1: \documentclass[11pt]{article}
2: %\documentclass{casabs}
3: \usepackage{amssymb}\usepackage{amsmath}
4: %\usepackage{latexsym}
5: \usepackage[all]{xy}
6: \usepackage[dvips]{graphicx}
7: \usepackage{pstricks}
8: \usepackage{pst-plot}
9:
10: %\usepackage{showlabels}
11:
12: \textwidth = 6.1 in
13: \textheight = 9in
14: \oddsidemargin = 0.0 in
15: \evensidemargin = 0.0 in
16: \topmargin = 0.0 in
17: \headheight = 0.0 in
18: \headsep = 0.0 in
19:
20:
21: \numberwithin{equation}{section}
22:
23: \def\be{\begin{equation}}
24: \def\ee{\end{equation}}
25: \def\ba{\begin{align}}
26: \def\ea{\end{align}}
27:
28: \def\p{\partial}
29:
30: \def\yboxit#1#2{\vbox{\hrule height #1 \hbox{\vrule width #1
31: \vbox{#2}\vrule width #1 }\hrule height #1 }}
32: \def\fillbox#1{\hbox to #1{\vbox to #1{\vfil}\hfil}}
33: \def\ybox{{\lower 1.3pt \yboxit{0.4pt}{\fillbox{8pt}}\hskip-0.2pt}}
34: %
35: % degree of W
36: \def\d{\kappa}
37: \def\hSigma{{\hat\Sigma}}
38: %
39: \def\CY#1{{\rm CY$_{#1}$}}
40: \def\Coh{{\bf Coh\ }}
41: \def\Cat{{\bf Cat}}
42: \def\l{\left}
43: \def\r{\right}
44: \def\mapr{\mathop{\longrightarrow}\limits}
45: \def\grade{\varphi}
46: \def\comments#1{}
47: \def\cc{{\rm c.c.}}
48: \def\cof{{\rm cof~}}
49: \def\Nbar{{\bar N}}
50: \def\NRR{{N_{RR}}}
51: \def\NNS{{N_{NS}}}
52: \def\tM{\tilde M}
53: \def\ib{{\bar i}}
54: \def\jb{{\bar j}}
55: \def\kb{{\bar k}}
56: \def\lb{{\bar l}}
57: \def\mb{{\bar m}}
58: \def\nb{{\bar n}}
59: \def\ab{{\bar a}}
60: \def\bb{{\bar b}}
61: \def\cb{{\bar c}}
62: \def\Cb{{\bar C}}
63: \def\db{{\bar d}}
64: \def\Dbar{{\bar D}}
65: \def\zb{{\bar z}}
66: \def\Zb{{\bar Z}}
67: \def\phib{{\bar \phi}}
68: \def\taub{{\bar \tau}}
69: \def\QC{\Bbb{C}}
70: \def\QH{\Bbb{H}}
71: \def\QM{\Bbb{M}}
72: \def\QR{\Bbb{R}}
73: \def\QX{\Bbb{X}}
74: \def\QZ{\Bbb{Z}}
75: \def\p{\partial}
76: \def\delbar{{\bar\partial}}
77: \def\tilp{\tilde\partial}
78: \def\eps{\epsilon}
79: \def\half{\frac{1}{ 2}}
80: \def\Tr{{{\rm Tr~ }}}
81: \def\tr{{\rm tr\ }}
82: \def\Re{{\rm Re\hskip0.1em}}
83: \def\Im{{\rm Im\hskip0.1em}}
84: \def\even{{\rm even}}
85: %\def\odd{{\rm odd}}
86: \def\lcm{{\rm lcm}}
87: \def\diag{{\rm diag}}
88: \def\bra#1{{\langle}#1|}
89: \def\ket#1{|#1\rangle}
90: \def\bbra#1{{\langle\langle}#1|}
91: \def\kket#1{|#1\rangle\rangle}
92: \def\vev#1{\langle{#1}\rangle}
93: \def\bigvev#1{\bigg\langle{#1}\bigg\rangle}
94: \def\Dslash{\rlap{\hskip0.2em/}D}
95: \def\intersect{\cdot}
96: \def\cA{{\cal A}}
97: \def\cC{{\cal C}}
98: \def\cD{{\cal D}}
99: \def\cE{{\cal E}}
100: \def\CF{{\cal F}}
101: \def\cF{{\cal F}}
102: \def\cG{{\cal G}}
103: \def\cH{{\cal H}}
104: \def\cI{{\cal I}}
105: \def\cK{{\cal K}}
106: \def\cL{{\cal L}}
107: \def\cM{{\cal M}}
108: \def\cN{{\cal N}}
109: \def\cO{{\cal O}}
110: \def\cP{{\cal P}}
111: \def\cQ{{\cal Q}}
112: \def\cS{{\cal S}}
113: \def\cT{{\cal T}}
114: \def\cU{{\cal U}}
115: \def\cV{{\cal V}}
116: \def\cW{{\cal W}}
117: \def\cX{{\cal X}}
118: \def\Det{{\rm Det}}
119: \def\CP{{\cal P}}
120: \def\CM{{\cal M}}
121: \def\CN{{\cal N}}
122: \def\CR{{\cal R}}
123: \def\CD{{\cal D}}
124: \def\CF{{\cal F}}
125: \def\CJ{{\cal J}}
126: \def\CP{{\cal P }}
127: \def\CL{{\cal L}}
128: \def\CV{{\cal V}}
129: \def\CO{{\cal O}}
130: \def\CZ{{\cal Z}}
131: \def\CE{{\cal E}}
132: \def\CG{{\cal G}}
133: \def\CH{{\cal H}}
134: \def\CC{{\cal C}}
135: \def\CB{{\cal B}}
136: \def\CS{{\cal S}}
137: \def\CA{{\cal A}}
138: \def\CC{{\cal C}}
139: \def\CK{{\cal K}}
140: \def\CP{{\cal P }}
141: \def\CQ{{\cal Q }}
142: \def\CE{{\cal E }}
143: \def\CV{{\cal V }}
144: \def\CZ{{\cal Z }}
145: \def\CS{{\cal S }}
146: \def\CY{{\cal Y }}
147: \def\CW{{\cal W}}
148: \def\CI{{\cal I}}
149: \def\ad#1#2{\frac{\delta}{\delta\sigma^{#1}(#2)}}
150: \def\ppt{\frac{\partial}{\partial t}}
151: \def\comment#1{[#1]}
152: \def\nl{\hfill\break}
153: \def\ap{\alpha'}
154: \def\floor#1{{#1}}
155: \def\sgn{{\rm sgn\ }}
156: \def\P{\BP}
157: \def\I{I}
158: \def\IA{IA}
159: \def\II{\relax{I\kern-.10em I}}
160: \def\IIa{{\II}a}
161: \def\IIb{{\II}b}
162: \def\TeV{{\rm TeV}}
163: \def\AdS{{\rm AdS}}
164: %
165: \def\imp{$\Rightarrow$}
166: \def\IZ{\relax\ifmmode\mathchoice
167: {\hbox{\cmss Z\kern-.4em Z}}{\hbox{\cmss Z\kern-.4em Z}}
168: {\lower.9pt\hbox{\cmsss Z\kern-.4em Z}}
169: {\lower1.2pt\hbox{\cmsss Z\kern-.4em Z}}\else{\cmss Z\kern-.4em
170: Z}\fi}
171: \def\IB{\relax{\rm I\kern-.18em B}}
172: \def\IC{{\relax\hbox{$\inbar\kern-.3em{\rm C}$}}}
173: \def\ID{\relax{\rm I\kern-.18em D}}
174: \def\IE{\relax{\rm I\kern-.18em E}}
175: \def\IF{\relax{\rm I\kern-.18em F}}
176: \def\IG{\relax\hbox{$\inbar\kern-.3em{\rm G}$}}
177: \def\IGa{\relax\hbox{${\rm I}\kern-.18em\Gamma$}}
178: \def\IH{\relax{\rm I\kern-.18em H}}
179: \def\II{\relax{\rm I\kern-.18em I}}
180: \def\IK{\relax{\rm I\kern-.18em K}}
181: \def\IN{\relax{\rm I\kern-.18em N}}
182: \def\IP{\relax{\rm I\kern-.18em P}}
183: %\def\IX{\relax{\rm X\kern-.01em X}}
184: %this doesn't work
185: \def\IX{{\bf X}}
186: %
187: \def\inbar{\,\vrule height1.5ex width.4pt depth0pt}
188: \def\mod{{\rm\; mod\;}}
189: \def\ndt{\noindent}
190: \def\p{\partial}
191: \def\pab{\pb_{\bar A} }
192: \def\pb{{\bar \p}}
193: \def\pgp{\pb g g^{-1}}
194: \font\cmss=cmss10 \font\cmsss=cmss10 at 7pt
195: \def\IR{\relax{\rm I\kern-.18em R}}
196: \def\pbar{\bar{\p}}
197: \def\qmvw{\CM_{\vec \zeta}(\vec v, \vec w) }
198: \def\sdtimes{\mathbin{\hbox{\hskip2pt\vrule
199: height 4.1pt depth -.3pt width .25pt\hskip-2pt$\times$}}}
200: \def\im{{\rm im\ }}
201: \def\ker{{\rm ker\ }}
202: \def\cok{{\rm cok\ }}
203: \def\End{{\rm End\ }}
204: \def\vol{{\rm vol}}
205: \def\id {{\bf 1}}
206: \def\ch{{\rm ch}}
207: \def\chern{{\rm c}}
208: \def\rank{{\rm rank}}
209: \def\aroof{{\rm \hat A}}
210: %
211: \def\CP{\IP}
212: \def\CM{{\cal M}}
213: \def\CN{{\cal N}}
214: \def\CZ{{\cal Z}}
215: %
216: \def\BR{\IR}
217: \def\BZ{\IZ} % for now
218: \def\BP{\IP}
219: \def\BR{\IR}
220: \def\BC{\IC}
221: \def\BM{\QM}
222: \def\BH{\QH}
223: \def\BX{\QX}
224: %
225: \def\ls{l_s}
226: \def\ms{m_s}
227: \def\gs{g_s}
228: \def\lp10{l_P^{10}}
229: \def\lp11{l_P^{11}}
230: \def\R11{R_{11}}
231:
232: \def\ch{{\rm ch}}
233:
234: \def\codim{{\mathop{\rm codim}}}
235: \def\cok{{\rm cok}}
236: \def\coker{{\mathop {\rm coker}}}
237:
238: \def\diff{{\rm diff}}
239: \def\Diff{{\rm Diff}}
240: \def\Det{{\rm Det}}
241: \def\Tr{{\rm Tr}}
242:
243: \def\gpg{g^{-1} \p g}
244: \def\ib{{\bar i}}
245: \def\zb {\bar{z}}
246: \def\xb {\bar{x}}
247: \def\mb {\bar{m}}
248: \def\Jb {\bar{J}}
249: \def\Xb {\bar{X}}
250: \def\Hb {\bar{H}}
251: \def\Gb {\bar{G}}
252: \def\jb {\bar{j}}
253: \def\psib{\bar{\psi}}
254: \def\wb {\bar{w}}
255: \def\qb {\bar{q}}
256: \def\G{\Gamma}
257:
258: %% MORE MACROS
259:
260: \font\manual=manfnt
261: \def\dbend{\lower3.5pt\hbox{\manual\char127}}
262: \def\danger#1{\smallskip\noindent\rlap\dbend%
263: \indent{\bf #1}\par\vskip-1.5pt\nobreak}
264: \def\c{\cdot}
265: \def\IZ{\relax\ifmmode\mathchoice
266: {\hbox{\cmss Z\kern-.4em Z}}{\hbox{\cmss Z\kern-.4em Z}}
267: {\lower.9pt\hbox{\cmsss Z\kern-.4em Z}} {\lower1.2pt\hbox{\cmsss
268: Z\kern-.4em Z}}\else{\cmss Z\kern-.4em Z}\fi}
269: \def\half {\frac{1}{ 2}}
270: \def\sdtimes{\mathbin{\hbox{\hskip2pt\vrule height 4.1pt depth -.3pt
271: width .25pt \hskip-2pt$\times$}} }
272:
273: \def\p{\partial}
274: \def\pb{\bar{\partial}}
275: \def\bar{\overline}
276: \def\CS{{\cal S}}
277: \def\CN{{\cal N}}
278: \def\CH{{\cal H}}
279: \def\rt2{\sqrt{2}}
280: \def\irt2{\frac{1}{\sqrt{2}}}
281: \def\rtt{\sqrt{3}}
282: \def\hrtt{\frac{\sqrt{3}}{ 2}}
283: \def\t{\tilde}
284: \def\ndt{\noindent}
285: \def\s{\sigma}
286: \def\b{\beta}
287: \def\a{\alpha}
288: \def\w{\omega}
289: \def\psid{\psi^{\dagger}}
290:
291: \def\lieg{{\bf \underline{g}}}
292: \def\liet{{\bf \underline{t}}}
293: \def\mod{{\rm mod}}
294: \def\Det{{\rm Det}}
295:
296:
297: %% MORE MACROS
298: \font\cmss=cmss10
299: \font\cmsss=cmss10 at 7pt
300: \def\IL{\relax{\rm I\kern-.18em L}}
301: \def\IH{\relax{\rm I\kern-.18em H}}
302: \def\IR{\relax{\rm I\kern-.18em R}}
303: \def\inbar{\vrule height1.5ex width.4pt depth0pt}
304: \def\IC{\relax\hbox{$\inbar\kern-.3em{\rm C}$}}
305: \def\rlx{\relax\leavevmode}
306: %%
307: \def\ZZ{\rlx\leavevmode\ifmmode\mathchoice{\hbox{\cmss Z\kern-.4em Z}}
308: {\hbox{\cmss Z\kern-.4em Z}}{\lower.9pt\hbox{\cmsss Z\kern-.36em Z}}
309: {\lower1.2pt\hbox{\cmsss Z\kern-.36em Z}}\else{\cmss Z\kern-.4em
310: Z}\fi}
311: %%% misc.
312: \def\IZ{\relax\ifmmode\mathchoice
313: {\hbox{\cmss Z\kern-.4em Z}}{\hbox{\cmss Z\kern-.4em Z}}
314: {\lower.9pt\hbox{\cmsss Z\kern-.4em Z}}
315: {\lower1.2pt\hbox{\cmsss Z\kern-.4em Z}}\else{\cmss Z\kern-.4em
316: Z}\fi}
317:
318: %% MORE MACROS
319: \def\codim{{\mathop{\rm codim}}}
320: \def\cok{{\rm cok}}
321: \def\coker{{\mathop {\rm coker}}}
322:
323: \def\ch{{\rm ch}}
324: \def\diff{{\rm diff}}
325: \def\Diff{{\rm Diff}}
326: \def\Det{{\rm Det}}
327: \def\Tr{{\rm Tr}}
328: \def\gpg{g^{-1} \p g}
329: \def\ib{{\bar i}}
330: \def\zb {\bar{z}}
331: \def\nb {\bar{n}}
332: \def\cb {\bar{c}}
333: \def\wb {\bar{w}}
334: \def\qb {\bar{q}}
335: \def\G{\Gamma}
336:
337:
338: \font\manual=manfnt
339: \def\dbend{\lower3.5pt\hbox{\manual\char127}}
340: \def\danger#1{\smallskip\noindent\rlap\dbend%
341: \indent{\bf #1}\par\vskip-1.5pt\nobreak}
342: \def\c{\cdot}
343: \def\IZ{\relax\ifmmode\mathchoice
344: {\hbox{\cmss Z\kern-.4em Z}}{\hbox{\cmss Z\kern-.4em Z}}
345: {\lower.9pt\hbox{\cmsss Z\kern-.4em Z}} {\lower1.2pt\hbox{\cmsss
346: Z\kern-.4em Z}}\else{\cmss Z\kern-.4em Z}\fi}
347: \def\half {\frac{1}{ 2}}
348: \def\sdtimes{\mathbin{\hbox{\hskip2pt\vrule height 4.1pt depth -.3pt
349: width .25pt \hskip-2pt$\times$}} }
350:
351: \def\pb{\bar{\partial}}
352: \def\bar{\overline}
353:
354: \def\rt2{\sqrt{2}}
355: \def\irt2{\frac{1}{\sqrt{2}}}
356: \def\rtt{\sqrt{3}}
357: \def\hrtt{\frac{\sqrt{3}}{ 2}}
358: \def\t{\tilde}
359: \def\T{\widetilde}
360: \def\ndt{\noindent}
361: \def\s{\sigma}
362: \def\bB {{\bf B}}
363: \def\sNSNS{\text{\tiny{NSNS}}}
364: \def\sRR{\text{\tiny{RR}}}
365: \def\sNS{\text{\tiny{NS}}}
366: \def\sR{\text{\tiny{R}}}
367: \def\sId{\text{\tiny{Id}}}
368: \def\scont{\text{\tiny{cont}}}
369: \def\sdisc{\text{\tiny{disc}}}
370:
371: \input{epsf}
372:
373: \usepackage{epsfig}
374:
375: \title{\Large{\bf Non-compact Gepner Models, Landau-Ginzburg Orbifolds and
376: Mirror Symmetry}} \author{Sujay K. Ashok$^{a,b}$, Raphael Benichou$^{c}$
377: and Jan Troost$^{c}$ } \date{}
378: \begin{document}
379: \maketitle
380:
381: \begin{center}
382: $^{a}$Institute of Mathematical Sciences\\
383: C.I.T Campus, Taramani\\
384: Chennai, India 600113\\
385: $^{b}$Perimeter Institute for Theoretical Physics\\
386: Waterloo, Ontario, ON N$2$L$2$Y$5$, Canada \\
387: $^{c}$Laboratoire de Physique Th\'eorique\footnote{Unit\'e Mixte du CRNS et
388: de l'Ecole Normale Sup\'erieure associ\'ee \`a l'universit\'e Pierre et
389: Marie Curie 6, UMR
390: 8549. Preprint LPTENS-07/51.}, Ecole Normale Sup\'erieure \\
391: $24$ Rue Lhomond Paris $75005$, France
392: \end{center}
393:
394: \begin{abstract}
395: We study non-compact Gepner models that preserve sixteen
396: or eight
397: supercharges in type II string theories. In particular, we develop an
398: orbifolded Landau-Ginzburg description of these models analogous to the
399: Landau-Ginzburg formulation of compact Gepner models. The Landau-Ginzburg
400: description provides an easy and direct access to the geometry of the
401: singularity associated to the non-compact Gepner models. Using these tools,
402: we are able to give an intuitive account of the chiral rings of the models,
403: and of the massless moduli in particular. By studying orbifolds of the
404: singular linear dilaton models, we describe mirror pairs of non-compact
405: Gepner models by suitably adapting the Greene-Plesser
406: construction of
407: mirror pairs for the compact case. For particular models, we
408: take a large level, low curvature limit in which we can
409: analyze corrections to a flat space orbifold approximation of the
410: non-compact Gepner models. This gives rise to a counting of moduli which
411: differs from the toric counting in a subtle way.
412: \end{abstract}
413:
414: \section{Introduction}
415: Mirror symmetry for Calabi-Yau $3$-folds is a subject of great interest
416: to physicists as well as mathematicians \cite{Yau:1998yt}\cite{Greene:1997ty}.
417: Mirror pairs were first exhibited \cite{Greene:1990ud}
418: by studying orbifolds of the quintic at the Gepner point \cite{Gepner:1987qi}
419: in the Calabi-Yau moduli space. This was explored in more detail in
420: \cite{Candelas:1990rm} which contained the first mathematical predictions from
421: mirror symmetry. Many generalizations were found using the Landau-Ginzburg formulation that were applicable in non-geometric regimes \cite{Berglund:1991pp}.
422:
423: In this paper, we wish to extend the study of mirror symmetry to non-compact
424: Gepner models. Various approaches towards this problem have been followed.
425: Toric Calabi-Yau manifolds have been very well studied in the literature
426: following the construction of mirror pairs for hypersurfaces in toric
427: varieties \cite{Batyrev} and for non-compact Calabi-Yau manifolds (see
428: \cite{Chiang:1999tz} and references therein). In the physics literature, this
429: has been reformulated using a gauged linear sigma model \cite{Witten:1993yc,
430: Hori:2000kt, Hori:2002cd}. Progress has also been made for more general
431: $c=9$ theories via conformal field theory techniques \cite{Eguchi:2004yi}. In
432: particular, the work on T-duality for the $N=2$ cigar and the Liouville
433: conformal field theories is relevant in this context \cite{Giveon:1999px,
434: Hori:2001ax, Tong:2003ik, Israel:2004jt}.
435:
436: One of the gaps we aim to fill in this paper is to clarify the link between
437: non-compact Gepner models and their Landau-Ginzburg description. For compact
438: Calabi-Yau manifolds that are hypersurfaces in weighted projective space, the
439: Landau-Ginzburg description is closely tied to the geometry
440: \cite{Greene:1988ut}. These are two of the many phases in a gauged linear
441: sigma-model description \cite{Witten:1993yc}. The Landau-Ginzburg model is also extremely useful for providing a simple setting to do computations. For instance, it has served well in the
442: past for establishing the link between compact Gepner models and compact
443: Calabi-Yau manifolds \cite{Greene:1988ut} as well as to provide the first list
444: of Calabi-Yau hypersurfaces in $W\IC\IP^4$ \cite{Candelas:1989hd}.
445:
446: Our approach in this paper is to stress the ingredients of the analysis
447: of Gepner models that survive the transition from going from the compact to
448: the non-compact case. We alo point out the characteristics that differ
449: in the two cases.
450: Moreover we study in detail the orbifolded Landau-Ginzburg description of these models
451: \cite{Vafa:1989xc, Intriligator:1990ua}. This allows us to
452: compare our models with geometric backgrounds which can be used as
453: internal spaces for type II strings on $\IR^{3,1}$.
454:
455: In section \ref{noncompactgepner} we will review the asymptotic
456: partition function of non-compact Gepner models \cite{Eguchi:2000tc,Hanany:2002ev,Murthy:2003es,Eguchi:2004yi,Israel:2004ir}. The asymptotic partition
457: function, which is proportional to the divergent volume of space-time,
458: otherwise behaves much as
459: the partition function in the compact case. In particular we show here that
460: the $\beta$-method of Gepner \cite{Gepner:1987qi} for constructing modular
461: invariants can be adapted to the non-compact case. We then start the
462: discussion of the localized spectrum
463: \cite{Hanany:2002ev,Eguchi:2004yi,Israel:2004ir}, which is
464: characteristic of non-compact models.
465:
466: In section \ref{LG} we will link the non-compact Gepner models to
467: Landau-Ginzburg models. We briefly remark on the difference between the
468: compact and non-compact Landau-Ginzburg model \cite{Ooguri:1995wj}. For
469: instance, the former gives rise to a unital chiral ring, while the latter
470: gives rise to a ring without unit element. We continue the analysis in
471: section \ref{orbifoldedLG} with a discussion of the orbifolds of the
472: Landau-Ginzburg models that are necessary to implement the GSO projection,
473: and other orbifold groups. We will see that the formalism for counting chiral
474: ring elements largely carries over from the compact case
475: \cite{Vafa:1989xc,Intriligator:1990ua}. However, it will become intuitive
476: that the Landau-Ginzburg potentials with negative powers exclude some of the
477: (anti-)chiral ring elements as the potential renders them non-normalizable.
478: Continuing the formal counting exercise will turn out to be useful
479: nevertheless. It gives rise to a natural picture of mirror symmetry in
480: conformal field theories, that is strongly reminiscent of its compact
481: counterpart \cite{Greene:1990ud, Eguchi:2004yi}. We will discuss this point
482: in section \ref{CFTmirror}.
483:
484: In section \ref{examples} we provide new examples of mirror conformal field
485: theories. The techniques developed to analyze orbifolded Landau-Ginzburg
486: models come in handy when treating these more complicated models. The mirror
487: pairs of conformal field theories are best understood as linear dilaton
488: backgrounds with $N=2$ superconformal symmetry. These %in turn
489: can be deformed or resolved, to give rise to perturbatively well-defined mirror string backgrounds. Finally, in section \ref{geometrymirror}, we discuss examples in which we can relate the mirror conformal field theories to geometries. It will turn out that we can approximate certain conformal field theories at large level with orbifold singularities that admit a toric description. At infinite level, we find agreement between the conformal field theory and the geometric results. At finite level, we find that the conformal field theory description takes into account various modifications to the background that can lift some moduli. In section \ref{conclusions} we discuss further applications of our results.
490:
491: \section{Non-compact Gepner Models}
492: \label{noncompactgepner}
493: We start out with a rather brief but technical review of non-compact Gepner models (see
494: e.g. \cite{Eguchi:2000tc,Murthy:2003es,Eguchi:2004yi,Israel:2004ir})
495: in order to clarify the fact that the asymptotic partition function of
496: non-compact Gepner models can be constructed using the same
497: tools as in the compact case.
498: We can then lay bare properties of the models along the same lines as in the
499: compact Gepner models. In the second part of this section we give a preview
500: of the ingredients that go into analyzing the localized part of the partition
501: function.
502:
503: \subsection{The Asymptotic Non-compact Gepner Models}
504: \label{asymptotic}
505:
506:
507: For simplicity we will work with type II string theory on $\IR^{3,1}$ times an
508: internal (non-compact) conformal field theory of central charge
509: nine\footnote{The constructions can be extended to lower-dimensional flat
510: spaces and to heterotic string theories, with little effort and lots of
511: indices.}. The internal conformal field theory is built from a product of
512: $N=2$ superconformal field theories. These can be split into three classes
513: depending on whether their central charge is smaller, larger
514: than or equal to
515: three.
516:
517: \begin{itemize}
518: \item[\_]The minimal $N=2$ superconformal field theories (see
519: e.g. \cite{Gepner:1986hr})
520: have central
521: charge smaller than three. They can be
522: viewed as coset conformal field theories of the form $SU(2)_{k-2} \times
523: U(1)_2 / U(1)_k$ of central charge $c_{MM}=3 -\frac{6}{k}$. The level $k$ is
524: the supersymmetric level of the total $SU(2)$ current algebra present in the
525: parent $N=1$ Wess-Zumino-Witten model. It is a positive integer
526: larger than or equal to two.
527:
528: Since the minimal $N=2$ superconformal models
529: have been reviewed frequently
530: (\cite{Greene:1996cy,Maldacena:2001ky}), and since they are
531: standard in the construction of Gepner models \cite{Gepner:1987qi}, we only
532: very briefly recall some of their properties. The primaries of the model can
533: be labeled by three quantum numbers: the spin $j$ under the $SU(2)_{k-2}$
534: current algebra, the $\IZ_{2k}$ valued chiral momentum $n$ under the $U(1)_k$
535: current algebra and the $\IZ_4$ valued chiral momentum $s$ labeling a
536: representation of the $U(1)_2$ current algebra. They satisfy the selection
537: rule: $2j\equiv n+s\ [2]$. We moreover have an equivalence between the
538: following representations: $(j,n,s) \equiv (\frac{k-2}{2}-j,n-k,s+2)$. The
539: left-moving $U(1)_R$ charge $Q_{MM}$ for a primary is equal to $Q_{MM}=
540: \frac{n}{k} + \frac{s}{2}$.
541:
542: \item[\_]The second class of theories has central charge
543: larger than three. An example in this class is an $N=2$ linear dilaton theory
544: with a slope such that the central charge is equal to $c=3+\frac{6}{l}$ with
545: positive and real values for the parameter $l$.
546: The superconformal algebra with central charge larger than three has
547: continuous unitary representations which are conveniently labeled
548: \cite{Dixon:1989cg} by a Casimir $j=1/2 + ip $ where $p \in \IR^+$, by an
549: integer momentum $2m \in \IZ$ and a $\IZ_4$ fermionic quantum number $s$. The
550: left-moving R-charge $Q_{nc}$ of a primary is $Q_{nc} = \frac{2m}{l}
551: +\frac{s}{2}$. (See e.g. \cite{Israel:2004jt} for a detailed discussion.)
552:
553: \item[\_]The $N=2$ superconformal
554: algebra with $c=3$ is exceptional, and can be represented for instance by free
555: scalars (which can realize compact or non-compact target space directions).
556: \end{itemize}
557:
558:
559: In order to mimic the Gepner construction, we will assume that in the case
560: $c>3$, the chiral
561: algebra has some more structure (see e.g. \cite{Eguchi:2000tc}). For simplicity we will
562: work under the assumption that the parameter $l$ is a positive integer\footnote{It is sufficient to suppose that the central
563: charge is parameterized by a positive fractional level $l$ \cite{Eguchi:2000tc}.}.
564: We can then add to
565: the chiral algebra the generator of spectral flow on the $N=2$ superconformal
566: algebra by $2 l$ units. The characters of the extended $N=2$ superconformal
567: algebra in the continuous representations are given by:
568: \begin{eqnarray}
569: Ch_{cont} (j,2m,s) &=& q^{\frac{p^2}{l}} \frac{1}{\eta^3(\tau)}
570: \Theta_{s,2}(\tau)
571: \Theta_{2m,l} (\tau).
572: \end{eqnarray}
573: It is crucial to us that the modular transformation properties of the
574: characters hinge upon the presence of the
575: $\theta$-functions at levels $2$ and $l$.
576:
577:
578: \subsubsection{Levels And Charge Lattice}
579:
580: We observe that the characters of the $N=2$ minimal models at
581: level $k$ transform as $\theta$-functions at levels $2$ and $-k$,
582: while the extended characters of the $N=2$ models with central charge
583: $c=3+\frac{6}{l}$ transform as $\theta$-functions at level $2$ and
584: $+l$. We thus note a first important sign difference
585: in the transformation rules of the characters. Modular
586: invariants in the continuous sector of the theory can be based on
587: modular invariants of $\theta$-functions. For one $U(1)_k$ current algebra
588: at level $k$
589: these are well-known to correspond to the divisors of $k$ via
590: orbifolding of the diagonal modular invariant. For a product of
591: $\theta$-functions, the analysis is more complicated,
592: but a large class of modular invariants can be constructed by
593: taking products of modular invariants of the factors, and then
594: orbifolding.
595:
596: In order to write down the modular invariant partition functions,
597: it is useful to introduce a charge lattice for the various $U(1)$
598: current algebras in
599: the theory. We introduce a vector of levels
600: $(2,2, \dots, 2; k_1, \dots , k_p ; l_1, \dots, l_q)$
601: where $p$ is the number of
602: minimal model factors, $q$ is the number of non-compact factors, while
603: the number of fermionic levels equal to $2$ is $1+q+p$. Indeed, we work
604: in light-cone gauge on $\IR^{3,1}$ such that there is one complex fermion
605: associated to the flat space directions (and there is
606: one complex fermion per factor
607: model).
608: The charge lattice is periodic. In each direction, the periodicity is twice
609: the level.
610: A point in the lattice is defined by a charge vector $r= (s_0,s_1, \dots,
611: s_{p+q}; n_1,\dots,n_p; 2m_1, \dots, 2m_q)$ where we used the
612: notation
613: $s_0$ for the charge of the flat
614: space fermions under the $U(1)_2$ current algebra, and
615: similarly for the other fermions, while we copy
616: the traditional notation for the
617: chiral momentum quantum numbers of compact and non-compact factors
618: that we introduced above (including their normalization).
619: The scalar product on the charge lattice is defined as follows:
620: \begin{eqnarray}
621: r^{(1)} \cdot r^{(2)} &=& -\frac{s_0^{(1)} s_0^{(2)}}{4} - \frac{s_1^{(1)} s_1^{(2)}}{4} - \dots
622: - \frac{s_{p+q}^{(1)} s_{p+q}^{(2)}}{4}
623: \nonumber \\
624: & & + \frac{n_1^{(1)} n_1^{(2)}}{2k_1} + \dots + \frac{n_p^{(1)} n_p^{(2)}}{2k_p}
625: \nonumber \\
626: & & - \frac{(2m_1^{(1)})(2m_1^{(2)})}{2l_1} -\dots - \frac{(2m_q^{(1)})(2m_q^{(2)})}{2l_q}\,.
627: \end{eqnarray}
628: The contribution of the chiral momenta corresponding to the non-compact factors comes with an opposite sign from those of the compact factors. The all-important signature of the
629: quadratic form is therefore $(-,\dots,-; +, \dots, + ; - , \dots, -)$.
630:
631: \subsubsection{A Canonical Vector}
632:
633: We introduce the following vector $\beta_0$ in the charge lattice\footnote{We
634: take the entries for the fermions to be minus one, in order to accord with
635: the convention that the left-moving $U(1)_R$ charge
636: is given for instance for a minimal
637: model factor by $Q_{MM}=\frac{n}{k}+\frac{s}{2}$.}:
638: \begin{eqnarray}
639: \beta_0 &=& (-1,-1, \dots, -1; 1,1,\dots,1; -1,-1, \dots,-1).
640: \end{eqnarray}
641: We have that the left-moving R-charge $Q$ for a primary state with charge
642: vector $r$ is equal to $Q=2 \beta_0 \cdot r$. We moreover have that
643: \begin{eqnarray}
644: \beta_0 \cdot \beta_0 &=& -\frac{1+p+q}{4} + \sum_{i=1}^p \frac{1}{2k_i} - \sum_{j=1}^q \frac{1}{2l_i}
645: \nonumber \\
646: &=& -1
647: \end{eqnarray}
648: where we used that the total central charge of the light-cone gauge conformal field theory is equal to
649: \be
650: 12 = 3(1+p+q) - \sum_{i=1}^p \frac{6}{k_i} + \sum_{j=1}^q \frac{6}{l_i} \,.
651: \ee
652: In summary, the vector $\beta_0$ is useful to measure the R-charge, and
653: squares to minus one.
654:
655: For the right-movers we will always take identical conventions to the
656: left-movers. In particular, the $N=2$ superconformal algebras have the
657: same structure constants. We will also define the right-moving
658: R-charge of the minimal model factors to be $\tilde{Q}_{MM} =
659: \frac{\tilde{n}}{k} + \frac{\tilde{s}}{2}$ and for the non-compact factors
660: $\tilde{Q}_{nc}=\frac{2 \tilde{m}}{k} + \frac{\tilde{s}}{2}$, while the charge vector
661: for the right-movers is $\tilde{r} = ( \tilde{s}_0, \dots, ; \dots ;
662: \dots, 2 \tilde{m}_{p+q})$. So, for the right-movers as well we have
663: that the total $U(1)_R$ charge is given by $\tilde{Q}= 2 \beta_0 \cdot
664: \tilde{r}$ with the same vector $\beta_0$.
665:
666:
667: \subsubsection{Products of $\theta$-functions}
668: We define the following notation for the product of characters of the
669: flat space fermions, the minimal model and the non-minimal $N=2$
670: superconformal field theories. Since the characters transform like
671: $\theta$-functions, we introduce the symbol:
672: \begin{eqnarray}
673: \Theta_{r} (\tau) &=& \Theta_{s_0,2}(\tau) \prod_{i=1}^p \chi_{j_i,n_i,s_i} (\tau)
674: \prod_{i=1}^q Ch_{cont}(j_{p+i},2m_i,s_{p+i}) (\tau)
675: \end{eqnarray}
676: where $r$ is the total charge vector, and the first factor corresponds
677: to the flat space fermions, while the following $p$ factors correspond to
678: minimal model characters,
679: and the final $q$ factors to the non-compact continuous
680: extended characters. In the $\Theta$ symbol we have left implicit the
681: labels corresponding to the levels, as well as those corresponding to
682: the Casimirs of the minimal and non-minimal models. The important point is that
683: the $\Theta$-functions transform as a product of ordinary $\theta$-functions
684: under modular transformations.
685:
686: We introduce now the first modular invariant partition function which is the
687: diagonal modular invariant:
688: \begin{eqnarray}
689: Z_{diag} &=& \sum_{r} \Theta_{r} (\tau) \Theta_r (\bar{\tau})\,,
690: \end{eqnarray}
691: where the diagonal sum over $r$ is over all inequivalent charges in the charge
692: lattice. Implicitly, we take a diagonal A-type modular invariant for the
693: Casimir invariants $j$ for all factors\footnote{It would be interesting to
694: study the more general modular invariants of type D and/or type E for the
695: compact factors.}. We have suppressed the divergent non-compact volume factor
696: in the formula for the asymptotic partition function.
697:
698: \subsubsection{Locality Orbifold}
699:
700: Now that we have set-up our theory in a way which is very analogous to
701: \cite{Gepner:1987qi}, we can follow that reference closely. Along the way, we
702: reformulate a few minor points in a more modern orbifold language.
703:
704: Locality in string theory
705: only allows for fermions having either
706: Neveu-Schwarz (NS) or Ramond (R) boundary conditions for all factors simultaneously for the
707: left- or the right-movers. We implement that locality constraint by
708: orbifolding the diagonal partition function by a diagonal $(\IZ_2)^{p+q}$ group.
709: The i'th $\IZ_2$ acts on a state with charge vectors $r$ and $\tilde{r}$ as
710: $(-1)^{(s_0+s_i+\tilde{s}_0 + \tilde{s}_i)/2}$ for $i=1, \dots, p+q$. The action
711: can be summarized by introducing vectors $\beta_i$ which have a $2$ as
712: the first and the $i+1$'th entry for $i=1, \dots, p+q$.
713: Then the action can equivalently be written as $(-1)^{\beta_i \cdot (r+\tilde{r})}$.
714: On an untwisted state with charges $s_0$ and $s_i$, the action is the
715: multiplication by a phase $(-1)^{s_0+s_i}$.
716:
717: The states invariant under $(\IZ_2)^{p+q}$ will be purely NS or purely R for
718: the left-movers. Since the partition function was diagonal, the same condition
719: will hold for the right-movers. The orbifold also introduces twisted states.
720: These have left- and right-moving charge vectors that differ by multiples of
721: the vectors $\beta_i$ \cite{Gepner:1987qi}. Indeed, since the fermion number
722: $s_i$ is defined modulo four, this introduces a twisted state sector for each
723: $\IZ_2$ orbifold factor, and moreover, since $\beta_i^2=2$, we have that the
724: twisted states we introduce in this fashion are also orbifold invariant
725: \cite{Gepner:1987qi}. Each of the $\IZ_2$ orbifolds introduces twisted
726: states, which renders the sum over left- and right-moving fermion numbers $s_i$
727: and $\tilde{s_i}$ independent, except for the fact that they need to be of the
728: same parity as the flat space fermion numbers $s_0$ and $\tilde{s}_0$
729: respectively. The above prescription is equivalent to the standard orbifold
730: procedure and produces a new modular invariant partition function. Note that
731: the flat space fermion quantum numbers $s_0$ and $\tilde{s}_0$ are still of
732: equal parity. We therefore only have NSNS and RR states at this point. The
733: sum over the left- and right-moving worldsheet fermion numbers $s_0$ and
734: $\tilde{s}_0$ will become decoupled after performing a final $\IZ_2$ GSO
735: projection, thus introducing fermions.
736:
737:
738: \subsubsection{Integer R-charge Orbifold}
739: \label{integerR}
740:
741: The standard GSO projection in string theory is based on the fact that
742: the partition function only has integer R-charges. In order to ensure
743: this condition in a Gepner model, we perform yet another orbifold. The
744: orbifold action on a state with charge $r$ and $\tilde{r}$ will be $exp
745: (2 \pi i \beta_0 \cdot (r + \tilde{r}))$. We first note that $\beta_0
746: \cdot \beta_i$ is an integer, such that the action of the orbifold on
747: $\beta_i$ twisted sectors is identical to the action on $\beta_i$ untwisted
748: sectors. The order of the $2 \beta_0$ orbifold is therefore the order
749: of the operator $e^{ 2 \pi i Q } $ in the untwisted theory (where
750: $Q$ is the total left-moving R-charge in light-cone gauge)\footnote{The theory
751: we start out with obeys the charge relation $r=\tilde{r}$.
752: Since $\beta_0 \cdot \beta_i \in \IZ$, we
753: still have the relation $e^{ 2 \pi i (Q+\tilde{Q})/2 } = e^{ 2 \pi i Q } $ after the
754: $(\IZ^{p+q}_2)$ orbifold.}. The order of
755: that orbifold is the smallest common divisor $d$ of all the levels in the
756: theory (including the fermionic level $2$ when $p+q$ is even and
757: not when $p+q$ is odd).
758:
759: It is clear in the untwisted sector that the orbifold forces the
760: left-moving (and right-moving) $U(1)_R$ charge to be integer. There
761: are also $d-1$ twisted sectors which have charge vectors which differ
762: by multiples of $ 2 \beta_0$. Since $\beta_0^2=-1$, we have that
763: these twisted sectors also have integer left- and right-moving
764: R-charges. Thus, the orbifold has provided us with a new modular
765: invariant partition function with integer R-charges for both left-
766: and right-movers. If we introduce the lattice $\Lambda$ generated
767: by the vectors $\beta_i$ and $2 \beta_0$, then we can write the
768: partition function of the theory
769: as\footnote{This partition function for non-compact models
770: is the analogue of $Z_0$ in
771: \cite{Greene:1990ud} for compact Gepner models.}:
772: \begin{eqnarray}
773: Z_0 &=& \sum_{r-\tilde{r} \in \Lambda} \Theta_r (\tau) \Theta_{\tilde{r}} (\bar{\tau})\,,
774: \end{eqnarray}
775: where we restrict the sum to invariant states, namely states obeying the
776: conditions
777: $ r \cdot \beta_i \in \IZ$ (purely NS or purely R) and
778: $ r \cdot 2 \beta_0 \in \IZ$ (integer R-charges). Again we have left implicit the diagonal
779: A-type invariant for the minimal models as well as the diagonal integral over
780: radial momenta for the compact factors that diverges like the volume of space-time.
781:
782: \subsubsection{The Standard GSO Projection}
783:
784: The partition function now has integer R-charges for all states, and can be
785: GSO projected in the same manner as the flat space partition function. We
786: project onto odd R-charges, i.e. we satisfy the condition $2 \beta_0 \cdot r
787: \in 2 \IZ +1$ as well as $2 \beta_0 \cdot \tilde{r} \in 2 \IZ+1$.
788: The charge difference between left- and right-movers for the twisted states is
789: proportional to an odd multiple of $\beta_0$. The twisted sectors are the NS-R
790: and R-NS sectors of the theory, which correspond to space-time fermions, and
791: contribute negatively to the space-time partition function thus implementing
792: space-time statistics. For type II theories, we have obtained an
793: asymptotic supersymmetric partition function in the Gepner formalism. Strictly
794: speaking we have a type IIB partition function, since we have made no
795: distinction between left- and right-movers. It can be transformed easily into
796: a type IIA partition function by flipping the chirality of the final
797: GSO projection for the right-movers in the R-sector.
798:
799: \subsubsection{Discrete Symmetries}
800: \label{symGroup}
801: In this subsection, we can follow \cite{Greene:1990ud} closely
802: since we have set up our model as in the compact case.
803: It is interesting to single out a particular symmetry group of the
804: model.
805:
806: The symmetry of the compact models contains a $\IZ_{k_i} \times \IZ_{k_i}$
807: group. Only the diagonal $\IZ_{k_i}$ subgroup has a non-trivial action on a
808: model with diagonal spectrum.
809: This subgroup acts as
810: \be
811: \Phi_{r,\tilde{r}} \rightarrow exp(2 \pi i
812: (n_i+\tilde{n}_i)/ 2 k_i) \Phi_{r,\tilde{r}}\,.
813: \ee
814: We can introduce a vector
815: \be
816: \gamma_i=(0,...,0;0,...,0,2,0,...,0; 0,...,0)
817: \ee
818: where $2$ is in the $i$'th entry after the first semi-column. The vector codes the action of the symmetry
819: group as follows:
820: \be
821: \Phi_{r,\tilde{r}} \rightarrow exp(\pi i \gamma_i \cdot (r +
822: \tilde{r}))\Phi_{r,\tilde{r}}\,.
823: \ee
824: We have a similar action in the non-compact theories. There is
825: a symmetry group $\IZ_{l_i} \times\IZ_{l_i}$, of which only the diagonal subgroup acts nontrivially on the
826: states. In the full Gepner model, we can think of the product of the diagonal
827: subgroups
828: \be
829: D=\prod_{i=1}^p \IZ_{k_i} \times \prod_{j=1}^q \IZ_{l_j}
830: \ee
831: as mapped into the charge lattice via the maps $\gamma_{i=1, \dots, p+q}$.
832:
833: Now we define the operator $g_0$ that acts by multiplication by $\exp(2\pi i
834: \beta_0 \cdot(r+\tilde{r}))$. It generates a group $\IZ_d$, that we used
835: previously to perform the integer R-charge orbifold.
836: This orbifold group $\IZ_d$ contains a subgroup
837: generated by $g_0^2$, which is a subgroup of
838: $D$. Indeed, the
839: element $g_0^2$ corresponds to a $\beta$-vector that has zero fermionic
840: entries. When the order $d$ of the group $\IZ_d$ is even, the subgroup
841: generated by $g_0^2$ has order $n=d/2$ and it has order $n=d$ when $d$ is odd
842: (see also \cite{Greene:1990ud}). The part of the diagonal symmetry group $D$ that
843: still acts after the projection onto integer R-charges is
844: \be
845: G=(\prod_{i=1}^p
846: \IZ_{k_i} \times \prod_{j=1}^q \IZ_{l_j})/\IZ_n\,.
847: \ee
848:
849: Following \cite{Greene:1990ud} we then define the maximal subgroup $H$ of $G$
850: which preserves supersymmetry in space-time. This is the subgroup
851: corresponding to all vectors $\beta_m$ in the charge lattice that satisfy the
852: equation
853: $2
854: \beta_m \cdot \beta_0 \in 2 \IZ$.
855: This condition ensures that the left- and right- moving R-charges only differ
856: by even integers, as required by supersymmetry.
857: If we write
858: \be
859: \beta_m = \sum_i c_m^i \gamma_i\,,
860: \ee
861: then this condition boils down to the condition that \cite{Greene:1990ud}
862: \be
863: \sum \frac{c_m^i}{k_i} + \sum \frac{c_m^j}{l_j} \in \IZ \,.
864: \ee
865: The subgroup $H$ generated by the vectors $\beta_m$ that satisfy this condition is the
866: maximal orbifold group consistent with space-time supersymmetry.
867:
868: Any subgroup $F$ of $H$ can be used to generate new supersymmetric orbifold
869: models. Precisely as in the compact case, modding out the original
870: model by the maximal subgroup $H$ generates the mirror model
871: (in the conformal field theory sense). Moreover, orbifolding by a subgroup $F$
872: will generate a model that is mirror to the orbifold of the original model by
873: $H/F$. This was argued in \cite{Greene:1990ud} for the compact case, and we
874: will show in section \ref{CFTmirror} that it is also true for the non-compact
875: models. Note that the mirror symmetry in the
876: conformal field theory that we have set up above holds for {\em
877: undeformed} models. We will see that deformations of a model are mapped to
878: resolutions of the mirror. We will discuss this important point in greater detail
879: later on.
880:
881: \subsection{The Deep-throat Region Of Non-compact Gepner Models}
882: \label{local}
883: Until now we have discussed only the continuous part of the spectrum of
884: the non-compact Gepner models.
885: The contribution of these states to the partition function is proportional to
886: the volume of the target space.
887:
888: A good starting point for the rest of our discussion will be to think
889: of the initial model as based on a product of $N=2$ superconformal
890: minimal models and $N=2$ linear dilaton theories.
891:
892: It is important to observe that although the exact torus partition
893: function exists for these conformal field theories,
894: they only describe the
895: asymptotic spectrum of the corresponding string
896: theory. Indeed, there is a region in space-time that is strongly
897: coupled (due to the linearly growing dilaton in one or several
898: directions). In that region, the one-loop spectrum is not a meaningful
899: quantity.
900:
901: Nevertheless, we can get a handle on possible deformations of that
902: singular string theory by working under the following hypotheses. We
903: look for local deformations in the strongly coupled region deep
904: in the throat(s).
905: Secondly, we focus on
906: marginal deformations of the worldsheet theory that
907: preserve supersymmetry in space-time. The $N=2$ superconformal algebra on the
908: worldsheet will be preserved, and the deformations will be based on chiral (or
909: anti-chiral) primaries. Morever, we suppose that the deformations cannot have
910: $U(1)_R$ quantum numbers that differ from those already appearing in the
911: asymptotic partition function, namely, the charges are quantized as in the
912: asymptotic partition function. We believe all of these are mild assumptions,
913: given space-time supersymmetry. (See e.g. \cite{Gukov:1999ya}
914: \cite{Shapere:1999xr} for similar reasonings, mostly from a space-time
915: perspective.)
916:
917: Using the fact that there is a map between $N=2$
918: superconformal algebra representations and reprentations of $SL(2,\IR)$
919: \cite{Dixon:1989cg}, we can
920: reformulate the above conditions as saying that the marginal operators in the
921: full theory should
922: be based on chiral primaries of dimension smaller or equal to one half in
923: the linear dilaton factors. The operator in the linear dilaton factor
924: should have a conformal dimension equal to $h=|m|/k$ where the $U(1)_R$
925: charge is given by $Q=2m/k$, and where $2m$ is an allowed $U(1)_R$ quantum
926: number given the (fixed) asymptotic partition function.
927: We get an upper cut-off: $2|m| \le k$ from the requirement of relevance on the
928: worldsheet.
929: Moreover, we want to study operators that are
930: normalizable at weak coupling.
931: Strict normalisability requires $2 |m|>1$. In particular
932: the value $2 m=0$ is excluded in the non-compact factors, since this would
933: correspond to a non-normalizable operator for a conformal field theory with a
934: non-compact target space.
935: In summary, the quantum number $2|m|$ has to lie in the range
936: $1 < 2|m| \le k$.
937: It should be noted that the operator with $2|m|=1$ is on the border
938: of being normalizable in the sense that it lies at the endpoint of a line of
939: delta-function normalizable operators \cite{Dixon:1989cg}. It will play a special role in what
940: follows, and we will call this type of operator almost-normalizable. We will flesh out the above analysis considerably in the following sections.
941:
942: \begin{figure} \centering
943: \includegraphics{defres}
944: \caption{
945: The asymptotic region in a non-compact Gepner model is identical to that of a linear dilaton conformal field theory. The strong coupling singularity can be cured either by turning on a Liouville potential, or by capping the cylinder with a cigar type deformation.
946: }
947: \label{defres}
948: \end{figure}
949:
950: \section{Landau-Ginzburg Models}
951: \label{LG}
952: One of our goals in this section is to obtain, in a simple manner, the chiral ring of localized operators in
953: the non-compact Gepner models we described in section \ref{noncompactgepner}.
954: For the compact case, this is most carried out by associating a Landau-Ginzburg model to
955: each of the factor conformal field theories. The underlying idea is that at low energies,
956: the Landau-Ginzburg model flows to the conformal field theory that corresponds to the minimal model.
957:
958: Furthermore, the GSO projection that we discussed for the non-compact Gepner
959: models maps to an orbifold of the Landau-Ginzburg theory. Therefore the
960: techniques derived in \cite{Vafa:1989xc, Intriligator:1990ua} to obtain the
961: spectrum in the Landau-Ginzburg models and orbifolds thereof will be crucial
962: for our analysis.
963:
964: In this section and the next, we give a similar Landau-Ginzburg
965: description of our non-compact Gepner models. We will find that much of the
966: technology used in the compact case can be used for the non-compact case
967: as well. However there are subtle differences in reading off the spectrum,
968: because some of the operators are not normalizable.
969:
970: \subsection{Landau-Ginzburg Potentials For Minimal Models}
971:
972: For $N=2$ superconformal minimal models, the flow between Landau-Ginzburg and
973: superconformal minimal models is well-studied
974: \cite{Martinec:1988zu, Vafa:1988uu} and we merely give a
975: brief
976: reminder. A Landau-Ginzburg model with $N=(2,2)$ supersymmetry and a chiral
977: superfield $\Phi$ with superpotential \be\label{simpleLG} W_{MM} = \Phi^k \ee
978: flows in the infrared to an $N=2$ superconformal minimal model. The chiral
979: ring of both models match one-to-one. The chiral unital ring of the Landau-Ginzburg
980: model is
981: $\, \, \IC[\Phi]/\partial_\Phi W$ which is linearly generated by the $k-1$
982: elements $\Phi^0, \Phi^1, \dots, \Phi^{k-2}$, which have (both left- and
983: right-moving) R-charge equal to $0,\frac{1}{k}, \dots , \frac{k-2}{k}$. These
984: match one-to-one to the chiral-chiral primaries of the diagonal minimal model.
985: Further evidence for this identification of the Landau-Ginzburg model fixed
986: point is provided by the matching of the elliptic genus of these models
987: \cite{Witten:1993jg}.
988:
989: The T-dual or mirror Landau-Ginzburg model based on an twisted chiral
990: superfield \cite{Rocek:1991ps} flows to the anti-diagonal minimal model. This
991: can be written as a
992: $\IZ_k$ orbifold of the model in
993: equation \eqref{simpleLG}. We will discuss such orbifolding methods to compute mirrors in later sections.
994:
995: We can summarize the chiral ring of the Landau-Ginzburg model by specifying a
996: Poincare polynomial
997: which is the trace over the chiral-chiral ring weighted
998: by the $U(1)$ R-charges \cite{Lerche:1989uy}:
999: \begin{eqnarray}
1000: Tr_{(c,c)} t^{Q} \tilde{t}^{\tilde{Q}} &=& 1+ (t \tilde{t})^{\frac{1}{k}} + (t \tilde{t})^{\frac{2}{k}}
1001: + \dots + (t \tilde{t})^{\frac{k-2}{k}}
1002: \nonumber \\
1003: & =& \frac{ 1- (t \tilde{t})^{\frac{k-1}{k}} }{ 1 - (t \tilde{t})^{\frac{1}{k}}}.
1004: \end{eqnarray}
1005: In the Ramond sector, this gives rise to a polynomial that keeps track of the
1006: R-charges
1007: of the Ramond-Ramond ground states:
1008: \begin{eqnarray}
1009: Tr_{RR} t^{Q} \tilde{t}^{\tilde{Q}} &=&(t \tilde{t})^{-\frac{1}{2}+\frac{1}{k}} + (t
1010: \tilde{t})^{-\frac{1}{2} +\frac{2}{k}}+ \dots + (t \tilde{t})^{+\frac{1}{2}-\frac{1}{k}}.
1011: \end{eqnarray}
1012: Note that the charges fill out the range from $-c/6$ to $+c/6$, and lie
1013: inside
1014: the interval $]-\half,+\half[$. See figure \ref{Rgroundstatescomp}.
1015: \begin{figure}
1016: \centering
1017: \includegraphics{Rgroundstatescomp}
1018: \caption{The R-charges of the ground states in the Ramond sector
1019: for a $N=2$ superconformal theory with central charge $c<3$.\label{Rgroundstatescomp}}
1020: \end{figure}
1021:
1022: \subsection{Non-compact Landau-Ginzburg Models}
1023:
1024: Now we want to discuss the link between the non-compact $N=2$ superconformal field
1025: theories and the IR fixed point of Landau-Ginzburg models with a superpotential of the form:
1026: \begin{eqnarray}
1027: W_{nc} &=& \Phi^{-l}
1028: \end{eqnarray}
1029: where $l$ is a positive integer. This was introduced\footnote{See also e.g.
1030: \cite{Ghoshal:1993qt, Hanany:1994fi} for an older and related use of
1031: these models in the context of two-dimensional gravity.} in
1032: \cite{Ooguri:1995wj} and we would like to understand how far we can argue for
1033: a formal analogy with the compact case, and if possible borrow the techniques
1034: that have been extensively used in that
1035: context. See also
1036: \cite{Hori:2000kt, Hori:2002cd}
1037: for a more detailed analysis of the renormalization group flow with fixed asymptotics.
1038: The central charge of this theory with fixed asympotics
1039: is $c= 3+ \frac{6}{l}$. %%
1040:
1041: It is natural to assume that the field $\Phi$ cannot take the value zero, and
1042: that moreover the point at infinity should remain a regular point. We can then
1043: associate to the model an operator ring $\, \IC [\Phi^{-1}]$ which should
1044: again be divided by the ideal generated by the derivative of the
1045: superpotential. This gives rise to a ring spanned by the $l+1$ elements
1046: $\Phi^{0}, \Phi^{-1}, \dots, \Phi^{-l}$. Because the target-space is
1047: non-compact, $\Phi^{0}$ is not normalizable. We exclude the operator from the
1048: ring. The ring of elements spanned by $\Phi^{-1}, \Phi^{-2}, \dots,
1049: \Phi^{-l}$ is a ring without unit element. This fact contrasts with the
1050: compact case, and it is associated to the non-existence of an $SL(2,R)$
1051: invariant ground state in the conformal field theory. The operators
1052: $\Phi^{-1}, \Phi^{-2}, \dots, \Phi^{-l}$ have R-charge
1053: $\frac{1}{l},\frac{2}{l},\dots,\frac{l}{l}$. We can match the operators
1054: $\Phi^{-2}, \dots, \Phi^{-l}$ onto the chiral ring of the diagonal linear
1055: dilaton theory with $N=2$ superconformal symmetry (under the assumptions of
1056: relevance and normalizability, as discussed in section \ref{local}). We
1057: moreover expect the operator $\Phi^{-1}$ to become the almost-normalizable
1058: chiral-chiral primary in the infra-red fixed point theory.
1059:
1060: Again, we can summarize the chiral-chiral spectrum in a Poincar\'e polynomial
1061: for the $(c,c)$ ring which for the case of
1062: (almost-)normalizable elements is:
1063: \begin{eqnarray}
1064: Tr_{(c,c)} t^{Q} \tilde{t}^{\tilde{Q}} &=& (t \tilde{t})^{\frac{1}{l}} + (t \tilde{t})^{\frac{2}{l}}
1065: + \dots + (t \tilde{t})^{\frac{l}{l}}
1066: \nonumber \\
1067: & =& (t \tilde{t})^{\frac{1}{l}} \frac{ 1- (t \tilde{t}) }{ 1 - (t \tilde{t})^{\frac{1}{l}}}.
1068: \end{eqnarray}
1069: For the strictly normalizable elements, we should use:
1070: \begin{eqnarray}
1071: Tr_{(c,c)} t^{Q} \tilde{t}^{\tilde{Q}} &=& (t \tilde{t})^{\frac{2}{l}} + (t \tilde{t})^{\frac{3}{l}}
1072: + \dots + (t \tilde{t})^{\frac{l}{l}}
1073: \nonumber \\
1074: &=& (t \tilde{t})^{\frac{2}{l}} \frac{ 1- (t \tilde{t})^{\frac{l-1}{l}} }{ 1 - (t \tilde{t})^{\frac{1}{l}}}.
1075: \end{eqnarray}
1076: Moreover, the normalizable Ramond ground states now have charges that go from $-1/2+1/l$ to
1077: $+1/2-1/l$ which again lie inside the interval $]-\half, +\half[$. The almost-normalizable ground state is an outlier at charge $-1/2$. It has a spectrally flowed partner at opposite charge $+1/2$. The Ramond ground states do not reach the charge $-c/6$ and $+c/6$. See figure
1078: \ref{Rgroundstatesnoncomp}.
1079: \begin{figure}
1080: \centering
1081: \includegraphics{Rgroundstatesnoncomp}
1082: \caption{The R-charges of the ground states in the Ramond sector
1083: for a $N=2$ superconformal theory with central charge $c>3$.
1084: There are almost-normalizable ground states at charges $\pm \frac{1}{2}$.}
1085: \label{Rgroundstatesnoncomp}
1086: \end{figure}
1087: The R-charges of the strictly normalizable
1088: Ramond-Ramond ground states can be coded in the
1089: polynomial:
1090: \begin{eqnarray}
1091: Tr_{RR} t^{Q} \tilde{t}^{\tilde{Q}} &=&(t \tilde{t})^{-\frac{1}{2}+\frac{1}{l}} + (t
1092: \tilde{t})^{-\frac{1}{2} +\frac{2}{l}}+ \dots + (t \tilde{t})^{+\frac{1}{2}-\frac{1}{l}}.
1093: \end{eqnarray}
1094:
1095: To summarize, we can associate a Landau-Ginzburg model to each factor of a
1096: non-compact Gepner model. The
1097: total superpotential will be the sum of the individual superpotentials. The
1098: Landau-Ginzburg models we discuss will be of the Fermat type. However, we note
1099: at this stage that there is a difference between the product
1100: of minimal and linear dilaton theories and the corresponding product of
1101: Landau-Ginzburg theories. Indeed, while the diagonal A-type minimal models are
1102: identified as infra-red fixed points of Landau-Ginzburg models, the
1103: non-compact
1104: Landau-Ginzburg model gives rise to a deformation of the
1105: $N=2$ linear dilaton theory.
1106: To the linear dilaton asymptotics, we add a deforming
1107: potential.
1108:
1109: So far we have studied simple Landau-Ginzburg theories (and their direct
1110: product theories). However, to provide Landau-Ginzburg analogues of the Gepner
1111: conformal field theories, we need to discuss orbifolded Landau-Ginzburg
1112: models.
1113:
1114: \section{Landau-Ginzburg Orbifolds}
1115: \label{orbifoldedLG}
1116:
1117: Our discussion of orbifolded Landau-Ginzburg models is mainly based on
1118: \cite{Vafa:1989xc, Intriligator:1990ua}. The orbifolding can arise due to the
1119: GSO projection in string theory, or due to a further geometric orbifolding of
1120: the resulting theory. In this section, we start out by discussing the orbifold
1121: action as being independent of possible actions on flat space factors,
1122: following \cite{Intriligator:1990ua}. We will comment on the relation to the
1123: full GSO projection (which also acts on the flat space factors)
1124: in the appendix. Since
1125: our discussion will be closely related to the long discussion in the
1126: literature of the compact case, we will briefly review that discussion and we
1127: will only treat in more detail the crucial differences that exist in the
1128: non-compact case.
1129:
1130: For each compact Landau-Ginzburg model with superpotential $\Phi_i^{k_i}$,
1131: there is a canonical diagonal action \be \Phi_i \rightarrow e^{ \frac{2 \pi i
1132: }{k_i} } \Phi_i \,, \ee which generates a $\IZ_{k_i}$ group. The exponent is
1133: determined by the $U(1)_R$ charge of the field $\Phi_i$. For every non-compact
1134: factor with superpotential $\Phi_j^{-l_j}$, the canonical action is \be \Phi_j
1135: \rightarrow e^{- \frac{2 \pi i}{l_j}} \Phi_j\,. \ee We act with the opposite
1136: phase, since the $U(1)_R$ charge of the field $\Phi_j$ is negative (becuase of
1137: the negative power in the superpotential). We introduce the charges $q_i$ for
1138: the fields $\Phi_i$ for both compact and non-compact factors,
1139: which are equal
1140: to \be \left(\frac{1}{k_1}, \frac{1}{k_2}, \dots ; -\frac{1}{l_1},
1141: -\frac{1}{l_2}, \dots\right)\,. \ee Then the above actions on the fields
1142: can be written as \be \Phi_i \rightarrow e^{2 \pi i q_i} \Phi_i\,. \ee The full
1143: symmetry group is then \be D=\prod_i \IZ_{k_i} \times \prod_j \IZ_{l_j}\,.
1144: \ee Note that it is identical to the group we identified previously in the
1145: context of non-compact Gepner models in section \ref{symGroup}.
1146:
1147: We study orbifolds of the theory by a subgroup of the above diagonal group $D$. For instance, we can choose to orbifold by a group generated by a single element, which has different weights in
1148: each of the factors, or by a product of such groups. For an element $h$ of the orbifold group, the action on each superfield can be written as
1149: \be
1150: \Phi_i \rightarrow e^{2 \pi i \Theta_i^h} \Phi_i\,.
1151: \ee
1152: The phases $\Theta_i^h$ parameterize the group elements $h$.
1153:
1154: The integer R-charge orbifold discussed in section \ref{integerR} is special,
1155: since it is part of the GSO projection. In this case, we orbifold by the
1156: group generated by $g_{0}$. The action of this operator on the superfields is coded as
1157: $\Theta_i^{g_{0}}=q_i$.
1158:
1159: Our main objective is to compute the number and charges of the
1160: chiral/chiral $(c,c)$ and anti-chiral/chiral $(a,c)$ states, as well as their
1161: $(a,a)$ and $(c,a)$ partners, in the orbifolded Landau-Ginzburg theory.
1162: In particular, those with
1163: $U(1)_R$ charges $\pm 1$ will lead to massless fields in
1164: spacetime. In the case where the geometric approximation is valid, marginal
1165: $(c,c)$ states correspond to complex structure deformations, and marginal
1166: $(a,c)$ states to K\"ahler moduli.
1167:
1168: A crucial observation is that for the Calabi-Yau examples we consider in this paper, $(c,c)$ states are in one-to-one correspondence with Ramond-Ramond ground states \cite{Lerche:1989uy}. More precisely, a Ramond-Ramond ground state becomes a $(c,c)$ state under the action of half-unit left-right symmetric spectral flow. Similarly, $(a,c)$ states are derived from Ramond-Ramond ground
1169: states by half-unit left-right antisymmetric spectral flow. It will prove a good strategy to work in the R-R sector rather than in the NS-NS sector. The procedure we will use is the following
1170: \cite{Intriligator:1990ua}:
1171: \begin{itemize}
1172: \item[\_]First identify the unprojected Ramond-Ramond ground states in each twisted sector and compute their R-charge.
1173: \item[\_]Then flow these states to the NS-NS sector, and keep those that are
1174: invariant under the orbifold action.
1175: \end{itemize}
1176: Let us study the two steps of this procedure in greater detail.
1177:
1178: \subsection{R-charges of Ramond-Ramond ground states}
1179: We first extend a result of reference \cite{Vafa:1989xc} for the twisted
1180: sector Ramond-Ramond ground states. Consider the sector of the theory twisted
1181: by $h$ (the following is also valid in the untwisted sector, with $h=1$).
1182: We define a particular Ramond-Ramond ground state $|0 \rangle^h_R$, which
1183: is the Ramond-Ramond ground state with the lowest left-moving R-charge for the
1184: compact factors, and the highest left-moving R-charge for the non-compact
1185: factors. The left-moving R-charges with respect to the individual factors of
1186: that
1187: Ramond-Ramond ground state is:
1188: \begin{align}\label{twistedRchargeleft}
1189: Q &= +\sum_{\Theta_i^h \notin Z} \left( \Theta_i^h - {[} \Theta_i^h {]} -
1190: \frac{1}{2} \right) \cr
1191: & \qquad+ \sum_{\Theta_i^h \in Z} (q_i - \frac{1}{2})\,,
1192: \end{align}
1193: and the right-moving R-charge is:
1194: \begin{align}\label{twistedRchargeright}
1195: \tilde{Q} &= -\sum_{\Theta_i^h \notin Z} \left( \Theta_i^h - {[} \Theta_i^h {]} -
1196: \frac{1}{2}\right)\cr
1197: & \qquad+ \sum_{\Theta_i^h \in Z} (q_i - \frac{1}{2})\,.
1198: \end{align}
1199: Here, the expression $[\Theta]$ is defined as the
1200: greatest integer smaller than $\Theta$ (for $\Theta$ not an integer). Remember
1201: that $\Theta_i^h$ is the phase that defines the action of $h$ on $\Phi_i$, and
1202: $q_i$ is the R-charge of $\Phi_i$.
1203:
1204: The arguments leading to the first line of these formulae are elaborate
1205: \cite{Vafa:1989xc}. We refer to that reference for details. Note that it is
1206: natural to look for the Ramond-Ramond ground state in the $h$-twisted sector
1207: by twisting the left-movers half-way in one direction, and twisting the
1208: right-movers half-way in the other direction, due to the symmetry between
1209: left- and right-movers in the original theory. The subtraction of the integer
1210: part of the twist arises because of the fact that we can otherwise find a
1211: state with smaller R-charge, between $-1/2$ and $+1/2$ that will have lower
1212: conformal dimension. See also figures \ref{Rgroundstatescomp} and
1213: \ref{Rgroundstatesnoncomp}. Furthermore, as argued in \cite{Vafa:1989xc}, the
1214: R-charges behave very much like fermion number, which is indeed lifted from
1215: $-1/2$ in the compact sector, by the contribution of $\Theta_i^h$. The crucial
1216: difference with the compact sector is only that for the non-compact sector we
1217: need to keep in mind that $\Theta_i^h$ is negative, and that therefore it is
1218: more natural to think of the state with fermion number or $U(1)_R$ charge
1219: moving {\em down} to a charge just below $+1/2$. That is precisely what is
1220: automatically coded in the above formula, since for $\Theta$ negative but
1221: bigger than $-1$, we have that $\Theta-[\Theta]-1/2=+1/2+\Theta$.
1222:
1223: The second line in each of these formulae \eqref{twistedRchargeleft} and \eqref{twistedRchargeright} arises from spectral flowing the untwisted factor NS-NS ground state to the R-R sector. Note that the
1224: superfield living in an untwisted factor can be given a nonzero vev. This is
1225: not possible in the twisted factor, since a constant does not satisfy the
1226: twisted boundary conditions. In this
1227: way, other Ramond-Ramond ground states can be generated in the same
1228: sector. For example, if the $i$-th factor is untwisted, the state
1229: $\Phi_i^{p}|0 \rangle_{R}^{h}$
1230: ($p$ integer) is another valid Ramond-Ramond ground state. It has left- and
1231: right-moving R-charge $(Q+p|q_i|,\tilde{Q}+p|q_i|)$, where we took into
1232: account the contribution of $\Phi_i^{p}$. In a compact factor, $p$ is
1233: restricted to the bounds $0 \le p \le k_i-2$. In a non-compact factor,
1234: strictly normalizable states have $2 \le -p \le l_i$, and almost-normalizable
1235: states have $p=-1$. These bounds follow from our discussion in section \ref{LG}.
1236:
1237: The upshot is that the formulas \ref{twistedRchargeleft} and \ref{twistedRchargeright},
1238: derived in \cite{Vafa:1989xc} for compact models, are also valid for non-compact models.
1239:
1240: Once this hurdle is ovecome, one can extend the reasoning of \cite{Intriligator:1990ua} to
1241: cover the non-compact case as well. We will not repeat the whole analysis
1242: here. Let us observe that all the information about the unprojected
1243: Ramond-Ramond ground states in the $h$-twisted sector is conveniently encoded
1244: in a Poincare polynomial equal to:
1245: \begin{align}
1246: \Tr_{R, \text{twisted, unprojected}} t^{Q} \tilde{t}^{\tilde{Q}} & =
1247: \left(\frac{t}{\tilde{t}} \right)^{ \sum_{\Theta_i^h \notin Z} ( \Theta_i^h - [ \Theta_i^h ] - 1/2)}
1248: \left(t \tilde{t} \right)^{\sum_{\Theta_i^h \in Z} (q_i-1/2) } \nonumber \\
1249: & \times \prod_{\begin{array}{c} \Theta_i^h \in Z \\ \scriptstyle{compact}\\ \scriptstyle{factors}
1250: \end{array}}
1251: \frac{1-(t\tilde{t})^{\frac{k_i-1}{k_i}}}{1-(t\tilde{t})^{\frac{1}{k_i}}}
1252: \prod_{\begin{array}{c} \Theta_i^h \in Z \\ \scriptstyle{noncompact}\\ \scriptstyle{factors}
1253: \end{array}}
1254: (t\tilde{t})^{\frac{2}{l_i}}\frac{1-(t\tilde{t})^{\frac{l_i-1}{l_i}}}{1-(t\tilde{t})^{\frac{1}{l_i}}}
1255: .
1256: \end{align}
1257:
1258: In the untwisted sector, the above analysis simplifies. The particular
1259: Ramond-Ramond ground state we discussed earlier had charges $(-c/6,-c/6)$. It
1260: is the state one gets by half-unit spectral flow of the NS-NS vacuum. All the
1261: superfields are untwisted, and can be given
1262: constant values. The relevant Poincar\'e
1263: polynomial is simply the product of the Poincar\'e polynomials of each
1264: factors.
1265:
1266: \subsection{Spectral flow to the NS-NS sector}
1267: In this way we can enumerate all the Ramond-Ramond ground states,
1268: and their R-charges. Then we can generate all the NS-NS (anti-)chiral states,
1269: by spectral flow.
1270:
1271: We obtain $(c,c)$ states by symmetric half-unit spectral flow. Their R-charges are easily computed from those of the Ramond-Ramond ground state: we add $(\frac{c}{6},\frac{c}{6})$ to the R-charges of the ground state. This operation does not change the twist: $h$-twisted Ramond-Ramond ground states flow to $h$-twisted $(c,c)$ states.
1272:
1273: In a similar way, we get $(a,c)$ states by anti-symmetric half-unit spectral
1274: flow. This time, the Ramond-Ramond ground state R-charges are shifted by
1275: $(-\frac{c}{6},\frac{c}{6})$. Note that anti-symmetric spectral flow changes the
1276: twist \cite{Vafa:1989xc} such that $h$-twisted Ramond-Ramond ground states flow to
1277: $h g_{0}$-twisted $(a,c)$ states.
1278:
1279: Among all these states, we keep only those that are invariant under the orbifold
1280: action. In the case of the integer R-charge orbifold, the selection is easy:
1281: we keep the states that have integer R-charges. But in other cases,
1282: the procedure is harder. We refer to \cite{Intriligator:1990ua} for a generic discussion of
1283: this projection.
1284:
1285: \subsection{Normalizability of non-compact twisted states}
1286:
1287: For the twisted Ramond-Ramond ground states, we encounter a subtle point and
1288: this is the crucial difference between the compact and the non-compact case.
1289: When we twist a coordinate, and restrict to constant modes, the field is set
1290: to zero. This will not give rise to a normalizable state when the
1291: Landau-Ginzburg potential has a negative power, since the potential blows up
1292: at zero. Although we could therefore immediately discard the states that have
1293: a non-compact twisted factor as not being contained in the normalizable
1294: spectrum, we advocate keeping an open mind -- we will continue the
1295: investigation into these non-normalizable states, in view of the possibility
1296: that we can tune the coefficient of the negative power potential to zero,
1297: which will render these states normalizable.
1298:
1299: Strictly speaking, the models where we do take into account such twisted
1300: sector deformations should be thought of as existing in the well-defined
1301: linear dilaton conformal field theories (which give rise to locally strongly
1302: coupled string theory backgrounds). We believe we provide ample justification
1303: for this procedure later on. In particular in section \ref{CFTmirror} we will
1304: find that these states are needed in the construction of mirror pairs of
1305: conformal field theories, by generalizing the Greene-Plesser prescription of
1306: finding mirrors by orbifolding. More precisely, states untwisted in the
1307: non-compact directions will be mapped in the mirror model to states twisted in
1308: the non-compact directions.
1309:
1310: To summarize, the full set of chiral and anti-chiral ring elements of the
1311: Landau-Ginzburg orbifold, including those arising in both the untwisted and
1312: twisted sectors, is interpreted as the spectrum of the linear dilaton theory.
1313: The Landau-Ginzburg model is a complex structure deformation of this linear
1314: dilaton background. This deformation removes from the spectrum the RR states
1315: twisted in the non-compact directions and the de-singularized string
1316: background only contains purely untwisted states. As we will see later on, in
1317: the mirror model, the deformation is mapped to a resolution of the theory. The
1318: spectrum of the resolved theory will, instead, only contain the states twisted
1319: in the non-compact directions.
1320:
1321: \section{Concrete Models}
1322: \label{examples}
1323: In this section we will apply what we have learned in the previous sections to
1324: some concrete models. We supplement the analysis of the
1325: asymptotic partition function of section \ref{asymptotic}
1326: with an analysis of the localized states (which we briefly touched upon
1327: in section \ref{local}). We start out with some
1328: observations on the nature of non-compact Gepner models, and how they differ
1329: from their compact counterparts.
1330: We will then compute the allowed deformations of $(c,c)$ and $(a,c)$ type in
1331: particular examples, in the language of orbifolded Landau-Ginzburg models. In the conformal field theory language, the conditions to be satisfied for chiral primaries have been written out in a series of papers \cite{Eguchi:2003ik, Eguchi:2004yi, Eguchi:2004ik}. We compare our results with the conformal field theory formalism in appendix \ref{CFTanalysis}.
1332:
1333:
1334:
1335: \subsection{Gepner Points At Large Level}
1336:
1337: We start out with some comments that allow us to single out some particularly
1338: interesting models.
1339: Compact Gepner models have factors with a central charge which is always
1340: smaller than three. A large level limit for all factors
1341: necessarily increases the central charge of the Gepner model, and
1342: therefore is not consistent with the criticality condition for string
1343: theory. Thus, compact Gepner models are necessarily at large curvature (and small
1344: volume).
1345:
1346: Non-compact Gepner models are of a different type. In particular we can have
1347: non-compact Gepner models that contain factor conformal field theories at
1348: central charge smaller {\it and} larger than three. We can therefore cancel
1349: off the difference in a large level limit, if we wish. Assuming that we
1350: demand the existence of a limit in which all levels are large, we immediately
1351: conclude that such a non-compact Gepner model necessarily has three
1352: factors (since the total central charge is $c=9=3\times 3$).
1353:
1354: There are two such classes of models (if we do not admit models with central
1355: charge
1356: precisely equal to three). One is where we have two minimal
1357: model factors and one non-compact factor, and the other has one
1358: minimal model factor and two non-compact factors. There are no other
1359: possibilities at central charge $c=9$. (When we consider a non-compact
1360: Gepner model at central charge $c=3.D$, with $D$ an integer different from
1361: $D=3$,
1362: there are other possibilities which can also easily be classified.)
1363:
1364: The class of models with two minimal models can be parameterized by the
1365: integer levels $k_1$ and $k_2$ of the minimal models, up to
1366: an initial choice of ADE modular invariant and a possible orbifold by a
1367: symmetry group.
1368: The level $l$ of the non-compact factor is then fixed to be $l
1369: = \frac{k_1 k_2}{k_1 + k_2}$ which can be integer or fractional.
1370:
1371:
1372: The other class of models is parameterized by two levels $l_1$ and
1373: $l_2$ for the non-compact models, and the combination $k= \frac{l_1
1374: l_2}{l_1+l_2}$ then needs to be integer, in order for the compact
1375: model at level $k$ to exist. The levels $l_i$ can a priori be
1376: fractional or real. As we have indicated before, we
1377: concentrate on the case where all levels are integer.
1378:
1379:
1380: \subsection*{Remarks}
1381: \begin{itemize}
1382: \item We note that a general class of models can be found by demanding
1383: $\frac{k_1 k_2}{k_1 + k_2}$ to be integer (with $k_{1}$ and $k_2$ positive
1384: integers). Suppose we isolate the
1385: greatest common divisor $d$ of $k_1=d \tilde{k}_1$ and $k_2= d \tilde{k}_2$
1386: (with $\tilde{k}_1$ and $\tilde{k}_2$ mutually prime). Then it can be shown
1387: that the level $ \frac{k_1 k_2}{k_1 + k_2}$ is integer if and only if $d$ is a
1388: divisor of $\tilde{k}_1+ \tilde{k}_2$. That easily generates a large class of
1389: models of which we will only study a few.
1390: \item We note that although the local curvature would seem to become small in
1391: the large level limit, this reasoning does not take into account the GSO
1392: orbifold that still needs to be performed. In specific examples it can be
1393: checked that the GSO orbifold recreates small radii in the geometry (see
1394: e.g. \cite{Israel:2004ir}\cite{Israel:2004jt} for a detailed discussion).
1395: We can therefore generically expect the large level limit to correspond to
1396: an {\em orbifolded} weakly curved background.
1397: \end{itemize}
1398:
1399: In the following we will mainly concentrate on the set of models that have
1400: the special property of allowing for a large level limit (although our
1401: formalism does apply more widely).
1402: We study in detail the set of models with three factors and levels $(2k,2k;k)$
1403: or $(k;2k,2k)$ for the minimal model and non-compact factors respectively, and
1404: orbifolds thereof.
1405: As a warm-up exercise however, we treat the instructive
1406: example at complex dimension $D=2$ with the compact model at level $k$ and the
1407: non-compact model at the same
1408: level $k$.
1409:
1410: \subsection{The $(k; k)$ Model}
1411:
1412: Let us apply the formalism for orbifolded Landau-Ginzburg models of section
1413: \ref{orbifoldedLG} to the case of two factors, one compact and one non-compact
1414: at equal levels $k$. We refer to the model as the $(k;k)$ model.
1415:
1416: The Landau-Ginzburg model has a potential
1417: \be
1418: W_{LG}=\Phi_1^k + \Phi_2^{-k}
1419: \ee
1420: for two chiral superfields $\Phi_{1}$ and $\Phi_2$. The orbifold group that is
1421: necessary to implement GSO is generated by
1422: \be
1423: g_{0}: (\Phi_1,\Phi_2) \rightarrow (e^{2 \pi i /k} \Phi_1,e^{- 2 \pi i /k} \Phi_2)\,.
1424: \ee
1425: We will be mostly interested in obtaining the numbers and types of
1426: deformations of the conformal field theory. In order to do this, one computes the left and right
1427: R-charges of the Ramond-Ramond ground states in all the twisted sectors using
1428: equations \eqref{twistedRchargeleft} and \eqref{twistedRchargeright}. Then, after
1429: spectral flow, one can compute the R-charges of the operators in the $(c,c)$
1430: and $(a,c)$ rings. For this $c=6$ theory, half-unit spectral flow amounts to adding $\pm
1431: 1$ unit of R-charge to the RR states.
1432:
1433: We recall that for the $(a,c)$ states, asymmetric spectral flow from the RR
1434: sector adds a twist by $g_{0}$.
1435: We tabulate the R-charges of the relevant states below, and indicate with a
1436: star the sectors in which
1437: non-zero constant modes can be given to the fields. We label by $\alpha$ the sector
1438: twisted by $g_{0}^{\alpha}$.
1439: \be
1440: \begin{array}{c|c|c|c}
1441: \alpha & RR & (c,c) & (a,c) \cr
1442: \hline
1443: 0 & (-1,-1)* & (0,0)* & (-1,+1) \cr
1444: 1 & (0,0) & (+1,+1) & (-2,0)* \cr
1445: 2 \le \alpha \le k-1 & (0,0) & (+1,+1)& (-1,+1) \cr
1446: \end{array}
1447: \ee
1448: We summarize the results:
1449: \begin{itemize}
1450: \item In the untwisted sector, the fields can have non-zero
1451: constant modes. The Ramond-Ramond ground state flows in the $(c,c)$ ring to the
1452: identity operator, with charges $(0,0)$. We moreover find $k-1$ marginal
1453: $(c,c)$ states in this sector. These
1454: arise from the monomials $\Phi_1^n \Phi_2^{-k+n}$ for $n=0,1, \dots, k-2$.
1455: They can be checked to be invariant under the orbifold projection.
1456: \item Asymmetric spectral flow from the RR untwisted sector gives
1457: $g_{0}$-twisted $(a,c)$ states. Therefore we get $k-1$ marginal $(a,c)$
1458: states in this sector. The fact that these states are marginal is a
1459: consequence of the special value of the central charge, $c=6$.
1460: \item For any value of the twist $\alpha=1,\dots,k-1$, we have no untwisted
1461: fields
1462: in the RR sector.
1463: Namely the fields $\Phi_1$ and $\Phi_2$ always have twisted boundary
1464: conditions in any twisted sector. Using the formula (\ref{twistedRchargeleft}), we find that the R-charges
1465: of the $\alpha$-twisted sector Ramond ground states is $(\alpha/k -0 -1/2) +
1466: (-\alpha/k+1-1/2) = 0$ on the left and $0$ on the right. We already see a
1467: phenomenon typical to our non-compact Gepner models. The twist contribution
1468: to the Ramond sector charges can cancel between compact and non-compact
1469: factors. We are in a special case, in which the twisted R-charges of all
1470: ground states are zero.
1471: \item After flowing symmetrically to the NS-NS sector, we find a single $(c,c)$ state with charges
1472: $(+1,+1)$ in each twisted sector. They give $k-1$ marginal $(c,c)$ states in total in the twisted
1473: sectors.
1474: \item Asymmetric spectral flow of the twisted RR ground states leads to $k-1$
1475: $(a,c)$ deformations, one in each sector (except for $\alpha=1$).
1476: \end{itemize}
1477:
1478: Therefore we have a total of $k-1$ marginal $(c,c)$ and $k-1$ marginal $(a,c)$
1479: states from the spectral flow of the untwisted RR sector ground states. In
1480: the Landau-Ginzburg model with potential $W_{LG}=\Phi_1^k+\Phi_2^{-k}$ these
1481: are the only admissible localized modes. The potential term $\Phi_2^{-k}$
1482: makes sure that the untwisted polynomials are allowed in the sense that they
1483: are normalizable at weak coupling, and have a mild behaviour at the $\Phi_2
1484: \approx 0$ end compared to the potential. These give rise to a $4(k-1)$
1485: real-dimensional moduli space of backgrounds in string theory with sixteen
1486: supercharges.
1487: In contrast, the twisted operators are not normalizable in the
1488: Liouville deformed model.
1489:
1490:
1491: \subsection{The $(2k,2k;k)$ Model.}\label{twoMM}
1492:
1493: For this slightly more complicated example, we list the full set of
1494: (unprojected)
1495: $(c,c)$
1496: and $(a,c)$ states and their charges, and then pick out those that are marginal (and invariant
1497: with respect to the orbifold projection). Again, we perform this
1498: exercise in the formalism of section \ref{orbifoldedLG} for orbifolded
1499: Landau-Ginzburg models.
1500: We consider a Landau-Ginzburg model with fields $\Phi_{1,2,3}$ and
1501: superpotential
1502: \be
1503: W_{LG} = \Phi_1^{2k} + \Phi_2^{2k} + \Phi_3^{-k}\,.
1504: \ee
1505: We perform the integer R-charge orbifold, generated by the $g_{0}$:
1506: \be
1507: g_{0}: (\Phi_1,\Phi_2,\Phi_3) \rightarrow (e^{2 \pi i /2k} \Phi_1, e^{2 \pi i
1508: /2k} \Phi_2, e^{- 2 \pi i/k} \Phi_3) \,.
1509: \ee
1510: In order to consider all the marginal operators, we will write down the
1511: relevant Poincare polynomials in each sector. First of all, in the untwisted
1512: sector we have the (unprojected, strictly
1513: normalizable) Poincare polynomial:
1514: \begin{eqnarray}
1515: \left( \frac{1-(t \tilde{t})^{(2k-1)/2k}}{1-(t
1516: \tilde{t})^{\frac{1}{2k}}}\right)^2 (t \tilde{t})^{\frac{2}{k}}
1517: \frac{1-(t \tilde{t})^{\frac{k-1}{k}}}{1-(t \tilde{t})^{\frac{1}{k}}}
1518: \end{eqnarray}
1519: which contains $(c,c)$ states only.
1520: When we label the twisted sectors by $\alpha=1,2, \dots, 2k-1$, we find that
1521: for $\alpha$ smaller than $k$ there are further twisted $(c,c)$ states in the
1522: NSNS
1523: sector determined by the polynomials:
1524: \be
1525: \left( \frac{t}{\tilde{t}} \right)^{-1/2} (t \tilde{t})^{3/2}
1526: \ee
1527: and when $\alpha$ is larger than $k$ by the polynomial
1528: \be
1529: \left( \frac{t}{\tilde{t}} \right)^{+1/2} (t \tilde{t})^{3/2},
1530: \ee
1531: while at $\alpha=k$ we find the polynomial
1532: \be
1533: (t\tilde{t})^{-1/k-1/2} (t \tilde{t})^{+3/2} (t \tilde{t})^{\frac{2}{k}}
1534: \frac{(1-(t \tilde{t}))^{\frac{k-1}{k}}}{(1-(t \tilde{t}))^{\frac{1}{k}}}.
1535: \ee
1536: where the last factor is due to the fact that $\Phi_3$ is untwisted in this sector.
1537: The twisted $(a,c)$ states are determined by the polynomials:
1538: \be
1539: \left( \frac{t}{\tilde{t}} \right)^{-2} (t \tilde{t})^0 .1
1540: \ee
1541: when $\alpha$ is smaller than $k$ and
1542: \be
1543: \left( \frac{t}{\tilde{t}} \right)^{-1} (t \tilde{t})^0 .1
1544: \ee
1545: when $\alpha$ is larger than $k$.
1546: When $\alpha=k$, we get
1547: \be
1548: (t\tilde{t})^{-1/k-1/2} \left(\frac{t}{ \tilde{t}}\right)^{-3/2} (t \tilde{t})^{\frac{2}{k}}
1549: \frac{(1-(t \tilde{t}))^{\frac{k-1}{k}}}{(1-(t \tilde{t}))^{\frac{1}{k}}}.
1550: \ee
1551: One can straightforwardly determine amongst these the states that are
1552: invariant under the orbifold action: they have integer R-charges.
1553:
1554: \subsection*{Marginal Deformations}
1555:
1556: We now want to look for exactly marginal deformations in these rings. These
1557: need to have left- and right R-charge equal to $\pm 1$. As for the $(k;k)$
1558: example, let us tabulate the R-charges of the ground states and their images
1559: under spectral flow. Once again, $\alpha$ labels the twist. A star indicates
1560: that some fields are untwisted and can be given nonzero vev's.
1561: \be
1562: \begin{array}{c|c|c|c}
1563: \alpha& RR & (c,c) & (a,c) \cr
1564: \hline
1565: 0 & (-3/2,-3/2)* & (0,0)* & (-1,1) \cr
1566: 1 & (-1/2,1/2) & (1,2) & (-3,0)* \cr
1567: 2 \le \alpha <k & (-1/2,1/2) & (1,2) & (-2,2) \cr
1568: k & (-1/k-1/2,-1/k-1/2)* &(-1/k+1,-1/k+1)* & (-2,2)\cr
1569: k+1 & (1/2,-1/2) &(2,1) & (-1/k-2,-1/k+1)*\cr
1570: \alpha>k+1 & (1/2,-1/2) & (2,1) & (-1,1)
1571: \end{array}
1572: \ee
1573: The table agrees with the Poincar\'e polynomials listed previously. We find
1574: the following marginal deformations:
1575:
1576: \begin{itemize}
1577: \item The untwisted $(c,c)$ marginal states are straightforwardly
1578: enumerated. They are given by invariant combinations of the $\Phi_i$ acting
1579: on the vacuum: $\Phi_1^a \Phi_2^b \Phi_3^{-c} |0 \rangle_{NS}$,
1580: with the bounds $0 \le a,b \le 2k-2$ and $2 \le c \le k$. The marginality
1581: condition is: $a+b+2c=2k$.\\
1582: Let's count these states. For a given $c$, each $a$ in the range $0 \le a \le
1583: 2k-2c$ gives exactly one solution. So the total number of states is:
1584: \be \sharp (c,c) = \sum_{c=2}^{k}(2k-2c+1) = (k-1)^2 \ee
1585:
1586: \item A further search for marginal $(c,c)$ states gives a negative result.
1587: When $\alpha$ is smaller than $k$, we have $(c,c)$ states that survive
1588: projection, but they are not marginal since they have charges $(1,2)$. When
1589: $\alpha=k$, the charges left and right can also never be both equal to one. When
1590: $\alpha$ is larger than $k$, the charges are $(2,1)$ which also never leads to
1591: marginality.
1592:
1593: \item Let us look for marginal $(a,c)$ states in the twisted sector. We get
1594: charges $(-2,2)$ when $\alpha-1$ is smaller than $k$, and therefore no marginal
1595: states in these sectors. When $\alpha-1$ is larger
1596: than $k$, we get charges $(-1,+1)$ which are marginal. So we get $(k-1)$
1597: $(a,c)$ states from the twisted sectors, labeled by $\{k+2,\ldots 2k\}$.
1598: There are no other marginal states in this theory.
1599:
1600: \end{itemize}
1601: In summary, we find $(k-1)^2$ untwisted marginal $(c,c)$ states, and we find
1602: $k-1$ marginal $(a,c)$ states. Once again, only those states that arise from
1603: the spectral flow of RR ground states with untwisted non-compact factors are
1604: retained in the theory deformed with the Liouville potential. So all
1605: the $(c,c)$ states are in the spectrum of the deformed theory, but none of the
1606: $(a,c)$
1607: states are.
1608:
1609: \subsubsection{Orbifolds}\label{twoMMorbifold}
1610:
1611: We will now consider orbifolds of the above model. As we will discuss in
1612: detail in the next section, this exercise is useful since it generates an
1613: infinite number of mirror theories. The logic will be analogous to the
1614: Greene-Plesser analysis for the Gepner point in the quintic Calabi-Yau. Under
1615: the hypothesis that the level $k$ is not prime, we can write the level as a
1616: product $k=k_1.k_2$ for two positive integers $k_1$ and $k_2$. Then we can
1617: perform a $\IZ_{k_1}$ orbifold of the three-factor model that we discussed
1618: above.
1619:
1620: The Landau-Ginzburg model has superpotential
1621: $$W_{LG}= \Phi_1^{2k} + \Phi_2^{2k} + \Phi_3^{-k}.$$
1622: The full symmetry group of $W_{LG}$ is $D=\IZ_{2k}
1623: \times \IZ_{2k} \times \IZ_{k}$, where each factor acts by phase
1624: multiplication on one of the superfields. The integer R-charge operator $g_{0}$ generates
1625: a $\IZ_{2k}$ subgroup. It acts as: \be g_{0}: (\Phi_1,\Phi_2,\Phi_3)
1626: \rightarrow (e^{\frac{2 i \pi}{2k}}\Phi_1,e^{\frac{2 i
1627: \pi}{2k}}\Phi_2,e^{-\frac{2 i \pi}{k}}\Phi_3) \,. \ee Now consider the
1628: $g_{1}$ operator with the following action: \be g_{1}:
1629: (\Phi_1,\Phi_2,\Phi_3) \rightarrow (e^{+\frac{2 i \pi
1630: k_2}{2k}}\Phi_1,e^{-\frac{2 i \pi k_2}{2k}}\Phi_2,\Phi_3) \,. \ee Earlier
1631: we orbifolded the Landau-Ginzburg model by the group generated by $g_{0}$. Consider the group of order
1632: $2k k_1$ generated by $g_{0}$ and $g_{1}$. We now orbifold the
1633: Landau-Ginzburg theory by that group. This is compatible with
1634: supersymmetry.
1635: Let us find marginal deformations in this orbifold model,
1636: following the Landau-Ginzburg method. First
1637: we tabulate the R-charges of the Ramond-Ramond ground states. We label the sector
1638: twisted by $g_{0}^{\alpha} g_{1}^{\beta}$ with $\alpha$ and $\beta$, within the
1639: bounds $0 \le \alpha \le 2k-1$ and $0 \le \beta \le k_1-1$. A star means that
1640: some fields are untwisted.
1641: \be
1642: \begin{array}{c|c|c}
1643: \alpha & \beta & RR \cr
1644: \hline
1645: \alpha=0 & \beta=0 & (-3/2,-3/2)* \cr
1646: 0 < \alpha <k & \beta k_2 < \alpha & (-1/2,1/2) \cr
1647: 0 < \alpha <k & \alpha < \beta k_2 & (1/2,-1/2) \cr
1648: k<\alpha & \beta k_2 <2k-\alpha & (1/2,-1/2) \cr
1649: k<\alpha & 2k-\alpha<\beta k_2 & (-1/2,1/2) \cr
1650: \alpha=0 & \beta \neq 0 & (-1/2-1/k,-1/2-1/k)* \cr
1651: \alpha=k & \beta & (-1/2-1/k,-1/2-1/k)* \cr
1652: \alpha=\beta k_2 & \beta \neq 0 & (-1/2+1/2k,-1/2+1/2k)* \cr
1653: \alpha=2k-\beta k_2 & \beta \neq 0 & (-1/2+1/2k,-1/2+1/2k)* \cr
1654: \end{array}
1655: \ee
1656: Let us count the marginal operators:
1657: \begin{itemize}
1658: \item In the untwisted sector, we find $(c,c)$ chiral primaries.
1659: Remember that in the model orbifolded by $g_{0}$, the $(k-1)^2$ $(c,c)$ states are labeled by three
1660: integers $a,b,c$, such that $0 \le a,b \le 2k-2$, $2 \le c \le k$ and
1661: $a+b+2c=2k$. The $g_{1}$ projection keeps only those that have $a \equiv b
1662: \ [2k_1]$.\\
1663: Let us work at given $c$. We want to count the number of solutions to the
1664: equation $a+b=2k-2c$, with $a \equiv b\ [2k_1]$. We write $b=a+2d k_1 $,
1665: with $d$ integer. Then we express $a$ and $b$ in terms of $c$ and $d$ only:
1666: $a=k-c-d k_1$, and $b=k-c+d k_1$. The bounds on
1667: $a$ and $b$ imply $-k+c \le d k_1 \le k-c$. Thus we have one solution for each
1668: integer $d$ between $\frac{c-k}{k_1}$ and $\frac{k-c}{k_1}$.\\
1669: So the total number of $(c,c)$ states is:
1670: \be \sharp (c,c) = \sum_{c=2}^{k} \left( \left\lfloor \frac{k-c}{k_1} \right\rfloor - \left\lceil
1671: \frac{c-k}{k_1} \right\rceil + 1 \right) \ee
1672: where $\lfloor x \rfloor$ is the largest integer smaller or equal to $x$, and
1673: $\lceil x \rceil$ is the smallest integer bigger or equal to $x$.\\
1674: This sum can be evaluated explicitly :
1675: \be \sharp (c,c) =
1676: 2 \left( (k_2-1)(k_1-1) + \sum_{\check{c}=2}^{k_2} (k_2-\check{c})k_1 \right) + k-1 =
1677: k_1 k_2^2 - 2k_2 +1 \ee
1678: \item In the sector twisted by $g_{0}^{\alpha\mp1}\ g_{1}^\beta$, we find
1679: $(c,a)$ and $(a,c)$ states\footnote{ The $\a\mp1$ occurs due to the shift in
1680: the labeling of the twisted sectors when flowing asymmetrically from the
1681: RR to the $(c,a)$ or $(a,c)$ sectors.}:
1682: \begin{itemize}
1683: \item If $0 \le \beta k_2 < \alpha <k$, or $\alpha > 2k- \beta k_2$,
1684: we find one marginal $(c,a)$ state. Let's count the number of such
1685: twisted sectors. It can be written as:
1686: \be \sharp(c,a) = \sum_{\alpha=1}^{k-1} \left( \left\lfloor \frac{\alpha}{k_2}
1687: \right\rfloor +1 \right)
1688: + \sum_{\alpha=k+1}^{2k-1} \left( k_1-1- \left\lfloor \frac{2k-\alpha}{k_2}
1689: \right\rfloor\right) \ee
1690: This sum can be computed explicitly as:
1691: \be \sharp(c,a) = \left( (k_2-1)\sum_{\hat{\alpha}=1}^{k_1} \hat{\alpha} \right) +
1692: \left( (k_1-1)(k-1) - (k_2-1)\sum_{\check{\alpha}=0}^{k_1-1} \check{\alpha}
1693: \right) = k_2 k_1^2 -2k_1+1 \ee
1694: \item Symmetrically, if $0 < \alpha < \beta k_2 $, or $k < \alpha < 2k-\beta k_2$,
1695: we find one $(a,c)$ state. The same counting shows that there are also $k_2 k_1^2 -2k_1+1$ such sectors.
1696: \item Eventually, if $\alpha=0$, $\alpha=k$ or $\beta k_2 = \pm \alpha \
1697: [2k]$, we find no marginal deformation.
1698: \end{itemize}
1699: \end{itemize}
1700:
1701: In summary, we have in this orbifold model:
1702: \begin{itemize}
1703: \item $k_1 k_2^2 -2k_2+1$ marginal $(c,c)$ states.
1704: \item $k_2 k_1^2 -2k_1+1$ marinal $(a,c)$ states.
1705: \end{itemize}
1706: Once again, only the $(c,c)$ states are present in the spectrum of the
1707: deformed theory.
1708:
1709:
1710: \subsection{The $(k;2k,2k)$ Model.} \label{twocigars}
1711: This example will be very similar to the previous one. For this reason our discussion will be brief and we will focus on marginal deformations. The Landau-Ginzburg model has superpotential
1712: \be
1713: W_{LG}= \Phi_1^k + \Phi_2^{-2k} + \Phi_3^{-2k}\,.
1714: \ee
1715: The integer R-charge operator $g_{0}$ acts on the superfields as:
1716: \be
1717: g_{0}: (\Phi_1,\Phi_2,\Phi_3) = (e^{\frac{2 i \pi}{k}}\Phi_1,e^{-\frac{2 i \pi}{2k}}\Phi_2,e^{-\frac{2 i \pi}{2k}}\Phi_3) \,.
1718: \ee
1719: We tabulate the R-charges of the Ramond ground states and their spectral
1720: flows.
1721: A star means that some fields are untwisted, and $\alpha$ labels the sector
1722: twisted by $g_{0}^{\alpha}$:
1723: \be
1724: \begin{array}{c|c|c|c}
1725: \alpha& RR & (c,c) & (a,c) \cr
1726: \hline
1727: 0 & (-3/2,-3/2)* & (0,0)* & (-2,2) \cr
1728: 1 & (1/2,-1/2) & (2,1) & (-3,0)* \cr
1729: 2 \le \alpha<k & (1/2,-1/2) & (2,1) & (-1,1) \cr
1730: k& (1/k-1/2,1/k-1/2)* &(1/k+1,1/k+1)* & (-1,1)\cr
1731: k+1& (-1/2,1/2) &(1,2) & (1/k-2,1/k+1)*\cr
1732: \alpha>k+1 & (-1/2,1/2) & (1,2) & (-2,2)
1733: \end{array}
1734: \ee
1735:
1736: \begin{itemize}
1737: \item In the untwisted sector, we find
1738: $(k-1)^2$ $(c,c)$ states.
1739: \item In the sector twisted by $g_{0}^\alpha$ ($0<\alpha<2k-1$), we find one
1740: $(a,c)$ state if $2\le \alpha \le k$, and one $(c,a)$ state if $k \le
1741: \alpha\le 2k-2$.
1742: \end{itemize}
1743: To summarize, this model has $(k-1)^2$ marginal $(c,c)$ deformations and $k-1$
1744: margial $(a,c)$ deformations.
1745:
1746:
1747: \subsubsection{Orbifolds}\label{twocigarsorbifold}
1748:
1749: We repeat the analysis of section \ref{twoMMorbifold}, and study the orbifolds
1750: of this model. The Landau-Ginzburg model has superpotential $W_{LG}= \Phi_1^k + \Phi_2^{-2k} + \Phi_3^{-2k}$. The full group of symmetries of $W_{LG}$ is $\IZ_k \times \IZ_{2k} \times \IZ_{2k}$, where each factor acts by phase multiplication on one superfield.
1751:
1752: The integer R-charge operator $g_{0}$ generates a $\IZ_{2k}$ subgroup and acts as follows:
1753: \be
1754: g_{0}: (\Phi_1,\Phi_2,\Phi_3) \rightarrow (e^{\frac{2 i \pi}{k}}\Phi_1,e^{-\frac{2 i \pi}{2k}}\Phi_2,e^{-\frac{2 i \pi}{2k}}\Phi_3)
1755: \ee
1756: Now we consider the $g_{1}$ operator with the following action:
1757: \be
1758: g_{1}: (\Phi_1,\Phi_2,\Phi_3) \rightarrow (\Phi_1,e^{-\frac{2 i \pi k_2}{2k}}\Phi_2,e^{+\frac{2 i \pi k_2}{2k}}\Phi_3)
1759: \ee
1760: As in the $(2k,2k;k)$ model, we assume that $k=k_1. k_2$, orbifold the theory
1761: by the subgroup generated by $g_{0}$ and $g_{1}$ and look for marginal
1762: deformations. The
1763: counting is very similar
1764: to the $(2k,2k;k)$ case. We find
1765: \begin{itemize}
1766: \item $k_1 k_2^2 -2k_2+1$ marginal $(c,c)$ states.
1767: \item $k_2 k_1^2 -2k_1+1$ marginal $(a,c)$ states.
1768: \end{itemize}
1769:
1770: \subsection{The $(3,3,3;2)$ Model}
1771:
1772: All our previous examples share an interesting feature. The marginal $(c,c)$
1773: operators are obtained by half-unit spectral flow of untwisted RR-ground
1774: states. On the other
1775: hand, the marginal $(a,c)$ operators are obtained by
1776: half-unit asymetric spectral flow of twisted RR-ground states. A direct
1777: consequence is that in these examples, the deformed theory only has
1778: $(c,c)$ moduli in its spectrum.
1779: This statement looks general, since untwisted ground states have equal left
1780: and right R-charges, and
1781: tend to flow to chiral operators. Similarly,
1782: twisted ground states have different left and right R-charges, and would be
1783: expected to flow to anti-chiral operators. However, we will show in the present example
1784: that there are exceptions to this rule.
1785:
1786: We consider the Gepner model with three compact factors at level 3, and one
1787: non-compact factor at level 2. The superpotential of the corresponding
1788: Landau-Ginzburg model is \be W=\Phi_1^3+\Phi_2^3+\Phi_3^3+\Phi_4^{-2}+\Phi_5^2
1789: \,, \ee and the model is orbifolded by the group generated by $g_{0}$: \be g_0
1790: : (\Phi_1,\Phi_2,\Phi_3,\Phi_4,\Phi_5)\rightarrow (e^{\frac{2i\pi}{3}}\Phi_1,
1791: e^{\frac{2i\pi}{3}}\Phi_2, e^{-\frac{2i\pi}{3}}\Phi_3, e^{-i\pi}
1792: \Phi_4,e^{i\pi}\Phi_5) \,. \ee The addition of $\Phi_5$ is the simplest means
1793: to ensure that the Calabi-Yau condition is maintained while setting all the
1794: phase factors in \cite{Intriligator:1990ua} to zero. Using the by now familiar
1795: techniques, we tabulate the R-charges of the ground states in various sectors:
1796: \be
1797: \begin{array}{c|c|c|c}
1798: \alpha & RR & (c,c) & (a,c) \\
1799: \hline
1800: 0 & (-3/2,-3/2)* \diamond & (0,0)* \diamond &(-1,1) \\
1801: 1 & (-1/2,1/2) & (1,2) & (-3,0)* \diamond \\
1802: 2 & (-1/2,-3/2)* & (1,0)* & (-2,2) \\
1803: 3 & (-1/2,-1/2)\diamond & (1,1)\diamond & (-2,0)* \\
1804: 4 & (-3/2,-1/2)* & (0,1)* & (-2,1) \diamond\\
1805: 5 & (1/2,-1/2) & (2,1) & (-3,1)* \\
1806: \end{array}
1807: \ee
1808: The star and the diamond respectively mean that the non-compact and the
1809: compact fields are untwisted. More precisely, $\Phi_{1,2,3}$
1810: can have zero modes in the $\alpha=0,3$ twisted sectors for the $(c,c)$ ring
1811: while they can have zero modes in the $\alpha=1,4$ twisted sectors in the
1812: $(a,c)$ ring. Similarly, $\Phi_{4,5}$ can have zero modes in the $\alpha=0,2,4$
1813: sectors in the $(c,c)$ ring, while it can have zero modes in the $\alpha=1,3,5$ sectors in the $(a,c)$ ring.
1814:
1815: >From this, we see that there are 2 $(c,c)$ moduli:
1816: \begin{align}
1817: \Phi_4^{-2}\ket{0}_{c,c}^{\alpha=0} \quad \text{and} \quad \ket{0}_{c,c}^{\alpha=3}
1818: \cr
1819: \end{align}
1820: and $2$ $(a,c)$ moduli:
1821: \be
1822: \ket{0}_{a,c}^{\alpha=0} \quad \text{and}\quad \Phi_4^{-2}\ket{0}_{a,c}^{\alpha=3} \,.
1823: \ee
1824: Notice that the $(c,c)$ modulus in the third twisted sector occurs in a sector
1825: in which the fields that corresponds to the non-compact direction, $\Phi_4$,
1826: is twisted. Moreover, the $(a,c)$ modulus in the third twisted sector appears
1827: while the non-compact direction is untwisted. As a consequence, the theory deformed by
1828: the Liouville potential has only one $(c,c)$ modulus in its spectrum, {\em
1829: and} it
1830: also has one $(a,c)$ modulus.
1831: (On the other hand, the resolved theory will have
1832: one $(a,c)$ modulus, together with one $(c,c)$ modulus.) It can be
1833: checked by direct calculation that this is consistent with the equations
1834: analyzed in \cite{Eguchi:2004yi}.
1835: The Gepner model analysis of this model is discussed briefly in appendix \ref{3332}.
1836:
1837:
1838: \section{Mirror Symmetry For Non-compact Gepner Models}\label{CFTmirror}
1839:
1840: In this section we address the question of identifying mirror pairs. In the
1841: case of compact Gepner models, when we specify the diagonal model as our
1842: starting point, we obtain the mirror model by modding out by the maximal
1843: discrete subgroup $H$ of the diagonal group $G$ that is
1844: consistent with space-time supersymmetry \cite{Greene:1990ud}. Subgroups $F$ of $H$ give rise to
1845: models that are mirror to models modded out by $H/F$.
1846:
1847: In the following we will argue that non-compact Gepner models behave very
1848: similarly, in their undeformed guise. In particular, we shall show that modding
1849: out by the maximal subgroup consistent with supersymmetry, we exchange $(c,c)$
1850: and $(c,a)$ deformations of the undeformed theory.
1851:
1852: A corollary of
1853: this statement is that a theory deformed by a given operator, will map after mirror symmetry to the mirror
1854: theory, deformed by the mirror operator. Thus, mirror symmetry applies to the deformed, regular theories as well.
1855:
1856: The main difference with the compact models is therefore that the mirror map
1857: includes the specification of the action of the mirror map on the deforming
1858: operator(s). Naturally, the specification is that one changes the right-moving
1859: R-charge of the deforming operator to find the mirror deformation. The mirror map extends to subgroups of $H$ as in the compact case.
1860:
1861:
1862: We believe it is best to illustrate the above general framework in a few
1863: examples. In the following section, we will then revisit these examples and
1864: see to what extent we can interpret mirror symmetry of the conformal field
1865: theories in a geometric framework.
1866:
1867: \subsection{Sixteen Supercharges}
1868:
1869: We recall that compact Gepner models at central charge $c=6$ lie in the moduli
1870: space of $K3$ compactifications of string theory. It is well-known
1871: \cite{Aspinwall:1994rg} that the mirror transform acts as an automorphism of
1872: the K3 moduli space and there are special points in the moduli space where
1873: there are fixed points. For our non-compact $(k;k)$ model at central charge
1874: $c=6$, we see that in its singular guise, the particular $8(k-1)$ deformations
1875: that we identified are indeed self-mirror. This can be seen in figure
1876: \ref{ALE}. Changing the sign of the right-moving R-charge exchanges the two
1877: individual sets of $4(k-1)$ deformation parameters that we distinguished
1878: previously. We note that this property of self-mirroring holds only for the
1879: singular model.\footnote{We do not advocate that the reader take the
1880: undeformed (well-defined) conformal field theory as a
1881: good description
1882: of the strongly coupled string background. We use it as a formal tool in
1883: arguing for the precise points of analogy and difference with compact Gepner
1884: models.}
1885:
1886: \begin{figure}
1887: \centering
1888: \includegraphics[scale=.92]{kkdef}
1889: \caption{Chiral-chiral and chiral-anti-chiral
1890: primaries in the $(k;k)$ model. Each dot corresponds to one chiral primary,
1891: identified by its non-compact left- and right-moving
1892: quantum numbers $2m$ and $2\tilde{m}$.\label{ALE}}
1893: \end{figure}
1894:
1895: As a consequence, we can discuss mirror symmetry for the weakly coupled deformed
1896: model. When we deform the $(k;k)$ model with a Liouville potential (consistent
1897: with sixteen supercharges), we are left with $4 (k-1)$ deformation parameters
1898: and a regular theory. It is mirror dual to the singular theory deformed by the
1899: winding potential (which geometrically gives rise to the cigar theory). The latter
1900: theory also has $4(k-1)$ deformation parameters, which can be mapped
1901: individually to their mirror images (using their worldsheet R-charges for each
1902: factor of the model).
1903:
1904: \subsection{Eight Supercharges}
1905:
1906: For compact Gepner models at central charge $c=9$, mirror symmetry maps one model onto another, exchanging $(c,c)$ and $(c,a)$ states. We will illustrate that this is the case for non-compact Gepner models as well. First we will treat a case in which we simply mod out by the maximal group, and thus
1907: obtain the mirror theory. Then we will show that one can mod out by a subgroup of the maximal subgroup and obtain mirror pairs. We will thus provide large classes of mirror non-compact Gepner models.
1908:
1909: \subsubsection{The $(k;2k,2k)$ Model}\label{zkoneorb}
1910:
1911: We note that the $(k;2k,2k)$ model, which is modded out only by the GSO
1912: $\IZ_{2k}$ group (in the Landau-Ginzburg formulation) has $(k-1)^2$ marginal
1913: $(c,c)$ deformations and $k-1$ marginal $(a,c)$ deformations, as discussed in
1914: section \ref{twocigars}. Orbifolds of this model were treated in section
1915: \ref{twocigarsorbifold}. The maximal group that we can divide out by
1916: consistent with supersymmetry is the orbifold group when $k_1=k$ and $k_2=1$.
1917: Then, substituting in the formulae for the massless moduli computed in that
1918: section, we get $k-1$ marginal $(c,c)$ deformations and $(k-1)^2$ marginal
1919: $(a,c)$ deformations. The orbifold indeed gives rise to the mirror theory.
1920: This is exactly analogous to the Greene-Plesser discussion of mirror conformal
1921: field theories associated to compact Calabi-Yau threefolds at a Gepner point.
1922:
1923: We want to compare these models to those in the literature. The $(k;2k,2k)$
1924: model modded out by the maximal group and deformed by the sum of Liouville
1925: potentials corresponds to the non-compact Calabi-Yau studied in
1926: \cite{Lerche:2000uy} to geometrically engineer pure $SU(k)$ gauge theory. The
1927: orbifold group restricts all the possible $(c,c)$ deformations to the $k-1$
1928: moduli studied in that paper which span the Coulomb branch of the $SU(k)$
1929: gauge theory. We will discuss the relation between our non-compact Gepner
1930: model, the associated Landau-Ginzburg model and this geometry in more detail
1931: in the next section.
1932:
1933: Moreover, we find that this model is mirror to a model that has no orbifold
1934: except the GSO projection, and deformed by the winding condensates in the two
1935: non-compact directions. That gives rise to the two-cigar model of
1936: \cite{Eguchi:2004ik}, as argued in that reference. From our perspective, we
1937: see that the double cigar deformations disallow all $(k-1)^2$ marginal $(c,c)$
1938: deformations in the unorbifolded model, and leaves only $k-1$ K\"ahler
1939: deformations. That matches precisely the analysis in \cite{Eguchi:2004ik}.
1940: Note that this gives a confirmation of our methodology: the counting in
1941: \cite{Eguchi:2004ik} is based on the spectrum identified by considering a
1942: regularized partition function \cite{Maldacena:2000kv, Israel:2004ir,
1943: Hanany:2002ev, Eguchi:2004yi}. Thus we find that the regularized partition
1944: function agrees with our intuitive arguments which find their basis in the
1945: Landau-Ginzburg model with negative power potentials (see sections \ref{LG}
1946: and \ref{orbifoldedLG}).
1947:
1948: As an example of the power of our simple description, we note that the model
1949: with Liouville deformations and no orbifold action is not directly related to
1950: the known models described in \cite{Lerche:2000uy, Eguchi:2004ik}.
1951: Nevertheless, we readily identify its mirror to be the orbifolded theory with
1952: cigar deformations in both the non-compact factors.
1953:
1954: \subsubsection{The Orbifolded $(2k,2k;k)$ Models}
1955:
1956: It should be clear now that the examples in sections \ref{twoMMorbifold} and
1957: \ref{twocigarsorbifold} give rise to two infinite classes of mirror
1958: non-compact Gepner models. We will illustrate this for one of the two classes
1959: since the other class behaves in almost every respect analogously.\footnote{We
1960: will signal the exception in the next section.}
1961:
1962: Recall that the $(2k,2k;k)$ model allows for a maximal $\IZ_k$
1963: orbifold
1964: consistent with supersymmetry. When the level $k$ is a product of two positive
1965: integers
1966: $k=k_1k_2$,
1967: we can orbifold by a non-maximal subgroup $F=\IZ_{k_1}$ which gives rise to a
1968: model which is mirror to a $\IZ_{k_2}$ orbifold of the same model. Indeed, the
1969: counting of $(k_1k_2^2-2k_2+1)$ marginal $(c,c)$ operators and
1970: $(k_2^2k_1-2k_1+1)$ marginal $(c,a)$ states in the first model (the
1971: $\IZ_{k_1}$ orbifold) is precisely the mirror of the counting of the model
1972: with $\IZ_{k_2}$ orbifold group, as can be seen by a exchanging the role of
1973: $k_1$ and $k_2$ (see figure \ref{mirrorsymmetryfigure}). Thus, we have very
1974: simply but explicitly demonstrated the existence of an infinite number of mirror pairs of singular non-compact Gepner models.
1975:
1976: We note that the model $(k;2k,2k)$ can be treated analogously. The proof
1977: of the generic fact that models modded out by
1978: a subgroup $F$ of $H$
1979: are mirror to models modded
1980: out by $H/F$ runs along very much the same lines as in the compact case \cite{Greene:1990ud}, for
1981: the undeformed theory. For the deformed theory we need to remember that the
1982: mirror map will mirror the deformation as well.
1983: \begin{figure}
1984: \centering
1985: \includegraphics[scale=.92]{mirrorsymmetryfigure}
1986: \caption{An explicit example: the $(2k,2k;k)$ models and its orbifolds by
1987: $\IZ_{k_1}$, with $k=30$.}
1988: \label{mirrorsymmetryfigure}
1989: \end{figure}
1990:
1991: \section{Mirror Non-compact Geometries and their Relation to $\,\IC^n/\Gamma$ Orbifolds}
1992: \label{geometrymirror}
1993: In this section we want to discuss some geometric realizations of the mirror
1994: map that we identified in the non-compact Gepner models above. We will see
1995: that we can identify various models with non-compact Calabi-Yau geometries,
1996: and approximate their mirror duals with abelian orbifolds of $\, \IC^3$ at
1997: large levels. At finite level, we find that the results of toric geometry
1998: acquire important modifications that lift certain
1999: moduli in our models.
2000:
2001: \subsection{The $(k;k)$ Model}
2002:
2003: In this model, the space-time background has sixteen supercharges. In
2004: analogy with the compact case, we identify the Calabi-Yau manifold that
2005: corresponds to the Landau-Ginzburg model by writing down the superpotential
2006: (augmented with the appropriate number of quadratic coordinates) in the non-compact weighted projective space $W\IC\IP^{4}$:
2007: \be
2008: w_1^k + w_2^2 + w_3^2 + c w_4^{-k} = 0.
2009: \ee
2010: The constant $c$ (multiplied by the coefficient of the first monomial)
2011: measures the strength of the Liouville deformation.
2012: By scaling the $\, \IC^\ast$ valued coordinate $w_4$ to one, we recuperate the
2013: equation:
2014: \begin{eqnarray}
2015: z_1^k + z_2^2 + z_3^2 + c &=& 0,
2016: \end{eqnarray}
2017: which describes an ALE space which is deformed by $c$ from its singular
2018: orbifold limit. We thus associate that geometry to the Landau-Ginzburg model, but we should
2019: keep in mind that this association is local, i.e. near the singularity.
2020: Asymptotically the spaces differ. We will see an example of the consequences of
2021: this difference later on.
2022: The matching of the $4(k-1)$ marginal deformations to the geometric moduli is
2023: well-known.
2024:
2025: \subsubsection{The Mirror Theory}
2026:
2027: The conformal field theory mirror to the above theory was argued to be the
2028: $(k;k)$ model with cigar deformation. In the appendix we recall, following
2029: \cite{Ooguri:1995wj, Kutasov:1995te}, that after a single T-duality, this
2030: model is mapped to a configuration of NS5-branes spread on a circle in a
2031: transverse plane. Moreover, using the explicit geometric description, it can
2032: be argued in great detail (see the appendix) that the NS5-branes spread on the
2033: circle map under that T-duality to an orbifold singularity of the type $\, \IC^2/\IZ_k$ at the tip
2034: of the cigar and the center of the minimal model disc. In particular, we see
2035: that the $\IZ_k$ orbifold that arises from the GSO projection on the side of
2036: the cigar mirror conformal field theory acts {\em geometrically} on the cigar
2037: and minimal model coordinates close to the tip of the cigar and the center of
2038: the disc.
2039:
2040: One should contrast this geometric action to the lack of such an action in the
2041: mirror geometry. Moreover, in this example, it becomes
2042: manifest that when the cigar/winding deformation is turned on in the singular
2043: theory, the deformation caps off the linear dilaton cylinder. As a consequence
2044: it gives rise to a (geometric) fixed point which allows for the localization
2045: of twisted sector states. That agrees with our prescription for keeping the
2046: $4(k-1)$ twisted sector states when turning on the cigar deformation.
2047:
2048: Note that there is a one-to-one match of the marginal supersymmetric
2049: deformations of the cigar times minimal model conformal field theory to the
2050: marginal supersymmetric deformations of the $\, \IC^2/\IZ_k$ orbifold (with
2051: $B$-field on the vanishing cycles) \cite{Ooguri:1995wj}.
2052:
2053: In the strictly infinite level limit, the match is expected, since then the
2054: cigar and disc flatten out completely. The two models become identical in that
2055: limit. However, at finite level $k$, the match is due to
2056: the rigidity of the sixteen supercharge construction, as will
2057: become clear when we discuss models
2058: with less supersymmetry.
2059:
2060: \subsection{The $(k;2k,2k)$ Model}
2061:
2062: Models with eight supercharges will show various new and interesting features.
2063: We will discuss two models in detail with increasing level of complexity. The
2064: model with levels $(k;2k,2k)$ and double Liouville deformation has a
2065: singularity that is well-described by a hypersurface in the
2066: (non-compact)weighted projective space of the form \cite{Eguchi:2004ik}:
2067: \begin{eqnarray}
2068: w_1^k + w_2^2 + w_3^2 + w_4^{-2k} + w_5^{-2k} &=& 0,
2069: \end{eqnarray}
2070: where the two coordinates $w_{4,5}$ are $\, \IC^\ast$ valued. The deforming
2071: monomials are in one-to-one correspondence with the $(k-1)^2$ marginal $(c,c)$
2072: deformations that we identified previously. After performing the $\IZ_{k}$
2073: orbifold, we find that only the complex structure
2074: deformations of the form $w_1^n (w_4 w_5)^{-k+n}$ for $n=0,1,\dots,k-2$ are
2075: invariant under the orbifold action. As mentioned earlier, this is precisely
2076: the geometry discussed in \cite{Lerche:2000uy, Hori:2002cd, Eguchi:2004ik}.
2077:
2078: \subsubsection{The Mirror Theory }
2079:
2080: Let us now turn to the mirror conformal field theory and see whether we can count the number of K\"ahler deformations of the mirror theory using geometric means. The first observation we make is that, just as in the case with sixteen supercharges, there is an infinite level limit in which the model flattens out, up to an overall orbifold action. We refer the reader to appendix \ref{NS5} for the basic arguments in favour of such a description. In that (strict) limit, the model, locally, near the tips of the cigars and the center of
2081: the minimal model disc becomes equivalent to the orbifold $\, \IC^3/\IZ_{2k}$. Again, we use that on this side of the mirror symmetry, the GSO projection acts geometrically and infer that the
2082: action of the $\IZ_{2k}$ on the three factors of $\, \IC^3$ is
2083: weighted
2084: as $\frac{1}{2k} (2k-1,2k-1,2)$. This orbifold is toric and the number of
2085: K\"ahler deformations of the geometry can be counted using toric geometry
2086: techniques\footnote{See, for instance, \cite{Aspinwall:1994ev, lustetal} and references
2087: therein for a review of these methods.} as follows.
2088:
2089: Consider the supersymmetric orbifold $\,\IC^3/\Gamma$, where the orbifold group of order $N$ is generated by
2090: \be \theta: (z_1,z_2, z_3) \longrightarrow (\omega^{a_1} z_1,\omega^{a_2} z_2, \omega^{a_3} z_3) \quad \text{such that}\quad \sum_{i}a_i = 0 \quad \text{mod} \quad N \,.
2091: \ee
2092: The counting of K\"ahler deformations proceeds as follows.
2093:
2094: \begin{itemize}
2095: \item Consider all powers of $\theta$ and list their exponents in multiples of
2096: $2 \pi i$ such that they fall in the range $1/N \times (0,N-1)$. Label them
2097: $(g_1, g_2, g_3)$.
2098:
2099: \item The K\"ahler moduli are in one to one correspondence with those powers
2100: of $\theta$ that satisfy the following conditions: \be\label{constraint}
2101: \sum_{i=1}^{n} g_i = 1 \quad \text{with} \quad 0\le g_i < 1 \,. \ee The
2102: power of $\theta$ tells you in which twisted sector the modulus appears.
2103: \end{itemize}
2104:
2105: Before we proceed further let us also briefly recall how one draws the toric
2106: diagram corresponding to the non-compact orbifold. This will turn out to be
2107: useful to compare and contrast the spectrum of these non-compact toric
2108: orbifolds with the Landau-Ginzburg computation of the spectrum of moduli for
2109: the non-compact Gepner models.
2110:
2111: Given an action of $\theta=(\theta_1, \theta_2, \theta_3)$ on $\,\IC^3$ as
2112: above, we first find basis vectors $\{D_1, D_2, D_3\}$ that satisfy \be
2113: \sum_{i=1}^{3}\theta_i \, (D_i)_a = 0 \quad \text{mod}\quad N \,. \ee These
2114: three basis vectors generate the toric fan. One solution is $(D_i)_3 = 1\
2115: \forall\ i$. One can find two other linearly independent solutions to this
2116: equation. Thus, neglecting the third coordinate of the vectors and plotting only the other two solutions to the above equation, one gets points on a plane. That this must be so follows from the Calabi-Yau condition. These points define the Newton polyhedron corresponding to the non-compact Calabi-Yau. The basis vectors $D_i$ form a cone over the polyhedron which allows one to draw the toric diagram for the Calabi-Yau threefold in the plane. Now, for each power of $\theta^{n}=(g^{(n)}_1,g^{(n)}_2,g^{(n)}_3)$ that satisfies the above two constraints (i.e. for every K\"ahler deformation),we add another vector (interior or boundary point in the toric diagram) $E_n$ given by
2117: \be
2118: E_n = \sum_{i=1}^{3}\, g^{(n)}_i D_i \,.
2119: \ee
2120:
2121: Let us apply the above algorithm to compute the K\"ahler moduli and draw the
2122: toric diagram, for our case where $\Gamma=\IZ_{2k}$ (i.e. $N=2k$) and \be
2123: (\theta_1, \theta_2, \theta_3) = \left(\frac{2k-1}{2k}, \frac{2k-1}{2k},
2124: \frac{2}{2k}\right)\,. \ee One can check that there are $k$ K\"ahler
2125: deformations which arise from the twisted sectors $k, k+1, \dots, 2k-1$. This
2126: is one more than the number of $(c,c)$ deformations we obtained from the
2127: Landau-Ginzburg computation in section \ref{twocigarsorbifold}. Let us draw
2128: the toric diagram corresponding to this orbifold. We choose a basis of vectors
2129: \be D_1 = (0,1,1) \quad D_2 =(0,-1,1) \quad D_3=(k,0,1) \,. \ee
2130: We add $k$ points, corresponding to the $k$ K\"ahler deformations, which are elements
2131: in the $(c,a)$ ring, in the twisted sectors, whose coordinates are given by
2132: \be E_{k}=(0,0,1)\,, \quad E_{k+1} = (1,0,1)\,, \quad E_{k+2} = (2,0,1)\,, \quad \ldots
2133: \quad E_{2k-1} = (k-1,0,1) \,. \ee
2134: \begin{figure}
2135: \centering
2136: \includegraphics[scale=.75]{toric.eps}
2137: \caption{The toric diagram for $\,\IC^3/\IZ_{2k}$ for $k=4$. Note hat $E_{k=4}$ corresponds to a boundary point while all other added points lie in the interior of the Newton polyhedron. We have also shown a possible triangulation of the polygon, corresponding to a particular resolution of the singularity.\label{toric}}
2138: \end{figure}
2139: These points are plotted in figure \ref{toric}.
2140: Since $E_k=\half(D_1+D_2)$ in our $\IZ_{2k}$ orbifold it follows that this
2141: point will always be on the boundary of the toric diagram.
2142:
2143: When we compare the
2144: two calculations of the moduli, we see that the difference arises in
2145: the $2k-1$st twisted sector of the GSO orbifold, since, in the conformal field
2146: theory, we did not find a marginal K\"ahler deformation in this twisted sector.
2147: Let us study in somewhat more detail how this difference comes about.
2148:
2149:
2150: It is well known that the points in the toric diagram can also be associated
2151: to exceptional divisors of the resolution of the orbifold. This is the twist-field-divisor map discussed in \cite{Aspinwall:1994ev}. The topology and
2152: intersection numbers of these divisors can be obtained in a straightforward
2153: manner by using the dual toric diagram. The boundary point corresponds to
2154: non-compact divisors and have the topology $\IP^1 \times\, \IC$. Therefore,
2155: the existence of the resolution corresponding to
2156: $E_k$ in the conformal field theory might seem
2157: surprising, given that the (strictly) normalizable deformations in the conformal field theory can be
2158: associated to compact cycles in the geometry. This can be understood as
2159: follows: the $k$th power of the orbifold action is trivial on one of the three
2160: complex directions in $\, \, \IC^3$, and creates a singularity that stretches along
2161: a complex line in $\, \, \IC^3$. The deformation is therefore akin to a K\"ahler
2162: deformation of a $\, \, \IC^2/\Gamma$ orbifold, embedded in $\, \IC^3$. However, it is
2163: important to note that the $\,\IC$ direction that is left invariant in the
2164: $k$th twisted sector, is the direction that is compactified
2165: in the conformal
2166: field theory by the addition of the Landau-Ginzburg
2167: potential.
2168: So in the
2169: conformal field theory, all the exceptional divisors become compact. A second
2170: effect of the compactification is that only $k-1$ of
2171: the deformations are
2172: linearly independent. The Landau-Ginzburg counting of chiral primaries gives an easy method
2173: to understand which of the exceptional divisors are chosen as a basis in the
2174: conformal field theory.
2175:
2176: \begin{figure}
2177: \centering
2178: \includegraphics[scale=0.75]{dual.eps}
2179: \caption{The dual toric diagram for $\,\IC^3/\IZ_{2k}$ with $k=4$. We use the same alphabet to denote the point in the toric diagram and the divisor corresponding to it in the dual toric diagram. Note that $E_{k}$ (which is $E_4$ in our case) is non-compact while all the other exceptional divisors are compact. \label{dual}}
2180: \end{figure}
2181:
2182: Note that the $(c,a)$ deformation which is the identity operator in the
2183: minimal model, sets the volume of the compact factor and simultaneously
2184: the volume of the two-cycle at the $\, \IC^2/\IZ_{2}$ singularity. The operator has
2185: charges $r=(0,0,0;0;k,k)$ and $\tilde{r} = (0,0,0;0;-k,-k)$ as can be seen
2186: from the Landau-Ginzburg model description, or appendix \ref{k2k2k}.
2187:
2188: \subsubsection{The $(2k,2k;k)$ Model}
2189:
2190: The toric abelian orbifold that is related to the $(2k,2k;k)$ model is the
2191: same as the one we discussed before. The differences with the previous model
2192: lie in the fact that we now compactify two coordinates, which reduces the
2193: number of K\"ahler moduli by two. By analyzing the spectrum of the conformal field theory as
2194: before, we find that in the toric diagram corresponding to the flat space
2195: orbifold, the twist fields associated to the divisors $E_{k}$ and $E_{k+1}$ are
2196: excluded in the conformal field theory.
2197:
2198: However, we gain a K\"ahler modulus in the $0$th twisted sector which sets the
2199: overall volume of the two compact factors. It is the identity operator in the
2200: minimal model factors, and the winding operator in the cigar factor. It has
2201: charges $r=(0,0,0;0,0;k)$ and $\tilde{r}=(0,0,0;0,0;-k)$ as can be seen from
2202: the Landau-Ginzburg description or from appendix \ref{2k2kk}. This is
2203: consistent with the picture developed in \cite{Giveon:1999px}: the effective
2204: string coupling at the tip of the cigar sets the volume of the resolved
2205: cycles.
2206:
2207: Note that in the previous example, we had a slight refinement of the picture
2208: developed in \cite{Giveon:1999px}. Namely, in the presence of the two
2209: non-compact directions, it is the modulus associated to the $\IZ_2$ orbifold
2210: singularity at the tips of the cigars that sets the volume of the internal compact space.
2211:
2212: \subsubsection{The $\IZ_{k_1}$ Orbifold Of The $(k;2k,2k)$ Model}
2213: In order to understand some more general features of the spectra of the
2214: conformal field theory and how they fit into the spectrum of the flat space
2215: orbifold approximation, let us study the orbifold models studied in sections
2216: \ref{twoMMorbifold} and \ref{twocigarsorbifold}.
2217:
2218: The flat space approximation to the conformal field theory is given by a $\,\IC^3/\Gamma$ orbifold, generated by the elements
2219: \begin{align}
2220: g_0 &= \frac{1}{2k}(2k-1,2k-1,2) \quad\text{and}\cr
2221: g_1 &= \frac{1}{2k}(k_2,2k-k_2,0) \,.
2222: \end{align}
2223: Using the algorithm discussed earlier, one can easily draw the toric diagram
2224: associated to this singularity as in figure \ref{orbtoric}.
2225: \begin{figure}
2226: \centering
2227: \includegraphics[scale=0.75]{orbdots1}
2228: \caption{The toric diagram for $\,\IC^3/(\IZ_{2k}\times \IZ_{k_1})$ with $k=6$ and $k_1=2$. \label{orbtoric}}
2229: \end{figure}
2230: \begin{figure}
2231: \centering
2232: \includegraphics[scale=0.75]{orbdots3}
2233: \caption{Spectrum of the $\IZ_{k_1}$ orbifold of the $(k;2k,2k)$ conformal field theory for $k=6$ and $k_1=2$ are denoted by the filled dots. The unfilled dots show those points in the flat space approximation which are not included in the conformal field theory. \label{orbthree}}
2234: \end{figure}
2235: \begin{figure}
2236: \centering
2237: \includegraphics[scale=0.75]{orbdots2}
2238: \caption{Spectrum of the $\IZ_{k_1}$ orbifold of the $(2k,2k;k)$ conformal field theory for $k=6$ and $k_1=2$ are denoted by the filled dots. The unfilled dots show those points in the flat space approximation which are not included in the conformal field theory. \label{orbtwo}}
2239: \end{figure}
2240: The spectrum of the exact conformal field theory has already been discussed in the main part of
2241: the paper. If we now plot the spectrum of the exact conformal field theory that corresponds to the $(k;2k,2k)$ model along the same lines, we get figure \ref{orbthree}. We have shown which elements of the flat space approximation get lifted in going to the exact conformal field theory description.
2242:
2243: \subsubsection{The $\IZ_{k_1}$ Orbifold Of The $(2k,2k;k)$ Model}
2244: For this model the flat space approximation is identical to the one discussed
2245: above\footnote{For the $(2k,2k;k)$ example, the canonical group element $g_0$
2246: is given by the inverse of the one we have written here but the group
2247: generated and the orbifold of $\,\IC^3$ associated to it is unchanged. }
2248: and it is drawn in figure \ref{orbtoric}. The conformal field theory analysis
2249: is, however, different and the K\"ahler deformations which are kept are shown
2250: in the figure \ref{orbtwo}.
2251:
2252:
2253: We mentioned before that in the toric diagram for the flat space orbifolds,
2254: the boundary points correspond to non-compact divisors. Nevertheless as we saw
2255: in the simpler orbifold example, in the exact conformal field theory
2256: description, some of these directions are compactified. These directions are
2257: those along which a potential of the form $\Phi^n$, $n > 0$ is turned on and
2258: which flow into a minimal model in the IR. However, the twist fields that correspond to divisors (via the twist-field-divisor map) which extend along directions that remain non-compact should be excluded in the conformal field theory as these do not lead to normalizable deformations. That this is
2259: so can be checked in our examples.
2260:
2261: For instance, in the $(2k,2k;k)$ model, in figure \ref{orbtwo}, the excluded
2262: points on the boundary of the toric diagram correspond to K\"ahler
2263: deformations which are in the twisted sectors $g_1^{\b}$ and
2264: $g_0^{k}g_1^{\b}$, with $\b = 1, \ldots k_1-1$. Their respective $U(1)$
2265: charges are given by \be \frac{1}{2k}(\b k_2,2k-\b k_2,0)\quad\text{and}\quad
2266: \frac{1}{2k}(k+\b k_2,k-\b k_2,0)\quad \text{for}\quad n\in \{1, \ldots, k_1-1
2267: \}\,. \ee As one can see, the twist fields are uncharged under the $U(1)$
2268: that acts on the non-compact direction and the divisors that correspond to
2269: these fields are non-compact. These are subsequently excluded from the
2270: conformal field theory spectrum.
2271:
2272: However, further work is required to understand in full generality which states of the flat space orbifold are retained in a given conformal field theory of the type studied in this article.
2273:
2274: \subsection{Relation to NS5-brane set-ups}
2275: The relation of the above orbifold approximations to NS5-brane set-ups has been discussed in the literature. These toric abelian orbifolds of $\, \, \IC^3$ can be
2276: mapped one-to-one to NS5-brane configurations that wrap holomorphic curves. The
2277: holomorphic curves can in turn be described by dimers that are
2278: systematically reconstructed from the abelian orbifold group. In the
2279: particular case above, one obtains hexagon tilings of the plane with
2280: labelings determined by the weights of the GSO projection and further
2281: orbifolding. They can be determined by straightforwardly generalizing
2282: the examples in \cite{Hanany:2005ve, Franco:2005rj}.
2283:
2284: In the sixteen supercharge case, one can show explicitly that the non-compact
2285: Gepner model captures the near-horizon doubly scaled limit of the backreacted
2286: NS5-brane geometry (as we have recalled in the appendix).
2287: In the case of eight supercharges, our non-compact Gepner
2288: models capture a near-horizon doubly scaled limit of the backreaction of
2289: NS5-branes wrapped on the holomorphic curves coded by the dimer corresponding
2290: to our non-compact Gepner model.
2291:
2292: \section{Conclusions and future directions}\label{conclusions}
2293:
2294: We have shown that the Gepner formalism for constructing modular invariant
2295: partition functions carries over to the asymptotic partition function of
2296: non-compact Gepner models. The analogy was then further developed in a
2297: discussion of the symmetry groups of the Gepner models, a classification of
2298: subgroups consistent with supersymmetry, and the discussion of how orbifolding
2299: by the maximal group can give rise to mirror models.
2300:
2301: Secondly, we discussed the deep throat region of the non-compact Gepner models
2302: and how to obtain the chiral primary states that are normalizable at weak
2303: coupling from the conformal field theory. We then discussed Landau-Ginzburg
2304: descriptions of both compact and non-compact Gepner models and extended the
2305: existing techniques to analyze Landau-Ginzburg orbifolds such that they
2306: applied to the non-compact models under discussion. This led to an intuitive
2307: understanding of which modes become normalizable, and which modes are lifted
2308: by a momentum or winding potential. The counting of deformations becomes very
2309: tractable in the Landau-Ginzburg formalism. For completeness, we matched it
2310: onto a more intricate conformal field theory counting.
2311:
2312: We used these results to argue that mirror symmetry can be implemented in non-compact Gepner models. When taking into account all possible deformations of the linear dilaton theory, it becomes analogous to the compact case. However, one always needs to keep in mind that the choice of deformation needs to be mirrored when discussing the deformed theories. As expected, we saw that in conformal field theory, mirror symmetry is implemented by a change in sign of the right-moving R-charge. The systematic treatment of symmetry groups allowed us to generate infinite classes of mirror pairs.
2313:
2314: Indeed, in non-compact Gepner models, one can cancel off positive and negative contributions to the central charges of the individual factors, thus allowing for infinite classes of models that also have a small curvature limit. In such small curvature limits, we argued that one recuperates flat space with an
2315: overall orbifold action. We identified such a limit, along with the orbifold, and showed that the conformal field theory matches with a flat space orbifold in the infinite level limit. At finite level, we identified subtle differences in the spectrum of the conformal field theory and the toric abelian orbifold singularity. It would be interesting to find a general rule that tells us, a priori, which modes of the toric orbifold are retained in the conformal field theory.
2316:
2317: There are a large number of future directions that one can pursue. For
2318: instance, one can generalize the Landau-Ginzburg models to models with
2319: fractional levels in the non-compact directions. One can also apply these
2320: conformal field theory techniques to describe the spectrum of chiral primaries
2321: and mirror theories for the heterotic string on non-compact Landau-Ginzburg
2322: orbifolds.
2323:
2324: Another direction that we had in mind while embarking upon this investigation
2325: is the following. We have an orbifold approximation to particular non-compact
2326: Gepner models. Setting fractional or regular (i.e. physical) branes at such
2327: toric orbifold singularities is one way to engineer interesting quiver gauge
2328: theories. The toric data allows us to compute the superpotential on the brane
2329: at the orbifold singularity using techniques that are very well developed.
2330: Extending these results to determine the worldvolume superpotentials for
2331: branes in non-compact Gepner models would be extremely interesting as such
2332: results are not yet available using the exact boundary state description of
2333: D-branes in these models.
2334:
2335: Furthermore, Seiberg duality is well-understood in the context of the toric
2336: quiver gauge theories \cite{Feng:2000mi, Beasley:2001zp, Cachazo:2001sg}. We
2337: would like to study whether one can understand Seiberg duality for D-branes in
2338: these almost toric spaces \cite{Eguchi:2003ik, Ashok:2005py,
2339: Fotopoulos:2005cn} microscopically, as in \cite{Murthy:2006xt,
2340: Ashok:2007sf}. The fact that we have a microscopic description of the
2341: near-horizon doubly scaled limit of these backgrounds as well as a tunable
2342: level, should give us further computational control.
2343:
2344: \section*{Acknowledgments}
2345: We are grateful to Agostino Butti, Eleonora Dell'Aquila, Bogdan Florea,
2346: Davide Forcella, Amihay Hanany and Ruben Minasian for helpful conversations.
2347: Our work was supported in part
2348: by the EU under the contract MRTN-CT-2004-005104.
2349:
2350: \appendix
2351:
2352: \section{Non-compact Gepner model analysis}\label{CFTanalysis}
2353:
2354: In this appendix we relate the counting of marginal deformations that we
2355: performed
2356: in a Landau-Ginzburg language to a more elaborate enumeration of states
2357: in a standard conformal field theory formalism.
2358: See \cite{Eguchi:2004ik}
2359: for a detailed discussion and further examples.
2360:
2361: A state in the $(k_1,...,k_p;l_1,...,l_q)$ model (restricted to the internal
2362: conformal field theory only) is associated to the left and right charges
2363: $$r=(s_1,...,s_{p+q};n_1,...,n_p;2m_1,...,2m_q)$$
2364: $$\tilde{r}=(\tilde{s}_1,...,\tilde{s}_{p+q};\tilde{n}_1,...,\tilde{n}_p;2\tilde{m}_1,...,2\tilde{m}_q)$$
2365: as well as the compact spins $j_1,...,j_{p}$ and the non-compact spins
2366: $j_{p+1},...,j_{p+q}$. The spins are the same on the left and on the right,
2367: since we consider diagonal partition functions. The $2 \beta_0$-orbifold
2368: imposes integral R-charges, and allows the difference of left and right
2369: charges $r-\tilde{r}$ to be an even multiple of the Gepner vector $\beta_0$.
2370: The orbifolds $\beta_i$ that align the periodicities of the fermions allow
2371: additional even differences between the left fermion numbers $s_i$ and the
2372: right fermion numbers $\tilde{s}_i$.
2373:
2374: Chiral primary operators are chiral primaries in each conformal field theory factor. From unitarity of the non-compact Gepner models (see appendix \ref{unitarity}) it follows that in each factor separately we satisfy the equation $\pm Q_i = 2 h_i$ for chiral (respectively anti-chiral) primaries and similarly for the right-movers. Marginality of the deformations implies that we need to satisfy the equations $$ Q = 2 \beta_0\cdot r = \pm 1 \quad \tilde{Q} = 2 \beta_0\cdot \tilde{r} = \pm 1. $$
2375:
2376: Solving this set of equations is straightforward but tedious because of the
2377: equivalences that exist for the minimal model quantum numbers:
2378: $$n_i \equiv n_i +2k_i,\quad s_i \equiv s_i+4,\quad (j,n,s) \equiv (\frac{k-2}{2}-j,n
2379: -k,s+2)$$
2380: The same kind of equivalences hold in the
2381: non-compact factor:
2382: $$2m_i \equiv 2m_i +2l_i,\quad s_i \equiv s_i+4,\quad (j,2m,s) \equiv (\frac{k+2}{2}-j,2m
2383: -k,s+2)$$
2384: That is one technical reason why the Landau-Ginzurg
2385: method is more efficient to count (anti)chiral operators.
2386: However it is possible to perform the counting in each individual example that we
2387: treated in the bulk of the paper. We do this analysis example by example.
2388:
2389: \subsection{The $(k;k)$ Model}
2390: In the $(k;k)$ model, the Gepner vector is $\beta_0=(-1,-1;1;-1)$. Looking for
2391: chiral primary operators, we find
2392: \begin{itemize}
2393: \item $k-1$ $(c,c)$ states in the untwisted sector with charges:
2394: $$ r=(0,0;n;k-n)=\tilde{r}, \quad j_1=n, \quad j_2=k-n, \quad 0 \le n \le k-2 $$
2395: \item $k-1$ $(a,c)$ states in the first twisted sector:
2396: $$ r=(0,0;-n;-k+n), \quad \tilde{r}=r-2\beta_0,\quad j_1=n, \quad j_2=k-n, \quad 0 \le n \le k-2 $$
2397: \item 1 $(c,c)$ state in each $\alpha$-twisted sector ($1 \le \alpha \le k-1$):
2398: $$ r=(0,0;\alpha-1;k-\alpha+1), \quad \tilde{r}=r-2\alpha \beta_0,\quad
2399: j_1=\alpha-1, \quad j_2=k-\alpha+1 $$
2400: \item 1 $(a,c)$ state in each $(\alpha+1)$-twisted sector ($1 \le \alpha \le k-1$):
2401: $$ r=(0,0;-\alpha+1;-k+\alpha-1), \quad \tilde{r}=r-2(\alpha+1) \beta_0, \quad
2402: j_1=\alpha-1, \quad j_2=k-\alpha+1 $$
2403: \end{itemize}
2404: For each $(c,c)$ state we also find a $(a,a)$ state
2405: and
2406: each $(c,a)$ state is similarly paired with an $(a,c)$ state.
2407: It is straightforward to match all these states with the ones we found more
2408: fluently in the bulk of the paper with the Landau-Ginzburg methods.
2409:
2410: For each state, described by its quantum numbers $r$, $\tilde{r}$, $j_1$ and
2411: $j_2$, we can write the corresponding closed string vertex
2412: operator:
2413: \be V_{r,\tilde{r}}^{j_1,j_2} = V^{j_1}_{n,s_1;\tilde{n},\tilde{s}_1}
2414: V^{j_2}_{2m,s_2;2\tilde{m},\tilde{s}_2} \,. \ee
2415: Here $V^{j_1}_{n,s_1;\tilde{n},\tilde{s}_1}$ is a vertex operator of
2416: quantum numbers $(j_1,n,s_1)$ in the minimal model,
2417: while $V^{j_2}_{2m,s_2;2\tilde{m},\tilde{s}_2}$ is a vertex operator
2418: of quantum numbers $(j_2,n,s_2)$ in the noncompact factor.
2419: If we denote the asymptotic coordinates along the cylinder as $\rho$ and
2420: $\theta$
2421: and the bosonized complex fermion by $H$, we have the asympotic expression for the
2422: vertex
2423: operators (see e.g. \cite{Hosomichi:2004ph}):
2424: \be V^{j}_{2m,s;2\tilde{m},\tilde{s}} = \exp\left[ \sqrt{\frac{2}{k}}
2425: \left( (j-1)\rho +im \theta_L +i\tilde{m} \theta_R \right)
2426: +isH_L+i\tilde{s}H_R \right] \,.\ee
2427:
2428: \subsection{The $(2k,2k;k)$ Model}
2429: \label{2k2kk}
2430: Let us briefly mention the subtlety that the $\beta_0$ vector of Gepner is defined in
2431: principle in all factors of the models. The reason we can consider
2432: its action separately in the internal factors only without encountering
2433: further difficulties lies in the fact that we first of all only consider even
2434: multiples of $\beta_0$, and that moreover an even multiple of $\beta_0$ is
2435: equivalent to the same vector with only non-zero internal entries, up to the
2436: vectors $\beta_i$. That is true for all three-factor models we consider (and
2437: it is therefore also true for the two-factor model when we incorporate two
2438: further flat directions).
2439:
2440: Having dispensed of that subtlety in comparing the two methods, we
2441: can again find the marginal deformations directly in the non-compact Gepner model:
2442: \begin{itemize}
2443: \item The $(k-1)^2$ untwisted $(c,c)$ states have quantum numbers:
2444: $$ r=(0,0,0;a,b;c)=\tilde{r}, \quad j_1=a, \quad j_2=b, \quad j_3=c$$
2445: $$ 0 \le a,b \le 2k-2, \quad 2 \le c \le k, \quad a+b+2c=2k $$
2446: \item The $k-1$ twisted $(a,c)$ states have quantum numbers:
2447: $$r=(0,0,0;-2k+\alpha,-2k+\alpha;k-\alpha),\quad \tilde{r}=r-2\alpha \beta_0$$
2448: $$j_1=2k-\alpha, \quad j_2=2k-\alpha, \quad j_3=\alpha-k, \quad k+2 \le \alpha \le 2k \,.$$
2449: where $\alpha$ labels the twisted sector in which these states live.
2450: \end{itemize}
2451: \noindent
2452: The corresponding closed string vertex operators are given by
2453: \be V_{r,\tilde{r}}^{j_1,j_2,j_3} = V^{j_1}_{n_1,s_1;\tilde{n}_1,\tilde{s}_1}
2454: V^{j_2}_{n_2,s_2;\tilde{n}_2,\tilde{s}_2}
2455: V^{j_3}_{2m,s_3;2\tilde{m},\tilde{s}_3}\,.\ee
2456:
2457: \subsection{The $(k; 2k, 2k)$ Model}\label{k2k2k}
2458:
2459: The marginal deformations are written as follows:
2460: \begin{itemize}
2461: \item The $(k-1)^2$ untwisted $(c,c)$ states have quantum numbers:
2462: $$ r=(0,0,0;a;b,c)=\tilde{r}, \quad j_1=a, \quad j_2=b, \quad j_3=c $$
2463: $$ 0 \le a \le k-2, \quad 2 \le b,c \le 2k, \quad 2a+b+c=2k $$
2464: \item The $k-1$ twisted $(a,c)$ states have quantum numbers:
2465: $$ r=(0,0,0;-k+\alpha;-\alpha,-\alpha), \quad \tilde{r}=r-2\alpha \beta_0$$
2466: $$j_1=k-\alpha, \quad j_2=\alpha, \quad j_3=\alpha, \quad 2 \le \alpha \le k $$
2467: where $\alpha$ labels the twisted sectors as before.
2468: \end{itemize}
2469: \noindent
2470: The corresponding closed string vertex operators are given by:
2471: \be V_{r,\tilde{r}}^{j_1,j_2,j_3} = V^{j_1}_{n,s_1;\tilde{n},\tilde{s}_1}
2472: V^{j_2}_{2m_1,s_2;2\tilde{m}_1,\tilde{s}_2}
2473: V^{j_3}_{2m_2,s_3;2\tilde{m}_2,\tilde{s}_3}\,.\ee
2474:
2475: For the orbifold models as well, one can perform the tedious exercise, thus
2476: affirming that the Landau-Ginzburg formalism is indeed more efficient.
2477:
2478: \subsection{The $(3,3,3;2)$ Model}\label{3332}
2479:
2480: There is an extra subtlety that we need to address for this model. From the
2481: conformal field theory point of view, it is easiest to
2482: ignore the quadratic Landau-Ginzburg model with a trivial infra-red fixed
2483: point,
2484: and to work with an even number of internal conformal field theory factors. It then follows that the vector
2485: \begin{equation}
2486: 2\beta_0=(-2,-2,-2,-2,-2;2,2,2;-2)
2487: \end{equation}
2488: in the full light-cone conformal field theory is not equivalent modulo the
2489: vectors $\beta_i$ to the vector $2 \gamma_0 =(0,0,0,0,0;2,2,2;-2)$ which
2490: represents the $g_0$ action on the Landau-Ginzburg internal conformal field
2491: theory. Yet, it can be show that the Landau-Ginzburg model does correctly
2492: count the conformal field theory chiral-chiral and chiral-anti-chiral states.
2493: Since it is only the $(c,a)$ state in the $3$-twisted sector that is crucial
2494: to us in this example, let us show how to identify that state in the conformal
2495: field theory. It corresponds to the state with charges $r=(0,0,0,0,0;0,0;2)$
2496: on the left, as indicated by the Landau-Ginzburg model. The right-moving
2497: charges are computed by observing that we are in the $3$-twisted sector. We
2498: obtain the charge (up to equivalences in the charge lattice)
2499: $\tilde{r}=(0,0,0,0,-2;0,0;0)$ (and we remain diagonal in the non-compact
2500: quantum number $j=1$). Indeed, the state with these charges is in the spectrum
2501: of the theory.
2502:
2503: Note that we had to adjust the precise identification of the state in the conformal field theory. This is typical of the model with an even number of factors. The final vertex operator is chiral and bosonic on the left, and anti-chiral and purely made of fermions on the right.
2504:
2505: \section{T-duality to NS5-branes revisited}\label{NS5}
2506: In this appendix, we recall the relation of the $(SU(2)/U(1) \times
2507: SL(2,R)/U(1))/\IZ_k$ coset model to the near-horizon geometry of a particular
2508: constellation of NS5-branes \cite{Ooguri:1995wj, Sfetsos:1998xd, Giveon:1999px}. It is useful to revisit this exercise because we will be able to explicitly identify the region of
2509: space-time in which the NS5-branes reside with the presence of a
2510: patch isomorphic to $\,\IC^2/\IZ_k$ in the T-dual. This fact is used as an
2511: argument in
2512: section \ref{geometrymirror}.
2513:
2514: We recall that the supergravity background generated by
2515: parallel NS5-branes stretching in
2516: the $x^{\mu=0,1,2,3,4,5}$-directions in the string frame is:
2517: \begin{eqnarray}
2518: ds^2 &=& \eta_{\mu\nu} dx^{\mu} dx^{\nu} + H(x^i) dx^i dx_i \nonumber \\
2519: e^{2 \Phi} &=& g_s^2 \, H(x^i) \nonumber \\
2520: H_{ijk} &=& - {\epsilon^l}_{ijk} \partial_l H(x^i)
2521: \end{eqnarray}
2522: where the harmonic function $H$ is determined by the positions of
2523: the NS5-branes $x^{i=6,7,8,9}_a$:
2524: \begin{eqnarray}
2525: H(x^i) &=& 1 + \sum_{a=1}^k \frac{\alpha'}{|x^i-x^i_a|^2}
2526: \end{eqnarray}
2527: We concentrate on $k$ NS5-branes spread evenly
2528: on a topologically trivial circle of coordinate
2529: radius $\rho_0$ in the $(x^6,x^7)$
2530: plane (see figure \ref{planeALF}). We recall from
2531: \cite{Sfetsos:1998xd, Israel:2005fn} that after the coordinate
2532: change $r \ge 0$ and $\theta \in {[} 0 , \pi/2 {]} $
2533: \begin{eqnarray}
2534: (x^6,x^7) &=& \rho_0 \cosh r \sin \theta \, (\cos \psi,\sin \psi) \nonumber \\
2535: (x^8,x^9) &=& \rho_0 \sinh r \cos \theta \, (\cos \phi,\sin \phi)\,
2536: \label{cartcoords}
2537: \end{eqnarray}
2538: and taking a near-horizon doubly scaled limit in which $\rho_0 \rightarrow 0$
2539: and $g_s \rightarrow 0$ and in which
2540: $\rho_0 g_s / \sqrt{\alpha'}$ (and $\alpha'$) is kept fixed \cite{Giveon:1999px},
2541: and after neglecting the localization of the NS5-branes on the circle,
2542: we obtain the NS5-brane background:
2543: \begin{eqnarray}
2544: ds^2 &=& dx^{\mu} dx_{\mu} + \alpha' k \,
2545: \left[ dr^2+d \theta^2 + \frac{\tanh^2 r \ d \phi^2 + \tan^2 \theta \ d \psi^2
2546: }{1+\tan^2 \theta \tanh^2 r} \right] \, ,
2547: \nonumber \\
2548: e^{ 2 \Phi} &=& \frac{g_\textsc{eff}^2 }{\cosh^2 r - \sin^2 \theta} \, ,
2549: \nonumber \\
2550: B &=& \frac{\alpha' k }{1+\tan^2 \theta \tanh^2 r} \ d \phi \wedge d \psi \,
2551: \label{tdualNS5geom}
2552: \end{eqnarray}
2553: where the effective string coupling constant is
2554: \begin{equation}
2555: g_\textsc{eff} = \frac{\sqrt{k \alpha'}g_s}{\rho_0}.
2556: \end{equation}
2557: We refer to \cite{Israel:2005fn} for the detailed calculation.
2558:
2559: A first observation to make is that the NS5-branes are located
2560: at $r=0$ and $\theta=\pi/2$. Moreover, the fact that the coordinate
2561: radius of the NS5-brane ring has gone to zero, has been compensated by the
2562: fact that the harmonic function and radial metric
2563: component has blown up
2564: and in such a way that the interior of the NS5-brane ring still has a
2565: proper size of order $\sqrt{k \alpha'}$. It is difficult to press together
2566: NS5-branes.
2567: That's an important aspect of the solution, since in the interior, near $r=0$
2568: and $\theta=0$, we simply find a portion of flat space.
2569:
2570: After T-duality in the angular direction $\psi$ around which
2571: the NS5-branes have been sprinkled, we found the T-dual geometry:
2572: \begin{eqnarray}
2573: ds^2 &=& \alpha' k \left[ d r^2 + \tanh^2 r \, \left( \frac{d\chi}{k}\right)^2
2574: + d\theta^2 + \mathrm{cotan}^2 \theta \, \left( \frac{d\chi}{k}-d\phi \right)^2 \right], \nonumber \\
2575: e^{2 \Psi}&=& \frac{g_\textsc{eff}}{k} \ \frac{1}{\cosh^2 r \ \sin^2 \theta}
2576: \end{eqnarray}
2577: where $\chi$ is an angle parameterizing the T-dual circle.
2578: The background is recognized as a vector $\IZ_k$ orbifold of the
2579: product of coset conformal field theory geometries
2580: $SU(2)_k /U(1) \, \times SL(2,R)_k /U(1)$.
2581:
2582: We want to add a couple of remarks to the analysis in \cite{Israel:2005fn}
2583: to which
2584: we refer the reader for further comments. In
2585: particular, we first note the singularity in both the metric
2586: and the dilaton at $\theta=0$. The origin of this singularity is the fact that
2587: we have performed an angular T-duality with a fixed point. Around the fixed
2588: point, as we have pointed out, the original background behaves like flat
2589: space. Thus, the T-dual behavior is recognized as the same type of
2590: singularity that one obtains in T-dualizing flat space in cylindrical
2591: coordinates. Since the original theory was regular near the origin, we expect the T-dual
2592: theory to behave well at this point as well.
2593:
2594: On the other hand, we observe that the region at $\theta= \pi/2$ and $r=0$
2595: where the NS5-branes resides has locally become identical to a $\,\IC^2/\IZ_k$
2596: orbifold. Thus, it is the orbifold singularity of order $k$ that codes the
2597: presence of the NS5-branes in the T-dual.
2598:
2599: We can make the above analysis more precise by performing the
2600: following mental exercise. Cut out from the $6-7$ plane a little disc at the
2601: origin. Topologically, the space transverse to the NS5-branes has become
2602: $\IR^3 \times S^1$, with NS5-branes spread on the $S^1$ (see figure \ref{planeALF}).
2603: \begin{figure}
2604: \centering
2605: \includegraphics[width=0.8\textwidth]{planeALF}
2606: \caption{NS5-branes spread on a topologically trivial circle (drawn on the left).
2607: When we cut out a little disc at the center, the configuration becomes topologically equivalent to NS5-branes spread on a topologically non-trivial circle (drawn on the right).}
2608: \label{planeALF}
2609: \end{figure}
2610: When we neglect the localization of the NS5-branes on the circle, the configuration is T-dual to
2611: an ALF space \cite{Ooguri:1995wj}\cite{Kutasov:1995te}, which in the particular case where the
2612: NS5-branes coincide in $\IR^3$ develops a $\,\IC^2/\IZ_k$ orbifold singularity.
2613: That reasoning is another version of the one above, which uses local
2614: fiberwise T-duality. Localizing the NS5-branes in
2615: the original geometry on the circle is known to be equivalent to taking into
2616: account worldsheet instanton corrections for the ALF space \cite{Gregory:1997te}\cite{Tong:2002rq}\cite{Okuyama:2005gx}, and this
2617: equivalence is conjectured to be valid for the NS5-branes on a topologically
2618: trivial circle as well \cite{Israel:2005fn}.
2619:
2620: We take away several facts from this analysis. Firstly, the $\IZ_k$
2621: orbifold projection acting on the $SU(2)/U(1) \times
2622: SL(2,R)/U(1)$ coset has fixed points and gives rise locally
2623: to a physical $\IC^2/\IZ_k$ singularity that is T-dual to the presence of
2624: $k$ NS5-branes.
2625: And secondly,
2626: in the geometry of the coset conformal field theory, the $\IZ_k$ GSO
2627: projection acts geometrically on the cigar coordinates,
2628: and this in contrast to the fact that the GSO projection in Gepner
2629: models does not orbifold the coordinates
2630: of weighted projective space. We recalled
2631: the above detailed example because we will use
2632: these facts as arguments in section \ref{geometrymirror} and they are particularly
2633: manifest in the above example.
2634:
2635: \section{The consequences of unitarity}\label{unitarity}
2636:
2637: We list in this appendix a number of properties that hold in unitary modules
2638: of the $N=2$ superconformal algebra, irrespective of the value of the central
2639: charge (and in particular also when the central charge is equal or greater
2640: than three). Our conventions for the $N=2$ superconformal algebra coincide
2641: with those of \cite{Polchinski:1998rr, Lerche:1989uy}.
2642: \subsection*{Properties}
2643: The following properties hold -- the proofs in the references that
2644: we give depends only on the unitarity of the
2645: representation spaces as we checked on a case-by-case basis:
2646: \begin{enumerate}
2647: \item All states in the NS-sector with conformal dimension $h$ and R-charge
2648: $Q$ satisfy the inequality $h \ge |Q|/2$ \cite{Lerche:1989uy}.
2649: \item
2650: An NS-sector field of conformal dimension $h$ and R-charge $Q$ is a
2651: chiral primary
2652: if and only if $h=Q/2$ and anti-chiral if and only if $h=-Q/2$
2653: \cite{Lerche:1989uy}.
2654: \item Chiral primary fields satisfy the inequality $h \le c/6$ \cite{Lerche:1989uy}.
2655: \item Ramond sector ground states have R-charges $Q$ in the range
2656: $ -c/6 \le Q \le +c/6$ \cite{Lerche:1989uy}.
2657: \item An NS-sector state $|\phi \rangle$
2658: satisfies the equations
2659: $G^+_{-l-1/2} | \phi \rangle = 0 = G^-_{l+1/2} |\phi \rangle$
2660: if and only if its conformal dimension
2661: $h$ and R-charge $Q$ satisfy the relation $h=(l+1/2) Q-c(l^2+l)/6$ \cite{Vafa:1989xc}.
2662: \item A chiral ring exists \cite{Lerche:1989uy}.
2663: \end{enumerate}
2664: \subsection*{Remarks}
2665: Note that we performed a non-trivial exercise. For instance, for conformal
2666: field theories with central charge $c<3$, it is argued in \cite{Lerche:1989uy}
2667: that there
2668: always exist a chiral primary field in the conformal field theory
2669: with conformal dimension $h=c/6$. That is a consequence of the existence of
2670: the unit operator in the theory (combined with spectral flow).
2671: In other words, there is a normalizable
2672: $SL(2,R)$ invariant ground state in these conformal field theories. That is
2673: not a consequence of unitarity only, and it does {\em not}
2674: hold generically for unitary $N=2$ superconformal field theories. In fact, the situation is
2675: subtle. For instance there are examples of bulk $N=2$
2676: superconformal field theories with central charge $c>3$ for which the unit
2677: operator does not exist when the conformal field theory is
2678: defined on a Riemann surface without boundary,
2679: while it does exist for instance on the boundary of a
2680: disc with particular boundary
2681: conditions (see e.g. \cite{Eguchi:2003ik}\cite{Israel:2004jt}).
2682: It is due to these kind of subtleties that it is useful to have a
2683: (partial) list of properties that we can indiscriminantly use for unitary
2684: $N=2$ theories with any central charge.
2685:
2686: \begin{thebibliography}{99}
2687:
2688: %\cite{Yau:1998yt}\cite{Greene:1997ty}
2689: \bibitem{Yau:1998yt}
2690: S.~T.~Yau,
2691: ``Mirror symmetry I,''
2692: %\href{http://www.slac.stanford.edu/spires/find/hep/www?irn=4308182}{SPIRES entry}
2693: {\it Providence, USA: AMS (1998) 444 p}
2694: %\cite{Greene:1997ty}
2695: \bibitem{Greene:1997ty}
2696: B.~Greene and S.~T.~Yau,
2697: ``Mirror symmetry II,''
2698: %\href{http://www.slac.stanford.edu/spires/find/hep/www?irn=4019032}{SPIRES entry}
2699: {\it Providence, USA: AMS (1997) 844 p. (AMS/IP studies in advanced mathematics. 1)}
2700:
2701: %\cite{Greene:1990ud}
2702: \bibitem{Greene:1990ud}
2703: B.~R.~Greene and M.~R.~Plesser,
2704: %``DUALITY IN CALABI-YAU MODULI SPACE,''
2705: Nucl.\ Phys.\ B {\bf 338}, 15 (1990).
2706: %%CITATION = NUPHA,B338,15;%%
2707:
2708: %\cite{Gepner:1987qi}
2709: \bibitem{Gepner:1987qi}
2710: D.~Gepner,
2711: %``Space-Time Supersymmetry in Compactified String Theory and Superconformal
2712: %Models,''
2713: Nucl.\ Phys.\ B {\bf 296} (1988) 757.
2714: %%CITATION = NUPHA,B296,757;%%
2715:
2716: %\cite{Candelas:1990rm}
2717: \bibitem{Candelas:1990rm}
2718: P.~Candelas, X.~C.~De La Ossa, P.~S.~Green and L.~Parkes,
2719: %``A pair of Calabi-Yau manifolds as an exactly soluble superconformal
2720: %theory,''
2721: Nucl.\ Phys.\ B {\bf 359} (1991) 21.
2722: %%CITATION = NUPHA,B359,21;%%
2723:
2724: %\cite{Berglund:1991pp}
2725: \bibitem{Berglund:1991pp}
2726: P.~Berglund and T.~Hubsch,
2727: %``A generalized construction of mirror manifolds,''
2728: Nucl.\ Phys.\ B {\bf 393}, 377 (1993)
2729: [arXiv:hep-th/9201014].
2730: %%CITATION = NUPHA,B393,377;%%
2731:
2732: \bibitem{Batyrev}
2733: V.~Batyrev,
2734: %"Dual Polyhedra and Mirror Symmetry for Calabi-Yau Hypersurfaces in Toric Varieties,"
2735: arXiv:alg-geom/9310003.
2736:
2737: %\cite{Chiang:1999tz}
2738: \bibitem{Chiang:1999tz}
2739: T.~M.~Chiang, A.~Klemm, S.~T.~Yau and E.~Zaslow,
2740: %``Local mirror symmetry: Calculations and interpretations,''
2741: Adv.\ Theor.\ Math.\ Phys.\ {\bf 3}, 495 (1999)
2742: [arXiv:hep-th/9903053].
2743: %%CITATION = 00203,3,495;%%
2744:
2745: %\cite{Witten:1993yc}
2746: \bibitem{Witten:1993yc}
2747: E.~Witten,
2748: %``Phases of N = 2 theories in two dimensions,''
2749: Nucl.\ Phys.\ B {\bf 403}, 159 (1993)
2750: [arXiv:hep-th/9301042].
2751: %%CITATION = NUPHA,B403,159;%%
2752:
2753: %\cite{Hori:2000kt}\cite{Hori:2002cd}\cite{Hori:2001ax}
2754: \bibitem{Hori:2000kt}
2755: K.~Hori and C.~Vafa,
2756: %``Mirror symmetry,''
2757: arXiv:hep-th/0002222.
2758: %%CITATION = HEP-TH/0002222;%%
2759: %\cite{Hori:2002cd}\cite{Hori:2001ax}
2760: \bibitem{Hori:2002cd}
2761: K.~Hori and A.~Kapustin,
2762: %``Worldsheet descriptions of wrapped NS five-branes,''
2763: JHEP {\bf 0211}, 038 (2002)
2764: [arXiv:hep-th/0203147].
2765: %%CITATION = JHEPA,0211,038;%%
2766: %\cite{Eguchi:2004yi}
2767: \bibitem{Eguchi:2004yi}
2768: T.~Eguchi and Y.~Sugawara,
2769: %``SL(2,R)/U(1) supercoset and elliptic genera of non-compact Calabi-Yau
2770: %manifolds,''
2771: JHEP {\bf 0405}, 014 (2004)
2772: [arXiv:hep-th/0403193].
2773: %%CITATION = JHEPA,0405,014;%%
2774:
2775: %\cite{Giveon:1999px}\cite{Eguchi:2004yi}
2776: \bibitem{Giveon:1999px}
2777: A.~Giveon and D.~Kutasov,
2778: %``Little string theory in a double scaling limit,''
2779: JHEP {\bf 9910} (1999) 034
2780: [arXiv:hep-th/9909110].
2781: %%CITATION = JHEPA,9910,034;%%
2782:
2783: %\cite{Hori:2001ax}
2784: \bibitem{Hori:2001ax}
2785: K.~Hori and A.~Kapustin,
2786: %``Duality of the fermionic 2d black hole and N = 2 Liouville theory as
2787: %mirror symmetry,''
2788: JHEP {\bf 0108}, 045 (2001)
2789: [arXiv:hep-th/0104202].
2790: %%CITATION = JHEPA,0108,045;%%
2791:
2792: %\cite{Tong:2003ik}
2793: \bibitem{Tong:2003ik}
2794: D.~Tong,
2795: %``Mirror mirror on the wall: On two-dimensional black holes and Liouville
2796: %theory,''
2797: JHEP {\bf 0304}, 031 (2003)
2798: [arXiv:hep-th/0303151].
2799: %%CITATION = JHEPA,0304,031;%%
2800:
2801: %\cite{Israel:2004jt}
2802: \bibitem{Israel:2004jt}
2803: D.~Israel, A.~Pakman and J.~Troost,
2804: %``D-branes in N = 2 Liouville theory and its mirror,''
2805: Nucl.\ Phys.\ B {\bf 710}, 529 (2005)
2806: [arXiv:hep-th/0405259].
2807: %%CITATION = NUPHA,B710,529;%%
2808:
2809: %\cite{Greene:1988ut}\cite{Vafa:1989xc}\cite{Intriligator:1990ua}
2810: \bibitem{Greene:1988ut}
2811: B.~R.~Greene, C.~Vafa and N.~P.~Warner,
2812: %``Calabi-Yau Manifolds and Renormalization Group Flows,''
2813: Nucl.\ Phys.\ B {\bf 324} (1989) 371.
2814: %%CITATION = NUPHA,B324,371;%%
2815: %\cite{Vafa:1989xc}\cite{Intriligator:1990ua}
2816:
2817: %\cite{Candelas:1989hd}
2818: \bibitem{Candelas:1989hd}
2819: P.~Candelas, M.~Lynker and R.~Schimmrigk,
2820: %``Calabi-Yau Manifolds in Weighted P(4),''
2821: Nucl.\ Phys.\ B {\bf 341}, 383 (1990).
2822: %%CITATION = NUPHA,B341,383;%%
2823:
2824: \bibitem{Vafa:1989xc}
2825: C.~Vafa,
2826: %``String Vacua and Orbifoldized L-G Models,''
2827: Mod.\ Phys.\ Lett.\ A {\bf 4}, 1169 (1989).
2828: %%CITATION = MPLAE,A4,1169;%%
2829: %\cite{Intriligator:1990ua}
2830: \bibitem{Intriligator:1990ua}
2831: K.~A.~Intriligator and C.~Vafa,
2832: %``LANDAU-GINZBURG ORBIFOLDS,''
2833: Nucl.\ Phys.\ B {\bf 339} (1990) 95.
2834: %%CITATION = NUPHA,B339,95;%%
2835:
2836: %\cite{Eguchi:2000tc}\cite{Murthy:2003es}\cite{Hanany:2002ev}\cite{Eguchi:2004yi}\cite{Israel:2004ir}
2837: \bibitem{Eguchi:2000tc}
2838: T.~Eguchi and Y.~Sugawara,
2839: %``Modular invariance in superstring on Calabi-Yau n-fold with A-D-E
2840: %singularity,''
2841: Nucl.\ Phys.\ B {\bf 577}, 3 (2000)
2842: [arXiv:hep-th/0002100].
2843: %%CITATION = NUPHA,B577,3;%%
2844:
2845: %\cite{Hanany:2002ev}
2846: \bibitem{Hanany:2002ev}
2847: A.~Hanany, N.~Prezas and J.~Troost,
2848: %``The partition function of the two-dimensional black hole conformal field
2849: %theory,''
2850: JHEP {\bf 0204}, 014 (2002)
2851: [arXiv:hep-th/0202129].
2852: %%CITATION = JHEPA,0204,014;%%
2853: %\cite{Murthy:2003es}
2854: \bibitem{Murthy:2003es}
2855: S.~Murthy,
2856: %``Notes on non-critical superstrings in various dimensions,''
2857: JHEP {\bf 0311}, 056 (2003)
2858: [arXiv:hep-th/0305197].
2859: %%CITATION = JHEPA,0311,056;%%
2860:
2861: %\cite{Israel:2004ir}
2862: \bibitem{Israel:2004ir}
2863: D.~Israel, C.~Kounnas, A.~Pakman and J.~Troost,
2864: %``The partition function of the supersymmetric two-dimensional black hole
2865: %and little string theory,''
2866: JHEP {\bf 0406}, 033 (2004)
2867: [arXiv:hep-th/0403237].
2868: %%CI
2869:
2870: %\cite{Ooguri:1995wj}
2871: \bibitem{Ooguri:1995wj}
2872: H.~Ooguri and C.~Vafa,
2873: %``Two-Dimensional Black Hole and Singularities of CY Manifolds,''
2874: Nucl.\ Phys.\ B {\bf 463}, 55 (1996)
2875: [arXiv:hep-th/9511164].
2876: %%CITATION = NUPHA,B463,55;%%
2877:
2878: %\cite{Maldacena:2000kv}
2879: \bibitem{Maldacena:2000kv}
2880: J.~M.~Maldacena, H.~Ooguri and J.~Son,
2881: %``Strings in AdS(3) and the SL(2,R) WZW model. II: Euclidean black hole,''
2882: J.\ Math.\ Phys.\ {\bf 42}, 2961 (2001)
2883: [arXiv:hep-th/0005183].
2884: %%CITATION = JMAPA,42,2961;%%
2885:
2886: %\cite{Gepner:1986hr}
2887: \bibitem{Gepner:1986hr}
2888: D.~Gepner and Z.~a.~Qiu,
2889: %``Modular Invariant Partition Functions for Parafermionic Field Theories,''
2890: Nucl.\ Phys.\ B {\bf 285} (1987) 423.
2891: %%CITATION = NUPHA,B285,423;%%
2892:
2893: %\cite{Greene:1996cy}
2894: \bibitem{Greene:1996cy}
2895: B.~R.~Greene,
2896: %``String theory on Calabi-Yau manifolds,''
2897: arXiv:hep-th/9702155.
2898: %%CITATION = HEP-TH/9702155;%%
2899:
2900: %\cite{Maldacena:2001ky}
2901: \bibitem{Maldacena:2001ky}
2902: J.~M.~Maldacena, G.~W.~Moore and N.~Seiberg,
2903: %``Geometrical interpretation of D-branes in gauged WZW models,''
2904: JHEP {\bf 0107} (2001) 046
2905: [arXiv:hep-th/0105038].
2906: %%CITATION = JHEPA,0107,046;%%
2907:
2908: %\cite{Dixon:1989cg}
2909: \bibitem{Dixon:1989cg}
2910: L.~J.~Dixon, M.~E.~Peskin and J.~D.~Lykken,
2911: %``N=2 Superconformal Symmetry and SO(2,1) Current Algebra,''
2912: Nucl.\ Phys.\ B {\bf 325}, 329 (1989).
2913: %%CITATION = NUPHA,B325,329;%%
2914:
2915: %\cite{Gukov:1999ya}
2916: \bibitem{Gukov:1999ya}
2917: S.~Gukov, C.~Vafa and E.~Witten,
2918: %``CFT's from Calabi-Yau four-folds,''
2919: Nucl.\ Phys.\ B {\bf 584} (2000) 69
2920: [Erratum-ibid.\ B {\bf 608} (2001) 477]
2921: [arXiv:hep-th/9906070].
2922: %%CITATION = NUPHA,B584,69;%%
2923: %\cite{Shapere:1999xr}
2924: \bibitem{Shapere:1999xr}
2925: A.~D.~Shapere and C.~Vafa,
2926: %``BPS structure of Argyres-Douglas superconformal theories,''
2927: arXiv:hep-th/9910182.
2928: %%CITATION = HEP-TH/9910182;%%
2929:
2930: %\cite{Martinec:1988zu}\cite{Vafa:1988uu}\cite{Witten:1993jg}
2931: \bibitem{Martinec:1988zu}
2932: E.~J.~Martinec,
2933: %``Algebraic Geometry and Effective Lagrangians,''
2934: Phys.\ Lett.\ B {\bf 217} (1989) 431.
2935: %%CITATION = PHLTA,B217,431;%%
2936: %\cite{Vafa:1988uu}
2937: \bibitem{Vafa:1988uu}
2938: C.~Vafa and N.~P.~Warner,
2939: %``Catastrophes and the Classification of Conformal Theories,''
2940: Phys.\ Lett.\ B {\bf 218} (1989) 51.
2941: %%CITATION = PHLTA,B218,51;%%
2942:
2943: %\cite{Witten:1993jg}
2944: \bibitem{Witten:1993jg}
2945: E.~Witten,
2946: %``On the Landau-Ginzburg description of N=2 minimal models,''
2947: Int.\ J.\ Mod.\ Phys.\ A {\bf 9}, 4783 (1994)
2948: [arXiv:hep-th/9304026].
2949: %%CITATION = IMPAE,A9,4783;%%
2950:
2951: %\cite{Rocek:1991ps}
2952: \bibitem{Rocek:1991ps}
2953: M.~Rocek and E.~P.~Verlinde,
2954: %``Duality, quotients, and currents,''
2955: Nucl.\ Phys.\ B {\bf 373}, 630 (1992)
2956: [arXiv:hep-th/9110053].
2957: %%CITATION = NUPHA,B373,630;%%
2958:
2959: %\cite{Lerche:1989uy}
2960: \bibitem{Lerche:1989uy}
2961: W.~Lerche, C.~Vafa and N.~P.~Warner,
2962: %``Chiral Rings in N=2 Superconformal Theories,''
2963: Nucl.\ Phys.\ B {\bf 324}, 427 (1989).
2964: %%CITATION = NUPHA,B324,427;%%
2965:
2966: %\cite{Ghoshal:1993qt}\cite{Hanany:1994fi}
2967: \bibitem{Ghoshal:1993qt}
2968: D.~Ghoshal and S.~Mukhi,
2969: %``Topological Landau-Ginzburg model of two-dimensional string theory,''
2970: Nucl.\ Phys.\ B {\bf 425} (1994) 173
2971: [arXiv:hep-th/9312189].
2972: %%CITATION = NUPHA,B425,173;%%
2973: %\cite{Hanany:1994fi}
2974: \bibitem{Hanany:1994fi}
2975: A.~Hanany, Y.~Oz and M.~Ronen Plesser,
2976: %``Topological Landau-Ginzburg Formulation And Integrable Structure Of 2-D
2977: %String Theory,''
2978: Nucl.\ Phys.\ B {\bf 425} (1994) 150
2979: [arXiv:hep-th/9401030].
2980: %%CITATION = NUPHA,B425,150;%%
2981:
2982: %\cite{Ghoshal:1995wm}
2983: \bibitem{Ghoshal:1995wm}
2984: D.~Ghoshal and C.~Vafa,
2985: %``C = 1 String As The Topological Theory Of The Conifold,''
2986: Nucl.\ Phys.\ B {\bf 453}, 121 (1995)
2987: [arXiv:hep-th/9506122].
2988: %%CITATION = NUPHA,B453,121;%%
2989:
2990: %\cite{Aspinwall:1994rg}
2991: \bibitem{Aspinwall:1994rg}
2992: P.~S.~Aspinwall and D.~R.~Morrison,
2993: %``String theory on K3 surfaces,''
2994: arXiv:hep-th/9404151.
2995: %%CITATION = HEP-TH/9404151;%%
2996:
2997: %\cite{Lerche:2000uy}
2998: \bibitem{Lerche:2000uy}
2999: W.~Lerche,
3000: %``On a boundary CFT description of nonperturbative N = 2 Yang-Mills
3001: %theory,''
3002: arXiv:hep-th/0006100.
3003: %%CITATION = HEP-TH/0006100;%%
3004:
3005: %\cite{Eguchi:2004ik}
3006: \bibitem{Eguchi:2004ik}
3007: T.~Eguchi and Y.~Sugawara,
3008: %``Conifold type singularities, N = 2 Liouville and SL(2,R)/U(1) theories,''
3009: JHEP {\bf 0501} (2005) 027
3010: [arXiv:hep-th/0411041].
3011: %%CITATION = JHEPA,0501,027;%%
3012:
3013: %\cite{Kutasov:1995te}
3014: \bibitem{Kutasov:1995te}
3015: D.~Kutasov,
3016: %``Orbifolds and Solitons,''
3017: Phys.\ Lett.\ B {\bf 383}, 48 (1996)
3018: [arXiv:hep-th/9512145].
3019: %%CITATION = PHLTA,B383,48;%%
3020:
3021: %\cite{Aspinwall:1994ev}
3022: \bibitem{Aspinwall:1994ev}
3023: P.~S.~Aspinwall,
3024: %``Resolution of orbifold singularities in string theory,''
3025: arXiv:hep-th/9403123.
3026: %%CITATION = HEP-TH/9403123;%%
3027:
3028: %\cite{Lust:2006zh}
3029: \bibitem{lustetal}
3030: D.~Lust, S.~Reffert, E.~Scheidegger and S.~Stieberger,
3031: %``Resolved toroidal orbifolds and their orientifolds,''
3032: arXiv:hep-th/0609014.
3033: %%CITATION = HEP-TH/0609014;%%
3034:
3035: %\cite{Hanany:2005ve}\cite{Franco:2005rj}
3036: \bibitem{Hanany:2005ve}
3037: A.~Hanany and K.~D.~Kennaway,
3038: %``Dimer models and toric diagrams,''
3039: arXiv:hep-th/0503149.
3040: %%CITATION = HEP-TH/0503149;%%
3041: %\cite{Franco:2005rj}
3042: \bibitem{Franco:2005rj}
3043: S.~Franco, A.~Hanany, K.~D.~Kennaway, D.~Vegh and B.~Wecht,
3044: %``Brane dimers and quiver gauge theories,''
3045: JHEP {\bf 0601} (2006) 096
3046: [arXiv:hep-th/0504110].
3047: %%CITATION = JHEPA,0601,096;%%
3048:
3049: %\cite{Feng:2000mi}\cite{Beasley:2001zp}\cite{Cachazo:2001sg}
3050: \bibitem{Feng:2000mi}
3051: B.~Feng, A.~Hanany and Y.~H.~He,
3052: %``D-brane gauge theories from toric singularities and toric duality,''
3053: Nucl.\ Phys.\ B {\bf 595} (2001) 165
3054: [arXiv:hep-th/0003085].
3055: %%CITATION = NUPHA,B595,165;%%
3056: %\cite{Beasley:2001zp}\cite{Cachazo:2001sg}
3057: \bibitem{Beasley:2001zp}
3058: C.~E.~Beasley and M.~R.~Plesser,
3059: %``Toric duality is Seiberg duality,''
3060: JHEP {\bf 0112} (2001) 001
3061: [arXiv:hep-th/0109053].
3062: %%CITATION = JHEPA,0112,001;%%
3063: %\cite{Cachazo:2001sg}
3064: \bibitem{Cachazo:2001sg}
3065: F.~Cachazo, B.~Fiol, K.~A.~Intriligator, S.~Katz and C.~Vafa,
3066: %``A geometric unification of dualities,''
3067: Nucl.\ Phys.\ B {\bf 628} (2002) 3
3068: [arXiv:hep-th/0110028].
3069: %%CITATION = NUPHA,B628,3;%%
3070:
3071: %\cite{Eguchi:2003ik}
3072: \bibitem{Eguchi:2003ik}
3073: T.~Eguchi and Y.~Sugawara,
3074: %``Modular bootstrap for boundary N = 2 Liouville theory,''
3075: JHEP {\bf 0401}, 025 (2004)
3076: [arXiv:hep-th/0311141].
3077: %%CITATION = JHEPA,0401,025;%%
3078:
3079: %\cite{Ashok:2005py}
3080: \bibitem{Ashok:2005py}
3081: S.~K.~Ashok, S.~Murthy and J.~Troost,
3082: %``D-branes in non-critical superstrings and minimal super Yang-Mills in
3083: %various dimensions,''
3084: Nucl.\ Phys.\ B {\bf 749}, 172 (2006)
3085: [arXiv:hep-th/0504079].
3086: %%CITATION = NUPHA,B749,172;%%
3087:
3088: %\cite{Fotopoulos:2005cn}
3089: \bibitem{Fotopoulos:2005cn}
3090: A.~Fotopoulos, V.~Niarchos and N.~Prezas,
3091: %``D-branes and SQCD in non-critical superstring theory,''
3092: JHEP {\bf 0510}, 081 (2005)
3093: [arXiv:hep-th/0504010].
3094: %%CITATION = JHEPA,0510,081;%%
3095:
3096: %\cite{Murthy:2006xt}
3097: \bibitem{Murthy:2006xt}
3098: S.~Murthy and J.~Troost,
3099: %``D-branes in non-critical superstrings and duality in N = 1 gauge theories
3100: %with flavor,''
3101: JHEP {\bf 0610}, 019 (2006)
3102: [arXiv:hep-th/0606203].
3103: %%CITATION = JHEPA,0610,019;%%
3104:
3105: %\cite{Ashok:2007sf}
3106: \bibitem{Ashok:2007sf}
3107: S.~K.~Ashok, S.~Murthy and J.~Troost,
3108: %``D-branes in unoriented non-critical strings and duality in SO(N) and Sp(N)
3109: %gauge theories,''
3110: JHEP {\bf 0706}, 047 (2007)
3111: [arXiv:hep-th/0703148].
3112: %%CITATION = JHEPA,0706,047;%%
3113:
3114:
3115: %\cite{Hosomichi:2004ph}
3116: \bibitem{Hosomichi:2004ph}
3117: K.~Hosomichi,
3118: %``N = 2 Liouville theory with boundary,''
3119: JHEP {\bf 0612} (2006) 061
3120: [arXiv:hep-th/0408172].
3121: %%CITATION = JHEPA,0612,061;%%
3122:
3123: %\cite{Sfetsos:1998xd}
3124: \bibitem{Sfetsos:1998xd}
3125: K.~Sfetsos,
3126: %``Branes for Higgs phases and exact conformal field theories,''
3127: JHEP {\bf 9901} (1999) 015
3128: [arXiv:hep-th/9811167].
3129: %%CITATION = JHEPA,9901,015;%%
3130:
3131: %\cite{Israel:2005fn}
3132: \bibitem{Israel:2005fn}
3133: D.~Israel, A.~Pakman and J.~Troost,
3134: %``D-branes in little string theory,''
3135: Nucl.\ Phys.\ B {\bf 722}, 3 (2005)
3136: [arXiv:hep-th/0502073].
3137: %%CITATION = NUPHA,B722,3;%%
3138:
3139: %\cite{Gregory:1997te}\cite{Tong:2002rq}\cite{Okuyama:2005gx}
3140: \bibitem{Gregory:1997te}
3141: R.~Gregory, J.~A.~Harvey and G.~W.~Moore,
3142: %``Unwinding strings and T-duality of Kaluza-Klein and H-monopoles,''
3143: Adv.\ Theor.\ Math.\ Phys.\ {\bf 1}, 283 (1997)
3144: [arXiv:hep-th/9708086].
3145: %%CITATION = 00203,1,283;%%
3146: %\cite{Tong:2002rq}\cite{Okuyama:2005gx}
3147: \bibitem{Tong:2002rq}
3148: D.~Tong,
3149: %``NS5-branes, T-duality and worldsheet instantons,''
3150: JHEP {\bf 0207}, 013 (2002)
3151: [arXiv:hep-th/0204186].
3152: %%CITATION = JHEPA,0207,013;%%
3153: %\cite{Okuyama:2005gx}
3154: \bibitem{Okuyama:2005gx}
3155: K.~Okuyama,
3156: %``Linear sigma models of H and KK monopoles,''
3157: JHEP {\bf 0508} (2005) 089
3158: [arXiv:hep-th/0508097].
3159: %%CITATION = JHEPA,0508,089;%%
3160:
3161: %\cite{Polchinski:1998rr}
3162: \bibitem{Polchinski:1998rr}
3163: J.~Polchinski,
3164: ``String theory. Vol. 2: Superstring theory and beyond,''
3165: %\href{http://www.slac.stanford.edu/spires/find/hep/www?irn=4634802}{SPIRES entry}
3166: {\it Cambridge, UK: Univ. Pr. (1998) 531 p}
3167:
3168: \end{thebibliography}
3169: \end{document}
3170:
3171: