math0104197/m6.tex
1: \documentclass[11pt]{article}
2: \setlength{\textwidth}{14cm}
3: \setlength{\oddsidemargin}{1cm}
4: 
5: \usepackage[dvips]{graphicx,color}
6: \usepackage{amssymb,latexsym,epsfig,epsf,xypic,amsmath}
7: \usepackage[mathscr]{euscript}
8: 
9: \makeatletter \@addtoreset{equation}{section} \makeatother
10: \renewcommand{\theequation}{\thesection.\arabic{equation}}
11: 
12: \newtheorem{Lemma}[equation]{Lemma}
13: \newtheorem{Theorem}[equation]{Theorem}
14: \newtheorem{Prop}[equation]{Proposition}
15: \newtheorem{Definition}[equation]{Definition}
16: \newtheorem{Corollary}[equation]{Corollary}
17: \newtheorem{Conj}[equation]{Conjecture}
18: \newenvironment{Proof}{\noindent\emph{Proof\ }}{\hfill$\square$\\}
19: \newenvironment{Remarks}{\noindent\textbf{Remarks}\ }{\\}
20: \newenvironment{Remark}{\noindent\textbf{Remark}\ }{\\}
21: 
22: \newcommand{\rt}[1]{\stackrel{#1\,}{\rightarrow}}
23: \newcommand{\Rt}[1]{\stackrel{#1\,}{\longrightarrow}}
24: \newcommand\A{\mathcal A}
25: \newcommand\db{\bar\partial}
26: \newcommand\R{\mathbb R}
27: \newcommand\C{\mathbb C\,}
28: \newcommand\Z{\mathbb Z}
29: \newcommand\Pee{\mathbb P}
30: \newcommand\PD{\mathrm{PD}\,}
31: \newcommand\OO{\mathscr O}
32: \newcommand\I{\mathcal I}
33: \newcommand\To{\longrightarrow}
34: \newcommand\ot{\leftarrow}
35: \newcommand\Ot{\longleftarrow}
36: \newcommand\into{\hookrightarrow}
37: \newcommand\res{\arrowvert_}
38: \newcommand\End{\mathrm{End}\,}
39: \newcommand\ip{{\mbox\,\_\hspace{-1.5mm}\shortmid\hspace{1.5pt}}}
40: \newcommand\ImO{\,\mathrm{Im}\,\Omega}
41: \newcommand\Jd{J\,\widetilde{\!d\theta\,}}
42: \newcommand\Jdh{J\,\widetilde{\!dh\,}}
43: \DeclareMathOperator\MCV{MCV}
44: \DeclareMathOperator\id{id}
45: \DeclareMathOperator\tr{tr}
46: \DeclareMathOperator\vol{vol}
47: \DeclareMathOperator\ind{ind}
48: 
49: \title{\textbf{Special Lagrangians, stable bundles and mean
50: curvature flow}}
51: \author{R. P. Thomas and S.-T. Yau}
52: \date{}
53: 
54: \begin{document}
55: 
56: \maketitle
57: \begin{abstract} \noindent
58: We make a conjecture about mean curvature flow of Lagrangian
59: submanifolds of Calabi-Yau manifolds, expanding on that of \cite{Th}.
60: We give new results about the stability condition in \cite{Th}, and
61: propose a Jordan-H\"older-type decomposition of (special) Lagrangians.
62: The main results are the uniqueness of
63: special Lagrangians in hamiltonian deformation classes of
64: Lagrangians, under mild conditions, and a proof of the conjecture
65: in some cases with symmetry: mean curvature flow converging to
66: Shapere-Vafa's examples of SLags.
67: \end{abstract}
68: 
69: 
70: \section{Introduction}
71: 
72: Fix a Calabi-Yau manifold $X$ with a holomorphic $(n,0)$-form
73: $\Omega$. In \cite{Th} a stability condition for Lagrangians in $X$
74: was described,
75: conjectured to be equivalent to the existence of a special Lagrangian
76: (SLag) in the hamiltonian deformation class of a fixed Lagrangian.
77: This was motivated by an infinite dimensional set-up
78: in which $U(1)$ gauge transformations act on the (infinite
79: dimensional) space of Lagrangians
80: with flat $U(1)$ connections on them. There is a natural complex
81: structure and symplectic form on this space and, ignoring issues
82: of integrability of these structures (see \cite{Th}),
83: the formal complexification of the $U(1)$ gauge transformations
84: gives hamiltonian deformations of the Lagrangian, with moment map
85: the $n$-form $\ImO\res L$. The stability condition was
86: motivated by an example of Joyce and the `angle criterion', in terms
87: of splittings of the Lagrangian into Seidel's graded Lagrangian
88: connect sums (as defined in Section \ref{index} below) and family
89: versions thereof, with a
90: certain phase inequality. This led to a conjecture, a sort of
91: globalised version of the angle criterion \cite{L}, \cite{N},
92: that the hamiltonian
93: deformation class of a Lagrangian should contain a SLag if and only
94: if the Lagrangian is stable; this SLag representative should then be
95: unique. Here
96: we expand on the conjecture and relate it to mean curvature flow. It
97: was verified for the simplest case $T^2$ in \cite{Th}; here we prove
98: it in a series of $n$-dimensional examples with symmetry (Theorem
99: \ref{faq}), and prove uniqueness of smooth SLags in hamiltonian deformation
100: classes whose Floer cohomology \cite{FO3} is defined (Theorem \ref{!}).
101: 
102: We write $\approx$ for ``in the same hamiltonian
103: deformation class as'', and use $\,\widetilde{\ }\,$ for the
104: isomorphism $T^*L\to TL$ induced by the metric on a Riemannian
105: manifold $L$. Restricting the Ricci-flat metric on a Calabi-Yau
106: manifold $(X,\Omega)$ to a Lagrangian submanifold $L$ we get an induced
107: volume form $\vol$ on $L$, and by a short calculation
108: \begin{equation} \label{vol}
109: \Omega\res L=e^{i\theta}\vol
110: \end{equation}
111: defines an $S^1$-valued function $\theta$ on $L$, the
112: \emph{phase function of} $L$. A \emph{grading} of $L$ is a lift of
113: $\theta$ to a real valued function. By Lagrangian we will always mean
114: graded Lagrangian (thus the Maslov class of the Lagrangian, which
115: is the class of $d\theta$ in $H^1(L;2\pi\mathbb{Z})$, is assumed to vanish,
116: and we have chosen a lift of $\theta$). $L$ is special Lagrangian
117: (SLag) if $\theta$ is a constant; equivalently, replacing $\Omega$
118: by $e^{-i\theta}\Omega$, $\ImO\res L\equiv0$.
119: An average, cohomological, measure of the phase of a homology class
120: $[L]$ is given by taking the phase of the complex number $\int_L\Omega$;
121: for $L$ graded with the variation of $\theta$ less than $2\pi$, this lifts
122: naturally to give a real number
123: $\phi(L)$, which is the phase of any SLag in the same homology class.
124: 
125: We should point out that as in \cite{Th}, we do not fully understand the
126: role of holomorphic discs in the theory. These are of course crucial in
127: the definition and hamiltonian deformation invariance of Floer
128: cohomology; until this is fully set up \cite{FO3} and all of its
129: expected properties (such as the spectral sequences of \cite{Oh2} and
130: \cite{P}) are proved and extended to the Calabi-Yau case,
131: some of the arguments below are necessarily conjectural; it will be
132: clear which ones. We also
133: deal exclusively with smooth (S)Lags; how to modify our constructions
134: to include singularities is an important question. Using only
135: (family) Lagrangian connect sums as the degenerations necessary
136: to describe stability of Lagrangians is also probably too restrictive,
137: studying other singularities and splittings may also be necessary; the
138: conjecture in this paper is probably just the first step in understanding
139: SLags in hamiltonian isotopy classes.
140: 
141: \noindent \textbf{Acknowledgements.} The symplectic ideas and
142: suggestions of Paul Seidel have been absolutely invaluable throughout
143: this work. We have also benefitted from comments from Kenji Fukaya,
144: Edward Goldstein, Spiro Karigiannis, Conan Leung, Jun Li, Elizabeth
145: Mann, Yong-Geun Oh and Xiao Wei Wang, and would like to thank Mike
146: Gage for the reference [An]. The first author is supported by a
147: Royal Society university research fellowship and by Imperial College,
148: London; the second author is supported by DOE grant DE-FG02-88ER35065
149: and NSF grant DMS-9803347.
150: 
151: 
152: \section{Mean curvature flow}
153: 
154: We first give a well-known geometric calculation which
155: we learnt from unpublished lectures of Rick Schoen on his work with
156: Jon Wolfson, but which dates back at least as far as \cite {HaL},
157: \cite{Oh1} and others.
158: 
159: \begin{Lemma} \label{mcf}
160: In the above notation, the mean curvature vector of the Lagrangian
161: $L\subset X$ is $MCV=\Jd$.
162: \end{Lemma}
163: 
164: \begin{Proof}
165: We want to show that for any vector $X$ tangent to $L$,
166: $X\theta=-\langle\MCV,JX\rangle$.
167: 
168: Picking an orthonormal basis of $T_pL$ and parallel transporting it
169: along rays in $L$ to a frame field $(e_i)$, $(e_i,Je_i)$ forms a local
170: basis for $TX$ around $p$. Letting $(f_j,g_j=-f_j\circ J)$ be the dual
171: basis of 1-forms, it is clear that at $p$,
172: $$
173: \Omega=e^{-i\theta}\bigwedge_j(f_j+ig_j),
174: $$
175: with $\theta$ the phase function of $L$. Since $\Omega$ is parallel,
176: $\nabla_X\Omega=0$ yields
177: \begin{eqnarray}
178: iX(\theta)\bigwedge_j(f_j+ig_j)&=&\sum_k(f_1+ig_1)\wedge\ldots\wedge
179: \nabla_X(f_k+ig_k)\wedge\ldots\wedge(f_n+ig_n) \nonumber \\
180: &=&\sum_k\left[\nabla_X(f_k+ig_k)\left({1\over2}(e_k-iJe_k)\right)\right]
181: \bigwedge_j(f_j+ig_j). \label{formula}
182: \end{eqnarray}
183: 
184: Taking covariant derivatives on the Calabi-Yau (i.e. not on $L$), we have
185: $$
186: -\langle\MCV,JX\rangle=-\langle\sum_i\nabla_{e_i}e_i,JX\rangle=
187: \sum_i\langle\nabla_{e_i}Je_i,X\rangle,
188: $$
189: since $J$ is both skew adjoint and parallel. But as $Je_i$ and
190: $X$ are orthogonal, this is
191: $$
192: -\sum_i\langle Je_i,\nabla_{e_i}X\rangle=\sum_i\langle e_i,J\nabla_Xe_i
193: \rangle,
194: $$
195: as we may choose $X$ to have zero bracket with the $e_i$s.
196: 
197: So comparing with (\ref{formula}) we are left with showing that
198: $$
199: \sum_k\left[\nabla_X(f_k+ig_k)\left({1\over2}(e_k-iJe_k)\right)\right]
200: =i\sum_i\langle e_i,J\nabla_Xe_i\rangle,
201: $$
202: i.e. that $(\nabla_X(f_k+ig_k))(e_k-iJe_k)=2i\langle e_k,J\nabla_Xe_k
203: \rangle$.
204: 
205: But $(f_k+ig_k)(e_k-iJe_k)=2$ is a constant, so the left hand side
206: is $-(f_k+ig_k)(\nabla_X(e_k-iJe_k))=-f_k(\nabla_X(-iJe_k))-
207: ig_k(\nabla_Xe_k)$; the other terms vanish as $\nabla_Xe_i$ was
208: chosen to be perpendicular to $L$. Using $\nabla_XJ=0$ and
209: recalling that $g_k=-f_k\circ J$, we obtain $2i\langle e_k,
210: J\nabla_Xe_k\rangle$.
211: \end{Proof}
212: 
213: Another simple but important result is how the phase $\theta$ and
214: volume form $\vol_L$ vary under a hamiltonian deformation
215: $J\,\widetilde{\!dh\,}$ of $L$. Such calculations appear in various forms
216: in \cite{Oh1}, \cite{Sm}, for instance; we give short geometric proofs for
217: completeness.
218: 
219: \begin{Lemma}
220: Under a hamiltonian deformation $\Jdh$ of a
221: Lagrangian $L$, we have
222: \begin{eqnarray}
223: {d\over dt}\theta\!&=&\!-\Delta_L(h), \label{h} \\ \nonumber
224: {d\over dt}\vol_L\!&=&\!-\langle d\theta,dh\rangle\vol_L.
225: \end{eqnarray}
226: \end{Lemma}
227: 
228: \begin{Proof}
229: Take real and imaginary parts of $e^{-i\theta}$
230: times the following:
231: \begin{eqnarray*}
232: i\dot\theta e^{i\theta}\vol_L\!\!\!&+&\!\!\!e^{i\theta}{d\over dt}\vol_L
233: ={d\over dt}(e^{i\theta}\vol_L)=(\mathcal L_{\Jdh}\Omega)\res L=
234: d(\Jdh\ip\Omega)\res L
235: \\ &=&\!\! id(e^{i\theta}\,\widetilde{\!dh\,}\ip\vol_L)=-e^{i\theta}
236: d\theta\wedge(\,\widetilde{\!dh\,}\ip\vol_L)-ie^{i\theta}d^*dh\vol_L.
237: \end{eqnarray*}
238: Setting $\Delta_L=d^*d$ $(=-\sum_i\partial^2_{x_i}$ in geodesic
239: coordinates) gives the result.
240: \end{Proof}
241: 
242: We next show that, given a suitable metric on the Lie
243: algebra $C^\infty(L,\R)/\R$, the gradient flow of the norm square
244: $-|m|^2$ of the moment map $m=\ImO\res L$ of \cite{Th} is mean
245: curvature flow. The following standard calculation, applicable in any
246: K\"ahler reduction picture, shows that the gradient flow of
247: $-|m|^2$ is given by $JX_{m^*}$, where $J$ is the complex structure,
248: $m^*$ is the element of the Lie algebra $C^\infty(L)$ corresponding to
249: the moment map $m=\ImO\res L$ in the dual of the Lie algebra under the
250: metric on $C^\infty(L)$, and $X_{m^*}$ is its induced action on the
251: space \{Lagrangians with flat $U(1)$ connections on them\}.
252: $$
253: X(-|m|^2)=-X(m(m^*))=-\omega(X,X_{m^*})=\langle X,JX_{m^*}\rangle.
254: $$
255: By the definition of the group action in \cite{Th}, this
256: deformation $JX_{m^*}$ is just the hamiltonian deformation of the
257: Lagrangian $L$ with hamiltonian function $m^*$ on $L$.
258: 
259: Choosing the volume form Re\,$\Omega\res L$ on $L$ to define an $L^2$
260: metric on $C^\infty(L)$ gives $m^*=\tan\theta$, since
261: $m=\ImO\res L=\sin\theta\,$vol$\,=\tan\theta\,$Re\,$\Omega\res L$.
262: Similarly using the induced Riemannian volume form vol gives
263: $m^*=\sin\theta$, while using
264: $$
265: {\sin\theta\over\theta}\,\mathrm{vol}
266: $$
267: as volume form on $L$ yields $m^*=\theta$.
268: Any of these are suitable for small phase
269: $\theta:\,L\to\R$, and give similar flows down which the moment map
270: decreases. The last one, however, is precisely mean curvature flow, by
271: Lemma \ref{mcf}.
272: 
273: This and the previous lemma show that under mean curvature flow,
274: the phase $\theta$ satisfies a (time dependent) heat equation while
275: the Riemannian volume form decreases (as usual):
276: \begin{eqnarray}
277: \dot\theta\!&=&\!-\Delta\,\theta, \label{thetadot} \\
278: {d\over dt}\vol_L\!&=&\!-|d\theta|^2\vol_L. \label{voldot}
279: \end{eqnarray}
280: We therefore obtain a maximum principle for $\theta$, whose range must
281: always decrease, but it is important to note that the Laplacian $\Delta$
282: is time dependent as the metric on $L$ used to define it varies.
283: 
284: From these follow a series of identities and estimates, many of which
285: we use later, but none are strong enough to
286: give long term existence of the mean curvature flow, and with good
287: reason. Mean curvature flow is a complicated and much-studied
288: subject (understood only in codimension 1, dimension 1 \cite{Gr}, and,
289: in special cases, in
290: two dimensions \cite{Wa}), with known examples of finite
291: time blow-up. While we might expect it to behave better for Lagrangians
292: (locally functions of one variable instead of $n$), examples in
293: Section \ref{eg} show similar phenomena. But in our examples
294: there will be a way round these problems, and we will be able to make
295: a conjecture about the flow which may help in its study.
296: 
297: 
298: \section{Connect sums and Floer gradings} \label{index}
299: 
300: The stability definition in \cite{Th} made extensive use of
301: graded Lagrangian connect sums \cite{S2}; a description
302: of these and their relationship to Floer cohomology will be
303: important again here, as will knowledge of the Floer index of
304: Lagrangian intersections. We fix our conventions and definitions
305: now; in some places these differ in orientation from some of
306: the mirror symmetry literature and \cite{S2};
307: the problem seems to be deciding on whether to use the standard
308: symplectic form $dxdy$ on $T^2$, or the equally standard
309: $dpdq=-dxdy$ considering it as the cotangent bundle of its
310: SYZ base $S^1$ (divided by a lattice) \cite{SYZ}.
311: 
312: \subsection{The connect sum}
313: 
314: Suppose we have two Lagrangians $L_1,\,L_2$ hamiltonian isotoped
315: to intersect transversally in a finite number of points. We will
316: work at one of these points $p$. There we can pick a local
317: Darboux chart with coordinates $(x_i,y_i)$ and symplectic
318: form $\sum_idx_i\wedge dy_i$ such that $L_2=\{y_i=0\}$ is the
319: $x$-axes, and
320: \begin{equation} \label{def}
321: L_1=\{y_i=\tan(\alpha)x_i\}
322: \end{equation}
323: for some $\alpha\in(0,\pi)$. (It would be more usual to use
324: $\alpha=\pi/2$, of course, but that situation can be moved to
325: this one by an obvious symplectic (shear) transformation).
326: 
327: Using $z_i=x_i+iy_i$ coordinates to set up the obvious isomorphism
328: to $\C^n$ (notice that this complex structure and that inherited from
329: $X$ may be different), the $L_i$ are
330: $$
331: L=e^{i\alpha}[0,\infty).S^{n-1}:=\{z_j=re^{i\alpha}a_j\,:\,r\in
332: [0,\infty),\,\mathbf a=(a_j)_{j=1}^n\in S^{n-1}\subset\R^n\subset\C^n\},
333: $$
334: where $\alpha$ is set to zero to give $L_2$.
335: 
336: So given a curve $\gamma$ in $\C$, we define a Lagrangian
337: $$
338: L_\gamma=\gamma.S^{n-1}=\{z_j=\gamma a_j\,:\,\mathbf a=(a_j)_{j=1}^n
339: \in S^{n-1}\subset\R^n\subset\C^n\}.
340: $$
341: Then $L_2$ is represented by $\gamma_2=[0,\infty)\subset\C$,
342: $L_1$ by $\gamma_1=e^{i\alpha}[0,\infty)\subset\C$, and $L_1\cup L_2$
343: by the V-shaped union of these curves.
344: 
345: In this notation the \emph{Lagrangian connect sum} $L_1\#L_2$ is
346: represented by any smoothing $\gamma=:\gamma_1\#\gamma_2$
347: of $\gamma_1\cup\gamma_2$ staying inside the cone
348: $\{re^{i\beta}\,:\,r>0,\,\beta\in[0,\alpha]\}$
349: which is $\gamma_1\cup\gamma_2$ outside a compact set, and a
350: smooth curve cutting off the cone at the origin. (So here
351: $\gamma$ is \emph{not} a connect sum of the curves $\gamma_i$
352: in the topological sense; we only use the notation because the
353: resulting Lagrangians are topological connect sums.)
354: 
355: We want to analyse the phase of such a connect sum; initially
356: in the complex structure we picked on $T_pX\cong\C^n$ using the
357: $x_i+iy_i$ coordinates and so, up to scale, $\Omega\res p=
358: dz_1\ldots dz_n$. Then the phase function of the Lagrangian
359: $L_\gamma$ associated to a curve $\gamma$ is easily calculated to
360: be $\theta(\gamma')+(n-1)\theta(\gamma)+N\pi$ for any $N\in\Z$
361: (where $\theta(z)\in(-\pi,\pi]$ is the phase of a
362: complex number $z=re^{i\theta(z)}$). Orienting $\gamma_2$
363: such that $\gamma_2'$ is a positive real number, and choosing
364: $L_2$ to have phase 0, corresponds to choosing $N=0$ and so grading
365: $L_1$ by
366: \begin{equation} \label{pa}
367: (-\pi+\alpha)+(n-1)\alpha=n\alpha-\pi.
368: \end{equation}
369: In particular, choosing $\alpha=\pi/n$, and setting, for any $c>0$,
370: $$
371: \gamma^{\ }_c=
372: \{re^{i\theta}\,:\,r^n=c\sin(n\theta),\ \theta\in(0,\pi/n)\}
373: $$
374: gives a SLag $L_\gamma$ which has a grading of phase identically
375: zero, asymptotic
376: to $L_1$ and $L_2$ at infinity, and this is precisely the local
377: model of the example of Joyce, Harvey and Lawlor used so
378: extensively in \cite{Th}. 
379: 
380: While this is not strictly of the form $\gamma_1\#\gamma_2$ as defined
381: above (it is only asymptotic to the $\gamma_i$, not equal outside
382: a compact set), by taking $c$ small we can make it as close as we like
383: to such a connect sum, all in the same hamiltonian deformation class,
384: and the construction of Joyce is indeed a hamiltonian deformation of a
385: connect sum as claimed in \cite{Th}.
386: 
387: We plot these SLag curves $\gamma_c\subset\C$ in Figure 1
388: as the light lines, converging as $c\to0$ to the V-shaped
389: $\gamma_1\cup\gamma_2$ (with $\alpha=\pi/n$). Then the dark lines
390: depict connect sums $L_1\#L_2$ for $\phi(L_2)=0$ and $\phi(L_1)=
391: \pm\epsilon$. If $\phi(L_1)<0$, the \emph{stable case} as described
392: in \cite{Th}, then we can choose the connect sum
393: such that the phase of $L_1\#L_2$ varies monotonically between its
394: values on $L_1$ and $L_2$, i.e. between $-\epsilon$ and 0.
395: 
396: \begin{figure}[h]
397: \center{
398: \input{hash.pstex_t}
399: \caption{$\gamma_1\#\gamma_2$, and the resulting phase function
400: $\theta_{L_1\#L_2}$, for $\phi(L_1)=\pm\epsilon$}}
401: \end{figure}
402: 
403: If, however, $\phi(L_1)>0$, the \emph{unstable case} in \cite{Th},
404: then the phase of $L_1\#L_2$ must initially decrease to move away
405: from $L_2$ before increasing to reach $L_1$ (i.e. $\gamma$ must cross
406: the light lines one way then the other), giving a phase
407: function which necessarily goes outside the range $(0,\epsilon)$
408: (see Figure 1).
409: This will be important to us later -- under mean curvature flow
410: we expect the phase function $\theta$ to evolve to a constant
411: in the stable case (under the heat equation (\ref{thetadot})) and
412: to a Heaviside step function (with values $0$ and $\epsilon$)
413: in the unstable case. This does not then contradict the maximum
414: principle as the unstable case has the described non-monotonic phase.
415: 
416: While this defines the symplectic connect sum in general by means
417: of our Darboux chart, the analysis of phases depended on the choice
418: of complex structure on $T_pX$, which may not have been the one
419: restricted from the Calabi-Yau $X$. In the general case we can fix
420: $\theta_p(L_2)=0$, without loss of generality, by rotating $\Omega$,
421: and pick local complex
422: coordinates $z_i=x_i+iy_i$ such that $\Omega_p=dz_1\ldots dz_n$ and,
423: at the level of tangent spaces at $p$, (the tangent space to) $L_2$
424: is at $y_i=0\ \forall i$. (The tangent space to) $L_1$ will be of the
425: form
426: \begin{equation} \label{l1}
427: L_1=\{z_i=re^{i\alpha_i}\},
428: \end{equation}
429: for some $\alpha_i$s that are \emph{no longer necessarily all the same}
430: in these coordinates. We are now connect summing Lagrangians of pointwise
431: phase $0$ and $\sum_{i=1}^n(\alpha_i)-\pi$ (compare (\ref{pa})), but
432: the resulting phase function will not be as simple as before -- it
433: is not pulled back from $\gamma$ but will vary over the $S^{n-1}$
434: fibres. Its average phase over the $S^{n-1}$s will have a similar
435: form to that in Figure 1, however, and in the case of all the
436: $\alpha_i$s being the same we get the earlier simpler picture.
437: 
438: The dependence of the hamiltonian deformation class of
439: $L_1\#L_2$ on the choice of scale of the neck at each intersection point
440: was described in (\cite{Th} Section 4) (in particular if there is
441: only one intersection point the class is uniquely defined). We should
442: also point out that the graded connect sum (when it exists)
443: is also independent of hamiltonian deformations of $L_1$ and $L_2$. While
444: the $L_i$s intersect transversely this is clear; we need only understand
445: what happens in crossing the codimension one wall of Lagrangians
446: intersecting in a double point (i.e. creating or cancelling two
447: intersection points). But it will be clear from the definition of grading
448: below that two such points must have grading differing by one, and
449: so the connect sum along both of them cannot be graded (\ref{grhash}).
450: 
451: 
452: \subsection{The grading on Floer cohomology}
453: 
454: The Floer cohomology group $HF^*(L_2,L_1;\C)$ \cite{FO3}
455: is the cohomology of a cochain complex made from a copy of $\C$
456: for each intersection point of two \emph{graded} Lagrangians
457: hamiltonian isotoped to intersect transversally.
458: The differential is defined by counting holomorphic
459: strips, with boundary in the Lagrangians, running from one intersection
460: point to another. It is a symplectic refinement of the
461: topological intersection theory of $L_1,\,L_2$, and as such is invariant
462: only under hamiltonian deformations of the $L_i$.
463: What is important to us is the grading of a particular transverse
464: intersection point, as defined in \cite{S2}, \cite{FO3}.
465: 
466: While this can be defined completely topologically, it is most easily
467: (and equivalently) defined via a complex structure. Again we
468: work at the level of tangent spaces, pick local coordinates and,
469: without loss of generality, take $L_2$ to have phase 0 and to be
470: the $x$-axes: $L_2=\{y_i=0\}$. Write $L_1$ as
471: $$
472: L_1=\{z_i=re^{i\alpha_i}\},
473: $$
474: \emph{where the $\alpha_i$s are all in} $(0,\pi)$. Then
475: $\sum\alpha_i=\theta_p(L_1)$ mod $\pi$ is the phase of $L_1$ up to
476: multiples of $\pi$, and the following integer
477: is the \emph{definition} of the Floer index of the point $p$:
478: \begin{equation} \label{maslov}
479: \ind_p(L_2,L_1):={1\over\pi}\left(\sum\alpha_i-\theta_p(L_1)\right).
480: \end{equation}
481: Notice therefore that $\ind_p(L_2,L_1)+\ind_p(L_1,L_2)=n$.
482: Applying the definition (\ref{maslov}) to the connect sums defined
483: in the last section (for which $\theta_p(L_1)=\sum\alpha_i-\pi$),
484: we recover a result of Seidel \cite{S2}:
485: \begin{equation} \label{grhash}
486: L_1\#L_2 \text{\emph{ exists as a graded connect sum if and only if }}
487: \ind_p(L_2,L_1)=1.
488: \end{equation}
489: (The only if
490: part follows from the independence of gradings and the Floer index
491: from the complex structure; we may therefore pick the complex
492: structure locally to have the form of the local model above.)
493: Given $L_1$ there is at most one choice of the grading on $L_2$
494: such that $\ind_p(L_2,L_1)=1$ at all intersection points $p$, so that
495: $L_1\#L_2$ can be graded.
496: 
497: In fact connect sums $L_1\#L_2$ whose own Floer cohomology is well
498: defined \cite{FO3} should correspond to Floer coclosed cochains, i.e.
499: elements of $HF^1(L_2,L_1)$,
500: mirror to the fact that extensions of sheaves $0\to E_1\to E\to E_2\to0$
501: correspond to elements of Ext$^1(E_2,E_1)$, as discussed in \cite{Th}.
502: 
503: We can also deal with the connect sums mentioned in \cite{Th}
504: which are \emph{relative} versions of the above construction;
505: $(n-r)$-dimensional connect sums carried out in a smooth family
506: over an $r$-dimensional base. Then the same Floer
507: index can be defined; there are now $r$ angles between the Lagrangians
508: that are zero, and $(n-r)$ whose signs can be computed to get the Floer
509: index. (The signs are constant over the family since the intersection
510: of the Lagrangians $L_1\cap L_2$ fibres over the base of the family
511: with fibres of constant dimension; an angle going to zero would cause a
512: fibre dimension to increase.)
513: 
514: 
515: \section{Uniqueness}
516: 
517: In finite dimensional symplectic quotient problems, convexity
518: properties of the moment map prove uniqueness of its zeros (modulo
519: the action of the real group) in a complexified group orbit.
520: Translating this into our terms is not quite possible, because there
521: are hamiltonian deformations of $L$ which are not given by the flow
522: of a \emph{fixed} hamiltonian on $L$. By this we mean $L_0,\ L_1$ are
523: deformations given by a constant hamiltonian $h\in C^\infty(L_0;\R)$
524: if the flow
525: \begin{equation} \label{flow}
526: f_t:\,L\to W, \quad {df\over dt}=\Jdh, \qquad t\in[0,1],
527: \end{equation}
528: takes $L_0=f_0(L)$ to $L_1=f_1(L)$.
529: All small deformations of a Lagrangian are of this form; for more
530: general deformations we have to use a
531: different proof of uniqueness of a SLag representative of a
532: hamiltonian deformation class (Proposition \ref{!} below), but for
533: these constant hamiltonian deformations
534: we describe the moment map proof to show how the formalism works.
535: 
536: \begin{Lemma} \label{?}
537: If two SLags $L_0,\ L_1$ are time-independent hamiltonian deformations
538: of each other, in the sense above, then $L_0=L_1$.
539: \end{Lemma}
540: 
541: \begin{Proof}
542: Without loss of generality we may take $\phi(L_0)=\phi(L_1)=0$. Then
543: we compute, down the flow (\ref{flow}),
544: $$
545: {d\over dt}\int_L h\ImO=\int h\,\mathcal L_{\Jdh}
546: \ImO=\int h\,d(\Jdh\ip\ImO)=\int\cos\theta\,dh\wedge*dh,
547: $$
548: where the last identity (equation (3.2) of \cite{Th}) is an easy
549: computation in local coordinates.
550: (We have abused notation and written $\ImO$ for $f_t^*\ImO$.)
551: 
552: So for $\theta$ lying in $(-\pi/2,\pi/2)$ this is always strictly
553: positive, but $\int_L h\ImO$ is zero at $t=0,\,1$. Thus
554: the two SLags must in fact coincide.
555: 
556: However, we must show that $\theta$ stays in this range if it starts
557: in it, and deal with the case when it is not so bounded. The way
558: to do this in fact proves the whole Lemma in one go anyway: pick a
559: maximum $x\in L$ of $h$. Then by (\ref{h}) $\dot\theta(x)\le0
560: \ \ \forall t$, but under the flow $\theta$ starts and ends at the
561: same value (the \emph{cohomologically} determined phase $\phi$ of the
562: SLags). So $\Delta h=0$ and all the second derivatives $h_{xx}\le0$
563: in any direction $\partial_x$ vanish. Similarly then all third
564: derivatives of $h$ must vanish (since we are at a maximum). Apply
565: the same procedure to the second derivatives $h_{xx}$ of $h$:
566: $\dot\theta_{xx}=-\Delta_Lh_{xx}\le0$ at $x$, since the other terms
567: $[\partial^2_x,\Delta_L]\,h$ involve derivatives of the metric times
568: third and lower order derivatives of $h$. Thus $h$'s 4th order
569: derivatives vanish, and so on.
570: 
571: To get an integral form of this, to show that
572: $h$ is in fact constant, it is enough to show that $h$ is constant in
573: a small ball around any global maximum $x$ (with $h(x)=0$, without loss
574: of generality). Consider geodesic balls $B_r$ of radius $r$ about $x$,
575: and their boundary spheres $S_r$. Fix a standard
576: unit-volume $(n-1)$-form $d\mu$ on the spheres $S_r$, so that the
577: volume form induced by the metric is $\partial_r\ip\vol=A(r)d\mu$,
578: where $r^{1-n}A(r)=c_1+O(r^2)$. For $r$ sufficiently small,
579: $d/dr\{r^{1-n}A(r)\}$ is
580: bounded by $c_2r$ (for some $c_2$ dependent only on the maximum of the
581: curvatures of $L$ in a neighbourhood of $x$ over time $t\in[0,T]$
582: of the flow). Therefore
583: $$
584: {d\over dr}\left\{{\int_{S_r}h\,A(r)d\mu\over r^{n-1}}\right\}=
585: r^{1-n}\int_{S_r}h_r\,A(r)d\mu+e,
586: $$
587: where $|e|\le c_3r^{2-n}\left|\int_{S_r}h\,A(r)d\mu\right|$.
588: Integrating over $r\in(0,R)$
589: for $R$ sufficiently small, and using the divergence theorem,
590: $$
591: f(R):=R^{1-n}\int_{S_R}h\,A(R)d\mu=
592: \int_0^Rr^{1-n}\left(\int_{B_r}\Delta h\vol\right)dr+E,
593: $$
594: where $|E|\le c_3\int_0^R|rf(r)|dr$.
595: Since $\Delta h=\dot\theta$, and $\int_0^T\dot\theta dt=0$
596: (where the hamiltonian deformation is over time $t\in[0,T]$,
597: and $\theta$ starts and ends at the same value), we see that
598: $$
599: \int_0^T|f(R)|dt\le c_3\int_0^T\int_0^R|rf(r)|dtdr
600: $$
601: for all small $R$. From this it follows that $f(r)\equiv0$.
602: That is, the average value of $h\le0$
603: over all small spheres surrounding $x$ (averaged over time
604: as the metric on $L$ varies) is zero, and $h$ must be identically
605: zero in a neighbourhhod of $x$.
606: \end{Proof}
607: 
608: However, we can do better by mirroring the algebro-geometric argument
609: that a non-zero map between stable bundles of the same slope is an
610: isomorphism, using the grading on Floer cohomology (\ref{maslov}).
611: This will appear to be slightly magical; the crux of the argument is
612: the hamiltonian isotopy invariance of Floer cohomology, provided
613: by precisely the holomorphic discs in the theory about which we
614: have had so little to say. Again we work in $n$ dimensions.
615: 
616: \begin{Theorem} \label{!}
617: Pick a connected graded Lagrangian $L$ whose obstructions
618: \emph{\cite{FO3}}
619: to the existence of its Floer cohomology vanish, and whose second
620: Stieffel-Whitney class $w_2$ is the restriction of a class
621: $\in H^2(X;\Z/2)$ on the whole manifold (for instance if $L$ is spin).
622: 
623: Then there can be at most one smooth special Lagrangian in the
624: hamiltonian deformation class of $L$.
625: 
626: In particular, SLag homology spheres are unique in their
627: hamiltonian deformation class in dimension 3 and above.
628: \end{Theorem}
629: 
630: \begin{Proof}
631: Since Floer cohomology is independent of hamiltonian deformations
632: \cite{FO3}, any two SLags $L_1,\,L_2$ in this same hamiltonian
633: deformation class satisfy
634: $$
635: HF^0(L_1,L_2)=H^0(L_1;\C)=\C,
636: $$
637: given that the zeroth order piece of $H^*(L)$ survives in $HF^*(L,L)$
638: for $L$ with Maslov class zero (\cite{FO3} Theorem E 1.7.4).
639: Thus there must be at least one intersection point $p$ of $L_1$ and $L_2$.
640: 
641: We first want to show that the (constant) phases of the $L_i$ are the same;
642: all we know a priori is that they differ by $r\pi$ for some $r\in\Z$.
643: Using a hamiltonian perturbation we may assume then that there is at least
644: one transverse intersection between the $L_i$s of Floer index 0, with the
645: phases of the $L_i$ at this point differing by $r\pi+\epsilon$. Thus writing,
646: locally, $L_2$ as the graph in $T^*L_1$ of $df$, $f$ has Morse index $r$
647: at the intersection point, so $r\ge0$. Similarly there is
648: a point of Floer index $n$ (i.e. a point of Floer index 0 when the roles
649: of $L_1$ and $L_2$ are reversed) which must correspond locally to $d$ of
650: a function of Morse index $n+r$; so $r\le0$ also. Therefore $r=0$ as
651: required.
652: 
653: So we have SLags $L_1,\,L_2$ of the same pointwise phase with at
654: least one intersection point which, if isolated, must have Floer index
655: (\ref{maslov}) zero. Thus $\theta_p=0$ in (\ref{maslov}), and the definition
656: (\ref{maslov}) of the index is therefore always positive (in fact between $0$ and
657: $n$), and zero only if the relative angles $\alpha_i=0\ \ \forall i$.
658: So the $L_i$ are tangent at $p$.
659: 
660: So there is no isolated transverse intersection point of Maslov class zero.
661: In fact, working in a small neighbourhood of the intersection, we may choose
662: coordinates such that $L_1$ is the graph in $T^*L_2$ of a closed one-form
663: $\sigma$ on $L_2$ which is also coclosed in a certain metric on $T^*L_2$
664: in a first order infinitesimal neighbourhood of $L_2$ (coclosedness is
665: the \emph{special} Lagrangian condition). In this small open set, write
666: $\sigma=df$, so that $d^*df=0$ and $f$ is harmonic; thus by the maximum
667: principle, it has no local maxima or minima. We want to show that $f$
668: is in fact constant. The Floer index (\ref{maslov})
669: of intersection points $df=0$ now reduces to the Morse index of $f$
670: at isolated critical points, but we also have to deal with degenerate
671: critical points of $f$. Assuming for a contradiction
672: that the critical set of $f$ is not all of $L_2$, we may perturb
673: $f$ inside any connected component of a small
674: neighbourhood of its critical set such that its value is unchanged on
675: the boundary, \emph{where it attains its global maximum and minimum},
676: and is Morse in the interior. (That we may take the extrema to be
677: on the boundary is the key point and a consequence of the maximum
678: principle.) We can then perturb $f$ further to
679: arrange its index 1 critical points to be lower (with respect to $f$)
680: than all higher index points (by general position arguments
681: \cite{Mi} Theorem 4.8) and then cancel any local minima with them
682: (\cite{Mi} Theorem 8.1). (There must be index 1 critical points if
683: there are any interior minima, by connectivity of our neighbourhood.)
684: 
685: The upshot is a hamiltonian perturbation of $L_1$, using this new
686: function, with no Floer index zero intersection points with $L_2$.
687: Thus $HF^0(L_1,L_2)=0$, a contradiction, so in fact
688: $f$ was locally constant and $L_1=L_2$.
689: 
690: The final statement follows from the fact that the obstructions of
691: \cite{FO3} live in $H^2(L)$, and homology spheres are spin.
692: \end{Proof}
693: 
694: As Donaldson pointed out, this proof is similar in flavour to proofs
695: of the Arnold conjecture. If the local situation (of all hamiltonian
696: deformations coming from a fixed function) held globally, the proof
697: would be `trivial', i.e. that of Lemma \ref{?} above. Even more
698: simply, if one SLag is a graph in the cotangent bundle of another,
699: we reduce the problem to the uniqueness of harmonic functions of
700: integral zero on $L$, i.e. to $H^0(L;\C)=\C$. To extend this
701: argument globally we need to replace de Rham cohomology $H^0(L;\C)$
702: by Floer cohomology $HF^0(L,L;\C)$.
703: 
704: 
705: \section{Analogues of some properties of sheaves}
706: 
707: In this section we discuss more
708: properties of (S)Lags that mirror those of holomorphic vector bundles
709: on Calabi-Yau manifolds.  As they rely heavily on Floer cohomology
710: arguments, many of the topics are necessarily treated informally and
711: unrigorously for now.
712: 
713: \subsection{Twisting by line bundles}
714: 
715: Any coherent sheaf can be twisted by a sufficiently positive
716: line bundle $\OO(N)$ so that it has sections; equivalently there are
717: homomorphisms to the bundle from any sufficiently negative line bundle.
718: If the sheaf has global support, this homomorphism is injective,
719: exhibiting $E$ as an extension
720: $$
721: 0\to\OO(-N)\to E\to Q\to0.
722: $$
723: One test of our notion of subobject of Lagrangians (in terms of connect
724: sums), then, is that there should be appropriate connect sums mirroring
725: this extension.
726: 
727: A line bundle $\mathcal L$ defines a \emph{spherical object} \cite{ST}
728: of the derived category of sheaves on a Calabi-Yau manifold $X$;
729: that is Ext$^i(\mathcal L,\mathcal L)=H^{0,i}(X)\cong H^*(S^n;\C)$
730: is $\C$ in dimensions $0$
731: and $n$, and zero otherwise. These should be mirror to Lagrangian
732: homology spheres; we will consider only spheres here so that we can
733: use the graded Dehn twists \cite{S2} around them. Negativity
734: compared to some other Lagrangian may not make sense in general
735: (intuitively, the Lagrangian
736: might be mirror not to a sheaf but to an object of the
737: derived category with Homs in negative degrees, etc.) but instead
738: we can consider only those spheres $L$ with only degree zero
739: intersection points (\ref{maslov}) with a fixed Lagrangian $L'$.
740: 
741: Then it is indeed true that we can exhibit $L$ as a subobject of
742: $L'$: denoting by $T_L$ the (graded) symplectic Dehn twist about
743: $L$, simply note that
744: $$
745: L'\approx T_L^{-1}T_LL'\approx L\#[L'\#(L[\,1\,])]
746: $$
747: expresses $L'$ as a connect sum of $L$ and something else. These relations
748: can be shown by grading similar results in \cite{S1}. In general this will
749: not destabilise $L'$ due to the phase of $L$ being so negative.
750: 
751: 
752: \subsection{Stability of (S)Lags} \label{ssl}
753: 
754: It is usual in correspondences between stable objects in algebraic
755: geometry and solutions of the corresponding moment map PDE for one
756: direction of the correspondence to be reasonably straightforward to
757: prove, namely that objects which satisfy the PDE are stable.
758: 
759: While we cannot prove this for SLags, we can show, for SLags
760: satisfying Floer cohomology restrictions as in Theorem \ref{!}
761: (in particular for spheres), that they cannot be destabilised
762: by other SLags. (To test for stability of sheaves it is sufficient
763: to test only with \emph{stable} subsheaves; if the conjecture of
764: \cite{Th} is true then similarly we could test for stability of
765: Lagrangians by connect summing only SLags; this would then
766: be enough to prove the general stability of SLags.)
767: 
768: The idea is that if $\phi(L_1)>\phi(L)$, with both $L_1$ and $L$
769: SLags, then the Floer index of any intersection point of $L_1$
770: and $L$ is strictly positive (\ref{maslov}), almost by definition.
771: But if $L_1$ were to destabilise $L$, i.e. $L=L_1\#L_2$ for some
772: $L_2$, then there should be canonical morphisms $HF^0(L_1,L)\ne0$
773: and $HF^0(L,L_2)\ne0$, a contradiction.
774: 
775: The morphism from $L_1\#L_2$ to $L_2$, by which we mean an element of
776: \begin{equation} \label{hom}
777: HF^0(L_1\#L_2,L_2),
778: \end{equation}
779: can be described as follows (the element of $HF^0(L_1,L_1\#L_2)$
780: is similar). We use the description of the connect sum in Section
781: \ref{index}. Choose a Morse function $f$ on $L_2$ which has
782: local maxima at intersection points $p$ with $L_1$,
783: and in local Darboux charts as in Section \ref{index}, is
784: pulled up from a linear function on $\gamma_2$. Let the function
785: have a unique local minimum elsewhere on $L_2$, and now use
786: this to hamiltonian deform $L_2$ off $L_1\#L_2$. By construction
787: $L_1$ and $L_1\#L_2$ now intersect at the critical points of $f$
788: only, with Floer index the Morse index of $f$. In terms of
789: Figure 1, as $f$ has a maximum on $\gamma_2$ at the vertex of
790: $\gamma_2$, it defines a hamiltonian deformation of $\gamma_2$
791: \emph{downwards}, away from the connect-sum neck. As $L_2$ only
792: intersects $L_1$ near these connect-sum necks, we can make
793: our charts small enough that $L_2$ now only intersects $L_1\#L_2$
794: where its hamiltonian deformation intersects the old $L_2$, i.e.
795: at the critical points of $f$.
796: 
797: We now have a unique index zero point of $(L_1\#L_2)\cap L_2$ at
798: the unique local minimum of $f$. What we require is that
799: this survives in the passage to cohomology of the cochain complex
800: to give $HF^0(L_2,L_1\#L_2)\cong\C$. For instance, if there are no
801: index one points (i.e. $f$ is a Morse function with only minima and
802: index $\ge2$ critical points) then this will clearly be the case.
803: More generally there is a spectral sequence analogous to Po\'zniak's
804: \cite{P} with
805: $$
806: \mathrm{coker}\,\{\bigoplus_i\C_{p_i}[-n]\to H^*(L_2)\}\ 
807: \Longrightarrow\ HF^*(L_1\#L_2,L_2)
808: $$
809: (with a certain bigrading)
810: converging to $HF^*(L_1\#L_2,L_2)$. Here the notation means that a
811: copy of $\C$ is mapped to $H^n(L_2)$ (i.e. it is in degree $n$)
812: for every intersection point $p_i$ via the Morse theory for $f$ (whose
813: maxima are at the $p_i$). Therefore the degree zero part also survives
814: if, for instance, $H^1(L_2)=0$. If $L_2$ is a sphere we can proceed
815: more directly by applying Seidel's exact sequence \cite{S3}.
816: 
817: Using similar methods on Lagrangians rather than SLags, we can cut
818: down on the number of possible destabilising Lagrangians $L_1$
819: we must check to conclude that a given $L$ is stable, rather
820: analogously to only checking for subsheaves of vector bundles
821: amongst those of lower rank. There are
822: no morphisms (non-zero elements of $HF^0(L_1,L)$) if the phase
823: of $L_1$, at an intersection point $p$, is greater than that of
824: $L$; the Floer index at $p$ is strictly positive.
825: So for $L_1$ to destabilise $L$ (and so $HF^0(L_1,L)\ne0$ for
826: Lagrangians satisfying the same conditions as above and in (\ref{!}),
827: e.g. homology spheres) we must have
828: $$
829: \inf_{x\in L_1}\theta_{L_1}(x)<\sup_{x\in L}\,\theta_L(x),
830: $$
831: and in fact the corresponding phase inequality
832: at each point of intersection.
833: 
834: Thus we do not have to check all Lagrangians $L_1,\,L_2$ for the stability
835: of $L'$ in \cite{Th}, just those whose phase function satisfies
836: $$
837: \inf_{x\in L_1}\theta_{L_1}(x)\le\sup_{x\in L}\,\theta_L(x)
838: \quad\text{and}\quad
839: \sup_{x\in L_2}\theta_{L_2}(x)\ge\inf_{x\in L}\,\theta_L(x),
840: $$
841: where we can in fact replace the left hand sides of these inequalities
842: by the inf (respectively sup) over all Lagrangians in the same hamiltonian
843: deformation class.
844: 
845: Assuming the conjecture in \cite{Th}, so that we need only check
846: SLag destabilisers to show that stability of $L_0$, we are reduced to
847: checking for destabilising subobjects amongst those $L_1,\ L_2$ whose
848: homology classes sum to $[L_0]$ and satisfy
849: \begin{equation} \label{ineq}
850: \sup_{L\approx L_0}\left(\inf_{x\in L}\,\theta_L(x)\right)\le\phi(L_1)
851: \le\phi(L_2)\le\inf_{L\approx L_0}\left(\sup_{x\in L}\,\theta_L(x)\right).
852: \end{equation}
853: 
854: 
855: \subsection{A Jordan-H\"older decomposition for Lagrangians} \label{JH}
856: 
857: In order to understand limits of mean curvature flow it will be useful
858: to have the following concept;
859: an analogue for Lagrangians of the Jordan-H\"older filtration of
860: sheaves (see \cite{HuL} 1.5, for instance).
861: 
862: \begin{Definition}
863: Given two graded Lagrangians $L_1,\,L$, write $L_1\le L$ if there
864: exists a graded Lagrangian $L_1'$ such that $L\approx L_1\#L_1'$.
865: We then also write $L/L_1$ for $L'_1$, and say that $L_1$ is
866: a subobject of $L$.
867: 
868: A Jordan-H\"older filtration of $L$ is a sequence of graded
869: Lagrangians $L_i$ such that
870: $$
871: L_1\le L_2\le\ldots\le L_k=L,
872: $$
873: and $L_i':=L_{i+1}/L_i$ is stable. The Jordan-H\"older decomposition
874: of $L$ is the the singular union
875: \begin{equation} \label{decomp}
876: L_1\cup(L_2/L_1)\cup\ldots\cup(L/L_{k-1}).
877: \end{equation}
878: \end{Definition}
879: 
880: In sheaf theory the Jordan-H\"older filtration need not be unique,
881: but the decomposition is. For smooth \emph{connected} Lagrangians,
882: with connected $L_i$ for all $i$, however, we expect
883: the filtration to be unique too; the difference is essentially that
884: while direct sum is an operation on bundles, we are proposing that
885: its mirror is the (singular) union of Lagrangians, and this
886: cannot give a smooth Lagrangian if there is non-zero Floer cohomology
887: between the two Lagrangians.
888: 
889: If we assume the conjectures of \cite{Th}
890: and Section \ref{conj} below, and the properties of Floer
891: cohomology \cite{FO3} for all of the above Lagrangians (e.g. if they
892: are homology spheres), we can demonstrate the existence and uniqueness
893: of the Jordan-H\"older filtration for a Lagrangian $L$ whose phase
894: function of $L$ satisfies $\sup\theta_L-\inf\theta_L<\pi$.
895: 
896: Without loss of generality we may assume (by rotating $\Omega$) that 
897: $\theta$ lies between $\pi/2-\epsilon$ and $-\pi/2+\epsilon$, for
898: some $\epsilon>0$. By the inequality (\ref{ineq}) above, then, any
899: $L_1\#L'$ destabilising it will satisfy $\phi(L_1),\phi(L')\in
900: (-\pi/2+\epsilon,\pi/2-\epsilon)$.
901: 
902: We choose such an $L_1$ of maximal phase, and, amongst other such $L_1$s
903: of the same phase, minimal $\int_{L_1}\mathrm{Re}\,\Omega$ (for the
904: purposes of this proof we will call this quantity cohomological volume).
905: This still need not specify $L_1$ uniquely though.
906: 
907: We claim that such an $L_1$ must
908: be stable by construction. Any subobject of $l\le L_1$ would also be
909: a subobject of $L$ and so by the construction of $L_1$ must either
910: have smaller phase, which is not possible since it destabilises $L_1$,
911: or equal phase and greater or equal cohomological volume.
912: But $L_1=l\#l'$, where $l'=L_1/l$ has phase
913: $\phi(l')=\phi(L_1)\in(-\pi/2,\pi/2)$ and so positive cohomological
914: volume $\int_{l'}\mathrm{Re}\,\Omega$. So the complex numbers
915: $$
916: \int_{L_1}\Omega=\int_l\Omega+\int_{l'}\Omega
917: $$
918: all have positive real part, implying that the cohomological volume
919: of $l$ is strictly less that that of $L_1$, a contradiction.
920: 
921: We then apply the same
922: procedure to $L'$, producing an $L_2\into L'$, and so on.
923: By construction $\phi(L'/L_2)\le\phi(L')\le\phi(L)<\pi/2-\epsilon$,
924: and there is a canonical morphism (\ref{hom}) in
925: $HF^0(L,L'/L_2)\ne0$, making $\phi(L'/L_2)\ge\inf_L\theta_L>-
926: \pi/2+\epsilon$ by (\ref{ineq}).
927: 
928: Thus, inductively, we get the same inequalities at each stage, and
929: the cohomological volume of $L'$ decreases strictly with each
930: decomposition $L\approx L_1\#\ldots\#L_n\#L'$. The cohomological
931: volume of any $l$ with phase $\phi(l)\in(-\pi/2+\epsilon,\pi/2-\epsilon)$
932: is greater than (or equal to in the SLag case) $\cos(\pi/2-\epsilon)
933: \int_l\vol$, by (\ref{vol}), where $\vol$ is its
934: Riemannian volume form. This is bounded below above zero,
935: so the process can have at most a finite number of steps.
936: 
937: This gives us the Jordan-H\"older filtration; next we consider
938: uniqueness when the $L_i$s are connected (assuming the conjectures
939: of \cite{Th} and Section \ref{conj} and some Floer cohomology).
940: Suppose that $L_1'\le L_2'\le\ldots\le L$ is another such connected
941: decomposition where we take the $L_i'$s to be SLag assuming the
942: conjecture of \cite{Th}. If $HF^0(L_1',L_1)\ne0$ then
943: by the proof of Theorem \ref{!} (which applies as $L_1'$ and $L_1$
944: have the same phase), $L_1$ and $L_1'$ are equal, and we pass to $L_2$.
945: 
946: If, however, $HF^0(L_1',L_1)=0$, then we claim that
947: $HF^0(L_1',L/L_1)\ne0$. Again this should follow from standard
948: facts about Floer cohomology, in particular a long exact sequence
949: $HF^*(L_1',L_1)\to HF^*(L_1',L)\to HF^*(L_1',L/L_1)\to HF^{*+1}
950: (L_1',L_1)$. For $L/L_1$ a sphere this is Seidel's exact sequence
951: (\cite{S3} Theorem 3.3), and in general one can establish it at
952: the level of chains by good choices of hamiltonian perturbations
953: as in Section \ref{ssl}; as usual the problem is in controlling
954: the differential, i.e. holomorphic discs.
955: 
956: Assuming this we may pass to $L_2$; inductively we eventually obtain that
957: $L_1'$ is isomorphic to one of the graded pieces $L_{i+1}/L_i$ of
958: the original filtration, and is a subobject of $L_{i+1}$ but not of
959: $L_i$. But this gives us a contradiction (in contrast to the sheaf
960: analogue), since we have that both $L_{i+1}\approx L_i\#L_1'$ and
961: $L_1'$ is a subobject of $L_{i+1}$. The first condition ensures that
962: there are representatives of the hamiltonian deformation classes
963: such that $L_{i+1}$ and $L_1'$ have no index $n$ intersection points
964: by the construction of (\ref{hom}), so that $HF^n(L_{i+1},L_1')=0$.
965: But this is $HF^0(L_1',L_{i+1})^*$, which cannot vanish by the second
966: condition. (It is here we use the connectivity condition, i.e. that
967: the connect sum $L_{i+1}=L_1'\#(L_{i+1}/L_1')$ is not a trivial
968: disjoint union. Without the connectivity condition the usual
969: proof (e.g. \cite{HuL} 1.5) that the Jordan-H\"older
970: \emph{decomposition} (rather than filtration)
971: of sheaves is unique applies to Lagrangians, now that we have proved
972: or assumed all (the mirror analogues of) the algebraic facts used for
973: sheaves in terms of Floer cohomology instead of Exts. \\
974: 
975: So in the simplest case of instability, such as the example of Joyce
976: considered in \cite{Th}, where $L=L_1\#L_2$ is the only relevant
977: decomposition of $L$ with $\phi(L_1)\ge\phi(L_2)$, the Jordan-H\"older
978: decomposition (\ref{decomp}) would be simply $L_1\cup L_2$
979: (where the $L_i$ are SLag
980: representatives of their classes). This, like all such
981: decompositions, is in the closure of the hamiltonian deformation
982: orbit of $L$ while not being in the orbit itself.
983: 
984: This should have relevance to the Schoen-Wolfson programme \cite{SW} to
985: find canonical representatives (in a fixed hamiltonian
986: deformation class) of Lagrangian homology classes using volume
987: minimisers and so SLags (they do not use a flow, but regularity results
988: to study minimising currents). Our conjecture (as in \cite{Th} and later
989: in Section \ref{conj}) should either provide a unique SLag in a
990: hamiltonian deformation class, or a number of SLags in a Jordan-H\"older
991: decomposition.
992: 
993: For instance in the example above of $L_1\#L_2$ in 2 dimensions
994: we would produce
995: SLags in the classes of $L_1$ and $L_2$, but we could also form
996: $L_2\#L_1$; this could then be stable (it is no longer destabilised by
997: either of the $L_i$; if the phases of the $L_i$ are sufficiently close
998: one can show that in fact nothing else destabilises it either)
999: and we should recover a SLag in this class (and so in the same
1000: homology class in two dimensions).
1001: 
1002: Since in two dimensions SLags are just holomorphic curves with respect
1003: to a different complex structure, this places heavy restrictions
1004: on stability. Take the $L_i$ above to be spheres in $K3$ surfaces.
1005: Then any holomorphic sphere is unique in its homology class (it has
1006: negative self intersection $-2$, so does not lie in a pencil). Any
1007: other homologous hamiltonian deformation class
1008: must therefore be unstable. Good examples are provided by taking a
1009: stable (SLag/holomorphic) sphere, and applying the square of a Dehn
1010: twist $T^2_{L_1}$ to it; this preserves homology classes but can change
1011: hamiltonian deformation classes. If it does it should produce an
1012: unstable Lagrangian with copies of $L_1$ in its Jordan-H\"older
1013: decomposition; this happens in all simple cases.
1014: $L_1\#L_2$ is taken to $L_2\#L_1$,
1015: for instance; only one of these can be stable, the other having
1016: a Jordan-H\"older decomposition $L_1\cup L_2$ in the simplest case.
1017: 
1018: More generally, instead of studying the action on individual (S)Lags
1019: of symplectomorphisms like $T^2_{L_1}$ above, we could try to study
1020: them all at once by studying the Lagrangian graph of the
1021: symplectomorphism in $X\times X$, and its mean curvature flow. This
1022: looks for minimal energy representatives of the hamiltonian isotopy
1023: class of a symplectomorphism, and breaks graphs up into correspondences
1024: representing singular maps (birational maps in the hyperk\"ahler case)
1025: with singularities concentrated in loci whose stability is affected by
1026: the symplectomorphism. For a Dehn twist $T_L$, for instance, we would
1027: expect to get the graph $\Delta$ of the identity, union $L\times L$.
1028: This also shows what the analogue of a Dehn twist $T_L$ should be when
1029: $L$ is not a sphere but a rational homology sphere (so that it is still
1030: spherical to complex coefficients, and so mirror to a twist on the
1031: derived category of sheaves on the mirror Calabi-Yau \cite{ST}). Namely
1032: $\Delta\cup L\times L$ is a Lagrangian correspondence in $X\times X$
1033: which should give an automorphism of the derived Fukaya category of $X$
1034: (by the usual Fourier-Mukai-type construction) \emph{not} induced by a
1035: symplectomorphism of $X$.
1036: 
1037: 
1038: \section{An example: families of affine quadrics} \label{eg}
1039: 
1040: Here we consider an example suggested to us by both Paul Seidel and
1041: Cumrun Vafa, used in \cite{SV} and \cite{KS}. Consider the
1042: affine algebraic variety $X^n$ given by
1043: $$
1044: \sum_{i=1}^nx_i^2=p(t)
1045: $$
1046: in $\C^n\times\C$, where $p$ is some polynomial in $t\in\C$ with only
1047: simple zeros. Denote by $\pi:\,X^n\to\C$ the projection to the $t$
1048: coordinate. Here we use the K\"ahler structure restricted from
1049: $\C^{n+1}$, and the nowhere-zero holomorphic volume form given by
1050: taking the Poincar\'e residue (\cite{GH} p 147) of the standard form
1051: $dx_{1\ldots n}dt:=dx_1\ldots dx_{n\,}dt$ on $\C^{n+1}$;
1052: this can be written as
1053: \begin{equation} \label{form}
1054: (-1)^{n+i+1}{dx_{1\ldots\hat{\imath}\ldots n\,}dt\res{X^n}\over2x_i}=
1055: {dx_1\ldots dx_n\res{X^n}\over\dot p(t)}
1056: \end{equation}
1057: for any $i$ (so where $x_i=0\ \forall i$ we can use the second expression).
1058: Here $\hat{\imath}$ means that we omit the $dx_i$ term from the wedge product.
1059: This is then \emph{not} parallel, and the metric
1060: we have chosen is \emph{not the Ricci-flat one}. Nonetheless it is
1061: a good explicit testing ground for the conjecture; we can still define
1062: $\theta$ as the phase of $\Omega\res L$ and SLags as having constant phase,
1063: of course we then use flow
1064: by the $\Jd$ vector, rather than mean curvature flow in
1065: this metric. While the two flows are similar and would be the same in the
1066: Ricci-flat metric, only the former has SLags as its stationary points
1067: (for the latter we get minimal submanifolds, which in this metric are
1068: not quite SLag). As Edward Goldstein pointed out to us, the $\Jd$ flow
1069: is the gradient flow of the weighted volume functional $\int_L|\Omega|
1070: \vol$ instead of $\int_L\vol$; everything proceeds analogously
1071: to before on weighting all vols by $|\Omega|$, as we shall see.
1072: 
1073: Each smooth fibre over $t\in\C$ is an affine quadric with a natural
1074: Lagrangian $S^{n-1}$
1075: `real' slice, namely the intersection of the fibre with the slice
1076: $$
1077: x_i\in\sqrt{p(t)}\,\R\quad\forall i.
1078: $$
1079: It is invariant under the obvious $O(n)$ action on $X^n$, and is the
1080: vanishing cycle of every singular fibre (i.e. the fibres over the
1081: roots of $p$). Therefore any path
1082: $\gamma:\,I\to\C$ ($I\ni u$ being some interval in $\R$) from one
1083: zero of $p$ to another lifts to give a canonical
1084: $O(n)$-invariant Lagrangian $n$-sphere $\gamma^n$, $S^{n-1}$-fibred
1085: over $\gamma$ except at the endpoints where it closes up. Also,
1086: any vector $\gamma'\partial_t$ in the base $\C\ni t$ lifts canonically
1087: to a vector
1088: \begin{equation} \label{lift}
1089: \gamma'\left(\partial_t+{\dot p\over2p}\sum x_i\partial_{x_i}\right)
1090: \end{equation}
1091: tangent to the infinitesimal Lagrangian $\gamma^n$ lying above $\gamma'$.
1092: Here $'$ denotes $d/du$. Note that $\gamma^1$ is a closed curve double
1093: covering $\gamma$, branched over $\gamma$'s endpoints. We will use this
1094: curve $\gamma^1$ later to study $\gamma^n$.
1095: 
1096: The phase function $\theta$ on $\gamma^n$ is also $O(n)$-invariant and
1097: so a function of $t\in\C$ which we may calculate at $x_1=\sqrt{p(t)},\,
1098: x_i=0\ \forall i\ge2$. Choosing a basis of tangent vectors to $\gamma^n$
1099: at this point,
1100: \begin{equation} \label{basis}
1101: \gamma'(\partial_t+{\dot p\over2p^{1/2}}\partial_{x_1}),\,
1102: \sqrt p\,\partial_{x_2},\ldots,\sqrt p\,\partial_{x_n},
1103: \end{equation}
1104: wedging them together and evaluating against the $(n,0)$-form
1105: (\ref{form}) gives
1106: $$
1107: \gamma'{\dot p\over2p^{1/2}}{(\sqrt p)^{n-1}\over\dot p}=
1108: {1\over2}\gamma'p^{n/2-1}.
1109: $$
1110: Therefore the phase function on $\gamma^n$ is given by
1111: \begin{equation} \label{pf}
1112: \theta:=\theta(\gamma^n)=\theta(\gamma')+\big({n\over2}-1\big)
1113: \theta(p(\gamma)),
1114: \end{equation}
1115: where $\theta(\gamma')$ is the usual angle of the path $\gamma$, and
1116: $\theta(p)$ is the phase of the complex number $p$ evaluated at
1117: $t=\gamma$. So
1118: $$
1119: d\theta=\left({d\theta(\gamma')\over du}+\big({n\over2}-1\big)
1120: {d\theta(p)\over du}\right)du,
1121: $$
1122: where $du$ is the pullback to $\gamma^n$ under the projection $\pi$
1123: of the corresponding 1-form on $\gamma(I)\subset\C$. 
1124: 
1125: Using the metric and the orthogonal basis (\ref{basis}) we see
1126: that
1127: $$
1128: \,\widetilde{\!du\,}={\gamma'(\partial_t+{\dot p\over2p^{1/2}}
1129: \partial_{x_1})\over|\gamma'(\partial_t+{\dot p\over2p^{1/2}}
1130: \partial_{x_1})|^2}
1131: $$
1132: at the point $x_1=\sqrt{p(t)},\,x_i=0\ \forall i\ge2$.
1133: 
1134: Therefore, by $O(n)$-invariance and the holomorphicity of the
1135: projection $\pi$, $\Jd$ is the canonical lift (\ref{lift}) of
1136: $$
1137: {d(\theta(\gamma')+\big({n\over2}-1\big)\theta(p))\over du}\,
1138: {\pi_*\!\left[J\gamma'(\partial_t+{\dot p\over2p^{1/2}}
1139: \partial_{x_1})\right]\over|\gamma'|^2(1+|\dot p|^2/4|p|)}=
1140: {{1\over|\gamma'|}{d\over du}(\theta(\gamma')+\big({n\over2}-1\big)\theta(p))
1141: \over1+|\dot p|^2/4|p|}\,
1142: i{\gamma'\over|\gamma'|}\partial_t.
1143: $$
1144: Denoting by $\mathbf t=\partial_u/|\gamma'|=\gamma'\partial_t/|\gamma'|$
1145: and $\mathbf n=i\mathbf t$ the unit tangent and normal
1146: vectors to $\gamma$ at a point $\gamma(u)$, the above is
1147: $$
1148: {\mathbf t[\theta(\gamma')+\big({n\over2}-1\big)\theta(p)]\over
1149: 1+|\dot p|^2/4|p|}\,\mathbf n.
1150: $$
1151: By the Cauchy-Riemann equations for the holomorphic function
1152: $\log p=\log|p|+i\theta(p)$, $\mathbf t\theta(p)=-\mathbf n\log|p|$,
1153: so that our flow is the lift to $X^n$ of the flow of $\gamma$ with vector
1154: \begin{equation} \label{result1}
1155: V^n={1\over1+|\dot p|^2/4|p|}(\MCV+\,
1156: (1-n/2)\mathbf n(\log|p|)\,\mathbf n),
1157: \end{equation}
1158: where $\MCV$ is the usual mean curvature vector of $\gamma$ in
1159: the flat metric on $\C$.
1160: 
1161: So we can reduce studying our flow to studying the flow
1162: of a curve $\gamma$ with fixed endpoints (at zeros of $p$), under
1163: the above vector field. We would like to relate this to mean curvature
1164: flow of $\gamma\subset\C$ in a different metric, and also to
1165: both our flow and the mean curvature flow for the double
1166: $\gamma^1$ of $\gamma$ in the
1167: double cover $X^1$ of $\C$ branched over the zeros of $p$.
1168: The advantage of this is that we now have a flow for a closed
1169: curve instead of a boundary value problem (but since the flow
1170: has $O(1)=\Z/2$ symmetry it is equivalent to a flow of the
1171: original curve $\gamma$ with fixed endpoints). We need the following
1172: lemma.
1173: 
1174: \begin{Lemma}
1175: Let $\langle\,.\,,\,.\,\rangle$ be the standard metric on $\C$, and
1176: $g$ a positive real-valued function on $\C$. Then with respect
1177: to the metric $g\langle\,.\,,\,.\,\rangle$, the mean curvature vector
1178: of a curve $\gamma\subset\C$ is, in terms of the standard mean curvature
1179: vector $\MCV$ (and
1180: calculating the unit normal $\mathbf n$ in the standard metric),
1181: $$
1182: {1\over g}\left(\MCV-{1\over2}\mathbf n(\log g)\,\mathbf n\right).
1183: $$
1184: \end{Lemma}
1185: 
1186: \begin{Proof}
1187: The endomorphism-valued 1-form $\Gamma$ defined by
1188: $$
1189: \Gamma_XY={1\over2g}((Xg)Y+(Yg)X-\langle X,Y\rangle\,\widetilde{\!dg\,})
1190: $$
1191: is symmetric and so defines a torsion-free connection on $\C$. It
1192: is easily checked to be orthogonal with respect to the metric
1193: $g\langle\,.\,,\,.\,\rangle$, and so gives its Levi-Civita connection
1194: $\nabla+\Gamma$ (where $\nabla$ is the usual connection on $\C$).
1195: Then $\langle\Gamma_{\gamma'}\gamma',\mathbf n\rangle$
1196: (where $\mathbf n$ is calculated in the original metric) is
1197: $-{1\over2g}|\gamma'|^2\mathbf n(g)=-{1\over2}|\gamma'|^2\mathbf n(\log g)$.
1198: 
1199: Since the unit normal to $\gamma$ in the new metric is
1200: $g^{-1/2}\mathbf n$, the new mean curvature vector is
1201: $$
1202: {g\langle\gamma''+\Gamma_{\gamma'}\gamma',g^{-1/2}\mathbf n\rangle\over
1203: g|\gamma'|^2}\,g^{-1/2}\mathbf n={1\over g}\left(\MCV-{1\over2}
1204: \mathbf n(\log g)\,\mathbf n\right),
1205: $$
1206: as claimed.
1207: \end{Proof}
1208: 
1209: Using this we can get a number of geometrically interesting flows which
1210: are equivalent to our original flow in $X^n$. Namely, using the result
1211: (\ref{result1}), the above Lemma, and the fact
1212: that locally (away from branch points) $X^1$ is conformally equivalent
1213: to $\C$ with its metric scaled by $g=1+|\dot p|^2/4|p|$ (by
1214: holomorphicity and (\ref{lift})), we can deduce the following.
1215: 
1216: Denote by $V^n$ the flow vector $\pi_*\Jd$ of the
1217: curve $\gamma\subset\C$ under our flow in $X^n$. Denote by
1218: $\MCV^1_g$ the flow vector of $\gamma\subset\C$ under mean curvature
1219: flow of $\gamma^1$ in $X^1$, with $X^1$'s natural metric scaled by
1220: a $\Z/2$-invariant function $g$ (and omit the
1221: $g$ in the notation if $g\equiv1$). And denote by $\MCV_g$
1222: the mean curvature vector of $\gamma$ in $\C$ with metric $g
1223: \langle\,.\,,\,.\,\rangle$.
1224: 
1225: Letting $\mathbf n$ be the unit normal to $\gamma$ calculated in the
1226: standard metric on $\C$, and letting $f={|p|^{n-1}\over|p|+
1227: |\dot p|^2/4}$\,, we have the following relations between the
1228: various flows:
1229: \begin{equation} \label{result2}
1230: V^n=f.\MCV^1_f=f.\MCV_{|p|^{n-2}},
1231: \end{equation}
1232: and
1233: \begin{equation} \label{result4}
1234: V^n=\MCV^1-\left[{1\over2}|p|^{2-n}\mathbf n(f)\right]\mathbf n
1235: =V^1-{1\over2}(n-1){\mathbf n(|p|)\over|p|+|\dot p|^2/4}\,\mathbf n.
1236: \end{equation}
1237: 
1238: The problem with the first two is that on $\C\supset\gamma$ the flow
1239: is \emph{not} parabolic, it has degeneracies at the end points. As
1240: $X^1$ is so closely modelled on $X^n$ (and is in fact canonically
1241: embedded in it), however, we might expect better on $X^1$. This is
1242: more or less true; the result is that writing (\ref{result4})
1243: in terms of the unit normal $\mathbf n^1$ \emph{on} $X^1$, we get
1244: 
1245: \begin{Theorem} \label{result!}
1246: $$
1247: V^n=\MCV^1-{1\over2}(n-1)\mathbf n^1(\log|p|)\,\mathbf n^1+
1248: {1\over2}\mathbf n^1(\log(|p|+|\dot p|^2/4))\,\mathbf n^1.
1249: $$
1250: \end{Theorem}
1251: 
1252: The last term is bounded (as near a zero of $p$, $\dot p\ne0$
1253: by nondegeneracy of $p$'s zeros) and so unimportant, we shall
1254: see, and the flow
1255: resulting from the first term is well understood. The second term
1256: is more curious; it is of the order of $\mathbf t^1\theta(p)\approx
1257: \mathbf t^1\theta((\gamma^1)')/2$ (where $'=\mathbf t^1$ denotes
1258: differentiation with respect to arclength on $X^1$) whenever we are
1259: close to a point where $\gamma$ emanates from a zero of $p$
1260: (so that $p\approx Ct$ and $\theta(p)\approx\theta(\gamma)
1261: \approx\theta(\gamma')\approx\theta((\gamma^1)')/2$).
1262: (The last approximation is of course
1263: not valid if $\gamma$ simply passes close to a zero of $p$; then
1264: the equation blows up quickly as a glance at (\ref{result1}) shows,
1265: $\gamma$ flowing to
1266: this zero and breaking across it as discussed below; in the stable
1267: case we will be able to rule out this behaviour and need only consider
1268: $\gamma$ ending at the zero.)
1269: 
1270: But this is half the curvature of $\gamma^1$, so we get an
1271: approximation to the first term again, and something like mean
1272: curvature flow for $\gamma^1\subset X^1$. In fact in a small
1273: neighbourhood of (the double cover of) a zero of $p$, in coordinates
1274: $(x,y)$ in which $\gamma^1$ is a graph $y(x)$, the evolution
1275: PDE is of the general shape
1276: $$
1277: y_t=y_{xx}+{y_x\over x},
1278: $$
1279: where by the $\mathbb Z/2$-symmetry $y_x\res{x=0}=0$, so the second
1280: term is approximately $y_{xx}$.
1281: So for some analysis we use this flow for $\gamma^1$, while for the rest
1282: we pass back to $n$-dimensions, and work with the phase function
1283: $\theta$ instead, giving a more standard (but $n$-dimensional)
1284: parabolic equation.
1285: 
1286: We first assert how the flow behaves, before proving it in the
1287: stable case in the next section (Theorem \ref{faq}).
1288: Note that any deformation of
1289: $\gamma$ is a hamiltonian deformation of $\gamma^n$ (and SLag
1290: $\gamma^n$s have no moduli) since the $\gamma^n$s are spheres. We
1291: picture what happens in Figures 2 and 4 in the 2 and 3
1292: dimensional cases respectively. The dots represents zeros
1293: of $p$ in both cases, and the epsilons and zeros are phases.
1294: 
1295: \begin{figure}[h]
1296: \center{
1297: \input{stab2.pstex_t}
1298: \caption{The two connect sums $L_1\#L_2$ (1,\,2) and
1299: $L_2\#L_1$ (3,\,4) in 2-dimensions}}
1300: \end{figure}
1301: 
1302: In two dimensions the curves $\gamma$ whose Lagrangians $\gamma^2$ have
1303: constant phase are the straight lines, as can be seen from (\ref{pf}).
1304: Curves such as those marked 1 and 4 in Figure 2 flow towards
1305: a straight line (of some \emph{non-zero} angle) corresponding to a
1306: SLag, whereas curve 2 flows up until it `hangs' on a zero of $p$ (in
1307: finite time), where, on restarting the flow for 2 different curves,
1308: the separate flows form a kink and in the limit converge to
1309: destabilising SLags $L_1,\ L_2$ of different phases.
1310: These unions of SLags of different phases are still
1311: stationary for the volume functional (satisfying the second
1312: order variational equations, just not the first order SLag equations),
1313: and in fact are minimising in odd dimensions (the angle criterion
1314: \cite{N}, \cite{L} makes minimality of the singular union locally
1315: equivalent to the above destabilising phase condition; in even
1316: dimensions reversing the order of the Lagrangians reverses the
1317: inequality and the configuration is not minimal, just stationary).
1318: 
1319: Again we see how the phase or angle criterion comes to bear; curves
1320: 1 and 2 are in the same homology class, but the two different phase
1321: signs give very different results. As noted before in Figure 1,
1322: this is related to the necessity of the phase to vary non-monotonically
1323: to form the unstable connect sum; in Figure 3 we plot
1324: the phases of the two connect sums, and with dotted lines
1325: their limits under the heat flow (\ref{newdot}) (this is the correct
1326: modification of (\ref{thetadot}) in the non Ricci-flat case).
1327: 
1328: \begin{figure}[h]
1329: \center{
1330: \input{phase.pstex_t}
1331: \caption{The phase $\theta_{L=L_1\#L_2}$ of (1) $(\theta_{L_1}\equiv
1332: -\epsilon)$ and (2) $(\theta_{L_1}\equiv\epsilon)$ respectively.}}
1333: \end{figure}
1334: 
1335: Drawing $\gamma$ the other way round the zero of $p$ gives something in
1336: the same homology class (the Dehn twist around the root of $p$ does not
1337: alter the homology class of the $S^1$
1338: fibre over $\gamma$), which is the opposite connect sum discussed in
1339: \cite{Th} -- once the phase inequality becomes unstable for one
1340: connect sum it becomes stable for the other.
1341: 
1342: The two connect sums are related by monodromy, as in \cite{Th}.
1343: Take a one parameter family of polynomials $p$ which rotates
1344: two zeros $z_1,\,z_2$ of $p$ around each other. Then under the
1345: resulting monodromy a curve joining $z_1$ to a
1346: third zero $z_3$ is taken from being `above' $z_2$ to
1347: being below it, thus turning one connect sum into the other. \\
1348: 
1349: The three dimensional picture is similar. In Figure 4 we plot the
1350: lines corresponding to SLags of phase zero, and connect sums $L_1\#L_2$
1351: for $\phi(L_2)=0$ and $\phi(L_1)=\mp\epsilon$ (curves 1 and 2).
1352: Again we see
1353: the same behaviour with the phases behaving as in the graphs
1354: in Figure 3 and the flow getting hung on a zero of $p$ in
1355: the unstable case, splitting the Lagrangian.
1356: 
1357: Reversing the order of the connect sum in this case involves taking the
1358: $S^2$ fibre once around the zero of $p$; this effects a Dehn twist,
1359: reversing its orientation. Thus although curve 3 appears to give a
1360: Lagrangian in the same homology class, it is not; the phase of $L_1$
1361: once we have been round the root of $p$ has shifted by $\pi$ and we get
1362: the connect sum $L_2\#(L_1[-1])$ discussed in \cite{Th}. As is also
1363: discussed there, this can be seen to be unstable.
1364: 
1365: \begin{figure}[h]
1366: \center{
1367: \input{stab3.pstex_t}
1368: \caption{The two connect sums $L_1\#L_2$ and
1369: $L_2\#(L_1[-1])$ in 3-dimensions}}
1370: \end{figure}
1371: 
1372: Things are not quite as simple as we have portrayed them if the
1373: initial curve has very large phase variation. It is quite possible
1374: for a curve corresponding to a \emph{stable} Lagrangian, which is
1375: nonetheless very far from being a SLag, to pass close to a zero of
1376: $p$ without the large negative curvature away from the zero that the
1377: SLags exhibit. It can then flow into the zero, the Lagrangian being
1378: split into unions of Lagrangians of which it was a connect sum,
1379: despite their phases being such that they do not destabilise it.
1380: This limit is in the closure of the hamiltonian deformation orbit
1381: of the original Lagrangian, but does not contradict stability.
1382: 
1383: (A similar often-ignored subtlety occurs with stable bundles:
1384: when moduli of semistable sheaves are created by using GIT on part of
1385: a Quot scheme, the orbits of stable sheaves are \emph{not} closed in
1386: Quot -- the closures contain gradeds coming from any extension of
1387: sheaves forming the sheaf, stable or not. It is only in the
1388: locus of points of Quot representing
1389: semistable sheaves that the stable orbits are closed.)
1390: 
1391: The point about stable
1392: Lagrangians is that \emph{this can be avoided} by choosing a hamiltonian
1393: deformation of the
1394: Lagrangian to have sufficiently small variation in phase $\theta$
1395: (or sufficiently small volume) that it cannot be split into
1396: destabilising Lagrangians of different phase (or higher total volume);
1397: this we discuss now.
1398: 
1399: 
1400: \section{The conjecture} \label{conj}
1401: 
1402: It is now clear what our conjecture should be. Fix a (graded) Lagrangian
1403: submanifold $L$ of a Calabi-Yau $n$-fold $X$, and choose
1404: the phase of $\Omega$ such that the cohomological phase $\phi(L)=0$.
1405: Suppose first that the variation in
1406: $L$'s phase function $\theta$ is sufficiently small in the sense that
1407: \begin{equation} \label{close}
1408: [\phi(L_1),\phi(L_2)]\not\subseteq(\inf_L\theta,\sup_L\theta),
1409: \end{equation}
1410: for all graded connect sums $L_1\#L_2\approx L$ (by this we mean
1411: either the pointwise connect sums of Section \ref{index} or one of
1412: the relative connect sums discussed in \cite{Th}). This condition
1413: (\ref{close})
1414: is preserved by the flow, by the maximum principle and equation
1415: (\ref{thetadot}), and so prohibits $L$ splitting up as a
1416: connect sum under the flow, as in the limit of flowing to such a
1417: splitting (\ref{close}) would be violated.
1418: 
1419: We can also usefully consider volume instead of phase. If the
1420: Riemannian volume of our Lagrangian $L$ is less than the
1421: cohomological volume of any decomposition:
1422: \begin{equation} \label{vclose}
1423: \vol(L)\le\int_{L_1}e^{-i\phi(L_1)}\Omega+\int_{L_2}e^{-i\phi(L_2)}
1424: \Omega,
1425: \end{equation}
1426: for all $L_1,\,L_2$ such that $L\approx L_1\#L_2$, then we again expect
1427: convergence of mean curvature flow to a SLag representative for $L$.
1428: This is also preserved under the flow, by (\ref{voldot}), and so
1429: precludes the flow splitting $L$ into $L_1\cup L_2$, as this splitting
1430: would necessarily have higher volume. (As the referee pointed out,
1431: this is a global condition on $L$, whereas (\ref{close}) appears
1432: to be a pointwise one. The global nature of (\ref{close}) is that the
1433: points concerned are the extrema of $\theta$ over all of $L$.)
1434: 
1435: \begin{Conj}
1436: If $L$ satisfies either of the conditions (\ref{close}) or
1437: (\ref{vclose}) then mean curvature flow for $L$ exists for all time
1438: and converges to a special Lagrangian in its hamiltonian deformation
1439: class; the unique SLag conjectured in \emph{\cite{Th}}.
1440: \end{Conj}
1441: 
1442: It is of course a consequence of this and the conjecture in \cite{Th}
1443: that some hamiltonian deformation of $L$ satisfies (\ref{close}) if
1444: and only if it is stable. The SLag should also be unique in its
1445: hamiltonian deformation class as in Theorem \ref{!}.
1446: If $L$ is stable but not close enough to
1447: being SLag that (\ref{close}) fails, then mean curvature flow
1448: can become singular in finite time, (locally) splitting the Lagrangian
1449: in the reverse of a connect sum operation (i.e. with a vanishing cycle
1450: which is an $S^{n-1}$, or an $S^{n-r}$-bundle over an
1451: $(r-1)$-dimensional base in the relative connect sum case). We might
1452: then conjecture that the resulting pieces are smooth so we can
1453: begin the process again until we get a
1454: decomposition into different phase SLags. Typically, in the simplest
1455: case, we would get $L_1\cup L_2$ (with $\phi(L_1)<\phi(L_2)$ by
1456: stability) which is not a hamiltonian deformation of $L$ (though it
1457: is in the closure of such deformations).
1458: 
1459: If $L$ is unstable, we would again expect such finite time
1460: singularities and SLag splittings. But if $L$'s phase variation,
1461: or volume, is sufficiently small, we can hope for convergence
1462: to the Jordan-H\"older decomposition of Section \ref{JH}. That is,
1463: while the volume of $L$ must be larger than the cohomological volume
1464: of its Jordan-H\"older decomposition, if it is less than any other
1465: decomposition then it can only flow to the former. Again we expect
1466: the flow to become singular in finite time, the limit
1467: (locally) splitting $L$ into pieces for which we restart the flow.
1468: This splitting of the Lagrangian is a manifestation of the well
1469: known finite-time dumb-bell singularities in mean curvature flow.
1470: 
1471: 
1472: \subsection*{Proof for our example}
1473: 
1474: Under our flow (\ref{result!}) in the Shapere-Vafa example
1475: the proofs of the evolution equations for the phase
1476: function $\theta$ (\ref{thetadot}) and volume (\ref{voldot})
1477: show that the equations are modified (as the metric is not quite
1478: Ricci-flat) to
1479: \begin{eqnarray}
1480: {d\over dt}\theta\!&=&\!-\Delta\,\theta+{\langle d\theta,d|\Omega|
1481: \rangle\over|\Omega|}, \label{newdot} \\
1482: {d\over dt}(|\Omega|\vol_L)\!&=&\!-|d\theta|^2(|\Omega|\vol_L).
1483: \label{nvdot}
1484: \end{eqnarray}
1485: when $|\Omega|=|\Omega/\!\vol\!|$ is not $\equiv1$.
1486: Therefore the maximum principle still holds for $\theta$, and
1487: the condition (\ref{close}) is again preserved by the flow. Similarly
1488: if we measure volume with respect to $|\Omega|\vol_L$ then this
1489: is decreasing and (\ref{vclose}) is preserved by the flow. We can
1490: now prove the appropriate version of our conjecture in this example.
1491: From the proof it will also be clear that the original conjecture
1492: could be proved in this case in the $O(n)$-invariant Ricci flat
1493: metric if we knew it,
1494: we would just not be able to be as explicit about the flow equations.
1495: 
1496: \begin{Theorem} \label{faq}
1497: Suppose that $\gamma$ is a curve in $\C$, with endpoints at zeros of
1498: $p$, and otherwise missing the zeros of $p$, such that its
1499: pointwise phase $\theta$ (\ref{pf}) satisfies (\ref{close})
1500: for all Lagrangians $L_i=\gamma^n_i,\ i=1,2,$ fibred over curves
1501: $\gamma_i$ in the base, and also $S-I:=\sup_\gamma\theta-
1502: \inf_\gamma\theta<2\pi/3$. Then the flow (\ref{result!}) exists for all
1503: time and converges in $C^\infty$ to a smooth curve whose phase
1504: function (\ref{pf}) is constant.
1505: \end{Theorem}
1506: 
1507: We break the proof up into existence of the flow (best dealt with
1508: at the level of $\gamma^n\subset X^n$), controlling the angle
1509: variation (using $\gamma\subset\C$) to ensure no $180^o$ kinks
1510: appear in $\gamma$, and using this to show the
1511: flow exists for all time (for which we use the double cover
1512: $\gamma^1\subset X^1$
1513: and $\theta$ on $\gamma^n$). We follow \cite{An}, in parts heavily
1514: modified to take care of the endpoints of $\gamma$. Finally
1515: we will show that the flow converges to a SLag.
1516: 
1517: \begin{Lemma} \label{exist}
1518: The flow (\ref{result!}) exists while the curvature of $\gamma^1$
1519: is bounded.
1520: \end{Lemma}
1521: 
1522: \begin{Proof}
1523: Firstly, short term existence of the flow, given any initial curve
1524: $\gamma\subset\C$ missing the zeros of $p$ except at its endpoints
1525: and \emph{such that $\gamma^1$ is $H^{2+\alpha}$} for some
1526: $\alpha>0$ (i.e. $\gamma^1$ has H\"older continuous curvature),
1527: is in fact most easily proved at the level of the $H^{2+\alpha}$
1528: Lagrangian $\gamma^n$; see \cite{An} for the method in 1 dimension
1529: (which easily generalises to $n$ dimensions), and \cite{Ch} for
1530: a similar $n$-dimensional result. This is also done
1531: in (\cite{Sm} Proposition 1.6) using results of Hamilton \cite{H},
1532: for instance.
1533: 
1534: While the curvature of the curve $\gamma^1\subset X^1$ is bounded,
1535: so is the norm of the flow vector (the last term in (\ref{result!})
1536: is always bounded, and the second term can be bounded by the
1537: curvature at an intermediate point by Taylor's theorem). So at
1538: any finite time $T$ the flow converges to a limit curve $\gamma_T^1$
1539: pointwise. Parametrising the curves by their arclength on $X^1$,
1540: their first and second derivatives as maps to $X^1$ are therefore
1541: bounded, which
1542: by Arzel\`a-Ascoli implies that for a subsequence of $t$ we have
1543: convergence in $C^1$ to a $C^1$ curve with bounded (weak)
1544: curvature. By the uniqueness of the limit, then, $\gamma_T^1\subset
1545: X^1$ has bounded curvature.
1546: 
1547: Bounds on (the derivative of) the phase of $\gamma^1$ give
1548: corresponding bounds on (the derivative of) the phase of
1549: $\gamma^n$ (via (\ref{result!}) for $n$ and $n=1$). So the phase
1550: function $\theta$ of $\gamma^n$ is also $C^0$ convergent to the
1551: phase of $\gamma_T^n$, and satisfies the parabolic equation
1552: (\ref{newdot}). Putting this into local coordinates and
1553: differentiating with respect to arclength $s$, we get a
1554: uniformly parabolic equation with bounded coefficients and a
1555: bounded solution $\theta_s$ on $t\in[0,T]$. By (\cite{LSU}
1556: Section III Theorem 10.1), then, $\theta_s$ is
1557: in fact $\alpha$-H\"older continuous for some $\alpha>0$,
1558: and $\gamma^1_T$ is $H^{2+\alpha}$. By the existence of the flow
1559: for $H^{2+\alpha}$ initial conditions, then, the flow
1560: exists for some time $t>T$.
1561: \end{Proof}
1562: 
1563: \begin{Lemma} \label{cone}
1564: $\limsup_{|s-s'|\to0}|\theta(\gamma'(s,t))-\theta(\gamma'(s',t))|<\pi$
1565: for all time $t$ for which the flow exists, where $s$ is arclength
1566: along $\gamma(\,.\,,t)$, and $\theta(\gamma')$ is the argument
1567: of $\gamma'\in\C$.
1568: \end{Lemma}
1569: 
1570: \begin{Proof}
1571: Working outside a fixed neighbourhood of the zeros of $p$ at the
1572: endpoints of $\gamma$, this follows from the bounds
1573: on $\theta=\theta(\gamma')+(n/2-1)\theta(p(\gamma))$ (\ref{pf})
1574: coming from the $\sup\theta-\inf\theta<2\pi/3$ assumption
1575: (preserved under the flow by the maximum principle (\ref{newdot})),
1576: as the variation of $\theta(p(\gamma))$ can be made arbitrarily small
1577: with $|s-s'|$. Since $\gamma$ must stay at a bounded distance from
1578: other zeros of $p$ by the condition (\ref{close}) and the
1579: maximum principle for $\theta$ (\ref{newdot}), we are left with
1580: proving the lemma in an arbitrarily small neighbourhood in $\C$
1581: of the endpoints of $\gamma$.
1582: 
1583: Unfortunately, in this region, the bounds we want for
1584: $\theta(\gamma')$ do not follow from the
1585: bounds we have for $\theta$, and in fact come only from
1586: comparison with known solutions. Draw the SLags of phase $S,\,I$
1587: emanating from a zero of $p$, i.e. the curves in $\C$ solving
1588: $\theta(\gamma')+(n/2-1)\theta(p(\gamma))=S$ or $I$.
1589: (Since this is an ODE, there is no problem in finding solutions
1590: and extending them to either infinity or another zero of $p$; see
1591: \cite{SV}.) In the tangent space to the zero of $p$ this gives a
1592: cone of angle $(S-I)/n<2\pi/3n$ which $\gamma$ lies inside and
1593: cannot cross either at $t=0$ or any later time in the flow. So in
1594: a sufficiently small neighbourhood of an endpoint of $\gamma$,
1595: we may bound $\theta(\gamma)$ inside a cone of angle less than
1596: $2\pi/3n$, and also take $\theta(p(\gamma))$ to be within any given
1597: $\epsilon$ of $\theta(\gamma)+C$ (since the zero
1598: of $p$ is nondegenerate; here $C$ is the phase of $\dot p$ at the
1599: zero of $p$). Thus $\theta(p(\gamma))$ can be
1600: bounded inside a similar cone, so that (\ref{pf}) bounds the
1601: variation of $\theta(\gamma')$ by $2\pi/3+(n/2-1)2\pi/3n<\pi$.
1602: 
1603: The bounds on $\theta(\gamma')$ imply that the curve does not
1604: spiral round its endpoints but moves away from them with nonzero
1605: derivative inside the above cone until it is outside the small
1606: neighbourhood employed above. So the remaining case to consider
1607: is if the curve can pass arbitrarily close to one of its own
1608: endpoints at some bounded-below arclength from its endpoints,
1609: i.e. if the cone of SLags above starting from a zero of $p$
1610: passes either side of that same zero at some nonzero arclength.
1611: Then there would be a Slag fibred over a curve starting and
1612: ending at the same root of $p$, with our Lagrangian $\gamma^n$
1613: the connect sum of this SLag and some other $\gamma_1^n$.
1614: But $\gamma^n$ cannot flow arbitrarily close to such a connect sum
1615: as its phase variation would approach at least
1616: the difference between the phase of the SLag and $\phi(\gamma^n_1)$,
1617: contradicting (\ref{close}).
1618: \end{Proof}
1619: 
1620: By Lemma \ref{exist} the flow exists for all time unless, as we
1621: suppose now, the curvature of $\gamma^1$ becomes unbounded in finite
1622: time. Then to get a contradiction we start by scaling as in \cite{An}.
1623: Pick $s_i,\,t_i,\ i=1,2,\ldots$ such
1624: that $t_i$ tends to the blow up time and the curvature $\kappa_i$ of
1625: $\gamma^1(s_i,t_i)=y_i$ is maximal
1626: over the curvatures of $\gamma^1(s,t)$ for all $s$, and all
1627: $t\le t_i$ (here we parameterise by arclength $s$ on $X^1$, centred
1628: at a zero $z$ of $p$, i.e. $\gamma^1\res{s=0}=z$ lies over an
1629: endpoint of $\gamma$).
1630: 
1631: How we handle the blow up depends on
1632: whether it happens at the branch points of $\gamma^1$ (i.e.
1633: the endpoints of $\gamma$), by
1634: which we mean $|s_i|=O(|\kappa_i^{-1}|)$, or in the interior
1635: $|s_i|\gg|\kappa_i^{-1}|$, due to the different nature of
1636: (\ref{result!}) at the branch points.
1637: We first deal with the interior where the flow is a perturbation
1638: of mean curvature flow and so can be handled by \cite{An}:
1639: 
1640: \begin{Lemma}
1641: Supposing that the curvature blows up as above, then
1642: $|s_i\kappa_i|$ is bounded.
1643: \end{Lemma}
1644: 
1645: \begin{Proof}
1646: Firstly, if after passing to a subsequence of $i\in\mathbb N$,
1647: and centring $s$ about the other branch point of $\gamma^1$
1648: if necessary, the blow up
1649: occurs at a finite distance $|s_i|>\epsilon>0$ from either branch
1650: point of $\gamma^1$ 
1651: then in this interior the flow (\ref{result!}) is a finite
1652: perturbation of mean curvature
1653: flow satisfying the conditions of \cite{An}, so a $180^o$ kink
1654: must appear in $\gamma^1$, contradicting Lemma \ref{cone}. So
1655: we need only deal with the case of $s_i\to0$ (by passing to a
1656: subsequence to concentrate around one of the two branch
1657: points, if necessary) while $r_i:=|s_i\kappa_i|\to\infty$.
1658: 
1659: Then we rescale as in \cite{An};
1660: \begin{equation} \label{scale}
1661: s\mapsto \kappa_is, \quad g\mapsto \kappa_ig, \quad
1662: t\mapsto \kappa_i^2(t-t_i),
1663: \end{equation}
1664: where $g$ is the metric on $X^1$. This rescaled flow for
1665: $\gamma^1_i$ has the same form as (\ref{result!}),
1666: $$
1667: \dot\gamma_i^1=\MCV-(n-1)\mathbf n(\log|p|_i^{1/2})\,\mathbf
1668: n+{1\over2}\mathbf n(\log(|p|_i+|\dot p|_i^2/4))\,\mathbf n,
1669: $$
1670: but with $|p|$ and $|\dot p|$ replaced by their pullbacks
1671: $|p|_i,\ |\dot p|_i$ to
1672: the new Riemannian surface (here it is important that $\dot p$
1673: is still computed in the old coordinates, \emph{then} pulled
1674: back). Therefore their gradients are scaled by $\kappa_i^{-1}$.
1675: $\mathbf n$ denotes the unit normal to $\gamma^1$ in the
1676: new metric on $X^1$. The curvature gets scaled by
1677: $\kappa_i^{-1}$ and so has a maximum, over $t\le0$, of 1 at
1678: $y_i$ (at time $t=0$). We want to show that the two perturbation
1679: terms on the right hand side of the above flow tend to zero as
1680: $i\to\infty$.
1681: 
1682: In the rescaled variables, work in a geodesic disc in $X^1$ of
1683: radius $r_i/2$ ($r_i=|\kappa_is_i|\to\infty$ as $i\to
1684: \infty$) about $y_i$. As $s_i\to0$, for
1685: $i$ sufficiently large this is within an arbitrarily small
1686: neighbourhood of $z$ in the original metric,
1687: in which $\gamma'$ varies within an angle $<\pi$ cone
1688: as in the proof of Lemma \ref{cone}, i.e. $(\gamma^1)'$
1689: varies within an angle $<\pi/2$ cone on the double cover
1690: $X^1$. So arclength $s$ on $\gamma^1$ and radial distance $r$
1691: in $X^1$ are equivalent metrics on $\gamma^1$ in this disc;
1692: $r_s:=\partial r/\partial s$ and $s_r=\partial s/\partial r$
1693: are both bounded.
1694: 
1695: As $y_i\in\gamma_i^1$ is of arclength $s_i\ge r_i$ from $z$ (the
1696: zero of $p$) at $s=0$, we deduce that all points of our disc
1697: are of distance $\ge cr_i/2$ from the zero of $p$ (for some
1698: constant $c>0$ fixed for all $i\gg1$) in the new metric.
1699: Thus, for $i$ large enough, we have
1700: $$
1701: |p|_i^{1/2}\ge C\kappa_i^{-1}(cr_i/2),
1702: $$
1703: where $C$ is a constant just less than
1704: the norm of the derivative of $p^{1/2}$ at the zero $z$
1705: in the original metric on $X^1$ ($p^{1/2}$ pulls back
1706: to a well defined function on $X^1$ with a simple zero
1707: at $z$). We can therefore bound
1708: $$
1709: |\dot\gamma_i^1-\MCV|\le(n-1){\kappa_i^{-1}\sup|d(p^{1/2})|
1710: \over C\kappa_i^{-1}(cr_i/2)}+{1\over2}\kappa_i^{-1}
1711: \sup|d\log(|p|+|\dot p|^2/4)|,
1712: $$
1713: where both sups are taken over small neighbourhoods of
1714: $z$ \emph{in the original metric on} $X^1$. As $i\to\infty$,
1715: $\kappa_i,\,r_i\to\infty$, so the above bound tends to zero,
1716: while the radius of the disc we are working on
1717: $r_i/2\to\infty$. It follows that in the limit we get
1718: mean curvature flow of a curve inside an infinite flat
1719: disc $\R^2$; see (\cite{An} Section 9) for how to pass to the
1720: limit to conclude that for this
1721: blow up to occur a $180^o$ kink must appear in the curve
1722: $\gamma^1$ (by which we mean the limsup in Lemma \ref{cone}
1723: is $\ge\pi$). But this contradicts Lemma \ref{cone}.
1724: 
1725: (We do not repeat Angenent's argument here as we will
1726: give a slightly harder, $n$-dimensional, version of
1727: it around the endpoints of $\gamma$ in Lemma \ref{blowup}
1728: below. The point is just that in the rescaling we can
1729: get rid of the last two terms of our flow to reduce to the
1730: results of \cite{An}.)
1731: \end{Proof}
1732: 
1733: The remaining case
1734: we must dismiss is that of $\kappa_i$ blowing up at points
1735: $y_i=\gamma^1(s_i,t_i)$ with $|s_i|<A/|\kappa_i|$ for some
1736: fixed $A$. Here we must work harder than \cite{An}.
1737: 
1738: \begin{Lemma} \label{blowup}
1739: The curvature of $\gamma^1$ does not blow up in finite time.
1740: \end{Lemma}
1741: 
1742: \begin{Proof}
1743: By Lemma \ref{cone} we know that there is an $A$ such that
1744: $|s_i|<A/|\kappa_i|$.
1745: We rescale variables as in (\ref{scale}), and work on
1746: a length $\kappa_i^{1/2}\to\infty$ interval (in the new metric)
1747: on $\gamma^1$ centred ($s=0$) at the zero $z$ of $p$. This is
1748: contained inside the ball of radius $\kappa_i^{-1/2}\to0$
1749: about $z$ in $X^1$ in the original metric, so for $i$ sufficiently
1750: large we can assume that $p(t)-Ct$ is arbitrarily small
1751: \emph{in $C^2$ norm} (here $t\in\C$ is the base parameter,
1752: \emph{not} time, $C=\dot p(z)$, and the same is true of any $C^r$
1753: norm; $r=2$ is the case of interest for us). We start by obtaining
1754: bounds on the polar angle of the curve and its tangent vector.
1755: We shall confuse functions on $\C$ with their pullbacks to $X^1$
1756: (so writing things like $p(\gamma^1)$ etc.).
1757: 
1758: Taking $i$ sufficiently large that the metric on the
1759: radius $\kappa_i^{1/2}$ disc about $z$ in $X^1$ is sufficiently
1760: close to being flat, define geodesic polar coordinates on $X^1$,
1761: $r^1:=|\gamma^1|,\ \theta^1:=\theta(\gamma^1)$ (which is
1762: $\theta(\gamma)/2$ to within a constant). Then we can
1763: assume that $\theta^1_s$ is arbitrarily $C^1$ close to
1764: \begin{equation} \label{sin}
1765: {1\over r}\sin(\theta(\gamma^1_s)-\theta^1),
1766: \end{equation}
1767: which is the exact formula for a flat metric and polar
1768: coordinates. This bounds $|r\theta^1_s|$. Since by
1769: construction the curvature of $\gamma^1$ is not more than
1770: one, i.e. $|(\theta(\gamma^1_s))_s|\le1$, we can bound
1771: $|\theta(\gamma^1_s)|\le s$.
1772: 
1773: Note that (\ref{sin}) in flat space gives us the differential equation
1774: $$
1775: f_s=\kappa_i-{\sin f\over r}
1776: $$
1777: for $f=\theta(\gamma^1_s)-\theta^1$, with $f(0)=0$ and $|\kappa_i|\le1$.
1778: This implies that $|f(s)|\le|s|$ (consider a point where the graph
1779: of $f$ crosses that of $\pm s$, where $|f_s|\ge1$, for a contradiction),
1780: so for $i$ sufficiently large that our polar coordinates are sufficiently
1781: close to flat coordinates we can deduce a bound on $|\theta(\gamma^1_s)-
1782: \theta^1|/s$. Thus $\theta^1/s$ is also bounded, and $\theta^1/r$ by
1783: the uniform comparison bounds of $r$ and $s$ given by the cone argument
1784: in Lemma \ref{cone}.
1785: 
1786: Instead of considering the equation (\ref{result!}) for $\gamma^1$,
1787: we analyse the equation (\ref{newdot}) for $\theta$ on $\gamma^n$.
1788: After rescaling it becomes
1789: \begin{equation} \label{eqn}
1790: \dot\theta=\Delta\theta+{\langle d\theta,d(|\Omega|_i)\rangle\over
1791: |\Omega|_i},
1792: \end{equation}
1793: where $|\Omega|_i$ is the pullback of $|\Omega|:=|\Omega/\!\vol\!|$ to
1794: $\gamma^1$ with
1795: its new metric. Note also that pulling functions up from $\gamma$ and
1796: taking their exterior derivative $d$ on either $\gamma^1$ or on
1797: $\gamma^n$ gives the same result via the obvious inclusion
1798: $\gamma^1\subset\gamma^n$ commuting with the projections to $\gamma$.
1799: Again we want to control this evolution equation as $i\to\infty$.
1800: 
1801: $|d\theta|=|\theta_s|=|\partial_s[\theta(\gamma^1_s)/2+(n/2-1)
1802: \theta(p(\gamma^1))]|$, so this is bounded by the estimates above,
1803: for $i$ sufficiently large that $\theta(p(\gamma^1))$ is $C^1$
1804: close to $\theta^1/2+$\,const. in the disc in which we are working.
1805: So we can bound the last term in (\ref{eqn}) by a constant times
1806: $$
1807: \kappa_i^{-1}{\sup|d\Omega|\over\inf|\Omega|},
1808: $$
1809: where the sup and inf are taken in the original metric over a small
1810: neighbourhood of $(0,\ldots,0,z)\in X^n$. This tends to zero as
1811: $i\to\infty$.
1812: 
1813: Computing the Laplacian on the space $\gamma^n$, with a radial
1814: coordinate $s$ and rotational symmetry about the origin $s=0$,
1815: makes (\ref{eqn})
1816: \begin{equation} \label{tat}
1817: \theta_t=\theta_{ss}+(n-1){R^i_s\over R^i}\theta_s+O(\kappa_i^{-1}),
1818: \end{equation}
1819: where $R^i=R^i(s)$ is the radius of the sphere $S^{n-1}$ at $s$ in the
1820: new metric.
1821: 
1822: To compute $R^i$, we use our $\gamma^1\subset X^1$ arclength coordinate
1823: $s$, the radial coordinate $r$ on $X^1$, and a radial coordinate
1824: $\rho$ on $\C$. For $i$ sufficiently large, for $s\le\kappa_i^{1/2}$
1825: in the new metric, we can approximate $p$ linearly about $z$ and
1826: so assume that $R^i$ is as close as we like
1827: to $\kappa_i\sqrt{|\dot p(z)|\rho}$ in $C^2$. Therefore $R^i_r$ is
1828: approximated by
1829: \begin{equation} \label{grig}
1830: R^i_r=\frac{R^i_\rho}{r_\rho}=\frac{R^i_\rho}{\sqrt{\kappa_i^2+(R^i_\rho)^2}}
1831: =\frac{1}{\sqrt{1+\frac{\kappa_i^2}{(R^i_\rho)^2}}}\approx
1832: \frac{1}{\sqrt{1+\frac{4\rho}{|\dot{p}(z)|}}}\ ,
1833: \end{equation}
1834: which is bounded and tends to $1$ in the interval
1835: $s\in[0,\kappa_i^{1/2})$. Similarly $R^i_{rr}$ can be taken to be
1836: arbitrarily small on the same interval, for $i$ sufficiently large.
1837: 
1838: Since $r_s$ is bounded, this gives bounds on $R^i_s$, implying that,
1839: on passing to a subsequence if necessary, the functions $R^i(s)$
1840: are convergent as $i\to\infty$ by the Arzel\`a-Ascoli theorem.
1841: 
1842: Note also that for all $i$, $R^i_s(0)=1$. But to preserve this
1843: in the limit, we must similarly bound $R^i_{ss}$. Differentiating
1844: $r^2_s+r^2(\theta^1_s)^2=1$ and $R^i_s=R^i_rr_s$ gives
1845: \begin{equation} \label{final}
1846: R^i_{ss}=R^i_{rr}r_s^2-R^i_r(r(\theta^1_s)^2+
1847: r^2\theta^1_s\theta^1_{ss}/r_s).
1848: \end{equation}
1849: We have bounded $R^i_{rr},\ r_s,\ r_s^{-1},\ r\theta^1_s$ and
1850: $\theta^1_s$; all this leaves is the last term in (\ref{final}).
1851: 
1852: We have approximated $\theta^1_s$ in $C^1$ by
1853: ${1\over r}\sin(\theta(\gamma^1_s)-\theta^1)$; differentiating
1854: approximates $r^2\theta^1_s\theta^1_{ss}/r_s$ as closely as
1855: we like (as $i\to\infty$) to
1856: $$
1857: {r\theta^1_s\over r_s}(\kappa_i-\theta^1_s)\cos(\theta(\gamma^1_s)
1858: -\theta^1)-\theta^1_s\sin(\theta(\gamma^1_s)-\theta^1),
1859: $$
1860: which we have bounded already.
1861: In conclusion, after passing to a subsequence if necessary, $R^i$ is
1862: $C^1$ convergent to some $R$ with $R_s(0)=1$, and the phase function
1863: $\theta^\infty$ of the limit curve $\gamma^1_\infty$
1864: (which exists by Arzel\`a-Ascoli
1865: since $|\gamma^1_s|=1$ and $|\gamma^1_t|$ is bounded by the bound on
1866: its curvature $\kappa$) satisfies the limit of (\ref{tat}):
1867: \begin{equation} \label{lap}
1868: \theta^\infty_t=\theta^\infty_{ss}+(n-1){R_s\over R}\theta^\infty_s.
1869: \end{equation}
1870: 
1871: But this is just the heat equation for $\theta$ on $\R^n$ with radial
1872: coordinate $s$ and the $O(n)$-invariant metric in which the radius of
1873: the $S^{n-1}$ fibre over $s$ is $R(s)$. By construction of the time
1874: rescaling (\ref{scale}) it exists for all time $t\le0$, and the
1875: solution $\theta^\infty$ is bounded. Also the metric is uniformly
1876: elliptic compared to the flat metric, as $R_r=1$ (\ref{grig}) and
1877: $r_s$ is bounded above and below away from zero. Therefore, by
1878: Moser's Harnack inequality \cite{Mo}, $\theta^\infty$ is in fact
1879: constant.
1880: 
1881: But for all $i$, max\,$\theta_s=1$ by construction, and passing to
1882: a subsequence if necessary the point
1883: where the maximum is obtained is convergent. To show then that
1884: max\,$\theta^\infty_s=1$, to get our contradiction, we
1885: need only know that $\theta_s$ is, say, \emph{uniformly (in $i$)}
1886: H\"older continuous. This is again a consequence of (\cite{LSU}
1887: Section III Theorem 10.1) as follows.
1888: By the boundedness of $\theta_s$ and the parabolic
1889: equation it satisfies (different for each $i$), $\theta_s$ is
1890: in fact $\alpha$-H\"older continuous for some $\alpha>0$, with
1891: $H^\alpha$ norm bounded by the bounds on the coefficients of
1892: the parabolic equations. But these are bounded uniformly in $i$,
1893: as a glance at (\ref{eqn}) confirms: the correction term tends to
1894: zero, and the Laplacian term (and its derivative with respect to
1895: $s$) is controlled by $C^2$ bounds on the metric which we provided
1896: above by bounding $R^i(s),\,R^i_s$ and $R^i_{ss}$.
1897: \end{Proof}
1898: 
1899: Finally we show that this infinite time flow converges using
1900: standard techniques (see for instance \cite{C} for a harder result).
1901: Notice that the same scaling proof (\ref{blowup}) that the
1902: curvature of $\gamma^1$ does not blow up in finite time shows the
1903: same for our now infinite time flow. So, using the $O(n)$ symmetry,
1904: the curvature of the metric on $\gamma^n$ stays uniformly bounded,
1905: and we have a $C^1$ bound on $\theta$. Therefore in the equation
1906: (\ref{newdot}) for $\theta$ on $\gamma^n$, which we rewrite as
1907: \begin{equation} \label{ntd}
1908: \dot\theta=-\Delta^\Omega\theta:=|\Omega|^{-1}*d(|\Omega|*d\theta),
1909: \end{equation}
1910: the coefficients have at least uniform $C^1$ bounds; $\theta$ then
1911: acquires a uniform $C^3$ bound by parabolic Schauder estimates (see
1912: \cite{LSU} III Theorem 12.1, for instance). Again by $O(n)$ symmetry
1913: we now get uniform $C^2$ bounds on $\gamma^n$'s curvature. And so it
1914: goes on, giving $C^\infty$ bounds and allowing us to extract a
1915: subsequence of times for which the flow converges in $C^\infty$ to
1916: a Slag.
1917: 
1918: To see that the flow converges without having to pass to
1919: a subsequence we need only show convergence of $\theta$ in $L^2$;
1920: this way no other subsequence of the flow can converge to
1921: a different limit in $C^\infty$. In fact we use an $L^2$-norm
1922: weighted by $|\Omega|$, and compute using (\ref{ntd})
1923: and (\ref{nvdot}):
1924: $$
1925: {d\over dt}\int_{\gamma^n}(\theta-\bar\theta)^2|\Omega|\vol=
1926: \int_{\gamma^n}\left\{2(\theta-\bar\theta)(-\Delta^\Omega\theta-
1927: {d\over dt}\bar\theta)-(\theta-\bar\theta)^2|d\theta|^2\right\}
1928: |\Omega|\vol,
1929: $$
1930: where $\bar\theta=\int\theta|\Omega|\vol/\int|\Omega|\vol$ is
1931: constant on $\gamma^n$ (but not in time). This is then bounded
1932: above by
1933: $$
1934: -2\int(\theta-\bar\theta)\Delta^\Omega(\theta-\bar\theta)
1935: |\Omega|\vol.
1936: $$
1937: It is easily checked that $\Delta^\Omega=d_{\,}^{*^\Omega}
1938: d$, where $d_{\,}^{*^\Omega}$ is the adjoint of $d$ with
1939: respect to the $L^2$-metric $\int_{\gamma^n}\langle\ \cdot\ ,
1940: \ \cdot\ \rangle|\Omega|\vol$ we are using. So its kernel is just
1941: the constants, and by the uniform $C^\infty$ bounds on the metric
1942: of $\gamma^n$ and $|\Omega|$ we can get a uniform lower bound
1943: $\lambda>0$ for its first nonzero eigenvalue. This then gives
1944: a bound $\int(\mu\Delta\mu)|\Omega|\vol\ge\lambda\int\mu^2|\Omega|\vol$
1945: for functions $\mu$ of integral zero. Setting $\mu=\theta-\bar\theta$
1946: gives
1947: $$
1948: {d\over dt}\int(\theta-\bar\theta)^2|\Omega|\vol\le-2\lambda
1949: \int(\theta-\bar\theta)^2|\Omega|\vol,
1950: $$
1951: which therefore tends to zero as required. Theorem \ref{faq} is proved. \\
1952: 
1953: We end by noting that one can get cone-type bounds similar to
1954: those of Lemma \ref{cone} on the tangent
1955: direction $\theta(\gamma^1_s)$ even as a curve $\gamma$
1956: representing an \emph{unstable} Lagrangian $\gamma^n$ approaches
1957: and breaks across a zero of $p$. Suppose that the initial phase
1958: variation of $\gamma^n$ is, without loss of generality,
1959: in some $(-\delta,\delta)$. Draw the cone with boundary the
1960: SLags emanating from the zero of $p$ with phase $-\delta,\ \pi+
1961: \delta$ (the straight lines emanating from the zeros of $p$ drawn
1962: in Figures 2 and 4 display the $\delta=0$ cone for dimensions 2
1963: and 3 respectively). Then in a
1964: sufficiently small neighbourhood of the zero, the variation in
1965: $\theta(p)$ can be taken to be less than $2\pi/n+2\delta+\epsilon$
1966: for any $\epsilon>0$. Since we are free to make $\delta$ slightly
1967: smaller without violating the initial bounds on phase, we can
1968: ensure that there is a neighbourhood of the zero of $p$, and a cone
1969: with vertex at $p$ whose walls $\gamma$ cannot cross, such that for
1970: the part of $\gamma$ lying in this neighbourhood, the variation
1971: $$
1972: \sup\theta(p(\gamma))-\inf\theta(p(\gamma))<2\pi/n+2\delta. 
1973: $$
1974: Comparing with (\ref{pf}) gives
1975: $$
1976: \sup\theta(\gamma')-\inf\theta(\gamma')<(n/2-1)(2\pi/n+2\delta)+
1977: 2\delta=(1-2/n)\pi+n\delta
1978: $$
1979: so that again no $180^o$ kinks can occur while this is less than
1980: $\pi$, i.e. for $\delta\le2\pi/n^2$. So again the analysis should
1981: be tractable in this case (more general spiralling around a zero
1982: would make matters worse). However, we have not carried out the
1983: analysis necessary to show that at the moment $\gamma$ reaches
1984: the zero of $p$ it is sufficiently smooth that the two resulting
1985: curves it splits into give $C^2$ Lagrangians (whose flow we could
1986: restart). \\
1987: 
1988: Of course by just studying the simple examples above we cannot
1989: hope to know how bad the singularities are that arise in finite
1990: time in the general case.
1991: Also, as mentioned in \cite{Th}, we should perhaps restrict to
1992: those Lagrangians whose Floer cohomology is well defined \cite{FO3}.
1993: This includes all homology spheres, however.
1994: 
1995: We should also point out the obvious fact that most of the
1996: evidence for our conjecture, other than perhaps the mirror
1997: symmetry and study of Joyce's examples in \cite{Th}, has been
1998: essentially one-dimensional (either
1999: for $T^2$ in \cite{Th}, or by symmetry reduction in the examples
2000: above). This is unrepresentative, essentially because the angles at
2001: which Lagrangians intersect (the $\alpha_i$s of (\ref{l1}))
2002: are all the same in this situation, and so are determined by the
2003: phase (their sum). So interesting phenomena, where degrees in
2004: Floer cohomology change (e.g. a Hom becomes an Ext$^i$ on the mirror
2005: while the phase remains fixed) are largely lost due to them being
2006: controlled entirely by the phase.
2007: 
2008: 
2009: \begin{thebibliography}{FO3}
2010: 
2011: \bibitem[An]{An} S. Angenent,
2012: \textit{Parabolic equations for curves on surfaces. I. Curves
2013: with $p$-integrable curvature.} Ann. of Math. (2) \textbf{132}
2014: (1990), 451--483.
2015: 
2016: \bibitem[C]{C} H. D. Cao,
2017: \textit{Deformation of K\"ahler metrics to K\"ahler-Einstein metrics on
2018: compact K\"ahler manifolds.} Invent. Math. \textbf{81} (1985),
2019: 359--372.
2020: 
2021: \bibitem[Ch]{Ch} K.-C. Chang,
2022: \textit{Heat flow and boundary value problem for harmonic maps.}
2023: Ann. Inst. H. Poincar\'e Anal. Non Lin\'eaire \textbf{6} (1989),
2024: 363--395.
2025: 
2026: \bibitem[FO3]{FO3} K. Fukaya, Y.-G. Oh, H. Ohta, and K. Ono,
2027: \textit{Lagrangian intersection Floer theory -- anomoly and obstruction},
2028: preprint.
2029: 
2030: \bibitem[GH]{GH}
2031: P. Griffiths and J. Harris,
2032: \textit{Principles of algebraic geometry.} Wiley, New York. 1978.
2033: 
2034: \bibitem[Gr]{Gr} M. Grayson,
2035: \textit{Shortening embedded curves},
2036: Ann. of Math. \textbf{129} (1989), 71--111.
2037: 
2038: \bibitem[H]{H}
2039: R. S. Hamilton,
2040: \textit{Three manifolds with positive Ricci curvature},
2041: Jour. Diff. Geom. \textbf{23} (1986), 69--96.
2042: 
2043: \bibitem[HaL]{HaL} F. R. Harvey and H. B. Lawson,
2044: \textit{Calibrated Geometries}, Acta Math. \textbf{148} (1982), 47--157.
2045: 
2046: \bibitem[HuL]{HuL}
2047: D. Huybrechts and M. Lehn,
2048: \textit{Geometry of moduli spaces of shaves.}
2049: Aspects in Mathematics Vol. E31, Vieweg. 2000.
2050: 
2051: \bibitem[KS]{KS}
2052: M. Khovanov and P. Seidel,
2053: \textit{Quivers, Floer cohomology, and braid group actions},
2054: J. Amer. Math. Soc. \textbf{15} (2002), 203--271.
2055: 
2056: \bibitem[L]{L} G. Lawlor,
2057: \textit{The angle criterion}, Inv. Math. \textbf{95} (1989), 437--446.
2058: 
2059: \bibitem[LSU]{LSU} O. A. Ladyzhenskaya, V. A. Solonikov and
2060: N. N. Ural'ceva,
2061: \textit{Linear and quasilinear equations of parabolic type},
2062: Transl. Math. Monographs \textbf{23}, AMS. 1968.
2063: 
2064: \bibitem[Mi]{Mi} J. Milnor,
2065: \textit{Lectures on the h-cobordism theorem}. Princeton University
2066: Press. 1965.
2067: 
2068: \bibitem[Mo]{Mo} J. Moser,
2069: \textit{A Harnack inequality for parabolic differential equations},
2070: Comm. Pure Appl. Math. \textbf{17} (1964) 101--134.
2071: 
2072: \bibitem[N]{N} D. Nance,
2073: \textit{Sufficient conditions for a pair of planes to be area minizing},
2074: Math. Ann. \textbf{279} (1989) 161--164.
2075: 
2076: \bibitem[Oh1]{Oh1} Y.-G. Oh,
2077: \textit{Mean curvature vector and symplectic topology of Lagrangian
2078: submanifolds in Einstein-K{\"a}hler manifolds}, Math. Z. \textbf{216}
2079: (1994), 471--482. 
2080: 
2081: \bibitem[Oh2]{Oh2} Y.-G. Oh,
2082: \textit{Floer cohomology, spectral sequence and the Maslov class of
2083: Lagrangian embeddings}, Internat. Math. Res. Notices \textbf{7} (1996),
2084: 305--346.
2085: 
2086: \bibitem[P]{P} M. Po\'zniak,
2087: \textit{Floer homology, Novikov rings and clean intersections}, Ph.\,D.
2088: thesis, University of Warwick, 1994.
2089: 
2090: \bibitem[SW]{SW} R. Schoen and J. Wolfson,
2091: \textit{Minimizing area among Lagrangian surfaces: the mapping problem.}
2092: J. Differential Geom. \textbf{58} (2001), 1--86. 
2093: 
2094: \bibitem[S1]{S1} P. Seidel,
2095: \textit{Lagrangian two-spheres can be symplectically knotted}, Jour.
2096: Diff. Geom. \textbf{52} (1999), 145--171.
2097: 
2098: \bibitem[S2]{S2} P. Seidel,
2099: \textit{Graded Lagrangian submanifolds}, Bull. Soc. Math. France.
2100: \textbf{128} (2000), 103--146.
2101: 
2102: \bibitem[S3]{S3} P. Seidel,
2103: \textit{Vanishing cycles and mutation}, European Congress of Mathematics,
2104: Vol. II (Barcelona, 2000), 65--85, Progr. Math., 202, Birkh{\"a}user, Basel, 2001.
2105: 
2106: \bibitem[ST]{ST} P. Seidel and R. P. Thomas,
2107: \textit{Braid group actions on derived categories of sheaves},
2108: Duke Math. J. \textbf{108} (2001), 37--108.
2109: 
2110: \bibitem[SV]{SV} A. Shapere and C. Vafa,
2111: \textit{BPS structure of Argyres-Douglas superconformal theories},
2112: preprint hep-th/9910182.
2113: 
2114: \bibitem[Sm]{Sm} K. Smoczyk,
2115: \textit{A canonical way to deform a Lagrangian submanifold},
2116: preprint dg-ga/9605005.
2117: 
2118: \bibitem[SYZ]{SYZ} A. Strominger, S.-T. Yau and E. Zaslow,
2119: \textit{Mirror Symmetry is T-Duality}, Nucl. Phys. \textbf{B479}
2120: (1996), 243--259.
2121: 
2122: \bibitem[Th]{Th} R. P. Thomas, 
2123: \textit{Moment maps, monodromy and mirror manifolds}, Symplectic geometry
2124: and mirror symmetry (Seoul, 2000), 467--498, World Sci. Publishing, 2001.
2125: 
2126: \bibitem[Wa]{Wa} M. T. Wang,
2127: \textit{Deforming area-preserving diffeomorphisms of surfaces by
2128: mean curvature flow}, Math. Res. Lett. \textbf{8} (2001), 651--661.
2129: 
2130: \end{thebibliography}
2131: 
2132: \small \noindent {\tt richard.thomas@ic.ac.uk} \newline
2133: \noindent Department of Mathematics, Imperial College, Huxley
2134: Building, 180 Queen's Gate, London, SW7 2BZ. UK. \newline\newline
2135: \noindent {\tt yau@math.harvard.edu} \newline
2136: \noindent Department of Mathematics, Harvard University, One
2137: Oxford Street, Cambridge MA 02138. USA.
2138: 
2139: \end{document}
2140: 
2141: 
2142: % Local Variables: 
2143: % TeX-parse-self: t
2144: % TeX-auto-save: t
2145: % mode: latex
2146: % TeX-master: "thesis"
2147: % End:
2148: