1: \documentclass[letterpaper,11pt]{amsart}
2:
3: \usepackage{amssymb}
4: \usepackage{eucal}
5: \usepackage{array}
6: \usepackage{epsfig}
7: \usepackage[all]{xy}
8: \CompileMatrices \numberwithin{equation}{section}
9: \SelectTips{cm}{}
10:
11: \title{More about vanishing cycles and mutation}
12: \author{Paul Seidel}
13: \date{October 5, 2000.}
14:
15: %
16: %------------------------ NON-MATH MACROS --------------------------
17: %
18: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
19:
20: \renewcommand{\subsection}{\vspace{1ex}
21: \hspace{-\parindent}%
22: \stepcounter{subsection}%
23: ({\sc \thesection\alph{subsection}})}
24:
25: \setlength{\extrarowheight}{4pt}
26:
27: \newcommand{\includefigure}[3]{%
28: \begin{figure}[#3]
29: \begin{center}
30: \epsfig{file=#2} \\ \caption{\label{fig:#1}}
31: \end{center}
32: \end{figure}}
33:
34: \newenvironment{myitemize}%
35: {\begin{itemize}\itemsep0.5em} {\end{itemize}}
36:
37: %
38: %----------------------- GENERIC MATH MACROS -----------------------
39: %
40: \newcommand{\R}{\mathbb{R}}
41: \newcommand{\Z}{\mathbb{Z}}
42: \newcommand{\C}{\mathbb{C}}
43: \newcommand{\Q}{\mathbb{Q}}
44: \newcommand{\N}{\mathbb{N}}
45: \newcommand{\iso}{\cong} %isomorphism sign
46: \newcommand{\htp}{\simeq} %homotopy sign
47: \newcommand{\smooth}{C^\infty}
48: \newcommand{\CP}[1]{\C {\mathrm P}^{#1}}
49: \newcommand{\RP}[1]{\R {\mathrm P}^{#1}}
50: \newcommand{\leftsc}{\langle}
51: \newcommand{\rightsc}{\rangle}
52: \newcommand{\Rgeq}{\R^{\scriptscriptstyle \geq 0}}
53: \newcommand{\Rleq}{\R^{\scriptscriptstyle \leq 0}}
54: \newcommand{\suchthat}{\; | \;}
55:
56: \newcommand{\id}{\mathrm{id}}
57: \newcommand{\ind}{\mathrm{ind}}
58: \newcommand{\re}{\mathrm{re}}
59: \newcommand{\im}{\mathrm{im}}
60: \renewcommand{\ker}{\mathrm{ker}}
61: \newcommand{\coker}{\mathrm{coker}}
62: \newcommand{\mymod}{\quad\text{mod }}
63: \newcommand{\Hom}{\mathrm{Hom}}
64: \newcommand{\End}{\mathrm{End}}
65:
66: \newcommand{\mo}{(M,\omega)}
67: \renewcommand{\o}{\omega}
68: \renewcommand{\O}{\Omega}
69: \newcommand{\Aut}{\mathrm{Aut}}
70: \newcommand{\Ham}{\mathrm{Ham}}
71: \newcommand{\Symp}{\mathrm{Symp}}
72: \newcommand{\Diff}{\mathrm{Diff}}
73:
74: \theoremstyle{plain}
75: \newtheorem{thm}{Theorem}[section]
76: \newtheorem{theorem}[thm]{Theorem}
77: \newtheorem{cor}[thm]{Corollary}
78: \newtheorem{corollary}[thm]{Corollary}
79: \newtheorem{lemma}[thm]{Lemma}
80: \newtheorem{prop}[thm]{Proposition}
81: \newtheorem{proposition}[thm]{Proposition}
82: \newtheorem{conjecture}[thm]{Conjecture}
83: \newtheorem{question}[thm]{Question}
84: \newtheorem{defn}[thm]{Definition}
85: \newtheorem{definition}[thm]{Definition}
86: \newtheorem{remark}[thm]{Remark}
87: \newtheorem{remarks}[thm]{Remarks}
88: \newtheorem{example}[thm]{Example}
89: \newtheorem{examples}[thm]{Examples}
90:
91: \theoremstyle{remark}
92: \newtheorem*{additional}{Additional references}
93: \newtheorem*{acknowledgments}{Acknowledgments}
94:
95: %
96: %-------------------- SPECIFIC MATH MACROS, PART I ----------------------
97: %
98: \newcommand{\Sympe}{\Symp^e}
99: \newcommand{\crit}{\mathrm{crit}}
100: \newcommand{\Phirel}{\Phi_{rel}}
101: \newcommand{\A}{\mathcal A}
102: \newcommand{\B}{\mathcal B}
103: \newcommand{\Cone}{\mathrm{Cone}}
104: \newcommand{\tL}{\widetilde{L}}
105: \newcommand{\tphi}{\tilde{\phi}}
106: \newcommand{\ttau}{\tilde{\tau}}
107: \newcommand{\J}{\mathbf{J}}
108: \newcommand{\Lag}{{\mathit{Lag}}^\rightarrow}
109: \renewcommand{\hom}{\mathit{hom}}
110: \newcommand{\Tw}{\mathrm{Tw}}
111: \newcommand{\Ob}{\mathrm{Ob}}
112: \newcommand{\JJ}{\mathcal{J}}
113: \newcommand{\RR}{\mathcal{R}}
114: \renewcommand{\SS}{\mathcal{S}}
115: %
116: %------------------ SPECIFIC MATH MACROS, PART II -----------------
117: %
118: \newcommand{\sing}{\mathrm{sing}}
119: \newcommand{\Coh}{\mathrm{Coh}}
120: \newcommand{\Morse}{\mathit{Morse}}
121: \newcommand{\fund}[1]{[\hspace{-0.15em}[ #1 ]\hspace{-0.15em}]}
122:
123: %
124: %--------------------------- THE PAPER ----------------------------
125: %
126: \begin{document}
127: \maketitle
128:
129: \section{Introduction}
130: %
131: Not surprisingly, this is a sequel to ``Vanishing cycles and mutation''
132: \cite{seidel00}. Notions from that paper will be used freely later on, but for
133: starters a short recapitulation seems appropriate. Let $D \subset \C$ be the
134: closed unit disc. An {exact Morse fibration} $\pi: E^{2n+2} \rightarrow D$ can
135: be described roughly as a family of $2n$-dimensional exact symplectic manifolds
136: parametrized by $D$, with singularities modelled on Morse-type critical points
137: of holomorphic functions. Such objects are suitable for developing
138: Picard-Lefschetz theory in a symplectic context. Suppose that $(E,\pi)$ comes
139: equipped with a relative Maslov map $\delta_{E/D}$, which is basically a
140: trivialization of $(\Lambda^{n+1}_{\C} TE)^{\otimes 2}$. Then one can associate
141: to it an algebraic invariant, the derived directed Fukaya category
142: $D^b\Lag(\widetilde{\Gamma})$, which encodes the symplectic geometry of the
143: vanishing cycles of $(E,\pi)$. It is constructed in three steps:
144:
145: \begin{myitemize}
146: \item[(i)]
147: Fix a base point $z_0 \in \partial D$, for definiteness $z_0 = -i$; the fibre
148: $M = E_{z_0}$ is an exact symplectic manifold. Then make an admissible choice
149: of paths \cite[Figure 2]{seidel00} from $z_0$ to the critical values of $\pi$.
150: This determines a {distinguished basis of vanishing cycles} in $M$, which is an
151: ordered family $\Gamma = (L_1,\dots,L_m)$ of embedded exact Lagrangian spheres
152: $L_i \subset M$ (there is one more bit of information, so-called framings of
153: the $L_i$, which we ignore for the moment). We will sometimes call such
154: families Lagrangian configurations in $M$. Any other distinguished basis,
155: arising from a different choice of paths, can be obtained from $\Gamma$ through
156: a sequence of Hurwitz moves.
157:
158: \item[(ii)]
159: $\delta_{E/D}$ induces a Maslov map $\delta_M$ on $M$, thus allowing one to
160: speak of graded Lagrangian submanifolds in $M$. By choosing gradings in an
161: arbitrary way, one lifts $\Gamma$ to a graded Lagrangian configuration
162: $\widetilde{\Gamma} = (\tL_1,\dots,\tL_m)$. Next one introduces the {directed
163: Fukaya category} $\A = \Lag(\widetilde{\Gamma})$, which is an
164: $A_\infty$-category with objects $\tL_1,\dots,\tL_m$. The spaces of morphisms
165: in $\A$ are essentially given by the natural cochain complexes underlying the
166: Floer cohomology groups $HF^*(\tL_i,\tL_k)$, but the ordering of the objects
167: also plays a role in the definition, which is what the word ``directed'' refers
168: to.
169:
170: \item[(iii)]
171: The {derived category} $D^b(\A)$ of $\A$ is no longer an $A_\infty$-category
172: but rather a category in the ordinary sense; in fact it is a triangulated
173: category, linear over $\Z/2$ since we are using $\Z/2$-coefficients. Its
174: objects are twisted complexes in $\A$, which are $A_\infty$-analogues of chain
175: complexes, and morphisms are chain homotopy classes of maps in a suitable
176: sense. Unlike $\A$ itself, $D^b(\A)$ is independent of the choice of $\Gamma$
177: (and of $\widetilde{\Gamma}$), hence is an invariant of $(E,\pi)$ and
178: $\delta_{E/D}$. The reason for this is a relation between Hurwitz moves of
179: $\Gamma$ and the purely algebraic notion of mutations of $\A$.
180: \end{myitemize}
181: %
182: All this is explained in more detail, although still without proofs, in
183: \cite{seidel00}. The present paper adds computations and some more far-reaching
184: conjectures. It is divided into four more or less independent parts. The first
185: part, formed by Sections \ref{sec:dimension-zero}--\ref{sec:examples},
186: introduces some classes of examples. Section \ref{sec:dimension-zero} concerns
187: the toy model case when the fibre dimension is zero. Section \ref{sec:examples}
188: explains how to construct exact Morse fibrations from algebro-geometric
189: objects, namely Lefschetz pencils or isolated hypersurface singularities. Also
190: included is a discussion of the mirror manifold of $\CP{2}$.
191:
192: The subject of the second part, spanning Sections
193: \ref{sec:tqft2}--\ref{sec:donaldson}, is Hochschild cohomology. The basic
194: observation is that like $D^b(\A)$, the Hochschild cohomology $HH^*(\A,\A)$ is
195: an invariant of $(E,\pi)$ and $\delta_{E/D}$. One would expect this to have a
196: more geometric meaning. Following a suggestion of Donaldson we propose a
197: conjecture in this direction, which relates Hochschild cohomology to the global
198: monodromy of the fibration. As a preliminary to this, Section \ref{sec:tqft2}
199: enlarges the topological quantum field theory framework from \cite[Section
200: 3]{seidel00} to include the Floer cohomology of symplectic automorphisms.
201:
202: The third part, Sections \ref{sec:morse}--\ref{sec:real}, concerns exact Morse
203: fibrations with a real involution. In that situation, and under certain
204: additional assumptions, the category $D^b(\A)$ should have a description in
205: terms of Morse theory on the real part. We state a precise conjecture, which is
206: a theorem at least in the lowest nontrivial dimension $n = 1$.
207:
208: Finally, Section \ref{sec:induction} is a reflection on the aspect of
209: dimensional reduction inherent in the theory. By this we mean that $\A$, which
210: is defined in terms of pseudo-holomorphic curves in the $2n$-dimensional
211: manifold $M$, produces invariants of the $(2n+2)$-dimensional object $(E,\pi)$.
212: In an attempt to push this further, we give a conjectural algorithm for
213: determining the Floer cohomology of certain Lagrangian spheres in $E$ through
214: computations in $D^b(\A)$. This, if correct, might be the first true
215: application of homological algebra methods in symplectic geometry.
216:
217: \begin{acknowledgments}
218: This paper owes much to Donaldson; several of the ideas presented here arose
219: during conversations with him. Givental and Kontsevich provided helpful
220: explanations concerning mirror symmetry.
221: \end{acknowledgments}
222:
223: \newpage
224:
225: \section{Branched covers\label{sec:dimension-zero}}
226: %
227: \subsection{}
228: In \cite{seidel00} it was tacitly assumed that all symplectic manifolds were of
229: dimension $>0$ (and that assumption will be resumed as soon as this section is
230: finished). The zero-dimensional case is rather trivial, but it still requires
231: some minor adaptations, which we now explain. Let's start with a basic
232: dictionary of zero-dimensional symplectic geometry:
233:
234: \begin{center}
235: \begin{tabular}{l||l}
236: compact symplectic manifold & finite set $M$ \\ \hline
237: Lagrangian sphere & subset $L \subset M$, $|L| = 2$
238: \\ \hline
239: Floer cohomology & $HF(L_1,L_2) = (\Z/2)^{L_1 \cap L_2}$
240: \\ \hline
241: Dehn twist & \begin{minipage}{16em} \vspace{2pt}
242: \raggedright
243: $\tau_L: M \rightarrow M$ swapping the two
244: points of $L$, leaving the rest fixed \vspace{2pt}
245: \end{minipage}
246: \\ \hline
247: exact Morse fibration & generic branched cover $\pi: E \rightarrow S$
248: \end{tabular}
249: \end{center}
250: Here a {\bf generic branched cover} is a proper map $\pi: E \rightarrow S$
251: between oriented surfaces with boundary, with $\pi^{-1}(\partial S) = \partial
252: E$, having finitely many critical points, which are locally modelled on $\pi(z)
253: = z^2$; and no two critical points may lie in the same fibre.
254:
255: The TQFT framework deserves a brief mention, if only for its simplicity. Let
256: $S$ be a compact oriented surface with boundary, $\Sigma \subset \partial S$ a
257: finite set of marked points, and $(E,\pi)$ a generic branched cover over $S^* =
258: S \setminus \Sigma$. Suppose moreover that we have a submanifold $Q \subset
259: \pi^{-1}(\partial S^*)$ such that $\pi|Q: Q \rightarrow \partial S^*$ is a
260: double cover; this corresponds to what in \cite{seidel00} was called a
261: Lagrangian boundary condition for $(E,\pi)$. The behaviour of $Q$ near a point
262: $\zeta \in \Sigma$ can be described by a pair $(L_{\zeta,+},L_{\zeta,-})$ of
263: Lagrangian zero-spheres in some fibre of $E$ close to $\zeta$. A continuous
264: section $u: S^* \rightarrow E$ of $\pi$ which satisfies $u(\partial S^*)
265: \subset Q$ singles out a point of $L_{\zeta,+} \cap L_{\zeta,-}$ for each
266: $\zeta$, and hence determines an element of $\bigotimes_\zeta HF(L_{\zeta,+},
267: L_{\zeta,-})$. Taking the sum over all sections defines an invariant
268: \begin{equation} \label{eq:rel}
269: \Phirel(E,\pi,Q) \in \bigotimes_{\zeta \in \Sigma} HF(L_{\zeta,+},L_{\zeta,-}).
270: \end{equation}
271: For trivial reasons, these invariants satisfy the gluing law formulated in
272: \cite[Section 3]{seidel00}. One can use them to find the correct
273: zero-dimensional analogues of various familiar maps in Floer theory. For
274: instance, the pair-of-pants product $HF(L_2,L_3) \otimes HF(L_1,L_2)
275: \rightarrow HF(L_1,L_3)$ is derived from an invariant \eqref{eq:rel} with $S =
276: D$, $\Sigma = \{\text{\em three points}\}$ and $E = S^* \times M$, together
277: with a boundary condition given by $L_1,L_2,L_3$ over the three connected
278: components of $\partial S^*$. Working this out explicitly shows that the
279: product takes a basis element $\bar{x} \otimes \bar{y} \in HF(L_2,L_3) \otimes
280: HF(L_1,L_2)$, $x \in L_2 \cap L_3$ and $y \in L_1 \cap L_2$, to $\bar{x} \in
281: HF(L_1,L_3)$ if $x = y$, and to zero otherwise. There is the same long exact
282: sequence in Floer cohomology as in \cite[Theorem 3.3]{seidel00} and again, the
283: more fundamental fact underlying it is a quasi-isomorphism
284: \[
285: \Cone(a: HF(L,L_2) \otimes HF(L_1,L) \rightarrow HF(L_1,L_2))
286: \xrightarrow{(h,b)} HF(L_1,\tau_L(L_2)).
287: \]
288: However, while $a$ and $b$ are as in higher dimensions, $h$ is the product
289: $HF(L,L_2) \otimes HF(L_1,L) \iso HF(L,\tau_L(L_2)) \otimes HF(L_1,L)
290: \rightarrow HF(L_1,\tau_L(L_2))$. Of course, both the result and proof are now
291: elementary!
292:
293: The next topic are gradings. We summarize their zero-dimensional theory in
294: another dictionary:
295:
296: \begin{center}
297: \begin{tabular}{l||l}
298: \begin{minipage}{9em} \vspace{2pt}
299: Maslov map on a \newline symplectic manifold \vspace{2pt}
300: \end{minipage}
301: & $\delta_M: M \rightarrow S^1$
302: \\ \hline
303: \begin{minipage}{9em} \vspace{2pt} \raggedright
304: graded Lagrangian sphere \vspace{2pt}
305: \end{minipage}
306: &
307: \begin{minipage}{18em} \vspace{2pt} \raggedright
308: $L \subset M$ with a map $\tL: L \rightarrow \R$ such
309: that $\exp(2\pi i \tL) = \delta_M|L$ \vspace{2pt}
310: \end{minipage}
311: \\ \hline
312: \begin{minipage}{9em} \vspace{2pt} \raggedright
313: graded Floer cohomology \vspace{2pt}
314: \end{minipage}
315: &
316: \begin{minipage}{18em} \vspace{2pt} \raggedright
317: in $HF^*(\tL_1,\tL_2)$, the degree of $\bar{x}$ is
318: $\tL_2(x) - \tL_1(x)$
319: \vspace{2pt}
320: \end{minipage}
321: \\ \hline
322: shifting the grading &
323: $\tL[\sigma] = \tL - \sigma$, $\sigma \in \Z$
324: \\ \hline
325: \begin{minipage}{9em} \vspace{2pt} \raggedright
326: graded symplectic automorphism \vspace{2pt}
327: \end{minipage}
328: &
329: \begin{minipage}{18em} \vspace{2pt}
330: $\phi: M \rightarrow M$ with a map $\tphi: M \rightarrow \R$ such that
331: $\exp(2\pi i \phi(x)) = \delta_M(x)/\delta_M(\phi^{-1}(x))$
332: \vspace{2pt}
333: \end{minipage}
334: \\ \hline
335: relative Maslov map &
336: \begin{minipage}{18em} \vspace{2pt} \raggedright
337: on $\pi: E \rightarrow S$, a trivialization
338: $\delta_{E/S}$ of $TE^{\otimes 2} \otimes (\pi^*TS)^{\otimes -2}$
339: \end{minipage}
340: \end{tabular}
341: \end{center}
342: %
343: The action of a graded symplectic automorphism on graded Lagrangian
344: submanifolds is defined by $(\tphi \tL)(x) = \tL(\phi^{-1}(x)) + \tphi(x)$. A
345: grading $\tL$ determines a preferred grading $\ttau_{\tL}$ of the associated
346: Dehn twist, characterized by
347: \[
348: \ttau_{\tL}(\tL) = \tL[1] \quad \text{ and } \quad \ttau_{\tL}(x) = 0 \text{
349: for } x \notin L;
350: \]
351: compare \cite[Equation (4)]{seidel00} for the first property. Graded Lagrangian
352: configurations $\widetilde{\Gamma} = (\tL_1,\dots,\tL_m)$ are defined as one
353: would expect. Their Hurwitz moves are
354: \begin{itemize}
355: \item
356: $\widetilde{\Gamma} \rightsquigarrow (\tL_1[\sigma_1],\dots,\tL_m[\sigma_m])$
357: for $\sigma_1,\dots,\sigma_m \in \Z$;
358: \item
359: $\widetilde{\Gamma} \rightsquigarrow c\widetilde{\Gamma} =
360: (\ttau_{\tL_1}(\tL_2),\dots,\ttau_{\tL_1}(\tL_m),\tL_1)$;
361: \item
362: $\widetilde{\Gamma} \rightsquigarrow r\widetilde{\Gamma} =
363: (\tL_1,\dots,\tL_{m-2},\ttau_{\tL_{m-1}}(\tL_m),\tL_{m-1})$.
364: \end{itemize}
365: Let $(E,\pi)$ be a generic branched cover over $D$, with a relative Maslov map
366: $\delta_{E/D}$. A drastically simplified version of the usual Picard-Lefschetz
367: argument produces from it a graded (with respect to the induced Maslov map
368: $\delta_M$) Lagrangian configuration in $M = E_{z_0}$, unique up to Hurwitz
369: equivalence.
370:
371: \begin{remark}
372: One can observe here two minor differences with respect to the
373: positive-dimensional situation. Firstly, $\ttau_{\tL}$ depends on $\tL$;
374: secondly, the Hurwitz equivalence class of a graded configuration contains
375: information that is lost if one forgets the gradings. Both phenomena go back to
376: the obvious fact that there is a $\Z^2$ ambiguity in the choice of grading for
377: a Lagrangian zero-sphere, as opposed to $\Z$ in higher dimensions.
378: \end{remark}
379:
380: The directed Fukaya category $\A = \Lag(\widetilde{\Gamma})$ attached to a
381: zero-dimensi\-onal graded configuration has objects $\tL_i$. The morphisms
382: $hom_\A(\tL_i,\tL_k)$ are $HF^*(\tL_i,\tL_k)$ for $i<k$, $\Z/2 \cdot
383: \id_{\tL_i}$ for $i = k$, and zero for $i>k$. The nontrivial compositions
384: $HF^*(\tL_j,\tL_k) \otimes HF^*(\tL_i,\tL_j) \rightarrow HF^*(\tL_i,\tL_k)$,
385: $i<j<k$, are given by the pair-of-pants product. Even though $\A$ is an
386: ordinary $\Z/2$-linear and $\Z$-graded category, we prefer to regard it as a
387: directed $A_\infty$-category where all composition maps $\mu_\A^d$ of order $d
388: \neq 2$ vanish. The details of the proof are slightly different than in the
389: higher-dimensional case, but the main result of \cite{seidel00} remains true,
390: which is that Hurwitz moves of $\widetilde{\Gamma}$ give rise to mutations of
391: $\A$; and one arrives at the familiar conclusion that
392: $D^b\Lag(\widetilde{\Gamma})$ is an invariant of $(E,\pi)$, $\delta_{E/D}$.
393:
394: Before going on to concrete examples, we need to recall some algebraic
395: terminology. Let $T$ be a quiver (an oriented graph). The path category
396: ${\mathcal P}T$ has one object for each vertex, and the space of morphisms
397: between two objects is the $\Z/2$-vector space freely generated by all paths in
398: $T$ going from one vertex to the other. One can further divide the morphism
399: spaces by some two-sided ideal, in order to kill certain designated morphisms.
400: This is usually referred to as describing a category by a ``quiver with
401: relations''. If the quiver is directed, which means that its vertices are
402: numbered $1,\dots,m$ such that there are no paths from the $i$-th one to the
403: $k$-th one unless $i<k$, ${\mathcal P}T$ and its quotients can again be seen as
404: special cases of directed $A_\infty$-categories, with morphisms only in degree
405: zero and vanishing composition maps of order $\neq 2$. Note that in this
406: situation our notion of derived category (defined through twisted complexes)
407: agrees with the classical one (defined by considering the categories as
408: algebras, and taking chain complexes of right modules over them).
409:
410: \subsection{}
411: Take the following quiver with $m$ vertices, with the arrows oriented in an
412: arbitrary way:
413: \begin{equation} \label{eq:bgp}
414: \xymatrix{
415: {\bullet} \ar[r] &
416: {\bullet} & \ar[l]
417: {\bullet} \ar[r] &
418: {\dots} &
419: {\bullet} \ar[l]
420: }
421: \end{equation}
422: and let $\A$ be its path category. It is a classical result that $D^b(\A)$ is
423: independent of the orientation of the arrows; what we will do is to explain
424: this geometrically. First of all, one can find a configuration of Lagrangian
425: zero-spheres $\Gamma = (L_1,\dots,L_m)$ in $M = \{1,\dots,m+1\}$ which, when
426: lifted to a graded configuration in the trivial way ($\delta_M = 0$ and $\tL_i
427: = 0$), has $\A$ as its directed Fukaya category. The way to do that is best
428: explained by an example:
429: \begin{equation} \label{eq:makebasis}
430: \begin{split}
431: & \xymatrix@R=-0.25em{
432: \bullet \ar[r] &
433: \bullet \ar[r] &
434: \bullet &
435: \bullet \ar[l] &
436: \bullet \ar[l] &
437: \bullet \ar[l] \ar[r] &
438: \bullet
439: \\
440: {\hspace{-1em} \{1,2\} \hspace{-1em}} &
441: {\hspace{-1em} \{1,3\} \hspace{-1em}} &
442: {\hspace{-1em} \{1,4\} \hspace{-1em}} &
443: {\hspace{-1em} \{5,4\} \hspace{-1em}} &
444: {\hspace{-1em} \{6,4\} \hspace{-1em}} &
445: {\hspace{-1em} \{7,4\} \hspace{-1em}} &
446: {\hspace{-1em} \{7,8\} \hspace{-1em}}
447: }
448: \\[0.5em]
449: \Gamma & = (\{1,2\},\{1,3\},\{7,4\},\{6,4\},\{5,4\},\{1,4\},\{7,8\}).
450: \end{split}
451: \end{equation}
452: Next one constructs an $(m+1)$-fold generic branched cover $\pi: E \rightarrow
453: D$ such that $\Gamma$ is one of its distinguished bases of vanishing cycles.
454: Inspection of \eqref{eq:makebasis} and an Euler characteristic computation show
455: that the total space $E$ is connected and in fact a disc. It is a consequence
456: of the classification of branched covers that there is just one such $(E,\pi)$
457: up to isomorphism, and moreover it admits precisely one homotopy class of
458: relative Maslov maps. Therefore all categories $D^b(\A)$ arise from the same
459: geometric situation, hence must be equivalent on grounds of the general theory.
460:
461: \subsection{}
462: Let $\pi: E \rightarrow D$ be a double cover branched along $2g+1 \geq 5$
463: points, so that the total space is a genus $g \geq 2$ surface with one boundary
464: component. Write $\iota$ for the nontrivial covering transformation, which is a
465: hyperelliptic involution of $E$. There is a unique homotopy class of relative
466: Maslov maps which are invariant under $\iota$; take $\delta_{E/D}$ in that
467: class. All vanishing cycles of $(E,\pi)$ are the same, and if we choose their
468: gradings in the most obvious way, the resulting directed Fukaya category $\A_g$
469: can be described by the quiver with relations
470: \[
471: \xymatrix{
472: {\bullet} \ar@/^1pc/[r]^{a_1} \ar@/_1pc/[r]_{b_1} &
473: {\bullet} \ar@/^1pc/[r]!<-1em,0em>^{a_2} \ar@/_1pc/[r]!<-1em,0em>_{b_2} &
474: {\bullet\; \dots\; \bullet}
475: \ar@/^1pc/[]!<1em,0em>;[r]^{a_{2g}}
476: \ar@/_1pc/[]!<1em,0em>;[r]_{b_{2g}} &
477: {\bullet}
478: } \qquad \qquad
479: \begin{cases}
480: b_{i+1}a_i = 0, & \\
481: a_{i+1}b_i = 0.
482: \end{cases}
483: \]
484: We will now make some further remarks concerning $\A_g$ and its derived
485: category. These may seem a bit unmotivated, but they will be at least partially
486: put into context later on; see Remark \ref{th:zero-spheres}. The following
487: definition is taken from \cite{seidel-thomas99}:
488:
489: \begin{defn} \label{def:spherical}
490: Let ${\mathcal C}$ be a triangulated category, linear over $\Z/2$ and such that
491: the spaces $\Hom^*_{\mathcal C}(-,-)$ are finite-dimensional. $C \in
492: \Ob\,{\mathcal C}$ is called spherical of dimension $(n+1)$ if
493: $\Hom^*_{\mathcal C}(C,C) \iso H^*(S^{n+1};\Z/2)$ and the composition
494: \[
495: \Hom^*_{\mathcal C}(X,C) \otimes \Hom^{n+1-*}_{\mathcal C}(C,X) \rightarrow
496: \Hom^{n+1}_{\mathcal C}(C,C) \iso \Z/2
497: \]
498: is a nondegenerate pairing for any $X \in \Ob\,{\mathcal C}$.
499: \end{defn}
500:
501: Under some additional assumptions on ${\mathcal C}$, which are satisfied for
502: derived categories of directed $A_\infty$-categories, one can associate to any
503: spherical object $C$ an exact self-equivalence of ${\mathcal C}$, the twist
504: functor $T_C$, which is well-defined up to isomorphism. If $C_1,C_2$ are
505: spherical with $\Hom^*_{\mathcal C}(C_1,C_2) = 0$ then their twist functors
506: commute; and if $\Hom^*_{\mathcal C}(C_1,C_2)$ is one-dimensional, one gets a
507: braid relation $T_{C_1}T_{C_2}T_{C_1} \iso T_{C_2}T_{C_1}T_{C_2}$.
508:
509: Denote the objects of $\A_g$ by $(X^1,\dots,X^{2g+1})$. The twisted complexes
510: $C_i = (X^i[1] \oplus X^{i+1}, \left(\begin{smallmatrix} 0 & 0 \\ a_i+b_i & 0
511: \end{smallmatrix}\right) )$, $1 \leq i \leq 2g$, are spherical objects of
512: dimension one in $D^b(\A_g)$, and moreover
513: \[
514: \dim\, \Hom^*_{D^b(\A_g)}(C_i,C_k) =
515: \begin{cases}
516: 0 & |i-k| \geq 2, \\
517: 1 & |i-k| = 1.
518: \end{cases}
519: \]
520: This implies that $T_{C_1},\dots,T_{C_{2g}}$ generate an action (in a weak
521: sense) of the braid group $B_{2g+1}$ on $D^b(\A_g)$. Now let $D^{per}(\A_g)$ be
522: the category defined like $D^b(\A_g)$ but using $\Z/2$-graded twisted
523: complexes. This is triangulated, and its two-fold shift functor $[2]$ is
524: isomorphic to the identity; we call it the periodic derived category of $\A_g$.
525: One can adapt the theory of spherical objects and twist functors to such
526: categories. In our case, this means that there are $C_i \in D^{per}(\A_g)$
527: defined as before, with the same properties. In addition there is now another
528: spherical object
529: \[
530: C_0 = (X^1 \oplus X^2[1] \oplus X^3 \oplus X^4[1],
531: \mbox{\smaller $\begin{pmatrix}
532: 0 & 0 & 0 & 0 \\
533: a_1 & 0 & 0 & 0 \\
534: 0 & b_2 & 0 & 0 \\
535: b_3b_2b_1 & 0 & a_3 & 0
536: \end{pmatrix}$} )
537: \]
538: The dimension of $\Hom^*_{D^{per}(\A_g)}(C_0,C_i)$ is one if $i = 4$ and zero
539: for all other $i>0$. The diagram representing this situation,
540: \[
541: \xymatrix{
542: & & & {C_0} \ar@{-}[d] \\
543: {C_1} \ar@{-}[r] &
544: {C_2} \ar@{-}[r] &
545: {C_3} \ar@{-}[r] &
546: {C_4} \ar@{-}[r] &
547: {C_5} \ar@{-}[r] &
548: \dots \ar@{-}[r] &
549: {C_{2g}}
550: }
551: \]
552: has the same structure as the configuration of curves in Wajnryb's presentation
553: of $R_{g,1}$, the mapping class group of a genus $g$ surface with one marked
554: point \cite{wajnryb83}; and one can ask
555:
556: \begin{question}
557: Let ${\mathcal C} \subset D^{per}(\A_g)$ be the full triangulated subcategory
558: generated by the $C_i$. Do $T_{C_0},\dots,T_{C_{2g}}$ generate an action of
559: $R_{g,1}$ on ${\mathcal C}$?
560: \end{question}
561:
562: \begin{additional}
563: The classical paper about the quiver \eqref{eq:bgp} is by Bernstein, Gelfand,
564: and Ponomarev \cite{bernstein-gelfand-ponomarev72}. The category $D^b(\A_g)$ is
565: closly related to those studied in \cite{khovanov-seidel98}, even though the
566: quiver presentation looks different. Those papers do not consider periodic
567: derived categories and therefore miss out on the additional object $C_0$.
568: \end{additional}
569:
570: \section{Examples from algebraic geometry\label{sec:examples}}
571: %
572: \subsection{}
573: Let $X$ be a smooth projective variety, $\xi \rightarrow X$ an ample line
574: bundle, and $\sigma_0,\sigma_1$ two holomorphic sections of $\xi$ which
575: generate a Lefschetz pencil of hypersurfaces $Y_z = \{ x \in X \suchthat
576: \sigma_0(x)/\sigma_1(x) = z\}$, $z \in \CP{1} = \C \cup \{\infty\}$. Suppose,
577: in addition to the Lefschetz condition, that $Y_\infty$ is smooth. What we want
578: to look at is, in principle, the holomorphic Morse function
579: \[
580: \sigma_0/\sigma_1 : X \setminus Y_\infty \longrightarrow \C.
581: \]
582: To obtain additional symplectic data, one chooses a metric on $\xi$, with
583: corresponding connection $A$, such that $\o_X = (i/2\pi)F_A$ is a K{\"a}hler
584: form. By a standard construction, $A$ and $\sigma_1$ give rise to a one-form
585: $\theta_X$ defined on $X \setminus Y_\infty$, such that $d\theta_X = \o_X$.
586: This still doesn't fit the definition of an exact Morse fibration, making some
587: further modifications necessary.
588:
589: Let $Z = Y_0 \cap Y_\infty$ be the base locus of the pencil, $G = \{ (z,x) \in
590: \CP{1} \times X \suchthat \sigma_0(x)/\sigma_1(x) = z\}$ its graph, and $p: G
591: \rightarrow \CP{1}$ the projection. $(G,p)$ is a Lefschetz fibration with
592: fibres $Y_z$, and contains a trivial subfibration $\CP{1} \times Z \subset G$.
593: Let $\O_G \in \Omega^2(G)$ be the pullback of $\o_X$; it is closed,
594: nondegenerate on the vertical tangent spaces $\ker(Dp) \subset TG$, and hence
595: defines a symplectic connection away from the critical points of $p$. Moreover,
596: the restriction of this connection to $\CP{1} \times Z \subset G$ is trivial.
597: Similarly, pulling back $\theta_X$ gives a one-form $\Theta_G$ defined on the
598: complement of $(\{\infty\} \times Y_\infty) \cup (\CP{1} \times Z) \subset G$,
599: with $d\Theta_G = \Omega_G$.
600:
601: After possibly rescaling $\sigma_0$, we may assume that $Y_z$ is smooth for
602: $|z| \geq 1$. Let $U \subset Y_0$ be a small open neighbourhood of $Z$. Using
603: the symplectic parallel transport in radial directions one constructs an
604: embedding $\Xi$,
605: \[
606: \xymatrix{
607: {D \times U} \ar[rr]^-{\Xi} \ar[dr] &&
608: {p^{-1}(D) \subset G} \ar[dl]^p \\
609: & {D} &
610: }
611: \]
612: with $\Xi\, |\, D \times Z$ and $\Xi\, |\,\{0\} \times U$ the obvious
613: inclusions, which has the property that $\Xi^*\Omega_G - \o_X|U$ vanishes on
614: each diameter $[-z;z] \times U$, $z \in S^1$. Let $h \in \smooth_c(U,[0;1])$ be
615: a function which is $\equiv 1$ on some smaller neighbourhood $U' \subset U$ of
616: $Z$. Take the map $H: p^{-1}(D) \rightarrow p^{-1}(D)$ which is the identity
617: outside $\im(\Xi)$ and satisfies $H\,\Xi(z,x) = \Xi((1-h(x))z,x)$.
618:
619: Set $E = p^{-1}(D) \setminus \Xi(D \times U'')$ for an even smaller open
620: neighbourhood $U'' \subset U'$ of $Z$, which we choose such that $\partial U''$
621: is smooth. Together with $\pi = p|E$, $\Omega = (H^*\Omega_G|E) \in
622: \Omega^2(E)$, $\Theta = (H^*\Theta_G|E) \in \Omega^1(E)$, and the given complex
623: structure near the critical points of $\pi$, this forms an exact Morse
624: fibration over $D$. Indeed, the required trivialization near $\partial_hE$ is
625: given by $\Xi$ itself; and the fact that $H$ contracts along the same radial
626: directions which are used to define $\Xi$ ensures that $\Omega$ equals
627: $\Omega_G$ on $\ker(D\pi) \subset TE$, hence is nondegenerate there.
628:
629: Suppose now that the canonical bundle $K_X$ satisfies
630: \begin{equation} \label{eq:relative-maslov}
631: K_X^{-2} \iso \xi^a \, \text{ holomorphically, for some $a \in \Z$.}
632: \end{equation}
633: Then $\sigma_1^a$ gives a trivialization of $K_X^{-2}$ over $X \setminus
634: Y_\infty$, determined up to multiplication with a constant. Since the
635: projection $G \rightarrow X$ identifies $E$ with a compact subset of $X
636: \setminus Y_\infty$, and $\O$ agrees with the pullback of $\o_X$ except in a
637: small neighbourhood of $\partial_hE$, one gets, at least up to homotopy, a
638: preferred relative Maslov map $\delta_{E/D}$. The category
639: $D^b\Lag(\widetilde{\Gamma})$ associated to $(E,\pi)$ and $\delta_{E/D}$
640: actually depends only on $X$ and on the Lefschetz pencil, and not on the other
641: choices made during the construction of $E$. That is because different choices
642: lead to exact Morse fibrations which are deformation equivalent in a suitable
643: sense. For the same reason, if $\xi$ is such that any two generic sections
644: define a Lefschetz pencil, the category is an invariant of the pair $(X,\xi)$.
645: %
646: \includefigure{degree2}{degree2.eps}{hb}%
647:
648: As a concrete example take $X = \CP{2}$, $\xi = {\mathcal O}(2)$. The fibre $M
649: = E_{z_0}$ of $(E,\pi)$ is, topologically, $\CP{1}$ with four small discs
650: removed. The relative Maslov map $\delta_{E/D}$ coming from
651: \eqref{eq:relative-maslov} induces a Maslov map $\delta_M$ on $M$, which is
652: characterized by having the same behaviour over each component of $\partial M$.
653: Figure \ref{fig:degree2} shows a distinguished basis $\Gamma = (L_1,L_2,L_3)$
654: of vanishing cycles. One can read off directly that the directed Fukaya
655: category $\A = \Lag(\widetilde{\Gamma})$, for a suitable choice of gradings, is
656: described by the quiver with relations
657: \begin{equation} \label{eq:cp2-quiver}
658: \xymatrix{
659: {\bullet} \ar@/^1pc/[rr]^{a_1} \ar@/_1pc/[rr]_{b_1} &&
660: {\bullet} \ar@/^1pc/[rr]^{a_2} \ar@/_1pc/[rr]_{b_2} &&
661: {\bullet}
662: } \qquad
663: \begin{cases}
664: a_2a_1 = b_2b_1, & \\
665: b_2a_1 = a_2b_1. &
666: \end{cases}
667: \end{equation}
668: Namely, the intersections $L_i \cap L_k$, $i<k$, consist of two points which
669: are essential, in the sense that they cannot be removed by an isotopy in $M$.
670: This implies that $hom_\A(\tL_i,\tL_k)$ is two-dimensional and $\mu^1_\A = 0$.
671: And there are four triangles in $M$ whose sides (in positive order) map to
672: $(L_1,L_2,L_3)$, giving rise to the nonzero products $\mu^2_\A(a_2,a_1)$,
673: $\mu^2_\A(b_2,a_1)$, $\mu_\A^2(a_2,b_1)$, $\mu^2_\A(b_2,b_1)$.
674:
675: \begin{remark}
676: Suppose that \eqref{eq:relative-maslov} holds but with $a \in \Q$, which is to
677: say $K_X^{-2c} \iso \xi^b$ for coprime $b,c$. Then $\sigma_1^b$ defines a
678: $c$-sheeted cyclic covering of $X \setminus Y_\infty$, hence also a covering of
679: $E$. The total space of the latter can be made into an exact Morse fibration
680: which then has a canonical relative Maslov map, so that our theory can be
681: applied to it. The $\Z/c$-action on $E$ by covering transformations induces a
682: self-equivalence $\sigma$ of $D^b\Lag(\widetilde{\Gamma})$ which satisfies
683: $\sigma^c \iso [1]$. One can view this as a category with ``fractional
684: gradings'', thinking of $\Hom(X,\sigma^iY)$ as ``the group $\Hom^{i/c}(X,Y)$ of
685: morphisms of degree $i/c$''.
686: \end{remark}
687:
688: \subsection{}
689: We need to extend the previous discussion to a degenerate situation. Let $X
690: \subset \CP{N}$ be a projective variety with singular set $X^\sing$, and
691: $\sigma_0,\sigma_1 \in (\C^{N+1})^*$ two linear forms which generate a pencil
692: $(Y_z)$ of hyperplane sections of $X$; so the line bundle we are looking at is
693: $\xi = {\mathcal O}_X(1)$. Suppose that the following conditions are satisfied:
694: $X^\sing \subset Y_\infty \setminus Y_0$; the base locus $Z = Y_0 \cap
695: Y_\infty$ is smooth; and $(Y_z)_{z \neq \infty}$ satisfies the same
696: nondegeneracy conditions as a Lefschetz pencil. Then the associated graph $p: G
697: \rightarrow \CP{1}$ is well-behaved except over $\infty \in \CP{1}$. The
698: construction made above, in which the fibre at infinity plays no role, goes
699: through without any changes, yielding an exact Morse fibration $(E,\pi)$ over
700: $D$. If we suppose moreover that $X$ is locally a complete intersection, so
701: that there is a well-defined canonical bundle $K_X$, then condition
702: \eqref{eq:relative-maslov} again ensures the existence of preferred relative
703: Maslov maps $\delta_{E/D}$.
704: %
705: \includefigure{mirrorp2}{mirrorp2.eps}{hb}%
706:
707: The example which we have in mind is $X = \{ x \in \CP{3} \suchthat x_0^3 =
708: x_1x_2x_3\}$, $\sigma_0(x) = x_1+x_2+x_3$, $\sigma_1(x) = x_0$. One can
709: identify $X \setminus Y_\infty$ with the affine hypersurface $x_1x_2x_3 = 1$,
710: and then $\sigma_0/\sigma_1$ becomes the function $x_1 + x_2 + x_3$, which
711: appears in Givental's work \cite[Theorem 5]{givental94} as the object mirror
712: dual to the projective plane. The next result exhibits the mirror phenomenon
713: within our framework.
714:
715: \begin{prop} \label{th:mirror}
716: Let $(E,\pi)$ be the exact Morse fibration constructed from
717: $(X,\sigma_0,\sigma_1)$, with relative Maslov map $\delta_{E/D}$ obtained from
718: \eqref{eq:relative-maslov}. Then for any graded distinguished basis of
719: vanishing cycles ones has
720: \[
721: D^b\Lag(\widetilde{\Gamma}) \iso D^b Coh({\mathbb P}^2)
722: \]
723: where $D^b Coh({\mathbb P}^2)$ is the derived category, in the classical sense,
724: of coherent sheaves on the projective plane over the field $\Z/2$.
725: \end{prop}
726:
727: The fibre $M = E_{z_0}$ is a torus with three small discs removed. Figure
728: \ref{fig:mirrorp2} shows a particular distinguished basis of vanishing cycles
729: $\Gamma = (L_1,L_2,L_3)$. In that picture, the Maslov map $\delta_M$ is that
730: given by the flat structure of $M$. One can therefore choose gradings such that
731: the groups $CF^*(\tL_i,\tL_k)$, $i<k$, are concentrated in degree zero. Then
732: $\A = \Lag(\widetilde{\Gamma})$ is given by the quiver with relations
733: \[
734: \xymatrix{
735: {\bullet} \ar@/^1.5pc/[rr]^{a_1} \ar[rr]^{a_2} \ar@/_1.5pc/[rr]_{a_3} &&
736: {\bullet} \ar@/^1.5pc/[rr]^{b_1} \ar[rr]^{b_2} \ar@/_1.5pc/[rr]_{b_3} &&
737: {\bullet}
738: } \qquad
739: \begin{cases}
740: b_i a_j = b_j a_i & i \neq j, \\
741: b_i a_i = 0.
742: \end{cases}
743: \]
744: A theorem of Beilinson \cite{beilinson78} says that $D^b(\A) \iso D^b
745: Coh({\mathbb P}^2)$. Since $D^b(\A)$ remains the same for all other
746: distinguished bases, the Proposition is proved. The same approach can be used
747: for the mirror partner of $\CP{1} \times \CP{1}$, with analogous results; I
748: have not checked any further cases.
749:
750: \begin{remarks}
751: (i) It is maybe helpful to mention one way of making drawings like Figure
752: \ref{fig:mirrorp2}. By projecting to one coordinate, each fibre
753: $(\sigma_0/\sigma_1)^{-1}(z)$ of $\sigma_0/\sigma_1: X \setminus Y_\infty
754: \rightarrow \C$ can be represented as a double covering of $\C^*$ branched over
755: three points. The vanishing cycles are preimages of paths in $\C^*$ joining two
756: branch points, and one finds them by looking at how the branch points come
757: together as one varies $z$.
758:
759: (ii) One sees from Figure \ref{fig:mirrorp2} that the Dehn twists $\tau_{L_i}$
760: can be represented by affine maps, at least away from a small neighbourhood of
761: the missing discs. It follows that any Lagrangian configuration $\Gamma' =
762: (L_1',L_2',L_3')$ which is Hurwitz equivalent to $\Gamma$ consists of curves
763: $L_i'$ isotopic in $M$ to straight lines. An easy argument using the topology
764: of $E$ %specifically $\pi_2(E) = 0$
765: shows that no two $L_i'$ can ever be isotopic to each other. Therefore
766: $HF^*(\tL_i',\tL_k')$, $i<k$, is concentrated in a single degree, for any
767: choice of gradings. This is related to the fact that mutations of the standard
768: exceptional collection in $D^b Coh({\mathbb P}^2)$ remain strongly exceptional;
769: see \cite{bondal-polishchuk94} for much more about this subject.
770: \end{remarks}
771:
772: \subsection{}
773: Let $f \in \C[x_1,\dots,x_{n+1}]$ be a polynomial with an isolated critical
774: point at the origin. A Morsification of $f$ is a smooth family of polynomials
775: $(f_t)_{0 \leq t < \rho}$ with $f_0 = f$ and such that the critical points of
776: $f_t$, $t>0$, which lie near the origin are nondegenerate. For sufficiently
777: small $0<\epsilon$, $0<\delta \ll \epsilon$, and $0 < t \ll \delta$, one finds
778: that
779: \[
780: X = \{ x \in \C^{n+1} \suchthat |x| \leq \epsilon, \; |f_t(x)| \leq \delta\}
781: \]
782: is a manifold with corners, and $f_t$ a holomorphic Morse function on it. One
783: can equip $X$ with the standard symplectic form on $\C^{n+1}$, and its
784: canonical bundle has a standard trivialization obtained from the constant
785: holomorphic volume form. After some modifications, this becomes an exact Morse
786: fibration $(E,\pi)$ over $D$ together with a canonical homotopy class of
787: relative Maslov maps $\delta_{E/D}$. We will not explain the details since the
788: procedure is similar to that for Lefschetz pencils. A deformation equivalence
789: argument shows that the resulting category $D^b\Lag(\widetilde{\Gamma})$ is
790: independent of all choices, and even of the Morsification, so that it is an
791: invariant of $f$.
792:
793: \begin{prop} \label{th:ade}
794: Let $f \in \C[x_1,x_2]$ be an ADE singularity. Then the category
795: $D^b\Lag(\widetilde{\Gamma})$ associated to it is equivalent to the derived
796: path category of the Dynkin quiver of the same type, with an arbitrary
797: orientation of the arrows.
798: \end{prop}
799:
800: A'Campo \cite[pp.\ 13--17]{acampo75b} has constructed particularly nice
801: Morsifications and distinguished bases $\Gamma = (L_1,\dots,L_m)$ for these
802: singularities. These have the properties that (i) there is a bijection between
803: the $L_i$ and the vertices of the corresponding Dynkin diagram, such that $L_i
804: \cap L_k$ consists of a single point if the $i$-th and $k$-th vertex are
805: connected by a line, and is empty otherwise; (ii) there are no $i<k<l$ with
806: $L_i \cap L_k \neq \emptyset$ and $L_k \cap L_l \neq \emptyset$. Figure
807: \ref{fig:e6} shows the $E_6$ case. One can choose gradings $\tL_i$ such that
808: $HF^*(\tL_i,\tL_k)$ is concentrated in degree zero for all $i<k$. Then
809: $\Lag(\widetilde{\Gamma})$ is the path category of the Dynkin quiver for a
810: particular orientation, which is such that each vertex is either a sink or a
811: source. Changing the orientation does not affect the derived category of the
812: path category, which completes the proof.
813: %
814: \includefigure{e6}{e6.eps}{ht}%
815:
816: We mention two more expected properties of the categories
817: $D^b\Lag(\widetilde{\Gamma})$ as invariants of singularities. Strictly speaking
818: these are conjectures, but the proofs should not be difficult.
819: \begin{myitemize}
820: \item ($\mu$-invariance)
821: They should remain the same when one deforms $f$ inside its $\mu = \text{\em
822: const.}$ stratum. Moreover, if $f$ is adjacent to $g$ then the category
823: belonging to $f$ should contain that of $g$ as a full triangulated subcategory.
824: \item (Stabilization)
825: The categories associated to $f(x)$ and $\hat{f}(x,y) = f(x) + y^2_1 + \dots +
826: y_q^2$ should be equivalent, for any $q$.
827: \end{myitemize}
828: For instance, the reader may have noticed that the categories of
829: $A_m$-singula\-ri\-ties in Proposition \ref{th:ade} are the same as those
830: arising from certain branched covers, which were considered in Section
831: \ref{sec:dimension-zero}b. This is an extreme case of the stabilization
832: property, since the branched covers can be viewed as Morsifications of the
833: one-dimensional singularities $f(x) = x^{m+1}$.
834:
835: \begin{additional}
836: Hori, Iqbal and Vafa \cite{hori-iqbal-vafa00} discuss mirror symmetry for Fano
837: varieties from a physics point of view. The construction of exact Morse
838: fibrations from singularities is similar to the definition of symplectic
839: monodromy in \cite[Section 6]{khovanov-seidel98}.
840: \end{additional}
841:
842: \section{Floer cohomology for automorphisms\label{sec:tqft2}}
843: %
844: \subsection{}
845: Let $(M,\o,\theta)$ be an exact symplectic manifold (which is, as always,
846: assumed to have contact type boundary). Floer cohomology associates to each
847: $\phi \in \Sympe(M)$ a pair of vector spaces $HF(\phi,-)$ and $HF(\phi,+)$.
848: These correspond to two ways of treating the fixed points near $\partial M$:
849: let $H \in \smooth(M,\R)$ be a function with $H|\partial M \equiv 0$, whose
850: Hamiltonian flow $(\phi_t^H)$ is equal to the Reeb flow on $\partial M$. To
851: define $HF(\phi,-)$ and $HF(\phi,+)$ one perturbs $\phi$ to $\phi \circ
852: \phi_t^H$ for some small $t<0$, respectively $t>0$. The difference is measured
853: by a long exact sequence
854: \begin{equation} \label{eq:boundary-term}
855: \xymatrix{
856: {HF(\phi,-)} \ar[r] & {HF(\phi,+)} \ar[r] & {H^*(\partial M;\Z/2).}
857: \ar@/^1.5em/[ll]_{ }
858: }
859: \end{equation}
860: For $\phi = \id_M$ this reduces to the usual cohomology exact sequence, with
861: $HF(\id_M,-) \iso H^*(M,\partial M;\Z/2)$ and $HF(\id_M,+) \iso H^*(M;\Z/2)$.
862: Other properties of $HF(\phi,\pm)$ are its invariance under isotopies within
863: $\Sympe(M)$; conjugation invariance, $HF(\phi_2\phi_1,\pm) \iso
864: HF(\phi_1\phi_2,\pm)$; and Poincar{\'e} duality, $HF(\phi,\pm) \iso
865: HF(\phi^{-1},\mp)^\vee$. If $M$ is equipped with a Maslov map $\delta_M$, a
866: grading $\tphi$ determines a $\Z$-grading on Floer cohomology, denoted by
867: $HF^*(\tphi,\pm)$. This satisfies $HF^*(\tphi[\sigma],\pm) \iso
868: HF^{*-\sigma}(\tphi,\pm)$ for $\sigma \in \Z$.
869:
870: Let $S$ be a closed oriented surface together with a finite set $\Sigma \subset
871: S$ of marked points, and $(E,\pi) = (E,\pi,\Omega,\Theta,J_0,j_0)$ an exact
872: Morse fibration over $S^* = S \setminus \Sigma$, with smooth fibres isomorphic
873: to $M$. Suppose that around each $\zeta \in \Sigma$ we have oriented local
874: coordinates $\psi_\zeta: D \rightarrow S$, such that there are commutative
875: diagrams
876: \begin{equation} \label{eq:local-structure}
877: \xymatrix{
878: {\Rleq \times T(\phi_\zeta)} \ar[d] \ar[rr]^-{\Psi_{\zeta}} &&
879: {E} \ar[d]^{\pi} \\
880: {D^*} \ar[rr]^-{\psi_{\zeta}|D^*} &&
881: {S^*.}
882: }
883: \end{equation}
884: Here $T(\phi_\zeta)$ is the mapping torus of $\phi_\zeta \in \Sympe(M)$,
885: meaning the quotient of $\R \times M$ by $(t,x) \sim (t-1,\phi_\zeta(x))$. The
886: map from $\Rleq \times T(\phi_\zeta)$ to $D^* = D \setminus \{0\}$ is
887: $(s,[t,x]) \mapsto \exp\,2\pi(s+it)$. The pullback of $\o$ to $\R \times M$ is
888: invariant under the $\Z$-action, hence descends to a two-form $\o_{\phi_\zeta}$
889: on $T(\phi_\zeta)$; and a choice of function $K \in \smooth_c(M \setminus
890: \partial M,\R)$ with $dK = \phi^*\theta - \theta$ yields a one-form
891: $\theta_{\phi_\zeta}$ on the same space. $\Psi_{\zeta}$ is an isomorphism of
892: differentiable fibre bundles between $\Rleq \times T(\phi_\zeta)$ and
893: $E|\psi_\zeta(D^*)$, satisfying $\Psi_\zeta^*\Omega = \o_{\phi_\zeta}$,
894: $\Psi_\zeta^*\Theta =\theta_{\phi_\zeta}$. By and large, this means that the
895: behaviour of $(E,\pi)$ around $\zeta$ is described by the exact symplectic
896: automorphism $\phi_\zeta$. One then has a relative invariant
897: \begin{equation} \label{eq:relative2}
898: \Phirel(E,\pi) \in \bigotimes_{\zeta \in \Sigma} HF(\phi_\zeta,\pm),
899: \end{equation}
900: where the signs $\pm$ can be chosen arbitrarily, subject to the restriction
901: that there must be at least one $\zeta$ labeled $+$ in each connected component
902: of $S$; in particular $\Sigma$ may not be empty. Changing an existing $-$ to a
903: $+$ gives a new relative invariant which is the image of the old one under the
904: map $HF(\phi_\zeta,-) \rightarrow HF(\phi_\zeta,+)$ from
905: \eqref{eq:boundary-term}. To define the relative invariant, one chooses a
906: suitable complex structure on $S^*$ and almost complex structure on $E$, and
907: considers the moduli spaces of pseudo-holomorphic sections.
908:
909: \begin{remark}
910: The asymmetry between $+$ and $-$ can be explained as follows. By definition of
911: an exact Morse fibration, one has a trivialization near $\partial E$. However,
912: the almost complex structures $J$ on $E$ used here do not have the obvious
913: product structure with respect to this trivialization; this is made necessary
914: by the perturbation of $\phi_\zeta$ near $\partial M$ which defines
915: $HF(\phi_\zeta,\pm)$. The convexity of $\partial E$ with respect to
916: $J$-holomorphic sections now becomes a more delicate matter, and it is there
917: that the sign question appears.
918: \end{remark}
919:
920: Suppose that $\zeta \neq \zeta'$ are points in $\Sigma$ with $\phi_{\zeta'} =
921: (\phi_{\zeta})^{-1}$, and look at a relative invariant $\Phirel(E,\pi)$ in
922: which $\zeta,\zeta'$ have different signs. The surface $\overline{S}$ obtained
923: by connect summing together $\zeta,\zeta'$ comes with marked points
924: $\overline{\Sigma}$ inherited from $\Sigma \setminus \{\zeta,\zeta'\}$, and
925: with an exact Morse fibration over $\overline{S} \setminus \overline{\Sigma}$.
926: The relative invariant of this new fibration (keeping the signs as before) can
927: be computed, if it is well-defined, by applying the Poincar{\'e} duality
928: pairing $HF(\phi_\zeta,\pm) \otimes HF(\phi_{\zeta'},\mp) \rightarrow \Z/2$ to
929: $\Phirel(E,\pi)$.
930:
931: This setup can be generalized and unified with that in \cite[Section
932: 3]{seidel00}. In the generalization one considers oriented compact surfaces $S$
933: with boundary, with marked points $\Sigma \subset S$ which may lie on the
934: boundary or in the interior, together with an exact Morse fibration $(E,\pi)$
935: over $S^* = S \setminus \Sigma$ having a Lagrangian boundary condition $Q$. The
936: structure of $(E,\pi)$ near a marked point in the interior remains as before,
937: while near boundary marked points $(E,\pi,Q)$ is as in \cite{seidel00}. One
938: then gets relative invariants of a mixed kind,
939: \begin{equation} \label{eq:mixed}
940: \Phirel(E,\pi,Q) \in \bigotimes_{\zeta \in \Sigma \cap \partial S}
941: HF(L_{\zeta,+},L_{\zeta,-}) \otimes \bigotimes_{\zeta \in \Sigma \setminus
942: \partial S} HF(\phi_\zeta,\pm).
943: \end{equation}
944: The restriction that there must be at least one $+$ now applies only to those
945: connected components of $S$ which have no boundary; and there are two different
946: gluing formulas, for boundary and interior marked points respectively.
947:
948: \subsection{}
949: We can now explain several kinds of maps between Floer cohomology groups as
950: special cases of the relative invariants \eqref{eq:relative2} and
951: \eqref{eq:mixed}.
952:
953: \begin{myitemize}
954: \item
955: Take an exact symplectic manifold $M$ and $\phi_1,\phi_2 \in \Sympe(M)$. Let $S
956: = S^2$ and $\Sigma = \{\text{\em 3 points}\}$. There is a unique exact Morse
957: fibration $(E,\pi)$ over $S^*$ which has no critical points and is even flat
958: (locally trivial), such that the symplectic monodromy around the three missing
959: points is respectively, $\phi_1$, $\phi_2$ and $(\phi_1\phi_2)^{-1}$. By
960: marking the points with suitable signs and using Poincar{\'e} duality, one
961: obtains from $\Phirel(E,\pi)$ a map $HF(\phi_1,\pm) \otimes HF(\phi_2,+)
962: \rightarrow HF(\phi_1\phi_2,\pm)$, which is the ``pair-of-pants'' product in
963: the Floer cohomology of symplectic automorphisms.
964:
965: \item
966: Let $L \subset M$ be an exact framed Lagrangian sphere. Then there is an exact
967: Morse fibration $(E,\pi)$ over $D$ with a single critical point, such that the
968: monodromy around $\partial D$ is $\tau_L$. If one extends this to $\C = S^2
969: \setminus \{\infty\}$ in such a way that it is flat near $\infty$, one gets an
970: invariant $\Phirel(E,\pi) \in HF(\tau_L^{-1},+)$.
971:
972: \item
973: Take the pair-of-pants product with $\phi_1 = \phi\tau_L$ and $\phi_2 =
974: \tau_L^{-1}$, and plug in the distinguished element of $HF(\tau_L^{-1},+)$
975: constructed above. This gives a canonical map $HF(\phi\tau_L,\pm) \rightarrow
976: HF(\phi,\pm)$, which we call $c$. Thanks to the gluing formula, $c$ can also be
977: described as the relative invariant of a certain exact Morse fibration over
978: $S^2 \setminus \{\text{\em 2 points}\}$ with a single critical point.
979:
980: \item
981: For any $\phi \in \Sympe(M)$, $\Rleq \times T(\phi^{-1})$ is an exact Morse
982: fibration over $D^*$. Denote by $(E,\pi)$ its restriction to $D \setminus
983: \{0,i\}$. Parallel transport along $\partial D$ starting at the base point $z_0
984: = -i$ yields a trivialization $E\,|\,\partial D \setminus \{i\} \iso (\partial
985: D \setminus \{i\}) \times M$. Given an exact Lagrangian submanifold $L \subset
986: M$, one can introduce the Lagrangian boundary condition which in this
987: trivialization is $(\partial D \setminus \{i\}) \times L$. Because of the
988: monodromy around $0$, $\Phirel(E,\pi,Q)$ takes values in $HF(\phi^{-1},\pm)
989: \otimes HF(\phi(L),L)$. Equivalently, one can see it as a map $d: HF(\phi,\pm)
990: \rightarrow HF(\phi(L),L)$.
991: \end{myitemize}
992:
993: \begin{theorem} \label{th:exact-sequence2}
994: Let $(M,\o,\theta)$ be an exact symplectic manifold, $L \subset M$ an exact
995: framed Lagrangian sphere, and $\phi \in \Sympe(M)$ an exact automorphism.
996: Suppose that $2c_1(M,L) \in H^2(M,L)$ is zero. Then there is a long exact
997: sequence, with maps $c$ and $d$ as defined above,
998: \begin{equation} \label{eq:sequence2}
999: \xymatrix{
1000: {HF(\phi\tau_L,\pm)} \ar[r]^-{c} &
1001: {HF(\phi,\pm)} \ar[d]^-{d} \\ &
1002: {HF(\phi(L),L).} \ar@/^1pc/[ul]
1003: }
1004: \end{equation}
1005: \end{theorem}
1006:
1007: If $M$ has a Maslov map $\delta_M$, and $\phi$, $L$ have gradings $\tphi$,
1008: $\tL$ then, with respect to the natural gradings of the groups in
1009: \eqref{eq:sequence2}, the maps $c,d$ have degree zero while the remaining one
1010: has degree one. The proof of the exact sequence is similar to that of
1011: \cite[Theorem 3.3]{seidel00}.
1012:
1013: \begin{example}
1014: By taking $\phi = \id_M$ resp.\ $\phi = \tau_L^{-1}$ one gets exact sequences
1015: \begin{align*}
1016: \xymatrix{
1017: {HF(\tau_L,+)} \ar[r]
1018: & {H^*(M;\Z/2)} \ar[d] \\ &
1019: {H^*(L;\Z/2),} \ar@/^1pc/[ul]
1020: }
1021: \qquad
1022: \xymatrix{
1023: {H^*(M;\Z/2)} \ar[r]
1024: & {HF(\tau_L^{-1},+)} \ar[d] \\ &
1025: {H^*(L;\Z/2).} \ar@/^1pc/[ul]
1026: }
1027: \end{align*}
1028: In the case when $M$ is a surface, this fits in with the results of direct
1029: computation in \cite{seidel96b}, which are that $HF(\tau_L,+) \iso
1030: H^*(M,L;\Z/2)$ and $HF(\tau_L^{-1},+) \iso H^*(M \setminus L;\Z/2)$. In fact,
1031: the same holds in all dimensions, as one can show by taking a closer look at
1032: the map $d$ in the exact sequence.
1033: \end{example}
1034: %
1035: \includefigure{skein2}{skein2.eps}{hb}%
1036:
1037: Now take an exact Morse fibration $(E,\pi)$ with arbitrary base $S$. To any
1038: loop $l: S^1 \rightarrow S \setminus S^\crit$ one can associate its symplectic
1039: monodromy $\rho_l$ and hence Floer cohomology groups, which are invariant under
1040: homotopies of $l$ that do not pass across $S^\crit$. We write $HF(l,\pm)$
1041: instead of $HF(\rho_l,\pm)$, in parallel with the notation $HF(c)$ adopted at
1042: the end of \cite[Section 3]{seidel00} for paths $c: [0;1] \rightarrow S$ with
1043: $c^{-1}(S^\crit) = \{0;1\}$. Theorem \ref{th:exact-sequence2} implies that if
1044: one has loops $l_-,l_+$ and a path $c$ as in Figure \ref{fig:skein2}, there is
1045: a long exact sequence
1046: \[
1047: \xymatrix{
1048: {HF(l_-)} \ar[r] & {HF(l_+)} \ar[r] & {HF(c).} \ar@/^1.5pc/[ll]^{ }
1049: }
1050: \]
1051: The connection with the original formulation is made by the Picard-Lefschetz
1052: theorem: $l_+$ differs from $l_-$ by a clockwise turn around $z \in S^\crit$,
1053: so that $\rho_{l_+}^{-1}\rho_{l_-}$ is the Dehn twist along the vanishing cycle
1054: arising from $z$.
1055:
1056: \section{Hochschild cohomology}
1057: %
1058: \subsection{}
1059: The Hochschild cohomology $HH^*(\A,\A)$ of a (small) $A_\infty$-category $\A$
1060: is a graded vector space, defined via the following cochain complex
1061: $CC(\A,\A)$. A cochain of degree $r$ is a sequence $h = (h^d)_{d \geq 0}$ of
1062: which each member $h^d$ is a family of graded linear maps
1063: \[
1064: \bigotimes_{i=1}^d hom_\A(X_i,X_{i+1}) \rightarrow hom_\A(X_1,X_{d+1})[r-d],
1065: \]
1066: one for each $(d+1)$-tuple $X_1,\dots,X_{d+1} \in \Ob\,\A$, satisfying
1067: $h^d(a_d,\dots,a_1) = 0$ whenever some $a_i$ is the identity map $\id_{X_i}$.
1068: In particular, $h^0$ consists of an element in $hom_\A^r(X,X)$ for each $X$.
1069: The differential is
1070: \[
1071: \begin{split}
1072: & (\partial h)^d(a_d,a_{d-1},\dots,a_1) = \\
1073: & \qquad
1074: = \sum_{i+j \leq d+1}
1075: \mu_\A^{d+1-j}(a_d,\dots,a_{i+j},h^j(a_{i+j-1},\dots,a_i),
1076: a_{i-1},\dots,a_1)
1077: \\
1078: & \quad \qquad
1079: + \sum_{i+j \leq d+1}
1080: h^{d+1-j}(a_d,\dots,a_{i+j},\mu_\A^j(a_{i+j-1},\dots,a_i),
1081: a_{i-1},\dots,a_1).
1082: \end{split}
1083: \]
1084: When $\A$ is directed, $CC(\A,\A)$ and hence $HH^*(\A,\A)$ are
1085: finite-dimensional.
1086:
1087: Hochschild cohomology has two related but different interpretations. One is in
1088: terms of first order deformations of the maps $\mu^d_\A$, and $\partial h = 0$
1089: appears there as the linearization of the structure equations of an
1090: $A_\infty$-category, extended by an $h^0$ term. The other involves the
1091: $A_\infty$-category $funct(\A,\A)$ of $A_\infty$-functors from $\A$ to itself,
1092: and it says that
1093: \begin{equation} \label{eq:functors}
1094: HH^*(\A,\A) = H(hom_{funct(\A,\A)}(\mathrm{Id}_\A,\mathrm{Id}_\A))
1095: \end{equation}
1096: is the space of $A_\infty$-natural transformations from the identity functor to
1097: itself. In the case where $\mu^d_\A = 0$ for $d \geq 3$, which means $\A$ is a
1098: dg-category, there is yet another approach which is essentially equivalent to
1099: \eqref{eq:functors} but more in line with classical homological algebra. Let
1100: $\A\!-\!mod\!-\!\A$ be the dg-category of dg-functors from $\A^{opp} \times \A$
1101: to chain complexes of vector spaces, and $D(\A\!-\!mod\!-\!\A)$ its derived
1102: category, defined by inverting quasi-isomorphisms as in \cite[Chapter
1103: 10]{bernstein-lunts94}. Take the dg-functor $\Delta_{\A} \in
1104: \Ob\,\A\!-\!mod\!-\!\A$ which assigns to $(X,Y) \in \Ob(\A^{opp} \times \A) =
1105: \Ob\,\A \times \Ob\,\A$ the complex $hom_\A(X,Y)$ (to motivate the notation, we
1106: should say that if $\A$ had only a single object $Z$, objects of
1107: $\A\!-\!mod\!-\!\A$ would be just dg-bimodules over the dg-algebra
1108: $hom_\A(Z,Z)$, and $\Delta_\A$ would be $hom_\A(Z,Z)$ considered as a bimodule
1109: over itself). Then
1110: \begin{equation} \label{eq:bimodule}
1111: HH^*(\A,\A) = \Hom^*_{D(\A\!-\!mod\!-\!\A)}(\Delta_\A,\Delta_\A).
1112: \end{equation}
1113: The proof goes by observing that the Hochschild complex coincides with
1114: $hom_{\A\!-\!mod\!-\!\A}(B_{\A}, \Delta_{\A})$ where $B_{\A}$, the bar
1115: construction, is quasi-isomorphic to $\Delta_{\A}$ and $K$-projective in the
1116: sense of \cite[Definition 10.12.2.1]{bernstein-lunts94}.
1117:
1118: \subsection{}
1119: An informal principle says that $HH^*(\A,\A)$ should be an invariant of
1120: $D^b(\A)$. It is unknown whether this is rigorously true, but there is a weaker
1121: result which is sufficient for our purpose.
1122:
1123: \begin{prop} \label{th:hochschild}
1124: Mutation of a directed $A_\infty$-category does not change its Hochschild
1125: cohomology.
1126: \end{prop}
1127:
1128: To prove this one goes through the list of moves \cite[Definition
1129: 5.2]{seidel00}. If $F: \A \rightarrow \B$ is a quasi-isomorphism, there are
1130: natural maps
1131: \[
1132: HH^*(\A,\A) \rightarrow H(hom_{funct(\A,\B)}(F,F)) \leftarrow HH^*(\B,\B),
1133: \]
1134: both of which can be shown to be isomorphisms by looking at the spectral
1135: sequence associated to the length filtration. Next, changing the gradings of
1136: the morphism groups $hom_\A(X^i,X^k)$ by some $\sigma_i-\sigma_k$ does not
1137: affect the Hochschild complex. Now consider $\A \rightsquigarrow c\A$. One can
1138: always find a directed dg-category ${\mathcal C}$ with a quasi-isomorphism
1139: ${\mathcal C} \rightarrow \A$, and then $c{\mathcal C}$ is quasi-isomorphic to
1140: $c\A$. Hence it is sufficient to prove $HH^*(c{\mathcal C},c{\mathcal C}) \iso
1141: HH^*({\mathcal C},{\mathcal C})$, which reduces the problem to the case of
1142: dg-categories. In that case there is a derived Morita equivalence between
1143: ${\mathcal C}$ and $c{\mathcal C}$ , which means that there are objects $P \in
1144: \Ob\,D(c{\mathcal C}\!-\!mod\!-\!{\mathcal C})$, $Q \in \Ob\,D({\mathcal
1145: C}\!-\!mod\!-\!c{\mathcal C})$ with
1146: \[
1147: P \otimes^{{\bf L}}_{{\mathcal C}} Q \iso \Delta_{c{\mathcal C}}, \quad
1148: Q \otimes^{{\bf L}}_{c{\mathcal C}} P \iso \Delta_{{\mathcal C}};
1149: \]
1150: for the definition of $\otimes^{{\bf L}}$ see again \cite[Chapter
1151: 10]{bernstein-lunts94}. The desired equality is a formal consequence of this
1152: and \eqref{eq:bimodule}. The argument for $\A \rightsquigarrow r\A$ is
1153: identical.
1154:
1155: \begin{additional}
1156: This section is essentially a review of known facts. $HH^*(\A,\A)$ is defined
1157: in \cite{penkava-schwarz95}; the $A_\infty$-categories $funct(-,-)$ in
1158: \cite{fukaya97}; and an account of derived Morita equivalences can be found in
1159: \cite{koenig-zimmermann98}. These are just sample references, and by no means
1160: the first ones historically.
1161: \end{additional}
1162:
1163: \section{Hochschild cohomology and global monodromy\label{sec:donaldson}}
1164: %
1165: \subsection{}
1166: What follows is my attempt to formulate an idea of Donaldson. Let $(E,\pi)$ be
1167: an exact Morse fibration over $D$ with a relative Maslov map $\delta_{E/D}$,
1168: $M$ the fibre over the base point with its induced Maslov map $\delta_M$, and
1169: $\mu \in \Sympe(M)$ the global monodromy, defined to be the symplectic parallel
1170: transport around $\partial D$ in positive direction. This comes with a
1171: canonical grading $\tilde{\mu}$, characterized by being zero near $\JJ_M \times
1172: \partial M$. After making an admissible choice of paths $(c_1,\dots,c_m)$, with
1173: corresponding distinguished basis $\Gamma = (L_1,\dots,L_m)$, one has that
1174: $\tilde{\mu}$ is isotopic to $\ttau_{L_1}\dots\ttau_{L_m}$ within the group of
1175: graded exact symplectic automorphisms. Choose gradings $\tL_i$ and let $\A =
1176: \Lag(\widetilde{\Gamma})$ be the directed Fukaya category. By Proposition
1177: \ref{th:hochschild}, $HH^*(\A,\A)$ is independent of the choice of $\Gamma$ and
1178: of the gradings $\widetilde{\Gamma}$.
1179:
1180: \begin{conjecture} \label{th:donaldson}
1181: There is a long exact sequence
1182: \begin{equation} \label{eq:donaldson}
1183: \xymatrix{
1184: {HF^*(\tilde{\mu},+)} \ar[r] &
1185: {H^*(E;\Z/2)} \ar[d] \\
1186: & {HH^*(\A,\A)} \ar[ul]
1187: }
1188: \end{equation}
1189: with the $\nwarrow$ map having degree one, and the others degree zero.
1190: \end{conjecture}
1191:
1192: It is elementary to check, using the classical Picard-Lefschetz formula, that
1193: the Euler characteristics add up in the right way. There is also an informal
1194: argument showing that, in a somewhat loose sense, $HF^*(\tilde{\mu},+)$ is
1195: ``constructed'' from nearly the same pieces as $HH^*(\A,\A)[-1]$, with
1196: $H^*(E;\Z/2)$ compensating for the small difference. We will now reproduce this
1197: argument, both because it seems to have been Donaldson's original motivation
1198: and because it uses nicely the two exact sequences in Floer cohomology. As in
1199: the discussion at the end of Section \ref{sec:tqft2}, it is convenient to use
1200: the notation $HF(l,\pm)$ and $HF(c)$ for Floer cohomology groups.
1201: %
1202: \includefigure{skd1}{skd1.eps}{hb}
1203:
1204: \begin{myitemize}
1205: \item[(i)]
1206: Let $l_0,\dots,l_m$ be the loops in $D \setminus D^{\crit}$ shown in Figure
1207: \ref{fig:skd1}. To each of them is associated a Floer group $HF(l_k,+)$. Our
1208: aim is to successively ``decompose'' $HF(l_m,+)$. To begin, by Theorem
1209: \ref{th:exact-sequence2} one has long exact sequences
1210: \[
1211: \xymatrix{
1212: {HF(l_k,+)} \ar[r] & {HF(l_{k-1},+)} \ar[r] &
1213: {HF(d_{1,k}),} \ar@/^1.5pc/[ll]^{ }
1214: }
1215: \]
1216: where the $d_{1,k}$ are as in Figure \ref{fig:skd2}. We interpret this as
1217: saying that $HF(l_m,+)$ is ``constructed'' by putting together the pieces
1218: $HF(l_0,+)$, $HF(d_{1,1})$, \dots, $HF(d_{1,m})$.
1219: %
1220: \includefigure{skd2}{skd2.eps}{hb}
1221:
1222: \item[(ii)]
1223: We next analyze $HF(d_{1,k})$ further. Take the paths $e_{i,k}$ and $e'_{i,k}$,
1224: $i<k$, from Figure \ref{fig:skd2}. \cite[Theorem 3.3]{seidel00} yields long
1225: exact sequences
1226: \[
1227: \xymatrix{
1228: {HF(d_{i,k})} \ar[r] & {HF(d_{i+1,k})} \ar[r] &
1229: {HF(e_{i,k}) \otimes HF(e'_{i,k}),} \ar@/^1.5pc/[ll]^{ }
1230: }
1231: \]
1232: so that $HF(d_{1,k})$ is ``built'' from $HF(d_{k,k})$ and $HF(e_{i,k}) \otimes
1233: HF(e'_{i,k})$, $1 \leq i < k$.
1234:
1235: \item[(iii)]
1236: A similar argument as in (ii) shows that $HF(e_{i,k}')$ can be ``decomposed''
1237: into $HF(e_{i,k})^\vee$ together with $HF(e_{i,j})^\vee \otimes HF(e_{j,k}')$
1238: for all $i < j < k$ (the dual vector space appears because the paths $e_{i,k}$,
1239: $e_{i,j}$ occur with reversed orientation). One can apply the same argument
1240: recursively to the second component $HF(e_{j,k}')$ of the tensor product, and
1241: this result in a different ``decomposition'' of $HF(e_{i,k}')$, of which the
1242: pieces are $(HF(e_{i_d,i_{d+1}}) \otimes \dots \otimes HF(e_{i_1,i_2}))^\vee$
1243: ranging over all $d \geq 1$ and $i = i_1 < i_2 < \dots < i_{d+1} = k$.
1244: \end{myitemize}
1245: %
1246: Putting all of this together yields a list of ``elementary pieces'' which make
1247: up $HF(l_m,+)$. At this point we revert to the conventional notation, so that
1248: $HF(l_m,+) = HF(\mu,+)$, $HF(l_0,+) = HF(\id_M,+) \iso H^*(M;\Z/2)$,
1249: $HF(d_{k,k}) = HF(L_k,L_k) \iso H^*(L_k;\Z/2)$, and $HF(e_{i,k}) =
1250: HF(L_i,L_k)$. We also restore the gradings, which were omitted up to now for
1251: the sake of simplicity. Then the ``elementary pieces'' are:
1252: \[
1253: \left\{
1254: \begin{split}
1255: & H^*(M;\Z/2); \\
1256: & (\Z/2)^m[-n\!-\!1]; \\
1257: & (\Z/2)^m[-1]; \text{ and} \\
1258: & HF^*(\tL_{i_1},\tL_{i_{d+1}})[-d\!-\!1] \otimes
1259: (HF^*(\tL_{i_d},\tL_{i_{d+1}}) \otimes \dots \otimes
1260: HF^*(\tL_{i_1},\tL_{i_2}))^\vee \\
1261: & \hspace{10em} \text{ for all $d \geq 1$ and $1\leq i_1 < \dots < i_{d+1} \leq m$.}
1262: \end{split}
1263: \right.
1264: \]
1265: The second and third piece are the result of splitting each $H^*(L_k;\Z/2)[-1]
1266: = \Z/2[-1] \oplus \Z/2[-n\!-\!1]$. Now $E$ is obtained up to homotopy
1267: equivalence by attaching $m$ cells of dimension $(n+1)$ to $M$. Hence the first
1268: two pieces can be ``put together'' again, in the same informal sense as before,
1269: to form $H^*(E;\Z/2)$. On the other hand the spectral sequence induced by the
1270: length filtration, which converges to $HH^*(\A,\A)[-1]$, has $E_1$ term equal
1271: to the direct sum of the remaining pieces (that is, all except the first two).
1272: Of course this sort of bookkeeping falls far short of a proof of Conjecture
1273: \ref{th:donaldson}, since it ignores all the maps in the long exact sequences,
1274: and equally the differentials in the spectral sequence.
1275:
1276: \subsection{}
1277: For a better understanding of Conjecture \ref{th:donaldson} one would need to
1278: know what the maps in it are. We outline here a possible description of one of
1279: them, namely $H^*(E;\Z/2) \rightarrow HH^*(\A,\A)$. This is pure speculation,
1280: supported mainly by the fact that the construction, apart from technicalities,
1281: seems natural.
1282:
1283: One may suppose that the $L_i$ are in generic position. Choose Morse functions
1284: $f$ on $M$ and $f_i$ on $L_i$, such that $f|L_i = f_i$ and the restriction of
1285: $f$ to a tubular neighbourhood of each $L_i$ is given by the sum of $f_i$ and a
1286: positive definite quadratic form on the normal bundle to $L_i$. Moreover,
1287: $f|\partial M = const.$ and the gradient should point outwards there. After
1288: choosing suitable Riemannian metrics, one gets Morse cohomology complexes
1289: $C_{Morse}(M)$ and $C_{Morse}(L_i)$ with $\Z/2$ coefficients, together with
1290: natural restriction maps $r_i: C_{Morse}(M) \rightarrow C_{Morse}(L_i)$.
1291: Enlarge $C_{Morse}(L_i)$ to a complex $\tilde{C}_{Morse}(L_i)$ by adding one
1292: generator in degree $-1$ whose boundary is the only zero-dimensional cocycle
1293: (the sum of all local minima of $f_i$). Composing $r_i$ and the inclusions
1294: $C_{Morse}(L_i) \hookrightarrow \tilde{C}_{Morse}(L_i)$ gives a map $\tilde{r}:
1295: C_{Morse}(M) \rightarrow \bigoplus_i \tilde{C}_{Morse}(L_i)$, whose mapping
1296: cone satisfies
1297: \begin{equation} \label{eq:morse-homology}
1298: H^*(\Cone(\tilde{r})) \iso H^*(E;\Z/2)[1].
1299: \end{equation}
1300: The point of using this rather strange model for the cohomology of $E$ is that
1301: there are natural maps
1302: \begin{equation} \label{eq:first-order}
1303: \tilde{C}_{Morse}(L_i) \rightarrow CC(\A,\A)[1], \quad C_{Morse}(M) \rightarrow
1304: CC(\A,\A).
1305: \end{equation}
1306: The first of these is a chain homomorphism; the second isn't, but the direct
1307: sum of the two is a chain map $\Cone(\tilde{r}) \rightarrow CC(\A,\A)[1]$,
1308: which in combination with \eqref{eq:morse-homology} gives the desired map
1309: $H^*(E;\Z/2) \rightarrow HH^*(\A,\A)$.
1310:
1311: Both maps in \eqref{eq:first-order} are constructed by the method of ``cutting
1312: down moduli spaces'' which is familiar from the definition of the cap product
1313: on Floer cohomology. We will describe the first map in detail, and the second
1314: one only briefly. Fix $i \in \{1,\dots,m\}$ and let $Z \subset L_i$ be the
1315: unstable manifold, under the gradient flow, of some critical point of $f_i$ of
1316: Morse index $p$. Suppose for the moment that $p>0$, so that $Z$ has positive
1317: codimension. Recall that the boundary operator $\mu_{\A}^1$ on $CF^*(\tL_{i_1},
1318: \tL_{i_2})$, $i_1<i_2$, is defined using pseudo-holomorphic maps $u: \R \times
1319: [0;1] \rightarrow M$ with boundary conditions $u(\R \times \{2-\nu\}) \subset
1320: L_{i_\nu}$. If $i_\nu = i$ for some $\nu \in \{1,2\}$, one can consider the
1321: subset of those maps $u$ which satisfy $u(0,2-\nu) \in Z$. If one assumes that
1322: the almost complex structures have been chosen generically (similar assumptions
1323: will be made tacitly later on) then counting isolated points in that subset
1324: yields a map
1325: \begin{equation} \label{eq:h1}
1326: h^1_Z: CF^*(\tL_{i_1},\tL_{i_2}) \rightarrow CF^{*+p}(\tL_{i_1},\tL_{i_2}).
1327: \end{equation}
1328: Extend this to the other Floer cochain groups in $\A$, that is to say to
1329: $CF^*(\tL_{i_1},\tL_{i_2})$ with $i_1,i_2 \neq i$, by setting $h_Z$ to be zero
1330: there. Now consider the composition maps of order $d \geq 2$, $\mu_\A^d:
1331: CF^*(\tL_{i_d},\tL_{i_{d+1}}) \otimes \dots \otimes CF^*(\tL_{i_1}, \tL_{i_2})
1332: \rightarrow CF^{*+2-d}(\tL_{i_1},\tL_{i_{d+1}})$ for $i_1 < \dots < i_{d+1}$.
1333: Their definition uses moduli spaces of pairs $(r,u)$, where $r \in \RR^{d+1}$
1334: is a point in the moduli space of discs with $d+1$ marked boundary point,
1335: $\SS^{d+1,*}_r$ is the corresponding disc with the marked points removed, and
1336: $u: \SS^{d+1,*}_r \rightarrow M$ is a map satisfying a pseudo-holomorphicity
1337: equation. $\partial \SS^{d+1,*}_r$ consists of intervals $I^{d+1}_{r,\nu}$, $1
1338: \leq \nu \leq d+1$, and the boundary conditions are
1339: \[
1340: u(I^{d+1}_{r,\nu}) \subset L_{i_\nu}.
1341: \]
1342: Assuming that $i_{\mu} = i$ for some $\mu$, one considers the moduli space of
1343: triples $(r,u,z)$, where $z$ is a point in $I^{d+1}_{r,\mu}$ such that $u(z)
1344: \in Z$. Counting isolated points in this moduli space defines a map
1345: \[
1346: h^d_Z: CF^*(\tL_{i_d},\tL_{i_{d+1}}) \otimes \dots \otimes CF^*(\tL_{i_1},
1347: \tL_{i_2}) \rightarrow CF^{*+p+1-d}(\tL_{i_1},\tL_{i_{d+1}}).
1348: \]
1349: Again, this is extended trivially to the cases when $i_1,\dots,i_{d+1} \neq i$.
1350: The sequence $(h_Z^0 = 0,h_Z^1,h_Z^2,\dots)$ is a Hochschild cochain $h_Z \in
1351: CC^{p+1}(\A,\A)$.
1352:
1353: The case $p = 0$, which we had excluded above, is considerably easier: one
1354: takes \eqref{eq:h1} to be the projection to the subspace spanned by the points
1355: of $L_{i_1} \cap L_{i_2}$ which lie on $Z$, and all other $h^d_Z$ to be zero.
1356: Finally there is a canonical element $\tilde{h} \in CC^0(\A,\A)$, defined by
1357: setting $\tilde{h}^0$ to be the unit element in $hom_\A^0(\tL_i,\tL_i) \iso
1358: \Z/2$, with all other components equal to zero. The assignment $Z \mapsto h_Z$,
1359: extended to $\tilde{C}_{Morse}(L_i)$ by mapping the generator of degree $-1$ to
1360: $\tilde{h}$, gives the first map in \eqref{eq:first-order}.
1361:
1362: The definition of the second map uses moduli spaces of triples $(r,u,z)$ with a
1363: point $z$ which can lie anywhere in $\SS^{d+1,*}_r$, such that $u(z)$ is in
1364: some stable manifold of $f$. The failure of this to be a chain map reflects the
1365: fact that $z$ can move to the boundary; in that case however $u(z)$ necessarily
1366: lies on some $L_i$, which provides the connection with the first map.
1367:
1368: \begin{remarks}
1369: (i) This discussion may seem rather abstract, but it has concrete computational
1370: implications, most strikingly for $n = 1$ when $M$ is a surface. In that case
1371: the $A_\infty$-category $\A$ is constructed from $\Gamma$ by a purely
1372: combinatorial count of immersed polygons, and the same is true for the maps
1373: \eqref{eq:first-order}. Suppose that Conjecture \ref{th:donaldson} is true and
1374: that our guess for the map $H^*(E;\Z/2) \rightarrow HH^*(\A,\A)$ which occurs
1375: in it is correct. Then one gets a combinatorial algorithm for computing the
1376: Floer cohomology $HF(\tau_{L_1}\dots\tau_{L_m},+)$ of an arbitrary product of
1377: Dehn twists in $M$ (along exact curves $L_i$ which admit gradings, to be
1378: precise).
1379:
1380: (ii) From a deformation theory point of view, what we have described is a first
1381: order infinitesimal deformation of $\A$ parame\-trized by the graded vector
1382: space $H^*(E;\Z/2)$. This was done by inserting one marked point into the
1383: Riemann surfaces which define $\mu^d_\A$. It seems natural to expect that using
1384: more marked points will allow one to extend this to higher order. This is
1385: interesting because in principle there are obstructions to such an extension,
1386: which are expressed by a natural graded Lie bracket on $HH^*(\A,\A)[1]$. If the
1387: extension to higher order is indeed possible, it would mean that the Lie
1388: bracket vanishes on the image of $H^*(E;\Z/2) \rightarrow HH^*(\A,\A)$.
1389: \end{remarks}
1390:
1391: \subsection{}
1392: To round off the discussion of Conjecture \ref{th:donaldson} we look at its
1393: implications in some specific cases. Suppose first that $(E,\pi)$ and
1394: $\delta_{E/D}$ come from a Lefschetz pencil satisfying
1395: \eqref{eq:relative-maslov} for some $a \in \Z$. Then $\mu$ can be isotoped to
1396: the identity but only within a group of symplectic automorphisms which is
1397: larger than $\Sympe(M)$, the isotopy being nontrivial along $\partial M$. Floer
1398: cohomology is not invariant under such isotopies. However, a more careful
1399: analysis shows that $HF(\mu,+) \iso HF(\id_M,-) \iso H^*(M,\partial M;\Z/2)$.
1400: More precisely, taking the canonical grading into account, this formula reads
1401: \[
1402: HF^*(\tilde{\mu},+) \iso H^{*+a-4}(M,\partial M;\Z/2).
1403: \]
1404: For $X = \CP{2}$, $\xi = {\mathcal O}(2)$, $\A$ has been determined in Section
1405: \ref{sec:examples}a, and one finds by explicit computation that $HH^r(\A,\A)
1406: \iso H^r(M,\partial M;\Z/2) \oplus H^r(E;\Z/2)$ for all $r$. This is compatible
1407: with an exact sequence \eqref{eq:donaldson} in which the horizontal arrow would
1408: be the zero map. There are reasons to believe that this arrow will in fact
1409: vanish for all Lefschetz pencils except the trivial one (the degree one pencil
1410: on $X = \CP{n+1}$).
1411: %
1412: \includefigure{mirror-monodromy}{mirmon.eps}{hb}
1413:
1414: Next consider the mirror dual of the projective plane, as in Section
1415: \ref{sec:examples}b. From the structure of the singular fibre $Y_\infty$ one
1416: sees that the global monodromy can be written as a product of Dehn twists and
1417: their inverses along disjoint curves, namely
1418: \[
1419: \mu = \prod_{i=1}^3
1420: (\tau_{B_i}\tau_{C_i}^{-1}\tau_{C_i'}^{-1}\tau_{C_i''}^{-1})
1421: \]
1422: where the curves are as shown in Figure \ref{fig:mirror-monodromy}. Using
1423: \cite{seidel96b} one computes that $HF^*(\tilde{\mu},+) = H^{*-2}(M \setminus
1424: \bigcup_{i=1}^3 C_i \cup C_i' \cup C_i'',\partial M;\Z/2)$. Comparing
1425: dimensions shows that $HH^r(\A,\A) \iso HF^{r+1}(\tilde{\mu},+) \oplus
1426: H^r(E;\Z/2)$, as in the previous example.
1427:
1428: As a third and final case, suppose that $(E,\pi)$ and $\delta_{E/D}$ come from
1429: a Morsification of an isolated hypersurface singularity. We conjecture that for
1430: all nontrivial singularities
1431: \begin{equation} \label{eq:acyclicity}
1432: HF(\mu) = 0;
1433: \end{equation}
1434: this can be checked easily in many cases, e.g.\ for weighted homogeneous
1435: singularities. From \eqref{eq:acyclicity} and Conjecture \ref{th:donaldson} it
1436: would follow that $HH^*(\A,\A) \iso H^*(E;\Z/2) = \Z/2$. Because of the role of
1437: Hochschild cohomology in deformation theory, this means that $\A$ is rigid: it
1438: admits no nontrivial first order deformations of any degree. As of now, there
1439: is no deeper explanation for this phenomenon.
1440:
1441: \begin{additional}
1442: For the cap product on Floer cohomology see \cite{le-ono97}, \cite{schwarz96}.
1443: Equation \eqref{eq:acyclicity}, if true, improves on old topological results of
1444: A'Campo \cite{acampo73b} and L{\^e} \cite{le74}.
1445: \end{additional}
1446:
1447: \section{Morse categories\label{sec:morse}}
1448: %
1449: \subsection{}
1450: Let $N^{n+1}$ be a compact manifold with codimension two corners. More
1451: precisely, we assume that $\partial N$ consists of three codimension one faces
1452: $\partial_- N$, $\partial_0 N$, $\partial_+ N$ (any of which may be empty) with
1453: $\partial_-N \cap \partial_+N = \emptyset$, and where the corners are
1454: $(\partial_- N \cup \partial_+N) \cap \partial_0 N$. Let $p: N \rightarrow
1455: [-1;1]$ be a Morse function satisfying $\partial_-N = p^{-1}(-1)$, $\partial_+N
1456: = p^{-1}(1)$, and such that $p|\partial_0N$, as a function on $\partial_0N$,
1457: has no critical points. In addition we require that if $x \in N$ is a critical
1458: point of $p$ with Morse index $i(x)$, then
1459: \begin{equation} \label{eq:self-indexing}
1460: \begin{cases}
1461: -1 < p(x) < 0 & \text{if $i(x) = 0$,} \\
1462: p(x) = 0 & \text{if $0<i(x)<n+1$,} \\
1463: 0 < p(x) < 1 & \text{if $i(x) = n+1$.} \end{cases}
1464: \end{equation}
1465: For $n = 1$ this is more or less the classical notion of self-indexing Morse
1466: function, but in higher dimensions it is far more restrictive. To any function
1467: $p$ with these properties one can associate a {\bf Morse category}
1468: $Morse(N,p)$. To do that, choose a Riemannian metric such that $\nabla p$ is
1469: parallel to $\partial_0N$. For two critical points $x,y$ of $p$, let ${\mathcal
1470: G}(x,y)$ be the space of unparametrized gradient trajectories going from $x$ to
1471: $y$. This is a smooth manifold; it is compact unless $i(x) = 0$ and $i(y) =
1472: n+1$, in which case there is a natural compactification to a manifold with
1473: boundary $\overline{\mathcal G}(x,y)$ such that
1474: \begin{equation} \label{eq:compactification}
1475: \partial\overline{\mathcal G}(x,y) = \bigsqcup_{0 < i(w) < n+1} {\mathcal
1476: G}(w,y) \times {\mathcal G}(x,w).
1477: \end{equation}
1478: These are all standard facts from Morse theory. What is important is that there
1479: is no need to choose the metric generically, since \eqref{eq:self-indexing}
1480: implies that $\nabla p$ is always Morse-Smale. A cobordism argument based on
1481: this fact shows that the topological type of the spaces ${\mathcal G}(x,y)$ and
1482: of their compactifications is independent of the metric.
1483:
1484: The objects of $\A = Morse(N,p)$ are the critical points of $p$, and the
1485: morphisms are $\hom_\A(x,y) = H_{-*}({\mathcal G}(x,y);\Z/2)$; note the sign in
1486: front of the grading. The composition maps
1487: \begin{equation} \label{eq:multiplication}
1488: H_{-*}({\mathcal G}(w,y)) \otimes H_{-*}({\mathcal G}(x,w)) \rightarrow
1489: H_{-*}({\mathcal G}(x,y))
1490: \end{equation}
1491: are as follows. If $x = w$ then $[point] \in H_0(point;\Z/2) = hom^0_\A(x,w)$
1492: acts as identity morphism, and similarly for $w = y$. If $i(x) = 0$, $0 < i(w)
1493: < n+1$, and $i(y) = n+1$, the composition comes from the inclusion of the
1494: boundary stratum \eqref{eq:compactification}, together with the fact that
1495: ${\mathcal G}(x,y) \hookrightarrow \overline{\mathcal G}(x,y)$ is a homotopy
1496: equivalence. In all remaining cases \eqref{eq:multiplication} is automatically
1497: zero, because the tensor product on the left hand side vanishes. Thus $\A$ is a
1498: $\Z/2$-linear and $\Z$-graded category. Ordering the critical points according
1499: to $p(x)$ shows that it can also be viewed as a directed $A_\infty$-category
1500: with $\mu^d_\A = 0$ for $d \neq 2$.
1501:
1502: Let $B(N,p) \subset N \setminus \partial N$ be the subspace of points whose
1503: gradient flow line exists for all time, which means that it converges to a
1504: critical point in both positive and negative direction. Define the fundamental
1505: object $\fund{B(N,p)} \in \Ob\,D^b(\A)$ to be the twisted complex
1506: \[
1507: (C,\delta_C) = \Big(\bigoplus_{dp(x) = 0} x[-i(x)], (\delta_{xy}) \Big),
1508: \]
1509: where $\delta_{xy} \in hom^1_\A(x[-i(x)],y[-i(y)])$ is the fundamental homology
1510: class $[{\mathcal G}(x,y)]$ if that space is compact and $x \neq y$, and zero
1511: otherwise. The generalized Maurer-Cartan equation reduces to
1512: $\mu^2_\A(\delta_C,\delta_C) = 0$, which is a consequence of
1513: \eqref{eq:compactification}. The notation $\fund{B(N,p)}$ is motivated by the
1514: fact that
1515: \begin{equation} \label{eq:sheafcohomology}
1516: \Hom^*_{D^b(\A)}(\fund{B(N,p)},\fund{B(N,p)}) \iso H^*(B(N,p);\Z/2).
1517: \end{equation}
1518: To see this one identifies $hom_{\Tw\,\A}(C,C)$ with the $E^1$ term of the
1519: spectral sequence, converging to $H^*(B(N,p);\Z/2)$, which arises from a
1520: decomposition of $B(N,p)$ determined by the gradient flow. There is only one
1521: nonzero differential in this spectral sequence, and that coincides with the
1522: differential on $hom_{\Tw,\A}(C,C)$, which completes the proof. If $N$ is
1523: closed, so that $B(N,p) = N$, the fundamental object $\fund{N}$ has another
1524: property. Namely, for any $X \in \Ob\,D^b(\A)$ the composition gives rise to a
1525: nondegenerate pairing
1526: \begin{equation} \label{eq:verdier}
1527: \Hom^*_{D^b(\A)}(X,\fund{N}) \otimes \Hom^{n+1-*}_{D^b(\A)}(\fund{N},X)
1528: \rightarrow
1529: \Hom^{n+1}_{D^b(\A)}(\fund{N},\fund{N}) \iso \Z/2.
1530: \end{equation}
1531:
1532: \begin{remarks}
1533: (i) Consider $B(N,p)$ as a space stratified by the stable manifolds of $\nabla
1534: p$. It was pointed out to me by Khovanov that the bounded derived category of
1535: sheaves of $\Z/2$-vector spaces, constructible with respect to this
1536: stratification, is equivalent to $D^b(\A)$; this appears to be a reformulation
1537: of a familiar result \cite{kapranov90}. The objects $x[-i(x)]$ correspond to
1538: the constant sheaves along the strata, and $\fund{B(N,p)}$ to the constant
1539: sheaf on all of $B(N,p)$. This explains the properties listed above:
1540: \eqref{eq:sheafcohomology} computes the cohomology of the constant sheaf, and
1541: \eqref{eq:verdier} is Verdier duality.
1542:
1543: (ii) The Morse categories for $p$ and $-p$ are not equivalent. However, they
1544: are Koszul dual, so that their derived categories are equivalent.
1545: \end{remarks}
1546:
1547: \subsection{}
1548: Assume now that $n = 1$, so that $p^{-1}(0)$ is a nodal curve on the surface
1549: $N$. Put a sign $\ominus,\oplus$ into each connected component of $N \setminus
1550: p^{-1}(0)$ that contains a local minimum respectively maximum of $p$. This
1551: diagram retains enough information about $p$ to reconstruct $\Morse(N,p)$.
1552: %
1553: \includefigure{triple-point}{triple.eps}{ht}%
1554:
1555: Let's say that two functions are related by a triple point move if the
1556: corresponding diagrams differ as shown in Figure \ref{fig:triple-point}. By a
1557: sequence of mutations that mimics the base changes in \cite[Volume\ 2, I \S 4]
1558: {arnold-gusein-zade-varchenko}, one can prove
1559:
1560: \begin{prop} \label{th:triple-point}
1561: Up to equivalence, a triple point move does not change the derived Morse
1562: category $D^bMorse(N,p)$.
1563: \end{prop}
1564:
1565: It would be interesting to find a topological interpretation of $HH^*(\A,\A)$
1566: for $\A = Morse(N,p)$. I have not succeeded in doing that, except for one
1567: special situation. Call $(N,p)$ cellular if all spaces ${\mathcal G}(x,y)$ are
1568: contractible (we are still assuming that $n = 1$; anyway cellularity is
1569: impossible in higher dimensions, except for trivial cases).
1570:
1571: \begin{prop} \label{th:cellular}
1572: If $(N,p)$ is cellular, $HH^*(\A,\A) \iso H^*(B(N,p);\Z/2)$.
1573: \end{prop}
1574:
1575: The first step in the proof is to note that the only possible nonzero groups
1576: are $HH^r(\A,\A)$ for $r = 0,1,2$. $HH^0(\A,\A) \iso H^0(B(N,p);\Z/2)$ is
1577: always true. From the cellularity one derives that $HH^1(\A,\A) \iso
1578: H^1(B(N,p);\Z/2)$, and also that the Euler characteristic of $HH^*(\A,\A)$
1579: equals that of $B(N,p)$, which completes the argument.
1580:
1581: \begin{additional}
1582: Objects similar to our Morse categories were introduced several years ago by
1583: Cohen, Jones and Segal \cite{cohen-jones-segal95} under the name flow
1584: categories. Proposition \ref{th:cellular} was inspired by A'Campo's paper
1585: \cite{acampo00}.
1586: \end{additional}
1587:
1588: \section{Real structures\label{sec:real}}
1589: %
1590: \subsection{}
1591: For the duration of this section, we extend the definition of exact Morse
1592: fibration by allowing each fibre to contain several critical points. This does
1593: not affect the theory seriously. The notion of admissible choice of paths
1594: remains the same; each path may give rise to several disjoint vanishing cycles,
1595: which one can place in arbitrary order in the distinguished basis. The
1596: important thing is that such bases are still unique up to Hurwitz moves.
1597:
1598: Let $(E,\pi) = (E,\pi,\Omega,\Theta,J_0,j_0)$ be an exact Morse fibration over
1599: $D$, with $j_0$ the standard complex structure on $D$, and $\delta_{E/D}$ a
1600: relative Maslov map. A real structure on it is an involution $\iota: E
1601: \rightarrow E$ such that
1602: \[
1603: \pi(\iota(x)) = \overline{\pi(x)},\;\; \iota^*\Omega = -\Omega,\;\;
1604: \iota^*\Theta = -\Theta,\;\; \iota^*J_0 = -J_0.
1605: \]
1606: There is always a $(j,J) \in \JJ_{E/D}$ with $j$ the standard complex structure
1607: on $D$ and such that $\iota^*J = -J$. Then the line bundle
1608: \begin{equation} \label{eq:delta}
1609: \Delta_{E/D} \,|\, \{(j,J)\} \times E = \Lambda^{n+1}(TE,J)^{\otimes 2} \otimes
1610: \pi^*(TD,j)^{\otimes -2}
1611: \end{equation}
1612: has a natural real involution $\iota_{(j,J)}$ which covers $\iota$. Call
1613: $\iota$ compatible with the relative Maslov map if $\delta_{E/D}\,|\,\{(j,J)\}
1614: \times E$ is homotopic to a trivialization of \eqref{eq:delta} which takes
1615: $\iota_{(j,J)}$ to complex conjugation in $\C$. This condition is independent
1616: of the choice of $(j,J)$.
1617:
1618: Given a real structure, one can consider $N = E^\iota$ with the Morse function
1619: $p = \pi|N: N \rightarrow [-1;1]$. $N$ is a manifold with corners of the kind
1620: considered in the previous section, with $\partial_-N \cup \partial_+N = (N
1621: \cap \partial_vE) = p^{-1}(\{\pm 1\})$ and $\partial_0N = N \cap \partial_hE$;
1622: and $p|\partial_0N$ has no critical points because the same holds for
1623: $\pi|\partial_hE$. We say that $\iota$ a good real structure if the critical
1624: points of $p$ satisfy \eqref{eq:self-indexing}, in which case there is a
1625: well-defined Morse category $Morse(N,p)$. We say that $\iota$ is complete if
1626: all critical points of $\pi$ lie on $N$.
1627:
1628: \begin{conjecture} \label{th:real}
1629: Let $(E,\pi)$ be an exact Morse fibration over $D$ with a good and complete
1630: real structure $\iota$, which is also compatible with a relative Maslov map
1631: $\delta_{E/D}$. Take an admissible choice of paths $(c_1,\dots,c_m)$, each of
1632: which lies in the lower half-disc $D \cap \{\im(z) \leq 0\}$, and let $\Gamma$
1633: be the corresponding distinguished basis. Then $\Lag(\widetilde{\Gamma})$, for
1634: a certain choice of grading $\widetilde{\Gamma}$, is quasi-isomorphic to
1635: $Morse(N,p)$.
1636: \end{conjecture}
1637:
1638: For $n = 1$ there is an elementary way of drawing the vanishing cycles which
1639: constitute $\Gamma$, starting from $(N,p)$; see e.g.\ \cite{acampo98}. Using it
1640: I have proved Conjecture \ref{th:real} in that dimension. There is also a
1641: general argument in support of it which, while falling short of a proof, is
1642: quite suggestive. Take an interval $[a;b] \subset [-1;1] = D \cap \R$ such that
1643: $[a;b] \cap D^\crit = \{a;b\}$. The symplectic parallel transport along $(a;b)$
1644: is $\iota$-equivariant, and hence preserves the real part $N$; in fact its
1645: restriction to $N$ is equal, up to reparametrization of the orbits, to the flow
1646: of $\nabla p$ with respect to some metric. Using property
1647: \eqref{eq:self-indexing} one can show that intersection points of the vanishing
1648: cycles arising from the endpoints $\{a,b\}$ are in one-to-one correspondence
1649: with gradient flow lines joining the corresponding critical points of $p$.
1650:
1651: \begin{remark}
1652: The combination of Conjecture \ref{th:real}, Proposition \ref{th:cellular} and
1653: Conjecture \ref{th:donaldson} leads to intriguing relations between the global
1654: monodromy of an exact Morse fibration on one hand, and its possible good and
1655: complete real structures on the other hand.
1656: \end{remark}
1657:
1658: Real structures are not the most general context in which there is a relation
1659: between directed Fukaya categories and Morse categories. Take an exact Morse
1660: fibration $(E,\pi)$ over $D$ with relative Maslov map $\delta_{E/D}$. A
1661: pseudo-real submanifold is a pair of embeddings
1662: \begin{equation} \label{eq:pseudo-real}
1663: \xymatrix{
1664: {N} \ar[d]_{p} \ar[r]^-{C} & E \ar[d]^{\pi} \\
1665: {[-1;1]} \ar[r]^{c} & D.
1666: }
1667: \end{equation}
1668: Here $c$ is a curve in $D$ with $c^{-1}(\partial D) = \{\pm 1\}$, meeting
1669: $\partial D$ transversally. $N$ is an $(n+1)$-dimensional manifold with corners
1670: as before. $C$ is an embedding transverse to $\partial_hE$ and $\partial_vE$,
1671: such that $C^{-1}(\partial_vE) = \partial_-N \cup \partial_+N$ and
1672: $C^{-1}(\partial_hE) = \partial_0N$; we require that $C^*\Omega$ is zero and
1673: $C^*\Theta$ is an exact one-form. $p$ is a Morse function such that
1674: $\partial_{\pm}N = p^{-1}(\pm 1)$. Furthermore, using the fact that $\im(C)$ is
1675: totally real with respect to $J$ for any $(j,J) \in \JJ_{E/D}$, one finds that
1676: there is a canonical homotopy class of sections of $S(\Delta_{E/D})$ over
1677: $\{(j,J)\} \times \im(C)$. If a section in that homotopy class can be chosen
1678: such that its composition with $\delta_{E/D}$ is a constant map $N \rightarrow
1679: S^1$, we say that our pseudo-real submanifold is compatible with the relative
1680: Maslov map. The notion of good pseudo-real submanifold is defined as before.
1681:
1682: \begin{conjecture}
1683: Suppose that $(N,p)$ is a good pseudo-real submanifold in $(E,\pi)$, compatible
1684: with $\delta_{E/D}$. Then $D^b\Lag(\widetilde{\Gamma})$ contains
1685: $D^bMorse(N,p)$ as a full triangulated subcategory. In particular there is an
1686: object $\fund{B(N,p)} \in \Ob\,D^b\Lag(\widetilde{\Gamma})$ whose endomorphism
1687: ring is $H^*(B(N,p);\Z/2)$.
1688: \end{conjecture}
1689:
1690: Again, it is not difficult to verify this for $n = 1$.
1691:
1692: \subsection{}
1693: As examples one can consider Lefschetz pencils as in Section
1694: \ref{sec:examples}a which are real, meaning that $X$ and $\xi$ have real
1695: involutions such that the composition of the two preserves $\sigma_0,\sigma_1$.
1696: Then the resulting exact Morse fibration has a natural real structure. To get
1697: the right kind of relative Maslov map, one must ask that the isomorphism
1698: \eqref{eq:relative-maslov} be compatible with the involutions on both sides. An
1699: example is $X = \CP{2}$, $\xi = {\mathcal O}(2)$, $\sigma_0(x) = x_0^2-x_2^2$,
1700: $\sigma_1(x) = x_0^2 + x_1^2 + x_2^2$, with complex conjugation. This gives
1701: rise to a complete and good real structure such that $(N,p)$ is $\RP{2}$ with
1702: its standard Morse function (having three critical points). One verifies easily
1703: that $Morse(N,p)$ is precisely given by the quiver with relations
1704: \eqref{eq:cp2-quiver}. Another possible application is to isolated hypersurface
1705: singularities which have real Morsifications. For instance, the Morsifications
1706: used in Section \ref{sec:examples}c for ADE plane curve singularities are real,
1707: and this provides another way of computing the associated categories.
1708:
1709: \begin{additional}
1710: The basic idea of relating vanishing cycles and real structures comes from work
1711: of A'Campo \cite{acampo75b} and Gusein-Zade \cite{gusein-zade74}.
1712: \end{additional}
1713:
1714: \section{Matching pairs and matching paths\label{sec:induction}}
1715: %
1716: \subsection{}
1717: Let ${\mathcal C}$ be a triangulated category, linear over $\Z/2$ and such that
1718: the spaces $\Hom^*_{\mathcal C}(X,Y)$ are finite-dimensional. Suppose that
1719: $(Y^1,\dots,Y^m)$ is a full exceptional collection in ${\mathcal C}$ such that
1720: for some $1 \leq i < m$, the following properties hold. Firstly
1721: \begin{equation} \label{eq:matching}
1722: \Hom^*_{\mathcal C}(Y^i,Y^{i+1}) = H^*(S^n;\Z/2)
1723: \end{equation}
1724: for some $n>0$. Secondly, if $a$ denotes the unique nontrivial morphism $Y^i
1725: \rightarrow Y^{i+1}$ of degree zero, the compositions
1726: \begin{align*}
1727: & a \circ -: \Hom^*_{\mathcal C}(Y^k,Y^i) \longrightarrow
1728: \Hom^*_{\mathcal C}(Y^k,Y^{i+1}) && \text{for $k<i$, and} \\
1729: & - \circ\, a: \Hom^*_{\mathcal C}(Y^{i+1},Y^l) \longrightarrow
1730: \Hom^*_{\mathcal C}(Y^i,Y^l) && \text{for $l>i+1$}
1731: \end{align*}
1732: are isomorphisms. We then call $(Y^i,Y^{i+1})$ a {\bf matching pair} of
1733: dimension $n$. Let $C \in \Ob\, {\mathcal C}$ be the cone of $a$, which is
1734: well-defined up to isomorphism.
1735:
1736: \begin{lemma} \label{th:matching-pair}
1737: $C$ is a spherical object of dimension $(n+1)$ in the sense of Definition
1738: \ref{def:spherical}.
1739: \end{lemma}
1740:
1741: For future use we need to mention one more fact about $C$. Assume that
1742: ${\mathcal C} = D^b(\A)$ is the derived category of some directed
1743: $A_\infty$-category $\A$, with $\Ob\,\A = \{X^1,\dots,X^m\}$, and moreover that
1744: $(Y^1,\dots,Y^m)$ is obtained from $(X^1,\dots,X^m)$ by mutation, that is to
1745: say by applying the transformations \cite[Equations (6) and (7)]{seidel00} and
1746: their inverses, plus possibly shifting each object by some amount. In this
1747: situation the following holds:
1748:
1749: \begin{lemma} \label{th:invariant}
1750: If $F: \A \rightarrow \A$ is a quasi-isomorphism which preserves the numbering
1751: of the objects, then $(D^bF)(C) \iso C$.
1752: \end{lemma}
1753:
1754: The proof is a general nonsense argument. Consider first a general triangulated
1755: category ${\mathcal C}$ as before with an exact self-equivalence $G$, and
1756: define $Inv(G) \subset \Ob\,{\mathcal C}$ to be the class of objects $X$ such
1757: that $G(X) \iso X$. This has the following properties: (i) if $X$ lies in
1758: $Inv(G)$ then so does $X[\sigma]$ for any $\sigma \in \Z$, as well as any
1759: object isomorphic to $X$; (ii) if $X,Y \in Inv(G)$ and there is a unique
1760: nonzero map $X \rightarrow Y$ of degree zero, then the cone over it lies in
1761: $Inv(G)$; (iii) if $X,Y \in Inv(G)$ then $T_X(Y)$, $T_X'(Y) \in Inv(G)$. Here
1762: $T_X(Y)$ is as in \cite[Section 5]{seidel00}, and $T_X'(Y)$ is the object which
1763: fits into an exact triangle
1764: \[
1765: T_X'(Y) \rightarrow Y \xrightarrow{ev^\vee} \Hom^*_{\mathcal C}(Y,X)^\vee
1766: \otimes X \rightarrow T_X'(Y)[1],
1767: \]
1768: where ${ev}^\vee$ is the transpose of the evaluation $ev: \Hom^*_{\mathcal C}
1769: (Y,X) \otimes Y \rightarrow X$. Now specialize to ${\mathcal C} = D^b(\A)$, $G
1770: = D^bF$. Then $X^i \in Inv(G)$, and by (iii) the same holds for the objects
1771: $Y^i$ of any mutated exceptional collection. Applying (ii) shows that $C \in
1772: Inv(G)$.
1773:
1774: \subsection{}
1775: Let $(E,\pi) = (E,\pi,\Omega,\Theta,J_0,j_0)$ be an exact Morse fibration over
1776: a compact base $S$, with $\dim\,E = 2n+2$. As explained in \cite[Section
1777: 3]{seidel00}, to any smooth path $c: [0;1] \rightarrow S$ with $c^{-1}(S^\crit)
1778: = \{0;1\}$ and $c'(0),c'(1) \neq 0$ one can associate two vanishing cycles
1779: $V_{c,0},V_{c,1}$, which are exact framed Lagrangian spheres in, say,
1780: $E_{c(1/2)}$. Now suppose that $c$ is an embedding, does not intersect
1781: $\partial S$, and that its vanishing cycles are isotopic, which means that
1782: there is a smooth family $I = (I_t)_{0 \leq t \leq 1}$ inside the space of all
1783: exact Lagrangian submanifolds of $E_{c(1/2)}$ joining $I_0 = V_{c,0}$ and $I_1
1784: = V_{c,1}$. Pairs $(c,[I])$ of this kind, where $[I]$ denotes the homotopy
1785: class of $I$ rel endpoints, are called {\bf matching paths}. The interest of
1786: this notion lies in a construction which I learnt from Donaldson. Roughly
1787: speaking (see below for a precise formulation) this produces from a matching
1788: path a submanifold of the total space $E$, which is Lagrangian in a suitable
1789: sense. The idea is easy to see in the case where $V_{c,0} = V_{c,1}$ and $I$ is
1790: the constant isotopy. By the definition of vanishing cycle, there are canonical
1791: $(n+1)$-dimensional balls $D_{c,\nu} \subset E$, $\nu = 0,1$, such that
1792: $\pi(D_{c,0}) = c([0;1/2])$, $\pi(D_{c,1}) = c([1/2;1])$, $\partial D_{c,\nu} =
1793: V_{c,\nu}$ and $\Omega|D_{c,\nu} = 0$. Under the assumption that $V_{c,0} =
1794: V_{c,1}$, the two balls fit together smoothly to give a closed submanifold of
1795: $E$. Fix a positive two-form $\beta$ on $S$. Then the base $\Omega^{(r)} =
1796: \Omega + r\cdot\pi^*\beta$ is symplectic for large positive $r$, and $D_{c,0}
1797: \cup D_{c,1}$ is Lagrangian with respect to it.
1798:
1799: In general, rather than choosing some $r$, it seems more natural to work with
1800: the entire family of forms $\Omega^{(r)}$. This requires a suitably adapted
1801: notion of Lagrangian submanifold. Let $\Lambda \subset \Rgeq \times (E
1802: \setminus \partial E)$ be a $(n+2)$-dimensional submanifold such that the
1803: projection $\Lambda \rightarrow \Rgeq$ is a proper submersion; this is the same
1804: as a smooth family of closed $(n+1)$-submanifolds $\Lambda^{(r)} \subset E$,
1805: $\Lambda^{(r)} = \Lambda \cap (\{r\} \times E)$. For simplicity assume that
1806: $H^1(\Lambda;\R) = 0$. Such a $\Lambda$ is called eventually Lagrangian if
1807: $\Omega^{(r)}|\Lambda^{(r)} = 0$ for $r \gg 0$. We can now state precisely the
1808: nature of the construction mentioned above: for any matching pair $(c,[I])$ it
1809: gives an eventually Lagrangian submanifold $\Lambda_{(c,[I])}$, unique up to
1810: isotopy.
1811:
1812: The first step is to trivialize $E$ over $c$, away from its endpoints, using
1813: symplectic parallel transport. The trivialization is the unique embedding
1814: $\Psi: (0;1) \times E_{c(1/2)} \rightarrow E$ which satisfies $\Psi(\{t\}
1815: \times E_{c(1/2)}) = E_{c(t)}$, $\Psi\,|\,\{1/2\} \times E_{c(1/2)} = \id$, and
1816: $\Psi^*\Omega = \Omega|E_{c(1/2)}$. Because $D_{c,0}$ and $D_{c,1}$ are
1817: likewise defined by parallel transport, they satisfy $\Psi^{-1}(D_{c,0}) =
1818: (0;1/2] \times V_{c,0}$, $\Psi^{-1}(D_{c,1}) = [1/2;1) \times V_{c,1}$. One can
1819: assume that the isotopy $(I_t)$ is constant near the endpoints. Then
1820: \[
1821: A = \{x_0\} \cup \{x_1\} \cup
1822: \bigcup_{t \in [0;1]} \Psi(\{t\} \times I_t),
1823: \]
1824: where $x_\nu \in E_{c(\nu)}$ are the critical points of $\pi$, is a closed
1825: submanifold of $E \setminus \partial E$ and contained in $\pi^{-1}c([0;1])$.
1826:
1827: By dualizing the normal vector field $\partial I_t/\partial t$ one gets a
1828: function $h_t$ (unique up to a constant) on each $I_t$. Extend the $h_t$ to
1829: functions $H_t$ on $E_{1/2}$ supported in a neighbourhood of $I_t$; these may
1830: be chosen such that $H_t = 0$ for $t$ close to $0$ or $1$. One can then find an
1831: $\alpha \in \Omega^1(E)$ vanishing near $\partial E$, which satisfies
1832: $\alpha\,|\,\ker(D\pi) = 0$ and $\Psi^*\alpha = H_t\,dt$. With respect to the
1833: modified symplectic forms $\Omega^{(r)} + d\alpha$, $r \gg 0$, $A$ is a
1834: Lagrangian submanifold. Moser's Lemma shows that $\Omega^{(r)} + d\alpha$ and
1835: $\Omega^{(r)}$ are diffeomorphic, and the diffeomorphism can be made to depend
1836: smoothly on $r$. Applying these diffeomorphisms to $A$ yields a smooth family
1837: of submanifolds of $E$, parametrized by $r \gg 0$, which are Lagrangian with
1838: respect to $\Omega^{(r)}$. One extends this arbitrarily to small values of $r$
1839: to complete the definition of $\Lambda_{(c,[I])}$.
1840:
1841: \begin{remarks}
1842: (i) We will often write $c$, $\Lambda_c$ instead of $(c,[I])$,
1843: $\Lambda_{(c,[I])}$, for the sake of brevity. This is fully justified only when
1844: $n = 1$: in that case each component of the space of exact Lagrangian circles
1845: on the surface $E_{c(1/2)}$ is simply-connected, so that $[I]$ contains no
1846: information.
1847:
1848: (ii) By definition each $\Lambda^{(r)}_c \subset E$, $r \in \Rgeq$, is
1849: diffeomorphic to $A$. Since $A$ is glued together from two balls, it is a
1850: homotopy $(n+1)$-sphere, but it may not be the standard one. To get standard
1851: spheres from the construction, one has to impose the additional condition that
1852: the composition
1853: \[
1854: \xymatrix{
1855: {S^n} \ar[r]^-{f_0} & {V_{c,0}} \ar[r] & {V_{c,1}} & \ar[l]_-{f_1} {S^n}
1856: }
1857: \]
1858: can be isotoped within $\Diff(S^n)$ to an element of $O(n+1)$. Here $f_0,f_1$
1859: are the framings of the vanishing cycles $V_{c,0},V_{c,1}$; and the middle
1860: arrow is the diffeomorphism determined, up to isotopy, by $(I_t)$.
1861: \end{remarks}
1862:
1863: Assume now that no connected component of $S$ is closed, so that $\beta \in
1864: \Omega^2(S)$ is exact. One then defines the Floer cohomology of two eventually
1865: Lagrangian submanifolds $\Lambda,\Lambda' \subset \R^{\geq 0} \times E$ by
1866: \begin{equation} \label{eq:eventually}
1867: HF(\Lambda,\Lambda') = HF(\Lambda^{(r)},(\Lambda')^{(r)})
1868: \end{equation}
1869: where the right hand side is with respect to $\Omega^{(r)}$, for some $r \gg
1870: 0$. This makes sense because $\Omega^{(r)}$ is exact, and because there is a
1871: natural class of compatible almost complex structures on $E$ such that
1872: pseudo-holomorphic curves do not reach out to $\partial E$. Moreover, one can
1873: show that the choice of $r$ does not affect \eqref{eq:eventually}.
1874:
1875: \subsection{}
1876: We will now explain the relation between the algebra and geometry introduced
1877: above. Let $(E,\pi)$ and $\delta_{E/D}$ be an exact Morse fibration over $D$
1878: with a relative Maslov map; the notation $M$ and $\delta_M$ is as usual. Let
1879: $c$ be a matching path. One can find an admissible choice of paths
1880: $(c_1,\dots,c_m)$ such that $c_i$ and $c_{i+1}$, for some $1 \leq i < m$, are
1881: as in Figure \ref{fig:matching}, and such that no other $c_j$ intersects $c$.
1882: In the corresponding distinguished basis $\Gamma = (L_1,\dots,L_m)$, $L_i$ and
1883: $L_{i+1}$ are isotopic. Now choose gradings $\widetilde{\Gamma} =
1884: (\tL_1,\dots,\tL_m)$ such that $\tL_i$ is isotopic to $\tL_{i+1}$, and let $\A
1885: = \Lag(\widetilde{\Gamma})$ be the directed Fukaya category. Then
1886: $(\tL_i,\tL_{i+1})$ is a matching pair in $D^b(\A)$. In fact one has
1887: \begin{align*}
1888: \Hom_{D^b(\A)}(\tL_i,\tL_{i+1}) & = H(hom_\A(\tL_i,\tL_{i+1})) \\ & =
1889: HF^*(\tL_i,\tL_{i+1}) \iso H^*(L_i;\Z/2)
1890: \end{align*}
1891: which proves \eqref{eq:matching}, and the other property is similarly easy. The
1892: informal idea is that the spherical object $C \in \Ob \, D^b(\A)$ associated to
1893: this matching pair ``represents'', in some sense, the eventually Lagrangian
1894: submanifold $\Lambda_c$. We will now make a concrete conjecture based on this
1895: philosophy.
1896: %
1897: \includefigure{matching}{matching.eps}{hb}%
1898:
1899: Suppose that $c,c'$ are two matching paths. For each of them one can find an
1900: admissible choice of paths of the kind considered above, but in general these
1901: choices cannot be the same for both paths. Therefore one gets two distinguished
1902: bases $\Gamma,\Gamma'$ of vanishing cycles, and two directed Fukaya categories
1903: $\A,\A'$ together with spherical objects $C \in \Ob \, D^b(\A)$ and $C' \in \Ob
1904: \, D^b(\A')$, constructed from matching pairs of objects in $\A,\A'$ in the way
1905: which we have explained. We know of course that $D^b(\A)$ and $D^b(\A')$ are
1906: equivalent, but for the present purpose we need to make this slightly more
1907: precise. There is a chain $\A = \A_0,\A_1,\dots,\A_r = \A'$ of directed
1908: $A_\infty$-categories, of which each is related to the preceding one in one of
1909: four possible ways: either $\A_{i+1}$ is quasi-isomorphic to $\A_i$; or it is
1910: obtained from $\A_i$ by shifting the grading of the morphism spaces; or else
1911: $\A_{i+1} = c\A$ or $r\A$. In the three last cases there are canonical
1912: equivalences $D^b(\A_i) \iso D^b(\A_{i+1})$; in the first case we choose an
1913: equivalence induced by an arbitrary quasi-isomorphism. Composing all of them
1914: gives an equivalence $G: D^b(\A) \rightarrow D^b(\A')$.
1915:
1916: \begin{conjecture} \label{th:induction}
1917: Let $\Lambda_c,\Lambda_{c'}$ be the eventually Lagrangian submanifolds
1918: associated to $c,c'$. Then
1919: \begin{equation} \label{eq:induction}
1920: HF(\Lambda_c,\Lambda_{c'}) \iso \bigoplus_{d \in \Z}
1921: \Hom^d_{D^b(\A')}(G(C),C').
1922: \end{equation}
1923: \end{conjecture}
1924:
1925: One can refine the statement by introducing suitable graded Floer cohomology
1926: groups for $(\Lambda_c,\Lambda_{c'})$ but we prefer to skip this, since it does
1927: not involve any really new ideas.
1928:
1929: The two admissible choices of paths which lead to $\Gamma,\Gamma'$ are not
1930: unique, and neither is the sequence of Hurwitz moves joining them; moreover, in
1931: the definition of $G$ we admitted arbitrarily chosen quasi-isomorphisms. It is
1932: implicitly part of the conjecture that the right hand side of
1933: \eqref{eq:induction} is independent of all these choices. The most
1934: doubtful-looking point is the bit about quasi-isomorphisms. Imagine for
1935: instance that $\A_i$ and $\A_{i+1}$ are directed Fukaya categories arising from
1936: the same Lagrangian configuration, but with different choices of almost complex
1937: structures. As mentioned in \cite[Section 6]{seidel00}, there is an explicit
1938: construction of quasi-isomorphisms between $\A_i$ and $\A_{i+1}$, using
1939: pseudo-holomorphic curves for a one-parameter family of choices of almost
1940: complex structures. One might think that for \eqref{eq:induction} to be
1941: correct, it would be necessary to choose the step between $D^b(\A_i)$ and
1942: $D^b(\A_{i+1})$ in the construction of $G$ to be such an analytically defined
1943: quasi-isomorphism. The answer to this objection is Lemma \ref{th:invariant},
1944: which says that choices of quasi-isomorphisms can never affect how objects like
1945: $C$ are mapped.
1946:
1947: To see what Conjecture \ref{th:induction} does, consider the case $n = 1$. Then
1948: $M$ is a surface, so that $\A$ and $\A'$ can be determined combinatorially, and
1949: so can $G: \A \rightarrow \A'$ (this is only true thanks to the freedom of
1950: choosing quasi-isomorphisms, which is why we were insisting on it). On the
1951: other hand, the left hand side of \eqref{eq:induction} is the Floer cohomology
1952: of Lagrangian two-spheres in the symplectic four-manifold $(E,\Omega^{(r)})$,
1953: which is difficult or even impossible to compute directly.
1954:
1955: \begin{remark} \label{th:zero-spheres}
1956: With some modifications, the discussion can be adapted to the trivial dimension
1957: $n = 0$. Here we only want to point out one example, namely the double branched
1958: cover $\pi: E \rightarrow D$ from Section \ref{sec:dimension-zero}c. Because
1959: the smooth fibre consists of two points, any path is a matching path, so that
1960: there is a multitude of one-dimensional spherical objects. This explains the
1961: geometric meaning of the objects $C_1,\dots,C_{2g} \in \Ob \, D^b(\A_g)$ which
1962: we wrote down earlier. By pushing the argument further, one can show that the
1963: zero-dimensional version of Conjecture \ref{th:induction} implies the algebraic
1964: formulae for geometric intersection numbers on $E$ which were derived, by a
1965: quite different argument, in \cite{khovanov-seidel98}.
1966: \end{remark}
1967:
1968: %
1969: %--------------------------- BIBLIOGRAPHY ---------------------------
1970: %
1971: \providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
1972: \begin{thebibliography}{VC}
1973: \bibitem[VC]{seidel00}
1974: P.~Seidel, \emph{Vanishing cycles and mutation}, Proceedings of the 3rd
1975: European Congress of Mathematics (Barcelona, 2000), Birkh{\"a}user, to appear.
1976:
1977: \bibitem{acampo00}
1978: N.~A'Campo, \emph{A combinatorial property of generic immersions of curves},
1979: Preprint.
1980:
1981: \bibitem{acampo73b}
1982: \bysame, \emph{Le nombre de lefschetz d'une monodromie}, Indag. Math.
1983: \textbf{35} (1973), 113--118.
1984:
1985: \bibitem{acampo75b}
1986: \bysame, \emph{Le groupe de monodromie du d{\'e}ploiement des singularit{\'e}s
1987: isol{\'e}es de courbes planes {I}}, Math. Ann. \textbf{213} (1975), 1--32.
1988:
1989: \bibitem{acampo98}
1990: \bysame, \emph{Real deformations and complex topology of plane curve
1991: singularities}, Annales Fac. Sci. Toulouse \textbf{8} (1999), 5--23.
1992:
1993: \bibitem{arnold-gusein-zade-varchenko}
1994: V.~I. Arnold, S.~M. Gusein-Zade, and A.~N. Varchenko, \emph{Singularities of
1995: differentiable maps}, Birkh{\"a}user, 1988.
1996:
1997: \bibitem{beilinson78}
1998: A.~Beilinson, \emph{Coherent sheaves on {$\mathbb{P}^n$} and problems of
1999: linear algebra}, Funct. Anal. Appl. \textbf{12} (1978), 214--216.
2000:
2001: \bibitem{bernstein-gelfand-ponomarev72}
2002: J.~Bernstein, I.~Gelfand, and V.~Ponomarev, \emph{Coxeter functors and
2003: {G}abriel's theorem}, Russian Math. Surveys \textbf{28} (1973), 17--32.
2004:
2005: \bibitem{bernstein-lunts94}
2006: J.~Bernstein and V.~Lunts, \emph{Equivariant sheaves and functors}, Lecture
2007: Notes in Mathematics, vol. 1578, Springer, 1994.
2008:
2009: \bibitem{bondal-polishchuk94}
2010: A.~Bondal and A.~Polishchuk, \emph{Homological properties of associative
2011: algebras: the method of helices}, Russian Math. Izvestiya \textbf{42} (1994),
2012: 219--260.
2013:
2014: \bibitem{cohen-jones-segal95}
2015: R.~L. Cohen, J.~D.~S. Jones, and G.~B. Segal, \emph{{F}loer's
2016: infinite-dimensional {M}orse theory and homotopy theory}, The {F}loer
2017: memorial volume (H.~Hofer, C.~Taubes, A.~Weinstein, and E.~Zehnder, eds.),
2018: Progress in Mathematics, vol. 133, Birkh{\"a}user, 1995, pp.~297--325.
2019:
2020: \bibitem{fukaya97}
2021: K.~Fukaya, \emph{Floer homology for three-manifolds with boundary {I}},
2022: Preprint, 1997.
2023:
2024: \bibitem{givental94}
2025: A.~Givental, \emph{Homological geometry and mirror symmetry}, Proceedings of
2026: the International Congress of Mathematics, Z{\"u}rich, vol.~1,
2027: Birkh{\"a}user, 1994, pp.~472--480.
2028:
2029: \bibitem{gusein-zade74}
2030: S.~M. Gusein-Zade, \emph{Dynkin diagrams for singularities of functions of two
2031: variables}, Functional Anal. Appl. \textbf{8} (1974), 295--300.
2032:
2033: \bibitem{hori-iqbal-vafa00}
2034: K.~Hori, A.~Iqbal, and C.~Vafa, \emph{{$D$}-branes and mirror symmetry},
2035: Preprint hep-th/0005247.
2036:
2037: \bibitem{kapranov90}
2038: M.~M. Kapranov, \emph{{M}utations and {S}erre functors on constructive
2039: bundles}, Functional Anal. Appl. \textbf{24} (1990), 155--156.
2040:
2041: \bibitem{khovanov-seidel98}
2042: M.~Khovanov and P.~Seidel, \emph{Quivers, {F}loer cohomology, and braid group
2043: actions}, Preprint Math.QA/0006056.
2044:
2045: \bibitem{koenig-zimmermann98}
2046: S.~Koenig and A.~Zimmermann (eds.), \emph{Derived equivalences for group
2047: rings}, Lecture notes in Math., vol. 1685, Springer, 1998.
2048:
2049: \bibitem{le74}
2050: D.~T. L{\^e}, \emph{La monodromie n'a pas de points fixes}, J. Fac. Sci. Univ.
2051: Tokyo Sect. IA \textbf{22} (1975), 409--427.
2052:
2053: \bibitem{le-ono97}
2054: H.~V. L{\^e} and K.~Ono, \emph{Cup-length estimates for symplectic fixed
2055: points},
2056: Contact and symplectic geometry (Cambridge, 1994), Cambridge Univ. Press,
2057: 1996, pp.~268--295.
2058:
2059: \bibitem{penkava-schwarz95}
2060: M.~Penkava and A.~Schwarz, \emph{{$A_\infty$}-algebras and the cohomology of
2061: moduli spaces}, Amer. Math. Soc. Translations, vol. 169, pp.~91--108,
2062: American Mathematical Society, 1995.
2063:
2064: \bibitem{schwarz96}
2065: M.~Schwarz, \emph{A quantum cup-length estimate for symplectic fixed points},
2066: Invent. Math. \textbf{133} (1998), 353--397.
2067:
2068: \bibitem{seidel96b}
2069: P.~Seidel, \emph{The symplectic {F}loer homology of a {D}ehn twist}, Math.
2070: Research Lett. \textbf{3} (1996), 829--834.
2071:
2072: \bibitem{seidel-thomas99}
2073: P.~Seidel and R.~Thomas, \emph{Braid group actions on derived categories of
2074: coherent sheaves}, Preprint math.AG/0001043. To appear in {\em Duke Math.\
2075: J}.
2076:
2077: \bibitem{wajnryb83}
2078: B.~Wajnryb, \emph{A simple presentation for the mapping class group of an
2079: orientable surface}, Israel J. Math. \textbf{45} (1983), 157--174.
2080: \end{thebibliography}
2081: \end{document}
2082: