math0409579/bz4.tex
1: \documentclass{amsart}
2: 
3: 
4: \usepackage{amsmath, amsthm, amsopn, amsfonts}
5: 
6: 
7: %\usepackage{showkeys}
8: 
9: \usepackage{color}
10: 
11: \usepackage{graphicx}
12: 
13: 
14: \usepackage{graphpap}
15: 
16: 
17: %\usepackage[english, francais]{babel}
18: 
19: 
20: 
21: 
22: 
23: %\textwidth= 14.5 cm
24: 
25: 
26: %\textheight= 20.5cm
27: 
28: 
29: %\hoffset = -1cm
30: 
31: 
32: % UNE MACRO POUR PLACER DU TEXTE SUR UNE BOITE. 
33: 
34: 
35: % SYNTAXE: \ecriture{boite}{
36: 
37: 
38: % \aat{coordhoriz}{coordvert}{texte}} la macro trace un quadrillage 
39: 
40: 
41: %de 50
42: 
43: 
44: % pour faciliter le placement du texte. On le supprime apres mise au
45: 
46: 
47: % point par la commande \pasdegrille en debut de fichier*
48: 
49: 
50: 
51: 
52: 
53: 
54: 
55: 
56: \def\finpreuve{\hfill $\Box$\par}
57: 
58: 
59: \def\pasdegrille{\let\grille = \pasgrille}
60: 
61: 
62: \def\ecriture#1#2{\setbox1=\hbox{#1}
63: 
64: 
65: \dimen1= \wd1
66: 
67: 
68: \dimen2=\ht1
69: 
70: 
71: \dimen3=\dp1
72: 
73: 
74: \grille #2 \box1 }
75: 
76: 
77: \def\aat#1#2#3{
78: 
79: 
80: \divide \dimen1 by 48
81: 
82: 
83: \dimen3=\dimen1
84: 
85: 
86: \multiply \dimen1 by #1
87: 
88: 
89: \advance \dimen1 by -\dimen3
90: 
91: 
92: \divide \dimen1 by 101
93: 
94: 
95: \multiply \dimen1 by 100
96: 
97: 
98: \divide \dimen2 by \count11
99: 
100: 
101: \multiply \dimen2 by #2 
102: 
103: 
104: \setbox0=\hbox{#3}\ht0=0pt\dp0=0pt
105: 
106: 
107:   \rlap{\kern\dimen1 \vbox to0pt{\kern-\dimen2\box0\vss}}\dimen1= \wd1
108: 
109: 
110: \dimen2=\ht1}
111: 
112: 
113: \def\pasgrille{
114: 
115: 
116: \count12= \dimen1 
117: 
118: 
119: \divide \count12 by 50
120: 
121: 
122: \divide \dimen2 by \count12
123: 
124: 
125: \count11 =\dimen2
126: 
127: 
128: \ 
129: 
130: 
131: \divide \dimen1 by 48
132: 
133: 
134: \setlength{\unitlength}{\dimen1}
135: 
136: 
137: \smash{\rlap{\ }}
138: 
139: 
140: \dimen1= \wd1
141: 
142: 
143: \dimen2=\ht1
144: 
145: 
146: }
147: 
148: 
149: \def\grille{
150: 
151: 
152: \count12= \dimen1 
153: 
154: 
155: \divide \count12 by 50
156: 
157: 
158: \divide \dimen2 by \count12
159: 
160: 
161: \count11 =\dimen2
162: 
163: 
164: \ 
165: 
166: 
167: \divide \dimen1 by 48
168: 
169: 
170: \setlength{\unitlength}{\dimen1}
171: 
172: 
173: \smash{\rlap{\graphpaper[1](0,0)(50, \count11)}}
174: 
175: 
176: \dimen1= \wd1
177: 
178: 
179: \dimen2=\ht1
180: 
181: 
182: }
183: 
184: 
185: 
186: 
187: 
188: \pasdegrille
189: 
190: 
191: %%%%
192: 
193: 
194: %%%
195: 
196: 
197: % FIN DE KLA
198: 
199: 
200: %MACRO POUR INCLURE DU TEXTE DANS UNE FIGURE
201: 
202: 
203: %
204: 
205: 
206: %
207: 
208: 
209: 
210: 
211: 
212: 
213: 
214: 
215: \usepackage{amssymb}
216: 
217: 
218: %\usepackage{draftcopy}
219: 
220: 
221: \ifx\pdfoutput\undefined
222: 
223: 
224: \usepackage{graphicx}  \DeclareGraphicsExtensions{.ps}
225: 
226: 
227: \def\Dessin#1{\begin{center}
228: 
229: 
230: \includegraphics{#1.mps}
231: 
232: 
233: \end{center}
234: 
235: 
236: }
237: 
238: 
239: \else
240: 
241: 
242: \usepackage{amsfonts}
243: 
244: 
245: \usepackage[pdftex]{graphicx}  \DeclareGraphicsExtensions{.pdf,.mps}
246: 
247: 
248: \def\Dessin#1{\begin{center}
249: 
250: 
251: \includegraphics{#1.mps}
252: 
253: 
254: \end{center}
255: 
256: 
257: }
258: 
259: 
260: \fi
261: 
262: 
263: 
264: 
265: 
266: \newcommand{\PSbox}[1]{\includegraphics[0in,0in][3in,3in]{#1}}
267: 
268: 
269: \usepackage[T1]{fontenc}
270: 
271: 
272: 
273: 
274: 
275: % Added by JW
276: 
277: 
278: \setlength{\marginparwidth}{.8in}
279: 
280: 
281: \newcommand{\mar}[1]{{\marginpar{\sffamily{\scriptsize #1}}}} 
282: 
283: 
284: \usepackage{epic, eepic}
285: 
286: 
287: % Have to temporarily widen margins to make marginal notes fit!
288: 
289: 
290: \setlength{\textheight}{8in} \setlength{\oddsidemargin}{0.35in}
291: 
292: 
293: \setlength{\evensidemargin}{0.35in} \setlength{\textwidth}{5.8in}
294: 
295: 
296: \setlength{\topmargin}{0.18in} \setlength{\headheight}{0.18in}
297: 
298: 
299: \setlength{\marginparwidth}{1.0in}
300: 
301: 
302: 
303: 
304: 
305: \newcommand{\abs}[1]{{\left\lvert{#1}\right\rvert}}
306: 
307: 
308: \newcommand{\norm}[1]{{\left\lVert{#1}\right\rVert}}
309: 
310: 
311: \newcommand{\bignorm}[1]{{\Bigl\lVert{#1}\Bigr\rVert}}
312: 
313: 
314: \newcommand{\pa}{{\partial}}
315: 
316: 
317: \newcommand{\h}{{\frac 12}}
318: 
319: 
320: \newcommand{\cinf}{{\mathcal{C}}^\infty}
321: 
322: 
323: 
324: 
325: 
326: \def\mathbbone{{\mathchoice {1\mskip-4mu \text{l}} {1\mskip-4mu \text{l}}
327: 
328: 
329: { 1\mskip-4.5mu \text{l}} { 1\mskip-5mu \text{l}}}}
330: 
331: 
332: 
333: 
334: 
335: \newcommand{\bc}{\mathbb C}
336: 
337: 
338: \newcommand{\br}{\mathbb R}
339: 
340: 
341: \newcommand{\bz}{\mathbb Z}
342: 
343: 
344: \newcommand{\im}{\operatorname{\rm Im}}
345: 
346: 
347: \newcommand{\mn}[1]{\Vert#1\Vert}
348: 
349: 
350: \newcommand{\op}{\operatorname{Op}}
351: 
352: 
353: \newcommand{\re}{\operatorname{\rm Re}}
354: 
355: 
356: \newcommand{\set}[1]{\left\{\,#1\,\right\}}
357: 
358: 
359: \newcommand{\w}[1]{\langle #1\rangle }
360: 
361: 
362: 
363: 
364: 
365: 
366: 
367: 
368: \def\squarebox#1{\hbox to #1{\hfill\vbox to #1{\vfill}}} 
369: 
370: 
371: \newcommand{\stopthm}{\hfill\hfill\vbox{\hrule\hbox{\vrule\squarebox 
372: 
373: 
374:                  {.667em}\vrule}\hrule}\smallskip} 
375: 
376: 
377: 
378: 
379: 
380: \pagestyle{headings}
381: 
382: 
383: 
384: 
385: 
386: 
387: 
388: 
389: 
390: 
391: 
392: 
393: 
394: 
395: 
396: 
397: 
398: %\pagestyle{plain}
399: 
400: 
401: 
402: 
403: 
404: \newcommand{\1}{{\bold 1}} 
405: 
406: 
407: \newcommand{\F}{{\mathcal F}} 
408: 
409: 
410: \newcommand{\CC}{{\mathbb C}}
411: 
412: 
413: \newcommand{\CI}{{\mathcal C}^\infty }
414: 
415: 
416: \newcommand{\HH}{{\mathcal H}}
417: 
418: 
419: \newcommand{\CIc}{{\mathcal C}^\infty_{\rm{c}} }
420: 
421: 
422: \newcommand{\CIb}{{\mathcal C}^\infty_{\rm{b}} }
423: 
424: 
425: \newcommand{\Op}{{\operatorname{Op}}}
426: 
427: 
428: %^{{w}}_h}}
429: 
430: 
431: \newcommand{\Oo}{{\mathcal O}} 
432: 
433: 
434: \newcommand{\Hh}{{\mathcal H}}
435: 
436: 
437: \newcommand{\pic}{{\mbox{Pic}}}
438: 
439: 
440: \newcommand{\Z}{{\mathbb Z}}
441: 
442: 
443: \newcommand{\Q}{{\mathbb Q}}
444: 
445: 
446: \newcommand{\RR}{{\mathbb R}}
447: 
448: 
449: \newcommand{\SP}{{\mathbb S}}
450: 
451: 
452: \newcommand{\ZZ}{{\mathbb Z}}
453: 
454: 
455: \newcommand{\NN}{{\mathbb N}}
456: 
457: 
458: \newcommand{\MO}{{\mathcal M}}
459: 
460: 
461: \newcommand{\U}{{\mathcal U}}
462: 
463: 
464: \newcommand{\N}{{\mathbb N}}
465: 
466: 
467: \newcommand{\vol}{\operatorname{vol}}
468: 
469: 
470: \newcommand{\rank}{\operatorname{rank}}
471: 
472: 
473: \newcommand{\itt}{\operatorname{it}}
474: 
475: 
476: \newcommand{\KKer}{\operatorname{ker}}
477: 
478: 
479: \newcommand{\supp}{\operatorname{supp}}
480: 
481: 
482: \newcommand{\Ran}{\operatorname{Ran}}
483: 
484: 
485: \newcommand{\UC}{U_{\circlearrowright}}
486: 
487: 
488: \newcommand{\QC}{Q_{\circlearrowright}}
489: 
490: 
491: \newcommand{\comp}{\operatorname{comp}}
492: 
493: 
494: \newcommand{\loc}{\operatorname{loc}}
495: 
496: 
497: \newcommand{\tr}{\operatorname{tr}}
498: 
499: 
500: \newcommand{\rest}{\!\!\restriction}
501: 
502: 
503: \newcommand{\ttt}{|\hspace{-0.25mm}|\hspace{-0.25mm}|}
504: 
505: 
506: \renewcommand{\Re}{\mathop{\rm Re}\nolimits}
507: 
508: 
509: \renewcommand{\Im}{\mathop{\rm Im}\nolimits}
510: 
511: 
512: 
513: 
514: 
515: % JW messed with the following:
516: 
517: 
518: \theoremstyle{plain}
519: 
520: 
521: \def\Rm#1{{\rm#1}}
522: 
523: 
524: \newtheorem{thm}{Theorem}
525: 
526: 
527: %   \renewcommand{\thethm}{\arabic{thm}}
528: 
529: 
530: \newtheorem{prop}{Proposition}[section]
531: 
532: 
533: %   \renewcommand{\theprop}{\arabic{prop}}
534: 
535: 
536: \newtheorem{cor}[prop]{Corollary}
537: 
538: 
539: %   \renewcommand{\thecor}{}
540: 
541: 
542: \newtheorem{lem}{Lemma}[section]
543: 
544: 
545: %   \renewcommand{\thelem}{\arabic{lem}}
546: 
547: 
548: %\numberwithin{lem}{section}
549: 
550: 
551: 
552: 
553: 
554: %\numberwithin{prop}{section}
555: 
556: 
557: \theoremstyle{definition}
558: 
559: 
560: \newtheorem{ex}{EXAMPLE}[section]
561: 
562: 
563: \newtheorem{exmple}{Example}[section]
564: 
565: 
566: \newtheorem{rem}{Remark}
567: 
568: 
569: \newtheorem{defn}[prop]{Definition} 
570: 
571: 
572: 
573: \numberwithin{equation}{section}
574: 
575: \newcommand{\thmref}[1]{Theorem~\ref{#1}}
576: \newcommand{\secref}[1]{Section~\ref{#1}}
577: \newcommand{\lemref}[1]{Lemma~\ref{#1}}
578: \newcommand{\exref}[1]{Example~\ref{#1}}
579: \newcommand{\corref}[1]{Corollary~\ref{#1}}
580: \newcommand{\propref}[1]{Proposition~\ref{#1}}
581: 
582: \def\bbbone{{\mathchoice {1\mskip-4mu \rm{l}} {1\mskip-4mu \rm{l}}
583: { 1\mskip-4.5mu \rm{l}} { 1\mskip-5mu \rm{l}}}}
584: 
585: 
586: \title[Eigenfunctions for partially rectangular billiards]
587: {Eigenfunctions for partially rectangular billiards}
588: \author[J. Marzuola]{Jeremy Marzuola}
589: \address{Mathematics Department, University of California \\
590: Evans Hall, Berkeley, CA 94720, USA}
591: \email{marzuola@math.berkeley.edu}
592: 
593: \usepackage{amssymb}
594: \usepackage{amsmath, amsthm, amsopn, amsfonts}
595: %\usepackage[dvips]{graphicx}
596: 
597: 
598: 
599: \def\11{{\rm 1~\hspace{-1.4ex}l} }
600: \def\R{\mathbb R}
601: \def\C{\mathbb C}
602: \def\Z{\mathbb Z}
603: \def\N{\mathbb N}
604: \def\E{\mathbb E}
605: \def\T{\mathbb T}
606: \def\O{\mathbb O}
607: \def\K{\mathbb K}
608: \def\cal{\mathcal}
609: 
610: 
611: \begin{document}    
612: 
613:    
614: \maketitle   
615:    
616: \section{Introduction}   
617: \label{in}
618: In this note, we further develop the methods of Burq-Zworski \cite{BZ3} 
619: to study eigenfunctions for billiards which have rectangular components: these include the Bunimovich 
620: billiard, the Sinai billiard, and the recently popular pseudointegrable billiards
621: \cite{Bog}.  The results are an application of a "black box" point of view as
622: presented in \cite{BZ2} by the same authors.  
623: 
624: 
625: 
626: \begin{figure}[ht]
627: \includegraphics[width=10cm]{sinai1.epsi}
628: \caption{Experimental images of eigenfunctions in a Sinai billiard
629: microwave cavity -- see {\tt http://sagar.physics.neu.edu}. We see
630: that there is always a non-vanishing presence near the boundary of the
631: obstacle as predicted by Theorem \ref{t:s} below.}
632: \label{fig:sin}
633: \end{figure}
634: 
635: 
636: By a partially rectangular billiard, we mean a connected planar domain,
637: $ \Omega $, with a piecewise smooth boundary, which contains a  rectangle,
638: $ R \subset \Omega $, such that if we decompose the boundary of 
639: $ R $, into pairs of parallel segments, $ \partial R = \Gamma_1 \cup \Gamma_2 $,
640: then $ \Gamma_i \subset \partial \Omega $, for at least one $ i$.
641: 
642: We show that for such billiards, 
643: the eigenfunctions of the Dirichlet, Neumann,
644: or periodic Laplacian cannot concentrate in closed sets in the interior of the rectangular part.
645: A combination of this elementary result with 
646: the now standard, but highly non-elementary, 
647: propagation results of Melrose-Sj\"ostrand \cite{MeSj} and Bardos-Lebeau-Rauch \cite{BLR},
648: can give further improvements -- see \cite{BZ2},\cite{BZ3}.
649: 
650: Here, we prove further non-concentration results, away from the obstacle in the Sinai billiard
651: (see Fig.1 and Theorem \ref{t:s}), and along certain trajectories
652: in pseudointegrable billiards, (see Fig.5 and Theorem \ref{t:j}).  
653: For recent motivation coming from the study of {\em quantum chaos}
654: we suggest \cite{Bog},\cite{BZ3},\cite{Do},\cite{Ze}, and references given there.
655: 
656: \medskip
657: \noindent
658: {\sc Acknowledgments.} This paper is a development of an unpublished work by N. Burq and M. Zworski, 
659: who treated the case of a square billiard rather than a general rectangular billiard.
660: The author is very grateful to Maciej Zworski and Nicolas Burq for allowing me to use results from their 
661: unpublished work, as well as many helpful conversations.  I would also like to thank Srinivas Sridhar for 
662: allowing the use the experimental images shown in Fig.\ref{fig:sin}.  
663: 
664: 
665: \section{Semiclassical Pseudodifferential Operators on a Torus}
666: \label{PDO}
667: 
668: In this section, we would like to discuss properties of Pseudodifferential Operators (PDO's) on a torus.  
669: To begin, we examine the nature of PDO's and their symbol classes.  In ${\mathbb R}^n$, we define a
670: Weyl quantization of an operator $a(x,hD)$ where $a \in {\mathcal S} ({\mathbb R}^{2n}), a=a(x, \xi)$ by:
671: $$a (x,hD) u(x) = \frac{1}{(2 \pi h)^n} \int_{{\mathbb R}^n} \int_{{\mathbb R}^n} e^{\frac{i}{h} \langle x-y, \xi \rangle}
672: a(x, \xi) u(y) dy d\xi ,$$
673: for $u \in {\mathcal S}$ and a symbol class by:  
674: $$S^k_\delta (m) = \{ a \in C^{\infty} ({\mathbb R}^{2n}) \| | \partial^\alpha a | \leq C_\alpha h^{-\delta |\alpha| -k} m 
675: \ \text{for all multi-indices} \ \alpha \},$$
676: where $m: {\mathbb R}^{2n} \to (0,\infty)$ is is an order function, i.e. there exist constants $C$, $N$ such that
677: $$ m(z) \leq C \langle z - w \rangle^N m(w).$$
678: We also have $$S^{-\infty}_\delta (m) = \bigcap_{k=-\infty}^{\infty} S^k_\delta (m).$$ 
679: For $k,\delta = 0$ we write simply $S (m)$. 
680: On a torus, however, $a$ and all its derivatives are bounded in the $x$ variable, thus for $h$ small and $k$ positive, we need not worry
681: about the derivatives in x, only those in $\xi$.  Also, for $k$ negative, provided
682: that we have the proper local regularity for our symbol $a$, this definition still works perfectly on a torus.  
683: 
684: 
685: Note also that we need only work with symbols that are periodic in the x-variable
686: with period determined by the dimensions of the torus.  In other words,
687: $a(x,\xi)=a(x+\gamma, \xi)$ for $\gamma \in (a{\mathbb Z}) \times (b{\mathbb Z})$, where $a, b \in {\mathbb R}$.  
688: Using this relation, we easily see the following propostion.
689: 
690: \begin{prop}
691: \label{p:3}  
692: If $a(x,\xi)$ is a periodic symbol in x with period $\gamma$, then $a(x,D) T_\gamma = T_\gamma a(x,D)$ where
693: $T_\gamma u(x) = u(x - \gamma)$.  
694: \end{prop}
695: \begin{proof}
696: We easily calculate:
697: 
698: \begin{eqnarray}
699: \label{eq:p3}
700: a(x,h D) T_\gamma (u) & = & \int_{{\mathbb R}^n} a(x,\xi) e^{\frac{i}{h} <x-y,\xi>} \hat{u} (y-\gamma) dy d\xi \\
701: &  = & \int_{{\mathbb R}^n} a(x-\gamma,\xi) e^{\frac{i}{h} <x-\gamma-{\tilde y},\xi>} \hat{u} (\tilde y) d{\tilde y} d\xi \\
702: &  = & T_\gamma(a(x,D) u)
703: \end{eqnarray}
704: \end{proof}
705: 
706: From the above proposition, it becomes clear that the properties of symbol classes in Euclidean space  
707: translate directly to properties of similarly defined symbol classes on a torus.  
708: For instance, we have the following result.
709: 
710: \begin{prop}
711: \label{p:4}
712: Given $u(x) = u(x+\gamma)$ where $\gamma$ is as above, and $u$ is $L^2$ on a torus, then $a(x,h D) u(x)$ is
713: $L^2$ on the torus, when $a \in {\mathcal S}_\delta (1)$, $0 \leq \delta \leq \frac{1}{2}$.
714: \end{prop}
715: \begin{proof}
716: Note that the condition on $a$ implies that it is $L^2$ bounded.
717: Given a function $u(x)$ which is periodic on a torus, we can write it as 
718: $\sum_\gamma T_\gamma u_0$ where $u_0 = \chi (x) u(x)$ and $\chi (x)$ is equal to 1 on a single copy
719: of the torus in the plane and 0 otherwise.  Note that no assumptions about the smoothness of $\chi(x)$ are made.  
720: Hence, $u_0 \in L^2$ and therefore, so is $a(x,h D) u_0$.  Then, 
721: $a(x,h D) u(x) = \sum_\gamma T_\gamma a(x,h D) u_0$.  The sum converges for each $x$ since $a(x,D) u_0$ will
722: have compact support and we have
723: $a(x,h D) u$ a periodic function that is $L^2$ on the torus. 
724: \end{proof}
725: 
726: Using similar techniques, we would like to develop the concept of a microlocal defect measure in this setting.
727: As shown in \cite{Zw}, we have the following theorem in Euclidean space:
728: \begin{thm}
729: \label{t:dm}
730: There exists a Radon measure $\mu$ on ${\mathbb R}^n$ and a sequence $h_j \rightarrow 0$ such that 
731: \begin{equation}
732: \langle a^w (x,h_j D) u(h_j), u(h_j) \rangle \rightarrow \int_{{\mathbb R}^{2n}} a(x,\xi) d\mu
733: \end{equation}
734: for all symbols $a \in S(1)$.
735: \end{thm}
736: 
737: We call $\mu$ a microlocal defect measure associated with the family $\{ u(h) \}_{0<h \leq h_0}$. 
738: Note that an $S(1)$ symbol on the torus corresponds to an $S(1)$ symbol on the plane, therefore this result proves
739: the existence of microlocal defect measures on a torus as well.  
740: 
741: \begin{proof}
742: 1.  Let $\{a_k \} \in C_c^\infty$ be dense in $C_0 ({\mathbb R}^{2n})$.  Select a sequence $h_j^1 \rightarrow 0$ such that
743: $$\langle a_1^w (x, h_j^1 D) u(h_j^1), u(h_j^1) \rangle \rightarrow \alpha_1.$$  Then, select a subsequence $\{ h_j^2 \} \subset \{ h_j^1 \}$
744: such that $$\langle a_2^w (x, h_j^2 D) u(h_j^2), u(h_j^2) \rangle \rightarrow \alpha_2.$$  Continue such that at the kth step, you take a subsequence
745: $\{ h_j^k \} \subset \{ h_j^{k-1} \}$ such that $$\langle a_k^w (x, h_j^k D) u(h_j^k), u(h_j^k) \rangle \rightarrow \alpha_k.$$  Then by
746: a diagonal argument, arrive at a sequence $h_j$ such that $$\langle a_k^w (x, h_j D) u(h_j), u(h_j) \rangle \rightarrow \alpha_k$$ for all $k=1,2,...$.
747: 
748: 2.  Define $\Phi(a_k) = \alpha_k$.  By a standard theorem on operator norms, we have for each $k$ that 
749: $$| \Phi (a_k) | = | \alpha_k | = \lim_{h_j \rightarrow \infty} | <a_k^w u(h_j), u(h_j) > | \leq
750: \limsup_{h_j \rightarrow \infty} C \| a_k^w \|_{L^2 \rightarrow L^2} \leq C \sup |a_k|.$$  The mapping $\Phi$ is bounded, linear
751: and densely defined, therefore uniquely extends to a bounded linear functional on $S(1)$, with the estimate $$ | \Phi (a) | \leq C \sup |a|$$
752: for all $a \in S(1)$.  The Riesz Representation Theorem therefore implies the existence of a (possibly complex valued) Radon measure on
753: ${\mathbb R}^{2n}$ such that $$\Phi (a) = \int_{{\mathbb R}^{2n}} a(x,\xi) d\mu.$$
754: \end{proof}
755: 
756: We now quote a general theorem about microlocal defect measures on Euclidean space which we can
757: then apply to a torus.  To state the propagation theorem in the form sufficient for our applications,
758: we follow \cite{BuIMRN}.
759: %\subsection{Semi-classical measures}
760: 
761: 
762: Let us consider a Riemannian manifold without boundary, $M$. By partitions of unity we can define semi-classical 
763: pseudo-differential operators $a(x, hD_x)$ associated to symbols $a(x, \xi)\in \CIc (T^* M)$. 
764: \par
765: Now we consider a sequence $(u_n)$ bounded in $L^2(M)$. satisfying
766: \begin{equation}\label{eq.equation}
767: (-h_n^2\Delta -1)u_n = 0.
768: \end{equation}  
769: Using~\eqref{eq.equation},
770: as in~\cite{GeLe93} (see also~\cite{BuIMRN}) we can prove the following result.
771: 
772: \begin{prop}
773: \label{prop2.1} There exist a subsequence $(n_{k})$ and a positive Radon  measure on $T^*M$, $\mu$ (a semi-classical measure for the sequence $(u_n)$), 
774: such that for any $a \in \CIc (T^* M)$
775: \begin{equation}
776: \label{eq2.8}\lim_{k\rightarrow + \infty}\left(a^w (x, h_{n_{k}}D_{x})u_{n_{k}}, u_{n_{k}}\right)_{ L^2(M)}= \langle \mu, a(x, \xi)\rangle.
777: \end{equation}
778: Furthermore this measure satisfies
779: \begin{enumerate}
780: \item The support of $\mu$ is included in the characteristic manifold:
781: \begin{equation}
782: \Sigma \stackrel{\rm{def}}{=} \{(x, \xi)\in T^*M; p(x,\xi)=\|\xi\|_x=1\}
783: \end{equation}
784: where  $\|\cdot \|_x$ is the norm for the metric at the point $x$,
785: \item The measure $\mu$ is invariant by the bicharacteristic flow (the flow of the Hamilton vector field of $p$):
786: \begin{equation}
787: \label{eq.inv}H_p \mu =0,
788: \end{equation}
789: \item For any $\varphi\in \CIc (T^*M)$, 
790: \begin{equation}
791: \label{eq.lim} \lim_{k \rightarrow + \infty} \|\varphi u_{n_k}\|^2 = \langle \mu, |\varphi|^2\rangle.
792: \end{equation}
793: \end{enumerate}
794: \end{prop}
795: The first two properties above are weak forms of the elliptic regularity and propagation of singularities results,  
796: whereas the last one states that there is no loss of $L^2$-mass at infinity in the $\xi$ variable.   
797: 
798: \begin{proof}
799: We will prove this proposition only for the case of a torus, but the methods are applicable to any manifold.
800: \bigskip
801: 
802: (0)  (Positivity)
803: We need to show that $a \geq 0$ implies $$\int_{{\mathbb T}^{2} \times {\mathbb R}^{2}} a(x, \xi) d\mu \geq 0.$$
804: Since $a \geq 0$, using the Garding inequality, we see that:  $$a^w (x,hD) \geq -Ch.$$
805: Let $h = h_j \rightarrow 0$, to see:
806: $$\int_{{\mathbb T}^{2} \times {\mathbb R}^{2}} a d\mu = \lim_{j \rightarrow \infty} \langle a^w (x,h_j D) u(h_j), u(h_j) \rangle \geq 0.$$
807: 
808: (1)  (Support of $\mu$)
809: Let $a$ be a smooth function such that $\text{supp} (a) \cap p^{-1} (1) = \emptyset$.  We must show 
810: $$\int_{{\mathbb T}^{2} \times {\mathbb R}^{2}} a d\mu = 0.$$
811: Select $\chi \in C^{\infty}_c ( {\mathbb T}^{2} \times {\mathbb R}^{2} )$ such that 
812: $$\text{supp} (a) \cap \text{supp} (\chi) = 0.$$
813: Then, $$a^w(x,hD) \left( ((p-1)^w+i \chi^w)^{-1} (p-1)^w \right) (x,hD) = a^w(x,hD) + {\mathcal O} (h^\infty)_{L^2 \rightarrow L^2}.$$
814: Apply $a^w (x,hD)$ to $u(h)$ to see that $a^w (x,hD) u(h) = o(1)$ and thus $\langle a^w (x,hD) u(h), u(h) \rangle \rightarrow 0$.
815: But, $$\langle a^w (x,hD) u(h_j), u(h_j) \rangle \rightarrow \int_{{\mathbb T}^{2} \times {\mathbb R}^{2}} a d\mu.$$
816: 
817: (2)  (Flow Invariance)
818: Select $a$ as above, then 
819: \begin{eqnarray} 
820: \langle [p^w,a^w] u(h), u(h) \rangle & = & \langle (p^w a^w -a^w p^w) u(h), u(h) \rangle \\
821:  & = & \langle a^w u(h), p^w u(h) \rangle - \langle p^w u(h), (a^w)^* u(h) \rangle \\
822:  & = & o(h) \ \text{as} \ h \rightarrow 0.
823: \end{eqnarray}
824: However, $[p^w,a^w] = \frac{h}{i} \{ p^w,a^w \} + O(h^2)$.  Hence, 
825: $$\langle [p^w,a^w] u(h), u(h) \rangle = \frac{h}{i} \langle \{ p,a \}^w u(h), u(h) \rangle + \langle o(h) u(h), u(h) \rangle.$$
826: As we let $h_j \rightarrow 0$, we get:  $$\int_{{\mathbb T}^{2} \times {\mathbb R}^{2}} \{p,a\} d\mu = 0.$$  
827: So, if $\Phi_t$ is the flow generated by the Hamiltonian vector field $H_p$, then 
828: $$\frac{d}{dt} \int_{{\mathbb T}^{2} \times {\mathbb R}^{2}} (\Phi_t* a) d\mu = 
829: \int_{{\mathbb T}^{2} \times {\mathbb R}^{2}} (H_p a)(\Phi_t) d\mu = 
830: \int_{{\mathbb T}^{2} \times {\mathbb R}^{2}} \{ p,a \} d\mu = 0.$$
831: Now, (3) follows easily by looking at the operator $|\varphi (x, \xi)|^2$ and applying the result about existence of a microlocal defect measure.
832: \end{proof}
833: 
834: 
835: \section{Partially rectangular billiards}
836: \label{prb}
837: 
838: In this section we will need to recall the basic control results \cite{Bu92},\cite{BZ2}
839: for rectagles, and the propagation results \cite{MeSj},\cite{BLR},\cite{BuIMRN},\cite{BuGe} for 
840: billiards. Since in the specific application presented in Section \ref{app} we only use propagation
841: away from the boundary, that is the only case we will review.
842: 
843: 
844: The following result from ~\cite{Bu92} is related to some earlier control results 
845: of Haraux~\cite{Ha} and  Jaffard~\cite{Ja}.\footnote{We remark that as noted in \cite{Bu92} 
846: the result holds for 
847: any product manifold $M= M_{x}\times M_{y}$, and the proof is essentially the same.}
848: \begin{prop}
849: \label{p:1} Let $\Delta$ be  the Dirichlet, Neumann, or periodic
850: Laplace operator on the rectangle $R= [0, 1]_{x} \times [0,a]_{y}$. 
851: Let $\omega_x$ be a non-empty open subset of $[0,1]$.  Then for any non-empty  $\omega \subset R$ of the form $
852: \omega= \omega_{x} \times [0,a]_{y}$, there 
853: exists $C$ such that for any solutions of
854: \begin{equation}
855: (\Delta -z) u =f \ \text{ on $R$}, \ u \rest_{\partial R}=0
856: \end{equation}
857: we have
858: \begin{equation}
859: \label{eq:6.12}\|u\|^2_{{L^2(R)}}\leq C \left(\|f \|^2_{H^{-1} (
860: [0,1]_{x}; L^2([0,a]_{y})) } +
861: \|u\rest _{\omega}\|^2_{{L^2(\omega)}} \right)
862: \end{equation}
863: \end{prop}
864: \begin{proof}
865: We will consider the Dirichlet case (the proof is the same in the other two 
866: cases) and 
867: decompose $u,f$ in terms of 
868: the basis of $L^2([0,a])$ formed by the Dirichlet eigenfunctions
869: $e_{k}(y)=  { \sqrt {{2}/a}}\sin(2k\pi y/a)$,
870: \begin{equation}
871: u(x,y)= \sum_{k}e_{k}(y) u_{k}(x), \qquad f(x,y)= \sum_{k}e_{k}(y) f_{k}(x).
872: \end{equation}
873: We get for $u_{k}, f_{k}$ the equation
874: \begin{equation}\label{estres.1}
875: \left(\Delta_{x}-\left(z+\left({2k\pi}/{a}\right)^2\right)\right)u_{k}= f_{k},\qquad u_{k}(0)=u_k(1)=0.
876: \end{equation}
877: We now claim that 
878: \begin{equation}
879: \label{eq:cont}
880: \|u_{k}\|^2_{{L^2([0,1]_{x})}}\leq C \left(\|f_k \|^2_{H^{-1}([0,1]_{x})} +
881: \|u_k \rest _{\omega_{x}}\|^2_{{L^2(\omega)}}\right), 
882: \end{equation}
883: from which, by summing the squares in $k$, we get~\eqref{eq:6.12}.
884: 
885: To see \eqref{eq:cont} we can use the propagation result above \ref{prop2.1} in dimension
886: one, but in this case an elementary calculation is easily available -- see
887: \cite{BZ3}.
888: \end{proof}
889: 
890: 
891: 
892: The following theorem is an easy consequence of Proposition \ref{p:1}:
893: 
894: \begin{thm}
895: \label{t:1}
896: Let $ \Omega $ be a partially rectangular billiard with the rectangular
897: part $ R \subset \Omega $, $ \partial R = \Gamma_1 \cup \Gamma_2 $, a decomposition 
898: into parallel components satisfying $ \Gamma_2 \subset \partial \Omega $. 
899: Let $ \Delta $ be the Dirichlet or Neumann Laplacian on $\Omega $. Then
900: for {\em any} neighbourhood of $ \Gamma_1 $ in $ \Omega $, $ V $, there exists $ C $
901: such that
902: \begin{equation}
903: \label{eq:t1}
904: - \Delta u = \lambda u \ \Longrightarrow \  \int_V | u ( x ) |^2 dx 
905: \geq \frac1C \int_R | u ( x ) |^2 dx \,, 
906: \end{equation}
907: that is, no eigenfuction can concentrate in $ R $ and away from $ \Gamma_1$.
908: \end{thm}
909: 
910: \begin{proof}
911: Let us take $x,y$ as 
912: the coordinates on the stadium, so that $x$ parametrizes $ \Gamma_2 \subset \partial 
913: \Omega $ and $ y$ parametrizes $ \Gamma_1 $, then  
914: \[ R = [0,1]_{x}\times [0,a]_{y} \,.\]
915: Let $\chi\in \CIc((0,1))$ be equal to $1$ on $[\varepsilon, 1- \varepsilon]$. Then $\chi(x) u(x,y)$ is a solution of
916: \begin{equation}
917: (\Delta-z)\chi u = [\Delta, \chi] u\text{ in $R$}
918: \end{equation} with the boundary conditions satisfied 
919: on $\partial R$. Applying Proposition~\ref{p:1}, we get 
920: \begin{equation}
921: \|\chi u\|_{L^2( R)}\leq C \left \| [ \Delta, \chi] u \|_{H^{-1}_{x}; L^2_{y}}+ 
922: \| u\rest _{\omega_{\varepsilon}}\|_{L^2( \omega_{\varepsilon})}\right)
923: \leq C' \| u\rest _{\omega_{\varepsilon}}\|_{L^2( \omega_{\varepsilon})} \,,
924: \end{equation}
925: where $\omega_{\varepsilon}$ is a neighbourhood of the support of $ \nabla\chi$.
926: Since a neighbourhood of $ \Gamma_1 $ in $ \Omega $ has to contain $ \omega_\varepsilon $
927: for some $ \varepsilon $, \eqref{eq:t1} follows.
928: \end{proof}
929: 
930: 
931: \section{Applications}
932: \label{app}
933: 
934: In \cite{BZ2} and \cite{BZ3}, Proposition \ref{p:1} is used to prove that in the
935: case of the Bunimovich billiard shown in Fig.\ref{fig:bath},
the states have nonvanishing 
936: density near the vertical boundaries of the rectangle. That follows from Theorem
937: \ref{t:1} which shows that we must have positive density in the wings of the
938: billiard, and the propagation result (in the boundary case) based on the fact that
939: any diagonal controls a disc geometrically (see  \cite[Section 6.1]{BZ2}; in fact 
940: we can use other control regions as shown in Fig.\ref{fig:bath}). 
941: Here we consider another case which accidentally generalizes a control theory 
942: result of Jaffard \cite{Ja}.
943: 
944: 
945: The Sinai billiard (see Fig.\ref{fig:sin}) 
946: is defined by removing a strictly convex open set, $ \mathcal O$, with 
947: a $ \CI $ boundary,
948:  from a flat torus, $ {\mathbb T}^2_{a,b} \stackrel{\rm{def}}{=}
949:  (a \SP^1 ) \times (b \SP^1 )$:
950: \[ S  \stackrel{\rm{def}}{=} {\mathbb T}^2_{a,b} \setminus {\mathcal O} \,.\]
951: The following theorem results by applying Theorem 1 to a torus with sides of 
952: arbitrary length.
953: 
954: \begin{figure}
955: \include{figure2}
956: \caption{Control regions in which eigenfunctions have positive mass and the
957: rectangular part for the Bunimovich stadium.}
958: \label{fig:bath}
959: \end{figure} 
960: 
961: 
962: \begin{thm}
963: \label{t:s}
964: Let $ V $ be any open neighbourhood of the convex boundary, $ \partial {\mathcal O} $, in 
965: a Sinai billiard, $ S $. If $ \Delta $ is the Dirichlet or Neumann
966: Laplace operator on $ S $ then there exists a constant, $ C = C ( V ) $, such that
967: \begin{equation}
968: \label{eq:ts}
969: - h^2 \Delta u = E(h) u \ \Longrightarrow \  \int_V | u ( x ) |^2 dx 
970: \geq \frac1C \int_S | u ( x ) |^2 dx \,, 
971: \end{equation}
972: for any $h$ and $|E(h) - 1| < \frac{1}{2}$.
973: \end{thm}
974: \begin{proof}
975: First note that we can easily limit ourselves to the case where our flat torus has
976: one side of length 1 and one side of length a.  Suppose that the result is not true, 
977: in other words, there exists a sequence of 
978: eigenfunctions $ u_n $, $ \| u _n \| = 1 $, with the corresponding 
979: eigenvalues $ \lambda_n \rightarrow \infty $,  such that $ \int_V | u_n ( x )|^2 dx
980: \rightarrow 0 $. 
981: 
982: We first observe that the only directions in the support of 
983: the corresponding semi-classical defect measure, $\mu$, have to be "rational", 
984: in other words, the trajectory must travel along a line of slope $\frac{m a}{n}$ where $m,n \in \NN$.  
985: The projection of a trajectory with an irrational direction is dense on the 
986: torus and hence must encounter the obstacle $ \partial {\mathcal O} $ (and 
987: consequently $V$). The propagation result recalled in Proposition \ref{prop2.1}, part (3),
988: gives a contradiction by choosing a proper test function $\phi$ which is nonzero on the support
989: of the measure $\mu$ resulting from our sequence of eigenfunctions (remark that we apply this result as long as the 
990: trajectory does not encounter the obstacle and consequently we need only the 
991: {\em interior} propagation).
992: 
993: Hence let us assume that there exists a rational direction in the support of 
994: the measure which then contains the periodic trajectory in that direction.  
995: As shown in Fig. \ref{f:3} we can find a maximal rectangular neighbourhood of 
996: the projection of that trajectory which avoids the obstacle.
997: 
998: \begin{figure}
999: \include{figure3}
1000: \caption{A maximal rectangle in a rational direction, avoiding the obstacle.
1001: Because the parrallelogram is certainly periodic and our region has uniform width,
1002: it is clear that the resulting rectangle is periodic.}
1003: \label{f:3}
1004: \end{figure} 
1005: 
1006: The rectangle can be described as $ R = [0,a_1]_{x_1} \times [0, b_1]_{y_1} $ with the the 
1007: $ y_1 $ coordinate parametrizing the trajectory. 
1008: Let $\epsilon$, $\delta >0$ be small.
1009: Let $ u $ be an eigenfunction in our sequence and define 
1010: $$ \chi = \{ \chi ( x_1 ) \in \CIc ({\mathbb T}^2) \ | \ \chi (x_1) = 1 \ \text{for all} \ x_1 \in (\epsilon,a_1-\epsilon), 
1011: \ \chi(x_1) = 0 \ \text{outside $(0,a_1)$} \}.$$  
1012: Note that we can then write $\chi = \chi(x,y)$ for $(x,y) \in {\mathbb T}^2_{1,a}$ as $x_1$ is simply a rotation and translation of 
1013: the standard coordinates.
1014: Then $ \chi ( x,y ) u ( x, y ) $ is a function on all of $ R $ satisfying the periodicity condition. 
1015: Let  $\Phi_\xi ( \nu ) = \Phi(\xi - \nu)$, where we define
1016: $$\Phi = \{ \Phi (\xi) \in \CIc ({\mathbb R}^2) \ | \ \Phi (\xi) = 1 \ \text{for} \ \xi \in B(0,\delta), 
1017: \ \Phi (\xi) = 0 \ \text{for} \ \xi \in {\mathbb R}^2 \setminus B(0,2 \delta) \}, $$ 
1018: where $B(0,\delta)$ is a ball centered at $0$ of radius $\delta$.
1019: Note, due to the compact support of this function in $\xi$,
1020: $\Phi_\xi (D)$ is in the symbol class $S(\langle \xi \rangle^{-N})$ for any $N$.
1021: Let $ \Delta_R $ is the (periodic) Laplacian on $ R $. Using Fourier decomposition
1022: we can arrange that $ [ \Delta_R , \Phi_\xi (D) ] = 0 $.  Since our eigenfunction is only defined on 
1023: ${\mathbb T}^2_{1,a} \setminus {\mathcal O}$, let us introduce a smooth function ${\chi}_0$ which is $0$ on a neighborhood $U$
of
1024: the obstacle and $1$ on ${\mathbb T}^2_{1,a} \setminus V$ where $U \subset V$.
 Choose $U$ and $V$ such that $\chi \chi_0 = \chi$.
1025: Hence, 
1026: \[  ( -h^2 \Delta_R - E(h) ) \Phi_\xi(D) \chi {\chi}_0 u = [ -h^2 \Delta_R , \Phi_\xi(D) \chi] u 
1027: =\Phi_\xi (D) [ -h^2 \Delta_R , \chi] \chi_0 u + {\mathcal O} ( h^{ \infty } ) 
1028: \,, \ \  \| u \| = 1 \,, 
1029: \]
1030: by the properties of $S(\langle \xi \rangle^{-N})$ operators acting on $L^2$ functions on a torus.
1031: As in the proof of Proposition \ref{p:1}, we now see that 
1032: \begin{equation}
1033: \label{eq:mic}
1034:  \| \Phi_\xi \chi {\chi}_0 u \|_{L^2} \leq C \int_\omega | \chi_0 u |^2 
1035: + {\mathcal O} ( h^{\infty } ) \,,
1036: \end{equation}
1037: where $ \omega $ is a neighbourhood of $ \nabla \chi $ (in
1038: the calculus of semi-classical pseudo-differential operators). 
1039: Since the semi-classical defect measure of $ \Phi_\xi \chi {\chi}_0 u $ (which is $|\Phi_\xi \chi {\chi}_0|^2\times \mu$) was assumed to be
1040: non-zero, \eqref{eq:mic} shows that the measure of $ {\chi}_0 u \rest_\omega $
1041: is non zero and consequently there is a point in the intersection of the supports of 
1042: $\mu$ and ${\chi}_0 u \rest_\omega$. But $\mu$ is invariant
1043: by the flow (as long as it does not intersect the obstacle) and hence, once we choose $\epsilon$, $\delta$ small enough 
1044: such that all the cut-offs above are very close to the
1045: boundary of $ R$, its support can be made intersect any neighbourhood of $ \partial {\mathcal O} $.
1046: \end{proof}
1047: 
1048: Now, from the above theorem, we see the following simple, but important consequences:
1049: 
1050: \begin{rem}
1051: Let $S= {\mathbb T}^2_{a,b} \setminus {\mathcal O}$ where $ {\mathcal O}$ is sufficiently smooth in the case
1052: of Neumann boundary conditions, but otherwise lacking restrictions.  Then, for $V$ any open neighborhood of 
1053: $\partial {\mathcal O}$, and $u$ a solution of $-h^2 \Delta u = E(h) u$ as above, then (4.1) is satisfied.  
1054: This follows from the above argument as the convexity of the obstacle was never used.
1055: Thus, the result holds for any obstacle (even connectedness is not assumed here)
1056: and is applicable to the special case of pseudointegrable billiards (see for instance
1057: \cite{Bog} for motivation and description).  In the next section, we use an argument similar to that above 
1058: in order to say even more about concentration along trajectories in specific pseudointegrable billiards. 
1059: By an elementary reflection principle, the result also holds for an obstacle inside a square with Dirichlet 
1060: or Neumann conditions on the boundary of the square.
1061: \end{rem}
1062: 
1063: \begin{rem}
1064: The proof above gives in fact the following estimate for any open neighbourhood, say $V$, of the obstacle:
1065: \begin{equation}
1066: \begin{gathered}
1067: \exists C; \forall u,\, f \in L^2(S) \text{ solutions of } (-\Delta + \lambda)u =f, \qquad u \rest_{\partial S} =0\\
1068: \|u\|_{L^2(S)}\leq C \left( \|f\|_{L^2(S)}+ \|u \11 _V\|_{L^2(V)}\right)
1069: \end{gathered} 
1070: \end{equation}
1071: and according to~\cite[Theorem 4]{BZ2}, this implies that the Schr{\"o}dinger equation in $S$ 
1072: is exactly controllable by $V$ in finite time. In fact, by working on the time evolution equation, 
1073: we could strengthen this result allowing an arbitrarily small time. 
1074: \end{rem}
1075: This latter result was previously known~\cite{Ja} 
1076: for the particular case $\Theta= \emptyset$ ($S= \mathbb{T}^2$) but the proof was 
1077: based on subtle results about Fourier series \cite{Ka}.
1078: 
1079: \def\cprime{$'$}
1080: 
1081: \begin{rem}
1082: As shown in \cite[Theorem 2$'$]{BZ2}, the results of Ikawa and G\'erard on 
1083: scattering by two convex obstacles (see \cite{BZ2} and references given there)
1084: give an estimate on the maximal concentration of an eigenfunction (or a 
1085: quasimode) on a closed orbit in a Sinai billiard. Let $ \chi \in \CI(S;[0,1])  $  be 
1086: supported in  a small neighbourhood of a closed transversally reflecting 
1087: orbit. Then  for any family $ ( - \Delta - \lambda ) u_\lambda = {\mathcal O}
1088: ( \lambda^{-\infty } ) $, $ \| u_\lambda \| = 1 $, 
1089: \[ C \int_S  | u ( x ) |^2 ( 1 - \chi ( x ) ) dx \geq \frac{1}{ \log \lambda } \,, \]
1090: that is a concentration on a closed trajectory, if at all possible, has to 
1091: be very weak.
1092: \end{rem}
1093: 
1094: \section{Pseudointegrable Billiards}
1095: \label{PIB}
1096: 
1097: We define a pseudointegrable billiard to be a plane polygonal billiard with corners whose angles are of the form $\frac{\pi}{n}$, 
1098: for any integer $n$ (see \cite{Bog1}).  
1099: In particular, we will be working with the billiard $P = {\mathbb T}^2_{a,b} \setminus { S}$ where
1100: $ S$ is a slit that is parrallel to a side of the torus but not a closed loop.  In Remark 1, 
1101: we point out that Theorem \ref{t:s} allows us to make statements about the $L^2$ mass
1102: of eigenfunctions in a neighborhood of the slit for pseudointegrable billiards.  For this particular type of billiard,
1103: it would be ideal to state that every eigenfunction must have non-zero mass in a small neighborhood of the edges of the slit (see Fig. \ref{f:4}).
1104: In this section, we prove a weaker result about non-concentration along certain classical
1105: trajectories in $P$ of semiclassical defect measures obtained from eigenfunctions $u$ such that $(-\lambda -\Delta)u = 0$ on $P$.  
1106:  
1107: 
1108: 
1109: \begin{figure}
1110: \include{figure4}
1111: \caption{A pseudointegrable billiard $P$ consisting of a torus with a slit, $S$ along which we have Dirichlet boundary conditions.
1112: We would like to show that eigenfunctions of the Laplacian on this torus must have concentration in the shaded regions $V_1$ and $V_2$.}
1113: \label{f:4}
1114: \end{figure} 
1115: 
1116: 
1117: As with the Sinai billiard, the classical behavior of trajectories must be taken into account in our treatment of this problem.   
1118: There cannot be concentration along trajectories that do not hit the slit as shown by Theorem \ref{t:s}.  If a trajectory 
1119: has irrational slope, it is dense in $P$, and thus has mass near the edges of the slit as in Section \ref{app}.
1120: Therefore, for our purpose, we concern ourselves only with rational trajectories which intersect the slit at some point.  As we are dealing with
1121: periodic boundary conditions, let us consider the
1122: plane tiled with copies of the billiard $P$.  
1123: 
1124: Assume that $P$ is oriented such that $S$ is parrallel to the $y$-axis.  
1125: Let $\gamma \in S^* (P)$ be a trajectory.  Given the natural projection
1126: $$ \pi_1: S^* (P) \rightarrow P,$$
1127: we take $\gamma' = \pi_1 (\gamma)$, or the physical path along which the trajectory travels. 
1128: Consider the projection
1129: $$\tilde{\pi} : \R^2 \rightarrow {\mathbb T}^2_{a,b}.$$
1130: We see that 
1131: $$\tilde{\pi} : \R^2 \setminus \tilde{S} \rightarrow P, \ \text{where} \ \tilde{S} = \tilde{\pi}^{-1} (S).$$
1132: Define $$\pi_2 : S^* (\R^2 \setminus \tilde{S} ) \rightarrow S^* (P)$$
1133: to be the obvious projection.  Let $\tilde{\gamma} = \pi_2^{-1} (\gamma)$.
1134: We can write $$\tilde{\gamma} = \displaystyle\cup_{j=1}^{\infty} \gamma_j,$$ where each $\gamma_j$ is a trajectory 
1135: in $ S^* (\R^2 \setminus \tilde{S})$.  We note that by construction, $\gamma_i \cap \gamma_j = \emptyset$ for $ i \neq j$.  To see this, assume that 
1136: $\gamma_i \cap \gamma_j = (x, \xi)$.  Then, $\gamma_i = \gamma_j$ as they would represent trajectories which travel 
1137: through the same point in the same direction.  Now, let $$\pi^*_1 : S^* (\R^2 \setminus \tilde{S}) \rightarrow \R^2 \setminus \tilde{S}.$$
1138: Select one trajectory from the above union, say $\gamma_1$.  Let $\gamma_1' = \pi^*_1 (\gamma_1)$. 
1139: We see that either $\gamma_1'$ is bounded in the $x$-direction
1140: or $\gamma_1'$ is unbounded in the $x$-direction.  Note that this property then holds for all $\gamma_j$, $j \in {\mathbb N}$.
1141: For a trajectory $\gamma$, if the resulting path $\gamma_1'$ is bounded in the $x$-direction,
1142: we say $\gamma$ is $x$-bounded.  We define $\gamma$ as $x$-unbounded if $\gamma_1'$ is unbounded in the $x$-direction.
1143: See Fig. \ref{f:7} for examples.  Now, we are prepared to state our theorem concerning the billiard $P$.
1144: 
1145: \begin{figure}
1146: \include{figure7}
1147: \caption{Some examples of $x$-bounded and $x$-unbounded trajectories.}
1148: \label{f:7}
1149: \end{figure}
1150: 
1151:   
1152: 
1153: \begin{thm}
1154: \label{t:j}
1155: Let $\gamma$ be an $x$-bounded trajectory on $P= \mathbb{T}^2 \setminus S$.
1156: If $ \Delta $ is the Dirichlet  Laplace operator on $ P $ then 
1157: there exists no microlocal defect measure obtained from the eigenfunctions  on $P$ 
1158: such that ${\rm{supp}} \  (d\mu) = \gamma$.
1159: \end{thm}
1160: 
1161: \begin{proof}
1162: 
1163: Let  $\gamma '$  be as above.  Let $V_\epsilon$ be an $\epsilon$ neighborhood of $\gamma'$.  
1164: If the theorem were false, we would have a sequence of eigenfunctions $u_n$, $\| u_n \|_{L^2} =1$ with the 
1165: property
1166: $$\displaystyle\int_{P \setminus V_\epsilon} | u_n |^2 dx \rightarrow 0,$$ for any $\epsilon.$ We show that
1167: this is impossible.
1168: 
1169: For each $u_n$, we have $(-\Delta - \lambda_n) u_n = 0$, $u_n|_S =0$, $u_n \in L^2(P)$.  
1170: Let $\tilde{\pi}$ be as above.
1171: We define the sequence
1172: $\tilde{u}_n = \tilde{\pi}^{-1} u_n$.  We have $(-\Delta - \lambda_n ) \tilde{u}_n = 0$,
1173: $\tilde{u}_n|_{\tilde{S}} = 0$, and $\tilde{u}_n \in L^2_{\text{per}} (\R^2 \setminus \tilde{S})$.
1174: 
1175: 
1176: If $\pi_2 : S^* (\R^2 \setminus \tilde{S}) \rightarrow S^* (P)$ is as above and 
1177: $\tilde{\gamma} = \pi_2^{-1} (\gamma)$, then $\tilde{u}_n \rightarrow d\tilde{\mu}$ with 
1178: $$\text{supp} \ (d\tilde{\mu}) = \tilde{\gamma} \subset S^* (\R^2 \setminus \tilde{S})\,. $$
1179: 
1180: Now, let $\pi_1^* : S^* (\R^2 \setminus \tilde{S}) \rightarrow \R^2 \setminus \tilde{S}$ be as above.
1181: Select one trajectory, say $\gamma_1$.  As 
1182: $\gamma_1$ is $x$-bounded, ${\gamma}_1' = \pi_1^* ( \gamma_1 )$ is contained in a strip in the plane which is 
1183: infinite in the $y$-direction and bounded in the $x$-direction.
1184: Thus, $\gamma_1'$ is contained in a strip, $C_0$, with minimal width in the $x$-direction.  Then,
1185: $\tilde{u}_n$ satisfies $(-\Delta -\lambda_n) \tilde{u}_n = 0$ on $C_0$, is periodic in the $y$-direction, and
1186: satisfies the following boundary conditions in the $x$-direction:  Dirichlet boundary conditions 
1187: along the slits that intersect the boundary of $C_0$ and periodic boundary conditions otherwise.
1188: 
1189: Without loss of generality, we can choose the $x$-coordinates such that the 
1190: boundaries of $C_0$ are
1191: $x=-R$ and $x=0$.  We can then reflect to a strip, say $\tilde{C}_1$, with boundaries $x=-R$ and $x=R$, by defining a new function on $\tilde{C}_1$ by:
1192: $$ \tilde{u}_n^{(1)} (x,y) =  \left\{ 
1193: \begin{array}{ll}
1194:  \ \ \tilde{u}_n (x,y)  &  x \in [-R,0], \\
1195:  -\tilde{u}_n (-x,y)  & x \in (0,R). 
1196: \end{array} \right.$$
1197: Note that $\tilde{u}_n$ is periodic with period $2R$.
1198: As a result, we have
1199: $$(-\Delta - \lambda_n) \tilde{u}_n^{(1)} = f_n^{(1)}$$ on $\tilde{C}_1$, where
1200: $$f_n^{(1)} = 2 u(0,y) {\delta}_{0} ' (x) - 2 u(R,y) {\delta}_{R} ' (x).$$
1201: We note that $ f_n^{(1)} $ is supported away from the slits, $ \tilde S $.
1202: 
1203: Define 
1204: $$\pi^{\sharp}_1 (x,y) = \left\{
1205:  \begin{array}{ll}
1206:  (x,y) & -R \leq x \leq 0, \\
1207:   ( -x,y) & 0 \leq x \leq R.
1208:   \end{array} \right.$$
1209: If $ \pi_1^\sharp \; : \; \tilde{C}_1 \rightarrow C_0 $, then $$ (\pi_1^\sharp)^{-1} ( \bigcup_j \gamma_j' ) $$ is again a  union 
1210: of paths resulting from disjoint trajectories.  
1211: Now, we iterate this procedure a finite number of times, 
1212: stopping the iteration when the disjoint trajectories in the lift intersect each slit only once.
1213: 
1214: 
1215: After each reflection, we restrict to a new minimal width strip, say $C_i$.  Let us call $\tilde{C}_i$ the strip resulting from the $i$th reflection.
1216: We define $\pi^{\sharp}_i :  \tilde{C}_i \rightarrow C_{i-1}$ for $1 \leq i < N$ such that
1217: $$\pi^{\sharp}_i (x,y) = \left\{
1218:  \begin{array}{ll}
1219:  (x,y) & (x,y) \in C_{i-1}, \\
1220:   ( 2 R_i -x,y) & (x,y) \in C_{i-1} '.
1221:   \end{array} \right.$$
1222: Here, $C_{i-1}'$ is defined as the reflected strip and $x=  R_{i-1}$ is the line of reflection for $\tilde{C}_i$.  We can subsequently
1223: define $f_n^{(i)}$ as a sum of delta functions resulting from jumps that occur after reflection, similar to $f_n^{(1)}$ above.
1224: We also have $\pi^N : \R^2 \rightarrow C_N$, the obvious projection that results after we tile the plane with copies of $C_N$.  
1225: So, we have:
1226: $$\R^2 \stackrel{\pi_N}{\rightarrow} C_N \subset \tilde{C}_N \stackrel{\pi^{\sharp}_{N}}{\rightarrow} C_{N-1} \subset \tilde{C}_{N-1}
1227: \stackrel{\pi^{\sharp}_{N-1}}{\rightarrow} ... 
1228: \stackrel{\pi^{\sharp}_{2}}{\rightarrow} C_1 \subset \tilde{C}_1 \stackrel{\pi^{\sharp}_{1}}{\rightarrow} C.$$
1229: Note that 
1230: \[ \pi_N^{-1} (\gamma_1') = \displaystyle\bigcup_j \gamma_{1,j}', \]
1231: where $\{ \gamma_{1,j}' \}$ is the set of all paths in $C_N$ generated by the trajectory $\gamma_1$ and the periodicity in $y$.
1232: 
1233: 
1234: \begin{figure}
1235: \include{figure6}
1236: \caption{This diagram describes how we "unfold" the eigenfunctions in order to derive a contradiction.}
1237: \label{fig:unf}
1238: \end{figure} 
1239: 
1240: After a finite number of reflections, we "unfolded" ${\gamma}_1 '$ to be a periodic line on a large strip, $C_N$, which does not 
1241: intersect a slit anywhere.  Now, let us choose $\Phi_\xi$, $\chi$, and $\chi_0$ as above in order to cut-off microlocally 
1242: on this strip around $\gamma_1$.  Again, recall that we can set $\chi \chi_0 = \chi$.  As $f_n^{(i)}$ is supported only in between the slits
1243: for each $i \in \mathbb{N}$, $1 \leq i \leq N$, by
1244: choosing $\Phi_\xi$ to commute with the periodic Laplacian, we have
1245: \[  ( -h^2 \Delta_R - E(h) ) \Phi_\xi \chi {\chi}_0 u_n = \Phi_\xi \chi f_n + [-h^2 \Delta_R , \Phi_\xi \chi ]  u 
1246: =\Phi_\xi  [ -h^2 \Delta_R , \chi ] {\chi}_0 u + {\mathcal O} ( h^{ \infty } ) 
1247: \,, \ \  \| u \| = 1 \,. 
1248: \]
1249: Thus, the result follows by contradiction from the proof of Theorem \ref{t:s}.  
1250: \end{proof}
1251: 
1252: \begin{rem}
1253: Though this result only shows non-concentration, the proof of Theorem \ref{t:s} can be used to show if $\gamma$ is an $x$-bounded trajectory and
1254: $u$ is an eigenfunction supported on $\gamma ' = \pi_1 (\gamma)$, then in fact there must be mass at the edges of the slits as desired.
1255: \end{rem}
1256: 
1257: \begin{rem}
1258: If instead of a torus, we had Dirichlet boundary conditions on the boundary of the rectangle as well as the slit, then 
1259: this non-concentration result can also be applied by an elementary reflection principle argument.  
1260: \end{rem}
1261: 
1262: \begin{thebibliography}{10}
1263: 
1264: %\bibitem{BS}
1265: %A. B\"acker, R. Schubert, and P. Stifter. 
1266: %\newblock On the number of bouncing ball modes in billiards.
1267: %\newblock {\em J. Phys. A: Math. Gen.} 30:6783-6795, 1997.
1268: 
1269: \bibitem{BLR} 
1270: C. Bardos, G. Lebeau and J. Rauch.
1271: \newblock 
1272: Sharp sufficient conditions for the observation, control, and stabilization of waves from the boundary.
1273: \newblock {\em SIAM J. Control Optim.} 30:1024--1065, 1992.
1274: 
1275: \bibitem{Bog}
1276: E. Bogomolny, U. Gerland, and C. Schmit. { Models of intermediate spectral statistics,} 
1277: {\em Phys. Rev. E} 59:1315-1318, 1999.
1278: 
1279: \bibitem{Bog1}
1280: E. Bogomolny and C. Schmit. { Structure of Wave Functions of Pseudointegrable Billiards,} 
1281: {\em Phys. Rev. Lett.} 92:244102, 2004.
1282: 
1283: \bibitem{Bu92}
1284: N.~Burq.
1285: \newblock
1286: Control for Schrodinger equations on product manifolds.
1287: \newblock{ \em Unpublished}, 1992
1288: 
1289: \bibitem{BuIMRN}
1290: N.~Burq.
1291: \newblock Semi-classical estimates for the resolvent in non trapping
1292:   geometries.
1293: \newblock {\em Int. Math. Res. Notices}, 5:221--241, 2002.
1294: 
1295: \bibitem{BuGe}
1296:          {N. Burq and P. G{\'e}rard},
1297: \newblock {Condition n{\'e}cessaire et suffisante pour la
1298:        contr{\^o}labilit{\'e} exacte des ondes}.
1299: \newblock{\em Comptes Rendus de L'Acad{{\'e}}mie des Sciences}, {749--752},{t.325, S{\'e}rie I}, 1996
1300: 
1301: 
1302: 
1303: \bibitem{BZ2}
1304: N.~Burq and M. Zworski.
1305: \newblock Geometric control in the presence of a black box.
1306: \newblock {\em JAMS}, to appear.
1307: 
1308: 
1309: \bibitem{BZ3}
1310: N.~Burq and M. Zworski.
1311: \newblock Bouncing ball modes and quantum chaos.
1312: \newblock {\em SIAM Review}, to appear.
1313: 
1314: 
1315: \bibitem{BuLe01}
1316: {N. Burq and G.Lebeau}.
1317: \newblock {Mesures de d\'efaut de compacit\'e, application au syst\`eme
1318:               de {L}am\'e}.
1319: \newblock{\em Ann. Sci. \'Ecole Norm. Sup. (4)}, No {34}, 817-870, 2001.
1320: 
1321: \bibitem{Do}
1322: H.~Donnelly.
1323: \newblock Quantum unique ergodicity.
1324: \newblock {\em Proc. Amer. Math. Soc.} 131:2945-2951, 2003.
1325: 
1326: \bibitem{Zw}
1327: L. C. Evans and M. Zworski.
1328: \newblock
1329: Lectures on Semiclassical Analysis.
1330: \newblock {\em Unpublished Lecture Series}, 2003.
1331: 
1332: \bibitem{GeLe93}
1333: {P. G{\'e}rard and E. Leichtnam}.
1334: \newblock {Ergodic Properties of Eigenfunctions for the {D}irichlet
1335:                   Problem}.
1336: \newblock{\em Duke Mathematical Journal,} No 71, 559--607, 1993
1337: 
1338: \bibitem{Ha}
1339: A. Haraux.
1340: \newblock S\'eries lacunaires et contr\^ole semi-interne des vibrations
1341:               d'une plaque rectangulaire.
1342: \newblock{\em J. Math. Pures Appl.} 68-4:457--465, 1989.
1343: 
1344: %\bibitem{Hor} L. H{\"o}rmander.
1345: %\newblock
1346: %The Analysis of Linear Partial Differential Operators. Vol. III, IV. 
1347: %\newblock Springer-Verlag, Berlin, 1985.
1348: 
1349: \bibitem{Ja}
1350: S. Jaffard.
1351: \newblock Contr\^ole interne exact des vibrations d'une plaque rectangulaire. 
1352: \newblock {\em Portugal. Math.} 47 (1990), no. 4, 423-429.
1353: 
1354: \bibitem{Ka} 
1355: J.P. Kahane.
1356: \newblock
1357: Pseudo-p\'eriodicit\'e et s\'eries de Fourier lacunaires.
1358: \newblock {\em Annales Sc. de l'Ecole Normale Sup\'erieure} 79, 1962.
1359: 
1360: 
1361: \bibitem{MeSj}
1362:  {R.B. Melrose and J. Sj{\"o}strand}.
1363: \newblock {Singularities of Boundary Value Problems {I \& II}},
1364: \newblock {\em Communications in Pure Applied Mathematics}, {31 \& 35}, {593- 617 \& 129-168}, {1978 \& 1982}.
1365: 
1366: 
1367: %\bibitem{Mi03}
1368: %L.~Miller. 
1369: %\newblock 
1370: %How violent are fast controls for {S}chr\"odinger
1371: %equation?
1372: %\newblock {\em preprint}, 2003.
1373: 
1374: \bibitem{Ze}
1375: S. Zelditch.
1376: \newblock 
1377: Quantum unique ergodicity.
1378: \newblock
1379: {\tt math-ph/0301035}
1380: 
1381: \end{thebibliography}
1382: \end{document}
1383: