirichlet-Neumann layers
1549: will differ significantly from the purely Dirichlet case
1550: studied in~\cite{DEK2,CEK,LL1,LL3},
1551: where the Gauss curvature of~$\gamma$ plays a crucial role.
1552: 
1553: %
1554: \subsection{Curved ambient space}
1555: %
1556: The results of the present paper extend to the case
1557: of strips embedded in an abstract two-dimensional Riemannian manifold~$\mathcal{A}$
1558: instead of the Euclidean plane.
1559: Indeed, it follows from~\cite{K1} (see also~\cite{K3} and~\cite[Sec.~5]{FK4})
1560: that the quadratic form~$Q_\eps$ has the same structure;
1561: the only difference is that in this more general situation
1562: $h_\eps$~is obtained as the solution of the Jacobi equation
1563: %
1564: \begin{equation*}
1565:   \partial_2^2 h_\eps + \eps^2\,K\,h_\eps = 0
1566:   \qquad\textrm{with}\qquad\left\{
1567:   \begin{aligned}
1568:     h_\eps(\cdot,0) &= 1 \,, \\
1569:     \partial_2 h_\eps(\cdot,0) &= -\eps\,\kappa \,,
1570:   \end{aligned}
1571:   \right.
1572: \end{equation*}
1573: %
1574: where~$K$ is the Gauss curvature of~$\mathcal{A}$.
1575: Here~$\kappa$ is the curvature of $\gamma:I\to\mathcal{A}$
1576: (it is in fact the geodesic curvature of~$\gamma$
1577: if the ambient space~$\mathcal{A}$ is embedded in~$\Real^3$).
1578: Consequently, up to higher-order terms in~$\eps$,
1579: the function $h_\eps$ coincides with the expression~\eqref{Jacobian}
1580: for the flat case $K=0$. Namely,
1581: $$
1582:   h_\eps(s,t) = 1 - \kappa(s) \, \eps \, t + \mathcal{O}(\eps^2)
1583:   \qquad\mbox{as}\qquad
1584:   \eps \to 0
1585:   \,.
1586: $$
1587: Following the lines of the proof in Section~\ref{Sec.proof},
1588: it is then possible to check that Theorems~\ref{Thm.mine}
1589: and~\ref{Thm.stronger} remain valid without changes.
1590: The curvature of the ambient space comes into the asymptotics
1591: via higher-order terms only.
1592: 
1593: %---------------------------%
1594: \subsection*{Acknowledgment}
1595: %---------------------------%
1596: The work has been supported by 
1597: the Czech Academy of Sciences and its Grant Agency
1598: within the projects IRP AV0Z10480505 and A100480501,
1599: and by the project LC06002 of the Ministry of Education,
1600: Youth and Sports of the Czech Republic.
1601: 
1602: %--------------%
1603: % BIBLIOGRAPHY %
1604: %--------------%
1605: %
1606: %\bibliography{bib}
1607: %\bibliographystyle{amsplain}
1608: %
1609: \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
1610: \providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR }
1611: % \MRhref is called by the amsart/book/proc definition of \MR.
1612: \providecommand{\MRhref}[2]{%
1613:   \href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2}
1614: }
1615: \providecommand{\href}[2]{#2}
1616: \begin{thebibliography}{10}
1617: 
1618: \bibitem{BF}
1619: D.~Borisov and P.~Freitas,
1620: \emph{Singular asymptotic expansions for Dirichlet eigenvalues
1621: and eigenfunctions of the Laplacian on thin planar domains},
1622: Ann. Inst. H. Poincar\'e Anal. Non Lin\'eaire, to appear;
1623: preprint on arXiv:0712.3245v1 [math.SP] (2007).
1624: 
1625: \bibitem{BMT}
1626: G.~Bouchitt\'e, M.~L. Mascarenhas, and L.~Trabucho, \emph{On the curvature and
1627:   torsion effects in one dimensional waveguides}, Control, Optimisation and
1628:   Calculus of Variations \textbf{13} (2007), no.~4, 793--808.
1629: 
1630: \bibitem{CEK}
1631: G.~Carron, P.~Exner, and D.~Krej\v{c}i\v{r}\'{\i}k, \emph{Topologically
1632:   nontrivial quantum layers}, J.~Math.\ Phys. \textbf{45} (2004), 774--784.
1633: 
1634: \bibitem{Davies}
1635: E.~B. Davies, \emph{Spectral theory and differential operators}, Camb. Univ
1636:   Press, Cambridge, 1995.
1637: 
1638: \bibitem{DKriz2}
1639: J.~Dittrich and J.~K{\v{r}}{\'{\i}}{\v{z}}, \emph{Curved planar quantum wires
1640:   with {D}irichlet and {N}eumann boundary conditions}, J. Phys. A \textbf{35}
1641:   (2002), L269--275.
1642: 
1643: \bibitem{DE}
1644: P.~Duclos and P.~Exner, \emph{{C}urvature-induced bound states in quantum
1645:   waveguides in two and three dimensions}, Rev. Math. Phys. \textbf{7} (1995),
1646:   73--102.
1647: 
1648: \bibitem{DEK2}
1649: P.~Duclos, P.~Exner, and D.~Krej\v{c}i\v{r}\'{\i}k, \emph{Bound states in
1650:   curved quantum layers}, Commun. Math. Phys. \textbf{223} (2001), 13--28.
1651: 
1652: \bibitem{ES}
1653: P.~Exner and P.~{\v S}eba, \emph{Bound states in curved quantum waveguides},
1654:   J.~Math.~Phys. \textbf{30} (1989), 2574--2580.
1655: 
1656: \bibitem{F}
1657: P.~Freitas,
1658: \emph{Precise bounds and asymptotics for the first
1659: Dirichlet eigenvalue of triangles and rhombi},
1660: J. Funct. Anal. \textbf{251} (2007), 376--398.
1661: 
1662: \bibitem{FK4}
1663: P.~Freitas and D.~Krej\v{c}i\v{r}\'{\i}k,
1664: \emph{Location of the nodal set for thin curved tubes},
1665:   Indiana Univ. Math. J.~\textbf{57} (2008), no. 1, to appear;
1666:   preprint on [math.SP/0602470] (2006).
1667: 
1668: \bibitem{FK1}
1669: \bysame,
1670: \emph{Instability results for the damped wave equation in unbounded domains},
1671: J.~Differential Equations \textbf{211} (2005), no.~1, 168--186.
1672: 
1673: \bibitem{FK3}
1674: \bysame, \emph{Waveguides with combined {D}irichlet and {R}obin boundary
1675:   conditions}, Math. Phys. Anal. Geom. \textbf{9} (2006), no.~4, 335--352.
1676: 
1677: \bibitem{Friedlander-Solomyak_2007a}
1678: L.~Friedlander and M.~Solomyak, \emph{On the spectrum of the {D}irichlet
1679:   {L}aplacian in a narrow strip}, preprint on arXiv:0705.4058v1 [math.SP]
1680:   (2007).
1681: 
1682: \bibitem{Friedlander-Solomyak_2007b}
1683: \bysame, \emph{On the spectrum of the {D}irichlet {L}aplacian in a narrow
1684:   strip, II}, preprint on arXiv:0710.1886v1 [math.SP] (2007).
1685: 
1686: \bibitem{GJ}
1687: J.~Goldstone and R.~L. Jaffe, \emph{Bound states in twisting tubes}, Phys.
1688:   Rev.~B \textbf{45} (1992), 14100--14107.
1689: 
1690: \bibitem{Grieser}
1691: D.~Grieser, 
1692: \emph{Thin tubes in mathematical physics, global analysis and spectral geometry}, 
1693: Analysis on Graphs and its Applications (Cambridge, 2007),
1694: Proceedings of Symposia in Pure Mathematics, Amer. Math. Soc., to appear.
1695: 
1696: \bibitem{JLP}
1697: E.~R. Johnson, M.~Levitin, and L.~Parnovski, \emph{Existence of eigenvalues of
1698:   a linear operator pencil in a curved waveguide -- localized shelf waves on a
1699:   curved coast}, Siam J. Math. Anal. \textbf{37} (2006), 1465--1481.
1700: 
1701: \bibitem{Karp-Pinsky_1988}
1702: L.~Karp and M.~Pinsky, \emph{First-order asymptotics of the principal
1703:   eigenvalue of tubular neighborhoods}, Geometry of random motion (Ithaca,
1704:   N.Y., 1987), Contemp. Math., vol.~73, Amer. Math. Soc., Providence, RI, 1988,
1705:   pp.~105--119.
1706: 
1707: \bibitem{K1}
1708: D.~Krej\v{c}i\v{r}\'{\i}k, \emph{Quantum strips on surfaces}, J.~Geom. Phys.
1709:   \textbf{45} (2003), no.~1--2, 203--217.
1710: 
1711: \bibitem{K3}
1712: \bysame, \emph{Hardy inequalities in strips on ruled surfaces}, J. Inequal.
1713:   Appl. \textbf{2006} (2006), Article ID 46409, 10 pages.
1714: 
1715: \bibitem{KKriz}
1716: D.~Krej\v{c}i\v{r}\'{\i}k and J.~K\v{r}\'{\i}\v{z}, \emph{On the spectrum of
1717:   curved quantum waveguides}, Publ.~RIMS, Kyoto University \textbf{41} (2005),
1718:   no.~3, 757--791.
1719: 
1720: \bibitem{LL1}
1721: Ch. Lin and Z.~Lu, \emph{Existence of bound states for layers built over
1722:   hypersurfaces in {$\mathbb{R}^{n+1}$}}, J.~Funct. Anal. \textbf{244} (2007),
1723:   1--25.
1724: 
1725: \bibitem{LL3}
1726: \bysame, \emph{Quantum layers over surfaces ruled outside a compact set},
1727:   J.~Math. Phys. \textbf{48} (2007), art. no. 053522.
1728: 
1729: \bibitem{OM}
1730: O.~Olendski and L.~Mikhailovska, \emph{Localized-mode evolution in a curved
1731:   planar waveguide with combined {D}irichlet and {N}eumann boundary
1732:   conditions}, Phys. Rev. E~ \textbf{67} (2003), art.~056625.
1733: 
1734: \bibitem{RS1}
1735: M.~Reed and B.~Simon, \emph{Methods of modern mathematical physics,
1736:   {I}.~{F}unctional analysis}, Academic Press, New York, 1972.
1737: 
1738: \bibitem{Schatzman_1996}
1739: M.~Schatzman, \emph{On the eigenvalues of the {L}aplace operator on a thin set
1740:   with {N}eumann boundary conditions}, Applicable Anal. \textbf{61} (1996),
1741:   293--306.
1742: 
1743: \end{thebibliography}
1744: %
1745: \end{document}
1746: