hep-th0404184/loc.tex
1: \documentclass[paper, 12pt, letterpaper]{JHEP} 
2: \usepackage{amsmath}
3: \usepackage{amssymb}
4: 
5: %1. Math symbols:
6: 
7: %1.1 AMS fonts:
8: 
9: \def\Bbb{\mathbb} \def\C{{\Bbb C}} \def\R{{\Bbb R}} \def\Z{{\Bbb Z}}
10: \def\H{{\Bbb H}} \def\Q{{\Bbb Q}} \def\P{{\Bbb P}} \def\ii{{\Bbb
11:     i}}\def\j{{\Bbb j}}\def\k{{\Bbb k}} \def\I{{\Bbb I}}\def\J{{\Bbb
12:     J}}\def\K{{\Bbb K}} \def\boldphi{\mbox{\boldmath $\phi$}}
13: \def\boldpsi{\mbox{\boldmath $\psi$}} \def\boldzeta{\mbox{\boldmath
14:     $\zeta$}}
15: 
16: 
17: 
18: %1.2 double line:
19: 
20: \font\mybbb=msbm10 at 8pt \font\mybb=msbm10 at 12pt
21: \def\bbb#1{\hbox{\mybbb#1}} \def\bb#1{\hbox{\mybb#1}}
22: 
23: \def\pRe{\bbb{R}} \def\Re {\bb{Re}} \def\S {\bb{S}} \def\T {\bb{T}}
24: %\def\P{\bb{P}}
25: %\def\Z {\bb{Z}}
26: 
27: 
28: 
29: %1.3 operatorname:
30: 
31: \def\Hom{\operatorname{Hom}} \def\Tors{\operatorname{Tors}}
32: \def\Ker{\operatorname{Ker}} \def\Spec{\operatorname{Spec}}
33: \def\Area{\operatorname{Area}} \def\Vol{{\rm Vol}} \def\vol {{\rm
34:     vol}} \def\ad{\operatorname{ad}} \def\tr{\operatorname{tr}}\def\Tr{\operatorname{Tr}}
35: \def\str{\operatorname{str}}
36: \def\Str{\operatorname{Str}}
37: \def\Pic{\operatorname{Pic}} \def\disc{\operatorname{disc}}
38: \def\cpl{\operatorname{cpl}} \def\Img{\operatorname{Im}}
39: \def\Rea{\operatorname{Re}} \def\Gr{\operatorname{Gr}}
40: \def\SO{\operatorname{SO}} \def\Sl{\operatorname{SL}}
41: \def\GO{\operatorname{O{}}} \def\SU{{\rm SU}}
42: \def\GU{\operatorname{U{}}} \def\Sp{\operatorname{Sp}}
43: \def\Spin{\operatorname{Spin}} \def\rank{\operatorname{rank}}
44: \def\Aff{\operatorname{Aff}} \def\diag{\operatorname{diag}}
45: \def\su{\rm{su(2)}}
46: 
47: 
48: %1.4 roman:
49: 
50: \def\ker{{\rm ker}} \def\coker{{\rm coker}} \def\rank{{\rm rank}}
51: \def\im{{\rm im\,}} \def\Im{{\rm Im\,}} \def\re{{\rm re\,}}
52: \def\Re{{\rm Re\,}} \def\tr{{\rm tr\, }} \def\dim{{\rm dim}}
53: \def\codim{{\rm codim}} \def\Tr{{\rm Tr\, }} \def\d{{\rm d }}
54: \def\Card{{\rm Card}} \def\li{{\rm linearly independent}} \def\ld{{\rm
55:     linearly dependent}} \def\deg{{\rm deg}} \def\rk{{\rm rk}} \def\det{{\rm det}}
56: \def\Div{{\rm Div}} \def\supp{{\rm supp}} \def\Gr{{\rm Gr}}
57: \def\End{{\rm End}} \def\Aut{{\rm Aut}}
58: \def\Spec{{\rm Spec}}
59: 
60: %1.5. script:
61: 
62: \def\cR{{\Scr R}} \def\cM{{\Scr M}} \def\cA{{\Scr A}} \def\cB{{\Scr
63:     B}} \def\cK{{\Scr K}} \def\cD{{\Scr D}} \def\cH{{\Scr H}}
64: \def\cT{{\Scr T}} \def\cL{{\Scr L}} \def\cF{{\Scr F}}
65: 
66: %1.6 gothic:
67: 
68: \font\gothics=ygoth at 10pt \font\gothicl=ygoth at 12pt
69: \def\gg#1{\hbox{\gothics#1}} \def\gG{\hbox{\gothicl G}} \def
70: \g{\hbox{\gothics g}}
71: 
72: %1.7 misc:
73: 
74: \def\id{\protect{{1 \kern-.28em {\rm l}}}}
75: 
76: 
77: %3. New commands
78: 
79: \newcommand{\be}{\begin{equation}} \newcommand{\ee}{\end{equation}}
80: \newcommand{\bea}{\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}}
81: \newcommand{\beann}{\begin{eqnarray*}}
82:   \newcommand{\eeann}{\end{eqnarray*}}
83: \newcommand{\bfig}{\begin{figure}} \newcommand{\efig}{\end{figure}}
84: \newcommand{\nn}{\nonumber}
85: \newcommand{\ba}{\begin{array}}\newcommand{\ea}{\end{array}}
86: 
87: 
88: %4. New theorems and their commands
89: 
90: \newtheorem{Proposition}{Proposition}[section]
91: \newtheorem{Definition}{Definition}[section]
92: \newtheorem{Theorem}{Theorem}[section]
93: \newtheorem{Lemma}{Lemma}[section]
94: \newtheorem{Corrolary}{Corrolary}[section]
95: 
96: \newcommand{\bp}{\begin{Proposition}}
97:   \newcommand{\ep}{\end{Proposition}}
98: \newcommand{\bt}{\begin{Theorem}} \newcommand{\et}{\end{Theorem}}
99: \newcommand{\bl}{\begin{Lemma}} \newcommand{\el}{\end{Lemma}}
100: \newcommand{\bc}{\begin{Corrolary}} \newcommand{\ec}{\end{Corrolary}}
101: 
102: 
103: %5. Karl
104: 
105: \def\omegat{\tilde{\omega}} \def\ot{\tilde{\omega}} \def\om{\omega}
106: \def\Qt{\tilde{Q}} \def\Vt{\tilde{V}} \def\Rt{\tilde{R}} \def\ep{\eps}
107: \def\CC{{\cal C}} \def\hn{{ \hat N}} \def\hq{{\hat Q}} \def\hp{{\hat
108:     \Phi}} \def\hs{{ \hat S}} \def\ha{{\hat A}} \def\cs{{\cal S}}
109: 
110: \author{
111: Manfred Herbst${}^a$, Calin-Iuliu Lazaroiu${}^b$ \\}
112: \author{~\\
113:      ${}^a$Department of Physics, CERN\\
114:      Theory Division\\
115:      CH-1211 Geneva 23\\
116:      Switzerland\\
117:      Manfred.Herbst@cern.ch\\{~}\\
118:      ${}^b$5243 Chamberlin Hall\\
119:      University of Wisconsin at Madison\\
120:      1150 University Ave\\
121:      Madison, Wisconsin 53706, USA\\
122:      calin@physics.wisc.edu\\
123:     }
124: 
125: 
126: \title{Localization and traces in open--closed topological Landau--Ginzburg models}
127: 
128: 
129: \abstract{We reconsider the issue of localization in open-closed B-twisted
130: Landau-Ginzburg models with arbitrary Calabi-Yau target. Through careful
131: analsysis of zero-mode reduction, we show that the closed model allows for a
132: one-parameter family of localization pictures, which generalize the standard
133: residue representation. The parameter $\lambda$ which indexes these pictures
134: measures the area of worldsheets with $S^2$ topology, with the residue
135: representation obtained in the limit of small area. In the boundary sector, we
136: find a double family of such pictures, depending on parameters $\lambda$ and
137: $\mu$ which measure the area and boundary length of worldsheets with disk
138: topology. We show that setting $\mu=0$ and varying $\lambda$ interpolates
139: between the localization picture of the B-model with a noncompact target space
140: and a certain residue representation proposed recently. This gives a complete
141: derivation of the boundary residue formula, starting from the explicit construction of the
142: boundary coupling. We also show that the various localization pictures are
143: related by a semigroup of homotopy equivalences.}
144: 
145: 
146: \preprint{MAD-TH-04-3\\  
147:           CERN--PH--TH/2004-070}
148: 
149: 
150: \begin{document}
151: 
152: 
153: \tableofcontents
154: 
155: \pagebreak
156: 
157: \vskip .6in
158: 
159: \section{Introduction}
160: \label{intro}
161: 
162: 
163: Closed topological Landau-Ginzburg models \cite{Vafa_LG} have been a useful
164: testing ground for string theory. They make direct contact with
165: the topological sector of rational conformal field theories 
166: through the Landau-Ginzburg approach to minimal models, and arise as phases of
167: $N=2$ string compactifications \cite{Witten_phases}.
168: Moreover, they give explicit realizations of the 
169: WDVV equations and examples of Frobenius manifolds, and have interesting relations with 
170: singularity theory.
171: 
172: 
173: In a similar vein, one can expect to learn important lessons about open
174: string theory by studying topological Landau-Ginzburg models in the presence
175: of D-branes (see \cite{us, fractional} for some recent results in this direction). 
176: While basic D-brane constructions were considered by many authors
177: (see \cite{Hori} and references therein), a systematic study has been
178: hampered by the lack of a reasonably general description of the boundary
179: coupling (the "Warner problem" \cite{Warner}).
180: 
181: Progress in removing this obstacle was made recently in  \cite{Kontsevich, Kap1,  Lerche,
182:   Kap2, Kap3} (see also \cite{Orlov}). These papers proposed a solution of the Warner problem for
183:   B-twisted Landau-Ginzburg models with target space $\C^n$ and for particular 
184: families
185:   of D-branes described by superbundles whose rank is constrained to be a
186:   power of two. In a slightly modified form, this 
187:   solution was generalized in \cite{coupling} by removing unnecessary 
188:   assumptions, thus giving the general form of the
189:   relevant boundary coupling. 
190: 
191: 
192: As in the closed string case, it is natural to translate the
193: physical data of open Landau-Ginzburg models into the language of 
194: singularity theory. For the closed string sector, a crucial step in this regard is
195: the localization formula of \cite{Vafa_LG}, which relates topological field
196: theory correlators to residues (see, for
197: example, \cite{Griffiths}). An open string version of this formula was  proposed in \cite{Kap2}, 
198: though a complete derivation based on the microscopic boundary coupling was not given. 
199: 
200: Given the boundary coupling constructed in \cite{coupling}, we shall re-consider this issue in the 
201: more general set-up of \cite{Labastida}, and give a complete derivation of
202: this localization formula. 
203: Another purpose of this paper is to extend the open and closed localization
204: formulae in a manner which reflects the basic intuition 
205: \cite{coupling} that the B-branes of Landau-Ginzburg models are the
206: result of tachyon condensation between the elementary branes of the B-type
207: sigma model, with tachyon condensation driven by the Landau-Ginzburg superpotential. 
208: As we shall show somewhere else, this allows one to make contact with the
209: string field theory approach advocated in a different context in 
210: \cite{Witten_CS, CIL2, CIL3, CIL4, CIL5, CIL6, CIL7, CIL8, CIL9, CIL10, Diaconescu}.
211: 
212: Perhaps surprisingly, this is non-trivial to
213: achieve already for the closed string sector. Indeed, the usual on-shell descriptions
214: of the space of bulk  observables differ markedly
215: between the two models. In the B-twisted sigma model, this space is described
216: as the ${\bar \partial}$-cohomology of the algebra of $(0,p)$-forms on the
217: target space, valued in holomorphic polyvector fields. In the Landau-Ginzburg
218: case, the space of on-shell bulk observables can be identified with the Jacobi ring
219: of the Landau-Ginzburg superpotential. As we shall see below, these two
220: descriptions are related in a subtle manner, namely by a one-parameter
221: family of "localization pictures" which interpolates between them. The
222: existence of such a family will be established by refining 
223: the localization argument of \cite{Vafa_LG}. 
224: The different pictures are indexed by a parameter $\lambda$, 
225: which roughly measures the area of 
226: worldsheets with $S^2$ topology.
227: The B-model description of the algebra of observables
228: arises in the limit when the worldsheet is collapsed to a point, while the 
229: Jacobi realization is recovered for very large areas. More
230: precisely, one can construct an off-shell model for each localization 
231: picture, with a reduced BRST operator whose cohomology reproduces the space
232: of on-shell bulk observables. The various pictures 
233: are related by a "homotopy flow", i.e. a one-parameter semigroup of operators 
234: homotopic to the identity. 
235: This flow induces the trivial action on BRST cohomology, thus identifying 
236: on-shell data between different pictures. 
237: 
238: Extending this construction to the boundary sector, we find a similar
239: description. Namely, we shall 
240: construct a family of localization pictures indexed by {\em two}
241: parameters $\lambda$ and $\mu$, which -- when real -- measure the area of a worldsheet with disk
242: topology and the length of its boundary. 
243: Taking $\mu=0$, the standard realization of observables in the open B-model
244: arises  for $\lambda\rightarrow 0$, while the LG description and residue formula 
245: of \cite{Kap2} are obtained in the opposite limit $\lambda\rightarrow +\infty$. The parameter
246: $\mu$ plays an auxiliary role, related to a certain boundary term which was
247: not included in \cite{coupling} since it is not essential for the
248: topological model. Contrary to previous proposals, we show that this
249: parameter can be safely set to zero, without affecting the localization data. 
250: Physically, localization on the critical set of $W$ is controlled by the 
251: bulk parameter $\lambda$, and is completely independent of the choice of
252: $\mu$. In fact, one must set $\mu=0$ in order to recover
253: \footnote{The authors of \cite{Kap2} propose a different limit, namely
254: $\mu\rightarrow +\infty$ with $\lambda=0$. It does not seem possible to achieve
255: localization on the critical set of $W$ in this limit.} the residue representation of 
256: \cite{Kap2}.
257: 
258: 
259: The paper is organized as follows. In Section 2, we review the bulk Lagrangian
260: and some of its basic properties, following \cite{Labastida}. In Section 3, we
261: discuss localization in the bulk sector. Through careful 
262: analysis, we show that one can localize on the
263: zero modes of the {\em sigma model} action, namely that part of the bulk action
264: which is independent of the Landau-Ginzburg superpotential $W$. This gives the
265: one-parameter family of localization pictures. We also give 
266: the geometric realization of these pictures, and the homotopy flow connecting them. 
267: In Section 5, we recall the boundary coupling given in  \cite{coupling} and
268: adapt it by adding a supplementary term
269: also suggested in \cite{Kap1, Lerche, Kap2} for a special case. This assures that
270: the coupling preserves a full copy of the $N=2$ topological algebra, when the
271: model is considered on a flat strip. Since the second generator of this
272: algebra is gauged when considering the model on a curved Riemann surface, this
273: condition does not play a fundamental role for unintegrated amplitudes, 
274: but the modified coupling is useful for comparison with \cite{Kap2}.
275: Section 5 constructs the boundary observables and correlators, and explains
276: how our model for the boundary  BRST operator arises in this approach. In Section 6, we
277: discuss localization in the boundary sector. As for the bulk, we proceed by
278: reducing to zero modes of the sigma model action. This gives a family 
279: of representations depending on two parameters $\lambda$ and $\mu$.
280: The second of these weights the contribution arising from the supplementary
281: boundary term. After describing boundary homotopy flows and
282: the associated geometric realization, we construct an off-shell representative
283: for the bulk-boundary map of \cite{CIL1} (see also \cite{Moore_Segal, Moore}), 
284: and use it to recover (an extension
285: of) the localization formula proposed in \cite{Kap2}, by taking the limit
286: $\lambda\rightarrow +\infty$ with $\mu=0$\footnote{In this limit, the
287:   supplementary term introduced in Section 5 does not contribute, so 
288: one can use the simplified boundary coupling of \cite{coupling}.}. 
289: Section 7 presents our conclusions. The Appendix gives the boundary 
290: conditions for the general D-brane coupling. 
291: 
292: 
293: 
294: \section{The bulk action}
295: \label{bulk}
296: 
297: The general formulation of closed B-type topological Landau-Ginzburg models
298: was given in \cite{Labastida}, extending the work of
299: \cite{Vafa_LG}. We take the target space to be a Calabi-Yau manifold $X$,
300: with the Landau-Ginzburg potential a holomorphic function $W\in H^0({\cal
301: O}_X)$.  Since any holomorphic function on a compact complex manifold is
302: constant, we shall assume that $X$ is non-compact.  
303: In the on-shell formulation, 
304: the Grassmann even worldsheet fields are 
305: the components $\phi^i$, $\phi^{\bar i}$ of the map $\phi:\Sigma \rightarrow
306: X$, while the G-odd fields are sections $\eta$, $\theta$ and $\rho$ of the
307: bundles $\phi^*({\bar T}X)$, $\phi^*(T^*X)$ and $\phi^*(TX)\otimes {\cal
308: T}^*\Sigma$ over the worldsheet $\Sigma$. Here ${\cal T}^*\Sigma$ is the
309: complexified cotangent bundle to $\Sigma$, while $TX$ and ${\bar T}X$ are the
310: holomorphic and antiholomorphic components of the complexified tangent bundle
311: ${\cal T}X$ to $X$. This agrees with the on-shell field content of 
312: the B-twisted sigma model \cite{Witten_mirror}.
313: 
314: 
315: \subsection{Action and BRST transformations}
316: 
317: To write the bulk action, we introduce new fields $\chi, {\bar
318:   \chi} \in \Gamma(\Sigma, \phi^*({\bar T}X))$ by the relations:
319: \bea
320: \eta^{\bar i}&=&\chi^{\bar i}+{\bar \chi}^{\bar i}\\
321: \theta_i&=&G_{i{\bar j}}(\chi^{\bar j}-{\bar \chi}^{\bar j})~~.
322: \eea
323: We shall also use the quantity $\theta^{\bar i}=G^{{\bar i}j}\theta_j$.
324: 
325: As in \cite{Labastida}, it is convenient to use an off-shell
326: realization of the BRST symmetry. For
327: this, consider an auxiliary G-even field ${\tilde F}$ transforming as a section of
328: $\phi^*({\cal T}X)$. Then the BRST transformations are:
329: \bea
330: \label{BRST}
331: \delta \phi^i=0~~&,&~~\delta \phi^{\bar i}=\chi^{\bar i}+{\bar \chi}^{\bar
332:   i}=\eta^{\bar i}\nn\\
333: \delta \chi^{\bar i}={\tilde F}^{\bar i}-\Gamma^{\bar
334:   i}_{{\bar j}{\bar k}}{\bar \chi}^{\bar j}\chi^{\bar k}~~&,&~~\delta {\bar \chi}^{\bar
335:   i}=-{\tilde F}^{\bar i}+\Gamma^{\bar i}_{{\bar j}{\bar
336:   k}}{\bar \chi}^{\bar j}\chi^{\bar k}~~\nn\\
337: \delta \rho^i_\alpha&=&2\partial_\alpha \phi^i~~\\
338: \delta {\tilde F}^i=i\varepsilon^{\alpha\beta} \left[D_\alpha
339:   \rho^i_\beta+\frac{1}{4}R^i_{j{\bar l}k}(\chi^{\bar l}+{\bar \chi}^{\bar l})
340: \rho^j_\alpha\rho^k_\beta\right]~~&,&~~\delta
341: {\tilde F}^{\bar i}=\Gamma^{\bar i}_{{\bar j}{\bar k}}{\tilde F}^{\bar j}(\chi^{\bar k}+{\bar \chi}^{\bar k})~~.\nn
342: \eea
343: These transformations are independent of
344:   $W$. Moreover, the transformations of $\phi$, $\eta$ and $\rho$ 
345: do not involve the auxiliary fields. In particular, we have  $\delta \eta^{\bar i}=0$.
346: These observations will be used in Section 4. 
347: 
348: Let us pick a Riemannian metric $g$ on the  worldsheet. The bulk action
349: of \cite{Labastida} takes the form:
350: \be
351: \label{Sbulk_decomp}
352: S_{bulk}=S_B+S_W
353: \ee
354: where:
355: \bea
356: \label{S_B}
357: \!\!\!\!\!S_B&=&\int_\Sigma {d^2\sigma \sqrt{g}}{\Big[ G_{i{\bar
358: j}}\left(g^{\alpha\beta} \partial_\alpha \phi^i \partial_\beta \phi^{\bar j}
359: -i\varepsilon^{\alpha\beta}\partial_\alpha \phi^i\partial_\beta\phi^{\bar
360: j}-\frac{1}{2}g^{\alpha\beta}\rho_\alpha^i D_\beta \eta^{\bar
361: j}-\frac{i}{2}\varepsilon^{\alpha\beta} \rho_\alpha^i D_\beta \theta^{\bar
362: j}-{\tilde F}^i{\tilde F}^{\bar j}\right)}\nn\\
363: \!\!\!\!&+&\frac{i}{4}\varepsilon^{\alpha\beta}R_{i{\bar l}k{\bar
364: j}}\rho^i_\alpha{\bar \chi}^{\bar l}\rho^k_\beta\chi^{\bar j}\Big] 
365: \eea
366: is the action of the B-twisted sigma model and $S_W=S_0+S_1$ is the potential-dependent term, with:
367: \bea
368: S_0&=&-\frac{i}{2} \int_{\Sigma}{d^2\sigma \sqrt{g} \left[D_{\bar i}
369: \partial_{\bar j}{\bar W} \chi^{\bar i}{\bar \chi}^{\bar j}-(\partial_{\bar
370: i}{\bar W}){\tilde F}^{\bar i}\right]}\\
371: S_1&=&-\frac{i}{2}\int_{\Sigma}{d^2\sigma \sqrt{g} \left[(\partial_i W){\tilde
372: F}^i + \frac{i}{4} \varepsilon^{\alpha\beta}D_i\partial_jW
373: \rho_\alpha^i\rho_\beta^j\right]}~~.
374: \eea
375: The quantity $\varepsilon^{\alpha\beta}=\frac{\epsilon^{\alpha\beta}}{\sqrt{g}}$ is
376: the Levi-Civita tensor, while $\epsilon^{\alpha\beta}$ is the associated density. We
377: have rescaled the Landau-Ginzburg potential $W$ by a factor of $\frac{i}{2}$
378: with respect to the conventions of \cite{Labastida} (the conventions for the
379: target space Riemann tensor and covariantized worldsheet derivative $D_\alpha$
380: are unchanged).  In $S_W$, we separated the term depending on $W$ from that
381: depending on its complex conjugate.
382: 
383: As shown in \cite{Labastida}, the topological sigma model action 
384: (\ref{S_B}) is BRST exact on closed Riemann surfaces.  Since in this
385: paper we shall allow $\Sigma$ to have a nonempty boundary, we must be careful
386: with total derivative terms. Extending the computation of \cite{Labastida} to
387: this case, we find:
388: \be
389: \label{exactS_B}
390: S_B+s=\delta V_B
391: \ee
392: where:
393: \be
394: V_B:=\int_{\Sigma}{d^2\sigma \sqrt{g} G_{i{\bar
395:   j}}\left(\frac{1}{2}g^{\alpha\beta}\rho^i_\alpha\partial_\beta \phi^{\bar
396:   j}-\frac{i}{2}\varepsilon^{\alpha\beta}\rho^i_\alpha
397:   \partial_\beta\phi^{\bar j}-{\tilde F}^i\chi^{\bar j}\right)}
398: \ee
399: and:
400: \be
401: \label{s}
402: s:=i\int_{\Sigma}{d^2\sigma
403:       \sqrt{g}\varepsilon^{\alpha\beta}\partial_\alpha(G_{{\bar i}j}\chi^{\bar
404:       i}\rho^j_\beta)}=i\int_\Sigma{d(G_{{\bar i}j}\chi^{\bar i}\rho^j)}~~.
405: \ee
406: Since total derivative  terms do not change physics on
407: closed Riemann surfaces, we are free to redefine the bulk 
408: sigma-model action by adding (\ref{s}) to $S_B$:
409: \be
410: {\tilde S}_B:=S_B+s=\delta V_B~~.
411: \ee
412: Accordingly, we shall use the modified bulk Landau-Ginzburg action:
413: \be
414: \label{tilde_S_bulk}
415: {\tilde S}_{bulk}=S_{bulk}+s={\tilde S}_B+S_0+S_1~~.
416: \ee
417: It is not hard to check that the term $S_0$ is BRST exact:
418: \be
419: \label{exactS_0}
420: S_0=\delta V_0
421: \ee
422: where:
423: \be
424: V_0=\frac{i}{4}\int_{\Sigma}{d^2\sigma \sqrt{g}\theta^{\bar i} \partial_{\bar
425: i}{\bar W}}~~.
426: \ee
427: Equations (\ref{exactS_B}) and (\ref{exactS_0}) are local, i.e. they hold for
428: the associated Lagrange densities without requiring integration by parts. Thus
429: both of these relations can be applied to bordered Riemann surfaces.
430: 
431: Since the boundary term (\ref{s}) is independent of the worldsheet metric, 
432: the bulk stress energy tensor has the form given in \cite{Labastida}:
433: \bea
434: T_{\mu \nu}&=&\frac{1}{2\sqrt{g}}\frac{\delta {\tilde S}_{bulk}}{\delta g^{\mu
435:     \nu}}=\frac{1}{2}G_{i{\bar j}}\left[\partial_\mu \phi^i\partial_\nu \phi^{\bar j}+
436: \partial_\nu \phi^i\partial_\mu \phi^{\bar j}-\frac{1}{2}\left(\rho_\mu^iD_\nu
437: \eta^{\bar j}+\rho_\nu^i D_\mu \eta^{\bar j}\right)\right]-\\
438: &-&\frac{1}{2}g_{\mu \nu}\left[G_{i{\bar j}}g^{\alpha\beta}(\partial_\alpha\phi^i\partial_\beta\phi^{\bar
439:     j}-\frac{1}{2} \rho_\alpha^i D_\beta \eta^{\bar j}-{\tilde F}^i{\tilde
440:     F}^{\bar j})-\frac{i}{2}\partial_i W {\tilde
441:     F}^i+\frac{i}{2}\partial_{\bar i}{\bar W}{\tilde
442:     F}^{\bar i}-\frac{i}{4}D_{\bar i}\partial_{\bar j}{\bar W}\theta^{\bar
443:     i}\eta^{\bar j}\right]~~.\nn
444: \eea
445: As explained in \cite{Labastida}, $T_{\mu\nu}$ is BRST exact only modulo
446: the equations of motion for the auxiliary fields:
447: \bea
448: \label{F_eom}
449: {\tilde F}^i=\frac{i}{2}G^{i{\bar j}}\partial_{\bar j}{\bar W}~~,~~
450: {\tilde F}^{\bar i}=-\frac{i}{2}G^{{\bar i}j}\partial_j W~~.
451: \eea
452: Imposing these equations, one finds:
453: \bea
454: T^{os}_{\mu\nu}&=&\frac{1}{2}G_{i{\bar j}}\left[\partial_\mu \phi^i\partial_\nu \phi^{\bar j}+
455: \partial_\nu \phi^i\partial_\mu \phi^{\bar j}-\frac{1}{2}\left(\rho_\mu^iD_\nu
456: \eta^{\bar j}+\rho_\nu^i D_\mu \eta^{\bar j}\right)\right]-\\
457: &-&\frac{1}{2}g_{\mu \nu}\left[G_{i{\bar j}}g^{\alpha\beta}(\partial_\alpha\phi^i\partial_\beta\phi^{\bar
458:     j}-\frac{1}{2} \rho_\alpha^i D_\beta \eta^{\bar j})+\frac{1}{4}G^{i{\bar j}}\partial_i W\partial_{\bar
459:     j}{\bar W}-\frac{i}{4}D_{\bar i}\partial_{\bar j}{\bar W}\theta^{\bar
460:     i}\eta^{\bar j}\right]~~.\nn
461: \eea
462: This obeys the BRST exactness condition \cite{Labastida}:
463: \be
464: \label{T_exact}
465: T^{os}_{\mu \nu}=\delta G_{\mu \nu}
466: \ee
467: where\footnote{The formula given for $G_{\mu \nu}$ in \cite{Labastida} seems
468:   to be missing a global prefactor of $-\frac{1}{2}$.}:
469: \be
470: \label{G_mu_nu}
471: G_{\mu\nu}=\frac{1}{4}\Big[G_{i{\bar j}}(\rho_\mu^i\partial_\nu \phi^{\bar j}+
472: \rho_\nu^i\partial_\mu \phi^{\bar j})-g_{\mu
473:   \nu}(G_{i{\bar j}}g^{\alpha\beta}\rho_\alpha^i\partial_\beta\phi^{\bar
474:   j}+\frac{i}{2}\theta^{\bar i}\partial_{\bar i}{\bar W})\Big]~~.
475: \ee
476: On an infinite flat cylinder, the supercharges: 
477: \be
478: \label{G_mu}
479: G_\mu:=\int{d\sigma_1 G_{0\mu}} 
480: \ee
481: generate symmetries $\delta_\mu=\{G_\mu, \cdot\}_P$ which together with $\delta=\{Q,\cdot\}_P$
482: and a supplementary nilpotent transformation $\delta':=\{M,\cdot\}_P$ 
483: form the topological algebra of
484: \cite{Labastida} (here  $\{\cdot, \cdot\}_P$ is the Poisson bracket of the
485: Hamiltonian formulation). When placing the model on a flat strip, the boundary
486: conditions break the symmetries $\delta'$ and $\delta_1$, but preserve the subalgebra
487: generated by $\delta$ and $\delta_0$. In the untwisted model, this subalgebra
488: corresponds to the usual B-type supersymmetry considered, for example, in \cite{Hori}.
489: 
490: 
491: 
492: 
493: 
494: \subsection{BRST variation of the bulk action and the topological Warner term}
495: 
496: It is not hard to check that the BRST variation of ${\tilde S}_{bulk}$ produces
497: a boundary term:
498: \be
499: \label{deltaSbulk}
500: \delta {\tilde S}_{bulk} =\delta S_1=\frac{1}{2} \int_{\partial
501:   \Sigma}{\rho^i\partial_i W}~~.
502: \ee
503: The presence of a non-zero right hand side in (\ref{deltaSbulk}) is known as
504:   the Warner problem \cite{Warner}.
505: 
506: 
507: \subsection{$\delta_0$-variation of the bulk action on flat Riemann surfaces}
508: 
509: When considered on a flat Riemann surface, our model has an enlarged symmetry
510: algebra which was originally described in \cite{Labastida}. In this paper we
511: shall need only the subalgebra obtained by considering an additional odd
512: generator $\delta_0$ beyond the BRST generator of (\ref{BRST}). In the
513: notations of \cite{Labastida}, we have $\delta_0=\{G_0,\}$, where
514: $G_0=G_z+G_{\bar z}$.  Using the results of \cite{Labastida}, one
515: finds:
516: \bea
517: \label{delta_0}
518: \delta_0\phi^i=\frac{1}{2}\rho^i_0~~&,&~~\delta_0\phi^{\bar i}=0\nn\\
519: \delta_0\rho_0^i=0~~&,&~~\delta_0\rho_1^i=-i{\tilde F}^i\\ 
520: \delta_0\eta^{\bar i}=\partial_0\phi^{\bar i}~~&,&~~\delta_0\theta^{\bar i}=-i\partial_1
521: \phi^{\bar i}\nn\\ 
522: \delta_0 {\tilde F}^i=0~~&,&~~\delta_0{\tilde F}^{\bar
523: i}=\frac{1}{2}(\partial_0\theta^{\bar i}+i\partial_1\eta^{\bar i})~~.\nn
524: \eea
525: In this subsection, we are assuming that the worldsheet metric is flat, namely
526: $g_{\alpha\beta}=\delta_{\alpha\beta}$ in the real coordinates $\sigma^0$ and
527: $\sigma^1$.
528: 
529: If $\Sigma$ is a cylinder, one finds $\delta_0{\tilde S}_{bulk}=0$. However,
530: the $\delta_0$ variation of ${\tilde S}_{bulk}$ gives a boundary term when the
531: model is considered on the strip or on the disk.  Let us take $\Sigma$ to be
532: infinite the strip given by $(\sigma^0,\sigma^1)\in \R\times [0,\pi]$. We find:
533: \be \delta_0 {\tilde S}_{bulk}=-\frac{1}{4}\int_{\partial \Sigma}{d\tau
534: \eta^{\bar i}\partial_{\bar i}{\bar W}}~~.  \ee Here $d\tau=d\sigma^0$ is the
535: length element along the boundary of $\Sigma$.
536: 
537: 
538: \section{Localization formula for correlators on the sphere}
539: \label{loc_bulk}
540: 
541: Let us consider zero-form bulk observables ${\cal O}$ which are independent
542: on the auxiliary fields ${\tilde F}^i$ or ${\tilde F}^{\bar i}$.  We are interested in the
543: sphere correlator of such observables:
544: \be
545: \label{correlator}
546: \langle {\cal O}\rangle_{\rm sphere} =\int{{\cal D}[\phi]{\cal D}[{\tilde F}]
547: {\cal D}[\eta]{\cal D}[\theta]{\cal D}[\rho]{~e^{-{\tilde S}_{bulk}}{\cal
548: O}}}~~,
549: \ee
550: where we assume that ${\cal O}$ is BRST closed.
551: 
552: In this section, we re-consider the localization formula for such correlators. 
553: Extending the original argument of \cite{Vafa_LG}, we will extract a
554: one-parameter family of representations of (\ref{correlator}) as a
555: finite-dimensional integral. The basic point is a follows. Since the B-model
556: piece of the Landau-Ginzburg action is BRST exact off-shell, the standard
557: argument of \cite{Witten_mirror} implies that we can localize on the zero
558: modes of the associated {\em sigma model}. Thus we shall localize on constant
559: maps, without requiring that such maps send the worldsheet to the critical
560: points of $W$. After this reduction, one finds that the Lagrangian density
561: of the model becomes BRST exact, so the resulting integral representation 
562: is insensitive to multiplying the Lagrangian density by a prefactor
563: $\lambda$. Since the former appears multiplied by the worldsheet area,
564: this prefactor measures the scale of the underlying $S^2$ worldsheet. Thus we
565: obtain a one-parameter family of localization formulae for our
566: correlator. Each such "localization picture" allows us to give a geometric
567: representation of genus zero data, thus providing a geometric model for
568: the off-shell state space, BRST operator and bulk trace. 
569: 
570: In this approach, the residue representation of \cite{Vafa_LG} is recovered in
571: the limit $\Re\lambda\rightarrow +\infty$, which forces the point-like image of
572: the worldsheet to lie on the critical set of $W$. Intuitively, this is the
573: limit of {\em large} worldsheet areas, the opposite of the
574: "microscopic" limit $\lambda\rightarrow 0$. 
575: Varying $\lambda$ allows one to interpolate between these
576: limits, thus connecting the "sigma-model like" and "residue-like" 
577: models of genus zero data. 
578: 
579: \subsection{Localization on $B$-model zero modes}
580: 
581: Since ${\tilde S}_B$ is BRST exact, we can replace the bulk
582: action with:
583: \be
584: {\tilde S}_{bulk}=t {\tilde S}_B+ S_W=t\delta V_B+S_W~~,
585: \ee
586: where $t$ is a complex parameter with $\Re t >0$ (so that the integral is
587: well-defined).  BRST invariance of the path integral together with BRST
588: closure of ${\cal O}$ imply that the resulting correlator is independent of
589: $t$. This means that (\ref{correlator}) can be computed in the limit
590: $\Re~t\rightarrow +\infty$, where the integral localizes on the zero-modes of
591: ${\tilde S}_B$. Since $\rho$ has no zero-modes on the sphere, we must
592: consider configurations for which $\rho^i_\alpha=0$ while $\phi$, $\eta$,
593: $\theta$ and ${\tilde F}$ are constant on the worldsheet.  For such configurations,
594: ${\tilde S}_B$ reduces to:
595: \be
596: {\tilde S}_B|_{\rm zero~modes}=-A G_{i{\bar j}}{\tilde F}^i{\tilde F}^{\bar j}~~,
597: \ee
598: with $G_{i{\bar j}}=G_{i{\bar j}}(\phi)$. Here $A$ is the area of the
599: worldsheet.  The contribution $S_W$ becomes:
600: \be
601: S_W|_{\rm zero~modes}=-\frac{i}{2}A\left[ D_{\bar i} \partial_{\bar j}{\bar W}
602: \chi^{\bar i}{\bar \chi}^{\bar j}- (\partial_{\bar i}{\bar W}){\tilde F}^{\bar
603: i}+ (\partial_i W){\tilde F}^i\right]~~.
604: \ee
605: Combining these expressions, we find the zero-mode reduction of the worldsheet
606: action:
607: \be
608: {\tilde S}_{0}:={\tilde S}_{bulk}|_{\rm zero~modes}=-A G_{i{\bar j}}{\tilde
609: F}^i{\tilde F}^{\bar j}- \frac{i}{2}A\left[ \frac{1}{2}D_{\bar i}
610: \partial_{\bar j}{\bar W} \theta^{\bar i}\eta^{\bar j}- (\partial_{\bar
611: i}{\bar W}){\tilde F}^{\bar i}+ (\partial_i W){\tilde F}^i\right]~~,
612: \ee
613: where we wrote $\chi$ and ${\bar \chi}$ in terms of $\eta$ and $\theta$.
614: The correlator (\ref{correlator}) reduces to an ordinary integral over the
615: zero-modes $\phi^i, \phi^{\bar i}$, ${\tilde F}^i, {\tilde F}^{\bar i}$ and
616: $\eta^{\bar i},\theta_i$:
617: \be
618: \label{correlator_0}
619: \langle {\cal O}\rangle_{\rm sphere} =\int{d\phi d{\tilde F} d\eta d\theta
620: {~e^{-{\tilde S}_{0}}{\cal O}}}~~.
621: \ee
622: On zero modes, the BRST generator (\ref{BRST}) takes the form:
623: \bea
624: \label{BRST_0}
625: \delta \phi^i=0~~&,&~~\delta \phi^{\bar i}=\eta^{\bar i}\nn\\ \delta
626: \eta^{\bar i}=0~~&,&~~\delta \theta^{\bar i}=2{\tilde F}^{\bar i}+\Gamma^{\bar
627: i}_{{\bar j}{\bar k}}\theta^{\bar j}\eta^{\bar k}~~\\ \delta {\tilde
628: F}^i=0~~&,&~~\delta {\tilde F}^{\bar i}=\Gamma^{\bar i}_{{\bar j}{\bar
629: k}}{\tilde F}^{\bar j}\eta^{\bar k}~~.\nn
630: \eea
631: In particular, we have:
632: \be
633: \label{tildeS_B_0_ex}
634: G_{i{\bar j}}{\tilde F}^i{\tilde F}^{\bar j}=\frac{1}{2}\delta [G_{i{\bar j}}{\tilde
635:   F}^i\theta^{\bar j}]~~,
636: \ee
637: which is the zero-mode remnant of equation (\ref{exactS_B}).  One can also
638: check directly that the reduced action is BRST-closed.
639: 
640: The integral over ${\tilde F}$ can be cast into Gaussian form through the change
641: of variables:
642: \be
643: {\tilde F}^i={\hat F}^i+\frac{i}{2}G^{i{\bar j}}\partial_{\bar j}{\bar W}~~,~~
644: {\tilde F}^{\bar i}={\hat F}^{\bar i}-\frac{i}{2}G^{{\bar i}j}\partial_j W~~.
645: \ee
646: Then the reduced action becomes: 
647: \be
648: {\tilde S}_{0}=-AG_{i{\bar j}}{\hat F}^i{\hat F}^{\bar j}+
649: \frac{1}{4}A \left[
650: -iD_{\bar i} \partial_{\bar j}{\bar W} \theta^{\bar
651:       i}\eta^{\bar j}+G^{i{\bar j}}(\partial_i W)(\partial_{\bar j}{\bar W})\right]~~.
652: \ee
653: Integrating over ${\hat F}$, we find:
654: \be
655: \label{correlator_0intmd}
656: \langle {\cal O}\rangle_{\rm sphere}=N \int{d\phi d\eta d\theta
657: e^{-\frac{A}{4} {\tilde L}_0}{\cal O}}~~,
658: \ee
659: where:
660: \be
661: \label{tildeL_0}
662: {\tilde L}_0:=-i D_{\bar i} \partial_{\bar j}{\bar W} \theta^{\bar
663:       i}\eta^{\bar j}+G^{i{\bar j}}(\partial_i W)(\partial_{\bar j}{\bar
664:       W})~~
665: \ee
666: plays the role of zero-mode Lagrange density. The prefactor in 
667: (\ref{correlator_0intmd})  has the form:
668: \be
669: N=\frac{(2\pi)^n}{ A^n \det(G_{i{\bar j}})}~~,
670: \ee
671: where $n$ is the complex dimension of the target space $X$. Since we
672: integrated out the fields ${\hat F}$, the BRST generator on
673: zero-modes reduces to:
674: \bea
675: \label{BRST_os_0}
676: \delta \phi^i=0~~&,&~~\delta \phi^{\bar i}=\eta^{\bar i}\nn\\
677: \delta \eta^{\bar i}=0~~&,&~~\delta \theta^{\bar
678:   i}=-iG^{{\bar i}j}\partial_jW+ \Gamma^{\bar i}_{{\bar j}{\bar
679:   k}}\theta^{\bar j}\eta^{\bar k}~~,
680: \eea
681: which is obtained from  (\ref{BRST_0}) by imposing the equations of motion:
682: \be
683: {\tilde F}^i=\frac{i}{2}G^{i{\bar j}}\partial_{\bar j}{\bar W}~~,~~
684: {\tilde  F}^{\bar i}=-\frac{i}{2}G^{{\bar i}j}\partial_jW~~.
685: \ee
686: For later reference, notice that the last relation in (\ref{BRST_os_0}) is equivalent with:
687: \be
688: \label{delta_theta_os}
689: \delta \theta_i=-i\partial_i W~~.
690: \ee
691: 
692: Using this form of the BRST transformations, one finds that the zero-mode
693: Lagrange density (\ref{tildeL_0}) is BRST exact:
694: \be
695: \label{exact_tildeL_0}
696: {\tilde L}_0=\delta {\tilde v}_0~~,
697: \ee
698: where:
699: \be
700: \label{tildev_0}
701: {\tilde v}_0=i\theta^{\bar i}\partial_{\bar i}{\bar W}~~.
702: \ee
703: Thus we can replace  (\ref{correlator_0intmd}) by:
704: \be
705: \label{correlator_final}
706: \langle {\cal O}\rangle_{\rm sphere} 
707: =N\int{d\phi d\eta d\theta {~e^{-\lambda {\tilde L}_0}{\cal O}}}~~,
708: \ee
709: where $\lambda$ is a complex parameter with positive real part. The integral
710: (\ref{correlator_final}) is independent of its value.
711: 
712: 
713: \subsection{The space of bulk observables and its cohomology}
714: \label{bulk_observables}
715: 
716: The observables of interest have the form:
717: \be
718: \label{bulk_obs}
719: {\cal O}_\omega(\sigma):=\omega^{j_1\dots j_q}_{{\bar i}_1\dots {\bar
720:     i}_p}(\phi(\sigma))\eta^{{\bar
721:     i}_1}(\sigma) \dots \eta^{{\bar i}_p}(\sigma)
722: \theta_{j_1}(\sigma) \dots \theta_{j_q}(\sigma)~~,
723: \ee
724: where $\omega:=\omega^{j_1\dots j_q}_{{\bar i}_1\dots {\bar i}_p}dz^{{\bar
725:   i}_1}\wedge \dots \wedge dz^{{\bar i}_p}\wedge \partial_{j_1}\wedge \dots \wedge \partial_{j_q}$
726: is a section of the bundle 
727: $\Lambda^p{{\bar T}^*X}\wedge \Lambda^q TX$. After reduction to B-model 
728: zero modes, we are left with:
729: \be
730: {\cal O}_\omega=\omega^{j_1\dots j_q}_{{\bar i}_1\dots {\bar
731:     i}_p}(\phi)\eta^{{\bar
732:     i}_1} \dots \eta^{{\bar i}_p}
733: \theta_{j_1}\dots \theta_{j_q}~~,
734: \ee
735: which can be identified with the polyvector-valued form $\omega$ upon setting:
736: \be
737: \label{ids_geom}
738: \eta^{\bar j}\equiv dz^{\bar j}~~,~~\theta_j\equiv \partial_j~~. 
739: \ee
740: 
741: The reduced BRST operator (\ref{BRST_os_0})  becomes:
742: \be
743: \label{delta_geom}
744: \delta\equiv {\bar \partial}+i_{\partial W}
745: \ee
746: where $i_{\partial_W}$ is the odd derivation of $\Lambda^*TX$ uniquely determined
747: by the conditions:
748: \be
749: i_{\partial W}(\partial_j)=-i\partial W(\partial_j)=-i\partial_j W~~.
750: \ee
751: Thus the cohomology of the differential superalgebra 
752: $({\cal H}, \delta)$, where ${\cal H}:=\Gamma(\Lambda^*{\bar T}^*X\wedge \Lambda^*TX)$, models
753: the algebra of bulk observables. 
754: The obvious  relations:
755: \be
756: {\bar \partial}^2=(i_{\partial W})^2={\bar
757:   \partial}i_{\partial_W}+i_{\partial_W}{\bar \partial}=0~~
758: \ee
759: show that $({\cal H},\delta)$ is a bicomplex. Hence the BRST cohomology is
760: computed by a spectral sequence $E_*$ whose second term equals:
761: \be
762: \label{E}
763: E_2:=H_{i_{\partial_W}}(H_{\bar \partial}({\cal H}))~~.
764: \ee
765: Since the target space is non-compact, we must of course specify a growth condition at
766: infinity. We shall take ${\cal H}$ to consist of those sections of 
767: the bundle $\Lambda^*{\bar T}^*X\wedge \Lambda^*TX$ whose coefficients have at most
768: polynomial growth. When the spectral sequence collapses to its second term,
769: the BRST cohomology reduces to (\ref{E}). A standard example is the case
770: $X=\C^n$, with $W$ a polynomial function of $n$ variables. 
771: Then the $\bar{\partial}$-Poincare Lemma implies 
772: that $H_{\bar \partial}({\cal H})$ coincides with the space
773: $\Gamma_{poly}(\Lambda^*TX)$ of polyvector fields with polynomial
774: coefficients.  In this case, the BRST cohomology reduces to the Jacobi ring
775: $\C[x_1\dots x_n]/\langle \partial_1 W\dots \partial_n W\rangle$, thereby
776: recovering a well-known result.  
777: 
778: 
779: 
780: 
781: 
782: \subsection{The geometric model}
783: \label{bulk_geom}
784: 
785: Let us translate (\ref{correlator_final}) into
786: classical mathematical language. Using (\ref{ids_geom}), we find:
787: \be
788: \label{tildeL_0form_0}
789: {\tilde L}_0\equiv i{\bar H}^{i}_{~{\bar j}}
790: dz^{\bar j}\wedge \partial_{i}+
791: G^{i{\bar j}}(\partial_iW)(\partial_{\bar j} {\bar W}) \in {\cal H}~~.
792: \ee
793: Here $H^i_{~{\bar j}}:=G^{i{\bar k}}{\bar H}_{{\bar k}{\bar j}}$, where 
794: $H_{ij}:=D_i\partial_j W$ is the Hessian of $W$. Consider the Hessian
795: operator:
796: \be
797: H=H^{\bar i}_{~j}dz^j\otimes \partial_{\bar i}\in Hom(TX,{\bar T}X)=T^*X\otimes
798: {\bar T}X~~,
799: \ee
800: whose complex conjugate has the form:
801: \be
802: {\bar H}={\bar H}^i_{~{\bar j}}dz^{\bar j}\otimes \partial_i\in Hom({\bar
803:   T}X,TX)={\bar T}^*X\otimes TX~~.
804: \ee
805: The quantity
806: ${\bar H}_a:={\bar H}^i_{~{\bar j}}dz^{\bar j}\wedge \partial_i\in {\bar T}^*X\wedge TX$ 
807: appearing in (\ref{tildeL_0form_0}) is the antisymmetric part of
808: ${\bar H}$.  On the other hand, the second term of (\ref{tildeL_0form_0}) is the norm of the
809: differential form $\partial W=\partial_iW dz^{i}$. This gives the
810: coordinate-independent version of (\ref{tildeL_0form_0}):
811: \be
812: {\tilde L}_0=i{\bar H}_{a}+||\partial W||^2~~.
813: \ee
814: Also note the representation:
815: \be
816: {\tilde v}_0=iG^{i{\bar j}}\partial_i\wedge \partial_{\bar j} {\bar W}~~.
817: \ee
818: It is now easy to see that (\ref{correlator_final}) becomes:
819: \be
820: \label{correlator_geom}
821: \Tr\omega:=\langle {\cal O}_\omega\rangle_{\rm sphere}=N\int_{X}{\Omega\wedge \left[
822: \Omega \lrcorner \left(e^{-\lambda {\tilde L}_0}\wedge \omega\right)\right]}~~,
823: \ee
824: where $\lrcorner$ denotes the total contraction of a form with a
825: polyvector. The linear functional $\Tr$ realizes the bulk trace of \cite{CIL1}. 
826: 
827: 
828: \paragraph{Observation}
829: The integral representation (\ref{correlator_geom}) allows us 
830: to give another (and completely rigorous) 
831: proof of $\lambda$-independence for $\delta
832: \omega=0$, with the assumption $\Re \lambda>0$. 
833: For this, we have to show that the $\lambda$-derivative of
834: (\ref{correlator_geom}) vanishes.
835: Since ${\tilde L}_0=\delta {\tilde v}_0$ and $\delta \omega=0$, this
836: derivative takes the form:
837: \be
838: \label{der_geom}
839: \frac{d}{d\lambda}\Tr \omega =
840: -\lambda N\int_{X}{\Omega\wedge \left[\Omega \lrcorner 
841: \delta \left(e^{-\lambda {\tilde L}_0}\wedge {\tilde v}_0\wedge \omega\right)\right]}~~.
842: \ee
843: Thus it suffices to show that 
844: $\int_{X}{\Omega\wedge \left[\Omega \lrcorner \delta \alpha\right]}$
845: vanishes for any $\alpha\in {\cal H}$ which decays exponentially at infinity
846: on $X$ (the exponential  decay for $\alpha=e^{-\lambda {\tilde L}_0}{\tilde
847:   v}_0\wedge \omega$
848: in (\ref{der_geom}) is due to the second term in (\ref{tildeL_0form_0})). 
849: Notice further that $\int_{X}{\Omega\wedge \left[\Omega
850:   \lrcorner \delta \alpha\right]}$
851: vanishes for degree reasons unless $\delta \alpha\in \Gamma(\Lambda^n{\bar T}^*X\wedge \Lambda^n
852: TX)$. Hence it is enough to show vanishing of $\int_{X}{\Omega\wedge \left[
853: \Omega \lrcorner \delta \alpha\right]}$ for an exponentially decaying $\alpha$
854: such that
855: $\delta \alpha\in \Gamma(\Lambda^n{\bar T}^*X\wedge \Lambda^n TX)$. In this
856: case, we obviously have $\delta\alpha={\bar \partial}\beta$ for some 
857: exponentially decaying $\beta\in \Gamma(\Lambda^{n-1}{\bar T}^*X\wedge
858: \Lambda^n TX)$ (this follows by noticing that the
859: image of $i_{\partial W}$ has vanishing intersection with the subspace 
860: $\Gamma(\Lambda^n{\bar T}^*X\wedge \Lambda^n TX)$). Therefore, we only need to show
861: that $\int_{X}{\Omega\wedge \left[\Omega \lrcorner {\bar \partial} \beta\right]}$
862: vanishes. This last fact follows from 
863: $\Omega\wedge \left[\Omega \lrcorner {\bar \partial}
864:   \beta\right]={\bar \partial}\left(\Omega\wedge \left[\Omega
865:     \lrcorner \beta\right]\right)$, since the boundary term
866: vanishes due to the exponential decay of $\beta$. 
867: The assumption $\Re\lambda>0$ is crucial,
868: since otherwise we cannot rely on exponential decay to conclude that the
869: boundary term vanishes.
870: 
871: 
872: \subsection{Localization pictures and homotopy flows}
873: \label{bulk_flow}
874: 
875: Expression (\ref{correlator_geom}) admits the following interpretation.
876: Consider the one-parameter semigroup of operators $U(\lambda)$ acting on
877: ${\cal H}$ through wedge multiplication by $e^{-\lambda{\tilde L}_0}$:
878: \be
879: U(\lambda)\omega:=e^{-\lambda{\tilde L}_0}\wedge \omega~~{\rm~for~all}~~\omega\in {\cal H}~~.
880: \ee
881: The semigroup is defined on the half-plane $\Delta:=\{\lambda\in
882: \C|\Re\lambda>0\}$, so that $U(\lambda)$ maps ${\cal H}$ into a subspace of
883: itself. Then (\ref{correlator_geom}) takes the form: 
884: \be
885: \label{Tr_flow}
886: \Tr\omega:=\Tr^B(U(\lambda)\omega)~~,
887: \ee
888: where $\Tr^B$ is the bulk trace of the B-twisted sigma model: 
889: \be
890: \Tr^B\omega:=N\int_{X}{\Omega\wedge \left(
891: \Omega \lrcorner \omega\right)}~~.
892: \ee
893: 
894: Since ${\tilde L}_0$ is BRST closed (${\tilde L}_0=\delta {\tilde v}_0$), each
895: operator $U(\lambda)$ is homotopy equivalent with the identity in the complex $({\cal H},
896: \delta)$:
897: \be
898: U(\lambda)=1+[\delta, W_\lambda]~~,
899: \ee
900: for some operator $W_\lambda$. In particular, $U(\lambda)$ is an endomorphism of our
901: complex, i.e. the following relation holds: 
902: \be
903: \label{endomorphism}
904: U(\lambda)\circ \delta=\delta \circ U(\lambda)~~. 
905: \ee
906: Such a semigroup will be called a {\em homotopy
907: flow}. It is clear that each $U(\lambda)$ is a quasi-isomorphism, i.e. 
908: induces an automorphism $U_*(\lambda)$ on the BRST cohomology $H_\delta({\cal
909:   H})$. Following relation (\ref{Tr_flow}), we define the {\em localization
910:   picture $\lambda$} by associating $\omega_\lambda:=U(\lambda)(\omega)\in
911: {\cal H}$ to each $\omega\in {\cal H}$ (then $\omega_\lambda$ is the representative of the "state"
912: $\omega$ in the picture $\lambda$). The representatives of this
913: picture belong to the subspace ${\cal H}_\lambda:=U(\lambda)({\cal
914:   H})\subset {\cal H}$. As in quantum mechanics, we have a representative for
915: any operator $T\in End({\cal H})$ in the localization picture $\lambda$:
916: \be
917: T_\lambda:=U(\lambda)\circ T \circ U(-\lambda)\in End({\cal H}_\lambda)~~,
918: \ee
919: where $\Re\lambda>0$ and $U(-\lambda)$ is defined as an operator from ${\cal
920:   H}_\lambda$ to ${\cal H}$. Relation (\ref{endomorphism})
921: shows that the BRST operator is "picture-independent" in the following sense: 
922: \be
923: Q_\lambda=Q|_{{\cal H}_\lambda}~~,
924: \ee
925: where in the right hand side we restrict both the domain and image 
926: to ${\cal H}_\lambda$. Relation (\ref{Tr_flow}) becomes:
927: \be
928: \Tr\omega=\Tr^B\omega_\lambda~~. 
929: \ee
930: 
931: For $\lambda=W=0$, we have $U(0)=Id_{\cal H}$ and we recover the familiar data
932: of the B-twisted sigma model. Namely, ${\cal H}$ provides a geometric model for the
933: off-shell state space, the Dolbeault operator ${\bar \partial}$ models the "localized
934: BRST operator" and $\Tr^B$ models the bulk trace of \cite{CIL1}. 
935: Turning on the Landau-Ginzburg superpotential $W$ and
936: performing localization as above with "worldsheet area" $\lambda$  leads to a
937: geometric model given by the triplet $({\cal H}, \delta, \Tr)$. This is
938: related to the triplet describing the B-twisted sigma model by the modification
939: $\delta={\bar \partial}+i_{\partial W}$ of the BRST operator, followed by the
940: homotopy flow $U(\lambda)$. 
941: 
942: Because varying $\lambda$ along the real axis amounts to changing the area of the
943: worldsheet, the operators $U(\lambda)$ implement a sort of
944: "renormalization group flow" connecting the point-like (UV) limit $\lambda=0$
945: with the large area (IR limit) $\lambda=+\infty$. Since the model is
946: topological, such a flow "does nothing" at the level of BRST cohomology, but
947: acts non-trivially off-shell.
948: 
949: 
950: \subsection{The residue formula for sphere correlators}
951: 
952: Since the integral (\ref{correlator_final}) is independent of $\lambda$, we
953: can compute its value for $\Re \lambda\rightarrow +\infty$ (more specifically,
954: we shall take $\lambda\rightarrow +\infty$ with $\lambda \in \R$). In this limit,
955: the second term in (\ref{tildeL_0}) forces the integral to localize on the
956: critical points of $W$, and
957: the Gaussian approximation  around these points 
958: becomes exact. For simplicity, we shall assume that the critical points of $W$
959: are isolated (the general case can be incorporated by a continuity argument). 
960: For simplicity, we shall also assume that the spectral sequence of Subsection
961: \ref{bulk_observables} collapses to its second term. 
962: 
963: Taking $\lambda\rightarrow +\infty$ with $\lambda \in \R$, we find that the correlator 
964: (\ref{correlator_final}) vanishes unless $\omega=f$ with $f$ a complex-valued 
965: function defined on $X$ \footnote{For this, notice that the bosonic Gaussian integral over
966:   fluctuations of $\phi$ around
967:   each critical point of $W$ produces a factor which is weighted by
968:   $\frac{1}{\lambda^n}$. Thus the fermionic Gaussian integral over $\theta$
969:   and $\eta$ must produce $n$ powers of $\lambda$ if one is to obtain a
970:   non-vanishing result in the limit $\lambda\rightarrow \infty$. This
971:   obviously requires that ${\cal O}_\omega$ contain no $\eta$'s or $\theta$'s, 
972:   so that the highest ($n$-th order term) in the expansion of $e^{+i\lambda D_{\bar i}
973:     \partial_{\bar j}{\bar W} \theta^{\bar i}\eta^{\bar j}}$ survives when
974:   performing the integral over $\eta$ and $\theta$. }. In this case, we obtain:
975: \bea &&\langle {\cal O}_f \rangle_{\rm sphere}=\nn\\
976: &=&\lim_{\lambda\rightarrow +\infty} \left(N\sum_{p\in {\rm Crit}
977: W}{\left[n!^2 (i\lambda)^n (-1)^{n(n-1)/2}\det ({\bar H}_{{\bar i}{\bar
978: j}})\right]\left[\frac{(2\pi)^n}{\lambda^n \det (H_{ij}) \det ({\bar H}_{{\bar
979: i}{\bar j}})}\det (G_{i{\bar j}})\right]}f|_{p}+O(1/\lambda)\right)\nn\\
980: &=&n!^2 (2\pi)^n (-1)^{n(n+1)/2}N\det (G_{i{\bar j}})\sum_{p\in {\rm Crit} W}{
981: \frac{1}{\det (H_{ij}(p))} f(p)}~~.  \eea
982: Since $\sum_{p\in {\rm Crit} W}{\frac{1}{\det (H_{ij}(p))} f(p)}
983: \propto \int_{X}{\frac{f(z)dz^1\wedge\dots \wedge dz^n }{\partial_1W\dots
984:   \partial_n W}}$ by residue theory \cite{Griffiths}, 
985: one recovers the following generalization of the well-known result of \cite{Vafa_LG}:
986: \bea
987: \langle {\cal O}_\omega \rangle_{\rm sphere}&=&0~~{\rm unless}~~\omega=f \\
988: \langle {\cal O}_f\rangle_{\rm sphere}&=& C
989: \int_{X}{\Omega \frac{f(z) }{\partial_1W\dots  \partial_n W}}~~.
990: \eea
991: Here $C$ is an uninteresting normalization constant. 
992: 
993: 
994: 
995: 
996: 
997: 
998: 
999: 
1000: \section{The boundary coupling}
1001: \label{boundary_coupling}
1002: 
1003: In this section we discuss the boundary coupling of our models. The
1004: construction is based on \cite{coupling}, with a certain modification which
1005: will prove useful later on. After recalling the basics of 
1006: superconnections, we construct the coupling in the form of \cite{coupling},
1007: with the addition of a term which insures $\delta_0$-invariance on a flat
1008: strip. While this does not affect the target space equations of motion,
1009: it will help us make contact with previous work on the subject. We also give
1010: the target space reflection of the $\delta_0$-invariance constraint. 
1011: 
1012: 
1013: 
1014: \subsection{Mathematical preparations}
1015: 
1016: Consider a complex superbundle $E=E_+\oplus E_-$ over $X$, and a
1017: superconnection \cite{Quillen} ${\cal B}$ on $E$. We let $r_\pm:=\rk E_\pm$. 
1018: The bundle of endomorphisms $End(E)$ is endowed with the 
1019: natural $\Z_2$ grading, with even and odd components:
1020: \bea
1021: End_+(E)&:=&~~~~~End(E_+)\oplus End(E_-)\\
1022: End_-(E)&:=& Hom(E_+,E_-)\oplus Hom(E_-,E_+)~~.
1023: \eea
1024: The superconnection ${\cal B}$  can be viewed 
1025: as a section of $[{\cal T}^*X\otimes End_+(E)]\oplus End_-(E)$. 
1026: In a local frame of $E$ compatible with the grading, this is a matrix:
1027: \be
1028: {\cal B}=\left[\ba{cc} A^{(+)}& F\\G & A^{(-)}\ea\right]
1029: \ee
1030: whose diagonal entries $A^{(\pm)}$ are connection one-forms on $E_\pm$, while 
1031: $F,G$ are elements of  $Hom(E_-,E_+)$ and $Hom(E_+,E_-)$. We 
1032: require that the superconnection has type $(0,\leq 1)$, i.e. the one-forms
1033: $A^{(\pm)}$ belong to $\Omega^{(0,1)}(End(E_\pm))$. The morphism $F$ should not be
1034: confused with the curvature form used below.
1035: 
1036: When endowed with the ordinary composition of morphisms, the space of sections
1037: $\Gamma(End(E))$ becomes an associative superalgebra. The space
1038: ${\cal H}_b:=\Omega^{(0,*)}(End(E))$ also carries an associative superalgebra structure,
1039: which is induced from $(\Omega^{(0,*)}(X),\wedge)$ and 
1040: $(\Gamma(End(E)),\circ)$ via the tensor product decomposition:
1041: \be
1042: \Omega^{(0,*)}(End(E))=\Omega^{(0,*)}(X)\otimes_{\Omega^{(0,0)}(X)} \Gamma(End(E))~~.
1043: \ee
1044: For decomposable
1045: elements $u=\omega\otimes f$ and $v=\eta\otimes g$, with
1046: homogeneous  
1047: $\omega,\eta$ and $f,g$, the associative product on ${\cal H}_b$ takes the form:
1048: \be
1049: uv=(-1)^{ \deg f~\rk \eta}(\omega\wedge \eta)\otimes (f\circ g)~~,
1050: \ee
1051: where $\deg$ denotes the grading of the superalgebra $End(E)$:
1052: \be
1053: \deg(f)=0\in \Z_2~~{\rm~if~}f\in End_+(E)~~,~~\deg(f)=1\in \Z_2~~{\rm~if~}f\in End_-(E)~~.
1054: \ee
1055: The total degree on ${\cal H}_b$ is given by:
1056: \be
1057: |\omega\otimes f|=\rk \omega +\deg f~~(mod~2)~~.
1058: \ee
1059: We also recall the supertrace on $End(E)$:
1060: \be
1061: \label{str}
1062: \str(f)=\tr f_{++}-\tr f_{--}~~,
1063: \ee
1064: where $f=\left[\ba{cc}f_{++}&f_{-+}\\f_{+-}&f_{--}\ea\right]$ is an
1065: endomorphism of $E$ with components $f_{\alpha\beta}\in Hom(E_\alpha,E_\beta)$
1066: for $\alpha,\beta=+,-$. This has the property:
1067: \be
1068: \label{str_cyc}
1069: \str(f\circ g)=(-1)^{\deg f \deg g}\str(g\circ f)~~
1070: \ee
1071: for homogeneous elements $f,g$. 
1072: 
1073: The twisted Dolbeault operator:
1074: \be
1075: \label{sc}
1076: {\bar {\cal D}}={\overline \partial} +{\cal B}=\left[\ba{cc} 
1077: {\bar \partial}+A^{(+)}& F\\G &{\bar \partial}+A^{(-)}\ea\right]
1078: \ee
1079: induces an odd derivation ${\bar \partial}+[{\cal B},\cdot]$ 
1080: of the superalgebra ${\cal H}_b$, where
1081: $[u,v]:=uv-(-1)^{|u||v|}vu$ is the supercommutator. 
1082: 
1083: The $(0,\leq 2)$ part of the superconnection's curvature has the form:
1084: \be
1085: \label{curvature}
1086: {\cal F}^{(0,\leq 2)}={\bar {\cal D}}^2={\overline \partial}
1087: {\cal B}+\frac{1}{2}[{\cal B}, {\cal B}]={\overline \partial}
1088: {\cal B}+{\cal B}{\cal B}=
1089: \left[\ba{cc} F^{(+)}_{(0,2)}+FG& {\bar \nabla} F\\{\bar \nabla} G &F^{(-)}_{(0,2)}+GF\ea\right]
1090: \ee
1091: where $F^{(\pm)}_{(0,2)}$ are the $(0,2)$ pieces of the curvature forms
1092: $F^{(\pm)}$ of $A^{(\pm)}$ and:
1093: \bea
1094: {\bar \nabla}F&=&{\bar \partial}F+
1095: A^{(+)}F+FA^{(-)}={\bar \partial}F+A^{(+)}\circ F-F\circ A^{(-)}~~\nn\\
1096: {\bar \nabla}G&=&{\bar \partial}G+A^{(-)}G+G A^{(+)}
1097: ={\bar \partial} G+A^{(-)}\circ G-G\circ A^{(+)}~~.
1098: \eea
1099: 
1100: We will use the the notations:
1101: \be
1102: \label{AD}
1103: A:=A^{(+)}\oplus A^{(-)}=\left[\ba{cc} A^{(+)}& 0\\0 & A^{(-)}\ea\right]~~,~~
1104: D:=\left[\ba{cc} 0& F\\G & 0\ea\right]
1105: \ee
1106: for the diagonal and off-diagonal parts of ${\cal B}$. Then $A$ is an
1107: connection one-form on $E$ (compatible with the grading), while
1108: $D$ is an odd endomorphism. We have ${\cal B}=A+D$ and:
1109: \be
1110: \label{02curvature}
1111: {\cal F}^{(0,\leq 2)}=F^{(0,2)}+{\bar \nabla}_A D +D^2~~.
1112: \ee
1113: Here $F^{(0,2)}=F^{(+)}_{(0,2)}+F^{(-)}_{(0,2)}$ is the $(0,2)$ part of the curvature of $A$ and 
1114: ${\bar \nabla}_A={\bar \partial}+[A,\cdot]$ is the Dolbeault operator twisted by $A$.  
1115: 
1116: \subsection{The boundary coupling}
1117: 
1118: Following \cite{coupling}, we define the partition function on a bordered and
1119: oriented Riemann surface $\Sigma$ by:
1120: \be
1121: \label{Z}
1122: Z:=\int{{\cal D}[\phi]{\cal D}[{\tilde F}]{\cal D}[\theta]{\cal D}[\rho]{\cal D}[\eta]
1123:   e^{-{\tilde S}_{bulk}} {\cal U}_1\dots {\cal U}_h}~~,
1124: \ee
1125: where $h$ is the number of holes and the factors ${\cal U}_a$ have the form:
1126: \be
1127: \label{calU}
1128: {\cal U}_a:=\Str Pe^{-\oint_{C_a}{d\tau_a M}}~~.
1129: \ee
1130: We are assuming that the boundary of $\Sigma$ is a disjoint union of 
1131: smooth circles $C_a$, associated with holes labeled by $a$. 
1132: The symbol $\Str$ denotes the supertrace on $GL(r_+|r_-)$, while $d\tau_a$ 
1133: stands for the length element along $C_a$ induced by the metric on the
1134: interior of $\Sigma$. The quantity $M$ is given by:
1135: \be
1136: \label{M}
1137: M=\left[\ba{cc} 
1138: {\hat {\cal A}}^{(+)}+\frac{i}{2}(FF^\dagger+G^\dagger G) &
1139: \frac{1}{2}\rho^i_0 \nabla_i F+\frac{i}{2}\eta^{\bar i}\nabla_{\bar i}G^\dagger
1140: \\
1141: \frac{1}{2} \rho^i_0 \nabla_i G +\frac{i}{2}\eta^{\bar i}\nabla_{\bar
1142:   i}F^\dagger &
1143: {\hat {\cal A}}^{(-)} +\frac{i}{2}(F^\dagger F+GG^\dagger)
1144: \ea\right]~~.
1145: \ee
1146: Here $\rho_0^i d\tau_a$ is the pull-back of $\rho^i$ to $C_a$ and:
1147: \be
1148: \label{calApm} 
1149: {\hat {\cal A}}^{(\pm)}:=A_{\bar i}^{(\pm)}{\dot \phi}^{\bar i} +\frac{1}{2}\eta^{\bar i}  
1150: F^{(\pm)}_{{\bar i} j}\rho_0^j
1151: \ee
1152: are connections on the bundles ${\cal E}_\pm$ obtained by pulling back
1153: $E_\pm$ to the boundary of $\Sigma$. The dot in (\ref{calApm}) stands for the
1154: derivative $\frac{d}{d\tau_a}$. Notice that $\nabla_i F=\partial_i F$ and
1155: $\nabla_i G=\partial_i G$ since $A$ is a $(0,1)$-connection.
1156: 
1157: We have:
1158: \be
1159: \label{M_dec}
1160: M={\hat {\cal A}}+\Delta+K
1161: \ee
1162: where: 
1163: \be
1164: \Delta:=\frac{1}{2}\rho_0^i\partial_i D~~,
1165: \ee
1166: \be
1167: K:=\frac{i}{2}\left(\eta^{\bar i}\nabla_{\bar i}D^\dagger +[D,D^\dagger]_+ \right)
1168: \ee
1169: and:
1170: \be
1171: {\hat {\cal A}}={\dot \phi}^{\bar i}A_{\bar i}+\frac{1}{2}F_{{\bar
1172:     i}j}\eta^{\bar i}\rho^j_0~~.
1173: \ee
1174: Here  $A$ is the direct sum connection on $End(E)$ introduced in
1175: (\ref{AD}). The first two terms in (\ref{M_dec}) agree with \cite{coupling},
1176: while the last term $K$ is added for comparison with \cite{Kap2}. As we shall
1177: see below, this term preserves BRST-invariance of the partition function
1178: (which is already preserved by the sum of the first two terms
1179: \cite{coupling}). As for the open B-model,
1180: adding $K$ insures invariance of the boundary coupling with respect to 
1181: the second generator $\delta_0$ of the $N=2$ topological algebra, thereby
1182: fixing an ambiguity familiar form Hodge theory\footnote{
1183: The symmetry generators $\delta$ and $\delta_0$ can be viewed as analogues of
1184: the operators ${\bar \partial}$ and ${\bar \partial}^\dagger$ of Hodge theory, as already pointed out in \cite{Witten_mirror}
1185: in the context of twisted B-models. The boundary coupling of \cite{coupling}
1186: is chosen to preserve BRST invariance of the partition function. This is
1187: ambiguous up to addition of 'exact' terms, an ambiguity which we can
1188: fix by requiring $\delta_0$-invariance of the partition function. }. 
1189: This modification has minor effects which can be safely
1190: ignored for most purposes 
1191: \footnote{As we shall see in the next section, the extra-term in the boundary
1192:   coupling can be used to 
1193: introduce a parameter $\mu$ characterizing boundary localization pictures. 
1194: For most practical purposes, this parameter can be set to zero, which amounts
1195: to neglecting the last term in (\ref{M_dec}). In
1196: particular, one must set $\mu$ to zero in order to recover the trace
1197: formula of \cite{Kap2}. It is the {\em bulk} parameter $\lambda$ which must be
1198: taken to infinity in order to recover the proposal of \cite{Kap2}.}. 
1199: 
1200: \subsection{The target space equations of motion}
1201: 
1202: To insure BRST invariance of the partition function 
1203: (\ref{Z}), we must choose the background superconnection
1204: ${\cal B}$ such that:
1205: \be
1206: \label{invar}
1207: \delta {\cal U}_a = \frac{1}{2}\left[\int_{C_a}{d\tau
1208:     \rho^i_0\partial_i W} \right] {\cal U}_a~~.
1209: \ee
1210: In this paper, we also require $\delta_0$-invariance of the partition function on the flat
1211: strip:
1212: \be
1213: \label{invar_0}
1214: \delta_0 {\cal U}_a = -\frac{1}{4}\left[\int_{C_a}{d\tau
1215:     \eta^{\bar i}\partial_i {\bar W}} \right] {\cal U}_a~~.
1216: \ee
1217: It is not hard to check the relations:
1218: \be
1219: \label{Qhol}
1220: \delta {\cal U}_a=-\Str \left[~I_a(\delta M)
1221:   Pe^{-\oint_{C_a}{d\tau_a M}}\right]
1222: \ee
1223: where:
1224: \bea
1225: I_a(\delta M)&=&\oint_{C_a}d\tau_a 
1226:   U_a^{-1}\Big(F_{{\bar i}{\bar j}}\eta^{\bar i}{\dot \phi}^{\bar
1227:       j}-\frac{1}{4} \nabla_k F_{{\bar i}{\bar j}}\eta^{\bar i}\eta^{\bar
1228:       j}\rho_0^k-{\dot \phi}^{\bar
1229:       i}\nabla_{\bar
1230:       i}D-\frac{1}{2}\rho^i_0\nabla_i(D^2)+\frac{1}{2}\eta^{\bar
1231:       i}\rho^j_0\nabla_j\nabla_{\bar i}D+\nn\\
1232: &&+\frac{1}{2}\eta^{\bar i}\eta^{\bar j}[F_{{\bar i}{\bar
1233:   j}},D^\dagger]+\eta^{\bar i}[\nabla_{\bar i}D,D^\dagger]+[D^2,D^\dagger]
1234: \Big)U_a~,~~~~\nn
1235: \eea
1236: and:
1237: \be
1238: \label{delta_0hol}
1239: \delta_0 {\cal U}_a=-\Str \left[~I_a(\delta_0 M)
1240:   Pe^{-\oint_{C_a}{d\tau_a M}}\right]
1241: \ee
1242: where:
1243: \bea
1244: I_a(\delta_0 M)&=&\frac{1}{4}\oint_{C_a}{d\tau_a 
1245: U_a^{-1}\left(\eta^{\bar i}\nabla_{\bar i}(D^\dagger)^2+[D,(D^\dagger)^2]\right)U_a}\nn\\
1246: &+&\frac{i}{2}\oint_{C_a}{d\tau_a 
1247: U_a^{-1}\left( -{\dot \phi}^i \nabla_i D^\dagger +\frac{1}{2}\eta^{\bar i}\nabla_{\bar
1248:   i}\nabla_j D^\dagger \rho_0^j+\frac{1}{2}\rho_0^i [D,\nabla_i D^\dagger]
1249:   \right) U_a}~~.~~~~\nn
1250: \eea
1251: Here $U_a(\tau_a)\in GL(r_+|r_-)$ is a certain invertible
1252:   operator\footnote{This should not be confused with the homotopy flow of
1253:   Subsection \ref{bulk_flow} !}) which plays
1254: the role of `parallel transport' defined by $M$ along $C_a$ (see
1255: \cite{coupling} for details). Namely $U_a(\tau_a)=U_a(\tau_a, 0)$, where:  
1256: \be
1257: \label{parr_tr}
1258: U_a(\tau_2,\tau_1):=Pe^{-\int_{\tau_1}^{\tau_2} M(\tau)d\tau}~~
1259: \ee
1260: if $\tau_2>\tau_1$. The origin of the proper length coordinate $\tau_a$ 
1261: along $C_a$ is chosen arbitrarily, while the orientation on $C_a$ is
1262: compatible with that of $\Sigma$. 
1263: The quantities   $F_{{\bar i}{\bar j}}$ etc. are the 
1264: $(0,2)$-components of the curvature of the direct sum connection $A$ introduced in (\ref{AD}).
1265: Notice the relations: 
1266: \be
1267: {\cal U}_a=\Str H_a(\tau)~~
1268: \ee
1269: where: 
1270: \be
1271: \label{holonomy}
1272: H_a(\tau)=U(\tau+l_a,\tau)
1273: \ee 
1274: are the "superholonomy operators" (here $l_a$ the length of $C_a$).
1275: 
1276: 
1277: Relations (\ref{invar}, \ref{invar_0}) and (\ref{Qhol},\ref{delta_0hol}) 
1278: show that the BRST and $\delta_0$-invariance conditions amount to:
1279: \bea
1280: \label{invariance_eqs}
1281: &&F_{{\bar i}{\bar j}}=0\\
1282: &&\nabla_{\bar i} D=0\\
1283: &&\nabla_i(D^2)=\partial_iW\\
1284: &&[D^\dagger,D^2]=0~~.
1285: \eea
1286: The first relation says that $A$ is integrable, so it defines a complex
1287: structure on the bundle $E$. The second condition means that $D\in End(E)$ is
1288: holomorphic with respect to this complex structure. The third equation 
1289: requires $D^2=c+W {\rm id}_E$, with $c$ a covariantly-constant endomorphism. Comparing with
1290: (\ref{02curvature}), we find that these first three conditions are equivalent with:
1291: \bea
1292: \label{eom}
1293: {\cal F}^{(0,\leq 2)}=c+W{\rm id}_E\Longleftrightarrow {\cal {\bar D}}^2=c+W{\rm id}_E~~.
1294: \eea
1295: This is the target space equation of motion for our open string background
1296: \cite{coupling}. Notice that (\ref{eom}) admit solutions only when
1297: $r_+=r_-$.
1298: 
1299: 
1300: For backgrounds satisfying the equation of motion, 
1301: the last condition in (\ref{invariance_eqs}) reads:
1302: \be
1303: [D^\dagger,D^2]=0 \Longleftrightarrow [D^\dagger,c]=0~~. 
1304: \ee
1305: This can be viewed as a partial "gauge-fixing" constraint,
1306: which is fulfilled, for example, if one takes $c$ to be proportional to the
1307: identity endomorphism (in which case the proportionality constant can be absorbed into $W$). 
1308: For simplicity, we shall take $c=0$ for the remainder of this paper. 
1309: 
1310: \section{Boundary observables and correlators}
1311: 
1312: As we  shall see in Section \ref{loc_boundary}, 
1313: the boundary conditions derived from the partition function (\ref{Z}) constrain $\theta$ in terms
1314: of $\eta$ along the boundary of the worldsheet. Hence it suffices to consider
1315: boundary observables of the form:
1316: \be
1317: \label{boundary_observable}
1318: {\cal O}_\alpha(\tau)=\alpha_{{\bar i}_1\dots {\bar i}_p}(\phi(\tau))\eta^{{\bar
1319:     i}_1}(\tau)\dots \eta^{{\bar i}_p}(\tau)~~,
1320: \ee
1321: where $\tau$ is a point on $\partial \Sigma$. Here $\alpha:=\alpha_{{\bar
1322:   i}_1\dots {\bar i}_p}dz^{{\bar i}_1}\wedge \dots \wedge dz^{{\bar i}_p}$ is a $(0,p)$ form
1323: valued in $End(E)$.
1324: 
1325: 
1326: Consider a collection of $m$ topological D-branes described by superbundles
1327: $E_a$ endowed with superconnections ${\cal B}_a$, such that the
1328: target space equations of motion are satisfied. The index $a$ runs from
1329: $1$ to $m$. Let $\Sigma$ be a Riemann
1330: surface with $m$ circle boundary components $C_a$, which we
1331: endow with the orientation induced from $\Sigma$.
1332: Choosing forms
1333: $\alpha^{(a)}_j\in \Omega^{(0,p^{(a)}_j)}(End(E_a))$ and points
1334: $\tau^{(a)}_1\dots \tau^{(a)}_{k_a}$ arranged in increasing cyclic order 
1335: along $C_a$, we are interested in the correlator:
1336: \bea
1337: \label{boundary_corr}
1338: &&\langle \prod_{a=1}^{m}\prod_{j_a=k_a}^{1}
1339: {\cal O}_{\alpha_{j_a}^{(a)}}(\tau^{(a)}_{j_a}) \rangle_\Sigma:=
1340: \int{{\cal D}[\phi]{\cal  D}[F]{\cal D}[\rho]{\cal D}[\eta]{\cal D}[\theta]}
1341: e^{-{\tilde  S}_{bulk}}\\
1342: &&\prod_{a=1}^{m}{\Str\left[
1343: {\cal O}_{\alpha^{(a)}_{k_a}}(\tau_{k_a}^{(a)})
1344: U_a(\tau^{(a)}_{k_a},\tau^{(a)}_{k_a-1})
1345: {\cal O}_{\alpha^{(a)}_{k_a-1}}(\tau^{(a)}_{k_a-1})\dots
1346: {\cal O}_{\alpha^{(a)}_1}(\tau_1^{(a)})
1347: U_a(\tau^{(a)}_1,\tau^{(a)}_{k_a})\right]}~~~,\nn
1348: \eea
1349: where we used the "parallel supertransport" operators defined in
1350: (\ref{parr_tr}). The integration domain in  (\ref{boundary_corr}) is
1351:   specified by the appropriate boundary conditions on the worldsheet fields,
1352:   which will be discussed in more detail below. 
1353: 
1354: Let us first consider a single operator insertion 
1355: ${\cal O}_\alpha$ along a circle boundary component $C$. In this case, the
1356: relevant factor in (\ref{boundary_corr}) is:
1357: \be
1358: \Str[H(\tau){\cal O}_\alpha(\tau)]~~.
1359: \ee
1360: We wish to compute the BRST variation of this quantity. From the relation:
1361: \be
1362: \label{deltaH}
1363: \delta H(\tau)=
1364: \left[\frac{1}{2}\int_{C}{\rho^i\partial_i W}\right] H(\tau)+[H(\tau), D(\tau)+A_{\bar
1365: i}(\tau)\eta^{\bar i}(\tau)]~~
1366: \ee
1367: we obtain:
1368: \be
1369: \label{deltaStr}
1370: \delta \Str[H(\tau){\cal O}_\alpha(\tau)]=
1371: \left[\frac{1}{2}\int_{C}{\rho^i\partial_i W}\right]\Str[H(\tau){\cal
1372:   O}_\alpha(\tau)]+\Str[H(\tau)\delta_b {\cal
1373:   O}_\alpha (\tau)]~~,
1374: \ee
1375: where:
1376: \be
1377: \label{delta_b}
1378: \delta_b{\cal O}_\alpha:=\delta {\cal O}_\alpha+[D+A_{\bar
1379: i}\eta^{\bar i}, {\cal O}_\alpha]~~.
1380: \ee
1381: Using the target space equations of motion, one easily checks that
1382: \footnote{Indeed, one has:
1383: \be
1384: \delta_b^2{\cal O}=\frac{1}{2}[F_{{\bar i}{\bar j}},{\cal O}]+\eta^{\bar
1385:   i}[\nabla_{\bar i}D,{\cal O}]+[D^2,{\cal O}]~~.
1386: \ee}
1387: $\delta_b$ squares to zero, so that it plays the role of an `effective' BRST operator
1388: in the boundary sector. Notice that $\delta_b$ arises naturally due to the
1389: second term in the BRST variation (\ref{deltaH}) of $H(\tau)$. 
1390: Using (\ref{delta_b}), we find the relation:
1391: \be
1392: \label{delta_b_loc}
1393: \delta_b{\cal O}_\alpha={\cal O}_{{\bar D}\alpha}
1394: \ee
1395: where ${\bar D}={\bar D}_{\cal B}$ is the Dolbeault operator on $\Omega^{(0,*)}(End(E))$
1396: twisted by the superconnection ${\cal B}$. 
1397: 
1398: 
1399: It is not hard to generalize (\ref{deltaStr}) to the case of $k$ 
1400: insertions along $C$:
1401: \bea
1402: \label{delta_Str_2}
1403: &&\delta \Str[{\cal O}_{\alpha_k}(\tau_k) U(\tau_k, \tau_{k-1})
1404: \dots {\cal O}_{\alpha_1}(\tau_1)U(\tau_1,\tau_k)]=\\
1405: &&~~~=\left[\frac{1}{2}\int_{C}{\rho^i\partial_i W}\right]
1406: \Str[{\cal O}_{\alpha_k}(\tau_k) U(\tau_k, \tau_{k-1})
1407: \dots {\cal O}_{\alpha_1}(\tau_1)U(\tau_1,\tau_k)]+\nn\\
1408: && ~~~~~~~+\sum_{j=1}^{k}{\Str[
1409: {\cal O}_{\alpha_k}(\tau_k)\dots 
1410: U(\tau_{j+1},\tau_j)\delta_b{\cal O}_{\alpha_j}(\tau_j)U(\tau_j,\tau_{j-1})
1411: \dots {\cal O}_{\alpha_1}(\tau_1)U(\tau_1,\tau_k) ]}~~.\nn
1412: \eea
1413: Applying this to (\ref{boundary_corr}), we find that the BRST variation of
1414: $e^{-{\tilde S}_{bulk}}$ is canceled by the first contribution in
1415: (\ref{delta_Str_2}), summed over circle boundary components. This gives:
1416: \bea
1417: \label{deltaStr_final}
1418: && \delta \left(e^{-{\tilde  S}_{bulk}}\prod_{a=1}^{m}{\Str\left[
1419: {\cal O}_{\alpha^{(a)}_{k_a}}(\tau_{k_a}^{(a)})
1420: U(\tau^{(a)}_{k_a},\tau^{(a)}_{k_a-1})
1421: \dots {\cal O}_{\alpha^{(a)}_1}(\tau_1^{(a)})
1422: U(\tau^{(a)}_1,\tau^{(a)}_{k_a})\right]}\right)=~~~\\
1423: &=&e^{-{\tilde S}_{bulk}}\sum_{a=1}^m\sum_{j_a=1}^{k_a}
1424: \Str\left[{\cal O}_{\alpha^{(a)}_{k_a}}(\tau_{k_a}^{(a)})
1425: U(\tau^{(a)}_{k_a},\tau^{(a)}_{k_a-1})
1426: \dots \delta_b {\cal O}_{\alpha^{(a)}_{j_a}}(\tau_{j_a}^{(a)})
1427: \dots {\cal O}_{\alpha^{(a)}_1}(\tau_1^{(a)})
1428: U(\tau^{(a)}_1,\tau^{(a)}_{k_a})\right]~~.\nn
1429: \eea
1430: Equation (\ref{deltaStr_final}) replaces the more familiar formula known from the open
1431: topological sigma model. Unlike the sigma model case, the left hand side 
1432: includes the factor $e^{-{\tilde S}_{bulk}}$, because its BRST variation does
1433: not vanish separately. Equation (\ref{deltaStr_final}) implies that the 
1434: correlator of $\delta_b$-closed boundary observables only depends on their
1435: $\delta_b$-cohomology class, and in particular such a correlator vanishes if
1436: one of the boundary observables is $\delta_b$-exact. Remember that ${\tilde S}_{bulk}={\tilde
1437:   S}_B+S_W$.
1438: BRST closure of ${\tilde S}_B$ implies that (\ref{deltaStr_final}) is equivalent with:
1439: \bea
1440: \label{deltaStr_modified}
1441: && \delta \left(e^{-S_W}\prod_{a=1}^{m}{\Str\left[
1442: {\cal O}_{\alpha^{(a)}_{k_a}}(\tau_{k_a}^{(a)})
1443: U(\tau^{(a)}_{k_a},\tau^{(a)}_{k_a-1})
1444: \dots {\cal O}_{\alpha^{(a)}_1}(\tau_1^{(a)})
1445: U(\tau^{(a)}_1,\tau^{(a)}_{k_a})\right]}\right)=~~~\\
1446: &=&e^{-S_W}\sum_{a=1}^m\sum_{j_a=1}^{k_a}
1447: \Str\left[{\cal O}_{\alpha^{(a)}_{k_a}}(\tau_{k_a}^{(a)})
1448: U(\tau^{(a)}_{k_a},\tau^{(a)}_{k_a-1})
1449: \dots \delta_b {\cal O}_{\alpha^{(a)}_{j_a}}(\tau_{j_a}^{(a)})
1450: \dots {\cal O}_{\alpha^{(a)}_1}(\tau_1^{(a)})
1451: U(\tau^{(a)}_1,\tau^{(a)}_{k_a})\right]~,\nn
1452: \eea
1453: a fact which will be used in Section \ref{loc_boundary}.
1454: 
1455: 
1456: 
1457: \paragraph{Observation} It is easy to extend the discussion above by including
1458: boundary condition changing observables, which in the present context have the
1459: form (\ref{boundary_observable}), but with $\alpha$ an element of 
1460: $\Omega^{(0,p)}(Hom(E_a,E_b))$. In this case, the operator 
1461: (\ref{delta_b}) is replaced by:
1462: \be
1463: \delta_b{\cal O}_\alpha:=\delta {\cal O}_\alpha+
1464: (D^{(b)}+A^{(b)}_{\bar i}\eta^{\bar i}){\cal O}_\alpha
1465: -(-1)^{\rk\alpha}{\cal O}_\alpha (D^{(a)}+A^{(a)}_{\bar i}\eta^{\bar i})
1466: \ee
1467: and ${\bar D}$ in relation (\ref{delta_b_loc}) becomes the Dolbeault operator
1468: on $\Omega^{(0,*)}(Hom(E_a,E_b))$, twisted by the superconnections 
1469: ${\cal B}_a$ and ${\cal B}_b$.
1470: 
1471: 
1472: \section{Localization formula for boundary correlators on the disk}
1473: \label{loc_boundary}
1474: 
1475: 
1476: We next discuss localization in the boundary sector. As for the bulk, we will
1477: proceed by localizing on sigma model zero-modes, thereby extracting a
1478: two-parameter family of localization formulae. The first index of this 
1479: family is the bulk parameter $\lambda$ of Section 3, while second parameter
1480: $\mu$ is associated with the last term in (\ref{M_dec}). These two parameters 
1481: measure the area and circumference length of a worldsheet with disk topology.  Each pair
1482: $(\lambda,\mu)$ defines a localization picture, and a certain
1483: off-shell representation of the boundary trace of \cite{Moore_Segal, Moore,
1484:   CIL1}. As we shall see below, 
1485: the various pictures are again related by a homotopy flow, and in particular
1486: the various representations of the boundary trace agree when reduced to the cohomology of
1487: $\delta_b$. In this approach, the appropriate 
1488: generalization of the residue representation of \cite{Kap2}
1489: is recovered in the limit $\lambda\rightarrow
1490: +\infty$ with $\mu=0$. 
1491: 
1492: The boundary conditions induced by the coupling (\ref{Z}) can be
1493: extracted by studying the Euler-Lagrange variations of the non-local action 
1494: $S_{eff}={\tilde S}_{bulk}-\ln {\cal U}$. These conditions are given
1495: explicitly in Appendix \ref{BCS}, where we also show that they are BRST
1496: invariant modulo the equations of motion for the auxiliary fields $F$. 
1497: The disk correlator of a collection of boundary observables ${\cal
1498:   O}_{\alpha_1}(\tau_1)\dots {\cal O}_{\alpha_k}(\tau_k)$ (with $\tau_1\dots
1499: \tau_k$ arranged in increasing cyclic order along the boundary) 
1500: is obtained by performing the relevant path integral while imposing 
1501: the  boundary conditions, which  cut out a subset ${\cal C}$ in field
1502: configuration space:
1503: \bea
1504: \label{disk_correlator}
1505: &&\langle {\cal O}_{\alpha_k}(\tau_k)\dots {\cal O}_{\alpha_1}(\tau_1)\rangle_{\rm disk}=\\
1506: &&~~~=\int_{\cal C}{{\cal D}[\phi]{\cal D}[{\tilde F}]
1507: {\cal D}[\eta]{\cal D}[\theta]{\cal D}[\rho]~e^{-{\tilde
1508:     S}_{bulk}}\Str[{\cal O}_{\alpha_k}(\tau_k) U(\tau_k, \tau_{k-1})
1509: \dots {\cal O}_{\alpha_1}(\tau_1)U(\tau_1,\tau_k)]}~~.\nn
1510: \eea
1511: In this section, we assume that ${\cal O}_{\alpha_j}$ are $\delta_b$-closed:
1512: \be
1513: \label{alpha_closure}
1514: \delta_b{\cal O}_{\alpha_j}=0\Longleftrightarrow {\bar D}_{\cal B}\alpha_j=0~~.
1515: \ee
1516: 
1517: \subsection{Localization on $B$-model zero-modes}
1518: 
1519: Remember that ${\tilde S}_{bulk}={\tilde S}_B+S_W$, where ${\tilde S}_B=\delta
1520: V_B$ is BRST exact. As in Section \ref{loc_bulk}, this allows us to replace 
1521: ${\tilde S}_{bulk}$ by $t{\tilde S}_B+S_W$, without changing the value of the
1522: correlator (\ref{disk_correlator}). Here $t$ is a complex variable with
1523: positive real part. Invariance of (\ref{disk_correlator}) under changes in $t$ 
1524: follows by differentiation with respect to this parameter upon using 
1525: $\delta_b$-closure of ${\cal O}_{\alpha_j}$ and equation (\ref{deltaStr_modified})
1526: \footnote{The path integral over ${\cal C}$ of the BRST exact term involved in
1527:   this argument vanishes
1528: because the boundary conditions determining ${\cal C}$ are preserved by the BRST transformations 
1529: up to terms which vanish by the equations of motion for ${\tilde F}$ (see Appendix \ref{BCS}).}. 
1530: 
1531: We can now take the limit $\Re t\rightarrow +\infty$
1532: to localize on the zero modes of ${\tilde S}_B$. This gives:
1533: \be
1534: \label{disk_correlator_loc}
1535: \langle {\cal O}_{\alpha_k}\dots {\cal
1536:   O}_{\alpha_1}\rangle_{\rm disk}=\langle {\cal O}_\alpha \rangle_{\rm
1537:   disk}=N\int_{{\cal C}_0}{d\phi d\eta d\theta
1538: e^{-\frac{A}{4}{\tilde L}_0}\Str[H_0{\cal O}_\alpha]}~~.
1539: \ee
1540: To arrive at this formula, 
1541: we noticed that the dependence of $\tau_j$ disappears on zero-modes, we 
1542: set $\alpha:=\alpha_k\wedge \dots \wedge \alpha_1$ and 
1543: integrated out the auxiliary fields ${\tilde F}$. 
1544: Also notice that $\delta_b {\cal O}_\alpha=0$ due to relations (\ref{alpha_closure}).
1545: The symbol $H_0$ denotes the
1546: restriction of the superholonomy factor $H$ to zero-modes:
1547: \be
1548: H_0=e^{-\frac{il}{2}k_0}~~,
1549: \ee
1550: where $l$ is the length of the disk's boundary and:
1551: \be
1552: \label{k0}
1553: k_0:=\eta^{\bar i}\nabla_{\bar i}D^\dagger +[D,D^\dagger]=\delta_b D^\dagger~~.
1554: \ee
1555: The symbol ${\cal C}_0$ denotes the subset of the space of zero modes cut out
1556: by the boundary conditions. Since we integrated out the auxiliary fields, this 
1557: subset is strictly BRST invariant:
1558: \be
1559: \delta {\cal C}_0\subset {\cal C}_0~~.
1560: \ee
1561: Using this property as well as $\delta$-exactness (\ref{exact_tildeL_0}) of ${\tilde L}_0$
1562: and $\delta_b$-exactness (\ref{k0}) of $k_0$, one checks
1563: \footnote{The proof requires the identity $\delta \Str B=\Str\delta_b B$ for any
1564:   quantity $B$ built out of zero modes. This holds because the supertrace of any
1565:   supercommutator vanishes.}  that
1566: (\ref{disk_correlator_loc}) is insensitive to rescalings of these
1567: quantities, and hence can be replaced with:
1568: \be
1569: \label{disk_correlator_lambda_mu}
1570: \langle {\cal O}_\alpha\rangle_{\rm disk}=N\int_{{\cal C}_0}{d\phi d\eta d\theta
1571: e^{-\lambda{\tilde L}_0}\Str[e^{-\mu k_0}{\cal O}_\alpha]}~~,
1572: \ee
1573: where $\lambda$ and $\mu$ are complex numbers such that $\Re \lambda>0$. The
1574: quantity (\ref{disk_correlator_lambda_mu}) is independent of the values of
1575: these two parameters.
1576: 
1577: To make (\ref{disk_correlator_lambda_mu}) explicit, we must describe the
1578: restriction to ${\cal C}_0$. The relevant boundary condition
1579: takes the form (see Appendix \ref{BCS}):
1580: \be
1581: \label{eta_theta}
1582: i(G_{i{\bar j}}\eta^{\bar j}+\theta_i){\cal U}=\Str[H(\tau)(\partial_i
1583: D+F_{i{\bar j}}\eta^{\bar j})]\Longleftrightarrow \Str[H(\tau)(\theta_i+iV_i)]=0~~,
1584: \ee
1585: where:
1586: \be
1587: V_i:=\partial_iD+(F_{i{\bar j}}-iG_{i{\bar j}})\eta^{\bar j}~~.
1588: \ee
1589: Equation (\ref{eta_theta}) instructs us to replace $\theta_i$ by $-iV_i$ under
1590: the supertrace in order to produce the desired restriction. 
1591: To implement these constraints, we shall use the quantity:
1592: \be
1593: \label{Pi}
1594: \Pi:=\frac{1}{n!}\epsilon^{i_1\dots i_n} (\theta_{i_1}
1595: 1_{End(E)}+iV_{i_1})\dots 
1596: (\theta_{i_n} 1_{End(E)}+iV_{i_n})~~.
1597: \ee
1598: Consider an $End(E)$-valued function $f$ of $\theta_i$: 
1599: \be
1600: f(\theta_1\dots \theta_n)=\sum_{p=0}^n\sum_{1\leq i_1<\dots <i_p\leq
1601:   n}{\theta_{i_1}\dots \theta_{i_p}f^{i_1\dots
1602:     i_p}}=\sum_{p=0}^n\frac{1}{p!}\theta_{i_1}\dots \theta_{i_p}f^{i_1\dots i_p}~~,
1603: \ee
1604: where $f_{i_1\dots i_p}\in End(E)$ with $f^{i_{\sigma(1)}\dots
1605:   i_{\sigma(p)}}=\epsilon(\sigma)f^{i_1\dots i_p}$ for all $\sigma\in
1606: \Sigma_p$ and in the last equality we use implicit summation over $i_1\dots
1607: i_p=1\dots n$. Here $\Sigma_p$ is the group of permutations on $p$ elements, while
1608: $\epsilon(\sigma)$ is the signature of the permutation $\sigma$. Then one
1609: checks the identity:
1610: \be
1611: \label{f_identity}
1612: \int{d\theta_1\dots d\theta_n 
1613: \Pi f(\theta_1\dots \theta_n)}=f(-iV_1\dots -iV_n)~~,
1614: \ee
1615: where the right hand side is defined by:
1616: \bea
1617: f(-iV_1\dots -iV_n)&:=&\sum_{p=0}^n\sum_{1\leq i_1<\dots <i_p\leq
1618:   n}{\frac{1}{p!}\sum_{\sigma\in \Sigma_p}
1619: \epsilon(\sigma)(-iV_{i_{\sigma(1)}})\dots (-iV_{i_{\sigma(p)}}) f^{i_1\dots
1620:   i_p}}\nn\\
1621: &=&\sum_{p=0}^n{\frac{1}{p!}
1622: (-iV_{i_1})\dots (-iV_{i_p}) f^{i_1\dots
1623:   i_p}}~~.
1624: \eea
1625: For later reference, we note the case $f=e^{q^i\theta_i}$, with $q^i$ some
1626: Grassmann-odd quantities depending on $\phi$ and $\eta$. 
1627: Then we have  $f_{i_1\dots i_p}=(-1)^{p(p+1)/2}q^{i_1}\dots q^{i_p}$
1628: and find $f(-iV_1\dots -iV_n)=e^{-iq^i V_i}$, using the fact that
1629: $q_j$ are mutually anti-commuting. This gives:
1630: \be
1631: \label{exp_identity}
1632: \int{d\theta_1\dots d\theta_n e^{q^i\theta_i}\Pi }=\int{d\theta_1\dots d\theta_n 
1633: \Pi e^{q^i\theta_i}}=e^{-iq^iV_i}~~.
1634: \ee
1635: 
1636: 
1637: Relation (\ref{f_identity}) shows that $\Pi$ is a sort of  `Poincare dual' of
1638: ${\cal C}_0$ on the supermanifold of
1639: field configurations. Using (\ref{exp_identity}), this observation allows us to write
1640: (\ref{disk_correlator_lambda_mu}) as an unconstrained integral over the space
1641: of sphere zero-modes :
1642: \be
1643: \label{disk_correlator_loc_bulk}
1644: \langle {\cal O}_\alpha\rangle_{\rm disk}=N\int{d\phi d\eta d\theta
1645: e^{-\lambda{\tilde L}_0}\Str[e^{-\mu k_0} 
1646: \Pi {\cal O}_\alpha ]}~~.
1647: \ee
1648: Employing equation (\ref{exp_identity}), we find:
1649: \be
1650: \label{disk_correlator_loc_boundary}
1651: \langle {\cal O}_\alpha\rangle_{\rm disk}=N\int{d\phi d\eta
1652: \Str[e^{-\mu k_0}e^{-\lambda{\tilde L}_0^b}{\cal O}_\alpha]}~~,
1653: \ee
1654: where:
1655: \bea
1656: \label{tildeL_0b}
1657: {\tilde L}_0^b:={\tilde L}_0|_{\theta_i\rightarrow -iV_i}&=&-D_{\bar i}\partial_{\bar
1658:   j}{\bar W}V^{\bar i}\eta^{\bar j}+G^{i{\bar j}}(\partial_iW)(\partial_{\bar
1659:   j}{\bar W})=\\
1660: &=&{\bar H}^i_{~\bar j}\eta^{\bar j}\partial_iD+\eta^{\bar
1661:   i}\eta^{\bar j}{\bar H}^k_{~{\bar i}}F_{k{\bar j}}+G^{i{\bar j}}(\partial_iW)(\partial_{\bar
1662:   j}{\bar W})~~.\nn
1663: \eea
1664: It is easy to check the relation:
1665: \be
1666: \delta_b V_i=\partial_i W~~,
1667: \ee
1668: which shows that the on-shell BRST variation (\ref{delta_theta_os}) of 
1669: $\theta_i$ agrees with the $\delta_b$-variation of $-iV_i$:
1670: \be
1671: \label{delta_delta_b}
1672: \delta_b(\theta_i 1_{End(E)}+iV_i)=0\Longleftrightarrow (\delta
1673: \theta_i)1_{End(E)}=-i\delta_b V_i~~
1674: \ee
1675: (this in particular implies that $\delta \Pi=0$). Using this equation and
1676: relation (\ref{exact_tildeL_0}), one finds that ${\tilde
1677:   L}_0^b$ is $\delta_b$-exact:
1678: \be
1679: \label{tildeL_0_b_ex}
1680: {\tilde L}_0^b=\delta_b {\tilde v}_0^b~~,
1681: \ee
1682: where:
1683: \be
1684: {\tilde v}_0^b={\tilde v}_0|_{\theta_i\rightarrow -iV_i}=V^{\bar i}\partial_{\bar i}{\bar W}~~.
1685: \ee
1686: Together with $\delta_b$-exactness of $k_0$, this can be used to give a direct
1687: proof of independence of (\ref{disk_correlator_loc_boundary}) of $\lambda$ and $\mu$.
1688: 
1689: We end by mentioning some useful properties of $V_i$. It is
1690: easy to compute the anticommutator:
1691: \be
1692: \label{V_commutator}
1693: [V_i,V_j]_{-}=\partial_i\partial_jW-\delta_b[\partial_i\partial_j (D+A_{\bar k}\eta^{\bar k})]~~.
1694: \ee
1695: Moreover, it is not hard to check the identity:
1696: \be
1697: \label{VO}
1698: [V_i, {\cal O}]=\partial_i(\delta_b{\cal O})-\delta_b(\partial_i{\cal O})
1699: \ee
1700: for any boundary observable ${\cal O}$ (as usual,  the quantity on the left hand
1701: side is a supercommutator). In particular, a $\delta_b$-closed
1702: boundary observable supercommutes with $V_i$ up to a $\delta_b$-exact term. 
1703: 
1704: \subsection{The space of boundary observables}
1705: \label{boundary_geom}
1706: 
1707: After reduction to zero-modes, each boundary insertion ${\cal O}_\alpha$ can
1708: be identified with the superbundle-valued differential form $\alpha$. This
1709: amounts to setting
1710: $\eta^{\bar i}\equiv dz^{\bar i}$, so the superalgebra ${\cal
1711:   H}_b:=\Omega^{(0,*)}(End(E))$ provides an off-shell model for the space of boundary
1712: excitations. Moreover, equation (\ref{delta_b_loc})
1713: identifies the boundary BRST operator $\delta_b$ with the operator 
1714: ${\bar D}_{\cal B}={\bar \nabla}_A+{\cal D}$ acting on ${\cal H}_b$,  
1715: where ${\bar \nabla}_A$ acts in the adjoint representation and 
1716: ${\cal D}=[D,\cdot]$. The
1717: target space equations of motion imply that $\delta_b$ squares to zero. Thus 
1718: ${\cal H}_b$ is a differential superalgebra.  To be precise, 
1719: we take ${\cal H}_b$ to consist of bundle-valued differential forms with at
1720: most polynomial growth at infinity. 
1721:  
1722: The target space equations of motion imply the relations:
1723: \be
1724: {\bar \partial}_A^2={\cal D}^2={\bar \partial}_A\circ {\cal D}+{\cal D}\circ
1725: {\bar \partial}_A=0~~,
1726: \ee
1727: which show that ${\cal H}_b$ is a bicomplex. Thus the boundary BRST cohomology is 
1728: computed by a spectral sequence $E^b_*$ whose second term has the form:
1729: \be
1730: \label{Eb}
1731: E^b_2=H_{\cal D}(H_{{\bar \partial}_A}({\cal H}_b))~~.
1732: \ee
1733: 
1734: 
1735: In the simple case $X=\C^n$, the holomorphic bundle  $E$ is the trivial
1736: superbundle of type $(r_+,r_-)$ and the spectral 
1737: sequence collapses to its second term. Then the BRST cohomology 
1738: coincides with the cohomology of ${\cal D}$ taken in the space of square matrices of dimension
1739: $r_++r_-$ whose entries are polynomial functions of $n$ complex variables.
1740: This recovers the result of \cite{Kap1}.
1741: 
1742: 
1743: 
1744: \subsection{The boundary-bulk and bulk-boundary maps}
1745: \label{bulk_bdry}
1746: 
1747: 
1748: The equivalent expressions (\ref{disk_correlator_loc_bulk}) and 
1749: (\ref{disk_correlator_loc_boundary}) allow us to extract an off-shell
1750: version of the boundary-bulk map of \cite{CIL1}:
1751: \be
1752: f_\mu({\cal O}_\alpha)=\Str[e^{-\mu k_0}\Pi {\cal O}_\alpha]~~.
1753: \ee
1754: This maps ${\cal H}_b$ to ${\cal H}$ and obeys:
1755: \be
1756: \langle {\cal O}_\alpha\rangle_{\rm disk}=\langle f_\mu({\cal
1757:   O}_{\alpha })\rangle_{\rm sphere}\Longleftrightarrow \Tr_b \alpha =\Tr
1758: f_\mu(\alpha)~~,
1759: \ee
1760: where we identified $\alpha$ with ${\cal O}_\alpha$. Here 
1761: $\Tr$ and $\Tr_b$ are the bulk and boundary traces determined by the 
1762: localization formulae (\ref{correlator_final}) and (\ref{disk_correlator_loc_boundary}):
1763: \bea
1764: \Tr\omega~&=&\langle {\cal O}_\omega\rangle_{\rm sphere}~=\int{d\phi d\eta d\theta e^{-\lambda {\tilde L}_0}{\cal
1765:     O}_\omega}\label{Tr}\\
1766: \Tr_b\alpha &=&\langle  {\cal O}_\alpha\rangle_{\rm disk}=
1767: \int{d\phi d\eta \Str[e^{-\mu k_0}e^{-\lambda {\tilde L}^b_0}{\cal
1768:     O}_\alpha]}~~.\label{Tr_o}
1769: \eea
1770: As in \cite{CIL1}, we can also define a bulk-boundary map $e$ through the adjunction formula:
1771: \be
1772: \label{adjunction}
1773: \Tr({\cal O}_\omega f_\mu({\cal O}_\alpha))=\Tr_b(e({\cal O}_\omega) {\cal O}_\alpha)~~.
1774: \ee
1775: From the relations above, we find:
1776: \be
1777: e({\cal O}_\omega):=e^{\lambda {\tilde L}_0^b}\int{d\theta_1\dots d\theta_n
1778:   e^{-\lambda {\tilde L}_0}{\cal O}_\omega \Pi}~~.
1779: \ee
1780: This maps ${\cal H}$ to ${\cal H}_b$. 
1781: 
1782: Using (\ref{delta_delta_b}), we find that $f_\mu$ and $e$ are
1783: compatible with the bulk and boundary BRST operators:
1784: \bea
1785: \delta\circ f_\mu &=& (-1)^n f_\mu\circ \delta_b ~~\label{delta_f}\\
1786: \delta_b \circ e ~&=& e \circ \delta~~.\label{delta_e}
1787: \eea
1788: To prove the second equation, we used the identity:
1789: \be
1790: \delta\int{d\theta_1\dots d\theta_n f(\theta_1\dots
1791:   \theta_n)}=\int{d\theta_1\dots d\theta_n \delta f(\theta_1\dots \theta_n)}~~,
1792: \ee
1793: which follows from (\ref{delta_theta_os}). 
1794: Relations (\ref{delta_f}) and (\ref{delta_e}) show that 
1795: $e$ and $f_\mu$ descend to well-defined maps $e_*$ and $f_*$ between the bulk and
1796: boundary BRST cohomologies (the latter are the maps considered in \cite{CIL1}). 
1797: Since $k_0$ is $\delta_b$-exact, one easily checks that $f_*$ is independent of $\mu$.
1798: 
1799: \subsection{A geometric model for the boundary trace}
1800: 
1801: As in Section \ref{loc_bulk}, we can use the identifications $\eta^{\bar
1802:   i}\equiv dz^{\bar i}$ and $\theta_i\equiv \partial_i$ to represent our formulae in
1803:   terms of standard geometric objects. We find:
1804: \be
1805: {\tilde L}_0^b\equiv{\bar H}\lrcorner(\partial D+F)+||\partial W||^2=\delta_b
1806:   {\tilde v}_0^b~~
1807: \ee
1808: with:
1809: \be
1810: {\tilde v}_0^b={\rm grad} {\bar W}\lrcorner ~(\partial D+F)~~
1811: \ee
1812: and:
1813: \be
1814: k_0={\bar \nabla}_{\cal B}D^\dagger={\bar \nabla}_A D^\dagger+[D,D^\dagger]~~.
1815: \ee
1816: The disk localization formula (\ref{disk_correlator_loc_boundary}) becomes:
1817: \be
1818: \label{Tr_bdry_geom}
1819: \Tr_b \alpha =N\int_{X}{\Omega\wedge \str[e^{-\lambda {\tilde L}_0^b}\wedge e^{-\mu
1820:   k_0} \wedge \alpha]}~~, 
1821: \ee
1822: while the quantity (\ref{Pi}) takes the form:
1823: \be
1824: \Pi=\frac{1}{n!}(\partial_1+iV_1)\wedge \dots \wedge (\partial_n+iV_n)~~,
1825: \ee
1826: with:
1827: \be
1828: V_i=\partial_i D+(F_{i{\bar j}}-iG_{i{\bar j}})dz^{\bar j}~~.
1829: \ee
1830: Defining $V=dz^i \otimes V_i$, we obtain:
1831: \be
1832: V=\partial D+ F-iG~~,
1833: \ee
1834: where $G:=G_{i{\bar j}}dz^{i}\otimes dz^{\bar j}$. Notice that here and above,
1835:   $\partial D$ is defined by $\partial D=dz^i\otimes \partial_i D$ (the order
1836:   matters since $D$ is odd). 
1837: 
1838: 
1839: \subsection{Boundary localization pictures and the homotopy flow}
1840: 
1841: As for the bulk sector, one can define a two-parameter semigroup of operators
1842: acting on ${\cal H}_b$ through:
1843: \be
1844: U_b(\lambda, \mu)(\alpha):=e^{-\lambda {\tilde L}_0^b}\wedge e^{-\mu
1845:   k_0} \wedge \alpha~~. 
1846: \ee
1847: The pair $(\lambda, \mu)$ is taken inside the domain: 
1848: \be
1849: \Delta_b:=\{(\lambda,\mu)\in \C^2 | \Re \lambda>0 \}~~.
1850: \ee
1851: Since both ${\tilde L}_0$ and ${\tilde L}_0^b$ are BRST-exact, each
1852: $U_b(\lambda,\mu)$ is homotopy-equivalent with the identity so
1853: this defines a homotopy flow. We let: 
1854: \be
1855: \Tr_b^B(\alpha):=N\int_{X}{\Omega\wedge \str(\alpha)}
1856: \ee
1857: denote the boundary trace of the B-twisted sigma model (viewed as a linear
1858: functional on the off-shell state space ${\cal H}_b$). Then equation (\ref{Tr_bdry_geom})
1859: becomes: 
1860: \be
1861: \Tr_b(\alpha)=\Tr_b^B(U_b(\lambda,\mu)(\alpha))~~. 
1862: \ee
1863: Again one can define localization pictures indexed by $\lambda$ and
1864: $\mu$. The boundary BRST operator satisfies: 
1865: \be
1866: U_b(\lambda,\mu)\circ\delta_b=\delta_b \circ U_b(\lambda, \mu)~~. 
1867: \ee
1868: 
1869: 
1870: 
1871: \subsection{Residue formula for boundary correlators on the disk}
1872: 
1873: As in Section \ref{loc_bulk}, we can use equation (\ref{disk_correlator_loc_bulk})
1874: to express boundary correlators in terms of generalized residues. For
1875: simplicity, we shall assume that the spectral sequence of Subsection \ref{boundary_geom}
1876: collapses to its second term. Setting $\mu=0$ in (\ref{disk_correlator_loc_bulk}) gives:
1877: \be
1878: \label{step1}
1879: \langle {\cal O}_\alpha\rangle_{\rm disk}=N\int{d\phi d\eta d\theta
1880: e^{-\lambda{\tilde L}_0}\Str[\Pi {\cal O}_\alpha]}~~.
1881: \ee
1882: We next take the limit $\lambda\rightarrow +\infty$ with $\lambda \in \R_+$. 
1883: As in Section \ref{loc_bulk}, this forces the integral to localize on the
1884: critical points of $W$, while the Gaussian approximation around these points
1885: becomes exact. Counting the powers of $\lambda$ produced by the bosonic and
1886: fermionic Gaussian integrals, we find that the correlator vanishes unless
1887: $\alpha=f$ with $f$ a section of $End(E)$. In this case, $\delta_b$-closure of
1888: ${\cal O}_f$ amounts to the conditions ${\bar \nabla}_Af=0$ and $[D,f]=0$, 
1889: and counting powers of $\lambda$ shows that the only contributions which
1890: survive in the limit come from those pieces of the factor 
1891: $\Pi$
1892: which are independent of $\theta$ and $\eta$. This gives:
1893: \bea
1894: \langle {\cal O}_\alpha\rangle_{\rm disk}&=&0~~{\rm unless}~~\alpha=f\in
1895: End(E)~~\\
1896: \langle {\cal O}_f\rangle_{\rm disk}&=& 
1897: \frac{C}{n!}\int_{X}{
1898: \Omega \frac{\Str[(i\partial D)^{\wedge n}f(z)]}{\partial_1W\dots \partial_nW}}~~,
1899: \eea
1900: where $C$ is the constant introduced in Section \ref{loc_bulk}. These
1901: expressions generalize the residue formula proposed in \cite{Kap2}. Notice,
1902: however, that the residue formula of \cite{Kap2} arises for $\mu=0$ and in the
1903: limit $\Re \lambda\rightarrow +\infty$. The limit proposed in \cite{Kap2} 
1904: (namely $\Re\mu\rightarrow +\infty$ with $\lambda=0$) does not suffice to
1905: localize the model's excitations unto the critical set of $W$. 
1906: 
1907: 
1908: \section{Conclusions}
1909: 
1910: 
1911: We gave a detailed and general discussion of localization in the bulk and boundary sectors
1912: of B-type topological Landau-Ginzburg models. In the bulk sector, we showed
1913: that careful reconsideration of the localization argument of \cite{Vafa_LG}
1914: leads to an entire family of localization formulae, parameterized by a complex
1915: number $\lambda$ of positive real part. When real, this parameter 
1916: measures the area of worldsheets with $S^2$ topology.
1917: The various "localization pictures" are related by a "homotopy flow"
1918: (a semigroup of operators homotopic to the identity), which implements
1919: rescalings of this area. The generalized
1920: localization argument leads to a one-parameter family of off-shell models for
1921: the bulk trace, extending the well-know result of \cite{Vafa_LG}. 
1922: The later is recovered for $\Re \lambda\rightarrow +\infty$, a degenerate
1923: limit which leads to the standard residue representation. 
1924: 
1925: In the boundary sector, a similar argument gives a family of localization
1926: formulae parameterized by complex variables $\lambda$ and $\mu$ subject to the
1927: condition $\Re \lambda>0$. When real, these parameters describe the area of a worldsheet with
1928: disk topology, respectively the length of its boundary.
1929: The boundary localization pictures are once again related by 
1930: a semigroup of homotopy equivalences, which implements rescaling of the disk's
1931: area and of the length of its boundary. This leads to a two-parameter family
1932: of off-shell models for the boundary trace. We also showed that the residue
1933: formula proposed in \cite{Kap2} arises in the limit $\lambda\rightarrow
1934: +\infty$ with $\mu=0$, and generalizes to the set-up of \cite{Labastida,
1935:   coupling}, which does not require constraints on the target space or on the
1936: rank of the holomorphic superbundle describing the relevant D-brane. In
1937: particular, this proves and generalizes the proposal of
1938: \cite{Kap2}, though the residue representation we have found arises in a limit which differs
1939: from previous proposals. The argument required to establish this result 
1940: is rather subtle, due to the complicated form of the boundary conditions.
1941: 
1942: \acknowledgments{The authors thank W. Lerche for collaboration in a related
1943:   project. C.~I.~L. thanks A. Kapustin and K. Hori for stimulating
1944:   conversations, and A. Klemm for support and interest in his work. }
1945: 
1946: 
1947: 
1948: \appendix
1949: 
1950: 
1951: \section{Euler-Lagrange variations and boundary conditions}
1952: \label{BCS}
1953: 
1954: Let us consider the Euler-Lagrange variations for our model.
1955: It is not hard to compute the variations of the bulk action:
1956: \bea
1957: \delta_\theta {\tilde S}_{bulk}&=&\frac{i}{2}\int_{\Sigma}{d^2\sigma \sqrt{g}
1958: \left[\varepsilon^{\alpha\beta}D_\beta\rho^i_\alpha+\frac{1}{2}G^{i{\bar
1959:       k}}D_{\bar
1960:   k}\partial_{\bar j} {\bar W}\eta^{\bar j}\right]\delta \theta_i}~~\label{bulk_EL_vars1}
1961: \\
1962: \delta_\eta {\tilde S}_{bulk}&=&\frac{1}{2}\int_{\Sigma}{d^2\sigma \sqrt{g}
1963: \left[g^{\alpha\beta} G_{{\bar i}j}D_\alpha\rho^j_\beta-\frac{i}{2}D_{\bar
1964:   i}\partial_{\bar j} {\bar W}\theta^{\bar j}\right]\delta\eta^{\bar
1965: i}}-\frac{1}{2}\int_{\partial\Sigma}{d\tau G_{{\bar i}j}(\rho_n^j+i\rho_0^j)\delta\eta^{\bar
1966: i}}~~~~~~~~~~\label{bulk_EL_vars2}\\
1967: \delta_\rho {\tilde S}_{bulk}&=&\frac{1}{2}\int_{\Sigma}{d^2\sigma \sqrt{g}
1968: \left[G_{i{\bar j}}
1969: (i\varepsilon^{\alpha\beta}D_\beta\theta^{\bar
1970:     j}+g^{\alpha\beta}D_\beta\eta^{\bar j})
1971: -\frac{1}{2}\varepsilon^{\alpha\beta}D_i\partial_j W
1972:   \rho_\beta^j\right]\delta \rho_\alpha^i}~~\nn\\
1973: &+&\frac{i}{2}\int_{\partial\Sigma}{d\tau
1974: G_{i{\bar j}}(\eta^{\bar j}+\theta^{\bar j})\delta \rho_0^i}~~\label{bulk_EL_vars3}\\
1975: \delta_\phi {\tilde S}_{bulk}&=&
1976: \int_{\Sigma}{d^2\sigma \sqrt{g}\left[
1977: -G_{{\bar i}j}\Delta\phi^j-\frac{i}{4}D_{\bar i}D_{\bar
1978:       j}\partial_{\bar k}{\bar W}\theta^{\bar j}\eta^{\bar
1979:       k}+\frac{i}{2}D_{\bar i}\partial_{\bar j}{\bar W}{\tilde F}^{\bar j}
1980: \right]\delta \phi^{\bar i}}\nn\\
1981: &+&\int_{\Sigma}{d^2\sigma \sqrt{g}\left[
1982: -G_{i{\bar j}}\Delta\phi^{\bar j}-\frac{i}{2}D_i\partial_j W {\tilde F}^j
1983: +\frac{1}{8}\varepsilon^{\alpha\beta}D_iD_j\partial_k W \rho_\alpha^j\rho_\beta^k
1984: \right]\delta \phi^i}\nn\\
1985: &+&\int_{\partial\Sigma}{d\tau G_{{\bar i}j}(\partial_n\phi^j+i{\dot \phi}^j)\delta
1986:   \phi^{\bar i}}+\int_{\partial\Sigma}{d\tau G_{i{\bar
1987:       j}}(\partial_n\phi^{\bar j}-i{\dot \phi}^{\bar j})\delta
1988:   \phi^i}~~.\label{bulk_EL_vars4}
1989: \eea
1990: For the boundary coupling ${\cal U}=Str H(0)$, we find:
1991: \be
1992: \delta {\cal U}=-\Str[H(0)I_C(\delta M)]~~,
1993: \ee
1994: where:
1995: \be
1996: I_C(\delta M)=\int_{0}^l{d\tau U(\tau)^{-1}\delta M(\tau) U(\tau)}~~.
1997: \ee
1998: For $\delta_M$ we substitute the Euler-Lagrange variations:
1999: \bea
2000: \delta_\theta M&=&0\\
2001: \delta_\eta M&=&-\left(\frac{1}{2}F_{{\bar
2002:       i}j}\rho_0^j+\frac{i}{2}\nabla_{\bar i}D^\dagger\right)\delta\eta ^{\bar i}\\
2003: \delta_\rho M&=&-\frac{1}{2}(\partial_iD+F_{i{\bar j}}\eta^{\bar j})\delta \rho_0^i~~
2004: \eea
2005: and: 
2006: \be
2007: U^{-1}\delta_\phi M U=\frac{d}{d\tau}(U^{-1}A_{\bar i}\delta \phi^{\bar
2008:   i}U)+U^{-1}(S_i\delta \phi^i+S_{\bar i}\delta \phi^{\bar i})U
2009: \ee
2010: with:
2011: \bea
2012: S_i&=&F_{i{\bar j}}{\dot \phi}^{\bar j}+\frac{1}{2}\partial_iF_{{\bar
2013:     j}k}\eta^{\bar j}\rho_0^k+\frac{1}{2}\rho_0^j\partial_i\partial_j
2014: D+\frac{i}{2}\eta^{\bar j}[F_{i{\bar j}},D^\dagger]+\frac{i}{2}[\partial_iD,D^\dagger]\\
2015: S_{\bar i}&=&F_{{\bar i}j}{\dot \phi}^j+\frac{1}{2}\nabla_{\bar i}F_{{\bar
2016:     j}k}\eta^{\bar j}\rho_0^k+\frac{1}{2}\rho_0^j[F_{{\bar
2017:     i}j},D]+\frac{i}{2}\eta^{\bar j}\nabla_{\bar i}\nabla_{\bar j}D^\dagger
2018: +\frac{i}{2}[D,\nabla_{\bar i}D^\dagger]~~.
2019: \eea
2020: This gives:
2021: \bea
2022: \delta_\theta {\cal U}&=&0\\
2023: \delta_\eta {\cal
2024:   U}&=&\int_{\partial\Sigma}{d\tau \Str[H(\tau)(\frac{i}{2}\nabla_{\bar
2025:     i}D^\dagger+\frac{1}{2}F_{{\bar i}j}\rho_0^j)]\delta\eta^{\bar i}}\\
2026: \delta_\rho {\cal U}&=& \frac{1}{2}\int_{\partial \Sigma}{d\tau
2027:   \Str[H(\tau)(\partial_iD+F_{i{\bar j}}\eta^{\bar j})]\delta \rho_0^i}\\
2028: \delta_\phi{\cal U}&=&-\int_{\partial \Sigma}{d\tau 
2029: \left(\Str[H(\tau)S_i]\delta\phi^i+\Str[H(\tau)S_{\bar i}]\delta \phi^{\bar i}\right)}~~.
2030: \eea
2031: To extract the boundary conditions, we write:
2032: \be
2033: e^{-{\tilde S}_{bulk}}{\cal U}=e^{-S_{eff}}~~,
2034: \ee
2035: where $S_{eff}={\tilde S}_{bulk}-\ln {\cal U}$ is viewed as a (non-local) worldsheet action. 
2036: Since we desire local equations of motion, the boundary contributions to
2037: (\ref{bulk_EL_vars1}-\ref{bulk_EL_vars4}) must cancel the variation of $\ln {\cal U}$:
2038: \be
2039: {\cal U} \delta{\tilde S}_{bulk}=\delta {\cal U}~~.
2040: \ee 
2041: Imposing this requirement, we find the boundary conditions:
2042: \bea
2043: G_{{\bar i}j}(\rho_n^j+i\rho_0^j){\cal U}&=&-\Str H(\tau)(i\nabla_{\bar i}D^\dagger +
2044: F_{{\bar i}j}\rho_0^j)\label{bc1}\\
2045: i(G_{i{\bar j}}\eta^{\bar j}+\theta_i){\cal U}&=&
2046: \Str H(\tau)(\partial_i D+F_{i{\bar j}}\eta^{\bar j})\label{bc2}\\
2047: G_{{\bar i}j}(\partial_n\phi^j+i{\dot \phi}^j){\cal U} &=&
2048: -\Str [H(\tau)S_{\bar i}]\label{bc3}\\
2049: G_{i{\bar j}}(\partial_n\phi^{\bar j}-i{\dot \phi}^{\bar j}){\cal U}&=&-
2050: \Str [H(\tau)S_i]~~.\label{bc4}
2051: \eea
2052: The Euler-Lagrange equations can be read off from the bulk contributions to 
2053: (\ref{bulk_EL_vars1}-\ref{bulk_EL_vars4}):
2054: \bea
2055: \epsilon^{\alpha\beta}D_\alpha \rho^i_\beta&=&\frac{1}{2}G^{i{\bar j}}D_{\bar
2056:   j}\partial_{\bar k}{\bar W}\eta^{\bar k}\\
2057: g^{\alpha\beta}D_\alpha\rho_\beta^i&=&\frac{i}{2}G^{i{\bar j}}D_{\bar j}\partial_{\bar
2058:   k}{\bar W}\theta^{\bar k}\\
2059: i\epsilon^{\alpha\beta}D_\beta \theta_i+g^{\alpha\beta}G_{i{\bar j}}D_\beta\eta^{\bar
2060:   j}&=&\frac{1}{2}\epsilon^{\alpha\beta}D_i\partial_jW
2061: \rho_\beta^j~~.
2062: \eea
2063: 
2064: It is not hard to see that the boundary conditions are BRST invariant modulo
2065: the equations of motion for $F$. For simplicity, we explain this for 
2066: condition (\ref{bc2}), which is of interest in Section
2067: \ref{loc_boundary}. Starting with (\ref{bc2}),
2068: one easily computes the BRST variations of the left and right hand sides:
2069: \bea
2070: \delta(LHS)&=&2iG_{i{\bar j}}{\tilde F}^{\bar j}{\cal U}+i(G_{i{\bar j}}\eta^{\bar
2071:     j}+\theta_i)\left[\frac{1}{2}\int_{\partial\Sigma}\rho^i \partial_i W\right]{\cal
2072:     U}~~\\
2073: \delta(RHS)&=&(\partial_i W){\cal U}+
2074: \left[\frac{1}{2}\int_{\partial\Sigma}\rho^k \partial_k W\right]
2075: \Str[H(\tau)(\partial_iD+F_{i{\bar
2076:     j}}\eta^{\bar j})]~~.
2077: \eea
2078: The two variations obviously agree if one uses equation (\ref{bc2}),
2079:     provided that the equation of motion ${\tilde F}^{\bar
2080:     i}=-\frac{i}{2}G^{{\bar i}j}\partial_j W$ holds.
2081: 
2082: 
2083: 
2084: 
2085: 
2086: 
2087: 
2088: \begin{thebibliography}{100}
2089: \bibitem{Vafa_LG}{C.~Vafa,
2090: ``Topological Landau-Ginzburg Models",
2091: Mod.\ Phys.\ Lett.\ A {\bf 6}, 337 (1991).}
2092: %%CITATION = MPLAE,A6,337;%%
2093: \bibitem{Witten_phases}{E.~Witten,
2094: ``Phases of N = 2 theories in two dimensions,''Nucl.\ Phys.\ B {\bf 403}, 159 (1993)
2095: [arXiv:hep-th/9301042].}
2096: %%CITATION = HEP-TH 9301042;%%
2097: \bibitem{Hori}{K.~Hori, A.~Iqbal and C.~Vafa,
2098: ``D-branes and mirror symmetry,''arXiv:hep-th/0005247.}
2099: %%CITATION = HEP-TH 0005247;%%
2100: \bibitem{Warner}{N.~P.~Warner,
2101: ``Supersymmetry in boundary integrable models,''
2102: Nucl.\ Phys.\ B {\bf 450}, 663 (1995)[arXiv:hep-th/9506064].}
2103: %%CITATION = HEP-TH 9506064;%%
2104: \bibitem{Kontsevich}{M. Kontsevich, {\em unpublished}.}
2105: \bibitem{Kap1}{A.~Kapustin and Y.~Li,
2106: ``D-branes in Landau-Ginzburg models and algebraic geometry,''
2107: JHEP {\bf 0312}, 005 (2003)[arXiv:hep-th/0210296].}
2108: %%CITATION = HEP-TH 0210296;%%
2109: \bibitem{Orlov}{D. Orlov, 
2110: "Triangulated categories of singularities and D-branes in Landau-Ginzburg
2111:   models", math.AG/0302304.}
2112: \bibitem{Lerche}{I.~Brunner, M.~Herbst, W.~Lerche and B.~Scheuner,
2113: ``Landau-Ginzburg realization of open string TFT,''
2114: arXiv:hep-th/0305133.}
2115: %%CITATION = HEP-TH 0305133;%%
2116: \bibitem{Kap2}{A.~Kapustin and Y.~Li,
2117: ``Topological correlators in Landau-Ginzburg models with boundaries,''
2118: arXiv:hep-th/0305136.}
2119: %%CITATION = HEP-TH 0305136;%%
2120: \bibitem{Kap3}{A.~Kapustin and Y.~Li,
2121: ``D-branes in topological minimal models: The Landau-Ginzburg approach,''
2122: arXiv:hep-th/0306001.}
2123: %%CITATION = HEP-TH 0306001;%%
2124: \bibitem{fractional}
2125: S.~K.~Ashok, E.~Dell'Aquila and D.~E.~Diaconescu,
2126: ``Fractional branes in Landau-Ginzburg orbifolds,''
2127: arXiv:hep-th/0401135.
2128: %%CITATION = HEP-TH 0401135;%%
2129: \bibitem{coupling}{C.~I.~Lazaroiu,
2130: ``On the boundary coupling of topological Landau-Ginzburg models,'' arXiv:hep-th/0312286.}
2131: %%CITATION = HEP-TH 0312286;%%
2132: \bibitem{Griffiths}{P. Griffiths and J. Harris: {\em Principles of Algebraic
2133:     Geometry}, Wiley, 1994. }
2134: \bibitem{Labastida}{J.~M.~F.~Labastida and P.~M.~Llatas,
2135: ``Topological matter in two-dimensions,''
2136: Nucl.\ Phys.\ B {\bf 379}, 220 (1992)[arXiv:hep-th/9112051].}
2137: %%CITATION = HEP-TH 9112051;%%
2138: 
2139: 
2140: \bibitem{Witten_CS}{E.~Witten,
2141: ``Chern-Simons gauge theory as a string theory,''
2142: Prog.\ Math.\  {\bf 133}, 637 (1995) [arXiv:hep-th/9207094].}
2143: %%CITATION = HEP-TH 0104200;%%
2144: \bibitem{CIL2}{C.~I.~Lazaroiu,
2145: ``Generalized complexes and string field theory,''
2146: JHEP {\bf 0106}, 052 (2001)[arXiv:hep-th/0102122].}
2147: %%CITATION = HEP-TH 0102122;%%
2148: \bibitem{CIL3}{C.~I.~Lazaroiu,
2149: ``Unitarity, D-brane dynamics and D-brane categories,''
2150: JHEP {\bf 0112}, 031 (2001)[arXiv:hep-th/0102183].}
2151: %%CITATION = HEP-TH 0102183;%%
2152: \bibitem{CIL4}{C.~I.~Lazaroiu,
2153: ``String field theory and brane superpotentials,''JHEP {\bf 0110}, 018 (2001)
2154: [arXiv:hep-th/0107162].}
2155: %%CITATION = HEP-TH 0107162;%%
2156: \bibitem{CIL5}{C.~I.~Lazaroiu,
2157: ``D-brane categories,'' Int.\ J.\ Mod.\ Phys.\ A {\bf 18}, 5299 (2003)
2158: [arXiv:hep-th/0305095].}
2159: %%CITATION = HEP-TH 0305095;%%
2160: \bibitem{CIL6}{C.~I.~Lazaroiu,
2161: ``Graded Lagrangians, exotic topological D-branes and enhanced  triangulated
2162: categories,''JHEP {\bf 0106}, 064 (2001)[arXiv:hep-th/0105063].}
2163: %%CITATION = HEP-TH 0105063;%%
2164: \bibitem{CIL7}{C.~I.~Lazaroiu, R.~Roiban and D.~Vaman,
2165: ``Graded Chern-Simons field theory and graded topological D-branes,''
2166: JHEP {\bf 0204}, 023 (2002) [arXiv:hep-th/0107063].}
2167: %%CITATION = HEP-TH 0107063;%%
2168: \bibitem{CIL8}{C.~I.~Lazaroiu and R.~Roiban,
2169: ``Holomorphic potentials for graded D-branes,''
2170: JHEP {\bf 0202}, 038 (2002)[arXiv:hep-th/0110288].}
2171: %%CITATION = HEP-TH 0110288;%%
2172: \bibitem{CIL9}{C.~I.~Lazaroiu,
2173: ``An analytic torsion for graded D-branes,''
2174: JHEP {\bf 0209}, 023 (2002)[arXiv:hep-th/0111239].}
2175: %%CITATION = HEP-TH 0111239;%%
2176: \bibitem{CIL10}{C.~I.~Lazaroiu and R.~Roiban,
2177: ``Gauge-fixing, semiclassical approximation and potentials for graded
2178: Chern-Simons theories,'' JHEP {\bf 0203}, 022 (2002)[arXiv:hep-th/0112029].}
2179: %%CITATION = HEP-TH 0112029;%%
2180: \bibitem{Diaconescu}{D.~E.~Diaconescu,
2181: ``Enhanced D-brane categories from string field theory,''
2182: JHEP {\bf 0106}, 016 (2001) [arXiv:hep-th/0104200].}
2183: \bibitem{Moore_Segal}{G. Moore and G. Segal, unpublished; 
2184: see http://online.kitp.ucsb.edu/online/mp01/}
2185: \bibitem{Moore}{G.~W.~Moore,
2186: ``Some comments on branes, G-flux, and K-theory,''
2187: Int.\ J.\ Mod.\ Phys.\ A {\bf 16}, 936 (2001)
2188: [arXiv:hep-th/0012007].}
2189: %%CITATION = HEP-TH 0012007;%%}
2190: \bibitem{CIL1}{C.~I.~Lazaroiu,
2191:  ``On the structure of open-closed topological field theory in two dimensions,''
2192: Nucl.\ Phys.\ B {\bf 603}, 497 (2001) [arXiv:hep-th/0010269].}
2193: %%CITATION = HEP-TH 0010269;%%
2194: \bibitem{Witten_mirror}{E.~Witten,
2195: ``Mirror manifolds and topological field theory,'', in " Essays on mirror
2196: manifolds",  pp 120--158, Internat. Press, Hong Kong, 1992.  [arXiv:hep-th/9112056].}
2197: %%CITATION = HEP-TH 9112056;%%
2198: 
2199: 
2200: \bibitem{us}{M.~Herbst, C.~I.~Lazaroiu and W.~Lerche,
2201: ``Superpotentials, A(infinity) relations and WDVV equations for open
2202: topological strings,'' arXiv:hep-th/0402110.}
2203: %%CITATION = HEP-TH 0402110;%%
2204: 
2205: \bibitem{Quillen}{D. Quillen, "Superconnections and the Chern character",
2206: Topology {\bf 24} (1985), no. 1, 89--95.}
2207: \end{thebibliography}
2208: \end{document}
2209: