hep-th0503173/SG.tex
1: %
2: % A. Bilal, S. Metzger
3: %
4: \documentclass[12pt,a4paper]{article}
5: \usepackage{amssymb}
6: \usepackage{epsfig}
7: \usepackage{color}
8: 
9: \setlength{\textwidth}{165mm} \setlength{\textheight}{230mm}
10: \oddsidemargin 0.mm \evensidemargin 0.mm \topmargin -1.9cm
11: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
12: \newcommand{\nsect}{\setcounter{equation}{0}
13: \def\theequation{\thesection.\arabic{equation}}\section}
14: \newcommand{\nappend}{\setcounter{equation}{0}
15: \def\theequation{\rm{A}.\arabic{equation}}\section*}
16: \newcommand{\appendixA}{\setcounter{equation}{0}
17: \def\theequation{\rm{A}.\arabic{equation}}\section*}
18: \newcommand{\appendixB}{\setcounter{equation}{0}
19: \def\theequation{\rm{B}.\arabic{equation}}\section*}
20: \newcommand{\appendixC}{\setcounter{equation}{0}
21: \def\theequation{\rm{C}.\arabic{equation}}\section*}
22: \newcommand{\appendixD}{\setcounter{equation}{0}
23: \def\theequation{\rm{D}.\arabic{equation}}\section*}
24: \newcommand{\appendixE}{\setcounter{equation}{0}
25: \def\theequation{\rm{E}.\arabic{equation}}\section*}
26: \newcommand{\appendixF}{\setcounter{equation}{0}
27: \def\theequation{\rm{F}.\arabic{equation}}\section*}
28: \newcommand{\appendixG}{\setcounter{equation}{0}
29: \def\theequation{\rm{G}.\arabic{equation}}\section*}
30: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
31: \def\bra#1{\left\langle #1\right|}
32: \def\half{\frac{1}{2}}
33: \def\Im{\mathop{\rm Im}}
34: \def\ket#1{\left| #1\right\rangle}
35: \def\sket#1{| #1 >}
36: \def\lie{\hbox{\it \$}}                          % Lie derivative symbol
37: \def\lineint{\oint \frac{d z}{2 \pi i}}
38: \def\modsq#1{| #1 |^2}
39: \def\NP#1#2#3{Nucl. Phys. {\bf #1} (19#2) #3}
40: \def\ov{\overline}
41: \def\partder#1#2{{\partial #1\over\partial #2}}
42: \def\PL#1#2#3{Phys. Lett. {\bf #1} (19#2) #3}
43: \def\PR#1#2#3{Phys. Rev. \underline{#1} (19#2) #3}
44: \def\PRL#1#2#3{Phys. Rev. Lett. \underline{#1} (19#2) #3}
45: \def\Re{\mathop{\rm Re}}
46: \def\secder#1#2#3{{\partial^2 #1\over\partial #2 \partial #3}}
47: \def\s2w{\sin^2 \theta_W}
48: \def\Tr{\mathop{\rm Tr}}
49: \def\und{\underline}
50: \def\VEV#1{\left\langle #1\right\rangle} \let\vev\VEV
51: \def\mbf#1{\hbox{\boldmath $#1$}}
52: %\relax
53: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
54: %   Other macros
55: %
56: \def\crbig{\\\noalign{\vspace {3mm}}}
57: \def\bigint{{\displaystyle\int}}
58: \def\und{\underline }
59: \def\be{\begin{equation}}
60: \def\ee{\end{equation}}
61: \def\ba{\begin{eqnarray}}
62: \def\ea{\end{eqnarray}}
63: \def\db{\bar{\partial}}
64: \def\del{\partial}
65: \def\w{\wedge}
66: \def\d{{\rm d}}
67: \def\k{\kappa}
68: \def\r{\rho}
69: \def\a{\alpha}
70: \def\A{\Alpha}
71: \def\b{\beta}
72: \def\B{\Beta}
73: \def\g{\gamma}
74: \def\G{\Gamma}
75: \def\dd{\delta}
76: \def\D{\Delta}
77: \def\e{\epsilon}
78: \def\E{\Epsilon}
79: \def\p{\pi}
80: \def\P{\Pi}
81: \def\c{\chi}
82: \def\C{\Chi}
83: \def\th{\theta}
84: \def\Th{\Theta}
85: \def\m{\mu}
86: \def\n{\nu}
87: \def\om{\omega}
88: \def\Om{\Omega}
89: \def\l{\lambda}
90: \def\L{\Lambda}
91: \def\s{\sigma}
92: \def\S{\Sigma}
93: \def\vphi{\varphi}
94: \def\cN{{\cal N}}
95: \def\cM{{\cal M}}
96: \def\V{\tilde V}
97: \def\cV{{\cal V}}
98: \def\ctV{\tilde{\cal V}}
99: \def\cL{{\cal L}}
100: \def\cR{{\cal R}}
101: \def\cA{{\cal A}}
102: \def\cg{{\cal{G} }}
103: \def\cD{{\cal{D} } }
104: \def\tm{\tilde{m}}
105: \def\Ct{{\widetilde C}}
106: \def\Gt{{\widetilde G}}
107: \def\At{{\widetilde A}}
108: \def\Rt{{\widetilde R}}
109: \def\Rh{{\hat R}}
110: \def\eh{{\hat\eta}}
111: \def\et{{\widetilde\epsilon}}
112: \def\ot{{\widetilde\omega}}
113: \def\ett{{\widetilde\eta}}
114: \def\Ah{\hat A}
115: \def\Ft{{\widetilde F}}
116: \def\St{{\widetilde S}}
117: \def\Sh{\hat S}
118: \def\t{\theta}
119: \def\rt{{\tilde r}}
120: \def\bs{\bigskip}
121: \def\no{\noindent}
122: \def\hb{\hfill\break}
123: \def\qq{\qquad}
124: \def\bl{\bigl}
125: \def\br{\bigr}
126: \def\delb{\overline\del}
127: \def\IR{\relax{\rm I\kern-.18em R}}
128: 
129: 
130: \def\np {  Nucl. Phys. }
131: \def \pl { Phys. Lett. }
132: \def \mpl { Mod. Phys. Lett. }
133: \def \prl { Phys. Rev. Lett. }
134: \def \pr  { Phys. Rev. }
135: \def \ap  { Ann. Phys. }
136: \def \cmp { Commun. Math. Phys. }
137: \def \ijmp { Int. J. Mod. Phys. }
138: 
139: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
140: 
141: \def \ha {{1\over 2}}
142: \def \third {{1\over 3}}
143: \def \ov {\over}
144: 
145: %\def\inv{^{\raise.15ex\hbox{${\scriptscriptstyle -}$}\kern-.05em 1}}
146: 
147: \def\tr{\,{\rm tr}\,}
148: \def\str{\,{\rm str}\,}
149: \def\ap{\alpha'}
150: \def\at{{\tilde \alpha}'}
151: \def\a{\alpha}
152: \def\b{\beta}
153: \def\g{\gamma}
154: \def\G{\Gamma}
155: \def\d{{\rm d}}
156: \def\dh{\hat\d}
157: \def\e{\epsilon}
158: \def\p{\psi}
159: \def\c{\chi}
160: \def\cb{{\overline\chi}}
161: \def\th{\theta}
162: \def\m{\mu}
163: \def\n{\nu}
164: \def\r{\rho}
165: \def\l{\lambda}
166: \def\lb{{\overline\lambda}}
167: \def\k{\kappa}
168: \def\s{\sigma}
169: \def\o{\omega}
170: \def\f{\phi}
171: \def\vf{\varphi}
172: \def\O{\Omega}
173: \def\ub{{\overline u}}
174: \def\ks{{k \kern-.5em /}}
175: \def\es{{\e \kern-.4em /}}
176: \def\ds{{\partial \kern-.5em /}}
177: \def\Ds{{D \kern-.6em /}}
178: \def\gt{{\tilde g}}
179: \def\ct{{\tilde c}}
180: \def\gh{{\hat g}}
181: \def\ch{{\hat c}}
182: \def\hh{{\hat h}}
183: \def\R{{\cal R}}
184: \def\ph{\hat\psi}
185: 
186: \def\bs{\bigskip}
187: \def\no{\noindent}
188: \def\hb{\hfill\break}
189: \def\qq{\qquad}
190: \def\bl{\bigl}
191: \def\br{\bigr}
192: \def \ha {{1\over 2}}
193: \def \ov {\over}
194: \def\hN{\hat{N}}
195: \def\hg{\hat{g}}
196: \def\zb{\bar{z}}
197: \def\mD{\mathcal{D}}
198: \def\zt{\tilde{z}}
199: \def\ztb{\bar{\tilde{z}}}
200: 
201: \def\wt{\widetilde}
202: 
203: \renewcommand{\theequation}{\thesection.\arabic{equation}}
204: %
205: %
206: %
207: \begin{document}
208: %\draft
209: %\preprint
210: %
211: %
212: %
213: %
214: \begin{titlepage}
215: \begin{flushright}
216: LPTENS-05/11\\
217: LMU-ASC 23/05\\
218: hep-th/0503173
219: \end{flushright}
220: \vskip 2.5cm
221: 
222: \begin{center}{\Large\bf Special geometry of local Calabi-Yau manifolds} \\
223: \vspace{2mm} {\Large\bf and superpotentials}
224: \vspace{2mm} {\Large \bf from holomorphic matrix models}
225: \vskip 1.5cm {\bf Adel Bilal$^{1}$ and Steffen
226: Metzger$^{1,2}$}
227: 
228: \vskip.3cm $^1$ Laboratoire de Physique Th\'eorique,
229: \'Ecole Normale Sup\'erieure - CNRS\\
230: 24 rue Lhomond, 75231 Paris Cedex 05, France
231: 
232: \vskip.3cm $^2$ Arnold-Sommerfeld-Center for Theoretical
233: Physics,\\
234: Department f\"ur Physik, Ludwig-Maximilians-Universit\"at\\
235: Theresienstr. 37, 80333 Munich, Germany\\
236: 
237: \vskip.3cm {\small e-mail: {\tt adel.bilal@lpt.ens.fr,
238: metzger@physique.ens.fr}}
239: \end{center}
240: \vskip .5cm
241: 
242: \begin{center}
243: {\bf Abstract}
244: \end{center}
245: \begin{quote}
246: We analyse the (rigid) special geometry of a class of local
247: Calabi-Yau manifolds given by hypersurfaces in $\mathbb{C}^4$ as
248: $W'(x)^2+f_0(x)+v^2+w^2+z^2=0$, that arise in the study of the
249: large $N$ duals of four-dimensional $\mathcal{N}=1$ supersymmetric
250: $SU(N)$ Yang-Mills theories with adjoint field $\Phi$ and
251: superpotential $W(\Phi)$. The special geometry relations are
252: deduced from the planar limit of the corresponding holomorphic
253: matrix model. The set of cycles is split into a bulk sector, for
254: which we obtain the standard rigid special geometry relations, and
255: a set of relative cycles, that come from the non-compactness of
256: the manifold, for which we find cut-off dependent corrections to
257: the usual special geometry relations. The (cut-off independent)
258: prepotential is identified with the (analytically continued) free
259: energy of the holomorphic matrix model in the planar limit. On the
260: way, we clarify various subtleties pertaining to the saddle point
261: approximation of the holomorphic matrix model. A formula for the
262: superpotential of IIB string theory with background fluxes on
263: these local Calabi-Yau manifolds is proposed that is based on
264: pairings similar to the ones of relative cohomology.
265: \end{quote}
266: 
267: \end{titlepage}
268: %
269: %
270: %
271: \setcounter{footnote}{0}
272: \setlength{\baselineskip}{.7cm}
273: \newpage
274: %
275: % BODY
276: %
277: \newtheorem{proposition}{Proposition}[section]
278: \newtheorem{theorem}[proposition]{Theorem}
279: \newtheorem{definition}[proposition]{Definition}
280: \newtheorem{conjecture}[proposition]{Conjecture}
281: 
282: 
283: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
284: \section{Introduction\label{Intro}}
285: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
286: Compactifications of string theory on Calabi-Yau manifolds have
287: been studied for almost two decades \cite{CHSW85}. One
288: particularly appealing property of Calabi-Yau compactifications is
289: that the special geometry structure of the effective supergravity
290: theory \cite{dWvP} can be understood from the fact that the
291: (K\"ahler and complex structure) moduli spaces of a Calabi-Yau
292: manifold are special K\"ahler manifolds. What is more, the
293: prepotential that corresponds to the complex structure
294: deformations can be expressed in terms of period integrals on the
295: Calabi-Yau space \cite{CO90}. For these periods one can deduce the
296: Picard-Fuchs differential equations and so get interesting
297: physical quantities from the solutions of these equations. In
298: general the computation of the K\"ahler part is more complicated
299: as in this case the period integrals are corrected by world-sheet
300: instantons. However, mirror symmetry comes to the rescue since the
301: K\"ahler moduli space of a Calabi-Yau $X$ can be mapped to the
302: complex structure moduli space of its mirror $\tilde X$. The
303: K\"ahler prepotential can then be computed using the mirror map.
304: 
305: To be more precise, if $\mathcal{F}$ is the prepotential on the
306: moduli space of complex structures of a Calabi-Yau
307: manifold $X$, one has the special geometry relations \cite{CO90}\\
308: \parbox{14cm}{
309: \begin{eqnarray}
310: X^I&=&\int_{\G_{A^I}}\O\ ,\nonumber\\
311: \mathcal{F}_I\equiv{\partial\mathcal{F}\over\partial
312: X^I}&=&\int_{\G_{B_I}}\O\ .\nonumber
313: \end{eqnarray}}\hfill\parbox{8mm}{\begin{eqnarray}\label{SG}\end{eqnarray}}\\
314: Here $\O$ is the unique holomorphic $(3,0)$-form on the
315: Calabi-Yau, $I\in\{0,\ldots ,h^{2,1}\}$ and
316: $\{\G_{A^I},\G_{B_J}\}$ is a symplectic basis of $H_3(X)$. The
317: homogeneous function $\mathcal{F}$ can be obtained from
318: $2\mathcal{F}=X^I\mathcal{F}_I$.
319: 
320: Triggered by the success of Seiberg and Witten in solving
321: four-dimensional gauge theories \cite{SW94} it became apparent
322: that {\it local} Calabi-Yau manifolds are also quite useful to
323: extract important information about four-dimensional physics
324: \cite{KLMVW96}. These are {\it non-compact} K\"ahler manifolds
325: with vanishing first Chern class and the idea is that they are
326: local models of proper Calabi-Yau manifolds. They also appeared in
327: the context of geometrical engineering and local mirror symmetry,
328: see e.g. \cite{KKV96}, where they are constructed from toric
329: geometry, and especially the analysis of the topological string
330: with these target manifolds has led to a variety of interesting
331: results \cite{AKMV03, ADKMV03}, see \cite{M04} for a review.
332: 
333: More recently a different class of local Calabi-Yau manifolds
334: appeared \cite{KKLM99}, \cite{CIV01}. These are given as
335: deformations of the space
336: $\mathcal{O}(-2)\oplus\mathcal{O}(0)\rightarrow \mathbb{CP}^1$ by
337: a polynomial $W$. Instead of a single $\mathbb{CP}^1$, the
338: deformed manifold, let us call it $Y$, contains $n$ two-spheres if
339: the degree of $W$ is $n+1$. These spaces can be taken through a
340: geometric transition \cite{CIV01}, similar to the conifold
341: transition of Gopakumar and Vafa \cite{GV98}. The resulting space,
342: we call it $X$, is a hypersurface in $\mathbb{C}^4$ described by
343: \begin{equation}
344: W'(x)^2+f_0(x)+v^2+w^2+z^2=0\ ,\label{deformed}
345: \end{equation}
346: where $(v,w,x,z)\in\mathbb{C}^4$ and $f_0(x)$ is a polynomial of degree $n-1$.\\
347: An obvious and important question to ask is whether we can find
348: special geometry for these manifolds as well. In fact, a local
349: Calabi-Yau manifold also comes with a holomorphic $(3,0)$-form
350: $\O$ and we want to check whether its integrals over an
351: appropriate basis of three-cycles satisfy (\ref{SG}). Clearly, the
352: naive special geometry relations need to be modified since our
353: local Calabi-Yau manifold $X$ now contains a non-compact
354: three-cycle $\G_{\hat B}$ and the integral of $\O$ over this cycle
355: is divergent. This can be remedied by introducing a cut-off
356: $\L_0$, but then the integral over the regulated cycle is cut-off
357: dependent whereas the prepotential is expected to be independent
358: of any cut-off. The question therefore is how the relation
359: \begin{equation}
360: \int_{\G_{B_I}}\O\stackrel{?}{=}{\partial\mathcal{F}\over\partial
361: X^I}
362: \end{equation}
363: should be modified to make sense on local Calabi-Yau manifolds.
364: Related issues have been addressed recently in \cite{DOV04}.
365: 
366: Mathematically the relation between the spaces $Y$ and $X$ is
367: obvious from the fact that both are related to the singular space
368: \begin{equation}
369: W'(x)^2+v^2+w^2+z^2=0\ .\label{singular}
370: \end{equation}
371: $X$ is simply the deformation\footnote{Usually we use the word
372: ``deformation'' in an intuitive sense, but what is meant here is
373: the precise mathematical term.} and $Y$ is nothing but the small
374: resolution of all the singularities in (\ref{singular}). Following
375: previous work in \cite{GV98,V01} it was shown in \cite{CIV01} that
376: this geometric transition has a beautiful physical interpretation
377: in type IIB string theory. Starting from the manifold $Y$ one can
378: wrap $N_i$ D5-branes around the $i$-th $\mathbb{CP}^1$ to get an
379: effective $\mathcal{N}=1$ $U(N)$ theory with an adjoint field
380: $\Phi$ in a vacuum that breaks the gauge group as $U(N)\rightarrow
381: U(N_1)\times\ldots\times U(N_{n})$, where $N:=\sum_iN_i$. After
382: the transition the branes disappear and we are left with a dual
383: $\mathcal{N}=1$ $U(1)^{n}$ theory with background flux three-form
384: $H$ with $\int_{\G_{A^i}}H=N_i$. The effective superpotential of
385: the dual gauge theory can be calculated from the formula
386: \cite{CIV01}
387: \begin{equation}
388: W_{eff}=\sum_{i=1}^{n}\left(\int_{\G_{A^i}}H\int_{\G_{B_i}}\O-\int_{\G_{B_i}}H\int_{\G_{A^i}}\O\right)\
389: .\label{Wintro}
390: \end{equation}
391: The $\G_{A^i},\G_{B_i}$ form a symplectic basis of three-cycles
392: and Eq. (\ref{Wintro}) obviously is invariant under symplectic
393: changes of basis which include the ``electric magnetic" duality
394: transformations $\G_{A^i}\rightarrow\G_{B_i},\
395: \G_{B_i}\rightarrow-\G_{A^i}$. Note that we do not write the
396: right-hand side as $\int_X H\w\O$ as it is not clear whether the
397: Riemann bilinear relation can be extended to non-compact
398: Calabi-Yau manifolds without modification. In \cite{CIV01} the
399: $\G_A$-cycles where taken to be compact and the $\G_B$-cycles all
400: non-compact. But $\int_{\G_{B_i}}\O$ contains a term that diverges
401: polynomially together with a term with a logarithmic divergence.
402: The latter has a nice interpretation in terms of the $\b$-function
403: of the gauge theory but the polynomial divergence has not been
404: understood. One of the goals of this note is to shed some light on
405: this aspect.
406: 
407: In a series of influential papers \cite{DV02}, Dijkgraaf and Vafa
408: reviewed these local Calabi-Yau manifolds and showed that the
409: field theory corresponding to branes wrapped on $\mathbb{CP}^1$s
410: in $Y$, which is holomorphic Chern-Simons theory \cite{Wi95},
411: reduces to a {\it holomorphic} matrix model. In fact, as will be
412: discussed below, the structure of the space (\ref{deformed}) is
413: essentially captured by a Riemann surface and a very similar
414: Riemann surface appears in the planar limit of the matrix model.
415: This is why one can learn something about the local Calabi-Yau
416: manifold from an analysis of the well-understood matrix model.
417: Specifically we are interested in understanding the detailed form
418: of the special geometry relations on local Calabi-Yau manifolds
419: from the analysis in the holomorphic matrix model.
420: 
421: The holomorphic matrix model is similar to the hermitian one, but
422: its potential $W(x)$ is defined on the complex plane, has complex
423: coefficients and the integration is performed over complex
424: $\hN\times \hN$ matrices with eigenvalues that are constrained to
425: lie on some path $\g$ in the complex plane. The precise definition
426: and solution involve various subtleties, many of which have been
427: addressed in \cite{La03}, and others will be clarified in this
428: note. The planar limit of the free energy of the matrix model is
429: given, as usual, by a saddle point approximation. We show that
430: saddle point solutions exist only for an appropriate choice of the
431: path $\g$, which is determined self-consistently in such a way
432: that {\it all} critical points of $W(x)$ appear as {\it stable}
433: critical points along the path! For the case of finite $\hN$ this
434: can be seen from an approximate solution of the saddle point
435: equations. In the planar limit one usually constructs the
436: eigenvalue density $\r_0(s)$ ($s$ is a real parameter along $\g$)
437: from the Riemann surface that appears in this limit. As a matter
438: of fact, every Riemann surface that arises in the planar limit of
439: a matrix model leads to a {\it real} density $\r_0(s)$. A way to
440: see this is to note that the filling fractions, i.e. the numbers
441: of eigenvalues in certain domains of $\mathbb{C}$, can be
442: calculated as real integrals over $\r_0(s)$. One can also turn the
443: argument around and construct a $\r_0(s)$ from an arbitrary
444: hyperelliptic Riemann surface. In general this $\r_0(s)$ will be
445: complex and one obtains constraints on the moduli of the surface
446: from the condition that $\r_0(s)$ should be real. Once we fix the
447: filling fractions, the moduli of the Riemann surface are in fact
448: uniquely determined. This, in turn, gives the positions of the
449: cuts $\mathcal{C}_i$ which support the eigenvalues and as the cuts
450: have to lie on the path $\g$ we get conditions
451: for the path.\\
452: Coupling the filling fractions to sources then gives a planar free
453: energy $F_0(J_i)$, and its Legendre transform
454: $\mathcal{F}_0(\tilde S_i)$ is the candidate prepotential. In
455: fact, the $\tilde S_i$ are related to the filling fractions in
456: such a way that they are given by the period integrals over the
457: (compact) $\a_i$-cycles on the Riemann surface and the
458: ${\partial\mathcal{F}_0\over\partial \tilde S_i}(\tilde S_j)$ can
459: be shown to be integrals over the corresponding (compact)
460: $\b_i$-cycles. These properties can in fact be generalised to
461: arbitrary hyperelliptic Riemann surfaces by analytically
462: continuing the $\tilde S_i$ to complex values. The prepotential
463: then still has the same form, but now it depends on complex
464: variables (and then it can no longer be interpreted as the planar
465: limit of the free energy of a matrix model). This proves the
466: standard special geometry relations for the standard cycles. The
467: same methods allow us to derive the modifications of the special
468: geometry relations for the relative cycles appearing in the setup.
469: Indeed, the non-compact period integrals contain, in addition to
470: the derivatives of the prepotential, a polynomial and a
471: logarithmical cut-off dependence and can therefore {\it not} be
472: written as a derivative of the prepotential. While the logarithmic
473: divergence is interpreted as related to the $\b$-function of the
474: dual gauge theory, the polynomial divergence has no counterpart
475: and should not appear in the effective superpotential. This will
476: be achieved by defining appropriate pairings similar to the ones
477: appearing in relative cohomology.
478: 
479: The analysis in the matrix model and the derivation of the special
480: geometry relations show that it is useful to work with a
481: symplectic basis of (relative) one-cycles on the Riemann surface
482: which consists of $n-1$ compact cycles $\a^i$ and the $n-1$
483: corresponding compact cycles $\b_i$, together with two (relative)
484: cycles $\hat\a$ and $\hat\b$, where only $\hat\b$ is non-compact.
485: Indeed, then one can perform symplectic transformations in the set
486: $\{\a^i,\b_j\}$ maintaining the usual special geometry relations.
487: However, once the relative cycle $\hat\b$ is combined with other
488: cycles the special geometry relations are modified. Quite
489: importantly the transformed prepotential always stays finite and
490: cut-off independent.
491: 
492: This paper is organised as follows. In the next section we explain
493: the structure of the local Calabi-Yau spaces we are considering.
494: In particular we review how the set of three-cycles in $X$ maps to
495: the set of relative one-cycles on a Riemann surface with marked
496: points. Section three deals with holomorphic matrix models, where
497: the potential of the model is chosen to correspond to the $W(x)$
498: of the Calabi-Yau manifold. We start with a short exposition of
499: general facts from holomorphic matrix models and then discuss how
500: to deal with the above-mentioned subtleties. We explain how
501: special geometry arises from the matrix model and how the
502: modifications for the non-compact cycles can be derived.
503: Furthermore, we discuss the properties of the prepotential and how
504: electric-magnetic duality is implemented. In section four we
505: propose a formula for the effective superpotential of IIB string
506: theory on these local Calabi-Yau manifolds. It contains the
507: above-mentioned pairings that are similar to the ones appearing in
508: relative cohomology and provides a precise reformulation of the
509: formulae found in \cite{CIV01} and \cite{DV02}. Section five
510: contains our conclusions.
511: 
512: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
513: \section{Local Calabi-Yau Manifolds and hyperelliptic Riemann surfaces}
514: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
515: \setcounter{equation}{0} Let then $(v,w,x,z)\in\mathbb{C}^4$,
516: $W(x)$ a polynomial of degree $n+1$ with $W(0)=g_0$, leading
517: coefficient one, and non-degenerate critical points, i.e. if
518: $W'(p)=0$ then $W''(p)\neq0$. Furthermore let $f_0(x)$ a
519: polynomial of degree $n-1$. In this note we are only interested in
520: local Calabi-Yau manifolds $X$ described by the equation
521: \begin{equation}
522: F(v,w,x,z)\equiv W'(x)^2+f_0(x)+v^2+w^2+z^2=0\ .\label{locCY}
523: \end{equation}
524: In particular, we want to see how the special geometry relations
525: (\ref{SG}) have to be modified in this case.
526: 
527: The holomorphic three-form on $X$ is given as\footnote{If
528: $F(v,w,x,z)$ is a holomorphic function on $\mathbb{C}^4$ then $\d
529: F$ is perpendicular to the hypersurface $F=0$. From the
530: holomorphic four-form $\d v\w\d w\w\d x\w\d z$ on $\mathbb{C}^4$
531: one defines the holomorphic three-form $\O$ on $F=0$ as the form
532: that satisfies $\d v\w\d w\w\d x\w\d z=\O\w\d F$.} \cite{CIV01},
533: \cite{AGLV}
534: \begin{equation}
535: \O={\d v\w\d w\w\d x\over{2z} }\ .
536: \end{equation}
537: Because of the simple dependence of the surface (\ref{locCY}) on
538: $v$ and $w$, every three-cycle of the space (\ref{locCY}) can be
539: understood \cite{KLMVW96} as a fibration of a two-sphere over a
540: line segment in the hyperelliptic Riemann surface $\S$,
541: \begin{equation}
542: y^2=W'(x)^2+f_0(x)=\prod_{i=1}^{n}(x-a_i^+)(x-a_i^-)\
543: ,\label{Riemannsurface}
544: \end{equation}
545: of genus $\gh=n-1$, see \cite{L96} for a review. $\S$ is a
546: two-sheeted covering of the complex plane where the two sheets are
547: connected by $n$ cuts between the points $a_i^-$ and $a_i^+$. Our
548: conventions are such that if $y_0$ is the branch of the Riemann
549: surface with $y_0(x)\sim W'(x)$ for $|x|\rightarrow \infty$, then
550: $y_0$ is defined on the upper sheet and $y_1=-y_0$ on the lower
551: one. For compact three-cycles the line segment connects two of the
552: branch points of the curve and the volume of the $S^2$-fibre
553: depends on the position on the base line segment. At the end
554: points of the segment one has $y^2=0$ and the volume of the sphere
555: shrinks to zero size. Non-compact three-cycles on the other hand
556: are fibrations of $S^2$ over a half-line that runs from one of the
557: branch points to infinity on the Riemann surface. Integration over
558: the fibre is elementary and gives
559: \begin{equation}\label{fibration}
560: \int_{S^2}\O=\pm 2\pi i\ y(x)\d x\ ,
561: \end{equation}
562: (the sign ambiguity will be fixed momentarily) and thus the
563: integral of the holomorphic $\O$ over a three-cycle is reduced to
564: an integral of $\pm 2\pi i y\d x$ over a line segment in $\S$.
565: Clearly, the integrals over line segments that connect two branch
566: points can be rewritten in terms of integrals over compact cycles
567: on the Riemann surface, whereas the integrals over non-compact
568: three-cycles can be expressed as integrals over a line that links
569: the two infinities on the two complex sheets. In fact, the
570: one-form
571: \begin{equation}
572: \zeta:=y\d x
573: \end{equation}
574: is meromorphic and diverges at infinity (poles of order $n+2$) on
575: the two sheets and therefore it is well-defined only on the
576: Riemann surface with the two infinities, we denote them by $Q$ and
577: $Q'$, removed. This surface with two points cut out is called
578: $\hat\S$. We are naturally led to consider the relative
579: homology\footnote{Let $M$ be a manifold and $N$ a submanifold of
580: $M$ and $C_j(M),C_j(N)$ the set of $j$-chains in $M$ and $N$,
581: respectively. One defines the group $C_j(M,N)$ of equivalence
582: classes of $j$-chains $c_j\in C_j(M)$, where $[c_j]:=c_j+C_j(N)$.
583: Then two chains are equivalent if they differ only by a chain in
584: $N$. As usual $H_j(M,N)=Z_j(M,N)/B_j(M,N)$, where
585: $Z_j(M,N):=\{[c_j]\in C_j(M,N):\partial[c_j]=[0]\}$ and
586: $B_j(M,N):=\partial C_{j+1}(M,N)$. Note that a representative
587: $c_j$ of an element in $H_j(M,N)$ may have a boundary as long as
588: the boundary lies in $N$.} $H_1(\S,\{Q,Q'\})$. This group contains
589: both, all the compact cycles, as well as cycles connecting $Q$ and
590: $Q'$ on the Riemann surface. If we take, for example,
591: $W(x)={x^2\over2}$ and $f_0(x)=-\m$ the surface (\ref{locCY}) is
592: nothing but the deformed conifold, which is $T^*S^3$. This space
593: contains two three-cycles, the compact base $\G_{\hat\a}\cong
594: S^3$, which maps to the compact one-cycle $\hat\a$ surrounding the
595: cut of the surface $y^2=x^2-\m$, and the non-compact fiber
596: $\G_{\hat\b}:=T_p^*S^3$, which maps to the non-compact one-cycle
597: $\hat\b$ which runs from $Q'$, i.e. infinity on the lower sheet
598: through the cut to $Q$, i.e. infinity on the upper sheet. This can
599: be generalised readily for arbitrary polynomials $W,f_0$ and one
600: finds a one-to-one correspondence between the (compact and
601: non-compact) three-cycles in
602: (\ref{locCY}) and $H_1(\S,\{Q,Q'\})$.\\
603: \begin{figure}[h]
604: \centering
605: \includegraphics[width=1.0\textwidth]{ABcycles.eps}\\
606: \caption[]{A symplectic choice of compact $A$- and non-compact
607: $B$-cycles for $n=3$. Note that the orientation of the two planes
608: on the left-hand side is chosen such that both normal vectors
609: point to the top. This is why the orientation of the $A$-cycles is
610: different on the two planes. To go from the representation of the
611: Riemann surface on the left to the one on the right one has to
612: flip the upper plane.} \label{ABcycles}
613: \end{figure}
614: There are various symplectic bases of this relative homology
615: group. One such basis is $\{A^i, B_j\}$, with $i,j=1,\ldots n$,
616: where the one-cycle $A^i$ runs around the $i$-th cut and the
617: relative one-cycles $B_j$ are all non-compact and run from $Q'$
618: through the $j$-th cut to $Q$. This is the choice of cycles used
619: in $\cite{CIV01}$ and it is shown in Fig.\ref{ABcycles}.
620: 
621: Another useful symplectic basis is the set
622: $\{\a^i,\b_j,\hat\a,\hat\b\}$, with $\gh=n-1$ compact cycles
623: $\a^i$ and $\gh=n-1$ compact cycles $\b_i$, with intersection
624: numbers $\a^i\cap\b_j=\delta^i_j$, together with one compact cycle
625: $\hat\a$ and one non-compact cycle $\hat\b$, with
626: $\hat\a\cap\hat\b=1$, see Fig.\ref{albecycles}. Note that although
627: these bases are equivalent, since one can be obtained from the
628: other by a symplectic transformation, the second basis is much
629: more useful for our purpose. This is because it contains only one
630: non-compact cycle and the new features coming from the
631: non-compactness of the space should be contained entirely in the
632: corresponding integral. Finally, we take $\G_{A^i},\G_{B_j}$ to be
633: the $S^2$-fibrations over $A^i,B_j$ and $\G_{\a^i},\G_{\b_j}$
634: $S^2$-fibrations over $\a^i,\b_j$.
635: \begin{figure}[h]
636: \centering
637: \includegraphics[width=1.0\textwidth]{albecycles.eps}\\
638: \caption[]{A symplectic set of cycles containing only one
639: non-compact cycle $\hat\b$. The cycle $\a^i$ surrounds $i$ of the
640: cuts, whereas the cycle $\b_i$ runs from cut $i$ to cut $i+1$ on
641: the upper sheet and from cut $i+1$ to cut $i$ on the lower one. As
642: before one has to flip the upper plane to find the representation
643: of the surface on the right.} \label{albecycles}
644: \end{figure}
645: 
646: So the problem effectively reduces to calculating the
647: integrals\footnote{The sign ambiguity of (\ref{fibration}) has now
648: been fixed, since we have made specific choices for the
649: orientation of the cycles. Furthermore, we use the (standard)
650: convention that the cut of $\sqrt{x}$ is along the negative real
651: axis of the complex $x$-plane. Also, on the right-hand side we
652: used that the integral of $\zeta$ over the line segment is $1\over
653: 2$ times the integral over a closed cycle $\g$.}
654: \begin{equation}
655: \int_{\G_{\g}}\O=-i\pi\int_{\g}\zeta\ \ \ \mbox{for}\ \ \
656: \g\in\{\a^i,\b_j,\hat\a,\hat\b\}\ .\label{integrals}
657: \end{equation}
658: For $\gh=1$ they can be reduced to various combinations of
659: elliptic integrals of the three kinds.
660: 
661: As we mentioned already, we expect new features to be contained in
662: the integral $\int_{\hat\b}\zeta$, where $\hat\b$ runs from $Q'$
663: on the lower sheet to $Q$ on the upper one. Indeed, it is easy to
664: see that this integral is divergent. It will be part of our task
665: to understand and properly treat this divergence. As usual, we
666: will regulate the integral and we have to make sure that physical
667: quantities do not depend on the regulator and remain finite once
668: the regulator is removed. Usually this is achieved by simply
669: discarding the divergent part. Instead, we want to give a more
670: intrinsic geometric prescription that will be similar to standard
671: procedures in relative cohomology. To render the integral finite
672: we simply cut out two ``small'' discs around the points $Q,Q'$. If
673: $x,x'$ are coordinates on the upper and lower sheet respectively,
674: we restrict ourselves to $|x|\leq \L_0$, $|x'|\leq\L_0$,
675: $\L_0\in\mathbb{R}$. Furthermore we take the cycle $\hat\b$ to run
676: from the point $\L_0'$ on the real axis of the lower sheet to
677: $\L_0$ on the real axis of the upper sheet. (Actually we could
678: take $\L_0$ and $\L_0'$ to be complex. We will come back to this
679: point later on.)
680: 
681: 
682: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
683: \section{Holomorphic Matrix Models and Special Geometry}
684: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
685: \setcounter{equation}{0}
686: 
687: Our goal is to relate the integrals (\ref{integrals}) to the
688: prepotential $\mathcal{F}_0$. It turns out that in order to
689: address this problem it is useful to perform calculations in the
690: matrix model that corresponds to our local Calabi-Yau manifold.
691: Indeed, the analysis of Dijkgraaf and Vafa tells us \cite{DV02}
692: that one should identify the prepotential and the planar limit of
693: the free energy of the holomorphic matrix model with potential
694: $W(x)$. Therefore, our goal will be to find the special geometry
695: relation in the holomorphic matrix model and to see how the
696: integrals (\ref{integrals}) over the cycles
697: $\a^i,\b_j,\hat\a,\hat\b$ are related to the planar limit of the
698: free energy.\\
699: One should note, however, that in the matrix model the filling fractions $S_i$, related to the
700: integrals over the $A$-cycles, are manifestly real, even though
701: $W(x)$ has arbitrary complex coefficients. So, strictly speaking,
702: the matrix model does {\it not} explore the full moduli space of
703: the Calabi-Yau manifold. Nevertheless, we will see that all
704: relevant formulae can be immediately continued to complex values
705: of the $S_i$ and, in particular, the
706: special geometry relations continue to be true.
707: 
708: \subsection{The Holomorphic Matrix Model}
709: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
710: The proper definition of the holomorphic matrix model is somewhat
711: more subtle than the one of the hermitian matrix model. Many of
712: these subtleties were nicely addressed in \cite{La03} and we will
713: briefly review them here. The usual identification of the planar
714: limit with the saddle point approximation involves even more
715: subtleties which we will have to clarify in this subsection.
716: Particular attention is paid to the dependence of the free energy
717: on the various parameters.
718: 
719: \subsubsection{The partition function and convergence properties}
720: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
721: We begin by defining the partition function of the {\it
722: holomorphic} one-matrix model following \cite{La03}. In order to
723: do so, one chooses a smooth path
724: $\g:\mathbb{R}\rightarrow\mathbb{C}$ without self-intersection,
725: such that $\dot{\g}(u)\neq 0\ \forall u\in\mathbb{R}$ and
726: $|\g(u)|\rightarrow \infty$ for $u\rightarrow\pm\infty$. Consider
727: the ensemble $\G(\g)$ of\footnote{We reserve the letter $N$ for
728: the number of colours in a $U(N)$ gauge theory. It is important to
729: distinguish between $N$ in the gauge theory and $\hat{N}$ in the
730: matrix model.} $\hN\times\hN$ complex matrices $M$ with spectrum
731: spec$(M)=\{\l_1,\ldots\l_{\hN}\}$ in\footnote{Here and in the
732: following we will write $\g$ for both the function and its image.}
733: $\g$ and distinct eigenvalues,
734: \begin{equation}
735: \Gamma(\gamma):=\{M\in\mathbb{C}^{\hN\times\hN}:\mbox{spec}(M)\subset\gamma,\
736: \mbox{all}\ \l_m \ \mbox{distinct}\}\ .
737: \end{equation}
738: The holomorphic measure on $\mathbb{C}^{\hat{N}\times \hat{N}}$ is
739: just $\d M\equiv\wedge_{p,q}\d M_{pq}$ with some appropriate sign
740: convention. The (super-)potential is
741: \begin{equation}
742: W(x):=g_0+\sum_{k=1}^{n+1} {g_k\over k} x^k,\ \ \ g_{n+1}=1\ .
743: \end{equation}
744: Without loss of generality we have chosen $g_{n+1}=1$. The only
745: restriction for the other complex parameters $\{g_k\}_{k=0,\ldots
746: n}$, collectively denoted by $g$, comes from the fact that the $n$
747: critical points $\m_i$ of $W$ should not be degenerate, i.e.
748: $W''(\m_i)\neq0$ if $W'(x)=\prod_{i=1}^n(x-\m_i)$. Then the
749: partition function of the holomorphic one-matrix model is
750: \begin{equation}
751: Z(\Gamma(\g),g,g_s,\hN):=C_{\hN}\int_{\Gamma(\g)}\d M \
752: \exp\left(-{1\over g_s} \tr W(M)\right),
753: \end{equation}
754: where $g_s$ is a positive coupling constant and $C_{\hN}$ is some
755: normalisation factor. To avoid cluttering the notation we will
756: omit the dependence on $\g$ and $g$ and write
757: $Z(g_s,\hN):=Z(\G(\g),g,g_s,\hN)$. As usual one diagonalises $M$
758: and performs the integral over the diagonalising matrices. The
759: constant $C_{\hN}$ is chosen in such a way that one arrives at
760: \begin{equation}
761: Z(g_s,\hN)={1\over{\hat{N}!}}\int_\gamma\d\lambda_1\ldots\int_\gamma\d\lambda_{\hat{N}}
762: \exp\left({-\hN^2S(g_s,\hN;\l_m)}\right)=:e^{-F(g_s,\hN)}\ ,
763: \end{equation}
764: where
765: \begin{equation}
766: S(g_s,\hN;\l_m)={1\over\hN^2g_s}\sum_{m=1}^{\hN}W(\l_m)-{1\over\hN^2}\sum_{p\neq
767: q}\ln(\l_p-\l_q)\ .
768: \end{equation}
769: See \cite{La03} for more details.
770: 
771: The convergence of the $\l_m$ integrals depends on the polynomial
772: $W$ and the choice of the path $\g$. For given $W$ the asymptotic
773: part of the complex plane ($|x|$ large) can be divided into
774: convergence domains $G_k^{(c)}$ and divergence domains
775: $G_k^{(d)}$, $k=1,\ldots n+1$, where $e^{-{1\over g_s}W(x)}$
776: converges, respectively diverges as $|x|\rightarrow\infty$. The
777: path $\g$ has to be chosen \cite{La03} to go from some convergence
778: domain $G_k^{(c)}$ to some other $G_l^{(c)}$, with $k\neq l$; call
779: such a path $\g_{kl}$, see Fig.\ref{convergence1}.
780: \begin{figure}[h]
781: \centering
782: \includegraphics{convergence1.eps}\\
783: \caption[]{Example of convergence and divergence domains for $n=2$
784: and a possible choice of $\g_{21}$. Because of holomorphicity the
785: path can be deformed without changing the partition function, for
786: instance one could use the path $\tilde\g_{21}$ instead.}
787: \label{convergence1}
788: \end{figure}
789: Then the value of the partition function depends only on the pair
790: $(k,l)$ and, because of holomorphicity, is not sensitive to
791: deformations of $\g_{kl}$. In particular, instead of $\g_{kl}$ we
792: can make the equivalent choice
793: \begin{equation}
794: \tilde\g_{kl}=\g_{p_1p_2}\cup\g_{p_2p_3}\cup\ldots\cup\g_{p_{n-1}p_{n}}\cup\g_{p_{n}p_{n+1}}\
795: \ \ \mbox{with}\ p_1=k,\ p_{n+1}=l ,\label{tildegamma}
796: \end{equation}
797: as shown in Fig.\ref{convergence1}. Here we split the path into
798: $n$ components, each component running from one convergence
799: sector to another. Again, due to holomorphicity we can choose the
800: decomposition in such a way that every component $\g_{p_ip_{i+1}}$
801: runs through one of the $n$ critical points of $W$ in
802: $\mathbb{C}$, or at least comes close to it. This choice of
803: $\tilde\g_{kl}$ will turn out to be very useful to understand the
804: saddle point approximation discussed
805: below.
806: Hence, the partition function and the free energy depend on the
807: pair $(k,l),g,g_s$ and $\hN$. Of course, one can always relate the
808: partition function for arbitrary $(k,l)$ to one with $(k',1)$, $k'=k-l+1$ mod
809: $n$, and
810: redefined coupling constants $g$.
811: 
812: \subsubsection{Matrix model technology}
813: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
814: Next, we need to recall some standard technology adapted to the
815: holomorphic matrix model. We first assume that the path $\g$
816: consists of a single connected piece. The case (\ref{tildegamma})
817: will be discussed later on. Let $s$ be the length coordinate of
818: the path $\g$, centered at some point on $\g$, and let $\l(s)$
819: denote the parameterisation of $\g$ with respect to this
820: coordinate. Then, for an eigenvalue $\l_m$ on $\g$, one has
821: $\l_m=\l(s_m)$ and the partition function can be rewritten as
822: \begin{equation}\label{part(s)}
823: Z(g_s,\hN)={1\over{\hN!}}\int_{\mathbb{R}}\d
824: s_1\ldots\int_{\mathbb{R}}\d
825: s_{\hat{N}}\prod_{l=1}^{\hat{N}}\dot{\l}(s_l)\exp\left(-\hN^2S(g_s,\hN;\l(s_m))\right)\
826: .
827: \end{equation}
828: The spectral density is defined as
829: \begin{equation}
830: \r(s,s_m):={1\over \hat{N}}\sum_{m=1}^{\hat{N}}\delta(s-s_m)\ ,
831: \end{equation}
832: so that $\r$ is normalised to one,
833: $\int_{-\infty}^\infty\r(s,s_m)\d s=1$. The normalised trace of
834: the resolvent of the matrix $M$ is given by
835: \begin{equation}\label{omrho}
836: \o (x,s_m):={1\over \hat{N}}\tr {1\over {x-M}}={1\over
837: \hat{N}}\sum_{m=1}^{\hat{N}}{1\over {x-\l(s_m)}}=\int\d
838: s{\r(s,s_m)\over {x-\l(s)}}\ ,
839: \end{equation}
840: for $x\in\mathbb{C}$. Following \cite{La03} we decompose the
841: complex plane into domains $D_i$, $i\in\{1,\ldots, n\}$, with
842: mutually disjoint interior, $( \cup_i\overline{D}_i=\mathbb{C},\ \
843: D_i\cap D_j=\emptyset\ \ \mbox{for}\ i\neq j)$. These domains are
844: chosen in such a way that $\g$ intersects each $\overline{D}_i$
845: along a single line segment $\D_i$, and $\cup_i\D_i=\g$.
846: Furthermore, $\m_i$, the $i$-th critical point of $W$, should lie
847: in the interior of $D_i$. One defines
848: \begin{equation}
849: \chi_i(M):=\int_{\partial D_i} {\d x\over 2\pi i}{1\over{x-M}}\ ,
850: \end{equation}
851: (which projects on the space spanned by the eigenvectors of $M$
852: whose eigenvalues lie in $D_i$), and the filling fractions
853: $\tilde\s_i(\l_m):={1\over\hN}\tr\chi_i(M)$ and
854: \begin{equation}
855: \s_i(s_m):=\tilde\s_i(\l(s_m))=\int\d s\
856: \r(s,s_m)\chi_i(\l(s))=\int_{\partial D_i}{\d x\over 2\pi i}\
857: \o(x,s_m)\ ,\label{fillingfraction}
858: \end{equation}
859: (which count the eigenvalues in the domain $D_i$, times $1/\hN$). Obviously
860: \begin{equation}\label{sumnu}
861: \sum_{i=1}^{n}\s_i(s_m)=1\ .
862: \end{equation}
863: 
864: One can apply standard methods (e.g. the ones of \cite{K99}) to
865: derive the loop equations of the holomorphic matrix model,
866: \begin{equation}\label{loop}
867: \langle \o(x,s_m)^2\rangle-{1\over
868: t}W'(x)\langle\o(x,s_m)\rangle-{1\over 4t^2}\langle
869: f(x,s_m)\rangle=0\ .
870: \end{equation}
871: Here
872: \begin{equation}
873: t=g_s\hN\label{t}
874: \end{equation}
875: is the quantity that will be held fixed in the planar limit below,
876: \begin{equation}
877: f(x,s_m):=-{4t\over{\hN}}\sum_{m=1}^{\hat{N}}{{W'(x)-W'(\l(s_m))}\over{x-\l(s_m)}}=-4t\int
878: \d s\ \r(s,s_m){{W'(x)-W'(\l(s))}\over{x-\l(s)}}\ ,
879: \end{equation}
880: and the expectation value is defined for a $h(\l_m)=h(\l(s_m))$ as
881: usual:
882: \begin{equation}
883: \langle h(\l_m)\rangle:={1\over Z(g_s,\hN)}\cdot{1\over
884: \hN!}\int_{\g}\d\l_1\ldots\int_{\g}\d\l_{\hat{N}}\ h(\l_m)
885: \exp\left(-\hN^2S(g_s,\hN;\l_m)\right)\ .
886: \end{equation}
887: 
888: It will be useful to define an effective action as
889: \begin{eqnarray}\label{effaction}
890: &&S_{eff}(g_s,\hN;s_m):=S(g_s,\hN;\l(s_m))-{1\over\hN^2}\sum_{m=1}^{\hN}\ln(\dot\l(s_m))\nonumber\\
891: &&=\int\d s\ \r(s;s_m)\left({1\over
892: t}W(\l(s))-{1\over{\hN}}\ln(\dot{\l}(s))-\mathcal{P}\int\d s' \
893: \r(s';s_p)\ln(\l(s)-\l(s'))\right)\ \ \ \ \ \ \ \ \ \
894: \end{eqnarray}
895: so that
896: \begin{equation}
897: Z(g_s,\hN)={1\over {\hN!}}\int\d s_1\ldots\int\d s_{\hat{N}}\
898: \exp\left(-\hN^2 S_{eff}(g_s,\hN;s_m)\right)\ .
899: \end{equation}
900: Note that the principal value is defined as
901: \begin{equation}\label{Pvalue}
902: \mathcal{P}\
903: \ln\left(\l(s)-\l(s')\right)={1\over2}\lim_{\e\rightarrow0}\left[\ln\left(\l(s)-\l(s')+
904: i\e\dot\l(s)\right)+\ln\left(\l(s)-\l(s')-i\e\dot\l(s)\right)\right]\
905: .
906: \end{equation}
907: The equations of motion corresponding to this effective action,
908: ${\delta S_{eff}\over\delta s_m}=0$, read
909: \begin{equation}\label{eommatrix}
910: {1\over t}W'(\l(s_m))={2\over{\hat{N}}}\sum_{p=1,\ p\neq
911: m}^{\hN}{1\over{\l(s_m)-\l(s_p)}}+{1\over{\hat{N}}}{\ddot{\l}(s_m)\over\dot{\l}(s_m)^2}\
912: .
913: \end{equation}
914: Using these equations of motion one can show that
915: \begin{eqnarray}\label{eomloop}
916: &&\o(x,s_m)^2-{1\over t}W'(x)\o(x,s_m)-{1\over 4 t^2}f(x,s_m)+\nonumber\\
917: &&+{1\over\hN}{\d\over\d x}\o(x,s_m)+{1\over\hN^2}
918: \sum_{m=1}^{\hN}{\ddot\l(s_m)\over\dot\l(s_m)^2}{1\over{x-\l(s_m)}}=0\
919: .
920: \end{eqnarray}
921: \\
922: \underline{Solutions of the equations of motion}\\
923: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
924: Note that in general the effective action is a complex function of
925: the real $s_m$. Hence, in general, i.e. for a generic path
926: $\g_{kl}$ with parameterisation $\l(s)$, there will be no solution
927: to (\ref{eommatrix}). One clearly expects that the existence of
928: solutions must constrain the path $\l(s)$ appropriately. Let us
929: study this in more detail.\\
930: Recall that we defined the domains
931: $D_i$ in such a way that $\m_i\subset D_i$. Let $\hN_i$ be the
932: number of eigenvalues $\l(s_m)$ which lie in the domain $D_i$, so
933: that $\sum_{i=1}^n\hN_i=\hN$, and denote them by $\l(s_a^{(i)})$,
934: $a\in\{1,\ldots\hN_i\}$.\\
935: Solving the equations of motion in general is a formidable
936: problem. To get a good idea, however, recall the picture of
937: $\hN_i$ fermions filled into the $i$-th ``minimum" of ${1\over
938: t}W$ \cite{K91}. For small $t$ the potential is deep and the
939: fermions are located not too far from the minimum, in other words
940: all the eigenvalues are close to $\m_i$. To be more precise
941: consider (\ref{eommatrix}) and drop the last term, an
942: approximation that will be justified momentarily. Let us take $t$
943: to be small and look for solutions\footnote{One might try the
944: general ansatz $\l(s_a^{(i)})=\m_i+\e\delta\l_a^{(i)}$ but it
945: turns out that a solution can be found only if $\e\sim\sqrt{t}$.}
946: $\l(s_a^{(i)})=\m_i+\sqrt{t}\delta\l_a^{(i)}$, where
947: $\delta\l_a^{(i)}$ is of order one. So, we assume that the
948: eigenvalues $\l(s_a^{(i)})$ are not too far from the critical
949: point $\m_i$. Then the equation reads
950: \begin{equation}
951: W''(\m_i)\delta\l_a^{(i)}={2\over\hN}\sum_{b=1,\ b\neq
952: a}^{\hN_i}{1\over{\delta\l^{(i)}_a-\delta\l_b^{(i)}}}+o(\sqrt{t})\
953: ,
954: \end{equation}
955: so we effectively reduced the problem to finding the solution for
956: $n$ distinct quadratic potentials. If we set $z_a:=\sqrt{\hN
957: W''(\m_i)\over2}\delta\l_a^{(i)}$ and neglect the
958: $o(\sqrt{t})$-terms this gives
959: \begin{equation}
960: z_a=\sum_{b=1,\ b\neq a}^{\hN_i}{1\over {z_a-z_b}}\ ,
961: \end{equation}
962: which can be solved explicitly for small $\hN_i$. It is obvious
963: that $\sum_{a=1}^{\hN_i}z_a=0$, and one finds that there is a
964: unique solution (up to permutations) with the $z_a$ symmetrically
965: distributed around 0 on the real axis. This justifies a posteriori
966: that we really can neglect the term proportional to the second
967: derivative of $\l(s)$, at least to leading order. Furthermore,
968: setting $W''(\m_i)=|W''(\m_i)|e^{i\phi_i}$ one finds that the
969: $\l(s_a^{i})$ sit on a tilted line segment around $\m_i$ where the
970: angle of the tilt is given by $-\phi_i/2$. This means for example
971: that for a potential with $W'(x)=x(x-1)(x+1)$ the eigenvalues are
972: distributed on the real axis around $\pm1$ and on the imaginary
973: axis around 0. Note further that, in general, the reality of $z_a$
974: implies that
975: ${W''(\m_i)\over2}\left({\delta\l^{(i)}_a}\right)^2>0$ which tells
976: us that, close to $\m_i$, $W(\l(s))-W(\m_i)$ is real with a {\it
977: minimum} at $\l(s)=\m_i$.\\
978: So we have found that the path $\g_{kl}$ has to go through the
979: critical points $\m_i$ with a tangent direction fixed by the phase
980: of the second derivative of $W$. On the other hand, we know that
981: the partition function does not depend on the form of the path
982: $\g_{kl}$. Of course, there is no contradiction: if one wants to
983: compute the partition function from a saddle point expansion, as
984: we will do below, and as is implicit in the planar limit, one has
985: to make sure that one expands around solutions of
986: (\ref{eommatrix}) and the existence of these solutions imposes
987: conditions on how to choose the path $\g_{kl}$. From now on we
988: will assume that the path is chosen in such a way that it
989: satisfies all these constraints. Furthermore, for later purposes
990: it will be useful to use the path $\tilde\g_{kl}$ of
991: (\ref{tildegamma}) chosen such that its part $\g_{p_ip_{i+1}}$
992: goes through all $\hN_i$ solutions $\l_a^{(i)}$,
993: $a=1,\ldots\hN_i$, and lies entirely in $D_i$, see
994: Fig.\ref{curvesinD}.
995: \begin{figure}[h]
996: \centering
997: \includegraphics{curvesinD.eps}\\
998: \caption[]{For the cubic potential of Fig.\ref{convergence1} we
999: show the choice of the domains $D_1$ and $D_2$ and of the path
1000: $\tilde\g_{21}$ with respect to the two critical points, as well
1001: as the cuts that form around these points.} \label{curvesinD}
1002: \end{figure}\\
1003: It is natural to assume that these properties together with the
1004: uniqueness of the solution (up to permutations) extend to higher
1005: numbers of $\hN_i$ as well. Of course once one goes beyond the
1006: leading order in $\sqrt{t}$ the eigenvalues are no longer
1007: distributed on a straight line, but on a line segment that is
1008: bent in general and that might or might not pass through $\m_i$.\\
1009: \\
1010: \underline{The large $\hN$ limit}\\
1011: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1012: We are interested in the large $\hN$ limit of the matrix model
1013: free energy. It is well known that the expectation values of the
1014: relevant quantities like $\r$ or $\o$ have expansions of the form
1015: \begin{equation}
1016: \langle\r(s,s_m)\rangle=\sum_{I=0}^{\infty}\r_I(s)\hN^{-I}\ \ \ ,\
1017: \ \ \langle\o(x,s_m)\rangle=\sum_{I=0}^{\infty}\o_I(x)\hN^{-I}\
1018: .\label{omexp}
1019: \end{equation}
1020: Clearly, $\o_0(x)$ is related to $\r_0(s)$ by the large $\hN$
1021: limit of (\ref{omrho}), namely
1022: \begin{equation}\label{omegaofrho}
1023: \o_0(x)=\int\d s{\r_0(s)\over{x-\l(s)}}\ .
1024: \end{equation}
1025: We saw already that an eigenvalue ensemble that solves the
1026: equations of motion is distributed along line segments around the
1027: critical points $\m_i$. In the limit $\hN\rightarrow \infty$ this
1028: will turn into a continuous distribution on the segments
1029: $\mathcal{C}_i$, through or close to the critical points of $W$.
1030: Then $\r_0(s)$ has support only on these $\mathcal{C}_i$ and
1031: $\o_0(x)$ is analytic in $\mathbb{C}$ with cuts $\mathcal{C}_i$.
1032: Conversely, $\r_0(s)$ is given by the discontinuity of $\o_0(x)$
1033: across its cuts:
1034: \begin{equation}
1035: \r_0(s):=\dot{\l}(s)\lim_{\e\rightarrow0}{1\over2\pi i}
1036: [\o_0(\l(s)-i\e\dot{\l}(s))-\o_0(\l(s)+i\e\dot{\l}(s))]\label{solution}\
1037: .
1038: \end{equation}
1039: 
1040: The planar limit we are interested in is $\hN\rightarrow\infty$,
1041: $g_s\rightarrow0$ with $t=g_s\hN$ held fixed. Hence we rewrite all
1042: $\hN$ dependence as a $g_s$ dependence and consider the limit
1043: $g_s\rightarrow 0$. Then, the equation of motion (\ref{eommatrix})
1044: reduces to
1045: \begin{equation}\label{eomplanar}
1046: {1\over t}W'(\l(s))=2\mathcal{P}\int \d s'\
1047: {\r_0(s')\over{\l(s)-\l(s')}}\ .
1048: \end{equation}
1049: Note that this equation is only valid for those $s$ where
1050: eigenvalues exist, i.e. where $\r_0(s)\neq 0$. In principle one
1051: can use this equation to compute the planar eigenvalue
1052: distribution $\r_0(s)$ for given $W'$.\\
1053: \\
1054: \underline{Riemann surfaces and planar solutions}\\
1055: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1056: The leading term in the expansion (\ref{omexp}) for
1057: $\langle\o(s,s_m)\rangle$ can be calculated from a saddle point
1058: approximation, where the $\{s_m\}$ are given by a solution
1059: $\{s_m^*\}$ of (\ref{eommatrix}): $\o_0(x)=\o(x;s_m^*)$. This is
1060: true for all ``microscopic'' operators, i.e. operators that do not
1061: modify the saddle point equations (\ref{eommatrix}). (Things would
1062: be different for ``macroscopic''operators like
1063: $e^{\hN\sum_{p=1}^{\hN}V(\l_p)}$.) In particular, this shows that
1064: expectation values factorise in the large $\hN$ limit, and the
1065: loop equation (\ref{loop}) reduces to the algebraic equation
1066: \begin{equation}\label{algconstraint}
1067: \o_0(x)^2-{1\over t}W'(x)\o_0(x)-{1\over 4t^2}f_0(x)=0\ ,
1068: \end{equation}
1069: where
1070: \begin{equation}\label{f0matrix}
1071: f_0(x)=-4t\int \d s \ \r_0(s){{W'(x)-W'(\l(s))}\over{x-\l(s)}}
1072: \end{equation}
1073: is a polynomial of degree $n-1$ with leading coefficient $-4t$.
1074: Note that this coincides with the planar limit of equation
1075: (\ref{eomloop}). If we define
1076: \begin{equation}\label{y0omega0}
1077: y_0(x):=W'(x)-{2t}\o_0(x)\ ,
1078: \end{equation}
1079: then $y_0$ is one of the branches of the algebraic curve
1080: \begin{equation}
1081: y^2=W'(x)^2+f_0(x)\ ,\label{algcurve}
1082: \end{equation}
1083: as can be seen from (\ref{algconstraint}). On this curve we use
1084: the same conventions as in section 2, i.e. $y_0(x)$ is defined on
1085: the upper sheet and cycles and orientations are chosen as in
1086: Fig.\ref{ABcycles} and Fig.\ref{albecycles}.
1087: 
1088: Solving a matrix model in the planar limit means to find a
1089: normalised, real, non-negative $\r_0(s)$ and a path $\tilde
1090: \g_{kl}$ which satisfy (\ref{omegaofrho}), (\ref{eomplanar}) and
1091: (\ref{algconstraint}/\ref{f0matrix}) for a given potential $W(z)$
1092: and a given asymptotics $(k,l)$ of $\g$.
1093: 
1094: Interestingly, for any algebraic curve (\ref{algcurve}) there is a
1095: contour $\tilde \g_{kl}$ supporting a {\it formal} solution of the
1096: matrix model in the planar limit. To construct it start from an
1097: arbitrary polynomial $f_0(x)$ or order $n-1$, with leading
1098: coefficient $-4t$, which is given together with the potential
1099: $W(x)$ of order $n+1$. The corresponding Riemann surface is given
1100: by (\ref{algcurve}), and we denote its branch points by $a_i^\pm$
1101: and choose branch-cuts $\mathcal{C}_i$ between them. We can read
1102: off the two solutions $y_0$ and $y_1=-y_0$ from (\ref{algcurve}),
1103: where we take $y_0$ to be the one with a behaviour
1104: $y_0\stackrel{x\rightarrow\infty}{\rightarrow}+W'(x)$. $\o_0(x)$
1105: is defined as in (\ref{y0omega0}) and we choose a path
1106: $\tilde\g_{kl}$ such that $\mathcal{C}_i\subset \tilde\g_{kl}$ for
1107: all $i$. Then the formal planar spectral density satisfying all
1108: the requirements can be determined from (\ref{solution}) (see
1109: \cite{La03}). However, in general, this will lead to a {\it
1110: complex} distribution $\r_0(s)$. This can be understood from the
1111: fact the we constructed $\rho_0(s)$ from a completely arbitrary
1112: hyperelliptic Riemann surface. However, in the matrix model the
1113: algebraic curve (\ref{algcurve}) is {\it not} general, but the
1114: coefficients $\a_k$ of $f_0(x)$ are constraint. This can be seen
1115: by computing the filling fractions
1116: \begin{equation}\label{nustar}
1117: \n_i^*:=\langle\s_i(s_m)\rangle={1\over2\pi i}\int_{\partial
1118: D_i}\o_0(x)\d x={1\over4\pi it}\int_{A^i}y_0(x)\d
1119: x=\int_{\g^{-1}(\mathcal{C}_i)}\r_0(s)\d s\ ,
1120: \end{equation}
1121: which must be real and non-negative. Here we used the fact that
1122: the $D_i$ were chosen such that $\g_{p_ip_{i+1}}\subset D_i$ and
1123: therefore $\mathcal{C}_i\subset D_i$, so for $D_i$ on the upper
1124: plane, $\partial D_i$ is homotopic to $-A^i$. Hence, ${\rm
1125: Im}\left(i\int_{A^i}y(x)\d x\right)=0$ which constrains the
1126: $\a_k$. We conclude that to construct distributions $\r_0(s)$ that
1127: are relevant for the matrix model one can proceed along the lines
1128: described above, but one has to impose the additional constraint
1129: that $\r_0(s)$ is {\it real}. As for finite $\hN$, this will
1130: impose conditions on the possible paths $\tilde\g_{kl}$ supporting
1131: the eigenvalue distributions.
1132: 
1133: To see this, we assume that the coefficients $\a_k$ in $f_0(x)$
1134: are small, so that the lengths of the cuts are small compared to
1135: the distances between the different critical points:
1136: $|a_i^+-a_i^-|\ll|\m_i-\m_j|$. Then in first approximation the
1137: cuts are straight line segments between $a_i^+$ and $a_i^-$. For
1138: $x$ close to the cut $\mathcal{C}_i$ we have
1139: $y^2\approx(x-a_i^+)(x-a_i^-)\prod_{j\neq
1140: i}(\m_i-\m_j)^2=(x-a_i^+)(x-a_i^-)(W''(\m_i))^2$. If we set
1141: $W''(\m_i)=|W''(\m_i)|e^{i\phi_i}$ and
1142: $a_i^+-a_i^-=r_ie^{i\psi_i}$, then, on the cut $\mathcal{C}_i$,
1143: the path $\g$ is parameterised by
1144: $\l(s)={a_i^+-a_i^-\over|a_i^+-a_i^-|}s=se^{i\psi_i}$, and we find
1145: from (\ref{solution})
1146: \begin{eqnarray}
1147: \r_0(s)&=&{1\over2\pi t }\sqrt{|\l(s)-a_1^+|}\sqrt{|\l(s)-a_1^-|}\ |W''(\m_i)|e^{i(\phi_i+2\psi_i)}\nonumber\\
1148: &=& {1\over2 \pi t }\sqrt{|\l(s)-a_1^+|}\sqrt{|\l(s)-a_1^-|}\
1149: W''(\m_i)(\dot\l(s))^2\ .
1150: \end{eqnarray}
1151: So reality and positivity of $\r_0(s)$ lead to conditions on the
1152: orientation of the cuts in the complex plane, i.e. on the path
1153: $\g$:
1154: \begin{equation}\label{condition}
1155: \psi_i=-{\phi_i/2}\ \ \ \ \ \ ,\ \ \ \ \
1156: W''(\m_i)(\dot\l(s))^2>0\ .
1157: \end{equation}
1158: These are precisely the conditions we already derived for the case
1159: of finite $\hN$. We see that the two approaches are consistent
1160: and, for given $W$ and fixed $\hN_i$ respectively $\n_i^*$, lead
1161: to a unique\footnote{To be more precise the path $\tilde\g_{kl}$
1162: is not entirely fixed. Rather, for every piece
1163: $\tilde\g_{p_ip_{i+1}}$ we have the requirement that
1164: $\mathcal{C}_i\subset\tilde\g_{p_ip_{i+1}}$.} solution
1165: $\{\l(s),\r_0(s)\}$ with real and positive eigenvalue
1166: distribution. Note again that the requirement of reality and
1167: positivity of $\r_0(s)$ constrains the phases of $a_i^+-a_i^-$ and
1168: hence the coefficients $\a_k$ of $f_0(x)$.
1169: 
1170: \subsubsection{The saddle point approximation for the partition function}
1171: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1172: Recall that our goal is to find a relation between the planar
1173: limit ($t=g_s\hN$ fixed, $g_s\rightarrow0$) of the free energy of
1174: the matrix model and the period integrals of $y(x)\d x$ on the
1175: corresponding Riemann curve. Since the standard planar free energy
1176: $F_0(t)$ depends on $t$ only it cannot appear in relations like
1177: (\ref{SG}), and one has to introduce a set of sources $J_i$ to
1178: have a free energy that depends on more variables. In this
1179: subsection we evaluate this source dependent free energy and its
1180: Legendre transform $\mathcal{F}_0(t,S)$ in the planar limit using
1181: a saddle point expansion.
1182: 
1183: We start by coupling sources to the filling
1184: fractions,\footnote{Note that $\exp\left(-{\hN^2\over
1185: t}\sum_{i=1}^{n-1}J_i\s_i(s_m)\right)$ looks like a macroscopic
1186: operator that changes the equations of motion. However, because of
1187: the special properties of $\s_i(s_m)$ we have
1188: ${\partial\over\partial s_n}\s_i(s_m)={1\over\hN}\int_{\partial
1189: D_i}{\d x\over2\pi i}{\dot\l(s_n)\over(x-\l(s_n))^2}$. In
1190: particular, for the path $\tilde \g_{kl}$ that will be chosen
1191: momentarily and the corresponding domains $D_i$ the eigenvalues
1192: $\l_m$ cannot lie on $\partial D_i$. Hence,
1193: ${\partial\over\partial s_n}\s_i(s_m)=0$ and the equations of
1194: motion remain unchanged.}
1195: \begin{eqnarray}\label{partsources}
1196: Z(g_s,\hN,J)&:=&{1\over{\hN!}}\int_{\g}\d \l_1\ldots\int_{\g}\d
1197: \l_{\hat{N}}\exp\left(-\hN^2S(g_s,\hN;\l_m)-{\hN\over
1198: g_s}\sum_{i=1}^{n-1}J_i\tilde\s_i(\l_m)\right)\nonumber\\
1199: &=&\exp\left(-F(g_s,\hN,J)\right)\ .
1200: \end{eqnarray}
1201: where $J:=\{J_1,\ldots,J_{n-1}\}$. Note that because of the
1202: constraint $\sum_{i=1}^{n}\tilde\s_i(\l_m)=1$,
1203: $\tilde\s_{n}(\l_m)$ is not an independent quantity and we can
1204: have only $n-1$ sources. This differs from the treatment in
1205: \cite{La03} and has important consequences, as we will see in the
1206: next section. We want to evaluate this partition function for
1207: $\hN\rightarrow\infty$, $t=g_s\hN$ fixed, from a saddle point
1208: approximation. We therefore use the path $\tilde\g_{kl}$ from
1209: (\ref{tildegamma}) that was chosen in such a way that the equation
1210: of motion (\ref{eommatrix}) has solutions $s_m^*$ and, for large
1211: $\hN$, $\mathcal{C}_i\subset\g_{p_ip_{i+1}}$. It is only then that
1212: the saddle point expansion converges and makes sense. Obviously
1213: then each integral $\int_{\g}\d\l_m$ splits into a sum
1214: $\sum_{i=1}^{n}\int_{\g_{p_ip_{i+1}}}\d\l_m$. Let
1215: $s^{(i)}\in\mathbb{R}$ be the length coordinate on
1216: $\g_{p_{i}p_{i+1}}$, so that $s^{(i)}$ runs over all of
1217: $\mathbb{R}$. Furthermore, $\tilde\s_i(\l_m)$ only depends on the
1218: number $\hN_i$ of eigenvalues in $\tilde\g_{kl}\cap
1219: D_i=\g_{p_ip_{i+1}}$. Then the partition function
1220: (\ref{partsources}) is a sum of contributions with fixed $\hN_i$
1221: and we rewrite is as
1222: \begin{equation}\label{Z(J)}
1223: Z(g_s,N,J)=\sum_{\stackrel{\sum_i\hN_i=\hN}{\hN_1,\ldots,\hN_{n}}}Z(g_s,\hN,\hN_i)e^{-{1\over
1224: g_s}\sum_{i=1}^{n-1}J_i\hN_i}\ ,
1225: \end{equation}
1226: where now
1227: \begin{eqnarray}
1228: &&\hspace{-0,8cm}Z(g_s,\hN,\hN_i)=\nonumber\\
1229: &&\hspace{-0,8cm}={1\over
1230: \hN_1\ldots\hN_n!}\int_{\g_{p_1p_2}}\d\l^{(1)}_1\ldots
1231: \int_{\g_{p_1p_2}}\d\l^{(1)}_{\hN_1}\ldots\int_{\g_{p_np_{n+1}}}
1232: \d\l^{(n)}_1\ldots\int_{\g_{p_np_{n+1}}}\d\l^{(n)}_{\hN_n}\exp\left(-{t^2\over
1233: g_s^2}S(g_s,t;\l_k^{(i)})\right)\nonumber\\
1234: &&\hspace{-0,8cm}=:\exp\left(\tilde{\mathcal{F}}(g_s,t,\hN_i)\right)
1235: \end{eqnarray}
1236: is the partition function with the additional constraint that
1237: precisely $\hN_i$ eigenvalues lie on $\g_{p_ip_{i+1}}$. Note that
1238: it depends on $g_s,t=g_s\hN$ and $\hN_1,\ldots\hN_{n-1}$ only, as
1239: $\sum_{i=1}^n\hN_i=\hN$. Now that these numbers have been fixed,
1240: there is precisely one solution to the equations of motion, i.e. a
1241: {\it unique} saddle-point configuration, up to permutations of the
1242: eigenvalues, on {\it each} $\g_{p_ip_{i+1}}$. These permutations
1243: just generate a factor $\prod_i \hN_i!$ which cancels the
1244: corresponding factor in front of the integral. As discussed above,
1245: it is important that we have chosen the $\g_{p_ip_{i+1}}$ to
1246: support this saddle point configuration close to the critical
1247: point $\m_i$ of $W$. Moreover, since $\g_{p_ip_{i+1}}$ runs from
1248: one convergence sector to another and by (\ref{condition}) the
1249: saddle point really is dominant (stable), the ``one-loop" and
1250: other higher order contributions are indeed subleading as
1251: $g_s\rightarrow 0$ with $t=g_s\hN$ fixed. This is why we had to be
1252: so careful about the choice of our path $\g$ as being composed of
1253: $n$ pieces $\g_{p_ip_{i+1}}$. In the planar limit
1254: $\n_i:={\hN_i\over\hN}$ is finite, and
1255: $\tilde{\mathcal{F}}(g_s,t,\n_i)={1\over
1256: g_s^2}\tilde{\mathcal{F}}_{0}(t,\n_i)+\ldots$. The saddle point
1257: approximation gives
1258: \begin{equation}\label{F0Seff}
1259: \tilde{\mathcal{F}}_{0}(t,\n_i)= -t^2
1260: S_{eff}(g_s=0,t;s_a^{(j)*}(\n_i))\ ,
1261: \end{equation}
1262: where (cf. (\ref{effaction})) $S_{eff}(g_s=0,t;s_a^{(j)*}(\nu_i))$
1263: is meant to be the value of $S(g_s=0,t;\l(s_a^{(j)*}(\nu_i)))$,
1264: with $\l(s_a^{(j)*}(\nu_i))$ the point on $\g_{p_ip_{i+1}}$
1265: corresponding to the unique saddle point value $s_a^{(j)*}$ with
1266: fixed fraction $\nu_i$ of eigenvalues $\l_m$ in $D_i$. Note that
1267: the ${1\over\hN^2}\sum\dot\l(s_m)$-term in (\ref{effaction})
1268: disappears in the present planar limit. Furthermore, to evaluate
1269: the ``one-loop" term one has to compute the logarithm of the
1270: determinant of an $\hN\times\hN$ matrix which gives a contribution
1271: to $\mathcal{F}(g_s,t,\n_i)$ of order $\hN\sim g_s^{-1}$, as well
1272: as an irrelevant constant $c(\hN)=-{\hN\over2}\log\hN$. The latter
1273: can be absorbed in the overall normalisation of $Z$.
1274: 
1275: It remains to sum over the $\hN_i$ in (\ref{Z(J)}). In the planar
1276: limit these sums are replaced by integrals:
1277: \begin{eqnarray}
1278: Z(g_s,t,J) &=&\int_0^1\d\n_1\ldots\int_0^1\d
1279: \n_{n}\ \delta\left(\sum_{i=1}^{n}\n_i-1\right)\nonumber\\
1280: &&\exp\left[-{1\over
1281: g_s^2}\left(t\sum_{i=1}^{n-1}J_i\n_i-\tilde{\mathcal{F}}_0(t,\n_i)\right)+
1282: c(\hN)+o(g_s^{-1})\right]\ .
1283: \end{eqnarray}
1284: 
1285: Once again, in the planar limit, this integral can be evaluated
1286: using the saddle point technique and for the source-dependent free
1287: energy $F(g_s,t,J)={1\over g_s^2} F_0(t,J)+\ldots$ we find
1288: \begin{equation}\label{F0}
1289: F_0(t,J)=\sum_{i=1}^{n-1}J_i\
1290: t\n_i^*-\tilde{\mathcal{F}}_0(t,\n_i^*)\ ,
1291: \end{equation}
1292: where $\n_i^*$ solves the new saddle point equation,
1293: \begin{equation}
1294: tJ_i={\partial\tilde{\mathcal{F}}_0\over\partial\n_i}(t,\n_j)\
1295: .
1296: \end{equation}
1297: This shows that $F_0(t,J)$ is nothing but the Legendre transform
1298: of $\tilde{\mathcal{F}}_0(t,\n_i^*)$ in the $n-1$ latter
1299: variables. If we define
1300: \begin{equation}\label{Snustar}
1301: S_i:=t\n_i^*\ ,\ \mbox{for}\ i=1,\ldots,n-1,
1302: \end{equation}
1303: we have the inverse relation
1304: \begin{equation}\label{SJ}
1305: S_i={\partial F_0\over\partial J_i}(t,J)\ ,
1306: \end{equation}
1307: and with $\mathcal{F}_0(t,S):=\tilde{\mathcal{F}}_0(t,{S_i\over
1308: t})$, where $S:=\{S_1,\ldots,S_{n-1}\}$, one has from (\ref{F0})
1309: \begin{equation}
1310: \mathcal{F}_0(t,S)=\sum_{i=1}^{n-1}J_iS_i-F_0(t,J)\ ,
1311: \end{equation}
1312: where $J_i$ solves (\ref{SJ}). From (\ref{F0Seff}) and the
1313: explicit form of $S_{eff}$, Eq.(\ref{effaction}) with
1314: $\hN\rightarrow\infty$, we deduce that
1315: \begin{equation}\label{mathcalF0}
1316: \mathcal{F}_0(t,S)=t^2\mathcal{P}\int\d s\int\d s'\
1317: \ln(\l(s)-\l(s'))\r_0(s;t,S_i)\r_0(s';t,S_i)-t\int\d s\
1318: W(\l(s))\r_0(s;t,S_i)\ ,
1319: \end{equation}
1320: where $\r_0(s;t,S_i)$ is the eigenvalue density corresponding to
1321: the saddle point configuration $s_a^{(i)*}$ with
1322: ${\hN_i\over\hN}=\n_i$ fixed to be $\n_i^*={S_i\over t}$. Hence it
1323: satisfies
1324: \begin{equation}\label{intrhoC}
1325: t\int_{\g^{-1}(\mathcal{C}_i)}\r_0(s;t,S_j)\d s=S_i\ \ \mbox{for}\
1326: i=1,2,\ldots n-1\ ,
1327: \end{equation}
1328: and obviously
1329: \begin{equation}\label{intrhoCn}
1330: t\int_{\g^{-1}(\mathcal{C}_n)}\r_0(s;t,S_j)\d
1331: s=t-\sum_{i=1}^{n-1}S_i\ .
1332: \end{equation}
1333: Note that the integrals in (\ref{mathcalF0}) are convergent and
1334: $\mathcal{F}_0(t,S)$ is a well-defined function.
1335: 
1336: \subsection{Special Geometry Relations}
1337: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1338: After this rather detailed study of the planar limit of
1339: holomorphic matrix models we now turn to the derivation of the
1340: special geometry relations for the Riemann surface
1341: (\ref{Riemannsurface}) and hence the local Calabi-Yau
1342: (\ref{locCY}). Recall that in the matrix model the $S_i=t\n_i^*$
1343: are real and therefore $\mathcal{F}_0(t,S)$ of Eq.
1344: (\ref{mathcalF0}) is a function of {\it real} variables. This is
1345: reflected by the fact that one can generate only a subset of all
1346: possible Riemann surfaces (\ref{Riemannsurface}) from the planar
1347: limit of the holomorphic matrix model, namely those for which
1348: $\n_i^*={1\over4\pi it}\int_{A^i}\zeta$ is real (recall $\zeta=y\d
1349: x$). We are, however, interested in the special geometry of the
1350: most general surface of the form (\ref{Riemannsurface}), which can
1351: no longer be understood as a surface appearing in the planar limit
1352: of a matrix model. Nevertheless, for any such surface we can apply
1353: the formal construction of $\r_0(s)$, which will in general be
1354: complex. Then one can use this complex ``spectral density" to
1355: calculate the function $\mathcal{F}_0(t,S)$ from
1356: (\ref{mathcalF0}), that now depends on {\it complex} variables.
1357: Although this is {\it not} the planar limit of the free energy of
1358: the matrix model, it will turn out to be the prepotential for the
1359: general hyperelliptic Riemann surface (\ref{Riemannsurface}) and
1360: hence of the local Calabi-Yau manifold (\ref{locCY}).
1361: 
1362: \subsubsection{Rigid special geometry}
1363: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1364: Let us then start from the general hyperelliptic Riemann surface
1365: (\ref{Riemannsurface}) which we view as a two-sheeted cover of the
1366: complex plane (cf. Figs.\ref{ABcycles},\ref{albecycles}), with its
1367:  cuts $\mathcal{C}_i$ between $a_i^-$ and $a_i^+$. We choose
1368: a path $\g$ on the upper sheet with parameterisation $\l(s)$ in
1369: such a way that $\mathcal{C}_i\subset\g$. The complex function
1370: $\r_0(s)$ is determined from (\ref{solution}) and
1371: (\ref{y0omega0}), as described above. We define the complex
1372: quantities
1373: \begin{equation}
1374: S_i:={1\over4\pi
1375: i}\int_{A^i}\zeta=t\int_{\g^{-1}(\mathcal{C}_i)}\r_0(s) \ \ \
1376: \mbox{for}\ i=1,\ldots,n-1\ ,
1377: \end{equation}
1378: and the prepotential $\mathcal{F}_0(t,S)$ as in (\ref{mathcalF0})
1379: (of course, $t$ is $-{1\over4}$ times the leading coefficient of
1380: $f_0$ and it can now be complex as well).
1381: 
1382: Following \cite{La03} one defines the ``principal value of $y_0$"
1383: along the path $\g$ (c.f. (\ref{Pvalue}))
1384: \begin{equation}
1385: y_0^p(s):={1\over2}\lim_{\e\rightarrow0}[y_0(\l(s)+i\e\dot{\l}(s))+y_0(\l(s)-i\e\dot{\l}(s))]\
1386: .
1387: \end{equation}
1388: For points $\l(s)\in\g$ outside $\mathcal{C}:=\cup_i\mathcal{C}_i$
1389: we have $y_0^p(s)=y_0(\l(s))$, while $y_0^p(s)=0$ on
1390: $\mathcal{C}$. With
1391: \begin{equation}
1392: \phi(s):=W(\l(s))-2t\mathcal{P}\int\d
1393: s'\ln(\l(s)-\l(s'))\r_0(s';t,S_i)
1394: \end{equation}
1395: one finds, using (\ref{omegaofrho}), (\ref{y0omega0}) and
1396: (\ref{Pvalue}),
1397: \begin{equation}\label{phiy0}
1398: {d\over ds}\phi(s)=\dot{\l}(s)y_0^p(s)\ .
1399: \end{equation}
1400: The fact that $y_0^p(s)$ vanishes on $\mathcal{C}$ implies
1401: \begin{equation}
1402: \phi(s)=\xi_i:=\mbox{constant on}\ \mathcal{C}_i\ .
1403: \end{equation}
1404: Integrating (\ref{phiy0}) between $\mathcal{C}_i$ and
1405: $\mathcal{C}_{i+1}$ gives
1406: \begin{equation}
1407: \xi_{i+1}-\xi_i=\int_{a^+_i}^{a^-_{i+1}}\d\l \ y_0(\l)={1\over
1408: 2}\int_{\b_i}\d x \ y(x)={1\over 2}\int_{\b_i}\zeta\ .
1409: \end{equation}
1410: From (\ref{mathcalF0}) we find for $i<n$
1411: \begin{equation}\label{xinxii}
1412: {\partial\over\partial S_{i}}\mathcal{F}_0(t,S)=-t\int\d
1413: s{\partial\r_0(s;t,S_j)\over\partial S_i}\phi(s)=\xi_n-\xi_i\ .
1414: \end{equation}
1415: To arrive at the last equality we used that $\r_0(s)\equiv0$ on
1416: the complement of the cuts, while on the cuts $\phi(s)$ is
1417: constant and we can use (\ref{intrhoC}) and (\ref{intrhoCn}).
1418: Then, for $i<n-1$,
1419: \begin{equation}\label{F0S}
1420: {\partial\over\partial
1421: S_{i}}\mathcal{F}_0(t,S)-{\partial\over\partial
1422: S_{i+1}}\mathcal{F}_0(t,S)={1\over 2}\int_{\b_i}\zeta\ .
1423: \end{equation}
1424: For $i=n-1$, on the other hand, we find
1425: \begin{equation}
1426: {\partial\over\partial
1427: S_{n-1}}\mathcal{F}_0(t,S)=\xi_{n}-\xi_{n-1}={1\over
1428: 2}\int_{\b_{n-1}}\zeta\ .
1429: \end{equation}
1430: We change coordinates to
1431: \begin{equation}
1432: \tilde S_i:=\sum_{j=1}^iS_j\ ,\label{tildeS}
1433: \end{equation}
1434: and find the rigid special geometry relations\footnote{See for
1435: example \cite{CRTP97} for a nice and detailed discussion of the
1436: difference between special and rigid special geometry.}
1437: \begin{eqnarray}
1438: \tilde S_i&=&{1\over 4\pi i }\int_{\a_{i}}\zeta\ \label{sgrela},\\
1439: {\partial\over\partial \tilde S_{i}}\mathcal{F}_0(t,\tilde
1440: S)&=&{1\over2}\int_{\b_{i}}\zeta\ .\label{sgrelb}
1441: \end{eqnarray}
1442: for $i=1,\ldots,n-1$. Note that the basis of one-cycles that
1443: appears in these equations is the one shown in
1444: Fig.\ref{albecycles} and differs from the one used in \cite{La03}.
1445: The origin of this difference is the fact that we introduced only
1446: $n-1$ currents $J_i$ in the partition function
1447: (\ref{partsources}).\\
1448: Next we use the same methods to derive the relation between the
1449: integrals of $\zeta$ over the cycles $\hat \a$ and $\hat \b$ and
1450: the planar free energy.
1451: 
1452: \subsubsection{Integrals over relative cycles}
1453: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1454: The first of these integrals encircles all the cuts, and by
1455: deforming the contour one sees that it is given by the residue of
1456: the pole of $\zeta$ at infinity, which is determined by the
1457: leading coefficient of $f_0(x)$:
1458: \begin{equation}
1459: {1\over4\pi i}\int_{\hat\a}\zeta=t\ .\label{SGhat1}
1460: \end{equation}
1461: The cycle $\hat\beta$ starts at infinity of the lower sheet, runs
1462: to the $n$-th cut and from there to infinity on the upper sheet.
1463: The integral of $\zeta$ along $\hat\b$ is divergent, so we
1464: introduce a (real) cut-off $\L_0$ and instead take $\hat\b$ to run
1465: from $\L_0'$ on the lower sheet through the $n$-th cut to $\L_0$
1466: on the upper sheet. We find
1467: \begin{eqnarray}
1468: {1\over2}\int_{\hat\b}\zeta=\int_{a^+_{n}}^{\L_0}
1469: y_0(\l)\d\l&=&\phi(\l^{-1}(\L_0))-\phi(\l^{-1}(a^+_{n}))=\phi(\l^{-1}(\L_0))-\xi_{n}\nonumber\\
1470: &=&W(\L_0)-2t\mathcal{P}\int\d s'\ln(\L_0-\l(s'))\r_0(s';t,\tilde
1471: S_i)-\xi_{n}.
1472: \end{eqnarray}
1473: On the other hand we can calculate
1474: \begin{equation}\label{xin}
1475: {\partial\over\partial t}\mathcal{F}_0(t,\tilde S)=-\int\d s\
1476: \phi(\l(s)){\partial\over\partial t}[t\r_0(s;t,\tilde
1477: S_i)]=-\sum_{i=1}^n\xi_i\int_{\g^{-1}(\mathcal{C}_i)}\d s\
1478: {\partial\over\partial t}[t\r_0(s;t,\tilde S_i)]=-\xi_{n}\ ,
1479: \end{equation}
1480: (where we used (\ref{intrhoC}) and (\ref{intrhoCn})) which leads
1481: to
1482: \begin{eqnarray}
1483: {1\over2}\int_{\hat \b}\zeta&=&{\partial\over\partial
1484: t}\mathcal{F}_0(t,\tilde
1485: S)+W(\L_0)-2t\mathcal{P}\int\d s'\ln(\L_0-\l(s'))\r_0(s';t,\tilde S_i)\nonumber\\
1486: &=&{\partial\over\partial t}\mathcal{F}_0(t,\tilde
1487: S)+W(\L_0)-t\log\L_0^2+o\left({1\over\L_0}\right)\ .\label{SGhat2}
1488: \end{eqnarray}
1489: Together with (\ref{SGhat1}) this looks very similar to the usual
1490: special geometry relation. In fact, the cut-off independent term
1491: is the one one would expect from special geometry. However, the
1492: equation is corrected by cut-off dependent terms. The last terms
1493: vanishes if we take $\L_0$ to infinity but there remain two
1494: divergent terms which we want to interpret in section
1495: \ref{superpot}.\footnote{Of course, one could define a cut-off
1496: dependent function $\mathcal{F}^{\L_0}(t,\tilde
1497: S):=\mathcal{F}_0(t,\tilde S)+tW(\L_0)-{t^2\over2}\log\L_0^2$ for
1498: which one has
1499: ${1\over2}\int_{\hat\b}\zeta={\partial\mathcal{F}^{\L_0}\over\partial
1500: t}+o\left({1\over\L_0}\right)$ similar to \cite{DOV04}. Note,
1501: however, that this is not a standard special geometry relation due
1502: to the presence of the $o\left({1\over\L_0}\right)$-terms.
1503: Furthermore $\mathcal{F}^{\L_0}$ has no interpretation in the
1504: matrix model and is divergent as $\L_0\rightarrow\infty$.}
1505: 
1506: \subsubsection{Homogeneity of the prepotential}
1507: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1508: The prepotential on the moduli space of complex structures of a
1509: {\it compact} Calabi-Yau manifold is a holomorphic function that
1510: is homogeneous of degree two. On the other hand, the structure of
1511: the local Calabi-Yau manifold (\ref{locCY}) is captured by a
1512: Riemann surface and it is well-known that these are related to
1513: rigid special geometry. The prepotential of rigid special
1514: manifolds does not have to be homogeneous (see for example
1515: \cite{CRTP97}), and it is therefore important to explore the
1516: homogeneity structure of $\mathcal{F}_0(t,\tilde S)$. The result
1517: is quite interesting and it can be written in the form
1518: \begin{equation}\label{homogeneity}
1519: \sum_{i=1}^{n-1}\tilde S_i{\partial\mathcal{F}_0\over\partial
1520: \tilde{S_i}}(t,\tilde S_i)+t{\partial\mathcal{F}_0\over\partial
1521: t}(t,\tilde S_i)=2\mathcal{F}_0(t,\tilde S_i)+t\int\d s\
1522: \r_0(s;t,\tilde S_i)W(\l(s))\ .
1523: \end{equation}
1524: To derive this relation we rewrite Eq.(\ref{mathcalF0}) as
1525: \begin{eqnarray}
1526: 2\mathcal{F}_0(t,\tilde S_i)&=&-t\int\d s\ \r_0(s;t,\tilde
1527: S_i)\left[\phi(s)+W(\l(s))\right]\nonumber\\
1528: &=&-t\int\d s\ \r_0(s;t,\tilde
1529: S_i)W(\l(s))+\sum_{i=1}^{n-1}(\xi_n-\xi_i)S_i-t\xi_n\ .
1530: \end{eqnarray}
1531: Furthermore, we have $\sum_{i=1}^{n-1}\tilde
1532: S_i{\partial\mathcal{F}_0\over\partial\tilde S_i}(t,\tilde
1533: S_i)=\sum_{i=1}^{n-1} S_i{\partial\mathcal{F}_0\over\partial
1534: S_i}(t, S_i)=\sum_{i=1}^{n-1}S_i(\xi_n-\xi_i)$, where we used
1535: (\ref{xinxii}). The result then follows from (\ref{xin}).\\
1536: Of course, the prepotential was not expected to be homogeneous,
1537: since already for the simplest example, the conifold,
1538: $\mathcal{F}_0$ is known to be non-homogeneous (see section
1539: \ref{conifold}). However, Eq.(\ref{homogeneity}) shows the precise
1540: way in which the homogeneity relation is modified on the local
1541: Calabi-Yau manifold (\ref{locCY}).
1542: 
1543: \subsubsection{Duality transformations}
1544: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1545: The choice of the basis $\{\a^i,\b_j,\hat\a,\hat\b\}$ for the
1546: (relative) one-cycles on the Riemann surface was particularly
1547: useful in the sense that the integrals over the compact cycles
1548: $\a^i$ and $\b_j$ reproduce the familiar rigid special geometry
1549: relations, whereas new features appear only in the integrals over
1550: $\hat\a$ and $\hat\b$. In particular, we may perform any
1551: symplectic transformation of the compact cycles $\a^i$ and $\b_j$,
1552: $i,j=1,\ldots n-1$, among themselves to obtain a new set of
1553: compact cycles which we call $a^i$ and $b_j$. Such symplectic
1554: transformations can be generated from (i) transformations that do
1555: not mix $a$-type and $b$-type cycles, (ii) transformations
1556: $a^i=\a^i,\ b_i=\b_i+\a^i$ for some $i$ and (iii) transformations
1557: $a^i=\b_i$, $b_i=-\a^i$ for some $i$. (These are analogue to the
1558: trivial, the $T$ and the $S$ modular transformations of a torus.)
1559: For transformations of the first type the prepotential
1560: $\mathcal{F}$ remains unchanged, except that it has to be
1561: expressed in terms of the new variables $s_i$, which are the
1562: integrals of $\zeta$ over the new $a^i$ cycles. Since the
1563: transformation is symplectic, the integrals over the new $b_j$
1564: cycles then automatically are the derivatives
1565: ${\partial\mathcal{F}_0(t,s)\over\partial s_i}$. For
1566: transformations of the second type the new prepotential is given
1567: by $\mathcal{F}_0(t,\tilde S_i)+i\pi\tilde S_i^2$ and for
1568: transformations of the third type the prepotential is a Legendre
1569: transform with respect to $\int_{a^i}\zeta$. In the corresponding
1570: gauge theory the latter transformations realise electric-magnetic
1571: duality. Consider e.g. a symplectic transformation that exchanges
1572: all compact $\a^i$-cycles with all compact $\b_i$ cycles:
1573: \begin{equation}
1574: \left(\begin{array}{c}\a^i\\\b_i\end{array}\right)\rightarrow\left(\begin{array}{c}a^i\\
1575: b_i\end{array}\right)=\left(\begin{array}{c}\b_i\\-\a^i\end{array}\right),
1576: \ \ \ i=1,\ldots n-1 \ .
1577: \end{equation}
1578: Then the new variables are the integrals over the $a_i$-cycles
1579: which are
1580: \begin{equation}
1581: \tilde\pi_i:={1\over2}\int_{\b_i}\zeta={\partial\mathcal{F}_0(t,\tilde
1582: S)\over\partial \tilde S_i}
1583: \end{equation}
1584: and the dual prepotential is given by the Legendre transformation
1585: \begin{equation}
1586: \mathcal{F}_{D}(t,\tilde\pi):=\sum_{i=1}^{n-1}\tilde
1587: S_i\tilde\pi_i-\mathcal{F}_0(t,\tilde S)\ ,
1588: \end{equation}
1589: such that the new special geometry relation is
1590: \begin{equation}
1591: {\partial\mathcal{F}_D(t,\tilde\pi)\over\partial\tilde\pi_i}=\tilde
1592: S_i={1\over4\pi i}\int_{\a^i}\zeta\ .
1593: \end{equation}
1594: Comparing with (\ref{F0}) one finds that
1595: $\mathcal{F}_D(t,\tilde\pi)$ actually coincides with $F_0(t,J)$
1596: where $J_i-J_{i+1}=\tilde\pi_i$ for $i=1,\ldots n-2$ and
1597: $J_{n-1}=\tilde\pi_{n-1}$.
1598: 
1599: Next, let us see what happens if we also include symplectic
1600: transformations involving the relative cycles $\hat\a$ and
1601: $\hat\b$. An example of a transformation of type (i) that does not
1602: mix $\{\a,\hat\a\}$ with $\{\b,\hat\b\}$ cycles is the one from
1603: $\{\a^i,\b_j,\hat\a,\hat\b\}$ to $\{A^i,B_j\}$, c.f.
1604: Figs.\ref{ABcycles},\ref{albecycles}. This corresponds to
1605: \begin{eqnarray}
1606: \bar S_1&:=&\tilde S_1\ ,\nonumber\\
1607: \bar S_i&:=&\tilde S_i-\tilde S_{i-1}\ \ \ \mbox{for}\ \ i=2,\ldots n-1\ ,\label{barS}\\
1608: \bar S_n&:=&t-\tilde S_{n-1}\ ,\nonumber
1609: \end{eqnarray}
1610: so that
1611: \begin{equation}
1612: \bar S_i={1\over 4\pi i}\int_{A^i}\zeta\ .
1613: \end{equation}
1614: The prepotential does not change, except that it has to be
1615: expressed in terms of the $\bar S_i$. One then finds for
1616: $B_i=\sum_{j=i}^{n-1}\b_j+\hat\b$
1617: \begin{equation}
1618: {1\over2}\int_{B_i}\zeta={\partial\mathcal{F}_0(t,\bar
1619: S)\over\partial \bar S_i}+W(\L_0)-\left(\sum_{j=1}^n\bar
1620: S_i\right)\log\L_0^2+o\left({1\over\L_0}\right)\ .
1621: \end{equation}
1622: We see that as soon as one ``mixes" the cycle $\hat\b$ into the
1623: set $\{\b_i\}$ one obtains a number of relative cycles $B_i$ for
1624: which the special geometry relations are corrected by cut-off
1625: dependent terms. An example of transformation of type (iii) is
1626: $\hat\a\rightarrow\hat\b$, $\hat\b\rightarrow-\hat\a$. Instead of
1627: $t$ one then uses
1628: \begin{equation}
1629: \hat\pi:={\partial\mathcal{F}_0(t,\tilde S)\over\partial
1630: t}=\lim_{\L_0\rightarrow\infty}\left[{1\over2}\int_{\hat\b}\zeta-W(\L_0)+t\log\L_0^2\right]
1631: \end{equation}
1632: as independent variable and the Legendre transformed prepotential
1633: is
1634: \begin{equation}
1635: \hat{\mathcal{F}}(\hat\pi,\tilde
1636: S):=t\hat\pi-\mathcal{F}_0(t,\tilde S)\ ,
1637: \end{equation}
1638: so that now
1639: \begin{equation}
1640: {\partial\hat{\mathcal{F}}(\hat\pi,\tilde
1641: S)\over\partial\hat\pi}=t={1\over4\pi i}\int_{\hat\a}\zeta\ .
1642: \end{equation}
1643: Note that the prepotential is well-defined and independent of the
1644: cut-off in all cases (in contrast to the treatment in
1645: \cite{DOV04}). The finiteness of $\hat{\mathcal{F}}$ is due to
1646: $\hat\pi$ being the {\it corrected, finite} integral over the
1647: relative $\hat\b$-cycle.\\
1648: Note also that if one exchanges all coordinates simultaneously,
1649: i.e. $\a^i\rightarrow\b_i,\
1650: \hat\a\rightarrow\hat\b,\b_i\rightarrow-\a^i,\
1651: \hat\b\rightarrow-\hat\a$, one has
1652: \begin{equation}
1653: \hat{\mathcal{F}}_D(\hat\pi,\tilde\pi):=t\hat\pi+\sum_{i=1}^{n-1}\tilde
1654: S_i\tilde\pi_i-\mathcal{F}_0(t,\tilde S_i)\ .
1655: \end{equation}
1656: Using the generalised homogeneity relation (\ref{homogeneity})
1657: this can be rewritten as
1658: \begin{equation}
1659: \hat{\mathcal{F}}_D(\hat\pi,\tilde\pi)=\mathcal{F}_0(t,\tilde
1660: S_i)+t\int\d s\ \r_0(s;t,\tilde S_i)W(\l(s))\ .
1661: \end{equation}
1662: 
1663: It would be quite interesting to understand the results of this
1664: chapter concerning the parameter spaces of local Calabi-Yau
1665: manifolds in a more geometrical way in the context of (rigid)
1666: special K\"ahler manifolds along the lines of \cite{CRTP97}.
1667: 
1668: 
1669: \section{The superpotential}\label{superpot}
1670: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1671: \setcounter{equation}{0} Adding a background three-form flux to
1672: type IIB strings on a local Calabi-Yau manifold generates an
1673: effective superpotential and breaks the $\mathcal{N}=2$
1674: supersymmetry of the effective action to $\mathcal{N}=1$. Starting
1675: from the usual formula for the effective superpotential
1676: \cite{CIV01} and performing a change of basis one arrives at
1677: \begin{equation}\label{Weff}
1678: W_{eff}=\sum_{i=1}^{n-1}\left(\int_{\G_{\a^i}}H\int_{\G_{\b_i}}\O-\int_{\G_{\b_i}}
1679: H\int_{\G_{\a^i}}\O\right)+\left(\int_{\G_{\hat\a}}H\int_{\G_{\hat\b}}\O-\int_{\G_{\hat\b}}
1680: H\int_{\G_{\hat\a}}\O\right)\ .
1681: \end{equation}
1682: As explained before, the integrals over three-cycles reduce to
1683: integrals over the one-cycles in $H_1(\S,\{Q,Q'\})$ on the Riemann
1684: surface $\S$. But this implies that the divergent terms in
1685: (\ref{SGhat2}) are quite problematic, as they lead to a divergence
1686: of the superpotential which has to be removed for the potential to
1687: make sense.
1688: 
1689: \subsection{Pairings on Riemann surfaces with marked points}
1690: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1691: To understand the divergence somewhat better we will study the
1692: meromorphic one-form $\zeta$ in more detail. First of all we
1693: observe that the integrals $\int_{\a^i}\zeta$ and
1694: $\int_{\b_j}\zeta$ only depend on the cohomology class $[\zeta]\in
1695: H^1(\hat\S)$, whereas $\int_{\hat\b}\zeta$  (where $\hat\b$
1696: extends between the poles of $\zeta$, i.e. from $\infty'$ on the
1697: lower sheet, corresponding to $Q'$, to $\infty$ on the upper
1698: sheet, corresponding to $Q$,) is not only divergent, it also
1699: depends on the representative of the cohomology class, since for
1700: $\tilde\zeta= \zeta+\d\rho$ one has
1701: $\int_{\hat\b}\tilde\zeta=\int_{\hat\b}\zeta+
1702: \int_{\partial\hat\b}\r\left(\neq\int_{\hat\b}\zeta\right)$. Note
1703: that the integral would be independent of the choice of the
1704: representative if we constrained $\r$ to be zero at
1705: $\partial\hat\b$. But as we marked $Q,Q'$ on the Riemann surface,
1706: $\r$ is allowed to take finite or even infinite values at these
1707: points and therefore the integrals differ in general.
1708: 
1709: The origin of this complication is, of course, that our cycles are
1710: elements of the {\it relative} homology group $H_1(\S,\{Q,Q'\})$.
1711: Then, their is a natural map
1712: $\langle.,.\rangle:H_1(\S,\{Q,Q'\})\times
1713: H^1(\S,\{Q,Q'\})\rightarrow \mathbb{C}$. $H^1(\S,\{Q,Q'\})$ is the
1714: relative cohomology group dual to $H_1(\S,\{Q,Q'\})$. In general,
1715: on a manifold $M$ with submanifold $N$, elements of relative
1716: cohomology can be defined as follows (see for example
1717: \cite{KL87}). Let $\O^k(M,N)$ be the set of $k$-forms on $M$ that
1718: vanish on $N$. Then $H^k(M,N):=Z^k(M,N)/B^k(M,N)$, where
1719: $Z^k(M,N):=\{\o\in\O^k(M,N):\d\o=0\}$ and
1720: $B^k(M,N):=\d\O^{k-1}(M,N)$. For $[\hat\G]\in H_k(M,N)$ and
1721: $[\eta]\in H^k(M,N)$ the pairing is defined as
1722: \begin{equation}
1723: \langle\hat\G,\eta\rangle:=\int_{\hat \G}\eta\ .
1724: \end{equation}
1725: This does not depend on the representative of the classes, since
1726: the forms are constraint to vanish on $N$.\\
1727: Now consider $\xi\in\O^k(M)$ such that $i^*\xi=\d\phi$, where $i:N\rightarrow M$
1728: is the inclusion mapping. Note that $\xi$ is not
1729: a representative of an element of relative cohomology, as it does not vanish on
1730: $N$. However, there is another representative in its cohomology class
1731: $[\xi]\in H^k(M)$, namely $\xi_{\phi}=\xi-\d\phi$ which now is also a
1732: representative of
1733: $H^k(M,N)$. For elements $\xi$ with this property we can extend the definition
1734: of the pairing to
1735: \begin{equation}
1736: \langle \hat\G,\xi\rangle:=\int_{\hat \G}\left(\xi-\d\phi\right)\
1737: .
1738: \end{equation}
1739: 
1740: Clearly, the one-form $\zeta=y\d x$ on $\hat\S$ is not a
1741: representative of an element of $H^1(\S,\{Q,Q'\})$. According to
1742: the previous discussion, one might try to find
1743: $\zeta_{\varphi}=\zeta-\d\varphi$ where $\varphi$ is chosen in
1744: such a way that $\zeta_{\varphi}$ vanishes at $Q,Q'$, so that in
1745: particular $\int_{\hat\b}\zeta_{\varphi}=\mbox{finite}$. In other
1746: words we would like to find a representative of $[\zeta]\in
1747: H^1(\hat\S)$ which is also a representative of $H^1(\S,\{Q,Q'\})$.
1748: Unfortunately, this is not possible, because of the logarithmic
1749: divergence, i.e. the simple poles at $Q,Q'$, which cannot be
1750: removed by an exact form. The next best thing we can do instead is
1751: to determine $\varphi$ by the requirement that
1752: $\zeta_{\varphi}=\zeta-\d\varphi$ only has simple poles at $Q,Q'$.
1753: Then we define the pairing
1754: \begin{equation}\label{pair}
1755: \left\langle\hat\b,\zeta\right\rangle:=\int_{\hat\b}\left(\zeta-\d\varphi\right)=
1756: \int_{\hat\b}\zeta_{\varphi}\ ,
1757: \end{equation}
1758: which diverges only logarithmically. To regulate this divergence
1759: we introduce a cut-off as before, i.e. we take $\hat\b$ to run
1760: from $\L_0'$ to $\L_0$. We will have more to say about this
1761: logarithmic divergence in the next section. So although
1762: $\zeta_{\varphi}$ is not a representative of a class in
1763: $H^1(\S,\{Q,Q'\})$ it is as close as we can get.
1764: 
1765: We now want to determine $\varphi$ explicitly. To keep track of
1766: the poles and zeros of the various terms it is useful to apply the
1767: theory of divisors, as explained e.g. in \cite{FK}. Let
1768: $P_1,\ldots P_{2\gh+2}$ denote those points on the Riemann surface
1769: of genus $\gh=n-1$ that correspond to the zeros of $y$ (i.e. to
1770: the $a_i^\pm$). Close to $a_i^\pm$ the good coordinates are
1771: $z_i^\pm=\sqrt{x-a_i^\pm}$. This shows that the divisor of $y$ is
1772: ${P_1\dots P_{2\gh+2}\over Q^{\gh+1}Q'^{\gh+1}}$, which simply
1773: states that $y$ has simple zeros at $P_1,\ldots P_{2\gh+2}$ and
1774: poles of order up to $\gh+1$ at $Q$ and $Q'$. Let $R,R'$ be the
1775: points on the Riemann surface that correspond to zero on the upper
1776: and lower sheet, respectively, then the divisor of $x$ is given by
1777: ${RR'\over QQ'}$. Finally the divisor of $\d x$ is ${P_1\dots
1778: P_{2\gh+2}\over Q^2Q'^2}$, since close to $a_i^\pm$ one has $\d
1779: x\sim z_i^\pm\d z_i^\pm$, and obviously $\d x$ has double poles at
1780: $Q$ and $Q'$. To study the leading poles and zeros of combinations
1781: of these quantities one simply has to multiply the corresponding
1782: divisors. In particular, the divisor of ${\d x\over y}$ is
1783: $Q^{\gh-1}Q'^{\gh-1}$ and for $\gh\geq 1$ it has no poles. Also,
1784: the divisor of $\zeta=y^2{\d x\over y}$ is ${P_1^2\ldots
1785: P_{2\gh+2}^2\over Q^{\gh+3}Q'^{\gh+3}}$ showing that $\zeta$ has
1786: poles of order
1787: $\gh+3,\gh+2,\ldots 2,1$ at $Q$ and $Q'$.\\
1788: Consider now $\varphi_k:={x^k\over y}$ with $\d\varphi_k={{k
1789: x^{k-1}\d x}\over y}-{x^k {y^2}'\d x\over2 y^3}$. For $x$ close to
1790: $Q$ or $Q'$ the leading term of this expression is
1791: $\pm(k-\gh-1)x^{k-\gh-2}\d x.$ This has no pole at $Q,Q'$ for
1792: $k\leq \gh$, and for $k=\gh+1$ the coefficient vanishes, so that
1793: we do not get simple poles at $Q,Q'$. This is as expected as
1794: $\d\varphi_k$ is exact and cannot contain poles of first order.
1795: For $k\geq\gh+2=n+1$ the leading term has a pole of order $k-\gh$
1796: and so $\d\varphi_k$ contains poles of order
1797: $k-\gh,k-\gh-1,\ldots2$ at $Q,Q'$. Note also that at $P_1,\ldots
1798: P_{2\gh+2}$ one has double poles for all $k$ (unless a zero of $y$
1799: occurs at $x=0$). Next, we set
1800: \begin{equation}
1801: \varphi={\mathcal{P}\over y} ,
1802: \end{equation}
1803: with $\mathcal{P} $ a polynomial of order $2\gh+3$. Then $\d
1804: \varphi$ has poles of order $\gh+3,\gh+2,\ldots 2$ at $Q,Q'$, and
1805: double poles at the zeros of $y$ (unless a zero of $\mathcal{P}_k$
1806: coincides with one of the zeros of $y$). From the previous
1807: discussion it is clear that we can choose the coefficients in
1808: $\mathcal{P}$ such that $\zeta_{\varphi}=\zeta-\d\varphi$ only has
1809: a simple pole at $Q,Q'$ and double poles at $P_1,P_2,\ldots
1810: P_{2\gh+2}$. Actually, the coefficients of the monomials $x^k$ in
1811: $\mathcal{P}$ with $k\leq\gh$ are not fixed by this requirement.
1812: Only the $\gh+2$ highest coefficients will be determined, in
1813: agreement with the fact that we cancel the $\gh+2$ poles of order
1814: $\gh+3,\ldots2$.
1815: 
1816: It remains to determine the polynomial $\mathcal{P}$ explicitly. The part of $\zeta$ contributing to the poles of order $\geq2$ at $Q,Q'$ is easily seen to be $\pm W'(x)\d x$ and we obtain the condition
1817: \begin{equation}
1818: W'(x)-\left(\mathcal{P}(x)\over\sqrt{W'(x)^2+f(x)}\right)'=
1819: o\left(1\over x^2\right)\ .
1820: \end{equation}
1821: Integrating this equation, multiplying by the square root and developing the square root leads to
1822: \begin{equation}
1823: W(x)W'(x)-{2t\over {n+1}}x^{n}-\mathcal{P}(x)=cx^{n}+
1824: o\left(x^{n-1}\right)\ ,
1825: \end{equation}
1826: where $c$ is an integration constant.
1827: We read off
1828: \begin{equation}\label{varphi}
1829: \varphi(x)={W(x)W'(x)-\left({2t\over {n+1}}+c\right)x^{n}+
1830: o\left(x^{n-1}\right)\over y}\ ,
1831: \end{equation}
1832: and in particular, for $x$ close to infinity on the upper or lower sheet,
1833: \begin{equation}
1834: \varphi(x)\sim\pm\left[W(x)-c+
1835: o\left(1\over x\right)\right]\ .\label{philarge}
1836: \end{equation}
1837: The arbitrariness in the choice of $c$ has to do with the fact
1838: that the constant $W(0)$ does not appear in the description of the
1839: Riemann surface. In the sequel we will choose $c=0$, such that the
1840: full $W(x)$ appears in (\ref{philarge}). As is clear from our
1841: construction, and is easily verified explicitly, close to $Q,Q'$
1842: one has $\zeta_{\varphi}\sim\left(\mp{2t\over x}+o\left({1\over
1843: x^2}\right)\right)\d x$.
1844: 
1845: With this $\varphi$ we find
1846: \begin{equation}\label{intzetaphi}
1847: \int_{\hat\b}\zeta_{\varphi}=\int_{\hat\b}\zeta-\int_{\hat\b}\d
1848: \varphi=
1849: \int_{\hat\b}\zeta-\varphi(\L_0)+\varphi(\L_0')=\int_{\hat\b}\zeta-2\left(W(\L_0)+o\left(1\over\L_0\right)\right)\
1850: .
1851: \end{equation}
1852: Note that, contrary to $\zeta$, $\zeta_{\varphi}$ has poles at the
1853: zeros of $y$, but these are double poles and it does not matter
1854: how the cycle is chosen with respect to the location of these
1855: poles (as long as it does not go right through the poles). Note
1856: also that we do not need to evaluate the integral of
1857: $\zeta_{\varphi}$ explicitly. Rather one can use the known result
1858: (\ref{SGhat2}) for the integral of $\zeta$ to find from
1859: (\ref{intzetaphi})
1860: \begin{equation}\label{resultpairing}
1861: {1\over2}\left\langle\hat\b,\zeta\right\rangle={1\over2}\int_{\hat\b}\zeta_{\varphi}={\partial\over\partial
1862: t}\mathcal{F}_0(t,\tilde
1863: S)-t\log\L_0^2+o\left({1\over\L_0}\right)\ .
1864: \end{equation}
1865: 
1866: Finally, let us comment on the independence of the representative
1867: of the class $[\zeta]\in H^1(\hat\S)$. Suppose we had started from
1868: $\tilde\zeta:=\zeta +\d\rho$ instead of $\zeta$. Then determining
1869: $\tilde\varphi$ by the same requirement that
1870: $\tilde\zeta-\d\tilde\varphi$ only has first order poles at $Q$
1871: and $Q'$ would have led to $\tilde\varphi=\varphi+\rho$ (a
1872: possible ambiguity related to the integration constant $c$ again
1873: has to be fixed). Then obviously
1874: \begin{equation}
1875: \left\langle\hat\b,\tilde\zeta\right
1876: \rangle=\int_{\hat\b}\tilde\zeta-\int_{\partial\hat\b}\tilde\varphi=\int_{\hat\b}\zeta-\int_{\partial\hat\b}\varphi=\left\langle\hat\b,\zeta\right\rangle\
1877: ,
1878: \end{equation}
1879: and hence our pairing only depends on the cohomology class
1880: $[\zeta]$.
1881: 
1882: \subsection{The superpotential revisited}
1883: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1884: At last we turn to the effective superpotential $W_{eff}$ of the
1885: low energy gauge theory given by the integrals of the three-forms
1886: $\O$ and $H$ over the three-cycles of the Calabi-Yau manifold
1887: (c.f. Eq.(\ref{Weff})). Following \cite{CIV01} and \cite{DV02} we
1888: define for the integrals of $H$ over the cycles $\G_A$ and $\G_B$:
1889: \begin{equation}
1890: N_i=\int_{\G_{A^i}}H\ \ ,\ \ \tau_i=
1891: \left\langle\G_{B_i},H\right\rangle\ \ \ \ \mbox{for}\ \
1892: i=1,\ldots n\ .
1893: \end{equation}
1894: It follows for the integrals over the cycles $\G_{\a}$ and $\G_{\b}$
1895: \begin{eqnarray}
1896: \tilde N_i=\int_{\G_{\a^i}}H=\sum_{j=1}^iN_j\ \ &,&\ \ \tilde
1897: \tau_i=\int_{\G_{\b_i}}H=\tau_i-\tau_{i+1}\ \ \ \ \mbox{for}\ \
1898: i=1,\ldots n-1\
1899: ,\nonumber\\
1900: N=\sum_{i=1}^nN_i= \int_{\G_{\hat\a}}H\ \ &,&\ \
1901: \tilde\tau_0=\left\langle\G_{\hat\b},H\right\rangle=\tau_n\ .
1902: \end{eqnarray}
1903: For the non-compact cycles, instead of the usual integrals, we use
1904: the pairings of the previous section. On the Calabi-Yau, the
1905: pairings are to be understood e.g. as $\tau_i=-i\pi\left\langle
1906: B_i,h\right\rangle$, where $\int_{S^2}H=-2\pi i h$ and $S^2$ is
1907: the sphere in the fibre of $\G_{B_i}\rightarrow B_i$. Note that
1908: this implies that the $\tau_i$ as well as $\tilde\tau_0$ have (at
1909: most) a logarithmic divergence, whereas the $\tilde\tau_i$ are
1910: finite. We propose that the superpotential should be defined as
1911: \begin{eqnarray}\label{superpotential}
1912: W_{eff}&=&\sum_{i=1}^{n-1}\left(\int_{\G_{\a^i}}H\int_{\G_{\b_i}}\O-\int_{\G_{\b_i}}
1913: H\int_{\G_{\a^i}}\O\right)+\left(\int_{\G_{\hat\a}}H\cdot\left\langle\G_{\hat\b},
1914: \O\right\rangle-\left\langle\G_{\hat\b},H\right\rangle\int_{\G_{\hat\a}}\O\right)\nonumber\\
1915: &=&-i\pi \sum_{i=1}^{n-1}\left(\tilde N_i\int_{\b_i}\zeta-
1916: \tilde\tau_i\int_{\a^i}\zeta\right)-i\pi \left(N
1917: \left\langle\hat\b,\zeta\right\rangle-\tilde\tau_0\int_{\hat\a}\zeta\right)\
1918: .
1919: \end{eqnarray}
1920: This formula is very similar to the one advocated for example in
1921: \cite{LMW02}, but now the pairing (\ref{pair}) is to be used. Note
1922: that Eq.(\ref{superpotential}) is invariant under symplectic
1923: transformations on the basis of (relative) three-cycles on the
1924: local Calabi-Yau manifold, resp. (relative) one-cycles on the
1925: Riemann surface, provided one uses the pairing (\ref{pair}) for
1926: every relative cycle. These include $\a^i\rightarrow\b_i,\
1927: \hat\a\rightarrow\hat\b,\ \b_i\rightarrow-\a^i,\
1928: \hat\b\rightarrow-\hat\a$, which acts as electric-magnetic
1929: duality. Using the special geometry relations (\ref{sgrela}),
1930: (\ref{sgrelb}) for the standard cycles and (\ref{SGhat1}),
1931: (\ref{resultpairing}) for the relative cycles, we obtain
1932: \begin{eqnarray}
1933: -{1\over2\pi i}W_{eff}&=&\sum_{i=1}^{n-1}\tilde
1934: N_i{\partial\over\partial \tilde S_i}\mathcal{F}_0(t,\tilde
1935: S_1,\ldots,\tilde S_{n-1})-2\pi
1936: i\sum_{i=1}^{n-1}\tilde\tau_i\tilde
1937: S_i\nonumber\\&&+N{\partial\over\partial t}\mathcal{F}_0(t,\tilde
1938: S_1,\ldots,\tilde S_{n-1})-\left(N\log \L^2_0+2\pi
1939: i\tilde\tau_0\right)t+o\left({1\over\L_0}\right) \
1940: .\nonumber\\\label{WL0}
1941: \end{eqnarray}
1942: The limit $\L_0\rightarrow \infty$ can now be taken provided
1943: \begin{equation}
1944: N\log\L_0^2+2\pi i\tilde\tau_0=N\log\L^2+2\pi i\tau\label{renorm}
1945: \end{equation}
1946: with finite $\L$ and $\tau$. Indeed, $\tilde\tau_0$ is the only
1947: flux number in (\ref{WL0}) that depends on $\L_0$, and its
1948: divergence is logarithmic because of its definition as a pairing
1949: $\left\langle\hat\b,h\right\rangle$. It is, of course, its
1950: interpretation as the $SU(N)$ gauge coupling constant
1951: $\tilde\tau_0={\t_0\over 2\pi}+{4\pi i\over g_0^2}$ which ensures
1952: the exact cancellation of the $\log\L_0$ -terms.
1953: 
1954: Eq.(\ref{WL0}) can be brought into the form of \cite{DV02} if we
1955: use the coordinates $\bar S_i$, as defined in (\ref{barS}) and
1956: such that $\bar S_i={1\over4\pi i}\int_{A^i}\zeta$ for all
1957: $i=1,\ldots n$. We get
1958: \begin{eqnarray}
1959: -{1\over2\pi i}W_{eff}&=&\sum_{i=1}^{n} N_i{\partial\over\partial
1960: \bar
1961: S_i}\mathcal{F}_0(\bar S)\nonumber\\
1962: &&-\sum_{i=1}^{n-1}\bar S_i\left(2\pi
1963: i\sum_{j=i}^{n-1}\tilde\tau_j+N\log\L^2+2\pi i\tau\right)-\bar
1964: S_{n}(2\pi i\tau+N\log\L^2)\ .\nonumber\\
1965: \end{eqnarray}
1966: Setting
1967: \begin{equation}
1968: 2\pi i\sum_{j=i}^{n-1}\tilde\tau_j+N\log\L^2=N_i\log\L_i^2\ \ \
1969: \mbox{for}\ i\in\{1,\ldots n-1\}
1970: \end{equation}
1971: and $\L_n:=\L$ we arrive at
1972: \begin{equation}\label{DVsuperpotential}
1973: -{1\over2\pi i}W_{eff}=\sum_{i=1}^{n}\left[
1974: N_i{\partial\over\partial \bar S_i}\mathcal{F}_0(\bar S)-\bar
1975: S_i\log\L_i^{2N_i} -2\pi i\bar S_i\tau\right]\ .
1976: \end{equation}
1977: This coincides with the corresponding formula in \cite{DV02}
1978: provided we identify ${\partial\mathcal{F}_0(S)\over\partial \bar
1979: S_i}$ with ${\partial\mathcal{F}_0^{pert}(S)\over\partial \bar
1980: S_i}+\bar S_i\log\bar S_i$. Indeed, $\mathcal{F}_0^{pert}$ was the
1981: perturbative part of the free energy of the matrix model and it
1982: was argued in \cite{DV02} that the $S\log S$ term comes from the
1983: measure. Here instead, $\mathcal{F}_0$ is computed in the exact
1984: planar limit of the matrix model, including perturbative and
1985: non-perturbative terms and therefore the $S_i\log S_i$-terms are
1986: already included.\footnote{The presence of $\bar S_i\log \bar S_i$
1987: in ${\partial\mathcal{F}_0\over\partial\bar S_i}$ and hence in
1988: $\int_{B_i}\zeta$ can be easily proven by monodromy arguments
1989: \cite{CIV01}.}
1990: 
1991: Finally, note that we could have chosen $\hat\b$ to run from a
1992: point $\L_0'=|\L_0|e^{i\t/2}$ on the lower sheet to a point
1993: $\L_0=|\L_0|e^{i\t/2}$ on the upper sheet. Then one would have
1994: obtained an additional term $-it\t$, on the right-hand side of
1995: (\ref{resultpairing}), which would have led to
1996: \begin{equation}
1997: \tau\rightarrow\tau+N{\t\over2\pi}
1998: \end{equation}
1999: in (\ref{DVsuperpotential}), as expected.
2000: 
2001: \subsection{Example: the conifold}\label{conifold}
2002: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2003: Next we want to illustrate our general discussion by looking at
2004: the simplest example, i.e. $n=1$. If we take $W={x^2\over2}$ and
2005: $f(x)=-\m=-4t$, $\mu\in\mathbb{R}^+$, the local Calabi-Yau is
2006: nothing but the deformed conifold,
2007: \begin{equation}
2008: x^2+v^2+w^2+z^2-\m=0\ .
2009: \end{equation}
2010: As $n=1$ the corresponding Riemann surface has genus zero. Then
2011: \begin{equation}
2012: \zeta=y\d x=\left\{\begin{array}{c}\ \ \sqrt{x^2-4t}\ \d x\ \ \
2013: \mbox{on the upper sheet}\\-\sqrt{x^2-4t}\ \d x\ \ \ \mbox{on the
2014: lower sheet}\end{array}\right.\ .
2015: \end{equation}
2016: We have a cut $\mathcal{C}=[-2\sqrt{t},2\sqrt{t}]$ and take
2017: $\l(s)=s$ to run along the real axis. The corresponding $\r_0(s)$
2018: is immediately obtained from (\ref{solution}) and (\ref{y0omega0})
2019: and yields the well-known $\r_0(s)={1\over2\pi t}\sqrt{4t-s^2}$,
2020: for $s\in[-2\sqrt{t},2\sqrt{t}]$ and zero otherwise, and from
2021: (\ref{mathcalF0}) we find the planar free energy
2022: \begin{equation}
2023: \mathcal{F}_0(t)={t^2\over2}\log t-{3\over4}t^2\
2024: .\label{F0conifold}
2025: \end{equation}
2026: Note that $t\int\d s\,\r_0(s)W(\l(s))={t^2\over2}$ and
2027: $\mathcal{F}_0$ satisfies the generalised homogeneity relation
2028: (\ref{homogeneity})
2029: \begin{equation}
2030: t{\partial\mathcal{F}_0\over\partial
2031: t}(t)=2\mathcal{F}_0(t)+{t^2\over2}\ .
2032: \end{equation}
2033: 
2034: Obviously one has $\zeta=-2t{\d x\over
2035: y}+\d\left(xy\over2\right)$, which would correspond to
2036: $\varphi={xy\over2}$. Comparing with (\ref{varphi}) this would
2037: yield $c=t$. The choice $c=0$ instead leads to
2038: $\varphi={xy\over2}+t{x\over y}$ and $\zeta=-2t{\d x\over
2039: y}+4t^2{\d x\over y^3} +\d\varphi$. The first term has a pole at
2040: infinity and leads to the logarithmic divergence, while the second
2041: term has no pole at infinity but second order poles at
2042: $\pm2\sqrt{t}$. One has
2043: \begin{eqnarray}
2044: \int_{\hat\a}\zeta&=&4\pi i t=4\pi i\bar S\\
2045: \int_{\hat\b}\zeta&=&\L_0\sqrt{\L_0^2-4t}-4t\log\left({\L_0\over2\sqrt t}+\sqrt{{\L_0^2\over4t}-1}\right)\\
2046: 2\varphi(\L_0)&=&\L_0\sqrt{\L_0^2-4t}+2t{\L_0\over\sqrt{\L_0^2-4t}}\
2047: .
2048: \end{eqnarray}
2049: Then
2050: \begin{eqnarray}
2051: {1\over2}\left\langle\hat\b,\zeta\right\rangle
2052: &=&t\log\left({4t\over\L_0^2}\right)-2t\log\left(1+\sqrt{{1-{4t\over\L_0^2}}}\right)-t{1\over\sqrt{1-{4t\over\L_0^2}}}\nonumber\\
2053: &=&{\partial\mathcal{F}_0(t)\over \partial
2054: t}-t\log\L_0^2+o\left({1\over\L_0^2}\right)\ ,
2055: \end{eqnarray}
2056: where we used the explicit form of $\mathcal{F}_0(t)$,
2057: (\ref{F0conifold}). Finally, in the present case,
2058: Eq.(\ref{superpotential}) for the superpotential only contains the
2059: relative cycles,
2060: \begin{equation}
2061: W_{eff}=-i\pi
2062: N\left\langle{\hat\b},\zeta\right\rangle+i\pi\tilde\tau_0\int_{\hat\a}\zeta
2063: \end{equation}
2064: or ($\bar S=t$)
2065: \begin{eqnarray}
2066: -{1\over2\pi i}W_{eff}&=&N\left(t\log t-t-t\log\L_0^2\right)-2\pi i\tilde\tau_0t+o\left({1\over\L_0^2}\right)\nonumber\\
2067: &=&\bar S\log\left({\bar S^N\over\L^{2N}}\right)-\bar SN -2\pi
2068: i\tau \bar S+o\left({1\over\L_0^2}\right)\ .
2069: \end{eqnarray}
2070: Sending now $\L_0$ to infinity, we get a finite effective
2071: superpotential of Veneziano-Yankielowicz type.\footnote{Of course,
2072: to get the form of \cite{CIV01} the $-2\pi i\tau\bar S$-term can
2073: be absorbed by redefining $\L=\left(\tilde\L
2074: e^{-{2\pi i\tau}\over 3N}\right)^{3/2}$.}\\
2075: 
2076: 
2077: 
2078: 
2079: %%%%%%%%%%%%%%%%%%%%%%%
2080: \section{Conclusions}
2081: %%%%%%%%%%%%%%%%%%%%%%%
2082: \setcounter{equation}{0}
2083: In this note we analysed the special
2084: geometry relations on local Calabi-Yau manifolds of the form
2085: \begin{equation}
2086: W'(x)^2+f_0(x)+v^2+w^2+z^2=0\ .\label{locCYconc}
2087: \end{equation}
2088: The space of compact and non-compact three-cycles on this manifold
2089: maps to the relative homology group $H_1(\S,\{Q,Q'\})$ on a
2090: Riemann surface $\S$, given by $y^2=W'(x)^2+f_0(x)$, with two
2091: marked points $Q,Q'$. We have shown that it is useful to split the
2092: elements of this set into a set of compact cycles $\a^i$ and
2093: $\b_i$ and a set containing the compact cycle $\hat\a$ and the
2094: non-compact cycle $\hat\b$ which together form a symplectic basis.
2095: The corresponding three-cycles on the Calabi-Yau manifold are
2096: $\G_{\a^i},\G_{\b_j},\G_{\hat\a},\G_{\hat\b}$. This choice of
2097: cycles is appropriate since the properties that arise from the
2098: non-compactness of the manifold are then captured entirely by the
2099: integral of the holomorphic three-form $\O$ over the non-compact
2100: three-cycle $\G_{\hat\b}$ which corresponds to $\hat\b$. Indeed,
2101: one finds the following relations
2102: \begin{eqnarray}
2103: -{1\over2\pi i}\int_{\G_{\a^i}}\O&=&2\pi i\tilde S_i\ ,\label{intalpha}\\
2104: -{1\over2\pi i}\int_{\G_{\b_i}}\O&=&{\partial\mathcal{F}_0(t,\tilde S)\over\partial\tilde S_i}\ ,\label{intbeta}\\
2105: -{1\over2\pi i}\int_{\G_{\hat\a}}\O&=&2\pi it\ ,\\
2106: -{1\over2\pi i}\int_{\G_{\hat\b}}\O&=&
2107: {\partial\mathcal{F}_0(t,\tilde S)\over\partial t}+W(\L_0)-t\log
2108: \L_0^2+o\left({1\over\L_0}\right)\ .\label{intbetahat}
2109: \end{eqnarray}
2110: In the last relation the integral is understood to be over the
2111: regulated cycle $\G_{\hat\b}$ which is an $S^2$-fibration over a
2112: line segment running from the $n$-th cut to the cut-off $\L_0$.
2113: Clearly, once the cut-off is removed, the last integral diverges.
2114: To get rid of the polynomial divergence we introduced a pairing on
2115: (\ref{locCYconc}) defined as
2116: \begin{equation}\label{pairCY}
2117: \left\langle\G_{\hat\b},\O\right\rangle:=\int_{\G_{\hat\b}}\left(\O-\d\Phi\right)=
2118: (-i\pi)\int_{\hat\b}\left(\zeta-\d\varphi\right)\ ,
2119: \end{equation}
2120: where
2121: \begin{equation}
2122: \Phi:={W(x)W'(x)-{2t\over{n+1}}x^n\over W'(x)^2+f_0(x)}\cdot{\d
2123: v\w\d w\over 2z}
2124: \end{equation}
2125: is such that
2126: $\int_{\G_{\hat\b}}\d\Phi=-i\pi\int_{\hat\b}\d\varphi$. This
2127: pairing is very similar in structure to the one appearing in the
2128: context of relative (co-)homology and we proposed that one should
2129: use this pairing so that Eq.(\ref{intbetahat}) is replaced by
2130: \begin{equation}
2131: -{1\over2\pi
2132: i}\left\langle\G_{\hat\b},\O\right\rangle={\partial\mathcal{F}_0(t,\tilde
2133: S)\over\partial t}-t\log \L_0^2+o\left({1\over\L_0}\right)\ .
2134: \end{equation}
2135: At any rate, whether one uses this pairing or not, the integral
2136: over the non-compact cycle $\G_{\hat\b}$ is {\it not} just given
2137: by the derivative of the prepotential with respect to $t$.\\
2138: The set of cycles $\{\a^i,\hat\a,\b_i,\hat\b\}$ is particularly
2139: convenient since we can perform arbitrary symplectic (duality)
2140: transformations in $\{\a^i,\b_j\}$ without changing the structure
2141: of the special geometry relations (\ref{intalpha}),
2142: (\ref{intbeta}). However, once we mix $\b_i$- and $\hat\b$-cycles,
2143: more special geometry relations are modified by cut-off dependent
2144: terms.
2145: 
2146: Furthermore, we reconsidered the effective superpotential that
2147: arises if we compactify IIB string theory on (\ref{locCYconc}) in
2148: the presence of a background flux $H$. We emphasize that, although
2149: the commonly used formula $W_{eff}=\int\O\w H$ is very elegant, it
2150: should rather be considered as a mnemonic for
2151: \begin{equation}\label{Wconc}
2152: W_{eff}=\sum_{i=1}^{n-1}\left(\int_{\G_{\a^i}}H\int_{\G_{\b_i}}\O-\int_{\G_{\b_i}}
2153: H\int_{\G_{\a^i}}\O\right)+\left(\int_{\G_{\hat\a}}H\left\langle\G_{\hat\b},
2154: \O\right\rangle-\left\langle\G_{\hat\b},H\right\rangle\int_{\G_{\hat\a}}\O\right)
2155: \end{equation}
2156: because the Riemann bilinear relations do not necessarily hold on
2157: non-compact Calabi-Yau manifolds. We noted that Eq.(\ref{Wconc})
2158: is invariant under symplectic transformations of the basis of the
2159: (relative) 3-cycles, provided one uses the pairing (\ref{pairCY})
2160: whenever a relative cycle appears. Some of these transformations
2161: act as electric-magnetic duality in the $U(1)^n$ gauge theory. By
2162: manipulating (\ref{Wconc}) one obtains both the explicit results
2163: of \cite{CIV01} and the more formal ones of \cite{DV02}. Although
2164: the introduction of the pairing did not render the integrals of
2165: $\O$ and $H$ over the $\G_{\hat\b}$-cycle finite since they are
2166: still logarithmically divergent, these divergences cancel in
2167: (\ref{Wconc}) and the effective superpotential is well-defined.
2168: 
2169: To derive these results we used the holomorphic matrix model as a
2170: technical tool to find the explicit form of the prepotential. On
2171: the way we have clarified several points related to the saddle
2172: point expansion of the holomorphic matrix model. We showed that
2173: although the partition function is independent of the choice of
2174: the path $\g$ appearing in the matrix model, one has to choose a
2175: specific path once one wants to evaluate the free energy from a
2176: saddle point expansion. Since the spectral density $\r_0(s)$ of
2177: the holomorphic matrix model is real by definition we found that
2178: the cuts that form around the critical points of the
2179: superpotential $W$ have specific orientations given by the second
2180: derivatives $W''$ at the critical points. A path $\g$ that is
2181: consistent with the saddle point expansion has then to be chosen
2182: in such a way that all the cuts lie on $\g$. This guarantees that
2183: one expands around a configuration for which the first derivatives
2184: of the effective action indeed vanish. To ensure that saddle
2185: points are really stable we were led to choose $\g$ to consist of
2186: $n$ pieces where each piece contains one cut and runs from
2187: infinity in one convergence domain to infinity of another domain.
2188: Then the ``one-loop" term is a convergent, subleading Gaussian
2189: integral. Using these results for the saddle point expansion of
2190: the matrix model we then determined the free energy of the model
2191: in the planar limit $\mathcal{F}_0(t,\tilde S_i)$. Here the
2192: $\tilde S_i$ fix the fraction of eigenvalues that sit close to the
2193: $i$-th critical point. The Riemann surfaces that appear in the
2194: planar limit of the matrix model only are a subset of the more
2195: general surfaces one obtains from the local Calabi-Yau manifolds,
2196: since the $\tilde S_i$ are manifestly real. We proved the
2197: (modified) special geometry relations in terms of
2198: $\mathcal{F}_0(t,\tilde S_i)$ for these Riemann surfaces. These
2199: relations can then be ``analytically continued" to complex values
2200: of $t$ and $\tilde S_i$, and we used the same
2201: $\mathcal{F}_0(t,\tilde S_i)$ to prove the modified special
2202: geometry relations for the general hyperelliptic Riemann surface
2203: (\ref{Riemannsurface}). One should note, however, that once $t$
2204: and $\tilde S_i$ are taken to be complex, $\mathcal{F}_0(t,\tilde
2205: S_i)$ still is the prepotential but it loses its interpretation as
2206: the planar limit of the free energy of a a matrix model.
2207: 
2208: \vskip 17.mm \noindent {\bf\large Acknowledgements} \vskip 3.mm
2209: 
2210: \noindent Steffen Metzger gratefully acknowledges support by the
2211: Gottlieb Daimler- und Karl Benz-Stiftung as well as by the
2212: Studienstiftung des deutschen Volkes. We would like to thank Jan
2213: Troost and Volodya Kazakov for helpful discussions.
2214: 
2215: \vskip 2.cm
2216: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2217: \begin{thebibliography}{99}
2218: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2219: 
2220: \bibitem{CHSW85}
2221: P. Candelas , G.T. Horowitz, A. Strominger and E. Witten, {\it
2222: Vacuum configurations for superstrings}, Nucl. Phys. {\bf B258},
2223: (1985) 46
2224: 
2225: \bibitem{dWvP}
2226: B. de Wit and A. Van Proeyen, {\it  Potentials and symmetries of
2227: general gauged $\mathcal{N}=2$ supergravity - Yang-Mills models},
2228: Nucl. Phys. {\bf B245} (1984) 89; B. de Wit, P. Lauwers and A. Van
2229: Proeyen, {\it Lagrangians of $\mathcal{N}=2$ supergravity - matter
2230: systems}, Nucl. Phys. {\bf B255} (1985) 569; E. Cremmer, C.
2231: Kounnas, A. Van Proeyen, J.P. Derendinger, S. Ferrara, B. de Wit
2232: and L. Girardello, {\it Vector multiplets coupled to
2233: $\mathcal{N}=2$ supergravity: superhiggs effect, flat potentials
2234: and geometric structure}, Nucl. Phys. {\bf B250} (1985) 385
2235: 
2236: \bibitem{CO90}
2237: P. Candelas and X. de la Ossa, {\it Moduli space of Calabi-Yau
2238: manifolds}, Nucl. Phys. {\bf B355} (1991) 455
2239: 
2240: \bibitem{SW94}
2241: N. Seiberg and E. Witten, {\it Electric-Magnetic Duality, Monopole
2242: Condensation, and Confinement in $\mathcal{N}=2$ Supersymmetric
2243: Yang-Mills Theory}, Nucl. Phys. {\bf B426} (1994) 19,
2244: Erratum-ibid. {\bf B430} (1994) 485, {\tt hep-th/9407087}; {\it
2245: Monopoles, Duality and Chiral Symmetry Breaking in $\mathcal{N}=2$
2246: Supersymmetric QCD}, Nucl. Phys. {\bf B431} (1994) 484, {\tt
2247: hep-th/9408099}
2248: 
2249: \bibitem{KLMVW96}
2250: A. Klemm, W. Lerche, P. Mayr, C. Vafa and N. Warner, {\it
2251: Self-Dual Strings and $\mathcal{N}=2$ Supersymmetric Field
2252: Theory}, Nucl. Phys. {\bf B477} (1996) 746, {\tt hep-th/9604034}
2253: 
2254: \bibitem{KKV96}
2255: S. Katz, A. Klemm and C. Vafa, {\it Geometric Engineering of Quantum Field Theories}, Nucl. Phys. {\bf B497} (1997)
2256: 173, {\tt hep-th/9609239}
2257: 
2258: \bibitem{AKMV03}
2259: M. Aganagic, A. Klemm, M. Mari\~{n}o and C. Vafa, {\it The
2260: Topological Vertex}, Commun. Math. Phys. {\bf 254} (2005) 425,
2261: {\tt hep-th/0305132}
2262: 
2263: \bibitem{ADKMV03}
2264: M. Aganagic, R. Dijkgraaf, A. Klemm, M. Mari\~{n}o and C. Vafa, {\it Topological Strings and Integrable
2265: Hierarchies}, {\tt hep-th/0312085}
2266: 
2267: \bibitem{M04}
2268: M. Mari\~no, {\it Chern-Simons Theory and Topological Strings}, {\tt hep-th/0406005}
2269: 
2270: \bibitem{KKLM99}
2271: S. Kachru, S. Katz, A. Lawrence, J. McGreevy, {\it Open string
2272: instantons and superpotentials}, Phys. Rev. {\bf D62} (2000)
2273: 026001, {\tt hep-th/9912151}
2274: 
2275: \bibitem{CIV01}
2276: F. Cachazo, K.A. Intriligator and C. Vafa, {\it A Large $N$
2277: Duality via a Geometric Transition}, Nucl. Phys. {\bf B603} (2001)
2278: 3, {\tt hep-th/0103067}
2279: 
2280: \bibitem{GV98}
2281: R. Gopakumar and C. Vafa, {\it M-theory and topological
2282: strings-I}, {\tt hep-th/9809187}; {\it M-theory and topological
2283: strings-II}, {\tt hep-th/9812127}; {\it On the gauge
2284: theory/geometry correspondence}, Adv. Theor. Math. Phys. {\bf 3}
2285: (1999) 1415, {\tt hep-th/9811131}
2286: 
2287: \bibitem{DOV04}
2288: U.H. Danielsson, M.E. Olsson and M. Vonk, {\it Matrix models, 4D
2289: black holes and topological strings on non-compact Calabi-Yau
2290: manifolds}, JHEP {\bf 0411} (2004) 007, {\tt hep-th/0410141}
2291: 
2292: \bibitem{V01}
2293: C. Vafa, {\it Superstrings and topological strings at large $N$},
2294: J. Math. Phys. {\bf 42}, (2001) 2798, {\tt hep-th/0008142}
2295: 
2296: \bibitem{DV02}
2297: R. Dijkgraaf and C. Vafa, {\it Matrix Models, Topological Strings, and Supersymmetric Gauge Theories}, Nucl. Phys.
2298: {\bf B644} (2002) 3, {\tt hep-th/0206255}; {\it On Geometry and Matrix Models}, Nucl. Phys. {\bf B644} (2002) 21,
2299: {\tt hep-th/0207106}; {\it A Perturbative Window into Non-Perturbative Physics}, {\tt hep-th/0208048}
2300: 
2301: \bibitem{Wi95}
2302: E. Witten, {\it Chern-Simons Gauge Theory as a String Theory},
2303: Prog. Math. {\bf 133} (1995) 637, {\tt hep-th/9207094}
2304: 
2305: \bibitem{La03}
2306: C.I. Lazaroiu, {\it Holomorphic matrix models}, JHEP {\bf 0305}
2307: (2003) 044, {\tt hep-th/0303008}
2308: 
2309: \bibitem{AGLV}
2310: V.I. Arnold et. al. {\it Singularity Theory}, Springer-Verlag,
2311: Berlin Heidelberg 1998
2312: 
2313: \bibitem{L96}
2314: W. Lerche, {\it Introduction to Seiberg-Witten Theory and its
2315: Stringy Origin}, {\tt hep-th/9611190}
2316: 
2317: \bibitem{K99}
2318: I.K. Kostov, {\it Conformal Field Theory Techniques in Random
2319: Matrix Models}, {\tt hep-th/9907060}
2320: 
2321: \bibitem{K91}
2322: I.R. Klebanov, {\it String theory in two dimensions}, Trieste
2323: Spring School (1991) 30, {\tt hep-th/9108019}
2324: 
2325: \bibitem{CRTP97}
2326: B. Craps, F. Roose, W. Troost and A. Van Proyen, {\it What is
2327: Special K\"ahler Geometry?}, Nucl. Phys. {\bf B503} (1997) 565,
2328: {\tt hep-th/9703082}
2329: 
2330: \bibitem{KL87}
2331: M. Karoubi and C. Leruste, {\it Algebraic Topology via
2332: Differential Geometry}, Cambridge University Press, Cambridge 1987
2333: 
2334: \bibitem{FK}
2335: H.M. Farkas and I. Kra, {\it Riemann Surfaces}, Springer Verlag,
2336: New York 1992
2337: 
2338: \bibitem{LMW02}
2339: W. Lerche, P. Mayr and N. Warner,  {\it Holomorphic N=1 Special
2340: Geometry of Open-Closed Type II Strings}, {\tt hep-th/0207259};
2341: {\it N=1 Special Geometry, Mixed Hodge Variations and Toric
2342: Geometry}, {\tt hep-th/0208039}; W. Lerche, {\it Special Geometry
2343: and Mirror Symmetry for Open String Backgrounds with N=1
2344: Supersymmetry}, {\tt hep-th/0312326}
2345: 
2346: 
2347: 
2348: \end{thebibliography}
2349: \end{document}
2350: