math0302234/HKL.tex
1: %%%%%%%%
2: 
3: %%%%%
4: 
5: %%%%%%%%%%%%%
6: %\documentclass[10pt,oneside,amscd,amssymb,epsf]{amsart}
7: % epsf not used
8: \documentclass[10pt,oneside,amscd,amssymb,epsf]{amsart}
9: 
10: % delete 'oneside' for final version
11: 
12: % \usepackage[notcite]{showkeys}
13: \usepackage{amsthm}
14: \usepackage{graphicx}
15: \usepackage{pb-diagram, Bdiagx}
16: \usepackage{psfrag}
17: 
18: %\usepackage{showkeys}
19: 
20: % Delete the next two lines for single spaced copy.
21: 
22: %\parskip 1.75\parskip plus 3pt minus 1pt
23: %\renewcommand{\baselinestretch}{1.5}  (this is from submitted copy)
24: 
25: \newcommand{\TryPackage}[3]{\IfFileExists{#1.sty}{\usepackage{#1}
26: #2}{#3}}
27: \TryPackage{mathrsfs}{\renewcommand{\mathcal}{\mathscr }}{%
28:         % else try euler fonts
29:         \TryPackage{eucal}{}{}}
30: 
31: %\pagestyle{plain}
32: \pagenumbering{arabic}
33: %\footskip .5in
34: %\textheight 8.30in
35: %\headheight .4in
36: 
37: % switch to 6 in for final copy
38: \textwidth 6in
39: %\textwidth 4.5in
40: 
41: \textheight 8.5in
42: \evensidemargin .2in
43: \oddsidemargin .2in
44: \topmargin .25in
45: \headsep .2in
46: \headheight 0.2in
47: \footskip .5in
48: 
49: \begin{document}
50: \title[Calderon Projector for the Hessian ...]{Calderon Projector for
51: the
52: Hessian of the perturbed Chern-Simons function on a 3-manifold with
53: boundary}
54: 
55: \author{Benjamin Himpel}
56: \address{Department of Mathematics, Indiana University, Bloomington, IN
57: 47405}
58: \email{bhimpel@indiana.edu}
59: \urladdr{http://mypage.iu.edu/$\sim$bhimpel}
60: 
61: \author{Paul Kirk}
62: \address{Department of Mathematics, Indiana University, Bloomington, IN
63: 47405}
64: \email{pkirk@indiana.edu}
65: \urladdr{http://mypage.iu.edu/$\sim$pkirk}
66: 
67: \author{Matthias Lesch}
68: \address{Universit\"at zu K\"oln, Mathematisches Institut, Weyertal
69: 86--90,
70: D--50931 K\"oln}
71: \email{lesch@mi.uni-koeln.de}
72: \urladdr{http://www.mi.uni-koeln.de/$\sim$lesch}
73: 
74: \newcommand{\cormark}{\marginpar{$\Leftarrow$}}
75: \date{Feb. 18, 2003}
76: \begin{abstract} {The existence and continuity for the Calder\'on
77: projector of the perturbed odd signature  operator on a 3-manifold is
78: established. As an application we give a new proof of a result of Taubes
79: relating the mod 2 spectral flow of a family of operators on a  homology
80: 3-sphere  with the difference in local intersection numbers of the
81: character varieties  coming from a Heegard decomposition. }
82: \end{abstract}
83: 
84: \thanks{The second named author gratefully acknowledges the support of
85: the
86: National Science Foundation under   grant no. DMS-9971020.}
87: 
88: \maketitle
89: 
90: % MACROS
91: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
92: % Macros included by Matthias
93: \renewcommand{\tilde}{\widetilde}
94: \newcommand{\Cald}{Calder\'on projector}
95: \newcommand{\cV}{{\mathcal V}}
96: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
97: 
98: \newcommand{\const}{\mbox{const}}
99: \newcommand{\lto}{\longrightarrow}
100: % Greek
101: \newcommand{\al}{\alpha}
102: \newcommand{\be}{\beta}
103: \newcommand{\ga}{\gamma}
104: \newcommand{\de}{\delta}
105: \newcommand{\ep}{\epsilon}
106:        \renewcommand{\th}{\theta}
107: \newcommand{\la}{\lambda}
108: \newcommand{\La}{\Lambda}
109: \newcommand{\om}{\omega}
110: \newcommand{\si}{\sigma}
111: \newcommand{\Ga}{\Gamma}
112: \newcommand{\Om}{\Omega}
113: \newcommand{\Si}{\Sigma}
114: \newcommand{\etab}{\tilde{\eta}}
115: \newcommand{\tsig}{\tilde{\sigma}}
116: \newcommand{\hal}{\hat{\alpha}}
117: 
118: \newcommand{\grad}{\mbox{grad}}
119: \newcommand{\gra}{\mbox{grad}^\sim}
120: 
121: \newcommand\tr{\mathop {\rm tr}}
122: 
123: \newcommand{\tlam}{\widetilde \lambda}
124: 
125: % Black-board bold
126: \newcommand{\zz}{{\mathbb Z}}
127: \newcommand{\rr}{{\mathbb R}}
128: \newcommand{\cc}{{\mathbb C}}
129: \newcommand{\qq}{{\mathbb Q}}
130: \newcommand{\hh}{{\mathbb H}}
131: 
132: % Caligraphic
133: \newcommand{\cA}{{\mathcal A}}
134: \newcommand{\cU}{{\mathcal U}}
135: \newcommand{\cE}{{\mathcal E}}
136: \newcommand{\cD}{{\mathcal D}}
137: \newcommand{\cC}{{\mathcal C}}
138: \newcommand{\cL}{{\mathcal L}}
139: \newcommand{\cB}{{\mathcal B}}
140: \newcommand{\cF}{{\mathcal F}}
141: \newcommand{\cG}{{\mathcal G}}
142: \newcommand{\cM}{{\mathcal M}}
143: \newcommand{\cH}{{\mathcal H}}
144: \newcommand{\cT}{{\mathcal T}}
145: \newcommand{\cO}{{\mathcal O}}
146: \newcommand{\cP}{{\mathcal P}}
147: % Hats
148: \newcommand{\hA}{\widehat{A}}
149: \newcommand{\hB}{\widehat{B}}
150: 
151: % Tildes
152: \newcommand{\tB}{{\widetilde{\mathcal B}}}
153: \newcommand{\tM}{{\widetilde{\mathcal M}}}
154: \newcommand{\tC}{{\widetilde{C}}}
155: 
156: % Gothic
157: %\newcommand{\hh}{{\mathfrak h}}
158: %\newcommand{\hhp}{{{\mathfrak h}^\perp}}
159: % Operators
160: \newcommand{\im}{\operatorname{im}}
161: \newcommand{\Map}{\operatorname{Map}}
162: \newcommand{\hol}{\operatorname{{\it hol}}}
163: \newcommand{\proj}{\operatorname{Proj}}
164: \newcommand{\Mas}{\operatorname{Mas}}
165: \newcommand{\Spec}{\operatorname{Spec}}
166: \newcommand{\Hom}{\operatorname{Hom}}
167: \newcommand{\Hol}{\operatorname{Hol}}
168: \newcommand{\crit}{\operatorname{Crit}}
169: \newcommand{\hess}{\operatorname{Hess}}
170: \newcommand{\ind}{\operatorname{ind}}
171: \newcommand{\image}{\operatorname{image}}
172: \newcommand{\ad}{\operatorname{ad}}
173: \newcommand{\End}{\operatorname{End}}
174: \newcommand{\Sign}{\operatorname{Sign}}
175: \newcommand{\SF}{\operatorname{SF}}
176: \newcommand{\Id}{\operatorname{Id}}
177: \def\eqref#1{(\ref{#1})}
178: 
179: \newcommand{\del}{\partial}
180: 
181: %\def\image{\text{\rm image}}
182: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
183: % Makros by Matthias 2002-Jan
184: \newcommand{\chern}{\operatorname{ch}}
185: \renewcommand{\eqref}[1]{\textnormal{(\ref{#1})}}
186: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
187: 
188: %\newcommand{\bbl}{{[\![}}
189: %\newcommand{\bbr}{{]\!]}}
190: %\newcommand{\bbA}{{\bbl A \bbr}}
191: %\newcommand{\minA}{{\widehat{A}^-}}
192: %\newcommand{\maxA}{{\widehat{A}^+}}
193: %\newcommand{\minB}{{\widehat{B}^-}}
194: %\newcommand{\maxB}{{\widehat{B}^+}}
195:        \newcommand{\opX}{{\bar{X}}}
196: 
197:    \theoremstyle{definition}
198: 
199: \newtheorem{defn}{Definition}[section]
200: \theoremstyle{plain}
201: \newtheorem{lemma}[defn]{Lemma}
202: \newtheorem{thm}[defn]{Theorem}
203: \newtheorem{prop}[defn]{Proposition}
204:        \newtheorem{cor}[defn]{Corollary}
205:      \theoremstyle{definition}
206: \newtheorem{remark}[defn]{Remark}
207: 
208: \numberwithin{equation}{section}
209: 
210: \section{Introduction}
211: 
212: The purpose of this article is to establish the existence and continuity
213: of the Calder\'on projectors for families of perturbed signature
214: operators
215: on a 3-manifold with boundary.  We explain briefly the motivation.
216: 
217: The Morse theory  of the Chern-Simons function on the space of $SU(n)$
218: connections over a 3-manifold leads to the construction of topological
219: invariants, notably Taubes' construction of Casson's invariant
220: \cite{taubes}, Floer's instanton homology \cite{floer}, and more
221: recently
222: $SU(3)$ and $SU(n)$ extensions of Casson's invariant \cite{herald-boden,
223: BHK-int, CLM, herald2}.  These invariants are defined using the Morse
224: index of the Hessian of the Chern-Simons function at a critical point.
225: To
226: make this rigorous Taubes showed (in the $SU(2)$ setting) how to
227: understand this ill-defined Morse index in terms of the spectral flow of
228: the odd signature operator coupled to a path of connections. Most
229: importantly for the present article, he also showed how to construct
230: perturbations of the Chern-Simons function in order to make it suitably
231: non-degenerate.
232: 
233: Taubes' ideas form the blueprint for   generalizations of Casson's
234: invariant. In particular the definition of the generalized Casson
235: invariants involve the spectral flow of 1-parameter families of
236: perturbed
237: odd signature operators.  These perturbed odd signature operators (which
238: we describe in detail below) are compact perturbations of Dirac
239: operators,
240: but they are not differential operators.
241: 
242: With the construction of a Morse-theoretic topological invariant
243: established, the problem arises of computing the invariant to discover
244: what information it contains.  Casson's surgery formula and Floer's
245: exact
246: triangle are examples of computational formulas which make it possible
247: to
248: compute their invariants. In general one wishes to have surgery formulas
249: for the generalized Casson invariants.  This leads to the problem of
250: computing the spectral flow of the Hessian of the Chern-Simons function between
251:   critical points in terms of cut--and--paste  data for the underlying
252: 3-manifold.
253: 
254: There is a good cut--and--paste theory for the spectral flow of Dirac
255: operators which make it possible to carry out  calculations.
256: Nicolaescu's article \cite{nico} gives a splitting formula for the
257: spectral flow of a family of Dirac operators. This formula equates the
258: spectral flow to an infinite--dimensional Maslov index corresponding to
259: the
260: associated paths of Calder\'on projectors coming from two parts in a
261: decomposition of the manifold. (Roughly speaking, the \Cald\  for a
262: Dirac
263: operator $D$ is the projection in
264: $L^2(\partial X)$ onto the subspace of boundary values of solutions to
265: the equation $D\phi=0$.)
266:     Nicolaescu's result, combined with some techniques introduced in
267: \cite{daniel-kirk}, sufficed to compute $SU(3)$ Casson invariants for
268: certain Seifert fibered homology 3-spheres in
269: \cite{BHKK,CLM}, where one could avoid using perturbations (at least
270: near
271: the reducible critical points). However, the presence of singularities in the flat moduli space 
272: complicates calculations for more general
273: homology spheres, requiring the use of perturbations. To apply the powerful
274: methods of \cite{nico} one needs an extension of the results concerning
275: the existence and continuity of the Calder\'on projectors for families
276: of
277: operators more general than Dirac operators; in fact general enough to
278: include the perturbed odd signature operator. This is the task
279: accomplished in this article. Our main result is the following, which
280: combines Theorems
281: \ref{continuityCald} and \ref{Poissonop-a}.
282: 
283: \bigskip
284: 
285: \noindent{\bf Theorem.} {\em Consider the class $\cD$ of Dirac operators
286: on $E\to X$ for a manifold with boundary $X$ which are in cylindrical
287: form
288: near
289: $\partial X$,  and the class $\mathcal{V}$ of auxiliary operators
290: satisfying the pseudolocality property
291: \textnormal{\eqref{pseudolocality}}.  Then for $D\in \cD$ and $V\in
292: \mathcal{V}$ the Calder\'on projector of $D+V$ is well defined.
293: 
294: We say that a family $D_t\in\cD$ varies continuously if all coefficients
295: of $D_t$ (in any local chart) vary continuously. Moreover, we equip
296: $\mathcal{V}$ with the usual norm topology of $\cB(L^2(X,E))$. Then for
297: each
298: $n\in\mathbb{Z}_+$ the map
299: \[ \bigl\{(D,V)\in\cD\times\mathcal{V}\,\big|\, \dim Z_0(  D+
300: V)=n\bigr\}
301:       \longrightarrow \cB(L^2(\Sigma,E)), \quad (D,V)\mapsto C(  D+
302: V)
303: \] taking a pair $(D,V)$ to the  Calder\'on projector $C$  for the operator
304: $D+V$ is continuous. }
305: 
306: \smallskip
307: 
308:   In the statement, $Z_0(D+V)$   is a vector space isomorphic to $K\oplus
309: K$, where
310: $K$ denotes the space of sections $\phi$ which vanish on a collar of the
311: boundary and which satisfy  $(D+V)\phi=0$.  For Dirac operators the
312: unique continuation property implies that
313: $Z_0=0$. A novel feature of our approach is that we do not require
314: the unique continuation property to hold.
315: 
316: \bigskip
317:  
318: We are chiefly concerned with the application of this result to the
319: perturbed odd signature operator arising from the Morse theory of the
320: Chern-Simons function, as this will be  used crucially in a forthcoming
321: article \cite{BHK-pqr}. (In fact the completion of that article was held
322: up by  the need for the main result of the present article.)
323: Unfortunately
324: there is no suitably precise exposition of the various constructions
325: which
326: arise in this context and so we include  the details necessary  to
327: set up
328: the problem  precisely and also for the convenience of the reader.  One
329: of our
330: purposes  is to explicitly provide justification for comments like ``in
331: the non-generic
332: case one must perturb, but the arguments carry through,''    in the
333: context
334: of cut--and--paste techniques for spectral flow.
335: 
336: It is our hope that in addition to providing the proof of the theorem
337: stated
338: above this article complements  the  exposition \cite{Herald1, herald2}.
339: Thus in Section
340: \ref{set-up} we  explain the approach used in the Morse-theoretic study
341: of
342: the Chern-Simons function, with an emphasis on   how the perturbations
343: are constructed and on the effect  of perturbing on the Hessian of the
344: Chern-Simons function. We also include proofs of   some  (known) results
345: concerning the pseudolocality of the perturbation (Proposition
346: \ref{lem2.1}) and an estimate concerning the  size of the perturbation
347: (Lemma \ref{estimate}) that we require  to apply the result stated above
348: to the context of the  Hessian of the perturbed Chern-Simons function.
349: 
350: Theorem \ref{thm5.3} is our desired application. In this statement,
351: $D_{A,f}$ denotes the
352: perturbed twisted odd signature operator coupled to a connection $A$
353: and a perturbation $f$. It is
354: essentially the Hessian of the perturbed Chern-Simons function
355: $cs+f:\cA\to \mathbb{R}$.
356: 
357: \bigskip
358: 
359: \noindent{\bf Theorem \ref{thm5.3}.} {\em Let $X$ be a compact
360: 3-manifold
361: with boundary and $\cA$ the space of   connections in a principal
362: $SU(n)$
363: bundle over $X$. Then there exists a neighborhood
364: $U\subset \cA$ of the space of flat connections, and an
365: $\epsilon>0$  so that if
366: $B_{\cP}(0,\epsilon)$ denotes the
367: $\epsilon$ ball around $0$ in the space of perturbations $\cP$,  the map
368: $U\times B_{\cP}(0,\epsilon)\to \cB(L^2(\Sigma,E))$ sending a pair
369: $(A,f)$ to the Calder\'on projector $P_{A,f}$ of $D_{A,f}$   is
370: continuous.}
371: 
372: \bigskip
373: 
374: We finish the article   by giving a new proof in Section
375: \ref{application}, using our methods, of the main step in Taubes'
376: argument
377: that his invariant
378: equals Casson's invariant.
379: 
380: \bigskip
381: 
382: \noindent{\bf Theorem} (Taubes){\bf .} {\em The mod 2 Spectral flow of
383: the
384: perturbed odd signature operator between two generic perturbed flat
385: connections on a homology 3-sphere equals zero if and only if the
386: corresponding  local intersection numbers of the $SU(2)$ character
387: varieties of the Heegard handlebodies in the $SU(2)$ character variety
388: of
389: the Heegard surface are equal. }
390: 
391: \bigskip
392: 
393:   The strategy of our proof is to use an adiabatic stretching of a collar
394: of the Heegard surface  and symplectic reduction to reduce the
395: infinite--dimensional Maslov index (which equals the sign Taubes assigns
396: to
397: a critical point) to a finite--dimensional Maslov index in  the character
398: variety of a Heegard surface  (which determines Casson's sign). Having
399: established in  Theorem
400: \ref{thm5.3} that the Calder\'on projectors exist and vary  nicely, the
401: first step in the argument uses this and Nicolaescu's theorem
402: \cite{nico} to identify spectral flow with  the infinite--dimensional
403: Maslov index of the  Calder\'on projectors. The rest of the argument is
404: a study of the effect of stretching and a corresponding symplectic
405: reduction,  as analyzed in
406: \cite{daniel-kirk}.
407: 
408: The authors thank  H. Boden, G. Daskalopoulos, C. Herald, T. Mrowka,
409: and K.
410: Wojciechowski for helpful discussions.
411: 
412: \section{basic set-up}\label{set-up}
413: \subsection{Connections on a 3-manifold}
414: 
415: We begin by recalling the  spaces and operators of interest.
416: 
417: Let $X$ be a compact 3-manifold.  Fix a principal $SU(n)$ bundle  $P$
418: over
419: $X$. We associate two vector bundles to $P$: the $\cc^n$ vector bundle
420: $E\to X$ associated   to the standard representation
421: $SU(n)\to GL(\cc^n)$, and the Lie algebra bundle
422: $\ad P\to X$ associated to the adjoint representation $SU(n)\to
423: GL(su(n))$
424: of
425: $SU(n)$ on its Lie algebra $su(n)$.  Notice that the bundle $P$ is
426: trivializable since
427: $\pi_1(SU(n))$ and $\pi_2(SU(n))$  are trivial.
428: 
429: We let $\cA$ denote the space of smooth connections on $P$. A connection
430: $A\in \cA$ defines  a covariant derivative
431: $$d_A:C^\infty(X;E)\to \Om^1(X;E).$$ It is  convenient to view $\ad
432: P$
433: as a subbundle of  $\End(E)$. Then the covariant derivative on $\ad P$
434: associated to $A$  (which we also denote by $d_A:C^\infty(X,\ad P)\to
435: \Om^1(X;\ad P)$)  satisfies
436: $$(d_A(\phi))(e)=d_A(\phi(e))-\phi(d_A(e))\  \text{ for } \
437: \phi\in C^\infty(X;\ad P), \ e\in C^\infty(X;E).$$ The covariant
438: derivatives    extend  to $p$-forms $d_A:\Om^p(X; E)\to
439: \Om^{p+1}(X; E)$  and  $d_A:\Om^p(X; \ad P)\to
440: \Om^{p+1}(X; \ad P)$ by requiring the graded Leibnitz rule to hold. The
441: operator $d^2_A:\Omega^p(X;E)\to\Omega^{p+2}(X;E)$ has order zero and
442: is
443: given by multiplication by the   curvature form  $F(A)\in
444: \Omega^2(X;ad P)$.
445: 
446:      The space
447: $\cA$ is an affine space modeled  on  the space of Lie algebra valued
448: 1-forms
449: $\Om^1(X;\ad P)$: Fixing a connection $A_0$ yields an identification
450: \begin{equation}\label{eq2.1}\cA\xrightarrow\cong \Om^1(X;\ad P),
451: \ A\mapsto A-A_0.\end{equation} It is sometimes convenient to  fix a
452: trivialization $P\cong X\times SU(n)$; then we can take $A_0$ to be the
453: trivial (product) connection which we denote by $\Theta$. The space
454: $\cA$
455: is acted on by the   group of (smooth) gauge transformations
456: $\cG=C^\infty(X;P\times_{Ad} SU(n))$ in the usual way: $d_{gA}(e)=
457: g(d_A(g^{-1}(e)))$.
458: 
459: We fix a Riemannian metric on $X$. If the boundary of $X$ is non-empty
460: we
461: will always assume that a collar of the boundary $\del X$ in $X$ is
462: isometric to $[0,1)\times \del X$.
463: 
464: The $L^2$ inner product on
465: $\Om^i(X;\ad P)$ is then defined in terms of the Hodge $*$ operator on
466: $X$ by the formula
467: \begin{equation}\label{eq2.2}\langle b,c\rangle_{L^2}=-\int_X tr(b\wedge
468: *c).\end{equation}
469: 
470: For analytical reasons one must at times complete the spaces $\cA$ and
471: $\cG$ in appropriate Sobolev norms.
472:      Thus we fix some $s\ge 1$ and take
473: $\cA_s$ to be the
474: $L^2_s$ completion of $\cA\cong \Om^1(X;\ad P)$, and $\cG_{s+1}$ to be
475: the
476: closure of
477: $\cG$ in the
478: $L^2_{s+1}$ completion of $C^\infty(\End(E))$.  Here $L^2_s$ denotes
479:      the completion in the Sobolev norm corresponding to sections with
480: $s$
481: derivatives in $L^2$, extended in the standard way to all
482: $s\in \rr$. Then
483: $\cG_{s+1}$ acts smoothly on $\cA_{s}$ (see \cite{FU}). The restriction
484: $s\ge 1$ guarantees that the gauge transformations are continuous.
485: 
486: \subsection{The Chern-Simons function} The Chern-Simons function
487: $cs:\cA\to
488: \rr$ is defined in terms of a trivialization $P\cong X\times SU(n)$ as
489: follows. To a connection
490: $A\in \cA$ one can assign the $su(n)$-valued 1-form
491: $$a:=A-\Theta\in
492: \Om^1(X;\ad P)\cong\Om^1(X)\otimes su(n).$$ One then defines
493: $$cs(A)= \frac{1}{8\pi^2}\int_X \tr(da \wedge a +\tfrac{2}{3} a\wedge
494: a\wedge a).$$ In this expression (as in Equation \eqref{eq2.2}) the
495: wedge
496: product means the wedging of the differential forms and matrix
497: multiplication of the
498: $su(n)$  coefficients. Note that the definition (and value) of the
499: Chern-Simons function depends on the choice of trivialization of
500: $P$. If $X$ is {\it closed}, then the reduction of $cs(A)$ modulo
501: $\zz$ is independent of the choice of trivialization, and is gauge
502: invariant. In fact, $cs(g\cdot A)= cs(A) + \text{degree}(g) $, where the
503: degree is defined using the fact that
504: $H^3(SU(n);\zz)=\zz$.  Thus when $X$ is closed, one considers the
505: Chern-Simons function as a circle valued function on the quotient
506: $cs:\cA/\cG\to \rr/\zz=S^1$.
507: 
508: In \cite{taubes}, Taubes gave an interpretation of Casson's invariant of
509: homology 3-spheres by studying the Morse theory of the Chern-Simons
510: function. This was generalized in much subsequent work, including the work of
511: Floer
512: \cite{floer} and Herald-Boden \cite{herald-boden}.
513: Thus we look next   at the differential topology of
514: $\cA$,
515: $\cG$,  and $cs$.
516: 
517: The tangent space of $\cA$ at a connection $A$ is canonically identified
518: with $\Om^1(X;\ad P)$.  It is not hard to identify the tangent space of
519: $\cG$  at the identity with $\Om^0(X;\ad P)$. Fixing a connection $A\in
520: \cA$ yields  the {\em action map}
521: $$g_A:\cG\to \cA, \  g\mapsto g\cdot A$$ The differential of this map is
522: just the map $d_A:\Om^0(X;\ad P)\to \Om^1(X;\ad P)$. Precisely, the
523: diagram
524: \begin{equation}\label{eq2.3}
525: \begin{diagram}\dgmag{900}\dgsquash
526: \node{T_1\cG}\arrow{e,t}{dg_A}\arrow{s,l}{\cong}
527:           \node{T_A\cA}\arrow{s,l}{\cong}\\
528: \node{\Om^0(X;\ad P)}\arrow{e,t}{d_A}\node{\Om^1(X;\ad P)}
529: \end{diagram}\end{equation} commutes. Thus the tangent space to the
530: quotient $\cA/\cG$ at $[A]$,
531: $T_{[A]}\cA/\cG$ can be identified with the cokernel of $d_A$ (at least
532: at
533: smooth points of $\cA/\cG$: making this precise requires looking at
534: Sobolev completions and dealing with quotient singularities but we will
535: not emphasize this point here).
536: 
537: \subsection{The Hessian of the Chern-Simons function}\label{sec2.3} {\em
538: In this section
539: we assume that $X$ is a closed 3-manifold}.  This is because the
540: differential operators which arise from the Morse theory of the
541: Chern-Simons function  have a simple  description on a closed manifold
542: which is adequate for our purposes, even though we are ultimately
543: interested in cutting and pasting questions. This section could be
544: written
545: in the context of manifolds with boundary, but in that case the proper
546: way
547: to study the Chern-Simons function is as a section of a determinant
548: bundle
549: as in \cite{RSW} and \cite{Herald1}. (The material of this subsection is
550: well known. Standard references include \cite{taubes,
551: donaldson-kronheimer, FU}.   We include it to set up notation and to
552: prepare the reader for
553: the section on perturbations.)
554: 
555: The differential
556: $dcs_A:T_A\cA\to
557: \rr$ of the Chern-Simons function is computed (see \cite{taubes}) {\em
558: for
559: $X$ closed} as:
560: \[ dcs_A(b)=\frac{1}{4\pi^2}\int_X \tr(F(A)\wedge b),\] with $F(A)$
561: the curvature 2-form of $A$. (Recall that
562: in
563: a  trivialization,
564:      $F(A) =da+a\wedge a$.)
565:      From this one concludes that the critical points of $cs:\cA\to\rr$
566: are
567: those
568: $A$ such that $F(A)=0$, i.e. the {\em flat connections}.
569: 
570: The Hessian of $cs:\cA\to\rr$, $\hess_A(cs):T_A\cA\times T_A\cA\to
571: \rr$  is defined for any connection  $A\in\cA$ using the affine
572: structure
573: of $\cA$. It can be computed explicitly when
574: $X$ is closed (\cite{taubes}):
575: \[
576: \hess(cs)_A(b,c)=-\frac{1}{4\pi^2}\langle *d_Ab,c\rangle.
577: \]
578: 
579: The $L^2$ inner product can be viewed as a Riemannian metric on the
580: infinite--dimensional manifold $\cA$ (or $\cA_s$). Identifying
581: vectors and covectors via 
582: this metric  allows us to define the gradient vector field of
583: $cs$
584:      by the rule $\langle\grad(cs)_A,b\rangle_{L^2}=dcs_A(b)$. Similarly
585: the
586: linearization $H_A:T_A\cA\to T_A\cA$ of $\grad(cs)$ is obtained
587: from the Hessian by the rule
588: $\langle H_A(b),c\rangle_{L^2}=\hess(cs)_A(b,c)$.  From the above
589: formulas
590: one concludes that $$\grad(cs)_A= -\tfrac{1}{4\pi^2}*F(A)$$ and
591: $$H_A(b)= -\tfrac{1}{4\pi^2}*d_A(b).$$
592: 
593: The self--adjoint operator $H_A$ is not elliptic because of the gauge
594: invariance of
595: $cs$. For example, at a flat connection $A$, $d_Ad_A=F(A)=0$ , and so
596: $H_A(d_A(b))=0$ for all $b\in \Om^0(X;\ad P)$. In other words the kernel
597: of $H_A$ contains the infinite--dimensional subspace of tangent vectors
598: to
599: the
600: $\cG$ orbit
601:      through $A$. This causes problems if one wishes to consider the
602: Morse
603: theory of $cs$.
604: 
605: As explained in \cite{taubes} (and is typical in gauge theory), the
606: correct
607: way to resolve this lack of ellipticity is to consider $\grad(cs)$
608: as a
609: vector field   on the quotient
610: $\cA/\cG$, and thus $H_A$ as a family of elliptic self--adjoint
611: operators on
612: $T_{[A]}\cA/\cG$.    Using the slice theorem  and Diagram
613: \eqref{eq2.3} one identifies
614: $T_{[A]}\cA/\cG$ (one has to take care at the singular points) as the
615: orthogonal complement of the image of
616: $d_A:\Om^0(X;\ad P)\to \Om^1(X;\ad P)$, in other words as the kernel of
617: the adjoint $d_A^*:\Om^1(X;\ad P)\to \Om^0(X;\ad P)$.  The Bianchi
618: identity implies that
619: $d_A^*(\grad(cs)(A))=-\tfrac{1}{4\pi^2}d_A^*(*F(A))=0$. If $A$ is a
620: critical point of $cs$ (i.e.~a flat connection), then
621: $$d_A^*(H_A(b))=\tfrac{1}{4\pi^2}*d_A**d_A(b)=
622: \tfrac{1}{4\pi^2}*d_Ad_A(b)=\tfrac{1}{4\pi^2}*F(A)(b)=0.$$
623: But $H_A(b)$
624: is
625: not necessarily in the kernel of $d_A^*$ when $A$ is not a critical
626: point.
627: This motivates setting
628: $$\tilde{H}_A :\ker d_A^*\to \ker d_A^*, \ \tilde{H}_A(b)=
629: -\frac{1}{4\pi^2}\proj_{\ker d_A^*} (*d_A(b))$$ where $\proj$ denotes the $L^2$
630: orthogonal
631: projection. The operator $\tilde{H}_A$ agrees with the restriction of
632: $H_A$ to $\ker d_A^*$ if the connection $A$ is flat.
633: 
634: Then $\tilde{H}_A$ is  Fredholm  (i.e. its extension to
635: $L^2_s(\Omega^1(X;\ad P))\cap \ker d_A^*\to L^2_{s-1}(\Omega^1(X;\ad
636: P))\cap \ker d_A^*$ is Fredholm). Moreover, it is self--adjoint with
637: compact resolvent with respect to the $L^2$ inner product restricted to
638: $\ker d_A^*$. The spaces $\ker d_A^*$ form a smooth vector bundle over
639: each orbit-type stratum of
640: $\cA/\cG$. In particular, Taubes' idea was to use  the spectral flow
641:    of
642: a
643: family of operators $\tilde{H}_{A_t}$ as a substitute for the difference
644: in Morse index of critical points $A_0$ and $A_1$ of the Chern-Simons
645: function, and thereby to define an Euler characteristic for
646: $\cA/\cG$.
647: 
648: (The {\em spectral flow} $\SF(H_t)$ of a 1-parameter family of
649: self--adjoint operators $H_t$ is
650: the integer defined to be the   number of eigenvalues  that cross zero, counted with sign.
651: Some conventions need to be set if the
652: operators in question have kernel at the endpoints of the path, and one
653: needs to
654: establish that the concept is well defined  for appropriately
655: continuous  paths of
656: unbounded operators. A careful construction appropriate for our
657: purposes can can be
658: found in
659: \cite{booss-lesch-phillips}. The basic properties and conventions are
660: explained in
661: \cite{kirk-lesch}.)
662: 
663: Taubes observed that the spectral flow of a family $\tilde{H}_{A_t}$
664: equals the spectral flow of a slightly better behaved family of
665: operators.
666: Consider, for any connection $A$, the twisted de Rham sequence
667: \begin{equation}\label{eq2.4} 0\to \Om^0(X;\ad
668: P)\xrightarrow{d_A}\Om^1(X;\ad P)\xrightarrow{d_A}\Om^2(X;\ad
669: P)\xrightarrow{d_A}\Om^3(X;\ad P)\to 0.
670: \end{equation} The sequence \eqref{eq2.4} is an  elliptic  complex when
671: $A$ is flat.   It is not a complex at a non-flat connection, but can be
672: made into a complex if one substitutes
673: $d'_A:=-4\pi^2 *\tilde{H}_A\circ\proj_{\ker
674: d_A^*}$ for
675: $d_A:\Omega^1(X;\ad P)\to \Omega^2(X;\ad P)$.   When $A$ is flat,
676: $d'_A=d_A$.
677: 
678: Whether or not
679: $A$ is flat, the   {\it odd signature operator}  (obtained essentially
680: by folding up the sequence \eqref{eq2.4})
681:      \begin{equation}
682: \label{foldedupop}\begin{split} &D_A:\Om^0(X;\ad P)\oplus \Om^1(X;\ad
683: P)\to \Om^0(X;\ad P)\oplus
684: \Om^1(X;\ad P)\\ &D_A(b,c)=(d_A^* c, *d_A c+d_A b)
685: \end{split}\end{equation} is a self--adjoint Dirac operator in the
686: sense of   e.g. \cite{booss-w}
687:   for any
688: connection $A$.
689: 
690: One also has a  self--adjoint operator $D'_A(b,c)=(d_A^* c, *d'_A c+
691: d_Ab)$. For $A$ flat, $D_A=D'_A$. For a general irreducible connection
692: $A$,
693: $D_A-D'_A$ is a   relatively compact operator. 
694: In fact, letting $\Delta_A^{-1}:\Om^0(X;\ad P)\to \Om^0(X;\ad P)$ denote the inverse of $d_A^*d_A$  (which is invertible for $A$ irreducible) and observing   that
695: $\proj_{\ker
696: d_A^*}x= x-d_A\Delta_A^{-1}d_A^*x$, we see that 
697: $$(D_A-D'_A)(b,c)= \big( *F(A)\Delta_A^{-1} d_A^*- d_A\Delta_A^{-1}*F(A)+d_A\Delta_A^{-1}*F(A)d_A\Delta_A^{-1}d_A^*\big)(c).$$
698:  (see \cite{taubes} for details).     
699: 
700: 
701:   By decomposing
702: $\Om^0(X;\ad P)\oplus
703: \Om^1(X;\ad P)$ as $\Om^0(X;\ad P)\oplus  \text{Image }d_A\oplus\ker
704: d_A^*
705: $, one sees that   the spectrum of
706: $D'_A$ is the disjoint union of the spectrum of $\tilde{H}_A$ and a
707: symmetric (with respect to $\lambda\to-\lambda$) spectrum.
708: A simple
709: argument then shows that given a path $A_t$ of (say irreducible)
710: connections joining two critical points of $cs$,   the spectral flow of
711: the family
712: $\tilde{H}_{A_t}$ (which plays the role of the difference in  Morse
713: indices of $cs$ at $A_0$ and
714: $A_1$) equals the spectral flow of the family $D_{A_t}$. What is gained
715: is that the spectral flow of $D_{A_t}$ makes sense whether or not $A_t$
716: is a path of irreducible connections, i.e.~even when the subspaces
717: $\ker d_A^*\subset \Omega^1(X; \ad P)$ do not vary continuously.
718: 
719: \subsection{Perturbations} We come now to the construction of
720: perturbations and their effect on
721: the Hessian of the Chern-Simons  function.
722:    The critical point set of
723: $cs$ may be complicated and it is often necessary to perturb the
724: Chern-Simons function to make it as nice as possible.
725: 
726:      Making precise what ``nice" means can be quite difficult in the case
727: of
728: $SU(n)$ for $n>2$, or when the homology of $X$ is complicated. In the
729: case
730: of $SU(2)$ and $X$ a homology sphere  Taubes replaced $cs$ by a function
731: $cs+h$ whose restriction to the top stratum of
732: $\cA/\cG$ (i.e. the irreducible connections; see Section
733: \ref{application})   is Morse in a suitable sense, and then used this to
734: define a topological invariant in the spirit of the Poincar\'e-Hopf
735: theorem,  which he showed equals Casson's invariant. In
736: \cite{herald-boden} a similar analysis is carried out for $SU(3)$ and
737: $X$
738: a homology sphere leading to a definition of an $SU(3)$ Casson
739: invariant;
740: in that case in addition to being Morse on the top stratum of
741: $\cA/\cG$,
742: $cs+h$ must be suitably non-degenerate in higher codimensional strata.
743: In
744: more recent work \cite{herald2} Herald shows how with a suitable choice
745: of
746: perturbation
747: $h$ one can define versions of Casson's invariant for rational homology
748: spheres and $SU(n)$, $n=2,3,4,5$.
749: 
750: There are slightly different approaches to constructing a suitable
751: family
752: of perturbations $h$, but they are similar and our method will apply to
753: each approach.  We mention that the basic requirements that $h$ should
754: satisfy are:
755: \begin{enumerate}
756: \item The perturbation $h$ should be a smooth function on the Hilbert
757: manifold $\cA_1/\cG_2$. 
758: \item The operator 
759: defined by the Hessian of $cs +h$  at $A$ should be a
760: compact perturbation of
761: $\tilde{H}_A$.
762: \item The family of admissible $h$ should be large enough so that one
763: can
764: prove various general position results about $cs+h$ on the strata and
765: their normal bundles of $\cA/\cG$.
766: \end{enumerate}
767: 
768: The following types of perturbations,   due to Taubes and Floer,
769: accomplish these goals, as explained in
770: \cite{herald2}.
771: 
772: First, fix a collection of smooth embeddings $\gamma_i:D^2\times S^1\to
773: X,\ i=1,\cdots,N$. Taubes \cite{taubes} requires that the embeddings
774: have
775: disjoint images, whereas   Floer \cite{floer} and Herald
776: \cite{herald2}  assume that   $\gamma_i(x,1)=\gamma_1(x,1)$ for all
777: $i$ and $x\in D^2$ and that the derivatives of the $\gamma_i$ in the
778: $S^1$
779: direction  at points in $D^1\times \{1\}$ are independent of $i$. (In
780: fact, nothing is lost by assuming that there is an interval $I$ around
781: $1\in S^1$ so that $\gamma_i(x,u)$ is independent of $i$ for $u\in I$.)
782: The results of this article work equally well in   each setting. The
783: difference between them is that it is easier to show that the third
784: condition above holds for   perturbations of the type Floer considers
785: since  the base point issues are easier to control. We will assume, to
786: keep the notation under control, that we are using the approach of
787: Floer,
788: i.e. the common basepoint approach.
789:     We do not address the third condition in this article.  In the
790: language
791: of
792: \cite{Herald1,herald2,herald-boden} the third  condition holds if the
793: set
794: of perturbations is {\em abundant}.
795: 
796: Next,  let $\cP$ denote the set of $C^r$ functions (for some fixed large
797: $r$),
798: $f:SU(n)^N\to \rr$, invariant under the conjugation action of
799: $SU(n)$, i.e. $f(rg_1r^{-1},\cdots,rg_Nr^{-1})=f(g_1,\cdots,g_N)$.
800: 
801: Now let $P|_{D^2}$ denote the restriction of the principal bundle $P$ to
802: the disc $D^2=\gamma_i(D^2\times\{1\})$. Given a smooth connection $A\in
803: \cA$, consider the map
804: $\Phi_A:P|_{D^2}\to SU(n)^N$ defined  by the rule
805: $\Phi_A(p)=(g_1,\cdots ,g_N)$ where $pg_i\in P_x$ is the endpoint of
806: the
807: parallel lift of  the loop $\gamma_i(x,u), \ u\in [0,2\pi]$  with
808: respect
809: to
810: $A$ starting at
811: $p$ (where $x\in D^2$ is the image of $p$ via the bundle projection).
812: Then
813: $\Phi_A(pr)=(r^{-1}g_1 r,\cdots,r^{-1}g_N r)$ for $r\in SU(n)$. It
814: follows
815: that if $f\in \cP$, $f\circ \Phi_A$ descends to a well-defined function
816: on
817: $D^2$.
818: 
819: If the principal bundle is trivialized, $P\cong X\times SU(n)$, then one
820: can view parallel lifting as the holonomy, a function from loops to
821: $SU(n)$, and in our context this allows us to define
822: $$\Hol:\cA\times D^2\to SU(n)^N, \ \ (A,x)\mapsto
823: (\Hol_{\gamma_1(x,-)}(A),\cdots \Hol_{\gamma_N(x,-)}(A)).$$
824:      Thus we denote the function $f\circ \Phi_A$  by
825: $x\mapsto f(\Hol(A,x))$. Notice  that $f(\Hol(A,x))$ is independent of
826: the
827: trivialization, and is unchanged if we change $A$ by a gauge
828: transformation.
829: 
830: Fix a smooth positive cut--off function $\eta$ on $D^2$ which vanishes
831: near
832: the boundary. The space of {\em admissible perturbations} is defined to
833: be
834: the space of functions of the form
835: \begin{equation}\label{eq2.5}h_f:\cA\to \rr, \ A\mapsto
836: \int_{D^2}f(\Hol(A,x))\eta(x)\ d^2x.\end{equation}
837:      Thus we view  the admissible perturbation functions as being
838: parameterized by the (Fr{\'e}chet) space $\cP$.
839: 
840: Taubes and Floer show that $h_f$ extends to a smooth function on
841: $\cA_2$.  Note that $d(cs +h_f)_A= dcs_A + d({h_f})_A $ and
842: $\hess(cs+h_f)_A=\hess(cs)_A+\hess(h_f)_A$.  Thus one can define
843: $$Q_{A,f}:T_A\cA\cong\Om^1(X;\ad P)\to T_A\cA $$ by  the rule
844: $$\langle
845: Q_{A,f}(b),c\rangle=\hess(h_f)_A(b,c)=\tfrac{\partial^2}{\partial t 
846: \partial s}|_{s=t=0}\ h_f(A+sb +tc).$$  Then setting $d_{A,f}:=
847: d_A-4\pi^2
848: *Q_{A,f}$ we see that $-\tfrac{1}{4\pi^2} *d_{A,f}$
849:      is the operator defined by the Hessian of $cs+h_f$ at $A$.
850: 
851: A connection $A$ is called {\em $f$-perturbed flat} if it is a critical
852: point of $cs +h_f$  or, equivalently,  if $*F(A)=4\pi^2
853: \grad(h_f)_A$. As in Section \ref{sec2.3}, we form the sequence
854: \begin{equation}\label{eq2.6}0\to \Om^0(X;\ad P)\xrightarrow{d_A}
855: \Om^1(X;\ad P)\xrightarrow{d_{A,f}}
856:      \Om^2(X;\ad P)\xrightarrow{d_A}
857:       \Om^3(X;\ad P)\to 0.\end{equation} The sequence \eqref{eq2.6} is a
858: complex if $A$ is
859: $f$-perturbed flat, and for any connection $A$ and perturbation $f$ the
860: operator obtained from  \eqref{eq2.6} (compare to \eqref{foldedupop}),
861: \begin{equation}
862: \begin{split} &D_{A,f}:\Om^0(X;\ad P)\oplus \Om^1(X;\ad P) \to
863: \Om^0(X;\ad
864: P)\oplus
865: \Om^1(X;\ad P),\\
866: &D_{A,f}(b,c) =(d_A^* c ,*d_{A,f} c  + d_A b )
867: \end{split}
868: \end{equation}
869: is a self--adjoint operator called the {\em perturbed
870: twisted odd signature operator}.
871: 
872: The operator $*Q_{A,f}$ is independent of the choice of
873: Riemannian metric on $X$ since the definition of
874: $\hess(h_f)_A(b,c)$ does not use the metric, and since
875: $$\hess(h_f)_A(b,c)=\langle Q_{A,f}(b),c\rangle=-\int_X\tr(c\wedge
876: *Q_{A,f}(b)).
877: $$ It follows that $d_{A,f}$ is metric independent. In particular, if
878: $A$
879: is $f$-perturbed flat, then the cohomology of \eqref{eq2.6}, and hence
880: the
881: dimension of the kernel of $D_{A,f}$, is independent of the metric on
882: $X$.
883: 
884: We will give a proof  in Lemma \ref{estimate}  of Taubes' observation
885: that for any
886: $A$ and $f$, the operator
887: $D_{A,f}$ is Fredholm and in fact a compact perturbation of
888: $D_A$. Assuming this,  we conclude the following.
889: 
890: \begin{lemma}\label{independence} Suppose that $f_0$ and $f_1$ are
891: perturbations,
892: $A_0$ and
893: $A_1$ are connections so that $A_0$ is $f_0$ perturbed flat and
894: $A_1$ is
895: $f_1$ perturbed flat.  If $f_t$ is a path of perturbations from $f_0$ to
896: $f_1$ and
897: $A_t$ is a path of connections from $A_0$ to $A_1$ then
898: the spectral flow $\SF(D_{A_t,f_t})$ is independent of the choice of
899: paths
900: $A_t$ and
901: $f_t$ and independent of the  choice of Riemannian metric.
902: \end{lemma}
903: \begin{proof}
904: Since the space of connections and the space of perturbations are
905: contractible and the spectral flow is a homotopy invariant,
906: $\SF(D_{A_t,f_t})$ is independent of the choice of paths $A_t$ from
907: $A_0$ to $A_1$ and path $f_t$ from $f_0$ to $f_1$. Since the dimensions
908: of
909: the kernels of $D_{A_0,f_0}$ and $D_{A_1,f_1}$ are independent of the
910: Riemannian metric, it similarly follows that   $\SF(D_{A_t,f_t})$ is
911: independent of the choice of Riemannian metric.
912: \end{proof}
913: 
914: \subsection{Locality of $Q_{A,f}$}\label{secloc}
915:      The operator $D_{A,f}$ is not a differential operator, but it has
916: the
917: property that it differs from a differential operator by an operator
918: which
919: is  ``localized'' in the neighborhood of the solid tori used to
920: construct the perturbations.
921: 
922: Write
923: $$D_{A,f}(b,c)= D_A(b,c) + (0, -4\pi^2 Q_{A,f}(c))$$ so that $D_{A,f}$
924: differs from the Dirac operator $D_A$ by
925: $-4\pi^2 Q_{A,f}$.
926: 
927: \begin{prop}\label{lem2.1} Let $S\subset X$ denote the union of the
928: images
929: of the embeddings $\gamma_i$, $S=\cup_i\gamma_i(D^2\times S^1)$.
930: \begin{enumerate}
931: \item If $b,b'\in \Om^1(X;\ad P)$ and $b|_{S}=b'|_{S}$, then
932: $Q_{A,f}(b)=Q_{A,f}(b')$.
933: \item For any $b\in\Om^1(X;\ad P)$, $Q_{A,f}(b)$ vanishes outside of
934: $S$.
935: \item If $A,A'$ are connections and $A|_S=A'|_S$, then
936: $Q_{A,f}(b)=Q_{A',f}(b)$ for all $b\in\Om^1(X;\ad P)$.
937: \end{enumerate}
938: 
939: \end{prop}
940: \begin{proof} Suppose that  $b\in\Om^1(X;\ad P)$ vanishes on $S$.  Then
941: for each
942: $c\in\Om^1(X;\ad P)$,
943: $$\langle Q_{A,f}(b),c\rangle=\hess(h_f)_A(b,c)=
944: \tfrac{\partial^2}{\partial s  \partial
945: t}|_{s=t=0}\int_{D^2}f(\Hol(A+sb+tc,x))\eta(x)d^2x=0,$$
946: since the parallel lifts of the loops $\gamma_i(x,u), \ u\in[0,2\pi]$
947: with
948: respect to the connection $A+sb+tc$ are independent of $s$.  Hence
949: $Q_{A,f}(b)=0$. Part (1) now follows since $Q_{A,f}$ is linear and
950: $b-b'$
951: vanishes on $S$.
952: 
953: For part (2), Let $b\in\Om^1(X;\ad P)$ be arbitrary. Let $U\subset X\setminus S$
954:      and choose $c\in\Om^1(X;\ad P)$ supported in $U$.  Then
955:      $$\langle Q_{A,f}(b),c\rangle=\langle b,Q_{A,f}(c)\rangle=0$$
956:      by the same argument given for part (1) since $c$ vanishes on
957:      $S$.  Since the $L^2$ inner product of $Q_{A,f}(b)$ with $c$
958:      vanishes for every $c$ supported on $U$, $Q_{A,f}(b)$ must vanish on
959:      $U$.
960: 
961: Part (3) is clear.
962: 
963: \end{proof}
964: 
965:      Proposition \ref{lem2.1} implies that  the
966: restriction of
967: $D_{A,f}$  to
968:      a codimension 0 submanifold whose boundary misses
969: $S=\cup_i$Image$(\gamma_i)$ is well--defined. To be precise, suppose
970: that  $Y\subset X$ is a compact
971:      codimension 0 submanifold. We consider two cases:
972: $S\subset$interior$(Y)$
973:      and $S\subset X\setminus Y$.
974: 
975:      If $S\subset$interior$(Y)$, $A$ is any connection on $Y$ and
976:        $c$ is a 1-form  on $Y$ with values in $\ad
977:      P|_Y$, extend $A$ and $c$ arbitrarily to $\tilde{A}, \tilde{c}$ on
978: $X$.
979: Then
980:      $Q_{\tilde{A},f}(\tilde{c})$ vanishes outside of $S$ and is
981: independent
982: of the choice of extensions. Thus $Q_{A,f}$ is well-defined as an
983: operator
984: on $Y$.
985:      Since $D_A$ is a differential operator it is locally defined and
986: hence
987: has a well defined restriction to $Y$. It follows that
988: $D_{A,f}=D_A-4\pi^2
989: Q_{A,f}$ is well defined on $Y$.
990: 
991: If $Y\subset X\setminus S$, then a similar argument shows that $D_{A,f}$
992: restricts
993: to the differential operator $D_A$ on $Y$.
994: 
995: \subsection{An estimate}  We next give an  argument to show that
996: $D_{A,f}$ differs from $D_A$ by an operator which is bounded on $L^2$.
997: This result is essentially contained in Taubes' article \cite{taubes}.
998: See
999: also
1000: \cite[Proposition 2.8]{herald-boden}.
1001: 
1002: \begin{lemma} \label{estimate} There exists a constant
1003: $C$   so that for all $A\in
1004: \cA, f\in\cP$, and $b\in \Om^1(X;\ad P)$
1005: $$\|Q_{A,f}(b)\|_{L^2}\leq C \|f\|_{C^2}\|b\|_{L^2}.$$ In particular,
1006: $D_{A,f}$ is a compact perturbation of
1007: $D_A$, viewed as operators $L^2_1\to L^2$.
1008: \end{lemma}
1009: \begin{proof}
1010: 
1011: Denote $\Hol(A+sb+tc,x)$ by $H_x(s,t)$,  so
1012: $H_x:\rr^2\to  SU(n)^N$, and let
1013: $f:SU(n)^N\to \rr$ be invariant under the conjugation action of $SU(n)$.
1014: Then
1015: $$
1016: \hess(h_f)_A (b,c) = \left.\tfrac{\partial^2}{\partial
1017:       s\partial t}\right|_{s=t=0} \int_{D^2} f(H_x(s,t))\eta(x) d^2x =
1018: \int_{D^2}
1019: \left.\tfrac{\partial^2}{\partial
1020:       s\partial t}\right|_{s=t=0}  f(H_x(s,t)) \eta(x)d^2x.
1021: $$
1022: 
1023: We may think of $H_x:\rr^2\to SU(n)^N\subset
1024: \cc^{n^2N}$ and  extend  $f$  to a smooth function
1025: $\cc^{n^2N}\to \rr$.    By the chain and the product rule  for vector
1026: valued functions  we have
1027: \begin{eqnarray*}
1028: \lefteqn{\left.\tfrac{\partial^2}{\partial
1029:       s\partial t}\right|_{s=t=0}  f(H_x(s,t))}\\ & = & D_{H_x(0,0)} f
1030: \circ
1031: \left.\tfrac{\partial^2}{\partial
1032:       s\partial t}\right|_{s=t=0} H_x(s,t) +
1033:       \left(\left.\tfrac{\partial}{\partial s}\right|_{s=0}
1034:       H_x(s,0)\right)^T \circ D_{H_x(0,0)}^2 f \circ
1035: \left.\tfrac{\partial}{\partial t}\right|_{t=0}
1036:       H_x(0,t)
1037: \end{eqnarray*}
1038: 
1039: Then
1040: \begin{eqnarray*}
1041: \lefteqn{\left|\left.\tfrac{\partial^2}{\partial
1042:       s\partial t}\right|_{s=t=0}  f(H_x(s,t))\right|}\\
1043: & \leq & \|D_{H_x(0,0)} f \| \cdot \|
1044: \left.\tfrac{\partial^2}{\partial
1045:       s\partial t}\right|_{s=t=0} H_x(s,t) \|+
1046:       \|\left.\tfrac{\partial}{\partial s}\right|_{s=0}
1047:       H_x(s,0)\| \cdot \|D_{H_x(0,0)}^2 f\| \cdot \|
1048: \left.\tfrac{\partial}{\partial t}\right|_{t=0}
1049:       H_x(0,t)\|
1050: \end{eqnarray*}
1051: 
1052: Given  $c\in \Om^1(X;\ad P)$, $i\in \{1,\cdots ,N\}$,  and
1053: $x\in D^2$,  let
1054: $\tilde c_{i,x}:[0,1]\to su(n)$ be defined as follows.  Trivialize $\ad
1055: P$
1056: over the disc $D^2\subset X$, so $\ad P|_{D^2}\cong D^2\times su(n)$.
1057: Parallel  translation  by $A$ along the path $g_{i,x}:[0,1]\to X,
1058: t\mapsto\gamma_i(x,e^{2\pi i t})$ trivializes $g_{i,x}^*(\ad P)$ over
1059: $[0,1]$, and this way the pullback $g_{i,x}^*(c)$ is viewed as a 1-form
1060: with
1061: $su(n)$ coefficients. Contracting with the unit speed vector field
1062: $\tfrac{\del}{\del t}$ yields the function
1063: $\tilde c_{i,x}:[0,1]\to su(n)$, i.e.~
1064: $\tilde c_{i,x}=g_{i,x}^*(c)(\tfrac{\del}{\del t})$.
1065: 
1066: We have the estimate
1067: $$\int_{D^2} \int_0^1 |\tilde c_{i,x}(\nu)|\eta(x)\ d\nu\ d^2x
1068: \leq \int_{D^2} \int_0^1 |c(\nu)|\eta(x) \ d\nu\ d^2x\leq C_1\|
1069: c\|_{L^1(S^1\times D^2)}\leq C\| c\|_{L^1(X)}$$ for some constants $C_1,
1070: C$ independent of $c$.
1071: 
1072: Given a  connection $B$ let
1073: $h_{i,x}(B)$ denote the holonomy of $B$ around the loop
1074: $\gamma_i(x,-)$. Thus
1075: $$H_x(s,t)=(h_{1,x}(A+s b+t c),
1076: \cdots,h_{1,x}(A+s b+t c)).$$   Lemma 2.6    of
1077: \cite{herald-boden} establishes the formulas:
1078: \begin{eqnarray*}
1079: \left.\tfrac{d}{dt}\right|_{t=0} h_{i,x}(A+t b) & = & h_{i,x}(A)
1080: \int_0^1
1081: \tilde b_{i,x}(\nu) d\nu\\
1082: \left.\tfrac{\partial^2}{\partial s\ \partial t}\right|_{s=t=0}
1083: h_{i,x}(A+s b+tc) & = &
1084: h_{i,x}(A)
1085: \int_0^1\int_0^\nu (\tilde b_{i,x}(\nu)\tilde c_{i,x}(\mu) + \tilde
1086: c_{i,x}(\nu)\tilde b_{i,x}(\mu)) d\mu d\nu.
1087: \end{eqnarray*}
1088: 
1089: Notice that $h_{i,x}(A)\in SU(n)$, and so is bounded independently of
1090: $A$.  Thus  (letting $C$ denote possibly different constants)
1091: $$
1092: \int_{D^2}  |  \left.\tfrac{d}{dt}\right|_{t=0} H_x(s,0) | \ \eta(x)\
1093: d^2x
1094: \leq\sum_{i=1}^N C\|b\|_{L^1(X)} = NC\|b\|_{L^1(X)}\leq
1095: C\|b\|_{L^2(X)}.$$
1096: The last inequality follows from Cauchy-Schwarz.
1097: 
1098: Similarly
1099: \begin{eqnarray*}
1100:      \lefteqn{\int_{D^2} | \left.\tfrac{\partial^2}{\partial
1101:       s\partial t}\right|_{s=t=0} H_x(s,t)  | \ \eta(x) \ d^2x }\\
1102: & \leq  & C \int_{D^2} \int_0^1 |b(\nu)|d\nu \int_0^1|c(\mu)|d\mu d^2x
1103:    \leq C \left\|\int_0^1 |b(\nu)|d\nu\right\|_{L^2(D^2)} \cdot
1104: \left\|\int_0^1|c(\mu)|d\mu\right\|_{L^2(D^2)}\\
1105: & \leq &  C \| b\|_{L^2(S^1 \times D^2)} \cdot \| c \|_{L^2(S^1 \times
1106: D^2)}
1107: \leq C
1108: \|b\|_{L^2(X)}
1109: \cdot \|c\|_{L^2(X) }.
1110: \end{eqnarray*}
1111: 
1112: Also
1113: $$ \|D_{H_x(0,0)} f\| + \|D^2_{H_x(0,0)} f\| \leq C \|f\|_{C^2}.
1114: $$
1115: Thus for any $b,c\in \Om^1(X;\ad P)$,
1116: $$  \bigl|\hess(h_f)_A (b,c)\bigr|  \leq C \|f\|_{C^2}
1117: \|b\|_{L^2} \|c\|_{L^2},
1118: $$
1119: and hence $\|Q_{A,f}\|\le C \|f\|_{C^2}$.
1120: 
1121: For the  last statement of Lemma \ref{estimate}, the difference
1122: $D_{A,f}-D_A:L^2_1\to L^2$ equals the restriction  of
1123: $-4\pi^2Q_{A,f}$ to $L^2_1$. This restriction  can  be expressed as the
1124: composite of the compact inclusion
1125: $L^2_1\subset L^2$ and the bounded operator $-4\pi^2Q_{A,f}$, and hence
1126: is
1127: compact since the compact operators form a 2-sided ideal.
1128: \end{proof}
1129: 
1130: \section{The Poisson operator and the \Cald\ for perturbed Dirac
1131: operators
1132: on manifolds with boundary} A compact manifold with non-empty boundary
1133: always embeds  as a codimension 0 submanifold of a closed manifold
1134: (e.g.~
1135: its double) in such a way that a given principal bundle and $SU(n)$
1136: connection extend. It follows from Proposition \ref{lem2.1} that given:
1137: \begin{enumerate}
1138: \item a 3-manifold with boundary $X$,
1139: \item a principal bundle $P\to X$,
1140: \item a connection $A$ on $P$,
1141: \item a collection of embeddings $\gamma_i:D^2\times S^1\to
1142: $interior$(X)$, and
1143: \item a function $f\in \cP$
1144: \end{enumerate} one can unambiguously define operators $Q_{A,f}$ and
1145: $D_{A,f}$ on
1146: $X$.  In this context we will assume that the connection $A$ is in
1147: cylindrical form near the boundary. This means that in a collar
1148: neighborhood
1149: $[0,\epsilon)\times \partial X$  of $\partial X$, $A$ is the pullback
1150: of a
1151: connection $a$ on $\partial X$ with respect to the projection
1152: $[0,\epsilon)\times \partial X\to \partial X$ to the second factor. Then
1153: the operator $D_{A,f}$ equals $D_A$ on the collar, and this implies that
1154: $D_{A,f}$ takes the Atiyah--Patodi--Singer form (for details see
1155: \cite[Definitions 2.2 and 2.3]{BHKK})
1156: \begin{equation} D_{A,f}|_{[0,\epsilon)\times \partial
1157: X}=D_{A}|_{[0,\epsilon)\times
1158: \partial X}=
1159: \gamma(\tfrac{\partial}{\partial u} + S_a). \label{G4.1}
1160: \end{equation}
1161: 
1162: Here $S_a:\Om^*(\partial X;\ad P)\to \Om^*(\partial X;\ad P)$ is the
1163: tangential operator, given by the formula
1164: \begin{equation}\label{eq4.45}S_a(\alpha_0,\alpha_1,\alpha_2)=(*d_a\al_1
1165: ,
1166: -*d_a\al_0-d_a*\al_2, d_a*\al_1)
1167: \end{equation}
1168:     and $\gamma$ is the bundle isomorphism given by the formula
1169: \begin{equation}\label{eq4.4}
1170: \gamma(\alpha_0,\alpha_1,\alpha_2)
1171: =(-*\al_2,*\al_1,*\al_0)\end{equation}
1172: (in the  formulas \eqref{eq4.4} and \eqref{eq4.45}, $*$ refers to the
1173: Hodge
1174: $*$ operator for the metric on $\partial X$). The bundle isomorphism
1175: $\gamma$ satisfies $\gamma^2=-\Id$.
1176: 
1177: %%MARKE
1178: \subsection{The invertible double}
1179: 
1180: So far we have shown that $D_{A,f}$ is the sum of a Dirac operator $D_A$
1181: and  an  $L^2$--bounded operator $-4\pi^2 Q_{A,f}$. The operator
1182: $Q_{A,f}$
1183: is not a differential operator but it has certain locality properties as
1184: explained in Section \ref{secloc}. On a manifold with boundary we
1185: arranged
1186: so that $D_{A,f}$ has a product form in a collar of the  boundary.
1187: 
1188: In the study of Dirac operators on manifolds with boundary, their
1189: \Cald\ and their Poisson operator play a central role (cf.
1190: \cite{booss-w}).
1191: However, since $Q_{A,f}$ is far from being a differential operator,
1192: existence of the \Cald\ and of the Poisson operator is a priori not
1193: clear.
1194: We therefore treat the existence question and related issues. For the
1195: convenience of the reader our presentation will be more or less
1196: self--contained.
1197: 
1198: Let $X$ be a compact connected Riemannian manifold with boundary
1199: $\partial
1200: X=\Sigma$ and $D$ a self--adjoint Dirac type operator on the Hermitian
1201: vector bundle
1202: $E\to X$. We assume that we are in a product situation, i.e. that there
1203: is
1204: a collar $U=[0,\epsilon)\times \Sigma$ of $\Sigma$ in which $D$ takes
1205: the
1206: form
1207: \begin{equation}
1208:       D=\gamma\bigl(\frac{d}{dx}+B\bigr),
1209: \label{G4.10}
1210: \end{equation} where $\gamma$ is a bundle endomorphism and $B$ is a
1211: first
1212: order self--adjoint elliptic differential operator on $E|\Sigma$. One
1213: has
1214: the fundamental relations
1215: \begin{equation}
1216:      \gamma^2=-\Id,\quad \gamma^*=-\gamma,\quad\text{and}\quad
1217:      B\gamma=-\gamma B.\label{G4.11}
1218: \end{equation} Note that $D_A$ is of this form \eqref{G4.1}.
1219: 
1220: To model $Q_{A,f}$ we assume abstractly that we are given an auxiliary
1221: self--adjoint bounded linear operator $V$ in $L^2(X,E)$. In most cases
1222: $V$
1223: will be a pseudo--differential operator of order $0$, but this is not
1224: assumed here. More importantly, we assume the following locality
1225: property
1226: (cf. Sec. \ref{secloc}).
1227: \begin{equation}
1228: \text{There is a compact domain
1229: $S\subset X\setminus [0,\epsilon)\times \Sigma$ such that $M_\varphi
1230: V=0$
1231: for all $\varphi\in C^\infty_0(X\setminus S)$.}
1232: \label{pseudolocality}
1233: \end{equation}
1234: Here $M_\varphi$ denotes the operator of multiplication by $\varphi$.
1235: Note that by self--adjointness we automatically also have
1236: \begin{equation}
1237:       VM_\varphi=0\quad\text{for}\quad \varphi\in C^\infty_0(X\setminus
1238: S).
1239:       \label{G4.13}
1240: \end{equation} Moreover, if $f,g\in L^2(E)$ such that $f|W=g|W$ in a
1241: neighborhood $W$ of $S$ then choosing $\psi\in C^\infty_0(W)$ with
1242: $\psi\equiv 1$ in a neighborhood of $S$ we find
1243: \begin{equation} V(f-g)=V((1-\psi)(f-g))=(VM_{1-\psi})(f-g)=0.
1244: \label{G4.14}
1245: \end{equation}
1246: 
1247: The property \eqref{pseudolocality} implies that the restriction of
1248: $D+V$
1249: to the collar $U$ equals $D$ and hence takes the  product form
1250: \eqref{G4.10}. This and self--adjointness is all that is required to
1251: prove
1252: Green's formula (see \cite{booss-w})
1253: \begin{equation}
1254: \langle (D+V) f,g\rangle_{L^2(X,E)}-
1255: \langle f, (D+V) g\rangle_{L^2(X,E)}=\langle f|_\Sigma,
1256: \gamma(g|_\Sigma)\rangle_{L^2(\Sigma,E)}.
1257: \end{equation}
1258: 
1259: An important tool for studying elliptic boundary problems is the {\em
1260: invertible double construction} and the associated Poisson operator.
1261: Since
1262: it will be important in the sequel we recall the main facts (cf.
1263: \cite{booss-w}).
1264: 
1265: Let $\tilde X:=X_+\cup_{\Sigma} X_-$ where $X_\pm=X$. Let
1266: $\tilde E\to \tilde X$ be the vector bundle obtained by gluing
1267: $E$ with gluing function $\gamma$ over $\Sigma$. More precisely,
1268: consider
1269: the bundle $E\times\{\pm\} \to X\times\{\pm\}$. Then
1270: $$\tilde{X} =X\times\{\pm\}/\sim,  \text{ where } (x,+)\sim (y,-) \text{
1271: if and only if }
1272:     x=y\text{ and }
1273:     x\in\Sigma.$$  We write $X_\pm:=X\times\{\pm\}$.  Also,
1274: $$\tilde{E}=E\times\{\pm\}/\sim\text{ where } (x,v,+)\sim (y,w,-)
1275: \text{ if
1276: and only if }
1277:     x=y\in\Sigma \text{ and }  \gamma(v)=w.
1278:      $$  Denote by $m:\tilde X\to\tilde X$ the reflection interchanging
1279: $X_+$
1280: and $X_-$, i.e.  $m(x,\pm)=(x,\mp)$. Then $m$ is covered by a
1281: bundle
1282: map
1283: $\tilde{m}:\tilde{E}\to \tilde{E}$ defined by
1284: $\tilde{m}:(x,v,\pm)=(x,\mp
1285: v,\mp)$.  Notice that the restriction of $\tilde{m}$ to $
1286: \tilde{E}|\Sigma$
1287: equals $\gamma$ since if $x\in \Sigma,$
1288: $\tilde{m}:(x,v,\pm)=(x,\mp v, \mp)\sim (x, \gamma(v),\pm).$ The map
1289: $\tilde{m}$
1290: preserves the Hermitian metric on $E$  and
1291: $\tilde{m}^2=-\Id$.
1292: 
1293: The maps $m$ and  $\tilde{m}$  induce a unitary map
1294: \begin{equation}
1295: \Phi:L^2(\tilde X,\tilde E)\to L^2(\tilde X,\tilde E),\quad
1296: \Phi f(x):=\tilde{m} ( f(m(x)))
1297: \end{equation}
1298: satisfying
1299: $\Phi^2=-\Id$.  Thus $\Phi$ is defined by the diagram
1300: \[\begin{diagram}
1301: \node{\tilde{E}}\node{\tilde{E}}\arrow{w,t}{\tilde{m}}\\
1302: \node{\tilde{X}}\arrow{n,r}{\Phi
1303: f}\arrow{e,t}{m}\node{\tilde{X}}\arrow{n,r}{f}
1304: \end{diagram}\]
1305: 
1306: Then $D$ extends naturally to a Dirac operator $\tilde D$
1307: on
1308: $\tilde E$ satisfying
1309: \begin{equation}
1310:        \Phi^* \tilde D\Phi=-\tilde D.\label{G4.15}
1311: \end{equation} Similarly, in view of \eqref{pseudolocality}, $V$ extends
1312: to a bounded operator $\tilde V$ satisfying $\Phi^*\tilde V\Phi=-\tilde
1313: V$
1314: which has the same locality properties as $V$ with
1315: $\tilde S=S\cup m(S)$ instead of $S$.
1316: 
1317: Alternatively, we may view $\tilde D$ as a self--adjoint realization of
1318: a
1319: well--posed boundary value problem of $D\oplus (-D)$ acting on the
1320: bundle
1321: $E\oplus E$ over $X$. Namely, putting
1322: \begin{equation}
1323:           \begin{split}
1324:               \Psi:&L^2(\tilde X,\tilde E)\longrightarrow L^2(X,E\oplus
1325: E)\\
1326:                    &\Psi f:= f|X_+\oplus (\Phi f)|X_+
1327:           \end{split}\label{G4.16}
1328: \end{equation} we find
1329: \begin{equation}
1330:         \Psi\tilde D\Psi^{-1}=D\oplus (-D)
1331: \end{equation} and
1332: \begin{equation}
1333:       \Psi(\operatorname{dom}\tilde D)=\Psi(L^2_1(\tilde E))=
1334:         \bigl\{f\in L^2_1(X,E\oplus E)\,\big|\, (f|\Sigma)_2=\gamma
1335: (f|\Sigma)_1\bigr\}.\label{G4.17}
1336: \end{equation} It is indeed not difficult to show that
1337: $(f|\Sigma)_2=\gamma (f|\Sigma)_1$ is a well--posed boundary condition
1338: for
1339: $D\oplus (-D)$. The boundary operator of this boundary condition,
1340: an orthogonal projection in the pseudodifferential
1341: Grassmannian (see \cite{booss-w}),
1342: is given by
1343: \begin{equation}
1344:        \frac 12\begin{pmatrix} \Id & \gamma\\ -\gamma &
1345: \Id\end{pmatrix}.\label{G4.18}
1346: \end{equation}
1347: 
1348: We introduce the usual notations for restriction, extension, and trace
1349: operators:
1350: 
1351: \begin{defn}  Let $r^\pm$ be restriction of sections on $\tilde X$ to
1352: $X_\pm$ and
1353: $e^\pm$ extension by $0$ from $X_\pm$. $\varrho^\pm,\varrho$ denote the
1354: trace maps
1355: \begin{equation}\begin{split}
1356:        &L^2_s(X_\pm,E)\longrightarrow L^2_{s-1/2}(\Sigma,E),\\
1357:        &L^2_s(\tilde X,\tilde E)\longrightarrow L^2_{s-1/2}(\Sigma,E),
1358:                    \end{split}\quad s>1/2.\label{G4.19}
1359: \end{equation} Similarly, for $0<|t|<\epsilon$ we denote by $\varrho_t$
1360: the trace map which restricts sections to $\{t\}\times\Sigma$. For
1361: $\varrho_t$
1362: \eqref{G4.19} holds accordingly.
1363: 
1364: Note that, by the locality property \eqref{pseudolocality} of $V$, we
1365: have
1366: $r^+\tilde V=Vr^+.$
1367: \end{defn}
1368: 
1369: For convenience we abbreviate $T:=D+V$ resp. $\tilde T:=\tilde D+\tilde
1370: V$. The pseudolocality property of $\tilde V$ implies  that $\tilde T$
1371: has
1372: well--defined restrictions to $X_{\pm}$.  The following fact is
1373: well--known:
1374: 
1375: \begin{lemma}\label{ML-S4.3} For arbitrary $s\in [0,\infty)$ the trace
1376: map
1377: $\varrho^+$ extends to a bounded linear map
1378: \[   \ker T\cap L^2_s(X,E)\to L^2_{s-1/2}(\Sigma,E).\] More precisely,
1379: if
1380: $f\in \ker T\cap L^2_s(X,E)$ then
1381: $\varrho^+(f)=\lim\limits_{t\to 0+}\varrho_t(f)$ exists in
1382: $L^2_{s-1/2}(\Sigma,E)$ and the so defined $\varrho^+$ is bounded.
1383: \end{lemma}
1384: With some care the lemma could even be stated for all $s\in\mathbb{R}$.
1385: However,
1386: Sobolev spaces of negative order on manifolds with boundary are a
1387: nuisance.
1388: Since we will not need them we content ourselves to the case $s\ge 0$.
1389: \begin{proof} For $V=0$ the Lemma is well--known
1390: (\cite[Thm. 13.1]{booss-w}, cf. also \cite[Chap. I]{lions-magenes}).
1391: If $Tf=0$ then $Df=0$ in a
1392: collar of $\Sigma$. Since the result is local in a collar of
1393: $\Sigma$, we reach the conclusion.\end{proof}
1394: 
1395: In light of Lemma \ref{ML-S4.3} we can safely introduce  the following
1396: notation.
1397: 
1398: \begin{defn} Set
1399: \begin{align*}
1400:           Z_\pm^s&:=\bigl\{ f\in L^2_s(X_\pm,E)\,\big|\, \tilde
1401: T|_{X_\pm}
1402: f=0\bigr\},\quad s\ge 0,\\
1403:           \La_\pm^s&:=\varrho^\pm Z_\pm^{s+1/2},\quad s\ge -1/2,\\
1404:           Z_0&:=\bigl\{f\in L^2_1(\tilde X,\tilde E)\,\big|\, \tilde T
1405: f=0\bigr\}.
1406: \end{align*}
1407: \end{defn} Note that from elliptic regularity we immediately conclude
1408: that
1409: if
1410: $f\in L^2(\tilde X,\tilde E)$ and $\tilde Tf=0$ then
1411: $f\in L^2_1(\tilde X,\tilde E)$. Since $V$ is only assumed to be
1412: $L^2$--bounded, this cannot be improved. Also by elliptic theory,
1413: $\tilde T$ is a Fredholm operator (it is a relatively compact
1414: perturbation
1415: of $\tilde D$) and hence $\dim Z_0<\infty$.
1416: 
1417: \begin{prop}\label{ML-S4.5} If $f\in Z_0$ then $f$ vanishes in a collar
1418: of
1419: $\Sigma$ and hence $Z_0=r^+ Z_0\oplus r^-Z_0$.
1420: 
1421: In particular, $\tilde D$ is an invertible operator,   called the
1422: \emph{invertible double of $D$}.
1423: \end{prop}
1424: \begin{proof} We apply Green's formula and find
1425: \begin{equation}\begin{split}
1426:         \| \varrho f\|^2&=-\langle\varrho^+r^+f,\gamma \gamma
1427: \varrho^+r^+f\rangle= -\langle\varrho^+r^+f,\gamma \varrho^+r^+(\Phi
1428: f)\rangle\\
1429:              &= -\langle Tf,\Phi f\rangle_{L^2(X_+,E)}+\langle f,T\Phi
1430: f\rangle_{L^2(X_+,E)}=0
1431: \end{split}
1432: \end{equation} since $f\in Z_0$.
1433: 
1434: Thus we have $\varrho f=0$. Since the Dirac operator has the weak unique
1435: continuation property (cf. \cite{booss-ucp}) this implies that $f$
1436: vanishes
1437: in a collar neighborhood of $\Sigma$. If $V=0$ then the weak unique
1438: continuation property and the connectedness of $X$ imply together with
1439: the
1440: proven vanishing statement for $f\in Z_0$ that $\tilde D$ is
1441: invertible.\end{proof}
1442: 
1443: \begin{defn} We say that $D+V$ has the {\em weak unique continuation
1444: property with respect to $\partial X$} if $f=0$ is the only element of
1445: $Z_+^{1/2}$ with $\varrho^+f=0$.
1446: \end{defn}
1447: 
1448: In general we cannot expect $D+V$ to have the weak unique continuation
1449: property. However, for small $V$ it is true:
1450: 
1451: \begin{prop}\label{WUCP} The operator $D+V$ has the weak unique
1452: continuation property with respect to $\partial X$ if and only if
1453: $\tilde
1454: D+\tilde V$ is invertible. This is in particular the case if  $\|\tilde
1455: D^{-1}\tilde V\|<1$.
1456: \end{prop}
1457: \begin{remark} Booss--Bavnbek, Marcolli, and Wang \cite{BMW} prove
1458: the weak unique continuation property for a class of perturbed Dirac
1459: operators
1460: (perturbations may be nonlinear)
1461: arising in Seiberg--Witten theory. However, our perturbations
1462: are in general not admissible in their sense (cf. \cite[Def. 2.5]{BMW}).
1463: An operator $\mathfrak{P}$ is admissible in the sense of \cite{BMW} if
1464: there is a locally bounded function $P$ such that
1465: $|\mathfrak{P}u(x)| \le P(u,x)|u(x)|$. This implies in particular that
1466: $\mathfrak{P}u(x)=0$ if $u(x)=0$, i.e.
1467: $\mathfrak{P}$ is a local operator. If $\mathfrak{P}$ is a linear
1468: operator
1469: on smooth sections then this locality automatically implies that
1470: $\mathfrak{P}$
1471: is a differential operator of order $0$ (i.e. a bundle endomorphism). So
1472: the point in \cite{BMW} is that $\mathfrak{P}$ is allowed to be
1473: nonlinear.
1474: \end{remark}
1475: 
1476: \noindent{\it Proof of Proposition \ref{WUCP}.} In view of \eqref{G4.15}
1477: and Prop.
1478: \ref{ML-S4.5} the weak unique continuation
1479: property with respect to $\partial X$ implies $Z_0=0$ and hence the
1480: invertibility of $\tilde D+\tilde V$.
1481: 
1482: Conversely, assume that $\tilde D+\tilde V$ is invertible and consider
1483: $f\in Z_+^{1/2}$ with $\varrho^+f=0$. Then $e^+f\in Z_0=\{0\}$ and hence
1484: $f=0$.
1485: 
1486:     $\|\tilde D^{-1}\tilde V\|<1$ implies the invertibility of
1487: $\tilde D+\tilde V$ by means of the Neumann series.
1488: \qed
1489: 
1490: \subsection{Existence of Poisson operator and \Cald}\label{sec41}
1491: 
1492: Let $P_{Z_0}$ be the orthogonal projection onto $Z_0$ and denote by
1493: $\varrho^*$ the $L^2$--dual of $\varrho$. In other words, for a
1494: distributional section $f$ of $\tilde E|\Sigma$ and a test function
1495: $\varphi\in C^\infty(\tilde X,\tilde E)$ one has $(\varrho^*
1496: f,\varphi):=(f,\varrho
1497: \varphi)$.
1498: In view of Proposition \ref{ML-S4.5} one
1499: infers from this formula immediately that
1500: $P_{Z_0}\varrho^*=0$.
1501: 
1502: Furthermore, denote by $\tilde G$ be the pseudoinverse of $\tilde T$,
1503: i.e.
1504: \begin{equation}
1505:       \tilde Gf:=(\Id-P_{Z_0})(\tilde T+P_{Z_0})^{-1}f=
1506:       \begin{cases} \tilde T^{-1}f,& f\in Z_0^\perp,\\
1507:                             0,& f\in Z_0.
1508:       \end{cases}\label{G4.20}
1509: \end{equation} Note that since $\tilde T$ maps $L^2_s\to L^2_{s-1}, 0\le
1510: s\le 1$ the pseudoinverse maps $L^2_s\to L^2_{s+1}, -1\le s\le 0$.
1511: Together with the mapping properties of the trace map we infer that
1512: $\tilde G\varrho^*$ maps $L^2_s(\Sigma,E)$ continuously into
1513: $L^2_{s+1/2}(\tilde X,\tilde E)$ for each $-1/2\le s<0$. It is a
1514: fundamental fact that for $r^\pm \tilde D^{-1}\varrho^*$ this can be
1515: improved. Namely, one has:
1516: 
1517: \begin{thm}\label{Poissonop-a} $r^\pm\tilde D^{-1}\varrho^*$ maps
1518: $L^2_s(\Sigma,E)$ continuously into $L^2_{s+1/2}(X_\pm,E)$ for all
1519: $s\ge -1/2$.
1520: \end{thm}
1521: 
1522: For a proof see \cite{booss-w}. It will also follow from our discussion
1523: of
1524: the continuous dependence of the \Cald\ below. First we want to show
1525: that
1526: the previous theorem easily extends to hold for $\tilde T$ and hence a
1527: \Cald\ can be constructed for $\tilde T$.
1528: 
1529: \begin{thm}\label{Poissonop-b}
1530: $r^\pm\tilde G\varrho^*$ maps $L^2_s(\Sigma,E)$ continuously into
1531: $Z_\pm^{s+1/2}$ for $-1/2\le s\le 1/2$.
1532: 
1533: Furthermore, we have the following resolvent identity relating $\tilde
1534: G$
1535: and $\tilde D^{-1}$:
1536: \[\tilde G=\tilde D^{-1}-\tilde G \tilde V\tilde D^{-1}-P_{Z_0}\tilde
1537: D^{-1}.
1538: \]
1539: \end{thm}
1540: \begin{proof} The resolvent identity follows from
1541: \begin{equation}\begin{split}
1542:        \tilde G\tilde V\tilde D^{-1}= \tilde G(\tilde T-\tilde D)\tilde
1543: D^{-1}\\
1544:                  = (\Id-P_{Z_0})\tilde D^{-1}-\tilde G.
1545: \end{split}\end{equation} The operator $P_{Z_0}\tilde D^{-1}$ is
1546: bounded
1547: from $L^2_s(\tilde X,\tilde E)$ to  $L^2_1(\tilde X,\tilde E)$ for $s\ge
1548: -1$  (see    Prop. \ref{ML-S4.5} and the comment preceding it).
1549: Moreover,
1550: $\tilde G\tilde V\tilde D^{-1}$ maps
1551: $L^2_{-1}(\tilde X,\tilde E)$ continuously into $L^2_1(\tilde X,\tilde
1552: E)$.
1553:    From Theorem \ref{Poissonop-a} and the resolvent identity
1554: we thus conclude that
1555: $r^\pm\tilde G\varrho^*$ maps $L^2_s(\Sigma,E)$ continuously into
1556: $L^2_{s+1/2}(X_\pm,E)$.
1557: 
1558: Moreover, since
1559: $\tilde T \tilde G\varrho^*=(\Id-P_{Z_0})\varrho^*=\varrho^*$ and
1560: $\varrho^*$ is supported on $\Sigma$, we find that the operator actually
1561: maps into $Z_\pm^{s+1/2}$.
1562: \end{proof}
1563: 
1564: The restriction  $-1/2\le s\le 1/2$ in Theorem \ref{Poissonop-b}
1565: is due to the low regularity
1566: assumptions on $V$. If $V$ is a pseudodifferential operator, then
1567: Theorem \ref{Poissonop-a} holds verbatim for $\tilde T$ instead of
1568: $\tilde D$.
1569: The previous proof can easily be modified to this case.
1570: 
1571: With Theorem \ref{Poissonop-b} and Lemma  \ref{ML-S4.3}  in place we can
1572: make the following definitions.
1573: 
1574: \begin{defn}\begin{enumerate}
1575: \item Define $$K_\pm:=K(T)_\pm:=\pm r^\pm\tilde
1576: G\varrho^*\gamma:L^2_s(\Sigma,E)\to Z^{s+1/2}_{\pm}, \ -1/2\leq s\leq
1577: 1/2.$$
1578: Then $K:=K_+$ is called the {\em Poisson operator} for
1579: $T=D+V$.
1580: \item Define the  projection $$C_\pm:=C_\pm(T):=\varrho^\pm
1581: K(T)_\pm:L^2_s(\Sigma,E)\to
1582: L^2_s(\Sigma,E), \ -1/2\leq s\leq 1/2.$$   Then $C:=C_+$ is called the {\em
1583: \Cald\ }of
1584: $T$.
1585: \item Taking $s=0$, the subspace $$\La :=\La(T):=
1586: \La^0_+(T) =\operatorname{range}C_+(T)\subset
1587: L^2(\Sigma,E)
1588: $$ is called the {\em Cauchy data space} of $T$.
1589: \end{enumerate}
1590: \end{defn}
1591: 
1592: \begin{prop}\label{ML-S4.9} $C_\pm(D+V)$ are complementary orthogonal
1593: projections in $L^2(\Sigma,E)$ with
1594: $\operatorname{range}C_\pm=\La_\pm^0$.
1595: 
1596: Furthermore, $\gamma C_\pm\gamma^*=C_\mp$. In other words the subspaces
1597: $\La_\pm^0$ are Lagrangian subspaces with respect to the symplectic form
1598: $\omega(x,y)=\langle x,\gamma y\rangle_{L^2}.$
1599: \end{prop}
1600: 
1601: Recall that a Lagrangian subspace is a subspace $V\subset L^2(\Sigma,E)$
1602: satisfying $\gamma V=V^\perp$.
1603: 
1604: \begin{proof}
1605: $K_\pm$ maps $L^2(\Sigma,E)$ into $Z_\pm^{1/2}$ by Theorem
1606: \ref{Poissonop-b}. Hence in view of Lemma \ref{ML-S4.3} we find that
1607: $C_\pm$ is
1608: $L^2$--bounded with
1609: $\operatorname{range} C_\pm\subset \La_\pm^0$.
1610: 
1611: We show  $\La_+^0\perp \La_-^0$: pick $\xi\in \La_+^0,
1612: \eta\in \La_-^0$ and choose $f\in Z_+^{1/2}, g\in Z_-^{1/2}$ with
1613: $\xi=\varrho^+f, \eta=\varrho^-g$. Then Green's formula yields
1614: \begin{equation}
1615:        \begin{split}
1616: (\xi,\eta)&=(\varrho^+f,\varrho^-g)=-(\varrho^+f,\gamma\varrho^+(\Phi
1617: g))\\
1618:          &=-(T f,\Phi g)_{L^2(X_+,E)}+(f,T\Phi g)_{L^2(X_+,E)}=0.
1619:        \end{split}
1620: \end{equation} Hence $\La_+^0\perp \La_-^0$.
1621: 
1622: Next we show that $C_++C_-=\Id$: pick $\xi\in L^2(\Sigma,E)$ and $f\in
1623: C^\infty(\tilde X,\tilde E)$.   Green's formula gives
1624: \begin{equation}\begin{split}
1625:             ((C_++C_-)\xi,\gamma\varrho f)_\Sigma
1626: &= (\varrho^+r^+\tilde
1627: G\varrho^*\gamma \xi,\gamma\varrho f)_\Sigma-(\varrho^-r^-\tilde
1628: G\varrho^*\gamma\xi,\gamma\varrho f)_\Sigma\\
1629:       &=(TK_+\xi,f)_{X_+}-(\tilde G\varrho^*\gamma\xi,Tf)_{X_+}+
1630:         (TK_-\xi,f)_{X_-}-(\tilde G\varrho^*\gamma\xi,\tilde Tf)_{X_-}\\
1631:       &=-(\tilde G\varrho^*\gamma\xi,\tilde
1632: Tf)_X=-((\Id-P_{Z_0})\varrho^*\gamma
1633: \xi,f)_X\\
1634:       &=(-\gamma\xi,\varrho f)_\Sigma=(\xi,\gamma\varrho f)_\Sigma.
1635:            \end{split}
1636: \end{equation}
1637: Since $\gamma\varrho: C^\infty(\tilde X,\tilde E)\to C^\infty (\Sigma;
1638: E)$ is surjective, $(C_++C_-)\xi=\xi$.
1639: 
1640: In sum, $C_\pm$ are $L^2$--bounded with orthogonal range $\subset
1641: \La_\pm^0$
1642: and $C_++C_-=\Id$. Hence $C_\pm$ must be the orthogonal projections onto
1643: $\La_\pm^0$.
1644: 
1645: Finally, consider $\xi=\varrho^+f\in \La_+^0, f\in Z_+^{1/2}$. Then, by
1646: \eqref{G4.15} we have $\Phi f\in Z_-^{1/2}$ and thus
1647: $\gamma \xi=\varrho^-(\Phi f)\in \La_-^0$. Hence $\gamma$ interchanges
1648: $\La_+^0$ and $\La_-^0$. Since $\La_+^0\perp \La_-^0$ the spaces
1649: $\La_\pm^0$ must
1650: be
1651: Lagrangian and we are done.
1652: \end{proof}
1653: 
1654: \subsection{Continuous dependence of the \Cald}\label{sec42} Let us
1655: first
1656: describe the set--up in which we want to study the dependence of
1657: the \Cald\ on its input data:  We consider the class $\cD$ of Dirac
1658: operators on $E$ which are in cylindrical form near $\Sigma$ and the
1659: class
1660: $\mathcal{V}$ of auxiliary operators satisfying
1661: \textnormal{\eqref{pseudolocality}}. We say that a family $D_t\in\cD$
1662: varies continuously if all coefficients of $D_t$ (in any local chart)
1663: vary continuously. Moreover, we equip
1664: $\cV$ with the usual norm topology of $\cB(L^2(X,E))$.
1665: Finally, we put $(\cD\times \cV)(n)
1666: :=\bigl\{(D,V)\in\cD\times\mathcal{V}\,\big|\, \dim Z_0( D+V)=n\bigr\}$.
1667: 
1668: We start with some elementary statements on the continuous dependence of
1669: $P_{Z_0}$ and $\tilde G$:
1670: 
1671: \begin{prop}\label{contGtilde} The maps
1672: \begin{equation}
1673: \begin{split}P_{Z_0}, \tilde G&:(\cD\times\cV)(n)\to \cB(L^2(\tilde
1674: X,\tilde E),
1675: L^2_1(\tilde X,\tilde E)),\\
1676:      &P_{Z_0}(D,V):=\proj(\ker(\tilde D+\tilde V)),\\
1677:      &\tilde G(D,V):=(\Id-P_{Z_0}(D,V))(\tilde D+\tilde
1678: V+P_{Z_0}(D,V))^{-1},
1679:    \end{split}\label{ML-G3.21}
1680: \end{equation}
1681: are continuous.
1682: \end{prop}
1683: Note that this statement is slightly stronger than the continuity
1684: of $P_{Z_0}, \tilde G$ as maps into $\cB(L^2(\tilde X,\tilde E))$.
1685: 
1686: \begin{proof} Operators in $\cD$ are in cylindrical form near $\Sigma$,
1687: hence
1688: $\cD\times \cV\to \cB(L^2_1(\tilde X,\tilde E),L^2(\tilde X,\tilde E)),
1689: (D,V)\mapsto \tilde
1690: D+\tilde V$ is continuous. Thus also
1691: $\mathbb{C}\times\cD\times \cV\to \cB(L^2_1(\tilde X,\tilde
1692: E),L^2(\tilde
1693: X,\tilde E)), (\lambda,D,V)\mapsto \tilde
1694: D+\tilde V-\lambda$ is continuous. For a fixed pair $(D_0,V_0)\in
1695: (\cD\times\cV)$ there is an $\varepsilon>0$ such that $0$ is the only
1696: spectral point of $\tilde D_0+\tilde V_0$ in the closed disc
1697: $\{z\in\mathbb{C}\,|\,
1698: |z|\le \varepsilon\}$. By continuity this remains true for $(D,V)$ in
1699: a sufficiently small neighborhood of $(D_0,V_0)$ in
1700: $(\cD\times\cV)(n)$. For these $(D,V)$ we have
1701: \begin{equation}
1702:       P_{Z_0}(D,V)=\frac{1}{2\pi i}\oint_{|z|=\varepsilon}(\lambda-\tilde
1703:       D-\tilde V)^{-1}d\lambda.
1704: \end{equation}
1705: Since the inversion is continuous this proves the continuity statement
1706: on
1707: $P_{Z_0}$.
1708: 
1709: The continuity statement on $\tilde G(D,V)$ now follows immediately from
1710: the defining formula \eqref{ML-G3.21}.
1711: \end{proof}
1712: 
1713: The resolvent identity in Theorem
1714: \ref{Poissonop-b} shows that for $(D,V)\in(\cD\times \cV)(n)$
1715: the Poisson operator and hence $C_+(D+V)$
1716: depend continuously on $V$: namely $K_+(D+V)=K_+(D)-r^+(\tilde
1717: G(D,V)\tilde
1718: V\tilde D^{-1}+P_{Z_0}(D,V)\tilde D^{-1})\varrho^*\gamma$ and by the
1719: previous Proposition
1720: $$(D,V)\mapsto r^+(\tilde G(D,V)\tilde V\tilde D^{-1}
1721: +P_{Z_0}(D,V)\tilde D^{-1})\varrho^*\gamma$$
1722: is continuous
1723: $(\cD\times \cV)(n)\to \cB(L^2_{-1/2}(\Sigma,E),L^2_1(X_+,E))$.
1724: However, the dependence on
1725: $D$ itself is more subtle. In order to explore
1726: this we have to look a bit closer at the construction of the \Cald.
1727: 
1728: Fix a $c\in \mathbb{R}$ with $c\not\in\operatorname{spec} B$.
1729: We abbreviate the spectral projections
1730: $P_{>c}:=1_{(c,\infty)}(B),P_{\le c}:=1_{(-\infty,c]}(B)$ etc. Then we
1731: set,
1732: for $\xi\in L^2_s(\Sigma,E)$,
1733: \begin{equation}
1734:     Q(B)\xi(x):=e^{-xB}P_{\ge c}\xi, \quad x\ge 0.
1735: \end{equation}
1736: 
1737: We first need to study the mapping properties of $Q(B)$ with respect
1738: to Sobolev norms. The results we are going to present are fairly
1739: standard
1740: applications of the spectral theorem. However, to be as self--contained
1741: as possible we   present the details.
1742: 
1743: Let  $\epsilon >0$, $\epsilon <1$ and fix a smooth cut--off
1744: function $\varphi:\rr\to [0,1]$ with
1745: $\varphi(x)=1$ for $ |x| \le\epsilon/2$  and $\varphi(x)=0$ for $|x|\ge
1746: \epsilon$.
1747: Let $\psi:\rr\to \rr$ be any function in $
1748: C^\infty_0((-\epsilon,\epsilon)\setminus\{0\})$,  for example
1749: $\psi=\varphi'$.
1750: In the following proposition we only  use the restrictions of $\varphi$
1751: and $\psi$ to
1752: $\{x\ge 0\}$.
1753: 
1754: \begin{prop}\label{ML-S3.11}\hfill
1755: 
1756: \textnormal{1. } For all
1757: $s,s'\in\mathbb{R}$ we have $e^+\psi
1758: Q(B)\in\cB(L^2_s(\Sigma,E),L^2_{s'}(\tilde X,\tilde E))$.
1759: 
1760: \textnormal{2.}
1761: For $-1/2\le s\le 1/2$ the operator
1762: $\varphi Q(B)$ maps
1763: $L^2_s(\Sigma,E)$ continuously into $L^2_{s+1/2}(X_+,\tilde E)$.
1764: \end{prop}
1765: \begin{proof} First we need to recall the definition of Sobolev spaces
1766: on a  manifold with boundary.  Since $B$ is elliptic
1767: the Sobolev norms on $L^2_s(\Sigma,E)$ can be defined using
1768: $\Id+|B|$:
1769: \begin{equation}
1770:        \langle f,g\rangle_{L^2_s}:=\langle (\Id+|B|)^s f,(\Id+|B|)^s
1771: g\rangle_{L^2}
1772: \end{equation}
1773: for $f,g\in L^2_s(\Sigma,E)$. Furthermore, for a nonnegative integer
1774: $k\in \mathbb{Z}_+$ a Sobolev norm for $L^2_k(\mathbb{R}_+\times
1775: \Sigma,E)$
1776: is given by
1777: \begin{equation}
1778:       \|f\|_{L^2_k}^2=\sum_{j=0}^k \int_0^\infty
1779: \|\partial_x^j(\Id+|B|)^{k-j}f(x)\|_{L^2(\Sigma,E)}^2dx.\label{Sobnorm}
1780: \end{equation}
1781: The Sobolev norms on $X_+$ are obtained by  patching together  the
1782: usual interior Sobolev norms with the norms \eqref{Sobnorm}.
1783: 
1784: 1. By complex interpolation it suffices to prove the claim for $s'\in
1785: \mathbb{Z}_+$. Applying the Leibniz rule to
1786: $\partial^j(\Id+|B|)^{k-j}\psi
1787: Q(B)$ we see that in view of \eqref{Sobnorm} it is in fact sufficient
1788: to show that for any $k\in\mathbb{Z}_+$ the operator $(\Id+|B|)^k\psi
1789: Q(B)$ is bounded $L^2_s(\Sigma,E)\to L^2(X_+,E)$.
1790: 
1791: Furthermore, for $\xi\in L^2_s(\Sigma,E)$
1792: we have $(\Id+|B|)^{s}\xi\in L^2(\Sigma,E)$
1793: and 
1794: \[(\Id+|B|)^k\psi Q(B)\xi=(\Id+|B|)^{k-s}\psi
1795: Q(B)((\Id+|B|)^{s}\xi)\]
1796: so that  it suffices to prove the $L^2$--boundedness of
1797: $(\Id+|B|)^k\psi Q(B)$ for all $k$.
1798: 
1799: Next, since $\psi\in C^\infty_0(0,\epsilon)$ there is an $x_0>0$
1800: such that $\psi(x)=0$ for $x\le x_0$. For $x_0\le x\le\epsilon$ the
1801: spectral
1802: theorem gives
1803: \begin{equation}
1804:      \|(\Id+|B|)^ke^{-xB}P_{\ge c}\|_{L^2(\Sigma,E)}
1805: =\sup_{\lambda\ge c}|(1+\lambda)^ke^{-x\lambda}|\le C(k,x_0,c,\epsilon)
1806: \end{equation}
1807: and thus for $\xi\in L^2(\Sigma,E)$
1808: \begin{equation}\begin{split}
1809:        \|\psi (\Id+|B|)^kQ(B)\xi\|_{L^2(X_+,E)}^2&\le \|\psi\|_\infty^2
1810: \int_{x_0}^\epsilon
1811:               \|(\Id+|B|)^ke^{-xB}P_{\ge c}\xi\|_{L^2(\Sigma,E)}^2dx\\
1812:        &\le \epsilon \|\psi\|^2 C(k,x_0,c,\epsilon)^2 \|\xi\|^2.
1813:                 \end{split}
1814: \end{equation}
1815:   The proof of 1. now follows since $\psi$ has compact support in $(0,\infty)$
1816: and hence in this case extension by $0$ 
1817: is bounded $L^2_{s'}(X_+)\to L^2_{s'}(\tilde{X})$ for any $s'$.
1818: 
1819: \smallskip
1820: 
1821: 2. Let $(E_\lambda)_{\lambda\in\mathbb{R}}$ be the spectral resolution
1822: of
1823: $B$. Then we estimate for $\xi\in L^2(\Sigma,E)$
1824: \begin{equation}\label{ML-basicestimate}
1825:     \begin{split}
1826:         \|\varphi&(\Id+|B|)^{1/2}Q(B)\xi\|
1827: \le \int_0^\infty \| (\Id+|B|)^{1/2}\varphi(x)e^{-xB} P_{\ge
1828: c}(B)\xi\|^2dx\\
1829:      &\le \int_c^\infty \int_0^\epsilon (1+|\lambda|) e^{-2x\lambda}dx
1830: d\langle
1831:         E_\lambda \xi,\xi\rangle\\
1832:       &\le C(c,\epsilon) \|\xi\|^2.
1833:     \end{split}
1834: \end{equation}
1835: 
1836: For $\xi\in L^2_{-1/2}(\Sigma,E)$ we can now estimate using
1837: \eqref{ML-basicestimate}
1838: \begin{equation}\begin{split}
1839:       \|\varphi Q(B)\xi\|_{L^2(X_+,E)}&=\|\varphi
1840:       (\Id+|B|)^{1/2}Q(B)(\Id+|B|)^{-1/2}\xi\|_{L^2(X_+,E)}\\
1841:         &\le \|\varphi (\Id+|B|)^{1/2}Q(B)\|_{L^2(\Sigma,E)\to
1842:       L^2(X_+,E)}\|(\Id+|B|)^{-1/2}\xi\|_{L^2(\Sigma,E)}
1843:                 \end{split}
1844: \end{equation}
1845: proving the continuity of $\varphi Q(B)$ from $L^2_{-1/2}(\Sigma,E)$
1846: to $L^2(X_+,E)$.
1847: 
1848: Next consider $\xi\in L^2_{1/2}(\Sigma,E)$. Then analogously we find
1849: (taking
1850: $\psi=\varphi'$ and applying part 1)
1851: \begin{equation}\begin{split}
1852:       \|\partial_x\varphi Q(B)\xi\|_{L^2(X_+,E)}&\le \|\varphi'
1853:       Q(B)\xi\|_{L^2(X_+,E)}+ \|\varphi
1854:       B(\Id+|B|)^{-1/2}Q(B)(\Id+|B|)^{1/2}\xi\|_{L^2}\\
1855:      &\le C_1 \|\xi\|_{L^2}+C_2\|(\Id+|B|)^{1/2}\xi\|_{L^2}
1856:                 \end{split}
1857: \end{equation}
1858: and
1859: \begin{equation}\begin{split}
1860:       \|\varphi (\Id+|B|)Q(B)\xi\|_{L^2(X_+,E)}
1861:          &= \|\varphi (\Id+|B|)^{1/2}Q(B)(\Id+|B|)^{1/2}\xi\|_{L^2}\\
1862:      &\le C(c,\epsilon)\|(\Id+|B|)^{1/2}\xi\|_{L^2}.
1863:                 \end{split}
1864: \end{equation}
1865: In view of \eqref{Sobnorm} this proves the continuity of $\varphi Q(B)$
1866: from $L^2_{1/2}(\Sigma,E)$ to $L^2_1(X_+,E)$.
1867: 
1868: The complex interpolation method \cite[Sec. 4.2]{taylor} now yields the
1869: assertion for
1870: all $s\in [-1/2,1/2]$.
1871: \end{proof}
1872: 
1873: \newcommand{\Ell}{\operatorname{Ell}}
1874: \begin{prop}\label{ML-continuity} Let $\Ell^1(\Sigma,E)$ be the set
1875: of first order self--adjoint elliptic differential operators on
1876: $E|\Sigma$.
1877: Then for $\psi\in C^\infty_0((-\epsilon,\epsilon)\setminus{0})$ the map
1878: \[
1879:     \Ell^1(\Sigma,E)_c:=\{B\in\Ell^1(\Sigma,E)\,|\,
1880: c\not\in\operatorname{spec} B\}
1881:    \longrightarrow \cB(L^2(\Sigma,E),L^2(X_+,E)),\quad B\mapsto \psi Q(B)
1882: \]
1883: is continuous.
1884: \end{prop}
1885: \begin{proof} Again let $x_0>0$ be such that $\psi(x)=0$ for $0\leq x
1886: \le x_0$.
1887: By the spectral theorem the map
1888: \begin{equation}
1889:      [x_0,\epsilon]\times \Ell^1(\Sigma,E)_c\longrightarrow
1890: \cB(L^2(\Sigma,E)),
1891:     \quad (x,B)\mapsto e^{-xB} P_{\ge c}(B)
1892: \end{equation}
1893: is continuous.
1894: 
1895: Now fix $B_0\in \Ell^1(\Sigma,E)_c$ and a $\delta>0$. By compactness
1896: there is a neighborhood $W$ of $B_0$ such that for $
1897: B\in W$ and all $x_0\le x\le \epsilon$
1898: \begin{equation}
1899:      \|e^{-xB}P_{\ge c}(B)-e ^{-xB_0}P_{\ge c}(B_0)\|_{L^2(\Sigma,E)}\le
1900:      \delta/\sqrt{\epsilon}.
1901: \end{equation}
1902: Thus for $\xi\in L^2(\Sigma,E)$
1903: \begin{equation}
1904:     \|\psi Q(B)\xi-\psi Q(B_0)\xi\|^2\le C^2\int_{x_0}^\epsilon
1905:      \|e^{-xB}P_{\ge c}(B)-e ^{-xB_0}P_{\ge c}(B_0)\|_{L^2(\Sigma,E)}^2dx
1906:       \|\xi\|^2
1907:     \le C^2\delta^2\|\xi\|^2.
1908: \end{equation} 
1909: Thus $  \|\psi Q(B) -\psi Q(B_0) \| \le C\delta$,    completing the
1910: proof.
1911: \end{proof}
1912: 
1913: Let
1914: \begin{equation}\label{defapproxCald}
1915:        R(B)\xi(x):=\begin{cases}
1916:                   -e^{-xB}P_{\ge c}\gamma \xi  & \text{for }x\ge 0, \\
1917:                    e^{-xB}P_{\le c}\gamma \xi  & \text{for } x<0,
1918:      \end{cases}
1919: \end{equation}
1920: and define the {\em approximate    Poisson operator
1921: }   $K^0_\pm:=\pm r^\pm \varphi R(B)\gamma$.
1922: 
1923: Applying Proposition \ref{ML-S3.11} and \ref{ML-continuity}
1924: (and its obvious counterparts for $x\le 0$) to $R(B)$ gives:
1925: \begin{prop}\label{approxCald}\textnormal{1.} If $\psi\in
1926: C^\infty_0((-\epsilon,\epsilon)\setminus\{0\})$ then for all
1927: $s,s'\in\mathbb{R}$,
1928: $$\psi
1929: R(B)\in\cB(L^2_s(\Sigma,E),L^2_{s'}(\tilde X,\tilde E)).$$  Moreover,
1930: as an element of $\cB(L^2(\Sigma,E),L^2(\tilde X,\tilde E))$,
1931: $\psi R(B)$ depends continuously on the coefficients of $B$ as long
1932: as $c\not\in\operatorname{spec} B$.
1933: 
1934: \textnormal{2.} The map  $\varphi R(B)$ maps
1935: $L^2(\Sigma,E)$ continuously into
1936: $L^2_{1/2}(X_-\amalg X_+,\tilde E)$. 
1937: %Moreover, $\varphi R(B)\in
1938: %\cB(L^2(\Sigma,E),L^2_{1/2}(X_-\dot\cup X_+,\tilde E))$ depends
1939: %continuously on
1940: %the coefficients of $B$.
1941: \end{prop}
1942: 
1943: If $f\in C^\infty(\tilde X,\tilde E)$ then Green's formula gives
1944: \begin{equation}\begin{split}
1945:        (\varphi R\xi,\tilde T f)_{\tilde X}&=(r^+\varphi R\xi,r^+\tilde
1946: Tf)_{X^+}+(r^-\varphi R\xi,r^-\tilde
1947:        Tf)_{X^-}\\
1948:        &= (-\gamma P_{\ge c}\gamma\xi,\varrho f)_\Sigma-(\gamma P_{\le
1949:        c}\gamma\xi,\varrho f)_\Sigma+(S(B)\xi,f)_{\tilde X}\\
1950:        &=((\varrho^*+S(B))\xi,f)_{\tilde X}
1951:        \end{split}
1952: \end{equation}
1953: with $S(B)\xi=\varphi'\gamma R\xi$. In view of Prop.
1954: \ref{approxCald} we have in particular that $S(B)\in
1955: \cB(L^2(\Sigma,E),L^2(\tilde X,\tilde E))$ and that it depends
1956: continuously on the coefficients of $B$. Note furthermore, that
1957: $S(B)$ maps in fact into
1958: $L^2_{s,\textrm{comp}}(\tilde X\setminus\Sigma,\tilde E)$ for all $s\ge
1959: 0$. Hence we
1960: have
1961: shown that
1962: \begin{equation}
1963:       \tilde T\varphi R=\varrho^*+S(B).\label{ML-G4.21}
1964: \end{equation}
1965: 
1966: The definition of $K_\pm^0$ immediately gives
1967: \begin{equation}
1968:          \begin{split}
1969:            \varrho^+K_+^0\xi&:=\lim_{t\to 0+}\varrho_tK_+^0\xi=P_{\ge
1970: c}\xi,\\
1971:            \varrho^-K_-^0\xi&:=\lim_{t\to 0-}\varrho_tK_-^0\xi=P_{\le
1972: c}\xi.
1973:          \end{split}\label{ML-G4.22}
1974: \end{equation} Note that we assumed $c\not\in\operatorname{spec} B$.
1975: 
1976: \eqref{ML-G4.21} and \eqref{ML-G4.22} allow to express
1977: $K_\pm(\tilde D+\tilde V)$ and $C_\pm(\tilde D+\tilde V)$ in terms of
1978: the
1979: operator $R(B)$. Namely, we have in view of \eqref{ML-G4.21}
1980: \begin{equation}
1981:         (\Id-P_{Z_0})\varphi R=\tilde G\tilde T\varphi R=
1982:            \tilde G\varrho^*+\tilde G S(B)
1983: \end{equation} and hence
1984: \begin{equation}\begin{split}
1985:        K_+ &=r^+\tilde G\varrho^*\gamma\\
1986:            &=r^+(\Id-P_{Z_0})\varphi R\gamma-r^+\tilde GS(B)\gamma\\
1987:            &=K_+^0-r^+P_{Z_0}\varphi R\gamma-r^+\tilde GS(B)\gamma.
1988:           \end{split}
1989: \end{equation} Similarly,
1990: \begin{equation}
1991:        K_-=K_-^0+r^-P_{Z_0}\varphi R\gamma+r^-\tilde GS(B)\gamma.
1992: \end{equation} Finally, this gives     the formulas:
1993: \begin{equation}\begin{split}
1994:        C_+&=P_{\ge c}(B)-\varrho^+\tilde G S(B)\gamma,\\
1995:        C_-&=P_{\le c}(B)+\varrho^-\tilde G S(B)\gamma.
1996: \end{split}\label{ML-G4.26}
1997: \end{equation}
1998: In view of Proposition \ref{contGtilde} and Proposition \ref{approxCald}
1999: the operator
2000: $\varrho^\pm\tilde G S(B)\gamma$ maps $L^2(\Sigma,E)$ continuously into
2001: $L^2_{1/2}(\Sigma,E)$ and, as long as $\dim Z_0$ is constant, it
2002: depends continuously on $D$ and $V$.
2003: 
2004: Since $c\not\in\operatorname{spec}(B)$ the operators $P_{\ge c}(B),
2005: P_{\le
2006: c}(B)$ depend continuously on $B$. Note that the choice of $c$ is
2007: irrelevant. To show the continuity of
2008: $C_+$ at $D$ and $V$ we choose a $c$ not in $\operatorname{spec}(B)$.
2009: Then
2010: we will have $c\not\in\operatorname{spec}(B)$ in a full neighborhood of
2011: $D$ and
2012: $V$. Of course, $S(B)$ also depends on $c$.
2013: 
2014: Summing up we have proved:
2015: 
2016: \begin{thm}\label{continuityCald} We consider the class $\cD$ of Dirac
2017: operators on $E$ which are in cylindrical form near $\Sigma$ and the
2018: class
2019: $\mathcal{V}$ of auxiliary operators satisfying
2020: \textnormal{\eqref{pseudolocality}}. We say that a family $D_t\in\cD$
2021: varies continuously if all coefficients of $D_t$ (in any local chart)
2022: vary
2023: continuously. Moreover, we equip
2024: $\mathcal{V}$ with the usual norm topology of $\cB(L^2(X,E))$. Then for
2025: each
2026: $n\in\mathbb{Z}_+$ the map
2027: \[ \bigl\{(D,V)\in\cD\times\mathcal{V}\,\big|\, \dim Z_0( D+V)=n\bigr\}
2028:       \longrightarrow \cB(L^2(\Sigma,E)), \quad (D,V)\mapsto C( D+ V)
2029: \] is continuous.
2030: \end{thm}
2031: 
2032: Note also that in view of \eqref{ML-G4.26} the operator $C-P_{\ge
2033: 0}(B)$
2034: maps $L^2$ continuously into $L^2_{1/2}$, i.e. `is  an operator of order
2035: $-1/2$'. If $V$ is a pseudodifferential operator then one easily shows
2036: that $\varrho^\pm \tilde G S(B)$ is a smoothing operator and we recover,
2037: though with a different argument, the well--known fact that $C-P_{\ge
2038: 0}(B)$ is smoothing
2039: \cite[Prop. 2.2]{Scott}, \cite[Prop. 4.1]{Grubb}.
2040: 
2041: %%MARKE
2042: \subsection{Application to the perturbed Dirac operator}
2043: 
2044: We turn to our specific operator $D_{A,f}=D_A-4\pi^2 Q_{A,f}$.   The
2045: discussion in Section \ref{secloc} shows that
2046: $Q_{A,f}$ satisfies \eqref{pseudolocality}. Hence the general results of
2047: Sections \ref{sec41} and \ref{sec42} apply. In particular there exist
2048: the
2049: Poisson operator $K_+(D_{A,f})$ and the Calder{\'o}n projector
2050: $P_{A,f}:=C_+(D_{A,f})$. Next we show that in a neighborhood of the flat
2051: connections and for ``small'' $f$  the weak unique continuation property
2052: with respect to $\partial X$ holds for $D_{A,f}$.
2053: 
2054: \begin{lemma}\label{nbdU}  There is an open neighborhood
2055: $U\subset
2056: \cA$ of the set of flat connections and a constant $C$ (which depends
2057: only
2058: on $X$ as a Riemannian manifold) so that given any
2059: $A\in U$,
2060: $$\|\tilde D_{A}^{-1}\|\leq C.$$
2061:      \end{lemma}
2062:      \begin{proof} Given a flat connection
2063: $A$ and a gauge transformation $g$ over $X$ the operators $D_A$ and
2064: $D_{g\cdot A}$ are unitarily equivalent (i.e. $D_{gA}(a)=gD_A
2065: g^{-1}(a)$),
2066: and this equivalence extends to $\tilde D_A$ since $\cG$ acts on the Lie
2067: algebra coefficients and $\gamma=\pm *$ acts on the differential forms,
2068: thus $g_x\gamma_x(v)=\gamma_x(g_x(v))$ for $x\in \Sigma$. Hence the
2069: spectra of
2070: $\tilde{D}_A$ and $\tilde{D}_{gA}$ coincide.
2071: 
2072: The space $\cM$ of flat connections modulo gauge  transformations on $X$
2073: is compact, being homeomorphic to the  compact  real algebraic variety
2074: $\Hom(\pi_1X,SU(n))/\text{conjugation}$. Thus   there is a lower bound,
2075: say $k$, on the absolute value of the smallest eigenvalue of
2076: $\tilde{D}_A$ as $A$ varies in the set of flat connections on $X$.  It
2077: follows that the largest eigenvalue of $\tilde{D}_A^{-1}$ is bounded by
2078: $1/k$, and hence $\|\tilde{D}_A^{-1}\|\leq 1/k$. Set
2079: $C=2/k$, then by continuity the set $U=\{A\in\cA\ |\
2080:      \|\tilde{D}_A^{-1}\|\leq C\}$ is a neighborhood of the set of flat
2081: connections.
2082: \end{proof}
2083: 
2084: \begin{prop}\label{lem4.2} There is an $\epsilon>0$ so that given any
2085: connection $A$ in the neighborhood $U$ of Lemma \textnormal{\ref{nbdU}},
2086: and any
2087:      $f\in\cP$ satisfying $\|f\|_{C^r}<\epsilon$, then
2088: $\tilde D_A-4\pi^2\tilde Q_{A,f}$ is invertible and hence has the weak
2089: unique continuation property with respect to
2090: $\partial X$.
2091: \end{prop}
2092: \begin{proof}  The Neumann series shows that $\tilde D_A-4\pi^2\tilde
2093: Q_{A,f}$ is invertible  if $\|\tilde D_A^{-1} 4\pi^2 \tilde
2094: Q_{A,f}\|<1$.
2095: 
2096: By Lemmas \ref{nbdU} and
2097: \ref{estimate} there exists a constant $C$ independent of $A$ and $f$
2098: such that
2099: $$\|\tilde D_A^{-1}4\pi^2 \tilde Q_{A,f}\|\le 4\pi^2 \|\tilde D_A^{-1}\|
2100: \|\tilde Q_{A,f}\|\le C\|f\|_{C^2}.$$ Hence we may choose
2101: $\epsilon:=\tfrac{1}{C}$. Proposition \ref{WUCP} then sows that
2102: $D_{A,f}$
2103: has the weak unique continuation property.
2104: \end{proof}
2105: 
2106: \begin{thm}\label{thm5.3} The map
2107: $U\times B_{\cP}(0,\epsilon)\to \cB(L^2(\Sigma,E))$ sending a pair
2108: $(A,f)$ to the Calder\'on projector $C_{A,f}$ of $D_{A,f}$ is
2109: continuous,
2110: and hence the Cauchy data spaces $\Lambda_{A,f}:=
2111: \operatorname{range}(C_{A,f})$ vary continuously (in the gap metric on
2112: the
2113: Grassmannian of Lagrangian subspaces of $L^2(\Sigma,E)$) with respect to
2114: $A,f$.
2115: \end{thm}
2116: \begin{proof} By Prop. \ref{lem4.2} we have $Z_0(\tilde D_A-4\pi^2\tilde
2117: Q_{A,f})=0$ for all $(A,f)\in \cA\times B_{\cP}(0,\epsilon)$. Thus the
2118: claim follows from Theorem \ref{continuityCald}.
2119: \end{proof}
2120: 
2121: \section{An application}\label{application} Theorem \ref{thm5.3} allows
2122: us
2123: to use the technology developed in \cite{nico} and \cite{daniel-kirk}
2124: to
2125: study how the spectral flow of a path of operators
2126: $D_{A_t,f_t}$ on a closed 3-manifold behaves with respect to a
2127: decomposition along a separating surface. In particular, the  proof of
2128: Nicolaesu's adiabatic limit theorem and Lemma 3.2 of \cite{daniel-kirk},
2129: which consider the effect of stretching the collar of $X$ on the Cauchy
2130: data spaces,  do not depend on the operators being Dirac operators, but
2131: only on the fact that they have the Atiyah--Patodi--Singer form on a
2132: collar.  Thus     Theorem 5.1 of
2133: \cite{daniel-kirk}, and all its consequences (Section 6 of
2134: \cite{daniel-kirk}) are true for the operators
2135: $D_{A,f}$.
2136: 
2137: This will be used crucially in forthcoming   calculations of the
2138: $SU(3)$ Casson invariant for Brieskorn spheres
2139: \cite{BHK-pqr}.  We finish this article with a useful  application of
2140: our
2141: main result. We outline a different proof of a result of Taubes which is
2142: the main ingredient in his proof that his invariant equals Casson's
2143: $SU(2)$
2144: invariant for homology 3-spheres, and is the starting point for proofs
2145: of
2146: various approaches to problems known as the  Atiyah--Floer conjecture.
2147: It is also  one of the simplest
2148: non-trivial examples to which this machinery can be applied, since
2149: this situation avoids the delicate problem of singularities in the
2150: moduli space.
2151: 
2152: \bigskip
2153: 
2154:    First, we set up  some notation. Let $X$ be
2155: a homology 3-sphere.  Let
2156: $P\to X$ be  the trivial principal
2157: $SU(2)$ bundle  and let $\cA(X)$ denote the ($L^2_1$ completion of the)
2158: space of
2159: $SU(2)$ connections on
2160: $P$. Denote by $\cG(X)$ the ($L^2_2$ completion of the) gauge group
2161: $\Map(X,SU(2))$.  Then let $\cA^*(X)$ be the subspace of $\cA(X)$
2162: consisting of irreducible connections, that is, those connections whose
2163: stabilizer in  $\cG(X)$ is equal to  the center $\{\pm \Id \}
2164: \subset\cG(X)$. Thus $\cA^*(X)/\cG(X)$ is a smooth Hilbert manifold.
2165: 
2166:     Suppose that
2167: $X$ is given a Heegard decomposition
2168: $X=X_1\cup_\Sigma X_2$ of genus $g>2$. Choose    embeddings $\gamma_i$
2169: of
2170: solid tori so that $S=\cup_i\gamma_i(D^2\times S^1)$   is contained in
2171: the
2172: interior of
2173: $X_2$, and  choose   $f\in \cP$ so   that  the restriction of $cs+h_f$
2174: to
2175: the space $\cA^*(X)$ of irreducible connections  has only
2176: non-degenerate
2177: critical points (i.e.~the kernel of
2178: $D_{A,f}$ is zero at every irreducible perturbed flat connection $A$).
2179: In
2180: this context the fact  that such $f$ are dense in
2181:    any  neighborhood of $0\in\cP$   is proven in
2182: \cite{taubes} and
2183: \cite{floer}. We will  restrict $f$ further below.
2184: 
2185: It follows that the critical points of
2186: $cs+h_f:\cA^*(X)/\cG(X)\to \rr/\zz$ are finite and isolated.
2187:     Let $\cM_f^*(X)$ denote the critical set, i.e.~ the moduli space of
2188: irreducible perturbed flat connections on $X$:
2189: $$ \cM_f^*(X)=\{A\in\cA^*(X)\ | \ d(cs+h_f)(A)=0\}/\cG(X).$$
2190: 
2191:    To a pair $[A_0], [A_1]\in
2192: \cM^*_f(X)$, Taubes assigns the sign
2193: $$\Sign_T([A_0],[A_1])= (-1)^{\SF(D_{A_t,f})} $$ where $A_t$ is any path
2194: in $\cA^*(X)$ joining lifts of $[A_0]$ and $[A_1]$. In fact an
2195: application
2196: of the  index theorem shows that the mod $8$ reduction of
2197: $\SF(D_{A_t,f})$
2198: depends only on the gauge equivalence classes $[A_0]$ and $[A_1]$.
2199: 
2200: Restricting irreducible connections to the pieces of the  Heegard
2201: decomposition leads to a diagram (with the obvious notation)
2202: \[
2203: \begin{diagram}\dgsquash[3/4]
2204: \node[2]{\cM^*(X_1)}\arrow{se}\\
2205: \node{\cM^*_f(X)}\arrow{ne}\arrow{se}\node[2]{\cM^*(\Sigma)}\\
2206: \node[2]{\cM^*_f(X_2)}\arrow{ne}
2207: \end{diagram}
2208: \] All arrows are embeddings. The space $\cM^*(\Sigma)$ is a smooth
2209:      $6g-6$ dimensional  symplectic manifold. (We will describe the
2210: symplectic structure below.)    The subspaces
2211: $\cM^*(X_1)$ and
2212: $\cM_f^*(X_2)$ are smooth  and
2213: orientable $3g-3$ dimensional Lagrangian  submanifolds
2214: (\cite{taubes,Herald1}). They intersect transversely in the compact
2215: $0$--dimensional manifold $\cM_f^*(X)$ for   appropriate $f$,
2216: namely those perturbations $f$ so that $cs+h_f$ has only
2217: non-degenerate
2218: critical points in $\cA^*(X)$.   Notice that since
2219: $S$ lies in the interior of
2220: $X_2$,  perturbed flat connections restrict to flat connections on $X_1$
2221: and $\Sigma$.
2222: 
2223: We also assume that $\|f\|_{C^2}$ is small enough so  the following
2224: requirements  hold:
2225: \begin{enumerate}
2226: \item The space $\cM^*_f(X_2)$ is diffeomorphic to the flat (i.e.
2227: unperturbed) moduli space $
2228: \cM^*(X_2)$. This is possible since, as $f$ varies, $\cM^*_f(X_2)$ varies
2229: by
2230: a (Legendrian) cobordism (\cite{Herald1}) and so for small enough $f$
2231: the
2232: cobordism is a product. (Some care must be taken since $\cM^*_f(X_2)$
2233: is non-compact,
2234: but the results of \cite{Herald1}  can be used to reach this conclusion.
2235: Alternatively, Taubes shows that the perturbations can be chosen to fix
2236: the reducible
2237: stratum.)
2238: \item The perturbation should be small enough so that $\cM^*_f(X_2)$
2239: lies
2240: in the neighborhood $U$ specified by Lemma \ref{nbdU}. That this  is
2241: possible follows again from  \cite{Herald1}.
2242: \item The perturbation should be small enough so that  the requirement
2243: $\|f\|<\epsilon$ of  Proposition \ref{lem4.2} holds on $X_2$.
2244: \end{enumerate}
2245: 
2246: Then Theorem
2247: \ref{thm5.3} implies that the map taking $A\in \cM^*_f(X_2)$ to the
2248: Cauchy data space $\Lambda_{A,f}$ for $D_{A,f}$ is continuous.
2249: 
2250:     In this language Casson assigns the sign (after  orienting
2251:     the  moduli spaces $\cM^*(X_1),\ \cM_f^*(X_2)$ and $\cM^*(\Sigma)$)
2252: \[\Sign_C([A_0],[A_1])=\begin{cases} 1 & \text{if } 
2253:  \cM^*(X_1)\cdot_{[A_1]}\cM_f^*(X_2)=\cM^*(X_1)\cdot_{[A_0]}\cM_f^*(X_2)
2254:  \text{ and,}\\
2255:  -1  &  \text{otherwise}, 
2256: \end{cases}\]
2257: where
2258: $\cM^*(X_1)\cdot_{[A_1]}\cM_f^*(X_2)\in \{\pm 1\} $ denotes the local
2259: transverse
2260: intersection  number of $\cM^*(X_1)$ and $\cM_f^*(X_2)$ in
2261: $\cM^*(\Sigma)$
2262: at
2263: $[A_1]$.   
2264: 
2265: Taubes' proof that his invariant equals Casson's invariant (up to an
2266: overall sign which depends on a method for fixing particular
2267: orientations
2268: of all the moduli spaces)  reduces to the assertion that
2269: $\Sign_T=\Sign_C$. Notice that we are looking at relative signs, so that
2270: the  validity of $\Sign_T=\Sign_C$ is independent of the choice of
2271: orientations of the moduli  spaces.
2272: 
2273: We begin with a lemma whose proof appears  to be well  known to those
2274: who
2275: know it well  (e.g. \cite{dostoglou-salomon}), although  we were  unable
2276: to  find  an explicit statement in the literature.  The
2277:      statement is implicit in \cite{daska},  but is described in the
2278: language of stable holomorphic bundles.  We are indebted to G.
2279: Daskalopoulos for help with  the argument.
2280: 
2281: \begin{lemma}\label{lemma6.1} The  space $\cF^*(\Sigma)$ of irreducible
2282: flat
2283: $SU(2)$ connections over a closed 2-manifold $\Sigma$ of genus greater
2284: than 2 is   simply connected. The moduli space
2285: $\cM^*(\Sigma)$ of such connections is also  simply connected.
2286: 
2287: \end{lemma}
2288: 
2289: \begin{proof}   Before we start the proof, we remark that in the
2290: statement
2291: of  Lemma \ref{lemma6.1}, the space $\cF^*(\Sigma)$ can mean either the
2292: smooth flat irreducible connections, or the $L^2_r$ completion of this
2293: space for any $r$ large enough so that the
2294: $L^2_{r+1}$ completion of the gauge group consists of continuous gauge
2295: transformations,   e.g.~$r>0$.
2296: 
2297: We first set up notation.  We consider the trivial rank 2 smooth
2298: vector bundle
2299: $E=\cc^2\times\Sigma\to \Sigma$  endowed with a Hermitian
2300: metric.  We fix a holomorphic structure on
2301: $\Sigma$.  The space $\cA$ of
2302: $SU(2)$ connections on $E$  is  
2303: identified with the space  of holomorphic structures with
2304: holomorphically trivial determinant on
2305: $E$ via the standard construction (see
2306: \cite{Atiyah-Bott}) since $\Sigma$ is 1 complex--dimensional and so every
2307: structure is integrable. We denote by   $\cA^s$  the subspace of stable
2308: holomorphic structures on $E$ (see \cite{Atiyah-Bott}).  We take the
2309: $L^2_r$ completions of these spaces.
2310: 
2311: Denote by $\cF \subset \cA$ the subspace of flat connections, and
2312: by $\cF^* $ the subspace of irreducible flat connections.   We
2313:   take $\cG$ to be the ($L^2_{r+1}$
2314: completion of the) $SU(2)$ gauge group
2315: $\text{Map}(\Sigma,SU(2))$.
2316: 
2317: The slice theorem (see e.g.~\cite{donaldson-kronheimer}) shows that
2318: the
2319: quotient map $\cA\to
2320: \cA /\cG $, when restricted to the irreducible connections
2321: $\cA^*$    gives a locally trivial map, and hence a fibration
2322: $\cA^*\to \cA^*/\cG$ with fiber
2323: $\cG/\pm\Id$. Restricting to the flat irreducible connections
2324: gives the  fibration
2325: \begin{equation}\label{eq6.8}\cG /{\pm\Id}\to
2326: \cF^*\to \cM^*.
2327: \end{equation}
2328: 
2329: That $\cM^*$ is simply connected follows immediately from  \cite[Theorem
2330: 7.1]{daska} and the fact
2331:  that  $\cM^*$ is
2332: homeomorphic to
2333: the space of stable rank 2 holomorphic bundles over $\Sigma$ with
2334: trivial
2335: determinant (see \cite{narasimhan-sheshadri, Atiyah-Bott}).
2336: 
2337: The calculations in \cite{daska} show that
2338: $\pi_2(\cM^*)=\zz\oplus\zz/2$ and it is not  hard to show  that
2339: $\pi_1(\cG/{\pm\Id})=\zz\oplus\zz/2$.  If we knew that   the
2340: map induced by the connecting homomorphism  in the long exact sequence
2341: of
2342: homotopy groups for the fibration \eqref{eq6.8} were an isomorphism (or
2343: merely surjective), then the  result would follow from  this long exact
2344: sequence
2345: since
2346: $\pi_1(\cM^*)=0$.  This is true, in fact the results of \cite{daska},
2347: though stated for $U(2)$ instead of $SU(2)$, are equally valid for
2348: $SU(2)$.
2349: 
2350: Alternatively, we can show that  $ \pi_1(\cF^*)$ is zero using the
2351: result
2352: of \cite{rade}. Fix a flat irreducible connection $A\in \cF^* $. Let
2353: $\gamma:S^1\to \cF$ be a loop of flat connections. Corollary 2.7 of
2354: \cite{daskalopoulos-uhlenbeck}  shows that
2355: $\cA^s $ is simply connected, and since
2356: $\cF^*\subset \cA^s$ (see \cite{Atiyah-Bott})  
2357: $\gamma$ bounds a disc
2358: $\bar{\gamma}:D^2\to \cA^s $. The main result of \cite{rade}
2359: asserts that
2360: the Yang-Mills flow on $\cA$, restricted to
2361: $\cA^s$, defines a strong deformation retraction of $\cA^s$ to
2362: $\cF^*$. Thus $\gamma$ bounds a disc in $\cF^*$, completing the proof.
2363: 
2364: \end{proof}
2365: 
2366: The space of  irreducible $SU(2)$ representations of a free group is
2367: easily proven to be path connected. Thus the space
2368: $\cM^*(X_1)$, homeomorphic to the space of conjugacy classes of
2369: irreducible representations of $\pi_1(X_1)$,   is path connected. There
2370: is
2371: a fibration
2372: $$\cG(X_1)/\pm \Id\to\cF^*(X_1)\to \cM^*(X_1)$$ as in
2373: the
2374: proof of Lemma \ref{lemma6.1}. The gauge group
2375: $\cG(X_1)=\text{Map}(X_1,SU(2))$ is path connected since $X_1$ has the
2376: homotopy type of a 1-complex. Thus $\cF^*(X_1)$ is path connected.
2377: The same argument shows that
2378:     the space of irreducible perturbed flat connections on
2379: $X_2$ is
2380:     also path connected, since   the  perturbation was chosen small
2381: enough
2382: so that $\cM^*_f(X_2)$ is diffeomorphic to the unperturbed
2383: moduli space $\cM^*(X_2)$.
2384: 
2385: Combining the facts of the preceding paragraph, Lemma \ref{lemma6.1},
2386: and the observation that
2387: since the space of connections on a manifold is contractible, any family
2388: of connections on a submanifold of
2389: $X$ extends to a family of connections on all of $X$,  we obtain the
2390: following useful proposition.
2391: 
2392: \begin{prop} Given perturbed flat irreducible $SU(2)$ connections $A_0,
2393: A_1$ on the closed 3-manifold $X$, there exists a 2-parameter family
2394: $A_{s,t},\ s,t\in[0,1]$    of irreducible connections on $X$ satisfying:
2395: \begin{enumerate}
2396: \item $A_{0,t}=A_0$  and $A_{1,t}=A_1$ for all
2397: $t\in[0,1]$.
2398: \item $ A_{s,0} $ restricts to a path  of irreducible flat connections
2399: on
2400: $X_1$.
2401: \item $A_{s,1}$ restricts to a path of irreducible perturbed flat
2402: connections on
2403: $X_2$.
2404: \item $A_{s,t}$ restricts to an irreducible  flat connection
2405: $a_{s,t}$ on
2406: $\Sigma$ for all  $s,t$.
2407: \item The restriction of $A_{s,t}$ to a collar of $\Sigma$ is in product
2408: form:
2409: $A_{s,t}|_{\Sigma\times[-1,1]}=\tfrac{\del}{\del u} + a_{s,t}$.
2410: \qed\end{enumerate}
2411: \end{prop}
2412: 
2413: We now give an argument that $\Sign_T=\Sign_C$.  The strategy is to
2414: identify
2415: $\Sign_T$ with an infinite--dimensional Maslov index of the Cauchy data
2416: spaces, then
2417: to use  homotopy and stretching arguments to symplectically reduce to a
2418: finite--dimensional Maslov index which we identify with $\Sign_C$. We
2419: recommend that
2420: the reader who is unfamiliar with the type of argument we use look at
2421: the ``user's
2422: guide'' section of
2423: \cite{daniel-kirk}.
2424: 
2425: Let $\La^1_{s,t}$ denote the  Cauchy data space for the restriction of
2426: the
2427: operator
2428: $D_{A_{s,t},f}$ to $X_1$. Similarly define $\La^2_{s,t}$. These vary
2429: continuously by Theorem \ref{thm5.3}.  Note that in contrast to Section 3, the superscripts 
2430: $1$ and $2$ refer to the Cauchy data spaces for the two parts of the Heegard decomposition, 
2431: not to the Sobolev indices. Thus $\La^1_{s,t}$ and $\La^2_{s,t}$ are continuous families of Lagrangian subspaces of $L^2(\Om^*_\Sigma(\ad P))$. Moreover, for each $(s,t)$, $(\La^1_{s,t},\La^2_{s,t})$ is  a Fredholm pair.  Hence the Maslov index $\Mas(\La^1_\alpha,\La^2_\al)\in\zz$ is defined
2432: for any path $\al:[0,1]\to [0,1]\times[0,1]$ (for details see \cite{kirk-lesch}).
2433: 
2434:     Nicolaescu's theorem  says that the spectral flow of the family
2435: $D_{A_{s,0},f},
2436: \ s\in[0,1]$ equals the (infinite--dimensional) Maslov index
2437: $\Mas(\La^1_{s,0},\La^2_{s,0})$ (see \cite{nico,kirk-lesch}).
2438: Hence
2439: $$\Sign_T([A_0],[A_1])=(-1)^{\SF(D_{A_{s,0}})}
2440: =(-1)^{\Mas(\Lambda^1_{s,0},\Lambda^2_{s,0})}.$$
2441: We will
2442: use
2443: the homotopy properties of the Maslov index to find a simpler,
2444: finite--dimensional expression  for $\Mas(\La^1_{s,0},\La^2_{s,0})$ which
2445: can  be compared to the difference in intersection numbers.
2446: 
2447: The reduction  will be based on the following observations. Since the
2448: connection $a_{s,t}$ on $\Sigma$ is flat for all $s,t$,  the kernel of
2449: $S_{a_{s,t}}$ is identified via the Hodge theorem with  the cohomology
2450: $H^*(\Sigma;\ad P_{a_{s,t}})$ of the twisted de Rham complex
2451: \begin{equation}\label{eq6.10}  0\to \Omega^0(\Sigma,\ad
2452: P_{a_{s,t}})\xrightarrow{d_{a_{s,t}}}
2453: \Omega^1(\Sigma,\ad P_{a_{s,t}})\xrightarrow{d_{a_{s,t}}}
2454: \Omega^2(\Sigma,\ad P_{a_{s,t}})\to 0.
2455: \end{equation}
2456: 
2457: Since $a_{s,t}$ is irreducible,
2458: $H^0(\Sigma;\ad P_{a_{s,t}})=0$ (i.e.~using \eqref{eq2.3} for $\Sigma$)
2459: and so
2460: $H^2(\Sigma;\ad P_{a_{s,t}})=0$ by Poincar\'e duality.  Thus the kernel
2461: of
2462: $S_{a_{s,t}}$ is just
2463: $H^1(\Sigma;\ad P_{a_{s,t}})$. But this   is canonically identified
2464: with the tangent space of
2465: $\cM^*(\Sigma)$ at $[a_{s,t}]$, since the image of
2466: $d_{a_{s,t}}:\Omega^0(\Sigma,\ad P_{a_{s,t}})\to
2467: \Omega^1(\Sigma,\ad P_{a_{s,t}})$ is the tangent space of the
2468: $\cG(\Sigma)$
2469: orbit and   $d_{a_{s,t}}:\Omega^1(\Sigma,\ad P_{a_{s,t}})\to
2470: \Omega^2(\Sigma,\ad P_{a_{s,t}})$ is the linearization at $a_{s,t}$ of
2471: the
2472: curvature map
2473: $a\mapsto F(a)$.
2474: 
2475:     Via this identification, the cup product (or wedge product if one
2476: identifies cohomology with  harmonic forms)
2477: $$H^1(\Sigma;\ad P_{a_{s,t}})\times H^1(\Sigma;\ad P_{a_{s,t}})
2478: \xrightarrow{\cup}H^2(\Sigma;\rr)=\rr$$ defines a symplectic structure
2479: on
2480: $\cM^*(\Sigma)$. In particular, since
2481: $\cM^*(\Sigma)$ is a smooth $6g-6$ dimensional manifold,
2482: $\ker S_{a_{s,t}}=H^1(\Sigma;\ad P_{a_{s,t}})$ is a smoothly varying
2483: family of symplectic subspaces of $L^2(\Omega^*_\Sigma(\ad P))$.  For
2484: convenience we use the notation
2485: $$H^1_{s,t}(\Sigma)=H^1(\Sigma;\ad P_{a_{s,t}}).$$
2486: 
2487: Let $P^+_{s,t}$ denote the ($L^2$ closure of the) positive eigenspan of
2488: the
2489: tangential operator
2490: $S_{a_{s,t}}$, and let  $P^-_{s,t}$ denote its negative eigenspan.
2491: Thus
2492: we have a smoothly varying (in $s,t$) decomposition
2493: $$L^2(\Omega^*_\Sigma(\ad P))=P^-_{s,t}\oplus H^1_{s,t}(\Sigma)\oplus
2494: P^+_{s,t}.$$
2495: 
2496: \begin{lemma} \label{lem6.1} For each $s\in [0,1]$, let
2497: $T^1_{s,0}\subset
2498: H^1_{s,0}(\Sigma)$  be the   subspace defined  by
2499: $$T^1_{s,0}=\text{\rm image }H^1(X_1;\ad P_{a_{s,0}})\to H^1(\Sigma;\ad
2500: P_{a_{s,0}}).$$ Then the spaces $T^1_{s,0}$ are Lagrangian  and vary
2501: continuously. Via the identification  $T_{a_{s,0}}\cM^*(\Sigma)\cong
2502: H^1_{s,0}(\Sigma)$ the subspace $T^1_{s,0}$ corresponds to the tangent
2503: space at $a_{s,0}$ of the submanifold $\cM^*(X_1)$ of $\cM^*(\Sigma)$.
2504: 
2505: Moreover
2506: $$\SF(D_{A_{s,0},f})=\Mas(\La^1_{s,0},\La^2_{s,0})=
2507: \Mas(P^+_{s,0}\oplus T^1_{s,0},\La^2_{s,0}).$$
2508: \end{lemma}
2509: 
2510: \begin{proof}The statement of Lemma \ref{lem6.1}  refers only to the
2511: parameters $(s,0), s\in [0,1]$; in other words, throughout the proof the
2512: ``$t$'' parameter is zero.
2513: 
2514: The fact  that $T^1_{s,0}$ is Lagrangian is a  standard fact in
2515: geometric
2516: topology whose easy proof (based on Poincar\'e duality) we leave to the
2517: reader.  The identification of
2518: $T^1_{s,0}$ with the tangent space of $\cM^*(X_1)$ is also standard; one
2519: of many proofs is to consider the map of  de Rham complexes
2520: $\Omega^*(X_1;\ad P)\to \Omega^*(\Sigma;\ad P)$ in light of the diagram
2521: \eqref{eq2.3}.
2522:    Since $\cM^*(X_1)$ is a
2523: smooth submanifold, it follows that $T^1_{s,0}$ varies smoothly. Thus we
2524: focus on the statements about the Maslov index.
2525: 
2526: It follows from the adiabatic limit theorem of \cite{nico} and its
2527: refinement in
2528: \cite{daniel-kirk} that   replacing the collar of $X_1$ by an
2529: increasingly
2530: longer collar $\Sigma\times [-R,0]$ gives a homotopy of the path
2531: $s\mapsto
2532: \La^1_{s,0}$ to
2533: $s\mapsto P^+_{s,0}\oplus T^1_{s,0}$.  This is a consequence of the
2534: fact that
2535: there are no
2536: $L^2$ solutions to $D_{A_{s,0},f}\phi=0$ on  $(\Sigma\times
2537: (-\infty,0])\cup
2538: X_1$, which follows from \cite{APS} and two facts:
2539: \begin{enumerate}
2540: \item  The  restriction map
2541: $H^1(X_1;\ad P_{a_{s,0}})\to H^1(\Sigma;\ad P_{a_{s,0}}) $  is
2542: injective.
2543: \item The space $T^1_{s,0}$  equals  the {\em limiting values of
2544: extended
2545: $L^2$ solutions} in the sense of \cite{APS}. This was first observed by
2546: Yoshida \cite{yoshida} and in the present notation a proof can be found
2547: in
2548: \cite{daniel-kirk} and \cite{kirk-lesch}.
2549: \end{enumerate}
2550: 
2551: Let $r\mapsto \La^1_{s,0}(r)$ denote this homotopy. To be explicit,
2552: \[
2553: \La^1_{s,0}(r)=\begin{cases}\text{the Cauchy data space for   }
2554: D_{A_{s,0},f}
2555: \text{ on }(\Sigma\times [-\tfrac{1}{1-r},0])\cup X_1&\text{if } r<1,\\
2556: P^+_{s,0}\oplus T^1_{s,0}&\text{if } r=1.
2557: \end{cases}
2558: \]
2559: 
2560: This is not a homotopy rel endpoints, and so to see why the Maslov
2561: indices
2562: nevertheless agree, it suffices to show that  the dimensions of the
2563: intersections  $\La^1_{0,0}(r)\cap
2564: \La^2_{0,0}$ and
2565: $\La^1_{1,0}(r)\cap
2566: \La^2_{1,0}$ are independent of $r$, in fact these intersections are
2567: zero
2568: for
2569: all $r$.
2570: 
2571: For $r<1$ these intersections are isomorphic to the kernel of
2572:     the extension of $D_{A_{i,0},f}$ $(i=0,1)$ to $X(r)=(\Sigma\times
2573: [-\tfrac{1}{1-r},0])\cup X_1\cup X_2$, i.e. $X$ with a long collar
2574: inserted near $\Sigma$.  But the dimension of the kernel of
2575: $D_{A_{i,0},f}$ is independent of the choice of  Riemannian metric on
2576: $X$
2577: since $A_{i,0}$ is perturbed flat: it is isomorphic to the cohomology of the
2578: complex
2579: \eqref{eq2.6}, which is zero by the assumption that each perturbed flat
2580: connection is non-degenerate.
2581: 
2582: For $r=1$, we need to see that $(P^+_{0,0}\oplus T^1_{0,0})\cap
2583: \La^2_{0,0}$ and
2584: $(P^+_{1,0}\oplus T^1_{1,0})\cap \La^2_{1,0}$ are both zero. The
2585: quotient
2586: $$\frac{(P^+_{0,0}\oplus \ker S_{a_{0,0}}) \cap \La^2_{0,0}}{{
2587: {P^+_{0,0} \cap
2588: \La^2_{0,0}}}}\subset \ker S_{a_{0,0}} $$ is just $T^2_{0,0}$
2589: (\cite{APS,kirk-lesch})
2590: and the fact that each perturbed flat connection is non-degenerate
2591: implies
2592: that $T^1_{0,0}\cap T^2_{0,0}=0$. Moreover,  $ P^+_{0,0} \cap \La^2_{0,0}=0$ since 
2593: any section in this intersection is the restriction to the boundary of a solution to $D_{A_{0,0},f}\phi=0$ which extends to an $L^2$ solution on 
2594: $(\Sigma\times
2595: (-\infty,0])\cup
2596: X_1$.   The only such $\phi$ is zero, as we argued above. 
2597: Hence
2598: $(P^+_{0,0}\oplus T^1_{0,0})\cap \La^2_{0,0}=0$. The other endpoint is
2599: handled
2600: identically.
2601: 
2602: Since  the Maslov index is unchanged when the paths of Lagrangians are
2603: homotoped by a
2604: homotopy that preserves the dimension of the intersections at the
2605: endpoints
2606: (\cite{kirk-lesch}),
2607: $$\Mas( \La^1_{s,0},\La^2_{s,0})=\Mas(P^+_{s,0}\oplus
2608: T^1_{s,0},\La^2_{s,0}),$$ as
2609: desired.
2610: \end{proof}
2611: 
2612: The spaces $H^1_{s,t}=\ker S_{s,t}$ form a symplectic vector bundle over
2613: the square
2614: $[0,1]\times[0,1]$. In fact, this is just the pullback of the tangent
2615: bundle  $T_*\cM^*(\Sigma)$ via the map $[0,1]\times [0,1]\to
2616: \cM^*(\Sigma)$
2617: given by $(s,t)\mapsto [a_{s,t}]$.
2618: 
2619:     We have a Lagrangian subbundle
2620: $T^1_{s,0}$ defined over the top edge of this square, and also the 2
2621: sides
2622: since
2623: $A_{0,t}$ is independent of $t$ and $A_{1,t}$ is independent of $t$.
2624: Extend
2625: this arbitrarily (e.g.~ pull back via a retraction  of the square to
2626: its
2627: three sides) to a Lagrangian subbundle
2628: $T^1_{s,t}$ over the entire square. Then the 2-parameter family
2629: $P^+_{s,t}\oplus T^1_{s,t}$  provides a homotopy which shows that
2630: \begin{equation}\label{eq6.1}
2631: \Mas(P^+_{s,0}\oplus T^1_{s,0},\La^2_{s,0})=\Mas(P^+_{s,1}\oplus
2632: T^1_{s,1},\La^2_{s,1}).
2633: \end{equation} Together with Lemma \ref{lem6.1}, \eqref{eq6.1} implies
2634: that
2635: \begin{equation}\label{eq6.2} SF(D_{A_{s,0},f})= \Mas(P^+_{s,1}\oplus
2636: T^1_{s,1},\La^2_{s,1}).
2637: \end{equation} An argument very similar to the proof of Lemma
2638: \ref{lem6.1}, this time taking the adiabatic limit on $X_2$,  shows that
2639: \begin{equation}\label{eq6.3}
2640:      \Mas(P^+_{s,1}\oplus T^1_{s,1},\La^2_{s,1})=\Mas(P^+_{s,1}\oplus
2641: T^1_{s,1},P^-_{s,1}\oplus T^2_{s,1}),
2642: \end{equation} using the fact that the path of connections $A_{s,1}$ is
2643: perturbed flat on $X_2$.  Moreover, $T^2_{s,1}\subset H^1_{s,1}(\Si)$ is
2644: isomorphic to image of the differential of the inclusion
2645: $\cM_f^*(X_2)\to \cM^*(\Sigma)$ at $A_{s,1}$, i.e.~
2646: $T^2_{s,1}=\text{image}\big( T_{A_{s,1}}(\cM_f^*(X_2))\to
2647: T_{a_{s,1}}(\cM^*(\Sigma))\big)$.
2648:        Combining \eqref{eq6.2} and \eqref{eq6.3} and the facts that  the
2649: Maslov index is additive with respect to direct sum of symplectic spaces
2650: and that $P^+_{s,1}$ and $P^-_{s,1}$ are always orthogonal, one arrives
2651: at
2652: the conclusion
2653: \begin{equation}
2654: \label{eq6.5}
2655: \SF(D_{A_{s,0},f})=\Mas( T^1_{s,1},  T^2_{s,1}).
2656: \end{equation}
2657: 
2658: The spectral flow $\SF(D_{A_{s,0},f})$ depends only on $f$ and the
2659: endpoints
2660: $A_0$ and
2661: $A_1$ of $A_t$; see  Lemma
2662: \ref{independence}. Moreover its reduction modulo 2 (in fact modulo 8)
2663: only
2664: depends on the gauge equivalence classes $[A_0]$ and $[A_1]$.
2665: Equation \eqref{eq6.5} has identified $\SF(D_{A_{s,0},f})$
2666:   with a finite--dimensional Maslov index  corresponding to
2667: the tangent spaces of Lagrangian submanifolds of  a symplectic
2668: manifold. The
2669: following   lemma  identifies the mod 2 reduction of this Maslov index
2670: with
2671: the difference in intersection numbers of  $\cM^*(X_1)$ and
2672: $\cM^*(X_2)$ at
2673: $[A_0]$ and $[A_1]$.
2674: 
2675:     \begin{lemma}\label{lem6.2} Let $A$ and $B$ be  smooth  oriented
2676: Lagrangian submanifolds of  a (finite--dimensional) symplectic manifold
2677: $W$.   Let $p$  and $q$ be two transverse intersection points of
2678: $A$ and $B$. Let $\gamma_A$ be a path in $A$ from $p$ to $q$ and
2679: $\gamma_B$ a path in
2680: $B$ from $p$ to  $q$.  Suppose that  $F(s,t)$ is a homotopy rel
2681: endpoints
2682: in $W$ from
2683: $\gamma_A$ to  $\gamma_B$.  Let $T^A_{s,0}$ be the Lagrangian subbundle
2684: of
2685: the restriction of
2686: $T_*W$ to $\gamma_A$ consisting of tangent vectors to $A$, and similarly
2687: let
2688: $T^B_{s,1}$ be the Lagrangian subbundle of the restriction of
2689: $T_*W$ to $\gamma_B$ consisting of tangent vectors to $B$.
2690: 
2691: Then $T^A$ extends over the homotopy, and if $T^A_{s,t}$ is any
2692: extension
2693: (with
2694: $T^A_{0,s}$ and $T^A_{1,t}$ independent of $t$),  then
2695: $\Mas(T^A_{s,1},T^B_{s,1})$  is even if and  only if  intersection
2696: number
2697: of $A$  and $B$ at $p$ and $q$ coincide.
2698: \end{lemma}\qed
2699: 
2700: The straightforward proof can easily be constructed by musing on the
2701: following picture, and is left to the reader.
2702: \medskip
2703: \begin{center}
2704: \psfrag{TAs,t}{$T^A_{s,t}$}
2705: \psfrag{(s,t)}{$(s,t)$}
2706: \psfrag{A}{$A$}
2707: \psfrag{B}{$B$}
2708: \psfrag{p}{$p$}
2709: \psfrag{q}{$q$}
2710: \includegraphics{proof44.eps}
2711: \end{center}
2712: \medskip
2713: Lemma \ref{lem6.2} and Equation \eqref{eq6.5}  then imply Taubes'
2714: theorem.
2715: \begin{thm}[Taubes]$\Sign_C=\Sign_T$, and hence Casson's   and
2716: Taubes' invariants are equal up to an overall sign (which may depend
2717: on the
2718: homology 3-sphere).\end{thm}\qed
2719: 
2720:    Taubes proves more, namely he gives a construction  for specifying an
2721: overall sign
2722: so that  in fact  his invariant equals Casson's invariant for all
2723: homology 3-spheres.
2724: 
2725: \vfill\eject
2726: \begin{thebibliography}{99}
2727: 
2728: \bibitem{Atiyah-Bott} \textsc{Atiyah, M. F.} and\textsc{ Bott, R.} {\em
2729: The Yang-Mills equations over Riemann surfaces.}  Philos. Trans. Roy.
2730: Soc.
2731: London Ser. A  308  (1983),  no. 1505, 523--615.
2732: 
2733: \bibitem{APS}  \textsc{Atiyah, M. F.},  \textsc{Patodi, V. K.} and
2734: \textsc{
2735: Singer, I. M.} {\em Spectral asymmetry and Riemannian geometry. I.}
2736: Math.
2737: Proc. Cambridge Philos. Soc.  77  (1975), 43--69.
2738: 
2739: \bibitem{herald-boden}\textsc{Boden, Hans U.} and \textsc{Herald,
2740: Christopher M.} {\em The
2741: ${\rm SU}(3)$ Casson invariant for integral homology $3$-spheres.} J.
2742: Differential Geom.  50  (1998),  no. 1, 147--206.
2743: 
2744: \bibitem{BHK-int}\textsc{Boden, Hans U.},
2745: \textsc{Herald, Christopher M.},  and \textsc{ Kirk, Paul} {\em An
2746: integer
2747: valued ${\rm SU}(3)$ Casson invariant.} Math. Res. Lett. 8 (2001), no.
2748: 5-6, 589--603.
2749: 
2750: \bibitem{BHKK}\textsc{Boden, Hans U.},
2751: \textsc{Herald, Christopher M.},  \textsc{ Kirk, Paul}, and
2752: \textsc{Klassen, Eric P.} {\em   Gauge theoretic invariants of Dehn
2753: surgeries on knots.}  Geom. Topol.  5  (2001), 143--226
2754: 
2755: \bibitem{BHK-pqr}\textsc{Boden, Hans U.},
2756: \textsc{Herald, Christopher M.},  and \textsc{ Kirk, Paul} {\em The
2757: SU(3)
2758: Casson Invariant for Brieskorn spheres.} In preparation.
2759: 
2760: \bibitem{booss-ucp} \textsc{Booss--Bavnbek, B.}
2761: {\em The unique continuation property for Dirac operators revisited.}
2762: Geometry and topology: Aarhus (1998), Contemp. Math. 258 (2000), 21--32.
2763: 
2764: \bibitem{booss-lesch-phillips} \textsc{Booss--Bavnbek, B.}, \textsc{
2765: Lesch, Matthias},  and \textsc{Phillips, John}
2766:    {\em Unbounded Fredholm operators and
2767: spectral flow.} Preprint, math.FA/0108014
2768: 
2769: \bibitem{BMW} \textsc{Booss--Bavnbek, B.}, \textsc{Marcolli, M.},
2770: and \textsc{Wang, B.L.}
2771: {\em Weak UCP and perturbed monopole equations.}
2772: Preprint, math.DG/0203171
2773: 
2774: \bibitem{booss-w} \textsc{Booss-Bavnbek, Bernhelm}, and
2775: \textsc{Wojciechowski, Krzysztof P.} Elliptic boundary problems for
2776: Dirac
2777: operators. Mathematics: Theory \& Applications. Birkh\"{a}user Boston,
2778: Inc., Boston, MA, 1993. xviii+307 pp.
2779: 
2780: \bibitem{CLM} \textsc{Cappell, S.}, \textsc{Lee, R.}, and
2781: \textsc{Miller,
2782: E. Y.} {\em  A perturbative SU(3) Casson invariant.} Comment. Math.
2783: Helv. 77 (2002), no. 3, 491-523
2784: 
2785: \bibitem{daniel-kirk}\textsc{Daniel, A. M.} and \textsc{ Kirk, Paul}
2786: {\em
2787: A general splitting formula for the spectral flow.} With an appendix by
2788: \textsc{K. P. Wojciechowski}.  Michigan Math. J.  46  (1999),  no. 3,
2789: 589--617.
2790: 
2791: \bibitem{daska} \textsc{Daskalopoulos, Georgios D.}   {\em The topology
2792: of
2793: the space of stable bundles on a compact Riemann surface.}  J.
2794: Differential
2795: Geom.  36  (1992),  no. 3, 699--746.
2796: 
2797: \bibitem{daskalopoulos-uhlenbeck} \textsc{Daskalopoulos, Georgios D.}
2798: and
2799: \textsc{ Uhlenbeck, Karen K.} {\em  An application of transversality to
2800: the
2801: topology of the moduli space of stable bundles.}  Topology  34  (1995),
2802: no. 1, 203--215.
2803: 
2804: \bibitem{donaldson-kronheimer} \textsc{Donaldson, S. K.} and
2805: \textsc{Kronheimer, P. B.} The geometry of four-manifolds. Oxford
2806: Mathematical Monographs. Oxford Science Publications. The Clarendon
2807: Press,
2808: Oxford University Press, New York, 1990. x+440 pp.
2809: 
2810: \bibitem{dostoglou-salomon}\textsc{Dostoglou, Stamatis} and
2811: \textsc{Salamon, Dietmar A.}  {\em  Self-dual instantons and holomorphic
2812: curves.}  Ann. of Math. (2)  139  (1994),  no. 3, 581--640.
2813: 
2814: \bibitem{floer}\textsc{Floer, Andreas} {\em An instanton-invariant for
2815: $3$-manifolds.}  Comm. Math. Phys.  118  (1988),  no. 2, 215--240.
2816: 
2817: \bibitem{Herald1}\textsc{Herald, Christopher M.} {\em Legendrian
2818: cobordism
2819: and Chern-Simons theory on $3$-manifolds with boundary. } Comm. Anal.
2820: Geom.  2  (1994),  no. 3, 337--413
2821: 
2822: \bibitem{herald2}\textsc{Herald, Christopher M.} {\em  } in preparation.
2823: 
2824:      \bibitem{FU} \textsc{Freed, Daniel S.} and \textsc{Uhlenbeck, Karen
2825: K.}
2826: Instantons and four-manifolds. Second edition. Mathematical Sciences
2827: Research Institute Publications, 1. Springer-Verlag, New York, 1991.
2828: xxii+194 pp.
2829: 
2830: \bibitem{Grubb} \textsc{Grubb, Gerd} {\em Trace expansions for
2831: pseudodifferential boundary problems for Dirac--type operators and more
2832: general systems.} Ark. Math. 37 (1999), 45--86.
2833: 
2834: \bibitem{kirk-lesch}  \textsc{Kirk, Paul} and \textsc{Lesch, Matthias}
2835: {\em
2836: The eta-invariant, Maslov index, and spectral flow for Dirac-type
2837: operators
2838: on manifolds with boundary.} Preprint,  math.DG/0012123
2839: 
2840: \bibitem{lions-magenes} \textsc{Lions, J. } and \textsc{Magenes, E.}
2841: {\em Non--homogeneous boundary problems and applications.}
2842: Springer--Verlag, New York, 1972.
2843: 
2844: \bibitem{narasimhan-sheshadri} \textsc{Narasimhan, M. S.} and \textsc{
2845: Seshadri, C. S.} {\em  Stable and unitary vector bundles on a compact
2846: Riemann surface.}  Ann. of Math. (2)  82  1965 540--567.
2847: 
2848: \bibitem{nico} \textsc{Nicolaescu, Liviu I.} {\em  The Maslov index, the
2849: spectral flow, and decompositions of manifolds.}  Duke Math. J.  80
2850: (1995),  no. 2, 485--533.
2851: 
2852: \bibitem{rade} \textsc{Rade, Johan} {\em On the Yang-Mills heat
2853: equation in two and three dimensions.} J. Reine Angew. Math. 431
2854: (1992), 123--163.
2855: 
2856: \bibitem{RSW}  \textsc{Ramadas, T. R.},  \textsc{Singer, I. M.}, and
2857: \textsc{Weitsman, J.} {\em Some comments on Chern-Simons gauge theory.}
2858: Comm. Math. Phys.  126  (1989),  no. 2, 409--420.
2859: 
2860: \bibitem{Scott} \textsc{Scott, S. G.} {\em Determinants of Dirac
2861: boundary
2862: value problems over odd--dimensional manifolds.} Commun. Math. Phys. 173
2863: (1995), 43--76
2864: 
2865: \bibitem{taubes} \textsc{Taubes, Clifford Henry} {\em Casson's invariant
2866: and gauge theory.} J. Differential Geom. 31 (1990), no. 2, 547--599.
2867: 
2868: \bibitem{taylor} \textsc{Taylor, Michael E.} {\em Partial differential
2869: equations.} Applied Mathematical Sciences Vol 115, Springer--Verlag,
2870: New York, 1996.
2871: 
2872: \bibitem{yoshida} \textsc{Yoshida, Tomoyoshi} {\em Floer homology and
2873: splittings of manifolds.}  Ann. of Math. (2)  134  (1991),  no. 2,
2874: 277--323.
2875: 
2876: \end{thebibliography}
2877: 
2878: \end{document}
2879: 
2880: