1: %version December, 2004
2: %
3: \documentclass[12pt]{book}
4: \usepackage{amsmath,graphicx}
5: \usepackage{multirow}
6: \usepackage{amsfonts}
7: \usepackage{amssymb}
8: \usepackage{fancyhdr}
9: %\usepackage{cite}
10: %\usepackage{showkeys}
11: \oddsidemargin 15pt
12: \evensidemargin 0pt
13: \textwidth 6in
14: \textheight 9in
15:
16: \pagestyle{fancy}
17: \renewcommand{\chaptermark}[1]{\markboth{#1}{}}
18: \renewcommand{\sectionmark}[1]{\markright{\thesection\ #1}}
19: \fancyhf{} % delete current setting for header and footer
20: \fancyhead[LE,RO]{\bfseries\thepage}
21: \fancyhead[LO]{\bfseries\rightmark}
22: \fancyhead[RE]{\bfseries\leftmark}
23: \renewcommand{\headrulewidth}{0.5pt}
24: \renewcommand{\footrulewidth}{0pt}
25: \addtolength{\headheight}{4.5pt} % make space for the rule
26: \fancypagestyle{plain}{\fancyhead{} % get rid of headers on plain pages
27: \renewcommand{\headrulewidth}{0pt} % and the line
28: }
29: \parskip 3pt plus 1pt
30:
31: \def\hybrid{\topmargin -20pt \oddsidemargin 0pt
32: \headheight 0pt \headsep 15pt
33: % \textwidth 6.5in % US paper
34: % \textheight 9in % US paper
35: \textwidth 6.25in % A4 paper
36: \textheight 8.6 in % A4 paper
37: \marginparwidth .875in
38: \parskip 5pt plus 1pt \jot = 1.5ex}
39: % The default is set to be hybrid
40: \hybrid
41:
42: \numberwithin{equation}{chapter}
43: \numberwithin{table}{section}\setlength{\multlinegap}{25pt}
44:
45: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
46: % abbreviate environments
47: \newcommand{\beq}{\begin{equation}}
48: \newcommand{\eeq}{\end{equation}}
49: %\newcommand{\bi}{\begin{itemize}}
50: %\newcommand{\ei}{\end{itemize}}
51: \newcommand{\bea}{\begin{eqnarray}}
52: \newcommand{\eea}{\end{eqnarray}}
53: \newcommand{\ba}{\begin{array}}
54: \newcommand{\ea}{\end{array}}
55: \newcommand{\bt}{\begin{tabular}}
56: \newcommand{\et}{\end{tabular}}
57: \newcommand{\bc}{\begin{center}}
58: \newcommand{\ec}{\end{center}}
59: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
60: % abbreviate Greek
61: \newcommand{\ax}{\alpha}
62: \newcommand{\bx}{\beta}
63: \newcommand{\cx}{\gamma}
64: \newcommand{\dx}{\delta}
65: \newcommand{\ox}{\omega}
66: \newcommand{\lx}{\lambda}
67: \newcommand{\ab}{\bar\alpha}
68: \newcommand{\bb}{\bar\beta}
69: \newcommand{\cb}{\bar\gamma}
70: \newcommand{\db}{\bar\delta}
71: \newcommand{\Sx}{\Sigma}
72: \newcommand{\Lx}{\Lambda}
73: \newcommand{\Ox}{\Omega}
74: \newcommand{\Dx}{\Delta}
75: \newcommand{\Gx}{\Gamma}
76: \newcommand{\Oxb}{\bar{\Omega}}
77: %
78: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
79: % Cal
80: \newcommand{\cC}{\mathcal{C}}
81: \newcommand{\cD}{\mathcal{D}}
82: \newcommand{\cL}{\mathcal{L}}
83: \newcommand{\cS}{\mathcal{S}}
84: \newcommand{\cK}{\mathcal{K}}
85: \newcommand{\cN}{\mathcal{N}}
86: \newcommand{\cW}{\mathcal{W}}
87: \newcommand{\cG}{\mathcal{G}}
88: \newcommand{\cA}{\mathcal{A}}
89: \newcommand{\cH}{\mathcal{H}}
90: \newcommand{\cB}{\mathcal{B}}
91: \newcommand{\cF}{\mathcal{F}}
92: \newcommand{\cI}{\mathcal{I}}
93: \newcommand{\cJ}{\mathcal{J}}
94: \newcommand{\cR}{\mathcal{R}}
95: \newcommand{\Ac}{\mathcal{A}}
96: \newcommand{\cV}{\mathcal{V}}
97: \newcommand{\Bc}{\mathcal{B}}
98: \newcommand{\KK}{\mathcal{K}}
99: \newcommand{\MM}{\mathcal{M}}
100: \newcommand{\cM}{\mathcal M}
101: \newcommand{\cQ}{\mathcal Q}
102: \newcommand{\cO}{\mathcal{O}}
103: \newcommand{\OO}{\mathcal{O}}
104: \newcommand{\Vw}{{\mathcal K}_w\vphantom{{\mathcal V}_w}}
105: \newcommand{\Gw}{{\mathcal G}_w\vphantom{{\mathcal G}_w}}
106: %\newcommand{\Vw}{\mathcal K}
107: %\newcommand{\Gw}{\mathcal G}
108:
109: \newcommand{\bfA}{\mathbf{A}}
110: \newcommand{\bfC}{\mathbf{C}}
111: \newcommand{\bfG}{\mathbf{G}}
112:
113: \newcommand{\IF}{\text{Im}\, \mathcal{F}}
114: \newcommand{\IM}{\text{Im}\, \mathcal{M}}
115: \newcommand{\RF}{\text{Re}\, \mathcal{F}}
116: \newcommand{\RM}{\text{Re}\, \mathcal{M}}
117: \newcommand{\I}{\text{Im}}
118: \newcommand{\R}{\text{Re}}
119:
120: \newcommand{\Kcs}{K^{\text{cs}}}
121: \newcommand{\Kks}{K^{\text{ks}}}
122: \newcommand{\pev}{{\varphi^{ev}}}
123: \newcommand{\pevb}{{\bar \varphi^{ev}}}
124: \newcommand{\podd}{{\varphi^{odd}}}
125: \newcommand{\poddb}{{\bar \varphi^{odd}}}
126: \newcommand{\fuh}{{\mathcal{\hat U}}}
127: \newcommand{\fu}{{\mathcal{U}}}
128: \newcommand{\fe}{{\mathcal{E}}}
129: \newcommand{\feh}{{\mathcal{\hat E}}}
130: \newcommand{\fa}{{\mathcal{ A}}}
131: \newcommand{\volume}{\text{{\small} vol}\, }
132:
133: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
134: % indices
135: \newcommand{\bi}{{\bar \imath}}
136: \newcommand{\ib}{{\bar\imath }}
137: \newcommand{\jb}{{\bar \jmath }}
138: \newcommand{\bj}{{\bar\jmath}}
139: \newcommand{\bk}{\bar{k}}
140: \newcommand{\bl}{\bar{l}}
141: \newcommand{\Kh}{{\hat{K}}}
142: \newcommand{\Lh}{{\hat{L}}}
143: \newcommand{\Ah}{{\hat{A}}}
144: \newcommand{\Bh}{{\hat{B}}}
145: \newcommand{\Ch}{{\hat{C}}}
146: \newcommand{\Dh}{{\hat{D}}}
147: \newcommand{\Mh}{{\hat{M}}}
148: \newcommand{\Nh}{{\hat{N}}}
149: \newcommand{\kh}{{\hat{k}}}
150: \newcommand{\ah}{{\hat{a}}}
151: \newcommand{\bh}{{\hat{b}}}
152: \newcommand{\ch}{{\hat{c}}}
153: \newcommand{\lh}{{\hat{l}}}
154: \newcommand{\mh}{{\hat{m}}}
155: \newcommand{\nh}{{\hat{n}}}
156: \newcommand{\Mext}{{\mathbb{M}^{3,1}}}
157: \newcommand{\Mint}{{Y}}
158:
159: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
160: % Dan's macros
161: \DeclareMathOperator{\SU}{\mathit{SU}}
162: \DeclareMathOperator{\SO}{\mathit{SO}}
163: \DeclareMathOperator{\Symp}{\mathit{Sp}}
164: \DeclareMathOperator{\Spin}{\mathit{Spin}}
165: \DeclareMathOperator{\so}{\mathit{so}}
166: \DeclareMathOperator{\su}{\mathit{su}}
167: \DeclareMathOperator{\symp}{\mathit{sp}}
168: \DeclareMathOperator{\spin}{\mathit{spin}}
169: \DeclareMathOperator{\GL}{\mathit{GL}}
170: \DeclareMathOperator{\SL}{\mathit{SL}}
171:
172: \DeclareMathOperator{\vol}{vol}
173:
174: \newcommand{\rep}[1]{\mathbf{#1}}
175:
176: \newcommand{\dd}{d}
177: \newcommand{\ii}{\mathrm{i}}
178:
179: \newcommand{\bbZ}{\mathbb{Z}}
180: \newcommand{\bbR}{\mathbb{R}}
181: \newcommand{\bbC}{\mathbb{C}}
182: \newcommand{\bbP}{\mathbb{P}}
183: \newcommand{\bbF}{\mathbb{F}}
184:
185:
186: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
187: % misc
188: \newcommand{\CY}{Calabi--Yau}
189: \newcommand{\half}{\frac12}
190: \newcommand{\quart}{\frac14}
191: \newcommand{\nn}{\nonumber}
192: %\def\Mnote#1{{\bf[MG: #1]}}
193: \def\Tnote#1{{\bf[TG: #1]}}
194: %\def\Hnote#1{{\bf[HJ: #1]}}
195: \def\Jnote#1{{\bf[JL: #1]}}
196: % \newcommand{\IM}{\textrm{Im} \,}
197: \newcommand{\RE}{\textrm{Re} \,}
198: \newcommand{\?}{{\bf [??]}}
199: %\newcommand{\xx}{{\bf [xx]}}
200: \newcommand{\addref}{{\bf [add ref]}}
201: \newcommand{\cref}{{\bf [check ref]}}
202: \newcommand{\chec}{{\bf [check]}}
203: \newcommand{\park}{{\bf [the following text/formulas are just being parked here]}}
204: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
205: % Jan's macros
206: \newcommand{\M}{M}
207: \newcommand{\N}{\Theta}
208: \newcommand{\Weff}{W^{\rm (eff)}}
209: \newcommand{\Weffb}{\bar W^{\rm (eff)}}
210: \newcommand{\Y}{Y}
211: \newcommand{\G}{\mathcal{I}}
212: \newcommand{\CHI}{\mathcal{I}}
213: \newcommand{\f}{}
214: \newcommand{\Jc}{J_{\rm c}}
215: \newcommand{\Omegac}{\Omega_{\rm c}}
216: \newcommand{\cc}{c}
217: \newcommand{\CC}{C}
218: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
219: % Mariana's macros
220: \newcommand{\hW}{\hat{W}}
221: \newcommand{\hK}{\hat{K}}
222: \newcommand{\hWb}{\hat{\bar W}}
223: \newcommand{\hphi}{{\phi}}
224: \newcommand{\Gt}{G^{(3)}\vphantom{G}}
225: \newcommand{\Gtb}{\bar{G}^{(3)}\vphantom{\bar G}}
226: \newcommand{\ha}{\hat{a}}
227: \newcommand{\hb}{\hat{b}}
228: \newcommand{\hab}{\hat{ a}}
229: \newcommand{\hbb}{\hat{ b}}
230: \newcommand{\cha}{\chi_{\hat{a}}}
231: \newcommand{\chab}{\bar{\chi}_{\hat{a}}}
232: \newcommand{\chb}{\chi_{\hat{b}}}
233: \newcommand{\chbb}{\bar{\chi}_{\hat{b}}}
234: \newcommand{\Dth}{\rm (D3)}
235: \newcommand{\eff}{\rm (eff)}
236: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
237: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
238: % Hans' macros - Beginning
239: \newcommand{\tr}{\mathrm{Tr}\:}
240: \newcommand{\id}{\mathbf{1}}
241: \newcommand{\com}[2]{\big[ {#1},{#2} \big]}
242: \newcommand{\lie}[2]{\left[ {#1},{#2} \right]}
243: \newcommand{\ins}[1]{\mathrm{i}_{#1}}
244: \newcommand{\D}{\mathrm{D}}
245: \newcommand{\Kw}{\mathcal{K}_w} % Warped CY volume
246: \newcommand{\Em}{\varphi} % Embedding map of the world-volume
247: \newcommand{\WV}{\mathcal{W}} % Worldvolume
248: \newcommand{\FD}{F}
249: \newcommand{\FA}{F_\mathrm{A}}%\vphantom{F_\mathrm{A}}} % Field strength - Brane
250: \newcommand{\norm}[1]{\lVert #1\rVert}
251: \newcommand{\Riem}[4]{R_{#1\hphantom{#2}#3#4}^{\hphantom{#1}#2}}
252: % Hans' macros - End
253: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
254:
255:
256: \begin{document}
257:
258:
259: \begin{titlepage}
260: \begin{center}
261:
262: \hfill hep-th/0507153\\
263: \hfill ZMP-HH/05-16\\
264: \vskip 1.5cm
265:
266: {\large \bf The effective action of type II Calabi-Yau orientifolds}
267: \footnote{%
268: This article is based on the Ph.D.~thesis of the author.}\\
269:
270:
271: \vskip 1cm
272:
273:
274:
275: {\bf Thomas W.\ Grimm }\footnote{%
276: From September 1, 2005: Department of Physics, University of Wisconsin, Madison WI 53706, USA } \\
277: \vskip 0.5cm
278:
279: {\em II. Institut f{\"u}r Theoretische Physik\\
280: Universit{\"a}t Hamburg, Luruper Chaussee 149\\
281: D-22761 Hamburg, Germany}\\
282: \vskip 5pt
283: and
284: \vskip 5pt
285: {\em Zentrum f\"ur Mathematische Physik \\
286: Universit\"at Hamburg,
287: Bundesstrasse 55\\
288: D-20146 Hamburg, Germany}\\
289:
290:
291: \vskip 10pt
292:
293: {\tt thomas.grimm@desy.de} \\
294:
295: \end{center}
296:
297: \vskip .5cm
298:
299: \begin{center} {\bf ABSTRACT } \end{center}
300: %\vspace{-2mm}
301:
302: \noindent
303:
304: This article first reviews the calculation of
305: the $N = 1$ effective action for generic type IIA and type IIB Calabi-Yau orientifolds
306: in the presence of background fluxes by using a Kaluza-Klein reduction. The K\"ahler potential,
307: the gauge kinetic functions and the flux-induced superpotential are
308: determined in terms of geometrical data of the Calabi-Yau orientifold and the background fluxes.
309: As a new result, it is shown that the chiral description directly relates to Hitchin's generalized geometry
310: encoded by special odd and even forms on a threefold, whereas a dual formulation with several
311: linear multiplets makes contact to the underlying $N=2$ special geometry.
312: In type IIB setups, the flux-potentials can be expressed
313: in terms of superpotentials, D-terms and, generically, a
314: massive linear multiplet. The type IIA superpotential depends on all geometric
315: moduli of the theory. It is reviewed, how type IIA orientifolds arise as a special
316: limit of M-theory compactified on specific $G_2$ manifolds by matching the effective actions. In a similar
317: spirit type IIB orientifolds are shown to descend from F-theory
318: on a specific class of Calabi-Yau fourfolds.
319: In addition, mirror symmetry for Calabi-Yau orientifolds
320: is briefly discussed and it is shown that the $N = 1$ chiral coordinates linearize the appropriate
321: instanton actions.
322:
323:
324: \vfill
325:
326:
327: \end{titlepage}
328:
329:
330: \vspace*{10cm}
331: \begin{center}
332: {\bf \large Acknowledgments}
333: \end{center}
334: This article is based on my Ph.D.~thesis.
335: First of all I would like to express my deep gratitude to my supervisor Prof.~Jan Louis
336: for his continuous support, expert advises and encouragement.
337: The collaboration with Mariana Gra\~na, Hans Jockers, Frederic Schuller and Mattias Wohlfarth
338: was very enjoyable and fruitful. I esspecially like to thank my office
339: mates Iman Benmachiche, Olaf Hohm, Hans Jockers, Andrei Micu and Anke Knauf for providing a very
340: delighting athmosphere and the numerous discussions about physiscs
341: and beyond. I am also indebted to David Cerde\~no, Vincente Cort\'es, Frederik Denef, Sergei Gukov, Henning Samtleben,
342: Sakura Sch\"afer-Nameki, Shamit Kachru, Boris~K\"ors, Paolo Merlatti, Thorsten Pr\"ustel,
343: Waldemar Schulgin, Silvia Vaul\`a and
344: Martin Weidner for various discussions and correspondence. I am grateful to my lovely girlfriend,
345: for supporting me through the last years.
346:
347: \vspace{.2cm}
348: This work is supported by the DFG -- The German Science Foundation,
349: the DAAD~--~the German Academic Exchange Service, and the European RTN Program MRTN-CT-2004-503369.
350:
351:
352:
353: \thispagestyle{plain}
354:
355:
356: \tableofcontents
357:
358: \renewcommand{\thetable}{\arabic{chapter}.\arabic{table}}
359:
360: \chapter{Introduction}
361:
362: The Standard Model of particle physics extended by massive neutrinos
363: has been tested to a very high precision and is believed
364: to correctly describe the known elementary particles and their interactions.
365: Experimentally, the only missing ingredient is the scalar Higgs particle, which gives
366: masses to the leptons and quarks, once it acquires a vacuum expectation
367: value. The Standard Model provides a realistic model of a renormalizable
368: gauge theory. Despite its impressive success there are also various theoretical
369: drawbacks, such as the large number of free parameters, the hierarchy and naturalness problem as
370: well as the missing unification with gravity. These indicate that it cannot be viewed
371: as a fundamental theory, but rather should arise as an effective description.
372:
373:
374: A natural extension of the Standard Model is provided by supersymmetry, which serves as a
375: fundamental symmetry between bosons and fermions. Supersymmetry predicts a superpartner
376: for all known particles and thus basically doubles the particle content of the theory.
377: However, none of the superpartners was ever detected in an accelerator experiment, which
378: implies that supersymmetry is appearing in its broken phase. The supersymmetric Standard
379: Model solves some of the problems of the Standard Model \cite{revSusy}.
380: Even in its (softly) broken phase it forbids large quantum corrections to scalar masses.
381: This allows the Higgs mass to remain to be of order the weak scale also in a theory
382: with a higher mass scale. Furthermore, assuming the supersymmetric Standard Model to be valid
383: up to very high scales, the renormalization group flow predicts a unification of all
384: three gauge-couplings. This supports the idea of an underlying theory
385: relevant beyond the Standard Model scales. However, it remains to unify
386: these extensions with gravity.
387:
388: On the other hand, we know that General Relativity
389: links the geometry of spacetime with the distribution of the matter densities. Einsteins
390: theory is very different in nature. It is a classical theory which is
391: hard to quantize due to its ultra-violet divergences (see however \cite{LQG}).
392: This fact constraints its range of validity to
393: phenomena, where quantum effects are of negligible importance. However, there is no
394: experimental evidence which contradicts large scale predictions based on General
395: Relativity.
396:
397: Facing these facts General Relativity and the Standard Model seems to be incompatible,
398: in the sense that neither of them allows to naturally adapt the other. This becomes
399: important in regimes where both theories have to be applied in order to describe the
400: correct physics. Early time cosmology or physics of black holes are only
401: two regimes where the interplay of quantum and gravitational effects become important.
402: To nevertheless approach this theoretically interesting questions one might
403: hope for a fundamental quantum theory combining the Standard Model and
404: General Relativity. Until now one does not know what this unifying theory
405: is, but one has at least one possible candidate. This theory is
406: known as String Theory, which was studied intensively from various directions in
407: the last thirty years. A comprehensive introduction to the
408: subject can be found in \cite{GSWbook,JPbook,Zwiebach}.
409:
410: Perturbative String Theory is a quantum theory of one-dimensional extended objects which
411: replace the ordinary point particles. These fundamental strings can appear in various vibrational
412: modes which at low energies are identified with different particles. The characteristic
413: length of the string is $\sqrt{\alpha'}$, where $\alpha'$ is the Regge slope.
414: Hence, the extended nature of the strings only becomes apparent close to the string
415: scale $1/\sqrt{\alpha'}$. The string spectrum naturally includes a mode corresponding to the graviton.
416: This implies that Sting Theory indeed includes gravity and as we will further discuss below
417: reduces to Einsteins theory at low energies. It most likely provides a renormalizable
418: quantum theory of gravity around a given background. It avoids the ultra-violet divergences of graviton
419: scattering amplitudes in field theory by smearing out the location of the interactions.
420:
421:
422: The extended nature of the fundamental strings poses strong consistency constraints on the
423: theory. Non-tachyonic String Theories (Superstring Theories) require space-time supersymmetry
424: and predict a ten-dimensional space-time at weak coupling.
425: Altogether there are only five consistent String Theories, which are called
426: type IIA, type IIB, heterotic $SO(32)$ and $E_8 \times E_8$ and type I. These theories are connected
427: by various dualities and one may eventually hope to unify all of them into one fundamental theory \cite{Dual, JPbook}.
428:
429: As striking a proper formulation of such a fundamental theory might be, much of its uniqueness and beauty could
430: be spoiled in attempting to extract four-dimensional results. This is equally true for the five String Theories
431: formulated in ten dimensions. One approach to reduce String Theory from ten to four space-time dimensions
432: is compactification on a geometric background of the form $\Mext \times \Mint$. $\Mext$ is
433: identified with our four-dimensional world, while $\Mint$ is chosen to be small and compact,
434: such that these six additional dimensions are not visible in experiments. This however induces a
435: high amount of ambiguity, since String Theory allows for various consistent choices of $\Mint$.
436: Eventually one would hope to find a String Field Theory formulated in ten dimensions, which
437: resolves this ambiguity and dynamically chooses a certain background. %also describes all sting states at once.
438: However, such a theory is still lacking and one is forced to take a sideway
439: to find and explore consistent string backgrounds.
440:
441: For a given background, the ten-dimensional theory can then be reduced to four dimensions by a
442: Kaluza-Klein compactification \cite{KaluzaKlein} (for a review on Kaluza-Klein reduction see
443: e.g.\ \cite{KK-review}). This amounts to expanding
444: the fields into modes of $\Mint$ and results in a full tower of Kaluza-Klein modes for
445: each of the string excitations. Additionally there are winding modes corresponding to
446: strings winding around cycles in $\Mint$.
447: Generically it is hard and phenomenologically not interesting
448: to deal with these infinite towers of modes and an effective description is needed.
449:
450:
451: In order to extract an effective formulation one may first integrate out the
452: massive string excitations with masses of order $1/\sqrt{\alpha'}$. This is possible
453: due to the fact that the string scale $1/\sqrt{\alpha'}$ is usually set to be of order the
454: Planck scale such that gravity couples with Newtonian strength.
455: In the point-particle limit $\alpha' \rightarrow 0$ the effective theory
456: describing the massless string modes is a supergravity theory (see e.g.~\cite{GSWbook,JPbook}).
457: It can be constructed by calculating string scattering amplitudes for massless states.
458: One then infers an effective action for these fields encoding the same
459: tree level scattering vertices. An example is the three-graviton scattering amplitude
460: in String Theory, which in an effective description can be equivalently obtained from the
461: ten-dimensional Einstein-Hilbert term.
462: Repeating the same reasoning for all other massless string modes
463: yields a ten-dimensional supergravity theory for each of the five String Theories.
464:
465: In a similar spirit one can also extract an effective Kaluza-Klein theory. For a
466: compact internal manifold $\Mint$ the first massive Kaluza-Klein
467: modes have a mass of order $1/R$, where $R$ is the `average radius' of $\Mint$. Hence,
468: choosing $\Mint$ to be sufficiently small these modes become heavy and can be integrated
469: out. On the other hand, $\Mint$ has to be large enough that winding modes of length $\sqrt{\alpha'}$
470: can be discarded. Together for $p$ being the characteristic momentum of the lower-dimensional fields
471: an effective description of the massless modes is valid in the regime $1/p \gg R \gtrsim \sqrt{\alpha'}$.
472:
473:
474: The structure of the four-dimensional theory obtained by such a reduction highly depends
475: on the chosen internal manifold. The properties of $\Mint$ determine the amount of supersymmetry
476: and the gauge-groups of the lower-dimensional theory. Generically one insists that $\Mint$
477: preserves some of the ten-dimensional supersymmetries. This is due to the fact that string
478: theory on supersymmetric backgrounds is under much better control and various consistency
479: conditions are automatically satisfied. It turns out that looking for a supersymmetric theory
480: with a four-dimensional Minkowski background the internal manifold has to be a
481: Calabi-Yau manifold \cite{Greene}. From a phenomenological point of view the resulting low-energy
482: supergravity theories need to include gauged matter fields filling the spectrum of the desired
483: gauge theory such as the supersymmetric Standard Model. However, parameters like the size and
484: shape of the compact space appear as massless neutral scalar fields in four dimensions.
485: They label the continuous degeneracy of consistent backgrounds
486: $\Mint$ and are generically not driven to any particular value;
487: they are moduli of the theory. In a Standard Model-like vacuum these moduli have to be massive, such
488: that they are not dynamical in the low-energy effective action. Therefore one needs to identify
489: a mechanism in String Theory which induces a potential for these scalars. As it is
490: well-known for supergravity theories this potential can provide at the same time
491: a way to spontaneously break supersymmetry.
492:
493: To generate a moduli-dependent potential in a consisted String Theory setup is a non-trivial task
494: and requires further refinements of the standard compactifications. Recently, much effort was
495: made to establish controllable mechanisms to stabilize moduli fields in type II
496: String Theory. The three most popular approaches are the inclusion of background
497: fluxes \cite{Strominger1}--\cite{TGL2}, instanton corrections \cite{Witten,HM,non-pert,Curio} and
498: gaugino condensates \cite{gaugino,non-pert}.
499: This raised the hope to find examples of string vacua with all moduli being
500: fixed \cite{KKLT,KachruK,TGL2,DSFGK}. Moreover, phenomenologically interesting scenarios
501: for particle physics and cosmology can be constructed within these setups \cite{reviewPP,reviewcosmo}.
502:
503: In contrast to $E_8 \times E_8$ and $SO(32)$ heterotic String Theory and type I
504: strings both Type II String Theory do not consist of non-Abelian gauge-groups
505: in their original formulation. Thus most of the model building was first concentrated
506: on the heterotic String Theory as well as type I strings.
507: This has changed after the event of the D-branes \cite{JP,JPbook,D-branes,AD},
508: which naturally induce non-trivial
509: gauge theories. It turned out that compactifications with space-time filling D-branes
510: combined with moduli potentials due to fluxes or non-perturbative
511: effects provide a rich arena for model building in particle physics as well as cosmology
512: \cite{reviewPP,reviewcosmo}. One of the reasons is that consistent setups
513: with D-branes and fluxes generically demand a generalization of the Kaluza-Klein Ansatz to so-called warped
514: compactifications \cite{BB1,Verlinde,GSS,GKP}. Remarkably, these compactifications provide
515: a String Theory realization of models with large hierarchies \cite{Verlinde,DRS,Mayr,GSS,GKP}
516: as they were first suggested in \cite{RS}.
517:
518: One of the major motivation of this work it to analyze the low energy
519: dynamics of the (bulk) supergravity moduli fields within a brane world setup
520: with a non-vanishing potential. Hence, we will more carefully introduce the basic
521: constituents in the following.
522:
523:
524: \section{Compactification and moduli stabilization}
525: \label{geom_red}
526:
527:
528: Sting Theory is consistently formulated in a ten-dimensional space-time.
529: In order in order to make contact with our four-dimensional observed world
530: one is forced to assign six of these dimensions to an invisible sector.
531: This can be achieve by choosing these dimensions to be
532: small and compact and not detectable in present experiments.
533: Even though the additional dimensions are not observed directly,
534: they influence the resulting four-dimensional physics in a crucial way.
535:
536: The idea of geometric compactification is rather old
537: and goes back to the work of Kaluza and Klein in 1920 considering
538: compactification of five-dimensional gravity on a circle \cite{KaluzaKlein}. They
539: aimed at combining gravity with $U(1)$ gauge theory in a higher-dimensional
540: theory. Through our motivations have changed, the techniques are very similar
541: and can be generalized to the reduction from ten to four dimensions.
542:
543: In the Kaluza-Klein reduction one starts by specifying an Ansatz for the background
544: space-time \cite{KK-review}. Topologically it is assumed to be a manifold of the product structure
545: \beq \label{product_Ansatz}
546: \cM_{10} = \Mext \times \Mint\ ,
547: \eeq
548: where $\Mext$ represent the four observed non-compact dimensions and $\Mint$ correspond
549: to the compact internal manifold. On this space one specifies a block-diagonal
550: background metric
551: \beq \label{background_metric}
552: ds^2 = g^{(4)}_{\mu \nu}(x)\, dx^\mu dx^\nu + g^{(6)}_{mn}(y)\, dy^m dy^n
553: %\left(\begin{array}{cc} g^{(4)}_{\mu \nu}(x) & 0 \\ 0 &g^{(6)}_{mn}(y) \end{array} \right)\ ,
554: \eeq
555: where $g^{(4)}_{\mu \nu}$ is a four-dimensional Minkowski metric and $g^{(6)}_{ab}$ is
556: the metric on the compact internal subspace. More generally, one can include
557: a nontrivial warp factor $e^{2A(y)}$ depending on the internal coordinates $y$
558: into the Ansatz \eqref{background_metric}. This amounts to replacing
559: $g^{(4)}_{\mu \nu}(x)$ with $e^{2A(y)} g^{(4)}_{\mu \nu}(x)$ which is
560: the most general Ansatz for a Poincar\'e invariant four-dimensional metric \cite{deWitSD,Strominger1,BB1,Verlinde,GSS,GKP}.
561: The functional form of the warp factor is then determined by demanding the
562: background Ansatz to be a solution of the supergravity theory. It becomes
563: a non-trivial function in the presence of localized sources such as D-branes.
564: However, for simplicity we will restrict ourselves to the Ansatz \eqref{background_metric}
565: in the following.
566:
567: The lower-dimensional theory is obtained by expanding all fields into modes of
568: the internal manifold $\Mint$. As an illustrative example we discuss
569: the Kaluza-Klein reduction of a ten-dimensional scalar $\Phi(x,y)$ fulfilling
570: the ten-dimensional Laplace equation $\Delta_{10} \Phi = 0$ \cite{KK-review}. Using the Ansatz \eqref{background_metric}
571: the Laplace operator splits as $\Delta_{10}=\Delta_4 + \Delta_6$ and we may apply the
572: fact that $\Delta_6$ on a compact space has a discrete spectrum. The coefficients arising in
573: the expansion of $\Phi(x,y)$ into eigenfunctions of $\Delta_6$
574: are fields depending only on the coordinates of $\Mext$. From a four dimensional
575: point of view the term $\Delta_6 \Phi$ thus appears as a mass term.
576: One ends up with an infinite tower of massive states with masses
577: quantized in terms of $1/R$, where $R$ is the `radius' of $\Mint$ such that
578: $\text{Vol}(\Mint)$ is of order $R^6$. Choosing the internal manifold to be small
579: enough the massive Kaluza-Klein states become heavy and can be integrate
580: out. The resulting effective theory encodes the dynamics of the four-dimensional fields
581: associated with the massless Kaluza-Klein modes satisfying
582: \beq \label{zero_modes}
583: \Delta_6 \Phi(x,y)\ =\ 0\ .
584: \eeq
585: In chapter \ref{TypeII} we review how this procedure can be generalized to all other
586: fields present in the ten-dimensional supergravity theories. This also includes the metric itself \cite{KK-review}.
587: Equation \eqref{background_metric} specifies the ten-dimensional background metric and a gravity theory
588: describes variations around this Ansatz.
589: In the non-compact dimensions these correspond to the four-dimensional graviton and
590: the effective action reduces to the standard Einstein-Hilbert term for the metric.
591: The situation changes for the internal part of the metric. Massless fluctuations of $g_{mn}(y)$
592: around its background value, such as changes of the size and shape of $\Mint$, correspond to scalar and vector
593: fields in four-dimensions.
594: %Zero modes of the corresponding constraint equations
595: %replacing \eqref{zero_modes} encode variations around the background metric \eqref{product_Ansatz}.
596: %These correspond to the four-dimensional graviton as well as massless deformations of the
597: %internal metric $g_{mn}(y)$.
598: As a result the four-dimensional
599: theory consists of a huge set of scalar and vector fields arising
600: as coefficients in the expansion of the ten-dimensional fields into zero modes of $\Mint$.
601: In order that the four dimensional theory inherits some of the supersymmetries
602: of the underlying ten-dimensional supergravity theory one restricts to background
603: manifolds with structure group in $SU(3)$ such as Calabi-Yau manifolds or six-tori.
604: This implies that the Kaluza-Klein modes reside in supermultiplets with dynamics encoded
605: by a supergravity theory.
606:
607:
608: As already remarked above
609: every compactification induces a set of massless neutral
610: scalars called moduli. In Calabi-Yau compactifications
611: it typically consists of more then 100 scalar fields parameterizing the
612: geometry of $\Mint$, which is clearly in conflict with the known particle spectrum.
613: It is a long-standing problem to find a
614: mechanism within String Theory to generate a potential for these fields.
615: Finding such a potential will fix
616: their values in a vacuum and make them sufficiently massive such
617: that they can be discarded from the observable spectrum.
618: Above we already listed the three most popular possibilities to generate such a
619: potential: background fluxes, instanton corrections
620: and gaugino condensation. Let us now focus our attention to the first mechanism,
621: since fluxes will play a major role in this work.
622:
623: To include background fluxes amounts to allowing for non-trivial vacuum expectation value of certain field
624: strengths \cite{Strominger1}--\cite{TGL2}. Take as an example a tensor field $B_2$. If its field strength
625: $H_3 = dB_2$ admits a background flux $H_3^{\rm flux} = \big<dB_2 \big>$, the kinetic term of
626: $B_2$ yields a contribution \cite{GKP}
627: \beq \label{flux_pot}
628: \int_{\cM_{10}} H_3^{\rm flux} \wedge * H_3^{\rm flux}\ ,
629: \eeq
630: which via the Hodge-$*$ couples to the metric and its deformations. Insisting on four-dimensional
631: Poincar\'e invariance of the background, non-trivial fluxes can only be induced on internal three-cycles
632: $\gamma$.
633: %In string theory they are quantized as
634: %\beq
635: % \frac{1}{(2\pi)^2 \alpha'} \int_\gamma H_3^{\rm flux} \in \bbZ\ , \qquad \gamma \in H_3(\Mint,\bbZ) \ .
636: %\eeq
637: The terms \eqref{flux_pot} induce a non-trivial
638: potential for the deformations of the internal metric $g_{mn}(y)$ which generically
639: stabilizes the corresponding moduli fields at a scale $m_{\rm flux} \sim \alpha'/R^3$ \cite{GKP,DWG}.
640:
641:
642: %Even though moduli stabilization with fluxes is rather controllable, t
643: There are at least
644: two further important points to remark. Firstly, note that in general one is not completely free to
645: choose the fluxes, but rather has to obey certain consistency conditions.
646: Fluxes generically induce a charge which has to be canceled on a compact space.
647: Hence, the setup needs to be enriched by objects carrying a negative charge \cite{GKP}.
648: Secondly, it is usually the case that fluxes do not stabilize all moduli of the theory.
649: In order to induce a potential for the remaining fields, one needs to include
650: non-perturbative effects such as instantons and gaugino condensates.
651: Various recent work \cite{non-pert} is intended to get some deeper insight
652: into the nature of these corrections.
653:
654: % To find consistet setups with all moduli fixed
655: %by a combination of these mechanisms remains a non-trivial task \cite{D}.
656:
657:
658:
659:
660: \section{Brane World Scenarios}
661: \label{braneworlds}
662:
663: In the middle of the 90's, the discovery of the D-brane opened
664: a new perspective for String Theory \cite{JP}. On the one hand, D-branes where
665: required to fill the conjectured web of string dualities \cite{Dual,JPbook}.
666: Their appearance supports the hope for a more fundamental underlying theory
667: unifying all the known String Theories. Moreover, they led to the conjecture of various
668: new connections between String Theories and supersymmetric gauge theories,
669: such as the celebrated AdS/CFT correspondence \cite{reviewAdSCFT}.
670: From a direct phenomenological point of view, they opened a whole new arena
671: for model building \cite{reviewPP}, since they come equipped with a gauge theory.
672:
673: More precisely, D-branes are extended objects defined as subspaces of the ten-dimensional
674: space-time on which open strings can end \cite{JP,JPbook,D-branes,AD}. Open strings with both ends on the same D-brane
675: correspond to an $U(1)$ gauge field in the low energy effective action. This gauge group gets enhanced
676: to $U(\cN)$ when putting a stack of $\cN$ D-branes on top of each other. At lowest order this
677: induces a Yang-Mills gauge theory in the low-energy effective action.
678: This fact allows to construct phenomenologically attractive
679: models from space-time filling D-branes consistently included in a compactification
680: of type II String Theory \cite{reviewPP}. The basic idea is that the Standard Model,
681: or rather its supersymmetric extensions, is realized on a stack of space-time filling D-branes.
682: The matter fields arise from dynamical excitations of the brane around its background configuration.
683: This is similar to the situation in standard compactifications discussed in the beginning of the
684: previous section, where moduli fields parameterize fluctuations of the background
685: metric on $\Mint$. The crucial difference is that fluctuations of the D-branes are charged under
686: the corresponding gauge group and can yield chiral fermions in topologically non-trivial
687: configuration \cite{reviewPP}.
688:
689: In addition to the applications in Particle Physics, D-branes can serve as essential
690: ingredients to construct cosmological models. Their non-perturbative nature can be
691: used to circumvent the no-go theorem excluding the possibility of de Sitter vacua
692: in String Theory \cite{KKLT,BKQ,reviewcosmo}.
693: Furthermore, similar to the fundamental string, D-branes are dynamical objects, which can
694: move through the ten-dimensional ambient space. In certain circumstances this
695: dynamical behavior was conjectured to be linked to a cosmological evolution \cite{reviewcosmo}.
696:
697: There are basically three steps to extract phenomenological data from brane world
698: scenarios. Firstly, one has to actually construct consistent examples
699: yielding the desired gauge groups, field content and amount of supersymmetry.
700: Secondly, to determine the dynamics of the theory one needs to evaluate the
701: low energy effective action of the brane excitations and the gauge neutral
702: bulk moduli. This can then be combined with the approach
703: to generate potentials by a flux-background and non-perturbative effects.
704:
705:
706: The resulting theory may exhibit various phenomenologically interesting features.
707: As briefly discussed in section \ref{geom_red} it can yield moduli stabilization in the vacuum.
708: Moreover, if the vacuum breaks supersymmetry this generically results in a set of soft supersymmetry
709: breaking terms for the charged matter fields on the D-branes (see ref. \cite{KL,BIM} for
710: a generic string inspired supergravity analysis).
711: These can be computed from the effective low energy action
712: as it has been carried out in refs.\ \cite{Soft1,GGJL,Soft2}.
713: On the other hand, anti-branes (or brane fluxes)
714: can be used to generate a positive cosmological constant \cite{KKLT,BKQ}.
715:
716: Even though this general approach sounds promising, it is extremely hard to
717: address all these issues at once. Hence, one is usually forced to
718: either concentrate on specific models or on one or the other ingredient
719: to develop techniques and to extract general results.
720:
721: As an example, one can already check if space-time filling D-branes and fluxes
722: alone can be consistently included in a compactification.
723: Namely, since D-branes are charged under certain fields of the
724: bulk supergravity theory they contribute a source term in the Bianchi identities
725: of these fields \cite{JP,JPbook,D-branes,AD}. This is similarly true for non-trivial
726: background fluxes. One can next apply the Gau\ss~law for the compact internal space such that
727: consistency requires internal sources to cancel. In this respect D-branes
728: are the higher dimensional analog of say positively charged particles.
729: Putting such a particle in a compact space, the field lines have to end somewhere
730: and we have to require for negative sources. In String Theory these negative sources are
731: either appropriately chosen anti-D-branes or `orientifold planes' \cite{JP,AD}.
732: Even though it is possible to construct consistent
733: scenarios with D-branes and anti-D-branes only, one may further insist to keep a
734: $D=4$ supergravity theory. This is mainly due
735: to the fact that these models are under much better control
736: and are not plagued by instabilities.
737: This favors the inclusion of appropriate orientifold planes, since there
738: negative tension cancels the run-away potentials for the moduli induced by D-branes.
739: In figure \ref{braneworld_fig} we schematically picture some ingredients of a
740: brane-world model.
741:
742: \begin{figure}[h]
743: \begin{center}
744: \includegraphics[height=5cm]{braneworld.eps}
745: \caption{\textit{Brane-world scenario on $\Mext \times \Mint$
746: with space-time filling D-branes, orientifold planes and background fluxes.}}
747: \label{braneworld_fig}
748: \end{center}
749: \end{figure}
750:
751: Orientifold planes arise in String Theories constructed
752: form type II strings by modding out world-sheet parity plus a geometric
753: symmetry $\sigma$ of $\Mext \times \Mint$ \cite{JP,AD}. On the level of the full String Theory
754: this implies that non-orientable string world-sheets, such as the Klein bottle
755: or the M\"obius strip, are allowed.
756: Focusing on the effective action orientifolds break part or all of
757: the supersymmetry of the low-energy theory. By imposing
758: appropriate conditions on the orientifold projection and the included D-branes
759: the setup can be adjusted to preserve exactly half of the original supersymmetry.
760:
761: {}From a phenomenological point of view spontaneously broken $\cN=1$
762: theories are of particular interest.
763: Starting from type II String Theories in ten space-time dimensions,
764: one can compactify on Calabi-Yau
765: threefolds to obtain $\cN=2$ theories in four dimensions.
766: This $\cN=2$ is further broken to $\cN=1$ if in addition
767: background D-branes and orientifold planes are present \cite{Ori,GKP,AAHV,BBKL,BH}.
768: The presence of background fluxes or other
769: effects generating a potential results in a
770: spontaneously broken $\cN=1$ theory \cite{Bachas}--\cite{TGL2}.
771: To examine this setup on the level of the effective action is one of the
772: motivations for this work. Note that all these brane world scenarios are conjectured
773: to admit a higher dimensional origin in a more fundamental theory, which we briefly introduce next.
774: However, it is important to keep in mind that this unifying theory is much
775: less understood then the five String Theories.
776:
777:
778: \subsection{From dualities to M- and F-theory}
779:
780: At the first glance in seems as if we have to choose one or
781: the other String Theory in which we aim to construct a specific
782: model. However, it turns out that many of these choices are
783: actually equivalent and linked by various dualities \cite{Dual,JPbook}. The full set of
784: dualities forms a interlocking web between all five String
785: Theories (see figure \ref{web}).
786:
787: \begin{figure}[h]
788: \begin{center}
789: \includegraphics[height=5cm]{dual_web.eps}
790: \caption{\textit{The duality web of String Theories.}}
791: \label{web}
792: \end{center}
793: \end{figure}
794:
795: As an example type IIA compactified on a circle of radius $R$
796: is shown to be equivalent to type IIB on a circle of radius $1/R$ \cite{JPbook,revT-dual}.
797: This duality is termed T-duality and relates two String Theories
798: at weak string coupling \cite{revT-dual}. There are also strong/weak dualities
799: such as S-duality, which is a symmetry of the type IIB String
800: Theory \cite{Dual}. Both of these dualities can be generalized and applied
801: to standard Calabi-Yau compactifications as well as
802: brane-world scenarios.
803:
804: A prominent example is mirror symmetry which can be interpreted as performing
805: several T-dualities \cite{SYZ}.
806: It associates to each Calabi-Yau manifold $Y$ a corresponding mirror Calabi-Yau $\tilde Y$ \cite{Mirror}.
807: Within the framework of String Theory it can be argued that
808: type IIA compactified on $Y$ is fully equivalent to type IIB strings on
809: $\tilde Y$. From a mathematical point of view mirror symmetry
810: exchanges the odd cohomologies of $Y$ with the even cohomologies of $\tilde Y$
811: and vice versa. Even stronger it suggests that the moduli spaces of the
812: two Calabi-Yau manifolds are identified. Remarkably, in specific
813: examples this allows to calculate stringy corrections to the theory on $Y$
814: from geometrical data of $\tilde Y$. Mirror symmetry can be generalized
815: to setups with D-branes \cite{Aspinwall} and eventually should identify type IIA and type IIB
816: brane world scenarios. This raises various non-trivial questions such
817: as in which way mirror symmetry applies to flux compactifications \cite{Vafa_NCY,GLMW}.
818:
819: Let us also introduce S-duality in slightly more detail \cite{Dual,JPbook}. Type
820: IIB String Theory contains in addition to the fundamental string
821: also a D-string (D1-brane).
822: It can now be argued that the theory where the fundamental string
823: is at low coupling $g_s$, and hence the D-string is very heavy, is
824: dual to a theory at $1/g_s$ with the role of both strings exchanged.
825: Carefully identifying the fields, S-duality is also shown to be a
826: symmetry of the corresponding type IIB low-energy effective action.
827: This strong/weak duality is actually part of a larger symmetry group
828: $Sl(2,\bbZ)$. It has been suggested in \cite{Vafa} that this duality
829: group admits a geometric interpretation in terms of two additional
830: toroidal dimensions. This twelve dimensional construction was named
831: F-theory. The additional two dimensions are necessarily a compact torus,
832: which however in compactifications can be non-trivially fibered over
833: the compactification manifold. This naturally applies to type IIB
834: brane-world scenarios, which generically admit backgrounds corresponding
835: these non-trivial compactifications \cite{Sen,DRS,GKP}.
836:
837: The existence of these various dualities suggests that the ten-dimensional
838: String Theories are actually just different limits of a more fundamental
839: theory \cite{Dual} as pictured in figure \ref{web}.
840: This mysterious theory unifying all five String Theories was named
841: M-theory. In general, not much is known about its actual
842: formulation and the required structures
843: are far less understood then the one for String Theory. However,
844: there are certain regimes in which one believes to find some
845: hints of its existence. This also includes the existence
846: D-branes, which fit into this picture as they occur from
847: higher-dimensional objects termed M-branes. There also is
848: a unique supergravity theory in eleven dimensions \cite{CJS},
849: which is interpreted to be the low-energy limit of M-theory.
850: In the final chapter of this article it will be this
851: low-energy theory which allows us lift the orientifold
852: compactifications to M-theory.
853:
854:
855:
856: \subsection{Topics and outline of this article}
857:
858: After this brief general introduction let us now turn to the actual topics
859: of this article. As just discussed, an essential step to extract
860: phenomenological properties of string vacua with
861: (spontaneously broken) $N=1$ supersymmetry in brane world
862: scenarios is to determine the low energy effective
863: action. In this work we focus
864: on type IIA and IIB String Theory compactified on generic Calabi-Yau
865: orientifolds and determine their low energy effective action in
866: terms of geometrical data of the Calabi-Yau orientifold and
867: the background fluxes. We include D-branes
868: for consistency, but freeze their matter fields (and moduli) concentrating
869: on the couplings of the bulk moduli. We also provide a detailed discussion
870: of the resulting $N=1$ moduli space in the chiral and the dual linear
871: multiplet description and check mirror symmetry in the large volume--large complex
872: structure limit. Moreover, we show at the level of the effective actions
873: that Calabi-Yau orientifolds with fluxes admit a natural embedding
874: into F- and M-theory compactifications.
875:
876: In \textit{chapter \ref{TypeII}} we first briefly review standard Calabi-Yau
877: compactifications of type IIA and type IIB supergravity and discuss the resulting
878: $N=2$ supergravity action. In doing so we focus
879: on the geometry of the moduli space $\cM^{\rm SK} \times \cM^{\rm Q}$ spanned by
880: the scalars of the $N=2$ supergravity theory. Supersymmetry constrains
881: it to locally admit this product form, where $\cM^{\rm SK}$ is a
882: special K\"ahler manifold and $\cM^{\rm Q}$ is a quaternionic manifold.
883: Furthermore, we introduce $N=2$ mirror symmetry on the level of
884: the effective action and present a somewhat non-standard construction
885: of the mirror map between the IIA and IIB quaternionic moduli spaces
886: reproducing the results of \cite{BGHL}.
887:
888: In \textit{chapter \ref{effective_actO}} we immediately turn to the
889: compactification of type II theories on Calabi-Yau orientifolds.
890: We start with a more detailed introduction to setups with
891: D-branes and orientifold planes and comment on consistency and
892: supersymmetry conditions. As already mentioned in section \ref{braneworlds}
893: orientifold planes are essential ingredients to obtain supersymmetric theories in
894: brane-world compactifications. They arise in String Theories modded
895: out by a geometrical symmetry $\sigma$ of $\Mext \times \Mint$
896: in addition to the world-sheet parity operation. We demand
897: $\Mint$ to be a generic Calabi-Yau manifold admitting an
898: isometric involutive symmetry $\sigma$. It turns out that in order to preserve
899: $N=1$ supersymmetry $\sigma$ has to be a holomorphic
900: map in type IIB and an anti-holomorphic map in type IIA
901: compactifications. Taking into account further properties of $\sigma$
902: one finds three supersymmetric setups \cite{AAHV,BH}: (1) IIB orientifolds with
903: $O3/O7$ planes, (2) IIB orientifolds with $O5/O9$ planes and
904: (3) IIA orientifolds with $O6$ planes.
905:
906: The spectrum of these
907: theories was first determined in \cite{BH}.
908: However, the effective action was only computed
909: for special cases of type IIB Calabi-Yau orientifolds with $O3/O7$
910: planes \cite{GKP,BBHL}. In \cite{TGL1} we generalized these results and
911: also included an analysis of $O5/O9$ setups.
912: For type IIA brane-world scenarios the calculation of the low energy
913: supergravity theory was mainly concerned with orbifolds of six-tori \cite{CP,reviewPP} for
914: which conformal field theory techniques can be applied. Complementary,
915: the dynamics of the bulk theory can extracted for general type IIA Calabi-Yau orientifolds by
916: using a Kaluza-Klein reduction as shown in our publication \cite{TGL2}.
917: In chapter \ref{effective_actO} we review the first parts of refs. \cite{TGL1,TGL2}
918: and determine the $N=1$ effective action of all three setups. We extract
919: the K\"ahler potential and the gauge-kinetic couplings by first assuming that
920: no background fluxes are present. The $N=1$ moduli space is shown to
921: be a local product $\tilde \cM^{\rm SK} \times \tilde \cM^{\rm Q}$, where
922: $\tilde \cM^{\rm SK}$ is a special K\"ahler manifold inside $\cM^{\rm SK}$
923: and $\tilde \cM^{\rm Q}$ is a K\"ahler manifold inside the quaternionic
924: manifold $\cM^{\rm Q}$.
925:
926: We end chapter \ref{effective_actO} with a discussion of
927: mirror symmetry for Calabi-Yau orientifolds and determine
928: the necessary conditions
929: on the involutive symmetries of the mirror IIA and IIB orientifold theories.
930: By specifying two types of
931: special coordinates on the IIA side, we are able to identify the large complex
932: structure limit of IIA orientifolds with the large volume limits of IIB orientifolds
933: with $O3/O7$ and $O5/O9$ planes.
934:
935: In \textit{chapter \ref{lin_geom_of_M}} we present a more detailed analysis of
936: the $N=1$ moduli space geometry of Calabi-Yau orientifold compactifications \cite{TGL1,TGL2}.
937: The special K\"ahler manifold $\tilde \cM^{\rm SK}$ inherits its geometrical
938: structure directly from $N=2$, such that we focus our attention
939: to the K\"ahler manifold $\tilde \cM^Q$ inside the quaternionic space.
940: We show that the definition of the K\"ahler coordinates as well as
941: certain no-scale type conditions can be more easily understood
942: in terms of the `dual' formulation where some chiral multiplets of the Calabi-Yau
943: orientifold are replaced by linear multiplets. A linear multiplet consists of
944: a real scalar and an anti-symmetric two-tensor as bosonic fields. In the massless case
945: this two-tensor is dual to a second real scalar and one is led back to the chiral description.
946: In order to do set the
947: stage for the orientifold analysis we first
948: review $N=1$ supergravity with several linear multiplets following \cite{BGG}.
949: The transformation into linear multiplets corresponds to a Legendre transformation of
950: the K\"ahler potentials and coordinates. In the dual picture
951: the characteristic functions for type IIB orientifolds take a particularly
952: simple form. Moreover, in type IIA orientifolds the Legendre transform is
953: essential to make contact with the underlying $N=2$ special geometry.
954: As a byproduct we determine an entire new class of no-scale
955: K\"ahler potentials which in the chiral formulation
956: can only be given implicitly as the solution of some constraint equation.
957: These new insights will enable us to give an direct construction of the K\"ahler
958: manifold $\tilde \cM^{\rm Q}$ in analogy to the moduli space of supersymmetric Lagrangian submanifolds \cite{Hitchin2}.
959: Moreover, this sets the stage to generalize the reduction to orientifolds of certain
960: non-Calabi-Yau manifolds introduced in \cite{HitchinGCM,Gualtieri}.
961:
962: In \textit{chapter \ref{fluxesAB}} we redo the Kaluza-Klein compactification by additionally allowing
963: for non-trivial background fluxes. For $O3/O7$ orientifolds this amounts to
964: a generalization of the analysis presented in \cite{GKP,BBHL} and
965: confirms that the Gukov-Vafa-Witten superpotential \cite{GVW} encodes
966: the potential due to background fluxes. However, we show that
967: for orientifolds with $O5/O9$ planes background fluxes generically
968: result in a non-trivial superpotential, $D$-terms as well as
969: a direct mass term for a linear multiplet.
970: Following this observation, supergravity theories with massive linear
971: multiplets coupled to vector and chiral multiplets where further analyzed
972: in \cite{mass_tensors}. Surprisingly, in type IIA orientifolds with background fluxes
973: the superpotential depends on all (bulk) moduli fields of the theory.
974: In \cite{KachruK} an equivalent observation was made for the underlying $N=2$
975: theory. This suggests that all geometric moduli can be stabilized in a supersymmetric
976: vacuum \cite{KachruK,TGL2}. In ref.~\cite{DeWGKT} this
977: was shown to be possible at large volume and small string coupling (see also \cite{VZ}).
978:
979: The IIA superpotential is expected to receive non-perturbative corrections
980: from world-sheet as well as D-brane instantons. In the final section of
981: chapter \ref{fluxesAB} we derive that for supersymmetric type IIA and type IIB instantons
982: the respective actions are linear in the chiral coordinates and therefore can result in holomorphic
983: corrections to the superpotential.
984:
985: In \textit{chapter \ref{M-F-embedding}} we embed type IIB and type IIA
986: orientifolds into F- and M-theory compactifications.
987: Orientifolds with $O3/O7$-planes can be obtained as a limit of
988: F-theory compactified on elliptically fibered Calabi-Yau fourfolds \cite{Sen}.
989: We check this correspondence on the level of the effective action
990: by first compactifying M-theory on a specific Calabi-Yau fourfold and comparing
991: the result with the effective action of $O3/O7$ orientifolds compactified
992: on a circle to $D=3$. The low energy effective action of
993: M-theory compactifications on Calabi-Yau four-folds
994: was determined in \cite{HL,BHS} and we use their results in
995: a slightly reformulated version. Moreover, it turns out
996: that this duality is best understood in the dual pictures
997: where three-dimensional vector multiplets are kept in the
998: spectrum and the K\"ahler potential is an explicit function of the
999: moduli. We determine simple solutions to the
1000: fourfold consistency conditions for which we find perfect matching between the
1001: orientifold and M-theory compactifications. This correspondence can be lifted
1002: to $D=4$ where M-theory on an elliptically fibered Calabi-Yau fourfold descents
1003: to an F-theory compactification.
1004:
1005: We end this chapter by also discussing the embedding of type IIA
1006: orientifolds into a specific class of $G_2$ compactifications of M-theory
1007: as suggested in \cite{KMcG}.
1008: Restricting the general results of \cite{PT,HM,Hitchin1,GPap,BW} to a specific
1009: $G_2$ manifold and neglecting the contributions arising from the singularities
1010: we show agreement between the low energy effective
1011: actions \cite{TGL2}. In \cite{TGL2} we discovered that only parts of the
1012: orientifold flux superpotential decent from fluxes in an M-theory
1013: compactifications on manifolds with $G_2$ holonomy.
1014: However, as we will argue one of the missing terms is
1015: generated on $G_2$ structure manifolds with non-trivial
1016: fibrations. However, the higher-dimensional origin of the term
1017: involving the mass parameter of massive type IIA supergravity
1018: remains mysterious.
1019:
1020: This article is mainly based on the publications
1021: \cite{TGL1} and \cite{TGL2} of the author.
1022: However, we also present various new results.
1023: Namely, it turns out to be possible to reformulate the results of
1024: \cite{TGL1,TGL2} in a very elegant and powerful way adapted to
1025: Hitchin's analysis of special even and odd forms on six-manifolds \cite{HitchinGCM,Hitchin1}.
1026: This allows for a better understanding of the N=1 moduli
1027: space inside the quaternionic manifold and
1028: suggests a generalization to non-Calabi-Yau orientifolds.
1029: Moreover, we included a detailed analysis of
1030: the orientifold limit of the F-theory embedding of type IIB orientifolds.
1031: In addition we identify the higher-dimensional origin of a second flux term of the
1032: IIA orientifold superpotential being due to a non-trivial fibration of a
1033: $G_2$ structure manifold.
1034:
1035:
1036: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1037: %
1038: % Type II on Calabi-Yau
1039: %
1040: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1041:
1042:
1043:
1044: \chapter{Calabi-Yau compactifications of Type II theories}
1045: \label{TypeII}
1046:
1047:
1048: In this section we review compactifications of type IIA and type IIB
1049: supergravity on a Calabi-Yau manifold $Y$. These lead to $N=2$ supergravity
1050: theories in four dimensions expressed in terms of the characteristic data
1051: of the Calabi-Yau space. We start our discussion with some mathematical preliminaries.
1052: In section \ref{CY-mfds} we introduce Calabi-Yau manifolds and give a short
1053: description of their moduli spaces. In a next step we turn to compactifications
1054: of IIA and IIB supergravity on Calabi-Yau manifolds in section \ref{revIIB} and \ref{revIIA}.
1055: Finally, in section \ref{revMirror} we give a brief account of $N=2$ mirror symmetry
1056: applied at the level of the effective action. The mirror map for the quaternionic
1057: moduli spaces will be constructed.
1058:
1059: \section{Calabi-Yau manifolds and their moduli space}
1060: \label{CY-mfds}
1061:
1062: String theory is consistently formulated in a ten-dimensional target space.
1063: In order to reduce to a four-dimensional observable world, we choose
1064: the background to be of the form $\cM_{10}=\Mext \times \Mint$ as
1065: already given in \eqref{product_Ansatz}.
1066: Here $\Mint$ is a compact six-dimensional manifold, which, in principle, we are
1067: free to choose. Due to this Ansatz, the Lorentz group of $\cM_{10}$ decomposes
1068: as $SO(9,1)\rightarrow SO(3,1)\times SO(6)$, where $SO(6)$ is the generic
1069: structure group of a sixfold.
1070: However, demanding $\Mint$ to preserve the minimal amount of supersymmetry
1071: one has to pick a manifold with structure group $SU(3)$.
1072: They admit one globally defined spinor $\eta$, since the $SO(6)$ spinor
1073: representation $\bf{4}$ decomposes to $\bf{1}\oplus\bf{3}$.
1074: Further demanding this spinor $\eta$ to be covariantly constant
1075: reduces the class of background manifolds to manifolds with
1076: $SU(3)$ holonomy \cite{GSWbook}. These spaces are called Calabi-Yau manifolds and
1077: are complex K\"ahler manifolds, which are in addition Ricci flat \cite{Huebsch}.
1078:
1079: In terms of $\eta$ one can globally define a covariantly constant
1080: two-from $J$ (the K\"ahler form) and a three-form $\Omega$ (the holomorphic three-form).
1081: For a fixed complex structure these
1082: fulfill the algebraic conditions
1083: \beq
1084: J \wedge J \wedge J\ \propto\ \Omega \wedge \bar \Omega\ , \qquad J \wedge \Omega = 0\ .
1085: \eeq
1086: where the proportionality factor depends on the normalization of $\Omega$ with respect to $J$.
1087: Performing a Kaluza-Klein reduction on the background \eqref{product_Ansatz}
1088: the massless four-dimensional fields arise as the zero modes of
1089: the internal Laplacian \eqref{zero_modes} \cite{GSWbook,JPbook}. These zero modes are in one-to-one correspondence
1090: with harmonic forms on $Y$ and thus their multiplicity is counted by the dimension
1091: of the non-trivial cohomologies of the Calabi-Yau manifold. The Calabi-Yau condition poses
1092: strong constraints on the Hodge decomposition of the cohomology groups.
1093: The only non-vanishing cohomology groups are the even and odd cohomologies
1094: \bea \label{odd_even_cohom}
1095: H^{ev}& =& H^{(0,0)} \oplus H^{(1,1)} \oplus H^{(2,2)} \oplus H^{(3,3)}\ , \\
1096: H^{odd}& =& H^{(3,0)} \oplus H^{(2,1)} \oplus H^{(1,2)} \oplus H^{(0,3)}\ . \nn
1097: \eea
1098: Their dimensions $h^{(p,q)}=\dim H^{(p,q)}$ can be summarized in the Hodge diamond
1099: as follows
1100: \beq \label{hodge_diamond}
1101: \begin{array}{c}
1102: h^{(0,0)}\\
1103: h^{(1,0)} \qquad h^{(0,1)}\\
1104: h^{(2,0)}\qquad h^{(1,1)} \qquad h^{(0,2)}\\
1105: h^{(3,0)}\qquad h^{(2,1)}\qquad h^{(1,2)} \qquad h^{(0,3)}\\
1106: h^{(3,1)}\qquad h^{(2,2)} \qquad h^{(1,3)}\\
1107: h^{(3,2)} \qquad h^{(2,3)}\\
1108: h^{(3,3)}
1109: \end{array}\ =\
1110: \begin{array}{c}
1111: 1\\
1112: 0 \qquad 0\\
1113: 0\qquad h^{(1,1)} \qquad 0\\
1114: 1\qquad h^{(2,1)}\qquad h^{(2,1)} \qquad 1\\
1115: 0\qquad h^{(1,1)} \qquad 0\\
1116: 0 \qquad 0\\
1117: 1
1118: \end{array}\ .
1119: \eeq
1120: Let us introduce a basis for the different cohomology
1121: groups by always choosing the unique harmonic representative in
1122: each cohomology class. The basis of harmonic $(1,1)$-forms
1123: we denote by $\omega_A$ with dual harmonic $(2,2)$-forms
1124: $\tilde \omega^A $ which form a basis of $H^{(2,2)}(Y)$.
1125: $(\alpha_{\hat K}, \beta^{\hat L})$ are harmonic three-forms
1126: and form a real, symplectic basis on $H^{(3)}(Y)$. Together
1127: the non-trivial intersection numbers are summarized as
1128: \beq \label{int-numbers1}
1129: \int_Y \omega_A \wedge \tilde \omega^B = \delta_A^B\ , \qquad
1130: \int_Y \alpha_{\hat K} \wedge \beta^{\hat L} = \delta^{\hat L}_{\hat K}\ ,
1131: \eeq
1132: with all other intersections vanishing. Finally, we denote by $\vol(Y)$ the
1133: harmonic volume $(3,3)$-form of the Calabi-Yau space.
1134: In Table \ref{CYbasis} we summarize
1135: the non-trivial cohomology groups on $Y$ and denote their basis elements.
1136:
1137: \begin{table}[h]
1138: \begin{center}
1139: \begin{tabular}{| c | c | c |} \hline
1140: \rule[-0.3cm]{0cm}{0.8cm} cohomology group&
1141: dimension & basis
1142: \\ \hline
1143: \rule[-0.3cm]{0cm}{0.8cm} $H^{(1,1)}$ &
1144: $h^{(1,1)}$ & $\omega_A$
1145: \\ \hline
1146: \rule[-0.3cm]{0cm}{0.8cm} $H^{(2,2)}$ & $h^{(1,1)}$ & $\tilde \omega^A$
1147: \\ \hline
1148: \rule[-0.3cm]{0cm}{0.8cm} $H^{(3)}$ & $2h^{(2,1)}+2$ &
1149: $(\alpha_{\hat K},\beta^{\hat L})$ \\ \hline
1150: \rule[-0.3cm]{0cm}{0.8cm} $H^{(2,1)}$ &
1151: $h^{(2,1)}$ &
1152: $\chi_K$
1153: \\ \hline
1154: \rule[-0.3cm]{0cm}{0.8cm} $H^{(3,3)}$ &
1155: $1$&
1156: $\vol$
1157: \\ \hline
1158: \end{tabular}
1159: \caption{\small \label{CYbasis}
1160: \textit{Cohomology groups on $Y$ and their basis elements.}}
1161: \end{center}
1162: \end{table}
1163:
1164: In sections \ref{revIIA} and \ref{revIIB} we explain how these harmonics
1165: yield four-dimensional massless fields, when expanding the ten-dimensional
1166: supergravity forms. Furthermore, there are additional massless modes arising as
1167: deformations of the metric $g_{i\jb}$.
1168: Considering variations $R_{mn}(g+\delta g)$ of the Ricci-tensor which
1169: respect the Ricci-flatness condition $R_{mn}=0$ forces $\delta g$ to satisfy a
1170: differential equation (the Lichnerowicz equation).
1171: Solutions to this equation can be identified in case of a Calabi-Yau manifold
1172: with the harmonic $(1,1)$- and $(2,1)$-forms, which parameterize K\"ahler structure and
1173: complex structure deformations of $Y$ \cite{Tian, CdO, Huebsch}. The deformations of the
1174: K\"ahler form $J = i {g}_{i\bj}\, dy^i \wedge d\bar y^\jb$
1175: give rise to $h^{(1,1)}$ real scalars $v^{A}$ and one expands \footnote{%
1176: Globally only those deformations are allowed which keep the volume
1177: of $Y$ as well as its two- and four-cycles positive, i.e.\
1178: $\int_Y J \wedge J \wedge J \ge 0$, $\int_{S_4} J \wedge J \ge 0$ and $\int_{S_2} J \ge 0$.
1179: These conditions are preserved under positive rescalings of the fields $v^A$, such that
1180: they span a $h^{(1,1)}-$dimensional cone.}
1181: \beq\label{def-v}
1182: g_{i\bj} + \delta g_{i\bj} = -i\, J_{i\bj} = -i\, v^{A}\, (\omega_{A})_{i\bj} \ , \qquad A = 1, \ldots, h^{(1,1)}\ .
1183: \eeq
1184: These real deformations are complexified by the $h^{(1,1)}$ real scalars $b^A(x)$
1185: arising in the expansion of the B-field present in both type II string theories.
1186: More precisely we introduce the complex fields
1187: \beq \label{def-t_II}
1188: t^A = b^A + i\, v^A\ ,
1189: \eeq
1190: which parameterize the $h^{(1,1)}-$dimensional complexified K\"ahler cone \cite{CdO}.
1191:
1192: The second set of deformations are variations of the complex structure
1193: of $Y$. They are parameterized by complex scalar fields $z^{K}$
1194: and are in one-to-one correspondence with harmonic
1195: $(1,2)$-forms
1196: \beq\label{cs}
1197: \delta{g}_{ij} = \frac{i}{||\Omega||^2}\, \bar z^{K}
1198: (\bar \chi_{K})_{i\ib\bj}\,
1199: \Omega^{\ib\bj}{}_j \ , \quad K=1,\ldots,h^{(1,2)}\ ,
1200: \eeq
1201: where $\Omega$ is the holomorphic (3,0)-form,
1202: $\bar\chi_{K}$ denotes a basis of $H^{(1,2)}$ and we abbreviate
1203: $||\Omega||^2\equiv \frac1{3!}\Omega_{ijk}\bar\Omega^{ijk}$.
1204:
1205: Together the complex scalars $z^K$ and $t^A$ span the geometric moduli
1206: space of the Calabi-Yau manifold. It is shown to be locally a product
1207: \beq \label{geom-mod}
1208: \cM^{\rm cs} \times \cM^{\rm ks}\ ,
1209: \eeq
1210: where both factors are special K\"ahler manifolds of complex dimension
1211: $h^{(2,1)}$ and $h^{(1,1)}$ respectively. To make that more precise let us first
1212: discuss $\cM^{\rm cs}$. Its metric $G_{K \bar L}$ is given by \cite{Tian,Strominger2,CdO}
1213: \beq \label{chi_barchi}
1214: G_{K \bar L} = -\frac{\int_Y \chi_K \wedge \bar \chi_{ L}}{\int_Y \Ox \wedge \bar \Ox} \ ,
1215: \eeq
1216: where $\chi_K$ is related to the
1217: variation of the three-form
1218: $\Omega$ via Kodaira's formula
1219: \beq \label{Kod-form}
1220: \chi_K(z,\bar z) = \partial_{z^K} \Omega(z)+ \Omega(z)\, \partial_{z^K}\Kcs \ .
1221: \eeq
1222: With the help of \eqref{Kod-form} one shows that $G_{K \bar L}$ is a K\"ahler manifold,
1223: since we can locally find complex coordinates $z^K$ and a function
1224: $K(z,\bar z)$ such that
1225: \beq\label{csmetric}
1226: G_{K \bar L} = {\partial}_{z^K}\partial_{\bar z^L}\ \Kcs\ , \qquad \Kcs = -\ln\Big[ i \int_Y \Ox \wedge \bar \Ox\Big]
1227: = -\ln i\Big[\bar Z^\Kh\mathcal{F}_\Kh - Z^\Kh\bar{\mathcal{F}}_\Kh \Big]
1228: \ ,
1229: \eeq
1230: where the holomorphic periods $Z^\Kh, \mathcal{F}_\Kh$ are defined as
1231: \beq \label{pre-z}
1232: % \cF(Z) = \tfrac{1}{2} Z^\Kh \cF_{\Kh}\ ,\quad
1233: Z^\Kh(z) = \int_Y \Omega(z) \wedge \beta^\Kh\ , \qquad
1234: \cF_\Kh(z) = \int_Y \Omega(z) \wedge \alpha_\Kh\ ,
1235: \eeq
1236: or in other words $\Omega$ enjoys the expansion
1237: \beq\label{Omegaexp}
1238: \Omega(z) = Z^\Kh(z)\, \alpha_\Kh - \cF_\Kh(z)\, \beta^\Kh\ .
1239: \eeq
1240: The K\"ahler manifold $\cM^{\rm cs}$ is furthermore special K\"ahler,
1241: since $\cF_\Kh$ is the first derivative with respect to $Z^\Kh$ of
1242: a prepotential
1243: $\cF = \frac{1}{2} Z^\Kh \cF_\Kh$. This implies that $G_{K \bar L}$
1244: is fully encoded in the holomorphic function $\cF$.
1245:
1246: Note that $\Omega$ is only defined up to complex rescalings by a holomorphic function
1247: $e^{-h(z)}$ which via \eqref{csmetric}
1248: also changes the K\"ahler potential by a K\"ahler transformation
1249: \beq\label{crescale}
1250: \Omega\to\Omega\, e^{-h(z)}\ , \qquad \Kcs\to\Kcs + h +\bar h\ .
1251: \eeq
1252: This symmetry renders one of the periods (conventionally
1253: denoted by $Z^0$) unphysical
1254: in that one can always choose to fix a K\"ahler gauge and set $Z^0 = 1$.
1255: The complex structure
1256: deformations can thus be identified with the remaining
1257: $h^{(1,2)}$ periods $Z^K$ by defining the special coordinates
1258: $z^K = {Z^K}/{Z^0}$. A more
1259: detailed discussion of special geometry can be found in appendix \ref{specialGeom}.
1260:
1261: Let us next turn to the second factor in \eqref{geom-mod} spanned by the complexified
1262: K\"ahler deformations $t^A$. The metric on $\cM^{\rm ks}$ is given by \cite{Strominger,CdO}
1263: \bea \label{Kmetric}
1264: G_{A B} = \frac{3}{2\KK}
1265: \int_{Y}\omega_A \wedge *\omega_B = -\frac{3}{2}\left( \frac{\KK_{AB}}{\KK}-
1266: \frac{3}{2}\frac{\KK_A \KK_B}{\KK^2} \right) = \partial_{t^a} \partial_{\bar t^B} K^{\rm ks}\ ,
1267: \eea
1268: where $*$ is the six-dimensional Hodge-$*$ on $Y$ and the K\"ahler potential $K^{\rm ks}$
1269: is given by
1270: \beq \label{Kpot_ks}
1271: K^{\rm ks} = - \ln \big[\tfrac{i}{6} \cK_{ABC}(t-\bar t)^A (t-\bar t)^B (t-\bar t)^C \big] = - \ln \tfrac{4}{3} \cK\ ,
1272: \eeq
1273: where $\frac16 \cK$ is the volume of the Calabi-Yau manifold.
1274: We abbreviated the intersection numbers as follows
1275: \bea\label{int-numbers}
1276: \KK_{ABC} &=& \int_{Y}\omega_A \wedge \omega_B \wedge \omega_C\ , \qquad \qquad
1277: \KK_{AB} = \int_{Y}\omega_A \wedge \omega_B \wedge J
1278: = \KK_{ABC}v^C\ , \\
1279: \KK_{A} &=& \int_{Y}\omega_A \wedge J \wedge J
1280: =\KK_{ABC}v^Bv^C \ , \qquad
1281: \KK = \int_{Y}J \wedge J \wedge J
1282: =\KK_{ABC}v^Av^Bv^C \ ,\nonumber
1283: \eea
1284: with $J = v^A \omega_A $ being the K\"ahler form of $Y$.
1285: The manifold $\cM^{\rm ks}$ is once again special K\"ahler, since
1286: $K^{\rm ks}$ given in \eqref{Kpot_ks} can be derived from a single holomorphic
1287: function $f(t)=-\frac{1}{6} \cK_{ABC} t^A t^B t^C$ via \eqref{Kinz}.
1288:
1289:
1290:
1291:
1292: \section{Type IIA on Calabi-Yau manifolds}
1293: \label{revIIA}
1294:
1295: Let us now apply these tools in Calabi-Yau compactifications of type IIA supergravity
1296: following \cite{BCF,FS}.
1297: This theory is the maximally supersymmetric theory in ten spacetime dimensions,
1298: which posses two gravitinos of opposite chirality. It is naturally obtained as the low energy limit of
1299: type IIA superstring theory. Thus the supergravity spectrum consists of the massless string modes.
1300: The bosonic fields are the dilaton $\hat \phi$, the ten-dimensional metric $\hat g$ and the two-form
1301: $\hat B_2$ in the NS-NS sector,
1302: while the one- and three-forms $\hat C_1,\hat C_3$ arise in
1303: the R-R sector.\footnote{We use a `hat'
1304: to denote ten-dimensional quantities and omit it for
1305: four-dimensional fields.} Using form notation (our conventions are summarized in appendix~\ref{conventions})
1306: the corresponding ten-dimensional type IIA supergravity action in the
1307: Einstein frame is given by \cite{JPbook}
1308: \bea \label{10dact}
1309: S^{(10)}_{IIA} &=& \int -\tfrac{1}{2}\hat R*\mathbf{1} -\tfrac{1}{4} d\hat \phi\wedge * d\hat \phi
1310: -\tfrac{1}{4} e^{-\hat \phi}\hat H_3 \wedge *\hat H_3
1311: -\tfrac{1}{2} e^{\frac{3}{2} \hat \phi}\hat F_2 \wedge *\hat F_2 \nn \\
1312: && -\tfrac{1}{2} e^{\frac{1}{2} \hat \phi}\hat F_4 \wedge *\hat F_4
1313: -\tfrac{1}{2} \hat B_2 \wedge \hat F_4 \wedge \hat F_4\ ,
1314: \eea
1315: where the field strengths are defined as
1316: \bea \label{defHFF1}
1317: \hat H_3 = d \hat B_2\ , \quad \hat F_2 = d\hat C_1\ , \quad
1318: \hat F_4 = d\hat C_3 - \hat C_1 \wedge \hat H_3\ .
1319: \eea
1320: In order to dimensionally reduce type IIA to a four-dimensional
1321: theory, we make the product Ansatz $\Mext \times \Mint$ and perform a Kaluza-Klein
1322: reduction. Since $Y$ is a Calabi-Yau manifold it posses one covariantly constant spinor $\eta$.
1323: Decomposing the two ten-dimensional gravitinos into $\eta$ times some four-dimensional spinor
1324: leads to two gravitinos in $D=4$. Hence, compactifying type IIA supergravity on a Calabi-Yau
1325: threefold $Y$ results in an $N=2$ theory in four space-time dimensions and the zero modes of
1326: $Y$ have to assemble into massless $N=2$ multiplets.
1327: These zero modes are in one-to-one correspondence with harmonic forms
1328: on $Y$ and thus their multiplicity is counted by the dimension
1329: of the non-trivial cohomologies of the Calabi-Yau manifold.
1330: For the dimensional reduction
1331: one chooses a block diagonal Kaluza-Klein Ansatz for the
1332: ten-dimensional background metric
1333: \beq \label{lineel}
1334: ds^2\ = \ \eta_{\mu \nu}(x)\, dx^\mu dx^\nu + g_{i \jb}(y)\, dy^i dy^\jb\ ,
1335: \eeq
1336: where $\eta_{\mu \nu},\mu,\nu=0,\ldots,3$
1337: is a four-dimensional Minkowski metric and
1338: $g_{i \jb},i,\jb=1 \ldots 3$ is a Calabi-Yau metric. Part of the four dimensional fields
1339: arise as variations around this background metric. They correspond to the four-dimensional
1340: graviton and the geometric deformations $v^A(x)$ and $z^K(x)$ defined in \eqref{def-v} and \eqref{cs}.
1341: Variations of the off-diagonal entries of this metric vanish due to the fact that $Y$ does
1342: not admit harmonic one-forms. Accordingly we expand
1343: the ten-dimensional gauge-potentials introduced in \eqref{defHFF1} in
1344: terms of harmonic
1345: forms on $Y$
1346: \bea \label{fieldexp}
1347: \hat C_1 &=& A^0(x)\ ,\qquad
1348: \hat B_2\, = \, B_2(x) + b^A(x) \, \omega_A\ ,\quad
1349: A\ =\ 1,\ldots, h^{(1,1)}\ ,\\
1350: \hat C_3 &=&
1351: A^A(x) \wedge \omega_A + \,
1352: \xi^\Kh(x)\, \alpha_\Kh - \tilde \xi_\Kh(x)\, \beta^\Kh\ , \quad \Kh\ =\ 0,\ldots, h^{(2,1)}\ . \nn
1353: \eea
1354: Here $b^A,\xi^\Kh,\tilde \xi_\Kh$ are four-dimensional scalars,
1355: $A^0,A^A$ are one-forms and $B_2$ is a two-form.
1356: The ten-dimensional one-form $\hat C_1$
1357: only contains a four-dimensional one-form $A^0$ in the expansion
1358: \eqref{fieldexp} since a Calabi-Yau
1359: threefold has no harmonic one-forms.
1360:
1361:
1362: The geometric deformations $v^A,z^K$ together with the fields defined in
1363: the expansions \eqref{fieldexp} assemble into a gravity multiplet $(g_{\mu\nu},A^0)$,
1364: $h^{(1,1)}$ vector multiplets $(A^A, v^A, b^A)$,
1365: $h^{(2,1)}$
1366: hypermultiplets $(z^K,\xi^K,\tilde \xi_K)$ and one tensor multiplet
1367: $(B_2,\phi,\xi^0,\tilde \xi_0)$ where we only give the bosonic
1368: components.
1369: Dualizing the two-form $B_2$ to a scalar $a$ results in one
1370: further hypermultiplet. We summarize the bosonic spectrum in
1371: table~\ref{tab-compIIAspec}.
1372:
1373: \begin{table}[h]
1374: \begin{center}
1375: \begin{tabular}{| c | c | c |} \hline
1376: \rule[-0.3cm]{0cm}{0.8cm} gravity multiplet &
1377: $1$ & {\small $(g_{\mu \nu},A^0)$}
1378: \\ \hline
1379: \rule[-0.3cm]{0cm}{0.8cm} vector multiplets &
1380: $h^{(1,1)}$ & {\small $(A^{A}, v^A,b^A)$}\\ \hline
1381: \rule[-0.3cm]{0cm}{0.8cm} hypermultiplets &
1382: $h^{(2,1)}$ &
1383: {\small $(z^K,\xi^K,\tilde \xi_K)$}\\ \hline
1384: \rule[-0.3cm]{0cm}{0.8cm} tensor multiplet &
1385: 1 &
1386: {\small $(B_2,\phi,\xi^0,\tilde \xi_0)$} \\ \hline
1387:
1388: \end{tabular}
1389: \caption{\small \label{tab-compIIAspec}
1390: \textit{ $N=2$ multiplets for Type IIA supergravity compactified on a Calabi-Yau manifold.}}
1391: \end{center}
1392: \end{table}
1393: In order to display the low energy effective action in the standard
1394: $N=2$ form one needs to redefine the field variables slightly.
1395: One combines the real scalars $v^A, b^A$ into complex fields
1396: $t^A$ as done in \eqref{def-t_II} and defines a four-dimensional
1397: dilaton $D$ according to
1398: \beq \label{4d-dilaton}
1399: e^{D} = e^{\phi} (\cK/6)^{-\frac{1}{2}}\ ,
1400: \eeq
1401: where $\cK$ is defined in \eqref{int-numbers}. Note that $v^A$, and hence the volume $\cK/6 = \text{Vol}_S(Y)$,
1402: are evaluated in string frame. In this frame the ten-dimensional
1403: Einstein-Hilbert term takes the form $\int \frac{1}{2} e^{-2\hat \phi} R * \mathbf{1}$
1404: and $J=v^A \omega_A$ is obtained from the internal part of this string frame metric.
1405: The kinetic term for the ten-dimensional Einstein frame metric reads
1406: $\int \frac{1}{2} R * \mathbf{1}$ and hence $J$ is related to $J_E$ in the
1407: Einstein frame via $J = e^{\phi/2}J_E$.
1408: Inserting the field expansions \eqref{fieldexp} into \eqref{defHFF1}, \eqref{10dact},
1409: reducing the Riemann scalar $R$ by including the complex and K\"ahler
1410: deformations and performing a Weyl rescaling to the standard Einstein-Hilbert term,
1411: one ends up with the four-dimensional $N=2$ effective action
1412: \cite{N=2review,BCF,FS}
1413: \bea \label{IIA-4}
1414: S^{(4)}_{\rm IIA} & = &\int -\tfrac12 R * \mathbf{1}
1415: + \tfrac12 \I \cN_{\Ah \Bh}\, F^{\Ah} \wedge *
1416: F^{\Bh}
1417: + \tfrac12 \R \cN_{\Ah \Bh}\, F^{\Ah} \wedge F^{\Bh} \\
1418: & & - G_{A B}\, dt^A \wedge * d\bar t^B
1419: - h_{uv}\, d\tilde q^u \wedge * d\tilde q^v \ , \nn
1420: \eea
1421: where $F^{\Ah} = dA^{\Ah}$. The couplings of the vector multiplets in the action
1422: \eqref{IIA-4} are encoded by the metric $G_{A B}$ and the complex matrix
1423: $\cN_{\Ah \Bh}$. $G_{A B}$ only depends on the moduli $t^A$ (or rather
1424: their imaginary parts) and is defined in \eqref{Kmetric} and \eqref{Kpot_ks}.
1425: The gauge-kinetic coupling matrix $\cN_{\Ah \Bh}$ also depends on
1426: the scalars $t^A$ and is given explicitly in \eqref{def-cN}.
1427: It can be calculated from the same holomorphic prepotential like $G_{AB}$ as
1428: explained in appendix \ref{specialGeom}.
1429:
1430: Next let us turn to the couplings of the hypermultiplet sector which are encoded in the
1431: quaternionic metric $h_{uv}$. From the Kaluza-Klein reduction one obtains \cite{FS}
1432: \bea \label{q-metr}
1433: h_{uv}\, d\tilde q^u\, d\tilde q^v &=& (dD)^2 + G_{K \bar L}\, dz^K d\bar z^L
1434: +\tfrac{1}{4}e^{4D}\big(da -(\tilde \xi_\Kh d\xi^\Kh - \xi^\Kh d\tilde \xi_\Kh) \big)^2 \\
1435: && - \tfrac{1}{2} e^{2D} (\text{Im}\; \cM)^{-1\ \Kh \Lh}
1436: \big(d\tilde \xi_\Kh - \cM_{\Kh \Nh} d\xi^\Nh \big)
1437: \big(d\tilde \xi_\Lh - \bar
1438: \cM_{\Lh \Mh} d\xi^\Mh \big)\ ,\nn
1439: \eea
1440: where $G_{K \bar L}$ is the metric on the space of complex structure deformations given in
1441: \eqref{chi_barchi} and \eqref{csmetric}. The complex coupling matrix $\cM_{\Kh \Lh}$
1442: appearing in \eqref{q-metr} depends on the complex structure deformations $z^K$ and is defined as
1443: \cite{CDAF}
1444: \bea \label{defM}
1445: \int \alpha_\Kh \wedge * \alpha_\Lh&=&-(\text{Im}\; \cM +(\text{Re}\; \cM)
1446: (\text{Im}\; \cM)^{-1}(\text{Re}\; \cM))_{\Kh \Lh}\ , \nn\\
1447: \int \beta^\Kh \wedge * \beta^\Lh &=&-(\text{Im}\; \cM)^{-1\ \Kh \Lh}\ , \\
1448: \int \alpha_\Kh\wedge * \beta^\Lh &=&
1449: -((\text{Re}\; \cM)(\text{Im}\; \cM)^{-1})_{\Kh}^\Lh\ . \nn
1450: \eea
1451: It can be calculated from the periods \eqref{pre-z} by using equation \eqref{gauge-c}.
1452: Thus also in the hypermultiplet sector all couplings are determined
1453: by a holomorphic prepotential and such metrics have been called dual or special
1454: quaternionic \cite{CFGi,FS}.
1455:
1456: As we have just reviewed the $N=2$ moduli space
1457: has the local product structure
1458: \beq \label{N=2modsp}
1459: \cM^{\rm SK} \times \cM^{\rm Q}\ ,
1460: \eeq
1461: where $\cM^{\rm SK}=\cM^{\rm ks}$ is the special K\"ahler manifold spanned
1462: by the scalars in the vector multiplets or in other words
1463: the (complexified) deformations of the Calabi-Yau K\"ahler form
1464: and $\cM^{\rm Q}$ is a dual quaternionic manifold spanned by
1465: the scalars in the hypermultiplets.
1466: $\cM^{\rm Q}$ has a special K\"ahler submanifold spanned by the
1467: complex structure deformations $\cM^{\rm cs}$.
1468:
1469: This ends our short review of Calabi-Yau compactifications of type IIA
1470: supergravity. There is a second $N=2$ supersymmetric theory in
1471: ten dimensions which is the low energy effective theory of type IIB
1472: string theory. Reviewing the Calabi-Yau reduction of this theory will be
1473: the task of the next section.
1474:
1475:
1476: \section{Type IIB on Calabi-Yau manifolds}
1477: \label{revIIB}
1478:
1479: Now we turn to the review of type IIB compactifications
1480: on Calabi-Yau spaces \cite{BGHL}.
1481: Type IIB supergravity is maximal supersymmetric in ten dimensions
1482: and possesses two gravitinos of the same chirality. It
1483: consists of the same NS-NS fields as type IIA: the scalar
1484: dilaton $\hat \phi$, the metric $\hat g$ and a two-form $\hat B_2$.
1485: In the R-R sector type IIB consists of even forms,
1486: the axion $\hat C_0$, a two-form $\hat C_2$ and a
1487: four-form $\hat C_4$.
1488: The low energy effective action in the $D=10$
1489: Einstein frame is given by
1490: \cite{JPbook}
1491: \begin{eqnarray}\label{10d-lagr}
1492: S^{(10)}_{IIB}&=&
1493: \int -\tfrac{1}{2} \hat R * \mathbf{1} - \tfrac{1}{4} d\hat \phi\wedge *d \hat \phi
1494: -\tfrac{1}{4} e^{-\hat \phi} \hat H_3 \wedge* \hat H_3 \\
1495: &&- \tfrac{1}{4} e^{2\hat \phi} \hat F_1 \wedge * \hat F_1 -
1496: \tfrac{1}{4} e^{\hat \phi} \hat F_3 \wedge * \hat F_3 -
1497: - \tfrac{1}{8}\hat F_{5} \wedge *\hat F_{5}
1498: -\tfrac{1}{4} \hat C_4 \wedge \hat H_3 \wedge \hat F_3\ ,
1499: \nonumber
1500: \end{eqnarray}
1501: with the field strengths defined as
1502: \begin{eqnarray}
1503: \hat H_3 \ =\ d \hat B_2\ , \quad \hat F_1 = d\hat C_0\ ,\quad
1504: \hat F_{q+1}\ =\ d \hat C_q - \hat C_{q-2} \wedge \hat H_3\ ,\quad q=2,4\ . \label{fieldstr}
1505: \end{eqnarray}
1506: The self-duality condition $\hat F_5=*\hat F_5$ is
1507: imposed at the level of the equations of motion.
1508:
1509: As in the type IIA compactifications discussed in the previous section
1510: we use the Ansatz \eqref{lineel}
1511: for the ten-dimensional background metric. Fluctuations around this background
1512: metric are parameterized by the four-dimensional graviton $g_{\mu \nu}$ and
1513: the geometric deformations of the Calabi-Yau metric. More precisely, we find
1514: $h^{(1,1)}$ real K\"ahler structure deformations $v^A$
1515: introduced in \eqref{def-v} and $h^{(2,1)}$ complex structure deformations $z^K$
1516: introduced in \eqref{cs}.
1517: The type IIB gauge potentials appearing in the Lagrangian
1518: \eqref{10d-lagr} are similarly
1519: expanded in terms of harmonic forms on $Y$ according to
1520: \begin{eqnarray}\label{CYexpansion}
1521: \hat B_2 &=& B_2(x) + b^A(x)\, \omega_A\ , \qquad
1522: \hat C_2\ =\ C_2(x) + c^A (x)\,\omega_A\ , \quad A=1,\ldots,h^{(1,1)}\ , \\
1523: \hat C_4 &=& D_2^A(x) \wedge \omega_A + V^{\hat K}(x) \wedge
1524: \alpha_{\hat K} - U_{\hat K}(x) \wedge \beta^{\hat K} +
1525: \rho_A(x)\, \tilde \omega^A\ ,
1526: \quad \hat K=0,\ldots,h^{(1,2)}.\nonumber
1527: \label{full-exp}
1528: \end{eqnarray}
1529: The four-dimensional fields appearing in the expansion \eqref{CYexpansion}
1530: are the scalars $b^A(x)$, $c^A(x)$ and $\rho_A(x)$,
1531: the one-forms $V^{\hat K}(x)$ and
1532: $U_{\hat K}(x)$ and the two-forms $B_2(x),C_2(x)$ and $D_2^A(x)$.
1533: The self-duality condition of $\hat F_5$ eliminates half of the
1534: degrees of freedom in $\hat C_4$ and in this section we choose to eliminate
1535: $D^A_2$ and $U_{\hat K}$ in favor of $\rho_A$ and $V^{\hat K}$.
1536: Finally, the two type IIB scalars $\hat \phi, \hat C_0$ also appear as
1537: scalars in $D=4$ and therefore we drop the hats henceforth
1538: and denote them by $\phi, C_0$.
1539:
1540: In summary the massless $D=4$ spectrum consists of
1541: the gravity multiplet with bosonic components $(g_{\mu \nu}, V^0)$,
1542: $h^{(2,1)}$ vector multiplets with bosonic components $(V^{K}, z^{K})$,
1543: $h^{(1,1)}$ hypermultiplets with bosonic components
1544: $(v^A, b^A, c^A, \rho_A)$
1545: and one double-tensor multiplet \cite{BVT} with bosonic components
1546: $(B_2, C_2, \phi, C_0)$ which can be dualized to an additional (universal)
1547: hypermultiplet. The four-dimensional spectrum is
1548: summarized in Table \ref{tab-compIIBspec}.
1549:
1550: \begin{table}[h]
1551: \begin{center}
1552: \begin{tabular}{| c | c | c |} \hline
1553: \rule[-0.3cm]{0cm}{0.8cm} gravity multiplet &
1554: $1$ & {\small $(g_{\mu \nu},V^0)$}
1555: \\ \hline
1556: \rule[-0.3cm]{0cm}{0.8cm} vector multiplets &
1557: $h^{(2,1)}$ & {\small $(V^{K}, z^{K})$}\\ \hline
1558: \rule[-0.3cm]{0cm}{0.8cm} hypermultiplets &
1559: $h^{(1,1)}$ &
1560: {\small $(v^A, b^A, c^A, \rho_A)$
1561: }\\ \hline
1562: \rule[-0.3cm]{0cm}{0.8cm} double-tensor multiplet &
1563: 1 &
1564: {\small $(B_2, C_2,\phi,C_0)$
1565: }\\ \hline
1566:
1567: \end{tabular}
1568: \caption{\small \label{tab-compIIBspec}
1569: \textit{ $N=2$ multiplets for Type IIB supergravity compactified on a Calabi-Yau manifold.}}
1570: \end{center}
1571: \end{table}
1572:
1573: The $N=2$ low energy effective action is computed by inserting
1574: \eqref{fieldstr} and \eqref{CYexpansion} into the action \eqref{10d-lagr}
1575: and integrating over the Calabi-Yau manifold.
1576: For the details we refer the reader to the literature
1577: \cite{BGHL,Michelson,DallAgata,LM} and only recall the results here.
1578: One finds
1579: \begin{eqnarray}\label{N=2}
1580: S_{IIB}^{(4)} &=& \int - \tfrac{1}{2} R *\! {\bf 1}
1581: + \tfrac{1}{4} \RE\cM_{\hat K \hat L} {F}^{\hat K} \wedge {F}^{\hat L} +
1582: \tfrac{1}{4} \IM_{\hat K\hat L} {F}^{\hat K} \wedge * {F}^{\hat L}\nonumber\\
1583: &&- G_{K L} d z^K \wedge * d \bar{z}^{L}
1584: - G_{AB} d t^A \wedge * d \bar t^B
1585: - d D \wedge * d D - \tfrac{1}{24} e^{2 D} \cK \dd l \wedge * \dd l \nonumber\\
1586: &&
1587: - \tfrac{1}{6} e^{2D} \cK G_{AB}
1588: \big(\dd c^A - l \dd b^A \big)\wedge * \big( \dd c^B - l \dd b^B
1589: \big)\\
1590: && - \tfrac{3}{8\cK} e^{2D} G^{AD} \big( \dd \rho_A -
1591: \cK_{ABC} c^B \dd b^C \big) \wedge\! *\big( \dd \rho_D -
1592: \cK_{DEF} c^E \dd b^F \big) \nonumber \\
1593: && -\tfrac{1}{4}e^{-4D} \dd B_2 \wedge * \dd B_2 - \tfrac{1}{24}
1594: e^{-2D} \cK
1595: \big( \dd C_2 - l \dd B_2 \big) \wedge *\big( \dd C_2 - l \dd B_2 \big)
1596: \nn \\
1597: && - \tfrac{1}{2} dC_2 \wedge \big( \rho_A \dd b^A - b^A d\rho_A \big)
1598: +\tfrac{1}{2} dB_2 \wedge c^A d \rho_A - \tfrac{1}{4}\cK_{ABC} c^A c^B dB_2 \wedge \dd b^C \nonumber \ ,
1599: \end{eqnarray}
1600: where $F^{\hat K}=dV^{\hat K}$.
1601: The gauge kinetic matrix $\cM_{\hat K\hat L}$ is related to the metric
1602: on $H^3(Y)$ and given in \eqref{defM}. The metric $G_{K L}(z,\bar z)$ which appears in \eqref{N=2}
1603: is the metric on the space of complex structure deformations given in \eqref{csmetric}.
1604: It is a special K\"ahler metric in that it is entirely determined by the holomorphic prepotential $\mathcal{F}(z)$
1605: \cite{Strominger2,CdO}. On the other hand, the metric $G_{AB}$ in \eqref{N=2} is the metric
1606: on the space of K\"ahler deformations defined in \eqref{Kmetric}.
1607:
1608: In order to entirely express \eqref{N=2} in terms
1609: of vector- and hypermultiplets we dualize the
1610: $D=4$ two-forms $B_2,C_2$ to scalar fields. This can be done, since $B_2$ and $C_2$ are massless
1611: and posses the gauge symmetries $C_2 \rightarrow C_2 + d\Lambda_1$ and $B_2 \rightarrow B_2 + d\tilde \Lambda_1$.
1612: Let us first dualize $C_2$.
1613: We replace $dC_2$ with $D_3$ and add the Lagrange multiplier $\frac12 h\, dD_3$ such
1614: that the differentiation with respect to $h$ yields $dD_3=0$. Locally this ensures
1615: that $D_3=dC_2$ for some two-form $C_2$. The terms in \eqref{N=2} involving $D_3$ are simply
1616: \beq
1617: \cL_{C_2} = - \tfrac{g}{4}
1618: \big( D_3 - C_0\, \dd B_2 \big) \wedge * \big( D_3 - C_0\, \dd B_2 \big)
1619: - \tfrac{1}{4} D_3 \wedge J_1 +\tfrac12 D_3 \wedge dh\ ,
1620: \eeq
1621: where we abbreviated $g = \frac{1}{6} e^{-2D} \cK$ and $J_1 = \rho_A \dd b^A - b^A d\rho_A$.
1622: Now we can consistently eliminate $D_3$ in favor of $h$ by its equation of motion. The
1623: dualized Lagrangian takes the form
1624: \beq \label{dual_h}
1625: \cL_{h} = - \tfrac{1}{4 g} \big(dh - \tfrac12 J_1 \big) \wedge *
1626: \big(dh - \tfrac12 J_1 \big)
1627: + \tfrac12 C_0\, dB_2 \wedge \big(dh - \tfrac12 J_1 \big)\ .
1628: \eeq
1629: Similarly we can dualize the two-from $B_2$ into a scalar $\tilde h$. Having replaced
1630: $C_2,B_2$ by $h,\tilde h$ in \eqref{N=2} the effective action can be written in the standard
1631: $N=2$ form \cite{BaggerW,dWvP,N=2review}
1632: \begin{eqnarray}
1633: S_{IIB}^{(4)} & = & \int -\tfrac{1}{2}R *{\mathbf 1}
1634: + \tfrac{1}{4} \RE\cM_{\Kh\Lh} {F}^\Kh \wedge {F}^\Lh + \tfrac{1}{4} \I
1635: \cM_{\Kh\Lh} {F}^\Kh \wedge * {F}^\Lh\nonumber\\
1636: &&\qquad - G_{KL} \dd z^K \wedge *\dd \bar{z}^{L}
1637: - h_{pq}\, \dd \tilde q^{p} \wedge * \dd \tilde q^{q} \ ,
1638: \label{action3}
1639: \end{eqnarray}
1640: where $q^{p}$ denote the coordinates for all
1641: $h^{(1,1)}+1$ hypermultiplets. The metric $h_{pq}$ is a quaternionic
1642: metric explicitly given by \cite{FS}
1643: \begin{align} \label{q-metrB}
1644: h_{pq}\, d\tilde q^p\, d\tilde q^q &= (d D)^2 + G_{AB} d t^A d\bar t^B
1645: + \tfrac{1}{24} e^{2 D} \cK (\dd C_0)^2 \nonumber\\
1646: & + \tfrac{1}{6} e^{2D} \cK G_{AB} \big(\dd c^A - C_0\, \dd b^A \big)\big( \dd c^B - C_0\, \dd b^B \big) \\
1647: & + \tfrac{3}{8\cK} e^{2D} G^{AD} \big( \dd \rho_A -
1648: \cK_{ABC} c^B \dd b^C \big) \big( \dd \rho_D -
1649: \cK_{DEF} c^E \dd b^F \big) \nonumber \\
1650: & + \tfrac{3}{2 \cK} e^{2D}\big(dh - \tfrac12 (\rho_A \dd b^A - b^A d\rho_A) \big)^2 \nn\\
1651: & + \tfrac12 e^{4D} \big(d\tilde h + C_0\, dh + c^A d\rho_A + \tfrac12 C_0\, (\rho_A \dd b^A - b^A d\rho_A)
1652: - \tfrac{1}{4}\cK_{ABC} c^A c^B \dd b^C \big)^2 . \nn
1653: \end{align}
1654: In summary the scalar moduli space $\cM^{\rm SK} \times \cM^{\rm Q}$ of the $N=2$ theory is
1655: the product of a quaternionic manifold $\cM^{\rm Q}$ spanned by the scalars
1656: $q^{p}$ with metric \eqref{q-metrB} and a special
1657: K\"ahler manifold $\cM^{\rm SK} = \cM^{\rm cs}$ spanned by the scalars $z^K$.
1658: The complexified K\"ahler structure deformations span a special K\"ahler manifold $\cM^{\rm ks}$
1659: inside $\cM^{\rm Q}$. In \cite{FS} it was shown that the quaternionic space
1660: can be constructed from the prepotential of $\cM^{\rm ks}$ such that
1661: $\cM^{\rm Q}$ is a special quaternionic manifold.
1662:
1663: This ends our brief summary of type IIB compactified on
1664: Calabi-Yau threefolds and its $N=2$ low energy effective action.
1665: We have seen that the effective actions of the type IIA and type
1666: IIB indeed take the standard $N=2$ form. In both cases the metrics
1667: on the special K\"ahler and special quaternionic manifold are encoded by
1668: a corresponding prepotential. However, the role of the K\"ahler and complex
1669: structure deformations is exchanged in type IIA and type IIB compactifications.
1670: As we will discuss momentarily this can be traced back to an underlying symmetry
1671: which finally enables us to identify both effective theories in the
1672: large volume -- large complex structure limit.
1673:
1674:
1675: \section{N=2 Mirror symmetry}
1676: \label{revMirror}
1677:
1678: In this section we briefly discuss mirror symmetry for Calabi-Yau
1679: compactifications \cite{Mirror}. From a mathematical point of view, mirror
1680: symmetry is a duality in the moduli space of Calabi-Yau manifolds.
1681: It states that for a given Calabi-Yau manifold $Y$, there exists
1682: a mirror Calabi-Yau $\tilde Y$ such that
1683: \beq \label{Hod_id}
1684: h^{(1,1)}(Y) = h^{(2,1)}(\tilde Y)\ , \qquad h^{(2,1)}(Y) = h^{(1,1)}(\tilde Y)\ .
1685: \eeq
1686: Applied to the Hodge diamond \eqref{hodge_diamond} this amounts to a reflection along the
1687: diagonal. In other words, mirror symmetry identifies the odd and even cohomologies \eqref{odd_even_cohom}
1688: of two topological distinct Calabi-Yau spaces
1689: \beq \label{cohom_id}
1690: H^{ev}(Y)\ \cong \ H^{odd}(\tilde Y)\ ,\qquad H^{odd}( Y)\ \cong \ H^{ev}(\tilde Y)\ .
1691: \eeq
1692: Moreover, it is much stronger than that, since
1693: it also implies an identification of the moduli spaces of deformations of $Y$ and $\tilde Y$.
1694: As given in \eqref{geom-mod} the geometric moduli space of a Calabi-Yau manifold
1695: is a local product of two special K\"ahler manifolds $\cM^{\rm ks}$ and $\cM^{\rm cs}$.
1696: Their complex dimensions are exactly given by $h^{(1,1)}$ and $h^{(2,1)}$. Motivated
1697: by \eqref{Hod_id} mirror symmetry conjectures the identifications
1698: \beq \label{Modspace_id}
1699: \cM^{\rm ks}(Y)\ \equiv\ \cM^{\rm cs}(\tilde Y)\ , \qquad
1700: \cM^{\rm cs}(Y)\ \equiv\ \cM^{\rm ks}(\tilde Y)\ ,
1701: \eeq
1702: as special K\"ahler manifolds. Recall that the geometry of $\cM^{\rm cs}(Y)$
1703: and $\cM^{\rm cs}(\tilde Y)$ are encoded in the variations of the holomorphic
1704: three-forms $\Omega$ and $\tilde \Omega$ of the two Calabi-Yau manifolds
1705: $Y$ and $\tilde Y$. These can be expanded in a real symplectic basis of $H^{3}(Y)$ and
1706: $H^{3}(\tilde Y)$ respectively
1707: \beq
1708: \Omega(z) = Z^\Kh \alpha_\Kh - \cF_\Kh \beta^\Kh\ , \qquad
1709: \tilde \Omega(\tilde z) =\tilde Z^\Ah \alpha_\Ah - \tilde \cF_\Ah \beta^\Ah\ ,
1710: \eeq
1711: Under the large volume mirror map
1712: the coordinates on the two manifolds $\cM^{\rm ks}(Y)$ and $\cM^{\rm cs}(\tilde Y)$ as
1713: well as $\cM^{\rm cs}(Y)$ and $\cM^{\rm ks}(\tilde Y)$
1714: are identified as
1715: \beq \label{mirror-map}
1716: t^A = {\tilde Z^A(\tilde z)}/{\tilde Z^0(\tilde z)} \ , \qquad {Z^K(z)}/{Z^0(z)} = \tilde t^K
1717: \eeq
1718: where $t^A$ and $\tilde t^K$ are the complexified K\"ahler deformations of $Y$ and $\tilde Y$.
1719: Equation \eqref{mirror-map} implies that $t^A,\tilde t^K$ are identified with special coordinates
1720: on $\cM^{\rm cs}$. Furthermore, recall that due to the special K\"ahler property the metric on both
1721: moduli spaces is encoded by a prepotential. Applying \eqref{Modspace_id} it follows that
1722: these prepotentials $f_Y(t)$ and $f_{\tilde Y}(\tilde z)$ as well as
1723: $f_Y(z)$ and $f_{\tilde Y}(\tilde t)$ are identified under mirror symmetry.
1724: One immediately notices, that this can not be the full truth, since $\cM^{\rm ks}$ and $\cM^{\rm cs}$ have
1725: a different structure. $\cM^{\rm ks}$ is a cone and admits the
1726: simple prepotential $f(t) = - \frac16 \cK_{ABC} t^A t^B t^C$, while the metric
1727: on $\cM^{\rm cs}$ is determined in terms of the (generically complicated)
1728: periods of the holomorphic three-form. Hence, one expects corrections to
1729: $f_Y(t)$ and $f_{\tilde Y}(\tilde t)$. These corrections get a physical interpretation
1730: as soon as mirror symmetry is embedded in string theory.
1731: They are due to strings wrapping two-cycles in $Y$ called world-sheet instantons.
1732: Schematically one identifies
1733: \beq
1734: f_{Y}(t) = t^3 + \mathcal{O}(e^{-t})=f_{\tilde Y}(\tilde z)\ ,\qquad f_Y(z)
1735: = \tilde t^3 + \mathcal{O}(e^{-\tilde t})=f_{\tilde Y}(\tilde t)\ .
1736: \eeq
1737: One can also turn the argument around and use mirror symmetry as a very powerful tool to calculate the world-sheet
1738: instanton corrections $\mathcal{O}(e^{-t})$ as done in the pioneering paper \cite{CdOGP}.
1739: In most cases this is much simpler then a direct calculation of the world-sheet instanton
1740: contributions.
1741:
1742: The most prominent applications of mirror symmetry in string theory
1743: is the identification of type IIA string theory compactified on $Y$ with
1744: type IIB string theory compactified on $\tilde Y$. It matches
1745: the full string theories including their low energy limits
1746: and supersymmetric D-brane states. With the material presented in this chapter we can check
1747: it on the level of the effective action by comparing \eqref{IIA-4} with
1748: \eqref{action3}. This amounts to matching the moduli spaces of the
1749: corresponding four-dimensional $N=2$ theories which take the standard $N=2$
1750: form \eqref{N=2modsp}. Since we already discussed the special K\"ahler part in \eqref{N=2modsp},
1751: let us now concentrate on the quaternionic manifolds $\cM^{\rm Q}(Y)$
1752: and $\cM^{\rm Q}(\tilde Y)$. In accordance with \eqref{cohom_id} and \eqref{mirror-map}
1753: one identifies the basis elements $(1, \omega_K,\tilde \omega^K, \vol(Y))$ of $H^{ev}(Y)$ with
1754: the basis $(\alpha_\Kh,\beta^\Kh)$ of $H^{odd}(\tilde Y)$ as
1755: \beq \label{basis_id}
1756: 1\leftrightarrow \alpha_0\ ,\quad \omega_K \leftrightarrow \alpha_K\ ,\quad \vol(Y) \leftrightarrow \beta^0\ ,
1757: \quad \tilde \omega^K \leftrightarrow \beta^K\ .
1758: \eeq
1759: We will work in this basis in the following.
1760: Let us now construct the explicit map for the quaternionic coordinates by using
1761: a slightly non-standard argument. We intend to apply the fact, that the fields of
1762: the quaternionic space describe the coupling to D-branes, which are extended
1763: non-perturbative objects present in both type II string theories. We will discuss the
1764: low energy dynamics and supersymmetry conditions of these objects more carefully in section
1765: \ref{D-branes}. All we need for constructing the mirror map for the quaternionic spaces
1766: is there coupling to the R-R forms in the supergravity theory and some information
1767: about supersymmetric branes in type IIA and type IIB string theory.
1768: It will become clear in section \ref{D-branes}, that the only supersymmetric Euclidean
1769: D-branes wrapping a cycle in a Calabi-Yau manifold are $D2$
1770: branes in Type IIA and $D(-1)$, $D1$, $D3$ and $D5$ branes in type IIB.
1771: The Chern-Simons action describes the coupling of the brane world-volume to the forms
1772: \beq \label{CS_coupling}
1773: \text{IIA:}\quad (\sum_{p\ even} \hat C_{p} \wedge e^{-\hat B_2})_3\ , \qquad
1774: \text{IIB:}\quad (\sum_{p\ odd} \hat C_{p} \wedge e^{-\hat B_2})_q\ ,\
1775: q = 0,2,4,6\ ,
1776: \eeq
1777: where $\hat C_p$ and $\hat B_2$ are the ten-dimensional R-R and NS-NS forms introduced in
1778: section \ref{revIIA} and \ref{revIIB}. By $(\ldots)_q$ we indicate that we
1779: only consider the $q-$form appearing in the sum of forms inside the parenthesis.
1780: Supersymmetry implies that the Euclidean D-branes, wrap cycles which are dual
1781: to harmonic forms. But the only odd harmonic forms are $(\alpha_{\hat K},\beta^{\hat K})$, while the even
1782: harmonic forms are
1783: $(1,\omega_K, \tilde \omega^K,\vol(Y))$.
1784: Next we match the Chern-Simons couplings \eqref{CS_coupling} for IIA and IIB Euclidean
1785: D-branes. We decompose \eqref{CS_coupling} on the respective cohomology basis elements
1786: by using the expansions \eqref{CYexpansion} of $\hat B_2$ and the R-R forms $\hat C_0,\hat C_2,\hat C_4$
1787: as well as the expansion \eqref{fieldexp} of $\hat C_3$.
1788: Applying the identification \eqref{basis_id} of the basis forms we find for the coefficients
1789: of $\alpha_\Kh$ and $(1,\omega_K)$ that
1790: \beq
1791: \xi^0 = C_0\ ,\qquad \xi^K = c^K - C_0\ b^K\ .
1792: \eeq
1793: Identifying the coefficients of $\beta^\Kh$ and $(\tilde \omega^K, \vol(Y))$
1794: yields higher powers in $\hat B_2$ and we find \footnote{We have replaced $\int C_6$ by
1795: $h + \tfrac12 \rho_A b^A$.
1796: This can be done since $C_6$ is dual to $C_2$, which was dualized to $h$ in \eqref{dual_h}.}
1797: \bea
1798: \tilde \xi_K &=& \rho_K - \cK_{KLM} c^L b^M + \tfrac{1}{2} C_0\ \cK_{KLM} b^L b^M\ , \\
1799: \tilde \xi_0 &=& h - \tfrac{1}{2} \rho_K b^K + \tfrac{1}{2} \cK_{KLM} c^K b^L b^M
1800: - \tfrac{1}{6}C_0\ \cK_{KLM} b^K b^L b^M\ .\nn
1801: \eea
1802: It remains to identify the space-time two-forms from the
1803: NS-NS sectors. Since $B_2^A$ and $B^B_2$ are the only remaining two-forms in the spectrum, we
1804: are forced to set $B_2^A = B^B_2$. Dualized into scalars this amounts to
1805: \beq
1806: a = 2\tilde h + C_0\, h + \rho_K(c^K - C_0\, b^K)
1807: \eeq
1808: Thus, by using the explicit form of the Chern-Simons coupling to D-branes,
1809: one can infer the mirror map for the coordinates on the quaternionic space.
1810: Of course, that the established map indeed transforms $h^A_{uv}$ given in \eqref{q-metr}
1811: into $h^B_{uv}$ given in \eqref{q-metrB} can be checked by direct calculation as done in \cite{BGHL}.
1812:
1813: This ends our review section on Calabi-Yau compactifications of type IIA
1814: and type IIB supergravity. We now turn to their orientifold versions which
1815: break $N=2$ to $N=1$. The aim of the next chapter is
1816: to determine the characteristic data of the resulting supergravity theory.
1817:
1818:
1819:
1820:
1821: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1822: %
1823: % Chapter: Effective actions Orientifolds
1824: %
1825: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1826:
1827:
1828:
1829: \chapter{Effective actions of Type II Calabi-Yau orientifolds}
1830: \label{effective_actO}
1831:
1832: In this chapter we discuss the four-dimensional low energy effective supergravity theory
1833: obtained by compactifying type IIA and type IIB string theory on Calabi-Yau orientifolds.
1834: Before entering the calculations we review some aspects of D-branes and orientifolds
1835: in section \ref{D-branes}. In particular, we introduce the low energy effective action
1836: for D-branes. Later on this will allow us to comment on corrections due to wrapped
1837: Euclidean D-branes to the bulk supergravity theory.
1838: %in sections \ref{IIB_orientifolds}, \ref{IIA_orientifolds} and \ref{non-pert_sup}.
1839: As we already explained in section \ref{braneworlds} the inclusion of
1840: space-time filling D-branes is essential for consistency. However, we freeze their moduli
1841: and matter fields such that they do not appear in the low energy effective action.\footnote{%
1842: This restriction was weakened e.g.~in \cite{GGJL,JL}, where the coupling to $D3$- and $D7$-bane moduli
1843: was determined by using the low energy effective action of the $D$ branes.}
1844: In a next step we turn to the main issue of this chapter and perform
1845: a Kaluza-Klein reduction by implementing the orientifold conditions and extract the
1846: resulting $N=1$ supergravity theory (sections \ref{oprojections} -- \ref{IIA_orientifolds}).
1847: Specifically we determine the K\"ahler potential and the gauge-kinetic
1848: coupling functions encoding the low energy effective theory. We end our analysis by checking
1849: mirror symmetry in the large complex structure and large volume limit in section \ref{Mirror_orientioflds}.
1850: A derivation of a flux induced superpotential and possible gaugings will be presented in
1851: chapter \ref{fluxesAB}.
1852:
1853:
1854: \section{D-branes and orientifolds}
1855: \label{D-branes}
1856:
1857: In this section we provide more details on D-branes and orientifolds
1858: as used in the construction of brane-world scenarios.
1859: As already mentioned in section \ref{braneworlds}
1860: brane world scenarios are currently one of the promising approaches
1861: to construct phenomenologically interesting models from
1862: string compactification \cite{reviewPP}. They consist of space-time filling
1863: D-branes serving as source for Abelian or non-Abelian gauge theories.
1864: String theory implies a low energy effective action for this gauge theory
1865: as well as the couplings to the bulk fields introduced in chapter \ref{TypeII}.
1866: More precisely, the gauge theory and the coupling to the NS-NS fields $\hat \phi$, $\hat g$ and $\hat B_2$
1867: is captured by the Dirac-Born-Infeld action. The most simple example is provided by a single
1868: $Dp$-brane, which admits an $U(1)$ gauge theory on its $p+1$-dimensional world-volume. The corresponding
1869: bosonic part of the Dirac-Born-Infeld action reads in string frame \cite{D-branes,JPbook}
1870: \begin{equation} \label{DBI}
1871: S_{\text{DBI}}^{\text{sf}}=-T_p\int_\WV d^{p+1}\xi\:e^{-\hat \phi}
1872: \sqrt{-\det\left(\Em^*(\hat g+\hat B_2)_{\hat \mu \hat \nu}+2\pi\alpha' \FD_{\hat \mu \hat \nu}\right)}\ ,
1873: \end{equation}
1874: where $T_p$ denotes the brane tension. The integral is taken over the $p+1$-dimensional
1875: world-volume $\WV$ of the $Dp$-brane, which is embedded in the ten-dimensional space-time
1876: manifold $\cM_{10}$ via the map $\Em:\WV\hookrightarrow \cM_{10}$.
1877: The Dirac-Born-Infeld action \eqref{DBI} contains an $U(1)$ field strength
1878: $\FD_{\hat \mu \hat \nu} = 2\partial_{[\hat \mu}A_{ \hat \nu]}$, which describes the
1879: $U(1)$ gauge theory to all orders in $\alpha' \FD$ \cite{Leigh:jq}.
1880: To leading order, the gauge theory reduces to an $U(1)$ gauge theory on the world-volume
1881: $\WV$ of the brane. The dynamics of the $Dp$-brane is encoded in the embedding map $\Em$. Fluctuations
1882: around a given $\varphi$ are parameterized by charged scalar fields, which provide the matter
1883: content of the low-energy effective theory.
1884:
1885: Since $Dp$-branes also carry R-R charges \cite{JP}, they couple as extended objects to appropriate
1886: R-R forms of the bulk, namely the $p+1$-dimensional world-volume couples naturally to the R-R form $\hat C_{p+1}$.
1887: Moreover, generically $D$-branes contain lower dimensional $D$-brane charges, and hence interact also with
1888: lower degree R-R forms \cite{Douglas:1995bn}.
1889: All these couplings to the bulk are implemented in the Chern-Simons action
1890: \beq \label{CSaction}
1891: S_{\text{CS}}=\mu_p\int_\WV\Em^*\Big(\sum_q \hat C_{q} \wedge e^{-\hat B_2}\Big) \wedge e^{2\pi\alpha'\FD}\ ,
1892: \eeq
1893: where $\mu_p$ is the charge of the D-branes.
1894: The lowest order terms in \eqref{CSaction} in the R-R fields are topological and represent the
1895: R-R tadpole contributions to the low energy effective action. Additionally,
1896: \eqref{CSaction} encodes the coupling of the gauged matter fields arising from
1897: perturbations of $\varphi$ to the R-R forms.
1898: The effective actions \eqref{DBI} and \eqref{CSaction} can be generalized to
1899: stacks of D-branes \cite{Myers}. This gives rise to non-Abelian gauge theories
1900: and appropriate (intersecting) embeddings can yield Standard Model like gauge theories \cite{reviewPP}.
1901:
1902: Generic brane world scenarios lead to non-supersymmetric low energy theories, which
1903: are plagued by various instabilities due to runaway potentials for the bulk moduli.
1904: In contrast, supersymmetric setups are under much better control and are therefore phenomenologically
1905: favored.
1906: However, the aim to preserve some supersymmetry poses strong conditions on the D-branes present
1907: in the setup. D-branes which preserve half of the original supersymmetries
1908: are called BPS branes and the corresponding supersymmetry conditions BPS conditions.
1909: In brane-world setups with a ten-dimensional background space-time
1910: of the form $\Mext \times \Mint$ two types of branes will be of importance which
1911: preserve four dimensional Poncar\'e invariance.
1912: Firstly, one includes D-branes filling the space-time $\Mext$ and wrapping a cycle in
1913: the manifold $\Mint$. These provide the gauge theory and matter fields just discussed.
1914: Secondly, one might add Euclidean D-branes (called D-instantons) solely wrapping
1915: a cycle in $\Mint$. They induce corrections to the supergravity theory and
1916: their effects can be useful to stabilize bulk moduli.
1917: The BPS conditions for both types of
1918: branes demand that the brane tensions $T_p$ and charges $\mu_p$ are
1919: equal. This ensures stability since the net force between BPS branes
1920: vanishes \cite{JP}. Moreover, there are conditions
1921: on the cycles $\Lambda_{Dp}$ in $\Mint$ wrapped by the branes.
1922: In \cite{BBS} it was shown that in a purely metric background with
1923: $\Mint$ being a Calabi-Yau manifold the only allowed cycles are
1924: special Lagrangian submanifolds of $\Mint$ in
1925: type IIA and holomorphic submanifolds in type
1926: IIB. More precisely special Lagrangian submanifolds are
1927: three-cycles $\Lambda^{(3)}$ in $Y$ for which
1928: \beq \label{spLagr-C}
1929: \vol(\Lambda^{(3)})=\tilde \varphi^*( \R \Omega)\ ,\qquad \tilde \varphi^*( \I \Omega) = 0\ , \qquad \tilde \varphi^* J = 0\ ,
1930: \eeq
1931: where $\vol(\Lambda^{(3)})=\det^{1/2}(\tilde \varphi^* g)\, d^3\lambda$
1932: is the volume form on $\Lambda^{(3)}$, $J$ and $\Omega$ are the
1933: K\"ahler form and holomorphic three-form of $Y$ as in chapter \ref{TypeII}
1934: and $\tilde \varphi$ defines the embedding of the D-brane into $Y$.
1935: On the other hand, holomorphic submanifolds are even-dimensional cycles $\Lambda^{(2)},\Lambda^{(4)}$ in
1936: $Y$ satisfying
1937: \beq \label{holom-C}
1938: \vol(\Lambda^{(2)})=\tilde \varphi^*(J)\ ,\qquad \vol(\Lambda^{(4)})=\tfrac{1}{2} \tilde \varphi^*(J \wedge J)\ ,
1939: \qquad \varphi^*(\Omega) = 0\ .
1940: \eeq
1941: It can be shown that the conditions \eqref{spLagr-C} and \eqref{holom-C} ensure that such cycles
1942: minimizes their volume in their homology classes (see e.g. \cite{BBS}).
1943:
1944: These conditions have to be adjusted as soon as one allows a non-trivial
1945: background of supergravity forms \cite{MMMS,CU}. As an example, the
1946: BPS conditions on the volume of the cycles in the presence of a non-trivial $\hat B_2$ field are given by \cite{MMMS}
1947: \bea \label{calcond}
1948: \text{IIA:} \qquad \vol_{DBI}(\Lambda^{(3)}_{Dp}) &=& e^{-i\theta_{Dp}}\ \tilde \varphi^* \big( \Omega \big)\ ,\\
1949: \text{IIB:} \qquad \vol_{DBI}(\Lambda^{(q)}_{Dp}) &=&
1950: e^{-i\theta_{Dp}}\ \tilde\varphi^* \big(e^{-\hat B_2 + i J} \big)_q\ , \quad q=2,4,6\ , \nn
1951: \eea
1952: where $\vol_{DBI}(\Lambda^{(q)}_{Dp}) = \det^{1/2}(\tilde \varphi^*[g + \hat B_2])\, d^q\lambda$ is the Dirac-Born-Infeld
1953: volume form on $\Lambda^{(q)}_{Dp}$.
1954: $e^{i\theta_{Dp}}$ denotes a constant phase which will be determined below.
1955: The BPS conditions involving the volume elements split into real and imaginary
1956: parts, where the imaginary part has to vanish on $\Lambda^{(q)}_{Dp}$ by using reality
1957: of $\vol_{DBI}(\Lambda^{(q)}_{Dp})$. The cycles $\Lambda^{(q)}_{Dp}$ satisfying the conditions
1958: \eqref{calcond} are called calibrated with respect to the form $e^{-i\theta_{Dp}}\,\Omega$ in type IIA
1959: and calibrated with respect to $e^{-i\theta_{Dp}}\, e^{-\hat B_2 + i J}$ in type IIB.
1960: In a setup with several D-branes some supersymmetry is preserved as
1961: soon as all D-branes are calibrated with respect to the same form.
1962: However, as we already explained in section \ref{braneworlds} this
1963: is not the end of the story, since consisted supersymmetric theories
1964: have to include negative tension objects such as orientifold planes \cite{GKP}.
1965:
1966: Similar to D-branes, orientifold planes are hyper-planes
1967: of the ten-dimensional background. They arise in string theories
1968: which contain non-orientable world-sheets. Orientifold theories can
1969: be constructed by starting from a closed string theory such as type
1970: IIA or type IIB strings and dividing out a symmetry group \cite{AD,Ori} \footnote{%
1971: As usual, dividing out a symmetry can be understood as a gauge fixing.}
1972: \beq \label{osym}
1973: G \cup S\Omega_p,
1974: \eeq
1975: where $G$ is a group of target space symmetries and $\Omega_p$ is the
1976: world-sheet parity, exchanging left and right movers. $S$
1977: contains operations, which render $S\Omega_p$ to be a
1978: symmetry of the string theory. For orientifolds \eqref{osym} consists
1979: of evidently perturbative symmetries of the string theory, which can be imposed
1980: order by order in perturbation theory and are believed to be unbroken also
1981: non-perturbatively. Specifically this implies that the orientifold projection
1982: can be consistently imposed in a low energy description.
1983: The orientifold planes are the hyper-surfaces left invariant by $S$.
1984: They naturally couple to the R-R forms and thus carry a charge. Moreover,
1985: they can have negative tension.\footnote{%
1986: Note that orientifold
1987: planes are to lowest order non-dynamical in string theory. This is not anymore true
1988: to higher orders as can be inferred from their F-theory interpretation \cite{Sen}.}
1989: This allows to construct consisted D-brane setups with some fraction of supersymmetry preserved.
1990: More precisely, in a background $\Mext \times \Mint$ orientifold planes
1991: wrap cycles in $\Mint$ arising as the fix-point set of $S$. If these
1992: are calibrated with respect to the same form as the cycles wrapped
1993: by the D-branes in the setup, the brane-orientifold setup can preserves some
1994: supersymmetry. We will comment on these conditions later on in this chapter.
1995:
1996: Before we define the precise orientifold projections relevant for this work in section
1997: \ref{oprojections}, let us first collect some possible symmetry
1998: operations allowed in $S$. In the simplest example $S$ only consists
1999: of a target space symmetry $\sigma:\cM_{10}\rightarrow \cM_{10}$,
2000: such that $\Omega_p \sigma $ is a symmetry of the
2001: underlying string theory. This will be the case for IIB orientifolds with $O5$
2002: or $O9$ planes. However, type IIB admits a second perturbative symmetry operation
2003: denoted by $(-1)^{F_L}$, where $F_L$ is the space-time fermion number in the left-moving
2004: sector.
2005: Under the action of $(-1)^{F_L}$ R-NS and R-R states are odd
2006: while NS-R and NS-NS states are even. Orientifolds with $O3$ and/or $O7$ planes
2007: arise from projections of the form $(-1)^{F_L} \Omega_p \sigma$ as we will argue
2008: below. In summary let us display the transformation behavior of the massless bosonic
2009: states under these two operations \cite{JPbook,AD}
2010: \beq \label{transf-AB}
2011: \begin{array}{cllll}
2012: \Omega_p:& \qquad \text{even:}\quad &\hat \phi, \ \hat g,\ \hat C_1,\ \hat C_2 ,
2013: \qquad& \text{odd:}\quad &\hat C_0, \ \hat B_2 ,\ \hat C_3, \ \hat C_4 \ ,\\
2014: (-1)^{F_L}:& \qquad \text{even:}\quad &\hat \phi, \ \hat g,\ \hat B_2 ,
2015: \qquad& \text{odd:}\quad &\hat C_0,\ \hat C_1,\ \hat C_2 ,\ \hat C_3, \ \hat C_4 \ ,
2016: \end{array}
2017: \eeq
2018: where we have also displayed the transformation properties of the type IIA forms.
2019: With these transformations at hand one easily checks that $\Omega_p$ as well as
2020: $(-1)^{F_L}$ are symmetries of the ten-dimensional type IIB supergravity action.
2021: This is in contrast to type IIA. By using \eqref{transf-AB} one immediately notices that
2022: $\Omega_p$, \,$(-1)^{F_L}$ and $(-1)^{F_L}\Omega_p$ alone are no symmetries
2023: of the type IIA effective action \eqref{10dact}. However, orientifolds with
2024: $O6$ planes arise if $S$ includes $(-1)^{F_L}\Omega_p$ as well as some appropriatly
2025: chosen target space symmetry which ensures that $S\Omega_p$ leaves \eqref{10dact} invariant.
2026: Let us now make this more explicite by properly defining the Calabi-Yau orientifold
2027: projections.
2028:
2029: \section{Orientifold projections} \label{oprojections}
2030:
2031: After this brief introduction we are now in the position to specify the orientifolds
2032: under consideration and give an explicit definition
2033: of the orientifold symmetry group \eqref{osym}.
2034: We start from type II string theory and compactify
2035: on a Calabi-Yau threefold $Y$. In addition we
2036: mod out by orientation reversal of the string
2037: world-sheet $\Omega_p$ together with an `internal'
2038: symmetry $\sigma$ which acts solely on $Y$
2039: but leaves the $D=4$ Minkowskian space-time untouched.
2040: We will restrict ourselves to involutive symmetries ($\sigma^2 = 1$) of $Y$
2041: and thus set $G$ in \eqref{osym} to be empty.\footnote{Calabi-Yau manifolds have only
2042: discrete isometries. For example in the case of the quintic,
2043: $\sigma$ could act
2044: by permuting the coordinates such that
2045: the defining equation is left invariant.
2046: A classification of all possible involutions
2047: of the quintic can be found in ref.\ \cite{BH}.}
2048: This avoids the appearance of further twisted
2049: sectors as they appear in general orbifold models \cite{DHVW}. In
2050: a next step we have to specify additional properties of $\sigma$ and the complete
2051: operation $S\Omega_p$ in order that it provides a symmetry of the
2052: string theory under consideration. To do that we discuss the type IIA and type IIB
2053: case in turn.
2054:
2055: \subsubsection{Type IIB orientifolds}
2056:
2057: Let us start with type IIB Calabi-Yau orientifolds and define the orientifold projections following
2058: \cite{Sen,DP,AAHV,BH}.
2059: %We are then in the position to check if these projections
2060: %are a symmetry of the ten-dimensional supergravity action
2061: %for the corresponding string theory.
2062: Later on, in section \ref{IIB_orientifolds} we show that gauge-fixing these symmetries indeed
2063: leads to an $N=1$ supergravity theory.
2064: In type IIB consistency requires
2065: $\sigma$ to be an isometric and holomorphic involution of $Y$ \cite{AAHV,BH}.
2066: A holomorphic isometry leaves both the metric
2067: and the complex structure of the Calabi-Yau manifold invariant.
2068: As a consequence also the K\"ahler form $J$
2069: is invariant such that
2070: \beq
2071: \sigma^* J = J \ ,
2072: \eeq
2073: where $\sigma^*$ denotes the pull-back of the map $\sigma$.
2074: Hence in our analysis we focus on the class of Calabi-Yau threefolds which
2075: admit such an involution but within this class we leave the
2076: threefolds arbitrary. Since the involution is holomorphic
2077: it respects the Hodge decomposition \eqref{odd_even_cohom} and we find
2078: in particular $\sigma^* H^{(3,0)} = H^{(3,0)}$. Picking the
2079: holomorphic three-form $\Omega$ as an representative of $H^{(3,0)}$
2080: and using that $(\sigma^*)^2 =\text{id}$ one is left with two
2081: possible actions
2082: \beq \label{Omegatransf}
2083: (1)\quad O3/O7: \quad \sigma^* \Omega = - \Omega\ ,\qquad \qquad (2)\quad O5/O9: \quad \sigma^* \Omega = + \Omega\ .
2084: \eeq
2085: Correspondingly, depending on the transformation properties of $\Omega$
2086: two different symmetry operations $\mathcal{O}=S\Omega_p$
2087: are possible \cite{Sen,DP,AAHV,BH} \footnote{%
2088: The factor $(-1)^{F_L}$ is included in $\mathcal{O}_{(1)}$
2089: to ensure that $\OO_{(1)}^2=1$ on all states.}
2090: \beq \label{o3-projection}
2091: \mathcal{O}_{(1)} = (-1)^{F_L} \Omega_p \, \sigma\ ,\qquad
2092: \mathcal{O}_{(2)} = \Omega_p \, \sigma
2093: \eeq
2094: where $\Omega_p$ is the world-sheet parity, $F_L$ is the space-time fermion number
2095: in the left-moving sector introduced at the end of section \ref{D-branes}.
2096: This specifies the operation $S \Omega_p$ in \eqref{osym} and,
2097: since $G$ is empty, the complete orientifold projection. We are now in the position to
2098: check if the orientifold projections are indeed a symmetry of the bosonic ten-dimensional
2099: type IIB supergravity action \eqref{10d-lagr}. We will do this check by concentrating only
2100: on some of the terms in \eqref{10d-lagr} keeping in mind that the analysis for the
2101: remaining terms is analoge. The background
2102: $\cM'=\Mext \times \sigma(Y)$ denotes the image of $\cM=\Mext\times Y$ under the geometric action
2103: $\sigma$. Also inserting the $\sigma$-transformed fields into \eqref{10d-lagr} one infers
2104: \footnote{Here we have used \eqref{wedge*comp}
2105: in order to give the component expression of the kinetic terms in \eqref{10d-lagr}.}
2106: \beq \label{tranfact}
2107: S^{(10)}_{IIB'}\ =\
2108: \int_{\cM'}\big( -\tfrac{1}{2} \hat R_{g'} *{'} \mathbf{1}
2109: - \tfrac{1}{4}g'^{MN} (\partial_M \hat \phi')(\partial_N \hat \phi') *{'} \mathbf{1} - \ldots
2110: -\tfrac{1}{4} \hat C_4' \wedge \hat H_3' \wedge \hat F_3'\big)\ ,
2111: \eeq
2112: where $g'=\sigma^* g,\ \hat \phi'=\sigma^*\hat \phi$ etc.\ and the dots denote
2113: terms transforming similar to the kinetic term of $\phi'$. The Hodge star $*'$ is evaluated on
2114: the manifold $\cM'$ with metric $g'$. Now we apply the properties of the involution. Since
2115: $\sigma$ is an isometry we find $g=g'$ and due to the holomorphicity of $\sigma$ we can deduce
2116: that the ten-dimensional volume element $*'\mathbf{1}$ does not change sign in going from $\cM'$ to $\cM$.\footnote{%
2117: Holomorphic maps do not change the orientation of $M$.} This ensures that the
2118: Einstein-Hilbert term takes the from $\int_{\cM'} \sigma^* (- \frac{1}{2} R * \mathbf{1})$ and by applying
2119: \eqref{int-form1} and \eqref{int-form2} is invariant under the isometric map $\sigma$.
2120: A similar reasoning applies to all other terms in \eqref{tranfact} and one concludes
2121: that the effective action \eqref{10d-lagr} is indeed unchanged by $\sigma$.
2122: Combined with the invariance of \eqref{10d-lagr} under the world-sheet parity $\Omega_p$ and $(-1)^{F_L}$
2123: one infers that the orientifold operations \eqref{o3-projection} are symmetries of the effective theory.
2124:
2125:
2126:
2127:
2128: The fix-point set of the involutions $\sigma$ in \eqref{o3-projection} determines
2129: the location of the orientifold planes.
2130: Modding out by $\mathcal{O}_{(1)}$
2131: leads to the possibility of having $O3$- and $O7$-planes
2132: while modding out by $\mathcal{O}_{(2)}$ allows
2133: $O5$- and $O9$-planes. To see this, recall that
2134: the four-dimensional Minkowski space is left invariant by
2135: $\sigma$ such that the orientifold planes are necessarily space-time filling.
2136: Using the fact that $\sigma$ is holomorphic they
2137: have to be even-dimensional (including the time direction) which
2138: selects $O3$-, $O5$-, $O7$- or $O9$-planes as the only possibilities.
2139: The actual
2140: dimensionality of the orientifold plane is then determined
2141: by the dimensionality of the fix-point set of $\sigma$ in $Y$.
2142: In order to determine this dimensionality we need the induced
2143: action of $\sigma$ on the tangent space at any point
2144: of the orientifold plane.
2145: Since one can always choose $\Omega \propto dy^1 \wedge dy^2 \wedge dy^3$
2146: we see that for $\sigma^* \Omega = - \Omega$
2147: the internal part of the orientifold plane is
2148: either a point or a surface of complex dimension two.
2149: Together with the space-time filling part we thus can have
2150: $O3$- and/or $O7$-planes.
2151: The same argument can be repeated for $\sigma^* \Omega = \Omega$
2152: which then leads to the possibility of
2153: $O5$- or $O9$-planes. There are no models with $O5$ and $O9$ planes, since
2154: the appearance of a $O9$ plane implies that the complete background $\cM_{10}$
2155: consist of fix-points of $\sigma=\text{id}$.
2156: The case of $O9$ planes is special and coincides with type I if one
2157: introduces the appropriate number of $D9$-branes to cancel tadpoles.
2158:
2159: Since the involution $\sigma$ is holomorphic the fix-point set of the involution
2160: are holomorphic cycles $\Lambda_{Op}$. This implies that they are calibrated
2161: with respect to the forms $1$ and $J \wedge J$ in orientifolds with $O3/O7$ planes
2162: and with respect to $J$ or $J\wedge J\wedge J$ in orientifolds with $O5$ or $O9$ planes.
2163: More precicely, one finds that the volume forms on $\Lambda_{Op}$ equals the pull-back
2164: of $e^{iJ}$ to the cycle \footnote{%
2165: Here we abbreviate the formal sum of $(q,q)$-forms
2166: $ e^{iJ} = 1 + iJ + \frac{1}{2!}J \wedge J - \frac{i}{3!} J \wedge J\wedge J$.}
2167: \beq \label{cal_sOp}
2168: \vol(\Lambda_{Op}) = e^{-i\theta_{Op}}\, e^{iJ} \big|_{\Lambda_{Op}}\ , \qquad
2169: \theta_{O3/7} = 0\ ,\quad \theta_{O5}=\tfrac{\pi}{2}\ ,\quad \theta_{O9}=-\tfrac{\pi}{2}\ ,
2170: \eeq
2171: where the phase depends on the type of orientifold planes in the setup. Furthermore
2172: one has $\Omega|_{\Lambda_{Op}}=0$. Cycles fulfilling these conditions minimize their volume within
2173: their homology class. Note that similar to \eqref{calcond} this condition has to be modified
2174: in the presence of a $\hat B_2$ field. In this case the form which
2175: calibrates the supersymmetric cycles is $e^{-\hat B_2+iJ}$.
2176: Let us check whether the fix-point sets $\Lambda_{Op}$ of $\sigma$
2177: remain calibrated.
2178: In the two orientifold setups only fields are kept in
2179: the spectrum which are invariant under the respective
2180: projection $\mathcal{O}_{(1/2)}$ given in \eqref{o3-projection}. Thus, by using \eqref{transf-AB}
2181: one infers that $\hat B_2$ has to transform
2182: as $\sigma^* \hat B_2 = - \hat B_2$ for both orientifold projections.
2183: This implies that $\hat B_2$ restricted to the fix-point set of
2184: $\sigma$ vanishes. \footnote{%
2185: Denoting $\rho^* \hat B_2 = \hat B_2|_{\Lambda_{Op}}$ the pull-back to the
2186: fix-point set $\Lambda_{Op}$ of $\sigma$ it follows
2187: $-\rho^*\hat B_2 = \rho^* (\sigma^* \hat B_2) = (\sigma \circ \rho)^* \hat B_2 = \rho^* \hat B_2$
2188: such that $\rho^*\hat B_2=0$.}
2189: One concludes that the cycles $\Lambda_{Op}$ remain
2190: calibrated with respect to the generalized calibration form, i.e.\
2191: \beq\label{cal-Op}
2192: \vol_{DBI}(\Lambda_{Op}) = e^{-i\theta_{Op}} e^{-\hat B_2+iJ} \big|_{\Lambda_{Op}}\ ,
2193: \eeq
2194: where $\theta_{Op}$ is as given in \eqref{cal_sOp} and $\vol_{DBI}(\Lambda_{Op})$ is defined as
2195: in \eqref{calcond}.
2196: At this point, one can compare the calibration condition \eqref{cal-Op} for the
2197: orientifold planes with the one for the $Dp$-branes given in \eqref{calcond}.
2198: In order to preserve some supersymmetry all orientifold planes and
2199: D-branes, have to be calibrated with respect to the same form. This implies that
2200: the phases $\theta_{Dp}$ in \eqref{calcond} have to coincide with $\theta_{Op}$ given
2201: in \eqref{cal_sOp} (see also \cite{JL} for the case of $D3/D7$ branes). This is
2202: equivalently true for $Dq$-instantons wrapping $q+1$-cycles in $Y$. In supersymmetric setups
2203: with $O(q+3)$ planes one has to set $\theta_{Dq}=\theta_{O(q+3)}$, where $e^{i\theta_{Dq}}$
2204: is the phase in the D-instanton calibration condition.
2205:
2206:
2207:
2208:
2209:
2210: \subsubsection{Type IIA orientifolds}
2211:
2212: Let us now turn to the type IIA Calabi-Yau orientifolds.
2213: In contrast to type IIB the orientifold projection has
2214: to include an anti-holomorphic and isometric involution $\sigma$
2215: in order to preserve $N=1$ supersymmetry \cite{AAHV,BBKL,BH}.
2216: Hence, the K\"ahler form on $Y$ transforms as
2217: \beq \label{constrJ}
2218: \sigma^* J\ =\ -J\ ,
2219: \eeq
2220: since $\sigma$ preserves the metric but yields a minus sign when applied to
2221: the complex structure.
2222: The complete projection takes the form
2223: \beq \label{oproj}
2224: \mathcal{O} = (-1)^{F_L} \Omega_p \sigma\ .
2225: \eeq
2226: In addition to the condition \eqref{constrJ}
2227: compatibility of $\sigma$ with the Calabi-Yau condition
2228: $\Omega \wedge \bar \Omega \propto J \wedge J \wedge J$
2229: implies that $\sigma$ also acts non-trivially on the three-form $\Omega$ as
2230: \beq \label{constrO}
2231: \sigma^* \Omega\ =\ e^{2i\theta} \bar \Omega \ ,
2232: \eeq
2233: where $e^{2i\theta}$ is a constant phase and we included a factor 2 for later convenience.
2234: Similar to the type IIB case we can check that the
2235: projection $\mathcal{O}$ is a symmetry of the type IIA supergravity
2236: action \eqref{10dact}. Note however, that $(-1)^{F_L} \Omega_p$ alone is not a symmetry
2237: of type IIA. Using \eqref{transf-AB} this can be already inferred from the fact that the kinetic and
2238: topological terms in \eqref{10dact} transform with a different sign.
2239: On the other hand, under the action of the involution $\sigma$ the effective action changes as
2240: \beq
2241: S^{(10)}_{IIA'}\ =\
2242: \int_{\cM'}\big( -\tfrac{1}{2} \hat R_{g'} *{'} \mathbf{1}
2243: - \tfrac{1}{4}g'^{MN} (\partial_M \hat \phi')(\partial_N \hat \phi') *{'} \mathbf{1} \ldots
2244: -\tfrac{1}{2} \hat B_2' \wedge \hat F_4' \wedge \hat F_4' \big)\ ,
2245: \eeq
2246: where as in \eqref{tranfact} we have set $g'=\sigma^* g,\ \hat \phi'=\sigma^*\hat \phi$ etc.\ and the Hodge star $*'$ is on
2247: the manifold $\cM'=\Mext \times \sigma(Y)$ with metric $g'$. Using the fact that $\sigma$ is
2248: an anti-holomorpic isometric involution it changes the sign of the volume element
2249: $*\mathbf{1} \sim \vol(\Mext) \wedge J' \wedge J' \wedge J'$, such that $*'\mathbf{1}=-*\mathbf{1}$.
2250: From equations \eqref{int-form1} and \eqref{int-form2} one finds that the topological term transforms with a minus sign
2251: while the kinetic terms remain invariant. This extra sign cancels the minus from the action of
2252: $(-1)^{F_L} \Omega_p$ such that $\mathcal{O}$ is indeed a symmetry of \eqref{10dact}.
2253: In section \ref{IIA_orientifolds} we show that gauge-fixing this symmetry results in an $N=1$ supergravity
2254: theory.
2255:
2256: Type IIA orientifolds with anti-holomorphic involution generically contain $O6$ planes. This
2257: is due to the fact, that the fixed point set of $\sigma$ in $Y$ are three-cycles
2258: $\Lambda_{O6}$ supporting the internal part of the orientifold planes.
2259: These cycles are special Lagrangian submanifolds of $Y$ as an
2260: immediate consequences of \eqref{constrJ} and \eqref{constrO}
2261: which implies \cite{HitchinLec}
2262: \beq \label{OLagr}
2263: J|_{\Lambda_{O6}} = 0\ , \qquad \I(e^{-i\theta}\Omega)|_{\Lambda_{O6}} = 0\ .
2264: \eeq
2265: In other words, they are calibrated with respect to
2266: %$\R(e^{-U-i\theta}\Omega)$
2267: $\R(e^{-i\theta}\Omega)$
2268: \beq \label{calibr-O6}
2269: \rm{vol}(\Lambda_{O6})\sim \R(e^{-i\theta}\Omega)\ ,
2270: \eeq
2271: where the overall normalization of $\Omega$ will be determined
2272: in \eqref{Omeganorm}. Once again this poses conditions on
2273: additional D-branes in the setup, if they are demanded to preserve the same
2274: supersymmetry. More precicely, BPS branes have to be calibrated
2275: with respect to the same form as the orientifold planes. This implies
2276: by comparing \eqref{calcond} with \eqref{calibr-O6} that
2277: $\theta_{D6}=\theta$ for space-time filling $D6$-branes wrapping a
2278: three-cycle in $Y$. A similar condition $\theta_{D2}=\theta$ has
2279: to hold for supersymmetric $D2$-instantons wrapping a three-cycle in $Y$.
2280:
2281:
2282: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2283: %
2284: % IIB orientifolds
2285: %
2286: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2287:
2288:
2289: \section{Type IIB Calabi-Yau orientifolds \label{IIB_orientifolds}}
2290:
2291:
2292: In this section we impose the projection \eqref{o3-projection}
2293: on the type IIB theory and derive the massless spectrum
2294: (section~\ref{o3-spectrum}) and its
2295: low energy $N=1, D=4$ effective supergravity action
2296: (section~\ref{O37_effective_act}).
2297: This generalizes similar derivations already performed in refs.\
2298: \cite{GKP,BBHL}.
2299: We restrict our analysis to the bosonic fields of the compactification
2300: keeping in mind that the couplings of the
2301: fermionic partners are fixed by supersymmetry. Furthermore, we
2302: include space-time filling D-branes for consistency but fix their
2303: moduli, such that they do not appear in the low energy effective
2304: action. The compactification we perform is closely related to the compactification
2305: of type IIB string theory on Calabi-Yau threefolds reviewed
2306: in chapter \ref{TypeII}. The orientifold projection
2307: \eqref{o3-projection}
2308: truncates the massless spectrum from $N=2$ to $N=1$
2309: multiplets and also leads to a modification of the couplings
2310: which render the low energy effective theory
2311: compatible with $N=1$ supergravity.
2312: Such truncation procedures from $N=2$ to $N=1$ supergravity has been carried
2313: out from a purely supergravity point of view
2314: in refs.\ \cite{ADAF}.
2315:
2316: \subsection{The $N=1$ spectrum \label{o3-spectrum}}
2317: %
2318: Before computing the effective action let us first
2319: determine the massless spectrum when
2320: the orientifold projection is taken into account and see how
2321: the fields assemble in $N=1$ supermultiplets \cite{BH}.
2322: In the
2323: four-dimensional compactified theory
2324: only states invariant under the projection are kept.
2325: Using equation \eqref{transf-AB} one immediately infers that
2326: the scalars $\hat \phi,\hat l$, the metric $\hat g$ and the
2327: four-form $\hat C_4$ are even under $(-1)^{F_L} \Omega_p$
2328: while both two forms $\hat B_2, \hat C_2$ are odd.
2329: Using \eqref{o3-projection} this implies that the invariant
2330: states have to obey
2331: \begin{equation} \label{fieldtransfB}
2332: \begin{array}{lcl}
2333: \\
2334: \sigma^* \hat \phi &=& \ \hat \phi\ , \\
2335: \sigma^* \hat g &=& \ \hat g\ , \\
2336: \sigma^* \hat B_2 &=& - \hat B_2\ ,
2337: \end{array}
2338: \hspace{1cm}
2339: \begin{array}{lcl}
2340: \multicolumn{3}{c}{ \underline{O3/O7}} \\[2ex]
2341: \sigma^* \hat C_0 &=& \ \ \hat C_0\ , \\
2342: \sigma^* \hat C_2 &=& - \hat C_2\ , \\
2343: \sigma^* \hat C_4 &=& \ \ \hat C_4\ ,
2344: \end{array}
2345: \hspace{1cm}
2346: \begin{array}{lcl}
2347: \multicolumn{3}{c}{ \underline{O5/O9}} \\[2ex]
2348: \sigma^* \hat C_0 &=& - \hat C_0\ , \\
2349: \sigma^* \hat C_2 &=& \ \ \hat C_2\ , \\
2350: \sigma^* \hat C_4 &=& - \hat C_4\ ,
2351: \end{array}
2352: \end{equation}
2353: where the first column is identical for both involutions $\sigma$
2354: in \eqref{o3-projection}.
2355: Since $\sigma$ is a holomorphic involution the cohomology groups $H^{(p,q)}$
2356: (and thus the
2357: harmonic $(p,q)$-forms) split into two eigenspaces
2358: under the action of $\sigma^*$
2359: \beq\label{H3split}
2360: H^{(p,q)} =
2361: H^{(p,q)}_+\oplus H^{(p,q)}_-\ .
2362: \eeq
2363: $H^{(p,q)}_+$ has dimension $h_+^{(p,q)}$ and denotes
2364: the even eigenspace of $\sigma^*$ while
2365: $H^{(p,q)}_-$ has dimension $h_-^{(p,q)}$ and denotes
2366: the odd eigenspace of $\sigma^*$.
2367: The Hodge $*$-operator commutes with $\sigma^*$ since $\sigma$ preserves the
2368: orientation and the metric of the Calabi-Yau manifold and thus the Hodge
2369: numbers obey $h^{(1,1)}_\pm=h^{(2,2)}_\pm$. Holomorphicity of $\sigma$
2370: further implies $h^{(2,1)}_\pm = h^{(1,2)}_\pm$ while
2371: \eqref{Omegatransf} leads to
2372: $h^{(3,0)}_+ = h^{(0,3)}_+=0, h^{(3,0)}_- = h^{(0,3)}_-=1$ for $O3/O7$ orientifolds
2373: and $h^{(3,0)}_+ = h^{(0,3)}_+=1, h^{(3,0)}_- = h^{(0,3)}_-=0$ for $O5/O9$ orientifolds.
2374: Furthermore, the volume-form which is proportional
2375: to $\Omega\wedge\bar\Omega$ is invariant under $\sigma^*$ and thus one has
2376: $h^{(0,0)}_+ = h^{(3,3)}_+=1, h^{(0,0)}_- = h^{(3,3)}_-=0$.
2377: We summarize the non-trivial cohomology groups including
2378: their basis elements in table~\ref{CYObasisB}.
2379: \begin{table}[h]
2380: \begin{center}
2381: \begin{tabular}{|c || c | c || c| c || c | c |} \cline{1-7}
2382: setup &\multicolumn{2}{|c||}{\rule[-0.3cm]{0cm}{0.8cm} cohomology group} &
2383: \multicolumn{2}{|c||}{dimension} & \multicolumn{2}{|c|}{basis}
2384: \\ \cline{1-7}
2385: \multirow{3}{1.2cm}[-.3cm]{$O3/O7$ $\text{\ \ and}$ $O5/O9$} &\rule[-0.3cm]{0cm}{0.8cm} $H^{(1,1)}_+$ & $H^{(1,1)}_-$ &
2386: $h^{(1,1)}_+$ & $h^{(1,1)}_- $ & $\omega_\alpha$ & $\omega_a$
2387: \\ \cline{2-7}
2388: &\rule[-0.3cm]{0cm}{0.8cm} $H^{(2,2)}_+$ & $H^{(2,2)}_-$ & $h^{(1,1)}_+$ & $h^{(1,1)}_-$ &
2389: $\tilde \omega^\alpha$ & $\tilde \omega^a$
2390: \\ \cline{2-7}
2391: &\rule[-0.3cm]{0cm}{0.8cm} $H^{(2,1)}_+$ & $H^{(2,1)}_-$
2392: & $h^{(2,1)}_+$ & $h^{(2,1)}_-$ &
2393: $\chi_\kappa$ & $\chi_k$
2394: \\ \hline
2395: O3/O7&\rule[-0.3cm]{0cm}{0.8cm} $H^{(3)}_+$ & $H^{(3)}_-$ & $2h^{(2,1)}_+$ & $2h^{(2,1)}_-+2$ &
2396: $(\alpha_{\kappa},\beta^{\lambda})$ & $(\alpha_{\hat k},\beta^{\hat l})$
2397: \\ \hline
2398: O5/O9&\rule[-0.3cm]{0cm}{0.8cm} $H^{(3)}_+$ & $H^{(3)}_-$ & $2h^{(2,1)}_+ +2$ & $2h^{(2,1)}_-$ &
2399: $(\alpha_{\kappa},\beta^{\lambda})$ & $(\alpha_{\hat k},\beta^{\hat l})$ \\ \hline
2400: \end{tabular}
2401: \caption{\small \label{CYObasisB}
2402: \textit{Cohomology groups and their basis elements.}}
2403: \end{center}
2404: \end{table}
2405:
2406:
2407:
2408:
2409:
2410: The four-dimensional invariant spectrum
2411: is found by using the Kaluza-Klein expansion
2412: given in eqs.\ \eqref{def-v}, \eqref{cs} and \eqref{CYexpansion}
2413: keeping only the fields which in addition obey \eqref{fieldtransfB}.
2414: We see immediately that the $D=4$ scalar field arising from
2415: $\hat\phi$ remains in the spectrum for both setups and as before we denote it by
2416: $\phi$. Since $\sigma^*$ leaves
2417: the K\"ahler form $J$ invariant only the
2418: $h_+^{(1,1)}$ even K\"ahler deformations $v^\alpha$ remain in the spectrum
2419: and we expand
2420: \beq\label{transJo}
2421: J = v^{\alpha}\, \omega_{\alpha} \ ,\qquad
2422: \alpha = 1,\ldots, h_+^{(1,1)}\ ,
2423: \eeq
2424: where $\omega_\alpha$ denotes a basis of $H^{(1,1)}_+$.
2425: {} Similarly, from eq.\ \eqref{cs} we infer that the
2426: invariance of the metric together with
2427: \eqref{Omegatransf} implies that the complex structure deformations
2428: kept in the spectrum correspond to elements in $H^{(1,2)}_-$ for $O3/O7$
2429: setups and to elements of $H^{(1,2)}_+$ for $O5/O9$. Hence, \eqref{cs} is
2430: replaced by
2431: \bea\label{cso}
2432: O3/O7:\quad \delta{g}_{ij} = \frac{i}{||\Omega||^2}\, \bar z^{k}
2433: (\bar \chi_{ k})_{i\ib\bj}\,
2434: \Omega^{\ib\bj}{}_j \ ,\quad k=1,\ldots,h_-^{(1,2)}\ , \\ \qquad
2435: O5/O9:\quad \delta{g}_{ij} = \frac{i}{||\Omega||^2}\, \bar z^{\kappa}
2436: (\bar \chi_{\kappa})_{i\ib\bj}\,
2437: \Omega^{\ib\bj}{}_j \ , \quad \kappa =1,\ldots,h_+^{(1,2)}\ ,\nn
2438: \eea
2439: where $\bar\chi_{k}\ (\bar \chi_{\kappa})$ denotes a basis of $H^{(1,2)}_-\ (H^{(1,2)}_+)$.\footnote{%
2440: In ref.\ \cite{BH} it is further shown that the
2441: $h_\pm^{(1,2)}$ deformations form a smooth submanifold
2442: of the Calabi-Yau moduli space.}
2443:
2444: {}From eqs.\ (\ref{fieldtransfB}) we learn that in the expansion of
2445: $\hat B_2$ only odd elements are kept. Thus, for both orientifold setups we
2446: have
2447: \beq \label{exp-B}
2448: \hat B_2 = b^a\, \omega_a\ ,\qquad a=1,\ldots, h_-^{(1,1)}\ ,
2449: \eeq
2450: where $\omega_a$ is a basis of $H^{(1,1)}_-$.
2451: The orientifold projections differ in the R-R sector. For $O3/O7$ orientifolds
2452: $\hat C_2$ is odd and $\hat C_4$ is even. Therefore the expansion \eqref{CYexpansion}
2453: is replaced by
2454: \beq\label{exp1}
2455: \hat C_2\ =\ c^a\, \omega_a\ , \qquad
2456: \hat C_4\ =\ D_2^\alpha\wedge \omega_\alpha
2457: + V^{\kappa}\, \wedge \alpha_{\kappa}
2458: + U_{\kappa}\wedge\beta^{\kappa}+
2459: \rho_\alpha\ \tilde \omega^\alpha\ ,
2460: \eeq
2461: where $\tilde\omega^\alpha$ is a basis
2462: of $H^{(2,2)}_+$ which is dual to $\omega_\alpha$, and
2463: $(\alpha_{\kappa}, \beta^{\kappa})$ is a real, symplectic
2464: basis of $H^{(3)}_+ = H^{(1,2)}_+ \oplus H^{(2,1)}_+$
2465: (c.f.\ table~\ref{CYObasisB}). From \eqref{fieldtransfB} we find that
2466: the axion $\hat C_0$ remains in the spectrum and we denote
2467: the corresponding four-dimensional field by $C_0$.
2468: Note that the two $D=4$ two-forms $B_2$ and $C_2$ present in the $N=2$
2469: compactification (see \eqref{CYexpansion})
2470: have been projected out and in the expansion of $\hat B_2$ and $\hat C_2$
2471: only the scalar fields $c^a, b^a$ appear.
2472: The non-vanishing of $c^a,b^a$ and $V^\kappa$ is closely related to the
2473: appearance of $O7$-planes. To understand this in more detail
2474: we recall, that $O3$-planes appear
2475: when the fix-point set of $\sigma$ is zero-dimensional in $Y$
2476: or in other words all tangent vectors at this point are odd under
2477: the action of $\sigma$.
2478: This in turn implies that locally
2479: two-forms are even under $\sigma^*$, while three-forms
2480: are odd. However, this is incompatible
2481: with the expansions given in \eqref{exp1} for
2482: non-vanishing $b^a,c^a$ and $V^\kappa$.
2483: For a setup also including $O7$-planes we locally
2484: get the correct transformation behavior,
2485: so that harmonic forms in $H^{(1,1)}_-$ and
2486: $H^{(2,1)}_+$ can be supported.
2487:
2488: For $O5/O9$ orientifolds the $\mathcal{O}_{(2)}$-invariant R-R forms transform exactly
2489: with the opposite sign under $\sigma$. Thus,
2490: the expansion \eqref{CYexpansion} reduces to
2491: \beq\label{expO5}
2492: \hat C_2\ =\ C_2+ c^\alpha\ \omega_\alpha\ ,
2493: \qquad \hat{C}_4 \ =\ D_{2}^a \wedge \omega_a + V^{k} \wedge
2494: \alpha_{k} - U_{k} \wedge \beta^{k} +
2495: \rho_a\, \tilde \omega^a\ .
2496: \eeq
2497: In this case the axion $\hat C_0$ is projected out and replaced by
2498: the $D=4$ antisymmetric tensor $C_2(x)$. As a consequence the $N=1$
2499: spectrum contains a `universal' linear multiplet $(\phi,C_2)$ which in the massless
2500: case can be dualized to a chiral multiplet.
2501: As for Calabi-Yau compactifications
2502: imposing the self-duality on $\hat F_5$
2503: eliminates half of the degrees of freedom in the expansions \eqref{exp1} and \eqref{expO5}
2504: of $\hat C_4$. For the one-forms $V^{\cdot},U_{\cdot}$ this corresponds to
2505: the choice of electric versus magnetic gauge potentials.
2506: On the other hand choosing the two forms $D_2^{\cdot}$
2507: or the scalars $\rho_{\cdot}$ determines
2508: the structure of the $N=1$ multiplets to be either a linear or a chiral
2509: multiplet and in chapter \ref{lin_geom_of_M} we discuss both cases.
2510:
2511: Altogether the resulting $N=1$ fields for the two setups
2512: assembles into a gravitational
2513: multiplet, $h_\pm^{(2,1)}$ vector multiplets and
2514: $(h_\mp^{(2,1)}+ h^{(1,1)}+1)$ chiral multiplets
2515: and are
2516: summarized in table~\ref{N=1spectrumtab} \cite{BH,TGL1}.
2517:
2518:
2519: \begin{table}[h]
2520: \begin{center}
2521: \begin{tabular}{|c|c|c||c|c|} \hline
2522: \rule[-0.3cm]{0cm}{0.8cm} & \multicolumn{2}{|c||}{$O3/O7$} & \multicolumn{2}{c|}{$O5/O9$}\\ \hline
2523: \rule[-0.3cm]{0cm}{0.85cm}
2524: gravity multiplet&1&$g_{\mu \nu} $ &1& $g_{\mu \nu}$ \\ \hline
2525: \rule[-0.3cm]{0cm}{0.85cm}
2526: vector multiplets& $\ h_+^{(2,1)}\ $& $V^{\lambda} $& $\ h_-^{(2,1)}\ $ & $V^{k} $ \\ \hline
2527: \rule[-0.3cm]{0cm}{0.85cm}
2528: \multirow{3}{30mm}[-3.5mm]{chiral multiplets} & $h_-^{(2,1)}$& $z^{k} $ & $h_+^{(2,1)}$& $z^{\lambda} $\\ \cline{2-5}
2529: \rule[-0.3cm]{0cm}{0.85cm}
2530: & $ h^{(1,1)}_-$ &$( b^a, c^a)$ & $h^{(1,1)}_+$& $( v^\alpha, c^\alpha)$\\ \cline{2-5}
2531: \rule[-0.3cm]{0cm}{0.85cm}
2532: & 1 & $(\phi,l)$ && \\ \hline
2533: \rule[-0.3cm]{0cm}{0.85cm}
2534: \multirow{3}{44mm}[2mm]{chiral/linear multiplets } & $h^{(1,1)}_+$& $( v^\alpha, \rho_\alpha )$ & $h^{(1,1)}_-$&
2535: $( b^a, \rho_a )$\\ \cline{2-5}
2536: \rule[-0.3cm]{0cm}{0.85cm} & && 1 & $(\phi,C_2)$ \\
2537: \hline
2538: \end{tabular}
2539: \caption{\label{N=1spectrumtab} $N =1$ spectrum of Type IIB orientifold compactifications.}
2540: \end{center}
2541: \end{table}
2542:
2543: Compared to the $N=2$ spectrum of the Calabi-Yau compactification
2544: given in table~\ref{tab-compIIBspec} we see that
2545: the graviphoton `left' the gravitational multiplet
2546: while the $h^{(2,1)}$ $N=2$ vector multiplets decomposed
2547: into $h_\pm^{(2,1)}$ $N=1$ vector multiplets plus $h_\mp^{(2,1)}$
2548: chiral multiplets. Furthermore, the $h^{(1,1)}+1$ hypermultiplets
2549: lost half of their physical degrees of freedom and are reduced
2550: into $h^{(1,1)}+1$ chiral multiplets. This is
2551: consistent with the theorem of \cite{AM,ADAF} where it was shown that
2552: any K\"ahler submanifold of a quaternionic manifold
2553: can have at most half of its (real) dimension.
2554:
2555:
2556:
2557:
2558:
2559: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2560: %
2561: % IIB: O3/O7 effective action
2562: %
2563: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2564:
2565: \subsection{The effective action \label{O37_effective_act}}
2566:
2567: In following we derive the effective actions encoding the
2568: dynamics of the $N=1$ multiplets of the type IIB orientifold
2569: theories. However, before entering the actual computations
2570: a cautionary note is in order.
2571: In the presence of localized sources such as orientifold planes and
2572: D-branes as well as in the presence of non-trivial background fluxes
2573: the product Ansatz \eqref{lineel} for the metric is strictly speaking not anymore
2574: suitable. This is due to the fact that the supergravity theory with source
2575: terms and fluxes does not have the background metric \eqref{lineel} as a solution
2576: \cite{BB1,Verlinde,GSS,GKP}. As deviation from the standard Calabi-Yau compactifications
2577: a non-trivial warp factor $e^{-2A}$ has to be included into the
2578: Ansatz for the metric \eqref{lineel} such that \cite{GKP,GP}
2579: \beq
2580: ds^2=e^{2A(y)} {g_{\mu \nu}}(x) dx^{\mu} dx^{\nu}+ e^{-2A(y)}
2581: {g_{i\bj}}(y) dy^i d\bar y^{\bj}\ .
2582: \label{warpmetric}
2583: \eeq
2584: However, in this work we perform our analysis in the
2585: unwarped Calabi-Yau manifold since in the large radius limit
2586: the warp factor approaches one and the metrics
2587: of the two manifolds coincide \cite{GKP,FP}.
2588: This in turn also implies that the metrics on the moduli space
2589: of deformations agree and as a consequence the kinetic terms
2590: in the low
2591: energy effective actions are the same. The difference appears
2592: in the potential when some of the
2593: Calabi-Yau zero modes are rendered massive. However,
2594: the regime $e^{2A(y)} \approx 1$ should be understood as a
2595: special limit and it would be desirable to generalize compactifications
2596: to warped backgrounds \eqref{warpmetric}.
2597:
2598: Let us now turn to the derivation of the four-dimensional effective action
2599: by redoing the Kaluza-Klein reduction of the ten-dimensional
2600: type IIB action given in \eqref{10d-lagr} for the truncated orientifold
2601: spectrum.
2602:
2603: \subsubsection{The reduction of the $N=2$ vector sector}
2604:
2605: We first consider the reduction of the vector sector
2606: of the $N=2$ supergravity theory obtained by type IIB
2607: Calabi-Yau compactification. As discussed in section \ref{revIIB} the
2608: four-dimensional bosonic components of the vector multiplets
2609: are $(z^K,V^K)$. The complex scalars $z^K$ parameterize the
2610: complex structure deformations of $Y$. Under the
2611: orientifold projection these $N=2$ multiplets split into
2612: chiral multiplets with bosonic components $(z^k)$ and vector
2613: multiplets $(V^\lambda)$ for $O3/O7$ orientifolds and chiral multiplets $(z^\lambda)$
2614: and vectors $(V^k)$ in $O5/O9$ orientifolds. Since the reduction of the vector
2615: sector is very similar for both the $O3/O7$ and $O5/O9$ case
2616: we will first concentrate on the first case
2617: and later give a rule how to translate these results to
2618: $O5/O9$ orientifolds.
2619:
2620: %In order to not overload the notation
2621: %we now choose the same indices for the $O3/O7$ and $O5/O9$ case.
2622: %In other words, the range of the indices is
2623: %\bea \label{indices1}
2624: % O3/O7:\quad \kappa = 1\ldots h^{(2,1)}_+\ , \quad k = 1\ldots h^{(2,1)}_-\ , \quad \hat k = 0\ldots h^{(2,1)}_-\ ,\\
2625: % O5/O9:\quad \kappa = 1\ldots h^{(2,1)}_-\ , \quad k = 1\ldots h^{(2,1)}_+\ , \quad \hat k = 0\ldots h^{(2,1)}_+\ ,\nn
2626: %\eea
2627: %where the hatted indices run over one more value. This is consistent with the fact that by
2628: %\eqref{cso} the three-form $\Omega$ is in $H^{(3)}_-$ for $O3/O7$ setups and in $H^{(3)}_+$ for
2629: %$O5/O9$ setups.
2630:
2631: Due to the split of the cohomology
2632: $H^{(3)}= H^{(3)}_+\oplus H^{(3)}_-$ the real symplectic basis
2633: $(\alpha_\Kh , \beta^\Lh)$
2634: of $H^{(3)}$ can be split into $(\alpha_{\kappa} , \beta^{\lambda})$
2635: of $H^{(3)}_+$ and $(\alpha_{\hat k} , \beta^{\hat l})$
2636: of $H^{(3)}_-$.
2637: % where the upper sign is for $O3/O7$ and the lower for $O5/O9$.
2638: Eqs.\ \eqref{int-numbers1} continue to hold which implies
2639: that both basis are symplectic and obey
2640: \begin{equation}\label{sbasiso}
2641: \int \alpha_{\kappa} \wedge \beta^{\lambda}
2642: = \delta^{\lambda}_{\kappa}\ ,
2643: \qquad
2644: \int \alpha_{\hat k} \wedge \beta^{\hat l}
2645: = \delta^{\hat l}_{\hat k}\ ,
2646: \end{equation}
2647: with all other intersections vanishing. Since $\hat C_4$ is even under $\sigma^*$
2648: the expansion \eqref{exp1} led to $h^{(3)}_+ = h^{(2,1)}_+$ vectors
2649: $V^\kappa$. The three-form $\Omega$ is odd under $\sigma^*$ and thus has to be expanded
2650: in a basis of $H^{(3)}_-$ according to
2651: \begin{equation}
2652: \Omega(z^k) = Z^{\hat k}\alpha_{\hat k} - \mathcal{F}_{\hat k}
2653: \beta^{\hat k}\ ,
2654: \label{cond-1}
2655: \end{equation}
2656: while the other periods $(Z^{\kappa},\mathcal{F}_{\kappa})$
2657: vanish
2658: \beq
2659: Z^{\kappa}|_{z^{\kappa}=0}= \int_Y \Omega\wedge\beta^\kappa = 0 \ , \qquad
2660: \mathcal{F}_{\kappa}\big|_{z^{\kappa}=0}
2661: = \int_Y \Omega\wedge\alpha_\kappa = 0 \ .
2662: \eeq
2663: As a consequence the metric on the space
2664: of complex structure deformations reduces to
2665: \beq\label{csmetrico}
2666: G_{ kl} =
2667: \frac{\partial}{\partial z^{k}}
2668: \frac{\partial}{\partial\bar z^{l}}
2669: \ K_{\rm cs}\ , \qquad
2670: K_{\rm cs} = -\ln\Big[ - i \int_Y \Ox \wedge \bar \Ox\Big]
2671: = -\ln i\Big[ Z^{\hat k} \bar \cF_{\hat k} - \bar Z^{\hat k} {\mathcal{F}}_{\hat k} \Big]
2672: \ ,
2673: \eeq
2674: replacing \eqref{csmetric}. The reduction of the kinetic terms for
2675: the $N=2$ vector sector thus yields \cite{TGL1}
2676: \beq \label{red-vector}
2677: S^{(4)\, vec}_{O3/O7}\ =\ \int - G_{k { l}} \; dz^{k} \wedge *d\bar z^{l}
2678: +\tfrac{1}{4}\text{Im}\; \cM_{\kappa \lambda}\;
2679: F^{\kappa}\wedge *F^{\lambda}
2680: +\tfrac{1}{4}\text{Re}\; \cM_{\kappa \lambda}\;
2681: F^{\kappa}\wedge F^{\lambda}\ ,
2682: \eeq
2683: where $F^\lambda = dV^\lambda$. Recall that the vectors $V^k$ as well
2684: as the graviphoton are projected out by the orientifold projection \eqref{o3-projection}
2685: and do not appear in \eqref{red-vector}.
2686: The coupling matrix $\cM_{\kappa \lambda}(z^k)$ in front of the remaining vectors
2687: $V^\kappa$ is evaluated on the subspace where $z^\kappa=0$ and thus depends on
2688: $z^k$ only. The analysis for $O5/O9$ orientifolds is in complete anology to the
2689: $O3/O7$ case, with the difference that the vectors $V^k$ and scalars $z^\lambda$
2690: remain in the spectrum while $V^\lambda$ and $z^k$ is projected out.
2691: The equations \eqref{sbasiso} -- \eqref{red-vector} can be translated to this second case by replacing
2692: the indices $k,l \rightarrow \kappa, \lambda$, $\kh \rightarrow \hat \kappa$ and
2693: $\kappa,\lambda \rightarrow k,l$.
2694: This is consistent with the fact that by
2695: \eqref{cso} the three-form $\Omega$ is in $H^{(3)}_-$ for $O3/O7$ setups and in $H^{(3)}_+$ for
2696: $O5/O9$ setups.
2697:
2698: \subsubsection{The reduction of the $N=2$ quaternionic sector}
2699:
2700: Similar to the vector sector, we now perform the reduction
2701: of the hypermultiplet couplings \eqref{q-metrB}. One computes
2702: the four-dimensional effective action by redoing the Kaluza-Klein reduction
2703: of the ten-dimensional type IIB action given in \eqref{10d-lagr} for the
2704: truncated orientifold spectrum. Equivalently, one can impose the
2705: orientifold constrains on the four-dimensional $N=2$ effective action
2706: \eqref{action3}. In type IIB the metric on the quaternionic manifold depends on the
2707: complexified K\"ahler deformations $t$ and the dilaton and is obtained from
2708: the intersection numbers in the even cohomologies. Hence, in order to perform the reduction
2709: to $N=1$ we first need to reconsider the structure of the metrics \eqref{Kmetric} and
2710: the intersection numbers \eqref{int-numbers} for the orientifold.
2711:
2712: Note that $\sigma^*J=J$ and $\sigma^*\hat B_2=-\hat B_2$ holds for both IIB orientifold
2713: projections. This implies that the constraints on the space of
2714: K\"ahler structure deformations are the same for $O3/O7$ as well as $O5/O9$ setups.
2715: Let us discuss them in the following.
2716: Corresponding to the decomposition
2717: $H^{(1,1)}=H^{(1,1)}_+\oplus H^{(1,1)}_-$ also
2718: the harmonic (1,1)-forms $\omega_A$
2719: split into
2720: $\omega_A = (\omega_\alpha, \omega_a)$
2721: %A= 1,\ldots,h^{(1,1)}
2722: %\alpha = 1,\ldots,h^{(1,1)}_+
2723: %a = 1,\ldots,h^{(1,1)}_-
2724: such that
2725: $\omega_\alpha$ is a basis of
2726: $H^{(1,1)}_+$ and
2727: $\omega_a$ is a basis of
2728: $H^{(1,1)}_-$. This in turn results in a decomposition of the intersection
2729: numbers $\KK_{ABC}$ given in \eqref{int-numbers}.
2730: Under the orientifold projection
2731: only $\KK_{\alpha\beta\gamma}$ and $\KK_{\alpha bc}$ can be non-zero
2732: while $\KK_{\alpha \beta c}= \KK_{abc} =0$ has to hold.
2733: Since the K\"ahler-form $J$ is invariant
2734: we also conclude from \eqref{int-numbers}
2735: that $\KK_{\alpha b}=0=\KK_{a}$. To summarize,
2736: keeping only the invariant intersection numbers results in
2737: \begin{eqnarray}\label{constr}
2738: \KK_{\alpha \beta c}= \KK_{abc} =\KK_{\alpha b}=\KK_{a}=0\ ,
2739: \end{eqnarray}
2740: while all the other intersection numbers can be non-vanishing.\footnote{From
2741: a supergravity point of view this
2742: has been also observed in refs.\ \cite{ADAF}.}
2743: Inserting \eqref{constr} into \eqref{Kmetric} we derive
2744: \begin{eqnarray} \label{splitmetr}
2745: G_{\alpha \beta}=
2746: -\frac{3}{2}\left( \frac{\KK_{\alpha \beta}}{\KK}-
2747: \frac{3}{2}\frac{\KK_\alpha \KK_\beta}{\KK^2} \right)\ , \qquad
2748: G_{a b}=-\frac{3}{2} \frac{\KK_{a b}}{\KK}\ , \qquad
2749: G_{\alpha b}\ =\ G_{a \beta}\ =\ 0\ ,
2750: \end{eqnarray}
2751: where
2752: \begin{equation}\label{intO3}
2753: \KK_{\alpha\beta}=\KK_{\alpha\beta\gamma}\; v^\gamma\ ,
2754: \quad \
2755: \KK_{ab}=\KK_{ab\gamma}\; v^\gamma\ ,\quad
2756: \KK_{\alpha}=\KK_{\alpha \beta\gamma}\; v^\beta v^\gamma\ ,
2757: \quad \KK=\KK_{\alpha \beta \gamma} \; v^\alpha v^\beta v^\gamma
2758: \ .
2759: \end{equation}
2760: %
2761: We see that the metric $G_{AB}$ given in \eqref{Kmetric}
2762: is block-diagonal with respect to the
2763: decomposition $H^{(1,1)}=H^{(1,1)}_+\oplus H^{(1,1)}_-$.
2764: For later use let us also record the inverse metrics
2765: \begin{eqnarray}\label{Ginvers}
2766: G^{\alpha \beta}
2767: = -\frac{2}{3} \KK \KK^{\alpha \beta} + 2 v^\alpha v^\beta\ ,
2768: \qquad
2769: G^{a b} =
2770: - \frac{2}{3} \KK \KK^{a b}\ ,
2771: \end{eqnarray}
2772: where $\KK^{\alpha \beta}$ and $\KK^{a b}$ are the inverse
2773: of $\KK_{\alpha \beta}$ and
2774: $\KK_{a b}$, respectively.
2775:
2776: The $N=2$ hypermultiplet couplings are reduced by inserting \eqref{constr} - \eqref{Ginvers}
2777: and truncating to the orientifold spectrum as summarized in table \ref{N=1spectrumtab}.
2778: Since this the orientifold spectrum of $O3/O7$ setups differs from the one
2779: of $O5/O9$ setups, one obtains two different effective actions.
2780: Together with the standard Einstein-Hilbert term and the contributions
2781: from the reduction of the $N=2$ vectors \eqref{red-vector} one finds after
2782: Weyl rescaling \cite{TGL1}
2783: \begin{eqnarray} \label{S_scalarO3}
2784: S^{(4)}_{O3/O7} &=&\int -\tfrac{1}{2} R *\mathbf{1}
2785: - G_{k \bar l} \; dz^{k} \wedge *d\bar z^{l}
2786: -G_{\alpha \beta} \; dv^\alpha \wedge *dv^\beta - G_{ab}\; db^a \wedge * db^b \nonumber \\
2787: && - dD \wedge * dD
2788: -\tfrac{1}{24}e^{2 D}\cK\, dl \wedge * dl - \tfrac{1}{6}e^{2D} \cK
2789: G_{ab}\left(dc^a-l db^a \right) \wedge *\left(dc^b-l db^b \right)
2790: \nonumber \\
2791: &&-\tfrac{3}{8 \cK}e^{2D} G^{\alpha \beta} \Big(d\rho_\alpha-
2792: \KK_{\alpha a b} c^a db^b \Big)
2793: \wedge
2794: *\Big(d\rho_\beta - \KK_{\beta cd} c^c db^d
2795: \Big) \nn \\
2796: &&+\tfrac{1}{4}\text{Im}\; \cM_{\kappa \lambda}\;
2797: F^{\kappa}\wedge *F^{\lambda}
2798: +\tfrac{1}{4}\text{Re}\; \cM_{\kappa \lambda}\;
2799: F^{\kappa}\wedge F^{\lambda}\ ,
2800: \end{eqnarray}
2801: and
2802: \begin{eqnarray} \label{S_scalarO5}
2803: S^{(4)}_{O5/O9}&=&\int -\tfrac{1}{2} R *\mathbf{1}
2804: - G_{\kappa \bar \lambda} \; dz^{\kappa} \wedge *d\bar z^{\lambda}
2805: - G_{\alpha \beta} \; dv^\alpha \wedge *dv^\beta
2806: \nn \\
2807: &&- G_{ab}\; db^a \wedge * db^b -d D \wedge * dD - \tfrac{1}{6}e^{2D} \cK G_{\alpha \beta}\; dc^\alpha \wedge * dc^\beta
2808: \nn \\
2809: &&- \tfrac{3}{2\KK}e^{2D}(dh+\tfrac{1}{2}(d\rho_a b^a -\rho_a db^a))
2810: \wedge *(dh+\tfrac{1}{2}(d\rho_a b^a - \rho_a db^a))\nonumber \\
2811: &&- \tfrac{3}{8 \cK}e^{2D}G^{ab}(d \rho_a - \KK_{ac\alpha} c^\alpha db^c)
2812: \wedge *(d \rho_b - \KK_{bd\beta} c^\beta db^d)\ . \nn\\
2813: &&+\tfrac{1}{4}\text{Im}\; \cM_{k l}\;
2814: F^{k}\wedge *F^{l}
2815: +\tfrac{1}{4}\text{Re}\; \cM_{k l}\;
2816: F^{k}\wedge F^{l}\ ,
2817: \end{eqnarray}
2818: where we have expressed the result in a chiral basis and used the index conventions given in table \ref{CYObasisB}. In
2819: contrast to ref. \cite{TGL1} we have expressed the effective actions
2820: in terms of the string frame K\"ahler structure deformations $v^\alpha$ and
2821: the four-dimensional dilaton
2822: \beq \label{def-D_B}
2823: e^D = e^\phi\ (\cK/6)^{-1/2}\ ,
2824: \eeq
2825: where $e^\phi$ is the ten-dimensional dilaton.
2826: This ends our computation of the orientifold bulk action. In remains to cast
2827: \eqref{S_scalarO3} and \eqref{S_scalarO5} into the standard $N=1$ form.
2828:
2829:
2830: \subsection{The K\"ahler potentials and gauge-couplings}
2831: \label{Kpo_gaugeIIB}
2832:
2833: Our next task is to transform the actions \eqref{S_scalarO3} and \eqref{S_scalarO5} into the standard
2834: $N=1$ supergravity form with chiral multiplets where it is
2835: expressed in terms of a K\"ahler potential $K$,
2836: a holomorphic superpotential $W$ and the holomorphic gauge-kinetic coupling
2837: functions $f$ as follows \cite{WB,GGRS}
2838: \beq\label{N=1action}
2839: S^{(4)} = -\int \tfrac{1}{2}R * \mathbf{1} +
2840: K_{I \bar J} DM^I \wedge * D\bar M^{\bar J}
2841: + \tfrac{1}{2}\text{Re}f_{\kappa \lambda}\
2842: F^{\kappa} \wedge * F^{\lambda}
2843: + \tfrac{1}{2}\text{Im} f_{\kappa \lambda}\
2844: F^{\kappa} \wedge F^{\lambda} + V*\mathbf{1}\ ,
2845: \eeq
2846: where
2847: \beq\label{N=1pot}
2848: V=
2849: e^K \big( K^{I\bar J} D_I W {D_{\bar J} \bar W}-3|W|^2 \big)
2850: +\tfrac{1}{2}\,
2851: (\text{Re}\; f)^{-1\ \kappa\lambda} D_{\kappa} D_{\lambda}
2852: \ .
2853: \eeq
2854: Here the $M^I$ collectively denote all
2855: complex scalars in chiral multiplets present in the theory and
2856: $K_{I \bar J}$ is a K\"ahler metric satisfying
2857: $ K_{I\bar J} = \partial_I \bar\partial_{\bar J} K(M,\bar M)$.
2858: The scalar potential is expressed in terms of the
2859: K\"ahler-covariant derivative $D_I W= \partial_I W +
2860: (\partial_I K) W$.
2861:
2862: In the reduction we did not find any scalar potential, such that one immediately concludes
2863: $W=0$ and $D_\kappa=0$. Next we need to find a complex structure on the space of
2864: scalar fields such that the metrics computed in \eqref{S_scalarO3} and \eqref{S_scalarO5}
2865: are manifestly K\"ahler.
2866:
2867: \subsubsection{The K\"ahler potential: $O3/O7$ setups}
2868:
2869: As we saw in \eqref{csmetrico} the
2870: complex structure deformations $z^{k}$ are already good K\"ahler
2871: coordinates with $G_{k\bar l}$ being the appropriate K\"ahler metric.
2872: For the remaining fields the definition of
2873: the K\"ahler coordinates is not so obvious.
2874: Guided by refs.\ \cite{HL,BBHL} we define
2875: \beq \label{def-coordsO3}
2876: \fe - i\, \fa = i\tau + iG^a \omega_a - T_\alpha \tilde \omega^\alpha
2877: \eeq
2878: where
2879: \beq \label{def-A}
2880: \pev = \fe + i\, \feh =e^{-\phi} e^{-\hat B_2 + iJ}\ ,\qquad \fa = e^{-\hat B_2} \wedge \sum_{q=0,2,4,6} \hat C_q |_{scalar}\ ,
2881: \eeq
2882: are sums of even forms.
2883: In \eqref{def-A} we have defined $\hat C_q |_{scalar}$ to be the part of $\hat C_q$ yielding scalars in $D=4$,
2884: e.g.~$\hat C_4|_{scalar} = \rho_\alpha\, \tilde \omega^\alpha$. Expanding all the forms in
2885: \eqref{def-coordsO3} by using \eqref{def-A},\eqref{exp-B} and \eqref{exp1}
2886: the coordinates take the form \cite{TGL1}
2887: \bea \label{tau}
2888: \tau &=& C_0+ie^{-\phi} \ , \qquad G^a =c^a-\tau b^a\ ,\nn \\
2889: T_\alpha &=& i( \rho_\alpha - \tfrac{1}{2} \cK_{\alpha ab}c^a b^b) + \tfrac{1}{2} e^{-\phi} \cK_\alpha
2890: - \zeta_\alpha\ ,
2891: \eea
2892: where\footnote{The definition of $\zeta_\alpha$ is unique up to a constant
2893: which does not enter into the metric. The possibility of a non-zero constant
2894: is important for the formulation in terms of linear multiplets in
2895: section~\ref{IIB_lin}.}
2896: \beq\label{zetadef}
2897: \KK_{\alpha} = \KK_{\alpha\beta\gamma} v^\beta v^\gamma \ ,\qquad
2898: \zeta_\alpha = -\frac{i}{2(\tau-\bar \tau)}\ \KK_{\alpha b c}G^b (G- \bar G)^c\ .
2899: \eeq
2900: In ref.~\cite{TGL1} it was checked explicitly that in terms of these coordinates
2901: the metric of \eqref{S_scalarO3} is K\"ahler with the K\"ahler potential \cite{TGL1}
2902: \beq \label{kaehlerpot-O7-1}
2903: K = K_{\rm cs}(z,\bar z)
2904: + K^{\rm Q}(\tau,T,G)\ , \qquad K_{\rm cs} = - \text{ln}\Big[-i\int_Y \Omega(z) \wedge \bar \Omega(\bar z) \Big]\ ,
2905: \eeq
2906: and
2907: \beq\label{kaehlerpot-Kk}
2908: K^{\rm Q} = -\text{ln}\big[-i(\tau - \bar \tau)\big]
2909: - 2 \text{ln}\big[ \text{Vol}_E(\tau,T,G)\big]=-\ln\big[2 e^{-4D}\big] \ ,
2910: \eeq
2911: where we have used \eqref{def-D_B} in order to evaluate the last equality.
2912: The Einstein frame volume
2913: $\text{Vol}_E (Y) = \frac{1}{6} e^{-\frac{3}{2} \phi} \KK_{\alpha\beta\gamma} v^\alpha v^\beta v^\gamma$
2914: in \eqref{kaehlerpot-Kk}
2915: should be understood
2916: as a function of the K\"ahler coordinates $(\tau,T,G)$
2917: which enter by solving (\ref{tau}) for $e^{-\phi/2} v^\alpha$ in terms of $(\tau,T,G)$.
2918: Unfortunately this solution cannot be given explicitly and therefore $\text{Vol}_E$ is known
2919: only implicitly via $e^{-\phi/2} v^\alpha(\tau,T,G)$.\footnote{This is in complete analogy
2920: to the situation encountered in compactifications
2921: of M-theory on Calabi-Yau fourfolds studied in \cite{HL}.
2922: This is no coincidence and can be understood from the fact
2923: that this theory can be lifted to F-theory on Calabi-Yau fourfolds
2924: which in a specific limit is related to orientifold
2925: compactifications of type IIB \cite{Sen}. In section \ref{F-theory}
2926: we make this more explicit by checking this correspondence on the level of the
2927: effective actions.}
2928: In chapter \ref{lin_geom_of_M} we show that the definition of the
2929: K\"ahler coordinates \eqref{tau} and the K\"ahler potential
2930: \eqref{kaehlerpot-O7-1} can be understood somewhat
2931: more conceptually in a dual formalism using linear multiplets
2932: $L^\alpha$ instead of the chiral multiplets $T_\alpha$.
2933:
2934: Let us return to the K\"ahler potential (\ref{kaehlerpot-O7-1}).
2935: $K_{cs}$ and the first term in \eqref{kaehlerpot-Kk} are the standard
2936: K\"ahler potentials for the complex structure deformations
2937: and the dilaton, respectively.
2938: $\text{Vol}_E(\tau,G,T)$ also depends on $\tau$ and therefore the metric mixes $\tau$
2939: with $T_\alpha$ and $G^a$. It is block diagonal in the
2940: complex structure deformations which do not mix with the other scalars.
2941: Hence, the moduli space locally has the form
2942: \beq \label{modulispaceO3}
2943: \cM_{N=1} = \tilde \cM^{\rm SK} \times\, \tilde \cM^{\rm Q}\ ,
2944: \eeq
2945: where each factor is a K\"ahler manifold. The manifold $ \tilde \cM^{\rm SK}$
2946: has complex dimension $h_-^{(1,2)}$ and is a special K\"ahler manifold in
2947: that $K_{\rm cs}$ satisfies \eqref{csmetrico}. It parameterizes the complex
2948: structure deformations of $Y$ respecting the orientifold constraint \eqref{Omegatransf}.
2949: On the other hand, $\tilde \cM^{\rm Q}$ is a $h^{(1,1)}+1$-dimensional K\"ahler manifold inside
2950: the quaternionic manifold $\cM^{\rm Q}$. Local coordinates are given by the
2951: fields $\tau, G^a,T_\alpha$ arising in the expansion \eqref{def-coordsO3}.
2952: Also the K\"ahler potential $K^{\rm Q}(\tau, G,T)$ fulfills
2953: special properties inherited from the underlying
2954: special quaternionic manifold.
2955: To see this, let us bring $K^{\rm Q}$ in a slightly different form.
2956: Using the explicit expansion \eqref{def-coordsO3} of $\pev$ one checks
2957: that up to a trivial K\"ahler transformation the K\"ahler potential
2958: \eqref{kaehlerpot-Kk} can be rewritten as
2959: \beq \label{def-Phi}
2960: K^{\rm Q} = -2\ln\, \Phi_B(\fe)\ , \qquad \Phi_B(\fe)\equiv i\big<\pev, \pevb \big>\ ,
2961: \eeq
2962: where $\pev=\fe+i\feh$ is defined in \eqref{def-A} and $\feh(\fe)$ has
2963: to be evaluated. In \eqref{def-Phi}
2964: we abbreviated the skew-symmetric product $\big<\varphi, \psi \big>$
2965: for two sums of even forms $\varphi=\varphi_0+\varphi_2 + \varphi_4 + \varphi_6$
2966: and $\psi=\psi_0+\psi_2+\psi_4+\psi_6$ as \cite{HitchinGCM}
2967: \beq \label{symp-form}
2968: \big<\varphi, \psi \big> = \int_Y\ \sum_{m} (-1)^{m} \varphi_{2m} \wedge \psi_{\,6-2m}\ .
2969: %= \int_Y \varphi_0\wedge \psi_6 - \varphi_2\wedge \psi_4 + \varphi_4\wedge \psi_2
2970: % -\varphi_6\wedge \psi_0\ .
2971: \eeq
2972: The function $\Phi_B$ can be identified with Hitchins functional on
2973: a generalized complex manifold \cite{HitchinGCM} evaluated for the simple form $\pev$ defined
2974: in \eqref{def-A} (see \cite{GLprep} for more details).
2975: We discuss the geometry of $\tilde \cM^{\rm Q}$ in greater detail in section \ref{geom_of_modspace}.
2976:
2977: Although not immediately obvious from its definition
2978: $K^{\rm Q}$ obeys a no-scale type condition in that it satisfies
2979: \beq\label{NScond}
2980: \frac{\partial K}{\partial N^I}\, (K^{-1})^{ I\bar J}\,
2981: \frac{\partial K}{\partial \bar N^{\bar J}} = 4\ ,
2982: \eeq
2983: where $N^I = (\tau,G^a,T_\alpha)$.\footnote{For $G^a=0$
2984: this has already been observed in \cite{GKP,BBHL,DWG,DAFT}.}
2985: This equality can be shown by direct computation as done in \cite{TGL1}.
2986: Alternatively, it can be deduced from the fact that $\Phi_B$ defined in
2987: \eqref{def-Phi} is homogeneous of degree two, i.e.~$\Phi_B(a\, \fe) = a^2\, \Phi_B(\fe)$ for all
2988: $a\in \bbR $ \cite{HitchinGCM}. Using \eqref{def-coordsO3} a
2989: simple calculation shows that $K^{\rm Q}=-2\ln \Phi_B$ satisfies \eqref{NScond}.
2990: From \eqref{N=1pot} we see that
2991: \eqref{NScond} implies $V\ge 0$ which we also show
2992: in the linear multiplet formalism in section \ref{IIB_lin}.
2993: For $\tau=\text{const.}$ the right hand side of \eqref{NScond} is found to
2994: be equal to $3$ as it is the case in the standard no-scale K\"ahler potentials
2995: of \cite{NS}.
2996:
2997: Let us relate \eqref{kaehlerpot-O7-1} to the known K\"ahler potentials
2998: in the literature.
2999: First of all, for $G^a=0$ and thus
3000: $T_\alpha=i \rho_\alpha + \frac{1}{2}\KK_\alpha $
3001: the K\"ahler potential
3002: \eqref{kaehlerpot-O7-1} reduce to the one given in \cite{BBHL}.
3003: Secondly, for one overall K\"ahler modulus $v$ parameterizing the volume
3004: (i.e. for $h^{(1,1)}_+=1$, $T_{\alpha}\equiv T$)
3005: the K\"ahler potential $K^{\rm Q}$ reduces to
3006: $K=-3 \text{ln}( T +\bar T )$
3007: which coincides with the K\"ahler potential
3008: determined in \cite{GKP}.
3009:
3010: Before we turn to the discussion of the $O5/O9$ case let us note that
3011: $K$ is invariant under the $SL(2,\bbR)$ transformations inherited from the
3012: ten-dimensional type IIB theory. In the orientifold theory
3013: this symmetry acts on $\tau$ by fractional linear transformations
3014: exactly as in $D=10$ and transforms $(b^a,c^a)$ as a doublet, such that
3015: \beq \label{Sl2}
3016: \tau \to \frac{a\tau + b}{c\tau +d}\ ,\qquad G^a \to \frac{G^a}{c\tau +d}\ , \qquad ad-bc=1\ .
3017: \eeq
3018: Under the $SL(2,{\bf R})$ only the second term of $K$
3019: given in \eqref{kaehlerpot-Kk}
3020: transforms but this transformation is just a K\"ahler transformation.
3021: This can be seen from \eqref{tau} and the fact that $e^{-\phi/2} v^\alpha$ and $z^k$ are invariant.
3022: This symmetry reduces to $SL(2,\bbZ)$ in the full string theory, which is nothing
3023: but the invariance group of a two-torus. This torus becomes part of the space-time
3024: in the formulation of `F-theory' \cite{Vafa}. We discuss in section \ref{F-theory}
3025: the embedding of $O3/O7$ orientifolds into this theory on the level of the effective action.
3026:
3027:
3028: \subsubsection{The K\"ahler potential: $O5/O9$ setups}
3029:
3030: In the action \eqref{S_scalarO5} we immediately see that the complex
3031: structure deformations $z^\kappa$ are again already good K\"ahler coordinates.
3032: For the remaining fields we find the appropriate K\"ahler coordinates
3033: to be
3034: \beq \label{def-coordsO5}
3035: \feh - i \fa\ =\ t^\alpha \omega_\alpha - A_b\, \tilde \omega^b - S\, \text{vol}(Y)\ ,
3036: \eeq
3037: where $\feh=\I\, \pev$ and $\fa$ are defined in \eqref{def-A} and we have used that in
3038: $O5/O9$ setups the axion $C_0$ gets projected out. Furthermore, we denoted by
3039: $\text{vol}(Y)=\cK^{-1} J\wedge J \wedge J$ the to one normalized volume form of $Y$.
3040: Using
3041: the expansions \eqref{transJo}, \eqref{exp-B}
3042: and \eqref{expO5} we obtain the explicit expressions \cite{TGL1}
3043: \begin{eqnarray}\label{Kcoord}
3044: t^\alpha &=& e^{-\phi} v^\alpha - i c^\alpha\ , \qquad A_a \ =\ \N_{ab} b^b + i \rho_a \ ,\\
3045: S & = & \tfrac{1}{6} e^{-\phi}\, \cK + i h
3046: - \tfrac{1}{4} (\text{Re}\N^{-1})^{ab} A_a (A+\bar A)_b\ , \nn
3047: \end{eqnarray}
3048: where we inserted
3049: \beq\label{Ndef}
3050: \N_{ab}(t) \equiv \KK_{ab \alpha} t^\alpha \ ,\qquad \int C_6 = h+\tfrac{1}{2} \rho_a b^a\ .
3051: \end{equation}
3052: The matrix $\N_{ab}$ depends holomorphically on the coordinates $t^\alpha$ which
3053: ensures that $\tilde \cM^{\rm Q}$ is K\"ahler \cite{FS,HL}.
3054: In the variables given in \eqref{Kcoord} the K\"ahler potential reads \cite{TGL1}
3055: \beq \label{O5-Kaehlerpot}
3056: K \ =\ K_{cs}(z,\bar z) + K^{\rm Q}(S,t,A)\ , \qquad K_{cs} = -\text{ln}\Big[-i\int\Omega \wedge \bar \Omega \Big]
3057: \eeq
3058: with
3059: \bea \label{kaehlerpot-KkO5}
3060: K^{\rm Q}& =& - \text{ln}\Big[\tfrac{1}{48}\KK_{\alpha \beta \gamma}(t+\bar t)^\alpha
3061: (t+\bar t)^\beta (t+\bar t)^\gamma \Big]\nn \\
3062: && - \text{ln}\Big[S + \bar S + \tfrac{1}{4} (A + \bar A)_a (\text{Re}\N^{-1})^{ab}
3063: (A + \bar A)_b \Big]\\
3064: &=& -\ln\big[2 e^{-4D}\big]\ . \nn
3065: \eea
3066: where we used \eqref{def-D_B}.
3067: The check that $K$ indeed reproduces \eqref{S_scalarO5} is straightforward, since
3068: \eqref{O5-Kaehlerpot} is closely related to the quaternionic
3069: `K\"ahler potential' given in \cite{FS} and we can make use
3070: of their results.\footnote{Note however,
3071: that the complex structure changed non-trivially.
3072: In \cite{FS} the standard $t \sim v + i b$ formed complex coordinates.}
3073: The same reference already observed
3074: that for a holomorphic matrix $\N$ the quaternionic geometry is also
3075: K\"ahler. This situation was also found in compactifications
3076: of the heterotic string to $D=3$ on a circle \cite{HL}.
3077:
3078: {}From \eqref{O5-Kaehlerpot} we infer that the $N=1$ moduli space admits
3079: the local product structure $\tilde \cM^{\rm SK}\times \tilde \cM^{\rm Q}$
3080: similar to \eqref{modulispaceO3}. However, in $O5/O9$ orientifolds $\tilde \cM^{\rm SK}$
3081: is a special K\"ahler manifold spanned by the $h^{(2,1)}_+$ complex scalars $z^\kappa$, which
3082: are the ones projected out in $O3/O7$ orientifolds. $\tilde \cM^{\rm Q}$ is spanned by the complex scalars
3083: $S,t^\alpha,A_a$ and thus is of complex dimension $h^{1,1} + 1$ as in $O3/O7$ setups.
3084: Furthermore, also $K^{\rm Q}$ for orientifolds with $O5/O9$ planes can be rewritten in terms
3085: of the functional $\Phi_B(\feh)$ as
3086: \beq \label{def-PhiO5}
3087: K^{\rm Q}\ =\ -2 \ln \Phi_B(\feh)\ ,\qquad \Phi_B(\feh ) \equiv i\big<\pev,\pevb \big>\ ,
3088: \eeq
3089: where $\pev=\fe(\feh)+i\feh$ are defined in \eqref{def-A}.
3090: The functional dependence of $K^{\rm Q}$ on $\pev$ is the same as in \eqref{def-Phi} for $O3/O7$ orientifolds.
3091: This can be understood from the fact that $\pev$ only depends on the NS-NS sector
3092: variables, which are the same in both types of orientifolds.
3093: Nevertheless, the local structure of $\tilde \cM^{\rm Q}$ is different for both orientifold theories.
3094: This becomes appearent when one expresses $K^{\rm Q}$ in terms of
3095: proper K\"ahler coordinates. In $O5/O9$ setups this corresponds to the fact that
3096: $\Phi_B$ is a function of $\feh$ as needed for \eqref{def-coordsO5}. Hence, in order
3097: to express $K^{\rm Q}$ in terms of the K\"ahler coordinates $S,t,A$ as in \eqref{kaehlerpot-KkO5}
3098: one evaluates $\fe(\feh)$. Let us end this discussion by remarking that
3099: $\Phi_B$ is also homogeneous of degree two in $\feh$, such that
3100: by using \eqref{def-coordsO5} one extracts a no-scale type condition equivalent to \eqref{NScond}.
3101:
3102:
3103: \subsubsection{The gauge-couplings: $O3/O7$ and $O5/O9$ setups}
3104:
3105: Our next task is to determine
3106: the gauge-kinetic coupling functions $f_{\kappa \lambda}$
3107: and show that they are holomorphic in the moduli. We do
3108: this only for $O3/O7$ orientifolds, since the result easily
3109: translates to the $O5/O9$ case. As explained in section
3110: \ref{O37_effective_act} this is achieved by an appropriate replacement of the
3111: indices.
3112: By comparing the actions \eqref{red-vector} and \eqref{N=1action} one finds
3113: \beq \label{gauge-couplingsO3}
3114: f_{\kappa \lambda}
3115: =-\tfrac{i}{2} \,
3116: \bar{\cM}_{\kappa \lambda}\Big|_{z^{\kappa}=0=\bar z^{\kappa}}\ ,
3117: \eeq
3118: where
3119: ${\cM}_{\kappa \lambda}$ is the
3120: $N=2$ gauge kinetic matrix given in \eqref{defM}
3121: evaluated at ${z^{\kappa}=\bar z^{\kappa}}=0$.
3122: Its holomorphicity in the complex structure deformations $z^k$ is not
3123: immediately obvious but can be shown by using
3124: \eqref{defM} and \eqref{gauge-c}.
3125: More precisely, from \eqref{defM} together with
3126: the decomposition of $H^{(3)}$ expressed by \eqref{H3split}
3127: and \eqref{sbasiso} we infer that ${\cM}_{\hat K \hat L}$
3128: is block diagonal or in other words
3129: ${\cM}_{\kappa \hat l} = 0.$ Multiplying ${\cM}_{\kappa \hat l}$
3130: with $X^{\hat l}$ and using $X^\lambda=0$ together with
3131: \eqref{gauge-c} we further conclude
3132: \beq\label{Fdiag}
3133: {\cF}_{\kappa \hat l}\Big|_{z^{\kappa}=0=\bar z^{\kappa}}=0\ .
3134: \eeq
3135: Finally inserting \eqref{cond-1} and \eqref{Fdiag}
3136: into \eqref{gauge-c}
3137: we arrive at \cite{TGL1}
3138: \begin{eqnarray}\label{fholo}
3139: f_{\kappa \lambda} (z^k)
3140: =-\tfrac{i}{2}
3141: \mathcal{F}_{\kappa \lambda}\Big|_{z^{\kappa}=0=\bar z^{\kappa}}\ ,
3142: \end{eqnarray}
3143: which is manifestly holomorphic since $\mathcal{F}_{\kappa \lambda}(z^k)$
3144: are holomorphic functions of the complex structure
3145: deformations $z^k$.
3146:
3147:
3148:
3149:
3150:
3151:
3152:
3153:
3154:
3155:
3156:
3157:
3158: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3159: %
3160: % IIA: The Effective action
3161: %
3162: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3163:
3164:
3165:
3166:
3167: \section{Type IIA Calabi-Yau orientifolds}
3168: \label{IIA_orientifolds}
3169:
3170: In this section we determine the $N=1$ supergravity action
3171: obtained by compactification of Type IIA string theory on
3172: a Calabi-Yau orientifold. The orientifold projection
3173: $\cO=(-1)^{F_L} \Omega_p \sigma$ was already defined in
3174: \eqref{oproj} and includes an anti-holomorphic isometric involution
3175: $\sigma$. In section \ref{spectrum-IIA} we extract the $N=1$ spectrum by
3176: identifying the fields invariant under $\cO$. The corresponding
3177: effective action is calculated in section \ref{eff_actIIA}.
3178: It is shown to be compatible with $N=1$ supersymmetry in section \ref{Kpo_gaugeIIA},
3179: where we determine the K\"ahler potential and gauge-kinetic coupling functions.
3180:
3181:
3182: \subsection{The $N=1$ spectrum \label{spectrum-IIA}}
3183:
3184: In order to determine the $\mathcal{O}$-invariant states let us recall
3185: that the ten-dimensional RR forms $\hat C_1$ and $\hat C_3$
3186: are odd
3187: under $(-1)^{F_L}$ while all other fields are even.
3188: Under the world-sheet
3189: parity $\Omega_p$ on the other hand $\hat B_2, \hat C_3$ are odd
3190: with all other fields being even.
3191: As a consequence the
3192: $\mathcal{O}$-invariant states have to satisfy
3193: \cite{BH}
3194: \begin{equation} \label{fieldtransf}
3195: \begin{array}{lcl}
3196: \sigma^* \hat \phi &=& \ \hat \phi\ , \\
3197: \sigma^* \hat g &=& \ \hat g\ , \\
3198: \sigma^* \hat B_2 &=& - \hat B_2\ ,
3199: \end{array}
3200: \hspace{2cm}
3201: \begin{array}{lcl}
3202: \sigma^* \hat C_1 &=& - \hat C_1\ , \\
3203: \sigma^* \hat C_3 &=& \ \hat C_3\ ,
3204: \end{array}
3205: \end{equation}
3206: while the deformations of the Calabi-Yau metric are constrained
3207: by \eqref{constrJ} and \eqref{constrO}.\footnote{%
3208: Following the argument presented in
3209: \cite{BH} we note that the involution does not change
3210: under deformations of $Y$.
3211: This is due to its involutive property
3212: and the fact that we identify involutions which differ
3213: by diffeomorphisms.
3214: Therefore we fix an involution and restrict the deformation space by demanding
3215: \eqref{constrJ} and \eqref{constrO}. }
3216:
3217:
3218: As we recalled in the previous section the massless modes are in one-to-one
3219: correspondence with the harmonic forms on $Y$. The space of harmonic forms
3220: splits under the involution $\sigma$ into even and odd eigenspaces
3221: \beq \label{cohom-split}
3222: H^p(Y)\ =\ H^p_+ \oplus H^p_-\ \ .
3223: \eeq
3224: Depending on the transformation properties given in \eqref{fieldtransf}
3225: the $\mathcal{O}$-invariant states reside either in $H^p_+$ or in $H^p_-$
3226: and as a consequence the number of states is reduced.
3227: We summarize all non-trivial cohomology groups including their basis elements
3228: in table \ref{CYObasis}.\\
3229:
3230: \begin{table}[h]
3231: \begin{center}
3232: \begin{tabular}{| c || c | c| c | c | c | c |} \hline
3233: \rule[-0.3cm]{0cm}{0.9cm} cohomology group & $\ H^{(1,1)}_+\ $ &
3234: $\ H^{(1,1)}_-\ $ & $\ H^{(2,2)}_+\ $ & $\ H^{(2,2)}_-\ $ & $\ H^{(3)}_+\ $ & $\ H^{(3)}_-\ $
3235: \\ \hline
3236: \rule[-0.3cm]{0cm}{0.8cm} dimension & $h^{(1,1)}_+$ & $h^{(1,1)}_- $
3237: & $h^{(1,1)}_-$ & $h^{(1,1)}_+$
3238: & $h^{(2,1)}+1$ & $h^{(2,1)}+1$
3239: \\ \hline
3240: \rule[-0.3cm]{0cm}{0.8cm} basis & $\omega_\alpha$ & $\omega_a$
3241: & $\tilde \omega^a$ & $\tilde \omega^\alpha$
3242: & $ a_{\Kh}$ & $b^{\Kh}$ \\ \hline
3243: \end{tabular}
3244: \caption{\small \label{CYObasis}
3245: \textit{Cohomology groups and their basis elements.}}
3246: \end{center}
3247: \end{table}
3248:
3249: $\omega_\alpha, \omega_a$ denote
3250: even and odd $(1,1)$-forms while
3251: $\tilde\omega^\alpha, \tilde\omega^a$ denote odd and even
3252: $(2,2)$-forms. The number of even $(1,1)$-forms is equal to the number
3253: of odd $(2,2)$-forms and vice versa since the
3254: volume form which is
3255: proportional to $J \wedge J \wedge J$ is odd and thus Hodge duality
3256: demands $h^{(1,1)}_+ = h^{(2,2)}_- ,\ h^{(1,1)}_- = h^{(2,2)}_+$.
3257: This can also be seen from the fact that the non-trivial
3258: intersection numbers are
3259: \beq \label{basis-int}
3260: \int \omega_\alpha \wedge \tilde \omega^\beta =
3261: \delta^{\beta}_\alpha\ ,\quad \alpha,\beta = 1, \ldots, h^{(1,1)}_+\ ,
3262: \qquad
3263: \int \omega_a \wedge \tilde \omega^b = \delta^{b}_a\ , \quad
3264: a,b=1,\ldots,h^{(1,1)}_-\ ,
3265: \eeq
3266: with all other pairings vanishing.
3267: {}From the volume-form being odd
3268: one further infers $h^{(3,3)}_+=0,$ $h^{(3,3)}_-=1$ and
3269: $h^{(0,0)}_+=1,\ h^{(0,0)}_-=0$.
3270:
3271:
3272: $H^{3}$ can be decomposed
3273: independently of the complex structure as
3274: $H^{3}=H^3_+ \oplus H^3_-$ where
3275: the (real) dimensions of both $H^3_+$ and $H^3_-$
3276: is equal and given by $h^{3}_+ =h^{3}_-= h^{(2,1)}+1$.
3277: Again this is a consequence of Hodge duality
3278: together with the fact that the volume-form is odd.
3279: It implies that for each element $a_\Kh \in H^3_+$
3280: there is a dual element $b^\Lh \in H^3_-$
3281: with the intersections
3282: \beq \label{basis_ab}
3283: \int a_\Kh \wedge b^\Lh = \delta^\Lh_\Kh \ , \qquad
3284: \Kh, \Lh = 0,\ldots, h^{(2,1)}\ .
3285: \eeq
3286: Compared to \eqref{int-numbers1} this amounts to a symplectic rotation
3287: such that all $\alpha$-elements are chosen to be even and
3288: all $\beta$-elements are chosen to be odd but with the intersection
3289: numbers unchanged.
3290: The orientifold projection breaks this symplectic invariance
3291: or in other words fixes a particular symplectic gauge
3292: which groups all basis elements into even and odd.
3293: This in turn implies that the basis $(a_\Kh,b^\Kh)$ is only one possible choice.
3294: However, since the calculation simplifies considerably for this basis, we first restrict
3295: to this special case and later give the general results with calculations summarized in
3296: section \ref{IIA_lin}.
3297:
3298:
3299:
3300: In the remainder of this subsection we determine the $N=1$ spectrum
3301: which survives the orientifold projections.
3302: Let us first discuss the K\"ahler moduli.
3303: From the eqs. \eqref{constrJ} and \eqref{fieldtransf} we see that both
3304: $J$ and $\hat B_2$ are odd and hence have to be expanded
3305: in a basis $\omega_a$ of odd harmonic $(1,1)$-forms
3306: \beq \label{expJB}
3307: J\ =\ v^a(x)\, \omega_a\ ,\qquad \hat B_2\ =\ b^a(x)\, \omega_a\ , \qquad a = 1,\ldots, h^{(1,1)}_-\ .
3308: \eeq
3309: In contrast to \eqref{fieldexp} the four-dimensional
3310: two-form $B_2$ gets projected out due to \eqref{fieldtransf} and the fact
3311: that $\sigma$ acts trivially on the flat dimensions.
3312: $v^a$ and $b^a$ are space-time scalars and
3313: as in $N=2$ they can be combined into complex coordinates
3314: \beq \label{def-t}
3315: t^a = b^a + i\, v^a\ , \qquad \Jc = B_2 + i J\ ,
3316: \eeq
3317: where we have also introduced the complexified
3318: K\"ahler form $\Jc$.
3319: We see that in terms of the field variables the same complex
3320: structure is chosen as in $N=2$ but the dimension of the K\"ahler moduli
3321: space is truncated from $h^{(1,1)}$ to $h^{(1,1)}_-$.
3322:
3323:
3324: The number of complex structure deformations is similarly reduced since
3325: \eqref{constrO} constrains the possible deformations.
3326: To see this one performs a symplectic rotation on
3327: \eqref{Omegaexp} and expands $\Omega$ in the basis of
3328: $H^p_+ \oplus H^p_-$, i.e.\ as\footnote{Let us stress that at this
3329: point all $N=2$ relations are still intact since \eqref{Omegapm}
3330: is just a specific choice of the standard $N=2$ basis \eqref{Omegaexp}.}
3331: \beq\label{Omegapm}
3332: \Omega(z) = Z^\Kh(z)\, a_\Kh - \cF_\Lh(z)\, b^\Lh\ .
3333: \eeq
3334: Inserted into \eqref{constrO} one finds
3335: \bea \label{Z=0}
3336: \I(e^{-i\theta} Z^\Kh)\ =\ 0\ , \qquad
3337: \R(e^{-i\theta} \cF_\Kh )\ = 0\ .
3338: \eea
3339: The first set of equations are $h^{(2,1)}+1$ real conditions
3340: for $h^{(2,1)}$ complex scalars $z^K$.
3341: One of these equations is redundant due to the
3342: scale invariance \eqref{crescale} of $\Omega$.
3343: More precisely, the phase of $e^{-h}$ can be used to
3344: trivially satisfy $\I(e^{-i\theta} Z^\Kh)= 0$ for one of the $Z^\Kh$.
3345: Thus $\I(e^{-i\theta} Z^\Kh)=0$
3346: projects out $h^{(2,1)}$ real scalars, i.e.\
3347: half of the complex structure deformations.
3348: Furthermore, in section \ref{eff_actIIA} we will see
3349: the remaining real complex structure deformations
3350: span a Lagrangian submanifold $\cM^{\rm cs}_\bbR$
3351: with respect to the K\"ahler form
3352: inside $\cM^{\rm cs}$.
3353: Note that the second set of equations in \eqref{Z=0}
3354: $\R (e^{-i\theta} \cF_\Kh ) = 0$
3355: should not be read as equations determining
3356: the $z^K$ but is a constraint on the periods (or equivalently
3357: the Yukawa couplings) of the Calabi-Yau
3358: which has to be fulfilled in order to admit an involutive symmetry
3359: with the property \eqref{constrO}.\footnote{This can also be seen
3360: as conditions arising in consistent truncations of
3361: $N=2$ to $N=1$ theories as discussed in ref.\ \cite{ADAF}.}
3362:
3363:
3364:
3365: As we have just discussed
3366: the complex rescaling \eqref{crescale}
3367: is reduced to the freedom of a real rescaling by \eqref{constrO}.
3368: %Thus we restrict to rescalings \eqref{crescale} with $f$ being real
3369: %when restricted to $\cM^{\rm cs}_\bbR$,
3370: Under these transformations $\Omega$ and the K\"ahler potential $\Kcs$
3371: change as
3372: \beq \label{real_K}
3373: \Omega\to\Omega\, e^{-\R(h)}\ , \qquad \Kcs\to\Kcs + 2 \R(h)\ ,
3374: \eeq
3375: when restricted to $\cM^{\rm cs}_\bbR$. This freedom can be used to set one of
3376: the $\R (e^{-i\theta}Z^{\Kh})$ equal to one and tells us
3377: that $\Omega$ depends only on $h^{(2,1)}$ real deformation parameters.
3378: However, it will turn out to be more
3379: convenient to leave this gauge freedom intact and define
3380: a complex `compensator' $C=re^{-i\theta}$ with the transformation property
3381: $C\to C e^{\R (h)}$.\footnote{This is reminiscent of the situation
3382: encountered in the computation of the entropy of $N=2$ black holes
3383: where it is also convenient to leave this scale invariance intact \cite{OSV}.}
3384: Later on we will relate $r$ to
3385: the inverse of the four-dimensional dilaton
3386: so that the scale invariant function $C\Omega$ depends on
3387: $h^{(2,1)}+1$ real parameters.
3388: Using \eqref{Omegapm} $C\Omega$ enjoys
3389: the expansion
3390: \beq \label{decompO}
3391: C \Omega\ =\ \R (C Z^\Kh)\, a_\Kh - i\I (C \cF_\Lh)\, b^\Lh\ .
3392: \eeq
3393:
3394:
3395:
3396: We are left with the expansion of the ten-dimensional fields
3397: $\hat C_1$ and $\hat C_3$ into harmonic forms.
3398: {}From \eqref{fieldtransf} we learn that $\hat C_1$ is odd
3399: and so together with the fact that
3400: $Y$ posses no harmonic one-forms
3401: and $\sigma$ acts trivially on the flat dimensions
3402: the entire $\hat C_1$ is projected out. This
3403: corresponds to the fact that the $N=2$ graviphoton $A^0$ is removed
3404: from the gravity multiplet,
3405: which in $N=1$ only consists of the metric $g_{\mu \nu}$ as
3406: bosonic component.
3407: Finally, $\hat C_3$ is even and thus can be expanded according to
3408: \beq \label{form-exp}
3409: \hat C_3 = \cc_3(x)
3410: + A^\alpha(x) \wedge \omega_\alpha + \CC_3\ ,\qquad
3411: \CC_3 \equiv \xi^\Kh(x)\, a_\Kh \ ,
3412: \eeq
3413: where $\xi^\Kh$ are $h^{(2,1)}+1$ real
3414: scalars, $A^\alpha$ are $h^{(1,1)}_+$ one-forms
3415: and $\cc_3$ is a three-form in four dimensions.
3416: $\cc_3$ contains no physical degree of freedom but as we will see
3417: in section~\ref{O6sup} corresponds to a
3418: constant flux parameter in the superpotential.
3419: The real scalars
3420: $\xi^\Kh$ have to combine with the
3421: $h^{(2,1)}$ real complex structure deformations
3422: and the dilaton to form chiral multiplets.
3423: In the next section we will find that the appropriate complex fields
3424: arise from the combination
3425: \beq\label{Omegacdef}
3426: \Omegac\ =\ \CC_3 + 2i\R(C\Omega) \ .
3427: \eeq
3428: Expanding $\Omegac$ in a basis \eqref{basis_ab} of $H^3_+(Y)$
3429: and using \eqref{decompO} and \eqref{form-exp} we have
3430: \beq\label{newO}
3431: \Omegac\ =\ 2 N^\Kh a_\Kh \ ,\qquad
3432: N^\Kh= \tfrac{1}{2} \int \Omega_c\wedge \beta^\Kh =
3433: \tfrac{1}{2}\big(\xi^\Kh + 2i \R (C Z^\Kh)\big)\ .
3434: \eeq
3435: Due to the orientifold projection the two three-forms
3436: $\Omega$ and $C_3$
3437: each lost half of their degrees of freedom and combined
3438: into a new complex three-form $\Omegac$.
3439: As we will show in more detail in the next section
3440: the `good' chiral coordinates in the $N=1$ orientifold
3441: are the periods of $C\Omega$ directly while in $N=2$
3442: the periods agree with the proper field variables only
3443: in special coordinates.
3444:
3445: Let us summarize the resulting $N=1$ spectrum.
3446: It assembles into a gravitational multiplet,
3447: $h^{(1,1)}_+$ vector multiplets and
3448: $(h^{(1,1)}_- + h^{(2,1)}+1)$ chiral multiplets.
3449: We list the bosonic parts of the $N=1$ supermultiplets in table
3450: \ref{N=1spectrumA} \cite{BH}. We see that the $h^{(1,1)}$ $N=2$
3451: vector multiplets split into $h^{(1,1)}_+$ $N=1$ vector multiplets and
3452: $h^{(1,1)}_-$ chiral multiplets while the
3453: $h^{(2,1)}+1$ hypermultiplets are reduced to $h^{(2,1)}+1$ chiral multiplets.
3454:
3455: \begin{table}[h]
3456: \begin{center}
3457: \begin{tabular}{|l|c|c|} \hline
3458: \rule[-0.3cm]{0cm}{0.8cm}
3459: multiplets& multiplicity & bosonic components\\ \hline\hline
3460: \rule[-0.3cm]{0cm}{0.8cm}
3461: gravity multiplet&1&$g_{\mu \nu} $ \\ \hline
3462: \rule[-0.3cm]{0cm}{0.8cm}
3463: vector multiplets& $h_+^{(1,1)}$& $A^{\alpha} $\\ \hline
3464: \rule[-0.3cm]{0cm}{0.8cm}
3465: {chiral multiplets} & $h_-^{(1,1)}$&
3466: $t^a$ \\ \hline
3467: \rule[-0.3cm]{0cm}{0.8cm}
3468: {chiral multiplets}
3469: & $ h^{(2,1)}+1$ &$ N^\Kh$\\
3470: \hline
3471: \end{tabular}
3472: \caption{\label{N=1spectrumA} \textit{$N =1$ spectrum of $O6$ orientifold compactification.}}
3473: \end{center}
3474: \end{table}
3475:
3476:
3477: \subsection{The effective action}
3478: \label{eff_actIIA}
3479:
3480: In this section we calculate the four-dimensional effective action of type
3481: IIA orientifolds by performing a Kaluza-Klein reduction of the
3482: ten-dimensional type IIA action \eqref{10dact} taking the
3483: orientifold constraints into account. Equivalently this amounts to
3484: imposing the orientifold projections on the $N=2$ action of
3485: section~\ref{revIIA}.
3486: %Additionally we discuss the reduced moduli space of K\"ahler and complex
3487: %structure deformations preserving \eqref{constrJ} and \eqref{constrO} in
3488: %somewhat more detail.
3489: Inserting \eqref{expJB}, \eqref{decompO}, \eqref{form-exp} into
3490: the ten-dimensional type IIA action \eqref{10dact} and performing a Weyl
3491: rescaling of the four-dimensional metric
3492: %$g_{\mu \nu} \rightarrow (\cK/6) g_{\mu \nu}$
3493: we find \cite{TGL2}
3494: \bea \label{act1}
3495: S^{(4)}_{O6} &=& \int -\tfrac{1}{2} R*\mathbf{1}
3496: - G_{a b}\, dt^a \wedge * d \bar t^b
3497: + \tfrac{1}{2} \text{Im}\, \cN_{\alpha \beta}\ F^\alpha \wedge * F^\beta
3498: + \tfrac{1}{2} \text{Re}\, \cN_{\alpha \beta}\ F^\alpha
3499: \wedge F^\beta \nn\\
3500: &&
3501: \quad -\, d D \wedge * dD -\, G_{K L}(q)\, dq^K \wedge * dq^L
3502: +\tfrac{1}{2} e^{2D}\, \text{Im}\, \cM_{ \Kh \Lh}\,
3503: d\xi^{\Kh} \wedge * d\xi^{\Lh} \ ,
3504: \eea
3505: where $F^\alpha = dA^\alpha$.
3506: Let us discuss the different couplings appearing in \eqref{act1}
3507: in turn.
3508: Apart from the standard Einstein-Hilbert term the first line arises
3509: from the projection of the $N=2$ vector multiplets action.
3510: As we already observed the orientifold projection reduces the number
3511: of K\"ahler moduli from $h^{(1,1)}$ to $h^{(1,1)}_-$ ($t^A\to t^a$)
3512: but leaves the complex structure on this component of the moduli space
3513: intact. Accordingly the metric $G_{ab}(t)$ is inherited from the
3514: metric $G_{AB}$ of the $N=2$ moduli space
3515: $\cM^{SK}$ given in \eqref{Kmetric}.
3516: Since the volume form is odd only intersection numbers with one or
3517: three odd basis elements in
3518: \eqref{int-numbers} can be non-zero and consequently one has
3519: \beq \label{van-int}
3520: \cK_{\alpha \beta \gamma} = \cK_{\alpha a b} = \cK_{\alpha a} = \cK_{\alpha} = 0\ ,
3521: \eeq
3522: while all other intersection numbers can be non-vanishing.\footnote{From a supergravity
3523: point of view this has been discussed also in \cite{ADAF}.}
3524: This implies that the metric $G_{AB}(t^A)$ of \eqref{Kmetric} is block
3525: diagonal and obeys
3526: \begin{eqnarray} \label{splitmetrIIA}
3527: G_{a b}=
3528: -\frac{3}{2}\left( \frac{\KK_{a b}}{\KK}-
3529: \frac{3}{2}\frac{\KK_a \KK_b}{\KK^2} \right)\ , \qquad
3530: G_{\alpha \beta}=-\frac{3}{2} \frac{\KK_{\alpha \beta}}{\KK}\ , \qquad
3531: G_{\alpha b}\ =\ 0\ ,
3532: \end{eqnarray}
3533: where
3534: \begin{equation}\label{intO6}
3535: \KK_{ab}=\KK_{abc}\; v^c\ ,
3536: \quad \
3537: \KK_{\alpha \beta}=\KK_{\alpha \beta a}\; v^a\ ,\quad
3538: \KK_{a}=\KK_{a b c}\; v^b v^c\ ,
3539: \quad \KK=\KK_{abc} \; v^a v^b v^c
3540: \ .
3541: \end{equation}
3542: In comparison to type IIB orientifolds the opposite intersection
3543: numbers vanish as can be seen by comparing \eqref{van-int} with \eqref{constr}.
3544: This is due to the fact that the K\"ahler form $J$ transforms in IIA and IIB
3545: orientifolds with a relative minus sign under the action of $\sigma$.
3546:
3547: The same consideration also truncates the $N=2$ gauge-kinetic
3548: coupling matrix $\cN_{\Ah \Bh}$ explicitly given in \eqref{def-cN}.
3549: Inserting \eqref{van-int} and \eqref{intO6} one arrives at
3550: \beq \label{def-N_alph_bet}
3551: \text{Re} \cN_{\alpha \beta} = - \cK_{\alpha \beta a} b^a\ , \qquad
3552: \text{Im} \cN_{\alpha \beta} = \cK_{\alpha \beta}\ ,\qquad \cN_{a \alpha}=\cN_{0 \alpha}=0\ .
3553: \eeq
3554: (The other non-vanishing matrix elements $\cN_{\ah\bh}$ arise in the potential
3555: \eqref{U-pot} once fluxes are turned on.)
3556:
3557: Let us now discuss the
3558: terms in the second line of \eqref{act1} arising from the
3559: reduction of the $N=2$ hypermultiplet action which is
3560: determined by the quaternionic metric \eqref{q-metr}.
3561: $D$ is the the four-dimensional dilaton defined in \eqref{4d-dilaton}.
3562: The metric $G_{KL}$ is inherited from the $N=2$
3563: K\"ahler metric $G_{K \bar L}(z,\bar z)$ given in \eqref{csmetric}
3564: and thus is
3565: the induced metric on the submanifold $\cM^{\rm cs}_\bbR$
3566: defined by the constraint \eqref{constrO}.
3567: More precisely, the complex structure deformations respecting \eqref{constrO}
3568: can be determined from \eqref{Kod-form}
3569: by considering infinitesimal variations of
3570: $\Omega$
3571: \beq
3572: \Omega(z + \delta z) \ =\ \Omega(z) + \delta z^K (\partial_{z^K} \Omega)_{z} \
3573: =\ \Omega(z) - \delta z^K( \Kcs_{z^K} \Omega - \chi_K)_z \ .
3574: \eeq
3575: Now we impose the condition that both
3576: $\Omega(z+\delta z)$ and $\Omega(z)$ satisfy \eqref{constrO}.
3577: This implies locally
3578: \beq \label{constr2}
3579: \delta z^K\, \partial_{z^K} \Kcs = \delta\bar z^K\, \partial_{\bar z^K} \Kcs\ , \qquad
3580: \delta z^K\sigma^* \chi_K = e^{2i \theta} \delta \bar z^K\bar \chi_K \ ,
3581: \eeq
3582: where $\partial_{z^K} \Kcs$ and $\chi_K$ are restricted to $\cM^{\rm cs}_\bbR$.
3583: Using the fact that $\Kcs$ is a K\"ahler potential and therefore $\partial_{z^K}\Kcs\neq 0$, we conclude from
3584: the first equation in \eqref{constr2} that for each $\delta z^K$ either the
3585: real or imaginary part has to be zero. This is consistent with the observation
3586: of the previous section that coordinates of $\cM^{\rm cs}_\bbR$ can be
3587: identified with
3588: the real or imaginary part of the complex structure deformations $z^K$.
3589: To simplify the notation we call these deformations collectively
3590: $q^K$ and denote the embedding map by
3591: $\rho:\cM^{\rm cs}_\bbR \hookrightarrow \cM^{\rm cs}$.
3592: Locally this corresponds to
3593: \beq \label{embmap1}
3594: \rho:\ q^K=(q^s,q^\sigma)\ \mapsto\ z^K=(q^s,iq^\sigma)\ ,
3595: \eeq
3596: for some splitting $z^K=(z^s,z^\sigma)$. In other words,
3597: the local coordinates on $\cM^{\rm cs}_\bbR$
3598: are $\R z^s=q^s$ and $\I z^\sigma = q^\sigma$ while $\I z^s=0=\R z^\sigma$.
3599: Using the second equation in \eqref{constr2}, the embedding
3600: map \eqref{embmap1} and the expression \eqref{chi_barchi} for the $N=2$ metric $G_{K\bar L}$ we also deduce that
3601: the K\"ahler form vanishes when pulled back to $\cM^{\rm cs}_\bbR$.
3602: In summary we
3603: have
3604: \beq \label{def-G}
3605: \rho^*(G_{K \bar L}\, dz^K d \bar z^L)\, \equiv\, G_{KL}(q)\, dq^K dq^L\ , \quad
3606: \rho^*(iG_{K \bar L}\, dz^K \wedge d \bar z^L)\, =\, 0\ .
3607: \eeq
3608: The first equation defines the induced metric while the second equation
3609: implies that $\cM^{\rm cs}_\bbR$
3610: is a Lagrangian submanifold of $\cM^{\rm cs}$ with respect to the
3611: K\"ahler-form.
3612:
3613:
3614:
3615: Finally, coming back to the action \eqref{act1}
3616: the matrix $\cM_{\Kh \Lh}$ is defined in analogy with
3617: \eqref{defM} as
3618: \beq \label{defM2}
3619: \int a_\Kh \wedge * a_\Lh = -\text{Im}\; \cM_{\Kh \Lh} \ , \qquad
3620: %+(\text{Re}\; \cM)
3621: % (\text{Im}\; \cM)^{-1}(\text{Re}\; \cM))_{\Kh \Lh}
3622: %\ , \qquad\qquad
3623: % \int a_\Kh\wedge * b^\Lh = 0\ , \nn\\
3624: \int b^\Kh \wedge * b^\Lh\ = \ -(\text{Im}\; \cM)^{-1\ \Kh \Lh}\ ,
3625: % -((\text{Re}\; \cM)(\text{Im}\; \cM)^{-1})_{\Kh}^\Lh\ . \nn
3626: \eeq
3627: where $\text{Im} \cM_{\Kh \Lh}$ can be given explicitly
3628: in terms of the periods by inserting \eqref{Z=0} into \eqref{gauge-c} \cite{TGL1}.
3629: %This yields
3630: %\bea \label{gauge-r}
3631: %\I \cM_{\Kh \Lh}=-\I \mathcal{F}_{\Kh \Lh}+2 \frac{(\text{Im}\; \mathcal{F})_{\Kh \Mh} \R (CZ^\Mh)
3632: % (\text{Im}\; \mathcal{F})_{\Lh \Nh}\R (CZ^\Nh) }{\R (CZ^\Nh)(\text{Im}\; \mathcal{F})_{\Nh\Mh}
3633: % \R (CZ^\Mh)}\ .
3634: %\eea
3635: Similarly one obtains $\R\cM_{\Kh \Lh}=0$
3636: consistent with the fact that \eqref{defM} implies that $\int a_\Kh\wedge * b^\Lh$ vanishes
3637: for the special basis $(a_\Kh,b^\Kh)$.
3638:
3639: This ends our discussion of
3640: the effective action obtained by applying the orientifold projection.
3641: The next step is to rewrite the action \eqref{act1}
3642: in the standard $N=1$ supergravity form
3643: which we turn to now.
3644:
3645:
3646: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3647: %
3648: % Kahler potential
3649: %
3650: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3651:
3652: \subsection{The K\"ahler potential and gauge-couplings}
3653: \label{Kpo_gaugeIIA}
3654:
3655:
3656: The standard $N=1$ supergravity the action is expressed
3657: in terms of a K\"ahler potential $K$, a holomorphic superpotential $W$
3658: and the holomorphic gauge-kinetic coupling
3659: functions $f$ as given in \eqref{N=1action}. Hence, our task is to
3660: find $K,f$ and $W$ for the type IIA orientifolds. As an immediate
3661: observation one finds that \eqref{act1} includes no potential, such
3662: that $W=0$ and $D_\alpha=0$. It is also not difficult to read
3663: off the gauge-kinetic coupling function $f_{\alpha \beta}$.
3664: Comparing \eqref{act1} with \eqref{N=1action}
3665: using \eqref{def-N_alph_bet} and \eqref{def-t}
3666: one infers
3667: \beq \label{gauge-A}
3668: f_{\alpha \beta}\ =\ -i \bar \cN_{\alpha \beta}\ =\ i \cK_{\alpha \beta a} t^a \ .
3669: \eeq
3670: As required by $N=1$ supersymmetry the $f_{\alpha \beta}$
3671: are indeed holomorphic. Note that they are linear in the $t^a$ moduli
3672: and do not depend on the complex structure and $\xi$-moduli.
3673:
3674: {}From \eqref{act1} we also immediately observe that
3675: the orientifold moduli space has the product structure
3676: \beq \label{direct-mod}
3677: \cM_{N=1}=\tilde\cM^{\rm SK} \times \tilde\cM^{\rm Q}\ .
3678: \eeq
3679: The first factor $\tilde\cM^{\rm SK}$ is a subspace of the
3680: $N=2$ moduli space $\cM^{\rm SK}$ with dimension $h^{(1,1)}_-$
3681: spanned by the complexified K\"ahler deformations $t^a$.
3682: The second factor $\tilde\cM^{\rm Q}$ is a subspace of the quaternionic
3683: manifold $\cM^{\rm Q}$ with
3684: dimension $h^{(2,1)} +1$
3685: spanned by the complex structure deformations $q^K$, the dilaton $D$
3686: and the scalars $\xi^{\hat K}$ arising from $C_3$.
3687: Let us discuss both factors in turn.
3688:
3689: As we already stressed earlier the metric $G_{ab}$
3690: of \eqref{act1} defined in \eqref{splitmetr} is a trivial
3691: truncation
3692: of the $N=2$ special K\"ahler metric \eqref{Kmetric} and therefore remains
3693: special K\"ahler. The K\"ahler potential is given by
3694: \beq \label{Kks}
3695: K^{\rm K}\ =\ - \ln \Big[\tfrac{i}{6}\cK_{a b c} (t -\bar t)^a (t -\bar t)^b (t -\bar t)^c \Big]
3696: \ = \ - \ln \Big[\tfrac{4}{3} \int_Y J \wedge J \wedge J\Big]\ ,
3697: \eeq
3698: where $J$ is the K\"ahler form in the string frame.
3699: Moreover, $K^{\rm K}$ can be obtained from
3700: the prepotential $f(t)=-\tfrac{1}{6} \cK_{abc}t^a t^b t^c$
3701: by using equation \eqref{Kinz}.
3702: It is well known that $K^{\rm K}$ obeys the standard no-scale condition
3703: \cite{NS}
3704: \beq \label{no-scale1}
3705: K_{t^a} K^{t^a \bar t^b} K_{ \bar t^b}\ =\ 3\ .
3706: \eeq
3707:
3708: The geometry of the second component $\tilde\cM^{\rm Q}$ in \eqref{direct-mod}
3709: is considerably more complicated. This is due to the fact that
3710: \eqref{newO} defines a new complex structure on the field space. In the following
3711: we sketch the calculation of the K\"ahler potential for the basis
3712: $(a_\Kh,b^\Kh)$ and only summarize the results for a generic symplectic basis.
3713: The details of this more involved calculation will be presented in section \ref{IIA_lin}.
3714:
3715: To begin with, let us define the compensator $C$ introduced in section \ref{spectrum-IIA} as
3716: \beq \label{def-C}
3717: C\ =\ e^{-D-i\theta} e^{\Kcs(q)/2}\ , \qquad C \rightarrow C e^{\R\, h(q)}\ ,
3718: \eeq
3719: where $\Kcs$ is the K\"ahler potential defined in \eqref{csmetric} restricted
3720: to the real subspace $\cM^{\rm cs}_{\bbR}$. We also displayed the transformation
3721: behavior of $C$ under real K\"ahler transformations \eqref{real_K}. With this at hand one
3722: defines the scale invariant variable
3723: \beq \label{l-def}
3724: l^\Kh \ =\ \R(C Z^\Kh(q))\ .
3725: \eeq
3726: Inserted into \eqref{act1} and using
3727: the Jacobian matrix encoding the change of variables $(e^D,q^K) \rightarrow l^\Kh$
3728: the second line \eqref{act1} simplifies as\footnote{%
3729: The calculation of this result can be found in section \ref{IIA_lin}.}
3730: \beq \label{IIAQ}
3731: \cL^{(4)}_{\rm Q} = 2 e^{2D}\, \text{Im}\, \cM_{\Kh \Lh}\,
3732: (dl^\Kh \wedge * dl^\Lh + \tfrac{1}{4} d\xi^{\Kh} \wedge * d\xi^{\Lh})\ .
3733: \eeq
3734: We see that the scalars $l^\Kh$ and $\xi^\Kh$ nicely combine
3735: into complex coordinates
3736: \beq \label{Ncoords}
3737: N^\Kh\ =\ \tfrac{1}{2}\xi^\Kh + i l^\Kh\
3738: =\ \tfrac{1}{2}\xi^\Kh + i \R(C Z^\Kh)
3739: = \tfrac{1}{2} \int \Omega_c\wedge b^\Kh
3740: \ ,
3741: \eeq
3742: which we anticipated in equation \eqref{newO}.
3743: The important fact to note here is that $\tilde\cM^{\rm Q}$
3744: is equipped with a new complex structure and the corresponding
3745: K\"ahler coordinates
3746: coincide with half of the periods of $\Omegac$.
3747: This is in contrast to the situation in $N=2$ where one of the periods
3748: ($Z^0$) is a gauge degree of freedom and the K\"ahler
3749: coordinates are the special coordinates $z^K = Z^K/Z^0$.
3750:
3751: In order to show that the metric in \eqref{IIAQ} is K\"ahler we need the
3752: explicit expression for the K\"ahler potential. Using \eqref{Z=0} in \eqref{gauge-c}
3753: one obtains straightforwardly
3754: \beq
3755: 2 e^{2D} \text{Im}\, \cM_{\Kh \Lh} = \partial_{N^\Kh} \partial_{\bar N^\Lh} K^{\rm Q}\ ,
3756: \eeq
3757: where
3758: \beq \label{KQsimple}
3759: K^{\rm Q} = -2 \ln\big[4i\cF(CZ)\big]\ , \qquad
3760: \cF\big(\R(CZ)\big) = \frac{i}2\, \R(C Z^\Kh)\, \I(C\cF_\Kh)\ .
3761: \eeq
3762: Alternatively, using \eqref{decompO} and $*\Omega =- i \Omega$
3763: one derives the integral representation
3764: \beq \label{intKQ}
3765: K^{\rm Q}\ = - 2\ln\Big[2\int_Y \R(C \Omega)\wedge *\R(C\Omega)\Big]=\ - \ln\, e^{-4D} \ ,
3766: \eeq
3767: where in the second equation we used \eqref{def-C} and \eqref{csmetric}. In
3768: the form \eqref{intKQ} the dependence of $K^{\rm Q}$ on the coordinates $N^\Kh$
3769: is only implicit and given by means of their definition \eqref{Ncoords}.
3770: Also $K^{\rm Q}$ obeys a no-scale type condition in that it
3771: satisfies
3772: \bea \label{no-scale2}
3773: K_{N^\Kh} K^{N^\Kh \bar N^\Lh} K_{\bar N^\Lh} = 4\ ,
3774: \eea
3775: which can be checked by direct calculation.
3776:
3777:
3778: The analysis so far started from the symplectic basis
3779: $(a_\Kh,b^\Kh)$ introduced in \eqref{basis_ab},
3780: determined the K\"ahler coordinates in \eqref{Ncoords}
3781: and derived the K\"ahler potential $K^{\rm Q}$
3782: in terms of the prepotential $\cF$ in \eqref{KQsimple} or as an
3783: integral representation in \eqref{intKQ}. Now we need to ask
3784: to what extent this result depends on the choice of
3785: the basis \eqref{basis_ab}. Or in other words let us redo
3786: the calculation starting from an arbitrary symplectic basis
3787: and determine the K\"ahler potential and the proper field variables
3788: for the corresponding orientifold theory.
3789: Let us first recall the situation
3790: in the $N=2$ theory reviewed in section \ref{revIIA}.
3791: The periods $(Z^\Kh,\cF_\Kh)$ defined in \eqref{pre-z}
3792: form a symplectic vector
3793: of $Sp(2h^{(1,2)}+2,\bf Z)$
3794: such that $\Omega$ given in \eqref{Omegaexp} and
3795: $\Kcs$ given in \eqref{csmetric} is manifestly invariant.
3796: The prepotential $\cF(Z) = \frac{1}{2} Z^\Kh \cF_\Kh$ on the other hand
3797: does depend
3798: on the choice of the basis $(\alpha_\Kh,\beta^\Kh)$
3799: and is not invariant.
3800:
3801: For $N=1$ orientifolds this situation is different
3802: since the orientifold projection \eqref{constrO} explicitly breaks the
3803: symplectic invariance.\footnote{A symplectic transformation $\cS$ preserve the
3804: form $\big<\alpha,\beta\big> = \int \alpha \wedge \beta$, such that
3805: $\big<\cS \alpha,\cS \beta \big> = \big< \alpha,\beta \big>$.
3806: On the other hand the anti-holomorphic involution satisfies
3807: $\big<\sigma^* \alpha,\sigma^* \beta \big>
3808: = - \big< \alpha,\beta \big>$.}
3809: This can also be seen from the form
3810: of the $N=1$ K\"ahler potential \eqref{KQsimple} which is expressed
3811: in terms of the non-invariant prepotential.
3812: One immediately concludes that the result \eqref{KQsimple} is
3813: basis dependent and $K^Q$ takes this simple form due to the special
3814: choice $a_\Kh \in H^{3}_+(Y)$ and $b^\Kh \in H^3_-(Y)$.\footnote{Note that this is in striking analogy to
3815: the background dependence of the B model partition function as discussed in \cite{BCOV,Witten2}.}
3816: On the other hand, the integral representation \eqref{intKQ} only implicitly depends
3817: on the symplectic basis through the definition of the coordinates $N^\Kh$.
3818: This suggest, that it is possible to generalize our results by allowing for
3819: an arbitrary choice of symplectic basis in the definition of the $N=1$ coordinates.
3820: More precisely, let us consider the generic basis $(\alpha_\Kh,\beta^\Lh)$,
3821: where we assume that the $h^3_+=h^{2,1}+1$ basis elements $(\alpha_k,\beta^\lambda)$
3822: span $H^3_+$ and the $h^3_-=h^{2,1}+1$ basis elements $(\alpha_\lambda,\beta^k)$ span $H^3_-$.
3823: In this basis the intersections \eqref{int-numbers1} take the form
3824: \beq \label{sp_alpha-beta}
3825: \int_Y \alpha_k \wedge \beta^l\ =\ \delta_k^l\ , \qquad
3826: \int_Y \alpha_\kappa \wedge \beta^\lambda\ =\ \delta_\kappa^\lambda\ ,
3827: \eeq
3828: with all other combinations vanishing.
3829: Applying the orientifold constraint \eqref{constrO} one concludes that
3830: the equations \eqref{Z=0} are replaced by
3831: \beq \label{Z=0gen}
3832: \I(C Z^k) = \R (C \cF_k )\ =\ 0\ , \qquad
3833: \R (C Z^\lambda) = \I(C \cF_\lambda)\ =\ 0\ .
3834: \eeq
3835: Correspondingly, the expansions \eqref{decompO} and \eqref{form-exp}
3836: take the form
3837: \bea \label{decompO2}
3838: C \Omega &=& \R (C Z^k) \alpha_k + i\I (C Z^\lambda ) \alpha_\lambda -
3839: \R (C \cF_\lambda) \beta^\lambda - i\I (C \cF_k) \beta^k\ ,\nn\\
3840: \CC_3 &=& \xi^k\, \alpha_k - \tilde \xi_\lambda\, \beta^\lambda\ ,
3841: \eea
3842: which implies that we also have to redefine the $N=1$ coordinates of
3843: $\tilde \cM^{\rm Q}$ in an appropriate way.
3844: In section~\ref{IIA_lin} we show that the
3845: new K\"ahler coordinates $(N^k,T_\lambda)$ are again determined by the periods of $\Omegac$ and given by
3846: \bea \label{Oexp}\label{def-NT}
3847: N^k &=&\tfrac{1}{2} \int \Omegac \wedge \beta^k \
3848: =\ \tfrac{1}{2}\xi^k + i \R(CZ^k)\ , \nn\\
3849: T_\lambda &=& i \int \Omegac \wedge \alpha_\lambda\ =\
3850: i\tilde \xi_\lambda - 2 \R (C \cF_\lambda) \ ,
3851: \eea
3852: where we evaluated the integrals by using \eqref{Omegacdef}
3853: and \eqref{decompO2}.
3854:
3855: The K\"ahler potential takes again the form \eqref{intKQ} but now
3856: depends on $N^k,T_\lambda$ and thus no longer simplifies to \eqref{KQsimple}.
3857: %A detailed discussion of this calculation is given in section \ref{IIA_lin}.
3858: Let us compare the situation to the original $N=2$ theory, which
3859: was formulated in terms of the
3860: $Z^\Kh$ or equivalently the special coordinates $z^K$. Holomorphicity
3861: in these coordinates played a central role in defining the prepotential
3862: encoding the special geometry of $\cM^{\rm cs}$ in $\cM^{\rm Q}$ (cf.~section
3863: \ref{revIIA}). In contrast, the $N=1$ orientifold constraints destroy this complex structure and force us
3864: to combine $\R(C\Omega)$ with the RR three-form $C_3$ into $\Omegac$. The
3865: K\"ahler coordinates are half of the periods of $\Omegac$
3866: but now in this more general case also the
3867: derivatives of $\cF$ can serve as coordinates as seen in \eqref{def-NT}.
3868: However, as it is shown in section
3869: \ref{IIA_lin}, $\R (C \cF_\lambda)$ and $e^{2D}\I (CZ^\lambda)$ are related
3870: by a Legendre transformation of the K\"ahler potential. Working with this transformed
3871: potential and the coordinates $\R(CZ^k)$ and $e^{2D}\I (CZ^\lambda)$ enables us
3872: to make contact to the underlying $N=2$ theory in its canonical formulation.
3873: From a supergravity point of
3874: view, this Legendre transformation corresponds to replacing the chiral multiplets
3875: $T_\lambda$ by linear multiplets as described in the next chapter.
3876: This is possible due to the translational isometries of $K$,
3877: which arise as a consequence of the $C_3$ gauge invariance
3878: and which render $K$ independent of
3879: the scalars $\xi$ and $\tilde\xi$.
3880: We show in section \ref{geom_of_modspace}
3881: that this also enables us to construct $\tilde \cM^{\rm Q}$ from $\cM^{\rm cs}_\bbR$
3882: similar to the moduli space of supersymmetric Lagrangian submanifolds in a
3883: Calabi-Yau space as described by Hitchin \cite{Hitchin2}.
3884: This also allows us to interpret the no-scale condition \eqref{no-scale2}
3885: geometrically.
3886:
3887: Let us summarize the results obtained so far. We found that the moduli
3888: space of $N=1$ orientifolds is indeed the product of two K\"ahler spaces
3889: with the K\"ahler potential
3890: \beq \label{N=1Kpot}
3891: K\ =\ K^{\rm K} + K^{\rm Q} = - \ln \Big[\tfrac{4}{3} \int_Y J \wedge J \wedge J\Big]
3892: - 2\ln\Big[2\int_Y \R(C \Omega)\wedge *\R(C\Omega)\Big]\ .
3893: \eeq
3894: The first term depends on the K\"ahler deformations of the orientifold
3895: while the second term is a function of the real complex structure
3896: deformations and the dilaton.
3897: The $N=1$ K\"ahler coordinates are obtained
3898: by expanding the complex combinations\footnote{This combination
3899: of forms has also appeared recently in ref.\ \cite{NOV}
3900: in the discussion of $D$-instanton couplings in the A-model.
3901: Here they appear as the proper chiral $N=1$ variables and as we will
3902: see in the next section they linearize the D-instanton action.}
3903: \beq \label{N=1coords}
3904: \Omegac\ =\ \CC_3 + 2i \R(C\Omega)\ ,\qquad
3905: \Jc\ =\ \hat B_2 + iJ \ ,
3906: \eeq
3907: in a real harmonic basis of $H^{3}_+(Y)$ and $H^{(1,1)}_-(Y)$ respectively.
3908: Note that $K$ does not depend on the scalars arising in the expansion of
3909: $\hat B_2$ and $\hat C_3$, such that the K\"ahler manifold admits a set of
3910: $h^{(1,1)}_- + h^{(2,1)}+1$ translational isometries. In other words
3911: $K$ consists of two functionals encoding the dynamics of the two-form $J$
3912: and the real three-form $\R(C\Omega)$.
3913: %\footnote{The functions
3914: %$V [ \R(C\Omega) ]= \int \R(C\Omega)\wedge* \R(C\Omega)$
3915: %and $V [J ]=\int J\wedge J \wedge J$ are known as Hitchins functionals \cite{Hitchin1}.
3916: %The orientifold constraints \eqref{constrJ} and \eqref{constrO}
3917: %restricts their domain to $J \in H^2_-(Y)$
3918: %and $\R(C\Omega) \in H^3_+(Y)$.}
3919: In type IIA orientifolds it is not difficult to rewrite $K^{\rm Q}$ in a form
3920: similar to \eqref{def-Phi}. Defining the odd form
3921: \beq \label{def-podd}
3922: \podd=\fu +i\, \fuh = C\Omega\ ,
3923: \eeq
3924: one finds
3925: \beq \label{symp-formodd}
3926: K^{\rm Q} = - 2 \ln \Phi_A(\fu) \ ,\qquad \Phi_A(\fu) \equiv i\big<\podd,\poddb \big>=i\int_Y \podd \wedge \poddb\ .
3927: \eeq
3928: The function $\Phi_A(\fu)$ is
3929: known as Hitchins functional for the real
3930: three-form $\fu$ \cite{Hitchin1,HitchinGCM}. The orientifold constraint \eqref{constrO}
3931: restricts its domain to $\fu \in H^3_+(Y)$. Applying the fact that
3932: $\Phi_A(\fu)$ is a homogeneous function of degree two $K^{\rm Q}$ obeys the no-scale type conditions
3933: \eqref{no-scale2}, \eqref{no-scale4}. This is independent of the chosen basis
3934: and can be also shown directly as done in section \ref{IIA_lin}.
3935: %Moreover, irrespective of the chosen basis the K\"ahler potential
3936: %obeys the no-scale type conditions \eqref{no-scale1} and
3937:
3938: The no-scale conditions are violated when further stringy
3939: corrections are included. $K$ receives additional contributions due
3940: to perturbative effects as well as world-sheet and $D2$ instantons.
3941: It is well-known that the combination $\Jc=\hat B_2 + i J$
3942: gives the proper coupling to the string world-sheet such that
3943: world-sheet instantons correct the holomorphic prepotential as
3944: $f(t) = -\frac{1}{6}\cK_{abc}t^a t^b t^c + O(e^{-t})$.
3945: Since we divided out the world-sheet parity these corrections also
3946: include non-orientable Riemann surfaces, such that the prepotential
3947: $f(t)$ consists of two parts $f(t) = f_{or}(t) + f_{unor}(t)$.
3948: The function $f_{or}$ counts holomorphic maps
3949: from orientable world-sheets to $Y$, while $f_{unor}$ counts holomorphic maps
3950: from non-orientable world-sheets to $Y$ \cite{BFM}.
3951: In the next section we show that $D2$ instantons naturally couple to the complex three-form
3952: $\Omegac$ and they are expected to correct
3953: $K^{\rm Q}$.
3954:
3955:
3956:
3957: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3958: %
3959: % Mirror symmetry
3960: %
3961: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3962:
3963: \section{Mirror symmetry \label{Mirror_orientioflds}}
3964:
3965: In this section we discuss mirror symmetry
3966: for Calabi-Yau orientifolds from the point of view of the effective
3967: action derived in the large volume limit. More precisely, we compare the $N=1$
3968: data for type IIB orientifolds on $\tilde Y/\sigma_B$ (section \ref{Kpo_gaugeIIB})
3969: with the data for type IIA orientifolds on $Y/\sigma_A$ (section \ref{Kpo_gaugeIIA}).
3970: Since we want to discuss mirror symmetry we choose $\tilde Y$
3971: to be the mirror manifold of $Y$. This implies that the
3972: non-trivial Hodge numbers $h^{(1,1)}$ and $h^{(2,1)}$ of $Y$ and $\tilde Y$
3973: satisfy $h^{(1,1)}(Y)=h^{(2,1)}(\tilde Y)$ and $h^{(2,1)}(Y)=h^{(1,1)}(\tilde Y)$ as already
3974: given in section \ref{revMirror} where we briefly introduced $N=2$ mirror symmetry.
3975: In orientifolds we also have to specify the
3976: involutions $\sigma_A$ and $\sigma_B$ which are identified under mirror symmetry. Since the
3977: discussion
3978: in this article is quite generic and never specified any involution
3979: $\sigma$ explicitly we also keep the discussion of mirror symmetry
3980: generic. That is we assume that there exists a mirror pair
3981: of manifolds $Y$ and $\tilde Y$ with a mirror pair of involutions
3982: $\sigma_A, \sigma_B$.
3983: Matching the number of $N=1$ multiplets summarized in table \ref{numberM}
3984: implies an orientifold version of
3985: \eqref{Hod_id},\footnote{For the sector of $\tilde \cM^{\rm Q}$ mirror
3986: symmetry
3987: is a constraint on the couplings rather than the Hodge
3988: numbers.} i.e.\
3989: \bea \label{matchchohm}
3990: O3/O7&: &\quad h^{1,1}_-(Y) = h^{2,1}_-(\tilde Y) \ , \qquad h^{1,1}_+(Y) = h^{2,1}_+(\tilde Y) \ ,\nn\\
3991: O5/O9&: &\quad h^{1,1}_-(Y) = h^{2,1}_+(\tilde Y) \ , \qquad h^{1,1}_+(Y) = h^{2,1}_-(\tilde Y) \ .
3992: \eea
3993: \begin{table}[h]
3994: \begin{center}
3995: \begin{tabular}{|l|c|c|c|} \hline
3996: \rule[-0.3cm]{0cm}{0.8cm}
3997: multiplets& IIA$_Y$ \ $O6$ & IIB$_{\tilde Y}$ \ $O3/O7$ & IIB$_{\tilde Y}$ \ $O5/O9$ \\ \hline\hline
3998: \rule[-0.3cm]{0cm}{0.9cm}
3999: {vector multiplets} & $h_+^{(1,1)}$ & $h_+^{(2,1)}$ & $h_-^{(2,1)}$ \\ \hline
4000: \rule[-0.3cm]{0cm}{0.9cm}
4001: chiral multiplets in $\tilde \cM^{\rm SK}$& $h_-^{(1,1)}$ & $h_-^{(2,1)}$ &
4002: $h_+^{(2,1)} $ \\ \hline
4003: \rule[-0.3cm]{0cm}{0.9cm}
4004: chiral multiplets in $\tilde \cM^{\rm Q}$&$h^{(2,1)} + 1$&$h^{(1,1)} + 1$ &
4005: $h^{(1,1)} + 1$ \\ \hline
4006: \end{tabular}
4007: \caption{ \textit{Number of $N=1$ multiplets of orientifold compactifications.}}\label{numberM}
4008: \end{center}
4009: \end{table}
4010:
4011:
4012:
4013:
4014:
4015:
4016: Our next task will be to match the couplings of the mirror theories.
4017: Since the effective actions on both sides
4018: are only computed in the large volume limit
4019: we can expect to find agreement only if we also take
4020: the large complex structure limit exactly as in the $N=2$ mirror
4021: symmetry.
4022: However, if one believes in mirror symmetry one can use the
4023: the geometrical results of the complex structure moduli space to
4024: `predict' the corrections to its mirror symmetric component.
4025: This is not quite as straightforward since the full $N=1$ moduli space is a
4026: lot more complicated than the underlying $N=2$ space \cite{BH}.
4027: Let us therefore start our analysis with the simpler situation of the
4028: special K\"ahler sectors $\tilde \cM^{\rm SK}_A,\, \tilde \cM_B^{\rm SK}$ in \eqref{direct-mod}
4029: and \eqref{modulispaceO3} and the vector multiplet couplings
4030: and postpone the analysis of $\tilde M^{\rm Q}_{A,B}$
4031: to section \ref{O3O7mirror}.
4032:
4033: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4034: \subsection{Mirror symmetry in $ \mathcal{M}^{\rm K}$} \label{mirrorMK}
4035: %%%
4036: Recall that the manifold $\tilde \cM^{\rm SK}_A$ is spanned by the
4037: complexified K\"ahler deformations $t^a$ preserving the constraint
4038: \eqref{constrJ}. Under mirror symmetry these moduli are mapped
4039: to the complex structure deformations which respect the constraint
4040: \eqref{Omegatransf}.
4041: In both cases the K\"ahler potential is merely a truncated
4042: version of the $N=2$ K\"ahler potential and one has
4043: \beq
4044: K^{\rm K}_A \ =\ - \ln \Big[\tfrac{4}{3} \int_Y J \wedge J \wedge J\Big]
4045: \quad \leftrightarrow \quad
4046: K^{\rm cs}_B\ =\ - \ln \Big[-i \int \Omega \wedge \bar \Omega \Big]\ .
4047: \eeq
4048: Both K\"ahler potentials can be expressed in terms of prepotentials
4049: $f_A(t), f_B(z)$ and in the large complex structure limit
4050: $f_B(z)$ becomes cubic and agrees with $f_A(t)$.
4051: Mirror symmetry therefore equates these prepotentials
4052: and exchanges $J^3$ with $\Omega\wedge\bar\Omega$
4053: exactly as in $N=2$
4054: \beq\label{mirrorK}
4055: f_A(t) = f_B(z) \ , \qquad J^3 \leftrightarrow
4056: \Omega\wedge\bar\Omega\ .
4057: \eeq
4058: In \cite{FMM} the $N=2$ version of this map was written into the form \footnote{The authors argued that
4059: this should be true also for mirror symmetry of certain non-Calabi-Yau backgrounds. }
4060: \beq \label{pure-spinor-map}
4061: e^{J_c}\ \leftrightarrow\ \Omega\ ,
4062: \eeq
4063: where $J_c$ is given in \eqref{N=1coords}.
4064: Thus for $\tilde \cM^{\rm SK}$ mirror symmetry is a truncated
4065: version of $N=2$ mirror symmetry. As we will see momentarily this also
4066: holds for the gauge kinetic couplings
4067: which depend holomorphically on the moduli spanning $\tilde \cM^{\rm SK}$.
4068:
4069: In type IIA the gauge-kinetic couplings
4070: are given in \eqref{gauge-A} and read
4071: $f_{\alpha \beta}(t) = i\cK_{\alpha\beta c} t^c$.
4072: The IIB couplings were determined in \eqref{fholo} to be
4073: \bea \label{gauge-B}
4074: f_{\alpha\beta}(z^a) = - {i} \bar \cM_{\alpha\beta}
4075: = - i\cF_{\alpha\beta}\ ,
4076: \eea
4077: where in order to not overload the notation we are using the same indices
4078: for both cases.\footnote{We rescaled the type IIB gauge bosons
4079: by $\sqrt 2$ in order to properly match the normalizations.}
4080: More precisely we are choosing
4081: \bea
4082: \alpha, \beta = 1, \ldots, h^{(2,1)}_+(\tilde Y)\ ,\qquad
4083: a, b = 1, \ldots, h^{(2,1)}_-(\tilde Y)\ , \qquad \textrm{for} \quad O3/O7\ ,\nn\\
4084: \alpha, \beta = 1, \ldots, h^{(2,1)}_-(\tilde Y)\ ,\qquad
4085: a, b = 1, \ldots, h^{(2,1)}_+(\tilde Y)\ ,\qquad \textrm{for} \quad
4086: O5/O9\ .
4087: \eea
4088: The matrix $\cF_{\alpha\beta}(z^a)$ is
4089: holomorphic and the second derivatives of the prepotential restricted
4090: to $\tilde \cM^{\rm K}_B$. In the large complex structure limit
4091: $\cF_{\alpha\beta}$ is linear
4092: in $z^a$ and therefore also agrees with the type IIA mirror
4093: couplings.
4094: Thus mirror symmetry implies the map $\cN_{\alpha \beta}(\bar t^a) = \cM_{\alpha\beta}(\bar z^a)$
4095: in both cases.
4096:
4097: This concludes our discussions of mirror symmetry
4098: for the chiral multiplets which span $\tilde\cM^{\rm SK}$.
4099: We have shown that
4100: the K\"ahler potential and
4101: the gauge-kinetic coupling functions
4102: agree in the large complex structure limit under mirror symmetry.
4103: In this sector the geometrical quantities on the type IIB side include
4104: corrections which are believed to
4105: compute world-sheet non-perturbative effects
4106: such as world-sheet instantons on the type
4107: IIA side.
4108: This is analogous to the situation
4109: in $N=2$ and may be traced back to the
4110: fact, that it is still possible to formulate a topological
4111: A model counting
4112: world-sheet instantons for Calabi-Yau orientifolds \cite{AAHV,BFM}.
4113:
4114: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4115: \subsection{Mirror symmetry in $ \cM^{\rm Q}$}
4116: %%%
4117: Let us now turn to the discussion of the K\"ahler manifolds $\tilde \cM^{\rm Q}_{A}$ and
4118: $\tilde \cM^{\rm Q}_{B}$ arising in
4119: the reduction of the quaternionic spaces.
4120: %As we already remarked above, the metrics on the
4121: %type IIB K\"ahler manifolds $\tilde \cM^{\rm Q}_B$ (for both projections in \eqref{constrOB})
4122: %are given in terms of `special coordinates'.
4123: On the IIA side the K\"ahler potential is given in \eqref{N=1Kpot}
4124: which is expressed in terms of the $h^{(2,1)}+1$ coordinates
4125: $(N^k,T_\lambda)$ defined in \eqref{Oexp}.
4126: In this definition we did not fix the scale invariance \eqref{real_K}
4127: $\Omega\to
4128: \Omega e^{-\R (h)}$ or in other
4129: words we defined the coordinates in terms of the scale invariant
4130: combination $C\Omega$. Somewhat surprisingly there seem to be two
4131: physically inequivalent ways to fix this scale invariance.
4132: In $N=2$ one uses the scale invariance to define special
4133: coordinates $z^K = Z^K/Z^0, z^0 = 1$ where $Z^0$ is the coefficient
4134: in front of the base element $\alpha_0$. The choice of $Z^0$
4135: is convention and
4136: due to the symplectic invariance any other choice would be
4137: equally good.
4138: However, as we already discussed in section 3.1 and 3.3 the
4139: constraint \eqref{constrO} breaks the symplectic invariance and
4140: $H^3$ decomposes into two eigenspaces $H^3_+\oplus H^3_-$.
4141: Thus in \eqref{decompO2} we have the choice to scale one of the $Z^k$
4142: equal to one or
4143: one of the $Z^\lambda$ equal to $i$.
4144: Denoting the corresponding basis element by $\alpha_0$,
4145: these two choices are characterized by
4146: $\alpha_0 \in H^{3}_+$ or $\alpha_0 \in H^{3}_-$.
4147: This choice identifies the dilaton direction inside the moduli space
4148: and therefore is crucial in identifying the type IIB
4149: mirror. This is related to the fact that in type IIB
4150: the dilaton reside in a chiral multiplet for $O3/O7$ orientifolds and in a
4151: linear multiplet for $O5/O9$ orientifolds as we make more explicit in section
4152: \ref{IIB_lin}. Let us discuss these two cases in turn.
4153:
4154:
4155: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4156: %
4157: % The Mirror of O3/O7
4158: %
4159: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4160:
4161: \subsubsection{The Mirror of IIB orientifolds with $O3/O7$ planes}
4162: \label{O3O7mirror}
4163:
4164: We first want to show that in the large complex structure limit
4165: $K^Q_A$ given in \eqref{intKQ} coincides with
4166: $K^{\rm Q}_B$ given in \eqref{kaehlerpot-Kk} for
4167: orientifolds with $O3/O7$ planes.
4168: It turns out that in order to do so we need to choose
4169: $\alpha_0 \in H^3_{+}$ and the dual basis element
4170: $\beta^0\in H^3_{-}$.
4171: It is convenient to keep track of this choice and therefore
4172: we mark the $\alpha$'s and $\beta$'s which contain $\alpha_0$
4173: and $\beta^0$ by putting a hat on the corresponding index.
4174: Thus we work in the basis $(\alpha_\kh,\beta^\lambda)$
4175: of $H^3_+$ and $(\alpha_\lambda,\beta^\kh)$
4176: of $H^{3}_-$. Therefore, we rewrite the combination $C\Omega$ as
4177: \beq
4178: C\Omega = g_A^{-1}(\textbf{1}\, \alpha_0 + q^k \alpha_k + iq^\lambda \alpha_\lambda) + \ldots\ ,
4179: \eeq
4180: where we introduced $g_A$ and the real special coordinates
4181: \beq \label{realspC1}
4182: g_A =\frac{1}{\R(CZ^0)}\ ,\qquad q^k = \frac{\R(CZ^k)}{\R(CZ^0)}\ , \qquad q^\lambda = \frac{\I(CZ^\lambda)}{\R(CZ^0)}\ .
4183: \eeq
4184: We also need to express the prepotential $\cF(Z)$
4185: in the special coordinates $q^k,q^\lambda$.
4186: In analogy to \eqref{def-f} one defines a function $f(q)$
4187: such that
4188: \beq \label{def-h(q)}
4189: \cF\big(\R[CZ^\kh],i\I[CZ^\lambda] \big)\ =\ i\big(\R[ CZ^0]\big)^2\ f(q^k,q^\lambda) \ .
4190: \eeq
4191: We are now in the position to rewrite the $N=1$ coordinates
4192: $N^\kh,T_\lambda$ given in
4193: \eqref{def-NT} in terms of $g_A$ and the special coordinates $q^K$.
4194: Inserting \eqref{realspC1}
4195: into \eqref{def-NT} one obtains
4196: \beq \label{c-in-q37}
4197: N^0\ =\ \tfrac{1}{2} \xi^0 + i g_A^{-1}\ , \qquad
4198: N^k\ =\ \tfrac{1}{2} \xi^k + i g_A^{-1} q^k \ , \qquad
4199: T_\lambda\ =\ i \tilde \xi_\lambda - 2 g_A^{-1} f_\lambda(q)\ ,
4200: \eeq
4201: where $f_\lambda$ is the first derivative of $f(q)$ with respect to $q^\lambda$.
4202:
4203: The final step is to specify $f(q)$ in the large complex structure
4204: limit.
4205: In this limit the $N=2$ prepotential is known to be
4206: \beq \label{N=2pre}
4207: \cF(Z) = \tfrac{1}{6} (Z^0)^{-1}{\kappa_{KLM} Z^K Z^L Z^M}\ .
4208: \eeq
4209: Inserted into the orientifold constraints
4210: \eqref{Z=0gen} one infers
4211: \beq \label{vankappa37}
4212: \kappa_{klm} = \kappa_{\kappa \lambda l} = 0 \ ,
4213: \eeq
4214: while $\kappa_{\kappa \lambda \mu}$ and $\kappa_{\kappa l m}$ can be non-zero.
4215: Using \eqref{vankappa37}, \eqref{def-h(q)} and \eqref{realspC1}
4216: we arrive at
4217: \beq\label{fori37}
4218: f(q)\ =\ - \tfrac{1}{6} \kappa_{\kappa \lambda \mu} q^\kappa q^\lambda q^\rho
4219: + \tfrac{1}{2} \kappa_{\kappa kl} q^\kappa q^k q^l\ .
4220: \eeq
4221:
4222: In order to continue
4223: we also have to specify the range the indices $k$ and $\lambda$
4224: take on the IIA side.
4225: A priori it is not fixed and can be changed by a symplectic transformation.
4226: Mirror symmetry demands
4227: \beq \label{na-nb}
4228: k = 1,\ldots, h^{(1,1)}_-(\tilde Y)\ , \qquad
4229: \lambda = 1,\ldots,h^{(1,1)}_+(\tilde Y)\ ,
4230: \eeq
4231: or in other words there have to be $h^{(1,1)}_-(\tilde Y)$
4232: basis elements $\alpha_k$ and $h^{(1,1)}_+(\tilde Y)$ basis elements
4233: $\beta^\lambda$ in $H^3_+(Y)$. In addition the
4234: non-vanishing couplings $\kappa_{\kappa \lambda \mu}$ and
4235: $\kappa_{\kappa l m}$
4236: have to be identified with
4237: $\cK_{\kappa \lambda \mu}$ and $\cK_{\kappa l m}$ appearing
4238: in the definition of the type IIB chiral coordinates \eqref{tau}.
4239: With these conditions fulfilled
4240: we can insert \eqref{fori37} into \eqref{c-in-q37} and compare with
4241: \eqref{tau}. This leads to the identification
4242: \beq
4243: N^{\kh} = (\tau, G^k) \qquad \textrm{and}\qquad
4244: T_{\lambda}^A = 2 T_{\lambda}^B\ ,
4245: \eeq
4246: which in terms of the Kaluza-Klein variables corresponds to
4247: \bea\label{phi=g}
4248: e^{\phi_B}&=& g_A \ ,\qquad q^\lambda\ =\ v^\lambda\ ,\qquad q^k\ =\ -b^k\ ,\nn\\
4249: \xi_0 &=& 2 C_0\ , \quad \xi^k=2(c^k-C_0 b^k)\ , \\
4250: \tilde \xi_\lambda &=& 2 \rho_\lambda - 2\cK_{\lambda kl}c^k b^l +
4251: C_0 \cK_{\lambda kl}b^k b^l\ .\nn
4252: \eea
4253: With these identifications one immediately shows
4254: $e^{D_A} = e^{D_B}$, where $e^{D_A}$ and
4255: $e^{D_B}$ are the four-dimensional dilatons of the type IIA and IIB theory.
4256: This implies that the K\"ahler potentials \eqref{intKQ} and \eqref{kaehlerpot-Kk}
4257: of the two theories coincide in the large volume --
4258: large complex structure limit. However, the corrections
4259: away from this limit cannot be properly understood
4260: from a pure supergravity analysis. It is clear that
4261: $K^{\rm Q}_A$ includes corrections of the mirror IIB
4262: theory but the precise nature of these corrections remains to be understood.
4263:
4264:
4265: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4266: %
4267: % The Mirror of O5/O9
4268: %
4269: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4270:
4271: \subsubsection{The Mirror of IIB orientifolds with $O5/O9$ planes}
4272: \label{O5O9mirror}
4273:
4274: In this section we check mirror symmetry for type IIB orientifolds with
4275: $O5/O9$ planes with complex coordinates and K\"ahler potential determined
4276: in section \ref{Kpo_gaugeIIB}.
4277: In order to find the same chiral data on the IIA side, we have to examine the
4278: case where $\alpha_0 \in H^3_{-}$. Therefore we choose a basis
4279: $(\alpha_k,\beta^{\hat \lambda})$ of $H^3_+$ and $(\alpha_{\hat \lambda},\beta^k)$
4280: of $H^{3}_-$. We rewrite the combination $C\Omega$ in this basis as
4281: \beq
4282: C\Omega = g_A^{-1}(i\, \alpha_0 + i q^\lambda \alpha_\lambda + q^k \alpha_k) + \ldots
4283: \eeq
4284: where we introduced the real special coordinates
4285: \beq \label{realspC2}
4286: g_A =\frac{1}{\I(CZ^0)}\ ,\qquad q^k = \frac{\R(CZ^k)}{\I(CZ^0)}\ , \qquad q^\lambda = \frac{\I(CZ^\lambda)}{\I(CZ^0)}\ .
4287: \eeq
4288: Let us also express the prepotential $\cF(Z)$ in terms of $q^k,q^\lambda$. As in $N=2$ one defines a
4289: function $f(q)$ such that
4290: \beq \label{def-h59}
4291: \cF\big(\R[CZ^k],i\I[CZ^{\hat\lambda}] \big) =- i\big(\I[ CZ^0]\big)^2\, f(q^k,q^\lambda) \ .
4292: \eeq
4293: We can now rewrite the $N=1$ coordinates $T_{\hat\lambda}, N^k$
4294: given in \eqref{def-NT} in terms of
4295: $q^k,q^\lambda$ and $g_A$ as
4296: \bea \label{c-in-q59}
4297: N^k &=& \tfrac{1}{2} \xi^k + i g^{-1}_A q^k \ , \qquad T_\lambda = i \tilde \xi_\lambda +2 g^{-1}_A f_\lambda(q)\ , \nn\\
4298: T_0 &=& i \tilde \xi_0 + 2 g^{-1}_A (2f(q)- f_\lambda q^\lambda - f_k q^k)\ ,
4299: \eea
4300: where $f_\lambda,f_k$ are the first derivatives of $f(q)$ with respect to $q^\lambda$ and $q^k$.
4301:
4302: Going to the large complex structure limit, the $N=2$ prepotential takes the form
4303: \eqref{N=2pre}. We split the indices as $K=(k,\hat \lambda)$ and apply the constraints
4304: \eqref{Z=0gen} to find that
4305: \beq \label{vankappa59}
4306: \kappa_{\kappa \lambda \mu} = \kappa_{\kappa k l} = 0 \qquad \kappa_{klm} \neq 0\ ,\qquad
4307: \kappa_{\kappa \lambda l} \neq 0\ .
4308: \eeq
4309: Using \eqref{vankappa59} and \eqref{def-h59} we can calculate $f(q)$ as
4310: \beq
4311: f(q) = \tfrac{1}{6} \kappa_{ k l m} q^k q^l q^m - \tfrac{1}{2} \kappa_{\kappa \lambda k} q^\kappa q^\lambda q^k\ .
4312: \eeq
4313: In order to match the chiral coordinates $T_0,T_\lambda,N^k$
4314: with the type IIB coordinates
4315: of \eqref{Kcoord} we need again to specify the range of the indices
4316: on the type IIA side. Obviously we need
4317: \beq \label{na-nb59}
4318: k=1, \ldots, h^{(1,1)}_+(\tilde Y)\ , \qquad \lambda= 1,\ldots, h^{(1,1)}_-(\tilde Y)\ ,
4319: \eeq
4320: which is the equivalent of \eqref{na-nb} with the plus and minus sign interchanged.
4321: Thus the non-vanishing intersections can be identified with
4322: $\cK_{klm}$ and $\cK_{\kappa\lambda k}$ on the IIB side.
4323: Inserting $f(q)$ back into the equations \eqref{c-in-q59} for the chiral
4324: coordinates $N^k,T_{\hat \lambda}$ and demanding \eqref{na-nb59} one can
4325: compare these to the type IIB coordinates \eqref{Kcoord}.
4326: One identifies
4327: \beq
4328: T_{\hat \lambda} = 2(S,A_\lambda)\ ,\qquad N^{k} = it^k \ .
4329: \eeq
4330: In terms of the Kaluza-Klein modes this amounts to the identification
4331: \bea
4332: g_A &=& e^{\phi_B}\ , \qquad q^k = -v^k\ , \qquad
4333: q^\lambda = b^\lambda\ ,\qquad
4334: \xi^k = -2 c^k\ , \nn \\
4335: \tilde \xi_\lambda &=& 2\rho_\lambda - 2 \cK_{\lambda \kappa l} c^l b^\kappa\ , \qquad
4336: \tilde \xi_0 = 2h + \cK_{l\lambda \kappa} c^l b^\lambda b^\kappa - \rho_\lambda b^\lambda\ .
4337: \eea
4338: With these identifications one shows again $e^{D_A} = e^{D_B}$ and as
4339: a consequence the K\"ahler potentials \eqref{intKQ} and \eqref{kaehlerpot-KkO5} agree
4340: in the large volume -- large complex structure limit.
4341:
4342:
4343: In summary, we found that it is indeed possible to obtain both type IIB
4344: setups as mirrors of the type IIA orientifolds.
4345: In analogy to \eqref{pure-spinor-map}
4346: we found by comparing \eqref{N=1coords} with \eqref{def-coordsO3} and \eqref{def-coordsO5} the mirror relation
4347: \bea\label{pure-spinor-map2}
4348: O3/O7: & \quad \podd\ \leftrightarrow\ \pev\ , &\qquad C_3 \leftrightarrow \fa\ ,
4349: \nn \\
4350: O5/O9: & \quad \podd\ \leftrightarrow\ -i \pev \ ,& \qquad C_3 \leftrightarrow \fa\ ,
4351: \eea
4352: where $\podd,\pev$ and $\fa$ are defined in \eqref{def-podd} and \eqref{def-A}.
4353: Furthermore, we found that the functionals $\Phi_A$ and $\Phi_B$ have to identified
4354: as
4355: \beq \label{mirror-hitchin}
4356: O3/O7: \quad \Phi_A(\fu) \leftrightarrow \ \Phi_B(\fe)\ , \qquad
4357: O5/O9: \quad \Phi_B(\fu) \leftrightarrow \ \Phi_B(\feh)\ ,
4358: \eeq
4359: such that the K\"ahler potentials are matched. However, the crucial role of the two
4360: definitions of special coordinates remains to be understood further.
4361:
4362: Let us close this chapter with a brief remark on the generalizations of this result.
4363: Formulated in this abstract fashion equations \eqref{pure-spinor-map2} and \eqref{mirror-hitchin}
4364: are expected to hold even for orientifolds of generalized complex manifolds. This includes certain
4365: $SU(3)$ structure manifolds, such as half-flat manifolds. This looks very promising and
4366: deserves further investigation \cite{GLprep}.
4367:
4368:
4369: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4370: %
4371: % Linear multiplets and the geometry of the moduli space
4372: %
4373: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4374:
4375:
4376: \chapter{Linear multiplets and the geometry of the moduli space}
4377: \label{lin_geom_of_M}
4378:
4379: In this chapter we explore the geometry of the $N=1$ moduli
4380: space in more detail. Our attempt is to get some deeper understanding
4381: of the properties of the K\"ahler manifolds obtained from the
4382: $N=2$ to $N=1$ reduction performed in the previous chapter. Recall
4383: that the orientifold moduli space is a direct product
4384: \beq \label{mod-spaceN=1}
4385: \tilde \cM^{\rm SK} \times \tilde \cM^{\rm Q}\ ,
4386: \eeq
4387: where $N=1$ supersymmetry demands each factor to be a K\"ahler manifold.
4388: $\tilde \cM^{\rm SK}$ is a submanifold of the $N=2$ special K\"ahler
4389: manifold $\cM^{\rm SK}$ parameterizing complex structure deformations
4390: in type IIB and complexified K\"ahler structure deformations in type IIA.
4391: As we have shown also $\tilde \cM^{\rm SK}$ is special K\"ahler, since
4392: it inherits its complex structure from $\cM^{\rm SK}$ and admits
4393: a K\"ahler metric obtained from a prepotential.
4394:
4395: The reduction of the hypermultiplet
4396: sector is more `radical' since it defines a K\"ahler manifold $\tilde \cM^{\rm Q}$
4397: inside of a quaternionic manifold $\cM^{\rm Q}$, which itself is not necessarily K\"ahler.
4398: This K\"ahler submanifold has half the dimension of the quaternionic space.
4399: In general it is a difficult
4400: mathematical problem to characterize K\"ahler manifolds inside quaternionic
4401: ones \cite{AM}. However, the quaternionic manifolds obtained by Calabi-Yau compactifications
4402: of type IIA or type IIB supergravity posses special properties. As shown
4403: in \cite{CFGi,FS} they can be constructed from special K\"ahler manifold $\cM^{\text{SK}}$ via the
4404: local c-map,
4405: \beq \label{c-map}
4406: \cM^{\rm SK}_{2n} \quad \xrightarrow{\text{c-map}}\quad \cM^{\rm Q}_{4n+4}\ ,
4407: \eeq
4408: where $2n$ and $4n+4$ are the real dimensions of $\cM^{\rm SK}$ and $\cM^{\rm Q}$.
4409: These quaternionic manifolds are termed special or dual quaternionic.
4410: One observes that their metric depends on only half of the bosonic fields in the
4411: hypermultiplets, or, in other words, on half of the quaternionic coordinates.
4412: More precisely, the components of the metrics \eqref{q-metr} and \eqref{q-metrB}
4413: on $\cM^{\rm Q}$ are functions of only NS-NS scalar fields $M^I_{\text{NS}}$.
4414: The second half are R-R scalar fields denoted by $M_{I\, \text{RR}}$ which appear in the
4415: quaternionic metrics only as a differential and hence posses Peccei-Quinn shift symmetries
4416: \beq
4417: M_{I\, \text{RR}} \rightarrow M_{I\, \text{RR}} + c_I\ ,
4418: \eeq
4419: for arbitrary constants $c_I$.
4420:
4421: The orientifold projection truncates half of the NS-NS fields and half of
4422: the R-R fields. $N=1$ supersymmetry forces the remaining fields to span a K\"ahler
4423: manifold $\tilde \cM^{\rm Q}$. Furthermore, it can be seen in tables \ref{N=1spectrumtab}
4424: and \ref{N=1spectrumA}
4425: that supersymmetry combines each NS-NS field $M^I_{\text{NS}}$ together with a R-R field
4426: $M_{I\, \text{RR}}$ into a chiral multiplet with bosonic
4427: components $M^I = (M^I_{\text{NS}}, M_{I\, \text{RR}})$ spanning $\tilde \cM^{\rm Q}$.
4428: The fact, that the R-R fields posses shift symmetries allows us to
4429: chose a set $M_{\alpha \, \text{RR}}$ and dualize them into two-tensors
4430: $D^{\alpha}_{2\, \text{RR}}$. This amounts to replacing
4431: the chiral multiplets $M^\alpha$ by linear multiples
4432: $L^\alpha=(M^\alpha_{\text{NS}},D^{\alpha}_{2\, \text{RR}})$, while keeping the
4433: remaining fields $M^a$ chiral. The manifold $\tilde \cM^{\rm Q}_{L^\alpha}$
4434: spanned by the real scalars $M^\alpha_{\text{NS}}$ and the complex scalars $M^a$
4435: still contains all the information about the full K\"ahler space $\tilde \cM^{\rm Q}$. In
4436: that one can construct $\tilde \cM^{\rm Q}$ starting from $\tilde \cM^{\rm Q}_{L^\alpha}$,
4437: \beq \label{dual-map}
4438: \tilde \cM^{\rm Q}_{L^\alpha} \quad \xrightarrow{\ \text{dualization of } D^\alpha_2\ }\quad \tilde \cM^{\rm Q}\ .
4439: \eeq
4440: This dualization procedure will be discussed in section \ref{linear_multiplets}.
4441: As we will explain there, the kinetic terms and couplings of the chiral and linear multiplets
4442: can be encoded by a single function $\tilde K$ being the Legendre transform of the K\"ahler potential.
4443: As an application we determine $\tilde K$ for all three orientifold setups. Firstly, in
4444: section \ref{IIB_lin} we apply the linear multiplet formalism to IIB orientifolds. Secondly,
4445: in section \ref{IIA_lin} we provide the missing calculation of the K\"ahler potential for $\tilde \cM^{\rm Q}$ for
4446: general IIA orientifolds. In this derivation we apply the techniques connected with
4447: the map \eqref{dual-map}.
4448:
4449:
4450: Finally, recall that the quaternionic space can be obtained from $\cM^{\rm SK}$
4451: via the local c-map construction \eqref{c-map}. In section \ref{geom_of_modspace}
4452: we construct the map
4453: \beq \label{N=1c-map}
4454: \cM^{\rm SK} \cap \tilde \cM^{\rm Q} \quad \xrightarrow{N=1 \text{ c-map} }\quad \tilde \cM^{\rm Q}\ ,
4455: \eeq
4456: which can be interpreted as the $N=1$ analog of the local c-map \eqref{c-map}.
4457: As we will show it is closely related to the dualization
4458: in \eqref{dual-map}, when specifying the right chiral fields $M^\alpha$ for dualization.
4459: This construction is inspired by the one presented in \cite{Hitchin2}, where
4460: the moduli space of Lagrangian submanifolds with $U(1)$ connection is discussed.
4461: Furthermore, it provides the basis to extend the analysis to non-Calabi-Yau orientifolds.
4462:
4463:
4464:
4465:
4466: \section{Linear multiplets and Calabi-Yau orientifolds\label{linear_multiplets}}
4467:
4468: In this section we rewrite the bulk effective action of type IIB and type IIA orientifolds
4469: using the linear multiplet formalism of ref.\ \cite{BGG}.
4470: In this way we will be able to understand the definition of the K\"ahler
4471: coordinates given in \eqref{tau}, \eqref{Kcoord} and \eqref{def-NT} as a superfield duality transformation
4472: and furthermore discover the no-scale properties of $K^{\rm Q}$
4473: somewhat more conceptually. In an analog three-dimensional situation this has
4474: also been observed in \cite{BHS}.
4475:
4476: Let us first briefly review $N=1$ supergravity coupled to $n$ linear multiplets
4477: $L^\alpha, \alpha=1,\ldots, n$ and
4478: $r$ chiral multiplets $N^A, A=1,\ldots,r$ following \cite{BGG}.
4479: Linear multiplets are defined by the constraint
4480: \beq\label{linearc}
4481: (D^2-8\bar R) L^\alpha = 0 = (\bar D^2-8R) L^\alpha\ ,
4482: \eeq
4483: where $D$ is the superspace covariant derivative and $R$ is the chiral
4484: superfield containing the curvature scalar.
4485: As bosonic components $L$ contains a real scalar field which we also
4486: denote by $L$ and the field strength of a
4487: two-form $D_2$.
4488: The superspace Lagrangian (omitting the gauge interactions) is given by
4489: \beq\label{actionL}
4490: S = - 3 \int E\, F(N^A,\bar N^A, L^\alpha)
4491: + \frac12 \int \frac{E}{R}\, e^{K/2}\ W(N)
4492: + \frac12 \int \frac{E}{R^\dagger}\, e^{K/2}\ \bar W(\bar N)
4493: \ ,
4494: \eeq
4495: where $E$ is the super-vielbein and $W$ the superpotential.
4496: The function $F$
4497: depends implicitly on the K\"ahler potential
4498: $K(N^A,\bar N^A, L^\alpha)$
4499: through the differential constraint\footnote{Strictly speaking
4500: $K(N^A,\bar N^A, L^\alpha)$ is not a K\"ahler potential
4501: but as we will see it determines the kinetic terms in the action.}
4502: \bea\label{Fcon}
4503: 1- \frac{1}{3}L^\alpha K_{L^\alpha} = F-L^\alpha F_{L^\alpha}\ ,
4504: \eea
4505: which ensures the correct normalization of the Einstein-Hilbert term.
4506: The subscripts on $K$ and $F$ denotes differentiation, i.e.\
4507: $K_{L^\alpha} = \frac{\partial K}{\partial L^\alpha},
4508: F_{L^\alpha} = \frac{\partial F}{\partial L^\alpha}$, etc.\ .
4509: Let us also define the kinetic potential $\tilde K$ and rewrite \eqref{Fcon} as
4510: \beq \label{kinpo-def}
4511: \tilde K = K - 3 F\ ,\qquad F = 1-\tfrac{1}{3} \tilde K_{L^\alpha} L^\alpha\ .
4512: \eeq
4513: Expanding \eqref{actionL} into components one finds that $\tilde K$
4514: determines the kinetic terms of the fields. More precisely,
4515: the (bosonic) component Lagrangian derived from \eqref{actionL}
4516: is found to be\footnote{This is a straightforward generalization
4517: of the Lagrangian for one linear multiplet given in \cite{BGG}.
4518: The potential for this case has also been given in \cite{HL}.}
4519: \bea\label{kinetic_lin}
4520: \cL &=& -\tfrac{1}{2}R*\mathbf{1} -
4521: \tilde K_{A\bar B}\, dN^A \wedge * d \bar N^{B}
4522: + \tfrac{1}{4} \tilde K_{L^\alpha L^\beta}\,
4523: dL^\alpha \wedge * dL^\beta - V * \mathbf{1}\nn\\
4524: && + \tfrac{1}{4} \tilde K_{L^\alpha L^\beta}\, dD^\alpha_2 \wedge * dD^\beta_2
4525: - \tfrac{i}2\, dD^\alpha_2 \wedge
4526: \big(\tilde K_{\alpha A}\, dN^A -\tilde K_{\alpha \bar A}\,d\bar N^A\big)
4527: \ ,
4528: \eea
4529: where
4530: \beq\label{Lsc}
4531: %\tilde K_{A\bar B} \equiv K_{A\bar B}- 3 L^\alpha \zeta^R_{\alpha,A\bar B}
4532: %\ , \qquad
4533: V = e^K \Big(\tilde K^{A \bar B}D_AW D_{\bar B}\bar W -
4534: (3- L^\alpha K_{L^\alpha}) |W|^2 \Big)\ .
4535: \eeq
4536: We see that the function
4537: $\tilde K(N,\bar N, L) = K - 3 F$ determines the kinetic terms of the fields
4538: $N^A$ and $L^\alpha$ as well as the couplings of the two-forms $D^\alpha_2$ to
4539: the chiral fields $N^I$. Note that only derivatives of $F_{L^\alpha}$ appear leaving a
4540: constant piece in $F_{L^\alpha}$ undetermined. This constant
4541: drops out from \eqref{Fcon}.
4542:
4543: In a next step we like to recover the
4544: standard $N=1$ effective action by dualizing the linear
4545: multiplets $L^\alpha$ into chiral multiplets $T_\alpha$.
4546: This establishes the map \eqref{dual-map}, which
4547: will be a useful tool in the remainder of this chapter.
4548: From here we can proceed in two ways.
4549: We can dualize the two-forms $D^\alpha_2$
4550: in components and show that the resulting action is
4551: K\"ahler by determining the K\"ahler potential and
4552: complex coordinates.
4553: This is done in appendix \ref{linm} and provides a simple,
4554: but somehow more tedious dualization procedure.
4555: However, performing the duality in superspace yields
4556: directly the proper K\"ahler coordinates $T_\alpha$ and
4557: K\"ahler potential $K(T,\bar T,N,\bar N)$.
4558:
4559: The duality transformation in superfields is
4560: performed in detail in \cite{BGG} and here we only repeat the
4561: essential steps.
4562: One first considers the linear multiplets $L^\alpha$ to be
4563: unconstrained real superfields and modifies the action
4564: \eqref{actionL} to read\footnote{We omit the superpotential
4565: terms here since they only depend on $N$ and play no role
4566: in the dualization.}
4567: \beq\label{actionX}
4568: S = - 3 \int E\, \Big(F(N^A,\bar N^A, L^\alpha) +
4569: 6 L^\alpha(T_\alpha + \bar T_\alpha) \Big) + \ldots\ ,
4570: \eeq
4571: where the $T_\alpha$ are chiral superfields and in order to be consistent
4572: with our previous conventions we have included a factor $6$
4573: in the second term.
4574: Variation with respect to $T_\alpha$ results in the constraint that $L^\alpha$ are linear multiplets
4575: and one arrives back at the action \eqref{actionL}.
4576: Variation with respect to the (unconstrained) $L^\alpha$ yields the
4577: equations\footnote{Notice that there is a misprint
4578: in the equivalent equation given in \cite{BGG}.}
4579: \beq \label{bGl}
4580: 6 (T_\alpha + \bar T_\alpha) + F_{L^\alpha}
4581: - \tfrac{1}{3} K_{L^\alpha}
4582: \big(F+ 6 L^\beta (T_\beta + \bar T_\beta)\big) =0 \ ,
4583: \eeq
4584: where we have used
4585: $\delta_{L} E = -\tfrac{1}{3} E K_{L^\alpha} \delta L^\alpha$.
4586: This equation determines
4587: $L^\alpha$ in terms of the chiral superfields $N^A,T_\alpha$ and is the looked
4588: for duality relation.
4589: However, depending on the specific form of $F$ and $K$
4590: one might not be able to solve \eqref{bGl} explicitly
4591: for $L^\alpha$ but instead only obtain an implicit
4592: relation $L^\alpha(N,\bar N, T+\bar T)$.
4593: Nevertheless one should
4594: insert $L^\alpha(N,\bar N, T+\bar T)$ back into \eqref{actionX}
4595: which then expresses the Lagrangian (implicitly) in terms
4596: of $T_\alpha$ and therefore defines a Lagrangian in the chiral superfield
4597: formalism.
4598: The unusual feature being that the explicit functional dependence is
4599: not known. A correctly normalized Einstein-Hilbert term is ensured by
4600: additionally imposing
4601: \beq \label{normeq}
4602: F(N,\bar N,L) + 6 L^{\alpha}(T_\alpha + \bar T_\alpha) = 1\ .
4603: \eeq
4604: Contracting \eqref{bGl} with $L^\alpha$ and using equation \eqref{normeq} one obtains
4605: \eqref{Fcon}. Thus $F$
4606: has to have the same functional dependence as before
4607: and therefore eqn.~\eqref{kinpo-def} is unmodified, but
4608: one should insert $L(N,\bar N,T+\bar T)$ implicitly
4609: determined by \eqref{bGl}. Using \eqref{normeq} the duality
4610: condition \eqref{bGl} can be cast into the form
4611: \beq \label{dual_coords}
4612: T_\alpha + \bar T_\alpha = \tfrac{1}{2}\tilde K_{L^\alpha}\ ,
4613: \eeq
4614: where $\tilde K$ is the kinetic potential defined in \eqref{kinpo-def}.
4615: We also like to rewrite the K\"ahler potential $K\big( L(N,\bar N, T +\bar T),N,\bar N\big)$
4616: in terms of $\tilde K$. Inserting \eqref{dual_coords} into \eqref{kinpo-def} one infers
4617: \beq \label{Kpot_dual}
4618: K(N,\bar N, T+\bar T) = \tilde K(N,\bar N,L) - 2(T_\alpha + \bar T_\alpha) L^\alpha\ ,
4619: \eeq
4620: where we removed a constant factor by means of a K\"ahler transformation.
4621: Equation \eqref{dual_coords} identifies $T_\alpha + \bar T_\alpha$ to be the canonical
4622: conjugate to $L^\alpha$ with respect to $\tilde K$, while by \eqref{Kpot_dual} the
4623: K\"ahler potential $K$ is the Legendre transform of $\tilde K$.
4624: The equations \eqref{dual_coords} and \eqref{Kpot_dual}
4625: characterize the map \eqref{dual-map} and can
4626: be equivalently obtained by a component field dualization as shown in appendix \ref{linm}.
4627: Before turning to the orientifold examples let us calculate the the bosonic
4628: effective action in terms of $\tilde K$ and the coordinates
4629: \beq \label{coordinates}
4630: N^A\ ,\qquad T_\alpha=i\tilde \xi_\alpha
4631: + \tfrac{1}{4} \tilde K_{L^\alpha}\ ,
4632: \eeq
4633: where $\tilde \xi_\alpha$ is the scalar dual to $D_2^\alpha$ and we have used \eqref{dual_coords}.
4634: Using the K\"ahler potential \eqref{Kpot_dual} one obtains
4635: \bea \label{dual_lagra}
4636: \cL &=& -\tfrac{1}{2}R*\mathbf{1} -
4637: \tilde K_{N^k\bar N^l}\, dN^k \wedge * d \bar N^{l}
4638: + \tfrac{1}{4} \tilde K_{L^\kappa L^\lambda}\,
4639: dL^\kappa \wedge * dL^\lambda - V * \mathbf{1} \\
4640: && + 4 \tilde K^{L^\kappa L^\lambda} \Big(d\tilde \xi_\kappa - \tfrac{1}2
4641: \I \big(\tilde K_{L^\kappa N^l}\,dN^l\big)\Big)\wedge *
4642: \Big(d\tilde \xi_\lambda - \tfrac{1}2
4643: \I \big(\tilde K_{L^\lambda N^k}\,dN^k\big)\Big) \ .\nn
4644: \eea
4645: where $\tilde K$ is the kinetic potential appearing in \eqref{Kpot_dual}.
4646: This is the dual Lagrangian to \eqref{kinetic_lin} as can be equivalently shown
4647: by component field dualization (equation \eqref{eff_act1}).
4648: We now give some explicit examples for this dualization, by applying it
4649: to the Calabi-Yau orientifolds studied in chapter \ref{effective_actO}.
4650:
4651:
4652: \subsection{Two simple examples: Type IIB orientifolds \label{IIB_lin}}
4653:
4654: \subsubsection{I.\quad $O3/O7$ orientifolds}
4655:
4656: Let us now restrict to simple potentials $K(N,\bar N, L)$ and
4657: $F(N,\bar N, L)$, which describe the correct kinematics
4658: for $O3/O7$ orientifolds. Our aim is to rewrite the action \eqref{S_scalarO3}
4659: in the linear multiplet formalism. As we are going to show this enables us to
4660: circumvent the implicit definition of the K\"ahler potential \eqref{kaehlerpot-O7-1}.
4661: In other words, replacing the chiral multiplets $T_\alpha$ with linear multiplets $L^\alpha$
4662: as just described allows us to give an explicit expression for $K$ in terms of $\tau,z$ and
4663: $L^\alpha$ \cite{TGL1}. This is achieved by the K\"ahler potential
4664: \beq\label{KL}
4665: K = K_0(N^A, \bar N^{A}) + \alpha\ln (\KK_{\alpha \beta \gamma} L^\alpha L^\beta L^\gamma)\ ,
4666: \eeq
4667: where we
4668: leave $K_0(N^A, N^{\bar A})$ and the normalization constant
4669: $\alpha$ arbitrary for the moment.
4670: Inserting \eqref{KL} into \eqref{Fcon} shows that possible solutions $F$ have the form
4671: \beq\label{FL}
4672: F=1 - \alpha + \tfrac{1}{3}\, L^\alpha \zeta^R_\alpha(N^A, \bar N^{ A}) \ ,
4673: \eeq
4674: where the real functions $\zeta^R_\alpha(N^A, \bar N^{A})$ are
4675: not further determined by \eqref{Fcon}. In that sense the
4676: $\zeta^R_\alpha(N^A, \bar N^{A})$ are additional input functions
4677: which determine the Lagrangian since they appear in the
4678: kinetic potential \eqref{kinpo-def}.
4679: Comparing \eqref{tau} with \eqref{dual_coords} by using \eqref{KL}
4680: and \eqref{FL} we are led to identify\footnote{Strictly
4681: speaking \eqref{dual_coords} only determines the real part
4682: of $T_\alpha$. The imaginary part can be found by comparing
4683: the explicit effective actions \eqref{S_scalarO3} and \eqref{dual_lagra}.}
4684: \beq\label{zetaid}
4685: \alpha = 1\ , \qquad L^\alpha = \tfrac{3}{2} e^{\phi}\, \frac{ v^\alpha}{\KK} \ , \qquad
4686: \zeta^R_\alpha = -\frac{i}{2(\tau-\bar \tau)} \KK_{\alpha b c} (G-\bar G)^b (G- \bar G)^c\ ,
4687: \eeq
4688: where $\zeta^R_\alpha = \zeta_\alpha+\bar \zeta_\alpha$ was already given in \eqref{zetadef}.
4689: Hence, we have shown that the definition of the K\"ahler coordinates in \eqref{tau}
4690: is nothing but the duality relation \eqref{dual_coords} obtained from the superfield
4691: dualization of the linear multiplets $L^\alpha$ to chiral multiplets $T_\alpha$.\footnote{%
4692: The case $\alpha=1$ is a somewhat special situation
4693: in that the function $F$ does not have a constant piece but only the term
4694: linear in $L^\alpha$.
4695: This in turn requires that the $\zeta^R_\alpha$ cannot be chosen zero but that they
4696: have at least a constant piece so that $F$ does not
4697: vanish. This constant is otherwise irrelevant since it
4698: drops out of all physical quantities.
4699: (In a slightly different context
4700: the case $\alpha=1$ has also been discussed in ref.\ \cite{Binetruy}.)}
4701: It remains to determine $K_0$. Comparing \eqref{KL} by using \eqref{zetaid}
4702: with \eqref{kaehlerpot-O7-1} one finds
4703: \beq\label{K0}
4704: K_0 = K_{\rm cs}(z,\bar z) -\text{ln}\big[-i(\tau - \bar \tau)\big] \ .
4705: \eeq
4706: In summary, the low energy effective action for $O3/O7$ orientifolds
4707: can be rewritten by using chiral multiplets $(z^k,\tau,G^a)$ and linear
4708: multiplets $L^\alpha$. This supergravity theory is determined (in the formalism of ref.\
4709: \cite{BGG} and apart from $W$ and $f$ which we can neglect for this discussion)
4710: by the independent functions $K$ and $F$ given in \eqref{KL} and \eqref{FL} together with
4711: \eqref{zetaid} and \eqref{K0}.
4712: Inserted into \eqref{kinpo-def} we determine the kinetic potential
4713: \beq\label{tildeK_O3}
4714: \tilde K(z, \tau, G,L) = K_{\rm cs}(z,\bar z) +
4715: \ln\Big(\frac{1}{2}\frac{\cK_{\alpha \beta \gamma} L^\alpha L^\beta L^\gamma}{l^0} \Big)
4716: - \frac{\cK_{\alpha a b} L^\alpha l^a l^b}{l^0}\ ,
4717: \eeq
4718: where we have defined $l^a = \I G^a$ and $l^0 = \I \tau$.
4719: In the dual formulation where the linear multiplets $L^\alpha$ are dualized
4720: to chiral multiplets $T_\alpha$ the Lagrangian is entirely determined
4721: by the K\"ahler potential given in \eqref{kaehlerpot-O7-1} with the `unusual'
4722: feature that it is not given explicitly in terms of the chiral
4723: multiplets but only implicitly via the constraint \eqref{dual_coords}.
4724: In this sense the orientifold compactifications
4725: (and similarly the compactifications of F-theory on elliptic Calabi-Yau
4726: fourfolds considered in \cite{HL} and section \ref{F-theory}) lead to
4727: a more general class of K\"ahler potentials
4728: then usually considered in supergravity.
4729: In fact the same feature holds for arbitrary $K_0$ and arbitrary $\zeta^R_\alpha$,
4730: such that also $O3/O7$ orientifolds with space-time filling $D3$ and $D7$
4731: branes fall into this class as shown in \cite{GGJL,JL}.
4732:
4733: Furthermore, these `generalized' K\"ahler potentials are all of
4734: `no-scale type' in that they lead to a positive semi-definite potential $V$.
4735: For $\alpha=1$ (and arbitrary $K_0$ and $\zeta_\alpha$)
4736: the K\"ahler potential \eqref{KL} obeys
4737: \beq
4738: L^\alpha K_{L^\alpha} = 3\ ,
4739: \eeq
4740: and hence the
4741: the second term in the potential \eqref{Lsc} vanishes leaving a positive semi-definite
4742: potential with a supersymmetric Minkowskian ground state.
4743: Since in the chiral formulation $K$ cannot even be given explicitly one can
4744: consider such $K$s as a `generalized' class of
4745: no-scale K\"ahler potentials.
4746: The analogous property has also been observed in refs.\ \cite{HL,BBHL,DAFT}.
4747: Finally note with what ease the no-scale property follows in the
4748: linear formulation compared to the somewhat involved computation
4749: in the chiral formulation performed in \cite{TGL1}.
4750:
4751: \subsubsection{II.\quad $O5/O9$ orientifolds}
4752:
4753: As second simple example let us dualize the
4754: effective action \eqref{S_scalarO5} of orientifolds with $O5/O9$ planes.
4755: In this case our motivation is slightly different, since in contrast
4756: to $O3/O7$ orientifolds, the K\"ahler potential is already given explicitly in
4757: terms of the K\"ahler coordinates. Recall however, that type IIB compactified
4758: on a Calabi-Yau naturally admits a double tensor multiplet $(\phi,C_0,B_2,C_2)$
4759: which is truncated to the linear multiplet $L=(\phi,C_2)$ by the $O5/O9$ orientifold
4760: projection. In section \ref{IIB_orientifolds} we dualized $C_2$ to a scalar $h$ and extracted
4761: the K\"ahler potential in the chiral picture. However,
4762: with the techniques presented above, we are now in the position to
4763: formulate this $N=1$ theory by keeping the linear multiplet $L$ \cite{TGL1}.
4764:
4765: Let us determine $\tilde K=K-3F$ encoding the couplings of the
4766: chiral and linear multiplets in \eqref{kinetic_lin}. As we will show
4767: in a moment the potential $K(N,\bar N, L)$ and the function $F(N,\bar N,L)$ are given by
4768: \beq\label{KpotL}
4769: K = K_0 + \text{ln}\, L\ , \qquad F = \tfrac{2}{3} + \tfrac{1}{3}\, L \zeta^R\ ,
4770: \eeq
4771: which is readily checked to be a solution of the normalization condition \eqref{Fcon}.
4772: Comparing equation \eqref{dual_coords} for $S$ by using the Ansatz \eqref{KpotL} with the
4773: definition \eqref{Kcoord} of $S$ one determines $L$ and $\zeta^R$ as
4774: \beq \label{Lzeta_O59}
4775: L = \tfrac{3}{2} e^{\phi}\cK^{-1}\ , \qquad
4776: \zeta^R = \tfrac{1}{4} (A+ \bar A)_a (\text{Re}\N^{-1})^{ab} (A+ \bar A)_b\ .
4777: \eeq
4778: Inserted back into \eqref{KpotL} indeed yields the K\"ahler potential \eqref{O5-Kaehlerpot}
4779: if we identify
4780: \beq \label{KLdetail}
4781: K_0 = K_{\rm cs}(z,\bar z)
4782: - \text{ln}\Big[\tfrac{1}{48}\KK_{\alpha \beta \gamma}(t+\bar t)^\alpha
4783: (t+\bar t)^\beta (t+\bar t)^\gamma \Big]\ ,
4784: \eeq
4785: Thus we have shown that the kinetic terms can consistently
4786: be described either in the chiral- or the linear multiplet formalism
4787: and we have determined the appropriate coordinates.
4788:
4789: Let us supplement our analysis with another formulation of the $O5/O9$ setups.
4790: Namely we like to dualize the chiral multiplet $S$
4791: as well as the chiral multiplets $A_a$ into a linear multiplets $L^0$ and $L^a$.
4792: As we will see, this will be a first case where $F(N,\bar N,L)$ is not linear
4793: in the linear multiplets $L^0,L^a$ in contrast to \eqref{FL} and \eqref{KpotL}.
4794: We will show momentarily that the K\"ahler potential still has the form
4795: \beq \label{K_dualA}
4796: K(z,t,L) = K_0(z,t) + \ln L^0\ ,
4797: \eeq
4798: where $K_0$ is the same as in \eqref{KLdetail}. $F$ can be deduced from
4799: equation \eqref{dual_coords}, which translates to
4800: \beq
4801: \tfrac{1}{2}\tilde K_{L^0} = S + \bar S\ , \qquad \tfrac{1}{2}\tilde K_{L^a} = A_a + \bar A_a
4802: \eeq
4803: Inserting \eqref{K_dualA} and the coordinates $S,A_a$ given in \eqref{Kcoord} one easily concludes that
4804: \beq \label{F_dualA}
4805: L^0=\tfrac{3}{2} e^\phi \frac{1}{\cK}\ ,\qquad L^a= \tfrac{3}{2} e^\phi \frac{b^a}{\cK}\ ,\qquad
4806: F = \tfrac{2}{3} - \tfrac{1}{3} (L^0)^{-1} \cK_{\alpha a b} (t^\alpha + \bar t^\alpha) L^a L^b\ .
4807: \eeq
4808: where $L^0$ is equal to $L$ in \eqref{Lzeta_O59}. Together with \eqref{K_dualA}
4809: this is consistent with the normalization equation \eqref{Fcon}.
4810: Inserting \eqref{K_dualA} and \eqref{F_dualA} into \eqref{kinpo-def} the kinetic potential reads
4811: \beq \label{tildeK_O5}
4812: \tilde K(z,t,A,L) = K_{\rm cs}(z,\bar z)
4813: - \ln\Big(\frac{1}{6} \frac{\cK_{\alpha \beta \gamma} l^\alpha l^\beta l^\gamma}{L^0} \Big)
4814: + 2 \frac{\cK_{\alpha a b} l^\alpha L^a L^b}{L^0}\ ,
4815: \eeq
4816: where we have defined $l^\alpha = \R\, t^\alpha$.
4817:
4818: Let us close this discussion by comparing this
4819: kinetic potential with the one obtained for $O3/O7$ orientifolds in \eqref{tildeK_O3}. They are identical
4820: under the identifications
4821: \beq
4822: \tilde K_{O3/O7}\ \rightarrow\ -\tilde K_{O5/O9}\ ,\qquad L^\alpha \rightarrow\ l^\alpha \ ,\qquad
4823: (l^a,l^0) \rightarrow\ (L^a,L^0)\ .
4824: \eeq
4825: Note however, that this is a rather drastic step, since we identify linear multiplets of the one
4826: theory with chiral multiplets of the other. It would be interesting to explore
4827: this duality in more detail. It corresponds in simple cases to two T-dualities and
4828: manifests itself by a rotation of the forms
4829: \beq
4830: \pev \ \rightarrow\ i\pev\ , \qquad (\fe, \feh)\ \rightarrow\ (-\feh, \fe)\ .
4831: \eeq
4832: This ends our discussion of IIB orientifolds. As we have seen, much of the
4833: underlying K\"ahler geometry can be directly analyzed by simply switching to
4834: the linear multiplet picture.
4835:
4836: \subsection{An involved example: Type IIA orientifolds \label{IIA_lin}}
4837:
4838: Let us now turn to a more involved application of the linear multiplet
4839: formalism or rather the Legendre transform method behind \eqref{dual_coords}
4840: and \eqref{Kpot_dual}. Namely, we will present a more detailed analysis
4841: of the moduli space $\tilde \cM^{\rm Q}$ for type IIA orientifolds \cite{TGL2}.
4842: Our aim is to show that the K\"ahler potential \eqref{intKQ} with coordinates
4843: $T_\lambda,N^k$ introduced in \eqref{Oexp} indeed encodes the correct
4844: low-energy dynamics of the theory obtained by Kaluza-Klein reduction.
4845: Furthermore, we show that $K^{\rm Q}$ always obeys a no-scale
4846: type condition equivalent to \eqref{no-scale2}.
4847:
4848: Let us start by performing the reduction of the ten-dimensional
4849: theory by using the general basis $(\alpha_\Kh,\beta^\Kh)$
4850: introduced in \eqref{decompO2}. It was chosen such that it splits on
4851: $H^3(Y)=H^{3}_+ \oplus H^3_-$ as
4852: \beq \label{basis1}
4853: (\alpha_k,\beta^\lambda) \in H^{3}_+(Y)\ , \qquad (\alpha_\lambda,\beta^k) \in H^{3}_-(Y)\ ,
4854: \eeq
4855: where both eigenspaces are spanned by $h^{2,1}+1$ basis vectors.
4856: As remarked above, we will only concentrate on the moduli space
4857: $\tilde \cM^{\rm Q}$, such that we can set $t^a=0$ and $A^\alpha=0$.
4858: Due to \eqref{fieldtransf}, the ten-dimensional three-form $\hat C_3$ is expanded in
4859: elements of $H^{3}_+(Y)$ as
4860: \beq
4861: \CC_3 = \xi^k(x)\, \alpha_k - \tilde \xi_\lambda(x)\, \beta^\lambda \ ,
4862: \eeq
4863: where $\xi^k, \tilde \xi_\lambda$ are $h^{2,1}+1$ real space-time scalars in
4864: four-dimensions. Inserting this Ansatz into the ten-dimensional effective
4865: action one finds
4866: \bea \label{act2}
4867: S^{(4)}_{\tilde \cM^{\rm Q}} &=& \int -\, d D \wedge * dD -\, G_{K L}(q)\, dq^K \wedge * dq^L
4868: +\tfrac{1}{2} e^{2D}\, \text{Im}\, \cM_{ k l}\,
4869: d\xi^{ k} \wedge * d\xi^{l} \\
4870: && + \tfrac{1}{2} e^{2D}\, (\text{Im}\, \cM)^{-1\ \kappa \lambda}
4871: \big(d\tilde \xi_\kappa - \text{Re}\, \cM_{\kappa l}\,
4872: d\xi^{l} \big)
4873: \wedge * \big(d\tilde \xi_\lambda-\text{Re}\, \cM_{\lambda k}\, d\xi^{ k} \big)\ , \nn
4874: \eea
4875: where compared to \eqref{act1} only the terms involving $\xi^{k},\tilde \xi_\lambda$ have
4876: changed. The metric $G_{K L}(q)$ was introduced in \eqref{def-G}
4877: and is the induced metric on the space of real
4878: complex structure deformations $\cM^{\rm cs}_\bbR$ parameterized by $q^K$.
4879: It remains to comment on the kinetic and coupling terms of the
4880: scalars $\xi^k, \tilde \xi_\lambda$. In the quaternionic metric
4881: \eqref{q-metr} of the $N=2$ theory they couple via the
4882: matrix $\cM_{\Kh \Lh}$ given in \eqref{defM}. Using the split of the symplectic basis
4883: $(\alpha_\Kh, \beta^\Kh)$ as given in \eqref{basis1} and the fact that by Hodge duality
4884: for a form $\gamma \in H^{3}_+$ one finds $ * \gamma \in H^{3}_-$ one concludes
4885: \beq
4886: \text{Re} \cM_{\kappa \lambda}(q) = \text{Re} \cM_{k l}(q) = \text{Im} \cM_{\kappa k}(q) = 0\ ,
4887: \eeq
4888: whereas $\text{Re} \cM_{k \lambda}, \text{Im} \cM_{\kappa \lambda}, \text{Im} \cM_{k l}$
4889: are generally non-zero on $\cM^{\text{cs}}_{\mathbb{R}}$. The explicit form of non-vanishing
4890: components can be obtained by restricting \eqref{gauge-c} to $\cM^{\rm cs}_\bbR$ and
4891: using the constraints \eqref{Z=0gen}.
4892:
4893: In order to combine the scalars $e^D,q^K$ with $\xi^k, \tilde \xi_\kappa$ into
4894: complex variables, we have to redefine these fields and rewrite the first two
4895: terms in \eqref{act2}. Thus we define the $h^{2,1}+1$ real coordinates
4896: \beq \label{lL-def}
4897: L^\lambda\ =\ - e^{2D}\, \I \big[C Z^\lambda(q) \big]\ ,\qquad
4898: l^k \ =\ \R\big[C Z^k(q)\big]\ ,
4899: \eeq
4900: which is consistent with the orientifold constraint
4901: \eqref{Z=0gen}. The additional factor of $e^{2D}$ was included in order
4902: to match the dilaton factors later on.
4903: Using \eqref{lL-def} one calculates the Jacobian matrix
4904: for the change of variables $(e^D,q^K)$ to $(l^k,L^\lambda)$ as
4905: explicitly done in \cite{TGL2}.
4906: It is then straight forward to
4907: rewrite \eqref{act2} by using the
4908: identities \eqref{spconst} of special geometry as
4909: \begin{align} \label{IIA1}
4910: S^{(4)}_{\tilde \cM^{\rm Q}} = & \int 2 e^{-2D} \text{Im} \cM_{\kappa \lambda}\, dL^\kappa \wedge * dL^\lambda
4911: + 2 e^{2D} \text{Im} \cM_{k l}\, dl^k \wedge * dl^l
4912: + \tfrac{e^{2D}}{2} \text{Im} \cM_{ k l}\,
4913: d\xi^{ k} \wedge * d\xi^{l} \nn\\
4914: &+ \tfrac{e^{2D}}{2} \, (\text{Im}\, \cM)^{-1\ \kappa \lambda}
4915: \big(d\tilde \xi_\kappa - \text{Re}\, \cM_{\kappa k}\,
4916: d\xi^{k} \big)
4917: \wedge * \big(d\tilde \xi_\lambda-\text{Re}\, \cM_{\lambda k}\, d\xi^{ k} \big)\ .
4918: \end{align}
4919: From \eqref{IIA1} one sees that the scalars $l^k$ and $\xi^k$ nicely combine
4920: into complex coordinates
4921: \beq
4922: N^k\ =\ \tfrac{1}{2}\xi^k + i l^k\ =\ \tfrac{1}{2}\xi^k + i \R(C Z^k)\ ,
4923: \eeq
4924: which corresponds to \eqref{def-NT}.
4925: In contrast, one observes that
4926: the metric for the kinetic terms of the
4927: scalars $\tilde \xi_\lambda$ is exactly the inverse of the one
4928: appearing in the kinetic terms of the scalar fields $L^\lambda$.
4929: Hence, comparing \eqref{IIA1} with \eqref{dual_lagra} on concludes that this
4930: action is obtained by dualizing a set of linear multiplets
4931: $(L^\lambda, D^\lambda_2)$ into chiral multiplets
4932: $(L^\lambda,\tilde \xi_\lambda)$. To extract $\tilde K(L, N,\bar N)$ we
4933: compare \eqref{IIA1} with \eqref{dual_lagra} and read off the metric
4934: \beq \label{lLmetric}
4935: \tilde K_{L^\kappa L^\lambda}\ =\ 8\, e^{-2D} \IM_{\kappa \lambda}\ , \quad
4936: \tilde K_{l^k l^l}\ =\ -8\, e^{2D} \IM_{k l} \ , \quad
4937: \tilde K_{L^\kappa l^l} \ =\ - 8\, \RM_{\kappa l}\ ,
4938: \eeq
4939: where we have used that the metric is independent of $\xi^k,\tilde \xi_\lambda$.
4940: This metric can be obtained from a kinetic potential of the form
4941: \beq \label{kinpo1}
4942: \tilde K(L,l)\ =\ - \ln\big[ e^{-4D} \big]+ 8e^{2D}\I \big[\rho^*\cF(CZ^k)\big]\ ,
4943: \eeq
4944: where $\cF$ is the prepotential of the special K\"ahler manifold $\cM^{\rm cs}$
4945: restricted to the real subspace $\cM^{\rm cs}_\bbR$. The map $\rho$ was given
4946: in \eqref{embmap1} and enforces the constraints \eqref{Z=0gen}. To show that $\tilde K$
4947: indeed yields the correct metric \eqref{lLmetric} one differentiates \eqref{kinpo1}
4948: with respect to $e^{-D},q^K$ and uses the inverse of the Jacobian matrix
4949: for the change of variables $(e^D,q^K)$ to $(l^k,L^\lambda)$. Applying equations
4950: \eqref{ML-hf} one finds its first derivatives
4951: \beq \label{first-der}
4952: \tilde K_{ L^\lambda} \ =\ - 8\, \R\big[C \cF_\lambda(q) \big] \qquad
4953: \tilde K_{l^k} \ = \ 8\, e^{2D}\, \I\big[C \cF_{k}(q) \big]\ .
4954: \eeq
4955: Repeating the procedure and differentiating \eqref{first-der}
4956: with respect to $e^{-D},q^K$ and using once again the inverse Jacobian
4957: one applies \eqref{def-M} to show \eqref{lLmetric}. Knowing \eqref{kinpo1}
4958: one can also extract the functions $F(L, N, \bar N)$ and $K(L,N,\bar N)$
4959: by applying \eqref{kinpo-def}. As we will show momentarily
4960: $K$ and $F = \frac13 (K- \tilde K)$ are given by
4961: \beq
4962: K(L,l) = - \ln \big[ e^{-4D} \big] \ , \qquad
4963: F(L,l) = - \tfrac{8}{3} e^{2D}\I \big[\rho^*\cF(CZ)\big]+\tfrac{1}{3}\ .
4964: \eeq
4965: It suffices to determine $K$ which expressed in the correct coordinates
4966: serves as the K\"ahler potential in the chiral description.
4967:
4968: As explained in the beginning of this section the actual K\"ahler potential of
4969: $\tilde \cM^{\rm Q}$ is the Legendre transform \eqref{Kpot_dual} of $\tilde K$ with
4970: respect to the variables $L^\lambda$. There we also found the explicit
4971: definition of the complex coordinates $T_\lambda$ combining $(L^\lambda,\tilde \xi_\lambda)$.
4972: Using \eqref{first-der} in \eqref{dual_coords} and fixing the normalization of the
4973: imaginary part of $T_\lambda$ by comparing \eqref{IIA1} with \eqref{dual_lagra}
4974: one finds
4975: \beq
4976: T_\lambda = i \tilde \xi_\lambda + \tfrac{1}{4}\tilde K_{L^\lambda}
4977: = i \tilde \xi_\lambda - 2\, \R\big(C F_\lambda\big) \ ,
4978: \eeq
4979: which coincides with \eqref{def-NT} already quoted in section \ref{Kpo_gaugeIIA}.
4980: To give an explicit expression for $K^{\rm Q}$ we insert
4981: equation \eqref{kinpo1} into \eqref{Kpot_dual}. Applying the $N=2$ identity
4982: $\cF=\frac12 Z^\Kh \cF_\Kh$, the constraint equations \eqref{Z=0gen}
4983: and \eqref{lL-def},\eqref{first-der} we rewrite
4984: \beq \label{K_lL}
4985: K^{\rm Q}= - \ln\big[e^{-4D} \big] + \tfrac{1}{2} (l^k \tilde K_{l^k} - L^\lambda \tilde K_{ L^\lambda})\ .
4986: \eeq
4987: It is possible to evaluate the terms appearing in the parentheses. In order to do that
4988: we combine the equations \eqref{lL-def} and \eqref{first-der} to the simple form
4989: \bea \label{lL}
4990: \R\big( C \Omega \big)\ =\ l^k \alpha_k + \tfrac{1}{8} \tilde K_{L^\lambda} \beta^\lambda\ ,\quad
4991: e^{2D} \I\big( C \Omega \big)\ =\ -L^\lambda \alpha_\lambda - \tfrac{1}{8} \tilde K_{l^k} \beta^k\ .
4992: \eea
4993: We now use equation \eqref{csmetric} and the definition \eqref{def-C} of $C$
4994: to calculate
4995: \beq \label{skconstr}
4996: 2 \int_Y \R( C\Omega) \wedge \I(C\Omega) = i \int_Y C\Omega \wedge \overline{C\Omega} = e^{-2D}\ .
4997: \eeq
4998: Inserting the equations \eqref{lL} into \eqref{skconstr} we find
4999: \bea \label{lL=4}
5000: L^\lambda \tilde K_{L^\lambda} - l^k \tilde K_{l^k} = 4\ .
5001: \eea
5002: Inserted back into \eqref{K_lL} we have shown that the K\"ahler potential
5003: has indeed the form \eqref{intKQ}.\footnote{By using the equation \eqref{skconstr} and $*\Omega=-i\Omega$
5004: it is straight forward to show $e^{-2D}=2\int \R(C\Omega)\wedge * \R(C\Omega)$}
5005: Moreover, \eqref{lL=4} directly translates into a no-scale type condition for $K^{\rm Q}$
5006: \bea \label{no-scale4}
5007: K_{w^\Kh} K^{w^\Kh \bar w^\Lh} K_{\bar w^\Lh} = 4\ ,
5008: \eea
5009: where $w^\Kh=(T_\kappa, N^k)$.
5010: In order to see this, one inserts the inverse K\"ahler metric \eqref{invKm1},
5011: the K\"ahler derivatives \eqref{Kder} and the derivatives of \eqref{lL=4} back into
5012: \eqref{lL=4}. In other words, we were able to translate one of
5013: the special K\"ahler conditions present in the underlying
5014: $N=2$ theory into a constraint on the geometry of
5015: $\tilde \cM^{\rm Q}$. Two non-trivial examples fulfilling
5016: \eqref{lL=4} are the $O3/O7$ and $O5/O9$ kinetic potentials
5017: \eqref{tildeK_O3} and \eqref{tildeK_O5}. They admit this
5018: simple form since instanton corrections are not taken into account.
5019:
5020:
5021:
5022: \section{The geometry of the moduli space}
5023: \label{geom_of_modspace}
5024:
5025: In this section we give an alternative formulation of
5026: the geometric structures of the moduli space $\tilde \cM^{\rm Q}$
5027: which is closely related the moduli space of
5028: supersymmetric Lagrangian submanifolds in a Calabi-Yau
5029: threefold \cite{Hitchin2}.\footnote{This
5030: analysis can equivalently be applied to the moduli space of
5031: $G_2$ compactifications of
5032: M-theory.} In this set-up also
5033: the no-scale conditions \eqref{no-scale2},
5034: \eqref{lL=4} are interpreted geometrically. This provides
5035: a more elegant description of the $N=1$ moduli space and
5036: its special properties. Moreover, we construct the $N=1$
5037: analog \eqref{N=1c-map} of the $N=2$ c-map \eqref{c-map}.
5038: Our analysis can serve as a starting point for the analysis
5039: of non-Calabi-Yau orientifolds by using the
5040: language of generalized complex manifolds invented by
5041: Hitchin \cite{HitchinGCM}.
5042:
5043: In section~\ref{IIA_orientifolds} we started from a $N=2$ quaternionic
5044: manifold $\cM^{\rm Q}$ and determined the submanifold
5045: $\tilde\cM^{\rm Q}$ by imposing the orientifold projection.
5046: $N=1$ supersymmetry ensured that this submanifold is K\"ahler.
5047: $\cM^{\rm Q}$ has a second but different K\"ahler submanifold
5048: $\cM^{\rm cs}$ which intersects with $\tilde\cM^{\rm Q}$
5049: on the real manifold $\cM^{\rm cs}_\bbR$.
5050: The c-map is in some sense the reverse operation where
5051: $\cM^{\rm Q}$ is constructed starting from $\cM^{\rm cs}$
5052: and shown to be quaternionic \cite{CFGi,FS}.
5053: In this section we analogously construct the K\"ahler manifold
5054: $\tilde\cM^{\rm Q}$ starting from $\cM^{\rm cs}_\bbR$.
5055:
5056: \begin{figure}[h]
5057: \begin{center}
5058: \includegraphics[height=4cm]{C-bundle.eps}
5059: \caption{\textit{The local moduli space $\cM_{\bbR} = \cM_{\bbR}^{\rm cs} \times \bbR$ in
5060: $\cM^{\rm cs} \times \bbC \simeq \cM^{\rm cs} \times H^{(3,0)}$.}}
5061: \label{loc_modspace}
5062: \end{center}
5063: \end{figure}
5064:
5065:
5066:
5067: In fact the proper starting point is not $\cM^{\rm cs}_\bbR$ but rather
5068: $\cM_\bbR=\cM_\bbR^{\rm cs} \times \bbR$ which is the local product of the
5069: moduli space
5070: of real complex structure deformations of a Calabi-Yau orientifold
5071: times the real dilaton direction. The $N=2$ analog of
5072: $\cM_\bbR$ is the extended moduli space
5073: $\hat\cM^{\rm cs} = \cM^{\rm cs} \times \bbC$ where $\bbC$
5074: is the complex line normalizing $\Omega$. The corresponding modulus
5075: can be identified with the complex dilaton \cite{Witten2}.
5076: The orientifold projection fixes the phase of the complex dilaton
5077: (it projects out the four-dimensional $B_2$) to be $\theta$ and thus reduces
5078: $\bbC$ to $\bbR$ (figure \ref{loc_modspace}).
5079: The local geometry of $\cM_\bbR$ is encoded in the variations of the real and imaginary part of the normalized holomorphic three-form $C\Omega$.
5080: This form naturally defines an embedding
5081: \beq
5082: E: \cM_\bbR \rightarrow V \times V^*
5083: = \ H^3_+(\mathbb{R}) \times H^3_-(\mathbb{R})\ .
5084: \eeq
5085: where $V =H^3_+(\mathbb{R})$ and we used the intersection
5086: form $\big<\alpha,\beta \big>=\int \alpha \wedge \beta $ on $H^{3}(Y)$
5087: to identify $V^* \cong H^3_-(\mathbb{R})$.
5088: $V \times V^*$ naturally admits a
5089: symplectic form $\cW$ and an indefinite metric $\cG$ defined as
5090: \bea
5091: \cW((\alpha_+,\alpha_-),(\beta_+,\beta_-)) = \big<\alpha_+, \beta_-\big> - \big<\beta_+ ,\alpha_-\big>\ , \nn\\
5092: \cG((\alpha_+,\alpha_-),(\beta_+,\beta_-)) = \big<\alpha_+, \beta_-\big> + \big<\beta_+ ,\alpha_-\big>\ ,
5093: \eea
5094: where $\alpha_\pm,\beta_\pm \in H^3_\pm(\mathbb{R})$.
5095:
5096: Now we construct $E$ in such a way that
5097: $\cM_{\bbR}$ is a Lagrangian submanifold of $V \times V^*$ with respect to
5098: $\cW$ and its metric is induced from $\cG$, i.e.\
5099: \beq\label{Lagr}
5100: E^*(\cW)=0 \ , \qquad E^*(\cG)=g
5101: \eeq
5102: where
5103: \beq \label{metrQ}
5104: \tfrac{1}{2} g = dD \otimes dD + G_{K L} d q^K \otimes d q^L\
5105: \eeq
5106: is the metric on $\cM_\bbR$ as determined in \eqref{act1}.
5107: As we are going to show momentarily $E$ is given by
5108: \bea \label{embmap}
5109: E(q^\Kh) = 2\,\big(\fu , -e^{2D} \fuh \big)\ ,
5110: \eea
5111: where $\fu+i\,\fuh = C\Omega$, $q^\Kh=(e^{-D},q^K)$ and
5112: $\Omega$ is evaluated at $q^K \in \cM^{\rm cs}_\bbR$.
5113: Additionally $E$ satisfies
5114: \beq \label{no-scale3}
5115: \cG(E(q^\Kh),E(q^\Kh)) = 4\ ,
5116: \eeq
5117: for all $q^K $. This implies that the image of all points in
5118: $\cM_\bbR$ have the same distance from the origin. Later on we will show that
5119: this translates into the no-scale condition \eqref{no-scale4}.
5120:
5121: Let us first show that the $E$ given in \eqref{embmap} indeed satisfies
5122: \eqref{Lagr} and \eqref{no-scale3}. The explicit calculation is straightforward
5123: and essentially included in the calculation presented in section \ref{IIA_lin}.\footnote{Formally one has to
5124: first evaluate $E_* (\partial_{Q^\Kh})$ and expresses the result in terms of the $(3,0)$-form $\Omega$
5125: and the $(2,1)$-forms $\chi_K$. One then uses that by definition of the pull-back
5126: $E^* \omega(\partial_{q^\Kh},\cdot ) = \omega(E_* (\partial_{q^\Kh}),\cdot)$ for a form $\omega$ on $V \times V^*$.
5127: Applied to $\cG$ and $\cW$ one finds that the truncation of the special K\"ahler potential
5128: \eqref{csmetric} and \eqref{chi_barchi} indeed imply \eqref{Lagr}. This calculation does not
5129: make use of any specific basis of $H^3_\pm$.} In order to connect with section
5130: \ref{IIA_lin} let us first recall how we applied the map \eqref{dual-map} to
5131: extract the chiral data of the $N=1$ moduli space.
5132: We started with a special K\"ahler manifold $\cM^{\rm sk}$ with metric determined
5133: in terms of a holomorphic prepotential $\cF(Z)$.
5134: Next we assumed that the $N=2$ theory with quaternionic space $\cM^{\rm Q}$ constructed
5135: via the local c-map \eqref{c-map} allows a reduction to $N=1$. Accordingly the section
5136: $\Omega(z)=Z^\Kh \alpha_\Kh - \cF_\Kh \beta^\Kh$ fulfills equation \eqref{Z=0gen} for some basis
5137: \beq \label{red-basis}
5138: (\alpha_k,\beta^\lambda) \in H^{3}_+ \ ,\qquad (\alpha_\lambda,\beta^k) \in H^{3}_-\ .
5139: \eeq
5140: Using this basis we found the kinetic potential $\tilde K(L,l)$ given in \eqref{kinpo1}, which
5141: explicitly depends on the prepotential $\cF$. It encodes the metric on
5142: $\tilde \cM^{\rm Q} \subset \cM^{\rm Q}$ via the K\"ahler potential \eqref{Kpot_dual}.
5143: On the other hand, equation \eqref{dual_coords} defines the complex structure on
5144: $\tilde \cM^{\rm Q}$.
5145:
5146: These steps can be translated into the language of this section. Namely,
5147: choosing the basis \eqref{red-basis} to expand the map $E$ defined in \eqref{embmap}
5148: one finds
5149: \beq \label{Eincoords}
5150: E(q^\Kh)\ =\ \big(2l^k \alpha_k + \tfrac{1}{4} \tilde K_{L^\lambda} \beta^\lambda,
5151: 2L^\lambda \alpha_\lambda + \tfrac{1}{4} \tilde K_{l^k} \beta^k\big)\ ,
5152: \eeq
5153: where $l^k,L^\lambda$ and $\tilde K_{L^\lambda}, \tilde K_{l^k}$ are functions of $q^\Kh$ as given in
5154: \eqref{lL-def} and \eqref{first-der}.
5155: We define coordinates $u^\Kh=(2l^k, \tfrac{1}{4}\tilde K_{L^\lambda})$ on $V$ and
5156: coordinates $v_\Kh=(\tfrac{1}{4}\tilde K_{l^k},-2L^\lambda)$ on $V^*$.
5157: In these coordinates the first two conditions in \eqref{Lagr} simply read
5158: \beq \label{Lagrc}
5159: E^*(du^\Kh \wedge dv_\Kh)=0\ , \qquad E^*(du^\Kh \otimes dv_\Kh) = g\ .
5160: \eeq
5161: {}From section \ref{IIA_lin} we further know that
5162: $\tilde K_{L^\kappa}, \tilde K_{l^k}$ are derivatives
5163: of a kinetic potential $\tilde K$ and thus we can evaluate $du^\Kh$
5164: and $dv_\Kh$ in terms of $l^k,L^\kappa$.
5165: Inserting the result into \eqref{Lagrc}
5166: the second equation can be rewritten as
5167: \beq
5168: \tfrac{1}{2} g\ =\ \tfrac{1}{4}
5169: \tilde K_{l^k l^l}\, dl^k \otimes dl^l -
5170: \tfrac{1}{4}
5171: \tilde K_{L^\kappa L^\lambda}\, dL^\kappa \otimes dL^\lambda\ ,
5172: \eeq
5173: while the first equation is trivially fulfilled due to the symmetry of $\tilde K_{l^k l^l}$
5174: and $\tilde K_{L^\kappa L^\lambda}$. This metric is exactly the one appearing in the action
5175: \eqref{IIA1} when using \eqref{lLmetric}. Expressing $g$ in coordinates
5176: $e^{D},q^K$ leads to \eqref{metrQ}, as we have already checked by going from
5177: \eqref{act2} to \eqref{IIA1} above.
5178: Furthermore, inserting \eqref{Eincoords} into \eqref{no-scale3}
5179: it exactly translates
5180: into the no-scale condition \eqref{lL=4}, which was shown in section
5181: \ref{IIA_lin} to be equivalent to \eqref{no-scale2}.
5182:
5183: We have just shown that $\cM_\bbR$ is a
5184: Lagrangian submanifold of $V \times V^*$.
5185: Identifying $T^*V \cong V \times V^*$ we conclude that $\cM_\bbR$
5186: can be obtained as the graph $(\alpha(u),u)$
5187: of a closed one-form $\alpha$. This implies that we can locally find
5188: a generating function $K': V \rightarrow \mathbb{R}$ such that $\alpha = dK'$.
5189: In local coordinates $(v_\Kh,u^\Kh)$ this amounts to
5190: \beq \label{v=K/u}
5191: v_\Kh(u) = \frac{\partial K'}{\partial u^{\Kh}}
5192: \eeq
5193: such that
5194: \beq
5195: - L^\kappa(u) = 2\, \frac{\partial K'(u)}{\partial \tilde K_{L^\kappa}}\ , \quad
5196: \tilde K_{l^k}(u) = 2\, \frac{\partial K'(u)}{\partial l^k}\ .
5197: \eeq
5198: These equations are satisfied if we define $K'$ in terms of $\tilde K$ as
5199: \bea \label{Legendre}
5200: 2 K'\ =\ \tilde K(L(u),l) - \tilde K_{L^\kappa}(u)\, L^\kappa(u)\ ,
5201: \eea
5202: which is nothing but the Legendre transform of $\tilde K$ with respect to $L^\kappa$.
5203: Later on we show that the function ${2}K'$ is identified with the K\"ahler potential
5204: $K$ given in \eqref{intKQ}.
5205:
5206: In order to do that, we now extend our discussion to the
5207: full moduli space $\tilde \cM^{\rm Q}$ including the scalars
5208: $\zeta^\Kh=(\xi^k,\tilde \xi_\kappa)$ parameterizing the
5209: three-form $\hat C_3$ in $H^{3}_+(\bbR)$. Locally one has
5210: \beq
5211: \tilde \cM^{\rm Q} = \cM_\bbR \times H^{3}_+(\bbR)\ .
5212: \eeq
5213: The tangent space at a point $p$ in $ \tilde \cM^{\rm Q}$ can be identified as
5214: \beq
5215: T_p \tilde \cM^{\rm Q} \cong H^3_+(\bbR)\oplus H^3_+(\bbR) \cong H^3_+(\bbR) \otimes \bbC\ ,
5216: \eeq
5217: where the first isomorphism is induced by the embedding $E$ given in \eqref{embmap}.
5218: This is a complex vector space and thus $\tilde \cM^{\rm Q}$ admits an
5219: almost complex structure $I$. In components it is given by
5220: \beq \label{def-I}
5221: I(\partial_{q^\Kh}) = (\partial u^\Lh/\partial q^\Kh)\, \partial_{\zeta^\Lh}\ ,\qquad
5222: I((\partial u^\Lh/\partial q^\Kh)\, \partial_{\zeta^\Lh})=-\partial_{q^\Kh}\ ,
5223: \eeq
5224: where we have used that $I$ is induced by the embedding map $E$. One can show that
5225: the almost complex structure $I$ is integrable, since
5226: \beq
5227: dw^\Kh = du^\Kh + i d \zeta^\Kh = (\partial u^\Lh/\partial q^\Kh) dq^\Kh + i d \zeta^\Kh\ ,
5228: \eeq
5229: are a basis of $(1,0)$ forms and $w^\Kh=u^\Kh+i\zeta^\Kh$ are complex coordinates on $\tilde \cM^{\rm Q}$.
5230: Using the definition of $u^\Kh$ one infers that as expected $w^\Kh = (N^k,T_\kappa)$.
5231: Moreover, one naturally extends the metric $g$ on $T \cM_\bbR$ to a hermitian metric
5232: on $T\tilde \cM^{\rm Q}$. The corresponding two-form is then given by
5233: \beq
5234: \tilde \omega(\partial_{\zeta^\Lh}, \partial_{q^\Kh}) = g(I\partial_{\zeta^\Lh},\partial_{q^\Kh})\ ,
5235: \qquad \tilde \omega(\partial_{\zeta^\Kh},\partial_{\zeta^\Lh})= \tilde \omega(\partial_{q^\Kh},\partial_{q^\Lh})=0\ .
5236: \eeq
5237: Using the definition \eqref{def-I} of the almost complex structure and
5238: equation \eqref{Lagr}, one concludes that $\tilde \omega$ is given by
5239: \beq \label{tildeo}
5240: \tilde \omega= dv_\Kh \wedge d\zeta^\Kh
5241: = 2i \frac{\partial^2 K'}{\partial w^\Kh \partial \bar w^\Lh} dw^\Kh \wedge d\bar w^\Lh \ ,
5242: \eeq
5243: where for the second equality we applied \eqref{v=K/u} and expressed the
5244: result in coordinates $w^\Kh=u^\Kh + i \zeta^\Kh$. Note that $K'$ is a function
5245: of $u^\Kh$ only, such that derivatives with respect to $w^\Kh$ translate to ones
5246: with respect to $u^\Kh$. Equation \eqref{tildeo} implies that $K^{\rm Q}=2K'$ is indeed
5247: the correct K\"ahler potential for the moduli space $\tilde \cM^{\rm Q}$.
5248:
5249: So far we restricted ourselves to type IIA orientifolds. However, by using the
5250: mirror map \eqref{pure-spinor-map2} one easily translates the above construction to IIB
5251: setups. In the IIB case the real manifold started with is simply the
5252: local product $\cM^B_\bbR = \cM^{\rm ks}_\bbR \times \bbR$, where $\cM^{\rm ks}_\bbR$
5253: is a real slice in the complexified K\"ahler cone $\cM^{\rm ks}$ and $\bbR$ parameterizes
5254: the four-dimensional dilaton direction. $\cM^{\rm ks}_\bbR$ is locally spanned by
5255: the fields $v^\alpha$ and $b^a$ introduced in section \ref{IIB_orientifolds}.
5256: Once again we aim to find the embedding map $E$
5257: \beq
5258: E: \cM^B_\bbR \rightarrow V \times V^*\ .
5259: \eeq
5260: In order to be more explicit we distinguish $O3/O7$ and $O5/O9$ setups and
5261: define
5262: \beq
5263: E_{O3/7}(q^\Kh)\ =\ 2\, (\fe,\ e^{2D_B} \feh)\ ,\qquad
5264: E_{O5/9}(q^\Kh)\ =\ 2\, (\feh,\ e^{2D_B} \fe)\ ,
5265: \eeq
5266: where $\fe+i\feh = e^{-\phi} e^{-B+iJ}$ and
5267: $q^{\Ah}=(e^{-D_B},v^\alpha,b^a)$. Correspondingly we need to set
5268: \beq
5269: V_{O3/7}=H^{ev}_+\ ,\quad V_{O3/7}^*=H^{ev}_- \qquad
5270: V_{O5/9}=H^{ev}_-\ ,\quad V_{O5/9}^*=H^{ev}_+\ ,
5271: \eeq
5272: where we have abbreviated \footnote{Recall that $H^{(0,0)}_- = H^{(3,3)}_-=0$ as discussed in section
5273: \ref{IIB_orientifolds}.}
5274: \beq \label{Heven}
5275: H^{ev}_+ = H^{(0,0)}_+ \oplus H^{(1,1)}_- \oplus H^{(2,2)}_+ \ ,\qquad
5276: H^{ev}_- = H^{(1,1)}_+ \oplus H^{(2,2)}_- \oplus H^{(3,3)}_+\ .
5277: \eeq
5278: Given a vector space $V$ of even forms, the identification of $V^*$ with the
5279: respective cohomology groups is done by using the intersection form $\big<\cdot,\cdot\big>$
5280: defined in \eqref{symp-form}.
5281: To check that $E_{O3/7}$ and $E_{O5/9}$ are defined correctly, one
5282: proceeds in full analogy to the type IIA case. Once again, the calculation
5283: simplifies considerably by using the existence of the kinetic potentials \eqref{tildeK_O3} and
5284: \eqref{tildeK_O5}.
5285:
5286: Let us summarize our results. We constructed the metric and complex structure of
5287: the K\"ahler manifold $\tilde \cM^{\rm Q} \subset \cM^{\rm Q}$ by specifying a map
5288: \beq
5289: E: \cM_\bbR \rightarrow V \times V^*\ ,
5290: \eeq
5291: where $\cM_{\bbR}$ parameterizes the real four-dimensional dilaton direction times
5292: certain deformations of the Calabi-Yau orientifold. $V$ is an appropriately chosen
5293: vector space
5294: \beq
5295: V_{IIA} = H^{odd}_+\ ,\qquad V_{IIB} = H^{ev}_\pm\ ,
5296: \eeq
5297: where $H^{odd}_+=H^3_+$ and $H^{ev}_\pm$ is given in \eqref{Heven}.
5298: More explicitly $E$ takes the form
5299: \beq
5300: E(q^\Kh)=2\, \big(\rho,- \hat \rho / \Phi_{A,B}\big)\ ,
5301: \eeq
5302: where $\Phi_{A,B}(\rho)$ is given in \eqref{def-Phi}, \eqref{def-PhiO5} and
5303: $\rho=(\fu,\fe,\feh)$ depending on the orientifold setup. In order
5304: to evaluate $\Phi_{A,B}(\rho)=e^{-2D}$ we use the definition of the four-dimensional dilaton
5305: \eqref{def-D_B}.
5306: Since $\cM_{\bbR}$ is embedded as a Lagrange submanifold in $V\times V^*$ it can be locally given
5307: by the graph of the one-form $dK'$. Moreover, since $E$ induces the metric on
5308: $\cM_{\bbR}$ and a complex structure on $\cM_{\bbR}\times V$ the function $2K'$ is nothing but the K\"ahler
5309: potential on the local moduli space $\tilde \cM^{\rm Q}=\cM_{\bbR}\times V$. Thus, the
5310: difficulty is to find the map $E$ or, by recalling \eqref{v=K/u}, the functional dependence
5311: $\hat \rho(\rho)$. This non-linear map
5312: \beq
5313: \rho\ \mapsto\ \hat \rho(\rho)\ ,
5314: \eeq
5315: lies at the heart of Hitchins approach to extract the geometry
5316: of even and odd forms on six-manifolds \cite{Hitchin1,HitchinGCM}.
5317: One may thus hope to generalize Calabi-Yau orientifolds to non-Calabi-Yau orientifolds \cite{GLprep}.
5318:
5319:
5320: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5321: %
5322: % NS-NS and R-R fluxes
5323: %
5324: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5325:
5326:
5327: \chapter{Calabi-Yau orientifolds with NS-NS and R-R background fluxes}
5328: \label{fluxesAB}
5329:
5330: In this chapter we redo the reduction of type IIB and type IIA on Calabi-Yau
5331: orientifolds by additionally allowing for non-trivial R-R and
5332: NS-NS background fluxes. As we will show, these fluxes
5333: result in non-trivial potentials for the supergravity fields and
5334: can lead to charged scalars or massive tensors.
5335:
5336:
5337: We first discuss the two type IIB setups. In section \ref{O3_wflux} we
5338: show that in orientifolds with $O3/O7$ planes fluxes introduce a superpotential only.
5339: More intriguingly, we point out in section \ref{O5_wflux} that
5340: $O5/O9$ setups with background flux in general admit a superpotential as well as
5341: a massive linear multiplet. Thus, additionally to the kinetic terms studied in
5342: section \ref{IIB_lin} we find $D-$terms and a direct mass
5343: term for a linear multiplet \cite{TGL1,mass_tensors}.
5344: In both IIB orientifold cases the induced potentials
5345: depend only on some but not all bulk moduli fields in the theory. In order
5346: to find potentials for the remaining moduli one has to take non-perturbative
5347: contributions into account. In \cite{Witten} it was argued that
5348: certain D-instantons induce corrections to the superpotential.
5349: To gain a better understanding of these corrections is subject
5350: of various recent work \cite{non-pert,DSFGK}.
5351: In section \ref{non-pert_sup} we do only a very moderate step and check if the
5352: resulting leading order superpotentials are holomorphic in the bulk coordinates.
5353: Assuming a generic form of such a superpotential one might achieve that all bulk fields
5354: are stabilized in the vacuum \cite{KKLT,non-pert,DSFGK}.
5355:
5356: In type IIA orientifolds the situation is slightly different. As we
5357: show in section \ref{O6sup} generic NS-NS and R-R background fluxes
5358: induce a superpotential which depends on all bulk moduli of the theory.
5359: Hence, appropriately chosen background fluxes could stabilize all geometric moduli
5360: in type IIA orientifolds. Additionally, one can attempt to include corrections due to
5361: non-perturbative effects. A brief discussion of superpotential contributions due
5362: to world-sheet or D-instantons can be found in section \ref{non-pert_sup}.
5363:
5364:
5365: \section{$O3/O7$ orientifolds: GVW superpotential}
5366: \label{O3_wflux}
5367:
5368: In this section we study $O3/O7$ orientifolds by also allowing
5369: background three-form fluxes $H_3$ and $ F_3$ on the
5370: Calabi-Yau manifold \cite{Michelson,TV,Mayr,GKP,TGL1}.
5371: The Bianchi identities together with the equations of motion imply
5372: that $H_3$ and $ F_3$ have to be harmonic three-forms.
5373: In orientifold compactifications they are further constrained
5374: by the orientifold
5375: projection. {}From \eqref{fieldtransfB}
5376: we see that for the projection given in \eqref{o3-projection}
5377: they both have to be odd under $\sigma^*$
5378: and hence are parameterized by elements of $H^{(3)}_-(Y)$.\footnote{This
5379: uses the fact that the exterior derivative on $Y$ commutes with $\sigma^*$.}
5380: It is convenient to combine the two three-forms into a complex
5381: $G_3$ according to
5382: \begin{equation}\label{fluxesB}
5383: G_3 = F_3 -\tau H_3\ , \qquad \tau= C_0 + i e^{- \phi}\ .
5384: \end{equation}
5385: $G_3$ can be explicitly expanded into a symplectic basis of $H^{(3)}_-$
5386: as
5387: \beq\label{G3exp}
5388: G_3 = m^{\hat k}\alpha_{\hat k} - e_{\hat k}\beta^{\hat k}\ , \qquad
5389: \hat k = 0,\ldots, h^{(1,2)}\ ,
5390: \eeq
5391: with $2(h^{(1,2)}_-+1)$ complex flux parameters
5392: \beq\label{mcomplex}
5393: m^{\hat k} = m^{\hat k}_F -\tau m^{\hat k}_H\ , \qquad
5394: e_{\hat k} = e_{\hat k}^F -\tau e_{\hat k}^H\ .
5395: \eeq
5396: However, in the following we do not need this explicit expansion and
5397: express our results in terms of $G_3$.
5398:
5399: The reduction of the IIB theory is performed by replacing
5400: \beq
5401: d\hat B_2 \ \rightarrow \ d\hat B_2 + H_3\ , \qquad d\hat C_2 \ \rightarrow \ d\hat C_2 + F_3\ ,
5402: \eeq
5403: in the field-strengths \eqref{fieldstr}. $H_3$ and $F_3$ are the
5404: background value of the field strengths
5405: $\hat F_3$ and $\hat H_3$ but do not effect
5406: $\hat F_5$ since the only possible terms would be of
5407: the form $H_3 \wedge C_2 $ or $B_2 \wedge F_3$
5408: but both $C_2$ and $B_2$ are projected
5409: out by the orientifold projection.\footnote{%
5410: We neglect subtleties appearing when $\hat B_2,\hat C_2$ do not arise with
5411: a derivative. These can be approached along the lines of \cite{DeWGKT}.}
5412: The only effect of non-trivial background fluxes is the appearance of
5413: a potential $V$. It is manifestly positive semi-definite and found to be
5414: \cite{TV,GKP,BBHL,DWG}
5415: \beq\label{potential37}
5416: V=
5417: %\frac{18i\ e^{4 \phi}}{\cK^{2}\int \Omega \wedge \bar \Omega }
5418: e^{K}
5419: \Big( \int \Omega \wedge \bar G_3 \int \bar \Omega \wedge G_3
5420: + G^{k { l}} \int \chi_{ k} \wedge G_3
5421: \int \bar \chi_{ l} \wedge \bar G_3 \Big)\ ,
5422: \eeq
5423: where $K$ is given in \eqref{kaehlerpot-O7-1}, $\chi_{k}$ is a basis of $H^{(2,1)}_-$ defined in \eqref{cso} and
5424: the background flux $G_3$ is defined in \eqref{fluxesB}.
5425: The details of the computation of $V$ can be found in \cite{GKP,TGL1}.
5426:
5427:
5428: Strictly speaking the additional term
5429: $\cL^{(4)}_{\text{top}} \sim
5430: %\frac{18}{\KK^2}
5431: \int_{Y}H_3 \wedge F_3 $
5432: arises in the Kaluza-Klein reduction.
5433: However, consistency of the compactifications requires its cancellation
5434: against Wess-Zumino like couplings of the orientifold planes
5435: to the R-R flux \cite{GKP}.
5436:
5437:
5438:
5439: Finally, one checks that the potential \eqref{potential37}
5440: can be derived from a superpotential $W$ via the expression
5441: given in \eqref{N=1pot} with vanishing $D$-term $D_\kappa=0$.
5442: For orientifolds with $c^a=b^a=0$
5443: $W$ was shown to be \cite{GVW,TV,GKP,BBHL,DWG}
5444: \begin{equation}
5445: W(\tau,z^k) = \int_{Y} \Omega \wedge G_3\ .
5446: \label{superpot}
5447: \end{equation}
5448: This continues to be the correct superpotential
5449: also if $c^a$ and $b^a$ are in the spectrum \cite{TGL1}, which is due to the
5450: fact that $K^{\rm Q}$ satisfies the no-scale condition \eqref{NScond}.
5451: This ends our analysis for $O3/O7$ setups. Surprisingly, for $O5/O9$ orientifolds
5452: the computation is more involved and forces us to once more apply and
5453: extend the linear multiplet techniques developed in chapter \ref{lin_geom_of_M}.
5454:
5455:
5456: \section{$O5/O9$ orientifolds: Gaugings and the massive linear multiplet}
5457: \label{O5_wflux}
5458:
5459: We now turn to the effective action of $O5/O9$ orientifolds with
5460: background fluxes. In order to detect the
5461: changes due to this non-trivial background, we proceed as in the $O3/O7$ case and
5462: first evaluate the field strengths (\ref{fieldstr}) including
5463: the possibility of background three-form fluxes
5464: $H_3$ and $F_3$.
5465: Since $\hat B_2$ and hence $H_3$ is odd
5466: it is again parameterized by $H^{(3)}_-$ while $\hat C_2$ and
5467: $F_3$ are even and therefore parameterized by $H^{(3)}_+$.
5468: As a consequence the
5469: explicit expansions of the background fluxes $H_3$ and $F_3$
5470: are given by
5471: \bea\label{mef}
5472: H_3&=& m^k_H \alpha_k - e_k^H \beta^k\ , \qquad k = 1, \ldots, h^{(2,1)}_-\ ,
5473: \nn\\
5474: F_3 &=& m_F^{\hat \kappa} \, \alpha_{\hat \kappa}
5475: - e^F_{\hat \kappa}\, \beta^{\hat \kappa} \ ,
5476: \qquad \hat\kappa = 0, \ldots, h^{(2,1)}_+ \ ,
5477: \eea
5478: where the $(m^k_H, e_k^H)$ are $2h^{(2,1)}_-$ constant flux
5479: parameters determining $H_3$ and
5480: $(m_F^{\hat \kappa}, e^F_{\hat \kappa})$ are $2h^{(2,1)}_++2$
5481: constant flux
5482: parameters corresponding to $F_3$.
5483: Inserting \eqref{exp-B}, \eqref{expO5} and \eqref{mef} into \eqref{fieldstr}
5484: we obtain
5485: \bea
5486: \hat H_3 &=& db^a \wedge \omega_a + m^k_H \alpha_{k} -
5487: e_{k}^H \beta^{k}\ ,\qquad
5488: \hat F_3 \ = \ dC_2+dc^\alpha \wedge \omega_\alpha + F_3\ ,\nonumber \\
5489: \hat F_5 &=& dD_2^a \wedge \omega_a + \tilde F^{k} \wedge \alpha_{k}
5490: - \tilde G_{k}\wedge \beta^{k}
5491: + d\rho_a \wedge \tilde \omega^a \\
5492: && - db^a \wedge C_2 \wedge \omega_a - c^\alpha db^a \omega_a \wedge
5493: \omega_\alpha\ ,\nn
5494: \eea
5495: where we defined
5496: \beq\label{Fcech}
5497: \tilde F^{k}= dV^{k} - m^{k}_H C_2\ , \qquad
5498: \tilde G_{k}=dU_{k} - e_{k}^H C_2\ .
5499: \eeq
5500: As in section \ref{IIB_orientifolds} the self-duality condition on $\hat F_5$
5501: is imposed by a Lagrange multiplier \cite{DallAgata} and we eliminate
5502: $D^a_{2}$ and $U_{k}$ by inserting their equations
5503: of motion into the action.
5504: After Weyl rescaling the four-dimensional metric with a factor $\KK/6$
5505: the ${N}=1$ effective action reads
5506: \bea\label{actiono5}
5507: S^{(4)}_{O5/O9} &=& \int -\tfrac{1}{2}R*\mathbf{1}-
5508: G_{\kappa\lambda} \; dz^{\kappa} \wedge *d\bar z^{\lambda}
5509: -G_{\alpha \beta} \; dv^\alpha \wedge *dv^\beta - G_{ab}\; db^a \wedge * db^b \nn \\
5510: && - \tfrac{e^{2D}}{6} \cK\, G_{\alpha \beta}\; dc^\alpha \wedge * dc^\beta
5511: - \tfrac{e^{-2D}}{24} \cK\, dC_2 \wedge * dC_2 -\tfrac{1}{4} dC_2 \wedge (\rho_a db^a - b^a d\rho_a) \nonumber \\
5512: &&
5513: - dD \wedge * dD
5514: - \tfrac{3e^{2D}}{8\KK} G^{ab}(d \rho_a - \KK_{ac\alpha} c^\alpha db^c)
5515: \wedge *(d \rho_b - \KK_{bd\beta} c^\beta db^d)- V*\mathbf{1}\nn \\[.1cm]
5516: &&
5517: + \tfrac{1}{4} \text{Re}\; \mathcal{M}_{k l}\; \tilde
5518: F^{k} \wedge \tilde F^{l} +
5519: \tfrac{1}{4} \text{Im}\; \mathcal{M}_{k l}\;
5520: \tilde
5521: F^{k} \wedge * \tilde F^{l} +
5522: \tfrac{1}{4} e_{k} (dV^{k} +
5523: \tilde F^{k})\wedge C_2\ ,\nn \\
5524: \eea
5525: where
5526: \begin{align}\label{pot5}
5527: V &=
5528: \frac{18i\ e^{4\phi}}{\cK^2 \int \Omega \wedge \bar \Omega }
5529: \left( \int \Omega \wedge F_3 \int \bar \Omega \wedge F_3
5530: + G^{\kappa \lambda} \int \chi_{\kappa} \wedge F_3
5531: \int \bar \chi_{\lambda} \wedge F_3 \right) \\
5532: &-
5533: \tfrac{9\, e^{2\phi}}{\KK^2} \Big[
5534: m^{k}_H\, (\text{Im} \MM)_{ k l}\, m^{l}_H +
5535: \big(e_{k}^H-(m_H \text{Re} \MM)_{k}\big) \big(\text{Im} \MM\big)^{-1 k l}
5536: \big(e_{l}^H-(m_H \text{Re} \MM)_{l}\big) \Big]\ . \nonumber
5537: \end{align}
5538: The derivation of this potential can be found in ref. \cite{TGL1}.\footnote{Note that in this class of
5539: orientifolds the topological term
5540: $\int_YH_3 \wedge F_3$ vanishes since there is no intersection between
5541: $H^{(3)}_+$ and $H^{(3)}_-$. Thus strictly speaking background
5542: D-branes have to be included in order to satisfy the tadpole cancellation
5543: condition.}
5544:
5545: The action \eqref{actiono5} has the standard one-form gauge invariance
5546: $V^k\to V^k+d\Lambda^k_0$
5547: but due to the modification in \eqref{Fcech}
5548: also a modified (St\"uckelberg) two-form gauge invariance given by
5549: \beq\label{2gauge}
5550: C_2\to C_2 +d\Lambda_1\ ,\qquad V^k\to V^k + m^k_H\Lambda_1\ .
5551: \eeq
5552: Thus for $m^k_H\neq 0$ one vector can be set to zero by an appropriate
5553: gauge transformation.
5554: This is directly related to the fact that \eqref{actiono5}
5555: includes mass terms proportional to $m^k_H$ for $C_2$ arising from
5556: \eqref{Fcech}. In this case gauge invariance requires
5557: the presence of Goldstone degrees of freedom which
5558: can be `eaten' by $C_2$.\footnote{Exactly the same situation
5559: occurs in Calabi-Yau compactifications of type IIB
5560: with background fluxes where both $B_2$ and $C_2$
5561: can become massive \cite{LM}.}
5562: Finally note that the last term in \eqref{actiono5} also
5563: includes a standard $D=4$ Green-Schwarz term $F^k\wedge C_2$.
5564:
5565:
5566: \subsection{Vanishing magnetic fluxes $m^k_H=0$}
5567:
5568: The next step is to show that $S^{(4)}_{O5/O9}$ is consistent
5569: with the constraints of $N=1$ supergravity. However, due to the
5570: possibility of $C_2$ mass terms this is not completely
5571: straightforward. A massive $C_2$
5572: is no longer dual to a scalar but rather to a vector.
5573: We find it more convenient to keep the
5574: massive tensor in the spectrum and discuss
5575: the $N=1$ constraints in terms of a massive linear multiplet.
5576: Before doing so, let us first discuss
5577: the situation $m^k_H= 0$ where $\tilde F^{l}= F^{l}$ holds.
5578: In this case the $C_2$ remains massless
5579: and can be dualized to a scalar field $h$ which together with
5580: the dilaton $ \phi$ combines to form a chiral multiplet
5581: $(\phi, h)$.
5582: Using the standard dualization procedure (see section \ref{revIIB})
5583: one obtains the effective action \eqref{S_scalarO5} plus the
5584: potential $V$ given in \eqref{pot5} evaluated at $m^k_H=0$.
5585: Furthermore, due to electric NS-NS fluxes the scalar $h$ is gauged
5586: and we have to replace in \eqref{S_scalarO5}
5587: \beq \label{hcov}
5588: dh\quad \rightarrow \quad Dh=d h - e_k^H V^k\ .
5589: \eeq
5590: Hence, $h$ couples non-trivially to the gauge fields as a direct consequence
5591: of the Green-Schwarz coupling $F^k\wedge C_2$
5592: in \eqref{actiono5}. In the dualized
5593: action the scalar $h$ then is charged
5594: under the $U(1)$ gauge transformation $h\to h + e_k^H \Lambda_0^k$ with $V^k \to V^k + d\Lambda_0^k$.
5595: Note that the gauge charges are set by the electric fluxes.
5596:
5597: The K\"ahler potential \eqref{O5-Kaehlerpot} with chiral coordinates \eqref{Kcoord}
5598: and the gauge-couplings \eqref{fholo} remain unchanged for the theory with
5599: $m^k_H=0$. However, due to the non-trivial electric NS-NS fluxes the
5600: covariant derivative of $h$ given in \eqref{hcov} translates into the
5601: covariant derivative $DS = dS - i e_{k} V^{k}$. It remains to cast
5602: the potential $V$ given in \eqref{pot5}, evaluated at $m^k_H=0$, into
5603: the standard $N=1$ supergravity form \eqref{N=1pot}.
5604: {}From eq.\ \eqref{hcov}
5605: we see that the axion is charged
5606: and as a consequence we
5607: expect a non-vanishing $D$-term in the potential. Recall
5608: the general formula
5609: for the $D$-term \cite{WB}
5610: \beq
5611: K_{I\bar J} \bar X^{\bar J}_k = i \partial_I D_k \ ,
5612: \eeq
5613: where $X^{I}$ is the Killing vector of the $U(1)$ gauge transformations
5614: defined as $\delta M^I = \Lambda^k_0 X_k^J \partial_J M^I$.
5615: Inserting \eqref{O5-Kaehlerpot}
5616: and \eqref{Kcoord} we obtain
5617: \begin{equation}
5618: D_k = - e_{k}^H\, \frac{\partial K}{\partial \bar S} =
5619: 3\, e_{k}^H\, e^{\phi}\cK^{-1} \ .
5620: \end{equation}
5621: Using also \eqref{gauge-couplingsO3} we arrive at the $D$-term contribution
5622: to the potential
5623: \begin{eqnarray}
5624: \tfrac{1}{2} (\text{Re}\; f)^{-1\ kl} D_k D_l = -
5625: \tfrac{9}{\cK^2} e^{2\phi} \; e_{k}^H\, (\text{Im}\; \MM)^{-1\ kl}\,
5626: e_{l}^H\ , \label{D-term}
5627: \end{eqnarray}
5628: which indeed reproduces the last term in \eqref{pot5} for
5629: $m^k_H=0$.
5630:
5631: The first term in \eqref{pot5} arises from the superpotential
5632: \begin{equation}\label{W5}
5633: W= \int_{Y} \Omega \wedge F_3\ ,
5634: \end{equation}
5635: which follows from a calculation analog to the $O3/O7$ case \cite{TGL1}.
5636: It is interesting that for this class of orientifolds
5637: the RR-flux $F_3$ results in a contribution to the superpotential while
5638: the NS-flux $H_3$ contributes instead to a $D$-term.
5639:
5640: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5641: %
5642: \subsection{Non-vanishing magnetic fluxes $m^k_H\neq 0$}
5643:
5644:
5645: Let us now turn to the case where both electric and magnetic fluxes are
5646: non-zero and the two-form $C_2$ is massive.
5647: In this case $C_2$ is dual to a massive vector or equivalently the massive
5648: linear multiplet is dual to massive vector multiplet.
5649: Here we do not discuss this duality but instead show how the couplings
5650: of a massive linear multiplet is consistent with the action
5651: \eqref{actiono5} \cite{mass_tensors}.
5652:
5653: In section \ref{IIB_lin} we already examined the kinetic terms and couplings
5654: for the $O5/O9$ theory in the presence of one tensor multiplet $L=(\phi,C_2)$.
5655: We found that they are determined in terms of the generalized K\"ahler potential
5656: and the function $F$ both given in \eqref{KpotL}.
5657: Let us now briefly discuss the situation of a massive
5658: linear multiplet coupled to $N=1$ vector- and chiral multiplets.
5659: For simplicity we discuss the situation in flat space
5660: and do not couple the massive linear multiplet to supergravity.
5661: However, we expect our results to generalize to the
5662: supergravity case. More details can be found in \cite{GGRS,mass_tensors}.
5663:
5664: As we already said, a linear multiplet $L$ contains a real scalar (also denote by $L$)
5665: and the field strength of a two-form $C_2$ as bosonic components. However,
5666: it does not contain the two-form itself
5667: which instead is a member of the chiral `prepotential' $\Phi$ defined
5668: as\footnote{We suppress the spinorial indices and use the convention
5669: $D\Phi \equiv D^\alpha\Phi_\alpha$,
5670: $\bar D\bar\Phi \equiv \bar D^{\dot{\alpha}}\bar\Phi_{\dot{\alpha}}$.}
5671: \beq
5672: L= D\Phi +\bar D \bar \Phi\ , \qquad \bar D\Phi = 0\ .
5673: \eeq
5674: This definition solves the constraint \eqref{linearc} (in flat space).
5675: The kinetic term for $L$ (or rather for $\Phi$) is given in
5676: \eqref{actionL} and a mass-term can be added via the
5677: chiral integral
5678: \beq\label{Lmasst}
5679: \cL_{m} = \tfrac14\int d^2\theta \Big[
5680: f_{kl}(N) (W^k - 2i m^k_H\Phi)(W^l - 2i m^l_H\Phi)
5681: + 2 e_k^H (W^k - i m^k_H\Phi)\Phi\Big] + {\rm h.c.}\ ,
5682: \eeq
5683: where $W^k= -\tfrac14 \bar D^2 DV^k$ are the chiral field strengths supermultiplets
5684: of the vector multiplets $V^k$ and $f_{kl}(N)$ are the gauge kinetic function
5685: which can depend holomorphically on the chiral multiplets $N$.
5686: $(m^k_H,e_k^H)$ are constant parameters which will turn out
5687: to correspond to the flux parameters defined in \eqref{mef}.
5688: The Lagrangian \eqref{Lmasst} is invariant under the standard
5689: one-form gauge invariance $V^k\to V^k +\Lambda_0^k + \bar \Lambda_0^k$
5690: ($\Lambda_0^k$ are chiral superfields)
5691: which leaves both $W^k$ and $\Phi$ invariant.
5692: In addition \eqref{Lmasst} has a two-form gauge invariance
5693: corresponding to \eqref{2gauge} given by
5694: \beq\label{linearg}
5695: \Phi \to \Phi +\tfrac{i}8 \bar D^2 D \Lambda_1\ ,\qquad
5696: V^k\to V^k + m^k_H \Lambda_1\ ,
5697: \eeq
5698: where $\Lambda_1 $ now is a real superfield.
5699: {}From \eqref{linearg} we see that one entire vector multiplet
5700: can be gauged away and thus plays the role of the Goldstone degrees
5701: of freedom which are `eaten' by the massive linear multiplet.
5702:
5703: In components one finds the bosonic action
5704: \beq
5705: \cL_{m} = -\tfrac{1}{2} \text{Re} f_{k l}\; \tilde
5706: F^{k} \wedge * \tilde F^{l} -
5707: \tfrac{1}{2} \text{Im} f_{k l}\;
5708: \tilde
5709: F^{k} \wedge \tilde F^{l} +
5710: \tfrac{1}{4} e_{k} (dV^{k} +
5711: \tilde F^{k})\wedge C_2 - V*\mathbf{1} \ ,
5712: \eeq
5713: where $\tilde F^{l}$ is defined exactly as in \eqref{Fcech}
5714: and the potential $V$ receives
5715: two distinct contributions
5716: \beq
5717: V=
5718: \tfrac{1}{2}\,
5719: (\text{Re} f)^{-1 kl} D_{k} D_{l} + 2\, m^k_H\text{Re} f_{kl}\, m^l_H\, L^2\ ,
5720: \qquad
5721: D_{k} = \big(e_k^H + 2\,\text{Im}f_{kl}\, m^l_H \big)\, L \ .
5722: \eeq
5723: The first term arises from eliminating the $D$-terms in
5724: the $U(1)$ field strength $W^k$ while the second term is a
5725: `direct' mass term for the scalar $L$.\footnote{Note that this second term
5726: is a contribution to the potential which is neither a $D$- nor an
5727: $F$-term but instead a `direct' mass term whose presence is enforced
5728: by the massive two-form.}
5729: Inserting the $D$-term yields a second contribution to the mass term
5730: and one obtains altogether
5731: \bea
5732: V&=&
5733: \tfrac12 \big[\big(e_k^H +2\text{Im} f_{kp}\, m^p \big)
5734: (\text{Re} f)^{-1 kl} \big( e_l^H +2 \text{Im} f_{lr}\, m^r\big)
5735: +4 m^k_H\, \text{Re} f_{kl}\, m^l_H\big] L^2\ . \qquad
5736: \eea
5737: Using \eqref{Lzeta_O59} and \eqref{gauge-couplingsO3}
5738: this precisely agrees with the second term in the potential \eqref{pot5}.
5739:
5740: As before the first term in \eqref{pot5} can be derived from the superpotential
5741: \eqref{W5}. This ends our discussion of
5742: type IIB orientifolds in a general NS-NS and R-R flux background.
5743: As we have seen, switching on fluxes yields a potential for only part of
5744: the moduli fields. This changes in IIA orientifolds to which we will
5745: turn now.
5746:
5747:
5748: \section{$O6$ orientifolds: Flux superpotentials}
5749: \label{O6sup}
5750:
5751: In this section we derive the effective action of type IIA orientifolds
5752: in the presence of background fluxes.
5753: For standard $N=2$ Calabi-Yau compactifications of type IIA a
5754: similar analysis is carried out in refs.\ \cite{LM,KachruK}.
5755: In order to do so
5756: we need to start from the ten-dimensional action of massive
5757: type IIA supergravity which differs from the action \eqref{10dact} in
5758: that the two-form $\hat B_2$ is massive. In the
5759: Einstein frame it is given by \cite{Romans}
5760: \bea \label{10dactm}
5761: S^{(10)}_{MIIA} &=& \int -\tfrac{1}{2}\hat R*\mathbf{1} -\tfrac{1}{4} d\hat \phi\wedge * d\hat \phi
5762: -\tfrac{1}{4} e^{-\hat \phi}\hat H_3 \wedge *\hat H_3
5763: -\tfrac{1}{2} e^{\frac{3}{2} \hat \phi}\hat F_2 \wedge *\hat F_2 \nn \\
5764: && -\tfrac{1}{2} e^{\frac{1}{2} \hat \phi}\hat F_4 \wedge *\hat F_4
5765: -\tfrac{1}{2} e^{\frac{5}{2} \hat \phi}\, (m^0)^2 * \mathbf{1} + \cL_{\rm top}\ ,
5766: \eea
5767: where
5768: \bea
5769: \cL_{\rm top}&=& -\tfrac{1}{2}\Big[ \hat B_2 \wedge d\hat C_3 \wedge d\hat C_3\
5770: -(\hat B_2)^2 \wedge d\hat C_3 \wedge d\hat C_1
5771: + \tfrac{1}{3}(\hat B_2)^3 \wedge (d\hat C_1)^2 \nn \\
5772: & & - \tfrac{m^0}{3}(\hat B_2)^3 \wedge d\hat C_3
5773: + \tfrac{m^0}{4}(\hat B_2)^4 \wedge d\hat C_1
5774: + \tfrac{(m^0)^2}{20}(\hat B_2)^5 \Big]\ ,
5775: \eea
5776: and the field strengths are defined as
5777: \bea \label{defHFF}
5778: \hat H_3 = d \hat B_2\ , \quad \hat F_2 = d\hat C_1+m^0 \hat B_2\ , \quad
5779: \hat F_4 = d\hat C_3 - \hat C_1 \wedge \hat H_3-\tfrac{m^0}{2}(\hat B_2)^2\ .
5780: \eea
5781: Compared to the analysis of the previous section we now include
5782: non-trivial background fluxes of the field strengths
5783: $F_2$, $H_3$ and $F_4$ on the Calabi-Yau orientifold.
5784: We keep the Bianchi identity and the equation of motion intact
5785: and therefore expand $F_2$, $H_3$ and $F_4$
5786: in terms of harmonic forms compatible with the orientifold
5787: projection. From \eqref{fieldtransf} we infer that $F_2$ is expanded in
5788: harmonic forms of $H^{2}_-(Y)$,
5789: $H_3$ in harmonic forms of $H^3_{-}(Y)$ and $F_4$ in harmonic forms
5790: of $H^{4}_+(Y)$.\footnote{As we observed in the previous section
5791: there is no $\hat C_1$
5792: due to the absence of one-forms on the orientifold.
5793: Nevertheless its field strength $F_2$
5794: can be non-trivial on the orientifold since $Y$ generically possesses
5795: non-vanishing harmonic two-forms.}
5796: Explicitly the expansions read
5797: \bea \label{fluxesA}
5798: H_3\, =\, q^\lambda \alpha_\lambda - p_k\, \beta^k\ , \quad F_2\, =\, -m^a \omega_a\ , \quad
5799: F_4\, =\, e_a\, \tilde \omega^a\ ,
5800: \eea
5801: where $(q^\lambda,p_k)$ are $h^{(2,1)}+1$ real NS flux parameters
5802: while $(e_a,m^a)$ are $2h^{1,1}_-$ real RR flux parameters.
5803: %By $m^0$ we denote the mass parameter $m$ of the massive type IIA theory which
5804: %we combine with the $h^{1,1}_-$ magnetic RR fluxes into a
5805: %vector $m^{\hat a}=(m^0,m^a)$.
5806: The harmonic forms $(\alpha_\lambda, \beta^k)$ are the elements of the real
5807: symplectic basis of $H^3_-$ introduced in \eqref{sp_alpha-beta}.
5808: The basis $\tilde \omega^a$ of
5809: $H^{(2,2)}_+$ is defined to be the dual basis of $\omega_a$ while the
5810: basis $\tilde \omega^\alpha$ denotes a basis of $H^{(2,2)}_-$ dual to $\omega_\alpha$.
5811:
5812: Inserting \eqref{expJB}, \eqref{form-exp} and \eqref{fluxesA} into
5813: \eqref{defHFF} we arrive at
5814: \bea \label{fieldst}
5815: \hat H_3 &=& db^a\wedge \omega_a + q^\lambda \alpha_\lambda - p_k\, \beta^k\ , \qquad \qquad
5816: \hat F_2 = (m^0 b^a + m^a)\, \omega_a\ ,\\
5817: \hat F_4 &=& dC_3 + dA^\alpha \wedge \omega_\alpha
5818: + d\xi^k \wedge \alpha_k -
5819: d\tilde \xi_\lambda \wedge \beta^\lambda +
5820: \big(b^a m^b -\tfrac12 m^0 b^a b^b\big)\,
5821: \cK_{abc}\tilde \omega^c + e_a\, \tilde \omega^a\ , \nn
5822: \eea
5823: where we have used $\omega_a \wedge \omega_b = \cK_{abc}\, \tilde \omega^c$.
5824: Now we repeat the KK-reduction of the previous section using the
5825: modified field strength
5826: \eqref{fieldst} and the action \eqref{10dactm} instead of \eqref{10dact}.
5827: This results in%
5828: \footnote{The action $S^{(4)}_{O6}$ is given in \eqref{act1}. However, due to the fact that
5829: we perform the Kaluza-Klein reduction in the generic basis introduced in \eqref{sp_alpha-beta} the kinetic
5830: terms for $\tilde \cM^{\rm Q}$ are replaced by \eqref{act2}. }
5831: \beq\label{Sflux}
5832: S^{(4)} = S^{(4)}_{O6} - \int \tfrac{g}{2}\, d\cc_3 \wedge * d\cc_3 + {h}\, d\cc_3 +
5833: U * \mathbf{1}\ ,
5834: \eeq
5835: where $S^{(4)}_{O6}$ is given in \eqref{act1}.
5836: $\cc_3$
5837: is the four-dimensional part of the ten-dimensional
5838: three-form $\hat C_3$ defined in \eqref{form-exp} and
5839: its couplings to the scalar fields are given by
5840: \beq
5841: g = e^{-4 \phi} \left(\tfrac{\cK}6
5842: \right)^3\ , \qquad h = e_a b^a + \tilde \xi_\lambda q^\lambda
5843: - \xi^k p_k + \tfrac{1}{2}\R \cN_{0 \ah}\, m^\ah \ ,
5844: \eeq
5845: where we denoted $m^\ah=(m^0,m^a)$. The potential term $U$ of \eqref{Sflux}
5846: is given by
5847: \beq \label{U-pot}
5848: U = \tfrac{9}{\cK^2} e^{2\phi} \int_Y H_3 \wedge * H_3
5849: - \tfrac{18}{\cK^2} e^{4\phi} \I \cN_{\ah \bh}\, m^\ah m^\bh
5850: +\tfrac{ 27 } {\cK^3} e^{4\phi} G^{ab}(e_a - \R \cN_{a\ah}\, m^\ah)(e_b - \R \cN_{b\bh}\, m^\bh)\ .
5851: \eeq
5852: The matrix $\cN_{\ah \bh} (t,\bar t)$ is defined to be the corresponding
5853: part of the $N=2$ gauge-coupling matrix \eqref{def-cN}
5854: restricted to $\tilde \cM^{\rm SK}$ by applying \eqref{van-int} and \eqref{splitmetr}.
5855:
5856: In four space-time dimensions
5857: $\cc_3$ is dual to a constant which plays the role of
5858: an additional electric flux $e_0$ in complete analogy with the
5859: situation in $N=2$ discussed in \cite{LM}.
5860: Eliminating $\cc_3$ in favor of $e_0$ by following \cite{LM} or \cite{BW}
5861: the potential takes the form \cite{TGL2}
5862: \beq \label{V-pot1}
5863: V = \tfrac{9}{\cK^2} e^{2\phi} \int H_3 \wedge * H_3
5864: - \tfrac{18}{\cK^2} e^{4\phi} (\tilde e_\ah - \cN_{\ah \ch}\, m^\ch) (\I \cN)^{-1\, \ah \bh}
5865: (\tilde e_\bh -\bar \cN_{\bh
5866: \ch}\, m^\ch)\ ,
5867: \eeq
5868: where we introduced the shorthand notation
5869: $\tilde e_\ah=(e_0 + \xi_\lambda q^\lambda-\xi^\kh p_\kh,e_a)$ and $m^\ah=(m^0,m^a)$.
5870: Note that in the presence of NS flux
5871: one can absorb $e_0$ by shifting the fields
5872: $\xi,\tilde \xi$. This corresponds to adding an integral form to
5873: $\CC_3$ as carefully discussed in \cite{BW}.
5874: However, for the discussion of mirror symmetry it is more convenient to
5875: keep the parameter $e_0$ explicitly in the action.
5876:
5877:
5878: In order to establish the consistency with $N=1$ supergravity
5879: one needs to rewrite $V$ given in \eqref{V-pot1} in terms of
5880: \eqref{N=1pot} or in other words we need express $V$ in terms
5881: of a superpotential $W$ and appropriate $D$-terms.
5882: From \eqref{Sflux} we infer that turning on fluxes does not
5883: charge any of the fields and therefore all $D$-terms have to vanish.
5884: In \cite{TGL2} it was checked that the potential \eqref{V-pot1} can be entirely expressed in
5885: terms of the superpotential
5886: \beq \label{superpot1}
5887: W\ =\ W^{\rm Q}(N,T) + W^{\rm K}(t)\ ,
5888: \eeq
5889: where
5890: \bea \label{superpot2}
5891: W^{\rm Q}(N^k,T_\lambda)& =& \int_Y \Omegac \wedge H_3\ =\
5892: - 2N^k p_k - i T_\lambda q^\lambda\ , \\
5893: W^{\rm K}(t^a) &=& e_0 + \int_Y \Jc \wedge F_4 - \tfrac{1}{2} \int_Y \Jc \wedge \Jc \wedge F_2
5894: - \tfrac{1}{6} m^0 \int_Y \Jc \wedge \Jc \wedge \Jc\ ,
5895: \nn\\
5896: &=& e_0 + e_a t^a + \tfrac{1}{2}\cK _{abc} m^a t^bt^c - \tfrac{1}{6} m^0 \cK _{abc} t^a t^bt^c\nn\ ,
5897: \eea
5898: and $\Omegac$ and $\Jc$ are defined in \eqref{N=1coords}. Using the
5899: definitions \eqref{symp-form} and \eqref{symp-formodd} of the skew-symmetric products $\big<\cdot,\cdot\big>$
5900: for even and odd forms $W$ is rewritten as
5901: \beq
5902: W = \big<e^{\Jc}, F \big> + \big<\Omegac,H_3 \big>\ , \qquad F = m^0 - F_2 - F_4 + F_6\ ,
5903: \eeq
5904: where we have defined $F_6$ via $e_0= \int_Y F_6$.
5905: We see that the superpotential is the sum of two terms.
5906: $W^{\rm Q}$ depends on the NS fluxes $(p_k,q^\lambda)$ of $H_3$ and the
5907: chiral fields $N^k,T_\lambda$ parameterizing the space $\tilde \cM^{\rm Q}$.
5908: $W^{\rm K}$ depends on the RR fluxes $(e_{\hat a}, m^{\hat b})$
5909: of $F_2$ and $F_4$ (together with $m^0$ and $e_0$) and
5910: the complexified K\"ahler deformations $t^a$ parameterizing
5911: $\cM^{\rm SK}$.
5912: We see that contrary to the type IIB case both types of moduli,
5913: K\"ahler and complex structure deformations appear in the superpotential
5914: suggesting the possibility that all moduli can be fixed in this set-up.
5915: This was resently shown to be the case in refs.~\cite{VZ,DeWGKT}.
5916:
5917:
5918: Let us end this section by comparing the R-R superpotentials of type IIA
5919: and type IIB orientifolds. Recall that for
5920: both IIB orientifold setups R-R fluxes induce superpotentials \eqref{superpot} and \eqref{W5}
5921: holomorphic in the complex structure deformations $z$. Hence, we compare
5922: \beq
5923: W_A(t) = \big<e^{\Jc}, F \big> \ , \qquad W_B(z) = \big<\Omega, F_3\big>\ ,
5924: \eeq
5925: where the skew-products are defined in \eqref{symp-form} and \eqref{symp-formodd}.
5926: As just discussed $F$ depends on $2h^{(1,1)}_- + 2$ RR fluxes $(e_\ah,m^\ah)$.
5927: To count the flux parameters labeling $F_3$ recall that it transforms differently in the
5928: two IIB orientifolds. $F_3$ sits in $H^{3}_-(\tilde Y)$ and is determined in terms of
5929: $2h^{(2,1)}_-+2$ real flux parameters for the $O3/O7$ case and sits in $H^{3}_+(\tilde Y)$
5930: depending on $2h^{(2,1)}_+ + 2$ real flux parameters
5931: for the $O5/O9$ case. Therefore, the number of flux parameters matches when choosing mirror
5932: involutions satisfying \eqref{matchchohm}. Exchanging \cite{FMM}
5933: \beq
5934: e^{\Jc}(t)\ \leftrightarrow\ \Omega(z) \ , \qquad F\ \leftrightarrow\ F_3\ ,
5935: \eeq
5936: as in equation \eqref{pure-spinor-map} the two superpotentials $W_A(t)$ and $W_B(z)$ get identified.
5937: In $N=2$ the mirror identification of the complex structure moduli space $\cM^{\rm cs}$ with the complexified K\"ahler
5938: moduli space $\cM^{\rm ks}$ can be used to calculate world-sheet instanton corrections to $\cM^{\rm ks}$.
5939: It would be interesting to generalize this to $N=1$ orientifold theories which allow additionally for
5940: non-oriented world-sheets as discussed at the end of section \ref{IIA_orientifolds}.
5941: In addition to world-sheet instantons also certain D-instantons induce correction terms
5942: to the superpotential. We will end this chapter by a few comments on their
5943: generic structure.
5944:
5945: \section{D-instanton corrections to the superpotentials}
5946: \label{non-pert_sup}
5947:
5948: Let us close this chapter by briefly discussing possible D-instanton
5949: corrections to the superpotentials \eqref{superpot}, \eqref{W5} and \eqref{superpot1}.
5950: They can arise from wrapping $D(p-1)$-branes around $p$-cycles
5951: $\Sigma_p$ \cite{BBS}. In addition to corrections of the K\"ahler potential
5952: D-instantons induce extra superpotential terms \cite{Witten}. These
5953: depend on brane moduli as well as bulk fields and found recent phenomenological
5954: application in moduli stabilization \cite{KKLT, non-pert, DSFGK}. It would be interesting to fully
5955: incorporate these effects and to understand the additional contributions due to
5956: non-orientable world-volumes. First steps into this direction are done in the
5957: recent works \cite{non-pert, DSFGK}.
5958: In this section we will take only a very moderate step and
5959: apply the calibration conditions to show that the D-brane action
5960: becomes linear in the bulk fields. This ensures holomorphicity of the induces
5961: superpotential terms when expressed in the proper K\"ahler variables of the
5962: respective orientifold setup.
5963:
5964: To make this more precise, recall that any correlation function
5965: is weighted by the string-frame
5966: world-volume action of the wrapped Euclidean $D(p-1)$-branes
5967: and thus includes a factor $e^{-S_{D(p-1)}}$ where
5968: \beq \label{instact}
5969: S_{D(p-1)} = i\mu_{p}\,
5970: \int_{\cW_{p}}\Big(d^{p} \lambda\ e^{- \hat \phi} \sqrt{\det\big({\varphi^*(\hat g+ \hat B_2) + \ell F}\big)}
5971: - i \Em^*\Big(\sum_q \hat C_{q} \wedge e^{-\hat B_2}\Big) \wedge e^{\ell \FD}\Big)\ .
5972: \eeq
5973: where $\ell=2\pi \alpha'$. This is the Euclidean analog of the
5974: Dirac-Born-Infeld action \eqref{DBI} plus the Chern-Simons action \eqref{CSaction}.
5975: $\cW_p$ is the world-volume of the $D(p-1)$-brane and $\varphi^*$ is the pull-back
5976: of the map $\varphi$ which embeds $\cW_p$ into Calabi-Yau orientifold $Y$,
5977: $\varphi:\cW_p \hookrightarrow Y$.
5978: We have chosen the R-R charge $\mu_p$ equal to the tension since
5979: the wrapped $D(p-1)$-branes must be BPS in order to preserve $N=1$ supersymmetry.
5980: In fact, as we already discussed in section \eqref{D-branes} there are
5981: additional condition arising from the requirement that the $Dp$-branes preserves
5982: the same supersymmetry that is left intact
5983: by the orientifold projections. This in turn implies
5984: that $O3/O7$ orientifolds can admit
5985: corrections from $D3$ instantons, $O5/O9$ setups from $D1$ and $D5$ instantons and
5986: $O6$ setups from $D2$ instantons. Moreover, these have to be calibrated
5987: with respect to the same forms as the internal parts of the orientifold planes.
5988:
5989: The calibration conditions for Euclidean $D(p-1)$-branes
5990: in a Calabi-Yau manifold have been derived in refs.\ \cite{BBS,MMMS}.
5991: Let us first apply their results to type IIA orientifolds with $O6$ planes.
5992: Recall that the unbroken supercharge has to be some linear combination
5993: $\epsilon=a^+ \epsilon_+ + a^- \epsilon_-$ of the two covariantly
5994: constant spinors $\epsilon_+$ and $\epsilon_-$ of the
5995: original $N=2$ supersymmetry. Let us denote the relative phase
5996: of $a^+$ and $a^-$ by $a^-/a^+=-ie^{i\theta_{D2}}$ while the
5997: absolute magnitude can be fixed by the normalization of $\Omega$.
5998: {} As forms $J$ and $\Omega$ have to obey the condition
5999: \beq
6000: J \wedge J \wedge J = \tfrac{3i}{2}e^{-2U} \Omega \wedge \bar \Omega\
6001: \eeq
6002: at every point in the moduli space. Note however, that $J$ depends on
6003: K\"ahler structure deformations $v^a$ while $\Omega$ is a function of
6004: the complex structure deformations $q^K$. Hence, $e^U$ is a non-trivial function
6005: of $v^a$ and $q^K$ and from $\int J^3 =\frac{3i}{2}e^{-2U}\int \Omega\wedge\bar\Omega$
6006: one infers
6007: \beq\label{Omeganorm}
6008: % \gamma_{ijk}\, \epsilon_+ =e^{-U} \Omega_{ijk} \epsilon_-\ , \qquad
6009: e^{U}=\sqrt{2}\, e^{\frac{1}{2}(K^{\rm K}-\Kcs)}\ ,
6010: \eeq
6011: where K\"ahler potential $K^{\rm K}(t)$
6012: is given in \eqref{Kks} while $\Kcs(q)$ is the restriction of the K\"ahler
6013: potential \eqref{csmetric} to the real slice $\cM^{\rm cs}_\bbR$.
6014: The existence of $\epsilon$ imposes constraints
6015: on the map $\varphi$. These BPS conditions read \cite{BBS,MMMS}
6016: \beq\label{sLagr-cond}
6017: \varphi^*(\Omega)\ =\ e^{U+i\theta_{D2}} \sqrt{\det\big({\varphi^*(\hat g+ \hat B_2) + \ell F_2}\big)}
6018: d^3 \lambda\ , \qquad
6019: \varphi^*\Jc + i 2\pi \alpha' F_2\ =\ 0\ ,
6020: \eeq
6021: where $\Jc$ is given in \eqref{def-t}.
6022: The second condition in \eqref{sLagr-cond} enforces
6023: $\varphi^*(J)=0$ as well as $\varphi^*\hat B_2 + \ell F_2 =0$, such that the first equation
6024: simplifies to
6025: \beq \label{cal1}
6026: \varphi^*\R( e^{-i\theta_{D2}}\Omega)\ =\ e^U \sqrt{\det\big(\varphi^*\hat g\big)} d^3 \lambda\ , \qquad
6027: \varphi^*\I( e^{-i\theta_{D2}}\Omega)\ =\ 0\ ,
6028: \eeq
6029: where we have used that the volume element on $\cW_3$ is real. For vanishing $F$ these conditions
6030: coincide with those displayed in equation \eqref{calcond}. Even in the general case
6031: \eqref{sLagr-cond} and \eqref{cal1} imply that the Euclidean $D2$ branes have to
6032: wrap special Lagrangian cycles in $Y$, which are calibrated with respect to
6033: $\R(e^{-U-i\theta_{D2}}\Omega)$.
6034: On the other hand, recall
6035: that the orientifold planes are located
6036: at the fixed points of the anti-holomorphic
6037: involution $\sigma$ in $Y$ which are
6038: special Lagrangian cycles calibrated
6039: with respect to $\R(e^{-U-i\theta}\Omega)$
6040: as was argued in eqs.\
6041: \eqref{OLagr} and \eqref{calibr-O6}.\footnote{$e^{-U}$ is the normalization factor which was left undetermined in \eqref{calibr-O6}.}
6042: Thus, in order for the D-instantons to
6043: preserve the same linear combination of the supercharges as the orientifold, we have to
6044: demand $\theta_{D2} =\theta$.
6045: Using this constraint and inserting the calibration conditions
6046: \eqref{cal1} back into \eqref{instact} one finds
6047: \beq \label{instact2}
6048: S_{D2} = i\mu_3 \,
6049: \int_{\cW_3} \big( \varphi^*\big[2\R( C\Omega) \big] - i \varphi^*(\hat C_3) \big)\ = \
6050: \, \int_{\cW_3} \varphi^*\Omegac \ ,
6051: \eeq
6052: where $C=\frac{1}{2} e^{-\phi-i\theta} e^{-U}$
6053: was defined in eqs.\ \eqref{def-C}, \eqref{4d-dilaton} and
6054: $\Omegac$ is given in \eqref{N=1coords}. The coefficients of $\Omegac$
6055: expanded in a basis of $H^{3}_+(Y)$
6056: are exactly the $N=1$ K\"ahler coordinates $(N^k,T_\lambda)$ introduced in \eqref{Oexp}. As a consequence the instanton action
6057: \eqref{instact2} is linear and thus holomorphic in these coordinates
6058: which shows that $D2$-instantons
6059: can correct the superpotential.
6060: Explicitly such corrections can be obtained by evaluating
6061: appropriate fermionic 2-point functions which are weighted
6062: by $e^{-S_{D2}}$ \cite{HM}. Applying \eqref{instact2}
6063: and keeping only the lowest term in the fluctuations
6064: of the instanton one obtains corrections of the form
6065: \beq
6066: W_{D3} \propto e^{-\int_{\Sigma_3} \Omegac}\ ,
6067: \eeq
6068: where $\Sigma_3$ is the three-cycle wrapped by the $D2$ instanton.
6069:
6070: This result can be lifted to M-theory by embedding Calabi-Yau orientifolds into
6071: compactifications on special $G_2$ manifolds.
6072: In this case the $D2$ instantons correspond
6073: to membranes wrapping three-cycles in the $G_2$ space
6074: which do not extend in the
6075: dilaton direction \cite{HM,KMcG}. The embedding of IIA
6076: orientifolds into $G_2$ manifolds and the comparison of the
6077: respective effective actions is the subject of section \ref{G2_embedding}.
6078:
6079: Let us next extend this observation to IIB orientifolds. For
6080: simplicity we set $F=0$ for these cases, since brane fluxes would
6081: correct the K\"ahler coordinates as discussed e.g.\ in \cite{JL}.
6082: Hence, the calibration
6083: conditions for the respective D$(p-1)$-instantons read \cite{MMMS}
6084: \beq \label{cal_IIB}
6085: \Em^*\big( e^{-B_2 + iJ} \big)_p = e^{i\theta_{D(p-1)}} \sqrt{\det {\varphi^*(\hat g+ \hat B_2)}}\ d^p \lambda\ ,
6086: \qquad p=2,4,6\ ,
6087: \eeq
6088: where $\big( e^{-B_2 + iJ} \big)_p$ denotes the $p$-form in the sum over even forms.
6089: In order that these instantons preserve the same supersymmetry as the orientifold planes
6090: we furthermore have to set $\theta_{D(p-1)} = \theta_{O(p+3)}$, where $\theta_{O(p+3)}$ is given in
6091: \eqref{cal_sOp}. Multiplying \eqref{cal_IIB} by $e^{-\phi}$ and comparing real and imaginary parts we find
6092: \beq \label{pull-e}
6093: \Em^*\fe_4 = e^{-\phi} \sqrt{\det {\varphi^*(\hat g+ \hat B_2)}}\ d^4 \lambda\ ,
6094: \eeq
6095: where $\fe_4$ is the four-form in $\fe$ defined in \eqref{def-A} and
6096: we have only displayed the equation for $D3$ instantons. Furthermore, by comparing \eqref{instact}
6097: and \eqref{def-A} one finds that $\int_{\cW_{p}} \Em^* \fa$ exactly reproduces the Chern-Simons action,
6098: since the vectors in the expansions of the R-R forms $C_p$ vanish when the pulled back to $\cW_p \subset Y$.
6099: Hence, together with \eqref{pull-e} we conclude that the instanton actions take the form
6100: \beq
6101: S_{D3}=i\mu_4 \int_{\cW_4} \Em^*\fe_4 - i \Em^* \fa = -i\mu_4\, T_\alpha\, \int_{\cW_4} \Em^* \tilde \omega^\alpha\ ,
6102: \eeq
6103: where the definition of $T_\alpha$ is given in \eqref{def-coordsO3}. This shows that also in type IIB
6104: orientifolds the $N=1$ K\"ahler coordinates defined in \eqref{def-coordsO3} and \eqref{def-coordsO5} linearize
6105: the instanton actions. By a similar reasoning as in the IIA case this ensures holomorphicity
6106: of instanton induced superpotentials in these coordinates.
6107:
6108: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6109: %
6110: % Embedding into M- and F-theory
6111: %
6112: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6113:
6114:
6115: \chapter{Embedding into M- and F-theory}
6116: \label{M-F-embedding}
6117:
6118: In this chapter we discuss the embedding of type IIA and
6119: type IIB orientifolds into compactifications of M- and F-theory.
6120: Let us first review the basic idea, by briefly introducing F- and
6121: M-theory in the limit needed for our considerations.
6122:
6123: F-theory provides a geometrical interpretation of the non-perturbative
6124: $Sl(2,\mathbb{Z})$ symmetry \eqref{Sl2} of type IIB string theory.
6125: Under this symmetry the complex dilaton $\tau$
6126: transforms in a non-trivial manner and
6127: can be interpreted as the complex structure modulus of
6128: a two-dimensional torus. In \cite{Vafa} this idea was
6129: put forward in arguing for a natural interpretation
6130: in terms of a twelve-dimensional F-theory.
6131: Compactifying this theory on a two-torus gives back type IIB in ten dimensions.
6132: However, in going to lower dimensions, this torus can be fibered over the
6133: internal manifold. Compactification of F-theory on such elliptically
6134: fibered manifolds $Y_{n+2} \rightarrow B_{n}$ is defined to be type IIB string
6135: theory compactified on the base $B_n$, with a complex dilaton field $\tau$
6136: varying over the internal manifold. One interesting case is when
6137: $Y_4$ is a elliptically fibered Calabi-Yau fourfold with base $B_3$.
6138: It was shown in \cite{Sen} that in a special limit which corresponds to
6139: a weak coupling limit of type IIB string theory the two-fold cover of $B_3$
6140: is a Calabi-Yau manifold. Furthermore, the compactification on $B_3$ corresponds to an
6141: orientifold compactification with $O7$ planes and $D7$ branes, which are
6142: located at points where the torus fibers become singular. This limit is called
6143: the orientifold limit
6144: \beq \label{orientifold_limit}
6145: \text{F-theory}\ /\ Y_4 \quad \xrightarrow[\text{limit}]{\text{\quad orientifold\quad }}
6146: \quad \text{Type IIB}\ /\ \mathcal{O}Y_6 \ .
6147: \eeq
6148: Section \ref{F-theory} is devoted to check this correspondence for the effective bulk
6149: actions of the two theories. However, since there is no known effective action
6150: for F-theory we will take a detour over M-theory compactified on $Y_4$.
6151: We compare the resulting three-dimensional effective action with the $D=3$ action
6152: obtained by compactifying the $O3/O7$ orientifold action on a circle. Later on
6153: we lift the correspondence to $D=4$ and compare it with \eqref{orientifold_limit}.
6154:
6155: In section \ref{G2_embedding} we discuss the embedding of Type IIA orientifolds into M-theory.
6156: Recall that type IIA supergravity can be obtained by compactifying
6157: 11-dimensional supergravity (the low energy limit of M-theory) on a circle.
6158: Correspondingly the $D=4,N=2$ theories arising in Calabi-Yau compactifications
6159: are lifted as
6160: \beq
6161: \text{Type IIA}\ /\ Y_6 \quad \cong \quad \text{M-theory}\ /\ S^1 \times Y_6 \ .
6162: \eeq
6163: Hence, the immediate question is to find some analog for
6164: the orientifold compactifications. In order to do that, one
6165: has to identify appropriate manifolds which upon compactification
6166: of M-theory (understood as 11-dimensional supergravity) yield a
6167: four-dimensional $N=1$ theory. Recalling that the number of supersymmetries
6168: is related to the number of covariantly constant spinors, the only possible
6169: candidates are seven-manifolds with structure group or holonomy $G_2$.
6170: This implies that the reduction of the $SO(7)$ spinor
6171: representation yields one singlet, which in the case of $G_2$ holonomy is furthermore
6172: covariantly constant with respect to the Levi-Cevita connection.
6173: It was argued in \cite{KMcG} that for a special class of
6174: $G_2$ manifolds $X$ the resulting four-dimensional theory coincides with
6175: the one of IIA Calabi-Yau orientifolds. Schematically one has
6176: \beq
6177: \text{Type IIA}\ /\ \mathcal{O}Y_6 \quad \cong \quad \text{M-theory}\ /\ X \ .
6178: \eeq
6179: In section \ref{G2_embedding} we verify this conjecture for a certain limit of the two
6180: theories. This enables us to match the $N=1$ characteristic functions determined in
6181: section \ref{Kpo_gaugeIIA} for IIA orientifolds with the one obtained for $G_2$ compactifications
6182: on $X$. As we will show, this includes the K\"ahler potential, the gauge-couplings as
6183: well as the flux superpotentials. In ref. \cite{TGL2} only part of the orientifold
6184: superpotentials were found to have an origin in an M-theory compactification on a
6185: manifold with $G_2$ holonomy. As we will show, the remaining terms are due
6186: to a non-trivial fibration of a manifold with $G_2$ structure introduced in \cite{CS,CCDLM}.
6187:
6188:
6189:
6190:
6191: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6192: %
6193: % F-Theory and CY IIB orientifolds
6194: %
6195: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6196:
6197: \section{F-theory and $O3/O7$ orientifolds}
6198: \label{F-theory}
6199:
6200: In this section we discuss the embedding of $O3/O7$ orientifolds
6201: into a F-theory compactification, which corresponds to the limit \eqref{orientifold_limit}.
6202: To analyze the two theories on the level of the effective bulk actions we start
6203: by compactifying M-theory on a Calabi-Yau four-fold.
6204: When shrinking the volume of the elliptic fiber the M-theory
6205: compactification on $Y_4$ is equivalent to an F-theory
6206: compactification on $Y_4$. We only perform this limit at the
6207: very end and rather compare the two theories in three
6208: dimensions. In order to do that we first briefly review
6209: compactifications of eleven-dimensional supergravity on
6210: Calabi-Yau fourfolds following \cite{HL,BHS}. We determine the effective action
6211: and characteristic functions encoding the supergravity theory.
6212: Next we compactify the four-dimensional effective action
6213: of $O3/O7$ orientifolds to three dimensions on a circle.
6214: We are then in the position to show, that the characteristic
6215: data of the two three-dimensional theories coincide if we
6216: choose a Calabi-Yau fourfold of the form
6217: \beq \label{def-Z}
6218: Y_4 = (Y \times T^2)/\hat \sigma\ ,
6219: \eeq
6220: where $Y$ is a Calabi-Yau threefold and $\hat \sigma = (\sigma,-1,-1)$.
6221: The involution $\hat \sigma$ acts as a holomorphic isometric involution on $Y$ and
6222: inverts both coordinates on $T^2$. Note that $Y_4$ generically admits singularities if
6223: $\sigma$ has a non-trivial fix-point set. These have to be smoothed out which
6224: yields additional moduli in the theory. The analog on the orientifold are
6225: moduli corresponding to D-branes and orientifold planes. However, since
6226: we only restricted to the bulk fields we will also freeze moduli arising in
6227: the process of smoothing out $Y_4$ defined in \eqref{def-Z}.
6228: Having matched the three-dimensional theories we comment on the lift to
6229: $D=4$. Finally, we also include a brief discussion on the lift
6230: of orientifold three-form flux $G_3$ to four-form flux $G_4$.
6231:
6232:
6233:
6234:
6235:
6236: \subsubsection{M-theory compactified on a Calabi-Yau fourfold}
6237:
6238: Let us start by summarizing compactification of M-theory on a
6239: Calabi-Yau fourfold by following the analysis of \cite{HL,BHS}.
6240: The low energy effective action of 11d supergravity is given by \cite{CJS}
6241: \bea \label{11act}
6242: S^{(11)}=\int - \tfrac{1}{2} R *\mathbf{1} - \tfrac{1}{4} F_4 \wedge * F_4
6243: -\tfrac{1}{12} C_3 \wedge F_4 \wedge F_4\ ,
6244: \eea
6245: where $F_4=dC_3$ is the field strength of $C_3$.
6246: The three-form $C_3$ together with the eleven-dimensional metric are the
6247: only bosonic fields in the low energy description of M-theory.
6248: Recall that the action \eqref{11act} is given to lowest order
6249: in $\kappa_{11}$. One-loop corrections associated to the sigma model anomaly of
6250: a $M5$-brane contribute additional terms to
6251: \eqref{11act} and induce a $C_3$ tadpole term
6252: $-\frac{\chi(Y_4)}{24}$ \cite{DLM,SVW}.
6253: This contribution can be canceled by considering setups
6254: with a certain number of background $M3$-branes or
6255: switched on background fluxes.
6256: However, for the moment we keep our analysis simple in sticking to
6257: the action \eqref{11act} without extra source terms.
6258:
6259: The fields of the three-dimensional theory arise from the expansion
6260: of the eleven-dimensional supergravity fields into harmonic forms.
6261: For a Calabi-Yau fourfold $Y_4$, the only non-vanishing cohomologies are
6262: given by
6263: \bea \label{Z-cohom}
6264: H^{0}(Y_4)& =& H^{(0,0)}\ , \qquad H^{2}(Y_4)\ =\ H^{(1,1)}\ ,\qquad H^{3}(Y_4)\ =\ H^{(2,1)} \oplus H^{(1,2)} \ ,\nn \\[.2cm]
6265: H^{4}(Y_4)&=&H^{(4,0)} \oplus H^{(3,1)} \oplus H^{(2,2)} \oplus H^{(1,3)} \oplus H^{(0,4)}\ ,
6266: \eea
6267: with their Hodge duals $H^{5}$, $H^{6}$ and $H^{8}$.
6268: Let us extract the spectrum obtained by expansion into harmonic basis forms of these
6269: cohomologies. This is done in analogy to the case of type II compactifications discussed in
6270: chapter \ref{TypeII}.
6271: The deformations of the metric of the fourfold respecting the
6272: Calabi-Yau condition split into two sets: $h^{(1,1)}(Y_4)$ real scalar K\"ahler
6273: structure deformations $M^\cA(x)$ and $h^{(3,1)}(Y_4)$ complex structure moduli $Z^\cK(x)$.
6274: Similar to \eqref{def-v} and \eqref{cs} for Calabi-Yau threefolds they parameterize
6275: the expansions
6276: \beq \label{deform_4}
6277: J_F \ =\ M^\cA(x) e_\cA\ ,\qquad
6278: \delta g_{\bi \bj} = -\frac{1}{3 ||\Omega ||^2} \bar \Omega_{F\, \bi}^{\ \ \ klm} Z^\cK(x)
6279: \Phi_{\cK\, klm\bj}
6280: \eeq
6281: where $J_F$ and $\Omega_F$ are the K\"ahler form and the holomorphic $(4,0)$-form on
6282: the Calabi-Yau fourfold. The harmonic forms $e_\cA, \cA=1,\ldots, h^{(1,1)}(Y_4)$
6283: form a basis of $H^{(1,1)}(Y_4)$, while $\Phi_\cK, \cK=1,\ldots, h^{(3,1)}(Y_4)$
6284: form a basis of $H^{(3,1)}(Y_4)$.
6285: Also $C_3$ is expanded into harmonic forms via the Kaluza-Klein Ansatz
6286: \beq \label{A3exp}
6287: C_3 = A^\cA(x)\wedge e_\cA + N^I(x)\, \Psi_I + \bar N^I(x)\, \bar \Psi_{ I}\ ,
6288: \eeq
6289: where $A^\cA(x)$ are vectors and $N^I(x)$ are complex scalars in three dimensions.
6290: The harmonic forms $\Psi_I,\bar \Psi_I,I=1,\ldots h^{(2,1)}$
6291: define a basis of $H^{3}(Y_4)$, which can be chosen to obey
6292: %
6293: %
6294: \footnote{This needs some words of justification. First, recall
6295: that for a complex manifold $Y_4$ the filtration $F^3(\MM) = H^{(3,0)}$,
6296: $F^2(\MM) = H^{(3,0)} \oplus H^{(2,1)}$,
6297: etc.\ can be shown to consist of holomorphic bundles $F^i(\MM)$ over the space of complex
6298: structure deformations. Since $H^{(3,0)}$ is empty
6299: for Calabi-Yau fourfolds, $H^{(2,1)}$ is a holomorphic
6300: bundle and one can locally choose a basis
6301: $\psi_I(Z),\ \partial_{\bar Z^\cK}\psi_I=0 $. Hence, the holomorphic derivative is expanded as
6302: %\beq \label{psiexp}
6303: $ \partial_{Z^\cK} \psi_I = (\sigma_\cK)^{J}_{I} \psi_J +
6304: (\lambda_\cK)^{\bar J}_{I} \bar \psi_{\bar J}, $
6305: %\eeq
6306: where $(\sigma_\cK)^{J}_{I},\ (\lambda_\cK)^{\bar J}_{I}$ are functions of $Z,\bar Z$.
6307: One can now show, that there exists a basis $\Psi^I = M_I^{\bar J} \bar \psi$ (for some real
6308: $M_I^{\bar J}$) which obeys \eqref{der_Psi}.
6309: In order that this is the case one has to demand: $\partial_{Z^\cK}\ln M^{\bar I}_J=A_{\cK J}^{\ \ \ I}$,
6310: $B_{\bar \cK I}^{\ \ \ \bar J}=(M^{-1})^{\bar J}_K M^{\bar L}_I (\bar \lambda_{\bar \cK})^{K}_{\bar L}$
6311: and $A_{\cK I}^{\ \ \ \bar J}=-(\sigma_\cK)^{J}_{I}$. A possible definition of $M^{\bar I}_J$ can
6312: be found in \cite{HL}.}
6313: %
6314: %
6315: \beq \label{der_Psi}
6316: \partial_{Z^{\cK}} \Psi_I = A_{\cK I}^{\ \ \ J} \Psi_J \ , \qquad
6317: \partial_{\bar Z^{\cK}} \Psi_I = B_{\bar \cK I}^{\ \ \ \bar J} \bar \Psi_J\ ,
6318: \eeq
6319: where $A_{\cK I}^{\ \ \ J}$ and $B_{\bar \cK I}^{\ \ \ \bar J}$ are model dependent
6320: functions of $Z$ and $\bar Z$. Differentiating these equations with respect
6321: to $Z^{\cK}$ and $\bar Z^{\cL}$ and comparing
6322: $\partial_{Z^\cK} \partial_{\bar Z^\cL} \Psi_I$ with $\partial_{\bar Z^\cL} \partial_{Z^\cK} \Psi_I$
6323: we extract the consistency conditions
6324: \beq \label{DE}
6325: \partial_{\bar Z^{\cK}} A_{\cL I}^{\ \ \ J} = B_{\bar \cK I}^{\ \ \ \bar L} \, \bar B_{\cL \bar L}^{\ \ \ J}\ ,
6326: \qquad
6327: \partial_{\bar Z^{\cK}} \bar B_{\cL \bar I}^{\ \ \ J} = A_{\bar \cK \bar I}^{\ \ \ \bar L} \bar B_{\cL \bar L}^{\ \ \ J}\ .
6328: \eeq
6329: In summary, the bosonic part of the $D=3,N=2$ supergravity spectrum obtained by compactification
6330: on a Calabi-Yau fourfold is displayed in table \ref{Mspectrum}.
6331:
6332: \begin{table}[h]
6333: \begin{center}
6334: \begin{tabular}{|l|c|c|c|} \hline
6335: \rule[-0.3cm]{0cm}{0.9cm}
6336: {gravity multiplet} & 1 & $g^{(3)}_{pq}$\\ \hline
6337: \rule[-0.3cm]{0cm}{0.9cm}
6338: {vector multiplets} & $h^{(1,1)}$ & $(M^\cA,A^\cA)$ \\ \hline
6339: \rule[-0.3cm]{0cm}{0.9cm}
6340: chiral multiplets & $h^{(3,1)} + h^{(2,1)}$ & $Z^\cK$, $N^I$
6341: \\ \hline
6342: \end{tabular}
6343: \caption{\textit{$D=3,N=2$ spectrum for M-theory
6344: on a Calabi-Yau fourfold.}}\label{Mspectrum}
6345: \end{center}
6346: \end{table}
6347:
6348: Also the calculation of the three-dimensional low energy effective action is similar to
6349: the analysis performed in chapter \ref{TypeII}.
6350: The field strength $F_4=dC_3$ is evaluated by using \eqref{A3exp} and \eqref{der_Psi}
6351: as
6352: \beq \label{fieldstr_4}
6353: F_4 = dA^\cA \wedge e_\cA + D N^I \Psi_I + D \bar N^I \bar \Psi_I\ ,
6354: \eeq
6355: with
6356: \beq
6357: DN^I = dN^I + (N^J A_{\cK J}^{\ \ \ I} + \bar N^J B_{\cK \bar J}^{\ \ \ I}) dZ^\cK \ , \qquad D\bar N^I = \overline{DN^I}
6358: \eeq
6359: Inserting \eqref{deform_4}, \eqref{fieldstr_4} and \eqref{A3exp} and performing the standard Weyl rescaling
6360: the effective action takes the form \cite{HL}
6361: \bea \label{F-theory_act}
6362: S^{(3)}_{F}&= &\int -\tfrac{1}{2} R - G_{\cK \cL}\, dZ^\cK \wedge * dZ^\cL
6363: - \tfrac{1}{2} d \ln \cV \wedge * d \ln \cV - \tfrac{1}{2} G_{\cA \cB}\, dM^\cA \wedge * dM^\cB\nn\\
6364: && - \tfrac{1}{2} \cG_{I\bar J}\ D N^I \wedge * D\bar N^J
6365: - \tfrac{1}{2} \cV^2\, G_{\cA \cB}\ dA^\cA \wedge dA^\cB \nn \\
6366: && + \tfrac{i}{4} d_{\cA I \bar J}\ dA^\cA \wedge (N^I D\bar N^J - \bar N^I D N^J)\ ,
6367: \eea
6368: where $G_{\cK \cL}$, $\cG_{I\bar J}$ and $G_{\cA \cB}$ are the metrics on $H^4$, $H^3$ and $H^2$ respectively
6369: and will be discussed in turn.
6370: Let us first comment on the complex structure and K\"ahler structure deformations.
6371: The higher-dimensional analog
6372: of \eqref{csmetric} is the metric $G_{\cK \cL}$ on the space of complex structure
6373: deformations of $Y_4$. It is K\"ahler and takes the form
6374: \beq
6375: G_{\cK \bar \cL} = \partial_{Z^\cK} \partial_{\bar Z^\cL} K^{\rm cs}_F\ , \qquad
6376: K^{\rm cs}_F=-\ln\big[\int_{Y_4} \Omega_F \wedge \bar \Omega_F \big]\ .
6377: \eeq
6378: In analogy to \eqref{Kmetric} and \eqref{int-numbers}
6379: we define on the space of
6380: $(1,1)$-forms intersection numbers $d_{\cA \cB \cC \cD}$ and
6381: a metric $G_{\cA \cB}$ via
6382: \beq \label{d_ABCD}
6383: d_{\cA \cB \cC \cD} = \int_{Y_4} e_\cA \wedge e_\cB \wedge e_\cC \wedge e_\cD\ , \qquad
6384: G_{\cA \cB} = \frac{1}{2 \cV} \int_{Y_4} e_\cA \wedge * e_\cB\ ,
6385: \eeq
6386: where $\cV = \frac{1}{4!}\int J_F \wedge J_F \wedge J_F \wedge J_F$ is the volume of the
6387: Calabi-Yau four-fold.
6388:
6389: In contrast to a Calabi-Yau threefold the four-dimensional manifold $Y_4$ admits a third non-trivial
6390: cohomology $H^{3}(Y_4)$ with metric $G_{I \bar J}$. It has non-vanishing intersections
6391: $d_{\cA I \bar J}$ with $H^2$ such that
6392: \beq \label{d_AIJ}
6393: d_{\cA I \bar J} = i\int_{Y_4} e_\cA \wedge \Psi_I \wedge \bar \Psi_{J}\ ,\qquad
6394: \cG_{I\bar J} = \frac{1}{4 \cV} \int_{Y_4} \Psi_I \wedge * \bar \Psi_{J}= - \frac{\,M^\cA d_{\cA I \bar J}}{4 \cV}\ ,
6395: \eeq
6396: where we have used $*\bar \Psi_I = -iJ_F \wedge \bar \Psi_I$ in order to evaluate
6397: the last equality.
6398: However, in general $\cG_{I\bar J}$ as well as $d_{\cA I \bar J}$ depend on the complex structure
6399: deformations $Z^\cK$, since their definition involves the forms $\Psi_I(Z,\bar Z)$.
6400: Hence, by using \eqref{der_Psi} we obtain differential equations for $d_{\cA I \bar J}$ and $\cG_{I\bar J}$,
6401: which read
6402: \beq \label{Dd}
6403: \partial_{Z^\cK} d_{\cA I \bar J} = A_{\cK I}^{\ \ \ K}\, d_{\cA K \bar J}\ ,\qquad
6404: \partial_{Z^\cK} \cG_{I\bar J} = A_{\cK I}^{\ \ \ K}\, \cG_{K\bar J}\ .
6405: \eeq
6406:
6407: Having determined the effective action \eqref{F-theory_act} we can now proceed in two ways. Either
6408: we dualize the vectors $A^{\cA}$ into scalars $P_\cA$ and combine them into chiral multiplets
6409: $T_\cA=(M^\cA,P_\cA)$. The K\"ahler potential of this $D=3, N=2$ theory was determined in \cite{HL}.
6410: It takes the form
6411: \beq
6412: K_F(Z,N,T) = - \ln\Big[ \int_{Y_4} \Omega_F \wedge \bar \Omega_F \Big] - 3 \ln \cV(T,N)\ ,
6413: \eeq
6414: where $\cV(T,N)$ is the volume of $Y_4$, which depends implicitly on the K\"ahler coordinates.
6415: This is indeed analog to the situation in type IIB orientifolds with $O3/O7$ planes.
6416: However,
6417: in section \ref{IIB_lin} we explored a way around this implicit definition by
6418: changing to the dual picture. In $D=4$ this amounts to by keeping linear multiplets
6419: $(L^\alpha,D^\alpha_2)$ in the spectrum, which allows to give $K$ as an explicit function
6420: of $L^\alpha$. As we will review momentarily, this is equivalently
6421: true for the $D=3$ theory \eqref{F-theory_act} and amounts to keeping the vector multiplets $(M^\cA, A^\cA)$ in
6422: the spectrum \cite{BHS}.
6423:
6424: General $D=3,N=2$ supergravity theories with vector and chiral multiplets
6425: are discussed e.g.\ in \cite{BHS}. To avoid a detailed review of their results we make contact with
6426: section \ref{linear_multiplets} by observing that the effective action \eqref{kinetic_lin} for chiral and
6427: linear multiplets in $D=4$ can be translated to $D=3$ chiral-vector setups
6428: by replacing $dD_2^\cA$ with $dA^\cA$.\footnote{Furthermore, one has to replace in the potential
6429: \eqref{Lsc} the factor $3$ by a $4$ \cite{HL}.} Using these identifications, one compares
6430: \eqref{kinetic_lin} with \eqref{F-theory_act} to find
6431: \beq
6432: L^\cA = \frac{M^\cA}{\cV}\ ,\qquad
6433: \tilde K_{L^\cA L^\cB} = - \tfrac{1}{2}\, \cV^2\, G_{\cA \cB}\ .
6434: \eeq
6435: The kinetic potential for the vector multiplet $(L^\cA, A^\cA)$ is found to be \cite{BHS}
6436: \beq \label{kin_F}
6437: \tilde K(L,N,Z) = - \ln\Big[ \int_{Y_4} \Omega_F \wedge \bar \Omega_F \Big]
6438: + \ln\big(d_{\cA \cB \cC \cD} L^\cA L^\cB L^\cC L^\cD \big) +
6439: L^\cA \zeta_A
6440: \eeq
6441: with
6442: \beq
6443: \zeta^R_A = \tfrac12 d_{A I \bar J} \bar N^I N^J+ \omega_{A I J} N^I N^J + \omega_{A \bar I \bar J} \bar N^I \bar N^J \ .
6444: \eeq
6445: The functions $\omega_{A \bar I \bar J}(Z,\bar Z)$ obey
6446: \beq \label{Domega}
6447: \partial_{\bar Z^\cK} \omega_{\cA \bar I \bar J} = B_{\bar \cK I}^{\ \ \ \bar K} d_{\cA J \bar K}\ ,
6448: \eeq
6449: but are otherwise unconstraint. It is now straight forward to check, that
6450: the effective action determined in terms of $\tilde K(L,N,Z)$ is
6451: indeed equivalent to \eqref{F-theory_act} up to a total derivative \cite{BHS}.\footnote{More precisely
6452: one finds $\tfrac{i}{4} d_{\cA I \bar J}\ (\bar N^I DN^J - N^I D \bar N^J) = \I(\tilde K_{L^\cA Q^m} d Q^m) +$ total derivative, where $Q^m=(N^I,Z^\cK)$.} This ends our review of the M-theory compactification. In order to compare
6453: \eqref{kin_O33} with the $O3/O7$ orientifold data, we first have to compactify the orientifold theory to
6454: three dimensions.
6455:
6456: \subsubsection{The $O3/O7$ orientifolds in three-dimensions}
6457:
6458: Let us now compactify the four-dimensional $O3/O7$ orientifold theory determined by
6459: \eqref{S_scalarO3} on a circle $S^1$. In order to do that we partly follow \cite{HL}, where general
6460: compactifications of $D=4,N=1$ theories are discussed.
6461: Due to the fact that $D=4$ chiral multiplets reduce to $D=3$ multiplets we turn our
6462: attention to the vectors $V^\kappa$ with kinetic terms \eqref{red-vector}. In three dimensions
6463: vectors are dual to scalars and for four supercharges the dynamics can be encoded by a K\"ahler
6464: or kinetic potential.
6465: The Kaluza-Klein reduction is performed by choosing the Ansatz
6466: \beq \label{3d-Ansatz}
6467: g^{(4)}_{\mu \nu} = \left(
6468: \begin{array}{cc}
6469: g^{(3)}_{pq} + r^2 A_p^0 A_q^0& r^2 A_q^0\\ r^2 A_p^0 & r^2
6470: \end{array}
6471: \right)\ , \qquad V^\kappa_\mu = (A^\kappa_p+A^0_p\, n^\kappa, n^\kappa)\ ,
6472: \eeq
6473: where $A^0_p,\ A^\kappa_p,\ p=1,2,3$ are vectors and $n^\kappa$ as well as $r$ (the radius of $S^1$) are
6474: scalars in three dimensions. The resulting $D=3$ theory posses chiral multiplets $(z^k,\tau,G^a,T_\alpha)$ and
6475: vector multiplets $(A^0,r)$ and $(A^\kappa,n^\kappa)$. Next we dualize the vectors $A^\kappa$
6476: into scalars $\tilde n_\kappa$
6477: by the standard Lagrange multiplier method (see section \ref{revIIB}). However, we keep the vector multiplet
6478: $(A^0,r)$ and denote $L={r}^{-1}$.
6479: The scalars $\tilde n_\kappa$ and $n^\kappa$ combine into complex scalars $D_\kappa$ via \cite{FS,HL}
6480: \beq
6481: D_\kappa = -f_{\kappa \lambda}(z)\, n^\lambda + i\, \tilde n_\kappa\ ,
6482: \eeq
6483: where $f_{kl}(z)$ are the gauge-couplings of the $O3/O7$ theory given in \eqref{fholo}. One next inserts the
6484: Ansatz \eqref{3d-Ansatz} into the $D=4$ orientifolds action \eqref{S_scalarO3} and performs a Weyl rescaling
6485: to obtain a $D=3$ effective action with standard Einstein-Hilbert term. Using the definition of $D_\kappa$
6486: this action is encoded by a kinetic
6487: potential
6488: \beq
6489: \tilde K_3 = - \ln\Big[ \int_Y \Omega \wedge \bar \Omega \Big] + K^{k}(\tau, G,T) + \ln (L) + L \zeta^R\ ,
6490: \eeq
6491: where $K^{k}(\tau, G,T)$ and $\zeta^R$ are given in \eqref{kaehlerpot-Kk} and \eqref{def-zR}.
6492: Replacing the chiral multiplets $T_\alpha$ by vector multiplets $(A^\alpha,L^\alpha)$ we
6493: apply \eqref{tildeK_O3} to rewrite the kinetic potential as
6494: \beq \label{kin_O33}
6495: \tilde K_3 = - \ln\Big[ \int_Y \Omega \wedge \bar \Omega \Big] %K^{\rm cs}(z)
6496: - \ln\big(-i(\tau -\bar \tau)\big) +
6497: \ln( \cK_{\alpha \beta \gamma} L^\alpha L^\beta L^\gamma) + \ln (L) +
6498: L^\alpha \zeta^R_\alpha + L \zeta^R\ ,
6499: \eeq
6500: where
6501: \beq \label{def-zR}
6502: \zeta^R_\alpha = -\frac{i}{2(\tau-\bar \tau)}\ \KK_{\alpha b c}(G-\bar G)^b (G- \bar G)^c\ , \quad
6503: \zeta^R = -\tfrac{1}{2} (D_k + \bar D_k) (\R f_{kl} )^{-1} (D_k + \bar D_k)\ .
6504: \eeq
6505: The function $\zeta^R_\alpha =\zeta_\alpha+\bar\zeta_\alpha$ was already given in \eqref{zetaid}.
6506: $\tilde K_3$ fully encodes the dynamics of the chiral multiplets $z^k,\tau,G^a,D_k$
6507: and the vector multiplets $(A^\alpha,L^\alpha)$ and $(A,L)$ in three-dimensions. This enables us
6508: to compare the orientifold theory with the M-theory compactification discussed at the
6509: beginning of this section.
6510:
6511: \subsubsection{F-theory embedding of $O3/O7$ orientifolds}
6512:
6513: In order to discuss the F-theory embedding of the $O3/O7$ bulk orientifold theory, we
6514: restrict to the simple fourfolds defined in \eqref{def-Z}. Working on these manifolds
6515: the $\hat \sigma$ invariant cohomologies split as
6516: \bea \label{cohom_splitF}
6517: H^2(Y_4) &=& H^{2}_+(Y) \oplus H^2_+(T^2) \ , \qquad H^3(Y_4)\ =\ H^{3}_+(Y) \oplus
6518: \big(H^{2}_-(Y)\wedge H^1_-(T^2)\big)\nn \\
6519: H^4(Y_4) &=& H^4_+ (Y) \oplus \big(H^3_-(Y)\wedge H^1_-(T^2)\big) \oplus \big(H^{2}_+(Y) \wedge H^2_+(T^2)\big)\ ,
6520: \eea
6521: where $H^q_\pm(Y)$ are the cohomology groups of $Y$ introduced in \eqref{H3split} and we denote
6522: by $H^1_-(T^2),\ H^2_+(T^2)$ the cohomologies of $T^2$. We denote a basis of the $T^2$-cohomologies
6523: by $\alpha^{(1,0)},\alpha^{(0,1)} \in H^1_-(T^2)$ and $\vol(T^2) \in H^2_+(T^2)$.\footnote{ Recall, that for $T^2$
6524: one finds $h^{(0,0)}_+=h^{(1,1)}_+=h^{(1,0)}_-=h^{(0,1)}_-=1$.}
6525: We next analyze the spectrum and couplings of the three-dimensional
6526: theory \eqref{F-theory_act} on the manifolds \eqref{def-Z}. Let us start with the complex structure deformations $Z^\cK$.
6527: {}From \eqref{cohom_splitF} one concludes, that the only $(3,1)$-forms in $H^4(Y_4)$ arise from
6528: the cohomology $H^{(2,1)}_-(Y)\wedge H^{(1,0)}_-(T^2)$ and $H^{(3,0)}_-(Y)\wedge H^{(0,1)}_-(T^2)$.
6529: Hence we set
6530: \beq
6531: Z^\cK \equiv (\tau,\ z^k)\ , \qquad \cK=0,\ldots, h^{2,1}_-(Y) \ .
6532: \eeq
6533: This is consistent with the fact that in F-theory the complex dilaton
6534: $\tau$ becomes the complex structure modulus
6535: of the torus fiber of the fourfold $Y_4$ given in \eqref{def-Z}. Hence, we
6536: will set $\alpha^{(1,0)}=dq - \tau dp$ and lift $\tau$
6537: to one of the complex structure deformations of $Y_4$.
6538: Moreover,
6539: in the orientifold limit the complex structure deformations of the orientifold $z^k$ are
6540: the complex structure deformations of the base of $Y_4$. On \eqref{def-Z} also
6541: the holomorphic four-form $\Omega_F$ splits as $\Omega_F=\Omega\wedge \alpha^{(1,0)}$,
6542: such that
6543: \beq \label{split_OO}
6544: \ln\Big[ \int_{Y_4} \Omega_F \wedge \bar \Omega_F \Big] = \ln\Big[-i\int_Y \Omega \wedge \bar \Omega \Big]
6545: + \ln\big[-i(\tau - \bar \tau) \big]\ ,
6546: \eeq
6547: where we have used $\int_{T^2} dq\wedge dp = 1$.
6548:
6549: The K\"ahler structure deformations of $Y_4$ assembled into the vector multiplets
6550: $(M^\cA/\cV, A^\cA)=(L^\cA, A^\cA)$. These split under the decomposition
6551: \eqref{cohom_splitF} into one modulus parameterizing the torus volume and
6552: $h^{(1,1)}_+$ K\"ahler structure deformations of $Y/\sigma$. In three dimensions
6553: this has an obvious counterpart in the orientifold theory, since an additional
6554: K\"ahler modulus $L=r^{-1}$ arose from the compactification on $S^1$.
6555: This leads us to identify
6556: \beq \label{id_L}
6557: L^\cA \equiv (L,L^\alpha)\ ,\qquad A^\cA \equiv (A^0,A^\alpha)\ ,\qquad \cA=0,\ldots, h^{1,1}_+(Y) \ .
6558: \eeq
6559: Note that this implies that one matches the volume modulus of $T^2$ with the inverse
6560: radius $L=r^{-1}$ of the $S^1$. Also the corresponding intersection numbers
6561: \eqref{d_ABCD} split on the manifolds \eqref{def-Z} as
6562: \beq
6563: d_{\cA \cB \cC \cD} \rightarrow d_{0 \alpha \beta \gamma}\ ,
6564: \eeq
6565: with all other intersections vanishing. Here we have chosen $e_0=\text{vol}(T^2)$
6566: to be the invariant volume form of $T^2$. This implies that in the kinetic potential
6567: \eqref{kin_F} the logarithm involving $L^\alpha$ splits as
6568: \beq \label{split_LLLL}
6569: \ln\big(d_{\cA \cB \cC \cD} L^\cA L^\cB L^\cC L^\cD \big) = \ln L +
6570: \ln\big(\cK_{\alpha \beta \gamma} L^\alpha L^\beta L^\gamma \big)\ ,
6571: \eeq
6572: where we have identified $d_{0\alpha \beta \gamma}=\cK_{\alpha \beta \gamma}$, being the
6573: intersections of $H^2_+(Y)$.
6574:
6575: Finally, the remaining chiral multiplets $N^I$ and the orientifold fields $G^a,D_\lambda$
6576: have to be matched
6577: \beq
6578: N^I \equiv (G^a,D_\lambda)\ , \qquad I = 1, \ldots, h^{(1,1)}_-(Y) + h^{(2,1)}_+(Y)\ .
6579: \eeq
6580: Once again, this is consistent with the split \eqref{cohom_splitF}
6581: of $H^3(Y_4)$. The intersection numbers $d_{A I \bar J }$ given in \eqref{d_AIJ}
6582: decompose as
6583: \beq
6584: d_{\cA I \bar J} \rightarrow d_{0\kappa \lambda}\ , d_{\alpha a b}\ ,
6585: \eeq
6586: while all other intersections vanish. Note however, that in general
6587: $d_{\cA I \bar J}$ depends on the complex structure moduli $Z^\cK$ and
6588: a naive identification $d_{\alpha a b} \cong \cK_{\alpha a b}$ can only
6589: be true up to a complex structure dependent part. To extract this dependence
6590: we can proceed in two ways. Either we compare the two kinetic potentials
6591: \eqref{kin_F} and \eqref{kin_O33} to determine
6592: $d_{\cA I\bar J}$ as well as $\omega_{\cA IJ}$ and check if the equations
6593: \eqref{Dd}, \eqref{Domega} and \eqref{DE} are obeyed. However, we choose
6594: a different route and look for simple solutions of the consistency conditions
6595: \eqref{DE}. Having determined $A_{\cK I}^{\ \ \ J}$ and $B_{\bar \cK I}^{\ \ \ \bar J}$
6596: we are in the position to solve \eqref{Dd}, \eqref{Domega} to determine $d_{\cA I \bar J}$
6597: and $\omega_{\cA I \bar J}$.
6598:
6599: To construct a simple solution to \eqref{DE} we start with a
6600: holomorphic functions $f_{IJ}(Z)$, which can arise e.g.~as gauge couplings of a
6601: supersymmetric theory. In terms of $f_{IJ}$ the equations \eqref{DE} are solved by
6602: \beq
6603: A_{\cK I}^{\ \ \ J} = - (\R f)^{-1\, JK}\, \partial_{Z^\cK} f_{KI} \ , \qquad
6604: B_{\bar \cK I}^{\ \ \ \bar J} = (\R f)^{-1\, JK}\, \partial_{\bar Z^\cK} \bar f_{KI}\ .
6605: \eeq
6606: Relevant for the orientifold embedding are the two special cases
6607: \beq
6608: f_{\kappa \lambda}(Z^k) = f_{\kappa \lambda}(z^k)\ , \qquad f_{00}(Z^0) = -i\tau\ ,
6609: \eeq
6610: where $f_{\kappa \lambda}$ are the gauge-couplings of the orientifold given in \eqref{fholo}
6611: and $-i\tau$ are the gauge-couplings of a gauge-theory on space-time filling $D3$ branes (see for example \cite{GGJL}).
6612: Not to surprisingly, these are exactly the right functions to match the kinetic
6613: potentials \eqref{kin_F} and \eqref{kin_O33}. Namely, consistent with
6614: \eqref{Dd} and \eqref{Domega} we identify
6615: \beq \label{couplings}
6616: d_{0 \kappa \lambda} = \omega_{0 \kappa \lambda} = (\R f)^{\kappa \lambda}\ , \qquad
6617: d_{\alpha a b} = \omega_{\alpha a b} = \frac{1}{\tau-\bar \tau} \cK_{\alpha a b}\ ,
6618: \eeq
6619: where $\cK_{\alpha a b}$ are the intersections on $Y$, which are independent of $\tau$ and $z^k$.
6620: Equations \eqref{split_OO}, \eqref{split_LLLL} and \eqref{couplings} imply
6621: that the kinetic potential of the M-theory compactification
6622: reduces to the one for $O3/O7$ orientifolds on the Calabi-Yau fourfold \eqref{def-Z}.
6623:
6624: The final step is to lift this correspondence to four dimensions. On the orientifold side
6625: this simply amounts to performing the decompactification limit $L_0=r_0^{-1} \rightarrow 0$,
6626: where $r_0$ arises in $r_0+r(x)$ as the background radius. Of course, the resulting theory
6627: coincides with the $D=4$ orientifold theory, if identifying the correct four-vectors.
6628: More subtle is the lift of the M-theory compactification, which is known as the F-theory limit.
6629: It amounts to shrinking the volume of the two-torus (identified in \eqref{id_L} with $L_0$)
6630: on an elliptically fibered Calabi-Yau fourfold. However, for the simple manifold \eqref{def-Z}
6631: this limes is rather straightforward and coincides with the decompactification limit for the
6632: orientifold.
6633:
6634:
6635: %Using kinetic potential \eqref{kin_O33} this amounts to
6636: %$\tilde K_{LL} dL\wedge *dL=-(r+r^0)^{-2} dr \wedge * dr \rightarrow 0$.
6637:
6638:
6639: In addition to the bulk theory one can allow for non-trivial four-form background flux
6640: $G_4=\big<dC_3\big>$ on $Y_4$. The theory will be changed by a non-vanishing
6641: potential, which is obtained from the Gukov-Vafa-Witten superpotential $\int \Omega_F \wedge G_4$.
6642: In order to relate it to the
6643: $O3/O7$ orientifold three-form flux $G_3$ given in \eqref{fluxesB}
6644: one locally writes \cite{DRS,GSS,GKP}
6645: \beq
6646: G_{4} = - \frac{G_{3} \wedge \alpha^{(0,1)}}{\tau - \bar \tau} + h.c. \ .
6647: \eeq
6648: This implies that the Gukov-Vafa-Witten superpotential reduces as
6649: \beq
6650: \int_{Y_4} \Omega_F \wedge G_4 = \int_Y \Omega \wedge G_3\ ,
6651: \eeq
6652: which coincides with the orientifold superpotential found in \eqref{superpot}.
6653:
6654: This ends our discussion of the F-theory embedding of $O3/O7$ orientifolds.
6655: Their are many directions for further research. It would be
6656: desirable to include $D7$ branes into the setup, which correspond to certain
6657: singularities on the Calabi-Yau fourfold. The naive fourfold given
6658: \eqref{def-Z} is only valid in the regime were moduli for D-branes and orientifold
6659: branes are frozen. F-theory compactifications provide powerful tools to
6660: approach regimes where these fields are included \cite{non-pert}.
6661: A second issue is to discuss moduli stabilization in those setups,
6662: resent results \cite{DSFGK} suggest that all moduli can be stabilized
6663: in F-theory compactifications by including fluxes and non-perturbative corrections.
6664:
6665:
6666:
6667: %In refs. \cite{HL2,BHS} it was argued
6668: %that after including certain higher order corrections to \eqref{11act} and
6669: %background flux $G_4$ one obtains a gauged supergravity in three dimensions.
6670: %Some of these results should have a four-dimensional counterpart.
6671:
6672:
6673:
6674: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6675: %
6676: % Type IIA orientifolds and special $G_2$
6677: %
6678: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6679:
6680:
6681: \section{Type IIA orientifolds and special $G_2$ manifolds}
6682: \label{G2_embedding}
6683:
6684: In this section we discuss the relationship between the type IIA
6685: Calabi-Yau orientifolds considered so far
6686: and $G_2$ compactifications of M-theory.
6687: In refs.\ \cite{KMcG} it was argued that for a specific class
6688: of $G_2$ compactifications $X$, type IIA orientifolds appear at special
6689: loci in their moduli space.
6690: More precisely,
6691: these $G_2$ manifolds have to be such that they admit the form
6692: \beq \label{spG_2}
6693: X\ =\ (Y \times S^1)/{\hat \sigma}\ ,
6694: \eeq
6695: where $Y$ is a Calabi-Yau threefold and $\hat \sigma = (\sigma,-1)$
6696: is an involution which inverts the coordinates of the circle $S^1$
6697: and acts as an anti-holomorphic isometric involution on $Y$.
6698: $\sigma$ and $\hat \sigma$ can have a non-trivial fix-point set
6699: and as a consequence $X$ is a singular $G_2$ manifold.
6700: In terms of the
6701: type IIA orientifolds the fix-points of $\sigma$ are the locations
6702: of the $O6$ planes in $Y$ and as we already discussed earlier cancellation of
6703: the appearing tadpoles require the presence
6704: of appropriate $D6$-branes. In this paper we froze all
6705: excitation of the $D6$-branes and only discussed the effective action
6706: of the orientifold bulk. In terms of $G_2$ compactification this
6707: corresponds to the limit where $X$ is smoothed out and all additional
6708: moduli arising in this process are frozen.
6709:
6710:
6711:
6712: The purpose of this section is to check the embedding of type IIA
6713: orientifolds into $G_2$ compactifications of M-theory
6714: at the level of the $N=1$ effective action.
6715: For orientifolds the effective action was derived in sections
6716: 3 and 4 and so as a first step we need to
6717: recall the effective action of M-theory
6718: (or rather eleven-dimensional supergravity) on smooth $G_2$ manifolds
6719: \cite{PT, HM, Hitchin1, GPap,BW}.
6720:
6721: The bosonic part of the eleven-dimensional
6722: supergravity theory was already given in equation \eqref{11act}.
6723: It encodes the dynamic of the bosonic components $g_{11}$ and $C_3$ of the supergravity multiplet.
6724: As in the reduction on Calabi-Yau manifolds one chooses the background metric
6725: to admit a block-diagonal form
6726: \beq \label{lin-el}
6727: ds^2 = ds^2_4(x) + ds^2_{G_2}(y)\ ,
6728: \eeq
6729: where $ds^2_4$ and $ds^2_{G_2}$ are the line elements of a Minkowski
6730: and a $G_2$ metric, respectively.
6731: The Kaluza-Klein Ansatz
6732: for the three-form $C_3$ reads
6733: \beq
6734: C_3 = c^i(x)\, \phi_i + A^\alpha(x) \wedge \omega_\alpha \ , \qquad i=1,\ldots,b^3(X)\ ,\quad \alpha = 1, \ldots, b^2(X)
6735: \eeq
6736: where $c^i$ are real scalars and $A^\alpha$ are one-forms in four space-time dimensions.
6737: The harmonic forms $\phi_i$ and $\omega_\alpha$ span a basis of
6738: $H^3(X)$ and $H^2(X)$, respectively.
6739: The $G_2$ holonomy allows for exactly one covariantly
6740: constant spinor which can be used to define a real, harmonic and
6741: covariantly constant
6742: three-form $\Phi$.\footnote{The covariantly constant
6743: three-form is the analog of the holomorphic three-form $\Omega$
6744: %(or the K\"ahler form $J$),
6745: on Calabi-Yau manifolds.}
6746: The deformation space of the $G_2$ metric has dimension $b^3(X)=\dim H^3(X,\bbR)$ and
6747: can be parameterized by expanding $\Phi$
6748: %($d\Phi=d*\Phi=0$)
6749: into the basis $\phi_i$ \cite{Joyce}
6750: \beq
6751: \Phi = s^i(x)\, \phi_i \ .
6752: \eeq
6753: One combines the real scalars
6754: $s^i$ and $c^i$ into complex coordinates according to
6755: \beq
6756: S^i = c^i + i s^i\ ,
6757: \eeq
6758: which form the bosonic components of $b^{3}(X)$ chiral multiplets.
6759: In addition the effective four-dimensional supergravity features
6760: $b^{2}(X)$ vector multiplets with the $A^\alpha$ as bosonic components.
6761: Due to the $N=1$ supersymmetry,
6762: the couplings of these multiplet are again expressed in terms of
6763: a K\"ahler potential $K_{G_2}$, gauge-kinetic
6764: coupling functions $f_{G_2}$ and a (flux induced) superpotential $W_{G_2}$.
6765: Let us discuss these functions in turn.
6766:
6767: The K\"ahler potential was found to be \cite{HM,Hitchin1,GPap,BW}
6768: \beq \label{G_2Kpo}
6769: K_{G_2}\ =\ - 3 \ln \big( \tfrac{1}{ \kappa^2_{11}} \tfrac{1}{7} \int_X \Phi \wedge * \Phi \big)\ ,
6770: \eeq
6771: where $\frac{1}{7}\int \Phi \wedge * \Phi = \text{vol}(X)$ is the volume of the $G_2$ manifold $X$.
6772: The associated K\"ahler metric is given by
6773: \beq \label{G_2Kmetr}
6774: \partial_{i}\bar \partial_{\bj} K_{G_2}\ =\ \tfrac{1}{4} \text{vol}(X)^{-1} \int_X \phi_i \wedge * \phi_j\ ,
6775: \qquad
6776: \partial_{i} K_{G_2}\ =\ \tfrac{i}{2} \text{vol}(X)^{-1} \int_X \phi_i \wedge * \Phi\ ,
6777: \eeq
6778: and obeys the no-scale type condition
6779: \beq
6780: (\partial_{i} K_{G_2})\, K_{G_2}^{i \jb}\, (\partial_{\bj} K_{G_2}) = 7\ .
6781: \eeq
6782:
6783: The holomorphic gauge coupling functions
6784: $f_{G_2}$ arise from the couplings of $C_3$ in
6785: \eqref{11act}. At the tree level they are linear in $S^i$
6786: and read \cite{HM,GPap}
6787: \beq \label{gauge-kinG}
6788: (f_{G_2})_{\alpha \beta} = \tfrac{i}{2 \kappa_{11}^2}\,
6789: S^i\int_X \phi_i \wedge \omega_\alpha \wedge \omega_\beta\ .
6790: \eeq
6791:
6792: Finally, non-vanishing background flux $G_4$ of $F_4 =dC_3$
6793: induces a scalar potential which via \eqref{N=1pot}
6794: can be expressed in terms of the superpotential
6795: \cite{Gukov,AS,BW}
6796: \beq \label{G_2supo}
6797: W_{G_2}\ =\ \tfrac{1}{4 \kappa_{11}^2}\int_X \big(\tfrac{1}{2} C_3 +i\Phi) \wedge G_4\ .
6798: \eeq
6799: (The factor $1/2$ ensures holomorphicity of $W_{G_2}$ in the
6800: coordinates $S^i$ and compensates the quadratic dependence on $C_3$
6801: \cite{BW}.)
6802:
6803:
6804: In order to compare the low energy effective theory of $G_2$
6805: compactifications
6806: % \eqref{G_2Kpo}, \eqref{gauge-kinG} and \eqref{G_2supo}
6807: with the one of the orientifold we first have to restrict to the special $G_2$ manifolds
6808: $X$ introduced in \eqref{spG_2}.
6809: This can be done by analyzing how the cohomologies of $X$ are related to the
6810: ones of $Y$. As in equation \eqref{cohom-split} we consider the splits $H^p(Y)=H^p_+ \oplus H^p_-$
6811: of the cohomologies into eigenspaces
6812: of the involution $\sigma$. Working on the $G_2$ manifold $X$ given in \eqref{spG_2}
6813: we thus find the $\hat \sigma$-invariant cohomologies
6814: \beq \label{splcoho}
6815: \begin{array}{cclcrcl}
6816: H^2(X) &=& H^2_+(Y)\ , &\ &
6817: H^3(X) &=& H^3_+(Y) \oplus \big[H^{2}_-(Y)\wedge H_-^1(S^1)\big] \ ,\Big. \\
6818: H^5(X) &=& H^4_-(Y)\wedge H_-^1(S^1)\ , &&
6819: H^4(X) &=& H^4_+(Y) \oplus \big[H^{3}_-(Y)\wedge H_-^1(S^1)\big] \ ,
6820: \end{array}
6821: \eeq
6822: where $H^2(X)$ and $H^5(X)$ as well as $H^3(X)$ and $H^4(X)$ are
6823: Hodge duals. $H_-^1(S^1)$ is the one-dimensional space containing
6824: the odd one-form of $S^1$. The split of $H^3(X)$ induces a split of the $G_2$-form
6825: $\Phi$ which is most easily seen by introducing locally an orthonormal basis
6826: $(e^1,\ldots,e^7) \in \Lambda^1(X)$ of one-forms.
6827: In terms of this basis one has \cite{Joyce,Hitchin1,CS}
6828: \beq\label{Phidecomp}
6829: \Phi
6830: % \ =\ e^{127} + e^{347} + e^{567} + e^{135} - e^{236} - e^{146} - e^{245}
6831: \ =\ J_M \wedge e^7 + \text{Re} \Omega_M \ ,
6832: \qquad
6833: *\Phi= \tfrac12 J_M\wedge J_M + \I \Omega_M\wedge e^7\ ,
6834: \eeq
6835: where
6836: \bea \label{defJO}
6837: J_M = e^1\wedge e^2 + e^3\wedge e^4 + e^5\wedge e^6\ , \quad
6838: \Omega_M = (e^1 + ie^2)\wedge(e^3+ie^4)\wedge(e^5+ie^6)\ .
6839: \eea
6840: Applied to the manifold \eqref{spG_2}
6841: we may interpret $e^7=dy^7$ as being the odd one-form
6842: along $S^1$. Since $\Phi$ is required to be invariant under
6843: $\hat\sigma$
6844: and $\sigma$ is anti-holomorphic the decomposition
6845: \eqref{Phidecomp} implies
6846: \beq \label{splitPhi}
6847: \hat \sigma^* J_M = - J_M\ , \qquad
6848: \hat \sigma^* \Omega_M = \bar \Omega_M\ .
6849: \eeq
6850: In terms of the
6851: basis vectors $e^1,\ldots,e^6$ this is ensured by choosing
6852: $e^4,e^5,e^6$ to be odd and
6853: $e^1,e^2,e^3$ to be even under $\sigma$.
6854: We see that $J_M$ and $\Omega_M$ satisfy
6855: the exact same conditions as the corresponding forms of the
6856: orientifold
6857: (c.f.\ \eqref{constrJ}, \eqref{constrO}) and thus have to be proportional to
6858: $J$ and $C\Omega$ used in section \ref{IIA_orientifolds}. In order to
6859: determine the exact relation
6860: it is necessary to fix their relative normalization.
6861: The relation between
6862: $J_M$ and the K\"ahler form $J$ in the string frame
6863: can be determined from the relation of the respective metrics.
6864: Reducing eleven-dimensional supergravity to type IIA supergravity in
6865: the string frame
6866: requires the line element \eqref{lin-el} of the eleven-dimensional
6867: metric to take the form
6868: \beq \label{metransatz}
6869: ds^2 = e^{-{2 \hat \phi}/{3}} ds_4^2(x) +
6870: e^{-{2 \hat \phi}/{3}} g_{(s)\, ab}\, dy^a dy^b + e^{{4 \hat \phi}/{3}} (dy^7)^2\ ,
6871: \eeq
6872: where $a,b=1,\ldots,6$.
6873: The factors $e^{\hat \phi}$ of the ten-dimensional dilaton are
6874: chosen such that the type IIA
6875: supergravity action takes the standard form with
6876: $g_{(s)}$ being the Calabi-Yau metric in string frame
6877: (see e.g.~\cite{JPbook}).
6878: Consequently we have to identify
6879: \beq\label{JM}
6880: J_M=e^{-{2 \hat \phi}/{3}} J\ .
6881: \eeq
6882:
6883: Similarly, using \eqref{defJO} we find that the normalization of $\Omega_M$ is given by
6884: \bea \label{norm}
6885: J_M \wedge J_M \wedge J_M = \frac{3i}{4}\, \Omega_M \wedge \bar \Omega_M\ .
6886: \eea
6887: Integrating over $Y$ and using \eqref{JM}, \eqref{Kks} and \eqref{csmetric} we obtain
6888: \bea \label{normO}
6889: \Omega_M = e^{-\hat \phi-i\theta}
6890: e^{\frac{1}{2}(\Kcs - K^{\rm K})}\, \Omega = \sqrt{8} C\Omega\ ,
6891: \eea
6892: where $C$ is given in \eqref{def-C}.
6893: The phase $e^{i\theta}$ drops out in \eqref{norm} such
6894: that we can choose it
6895: as in \eqref{constrO} in order to fulfill \eqref{splitPhi}.
6896: Inserting $J_M$ and $\Omega_M$ into equation \eqref{splitPhi}
6897: one arrives at
6898: \beq \label{Phi-o}
6899: \Phi = J \wedge d\tilde y^7 + \sqrt{8} \text{Re}(C\Omega) \ ,
6900: \eeq
6901: where we defined $d\tilde y^7 = e^{-\frac{2 \hat \phi}{3}} dy^7$. The form
6902: $d\tilde y^7$ is normalized such that $\int_{S^1} d\tilde y^7=2\pi R$ where
6903: the metric \eqref{metransatz} was used and
6904: $R$ is the $\phi$-independent radius of the internal circle.
6905: We also set $\kappa^2_{10}=\kappa_{11}^2 / 2\pi R = 1$ henceforth.
6906: Using \eqref{Phi-o}, \eqref{Phidecomp}
6907: and \eqref{def-C} we calculate
6908: \beq \label{voldec}
6909: \tfrac{1}{\kappa^2_{11}}\, \tfrac{1}{7} \int \Phi \wedge * \Phi
6910: = e^{-\frac{4\hat \phi}{3}} \, \tfrac{1}{6} \int J \wedge J \wedge J \ ,
6911: \eeq
6912: which equivalently can be obtained by applying the split
6913: $\text{vol}(X)=\text{vol}(Y)\cdot\text{vol}(S^1)$ of the $G_2$ volume when evaluated in the metric \eqref{metransatz}.
6914: Inserting \eqref{voldec} into \eqref{G_2Kpo} using \eqref{def-C}
6915: we obtain
6916: \beq\label{IIAori}
6917: K_{G_2}\ =\
6918: - \ln \Big[ \tfrac{1}{6} \int_Y J \wedge J \wedge J \Big]
6919: - 2\ln\Big[2\int_Y \R(C \Omega)\wedge *_6 \R(C\Omega)\Big]\ .
6920: \eeq
6921: Thus we find exactly the K\"ahler potential
6922: $K$ of the type IIA orientifold as given in \eqref{N=1Kpot}.\footnote{%
6923: In terms of the Hitchin functionals \cite{Hitchin1} recently discussed in
6924: \cite{DGNV,Nekrasov} the reduction of the
6925: $G_2$ K\"ahler potential \eqref{G_2Kpo} corresponds to
6926: the split of the seven-dimensional Hitchin functional to the
6927: two six-dimensional ones \ref{IIAori}.}
6928:
6929:
6930: In order to compare the gauge kinetic functions and the superpotential
6931: we also need to identify the
6932: K\"ahler coordinates of the two theories.
6933: $C_3$ splits under the decomposition \eqref{splcoho}
6934: of the cohomologies as\footnote{We have introduced a factor of $\sqrt{2}$ for later convenience.}
6935: \beq \label{spl-C}
6936: C_3 = \hat B_2 \wedge d\tilde y^7 + \sqrt{2} \hat C_3 \ ,
6937: \eeq
6938: where $\hat B_2$ is an odd two-form on $Y$
6939: and $\hat C_3$ an even three-form on $Y$.
6940: Combining \eqref{Phi-o} and \eqref{spl-C} using \eqref{N=1coords}
6941: one finds
6942: \beq \label{spl-CPhi}
6943: S^i \phi_i\ =\ C_3 + i\Phi\ =\ \Jc \wedge d\tilde y^7 + \sqrt{2}\, \Omegac\ .
6944: \eeq
6945: As discussed after
6946: \eqref{N=1coords} the coefficients arising in the expansions of $\Jc$ and $\Omegac$
6947: into the basis $(\alpha_k,\beta^\lambda)$ of $H^{3}_+(Y)$
6948: and $\omega_a$ of $H^2(Y)$ are exactly the orientifold coordinates and
6949: therefore we have to identify
6950: $S^a \cong t^a$ and $S^K \cong (N^k,T_\lambda)$.
6951: With this information at hand, it is not difficult to show that the
6952: gauge-kinetic couplings \eqref{gauge-kinG} coincide with \eqref{gauge-A}.
6953: One splits $\phi_a = \omega_a \wedge d\tilde y^7$ and obtains
6954: \beq
6955: (f_{G_2})_{\alpha \beta}
6956: = \tfrac{i}{2} S^a \int_Y \omega_a \wedge \omega_\alpha \wedge
6957: \omega_\beta\ \sim i t^a \cK_{a\alpha\beta} = (f_{OY})_{\alpha \beta}\
6958: ,
6959: \eeq
6960: where the precise factor depends on the normalization of
6961: the gauge fields.
6962:
6963: It remains to compare the flux induced superpotentials \eqref{G_2supo}
6964: with \eqref{superpot1}. Using the
6965: cohomology splits \eqref{splcoho} and \eqref{spl-C}
6966: the background flux
6967: splits accordingly as $G_4 = H_3 \wedge d\tilde y^7 + \sqrt{2} F_4$.
6968: Inserted into \eqref{G_2supo}
6969: using \eqref{spl-CPhi} we arrive at
6970: \beq
6971: W_{G_2} = \tfrac{1}{\sqrt{8}}\int_Y \Jc \wedge F_4 +
6972: \tfrac{1}{\sqrt{8}}\int_Y \Omegac \wedge H_3
6973: \eeq
6974: Compared to \eqref{superpot1} the superpotential $W_{G_2}$ only
6975: includes terms proportional to the fluxes $H_3$ and $F_4$.\footnote{%
6976: The term proportional to $e_0$ in \eqref{superpot2} can be absorbed
6977: into a redefinition of $\R t^a$ \cite{BW}.}
6978: An interesting question is to identify the remaining terms
6979: in \eqref{superpot1} which are likely to arise once manifolds with
6980: $G_2$ structure (instead of $G_2$ holonomy) are considered. The term
6981: due to $F_2$ arises in compactifications on fibered
6982: $G_2$ manifolds $X \rightarrow Y$ \cite{CS,CCDLM}. In our case we
6983: restrict to circle fibrations over the quotient $Y/\sigma$, where $Y$
6984: is a Calabi-Yau manifold. We introduce the projection $\pi:X \rightarrow Y$. The metric on such
6985: a manifold takes the form
6986: \beq
6987: g_{G_2} = \alpha \otimes \alpha + \pi^* g\ ,
6988: \eeq
6989: where $g$ is the metric on $Y$ and $d\alpha=\pi^* F_2$. This implies
6990: that $X$ has not anymore $G_2$ holonomy but rather $G_2$ structure with
6991: $d\Phi = F_2\wedge J$ being not closed. Following \cite{BJ} this
6992: induces a superpotential term of the form
6993: \beq \label{non-G_2}
6994: W = \int_X (dC_3 + i d\Phi) \wedge (C_3+i\Phi) + \ldots = \int_Y F_2 \wedge \Jc \wedge \Jc + \ldots\ ,
6995: \eeq
6996: where $\Phi$ and $C_3$ are given in \eqref{Phi-o} and \eqref{spl-C} with $d\tilde y^7 = \alpha$
6997: and $dC_3 = \hat B_2 \wedge F_2 + \ldots$. This reproduces exactly the $F_2$ superpotential term \eqref{superpot2}
6998: in type IIA orientifolds. It remains to reveal the origin
6999: the superpotential term linear in $m^0$. Unfortunately, this is less straightforward and
7000: is likely to involve more general $G_2$ manifolds \cite{Witt}.\footnote{We like to thank A. Micu for discussions on this point.}
7001: It would be nice to make this more explicit and
7002: to point out the relation to the Scherk-Schwarz constructions of massive IIA supergravity.
7003: % One obvious possibility is to choose
7004: % a fibred $G_2$ manifold with $d\alpha=\pi^* (m^0 J+ F_2)$. This yields the desired
7005: % terms in \eqref{non-G_2}. It would be nice to understand this correspondence
7006: % in more detail and to point out the relation to the Scherk-Schwarz constructions
7007: % of massive IIA supergravity.
7008:
7009: \chapter{Conclusions}
7010:
7011: In this work we determined the low energy effective action for type IIB and type IIA
7012: Calabi-Yau orientifolds in the presence of background fluxes.
7013: In our analysis we did not specify a particular Calabi-Yau manifold but
7014: merely demanded that it admits an isometric involution $\sigma$.
7015: Furthermore, in order to preserve $N=1$ supersymmetry $\sigma$ was chosen to
7016: be a holomorphic map in type IIB and an anti-holomorphic map in type IIA.
7017: Depending on the explicit action of $\sigma$ on the holomorphic three-form $\Omega$,
7018: we analyzed three distinct cases: (1) orientifolds with $O3/O7$-planes, (2)
7019: orientifolds with $O5/O9$-planes and (3) orientifolds with $O6$-planes.
7020: For each case we calculated the characteristic functions of the
7021: corresponding $N=1$ supergravity theories and discussed their
7022: generic properties.
7023:
7024: In chapter \ref{effective_actO} we restricted to the case where background fluxes are absent
7025: and no potential is generated. We computed the effective action by a Kaluza-Klein analysis
7026: valid in the large volume limit and determined the
7027: chiral variables, the K\"ahler potentials and the gauge kinetic functions for
7028: all three setups. We found that the moduli space of the $N=1$ theory inherits
7029: a product structure $\tilde \cM^{\rm SK} \times \tilde \cM^Q$ from
7030: the underlying $N=2$ theory obtained by ordinary Calabi-Yau
7031: compactification of type II theories. $\tilde \cM^{\rm SK}$ is
7032: a special K\"ahler manifold parameterized by the complex structure deformations
7033: in type IIB and by the complexified K\"ahler deformations in type IIA.
7034: For type IIB orientifolds the second component $\tilde \cM^Q$ is parameterized
7035: by the periods of the complex even form $\fe - i\, \fa$ for setups with
7036: $O3/O7$ planes and by the periods of $\feh - i\, \fa$ for setups with $O5/O9$. The
7037: form $\fe+ i\,\feh =e^{-\hat \phi }\, e^{-\hat B_2+iJ}$ comprises of the complexified
7038: K\"ahler deformations while $\fa$ is a sum of the even R-R forms defined in
7039: \eqref{def-A}. On the other hand, for type
7040: IIA orientifolds with $O6$ planes $\tilde \cM^{\rm Q}$
7041: is spanned by the periods of the complex three-form $\Omegac=C_3 + 2i \R C\Omega$
7042: containing the complex structure deformations of the Calabi-Yau orientifold.
7043: $\tilde \cM^{\rm Q}$ is a K\"ahler submanifold inside the quaternionic manifold
7044: with a K\"ahler potential encoding the dynamics of the even/odd
7045: forms of the respective orientifold setup.
7046: Finally we showed that in the large volume -- large complex structure limit
7047: one finds mirror symmetric effective actions if one compares
7048: type IIA and type IIB supergravity compactified on mirror manifolds
7049: and in addition chooses a set of `mirror involutions'.
7050: For $\tilde\cM^K$ mirror symmetry amounts to a truncated versions
7051: of $N=2$ mirror symmetry in that it still relates
7052: two holomorphic prepotentials. In this case the corrections computed
7053: by mirror symmetry are likely to be analogous to the situation in $N=2$.
7054: For $\tilde\cM^Q$
7055: the situation is more involved since the geometry of the moduli
7056: space changes drastically. Nevertheless we were able to show that
7057: mirror symmetry holds in the large volume - large complex structure limit.
7058: However, understanding
7059: the nature of the corrections computed by mirror symmetry
7060: appear to be more involved and certainly deserves further study.
7061: It is interesting to note that mirror symmetry can be understood as an exchange
7062: of the odd form $\Omegac$ with the even forms $\fe+i\, \fa$ or $\feh+i\, \fa$ in accord with
7063: \cite{FMM}. Two choices of special coordinates in $\Omegac$ single out the corresponding orientifold
7064: setup on the mirror side. It would be desirable to reveal
7065: the origin of this mapping and finally to generalize it to non-Calabi-Yau compactifications.
7066:
7067: In chapter \ref{lin_geom_of_M} we presented a more detailed investigation of
7068: the $N=1$ moduli space of Calabi-Yau orientifold compactifications.
7069: The special K\"ahler manifold $\tilde \cM^{\rm SK}$ inherits its geometrical
7070: structure directly from $N=2$, such that we focused on
7071: the K\"ahler manifold $\tilde \cM^Q$ inside the quaternionic space.
7072: It turned out that the definition of the K\"ahler coordinates as well as
7073: the no-scale type conditions on $\tilde \cM^{\rm Q}$ can be more easily understood
7074: in terms of the `dual' formulation where some chiral multiplets of the Calabi-Yau
7075: orientifold are replaced by linear multiplets. After a brief review of $N=1$ supergravity with
7076: several linear multiplets we reformulated all three orientifold setups by dualizing a certain set of
7077: chiral multiplets. The transformation into
7078: linear multiplets corresponds to a Legendre transformation of
7079: the K\"ahler potential and coordinates. The new kinetic potential of $O3/O7$ and $O5/O9$
7080: orientifolds takes a particularly simple form induced from a tree-level prepotential.
7081: In contrast for $O6$ orientifolds it is given in terms of a generic prepotential satisfying the
7082: orientifold constrains and generically includes correction corresponding to world-sheet instantons in
7083: type IIB. For orientifolds with $O6$ planes the Legendre transform was essential to make contact
7084: with the underlying $N=2$ special geometry. As a byproduct we determined an entire new class of no-scale
7085: K\"ahler potentials which in the chiral formulation
7086: can only be given implicitly as the solution of some constraint equation.
7087: We closed this chapter by giving an explicit construction of the K\"ahler
7088: manifold $\tilde \cM^{\rm Q}$ replacing the $N=2$ c-map. The space
7089: $\tilde \cM^{\rm Q}$ was shown to admit a geometric structure similar to the one of the moduli
7090: space of supersymmetric Lagrangian submanifolds \cite{Hitchin2}. This also
7091: provides the ground for a more general investigation of non-Calabi-Yau orientifolds.
7092: Namely, we found that the K\"ahler potential of $\tilde \cM^{\rm Q}$ is the
7093: logarithm of Hitchins functional for a generalized complex sixfold
7094: evaluated for the simple even and odd forms associated to the orientifold setup.
7095:
7096: In chapter \ref{fluxesAB} we repeated the Kaluza-Klein compactification by additionally allowing
7097: for non-trivial background fluxes.
7098: In the $O3/O7$ case the background fluxes induce a non-trivial scalar potential
7099: which is determined in terms of a superpotential previously given in
7100: \cite{GVW,TV,GKP,BBHL}. We also included the scalar fields $(b^a,c^a)$ arising from the two
7101: type IIB two-forms $B_2$ and $C_2$. We showed that in this case the potential
7102: is unmodified which can be traced to the no-scale property of the K\"ahler potential.
7103: For orientifolds with $O5/O9$ planes the influence of background fluxes is more
7104: involved. This is due to the fact that the space-time two-form $C_2$ arising in the
7105: expansion of the RR field $\hat C_2$ remains in the spectrum. It combines with
7106: the dilaton into a linear multiplet, which only if it is massless can be dualized to
7107: a chiral multiplet. However, generic NS three-form background fluxes render this
7108: form massive. We therefore first restricted our attention to the case were the mass
7109: term vanishes which occurs if the magnetic fluxes arising from
7110: the NS three-form $H_3$ are set to zero. In the resulting
7111: chiral description the axion dual to $C_2$ is gauged with the gauge charges set by the
7112: electric fluxes. The scalar potential now consists of two distinct contributions.
7113: The term which depends on the RR fluxes arising from
7114: $F_3$ is obtained from a (truncated) superpotential of the previous case
7115: whereas the second contribution depends on the electric fluxes of
7116: $H_3$ and arises from $D$-terms which are present due to the gauged isometry.
7117: Finally, we also analyzed non-vanishing magnetic fluxes
7118: in the NS sector which can be described by an $N=1$ theory including a massive linear
7119: multiplet coupled to vector and chiral multiplets.
7120: In this case the scalar potential additionally
7121: includes a direct mass term for the scalar
7122: in the linear multiplet which is neither a $D$- nor an $F$-term.
7123: For type IIA orientifolds all background fluxes induce a superpotential $W$
7124: which depends on all geometrical moduli. It splits into the sum of two terms
7125: with one term depending on the RR fluxes and the complexified
7126: K\"ahler form $J_c$ while the second term
7127: features the NS fluxes and $\Omegac$.
7128: Both terms are expected to receive non-perturbative corrections
7129: from world-sheet- and D-brane instantons.
7130: We showed that for supersymmetric type IIA and type IIB instantons the respective actions
7131: are linear in the chiral coordinates and thus can result in holomorphic
7132: corrections to $W$.
7133:
7134: In the last chapter \ref{M-F-embedding} we analyzed the embedding of type IIB and type IIA
7135: orientifolds into F- and M-theory compactifications.
7136: Orientifolds with $O3/O7$-planes can be obtained as a limit of
7137: F-theory compactified on elliptically fibered Calabi-Yau fourfolds \cite{Vafa,Sen}.
7138: To check this correspondence on the level of the effective action we
7139: took a sideway by first compactifying M-theory on a Calabi-Yau fourfold.
7140: This yields a three-dimensional $N=2$ supergravity theory determined in terms
7141: of the characteristic data of the Calabi-Yau fourfold. Restricting to a specific
7142: fourfold this effective theory can be compared to the one obtained by
7143: compactifying the effective action of $O3/O7$ orientifolds
7144: on a circle to $D=3$. We determined simple solutions to the
7145: fourfold consistency conditions for which we found perfect matching between the
7146: orientifold and M-theory compactifications. This correspondence can be lifted
7147: to $D=4$ where the M-theory on the elliptically fibered fourfold descends to an F-theory compactification.
7148: In our analysis we neglected contributions due to singularities of
7149: the Calabi-Yau fourfold. Smoothed out they yield additional moduli, which
7150: are identified with $D7$ or $O7$ moduli in the orientifold limit.
7151: In a next step one can attempt to include these into the analysis and
7152: later deform away from the orientifold limes. Non-trivial fibrations appear if
7153: the orientifold charges are not canceled locally and the F-theory picture becomes
7154: essential. Finally we also discussed the embedding of type IIA
7155: orientifolds into a specific class of
7156: $G_2$ compactification of M-theory. Neglecting the
7157: contributions arising from the singularities of the $G_2$ manifold
7158: we were able to show agreement between the low energy effective
7159: actions. Comparing the superpotentials we only discovered
7160: the terms which are due to four-form flux from in M-theory.
7161: However, relaxing the condition of $G_2$ holonomy we were able to
7162: identify one of the remaining terms as corresponding to a
7163: non-trivial fibration of a $G_2$ structure manifold. It remains to
7164: identify the counterpart of the orientifold
7165: superpotential term cubic in the complexified K\"ahler moduli. This
7166: term is propotional to the mass parameter of massive IIA supergravity
7167: and plays the essential role in moduli stabilization.
7168:
7169: Let us end our conclusions with some directions for further research.
7170: Firstly, it would be desirable to include D-brane matter fields into
7171: the orientifold setups. For type IIB setups with $D3$ and $D7$ branes this was done, for example,
7172: in refs. \cite{GGJL,JL}. An important task is to extend these results to
7173: type IIA orientifolds with space-time filling $D6$ branes. The knowledge
7174: of the full effective action enables to perform a calculation of soft supersymmetry
7175: breaking terms of semi-realistic D-brane scenarios.
7176:
7177: As already mentioned, a generalization to non-Calabi-Yau orientifolds is of
7178: particular interest \cite{GLprep}. Orientifolds allow for consistent $D=4$
7179: Minkowski or Anti-de Sitter vacua for which the internal manifold possesses
7180: non-trivial torsion. As we have argued, the orientifold projections specify
7181: a K\"ahler submanifold in the quaternionic $N=2$ moduli space with geometry encoded
7182: by special even and odd forms. The K\"ahler potential is Hitchins functional truncated
7183: by the projection. A similar analysis is likely to apply to orientifolds of
7184: generalized complex manifolds as introduced in \cite{HitchinGCM}.
7185:
7186: Brane worlds in orientifolds are a prominent arena for model building in
7187: particle physics and cosmology. However, finding a particular vacuum featuring the
7188: properties of our universe is a highly non-trivial task. One major step into this
7189: direction is to extract vacua with stabilized moduli fields. Assuming that this
7190: can be achieved, for example by background fluxes, one encounters a huge set of possible vacua
7191: labeled by different flux quantum numbers.
7192: In the pioneering paper \cite{Douglas} it was argued that a statistical analysis of this `landscape'
7193: could lead a deeper understanding of the vacuum structure of string theory.
7194: These considerations were mostly applied to type IIB orientifolds and
7195: certain M-theory vacua. It is an interesting task to generalize this to type
7196: IIA orientifolds. For early time cosmology a wave-function for flux vacua could
7197: yield an interesting attempt to approach quantum cosmological questions within
7198: the framework of string theory \cite{OVV}. It would be nice to relate these
7199: new developments in topological string theories to the results of $N=1$
7200: flux compactifications. Surprisingly various similarities appear, which hint
7201: to at least a formal relation.
7202:
7203:
7204:
7205:
7206: \chapter{Appendix}
7207:
7208: \appendix
7209:
7210:
7211:
7212: %\def\theequation{\normalsize \Alph{section}.\arabic{equation}}
7213: \renewcommand{\thesection}{\Alph{section}}
7214: \renewcommand{\theequation}{\Alph{section}.\arabic{equation}}
7215: % \renewcommand{\thechapter}{\}
7216: %\appendix
7217: %\chapter{Appendix}
7218: \section{Conventions}\label{conventions}
7219: In this appendix we summarize our conventions.
7220:
7221: \begin{itemize}
7222: \item
7223: The coordinates of the four-dimensional Minkowski space-time are
7224: denoted by $x^\mu, \mu=0,\ldots,3$.
7225: The corresponding metric is chosen to have signature $(-,+,+,+)$.
7226: The coordinates of the compact Calabi-Yau manifold $Y$
7227: are denoted by $y^i, \bar y^\bi,\ i,\bi=1,2,3$.
7228:
7229: \item
7230: $p$-forms are expanded into a real basis according to
7231: \beq
7232: A_p\ =\ \frac{1}{p!}\,
7233: A_{\mu_1 \ldots \mu_p} dx^{\mu_1}\wedge \ldots \wedge dx^{\mu_p}\ .
7234: \eeq
7235: \item
7236: $(p,q)$-forms are expanded into a complex basis as
7237: \beq
7238: A_{p,q} = \frac{1}{p!q!} A_{i_1 \ldots i_p \bi_1 \ldots \bi_q} dy^{i_1}\wedge \ldots \wedge dy^{i_p}
7239: \wedge d\bar y^{\bi_1}\wedge \ldots \wedge d\bar y^{\bi_q}\ .
7240: \eeq
7241: \item
7242: The exterior derivative is defined as
7243: \beq
7244: dA_p=\frac{1}{p!} \partial_\mu A_{\mu_1 \ldots \mu_p} dx^\mu\wedge dx^{\mu_1}\wedge \ldots \wedge dx^{\mu_p}\ .
7245: \eeq
7246: \item
7247: The field strength of a $p$-form $F_{p+1}=dA_p$
7248: is given by
7249: \beq
7250: F_{\mu_1 \ldots \mu_{p+1}} = (p+1)\, \partial_{[\mu_1}A_{\mu_2\ldots \mu_{p+1}]}\ .
7251: \eeq
7252: \item
7253: The inner product for real forms is defined by using the Hodge-$*$ operator. In
7254: components we have
7255: \beq \label{wedge*comp}
7256: \int F_p \wedge * F_p = \frac{1}{p!}\int F_{\mu_1 \ldots \mu_p} F^{\mu_1 \ldots \mu_p} *\mathbf{1}\ ,
7257: \eeq
7258: where $*\mathbf{1} = d^d x\, \sqrt{-g}$ is the $d$-dimensional measure.
7259:
7260: \item
7261: The Hodge-$*$ satisfies
7262: $** F_p = (-1)^{p(d-p)+\kappa} F_p$, where $\kappa=1$ for Lorentzian signature
7263: and $\kappa=0$ for Euclidean signature.
7264:
7265: \item
7266: Let $\sigma_1$ and $\sigma_2$ be an orientiation preserving and an orientation reversing map
7267: $\sigma_{1,2}: M \rightarrow M$, where $M$ is an $n$-dimensional manifold. Then one finds
7268: for a $n$-form $\omega$ on $M$ that
7269: \beq \label{int-form1}
7270: \int_{\sigma_1(M)} \omega = \int_M \sigma_1^*(\omega)\ , \qquad \int_{\sigma_2(M)} \omega = -\int_M \sigma_2^*( \omega)\ .
7271: \eeq
7272: However, if we choose $\omega_{M}=*\mathbf{1}$ to be the canonical volume form of $M$ then
7273: $\omega_{\sigma_1(M)}= \sigma_1^* (\omega_{M})$ and $\omega_{\sigma_2(M)}= -\sigma_2^*( \omega_{M})$,
7274: such that
7275: \beq \label{int-form2}
7276: \int_{\sigma_{1,2}(M)} \omega_{\sigma_{1,2}(M)} = \int_M \sigma_{1,2}^*(\omega_M)\ .
7277: \eeq
7278:
7279: \end{itemize}
7280:
7281:
7282:
7283:
7284: \section{N=2 supergravity and special geometry}
7285: \label{specialGeom}
7286:
7287: In this appendix we briefly summarize the $N=2$ special geometry
7288: of the Calabi-Yau
7289: moduli space. A more detailed discussion can be found,
7290: for example, in refs.\ \cite{CdO,Strominger2,Freed,N=2review,CRTV}.
7291: A special K\"ahler manifold $\cM$ is a Hodge-K\"ahler manifold (with line bundle $\cL$)
7292: of real dimension $2n$ with associated
7293: holomorphic flat $Sp(2n+2,\mathbb{R})$ vector bundle $\mathcal{H}$ over $\cM$. Furthermore
7294: there exists a holomorphic section $\Omega(z)$ of $\cL$ such that
7295: \beq\label{N=2KP}
7296: K(z,\bar z) = - \ln i \big<\Omega(z) , \bar \Omega(\bar z) \big>\ , \qquad
7297: \big<\Omega, \partial_{z^K} \Omega\big> = 0\ , \qquad K=1,\ldots n\ ,
7298: \eeq
7299: where $K$ is the K\"ahler potential of $\cM$ and
7300: $\big<\cdot,\cdot \big>$ is the
7301: symplectic product on the fibers.
7302: This is precisely what one encounters in the moduli space
7303: of the complex structure deformations of a Calabi-Yau manifold
7304: with $\Omega$ being the holomorphic three-form.
7305: In this case
7306: one is lead to set $n=h^{(2,1)}$ and identify
7307: the fibers of the associated
7308: $Sp$-bundle with $H^3(Y,\mathbb{C})$. The symplectic product is given by
7309: the intersections on $H^3(Y,\bbC)$ as
7310: \beq \label{sympl-f}
7311: \big<\alpha, \beta \big> = \int_Y \alpha \wedge \beta\ .
7312: \eeq
7313: The K\"ahler covariant derivatives of $\Omega$ are denoted by
7314: $\chi_K$ as explicitly given in \eqref{Kod-form}.
7315: In terms of the symplectic basis $(\alpha_\Kh, \beta^\Kh)$
7316: introduced in \eqref{int-numbers1} both
7317: $\Omega$ and $\chi_K$ enjoy the expansion
7318: \beq \label{def-chi}
7319: \Omega = Z^\Kh\, \alpha_\Kh - \cF_\Kh\, \beta^\Kh\ , \qquad
7320: \chi_K = \chi^\Lh_{K}\, \alpha_\Lh - \chi_{\Lh|K}\, \beta^\Lh\ .
7321: \eeq
7322: %Correspondingly $\Omega(z)$ transforms as given in
7323: %\eqref{crescale} under K\"ahler transformations. We also introduce
7324: %a section $\chi_K$ of $\cH\times \cL$ as
7325: %\beq
7326: % \chi_K = \partial_{z^K} \Omega + K_{z^K} \Omega\ ,
7327: %\eeq
7328: %such that $\chi_K$ is the covariant derivative of $\Omega$.
7329: %On the fibers of $\cH$ one
7330: %can choose a real symplectic basis $(\alpha_\Kh,\beta^\Kh)$ satisfying
7331: %\beq
7332: % \big< \alpha_\Kh, \beta^\Lh \big> =\delta^\Lh_\Kh\ , \quad
7333: % \big< \alpha_\Kh, \alpha_\Lh \big> =
7334: % \big< \beta^\Kh , \beta^\Lh \big> =0\ , \qquad \Kh,\Lh=0,\ldots,n\ .
7335: %\eeq
7336: %In this basis the sections $\Omega(z)$ and $\chi(z,\bar z)$ admit the expansion
7337: The holomorphic functions $Z^\Kh(z)$ and $\cF_\Kh(z)$
7338: are called the periods of $\Omega$, while $\chi^\Lh_{K}(z,\bar z)$ and $\chi_{\Lh|K}(z,\bar z)$
7339: are the periods of $\chi_K$. In terms of $Z^\Kh,\cF_\Kh$ the K\"ahler potential
7340: \eqref{N=2KP} can be rewritten as in \eqref{csmetric}.
7341:
7342: For every special K\"ahler manifold there exists
7343: a complex matrix $\cM_{\Kh \Lh}(z,\bar z)$ defined as
7344: \beq \label{def-M}
7345: \cM_{\Kh \Lh} = (\bar \chi_{\Kh|\bar M}\ \ \cF_\Kh) (\bar \chi^\Lh_{\bar M}\ \ Z^\Lh)^{-1}\ ,
7346: \eeq
7347: where $\chi^{\Lh}_{K}$ and $\chi_{\Lh| K}$ are given in \eqref{def-chi}.
7348: Furthermore, one extracts from \eqref{def-M} the
7349: identities
7350: \bea \label{ML-hf}
7351: \cF_{\Kh} = \cM_{\Kh \Lh} Z^\Lh\ , \qquad \chi_{\Lh| K} = \bar \cM_{\Lh \Mh} \chi^{\Mh}_{K}\ ,
7352: \eea
7353: which can be used to rewrite \eqref{N=2KP} as
7354: \bea \label{spconst}
7355: G_{M \bar N} & = & -2 e^{K} \chi^\Kh_M\, \I \cM_{\Kh \Lh}\, \bar \chi^\Lh_{\bar N}\ , \qquad
7356: 1 \ = \ -2 e^{K} Z^\Kh\, \text{Im}\, \cM_{\Kh \Lh}\,\bar Z^\Lh \ , \\
7357: 0 & = & -2 \bar \chi^\Kh_{\bar M}\, \text{Im}\, \cM_{\Kh \Lh}\,\bar Z^\Lh \ .\nn
7358: \eea
7359:
7360: If one assumes that
7361: the Jacobian matrix $\partial_{z^L}\big(Z^K/Z^0 \big)$ is invertible
7362: $\cF_\Kh$ is the derivative of a holomorphic prepotential $\cF$ with respect to the periods $Z^\Kh$.
7363: It is homogeneous of degree two and obeys
7364: \beq
7365: \cF = \tfrac{1}{2}Z^\Kh \cF_{\Kh}\ , \qquad \cF_\Kh =\partial_{Z^\Kh} \cF\ , \qquad
7366: \cF_{\Kh \Lh} =\partial_{Z^\Kh} \cF_{\Lh}\ ,
7367: \qquad \cF_{\Lh}= Z^\Kh \cF_{\Kh \Lh}\ ,
7368: \eeq
7369: which implies that $\cF_{\Kh \Lh}(Z)$ is invariant
7370: under rescalings of $Z^\Kh$.
7371: Notice that $\cF$ is only invariant under a restricted
7372: class of symplectic transformations
7373: and thus depends on the choice of symplectic basis.
7374:
7375: The complex matrix
7376: $\cM_{\Kh \Lh}$ defined in \eqref{def-M} can be rewritten in terms of the
7377: periods $Z^\Kh$ and the matrix $\cF_{\Kh\Lh}(Z)$ as
7378: \bea \label{gauge-c}
7379: \cM_{\Kh \Lh}=\overline{ \mathcal{F}}_{\Kh \Lh}+2i \frac{(\text{Im}\; \mathcal{F})_{\Kh \Mh} Z^\Mh
7380: (\text{Im}\; \mathcal{F})_{\Lh \Nh}Z^\Nh }{Z^\Nh(\text{Im}\; \mathcal{F})_{\Nh\Mh}
7381: Z^\Mh}\ .
7382: \eea
7383:
7384:
7385:
7386: Whenever the
7387: Jacobian matrix $\partial_{z^L}\big(Z^K/Z^0 \big)$ is invertible
7388: the $Z^\Kh$ can be viewed as projective coordinates of $\mathbb{P}_{h^{(2,1)}+1}$.
7389: Going to a special gauge, i.e.~fixing the K\"ahler transformations
7390: \eqref{crescale}, one introduces
7391: special coordinates $z^K$ by setting $z^K=Z^K/Z^0$.
7392: Due to the homogeneity of $\cF$ it is possible to define
7393: a holomorphic prepotential $f(z)$ which only depends on the special
7394: coordinates as
7395: \beq\label{def-f}
7396: \cF(Z) = (Z^0)^2 f(z)\ .
7397: \eeq
7398: In terms of $f$ the K\"ahler potential given in \eqref{N=2KP}
7399: reads
7400: \beq \label{Kinz}
7401: K\ =\ - \ln i|Z^0|^2 \big[2(f-\bar f)-(\partial_K\, f + \partial_{\bar K} \bar f)(z^K - \bar z^K) \big]\ .
7402: \eeq
7403:
7404: A special example of the situation just discussed is the moduli
7405: space spanned by the complexified K\"ahler deformations $t^A$ introduced
7406: in \eqref{4d-dilaton}. These fields can be interpreted as special coordinates on
7407: a special K\"ahler manifold $\cM^{\rm SK}(t,\bar t)$ \cite{CdO}.
7408: The K\"ahler potential of the metric $G_{AB}$ given
7409: in \eqref{Kmetric} is of the form \eqref{Kinz} with
7410: \beq \label{pre-K}
7411: f(t)=-\tfrac{1}6 \cK_{ABC} t^A t^B t^C\ .
7412: \eeq
7413: Furthermore, inserting \eqref{pre-K}
7414: into \eqref{gauge-c} using \eqref{def-f}
7415: one determine the gauge-couplings $\cN_{\Ah \Bh}(t,\bar t)$
7416: to be
7417: \bea \label{def-cN}
7418: \text{Re} \cN &=& \ \
7419: \left(\ba{cc}-\frac13 \cK_{ABC}b^A b^B b^C & \frac12 \cK_{ABC} b^B b^C \\
7420: \frac12 \cK_{ABC} b^B b^C & - \cK_{ABC}b^C \ea \right)\ , \nn \\
7421: \text{Im} \cN &=& -\frac{\cK}{6}
7422: \left(\ba{cc}1 + 4 G_{AB}b^A b^B & -4 G_{AB}b^B \\
7423: - 4 G_{AB}b^B & 4 G_{AB} \ea \right)\ , \nn \\
7424: (\text{Im} \cN)^{-1} &=& - \frac{6}{\cK}
7425: \left(\ba{cc}1 & b^A \\
7426: b^A & \frac14 G^{AB} + b^A b^B \ea \right)\ ,
7427: \eea
7428: where $G_{AB}$ is given in \eqref{Kmetric}.
7429:
7430:
7431: % In this appendix we briefly summarize $N=2$ supergravity in four space-time
7432: % dimensions. We consentrate on effective action for the bosonic fields and the geometry of
7433: % the corresponding $N=2$ moduli spaces. A more detailed disscussion also including possible gaugings
7434: % can be found, for example, in ref.~\cite{N=2review}.
7435:
7436: % The $N=2$ theories under consideration consist of a gravity multiplet with bosonic
7437: % components $(g_{\mu \nu},A^0)$, $n_v$ vector multiplets with bosonic components $(A^A,z^A)$
7438: % and $n_h$ hypermultiplets with bosonic components $(\tilde q^{u},\tilde q^{u+1},\tilde q^{u+2},\tilde q^{u+3})$.
7439: % The kinetic terms and couplings of these fields are encoded by the effective
7440: % action \cite{N=2review}
7441: % \bea \label{N=2act}
7442: % S^{(4)}_{ N=2} & = &\int-\ \tfrac12 R * \mathbf{1}\ -\ h_{uv}\, d\tilde q^u \wedge * d\tilde q^v \\
7443: % && - G_{K \bar L}\, dz^K \wedge * d\bar z^L + \tfrac14 \I \cM_{\Kh \Lh}\, F^{\Kh} \wedge *
7444: % F^{\Lh}
7445: % + \tfrac14 \R \cM_{\Kh \Lh}\, F^{\Kh} \wedge F^{\Lh}\ , \nn
7446: % \eea
7447: % where $A^\Kh=(A^0,A^K)$ and $F^{\Ah} = dA^{\Ah}$. The first line in \eqref{N=2act}
7448: % contains the standard Einstein-Hilbert term as well as the kinetic terms for
7449: % the hypermultiplets. $N=2$ supersymmetry demands $h_{uv}$ to be a quaternionic
7450: % metric such that the scalars $\tilde q^u$ span a quaternionic manifold $\cM^{\rm Q}_{4 n_h}$ of real
7451: % dimension $4n_h$ \cite{N=2review}. The second line in \eqref{N=2act} consists of
7452: % the kinetic terms for the vector multiplets. In $N=2$ supergravity the metric
7453: % $G_{K \bar L}$ has to be special K\"ahler implying that the scalars $z^K$ span
7454: % a special K\"ahler manifold $\cM^{\rm SK}_{2n_v}$ of real dimension $2n_v$ \cite{N=2review,CRTV,Freed}.
7455: % The complex gauge coupling matrix $\cM_{\Kh \Lh}(z,\bar z)$ is generically a non-trivial function
7456: % of $z^K,\bar z^K$, but by the constraints of $N=2$ supersymmetry independent of the
7457: % hypermultiplet scalars $\tilde q^u$. As we will disscuss momentarily $\cM_{\Kh \Lh}$ and
7458: % the metric $G_{K \hat L}$ are not independent, but rather linked in an intriquing
7459: % way. In summary the $N=2$ moduli space admits the local product form
7460: % $\cM^{\rm Q}_{4 n_h} \times \cM^{\rm SK}_{2n_v}$.
7461: % In the following we will concentrate on the special K\"ahler manifold
7462: % $\cM^{\rm SK}_{2n_v}$ obtained in Calabi-Yau compactifications of
7463: % type II supergravity theories. We refer the reader to \cite{N=2review} for a
7464: % detailed discussion of the quaternionic space $\cM^{\rm Q}_{4 n_h}$.
7465:
7466: % In compactifications of type IIA or type IIB supergravity on
7467: % Calabi-Yau manifolds the moduli space $\cM^{\rm SK}$
7468: % is spanned by the metric deformations of the Calabi-Yau metric \cite{CdO,Strominger2,CRTV}.
7469: % To make this more precise, let us summarize some
7470: % facts about special geometry.
7471: % %More precely, the manifold $\cM^{\rm SK}$ consists
7472: % %of the $n_v = h^{2,1}(Y)$ complex structure deformations $z^K$ in type IIB (c.f.~section \ref{revIIB}),
7473: % %while it consists of the $n_v = h^{1,1}(Y)$
7474: % %complexified K\"ahler deformations $t^A$ in type IIA (c.f.~section \ref{revIIA}).
7475: % %
7476: % %In this appendix we briefly summarize the $N=2$ special geometry
7477: % %of the Calabi-Yau
7478: % %moduli space. A more detailed discussion can be found,
7479: % %for example, in refs.\ \cite{CdO,Strominger2,Freed,N=2review,CRTV}.
7480: % A special K\"ahler manifold $\cM^{\rm SK}$ is a Hodge-K\"ahler manifold (with line bundle $\cL$)
7481: % of real dimension $2n$ with associated
7482: % holomorphic flat $Sp(2n+2,\mathbb{R})$ vector bundle $\mathcal{H}$ over $\cM^{\rm SK}$. Furthermore
7483: % there exists a holomorphic section $\Omega(z)$ of $\cL$ such that
7484: % \beq\label{N=2KP}
7485: % K(z,\bar z) = - \ln i \big<\Omega(z) , \bar \Omega(\bar z) \big>\ , \qquad
7486: % \big<\Omega, \partial_{z^K} \Omega\big> = 0\ , \qquad K=1,\ldots n\ ,
7487: % \eeq
7488: % where $K$ is the K\"ahler potential of $\cM^{\rm SK}$ and
7489: % $\big<\cdot,\cdot \big>$ is the
7490: % symplectic product on the fibers.
7491: % This is precisely what one encounters in the moduli space
7492: % of the complex structure deformations of a Calabi-Yau manifold
7493: % with $\Omega$ being the holomorphic three-form.
7494: % In this case
7495: % one is lead to set $n=h^{(2,1)}$ and identify
7496: % the fibers of the associated
7497: % $Sp$-bundle with $H^3(Y,\mathbb{C})$. The symplectic product is given by
7498: % the intersections on $H^3(Y,\bbC)$ as
7499: % \beq \label{sympl-f}
7500: % \big<\alpha, \beta \big> = \int_Y \alpha \wedge \beta\ .
7501: % \eeq
7502: % The K\"ahler covariant derivatives of $\Omega$ are denoted by
7503: % $\chi_K$ as explicitly given in \eqref{Kod-form}.
7504: % In terms of the symplectic basis $(\alpha_\Kh, \beta^\Kh)$
7505: % introduced in \eqref{int-numbers1} both
7506: % $\Omega$ and $\chi_K$ enjoy the expansion
7507: % \beq \label{def-chi}
7508: % \Omega = Z^\Kh\, \alpha_\Kh - \cF_\Kh\, \beta^\Kh\ , \qquad
7509: % \chi_K = \chi^\Lh_{K}\, \alpha_\Lh - \chi_{\Lh|K}\, \beta^\Lh\ .
7510: % \eeq
7511: % %Correspondingly $\Omega(z)$ transforms as given in
7512: % %\eqref{crescale} under K\"ahler transformations. We also introduce
7513: % %a section $\chi_K$ of $\cH\times \cL$ as
7514: % %\beq
7515: % % \chi_K = \partial_{z^K} \Omega + K_{z^K} \Omega\ ,
7516: % %\eeq
7517: % %such that $\chi_K$ is the covariant derivative of $\Omega$.
7518: % %On the fibers of $\cH$ one
7519: % %can choose a real symplectic basis $(\alpha_\Kh,\beta^\Kh)$ satisfying
7520: % %\beq
7521: % % \big< \alpha_\Kh, \beta^\Lh \big> =\delta^\Lh_\Kh\ , \quad
7522: % % \big< \alpha_\Kh, \alpha_\Lh \big> =
7523: % % \big< \beta^\Kh , \beta^\Lh \big> =0\ , \qquad \Kh,\Lh=0,\ldots,n\ .
7524: % %\eeq
7525: % %In this basis the sections $\Omega(z)$ and $\chi(z,\bar z)$ admit the expansion
7526: % The holomorphic functions $Z^\Kh(z)$ and $\cF_\Kh(z)$
7527: % are called the periods of $\Omega$, while $\chi^\Lh_{K}(z,\bar z)$ and $\chi_{\Lh|K}(z,\bar z)$
7528: % are the periods of $\chi_K$. In terms of $Z^\Kh,\cF_\Kh$ the K\"ahler potential
7529: % \eqref{N=2KP} can be rewritten as in \eqref{csmetric}.
7530:
7531: % For every special K\"ahler manifold there exists
7532: % a complex matrix $\cM_{\Kh \Lh}(z,\bar z)$ defined as
7533: % \beq \label{def-M}
7534: % \cM_{\Kh \Lh} = (\bar \chi_{\Kh|\bar M}\ \ \cF_\Kh) (\bar \chi^\Lh_{\bar M}\ \ Z^\Lh)^{-1}\ ,
7535: % \eeq
7536: % where $\chi^{\Lh}_{K}$ and $\chi_{\Lh| K}$ are given in \eqref{def-chi}.
7537: % It is precisely this matrix, which encondes the gauge-couplings of the vector
7538: % fields in the effective action \eqref{N=2act}.
7539: % Furthermore, one extracts from \eqref{def-M} the
7540: % identities
7541: % \bea \label{ML-hf}
7542: % \cF_{\Kh} = \cM_{\Kh \Lh} Z^\Lh\ , \qquad \chi_{\Lh| K} = \bar \cM_{\Lh \Mh} \chi^{\Mh}_{K}\ ,
7543: % \eea
7544: % which can be used to rewrite \eqref{N=2KP} as
7545: % \bea \label{spconst}
7546: % G_{M \bar N} & = & -2 e^{K} \chi^\Kh_M\, \I \cM_{\Kh \Lh}\, \bar \chi^\Lh_{\bar N}\ , \qquad
7547: % 1 \ = \ -2 e^{K} Z^\Kh\, \text{Im}\, \cM_{\Kh \Lh}\,\bar Z^\Lh \ , \\
7548: % 0 & = & -2 \bar \chi^\Kh_{\bar M}\, \text{Im}\, \cM_{\Kh \Lh}\,\bar Z^\Lh \ .\nn
7549: % \eea
7550:
7551: % If one assumes that
7552: % the Jacobian matrix $\partial_{z^L}\big(Z^K/Z^0 \big)$ is invertible
7553: % $\cF_\Kh$ is the derivative of a holomorphic prepotential $\cF$ with respect to the periods $Z^\Kh$.
7554: % It is homogeneous of degree two and obeys
7555: % \beq
7556: % \cF = \tfrac{1}{2}Z^\Kh \cF_{\Kh}\ , \qquad \cF_\Kh =\partial_{Z^\Kh} \cF\ , \qquad
7557: % \cF_{\Kh \Lh} =\partial_{Z^\Kh} \cF_{\Lh}\ ,
7558: % \qquad \cF_{\Lh}= Z^\Kh \cF_{\Kh \Lh}\ ,
7559: % \eeq
7560: % which implies that $\cF_{\Kh \Lh}(Z)$ is invariant
7561: % under rescalings of $Z^\Kh$.
7562: % Notice that $\cF$ is only invariant under a restricted
7563: % class of symplectic transformations
7564: % and thus depends on the choice of symplectic basis.
7565:
7566: % The complex matrix
7567: % $\cM_{\Kh \Lh}$ defined in \eqref{def-M} can be rewritten in terms of the
7568: % periods $Z^\Kh$ and the matrix $\cF_{\Kh\Lh}(Z)$ as
7569: % \bea \label{gauge-c}
7570: % \cM_{\Kh \Lh}=\overline{ \mathcal{F}}_{\Kh \Lh}+2i \frac{(\text{Im}\; \mathcal{F})_{\Kh \Mh} Z^\Mh
7571: % (\text{Im}\; \mathcal{F})_{\Lh \Nh}Z^\Nh }{Z^\Nh(\text{Im}\; \mathcal{F})_{\Nh\Mh}
7572: % Z^\Mh}\ .
7573: % \eea
7574:
7575:
7576:
7577: % Whenever the
7578: % Jacobian matrix $\partial_{z^L}\big(Z^K/Z^0 \big)$ is invertible
7579: % the $Z^\Kh$ can be viewed as projective coordinates of $\mathbb{P}_{h^{(2,1)}+1}$.
7580: % Going to a special gauge, i.e.~fixing the K\"ahler transformations
7581: % \eqref{crescale}, one introduces
7582: % special coordinates $z^K$ by setting $z^K=Z^K/Z^0$.
7583: % Due to the homogeneity of $\cF$ it is possible to define
7584: % a holomorphic prepotential $f(z)$ which only depends on the special
7585: % coordinates as
7586: % \beq\label{def-f}
7587: % \cF(Z) = (Z^0)^2 f(z)\ .
7588: % \eeq
7589: % In terms of $f$ the K\"ahler potential given in \eqref{N=2KP}
7590: % reads
7591: % \beq \label{Kinz}
7592: % K\ =\ - \ln i|Z^0|^2 \big[2(f-\bar f)-(\partial_K\, f + \partial_{\bar K} \bar f)(z^K - \bar z^K) \big]\ .
7593: % \eeq
7594:
7595: % A second more specific example of a situation just discussed is the moduli
7596: % space spanned by the complexified K\"ahler deformations $t^A$ introduced
7597: % in \eqref{4d-dilaton}. These fields can be interpreted as special coordinates on
7598: % a special K\"ahler manifold $\cM^{\rm SK}(t,\bar t)$ \cite{CdO}.
7599: % The K\"ahler potential of the metric $G_{AB}$ given
7600: % in \eqref{Kmetric} is of the form \eqref{Kinz} with
7601: % \beq \label{pre-K}
7602: % f(t)=-\tfrac{1}6 \cK_{ABC} t^A t^B t^C\ .
7603: % \eeq
7604: % Furthermore, inserting \eqref{pre-K}
7605: % into \eqref{gauge-c} using \eqref{def-f}
7606: % one determine the gauge-couplings $\cN_{\Ah \Bh}(t,\bar t)$
7607: % to be
7608: % \bea \label{def-cN}
7609: % \text{Re} \cN &=& \ \
7610: % \left(\ba{cc}-\frac13 \cK_{ABC}b^A b^B b^C & \frac12 \cK_{ABC} b^B b^C \\
7611: % \frac12 \cK_{ABC} b^B b^C & - \cK_{ABC}b^C \ea \right)\ , \nn \\
7612: % \text{Im} \cN &=& -\frac{\cK}{6}
7613: % \left(\ba{cc}1 + 4 G_{AB}b^A b^B & -4 G_{AB}b^B \\
7614: % - 4 G_{AB}b^B & 4 G_{AB} \ea \right)\ , \nn \\
7615: % (\text{Im} \cN)^{-1} &=& - \frac{6}{\cK}
7616: % \left(\ba{cc}1 & b^A \\
7617: % b^A & \frac14 G^{AB} + b^A b^B \ea \right)\ ,
7618: % \eea
7619: % where $G_{AB}$ is given in \eqref{Kmetric}.
7620:
7621:
7622: % % In the appendix \ref{CofK} we will also need another formulation
7623: % % of special geometry. Instead of introducing a line bundle $\cL$
7624: % % with holomorphic section $\Omega(z)$ transforming as in
7625: % % \eqref{crescale} we now switch to the associated $U(1)$ principal bundle $\mathcal{P}$.
7626: % % A section $\varphi(z,\bar z)$ of this bundle transforms under K\"ahler transformations as
7627: % % $\varphi \rightarrow e^{-i \I f(z)} \varphi$ and we define the connection as
7628: % % $(\partial_K + \tfrac12 \partial_K K) \varphi$ and
7629: % % $(\partial_{\bar K} - \tfrac12 \partial_{\bar K} K) \varphi$.
7630: % % Thus the to $\Omega \in \cL$ corresponding covariantly holomorphic
7631: % % section $V \in \mathcal{P}$ is given by
7632: % % \bea \label{def-V}
7633: % % V = e^{K/2} \Omega\ , \qquad V(z,\bar z) = \cL^\Kh(z,\bar z) \alpha_\Kh - \cM_\Kh(z,\bar z) \beta^\Kh\ .
7634: % % \eea
7635: % % Its covariant derivative we denote by
7636: % % \bea \label{fh-def}
7637: % % U_K \equiv (\partial_K + \tfrac12 \partial_K K) V(z,\bar z) = f^\Lh_{K} \alpha_\Lh - h_{\Lh|K} \beta^\Lh\ .
7638: % % \eea
7639: % % Comparing \eqref{fh-def} with equation \eqref{Kod-form} one identifies
7640: % % $U_{K} = e^{\frac{K}{2}} \chi_K(z)$. In terms of $U_K$ and $V$ the
7641: % % special K\"ahler conditions \eqref{N=2KP} and
7642: % % \eqref{chi_barchi} take the form
7643: % % \beq \label{fcontr}
7644: % % \big<U_K,\bar U_{L}\big> = -iG_{K \bar L}\ , \quad
7645: % % \big<V,\bar V\big> =- i\ , \quad \big<V,U_K\big> = 0\ ,\quad \big<U_K,U_L\big> = 0\ .
7646: % % \eeq
7647: % % One can now define a complex matrix $\cM$ as
7648: % % \beq \label{gauge-cL}
7649: % % \cM_{\Kh \Lh} = (\bar h_{\Kh|\bar M}\ \ \cM_\Kh) (\bar f^\Lh_{\bar M}\ \ \cL^\Lh)^{-1}\ ,
7650: % % \eeq
7651: % % and show that it transforms under $Sp(2n+2,\mathbb{R})$ in the appropriate way to serve as a
7652: % % gauge coupling matrix. This translates into equation \eqref{gauge-c}, when we assume the existence
7653: % % of a prepotential. Moreover, it follows from \eqref{gauge-cL} that
7654: % % \bea \label{ML-hf}
7655: % % \cM_{\Kh} = \cM_{\Kh \Lh} \cL^\Lh\ , \quad h_{\Kh| M} = \bar \cM_{\Kh \Lh} f^{\Lh}_{M}\ .
7656: % % \eea
7657: % % These equations can be used to give the component expressions of \eqref{fcontr}
7658: % % \bea \label{spconst}
7659: % % G_{M \bar N} & = & -2 f^\Kh_M\, \text{Im}\, \cM_{\Kh \Lh}\, \bar f^\Lh_{\bar N}\ , \qquad
7660: % % 1 \ = \ -2 \cL^\Kh\, \text{Im}\, \cM_{\Kh \Lh}\,\bar \cL^\Lh \ , \\
7661: % % 0 & = & -2 \bar f^\Kh_{\bar M}\, \text{Im}\, \cM_{\Kh \Lh}\,\bar \cL^\Lh \ .\nn
7662: % % \eea
7663:
7664:
7665:
7666: \section{Supergravity with several linear multiplets} \label{linm}
7667:
7668: In this appendix we briefly discuss the dualization of several massless linear
7669: multiplets to chiral multiplets. We only discuss the bosonic component fields and do not include possible couplings to vector
7670: multiplets. Our aim is to extract the K\"ahler potential
7671: for the $N=1,D=4$ supergravity theory with all linear multiplets replaced by chiral ones.
7672: Let us begin by recalling the effective action for a set of
7673: linear multiplets $(L^\lambda, D^\lambda_2)$ couplet to chiral multiplets
7674: $N^k$. It takes the form\footnote{This action can be obtained by a straight forward
7675: generalization of the action for one linear multiplet given in \cite{BGG}.}
7676: \bea\label{kinetic}
7677: \cL &=& -\tfrac{1}{2}R*\mathbf{1} -
7678: \tilde K_{N^k\bar N^l}\, dN^k \wedge * d \bar N^{l}
7679: + \tfrac{1}{4} \tilde K_{L^\kappa L^\lambda}\,
7680: dL^\kappa \wedge * dL^\lambda \nn\\
7681: && + \tfrac{1}{4} \tilde K_{ L^\kappa L^\lambda}\, dD^\kappa_2 \wedge * dD^\lambda_2
7682: - \tfrac{i}2\, dD^\lambda_2 \wedge
7683: \big(\tilde K_{L^\lambda N^k}\,dN^k -\tilde K_{L^\lambda \bar N^k}\,d\bar N^k\big)
7684: \ ,
7685: \eea
7686: where $\tilde K(L,N,\bar N)$ is a function of the scalars $L^\lambda$ and
7687: the chiral multiplets $N^k$. The kinetic potential $\tilde K$ is the
7688: analog of
7689: the K\"ahler potential in the sense that it encodes the dynamics of the linear
7690: and chiral multiplets. In order to dualize the linear multiplets $(L^\lambda, D^\lambda_2)$
7691: into chiral multiplets $(L^\lambda,\tilde \xi_\lambda)$ one replaces
7692: $dD_2^\lambda$ by the form $D_3^\lambda$ and adds the term
7693: \beq
7694: \cL \to \cL + \delta \cL\ , \qquad
7695: \delta \cL\ =\
7696: - 2\tilde \xi_\lambda\, dD^\lambda_3\ =\ - 2 D_3^\lambda \wedge d\tilde \xi_\lambda\ ,\
7697: \eeq
7698: where $\tilde \xi_\lambda(x)$ is a Lagrange multiplier. Eliminating $\tilde \xi_\lambda$
7699: one finds that $dD^\lambda_3=0$ such that locally $D_3^\lambda=dD_2^\lambda$ as required.
7700: Alternatively one can consistently eliminate $D^\lambda_3$ by inserting
7701: its equations of motion
7702: \beq
7703: *D_3^\kappa = 4 \tilde K^{L^\kappa L^\lambda}\Big(d\tilde \xi_\lambda + \tfrac{i}4
7704: \big(\tilde K_{L^\lambda N^k}\,dN^k -\tilde K_{L^\lambda \bar N^k}\,d\bar N^k\big)\Big)
7705: \eeq
7706: back into the Lagrangian \eqref{kinetic}.
7707: The resulting dual Lagrangian takes the form
7708: \bea \label{eff_act1}
7709: \cL &=& -\tfrac{1}{2}R*\mathbf{1} -
7710: \tilde K_{N^k\bar N^l}\, dN^k \wedge * d \bar N^{l}
7711: + \tfrac{1}{4} \tilde K_{L^\kappa L^\lambda}\,
7712: dL^\kappa \wedge * dL^\lambda \\
7713: && + 4 \tilde K^{L^\kappa L^\lambda} \Big(d\tilde \xi_\kappa - \tfrac{1}2
7714: \I \big(\tilde K_{L^\kappa N^l}\,dN^l\big)\Big)\wedge *
7715: \Big(d\tilde \xi_\lambda - \tfrac{1}2
7716: \I \big(\tilde K_{L^\lambda N^k}\,dN^k\big)\Big) \ .\nn
7717: \eea
7718: Since we intend to use these results in the effective action for Calabi-Yau
7719: orientifolds, we make a further simplification. We demand that the kinetic potential
7720: $\tilde K$ is only a function of $L^\lambda$ and the imaginary part of $N^k$, which we
7721: denote by $l^k=\I N^k$. This implies that all chiral fields $N^k$ admit a Peccei-Quinn
7722: shift symmetry acting on the real parts of $N^k$ as it is indeed the case for the
7723: orientifold setups. Thus the effective Lagrangian \eqref{eff_act1} simplifies to
7724: \bea \label{linaction}
7725: \cL &=& -\tfrac{1}{2}R*\mathbf{1} -
7726: \tfrac{1}{4}\tilde K_{l^k l^l}\, dN^k \wedge * d \bar N^l
7727: + \tfrac{1}{4} \tilde K_{L^\kappa L^\lambda}\,
7728: dL^\kappa \wedge * dL^\lambda \\
7729: && + 4 \tilde K^{L^\kappa L^\lambda}
7730: \Big(d\tilde \xi_\kappa + \tfrac{1}{4}\tilde K_{L^\kappa l^l}\, d\, \text{Re}N^l\Big)\wedge *
7731: \Big(d\tilde \xi_\lambda + \tfrac{1}{4}\tilde K_{L^\lambda l^k}\, d\, \text{Re}N^k\Big) \ .\nn
7732: \eea
7733: This $N=1$ Lagrangian is written completely in terms of chiral multiplets and therefore
7734: can be derived from a K\"ahler potential when choosing appropriate complex coordinates
7735: $N^k$ and $T_\lambda=(L^\lambda, \tilde \xi_\lambda)$.
7736: As we will see in a moment, a direct calculation yields that this K\"ahler potential
7737: is the Legendre transform of $\tilde K$ with respect to the scalars $L^\kappa$.
7738: It takes the
7739: form
7740: \beq \label{LegKP}
7741: K(T,N) = \tilde K(L, N - \bar N) - 2 (T_\kappa +\bar T_\kappa) L^\kappa
7742: \eeq
7743: where $L^\kappa(N,T)$ is a function of the complex fields $N^k,T_\lambda$. This
7744: dependence is implicitly given via the definition of the coordinates $T_\lambda$
7745: \beq\label{defT}
7746: T_\lambda = i\tilde \xi_\lambda + \tfrac{1}{4}\tilde K_{L^\lambda}\ .
7747: \eeq
7748: However, in order to calculate the K\"ahler metric, one only needs to determine
7749: the derivatives of $L^\kappa(N,T)$ with respect to
7750: $N^k,T_\lambda$. They are obtained by differentiating \eqref{defT} and simply read
7751: \beq \label{derL}
7752: {\partial L^\kappa}/{\partial T_\lambda} = 2 \tilde K^{L^\kappa L^\lambda}\ , \qquad
7753: {\partial L^\kappa}/{\partial N^l} = - \tfrac{1}{2i} \tilde K^{L^\kappa L^\lambda} \tilde K_{L^\lambda l^l}\ .
7754: \eeq
7755: Using these identities one easily calculates the first derivatives of the K\"ahler
7756: potential \eqref{LegKP} as
7757: \beq \label{Kder}
7758: K_{T_\alpha} = -2 L^\alpha\ , \qquad K_{N^A} = \tfrac{1}{2i} \tilde K_{l^A}\ .
7759: \eeq
7760: Applying the equations \eqref{derL} once more when differentiating \eqref{Kder}
7761: one finds the K\"ahler metric
7762: \bea \label{Km1}
7763: K_{T_\alpha \bar T_\beta} &=& -4 \tilde K^{L^\alpha L^\beta}\ , \quad
7764: K_{T_\alpha \bar N^A}\ =\ i \tilde K^{L^\alpha L^\beta} \tilde K_{L^\beta l^A}\ , \nn \\
7765: K_{N^A \bar N^B} &=& \tfrac{1}{4} \tilde K_{l^A l^B} - \tfrac{1}{4}
7766: \tilde K_{ l^A L^\alpha }\, \tilde K^{L^\alpha L^\beta}\, \tilde K_{L^\beta l^B}\ ,
7767: \eea
7768: with inverse
7769: \bea \label{invKm1}
7770: K^{T_\alpha \bar T_\beta} &=& - \tfrac{1}{4} \tilde K_{L^\alpha L^\beta}
7771: + \tfrac{1}{4} \tilde K_{ l^A L^\alpha }\, \tilde K^{l^A l^B} \, \tilde K_{L^\beta l^B}\ ,
7772: \nn \\
7773: K^{T_\alpha \bar N^B} & = & -i \tilde K^{l^A l^B}\, \tilde K_{ l^A L^\alpha }\ , \quad
7774: K^{N^A \bar N^B} \ = \ 4 \tilde K^{l^A l^B}\ .
7775: \eea
7776: Finally, one checks that $K(T,N)$ is indeed the K\"ahler potential for the chiral part of the
7777: Lagrangian \eqref{linaction}. This is done by plugging in the definition
7778: of $T_\kappa$ given in \eqref{defT} and the K\"ahler metric \eqref{Km1} into
7779: \beq
7780: \cL = -\tfrac{1}{2}R*\mathbf{1} - K_{M^I \bar M^J}\ dM^I \wedge * d\bar M^J\ ,
7781: \eeq
7782: where $M^I=(N^k, T_\lambda)$.
7783:
7784:
7785:
7786:
7787:
7788:
7789:
7790:
7791:
7792:
7793:
7794:
7795:
7796:
7797:
7798:
7799: \newpage
7800: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7801: \begin{thebibliography}{99}
7802:
7803: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
7804: % Introduction
7805: \bibitem{revSusy}
7806: For a review see, for example,
7807: H.~P.~Nilles, ``Supersymmetry, Supergravity And Particle Physics,''
7808: Phys.\ Rept.\ {\bf 110} (1984) 1;\\
7809: W.~Hollik, R.~R\"uckl and J.~Wess (eds.),
7810: ``Phenomenological aspects of supersymmetry'', Springer Lecture Notes, 1992;\\
7811: J.~A.~Bagger, ``Weak-scale supersymmetry: Theory and practice,'' arXiv:hep-ph/9604232; \\
7812: S.~P.~Martin, ``A supersymmetry primer,'' arXiv:hep-ph/9709356; \\
7813: J.~Louis, I.~Brunner and S.~J.~Huber,
7814: ``The supersymmetric standard model,'' arXiv:hep-ph/9811341,
7815: and references therein.
7816:
7817: \bibitem{LQG}
7818: A.~Ashtekar and J.~Lewandowski,
7819: ``Background independent quantum gravity: A status report,''
7820: Class.\ Quant.\ Grav.\ {\bf 21}, R53 (2004) [arXiv:gr-qc/0404018].
7821:
7822: \bibitem{GSWbook}
7823: M.~B.~Green, J.~H.~Schwarz and E.~Witten,
7824: ``Superstring Theory'', Vol. 1\& 2, Cambridge University Press,
7825: Cambridge, 1987.
7826:
7827: \bibitem{JPbook}
7828: J.~Polchinski,
7829: ``String Theory'', Vol. 1\& 2,
7830: Cambridge University Press, Cambridge, 1998.
7831:
7832: \bibitem{Zwiebach}
7833: B.~Zwiebach, ``A first course in string theory'',
7834: Cambridge University Press, Cambridge, 2004.
7835:
7836: \bibitem{Dual}
7837: For a review see, for example,
7838: P.~K.~Townsend,
7839: ``Four lectures on M-theory,''
7840: arXiv:hep-th/9612121;\\
7841: J.~H.~Schwarz,
7842: ``Lectures on superstring and M theory dualities,''
7843: Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 55B} (1997) 1
7844: [arXiv:hep-th/9607201]; \\
7845: A.~Sen,
7846: ``An introduction to non-perturbative string theory,''
7847: arXiv:hep-th/9802051;\\
7848: N.~A.~Obers and B.~Pioline,
7849: ``U-duality and M-theory,''
7850: Phys.\ Rept.\ {\bf 318} (1999) 113
7851: [arXiv:hep-th/9809039];\\
7852: E.~Kiritsis,
7853: ``Supersymmetry and duality in field theory and string theory,''
7854: arXiv:hep-ph/9911525;\\
7855: A.~Sen, ``An introduction to duality symmetries in string theory,''
7856: {\it Prepared for Les Houches Summer School: Session 76:
7857: Euro Summer School on Unity of Fundamental Physics:
7858: Gravity, Gauge Theory and Strings,
7859: Les Houches, France, 30 Jul - 31 Aug 2001},
7860: and references therein.
7861:
7862:
7863: \bibitem{KaluzaKlein}
7864: T.~Kaluza,
7865: ``On The Problem Of Unity In Physics,''
7866: Sitzungsber.\ Preuss.\ Akad.\ Wiss.\ Berlin (Math.\ Phys.\ ) {\bf 1921} (1921) 966;\\
7867: O.~Klein,
7868: ``Quantum Theory And Five-Dimensional Theory Of Relativity,''
7869: Z.\ Phys.\ {\bf 37}, 895 (1926)
7870: [Surveys High Energ.\ Phys.\ {\bf 5}, 241 (1986)].
7871:
7872:
7873: \bibitem{KK-review}
7874: For a review see, for example, M.~J.~Duff, B.~E.~W.~Nilsson and C.~N.~Pope,
7875: ``Kaluza-Klein Supergravity,''
7876: Phys.\ Rept.\ {\bf 130} (1986) 1;\\
7877: J.~M.~Overduin and P.~S.~Wesson,
7878: ``Kaluza-Klein gravity,''
7879: Phys.\ Rept.\ {\bf 283} (1997) 303
7880: [arXiv:gr-qc/9805018], and references therein.
7881:
7882: \bibitem{Greene}
7883: For a review see, for example, B.~R.~Greene,
7884: ``String theory on Calabi-Yau manifolds,''
7885: arXiv:hep-th/9702155, and references therein.
7886:
7887:
7888:
7889: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7890: % Flux
7891:
7892: \bibitem{Strominger1}
7893: A.~Strominger, ``Superstrings With Torsion,''
7894: Nucl.\ Phys.\ B {\bf 274}, 253 (1986).
7895:
7896: \bibitem{Bachas}
7897: C.~Bachas,
7898: ``A Way to break supersymmetry,''
7899: arXiv:hep-th/9503030;\\
7900: J.~Polchinski and A.~Strominger,
7901: ``New Vacua for Type II String Theory,''
7902: Phys.\ Lett.\ B {\bf 388} (1996) 736
7903: [arXiv:hep-th/9510227].
7904:
7905: \bibitem{BB1}
7906: K.~Becker and M.~Becker,
7907: ``M-Theory on Eight-Manifolds,''
7908: Nucl.\ Phys.\ B {\bf 477}, 155 (1996)
7909: [arXiv:hep-th/9605053].
7910:
7911:
7912: \bibitem{Michelson}
7913: J.~Michelson,
7914: ``Compactifications of type IIB strings to four dimensions with
7915: non-trivial classical potential,''
7916: Nucl.\ Phys.\ B {\bf 495} (1997) 127
7917: [arXiv:hep-th/9610151].
7918:
7919: \bibitem{Verlinde}
7920: H.~Verlinde,
7921: ``Holography and compactification,''
7922: Nucl.\ Phys.\ B {\bf 580} (2000) 264
7923: [arXiv:hep-th/9906182];\\
7924: C.~S.~Chan, P.~L.~Paul and H.~Verlinde,
7925: ``A note on warped string compactification,''
7926: Nucl.\ Phys.\ B {\bf 581}, 156 (2000)
7927: [arXiv:hep-th/0003236].
7928:
7929:
7930: \bibitem{GVW}S.~Gukov, C.~Vafa and E.~Witten,
7931: ``CFT's from Calabi-Yau four-folds,''
7932: Nucl.\ Phys.\ B {\bf 584}, 69 (2000)
7933: [Erratum-ibid.\ B {\bf 608}, 477 (2001)]
7934: [arXiv:hep-th/9906070].
7935:
7936:
7937: \bibitem{DRS}
7938: K.~Dasgupta, G.~Rajesh and S.~Sethi,
7939: ``M theory, orientifolds and G-flux,''
7940: JHEP {\bf 9908} (1999) 023
7941: [arXiv:hep-th/9908088].
7942:
7943:
7944: \bibitem{TV}
7945: T.~R.~Taylor and C.~Vafa,
7946: ``RR flux on Calabi-Yau and partial supersymmetry breaking,''
7947: Phys.\ Lett.\ B {\bf 474} (2000) 130
7948: [arXiv:hep-th/9912152].
7949:
7950:
7951: \bibitem{Mayr}
7952: P.~Mayr,
7953: ``On supersymmetry breaking in string theory and its realization in brane
7954: worlds,''
7955: Nucl.\ Phys.\ B {\bf 593} (2001) 99
7956: [arXiv:hep-th/0003198];\\
7957: P.~Mayr,
7958: ``Stringy world branes and exponential hierarchies,''
7959: JHEP {\bf 0011} (2000) 013
7960: [arXiv:hep-th/0006204].
7961:
7962:
7963:
7964: \bibitem{GSS}
7965: B.~R.~Greene, K.~Schalm and G.~Shiu,
7966: ``Warped compactifications in M and F theory,''
7967: Nucl.\ Phys.\ B {\bf 584} (2000) 480
7968: [arXiv:hep-th/0004103].
7969:
7970:
7971: \bibitem{GKP}
7972: S.~B.~Giddings, S.~Kachru and J.~Polchinski,
7973: ``Hierarchies from fluxes in string compactifications,''
7974: Phys.\ Rev.\ D {\bf 66} (2002) 106006
7975: [arXiv:hep-th/0105097].
7976:
7977:
7978: \bibitem{CKLT}
7979: G.~Curio, A.~Klemm, D.~L{\"u}st and S.~Theisen,
7980: ``On the vacuum structure of type II string compactifications on Calabi-Yau spaces with H-fluxes,''
7981: Nucl.\ Phys.\ B {\bf 609} (2001) 3
7982: [arXiv:hep-th/0012213];\\
7983: G.~Curio and A.~Krause,
7984: ``Four-flux and warped heterotic M-theory compactifications,''
7985: Nucl.\ Phys.\ B {\bf 602}, 172 (2001)
7986: [arXiv:hep-th/0012152];\\
7987: G.~Curio, A.~Klemm, B.~K{\"o}rs and D.~L{\"u}st,
7988: ``Fluxes in heterotic and type II string compactifications,''
7989: Nucl.\ Phys.\ B {\bf 620} (2002) 237
7990: [arXiv:hep-th/0106155].
7991:
7992: \bibitem{HL}
7993: M.~Haack and J.~Louis,
7994: ``Duality in heterotic vacua with four supercharges,''
7995: Nucl.\ Phys.\ B {\bf 575} (2000) 107
7996: [arXiv:hep-th/9912181]; \\
7997: ``M-theory compactified on Calabi-Yau fourfolds with background flux,''
7998: Phys.\ Lett.\ B {\bf 507} (2001) 296
7999: [arXiv:hep-th/0103068].
8000:
8001:
8002: \bibitem{BB2}
8003: K.~Becker and M.~Becker,
8004: ``Supersymmetry breaking, M-theory and fluxes,''
8005: JHEP {\bf 0107} (2001) 038
8006: [arXiv:hep-th/0107044].
8007:
8008: \bibitem{DallAgata}
8009: G.~Dall'Agata,
8010: ``Type IIB supergravity compactified on a Calabi-Yau manifold with H-fluxes,''
8011: JHEP {\bf 0111} (2001) 005
8012: [arXiv:hep-th/0107264].
8013:
8014: \bibitem{KST}
8015: S.~Kachru, M.~B.~Schulz and S.~Trivedi,
8016: ``Moduli stabilization from fluxes in a simple IIB orientifold,''
8017: JHEP {\bf 0310} (2003) 007
8018: [arXiv:hep-th/0201028].
8019:
8020: \bibitem{LM}
8021: J.~Louis and A.~Micu,
8022: ``Type II theories compactified on Calabi-Yau threefolds in the presence of background fluxes,''
8023: Nucl.\ Phys.\ B {\bf 635} (2002) 395
8024: [arXiv:hep-th/0202168].
8025:
8026: \bibitem{BBHL}
8027: K.~Becker, M.~Becker, M.~Haack and J.~Louis,
8028: ``Supersymmetry breaking and alpha'-corrections to flux induced potentials,''
8029: JHEP {\bf 0206} (2002) 060
8030: [arXiv:hep-th/0204254].
8031:
8032: \bibitem{FPo}
8033: S.~Ferrara and M.~Porrati,
8034: ``N = 1 no-scale supergravity from IIB orientifolds,''
8035: Phys.\ Lett.\ B {\bf 545} (2002) 411
8036: [arXiv:hep-th/0207135].
8037:
8038: \bibitem{DWG} O.~DeWolfe and S.~B.~Giddings,
8039: ``Scales and hierarchies in warped compactifications and brane worlds,''
8040: Phys.\ Rev.\ D {\bf 67}, 066008 (2003)
8041: [arXiv:hep-th/0208123].
8042:
8043: \bibitem{Soft1}
8044: M.~Gra\~na,
8045: ``MSSM parameters from supergravity backgrounds,''
8046: Phys.\ Rev.\ D {\bf 67}, 066006 (2003)
8047: [arXiv:hep-th/0209200];\\
8048: B.~K\"ors and P.~Nath,
8049: ``Effective action and soft supersymmetry breaking for intersecting D-brane
8050: models,''
8051: Nucl.\ Phys.\ B {\bf 681}, 77 (2004)
8052: [arXiv:hep-th/0309167];\\
8053: P.~G.~C\'amara, L.~E.~Ib\'a\~nez and A.~M.~Uranga,
8054: ``Flux-induced SUSY-breaking soft terms,''
8055: Nucl.\ Phys.\ B {\bf 689} (2004) 195
8056: [arXiv:hep-th/0311241].
8057:
8058: \bibitem{GGJL}
8059: M.~Gra\~na, T.~W.~Grimm, H.~Jockers and J.~Louis,
8060: %``Soft supersymmetry breaking in Calabi-Yau orientifolds with D-branes and
8061: %fluxes,''
8062: Nucl.\ Phys.\ B {\bf 690} (2004) 21
8063: [arXiv:hep-th/0312232].
8064:
8065: \bibitem{Soft2}
8066: A.~Lawrence and J.~McGreevy,
8067: ``Local string models of soft supersymmetry breaking,''
8068: arXiv:hep-th/0401034;\\
8069: ``Remarks on branes, fluxes, and soft SUSY breaking,''
8070: arXiv:hep-th/0401233;
8071: D.~L\"ust, P.~Mayr, R.~Richter and S.~Stieberger,
8072: ``Scattering of gauge, matter, and moduli fields from intersecting branes,''
8073: Nucl.\ Phys.\ B {\bf 696}, 205 (2004)
8074: [arXiv:hep-th/0404134];\\
8075: D.~L\"ust, S.~Reffert and S.~Stieberger,
8076: ``Flux-induced soft supersymmetry breaking in chiral type IIb orientifolds with
8077: D3/D7-branes,''
8078: arXiv:hep-th/0406092;\\
8079: P.~G.~C\'amara, L.~E.~Ib\'a\~nez and A.~M.~Uranga,
8080: ``Flux-induced SUSY-breaking soft terms on D7-D3 brane systems,''
8081: arXiv:hep-th/0408036;\\
8082: D.~L\"ust, S.~Reffert and S.~Stieberger,
8083: ``MSSM with soft SUSY breaking terms from D7-branes with fluxes,''
8084: arXiv:hep-th/0410074;\\
8085: A.~Font and L.~E.~Ib\'a\~nez,
8086: ``SUSY-breaking soft terms in a MSSM magnetized D7-brane model,''
8087: arXiv:hep-th/0412150;\\
8088: D.~L\"ust, P.~Mayr, S.~Reffert and S.~Stieberger,
8089: ``F-theory flux, destabilization of orientifolds and soft terms on
8090: D7-branes,''
8091: arXiv:hep-th/0501139;\\
8092: K.~Choi, A.~Falkowski, H.~P.~Nilles and M.~Olechowski,
8093: ``Soft supersymmetry breaking in KKLT flux compactification,''
8094: arXiv:hep-th/0503216.
8095:
8096:
8097: \bibitem{BKL}
8098: R.~Blumenhagen, D.~L{\"u}st and T.~R.~Taylor,
8099: ``Moduli stabilization in chiral type IIB orientifold models with fluxes,''
8100: Nucl.\ Phys.\ B {\bf 663} (2003) 319
8101: [arXiv:hep-th/0303016].
8102:
8103: \bibitem{BHS}
8104: M.~Berg, M.~Haack and H.~Samtleben,
8105: ``Calabi-Yau fourfolds with flux and supersymmetry breaking,''
8106: JHEP {\bf 0304} (2003) 046
8107: [arXiv:hep-th/0212255].
8108:
8109: \bibitem{KKLT}
8110: S.~Kachru, R.~Kallosh, A.~Linde and S.~P.~Trivedi,
8111: ``De Sitter vacua in string theory,''
8112: Phys.\ Rev.\ D {\bf 68} (2003) 046005
8113: [arXiv:hep-th/0301240].
8114:
8115:
8116: \bibitem{Ferrara}
8117: R.~D'Auria, S.~Ferrara, F.~Gargiulo, M.~Trigiante and S.~Vaula,
8118: ``N = 4 supergravity Lagrangian for type IIB on T**6/Z(2) in presence of
8119: fluxes and D3-branes,''
8120: JHEP {\bf 0306} (2003) 045
8121: [arXiv:hep-th/0303049];\\
8122: C.~Angelantonj, S.~Ferrara and M.~Trigiante,
8123: ``New D = 4 gauged supergravities from N = 4 orientifolds with fluxes,''
8124: JHEP {\bf 0310} (2003) 015
8125: [arXiv:hep-th/0306185];\\
8126: for a review see,
8127: L.~Andrianopoli, S.~Ferrara and M.~Trigiante,
8128: ``Fluxes, supersymmetry breaking and gauged supergravity,''
8129: arXiv:hep-th/0307139;\\
8130: C.~Angelantonj, S.~Ferrara and M.~Trigiante,
8131: ``Unusual gauged supergravities from type IIA and type IIB orientifolds,''
8132: Phys.\ Lett.\ B {\bf 582}, 263 (2004)
8133: [arXiv:hep-th/0310136].
8134: C.~Angelantonj, R.~D'Auria, S.~Ferrara and M.~Trigiante,
8135: ``K3 x T**2/Z(2) orientifolds with fluxes, open string moduli and critical points,''
8136: Phys.\ Lett.\ B {\bf 583}, 331 (2004)
8137: [arXiv:hep-th/0312019].
8138:
8139:
8140: \bibitem{BHK}
8141: M.~Berg, M.~Haack and B.~K{\"o}rs,
8142: ``An orientifold with fluxes and branes via T-duality,''
8143: Nucl.\ Phys.\ B {\bf 669} (2003) 3
8144: [arXiv:hep-th/0305183];\\
8145: M.~Berg, M.~Haack and B.~K\"ors,
8146: ``Brane/Flux Interactions in Orientifolds,''
8147: arXiv:hep-th/0312172.
8148:
8149: \bibitem{deAlwis:2003sn}
8150: S.~P.~de Alwis,
8151: ``On potentials from fluxes,''
8152: Phys.\ Rev.\ D {\bf 68} (2003) 126001
8153: [arXiv:hep-th/0307084];\\
8154: A.~Buchel,
8155: ``On effective action of string theory flux compactifications,''
8156: Phys.\ Rev.\ D {\bf 69}, 106004 (2004)
8157: [arXiv:hep-th/0312076];\\
8158: A.~Giryavets, S.~Kachru, P.~K.~Tripathy and S.~P.~Trivedi,
8159: ``Flux compactifications on Calabi-Yau threefolds,''
8160: JHEP {\bf 0404} (2004) 003
8161: [arXiv:hep-th/0312104];\\
8162: R.~Brustein and S.~P.~de Alwis,
8163: ``Moduli potentials in string compactifications with fluxes: Mapping the
8164: discretuum,''
8165: Phys.\ Rev.\ D {\bf 69}, 126006 (2004)
8166: [arXiv:hep-th/0402088].
8167:
8168: \bibitem{TGL1}
8169: T.~W.~Grimm and J.~Louis,
8170: ``The effective action of N = 1 Calabi-Yau orientifolds,''
8171: Nucl.\ Phys.\ B {\bf 699} (2004) 387
8172: [arXiv:hep-th/0403067].
8173:
8174: \bibitem{KachruK}
8175: S.~Kachru and A.~K.~Kashani-Poor,
8176: ``Moduli potentials in type IIA compactifications with RR and NS flux,''
8177: JHEP {\bf 0503} (2005) 066
8178: [arXiv:hep-th/0411279].
8179: %%CITATION = HEP-TH 0411279;%%
8180:
8181: \bibitem{TGL2}
8182: T.~W.~Grimm and J.~Louis,
8183: ``The effective action of type IIA Calabi-Yau orientifolds,''
8184: arXiv:hep-th/0412277, to appear in Nucl.\ Phys.\ B.
8185:
8186: % end: flux
8187: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8188:
8189:
8190: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8191: % non-pert
8192:
8193:
8194: \bibitem{Witten}
8195: E.~Witten,
8196: ``Non-Perturbative Superpotentials In String Theory,''
8197: Nucl.\ Phys.\ B {\bf 474} (1996) 343
8198: [arXiv:hep-th/9604030].
8199:
8200: \bibitem{HM}
8201: J.~A.~Harvey and G.~W.~Moore,
8202: ``Superpotentials and membrane instantons,''
8203: arXiv:hep-th/9907026.
8204:
8205: \bibitem{non-pert}
8206: For recent results see, for example,
8207: L.~G\"orlich, S.~Kachru, P.~K.~Tripathy and S.~P.~Trivedi,
8208: ``Gaugino condensation and nonperturbative superpotentials in flux
8209: compactifications,''
8210: arXiv:hep-th/0407130;
8211: R.~Blumenhagen, M.~Cvetic, F.~Marchesano and G.~Shiu,
8212: ``Chiral D-brane models with frozen open string moduli,''
8213: JHEP {\bf 0503} (2005) 050
8214: [arXiv:hep-th/0502095]\\
8215: G.~Curio, A.~Krause and D.~L\"ust,
8216: ``Moduli stabilization in the heterotic / IIB discretuum,''
8217: arXiv:hep-th/0502168;\\
8218: P.~K.~Tripathy and S.~P.~Trivedi,
8219: ``D3 brane action and fermion zero modes in presence of background flux,''
8220: arXiv:hep-th/0503072;\\
8221: R.~Kallosh, A.~K.~Kashani-Poor and A.~Tomasiello,
8222: ``Counting fermionic zero modes on M5 with fluxes,''
8223: arXiv:hep-th/0503138;\\
8224: P.~Berglund and P.~Mayr,
8225: ``Non-perturbative superpotentials in F-theory and string duality,''
8226: arXiv:hep-th/0504058;\\
8227: D.~L\"ust, S.~Reffert, W.~Schulgin and S.~Stieberger,
8228: ``Moduli stabilization in type IIB orientifolds. I: Orbifold limits,''
8229: arXiv:hep-th/0506090,
8230: and references therein.
8231:
8232: \bibitem{Curio}
8233: G.~Curio and A.~Krause,
8234: ``G-fluxes and non-perturbative stabilisation of heterotic M-theory,''
8235: Nucl.\ Phys.\ B {\bf 643}, 131 (2002)
8236: [arXiv:hep-th/0108220];\\
8237: M.~Becker, G.~Curio and A.~Krause,
8238: ``De Sitter vacua from heterotic M-theory,''
8239: Nucl.\ Phys.\ B {\bf 693}, 223 (2004)
8240: [arXiv:hep-th/0403027].
8241:
8242:
8243: \bibitem{gaugino}
8244: J.~P.~Derendinger, L.~E.~Ib\'a\~nez and H.~P.~Nilles,
8245: ``On The Low-Energy D = 4, N=1 Supergravity Theory Extracted From The D = 10,
8246: N=1 Superstring,''
8247: Phys.\ Lett.\ B {\bf 155} (1985) 65;\\
8248: M.~Dine, R.~Rohm, N.~Seiberg and E.~Witten,
8249: ``Gluino Condensation In Superstring Models,''
8250: Phys.\ Lett.\ B {\bf 156} (1985) 55.
8251: C.~P.~Burgess, J.~P.~Derendinger, F.~Quevedo and M.~Quiros,
8252: ``On gaugino condensation with field-dependent gauge couplings,''
8253: Annals Phys.\ {\bf 250} (1996) 193
8254: [arXiv:hep-th/9505171].
8255:
8256:
8257: %%CITATION = HEP-TH 0502095;%%
8258:
8259:
8260: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8261: % stab all moduli
8262:
8263:
8264: \bibitem{DSFGK}
8265: F.~Denef, M.~R.~Douglas, B.~Florea, A.~Grassi and S.~Kachru,
8266: ``Fixing All Moduli in a Simple F-Theory Compactification,''
8267: arXiv:hep-th/0503124.
8268:
8269:
8270: \bibitem{reviewPP}
8271: For a review see, for example,
8272: E.~Kiritsis,
8273: ``D-branes in standard model building, gravity and cosmology,''
8274: Fortsch.\ Phys.\ {\bf 52} (2004) 200
8275: [arXiv:hep-th/0310001];\\
8276: A.~M.~Uranga,
8277: ``Chiral four-dimensional string compactifications with intersecting
8278: D-branes,''
8279: Class.\ Quant.\ Grav.\ {\bf 20}, S373 (2003)
8280: [arXiv:hep-th/0301032];\\
8281: D.~L\"ust,
8282: ``Intersecting brane worlds: A path to the standard model?,''
8283: Class.\ Quant.\ Grav.\ {\bf 21} (2004) S1399
8284: [arXiv:hep-th/0401156];\\
8285: L.~E.~Ib\'a\~nez,
8286: ``The fluxed MSSM,''
8287: Phys.\ Rev.\ D {\bf 71}, 055005 (2005)
8288: [arXiv:hep-ph/0408064]; \\
8289: R.~Blumenhagen,
8290: ``Recent progress in intersecting D-brane models,''
8291: arXiv:hep-th/0412025;
8292: R.~Blumenhagen, M.~Cvetic, P.~Langacker and G.~Shiu,
8293: ``Toward realistic intersecting D-brane models,''
8294: arXiv:hep-th/0502005,
8295: and references therein.
8296:
8297:
8298: \bibitem{reviewcosmo}
8299: For a review see, for example,
8300: A.~Linde,
8301: ``Prospects of inflation,''
8302: arXiv:hep-th/0402051;\\
8303: V.~Balasubramanian,
8304: ``Accelerating universes and string theory,''
8305: Class.\ Quant.\ Grav.\ {\bf 21} (2004) S1337
8306: [arXiv:hep-th/0404075];\\
8307: C.~P.~Burgess,
8308: ``Inflationary String Theory?,''
8309: arXiv:hep-th/0408037;
8310: A.~D.~Linde, ``Particle Physics and Inflationary Cosmology,''
8311: arXiv:hep-th/0503203,
8312: and references therein.
8313:
8314: \bibitem{JP}
8315: A.~Sagnotti,
8316: ``Open Strings And Their Symmetry Groups,''
8317: arXiv:hep-th/0208020; \\
8318: J.~Dai, R.~G.~Leigh and J.~Polchinski,
8319: ``New Connections Between String Theories,''
8320: Mod.\ Phys.\ Lett.\ A {\bf 4} (1989) 2073;\\
8321: R.~G.~Leigh,
8322: ``Dirac-Born-Infeld Action From Dirichlet Sigma Model,''
8323: Mod.\ Phys.\ Lett.\ A {\bf 4} (1989) 2767;\\
8324: M.~Bianchi and A.~Sagnotti,
8325: ``On The Systematics Of Open String Theories,''
8326: Phys.\ Lett.\ B {\bf 247} (1990) 517;
8327: ``Twist Symmetry And Open String Wilson Lines,''
8328: Nucl.\ Phys.\ B {\bf 361} (1991) 519;\\
8329: P.~Horava,
8330: ``Strings On World Sheet Orbifolds,''
8331: Nucl.\ Phys.\ B {\bf 327} (1989) 461;\\
8332: J.~Polchinski,
8333: ``Dirichlet-Branes and Ramond-Ramond Charges,''
8334: Phys.\ Rev.\ Lett.\ {\bf 75} (1995) 4724
8335: [arXiv:hep-th/9510017];\\
8336: E.~G.~Gimon and J.~Polchinski,
8337: ``Consistency Conditions for Orientifolds and D-Manifolds,''
8338: Phys.\ Rev.\ D {\bf 54} (1996) 1667
8339: [arXiv:hep-th/9601038].
8340:
8341: \bibitem{D-branes}
8342: For a review see, for example,
8343: J.~Polchinski, ``Lectures on D-branes,'' arXiv:hep-th/9611050; \\
8344: C.~P.~Bachas, ``Lectures on D-branes,'' arXiv:hep-th/9806199;\\
8345: C.~V.~Johnson, ``D-brane primer,'' arXiv:hep-th/0007170,
8346: and references therein.
8347:
8348: \bibitem{AD}
8349: For a review see, for example,
8350: A.~Dabholkar,
8351: ``Lectures on orientifolds and duality,''
8352: [arXiv:hep-th/9804208];\\
8353: C.~Angelantonj and A.~Sagnotti,
8354: ``Open strings,''
8355: Phys.\ Rept.\ {\bf 371} (2002) 1
8356: [Erratum-ibid.\ {\bf 376} (2003) 339]
8357: [arXiv:hep-th/0204089], and references therein.
8358:
8359:
8360: \bibitem{RS}
8361: L.~Randall and R.~Sundrum,
8362: ``An alternative to compactification,''
8363: Phys.\ Rev.\ Lett.\ {\bf 83}, 4690 (1999)
8364: [arXiv:hep-th/9906064];\\
8365: L.~Randall and R.~Sundrum,
8366: ``A large mass hierarchy from a small extra dimension,''
8367: Phys.\ Rev.\ Lett.\ {\bf 83}, 3370 (1999)
8368: [arXiv:hep-ph/9905221].
8369:
8370:
8371: \bibitem{deWitSD}
8372: B.~de Wit, D.~J.~Smit and N.~D.~Hari Dass,
8373: ``Residual Supersymmetry Of Compactified D = 10 Supergravity,''
8374: Nucl.\ Phys.\ B {\bf 283} (1987) 165;\\
8375: J.~M.~Maldacena and C.~Nunez,
8376: ``Supergravity description of field theories on curved manifolds and
8377: a no go theorem,''
8378: Int.\ J.\ Mod.\ Phys.\ A {\bf 16} (2001) 822
8379: [arXiv:hep-th/0007018].
8380:
8381:
8382:
8383: \bibitem{reviewAdSCFT}
8384: For a review see, for example,
8385: O.~Aharony, S.~S.~Gubser, J.~M.~Maldacena, H.~Ooguri and Y.~Oz,
8386: ``Large N field theories, string theory and gravity,''
8387: Phys.\ Rept.\ {\bf 323} (2000) 183
8388: [arXiv:hep-th/9905111],
8389: and references therein.
8390:
8391: \bibitem{KL} V.~S.~Kaplunovsky and J.~Louis,
8392: ``Model independent analysis of soft terms in effective supergravity and in string theory,''
8393: Phys.\ Lett.\ B {\bf 306}, 269 (1993)
8394: [arXiv:hep-th/9303040].
8395:
8396: \bibitem{BIM}
8397: L.~E.~Ib\'a\~nez and D.~L\"ust,
8398: ``Duality anomaly cancellation, minimal string unification and the effective low-energy Lagrangian of 4-D strings,''
8399: Nucl.\ Phys.\ B {\bf 382} (1992) 305
8400: [arXiv:hep-th/9202046];\\
8401: A.~Brignole, L.~E.~Ib\'a\~nez and C.~Mu\~noz,
8402: ``Towards a theory of soft terms for the supersymmetric Standard Model,''
8403: Nucl.\ Phys.\ B {\bf 422}, 125 (1994)
8404: [Erratum-ibid.\ B {\bf 436}, 747 (1995)]
8405: [arXiv:hep-ph/9308271].
8406:
8407:
8408: \bibitem{BKQ}
8409: C.~P.~Burgess, R.~Kallosh and F.~Quevedo,
8410: ``de Sitter string vacua from supersymmetric D-terms,''
8411: JHEP {\bf 0310} (2003) 056
8412: [arXiv:hep-th/0309187].
8413:
8414: \bibitem{Ori}
8415: C.~Angelantonj, M.~Bianchi, G.~Pradisi, A.~Sagnotti and Y.~S.~Stanev,
8416: ``Chiral asymmetry in four-dimensional open- string vacua,''
8417: Phys.\ Lett.\ B {\bf 385} (1996) 96
8418: [arXiv:hep-th/9606169];\\
8419: M.~Berkooz and R.~G.~Leigh,
8420: ``A D = 4 N = 1 orbifold of type I strings,''
8421: Nucl.\ Phys.\ B {\bf 483} (1997) 187
8422: [arXiv:hep-th/9605049];\\
8423: G.~Aldazabal, A.~Font, L.~E.~Ib\'a\~nez and G.~Violero,
8424: ``D = 4, N = 1, type IIB orientifolds,''
8425: Nucl.\ Phys.\ B {\bf 536} (1998) 29
8426: [arXiv:hep-th/9804026];\\
8427: M.~Cvetic, G.~Shiu and A.~M.~Uranga,
8428: ``Chiral four-dimensional N = 1 supersymmetric type IIA orientifolds from
8429: intersecting D6-branes,''
8430: Nucl.\ Phys.\ B {\bf 615} (2001) 3
8431: [arXiv:hep-th/0107166];
8432:
8433:
8434:
8435:
8436: \bibitem{AAHV}
8437: B.~Acharya, M.~Aganagic, K.~Hori and C.~Vafa,
8438: ``Orientifolds, mirror symmetry and superpotentials,''
8439: [arXiv:hep-th/0202208].
8440:
8441:
8442: \bibitem{BBKL}
8443: R.~Blumenhagen, V.~Braun, B.~K\"ors and D.~L\"ust,
8444: ``Orientifolds of K3 and Calabi-Yau manifolds with intersecting D-branes,''
8445: JHEP {\bf 0207} (2002) 026
8446: [arXiv:hep-th/0206038];
8447: R.~Blumenhagen, V.~Braun, B.~K\"ors and D.~L\"ust,
8448: ``The standard model on the quintic,''
8449: arXiv:hep-th/0210083.
8450:
8451:
8452: \bibitem{BH}
8453: I.~Brunner and K.~Hori,
8454: ``Orientifolds and mirror symmetry,''
8455: JHEP {\bf 0411} (2004) 005
8456: [arXiv:hep-th/0303135];\\
8457: I.~Brunner, K.~Hori, K.~Hosomichi and J.~Walcher,
8458: ``Orientifolds of Gepner models,''
8459: arXiv:hep-th/0401137.
8460:
8461: \bibitem{revT-dual}
8462: For a review see, for example,
8463: E.~Alvarez, L.~Alvarez-Gaume and Y.~Lozano,
8464: ``An introduction to T duality in string theory,''
8465: Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 41} (1995) 1
8466: [arXiv:hep-th/9410237]; \\
8467: A.~Giveon, M.~Porrati and E.~Rabinovici,
8468: ``Target space duality in string theory,''
8469: Phys.\ Rept.\ {\bf 244} (1994) 77
8470: [arXiv:hep-th/9401139].
8471: and references therein.
8472:
8473: \bibitem{SYZ}
8474: A.~Strominger, S.~T.~Yau and E.~Zaslow,
8475: ``Mirror symmetry is T-duality,''
8476: Nucl.\ Phys.\ B {\bf 479} (1996) 243
8477: [arXiv:hep-th/9606040].
8478:
8479: \bibitem{Mirror}
8480: For a review see, for example,
8481: S.~Hosono, A.~Klemm and S.~Theisen,
8482: ``Lectures on mirror symmetry,''
8483: arXiv:hep-th/9403096;\\
8484: K.~Hori, S.~Katz, A.~Klemm, R.~Pandharipande, R.~Thomas, C.~Vafa, R.~Vakil and E.~Zaslow,
8485: ``Mirror symmetry'' (Clay Mathematics Monographs, Vol. 1,2).
8486: and references therein.
8487:
8488: \bibitem{Aspinwall}
8489: For a review see, for example, P.~S.~Aspinwall,
8490: ``D-branes on Calabi-Yau manifolds,''
8491: arXiv:hep-th/0403166.
8492: and references therein.
8493:
8494: \bibitem{Vafa_NCY}
8495: C.~Vafa,
8496: ``Superstrings and topological strings at large N,''
8497: J.\ Math.\ Phys.\ {\bf 42} (2001) 2798
8498: [arXiv:hep-th/0008142].
8499:
8500: \bibitem{GLMW}
8501: S.~Gurrieri, J.~Louis, A.~Micu and D.~Waldram,
8502: ``Mirror symmetry in generalized Calabi-Yau compactifications,''
8503: Nucl.\ Phys.\ B {\bf 654}, 61 (2003)
8504: [arXiv:hep-th/0211102].
8505:
8506: \bibitem{Vafa}
8507: C.~Vafa,
8508: ``Evidence for F-Theory,''
8509: Nucl.\ Phys.\ B {\bf 469} (1996) 403
8510: [arXiv:hep-th/9602022];\\
8511: D.~R.~Morrison and C.~Vafa,
8512: ``Compactifications of F-Theory on Calabi--Yau Threefolds -- I,''
8513: Nucl.\ Phys.\ B {\bf 473}, 74 (1996)
8514: [arXiv:hep-th/9602114];\\
8515: D.~R.~Morrison and C.~Vafa,
8516: ``Compactifications of F-Theory on Calabi--Yau Threefolds -- II,''
8517: Nucl.\ Phys.\ B {\bf 476}, 437 (1996)
8518: [arXiv:hep-th/9603161].
8519:
8520: \bibitem{Sen}
8521: A.~Sen,
8522: ``F-theory and Orientifolds,''
8523: Nucl.\ Phys.\ B {\bf 475} (1996) 562
8524: [arXiv:hep-th/9605150].
8525:
8526: \bibitem{CJS}
8527: E.~Cremmer, B.~Julia and J.~Scherk,
8528: ``Supergravity Theory In 11 Dimensions,''
8529: Phys.\ Lett.\ B {\bf 76} (1978) 409.
8530:
8531:
8532: %%%%%%%%%%%%% Outline
8533:
8534: \bibitem{BGHL}
8535: M.~Bodner and A.~C.~Cadavid,
8536: ``Dimensional Reduction Of Type IIb Supergravity And Exceptional Quaternionic Manifolds,''
8537: Class.\ Quant.\ Grav.\ {\bf 7} (1990) 829;\\
8538: R.~B\"ohm, H.~G\"unther, C.~Herrmann and J.~Louis,
8539: ``Compactification of type IIB string theory on Calabi-Yau threefolds,''
8540: Nucl.\ Phys.\ B {\bf 569} (2000) 229
8541: [arXiv:hep-th/9908007].
8542:
8543: \bibitem{CP}
8544: D.~Cremades, L.~E.~Ib\'a\~nez and F.~Marchesano,
8545: ``Yukawa couplings in intersecting D-brane models,''
8546: JHEP {\bf 0307}, 038 (2003)
8547: [arXiv:hep-th/0302105];\\
8548: M.~Cvetic and I.~Papadimitriou,
8549: ``Conformal field theory couplings for intersecting D-branes on
8550: orientifolds,''
8551: Phys.\ Rev.\ D {\bf 68} (2003) 046001
8552: [Erratum-ibid.\ D {\bf 70} (2004) 029903]
8553: [arXiv:hep-th/0303083];\\
8554: D.~L\"ust, P.~Mayr, R.~Richter and S.~Stieberger,
8555: ``Scattering of gauge, matter, and moduli fields from intersecting branes,''
8556: Nucl.\ Phys.\ B {\bf 696}, 205 (2004)
8557: [arXiv:hep-th/0404134].
8558:
8559:
8560: \bibitem{BGG}
8561: P.~Binetruy, G.~Girardi and R.~Grimm,
8562: ``Supergravity couplings: A geometric formulation,''
8563: Phys.\ Rept.\ {\bf 343} (2001) 255
8564: [arXiv:hep-th/0005225].
8565:
8566:
8567: \bibitem{Hitchin2}
8568: N. Hitchin, ``The moduli space of special Lagrangian submanifolds,''
8569: dg-ga/9711002.
8570:
8571:
8572: \bibitem{HitchinGCM}
8573: N.~Hitchin,
8574: ``Generalized Calabi-Yau manifolds,''
8575: Quart.\ J.\ Math.\ Oxford Ser.\ {\bf 54} (2003) 281
8576: [arXiv:math.dg/0209099].
8577:
8578:
8579: \bibitem{Gualtieri}
8580: M.~Gualtieri, `` Generalized complex geometry,''
8581: math.DG/0401221.
8582:
8583:
8584: \bibitem{mass_tensors}
8585: R.~D'Auria and S.~Ferrara,
8586: ``Dyonic masses from conformal field strengths in D even dimensions,''
8587: Phys.\ Lett.\ B {\bf 606} (2005) 211
8588: [arXiv:hep-th/0410051];\\
8589: J.~Louis and W.~Schulgin,
8590: ``Massive tensor multiplets in N = 1 supersymmetry,''
8591: Fortsch.\ Phys.\ {\bf 53} (2005) 235
8592: [arXiv:hep-th/0410149];\\
8593: U.~Theis,
8594: ``Masses and dualities in extended Freedman-Townsend models,''
8595: Phys.\ Lett.\ B {\bf 609} (2005) 402
8596: [arXiv:hep-th/0412177];\\
8597: S.~M.~Kuzenko,
8598: ``On massive tensor multiplets,''
8599: JHEP {\bf 0501} (2005) 041
8600: [arXiv:hep-th/0412190];\\
8601: R.~D'Auria, S.~Ferrara, M.~Trigiante and S.~Vaula,
8602: ``N = 1 reductions of N = 2 supergravity in the presence of tensor
8603: multiplets,''
8604: JHEP {\bf 0503} (2005) 052
8605: [arXiv:hep-th/0502219].
8606:
8607:
8608: \bibitem{DeWGKT}
8609: O.~DeWolfe, A.~Giryavets, S.~Kachru and W.~Taylor,
8610: ``Type IIA Moduli Stabilization,''
8611: arXiv:hep-th/0505160.
8612:
8613: \bibitem{VZ}
8614: G.~Villadoro and F.~Zwirner,
8615: ``N = 1 effective potential from dual type-IIA D6/O6 orientifolds with
8616: general fluxes,''
8617: JHEP {\bf 0506} (2005) 047
8618: [arXiv:hep-th/0503169].
8619:
8620:
8621:
8622: \bibitem{KMcG}
8623: S.~Kachru and J.~McGreevy,
8624: ``M-theory on manifolds of G(2) holonomy and type IIA orientifolds,''
8625: JHEP {\bf 0106} (2001) 027
8626: [arXiv:hep-th/0103223].
8627:
8628: \bibitem{PT}
8629: G.~Papadopoulos and P.~K.~Townsend,
8630: ``Compactification of D = 11 supergravity on spaces of exceptional holonomy,''
8631: Phys.\ Lett.\ B {\bf 357} (1995) 300
8632: [arXiv:hep-th/9506150].
8633:
8634: \bibitem{Hitchin1}
8635: N. Hitchin, ``The geometry of three-forms in six and seven dimensions,''
8636: math.DG/0010054.
8637:
8638:
8639: \bibitem{GPap}
8640: J.~Gutowski and G.~Papadopoulos,
8641: ``Moduli spaces and brane solitons for M theory compactifications on holonomy
8642: G(2) manifolds,''
8643: Nucl.\ Phys.\ B {\bf 615} (2001) 237
8644: [arXiv:hep-th/0104105].
8645:
8646:
8647: \bibitem{BW}
8648: C.~Beasley and E.~Witten,
8649: ``A note on fluxes and superpotentials in M-theory compactifications on
8650: manifolds of G(2) holonomy,''
8651: JHEP {\bf 0207} (2002) 046
8652: [arXiv:hep-th/0203061].
8653:
8654:
8655:
8656: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8657: %%
8658: %% Chapter 2
8659: %%
8660: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8661:
8662: \bibitem{Huebsch}
8663: T. H\"ubsch,
8664: ``Calbi-Yau manifolds -- A Bestiary for Physicists'',
8665: World Scientific Publishing (1994).
8666:
8667:
8668: \bibitem{Tian}
8669: G.~Tian, in ``Mathematical aspects of string theory'', p.629, S.-T.Yau (ed.),
8670: World Scientific, Singapore, 1987
8671:
8672:
8673: \bibitem{Strominger2}
8674: A.~Strominger,
8675: ``Special Geometry,''
8676: Commun.\ Math.\ Phys.\ {\bf 133} (1990) 163.
8677:
8678:
8679: \bibitem{CdO}
8680: P.~Candelas and X.~de la Ossa,
8681: ``Moduli Space Of Calabi-Yau Manifolds,''
8682: Nucl.\ Phys.\ B {\bf 355}, 455 (1991).
8683:
8684:
8685:
8686: \bibitem{Strominger}
8687: A.~Strominger,
8688: ``Yukawa Couplings In Superstring Compactification,''
8689: Phys.\ Rev.\ Lett.\ {\bf 55} (1985) 2547.
8690:
8691:
8692:
8693: \bibitem{BCF}
8694: M.~Bodner, A.~C.~Cadavid and S.~Ferrara,
8695: ``(2,2) Vacuum Configurations For Type Iia Superstrings: N=2 Supergravity
8696: Lagrangians And Algebraic Geometry,''
8697: Class.\ Quant.\ Grav.\ {\bf 8} (1991) 789.
8698:
8699:
8700: \bibitem{FS}
8701: S.~Ferrara and S.~Sabharwal,
8702: ``Dimensional Reduction Of Type II Superstrings,''
8703: Class.\ Quant.\ Grav.\ {\bf 6} (1989) L77;\\
8704: ``Quaternionic Manifolds For Type II Superstring Vacua Of Calabi-Yau Spaces,''
8705: Nucl.\ Phys.\ B {\bf 332} (1990) 317.
8706:
8707:
8708: \bibitem{N=2review}
8709: For a review of $N=2$ supergravity see, for example,
8710: L.~Andrianopoli, M.~Bertolini, A.~Ceresole, R.~D'Auria, S.~Ferrara, P.~Fre and T.~Magri,
8711: ``N = 2 supergravity and N = 2 super Yang-Mills theory on general scalar
8712: manifolds: Symplectic covariance, gaugings and the momentum map,''
8713: J.\ Geom.\ Phys.\ {\bf 23} (1997) 111
8714: [arXiv:hep-th/9605032].
8715:
8716:
8717: \bibitem{CDAF}
8718: H.~Suzuki,
8719: ``Calabi-Yau compactification of type IIB string and a mass formula of the
8720: extreme black holes,''
8721: Mod.\ Phys.\ Lett.\ A {\bf 11} (1996) 623
8722: [arXiv:hep-th/9508001];\\
8723: A.~Ceresole, R.~D'Auria and S.~Ferrara,
8724: ``The Symplectic Structure of N=2 Supergravity and its Central Extension,''
8725: Nucl.\ Phys.\ Proc.\ Suppl.\ {\bf 46} (1996) 67
8726: [arXiv:hep-th/9509160].
8727:
8728:
8729: \bibitem{CFGi}
8730: S.~Cecotti, S.~Ferrara and L.~Girardello,
8731: ``Geometry Of Type Ii Superstrings And The Moduli Of Superconformal Field
8732: Theories,''
8733: Int.\ J.\ Mod.\ Phys.\ A {\bf 4} (1989) 2475.
8734:
8735:
8736: \bibitem{BVT}
8737: F.~Brandt,
8738: ``New N = 2 supersymmetric gauge theories: The double tensor multiplet and
8739: its interactions,''
8740: Nucl.\ Phys.\ B {\bf 587} (2000) 543
8741: [arXiv:hep-th/0005086];\\
8742: U.~Theis and S.~Vandoren,
8743: ``N = 2 supersymmetric scalar-tensor couplings,''
8744: JHEP {\bf 0304} (2003) 042
8745: [arXiv:hep-th/0303048];\\
8746: G.~Dall'Agata, R.~D'Auria, L.~Sommovigo and S.~Vaula,
8747: ``D = 4, N = 2 gauged supergravity in the presence of tensor multiplets,''
8748: arXiv:hep-th/0312210.
8749:
8750:
8751: \bibitem{BaggerW}
8752: J.~Bagger and E.~Witten,
8753: ``Matter Couplings In N=2 Supergravity ,''
8754: Nucl.\ Phys.\ B {\bf 222} (1983) 1.
8755:
8756:
8757: \bibitem{dWvP}
8758: B.~de Wit and A.~Van Proeyen,
8759: ``Potentials And Symmetries Of General Gauged N=2 Supergravity - Yang-Mills
8760: Models,''
8761: Nucl.\ Phys.\ B {\bf 245} (1984) 89;\\
8762: B.~de Wit, P.~G.~Lauwers and A.~Van Proeyen,
8763: ``Lagrangians Of N=2 Supergravity - Matter Systems,''
8764: Nucl.\ Phys.\ B {\bf 255} (1985) 569.
8765:
8766: \bibitem{CdOGP}
8767: P.~Candelas, X.~C.~De La Ossa, P.~S.~Green and L.~Parkes,
8768: ``A Pair Of Calabi-Yau Manifolds As An Exactly Soluble Superconformal
8769: Theory,''
8770: Nucl.\ Phys.\ B {\bf 359} (1991) 21.
8771: %%CITATION = NUPHA,B359,21;%%
8772:
8773:
8774:
8775: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8776: %%
8777: %% Chapter 3
8778: %%
8779: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8780:
8781: \bibitem{JL}
8782: H.~Jockers and J.~Louis,
8783: ``The effective action of D7-branes in N = 1 Calabi-Yau orientifolds,''
8784: Nucl.\ Phys.\ B {\bf 705} (2005) 167
8785: [arXiv:hep-th/0409098];\\
8786: H.~Jockers and J.~Louis,
8787: ``D-terms and F-terms from D7-brane fluxes,''
8788: arXiv:hep-th/0502059.
8789:
8790: \bibitem{Leigh:jq}
8791: R.~G.~Leigh,
8792: ``Dirac-Born-Infeld Action From Dirichlet Sigma Model,''
8793: Mod.\ Phys.\ Lett.\ A {\bf 4} (1989) 2767.
8794:
8795:
8796: \bibitem{Douglas:1995bn}
8797: M.~R.~Douglas,
8798: ``Branes within branes,''
8799: arXiv:hep-th/9512077.
8800:
8801: \bibitem{Myers}
8802: R.~C.~Myers,
8803: ``Dielectric-branes,''
8804: JHEP {\bf 9912} (1999) 022
8805: [arXiv:hep-th/9910053].
8806:
8807:
8808: \bibitem{BBS}
8809: K.~Becker, M.~Becker and A.~Strominger,
8810: ``Five-branes, membranes and nonperturbative string theory,''
8811: Nucl.\ Phys.\ B {\bf 456} (1995) 130
8812: [arXiv:hep-th/9507158].
8813:
8814:
8815: \bibitem{MMMS}
8816: M.~Marino, R.~Minasian, G.~W.~Moore and A.~Strominger,
8817: ``Nonlinear instantons from supersymmetric p-branes,''
8818: JHEP {\bf 0001} (2000) 005
8819: [arXiv:hep-th/9911206].
8820:
8821:
8822: \bibitem{CU}
8823: J.~F.~G.~Cascales and A.~M.~Uranga,
8824: ``Branes on generalized calibrated submanifolds,''
8825: JHEP {\bf 0411} (2004) 083
8826: [arXiv:hep-th/0407132].
8827:
8828: \bibitem{DHVW}
8829: L.~J.~Dixon, J.~A.~Harvey, C.~Vafa and E.~Witten,
8830: ``Strings On Orbifolds,''
8831: Nucl.\ Phys.\ B {\bf 261} (1985) 678;\\
8832: L.~J.~Dixon, J.~A.~Harvey, C.~Vafa and E.~Witten,
8833: ``Strings On Orbifolds. 2,''
8834: Nucl.\ Phys.\ B {\bf 274} (1986) 285.
8835:
8836:
8837: \bibitem{DP}
8838: A.~Dabholkar and J.~Park,
8839: ``Strings on Orientifolds,''
8840: Nucl.\ Phys.\ B {\bf 477} (1996) 701
8841: [arXiv:hep-th/9604178].
8842:
8843:
8844: \bibitem{HitchinLec}
8845: See, for example, N.~Hitchin,
8846: ``Lectures on Special Lagrangian Submanifolds'',
8847: Lectures given at the ICTP School on Differential Geometry April 1999.
8848:
8849:
8850: \bibitem{OSV}
8851: A.~Strominger,
8852: ``Macroscopic Entropy of $N=2$ Extremal Black Holes,''
8853: Phys.\ Lett.\ B {\bf 383} (1996) 39
8854: [arXiv:hep-th/9602111].
8855:
8856:
8857: \bibitem{BFM}
8858: S.~Sinha and C.~Vafa,
8859: ``SO and Sp Chern-Simons at large N,''
8860: arXiv:hep-th/0012136;\\
8861: D.~E.~Diaconescu, B.~Florea and A.~Misra,
8862: ``Orientifolds, unoriented instantons and localization,''
8863: JHEP {\bf 0307} (2003) 041
8864: [arXiv:hep-th/0305021];\\
8865: V.~Bouchard, B.~Florea and M.~Marino,
8866: ``Counting higher genus curves with crosscaps in Calabi-Yau orientifolds,''
8867: arXiv:hep-th/0405083.
8868:
8869:
8870:
8871:
8872:
8873:
8874:
8875: \bibitem{AM}
8876: D.V.\ Alekseevsky, S.\ Marchiafava,
8877: ``Hermitian and K\"ahler Submanifolds of a Quaternionic K\"ahler Manifold,''
8878: Osaka J.\ Math.\ 38, 4 , (2001), 869.
8879:
8880:
8881: \bibitem{ADAF}
8882: L.~Andrianopoli, R.~D'Auria and S.~Ferrara,
8883: ``Supersymmetry reduction of N-extended supergravities in four dimensions,''
8884: JHEP {\bf 0203} (2002) 025
8885: [arXiv:hep-th/0110277];\\
8886: ``Consistent reduction of N = 2 $\to$ N = 1 four dimensional supergravity coupled to matter,''
8887: Nucl.\ Phys.\ B {\bf 628} (2002) 387
8888: [arXiv:hep-th/0112192].
8889:
8890:
8891: \bibitem{GP}
8892: M.~Gra\~na and J.~Polchinski,
8893: ``Supersymmetric three-form flux perturbations on AdS(5),''
8894: Phys.\ Rev.\ D {\bf 63} (2001) 026001
8895: [arXiv:hep-th/0009211];\\
8896: M.~Gra\~na and J.~Polchinski,
8897: ``Gauge / gravity duals with holomorphic dilaton,''
8898: Phys.\ Rev.\ D {\bf 65} (2002) 126005
8899: [arXiv:hep-th/0106014].
8900:
8901:
8902: \bibitem{FP}
8903: A.~R.~Frey and J.~Polchinski,
8904: ``N = 3 warped compactifications,''
8905: Phys.\ Rev.\ D {\bf 65} (2002) 126009
8906: [arXiv:hep-th/0201029].
8907:
8908:
8909: \bibitem{WB}
8910: J.~Wess and J.~Bagger,
8911: ``Supersymmetry And Supergravity,''
8912: Princeton University Press, Princeton, 1992.
8913:
8914: \bibitem{GGRS}
8915: S. J.\ Gates, M.\ T.\ Grisaru, M.\ Rocek and W.\ Siegel,
8916: ``Superspace: or one thousand and one lessons in supersymmetry,''
8917: Frontiers in Physics, 58, Benjamin/Cummings, 1983.
8918:
8919: \bibitem{GLprep}
8920: I.~Benmachiche, T.~W.~Grimm and J.~Louis, in preperation.
8921:
8922:
8923: \bibitem{DAFT}
8924: R.~D'Auria, S.~Ferrara and M.~Trigiante,
8925: ``c-map,very special quaternionic geometry and dual K\"ahler spaces,''
8926: Phys.\ Lett.\ B {\bf 587} (2004) 138
8927: [arXiv:hep-th/0401161];\\
8928: R.~D'Auria, S.~Ferrara and M.~Trigiante,
8929: ``Homogeneous special manifolds, orientifolds and solvable coordinates,''
8930: Nucl.\ Phys.\ B {\bf 693} (2004) 261
8931: [arXiv:hep-th/0403204].
8932:
8933:
8934:
8935: \bibitem{NS}
8936: E.~Cremmer, S.~Ferrara, C.~Kounnas and D.~V.~Nanopoulos,
8937: ``Naturally Vanishing Cosmological Constant In N=1 Supergravity,''
8938: Phys.\ Lett.\ B {\bf 133}, 61 (1983); \\
8939: J.~R.~Ellis, A.~B.~Lahanas, D.~V.~Nanopoulos and K.~Tamvakis,
8940: ``No - Scale Supersymmetric Standard Model,''
8941: Phys.\ Lett.\ B {\bf 134}, 429 (1984);\\
8942: R.~Barbieri, E.~Cremmer and S.~Ferrara,
8943: ``Flat And Positive Potentials In N=1 Supergravity,''
8944: Phys.\ Lett.\ B {\bf 163} (1985) 143.
8945:
8946:
8947: \bibitem{BCOV}
8948: M.~Bershadsky, S.~Cecotti, H.~Ooguri and C.~Vafa,
8949: ``Kodaira-Spencer theory of gravity and exact results for quantum string
8950: amplitudes,''
8951: Commun.\ Math.\ Phys.\ {\bf 165} (1994) 311
8952: [arXiv:hep-th/9309140].
8953:
8954:
8955: \bibitem{Witten2}
8956: E.~Witten,
8957: ``Quantum background independence in string theory,''
8958: arXiv:hep-th/9306122.
8959:
8960:
8961: \bibitem{NOV}
8962: N.~Nekrasov, H.~Ooguri and C.~Vafa,
8963: ``S-duality and topological strings,''
8964: JHEP {\bf 0410} (2004) 009
8965: [arXiv:hep-th/0403167].
8966:
8967:
8968: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8969: %
8970: % Chapter 4
8971: %
8972: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8973:
8974: \bibitem{Binetruy}
8975: P.~Binetruy,
8976: ``Dilaton, Moduli And String / Five-Brane Duality As Seen From
8977: Four-Dimensions,''
8978: Phys.\ Lett.\ B {\bf 315} (1993) 80 [arXiv:hep-th/9305069].
8979:
8980:
8981: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8982: %
8983: % Chapter 5
8984: %
8985: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
8986:
8987: \bibitem{Romans}
8988: L.~J.~Romans,
8989: ``Massive N=2a Supergravity In Ten-Dimensions,''
8990: Phys.\ Lett.\ B {\bf 169} (1986) 374.
8991:
8992:
8993: \bibitem{FMM}
8994: S.~Fidanza, R.~Minasian and A.~Tomasiello,
8995: ``Mirror symmetric SU(3)-structure manifolds with NS fluxes,''
8996: arXiv:hep-th/0311122; \\
8997: M.~Gra\~na, R.~Minasian, M.~Petrini and A.~Tomasiello,
8998: ``Supersymmetric backgrounds from generalized Calabi-Yau manifolds,''
8999: JHEP {\bf 0408} (2004) 046
9000: [arXiv:hep-th/0406137].
9001:
9002:
9003: \bibitem{Gukov}
9004: S.~Gukov,
9005: ``Solitons, superpotentials and calibrations,''
9006: Nucl.\ Phys.\ B {\bf 574} (2000) 169
9007: [arXiv:hep-th/9911011].
9008:
9009:
9010: \bibitem{AS}
9011: B.~S.~Acharya and B.~Spence,
9012: ``Flux, supersymmetry and M theory on 7-manifolds,''
9013: arXiv:hep-th/0007213.
9014:
9015:
9016:
9017:
9018:
9019: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
9020: %
9021: % Chapter 6
9022: %
9023: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
9024:
9025: \bibitem{DGNV}
9026: R.~Dijkgraaf, S.~Gukov, A.~Neitzke and C.~Vafa,
9027: ``Topological M-theory as unification of form theories of gravity,''
9028: arXiv:hep-th/0411073.
9029:
9030:
9031: \bibitem{Nekrasov}
9032: N.~Nekrasov,
9033: ``A la recherche de la m-theorie perdue. Z theory: Chasing m/f theory,''
9034: arXiv:hep-th/0412021.
9035:
9036:
9037: \bibitem{CS}
9038: S.~Chiossi, S.~Salamon,
9039: ``The intrinsic torsion of SU(3) and $G_2$ structures,''
9040: Differential Geometry, Valencia 2001, World Sci. Publishing, 2002, pp 115-133
9041: [arXiv:math.DG/0202282].
9042:
9043:
9044: \bibitem{CCDLM}
9045: G.~L.~Cardoso, G.~Curio, G.~Dall'Agata, D.~L\"ust, P.~Manousselis and
9046: G.~Zoupanos,
9047: ``Non-K\"ahler string backgrounds and their five torsion classes,''
9048: Nucl.\ Phys.\ B {\bf 652} (2003) 5
9049: [arXiv:hep-th/0211118]
9050:
9051:
9052:
9053: \bibitem{DLM}
9054: M.~J.~Duff, J.~T.~Liu and R.~Minasian,
9055: ``Eleven-dimensional origin of string / string duality: A one-loop test,''
9056: Nucl.\ Phys.\ B {\bf 452} (1995) 261
9057: [arXiv:hep-th/9506126].
9058: %%CITATION = HEP-TH 9506126;%%
9059:
9060:
9061: \bibitem{SVW}
9062: S.~Sethi, C.~Vafa and E.~Witten,
9063: ``Constraints on low-dimensional string compactifications,''
9064: Nucl.\ Phys.\ B {\bf 480}, 213 (1996)
9065: [arXiv:hep-th/9606122].
9066:
9067:
9068:
9069:
9070:
9071:
9072: \bibitem{Joyce}
9073: D.~D.~Joyce,
9074: ``Compact Riemannian 7-manifolds with Holonomy $G_2$, I./II.''
9075: J. Diff. Geom. {\bf 43} (1996) 291-328
9076:
9077:
9078:
9079:
9080: \bibitem{BJ}
9081: B.~S.~Acharya and B.~Spence,
9082: ``Flux, supersymmetry and M theory on 7-manifolds,''
9083: arXiv:hep-th/0007213;\\
9084: K.~Behrndt and C.~Jeschek,
9085: ``Fluxes in M-theory on 7-manifolds: G-structures and superpotential,''
9086: Nucl.\ Phys.\ B {\bf 694} (2004) 99
9087: [arXiv:hep-th/0311119];\\
9088: T.~House and A.~Micu,
9089: ``M-theory compactifications on manifolds with G(2) structure,''
9090: arXiv:hep-th/0412006;\\
9091: N.~Lambert,
9092: ``Flux and Freund-Rubin superpotentials in M-theory,''
9093: arXiv:hep-th/0502200.
9094:
9095: \bibitem{Witt}
9096: C.~Jeschek and F.~Witt,
9097: ``Generalised G(2)-structures and type IIB superstrings,''
9098: JHEP {\bf 0503} (2005) 053
9099: [arXiv:hep-th/0412280];\\
9100: F.~Witt,
9101: ``Generalised $G_2$-manifolds,''
9102: arXiv:math.dg/0411642.
9103:
9104: \bibitem{Douglas}
9105: M.~R.~Douglas,
9106: ``The statistics of string / M theory vacua,''
9107: JHEP {\bf 0305} (2003) 046
9108: [arXiv:hep-th/0303194].
9109:
9110: \bibitem{OVV}
9111: H.~Ooguri, C.~Vafa and E.~Verlinde,
9112: ``Hartle-Hawking wave-function for flux compactifications,''
9113: arXiv:hep-th/0502211.
9114:
9115:
9116: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
9117: % section 6
9118:
9119:
9120:
9121:
9122:
9123:
9124: %%%%%%%%%%%%%%%%%%%%%%%%%%
9125: % Appendices
9126:
9127:
9128: \bibitem{CRTV}
9129: B.~Craps, F.~Roose, W.~Troost and A.~Van Proeyen,
9130: ``What is special K\"ahler geometry?,''
9131: Nucl.\ Phys.\ B {\bf 503} (1997) 565
9132: [arXiv:hep-th/9703082].
9133:
9134:
9135: \bibitem{Freed}
9136: D.~S.~Freed,
9137: ``Special K\"ahler manifolds,''
9138: Commun.\ Math.\ Phys.\ {\bf 203}, 31 (1999)
9139: [arXiv:hep-th/9712042].
9140:
9141:
9142:
9143:
9144:
9145: %\bibitem{Acharya}
9146: % B.~S.~Acharya,
9147: % ``A moduli fixing mechanism in M theory,''
9148: % arXiv:hep-th/0212294.
9149: % %%CITATION = HEP-TH 0212294;%%
9150:
9151:
9152:
9153:
9154: \end{thebibliography}
9155: \end{document}
9156:
9157:
9158:
9159:
9160:
9161:
9162: