hep-th0612125/ccy.tex
1: \documentclass[12pt]{article}
2: \usepackage{epsfig}
3: %\input epsf
4: \usepackage{amsmath,amssymb,amsthm,amscd}
5: \usepackage{rotating}
6: 
7: % My default margin widths and so on unless overridden in the latex
8: %file
9: \setlength{\oddsidemargin}{0.25in}      % 1.25in left margin
10: \setlength{\evensidemargin}{0.25in}     % 1.25in left margin (even pages)
11: \setlength{\topmargin}{0.5in}           % 1in top margin
12: \setlength{\textwidth}{6.0in}           % 6.0in text - 1.25in rt  margin
13: \setlength{\textheight}{9in}            % Body ht for 1in margins
14: \addtolength{\topmargin}{-\headheight}  % No header, so compensate
15: \addtolength{\topmargin}{-\headsep}     % for header height and  separation
16: \setlength{\marginparwidth}{0.75in}
17: %\setlength{\marginparsep}(0.05 in}
18: % For my home printer
19: % \addtolength{\topmargin}{0.5 in}
20: 
21: 
22: 
23: \let\ex=\times          %\def\ShowPolyData{} %\def\ShowOtherConData{}
24: \def\IP{{\mathbb P}}
25: \def\re{{\rm e}}
26: \def\ZZ{{\mathbb Z}}
27: \def\IR{{\mathbb R}}
28: \def\IC{{\mathbb C}}
29: \def\bT{{\mathbb T}}
30: \def\IQ{{\mathbb Q}}
31: \def\bC{{\mathbb C}}
32: \def\IF{{\mathbb F}}
33: \def\WP{{\mathrm W\mathbb P}}
34: \let\D=\Delta \let\s=\sigma \let\Th=\Theta \let\S=\Sigma
35: \def\2{{1\over2}}
36: \let\<=\langle \let\>=\rangle
37: \def\new#1\endnew{{\bf #1}}
38: \def\ifundefined#1{\expandafter\ifx\csname#1\endcsname\relax}
39: \ifundefined{draftmode}\else    \input draftmode        \fi
40: \let\Msize=\footnotesize        \def\VL{\;\vrule\;}     \def\v{\nu^*}
41: \def\BM{\Msize\begin{matrix}}           \def\EM{\end{matrix}}
42: \def\MN M:#1 #2 N:#3 #4 {{(#1_{#2},#3_{#4})}}
43: \def\MNH M:#1 #2 N:#3 #4 H:#5,#6 [#7]{{(#1_{#2},#3_{#4})^{#5,#6}_{#7}}}
44: \newcommand{\um}{\phantom{-}}
45: \newcommand{\ds}{\displaystyle}
46: \newcommand{\cF}{{\cal F}}
47: \newcommand{\cM}{{\cal M}}
48: \newcommand{\cK}{{\cal K}}
49: \def\dd{\mathrm{d}}
50: \def\cI{\mathcal{I}}
51: \def\sO{\mathscr{O}}
52: 
53: 
54: \DeclareMathOperator{\rank}{rk} \DeclareMathOperator{\ch}{c}
55: \def\CY{Calabi--Yau }
56: \def\Vol{\mathrm{Vol}}
57: \def\conv{\mathrm{conv}}
58: \def\r{\mathrm{r}}
59: \newcommand{\ri}{{\rm i}}
60: \newcommand{\tr}{{\rm Tr}}
61: \newcommand{\zb}{\bar z}
62: \newcommand{\refb}[1]{(\ref{#1})}
63: \newcommand{\p}{\partial}
64: \newcommand{\half}{{1\over 2}}
65: \newcommand{\sectiono}[1]{\section{#1}\setcounter{equation}{0}}
66: \renewcommand{\theequation}{\thesection.\arabic{equation}}
67: 
68: 
69: %%Useful symbols%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
70: \def\ket{\rangle}
71: \def\bra{\langle}
72: \def\CA{{\cal A}}
73: \def\CB{{\cal B}}
74: \def\CC{{\cal C}}
75: \def\CD{{\cal D}}
76: \def\CE{{\cal E}}
77: \def\CF{{\cal F}}
78: \def\CG{{\cal G}}
79: \def\CH{{\cal H}}
80: \def\CI{{\cal I}}
81: \def\CJ{{\cal J}}
82: \def\CK{{\cal K}}
83: \def\CL{{\cal L}}
84: \def\CM{{\cal M}}
85: \def\CN{{\cal N}}
86: \def\CO{{\cal O}}
87: \def\CP{{\cal P}}
88: \def\CQ{{\cal Q}}
89: \def\CR{{\cal R}}
90: \def\CS{{\cal S}}
91: \def\CT{{\cal T}}
92: \def\CU{{\cal U}}
93: \def\CV{{\cal V}}
94: \def\CW{{\cal W}}
95: \def\CX{{\cal X}}
96: \def\CY{{\cal Y}}
97: \def\CZ{{\cal Z}}
98: 
99: %macros
100: \newcommand{\todo}[1]{{\em \small {#1}}\marginpar{$\Longleftarrow$}}
101: \newcommand{\labell}[1]{\label{#1}}
102: \newcommand{\bbibitem}[1]{\bibitem{#1}\marginpar{#1}}
103: \newcommand{\llabel}[1]{\label{#1}\marginpar{#1}}
104: \newcommand{\dslash}[0]{\slash{\hspace{-0.23cm}}\partial}
105: 
106: % macros for the conical defect paper
107: \newcommand{\sphere}[0]{{\rm S}^3}
108: \newcommand{\su}[0]{{\rm SU(2)}}
109: \newcommand{\so}[0]{{\rm SO(4)}}
110: %\newcommand{\sl}[0]{{\rm SL(2,R)}}
111: \newcommand{\bK}[0]{{\bf K}}
112: \newcommand{\bL}[0]{{\bf L}}
113: \newcommand{\bR}[0]{{\bf R}}
114: \newcommand{\tK}[0]{\tilde{K}}
115: \newcommand{\tL}[0]{\bar{L}}
116: \newcommand{\tR}[0]{\tilde{R}}
117: 
118: 
119: \newcommand{\btzm}[0]{BTZ$_{\rm M}$}
120: \newcommand{\ads}[1]{{\rm AdS}_{#1}}
121: \newcommand{\eds}[1]{{\rm EdS}_{#1}}
122: \newcommand{\sph}[1]{{\rm S}^{#1}}
123: \newcommand{\cosm}[0]{R}
124: \newcommand{\hdim}[0]{\bar{h}}
125: \newcommand{\bw}[0]{\bar{w}}
126: \newcommand{\bz}[0]{\bar{z}}
127: \newcommand{\be}{\begin{equation}}
128: \newcommand{\ee}{\end{equation}}
129: \newcommand{\bea}{\begin{eqnarray}}
130: \newcommand{\eea}{\end{eqnarray}}
131: \newcommand{\pat}{\partial}
132: \newcommand{\lp}{\lambda_+}
133: \newcommand{\bx}{ {\bf x}}
134: \newcommand{\bk}{{\bf k}}
135: \newcommand{\bb}{{\bf b}}
136: \newcommand{\BB}{{\bf B}}
137: \newcommand{\tp}{\tilde{\phi}}
138: \hyphenation{Min-kow-ski}
139: 
140: 
141: %%Commonly used constants and symbols%%%%%%%%%%%%%%%%%%%%%%%%%
142: \def\apr{\alpha'}
143: \def\a{\alpha}
144: \def\str{{str}}
145: \def\lstr{\ell_\str}
146: \def\gstr{g_\str}
147: \def\Mstr{M_\str}
148: \def\lpl{\ell_{pl}}
149: \def\Mpl{M_{pl}}
150: \def\varep{\varepsilon}
151: \def\del{\nabla}
152: \def\grad{\nabla}
153: \def\tr{\hbox{tr}}
154: \def\perp{\bot}
155: \def\half{\frac{1}{2}}
156: \def\p{\partial}
157: \def\perp{\bot}
158: \def\eps{\epsilon}
159: 
160: 
161: 
162: \begin{document}
163: \begin{titlepage}
164: {}~ \hfill\vbox{ \hbox{} }\break
165: 
166: 
167: \rightline{hep-th/0612125} \rightline{MAD-TH-06-12} \vskip 1cm
168: 
169: \centerline{\Large \bf Topological string theory on compact
170: Calabi-Yau:} \vskip 0.2 cm \centerline{\Large \bf modularity and
171: boundary conditions} \vskip 0.5 cm
172: 
173: %\centerline{\Large \bf Draft:} \vskip 0.2 cm \centerline{\Large \bf preliminary
174: %and unfinished} \vskip 0.5 cm
175: 
176: \long\def\symbolfootnote[#1]#2{\begingroup%
177: \def\thefootnote{\fnsymbol{footnote}}\footnote[#1]{#2}\endgroup}
178: 
179: %\renewcommand{\thefootnote}{\fnsymbol{footnote}}
180: \vskip 30pt \centerline{ {\large \rm Min-xin Huang
181: \symbolfootnote[1]{minxin@physics.wisc.edu},
182: Albrecht Klemm\symbolfootnote[7]{aklemm@physics.wisc.edu} and
183: Seth Quackenbush\symbolfootnote[2]{squackenbush@wisc.edu} }} \vskip .5cm \vskip
184: 30pt
185: \centerline{$^{\star}$ $^{\star\star}$ $^\dagger${\it Department of Physics, University of
186: Wisconsin}}
187: \centerline{$^{\star\star}${\it Department of Mathematics, University of
188: Wisconsin}}
189: \centerline{\it Madison, WI 53706, U.S.A.}
190: 
191: 
192: \setcounter{footnote}{0}
193: %\renewcommand{\thefootnote}{\arabic{footnote}}
194: \vskip 100pt
195: \begin{abstract}
196: The topological string partition function $Z(\lambda,t,\bar t)$
197: $=\exp(\lambda^{2g-2} F_g(t,\bar t))$ is calculated on a compact Calabi-Yau $M$. 
198: The $F_g(t,\bar t)$ fulfill the holomorphic anomaly
199: equations, which imply that $\Psi=Z$ transforms as a wave function
200: on the symplectic space $H^3(M,\mathbb{Z})$. This defines it
201: everywhere in the moduli space ${\cal M}(M)$ along with preferred
202: local coordinates. Modular properties of the sections $F_g$ as
203: well as local constraints from the 4d effective action allow us to
204: fix $Z$ to a large extent. Currently with a newly found 
205: gap condition at the conifold, regularity at the orbifold and the most naive bounds
206: from Castelnuovo's theory, we can provide the boundary data, which
207: specify $Z$, e.g. up to genus 51 for the quintic.
208: \end{abstract}
209: 
210: \end{titlepage}
211: \vfill \eject
212: 
213: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
214: 
215: \newpage
216: 
217: 
218: \baselineskip=16pt
219: 
220: \tableofcontents
221: 
222: 
223: \section{Outline}
224: 
225: Coupling topological matter to topological gravity is a key
226: problem in string theory. Conceptually most relevant is the
227: topological matter sector of the critical string as it arises e.g. in
228: Calabi-Yau compactifications. Topological string theory on
229: non-compact Calabi-Yau manifolds such as ${\cal O}(-3)\rightarrow
230: \mathbb{P}^2$ is essentially solved either by localization -
231: \cite{Klemm:1999gm} or large N-techniques~\cite{Aganagic:2003db}
232: and has intriguing connections to Chern-Simons
233: theory~\cite{Witten:1992fb}, open-closed string
234: duality~\cite{Gopakumar:1998ki}, matrix
235: models~\cite{Dijkgraaf:2002fc}, integrable hierarchies of
236: non-critical string theory~\cite{Aganagic:2003qj} and 2d
237: Yang-Mills theory~\cite{Vafa:2004qa}.
238: 
239: However, while local Calabi-Yau manifolds are suitable to study
240: gauge theories and more exotic field theories in 4d and specific
241: couplings to gravity, none of the techniques above extends to
242: compact Calabi-Yau spaces, which are relevant for important
243: questions in 4d quantum gravity concerning e.g. the properties of
244: 4d black holes~\cite{Ooguri:2004zv} and the wave function in mini
245: superspace~\cite{Ooguri:2005vr}.
246: 
247: 
248: Moreover, while the genus dependence is encoded in the
249: Chern-Simons and matrix model approaches in a superior fashion by
250: the ${1\over N^2}$-expansion, the moduli dependence on the
251: parameter $t$ is reconstructed locally and in a holomorphic limit,
252: typically by sums over partitions. This yields an algorithm, which
253: grows exponentially in the world-sheet degree or the
254: space-time instanton number.
255: 
256: As the total $F_g(t,\bar t)$ are modular invariant sections over
257: the moduli space ${\cal M}(M)$, they must be generated by a ring
258: of almost holomorphic modular forms.  This solves the dependence
259: on the moduli in the most effective way. In the following we will
260: show that space-time modularity, the holomorphic anomaly equations
261: of Bershadsky, Cecotti, Ooguri and Vafa, as well as boundary
262: conditions at various boundary components of the moduli space,
263: solve the theory very efficiently.
264: 
265: For compact (and non-compact) Calabi-Yau spaces mirror symmetry is
266: proven at genus zero. The modular properties that we need are also
267: established  at genus zero. Moreover it has been argued recently
268: that the  holomorphic anomaly recursions follow from categorical
269: mirror symmetry \cite{costello,kontsevich}. To establish  mirror
270: symmetry at higher genus, one needs merely to prove that the same
271: boundary data fix the $F_g(t,\bar t)$ in the $A$- and the
272: $B$-model.
273: 
274: 
275: \subsection{Extending the Seiberg-Witten approach to gravity}
276: Seiberg-Witten reconstructed the non-perturbative N=2 gauge gauge
277: coupling from meromorphic section over ${\cal M}(M)$ using their
278: modular properties and certain local  data from the effective
279: action at singular divisors of ${\cal M}(M)$. In~\cite{Huang}~we
280: reconsidered the problem of topological string on local Calabi-Yau
281: from the modular point of view and found that the singular
282: behaviour of the gravitational couplings is restrictive enough to
283: reconstruct them globally. This can be viewed as the most straightforward 
284: extension of the  Seiberg-Witten approach to gravitational
285: couplings.
286: 
287: Note that the problem of instanton counting in these cases is
288: solved either by geometric engineering, one of the techniques
289: mentioned above, or more directly by the localisation techniques
290: in the moduli space of gauge theory instantons by Nekrasov,
291: Nakajima et. al. It is nevertheless instructive to outline the
292: general idea in this simple setting\footnote{Maybe the simplest
293: example of the relation between modularity and the holomorphic
294: anomaly equations is provided by Hurwitz theory on elliptic curves
295: \cite{dijkgraafzagier}.}. We focused on the N=2 $SU(2)$
296: Seiberg-Witten case, but the features hold for any local
297: Calabi-Yau whose mirror is an elliptic curve with a meromorphic
298: differential~\cite{Aganagic:2006wq,Huang2}\footnote{With fairly
299: obvious generalizations for the cases where the mirror is a higher
300: genus curve. In this case the traditional modular forms of
301: subgroups $\Gamma$ of $\Gamma_0:=SL(2,\mathbb{Z})$ have to be
302: replaced by Siegel modular forms of subgroups of
303: $Sp(2g,\mathbb{Z})$~\cite{Aganagic:2006wq}.} and are as follows
304: \begin{itemize}
305: \item The genus $g$ topological string partition functions are given by
306: \begin{equation}
307: F^{(g)}(\tau, \bar \tau)=\xi^{2g-2} \sum_{k=0}^{3(g-1)}
308: \hat E_2^{k}(\tau,\bar \tau) c^{(g)}_k(\tau)\ .
309: \label{eq:generallocalform}
310: \end{equation}
311: Here $\hat E_2(\tau,\bar \tau):= E_2(\tau) + {6 i \over \pi(\bar \tau-\tau)}$ is the
312: modular invariant anholomorphic extension of the second Eisenstein series
313: $E_2(\tau)$ and the holomorphic `Yukawa coupling'
314: $\xi:=C^{(0)}_{ttt}={\partial \tau \over \partial t}$ is an
315: object of weight $-3$ under the modular $\Gamma\in \Gamma_0=SL(2,\mathbb{Z})$.
316: For example for pure $N=2$ $SU(2)$ gauge theory $\Gamma=\Gamma(2)$~\cite{Seiberg:1994rs}.
317: Modular invariance implies then that $c^{(g)}_k(\tau)$ are modular forms of
318: $\Gamma$ of weight $6(g-1)-2k$.
319: \item The simple anti-holomorphic dependence of (\ref{eq:generallocalform}) implies
320: that the only part in $F^{(g)}(\tau, \bar \tau)$ not fixed by the recursive anomaly
321: equations is the weight $6(g-1)$ holomorphic forms $c^{(k)}_0(\tau)$, which
322: are finitely generated as a weighted polynomial $c^{(g)}_0(\tau)=p_{6(g-2)}(k_1,\ldots,k_m)$
323: in the holomorphic generators $G_{k_1},\ldots,G_{k_m}$ of forms of $\Gamma$.
324: \item The finite data needed to fix the coefficients in $p_{6(g-2)}(k_1,\ldots,k_m)$
325: are provided in part by the specific leading behaviour of the $F^{(g)}$ at the conifold divisor
326: \begin{eqnarray} \label{thegap}
327: F^{(g)}_{\textrm{conifold}}=\frac{(-1)^{g-1}B_{2g}}{2g(2g-2)t_D^{2g-2}}+\mathcal{O}(t_D^0),
328: \end{eqnarray}
329: in special local coordinates $t_D$. The order of the leading term
330: was established in~\cite{BCOV}, the coefficient of the leading
331: term in \cite{Ghoshal:1995}, and the `gap condition,' i.e. the
332: vanishing of the following $2g-3$ negative powers in $t_D$
333: in~\cite{Huang}. This property in particular carries over to the
334: compact case and we can give indeed a string theoretic explanation
335: of the finding in~\cite{Huang}.
336: \item Further conditions are provided by the regularity of the
337: $F^{(g)}$ at orbifold points in ${\cal M}(M)$. These conditions
338: unfortunately turn out to be somewhat weaker in the global case
339: than in the local case.
340: \end{itemize}
341: 
342: Similar forms as (\ref{eq:generallocalform}) for the $F^{(g)}$ 
343: appear in the context of Hurwitz theory on elliptic 
344: curves~\cite{dijkgraafzagier}, of mirror symmetry in $K3$ 
345: fibre limits\cite{Marino:1998} and on rational 
346: complex surfaces~\cite{Hosono:1999qc,Hosono:2001gf}.
347: In the local cases, which have elliptic curves as mirror geometry,
348: we found~\cite{Huang,Huang2}, that the above conditions
349: (over)determine the unknowns in $p_{6(g-2)}(k_1,\ldots,k_m)$ and
350: solve the theory. This holds also for the gauge theories with
351: matter, which from geometric engineering point of view correspond
352: to local Calabi-Yau manifolds with several (K\"ahler)
353: moduli~\cite{Huang2}. Using the precise anholomorphic
354: dependence and restrictions from space-time modularity one can
355: iterate the holomorphic anomaly equation with an algorithm which
356: is exact in the moduli dependence and grows polynomially in
357: complexity with the genus.
358: 
359: Here we extend this approach further to compact Calabi-Yau spaces
360: and focus on the class of one K\"ahler moduli Calabi-Yau spaces
361: $M$ such as the quintic. More precisely we treat the class of one
362: modulus cases whose mirror $W$ has, parameterized by a suitable
363: single cover variable, a Picard-Fuchs system  with exactly three
364: regular singular points: The point of maximal monodromy, a
365: conifold point, and a point with rational branching. The latter can
366: be simply a $\mathbb{Z}_d$ orbifold point. This e.g. is the case
367: for the hypersurfaces where the string theory has an exact
368: conformal field theory description at this point in terms of an
369: orbifold of a tensor product of minimal $(2,2)$ SCFT field
370: theories, the so called the Gepner-model. For some complete
371: intersections there are massless BPS particles at the branch
372: locus, which lead in addition to logarithmic singularities.
373: 
374: We find a natural family of coordinates in which the conifold
375: expansion as well as the rational branched logarithmic
376: singularities exhibit the gap condition (\ref{thegap}). Despite
377: the fact that the modular group, in this case a subgroup of ${\rm
378: Sp}(h^3,\mathbb{Z})$, is poorly understood\footnote{Subgroups of
379: ${\rm Sp}(4,\mathbb{Z})$ in which the monodromy group of the
380: one-parameter models live have been recently specified
381: \cite{cyy}.}, we will see that the essential feature carry over to
382: the compact case. Modular properties, the ``gap condition'',
383: together with regularity at the orbifold, the leading behaviour
384: of the $F_g$ at large radius, and Castelenovo's Bound 
385: determine topological string on one modulus 
386: Calabi-Yau to a large extent.
387: 
388: 
389: 
390: \section{The topological B-model}
391: %---------------------------------------------------------
392: In this section we give a quick summary of the approach of
393: \cite{BCOVI,BCOV} to the topological B-model, focusing as fast as
394: possible on the key problems that need to be overcome: namely the
395: problem of integrating the anomaly equation efficiently and
396: the problem of fixing the boundary conditions.
397: 
398: 
399: \subsection{The holomorphic anomaly equations}
400: \label{holomorphicanomaly}
401: The definition of
402: $F^{(g)}$ is $F^{(g)}=\int_{ {\cal M}_g} \mu_g$ with measure on ${\cal M}_g$
403: \begin{equation}
404: \mu_g=\prod_{i=1}^{3g-3} {\rm d} m_i {\rm d} {\bar m}_{\bar \imath}
405: \left\langle \prod_{i,\bar \imath}
406: \int_\Sigma G_{zz} \mu^{(i)\, z}_{\bar z} {\rm d}^2 z
407: \int_\Sigma G_{\bar z \bar z} \mu^{(i)\, \bar z}_{z} {\rm d}^2 z \right\rangle \ .
408: \end{equation}
409: Here the Beltrami differentials $\mu^{(i)\, z}_{\bar z} {\rm d}
410: {\bar z}$ span $H^1(\Sigma,T\Sigma)$, the tangent space to ${\cal
411: M}_g$. The construction of the measure $\mu_g$ is strikingly
412: similar to the one for the bosonic string, once the BRST partner
413: of the energy-momentum tensor is identified with the
414: superconformal current $G_{zz}{\rm d } z$ and the ghost number
415: with the $U(1)$ charge \cite{MB}. $\left\langle \right\rangle$ is
416: to be evaluated in the internal $(2,2)$ SCFT, but it is easy to
417: see that it gets only contributions from the topological $(c,c)$
418: sector.
419: 
420: The holomorphic anomaly equation reads for $g=1$  \cite{BCOVI}
421: \begin{equation}
422: {\bar \partial}_{\bar k} \partial_m F^{(1)}={1\over 2}
423: {\bar C}_{\bar k}^{ij} C^{(0)}_{mij} + \left( {\chi\over 24} -1\right )
424: G_{\bar km}\ ,
425: \label{eq:anomalyg1}\
426: \end{equation}
427: where $\chi$ is the Euler number of the target space $M$, and for
428: $g>1$ \cite{BCOV}
429: \begin{equation}
430: {\bar \partial}_{\bar k} F^{(g)}={1\over 2}{\bar C}_{\bar k}^{ij}
431: \left(D_i D_j F^{(g-1)}+ \sum_{r=1}^{g-1}D_i F^{(r)}  D_j F^{(g-r)}\right)\ .
432: \label{eq:anomaly}
433: \end{equation}
434: The right hand side of the equations comes from the complex
435: co-dimension one boundary of the moduli space of the worldsheet
436: ${\cal M}_g$, which corresponds to pinching of handles. The key idea is  that $\bar \partial_{\bar
437: k} F^{(g)}=\int_{{\cal M}_g} \bar \partial \partial \lambda_g$,
438: where $\bar \partial \partial$ are derivatives on ${\cal M}_g$ so
439: that $\bar \partial_{\bar k} F^{(g)}=\int_{\partial {\cal M}_g}
440: \lambda_g$. The contribution to the latter integral is  from the
441: co-dimension one boundary $\partial {\cal M}_g$.
442: 
443: The first equation (\ref{eq:anomalyg1}) can be integrated using
444: special geometry up to a holomorphic function \cite{BCOVI}, which
445: is fixed by the consideration in Sect. \ref{boundaryconditions}.
446: 
447: The equations  (\ref{eq:anomaly}) are solved in BCOV using the
448: fact that due to
449: \begin{equation}
450: \bar D_{\bar i} \bar C_{\bar \jmath \bar k \bar l}=\bar D_{\bar \jmath} \bar C_{\bar
451: \imath \bar k \bar l}
452: \end{equation}
453: one can integrate
454: \begin{equation}
455: \bar C_{\bar \jmath \bar k \bar l}=e^{-2 K} \bar D_{\bar i}\bar D_{\bar \jmath}
456: \bar \partial_{\bar k} S
457: \end{equation}
458: as
459: \begin{equation}
460: S_{\bar \imath} =\bar \partial_{\bar \imath} S,\ \ \
461: S^j_{\bar \imath} =\bar \partial_{\bar \imath} S^j,\ \ \
462: \bar C_{\bar k}^{ij}= \partial_{\bar k} S^{ij}\ .
463: \label{eq:propagators}
464: \end{equation}
465: The idea is to write the right hand side of (\ref{eq:anomaly}) as
466: a derivative w.r.t. $\bar \partial_{\bar k}$. In the first step
467: one writes
468: \begin{equation}
469: \begin{array}{rl}
470: {\bar \partial}_{\bar k} F^{(g)}=&\ds{{\bar \partial}_{\bar k}
471: \left({1\over 2}S^{ij}
472: \left(D_i D_j F^{(g-1)}+ \sum_{r=1}^{g-1}D_i F^{(r)}  D_j F^{(g-r)}\right)\right)} \\[ 3 mm]
473: &\ds{-{1\over 2} S^{ij}{\bar \partial}_{\bar k}\left(D_i D_j F^{(g-1)}+
474: \sum_{r=1}^{g-1}D_i F^{(r)}  D_j F^{(g-r)}\right)\ .}
475: \end{array}
476: \label{int1:anomaly}
477: \end{equation}
478: With the commutator $R^l_{i\bar k j}=-\bar \partial_{\bar k}
479: \Gamma_{ij}^l=[D_i,\partial_{\bar k}]^l_j= G_{i\bar
480: k}\delta^l_j+G_{j\bar k}\delta^l_i-C^{(0)}_{ijm} {\bar C}_{\bar
481: k}^{ml}$ the  second term can be rewritten so that the $\bar
482: \partial_{\bar k}$ derivative acts in all terms directly on
483: $F^{(g)}$. Then using (\ref{eq:anomalyg1}, \ref{eq:anomaly}) with
484: $g'<g$ one can iterate the procedure, which produces an equation
485: of the form
486: \begin{equation}
487: {\bar \partial}_{\bar k} F^{(g)}=\bar \partial_{\bar k}
488: \Gamma^{(g)}(S^{ij},S^i,S, C^{(<g)}_{i_1,\ldots,i_n})\ ,
489: \end{equation}
490: where $\Gamma^{(g)}$ is a functional of  $S^{ij},S^i,S$ and $C^{(<g)}_{i_1,\ldots,i_n}$.
491: This implies that
492: \begin{equation}
493: F^{(g)}=\Gamma^{(g)}(S^{ij},S^i,S, C^{(<g)}_{i_1,\ldots,i_n})+c^{(g)}_0(t)\ ,
494: \label{fgformal}
495: \end{equation}
496: is a solution. Here $c^{(g)}_0(t)$ is the holomorphic ambiguity,
497: which is not fixed by the recursive procedure. It is holomorphic
498: in $t$ as well as modular invariant. The {\sl major conceptual
499: problem} of topological string theory on compact Calabi-Yau is to
500: find the {\sl boundary conditions} which fix $c^{(g)}_0(t)$. Note
501: that the problem is not well defined without the constraints from
502: modular invariance. Using the generalization of the gap condition
503: in Sect. \ref{boundaryconditions}, the behaviour of the orbifold
504: singularities in Sect. \ref{quinticorbifold} and Castelnuovo's
505: bound in Sect. \ref{quinticdbrane} we can achieve this goal to a
506: large extent.
507: 
508: 
509: Properties of the $\Gamma^{(g)}(S^{ij},S^i,S,
510: C^{(<g)}_{i_1,\ldots,i_n})$ are established using the auxiliary
511: action
512: \begin{equation}
513: Z=\int \dd x \dd \phi \exp(Y+\tilde W)
514: \label{auxilliaryaction}
515: \end{equation}
516: where
517: \begin{equation}
518: \begin{array}{rl}
519: \tilde W(\lambda,x,\phi,t,\bar t)=&\ds{\sum_{g=0}^\infty \sum_{m=0}^\infty
520: \sum_{n=0}^\infty {1\over m! n!} \tilde C^{(g)}_{i_1,\ldots,i_n,\phi^{m}}
521: x_{i_1}\ldots x_{i_n} \phi^m}\\
522: =&\ds{\sum_{g=0}^\infty  \sum_{n=0}^\infty
523: {\lambda^{2g -2}\over n!} C^{(g)}_{i_1,\ldots,i_n}x_{i_1}\ldots x_{i_n}(1-\phi)^{2-2g-n}}+\left({\chi\over 24}-1\right)
524: \log\left(1\over {1-\phi}\right) \ ,
525: \end{array}
526: \end{equation}
527: with $C^{(g)}_{i_1,\ldots,i_n}=D_{i_1}\ldots D_{i_n} F^{(g)}$  and
528: the ``kinetic term'' is given by
529: \begin{equation}
530: Y(\lambda , x, \phi;t,\bar t)=-{1\over 2 \lambda^2}
531: (\Delta_{ij}x^i x^j+ 2 \Delta_{i\phi}x^i \phi +\Delta_{\phi\phi}
532: \phi^2) + {1\over 2} \log\left(\det \Delta\over \lambda^2\right)\ .
533: \end{equation}
534: In \cite{BCOV} it was shown that $\exp(\tilde W)$ fulfills an
535: equation
536: \begin{equation}
537: {\partial \over \partial {\bar t_i}} \exp(\tilde W) =\left[ {\lambda ^2\over 2}
538: \bar C^{jk}_{\bar \imath} {\partial^2 \over \partial x^j \partial x^k} -
539: G_{\bar \imath j}  x^j {\partial \over \partial \phi} \right]\exp(\tilde W)\
540: \label{linearanomaly}
541: \end{equation}
542: that is equivalent to the holomorphic anomaly equations, by
543: checking the coefficients of the $\lambda$ powers, and $\exp(Y)$
544: fulfills
545: \begin{equation}
546: {\partial \over \partial {\bar t_{\bar \imath}}} \exp(Y) =\left[ -{\lambda ^2\over 2}
547: \bar C^{jk}_{\bar \imath} {\partial^2 \over \partial x^j \partial x^k} -
548: G_{\bar \imath j}  x^j {\partial \over \partial \phi} \right]\exp(Y) \
549:                 \end{equation}
550: implying that  $\Delta_{ij}$, $\Delta_{i\phi}$ and $\Delta_{\phi\phi}$ are the inverses
551: to the propagators $K^{ij}=-S^{ij}$, $K^{i\phi}:=-S^i$ and  $K^{\phi\phi}:=-2S$. A
552: saddle point expansion of $Z$ gives
553: \begin{equation}
554: \log(Z)=\sum_{g=2}^{\infty} \lambda^{2g-2}\left[F^{(g)}-
555: \Gamma^{(g)}(S^{ij},S^i,S, C^{(<g)}_{i_1,\ldots,i_n})\right]\ ,
556: \end{equation}
557: where  $\Gamma^{(g)}(S^{ij},S^i,S, C^{(<g)}_{i_1,\ldots,i_n})$ is
558: simply the Feynman graph expansion of the action
559: (\ref{auxilliaryaction}) with the vertices $\tilde
560: C^{(g)}_{i_1,\ldots,i_n,\phi^{m}}$ and the propagators above.
561: Moreover it can be easily shown that ${\partial \over
562: \partial_{\bar t_i}}Z=0$, which implies to all orders that
563: $F^{(g)}$ can be written as (\ref{fgformal}). This establishes the
564: reduction of the whole calculation to the determination of the
565: holomorphic modular invariant sections $c^{(g)}_0(t)\in {\cal
566: L}^{2g-2}$. However it also reflects the {\sl major technical
567: problem} in the approach of BCOV, namely that the procedure to
568: determine the recursive anholomorphic part {\sl grows
569: exponentially with the genus}. It has been observed in
570: \cite{Katz:1999} that in concrete cases the terms appearing in the
571: Feynman graph expansion are not functionally independent. This is a
572: hint for finitely generated  rings of anholomorphic modular
573: forms over ${\cal M}(M)$. Using the modular constraints
574: systematically in each integration step Yamaguchi and Yau
575: developed a recursive procedure for the quintic whose complexity grows
576: asymptotically only polynomially, see (\ref{growth}).
577: 
578: Since the B-model is 2d gravity coupled to 2d matter, let us
579: compare the situation with pure  2d gravity, where the objects of
580: interest are  correlation functions of $\tau_{d_i}=(2d_i+1)!!
581: c_1(L_i)^{d_i}$  which are forms on ${\overline {\cal M}_g}$
582: constructed from the descendent fields
583: \begin{equation}
584: F_g(t_0,t_1,\ldots )=\sum_{\{d_i\}} \langle \prod \tau_{d_i}\rangle_g
585: \prod_{r>0} {t_r^{n_r}\over n_r!}\ .
586: \end{equation}
587: Here $\{d_i\}$ are the set of all non-negative integers and  $n_r:={\rm Card}(i:d_i=r)$.
588: 
589: The linear second order differential equations (\ref{linearanomaly}) is the small
590: phase space analog of  the {\sl Virasoro constraints}
591: \begin{equation}
592: L_n Z=0, \qquad n\ge -1
593: \label{virasoroconstraints}
594: \end{equation}
595: on $Z=e^F$ with $F=\sum_{g=0}^\infty \lambda^{2 g-2} F_g$ the free
596: energy of 2d topological gravity~\cite{wittensurvey}. Indeed the
597: $L_n$ with $[L_n,L_m]=(n-m) L_{n+m}$ are second order linear
598: differential operators in the $t_i$. The {\sl non-linear KdV
599: Hierarchy}, which together with dilaton and string  equation are
600: equivalent to (\ref{virasoroconstraints})~\cite{wittensurvey}, and
601: correspond in the small phase space of the B-model to the
602: holomorphic anomaly equations
603: (\ref{eq:anomalyg1},\ref{eq:anomaly}). In the $A$-model approaches
604: to topological string on Calabi-Yau manifolds, such as relative
605: GW-theory, localisation or attempts to solve the theory via
606: massive $(2,2)$ models, the descendents are introduced according
607: to the details of the geometrical construction and then ``summed
608: away''.
609: 
610: The combinatorial cumbersome information in the descendent sums is
611: replaced in the B-model by the contraints from the modular group,
612: holomorphicity and boundary information from the effective 4d
613: action. As a consequence of this beautiful interplay between
614: space-time and world-sheet properties one needs only the small
615: phase space equations (\ref{eq:anomalyg1}, \ref{eq:anomaly},
616: \ref{linearanomaly}).
617: 
618: This approach requires the ability to relate various local
619: expansions of $F^{(g)}$ near the boundary of the moduli space.
620: Sensible local expansions (of terms in the effective action) are
621: in locally monodromy invariant coordinates. As explained
622: in~\cite{Aganagic:2006wq} these coordinates in various patches are
623: related by symplectic transformations on the phase space $H^3(M)$.
624: The latter extend as metaplectic transformations to the wave
625: function $\psi=Z$ of the topological string on the
626: Calabi-Yau~\cite{wittenwavefunction}, which defines the
627: transformation on the $F^{(g)}$. It will be important for us that
628: the real
629: polarisation~\cite{Verlinde:2004ck,Aganagic:2006wq,Gunaydin:2006bz}
630: defines an unique splitting (\ref{fgformal},\ref{fgdefinition}) of
631: local expansions of the $F^{(g)}$ in the anholomorphic modular
632: part determined by the anomaly equations and the holomorphic
633: modular part $c^{(g)}_0(t)$. Aspects of the wave functions
634: properties and the various polarizations have been further
635: discussed
636: in~\cite{Verlinde:2004ck,Aganagic:2006wq,Gunaydin:2006bz}.
637: 
638: 
639: 
640: 
641: 
642: %-------------------------------------------------------------
643: \subsection{Boundary conditions from light BPS states}
644: \label{boundaryconditions}
645: Boundaries in the moduli space ${\cal M}(M)$ correspond to degenerations
646: of the manifold $M$ and general properties of the effective action can be inferred
647: from the physics of the lightest states. More precisely the light states relevant
648: to the $F^{(g)}$ terms in the $N=2$ actions are the BPS  states.
649: Let us first discuss the boundary conditions for $F^{(1)}$ at the singular points
650: in  the moduli space.
651: \begin{itemize}
652: \item At the point of maximal unipotent monodromy in the mirror manifold
653: $W$, the K\"ahler areas, four, and  six volumes of the original
654: manifold $M$ are all large. Therefore the lightest string states
655: are the constant maps $\Sigma_g\rightarrow pt\in M$. For these
656: Kaluza-Klein reduction, i.e. a zero mode analysis of the A-twisted
657: non-linear $\sigma$-model is sufficient to calculate the leading
658: behaviour\footnote{The leading of $F^{(0)}$  at this point is
659: similarly calculated  and given in (\ref{prepotential}).} of
660: $F^{(1)}$ as~\cite{BCOVI}
661: \begin{equation}
662: F^{(1)}={t_i\over 24} \int c_2 \wedge J_i+ {\cal O}(e^{2 \pi i t})\ .
663: \label{F1}
664: \end{equation}
665: Here   $2 \pi i \,t_i={X^i\over X^0}$ are the canonical K\"ahler parameters,
666: $c_2$ is the second Chern class, and $J_i$ is the basis for the
667: K\"ahler cone dual to 2-cycles $C_i$ defining the $t_i:=\int_{C_i} \hat J= \int_{C_i} \sum_{i} t_i J_i$.
668: 
669: \item At the conifold divisor in the moduli space  ${\cal M}(W)$, $W$
670: develops a nodal singularity, i.e., a collapsing  cycle
671: with $S^3$ topology. As discussed in sect. \ref{quinticconifold} this
672: corresponds to the vanishing of the total volume of $M$. The leading
673: behaviour at this point is universally~\cite{Hosono:1994ax}
674: \begin{equation}
675: F^{(1)}= {1\over 12} \log(t_D)+O(t_D)\ .
676: \label{betaconifold}
677: \end{equation}
678: This leading behaviour has been physically explained as the effect of integrating out
679: a non-perturbative hypermultiplet, namely the  extremal black hole of~\cite{Strominger:1995cz}.
680: Its mass $\sim t_D$, see (\ref{massbh}), goes to zero at the conifold and it couples to
681: the $U(1)$ vector in the $N=2$ vectormultiplet, whose lowest component is the modulus $t_D$.
682: The factor $\frac{1}{12}$ comes from the gravitational one-loop $\beta$-function,  which describes the
683: running of the $U(1)$ coupling~\cite{Vafa:1995ta}.
684: A closely related  situation is the one of a shrinking lense space $S/G$.
685: As explained in \cite{GVbfg} one gets in this case several BPS hyper multiplets as the
686: bound states of wrapped D-branes, which modifies the  factor $\frac{1}{12}\rightarrow \frac{|G|}{12}$
687: in the one loop $\beta$-function (\ref{betaconifold}).
688: \item The gravitational $\beta$-function argument extends also to non-perturbative spectra arising at more
689: complicated singularities, e.g. with gauge symmetry enhancement and adjoint
690: matter~\cite{Klemm:1995kj}.
691: \end{itemize}
692: For the case of the one parameter families the above boundary information and the fact is
693: sufficient to fix the holopmorphic ambiguity in $F^{(1)}$.
694: 
695: To learn from  the effective action point of view about the higher
696: genus boundary behaviour, let us recall that the $F^{(g)}$ as in
697: $F(\lambda,t)=\sum_{g=1}^\infty \lambda^{2g-2} F^{(g)}(t)$ give
698: rise to the following term:
699: \begin{equation}
700: S^{N=2}_{1-loop}=\int {\rm d}^4 x R_+^2 F(\lambda ,t)\ ,
701: \label{N=2oneloop}
702: \end{equation}
703: where $R_+$ is the self-dual part of the curvature and we identify
704: $\lambda$ with $F_+$, the self-dual part of the graviphoton field
705: strength. As explained in \cite{GVI,GVII}, see \cite{MB} for a
706: review, the term is computed by a one-loop integral in a constant
707: graviphoton background, which depends only on the left
708: ($SO(4)=SU(2)_L\otimes SU(2)_R$) Lorentz quantum numbers of BPS
709: particles $P$ in the loop. The calculation is very similar to the
710: normal Schwinger-loop calculation. The latter computes the
711: one-loop effective action in an $U(1)$ gauge theory, which comes
712: from integrating out massive particles $P$ coupling to a constant
713: background $U(1)$ gauge field. For a self-dual background field
714: $F_{12}=F_{34}=F$ it leads to the following one-loop determinant
715: evaluation:
716: \begin{equation}
717: S^{S}_{1-loop}=\log\, \det \left(\nabla + m^2 + 2 e \, \sigma_L F\right)=
718: \int_{\epsilon}^\infty \frac{{\rm d} s}{s} \ds{{ {\rm Tr} (-1)^f \exp({- s m^2 }) \exp({-2 s e \sigma_L F})}\over
719: 4 \sin^2\ds{\left( s e F/2\right)}}\ .
720: \label{schwingerloop}
721: \end{equation}
722: Here the $(-1)^f$ takes care of the sign of the log of the
723: determinant depending on whether $P$ is a boson or a fermion, and
724: $\sigma_L$ is the Cartan element in the left Lorentz
725: representation of $P$. To apply this calculation to the $N=2$
726: supergravity case one notes, that the graviphoton field couples
727: to the mass, i.e., we have to identify $e=m$. The loop has two
728: $R_+$ insertions and an arbitrary, (for the closed string action
729: even) number, of graviphoton insertions. It turns out \cite{GVII}
730: that the only supersymmetric BPS states with the Lorentz quantum
731: numbers
732: \begin{equation}
733: \left[\left({\bf\frac{1}{2}},{\bf 0}\right)+2({\bf 0},{\bf 0})\right]\otimes {\cal R}
734: \label{lorentzrep}
735: \end{equation}
736: contribute to the loop. Here  ${\cal R}$ is an arbitrary Lorentz
737: representation of $SO(4)$. Moreover the two $R_+$ insertions are
738: absorbed by the first factor in the Lorentz representation
739: (\ref{lorentzrep}), and the coupling of the particles in the loop
740: to $F_+$ insertions in the $N=2$ evaluation works exactly  as in
741: the non-supersymmetric Schwinger-loop calculation above for $P$ in
742: the representation ${\cal R}$.
743: 
744: What are the microscopic BPS states that run in the loop? They are
745: related to non-perturbative RR states, which are the only charged
746: states in the Type II compactification. They come from branes
747: wrapping cycles in the Calabi-Yau, and as BPS states their masses
748: are proportional to their central charge (\ref{dbranecharge}). For
749: example, in the large radius in the type IIA string on $M$, the
750: mass is determined by integrals of complexified volume forms over
751: even cycles. E.g., the mass of a 2 brane wrapping a holomorphic
752: curve ${\cal C}_\beta\in  H^2(M,\mathbb{Z})$ is given by
753: \begin{equation}
754: m_\beta=\frac{1}{\lambda} \int_{ {\cal C}_\beta\in H^2(M,\mathbb{Z})} (iJ+B)=
755: \frac{1}{\lambda} 2 \pi i t\cdot \beta =:\frac{1}{\lambda}t_\beta\ .
756: \label{massd2}
757: \end{equation}
758: We note that $H^2(M,\mathbb{Z})$ plays here the role of the charge
759: lattice. In the type IIB picture the charge is given by integrals of the
760: normalized holomorphic $(3,0)$-form $\Omega$. In particular the mass of the extremal
761: black hole that vanishes at the conifold  is given by
762: \begin{equation}
763: m_{BH}=\frac{1}{\lambda \int_{A_D} \Omega} \int_{S^3} \Omega =:\frac{1}{\lambda}t_D\ ,
764: \label{massbh}
765: \end{equation}
766: where $A_D$ is a suitable non-vanishing cycle at the conifold. It
767: follows from the discussion in previous paragraph that with the
768: identification $e=m$ and  after a rescaling $s\rightarrow s
769: \lambda / e$ in (\ref{schwingerloop}), as well as absorbing $F$
770: into $\lambda$, one gets a result for (\ref{N=2oneloop})
771: \begin{equation}
772: F(\lambda, t)= \int_{\epsilon}^\infty \frac{{\rm d} s}{s} \ds{{
773: {\rm Tr} (-1)^f \exp({- s t }) \exp({-2 s  \sigma_L
774: \lambda})}\over 4 \sin^2\ds{\left( s \lambda /2\right)}}\ .
775: \label{n=2loop}
776: \end{equation}
777: Here $t$ are the regularized masses, c.f.
778: (\ref{massd2},\ref{massbh}) of the light particles $P$ that are
779: integrated out, $f$ is their spins in ${\cal R}$, and $\sigma_L$
780: is the Cartan element in the representation  ${\cal R}$.
781: 
782: At the large volume point one can related the relevant BPS states
783: actually to bound states of $D2$  with and infinite tower of $D0$
784: branes with quantized momenta along the $M$-theory circle.
785: Moreover the left spin content of the bound state can be related
786: uniquely to the genus of ${\cal C}_\beta$. This beautiful story
787: leads, as explained in \cite{GVI,GVII}, after summing over the to
788: momenta of the $D0$ states  to (\ref{schwingerloopd2d0}), which
789: together with Castelnuovo's bound for smooth curves leads to very
790: detailed and valuable boundary information as explained in Sec.
791: \ref{symplecticinvariants}. It is important to note that all
792: states are massive so that there are no poles in $F^{(g)}$ for
793: $g>1$. Hence the leading contribution is regular and can be
794: extracted from the constant $\beta=0$  contribution in
795: (\ref{schwingerloopd2d0}) as
796: \begin{eqnarray} \label{constantmap}
797: \lim_{t\rightarrow
798: \infty}F^{(g)}_{\textrm{A-model}}=\frac{(-1)^{g-1}B_{2g}B_{2g-2}}{2g(2g-2)(2g-2)!}\cdot \frac{\chi}{2}\ .
799: \label{constantmaps}
800: \end{eqnarray}
801: The moduli space of constant maps factors into tow components: the
802: moduli space of the world sheet curve $\Sigma_g$ and the location
803: of the image point, i.e. $M$. The measures on the components are
804: $\lambda^{3g-3}$ and $c_3(T_M)$ respectively, explaining the above
805: formula.
806: 
807: Let us turn to type IIB compactifications near the conifold. As it was checked
808: with the $\beta$-function  in~\cite{Vafa:1995ta} there is precisely one
809: BPS hypermultiplet with the Lorentz representation of the
810: first factor in (\ref{lorentzrep}) becoming massless at the conifold. In this case
811: the Schwinger-Loop calculation (\ref{n=2loop}) simply becomes
812: \begin{equation}
813: F(\lambda, t_D)=\int_{\epsilon}^\infty \frac{{\rm d} s}{s} \ds{{
814: \exp({- s t_D })} \over 4 \sin^2\ds{\left( s \lambda
815: /2\right)}}+{\cal O}(t_D^0) =\sum_{g=2}^\infty \left(\lambda\over
816: t_D\right)^{2g-2} \frac{(-1)^{g-1} B_{2g}}{2 g (2 g-2)}+{\cal
817: O}(t_D^0)\ . \label{gap}
818: \end{equation}
819: Since there are no other light particles, the above equation
820: (\ref{gap}) encodes all singular terms in the effective action.
821: There will be regular terms coming from other massive states. This
822: is precisely the gap condition.
823: 
824: 
825: 
826: %--------------------------------------------------------------------------------------
827: \section{Quintic}
828: \label{quintic}
829: 
830: We consider the familiar case of the quintic hypersurface in
831: $\mathbb{P}^4$. The topological string amplitudes $F^{(g)}$ were
832: computed up to genus 4 in \cite{BCOV, Katz:1999} using the
833: holomorphic anomaly equation, and fixing the holomorphic ambiguity
834: by various geometric data. It was also observed a long time ago
835: \cite{Ghoshal:1995} that the leading terms in $F^{(g)}$ around the
836: conifold point are the same as $c=1$ strings at the self-dual radius,
837: thus providing useful information for the holomorphic ambiguity. We
838: want to explore whether we can find coordinates in which 
839: the $F^{(g)}$ on the compact Calabi-Yau exhibit the gap structure 
840: around the conifold point that was recently found for local 
841: Calabi-Yau geometries~\cite{Huang}. In order to do this, it is useful to 
842: rewrite the topological string amplitudes
843: as polynomials \cite{Yamaguchi}. We briefly review the formalism
844: in \cite{Yamaguchi}.
845: 
846: The quintic manifold $M$ has one K\"ahler modulus $t$ and its mirror
847: $W$ has one complex modulus $\psi$ and is given by the 
848: equation~\footnote{Here for later convenience we use a slightly 
849: different notation from that of
850: \cite{Yamaguchi}. Their notation is simply related to ours by a
851: change of variable $\psi^5\rightarrow \psi$.}
852: \begin{equation}
853: W=x_1^5+x_2^5+x_3^5+x_4^5+ x_5^5-5\psi^\frac{1}{5}x_1x_2x_3x_4x_5=0\ .
854: \label{quintic} 
855: \end{equation} 
856: 
857: There is a relation between $t$ and $\psi$ known as the mirror 
858: map $t(\psi)$. The mirror map and the genus zero amplitude 
859: can be obtained using the Picard-Fuchs equation on $W$
860: \begin{eqnarray}
861: \{(\psi\partial_{\psi})^4-\psi^{-1}(\psi\partial_{\psi}-\frac{1}{5})
862: (\psi\partial_{\psi}-\frac{2}{5})(\psi\partial_{\psi}-\frac{3}{5})(\psi\partial_{\psi}-\frac{4}{5})\}\omega=0
863: \label{differentialoperator}
864: \end{eqnarray}
865: We can solve the equation as an asymptotic series around
866: $\psi\rightarrow \infty$.
867: %------------------------------------------------------------------------------------------------
868: \subsection{$\psi=\infty$ expansion and integer symplectic basis}
869: \label{integralsymplecticbasis}
870: Here we set the notation for the periods in the integer 
871: symplectic basis on $W$  and the relation of this basis to the 
872: D-brane charges on $M$. Eqs. (\ref{periods},\ref{prepotential}) and 
873: following, apply to all one parameter models.
874:  
875: The point $\psi=\infty$  has maximal unipotent monodromy and
876: corresponds to the large radius expansion of the mirror $M$
877: \cite{Candelas:1990rm}. In the variable $z={1\over 5^5\psi}$ and using the definitions
878: \begin{equation}
879: \omega(z,\rho):= \sum
880: _{n=0}^{\infty}\frac{\Gamma(5(n+\rho)+1)}{\Gamma^5(n+\rho+1)} z^{n
881: +\rho}\qquad D^k_\rho \omega:=\left.\frac{1}{(2\pi i)^k k!}\frac{\partial^k}{\partial^k \rho}\omega\right|_{\rho=0}
882: \end{equation}
883: one can write  the solutions~\cite{Hosono:1994ax}
884: \begin{equation}
885: \begin{array}{rl}
886: \omega_0&= \omega(z,0)=\sum _{n=0}^{\infty}
887: \frac{(5n)!}{(n!)^5(5^5\psi)^{n}}\\[ 2mm]
888: \omega_1&=D_\rho \omega(z,0)={1\over 2 \pi i} \left(\omega_0 \ln(z)+\sigma_1\right)\\[ 2mm]
889: \omega_2&= \kappa D^2_\rho \omega(z,\rho)-c \omega_0={\kappa\over 2\cdot  (2 \pi i)^2 }\left( \omega_0 \ln^2(z)+2 \sigma_1 \ln(z)+\sigma_2\right) \\[2 mm]
890: \omega_3&= \kappa D^3_\rho \omega(z,\rho)-c \omega_1+e \omega_0=\frac{\kappa}{6 \cdot  (2 \pi i)^3 }\left( \omega_0 \ln^3(z)+3 \sigma_1 \ln^2(z)+3 \sigma_2\ln(z)+\sigma_3\right)
891: \end{array}
892: \label{rawbasis}
893: \end{equation}
894: The constants $\kappa,c,e$ are topological intersection numbers,
895: see below. The $\sigma_k$ are also determined by
896: (\ref{differentialoperator}). To the first few orders,
897: $\omega_0=1+120 z+113400z^2+{\cal O}(z^3)$, $\sigma_1=770z+810225
898: z^2+{\cal O}(z^3)$, $\sigma_2=1150z+{4208175\over 2} z^2+{\cal
899: O}(z^3)$ and $\sigma_3=-6900z-{9895125\over 2}z^2+{\cal O}(z^3)$.
900: 
901: 
902: The solutions (\ref{rawbasis}) can be combined into the period vector
903: $\vec \Pi$ with respect to an integer symplectic basis\footnote{With $A^i\cap B_j=\delta^i_j$
904: and zero intersections otherwise.} $(A^i,B_j)$ of $H^3(W,\mathbb{Z})$
905: as follows \cite{Candelas:1990rm}:
906: \begin{equation}
907: \vec \Pi  =
908: \left(
909: \begin{array}{c}
910: \int_{B_1} \Omega\\
911: \int_{B_2} \Omega\\
912: \int_{A^1} \Omega\\
913: \int_{A^2} \Omega
914: \end{array}
915: \right)=
916: \left(
917: \begin{array}{c}
918: F_0\\
919: F_1\\
920: X_0\\
921: X_1\\
922: \end{array}
923: \right)=
924: \omega_0\left(
925: \begin{array}{c}
926: 2{\cal F}^{(0)} -t \partial_t {\cal F}^{(0)}\\
927: \partial_t {\cal F}^{(0)} \\
928: 1\\
929: t\\
930: \end{array}
931: \right)=\left(
932: \begin{array}{c}
933: \omega_3+c\, \omega_1+e\, \omega_0 \\
934: -\omega_2- a\, \omega_1+c \, \omega_0 \\
935: \omega_0\\
936: \omega_1 \\
937: \end{array}
938: \right)\ .
939: \label{periods}
940: \end{equation}
941: Physically, $t$ is the complexified area of a degree one curve and is related by the mirror map
942: \begin{eqnarray}
943: 2 \pi i t(\psi)&=&\int_{\cal C} (iJ+ B)=
944: {\omega_1\over \omega_0}=-\log(5^5\psi)+\frac{154}{625\psi}+\frac{28713}{390625\psi^{2}}+.. \label{mm}\\
945: {1\over z}=5^5 \psi&=&\frac{1}{q} + 770 + 421375\,q +
946: 274007500\,q^2 +\ldots \label{imm}
947: \end{eqnarray}
948: to $\psi$. In (\ref{imm}), we inverted (\ref{mm}) with  $q=e^{2 \pi i t}$.
949: The prepotential is given by
950: \begin{equation}
951: {\cal F}^{(0)}=-{\kappa\over 3!} t^3-{a\over 2}  t^2+ c t+\frac{e}{2}+ f_{inst}(q) \
952: \label{prepotential}
953: \end{equation}
954: where the instanton expansion $f_{inst}(q)$  vanishes in the large
955: radius  $q\rightarrow 0$ limit. The constants in
956: (\ref{periods},\ref{prepotential}) can be related to the classical
957: geometry of the mirror
958: manifold~\cite{Hosono:1994ax,Candelas:1990rm}. Denote by  $J\in
959: H_2(M,\mathbb{Z})$ the K\"ahler form which spans the one-dimensional
960: K\"ahler cone. Then $\kappa=\int_M J\wedge J\wedge J
961: $ is the classical triple intersection number, $c={1\over 24}
962: \int_M c_2(T_M) \wedge J$ is proportional to the  evaluation of
963: the second Chern class of the tangent bundle $T_M$ against $J$,
964: $e={\zeta(3)\over (2 \pi i)^3}\int_M c_3(T_M)$ is proportional to
965: the Euler number, and  $a$ is related to the topology of the divisor
966: $D$ dual to $J$. $A^1$ is topologically a $T_{\mathbb{R}}^3$,
967: while analysis at the conifold shows that the dual cycle $B_1$ has
968: the topology of an $S^3$.
969: 
970: The identification of the central charge formula for compactified
971: type II supergravity~\cite{Ceresole:1995jg} with the K-theory
972: charge of $D$-branes as objects $A\in
973: K(M)$~\cite{Minasian:1997mm,Cheung:1997az},
974: \begin{equation}
975: \vec Q\cdot \vec \Pi=-\int_M e^{-\hat J} {\rm ch}(A)\sqrt{{\rm td}( M )}=Z(A),
976: \label{dbranecharge}
977: \end{equation}
978: with $\hat J =t (i J+B)$ for the one-dimensional K\"ahler cone,
979: checks the $D$-brane interpretation of (\ref{periods})~\cite{Brunner:1999jq,Diaconescu:1999vp,Mayr:2000as}
980: in the $q\rightarrow 0$ limit on the classical terms $\kappa, a, c$, but misses the  ${\chi \zeta(3)\over (2 \pi i)^3}$
981: term. Based on their scaling with the area parameter $t$ the periods $(F_0,F_1,X_0,X_1)$
982: are identified with the masses of $(D_6,D_4,D_0,D_2)$ branes. For smooth intersections and
983: $D$ the restriction of the hyperplane class we can readily calculate $\kappa,a,c,\chi$
984: using the adjunction formula, see Appendix A.
985: 
986: %------------------------------------------------------------------------------------------------------
987: \subsection{Polynomial expansion of $F^{(g)}$}
988: \label{polynomialfg}
989: 
990: >From the periods,  or equivalently the prepotential ${\cal
991: F}^{(0)}$, we can compute the K\"ahler potential $K:=-\log
992: (i\left({\bar X}^{\bar a} F_a- X^{a} \bar F_{\bar a}\right))$ and
993: metric $G_{\psi\bar{\psi}}:=\partial_\psi {\bar\partial}_{\bar
994: \psi}\, K$ in the moduli space. The genus zero Gromov-Witten
995: invariants are obtained by expanding ${\cal F}^{(0)}$ in large
996: K\"ahler parameter in a power series in $q$, see Section
997: \ref{symplecticinvariants}.
998: 
999: 
1000: We use the notation of \cite{Yamaguchi} and introduce the following symbols:
1001: \begin{eqnarray}
1002: &&
1003: A_p:=\frac{(\psi\partial_{\psi})^pG_{\psi\bar{\psi}}}{G_{\psi\bar{\psi}}},
1004: ~~~B_p:=\frac{(\psi\partial_{\psi})^pe^{-K}}{e^{-K}}, ~~(p=1,2,3,\cdots) \nonumber \\
1005: && C:=C_{\psi\psi\psi}\psi^3, ~~~ X:=\frac{1}{1-\psi}=:-\frac{1}{\delta}
1006: \label{generators}
1007: \end{eqnarray}
1008: Here $C_{\psi\psi\psi}\sim \frac{\psi^{-2}}{1-\psi}$ is the three
1009: point Yukawa coupling, and $A:=A_1=-\psi\Gamma_{\psi\psi}^{\psi}$
1010: and $B:=B_1=-\psi\partial_\psi K$ are the Christoffel and K\"ahler
1011: connections. In the holomorphic limit $\bar{\psi}\rightarrow
1012: \infty$, the holomorphic part of the K\"ahler potential and the
1013: metric go like
1014: \begin{eqnarray}
1015: e^{-K}\sim \omega_0, ~~~~G_{\psi\bar{\psi}}\sim \partial_\psi t,
1016: \end{eqnarray}
1017: so in the holomorphic limit, the generators $A_p$ and $B_p$ are
1018: \begin{eqnarray}
1019: A_p=\frac{(\psi\partial_{\psi})^p(\partial_\psi t)}{\partial_\psi
1020: t}, ~~~B_p=\frac{(\psi\partial_{\psi})^p \omega_0}{ \omega_0}.
1021: \end{eqnarray}
1022: It is straightforward to compute them in an asymptotic series
1023: using the Picard-Fuchs equation. There are also some derivative
1024: relations among the generators,
1025: \begin{eqnarray}
1026: \psi\partial_\psi A_p=A_{p+1}-AA_p, ~~~ \psi\partial_\psi
1027: B_p=B_{p+1}-BB_p,~~~ \psi\partial_\psi X=X(X-1). \nonumber
1028: \end{eqnarray}
1029: The topological amplitudes in the ``Yukawa coupling frame'' are
1030: denoted as
1031: \begin{eqnarray}
1032: P_g:=C^{g-1}F^{(g)}, ~~~~
1033: P^{(n)}_g=C^{g-1}\psi^{n}C_{\psi^n}^{(g)}.
1034: \label{defPg}
1035: \end{eqnarray}
1036: The A-model higher genus Gromov-Witten invariants are obtained in
1037: the holomorphic limit $\bar{t}\rightarrow \infty$ with a familiar
1038: factor of $\omega_0$ as
1039: \begin{eqnarray} \label{10-14-2.8}
1040: F^{(g)}_{\textrm{A-model}} &=& \lim_{\bar{t}\rightarrow \infty}
1041: \omega_0^{2(g-1)}(\frac{1-\psi}{5\psi})^{g-1}P_g.
1042: \end{eqnarray}
1043: The $P^{(n)}_g$ are defined for $g=0, n\geq 3$, $g=1, n\geq 1$,
1044: and $g=2, n\geq 0$. We have the initial data
1045: \begin{eqnarray}
1046: P^{(3)}_{g=0}&=&1 \nonumber \\
1047: P^{(1)}_{g=1} &=&
1048: -\frac{31}{3}B+\frac{1}{12}(X-1)-\frac{1}{2}A+\frac{5}{3}
1049: \label{Pg01},
1050: \end{eqnarray}
1051: and using the Christoffel and K\"ahler connections we have the
1052: following recursion relation in $n$,
1053: \begin{eqnarray}
1054: P^{(n+1)}_g=\psi\partial_\psi
1055: P^{(n)}_g-[n(A+1)+(2-2g)(B-\frac{1}{2}X)]P^{(n)}_g.
1056: \end{eqnarray}
1057: 
1058: One can also use the Picard-Fuchs equation and the special
1059: geometry relation to derive the following relations among
1060: generators:
1061: \begin{eqnarray}
1062: B_4&=&2XB_3-\frac{7}{5}XB_2+\frac{2}{5}XB-\frac{24}{625}X \nonumber \\
1063: A_2&=& -4B_2-2AB-2B+2B^2-2A+2XB+XA+\frac{3}{5}X-1.
1064: \end{eqnarray}
1065: By taking derivatives w.r.t. $\psi$ one see that all higher $A_p$
1066: ($p\geq 2$) and $B_p$ ($p\geq 4$) can be written as polynomials of
1067: $A$, $B$, $B_2$, $B_3$, $X$. It is convenient to change variables
1068: from $(A,B,B_2,B_3,X)$ to $(u,v_1,v_2,v_3,X)$ as
1069: \begin{eqnarray}
1070: && B=u, ~~~ A=v_1-1-2u,~~~ B_2=v_2+uv_1, \nonumber \\
1071: && B_3=v_3-uv_2+uv_1X-\frac{2}{5}uX.
1072: \end{eqnarray}
1073: Then the  main result of \cite{Yamaguchi} is the following
1074: proposition.
1075: \\
1076: 
1077: \noindent{\bf Proposition}: Each $P_g$ ($g\geq 2$) is a degree
1078: $3g-3$ inhomogeneous polynomial of $v_1$, $v_2$, $v_3$, $X$, where
1079: one assigns the degree $1,2,3,1$ for $v_1, v_2, v_3, X$,
1080: respectively.
1081: \\
1082: 
1083: For example, at genus two the topological string amplitude is
1084: \begin{eqnarray}
1085: P_2 &=& \frac{25}{144} - \frac{625}{288} v_1 + \frac{25}{24} v_1^2
1086: - \frac{5}{24}v_1^3 - \frac{625}{36}v_2 + \frac{25}{6}v_1v_2 +
1087: \frac{350}{9}v_3 - \frac{5759}{3600}X  \nonumber \\ && -
1088: \frac{167}{720}v_1X + \frac{1}{6}v_1^2X - \frac{475}{12}v_2X +
1089: \frac{41}{3600}X^2- \frac{13}{288}v_1X^2 + \frac{X^3}{240}.
1090: \end{eqnarray}
1091: The expression for the $P_g$, $g=1,\ldots,12 $ for all models
1092: discussed in this paper can be obtained in a
1093: Mathematica readable form on \cite{webpage}, see ``Pgfile.txt.''
1094: 
1095: %--------------------------------------------------------------
1096: \subsection{Integration of the holomorphic anomaly equation}
1097: \label{integration}
1098: The anti-holomorphic derivative $\partial_{\bar{\psi}}B_p$ of
1099: $p\geq 2$ can be related to $\partial_{\bar{\psi}}B$
1100: \cite{Yamaguchi}. Assuming $\partial_{\bar{\psi}}A$ and
1101: $\partial_{\bar{\psi}}B$ are independent, one obtains two
1102: relations for $P^{(n)}_g$ from the BCOV holomorphic anomaly
1103: equation as the following:
1104: \begin{eqnarray}
1105: \frac{\partial P_g}{\partial u}&=&0 \label{BCOV1} \\
1106: (\frac{\partial}{\partial{v_1}}-u\frac{\partial }{\partial
1107: {v_2}}-u(u+X)\frac{\partial}{\partial{v_3}})P_g&=&-\frac{1}{2}(P^{(2)}_{g-1}+\sum_{r=1}^{g-1}P^{(1)}_rP^{(1)}_{g-r})
1108: \label{BCOV2}
1109: \end{eqnarray}
1110: The first equation (\ref{BCOV1}) implies there is no $u$
1111: dependence in $P_g$, as already taken into account in the main
1112: proposition of \cite{Yamaguchi}. The second equation (\ref{BCOV2})
1113: provides a very efficient way to solve $P_g$ recursively up to a
1114: holomorphic ambiguity. To do this, one writes down an ansatz for
1115: $P_g$ as a degree $3g-3$ inhomogeneous polynomial of $v_1$, $v_2$,
1116: $v_3$, $X$, plugs in equation (\ref{BCOV2}), and uses the lower
1117: genus results $P_r^{(1)}$, $P^{(2)}_r$ ($r\leq g-1$) to fix the
1118: coefficients of the polynomials. The number of terms in general
1119: inhomogeneous weighted polynomials in $(v_1$, $v_2$, $v_3,X)$ with
1120: weights $(1,2,3,1)$ is given by the generating function
1121: \begin{equation}
1122: {1\over {(1-x)^3 (1-x^2) (1-x^3)}}=\sum_{n=0}^\infty p(n) x^n\ .
1123: \end{equation}
1124: It is easy to see from the equation (\ref{BCOV2}) that the terms
1125: $v_1,\ldots v_1^{2g-4}$ vanish, which implies that the number of
1126: terms in  $P_g$ is
1127: \begin{equation}
1128: n_g= p(3 g-3)-(2 g-4),
1129: \end{equation}
1130: e.g. $n_g= 14, 62, 185, 435, 877, 1590, 2666, 4211, 6344$ for
1131: $g=1,\ldots,10$. Comparing with $\sum_{n=1}^\infty \tilde p(n)
1132: x^n={1\over (1-x)^5}$ it follows in particular that
1133: asymptotically
1134: \begin{equation}
1135: n_g \preceq (3g-3)^4 \ . \label{growth}
1136: \end{equation}
1137: 
1138: 
1139: Note that equation (\ref{BCOV2}) determines the term
1140: $\hat  P_g(v_1,v_2,v_3,X)$ in
1141: \begin{equation}
1142: P_g=:\hat P_g(v_1,v_2,v_3,X) + f^{(g)}(X)\
1143: \label{fgdefinition}
1144: \end{equation}
1145: completely. We can easily understand why the terms in the modular
1146: as well as holomorphic ambiguity $f^{(g)}(X)$ are not fixed, by
1147: noting that the equation (\ref{BCOV2}) does not change when we add
1148: a term proportional to $X^i$ to $P_g$. This ambiguity must have
1149: the form~\cite{Katz:1999,Yamaguchi}
1150: \begin{eqnarray} \label{10-14-2.16}
1151: f^{(g)}=\sum_{i=0}^{3g-3}a_iX^i\ .
1152: \end{eqnarray}
1153: The maximal power of $X$ is determined by (\ref{gap}). This
1154: follows from (\ref{defPg}) and the universal behaviour of $t_D\sim
1155: \delta+ {\cal O}(\delta^2)$ at the conifold (\ref{dualmirrormap}).
1156: Note in particular from (\ref{leadingconifold}) that $\hat  P_g$
1157: is less singular at that point. The minimal power in
1158: (\ref{10-14-2.16}) follows from (\ref{constantmaps}) for $\psi
1159: \sim \infty$ and the leading behaviour of the solutions in Sect.
1160: \ref{integralsymplecticbasis}.
1161: 
1162: We will try to fix these $3g-2$ unknown constants $a_i$
1163: ($i=0,1,2\cdots 3g-3$) by special structure of expansions of
1164: $F^{(g)}$ around the orbifold point $\psi\sim 0$ and the conifold
1165: point $\psi\sim 1$. Before proceeding to this, we note the
1166: constant term is fixed by the known leading coefficients in large
1167: complex structure modulus limit $\psi\sim \infty$ in
1168: \cite{Marino:1998, GVII, FP}.
1169: 
1170: The leading constant terms in the A-model expansion $\psi\sim
1171: \infty$ come from the constant map from the worldsheet to the
1172: Calabi-Yau (\ref{constantmap}). The large complex structure
1173: modulus behavior of $X$ is
1174: \begin{eqnarray}
1175: X\sim \frac{1}{\psi}\sim q=e^{2\pi it}.
1176: \end{eqnarray}
1177: So only the constant term $a_0$ in the holomorphic ambiguity
1178: contributes to leading term in A-model expansions
1179: (\ref{constantmap}), and is thus fixed. We still have $3g-3$
1180: coefficients $a_i$ ($i=1,2,\cdots, 3g-3$) to be fixed.
1181: 
1182: 
1183: 
1184: %------------------------------------------------------------------
1185: \subsection{Expansions around the orbifold point $\psi=0$}
1186: \label{quinticorbifold}
1187: To analyze the $F^{(g)}$ in a new region of the 
1188: moduli space we have to find the right choice of 
1189: polarization. To do this we analytically 
1190: continue the periods to determine the symplectic 
1191: pairing in the new region and pick the new 
1192: choice of conjugated varaibles.      
1193: 
1194: The solutions of the  Picard-Fuchs equation around the orbifold point$\psi\sim 0$
1195: are four power series solutions with the indices
1196: $\frac{1}{5},\frac{2}{5},\frac{3}{5},\frac{4}{5}$
1197: \begin{equation}
1198: \begin{array}{rl}
1199: \omega_{k}^{\rm orb}&=\ds{\psi^{\frac{k}{5}}\sum_{n=0}^\infty
1200: \frac{\left(\left[\frac{k}{5}\right]_n\right)^5}{\left[k\right]_{5 n}}\left(5^5 \psi\right)^n} \\[2 mm]
1201: &=\ds{-\frac{\Gamma(k)}{\Gamma^5\left(\frac{k}{5}\right)}\int_{{\cal
1202: C}_0} \frac{\dd s}{e^{2 \pi i s}-1}
1203: \frac{\Gamma^5\left(s+\frac{k}{5}\right)}{\Gamma(5 s+k)}
1204: \left(5^5 \psi\right)^{s+\frac{k}{5}} \ , \qquad k=1,\ldots, 4}\ .
1205: \label{orbifoldbasis}
1206: \end{array}
1207: \end{equation}
1208: \parbox{14cm} % Seitenbreite
1209: {
1210:    %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1211:    %% Abbildung contour  %%%%%%%%%%%%%%%%%%%%%%
1212:    %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1213: 
1214:    \begin{center}
1215:    \mbox{
1216:          % \begin{turn}{-90}  % Hier k"onnen Sie das Bild drehen
1217:                              % z.B. um -90 Grad
1218:              \epsfig{file=contour.eps,width=14cm}
1219:          % \end{turn}
1220: }
1221:    \end{center}
1222:    %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1223: }
1224: %
1225: \vskip -2mm
1226: \centerline{ $\qquad\qquad$ ${\cal C}_\infty$ for $|\psi|> 1$ $\qquad\qquad\qquad$ ${\cal C}_0$ for $|\psi|< 1$.}
1227: % nachfolgender Text
1228: \vskip 2mm
1229: 
1230: The Pochhammer symbol are defined as $[a]_n:=\frac{\Gamma(a+n)}{\Gamma(a)}$ and we
1231: normalized the first coefficient in  $\omega_{k}^{\rm orb}= \psi^\frac{k}{5}+{\cal O}(\psi^\frac{5+k}{5})$
1232: to one.  The expression in the first line is recovered from the integral representation
1233: by noting that the only poles inside ${\cal C}_0$ for which the integral
1234: converges for $|\psi|<0$ are from $g(s)=\frac{1}{\exp(2 \pi i s)-1}$, which behaves
1235: at $s^\eps_{-n}=n-\epsilon$, $n\in \mathbb{N}$ as $g(s^\eps_{-n})\sim-\frac{1}{2\pi i \epsilon}$.
1236: 
1237: Up to normalization this basis of solutions is canonically
1238: distinguished, as it diagonalizes the $\mathbb{Z}_5$ monodromy at
1239: $\psi=0$. Similar as for the $\mathbb{C}^3/\mathbb{Z}_3$
1240: orbifold~\cite{Aganagic:2006wq}, it can be viewed as a twist field
1241: basis. Here this basis is induced from the twist field basis of
1242: $\mathbb{C}^5/\mathbb{Z}_5$. As it was argued
1243: in~\cite{Aganagic:2006wq} for $\mathbb{C}^3/\mathbb{Z}_3$, this
1244: twist field basis provides the natural coordinates in which the
1245: $F^{(g)}$ near the orbifold point can be interpreted as generating
1246: functions for orbifold Gromov-Witten invariants. Following up on
1247: foundational work on orbifold Gromov-Witten theory \cite{CR} and
1248: examples in two complex dimensions~\cite{BGP} this prediction has
1249: been checked by direct computation of orbifold  Gromov-Witten
1250: invariants~\cite{CCIT} at genus zero. This provides a beautiful
1251: check on the global picture of mirror symmetry.
1252: 
1253: As explained in~\cite{Aganagic:2006wq} the relation between the
1254: large radius generating function of Gromov-Witten invariants and
1255: generating function of orbifold Gromov-Witten invariants is
1256: provided by the metaplectic transformation of the wave
1257: function~\cite{wittenwavefunction}. Since the change of the phase
1258: space variables from large radius to the orbifold the symplectic
1259: form is only invariant up to scaling one has to change the
1260: definition of the string coupling, which plays the role of
1261: $\hbar$ in the  metaplectic transformation.  These can be viewed
1262: as small phase space specialization of the  metaplectic
1263: transformation on the large phase space~\cite{GiventhalCoates}.
1264: 
1265: Using the modular invariance of the anholomorphic $\hat F^g(t,\bar
1266: t)$ it has been further shown in~\cite{Aganagic:2006wq} that this
1267: procedure of obtaining the transformed holomorphic wave function
1268: is simply equivalent to taking the holomorphic limit on the $\hat
1269: F^g(t,\bar t)$. This point was made~\cite{Aganagic:2006wq} for the
1270: local case. But the only point one  has to keep in mind for the
1271: global case is that $\hat F^g(t,\bar t)$ are globally well defined
1272: sections of the K\"ahler line bundle. i.e. one has to perform a
1273: K\"ahler transformation along with the holomorphic limit.
1274: 
1275: 
1276: We will now study  the transformation from the basis
1277: (\ref{periods}) to the basis (\ref{orbifoldbasis}) to make B-model
1278: prediction along the lines of~\cite{Aganagic:2006wq} with the
1279: additional K\"ahler transformation. Since the symplectic form
1280: $\omega$ on the moduli space is  invariant under monodromy and
1281: $\omega_k^{orb}$ diagonalizes the $\mathbb{Z}_5$ monodromy, we
1282: must have in accordance with the expectation from the orbifold
1283: cohomology $H^*(\mathbb{C}^5/\mathbb{Z}_5)=\mathbb{C}{\bf
1284: 1}_0\oplus\mathbb{C}{\bf 1}_1\oplus\mathbb{C}{\bf
1285: 1}_2\oplus\mathbb{C} {\bf 1}_3\oplus\mathbb{C}{\bf 1}_4$
1286: \begin{equation}
1287: \omega=\dd F_k\wedge \dd X_k= -\frac{5^4s_1}{6}\dd
1288: \omega^{orb}_4\wedge \dd \omega^{orb}_1+ \frac{5^4s_2}{2}\dd
1289: \omega^{orb}_3\wedge \dd \omega^{orb}_2\ .
1290: \end{equation}
1291: The rational factors above have been chosen to match constraints
1292: from special geometric discussed below.  Similarly the monodromy
1293: invariant K\"ahler potential must have the form
1294: \begin{equation}
1295: e^{-K}=\sum_{k=1}^4 r_k \omega_k^{\rm orb} \overline{\omega_k^{\rm orb}}\ .
1296: \label{eK}
1297: \end{equation}
1298: To obtain the $s_i,r_i$ by analytic continuation to the basis (\ref{periods}) we follow \cite{Candelas:1990rm} for the
1299: quintic and the generalisation in~\cite{Klemm:1992tx} for other cases and  note that the integral converges for
1300: $|\psi|>1$ due to the asymptotics of the $f(s)=\frac{\Gamma^5(s+k/5)}{\Gamma(5s+k)}$ term,
1301: when the integral is closed along ${\cal C}_\infty$~\cite{Candelas:1990rm}. At $s^\eps_{n}=-n-\epsilon$
1302: the  $g(s^\eps_{n})$ pole is compensated by the  $f(s^\eps_{n})$ zero and at
1303: $s^\eps_{n,k}= -n-k/5-\epsilon$ we note the expansions
1304: \begin{equation}
1305: \begin{array}{rl}
1306: g(s^\eps_{n,k})=&{\frac{\a^k}{1-\a^k}+
1307: \frac{2\pi i\a^k}{(1-\a^k)^2}\eps+
1308: \frac{(2\pi i)^2\a^k(1+\a^k)}{2(1-\a^k)^3}\eps^2+
1309: \frac{(2\pi i)^3\a^k(1+4\a^k+\a^{2k})}{6(1-\a^k)^4}\eps^3+{\cal O}(\eps^4)}\\[3 mm]
1310: f(s^\eps_{n,k})=&{\frac{\kappa \omega_0(n)}{\eps^4}+
1311: \frac{\kappa \sigma_1(n)}{\eps^3}+
1312: \frac{1}{\eps^2}\left(\frac{\kappa \sigma_2(n)}{2}+\frac{(2\pi i)^2c_{2J} \omega_0(n)}{24}\right)+}\\[3 mm]
1313: &{\frac{1}{\eps}\left(\frac{\kappa \sigma_3(n)}{6}+\frac{(2\pi i)^2 c_{2J}\sigma_1(n)}{24}+\chi \zeta(3)
1314: \omega_0(n)\right)+{\cal O}(\eps^0)}\\[3 mm]
1315: \left(5^5\psi\right)^{s^\eps_n}=&{z^n(1+\log(z)\eps+\frac{1}{2}\log(z)^2\eps^2+\frac{1}{6}\log(z)^3\eps^3+{\cal O}(\eps^4)})
1316: \end{array}.
1317: \end{equation}
1318: Here  $\alpha=\exp(2 \pi i/5)$. $\kappa=\int_MJ^3$,
1319: $c_{2J}=\int_M c_2 J$, and $\chi=\int_M c_3$ are calculated
1320: in App. \ref{intersection}~\footnote{In fact using the
1321: generalization of (\ref{orbifoldbasis}) in~\cite{Klemm:1992tx} it
1322: is easily shown that the combinatorics, which leads to
1323: (\ref{aconst}), are the same as the ones leading to the occurrence
1324: of the classical intersections here.}.  The $\omega_0(n)$,
1325: $\sigma_i(n)$ are coefficients of the series we encountered in
1326: sec. \ref{integralsymplecticbasis}. Performing the residue
1327: integration and comparing with (\ref{rawbasis},\ref{periods}) we
1328: get
1329: \begin{equation}
1330: \omega_k^{orb}=\frac{(2 \pi i)^4 \Gamma(k)}{\Gamma^5\left(\frac{k}{5}\right)}\left(
1331: \frac{\a^k F_0}{1-\a^k}-
1332: \frac{\a^k F_1}{(1-\a^k)^2}+
1333: \frac{5\a^k(\a^{2 k}-\a^k +1)X_0}{(1-\a^k)^4}+
1334: \frac{\a^k(8\a^{k}-3)X_1}{(1-\a^k)^3}\right)
1335: \label{matoi}
1336: \end{equation}
1337: It follows with $r_i=\frac{\Gamma^{10}\left(\frac{k}{5}\right)}{\Gamma^2(k)} c_i$
1338: that
1339: \begin{eqnarray}
1340: c_1&=&-c_4=\a^2 (1-\a)(2+\a^2+\a^3), \quad  c_3=-c_2=\a(2+\a-\a^2-2 \a^3)\nonumber \\
1341: s_1&=&s_2 =-\frac{1}{5^5(2\pi i)^3} \label{metaplectic}\\
1342: \left(\begin{array}{c}F_0\\ F_1\\X_0\\X_1 \end{array} \right)&=&
1343: \psi^{1/5}\frac{\a \Gamma^5\left(1\over 5\right)}{(2 \pi i)^4} \left(\begin{array}{c}
1344: (1-\a)(\a-1-\a^2)\\
1345: \frac{1}{5}(8-3 \a) (1-\a)^2\\
1346: (1-\a+\a^2)\\
1347: \frac{1}{5}(1-\a)^3 \end{array}\right)+{\cal O}(\psi^{2/5})\ . \label{leading}
1348: \end{eqnarray}
1349: Eq. (\ref{metaplectic}) implies that up to a rational rescaling of
1350: the orbifold periods the transformation of the wave functions from
1351: infinity to the orbifold is given by a metaplectic transformation
1352: with the same rescaling of the string coupling as for the
1353: $\mathbb{C}^3/\mathbb{Z}_3$ case in \cite{Aganagic:2003db}. Eq.
1354: (\ref{leading}) implies that there are no projective coordinates
1355: related to an  ${\rm Sp}(4,\mathbb{Z})$ basis, which would vanish
1356: at the orbifold. This means that there is no massless RR state in
1357: the K-theory charge lattice which  vanishes at the orbifold point.
1358: We note further that after rescaling of the orbifold periods the
1359: transformation (\ref{matoi}) can be chosen to lie in ${\rm
1360: Sp}(4,\mathbb{Z}[\a,\frac{1}{5}])$.
1361: 
1362: 
1363: We can define the analogue of mirror map at the orbifold point,
1364: \begin{eqnarray}
1365: s=\frac{\omega_2^{orb}}{\omega^{orb}_1}=\psi^{\frac{1}{5}}(1+\frac{13\psi}{360}+\frac{110069\psi^{2}}{9979200}
1366: +\mathcal{O}(\psi^{3}))
1367: \end{eqnarray}
1368: where  we use the notation $s$, as in~\cite{Aganagic:2006wq}, to
1369: avoid confusion with the mirror map in the large volume limit. We
1370: next calculate the genus zero prepotential at the orbifold point.
1371: For convenience let us rescale our periods $\hat \omega_{k-1}=
1372: 5^{3/2} \omega^{orb}_{k}$. The Yukawa-Coupling is transformed to
1373: the $s$ variables as
1374: \begin{equation}
1375: C_{sss}=\frac{1}{\hat \omega_0^2}
1376: \frac{5}{\psi^2(1-\psi)} \left(\frac{\partial \psi}{\partial s}\right)^3=
1377: 5 +\frac{5}{3} s^5 +\frac{5975}{6048} s^{10}+\frac{34521785}{54486432} s^{15}+{\cal O}(s^{20})  \ .
1378: \end{equation}
1379: A trivial consistency check of special geometry is that the genus
1380: zero prepotential $F^{(0)}=\int \dd s \int \dd s \int \dd s \
1381: C_{sss}$ appears in the periods $\hat \Pi^{orb}=( \hat
1382: \omega_0,\hat \omega_1,\frac{5}{2!} \hat \omega_2,-\frac{5}{3!}
1383: \hat \omega_3)^T$ as
1384: \begin{equation}
1385: \hat \Pi^{orb}=\hat \omega_0 \left(\begin{array}{c}
1386: 1\cr
1387: s\cr
1388: \partial_s F_{\textrm{A-orbf.}}^{(0)}\cr
1389: 2 F_{\textrm{A-orbf.}}^{(0)}- s\partial_s F_{\textrm{A-orbf.}}^{(0)}
1390: \end{array}
1391: \right)\ .
1392: \end{equation}
1393: This can be viewed also as a simple check on the lowest order 
1394: meta-plectic transformation of $\Psi$ which is just the Legendre 
1395: transformation. Note that the Yukawa coupling is 
1396: invariant under the $\mathbb{Z}_5$ which acts as
1397: $s\mapsto \alpha s$. $\mathbb{Z}_5$ implies further that there can be no
1398: integration constants, when passing from $C_{sss}$ to $F_0$ and the coupling
1399: $\lambda$ must transform with $\lambda \mapsto \alpha^\frac{3}{2} \lambda$ to
1400: render $F(\lambda,s,\bar s)$ invariant.
1401: 
1402: The holomorphic limit  $\bar{\psi}\rightarrow 0$  of K\"ahler
1403: potential and metric follows from (\ref{eK}) by extracting the
1404: leading anti-holomorphic behaviour. Denoting\footnote{This  is to
1405: make contact with the other one modulus cases. Of course if
1406: $a_1=a_2$ a log singularity appears and the formula does not
1407: apply.} by $a_k$ the leading powers of  $\omega_k^{orb}$  we find
1408: \begin{equation}
1409: \lim_{\bar \psi\rightarrow 0} e^{-K}=r_1 \bar \psi^{a_1}
1410: \omega^{orb}_1, \qquad \lim_{\bar \psi\rightarrow 0} G_{\psi\bar
1411: \psi}= \bar \psi^{a_2-a_1-1}
1412: \frac{r_2}{r_1}\left(\frac{a_2}{a_1}-1\right) \frac{\partial
1413: s}{\partial \psi} \ .
1414: \end{equation}
1415: Note that the constants and  $\bar \psi$ and its leading power
1416: are irrelvant for the holomorphic limit of the generators
1417: (\ref{generators})
1418: \begin{eqnarray}
1419: X&=& \frac{1}{1-\psi}=1+\psi+\psi^{2}+\mathcal{O}(\psi^{3})
1420: \nonumber \\ A&=& -\frac{4}{5}+\frac{13}{60}\psi+
1421: \frac{3551}{18144}\psi^2+\mathcal{O}(\psi^{3}) \nonumber \\
1422: B&=&
1423: \frac{1}{5}+\frac{1}{120}\psi+\frac{17}{4032}\psi^2+\mathcal{O}(\psi^{3})
1424: \nonumber \\
1425: B_2 &=& \frac{1}{25}+\frac{7}{600}\psi+\frac{1027}{100800}\psi^2
1426: +\mathcal{O}(\psi^{3}) \nonumber
1427: \\
1428: B_3 &=&
1429: \frac{1}{125}+\frac{43}{3000}\psi+\frac{1633}{72000}\psi^2+\mathcal{O}(\psi^{3}).
1430: \nonumber
1431: \end{eqnarray}
1432: 
1433: Using this information and integrating (\ref{Pg01}) we obtain the
1434: genus one free energy $F^{(1)}_{\textrm{A-orbf.}}=
1435: -\frac{s^5}{9}+\ldots$ The regularity of $F^{(1)}$, i.e. the
1436: absence of log terms, is expected as there are no massless BPS
1437: states at the Gepner-point. Because of this, the considerations in
1438: Sec.\ref{boundaryconditions} imply also that the higher genus
1439: amplitudes
1440: \begin{eqnarray} \label{fgorb}
1441: F^{(g)}_{\textrm{A-orbf.}} &=& \lim_{\bar{\tilde t}\rightarrow 0}
1442: (\hat \omega_0)^{2(g-1)}(\frac{1-\psi}{5\psi})^{g-1}P_g
1443: \end{eqnarray}
1444: have no singularity at the orbifold point $\psi\sim 0$. This is in
1445: accordance with the calculations in \cite{Katz:1999} and implies
1446: that $\frac{P_g}{\psi^{\frac{3}{5}(g-1)}}$ is regular at $\psi\sim
1447: 0$.
1448: 
1449: The situation for the higher genus amplitudes  for the compact
1450: ${\cal O}(5)$ constraint in $\mathbb{P}^4$ is considerably
1451: different from the one for the resolution ${\cal O}(-5)\rightarrow
1452: \mathbb{P}^4$ of the $\mathbb{C}^5/\mathbb{Z}_5$. In normal
1453: Gromov-Witten theory for genus $g>0$ on ${\cal O}(5)$ in
1454: $\mathbb{P}^4$ there is no bundle whose Euler class of its pullback
1455: from the ambient space $\mathbb{P}^4$ to the moduli space of
1456: maps gives rise to a suitable measure on ${\cal M}_{g,\beta}$ that
1457: counts the maps to the quintic. This is the same difficulties one
1458: has to face for higher genus calculation for the orbifold GW
1459: theory in ${\cal O}(5)$ in  $\mathbb{P}^4$, and it is notably
1460: different\footnote{Since the brane bound state cohomology at
1461: infinity can only be understood upon including higher genus
1462: information, see Sec. \ref{symplecticinvariants} and
1463: \ref{quinticdbrane}, the claims that one can learn essential
1464: properties about the D-branes of the quintic at small volume from
1465: the $\mathbb{C}^5/\mathbb{Z}_5$ orbifold might be overly
1466: optimistic.} from the equivariant GW theory on
1467: $\mathbb{C}^5/\mathbb{Z}_5$.
1468: 
1469: However we claim that our $F^{(g)}_{\textrm{A-orbf.}}$ predictions
1470: from the $B$-model computation contain the information about the
1471: light even RR states at the orbifold point in useful variables and
1472: could in principle be checked  in the A-model by some version of
1473: equivariant localisation. Below we give the first few order
1474: results. They are available to genus $20$ at \cite{webpage}.
1475: $$
1476: \begin{array}{rl}
1477: F^{(0)}_{\textrm{A-orbf.}}&=\frac{5\,s^3}{6} + \frac{5\,s^8}{1008} + \frac{5975\,s^{13}}{10378368} +
1478: \frac{34521785\,s^{18}}{266765571072}+\ldots\\ [2 mm]
1479: F^{(1)}_{\textrm{A-orbf.}}&= -\frac{s^5}{9} - \frac{163 \, s^{10}}{18144} - \frac{85031 \, s^{15}}{46702656} -
1480: \frac{6909032915\, s^{20}}{20274183401472} +\ldots \\[2 mm]
1481: F^{(2)}_{\textrm{A-orbf.}}&=\frac{155\,s^2}{18} - \frac{5\,s^7}{864} + \frac{585295\,s^{12}}{14370048} +
1482: \frac{1710167735\,s^{17}}{177843714048}+\ldots\\[2 mm]
1483: F^{(3)}_{\textrm{A-orbf.}}&=\frac{488305\,s^4}{9072} - \frac{3634345\,s^9}{979776} - \frac{1612981445\,s^{14}}{7846046208} -
1484:   \frac{2426211933305\,s^{19}}{116115777662976} +\ldots\\[2 mm]
1485: F^{(4)}_{\textrm{A-orbf.}}&=\frac{48550\,s}{567} + \frac{36705385\,s^6}{163296} + \frac{16986429665\,s^{11}}{603542016} +
1486:   \frac{341329887875\,s^{16}}{70614415872}+\ldots\\[2 mm]
1487: F^{(5)}_{\textrm{A-orbf.}}&=\frac{1237460905\,s^3}{224532} + \frac{108607458385\,s^8}{28740096} - \frac{2079654832074515\,s^{13}}{1553517149184} -
1488:   \frac{50102421421803185\,s^{18}}{438808843984896}+\ldots\\ [2 mm]
1489: \end{array}
1490: $$
1491: 
1492: The holomorphic ambiguity (\ref{10-14-2.16}) is a power series
1493: of $\psi$ starting from a constant term, so requiring
1494: $\frac{P_g}{\psi^{\frac{3}{5}(g-1)}}$ to be regular imposes
1495: \begin{equation}
1496: \lceil\frac{3}{5}(g-1)\rceil
1497: \end{equation}
1498: number of relations in $a_i$ in (\ref{10-14-2.16}), where
1499: $\lceil\frac{3}{5}(g-1)\rceil$ is the ceiling, i.e. the smallest
1500: integer greater or equal to $\frac{3}{5}(g-1)$. We note that the
1501: leading behaviour $\omega^{orb}_1\sim \psi^\frac{1}{d}$ with $d=5$
1502: for the quintic, which is typical for an orbifold point in compact
1503: Calabi-Yau, and which ``shields'' the singularity and diminishes
1504: the boundary conditions at the orbifold point from $g-1$ to
1505: $\lceil \frac{d-2}{d}(g-1)\rceil $. If this period is
1506: non-vanishing at $\psi=0$ and it is indeed simply a constant for
1507: all local cases~\cite{Huang2}, one gets $g-1$ conditions, which
1508: together with the gap condition at the conifold and the constant
1509: map information is already sufficient to completely solve the
1510: model. An example of this type is ${\cal O}(-3)\rightarrow
1511: \mathbb{P}^2$.
1512: 
1513: \subsection{Expansions around the conifold point $\psi=1$}
1514: \label{quinticconifold}
1515: An new feature of the conifold region is that there is an 
1516: choice in picking the polarization, but as we will show 
1517: the gap property is independent of this choice.       
1518: 
1519: A basis of solutions of the Picard-Fuchs equation around the
1520: conifold point $\psi-1=\delta\sim 0$ is the following
1521: \begin{equation}
1522: \vec \Pi_c=\left(
1523: \begin{array}{c}
1524: \omega_0^c\\
1525: \omega_1^c\\
1526: \omega_2^c\\
1527: \omega_3^c\\
1528: \end{array}
1529: \right)
1530: =\left(
1531: \begin{array}{l}
1532: 1+\frac{2\delta^3}{625}-\frac{83\delta^4}{18750}+\frac{757\delta^5}{156250}+\mathcal{O}(\delta^6)\\
1533: \delta-\frac{3\delta^2}{10}+\frac{11\delta^3}{25}-\frac{217\delta^4}{2500}+\frac{889\delta^5}{15625}+\mathcal{O}(\delta^6)\\
1534: \delta^2-\frac{23\delta^3}{30}+\frac{1049\delta^4}{1800}-\frac{34343\delta^5}{75000}+\mathcal{O}(\delta^6) \\
1535: \omega_1^c\log(\delta)-\frac{9\,d^2}{20} - \frac{169\,d^3}{450} + 
1536:   \frac{27007\,d^4}{90000} - 
1537:   \frac{152517\,d^5}{625000}+\mathcal{O}(\delta^6)
1538: \end{array}
1539: \right)
1540: \end{equation}
1541: Here we use the superscript ``c'' in the periods to denote them as
1542: solutions around the conifold point. We see that one of the
1543: solutions $\omega_1^c$ is singled out as it multiplies the log in
1544: the solution $\omega_3^c$. By a Lefshetz argument \cite{AGV} it
1545: corresponds to the integral over the vanishing $S^3$ cycle $B_1$
1546: and moreover a solution containing the log is the integral over
1547: dual cycle $A_1$. Comparing with
1548: (\ref{periods},\ref{dbranecharge}) shows in the Type IIA
1549: interpretation that the $D6$ brane becomes massless. To determine
1550: the symplectic basis we analytically continue the solutions
1551: (\ref{periods}) from $\psi=\infty$ and get
1552: \begin{equation}
1553: \left(
1554: \begin{array}{c}
1555: F_0\\[2 mm]
1556: F_1\\[2 mm]
1557: X_0\\[2 mm]
1558: X_1\\[2 mm]
1559: \end{array}
1560: \right)=
1561: \left(\begin{array}{rrrr}
1562: 0&{\sqrt{5}\over 2 \pi i}&0&0\\[1 mm]
1563: a-{11 i\over 2} g& b-{11i \over 2}h& c-{11i \over 2}r&0\\[1 mm]
1564: d&e&f&-{ \sqrt{5} \over (2 \pi i)^2}\\
1565: ig&ih& ir& 0\end{array}
1566: \right)\left(
1567: \begin{array}{c}
1568: \omega_0^c\\[2 mm]
1569: \omega_1^c\\[2 mm]
1570: \omega_2^c\\[2 mm]
1571: \omega_3^c\\[2 mm]
1572: \end{array}
1573: \right)
1574: \end{equation}
1575: Six of the real numbers
1576: $a,\ldots,r$ are only known numerically\footnote{ 
1577: $a= 6.19501627714957\ldots$,
1578: $b= 1.016604716702582\ldots$, 
1579: $c= -0.140889979448831\ldots$, 
1580: $d=1.07072586843016\ldots$, 
1581: $e= -0.0247076138044847\ldots$, 
1582: $g= 1.29357398450411\ldots$, 
1583: $h=\frac{ 2\,b\,g\,\pi -\left( {\sqrt{5}}\,d \right) }{2\,a\,\pi }$, 
1584: $r=\frac{5 + 16\,c\,g\,{\pi }^3}{16\,a\,{\pi }^3}$,
1585: $f=\frac{{\sqrt{5}}\,b + 8\,c\,d\,{\pi }^2}{8\,a\,{\pi }^2}$.}. 
1586: Nevertheless we can give the symplectic form exactly in the new basis 
1587: \begin{equation} 
1588: \dd F_k\wedge \dd X_k = -\frac{1}{(2 \pi i)^3} \left(\frac{5}{2} \dd \omega_2^c\wedge\dd \omega_0^c +  
1589: (-5)\dd \omega_3^c\wedge\dd \omega_1^c\right)\ .
1590: \label{symplecticconifold} 
1591: \end{equation} 
1592: 
1593: 
1594: The mirror map should be invariant under the conifold monodromy
1595: and vanishing at the conifold. The vanishing period has $D6$
1596: brane charge and is singled out to appear in the numerator of the
1597: mirror map. The numerator is not fixed up to the fact that 
1598: $\omega_3^c$ should not appear. The simplest mirror map compatible with 
1599: symplectic form (\ref{symplecticconifold}) is
1600: \begin{eqnarray} \label{dualmirrormap}
1601: t_D(\delta)&:=&\frac{\omega_1^c}{\omega_0^c}=\delta-\frac{3
1602: \delta^2}{10}+\frac{11\delta^3}{75}-
1603: \frac{9\delta^4}{100}+\frac{5839 t_D^5}{93750}
1604: +\mathcal{O}(t_D^6)\\
1605: \delta(t_D)&=&t_D+\frac{3 t_D^2}{10}+\frac{t_D^3}{30}+\frac{t_D^4}{200}+\frac{169 t_D^5}{375000}
1606: +\mathcal{O}(t_D^6)
1607: \end{eqnarray}
1608: We call this the dual mirror map and denote it $t_D$ to
1609: distinguish from the large complex structure modulus case.
1610: 
1611: 
1612: In the holomorphic limit $\bar{\delta}\rightarrow 0$, the Kahler
1613: potential and metric should behave as $e^{-K}\sim \omega_0^c$ and
1614: $G_{\delta\bar{\delta}}\sim \partial_\delta t_D$. We can find the
1615: asymptotic behavior of various generators,
1616: \begin{eqnarray}
1617: X&=& \frac{1}{1-\psi}=-\frac{1}{\delta}
1618: \nonumber \\
1619: A&=&
1620: -\frac{3}{5}-\frac{2}{25}\delta+\frac{2}{125}\delta^2-\frac{52}{9375}\delta^3+\mathcal{O}(\delta^4)
1621: \nonumber \\
1622: B&=&
1623: \frac{6}{625}\delta^2-\frac{76}{9375}\delta^3+\frac{611}{93750}\delta^4
1624: +\mathcal{O}(\delta^5)
1625: \nonumber \\
1626: B_2 &=&
1627: \frac{12}{625}\delta-\frac{16}{3125}\delta^2+\frac{82}{46875}\delta^3
1628: +\mathcal{O}(\delta^4)
1629: \nonumber \\
1630: B_3 &=&
1631: \frac{12}{625}+\frac{28}{3125}\delta-\frac{78}{15625}\delta^2
1632: +\mathcal{O}(\delta^3) \label{leadingconifold.}
1633: \end{eqnarray}
1634: Now we can expand
1635: \begin{eqnarray} \label{10-14-2.21}
1636: F^{(g)}_{\textrm{conifold}}=\lim_{\bar{\delta}\rightarrow
1637: 0}(\omega_0^c)^{2(g-1)}(\frac{1-\psi}{\psi})^{g-1}P_g
1638: \end{eqnarray}
1639: around the conifold point in terms of $t_D$ using the dual mirror
1640: map (\ref{dualmirrormap}). 
1641: 
1642: Remarkably it turns out that  
1643: shifts $\omega_0^c\rightarrow \omega_0^c+b_1\omega_1^c+b_2\omega_2^c$ 
1644: does not affect the structure we are interested in. The fact that the 
1645: $b_1$ shifts do not affect the amplitudes is reminiscent of the ${\rm
1646: SL}(2,\mathbb{C})$ orbit theorem~\cite{Schmidt,CKS} and is proven
1647: in App. \ref{sectioninvariance}. It is therefore reasonable to 
1648: state the results in the more general polarization and define $\hat \omega^c_0= \omega^c_0 
1649: + b_1 \omega^c_1 + b_2 \omega^c_2$. We first determine the genus
1650: $0$ prepotential checking consistency of the solutions with special 
1651: geometry. Defining $\hat t_D=\frac{\omega_1^c}{\hat \omega_0^c}$ and 
1652: $\Pi_{con}=(\hat \omega^c_0,\omega^c_1,\frac{5}{2}\omega^c_2,-5\omega^c_3)^T$ we get
1653: \begin{equation}
1654: \hat \Pi^{orb}=\hat \omega_0 \left(\begin{array}{c}
1655: 1\cr
1656: \hat t_D\cr
1657: 2 F_{\textrm{conif.}}^{(0)}- \hat t_D \partial_{\hat t_D} 
1658: F_{\textrm{conif.}}^{(0)}\cr
1659: \partial_{\hat t_D} F_{\textrm{conif.}}^{(0)}
1660: \end{array}
1661: \right)\ .
1662: \end{equation}
1663: We can now calculate the $F^{(g)}$ in the generalized polarization. Since the normalization of
1664: $\hat \omega_0^c$ is not fixed by the Picard-Fuchs equation, we 
1665: pick the  normalization $b_0=1$ that will be convenient later on\footnote{The $b_0$ dependence can be restored noting
1666: that each order in $\hat t_D$ is homogeneous in $b_i$.} Using the known expression for the ambiguity 
1667: at genus $2$, $3$ \cite{BCOV, Katz:1999,Yamaguchi} for $g=0-3$ and Castelnuovos bound for $g>3$ we 
1668: find the same interesting structure first observed in \cite{Huang}
1669: \begin{equation} 
1670: \begin{array}{rl} 
1671: F_{\textrm{conif.}}^{(0)}&=-\frac{5}{2} \log (\hat t_D) \hat t_D^2 +\frac{5}{12}\left( 1 - 6b_1 \right)\hat t^3_D + 
1672:   \left(\frac{5}{12} \left( b_1 - 3 b_2 \right)  -\frac{89}{1440}  - 
1673: \frac{5}{4}\,b_1^2\right) \hat t_D^4 + {\cal O}(\hat t_D^5)\\ [2 mm]
1674: F_{\textrm{conif.}}^{(1)}&=-\frac{\log (\hat t_D)}{12} + \left( \frac{233}{120} - \frac{113\,b_1}{12} \right) \,\hat t_D + 
1675:   \left(\frac{233\,b_1}{120} - \frac{113\,{b_1}^2}{24} - 
1676:      \frac{107 b_2}{12} -\frac{2681}{7200}\right) \,{\hat t_D}^2 + {{\cal O}(\hat t_D^3)}\\[2 mm]
1677: F_{\textrm{conif.}}^{(2)}&=\frac{1}{240 {\hat t_D}^2} - \left(\frac{120373}{72000}   + \frac{11413 b_2}{144}\right) + 
1678:   \left( \frac{107369}{150000} - \frac{120373\,b_1}{36000} + \frac{23533\,b_2}{720} - 
1679:      \frac{11413 b_1 b_2}{72} \right) \,\hat t_D + {{\cal O}(\hat t_D^2)}\\ [2 mm] 
1680: F_{\textrm{conif.}}^{(3)}&=\frac{1}{1008\,{\hat t_D}^4} - \left(
1681:    \frac{178778753}{324000000}   + \frac{2287087\,b_2}{43200} + 
1682:     \frac{1084235\,{b_2}^2}{864}\right) + {\cal O}(\hat t_D)\\[2 mm]
1683: F_{\textrm{conif.}}^{(4)}&=\frac{1}{1440\,{\hat t_D}^6} - \left( \frac{977520873701}{3402000000000}   
1684: + \frac{162178069379\,b_2}{3888000000} + 
1685:     \frac{5170381469\,{b_2}^2}{2592000} + \frac{490222589\,{b_2}^3}{15552}\right) + 
1686:   {\cal O}(\hat t_D) \cr.
1687: \end{array}
1688: \end{equation}
1689: As explained in sec. \ref{boundaryconditions} we expect this gap structure
1690: to be present at higher genus as in (\ref{thegap}), i.e.
1691: \begin{eqnarray}
1692: F^{(g)}_{\textrm{conifold}}&=&\frac{(-1)^{g-1}B_{2g}}{2g(2g-2)\hat t_D^{2g-2}}+\mathcal{O}(\hat t_D^0).
1693: \end{eqnarray}
1694: If this is true then it will impose $2g-2$ conditions on the
1695: holomorphic ambiguity (\ref{10-14-2.16}). We can further note that
1696: because the prefactor in (\ref{10-14-2.21}) goes like
1697: $\delta^{g-1}$, and the generator $X$ goes like $X\sim
1698: \frac{1}{\delta}$, the terms $a_iX^i$ in holomorphic ambiguity
1699: (\ref{10-14-2.16}) with $i\leq g-1$ do not affect the gap
1700: structure in (\ref{thegap}). Therefore the gap structure fixes the
1701: coefficients $a_i$ of $i=g, \cdots, 3g-3 $ in (\ref{10-14-2.16}),
1702: but not the coefficients $a_i$ with $i\leq g-1$. 
1703: Note that the choice of $b_i$, $i=0,1,2$ does not  affect the gap structure 
1704: at all.  In App. \ref{sectioninvariance} it is proven that the generators 
1705: of the modular forms in $P_g$ do not change under the shift $b_1$. 
1706: The effect of this shift is hence merely a K\"ahler gauge 
1707: transformation.   On the other hand we note that imposing the 
1708: invariance of the gap structure under the $b_2$ shift does not 
1709: give further conditions on the $a_i$, but it does affect 
1710: the subleading expansion.
1711: 
1712: 
1713: 
1714: 
1715: 
1716: \subsection{Fixing the holomorphic ambiguity: a summary of results}
1717: 
1718: Putting all the information together, let us do a counting of
1719: number of unknown coefficients. Originally we have $3g-2$
1720: coefficients in (\ref{10-14-2.16}). The constant map calculation
1721: \cite{Marino:1998, GVI, FP} in A-model large complex structure
1722: modulus limit $\psi\sim \infty$ fixes one constant $a_0$. The
1723: conifold expansion around $\psi\sim 1$ fixes $2g-2$ coefficients
1724: $a_i$ of $i=g, \cdots, 3g-3$, and the orbifold expansion around
1725: $\psi\sim 0$ further fixes $\lceil\frac{3(g-1)}{5}\rceil$
1726: coefficients. So the number of unknown coefficients at genus $g$
1727: is
1728: \begin{eqnarray}
1729: 3g-2-(1+2g-2+\lceil\frac{3(g-1)}{5}\rceil)=[\frac{2(g-1)}{5}].
1730: \label{boundarycountquintic}
1731: \end{eqnarray}
1732: This number is zero for genus $g=2,3$. So we could have computed
1733: the $g=2,3$ topological strings using this information, although
1734: the answers were already known. At $g\geq 4$ there are still
1735: some unknown constants. However, in the A-model expansion when we
1736: rewrite the Gromov-Witten invariants in terms of Gopakumar-Vafa
1737: invariants, one can in principle use the Castelnuovos bound to fix
1738: the $F^{(g)}$ up to genus 51. This will be shown in sect.
1739: \ref{quinticdbrane}.
1740: 
1741: 
1742: 
1743: \section{One-parameter Calabi-Yau spaces with three regular singular points}
1744: We  generalize the analysis for the quintic to other one K\"ahler
1745: parameter Calabi-Yau three-folds, whose mirror $W$ has a
1746: Picard-Fuchs-system with three regular singular points. Note that
1747: this type of CY has been completely classified~\cite{DoranMorgan}
1748: starting from the Riemann-Hilbert reconstruction of the
1749: Picard-Fuchs equations given the monodromies and imposing the
1750: special geometry property as well as integrality conditions on the
1751: solutions of the latter. There are 13 cases whose mirrors can be
1752: realized as hypersurfaces and complete intersections in weighted
1753: projective spaces with trivial fundamental group\footnote{We also 
1754: have some results on the cases with non-trivial fundamental group. 
1755: They are available on request.}, and are well-known
1756: in the mirror symmetry literature. A fourteenth case is
1757: related to a degeneration of a two parameter model as pointed out
1758: in~\cite{KKRS}. Five are obtained as a free discrete orbifold of
1759: the former CY.
1760: 
1761: 
1762: We focus on the $13$ former cases. In the notation of
1763: \cite{Katz:1999} these complete intersections of degree
1764: $(d_1,d_2,\cdots,d_k)$ in weighted projective spaces
1765: $\mathbb{P}^n(w_1,\dots,w_l)$, Calabi-Yau manifolds are
1766: abbreviated as $X_{d_1,d_2,\cdots,d_k}(w_1,\dots,w_l)$. For
1767: example, the familiar quintic manifold is denoted as $X_{5}(1^5)$.
1768: The list of 13 such examples is the following:
1769: \begin{eqnarray}
1770: &&X_{5}(1^5):
1771: \vec{a}=(\frac{1}{5},\frac{2}{5},\frac{3}{5},\frac{4}{5}),~~
1772: X_{6}(1^4,2):
1773: \vec{a}=(\frac{1}{6},\frac{2}{6},\frac{4}{6},\frac{5}{6}),~~
1774:  X_{8}(1^4,4):
1775: \vec{a}=(\frac{1}{8},\frac{3}{8},\frac{5}{8},\frac{7}{8}),\nonumber \\
1776: && X_{10}(1^3,2,5):
1777: \vec{a}=(\frac{1}{10},\frac{3}{10},\frac{7}{10},\frac{9}{10}),~~
1778: X_{3,3}(1^6):
1779: \vec{a}=(\frac{1}{3},\frac{1}{3},\frac{2}{3},\frac{2}{3}),\nonumber \\
1780: && X_{4,2}(1^6):
1781: \vec{a}=(\frac{1}{4},\frac{1}{2},\frac{1}{2},\frac{3}{4}),
1782: ~~X_{3,2,2}(1^7):
1783: \vec{a}=(\frac{1}{3},\frac{1}{2},\frac{1}{2},\frac{2}{3}),\nonumber \\
1784: && X_{2,2,2,2}(1^8):
1785: \vec{a}=(\frac{1}{2},\frac{1}{2},\frac{1}{2},\frac{1}{2}),~~
1786: X_{4,3}(1^5,2):
1787: \vec{a}=(\frac{1}{4},\frac{1}{3},\frac{2}{3},\frac{3}{4}),
1788: \nonumber \\
1789: &&X_{4,4}(1^4,2^2):
1790: \vec{a}=(\frac{1}{4},\frac{1}{4},\frac{3}{4},\frac{3}{4}),~~
1791: X_{6,2}(1^5,3):
1792: \vec{a}=(\frac{1}{6},\frac{1}{2},\frac{1}{2},\frac{5}{6}),
1793: \nonumber \\
1794: &&X_{6,4}(1^3,2^2,3):
1795: \vec{a}=(\frac{1}{6},\frac{1}{4},\frac{3}{4},\frac{5}{6}),~~
1796: X_{6,6}(1^2,2^2,3^2):
1797: \vec{a}=(\frac{1}{6},\frac{1}{6},\frac{5}{6},\frac{5}{6}).
1798: \nonumber
1799: \end{eqnarray}
1800: The examples satisfy the Calabi-Yau condition $\sum_i d_i=\sum_i
1801: w_i$ required by the vanishing of the first Chern class. The
1802: components of the vector $\vec{a}$ specify the Picard-Fuchs
1803: operators for the mirror manifolds with $h^{2,1}=1$,
1804: \begin{eqnarray} \label{PF-10-28-mh}
1805: \{(\psi\frac{\partial}{\partial
1806: \psi})^4-\psi^{-1}\prod_{i=1}^4(\psi\frac{\partial}{\partial
1807: \psi}-a_i)\}\Pi=0.
1808: \end{eqnarray}
1809: The indices of the Picard-Fuchs equation satisfy $\sum_{i=1}^4
1810: a_i=2$ and we have arranged $a_i$ in increasing order for later
1811: convenience.
1812: 
1813: There are three singular points in the moduli space. The maximally
1814: unipotent point is the large complex structure modulus limit
1815: $\psi\sim \infty$ that has three logarithmic solutions for the
1816: Picard-Fuchs equation. The conifold point is $\psi=1$ with three
1817: power series solutions and one logarithmic solution for the
1818: Picard-Fuchs equation. If the indices $a_i$ are not degenerate,
1819: the Picard-Fuchs equation around the orbifold point $\psi=0$ has
1820: four powers series solutions with the leading terms going like
1821: $\psi^{a_i}$. Each degeneration of the indices generates a
1822: logarithmic solution for the Picard-Fuchs equation around the
1823: orbifold point $\psi=0$.
1824: %-----------------------------------------------------------------------
1825: \subsection{The integration of the anomaly equation}
1826: We can straightforwardly generalize the formalism in
1827: \cite{Yamaguchi} to the above class of models.  The mirror map is
1828: normalized as $t= \log(\frac{\prod_i d_i^{d_i}}{\prod_i
1829: w_i^{w_i}}\psi)+\mathcal{O}(\frac{1}{\psi})$, so that the
1830: classical intersection number in the prepotential is $F^{(0)}=
1831: \frac{\kappa}{6}t^3+\cdots$ where $\kappa=\frac{\prod_i
1832: d_i}{\prod_i w_i}$. The generators of the topological string
1833: amplitudes are defined accordingly,
1834: \begin{eqnarray}
1835: &&A_p:=\frac{(\psi\partial\psi)^p
1836: G_{\psi\bar{\psi}}}{G_{\psi\bar{\psi}}},~~B_p:=\frac{(\psi\partial\psi)^p
1837: e^{-K}}{e^{-K}}, ~~(p=1,2,3,\cdots) \nonumber \\ &&
1838: C:=C_{\psi\psi\psi}\psi^3,~~X:=\frac{1}{1-\psi}.
1839: \end{eqnarray}
1840: Here the familiar three point Yukawa coupling is
1841: $C_{\psi\psi\psi}\sim \frac{\psi^{-2}}{1-\psi}$, and as in the
1842: case of the quintic we denote $A:=A_1$ and $B:=B_1$. The generators
1843: satisfy the derivative relations
1844: \begin{eqnarray}
1845: \psi\partial_\psi A_p=A_{p+1}-AA_p, ~~~ \psi\partial_\psi
1846: B_p=B_{p+1}-BB_p,~~~ \psi\partial_\psi X=X(X-1), \nonumber
1847: \end{eqnarray}
1848: and the recursion relations
1849: \begin{eqnarray}
1850: B_4 &=&
1851: (\sum_ia_i)XB_3-(\sum_{i<j}a_ia_j)XB_2+(\sum_{i<j<k}a_ia_ja_k)XB-(\prod_i
1852: a_i)X, \nonumber \\
1853: A_2 &=& -4B_2-2AB-2B+2B^2-2A+2XB+XA-r_0X-1,
1854: \end{eqnarray}
1855: where the first equation can be derived from the Picard-Fuchs
1856: equation (\ref{PF-10-28-mh}) and the second equation is derived
1857: from the special geometry relation up to a holomorphic ambiguity
1858: denoted by the constant $r_0$. One can fix the constant $r_0$ by
1859: expanding the generators around any of the singular points in the
1860: moduli space. For example, as we will explain, the asymptotic
1861: behaviors of various generators around the orbifold point are
1862: $A_p=(a_2-a_1-1)^p+\mathcal{O}(\psi)$,
1863: $B_p=a_1^p+\mathcal{O}(\psi)$, and $X=1+\mathcal{O}(\psi)$, so we
1864: find the constant is
1865: \begin{eqnarray} \label{constantr0-10-28}
1866: r_0=a_1(1-a_1)+a_2(1-a_2)-1.
1867: \end{eqnarray}
1868: The polynomial topological amplitudes $P_g$ are defined by
1869: $P^{(n)}_g:=C^{g-1}\psi^nC^{(g)}_{\psi^n}$, and satisfy the
1870: recursion relations with initial data
1871: \begin{eqnarray}
1872: P^{(3)}_{g=0}&=&1, \nonumber \\
1873: P^{(1)}_{g=1} &=&
1874: (\frac{\chi}{24}-2)B-\frac{A}{2}+\frac{1}{12}(X-1)+\frac{s_1}{2},
1875: \nonumber \\
1876: P^{(n+1)}_g &=& \psi\partial_\psi
1877: P^{(n)}_g-[n(A+1)+(2-2g)(B-\frac{X}{2})]P^{(n)}_g,
1878: \end{eqnarray}
1879: where $\chi$ is the Euler character of the Calabi-Yau space and
1880: $s_1=2 c-\frac{5}{6}$ is a constant that can be fixed by the
1881: second Chern class of the Calabi-Yau; see appendix A. We provide
1882: the list of constants for the 13 cases of one-parameter Calabi-Yau
1883: in the following table.
1884: 
1885: \begin{table}
1886: \begin{centering}
1887: \begin{tabular}{|r|r|r|r|r|r|r|r|}
1888: \hline CY &$X_5(1^5)$ & $X_6(1^4,2)$ & $X_8(1^4,4)$ &
1889: $X_{10}(1^3,2,5)$ & $X_{3,3}(1^6)$
1890: \\ \hline
1891: $\chi$ & $-200$ & $-204$ & $-296$ & $-288$ & $-144$ \\
1892: \hline $s_1$ & $\frac{10}{3}$ & $\frac{8}{3}$ & $\frac{17}{6}$ &
1893: $2$ & $\frac{11}{3}$ \\ \hline CY & $X_{4,2}(1^6)$ &
1894: $X_{3,2,2}(1^7)$ &$X_{2,2,2,2}(1^8)$ & $X_{4,3}(1^5,2)$ &
1895: $X_{4,4}(1^4,2^2)$  \\ \hline $\chi$ & $-176$ & $-144$  & $-128$ &
1896: $-156$ & $-144$  \\ \hline $s_1$  & $\frac{23}{6}$ &
1897: $\frac{25}{6}$ & $\frac{9}{2}$ & $\frac{19}{6}$ & $\frac{5}{2}$ \\
1898: \hline CY & $X_{6,2}(1^5,3)$ & $X_{6,4}(1^3,2^2,3)$ &
1899: $X_{6,6}(1^2,2^2,3^2)$ & &\\ \hline  $\chi$ & $-256$ & $-156$ &
1900: $-120$  &
1901: &\\ \hline  $s_1$ & $\frac{7}{2}$ & $\frac{11}{6}$ & $1$  & & \\
1902: \hline
1903: \end{tabular} \caption{Euler numbers and the constant $s_1$}
1904: \end{centering}
1905: \end{table}
1906: 
1907: After changing variables to a convenient basis from
1908: $(A,B,B_2,B_3,X)$ to $(u,v_1,v_2,v_3,X)$ as the following:
1909: \begin{eqnarray}
1910: &&B=u,~~A=v_1-1-2u,~~B_2=v_2+uv_1, \nonumber \\ &&
1911: B_3=v_3-uv_2+uv_1X-(r_0+1)u X. \nonumber
1912: \end{eqnarray}
1913: One can follow \cite{Yamaguchi} and use the BCOV formalism to show
1914: that $P_g$ is a degree $3g-3$ inhomogeneous polynomial of
1915: $v_1,v_2,v_3, X$ with the assigned degrees $1,2,3,1$ for $v_1,
1916: v_2, v_3, X$ respectively. The BCOV holomorphic anomaly equation
1917: becomes
1918: \begin{eqnarray}
1919: (\frac{\partial}{\partial{v_1}}-u\frac{\partial }{\partial
1920: {v_2}}-u(u+X)\frac{\partial}{\partial{v_3}})P_g
1921: =-\frac{1}{2}(P^{(2)}_{g-1}+\sum_{r=1}^{g-1}P^{(1)}_rP^{(1)}_{g-r}).
1922: \end{eqnarray}
1923: As in the case of the quintic, the holomorphic anomaly equation
1924: determines the polynomial $P_g$ up to a holomorphic ambiguity
1925: \begin{eqnarray}
1926: f=\sum_{i=0}^{3g-3} a_iX^i.
1927: \end{eqnarray}
1928: 
1929: \subsection{The boundary behaviour}
1930: The constant term $a_0$ is fixed by the known constant map
1931: contribution in the A-model expansion
1932: \begin{eqnarray}
1933: F^{g}_{\textrm{A-model}}=\lim _{t\rightarrow \infty}
1934: \omega_0^{2(g-1)}(\frac{1-\psi}{\kappa \psi})^{g-1}P_g,
1935: \end{eqnarray}
1936: where $\omega_0$ is the power series solution in the large complex
1937: structure modulus limit.
1938: 
1939: The main message here is that structure of the conifold expansion
1940: is universal. For all cases of Calabi-Yau spaces, the Picard-Fuchs
1941: equation around $z=\psi-1$ has four solutions that go like
1942: $\omega_0= 1+\mathcal{O}(z)$, $\omega_1= z+\mathcal{O}(z^2)$,
1943: $\omega_2=z^2+\mathcal{O}(z^3)$, and
1944: $\omega_4=\omega_1\log(z)+\mathcal{O}(z^4)$. We define a dual
1945: mirror map $t_D=\frac{\omega_1}{\omega_0}=z+\mathcal{O}(z^2)$,
1946: expand the topological string amplitudes in terms of $t_D$, and
1947: impose the following gap structure in the conifold expansion:
1948: \begin{eqnarray}
1949: F^{(g)}_{\textrm{conifold}} &=& \lim_{\bar{z}\rightarrow
1950: 0}\omega_0^{2(g-1)}(\frac{1-\psi}{\psi})^{g-1}P_g \nonumber \\
1951: &=&
1952: \frac{(-1)^{g-1}B_{2g}}{2g(2g-2)t_D^{2g-2}}+\mathcal{O}(t_D^0).
1953: \end{eqnarray}
1954: This fixes $2g-2$ coefficients in the holomorphic ambiguity.
1955: 
1956: On the other hand, we discover a rich variety of singularity
1957: structures around the orbifold point $\psi=0$. A natural
1958: symplectic basis of solutions of the Picard-Fuchs equation
1959: (\ref{PF-10-28-mh}) is picked out by the fractional powers of the
1960: leading terms. As in the case of quintic for our purpose we only
1961: need the first two solutions $\omega_0$, $\omega_1$. The leading
1962: behaviors are
1963: \begin{enumerate}
1964: \item If $a_2>a_1$, then we have
1965: $\omega_0=\psi^{a_1}(1+\mathcal{O}(\psi))$,
1966: $\omega_1=\psi^{a_2}(1+\mathcal{O}(\psi))$.
1967: \item If $a_2=a_1$, then we have
1968: $\omega_0=\psi^{a_1}(1+\mathcal{O}(\psi))$,
1969: $\omega_1=\omega_0(\log(\psi)+\mathcal{O}(\psi))$.
1970: \end{enumerate}
1971: In both cases we can define a mirror map around the orbifold point
1972: as $s=\frac{\omega_1}{\omega_0}$. Using the behaviors of
1973: K\"ahler potential $e^{-K}\sim \omega_0$ and metric
1974: $G_{\psi\bar{\psi}}\sim
1975: \partial_\psi s$ in the
1976: holomorphic limit, we can find the asymptotic expansion of various
1977: generators. In both cases $a_2>a_1$ and $a_2=a_1$, the leading
1978: behaviors are non-singular:
1979: \begin{eqnarray}
1980: A_p &=& (a_2-a_1-1)^p+\mathcal{O}(\psi) \nonumber \\
1981: B_p &=& a_1^p+\mathcal{O}(\psi).
1982: \end{eqnarray}
1983: This can be used to fix a holomorphic ambiguity
1984: (\ref{constantr0-10-28}) relating $A_2$ to other generators. On
1985: the other hand, since the constant (\ref{constantr0-10-28})  can
1986: be also derived at other singular points of the moduli space, this
1987: also serves as a consistency check that we have chosen the correct
1988: basis of solutions $\omega_0$, $\omega_1$ at the orbifold point.
1989: 
1990: The orbifold expansion of the topological string amplitudes are
1991: \begin{eqnarray}
1992: F^{(g)}_{\textrm{orbifold}}=\lim_{\bar{\psi}\rightarrow 0
1993: }\omega_0^{2(g-1)}(\frac{1-\psi}{\psi})^{g-1}P_g\sim
1994: \frac{P_g}{\psi^{(1-2a_1)(g-1)}}
1995: \end{eqnarray}
1996: We can expand the polynomial $P_g$ around the orbifold point
1997: $\psi=0$. Generically, $P_g$ is power series of $\psi$ starting
1998: from a constant term, so $F^{(g)}_{\textrm{orbifold}}\sim
1999: \frac{1}{\psi^{(1-2a_1)(g-1)}}$. Since $\sum_ia_i=2$ and $a_1$ is
2000: the smallest, we know $a_1\leq \frac{1}{2}$ and the topological
2001: string amplitude around orbifold point is generically singular.
2002: Interestingly, we find the singular behavior of the topological
2003: strings around the orbifold point is not universal and falls into
2004: four classes:
2005: 
2006: \begin{enumerate}
2007: \item This class includes all 4 cases of hypersurfaces and 4 other cases of complete intersections.
2008: They are the Calabi-Yau spaces $X_5(1^5)$, $X_6(1^4,2)$,
2009: $X_8(1^4,4)$, $X_{10}(1^3,2,5)$, $X_{3,3}(1^6)$,
2010: $X_{2,2,2,2}(1^8)$, $X_{4,4}(1^4,2^2)$, $X_{6,6}(1^2,2^2,3^2)$.
2011: For these cases the singularity at the orbifold point is cancelled
2012: by the series expansion of the polynomial $P_g$. The requirement
2013: of cancellation of singularity in turn imposes
2014: \begin{eqnarray}
2015: \lceil(1-2a_1)(g-1)\rceil
2016: \end{eqnarray}
2017: conditions on the holomorphic ambiguity in $P_g$. Taking into
2018: account the A-model constant map condition and the boundary
2019: condition from conifold expansion, we find the number of unknown
2020: coefficients at genus $g$ is
2021: \begin{eqnarray} \label{number1-10-28}
2022: [2a_1(g-1)].
2023: \end{eqnarray}
2024: Notice for the Calabi-Yau $X_{2,2,2,2}(1^8)$ the cancellation is
2025: trivial since in this case $a_1=\frac{1}{2}$ and topological
2026: strings around the orbifold point are generically non-singular, so
2027: this does not impose any boundary conditions.
2028: 
2029: 
2030: \item This class of Calabi-Yau spaces include $X_{4,2}(1^6)$ and
2031: $X_{6,2}(1^5,3)$. We find the singularity at the orbifold point is
2032: not cancelled by the polynomial $P_g$, but it has a gap structure
2033: that closely resembles the conifold expansion. Namely, when we
2034: expand the topological strings in terms of the orbifold mirror map
2035: $s=\frac{\omega_1}{\omega_0}$, we find
2036: \begin{eqnarray}
2037: F^{(g)}_{\textrm{orbifold}}=\frac{C_g}{s^{2(g-1)}}+\mathcal{O}(s^0)
2038: \end{eqnarray}
2039: where in our normalization convention, the constant is
2040: $C_g=\frac{B_{2g}}{2^{5g-4}(g-1)g}$ for the $X_{4,2}(1^6)$ model,
2041: and $C_g=\frac{B_{2g}}{3^{3(g-1)}4(g-1)g}$ for the
2042: $X_{6,2}(1^5,3)$ model. Since the mirror map goes like
2043: $s\sim \psi^{a_2-a_1}$, and the $P_g$ is a power series of
2044: $\psi$, this imposes
2045: \begin{eqnarray}
2046: \lceil 2(a_2-a_1)(g-1) \rceil
2047: \end{eqnarray}
2048: conditions on the holomorphic ambiguity in $P_g$. In both cases of
2049: $X_{4,2}(1^6)$ and $X_{6,2}(1^5,3)$, we have $a_2=\frac{1}{2}$, so
2050: the number of un-fixed coefficients is the same as the models in
2051: the first class, namely
2052: \begin{eqnarray}
2053: [2a_1(g-1)].
2054: \end{eqnarray}
2055: 
2056: \item This class of Calabi-Yau spaces include only the Calabi-Yau $X_{3,2,2}(1^7)$.
2057: For this model, the singularity around the orbifold point does not
2058: cancel, so there is no boundary condition imposed on the
2059: holomorphic ambiguity in $P_g$. At genus $g$ we are simply left
2060: with $g-1$ unknown coefficients.
2061: 
2062: \item This class of Calabi-Yau spaces include $X_{4,3}(1^5,2)$ and
2063: $X_{6,4}(1^3,2^2,3)$. For this class of models, the singularity at
2064: the orbifold point is partly cancelled. Specifically, we find
2065: $\frac{P_g}{\psi^{(1-2a_1)(g-1)}}$ is not generically regular at
2066: $\psi\sim 0$, but $\frac{P_g}{\psi^{(1-2a_2)(g-1)}}$ is always
2067: regular. This then imposes
2068: \begin{eqnarray}
2069: \lceil(1-2a_2)(g-1)\rceil
2070: \end{eqnarray}
2071: conditions on holomorphic ambiguity in $P_g$, and the number of
2072: unknown coefficients at genus $g$ is now
2073: \begin{eqnarray}
2074: [2a_2(g-1)].
2075: \end{eqnarray}
2076: 
2077: \end{enumerate}
2078: 
2079: It looks like we have the worst scenarios in the two models
2080: $X_{2,2,2,2}(1^8)$ and $X_{3,2,2}(1^7)$, where we essentially get
2081: no obvious boundary conditions at the orbifold point $\psi=0$.
2082: However after a closer examination, we find some patterns in the
2083: leading coefficients of the orbifold expansion as the followings.
2084: In our normalization convention, we find the leading constant
2085: coefficients of $X_{2,2,2,2}(1^8)$ model is
2086: \begin{eqnarray}
2087: F^{(g)}_{\textrm{orbifold}}=\frac{(21\cdot2^{2g-2}-5)(-1)^{g}B_{2g}B_{2g-2}}{2^{2g-3}g(2g-2)(2g-2)!}
2088: +\mathcal{O}(\psi)
2089: \end{eqnarray}
2090: whereas for $X_{3,2,2}(1^7)$ model the leading coefficients are
2091: \begin{eqnarray} \label{twoparticles-12-01}
2092: F^{(g)}_{\textrm{orbifold}}=
2093: \frac{(7\cdot2^{2g-2}-1)B_{2g}}{2^{4g-3}3^{2g-2}(g-1)g}
2094: \frac{1}{s^{2g-2}}+\mathcal{O}(\frac{1}{s^{2g-8}})
2095: \end{eqnarray}
2096: These leading coefficients provide one more boundary condition for
2097: the models $X_{2,2,2,2}(1^8)$ and $X_{3,2,2}(1^7)$, although this
2098: is not much significant at large genus. On the other hand, we
2099: observe the leading coefficients of $X_{2,2,2,2}(1^8)$ model are
2100: similar to the constant map contribution of Gromov-Witten
2101: invariants except the factor of $(21\cdot2^{2g-2}-5)$, whereas the
2102: leading coefficients of $X_{3,2,2}(1^7)$ model are similar to the
2103: conifold expansion except the factor of $(7\cdot2^{2g-2}-1)$.
2104: These non-trivial factors can not be simply removed by a different
2105: normalization of variables and therefore contain useful
2106: information. In fact in the latter case of $X_{3,2,2}(1^7)$ model,
2107: the factor of $(7\cdot2^{2g-2}-1)$ will motivate our physical
2108: explanations of the singularity in a moment.
2109: 
2110: 
2111: As shown in \cite{Katz:1999}, see also sec.
2112: \ref{symplecticinvariants}, for a fixed genus $g$, the
2113: Gopakumar-Vafa invariants $n^g_d$ are only non-vanishing when the
2114: degree $d$ is bigger than $a_g$, where $a_g$ is a model dependent
2115: number with weak genus dependence, in particular for large $g$ one
2116: has $d_{\textrm{min}}-1 = a_g \sim \sqrt{g}$. So long as the
2117: number of zeros in the low degree Gopakumar-Vafa invariants are
2118: bigger than the number of unknown coefficients that we determine
2119: above using all available boundary conditions, we have a
2120: redundancy of data to compute the topological strings recursively
2121: genus by genus, and are able to make non-trivial checks of our
2122: computations. For all of the 13 cases of one-parameter Calabi-Yau
2123: spaces, we are able to push the computation to very high genus. So
2124: far our calculations are limited only by the power of our
2125: computational facilities.
2126: 
2127: We now propose a ``phenomenological'' theory of the singularity
2128: structures at the point with rational branching. Our underlying philosophy is
2129: that a singularity of $F^{(g)}$ in the moduli space can only be
2130: generated if there are charged massless states near this point of
2131: moduli space. This is already familiar from the behaviors at
2132: infinity $\psi=\infty$ and conifold point $\psi=1$. At infinity
2133: $\psi=\infty$ the relevant charged states are massive $D2-D0$
2134: brane bound states and therefore $F^{(g)}$ are regular, whereas at
2135: the conifold point there is a massless charged state from a $D3$
2136: brane wrapping a vanishing 3-cycle, and this generates the gap
2137: like singularity at the conifold point as we have explained. We
2138: should now apply this philosophy to the much richer behavior at
2139: the orbifold point $\psi=0$. We discuss the four classes of models
2140: in the same order as mentioned above.
2141: \begin{enumerate}
2142: \item We argue for this class of models the $F^{(g)}$ are regular
2143: at the orbifold point because there is no massless charged state.
2144: A necessary condition would be that the mirror map parameter
2145: $s$ is non-zero at the orbifold point, since as we have
2146: learned there are D-branes wrapping cycles whose charge and mass
2147: are measured by $s$. This is clear for the complete
2148: intersection cases, namely $X_{3,3}(1^6)$, $X_{2,2,2,2}(1^8)$,
2149: $X_{4,4}(1^4,2^2)$, $X_{6,6}(1^2,2^2,3^2)$, because for these
2150: models the first two indices of the Picard-Fuchs equation is
2151: degenerate $a_2=a_1$, therefore the mirror map goes like
2152: \begin{eqnarray}
2153: s\sim \log{\psi}\rightarrow \infty
2154: \end{eqnarray}
2155: and we see that the D-branes are very massive and generate
2156: exponentially small corrections just like the situation at
2157: infinity $\psi=\infty$. As for the hypersurface cases $X_5(1^5)$,
2158: $X_6(1^4,2)$, $X_8(1^4,4)$, $X_{10}(1^3,2,5)$, we comment that
2159: although naively the mirror map goes like $s\sim
2160: \psi^{a_2-a_1}\rightarrow 0 $, there is a change of basis under
2161: which the generators are invariant as explained in Appendix
2162: \ref{sectioninvariance}, and which could make the periods goes
2163: like $\omega_0\sim\omega_1\sim \psi^{a_1}$, therefore the mirror
2164: map becomes finite at the orbifold point. This is also consistent
2165: with the basis at orbifold point we obtained by analytic
2166: continuation from infinity $\psi=\infty$. In fact we checked that the regularity 
2167: of $F^{(g)}$ is not affected when we take $s$ to be the ratio of generic 
2168: arbitrary linear combinations of the periods $\omega_i$, $i=0,1,2,3$.  
2169: 
2170: \item We argue this class of models have a conifold like structure
2171: because of the same mechanism we have seen for the conifold point
2172: $\psi=1$. This is consistent with the fact that the degenerate
2173: indices in these cases, e.g. models $X_{4,2}(1^6)$ and
2174: $X_{6,2}(1^5,3)$, are the middle indices, namely we have
2175: $a_2=a_3=\frac{1}{2}$. Therefore the Picard-Fuchs equation
2176: constrains one of periods to be a power series proportional to
2177: \begin{eqnarray}
2178: \omega_1\sim \psi^{\frac{1}{2}}
2179: \end{eqnarray}
2180: Since there is another period that goes like $\omega_0\sim
2181: \psi^{a_1}$, the mirror map goes like $s\sim
2182: \psi^{\frac{1}{2}-a_1}\rightarrow 0$ and it is not possible to
2183: change the basis in a way such that the mirror map is finite.
2184: Integrating out a charged nearly massless particle generates the
2185: gap like conifold singularity as we have explained.
2186: 
2187: \item For the model $X_{3,2,2}(1^7)$ we argue there are two
2188: massless states near the orbifold point. Since the middle indices
2189: also degenerate $a_2=a_3=\frac{1}{2}$, we can apply the same
2190: reasoning from the previous case and infer that the mirror map
2191: goes like $s\sim \psi^{\frac{1}{6}}$ and there must be at
2192: least one charged massless state from a D3 brane wrapping
2193: vanishing 3-cycle. However the situation is now more complicated.
2194: We do not find a gap structure in the expansion of
2195: $F^{(g)}_{\textrm{orbifold}}$, and the leading coefficients differ
2196: from the usual conifold expansion by a factor of $(7\cdot
2197: 2^{2g-2}-1)$ as observed in (\ref{twoparticles-12-01}). These can
2198: be explained by postulating that there are two massless particles
2199: in this case whose masses are $m$ and $2m$. This fits nicely with
2200: the $2^{2g-2}$ power in the intriguing factor of $(7\cdot
2201: 2^{2g-2}-1)$, and also explains the absence of gap structure by
2202: the possible interactions between the two massless particles.
2203: 
2204: \item Finally, we discuss the cases of models $X_{4,3}(1^5,2)$ and
2205: $X_{6,4}(1^3,2^2,3)$. The indices $a_i$ are not degenerate in
2206: these cases. What makes these models different from the
2207: hypersurface cases is the fact that the ratio $\frac{a_2}{a_1}$ is
2208: now not an integer in these cases. This makes it difficult to
2209: change the basis such that the mirror map is finite. We conjecture
2210: that the mirror map indeed goes like $s\sim \psi^{a_2-a_1}$ and
2211: therefore there exist massless particle(s) whose masses are
2212: proportional to $s$ and who are responsible for generating the
2213: singularity of $F^{(g)}$ at the orbifold point. This is consistent
2214: with our analytic continuation analysis in Appendix
2215: \ref{appendixD-12-06}. We note that a necessary non-trivial
2216: consequence of this scenario would be that the
2217: $F^{(g)}_{\textrm{orbifold}}$ is no more singular than
2218: $\frac{1}{s^{2g-2}}$, i.e. the product
2219: \begin{eqnarray}
2220: s^{2g-2}F^{(g)}_{\textrm{orbifold}}\sim
2221: \frac{P_g}{\psi^{(1-2a_2)(g-1)}}
2222: \end{eqnarray}
2223: should be regular. This is precisely what we observe
2224: experimentally.
2225: 
2226: 
2227: 
2228: 
2229: \end{enumerate}
2230: 
2231: 
2232: \section{Symplectic invariants at large radius}
2233: \label{symplecticinvariants} The coefficients of the large radius
2234: expansion of the $\cF^{(g)}=\lim_{{\bar t}\rightarrow \infty}
2235: F^{(g)}(t,\bar t)$  have an intriguing conjectural interpretation
2236: as symplectic invariants of $M$. First of all we have
2237: \begin{equation}
2238: \cF^{(g)}(q)=\sum_{\beta} r_\beta^{(g)} q^{\beta}\ ,
2239: \end{equation}
2240: where $r^{(g)}_\beta\in \mathbb{Q}$ are the {\sl Gromow-Witten
2241: invariants} of holomorphic maps. Secondly the {\sl Gopakumar-Vafa
2242: invariants}~\cite{GVII} count the cohomology of the $D_0-D_2$
2243: bound state moduli space, see also~\cite{Katz:1999},  and are
2244: related to the
2245: \begin{equation}
2246: \begin{array}{rcl}
2247: \cF(\lambda,t)&=&\ds{\sum_{g=0}^\infty \lambda^{2 g-2} \cF^{(g)}}(t)\\
2248: &=&\ds{{c(t)\over \lambda^2}+l(t)+\sum_{g=0}^\infty \sum_{\beta\in H_2(M,\ZZ)}\sum_{m=1}^\infty n_{\beta}^{(g)} {1\over m}
2249: \left(2 \sin {m \lambda \over 2}\right)^{2 g-2}  q^{\beta m}}\ .
2250: \end{array}
2251: \label{schwingerloopd2d0}
2252: \end{equation}
2253: Here $c(t)$ and $l(t)$ are some cubic and linear polynomials in $t$, which follow
2254: from the leading behaviour of $\cF^{(0)}$ and $\cF^{(1)}$  as explained in (\ref{prepotential})
2255: and (\ref{F1}).
2256: With $q_\lambda=e^{i\lambda}$ we can write a product form\footnote{Here we dropped the $\exp({c(t)\over \lambda^2}+l(t))$
2257: factor of the classical terms  at genus $0,1$.} for the partition function $Z^{\rm hol}=\exp(\cF^{\rm hol})$
2258: \begin{equation}
2259: Z^{\rm hol}_{\rm GV}(M,\lambda,q)=\prod_{\beta}\left[
2260: \left(\prod_{r=1}^\infty (1-q_\lambda^r q^\beta)^{r n_\beta^{(0)}}\right)
2261: \prod_{g=1}^\infty \prod_{l=0}^{2g-2}(1-q_\lambda^{ g-l-1} q^\beta)^{(-1)^{g+r}
2262: \left(2 g-2\atop l\right) n_\beta^{(g)}}\right]\
2263: \label{zhol}
2264: \end{equation}
2265: in terms of the {\sl Gopakumar-Vafa invariants} $n_\beta^{(g)}$. Based on the partition
2266: functions there is a conjectural relation of the latter to the {\sl Donaldson-Thomas invariants}
2267: $\tilde n_\beta^{(g)}$, which are invariants of the moduli space $I_k(M,\beta)$ of
2268: ideal sheaves $\cI$ on $M$.  Defining $Z^{\rm hol}_{\rm DT}(M,q_{\lambda},q)=
2269: \sum_{\beta,k\in {\ZZ}} {\tilde n}^{(k)}_{\beta} q_\lambda^k q^\beta$ one expects~\cite{DTGW}
2270: \begin{equation}
2271: Z^{\rm hol}_{\rm GV}(M,q_{\lambda},q)M(q_\lambda)^{\frac{\chi(M)}{2}}=Z^{\rm hol}_{\rm DT}(M,-q_{\lambda},q)\ ,
2272: \label{eq:GVDT}
2273: \end{equation}
2274: where the McMahon function  is defined as $M(q_\lambda):=\prod_{n\ge 0} {1\over (1-q_\lambda^n)^n}$.
2275: We will give below the information about the ${\cal F}^{(g)}$ in terms of the
2276: {\sl Gopakumar-Vafa invariants} and give a  more detailed account of the data
2277: of  the symplectic invariants on the webpage \cite{webpage}.
2278: 
2279: 
2280: \begin{table}
2281: \begin{centering}
2282: \begin{tabular}{|r|rrrrrr|}
2283: \hline
2284: g &d=1 &d=2 &d=3 &d=4 &d=5 &d=6  \\
2285: \hline
2286: \, 0& 2875& 609250& 317206375& 242467530000& \!229305888887625&\! 248249742118022000 \\
2287: 1& 0    &     0&    609250&   3721431625&  12129909700200&  31147299733286500 \\
2288: 2& 0    &     0&         0&       534750&     75478987900&    871708139638250 \\
2289: 3& 0    &     0&         0&         8625&       -15663750&     3156446162875  \\
2290: 4& 0    &     0&         0&            0&           49250&       -7529331750 \\
2291: 5& 0    &     0&         0&            0&            1100&         -3079125 \\
2292: 6& 0    &     0&         0&            0&              10&            -34500 \\
2293: 7& 0    &     0&         0&            0&               0&                0  \\
2294: \hline
2295: \end{tabular}
2296: \vskip 2 pt
2297: \begin{tabular}{|r|rrr|}
2298: \hline
2299: g &d=7 & d=8 &d=9 \\
2300: \hline
2301: 0& \!\!\! 295091050570845659250& \!\!\!\!\!\! 375632160937476603550000& \!\!\!\!\!\! 503840510416985243645106250\\
2302: 1& \!\!\! 71578406022880761750& \!\!\!\!\!\! 154990541752961568418125& \!\!\!\!\!\! 324064464310279585657008750 \\
2303: 2& 5185462556617269625& 22516841063105917766750& 81464921786839566502560125 \\
2304: 3& 111468926053022750& 1303464598408583455000& 9523213659169217568991500 \\
2305: 4& 245477430615250& 25517502254834226750& 507723496514433561498250 \\
2306: 5& -1917984531500& 46569889619570625& 10280743594493108319750 \\
2307: 6& 1300955250& -471852100909500& 30884164195870217250 \\
2308: 7& 4874000& 2876330661125&-135197508177440750 \\
2309: 8& 0& -1670397000& 1937652290971125 \\
2310: 9& 0& -6092500& -12735865055000\\
2311: 10& 0& 0& 18763368375\\
2312: 11& 0& 0& 5502750\\
2313: 12& 0& 0& 60375\\
2314: 13& 0& 0&  0\\
2315: \hline
2316: \end{tabular}
2317: \begin{tabular}{|r|rr|}
2318: \hline
2319: g &d=10 & d=11 \\
2320: \hline
2321: 0& \quad \,\,\,\,704288164978454686113488249750&  \quad \,\,\,\,\,   1017913203569692432490203659468875\\
2322: 1& 662863774391414096742406576300& 1336442091735463067608016312923750 \\
2323: 2& 261910639528673259095545137450& 775720627148503750199049691449750 \\
2324: 3& 52939966189791662442040406825 & 245749672908222069999611527634750 \\
2325: 4& 5646690223118638682929856600  &44847555720065830716840300475375\\
2326: 5& 302653046360802682731297875   & 4695086609484491386537177620000\\
2327: 6& 6948750094748611384962730     &267789764216841760168691381625\\
2328: 7& 40179519996158239076800       &7357099242952070238708870000\\
2329: 8& -25301032766083303150         &72742651599368002897701250\\
2330: 9& 1155593062739271425           &140965985795732693440000\\
2331: 10& -17976209529424700           &722850712031170092000 \\
2332: 11& 150444095741780              &-18998955257482171250\\
2333: 12& -454092663150                & 353650228902738500 \\
2334: 13&   50530375                   &-4041708780324500\\
2335: 14&  -286650                     &22562306494375\\
2336: 15&   -5700                      &-29938013250\\
2337: 16&    -50                       & -7357125\\
2338: 17&     0                        &  -86250\\
2339: 18&     0                        &      0\\
2340: \hline
2341: \end{tabular}
2342: \caption{BPS invariants $n^g_d$ on the Quintic hypersurface in $\mathbb{P}^4$. See also Table 3.}
2343: \end{centering}
2344: \end{table}
2345: 
2346: 
2347: 
2348: \subsection{Castelnuovo's theory and the cohomology of the BPS state moduli space}
2349: 
2350: Let us give checks of the numbers using techniques of algebraic
2351: geometry and the description of the BPS moduli space and its
2352: cohomology developed in \cite{GVII,Katz:1999}. The aim is to check
2353: the gap condition in various geometric settings, namely
2354: hypersurfaces and complete intersections in (weighted) projective
2355: spaces discussed before. According to \cite{GVII,Katz:1999} the
2356: BPS number of a given charge, i.e. degree $d$, can be calculated
2357: from cohomology of the moduli  space $\hat {\cal M}$ of a
2358: $D_2-D_0$ brane system. The latter is the fibration of the
2359: Jacobian $T^{2 \tilde g}$ of a genus ${\tilde g}$ curve over its
2360: moduli space of deformations ${\cal M}$ . Curves of arithmetic
2361: genus $g<{\tilde g}$ are degenerate curves, in the simplest case
2362: with $\delta={\tilde g}-g$ nodes. Their BPS numbers are calculated
2363: using the Euler numbers of relative Hilbert schemes ${\cal
2364: C}^{(i)}$ of the universal curve (${\cal C}^{(0)}={\cal M}$,
2365: ${\cal C}^{(1)}$ is the universal curve, etc) in simple situations
2366: as follows:
2367: \begin{equation}
2368: \begin{array}{rl}
2369: n_d^{g}&={\displaystyle n_d^{{\tilde g} -\delta}=(-1)^{{\rm
2370: dim}({\cal M})+\delta}\sum_{p=0}^\delta b( {\tilde g}-p,\delta-p)
2371: e({\cal C}^{(p)})},
2372: \\[ 3 mm]
2373: b(g,k)&={2\over k!} (g-1)\prod_{i=1}^{k-1} (2 g - (k+2) +i),\qquad   b(g,0)=0.
2374: \end{array}
2375: \label{bpsformula}
2376: \end{equation}
2377: 
2378: As explained in \cite{Katz:1999} curves in  projective spaces
2379: meeting the  quintic are either plane curves in $\mathbb{P}^2$,
2380: curves in $\mathbb{P}^3$, or $\mathbb{P}^4$. In all case one gets
2381: from Castelnuovo theory a bound on $g$, which grows for large $d$
2382: like  $g(d)\sim d^2$. For a detailed exposition of curves in
2383: projective space see~\cite{Harris}. Using this information one can
2384: determine which curves above is realized and contributes to the
2385: BPS numbers. These statements generalize to the  hypersurfaces and
2386: complete intersections with one K\"ahler modulus in weighted
2387: projective spaces. In particular the qualitative feature $g(d)\sim
2388: d^2$ of the bound for large $d$ carries over. We note for
2389: later convenience that to go from a smooth curve of genus $\tilde
2390: g$ to a curve with arithmetic genus $g=\tilde g-\delta$ by
2391: enforcing $\delta$ nodes we get from (\ref{bpsformula})
2392: \begin{equation}
2393: \begin{array}{rl}
2394: n^{\tilde g-1}_d&=(-1)^{{\rm dim}({\cal M})+1}\left( e({\cal C})+ (2\tilde g-2) e(\cal M)\right)\\
2395: n^{\tilde g-2}_d&=(-1)^{{\rm dim}({\cal M})+1}\left(e({\cal C}^{(2)})+(2 \tilde g-4) e({\cal C})+{1\over 2}(2 \tilde g-2) (2 \tilde g-5)e(\cal M)\right)\ . \\
2396: \end{array}
2397: \label{BPSspecial}
2398: \end{equation}
2399: 
2400: %----------------------------------------------------------------------
2401: \subsection{D-branes on the quintic}
2402: \label{quinticdbrane}
2403: One consequence of our global understanding of the $F^{(g)}$
2404: is that we can make detailed statements about the `number' of
2405: $D$-brane states for the quintic at large radius.
2406:  We focus on $d=5$, because
2407: there is a small numerical flaw in the analysis of
2408: \cite{Katz:1999}, while the right numerics confirms the gap
2409: structure quite significantly. In this case the complete
2410: intersection with multidegree $(1,1,5)$ is a plane curve with
2411: genus $\tilde g=(d-1)(d-2)/2=6$, while the other possibilities
2412: have at most genus $g=2$. Curves of $(g=\tilde g,d)=(6,5)$ are
2413: therefore smooth plane curves with $\delta=0$ and according to
2414: (\ref{bpsformula}) their BPS number is simply $n^g_d=(-1)^{{\rm
2415: dim} \cal M} e({\cal M})$. Since $d=5$ their moduli space is
2416: simply the moduli space of $\mathbb{P}^2$'s in $\mathbb{P}^4$,
2417: ${\cal M}$ is the Grassmannian\footnote{The space of
2418: $\mathbb{P}^k$'s in  $\mathbb{P}^n$ , which we call
2419: ${\mathbb{G}}(k,n)$, is also the space of $k+1$ complex
2420: dimensional subspaces in an $n+1$ dimensional complex vector
2421: space, which is often alternatively denote as $G(k+1,n+1)$.}
2422: ${\mathbb{G}}(2,4)$. Grassmannians ${\mathbb{G}}(k,n)$ have
2423: dimensions $(k+1)(n-k)$ and their Euler number can be calculated
2424: most easily by counting toric fixed points to be
2425: $\chi({\mathbb{G}}(k,n))=\left(n+1\atop k+1\right)$.  We get
2426: $n^6_5=(-1)^6 10=10$.
2427: 
2428: For the $(g,d)=(5,5)$ curves we have to determine the Euler number
2429: of the universal curve ${\cal C}$, which is a fibration $\pi:{\cal
2430: C} \rightarrow {\cal M}$ over ${\cal M}$. To get  an geometric
2431: model for ${\cal C}$ we consider the projection $\tilde \pi:{\cal
2432: C}\rightarrow X$. The fiber over a point $p\in X$ is the  set of
2433: $\IP^2$'s in $\IP^4$ which contain the point $p$. This is described
2434: as the space of $\mathbb{P}^1$'s in $\mathbb{P}^3$ i.e.
2435: ${\mathbb{G}}(1,3)$ with~\footnote{${\mathbb{G}}(1,3)$ is Plucker
2436: embedded in $\mathbb{P}^5$ as a quadric (degree 2). From the
2437: adjunction formula we also  get $\chi({\mathbb{G}}(1,3))=6$.}
2438: $\chi({\mathbb{G}}(1,3))=6$. As the fibration $\tilde \pi$ is
2439: smooth we obtain $e({\cal C})=\chi(X)
2440: \chi({\mathbb{G}}(1,3))=-200\cdot 6=-1200$. Applying  now
2441: (\ref{BPSspecial}) we get $n^{5}_5=(-1)^5(-1200+ (2 \cdot 6-2)
2442: 10)=1100$.
2443: 
2444: 
2445: The calculation of $n^4_4$ requires the  calculation  of  $e({\cal C}^{(2)})$. The model
2446: for ${\cal C}^{(2)}$ is constructed from the fibration  $\hat \pi:{\cal C}^{(2)}\rightarrow {\rm Hilb}^2(X)$
2447: as follows. A point in ${\rm Hilb}^2(X)$ are either two distinct points or one point of multiplicity $2$ with  distinct
2448: tangent direction.  In both cases the
2449: fiber over $P \in {\rm Hilb}^2(X)$ is an $\mathbb{P}^2$ passing though $2$ points in $\mathbb{P}^4$,
2450: which is a  $\mathbb{P}^2$. The fibration is smooth and it remains to calculate the Euler number
2451: of the basis.
2452: %Using the formula
2453: %\begin{equation}
2454: %\sum_{n=0}^\infty e({\rm Sym}^n X) q^n=\prod_{i=1}^\infty \left({1\over 1- q^n} \right)^{e(M)}
2455: %\end{equation}
2456: There are nice product formulas for the Euler number of symmetric
2457: products of surfaces modded out by $S_n$. For surfaces it is more
2458: cumbersome. We calculate the Euler number $e({\rm
2459: Sym}^2(X))=\left(-199\atop 2\right)$. ${\rm Hilb}^2(X)$ is the
2460: resolution of the  orbifold  ${\rm Sym}^2(X)$, which has the
2461: diagonal $X$ as fix point set. The resolution replaces each point
2462: in the fixed point set by $\mathbb{P}^2$. Simple surgery and  the
2463: smooth fibration structure of  ${\cal C}^{(2)}$ gives hence
2464: $e({\cal C}^{(2)})=3 (e({\rm Sym}^2(X)+(3-1)e(X))=58500,$ which by
2465: (\ref{BPSspecial}) yields  $n^{4}_5=58500+(2\cdot 6-4)(-1200)+35
2466: \cdot 10=49250$.
2467: 
2468: The approach becomes more difficult with the number of free points
2469: $\delta$ and at $\delta=4$ it is currently not know how to treat
2470: the singularities of the Hilbert scheme.
2471: 
2472: On the other hand smooth curves at the `edge' of the  Castelnuovo
2473: bound are of no principal problem. E.g., using adjunction for a
2474: smooth complete intersection of degree $(d_1,\ldots,d_r$) in a
2475: (weighted) projective space $\mathbb{WCP}^n(w_1,\ldots,w_{n+1})$
2476: in Appendix A, we calculate $\chi=(2-2g)$ and see that the degree
2477: 10 genus 16 curve is the complete intersection  $(1,2,5)$. The
2478: moduli space is calculated by counting the independent
2479: deformations of that complete intersection. The degree five
2480: constraint lies on the quintic, the linear constraint has five
2481: parameters. The identification by the $\mathbb{C}^*$ action of the
2482: ambient $\mathbb{P}^4$ shows that these parameters lie in a
2483: $\mathbb{P}^4$. This constraint allows to eliminate one variable
2484: from the generic quadratic constraint  which has hence $10$
2485: parameters and a $\mathbb{P}^9$ as moduli space. So we check in
2486: Tab. 2 the entry $n^{16}_{10}=(-1)^{13} 5\cdot 10=-50$.
2487: 
2488: Let us discuss the upper bound on the genus at which we can
2489: completely completely fix the $F_g$ given simply the bound
2490: (\ref{boundarycountquintic}). We claim that this bound is  $g\le
2491: 51$. At degree $20$ there is a smooth complete intersection curve
2492: $(1,4,5)$ of that genus. We first check that this is the curve of
2493: maximal genus in degree $20$. The Castelnuovos bound for curves in
2494: $\mathbb{P}^4$ shows that they have smaller genus~\cite{Harris}.
2495: We further see from the discussion in~\cite{Harris} that for
2496: curves in $\mathbb{P}^3$ not on quadric and a cubic, which would
2497: have the wrong degree, the Castelnuovos's bound is saturated for
2498: the complete intersection  $(1,4,5)$. For $g=51$
2499: (\ref{boundarycountquintic}) indicates that the gap, constant map
2500: contribution  and regularity at the orbifold fixes $131$ of the
2501: $151$ unknown coefficients in  (\ref{10-14-2.16}). The vanishing
2502: of $n^{51}_{d}=0$, $1\le d\le 19$ and the value of
2503: $n^{51}_{20}=(-1)^{ 4+ 34}
2504: \chi(\mathbb{P}^4)\chi(\mathbb{P}^{34}) =165$ for the Euler number
2505: of the moduli space of the smooth curve give us the rest of the
2506: data.
2507: 
2508: \parbox{14cm} % Seitenbreite
2509: {
2510:    %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2511:    %% Abbildung contour  %%%%%%%%%%%%%%%%%%%%%%
2512:    %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2513: 
2514:    \begin{center}
2515:    \mbox{
2516:          % \begin{turn}{-90}  % Hier k"onnen Sie das Bild drehen
2517:                              % z.B. um -90 Grad
2518:              \epsfig{file=castelnouvo.eps,width=14cm}
2519:          % \end{turn}
2520: }
2521:    \end{center}
2522:    %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2523: }
2524: %
2525: \vskip -2mm\noindent {\em Castelnuovo's bound for higher genus
2526: curves on the quintic. The dashed line correspond roughly (up to taking the floor) 
2527: to the number of coefficients in $f_g$ (\ref{10-14-2.16}) which are not fixed by
2528: constant map contribution, conifold and orbifold boundary
2529: conditions.}
2530: % nachfolgender Text
2531: \vskip 2mm
2532: 
2533: 
2534: It is of course no problem to calculate  form the B-model the
2535: higher genus amplitudes to arbitrary degree. For completeness we
2536: report the first nontrivial numbers for $g=18-20$ in Tab. 3.
2537: 
2538: \begin{table}
2539: \begin{centering}
2540: \begin{tabular}{|r|rr|}
2541: \hline
2542: d &g=18 &g=19\\
2543: \hline
2544: \vdots& \vdots   &   \vdots \\
2545: 11& 0 &  0  \\
2546: 12&-3937166500  & -13403500  \\
2547: 13& 285683687197594125   &   -2578098061480250            \\
2548: 14&  -95076957496873268057250  &    2730012506820193210000            \\
2549: 15&  6438165666769014564325336250  &            -342304337102629200272769700   \\
2550: 16&  15209062594213864261318125134875  &    15209062594213864261318125134875           \\
2551: \hline
2552: \end{tabular}
2553: 
2554: \begin{tabular}{|r|r|}
2555: \hline
2556: d &g=20 \\
2557: \hline
2558: \vdots& \vdots \\
2559: 11& 0 \\
2560: 12& 0 \\
2561: 13& 10690009494250 \\
2562: 14& -59205862559233156250 \\
2563: 15& 15368486208424999875838025 \\
2564: 16& -1036824730393980503709247290500 \\
2565: \hline
2566: \end{tabular}
2567: \caption{Some higher degree genus 18-20 BPS numbers for the quintic. Note
2568: that we can calculate all Donaldson-Thomas invariants for $d=1,\ldots,12$
2569: exactly.}
2570: \end{centering}
2571: \end{table}
2572: 
2573: \subsection{D-brane states on hypersurfaces in weighted projective space}
2574: 
2575: Similarly, for the sextic in $\mathbb{P}^4(1^4,2)$, the degree
2576: $(1,2,6)$ complete intersection curve has genus $g=10$ and degree
2577: $d={\prod_i d_i\over \prod_k w_k}=6$ in the weighted projective
2578: space . Its moduli space is $\mathbb{P}^3$ for the degree one
2579: constraint, i.e., we can eliminate $x_4$ form the quadric and the 7
2580: coefficeints of the monomials $x_1^2, x_1x_2, x_1x_3, x_2^2, x_2
2581: x_3 ,x_3^2,x_5$ form a $\mathbb{P}^6$. This yields
2582: $n^{10}_6=(-1)^9 4 \cdot 7=-28$.
2583: 
2584: There are further checks for the octic in $\mathbb{P}^4(1^4,4)$ BPS
2585: invariants. The complete intersection $(1^2,8)$
2586: has total degree $2$ and genus $g=3$. The two linear constraints
2587: describe a $\mathbb{P}^1$ in $\mathbb{P}^3$ i.e. an
2588: $\mathbb{G}(1,3)$ with Euler number $6$ and dimension $4$, which
2589: yields $n^3_2=6$. Similarly we have a $g=7$ complete intersection
2590: $(1,2,8)$ of degree $4$, whose moduli space is  $\mathbb{P}^3$
2591: times $\mathbb{P}^5$ hence $n^7_4=24$.
2592: 
2593: For the degree 10 hypersurface in $\mathbb{P}^4(1^3,2,5)$ we check the BPS invariants:
2594: {}From the degree $(1,1,10)$ hypersuface of degree $1$ complete intersection with $g=2$. The moduli
2595: space of the linear constraints are just the one of  point in
2596: $\mathbb{P}^2$, i.e.$\mathbb{P}^2$, hence $n^2_1=3$. The degree
2597: $(1,2,10)$ complete intersection with total degree $2$ and genus
2598: $4$ has the moduli space of the linear constraint, which is
2599: $\mathbb{P}^2$ and of the quadratic constraint is $\mathbb{P}^3$
2600: (from the coefficients of the monomials $x_1^2,x_1 x_2,
2601: x^2_2,x_4$), yielding $n_2^4=- 12$. Finally  the $(1,3,10)$
2602: complete intersection with genus $7$ and degree $3$, has a moduli
2603: space $\mathbb{P}^2$ times $\mathbb{P}^5$ (from the coefficients
2604: of the monomials $x_1^3,x_1^2 x_2, x_1 x^2_2,x_1 x_4,x_2^3 ,x_2
2605: x_4$) and $n_3^7=- 18$.
2606: 
2607: These checks in different geometrical situations
2608: establish quite impressively the universality
2609: of the gap structure at the conifold expansion.
2610: 
2611: 
2612: 
2613: \subsection{D-branes on complete intersections}
2614: Here we summarize our results on one modulus complete
2615: intersections in (weighted) projective space. More complete
2616: results are available in \cite{webpage}. Again we can check many
2617: BPS invariants associated to the smooth curves.
2618: 
2619: Let us check e.g.in table 7 the $n_9^{10}=15$. According
2620: to (\ref{chi}) we see that at degree $9$ there is a smooth genus
2621: $10$ curve, given by a complete intersection of multi degree $(1^2,3^2)$ in
2622: $\mathbb{P}^5$. Their moduli space is the Grassmannian
2623: $\mathbb{G}(3,5)$, which has Euler number $15$ and dimension $8$,
2624: hence  $n_9^{10}=15$. In a very  similar way it can be seen that
2625: the $n_8^9$ comes form a complete intersection curve of degree
2626:  $(4,2,1^2)$ with the same moduli
2627: space, so $n_8^9=15$ in table 8. Grassmannians related to complete
2628: intersection are also identified with the moduli spaces of the
2629: following smooth curves: The total degree six curve of genus seven
2630: in table 11 is a CI of multi degree $(1^2,3,4)$. Its moduli space
2631: is a $\mathbb{G}(2,4)$ explaining $n_6^7=10$. The degree four
2632: curve of genus five in table 12 is of multi degree $(1^2,4^2)$ and
2633: has moduli space $\mathbb{G}(1,3)$ yielding $n_4^5=6$.  The degree
2634: four curve of genus five in table 13 is of multi degree
2635: $(1^2,2,6)$ and has moduli space $\mathbb{G}(2,4)$ yielding
2636: $n_4^5=10$.  The degree two curve of genus three in table 14 is of
2637: multi degree $(1^2,4,6)$ and has moduli space
2638: $\mathbb{G}(0,2)=\mathbb{P}^2$ yielding $n_2^3=3$. The moduli
2639: space of the degree $4$ genus seven curve $(1,2,4,6)$ is an
2640: $\mathbb{P}^2$ times the moduli space $\mathbb{P}^4$ of quadrics
2641: in $\mathbb{WCP}^3(1^2,2^2)$, so that $n_4^7=(-1)^6 3\cdot 5=15$.
2642: For the degree $(6,6)$ complete intersection in
2643: $\mathbb{WCP}^3(1^2,2^2,3^2)$, see table 15, we have a degree one
2644: genus two intersection $(1^2,6^2)$, whose moduli space is a point
2645: hence $n_1^2=1$, a degree two genus four intersection $(1,2,6^2)$,
2646: whose moduli space is $\mathbb{P}^1$ times the moduli space
2647: $\mathbb{P}^2$ of quadrics in $\mathbb{WCP}^2(1,2^2)$ hence
2648: $n_2^4=-6$, a degree three  genus seven intersection $(1,3,6^2)$,
2649: whose moduli space is $\mathbb{P}^1$ times the moduli space
2650: $\mathbb{P}^4$ of cubics in $\mathbb{WCP}^4(1,2^2,3^2)$ hence
2651: $n_3^7=-10$ and finally  a degree four genus eleven intersection
2652: $(1,4,6^2)$, whose moduli space is  $\mathbb{P}^1$ times the
2653: moduli space $\mathbb{P}^5$ of quadrics in
2654: $\mathbb{WCP}^4(1,2^2,3^2)$ hence $n_4^{11}=12$.
2655: 
2656: 
2657: There are many further checks that are somewhat harder to perform.
2658: E.g. we notice that there is a genus one degree three curve in the
2659: $(3,2,2)$ CI in $\mathbb{P}^6$, which comes from a complete
2660: intersection $(1^4,3)$. Now the moduli space of this complete
2661: intersection in  $\mathbb{P}^6$ is $\mathbb{G}(2,6)$. However not
2662: all $\mathbb{P}^2$ parametrized by $\mathbb{G}(2,6)$, which
2663: contain the cubic, are actually in the two quadrics of the
2664: $(3,2,2)$ CICY. We can restrict to those $\mathbb{P}^2$, which
2665: fulfill these constraints, by considering the simultaneous zeros
2666: of sections of two rank six bundles of quadratic forms on the
2667: moving $\mathbb{P}^2$. These are a number of points, which is
2668: calculated by the integral of the product of the Chern classes of
2669: these rank 6 bundles over  $\mathbb{G}(2,6)$. Indeed we obtain,
2670: for example with ``Schubert'' \cite{Schubert}
2671: \begin{equation}
2672: n_3^1=(-1)^0 \int_{\mathbb{G}(2,6)} c^2_6({\rm Sym}(2,Q))=64,
2673: \end{equation}
2674: which confirms the corresponding entry in table 9.
2675: 
2676: 
2677: 
2678: \section{Conclusions}
2679: 
2680: In this paper we solve the topological string B-model on compact
2681: Calabi-Yau $M$ using the modularity of the $F_g$,
2682: the wave function transformation property of $Z$ and the boundary
2683: information imposed by effective action considerations. The method
2684: pushes the calculation to unprecedented high genus amplitudes.
2685: E.g. for the quintic the boundary condition count
2686: (\ref{boundarycountquintic}) together with the simplest vanishing
2687: arguments at large volume fixes the amplitudes up to genus $g=51$.
2688: Beyond that the  prime mathematical problem to overcome in this
2689: region of the moduli space is understand the degeneration of more
2690: then four points in the relative Hilbert scheme of the universal
2691: curves in a threefold\footnote{As a motivation and check for the
2692: task to develop the  theory of Hilbert schemes for 3folds we
2693: calcultated the invariants explicitly to high genus. For the
2694: quintic to genus 20 and for all other the results up genus 12 are
2695: available at \cite{webpage}.}. Similar problems have been
2696: encountered in~\cite{Gaiotto:2006wm}, where it was suggested to
2697: fix a very similar ambiguity to a anholomorphic ${\rm
2698: SL}(2,\mathbb{Z})$-modular elliptic index of a $D4-D2-D0$ brane
2699: system~\cite{Gaiotto:2006wm,deBoer:2006vg,DennefMoore}. There one  uses
2700: ${\rm SL}(2,\mathbb{Z})$ invariance of the index and a dual dilute gas
2701: approximation in $AdS_3\times S^2\times M$ to fix the coeffcients 
2702: of the ring of modular forms. The construction of the moduli-space of the  
2703: $D4-D2-D0$ brane system uses rational GW invariants and implies non-trivial 
2704: relations among them~\cite{Gaiotto:2006wm}. Such considerations 
2705: could in principal provide further boundary conditions 
2706: at large radius.
2707: 
2708: 
2709: Our sharpest tool is the global control of $Z$ over ${\cal
2710: M}(M)$ and we  expect that by a closer analysis of the RR-spectrum
2711: at the orbifold of compact Calabi-Yau, we will be able to recover
2712: at least the $\lceil \frac{2}{d}(g-1)\rceil$ conditions that one
2713: loses relative to the local cases~\cite{Huang2} and solve the
2714: model completely. We obtained not only the Gromov-Witten, the
2715: Donaldson-Thomas and Gopakumar-Vafa invariants at infinity, but
2716: also the local expansion at the conifold, the Gepner point and
2717: other more exotic singularities with one or more massless states.
2718: The leading singular terms in the effective action reflect the
2719: massless states. The branch locus of the 13 parameter 
2720: models has an intriguing variety of such light spectra and 
2721: we can learn from the effective action about the singularity 
2722: and vice versa. Stability properties of theses states have 
2723: been analysed in App. \ref{appendixD-12-06}.
2724: 
2725: 
2726: Most importantly our exact expansions do contain further
2727: detailed information of the towers massive RR-states at these
2728: points. We described them in natural local variables. The 
2729: information from different genera should be of great value for the
2730: study of stable even D-brane bound states on compact Calabi-Yau as
2731: it is the content of the supersymmetric index of~\cite{GVII},
2732: which is protected under deformations of the complex structure.
2733: Non-compact Calabi-Yau such as the resolution of
2734: $\mathbb{C}^n/G$, with $G\in {\rm SL}(n,\mathbb{C})$ have
2735: no complex moduli. The issue does not arise and the situation is
2736: better understood, see e.g. \cite{Bridgeland,Aspinwall} for reviews.
2737: 
2738: One can also use the  explicit expansions  to study  the
2739: integrable theories that have been associated to the local
2740: expansion of the topological string on Calabi-Yau manifold, such
2741: as the $c=1$ string at the conifold or the quiver gauge theories
2742: at the orbifold, matrix models e.g. at the ADE singularities and new 
2743: ones for more exotic singularities such as the  
2744: branch points of the complete intersections Calabi-Yau 
2745: manifolds that we discussed here.
2746: 
2747: 
2748: The ability to obtain the imprint of the BPS spectrum on the
2749: effective action everywhere on the moduli space is of
2750: phenomenological interest as flux compactifications drive the
2751: theory to attractor points inside the moduli space.
2752: 
2753: Our expression are governed by the representation of the modular
2754: group of the Calabi-Yau on almost holomorphic forms, which we
2755: explicitly constructed from the periods, without having much of an
2756: independent theory about them. The simpler case of the torus
2757: suggest that such forms and their extensions should play a role
2758: in the study of of virtually any physical amplitude --- open or closed--- 
2759: in compactifications on the Calabi-Yau space, even as conjectured 
2760: in the hypermultiplet sector~\cite{Rocek:2005ij}.
2761: 
2762: One may finally wonder whether the topological string B-model is
2763: an integrable theory that is genuinely associated to this new and
2764: barely explored class of modular forms on Calabi-Yau spaces moduli
2765: spaces, whereas most known integrable models are associated to
2766: abelian varieties. As it was noted
2767: in~\cite{Verlinde:2004ck,Aganagic:2006wq, Gunaydin:2006bz}~in the
2768: complex moduli space extended by the dilaton, called extended
2769: phase space, one has one has rigid special K\"ahler geometry and
2770: many aspects of the sympletic transformations and its metaplectic
2771: realization are easier understood in the extended phase space.
2772: There are two maps $\Phi^{(i)}: {\cal M}\rightarrow T^{(k)}_{IJ}$,
2773: $I,J=1,\ldots,\frac{h^3}{2}$ from the complex moduli space to
2774: tensors in the extended phase space on which
2775: $\left(\begin{array}{cc} A&B\\ C&D \end{array}\right)\in {\rm Sp}
2776: (h^3,\mathbb{Z})$ acts projectively like $T^{(k)}\mapsto (A
2777: T^{(k)}+B)(C T^{k}+D)^{-1}$. For the holomorphic object
2778: $\tau_{IJ}=\partial_I \partial_J F^{(0)}=:T^{(1)}_{IJ}(t)$, which is 
2779: mostly discussed in this
2780: context of the metaplectic transfomations~\cite{Verlinde:2004ck,
2781: Aganagic:2006wq,Gunaydin:2006bz}, ${\rm Im}(\tau)$ is indefinite,
2782: while for the non-holomorphic object ${\cal N}_{IJ}={\bar
2783: \tau}_{IJ}+ 2 i \frac{{\rm Im}\tau_{IK} X^K {\rm Im}\tau_{IL}
2784: X^L}{X^L{\rm Im}\tau_{KL} X^L} =:T^{(2)}_{IJ}(t,\bar t)$ comes
2785: from the kinetic term in the 10d action whose reduction involves
2786: the Hodge-star on $M$. It's imaginary  part ${\rm Im}({\cal N})>0$
2787: is the kinetic term for the vector multiplets and is  hence
2788: positive definite. In other words ${\rm Im}\Phi^{(2)}$ defines a
2789: map to the Siegel upper space. $\Phi^{(2)}$ should relate Siegel
2790: modular forms for admittedly very exotic subgroups~\cite{cyy} of
2791: ${\rm Sp}(4,\mathbb{Z})$ to Calabi-Yau amplitudes. Such Siegel
2792: modular forms for abelian varieties  are also associated to $N=2$
2793: Seiberg-Witten (gauge) theories, while the modular forms on
2794: Calabi-Yau studied in this paper underline $N=2$ exact terms in
2795: $N=2$ supergravity. The map $\Phi^{(2)}$ could be a manifestation
2796: of a gravity-gauge theory correspondence for 4d theories with
2797: $N=2$ supersymmetry.
2798: 
2799: 
2800: It is no principal problem to generalize this to multi-moduli Calabi-Yau as
2801: long as the Picard-Fuchs equations are known. These have different, more
2802: general singularities with interesting local effective actions. In
2803: K3 fibrations which have at least two moduli, the modular properties are much
2804: better understood and in fact the ambiguity in the fiber is
2805: complete fixed heterotic string calculations. Moreover these cases
2806: have $N=2$ field theory limits, which contain further information,
2807: which might be sufficient to solve these models \cite{GKMW}.
2808: 
2809: 
2810: 
2811: 
2812: %----------------------------------------------------------------------------------------
2813: 
2814: 
2815: \vspace{0.2in} {\leftline {\bf Acknowledgments:}}
2816: 
2817: We thank M.~Aganagic, V.~Bouchard, T.~Grimm, S.~Katz, M.~Kontsevich, M.~Marino, C.~Vafa,
2818: S.~T.~Yau and D.~Zagier for discussions. Sheldon Katz helped us with
2819: the verifications of the BPS numbers and Cumrun Vafa with remarks
2820: on the draft. Don Zagier's comments on~\cite{Huang} triggered many ideas here.
2821: We thank the MSRI in Berkeley and AK thanks in particular the
2822: Simons Professorship Program. MH/AK thank also the Simons Workshops
2823: in Mathematics and Physics 05/06 for its hospitality.
2824: 
2825: \appendix
2826: 
2827: \section{Appendices}
2828: %----------------------------------------------------------------------------------
2829: \subsection{Classical intersection calculations using the adjunction
2830: formula}
2831: \label{intersection}
2832: 
2833: The adjunction formula\footnote{See \cite{fulton} for a
2834: pedagogical account of these matters.} for the total Chern class
2835: of a  for dimension $m=n-r$ smooth complete intersections  $M$ of
2836: multi degree  $d_1,\ldots,d_r$ in a weighted projective space
2837: $\mathbb{WCP}^n(w_1,\ldots,w_{n+1})$ is
2838: \begin{equation}
2839: c(T_M)=\sum_i c_i(T_M)=
2840: {c(T_{\mathbb{WCP}})\over c({\cal N})}= {\prod_{i=1}^{n+1} (1+w_i K)\over \prod_{k=1}^r( 1+ d_k K)}=\sum_i c_i K^i\ ,
2841: \label{adjunction}
2842: \end{equation}
2843: where   $c(T_{\mathbb{WCP}})=\sum_i c_i(T_{\mathbb{WCP}})=\prod_{i=1}^{n+1} (1+w_i K)$
2844: is the total Chern class  of the weighted projective space, $K$ is its K\"ahler class and
2845: $c({\cal N})=\prod_{k=1}^r( 1+ d_k K)$ is  the  total Chern class of the normal bundle.
2846: 
2847: Integration of a top form $\omega=x J^m$ with $J=K|_M$ over $M$ is obtained by integration
2848: along the normal direction as
2849: \begin{equation}
2850: \int_{M} \omega=\int_{\mathbb{WCP}} \omega \wedge c_r({\cal N}) = {x\over \prod_{k=1}^{n+1} w_k} \prod_{k=1}^r d_k \ .
2851: \label{integrationalg}
2852: \end{equation}
2853: Here we used the normalization  $\int_{\mathbb{WCP}} K^n= {1\over \prod_{k=1}^{n+1} w_k}$.
2854: This  yields the first line below:
2855: \begin{eqnarray}
2856: \kappa&=&\int_{M} J^m={\prod_{k=1}^r d_k\over \prod_{i=1}^{r+1} w_i} \label{kappa}\\
2857: \chi  &=& \int_M c_3(T_M)={c_3\over \prod_{k=1}^{n+1} w_k} \prod_{k=1}^r d_k \label{chi}\\
2858:   c  &=& {1\over 24} \int_M c_2\wedge J={1\over 24} {c_2\over \prod_{k=1}^{n+1} w_k} \prod_{k=1}^r d_k \label{cconst}\\
2859:   a  &=& {1\over 2}  \int_M i_* c_1(D)\wedge J= {1 \over 2}
2860: \int_{\mathbb{WCP}} {c(T_M)\over (1+J)}\wedge J^{r+1}={{\rm coeff} \left({c(T_M)\over (1+J)},J^{m-1}\right)
2861: \over 2\cdot \prod_{k=1}^{n+1} w_k }\label{aconst}
2862: \end{eqnarray}
2863: Combining (\ref{adjunction},\ref{integrationalg}) one gets the line 2
2864: and 3. The leading $t$ terms in $F_0$ can be obtained by
2865: calculating $Z(M)$  using (\ref{dbranecharge}), while the last
2866: line follows from the calculation of $Z(D)$ assuming that the
2867: $D_4$-brane is supported on $D$ the restriction of the hyperplane
2868: class\footnote{$a$ is physically less relevant, as it does not
2869: affect the effective action. Its value $a={11\over 2}$ obtained
2870: for the quintic from  (\ref{aconst}) checks
2871: with~\cite{Candelas:1990rm}} of ${\mathbb{WCP}}$ to $M$ and the Gysins
2872: formula for smooth embeddings \cite{fulton}.
2873: 
2874: \subsection{Tables of Gopakumar-Vafa invariants}
2875: 
2876: We list the tables of BPS invariants for all the Calabi-Yau models
2877: computed in this paper.
2878: 
2879: 
2880: %-----------------------------SEXTIC----------------------------------------------------------
2881: 
2882: \begin{table}
2883: \begin{centering}
2884: \begin{tabular}{|r|rrrrr|}
2885: \hline
2886: g &d=1 &d=2 &d=3 &d=4 &d=5  \\
2887: \hline
2888: \, 0& \, 7884&6028452& \, 11900417220& \, 34600752005688& \,  24595034333130080\\
2889: 1& 0   &   7884&   145114704&  1773044322885&  17144900584158168\\
2890: 2& 0   &     0&        17496&    10801446444&    571861298748384\\
2891: 3& 0   &     0&          576&      -14966100&      1985113680408\\
2892: 4& 0   &     0&            6&       -47304  &       -21559102992\\
2893: 5& 0   &     0&            0&              0&           22232340\\
2894: 6& 0   &     0&            0&              0&              63072\\
2895: 7& 0   &     0&            0&              0&                  0\\
2896: \hline
2897: \end{tabular}
2898: \vskip  5 pt
2899: \begin{tabular}{|r|rrr|}
2900: \hline
2901: g &d=6 & d=7 &d=8 \\
2902: \hline
2903: 0& \!\!\! 513797193321737210316&  \!\!\!\!\!\!\!\!\!\!\!2326721904320912944749252&
2904: \!\!\!\!\!\!\!\!\!\!\!11284058913384803271372834984\\
2905: 1& \!\!\! 147664736456952923604& \!\!\!\!\!\!\!\!\!\!\!1197243574587406496495592&
2906: \!\!\!\!\!\!\!\!\!\!\!9381487423491392389034886369\\
2907: 2&  \!\!\! 13753100019804005556&  \!\!\!\!\!\!\!\!\!\!\! 233127389355701229349884&
2908: \!\!\!\!\!\!\!\!\!\!3246006977306701566424657380\\
2909: 3&    411536108778637626&   19655355035815833642912& 561664896373523768591774196\\
2910: 4&      1094535956564124&     628760082868148062854& 48641994707889298118864544\\
2911: 5&       -18316495265688&       3229011644338336680& 1863506489926528403683191\\
2912: 6&          207237771936&        -18261998133124302& 20758968356323626025164\\
2913: 7&            -583398600&           513634614205788& 10040615124628834206\\
2914: 8&               -146988&            -8041642037676& 1129699628821681740\\
2915: 9&                 -3168&               54521267292& -38940584273866593\\
2916: 10&                  -28&                 -43329384&  904511824896888\\
2917: 11&                    0&                  -110376& -12434437188576\\
2918: 12&                    0&                     0 &     76595605884 \\
2919: \hline
2920: \end{tabular}
2921: \caption{BPS invariants $n^g_d$ on the Sextic hypersurface in $\mathbb{P}^4(1^4,2)$.}
2922: \end{centering}
2923: \end{table}
2924: 
2925: 
2926: 
2927: 
2928: %----------------------------------Octic-------------------------------------------------
2929: \begin{table}
2930: \begin{centering}
2931: \begin{tabular}{|r|rrrrr|}
2932: \hline
2933: g &d=1 &d=2 &d=3 &d=4 &d=5  \\
2934: \hline
2935: 0& 29504& 128834912&   1423720546880& 23193056024793312&467876474625249316800 \\
2936: 1& 0   &      41312&     21464350592&  1805292092705856&101424054914016355712 \\
2937: 2& 0   &        864&       -16551744&    12499667277744&  5401493537244872896 \\
2938: 3& 0   &          6&         -177024&     -174859503824&    20584473699930496 \\
2939: 4& 0   &          0&               0&         396215800&     -674562224718848 \\
2940: 5& 0   &          0&               0&            301450&       12063928269056 \\
2941: 6& 0   &          0&               0&              4152&         -86307810432 \\
2942: 7& 0   &          0&               0&                24&             37529088 \\
2943: 8& 0   &          0&               0&                 0&               354048 \\
2944: $\vdots$&$\vdots$&$\vdots$&$\vdots$& $\vdots$& \\
2945: \hline
2946: \end{tabular}
2947: \caption{BPS invariants $n^g_d$ on the Octic hypersurface in $\mathbb{P}^4(1^4,4)$.}
2948: \end{centering}
2949: \end{table}
2950: 
2951: 
2952: 
2953: %---------------------------------Degree Ten--------------------------------------------
2954: \begin{table}
2955: \begin{centering}
2956: \begin{tabular}{|r|rrrr|}
2957: \hline
2958: g &d=1 &d=2 &d=3 &d=4  \\
2959: \hline
2960: 0& 231200& 12215785600& 1700894366474400& 350154658851324656000
2961: %& 89338191421813572850115680
2962: \\
2963: 1&    280&   207680960&  161279120326840& 103038403740897786400
2964: %& 59221844124053623534387208
2965: \\
2966: 2&      3&    -537976 &    1264588024791&   8495973047204168640
2967: %& 14756017196840260164213441
2968: \\
2969: 3&      0&       -1656&     -46669244594&     61218893443516800
2970: %&  1453221695196526165828470
2971: \\
2972: 4& 0     &         -12&        630052679&     -2460869494476896
2973: %&    34858775918056957341231
2974: \\
2975: 5& 0     &           0&         -1057570&       145198012290472
2976: %&     -116969829757480707582
2977: \\
2978: 6& 0     &           0&            -2646&        -5611087226688
2979: %&       21889605001143026505
2980: \\
2981: 7& 0     &           0&              -18&         125509540304
2982: %&        -2168811495068153226
2983: \\
2984: 8& 0     &           0&                0&         -1268283512
2985: %&          168060629266907127
2986: \\
2987: $\vdots$&$\vdots$&$\vdots$&$\vdots$&$\vdots$ \\
2988: \hline
2989: \end{tabular}
2990: \caption{BPS invariants $n^g_d$ on the degree 10  hypersurface in $\mathbb{P}^4(1^3,2,5)$.}
2991: \end{centering}
2992: \end{table}
2993: 
2994: %----------------------------------bicubic -------------------------------------------------
2995: \begin{table}
2996: \begin{centering}
2997: 
2998: \begin{tabular}{|r|rrrrrr|}
2999: \hline
3000: g &d=1 &d=2 &d=3 &d=4 &d=5 &d=6 \\
3001: \hline
3002: \, 0& 1053 & 52812 & 6424326  & 1139448384 & 249787892583& 62660964509532\\
3003: 1& 0   &  0 & 3402 &5520393  & 4820744484&3163476682080\\
3004: 2& 0   &  0 & 0 & 0 &  5520393& 23395810338\\
3005: 3& 0   &  0 &  0&  0& 0& 6852978 \\
3006: 4& 0   &  0 &  0&  0& 0&  10206\\
3007: 5& 0   &  0 &  0&  0& 0&  0 \\
3008: \hline
3009: \end{tabular}
3010: \begin{tabular}{|r|rrr|}
3011: \hline
3012: g &d=7 &d=8 &d=9  \\
3013: \hline
3014: 0& 17256453900822009 & 5088842568426162960  & 1581250717976557887945 \\
3015: 1&  1798399482469092 &  944929890853230501  &  473725044069553679454 \\
3016: 2&   42200615912499  &   50349477671013600  &   47431893998882182563 \\
3017: 3&      174007524240 &     785786604262830  &    1789615720312984368 \\
3018: 4&          -484542  &       2028116431098  &      21692992151427138\\
3019: 5&           158436 &           -784819773  &         36760497856020\\
3020: 6&                0 &               372762  &           -61753761036\\
3021: 7&                0 &               6318    &             -5412348\\
3022: 8&                0 &                0      &                39033\\
3023: 9&                0 &                0      &                1170\\
3024: 10&               0 &                0      &                 15\\
3025: 11      &         0 &                0      &                  0\\
3026: \hline
3027: \end{tabular}
3028: \caption{$n^g_d$ for the degree  (3,3) complete intersection in $\mathbb{P}^5$.}
3029: \end{centering}
3030: \end{table}
3031: 
3032: 
3033: \begin{table}
3034: \begin{centering}
3035: \begin{tabular}{|r|rrrrrr|}
3036: \hline
3037: g &d=1 &d=2 &d=3 &d=4 &d=5 &d=6\\
3038: \hline
3039: \, 0& 1280& 92288& 15655168& 3883902528& 1190923282176& 417874605342336  \\
3040: 1& 0&0& 0& -672& 16069888& 174937485184\\
3041: 2& 0& 0& 0& -8& 7680& 12679552\\
3042: 3& 0& 0& 0& 0& 0& 276864\\
3043: 4& 0& 0& 0& 0& 0& 0\\
3044: \hline
3045: \end{tabular}
3046: \begin{tabular}{|r|rrr|}
3047: \hline
3048: g &d=7 &d=8 &d=9\\
3049: \hline
3050: 0& \! 160964588281789696&  \!\! 66392895625625639488&  \!\! 28855060316616488359936 \\
3051: 1&  19078577926517760& 14088192680381290336& 9895851364631438617600\\
3052: 2&  494602061689344& 853657285175383648& 1137794220513866498304\\
3053: 3&  2016330670592& 14859083841009280& 49286012311292922368\\
3054: 4&  -285585152& 37334304102560& 679351051885623552\\
3055: 5& 591360& -46434384200& 1103462757073920\\
3056: 6& 7680& -8285120& -4031209095680\\
3057: 7&0& 67208& 370290688 \\
3058: 8& 0& 1520& -2270720 \\
3059: 9& 0& 15& -25600 \\
3060: 10&0& 0& 0\\
3061: \hline
3062: \end{tabular}
3063: 
3064: \caption{$n^g_d$ for the degree (4,2) complete intersection in $\mathbb{P}^5$.}
3065: \end{centering}
3066: \end{table}
3067: 
3068: \begin{table}
3069: \begin{centering}
3070: \begin{tabular}{|r|rrrrrr|}
3071: \hline
3072: g &d=1 &d=2 &d=3 &d=4 &d=5 &d=6\\
3073: \hline
3074: 0& 720& 22428& 1611504& 168199200& 21676931712& 3195557904564  \\
3075: 1& 0& 0& 64& 265113& 198087264& 89191835056\\
3076: 2& 0& 0& 0& 0& 10080& 180870120\\
3077: 3& 0& 0& 0& 0& 0& -3696  \\
3078: 4& 0& 0& 0& 0& 0& -56 \\
3079: 5& 0& 0& 0& 0& 0& 0 \\
3080: \hline
3081: \end{tabular}
3082: \begin{tabular}{|r|rrr|}
3083: \hline
3084: g &d=7 &d=8 &d=9\\
3085: \hline
3086: 0& \,\, 517064870788848& \, 89580965599606752& \, 16352303769375910848  \\
3087: 1& 32343228035424&10503104916431241& 3201634967657293024 \\
3088: 2& 315217101456& 280315384261560& 178223080602086784 \\
3089: 3& 199357344& 1430336342574& 2915033921871456 \\
3090: 4& 30240& 194067288& 8888143990672 \\
3091: 5& 0& 795339& -233104896  \\
3092: 6& 0& 0& 4857552 \\
3093: 7& 0& 0& 384 \\
3094: 8& 0& 0&  0\\
3095: \hline
3096: \end{tabular}
3097: \caption{$n^g_d$ for the (3,2,2) complete intersection in $\mathbb{P}^6$.}
3098: \end{centering}
3099: \end{table}
3100: 
3101: \begin{table}
3102: \begin{centering}
3103: \begin{tabular}{|r|rrrrrrr|}
3104: \hline
3105: g &d=1 &d=2 &d=3 &d=4 &d=5 &d=6& d=7\\
3106: \hline
3107: 0& \,\,  512&\, 9728& \,416256& \,25703936& \,1957983744&  \, 170535923200&  \,16300354777600   \\
3108: 1& 0& 0& 0& 14752& 8782848& 2672004608& 615920502784 \\
3109: 2& 0& 0& 0& 0& 0& 1427968& 2440504320\\
3110: 3& 0&0& 0& 0& 0& 0& 86016 \\
3111: 4&0& 0& 0& 0& 0& 0& 0  \\
3112: \hline
3113: \end{tabular}
3114: \begin{tabular}{|r|rrrr|}
3115: \hline
3116: g &d=8 &d=9 &d=10 &d=11\\
3117: \hline
3118: 0& \!\!\!\!\!  1668063096387072& \!\!\!\!\!\!\!\!\! 179845756064329728& \!\!\!\!\!\!\!\!\!20206497983891554816& \!\!\!\!\!\!\!\!\!2347339059011866069504 \\
3119: 1& \!\!\!\!\!\!\!\!\!123699143093152& \!\!\!\!\!\!\!\!\!22984995833484288&  \!\!\!\!\!\!\!\!\! 4071465816864581632&\!\!\!\!\!\!\!\!\! 698986176207439627264\\
3120: 2&  1628589698304& 702453851520512& 236803123487243776&68301097513852719616\\
3121: 3& 2403984384& 4702943495168& 4206537025629952& 2482415474680798208\\
3122: 4& -37632& 2449622016& 16316531089408& 29624281509824512\\
3123: 5& -672& 258048& 2777384448& 73818807399424\\
3124: 6& 0& 0& 4283904& 1153891840\\
3125: 7& 0& 0& 0& 26348544\\
3126: 8& 0& 0& 0& 0\\
3127: \hline
3128: \end{tabular}
3129: \caption{ $n^g_d$ for the degree (2,2,2,2) complete intersection in $\mathbb{P}^7$.}
3130: \end{centering}
3131: \end{table}
3132: 
3133: 
3134: \begin{table}
3135: \begin{centering}
3136: \begin{tabular}{|r|rrrrrr|}
3137: \hline
3138: g &d=1 &d=2 &d=3 &d=4 &d=5 &d=6 \\
3139: \hline
3140: 0& 1944& 223560& 64754568& 27482893704& 14431471821504& 8675274727197720\\
3141: 1&  0& 27& 161248& 381704265& 638555324400& 891094220317561\\
3142: 2& 0& 0& 0& 227448& 3896917776& 20929151321496\\
3143: 3& 0& 0& 0& 81& 155520& 75047188236\\
3144: 4& 0& 0& 0& 0& 5832& -40006768\\
3145: 5& 0& 0& 0& 0& 0& 26757\\
3146: 6& 0& 0& 0& 0& 0& 816\\
3147: 7& 0& 0& 0& 0& 0& 10\\
3148: 8&  0& 0& 0& 0& 0& 0\\
3149: \hline
3150: \end{tabular}
3151: \caption{ $n^g_d$ for the degree (4,3) complete intersection in $\mathbb{WCP}^5(1^5,2)$.}
3152: \end{centering}
3153: \end{table}
3154: 
3155: 
3156: \begin{table}
3157: \begin{centering}
3158: \begin{tabular}{|r|rrrrrr|}
3159: \hline
3160: g &d=1 &d=2 &d=3 &d=4 &d=5 &d=6 \\
3161: \hline
3162: 0& 3712& 982464& 683478144& 699999511744& 887939257620352& 1289954523115535040\\
3163: 1& 0& 1408& 6953728& 26841854688& 88647278203648& 266969312909257728\\
3164: 2& 0& 0& 3712& 148208928& 2161190443904& 17551821510538560\\
3165: 3& 0& 0& 0& -12432& 7282971392& 362668189458048\\
3166: 4&0& 0&0& 384& -14802048& 773557598272\\
3167: 5& 0& 0& 0& 6& -22272& -7046285440\\
3168: 6& 0& 0& 0& 0& 0& 6367872\\
3169: 7& 0& 0& 0& 0& 0& 11264\\
3170: 8& 0& 0& 0& 0& 0& 0\\
3171: \hline
3172: \end{tabular}
3173: \caption{$n^g_d$ for the degree (4,4) complete intersection in $\mathbb{WCP}^5(1^4,2^2)$.}
3174: \end{centering}
3175: \end{table}
3176: 
3177: 
3178: \begin{table}
3179: \begin{centering}
3180: \begin{tabular}{|r|rrrrrr|}
3181: \hline
3182: g &d=1 &d=2 &d=3 &d=4 &d=5 &d=6 \\
3183: \hline
3184: 0&4992&\!\!\!\!\!\! 2388768& \!\!\!\!\!\! 2732060032&\!\!\!\!\!\! 4599616564224& \!\!\!\!\!\!9579713847066240& \!\!\!\!\!\! 22839268002374163616 \\
3185: 1&0& -504& 1228032& 79275664800& 633074010435840& 3666182351842338408 \\
3186: 2& 0& -4& 14976& -13098688& 3921835430016& 128614837503143532\\
3187: 3& 0& 0& 0& 87376& -5731751168& 482407033529880\\
3188: 4&0& 0& 0& 1456& -7098624& -3978452463012 \\
3189: 5&0& 0& 0& 10& -59904& 1776341072 \\
3190: 6& 0& 0& 0& 0& 0& 18680344\\
3191: 7&0& 0& 0& 0& 0& -7176 \\
3192: 8& 0& 0& 0&  0& 0& -36\\
3193: 9&0& 0& 0& 0& 0& 0\\
3194: \hline
3195: \end{tabular}
3196: \caption{ $n^g_d$ for the degree (6,2) complete intersection in $\mathbb{WCP}^5(1^5,3)$.}
3197: \end{centering}
3198: \end{table}
3199: 
3200: \begin{table}
3201: \begin{centering}
3202: \begin{tabular}{|r|rrrrr|}
3203: \hline
3204: g &d=1 &d=2 &d=3 &d=4 &d=5  \\
3205: \hline
3206: 0& \ \ 15552&\ \  27904176& \ \ 133884554688&\ \  950676829466832& \ 8369111295497240640\\
3207: 1& 8& 258344& 5966034472& 126729436388624& 2512147219945401752\\
3208: 2&0& 128& 36976576& 4502079839576& 264945385369932352\\
3209: 3&0& 3& -64432& 15929894952& 9786781718701824\\
3210: 4& 0& 0& -48& -272993052& 42148996229312 \\
3211: 5& 0& 0& 0& 800065& -592538522344\\
3212: 6& 0& 0& 0& 1036& 14847229472\\
3213: 7& 0& 0& 0& 15& -148759496\\
3214: 8& 0& 0& 0& 0& 160128\\
3215: 9& 0& 0& 0& 0& 96\\
3216: 10& 0& 0& 0& 0& 0\\
3217: \hline
3218: \end{tabular}
3219: \caption{$n^g_d$ for the  degree (6,4) complete intersection in
3220: $\mathbb{WCP}^5(1^3,2^2,3)$.}
3221: \end{centering}
3222: \end{table}
3223: 
3224: \begin{table}
3225: \begin{centering}
3226: \begin{tabular}{|r|rrrr|}
3227: \hline
3228: g &d=1 &d=2 &d=3 &d=4 \\
3229: \hline
3230: 0&67104& 847288224& 28583248229280& 1431885139218997920 \\
3231: 1& 360& 40692096& 4956204918600& 616199133098321280\\
3232: 2& 1& 291328& 254022248925& 102984983365762128 \\
3233: 3& 0& -928& 1253312442& 6925290146728800\\
3234: 4& 0& -6& -39992931& 104226246583368\\
3235: 5& 0& 0& 867414& -442845743788\\
3236: 6&0& 0& -1807& 53221926192\\
3237: 7& 0& 0& -10& -3192574724\\
3238: 8& 0& 0& 0& 111434794\\
3239: 9& 0& 0& 0& -1752454\\
3240: 10& 0& 0& 0& 3054\\
3241: 11&0& 0& 0& 12\\
3242: 12& 0& 0& 0& 0\\
3243: \hline
3244: \end{tabular}
3245: \caption{$n^g_d$ for the degree (6,6) complete intersection in $\mathbb{WCP}^5(1^2,2^2,3^2)$.}
3246: \end{centering}
3247: \end{table}
3248: 
3249: 
3250: \newpage
3251: %-----------------------------------------------------------------
3252: \subsection{Invariance of the generators under a change of the basis}
3253: \label{sectioninvariance}
3254: We have seen the topological strings can be written as polynomials
3255: of the generators $v_1$ , $v_2$, $v_3$, and $X$. In the
3256: holomorphic limit, these generators can be computed from the first
3257: two solutions $\omega_0$, $\omega_1$ of the Picard-Fuchs equation.
3258: In the followings we prove that under an arbitrary linear change
3259: of basis of $\omega_0$ and $\omega_1$, these generators and
3260: therefore the topological strings are actually invariant. This is
3261: true anywhere in the moduli space. In particular, this partly
3262: explains why the gap structure in the conifold expansion is not
3263: affected by a change of basis of $\omega_0$ as we observed in
3264: all cases.
3265: 
3266: Since $X=\frac{1}{1-\psi}$ is independent of the basis $\omega_0$
3267: and $\omega_1$ , it is trivially invariant. In the holomorphic
3268: limit, The Kahler potential and metric go like $e^{-K}\sim
3269: \omega_0$ and $G_{\psi\bar{\psi}}\sim
3270: \partial_{\psi}t$, where $t=\frac{\omega_1}{\omega_0}$ is the mirror
3271: map. The generators $u$ and $v_i$ are related to $A_i$ and $B_i$,
3272: which we recall were defined as
3273: \begin{eqnarray} &&A_p:=\frac{(\psi\partial\psi)^p
3274: G_{\psi\bar{\psi}}}{G_{\psi\bar{\psi}}},~~B_p:=\frac{(\psi\partial\psi)^p
3275: e^{-K}}{e^{-K}}, ~~(p=1,2,3,\cdots)
3276: \end{eqnarray}
3277: So a different normalization of basis $\omega_0$ of $\omega_1$, as
3278: well as a change of basis in $\omega_1\rightarrow \omega_1+b_1
3279: \omega_0$ obviously do not change the generators $A_i$ and $B_i$,
3280: and therefore the generators $u$ and $v_i$ are also invariant.
3281: 
3282: We now tackle the remaining less trivial situation, namely a
3283: change of basis in $\omega_0$ as the following
3284: \begin{eqnarray} \label{basis-11-04-001}
3285: \omega_0\rightarrow \tilde{\omega}_0=\omega_0+b_1\omega_1
3286: \end{eqnarray}
3287: where $b_1$ is an arbitrary constant. We denote the K\"ahler
3288: potential, metric, mirror map and various generators in the new
3289: basis by a tilde symbol. It is straightforward to relate them to
3290: variables in the original basis. We find the following relations
3291: for the mirror map
3292: \begin{eqnarray}
3293: s=\frac{\omega_1}{\tilde{\omega}_0}=\frac{t}{1+b_1t}
3294: \nonumber \\
3295: \partial_{\psi}s=\frac{\partial_\psi t}{(1+b_1t)^2}
3296: \end{eqnarray}
3297: and the generators $A$ and $B$
3298: \begin{eqnarray} \label{ABrelation-11-04-002}
3299: \tilde{A}&=&\frac{\psi\partial_\psi\tilde{G}_{\psi\bar{\psi}}}{\tilde{G}_{\psi\bar{\psi}}}=A-\frac{2b_1\psi\partial_\psi t}{1+b_1\psi} \nonumber \\
3300: \tilde{B}&=&\frac{\psi\partial_\psi\tilde{\omega}_0}{\tilde{\omega}_0}=B+\frac{b_1\psi\partial_\psi
3301: t}{1+b_1\psi}
3302: \end{eqnarray}
3303: So we find the generators $A$ and $B$, as well as the generator
3304: $u=B$ are \textit{not} invariant under the change of basis
3305: (\ref{basis-11-04-001}). However, we recall the generator $v_1$ is
3306: defined as
3307: \begin{eqnarray}
3308: v_1=1+A+2B
3309: \end{eqnarray}
3310: Using the equations in (\ref{ABrelation-11-04-002}) we find the
3311: generator $v_1$ is invariant, namely $\tilde{v}_1=v_1$. To see
3312: $v_2$ and $v_3$ are invariant, we use the derivative relations
3313: \begin{eqnarray}
3314: \psi\partial_\psi v_1 &=& -v_1^2-2v_2-(1+r_0)X+v_1X \label{dr-11-04-0031} \\
3315: \psi\partial_\psi v_2 &=& -v_1v_2+v_3 \label{dr-11-04-0032}
3316: \end{eqnarray}
3317: where $r_0$ is a constant that appears in the relation of
3318: generator $A_2$ to lower generators. These derivative relations
3319: are exact and independent of the choice of the basis in asymptotic
3320: expansion. We have show that $v_1$ and $X$ in the first equations
3321: (\ref{dr-11-04-0031}) are invariant under a change of the basis
3322: (\ref{basis-11-04-001}), therefore the generator $v_2$ appearing
3323: on the right hand side must be also invariant. Applying the same
3324: logic to the second equation (\ref{dr-11-04-0032}) we find the
3325: generator $v_3$ are also invariant.
3326: 
3327: Our proof explains why a change of basis like
3328: (\ref{basis-11-04-001}) does not change the gap structure around
3329: the conifold point and seems to be related to the $SL_2$ orbit
3330: theorem of \cite{Schmidt,CKS}. Under a change of basis, the mirror map at the
3331: conifold point is
3332: $\tilde{t}_D=\frac{\omega_0t_D}{\tilde{\omega}_0}$, and has the
3333: asymptotic leading behavior $\tilde{t}_D\sim t_D\sim
3334: \mathcal{O}(\psi)$. Recall in the holomorphic limit, the conifold
3335: expansion is
3336: \begin{eqnarray}
3337: F^{(g)}_{\textrm{conifold}} =
3338: \omega_0^{2(g-1)}(\frac{1-\psi}{\psi})^{g-1}P_g(v_1,v_2,v_3,X) ,
3339: \end{eqnarray}
3340: As we have shown the generators $v_i$ and therefore $P_g$ are
3341: invariant, so in the new basis
3342: \begin{eqnarray}
3343: \tilde{F}^{(g)}_{\textrm{conifold}}=
3344: (\frac{\tilde{\omega}_0}{\omega_0})^{2(g-1)}
3345: F^{(g)}_{\textrm{conifold}}=(\frac{t_D}{\tilde{t}_D})^{2(g-1)}
3346: F^{(g)}_{\textrm{conifold}}
3347: \end{eqnarray}
3348: It is clear if there is a gap structure in one basis
3349: $F^{(g)}_{\textrm{conifold}}=\frac{(-1)^{g-1}B_{2g}}{2g(2g-2)t_D^{2g-2}}+\mathcal{O}(t_D^0)$,
3350: the same gap structure will be also present in the other basis,
3351: \begin{eqnarray}
3352: \tilde{F}^{(g)}_{\textrm{conifold}}=\frac{(-1)^{g-1}B_{2g}}{2g(2g-2)\tilde{t}_D^{2g-2}}+\mathcal{O}(\tilde{t}_D^0)
3353: \end{eqnarray}
3354: The asymptotic expansion in sub-leading terms $\mathcal{O}(t_D^0)$
3355: and $\mathcal{O}(\tilde{t}_D^0)$ will be different and can be
3356: computed by the relation between $t_D$ and $\tilde{t}_D$.
3357: 
3358: Around the conifold point there is another power series solution
3359: to the Picard-Fuchs equation that goes like $\omega_2\sim
3360: \mathcal{O}(\psi^2)$. We also observe that the gap structure is
3361: not affected by a change of the basis
3362: \begin{eqnarray}
3363: \omega_0\rightarrow \omega_0+b_2\omega_2
3364: \end{eqnarray}
3365: It appears to be much more difficult to prove this observation,
3366: since now the generators $v_i$ are not invariant under this change
3367: of basis. A proof of our observation would
3368: depend on the specific details of the polynomial $P_g$, and
3369: probably requires a deeper conceptual understanding of the
3370: conifold expansion. We shall leave this for future investigation.
3371: %------------------------------------------------------------------
3372: 
3373: 
3374: \subsection{Symplectic basis, vanishing cycles and massless particles}
3375: \label{appendixD-12-06}
3376: 
3377: We can study in more details the analytic continuation of the
3378: symplectic basis of the periods to the orbifold point $\psi=0$.
3379: For the four hypersurface cases and two other complete
3380: intersection models $X_{4,3}(1^5,2)$ and $X_{6,2}(1^3,2^2,3)$, the
3381: indices $a_i$ ($i=1,2,3,4$) of the Picard-Fuchs equation are not
3382: degenerate at the orbifold point, so there are $4$ power series
3383: with the leading behavior of $\psi^{a_i}$, and the analytic
3384: continuation procedure is similar to the quintic case. In our
3385: physical explanation of the singularity structure of higher genus
3386: topological string amplitudes in these models, we claim that for
3387: the hypersurface cases there is no stable massless charged state around
3388: the orbifold point, whereas for models $X_{4,3}(1^5,2)$ and
3389: $X_{6,2}(1^3,2^2,3)$ there are nearly massless charged states of
3390: mass $m\sim \psi^{a_2-a_1}$. Since the charge and the mass of a
3391: D-brane wrapping cycle is determined by the mirror map parameter,
3392: which is the ratio of two symplectic periods, it is only possible
3393: to have massless particles if there is a rational linear
3394: combination of the periods with the leading behavior of
3395: $\psi^{a_k}$, $k>1$.
3396: 
3397: For a complete intersection of degree $(d_1,\cdots,d_r)$ in
3398: weighted projective space $\mathbb{WCP}^n(w_1,\cdots,w_{n+1})$
3399: with non-degenerate indices $a_k$, the natural basis of solutions
3400: at the orbifold point is
3401: \begin{eqnarray}
3402: \omega_k^{\textrm{orb}}=\psi^{a_k}\frac{\prod_{i=1}^r\Gamma(d_ia_k)}{\prod_{i=1}^{n+1}\Gamma(w_ia_k)}
3403: \sum_{n=0}^{\infty}(c_0\psi)^n\frac{\prod_{i=1}^{n+1}\Gamma(w_i(n+a_k))}{\prod_{i=1}^r\Gamma(d_i(n+a_k))},~~~
3404: k=1,2,3,4
3405: \end{eqnarray}
3406: where
3407: $c_0=\frac{\prod_{i=1}^rd_i^{d_i}}{\prod_{i=1}^{n+1}w_i^{w_i}}$.
3408: We can write the sum of the series as a contour integral\footnote{Further useful
3409: properties of the periods of the one parameter models have been established in~\cite{Lazaroiu:2000jx}. }
3410: enclosing the positive real axis and analytically continue to the negative
3411: real axis to relate the above basis of solutions to the known
3412: symplectic basis at infinity $\psi=\infty$. We find the following
3413: relation, generalizing the result for quintic case,
3414: \begin{eqnarray} \label{D.113-12-07}
3415: \omega_k^{\textrm{orb}}&=&(2\pi
3416: i)^4\frac{\prod_{i=1}^r\Gamma(d_ia_k)}{\prod_{i=1}^{n+1}\Gamma(w_ia_k)}\{\frac{\alpha_k
3417: F_0}{1-\alpha_k}-\frac{\alpha_k F_1}{(1-\alpha_k)^2}
3418: +\frac{\alpha_k[\kappa(1+\alpha_k)-2a(1-\alpha_k)
3419: ]}{2(1-\alpha_k)^3}X_1 \nonumber \\
3420: && +\frac{\alpha_k[12c(1-\alpha_k)^2
3421: +\kappa(1+4\alpha_k+\alpha_k^2)]}{6(1-\alpha_k)^4}X_0
3422:   \}
3423: \end{eqnarray}
3424: here $\alpha_k=\exp(2\pi ia_k)$ and $\kappa$, $c$, $a$ are from
3425: the classical intersection calculations in Appendix
3426: \ref{intersection}. We find for all cases the symplectic form and
3427: Kahler potential have the same diagonal behavior as the case of
3428: the quintic
3429: \begin{eqnarray}
3430: \omega=dF_k\wedge dX_k=s_1 d\omega_1^{\textrm{orb}}\wedge
3431: d\omega_4^{\textrm{orb}}+s_2 d\omega_2^{\textrm{orb}}\wedge
3432: d\omega_3^{\textrm{orb}}
3433: \end{eqnarray}
3434: and $e^{-K}=\sum_{k=1}^4 r_k \omega_k^{\textrm{orb}}
3435: \overline{\omega_k^{\textrm{orb}}}$, for some constants $s_1, s_2$
3436: and $r_k$.
3437: 
3438: It is straightforward to invert the transformation
3439: (\ref{D.113-12-07}) and study the asymptotic behavior of the
3440: geometric symplectic basis $(F_0,F_1,X_0,X_1)$ around the orbifold
3441: point. Generically a linear combination of the symplectic periods
3442: $(F_0,F_1,X_0,X_1)$ is proportional to the power of $\psi$ with
3443: the lowest index, namely, for generic coefficients $c_1, c_2, c_3,
3444: c_4$ we have
3445: \begin{eqnarray}
3446: c_1X_0+c_2X_1+c_3F_1+c_4F_0\sim \omega_1^{\textrm{orb}}\sim
3447: \psi^{a_1}
3448: \end{eqnarray}
3449: In order for massless particles to appear at the orbifold point,
3450: there must be a $Sp(4,\mathbb{Z})$ transformation of the
3451: symplectic basis such that one of periods goes to zero faster that
3452: the generic situation, namely we should have integer coefficients $n_k\in \mathbb{Z}$
3453: satisfying
3454: \begin{eqnarray} \label{D.115}
3455: n_1X_0+n_2X_1+n_3F_1+n_4F_0\sim \omega_2^{\textrm{orb}}\sim
3456: \psi^{a_k},~~~~k>1
3457: \end{eqnarray}
3458: 
3459: We find (\ref{D.115}) is impossible for three of the hypersurface
3460: cases $X_5(1^5)$, $X_8(1^4,4)$ and $X_{10}(1^3,2,5)$, but possible
3461: for the Sextic hypersurface $X_6(1^4,2)$ and the complete
3462: intersection cases $X_{4,3}(1^5,2)$ and $X_{6,2}(1^3,2^2,3)$.
3463: Specifically, for the Sextic hypersurface $X_6(1^4,2)$, the
3464: condition for (\ref{D.115}) with $k=2$
3465: \begin{eqnarray}
3466: n_1+3n_3=0, ~~~ 3n_1+4(n_2+n_4)=0
3467: \label{allowed1}
3468: \end{eqnarray}
3469: whereas for the complete intersections  $X_{4,3}(1^5,2)$ and
3470: $X_{6,2}(1^3,2^2,3)$, the conditions for (\ref{D.115}) are
3471: \begin{eqnarray}
3472: n_1+3n_3=0, ~~~ n_1=4n_2+8n_4,
3473: \label{allowed2}
3474: \end{eqnarray}
3475: and
3476: \begin{eqnarray}
3477: n_1+2n_3=0, ~~~ n_2+\frac{7}{24}n_3+n_4=0.
3478: \label{allowed3}
3479: \end{eqnarray}
3480: respectively.
3481: 
3482: This fact that there are massless integer charged states possible
3483: in the models $X_{4,3}(1^5,2)$ and  $X_{4,3}(1^5,2)$ is entirely
3484: consistent with our physical picture, which explained the singular
3485: behaviour of the $F^{(g)}$ form the effective action point of
3486: view. What is very interesting is the fact that the period
3487: degeneration at the branch point of the hypersurface models,
3488: $X_5(1^5)$,  $X_6(1^4,2)$, $X_8(1^4,4)$ and  $X_{10}(1^3,2,5)$  
3489: is at genus zero very similar to the cases $X_{4,3}(1^5,2)$ and  $X_{6,2}(1^3,2^2,3)$.
3490: In particular the periods of the models have no logarithmic singularities. 
3491: The leading behaviour of the higher genus expansion, which we obtain 
3492: from the global properties, indicates that the BPS states,  
3493: which are possibly massless by (\ref{allowed1},\ref{allowed2},\ref{allowed3}),   
3494: are stable in  $X_{4,3}(1^5,2)$  and  $X_{6,2}(1^3,2^2,3)$ models, 
3495: but  not stable in the  $X_6(1^4,2)$ model.  
3496: 
3497: 
3498: %although it seems possible, the actual geometric cycles at the
3499: %orbifold point do not shrink faster than they generically do.
3500: 
3501: 
3502: \begin{thebibliography}{10}
3503: 
3504: \bibitem{Klemm:1999gm}
3505:   A.~Klemm and E.~Zaslow,
3506:    ``Local mirror symmetry at higher genus,''
3507:   arXiv:hep-th/9906046.
3508:   %%CITATION = HEP-TH 9906046;%%
3509: 
3510: 
3511: \bibitem{Aganagic:2003db}
3512:   M.~Aganagic, A.~Klemm, M.~Marino and C.~Vafa,
3513:    ``The topological vertex,''
3514:   Commun.\ Math.\ Phys.\  {\bf 254}, 425 (2005)
3515:   [arXiv:hep-th/0305132].
3516:   %%CITATION = HEP-TH 0305132;%%
3517: 
3518: \bibitem{Witten:1992fb}
3519:   E.~Witten,
3520:    ``Chern-Simons Gauge Theory As A String Theory,''
3521:   Prog.\ Math.\  {\bf 133}, 637 (1995)
3522:   [arXiv:hep-th/9207094].
3523:   %%CITATION = HEP-TH 9207094;%%
3524: 
3525: \bibitem{Gopakumar:1998ki}
3526:   R.~Gopakumar and C.~Vafa,
3527:    ``On the gauge theory/geometry correspondence,''
3528:   Adv.\ Theor.\ Math.\ Phys.\  {\bf 3}, 1415 (1999)
3529:   [arXiv:hep-th/9811131].
3530:   %%CITATION = HEP-TH 9811131;%%
3531: 
3532: \bibitem{Dijkgraaf:2002fc}
3533:   R.~Dijkgraaf and C.~Vafa,
3534:    ``Matrix models, topological strings, and supersymmetric gauge theories,''
3535:   Nucl.\ Phys.\ B {\bf 644}, 3 (2002)
3536:   [arXiv:hep-th/0206255].
3537:   %%CITATION = HEP-TH 0206255;%%
3538: 
3539: 
3540: 
3541: 
3542: \bibitem{Aganagic:2003qj}
3543:   M.~Aganagic, R.~Dijkgraaf, A.~Klemm, M.~Marino and C.~Vafa,
3544:    ``Topological strings and integrable hierarchies,''
3545:   Commun.\ Math.\ Phys.\  {\bf 261}, 451 (2006)
3546:   [arXiv:hep-th/0312085].
3547:   %%CITATION = HEP-TH 0312085;%%
3548: 
3549: \bibitem{Vafa:2004qa}
3550:   C.~Vafa,
3551:    ``Two dimensional Yang-Mills, black holes and topological strings,''
3552:   arXiv:hep-th/0406058.
3553:   %%CITATION = HEP-TH 0406058;%%
3554: 
3555: \bibitem{costello}
3556: K.~Costello,
3557: ``Topological conformal field theories and Calabi-Yau categories,''
3558: [math.QA/0412149].
3559: 
3560: \bibitem{kontsevich}
3561: M.~Kontsevich, ``TQFTs and geometry of pre-Frobenius manifolds'' and  ``$A_\infty$  categories
3562: and  their (co)homology theories,'' Lecures at the ``Erwin Schroedinger Institute,'' Vienna 2006.
3563: 
3564: \bibitem{Ooguri:2004zv}
3565:   H.~Ooguri, A.~Strominger and C.~Vafa,
3566:    ``Black hole attractors and the topological string,''
3567:   Phys.\ Rev.\ D {\bf 70}, 106007 (2004)
3568:   [arXiv:hep-th/0405146].
3569:   %%CITATION = HEP-TH 0405146;%%
3570: 
3571: \bibitem{Ooguri:2005vr}
3572:   H.~Ooguri, C.~Vafa and E.~P.~Verlinde,
3573:    ``Hartle-Hawking wave-function for flux compactifications,''
3574:   Lett.\ Math.\ Phys.\  {\bf 74}, 311 (2005)
3575:   [arXiv:hep-th/0502211].
3576:   %%CITATION = HEP-TH 0502211;%%
3577: 
3578: 
3579: %\cite{Huang:2006si}
3580: \bibitem{Huang}
3581:   M.~x.~Huang and A.~Klemm,
3582:   ``Holomorphic anomaly in gauge theories and matrix models,''
3583:   arXiv:hep-th/0605195.
3584:   %%CITATION = HEP-TH 0605195;%%
3585: 
3586: \bibitem{Huang2}
3587:   M.~x.~Huang and A.~Klemm,
3588:   to appear.
3589:   %%CITATION = HEP-TH 0605195;%%
3590: 
3591: \bibitem{Seiberg:1994rs}
3592:   N.~Seiberg and E.~Witten,
3593:    ``Electric - magnetic duality, monopole condensation, and confinement in N=2
3594:     supersymmetric Yang-Mills theory,''
3595:   Nucl.\ Phys.\ B {\bf 426}, 19 (1994)
3596:   [Erratum-ibid.\ B {\bf 430}, 485 (1994)]
3597:   [arXiv:hep-th/9407087].
3598:   %%CITATION = HEP-TH 9407087;%%.
3599: 
3600: 
3601: %\cite{Aganagic:2006wq}
3602: \bibitem{Aganagic:2006wq}
3603:   M.~Aganagic, V.~Bouchard and A.~Klemm,
3604:   ``Topological strings and (almost) modular forms,''
3605:   arXiv:hep-th/0607100.
3606:   %%CITATION = HEP-TH 0607100;%%
3607: 
3608: 
3609: 
3610: \bibitem{Hosono:1999qc}
3611:   S.~Hosono, M.~H.~Saito and A.~Takahashi,
3612:   %``Holomorphic anomaly equation and BPS state counting of rational  elliptic
3613:   %surface,''
3614:   Adv.\ Theor.\ Math.\ Phys.\  {\bf 3} (1999) 177
3615:   [arXiv:hep-th/9901151].
3616:   %%CITATION = HEP-TH 9901151;%%
3617: 
3618: \bibitem{Hosono:2001gf}
3619:   S.~Hosono, M.~H.~Saito and A.~Takahashi,
3620:   %``Relative Lefschetz Action and BPS State Counting,''
3621:   arXiv:math.ag/0105148.
3622:   %%CITATION = MATH-AG 0105148;%%
3623: 
3624: \bibitem{dijkgraafzagier} R.~Dijkgraaf, ``Mirror Symmetry and elliptic curves,'' in {\sl
3625: The moduli Space of Curves}, Progr. Math. {\bf 129}, 149;
3626: K.~Kaneko and D.~B.~Zagier, ``A generalized Jacobi theta functions
3627: and quasimodular forms,'' ibidem 165.
3628: 
3629: 
3630: %\cite{Candelas:1990rm}
3631: \bibitem{Candelas:1990rm}
3632:   P.~Candelas, X.~C.~De La Ossa, P.~S.~Green and L.~Parkes,
3633:    ``A pair of Calabi-Yau manifolds as an exactly soluble superconformal
3634:   theory,''
3635:   Nucl.\ Phys.\ B {\bf 359}, 21 (1991).
3636:   %%CITATION = NUPHA,B359,21;%%
3637: 
3638: %\cite{Yamaguchi:2004bt}
3639: \bibitem{Yamaguchi}
3640:   S.~Yamaguchi and S.~T.~Yau,
3641:   ``Topological string partition functions as polynomials,''
3642:   JHEP {\bf 0407}, 047 (2004)
3643:   [arXiv:hep-th/0406078].
3644:   %%CITATION = HEP-TH 0406078;%%
3645: 
3646: 
3647: %\cite{Bershadsky:1993ta}
3648: \bibitem{BCOVI}
3649:   M.~Bershadsky, S.~Cecotti, H.~Ooguri and C.~Vafa,
3650:    ``Holomorphic anomalies in topological field theories,''
3651:   Nucl.\ Phys.\ B {\bf 405}, 279 (1993)
3652:   [arXiv:hep-th/9302103].
3653:   %%CITATION = HEP-TH 9302103;%%
3654: 
3655: %\cite{Bershadsky:1993cx}
3656: \bibitem{BCOV}
3657:   M.~Bershadsky, S.~Cecotti, H.~Ooguri and C.~Vafa,
3658:    ``Kodaira-Spencer theory of gravity and exact results for quantum string
3659:   amplitudes,''
3660:   Commun.\ Math.\ Phys.\  {\bf 165}, 311 (1994)
3661:   [arXiv:hep-th/9309140].
3662:   %%CITATION = HEP-TH 9309140;%%
3663: 
3664: \bibitem{Hosono:1994ax}
3665:   S.~Hosono, A.~Klemm, S.~Theisen and S.~T.~Yau,
3666:    ``Mirror symmetry, mirror map and applications to complete intersection
3667:   Calabi-Yau spaces,''
3668:   Nucl.\ Phys.\ B {\bf 433}, 501 (1995)
3669:   [arXiv:hep-th/9406055].
3670:   %%CITATION = HEP-TH 9406055;%%
3671: 
3672: \bibitem{MB} K.~Hori, S.~Katz, A.~Klemm, R.Pandharipande, R.~Thomas, C.~Vafa, R.~Vavil and E. Zaslow,
3673: {\sl Mirror Symmetry}, AMS (2003).
3674: 
3675: \bibitem{Ceresole:1995jg}
3676:   A.~Ceresole, R.~D'Auria, S.~Ferrara and A.~Van Proeyen,
3677:    ``Duality transformations in supersymmetric Yang-Mills theories coupled to
3678:    supergravity,''
3679:   Nucl.\ Phys.\ B {\bf 444}, 92 (1995)
3680:   [arXiv:hep-th/9502072].
3681:   %%CITATION = HEP-TH 9502072;%% and the $K$-theory
3682: 
3683: \bibitem{Minasian:1997mm}
3684:   R.~Minasian and G.~W.~Moore,
3685:    ``K-theory and Ramond-Ramond charge,''
3686:   JHEP {\bf 9711}, 002 (1997)
3687:   [arXiv:hep-th/9710230].
3688:   %%CITATION = HEP-TH 9710230;%%
3689: 
3690: \bibitem{Cheung:1997az}
3691:   Y.~K.~Cheung and Z.~Yin,
3692:    ``Anomalies, branes, and currents,''
3693:   Nucl.\ Phys.\ B {\bf 517}, 69 (1998)
3694:   [arXiv:hep-th/9710206].
3695:   %%CITATION = HEP-TH 9710206;%%
3696: 
3697: \bibitem{Brunner:1999jq}
3698:   I.~Brunner, M.~R.~Douglas, A.~E.~Lawrence and C.~Romelsberger,
3699:    ``D-branes on the quintic,''
3700:   JHEP {\bf 0008}, 015 (2000)
3701:   [arXiv:hep-th/9906200].
3702:   %%CITATION = HEP-TH 9906200;%%.
3703: 
3704: 
3705: \bibitem{fulton} W.~Fulton, ``Intersection Theory,'' Springer (1998).
3706: 
3707: \bibitem{AGV} V.~I.~ Arnold, S.~M.~ Gusein-Zade, A.~N.~ Varchencko, ``Singularities of diefferential maps I\& II,''
3708: Birkh\"auser, Basel (1985).
3709: 
3710: 
3711: \bibitem{Diaconescu:1999vp}
3712:   D.~E.~Diaconescu and C.~Romelsberger,
3713:   %``D-branes and bundles on elliptic fibrations,''
3714:   Nucl.\ Phys.\ B {\bf 574}, 245 (2000)
3715:   [arXiv:hep-th/9910172].
3716:   %%CITATION = HEP-TH 9910172;%%
3717: 
3718: \bibitem{Mayr:2000as}
3719:   P.~Mayr,
3720:    ``Phases of supersymmetric D-branes on Kaehler manifolds and the McKay
3721:    correspondence,''
3722:   JHEP {\bf 0101}, 018 (2001)
3723:   [arXiv:hep-th/0010223].
3724:   %%CITATION = HEP-TH 0010223;%%
3725: 
3726: %\cite{Katz:1999xq}
3727: \bibitem{Katz:1999}
3728:   S.~Katz, A.~Klemm and C.~Vafa,
3729:   ``M-theory, topological strings and spinning black holes,''
3730:   Adv.\ Theor.\ Math.\ Phys.\  {\bf 3}, 1445 (1999)
3731:   [arXiv:hep-th/9910181].
3732:   %%CITATION = HEP-TH 9910181;%%
3733: 
3734: %\cite{Ghoshal:1995wm}
3735: \bibitem{Ghoshal:1995}
3736:   D.~Ghoshal and C.~Vafa,
3737:   ``C = 1 String As The Topological Theory Of The Conifold,''
3738:   Nucl.\ Phys.\ B {\bf 453}, 121 (1995)
3739:   [arXiv:hep-th/9506122].
3740:   %%CITATION = HEP-TH 9506122;%%
3741: 
3742: %\cite{Marino:1998pg}
3743: \bibitem{Marino:1998}
3744:   M.~Marino and G.~W.~Moore,
3745:   ``Counting higher genus curves in a Calabi-Yau manifold,''
3746:   Nucl.\ Phys.\ B {\bf 543}, 592 (1999)
3747:   [arXiv:hep-th/9808131].
3748:   %%CITATION = HEP-TH 9808131;%%
3749: 
3750: %\cite{Klemm:2005pd}
3751: \bibitem{Klemm:2005pd}
3752:   A.~Klemm and M.~Marino,
3753:    ``Counting BPS states on the Enriques Calabi-Yau,''
3754:   arXiv:hep-th/0512227.
3755:   %%CITATION = HEP-TH 0512227;%%
3756: 
3757: 
3758: %\cite{Gopakumar:1998ii}
3759: \bibitem{GVI}
3760:   R.~Gopakumar and C.~Vafa,
3761:   ``M-theory and topological strings. I,''
3762:   arXiv:hep-th/9809187.
3763: 
3764: \bibitem{GVII}
3765:   R.~Gopakumar and C.~Vafa,
3766:  ``M-theory and topological strings. II,''
3767:    hep-th/9812127.
3768:   %%CITATION = HEP-TH 9809187;%%
3769: 
3770: %\cite{Faber:2000ma}
3771: \bibitem{FP}
3772:   C.~Faber and R.~Pandharipande,
3773:    ``Hodge Integrals and Gromov-Witten theory,''
3774:   arXiv:math.ag/9810173.
3775:   %%CITATION = MATH-AG 0002112;%%
3776: 
3777: %\cite{GKMW}
3778: \bibitem{GKMW}
3779: T.~Grimm, A.~Klemm, M.~Marino and M.~Weiss, work in progress.
3780: 
3781: \bibitem{Vafa:1995ta}
3782:   C.~Vafa,
3783:    ``A Stringy test of the fate of the conifold,''
3784:   Nucl.\ Phys.\ B {\bf 447}, 252 (1995)
3785:   [arXiv:hep-th/9505023].
3786:   %%CITATION = HEP-TH 9505023;%%
3787: 
3788: 
3789: \bibitem{wittensurvey} E.~Witten, ``Two-Dimensional Gravity and intersection theory on  moduli spaces,''
3790: Surveys in Diff. Geometry {\bf 1} (1991) 243.
3791: 
3792: 
3793: 
3794: \bibitem{Strominger:1995cz}
3795:   A.~Strominger,
3796:    ``Massless black holes and conifolds in string theory,''
3797:   Nucl.\ Phys.\ B {\bf 451}, 96 (1995)
3798:   [arXiv:hep-th/9504090].
3799:   %%CITATION = HEP-TH 9504090;%%
3800: 
3801: 
3802: \bibitem{GVbfg} R. Gopakumar and C. Vafa, ``Branes and fundamental Groups,''
3803:   [arXiv:hep-th/9712048].
3804: 
3805: 
3806: 
3807: \bibitem{Klemm:1995kj}
3808:   A.~Klemm and P.~Mayr,
3809:    ``Strong coupling singularities and non-Abelian gauge symmetries in  N = 2
3810:   string theory,'' Nucl.\ Phys.\ B {\bf 469}, 37 (1996)
3811:   [arXiv:hep-th/9601014].
3812: 
3813: 
3814: 
3815: \bibitem{Klemm:1992tx} A.~Klemm and S.~Theisen,
3816:    ``Considerations of one modulus Calabi-Yau compactifications: Picard-Fuchs
3817:    equations, Kahler potentials and mirror maps,''
3818:   Nucl.\ Phys.\ B {\bf 389}, 153 (1993)
3819:   [arXiv:hep-th/9205041].
3820:   %%CITATION = HEP-TH 9205041;%%
3821: 
3822: \bibitem{DTGW} D.~Maulik, N.~Nekrasov, A.~Okounkov and R.~Pandharipande,
3823:         {\sl Gromov-Witten theory and Donaldson-Thomas theory I,}
3824:         [arXiv:math.AG/0312059].
3825: 
3826: 
3827: 
3828: \bibitem{Verlinde:2004ck}
3829:   E.~P.~Verlinde,
3830:   ``Attractors and the holomorphic anomaly,''
3831:   arXiv:hep-th/0412139.
3832:   %%CITATION = HEP-TH 0412139;%%
3833: 
3834: 
3835: 
3836: \bibitem{Gunaydin:2006bz}
3837:   M.~Gunaydin, A.~Neitzke and B.~Pioline,
3838:   ``Topological wave functions and heat equations,''
3839:   arXiv:hep-th/0607200.
3840:   %%CITATION = HEP-TH 0607200;%%
3841: 
3842: \bibitem{cyy} Y.-H. Chen, Y. Yang, N. Yui, ``Monodromy of Picard-Fuchs
3843: differential equations for Calabi-Yau threefolds,'' [arXiv:math.AG/0605675].
3844: 
3845: 
3846: 
3847: \bibitem{CR} W.~Chen and Y.~Ruan, ``Orbifold Gromow-Witten theory'', Orbifolds in mathmatics and physics
3848: Madison WI 2001, Contemp. Math., vol 310, Amer. Math. Soc., Providence, RI, 2002 25-85.
3849: 
3850: 
3851: \bibitem{BGP} J.~Bryan, T.~Graber, R.~Pandharipande, ``The orbifold quantum
3852: cohomology of $C^2/Z_3 $ and Hurwitz-Hodge integrals,'' [arXiv:math.AG/0510335].
3853: 
3854: \bibitem{CCIT} T.~Coates, A.~Corti, H.~Iritani and H.-H. Tseng, ``Wall-Crossing in Toric Gromow-Witten Theory I:
3855: Crepant Examples'', [arXiv:math.AG/0611550].
3856: 
3857: 
3858: \bibitem{wittenwavefunction} E.~Witten, ``Quantum background independence
3859: in string theory,'' Nucl. Phys. B {\bf 372}, 187 (1992), [arXiv:hep-th/9306122].
3860: 
3861: \bibitem{GiventhalCoates} A.~B.~Giventhal, ``Symplectic geometry of Frobenius structures,''
3862: Frobenius manifolds, Aspects. Math., E36, Vieweg, Wiesbaden,
3863: (91-112) 2004, T.~Coates, ``Giventhals Lagrangian Cone and
3864: $S^1$-Equivariant Gromow-Witten Theory,'' [arXiv:math.AG.0607808].
3865: 
3866: \bibitem{DoranMorgan} C.~Doran and J~Morgan, ``Mirror Symmetry and Integral Variations of
3867:     Hodge Structure Underlying One Parameter Families of Calabi-Yau Threefolds,'' [arXiv:math.AG/0505272].
3868: 
3869: \bibitem{Lazaroiu:2000jx}
3870:   C.~I.~Lazaroiu,
3871:   ``Collapsing D-branes in one-parameter models and small/large radius
3872:    duality,''
3873:   Nucl.\ Phys.\ B {\bf 605}, 159 (2001)
3874:   [arXiv:hep-th/0002004].
3875:   %%CITATION = HEP-TH 0002004;%%
3876: 
3877: \bibitem{KKRS} A.~Klemm, M.~Kreuzer, E.~Riegler and E.~Scheidegger,
3878:   ``Topological string amplitudes, complete intersection Calabi-Yau spaces  and
3879:   threshold corrections,''
3880:   JHEP {\bf 0505}, 023 (2005)
3881:   [arXiv:hep-th/0410018].
3882:   %%CITATION = HEP-TH 0410018;%%
3883: 
3884: \bibitem{Schmidt} W.~Schmidt, ``Variaton of Hodge structure: The singularities of the period mapping,''
3885:         Inventiones Math.{\bf 22} (1973) 211-319.
3886: 
3887: \bibitem{CKS} E.~ Cattani, A.~Kaplan and W.~Schmidt, ``Degenerarions of Hodge Structures,''
3888: Annals of Math. {\bf 123} (1986)  457-535.
3889: 
3890: \bibitem{Harris} J.~Harris, ``Curves in Projective Space,'' Sem. de Math. Sup. Les Presses de l'University
3891: de Monteral (1982).
3892: 
3893: \bibitem{webpage} {\tt http://uw.physics.wisc.edu/$\sim$strings/aklemm/highergenusdata/}
3894: 
3895: \bibitem{Bridgeland}  T.~Bridgeland, ``Derived categories of coherent sheaves,''
3896: Survey article for ICM 2006,  [arXiv:math.AG/0602129].
3897: 
3898: \bibitem{Aspinwall}
3899:   P.~S.~Aspinwall,
3900:   ``D-branes on Calabi-Yau manifolds,''
3901:   arXiv:hep-th/0403166.
3902:   %%CITATION = HEP-TH 0403166;%%
3903: 
3904: 
3905: \bibitem{Schubert} S.~Katz, S. Str{\o}mme and J-M. {\O}kland: ``Schubert: A Maple package
3906: for intersection theory and enumerative geometry,'' {\tt http://www.uib.no/People/nmasr/schubert/}.
3907: 
3908: \bibitem{Gaiotto:2006wm}
3909:   D.~Gaiotto, A.~Strominger and X.~Yin,
3910:   ``The M5-brane elliptic genus: Modularity and BPS states,''
3911:   arXiv:hep-th/0607010.
3912:   %%CITATION = HEP-TH 0607010;%%
3913: 
3914: \bibitem{deBoer:2006vg}
3915:   J.~de Boer, M.~C.~N.~Cheng, R.~Dijkgraaf, J.~Manschot and E.~Verlinde,
3916:   ``A farey tail for attractor black holes,''
3917:   JHEP {\bf 0611}, 024 (2006)
3918:   [arXiv:hep-th/0608059].
3919:   %%CITATION = HEP-TH 0608059;%% using the dilute gas approximation in $AdS_3\times S^2\times M$.
3920: 
3921: \bibitem{DennefMoore} F.~Dennef and G.~Moore, ``From OSV to OSV,'' talk available at
3922: {\tt http://strings06.itp.ac.cn/} and work in progress.
3923: 
3924: 
3925: \bibitem{Rocek:2005ij}
3926:   M.~Rocek, C.~Vafa and S.~Vandoren,
3927:   ``Hypermultiplets and topological strings,''
3928:   JHEP {\bf 0602} (2006) 062
3929:   [arXiv:hep-th/0512206].
3930:   %%CITATION = HEP-TH 0512206;%%.
3931: 
3932: \end{thebibliography}
3933: 
3934: 
3935: 
3936: 
3937: 
3938: \end{document}
3939: