1: \documentclass[11pt]{article}
2: \setlength{\textwidth}{14cm}
3: \setlength{\oddsidemargin}{1cm}
4:
5: \usepackage{amssymb,latexsym,epsfig,epsf,xypic}
6: \usepackage[mathscr]{euscript}
7:
8: \makeatletter \@addtoreset{equation}{section} \makeatother
9: \renewcommand{\theequation}{\thesection.\arabic{equation}}
10:
11: \newtheorem{Lemma}[equation]{Lemma}
12: \newtheorem{Theorem}[equation]{Theorem}
13: \newtheorem{Prop}[equation]{Proposition}
14: \newtheorem{Definition}[equation]{Definition}
15: \newtheorem{Corollary}[equation]{Corollary}
16: \newtheorem{Conj}[equation]{Conjecture}
17: \newenvironment{Proof}{\noindent\emph{Proof\ }}{\hfill$\square$\\}
18: \newenvironment{Remarks}{\noindent\textbf{Remarks}\ }
19:
20: \newcommand{\rt}[1]{\stackrel{#1\,}{\rightarrow}}
21: \newcommand{\Rt}[1]{\stackrel{#1\,}{\longrightarrow}}
22: \newcommand\A{\mathcal A}
23: \newcommand\db{\bar\partial}
24: \newcommand\R{\mathbb R}
25: \newcommand\C{\mathbb C\,}
26: \newcommand\Z{\mathbb Z}
27: \newcommand\Pee{\mathbb P}
28: \newcommand\PD{\mathrm{PD}\,}
29: \newcommand\OO{\mathscr O}
30: \newcommand\I{\mathcal I}
31: \newcommand\To{\longrightarrow}
32: \newcommand\ot{\leftarrow}
33: \newcommand\Ot{\longleftarrow}
34: \newcommand\into{\hookrightarrow}
35: \newcommand\res{\arrowvert_}
36: \newcommand\tr{\mathrm{tr}}
37: \newcommand\id{\mathrm{id}}
38: \newcommand\End{\mathrm{End}\,}
39: \newcommand\ip{{\mbox\,\_\hspace{-1.5mm}\shortmid\hspace{1.5pt}}}
40: \newcommand\ImO{\,\mathrm{Im}\,\Omega}
41:
42: \title{\textbf{Moment maps, monodromy and mirror manifolds}}
43: \author{R.\,P. Thomas}
44: \date{}
45:
46: \begin{document}
47:
48: \maketitle
49: \begin{abstract} \noindent
50: Via considerations of symplectic reduction, monodromy, mirror
51: symmetry and Chern-Simons functionals, a conjecture is proposed on
52: the existence of special Lagrangians in the
53: hamiltonian deformation class of a given Lagrangian submanifold of a
54: Calabi-Yau manifold. It involves a stability condition
55: for graded Lagrangians, and can be proved for the simple case of
56: $T^2$.
57: \end{abstract}
58:
59: \section{Introduction}
60: Just as explicit solutions of the Einstein and Hermitian-Yang-Mills
61: equations exist only on spaces that are either low dimensional,
62: non-compact and/or highly symmetric,
63: so the equations for special Lagrangian (SLag) cycles,
64: also important in physics, have the same properties.
65: Physically there are also similarities in that we have two first order
66: supersymmetric minimal energy equations (HYMs and SLag)
67: implying the more standard second order equations (YMs and minimal
68: volume equations).
69:
70: There are powerful existence results of Calabi and Yau (and more
71: recently Tian, Donaldson and others) for the Einstein equations, and of
72: Donaldson-Uhlenbeck-Yau for the HYM equations, so long as we are on a
73: K\"ahler (or projective) manifold; this often reduces
74: an infinite dimensional problem in PDEs to a finite dimensional
75: problem in linear algebra. Producing many K\"ahler-Einstein
76: (e.g. Calabi-Yau) manifolds becomes trivial, and dealing with
77: Hermitian-Yang-Mills connections requires only algebraic
78: computations; in both cases the complicated role of the K\"ahler form
79: and/or metric is almost removed. This can be thought of as possible
80: because of the existence of some infinite dimensional geometry recasting
81: the equations in terms of moment maps and symplectic reduction. A
82: similar situation for SLags would therefore be highly desirable.
83: In particular it might give a way of studying SLags using only
84: Lagrangians and symplectic geometry, much as HYM connections are
85: studied via stable bundles and algebraic geometry.
86:
87: This paper explores the mirror symmetry of holomorphic bundles
88: (on a Calabi-Yau $3$-fold $M$,
89: often referred to here as `the complex side') and Lagrangians
90: (on the mirror Calabi-Yau 3-fold $W$, 'the symplectic side', known
91: as the K\"ahler side in the physics literature). Many
92: people have worked and are still working on proving some kind of
93: direct correspondence between such objects given an
94: SYZ torus fibration \cite{SYZ}; see for example \cite{AP}, \cite{BBHM},
95: \cite{Ch}, \cite{Fu1}, \cite{Gr}, \cite{LYZ}, \cite{PZ}, \cite{Ty}, and
96: see \cite{MMM} for a review of this and many many more issues in mirror
97: symmetry. Here, however, we work purely formally without
98: reference to a particular pair of mirror manifolds, without
99: worrying about what mirror symmetry might rigorously mean, and we will
100: not try to give any explicit correspondence. Using mirror symmetry
101: merely as motivation, we point out some similar structures on
102: both sides of the mirror map. Under some conditions (in some
103: `large complex structure' or `semi-classical' or somesuch limit)
104: these structures might be genuinely dual; again it does not matter
105: if they are not in general. For instance, physics \cite{MMMS},
106: \cite{DFR} predicts that one should consider not the HYM equations and
107: slope but some perturbation of them away from the large complex
108: structure limit; however these equations also come from a moment map
109: and, conjecturally, a stability condition (for a discussion of such
110: matters see \cite{Le} or \cite{T3}). So while the slope and phase of
111: Lagrangians discussed below might not be exactly mirror to slope of
112: bundles, it should be mirror to something with analogous properties
113: and significance.
114:
115: Loosely, we would like to think of submanifolds in a fixed homology class
116: as mirror to connections on a fixed
117: topological complex bundle (with Chern classes mirror to the homology
118: class); then Lagrangians should correspond to holomorphic
119: connections (i.e. integrable connections; those with no
120: $(0,2)$-curvature) and special Lagrangians to those with HYM
121: curvature. These last two conditions should be stability conditions
122: for the group actions of hamiltonian deformations and complex gauge
123: transformations, respectively. The full picture is much more complicated,
124: involving triangulated categories and so forth, as envisaged some six
125: years ago in the seminal conjecture of Kontsevich \cite{K}; we can
126: ignore this in only using mirror symmetry
127: as motivation. It could be noted, however, that the functionals defined
128: below are additive under exact sequences of holomorphic vector bundles
129: and sums of Lagrangians, so should extend to the derived category
130: of coherent sheaves and the derived Fukaya category of Lagrangians
131: respectively.
132:
133: First note that while the connections side has a complex
134: structure and a complex gauge group involved, the Lagrangian side needs
135: complexifying. So motivated by Kontsevich \cite{K} and by physics
136: (e.g. \cite{SYZ}) we add in connections on the submanifolds (which
137: will later reduce to flat connections on Lagrangians). The dictionary
138: we are aiming towards, much of which is already standard, is the
139: following in the 3-dimensional case; all the terms used will be defined
140: in due course. \newpage
141:
142: \renewcommand\arraystretch{2}
143: $$
144: \begin{array}{|c|c|}
145: \hline \text{Complex side }M & \text{Symplectic side }W \\
146: \hline\hline \Omega=\Omega^{ }_M\in H^{3,0} & \omega=\omega^{ }_W\in
147: H^{1,1} \\ \hline
148: H^{\mathrm{ev}} & H^3 \\ \hline
149: \mathrm{Connections\ }A\mathrm{\ on\ a\ fixed\ }C^\infty\mathrm{\ complex}
150: & \mathrm{Submanifolds/cycles\ }L\mathrm{\ in\ a\ fixed
151: \ class} \vspace{-4mm} \\
152: \mathrm{bundle\ }E;\ \,v(E)=ch(E)\sqrt{\mathrm{Td\,}X}\in H^{\mathrm{ev}} &
153: [L]\in H^3,\mathrm{\ with\ a\ connection \ on\ }\C\times L \\ \hline
154: CS_\C(A=A_0+a)= \hspace{25mm} &
155: f_\C(A,L)=\int_{L_0}^L(F+\omega)^2 \hspace{2cm}\vspace{-3mm} \\
156: \hspace{6mm} \frac1{4\pi^2}\int_M\tr\left(\db_{A_0}a\wedge
157: a+\frac23a^{\wedge3}\right)\wedge\Omega &
158: \hspace{1cm}=\int_{L_0}^L(F^2+\omega^2)+2
159: \int_{L_0}^L\omega\wedge F \\ \hline
160: \mathrm{Critical\ points:\ }F_A^{0,2}=0 &
161: \mathrm{Critical\ points:\ }\omega\res L=0,\ F_A=0 \vspace{-4mm} \\
162: \mathrm{holomorphic\ bundles\ } &
163: \mathrm {Lagrangians\ +\ flat\ line\ bundles\ } \\ \hline
164: \text{Holomorphic Casson invariant \cite{T1}} &
165: \text{Counting SLags \cite{J}} \\ \hline
166: \mathrm{Gauge\ group} & U(1)\mathrm{\ gauge\ group\ on\ }L \\ \hline
167: \mathrm{Complexified\ gauge\ group} & \mathrm{Hamiltonian\
168: deformations} \\ \hline
169: \omega=\omega^{\ }_M\in H^{1,1} & \Omega=\Omega^{\ }_W\in H^{3,0} \\ \hline
170: \mathrm{Moment\ map\ \ }F_A\wedge\omega^{n-1} &
171: \mathrm{Moment\ map\ \,}\ImO\res L \\ \hline
172: \mathrm{Stability,\ slope\ }\mu=\frac1{\mathrm{rk\,}E}
173: \int\tr\,F_A\wedge\omega^{n-1} & \mathrm{Stability,\ slope\ }
174: \mu=\frac1{\mathrm{vol\,}(L)}\int_L\ImO \\ \hline
175: \end{array}
176: $$
177: (In the fourth line, $v(E)$ is the Mukai vector of $E$; in the last
178: line, vol$\,(L)$ is the cohomological volume measured with respect
179: to Re$\,\Omega$.) A SLag cycle (of phase $\phi$) in a Calabi-Yau
180: is a Lagrangian with Im\,$(e^{-i\phi}\Omega)\res L\equiv0$ \cite{HL};
181: then Re\,$(e^{-i\phi}\Omega)\res
182: L$ is the Riemannian volume form on $L$ induced by the Ricci-flat
183: metric on the Calabi-Yau. Obviously, rotating $\Omega$ by $e^{-i\phi}$
184: gives SLags in the more traditional sense of phase zero.
185: The part of the theory to do with SLags will apply in all dimensions;
186: it is only the functionals that are special to Calabi-Yau 3-folds.
187:
188: We will partially justify the above table, though the symplectic
189: structure and moment map give problems that will appear in due course.
190: However we can derive enough to arrive at
191: a conjecture about Lagrangians and SLags for which evidence will
192: be given by using monodromy and mirror symmetry from \cite{ST}
193: to interpret an example of Lawlor and Joyce \cite{J}. \\
194:
195:
196: \noindent \textbf{Acknowledgements.} The debt of any ex-student of
197: Simon Donaldson writing a paper on moment maps should be clear.
198: This work is also more immediately influenced by the papers
199: \cite{D},\,\cite{J},\,\cite{K}. In particular I was surprised to see the
200: Lagrangian condition coming from a moment map in \cite{D},\,\cite{H},
201: which does not fit into the
202: scheme I always supposed was true. So the purpose of this paper, apart
203: from trying to set a record for the number of m's in a title, is to
204: expand on that scheme and to try to get the \emph{special} condition
205: from a moment map instead. This paper was finished in the summer of 2000
206: and reported on in \cite{T3}; since then exciting new ideas have
207: appeared in physics \cite{Do} and mathematics \cite{KS} better explaining
208: mirror symmetry. I would like to thank S.-T. Yau, C. H.
209: Taubes and Harvard University for support, and Yi Hu,
210: Albrecht Klemm, Ivan Smith and Xiao Wei Wang for
211: useful conversations. Communications with Mike Douglas, Dominic Joyce,
212: Paul Seidel, S.-T. Yau and Eric Zaslow have been extremely influential.
213:
214:
215: \section{Chern-Simons-type functionals and critical points}
216:
217: Consider the space $\A$ of $(0,1)$-connections $A$ on a fixed complex
218: bundle on a Calabi-Yau 3-fold $M$. This infinite dimensional space
219: has a natural complex structure, with respect to which it admits a
220: holomorphic functional, Witten's holomorphic Chern-Simons functional
221: \cite{W1},\,\cite{DT},
222: $$
223: CS_\C(A=A_0+a)=\frac1{4\pi^2}\int_M\tr\left(\db_{A_0}a\wedge
224: a+\frac23a\wedge a\wedge a\right)\wedge\Omega,
225: $$
226: where $\Omega$ is the holomorphic (3,0)-form. It is infinitesimally
227: gauge-invariant (gauge transformations not homotopic to the identity
228: can give periods to $CS_\C$) and its gradient is $F_A^{0,2}$,
229: with zeros the integrable connections. That is, after dividing by
230: gauge equivalence (under which grad$\,CS_\C$ is invariant), the critical
231: points of $CS_\C$ form the space of holomorphic bundles of the same
232: topological type. As critical points of a functional, moduli of
233: holomorphic bundles have virtual dimension zero, and one might try to
234: make sense of counting them -- a holomorphic Casson invariant \cite{T1}.
235: This is independent of deformations of the complex structure, but can
236: have wall-crossing changes as the K\"ahler form varies. (This is because
237: we count only stable bundles, and the notion of stability depends on a
238: K\"ahler form.)
239:
240: On the other hand, on a different Calabi-Yau 3-fold $W$ (for instance
241: the mirror, in some situation where this makes sense),
242: Lagrangians are the critical points of a functional
243: too, on the space of all 3-dimensional submanifolds (or cycles):
244: $$
245: f_\R(L)=\int_{L_0}^L\omega\wedge\omega,
246: $$
247: where $\omega$ is the symplectic form on $W$. Here $L_0$ is a fixed cycle
248: in the same homology class, and we integrate over a 4-chain
249: with boundary $L-L_0$; the functional $f_\R$ is invariant under the choice
250: of different, homologous, 4-chains (picking non-homologous 4-chains can give
251: periods to $f_\R$). It is invariant under
252: deformations of $L$ pulled back from hamiltonian deformations of $W$
253: (deformations generated by vector fields $v$ on $W$ whose
254: contraction with $\omega$ is exact $v\ip\omega=dh$ at each point in
255: time) as $\int_L\omega\wedge dh=0$, and its
256: gradient is $\omega|_L$. Thus its critical points are Lagrangian
257: submanifolds. We would like to think of $f_\R$ as mirror
258: to $CS_\C$, but to do so we must complexify it.
259:
260: Thus we work on the space $\A$ of submanifolds $L$ of $W$ with $U(1)$
261: connections
262: $A$ on the trivial bundle $\C\times L$ on $L$. Notice these submanifolds
263: are not parameterised by a map of a real 3-manifold into $W$; we are only
264: interested in the image $L$. \emph{From now on we shall restrict
265: attention to smooth Lagrangian submanifolds.} Formally, we
266: consider the tangent space to $\A$ at a point $(A,L\subset W)$ to be
267: \begin{equation} \label{tsp}
268: \Omega^1(L;\R)\oplus \Omega^1(L;\R),
269: \end{equation}
270: at least for those $L$ with no $J$-invariant subspaces of its tangent
271: spaces ($J$ is the complex structure on $W$, and this is reasonable since
272: we are looking for Lagrangian submanifolds after all). The first factor
273: is the obvious tangent space to the connections on $L$; the second gives
274: deformations of $L$ via the vector fields produced by contracting with
275: the K\"ahler form $\omega$ on $W$. That is, we use the metric on $W$ to
276: map $\Omega^1(L)$ to $\Omega^1(W)|_L$, then use the isomorphism
277: provided by $\omega$ to get a vector field along $L$. Equivalently,
278: using the metric on $W$, we may think of one-forms on $L$ as tangent
279: vectors to $L$, then apply the complex structure $J$ on $W$ to give
280: $W$-vector fields on $L$. We denote this map from one-forms to normal
281: vector fields by
282: \begin{equation}\label{ip}
283: \Omega^1(L)\to TW\res L, \qquad \sigma\mapsto\sigma\ip\omega^{-1}.
284: \end{equation}
285: Connections on $L$ are carried along by the
286: vector field to connections on nearby cycles, and we are identifying
287: the space of $U(1)$ connections with $i\Omega^1(L;\R)$.
288:
289: There is a natural almost complex structure on $\A$, acting as
290: \renewcommand\arraystretch{1}
291: $$
292: J=\left(\begin{array}{cc} 0 & 1 \\ \!-1 & 0 \end{array}\right),
293: $$
294: with respect to the splitting (\ref{tsp}) of the tangent spaces. With
295: respect to this we claim to have the following holomorphic functional
296: $$
297: f_\C(A,L)=\int_{L_0}^L(F+\omega)^2=\int_{L_0}^L(F^2+\omega^2)\ +\ 2
298: \int_{L_0}^L\omega\wedge F.
299: $$
300: Here we have extended $A$ to a connection on the trivial bundle on the
301: whole of $W$ (restricting to a fixed connection $A_0$ on $L_0$, and to
302: $A$ on $L$) and taken its curvature form $F$. We have again picked
303: a 4-cycle bounding $L-L_0$; because $F$ and $\omega$ are closed the
304: resulting functional is independent of different homologous choices
305: of the 4-cycle, and in general well defined up to the addition of some
306: discrete periods. It is also (again) independent of hamiltonian isotopies
307: of $L$. Notice that the $\int_{L_0}^LF^2$ term is just the
308: \emph{real} Chern-Simons functional $CS_\R$ of the connection $A$ on $L$,
309: whose critical points are well known to be flat connections. As pointed
310: out to me by Eric Zaslow, the real and complex Chern-Simons
311: functionals already appear in \cite{W1} and \cite{Va} as possible
312: mirror partners (this is partially justified in \cite{LYZ}), but without
313: the terms in the symplectic form (and including instanton corrections
314: from holomorphic discs which we are ignoring for our rough analogy).
315: Asking for a real function to be equal to a complex one is possible
316: when one restricts attention to a real slice such as the space of
317: \emph{Lagrangian} submanifolds in $\A$; deforming within this space
318: the imaginary part of $f_\C$ remains constant and it reduces to
319: $CS_\R$. But allowing the imaginary counterparts to these real
320: deformations the right functional to consider is $f_\C$. Notice also
321: that if $\omega/2\pi$ is integral, so that we can pick a connection $B$
322: with curvature $-i\omega$, then the action functional
323: can be written in the more familiar looking Chern-Simons form
324: $$
325: f_\C(A,L)=\int_L(B+iA)\wedge d(B+iA)=\int_LCdC
326: $$
327: for the `complexified connection' $C=B+iA$ (a $\C^{\!\times}$-connection,
328: instead of a $U(1)$-connection.) This makes more contact with the
329: physics literature and allows one to extend the identification of
330: $CS_\R$ and $CS_\C$ in \cite{LYZ} to non Lagrangian sections, giving
331: complex values. Tian has informed me that he and Chen have also
332: considered the functional $f_\C(A,L)$ \cite{Ch}.
333:
334: Mirror symmetry should relate Lagrangians not just to bundles but the
335: whole derived category. For Riemann surfaces $C\subset M$, for instance,
336: there is a functional in \cite{DT}, \cite{W2} rather like $f_\R$
337: above:
338: $$
339: \int_{C_0}^C\Omega
340: $$
341: is formally holomorphic and has as critical points the \emph{holomorphic}
342: curves $C$. Similarly for four-manifolds $S\subset M$ with connections on
343: them the following functional (formally similar to $f_\C$)
344: $$
345: \int_{S_0}^S\tr\,F\wedge\Omega
346: $$
347: has critical points the holomorphic surfaces with flat connection on them.
348: Alternatively, as $CS_\C$ is additive under extensions of
349: bundles it does extend to the derived category. (Whether these two
350: approaches are compatible; i.e. whether or not the functional
351: associated to a curve or surface is the same as $CS_\C$ applied
352: to a locally free resolution of its structure sheaf, up to a constant,
353: seems to not have been worked out.)
354:
355: That $f_\C$ is holomorphic follows from the computation that the derivative
356: of $f_\C$ down $a\in\Omega^1(L)\oplus0$ (that only changes the connection
357: $A\mapsto A+\delta a$) is $\int_L2F\wedge ia+2\omega\wedge ia$,
358: while the derivative down $-Ja\in0\oplus\Omega^1(L)$, i.e. down the
359: vector field $a\ip\omega^{-1}$, is $\int_L2\omega\wedge a+2a\wedge F$.
360: The second expression is $-i$ times
361: the first, so the derivative is complex linear and $f$ is holomorphic.
362: Equivalently we are saying that $df_\C$ is the 2-form
363: $$
364: 2i\,(F+\omega)\ \oplus\ 2\,(F+\omega),
365: $$
366: which pairs with the tangent space (\ref{tsp}) by integration over $L$ to
367: give a form of type (1,0) on (\ref{tsp}).
368:
369: Thus critical points of the functional are Lagrangian cycles with flat line
370: bundles on them: exactly the basic building blocks of the objects proposed
371: in \cite{K} to be mirror dual to the holomorphic bundles that are the
372: critical points of $CS$. So this ties
373: in three well known moduli problems of virtual dimension zero (i.e. with
374: deformation theories whose Euler characteristic vanishes) -- flat bundles
375: on 3-manifolds, holomorphic bundles on Calabi-Yau 3-folds, and Lagrangians
376: (up to hamiltonian deformation) in symplectic 6-manifolds.
377:
378: So as mirror to \cite{T1} one would like to count Lagrangians
379: (up to hamiltonian deformations) plus flat line bundles on them, and
380: this is what Joyce's work \cite{J} has begun to tackle (in the rigid
381: case of $L$ being a homology sphere). Mirroring precisely the behaviour
382: of the holomorphic Casson invariant this count appears to be independent
383: of deformations of the K\"ahler form and to have wall-crossing changes
384: as the complex structure varies.
385:
386: \section{Gauge equivalence and moment maps}
387:
388: In fact what Joyce is proposing to count is \emph{special} Lagrangian
389: spheres with flat line bundles on them
390: (hence the otherwise anomalous dependence on the complex
391: structure), while \cite{T1} counts \emph{stable} bundles (i.e. by
392: Donaldson-Uhlenbeck-Yau, modulo the technicalities of polystable and
393: non-locally-free sheaves, we count \emph{Hermitian-Yang-Mills}
394: connections; hence the dependence on the K\"ahler form). (Tyurin
395: \cite{Ty} was perhaps the first to suggest
396: that the holomorphic Casson invariant should be related by mirror
397: symmetry to the real Casson invariant (here the $U(1)$ Casson
398: invariant) of SLag submanifolds.)
399:
400: The link should be, of course, that we want to consider holomorphic
401: connections on one side, up to complex gauge equivalence, and
402: Lagrangians on the other side, up to hamiltonian isotopy, and in both cases
403: we try to do this by picking distinguished representatives of equivalence
404: classes by the usual method of symplectic reduction.
405: Bringing in a K\"ahler structure on the complex side, we get a moment
406: map for the gauge group action, whose zeros give the HYM equations.
407: Dually, we would like to bring in the holomorphic 3-form on the
408: symplectic (K\"ahler) side, and get a complex group to act.
409: So again complexify by adding flat line bundles: consider the critical
410: points of the functional $f$ of the last section, i.e.
411: the space
412: $$
413: \mathcal Z=\{(L,A)\,:\,L\subset W\mathrm{\ is\ Lagrangian,\ }A
414: \mathrm{\ is\ a\ flat\ connection\ on\ }L\,\}
415: $$
416: (\emph{not} up to gauge equivalence). In fact consider this space
417: on a Calabi-Yau manifold $W$ of any dimension $n$. It has tangent space
418: $$
419: T_{(L,A)}\mathcal Z=Z^1(L)\oplus Z^1(L)
420: $$
421: ($Z^1(L)$ denotes closed real one-forms on $L$), the first being tangent
422: to the space of flat connections, the second giving
423: normal vector fields (by contracting with $\omega^{-1}$) preserving the
424: Lagrangian condition. We have an obvious almost complex structure
425: \begin{equation} \label{aha}
426: J=\left(\begin{array}{cc} 0 & 1 \\ -1 & 0 \end{array}\right).
427: \end{equation}
428:
429: Then the real group $C^\infty(L;\R)/\R$ acts as the Lie algebra to the
430: group of gauge transformations on
431: the flat line bundles (taking $d$ and adding to the connection)
432: whose complexification $C^\infty(L;\C)/\C$
433: acts complex linearly: the imaginary part $C^\infty(L;\R)/\R$
434: acts by hamiltonian deformations through the normal vector field
435: $$
436: h\mapsto dh\ip\omega^{-1}.
437: $$
438: Unfortunately, without using a metric this vector field is only
439: defined up to the addition of tangent vector fields to $L$; the map
440: (\ref{ip}) is really a map to $(TW\res L)/TL$ which we have lifted to
441: $TW\res L$ using the metric. (Equivalently we can extend
442: $h$ to a first formal neighbourhood of $L$ in different ways to get a
443: different vector field.) How we pick this alters how we carry the flat
444: connection along with $L$, and how the almost complex structure
445: (\ref{aha}) acts. For instance suppose we are in the rather artificial
446: case of $L$ being transverse to an SYZ $T^n$-fibration. Then
447: we can carry $L$ and the flat connection up the fibres
448: and identify the functions $C^\infty(L)$ from Lagrangian
449: to Lagrangian using the projection. Thus the group remains constant
450: as $L$ moves (effectively what we are doing is extending
451: functions from $L$ to a neighbourhood of $L$ in $W$ by pulling up
452: along the SYZ fibres). This does not work when $L$ branches over the
453: base of such a fibration. One can instead use the metric to define
454: normal vector fields, but then identifying the Lie algebra
455: $C^\infty(L)$ with a fixed $C^\infty(L_0)$ for all $L$ becomes difficult.
456:
457: This problem is perhaps not so surprising -- the moment the Lagrangian
458: has branching over the base of an SYZ fibration simple explicit
459: correspondences between Lagrangians and vector bundles (such as \cite{LYZ})
460: also break down due to our ignoring important holomorphic
461: disc instanton corrections that appear in the physics. For instance
462: recent work of Fukaya, Oh, Ohta and Ono \cite{FO3}, surveyed in \cite{Fu1},
463: show these provide the obstructions mirror to those of deformations
464: of holomorphic bundles \cite{T2} -- one should not in general consider
465: all (S)Lags (which are unobstructed) as mirror to holomorphic bundles,
466: but only those whose Floer cohomology (whose definition involves holomorphic
467: discs) is well defined.
468:
469: However, what is clear is the totality of the group action,
470: even if identifying individual elements causes problems, and this is
471: all we really need. For instance in the $K3$ (or $T^4$) case one can get
472: the same total group orbit, with a genuine
473: fixed group acting, by hyperk\"ahler rotating a construction due to
474: Donaldson \cite{D}. The end result is that
475: one considers parametrised Lagrangian embeddings $f$ from a Riemann
476: surface $L$ into the $K3$ such that the pullback of Re$\,\Omega$ is a
477: fixed symplectic form on $L$. Then the group of exact
478: symplectomorphisms of
479: $$
480: (L,f^*\mathrm{Re\,}\Omega)
481: $$
482: provide a symmetry group of the space of maps $f$, which also carries
483: a natural K\"ahler structure. Complexified orbits give hamiltonian
484: deformations, and the moment map is $m(f)=f^*\ImO$. The connection
485: with our construction is that after fixing a line bundle $\eta$
486: and connection with curvature $f^*\mathrm{Re\,}\Omega$, an infinitesimal
487: symplectomorphism $\phi$ induces a flat connection, via parallel transport
488: and pull back, on the bundle $\eta\otimes\phi^*\eta^*$. Globally the
489: action is different (this action has non-zero Lie bracket, for instance,
490: and a fixed group) but the total group orbit and the moment map
491: (see below) are the same.
492:
493: In general it is clear that the problem of identifying the group
494: for different embeddings of $L$ should be resolved by working with
495: the space of maps from a fixed $L_0$ to $W$, and enlarging the
496: group by including diffeomorphisms of $L_0$, giving a semi-direct
497: product of Diff$(L_0)$ and $U(1)$ gauge transformations on $L_0$.
498: Then the moment map for the diffeomorphism
499: part of the total group would be the Lagrangian condition as in
500: \cite{D}, and the problems we are encountering would come from
501: the fact that the group is a semi-direct product and not a product,
502: so that we cannot separate the two out and divide by them separately,
503: as in effect we have been trying to do. Unfortunately, I have not found the
504: correct formulation of the problem, but it is not so important for
505: follows.
506:
507: So we shall not worry too much about whether the complex structure
508: defined above is integrable, the group is fixed, or the symplectic
509: structure below is closed. In 1 complex dimension it is trivial, in
510: 2 we can use Donaldson's picture, and in 3 dimensions
511: we could either try to use an abstract SYZ fibration to deform and
512: identify Lagrangians transverse to it, or take everything
513: in this section as motivation for finding the stability condition for
514: Lagrangians of the next section.
515:
516: Fix a homology class of Lagrangians and multiply $\Omega$ by a unit
517: norm complex number so that $\int_L\ImO=0$.
518: We induce a symplectic structure on $\mathcal Z$ from $J$ and the
519: following metric on the tangent space
520: $$
521: \langle a,b\rangle=\int_La\wedge((b\ip\omega^{-1})\ip\ImO),
522: $$
523: for $a,\,b$ closed 1-forms. A computation in local coordinates
524: shows this is symmetric in $a$ and $b$; in fact it can be written as
525: \begin{equation} \label{lc}
526: \int_La\wedge(\widetilde b\ip\mathrm{\,Re\,}\Omega)=\int_L\cos\theta
527: \,(a\wedge*b),
528: \end{equation}
529: where $\,\widetilde{\ }\,$ is the isomorphism $T^*L\to TL$ set up by
530: the induced metric on $L$, $\Omega\res L=e^{i\theta}\mathrm{vol}_L$,
531: and vol$_L$ the Riemannian volume form on $L$ induced by the Ricci-flat
532: metric. Thus for Lagrangians with $\theta\in(-\pi/2,\pi/2)$, i.e.
533: those for which Re$\,\Omega$ restricts to a nowhere vanishing volume
534: form on $L$ and so are not too far from being SLag ($\theta\equiv0$),
535: this gives a non-degenerate metric.
536:
537: The symplectic form is invariant under the group action, and formally
538: the moment map is indeed $m(L,A)=\ImO$ in the dual $\Omega^n(L)_0$ of
539: the Lie algebra (i.e. $n$-forms on $L$ with integral zero). This follows
540: from the computation
541: $$
542: X\int_Lh\ImO=\int_Lh\,d\,(X\ip\ImO)=
543: \int_Ldh\wedge((\sigma\ip\omega^{-1})\ip\ImO),
544: $$
545: where $X=\sigma\ip\omega^{-1}$ is a normal vector field to the Lagrangian
546: $L$ down which we compute the derivative of the hamiltonian $\int_L
547: h\ImO=\langle m(L,A),h\rangle$ for the infinitesimal
548: action of $h$. Here have extended $h$ to a first-order neighbourhood
549: of $L\subset W$ so that it is constant in the direction of
550: $X=\sigma\ip\omega^{-1}$. Then the right hand side of the above
551: equation is the pairing using the symplectic form of $dh$ and $\sigma$,
552: as required.
553:
554: Infinitesimally we can see the moment map interpretation very
555: easily, and fitting naturally with the mirror bundle point of view.
556: Deformations of holomorphic connections $A$ modulo
557: complex gauge equivalence are given by a ker$\,\db_A$/im$\,\db_A$\,
558: first cohomology group, related to deformations
559: ker$\,\db_A\,\cap$\,ker$\,\db_A^*$ of the HYM equations
560: (modulo \emph{unitary} gauge transformations)
561: via Hodge theory, with the moment
562: map equation providing the $d^*=0$ slice to the imaginary part of the
563: linearised group action. Similarly, deformations of
564: Lagrangians are given by closed 1-forms
565: ker$\,d:\,\Omega^1(L;\R)\to\Omega^2(L;\R)$, so that dividing by hamiltonian
566: deformations we get
567: $$
568: H^1(L)=\mathrm{ker}\,d/\mathrm{im}\,d.
569: $$
570: If instead of dividing we impose the \emph{special}
571: condition, we get a ker$\,d^*$ slice
572: $$
573: H^1(L)=\mathrm{ker}\,d\cap\mathrm{ker}\,d^*,
574: $$
575: to the (imaginary) deformations (real deformations are given by
576: changing the flat $U(1)$ connection that can be incorporated into this).
577:
578:
579: \subsection*{A symplectic example}
580:
581: To motivate a guess at the correct definition of stability for
582: Lagrangians, we expand on an example of Lawlor and Joyce
583: (\cite{J} Sections 6 and 7, building on work of \cite{Ha}, \cite{L};
584: see also a similar example in \cite{SV} that is studied in
585: \cite{TY}), explaining its relevance to mirror symmetry, and giving a
586: simple example in algebraic geometry that mirrors it.
587:
588: First define the pointwise phase $\theta$ of a submanifold $L$: we may
589: write
590: $$
591: \Omega\res L=e^{i\theta}\mathrm{\,vol\,}
592: $$
593: where vol is the Riemannian volume form on $L$ induced by Yau's
594: Ricci-flat metric \cite{Y} on $W$. Thus vol provides a (local) orientation
595: for $L$, and reversing its sign alters the phase $\theta$ by $\pi$. A
596: SLag is a Lagrangian with \emph{constant} phase $\theta$.
597:
598: At first sight $\theta$ is multiply-valued; we always choose it to be a
599: fixed single-valued function to $\R$, lifting $e^{i\theta}:\,L\to S^1$
600: and thus providing the Lagrangian with a \emph{grading} as introduced by
601: Kontsevich \cite{K},\,\cite{S2}. Thus \emph{we only consider Lagrangians
602: of vanishing Maslov class} -- for a Calabi-Yau this is the winding class
603: $\pi_1(L)\to\pi_1(S^1)$ of the phase map
604: $$
605: L\Rt{e^{i\theta}}S^1,
606: $$
607: which of course vanishes for a SLag. (The definition of grading in
608: \cite{K}, \cite{S2} is topological and uses the universal
609: $\mathbb Z$-cover of the bundle of Lagrangian Grassmannians; here we
610: first pass to the $\mathbb Z/2$ orientation cover of the Grassmannian,
611: choosing an orientation of our Lagrangians,
612: and then use a complex structure to pass to the universal
613: $\mathbb Z$-cover of this. The two definitions
614: are of course equivalent.)
615:
616: Similarly we can define a kind of average phase $\phi=\phi(L)$ of a
617: submanifold (or homology class) $L\subset W$ by
618: $$
619: \int_L\Omega=A\,e^{i\phi(L)},
620: $$
621: for some real number $A$; we then use Re\,$(e^{-i\phi(L)}\Omega\res L)$
622: to orient $L$. Reversing the sign of $A$ alters the phase by $\pi$
623: and reverses the orientation. Again for a \emph{graded} Lagrangian
624: $L=(L,\theta)$, and \emph{we will always implicitly assume a grading},
625: $\phi(L)$ is canonically a real number (rather than $S^1$-valued).
626: Shifting the grading $[\,2n\,]:\,\theta\mapsto\theta+2n\pi$ gives a
627: similar shift to the phase $\phi(L)$.
628:
629: The terminology comes from the fact that if there is a submanifold
630: in the same homology class as $L$ that is SLag with respect to some
631: rotation of $\Omega$, then it is with respect to $e^{-i\phi(L)}\Omega$.
632: Slope, which we define as
633: $$
634: \mu(L):=\tan(\phi(L))=\frac1{\int_L\mathrm{Re\,}\Omega}\int_L\ImO,
635: $$
636: is defined independently of grading, is monotonic in $\phi$ in the
637: range $(-\pi/2,\pi/2)$, and is invariant under change of orientation
638: $\phi\mapsto\phi\pm\pi$. This agrees with the slope of a straight
639: line SLag in the case of $T^2$, as featured in \cite{PZ}, and we think
640: of it as mirror to the slope of a mirror sheaf, as is shown for tori
641: in \cite{PZ} (see \cite{DFR} for corrections in higher dimensions away
642: from the large complex structure limit).
643:
644: Joyce describes examples of SLags which we interpret as follows.
645: We have a family of Calabi-Yau 3-folds $W^t$ as $t$ ranges through (a
646: small open subset of) the moduli space of complex structures on $W$
647: \emph{with fixed symplectic structure}. That is, the holomorphic 3-form
648: $\Omega^t$ varies with $t$, but the K\"ahler form $\omega$ is fixed.
649: We also have a family of SLag homology 3-spheres
650: $L_1^t,\ L_2^t\subset W^t$ such that $L^t_1$ and $L^2_t$ intersect
651: at a point. If we choose a rotation of $\Omega^t$ such that $L_2^t$
652: always has phase $\phi_2^t\equiv0$ (this is possible locally at least;
653: in the family described later it will have to be modified slightly),
654: then we are interested as $t$ varies only in the complex number
655: $$
656: \int_{L_1^t}\Omega^t=R^te^{\phi^t_1}
657: $$
658: and its polar phase $\phi=\phi^t_1$; we plot this (i.e. the projection
659: from the complex structure moduli space to $\C$ via this map) in Figure 1.
660:
661: Then in Joyce's example, for $\phi<0$ (and $R^t>0$)
662: there is a SLag $L^t$ (of some phase $\phi^t$)
663: in the homology class $[L^t]=[L^t_1]+[L^t_2]$,
664: such that as $\phi\uparrow0$, this degenerates to a singular union
665: of SLags of the same phase $L^t=L^t_1\cup L^t_2$ and then disappears
666: for $\phi>0$.
667:
668: Most importantly, where $L^t$ exists as a smooth SLag ($\phi<0$) we have
669: the slope (and phase) inequality
670: \begin{equation} \label{ineq}
671: \mu^t_1<\mu^t_2, \qquad \mathrm{i.e.\ \,}\phi=\phi^t_1<\phi^t_2\equiv0;
672: \end{equation}
673: at $t=0,\ L^t$ becomes the singular union of $L^t_1$ and $L^t_2$,
674: with
675: $$
676: \mu^t_1=\mu^t_2\ (\phi^t_1=\phi^t_2);
677: $$
678: then there is no SLag in $L$'s homology class for
679: $$
680: \mu^t_1>\mu^t_2\ (\phi^t_1>\phi^t_2),
681: $$
682: \emph{though there is a Lagrangian}, of course -- the symplectic
683: structure has not changed. Though
684: we have been using slope $\mu$ in order to strengthen the analogy
685: with the mirror (bundle) situation, from now on we shall use only
686: the phase (lifted to $\R$ using the grading). While each is
687: monotonic in the other for small phase (as $\tan\phi=\mu$), slope
688: does not see orientation as phase does; reversing orientation adds
689: $\pm\pi$ to the phase but leaves $\mu$ unchanged. This is related to
690: the fact that we should really be working with complexes and so forth
691: on the mirror side (the bundle analogy is too narrow) and changing
692: orientation has no mirror analogue in terms of only stable bundles; it
693: corresponds to shifting (complexes of) bundles by one place
694: in the derived category. While slopes of bundles cannot go past
695: infinity (without moving degree in the derived category at least), for
696: Lagrangians they certainly can, and phase $\phi$ continues monotonically
697: upwards as its slope $\tan\phi$ becomes singular and then negative.
698:
699: Importantly, we can think of the various SLags as independent of time
700: when thought of as Lagrangians in the fixed symplectic manifold $W^t$:
701:
702: \begin{Lemma}
703: For $t>0$ the SLags $L^t$ are all in the same hamiltonian deformation
704: class. Similarly for $L_1^t,\ L^t_2$, and for $t<0$.
705: \end{Lemma}
706:
707: \begin{Proof}
708: Now choosing the phase of $\Omega^t$ such that $\phi(L^t)\equiv0$,
709: \begin{equation} \label{fred}
710: \int_L\frac{d}{dt}(\ImO^t)=\int_L\,\mathrm{Im}\,\dot\Omega^t=0.
711: \end{equation}
712: To show this deformation preserves the hamiltonian class of L,
713: we need to find a corresponding first order hamiltonian deformation
714: $dh\ip\omega^{-1}$ under which the change in $\ImO^t$,
715: $$
716: \mathcal L_{dh\_\hspace{-.4mm}\shortmid\hspace{1pt}\omega^{-1}}
717: (\ImO^t)\res L=d((dh\ip\omega^{-1})\ip\ImO^t)\res L,
718: $$
719: is $-$Im\,$\dot\Omega^t\res L$. But as Re\,$\Omega^t\res L$ is the induced
720: Riemannian volume form vol$^t$ on $L$, this means we want to solve
721: $$
722: -\mathrm{Im\,}\dot\Omega^t\res L=d(J(dh\ip\omega)\ip\mathrm{\,Re\,}
723: \Omega^t\res L)=d(\widetilde{dh}\ip\mathrm{\,vol}^t)=d(*dh)=\Delta(*h),
724: $$
725: where $J$ is the complex structure and $\,\widetilde{\ }\,$ is the
726: isomorphism $T^*L\to TL$ set up by the induced metric on $L$. So the
727: equation has a solution by the Fredholm alternative and (\ref{fred}).
728: \end{Proof}
729:
730: Thus for $\phi>0$ we consider the $L^t$s as the \emph{same} as
731: Lagrangian submanifolds (up to hamiltonian deformation) in the
732: \emph{fixed} symplectic manifold
733: $W^t$; it is only the SLag representative that changes as $\Omega^t$
734: varies. We think of this as mirror to a \emph{fixed} holomorphic bundle
735: in a fixed complex structure, with varying HYM connection as the mirror
736: K\"ahler form changes.
737:
738: \begin{Lemma} \label{lem} In the analogous 2-dimensional situation of
739: SLags in a $K3$ or abelian surface, the obstruction does not occur.
740: \end{Lemma}
741:
742: \begin{Proof}
743: Choose a real path of complex structures $W^t,\
744: t\in(-\epsilon,\epsilon)$ in complex structure moduli space such that
745: there is a nodal SLag $L^0=L_1^0\cup L_2^0$ in $W^0$. Without loss of
746: generality we can choose the
747: phase of $\Omega^t$ so that both $\omega$ and Im$\,\Omega^t$ pair to
748: zero on the homology class of $L^0$. Now hyperk\"ahler rotate the
749: complex structures so that instead the new Re$\,\Omega^t$ and
750: Im$\,\Omega^t$ pair to zero on the homology class of $L^0$ for all
751: $t$. $L^0$ is now a nodal holomorphic curve $C$ in the central $K3$.
752: We can understand deformations of $C$ via
753: deformations of the ideal sheaf $\mathcal J_C$, with obstructions in
754: $$
755: \mathrm{Ext}^2(\mathcal J_C,\mathcal J_C)\to H^{0,2}(W)\cong\C,
756: $$
757: where the arrow is the trace map and is an isomorphism by Serre
758: duality. Standard deformation theory shows the obstruction is
759: purely cohomological -- it
760: is the derivative of the $H^{0,2}$-component of the class
761: $$
762: [C]\in H^2(W)\cong H^{2,0}(W)\oplus H^{1,1}(W)\oplus H^{0,2}(W).
763: $$
764: But we have fixed this to remain zero by the phase condition,
765: so the curve deforms to all $t$ (really we should assume the family
766: is analytic in $t$ here and extend to $t\in\C$, or just work with
767: first order deformations). Hyperk\"ahler unrotating gives back a
768: family of SLags.
769: \end{Proof}
770:
771: There is a notion of connect summing Lagrangian submanifolds intersecting
772: in a single point (probably due to Polterovich) -- see for instance
773: Appendix A of \cite{S1} -- which
774: we claim gives the smoothings $L^t$ of the singular $L^0=L_1\cup L_2$.
775: This follows by comparing the local models \cite{J}, \cite{S1} for
776: the Lagrangians; see \cite{TY} where it is studied in more detail for
777: a related purpose, and our conventions (slightly different from those
778: of \cite{S2}) are described. While
779: topologically we are just connect summing $L_1$ and $L_2$ by removing a
780: small 3-ball containing the intersection point from each and gluing the
781: resulting boundary $S^3$s (there are two ways, depending on orientation),
782: symplectically the construction does not explicitly use orientations of
783: the submanifolds. (Effectively we are using their relative orientation --
784: the canonical orientation of the sum of the tangent spaces of $L_1,\ L_2$
785: at the intersection point given by the symplectic structure.)
786:
787: Giving $L_1$ and $L_2$ in that order produces a Lagrangian, well
788: defined up to hamiltonian isotopy (this will be shown in Section
789: \ref{Ko} in more generality; see (\ref{proj})),
790: $$
791: L_1\#L_2,
792: $$
793: with the singular union $L_1\cup L_2$ a limit point in the hamiltonian
794: isotopy class, which is \emph{not} itself hamiltonian isotopic
795: to $L_1\#L_2$ (we have seen a family of hamiltonian
796: deformations which has limit $L_1\cup L_2$, but the deformations
797: are singular at this limit).
798:
799: There is also an obvious notion of \emph{graded connect sum},
800: which is in fact what we shall always mean by $\#$. There is a unique
801: grading on $L_1$ compatible with a fixed grading on $L_2$ such that
802: we can give a (continuous) grading to the smoothing $L_1\#L_2$. In the
803: case of connect summing at multiple intersection points (Section
804: \ref{Ko}) there is \emph{at most} one such grading; in general the
805: graded connect sum may not exist.
806:
807: In $n$ dimensions, if $L_1$ and $L_2$ are graded such that $L_1\#L_2$
808: exists, then on reversing the order of the $L_i$, the graded sum that
809: exists is
810: \begin{equation} \label{+-}
811: L_2\#(L_1[2-n]) \quad\text{in the homology class}\quad
812: [L_2]+(-1)^n[L_1].
813: \end{equation}
814: Here $L[m]$ means the graded Lagrangian $L$ with its grading changed by
815: adding $m\pi$ to $\theta$, and the homology class of $L_1\#L_2$ is
816: $[L_1]+[L_2]$ using the orientations on the $L_i$s induced by the gradings.
817:
818: This is closely related, as we shall see, to Joyce's obstruction,
819: and the lack of it in dimension 2 (Lemma \ref{lem}). In 2 dimensions,
820: $L_1\#L_2$ and $L_2\#L_1$ are in the same homology class, though
821: by a result of Seidel \cite{S1} \emph{not in general in the same hamiltonian
822: isotopy class},
823: $$
824: L_1\#L_2\not\approx L_2\#L_1,
825: $$
826: importantly (we use $\approx$ to denote equivalence up to hamiltonian
827: deformations). For $t>0$ in the above family $L^t$ is in the constant
828: hamiltonian deformation class of $L_1\#L_2$, for $t<0$ it is in the
829: different class of $L_2\#L_1$, and at $t=0$ it is $L_1\cup L_2$ -- in
830: neither class but in the closure of both. (For
831: complex $t$ the symplectic structure is no longer constant like it is
832: for $t\in\R$, as one can see by following through the hyperk\"ahler
833: rotation; thus we do not get a contradiction to the above statement
834: by going round $t=0$ in $\C$.) In 3 dimensions, however, the
835: corresponding obvious choice for a SLag on the other side of the
836: $\pi^t_1=0$ wall, $L_2\#L_1[-1]$, is in the wrong homology class.
837:
838: In the case that the $L_i$ are Lagrangian spheres we can see this
839: by going round the wall
840: \begin{equation} \label{wall}
841: \phi(L_1^t)=0\simeq\phi(L_2^t),
842: \end{equation}
843: and using monodromy. In the 2-dimensional $K3$ or $T^4$ case this
844: works as follows.
845:
846: \begin{figure}[h]
847: \vspace{4mm}
848: \center{
849: \input{k3.pstex_t}
850: \caption{$\left(\int_{L_1}\Omega\right)$-space, as $\Omega$
851: on $K3$ varies, with polar coordinates $(R,\,\phi(L_1))$}}
852: \vspace{2mm}
853: \end{figure}
854:
855: Consider a disc in complex structure moduli space over which the
856: family of K\"ahler $K3$ surfaces (with constant K\"ahler form)
857: degenerates at the origin to a $K3$ with an ordinary double point
858: (ODP) with the Lagrangian $L_1\cong S^2$ as vanishing cycle.
859: A local model is the standard K\"ahler structure on $x^2+y^2+z^2=u$,
860: over the parameter $u$ in the unit disc in $\C$. Now base-changing
861: by pulling back to the double cover in $u$, $u\mapsto u^2$, we
862: get the 3-fold
863: $$
864: x^2+y^2+z^2=u^2,
865: $$
866: with a 3-fold ODP which has a small
867: resolution at the origin putting in a holomorphic sphere resolving
868: the central $K3$ fibre $u=0$. Choosing a nowhere-zero holomorphic
869: section $\Omega_u$ of the fibrewise $(2,0)$-forms
870: (using the fact that the relative canonical bundle of either family is
871: trivial), this restricts to zero on the exceptional $\Pee^1$ (which is
872: homologous to the vanishing cycle $L_1$). Therefore the function
873: \begin{equation} \label{zero}
874: \int_{L_1}\Omega_u
875: \end{equation}
876: \emph{has a simple zero at} $u=0$, i.e. it vanishes to order 1 in $u$.
877: (The same expression vanished only as $\sqrt u$ in the original family
878: with the singular fibre, and as such its sign was not well defined; this
879: is because the class $[L_1]$ was defined globally only up to the
880: monodromy $T_{L_1}[L_1]=-[L_1]$, i.e. up to a sign. In our new family
881: the monodromy action $T^2_{L_1}$ is trivial on homology so it makes sense
882: to talk about the homology class $[L_1]$ in any fibre, and (\ref{zero})
883: is single valued.)
884:
885: Then our loop of complex structures is given by taking the loop
886: $u=e^{it}$ and setting $\Omega^t=\Omega_{e^{it}}$. Pulling back the
887: K\"ahler form from the original family, we get a locally
888: trivial fibre bundle of symplectic manifolds over the
889: circle whose monodromy is the Dehn twist $T_{L_1}^2$ (since the
890: monodromy round the un-base-changed loop is $T_{L_1}$ \cite{S1}). As
891: the family no longer has a singular fibre this monodromy is trivial as
892: a diffeomorphism, but it is a result of \cite{S1},\,\cite{S2} that as a
893: symplectic automorphism \emph{it is non-trivial}. This is possible
894: because although the family is a locally trivial bundle of symplectic
895: manifolds over the punctured disc, over $u=0$ the symplectic form
896: becomes degenerate since it was pulled back via the resolution map.
897:
898: Measuring $[L_1]$ against $\Omega_u$ as in (\ref{zero}) we see a
899: principle familiar in physics (in issues of `marginal stability',
900: and taught to me by Eric Zaslow) -- we detect a monodromy,
901: like the degree 1 map $S^1\to\C^{\!\times}$ given by $t\mapsto\int_{L_1}
902: \Omega^t$ in this example, by counting wall crossing where a certain
903: real part (here $\int_{L_1}\ImO^t$, or the phase $\phi_1^t$) hits
904: $0\simeq\phi^t_2$.
905:
906: (Here we can no longer choose the phase of $\Omega$
907: such that $\phi^t_2=\phi(L^t_2)\equiv0$ in the whole family, as the
908: homology class of $L_2$ is not preserved in the family:
909: $$
910: [T_{L_1}^2L_2]=[L_2]+2[L_1].
911: $$
912: However, for a sufficiently small loop about the ODP, i.e. for
913: $\big|\int_{L_1}\Omega\big|$ sufficiently small, this will not
914: affect us much and we can write $\phi^t_2\simeq0$: we are only
915: interested in topological information like winding numbers and
916: $\phi^t_1$ crossing the wall at $\phi^t_2\simeq0$, which are
917: unaffected by small perturbations.)
918:
919: So instead of going through the $\phi(L_1^t)=\phi(L_2^t)\simeq0$
920: wall we can go round it. If the loop is sufficiently
921: small we do not encounter any more walls where the homology class
922: $[L_1]+[L_2]$ can be split into classes of the same phase to possibly
923: make the SLag a singular union of distinct SLags of equal phase. For
924: instance the wall at phase 0 does not extend past $u=0$ to phase
925: $\phi^t_1=\pi$ (even though there $\mu^t_1=0$) --
926: the phase of $L_1$ is not zero but $\pi$, and is only zero for $L_1$ with
927: the opposite orientation, so it does not exist as a SLag (e.g. in the
928: hyperk\"ahler rotated situation, we are saying there is no complex curve in
929: $L_1$'s homology class to possibly make $L$ the nodal union of $L_1$
930: and something else, there is only an anti-complex curve).
931: So we really can go round the wall; it ends at $u=0$.
932:
933: So this monodromy
934: description shows that on the other $t\uparrow2\pi$ side of the wall
935: the SLag deforming $L_2\cup L_1$ is in the hamiltonian deformation
936: class
937: \begin{equation} \label{giggs}
938: T_{L_1}^2L\,=\,T_{L_1}^2(L_1\#L_2)\,=\,T_{L_1}^2(T_{L_1}^{-1}L_2)\,
939: \approx\,T_{L_1}L_2\,\approx\,L_2\#L_1,
940: \end{equation}
941: as claimed (for the above equalities see \cite{S1}, \cite{S2}).
942:
943: Notice that the alternative connect sum description of the
944: above Lagrangian
945: \begin{equation} \label{decomp}
946: L_2\#L_1\,=\,T_{L_1}^2(L_1\#L_2)\,\approx\,T_{L_1}^2(L_1)\#
947: T_{L_1}^2(L_2)\,\approx\,L_1[-2]\#T_{L_1}^2(L_2),
948: \end{equation}
949: does not violate the phase inequality to (\ref{ineq}), as
950: $$
951: -2\pi+\epsilon\approx\phi(L_1[-2])<\phi(T_{L_1}^2(L_2))\simeq0.
952: $$
953: This is why it is important here to keep track of gradings --
954: assigning the phase
955: $\epsilon$ to $\phi(T_{L_1}^2(L_1))$ would give the opposite
956: inequality, but one would not be able to form the above graded
957: connect sum without also shifting
958: the phase of $T_{L_1}^2(L_2)$ by $-2\pi$.
959:
960: \begin{figure}[h]
961: \vspace{4mm}
962: \center{
963: \input{m5.pstex_t}
964: \caption{$\left(\int_{L_1}\Omega\right)$-space, as $\Omega$
965: on a 3-fold varies, with polar coordinates $(R,\,\phi(L_1))$}}
966: \end{figure}
967:
968: The 3-fold case (which Dominic Joyce has also, independently, considered)
969: is slightly different; we need only take a single Dehn twist
970: $T_{L_1}$ corresponding to the local family
971: $$
972: x^2+y^2+z^2=u,
973: $$
974: over $u\in\C$ to get a winding number one loop in the phase of $L_1$.
975: This is because
976: $$
977: T_{L_1}L_1\approx L_1[1-n]
978: $$
979: in dimension $n$, so in 3 dimensions the homology class $[L_1]$ is
980: preserved instead of being reversed. The corresponding picture is
981: displayed in Figure 2.
982:
983: Again there is a SLag on the other side of the $\phi=0$ wall, but
984: it is in the wrong homology class $[L_2]$:
985: \begin{equation} \label{cole}
986: T_{L_1}L\approx L_2.
987: \end{equation}
988: Analogously to (\ref{decomp}) this has a number of decompositions
989: as connect sums induced by monodromy,
990: $$
991: T_{L_1}(L_1\#L_2)\,\approx\,L_1[-2]\#(L_2\#(L_1[-1]))\,\approx\,L_2
992: \,\approx\,(L_1\#L_2)\#(L_1[\,1\,]),
993: $$
994: none of which violate the phase inequality (\ref{ineq}). The only other
995: obvious choice for a (S)Lag on the other side of the $\phi=0$ wall
996: (given the $K3$ result) is $T_{L_1}^2(L_1\#L_2)\approx L_2\#(L_1[-1])$;
997: this however is also in the wrong homology class, and in any case
998: does violate (\ref{ineq}) and so, by Joyce's analysis, should not
999: be represented by a SLag. Thinking of $T_{L_1}^2$ as rotating through
1000: $-4\pi$ in Figure 2, it is at roughly $-3\pi$ that the phase inequality
1001: (\ref{ineq}) gets violated, and the $-\pi$ rotation of $L_2$ splits
1002: as a SLag into the union of the $-\pi$ rotations of $(L_1\#L_2)$ and
1003: $L_1[\,1\,]$: these both have phase approximately zero.
1004:
1005:
1006: \subsection*{A holomorphic bundle example}
1007:
1008: These phenomena are similar to wall-crossing in bundle theory on the
1009: complex side -- in a real one-parameter family of K\"ahler forms, for
1010: fixed complex structure, stable holomorphic bundles for $t>0$ can become
1011: semistable at $t=0$ and unstable for $t<0$.
1012:
1013: An example that mirrors Joyce's is the following. Suppose we
1014: have two stable bundles (or coherent sheaves) $E_1$ and $E_2$ with
1015: $$
1016: \mathrm{Ext}^1(E_2,E_1)\cong\C.
1017: $$
1018: This is $H^1(E_1\otimes E_2^*)$ in the case of
1019: bundles and is the mirror \cite{K} of the one dimensional
1020: Floer cohomology $HF^*(L_2, L_1)\cong\C$ that is defined by the single
1021: intersection point of $L_1$ and $L_2$ (see Section \ref{Ko} for more
1022: details of this, and an explanation of why we are dealing with Ext$^1$
1023: and $HF^1$ here). We then form $E$ from this extension class
1024: \begin{equation} \label{E}
1025: 0\to E_1\to E\to E_2\to0.
1026: \end{equation}
1027: Take a family of K\"ahler forms $\omega^t$ such that
1028: $\mu^t(E_2)-\mu^t(E_1)$ is the same sign as $t$ (here $\mu^t(F)=
1029: c_1(F)\,.\,(\omega^t)^{n-1}/$\,rk\,$(F)$ is the slope of $F$ with
1030: respect to $\omega^t$). Supposing that the $E_i$ are stable for all
1031: $t\in(-\epsilon,\epsilon)$, we claim that $E$ is stable
1032: for sufficiently small $t>0$, while it is destabilised by $E_1$ for
1033: $t\le0$. Without loss of generality take $\mu^t(E_2)=\mu$ fixed,
1034: and $\mu^t(E_1)=\mu-t$. As $E_2$ is stable, for $t$ sufficiently small
1035: there are no subsheaves of $E_2$ of slope greater than $\mu-t$, so for
1036: any stable destabilising subsheaf $F$ of $E$, the composition
1037: $$
1038: F\into E\to E_2
1039: $$
1040: cannot be an injection (unless it is an isomorphism, but (\ref{E}) does
1041: not split. So $F\cap E_1\ne0$, and the quotient $Q=F/(F\cap E_1)$ has
1042: slope $\mu(Q)>\mu(F)>\mu-t$ by the stability of $F$ and
1043: instability of $E$. But $Q$ injects into $E_2$, which we know is
1044: impossible.
1045:
1046: In the 2-dimensional case, by Serre duality Ext$^1(E_1,E_2)
1047: \cong\,$Ext$^1(E_2,E_1)^*\cong\C$on $K3$ or $T^4$, so for $t<0$ we
1048: can instead form an extension
1049: \begin{equation} \label{ext}
1050: 0\to E_2\to E'\to E_1\to0,
1051: \end{equation}
1052: to give a new bundle $E'$ which is also stable, and has the same Mukai
1053: vector
1054: $$
1055: v(E')=v(E_1)+v(E_2);
1056: $$
1057: compare (\ref{+-}). At $t=0$ we take the (polystable)
1058: bundle
1059: $$
1060: E_1\oplus E_2.
1061: $$
1062: This is because the semistable extension (\ref{E}) no longer admits a
1063: Hermitian-Yang-Mills metric, but $E_1\oplus E_2$ does. Also, the algebraic
1064: geometry of the moduli problem shows that while a semistable bundle gets
1065: identified in the moduli space with the other (``S-equivalent'') sheaves
1066: in the closure of its gauge group orbit, there is a distinguished
1067: representative of its
1068: equivalence class -- the polystable direct sum (of the Jordan-H\"older
1069: filtration, which here is $E_1\oplus E_2$).
1070:
1071: Thus, while the HYM connections vary, the bundle
1072: has only 3 different holomorphic structures -- for $t>0,\ t=0,$ and $t<0$.
1073: Put another way (to spell out the analogy with the Lagrangians $L^t,\
1074: L_1,\ L_2$) as $\omega_t$ varies with $t>0$ we take different points in
1075: a fixed complexified gauge group orbit, and at $t=0$ we take as limit point
1076: something in a different orbit that is nonetheless in the closure of the
1077: $t>0$ (and $t<0$) orbit. The \emph{stable} deformations of the polystable
1078: $E_1\oplus E_2$ (which we are thinking of as the mirror of the singular
1079: union $L_1\cup L_2$, of course) are precisely (\ref{E}) for $t>0$ and
1080: (\ref{ext}) for $t<0$.
1081:
1082: In the 3-fold case, however, Serre duality gives Ext$^2(E_1,E_2)\cong\,
1083: $Ext$^1(E_2,E_1)^*\cong\C$ instead, and so no stable extension (\ref{ext}).
1084: In fact one would expect there to be no stable bundle with the right Chern
1085: classes; instead the one dimensional Ext$^2$ gives us a complex $E'$ in
1086: the derived category $D^b(M)$ fitting into an exact sequence of complexes
1087: $$
1088: 0\to E_2\to E'\to E_1[-1]\to0,
1089: $$
1090: where $E_1[-1]$ is $E_1$ shifted in degree by one place to the right as
1091: a complex. This has Mukai vector
1092: $$
1093: v(E')=v(E_2)-v(E_1),
1094: $$
1095: compare (\ref{+-}). Thus, just as in the
1096: case of SLags, as we pass through $t=0$ there is no natural stable
1097: object on the other side in the same homology class in 3 dimensions
1098: (though there is in 2 dimensions) and so an element of the appropriate
1099: moduli space disappears.
1100:
1101: In fact, as in the Lagrangian example, the natural stable object
1102: on the other side of the wall is $E_2$ if we consider monodromy.
1103: The mirror of the symplectic Dehn twists of above are described
1104: in \cite{ST} (in the case that the bundles $E_i$ are \emph{spherical}
1105: in the sense of \cite{ST}: Ext$^k(E_i,E_i)\cong H^k(S^n;\C)$; this is
1106: the natural mirror analogue of the
1107: $L_i$s being spheres). These are the twists $T_{E_1}$ of \cite{ST}
1108: on the derived category of the Calabi-Yau that act on the extension
1109: bundle $E$ of (\ref{E}) to give precisely the extension (\ref{ext}),
1110: $$
1111: T_{E_1}^2E=E'
1112: $$
1113: (compare (\ref{giggs})), as a short calculation using \cite{ST}
1114: shows. Similarly
1115: $$
1116: T_{E_1}E=E_2,
1117: $$
1118: the analogue of (\ref{cole}). (In both of these calculations it is
1119: important to calculate this monodromy in the derived
1120: category; in the $K3$ case the action of $T_{E_1}^2$ is trivial on
1121: K-theory and cohomology, and we cannot distinguish between (\ref{E})
1122: and (\ref{ext}), but they are very different as holomorphic bundles
1123: and as elements of the derived category.)
1124:
1125: The mirror wall crossing, with a SLag splitting into two and then
1126: disappearing, is interpreted in \cite{DFR} (and in \cite{SV} in a
1127: different case) as the
1128: state it represents decaying as we reach a point of `marginal
1129: stability'. Despite this dealing with only SLags (and so with only
1130: a priori \emph{stable} Lagrangians in our mathematical sense of
1131: stability), this suggestive language does in fact have something to say
1132: about the stability, in our sense of group actions, of (non-special)
1133: Lagrangians, by considering the nodal limit $L_1\cup L_2$ to be a
1134: \emph{semistable} Lagrangian.
1135:
1136: Thus the Lagrangian $L_1\#L_2$ (which
1137: \emph{always exists as a Lagrangian} as the complex
1138: structure varies with fixed K\"ahler form) becomes semistable
1139: at $t=0$ and is represented by something in a different orbit of the
1140: hamiltonian deformation symmetry group (but in the closure of the
1141: original orbit), and is unstable for $t<0$ so exists there only as a
1142: Lagrangian and \emph{not} as a SLag. This, and the bundle analogue
1143: described above, leads us to think of the Lagrangian $L_1$
1144: as destabilising $L=L_1\#L_2$ when $\phi(L_1)\ge\phi(L_2)$. This
1145: motivates the now obvious definition of stability in Section \ref{st};
1146: first we explain more about the connections to mirror symmetry, and
1147: generalisations to connect sums at more intersection points.
1148:
1149:
1150: \section{Relationship to Kontsevich's mirror conjecture} \label{Ko}
1151:
1152: The inspiration behind most of this paper is of course Kontsevich's
1153: mirror conjecture \cite{K}. In particular, Kontsevich proposes that
1154: the graded vector spaces Ext$^*$ and $HF^*$ should be isomorphic for
1155: mirror choices of bundles $E_i$ and graded Lagrangians $L_i$
1156: (or more exotic objects in their derived categories)
1157: $$
1158: HF^*(L_2,L_1)\cong\,\mathrm{Ext}^*(E_2,E_1);
1159: $$
1160: this corresponds to the equality of (graded) morphisms on both sides.
1161: Here $HF^*$ is Floer cohomology \cite{Fl} -- a symplectic refinement
1162: of the intersection number of $L_1$ and $L_2$ -- which can be $\mathbb
1163: Z$-graded for graded Lagrangians \cite{S2}, whenever it is defined
1164: \cite{FO3}, \cite{Fu1}. (More precisely it is the cohomology of a chain
1165: complex built out of the free vector space generated by the intersection
1166: points, with the differential defined by counting holomorphic
1167: discs with boundary in the Lagrangians running from one intersection
1168: point to another.) In mirror symmetry, and so in this paper, one
1169: should only really consider those Lagrangians whose Floer cohomology is
1170: well defined \cite{Fu1}.
1171:
1172: Thus the point of intersection of the $L_1$ and $L_2$ of the last
1173: section define the Floer cohomology $HF^*(L_2,L_1)\cong\C$, and the
1174: grading of \cite{S2} is designed specifically so that $L_1\#L_2$ can
1175: be graded precisely when the relative gradings of the $L_i$ force
1176: the Floer cohomology to be concentrated in degree 1; $HF^*(L_2,L_1)=
1177: HF^1(L_2,L_1)$. We then think of the connect sum $L_1\#L_2$ as being
1178: mirror to the extension (\ref{E}) defined by Ext$^1(E_2,E_1)\cong\C$.
1179: Fukaya, Seidel, and perhaps others have also proposed that Lagrangian
1180: connect sum should be mirror to extensions \cite{Fu2}, \cite{S3}.
1181:
1182: We also consider connect sums of Lagrangians intersecting at $n$
1183: points $p_i$. Then the connect sum is not unique up to hamiltonian
1184: deformation: $H^1$ is added to the Lagrangian as loops between the
1185: intersection points, giving additional deformations of its hamiltonian
1186: isotopy class. The upshot is that there is a scaling of the neck of
1187: the connect sum at each intersection point; we denote any such
1188: resulting Lagrangian by $L_1\#L_2$. Since we insist on all intersection
1189: points having Floer (Maslov) index one (so that the connect sum can be
1190: graded), the Floer differential vanishes in this case, and
1191: these scalings define a class in $HF^1(L_2,L_1)$.
1192:
1193: Deformations (up to those which are hamiltonian) as such a connect sum
1194: are given by the elements of
1195: $$
1196: H^1(L_1\#L_2)\cong H_{n-1}(L_1\#L_2)
1197: $$
1198: spanned by the $S^{n-1}$ vanishing cycles $S_i$ at the points of
1199: intersection $p_i\in L_1\cap L_2$. Given a particular connect sum,
1200: the deformation represented by $\sum_ia_iS_i$ simply scales the local
1201: gluing parameter in a Darboux chart around each $p_i$ by a factor
1202: $(1+a_i)$ (here $a_i$ is considered to be infinitesimal). Since
1203: the sum of these spheres
1204: separates $L_1\#L_2$ into $L_1\backslash\cup\{p_i\}$ and
1205: $L_2\backslash\cup\{p_i\}$ and so is zero in homology
1206: $$
1207: \sum_i[S_i]=\pm\partial[L_1\backslash\cup\{p_i\}]=
1208: \mp\partial[L_2\backslash\cup\{p_i\}]=0\in H_{n-1}(L_1\#L_2),
1209: $$
1210: the infinitesimal deformation represented by $\sum_iS_i$ is zero
1211: (it is pure hamiltonian) and dividing out gives the projectivisation
1212: \begin{equation} \label{proj}
1213: \Pee(\oplus_i\R_{p_i}).
1214: \end{equation}
1215: (Replace $\R$ by $\C$ when including flat bundles and their gluing
1216: parameters at the $p_i$s.) This
1217: explains the earlier claim that connect sums at one intersection
1218: point are uniquely defined up to hamiltonian deformations.
1219: More precisely, when holomorphic discs are
1220: taken into account and we consider only those Lagrangians whose Floer
1221: cohomology is defined \cite{FO3}, hamiltonian deformation classes of
1222: connect sums \emph{whose Floer cohomology can be defined}
1223: should be parameterised by $\Pee(HF^1(L_2,L_1))$.
1224: (On the mirror side \emph{isomorphism classes} of extensions of $E_2$
1225: by $E_1$ are parametrised by $\Pee\mathrm{\,Ext}^1(E_2,E_1)$.)
1226:
1227: We would then expect that the resulting connect sum has a canonical
1228: homomorphism from $L_1$; that is there should be a canonical element
1229: $$
1230: \id_{L_1}\in HF^0(L_1,L_1\#L_2)
1231: $$
1232: for any graded Lagrangians $L_i$ for which the graded connect sum
1233: exists. While a local model suggests this is true (see for instance
1234: \cite{TY}), a complete proof is
1235: still not available. This homomorphism we think of as expressing $L_1$
1236: as a \emph{subobject} of $L_1\#L_2$; i.e. as giving an injection.
1237: It should be emphasised that subobject does not make sense in a
1238: triangulated category such as the derived Fukaya category of Lagrangians;
1239: in the context of the derived category of sheaves, subobject
1240: only makes sense for an abelian category such as that of the sheaves
1241: themselves (i.e. complexes with cohomology in degree zero only).
1242: What we are proposing is that it also makes sense in the category of
1243: (complexes of sheaves mirror to) graded Lagrangians, and is vital to
1244: make definitions of stability (which involve such subobjects). While
1245: there are now more Homs to consider, in particular those of higher
1246: order (i.e. Homs to Lagrangians shifted in phase by some $2\pi n$),
1247: the targets of these Homs have higher phase and so do not disturb the
1248: definition of stability below -- this is seemingly a huge piece of luck
1249: that means we can extend the stability condition for bundles to all
1250: Lagrangians. For similar reasons, the many
1251: connect sum decompositions of the $L_i$s given in the last section
1252: also do not destabilise them.
1253:
1254: There are other operations, however, which can also be thought of
1255: as Ext$^1$-type extensions. For instance, taking the product of a
1256: single Lagrangian curve $L_1$ in $T^2$ with a (graded) connect sum
1257: $L_2\#L_3$ in another $T^2$, we get a
1258: Lagrangian $L_1\times(L_2\#L_3)$ in $T^4$ which is some kind of
1259: extension of the Lagrangians $L_1\times L_2$ and $L_1\times L_3$ in
1260: $T^4$. Supposing that the $L_i$s are mirror to some (complexes of)
1261: sheaves $E_i$, and that the connect sum $L_2\#L_3$ is mirror to an
1262: extension represented by an element $e\in$\,Ext$^1(E_3,E_2)$. Then
1263: by the K\"unneth formula for sheaf cohomology, we see that
1264: $L_1\times(L_2\#L_3)$ is indeed mirror to an extension
1265: $$
1266: \id\,\otimes e\in\mathrm{Hom\,}(E_1,E_1)\otimes\mathrm{Ext}^1
1267: (E_3,E_2)=\mathrm{Ext}^1(E_1\boxtimes E_3,E_1\boxtimes E_2),
1268: $$
1269: and so this sort of relative connect sum (which is not $\#$ on
1270: $T^4$: $L_1\times L_2$ and $L_1\times L_3$ do not intersect
1271: transversely) should also be considered.
1272:
1273: So we consider Lagrangians $L_1,\,L_2$ intersecting \emph{cleanly}
1274: (see e.g. \cite{S1} Definition 2.1), that is $N=L_1\cap L_2$ is a smooth
1275: submanifold, and $TN=TL_1\res N\cap TL_2\res N$. Basic results of
1276: Weinstein allow us to identify a neighbourhood of $N$ with a
1277: neighbourhood of the zero section $N$ in $T^*N\oplus E$, where the total
1278: space of $T^*N$ has its canonical symplectic structure, and
1279: $$
1280: E\equiv(TL_1\res N)/TN\,\oplus\,(TL_2\res N)/TN
1281: $$
1282: is the annihilator, under the symplectic form, of $TN\subset TX\res N$
1283: (to which the symplectic form therefore descends, making $E$ a
1284: symplectic bundle).
1285:
1286: Choosing a metric on $E$, compatible with its symplectic structure,
1287: such that its transverse
1288: subbundles $(TL_1\res N)/TN,\ (TL_2\res N)/TN$ are orthogonal, we can now
1289: perform the family connect sum of these, over the base $N$, since
1290: the local model in \cite{S1} is $O(n)$ invariant. As before we insist
1291: that this can be compatibly graded again denote it by $\#$; given a
1292: grading on $L_1$ there will be at most one grading on $L_2$ such that
1293: this graded relative connect sum exists.
1294:
1295: It should be noted that although such a clean intersection could be
1296: hamiltonian isotoped to be transverse, the resulting intersection points
1297: would not necessarily all be of Floer/Maslov index one, and so the
1298: pointwise graded connect sum could not be formed at every point; we
1299: would end up with an immersed Lagrangian. Studying which immersed
1300: Lagrangians should be included in the Fukaya category, and which
1301: embedded Lagrangians they should be considered equivalent to, is
1302: an important part of mirror symmetry and will need to be better understood
1303: to refine our conjecture. For instance forming extensions of bundles
1304: which also have nonzero homorphisms between them would appear to
1305: be mirror to forming connect sums between graded Lagrangians at index
1306: one intersection points, leaving the index zero intersection points
1307: immersed. In general one would like to consider two objects of the Fukaya
1308: category to be equivalent if their Floer cohomologies with any other
1309: objects are the same. This would include hamiltonian deformation
1310: equivalence, but also more exotic equivalences for immersed
1311: Lagrangians (thanks to Paul Seidel for pointing this out to me).
1312: A start in understanding the Floer cohomology of immersed
1313: Lagrangians is \cite{Ak}; in the present paper we are largely ignoring
1314: singularities.
1315:
1316:
1317: \section{Stability} \label{st}
1318:
1319: \begin{Definition}
1320: Take graded Lagrangians $(L_1,\theta_1)$ and $(L_2,
1321: \theta_2)$, hamiltonian isotoped to intersect cleanly, and such
1322: that the graded (relative) Lagrangian connect sums
1323: $(L_1\#L_2,\theta_1\#\theta_2)$ exist as above. Then a
1324: Lagrangian $L$ of Maslov class zero is said to be
1325: destabilised by the $L_i$ if it is hamiltonian isotopic to such an
1326: $L_1\#L_2$, and the phases (\emph{real} numbers, induced by the
1327: gradings) satisfy
1328: $$
1329: \phi(L_1)\ge\phi(L_2).
1330: $$
1331: If $L$ is not destabilised by any such $L_i$ then it is called stable.
1332: \end{Definition}
1333:
1334: \begin{Remarks}
1335: \begin{description} \item[\ \ $\bullet$\ \ ]
1336: There is an obvious notion of a flux homomorphism for isotopies
1337: of smooth Lagrangians, taking a deformation to an element of
1338: $H^1(L;\R)$ (and linearising to give the usual deformation theory of
1339: Lagrangians). Namely, take a deformation $\Phi_t(L)$ through a
1340: vector field $X_t,\ t\in[0,1]$ to the one form
1341: $$
1342: \int_0^1(X_t\ip\omega)dt\in H^1(L;\R).
1343: $$
1344: Alternatively, the homomorphism takes a loop $\gamma\subset L$,
1345: tracing out the 2-cycle $f(\gamma\times[0,1])$ in $W$ under the
1346: isotopy, to the real number $\int_{\gamma\times[0,1]}\omega$.
1347: (See Chapter 10 of \cite{MS} for the analogous map for
1348: symplectomorphisms.) If the isotopy $\Phi_t$ is hamiltonian, the
1349: flux is zero; the converse is also easily proved using the methods of
1350: (\cite{MS} Theorem 10.12): we may assume without loss of generality
1351: that the 1-form $\int_0^1X_t\ip\omega$ is identically zero in
1352: $\Omega^1(L)$. [To see this, write the 1-form as
1353: $d\phi$, and compose the deformation with
1354: the time one map of the hamiltonian flow with vector field
1355: $d\phi\ip\omega^{-1}$; this does not alter the flux in $H^1(L;\R)$ or
1356: the property of being hamiltonian.] Then let $\Sigma^s$ be the closed
1357: 1-form on $L$ defined by
1358: $$
1359: \Sigma^s=\int_0^sX_t\ip\omega dt,
1360: $$
1361: and let $\Psi^s_t$ be the corresponding flow through time $t$. Then
1362: the flow $\phi_t=\Psi^t_t\circ\Phi_t$ is the corresponding
1363: \emph{hamiltonian} flow from $\Phi_0(L)$ to $\Phi_1(L)$; see
1364: \cite{MS}.
1365:
1366: Thus it is not too hard to check if two Lagrangians
1367: are hamiltonian deformations of each other, at least through
1368: smooth Lagrangians, if we know they are deformations of each
1369: other as Lagrangians. This second condition, however, is harder
1370: to test, as the results of \cite{S1} demonstrate.
1371:
1372: \item[\ \ $\bullet$\ \ ] As mentioned in the last section,
1373: holomorphic discs are
1374: crucial in both mirror symmetry and Floer cohomology; thus one
1375: should perhaps restrict attention in the above definition to
1376: those Lagrangians whose Floer cohomology is defined \cite{FO3},
1377:
1378: \item[\ \ $\bullet$\ \ ]As pointed out to me by Conan Leung, this
1379: definition and the resulting conjecture below may only be reasonable
1380: close to the
1381: large complex structure limit point where the mirror symmetric
1382: arguments used to motivate the conjecture are most valid.
1383: \end{description}
1384: \end{Remarks}
1385:
1386: \begin{Conj}
1387: A Lagrangian of Maslov class zero has a special Lagrangian in its
1388: hamiltonian deformation class if and only if it is stable, and this
1389: SLag representative is unique.
1390: \end{Conj}
1391:
1392: Again, we have been vague about singularities: which we allow, and what
1393: hamiltonian deformation equivalence would mean for them. We might also
1394: want to restrict attention to those Lagrangians whose Floer cohomology
1395: exists \cite{FO3}, and whose Oh spectral sequence $H^*(L)\Rightarrow
1396: HF^*(L,L)$ \cite{Oh} degenerates; this will be discussed more in \cite{TY}.
1397: We might also want to restrict to Lagrangians whose phase function
1398: varies only by a certain bounded amount; in the example worked out in
1399: \cite{TY} this is required. In \cite{TY}
1400: it is shown there that the gradient of the
1401: norm-squared $|m|^2$ of the moment map can be taken to be the mean
1402: curvature vector of the Lagrangian, so mean curvature flow (which
1403: is hamiltonian for Maslov class zero) should
1404: converge to this SLag representative if the Lagrangian is stable
1405: and the phase satisfies certain bounds.
1406:
1407:
1408: \section{The 2-torus}
1409:
1410: Everything works rather simply on $T^2$; Grayson \cite{G}, building
1411: on work of Gage, Hamilton and others (e.g. \cite{GH}), has shown that
1412: mean curvature flow for curves (of Maslov class zero) converges to
1413: straight lines and so we get the mirror symmetric analogue
1414: of Atiyah's classification \cite{At} of sheaves on an elliptic curve
1415: -- they are basically all sums of stable sheaves. The only exceptions
1416: are the non-trivial extensions of certain
1417: sheaves by themselves; these correspond to thickenings of the
1418: corresponding special Lagrangian (giving fat SLags, as they are known
1419: in Britain, or multiply-wrapped cycles in physics speak).
1420:
1421: \begin{figure}[ht]
1422: \vspace{1mm}
1423: \center{\input{T2.pstex_t}}
1424: \caption{$L_1\# L_2$ and $L_2\#(L_1[1])$, equivalent SLags, and
1425: their mirror sheaves}
1426: \vspace{1mm}
1427: \end{figure}
1428:
1429: We give an example to demonstrate why one cannot form smooth
1430: \emph{unstable} Lagrangians on $T^2$ in Figure 3.
1431: First, giving $L_1$ and $L_2$ the gradings such that their
1432: phases are 0 and $\pi/4$, we expect $L_1\#L_2$ to be stable, and
1433: indeed we see it is hamiltonian deformation equivalent to the slope
1434: $1/2$ SLag mirror to the stable extension $E$ of $\OO$ by $\OO(p)$
1435: (where $p$ is a basepoint of $T^2$ with corresponding line bundle mirror
1436: to the diagonal SLag drawn).
1437:
1438: If one then tries to form an unstable SLag $L_2\#L_1$, the \emph{graded}
1439: connect sum does not exist -- the phase would become
1440: discontinuous. To form $L_2\#L_1$ we see from the diagram that we have to
1441: take the phase of $L_1$ to be $\pi$, thus reversing its orientation, and
1442: in fact forming $L_2\#(L_1[1])$.
1443: Then the stability inequality (\ref{ineq}) is not violated, and in fact
1444: this Lagrangian is stable and hamiltonian deformation equivalent to
1445: the SLag in $T^2$ represented by the vertical edge of the square (and so
1446: drawn with a little artistic license in Figure 3). Under the mirror
1447: map this corresponds to replacing the extension Ext$^1$ class by a Hom
1448: (as we have shifted complexes of sheaves by one place) and taking the
1449: cone of this in the derived
1450: category; this is the cokernel $\OO_p$ of Figure 3.
1451:
1452: As pointed out to me by Markarian and Polishchuk, one can play with
1453: lots of pictures of connect sums on tori to recover descriptions
1454: of certain moduli of sheaves, their special cycles (for instance
1455: where one connect-sum neck parameter goes to zero), and so forth,
1456: giving results similar to some of those in \cite{FO}.
1457:
1458: This example can be extended to show that we cannot form the graded
1459: connect sum $L_1\#L_2$ of any two Lagrangians (via a class in
1460: $HF^*(L_1,L_2)$) if $\phi(L_1)>\phi(L_2)$. Namely, replace
1461: $L_1$ and $L_2$ by their hamiltonian deformation equivalent SLag
1462: representatives, which are straight lines of constant phase $\theta_i=
1463: \phi(L_i)$. As Figure 3 shows, $L_1\#L_2$ can be compatibly
1464: graded about an intersection point if and only if we have the local
1465: inequalities
1466: $$
1467: \theta_2>\theta_1>\theta_2-\pi.
1468: $$
1469: Thus we require $\phi(L_2)>\phi(L_1)$. (We will explain this kind
1470: of phenomenon more generally in \cite{TY} in terms of the grading
1471: on Floer cohomology.) Each intersection point is
1472: Floer coclosed since the Floer grading is the same
1473: as the relative orientation of the Lagrangians, mod 2, and this
1474: is the same at each intersection point of the straight lines. So each
1475: possible connect sum of the SLags defines a class in $HF^*$, and
1476: any other connect sum, defined on hamiltonian deformations of $L_1$
1477: and $L_2$ by a class in $HF^*$, will be hamiltonian deformation
1478: equivalent to the appropriate connect sum of the SLags, and so
1479: satisfy the same phase inequality.
1480:
1481: If two smooth Lagrangians have the same phase then their
1482: representative SLags will either be the same or disjoint parallel
1483: SLags. Either way there are no connect sums (though as mentioned above
1484: to account for the mirror symmetry of bundles one should also include
1485: non-trivial thickenings of SLags in the Fukaya category).
1486:
1487: So unstable Lagrangians do not exist, and by the result of \cite{G}
1488: mentioned earlier, the conjecture is true on $T^2$.
1489:
1490: Thus complex dimension 1 is too simple -- in trying to make the phase
1491: of one Lagrangian become larger than the phase of another, the two must
1492: cross, thus reversing their relative orientations and changing the
1493: order of the connect sum. Far more complicated phenomena arise
1494: in 2 and 3 dimensions, however.
1495:
1496:
1497: \begin{thebibliography}{MMMS}
1498:
1499: \bibitem[Ak]{Ak} M. Akaho,
1500: \textit{An intersection theory of Lagrangian immersions}, preprint.
1501:
1502: \bibitem[At]{At} M. F. Atiyah,
1503: \textit{Vector bundles over an elliptic curve},
1504: Proc. London Math. Soc. \textbf{7} (1957), 414--452.
1505:
1506: \bibitem[AP]{AP} D. Arinkin and A. Polishchuk,
1507: \textit{Fukaya category and Fourier transform}, preprint
1508: math.AG/9811023.
1509:
1510: \bibitem[BBHM]{BBHM}
1511: C. Bartocci, U. Bruzzo, D. Hernandez Ruiperez and J. Munoz Porras,
1512: \textit{Mirror symmetry on $K3$ surfaces via Fourier-Mukai transform},
1513: Comm. Math. Phys. \textbf{195} (1998), 79--93.
1514:
1515: \bibitem[Ch]{Ch} J. Chen,
1516: \textit{Lagrangian sections and holomorphic $U(1)$-connections}, preprint.
1517:
1518: \bibitem[D]{D} S. K. Donaldson,
1519: \textit{Moment maps and diffeomorphisms}, Asian Jour. Math. \textbf{3}
1520: (1999), 1--16.
1521:
1522: \bibitem[DT]{DT} S. K. Donaldson and R. P. Thomas,
1523: \textit{Gauge theory in higher dimensions.} In: The Geometric Universe:
1524: Science, Geometry and the work of Roger Penrose, S. A. Huggett et al
1525: (eds), Oxford University Press, 1998.
1526:
1527: \bibitem[Do]{Do} M. R. Douglas,
1528: \textit{D-branes, Categories and N=1 Supersymmetry},
1529: preprint hep-th/0011017.
1530:
1531: \bibitem[DFR]{DFR} M. R. Douglas, B. Fiol, and C. R\"omelsberger,
1532: \textit{Stability and BPS branes}, preprint hep-th/0002037.
1533:
1534: \bibitem[FO]{FO}
1535: B. Feigin and A. Odesskij,
1536: \textit{Vector bundles on an elliptic curve and Sklyanin algebras},
1537: q-alg/9509021. In: Topics in quantum groups and finite-type invariants,
1538: B. Feigin et al (eds), AMS Transl. 185 (38), 65--84, 1998.
1539:
1540: \bibitem[Fl]{Fl} A. Floer,
1541: \textit{Morse theory for Lagrangian intersections},
1542: Jour. Diff. Geom. \textbf{28} (1988), 513--547.
1543:
1544: \bibitem[Fu1]{Fu1} K. Fukaya,
1545: \textit{Floer homology and Mirror symmetry I}, preprint (1999)
1546: http://www.kusm.kyoto-u.ac.jp/$\sim$fukaya/fukaya.html.
1547:
1548: \bibitem[Fu2]{Fu2} K. Fukaya,
1549: \textit{Mirror symmetry of Abelian variety and multi theta functions},
1550: preprint (2000) http://www.kusm.kyoto-u.ac.jp/$\sim$fukaya/fukaya.html.
1551:
1552: \bibitem[FO3]{FO3} K. Fukaya, Y.-G. Oh, H. Ohta, and K. Ono,
1553: \textit{Lagrangian intersection Floer theory -- anomoly and obstruction},
1554: book to appear in International Press.
1555:
1556: \bibitem[GH]{GH} M. Gage and R. Hamilton,
1557: \textit{The shrinking of convex plane curves by the heat equation},
1558: Jour. Diff. Geom. \textbf{23} (1986), 69--96.
1559:
1560: \bibitem[G]{G} M. Grayson,
1561: \textit{Shortening embedded curves},
1562: Ann. of Math. \textbf{129} (1989), 71--111.
1563:
1564: \bibitem[Gr]{Gr} M. Gross,
1565: \textit{Special Lagrangian Fibrations I: Topology.} In: Integrable
1566: systems and algebraic geometry (Kobe/Kyoto, 1997), 156--193,
1567: World Sci. Publishing, 1998.
1568:
1569: \bibitem[Ha]{Ha} F. R. Harvey,
1570: \textit{Spinors and Calibrations.} Academic Press, San Diego, 1990.
1571:
1572: \bibitem[HL]{HL} F. R. Harvey and H. B. Lawson,
1573: \textit{Calibrated Geometries}, Acta Math. \textbf{148} (1982), 47--157.
1574:
1575: \bibitem[H]{H} N. Hitchin,
1576: \textit{Lectures on Special Lagrangian Submanifolds},
1577: preprint math.DG/9907034.
1578:
1579: \bibitem[J]{J} D. Joyce,
1580: \textit{On counting special Lagrangian homology 3-spheres},
1581: preprint hep-th/9907013.
1582:
1583: \bibitem[K]{K} M. Kontsevich,
1584: \textit{Homological Algebra of Mirror Symmetry}, International Congress
1585: of Mathematicians, Z\"urich 1994. Birkh\"auser, 1995.
1586:
1587: \bibitem[KS]{KS} M. Kontsevich and Y. Soibelman,
1588: \textit{Homological mirror symmetry and torus fibrations}, preprint
1589: math.SG/0011041.
1590:
1591: \bibitem[L]{L} G. Lawlor,
1592: \textit{The angle criterion}, Inv. Math. \textbf{95} (1989), 437--446.
1593:
1594: \bibitem[Le]{Le} N. C. Leung,
1595: \textit{Mirror symmetry without corrections}, preprint math.DG/0009235.
1596:
1597: \bibitem[LYZ]{LYZ} N. C. Leung, S.-T. Yau and E. Zaslow,
1598: \textit{From Special Lagrangian to Hermitian-Yang-Mills via Fourier-Mukai
1599: Transform}, preprint math.DG/0005118.
1600:
1601: \bibitem[MS]{MS} D. McDuff and D. Salamon,
1602: \textit{Introduction to symplectic topology}, 2nd edition, Oxford
1603: University Press, 1998.
1604:
1605: \bibitem[MMM]{MMM} Y. Manin,
1606: \textit{Moduli, Motives, Mirrors}, preprint math.AG/0005144.
1607:
1608: \bibitem[MMMS]{MMMS} M. Marino, R. Minasian, G. Moore and A. Strominger,
1609: \textit{Nonlinear Instantons from Supersymmetric p-Branes},
1610: preprint hep-th/9911206.
1611:
1612: \bibitem[Oh]{Oh} Y.-G. Oh,
1613: \textit{Floer cohomology, spectral sequence and the Maslov class of
1614: Lagrangian embeddings}, Internat. Math. Res. Notices \textbf{7} (1996),
1615: 305--346.
1616:
1617: \bibitem[PZ]{PZ} A. Polishchuk and E. Zaslow,
1618: \textit{Categorical mirror symmetry: the elliptic curve},
1619: Adv. Theor. Math. Phys. \textbf{2} (1998), 443--470.
1620:
1621: \bibitem[S1]{S1} P. Seidel,
1622: \textit{Lagrangian two-spheres can be symplectically knotted}, Jour.
1623: Diff. Geom. \textbf{52} (1999), 145--171.
1624:
1625: \bibitem[S2]{S2} P. Seidel,
1626: \textit{Graded Lagrangian submanifolds}, Bull. Soc. Math. France.
1627: \textbf{128} (2000), 103--146.
1628:
1629: \bibitem[S3]{S3} P. Seidel, private communication.
1630:
1631: \bibitem[ST]{ST} P. Seidel and R. P. Thomas,
1632: \textit{Braid group actions on derived categories of sheaves},
1633: pre\-print math.AG/0001043.
1634:
1635: \bibitem[SV]{SV} A. Shapere and C. Vafa,
1636: \textit{BPS structure of Argyres-Douglas superconformal theories},
1637: preprint hep-th/9910182.
1638:
1639: \bibitem[SYZ]{SYZ} A. Strominger, S.-T. Yau and E. Zaslow,
1640: \textit{Mirror Symmetry is T-Duality}, Nucl. Phys. \textbf{B479}
1641: (1996), 243--259.
1642:
1643: \bibitem[T1]{T1} R. P. Thomas,
1644: \textit{A holomorphic Casson invariant for Calabi-Yau 3-folds, and
1645: bundles on K3 fibrations}, preprint math.AG/9806111.
1646:
1647: \bibitem[T2]{T2} R. P. Thomas,
1648: \textit{An obstructed bundle on a Calabi-Yau 3-fold},
1649: Adv. Theor. Math. Phys. \textbf{3} (1999).
1650:
1651: \bibitem[T3]{T3} R. P. Thomas,
1652: \textit{D-branes: Lagrangians, bundles and derived categories},
1653: to appear as a chapter in the book \textit{``Mirror Symmetry''},
1654: ed. C. Vafa and E. Zaslow. Available at
1655: http://www.ma.ic.ac.uk/$\sim$rpwt.
1656:
1657: \bibitem[TY]{TY} R. P. Thomas and S.-T. Yau,
1658: \textit{Special Lagrangians, stable bundles and mean curvature flow},
1659: preprint.
1660:
1661: \bibitem[Ty]{Ty} A. N. Tyurin,
1662: \textit{Non-abelian analogues of Abel's theorem}, ICTP preprint
1663: (1997).
1664:
1665: \bibitem[Va]{Va} C. Vafa,
1666: \textit{Extending Mirror Conjecture to Calabi-Yau with Bundles},
1667: Comm. Contemp. Math. \textbf{1} (1999), 65--70.
1668:
1669: \bibitem[W1]{W1} E. Witten,
1670: \textit{Chern-Simons gauge theory as a string theory},
1671: The Floer memorial volume, 637--678. Progr. Math., 133,
1672: Birkh\"auser, Basel, 1995.
1673:
1674: \bibitem[W2]{W2} E. Witten,
1675: \textit{Branes And The Dynamics Of QCD},
1676: Nucl. Phys. \textbf{B}507 (1997), 658--690.
1677:
1678: \bibitem[Y]{Y} S.-T. Yau,
1679: \textit{On the Ricci curvature of a compact K\"ahler manifold and the
1680: complex Monge-Ampere equation}, Comm. Pure Appl. Math.
1681: \textbf{31} (1978), 339--411.
1682:
1683: \end{thebibliography}
1684:
1685: \noindent {\tt thomas@math.harvard.edu} \newline
1686: \noindent Department of Mathematics, Harvard University, One
1687: Oxford Street, Cambridge MA 02138. USA.
1688:
1689: \end{document}
1690:
1691:
1692: % Local Variables:
1693: % TeX-parse-self: t
1694: % TeX-auto-save: t
1695: % mode: latex
1696: % TeX-master: "thesis"
1697: % End: