1: \documentclass[12pt]{article}
2:
3: \usepackage{graphicx}
4: \usepackage{color}
5: \usepackage{bm}
6:
7: \def\eq#1{(\ref{#1})}
8: \def\Eq#1{Eq.~(\ref{#1})}
9: \newcommand{\secn}[1]{Section~\rf{#1}}
10: \newcommand{\bra}[1]{\langle{#1}|}
11: \newcommand{\ket}[1]{|{#1}\rangle}
12: \newcommand{\braket}[2]{\langle{#1}|{#2}\rangle}
13: \newcommand{\tbl}[1]{Table~\ref{#1}}
14: \newcommand{\fig}{Fig.~\ref}
15: \newcommand{\nl}{\nonumber \\}
16: \def\zeroslash{\mathord{\not\mathrel{\hskip 0.04 cm 0}}}
17: \def\beq{\begin{equation}}
18: \def\eeq{\end{equation}}
19: \def\beqa{\begin{eqnarray}}
20: \def\eeqa{\end{eqnarray}}
21: \def\bet{\begin{tabular}}
22: \def\eet{\end{tabular}}
23: \def\del{\partial}\def\delb{\bar\partial}
24: \newcommand{\ex}[1]{{\rm e}^{#1}} \def\ii{{\rm i}}
25: \newcommand{\Tr}{{\rm Tr}}
26: \newcommand{\sect}[1]{\setcounter{equation}{0}\section{#1}}
27: \renewcommand{\theequation}{\thesection.\arabic{equation}}
28: \newcommand{\vs}[1]{\vspace{#1 mm}}
29: \renewcommand{\a}{\alpha}
30: \renewcommand{\b}{\beta}
31: \renewcommand{\c}{\gamma}
32: \renewcommand{\d}{\delta}
33: \renewcommand{\l}{\lambda}
34: \newcommand{\e}{\epsilon}
35: \newcommand{\ve}{{\vec{\e}}}
36: \newcommand{\hs}[1]{\hspace{#1 mm}}
37: \newcommand{\shalf}{\frac{1}{2}}
38: \newcommand{\bxi}{\mbox{\boldmath $\xi$}}
39: \newcommand{\bkp}{{\bf k}_\perp}
40: \newcommand{\bxp}{{\bf x}_\perp}
41: \newcommand{\byp}{{\bf y}_\perp}
42: \renewcommand{\Im}{{\rm Im}\,}
43: \def\one{{\hbox{ 1\kern-.8mm l}}}
44:
45:
46:
47: \renewcommand{\baselinestretch}{1.1}
48: \textwidth 160mm
49: \textheight 215mm
50: \topmargin -.05in
51: \oddsidemargin 5mm
52:
53: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
54:
55:
56: \usepackage[centertags]{amsmath}
57:
58: \usepackage{amsfonts}
59: \usepackage{amssymb}
60: \usepackage{amsthm}
61:
62: \begin{document}
63:
64: \begin{titlepage}
65:
66: \setcounter{page}{0}
67:
68: \begin{flushright}
69: {DFTT 16/2007}\\
70: {QMUL-PH-07-17}
71: \end{flushright}
72:
73: \vspace{0.6cm}
74:
75: \begin{center}
76: {\Large \bf New twist field couplings from the \\
77: partition function for multiply wrapped D-branes.} \\
78:
79: \vskip 0.8cm
80:
81: {\bf Dario Du\`o and Rodolfo Russo}\\
82: {\sl
83: Centre for Research in String Theory \\ Department of Physics\\
84: Queen Mary, University of London\\
85: Mile End Road, London, E1 4NS,
86: United Kingdom}\\
87:
88: \vskip .3cm
89:
90: {\bf Stefano Sciuto}\\
91: {\sl Dipartimento di Fisica Teorica, Universit\`a di Torino}\\
92: and {\sl INFN, Sezione di Torino}\\
93: {\sl Via P. Giuria 1, I-10125 Torino, Italy}\\
94:
95: \vskip 1.2cm
96:
97: \end{center}
98:
99: \begin{abstract}
100:
101: We consider toroidal compactifications of bosonic string theory with
102: particular regard to the phases (cocycles) necessary for a consistent
103: definition of the vertex operators, the boundary states and the
104: T-duality rules. We use these ingredients to compute the planar
105: multi-loop partition function describing the interaction among
106: magnetized or intersecting D-branes, also in presence of open string
107: moduli. It turns out that unitarity in the open string channel
108: crucially depends on the presence of the cocycles. We then focus on
109: the 2-loop case and study the degeneration limit where this partition
110: function is directly related to the tree-level 3-point correlators
111: between twist fields. These correlators represent the main ingredient
112: in the computation of Yukawa couplings and other terms in the
113: effective action for D-brane phenomenological models. By factorizing
114: the 2-loop partition function we are able to compute the 3-point
115: couplings for abelian twist fields on generic non-factorized tori,
116: thus generalizing previous expressions valid for the 2-torus.
117:
118: \end{abstract}
119:
120: \vfill
121:
122: \end{titlepage}
123:
124: \sect{Introduction}\label{Intro}
125:
126: The relation between the modular and the unitarity properties of open
127: string amplitudes have played a crucial role in deepening our
128: understanding of string theory. For instance,
129: Lovelace~\cite{Lovelace:1971fa} studied the modular transformation of
130: 1-loop non-planar open string amplitudes. By requiring that these
131: amplitudes do not contain cuts, he discovered the critical dimension
132: of bosonic string theory. More than twenty years later
133: Polchinski~\cite{Polchinski:1995mt} used the same modular
134: transformation between the open and the closed string channels on the
135: 1-loop partition function with Neumann and Dirichlet boundary
136: conditions. By doing so he was able to show that D-branes are actually
137: dynamical objects that have gravitational couplings with closed
138: strings.
139:
140: The same interplay between modular transformations and unitarity
141: properties exists also for higher loop amplitudes. For instance, the
142: open string diagram for the 2-loop planar partition function is a
143: Riemann surface with three boundaries and no handles. This surface
144: can be described either in the closed string channel as in
145: Fig.~\ref{2l}a, or in the open channel as in Fig.~\ref{2l}b. In the
146: first case, by unitarity, one should be able to decompose the result
147: into a $3$-vertex among closed strings and three boundary states. In
148: the open string parametrization, on the other hand, the same amplitude
149: should factorize into three open strings propagators and two $3$-point
150: vertices among open strings. This double description of the 2-loop
151: partition function is particularly interesting when different
152: left/right gluing conditions are imposed on the various boundaries.
153: In~\cite{Russo:2007tc} the bosonic contribution to the twisted
154: $g$-loop partition function was computed by using the closed string
155: description and then was modular transformed in the open string
156: channel. It was also shown that the factorization of the 2-loop
157: diagram provides an effective strategy to compute the couplings among
158: twists fields ($\sigma_\epsilon$), which is alternative to the
159: stress-energy tensor technique of~\cite{Dixon:1986qv}. From the
160: Conformal Field Theory (CFT) point of view, these $\sigma_\epsilon$'s
161: are operators implementing a change in the boundary conditions for the
162: (complexified) bosonic coordinates. The boundary conditions induced
163: by the $\sigma_\epsilon$'s are appropriate to describe open strings
164: stretched between D-branes with constant magnetic
165: fields~\cite{Abouelsaood:1986gd} or D-branes at
166: angles~\cite{Berkooz:1996km}. This kind of open strings is one of the
167: main ingredients in D-brane model building (for a recent review
168: see~\cite{Marchesano:2007de} and the references therein). The
169: $3$-twist field correlators mentioned above provide the non-trivial
170: part of the Yukawa couplings\footnote{Actually the twist field
171: couplings play the same role also in phenomenologically interesting
172: compactifications of Heterotic string
173: theory~\cite{Dixon:1986qv,Burwick:1990tu,Erler:1992gt,Stieberger:1992bj,Stieberger:1992vb}
174: }~\cite{Cremades:2003qj,Cvetic:2003ch,Abel:2003vv,Abel:2003yx,Lust:2004cx,Russo:2007tc}
175: and of other terms in the effective action generated by stringy
176: instantons~\cite{Blumenhagen:2006xt,Ibanez:2006da}.
177:
178: \begin{figure}
179: \begin{center}
180: \input g2.pstex_t
181: \end{center}
182: \caption{\label{2l}The twisted open string partition function with
183: three borders and no handles ({\em i.e.} on each boundary
184: different left/right gluing conditions are imposed). In~\ref{2l}a
185: the amplitude is depicted in the closed string channel, while
186: in~\ref{2l}b it is modular transformed and the open string
187: propagation is manifest.}
188: \end{figure}
189:
190: In this paper we generalize the results of~\cite{Russo:2007tc} in
191: various ways. We compute the bosonic contribution to the planar
192: $g$-loop partition function for open strings on generic tori with both
193: multiply wound D-branes and non-zero Wilson lines or open string
194: moduli. Then we use this result to derive an explicit expression for
195: the 3-twist field correlators valid beyond the case of 2-torus.
196: However, in our computations, we still keep a technical assumption:
197: the monodromy matrices characterizing the open strings stretched
198: between the D-branes must commute, see Eq.~\eq{coms}. This is always
199: the case when the gauge field strengths on the various branes are
200: themselves commuting or parallel. However this more stringent
201: condition is not necessary and we are able to compute the partition
202: function also for some setups involving oblique fluxes (according to
203: the nomenclature of~\cite{Bianchi:2005yz}). On the other hand the
204: assumption~\eq{coms} will restrict our final results on the twist field
205: couplings to the abelian case. The D-brane configurations studied in
206: this paper can be promoted to supersymmetry preserving setups, once
207: they are embedded in superstring theory. This is achieved simply by
208: imposing some constraints on the magnetic fluxes to satisfy the D-term
209: and the F-term conditions. Moreover the latter ones are often
210: automatically implied by our hypothesis~\eq{coms}, since it ensures
211: that there is a particular complex basis for the torus where almost
212: all magnetic fluxes are described by $(1,1)$-forms.
213:
214: As it was done in~\cite{Russo:2007tc}, we compute the twisted
215: partition function in the closed string channel by using the operator
216: formalism (see~\cite{Alvarez-Gaume:1988bg,DiVecchia:1988cy} for
217: detailed discussion). This approach requires particular care in
218: dealing with some phase factors present in the definition of the
219: string vertex operators and boundary states.
220: The origin of these phases is
221: well-known: in toroidal compactifications the logarithmic branch cut of
222: the bosonic Green function can sometime become visible and it is
223: necessary to compensate for this by adding to the string vertices a
224: phase known as cocycle (see for instance the paragraph ``A
225: technicality'' in Sect. 8.2 of~\cite{Polchinski:1998rq}). All cocycle
226: factors were ignored in~\cite{Russo:2007tc}, but this had no visible
227: effects because, as we will show, these phases were not crucial in the
228: particular examples considered there. However, in general, the
229: partition function is unitary, in the open string channel, only when
230: all phases have been taken into account. This comes as no surprise
231: because the partition function of Fig.~\ref{2l}a contains a sum over
232: all possible $3$-string vertices and the effect of the cocycles
233: becomes clearly visible since they yield relative phases between
234: different terms. The presence of these cocycles has some consequences
235: also on the precise formulation of the T-duality transformations. In
236: fact these transformations should preserve both the spectrum and the
237: interactions, including, in the latter case, the phases needed for
238: ensuring the locality of the interactions. This does not happen with
239: the naive version of the T-duality rules usually written, and we show
240: that it is necessary to introduce some cocycle phases also in the
241: T-duality transformation to recover a consistent picture.
242:
243: Then we study the degeneration limit of the $g=2$ partition function
244: in the open string channel, see Fig.~\ref{2l}b, and derive an explicit
245: form for the 3-twist field correlators on arbitrary tori. This
246: computation, in presence of multiply wrapped D-branes, provides a
247: concrete example showing that the cocycle phases are crucial for
248: unitarity already at the level of the 2-loop partition function.
249: Our result on the 3-twist field couplings generalizes previous
250: expressions valid for configurations that are completely factorized in
251: a product of
252: 2-tori~\cite{Cremades:2003qj,Cvetic:2003ch,Abel:2003vv,Abel:2003yx,Lust:2004cx,Russo:2007tc},
253: which requires that both the background geometry and the D-brane gauge
254: field strengths are non-trivial only along the same orthogonal $T^2$'s
255: and all ``off-diagonal'' entries are put to zero. Even in presence of
256: commuting fluxes this is a very particular situation, where the
257: so-called classical contribution to the 3-twist field correlators is
258: written in terms of Jacobi Theta-functions. In general, higher genus
259: Theta-functions appear and so the Yukawa couplings that arise in
260: non-factorized D-brane models will have a more complicated structure
261: and a richer moduli dependence than those derived so far for the
262: $T^2$. However, we still find the same separation between K\"ahler and
263: complex structure moduli and only the latter ones enter in the
264: expressions for the classical contribution (this in the magnetized
265: setup, of course the roles are switched in the language of D-branes at
266: angles).
267:
268: The structure of the paper is the following. In Section~\ref{t2n} we
269: briefly review the main features of the string dynamics in toroidal
270: compactifications. We pay particular care to the cocycle factors in
271: the definition of the vertices and of the boundary states. We show
272: that these phases enter in a non-trivial way also in the T-duality
273: transformation of the states with non-zero Kaluza-Klein or winding
274: numbers. In Section~\ref{SectionAmplitude} we revisit the computation
275: of the twisted partition function discussed in~\cite{Russo:2007tc} and
276: include the effect of the Wilson lines and the D-brane multiple
277: wrappings. The only hypothesis we still take is to restrict ourselves
278: to the case of commuting monodromy matrices (see Eq.~\eq{coms}). In
279: Section~\ref{tft2d} we focus on the $g=2$ case and study the unitarity
280: properties of the result. By doing so we derive the general expression
281: for the 3-point correlators between abelian twist fields. Finally some
282: technicalities are left to the Appendix.
283:
284:
285: \sect{Toroidal compactifications}\label{t2n}
286:
287: \subsection{Some properties of $T^{2d}$}
288:
289: A $2d$-dimensional real torus $T^{2d}$ is defined by a collection of
290: $2d$ vectors $a_M$ in the Euclidean space $\mathbb{R}^{2d}$: $T^{2d}$
291: is simply $\mathbb{R}^{2d}$ modulo the identification with integer
292: shifts along the $a_M$'s
293: \begin{equation}\label{lattice}
294: x \equiv x+2\pi \sqrt{\alpha'}\sum_{M=1}^{2d}c^Ma_M\,\,\,\,,
295: \,\,\,\,\forall x \in
296: \mathbb{R}^{2d}\,\,\,\mathrm{and}\,\,\,c^M \in \mathbb{Z}~.
297: \end{equation}
298: One can choose a Cartesian reference frame and write the components of
299: each vector $a_M$ as $a^a_M$. Then the metric on $T^{2d}$
300: is\footnote{In our conventions the coordinates with the indices
301: $M,N,\dots$ are parallel to the lattice defining the torus; they
302: have period $2\pi \sqrt{\alpha'}$ and are referred to as the
303: integral basis.} $G_{MN}= \sum_{a=1}^{2d}a^a_Ma^a_N$. The components
304: of each $a_M$ can be used to fill in the column vectors of a square
305: matrix $E^a_{\;M}\equiv a^a_M$; then, by construction, $E$ is a
306: vielbein matrix satisfying
307: \begin{equation}\label{vielbein}
308: G={}^tEE\,\,\,\,,\,\,\,\,1={}^tE^{-1}GE^{-1}~.
309: \end{equation}
310: The inverse matrix $E^{-1}$ instead has the dual vectors $\hat{a}^M$
311: as rows: $\sum_{a=1}^{2d}\hat{a}^M_{a} a^a_N=\delta^M_N$. Of course any
312: other matrix $E'=OE$ obtained by means of an orthogonal rotation $O$
313: of $E$ is a good vielbein matrix.
314: We can also introduce complex coordinates and the complex vielbein
315: $\mathcal{E}$, such that
316: \begin{equation}\label{Eps}
317: \mathcal{E}=SE\,\,\,\,,\,\,\,\,S=\frac{1}{\sqrt{2}}\left(
318: \begin{array}{cc}
319: 1 & \ii \\
320: 1 & -\ii
321: \end{array}
322: \right) ~,
323: \end{equation}
324: where all the four blocks of $S$ are proportional to the $d \times d$
325: identity matrix. Two sets of complex coordinates are inequivalent if
326: they cannot be connected by a unitary transformation. Thus the
327: $SO(2d)$ ambiguity in the definition of the $E$'s implies that on the
328: same real torus there is a set of inequivalent complex structures
329: which is parametrized by $SO(2d)/U(d)$. Notice that in the complex
330: coordinates the flat metric $\mathcal{G}$ is off-diagonal
331: \begin{equation}\label{offmetric}
332: \mathcal{G}= {}^t \mathcal{E}^{-1}G\mathcal{E}^{-1}=\bar{S} S^{-1}= \left(
333: \begin{array}{cc}
334: 0 & 1\\
335: 1 & 0
336: \end{array} \right)~.
337: \end{equation}
338:
339: In presence of a constant antisymmetric two-form, there is a
340: particular complex structure that plays a special role. Of course in
341: our applications this 2-form will be the gauge invariant combination
342: between the Kalb-Ramond $B$-field and $F$, the magnetic field on the
343: D-Branes: $\mathcal{F}=B+F$, thus we will indicate this antisymmetric
344: tensor with $\mathcal{F}$ already in this section (notice the
345: different convention with respect to \cite{Russo:2007tc}, as here the
346: factor of $2\pi \alpha'$ in front of $F$ is reabsorbed in the
347: definition of a dimensionless magnetic field).
348: Its expression in a real cartesian basis is
349: \begin{equation}\label{GF}
350: \mathcal{F}_c=\,{}^t\!E^{-1}\mathcal{F}E^{-1}=EG^{-1}\mathcal{F}E^{-1}~.
351: \end{equation}
352: and is related by a similarity transformation to the
353: combination $G^{-1}\mathcal{F}$ which will be the most relevant in the
354: following sections.
355: The antisymmetric matrix $ \mathcal{F}_c$ can be reduced to a
356: block-diagonal form
357: $\mathcal{F}_{\mathrm{b.d.}}$ by means of an orthogonal rotation $O_f$
358: (where the subscript is just to recall that this transformation
359: depends in general on $\mathcal{F}$)
360: \begin{equation}\label{Fbd}
361: \mathcal{F}_{\mathrm{b.d.}}=\left(
362: \begin{array}{cc}
363: 0 & \mathfrak{f}_d\\
364: -\mathfrak{f}_d & 0
365: \end{array}
366: \right)
367: =\,{}^t\!E_f^{-1}\mathcal{F}E_f^{-1}=E_fG^{-1}\mathcal{F}E_f^{-1}~,
368: \end{equation}
369: where $\mathfrak{f}_d$ is a $d \times d$ diagonal matrix with real
370: entries $\mathfrak{f}_{aa}$. The vielbein matrix $E_f=O_fE$ transforms
371: at the same time the metric $G$ into the identity and $\mathcal{F}$
372: into the block-diagonal matrix~\eq{Fbd}. Of course we can use the
373: vielbein matrix $E_f$ to introduce a particular set of complex
374: coordinates ($\mathcal{E}_f=SE_f$) which diagonalizes $G^{-1}\mathcal{F}$
375: \begin{equation}\label{F'}
376: \mathcal{F}^{(d)}=\mathcal{E}_fG^{-1}\mathcal{F}\mathcal{E}^{-1}_f=\left(
377: \begin{array}{cc}
378: -\ii \mathfrak{f}_d & 0\\
379: 0 & \ii \mathfrak{f}_d
380: \end{array} \right)=\mathcal{G}\,{}^t\!
381: \mathcal{E}_f^{-1}\mathcal{F}\mathcal{E}_f^{-1}~,
382: \end{equation}
383: where ${}^t\!\mathcal{E}_f^{-1}\mathcal{F}\mathcal{E}_f^{-1}$ is a
384: block-diagonal matrix. From~\eq{F'} it is easy to see that, in this
385: complex basis, ${\cal F}$ is a $(1,1)$-form. In the following we will
386: always use, as Cartesian basis, the one defined by the vielbeins $E_f$
387: or ${\cal E}_f$, thus we will drop the subscripts without risk of
388: ambiguities. Finally let us recall the definition of the complex
389: structure as the mixed tensor
390: \begin{equation}
391: \mathcal{I}=\ii dz \otimes \frac{\partial}{\partial z}-\ii
392: d\bar{z}\otimes \frac{\partial}{\partial \bar{z}}~.
393: \end{equation}
394: In the following sections we will need the expression of
395: $\mathcal{I}_a^{\phantom{a}b}$ in the integral basis, which is easily
396: derived by using the complex vielbein $\mathcal{E}$:
397: \begin{equation}\label{RealComplexStructure}
398: \mathcal{I}_M^{\phantom{M}N}=\left( {}^t\mathcal{E}
399: \right)_M^{\phantom{M}a}\mathcal{I}_a^{\phantom{a}b}\left(
400: {}^t\mathcal{E}^{-1} \right)_b^{\phantom{b}N}~.
401: \end{equation}
402:
403: \subsection{Closed strings on $T^{2d}$} \label{SectionVertex}
404:
405: The coordinates of the closed strings propagating on $T^{2d}$ with a
406: constant background $B$-field can be written as a sum of left and right
407: handed fields, and the world-sheet is described by a free CFT. Then we
408: have the usual mode expansion
409: \begin{equation}\label{stringcoordinates}
410: x^M_{\mathrm{cl}}(z,\bar{z}) =
411: \frac{X^M_{\mathrm{cl}}(z)+\tilde{X}^M_{\mathrm{cl}}(\bar{z})}{2},
412: \,\,\,\,\mathrm{where}\,\,\,\,\,
413: X^M_{\mathrm{cl}}(z) = x^M-\ii \sqrt{2\alpha'}\alpha_0^M \ln
414: z+\ii \sqrt{2\alpha'}\sum_{m \neq 0}\frac{\alpha_m^Mz^{-m}}{m},
415: \end{equation}
416: and the complexified world-sheet coordinates are
417: \begin{equation}\label{WScoordinates}
418: z=\ex{\tau+\ii \sigma}\,\,\,\,\,\,\mathrm{and}\,\,\,\,\,\,
419: \bar{z}=\ex{\tau-\ii \sigma}~,
420: \end{equation}
421: $\tau$ and $\sigma$ being the world-sheet time and spacial coordinates
422: respectively. Upon canonical quantization, the commutation relations
423: for the oscillators in Eq.~(\ref{stringcoordinates}) are
424: \begin{equation}\label{commutators}
425: \left[ \alpha_m^M,\alpha_{-n}^N \right]=n\delta_{m,n}G^{MN},
426: \,\,\,\,\,\, \mathrm{and}\,\,\,\,\,\, \left[ x^M,\alpha_0^N \right]=\ii
427: \sqrt{2\alpha'}G^{MN}~,
428: \end{equation}
429: and similarly in the right moving sector. The allowed winding and
430: Kaluza-Klein modes are encoded in the Narain lattice which depends on
431: the background fields $G$ and $B$ as follows
432: \begin{equation}\label{zeromodes}
433: \alpha_0^M=\frac{G^{MN}}{\sqrt{2}} \left[ \hat{n}_N+ \left(
434: G_{NN'}-B_{NN'} \right) \hat{m}^{N'}
435: \right]\,\,\,\,,\,\,\,\,\tilde{\alpha_0}^M=\frac{G^{MN}}{\sqrt{2}}
436: \left[ \hat{n}_N- \left( G_{NN'}+B_{NN'} \right) \hat{m}^{N'} \right],
437: \end{equation}
438: where $\hat{n}_M$ and $\hat{m}^M$ are respectively the Kaluza-Klein
439: and winding numbers operators.
440:
441: Let us now consider the interacting theory and pay particular
442: attention to the phases necessary when the target space is compact.
443: The basic building block is the tree-level coupling among three
444: generic closed strings (Reggeon vertex). Since there is no preferred
445: ordering of three points on a sphere, this vertex must be invariant
446: under the action of the permutation group exchanging any of the
447: punctures. The usual emission vertex valid in the uncompact space does
448: not have this property when it is naively generalized to the $T^{2d}$
449: case. This is a well known issue, related to the compactness of the
450: target space, and it is a consequence of the logarithmic branch cut of
451: the bosonic Green function. Similar problems arise also in the usual
452: formalism of the vertex operators describing the emission of
453: particular on-shell string states (see, for instance, Sect.~8.2
454: in~\cite{Polchinski:1998rq}). In order to eliminate this branch cut
455: one has to add suitable cocycle factors to the usual expression of the
456: vertex operators. Here we tackle this issue exactly in the same way by
457: generalising these cocycle factors to the Reggeon vertex formalism.
458:
459: In order to see that the usual $3$-string vertex for closed strings is
460: not invariant under the permutation of the external states when the
461: target space is a torus, it is sufficient to focus on the zero-modes
462: contribution. The explicit form of the full vertex $V_3$ can be found,
463: for instance, in \cite{Russo:2007tc} and its zero-mode part is simply
464: \begin{equation}\label{ExpLn}
465: \exp \left[\sum_{j>i=0}^2 \a^i_0 \ln(z_j-z_i)G \a^j_0 +
466: \sum_{j>i=0}^2 \tilde\a^i_0 \ln(\bar{z}_j-\bar{z}_i)G \tilde\a^j_0
467: \right],
468: \end{equation}
469: where the $z_i~(i=0,1,2)$ are the positions of the three punctures on
470: the sphere (which is represented as the compactified complex plane);
471: the upper index identifies one of the three external states, and all
472: space-times indices have been suppressed. The oscillator part of the
473: $3$-string vertex is invariant under the exchange of the strings $1
474: \leftrightarrow 2$, while the zero-mode contribution~(\ref{ExpLn})
475: gets a phase given by
476: \begin{equation}\label{phase}
477: \exp \left[ -\ii\pi \left( \a_0^1G\a_0^2-\tilde{\a}_0^1G\tilde{\a}_0^2
478: \right) \right] = \exp \left[ -\ii \pi \bigg(
479: \hat{n}_1\hat{m}_2+\hat{m}_1 \hat{n}_2 \bigg) \right],
480: \end{equation}
481: where we have used~\eq{zeromodes}. By using this result, we can build a new
482: invariant vertex with a cocycle factor that compensates for the
483: phase~(\ref{phase}). A possible choice for this cocyle factor is
484: \begin{equation}\label{vertex}
485: V^c_3=V_3 \exp \left[ \frac{\ii \pi}{2} \bigg(
486: \hat{n}_1 \hat{m}_2 -\hat{m}_1 \hat{n}_2 \bigg) \right]~.
487: \end{equation}
488: By using the conservation of the Kaluza-Klein and winding numbers of
489: the emitted strings, the vertex~(\ref{vertex}) is now easily shown to
490: be invariant under the full permutation group acting on the three
491: punctures.
492:
493: However, the cocycle factor added seems to break the vertex invariance
494: under T-Duality transformations. For instance, there is a particular
495: T-duality transformation that exchanges the Kaluza-Klein and winding
496: operators. If its effect could be simply written as $\hat{n}
497: \leftrightarrow \hat{m}$, as it is usually done, then we could have
498: that $V_3^c \to -V_3^c$ for certain external states. In order to
499: clarify this point, let us see the transformation properties of
500: $V_3^c$ under a generic T-duality transformation. These
501: transformations can be encoded in a $O(2d,2d,\mathbb{Z})$
502: matrix~\cite{Giveon:1994fu}
503: \begin{equation}\label{Tduality}
504: T= \left(
505: \begin{array}{cc}
506: a & b\\
507: c & d
508: \end{array}
509: \right)\,\,\,\,,\,\,\,\,\,\, \mathrm{with} \,\,\,\,\,\,
510: a\, {}^t\! b+ b\, {}^t\! a=0
511: \,\,\,,\,\,\,
512: c \,{}^t\! d+d\, {}^t\! c=0\,\,\,,\,\,\,
513: a\, {}^t\!d+b\,{}^t\!c=1~,
514: \end{equation}
515: where $a$, $b$, $c$ and $d$ are $2d \times 2d$ matrices and their
516: constraints follow from the group definition, namely
517: \begin{equation}\label{symplectic}
518: J=TJ\, {}^t T\,\,\,\,\,\,,\,\,\,\,\,\,
519: J=\left(
520: \begin{array}{cc}
521: 0 & 1\\
522: 1 & 0
523: \end{array}
524: \right).
525: \end{equation}
526: Let us observe that, as $J=J^{-1}$, by inverting the relation above,
527: one finds that also ${}^tT$ is an $O(2d,2d,\mathbb{Z})$ matrix with
528: similar constraints imposed on its entries. The duality acts as
529: follows on the geometrical background
530: \begin{equation}\label{E'}
531: G'+B'=\left[ a(G+B)+b \right] \left[ c(G+B)+d \right] ^{-1}
532: \end{equation}
533: and on the string oscillators
534: \begin{equation}\label{a'}
535: \alpha_n =T_+\alpha'_n ~~,~~~~
536: \tilde{\alpha}_n = T_-\tilde{\alpha'}_n~,
537: \end{equation}
538: where
539: \begin{eqnarray}\label{a''}
540: T_{\pm} =
541: \left[c(\pm G+ B)+ d \right] ^{-1}.
542: \end{eqnarray}
543: Notice that one can prove \cite{Giveon:1994fu} that
544: ${}^tT_{\pm}GT_{\pm}=G'$. Focusing now on the zero-modes, we can
545: derive from~(\ref{a'}) the action of the T-duality transformation on
546: the winding and Kaluza-Klein operators
547: \begin{equation}\label{n'm'}
548: \hat{n}={}^td\hat{n}'+{}^tb\hat{m}'\,\,\, ,
549: \,\,\,\,\hat{m}={}^tc\hat{n}'+{}^ta\hat{m}'
550: ~~~\Longleftrightarrow~~~
551: \hat{n}'=a\hat{n}+b\hat{m} \,\,\, ,
552: \,\,\,\, \hat{m}'=c\hat{n}+d\hat{m}
553: \end{equation}
554: Then the transformation law for the vertex~(\ref{vertex}) is
555: \begin{eqnarray}\label{V'}
556: && V^c_3 = V^{c'}_3\ex{\ii \pi \left[
557: {}^t\hat{n}_1'd^tc\hat{n}_2' + {}^t\hat{m}_1'b^ta\hat{m}_2' +
558: {}^t\hat{n}_1'd^ta\hat{m}_2' +
559: {}^t\hat{m}_1'b^tc\hat{n}_2' - {}^t\hat{n}_1'\hat{m}_2' \right]}=\\
560: \nonumber && = V^{c'}_3 \prod_{j=0}^2 \bigg\{ \ex{ \ii \pi
561: \left[ \sum_{M<N}\left( (\hat{n}'_j)_M (d {}^tc)^{MN} (\hat{n}'_j)_N +
562: (\hat{m}'_j)^M (b {}^ta)_{MN} (\hat{m}'_j)^N \right) +
563: {}^t\hat{m}'_jb {}^tc \hat{n}'_j \right]} \bigg\},
564: \end{eqnarray}
565: where in the last line we used of the conservation of windings and
566: momenta. Thus the vertex~(\ref{vertex}) is fully symmetric under
567: permutations of the external states, but is not invariant under the
568: T-Duality transformations~\eq{a'}. However, it is interesting to
569: notice that the phase (which is actually just a sign) generated by the
570: transformations~\eq{a'} can always be written as a product of three
571: terms each one depending only a single external state, as it is done
572: in the second line of~\eq{V'}. This means that the invariance of the
573: vertex~(\ref{vertex}) under T-Duality can be restored, provided that
574: we introduce the appropriate cocycle also in the T-duality
575: transformations, as a generalisation of the standard rules discussed
576: above and in~\cite{Giveon:1994fu}. In fact it is sufficient to
577: postulate that the closed string states transform according
578: to~\eq{n'm'} {\em and} also acquire the same phase in the curly
579: brackets of (\ref{V'})
580: \begin{equation}
581: \label{ts}
582: \ket{n,m} \to
583: \ex{ \ii \pi
584: \left[ \sum_{M<N}\left( (\hat{n}')_M (d {}^tc)^{MN} (\hat{n})_N +
585: (\hat{m}')^M (b {}^ta)_{MN} (\hat{m}')^N \right) +
586: {}^t\hat{m}'b {}^tc \hat{n}'\right]}
587: \; \ket{n',m'}~.
588: \end{equation}
589: We will see in the following section that this is indeed the case when
590: considering boundary states describing magnetised D-Branes as dual for
591: instance to configurations of purely Dirichlet or intersecting
592: D-Branes.
593:
594: It is not difficult to generalize the analysis above to the case of a
595: Reggeon vertex describing the interaction of many closed strings. This
596: can be obtained just by gluing together the 3-point
597: vertices~\eq{vertex} and the result is
598: \begin{equation}\label{Nvertex}
599: V^c_{N+1}=V_{N+1} \exp \left[ \frac{\pi \ii}{2} \sum_{j>i=1}^{N}
600: \bigg(\hat{n}_i \hat{m}_j -\hat{m}_i \hat{n}_j
601: \bigg) \right]~.
602: \end{equation}
603: Here $V_{N+1}$ is the standard vertex, where the cocycles are ignored
604: and, as before, the indices $i,\;j=0,1,\ldots,N$ label the $N+1$
605: external states.
606:
607: \subsection{The Boundary State for a wrapped magnetized D-Brane}\label{SectionBoundaryState}
608:
609: In this section we study, from the closed string point of view, the
610: space-filling magnetised D-Branes with generic wrapping numbers on the
611: torus cycles (see~\cite{Pesando:2005df,DiVecchia:2006gg} for previous
612: works on this subject). In the closed string sector D-Branes are
613: described by boundary states $\ket{B_F}$ that enforce an
614: identification between the left and right moving modes (for a review
615: of the boundary state formalism in the Ramond Neveu-Schwarz formalism
616: see~\cite{DiVecchia:1999rh,DiVecchia:1999fx}). For magnetized D-branes
617: we have
618: \begin{equation}\label{Identification}
619: \left[(G+\mathcal{F}) \alpha_n +
620: (G-\mathcal{F}) \tilde{\alpha}_{-n} \right]
621: \ket{B_F}=0,\,\,\,\,\,\,\forall n \in \mathbb{Z}~,
622: \end{equation}
623: where we used the gauge invariant combination of the Kalb-Ramond
624: $B$-field and the field strength $F$ on the D-brane:
625: $\mathcal{F}=B+F$. In the integral basis, the magnetic field is
626: quantized as a consequence of the compactness of the torus,
627: \begin{equation}\label{Fquantization}
628: F_{MN}=\frac{p_{MN}}{w_Mw_N}~,
629: \end{equation}
630: $p_{MN}$ being an integer matrix and $w_M$ being the wrapping numbers
631: of the D-Brane along the cycles of the torus. Clearly
632: Eq.~\eq{Identification}, being a set of linear constraints, fixes the
633: form of the boundary state up to an overall factor that can also
634: depend on the Kaluza-Klein and winding operators. We will now show
635: that the boundary state for a magnetised D-Brane does indeed contain
636: non-trivial phases that depend on the winding numbers and on the field
637: $F$; moreover it turns out that this phase is strictly related to the
638: phase that closed string states acquire under T-Duality.
639: Following~\cite{Callan:1988wz}, we consider the gauge field
640: contribution to the action in the string path-integral as an
641: interaction term that acts on the standard boundary state for an
642: unmagnetised D-Brane wrapping the $T^{2d}$. We want to derive the
643: dependence of $\ket{B_{F}}$ on the magnetic field by applying the
644: usual path ordered Wilson loop operator
645: $\mathrm{P}[\exp{(\frac{\ii}{2\pi\alpha'}\oint A dx)}]$ to
646: $\ket{B_{F=0}}$. The computation was performed in~\cite{Callan:1988wz}
647: in a Minkowski flat target space, while the novelty of the present
648: calculation\footnote{An analogous calculation is performed
649: in~\cite{Igor:2008}; we thank I. Pesando for letting us know
650: their results before publication.} is the compactness of the torus
651: wrapped by the D-Branes. As the non-zero modes contribution to the
652: boundary state is not affected by the shape of the target space, we
653: will mostly focus on the zero-modes. Of course, our aim is to
654: determine all the $F$-dependent terms that cannot be fixed
655: from~\eq{Identification}. In order to do so we have to pay some care
656: to the definition of the Wilson loop operator.
657:
658: \subsubsection{Gauge bundles and gauge invariant Wilson loop}
659:
660: Any gauge potential for~\eq{Fquantization} will involve a linear, and
661: thus non-periodic, function. Let us start from the simpler case
662: $w_M=1$, $\forall M$. As usual, we need to compensate for the
663: non-periodicity of $A$ by introducing a set of gauge transformations
664: $U_N$. Each $U_N$ encodes the gluing conditions for the gauge
665: potential between two copies of the torus that are adjacent along the
666: $N$-th direction in the covering space. So the gauge potential living
667: on a D-Branes world volume wrapping a cycle of the compactification
668: torus must satisfy
669: \begin{equation}\label{APeriodicity}
670: A_M\left( x+2\pi \sqrt{\alpha'}a_N \right)=U_N(x)\left( 2\pi
671: \alpha' \ii \partial_M + A_M(x) \right) U_N^{\dagger}(x)~,
672: \end{equation}
673: where $a_N$ denotes the $N$-th cycle of the torus. In the case under
674: analysis, all gauge transformations $U_N$ belong to $U(1)$ and so the
675: formula above can be further simplified. We choose not to do it, so as
676: to keep the equations~\eq{APeriodicity}--\eq{U'} valid also for the
677: non-abelian generalization that we will need once we reintroduce
678: multiple wrappings. The background gauge field (gauge bundle) is
679: properly defined by Eq.~\eq{Fquantization} together with the set of
680: $U_N$'s. In order to have a consistent bundle, the gluing matrices
681: must satisfy the overlap condition
682: \begin{equation}\label{consistency}
683: U_N^{\dagger}(x)U_M^{\dagger}\left( x+2\pi \sqrt{\alpha'}a_N \right)
684: U_N \left( x+2\pi \sqrt{\alpha'}a_M \right)U_M(x)=1~.
685: \end{equation}
686: All fields charged under the gauge potential have to obey periodicity
687: conditions similar to~\eq{APeriodicity}. For instance, fields
688: transforming in the fundamental ($\Phi$) or in the adjoint ($\Psi$)
689: representation must satisfy
690: \begin{equation}\label{FunPeriodicity}
691: \Phi\left( x+ 2\pi\sqrt{\alpha'}a_N \right)=U_N(x)\Phi(x)~~,~~~~
692: \Psi\left( x+2\pi \sqrt{\alpha'}a_N \right)=U_N(x)\Psi(x)U_N^{\dagger}(x)~.
693: \end{equation}
694: As a consequence of Eq.~(\ref{FunPeriodicity}), under a generic gauge
695: transformation $\gamma(x)$, the gluing matrices $U_N$ transform in
696: the following fashion
697: \begin{equation}\label{U'}
698: U_N(x) \rightarrow \gamma \left( x+ 2\pi \sqrt{\alpha'}a_N \right)
699: U_N(x) \gamma^{\dagger}(x)~.
700: \end{equation}
701: Notice that there are no restrictions on $\gamma(x)$ and in particular
702: it does not have to be periodic. So the form of the $U_N$'s can change
703: under a gauge transformation.
704:
705: We now focus on the definition of the Wilson loop operator we need
706: for the computation of the magnetized boundary state. Let us consider
707: a path $c$ connecting two points that are separated by the lattice
708: vector $\sum_{L=1}^{2d}m_La_L$, $m_L \in \mathbb{Z}$ (which means that
709: they are identified on the torus). In order to be consistent with our
710: convention for $\mathcal{F}$, this path will start from
711: $x+2\pi\sqrt{\alpha'}\sum_{L=1}^{2d}m_La_L$ and end in $x$. Then it is
712: clear that the naive path ordered Wilson loop operator is not gauge
713: invariant\footnote{and even depends on the initial point $x$ of the
714: path $c$.}, but
715: transforms as
716: \begin{equation}\label{Wilson'}
717: \mathrm{P}[\ex{\ii \int_c A dx}] \rightarrow \gamma(x)\;
718: \mathrm{P}[\ex{\ii \int_c A dx}] \; \gamma^{\dagger}\left( x+
719: 2\pi\sqrt{\alpha'}\sum_{L=1}^{2d} m_L a_L \right) ~.
720: \end{equation}
721: We can recover a gauge invariant object if we multiply the Wilson
722: loop~\eq{Wilson'} by a sequence of $U$'s which forms a discretized
723: version of the path $c$. By using~\eq{consistency} we can choose to
724: collect together all shifts along the direction $K=1$, then those
725: along $K=2$ and so on. In formulae we have
726: \begin{equation}\label{InvariantWilson}
727: \left[ \prod_{K=1}^{2d}\prod_{m=0}^{m_K-1}U_K \left(
728: x+2\pi \sqrt{\alpha'}\sum_{L=1}^{K-1}m_La_L+2\pi
729: \sqrt{\alpha'}ma_K \right)\right]
730: \mathrm{P}[\ex{\frac{\ii}{2\pi\alpha'} \int_c A dx}]~,
731: \end{equation}
732: where only the values $m_K\geq 1$ are relevant (if we have $m_K=0$ for
733: certain $K$, then the corresponding $U_K$ does not appear in the
734: product). By using~\eq{U'} and~\eq{Wilson'} (and~\eq{APeriodicity}),
735: one can check that~\eq{InvariantWilson} is invariant under an
736: arbitrary $U(1)$ gauge transformation $\gamma(x)$ (and does not depend
737: on the initial point $x$ of the path $c$).
738:
739: \subsubsection{Wrapped D-Branes as non abelian gauge bundles}
740:
741: A D-brane with multiple wrappings ($w_M >1$ for some $M$) is better
742: described~\cite{Guralnik:1997sy,Guralnik:1997th} in terms of a
743: non-trivial gauge bundle on the torus $T^{2d}$~\cite{tHooft:1981sz}.
744: In the D-brane language this amounts to considering a set of
745: $w=\prod_{M=1}^{2d}w_M$ coincident D-branes with the same gauge field
746: strength~\eq{Fquantization}, but with non-trivial transition matrices
747: $U_M$. The non-abelian character of the configuration is encoded in
748: the $U_M$'s that, as consequence of~\eq{FunPeriodicity}, glue together
749: the various D-branes in a single wrapped object. In absence of
750: magnetic fields, this can be easily seen by choosing as $U_M$ the
751: following $w_M \times w_M$ transition matrix
752: \begin{equation}\label{P}
753: U_M=P_{w_M \times w_M}=\left(
754: \begin{array}{cccccc}
755: 0 & 1 & 0 & 0 & \cdots & 0\\
756: 0 & 0 & 1 & 0 & \cdots & 0\\
757: 0 & 0 & 0 & 1 & \cdots & 0\\
758: \vdots & \vdots & \vdots & \vdots & \ddots & \vdots \\
759: 1 & 0 & 0 & 0 & \cdots & 0
760: \end{array} \right)~.
761: \end{equation}
762: Notice that, when $F=0$, a D-Brane wrapped $w_M$ times along the
763: $M$-th cycle of the torus can be smoothly deformed, at the classical
764: level, into $w_M$ coincident
765: branes~\cite{Polchinski:1996fm,Hashimoto:1996pd}. This means that
766: there is a family of flat gauge bundles interpolating between~\eq{P}
767: and $U_M=1$.
768:
769: In order to introduce the effect of the magnetic field
770: (\ref{Fquantization}) on the D-Brane, it turns out to be convenient to
771: choose a fundamental cell of the torus lattice where this field is in
772: a block-diagonal form. This can be always done~\cite{Griffiths} (see
773: also the Appendix for a more pedestrian proof), so from now on we take
774: \begin{equation}\label{Fblock}
775: F=\left(
776: \begin{array}{ccccc}
777: 0 & \frac{p_1}{W_1} & 0 & 0 & \cdots\\
778: -\frac{p_1}{W_1} & 0 & 0 & 0 & \cdots\\
779: 0 & 0 & 0 & \frac{p_2}{W_2} & \cdots\\
780: 0 & 0 & -\frac{p_2}{W_2} & 0 & \cdots\\
781: \vdots & \vdots & \vdots & \vdots & \ddots
782: \end{array} \right)~,
783: \end{equation}
784: where $p_{\alpha} \in \mathbb{Z}$ and $W_{\alpha} \in
785: \mathbb{Z}-\{0\}$, $\forall \alpha=1,\dots,d$. Notice that even if the
786: field $F$ now describes a direct product of $d$ $T^2$'s inside the
787: $T^{2d}$, the compactification is in general non factorizable as a
788: consequence of the form of the metric. The $p_{\alpha}$'s can be
789: interpreted as the Chern classes of the magnetic fields while the
790: $W_{\alpha}$'s are the products of the couples of wrapping numbers on
791: each of the $T^2$'s inside $T^{2d}$.
792:
793: Two comments are in order now. First, if $p_\alpha$ and $W_\alpha$ are
794: not co-prime, the configuration can again be smoothly deformed, at the
795: classical level, to a new configuration with co-prime $p'_\alpha$ and
796: $W'_\alpha$ and $p'_\alpha/W'_\alpha=p_\alpha/W_\alpha$. Second it is
797: not necessary to specify all the wrappings of the brane along each
798: cycle of the torus. In fact it is possible to show that configurations
799: of magnetized D-Branes with the same product of wrappings along the
800: pairs of cycles of the $T^2$'s inside the $T^{2d}$ defined by the form
801: of the field (\ref{Fblock}) are equivalent. Indeed the transition
802: matrices defining the gauge bundle of a brane wrapped along a $T^2$
803: are related by a gauge transformation if they describe branes with the
804: same Chern class $p$ and the same product of the wrappings $W$. Thus
805: we can choose to wrap the branes $W_{\alpha}$ times along the even
806: directions $x_M\equiv x_{2\alpha}$ only. We will also make the
807: following gauge choice for the gauge potential~(\ref{Fblock})
808: \begin{equation}\label{gauge}
809: A_{M \equiv 2\alpha}(x)=\frac{p_{\alpha}}{W_{\alpha}}x_{2\alpha-1}+
810: 2\pi\sqrt{\alpha'}C_{2\alpha-1}
811: \,\,\,\,\,\,\mathrm{and}\,\,\,\,\,\,
812: A_{M \equiv 2\alpha-1}(x)=2\pi\sqrt{\alpha'}C_{2\alpha}~,\,\,\,\,\,\,
813: \forall \alpha=1,\dots d\,,
814: \end{equation}
815: where we have introduced non-zero Wilson lines $2\pi
816: \sqrt{\alpha'}C_M$, with $C_M$ adimensional. In this way the non
817: abelian gauge bundle describing the magnetized wrapped D-Brane is
818: characterized by
819: \begin{eqnarray}\label{U}
820: U_{2\alpha}(x) &=& 1_{W_1 \times W_1} \otimes \ldots \otimes
821: P_{W_{\alpha} \times W_{\alpha}} \otimes 1_{W_{\alpha+1} \times
822: W_{\alpha+1}} \otimes \ldots \otimes 1_{W_{d} \times W_{d}}
823: \\ \nonumber
824: U_{2\alpha-1}(x) &=& 1_{W_1 \times W_1} \otimes \ldots \otimes
825: (Q_{W_{\alpha} \times W_{\alpha}})^{p_{\alpha}} \otimes 1_{W_{\alpha+1} \times
826: W_{\alpha+1}} \otimes \ldots \otimes 1_{W_{d} \times W_{d}}
827: ~\ex{^{\frac{\ii}{\sqrt{\alpha'}}\frac{p_{\alpha}}{W_{\alpha}}x_{2\alpha}}}\,,
828: \end{eqnarray}
829: where $Q_{W_{\alpha} \times W_{\alpha}}=\mathrm{diag}\left\{ 1,\ex{
830: 2\pi\ii /W_{\alpha}},\dots,\ex{2\pi \ii (W_{\alpha}-1)/W_{\alpha}}
831: \right\}$. Notice that the form of the transition matrices $U_M$ is
832: not affected by the presence of the Wilson-lines as for them the
833: relation (\ref{APeriodicity}) is trivially satisfied by the identity
834: gluing matrix.
835:
836: The generalization of~\eq{InvariantWilson} to this non-abelian setup
837: is straightforward and the operator we need to use to derive the
838: magnetized boundary state from the unmagnetized one reads
839: \begin{equation}\label{WilsonOperator}
840: {\cal O}_A = \Tr \left\{
841: \left[ \prod_{K=1}^{2d}\prod_{m=0}^{\hat{m}_K-1}U_K
842: \left(x_{\rm cl}+\sum_{L=1}^{K-1}2\pi
843: \sqrt{\alpha'}\hat{m}_La_L+2\pi \sqrt{\alpha'}m a_K \right)
844: \right] \mathrm{P}[\ex{\frac{\ii}{2\pi\alpha'} \int_c A dx_{\rm
845: cl}}] \right\}\,,
846: \end{equation}
847: where the $x_{\rm cl}^M$'s are the usual string
848: coordinates~\eq{stringcoordinates} and the $\hat{m}_M$'s are the
849: operators that read the winding numbers of the closed string states.
850: In this non-abelian generalization we have to put $U(x)$ at the right
851: hand of the sequence and then follow the order determined by the path
852: $c$.
853:
854: \subsubsection{Computation of the Boundary State}
855:
856: We can now compute the action of ${\cal O}_{A}$ on the unmagnetized
857: boundary state $\ket{B_{F,C}} = {\cal O}_A \ket{B_{A=0}}$. The
858: unmagnetized boundary state for a wrapped D-Brane is found from the
859: one of an unwrapped brane (see~\cite{DiVecchia:1999fx} and references
860: therein) by applying the same operator (\ref{WilsonOperator}) with the
861: choice
862: \begin{equation}
863: A_M=0\,\,\,\,\,\, \mathrm{and}\,\,\,\,\,\,
864: U_K=1_{w_1 \times w_1} \otimes \ldots \otimes P_{w_K \times
865: w_K} \otimes 1_{w_{K+1} \times
866: w_{K+1}} \otimes \ldots \times 1_{w_{2d} \times w_{2d}}~,
867: \end{equation}
868: with $P$ defined as in Eq.~(\ref{P}). In this case the trace
869: in~\eq{WilsonOperator} reads
870: \begin{equation}
871: \mathrm{Tr}\left[ \prod_{K=1}^{2d}U_K^{\hat{m}_K} \right]
872: \end{equation}
873: and it is different from zero only when the windings of the emitted
874: closed strings are integer multiples of the wrappings of the D-Brane
875: on each cycle of the torus. Hence only these states couple to the
876: wrapped D-brane, as expected, and we have
877: \begin{equation}\label{BF0}
878: \ket{B_{A=0}}=\sqrt{\mathrm{Det}(G+B)}\sum_{m^M \in Z} \prod_{n=1}^{\infty}
879: \ex{-\frac{1}{n}\tilde{\a}_n^{\dagger}G R_0\,\a_n^{\dagger}}
880: \ket{0;w_M m^M}~,
881: \end{equation}
882: where there is no sum understood over the repeated index $M$;
883: $R_0=(G-B)^{-1}(G+B)$ is the identification matrix between left and
884: right moving oscillators and depends on the geometric background of
885: the torus; finally the ket $\ket{0;w_Mm^M}$ represents the closed
886: string vacuum state with zero Kaluza-Klein momenta and winding numbers
887: equal to $w_M m^M$ for $M=1,2,...,2d$.
888:
889: It is easy to begin by turning on only the Wilson lines and to keep
890: vanishing magnetic fields on the D-brane world volume. We have just to
891: isolate the Wilson line contribution to $\mathcal{O}_A$
892: in~\eq{WilsonOperator} when acting on $\ket{B_{A=0}}$, namely
893: \begin{equation}\label{WL}
894: \ket{B_{C}}=
895: \ex{\frac{\ii}{\sqrt{\alpha'}} \int_c C \cdot dx}~
896: \ket{B_{A=0}}=\ex{2\pi \ii C \cdot \hat{m}}\ket{B_{A=0}}~.
897: \end{equation}
898: The explicit evaluation of the contributions of the magnetic fields
899: $F$ is longer. It can be split into the zero and non-zero mode part
900: of the string coordinate $x_{\rm cl}(z,\bar{z})$ defined
901: in~\eq{stringcoordinates}. Let us focus on the zero-mode part of the
902: computation, since the non-zero mode contribution has just the effect
903: of replacing $B$ with $\mathcal{F}$ in~(\ref{BF0}),
904: see~\cite{Callan:1988wz}. A first result is that the magnetized
905: boundary state couples only with the closed strings whose windings in
906: the $\alpha$-th $T^2$ (as defined by the form of the magnetic field
907: in~(\ref{Fblock})) are integer multiples of $W_{\alpha}$. This is
908: again a consequence of the trace in~\eq{WilsonOperator} where the
909: transition matrices are defined as in Eq.~(\ref{U}). It is convenient
910: to define the $2d \times 2d$ matrix
911: \begin{equation}\label{wma}
912: w=\left(
913: \begin{array}{ccc}
914: W_1 & 0 & \cdots\\
915: 0 & W_2 & \cdots\\
916: \vdots & \vdots & \ddots
917: \end{array} \right)~\otimes 1_{2\times 2}
918: \end{equation}
919: and use $w m$ to indicate the $2d$ column vector containing the
920: windings of the closed strings emitted by the magnetized D-Brane. We
921: can also see how the action of ${\cal O}_A$ yields the relation
922: between windings and Kaluza-Klein numbers
923: \begin{equation}\label{nFm}
924: \hat{n} = -F \hat{m}~,
925: \end{equation}
926: which is usually derived from the identification imposed by
927: Eq.~(\ref{Identification}) on the closed string zero-modes. For this,
928: it is sufficient to focus on the zero-modes contribution linear both
929: in the position operator $\hat{q}^M=\left(x^M+\tilde{x}^M \right)/2$
930: and in the oscillators $\a_0$ or $\tilde{\a_0}$. Using the form of $F$
931: in Eq.~(\ref{Fblock}) with the gauge choice~(\ref{gauge}) and the
932: transition matrices~\eq{U}, we can evaluate this contribution as
933: follows
934: \begin{eqnarray}
935: \nonumber \ket{B_F} & \sim &
936: \ex{\frac{\ii}{\sqrt{\alpha'}}\sum_{\alpha=1}^d
937: \frac{p_{\alpha}}{W_{\alpha}}\hat{m}_{2\alpha-1}\hat{q}_{2\alpha}}
938: \ex{-\ii \frac{\sqrt{\alpha'}}{2\sqrt{2}}
939: \sum_{\alpha=1}^d\int_{0}^{2\pi}d\sigma \frac{1}{2\pi
940: \alpha'}\frac{p_{\alpha}}{W_{\alpha}}(x_{2\alpha-1}+\tilde{x}_{2\alpha-1})
941: (\a_0^{2\alpha}-\tilde{\a}_0^{2\alpha})}\ket{0;wm}=\\
942: & = &
943: \ex{\ii
944: \sum_{\alpha=1}^d\frac{p_{\alpha}}{W_{\alpha}}\hat{m}_{2\alpha-1}
945: \frac{\hat{q}_{2\alpha}}{\sqrt{\alpha'}}}\ex{-\ii \sum_{\alpha=1}^n
946: \frac{p_{\alpha}}{W_{\alpha}}\frac{\hat{q}_{2\alpha-1}}{\sqrt{\alpha'}}
947: \hat{m}_{2\alpha}}\ket{0;wm}=
948: \ket{-Fwm,wm}~.
949: \end{eqnarray}
950: Notice that at this stage we can forget about the path-ordering in the
951: Wilson operator $\mathcal{O}_A$ and explicitly perform the integration
952: in the first line of the previous equation, as the zero-mode
953: contributions of the string fields entering the Wilson loop in
954: $\mathcal{O}_A$ commute with each other at different values of
955: $\sigma$. Finally let us consider the terms quadratic in the
956: zero-modes $\alpha_0$ and $\tilde{\alpha}_0$ that follow from the
957: standard Wilson loop exponential in Eq.~(\ref{WilsonOperator}),
958: \begin{eqnarray} \label{QuadraticZeroModes}
959: &&
960: \ex{-\frac{\ii}{2}\sum_{\alpha=1}^d\int_{0}^{2\pi}d\sigma
961: \frac{1}{2\pi}\frac{p_{\alpha}}{W_{\alpha}}\left(
962: \a_0^{2\alpha-1}\a_0^{2\alpha}-\tilde{\a}_0^{2\alpha-1}
963: \a_0^{2\alpha}-\a_{0}^{2\alpha-1}\tilde{\a}_0^{2\alpha}+
964: \tilde{\a}_0^{2\alpha-1}\tilde{\a}_0^{2\mu}\right) \sigma}
965: \ket{-Fwm;wm}=
966: \\ \nonumber &&
967: = \ex{-\ii \pi \sum_{\alpha=1}^d W_{\alpha}p_{\alpha}
968: m_{2\alpha-1}m_{2\alpha}}\ket{-Fwm;wm} =
969: \ex{-\ii\pi\sum_{M<N}\hat{m}^M F_{MN} \hat{m}^N}
970: \ket{-Fwm;wm}~.
971: \end{eqnarray}
972: The phase in~(\ref{QuadraticZeroModes}) could not have been deduced
973: just by looking at the constraints~(\ref{Identification}). So the
974: expression for the boundary state describing a magnetised D-Brane is
975: \begin{eqnarray}\label{WLBF}
976: \ket{B_{F,C}}= \sqrt{\mathrm{Det}(G+\mathcal{F})}
977: & \sum_{m \in \mathbb{Z}^{2d}} &\!
978: \ex{-\ii\pi\!\sum_{M<N}\hat{m}^MF_{MN}\hat{m}^N}
979: \, \ex{2\pi \ii C \hat{m}}
980: \\ \nonumber & \times & \!\!\!
981: \left[\prod_{n=1}^{\infty}\ex{-\frac{1}{n}\tilde{\a}_n^{\dagger}
982: G R\, \a_n^{\dagger}} \right]
983: \ket{-Fwm ,wm}~,
984: \end{eqnarray}
985: where the identification matrix $R$ is
986: \begin{equation}
987: \label{eq:imr}
988: R=(G-\mathcal{F})^{-1}(G+\mathcal{F})~.
989: \end{equation}
990: Even if the phase in Eq.~\eq{WLBF} has been calculated for a block
991: diagonal $F$, in the Appendix we will prove that Eq.~\eq{WLBF} holds
992: for a generic $F$ (see Eq.~\eq{Fquantization}), a part for possible
993: half integer shifts in the Wilson line.
994:
995: Let us now analyze how Eq.~\eq{WLBF} transforms under an
996: $O(2d,2d,\mathbb{Z})$ transformation~(\ref{Tduality}). Generically,
997: after the T-duality, we have a new magnetized D-brane with
998: \begin{equation}\label{Fduality}
999: F' = (aF - b)(-cF +d)^{-1}~,~~~~R'=T_-^{-1}R T_+,
1000: \end{equation}
1001: as it can be seen by using the relations in~(\ref{n'm'}); moreover,
1002: since $F$ is an antisymmetric matrix, also $F'$ is antisymmetric,
1003: thanks to constraints in~(\ref{Tduality}). If the combination
1004: $(-cF+d)$ in~\eq{Fduality} is not invertible, then the transformed
1005: D-brane will have some direction with Dirichlet boundary conditions.
1006: For instance, we can check that any magnetized brane can be easily
1007: related to a lower dimensional D-Brane at angles via T-Duality. First
1008: let us put the magnetic field in the block-diagonal form, as in
1009: Eq.~(\ref{Fblock}). Then, we T-dualize the even direction of each of
1010: the $T^2$'s defined by $F$ inside the $T^{2d}$, that is, we choose the
1011: following $O(2d,2d,\mathbb{Z})$ matrix
1012: \begin{equation}
1013: a=d= 1_{d \times d} \otimes \left(
1014: \begin{array}{cc}
1015: 1 & 0\\
1016: 0 & 0
1017: \end{array} \right)\,\,\,\,\,\,
1018: \mathrm{and}\,\,\,\,\,\,
1019: b=c= 1_{d \times d} \otimes \left(
1020: \begin{array}{cc}
1021: 0 & 0\\
1022: 0 & 1
1023: \end{array} \right)~.
1024: \end{equation}
1025: By using the action of the duality transformations of the Kaluza-Klein
1026: and winding operators, as in Eq.~(\ref{n'm'}), the phase
1027: (\ref{QuadraticZeroModes}) can be rewritten as
1028: \begin{equation}
1029: \ex{-\ii \pi
1030: \sum_{\alpha=1}^d\frac{p_{\alpha}}{W_{\alpha}}
1031: (W_{\alpha}^2)m_{2\alpha-1}m_{2\alpha}}
1032: = \ex{-\ii \pi \sum_{\alpha=1}^d\hat{m}'_{2\alpha}\hat{n}'_{2\alpha}}~.
1033: \end{equation}
1034: Notice that this phase exactly compensates for the one that the closed
1035: string states in the ket of~\eq{WLBF} acquire under T-Duality, as
1036: stated in Eq.~(\ref{ts}). Thus we see that the boundary states for
1037: purely geometrical configurations of D-Branes (like brane intersecting
1038: at angles) do not contain any non trivial phase depending on the
1039: emitted closed strings zero-modes, as expected. By reinstating the
1040: $g_s$ dependence and using the results of~\cite{Myers:1999ps}, it is
1041: possible to show that also the prefactor
1042: $\sqrt{\mathrm{Det}(G+\mathcal{F})}$ transforms into the one expected
1043: for the boundary state of a D-Brane at angle.
1044:
1045: It is also possible to transform any magnetized D-brane into a D-brane
1046: with Dirichlet boundary conditions along all the coordinates of the
1047: torus. In this case, the matrices $c$ and $d$ defining the T-duality
1048: are related to the magnetic field $F= c^{-1}d$. When $F$ is
1049: block-diagonal~(\ref{Fblock}),
1050: one can choose $c=w$ given by ~(\ref{wma}) and easily build integer matrices
1051: $a$ and $b$ satisfying Eqs.~(\ref{Tduality}) using the fact that
1052: the wrappings $W_\alpha$ and the Chern numbers
1053: $p_\alpha$ are coprime. Exactly as in the previous example, the phase of the
1054: magnetized boundary state cancels against the phase generated by the
1055: T-duality transformation~(\ref{ts}). Thus one recovers the standard
1056: form of a Dirichlet boundary state, where the identification matrix is
1057: simply $R=-1$.
1058:
1059: Finally the transformations on the closed strings
1060: zero-modes~(\ref{ts}) show explicitly that the Wilson lines
1061: in~(\ref{WLBF}) are related to the positions of the D-Brane if the
1062: dualized directions have Dirichlet boundary condition. For the case of
1063: the D-branes at angles, in each of the two-dimensional tori inside the
1064: $T^{2d}$, half of the components of the Wilson lines becomes positions
1065: and the remaining ones are still interpreted as residual Wilson lines
1066: on the dualized brane. When the D-brane is transformed into a point in
1067: the compact space, then all components of the Wilson lines are
1068: geometrized into positions of the dual D-brane.
1069:
1070: \sect{The twisted partition function}\label{SectionAmplitude}
1071:
1072: In this section we will compute the bosonic contribution to the open
1073: string twisted partition function for multiply wrapped D-branes. The
1074: planar partition function $Z_g(F)$ is obtained by starting from the
1075: tree-level vertex~(\ref{Nvertex}) and sewing the external legs with
1076: boundary states describing magnetized D-Branes with Wilson lines
1077: turned on~(\ref{WLBF}). From the worldsheet point of view, this means
1078: that we start with a sphere and cut out $ g+1$ boundaries
1079: representing the magnetised D-branes. Thus we are dealing with a
1080: Riemann surface of genus zero, with $g+1$ borders and no crosscaps; in
1081: the open string channel it corresponds to a $g$-loop diagram.
1082:
1083: We start from the result obtained in~\cite{Russo:2007tc} for D-branes
1084: without multiple wrappings and where all cocycle phases were
1085: ignored\footnote{We follow as much as possible the conventions of that
1086: paper.}. In the open string channel, one of the D-branes (whose
1087: identification matrix is indicated with $R_0$) is singled out as the
1088: external border of the diagram; thus it is natural to introduce the
1089: monodromy matrices ${\cal S}_\mu \equiv R_{0}^{-1} R_{\mu}$, with
1090: $\mu=1,\ldots,g$, whose eigenvalues are $\ex{\pm
1091: 2\pi\ii\e_\mu^\alpha}$
1092: ~($\alpha=1,2,..,d$).
1093: The only assumption we will make on the monodromy matrices
1094: $\mathcal{S}_{\mu}$ is that they commute with each other, namely that
1095: \begin{equation}\label{coms}
1096: \left[ \mathcal{S}_{\mu},\mathcal{S}_{\nu} \right]=0~.
1097: \end{equation}
1098: Notice that this does not imply that the identification matrices $R_i$
1099: with $i=0,\dots,g$ also commute with each other. Of course the
1100: converse holds and~\eq{coms} is implied by the requirement that
1101: $\left[R_i,R_j \right]=0$. By following the classification
1102: of~\cite{Bianchi:2005yz} this more restrictive constraint is related
1103: to configurations with parallel magnetic fluxes. However, while
1104: Eq.~\eq{coms} is invariant under the T-Duality, the constraint among
1105: the identification matrices $R$ is not, as they do not transform by
1106: a similarity transformation (see Eq.~\eq{Fduality}).
1107: Thus we will consider the slightly more
1108: general class of configurations satisfying~\eq{coms}. In this case, it
1109: is convenient to perform a T-Duality and transform the zero-th D-Brane
1110: into a purely Dirichlet D-Brane, i.e. with $R_0=-1$. As discussed in
1111: the previous section, this can be done with a T-duality having $c^{-1}
1112: d=F_0$. With this choice we have $\mathcal{S}_{\mu}=-R_{\mu}$ and the
1113: commutator above can be rewritten as
1114: \begin{equation}
1115: \label{parallel}
1116: \left[ G^{-1} {\cal F}_{\mu},G^{-1} {\cal F}_{\nu} \right]=0~,
1117: \end{equation}
1118: with $\mu=1,\ldots,g$ only. This implies that there exists a complex
1119: basis in which all the $G^{-1}\mathcal{F}_{\mu}$ are diagonal as in
1120: Eq.~(\ref{F'}).
1121: Moreover it implies:
1122: \begin{equation}
1123: \label{comF}
1124: \left[ G^{-1}(F_{\hat\mu}-F_g),G^{-1}(F_{\hat\nu}-F_g) \right]=0,~
1125: ~~~~~~\forall {\hat\mu},{\hat\nu}=1,\dots,g-1~;
1126: \end{equation}
1127: as a consequence, for a generic $G$ one can deduce that a fundamental
1128: cell of the lattice torus exists where all the field differences
1129: $(F_{\hat\mu}-F_g)$ are simultaneously block diagonal.
1130: At any time, we can use again the T-duality rules
1131: discussed in the previous section and go back to the original system
1132: with all magnetized D-branes.
1133:
1134:
1135: We start from Eq.~(3.29) of~\cite{Russo:2007tc} describing the
1136: $g$-loop partition function in the open string channel
1137: \begin{equation} \label{opchT}
1138: Z_g (F) \sim \left[ \prod_{\mu = 1}^{g} \sqrt{{\rm
1139: Det}\left(1-G^{-1} {\cal F}_\mu\right)}\right]
1140: \int\left[d Z \right]_g \;\mathcal{A}^{(0)}\;
1141: \prod_{\alpha=1}^d \left[ \ex{- \ii \pi \vec{\e^\alpha} \cdot \tau \cdot
1142: \vec{\e}^\alpha} \; \frac{\det \tau }{\det
1143: T_{\vec{\e}^\alpha} } \; {\cal R}_g \left(
1144: \vec{\e}^\alpha \cdot \tau \right) \right]\,,
1145: \end{equation}
1146: where $\left[d Z \right]_g$ is the untwisted ($F_i=0$) result and
1147: $\vec{\e}^\alpha$ collects in a vector of length $g$ all the twists
1148: ${\e}^\alpha$; $\tau$ and $T_{\vec{\e}^\alpha}$ are the standard and the
1149: twisted period matrix respectively.~~~
1150: $\mathcal{A}^{(0)}$ is the
1151: classical contribution to the partition function calculated in
1152: \cite{Russo:2007tc} in absence of the cocycle phases
1153: and setting all of the D-branes wrappings to one:
1154: \begin{equation}\label{claw1}
1155: \mathcal{A}^{(0)} =
1156: \sum\, \Delta\;
1157: \exp\!\left\{\!\pi\ii \sum_{\hat\mu,\hat\nu=1}^{g-1} \a_0^{\hat\mu} G {\cal
1158: S}_{\hat\mu}^{-1/2} \bm{D}_{\hat{\mu}\hat{\nu}}{\cal S}_{\hat{\nu}}^{1/2}
1159: ~\a_0^{\hat\nu}\right\} ~,
1160: \end{equation}
1161: where $\bm{D}_{\hat{\mu}\hat{\nu}}$ is a space-time matrix determined
1162: by the ${\cal S}_\mu$'s and the sum is over all the winding numbers
1163: that satisfy the Kronecker's deltas representing the
1164: identification~\eq{nFm} for each boundary state and the Kaluza-Klein
1165: and winding conservations (recall that closed strings emitted by the
1166: D-Brane with $R_0=-1$ are characterized by unconstrained Kaluza-Klein
1167: momenta and no winding numbers):
1168: \begin{equation}
1169: \label{Kdelta}
1170: \Delta = \left[
1171: \prod_{\mu=1}^g \delta\left(\hat{n}_{\mu}+
1172: F_{\mu} \hat{m}_{\mu}\right)\right]
1173: \delta\left(\sum_{i=0}^g \hat{n}_i\right)
1174: \delta\left(\sum_{\mu=1}^g \hat{m}_{\mu}\right)~.
1175: \end{equation}
1176: All the $\e$-dependent ingredients, including the function ${\cal
1177: R}_g$ and the matrix $\bm{D}$, are defined in~\cite{Russo:2007tc}
1178: and, of course, depend on the moduli of the Riemann surface. We do
1179: not need the precise form of all these ingredients, but only some
1180: properties that we will recall later in this section. Notice that the
1181: classical contribution~\eq{claw1} is nontrivial only for $g \geq 2$:
1182: for the annulus we have $\mathcal{A}^{(0)} = 1$ and Eq.~\eq{opchT}
1183: reduces to the partition function in the uncompact
1184: space~\cite{Bachas:1992bh}.
1185:
1186: In this section we complete~\eq{claw1} to include also the effects of
1187: the cocycle phases, multiple wrappings and open string moduli (Wilson
1188: lines and/or D-brane positions). The classical contribution is the
1189: only part of the partition function that is affected by this
1190: generalization, as it is clear from the form of the interaction
1191: vertex~(\ref{Nvertex}) and of the magnetized D-Branes boundary
1192: states~(\ref{WLBF}). Basically we need to include in the sewing
1193: procedure the cocycle factor in~(\ref{Nvertex}) and the
1194: phases~
1195: (\ref{WLBF}). Of course by
1196: following this approach we are effectively working in the closed
1197: string description. However the new contribution is independent of the
1198: world-sheet moduli and thus can be directly included in Eq.~\eq{opchT}
1199: which is written in the open string channel. It is then clear that,
1200: in the expression for $\mathcal{A}$, we will have the same exponential
1201: of Eq.~\eq{claw1} multiplied by some additional factors related to
1202: cocycles. So let us consider these new contributions: by using~\eq{Nvertex},
1203: ~\eq{nFm} and~\eq{WLBF}, one can see that all the phases from the cocycles and
1204: the Wilson lines yield
1205: \begin{eqnarray}\label{coco}
1206: && \mathrm{exp} \left\{ 2\pi \ii \left[ \sum_{\mu=1}^{g}
1207: C_{\mu} \hat{m}_{\mu} +Y_0\hat{n}_0 \right] \right\}
1208: \times \exp \left\{ \pi \ii \sum_{\mu=1}^{g}
1209: \sum_{M<N}\hat{m}_{\mu}^M (F_{\mu})_{MN}\, \hat{m}_{\mu}^N \right\}\\
1210: & \times & \nonumber \exp\left\{ \frac{\pi \ii}{2} \sum_{\nu>\mu=1}^{g}
1211: \left(\hat{m}_\mu F_\mu \hat{m}_\nu + \hat{m}_\mu
1212: F_\nu \hat{m}_\nu \right)
1213: \right\} ~.
1214: \end{eqnarray}
1215: where $Y_0$ encodes the position of the zero-th D-Brane which is
1216: point-like along the torus directions. By using~\eq{Kdelta}, we can
1217: eliminate $\hat m_g$ from these sums. Then it is easy to see that we
1218: can rewrite the dependence on the open string moduli as follows:
1219: \begin{equation}\label{wld}
1220: \exp\left\{2\pi\ii\sum_{\hat{\mu}=1}^{g-1} \left[ C_{\hat{\mu}}- C_{g}+
1221: Y_0(F_{\hat{\mu}}-F_g) \right]
1222: \hat{m}_{\hat{\mu}}\right\} \equiv
1223: \ex{2\pi\ii \sum\limits_{\hat{\mu}=1}^{g-1} {}^t\!\rho_{\hat{\mu}}
1224: \hat{m}_{\hat{\mu}}}\,.
1225: \end{equation}
1226: The second exponential in~\eq{coco} comes from the boundary states;
1227: using the conservation of the winding numbers, the exponent can be
1228: rewritten as follows:
1229: \begin{equation}
1230: \sum_{\mu=1}^g\sum_{M<N} \hat{m}_{\mu} ^M (F_{\mu})_{MN}
1231: \hat{m}_{\mu}^N = \sum_{\hat{\mu}=1}^{g-1}\sum_{M<N}
1232: \hat{m}_{\hat{\mu}}^M (F_{\hat{\mu}})_{MN} \hat{m}_{\hat{\mu}}^N
1233: - \sum_{\hat{\mu},\hat{\nu}=1}^{g-1} \sum_{M<N}
1234: \hat{m}_{\hat{\mu}}^M (F_g)_{MN} \hat{m}_{\hat{\nu}}^N~.
1235: \end{equation}
1236: By combining this contribution with the
1237: last exponent in Eq.~\eq{coco} and using~\eq{Kdelta} we get
1238: \begin{equation}\label{e1final}
1239: \exp\left[-\pi\ii \sum_{\hat{\mu},\hat{\nu}=1}^{g-1}
1240: \sum_{M<N} \hat m_{\hat{\mu}}^M
1241: (F_{\hat{\mu}\hat{\nu}})_{MN} \hat m_\nu^N \right] ~,
1242: \end{equation}
1243: where
1244: \begin{equation}\label{Fmunu}
1245: F_{\hat{\mu}\hat{\nu}}= F_{\hat{\nu}\hat{\mu}}=F_{\hat{\nu}}-F_g
1246: ~~,~~~\mbox{for}~\hat\nu\geqslant\hat\mu.
1247: \end{equation}
1248: Observe that in order to obtain this final form we have changed the sign of the combination $\sum_{M<N}\hat{m}^M_{\mu}(F_{\mu})_{MN}\hat{m}^N_{\mu}$ as each term of the sum is integer (see the Appendix).
1249:
1250: Thus the total contribution from the
1251: various phase factors is just the product of~\eq{wld}
1252: and~\eq{e1final}. This expression has no dependence on the metric of
1253: the torus and, in particular, Eq.~\eq{e1final} provides just some
1254: relative signs between contributions related to different values of
1255: $\hat{m}$.
1256: Moreover it depends only on the differences $(F_{\hat{\nu}}-F_g)$;
1257: therefore, thanks to Eq.~\eq{comF}, we can always consider
1258: $F_{\hat{\mu}\hat{\nu}}$ block diagonal.
1259:
1260: Let us now reconsider the exponential in~\eq{claw1}. By
1261: using~\eq{zeromodes}, \eq{nFm}, \eq{parallel} and the various
1262: conservation laws we can rewrite it as follows
1263: \begin{equation}
1264: \label{erew}
1265: \exp \left\{ \frac{\pi \ii}{2}
1266: \sum_{\hat{\mu},\hat{\nu}=1}^{g-1}{}^t\hat{m}_{\hat{\mu}}
1267: G\left[ 1- \left( G^{-1}\mathcal{F}_{\hat{\mu}}\right)^2
1268: \right]^{\frac{1}{2}}
1269: \bm{D}_{\hat{\mu}\hat{\nu}}(\mathcal{S})
1270: \left[ 1- \left(
1271: G^{-1}\mathcal{F}_{\hat{\nu}} \right)^2
1272: \right]^{\frac{1}{2}} \hat{m}_{\hat{\nu}} \right\}~,
1273: \end{equation}
1274: where we explicitly remind that $\bm{D}$ is a function of the
1275: space-time matrices ${\cal S}$. It is convenient to rewrite~\eq{erew}
1276: in the complex basis defined by the vielbein $\mathcal{E}$, where the
1277: $G^{-1}\mathcal{F}_i$'s are diagonal and denoted by
1278: $\mathcal{F}^{(d)}_i$ as in Eq.~(\ref{F'}):
1279: \begin{equation}\label{erew2}
1280: \mathrm{exp}\left\{ \frac{\pi\ii}{2} \sum_{\hat{\mu},\hat{\nu}=1}^{g-1}
1281: {}^t\hat{m}_{\hat{\mu}} \left[{}^t\!\mathcal{E}
1282: \mathcal{G}\sqrt{\left( 1-(\mathcal{F}_{\hat{\mu}}^{(d)})^{2} \right)
1283: \left( 1-(\mathcal{F}^{(d)}_{\hat{\nu}})^2 \right)} \left(
1284: \begin{array}{cc}
1285: \mathcal{D}_{\hat{\mu}\hat{\nu}}(\epsilon) & 0\\
1286: 0 & \mathcal{D}_{\hat{\mu}\hat{\nu}}(-\epsilon)
1287: \end{array} \right)\mathcal{E}\right]
1288: \hat{m}_{\hat{\nu}}\right\}~,
1289: \end{equation}
1290: where now each $\mathcal{D}_{\hat{\mu}\hat{\nu}}$ is $d\times d$
1291: diagonal matrix that depends on the eigenvalues of the ${\cal
1292: S}_\mu$'s.
1293:
1294: The square
1295: parenthesis in~\eq{erew2} is contracted with a symmetric combination
1296: of $\hat{m}$, so we can symmetrize it. Then, by using~\eq{offmetric}, one
1297: can easily check that~\eq{erew2} is equal to
1298: \begin{equation}\label{erew3}
1299: \mathrm{exp}\left\{ \frac{\pi\ii}{2} \sum_{\hat{\mu},\hat{\nu}=1}^{g-1}
1300: {}^t\hat{m}_{\hat{\mu}} \left[{}^t\!\mathcal{E}
1301: \mathcal{G}\sqrt{\left( 1-(\mathcal{F}^{(d)}_{\hat{\mu}})^2 \right)
1302: \left( 1-(\mathcal{F}^{(d)}_{\hat{\nu}})^2 \right)} \left(
1303: \begin{array}{cc}
1304: \hat{\tau}_{\hat{\mu}\hat{\nu}} & 0\\
1305: 0 & \hat{\tau}_{\hat{\nu}\hat{\mu}}
1306: \end{array} \right)\mathcal{E}\right]
1307: \hat{m}_{\hat{\nu}}\right\}~,
1308: \end{equation}
1309: where the $d\times d$ diagonal matrix\footnote{Our
1310: $\hat{\tau}$ is equal to the $\tau$ of~\cite{Antoniadis:2005sd}; we
1311: indicate it with a different symbol in order to avoid confusion with
1312: the standard period matrix.}
1313: $\hat\tau_{\hat{\mu}\hat{\nu}}$
1314: is given by:
1315: \begin{equation}
1316: \label{hattau}
1317: \hat\tau_{\hat{\mu}\hat{\nu}} \equiv \frac 12 \left[
1318: \mathcal{D}_{\hat{\mu}\hat{\nu}}(\e) +
1319: \mathcal{D}_{\hat{\nu}\hat{\mu}}(-\e)\right]~.
1320: \end{equation}
1321:
1322: Expressing $\mathcal{E}$ in terms of the real vielbein $E$
1323: ($\mathcal{E}= S E$, with $S$ given in Eq.~\eq{Eps}) we can go to the
1324: real basis,
1325: where Eq.~\eq{erew3} reads:
1326: \begin{equation}\label{erew3'}
1327: \mathrm{exp}\left\{ \frac{\pi\ii}{2} \sum_{\hat{\mu},\hat{\nu}=1}^{g-1}
1328: {}^t\hat{m}_{\hat{\mu}} \left[{}^t\!E
1329: \sqrt{\left( 1-(\mathcal{F}^{(d)}_{\hat{\mu}})^2 \right)
1330: \left( 1-(\mathcal{F}^{(d)}_{\hat{\nu}})^2 \right)} \left(
1331: \begin{array}{cc}
1332: \hat{\tau}^S & \ii\hat{\tau}^A\\
1333: -\ii\hat{\tau}^A& \hat{\tau}^S
1334: \end{array} \right)_{\hat{\mu}\hat{\nu}} E\right]
1335: \hat{m}_{\hat{\nu}}\right\}~,
1336: \end{equation}
1337: where $\hat\tau^S$ and $\hat{\tau}^A$ are the symmetric
1338: and the antisymmetric part of $\hat{\tau}$, in the
1339: exchange of $\hat{\mu},\;\hat{\nu}$.
1340: As $\hat{\tau}$ is purely imaginary and $Im~ \hat{\tau}^S$ is
1341: positive definite because of the
1342: Riemann bilinear identities~\cite{Antoniadis:2005sd} and moreover
1343: $\left( \sqrt{ 1-(\mathcal{F}^{(d)}_{\hat{\mu}})^2} \right)_{ab}=
1344: \delta_{ab}~|\left(1-\mathcal{F}^{(d)}_{\hat{\mu}}\right)_{aa}|$,
1345: the convergence of the series in Eq.~\eq{claw1} is assured.
1346:
1347: Following \cite{Bianchi:2005sa} we can rewrite the Born-Infeld square
1348: roots above in yet another way by
1349: using another important consequence of the
1350: Riemann bilinear identities~\cite{Antoniadis:2005sd}:
1351: \begin{equation}\label{C}
1352: C_{\hat{\mu}\hat{\nu}} =C_{\hat{\nu}\hat{\mu}} \equiv
1353: \frac{1}{2}\left[ \mathcal{D}_{\hat{\mu}\hat{\nu}}(\epsilon)-
1354: \mathcal{D}_{\hat{\nu}\hat{\mu}}(-\epsilon)\right] =
1355: \ii \frac{\sin(\pi \epsilon_{\hat{\mu}})
1356: \sin(\pi \epsilon_{\hat{\nu}}-\pi\epsilon_{g})}
1357: {\sin(\pi \epsilon_g)},\,\,\,\,\,\,
1358: \hat{\nu}\geqslant \hat{\mu}~,
1359: \end{equation}
1360: where also $C_{\hat{\mu}\hat{\nu}}$ and the sines are $d\times d$
1361: diagonal matrices whose entries depend on the different
1362: values of $\e$. This will be useful for the analysis of the
1363: degeneration limit in the next section. The $2d\times 2d$ matrix~
1364: $\mathrm{diag}\left\{ \sin(\pi \epsilon_{\mu}),\sin(\pi\epsilon_{\mu})
1365: \right\}$~ can be written as\footnote{We will not keep track of the
1366: sign choices for the square roots: they clearly depend on whether
1367: each $\e_\mu$ is negative or positive.}:
1368: \begin{equation} \label{tru}
1369: \left(
1370: \begin{array}{cc}
1371: \sin(\pi \epsilon_{\mu}) & 0\\
1372: 0 & \sin(\pi \epsilon_{\mu})
1373: \end{array} \right)
1374: =\sqrt{\frac{1}{1-(\mathcal{F}^{(d)}_{\mu})^2}}~.
1375: \end{equation}
1376: This can be checked by rewriting the sine in terms of exponentials
1377: which are directly related to the components of ${\cal S}$ in the
1378: complex basis: $\sin(\pi \epsilon^\a_{\mu})=[\sqrt{2-{\cal
1379: S}_\mu^{-1}-{\cal S}_\mu}\,]_{\a\a}/2,~~\a=1,2,..,d $. Also, using the same
1380: procedure, we have
1381: \begin{equation}\label{tru1}
1382: \left(
1383: \begin{array}{cc}
1384: \sin( \pi \epsilon_{\hat{\nu}}-\pi \epsilon_g) & 0\\
1385: 0 & \sin( \pi \epsilon_{\hat{\nu}}-\pi \epsilon_g)
1386: \end{array} \right)=\frac{ \left(
1387: \begin{array}{cc}
1388: \ii & 0\\
1389: 0 & -\ii
1390: \end{array} \right)
1391: (\mathcal{F}^{(d)}_g-\mathcal{F}^{(d)}_{\hat{\nu}})}
1392: {\sqrt{\left( 1-(\mathcal{F}^{(d)}_g)^2\right)
1393: \left( 1-(\mathcal{F}^{(d)}_{\hat{\nu}})^2 \right)}}~,
1394: \end{equation}
1395: with $\mathcal{F}^{(d)}_{\mu}$ defined as in Eq.~(\ref{F'}). From~\eq{tru}
1396: and~\eq{tru1} one can see that
1397: \begin{equation}\label{Cfinal}
1398: C_{\hat{\mu}\hat{\nu}}=C_{\hat{\nu}\hat{\mu}}= \frac{ \left(
1399: \begin{array}{cc}
1400: 1 & 0\\
1401: 0 & -1
1402: \end{array} \right)
1403: \left( \mathcal{F}^{(d)}_{\hat{\nu}}-\mathcal{F}^{(d)}_g\right)}
1404: {\sqrt{\left( 1-(\mathcal{F}^{(d)}_{\hat{\mu}})^2 \right)
1405: \left( 1-(\mathcal{F}^{(d)}_{\hat{\nu}})^2\right)}}~~,
1406: \,\,\,\,\,\,\mbox{for}~~~
1407: \hat{\nu}\geqslant \hat{\mu}~.
1408: \end{equation}
1409: Thus we can eliminate the square roots in~\eq{erew3} in favor of $C$;
1410: then it is convenient to decompose $\hat\tau$ into its symmetric
1411: ($\hat{\tau}^S$) and the antisymmetric ($\hat{\tau}^A$) parts.
1412: Hence we can use Eq.~(\ref{F'}) to rewrite the
1413: diagonal fields $ \mathcal{F}^{(d)}$
1414: in terms of the $\mathcal{F}$'s, take advantage of the identity
1415: $\mathcal{F}_\mu-\mathcal{F}_\nu = F_\mu-F_\nu $, insert the
1416: contribution of the cocycles given by Eq.~\eq{e1final} and finally get the
1417: classical contribution to the twisted partition function
1418: describing wrapped D-branes on an generic $T^{2d}$ :
1419: \begin{eqnarray}
1420: \mathcal{A} & =& \sum \Delta \;
1421: \mathrm{exp}\left\{{\pi\ii}
1422: \sum_{\hat{\mu},\hat{\nu}=1}^{g-1} \sum_{M<N}
1423: {}^t\hat{m}_{\hat{\mu}}^M (F_{\hat{\mu}\hat{\nu}})_{MN}
1424: \hat{m}_{\hat{\nu}}^N\right\}\;
1425: \ex{2\pi \ii \sum\limits_{\hat{\mu}=1}^{g-1} {}^t\!\rho_{\hat\mu} \hat
1426: m_{\hat\mu}}
1427: \label{awv1} \\ \nonumber &\times&
1428: \mathrm{exp}
1429: \left\{\frac{\pi\ii}{2} \sum_{\hat{\mu},\hat{\nu}=1}^{g-1}
1430: {}^t\hat{m}_{\hat{\mu}}
1431: {}^t\mathcal{E} \left[ \left(
1432: \begin{array}{cc}
1433: \frac{\hat\tau^S}{C} & 0\\
1434: 0 & -\frac{\hat\tau^S}{C}
1435: \end{array} \right)_{\hat{\mu}\hat{\nu}} \!\!\! +
1436: \left(
1437: \begin{array}{cc}
1438: \frac{\hat\tau^A}{C} & 0\\
1439: 0 & \frac{\hat\tau^A}{C}
1440: \end{array} \right)_{\hat{\mu}\hat{\nu}}
1441: \right] {}^t\mathcal{E}^{-1} F_{\hat{\mu}\hat{\nu}}
1442: \hat{m}_{\hat{\nu}} \right\}~.
1443: \end{eqnarray}
1444: where $ F_{\hat{\mu}\hat{\nu}}$ is the matrix defined in
1445: Eq.~\eq{Fmunu}\footnote{The inverse of $C$ must be meant only with
1446: respect to the Lorentz indeces, at fixed $\hat{\mu}$ and
1447: $\hat{\nu}$.}. Then the full partition function is simply given
1448: by~\eq{opchT}, where $\mathcal{A}^{(0)}$ is substituted by the
1449: complete expression above.
1450:
1451: If we restrict ourselves to the case of a factorizable torus
1452: $T^{2d}=(T^2)^d$, then~\eq{awv1} agrees\footnote{Apart from some
1453: factors of two.} with the results of~\cite{Antoniadis:2005sd}. In
1454: order to make contact with their setup it is first useful to perform a
1455: T-Duality in such a way that the singled-out boundary with
1456: identifications encoded in the $R_0$ matrix is transformed back to a
1457: magnetized D-Brane. From~\eq{Fduality} we can compute the
1458: transformation of the difference between two gauge field strengths
1459: \begin{equation}
1460: \label{fdiff}
1461: F'_i-F'_j = (F_j\,{}^t c + \,{}^t d)^{-1} (F_j-F_i) (cF_i-d)^{-1}~.
1462: \end{equation}
1463: In particular the duality we performed to put $R_0=-1$ had $d=cF_0$.
1464: Then by combining Eq.~(\ref{fdiff}) with the transformation of the
1465: winding numbers (\ref{n'm'}), the amplitude~\eq{awv1} reduces to the
1466: following product of terms, each one related to a single $T^2$:
1467: \begin{eqnarray}
1468: \mathcal{A}^w_{(T^2)^d} \!\!\!& =& \! \sum \Delta \;
1469: \mathrm{exp}\left\{{\pi\ii}
1470: \sum_{\hat{\mu},\hat{\nu}=1}^{g-1} \sum_{M<N}
1471: {}^t\hat{m}_{\hat{\mu}}^M (F_{\hat\mu\hat\nu})_{MN}
1472: \hat{m}_{\hat{\nu}}^N\right\}\;
1473: \ex{2\pi \ii \sum\limits_{\hat{\mu}=1}^{g-1} {}^t\!\rho_{\hat\mu} \hat
1474: m_{\hat\mu}}
1475: \nonumber \\ &\times& \label{t2fact}
1476: \prod_{\a=1}^d\mathrm{exp}\left\{\frac{\pi}{2}
1477: \sum_{\hat{\mu},\hat{\nu}=1}^{g-1} {}^t \hat m^\a_{\hat{\mu}}\left[
1478: \left(\frac{\tau_S(\epsilon^\a)}{C(\epsilon^\a)}\right)_{\hat{\mu}\hat\nu}
1479: \!\! \mathcal{I}^{(\a)} +\ii \left(
1480: \frac{\tau_A(\epsilon^\a)}{C(\epsilon^\a)}\right)_{\hat{\mu}\hat\nu}\right]
1481: F^{(\a)}_{\hat{\mu}\hat\nu} \hat{m}^\a_{\hat\nu} \right\} ~,
1482: \end{eqnarray}
1483: where now
1484: \begin{equation}
1485: F_{\hat{\mu}\hat{\nu}}=F_{\hat{\nu}\hat{\mu}}=(F_0-F_{\hat{\mu}})(F_0-F_g)^{-1}
1486: (F_{\hat{\nu}}-F_g)
1487: ~~,~~~\mbox{for}~\hat\nu\geqslant\hat\mu,
1488: \end{equation}
1489: and $\mathcal{I}^{(\a)}$ is the complex structure of each
1490: of the $T^2$'s defined as Eq.~(\ref{RealComplexStructure}). This can
1491: be compared with Eq.~(A.28) of~\cite{Antoniadis:2005sd}. The norm of
1492: the vector $v_i$ appearing there is related to the Born-Infeld square
1493: roots: ${|v_i U|}/{\sqrt{U_2 T_2}} = w_{\hat{\mu}} \sqrt{\left(
1494: 1-(\mathcal{F}^{(d)}_{\hat{\mu}})^2 \right)}$. One can use this
1495: in~\eq{Cfinal} and~\eq{t2f} together with the explicit expression for
1496: the $T^2$ complex structure~\eq{cst2} to check that the two results
1497: are related by a further T-duality that exchanges $T\leftrightarrow
1498: -1/U$.
1499:
1500: \sect{Twist fields couplings on $T^{2d}$.} \label{tft2d}
1501:
1502: In this section we will focus on the vacuum diagram with three
1503: boundaries (i.e. $g=2$) and study the degeneration limit where all
1504: three open string propagators in Fig.~\ref{2l}b become long and thin.
1505: In this situation the partition function factorizes in two tree-level
1506: 3-point correlators between twist fields. This result provides the
1507: main contribution for the computation of the Yukawa couplings in
1508: string phenomenological models and of other terms in the effective
1509: action generated by stringy
1510: instantons~\cite{Blumenhagen:2006xt,Ibanez:2006da}. For $g=2$ the only
1511: non-vanishing entry for $\hat\tau$ is clearly $\hat\tau^S_{11}$, which
1512: has been computed, at leading order in this degeneration limit,
1513: in~\cite{Russo:2007tc}. Again the novelty of the present computation
1514: resides in the analysis of the classical part~\eq{awv1}, while all
1515: other ingredients of the partition function~\eq{opchT} factorize
1516: exactly as before, since they do not depend on the wrapping numbers or
1517: the Wilson lines. In the degeneration limit under study we
1518: have~\cite{Russo:2007tc} that $\mathcal{D}_{11}(\epsilon) \rightarrow
1519: 0$ from which
1520: \begin{equation}
1521: \left( \frac{\hat{\tau}^s(\epsilon)}{C(\epsilon)} \right)_{11}
1522: \rightarrow -1_{d\times d}
1523: \end{equation}
1524: thus the exponential in the second line of Eq.~(\ref{awv1}) becomes
1525: \begin{equation}\label{e2finalg2}
1526: \mathrm{exp}\left\{\frac{\pi}{2}{}^t \hat{m}_1 \mathcal{I}(F_2-F_1)
1527: \hat{m}_1 \right\}~,
1528: \end{equation}
1529: where we have introduced the complex structure of the torus in the
1530: integral basis as in Eq.~(\ref{RealComplexStructure}). In order to
1531: write the final form of the amplitude as a sum over unconstrained
1532: integers, it is necessary to solve the conservations (\ref{Kdelta}),
1533: which, in the case under study, become
1534: \begin{equation}
1535: \hat{n}_1=-F_1\hat{m}_1\,\,\,,\,\,\,\hat{n}_2=-F_2\hat{m}_2\,\,\,,\,\,\,
1536: \hat{m}_1+\hat{m}_2=0~.
1537: \end{equation}
1538: Of course the solutions must have integer Kaluza-Klein and windings
1539: numbers, so there must exist a minimal\footnote{By minimal we mean
1540: that any other matrix with the same property is an integer
1541: multiple of $H$.} integer invertible matrix $H$ such that $F_1 H$
1542: and $F_2 H$ are integer matrices and the solution can be written as
1543: $\hat{m}_1=Hh$, with $h \in \mathbb{Z}^{2d}$. Then we define
1544: $\mathcal{I}'={}^t\!H\mathcal{I}\,{}^t\!H^{-1}$, which is still a
1545: complex structure, and $F\equiv {}^t\!H(F_2-F_1)H$; so the
1546: degeneration limit of the amplitude (\ref{awv1}) in our $g=2$ case is
1547: \begin{equation}\label{Azm2}
1548: \mathcal{A}= \sum_{h \in \mathbb{Z}^{2d}} \mathrm{exp}
1549: \left\{ \frac{\pi}{2}\,\left[{}^t\!h \mathcal{I}'F h+
1550: \ii \sum_{M<N} h^M F_{MN} h^N \right] \right\}\times \mathrm{exp}
1551: \left\{ 2\pi \ii \,{}^t\!\rho_1 Hh \right\} ~,
1552: \end{equation}
1553: with a possible half-integer shift of $\rho_1$ (see the Appendix for
1554: further details).
1555:
1556: By unitarity it must be possible to rewrite the result in~(\ref{Azm2})
1557: as a sum where each term is the product of two functions that are one
1558: the complex conjugate of the other. Each function represents the
1559: classical contribution to the coupling among three twist fields, while
1560: the quantum contribution follows from the factorization of the other
1561: terms in the partition functions~\eq{opchT}, see~\cite{Russo:2007tc}.
1562: The presence of the sum is due to the fact that the vacuum describing an
1563: open string stretched between two magnetized D-branes has a finite
1564: degeneracy~\cite{Abouelsaood:1986gd} in compact spaces. So each string
1565: state has a number of replica and the various terms in the sum
1566: describe the couplings between these different copies of the twist
1567: fields (of course this is exactly what we need, in phenomenological
1568: models, to describe the Yukawa couplings for different families).
1569: Let us see how this works in the simple case of the 2-torus.
1570:
1571: \subsection{The two-torus example}
1572:
1573: For a generic tilted torus the metric can be written as a function of
1574: the K\"ahler and complex structure moduli
1575: \begin{equation}\label{G}
1576: G=\frac{T_2}{U_2} \left(
1577: \begin{array}{cc}
1578: 1 & U_1\\
1579: U_1 & |U|^2
1580: \end{array} \right)
1581: ={}^t \mathcal{E} \left(
1582: \begin{array}{cc}
1583: 0 & 1\\
1584: 1 & 0
1585: \end{array} \right)
1586: \mathcal{E},
1587: \end{equation}
1588: having defined the complex structure as $U=U_1+\ii U_2$ and the
1589: K\"ahler form as $T=T_1+\ii T_2$. Thus the vielbein (\ref{offmetric})
1590: reads
1591: \begin{equation}\label{t2v}
1592: \mathcal{E}=\sqrt{\frac{T_2}{2U_2}}\left(
1593: \begin{array}{cc}
1594: 1 & U\\
1595: 1 & \bar{U}
1596: \end{array} \right)
1597: \end{equation}
1598: from which one can write the explicit form of the $T^2$ complex
1599: structure in the integral basis
1600: \begin{equation}\label{cst2}
1601: \mathcal{I}={}^t \mathcal{E} \left(
1602: \begin{array}{cc}
1603: \ii & 0\\
1604: 0 & -\ii
1605: \end{array} \right)
1606: {}^t \mathcal{E}^{-1}=-\frac{1}{U_2}\left(
1607: \begin{array}{cc}
1608: U_1 & -1\\
1609: |U|^2 & -U_1
1610: \end{array} \right),
1611: \end{equation}
1612: The magnetic fields on the two magnetized D-Branes are identified by the
1613: Chern numbers $p_i$ and the products of the wrappings along the
1614: two cycles of the torus $W_i$:
1615: \begin{equation}\label{t2f}
1616: F_i= \left(
1617: \begin{array}{cc}
1618: 0 & \frac{p_i}{W_i}\\
1619: -\frac{p_i}{W_i} & 0
1620: \end{array} \right),~~~~i=1,2.
1621: \end{equation}
1622: One can easily check that the matrix $H$ is simply proportional to the
1623: $2 \times 2$ identity, namely $H=W_1W_2/\delta \otimes 1_{2 \times
1624: 2}$, where $\delta=\mathrm{G.C.D}\left\{ W_1,W_2\right\}$. This
1625: implies that Eq.~(\ref{Azm2}) can be put in the following form
1626: \footnote{In the configuration of \cite{Russo:2007tc} one gets $I>0$,
1627: otherwise one should write $|I|$, because of the note before
1628: Eq.~\eq{tru}.}
1629: \begin{equation}\label{n422}
1630: \sum_{h_1,h_2 \in \mathbb{Z}}\exp \left\{ -\frac{\pi}{2}\frac{I}{U_2}
1631: \left[ h_1^2+|U|^2h_2^2+2Uh_1h_2 \right] +
1632: 2\pi \ii \frac{1}{\delta}\mathcal{C}_M h^M \right\},
1633: \end{equation}
1634: with
1635: \begin{equation}\label{I}
1636: I=\frac{W_1^2W_2^2}{\delta^2}
1637: \left(\frac{p_2}{W_2}-\frac{p_1}{W_1}\right)=
1638: I_{21}\frac{W_1W_2}{\delta^2} ~,
1639: \end{equation}
1640: where we introduced the intersection numbers
1641: \begin{equation}\label{Iij}
1642: I_{ij}=p_iW_j-p_jW_i
1643: \end{equation}
1644: and
1645: \begin{equation}
1646: \mathcal{C}_M=W_1W_2\left[(F_1-F_2)Y_0+(C_1-C_2)\right]_M~.
1647: \end{equation}
1648: We want to perform a T-Duality in such a way that also the zero-th D-Brane
1649: is magnetized. The intersection numbers are invariant under this
1650: operation, while the form of the matrix $I$ is modified due to the
1651: transformation of the wrapping numbers. Recalling that the T-Duality
1652: relating a Dirichlet to a magnetized brane is encoded in an
1653: $O(2,2,\mathbb{Z})$ matrix of the type in Eq.~(\ref{Tduality})
1654: with\footnote{We indicate with a tilde the quantities in the picture
1655: with a magnetized zero-th D-Brane.}
1656: \begin{equation}\label{zeroDuality}
1657: d=\left(
1658: \begin{array}{cc}
1659: 0 & -p_0\\
1660: p_0 & 0
1661: \end{array} \right)\,\,\,\,\mathrm{and}\,\,\,\,c=\tilde{w}_0 \equiv
1662: \tilde{W}_0 \otimes 1_{2\times 2}~,
1663: \end{equation}
1664: it is possible to show, combining the invariance of $I_{21}$ with
1665: Eq.~(\ref{fdiff}), that $W_1 \rightarrow I_{01}$ and $W_2 \rightarrow
1666: I_{20}$. Thus
1667: \begin{equation}\label{I'}
1668: I
1669: =\frac{I_{01}I_{21}I_{20}}{\delta^2}
1670: \end{equation}
1671: with $\delta=\mathrm{G.C.D.}\left\{ I_{20},I_{10}\right\}
1672: =\mathrm{G.C.D.}\left\{ I_{20},I_{10},I_{21}\right\}$, since we can make
1673: use of the property
1674: $I_{21}\tilde{W}_0+I_{20}\tilde{W}_1+I_{01}\tilde{W}_2=0$. Under the
1675: same duality the open string moduli transform as follows
1676: \begin{equation}
1677: C_{\mu}=\frac{\tilde{C}_{\mu}}{\tilde{w}_0(F_{0}-F_{\mu})}
1678: \,\,\,\,\,\,\mathrm{and} \,\,\,\,\,\,Y_0=\tilde{w}_0\tilde{C}_0
1679: \end{equation}
1680: hence
1681: \begin{equation}
1682: {\mathcal{C}}_M=
1683: \tilde{W}_0I_{12}\tilde{C}^{(0)}_M+\tilde{W}_1I_{20}\tilde{C}^{(1)}_M+
1684: \tilde{W}_2I_{01}\tilde{C}^{(2)}_M.
1685: \end{equation}
1686: where the superscript $(i)$ indicates the three boundaries and the
1687: subscript $M$ is the Lorentz index.
1688:
1689: The configurations studied~\cite{Russo:2007tc} had $\delta$ and all
1690: $\tilde{W}_i$ equal to $1$. In this case $I$ is always an even number,
1691: since it is the product of three integers that sum to zero. Thus the
1692: contribution of the cocycles in~\eq{Azm2} is irrelevant and~\eq{n422}
1693: can be rewritten as it was done in~\cite{Russo:2007tc}. We choose not
1694: to do that here, because it is easier to deal always with~\eq{n422}
1695: without treating the case $\tilde{W}_i=1$ separately. In order to
1696: factorize the amplitude above
1697: and find the Yukawa couplings corresponding to the states of the open
1698: strings stretched between pairs of D-Branes with different magnetic
1699: fields on their world volume, it is necessary to first perform a
1700: Poisson resummation on the integer $h_1$
1701: \begin{equation}
1702: \sum_{h_1=-\infty}^{+\infty}\ex{-\pi Ah_1^2+2\pi
1703: h^1As}=\frac{1}{\sqrt{A}}
1704: \ex{\pi As^2}\sum_{h_1=-\infty}^{+\infty}
1705: \ex{-\pi \frac{h_1^2}{A}-2\pi \ii h_1s}~,~~~~~A>0.
1706: \end{equation}
1707: This yields a new form of the same amplitude which is easy to
1708: factorize once we introduce a new pair of integers, $r$ and $k$,
1709: through the relations
1710: \begin{equation}
1711: h_1= rI+l =I \left( r+\frac{l}{I} \right)
1712: \,\,\,\,\,\,\mathrm{and}\,\,\,\,\,\, h_2=k-r~,
1713: \end{equation}
1714: where $l=1,\dots,I$. It is manifest that in this way both the former
1715: and the latter pair of integers range in the whole $\mathbb{Z}$.
1716: Notice that this is ensured by summing over the additional integer $l$
1717: as well. Simple algebraic manipulations then lead to the product of
1718: two Jacobi Theta-functions as follows
1719: \begin{equation}
1720: \mathcal{A}=\sqrt{\frac{2U_2}{I}}\sum_{l=1}^{I} \vartheta \left[
1721: \begin{array}{c}
1722: \frac{l}{I}-\frac{1}{I}\frac{\mathcal{C}_1}{\delta}\\
1723: \frac{\mathcal{C}_2}{\delta}
1724: \end{array} \right] \left( 0 \big| IU \right) \times \vartheta \left[
1725: \begin{array}{c}
1726: \frac{l}{I}-\frac{1}{I}\frac{\mathcal{C}_1}{\delta}\\
1727: -\frac{\mathcal{C}_2}{\delta}
1728: \end{array} \right] \left( 0 \big| -I\bar{U} \right)
1729: \end{equation}
1730: This result generalizes the one of ref.\cite{Russo:2007tc} and is in
1731: agreement\footnote{The apparent mismatch related to
1732: the presence of the wrapping numbers of the D-Branes in the Wilson
1733: lines dependence of the couplings is resolved by checking that
1734: $W_iC^{(i)}_1\,,\,W_iC^{(i)}_2 \in [0,1]$ as well as the parameters
1735: $\epsilon_i$ and $\theta_i$ defined in \cite{Cremades:2003qj}.} with
1736: the Section 3.1.3 of \cite{Cremades:2003qj}, as we find that, if the
1737: intersection numbers $I_{ij}$ are not coprime, one has
1738: $I_{20}I_{01}I_{21}/\delta^2$ non vanishing Yukawa couplings, labeled
1739: by the integer $l$. Indeed, in the dual picture involving intersecting
1740: D-Branes, every open string living in the intersection between two
1741: fixed D-Branes, say for instance $i$ and $j$, will only couple to
1742: $|I_{jk}I_{ik}|/\delta^2$ strings from the intersections between the
1743: D-Brane $k$ and the D-Branes $i$ and $j$ \cite{Cremades:2003qj}.
1744:
1745: \subsection{Twist fields couplings on a generic $T^{2d}$ compactification}
1746:
1747: In order to factorize the classical contribution to the partition
1748: function~(\ref{Azm2}) in the most general case analyzed in Section
1749: \ref{SectionAmplitude} and read the corresponding twist field
1750: couplings, we need to use the properties of the complex structure. Let
1751: us first decompose ${\cal I}'$ in terms of its $d \times d$ blocks
1752: \begin{equation}
1753: \mathcal{I}'=\left(
1754: \begin{array}{cc}
1755: A & B\\
1756: C & D
1757: \end{array} \right)~.
1758: \end{equation}
1759: Hence it is simple to check that ${\cal I'}^2=-1$ yields
1760: \begin{equation}
1761: \label{ABCD}
1762: AB=-BD~~,~~~ \mbox{and}~~ A^2=-(1+BC)~.
1763: \end{equation}
1764: Then we choose a basis for the torus lattice where the combination
1765: ${}^t\!H(F_2-F_1)H$ takes the following form
1766: \begin{equation}\label{Fblock1}
1767: F=\left(
1768: \begin{array}{cc}
1769: 0 & \hat{F}\\
1770: -\hat{F} & 0
1771: \end{array} \right).
1772: \end{equation}
1773: This can be done by putting the matrix $F$ in the form of
1774: Eq.~(\ref{Fblock}) (thanks again to the result of the Appendix) and by
1775: a suitable relabeling of the rows and the columns. Notice that this
1776: relabeling does not affect the form of the amplitude to be factorized.
1777: For sake of brevity, we also introduce the $2d$-components vector
1778: $\beta=\,{}^tH\rho_1$. Then the general amplitude to be factorized has
1779: the following form\footnote{In the following $i$ indicates a $d\times
1780: d$ imaginary matrix: $i\equiv \ii\, 1_{d\times d}$.}
1781: \begin{equation}\label{Azm3}
1782: \sum_{h_i \in \mathbb{Z}^d} \exp \left\{ \frac{\pi}{2}
1783: \left[ \left( {}^t\!h_1\, {}^t\!h_2 \right) \left(
1784: \begin{array}{cc}
1785: -B\hat{F} & (i +A)\hat{F}\\
1786: (i-D)\hat{F} & C\hat{F}
1787: \end{array} \right) \left(
1788: \begin{array}{c}
1789: h_1\\
1790: h_2
1791: \end{array} \right) \right]+2\pi \ii \,
1792: {}^t\!\beta_1h_1+2\pi \ii \,{}^t\!\beta_2 h_2 \right\}~.
1793: \end{equation}
1794: As next step we need to perform a Poisson resummation
1795: \begin{equation}\label{Poisson}
1796: \sum_{h_1 \in \mathbb{Z}^d}\ex{-\pi \,{}^t\!h_1\mathcal{B}h_1+
1797: 2\pi\,{}^t\!h_1\mathcal{B}s}=\frac{1}{\sqrt{\mathrm{Det}\mathcal{B}}}
1798: \ex{\pi \,{}^t\!s\mathcal{B}s}\sum_{h_1 \in \mathbb{Z}^d}
1799: \ex{-\pi \,{}^t\!h_1\mathcal{B}^{-1}h_1-2\pi \ii \,{}^t\!h_1s}
1800: \end{equation}
1801: on the first $d$ components of $h$ which we indicate with $h_1$. So in
1802: our case we have
1803: \begin{equation}
1804: \mathcal{B}=\frac{1}{2} \hat{F}\,{}^t\!B \,\,\,\,\,\,
1805: \mathrm{and} \,\,\,\,\,\,
1806: s=\frac{1}{2}\mathcal{B}^{-1}(A+i)\hat{F}h_2+
1807: \ii \mathcal{B}^{-1}\beta_1~.
1808: \end{equation}
1809: The first exponential in the r.h.s. of the Poisson resummation formula
1810: (\ref{Poisson}) yields a quadratic term in the vector $h_2$, which can
1811: be combined with a similar contribution present in the initial
1812: expression~(\ref{Azm3})
1813: \begin{eqnarray}
1814: \nonumber && \frac{\pi}{4} \,{}^t\!h_2 \hat{F} (i+\,{}^t\!A)\,
1815: {}^t\!\mathcal{B}^{-1}(i+A)\hat{F}h_2+
1816: \frac{\pi}{2}\,{}^t\!h_2 C\hat{F} h_2=\\
1817: && = \frac{\pi}{2}\,{}^t\!h_2
1818: \left[ \hat{F}(i+\,{}^t\!A)\hat{F}^{-1}B^{-1}(i+A)\hat{F}+
1819: C\hat{F} \right]h_2~.
1820: \end{eqnarray}
1821: Recalling that $\mathcal{I}'\hat{F}$ is a symmetric matrix, it is not difficult
1822: to see that $A\hat{F}=-\hat{F}\,{}^t\!D$. Some algebraic
1823: manipulations involving these identities simplify the previous
1824: expression into
1825: \begin{equation}
1826: \ii \pi \,{}^t\!h_2 B^{-1}(i+A)\hat{F}h_2.
1827: \end{equation}
1828: Hence the Poisson resummation performed on Eq.~(\ref{Azm3}) gives
1829: \begin{eqnarray}
1830: \nonumber \frac{1}{\sqrt{\mathrm{Det}(2B\hat{F})}} &&
1831: \sum_{h_i \in \mathbb{Z}^d}
1832: \exp \left\{ \ii \pi \left[ \,{}^t\!h_2B^{-1}(i+A)
1833: \hat{F}h_2+2\ii \,{}^t\!\beta_1\,{}^t\!B^{-1}
1834: \hat{F}^{-1}\beta_1+ \right. \right. \\
1835: \nonumber && +\,{}^t\!h_2\hat{F}(i+\,{}^t\!A)\,{}^t\!
1836: \hat{F}^{-1}B^{-1}\beta_1+\,{}^t\!\beta_1
1837: \hat{F}^{-1}B^{-1}(i+A)\hat{F}h_2\\
1838: \nonumber && +2\,{}^t\!\beta_2h_2+2\ii \,{}^t\!h_1
1839: \,{}^t\!B^{-1}\hat{F}^{-1}h_1-2\,{}^t\!h_2B^{-1}(i+A)h_1-\\
1840: && \left. \left. 4\ii \,{}^t\!h_1\,{}^t\!B^{-1}
1841: \hat{F}^{-1}\beta_1 \right] \right\}~.
1842: \end{eqnarray}
1843: In the following manipulations we will focus on the expression in the
1844: square brackets only. It is useful to observe that the by redefining
1845: \begin{equation}
1846: \gamma_1=\hat{F}^{-1}\beta_1 \,\,\,\,\,\,
1847: \mathrm{and} \,\,\,\,\,\, k=\hat{F}^{-1}h_1
1848: \end{equation}
1849: and making use again of the identities mentioned earlier, involving
1850: the entries of the complex structure and of $\hat{F}$, one can rewrite the
1851: content of the square brackets above as
1852: \begin{eqnarray}
1853: \nonumber && {}^t\!h_2B^{-1}(i+A)\hat{F}h_2+2\ii \,
1854: {}^t\!\gamma_1B^{-1}\hat{F}\gamma_1+\,{}^t\!
1855: h_2B^{-1}(i+A)\hat{F}\gamma_1+\\
1856: \nonumber && +\,{}^t\!\gamma_1B^{-1}(i+A)
1857: \hat{F}h_2+2\,{}^t\!\beta_2h_2+2\ii \,{}^t\!kB^{-1}\hat{F}k-\\
1858: && -2\,{}^t\!h_2B^{-1}(i+A)\hat{F}k-4\ii \,{}^t\!kB^{-1}\hat{F}\gamma_1 ~.
1859: \end{eqnarray}
1860: As $k$ in general is no longer a column of integers, it is convenient
1861: to write it
1862: distinguishing its integer part from the remainder
1863: \begin{equation}
1864: k= r+\hat{F}^{-1}l ~,
1865: \end{equation}
1866: with $r \in \mathbb{Z}^d$ and $l_{\alpha} \in
1867: [1,\hat{F}_{\alpha\alpha}]$. Thus the expression above reads
1868: \begin{eqnarray}
1869: \nonumber && 2 \,{}^t\!\left( r+\hat{F}^{-1}l-\gamma_1 \right)
1870: \ii B^{-1}\hat{F} \left( r+\hat{F}^{-1}l-\gamma_1 \right)+\,{}^t\!h_2
1871: B^{-1}(i+A)\hat{F}h_2+\\
1872: && -2\,{}^t\!h_2B^{-1}(i+A)\hat{F}
1873: \left( r+\hat{F}^{-1}l-\gamma_1 \right) +2\,{}^t\!\beta_2h_2~.
1874: \end{eqnarray}
1875: Finally, defining $s=r-h_2 \in \mathbb{Z}^d$, one has
1876: \begin{eqnarray}
1877: \nonumber {}^t\!\left( r+\hat{F}^{-1}l-\gamma_1 \right)
1878: B^{-1}(i-A)\hat{F} \left( r+\hat{F}^{-1}l-\gamma_1 \right) +
1879: 2\,{}^t\! \left( r+\hat{F}^{-1}l-\gamma_1 \right) \beta_2+\\
1880: {}^t\!\left( s+\hat{F}^{-1}l-\gamma_1 \right)B^{-1}(i+A)\hat{F}
1881: \left( s+\hat{F}^{-1}l-\gamma_1 \right)-2\,{}^t\!
1882: \left( s+\hat{F}^{-1}l-\gamma_1 \right) \beta_2~.
1883: \end{eqnarray}
1884: Thus the factorized amplitude reads
1885: \begin{eqnarray}\label{FactAmpl}
1886: \nonumber \mathcal{A}=
1887: \sum_{l_{\alpha}=1}^{\hat{F}_{\alpha\alpha}}
1888: \frac{1}{\sqrt{\mathrm{Det}(2B\hat{F})}} && \vartheta \left[
1889: \begin{array}{c}
1890: \hat{F}^{-1}(l-\beta_1)\\
1891: \beta_2
1892: \end{array} \right] \left( 0 \big| B^{-1}(i-A)\hat{F} \right) \times\\
1893: && \vartheta \left[
1894: \begin{array}{c}
1895: \hat{F}^{-1}(l-\beta_1)\\
1896: -\beta_2
1897: \end{array} \right] \left( 0 \big| B^{-1}(i+A)\hat{F} \right)
1898: \end{eqnarray}
1899: written in terms of $d$-dimensional Theta-functions
1900: \begin{equation}
1901: \vartheta \left[
1902: \begin{array}{c}
1903: a\\ b
1904: \end{array} \right] \left(\nu | \tau \right)=
1905: \sum_{h \in \mathbb{Z}^d} \mathrm{exp}\left[ \pi \ii \,
1906: {}^t\!(h+a)\tau(h+a)+2\pi \ii\,{}^t\!(\nu+b)(h+a) \right]~.
1907: \end{equation}
1908: Notice that the function in the second line of the
1909: Eq.~(\ref{FactAmpl}) is indeed the complex conjugate of the one in the
1910: first line, as $\hat{F}$, $A$ and $B$ are real $d \times d$ matrices
1911: since $\mathcal{I}$ in Eq.~(\ref{RealComplexStructure}) is
1912: real. This function is to be interpreted as the classical contribution
1913: to the Yukawa couplings for three twisted states arising in a generic
1914: $T^{2d}$ compactification of string theory with magnetised space
1915: filling D-Branes. The sum in front of the couplings reveals their
1916: multiplicity, given by $\mathrm{Det}\hat{F}=
1917: \prod_{\alpha=1}^d\hat{F}_{\a\a}$.
1918: Finally for the sake of
1919: completeness let us write the expression for the correlator between
1920: three twist fields (fixing one $l$, i.e. one particular coupling)
1921: including also its quantum contribution in the $\mathcal{N}=1$
1922: supersymmetric configuration
1923: $\epsilon^\a_1+\epsilon^\a_2+\epsilon^\a_3=1$\footnote{We use the
1924: notations of Sect.4 of \cite{Russo:2007tc}.}:
1925: \begin{eqnarray} \nonumber
1926: \langle \sigma_{\epsilon_1}\sigma_{\epsilon_2}
1927: \sigma_{\epsilon_3}\rangle & = & \prod_{i=1}^3
1928: \prod_{\a=1}^d \left[ \frac{\Gamma(1-\epsilon^\a_i)}{\Gamma(\epsilon^\a_i)}
1929: \right]^{\frac{1}{4}}
1930: \left( \mathrm{Det}(2B) \right)^{-\frac{1}{4}}\\
1931: & \times & \vartheta\left[
1932: \begin{array}{c}
1933: \hat{F}^{-1}(l-\beta_1)\\
1934: \beta_2
1935: \end{array} \right] \left( 0 \big| B^{-1}(i-A)\hat{F} \right)~.
1936: \end{eqnarray}
1937: We can check that this result is in agreement with the literature
1938: considering in particular the multiplicity of the Yukawa couplings in
1939: the case of parallel fluxes, i.e. when all of the boundaries are
1940: magnetized D-Branes with magnetic fields of the type (\ref{Fblock})
1941: put in the form (\ref{Fblock1}). Notice that this setup can always be
1942: T-dualized into a configuration of intersecting D-Branes on a
1943: $2d$-dimensional torus that is not geometrically a direct product of
1944: $d$ $T^2$'s. However in the counting of the non vanishing 3-point
1945: correlators between twisted states the metric of the torus is not
1946: involved, and thus we expect this multiplicity is equal to the one
1947: already calculated~\cite{Cremades:2003qj} in fully factorizable models
1948: of D-Branes at angles on $(T^2)^d$. Indeed following the steps of the
1949: previous subsection it is not difficult to see that $H=w_1w_2/\delta$,
1950: $\delta$ being a diagonal matrix whose eigenvalues are the G.C.D.'s of
1951: the entries of $w_1$ and $w_2$ as in Eq.~(\ref{wma}). Then by
1952: generalizing also the definition (\ref{Iij}) into a diagonal matrix,
1953: with the same structure and the intersection numbers as entries, one
1954: has $\hat{F}=I_{21}w_1w_2/\delta^2$. Upon the straightforward
1955: generalization of the T-Duality in Eq.~(\ref{zeroDuality}) $\hat{F}
1956: \rightarrow I_{21}I_{20}I_{01}/\delta^2$ where $\delta$ now contains
1957: the G.C.D.'s of the entries of the three intersection numbers and
1958: \begin{equation}
1959: \mathrm{Det}\hat{F}=
1960: \prod_{\alpha=1}^d\frac{(p_{2\alpha}W_{0\alpha}-p_{0\alpha}W_{2\alpha})
1961: (p_{0\alpha}W_{1\alpha}-p_{1\alpha}W_{0\alpha})
1962: (p_{2\alpha}W_{1\alpha}-p_{2\alpha}W_{2\alpha})}{\delta^2_{\alpha\alpha}}
1963: \end{equation}
1964: This number agrees with the product of the multiplicity of Yukawa
1965: couplings in each of the $T^2$'s inside the $T^{2d}$ defined by the
1966: form of the magnetic fields (see \cite{Cremades:2003qj}).
1967:
1968: \vspace{1cm}
1969:
1970: \noindent {\large {\bf Acknowledgements}}
1971:
1972: \vspace{3mm}
1973:
1974: \noindent
1975: We wish to thank Giulio Bonelli, Paolo Di Vecchia, Francisco Morales,
1976: Igor Pesando, Sanjaye Ramgoolam, Alessandro Tanzini and Steve Thomas
1977: for useful discussions and comments. This work is partially supported
1978: by the European Community's Human Potential under contract
1979: MRTN-CT-2004-005104 and MRTN-CT-2004-512194, and by the Italian MIUR
1980: under contract PRIN 2005023102. ~R. R. and S. Sc. thank the
1981: Galileo Galilei Institute for Theoretical Physics, and S. Sc. thank the
1982: Queen Mary University, for hospitality during the completion of this work.
1983:
1984: \appendix
1985:
1986: \sect{Generic and block-diagonal integer matrices}\label{AppendixFTransf}
1987:
1988: A generic integer $2d \times 2d$ antisymmetric matrix $M$ can always
1989: be put in a block-diagonal form by means of a transformation of the
1990: type $M \rightarrow {}^tOMO$, $O$ being a unimodular integer matrix.
1991:
1992: In order to show that this is indeed the case one can observe that any
1993: antisymmetric matrix:
1994: \begin{equation}\label{M}
1995: M=\left(
1996: \begin{array}{cc}
1997: A & B\\
1998: -{}^tB & C
1999: \end{array} \right)
2000: \end{equation}
2001: where
2002: $A$ and $C$ are $2k \times 2k$ and $2(d-k) \times 2(d-k)$
2003: antisymmetric matrices and $B$ is a rectangular $2k \times 2(d-k)$
2004: matrix, can be block diagonalized by
2005: \begin{equation}\label{O}
2006: O=\left(
2007: \begin{array}{cc}
2008: 1_{2k} & -A^{-1}B\\
2009: 0 & 1_{2(d-k)}
2010: \end{array} \right),
2011: \end{equation}
2012: getting:
2013: \begin{equation}\label{OMO}
2014: {}^tOMO=\left(
2015: \begin{array}{cc}
2016: A & 0\\
2017: 0 & {}^tBA^{-1}B+C
2018: \end{array} \right)
2019: \end{equation}
2020: Notice that each block of ${}^tOMO$ is also an antisymmetric matrix
2021: and that the determinant of the matrix $O$ is one. Since the form of
2022: ${}^tOMO$ in Eq.~(\ref{OMO}) is independent of the choice of $k$, one
2023: can use an iterative procedure to obtain the final block diagonal
2024: form
2025: always choosing $k=1$.
2026: Using this procedure $d-1$ times one finds
2027: \begin{equation}\label{M''}
2028: {}^tO M O
2029: =\left(
2030: \begin{array}{cccccc}
2031: a_1 & 0 & 0 & 0 & \cdots & \cdots\\
2032: 0 & a_2 & 0 & 0 & \cdots &\cdots\\
2033: \vdots & \vdots & \vdots & \vdots & \vdots & \vdots\\
2034: 0 & 0 & 0 & 0 & \cdots & a_d \\
2035: \end{array} \right) \otimes
2036: \left(
2037: \begin{array}{cc}
2038: 0 & 1 \\
2039: -1 & 0
2040: \end{array} \right),
2041: \end{equation}
2042: where: $O=O_1 O_2...O_{d-1}$.
2043: By a suitable permutation of the rows and of the columns of the matrix above
2044: one can rewrite the transformed matrix as
2045: \begin{equation}\label{M'}
2046: M'= \left(
2047: \begin{array}{cc}
2048: 0 & \tilde{M}\\
2049: -\tilde{M} & 0
2050: \end{array} \right),
2051: \end{equation}
2052: where $\tilde{M}=\mathrm{diag}\{a_1,a_2, \dots \,a_d \}$.
2053:
2054: If the matrix $M$ has integer elements, the transformed matrix will be
2055: integer only if $\mathrm{Det}A=1$. We will now show that it is always
2056: possible to reduce to this case at any step of the iterative
2057: procedure. Obviously, if at least one element of the matrix $M$ is
2058: $1$, it is enough to suitably relabel the rows and the columns.
2059: Otherwise, if $\mathrm{Det}A \neq 1$ we can distinguish various cases.
2060: The simplest one is when in a row two elements are coprime. By
2061: relabeling of the rows and the columns it is possible to put these
2062: elements ($a$ and $b$) in the first row as follows:
2063: \begin{equation}\label{M2}
2064: M
2065: =\left(
2066: \begin{array}{cccc}
2067: 0 & a & b & \cdots\\
2068: -a & 0 & c & \cdots\\
2069: -b & -c & 0 & \cdots\\
2070: \vdots & \vdots & \vdots & \ddots
2071: \end{array} \right).
2072: \end{equation}
2073: Then it is easy
2074: to see that the unimodular integer matrix
2075: \begin{equation}\label{O1}
2076: Q=\left(
2077: \begin{array}{ccccc}
2078: 1 & 0 & 0 & 0 & \cdots\\
2079: 0 & x & -b & 0 & \cdots\\
2080: 0 & y & a & 0 & \cdots\\
2081: 0 & 0 & 0 & 1 & \cdots\\
2082: \vdots & \vdots & \vdots & \vdots & \ddots
2083: \end{array} \right),
2084: \end{equation}
2085: with $x$, $y$ solution of the Diophantine equation $ax+by=1$,
2086: transforms $M$ into:
2087: \begin{equation}
2088: M \rightarrow {}^t Q M Q = \left(
2089: \begin{array}{cccc}
2090: 0 & 1 & 0 & \cdots\\
2091: -1 & 0 & c & \cdots\\
2092: 0 & -c & 0 & \cdots\\
2093: \vdots & \vdots & \vdots & \ddots
2094: \end{array} \right),
2095: \end{equation}
2096: which has $\mathrm{Det}A=1$.
2097:
2098: If, instead, there is no row with two coprime elements, but in
2099: Eq.~\eq{M2} $a\neq \pm b$ and their greatest common denominator $d$ is
2100: different from $1$, one can repeat the previous step with the matrix:
2101: \begin{equation}\label{O2}
2102: Q'=\left(
2103: \begin{array}{ccccc}
2104: 1 & 0 & 0 & 0 & \cdots\\
2105: 0 & x & -\frac{b}{d} & 0 & \cdots\\
2106: 0 & y & \frac{a}{d} & 0 & \cdots\\
2107: 0 & 0 & 0 & 1 & \cdots\\
2108: \vdots & \vdots & \vdots & \vdots & \ddots
2109: \end{array} \right),
2110: \end{equation}
2111: where now $x$ and $y$ solve the Diophantine equation $ax+by=d$.
2112: This yields
2113: \begin{equation}
2114: M \rightarrow \,{}^t\!Q'MQ'=\left(
2115: \begin{array}{cccc}
2116: 0 & d & 0 & \cdots\\
2117: -d & 0 & c & \cdots\\
2118: 0 & -c & 0 & \cdots\\
2119: \vdots & \vdots & \vdots & \ddots
2120: \end{array} \right).
2121: \end{equation}
2122: One has to apply this procedure (which does not change the other
2123: elements of the first row, from the 4th column on) till the matrix is
2124: reduced to one of the following cases: either in the first row of the
2125: transformed $M$ there are two coprime non-vanishing entries, then one
2126: can use the matrix $Q$ to obtain $a_1=1$; or all of the non-zero
2127: elements there coincide with $\pm d$. This is the case if for instance
2128: in the original matrix $M$ the first row contained elements which were
2129: all multiples of $d$. If there is any other row in the transformed
2130: matrix with two different non-zero elements $a \neq \pm b$ for which
2131: $d$ is not a divisor, then, by exchanging rows and columns among
2132: themselves, it is possible to bring this as the first row and reapply
2133: the transformations encoded in $Q$ or $Q'$. Otherwise one can have two
2134: possible forms for the transformed matrix. One possibility is that
2135: all of the elements of $M$ are integer multiples of $d$. In this case
2136: the common divisor $d$ can be factored out to reduce to the case with
2137: $\mathrm{Det} A=1$.
2138:
2139: The other possibility is that the matrix has diagonal blocks, in
2140: which all the elements are multiple of different integers $d_i$.
2141: If all the non trivial blocks are $2\times 2$, we have got our aim; otherwise
2142: we can factor out $d_i$ from the block of larger
2143: dimensions; for each of them we are again reduced to the case with
2144: $\mathrm{Det} A=1$. Repeating the procedure, if needed, we finally
2145: end to the matrix:
2146: \begin{equation}
2147: M_1=\left(
2148: \begin{array}{ccccc}
2149: 0 & d_1 & 0 & 0 & \cdots\\
2150: -d_1 & 0 & 0 & 0 & \cdots\\
2151: 0 & 0 & 0 & d_2 & \cdots\\
2152: 0 & 0 & -d_2 & 0 & \cdots\\
2153: \vdots & \vdots & \vdots & \vdots & \ddots
2154: \end{array} \right)
2155: \end{equation}
2156: with all integer elements.
2157:
2158: Although this already is the final form we are after, for sake of
2159: completeness we recall that the normal form of the initial
2160: matrix~(\ref{M''}), as discussed in \cite{Griffiths}, has the further
2161: property that $a_{\a+1}/a_\a \in \mathbb{N}$, $\forall\a$. In order to
2162: achieve this (even if it is not strictly necessary for the
2163: computations considered here) one can use the following transformation
2164: \begin{equation}
2165: Q''=\left(
2166: \begin{array}{ccccc}
2167: 1 & 0 & 0 & 0 & \cdots\\
2168: 0 & 1 & 0 & 0 & \cdots\\
2169: 1 & 0 & 0 & 1 & \cdots\\
2170: 0 & 0 & 1 & 0 & \cdots\\
2171: \vdots & \vdots & \vdots & \vdots & \ddots
2172: \end{array} \right)
2173: \end{equation}
2174: that mixes the $d_i$'s and gives back a form that can be reduced by
2175: means of either $Q$ or $Q'$. Following this procedure, it is possible
2176: to convince oneself that the final matrix~(\ref{M''}) entries satisfy
2177: the property mentioned above, since one actually ends the repeated
2178: application of $Q$, $Q'$, and $Q''$ only if in the first $2 \times 2$
2179: block there is a one, or if the matrix is proportional to an integer
2180: as a whole.
2181:
2182: Coming back to our main problem, we can now apply the procedure
2183: outlined in this Appendix to rewrite the quantized magnetic
2184: field~(\ref{Fquantization}) on a generic magnetized D-Brane in the
2185: form~(\ref{Fblock}). It is first convenient to define an integer
2186: matrix $P$ associated to the magnetic field~(\ref{Fquantization})
2187: \begin{equation}
2188: \label{p}
2189: P=F \times \mathrm{m.c.m}\left\{ w_M w_N,\, M\neq N,~~\,\forall
2190: M,N=1, \dots,2d
2191: \right\}
2192: = \omega F.
2193: \end{equation}
2194: where $\omega$ is the minimum common multiple of all the pairs of
2195: wrappings that appear in the denominators of
2196: Eq.~(\ref{Fquantization}). The integer matrix $P $ can now be
2197: transformed into a block-diagonal form as in Eq.~(\ref{M''}) by means
2198: of an integer unimodular matrix $O$ that preserves the lattice of the
2199: torus as
2200: \begin{equation}
2201: \label{cell}
2202: P \rightarrow \,{}^t\!OPO=
2203: \omega\,{}^t\!OFO=\omega F_{\mathrm{block}}.
2204: \end{equation}
2205: Hence $F_{\mathrm{block}}$ will have the same form as Eq.~(\ref{M''}) with
2206: rational entries whose numerators and denominators could still have
2207: factors in common. By expurgating these factors one exactly recovers
2208: the form in Eq.~(\ref{Fblock}).
2209: First of all we will show that the phase factor in the boundary state~\eq{WLBF}
2210: is not affected by the change of the fundamental cell in the lattice
2211: torus performed in~\eq{cell}. It reads:
2212: \begin{equation}\label{ph}
2213: \mathrm{Ph}=\mathrm{exp}\left[\ii \pi\sum_{M<N}
2214: \hat{m}^MF_{MN}\hat{m}^N \right]
2215: =\mathrm{exp}\left[\frac{ \ii \pi}{\omega}\sum_{M<N}
2216: \hat{m}^MP_{MN}\hat{m}^N \right]
2217: \end{equation}
2218: where $P$ is the integer matrix defined in~\eq{p}, that we write in
2219: the form of Eq.~\eq{M}.
2220: To write it as a block diagonal matrix, we use the techniques just discussed; focusing at first on the simplest case with
2221: \begin{equation}\label{A}
2222: A =\left(
2223: \begin{array}{cc}
2224: 0 & 1\\
2225: -1 & 0
2226: \end{array} \right),
2227: \end{equation}
2228: let us consider how the phase $\mathrm{Ph}$ of~\eq{ph} transforms under the substitution:
2229: $\hat{m}=O \hat{m}'$, with
2230: $O$ as in Eq.~(\ref{O}).
2231: One gets:
2232: \begin{eqnarray}\label{ph'}
2233: \nonumber \mathrm{Ph}=\mathrm{exp} \left[ \frac{\ii \pi}{\omega}\left(
2234: \hat{m}'_1 A_{12}\hat{m}'_2+
2235: \sum_{j=3}^{2d}B_{2j}\hat{m}'_jA_{12}\hat{m}'_2
2236: -\hat{m}'_1A_{12}\sum_{j=3}^{2d}B_{1j}\hat{m}'_j
2237: \right. \right.\\
2238: \nonumber -\sum_{j,k=3}^{2d}B_{2j}\hat{m}'_jA_{12}B_{1k}\hat{m}'_k+
2239: \sum_{j=3}^{2d}\hat{m}'_1B_{1j}\hat{m}'_j+
2240: \sum_{j,k=3}^{2d}B_{2j}\hat{m}'_jB_{1k}\hat{m}'_k\\
2241: \left. \left. +\sum_{j=3}^{2d}\hat{m}'_2B_{2j}\hat{m}'_j-
2242: \sum_{j,k=3}^{2d}B_{1j}\hat{m}'_jB_{2k}\hat{m}'_k +
2243: \sum_{k>j=3}^{2d}\hat{m}'_jC_{jk}\hat{m}'_k \right) \right]
2244: \end{eqnarray}
2245: By remembering that all the winding numbers $\hat{m}^N$ must be
2246: integer multiples of the corresponding wrapping numbers $w_N$, one can
2247: check that, in spite of the denominator $\omega$, all the terms in the
2248: exponent are integer multiples of $\ii \pi$. In fact combining the
2249: form of the matrices~(\ref{O}) and~(\ref{p}), one finds the following
2250: expressions for $\hat{m}=O\hat{m}'$
2251: \begin{eqnarray}
2252: \nonumber \hat{m}_1 & = & \hat{m}'_1+\frac{\omega}{w_2}
2253: \sum_{i=3}^{2d}p_{2i}\frac{\hat{m}'_i}{w_i}\\
2254: \nonumber \hat{m}_2 & = & \hat{m}'_2-\frac{\omega}{w_1}
2255: \sum_{i=3}^{2d}p_{1i}\hat{m}'_i\\
2256: \hat{m}'_i & = & \hat{m}_i
2257: \end{eqnarray}
2258: In this case, in order to have the matrix $A$ in the form~(\ref{A}),
2259: it is necessary
2260: that $p_{12}=1$ and $\omega=w_1w_2$, hence, since from the last line
2261: of the previous equation $\hat{m}'_i/w_i$ must be integer, it is also
2262: true, in the first and second line, that the transformed winding
2263: numbers $\hat{m}^{'N}$ are integer multiples of the wrapping numbers
2264: $w_N,~ \forall N=1,2,...,2d$.
2265: Thus each of the terms in the sum~(\ref{ph'}) is an integer number, as
2266: one can check for instance considering the first term of the second
2267: line in Eq.~(\ref{ph'}); writing $\hat{m}^{'N}=m^{'N}w_N$ with $m^{'N}
2268: \in \mathbb{Z}$ one gets:
2269: \begin{equation}
2270: \frac{1}{\omega}B_{2j}\hat{m}'_jA_{12}B_{1k}\hat{m}'_k=
2271: \frac{1}{\omega} \frac{\omega p_{2j}}{w_2w_j}m'_jw_j \frac{\omega
2272: p_{1k}}{w_1w_k}m'_kw_k=p_{2j}m'_jp_{1k}m'_k \in \mathbb{Z}.
2273: \end{equation}
2274: So we can freely change the sign of each term in Eq.~(\ref{ph'}), obtaining
2275: \begin{eqnarray}\label{phf}
2276: \mathrm{Ph} & = & \mathrm{exp} \left[ \ii \pi \left(
2277: \sum_{M<N} \hat{m}^{'M}({}^tO FO)_{MN}
2278: \hat{m}^{'N}+\frac{1}{\omega}
2279: \sum_{j=3}^{2d}B_{1j}B_{2j} \hat{m}_j^{'2} \right) \right]=\\
2280: & = & \mathrm{exp} \left[ \ii \pi \left( \sum_{M<N} \hat{m}^{'M}({}^tO
2281: FO)_{MN}
2282: \hat{m}^{'N}+
2283: \sum_{j=3}^{2d}p_{1j}p_{2j} \frac{\hat{m}'_j}{w_j} \right) \right],
2284: \end{eqnarray}
2285: where we have used the explicit expression of $B_{1j}$ and $B_{2j}$ in
2286: terms of the Chern numbers $p_{1j}$ and $p_{2j}$ and of the winding
2287: numbers; moreover we have taken into account the fact that $(m'_j)^2$
2288: has the same parity (even/odd) as $m'_j=\hat{m}'_j/w_j$. Thus the
2289: phase factor can be written in terms of the transformed field ${}^tO
2290: FO$ and of the transformed winding numbers $\hat{m}'$'s with the same
2291: functional form as the original one~(\ref{ph}), with a half-integer
2292: shift of the Wilson line when $p_{1j}p_{2j}$ is odd.
2293:
2294: If ${}^tO FO$ is already block diagonal, we have ended our job,
2295: otherwise we have to repeat the procedure. In an analogous fashion, if
2296: the entries of $A$ are not equal to one, one can check that the
2297: transformations related to the matrices in Eq.~(\ref{O1})
2298: and~(\ref{O2}), involved in reducing $A$ to the form considered in the
2299: previous example, also preserve the form of the phase factor up to
2300: half-integer Wilson lines. With similar manipulations it is also
2301: possible to prove, in a basis in which $(F_2-F_1)$ is block-diagonal,
2302: that the phase factors in Eq.~\eq{Azm2} follow from those in
2303: Eq.~(\ref{e2finalg2}). As usual, one has to introduce $h \in
2304: \mathbb{Z}^{2d}$ by using $\hat{m}_1=Hh$; then it is possible to check
2305: that the combination $\sum_{M<N}(Hh)^M (F_2-F_1)_{MN} (Hh)^N$ is
2306: equal, modulus two, to $\sum_{M<N} h^M\left[H(F_2-F_1)H\right]_{MN}
2307: h^N$, apart from terms quadratic in $h^M$ that can be reabsorbed into
2308: a half-integer shift of the Wilson lines.
2309:
2310: Finally we mention that the transformations discussed in this Appendix
2311: do not affect the other contributions to the amplitude, in the
2312: effective field theory limit that we consider for the factorization,
2313: if one suitably redefines the complex structure in
2314: Eq.~(\ref{e2finalg2}). Indeed the combination
2315: ${}^t\hat{m}_1\mathcal{I}F\hat{m}_1$ can be rewritten, by redefining
2316: $\hat{m}_1=O\hat{m}'_1$, as
2317: \begin{equation}
2318: {}^t\!\hat{m}'_1\,{}^t\!O\mathcal{I}FO\hat{m}'_1=
2319: {}^t\hat{m}'_1\,{}^t\!O\mathcal{I}{}^t\!O^{-1}\,{}^t\!OFO\hat{m}'_1
2320: =\,{}^t\!\hat{m}'_1\mathcal{\hat{I}}F_{\mathrm{block}}\hat{m}'_1,
2321: \end{equation}
2322: $\mathcal{\hat{I}}$ still being a good complex structure and
2323: $F_{\mathrm{block}}$
2324: being in the form~(\ref{M''}).
2325:
2326: \begin{thebibliography}{10}
2327:
2328: \bibitem{Lovelace:1971fa}
2329: C.~Lovelace,
2330: \newblock Phys. Lett. {\bf B34}, 500 (1971).
2331: %%CITATION = PHLTA,B34,500;%%
2332:
2333: \bibitem{Polchinski:1995mt}
2334: J.~Polchinski,
2335: \newblock Phys. Rev. Lett. {\bf 75}, 4724 (1995), hep-th/9510017.
2336: %%CITATION = HEP-TH/9510017;%%
2337:
2338: \bibitem{Russo:2007tc}
2339: R.~Russo and S.~Sciuto,
2340: \newblock JHEP {\bf 04}, 030 (2007), hep-th/0701292.
2341: %%CITATION = HEP-TH/0701292;%%
2342:
2343: \bibitem{Dixon:1986qv}
2344: L.~J. Dixon, D.~Friedan, E.~J. Martinec, and S.~H. Shenker,
2345: \newblock Nucl. Phys. {\bf B282}, 13 (1987).
2346: %%CITATION = NUPHA,B282,13;%%
2347:
2348: \bibitem{Abouelsaood:1986gd}
2349: A.~Abouelsaood, J.~Callan, Curtis~G., C.~R. Nappi, and S.~A. Yost,
2350: \newblock Nucl. Phys. {\bf B280}, 599 (1987).
2351: %%CITATION = NUPHA,B280,599;%%
2352:
2353: \bibitem{Berkooz:1996km}
2354: M.~Berkooz, M.~R. Douglas, and R.~G. Leigh,
2355: \newblock Nucl. Phys. {\bf B480}, 265 (1996), hep-th/9606139.
2356: %%CITATION = HEP-TH 9606139;%%
2357:
2358: \bibitem{Marchesano:2007de}
2359: F.~Marchesano,
2360: \newblock Fortsch. Phys. {\bf 55}, 491 (2007), hep-th/0702094.
2361: %%CITATION = HEP-TH/0702094;%%
2362:
2363: \bibitem{Burwick:1990tu}
2364: T.~T. Burwick, R.~K. Kaiser, and H.~F. Muller,
2365: \newblock Nucl. Phys. {\bf B355}, 689 (1991).
2366: %%CITATION = NUPHA,B355,689;%%
2367:
2368: \bibitem{Erler:1992gt}
2369: J.~Erler, D.~Jungnickel, M.~Spalinski, and S.~Stieberger,
2370: \newblock Nucl. Phys. {\bf B397}, 379 (1993), hep-th/9207049.
2371: %%CITATION = HEP-TH 9207049;%%
2372:
2373: \bibitem{Stieberger:1992bj}
2374: S.~Stieberger, D.~Jungnickel, J.~Lauer, and M.~Spalinski,
2375: \newblock Mod. Phys. Lett. {\bf A7}, 3059 (1992), hep-th/9204037.
2376: %%CITATION = HEP-TH 9204037;%%
2377:
2378: \bibitem{Stieberger:1992vb}
2379: S.~Stieberger,
2380: \newblock Phys. Lett. {\bf B300}, 347 (1993), hep-th/9211027.
2381: %%CITATION = HEP-TH 9211027;%%
2382:
2383: \bibitem{Cremades:2003qj}
2384: D.~Cremades, L.~E. Ibanez, and F.~Marchesano,
2385: \newblock JHEP {\bf 07}, 038 (2003), hep-th/0302105.
2386: %%CITATION = HEP-TH 0302105;%%
2387:
2388: \bibitem{Cvetic:2003ch}
2389: M.~Cvetic and I.~Papadimitriou,
2390: \newblock Phys. Rev. {\bf D68}, 046001 (2003), hep-th/0303083.
2391: %%CITATION = HEP-TH 0303083;%%
2392:
2393: \bibitem{Abel:2003vv}
2394: S.~A. Abel and A.~W. Owen,
2395: \newblock Nucl. Phys. {\bf B663}, 197 (2003), hep-th/0303124.
2396: %%CITATION = HEP-TH 0303124;%%
2397:
2398: \bibitem{Abel:2003yx}
2399: S.~A. Abel and A.~W. Owen,
2400: \newblock Nucl. Phys. {\bf B682}, 183 (2004), hep-th/0310257.
2401: %%CITATION = HEP-TH 0310257;%%
2402:
2403: \bibitem{Lust:2004cx}
2404: D.~Lust, P.~Mayr, R.~Richter, and S.~Stieberger,
2405: \newblock Nucl. Phys. {\bf B696}, 205 (2004), hep-th/0404134.
2406: %%CITATION = HEP-TH 0404134;%%
2407:
2408: \bibitem{Blumenhagen:2006xt}
2409: R.~Blumenhagen, M.~Cvetic, and T.~Weigand,
2410: \newblock Nucl. Phys. {\bf B771}, 113 (2007), hep-th/0609191.
2411: %%CITATION = HEP-TH/0609191;%%
2412:
2413: \bibitem{Ibanez:2006da}
2414: L.~E. Ibanez and A.~M. Uranga,
2415: \newblock JHEP {\bf 03}, 052 (2007), hep-th/0609213.
2416: %%CITATION = HEP-TH/0609213;%%
2417:
2418: \bibitem{Bianchi:2005yz}
2419: M.~Bianchi and E.~Trevigne,
2420: \newblock JHEP {\bf 08}, 034 (2005), hep-th/0502147.
2421: %%CITATION = HEP-TH 0502147;%%
2422:
2423: \bibitem{Alvarez-Gaume:1988bg}
2424: L.~Alvarez-Gaume, C.~Gomez, G.~W. Moore, and C.~Vafa,
2425: \newblock Nucl. Phys. {\bf B303}, 455 (1988).
2426: %%CITATION = NUPHA,B303,455;%%
2427:
2428: \bibitem{DiVecchia:1988cy}
2429: P.~Di~Vecchia {\em et~al.},
2430: \newblock Nucl. Phys. {\bf B322}, 317 (1989).
2431: %%CITATION = NUPHA,B322,317;%%
2432:
2433: \bibitem{Polchinski:1998rq}
2434: J.~Polchinski,
2435: \newblock Cambridge, UK: Univ. Pr. (1998) 402 p.
2436:
2437: \bibitem{Giveon:1994fu}
2438: A.~Giveon, M.~Porrati, and E.~Rabinovici,
2439: \newblock Phys. Rept. {\bf 244}, 77 (1994), hep-th/9401139.
2440: %%CITATION = HEP-TH 9401139;%%
2441:
2442: \bibitem{Pesando:2005df}
2443: I.~Pesando,
2444: \newblock (2005), hep-th/0505052.
2445: %%CITATION = HEP-TH 0505052;%%
2446:
2447: \bibitem{DiVecchia:2006gg}
2448: P.~Di~Vecchia, A.~Liccardo, R.~Marotta, F.~Pezzella, and I.~Pesando,
2449: \newblock (2006), hep-th/0601067.
2450: %%CITATION = HEP-TH 0601067;%%
2451:
2452: \bibitem{DiVecchia:1999rh}
2453: P.~Di~Vecchia and A.~Liccardo,
2454: \newblock NATO Adv. Study Inst. Ser. C. Math. Phys. Sci. {\bf 556}, 1 (2000),
2455: hep-th/9912161.
2456: %%CITATION = HEP-TH 9912161;%%
2457:
2458: \bibitem{DiVecchia:1999fx}
2459: P.~Di~Vecchia and A.~Liccardo,
2460: \newblock (1999), hep-th/9912275.
2461: %%CITATION = HEP-TH 9912275;%%
2462:
2463: \bibitem{Callan:1988wz}
2464: J.~Callan, Curtis~G., C.~Lovelace, C.~R. Nappi, and S.~A. Yost,
2465: \newblock Nucl. Phys. {\bf B308}, 221 (1988).
2466: %%CITATION = NUPHA,B308,221;%%
2467:
2468: \bibitem{Igor:2008}
2469: P.~Di~Vecchia, A.~Liccardo, R.~Marotta, I.~Pesando, and F.~Pezzella,
2470: \newblock (2007), arXiv:0709.4149 [hep-th].
2471: %%CITATION = ARXIV:0709.4149;%%
2472:
2473: \bibitem{Guralnik:1997sy}
2474: Z.~Guralnik and S.~Ramgoolam,
2475: \newblock Nucl. Phys. {\bf B499}, 241 (1997), hep-th/9702099.
2476: %%CITATION = HEP-TH/9702099;%%
2477:
2478: \bibitem{Guralnik:1997th}
2479: Z.~Guralnik and S.~Ramgoolam,
2480: \newblock Nucl. Phys. {\bf B521}, 129 (1998), hep-th/9708089.
2481: %%CITATION = HEP-TH/9708089;%%
2482:
2483: \bibitem{tHooft:1981sz}
2484: G.~'t~Hooft,
2485: \newblock Commun. Math. Phys. {\bf 81}, 267 (1981).
2486: %%CITATION = CMPHA,81,267;%%
2487:
2488: \bibitem{Polchinski:1996fm}
2489: J.~Polchinski, S.~Chaudhuri, and C.~V. Johnson,
2490: \newblock (1996), hep-th/9602052.
2491: %%CITATION = HEP-TH/9602052;%%
2492:
2493: \bibitem{Hashimoto:1996pd}
2494: A.~Hashimoto,
2495: \newblock Int. J. Mod. Phys. {\bf A13}, 903 (1998), hep-th/9610250.
2496: %%CITATION = HEP-TH/9610250;%%
2497:
2498: \bibitem{Griffiths}
2499: P.~Griffiths and J.~Harris,
2500: \newblock {\em Principles of Algebraic Geometry} (John Wiley \& Sons Inc,
2501: 1994).
2502:
2503: \bibitem{Myers:1999ps}
2504: R.~C. Myers,
2505: \newblock JHEP {\bf 12}, 022 (1999), hep-th/9910053.
2506: %%CITATION = HEP-TH/9910053;%%
2507:
2508: \bibitem{Bachas:1992bh}
2509: C.~Bachas and M.~Porrati,
2510: \newblock Phys. Lett. {\bf B296}, 77 (1992), hep-th/9209032.
2511: %%CITATION = HEP-TH 9209032;%%
2512:
2513: \bibitem{Antoniadis:2005sd}
2514: I.~Antoniadis, K.~S. Narain, and T.~R. Taylor,
2515: \newblock Nucl. Phys. {\bf B729}, 235 (2005), hep-th/0507244.
2516: %%CITATION = HEP-TH 0507244;%%
2517:
2518: \bibitem{Bianchi:2005sa}
2519: M.~Bianchi and E.~Trevigne,
2520: \newblock JHEP {\bf 01}, 092 (2006), hep-th/0506080.
2521: %%CITATION = HEP-TH/0506080;%%
2522:
2523: \end{thebibliography}
2524:
2525:
2526: \end{document}
2527: