1:
2: %Version of Jan. 9, 2001
3:
4:
5:
6:
7: \documentclass[10pt]{article}
8:
9: \usepackage{amsmath}
10: \usepackage{amsthm}
11: \usepackage{amssymb}
12: \usepackage{amsfonts}
13: \usepackage{epsfig}
14:
15: \input xypic
16:
17:
18: \relpenalty=10000
19: \binoppenalty=10000
20: \tolerance=50000
21:
22:
23: \newtheorem{The}{Theorem}[section]
24: \newtheorem{Cor}[The]{Corollary}
25: \newtheorem{Pro}[The]{Proposition}
26: \newtheorem{Exa}[The]{Example}
27: \newtheorem{Rem}[The]{Remark}
28: \newtheorem{Lem}[The]{Lemma}
29: \newtheorem{Def}[The]{Definition}
30: \newtheorem{Con}[The]{Conjecture}
31:
32: \def\proof{\vspace{2ex}\noindent{\bf Proof.} }
33:
34: \def\endproof{\relax\ifmmode\expandafter\endproofmath\else
35: \unskip\nobreak\hfil\penalty50\hskip.75em\hbox{}\nobreak\hfil\bull
36: {\parfillskip=0pt \finalhyphendemerits=0 \bigbreak}\fi}
37: \def\endproofmath$${\eqno\bull$$\bigbreak}
38: \def\bull{\vbox{\hrule\hbox{\vrule\kern3pt\vbox{\kern6pt}\kern3pt\vrule}\hrule}}
39:
40: \def\cancel#1#2{\ooalign{$\hfil#1\mkern1mu/\hfil$\crcr$#1#2$}}
41:
42: \def\dirac{\mathpalette\cancel\partial}
43: \def\Dirac{\mathpalette\cancel D}
44:
45:
46: \newcommand{\C}{{\mathbb C}}
47: \newcommand{\R}{{\mathbb R}}
48: \newcommand{\Z}{{\mathbb Z}}
49: \newcommand{\Q}{{\mathbb Q}}
50: \renewcommand{\P}{{\mathbb P}}
51: \newcommand{\AAA}{{\mathbb A}}
52: \newcommand{\A}{{\cal A}}
53: \newcommand{\E}{{\cal E}}
54: \newcommand{\G}{{\cal G}}
55: \newcommand{\M}{{\cal M}}
56: \newcommand{\B}{{\cal B}}
57:
58: \newcommand{\ind}{\mathfrak{i}}
59: \renewcommand{\j}{\mathfrak{j}}
60:
61: \newcommand{\la}{\langle}
62: \newcommand{\ra}{\rangle}
63: \newcommand{\ba}{\begin{eqnarray}}
64: \newcommand{\na}{\end{eqnarray}}
65: \newcommand{\beq}{\begin{equation}}
66: \newcommand{\eeq}{\end{equation}}
67: \newcommand{\s}{\mathfrak{s}}
68: \renewcommand{\t}{\mathfrak{t}}
69: \newcommand{\spinc}{\mathrm{Spin}^c}
70:
71:
72: \title{Exact triangles in monopole homology and the Casson-Walker invariant}
73: \author{Matilde Marcolli and Bai-Ling Wang}
74: \date{}
75:
76: \begin{document}
77: \maketitle
78:
79: \tableofcontents
80:
81: \section{Introduction}
82:
83: The purpose of this paper is to give a general outline of the problem
84: of the exact triangles in Seiberg--Witten--Floer theory. We present
85: here the most general case, where the problem consists of producing a
86: surgery formula relating the monopole homology of a compact oriented
87: 3--manifold $Y$ with an embedded knot $K$, and the monopole homologies
88: of some 3--manifolds obtained by Dehn surgery on $K$.
89:
90: In the series of papers \cite{CMW} \cite{MW2} \cite{MW3} \cite{MW4},
91: we studied the problem in the case of an integral homology
92: 3-sphere $Y$, and the 3--manifolds $Y_1$
93: and $Y_0$ obtained by Dehn surgery on $K$ with framing $1$ and $0$,
94: respectively.
95:
96: The results of \cite{CMW} \cite{MW2} \cite{MW3} \cite{MW4} are, at
97: this stage, still to be considered as ``work in progress'', where some
98: of the proofs need more rigorous presentations. The main
99: result of that series of papers is that the
100: Seiberg-Witten-Floer homologies of $Y$ and of the manifolds $Y_1$
101: and $Y_0$ are related by an exact triangle
102: {\small
103: \[
104: \diagram
105: HF^{SW}_{*} (Y_1, g_1) \rto^{w^1_{*}}
106: & HF^{SW}_{*}(Y, g_0, \mu)
107: \dlto_{w^0_{*}} \\
108: \bigoplus_{k\in \Z} HF^{SW}_{(*)}(Y_0, \s_k) \uto_{w_{(*)}}
109: & \\
110: \enddiagram
111: \] }
112: In this triangle the maps $w_*^1, w_*^0$ and $w_{(*)}$ are induced
113: by the surgery cobordisms connecting $Y_1$ and $Y$,
114: $Y$ and $Y_0$, $Y_0$ and $Y_1$, respectively, and $\mu$ is the surgery
115: perturbation that simulates the effect of Dehn surgery.
116:
117: In this paper, we explain how, following the same strategy for the
118: proof of the surgery formula which we have introduced in the previous
119: work, we may be able to extend this exact triangle to the more
120: general case of any closed oriented 3-manifold with a smoothly
121: embedded knot.
122:
123: In the last section of this paper we give a topological application of
124: the kind of arguments that lead to the proof of the ``geometric
125: triangles'', namely the surgery formula for monopole homology viewed
126: at the level of generators. We show that a modified
127: version of the Seiberg-Witten invariant agrees with the Casson-Walker
128: invariant, for any closed and oriented rational homology 3-sphere.
129:
130: Let Y be a closed oriented 3-manifold,
131: with a smooth embedded knot $K$, let $\nu(K)$
132: be the tubular neighbourhood of $K$ in $Y$.
133: Choose an identification of $\nu(K)$ with
134: $D^2\times S^1$:
135: \ba
136: \nu(K) \cong D^2\times S^1,
137: \label{frame}
138: \na
139: where $K$ is mapped to the core of the
140: solid torus $D^2\times S^1$. Under the identification (\ref{frame}),
141: on the boundary $T^2$, we fix a basis $m, l$ of $H_1(T^2, \Z)$ such that
142: $l$ is the longitude (parallel to $K$ under the identification (\ref{frame}))
143: and $m$ is the right-handed meridian (intersecting $l$ once), the orientation
144: determined by $m\wedge l$ coincides with
145: the orientation induced from $Y$. The corresponding
146: longitude and meridian in the knot complement
147: $ V = Y\backslash \nu (K) $
148: are denoted by $m', l'$ respectively.
149: Similarly, let $m''$ and $l''$ be the meridian
150: and longitude in the tubular neighbourhood of the knot $\nu(K)$. The
151: meridian $m''$ bounds a disk $D^2$ in $\nu(K)$, and $l''$ generates
152: $H_1(\nu(K),\Z)$ and parallels to $K$. Let $p$ and $q$ be two relatively prime
153: integers, the Dehn surgery with coefficient $p/q\in \Q\cup \{\infty\}$
154: on $K$ is the operation of removing $\nu(K)$ and gluing in $D^2\times S^1$
155: by an orientation reversing diffeomorphism $f_{p/q}$
156: of $T^2$ that satisfies
157: \[
158: f_{p/q} (m'') = pm'-ql'.
159: \]
160: The resulting manifold is denoted by $Y_{p/q}$. Note that in general
161: $Y_{p/q}$ depends on the choice of the identification (\ref{frame}).
162:
163: Let $\s$ be a $\spinc$ structure on $Y$. We shall see that, with a
164: suitable choice of metrics and perturbation, $(Y, \s)$ has
165: non-empty monopole moduli space only if $\s|_{\nu(K)}$
166: has trivial determinant, hence we shall always assume that
167: the $\spinc$ structure $\s$ is trivial around $K$.
168: If $K$ represents a trivial homology class
169: in $H_1(Y, \Z)$, then there is only one $\spinc$ structure on
170: $Y$ which agrees with $\s$ over $Y-\nu(K)$ and $\nu(K)$.
171: Suppose that $K$ represents a torsion element of order $n$ in
172: $H_1(Y, \Z)$, which means,
173: \ba
174: \label{order:n}
175: \displaystyle{
176: \frac {|\hbox{Torsion}( H_1(Y, \Z))|}{|\hbox{Torsion}( H_1(Y-\nu(K), \Z))|}}
177: =n.
178: \na
179: In other words, $n$ is the minimal positive integer such
180: that $n[K]$ is homologous to zero in $H_1(Y, \Z)$.
181: Then there exists a $\Z_n$-family of $\spinc$ structures
182: $
183: \s\otimes L_k (k=1, \cdots, n)
184: $
185: which agree with $\s$ over $Y-\nu(K)$ and $\nu(K)$, where
186: $L_k$ is a complex line bundle whose Euler
187: class is given by $kPD([K])$. If $[K] \neq 0$ in $H_1(Y, \Q)$,
188: then there exists a $\Z$-family of $\spinc$ structures
189: $
190: \s\otimes L_k (k\in \Z)
191: $
192: which agree with $\s$ over $Y-\nu(K)$ and $\nu(K)$, where
193: $L_k$ is a complex line bundle whose Euler
194: class is given by $kPD([K])$.
195:
196: Let $Y_1$ be the $(+1)$-surgery on $K$,
197: and $Y_0$ be the 0-surgery on $K$,
198: we can consider separately the following cases.
199:
200:
201:
202: \noindent $\bullet$
203: First we assume that $Y$ is a rational homology 3-sphere with
204: a smoothly embedded knot $K$ representing
205: a torsion element of order $n$ in $H_1(Y, \Z)$ in the sense of
206: (\ref{order:n}). Then $Y_1$ is a rational homology
207: 3-sphere, and $Y_0$ is a rational homology $S^1\times S^2$.
208: Let $\s$ be a $\spinc$ structure on $Y$, which is
209: trivial on $\nu(K)$. Gluing
210: the $\spinc$ structures $s|_{Y-\nu(K)}$ and $\s|_{\nu(K)}$
211: along $T^2$ via different gauge transformations
212: on $T^2$ results in
213: a $\Z_n$-family of $\spinc$ structures on $Y$ and $Y_1$,
214: and a $\Z$-family of $\spinc$ structures on $Y_0$.
215: Without confusion, thinking $K$ as the core of
216: the attaching solid torus $\nu(K)$, we denote these structures
217: on $Y, Y_1$ and $Y_0$ all by
218: $ \s\otimes L_k (k\in Z),$
219: where $L_k$ is a complex line bundle of Euler
220: class $kPD([K])$.\, With this notation, it is understood
221: that for $Y$ and $Y_1$, and with $n[K]=0$, there is only a
222: $\Z_n$-family of $\spinc$ structures.
223: Then, with a careful choice of metrics and perturbations
224: on $Y, Y_1$ and $Y_0$ as in Part I \cite{CMW}, we will
225: obtain the following exact triangles for the Seiberg-Witten-Floer
226: homologies:
227: {\small
228: \[
229: \diagram
230: HF^{SW}_{*} (Y_1, \s\otimes L_m, g_1) \rto^{w^1_{*}}
231: & HF^{SW}_{*}(Y,\s\otimes L_m, g, \mu)
232: \dlto_{w^0_{*}} \\
233: \bigoplus_{k\in \Z} HF^{SW}_{(*)}(Y_0, \s\otimes L_{nk+p}) \uto_{w_*}
234: & \\
235: \enddiagram
236: \] }
237: which holds for any fixed choice of $m$ and $p$ in $\{ 0,\ldots, n-1
238: \}$, and for a corresponding choice of perturbation, we have
239: {\small
240: \[
241: \diagram
242: \bigoplus_{k=1}^n HF^{SW}_{*} (Y_1, \s\otimes L_k, g_1) \rto^{w^1_{*}}
243: & \bigoplus_{k=1}^n HF^{SW}_{*}(Y,\s\otimes L_k, g, \mu)
244: \dlto_{w^0_{*}} \\
245: \bigoplus_{k\in \Z} HF^{SW}_{(*)}(Y_0, \s\otimes L_k) \uto_{w_*}
246: & \\
247: \enddiagram
248: \] }
249: In both cases the homomorphisms $w^1_{*}, w^0_{*}$ and
250: $w_*$ are induced by the surgery cobordisms.
251: These exact triangles generalize the results of
252: Part I-IV \cite{CMW} \cite{MW2}\cite{MW3} \cite{MW4},
253: in the sense that the above exact triangle reduces to the
254: exact triangle for an integral homology 3-sphere
255: when $n=1$.
256:
257: \noindent $\bullet$
258: Now we assume that $(Y, \s)$ is a closed oriented
259: 3-manifold of $b_1(Y)>0$ and with a $\spinc$ structure $\s$.
260: Let $K$ be a knot representing
261: a torsion homology class in $H_1(Y, \Z)$ of order $n$
262: in the sense of (\ref{order:n}).
263: If $\s$ has non-trivial determinant in the sense
264: that $c_1(\s) \neq 0$ in $H^2(Y, \Q)$, then
265: the Seiberg-Witten-Floer homology
266: for $(Y, \s)$ is $\Z_{2\ell}$-graded, where
267: $2\ell$ is the multiplicity of $c_1(\s)$ in
268: $H^2(Y, \Z)/\hbox{Torsion}$, i.e,
269: $c_1(\s) (H_2(Y, \Z)/\hbox{Torsion}) = 2\ell$. The gluing of the
270: $\spinc$ structures corresponding to these three different
271: surgeries gives rise to a $\Z_n$-family of $\spinc$
272: structures on $Y, Y_1$ and a $\Z$-family of
273: $\spinc$ structures on $Y_0$, which
274: we denote by
275: $ \s\otimes L_k \qquad (k\in \Z),$
276: with the convention as before. Then, for any
277: $\spinc$ structure $\s\otimes L_k$ ($k=1, \cdots, n)$ on $Y$
278: and $Y_1$, the corresponding Seiberg-Witten-Floer
279: homologies are all $\Z_{2\ell}$-graded, while
280: for any $\spinc$ structure $\s\otimes L_k$ ($k\in \Z$) on $Y_0$,
281: the Seiberg-Witten-Floer
282: homology $HF^{SW}_* (Y_0, \s\otimes L_k)$ is
283: $\Z_{\ell_{[k]}}$-graded, where $\ell_{[k]}$ is the greatest common
284: factor in $2\ell$ and $2k$. Similar to Part IV \cite{MW4},
285: the $\Z_{\ell_{[k]}}$-graded homology $HF^{SW}_* (Y_0, \s\otimes L_k)$
286: can be lifted to a $\Z_{\ell}$-graded homology
287: $HF^{SW}_{(*)} (Y_0, \s\otimes L_k)$. For this case, we derive in this
288: paper the following exact triangles:
289: {\small
290: \ba
291: \diagram
292: HF^{SW}_{*} (Y_1, \s\otimes L_m, g_1) \rto^{w^1_{*}}
293: & HF^{SW}_{*}(Y, \s\otimes L_m, g, \mu)
294: \dlto_{w^0_{*}} \\
295: \bigoplus_{k\in \Z} HF^{SW}_{(*)}(Y_0, \s\otimes L_{nk+p}) \uto_{w_{(*)}}.
296: & \\
297: \enddiagram
298: \label{b_1>0:n:non-torsion:mp}
299: \na }
300: which holds for any choice of $m$ and $p$ in $\{ 0,\ldots, n-1 \}$,
301: and for a corresponding choice of perturbation, we have
302: {\small
303: \ba
304: \diagram
305: \bigoplus_{k=1}^n HF^{SW}_{*} (Y_1, \s\otimes L_k, g_1) \rto^{w^1_{*}}
306: & \bigoplus_{k=1}^n HF^{SW}_{*}(Y, \s\otimes L_k, g, \mu)
307: \dlto_{w^0_{*}} \\
308: \bigoplus_{k\in \Z} HF^{SW}_{(*)}(Y_0, \s\otimes L_k) \uto_{w_{(*)}}.
309: & \\
310: \enddiagram
311: \label{b_1>0:n:non-torsion}
312: \na }
313:
314: \noindent $\bullet$
315: If $\s$ is a torsion $\spinc$ structure on $Y$, then for any
316: $k=1, \cdots, n$,
317: $HF^{SW}_{*}(Y, \s\otimes L_k, \Z[[t]])$
318: and $HF^{SW}_{*}(Y_1, \s\otimes L_k, \Z[[t]])$ are $\Z$-graded
319: with $\Z[[t]]$-coefficient. The reduced
320: versions $HF^{SW}_{*}(Y, \s\otimes L_k)$
321: and $HF^{SW}_{*}(Y_1, \s\otimes L_k)$ are obtained by setting
322: $t=0$. The Seiberg-Witten-Floer homology
323: $HF^{SW}_*(Y_0, \s\otimes L_k)$ is $\Z_{2k}$-graded, and
324: can be lifted to a $\Z$-graded version, denoted by
325: $HF^{SW}_{(*)}(Y_0, \s\otimes L_k)$. Then the
326: exact triangles take the same form as (\ref{b_1>0:n:non-torsion:mp})
327: and (\ref{b_1>0:n:non-torsion}).
328:
329: \noindent $\bullet$
330: The remaining case is when a smoothly embedded
331: knot $K$ represents a non-trivial
332: homology class in $H_1(Y, \Q)$. Let $n$ be the
333: minimal positive integer such that there is
334: a 2-cycle in $H_2(Y, \Z)$ intersecting
335: $K$ at $n$ points.
336: Then the $0$-surgery on $Y$ along $K$ yields $Y_0$
337: satisfying $b_1(Y_0) = b_1(Y) -1$. More precisely,
338: $H_1(Y_0, \Z)$ is obtained
339: by replacing the $\Z$-component $\Z \la [K]\ra$ of $H_1(Y, \Z)$
340: by $\Z_n \la [m'']= [l']\ra$. Notice that
341: $Y$ can be thought of as the manifold obtained by
342: $0$-surgery on $m'\subset Y-\nu(K) \subset Y_0$,
343: and $Y_1$ can be thought as the result
344: of $(+1)$-surgery on
345: $m'\subset Y-\nu(K) \subset Y_0$. Since $m'$ represents
346: a torsion element of order $n$ in $H_1(Y_0, \Z)$ in the
347: sense of (\ref{order:n}), we have
348: the exact triangles for $(Y, \s, K)$ obtained from
349: the corresponding exact triangles for $(Y_0, \s, m')$
350: in the form of (\ref{b_1>0:n:non-torsion:mp})
351: and (\ref{b_1>0:n:non-torsion}). Thus,
352: it is enough to establish the exact triangle
353: for a general closed oriented 3-manifold
354: $Y$ with a smoothly embedded knot representing
355: a torsion element of order $n$ in $H_1(Y, \Z)$.
356:
357:
358:
359: We now summarize the main theorem of this paper.
360:
361: \begin{The}\label{main:theorem}
362: Let $(Y, \s)$ be a closed oriented 3-manifold with
363: a $\spinc$ structure which is trivial around a
364: smoothly embedded knot $K$. Assume that $K$ represents
365: a torsion element of order $n$ in $H_1(Y, \Z)$ in the sense
366: of (\ref{order:n}). Assume that the
367: canonical framing of $K$ is given by
368: an identification of $D^2\times S^1$ with the
369: tubular neighbourhood $\nu(K)$ such that $K$ is given
370: by $\{0\} \times S^1$. Here the parallel simple curve
371: on $T^2$ provides the longitude of $K$ and the right handed meridian
372: is given by $\partial (D^2) \times \{ pt\}$. The orientation
373: determined by $m\wedge l$ coincides with the orientation induced from $Y$.
374: Let $Y_1$ and $Y_0$ be the manifolds obtained, respectively,
375: by $(+1)$ and $0$ surgery along $K$ in $Y$.
376: With a careful choice of metrics and perturbations, we obtain the
377: following exact triangle induced
378: by the surgery cobordisms after possible grading shifts:
379: {\small
380: \[
381: \diagram
382: \bigoplus_{k=1}^n HF^{SW}_{*} (Y_1, \s\otimes L_k) \rto^{w^1_{*}}
383: & \bigoplus_{k=1}^nHF^{SW}_{*}(Y, \s\otimes L_k)
384: \dlto_{w^0_{*}} \\
385: \bigoplus_{k\in \Z} HF^{SW}_{(*)}(Y_0, \s\otimes L_k) \uto_{w_{(*)}}
386: & \\
387: \enddiagram
388: \] }
389: Here $\s\otimes L_k$ is the $\spinc$ structure obtained
390: by tensoring a $\spinc$ structure $\s$ with a complex line
391: bundle $L_k$ of Euler class $kPD([K])$. Moreover, for any
392: fixed $m, p\in \{0, \cdots, n-1\}$, we have the
393: following more refined version of the exact triangle:
394: {\small
395: \[
396: \diagram
397: HF^{SW}_{*} (Y_1, \s\otimes L_m) \rto^{w^1_{*}}
398: & HF^{SW}_{*}(Y, \s\otimes L_m)
399: \dlto_{w^0_{*}} \\
400: \bigoplus_{k\in \Z} HF^{SW}_{(*)}(Y_0, \s\otimes L_{nk+p}) \uto_{w_{(*)}}
401: & \\
402: \enddiagram
403: \] }
404: Here again the maps are induced by the surgery cobordisms, possibly
405: after a shift in the grading.
406: \end{The}
407:
408: The major technical steps required in the proof are described in
409: \cite{CMW}\cite{MW2}\cite{MW3}\cite{MW4}. Thus, in this paper,
410: we shall address only those issues that are relevant to
411: this general exact triangle, while we refer to the previous papers for the
412: general setting and results.
413:
414: In the last section of the paper, we show that a suitably modified
415: version of the Seiberg--Witten invariant of a rational homology
416: 3-sphere agrees with the Casson--Walker invariant.
417: For any rational homology 3-sphere $(Y, \s, g)$ with
418: a $\spinc$ structure $\s$ and a Riemannian metric,
419: the counting of the irreducible Seiberg-Witten
420: monopoles defines the Seiberg-Witten invariant
421: \[
422: SW_Y(\s, g) = \# \bigl(\M_Y^*(\s, g)\bigr),
423: \]
424: where each irreducible monopole in $\M_Y^*(\s, g)$ has
425: a natural orientation from the linearization of the Seiberg-Witten
426: equations. As studied in \cite{MW1}, $SW_Y(\s, g)$ depends on the metric and
427: perturbation used in the definition, in order to obtain a topological
428: invariant, we can modify $SW_Y(\s, g)$ by a metric and perturbation
429: dependent correction term as follows. Choose any four manifold $X$
430: with boundary $Y$, such that $X$ is endowed with a
431: cylindrical-end metric modeled on $(Y, g_Y)$. Choose
432: a $\spinc$ structure $\s_X$ on $X$ which agrees with
433: $\s$ on $Y$ over the end, and choose a connection $A$
434: on $(X, \s_X)$ which extends the unique reducible
435: $\theta_{\s}$ on $(Y, \s)$. Then we set
436: \ba
437: \xi_Y(\s, g) = Ind_\C (\Dirac^X_A) -\displaystyle{\frac{1}{8} }
438: \bigl( c_1(\s_X)^2 -\sigma (X)\bigr),
439: \label{correction:term}
440: \na
441: where $Ind_\C (\Dirac^X_A)$ is the complex index of the
442: Dirac operator on $(X, \s_X)$ twisted with
443: the extending $\spinc$ connection $A$ and $\sigma (X)$ is
444: the signature of $X$. By the Atiyah-Patodi-Singer index theorem,
445: $\xi_Y(\s, g)$ is independent of the choice of
446: $(X, \s_X)$ and $A$, actually, $\xi_Y(\s, g)$ can be expressed as a
447: combination of the Atiyah-Patodi-Singer eta invariants for the
448: Dirac operator and signature operator on $(Y, \s)$:
449: \[
450: \xi_Y(\s, g) = \displaystyle{-
451: \frac 14 \eta^{\dirac_{\theta_{\s}}}_Y (0) -
452: \frac 18 \eta^{sign}_Y(0).}
453: \]
454:
455: The modified version of the Seiberg-Witten invariant is defined
456: as
457: \ba
458: \hat{SW}_Y(\s) = SW_Y (\s, g) - \xi_Y(\s, g).
459: \label{hat:SW}
460: \na
461: Then we prove the following equivalence between
462: $\hat{SW}_Y$ and the Casson-Walker invariant.
463:
464: \begin{The}
465: Let $Y$ be a rational homology 3-sphere. Then,
466: \[
467: \sum_{\s\in \spinc (Y)} \hat{SW}_Y(\s) = \displaystyle{\frac{1}{2}}
468: |H_1(Y, \Z)| \lambda (Y),
469: \]
470: where $\lambda (Y)$ is the Casson-Walker invariant of $Y$ (cf. \cite{Walker}).
471: \end{The}
472:
473:
474:
475:
476:
477:
478:
479: \vskip .2in
480:
481: {\bf Acknowledgements} BLW likes to acknowledge the paper
482: of Ozsv\'ath and Szab\'o \cite{OS} on the theta divisor and the Casson-Walker
483: invariant which leads to his proof of the equivalence
484: of $\hat{SW}_Y$ and the Casson-Walker invariant, hence
485: proving the conjecture formulated in \cite{OS} on the equivalent
486: between $\hat{SW}_Y$
487: and their $\hat\theta$ invariant for all rational homology 3-sphere.
488: BLW is partially supported by Australia Research Council Fellowship.
489:
490:
491:
492: \section{The geometric triangle}
493:
494: In this section, we identify the
495: monopoles on $Y$ with monopoles on $Y_1$ and $Y_0$. Suppose given a
496: smoothly embedded knot
497: $K$ in $(Y, \s)$, which represents a torsion element of order $n$ in
498: $H_1(Y, \Z)$. We can split $Y$ along a torus as in
499: \cite{CMW},
500: \[
501: Y= V\cup_{T^2} \nu(K).
502: \]
503: We choose a metric on $Y$ with a long cylinder
504: $[-r, r] \times T^2$, and denote the resulting manifold
505: as
506: \[
507: Y(r) = V\cup_{T^2} ([-r, r] \times T^2) \cup_{T^2} \nu(K).
508: \]
509: We additionally require that the chosen metric on $Y$ satisfies the
510: condition of Lemma
511: 3.18 in \cite{CMW} in the neighbourhood $\nu(K)$ of the knot:
512: it has non-negative scalar curvature, strictly
513: positive away from the boundary. This induces a natural
514: metric on $Y_0(r)$. On $Y_1(r)$, we need to choose a metric which
515: agrees with the original metric on $Y (r)$ when restricted
516: to the knot complement $V$. The induced
517: metric from $Y_1(r)$ in the torus neighbourhood of $\nu(K)$
518: is the metric described in Lemma 3.21 \cite{CMW}.
519:
520:
521: With this choice of the metric, the moduli space of monopoles
522: on $(Y, \s)$ is non-empty only if $\s|_{\nu(K)} $ is trivial.
523: Gluing the two $\spinc$ structures $\s|_V$ and $\s|_{\nu(K)}$
524: along $T^2$ by a gauge transformation on $T^2$ gives rise to a
525: $\Z_n$-family of $\spinc$ structures on $Y$ and $Y_1$, and to
526: a $\Z$-family of $\spinc$ structures on $Y_0$. The resulting
527: $\spinc$ structures can be classified as the result of tensoring the
528: original $\spinc$ structure $\s$ with complex
529: line bundles $L_k (k\in \Z)$ whose
530: Euler class is given by $kPD([K])$. The gluing
531: theorem for 3-dimensional monopoles and the
532: perturbation $\mu$ on
533: $\nu(K)$, ``simulating the effect of surgery'',
534: provide the decomposition
535: of the moduli space for
536: \[
537: \cup_{k\in \{ 1, \cdots, n\}} \M_Y (\s\otimes L_k).
538: \]
539:
540: \begin{The}
541: \label{decomposition}
542: With the choice of perturbations and metrics on $Y, Y_1$ and
543: $Y_0$ described above, we have the following relation between the
544: critical sets of the Chern-Simons-Dirac functional
545: on the manifolds $Y, Y_1$ and $Y_0$:
546: \[\begin{array}{lll}
547: &&\bigcup_{k\in \{ 1, \cdots, n\}} \M_Y (\s\otimes L_k)\\[2mm]
548: &=& \bigcup_{k\in \{ 1, \cdots, n\}} \M_{Y_1} (\s\otimes L_k)
549: \cup \bigcup_{k\in \Z} \M_{Y_0} (\s\otimes L_k).
550: \end{array}
551: \]
552: \end{The}
553: \begin{proof}
554: First we assume that $Y$ is a rational homology 3-sphere. When
555: stretching the neck in $Y(r)$, as $r\to \infty$, we get two manifolds,
556: each endowed with an infinite cylindrical end,
557: \[
558: V(\infty) = V\cup_{T^2} ([0, \infty) \times T^2)\]\[
559: \nu(K) (\infty) = \nu(K) \cup_{T^2} ((-\infty, 0] \times T^2).
560: \]
561: The Seiberg-Witten monopole moduli space of $Y(r)$, for sufficiently large
562: $r$, can be described in terms of the
563: moduli spaces of $V(\infty)$ and $\nu(K) (\infty)$ as
564: analyzed in \cite{CMW}.
565:
566: Notice that the moduli spaces of flat connections on $T^2$, modulo
567: the subgroups of the gauge transformations on $T^2$ which
568: can be extended to the whole manifolds $V$ or $\nu(K)$, define
569: the following character varieties:
570: \[
571: \chi_0(T^2, V)
572: = H^1(T^2, \R)/2n\Z \la PD([l])\ra,
573: \]\[
574: \chi_0(T^2, \nu(K))
575: = H^1(T^2, \R)/2\Z \la PD([m])\ra.
576: \]
577: Thus the character variety $\chi_0(T^2, Y)$ is
578: $ \chi_0(T^2, Y) = H^1(T^2, \R).$
579: The covering maps between these character varieties are illustrated as
580: follows
581: {\small \[
582: \diagram
583: 2\Z \langle PD([m]) \rangle \drto & & 2n\Z \langle PD([l]) \rangle\dlto \\
584: & \chi_0(T^2,Y) \dlto^{\pi_1}\drto^{\pi_2} & \\
585: \chi_0(T^2,V)\drto & & \chi_0(T^2,\nu(K))\dlto \\
586: & \chi(T^2) & \\
587: \enddiagram
588: \]}
589: where the maps $\pi_i$ are the covering maps with fibers as indicated.
590:
591: Based on the analysis of the space $\M_V^* (\s|_V)$ of irreducible
592: monopoles on a 3-manifold with
593: a cylindrical end modelled on $T^2$, as in \cite{CMW},
594: we see that the asymptotic limit map defines a continuous
595: map:
596: \[
597: \M_V^* (\s|_V) \stackrel{\partial_{\infty}}{\to}
598: \chi_0(T^2,V).
599: \]
600: The reducibles on $V$ embed in the character variety $\chi_0(T^2,V)$.
601: Then, for a sufficiently large $r$, the gluing theorem gives a
602: diffeomorphism:
603: \[
604: \#_Y : {\cal M}^*_V\backslash \partial_\infty^{-1}(U_\theta)
605: \times_{\chi_0(T^2,Y)}\chi (\nu(K))
606: \to\bigcup_{k=1}^n {\cal M}^*_{Y(r)}(\s\otimes L_k),
607: \]
608: here $U_\Theta$ is a small neighbourhood of the ``bad points'' $\Theta$
609: in $\chi_0(T^2, Y)$ where the twisted Dirac operator
610: on $T^2$ has non-trivial kernel, and $ \chi (\nu(K))$
611: are the reducible lines in $\chi_0(T^2,Y)$. The above
612: fibred product is obtained (cf.\cite{CMW}) by taking the
613: pullbacks of the images of the boundary value maps under the
614: projections
615: {\small
616: \[
617: \diagram
618: & \chi_0(T^2,Y) \dlto^{\pi_1}\drto^{\pi_2} & \\
619: \chi_0(T^2,V) & & \chi_0(T^2,\nu(K)). \\
620: \enddiagram
621: \]}
622:
623:
624: Let $(u, v)$ be the coordinates on
625: $\chi_0(T^2,Y) \cong H^1(T^2, \R)$ corresponding
626: to the holonomy around the longitude $l$ and
627: the meridian $m$ respectively. In the gluing map $\#_Y$ above,
628: $\chi(\nu(K))$ corresponds to the lines
629: $\{v=2k, k=1, \cdots, n\}$. For each line $\{v=2k\}$,
630: the image of the gluing map gives
631: a diffeomorphism onto $\M^*_{Y(r)} (\s\otimes L_k)$.
632:
633: For each $\spinc$ structure $\s\otimes L_k$, there is a
634: unique reducible monopole
635: on $Y(r)$, which is given by the
636: intersection of $\chi(V, \s)$, the flat connections
637: on $(V, \s)$, with the line $\{ v= 2k\}$ in $\chi_0(T^2, Y)$:
638: \[
639: \theta_Y(k) = \{ u=u(\s)\} \times_{\chi_0(T^2, Y)} \{ v= 2k\}.
640: \]
641: Here $u(\s)$ is the holonomy of the flat connections
642: in $\chi(V, \s)$ around the longitude $l'$.
643:
644: Similarly, we have gluing maps for monopoles on $Y_1(r)$ and
645: $Y_0(r)$, respectively.
646: In the gluing map $ \#_{Y_1}$
647: for $ {\cal M}^*_{Y_1(r)}(\s\otimes L_k)$,
648: \[
649: \#_{Y_1} : {\cal M}^*_V \backslash
650: \partial_\infty^{-1}(U_\theta)\times_{\chi_0(T^2,Y_1)}\chi (\nu(K))
651: \to\bigcup_{k=1}^n {\cal M}^*_{Y_1(r)}(\s\otimes L_k),
652: \]
653: $\chi (\nu(K))$ is identified with $v-u = 2k+1$.
654: Similarly, we have the gluing map $ \#_{Y_0}$
655: for $ {\cal M}^*_{Y_0(r)}(\s\otimes L_k)$,
656: \[
657: \#_{Y_0} : {\cal M}^*_V \backslash
658: \partial_\infty^{-1}(U_\theta)\times_{\chi_0(T^2,Y_0)}\chi (\nu(K))
659: \to \bigcup_{k\in\Z}{\cal M}^*_{Y_0(r)}(\s\otimes L_k),
660: \]
661: where $\chi (\nu(K))$ is given by $ u = 2k$. For each $k\in \{1,
662: \cdots, n\}$, the reducible monopole for $(Y_1, \s\otimes L_k)$,
663: consists of the unique point
664: \[
665: \theta_{Y_1}(k) = \{ u=u(\s)\} \times_{\chi_0(T^2, Y)} \{ v=u+ 2k+1\}.
666: \]
667:
668: For $Y_0$ with a non-trivial $\spinc$ structure
669: $\s\otimes L_k$ ($k\in \Z, k\neq 0$), the set of reducibles is empty
670: for any generic perturbation, and for
671: $\s\otimes L_0= \s$, it consists of one circle of reducibles
672: $u=0$ in the cylinder $\chi_0(T^2,Y_0) = \chi_0(T^2,V)$, which
673: can be perturbed away by introducing a small perturbation as in
674: Theorem 6.13 \cite{CMW}.
675:
676: Now we apply the perturbation to simulate the effect of Dehn
677: surgery. This amounts to a careful choice of perturbation as
678: in Section 6 \cite{CMW}, which we now briefly describe.
679:
680: Choose a compactly supported 2-form $\mu$ representing the generator
681: of $ H^2_{cpt} (D^2 \times S^1)$, defined as in Lemma 3.18 \cite{CMW},
682: such that we have
683: \ba
684: \int_{D^2 \times \{pt \}}\mu = 1 \label{normalization}
685: \na
686: for any point on $S^1$. Under the isomorphism $H^2_{cpt} (\nu(K))
687: \cong H_1(\nu(K))$, given by Poincar\'e duality, this form corresponds
688: to the generator $[\mu]= PD_{\nu(K)}(l)$.
689: The class of $\mu$ in $H^2( D^2 \times S^1)$
690: is trivial, and we can write $\mu= d\nu$, where $\nu$ is a 1-form
691: satisfying $\int_{S^1 \times \{pt \}}\nu = 1$, i.e.
692: $\nu= PD_{T^2}(l)$.
693: Choose on $\nu(K)$ a metric as in Lemma 3.18 \cite{CMW}.
694:
695: Fix a $U(1)$-connection $A_0$ representing the trivial connection on $T^2$,
696: For any $U(1)$-connection $A$, define
697: $T_A$ to be
698: \[
699: T_A(z)= -i\int_{\{z\in D^2\} \times S^1} (A-A_0).
700: \]
701:
702: For any given $\epsilon >0$, we can choose a function $\hat f:\R\to\R$ with
703: the following properties.
704:
705: (a) $\hat f$ is continuously differentiable on $(-1,1)$ and satisfies the
706: periodicity $\hat f(t+2)=\hat f(t)$
707:
708: (b) the derivative $\hat f'$ has range $\hat f'(t)\in [-1,1]$ for all $t\in
709: [-1,1]$, and satisfies $\hat f'(1-t) = \hat f'(1+t)$ for $t\in\R$.
710:
711: (c) the following estimate holds:
712: $\sup_{t\in [-1+\epsilon, 1-\epsilon]} | \hat f'(t)-t | <
713: \epsilon.$
714:
715: Now, for the $\spinc$ structure $\s\otimes L_k$ ($k=1, \cdots, n$),
716: consider the function $f_k'(t)= \hat f'(t+1)+2k$ and define a
717: perturbation of the Seiberg-Witten equations
718: on $(Y, \s\otimes L_k)$ in the following
719: way:
720: \ba
721: \left\{\begin{array}{l}
722: F_A = *\sigma(\psi,\psi) + f'_k(T_A) \mu \\[2mm]
723: \dirac_A (\psi ) = 0
724: \end{array}
725: \right. .
726: \label{p SW}
727: \na
728:
729: With respect to the chosen metric on $\nu(K)$, with sufficiently large
730: positive scalar curvature on the support of $\mu$ as specified in
731: Lemma 3.18 \cite{CMW},
732: the only solutions
733: of the perturbed monopole equations are reducibles $(A, 0)$, that satisfy
734: \ba
735: F_A = f_k'(T_A)\mu.
736: \label{deformflat}
737: \na
738:
739: In addition to this surgery perturbation, we consider another
740: perturbation of the Seiberg--Witten equations on the
741: tubular neighbourhood $\nu(K)$ in $Y$, $Y_1$, and $Y_0$. This
742: perturbation has the effect of producing a global shift in the
743: character variety to avoid the bad points on $H^1(T^2, \R)$ when
744: we deform the unperturbed geometric triangles in
745: $\chi (T^2)$ to the perturbed geometric triangles in
746: $\chi (T^2)$.
747:
748: Let $\mu$ be a compactly supported 2-form in $D^2\times S^1$
749: satisfying (\ref{normalization}). Let $\eta >0$ be some small real
750: parameter. Consider an additional perturbation
751: \ba F_A= *\sigma(\psi,\psi) \pm\eta \mu \label{eta:pert} \na
752: of the curvature equation on $\nu(K)$ inside $Y$ and inside $Y_0$.
753:
754: This perturbation has the effect of shifting the asymptotic values by
755: an amount $\eta$. We choose the sign so that the line of reducibles on
756: $\nu(K)\subset Y$ for $\s\otimes L_k$
757: becomes $\{ (u,v) | v=2k +\eta \}$, the line of reducibles on
758: $\nu(K)\subset Y_1$ for $\s\otimes L_k$
759: remains the same $\{ (u,v) | v-u =2k+ 1 \}$, and the line
760: of reducibles on $\nu(K)\subset Y_0$ for $\s\otimes L_k$
761: becomes $\{ (u,v) | u=2k+ \eta \}$.
762:
763: On $\nu(K)$ inside $Y$ for $\s\otimes L_k$
764: we shall consider the perturbed curvature
765: equation
766: \ba F_A= *\sigma(\psi,\psi) +
767: (f'_k(T_A-\eta)+\eta)\mu. \label{eta:mu:pert} \na
768:
769: Therefore, we can partition
770: the moduli spaces for $(Y, \s\otimes L_k)$ ($k=1,\cdots, n$) into the union
771: of the moduli spaces for $(Y_1, \s\otimes L_k)$ ($k=1,\cdots, n$)
772: and $(Y, \s\otimes L_k)$ ($k\in \Z)$, as in Theorem 6.3
773: \cite{CMW}. This completes the proof of the
774: theorem for the case of rational homology 3-sphere $Y$ with
775: a knot $K$ representing a torsion element of order n in $H_1(Y, \Z)$.
776:
777: For a general 3-manifold $Y$ with a smoothly embedded
778: knot $K$ representing a torsion element of order n
779: in $H_1(Y, \Z)$, the proof is essentially the same as the case
780: of rational homology spheres discussed above, and we omit the details
781: here.
782:
783: \end{proof}
784:
785: The perturbation can be illustrated as in Figure
786: \ref{deform:triangle1} where $n=4$. From now on, we will
787: use the following notations to denote the reducibles lines
788: for $Y$, $Y_1$ and $Y_0$ repectively:
789: \ba\begin{array}{c}
790: L_Y(\s_k)= \{ (u,v) | \, v=f'_k(u-\eta) +\eta \}\\[2mm]
791: L_{Y_1}(\s_k)=\{ (u,v) | \, v=u+2k+1 \}\\[2mm]
792: L_{Y_0}(\s_k)= \{ (u,v) | \, v=2k+\eta \}.
793: \end{array}
794: \label{reducible:lines}
795: \na
796:
797: \begin{figure}[ht]
798: \epsfig{file= ex5fig0.eps,angle= 0}
799: \caption{The perturbed geometric triangle as in Theorem \ref{decomposition}}
800: \label{deform:triangle1}
801: \end{figure}
802:
803:
804:
805: With a more careful study of the perturbed geometric triangles, we have
806: the following decomposition of 3-dimensional monopoles
807: on $Y$ under the Dehn surgery.
808:
809: \begin{The}
810: With the perturbations and metrics on $Y, Y_1$ and
811: $Y_0$ as in Theorem \ref{decomposition}, and for any
812: fixed $m, p \in \{0, \cdots, n-1\}$, there exists
813: a further perturbation on $Y$,
814: such that we have the following relation between the
815: critical sets of the Chern-Simons-Dirac functional
816: on the manifolds $Y, Y_1$ and $Y_0$:
817: \ba \label{split-mp}
818: \M_Y (\s\otimes L_m)
819: = \M_{Y_1} (\s\otimes L_m)
820: \cup \bigcup_{k\in \Z} \M_{Y_0} (\s\otimes L_{nk+p}).
821: \na
822: \label{refine:decomposition}
823: \end{The}
824: \begin{proof}
825: In the proof of Theorem \ref{decomposition}, we know
826: that the perturbed Seiberg-Witten monopoles on $Y, Y_1$
827: and $Y_0$ are given by the following gluing models (here
828: we assume that $Y$ is a rational homology 3-sphere):
829: \[
830: {\cal M}^*_{Y(r)}(\s\otimes L_k)
831: \cong {\cal M}^*_V\backslash \partial_\infty^{-1}(U_\theta)
832: \times_{\chi_0(T^2,Y)} \{ v=2k \},
833: \]
834: \[
835: {\cal M}^*_{Y_1(r)}(\s\otimes L_k)
836: \cong {\cal M}^*_V\backslash \partial_\infty^{-1}(U_\theta)
837: \times_{\chi_0(T^2,Y_1)} \{ v= u+2k+1\},
838: \]
839: \[
840: {\cal M}^*_{Y_0(r)}(\s\otimes L_k)
841: \cong {\cal M}^*_V\backslash \partial_\infty^{-1}(U_\theta)
842: \times_{\chi_0(T^2,Y_0)} \{ u=2k\}.
843: \]
844: Note that the additional perturbation (\ref{eta:pert}) of the
845: curvature equation on $\nu(K)$ inside $Y$ and inside $Y_0$
846: introduce a shift of coordinates $(u, v)$
847: to $(u+\eta, v+\eta)$. We can introduce these new coordinates,
848: still denoted by $(u, v)$. Then the reducible line for $V$
849: is given by $u = -\eta$ in $H^1(T^2, \R)$.
850:
851: We will show that there exists a further surgery
852: perturbation on $Y$ that suits the purpose of
853: identifying monopoles on $Y, Y_1$ and
854: $Y_0$ as stated in the Theorem.
855: Without loss of the generality,
856: after possible coordinates change, we can assume that
857: $\s\otimes L_m = \s\otimes L_0$.
858: Fix $p\in \{0, \cdots, n-1\}$. We will construct a function
859: $f_0': \R \to \R$,
860: which depends on a small $\epsilon >0$, such that,
861: as $\epsilon \to 0$, the curve $v=f_0'(u)$ approaches
862: the union of lines
863: \[ L_{Y_1}(\s\otimes L_0)\cup \bigcup_{k\in \Z} L_{Y_0}(\s\otimes
864: L_{nk+p}) \]
865: where
866: $ L_{Y_1}(\s\otimes L_0)=\{ v=u+1\},
867: L_{Y_0}(\s\otimes L_{nk+p})= \{ u= 2nk+2p\}.$
868:
869: We identify $\chi_0(T^2, V)$ with the fundamental domain
870: \[
871: \{ u\in \R\} \times \{ 0\le v < 2n \},
872: \]
873: in $H^1(T^2, \R)$. The asymptotic values
874: $\partial_\infty (\M^*_V) \subset \chi_0(T^2, V)$ can
875: be lifted to $H^1(T^2, \R)$ periodically. Using this
876: $2n$--periodicity, we only need to construct
877: a function $f_0: [-1, 2n-1] \to [0, 2n]$,
878: which depends on $\epsilon$, such that, for any given
879: $\epsilon\leq \epsilon_0$, we have
880: \[
881: \sup_{t\in [-1, 2n-1]\backslash [-\epsilon , \epsilon]}
882: |f'_0 (t) - t | < \epsilon.
883: \]
884: Such function can be easily constructed as in the
885: proof of Theorem \ref{decomposition}. Then
886: over $[2nk-1, 2nk+2n-1]$, $f_0'(t)$ is defined
887: to be $f'_0 (t)+2nk$. See Figure
888: \ref{deform:triangle2} where the perturbation is
889: illustrated in the cases with $n=4, m=p=0$ and
890: with $n=4, m=0, p=2$, respectively. The general case
891: can be proved by a similar method.
892: \end{proof}
893:
894:
895: \begin{figure}[ht]
896: \epsfig{file= ex5fig1.eps,angle= 0}
897: \caption{The perturbed geometric triangle as in Theorem
898: \ref{refine:decomposition}
899: \label{deform:triangle2}}
900: \end{figure}
901:
902:
903: \section{The relative gradings}
904:
905: In the previous section, we have the following decomposition:
906: \[
907: \M_{Y, \mu} (\s\otimes L_m)
908: = \M_{Y_1} (\s\otimes L_m)
909: \cup \bigcup_{k\in \Z} \M_{Y_0} (\s\otimes L_{nk+p}) \]
910: for any fixed $m, p\in \{0, \cdots, n-1\}$.
911: Assume that $Y$ is a rational homology 3-sphere.
912: First we fix a grading on $\M_{Y, \mu} (\s\otimes L_0)$ defined
913: in terms of the spectral flow of the linearization operator
914: for the 3-dimensional Seiberg-Witten equations along a path
915: connecting an irreducible monopole in $\M_{Y,\mu} (\s\otimes L_0)$
916: to the unique reducible $\theta_Y(0)$ in the configuration
917: space for $(Y, \s\otimes L_0)$. Then
918: the analysis of the relative grading in Part I section 7 \cite{CMW}
919: can be applied to induce a compatible grading
920: on $\M_{Y_1} (\s\otimes L_0) \cup
921: \bigcup_{k\in \Z} \M_{Y_0} (\s\otimes L_{nk+p})$ as follows, cf.
922: Proposition 7.3 -- Corollary 7.7 in \cite{CMW}.
923:
924: \begin{Pro}\label{relative:grading0} Let $Y$ be a rational homology 3-sphere.
925: For any fixed $p \in \{0, \cdots, n-1\}$, the Floer complexes
926: \[
927: C_*(Y, \s\otimes L_0, \mu) = \oplus _{a\in \M_{Y,\mu} (\s\otimes L_0)}
928: \Z \la a \ra,
929: \]
930: \[
931: C_*(Y_1, \s\otimes L_0) = \oplus _{a\in \M_{Y_1} (\s\otimes L_0)} \Z
932: \la a \ra,
933: \]
934: \[
935: C_*(Y_0, \s\otimes L_{nk+p}) = \oplus _{a\in \M_{Y_0}
936: (\s\otimes L_{nk+p})} \Z \la a\ra,
937: \]
938: have a compatible relative grading of generators in the
939: following sense.
940: \begin{enumerate}
941: \item Suppose given two irreducible critical points $a, b$
942: in $\M^*_{Y_1}(\s\otimes L_0)$, and the corresponding elements
943: $a^\epsilon, b^\epsilon$ in $\M^*_{Y,\mu}(\s\otimes L_0)$
944: under the above decomposition (\ref{split-mp}). Then
945: \[
946: \deg_{Y,\mu}(a^\epsilon)-\deg_{Y,\mu}(b^\epsilon)=
947: \deg_{Y_1}(a)-\deg_{Y_1}(b).
948: \]
949: \item Suppose given two monopoles $a, b$
950: in $\M^*_{Y_0}(\s\otimes L_{nk+p})$, and the corresponding elements
951: $a^\epsilon, b^\epsilon$ in $\M^*_{Y, \mu}(\s\otimes L_0)$
952: under the above decomposition (\ref{split-mp}). Then
953: \[
954: \deg_{Y_0,\s_k}(a)-
955: \deg_{Y_0,\s_k}(b) = \deg_{Y,\mu}(a^\epsilon)-
956: \deg_{Y,\mu}(b^\epsilon) \ \ \hbox{ mod } (2nk+2p).
957: \]
958: Therefore, the grading $\deg_{Y,\mu}$ defines a $\Z$-valued lift
959: of the $\Z_{2nk+2p}$-valued relative index on
960: $\M^*_{Y_0}(\s\otimes L_{nk+p})$ under the decomposition
961: of $\M^*_{Y, \mu}(\s\otimes L_0)$.
962: \end{enumerate}
963:
964: \end{Pro}
965:
966: For a general 3-manifold $(Y, \s)$ with $b_1(Y)> 0$
967: and a knot $K$ representing
968: a torsion element of order $n$, if $\s$ is a torsion
969: $\spinc$ structure, we know that,
970: for any $k \in \{0, \cdots, n-1\}$, the $\spinc$ structure
971: $\s\otimes L_k$ also has a torsion class $c_1(\s\otimes L_k)$. Thus,
972: after a small perturbation to get rid of the $(S^1)^{b_1(Y)}$-family
973: of reducibles, we have a $\Z$-graded
974: $$\M^*_{Y, \mu}(\s\otimes L_k) \cong \M_{Y,\mu}(\s\otimes L_k).$$
975: Then it is easy to see that the results
976: in Part I section 7 \cite{CMW} hold in this case as well
977: without any substantial change.
978:
979: Now assume that $c_1(\s)$ is a non-torsion element, with multiplicity $2\ell$
980: in $H^2(Y, \Z)/\hbox{Torsion}$. Then, for
981: any $k \in \{0, \cdots, n-1\}$, the set of generators
982: $$\M^*_{Y, \mu}(\s\otimes L_k) \cong \M_{Y, \mu}(\s\otimes L_k)$$
983: is $2\ell$-graded, and so is
984: $$\M^*_{Y_1}(\s\otimes L_k) \cong \M_{Y_1}(\s\otimes L_k).$$
985: Then for any $k\in \Z$, any non-empty moduli space
986: $$\M^*_{Y,\mu}(\s\otimes L_k) \cong \M_{Y, \mu}(\s\otimes L_k)$$
987: is $\Z_{2\ell_{[k]}}$-graded, where $2\ell_{[k]}$
988: is the maximum common factor of $2\ell$ and $2k$.
989: We can choose a relative grading on $\M^*_{Y, \mu}(\s\otimes L_0)$
990: by the spectral flow of the linearization operator along
991: a path connecting any irreducible monopole
992: in $\M^*_{Y}(\s\otimes L_0)$ to a fixed monopole
993: $a_0$ in $\M^*_{Y, \mu}(\s\otimes L_0)$.
994: Note that this grading is $2\ell$-graded. Then the analysis
995: in Part I section 7 can also be applied to obtain the
996: following proposition.
997:
998: \begin{Pro}\label{relative:grading1}
999: For $p \in \{0, \cdots, n-1\}$ and $k\in \Z$, the Floer complexes
1000: $C_*(Y, \s\otimes L_0), C_*(Y_1, \s\otimes L_0)$, and
1001: $C_*(Y_0, \s\otimes L_{nk+p})$ have a compatible
1002: relative grading of generators in the following sense.
1003: \begin{enumerate}
1004: \item Suppose given two irreducible critical points $a, b$
1005: in $\M^*_{Y_1}(\s\otimes L_0)$, and the corresponding elements
1006: $a^\epsilon, b^\epsilon$ in $\M^*_{Y, \mu}(\s\otimes L_0)$,
1007: then
1008: \[
1009: \deg_{Y,\mu}(a^\epsilon)-\deg_{Y,\mu}(b^\epsilon)=
1010: \deg_{Y_1}(a)-\deg_{Y_1}(b),
1011: \]
1012: as a $\Z_{\ell}$-valued function.
1013: \item Suppose given two monopoles $a, b$
1014: in $\M^*_{Y_0}(\s\otimes L_{nk+p})$, and the corresponding elements
1015: $a^\epsilon, b^\epsilon$ in $\M^*_{Y, \mu}(\s\otimes L_0)$,
1016: then
1017: \[
1018: \deg_{Y_0,\s_k}(a)-
1019: \deg_{Y_0,\s_k}(b) = \deg_{Y,\mu}(a^\epsilon)-
1020: \deg_{Y,\mu}(b^\epsilon) \ \ \hbox{ mod } (2\ell_{[nk+p]}).
1021: \]
1022: Thus, the induced grading $ \deg_{Y,\mu}$ on
1023: $\M^*_{Y_0}(\s\otimes L_{nk+p})$, under the
1024: decomposition (\ref{split-mp}), defines a choice of a
1025: $\Z_{2\ell}$-valued lifting of the $\Z_{2\ell_{[nk+p]}}$-valued
1026: relative index.
1027: \end{enumerate}
1028: \end{Pro}
1029:
1030:
1031:
1032: In the rest of this section we discuss the induced
1033: gradings on the various $\M_Y^*(\s\otimes L_m)$, for $m \in \{1,
1034: \cdots, n-1\}$.
1035: Notice that, in the case of a rational homology 3-sphere $Y$,
1036: the perturbed line $L_{Y}(\s\otimes L_m)$ is a
1037: deformation of the line $L_{Y}(\s\otimes L_0)$. Along this
1038: deformation, the corresponding parameterized
1039: reducibles form a path
1040: connecting $\theta_Y(0)$ to $\theta_Y(m)$. As the following Lemma
1041: shows, this deformation
1042: can be realized as a perturbation of the monopole equations
1043: for $(Y, \s\otimes L_0)$.
1044:
1045: \begin{Lem}
1046: For any $m \in \{1, \cdots, n-1\}$, there exists a perturbation
1047: $\mu_m$ for the Seiberg-Witten monopole equations
1048: on $(Y, \s\otimes L_0)$,
1049: such that there is a diffeomorphism:
1050: \[
1051: \M_{Y, \mu_m} (\s\otimes L_0) \cong \M_Y(\s\otimes L_m).
1052: \]
1053: Moreover, if $Y$ is a rational homology 3-sphere,
1054: the grading on $\M_Y^*(\s\otimes L_m)$
1055: induces a grading on $\M_{Y, \mu_m}^* (\s\otimes L_0)$. Under the
1056: above identification,
1057: the resulting grading differs from the original grading
1058: on $\M_{Y}^* (\s\otimes L_0)$ by
1059: the wall-crossing formulae studied in
1060: \cite{MW1} for $(Y, \s\otimes L_0)$. If $b_1(Y)> 0$,
1061: then the induced grading on $\M_Y(\s\otimes L_m)$
1062: from $\M_{Y, \mu_m} (\s\otimes L_0)$ agrees with the
1063: relative grading on $\M_Y(\s\otimes L_m)$.
1064: \label{pert-0m}
1065: \end{Lem}
1066:
1067: \begin{proof}
1068: The first claim follows from the gluing models
1069: of the monopoles in $\M_{Y} (\s\otimes L_0)$
1070: and $ \M_Y (\s\otimes L_m)$. Then, using the results in \cite{MW1}, we
1071: know that, in the
1072: case of the rational homology 3-sphere,
1073: the spectral flow of the twisted Dirac operator
1074: along the path of reducibles gives the
1075: index shift on
1076: $\M_{Y}^* (\s\otimes L_0)$ and $\M_{Y, \mu_m}^* (\s\otimes L_0)$
1077: according to the wall-crossing formulae derived in \cite{MW1}. Again
1078: by the results of \cite{MW1},
1079: for $Y$ with $b_1(Y)>0$, the induced
1080: grading from the parametrized spectral flow is same as the
1081: original relative index on $ \M_Y (\s\otimes L_m)$.
1082: \end{proof}
1083:
1084: Thus, Lemma \ref{pert-0m} provides a consistent way of assigning a
1085: choice of absolute grading on the various $\M_Y (\s\otimes L_m)$.
1086: The degree shift of Lemma \ref{pert-0m} can be described as follows.
1087: Let $\theta_Y(t)$ be the path of reducibles
1088: on $(Y, \s\otimes L_0) $ for the family of perturbations
1089: connecting $\M_{Y}^* (\s\otimes L_0)$
1090: to $\M_{Y, \mu_m}^* (\s\otimes L_0)$,
1091: then the wall crossing formulae in \cite{MW1}
1092: tell us that
1093: the index shift is given by the complex spectral flow
1094: \ba
1095: SF_\C(\dirac_{\theta_Y(t)}).
1096: \label{index:shift}
1097: \na
1098:
1099:
1100: \section{Geometric limits and the holomorphic triangles}
1101:
1102:
1103: \subsection{Surgery cobordisms}
1104:
1105: We first briefly describe the surgery cobordisms from
1106: $Y_1$ to $Y$, from $Y$ to $Y_0$, and from $Y_0$ to $Y_1$,
1107: respectively, as in \cite{MW3}. The cobordism $W_1$, from
1108: $Y_1$ to $Y$, is obtained by removing from the trivial cobordism
1109: $Y_1\times [0, 1]$ an $S^1\times D\cong \nu(K)\times \{ 1\}$, where $D$ is
1110: a disk, and $\nu(K)$ is the tubular neighbourhood of the knot in
1111: $Y_1$, and then attaching a 2-handle with framing $-1$.
1112: We denote by $D_1$ the core disk of the 2-handle in $W_1$.
1113: Similarly, the cobordism $W_0$, form $Y$ to $Y_0$, is obtained by removing
1114: from the trivial cobordism $Y_0\times [0, 1]$ an $S^1\times D\cong
1115: \nu(K)\times \{ 0\}$ and attaching a 2-handle with framing zero.
1116: We denote by $D_0$ the core disk of the 2-handle in $W_0$.
1117: Attaching the two-handle has the effect of modifying the
1118: boundary component $Y_1\times \{ 1\}$ in the trivial cobordism to the
1119: boundary component $Y\times \{ 1 \}$ in the non-trivial cobordism
1120: $W_1$, or, respectively, the
1121: boundary component $Y_0\times \{ 0 \}$ in the trivial cobordism
1122: to the boundary component $Y\times \{ 0 \}$ in $W_0$.
1123: The cobordism $\bar W_2$ connecting $Y_0$ and $Y_1$,
1124: satisfies the relation
1125: $$ \bar W = \bar W_2 \# \C \P^2, $$
1126: where $\bar W$ is the composite cobordism $\bar W=\bar W_0 \cup_Y \bar
1127: W_1$.
1128:
1129: We assume that the 3-manifolds $Y_1$, $Y$, and $Y_0$ are endowed with metrics
1130: with a long cylinder $T^2\times [-r,r]$, as specified in
1131: section 2 (see also \cite{CMW}). We consider the manifolds $W_1$ and
1132: $W_0$ endowed
1133: with infinite cylindrical ends $Y_1 \times (-\infty,
1134: -T_0]$ and $Y\times [T_0,\infty)$, and $Y_0 \times
1135: [T_0,\infty)$ and $Y\times (-\infty, -T_0]$, respectively.
1136: As in \cite{MW3}, we can decompose the cobordisms $W_i$ as
1137: \ba
1138: W_i= V\times \R \cup_{T^2\times \R} T^2\times [-r,r]\times \R
1139: \cup_{T^2\times \R} W_i(\nu(K)).
1140: \label{split:cobord}
1141: \na
1142: The region $W_i(\nu(K))$ has the following property. There
1143: is a compact set ${\cal K}$ in $W_i$ such that the intersection
1144: ${\cal K} \cap W_i(\nu(K))$ is obtained by attaching a 2-handle
1145: $D\times D$ to the product $\nu(K)\times [-T_0, T_0]$, and, outside
1146: of ${\cal K}$, the region ${\cal K}^c \cap W_i(\nu(K))$ consists of product
1147: regions $\nu(K)\times [ T_0 ,\infty)$ and $\nu(K)\times (-\infty, -T_0]$,
1148: and $T^2\times [r_0,r] \times [-T_0,T_0]$.
1149:
1150: As in \cite{MW3}, consider an interior point $x_i$
1151: contained in the core disk of the 2-handle, $x_i \in D_i$,
1152: and we denote by $\hat W_i$ the
1153: punctured cobordism $\hat W_i = W_i \backslash \{ x_i \}$. Similarly,
1154: we can consider the punctured manifold
1155: $$ \hat W_i(\nu(K)) = W_i(\nu(K))\backslash \{ x_i \}. $$
1156: In the manifolds $\hat W_i(\nu(K))$, endowed with an extra
1157: asymptotic end of the form $S^3 \times [0,\infty)$ at
1158: the puncture, we can identify a product region
1159: \ba
1160: {\cal V}=\nu(K)_{r_0}\times \R \cong D\times (D_i\backslash \{ x_i
1161: \}).
1162: \label{prodW}
1163: \na
1164: Thus, we identify the manifold $W_i$ with a
1165: connected sum
1166: $$ W_i = \hat W_i \# Q_i, $$
1167: with a long cylindrical neck $S^3\times [-T(r),T(r)]$, and with $Q_i$ a
1168: 4-ball, where $S^3$ is decomposed as the union of
1169: two solid tori in the standard way,
1170: $S^3=\nu \cup \tilde \nu$, with $\nu\cong \tilde\nu \cong D\times
1171: S^1$. Then the product region ${\cal V}$ of (\ref{prodW})
1172: in $W_i$ identifies the standard solid torus $\nu$ in $S^3$ with the
1173: neighbourhood
1174: $\nu(K)$ of the knot $K$ in $Y$. Similarly, there is a product
1175: region $\tilde {\cal V}$ which identifies the other solid torus
1176: $\tilde\nu$ in $S^3$ with the
1177: tubular neighbourhood $\nu(K)$ in $Y_i$, after the surgery.
1178: The resulting punctured cobordism can be written as
1179: \ba
1180: \hat W_i(r)=( V_r\times \R) \cup {\cal V}(r) \cup \tilde {\cal V}(r).
1181: \label{hat:W(r)}
1182: \na
1183:
1184: We now impose a choice of metrics and perturbations for
1185: the Seiberg-Witten equations
1186: on the cobordisms as in subsection 2.2 of \cite{MW3}. Then we can adopt the
1187: results of \cite{MW2}\cite{MW3} to understand
1188: the asymptotic limits of finite energy monopoles, under the
1189: splitting of the punctured cobordisms as $r\to\infty$, as in
1190: (\ref{hat:W(r)}).
1191:
1192:
1193: \subsection{Geometric limits and holomorphic triangles}
1194:
1195: We assume that $Y$ and $Y_1$ are endowed with the
1196: $\spinc$ structure $\s\otimes L_0$. The other $\spinc$ structures can
1197: be studied analogously.
1198:
1199: On the surgery cobordism $W_1(r)$, there
1200: is a $\Z$-family of $\spinc$ structures whose
1201: restrictions to the two ends agree with
1202: $\s\otimes L_0$ on $Y$ and $Y_1$ respectively.
1203: We will study the moduli spaces of monopoles on
1204: $W_1(r)$ with asymptotic values in
1205: $\M_{Y_1}(\s\otimes L_0)$ and
1206: $\M_{Y, \mu}(\s\otimes L_0)$ at the two ends,
1207: where $\mu$ is the surgery perturbation defined by $f_0'$ as
1208: in the proof of Theorem \ref{refine:decomposition}.
1209: In the case $b_1(Y)>0$, we only consider
1210: the components of minimal energy, as defined in \cite{MW3} \cite{MW4},
1211: among all the possible moduli spaces with different $\spinc$
1212: structures and with the given asymptotic values.
1213: In particular, for a rational homology 3-sphere $Y$,
1214: we only consider the moduli spaces of minimal
1215: dimension among the $\Z$-family of $\spinc$ structures.
1216:
1217: With this convention understood, we denote by
1218: $\M^{W_1}(a_1, a)$
1219: the moduli space with asymptotic values
1220: $a_1\in\M_{Y_1}(\s\otimes L_0)$ and $a\in\M_{Y,\mu}(\s\otimes L_0)$.
1221: Similarly, we denote by $\M^{W_0}(a, a_0)$
1222: the moduli space with asymptotic values
1223: $a\in\M_{Y,\mu}(\s\otimes L_0)$ and
1224: $a_0 \in \M_{Y_0}(\otimes L_{nk+p})$ for $k\in \Z$, for a
1225: fixed $p\in\{0, \ldots, n-1\}$. Under a generic choice of the perturbation,
1226: all these moduli spaces $\M^{W_1}(a_1,a)$ and
1227: $\M^{W_0}(a,a_0)$ are cut out transversely and of the expected
1228: dimension.
1229:
1230: The convergence and gluing arguments developed in \cite{MW1} can be
1231: applied to this case as well, to give the following
1232: compactifications of $\M^{W_1}(a_1, a)$ and $\M^{W_0}(a, a_0)$.
1233:
1234: \begin{Pro}\label{compactification}
1235: Suppose that $\M^{W_1}(a_1, a)$ is non-empty, then $\M^{W_1}(a_1, a)$
1236: admits a compactification to a manifold
1237: with corners, where the
1238: codimension $1$ boundary strata consist of
1239: \ba
1240: \label{compact:W1}
1241: \begin{array}{c}
1242: \bigcup_{c\in \M_{Y, \mu}^*(\s\otimes L_0)}
1243: \M^{W_1}(a_1,c)\times \hat \M_{Y,\mu}(c,a)
1244: \\[2mm]
1245: \cup \bigcup_{c_1 \in \M_{Y_1}^*(\s\otimes L_0)} \hat\M_{Y_1}(a_1,c_1)\times
1246: \M^{W_1}(c_1,a),
1247: \end{array}
1248: \na
1249: and with extra components
1250: \ba
1251: \begin{array}{c}
1252: \hat \M_{Y_1}(a_1,\theta_1)\times U(1) \times
1253: \M^{W_1}(\theta_1,a) \\[2mm]
1254: \M^{W_1}(a_1,\theta)\times U(1) \times\hat
1255: \M_{Y,\mu}(\theta,a),
1256: \end{array}
1257: \label{red:compact:W1}
1258: \na
1259: when splitting through the reducibles $\theta_1$ and
1260: $\theta$ in $ \M_{Y_1}^*(\s\otimes L_0)$ and
1261: $\M_{Y, \mu}^*(\s\otimes L_0)$ respectively .
1262: We also have the similar compactification for
1263: $\M^{W_0} (a, a_0)$.
1264: \end{Pro}
1265:
1266:
1267: Now we can describe the geometric limits of monopoles
1268: in $\M^{W_1}(a_1, a)$ and $\M^{W_0} (a, a_0)$
1269: when stretching $r\to \infty$ inside
1270: \[
1271: \hat W_i(r)= V_r\times \R \cup {\cal V}(r) \cup \tilde {\cal V}(r).
1272: \]
1273: We only describe the case of $\hat W_1(r)$. With
1274: similar arguments we have the corresponding
1275: geometric limits for $\hat W_0(r)$.
1276: The proof of the results stated below on these geometric limits
1277: follows from the same arguments of \cite{MW2}\cite{MW3}.
1278:
1279: Consider elements $a_i^{(1)}\in \M_{Y_1}^*(\s\otimes L_0)$
1280: and $a_j(\epsilon) \in \M_{Y,\mu}^*(\s\otimes L_0)$,
1281: which we can write as
1282: $$ a_i^{(1)}=[(A_i^-,\psi_i^-)\#(a_{\infty,i}^-,0)] $$
1283: $$ a_j(\epsilon)= [(A_j^+(\epsilon),\psi_j^+(\epsilon))\#
1284: (a_{\infty,j}^+(\epsilon),0)], $$
1285: with
1286: $$[A_i^-,\psi_i^-], \, \, [A_j^+(\epsilon),\psi_j^+(\epsilon)]\in \,
1287: \, \M_V^*(\s\otimes L_0|_V)$$
1288: $$ a_{\infty,i}^-\in \, \, L_{Y_1}(\s\otimes L_0),
1289: a_{\infty,j}^+(\epsilon)\in \, \, L_{Y,\mu}(\s\otimes L_0).$$
1290:
1291: \begin{The} (Proposition 3.1 and Remark 6.2 in \cite{MW3})
1292: Assume that $\M^{W_1(r)}\bigl(a_i^{(1)}, a_j(\epsilon)\bigr)$
1293: is non-empty for all
1294: sufficiently large $r$. Then a
1295: family of solutions $[\A_1(r), \Psi_1(r)]$ in
1296: $\M^{W_1(r)}\bigl(a_i^{(1)}, a_j(\epsilon)\bigr)$
1297: defines the following geometric limits on
1298: $ V_r\times \R \cup {\cal V}(r) \cup \tilde {\cal V}(r)$
1299: as $r\to \infty$.
1300:
1301: \noindent
1302: (a). A finite energy solution $[\A',\Psi']^\epsilon$ of the perturbed
1303: Seiberg-Witten equations on $V \times \R$, with a radial limit
1304: $a_\infty(\epsilon)$ in $\partial_\infty (\M_V^*)\subset
1305: \chi_0(T^2,V)$, and with temporal limits $[A,\psi]_1^\epsilon$ and
1306: $[\tilde A, \tilde \psi]_1^\epsilon$ in
1307: $ \partial_\infty^{-1}(a_\infty(\epsilon)) \subset \M_V^*. $
1308:
1309: \noindent
1310: (b). Two paths $[A(t),\psi(t)]_1^\epsilon$ in $\M_V^*$, for
1311: $t\in [-1,0)$ and $t\in (0,1]$, with
1312: \[
1313: \begin{array}{lll}
1314: [A(-1),\psi(-1)]_1^\epsilon = [A_i^-, \psi_i^-], &\quad&
1315: \lim_{t\to 0-} [A(t),\psi(t)]_1^\epsilon =
1316: [A,\psi]_1^\epsilon \\[2mm]
1317: [A(1),\psi(1)]_1^\epsilon = [ A_j^+(\epsilon) ,
1318: \psi_j^+(\epsilon)], &\quad&
1319: \lim_{t\to 0+} [A(t),\psi(t)]_1^\epsilon = [\tilde A,
1320: \tilde \psi]_1^\epsilon
1321: \end{array}\]
1322: These paths induce a continuous, piecewise smooth path
1323: $a_1^\epsilon(t)$ on $\partial_\infty (\M_V^*)$
1324: satisfying $ a_1^\epsilon(t)=\partial_\infty
1325: \bigl([A(t),\psi(t)]_1^\epsilon\bigr)$, with
1326: $$ a_1^\epsilon (-1)=a_{\infty,i}^- \ \ a_1^\epsilon
1327: (0)=a_\infty(\epsilon) \ \ a_1^\epsilon (1)=
1328: a_{\infty,j}^+(\epsilon). $$
1329: As $\epsilon\to 0$, these geometric limits define paths
1330: $[A(t),\psi(t)]$ and $a(t)$ with
1331: $$ a(-1)=a_{\infty,i}^-, \ \ a(0)=a_\infty, \ \
1332: a(1)=a_{\infty,j}^+, $$
1333: in $\partial_\infty (\M_V^*)\subset
1334: \chi_0(T^2,V)$, and $a_\infty=\lim_\epsilon a_\infty(\epsilon)$.
1335:
1336: (c) There is a holomorphic triangle in $H^1(T^2,\R)$ with
1337: vertices at
1338: $$ a^-_{\infty,i}, \vartheta_1, a^+_{\infty,j}(\epsilon) $$
1339: and sides given by parameterized arcs along the lines
1340: $L_{Y_1}(\s\otimes L_0)$, $L_{Y,\mu}(\s\otimes L_0)$ and
1341: $ \{ a_1^\epsilon(t) \}\subset \partial_\infty(\M_V^*).$
1342: Here we denote
1343: $ \vartheta_1 = L_{Y_1}(\s\otimes L_0)\cap L_{Y,\mu}(\s\otimes
1344: L_0).$
1345: \label{geom:lim1}
1346: \end{The}
1347:
1348: This Theorem shows that the
1349: moduli space $\M^{W_1(r)} \bigl(a_i^{(1)}, a_j(\epsilon)\bigr)$
1350: is characterized by the geometric limits on $V\times \R$ from (a) and
1351: the holomorphic triangles in (c). Two typical
1352: holomorphic triangles
1353: for $\M^{W_1(r)} \bigl(a_i^{(1)}, a_j(\epsilon)\bigr)$
1354: or $\M^{W_0(r)} \bigl(a_i(\epsilon), a_j^{(0)}\bigr)$
1355: are illustrated in
1356: Figure \ref{holomorphic:triangles} where $n=4, m=k=0$ and $p=2$.
1357: Here the points $\vartheta_i$ are the intersection points
1358: $$ \vartheta_1 = L_{Y_1}(\s\otimes L_0) \cap L_{Y,\mu}(\s\otimes
1359: L_0)$$
1360: for $W_1$ and
1361: $$ \vartheta_0 = L_{Y,\mu}(\s\otimes L_0) \cap L_{Y_0}(\s\otimes
1362: L_{nk+p}) $$
1363: in the case of $W_0$.
1364: In other words, $\vartheta_i$ is the restriction to
1365: $T^2=\partial
1366: \nu=\partial \tilde \nu$ of the unique reducible point $\theta_{S^3}$
1367: at the puncture in the cobordism.
1368:
1369: \begin{figure}[ht]
1370: \epsfig{file=ex5fig2.eps,angle= 0}
1371: \caption{Holomorphic triangles for $W_1(r)$ and $W_0(r)$}
1372: \label{holomorphic:triangles}
1373: \end{figure}
1374:
1375:
1376:
1377: In order to glue back the
1378: geometric limits and the holomorphic triangles, there are some
1379: admissible conditions for the geometric limits on $V\times \R$,
1380: which are studied in section 6.1, section 6.2 and
1381: section 6.3 \cite{MW3}.
1382: As in section 6 \cite{MW3}, assuming that the element
1383: $a_\infty$ is away from the bad
1384: points in the character variety of $T^2$, we denote by
1385: $\hat\M_{V\times \R}(a_\infty)$ the {\em balanced energy} moduli space
1386: of the Seiberg-Witten equations on $V\times \R$ with
1387: asymptotic value $a_\infty$ in the radial direction. Then we have
1388: $$ \hat \M_{V\times \R}(a_\infty) = \bigcup_{[A,\psi],[\tilde A,\tilde
1389: \psi]}\hat \M_{V\times \R}( [A,\psi],[\tilde A,\tilde
1390: \psi], a_\infty), $$
1391: with
1392: $$ [A,\psi],[\tilde A,\tilde\psi] \in \partial_\infty^{-1}(a_\infty)
1393: \subset \M^*_V. $$
1394: Each moduli space
1395: $$ \M_{V\times \R}( [A,\psi],[\tilde A,\tilde
1396: \psi], a_\infty), $$
1397: for a fixed choice of $[A,\psi]$ and $[\tilde A,\tilde
1398: \psi]$ in $\partial_\infty^{-1}(a_\infty)$ in $\M_V^*$, is a smooth
1399: finite dimensional oriented manifold of the expected dimension, where
1400: the orientation is given by the corresponding determinant line bundle
1401: of the linearization operator for the monopole equations on
1402: $V\times \R$.
1403:
1404: Remember that we are only considering the components with minimal
1405: energy or minimal dimension among all the moduli spaces
1406: of finite energy monopoles with fixed asymptotic values.
1407: Recall that there is a notion of admissible triples
1408: (cf. Definition 6.4 \cite{MW3}), which
1409: singles out those elements $([A,\psi],[\tilde A,\tilde \psi], a_\infty)$
1410: which arise as part of the geometric limits of solutions in
1411: $\M^{W_1(r)}(a_1,a)$
1412: (or $\M^{W_0(r)}(a_1,a_0)$, or $\M_{Y(r)\times \R}(a,b)$ etc).
1413: For example, in the case of $\M^{W_1(r)}(a_1,a)$, with
1414: $$ a_1=[(A^-,\psi^-)\#(a^-,0)] $$
1415: $$ a=[(A^+,\psi^+)\#(a^+,0)], $$
1416: an admissible triple $([A,\psi],[\tilde A,\tilde \psi], a_\infty)$
1417: must satisfy the following conditions:
1418: there exists a smooth regular
1419: parameterization $a(t)$, for $t\in [-1,1]$ of the path in
1420: $\partial_\infty(\M_V^*)$ connecting $a^-$ and $a^+$, such that
1421: $a(0)=a_\infty$, and corresponding smooth paths $[A(t),\psi(t)]$ in
1422: $\M_V^*$, for $t\in [-1,0)$ and $t\in (0,1]$, satisfying
1423: $\partial_\infty[A(t),\psi(t)]=a(t)$, and with
1424: $$ [A(-1),\psi(-1)]=[A^-,\psi^-] \ \ \hbox{ and
1425: } \ \ \lim_{t\to
1426: 0_-}[A(t),\psi(t)]=[A,\psi] $$
1427: $$ \lim_{t\to 0_+}[A(t),\psi(t)]=[\tilde A,\tilde \psi] \ \ \hbox{ and
1428: } \ \ [A(1),\psi(1)]=[A^+,\psi^+]. $$
1429: By the results of \cite{MW3}, we know that the possible choices of
1430: $a_\infty$ and of the admissible data in $\hat\M_{V\times
1431: \R}(a_\infty)$ are uniquely determined by the inequivalent holomorphic
1432: triangles $\Delta$ with vertices
1433: $\{ a^-,\vartheta_1, a^+ \}$ and sides along the union of
1434: Lagrangians $\ell \cup \ell_1 \cup \ell_\mu$, with $\ell$ defined by
1435: the asymptotic values $\partial_\infty(\M_V^*)$. We denote by
1436: $\Xi_{W_1} (a_1, a)$ the set of such inequivalent holomorphic
1437: triangles.
1438: Moreover, under the identification of the choice of admissible triples
1439: $([A,\psi],[\tilde A,\tilde \psi], a_\infty)$
1440: with the choice of inequivalent oriented holomorphic triangles in
1441: $\Xi_{W_1} (a_1, a)$, the gluing map
1442: gives an orientation preserving diffeomorphism
1443: \ba \label{glue:W_1}
1444: \#_{W_1}:
1445: \bigcup_{([A,\psi],[\tilde A,\tilde \psi], a_\infty)\in \Xi_{W_1} (a_1, a)}
1446: \M_{V\times\R}( [A,\psi],[\tilde A,\tilde \psi], a_\infty)
1447: \to
1448: \M^{W_1}(a_1,a).
1449: \na
1450:
1451: Similarly, there are orientation preserving diffeomorphisms
1452: given by the gluing maps for $W_0(r)$, $(Y(r)\times \R, \s\otimes L_0)$,
1453: $(Y_1(r)\times \R, \s\otimes L_0)$ and $(Y_0(r)\times \R, \s\otimes
1454: L_{nk+p})$ defined over the set of admissible triples, which are in turn
1455: determined by the corresponding inequivalent oriented holomorphic triangles
1456: or holomorphic discs. The corresponding sets
1457: are denoted by
1458: $\Xi_{W_0}(a, a_0)$, $\Xi_{Y}(a, b)$, $\Xi_{Y_1}(a_1, b_1)$
1459: and $\Xi_{Y_0}(a_0, b_0)$, respectively.
1460: We summarize all the gluing theorems for the
1461: geometric limits and holomorphic triangles (discs) as follows:
1462:
1463: \begin{The} \label{final:glue}
1464: Suppose given a pair of monopoles $a^{(1)}, b^{(1)}$ in
1465: $\M_{Y_1}(\s\otimes L_0)$, and a pair of monopoles
1466: $a(\epsilon), b(\epsilon)$ in $\M_{Y_1, \mu}(\s\otimes L_0)$,
1467: where the surgery perturbation $\mu$ determined by $f_0'$ depends on a
1468: small parameter $\epsilon$. Suppose given a pair of monopoles
1469: $a^{(0)}, b^{(0)}$ in $\M_{Y_0}(\s\otimes L_{nk+p})$
1470: for some $k\in \Z$ and $p\in \{0, \cdots, n-1\}$. Then, for
1471: sufficiently large $r$,
1472: the gluing maps give the following
1473: orientation preserving diffeomorphisms
1474: \[\begin{array}{l}
1475: \bigcup_{\Delta\in \Xi_{Y} (a(\epsilon),
1476: b(\epsilon))}
1477: \M_{V\times\R}( \Delta)
1478: \stackrel{\#_{Y(r)\times \R}}{\longrightarrow}
1479: \M_{Y\times\R} (a(\epsilon), b(\epsilon)),
1480: \\
1481: \bigcup_{\Delta\in \Xi_{Y_1}(a^{(1)},
1482: b^{(1)})}
1483: \M_{V\times\R}( \Delta)
1484: \stackrel{\#_{Y_1(r)\times \R}}{\longrightarrow} \M_{Y_1\times\R} (a^{(1)},
1485: b^{(1)}),
1486: \\
1487: \bigcup_{\Delta\in \Xi_{Y_0
1488: }(a^{(0)}, b^{(0)})}\M_{V\times\R}( \Delta)
1489: \stackrel{\#_{Y_0(r)\times \R}}{\longrightarrow}
1490: \M_{Y_0\times\R} (a^{(0)}, b^{(0)}),
1491: \\
1492: \bigcup_{\Delta\in \Xi_{W_1} (a^{(1)},
1493: a(\epsilon))}
1494: \M_{V\times\R}( \Delta)
1495: \stackrel{\#_{W_1}}{\longrightarrow} \M^{W_1}(a^{(1)},
1496: a(\epsilon)),
1497: \\
1498: \bigcup_{\Delta\in \Xi_{W_0}
1499: (a(\epsilon), a^{(0)})}
1500: \M_{V\times\R}( \Delta)
1501: \stackrel{\#_{W_0}}{\longrightarrow} \M^{W_0} (a(\epsilon), a^{(0)}),
1502: \end{array}\]
1503: where, for simplicity, we denoted a triple $([A,\psi],[\tilde A,\tilde
1504: \psi], a_\infty)$ by $\Delta$.
1505: \end{The}
1506:
1507:
1508:
1509:
1510:
1511: \section{Proof of exactness and the surgery triangle}
1512:
1513: In this section, we will prove the main theorem of this paper. Notice
1514: that we only prove the refined exact triangle in Theorem
1515: \ref{main:theorem} for $m=0$. The arguments for this case can be
1516: adapted to give an analogous proof for $m\neq 0$.
1517:
1518: \subsection{The chain homomorphisms}
1519:
1520: Recall that the moduli spaces for $W_1$ and $W_0$ which we consider
1521: are only the components of minimal energy (or dimension). They are
1522: smooth and oriented manifolds of the expected dimension, and they can
1523: be compactified according to Proposition \ref{compactification}. In
1524: particular, whenever one such moduli space is 0-dimensional, we have a
1525: counting of points with the orientation. Thus,
1526: as in \cite{MW3}, we can define the chain homomorphisms
1527: between the Floer chain groups of $(Y, \mu, \s\otimes L_0)$,
1528: $(Y_1, \s\otimes L_0)$ and $(Y_0, \s\otimes L_{nk+p})$ as follows.
1529:
1530: \begin{Def}
1531: We define the map $w_*^1: C_*(Y_1)\to C_*(Y,\mu)$ by assigning the
1532: matrix elements
1533: $$ \la a , w_*^1(a_1) \ra =\# \M^{W_1} (a, a_1),$$
1534: where the right hand side is a counting of points with the orientation
1535: if $\M^{W_1} (a, a_1)$ is 0-dimensional and non-empty, and it is
1536: $\la a , w_*^1(a_1) \ra =0$ otherwise.
1537: Similarly, we define the map
1538: $w_*^0: C_*(Y,\mu)\to \oplus_k C_{(*)}(Y_0, \s\otimes L_{nk+p})$,
1539: with the the matrix coefficients
1540: $$ \la a_0 , w_*^0(a) \ra =\# \M^{W_0}(a, a_0) $$
1541: if $\M^{W_0}(a, a_0))$ is 0-dimensional and non-empty, and
1542: $ \la a_0 , w_*^0(a) \ra =0$ otherwise.
1543: Here the chain groups $C_{(*)}(Y_0, \s\otimes L_{nk+p})$ are
1544: equipped with the lifting of the relative
1545: grading on $\M_{Y_0}(\s\otimes L_{nk+p})$ as described in
1546: Proposition \ref{relative:grading1}.
1547: \label{w10}
1548: \end{Def}
1549:
1550: Notice that on $W_1$ there is a $\Z$-family of
1551: $\spinc$ structures which agree with
1552: $\s\otimes L_0$ when restricted to the two
1553: ends. From Lemma 4.1 in \cite{MW3}, the choice of components with
1554: minimal energy and dimension essentially
1555: excludes other $\spinc$ structures when
1556: we consider the moduli space $\M^{W_1}(a_1, a)$ with fixed
1557: asymptotic values $a_1 \in \M_{Y_1}^*(\s\otimes L_0)$ and
1558: $a\in \M^*_{Y, \mu}(\s\otimes L_0)$.
1559:
1560: In the case of $b_1(Y)>0$ and $c_1(\s\otimes L_0)$ is non-torsion,
1561: when both $\M_{Y, \mu}(\s\otimes L_0)$ and
1562: $\M_{Y_1}(\s\otimes L_0)$ are $\Z_{2\ell}$-graded, with
1563: $2\ell$ the multiplicity of $c_1(\s\otimes L_0)$ in
1564: $H^2(Y, \Z)/\hbox{Torsion}$, then Proposition \ref{relative:grading1}
1565: ensures that there exists
1566: a compatible grading on $\M_{Y_1}(\s\otimes L_0)$
1567: and a $\Z_{2\ell}$--lifting of the relative grading
1568: on any $\M_{Y_0}(\s\otimes L_{nk+p})$. For those
1569: $\spinc$ structures $\s\otimes L_0$ with torsion $c_1(\s\otimes L_0)$,
1570: the corresponding Floer homology is defined
1571: to be
1572: \[
1573: HF_*^{SW}(Y, \s\otimes L_0, \Z[[t]])|{t=0},
1574: \]
1575: as described in \cite{MW4}. Then it is easy to see that
1576: the choice of components with minimal energy and dimension makes
1577: $w_*^1$ and $w_*^0$ well-defined.
1578:
1579: Then, with the help of the compactifications in
1580: Proposition \ref{compactification},
1581: the proofs of Lemma 4.3 and Lemma 4.4 in \cite{MW3}
1582: go through without any substantial change to give the following
1583: Lemma.
1584:
1585: \begin{Lem}
1586: \begin{enumerate}
1587: \item The maps $w_*^i$ are chain homomorphisms.
1588: \item Suppose given $a_1$ in $\M_{Y_1}^*(\s\otimes L_0)$
1589: and $a_0$ in $\M_{Y_0}(\s\otimes L_{nk+p})$, such that the relative
1590: index induced from $\M_{Y, \mu}^*(\s\otimes L_0)$ as in
1591: Proposition \ref{relative:grading1} is zero, then the
1592: composite map $w^0_* \circ w^1_*$ is given by
1593: \[
1594: \la w^0_* \circ w^1_*(a_1), a_0 \ra =
1595: \# \M^{W} (a_1, a_0).
1596: \]
1597: Here we use the same convention on the choice of the
1598: moduli spaces for $W =W_1\#_Y W_0$.
1599: \end{enumerate}
1600: \label{chain:homo}
1601: \end{Lem}
1602:
1603: Thus, we have obtained a sequence of
1604: chain complexes induced by the surgery
1605: cobordisms:
1606: \[
1607: 0\to C_*(Y_1, \s\otimes L_0)\stackrel{w^1_*}{\longrightarrow}
1608: C_*(Y,\mu, \s\otimes L_0)\stackrel{w^0_*}{\longrightarrow}
1609: \oplus_{k\in \Z} C_{(*)}(Y_0,\s\otimes L_{nk+p}) \to 0,
1610: \]
1611: for any fixed $p\in \{0, \cdots, n-1\}$.
1612: We shall prove that this sequence is exact and that the corresponding
1613: exact triangle is the surgery triangle, namely the connecting
1614: homomorphism is also defined via a surgery cobordism. The gluing Theorem
1615: \ref{final:glue}
1616: for the admissible geometric limits on $V\times \R$
1617: and the corresponding holomorphic triangles (or discs)
1618: will play a crucial role in the proof of these statements.
1619:
1620:
1621:
1622: \subsection{Proof of exactness}
1623: We first prove that $w^1_*$ is injective and
1624: $w^0_*$ is surjective, then we show that as in \cite{MW3},
1625: $w^0_* \circ w^1_* =0$. This, together with the specific properties of
1626: the maps $w^1_*$ and $w^0_*$ will be sufficient to prove the exactness
1627: in the middle term of the sequence:
1628: \[ 0\to C_*(Y_1, \s\otimes L_0)\stackrel{w^1_*}{\longrightarrow}
1629: C_*(Y,\mu, \s\otimes L_0)\stackrel{w^0_*}{\longrightarrow}
1630: \oplus_{k\in \Z} C_{(*)}(Y_0,\s\otimes L_{nk+p}) \to 0. \]
1631:
1632: We can partition the moduli space
1633: $\M_{Y, \mu} (\s\otimes L_0)$ according to the
1634: decomposition (\ref{split-mp}) of Theorem \ref{refine:decomposition},
1635: \[
1636: \M^*_{Y, \mu} (\s\otimes L_0) \cong
1637: \M^*_{Y_1} (\s\otimes L_0) \cup \bigcup_{k\in \Z}
1638: \M_{Y_0} (\s\otimes L_{nk+p}),
1639: \]
1640: where $\mu$ represents the surgery perturbation determined by $f_0'$
1641: as in the proof of Theorem \ref{refine:decomposition}. Here
1642: $f_0'$ depends on a small parameter $\epsilon >0$. That is,
1643: we identify $\M^*_{Y, \mu} (\s\otimes L_0) $ with
1644: a collection of points
1645: \ba
1646: {\cal M}_{Y,\mu}^* (\s\otimes L_0)
1647: = \{ a^{(1)}_i(\epsilon) \}_{i=1,\ldots r} \cup \{
1648: a^{(0)}_j(\epsilon) \}_{j=r+1,\ldots, s },
1649: \label{partition}
1650: \na
1651: so that, as we let $\epsilon \to 0$, the points
1652: $\{ a^{(1)}_i(\epsilon) \}_{i=1,\ldots r} $
1653: get identified with the corresponding elements
1654: $\{ a^{(1)}_i\}_{i=1,\ldots r}$
1655: in $ \M^*_{Y_1} (\s\otimes L_0)$ and, similarly, the points
1656: $\{ a^{(0)}_j(\epsilon) \}_{j=r+1,\ldots, s }$
1657: get identified with the corresponding elements
1658: $\{ a^{(0)}_j \}_{j=r+1,\ldots, s }$
1659: in $ \cup_{k\in \Z} \M_{Y_0} (\s\otimes L_{nk+p})$.
1660:
1661: \begin{Lem}\label{inj:surj} The coefficients of the maps $w_*^1$ and
1662: $w_*^0$ satisfy
1663: \begin{enumerate}
1664: \item $ \la a^{(1)}_i(\epsilon), w_*^1(a^{(1)}_j \ra =\delta_{ij}$;
1665: \item $ \la a^{(0)}_i, w_*^0(a^{(0)}_j(\epsilon) \ra
1666: =\delta_{ij}$. \end{enumerate}\noindent
1667: Thus, $w_*^1$ is injective and $w_*^0$
1668: is surjective.
1669: \end{Lem}
1670: \begin{proof} Using the gluing Theorem \ref{final:glue}, we can
1671: describe the moduli space
1672: $\M^{W_1}\bigl(a_i^{(1)}, a_j^{(1)}(\epsilon)\bigr)$ for
1673: two monopoles $a_i^{(1)}, a_j^{(1)}(\epsilon)$ of
1674: relative index $0$. With our convention
1675: on the choice of components for these moduli spaces, we know
1676: that $\M^{W_1}\bigl(a_i^{(1)}, a_j^{(1)}(\epsilon)\bigr)$
1677: is zero-dimensional, if non-empty, and obtained
1678: as the gluing of the admissible geometric limits
1679: on $V\times \R$ and the corresponding holomorphic
1680: triangles.
1681:
1682: As we let $\epsilon \to 0$, it is easy to see that the holomorphic
1683: triangles degenerate to certain holomorphic discs, and
1684: the admissible geometric limits
1685: on $V\times \R$ for $\M^{W_1}\bigl(a_i^{(1)}, a_j^{(1)}(\epsilon)\bigr)$
1686: are identified with the admissible geometric limits
1687: on $V\times \R$ for $\M_{Y_1\times \R}(a_i^{(1)}, a_j^{(1)})$.
1688: Since the relative index of $a_i^{(1)}, a_j^{(1)}$ is
1689: zero, $\M_{Y_1\times \R}(a_i^{(1)}, a_j^{(1)})$,
1690: being a zero-dimensional moduli space of minimal energy,
1691: is empty unless $a_i^{(1)}=a_j^{(1)}$, in which case
1692: $\M_{Y_1\times \R}(a_i^{(1)}, a_i^{(1)})$ consists of a unique
1693: solution. This proves that
1694: \[
1695: \la a^{(1)}_i (\epsilon), w_*^1(a^{(1)}_j\ra =\delta_{ij}.
1696: \]
1697: Similarly, we obtain
1698: $\la a^{(0)}_i, w_*^0(a^{(1)}_j(\epsilon) \ra =\delta_{ij}.$\end{proof}
1699:
1700: Now we prove the exactness in the middle term. We proceed
1701: as in \cite{MW3} to show that $w_*^0\circ w_*^1=0$, which, together
1702: with Lemma \ref{inj:surj} is
1703: sufficient to establish the exact triangle. Again, we
1704: will use heavily the gluing theorem \ref{final:glue}
1705: to analyze the moduli spaces on the
1706: cobordisms. The following Lemma is the direct
1707: consequence of the results of Lemma \ref{chain:homo} and Lemma
1708: \ref{inj:surj}.
1709:
1710: \begin{Lem}
1711: Suppose given $a^{(1)}_i\in \M_{Y_1}(\s\otimes L_0)$ and
1712: $a^{(0)}_j(\epsilon) \in \M_{Y, \mu}(\s\otimes L_0)$ which
1713: corresponds to $a^{(0)}_j \in \cup_{k\in \Z} \M_{Y_0}(\s\otimes L_{nk+p})$.
1714: The coefficients of
1715: the composition map $w_*^0\circ w_*^1$ satisfy
1716: \[
1717: \la a^{(0)}_j, w_*^0\circ w_*^1 (a^{(1)}_i) \ra
1718: = \la a^{(0)}_j(\epsilon), w_*^1 (a^{(1)}_i) \ra
1719: + \la a^{(0)}_j, w_*^0 (a^{(1)}_i(\epsilon)) \ra.\]
1720: \end{Lem}
1721:
1722: Since the coefficient $\la a^{(0)}_j(\epsilon), w_*^1 (a^{(1)}_i) \ra$
1723: is given by the counting of monopoles
1724: in $\M^{W_1} (a^{(1)}_i, a^{(0)}_j(\epsilon) )$ with the orientation,
1725: and $\la a^{(0)}_j, w_*^0 (a^{(1)}_i(\epsilon)) \ra$ is given by the counting
1726: of monopoles
1727: in $\M^{W_0} (a^{(1)}_i(\epsilon), a^{(0)}_j )$ with the orientation,
1728: the gluing Theorem \ref{final:glue}, for the admissible geometric limits
1729: on $V\times \R$ and the corresponding holomorphic
1730: triangles, yields the following Lemma (cf. Theorem 6.9 \cite{MW3}).
1731:
1732: \begin{Lem}
1733: \label{W_1:W_0}
1734: For small enough $\epsilon$ and large $r\ge r_0$, there
1735: is an orientation reversing diffeomorphism
1736: \[
1737: \M^{W_1(r)}\bigl(a_i^{(1)}, a_j^{(0)}(\epsilon) \bigr)
1738: \cong \M^{W_0(r)} \bigl(a_i^{(1)}(\epsilon), a_j^{(0)}\bigr).
1739: \]
1740: Hence, we have $ w_*^0\circ w_*^1 =0$.
1741: \end{Lem}
1742:
1743: The Lemma corresponds to the fact that, in the two cases, the same
1744: triangles are counted with the reverse orientation.
1745:
1746: With these Lemmata at hand, the arguments in Section 6.5 \cite{MW3}
1747: yield the following exact triangle for any $p \in \{0, \cdots, n-1\}$:
1748: {\small
1749: \ba
1750: \diagram
1751: HF^{SW}_{*} (Y_1, \s\otimes L_0) \rto^{w^1_{*}}
1752: & HF^{SW}_{*}(Y, \s\otimes L_0)
1753: \dlto_{w^0_{*}} \\
1754: \bigoplus_{k\in \Z} HF^{SW}_{(*)}(Y_0, \s\otimes L_{nk+p})
1755: \uto_{\Delta_{(*)}}
1756: & \\
1757: \enddiagram
1758: \label{triangle}
1759: \na }
1760:
1761: After a possible shift of relative index, we obtain the general
1762: exact triangle for any $m, p \in \{0, \cdots, n-1\}$:
1763: {\small
1764: \[
1765: \diagram
1766: HF^{SW}_{*} (Y_1, \s\otimes L_m) \rto^{w^1_{*}}
1767: & HF^{SW}_{*}(Y, \s\otimes L_m)
1768: \dlto_{w^0_{*}} \\
1769: \bigoplus_{k\in \Z} HF^{SW}_{(*)}(Y_0, \s\otimes L_{nk+p})
1770: \uto_{\Delta_{(*)}}
1771: & \\
1772: \enddiagram
1773: \] }
1774: This completes the proof of our main result, Theorem \ref{main:theorem},
1775: except for the claim that the connecting homomorphisms
1776: $\Delta_{(*)}$ are induced by the surgery cobordism
1777: connecting $Y_0$ to $Y_1$. We shall discuss this statement in the next
1778: subsection.
1779:
1780:
1781: \subsection{The surgery triangle}
1782:
1783: To give a precise description of the
1784: connecting homomorphism $\Delta_{(*)}$,
1785: we need to study the discrepancy between
1786: the boundary operator
1787: $\partial_Y$ of the Floer complex $C_*(Y,\s\otimes L_0, \mu)$
1788: and the operator
1789: $\partial_{Y_1}\oplus \bigoplus_{k\in \Z}
1790: \partial_{Y_0,k}$ on
1791: $$ C_*(Y_1, \s\otimes L_0) \oplus
1792: \bigoplus_{k\in \Z} C_{(*)}(Y_0,\s\otimes L_{nk+p}). $$
1793: Let us identify again the points of $\M_{Y,\mu}^*(\s\otimes L_0)$
1794: as in (\ref{partition}) with
1795: $$ {\cal M}_{Y,\mu}^* (\s\otimes L_0)
1796: = \{ a^{(1)}_i(\epsilon) \}_{i=1,\ldots r} \cup \{
1797: a^{(0)}_j(\epsilon) \}_{j=r+1,\ldots, s }.$$
1798: Then we have the following Lemma which
1799: gives the connecting homomorphism $\Delta_{(*)}$.
1800:
1801: \begin{Lem}
1802: Suppose given a cycle in $\sum_i x_i a^{(0)}_i$ in
1803: $C_{(*)}(Y_0,\s\otimes L_{nk+p})$.
1804: The image of $\sum_i x_i a^{(0)}_i$ under the
1805: connecting homomorphism $\Delta$ is given by
1806: $$ \Delta(\sum_i x_i a^{(0)}_i)= \sum_{i, j} x_i
1807: \#\bigl(\M_{Y\times \R} (a_i^{(0)}(\epsilon), a_j^{(1)}(\epsilon))\bigr)
1808: a_j^{(1)}.$$
1809: \end{Lem}
1810: \begin{proof} (Lemma 7.1, Lemma 7.2 in \cite{MW3})
1811: Using the diagram chasing, Lemma \ref{inj:surj} and Lemma
1812: \ref{W_1:W_0}, we see that we have
1813: \[
1814: \begin{array}{lll}
1815: && \partial_{Y,\mu}(\sum_i x_i
1816: a^{(0)}_i(\epsilon)) \\[2mm]
1817: &=& \sum_{i,j} x_i
1818: \#\bigl(\M_{Y\times\R}(a_i^{(0)}(\epsilon),a_j^{(1)}(\epsilon))\bigr)
1819: a_j^{(1)}(\epsilon) \\[2mm]
1820: &&- \sum_{i,j,k} x_i
1821: \#\bigl(\M_{Y\times\R}(a_i^{(0)}(\epsilon),a_j^{(1)}(\epsilon))\bigr)
1822: \#\bigl(\M^{W_0}(a_j^{(1)}(\epsilon), a_k^{(0)}))\bigr)
1823: a_k^{(0)}(\epsilon)\\[2mm]
1824: &=& \sum_{i,j} x_i
1825: \#\bigl(\M_{Y\times\R}(a_i^{(0)}(\epsilon),a_j^{(1)}(\epsilon))\bigr)
1826: a_j^{(1)}(\epsilon) \\[2mm]
1827: &&+ \sum_{i,j,k} x_i
1828: \#\bigl(\M_{Y\times\R}(a_i^{(0)}(\epsilon),a_j^{(1)}(\epsilon))\bigr)
1829: \#\bigl(\M^{W_1}(a_j^{(1)}, a_k^{(0)})(\epsilon))\bigr)
1830: a_k^{(0)}(\epsilon).
1831: \end{array}
1832: \]
1833: By comparing this expression with
1834: $$
1835: \begin{array}{lll}
1836: &&w_*^1( \sum_i x_i \#\bigl( \M_{Y\times \R} (a_i^{(0)}(\epsilon),
1837: a_j^{(1)}(\epsilon))\bigr) a_j^{(1)}) \\[2mm]
1838: &=&\sum_j x_j \#\bigl( \M_{Y\times \R} (a_j^{(0)}(\epsilon),
1839: a_i^{(1)}(\epsilon))\bigr) a_i^{(1)}(\epsilon) \\[2mm]
1840: && +\sum_{i, j, k} x_i \#\bigl(\M_{Y\times \R}(a_i^{(0)}(\epsilon),
1841: a_j^{(1)}(\epsilon))\bigr) \#\bigl(\M^{W_1}(a_j^{(1)}, a_k^{(0)}(\epsilon))
1842: a_k^{(0)}(\epsilon)\bigr),
1843: \end{array} $$
1844: we obtain
1845: $$ \Delta(\sum_i x_i a^{(0)}_i)= \sum_{i, j} x_i
1846: \#\bigl(\M_{Y\times \R} (a_i^{(0)}(\epsilon), a_j^{(1)}(\epsilon)\bigr)
1847: a_j^{(1)}.$$
1848: This completes the proof.
1849: \end{proof}
1850:
1851: In next proposition, we show that the connecting homomorphism
1852: in the exact sequence can also be described as a map $w_{(*)}$,
1853: induced by a surgery cobordism $\bar W_2$ connecting
1854: $Y_0$ to $Y_1$, which satisfies $W_1 \#_Y W_0 = W_2 \# \bar{ \C\P^2}$.
1855: The resulting diagram
1856: $$ C_*(Y_1, \s\otimes L_0) \stackrel{w_*^1}{\to} C_*(Y,\s\otimes L_0,\mu)
1857: \stackrel{w^0_*}{\to} \oplus_k
1858: C_{(*)}(Y_0,\s\otimes L_{nk+p}) \stackrel{w_{(*)}}{\to} C_*(Y_1,
1859: \s\otimes L_0)[-1] $$
1860: is therefore a distinguished triangle, the surgery triangle.
1861:
1862: \begin{Pro} (Proposition 7.3 \cite{MW3})
1863: The connecting homomorphism $\Delta_{(*)}$
1864: in the exact triangle is given by
1865: the following expression,
1866: $$ \Delta_{(*)}(\sum_i x_i a^{(0)}_i)=
1867: w_{(*)}(\sum_i x_i a^{(0)}_i),
1868: $$
1869: for any cycle $\sum_i x_i a^{(0)}_i$ in
1870: $\oplus_k C_{(*)}(Y_0,\s\otimes L_{nk+p})$ for any
1871: fixed $p \in \{0, \cdots, n-1\},$ where
1872: $$ w_{(*)} : \oplus_k C_{(*)}(Y_0,\s\otimes L_{nk+p})
1873: \to C_*(Y_1, \s\otimes L_0)[-1]$$
1874: is the homomorphism defined by counting solutions in the
1875: zero-dimensional components of the moduli spaces
1876: $$ \M^{\bar W_2}(a_j^{(0)},a_i^{(1)}), $$
1877: over the cobordism $\bar W_2$.
1878: \end{Pro}
1879: \begin{proof}
1880: The argument is exactly the same as in the proof of Proposition
1881: 7.3 in \cite{MW3}, therefore we shall omit the proof here.
1882: \end{proof}
1883:
1884:
1885:
1886:
1887:
1888:
1889: \section{Seiberg--Witten and Casson--Walker invariant}
1890:
1891: In this section, we derive the relation between the
1892: topologically invariant version of the Seiberg-Witten invariant and
1893: the Casson-Walker invariant for rational homology 3-spheres. Together
1894: with the equivalence between the Casson-Walker invariant
1895: and the theta invariant introduced by Ozsv\'ath and Szab\'o in
1896: \cite{OS}, our result proves their
1897: conjecture relating the Seiberg-Witten invariant and their
1898: theta invariant.
1899:
1900: Let Y be a rational homology 3-sphere with a smoothly embedded knot $K$
1901: representing a torsion element of order $n$ in $H_1(Y, \Z)$,
1902: \[
1903: \displaystyle{
1904: \frac {|H_1(Y, \Z)|}{|\hbox{Torsion}( H_1(Y-\nu(K), \Z))|}}
1905: =n,
1906: \]
1907: and endowed with the canonical framing $(m, l)$ in
1908: a fixed identification:
1909: $\nu (K)\cong D^2 \times S^1$. Let $p$ and $q$ be relatively prime
1910: integers. The Dehn surgery with coefficient $p/q \in \Q \cup\{\infty\}$
1911: on $K$ gives rise to another closed manifold $Y_{p/q}$.
1912:
1913: Denote by $\spinc (V)$ the set of equivalence classes of $\spinc$
1914: structures on $V= Y\backslash \nu (K)$ with trivial restriction
1915: to the boundary $T^2$. Then, for any $Y_{p/q}$, there
1916: is a surjective map:
1917: \[
1918: \iota_{Y_{p/q}}: \qquad \spinc (Y_{p/q}) \to \spinc (V),
1919: \]
1920: where, for any $\s\in \spinc (Y_{p/q})$, $\iota_{Y_{p/q}} (\s)$
1921: is given by the restriction to $V\subset Y_{p/q}$. The fiber of
1922: $\iota_{Y_{p/q}}$ is given by a cyclic group
1923: generated by the Poincar\'e dual of the core
1924: of $Y_{p/q}\backslash V$.
1925: Formally, for $\iota_{Y}(\s)$ with $\s \in \spinc (Y)$, we identify the fiber
1926: of $\iota_{Y}$ with
1927: the following set of $\spinc$ structures
1928: \[
1929: \iota_{Y}^{-1}(\iota_{Y}(\s)=\bigcup_{m=0, \cdots, n-1}
1930: \{\s\otimes L_m| c_1(L_m) = mPD([K]) \in H^2(Y, \Z) \}.
1931: \]
1932: Similarly the fiber of $\iota_{Y_{p/q}}$ is given by
1933: \[
1934: \bigcup_{m=0, \cdots, np-1} \{\s\otimes L_m| c_1(L_m)
1935: = mPD([K]) \in H^2(Y_{p/q}, \Z) \},
1936: \]
1937: and the fiber of $\iota_{Y_{0}}$ is given by
1938: \[
1939: \bigcup_{ m\in \Z}\{\s\otimes L_m| c_1(L_m)
1940: = mPD([K]) \in H^2(Y_0, \Z) \}.
1941: \]
1942: Here we use the same notation $\s$ on $Y_{p/q}$ ($Y$, or $Y_0$)
1943: as the corresponding $\spinc$ structure obtained by gluing
1944: $\s \in \spinc (V)$ with the trivial $\spinc$ structure on $\nu(K)$
1945: by the trivial gauge transformation on $T^2$. We hope this
1946: notation will not cause any confusion.
1947:
1948: Assume that $V$ and $\nu(K)$
1949: are equipped with a metric with a cylindrical end modeled on $T^2$.
1950: Let $\s$ be a $\spinc$ structure on $Y$. By the
1951: result of \cite{CMW} on the moduli space of finite energy monopoles on
1952: $(V, \iota_Y (\s))$,
1953: we know that
1954: the irreducible part, denoted $\M^*_V(\s)$, is a smooth,
1955: oriented 1-dimensional manifold. The
1956: asymptotic values along the cylindrical end define a
1957: continuous map:
1958: \[
1959: \partial_\infty: \qquad \M^*_V(\s) \rightarrow \chi_0 (T^2, V),
1960: \]
1961: where $\chi_0 (T^2, V)$ is a $\Z\times \Z_n$-covering of
1962: the character torus $\chi (T^2)$. Sometime it is convenient
1963: to compose the above asymptotic value map with this
1964: covering map and define a boundary value map:
1965: \ba
1966: \partial_\infty: \qquad \M^*_V(\s) \rightarrow \chi (T^2).
1967: \label{partial:infty}
1968: \na
1969:
1970: Notice that the reducible part $\chi (V)$ of the moduli space on
1971: $(V, \iota_Y (\s))$
1972: is an embedded circle $\chi (V) \subset \chi_0 (T^2, V)$ under the
1973: asymptotic value map. This becomes a circle of
1974: multiplicity $n$ in $\chi (T^2)$. There is a ``bad point''
1975: in $\chi (T^2)$, given by the flat connections such that the
1976: corresponding twisted Dirac operator has a non-trivial kernel.
1977: We can endow $\chi (T^2)$ with a coordinate system $(u, v)$ defined
1978: by the holonomy around the longitude $l$ and the meridian $m$,
1979: respectively, so that the bad point corresponds to $(u, v) =(1,
1980: 1)$. The reducible circle $\chi (V)$, with the holonomy around the
1981: longitude $l$ of order $n$, is given by $u = u(\s)$, with
1982: $u(\s) \in \{0, 2/n, \cdots, 2(n-1)/n\}$. After a suitable perturbation,
1983: and a corresponding shift of coordinates, as discussed in \cite{CMW},
1984: we can assume that the bad point does not lie on any of these $n$
1985: possible circles $u = u(\s)$ of reducibles $\chi (V)$.
1986:
1987: From the result in \cite{CMW}, we know
1988: that, under the map $\partial_\infty$ in (\ref{partial:infty}), the
1989: boundary points $\partial ( \M^*_V(\s))$ are
1990: either mapped to the bad point in $\chi(T^2)$ or mapped
1991: to the reducible circle $u= u(\s)$ on $\chi(T^2)$.
1992:
1993: Let $\chi (\nu(K) \subset Y_{p/q})$ be the reducible circle
1994: on $\nu(K) \subset Y_{p/q}$, which maps to a closed curve on $\chi(T^2)$
1995: with slope $p/q$ in the $(u, v)$-coordinates: parallel to $pv =qu$.
1996: Looking at the induced Spin structure on $T^2 \subset Y_{p/q}$,
1997: we know that the curve $\chi (\nu(K) \subset Y_{p/q})$ goes through
1998: $(0, 1)$ if $q$ is odd or goes through
1999: $(0, 0)$ if $q$ is even, cf.\cite{CMW}. Again, after a suitable
2000: perturbation as in (\ref{eta:pert}), and the corresponding shift of
2001: coordinates, we can assume that this $p/q$-curve is away from the bad
2002: point on $\chi(T^2)$ and does not meet $u= u(\s)$
2003: along the coordinate line $v=0$. Then we know that
2004: $u= u(\s)$ intersects $\chi (\nu(K) \subset Y_{p/q})$
2005: inside $\chi(T^2)$ at $p$ points, which
2006: are denoted by $\theta_1, \cdots, \theta_p$, ordered according the
2007: orientation of $ u= u(\s) \subset \chi (T^2)$. They can be
2008: lifted to $pn$ points in $\chi_0(T^2, V)$. We denote these points by
2009: $\theta_1^{(k)}, \cdots,
2010: \theta_p^{(k)}$, ($k=0, 1, \cdots, n-1$)
2011: according to the order. Denote by $\theta_0$ the intersection point of
2012: $u= u(\s)$ with $v=0$ in $\chi (T^2)$. This
2013: can be lifted to $n$-points $\theta_0^{(0)},
2014: \theta_0^{(1)}, \cdots, \theta_0^{(n-1)}$
2015: on $\chi_0(T^2, V)$.
2016: Moreover, we can assume that the map $\partial_\infty$ in
2017: (\ref{partial:infty})
2018: is transverse to the curves $u= u(\s)$, $v=0$ and
2019: $\chi (\nu(K) \subset Y_{p/q})$, by a suitable perturbation of the
2020: Seiberg-Witten equations on $V$ as in \cite{CMW}. We can also assume
2021: that the image
2022: $\partial_\infty\M^*_V(\s)$ does not meet the points $\theta_0,
2023: \theta_1, \cdots \theta_p$ in $\chi (T^2)$, again
2024: by suitable perturbation, as discussed in \cite{CMW}.
2025:
2026: Let $I$ be any open interval in
2027: \[
2028: \chi (V) = \{ u = u(\s)\} \subset \chi_0(T^2, V).
2029: \]
2030: We denote by $SF_\C (\dirac^V_I)$
2031: the complex spectral flow of Dirac operator
2032: on $V$ twisted with the path of reducible connections $I$ on $V$. From
2033: the analysis in \cite{CMW} and \cite{MW1},
2034: we know that
2035: \ba
2036: \# \bigl( \partial_\infty|_{\partial \M^*_V(\s)} \bigr)^{-1}
2037: (I) = SF_\C (\dirac^V_I).
2038: \label{boundary:on:I}
2039: \na
2040: For convenience, we define
2041: \ba
2042: SF_\C (\dirac^V_{[\theta_i, \theta_j]}) = \sum_{k=0}^{n-1}
2043: SF_\C (\dirac^V_{[\theta_i^{(k)}, \theta_j^{(k)}]}).
2044: \label{spectralflow:sum}
2045: \na
2046:
2047: With this notation understood, we can state the following
2048: proposition relating the Seiberg-Witten invariants on $Y_{p/q}$,
2049: $Y$ and $Y_0$.
2050:
2051: \begin{Pro}\label{p/q-formulae:SW}
2052: Consider generic compatible small perturbations of the Seiberg-Witten
2053: equations
2054: on $Y_{p/q}$, $Y$ and $Y_0$, such that the map $\partial_\infty$ as in
2055: (\ref{partial:infty})
2056: is transverse to the curves $u= u(\s)$, $v=0$ and
2057: $\chi (\nu(K) \subset Y_{p/q})$ and misses the points
2058: $\theta_0, \theta_1, \cdots, \theta_p$ in $\chi (T^2)$. Then
2059: we have the following relation:
2060: \[\begin{array}{lll}
2061: &&\sum_{k=0}^{pn-1} SW_{Y_{p/q}} (\s\otimes L_k, g_{Y_{p/q}})\\[2mm]
2062: &=& p\sum_{k=0}^{n-1} SW_{Y} (\s\otimes L_k, g_{Y}) +
2063: q \sum_{k\in \Z} SW_{Y_0} (\s\otimes L_k)\\[2mm]
2064: && + \sum_{i=1}^p SF_\C (\dirac^V_{[\theta_0, \theta_i]}).
2065: \end{array}
2066: \]
2067: \end{Pro}
2068: \begin{proof} By the gluing theorem for 3-dimensional monopoles as in
2069: \cite{CMW}, we have
2070: \[
2071: \bigcup_{k=0}^{pn-1} \M^*_{Y_{p/q}} (\s\otimes L_k)
2072: = \M^*_V (\s) \times_{\chi (T^2)} \chi (\nu(K) \subset Y_{p/q}),
2073: \]
2074: where $\chi (\nu(K) \subset Y_{p/q})$ is the $p/q$-curve on
2075: $\chi (T^2)$. Thus, we obtain
2076: \ba
2077: \sum_{k=0}^{pn-1} SW_{Y_{p/q}} (\s\otimes L_k, g_{Y_{p/q}})
2078: = \# \bigl( \M^*_V (\s) \times_{\chi (T^2)} \chi (\nu(K) \subset Y_{p/q})
2079: \bigr).
2080: \label{p:q}
2081: \na
2082: Notice that the set
2083: $
2084: \{\theta_1^{(k)}, \cdots, \theta_p^{(k)}: k = 0, 1, \cdots, n-1 \}
2085: $
2086: consists of the unique reducible monopole for each
2087: $(Y_{p/q}, \s\otimes L_k)$.
2088:
2089:
2090: Similarly, we have
2091: \ba
2092: \sum_{k=0}^{n-1} SW_{Y} (\s\otimes L_k, g_{Y})
2093: = \# \bigl( \M^*_V (\s) \times_{\chi (T^2)} \{ v=0\}\bigr).
2094: \label{Y}
2095: \na
2096: Here the reducible set consists of
2097: $
2098: \{ \theta_0^{(0)}, \theta_0^{(1)}, \cdots, \theta_0^{(n-1)} \}.
2099: $
2100:
2101: In order to avoid the circle of reducibles on $(Y_0, \s \otimes L_0)$,
2102: we need to introduce a small perturbation such that
2103: $\chi (\nu(K)\subset Y_0)$ on $\chi (T^2)$ is a small parallel
2104: shifting of $u = u(\s)$ such that the bad point is not
2105: contained in the narrow strip bounded by these
2106: two parallel curves. We denote this small shift
2107: of $u = u(\s)$ by $u = u(\s) +\eta$, where $\eta$ is a sufficiently small
2108: positive number. This can be achieved by a perturbation of the
2109: equations as in \cite{CMW}. Then we have
2110: \ba
2111: \sum_{k\in \Z} SW_{Y_0} (\s\otimes L_k)
2112: = \# \bigl( \M^*_V (\s) \times_{\chi (T^2)} \{ u= u(\s) +\eta\}\bigr).
2113: \label{Y_0}
2114: \na
2115:
2116: In order to compare the three countings in (\ref{p:q}) -- (\ref{Y_0}), we
2117: need to choose an oriented 2-chain $ C$ in $\chi (T^2)$ whose
2118: boundary 1-chain is given by
2119: \[\begin{array}{lll}
2120: &&\chi (\nu(K)\subset Y_{p/q}) - p \chi (\nu(K)\subset Y) -q
2121: \chi (\nu(K)\subset Y_0) \\[2mm]
2122: &=& \chi (\nu(K)\subset Y_{p/q}) - p \{ v= 0\}
2123: - q\{ u= u(\s) +\eta\},
2124: \end{array}
2125: \]
2126: and such that $C$ does not contain the bad point in $\chi (T^2)$. Then,
2127: counting the boundary points of $\partial_\infty^{-1}(C)$, as a
2128: 0-chain, we obtain
2129: \ba
2130: \begin{array}{lll}
2131: &&\# \bigl( \partial_\infty^{-1} (\chi (\nu(K)\subset Y_{p/q}) \bigr)
2132: \\[2mm]
2133: &=& p \# \bigl( \partial_\infty^{-1} (\{ v= 0\}) \bigr)
2134: + q \# \bigl( \partial_\infty^{-1}(\{ u= u(\s) +\eta\}) \bigr)\\[2mm]
2135: && + \# \bigl( \partial_\infty\mid_{\partial (\M^*_V(\s)) }\bigr)^{-1} (C).
2136: \label{count:C}
2137: \end{array}
2138: \na
2139:
2140: As $C$ does not contain the bad points, we know
2141: that the possible points of $\partial_\infty (\partial (\M^*_V(\s)) )
2142: \cap C$ all lie on the curve $u = u(\s)$, away from
2143: the points $\theta_0, \theta_1, \cdots, \theta_p$. It is easy to see
2144: that $C$ covers the intervals of $u = u(\s)$ between two consecutive
2145: points $\theta_i$ with different multiplicities:
2146: the multiplicities are
2147: $ p, p-1, \cdots, 1, 0$,
2148: for the intervals
2149: \[
2150: [\theta_0, \theta_1], [\theta_1, \theta_2],
2151: \cdots, [\theta_{p-1}, \theta_p], [\theta_p, \theta_0],
2152: \]
2153: respectively.
2154: By the identity (\ref{boundary:on:I}) and the definition
2155: (\ref{spectralflow:sum}), we know that
2156: \ba
2157: \# \bigl( \partial_\infty|_{\partial (\M^*_V(\s))} \bigr)^{-1}
2158: (C) = \sum_{i=1}^p SF_\C (\dirac ^V_{[\theta_0, \theta_i]}).
2159: \label{count:u(s)}
2160: \na
2161: Combining all the identities in (\ref{p:q}), (\ref{Y}),
2162: (\ref{Y_0}), (\ref{count:C}) and (\ref{count:u(s)}),
2163: we obtain the proof of the proposition.
2164: \end{proof}
2165:
2166: The Seiberg-Witten invariant for any rational homology 3-sphere
2167: depends on metric and perturbation (cf.\cite{MW1}). We now consider
2168: the correction term (\ref{correction:term}) as defined in the
2169: introduction. We have the following proposition relating the
2170: correction terms for $Y_{p/q}$ and $Y$.
2171:
2172: \begin{Pro}\label{p/q-formulae:correction}
2173: \begin{enumerate}
2174: \item For any rational homology 3-sphere $Y$ with a $\spinc$ structure
2175: $\s$ and a Riemannian metric $g_Y$,
2176: \[
2177: \hat{SW}_Y (\s) = SW_Y (\s, g_Y) - \xi (\s, g_Y)
2178: \]
2179: is a well-defined topological invariant.
2180: \item For any relatively prime integers $p$ and $q$, a positive integer $n$,
2181: and $u\in \{0, 2/n,\ldots, 2(n-1)/n \}$, we have that
2182: \[
2183: \sum_{k=0}^{pn-1} \xi_{Y_{p/q}} (\s\otimes L_k, g_{Y_{p/q}})
2184: - p \sum_{k=0}^{n-1} \xi_{Y} (\s\otimes L_k, g_{Y}) -
2185: \sum_{i=1}^p SF_\C (\dirac ^V_{[\theta_0, \theta_i]}) \]
2186: is independent of the
2187: manifold $Y$ and depends only on $p, q, n$, and
2188: $u(\s) \in \{0, 2/n, \ldots, 2(n-1)/n\}.$
2189: \end{enumerate}
2190: \end{Pro}
2191: \begin{proof} Claim (1) follows from the wall-crossing formulae
2192: in \cite{MW1} and the Atiyah-Patodi-Singer index theorem.
2193: The proof of claim (2) is analogous to the proof of Proposition
2194: 7.9 in \cite{OS}. We adapt their arguments to our situation.
2195: We write the standard surgery cobordism between
2196: $S^3$ and the Lens space $L(p, q)$ as
2197: \[
2198: W(S^3, L(p, q)) = \bigl( [0, 1] \times S^1\times D^2 \bigr)
2199: \cup_{ [0, 1] \times S^1 \times S^1} X_{p/q},
2200: \]
2201: Then the surgery cobordism between $Y$ and $Y_{p/q}$
2202: can be identified as
2203: \[
2204: W_{p/q} = \bigl( [0, 1] \times V\bigr)
2205: \cup_{ [0, 1] \times S^1 \times S^1} X_{p/q}.
2206: \]
2207: We fix a metric on $W_{p/q}$ which respects the product structure
2208: $[0, 1] \times V$ and $[0, 1] \times S^1 \times S^1$,
2209: and agrees with $g_Y$ and $g_{Y_{p/q}}$ on the boundaries
2210: $Y$ and $Y_{p/q}$, respectively.
2211:
2212: For a $\spinc$ structure $\s\otimes L_i^{(m)}$
2213: in $\{\s\otimes L_k: k=0, \cdots, pn-1\}$ on
2214: $Y_{p/q}$, whose reducible monopole corresponds
2215: to $\theta_i^{(m)}$ (with $i\in \{1, \cdots, p\}$ and
2216: $m\in \{ 0, \cdots n-1\}$), we consider the $\spinc$ structure
2217: $\s\otimes L_m$ on $Y$ whose reducible monopole is
2218: $\theta_0^{(m)}$. Then we claim that
2219: \ba\label{xi:m:i}
2220: \xi_{Y_{p/q}} (\s\otimes L_i^{(m)}, g_{Y_{p/q}})
2221: - \xi_{Y} (\s\otimes L_m, g_Y) -SF_\C
2222: (\dirac^V_{[\theta_0^{(m)}, \theta_i^{(m)}]})\na
2223: is independent of $Y$ and depends only on $p, q, n$ and on
2224: $u(\s) \in \{0, 2/n, \cdots, 2(n-1)/n\}.$
2225:
2226: To prove this claim, we choose a $\spinc$ structure $\tilde \s$
2227: on $W_{p/q}$ whose restriction to $Y$ and $Y_{p/q}$ is given by
2228: $\s\otimes L_m$ and $\s\otimes L_i^{(m)}$, respectively, and such that
2229: $c_1(\tilde\s)^2 =1$. On $(W_{p/q}, \tilde \s)$, we choose
2230: a connection $A$, whose restriction to
2231: $V\times [0, 1]$ is the path of reducibles connecting
2232: $\theta_0^{(m)}$ to $\theta_i^{(m)}$ along the curve
2233: $\chi (V) \subset \chi_0(T^2, V)$. Then we have
2234: \ba\begin{array}{lll}
2235: && \xi_{Y_{p/q}} (\s\otimes L_i^{(m)}, g_{Y_{p/q}})
2236: - \xi_{Y} (\s\otimes L_m, g_Y) \\[2mm]
2237: &=& Ind_\C (\Dirac_A^{W_{p/q}}) - \displaystyle {
2238: \bigl( \frac {c_1(\tilde \s)^2 -\sigma (W_{p/q}) }{8}\bigr)}\\[2mm]
2239: &=& Ind_\C (\Dirac_A^{W_{p/q}})\\[2mm]
2240: &=& Ind_\C (\Dirac_A^{[0,1]\times V}) + Ind_\C (\Dirac_A^{X_{p/q}})
2241: \end{array}\label{splitting}
2242: \na
2243: where the third equality follows from the
2244: splitting principle for the index, as the
2245: Dirac operator has no kernel on the various boundaries and
2246: corners \cite{CLM}. Notice that we have
2247: \[
2248: Ind_\C (\Dirac_A^{[0,1]\times V}) =
2249: SF_\C (\dirac^V_{[\theta_0^{(m)}, \theta_i^{(m)}]}),
2250: \]
2251: and the connection $A|_{X_{p/q}}$ extends to connection $A_0$
2252: on $W(S^3, L(p, q))$ by a flat connection, whose index on
2253: $[0,1] \times S^1 \times D^2$ satisfies
2254: \[
2255: Ind_\C (\Dirac_{A_0}^{[0,1] \times S^1 \times D^2} )=0.
2256: \]
2257: In fact, we can choose the metric on $W(S^3, L(p, q))$ with
2258: a positive scalar curvature metric on $[0,1] \times S^1 \times D^2$.
2259: Therefore, we have
2260: \[
2261: Ind_\C (\Dirac_A^{X_{p/q}}) =
2262: Ind_\C (\Dirac_{A_0}^{W(S^3, L(p, q))}),
2263: \]
2264: which depends only on $p, q, n$ and $u(\s)$, and so does the quantity
2265: \ba
2266: \label{splitting2}
2267: \xi_{Y_{p/q}} (\s\otimes L_i^{(m)}, g_{Y_{p/q}})
2268: - \xi_{Y} (\s\otimes L_m, g_Y)-
2269: SF_\C (\dirac^V_{[\theta_0^{(m)}, \theta_i^{(m)}]}).
2270: \na
2271: When summing the identity (\ref{splitting2})
2272: over $i\in \{1, \cdots, p\}$ and $m\in \{0, \cdots, n-1\}$,
2273: notice that the term $\xi_{Y} (\s\otimes L_m, g_Y)$ is independent of
2274: $i\in \{1, \cdots, p\}$,
2275: hence we obtain the proof of the claim (2) by using the definition
2276: (\ref{spectralflow:sum}).
2277: \end{proof}
2278:
2279: With these two propositions in place, we now have the following surgery
2280: formula for the modified version of the Seiberg-Witten invariant.
2281:
2282: \begin{The}\label{p/q-formulae:hatSW}
2283: Given any two relatively prime integers $p$ and $q$, a positive
2284: integer $n$ and $u\in \{0, 2/n, 2(n-1)/n\}$, there is a rational valued
2285: function $s(p, q, n, u)$, depending only on $p, q, n$ and $u$,
2286: satisfying the
2287: following property. Let $Y$ be a rational homology 3-sphere with a smoothly
2288: embedded knot and a canonical framing $(m, l)$
2289: such that $\nu (K)\cong D^2\times S^1$. Assume that
2290: $K$ represents a torsion element of order $n$ in $H_1(Y, \Z)$. Let
2291: $\s$ be a $\spinc$ structure on $Y$. Then we have
2292: \[
2293: \begin{array}{lll}
2294: &&\sum_{k=0}^{pn-1} \hat{SW}_{Y_{p/q}} (\s\otimes L_k)\\[2mm]
2295: &=& p\sum_{k=0}^{n-1} \hat{SW}_{Y} (\s\otimes L_k) +
2296: q \sum_{k\in \Z} SW_{Y_0} (\s\otimes L_k)\\[2mm]
2297: && + s(p, q, n, u). \end{array}
2298: \]
2299: \end{The}
2300: \begin{proof} Following from Proposition \ref{p/q-formulae:correction},
2301: we know that
2302: \ba
2303: \sum_{k=0}^{pn-1} \xi_{Y_{p/q}} (\s\otimes L_k, g_{Y_{p/q}})
2304: -p \sum_{k=0}^{n-1} \xi_{Y} (\s\otimes L_k, g_{Y}) -
2305: \sum_{i=1}^p SF_\C (\dirac ^V_{[\theta_0, \theta_i]})
2306: \label{s(p,q,n,u)}
2307: \na
2308: depends only on $p, q, n$ and
2309: $u=u(\s) \in \{0, 2/n, \cdots, 2(n-1)/n\}$. We denote this term by
2310: $s(p, q, n, u)$. By subtracting (\ref{s(p,q,n,u)}) from the surgery
2311: formula for the Seiberg-Witten invariants in Proposition
2312: \ref{p/q-formulae:SW}, we obtain the proof of this theorem.
2313: \end{proof}
2314:
2315: Now we can establish the equivalence between
2316: the modified version of the Seiberg-Witten invariant $\hat{SW}$
2317: and the Casson-Walker invariant for rational homology 3-spheres.
2318:
2319: \begin{The}
2320: For any rational homology 3-sphere $Y$, we have
2321: \[
2322: \sum_{\s\in \spinc (Y)} \hat {SW}_Y (\s)
2323: = \displaystyle{\frac 12} |H_1(Y, \Z)| \lambda (Y)
2324: \]
2325: where $\lambda (Y)$ is the Casson-Walker invariant.
2326: \end{The}
2327: \begin{proof}
2328: We first derive the surgery formula for the invariant
2329: $\sum_{\s\in \spinc (Y)} \hat {SW}_Y (\s)$ from
2330: Theorem \ref{p/q-formulae:hatSW} and the Seiberg-Witten invariant for
2331: $Y_0$ (a rational homology $S^1\times S^2$, i.e., $b_1(Y_0)=1$) (see
2332: \cite{MT} \cite{Don}):
2333: \ba\label{surgery:formuale:SW}
2334: \begin{array}{lll}
2335: &&\sum_{\s\in \spinc (Y_{p/q})}
2336: \hat{SW}_{Y_{p/q}} (\s)\\[2mm]
2337: &=& p\sum_{\s\in \spinc (Y)} \hat{SW}_{Y} (\s) +
2338: q \sum_{j=0}^\infty a_j j^2 + |H_1(Y, \Z)| s(p, q, n) \end{array}
2339: \na
2340: where $s(p, q, n)= \sum_u s(p, q, n, u)/n$ and $a_j$ is the coefficient
2341: of the symmetrized Alexander polynomial of $Y_0$,
2342: \[
2343: A(t) = |\hbox{Torsion} (H_1(Y_0, \Z))| + \sum_{j=1}^\infty a_j (t^j + t^{-j})
2344: \]
2345: normalized such that
2346: \[
2347: A(1) = |\hbox{Torsion} (H_1(Y_0, \Z))|.
2348: \]
2349: Set $\bar \lambda (Y) = \displaystyle{\frac 12}|H_1(Y, \Z)| \lambda (Y)$
2350: as the normalized Casson-Walker invariant. Then
2351: the surgery formula in \cite{Walker}
2352: for $\bar \lambda (Y)$ can be expressed as (cf. \cite{OS}):
2353: \ba\label{surgery:formuale:CW}
2354: \begin{array}{lll}
2355: \bar \lambda (Y_{p/q}) &=& p \bar \lambda (Y) + q \sum_{j=0}^\infty a_j j^2
2356: \\[2mm]
2357: && + |H_1(Y, \Z)| \bigl( \displaystyle{
2358: \frac{q(n^2-1)}{12n^2} -\frac{p s(p, q)}{2}}\bigr).
2359: \end{array}\na
2360: Here $s(p, q)$ is the Dedekind sum of relatively prime integers $p$
2361: and $q$ (cf. \cite{Walker}).
2362: Comparing (\ref{surgery:formuale:SW}) and (\ref{surgery:formuale:CW}),
2363: we only need to show that
2364: \ba
2365: s(p, q, n) = \displaystyle{
2366: \frac{q(n^2-1)}{12n^2} -\frac{p s(p, q)}{2}}.
2367: \label{s(p,q,n)}
2368: \na
2369:
2370: Since $s(p,q,n)$ is independent of the manifold $Y$, we can choose
2371: some examples that can be computed explicitly, and use them to
2372: identify the coefficient $s(p,q,n)$.
2373: The Lens space $L(p, q)$ can be obtained by a $p/q$-surgery
2374: on an unknot in $S^3$. The calculation of Nicolaescu \cite{Nic}
2375: for $L(p, q)$ gives us that
2376: \[
2377: \sum_{\s\in \spinc (L(p, q))} \hat{SW}_{L(p, q)}(\s) =
2378: -\displaystyle{
2379: \frac{ps(p, q)}{2}}.
2380: \]
2381: This implies that (\ref{s(p,q,n)}) holds for $n=1$. Now
2382: we can prove (\ref{s(p,q,n)}) by induction on $n$.
2383: This is exactly the same argument as in the proof of Theorem 7.5 in
2384: \cite{OS} on the equivalence of their theta invariant and the
2385: Casson-Walker invariant. The example is the Seifert manifold
2386: $M(n, 1; -n, 1; q, -p)$, obtained by $p/q$ surgery on a knot of order $n$ in
2387: $L(n, 1) \# \overline{L(n, 1)}$. By Kirby calculus it is possible to show
2388: that $M(n, 1; -n, 1; q, -p)$ can be obtained as
2389: $(-n)$-surgery on a knot in the Lens space $L(pn-q, q)$, and
2390: can be obtained as a sequence of surgeries on knots of order less than $n$,
2391: see the proof of Theorem 7.5 in \cite{OS} for details.
2392: \end{proof}
2393:
2394:
2395: \begin{thebibliography}{999999}
2396:
2397: \bibitem{CLM} S. Cappell, R. Lee, E. Miller, {\em
2398: Self-adjoint elliptic operators and manifold decompositions,
2399: part II: Spectral flow and Maslov index.} Comm. Pure Appl. Math.,
2400: 49:869-909, 1996.
2401:
2402: \bibitem{CMW} A. Carey, M. Marcolli, B.L. Wang, {\em Exact triangles
2403: in Seiberg-Witten Floer theory. Part I: the geometric triangle},
2404: preprint, math.DG/9907065.
2405:
2406: \bibitem{Don} S. Donaldson, {\em Topological field theories and formulae
2407: of Casson and Meng-Taubes, } Proceedings of the Kirbyfest
2408: (Berkeley, CA, 1998), 87--102 (electronic), Geom.
2409: Topol. Monogr., 2, Geom. Topol., Coventry, 1999
2410:
2411:
2412: \bibitem{MW1} M. Marcolli, B.L. Wang, {\em Equivariant
2413: Seiberg-Witten Floer theory}, preprint, dg-ga/9606003.
2414:
2415: \bibitem{MW2} M. Marcolli, B.L. Wang, {\em Exact triangles
2416: in Seiberg-Witten Floer theory. Part II: geometric limits of flow
2417: lines}, preprint, math.DG/9907080.
2418:
2419: \bibitem{MW3} M. Marcolli, B.L. Wang, {\em Exact triangles
2420: in Seiberg-Witten Floer theory. Part III: proof of exactness},
2421: preprint, math.DG/0009157.
2422:
2423: \bibitem{MW4} M. Marcolli, B.L. Wang, {\em Exact triangles
2424: in Seiberg-Witten Floer theory. Part IV: $\Z$-graded monopole
2425: homology}, preprint, math.DG/0009159.
2426:
2427: \bibitem{MT} G. Meng, C. Taubes, {\em $\underline{\rm SW}=$ Milnor torsion},
2428: Math. Res. Lett. 3 (1996), no. 5, 661--674.
2429:
2430: \bibitem{Nic} L. Nicolaescu, {\em Seiberg-Witten theoretic invariants
2431: of lens spaces}, preprint, math.DG/9901071.
2432:
2433: \bibitem{OS} P. Ozsv\'ath, Z. Szab\'o, {\em The theta divisor and
2434: the Casson-Walker invariant}, preprint, math.DG/0006194.
2435:
2436: \bibitem{Walker} K. Walker, {\em An extension of Casson's invariant},
2437: Ann. Math. Studies, Vol. 126, Princeton Univ. Press, 1992.
2438:
2439:
2440: \end{thebibliography}
2441:
2442:
2443: \noindent {\bf Matilde Marcolli}\par
2444: \noindent Max--Planck--Institut f\"ur Mathematik, Vivatsgasse 7,
2445: D-53111 Bonn, Germany. \par
2446: \noindent marcolli\@@mpim-bonn.mpg.de
2447:
2448: \vskip .2in
2449:
2450: \noindent {\bf Bai-Ling Wang}\par
2451: \noindent Department of Pure
2452: Mathematics, University of Adelaide, Adelaide SA 5005, Australia. \par
2453: \noindent bwang\@@maths.adelaide.edu.au
2454:
2455: \end{document}
2456:
2457:
2458:
2459: