1: \newif\iffigs\figstrue
2: % Uncomment the next line if you don't want the figures:
3: %\figsfalse
4:
5: \documentclass[paper, 12pt, letterpaper]{JHEP}
6:
7:
8: \def\Bbb{\bf}
9: \def\C{{\Bbb C}}
10: \def\R{{\Bbb R}}
11: \def\Z{{\Bbb Z}}
12: \def\H{{\Bbb H}}
13:
14: \def\Hom{\operatorname{Hom}}
15: \def\Tors{\operatorname{Tors}}
16: \def\Ker{\operatorname{Ker}}
17: \def\Spec{\operatorname{Spec}}
18: \def\Area{\operatorname{Area}}
19: \def\Vol{\operatorname{Vol}}
20: \def\ad{\operatorname{ad}}
21: \def\tr{\operatorname{tr}}
22: \def\Pic{\operatorname{Pic}}
23: \def\disc{\operatorname{disc}}
24: \def\cpl{\operatorname{cpl}}
25: \def\Img{\operatorname{Im}}
26: \def\Rea{\operatorname{Re}}
27: \def\Gr{\operatorname{Gr}}
28: \def\SO{\operatorname{SO}}
29: \def\Sl{\operatorname{SL}}
30: \def\GO{\operatorname{O{}}}
31: \def\SU{\operatorname{SU}}
32: \def\GU{\operatorname{U{}}}
33: \def\Sp{\operatorname{Sp}}
34: \def\Spin{\operatorname{Spin}}
35: \def\rank{\operatorname{rank}}
36: \def\bearray{\begin{eqnarray}}
37: \def\eearray{\end{eqnarray}}
38: \def\bearraynn{\begin{eqnarray*}}
39: \def\eearraynn{\end{eqnarray*}}
40: \def\bfig{\begin{figure}}
41: \def\efig{\end{figure}}
42: \def\Aff{\operatorname{Aff}}
43: \def\diag{\operatorname{diag}}
44: \def\opeq#1{\advance\lineskip#1 \advance\baselineskip#1
45: \advance\lineskiplimit#1}
46: \def\eqalignsq#1{\null\,\vcenter{\opeq{2.5\jot}\mathsurround=0pt
47: \everycr={}\tabskip=0pt\offinterlineskip
48: \halign{\strut\hfil$\displaystyle{##}$&$\displaystyle{{}##}$\hfil
49: \crcr#1\crcr}}\,\null}
50: \def\eqalign#1{\null\,\vcenter{\opeq{2.5\jot}\mathsurround=0pt
51: \everycr={}\tabskip=0pt
52: \halign{\strut\hfil$\displaystyle{##}$&$\displaystyle{{}##}$\hfil
53: \crcr#1\crcr}}\,\null}
54:
55: \def\sm{$\sigma$-model}
56: \def\nlsm{non-linear \sm}
57: \def\smm{\sm\ measure}
58: \def\CY{Calabi--Yau}
59: \def\LG{Landau-Ginzburg}
60: \def\cR{{\Scr R}}
61: \def\cM{{\Scr M}}
62: \def\cA{{\Scr A}}
63: \def\cB{{\Scr B}}
64: \def\cK{{\Scr K}}
65: \def\cD{{\Scr D}}
66: \def\cH{{\Scr H}}
67: \def\cT{{\Scr T}}
68: \def\cL{{\Scr L}}
69: \def\cF{{\Scr F}}
70:
71:
72: \def\spnh{\Spin(32)/\Z_2}
73:
74: \def\Pf{{\em Proof: }}
75:
76: \def\ker{{\rm ker}}
77: \def\coker{{\rm coker}}
78: \def\rank{{\rm rank}}
79: \def\im{{\rm im}}
80: \def\dim{{\rm dim}}
81: \def\codim{{\rm codim}}
82: \def\Card{{\rm Card}}
83: \def\li{{\rm linearly independent}}
84: \def\ld{{\rm linearly dependent}}
85: \def\deg{{\rm deg}}
86: \def\det{{\rm det}}
87: \def\Div{{\rm Div}}
88: \def\supp{{\rm supp}}
89: \def\Gr{{\rm Gr}}
90: \def\End{{\rm End}}
91: \def\Aut{{\rm Aut}}
92:
93:
94: \newtheorem{Proposition}{Proposition}[section]
95: \newtheorem{Definition}{Definition}[section]
96: \newtheorem{Theorem}{Theorem}[section]
97: \newtheorem{Lemma}{Lemma}[section]
98: \newtheorem{Corrolary}{Corrolary}[section]
99:
100:
101:
102: \newcommand{\be}{\begin{equation}}
103: \newcommand{\ee}{\end{equation}}
104: \newcommand{\bea}{\begin{eqnarray}}
105: \newcommand{\eea}{\end{eqnarray}}
106: \newcommand{\we}{\wedge}
107: \newcommand{\bp}{\begin{Proposition}}
108: \newcommand{\ep}{\end{Proposition}}
109: \newcommand{\bt}{\begin{Theorem}}
110: \newcommand{\et}{\end{Theorem}}
111: \newcommand{\bl}{\begin{Lemma}}
112: \newcommand{\el}{\end{Lemma}}
113: \newcommand{\bc}{\begin{Corrolary}}
114: \newcommand{\ec}{\end{Corrolary}}
115: \newcommand{\nn}{\nonumber}
116:
117:
118: %EOF
119: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
120: \newcommand{\rf}[1]{(\ref{#1})}
121: \renewcommand{\theequation}{\thesection.\arabic{equation}}
122: %\renewcommand{\thefootnote}{\fnsymbol{footnote}}
123:
124: %\def\appendix#1{
125: % \addtocounter{section}{1} \setcounter{equation}{0}
126: %\renewcommand{\thesection}{\Alph{section}} \section*{Appendix}
127: %\addcontentsline{toc}{section}{Appendix \thesection\ \ \ #1} }
128:
129: \newcommand{\newsection}{ % Numeration of eqs. is automatic
130: \setcounter{equation}{0}
131: \section}
132: \newcommand{\non}{\nonumber \\*}
133:
134: \newcommand{\eq}[1]{Eq.~(\ref{#1})}
135: \newcommand{\nonu}{\nonumber}
136: \def \ov {\over }
137: \def\bea{\begin{eqnarray}}
138: \def\eea{\end{eqnarray}}
139: \def \ci {\cite}
140: \def \de{\partial}
141: \def \x{{\bf x}}
142: \def\y{{\bf y}}
143: \def\k{{\bf k}}
144: \def\LB{\left(}
145: \def\RB{\right)}
146: \def\be{\begin{equation}}
147: \def\ee{\end{equation}}
148: \def\ba{\begin{eqnarray}}
149: \def\ea{\end{eqnarray}}
150: \def\la{\label}
151: \def \bi{\bibitem}
152: \def \Tr {{\rm tr}}
153: \def\O{{\cal O}}
154: \def\C{{\cal C}}
155: \def\LA{\langle}
156: \def\RA{\rangle}
157: \def\X{{\bar X}}
158: \def\a{\alpha}
159: \def\b{\beta}
160: \def\e{\epsilon}
161: \def\A{{\cal A}}
162: \def\n{{\nabla}}
163: \def\G{{\Gamma}}
164: \def\1{{{(1)}}}\def\2{{{(2)}}}\def\3{{{(3)}}}
165:
166:
167:
168: \font\mybbb=msbm10 at 8pt
169: \font\mybb=msbm10 at 12pt
170: \def\bbb#1{\hbox{\mybbb#1}}
171: \def\bb#1{\hbox{\mybb#1}}
172: \def\pRe{\bbb{R}}
173: \def\Re {\bb{R}}
174: \def\Z {\bb{Z}}
175: \def\q{\bb{a}}
176: \def\id{\protect{{1 \kern-.28em {\rm l}}}}
177: \def\cn{{\cal N}}
178:
179: \def\p#1{{\phi{}^{(#1)}}}
180: \def\hp#1{{{\phi'}{}^{(#1)}}}
181: \def\tp#1{{{\tilde\phi}{}^{(#1)}}}
182: \def\c#1{{c{}^{(#1)}}}
183: \def\hc#1{{{c'}{}^{(#1)}}}
184: \def\tc#1{{{\tilde c}{}^{(#1)}}}
185:
186: %\def\boldphi{{\phi\hspace{-6.62pt}\phi\hspace{-6.62pt}\phi}}
187: \def\boldphi{\mbox{\boldmath $\phi$}}
188:
189:
190: %%\boldwe is used in the definition of \bwe. It can also be used separately.
191: \def\boldwe{\mbox{\boldmath $\wedge$}}
192: \def\bwe{{{{\boldwe\hspace{-8.9pt}\boldwe}\hspace{-8.8pt}\boldwe}
193: \hspace{-8.8pt}\boldwe}}
194: \def\bweft{{{{\boldwe\hspace{-8.1pt}\boldwe}\hspace{-8pt}\boldwe}
195: \hspace{-8pt}\boldwe}}
196:
197:
198: \def\pb#1{{{\boldphi}{}^{(#1)}}}
199: \def\hpb#1{{{\bf\boldphi'}{}^{(#1)}}}
200: \def\tpb#1{{{\bf\tilde\boldphi}{}^{(#1)}}}
201: \def\cb#1{{{\bf c}{}^{(#1)}}}
202: \def\hcb#1{{{\bf c'}{}^{(#1)}}}
203: \def\tcb#1{{{\bf\tilde c}{}^{(#1)}}}
204:
205:
206: \font\gothics=ygoth at 10pt
207: \font\gothicl=ygoth at 12pt
208: \def\gg#1{\hbox{\gothics#1}}
209:
210: \def\gG{\hbox{\gothicl G}}
211:
212: \font\frakl=yfrak at 12pt
213: \font\fraks=yfrak at 10pt
214:
215: \def\fG{\hbox{\frakl G}}
216:
217: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
218:
219:
220:
221:
222: \usepackage{graphics}
223:
224:
225: \title{Graded Chern-Simons field theory and graded topological
226: D-branes}
227:
228: \author{C.~I.~Lazaroiu, R. Roiban and D. Vaman
229: \\C.~N.~Yang Institute for Theoretical Physics\\
230: SUNY at Stony BrookNY11794-3840, U.S.A.\\
231: calin, roiban, dvaman@insti.physics.sunysb.edu
232: }
233:
234:
235: \abstract{We discuss graded D-brane systems of the topological A model
236: on a Calabi-Yau threefold,
237: by means of their string field theory. We give a detailed analysis
238: of the extended string field action, showing that it
239: satisfies the classical master equation, and construct
240: the associated BV system. The analysis is entirely general
241: and it applies to any collection of D-branes (of distinct grades)
242: wrapping the same special Lagrangian cycle, being valid in
243: arbitrary topology. Our discussion employs a
244: $\Z$-graded version of the covariant BV formalism, whose
245: formulation involves the concept of
246: {\em graded supermanifolds}.
247: We discuss this formalism in detail
248: and explain why $\Z$-graded supermanifolds are necessary for a correct
249: geometric understanding of BV systems.
250: For the particular case of graded D-brane pairs,
251: we also
252: give a direct construction of the master action, finding complete agreement
253: with the abstract formalism. We analyze formation
254: of acyclic composites and show that, under certain topological assumptions,
255: all states resulting from the condensation process of a pair of branes
256: with grades differing by one unit are BRST trivial and thus the
257: composite can be viewed as a closed string vacuum. We prove
258: that there are {\em six} types of pairs which must
259: be viewed as generally inequivalent. This contradicts the
260: assumption that `brane-antibrane' systems exhaust the nontrivial
261: dynamics of topological A-branes with the same geometric support.
262: }
263:
264:
265: \preprint{YITP-SB 01-37}
266:
267:
268:
269:
270: \begin{document}
271:
272:
273: \tableofcontents
274:
275: \pagebreak
276:
277: \section{Introduction}
278: ${}$
279:
280: The issue of D-brane composite formation plays a central role in the
281: study of Calabi-Yau compactifications of open superstrings. It has
282: gradually become clear \cite{Douglas_Kontsevich, com1, com3,
283: Aspinwall, Diaconescu, sc, Oz_superconn, Oz_triples} that a proper
284: understanding of D-brane condensation processes holds the key to
285: disentangling the relevance of the derived category of coherent
286: sheaves and its A-model analogue, the derived category of Fukaya's
287: category \cite{Fukaya, Fukaya2, Kontsevich_recent} (or rather, a
288: generalization thereof \cite{sc}). Since D-brane condensation
289: involves off-shell string dynamics, the most systematic approach to
290: this problem is through the methods of open string field theory. In
291: fact, it seems that `mirror symmetry with open strings' is best
292: formulated as a (quasi)-equivalence of open string field theories with
293: D-branes, an overarching off-shell version which would subsume all
294: previous statements. A fundamental ingredient in this program is the
295: observation of \cite{Douglas_Kontsevich} that the correct description
296: of D-branes in Calabi-Yau compactifications involves an extra datum,
297: an integer-valued quantity called the D-brane's grade. As pointed out
298: in \cite{com3}, the procedure of including graded D-branes admits a
299: general string field theoretic description (the so-called `shift
300: completion' of a D-brane category). Taking this fact into account
301: leads to a concrete presentation of the {\em extended} moduli space
302: of open string vacua, an approach which affords computation of various
303: physical quantities away from points of the geometric moduli space.
304: Moreover, the papers \cite{com3, Diaconescu} and \cite{sc} gave an
305: explicit description of certain subsectors of the relevant string
306: field theories for the case of Calabi-Yau threefolds.
307:
308: The purpose of this paper is to continue the analysis (initiated in
309: \cite{sc}) of a string field-theoretic model relevant for the dynamics
310: of A-type \cite{Ooguri} graded {\em topological} D-branes of a
311: Calabi-Yau threefold compactification. In \cite{sc}, a sector of the
312: string field theory of such objects was described in terms of a
313: $\Z$-graded version of Chern-Simons field theory living on a special
314: Lagrangian cycle. The model captures the off-shell dynamics of
315: topological branes of arbitrary grades wrapped over the cycle. As
316: sketched in that paper, the classical moduli space of vacua of this
317: theory allows one to recover a piece of the extended boundary moduli
318: space, an object which plays a crucial role in the homological mirror
319: symmetry program \cite{Kontsevich}. More precisely, points of the
320: extended moduli space can be described as {\em generalized D-brane
321: composites}, obtained by condensing boundary condition changing
322: states between {\em graded} D-branes. This suggest that one can
323: extract physical information about such points by studying the quantum
324: dynamics of the resulting theory.
325:
326: Since the theory written down in \cite{sc} involves higher rank forms,
327: its quantization must deal with the issue of reducible gauge algebras.
328: Therefore, a correct analysis of this theory requires the full force
329: of the Batalin-Vilkovisky formalism. The purpose of the present paper
330: is to carry out the classical part of this analysis, thus providing a
331: precise starting point for a study of quantum dynamics. Applying such
332: methods, we will be able to recover the extended action already
333: written down in \cite{sc}, which plays the role of (classical) master
334: action for our theory. This enables us to show that the model
335: constructed in \cite{sc} is a consistent starting point for a
336: quantum-mechanical analysis. As an application, we give a detailed
337: discussion of graded D-brane pairs, thus obtaining a realization of
338: ideas proposed in \cite{Vafa_cs}, though in a somewhat different
339: context.
340:
341: Our investigation reveals that there are six types of D-brane pairs
342: which are physically inequivalent in general.
343: This confirms the $\Z_6$ periodicity
344: of the D-brane grade suggested in \cite{Douglas_Kontsevich} from
345: worldsheet considerations and contradicts the assumption that
346: `brane-antibrane' systems exhaust the nontrivial dynamics of
347: topological A-branes with the same geometric support.
348:
349: The paper is organized as follows. We start in Section 2 by
350: recalling the construction of the classical string field action of
351: \cite{sc}. In section 3, we discuss the associated extended string
352: field theory, giving a detailed presentation of the structures
353: involved. Though this section is somewhat technical, a clear
354: description of these structures is crucial for a correct understanding
355: of latter work. The central objects are the so-called extended
356: boundary product and extended bilinear form introduced in \cite{sc},
357: whose construction we explain in detail. In Section 4, we proceed to
358: show that the extended action discussed in Section 3 satisfies the
359: classical master equation with respect to the antibracket induced by a
360: certain odd symplectic form (which coincides with the extended
361: bilinear form up to sign factors and a shift of grading). This
362: establishes the fact that the the extended action plays the role of
363: classical BV action for our systems. The proof, which is completely
364: general, makes use of the geometric version of the BV formalism, as
365: discussed, for example, in \cite{Kontsevich_Schwarz}. Our approach is
366: in fact a certain modification of usual covariant framework, which
367: differs from the latter by incorporating the ghost grading. This
368: modified formalism, which is necessary for a correct description of BV
369: systems, involves the concept of {\em graded supermanifolds} which was
370: recently introduced in \cite{Voronov}. We therefore start with a
371: general exposition of the geometric $\Z$-graded formalism, which is of
372: independent interest for foundational studies of BV quantization. We
373: then apply this framework to the extended string field theory of
374: Section 3. This allows us to give a complete construction of the
375: relevant BV system, and a very concise and general proof of the fact
376: that the extended action satisfies the classical master equation. We
377: also identify the classical gauge which leads to the unextended string
378: field theory. After discussing the form of the BRST operator in this
379: gauge, we proceed with a discussion of the particular case of graded
380: D-brane pairs. Section 5 recalls the open string interpretation of the
381: various data discussed in Section 2, and gives their concrete
382: description for such systems. In Section 6, we consider
383: composite formation for D-brane pairs and the issue of acyclic
384: condensates. After discussing the worldsheet BRST cohomology and
385: explaining how it distinguishes between the various pairs, we explain
386: the interpretation of string field vacua as points of the extended
387: moduli space, and give an explicit construction of acyclic composites
388: for certain graded D-brane pairs whose relative grading equals one.
389: This gives a concrete realization of suggestions made in
390: \cite{Vafa_cs}.
391:
392: While the geometric BV framework of Section 4 is extremely powerful
393: and general, it does not explain the origin of the various components
394: of the extended string field. To gain some insight into this issue, we
395: proceed in Section 7 with a direct construction of the BV action for
396: the case of graded pairs. We show that the more familiar
397: component approach leads to an action which can be viewed as the
398: expression of the extended action of Section 3 in a particular system
399: of linear coordinates, and show how the various ghosts and antifields
400: arise in standard manner by performing the BV-BRST resolution of our
401: (closed, but generally reducible) gauge algebras. We also give a
402: discussion of the relation between pairs with arbitrarily high
403: relative grade. Section 8 connects these results with the geometric
404: approach of Section 4, and in particular gives a very concise
405: formulation of the BV-BRST algorithm. This synthetic description
406: should be useful for understanding the structure of the gauge algebra
407: of systems containing more than two graded branes. We end in Section
408: 9 by presenting our conclusions and a few directions for further
409: research.
410:
411:
412: \section{A string field theory for graded topological D-branes}
413:
414: This section describes the string field theory of a collection of
415: graded topological D-branes wrapping the same special Lagrangian cycle
416: of a Calabi-Yau threefold. This theory was written down in \cite{sc}
417: starting from a worldsheet analysis and using the framework developed
418: in \cite{com1, com3}.
419:
420: \subsection{Graded D-branes}
421:
422: We start by recalling some basic concepts introduced in
423: \cite{Douglas_Kontsevich} (see also \cite{Aspinwall} and \cite{sc}).
424: We are interested in so-called {\em graded topological D-branes} of
425: the A-model compactified on a Calabi-Yau threefold $X$. Recall from
426: \cite{Witten_CS} that an {\em ungraded} topological D-brane can be
427: described as a pair $(L,E)$ where $L$ is a (connected) special
428: Lagrangian cycle of $X$ and $E$ is a flat vector bundle on
429: $L$\footnote{We remind the reader that this is a vector bundle endowed
430: with a flat structure, i.e. the equivalence class of a family of
431: local trivializations whose transition functions are constant.
432: Specifying a flat structure amounts to giving a gauge-equivalence
433: class of flat connections.}. The generalization to {\em graded}
434: D-branes \cite{Douglas_Kontsevich, sc} is obtained by replacing $L$
435: with a graded version \cite{Seidel}, which for practical purposes
436: amounts to fixing an integer $n$ (the brane's {\em grade}) and a
437: certain orientation of $L$ which depends on $n$. As discussed in
438: \cite{Douglas_Kontsevich, Aspinwall, sc}, the worldsheet $U(1)$ charge
439: in the boundary condition changing sectors between two graded D-branes
440: wrapping $L$ is then shifted by $\pm n$, while the boundary products
441: in various sectors contain extra signs. Moreover, the boundary metric
442: receives grade-dependent signs in boundary condition changing sectors,
443: due to the fact that the relevant orientation of $L$ depends on the
444: branes' grades.
445:
446:
447:
448: \subsection{Graded Chern-Simons field theory as a string field
449: theory for graded D-branes}
450:
451:
452: As explained in detail in \cite{sc}, it is possible to describe
453: certain graded D-brane systems through a generalization of
454: Chern-Simons field theory. To be specific, we consider a collection of
455: graded topological A-type branes wrapping {\em the same} special
456: Lagrangian cycle $L$ in $X$. The main assumption for what follows is
457: that no two D-branes of this collection have the same grade. This
458: allows one to label the branes by their grades, which form a finite or
459: infinite set of integers. If $a_n$ denotes the D-brane of grade $n$,
460: then we denote by $E_n$ its underlying bundle. This notation includes
461: the specification of a flat structure (flat connection) on each $E_n$.
462:
463: With these hypotheses, it was shown in \cite{sc} that the string field
464: theory of the system is a graded version of Chern-Simons field theory,
465: which generalizes both the usual Chern-Simons description of
466: \cite{Witten_CS} and the supergroup Chern-Simons proposal of
467: \cite{Vafa_cs}. This describes the dynamics of so-called `degree one
468: graded connections' \cite{Bismut_Lott} on the graded (super-)bundle
469: ${\bf E}=\oplus_{n}{E_n}$. To define the theory, one must first
470: describe the so-called {\em total boundary space} ${\cal H}$ of
471: \cite{sc}, which consists of the off-shell states of open strings
472: stretching between our branes.
473:
474: \subsubsection{The total boundary algebra}
475:
476: We start by considering the algebra ${\cal E}$ of endomorphisms of
477: ${\bf E}$. This admits a natural $\Z$-grading induced by the bundle
478: decomposition: \be
479: \label{dec}
480: End({\bf E})=\oplus_{k}{End_k({\bf E})}~~, \ee where: \be End_k({\bf
481: E})=\oplus_{n-m=k}{Hom(E_m, E_n)}~~. \ee If $f$ is a morphism from
482: $E_m$ to $E_n$, then its degree with respect to this grading is given
483: by: \be \Delta(f)=n-m~~. \ee It is easy to see that the composition
484: of morphisms is homogeneous of degree zero with respect to this
485: grading: \be \Delta(f\circ g)=\Delta(f)+\Delta(g)~~. \ee It follows
486: that ${\cal E}$ (with multiplication given by composition of
487: morphisms) forms a graded associative algebra with respect to the
488: grading induced by $\Delta$.
489:
490: The next step is to consider the tensor product ${\cal
491: H}=\Omega^*(L)\otimes {\cal E}$ between the exterior algebra of $L$
492: and the endomorphism algebra ${\cal E}$. Since both algebras are
493: $\Z$-graded, ${\cal H}$ is endowed with gradings $rk$ and $\Delta$
494: induced from its components, as well as with the total grading
495: $|~.~|=rk +\Delta$. On decomposable elements $u=\rho\otimes f$, these
496: are given by: \be rk u=rk
497: \rho~~,~~\Delta(u)=\Delta(f)~~,~~|u|=rk\rho+\Delta(f)~~. \ee An
498: arbitrary element $u$ of ${\cal H}$ can be viewed as a matrix
499: $u=(u_{mn})$, where $u_{mn}\in \Omega^*(L)\otimes \Gamma(Hom(E_m,
500: E_n))$ is a bundle-valued form. Then $\Delta(u_{mn})=n-m$ and
501: $|u_{mn}|=rk u_{mn}+n-m$~~. The space ${\cal H}$ is also endowed with
502: a canonical multiplication $\bullet$ (the `total boundary product' of
503: \cite{sc}), induced from the multiplicative structure of its tensor
504: components. On decomposable elements $u=\rho\otimes f$ and
505: $v=\eta\otimes g$, this is given by: \be
506: \label{dot}
507: u\bullet v=(-1)^{\Delta(f)rk \eta}(\rho\wedge \eta)\otimes (f\circ
508: g)~~. \ee Up to the sign prefactor, the right hand side is simply the
509: usual wedge product of $End({\bf E})$ -valued forms, which includes
510: composition of bundle morphisms on the coefficients: \be u\wedge
511: v=(\rho\wedge \eta)\otimes (f\circ g)~~. \ee This allows us to write
512: (\ref{dot}) in a slightly more familiar form: \be u\bullet
513: v=(-1)^{\Delta(u)rk v}u\wedge v~~. \ee
514:
515: The boundary product is homogeneous of degree zero with respect to the
516: total grading: \be |u\bullet v|=|u|+|v|~~. \ee Hence ${\cal H}$
517: becomes a graded associative algebra when endowed with the grading
518: $|~.~|$ and the product $\bullet$. This well-known construction of a
519: graded associative structure on the tensor product from similar
520: structures on components is usually denoted by: \be {\cal
521: H}=\Omega^*(L){\hat \otimes} {\cal E}~~, \ee where the hat above the
522: tensor product indicates that the multiplication and grading on the
523: resulting space are constructed in the canonical manner discussed
524: above.
525:
526:
527: A supplementary datum is provided by the existence of a natural
528: differential $d$ on the product algebra ${\cal H}$. This is the
529: exterior differential on $End({\bf E})$-valued forms, twisted with the
530: direct sum connection $A=\oplus_n{A_n}$. Flatness of $A_n$ assures
531: that $d^2=0$, and definition (\ref{dot}) implies that $d$ acts as a
532: graded derivation of the product $\bullet$ (with respect to the total
533: grading): \be d(u\bullet v)=(du)\bullet v+(-1)^{|u|}u\bullet (dv)~~.
534: \ee Moreover, one has: \be |du|=|u|+1~~. \ee
535:
536:
537: The final element needed in the construction is a `trace' on ${\cal
538: H}$. This is induced by the natural traces on $\Omega^*(L)$ and
539: ${\cal E}$, which are defined as follows. For complex-valued forms
540: $\rho\in \Omega^*(L)$, we define: \be
541: Tr_\Omega(\rho)=\int_{L}{\rho}~~, \ee while for $End({\bf E})$-valued
542: morphisms $f$ we have the supertrace of \cite{Quillen}: \be
543: \label{str}
544: str(f)=\sum_{m}{(-1)^{m}tr_{m}(f_{mm})}~~, \ee where $tr_{m}$ is the
545: fiberwise trace in the bundle $End(E_m)$. Note that $str(f)$ is a
546: complex-valued function defined on $L$. When all components $E_n$ of
547: ${\bf E}$ have grades $n$ of the same parity (which in the language of
548: \cite{Quillen} amounts to taking the even or odd component of ${\bf
549: E}$ to be the zero bundle), the supertrace (\ref{str}) reduces to
550: $\pm$ the ordinary fiberwise trace on the bundle $End({\bf E})$.
551:
552: Both traces are graded-symmetric with respect to the natural degrees
553: on their spaces of definition: \be
554: \label{traces}
555: Tr_\Omega(\rho\wedge \eta)=(-1)^{rk\rho rk\eta}Tr_{\Omega}(\eta\wedge
556: \rho) ~~,~~ str(f\circ g)=(-1)^{\Delta(f)\Delta(g)}str(g\circ f)~~.
557: \ee Since $Tr_\Omega(u\wedge v)$ and $str(f\circ g)$ vanish unless $rk
558: \rho+rk \eta=3$ (remember that $L$ is a 3-cycle !), respectively
559: $\Delta(f)+\Delta(g)=0$, the graded symmetry properties are equivalent
560: with: \be Tr_\Omega(\rho\wedge \eta)=Tr_{\Omega}(\eta\wedge
561: \rho)~~{\rm~and~}~~ str(f\circ g)=(-1)^{\Delta(f)}str(g\circ f)~~.
562: \ee Using (\ref{traces}), we define a trace on ${\cal H}$ which on
563: decomposable elements $u=\rho\otimes f$ is given by: \be Tr_{\cal
564: H}(u)=\int_{L}{str(f)\rho}~~. \ee It is easy to check that this is
565: graded-symmetric: \be Tr_{\cal H}(u\bullet v)=(-1)^{|u||v|}Tr_{\cal
566: H}(v\bullet u)~~. \ee It immediately follows that the
567: (nondegenerate) bilinear form on ${\cal H}$, defined through: \be
568: \langle u , v \rangle:=Tr_{\cal H}(u\bullet v)=\int_{L}{str(u\bullet
569: v)}~~, \ee is graded-symmetric as well: \be \langle u , v
570: \rangle=(-1)^{|u||v|}\langle v, u \rangle~~. \ee It is clear that
571: $\langle u, v\rangle$ vanishes on bi-homogeneous elements unless both
572: of the conditions $rk u+rk v=3$ and $\Delta(u)+\Delta(v)=0$ are
573: satisfied. It follows that non-vanishing of $\langle u, v\rangle$
574: requires $|u|+|v|=3$, for elements homogeneous with respect to the
575: total degree. Due to this selection rule, the graded symmetry property
576: is in fact equivalent with: \be \langle u , v \rangle=\langle v, u
577: \rangle~~. \ee The last properties we shall need are invariance of
578: the bilinear form with respect to the total boundary product and
579: differential: \be \langle du, v\rangle+(-1)^{|u|}\langle u,
580: dv\rangle=0~~,~~ \langle u\bullet v, w\rangle=\langle u, v\bullet
581: w\rangle~~. \ee These follows easily upon using the properties of the
582: differential and supertrace. We end by noting that the trace and
583: bilinear form can be expressed in more familiar language as follows:
584: \be Tr_{\cal H}(u)=\int_{L}{str(u)}~~,~~\langle u, v
585: \rangle=\int_{L}{(-1)^{\Delta(u)rk v} str(u\wedge v)}~~, \ee if one
586: extends the supertrace to bundle-valued forms through: \be
587: str(\rho\otimes f):=str(f)\rho~~. \ee
588:
589:
590:
591: \subsubsection{The string field action and its gauge algebra}
592:
593: The string field theory of \cite{sc} is described by the action: \be
594: \label{action}
595: S(\phi)=\int_{L}{str\left[\frac{1}{2}\phi\bullet d\phi+\frac{1}{3}
596: \phi\bullet\phi\bullet\phi\right]}~~. \ee This is defined on the
597: component ${\cal H}^1=\{\phi\in {\cal H}||\phi|=1\}$ of the total
598: boundary space. It can also be written in the form: \be
599: S(\phi)=\frac{1}{2}\langle \phi, d\phi\rangle +\frac{1}{3}\langle
600: \phi,\phi\bullet \phi\rangle~~, \ee which is discussed, for example,
601: in \cite{com1}. As explained in more detail below, the physical
602: interpretation of the defining data is as follows. $d$ plays the role
603: of a `total worldsheet BRST charge' on a certain collection of
604: (topological) open string sectors. The product $\bullet$ is a total
605: open string product for that collection and the bilinear form $\langle
606: .,.\rangle$ is a `total BPZ form' (total topological metric).
607:
608:
609: The action (\ref{action}) is invariant with respect to infinitesimal
610: gauge transformations of the form: \be
611: \label{gauge}
612: \phi\rightarrow
613: \phi+\delta_\alpha\phi=\phi-d\alpha-[\phi,\alpha]_\bullet~~, \ee where
614: $\delta_\alpha\phi=-d\alpha-[\phi,\alpha]_\bullet$, with $\alpha\in
615: {\cal H}^0=\oplus_{\tiny
616: \begin{array}{c}m,n\\m\geq n\end{array}}
617: \Gamma(Hom(E_m,E_n))\otimes\Omega^{m-n}(L)$ a charge zero element
618: $(|\alpha|=0$) of ${\cal H}$. In these relations, $[.,.]_\bullet$
619: denotes the graded commutator in the total boundary algebra, which on
620: arbitrary homogeneous elements is given by: \be
621: \label{comdot}
622: [u,v]_\bullet=u\bullet v -(-1)^{|u||v|}v\bullet u~~. \ee It is easy
623: to check that: \be
624: \delta_\alpha\delta_\beta\phi-\delta_\beta\delta_\alpha\phi=\delta_{
625: [\alpha,\beta]_\bullet}\phi~~, \ee so that the Lie algebra of
626: transformations of the form (\ref{gauge}) closes off-shell. Note that
627: $[\alpha,\beta]_\bullet=\alpha\bullet\beta-\beta\bullet \alpha$ since $\alpha$
628: and $\beta$ have vanishing $U(1)$ charge. In fact, relation
629: (\ref{comdot}) shows that our gauge transformations give a
630: representation of the (infinite-dimensional) Lie algebra ${\bf
631: g}=({\cal H}^0, [.,.]_\bullet)$, which is a subalgebra of the graded
632: Lie algebra $({\cal H}, [.,.]_\bullet)$. The gauge group ${\cal G}$
633: results by exponentiation of ${\bf g}$. This infinite-dimensional
634: group is rather exotic, since its generators are higher rank forms on
635: the cycle $L$. Note that we insist on considering {\em all} elements
636: of ${\cal H}^0$ as generators of this group, even though we start with
637: a particular background flat connection on $E=\oplus_{n}E_n$ which
638: happens to split as a direct sum $\oplus_{m}A_m$. Even for a direct
639: sum background, one {\em cannot} restrict to the `diagonal' subalgebra
640: $\oplus_{m} \Gamma(Hom(E_m,E_m))(L)$ of ${\bf g}$. Inclusion of
641: non-diagonal generators is necessary for consistency of the string
642: field interpretation, since such a decomposition of the connection is
643: accidental, and one can deform away from direct sum backgrounds by
644: condensing boundary condition changing states \cite{sc}.
645:
646:
647: \subsection{The open string interpretation}
648:
649: The precise interpretation of (\ref{action}) in terms of open A-type
650: strings arises upon applying the general formalism discussed in
651: \cite{com1, com3}. For this, one considers the decomposition: \be
652: \label{decomposition}
653: {\cal H}=\oplus_{m,n}{{\cal H}_{nm}}~~, \ee where ${\cal
654: H}_{nm}=\Omega^*(L)\otimes \Gamma(Hom(E_m, E_n))$ and notices that
655: the product $\bullet$ and differential $d$ are compatible with it in
656: the sense that $\bullet$ vanishes on ${\cal H}_{kn'}\times {\cal
657: H}_{nm}$ if $n\neq n'$ and takes ${\cal H}_{kn}\times {\cal H}_{nm}$
658: into ${\cal H}_{km}$, while $d$ takes ${\cal H}_{nm}$ into ${\cal
659: H}_{nm}$. This implies that the collection of spaces ${\cal
660: H}_{nm}$ can be viewed as the morphism spaces of a differential
661: graded category built on the objects $a_n$ (which, due to our
662: assumption, are in bijection with the grades $n$ present in the
663: system). The category interpretation results upon defining $Hom(a_m,
664: a_n):={\cal H}_{nm}$. One can further check that the bilinear form
665: $\langle . , .\rangle$ is compatible with the decomposition
666: (\ref{decomposition}), in the sense that it vanishes on ${\cal
667: H}_{m'n'}\times {\cal H}_{nm}$ unless $n'=n$ and $m'=m$. Then the
668: general discussion of \cite{com1} suggests that ${\cal H}_{nm}$ should
669: be interpreted as the (off-shell) state space of open topological
670: strings stretching from $a_m$ to $a_n$. This interpretation is indeed
671: valid, and can be recovered from the topological A-model as discussed
672: in \cite{sc}.
673:
674: For example, the fact that the natural grading on ${\cal H}_{nm}$ is
675: given by $|~.~|$ implies that the worldsheet $U(1)$ charge of states
676: for the string stretching from $a_m$ to $a_n$ is given by: \be
677: \label{ghost_mn}
678: |u|=rk u+n-m~~, \ee which agrees with the observation of
679: \cite{Douglas_Kontsevich} that the $U(1)$ charge is shifted in
680: boundary condition changing sectors connecting two graded topological
681: D-branes. A direct construction of the string field theory
682: (\ref{action}) can be found in the paper \cite{sc}, which takes the
683: sigma model perspective as a starting point.
684:
685:
686:
687: \section{The extended string field theory}
688:
689:
690: The theory (\ref{action}) can be extended in a manner reminiscent of
691: that discussed in \cite{Gaberdiel}. This extension was already written
692: down in \cite{sc}, which gave a very short discussion of its
693: structure. Here we give a more complete exposition.
694:
695: Since we start with an action based on a graded super-bundle, the
696: various objects involved in the extension procedure are somewhat
697: subtle and we shall give a careful discussion of their construction.
698: We warn the reader that a cursory reading of the present section may
699: lead to serious misunderstanding of our sign conventions.
700:
701:
702: \subsection{The extended boundary data}
703:
704: In order to formulate the extended theory, we must define an {\em
705: extended boundary algebra} $({\cal H}_e,d, *)$, a differential
706: graded associative algebra which extends
707: $({\cal H},d,\bullet)$. We
708: also need an {\em extended topological metric} $\langle .,.
709: \rangle_e$, a graded-symmetric nondegenerate bilinear form on ${\cal
710: H}_e$ which extends the BPZ form $\langle .,.\rangle$. We shall
711: consider the three elements $*$, $d$ and $\langle . , . \rangle_e$ in
712: turn.
713:
714: The extended boundary algebra $({\cal H}_e, *)$ is obtained (as in
715: \cite{Gaberdiel}) by considering a (complex) Grassmann algebra $G$ and
716: constructing the graded associative algebra ${\cal
717: H}_e=\Omega^*(L){\hat \otimes}{\cal E} {\hat \otimes}G$. The
718: extended boundary product $*$ will be the standard product on this
719: algebra, to be discussed below. The Grassmann algebra $G$
720: \footnote{All of the constructions of this paper can in fact be
721: carried out with an {\em arbitrary} commutative superalgebra $G$
722: with unit.} (whose elements we denote by $\alpha, \beta \dots$)
723: comes endowed with the Grassmann degree $g$ and a multiplication which
724: we write as juxtaposition. We allow $G$ to have any number of odd
725: generators, which we denote by $\xi^\mu$. An element of $G$ has the
726: form: \be
727: \label{Ggens}
728: \alpha=\alpha_0+\sum_{k\geq 1}
729: \sum_{\mu_1<\dots<\mu_k}{\alpha_{\mu_1\dots
730: \mu_k}\xi^{\mu_1}...\xi^{\mu_k}}~~, \ee where $\alpha_0$ and
731: $\alpha_{\mu_1..\mu_k}$ are complex numbers. We note the existence of
732: an evaluation map $ev_G$ from $G$ to $\C$, which projects out the odd
733: generators: \be
734: \label{evG}
735: ev_G(\alpha)=\alpha_0~~. \ee
736:
737: Since we shall tensor with $G$, we will encounter various
738: $\Z_2$-valued degrees, for which we use the following convention. We
739: let $\Z_2=\Z/2\Z=\{{\hat 0}, {\hat 1}\}$, where ${\hat 1}$ is the
740: unit. For an element $t\in \Z_2$, we define the power $(-1)^t$ by
741: picking any representative for $t$ in the covering space $\Z$; this is
742: clearly well-defined since $(-1)^2=+1$. If $n$ is an integer and $t$
743: is an element of $\Z_2$, then $nt\in \Z_2$ is the product of $t$ with
744: the mod 2 reduction of $n$.
745:
746: Associativity of ${\hat {\otimes}}$ allows us to view ${\cal H}_e$ as
747: either of the tensor products ${\cal H}{\hat \otimes }G$ or
748: $\Omega^*(L){\hat \otimes}{\hat {\cal E}}$, where ${\hat {\cal
749: E}}:={\cal E}{\hat {\otimes}} G$. It will be useful to discuss the
750: multiplicative structure of ${\cal H}_e$ from both of these
751: perspectives. For this, we first recall the multiplicative structure
752: on the factor ${\hat {\cal E}}$ of the second presentation.
753:
754: \subsubsection{The algebra ${\hat {\cal E}}$ of Grassmann-valued sections of
755: $End({\bf E})$}
756:
757: The space ${\hat {\cal E}}$ is the graded associative algebra of
758: Grassmann-valued sections of $End({\bf E})$. If ${\hat f}=f\otimes
759: \alpha$ and ${\hat g}=g \otimes \beta$ are two decomposable elements
760: of ${\hat {\cal E}}$ (with $f,g$ elements of ${\cal E}$ and $\alpha,
761: \beta$ elements of $G$), then their canonical multiplication is given
762: by: \be
763: \label{hatfg}
764: {\hat f}{\hat g}= (-1)^{g(\alpha)\Delta(g)}(f\circ g)\otimes
765: (\alpha\beta)~~. \ee
766:
767: One can extend the composition of morphisms $\circ$ from ${\cal E}$ to
768: a naive multiplication on ${\hat {\cal E}}$ given by: \be
769: \label{hatfg_circ}
770: {\hat f}\circ {\hat g}=(f\circ g)\otimes (\alpha\beta)~~. \ee In
771: terms of this naive product, the defining relation (\ref{hatfg})
772: becomes: \be {\hat f}{\hat g}=(-1)^{g({\hat f})\Delta({\hat g})}{\hat
773: f}\circ {\hat g}~~, \ee where $g({\hat f})=g(\alpha)$ and
774: $\Delta({\hat g})=\Delta(g)$ are the degrees induced on ${\hat {\cal
775: E}}$ by the $\Z_2$-grading of $G$ and the $\Z$-grading of ${\cal
776: E}$. Note that $g({\hat f})$ is simply the Grassmannality of ${\hat
777: f}$, while $\Delta({\hat g})=n-m$ if ${\hat g}$ is a morphism from
778: $E_m$ to $E_n$.
779:
780: The algebra ${\hat {\cal E}}$ is endowed with the total $\Z_2$-valued
781: grading induced by the sum $\sigma=\Delta~(mod~2)~+g$ between the
782: mod~2 reduction of $\Delta$ and the Grassmann degree $g$ on $G$: \be
783: \sigma(f\otimes \alpha)=\Delta(f)~(mod~2)~+g(\alpha)~~. \ee The
784: product (\ref{hatfg}) is homogeneous of degree zero with respect to
785: this grading: \be \sigma({\hat f}{\hat g})=\sigma({\hat
786: f})+\sigma({\hat g})~~. \ee
787:
788: The supertrace $str$ on ${\cal E}$ extends to a functional $str_e$ on
789: ${\hat {\cal E}}$, which associates a {\em Grassmann-valued} function
790: $str_e({\hat f})$ (defined on $L$) to each Grassmann-valued section of
791: ${\hat f}$ of $End({\bf E})$. On decomposable elements ${\hat
792: f}=f\otimes \alpha$, this `extended supertrace' is given by: \be
793: str_e(f\otimes \alpha):=str(f)\otimes \alpha~~. \ee
794:
795: It is easy to check that the extended supertrace has the property: \be
796: str_e({\hat f}{\hat g})=(-1)^{\sigma({\hat f})\sigma({\hat g})}
797: str_e({\hat g}{\hat f})~~. \ee
798:
799: \subsubsection{The multiplicative structure on ${\cal H}_e$}
800:
801: A decomposable element ${\hat u}$ of ${\cal H}_e$ can be presented as:
802: \be \label{dec1} {\hat u}=\rho\otimes f\otimes \alpha=u\otimes
803: \alpha=\rho\otimes {\hat f}~~, \ee where $\rho$ is a (complex-valued)
804: form on $L$, $f$ is an endomorphism of ${\bf E}$ and $\alpha$ is an
805: element of the Grassmann algebra $G$. In this relation, we defined
806: $u=\rho\otimes f$ (an element of ${\cal H}=\Omega^*(L)\otimes {\cal
807: E}$) and ${\hat f}=f\otimes \alpha$ (an element of ${\hat {\cal E}}=
808: {\cal E}\otimes G$). If ${\hat v}=\eta\otimes g\otimes \beta=v\otimes
809: \beta=\eta\otimes {\hat g}$ is another element of ${\cal H}_e$ (with
810: $v=\eta \otimes g$ and ${\hat g}=g\otimes \beta$), then the canonical
811: product $*$ in ${\cal H}_e$ is given by \footnote{It might be useful
812: to show that the product $*$ is indeed associative. Consider degree
813: one Grassmann valued forms in ${\cal H}_e$: $\hat a*(\hat b*\hat
814: c)=\hat a*(-1)^{(1-rk b )rkc+ g(b)\Delta(c)}\hat b \hat
815: c=(-1)^{(1-rk b )rk c+ g(b)\Delta(c)+(1-rk a )(rk b+rk
816: c)+g(a)(\Delta(b)+\Delta(c))} \hat a\hat b\hat c$ . On the other
817: hand $(\hat a*\hat b)*\hat c=(-1)^{(1-rk a )rk b + g(a)\Delta
818: b}(\hat a\hat b)*\hat c= (-1)^{(1-rk a)rk b+ g(a)\Delta(b)+(2-rk
819: a-rk b)rk c+(g(a)+g(b))\Delta(c)}\hat a\hat b\hat c$. Since the
820: signs are the same in both cases the product $*$ satisfies
821: associativity. }: \bea
822: \label{star1}
823: {\hat u}*{\hat v}&=&(-1)^{g(\alpha)|v|}(u\bullet v)\otimes
824: (\alpha\beta)
825: \nonu\\
826: &=&(-1)^{(g(\alpha)+\Delta(f))rk \eta}(\rho\wedge \eta)\otimes
827: ({\hat f}{\hat g})\nonu\\
828: &=& (-1)^{g(\alpha)|v|+\Delta(f)rk \eta}(\rho \wedge \eta) \otimes
829: (f\circ g)\otimes (\alpha\beta)~~. \eea The first two equations
830: correspond to viewing ${\cal H}_e$ as ${\cal H}{\hat \otimes }G$ and
831: $\Omega^*(L){\hat \otimes}{\hat {\cal E}}$ respectively. The last
832: treats ${\cal H}_e$ as the triple tensor product $\Omega^*(L){\hat
833: \otimes} {\cal E}{\hat \otimes }G$.
834:
835: The extended boundary space is equipped with the total ($\Z_2$-valued)
836: degree $deg$ induced from its components: \be deg({\hat u})= (rk
837: \rho+\Delta(f))~(mod~2)~+g(\alpha)=|u|~(mod~2)~+g(\alpha)= rk
838: \rho~(mod~2)~+\sigma({\hat f})~~, \ee on elements ${\hat u}$ of the
839: form (\ref{dec1}). The extended boundary product (\ref{star1}) is
840: homogeneous of degree zero with respect to this grading: \be deg
841: ({\hat u}*{\hat v})=deg ({\hat u})+deg ({\hat v})~~. \ee
842:
843:
844: \subsubsection{The trace and bilinear form on ${\cal H}_e$}
845: The extended boundary space is endowed with a trace $Tr_e$ which on
846: decomposable elements ${\hat u}=\rho\otimes f\otimes \alpha$ takes the
847: form: \be Tr_e({\hat u})=\int_{L}{\rho~str(f)~\alpha}=
848: \int_{L}{\rho~str_e({\hat f})}~~= Tr_{\cal H}(u)\otimes \alpha~~. \ee
849: This associates an element of $G$ to every element of ${\cal H}_e$. It
850: is easy to check that the extended trace is graded-symmetric with
851: respect to the total degree: \be
852: \label{tre}
853: Tr_e({\hat u}*{\hat g})=(-1)^{deg {\hat u}~deg {\hat v}} Tr_e({\hat
854: v}*{\hat u})~~. \ee One can also write: \be Tr_e({\hat
855: u})=\int_{L}{str_e({\hat u})}~~, \ee upon extending $str_e$ to
856: ${\cal H}_e$ through: \be str_e(\rho\otimes {\hat f})=\rho~str_e({\hat
857: f})~~. \ee
858:
859:
860: We next introduce a (nondegenerate) bilinear form on ${\cal H}_e$
861: through: \be \langle {\hat u}, {\hat v}\rangle_e:=Tr_e({\hat u}*{\hat
862: v}) =\int_{L}{str_e({\hat u}*{\hat v})}~~. \ee Property (\ref{tre})
863: assures graded-symmetry of this form with respect to the total degree
864: on the extended boundary space: \be \langle {\hat u}, {\hat
865: v}\rangle_e=(-1)^{deg {\hat u}~deg {\hat v}} \langle {\hat v}, {\hat
866: u}\rangle_e~~. \ee It is also easy to check invariance of the
867: extended bilinear form with respect to the extended boundary product:
868: \be \langle {\hat u}*{\hat v},{\hat w}\rangle_e=\langle {\hat u},
869: {\hat v}*{\hat w}\rangle_e~~. \ee We finally note the worldsheet
870: charge selection rule: \be
871: \label{dc}
872: \langle {\hat u},{\hat v}\rangle_e=0~~{\rm~unless~}~|{\hat u}| +|{\hat
873: v}|=3~~. \ee Since the supertrace only couples the component
874: $u_{mn}$ to $v_{nm}$, one has in fact separate selection rules for the
875: gradings $\Delta$ and $rk$: \be
876: \label{selrules}
877: \langle {\hat u},{\hat v}\rangle_e=0~~{\rm~unless~}~rk{\hat u}
878: +rk{\hat v}=3~ {\rm~and~}~\Delta({\hat u})+\Delta({\hat v})=0~~, \ee
879: for bi-homogeneous elements $u$ and $v$.
880:
881:
882: \subsubsection{Expression of the extended product in terms
883: of the wedge product and `twisted wedge product' of Grassmann-valued
884: forms with coefficients in $End({\bf E})$}
885:
886: It is possible to express the extended boundary data discussed above
887: in somewhat more familiar language as follows. Upon regarding ${\cal
888: H}_e$ as the tensor product $\Omega^*(L){\hat \otimes}{\hat {\cal
889: E}}$, one has the usual wedge product: \be
890: \label{wedge}
891: {\hat u}\wedge {\hat v}=(\rho\wedge \eta)\otimes ({\hat f}\circ{\hat
892: g})~~, \ee which uses the composition (\ref{hatfg_circ}) of
893: Grassmann-valued bundle morphisms. One can also define a `twisted
894: wedge product' by\footnote{The twisted wedge product $\bweft$ was used
895: in \cite{sc}, where it was denoted simply by $\wedge$, since it is
896: the natural extension of the wedge product of bundle-valued forms to
897: the graded case. In the present paper, we reserve the notation
898: $\wedge$ (later simply written as juxtaposition) for the usual wedge
899: product built with the ordinary composition of morphisms.}: \be
900: \label{hatuv_wedge}
901: {\hat u}\bwe{\hat v}=(\rho\wedge \eta)\otimes ({\hat f}{\hat g})~~.
902: \ee The idea behind this definition is that, since ${\hat u}$ and
903: ${\hat v}$ are forms with Grassmann-valued coefficients in a {\em
904: graded} bundle, it is natural to consider a `wedge product' which
905: includes multiplication with respect to the natural product
906: (\ref{hatfg}) on the coefficient algebra ${\hat {\cal E}}$.
907:
908: It is easy to see that: \be {\hat u}\bwe{\hat v}=(-1)^{g({\hat
909: u})\Delta({\hat v})}{\hat u}\wedge {\hat v} \ee
910:
911:
912:
913: With these definitions, equation (\ref{star1}) gives: \be
914: \label{star}
915: {\hat u}*{\hat v}=(-1)^{(g({\hat u})+\Delta({\hat u}))rk {\hat v}}
916: {\hat u}\bwe {\hat v}=(-1)^{g({\hat u})|{\hat v}|+\Delta({\hat u})rk
917: {\hat v}} {\hat u}\wedge {\hat v}~~, \ee where we extended the
918: grades $\Delta, g$ and $rk$ to ${\cal H}_e$ in the obvious manner: $
919: \Delta(\rho\otimes f\otimes \alpha):=\Delta(f)$, $g(\rho\otimes
920: f\otimes \alpha):=g(f)$, $rk(\rho\otimes f\otimes \alpha):=rk\rho$.
921:
922:
923: As in the previous section, the product (\ref{hatuv_wedge}) allows for
924: a formulation of the extended algebraic structure in perhaps more
925: familiar language. Upon viewing ${\cal H}_e$ as $\Omega^*(L)\otimes
926: {\hat {\cal E}}$, one can locally expand its elements in the form: \be
927: \label{gvf}
928: {\hat u}= \sum_{k=0}^3{dx^{\alpha_1}\wedge ... \wedge
929: dx^{\alpha_k}{\bf U}^{(k)}_{\alpha_1..\alpha_k}(x)} \ee where the
930: coefficients ${\bf U}_{\alpha_1..\alpha_k}(x)$ are Grassmann-valued
931: sections of $End({\bf E})$, i.e. elements of ${\hat {\cal E}}$. Then
932: the twisted wedge product (\ref{hatuv_wedge}) reads: \be
933: \label{doo}
934: {\hat u}\bwe {\hat v}=\sum_{k,l=0}^3{ dx^{\alpha_1}\wedge ... \wedge
935: dx^{\alpha_k}\wedge dx^{\beta_1}\wedge ... \wedge dx^{\beta_l}({\bf
936: U}^{(k)}_{\alpha_1..\alpha_k}(x) {\bf
937: V}^{(l)}_{\beta_1..\beta_l}(x))}~~, \ee {\em where the product
938: ${\bf U}^{(k)}_{\alpha_1..\alpha_k}(x) {\bf
939: V}^{(l)}_{\beta_1..\beta_l}(x)$ is defined as in (\ref{hatfg})}.
940:
941:
942:
943: We finally note that one can also define a naive extension of the
944: product $\bullet$ of (\ref{dot}) to Grassmann-valued forms with
945: coefficients in $End({\bf E})$: \be \label{bullet_extended} {\hat
946: u}\bullet {\hat v}=(u\bullet v)\otimes ({\hat f}{\hat
947: g})=(-1)^{\Delta({\hat u}) rk {\hat v}}{\hat u}\bwe {\hat
948: v}=(-1)^{\Delta({\hat u}) rk {\hat v}+g({\hat u})\Delta({\hat
949: v})}{\hat u}\wedge {\hat v}~~, \ee where $({\hat f}{\hat g})$ is
950: again defined as in (\ref{hatfg}). With this definition, one obtains:
951: \be {\hat u}*{\hat v}=(-1)^{g({\hat u})rk {\hat v}}{\hat u}\bullet
952: {\hat v}~~. \ee
953:
954:
955:
956: \subsubsection{The extended differential}
957:
958: The differential on $\Omega^*(L, End(E))$ extends to ${\cal H}_e$ in
959: the obvious manner: \be
960: \label{de}
961: d{\hat u}=du\otimes \alpha~~, \ee on decomposable elements ${\hat u}=u
962: \otimes \alpha$. The symbol $d$ in the second equality is the
963: differential on ${\cal H}$. The differential (\ref{de}) is a graded
964: derivation of $*$, with respect to the total degree: \be d({\hat
965: u}*{\hat v})=(d{\hat u})*{\hat v}+(-1)^{deg {\hat u}} {\hat
966: u}*d{\hat v}~~. \ee It is also easy easy to check that: \be \langle
967: d{\hat u},{\hat v}\rangle_e=-(-1)^{deg {\hat u}}\langle {\hat u},
968: d{\hat v} \rangle_e~~. \ee
969:
970: \subsection{The extended action and restricted odd symplectic form}
971:
972: The data of the previous subsection allows one to write an extended
973: action on the subspace ${\cal H}_e^1=\{ {\hat \phi}\in {\cal H}_e |
974: deg {\hat \phi}={\hat 1}\}$ of ${\cal H}_e$: \be
975: \label{extended_action}
976: S_e({\hat \phi})= \int_L{ str_e\left[\frac{1}{2}{\hat \phi} *d{\hat
977: \phi}+ \frac{1}{3}{\hat \phi}*{\hat \phi}*{\hat \phi}
978: \right]}~~. \ee This action was written down in \cite{sc} by
979: analogy with the general extension procedure discussed in
980: \cite{Gaberdiel, Zwiebach_open} \footnote{As the latter translates in
981: our conventions and with certain modifications.}. It is one of the
982: purposes of this paper to show that $S_e$ plays the role of BV action
983: for the string field theory (\ref{action}).
984:
985:
986: On ${\cal H}_e^1$, the product (\ref{star}) agrees with the
987: multiplication $\bullet$ of (\ref{dot}) precisely when the worldsheet
988: $U(1)$ charge of the first factor is odd, in which case its Grassmann
989: degree is even (see equation (\ref{star1})). If we extend the
990: evaluation map (\ref{evG}) to a map from ${\cal H}_e$ to ${\cal H}$
991: defined through: \be ev_G(u\otimes \alpha):=ev_G(\alpha)u=\alpha_0u~~,
992: \ee we find that: \be {\cal H}^1=ev_G(M_0)~~, \ee where $M_0$ is the
993: subspace of ${\cal H}_e^1$ given by: \be
994: \label{M0}
995: M_0=\{{\hat \phi}\in {\cal H}_e^1||{\hat \phi}|=1\}~~. \ee This
996: implies that the restriction of $S_e$ to $M_0$ is related to the
997: unextended action of (\ref{action}) through: \be
998: \label{restriction}
999: ev_G(S_e({\hat \phi}))=S(ev_G({\hat \phi}))~{\rm~for~}
1000: {\hat \phi}\in M_0~~. \ee
1001:
1002: \noindent For later reference, we note that the restriction:
1003: \be
1004: \label{omega0}
1005: \omega_0:=\langle . , .\rangle_e|_{{\cal H}_e^1\times {\cal H}_e^1}
1006: \ee of the extended bilinear form to the subspace ${\cal H}_e^1$ is an
1007: {\em antisymmetric} nondegenerate bilinear form whose values are
1008: Grassmann-odd numbers. This follows from the properties of $\langle .
1009: , . \rangle_e$ discussed in the previous section. Let us express
1010: $\omega_0$ in terms of the wedge products defined in (\ref{wedge}) and
1011: (\ref{hatuv_wedge}). Upon using relation (\ref{star}) and the fact
1012: that $deg {\hat u} =deg {\hat v} ={\hat 1}$, one obtains: \be
1013: \label{omega0_rescaling}
1014: \omega_0({\hat u}, {\hat v})=(-1)^{(1-rk {\hat u})rk {\hat
1015: v}}\int_{L}{ str_e({\hat u} \bwe {\hat v})}~~, \ee which in view
1016: of the selection rules (\ref{selrules}) also reads: \be
1017: \label{omega0_wedge}
1018: \omega_0({\hat u}, {\hat v})=(-1)^{rk {\hat v}} \int_{L}{str_e({\hat
1019: u}\bwe {\hat v})}=(-1)^{rk {\hat v}+g({\hat u}) \Delta({\hat v})}
1020: \int_{L}{str_e({\hat u}\wedge {\hat v})}~~. \ee Moreover, we notice
1021: that $\omega_0({\hat u}, {\hat v})$ vanishes unless $g({\hat
1022: u})+g({\hat v})$ is odd (this follows from the constraints $deg
1023: {\hat u}=deg {\hat v}=odd$ and the selection rule (\ref{dc})); this
1024: establishes that $\omega_0$ takes Grassmann-odd values. These
1025: observations will be useful in Section 7.
1026:
1027: \subsection{Superspace formulation of the extended action}
1028:
1029: The extended action (\ref{extended_action}) can be formulated in
1030: superspace language as follows. Consider the supermanifold ${\cal
1031: U}:=\Pi TL$ obtained by applying parity reversal on the fibers of
1032: the tangent bundle of $L$. Superfunctions defined on ${\cal L}$ are
1033: (Grassmann-valued) superfields on $L$, with odd superspace coordinates
1034: $\theta^j$ associated with the tangent vectors
1035: $\partial_j=\frac{\partial}{\partial x^j}$ defined by a coordinate
1036: system $\{x^j\}_{j=1..3}$ on $L$.
1037:
1038: Grassmann-valued forms with coefficients in $End({\bf E})$ can be put
1039: into correspondence with superfields valued in $End({\bf E})$ upon
1040: identifying ${\hat u}$ of equation (\ref{gvf}) with: \be
1041: \label{superfield}
1042: {\bf U}(x,\theta)=\sum_{k=0}^3{\theta^{\alpha_1}...\theta^{\alpha_k}
1043: {\bf U}^{(k)}_{\alpha_1...\alpha_k}(x)}~~. \ee In order to
1044: translate the action (\ref{extended_action}) in superspace language,
1045: one must use the somewhat unusual convention that the components ${\bf
1046: U}^{(k)}_{\alpha_1...\alpha_k}(x)\in {\hat {\cal E}}$ of superfields
1047: of the form (\ref{superfield}) are multiplied with the product
1048: (\ref{hatfg}) and that the sign obtained when commuting $\theta^j$
1049: with such a component \ is $(-1)^{\sigma({\bf U})}$. With these
1050: conventions, one can check \cite{sc} that multiplication of
1051: superfields reproduces the product (\ref{star}) for the associated
1052: forms, which allows one to write the extended action as: \be
1053: \label{extended_action_superspace}
1054: S_e({\bf \Phi})= \int_L d^3x\int{d^3\theta str_e\left[\frac{1}{2}{\bf
1055: \Phi} {\bf D}{\bf \Phi}+ \frac{1}{3}{\bf \Phi}{\bf \Phi}{\bf
1056: \Phi} \right]}~~, \ee where ${\bf \Phi}$ is the superfield
1057: associated with ${\hat \phi}$ and ${\bf D}=\theta^j{\partial_j}$. In
1058: this paper, we shall only use the differential form language of
1059: (\ref{extended_action}).
1060:
1061: \subsection{The underlying superbundle and the physical role of $\Z$-grading}
1062:
1063: It is clear from our construction that the extended boundary product
1064: $*$ (and thus the extended action (\ref{extended_action})) depend only
1065: on the mod two reduction of the relative D-brane grade $\Delta$. To
1066: formalize this, let us consider the reduction $E=E_{even}\oplus
1067: E_{odd}$ of the $\Z$-grading of ${\bf E}$, where: \bea
1068: E_{even}=\oplus_{n=even}{E_n}~~,~~E_{odd}=\oplus_{n=odd}{E_n}~~. \eea
1069: This allows us to view ${\bf E}$ as a superbundle \cite{Quillen},
1070: while forgetting the finer data associated with the $\Z$-grading. It
1071: is clear that the boundary product, extended boundary product and
1072: extended action depend only on this superbundle structure. In
1073: particular, one has only two classes of extended actions. The first
1074: corresponds to the case $E_{even}=0$ or $E_{odd}=0$ and can be
1075: recognized as the extended Chern-Simons action coupled to the bundle
1076: ${\bf E}=E_{odd}$ or ${\bf E}=E_{even}$. The second corresponds to
1077: $E_{even}\neq 0$ and $E_{odd}\neq 0$ and can be viewed as an extended
1078: version of the `supergroup Chern-Simons action' \cite{supergroupCS}
1079: coupled to ${\bf E}=E_{odd}\oplus E_{even}$. This agrees with ideas
1080: proposed in \cite{Vafa_cs}.
1081:
1082: The D-brane grade plays the role of specifying the finer $\Z$-grading
1083: given by the worldsheet $U(1)$ charge. It is this piece of data which
1084: defines the subspace $\{{\hat \phi}\in {\cal H}^1_e||{\hat \phi}|=1\}$
1085: on which the extended action (\ref{extended_action}) reduces to the
1086: unextended functional (\ref{action}). As we shall see in the next
1087: section, the extended theory can be viewed as a classical BV system,
1088: with $S_e$ playing the role of tree-level master action. From the BV
1089: perspective, the choice of D-brane grading is what specifies both the
1090: BV ghost number and the so-called {\em classical gauge}. In
1091: particular, two theories which have the same underlying superbundle
1092: but distinct choices of $\Z$-grading have the same tree-level BV
1093: actions but correspond to different choices for these two pieces of
1094: data. Since the ghost grading is physically relevant (in particular,
1095: as we recall at the end of Section 4.1.3, it specifies the algebra of
1096: classical on-shell gauge-invariant observables, given the other
1097: tree-level BV data), different choices of D-brane grading lead to {\em
1098: different} physical theories, in spite of the fact that the
1099: difference may not be manifest in the BV action itself. This is the
1100: crucial conceptual distinction between our approach and the proposals
1101: of \cite{Vafa_cs}.
1102:
1103:
1104:
1105: \section{The extended action as a classical master action}
1106:
1107: In this section we show that the extended action satisfies the
1108: classical master equation with respect to a BV bracket induced by an
1109: odd symplectic form associated to the extended bilinear form. We also
1110: show that $S_e$ reduces to the unextended action $S$ in a certain
1111: `classical gauge'. These extremely general results are valid for an
1112: arbitrary collection of graded branes (of distinct grades) wrapping
1113: the cycle $L$, and hold for any topology of the cycle and background
1114: connection.
1115:
1116: Our approach uses a certain variant of the geometric BV framework
1117: developed in \cite{Witten_antibracket, Henneaux_geom, Khudaverdian}
1118: and \cite{Kontsevich_Schwarz,Schwarz_geom, Schwarz_semiclassical,
1119: Schwarz_symms, Schwarz_superanalogues}. This formalism has the
1120: advantage that it is computationally compact and well-adapted to
1121: topologically-nontrivial situations. In fact, it turns out that the
1122: current version of the geometric BV formalism is incomplete, and we
1123: shall have to extend it in an appropriate manner. The problem is that
1124: the geometric description presented in the references just cited does
1125: not keep track of the BV ghost number. Indeed, the geometric formalism
1126: is usually discussed in terms of a P-manifold, i.e. a supermanifold
1127: endowed with an odd symplectic form. While this correctly considers
1128: the Grassmann parity of various BV fields, it fails to account for the
1129: ghost grading. This $\Z$-grading on the space of superfunctions plays
1130: a crucial role in the bottom-up (or homological) approach to BV
1131: quantization \cite{FHST, FH, Stasheff_bv, Henneaux_lectures, Gomis}
1132: and in many questions of direct physical significance. For example, it
1133: is a central result of the BV formalism that the BRST cohomology in
1134: {\em ghost} degree zero computes the space of on-shell gauge-invariant
1135: observables of the system. Since the current geometric formulation
1136: does not consider the ghost grading, it does not allow for a
1137: description of this (and other) fundamental results. In particular,
1138: two {\em distinct} BV systems can have the same supermanifold
1139: interpretation and the same BV action, so they cannot be distinguished
1140: by the current geometric approach. This can lead to confusion when
1141: applied to our models. To avoid such problems, one must refine the
1142: geometric formulation by explicitly including the ghost grading. This
1143: can be done with the help of $\Z$-{\em graded supermanifolds}, which
1144: were recently discussed in \cite{Voronov}. We start with a brief
1145: account of graded supermanifolds and continue by presenting a
1146: $\Z$-graded version of the geometric BV formalism. We then apply it to
1147: our theories in order to obtain a complete description of the
1148: associated BV systems. Most of this section is formulated in an
1149: entirely general manner, and may be of independent interest for
1150: foundational studies of BV quantization.
1151:
1152: \subsection{Covariant description of classical BV systems in the graded
1153: supermanifold approach}
1154:
1155: \subsubsection{Supermanifold conventions}
1156:
1157: We remind the reader that there are two major proposals for a rigorous
1158: definition of supermanifolds, the so-called Berezin-Konstant
1159: \cite{Berezin} and DeWitt-Rogers \cite{DeWitt, Rogers} theories. The
1160: major difference between the two is that the definition of a
1161: DeWitt-Rogers supermanifold requires the choice of an auxiliary
1162: Grassmann algebra $G$, the `algebra of constants'. Berezin's approach
1163: is based on an `intrinsic' sheaf of superalgebras, which leads to a
1164: formulation in terms of ringed spaces (`superschemes'). In this
1165: theory, the manifold has only even points, while the odd coordinates
1166: appear as a form of `algebraic fuzz'. By contrast, the DeWitt-Rogers
1167: theory constructs supermanifolds which possess both even and odd
1168: points, thus leading to a geometrization of the odd directions; this
1169: geometric description of odd coordinates depends on the algebra of
1170: constants $G$. It is a basic result that the set of $G$-valued points
1171: (defined in a manner similar to that employed in scheme theory) of a
1172: Berezin supermanifold defines a DeWitt-Rogers supermanifold
1173: \footnote{More precisely, it defines a so-called
1174: $H^\infty$-supermanifold \cite{Rogers}.} (though not every
1175: DeWitt-Rogers supermanifold can be obtained in this manner
1176: \cite{Rogers}). Since the extended theory (\ref{extended_action})
1177: incorporates the auxiliary Grassmann algebra $G$, we shall employ the
1178: formalism due to DeWitt and Rogers. Thus all of our supermanifolds
1179: are understood in the DeWitt-Rogers sense \footnote{The formalism we
1180: use can in fact be applied to so-called $G^\infty$-supermanifolds,
1181: which are a generalization of $H^\infty$-supermanifolds
1182: \cite{Rogers}.}.
1183:
1184: Given a (complex) DeWitt-Rogers supermanifold $M$ (modeled over the
1185: algebra of constants $G$), its tangent space at a point $p$ is a
1186: super-bimodule $T_pM$ over $G$ (see Appendix A), whose left and right
1187: module structures are compatible: \be \alpha
1188: X_p=(-1)^{\epsilon_\alpha\epsilon(X_p)}X_p\alpha~~, \ee for $\alpha$
1189: an element of $G$ and $X_p$ an element of $T_pM$. We make the
1190: convention that $\epsilon$ denotes the $\Z_2$-degree of an element in
1191: the space to which it belongs. Thus $\epsilon_\alpha$ is the
1192: Grassmann parity of $\alpha$, while $\epsilon(X_p)$ is the parity of
1193: $X_p$ with respect to the $\Z_2$-grading on $T_pM$. The disjoint union
1194: of the tangent spaces gives the tangent bundle $TM$.
1195:
1196: Globally defined $G$-valued functions on $M$ form a commutative
1197: $\Z_2$-graded ring ${\cal F}(M,G)$ with respect to pointwise
1198: multiplication, with: \be \epsilon_F={\hat 0}~{\rm~if~}F(p)\in
1199: G_0~{\rm~for~all~}p\in M~~,~~ \epsilon_F={\hat 1}~{\rm~if~}F(p)\in
1200: G_1~{\rm~for~all~}p\in M~~. \ee This is also a left- and right-
1201: $\Z_2$-graded algebra over the ring of constants $G$. Left and right
1202: derivations of this algebra give so-called left and right vector
1203: fields on $M$ (this is discussed in more detail in Appendix B). The
1204: spaces of left/right vector fields are graded in the obvious manner,
1205: with even and odd derivations corresponding to even and odd vector
1206: fields. It is customary to identify left and right derivations, and
1207: we shall do so in the following (see Appendix B for details of this
1208: construction). This allows us to speak simply about vector fields.
1209: With this convention, a vector field $X$ can act both to the left (as
1210: a left derivation) and to the right (as a right derivation), with the
1211: two actions related by applying the sign rule. We shall indicate the
1212: left and right actions by superscript arrows pointing respectively to
1213: the right and left. For every function $F$ we thus have: \be
1214: \label{done}
1215: F\stackrel{\leftarrow}{X}=
1216: (-1)^{\epsilon_F\epsilon_X}\stackrel{\rightarrow}{X}F:=dF(X)~~, \ee
1217: where $dF$ is by definition the differential of $F$. This is a
1218: complex-linear function defined on the space of vector fields, which
1219: is also $G$-linear in the obvious sense: \be
1220: \label{lr_linear}
1221: dF(X\alpha)=dF(X)\alpha~~,~~dF(\alpha~X)=(-1)^{\epsilon_\alpha\epsilon_F}
1222: \alpha~dF(X)~~. \ee It induces $G$-linear functionals $d_pF$ on each
1223: of the tangent spaces $T_pM$.
1224:
1225: The space of vector fields is endowed with a Lie bracket, which in
1226: terms of the action on functions is given by: \be
1227: F\stackrel{\longleftarrow}{[X,Y]}:=-
1228: F(\stackrel{\leftarrow}{X}\stackrel{\leftarrow}{Y}-
1229: (-1)^{\epsilon_X\epsilon_Y}\stackrel{\leftarrow}{Y}
1230: \stackrel{\leftarrow}{X}) \Longleftrightarrow
1231: \stackrel{\longrightarrow}{[X,Y]}F=
1232: (\stackrel{\rightarrow}{X}\stackrel{\rightarrow}{Y}-
1233: (-1)^{\epsilon_X\epsilon_Y}
1234: \stackrel{\rightarrow}{Y}\stackrel{\rightarrow}{X})F~~. \ee This
1235: operation satisfies $\epsilon_{[X,Y]}=\epsilon_X+\epsilon_Y$ and is
1236: graded-symmetric and $G$-bilinear: \be
1237: \label{Lie}
1238: \left[X,Y\right]~~=-(-1)^{\epsilon_X\epsilon_Y}\left[Y,X\right]~~,~~
1239: \left[\alpha X,Y\beta\right] =~~~ \alpha \left[X,Y\right]\beta~~. \ee
1240: It also satisfies the graded Jacobi identity: \be
1241: [[X,Y],Z]+(-1)^{\epsilon_X(\epsilon_Y+\epsilon_Z)}[[Y,Z],X]+
1242: (-1)^{\epsilon_Z(\epsilon_X+\epsilon_Y)}[[Z,X],Y]=0~~. \ee Endowed
1243: with this commutator, the space of vector fields becomes a
1244: $\Z_2$-graded Lie algebra (Lie superalgebra).
1245:
1246:
1247: The space of functionals $\eta(X)$ obeying $G$-linearity constraints
1248: of the type (\ref{lr_linear}) forms a $\Z_2$-graded $G$-bimodule in
1249: the obvious manner. This is the space $\Omega^1(M)$ of one-forms on
1250: $M$. One defines higher rank forms with the help of the wedge product:
1251: \be
1252: \label{fwedge}
1253: \rho\wedge \eta=\rho\otimes
1254: \eta-(-1)^{\epsilon_\rho\epsilon_\eta}\eta\otimes \rho~~, \ee which
1255: has the graded symmetry property: \be \rho\wedge
1256: \eta=(-1)^{\epsilon_\rho\epsilon_\eta+1}\eta\wedge \rho~~. \ee Upon
1257: taking iterated wedge products one obtains forms of arbitrary ranks
1258: and (\ref{fwedge}) extends to such forms in the obvious fashion. One
1259: also has an exterior differential, obtained by extending (\ref{done}).
1260: In particular, a two-form $\omega(X,Y)$ on $X$ is a $G$-valued
1261: complex-bilinear functional on vector fields which has the properties:
1262: \bea
1263: \label{forms}
1264: \epsilon_{\omega(X,Y)}~~~
1265: &=&~~~~~\epsilon_X+\epsilon_Y+\epsilon_\omega~~\nn\\
1266: \omega(\alpha X,Y\beta)&=&(-1)^{\epsilon_\alpha \epsilon_\omega}\alpha
1267: \omega(X,Y)\beta~~\\
1268: \omega(X,Y)~~&=&(-1)^{\epsilon_X\epsilon_Y+\epsilon_\omega}\omega(Y,X)~~.\nn
1269: \eea The quantity $\epsilon_\omega\in \{{\hat 0},{\hat 1}\}$ defines
1270: its parity: $\omega$ is even if $\epsilon_\omega={\hat 0}$ and odd if
1271: $\epsilon_\omega={\hat 1}$. A two-form $\omega$ is called {\em
1272: symplectic} if it is nondegenerate and closed $(d\omega=0$).
1273:
1274:
1275: Local coordinates give independent Grassmann-valued functions $z^a$
1276: defined on an open subset of $M$, whose parities we denote by
1277: $\epsilon_a:=\epsilon(z_a)$. Given such coordinates, one has
1278: locally-defined vector fields
1279: $\partial^l_a=(-1)^{\epsilon_a}\partial^r_a$ (of parity $\epsilon_a$),
1280: which are uniquely determined by: \be
1281: \stackrel{\rightarrow}{\partial^l_a}z^b=
1282: z^b\stackrel{\leftarrow}{\partial^r_a}=\delta^b_a~~. \ee Their action
1283: on a function $F$ defines its left and right derivatives: \be
1284: \stackrel{\rightarrow}{\partial^l_a}F= \frac{\partial_lF}{\partial
1285: z^a}=(-1)^{\epsilon_F\epsilon_a}dF(\partial^l_a)~~,~~
1286: F\stackrel{\leftarrow}{\partial^r_a}=\frac{\partial_rF}{\partial z^a}=
1287: dF(\partial^r_a)~~. \ee For the coordinate functions one obtains: \be
1288: dz^a(\partial^r_b)=(-1)^{\epsilon_a}dz^a(\partial^l_b)=\delta^a_b~~.
1289: \ee This allows us to write $dF$ in the form: \be
1290: \label{dF}
1291: dF=\frac{\partial_rF}{\partial
1292: z^a}dz^a=dz^a\frac{\partial_lF}{\partial z^a}
1293: =(-1)^{\epsilon_a(\epsilon_F+1)} \frac{\partial_lF}{\partial
1294: z^a}dz^a~~. \ee Given a vector field $X$, one can expand it locally
1295: as: \be X=X^a_l\partial^l_a=\partial^r_aX^a_r~~, \ee which defines its
1296: left and right coefficients $X^a_l$ and $X^a_r$ (=locally defined
1297: Grassmann-valued functions). Equation (\ref{dF}) then gives: \be
1298: dF(X)=\frac{\partial_rF}{\partial z^a}X^a_r=
1299: (-1)^{\epsilon_X\epsilon_F}X^l_a\frac{\partial_lF}{\partial z^a}~~.
1300: \ee Let us next consider the local expression of an {\em odd}
1301: symplectic form $\omega$. If one defines its coefficients through: \be
1302: \label{omega_components}
1303: \omega_{ab}:=\omega(\partial^r_a, \partial^r_b)=
1304: (-1)^{\epsilon_a+\epsilon_b}\omega(\partial^l_a, \partial^l_b)
1305: =(-1)^{\epsilon_a}\omega(\partial^l_a, \partial^r_b)
1306: =(-1)^{\epsilon_b}\omega(\partial^r_a, \partial^l_b)~~, \ee then it is
1307: easy to check that: \be \omega=-\frac{1}{2}\omega_{ab}dz^b\wedge
1308: dz^a~~ \ee (note the reversed order in the wedge product). Its value
1309: on an arbitrary pair of vector fields then follows from the
1310: bi-linearity property listed in (\ref{forms}): \be
1311: \omega(X,Y)=(-1)^{(\epsilon_X+\epsilon_a)\epsilon_b}\omega_{ab}X^a_rY^b_r~~.
1312: \ee Definition (\ref{omega_components}) and relations (\ref{forms})
1313: imply the properties: \be
1314: \epsilon(\omega_{ab})=\epsilon_a+\epsilon_b+1~~,~~
1315: \omega_{ab}=-(-1)^{\epsilon_a\epsilon_b}\omega_{ba}~~. \ee
1316:
1317: \subsubsection{$\Z$-graded supermanifolds}
1318:
1319: The collection ${\cal F}=({\cal F}(U,G))$ of $G$-valued functions
1320: defined on open subsets $U$ of $G$ forms a sheaf of superalgebras with
1321: respect to the $\Z_2$-grading given by $\epsilon$. A {\em
1322: $\Z$}-graded supermanifold \cite{Voronov} is a supermanifold endowed
1323: with a $\Z$-grading $s$ on this sheaf. This $\Z$-grading is required
1324: to be compatible with pointwise multiplication $s(FG)=s(F)+s(G)$ and
1325: with restriction from an open set to its open subsets. We also require
1326: $s(\alpha F)=s(F\alpha)=s(F)$ for $\alpha\in G$. The $\Z_2$-grading
1327: $\epsilon$ need not be the mod~2 reduction of $s$; in fact, this is
1328: almost never the case if one works with DeWitt-Rogers
1329: supermanifolds\footnote{The reason is that $\epsilon$ must satisfy
1330: $\epsilon(F\alpha)=\epsilon_F+\epsilon_\alpha$, while $s$ satisfies
1331: $s(F\alpha)=s(F)$. Hence $\epsilon =s~(mod~2)$ would require
1332: $\epsilon_\alpha=0$ for all $\alpha\in G$, which is only possible if
1333: $G$ has no odd generators.}.
1334:
1335: A $\Z$-grading on ${\cal F}$ can be specified by giving an atlas
1336: $\{(U, z^a_U)\}$ of local coordinates and picking integer grades
1337: $s_a^U$ for $z^U_a$ such that the change of coordinates from $z^U_a$
1338: to $z^V_a$ (when $U$ intersects $V$) is compatible with these degrees.
1339: For simplicity, let us restricts the elements of ${\cal F}(U,G)$ to be
1340: functions which are polynomial in coordinates\footnote{One can also
1341: consider formal power series, which gives a formal $\Z$-graded
1342: supermanifold. If one wishes to extend this beyond formal power
1343: series, one has to deal with issues of convergence, which we wish to
1344: avoid.}. Such a function has the form: \be
1345: \label{function}
1346: F(p)=\sum_{k=1}^N\sum_{a_1..a_k}{\alpha_{a_1...a_k}
1347: z^{a_1}_U(p)...z^{a_k}_U(p)}\Leftrightarrow
1348: F=\sum_{k=1}^N\sum_{a_1..a_k}{\alpha_{a_1...a_k}
1349: z^{a_1}_U...z^{a_k}_U}~~, \ee where $\alpha_{a_1..a_k}$ are elements
1350: of $G$ and the sum runs over monomials of degree smaller than some
1351: positive integer $N$. We extend $s_a$ to a $\Z$-grading on ${\cal
1352: F}(U,G)$ by declaring that
1353: $s(z^{a_1}...z^{a_k})=s_{a_1}+...+s_{a_k}$. A function
1354: (\ref{function}) is $s$-homogeneous of degree $\sigma$ if all of the
1355: monomials appearing in its expansion satisfy
1356: $s(z^{a_1}...z^{a_k})=\sigma$. It is clear that this grading is
1357: compatible with pointwise multiplication and restriction to open
1358: subsets of $U$, and satisfies $s(\alpha F)=s(F\alpha)=s(F)$.
1359:
1360: If $V$ is another coordinate neighborhood in the distinguished atlas
1361: (such that $U$ intersects $V$), then on the intersection $U\cap V$ one
1362: can express $z^a_V$ as: \be z^a_V=z^a_V(z_U)~~, \ee where we assume
1363: that the transition functions are polynomial. The compatibility
1364: condition requires that $s_a$ coincide with the degree of the function
1365: $z^a_V(z_U)$, defined with respect to the coordinates $z^U$. This
1366: assures us that the degree of a function in ${\cal F}(U\cap V,G)$ does
1367: not depend on the coordinates in which it is computed, and thus we
1368: have a well-defined $\Z$-grading on the sheaf ${\cal F}$\footnote{To
1369: globalize this argument one needs to assume the existence of an
1370: appropriate partition of unity etc.}.
1371:
1372:
1373: Notice that the Grassmann coefficients $\alpha$ play no role the
1374: grading $s$, i.e. one can formally write $s(\alpha_{a_1..a_k})=0$.
1375: As mentioned above, the Grassmann grading $\epsilon_F$ of a function
1376: $F$ need not coincide with the mod~2 reduction of its $\Z$-grading.
1377: For example, if $F=\alpha z^{a_1}...z^{a_k}$, then
1378: $\epsilon_F=\epsilon_\alpha+\epsilon_{a_1}+...+\epsilon_{a_k}$, but
1379: $s_F=s_{a_1}+..+s_{a_k}$, so that $s_F (mod~2)$ may differ from
1380: $\epsilon_F$ even if one chooses $s_a$ such that $s_a
1381: (mod~2)=\epsilon_a$. This mismatch between $\Z_2$-grading and
1382: $\Z$-grading is due to the presence of the Grassmann algebra of
1383: constants $G$, and thus is an inescapable feature of working with
1384: DeWitt-Rogers supermanifolds. The $\Z$ and $\Z_2$-gradings $s$ and
1385: $\epsilon$ must be viewed as independent pieces of data.
1386:
1387: The integer grading on Grassmann-valued functions allows us to
1388: introduce $\Z$-gradings on the spaces of vector fields and
1389: differential forms. Since vector fields $X$ are complex-linear maps
1390: from ${\cal F}$ to itself, we shall say that $X$ is $s$-homogeneous of
1391: degree $\sigma$ if: \be
1392: \label{sX}
1393: s(F\stackrel{\leftarrow}{X})=s(F)+s_X~~ \ee for some integer $s_X$
1394: which defines the $s$-degree of $X$. It is clear that
1395: $s([X,Y])=s_X+s_Y$.
1396:
1397: We shall say that a local coordinate system is {\em $s$-homogeneous}
1398: if the associated coordinate functions $z^a$ are $s$-homogeneous
1399: elements of ${\cal F}$. In this case, we denote $s(z^a)$ by $s_a$.
1400: Given an $s$-homogeneous coordinate system, it is clear that
1401: $s(F\stackrel{\leftarrow}{\partial^r_a})=
1402: s(\stackrel{\rightarrow}{\partial^l_a}F) =s(F)-s_a$, which implies:
1403: \be s(\partial^r_a)=s(\partial^l_a)=-s_a~~. \ee This allows us to
1404: introduce a $\Z$-grading on the tangent spaces $T_pM$ by using
1405: $X_p=(\partial^r_a)_pX^r_a(p)$ and the rule $s(X^r_a(p))=0$ (since
1406: $X^a_r(p)$ is a Grassmann constant, i.e. an element of $G$). It is
1407: clear that this grading is independent of the choice of
1408: $s$-homogeneous coordinates; it can be defined more invariantly by
1409: considering localization of vector fields. This $\Z$-grading, as well
1410: as the $\Z$-grading (\ref{sX}) on vector fields, have no direct
1411: relation to the $\Z_2$-grading $\epsilon$.
1412:
1413: If $\eta$ is a linear functional on vector fields, then we define its
1414: $s$-degree through: \be s(\eta(X))=s(X)+s_\eta~~. \ee In
1415: $s$-homogeneous coordinates, the relation
1416: $dz^a(\partial^r_b)=\delta^a_b$ implies: \be s(dz^a)=s(z^a)=s_a~~.
1417: \ee Moreover, we obtain: \be
1418: s(dF)=s_F~~,{\rm~i.e.~~~}s(dF(X))=s_F+s_X~~, \ee and
1419: $s(\frac{\partial_lF}{\partial z^a})=s(\frac{\partial_rF}{\partial
1420: z^a})= s_F-s_a$.
1421:
1422: This grading extends to multilinear forms in the obvious manner. For
1423: a two-form, we have: \be s(\omega(X,Y))=s(X)+s(Y)+s_\omega~~ \ee
1424: (notice that $\omega(X,Y)$ is an element of ${\cal F}(M,G)$). In
1425: $s$-homogeneous local coordinates, this gives: \be
1426: s(\omega_{ab})=s_\omega -s_a-s_b~~, \ee where $\omega_{ab}$ is the
1427: function $z\rightarrow \omega_{ab}(z)$.
1428:
1429: \subsubsection{Basics of the geometric framework}
1430:
1431: We now give a brief outline of a $\Z$-graded version of the geometric
1432: BV formalism. This is a supergeometric version of the symplectic
1433: formalism of Hamiltonian mechanics, endowed with the supplementary
1434: data of an integer-valued grading (the ghost grading). Its starting
1435: point is a {\em graded P-manifold}, i.e. a graded DeWitt-Rogers
1436: supermanifold $M$ (modeled\footnote{In our case $n$ will be infinite,
1437: as always in field theory. We shall neglect the well-known problems
1438: with infinite-dimensional supermanifolds (see, for example,
1439: \cite{Schmitt}). In fact, we shall later apply this formalism to
1440: linear supermanifolds only, for which the treatment can be made
1441: rigorous in terms of Banach supermanifolds. The condition `modeled
1442: on $\R^{n,n}$' means that one has an equal number of even and odd
1443: coordinates; in our application, this can be formulated in terms of
1444: countable coordinate frames.} on $\R^{n,n}$), endowed with an {\em
1445: odd} symplectic form $\omega$ which is $s$-homogeneous of degree
1446: $s_\omega=-1$
1447:
1448: Given a P-manifold, the odd symplectic form allows one to define a
1449: (right) Hamiltonian vector field $Q_F$ associated with an arbitrary
1450: (Grassmann-valued) function $F$ on $M$: \be dF(X)=\omega(Q_F,X)~~.
1451: \ee This equation is sensible since both the left and right hand sides
1452: are linear with respect to the right $G$-module structure on vector
1453: fields; non-degeneracy of $\omega$ assures the existence of a unique
1454: solution.
1455:
1456: It is clear from this definition that: \be Q_{\alpha
1457: F}=(-1)^{\epsilon_\alpha}\alpha Q_F~~,~~Q_{F\alpha}=Q_F\alpha~~, \ee
1458: for any Grassmann constant $\alpha$.
1459:
1460: Since the symplectic form satisfies $\epsilon_\omega={\hat 1}$ and
1461: $s_\omega=-1$, the vector field $Q_F$ has parity
1462: $\epsilon_{Q_F}=\epsilon_F+{\hat 1}$ and ghost number $s_{Q_F}=s_F+1$.
1463: In local coordinates $z^a$, one has the expansions
1464: $Q_F=Q^a_{F,l}\partial^l_a=\partial^r_aQ^a_{F,r}$, with the
1465: components: \be
1466: Q^a_{F,l}=(-1)^{\epsilon_F+\epsilon_a+1}\frac{\partial_lF}{\partial
1467: z^b} \omega^{ba}~~,~~
1468: Q^a_{F,r}=(-1)^{\epsilon_a+1}\omega^{ab}\frac{\partial_rF}{\partial
1469: z^b}~~, \ee where $\omega^{ab}$ is the inverse of the matrix
1470: $\omega_{ab}$: \be
1471: \omega^{ab}\omega_{bc}=\omega_{cb}\omega^{ba}=\delta^a_c~~. \ee Note
1472: the properties: \be \epsilon(\omega_{ab})=\epsilon_a+\epsilon_b+1~~,
1473: ~~s(\omega^{ab})=s_a+s_b+1~~,~~
1474: \omega^{ab}=-(-1)^{(\epsilon_a+1)(\epsilon_b+1)}\omega_{ba}~~. \ee
1475:
1476: Given two functions $F,G$ on $M$\footnote{We sometimes use the symbol
1477: $G$ to denote a Grassmann-valued function on $M$. This should not
1478: be confused with the underlying Grassmann algebra, which is denoted
1479: by the same letter.}, we define their {\em odd Poisson bracket}
1480: (antibracket) through: \be
1481: \label{antibracket}
1482: \{F,G\}=-\omega(Q_F, Q_G)=-dF(Q_G)=(-1)^{(\epsilon_F+1)(\epsilon_G+1)}
1483: dG(Q_F)= \frac{\partial_rF}{\partial z^a}\omega^{ab}
1484: \frac{\partial_lG}{\partial z^b}~~. \ee One has
1485: $\epsilon_{\{F,G\}}=\epsilon_F+\epsilon_G+1$ and
1486: $s(\{F,G\})=s_F+s_G+1$. It is easy to check the properties: \bea
1487: \{FG,H\}=F\{G,H\}+(-1)^{\epsilon_F\epsilon_G}G\{F,H\}~~&,&~~
1488: \{F,GH\}=\{F,G\}H+(-1)^{\epsilon_G\epsilon_H}\{F,H\}G~~\nn\\
1489: \{\alpha F,G\beta\}=\alpha\{F,G\}\beta~~~&,&~~
1490: \{F,G\}=-(-1)^{(\epsilon_F+1)(\epsilon_G+1)}\{G,F\}~~. \eea as well
1491: as the odd graded Jacobi identity: \be
1492: \{\{F,G\},H\}+(-1)^{(\epsilon_F+1)(\epsilon_G+\epsilon_H)}\{\{G,H\},F\}+
1493: (-1)^{(\epsilon_H+1)(\epsilon_F+\epsilon_G)}\{\{H,F\},G\}~~. \ee
1494:
1495: In particular, the space of $G$-valued functions on $M$ forms an odd
1496: Lie superalgebra \footnote{An odd Lie superalgebra \cite{Manin} is
1497: simply the parity change of a Lie superalgebra. This is obtained by
1498: reversing the parity of all elements, while leaving the Lie bracket
1499: unchanged. Together with pointwise multiplication of functions, the
1500: antibracket endows the space ${\cal F}(M,G)$ with the structure of a
1501: so-called {\em odd Poisson algebra} or {\em Gerstenhaber algebra}.}
1502: with respect to the antibracket. Equation (\ref{antibracket}) shows
1503: that: \be
1504: \label{act}
1505: F\stackrel{\leftarrow}{Q_G}=-\{F,G\}\Longleftrightarrow
1506: ~\stackrel{\rightarrow}{Q_G}F=-(-1)^{\epsilon_F(\epsilon_G+1)}\{F,G\}~~.
1507: \ee Together with the Jacobi identity, this implies: \be
1508: \label{morphism}
1509: Q_{\{F,G\}}=[Q_F,Q_G]~~. \ee Thus the map $F\rightarrow Q_F$ acts as
1510: an `odd morphism' (the composition of a morphism of $\Z_2$-graded Lie
1511: algebras with parity change). This translates the odd Lie
1512: superalgebra language appropriate for functions into the $\Z_2$-graded
1513: Lie algebra language relevant for vector fields.
1514:
1515: In the context of BV quantization, the antibracket is interpreted as
1516: the BV bracket. For any function $F$, one has\footnote{Note that
1517: (\ref{antibracket}) implies $\omega(Q_F,Q_F)=-dF(Q_F)=0$ if $F$ is
1518: odd.}: \be \{F,F\}=-\omega(Q_F, Q_F)=-dF(Q_F)~~,~~
1519: Q_{\{F,F\}}=[Q_F,Q_F]~~. \ee Hence given an action (even function of
1520: ghost degree zero) $S_{BV}$ on our supermanifold, the classical master
1521: equation can be written in the equivalent forms: \be
1522: \label{master}
1523: \{S_{BV}, S_{BV}\}=0\Leftrightarrow QS_{BV}=0\Leftrightarrow
1524: \omega(Q,Q)=0\Leftrightarrow [Q,Q]=0\Leftrightarrow
1525: Q^2(F)=0~{~\rm~for~all~}F~~, \ee where we defined $Q:=Q_{S_{BV}}$. It
1526: is clear that $\epsilon(Q_F)={\hat 1}$ and $s(Q_F)=+1$. In
1527: particular, a classical BV system defines a so-called {\em
1528: QP-manifold} \cite{Kontsevich_Schwarz}, i.e. a P-manifold endowed
1529: with an odd nilpotent vector $Q$ field which preserves the odd
1530: symplectic form.
1531:
1532: We remind the reader that an odd vector field $Q$ on a supermanifold
1533: is called {\em nilpotent} (or {\em homological}) if it satisfies
1534: $[Q,Q]=0\Leftrightarrow Q^2F=0{\rm~for~all~}F$. Given such a vector
1535: field, the space ${\cal F}(M,G)$ of globally defined Grassmann-valued
1536: functions (viewed as a complex vector space) becomes a complex with
1537: respect to the differential $Q$. If the underlying supermanifold is
1538: $\Z$-graded, and if $Q$ has $\Z$-degree equal to $+1$, then $({\cal
1539: F}(M,G),Q)$ is a $\Z$-graded cochain complex. The main result of the
1540: bottom-up approach to BV quantization is that, for $Q=Q_{BV}$, the
1541: cohomology of this complex in integer degree zero computes the space
1542: of gauge-invariant functionals on the shell of the associated
1543: classical action (the relation between the BV action and the classical
1544: action is described in geometric terms in the next section). Since the
1545: space of such observables has direct physical meaning, it is clear
1546: that two BV systems which have different ghost gradings (but the same
1547: underlying manifold and odd symplectic form) must be considered as
1548: distinct. Otherwise, there would be no clear way of recovering the
1549: classical data from the geometric formalism -- in particular, one
1550: would reach the paradox that two classical systems with very different
1551: algebras of on-shell gauge-invariant observables are equivalent,
1552: provided that they differ `only' by the choice of ghost grading, a
1553: statement which is clearly incorrect. This observation justifies the
1554: need for a $\Z$-graded geometric formalism, and is crucial for a
1555: correct understanding of graded D-brane systems. We believe that a
1556: correct geometric description of BV systems must systematically
1557: consider the ghost grading. Below, we limit ourselves to
1558: re-formulating some basic results of the geometric framework (which
1559: will be needed in our application) in the graded manifold language.
1560:
1561: \subsubsection{Gauges and BRST transformations}
1562:
1563: In the geometric formalism, a {\em gauge} corresponds to the choice of
1564: a Lagrangian sub-supermanifold of $M$, i.e. a sub-supermanifold ${\cal
1565: L}$ whose total dimension is half of the total dimension of $M$
1566: \footnote{If a supermanifold is modeled on $\R^{p|q}$, then its total
1567: dimension is $p+q$.} and with the property that $\omega$ restricts
1568: to zero on ${\cal L}$. To make contact with the bottom-up approach,
1569: one must also assume that ${\cal L}$ is $s$-{\em homogeneous}, i.e.
1570: its tangent bundle $T{\cal L}$ decomposes as a direct sum of
1571: $s$-homogeneous subbundles of $TM|_{\cal L}$: \be T{\cal
1572: L}=\oplus_{s\geq 0}{TM|_{\cal L}(s)}~~, \ee where $TM|_{\cal L}(s)$
1573: is the subbundle of $TM|_{\cal L}$ consisting of elements of ghost
1574: degree equal to $s$ \footnote{It is easy to see that $TM=T_{+}M\oplus
1575: T_{-}M$ (where $T_{\pm}M=\oplus_{\pm s\geq 0}{T(s)}$) gives a
1576: Lagrangian decomposition of $TM$; this follows from the condition
1577: $s_\omega=-1$. An $s$-homogeneous Lagrangian submanifold is an
1578: integral submanifold for the Frobenius distribution $T_{+}M$. This
1579: distribution is clearly integrable, since $s([X,Y])=s(X)+s(Y)\geq 0$
1580: for all vector fields $X,Y$ satisfying $s(X), s(Y)\geq 0$.}.
1581:
1582: The path integral in this gauge is given by integrating
1583: $e^{-\frac{i}{\hbar }S_{BV}}|_{\cal L}=e^{-\frac{i}{\hbar }S_{\cal
1584: L}}$ along ${\cal L}$, where $S_{\cal L}:=S_{BV}|_{\cal L}$ is the
1585: restriction of $S_{BV}$ to ${\cal L}$. This is a global version of the
1586: usual description of gauges in terms of fields and antifields and
1587: gauge-fixing fermions. We shall follow common practice and omit the
1588: word `super' when talking about submanifolds of a supermanifold. We
1589: remind the reader that an odd vector field on $M$ can be viewed as an
1590: odd section of $TM$ or an even section of the parity changed bundle
1591: $\Pi TM$. It is sometimes convenient to work with even sections only,
1592: in which case the parity of a vector field is made clear by the
1593: presence or absence of parity change on the underlying bundle. We
1594: shall sometimes use this convention in what follows. Therefore, a
1595: section of a bundle will always mean an even section unless explicitly
1596: stated otherwise.
1597:
1598:
1599: Let us recall from \cite{Kontsevich_Schwarz} how the BRST
1600: transformations of the gauge-fixed action are realized in the
1601: geometric formalism. Choosing a gauge ${\cal L}$, one constructs a
1602: symmetry of $S_{\cal L}$ as follows. Upon restricting $Q$ to ${\cal
1603: L}$, one obtains a section of the bundle $\Pi TM|_{\cal L}$. In
1604: order to produce an (odd) vector field on ${\cal L}$, one considers
1605: the decomposition $TM|_{\cal L}=T{\cal L}\oplus N{\cal L}$, where: \be
1606: N{\cal L}=\oplus_{s<0}{TM|_{\cal L}(s)}~~. \ee It is clear from the
1607: condition $s_\omega=-1$ that this is a Lagrangian splitting of the
1608: restricted tangent bundle $TM|_{\cal L}$, i.e. $\omega$ vanishes on
1609: $T{\cal L}\times T{\cal L}$ and $N{\cal L}\times N{\cal L}$ and is
1610: non-degenerate on $T{\cal L}\times N{\cal L}$ and on $N{\cal L}\times
1611: T{\cal L}$. As explained in more detail below, this decomposition of
1612: $TM|_{\cal L}$ is related to the field-antifield split of the local
1613: formalism.
1614:
1615: One has a similar decomposition $\Pi TM|_{\cal L}=\Pi T{\cal L}\oplus
1616: \Pi N{\cal L}$ of the parity changed bundle. If $T, R$ are the
1617: associated projectors of $\Pi TM|_{\cal L}$ onto $\Pi T{\cal L}$ and
1618: $\Pi N{\cal L}$, then the relations: \be
1619: \label{qgen}
1620: q:=TQ|_{\cal L}~~,~~q^*:=RQ|_{\cal L} \ee define sections of $\Pi
1621: T{\cal L}$ and $\Pi N {\cal L}$ which give a decomposition of $Q$ on ${\cal
1622: L}$: \be
1623: \label{Qdec}
1624: Q|_{\cal L}=q+q^*~~. \ee It is clear that the operators $T$ and $R$
1625: are $s$-homogeneous of degree zero, so that the vector fields $q$ and
1626: $q^*$ are $s$-homogeneous of degree $+1$. More generally, vectors
1627: $u\in TM|_{\cal L}$ decompose as $u={\bf u}+{\bf u}^*$, with ${\bf
1628: u}\in T{\cal L}$ and ${\bf u}^*\in N{\cal L}$. Upon using this in
1629: the defining relation for $Q$, one obtains: \be dS_{BV}(u)=\omega(u,
1630: Q)=\omega({\bf u}, q^*)+ \omega({\bf u}^*, q)~~. \ee which combines
1631: with $dS_{BV}(u)=dS_{BV}({\bf u})+dS_{BV}({\bf u}^*)= dS_{\cal L}({\bf
1632: u})+dS_{BV}({\bf u}^*)$ to give: \be
1633: \label{first}
1634: dS_{\cal L}({\bf u})=\omega({\bf u}, q^*)~~,~~ dS_{BV}({\bf
1635: u}^*)=\omega({\bf u}^*, q)~~. \ee The second equation shows that
1636: the first order term in the Taylor expansion of $S_{BV}$ in antifields
1637: is proportional with $q$. The first implies that the value of $q^*$ at
1638: a point $p$ in ${\cal L}$ vanishes precisely when $p$ is critical for
1639: $S_{\cal L}$. Hence the critical set of $S_{\cal L}$ is the locus
1640: where $Q$ is tangent to ${\cal L}$. Combining this with equation
1641: (\ref{Qdec}) and using the nilpotence of $Q$ shows that $q$ squares to
1642: zero `on the shell of $S_{\cal L}$': \be [q,
1643: q]=0~~{\rm~on~}~Crit(S_{\cal L})~~. \ee
1644:
1645: We finally note from (\ref{qgen}) and (\ref{master}) that $q$
1646: generates a symmetry of the gauge fixed action: \be
1647: \stackrel{\rightarrow}{q}S_{\cal L}=0~~. \ee It is clear that $q$ is
1648: the BRST generator in the gauge $\cal L$.
1649:
1650:
1651:
1652: \subsubsection{The coordinate description}
1653:
1654:
1655: We now sketch how the local description arises in the geometric
1656: formalism. Given a gauge ${\cal L}$, one can locally identify $M$ and
1657: the total space of the bundle $N{\cal L}$. One chooses
1658: $s$-homogeneous coordinates $z^\alpha$ and $z^*_\alpha$ along ${\cal
1659: L}$ and the fiber of $N{\cal L}$ such that $(z^*_\alpha, z^\alpha)$
1660: are Darboux coordinates for $\omega$, i.e. : \be
1661: \epsilon(z^*_\alpha)+\epsilon(z^\alpha)={\hat 1}~~{\rm~and~}~~
1662: \omega_{\alpha^*\beta}=-\omega_{\beta\alpha^*}=\delta_{\alpha,\beta}~~.
1663: \ee and such that $s(z^*_\alpha)+s(z^\alpha)=-1$\footnote{Note that
1664: one need not assume $\epsilon(z^\alpha)=s(z^\alpha) (mod~2)$ and
1665: $\epsilon(z^*_\alpha)=s(z^*_\alpha) (mod~2)$. In fact, this is
1666: impossible to arrange if the classical action $S$ contains
1667: Grassmann-odd variables.}.
1668:
1669:
1670: In this case, the odd symplectic form reduces to: \be
1671: \label{Darboux}
1672: \omega=dz^*_\alpha\wedge dz^\alpha~~. \ee One can identify $z^\alpha$
1673: and $z^*_\alpha$ with the fields and antifields of the traditional
1674: formalism. In such coordinates, the BV bracket has the familiar form:
1675: \be
1676: \label{Darboux_bracket}
1677: \{F,G\}= \frac{\partial_r F}{\partial z^\alpha} \frac{\partial_l
1678: G}{\partial z^*_\alpha}- \frac{\partial_r F}{\partial z^*_\alpha}
1679: \frac{\partial_l G}{\partial z^\alpha} ~~, \ee and the Lagrangian
1680: manifold ${\cal L}$ is locally described by the equations
1681: $z^*_\alpha=0$.
1682:
1683:
1684:
1685: \subsubsection{The geometric meaning of BV `quantization'}
1686:
1687: The procedure of BV `quantization' translates as follows. One starts
1688: with a so-called `classical gauge' ${\cal L}$, and with a {\em
1689: classical action} $S_{\cal L}=S_{BV}|_{\cal L}$ defined on ${\cal
1690: L}$. This gauge is typically not convenient for the purpose of
1691: quantization, in that $S_{\cal L}$ is degenerate (has degenerate
1692: Hessian) on ${\cal L}$, which in field-theoretic applications leads to
1693: `infinite factors' in the path integral and the impossibility of
1694: defining propagators. A degenerate Hessian signals the existence of
1695: flat directions for $S$, which in conveniently chosen local
1696: coordinates $z=(z^a)$ means that $S$ depends only on a subset
1697: $x=(z^j)$ of $(z^a)$, which in practice is the subset of coordinates
1698: with ghost degree $s(z^j)=0$. To avoid such problems, one picks
1699: another gauge ${\cal L}'$ such that the Hessian of $S_{\cal L'}$ is
1700: nondegenerate. This allows for the definition of propagators in the
1701: new gauge, which provides a starting point for perturbative
1702: renormalization of the associated path integral.
1703:
1704: The BV procedure can be seen as a systematic approach to performing
1705: the change of gauge from ${\cal L}$ to ${\cal L}'$. This is done by
1706: first extending the classical action $S_{\cal L}$ to the BV action
1707: $S_{BV}$, and then restricting the latter to ${\cal L}'$ to obtain a
1708: candidate for a meaningful definition of the path integral (this
1709: process can be described locally in the traditional language of
1710: gauge-fixing fermions). In its most general form \cite{Schwarz_geom},
1711: the central result is that two gauges ${\cal L}_1$ and ${\cal L}_2$
1712: are physically equivalent provided that their bodies (even parts) are
1713: homologous in the body of $M$ and that the BV action satisfies the
1714: quantum generalization of the classical master equation. It is
1715: important to realize, however, that this statement need not be valid
1716: in more general situations. There is no reason to expect that
1717: `topologically inequivalent' gauges ${\cal L}_1$ and ${\cal L}_2$ lead
1718: to equivalent path integrals.
1719:
1720:
1721: \subsubsection{The BV algorithm}
1722:
1723: The full BV data is rarely known in practical applications. In a
1724: typical situation, one is only given the action $S:=S_{\cal L}$ in a
1725: classical gauge. Since the classical gauge is degenerate, one has no
1726: apriori knowledge of any of the data $M$, $\omega$ or $S_{BV}$. In
1727: this case, one recovers a BV system $(M,\omega, S_{BV}, {\cal L})$
1728: (such that $S_{BV}|_{\cal L}=S$) in the following constructive manner.
1729: First, one has to decide on a gauge algebra, i.e. an algebra of
1730: symmetries of $S$. It should be stressed that the choice of gauge
1731: algebra is not uniquely determined by $S$, since one can insist to
1732: choose a strict subalgebra of the maximal algebra of gauge symmetries
1733: of the classical action\footnote{We shall encounter this phenomenon in
1734: Section 7, when discussing the BV action for D-brane pairs with
1735: relative grading greater than one.}; this choice is dictated by the
1736: physical interpretation of the model. Then one constructs a BV system
1737: through the following two-step procedure:
1738:
1739: \
1740:
1741: (1) Perform the BRST extension of $S(x)$, by applying the BRST
1742: procedure for the given gauge algebra (the precise choice of algebra
1743: influences the result of this step). This enlarges the classical
1744: system from the set of classical fields $x$ to the set of BV fields
1745: $z^\alpha=(x, c_1, c_2,..)$, where $c_k$ are ghosts at generation $k$.
1746: It also produces a nilpotent odd vector field $q$ (the BRST generator)
1747: acting on the enlarged collection of fields. The number of ghost
1748: generations is dictated by the degree of reducibility of the gauge
1749: algebra. After extracting the complete set of ghosts $c_k$, introduce
1750: antifields $z^*_\alpha=(x^*, c_1^*,c_2^*,..)$ for the classical fields
1751: and ghosts. The BV fields and antifields $(z^\alpha, z^*_\alpha)$ are
1752: identified with Darboux coordinates of $M$. This allows one to
1753: locally recover both the supermanifold $M$ and the odd symplectic form
1754: $\omega$ upon using relation (\ref{Darboux}). The Lagrangian
1755: submanifold ${\cal L}$ which defines the classical gauge is recovered
1756: through the equation $z^*_\alpha=0$. The fibers of the Lagrangian
1757: complement $N$ of $T{\cal L}$ are locally defined by the directions
1758: $z^\alpha=ct$.
1759:
1760: Knowledge of the correct collection of BV fields and antifields allows
1761: one to enlarge the degenerate classical action $S(x)$ by adding the
1762: so-called first order action $S_1(z^*,z)=z^*_\alpha q^\alpha(z)$. $S$
1763: and $S_1$ are the first two terms in the expansion of the BV action in
1764: antifields.
1765:
1766: \
1767:
1768: (2) The odd symplectic form $\omega$ recovered in the first step
1769: defines the BV bracket $\{.,.\}$. To recover the full BV action, one
1770: must solve the master equation $\{S_{BV}, S_{BV}\}=0$ with the ansatz
1771: $S_{BV}=\sum_{k\geq 0}{S_k}$, where $S_k$ is the $k^{th}$ order term
1772: in the Taylor expansion in antifields. The first two terms $S_0=S$ and
1773: $S_1$ are known from Step (1).
1774:
1775:
1776: \subsection{Realization of the geometric data and check of the master equation}
1777:
1778: \subsubsection{The graded P-manifold}
1779:
1780: \paragraph{The supermanifold}
1781:
1782: Remember that the unextended total boundary space ${\cal H}$ is a
1783: $\Z$-graded vector space with respect to the worldsheet $U(1)$ charge
1784: $|~.~|$. The mod $2$ reduction of $|~.~|$ makes ${\cal H}$ into a
1785: vector superspace ($\Z_2$-graded vector space): \be {\cal H}={\cal
1786: H}_{even}\oplus {\cal H}_{odd}~~, \ee where: \bea {\cal
1787: H}_{even}&=&\{u\in {\cal H}||u|=even\}= \oplus_{k+n-m=even~}
1788: {\Omega^k(L)\otimes \Gamma(Hom(E_m,E_n))}~~,\nn\\
1789: {\cal H}_{odd}~&=&\{u\in {\cal H}||u|=odd~~\}=\oplus_{k+n-m=odd~~}
1790: {\Omega^k(L)\otimes \Gamma(Hom(E_m,E_n))}~~. \eea
1791:
1792:
1793: To construct the supermanifold relevant for the geometric BV
1794: formalism, we consider a new $\Z$-grading $s$ on ${\cal H}$ which is
1795: related to $|~.~|$ by: \be s(u)=1-|u|~~. \ee The vector space ${\cal
1796: H}$ endowed with this grading will be denoted by ${\tilde {\cal
1797: H}}$. The mod~2 reduction of $s$ makes ${\tilde {\cal H}}$ into a
1798: vector superspace: \be {\tilde {\cal H}}={\tilde {\cal
1799: H}}_{even}\oplus {\tilde {\cal H}}_{odd}~~, \ee where ${\tilde
1800: {\cal H}}_{even}={\cal H}_{odd}$ and ${\tilde {\cal H}}_{odd~}
1801: ={\cal H}_{even}$.
1802:
1803: It is clear that the superspaces ${\cal H}$ and ${\tilde {\cal H}}$
1804: differ by parity change. They define infinite-dimensional complex
1805: linear supermanifolds (in the sense of Berezin), which we denote by
1806: $L({\tilde {\cal H}})$ and $L({\cal H})$. Their $G$-valued points
1807: define DeWitt-Rogers supermanifolds: \be M:=L({\tilde {\cal
1808: H}})(G)=({\tilde {\cal H}}\otimes G)^0= {\tilde {\cal
1809: H}}_e^0={\cal H}_e^1~~,~~ P=L({\cal H})(G)=({\cal H}\otimes
1810: G)^0={\cal H}_e^0~~, \ee where we defined ${\tilde {\cal
1811: H}}_e:={\tilde {\cal H}}\otimes G$. In these equations, ${\cal
1812: H}_e$ and ${\tilde {\cal H}}_e$ are viewed as vector superspaces,
1813: and we used the obvious relation: \be {\tilde {\cal H}}_e=\Pi {\cal
1814: H}_e~~. \ee
1815:
1816: \paragraph{The graded manifold structure}
1817:
1818: We will mainly be interested in the linear supermanifold $M$, on which
1819: we now introduce a structure of $\Z$-graded manifold. For this,
1820: consider a homogeneous basis $e_a$ of ${\cal H}$, with
1821: $|e_a|:=|a|\Rightarrow s(e_a)=s_a:=1-|a|$. Then every element ${\hat
1822: \phi}$ of ${\cal H}^1_e$ has the expansion: \be {\hat
1823: \phi}=\sum_{a}{e_a\otimes {\hat \phi}^a}~~, \ee with ${\hat
1824: \phi}^a\in G$ and $g({\hat \phi}^a)=deg {\hat \phi}-|e_a| (mod~2)=
1825: 1-|e_a| (mod~2)=s_a (mod~2)$. This allows us to define maps
1826: $z^a:{\cal H}_e\rightarrow G$ through $z^a({\hat \phi})={\hat
1827: \phi}^a$, which have parities $\epsilon_a=g_a=s_a (mod~2)$ as
1828: $G$-valued functions. They give (global) coordinates on the
1829: supermanifold $M$. Coordinates for $M$ obtained in this manner will be
1830: called {\em homogeneous linear coordinates}.
1831:
1832:
1833: The collection of homogeneous linear coordinates associated to all
1834: homogeneous bases $(e_a)$ of ${\cal H}$ forms a distinguished atlas
1835: for our supermanifold. Such coordinates are endowed not only with a
1836: $\Z_2$-degree $\epsilon_a=\epsilon(z^a)$, but also with a $\Z$-valued
1837: degree $s(z^a):=s_a=s(e_a)$, such that $\epsilon_a=s_a (mod~2)$. As
1838: explained in subsection 4.2., this can be used to define a
1839: $\Z$-grading $s$ on the sheaf ${\cal F}$ of $G$-valued functions, if
1840: one restricts the latter to consist of functions which are polynomial
1841: in coordinates. This grading on ${\cal F}$ will play the role of ghost
1842: grading in the BV formalism.
1843:
1844: \paragraph{Vector fields as nonlinear operators}
1845: Since $M$ is a linear supermanifold, vector fields on $M$ can be
1846: viewed as maps of the form $X:{\hat \phi}\rightarrow ({\hat \phi},
1847: {\bf X}({\hat \phi}))\in M\times {\tilde {\cal H}}_e$, where ${\bf X}$
1848: is a (generally nonlinear) operator from $M$ to ${\tilde {\cal H}}_e$.
1849: Even ($\epsilon_X={\hat 0}$) and odd ($\epsilon_X={\hat 1}$) vector
1850: fields correspond to operators ${\bf X}$ from $M$ to ${\cal H}_e^1$
1851: and $M$ to ${\cal H}_e^0$ respectively.
1852:
1853: In homogeneous linear coordinates $z^a$, the vector fields
1854: $\partial^r_a$ correspond to the constant operators ${\hat
1855: \phi}\rightarrow e_a\otimes 1_G$. One has $s(\partial^r_a)=-s_a$ and
1856: $\epsilon(\partial^r_a)=\epsilon_a=s_a (mod~2)$. For an arbitrary
1857: vector field $X=\partial^r_aX^a_r$, one has ${\bf X}({\hat
1858: \phi})=e_a\otimes X^a_r({\hat \phi})$ (with $X^a_r({\hat \phi})\in
1859: G$). Since $z^a({\bf X}({\hat \phi}))=X^a_r({\hat \phi})$, the vector
1860: field $X$ can be recovered from the operator ${\bf X}$ through the
1861: relation: \be (z^a\stackrel{\leftarrow}{X})({\hat \phi})=X^a({\hat
1862: \phi})= z^a({\bf X}({\hat \phi}))~~. \ee Note that one can expand:
1863: \be {\bf X}({\hat \phi})={\bf X}_0+{\bf X}_1({\hat \phi})+ {\bf
1864: X}_2({\hat \phi}, {\hat \phi})+{\bf X}_3({\hat \phi}, {\hat \phi},
1865: {\hat \phi})+...~~, \ee where ${\bf X}_k: ({\tilde {\cal H}}\otimes
1866: G)^{\otimes k}\rightarrow {\tilde {\cal H}}\otimes G$ are
1867: $G$-multilinear operators. It is not hard to check that the vector
1868: field $X$ is $s$-homogeneous of degree $\sigma$ if and only if ${\bf
1869: X}_k$ are $s$-homogeneous of degree $-\sigma$, i.e.: \be s({\bf
1870: X}_k({\hat \phi}_1, ..., {\hat \phi}_k))= s({\hat
1871: \phi}_1)+...+s({\hat \phi}_k)-\sigma~~. \ee
1872:
1873: By localizing the description of vector fields to a point ${\hat
1874: \phi}$ of $M$, one obtains the identification: \be T_{\hat
1875: \phi}M={\tilde {\cal H}}\otimes G={\tilde {\cal H}}_e~~, \ee which
1876: holds both as an isomorphism of right G-supermodules and as an
1877: isomorphism of $\Z$-graded vector spaces (the $\Z$-grading on both
1878: sides being given by the ghost degree $s$). It follows that the total
1879: space of the tangent bundle to $M$ can be identified with: \be
1880: \label{tb}
1881: TM=M\times {\tilde {\cal H}}_e~~. \ee The left G-supermodule
1882: structure on $T_{\hat \phi}M$ is defined through\footnote{Note that
1883: the exponent in (\ref{lrmod}) involves $\epsilon(X_{\hat \phi})= deg
1884: X_{\hat \phi}+{\hat 1}$ and {\em not} $deg X_{\hat \phi}$. This is
1885: due to the presence of parity reversal in the isomorphism of
1886: superspaces $T_{\hat \phi}M={\tilde {\cal H}}_e=\Pi{\cal H}_e$.}:
1887: \be
1888: \label{lrmod}
1889: \alpha X_{\hat \phi}= (-1)^{\epsilon_\alpha\epsilon(X_{\hat
1890: \phi})}X_{\hat \phi}\alpha~~, \ee where $\alpha\in G$ and
1891: $\epsilon_\alpha=g(\alpha)$.
1892:
1893:
1894: \paragraph{The odd symplectic form}
1895:
1896: The extended bilinear form $\langle . , . \rangle_e$ on ${\cal H}_e$
1897: allows us to define the following two-form on $M$: \be
1898: \label{omega}
1899: \omega_{\hat \phi}(X_{\hat \phi}, Y_{\hat \phi}):=
1900: (-1)^{\epsilon(X_{\hat \phi})}\langle X_{\hat \phi}, Y_{\hat
1901: \phi}\rangle_e~~, \ee where ${\hat \phi}$ is a point in $M$ and
1902: $X_{\hat \phi}, Y_{\hat \phi}\in T_{\hat \phi}M={\tilde {\cal
1903: H}}\otimes G$ are tangent vectors to $M$ at ${\hat \phi}$. It is
1904: easy to check that $\omega$ is an {\em odd symplectic form} on $M$,
1905: and that $s_\omega=-1$. The last statement follows upon choosing a
1906: homogeneous basis $s_a$ of ${\cal H}$ and considering the coefficients
1907: of $\omega$ in this basis: \be \omega_{ab}=(-1)^{\epsilon_a}\langle
1908: e_a, e_b\rangle=ct~~, \ee where we used the identification
1909: $\partial^r_a=e_a\otimes 1_G$ and the fact that $\langle e_a\otimes
1910: 1_G, e_b\otimes 1_G\rangle_e= \langle e_a, e_b\rangle$. The selection
1911: rules (\ref{selrules}) allow us to restrict to the case
1912: $|e_a|+|e_b|=3\Leftrightarrow s_a+s_b=-1$. In this case, one has
1913: $s(\omega_{ab})=-s_a-s_b-1$, where we used the fact that $\omega_{ab}$
1914: is a constant and thus $s(\omega_{ab})=0$. This implies that
1915: $s_\omega=-1$. We conclude that $(M, \omega)$ is a (DeWitt-Rogers)
1916: graded P-manifold. Note that the restriction of $\omega$ to the even
1917: component $TM^0\approx {\cal H}_e^1$ of the tangent bundle coincides
1918: with the form $\omega_0=\langle ., .\rangle_e|_{M\times M}$ of
1919: equation (\ref{omega0}). In fact, $\omega$ is completely determined
1920: by this restriction and by the requirement that it must be an odd
1921: form. This observation will be useful in Section 7.
1922:
1923: The extended action (\ref{extended_action}) is an (even) function
1924: defined on $M$. To check that its ghost degree equals zero, we first
1925: notice that the operator $d$ and extended boundary product $*$ satisfy
1926: $s(d{\hat \phi})=s({\hat \phi})-1$ and $s({\hat \phi}_1*{\hat \phi}_2)
1927: =s({\hat \phi}_1)+s({\hat \phi}_2)-1$. Since $d$ and $*$ are right
1928: $G$-linear and bilinear respectively, they define a nonlinear operator
1929: ${\bf W}({\hat \phi})=\frac{1}{2}d{\hat \phi}+ \frac{1}{3}{\hat
1930: \phi}*{\hat \phi}$ which is $s$-homogeneous of degree $-1$. This in
1931: turn defines a vector field $W$ of ghost degree $+1$ on $M$. On the
1932: other hand, the identity operator ${\bf I}:{\hat \phi}\rightarrow
1933: {\hat \phi}$ defines a vector field $I$ of ghost degree zero. It is
1934: then clear that $S_e$ can be written in the purely geometric form: \be
1935: S_e=\omega(I, W)~~, \ee which obviously has ghost degree zero (since
1936: $\omega$ has ghost degree $-1$). We are now ready to apply the
1937: geometric formalism to the system $(M, S_e, \omega)$ in order to show
1938: that the extended action satisfies the classical master equation.
1939:
1940:
1941: \subsubsection{The odd vector field $Q$ }
1942:
1943: As discussed above, the odd Hamiltonian vector field $Q:=Q_{S_e}$ can
1944: be viewed as a (non-linear) map ${\bf Q}$ from $M$ to $P$. To obtain
1945: an explicit formula for ${\bf Q}$, we compute the variation of
1946: $S_e({\hat \phi})$ under an infinitesimal change of ${\hat \phi}$: \be
1947: \delta S_e({\hat \phi})=\langle d{\hat \phi}+\frac{1}{2}[{\hat
1948: \phi},{\hat \phi}]_*, \delta {\hat \phi} \rangle_e= -\omega_{\hat
1949: \phi}( d{\hat \phi}+\frac{1}{2}[{\hat \phi},{\hat \phi}]_*, \delta
1950: {\hat \phi})~~, \ee which means that the differential of $S_e$ has the
1951: form: \be dS_e(X)=\omega(Q, X)~~, \ee with: \be
1952: \label{bfQ}
1953: {\bf Q}({\hat \phi})=-(d{\hat \phi}+\frac{1}{2}[{\hat \phi},{\hat
1954: \phi}]_*)~~. \ee
1955:
1956: \subsubsection{Check of the master equation}
1957:
1958: Relation (\ref{bfQ}) implies: \be \omega(Q,Q)({\hat \phi})=
1959: \omega_{\hat \phi}({\bf Q}({\hat \phi}), {\bf Q}({\hat \phi}))
1960: =-\langle d{\hat \phi}, d{\hat \phi}\rangle_e- 2\langle d{\hat \phi},
1961: {\hat\phi}^{*2}\rangle_e- \langle {\hat \phi}^{*2}, {\hat
1962: \phi}^{*2}\rangle_e~~, \ee where we use the notation $\phi^{*n}$ to
1963: indicate the $n^{th}$ power of $\phi$ computed with the product $*$.
1964: It is easy to check that all three terms vanish upon using the
1965: properties of $\omega$ and the condition $deg {\hat \phi}={\hat 1}$.
1966: Indeed: \be \langle d{\hat \phi}, d{\hat \phi}\rangle_e= \langle {\hat
1967: \phi}, d^2{\hat \phi}\rangle_e=0~~, \ee \bea \langle d{\hat \phi},
1968: {\hat \phi}^{*2}\rangle_e= \langle{\hat \phi}, (d{\hat \phi})*{\hat
1969: \phi}\rangle_e &-&\langle{\hat \phi}, {\hat \phi}*d{\hat
1970: \phi}\rangle_e= -\langle d{\hat \phi})*{\hat \phi}, {\hat
1971: \phi}\rangle_e
1972: -\langle {\hat \phi}^{*2}, d{\hat \phi}\rangle_e=~~~~~~\nn\\
1973: &=&-2\langle d{\hat \phi}, {\hat \phi}^{*2}\rangle_e\Longrightarrow
1974: \langle d{\hat \phi}, {\hat \phi}^{*2}\rangle_e=0~~, \eea and finally:
1975: \bea \langle{\hat \phi}^{*2}, {\hat \phi}^{*2}\rangle_e
1976: &=&\langle{\hat \phi}, {\hat \phi}^{*3}\rangle_e=- \langle{\hat
1977: \phi}^{*3}, {\hat \phi}\rangle_e= -\langle{\hat \phi}^{*2}, {\hat
1978: \phi}^{*2}\rangle_e~~ \Longrightarrow \langle{\hat \phi}^{*2}, {\hat
1979: \phi}^{*2}\rangle_e=0~~.\nn \eea We conclude that $\omega({\bf Q},
1980: {\bf Q})=0$, which in view of equations (\ref{master}) implies that
1981: $S_e$ satisfies the classical master equation $\{S_e,S_e\}=0$, with
1982: respect to the BV bracket induced by $\omega$.
1983:
1984: \subsubsection{The classical gauge}
1985:
1986: Let us consider the decompositions: \bea
1987: M=\oplus_{s}{M_s}~~,~~P=\oplus_{s}{P_s}~~, \eea where $M_s={\tilde
1988: {\cal H}}^s\otimes G_{s~(mod~2)}= {\cal H}^{1-s}\otimes
1989: G_{s~(mod~2)}$, $P_s={\tilde {\cal H}}^s\otimes G_{(1-s)~(mod~2)}=
1990: {\cal H}^{1-s}\otimes G_{(1-s)~(mod~2)}$ denote the the subspaces of
1991: $M$ and $P$ spanned by elements ${\hat \phi}$ of ghost number $s$.
1992: The space $M_0={\tilde {\cal H}}^0\otimes G_0= {\cal H}^1\otimes G_0$
1993: was considered in equation (\ref{M0}) of Section 3.2. As mentioned
1994: there, the unextended string field $\phi$ can be related to the
1995: component ${\hat \phi}_0$ of ${\hat \phi}$ along this subspace: \be
1996: {\hat \phi}_0=\phi\otimes 1_G+{\hat \phi}_0^\mu\xi_\mu~~, \ee where
1997: $\xi^\mu$ are the odd generators of $G$ (see relation \ref{Ggens}).
1998: Since it is easy to include the evaluation map $ev_G$ in relations
1999: such as (\ref{restriction}), we shall denote ${\hat \phi}_0$ by $\phi$
2000: in order to simplify notation. This allows us to view the unextended
2001: string field as the component of ${\hat \phi}$ along $M_0$.
2002:
2003: The selection rule (\ref{dc}) for the symplectic form implies that the
2004: subspaces: \bea
2005: \label{lagr_dec}
2006: {\cal L}&=&\{{\hat \phi}\in M|s({\hat \phi})\geq 0\}= \oplus_{s\geq
2007: 0}{M_s}
2008: ~~\nn\\
2009: {\cal N}&=&\{{\hat \phi}\in M|s({\hat \phi})<0\}=\oplus_{s<0}{M_s}
2010: \eea give a Lagrangian decomposition of $(M, \omega)$ (viewed as an
2011: odd symplectic vector space). In particular, ${\cal L}$ is a
2012: Lagrangian submanifold of $M$, and we can identify $T{\cal L}={\cal
2013: L}\times {\cal L}$ and $N{\cal L}={\cal L}\times {\cal N}$ (then
2014: $TM|_{\cal L}(s)={\cal L}\times M_s$).
2015:
2016:
2017:
2018: The same selection rule shows that the restriction of $S_e$ to ${\cal
2019: L}$ coincides with $S$ (up to application of $ev_G$). It follows
2020: that $S_e$ plays the role of BV action for the classical action $S$,
2021: while ${\cal L}$ describes the associated `classical gauge'. Since
2022: $S_e|_{\cal L}=S$ depends only on the component of ${\hat \phi}$ along
2023: $M_0$, the classical action $S_{\cal L}$ has degenerate Hessian on
2024: ${\cal L}$. Note that the classical gauge ${\cal L}$ is entirely
2025: determined by the ghost grading $s$.
2026:
2027:
2028: Following standard BV procedure, we define BV {\em fields}
2029: $\boldphi\in {\cal L}$ and {\em antifields} $\boldphi^*\in {\cal N}$
2030: as the components of ${\hat \phi}$ along ${\cal L}$ and ${\cal N}$.
2031: This gives the decomposition ${\hat \phi}=\boldphi+\boldphi^*$. The
2032: component $\phi\in M_0$ of $\boldphi$ is the unextended string field,
2033: while the higher components $c_{s}\in M_{s}$ ($s\geq 1$) play the role
2034: of ghosts. The highest component $\phi^*\in M_{-1}$ of $\boldphi^*$
2035: is the antifield of $\phi$, while the lower components $c^*_{s}\in
2036: M_{-1-s}$ $(s\geq 1$) are the antifields of $c_s$. Hence one has the
2037: decomposition: \be
2038: \label{fa_expansion}
2039: {\hat \phi}=... +c^*_2+ c^*_1+\phi^*+\phi+c_1+c_2+...~~, \ee which we
2040: also write in the form: \be {\hat \phi}=\oplus_{s}{\phi_s}~~, \ee
2041: where $\phi_0:=\phi$, $\phi_{-1}=\phi^*$, $\phi_s=c_s$ and
2042: $\phi_{-1-s}=c_s^*$ for $s\geq 1$. We have $\boldphi=\oplus_{s\geq
2043: 0}{\phi_s}$ and $\boldphi^*=\oplus_{s<0}{\phi_s}= \oplus_{s\geq
2044: 0}{\phi^*_s}$. Note the relations: \be
2045: \phi_{-1-s}=\phi^*_s~{\rm~for~}s\geq 0~,~ rk \phi^*_s + rk
2046: \phi_s=3~,~\Delta(\phi^*_s)+\Delta(\phi_s)=0~,~ g(\phi^*_s) +
2047: g(\phi_s)={\hat 1}~~, \ee which are due to the selection rules for
2048: $\omega$.
2049:
2050:
2051: \subsubsection{BRST transformations in the classical gauge}
2052:
2053: Applying parity change to the bundle decomposition $TM|_{\cal
2054: L}=T{\cal L}\oplus T{\cal N}$ gives $\Pi TM|_{\cal L}=\Pi T{\cal
2055: L}\oplus \Pi N{\cal L}$, with $\Pi T{\cal L}={\cal L}\times \Pi
2056: {\cal L}$ and $\Pi N{\cal L}={\cal L}\times \Pi {\cal N}$, with $\Pi
2057: {\cal L}$ and $\Pi {\cal N}$ given by the following complementary
2058: subspaces of ${\cal H}_e$: \be
2059: \label{pN}
2060: \Pi {\cal L}=\oplus_{s\geq 0}{P_s}~~,~~ \Pi {\cal
2061: N}=\oplus_{s<0}{P_s}~~, \ee where used the obvious identity $\Pi
2062: M_s=P_s$. Since ${\cal L}$ is a vector space and the bundle $\Pi
2063: T{\cal L}={\cal L}\times \Pi {\cal L}$ is trivial, the odd vector
2064: field $q$ of (\ref{qgen}) can be identified with an even nonlinear
2065: operator ${\bf q}$ from ${\cal L}$ to $\Pi {\cal L}$. Relation
2066: (\ref{qgen}) then translates as: \be
2067: \label{BRST}
2068: {\bf q}(\boldphi)= T({\bf Q}(\boldphi))=
2069: -T\left(d\boldphi+\frac{1}{2}[\boldphi,\boldphi]_*\right)~~, \ee where
2070: $T$ is the projector of $P$ onto $\Pi {\cal L}$, taken parallel with
2071: the subspace $\Pi {\cal N}$ (the projector on parity changed BV
2072: fields).
2073:
2074: \subsubsection{Expansion of $S_e$ in antifields}
2075:
2076: We end this section by writing the extended action in a form which
2077: will be useful later. It is easy to check that: \be
2078: \label{Sexp}
2079: S_e({\hat \phi})=S(\phi)-\langle \boldphi^*, {\bf Q}
2080: (\boldphi)\rangle_e+\langle \boldphi,\boldphi^*~*\boldphi^*\rangle_e=
2081: S(\phi)-\langle \boldphi^*, {\bf q} (\boldphi)\rangle_e+\langle
2082: \boldphi,\boldphi^*~*\boldphi^*\rangle_e~~. \ee This expression
2083: corresponds to the expansion of $S_e$ in antifields around the
2084: classical gauge ${\cal L}$.
2085:
2086: \section{Graded D-brane pairs \label{sec:gradedpairs}}
2087:
2088: \subsection{`Component' description of graded pairs}
2089:
2090: If one restricts to the case of graded D-brane pairs, our string field
2091: theory can be described as follows (figure 1). Using the labels $a,b$
2092: to denote the associated graded branes, with underlying flat bundles
2093: $E_a$ and $E_b$ living on $L$, one can take the grade of $a$ to be
2094: zero without loss of generality. If $n$ denotes the grade of $b$,
2095: then the relative grade $grade(b)-grade(a)$ is $n$. The theory has
2096: four boundary sectors, which we denote by $Hom(a,a)$, $Hom(b,b)$ (the
2097: diagonal sectors) and $Hom(a,b)$, $Hom(b,a)$ (the off-diagonal, or
2098: boundary condition changing sectors). They are the off-shell spaces of
2099: states for strings stretching from $a$ to $a$, $b$ to $b$, $a$ to $b$
2100: and $b$ to $a$ respectively. The localization arguments of
2101: \cite{Witten_CS}, combined with the shift of $U(1)$ charge discussed
2102: in \cite{Douglas_Kontsevich}, lead to the identifications: \bea
2103: \label{homs}
2104: Hom^k(a,a)=\Omega^k(End(E_a)) &,& Hom^k(b,b)=
2105: \Omega^k(End(E_b)) \nn\\
2106: Hom^k(a,b)=\Omega^{k-n}(Hom(E_a,E_b)) &,& Hom^k(b,a)=\Omega^{k+n}
2107: (Hom(E_b,E_a))~, \eea where $k$ is the worldsheet $U(1)$ charge. The
2108: relation between the charge $|~.~|$ and form rank is given by
2109: (\ref{ghost_mn}): \be |u_{AB}|=rk u_{AB}+grade(B)-grade(A)~~{\rm~for~}
2110: u_{AB}\in Hom(A, B)~, \ee for all $A,B\in \{a,b\}$. In equations
2111: (\ref{homs}), we let $k$ take any integer value and define the space
2112: of forms of negative ranks (or ranks greater than three) on $L$ to be
2113: zero. Thus non-vanishing states in the sectors $Hom(a,b)$ and
2114: $Hom(b,a)$ have charges $k=n,n+1,n+2$ or $n+3$ and
2115: $k=-n,-n+1,-n+2,-n+3$ respectively. It follows that the total
2116: boundary space ${\cal H}=Hom(a,a)\oplus Hom(b, a)\oplus Hom(a,b)\oplus
2117: Hom(b,b)$ has non-vanishing components of charges $k=\{-n,\dots
2118: ,\,-n+3\} \cup \{n,\dots ,\, n+3\}\cup \{0,\dots,\,3\}$.
2119:
2120: \hskip 1.0 in
2121: \begin{center}
2122: \scalebox{0.6}{\input{pair.pstex_t}}
2123: \end{center}
2124: \begin{center}
2125: Figure 1. {\footnotesize Boundary sectors for a pair of graded
2126: D-branes wrapping the same special Lagrangian cycle. The two
2127: D-branes $a$ and $b$ are thickened out for clarity, though their
2128: (classical) thickness is zero.}
2129: \end{center}
2130:
2131:
2132: The various boundary sectors carry BRST operators $d_{aa}, d_{ab},
2133: d_{ba}$ and $d_{bb}$ given by the covariant differentials twisted with
2134: the flat connections $A_a$ and $A_b$. One also has a modification of
2135: the boundary product: \be
2136: \label{cdot}
2137: u_{BC}\bullet u_{AB}=(-1)^{[grade(B)-grade(C)]\,rk u_{AB}}
2138: \,u_{BC}\wedge u_{AB}~~, \ee for all $A,B,C\in \{a,b\}$.
2139:
2140: Upon using this data, one can write the explicit form of the string
2141: field action (\ref{action}) for such a system. In this case, the
2142: graded bundle is ${\bf E}=E_a\oplus E_b$, and the string field $\phi$
2143: is a (worldsheet charge one) element of the total boundary space
2144: ${\cal H}=\Omega^*(L)\otimes \Gamma(End({\bf E}))$. It can be
2145: represented as a matrix of bundle-valued forms: \be
2146: \phi=\left[\begin{array}{cc} \phi^{(1)}_{aa}~~~&~~~\phi^{(1+n)}_{ba}\\
2147: \phi^{(1-n)}_{ab}&\phi^{(1)}_{bb}\end{array}\right]~~,
2148: \label{eq:morphisms}
2149: \ee with: \bea \phi^{(1)}_{aa}\in
2150: Hom^1(a,a)=\Omega^1(End(E_a))&,&\phi^{(1)}_{bb}\in Hom^1(b,b)=
2151: \Omega^1(End(E_b))\nn\\
2152: \phi^{(1-n)}_{ab}\in Hom^1(a,b)=\Omega^{1-n}(Hom(E_a,E_b)) &,&
2153: \phi^{(1+n)}_{ba}\in Hom^1(b,a)=\Omega^{1+n}(Hom(E_b,E_a))\nn~,~~~~~
2154: \eea where the superscripts in round brackets indicate form rank.
2155:
2156: Given two states $u,v\in {\cal H}$, represented by the matrices
2157: $u=\left[\begin{array}{cc} u_{aa}&u_{ba}\\
2158: u_{ab} & u_{bb}\end{array}\right]$ and
2159: $v=\left[\begin{array}{cc} v_{aa}&v_{ba}\\
2160: v_{ab} & v_{bb}\end{array}\right]$, their boundary product
2161: (\ref{dot}) is given by: \be u\bullet v:=\left[\begin{array}{ccc}
2162: u_{aa}\bullet v_{aa}+u_{ba}\bullet v_{ab}
2163: &{~}& u_{aa}\bullet v_{ba} +u_{ba}\bullet v_{bb}\\
2164: u_{ab}\bullet v_{aa}+ u_{bb}\bullet v_{ab} &{~}& u_{bb}\bullet
2165: v_{bb} + u_{ab}\bullet v_{ba}
2166: \end{array}\right]~~.
2167: \label{cdot'}
2168: \ee We also have the fiberwise supertrace on $End({\bf E})$: \be
2169: str\left[\begin{array}{cc} u_{aa}&u_{ba}\\
2170: u_{ab} &
2171: u_{bb}\end{array}\right]=tr_{a}(u_{aa})+(-1)^ntr_{b}(u_{bb})~~,
2172: \ee and the total worldsheet BRST operator: \be
2173: d\left[\begin{array}{cc} u_{aa}&u_{ba}\\
2174: u_{ab} & u_{bb}\end{array}\right]=
2175: \left[\begin{array}{cc} d_{aa}u_{aa}&d_{ba}u_{ba}\\
2176: d_{ab}u_{ab} & d_{bb}u_{bb}\end{array}\right]~~, \ee as well as
2177: the boundary bilinear form: \be \langle u, v
2178: \rangle=\int_{L}{str(u\bullet v)}=\int_{L}{\left[~tr_a(u_{aa}\bullet
2179: v_{aa}+ u_{ba}\bullet v_{ab})+(-1)^n tr_b(u_{ab}\bullet
2180: v_{ba}+u_{bb}\bullet v_{bb})~\right]}~~, \ee where $tr_a$ and
2181: $tr_b$ denote the fiberwise trace on the bundles $End(E_a)$ and
2182: $End(E_b)$.
2183:
2184: With these notations, the string field action (\ref{action}) expands
2185: as: \bea
2186: \label{action_pair}
2187: S(\phi)&=&\int_{L}{str\left[\,{1\ov 2}\phi \bullet d
2188: \phi +\frac{1}{3}\phi\bullet \phi\bullet \phi \right]} \\
2189: =&\frac{1}{2}&\int_{L}{\left[tr_a(\phi_{aa}\bullet d\phi_{aa}+
2190: \phi_{ba}\bullet d\phi_{ab})+
2191: (-1)^n~tr_b(\phi_{bb}\bullet d\phi_{bb}+\phi_{ab}\bullet d\phi_{ba})\right]}+\nn\\
2192: &\frac{1}{3}&\int_{L}{\left[tr_a(\phi_{aa}\bullet \phi_{aa}\bullet
2193: \phi_{aa}+ \phi_{aa}\bullet \phi_{ba}\bullet
2194: \phi_{ab}+\phi_{ba}\bullet \phi_{bb}\bullet \phi_{ab}
2195: +\phi_{ba}\bullet \phi_{ab}\bullet \phi_{aa})\right]}+\nn\\
2196: &\frac{1}{3}&\int_{L}{(-1)^n\left[tr_b(\phi_{bb}\bullet
2197: \phi_{bb}\bullet \phi_{bb}+ \phi_{bb}\bullet \phi_{ab}\bullet
2198: \phi_{ba}+\phi_{ab}\bullet \phi_{aa}\bullet \phi_{ba}
2199: +\phi_{ab}\bullet \phi_{ba}\bullet \phi_{bb})\right]} ~~.\nonumber
2200: \eea Since the rank of a form on the three-cycle $L$ lies between $0$
2201: and $3$, one can distinguish the cases $n=-2,-1,0,1,2$ as well as the
2202: `diagonal' case $|n|\geq 3$.
2203: %In the last situation, the boundary condition changing sectors
2204: %reduce to the null vector space, so one has two independent topological
2205: %D-branes (no nonzero boundary condition changing states).
2206: Notice that one can easily translate from negative to positive $n$ by
2207: reversing the roles of $a$ and $b$, so we can further restrict to
2208: $n=0,1$ or $2$. The case $n=0$ gives the usual Chern-Simons theory on
2209: the direct sum bundle $E_a\oplus E_b$. The case $n\geq 3$ gives a
2210: graded sum of two Chern-Simons theories. Hence the interesting cases
2211: are $n=1$ and $n=2$.
2212:
2213: \subsection{Comparison of worldsheet BRST cohomologies}
2214:
2215: Let us compare the physical (charge one) worldsheet BRST cohomology
2216: for the cases $n=0,1,2$. The cohomology of the BRST operator $d$ has
2217: the direct sum decomposition: \be H^1_d({\cal H})=H^1(End(E_a))\oplus
2218: H^1(End(E_b))\oplus H^{1+n}(Hom(E_b,E_a)) \oplus
2219: H^{1-n}(Hom(E_a,E_b))~~. \ee Hence it suffices to compare the
2220: `off-diagonal components' $H^1_{od}=H^{1+n}(Hom(E_b,E_a))\oplus
2221: H^{1-n}(Hom(E_a,E_b))$. Since $H^{1+n}(Hom(E_b,E_a))\approx
2222: H^{2-n}(Hom(E_a, E_b))$ by Poincar\'e duality, one obtains:
2223:
2224: \
2225:
2226: (1) ($n=0$) $H^1_{od}\approx H^1(Hom(E_a,E_b))\oplus
2227: H^2(Hom(E_a,E_b))$.
2228:
2229: (2) ($n=1$) $H^1_{od}\approx H^0(Hom(E_a,E_b))\oplus
2230: H^1(Hom(E_a,E_b))$.
2231:
2232: (3) ($n=2$) $H^1_{od}\approx H^0(Hom(E_a,E_b))$.
2233:
2234: (4) ($n\geq 3$) $H^1_{od}\approx 0$.
2235:
2236: \
2237:
2238: It is clear from this that $H^1_d({\cal H})$ will generally be
2239: different for various $n$. For the simple case when $E_a\approx
2240: E_b={\cal O}_L$ (the trivial line bundle on $L$), endowed with the
2241: trivial flat structures, one obtains:
2242:
2243: \
2244:
2245: (1) ($n=0$) $dim H^1_{od}=2b_1(L)$.
2246:
2247: (2) ($n=1$) $dim H^1_{od}=1+b_1(L)$.
2248:
2249: (3) ($n=2$) $dim H^1_{od}=1$.
2250:
2251: (4) ($n\geq 3$) $dim H^1_{od}=0$~~,
2252:
2253: \
2254:
2255: \noindent where $b_1(L)$ is the first Betti number of $L$
2256: (we assume that $L$ is connected, so that $dim H^0(L)=1$). Since $L$
2257: is compact, the dimension $dim H^1_d({\cal H})=2b_1(L)+dim H^1_{od}$
2258: counts the number of independent physical degrees of freedom. It
2259: follows that our theories are generally inequivalent. The case
2260: $b_1(L)\geq 2$ (for example a special Lagrangian 3-torus, with
2261: $b_1(L)=3$) allows us to distinguish between all four classes based on
2262: this simple argument. Including the `conjugate' cases $n<0$, we
2263: conclude that there are in general {\em six} distinct types of D-brane
2264: pairs, which shows that the $\Z$-valued D-brane grade has physical
2265: consequences. This is true even though D-brane pairs whose relative
2266: gradings coincide modulo two (such as the pairs with $n=-1$, $n=+1$,
2267: or the pairs $n=-2$, $n=0$ and $n=2$) have the same classical master
2268: action. As mentioned above, the difference between such theories can
2269: be understood at the BV level as resulting from inequivalent choices
2270: for the ghost grading and classical gauge.
2271:
2272:
2273: \section{Composite formation and acyclicity}
2274:
2275: \subsection{Vacuum shifts and D-brane composites}
2276:
2277: As discussed in some detail in \cite{sc} (upon following the general
2278: framework of \cite{com1, com3}), D-brane composite formation can be
2279: described by the simple device of shifting the string vacuum. This
2280: results from the observation \cite{com1} that a vacuum shift will
2281: generally break the decomposition of the total boundary space ${\cal
2282: H}$ into boundary sectors, thereby forcing a change in our D-brane
2283: interpretation\footnote{This process admits an elegant mathematical
2284: description in the language of differential graded categories
2285: \cite{com1}, but no familiarity with category theory is required for
2286: reading the present paper.}. In the case of graded D-brane pairs,
2287: this process can be described as follows. Suppose that we are given a
2288: (worldsheet charge one) solution $\phi$ to the string field equations
2289: of motion: \be
2290: \label{eom}
2291: \frac{\delta S}{\delta \phi}=0\Longleftrightarrow d\phi+\phi\bullet
2292: \phi=0~~, \ee and with the property that at least one of the
2293: components $\phi^{(1-n)}_{ab}$ and $\phi^{(1+n)}_{ba}$ is nonzero. If
2294: one shifts the string vacuum through $\phi$, then the BRST operator
2295: around the new vacuum is given by: \be
2296: \label{d'}
2297: d'_\phi~u=du+[\phi,u]_\bullet=du+\phi\bullet u -(-1)^{|u|}u\bullet
2298: \phi ~~, \ee where $[u,v]_\bullet= u\bullet v-(-1)^{|u||v|}v\bullet u$
2299: stands for the {\em graded} commutator with respect to the boundary
2300: product $\bullet$ and $U(1)$ charge $|~.~|$, and $u$ is an element of
2301: the total boundary space ${\cal H}$. The string field equations of
2302: motion (\ref{eom}) are equivalent with the condition that $d'_\phi$
2303: squares to zero. The important observation is that $d'_\phi$ does not
2304: preserve the original boundary sectors $Hom(a,a)$, $Hom(a,b)$ etc, due
2305: to non-vanishing of either $\phi^{(1-n)}_{ab}$ or $\phi^{(1+n)}_{ba}$.
2306: Hence one cannot interpret the new background as a collection of two
2307: D-branes. In fact, there is no decomposition of ${\cal H}$ into new
2308: boundary sectors which would satisfy the basic conditions for such an
2309: interpretation (these conditions can be formulated \cite{com1} as the
2310: existence of a category structure compatibile with $d'_\phi$ and with
2311: the boundary product and bilinear form). Hence the new vacuum must be
2312: interpreted as a single object, namely as a composite of the original
2313: branes. In this case, ${\cal H}$ is viewed as the boundary sector of
2314: strings `stretching from this composite to itself', though whether
2315: such an interpretation can be implemented in some sort of sigma model
2316: requires a case by case analysis. The issue of some sigma model
2317: representation for such composites is secondary for our purpose, since
2318: we are interested in {\em off-shell} string dynamics, for which it is
2319: natural to take the string field theory approach as fundamental. From
2320: our perspective, {\em all} D-brane composites constructed in this
2321: manner are equally `fundamental', and their inclusion in the theory is
2322: a dynamical requirement. Whether the entirety of this dynamics admits
2323: some open sigma model interpretation is irrelevant for our
2324: considerations.
2325:
2326:
2327: \subsection{Graded D-brane composites as a
2328: representation of the extended moduli space}
2329:
2330: Let us consider an ungraded topological D-brane $a$ wrapping $L$, i.e.
2331: a flat bundle $E$ on $L$, with underlying connection $A$. The string
2332: field theory for strings whose endpoints lie on $a$ is described
2333: \cite{Witten_CS} by the Chern-Simons field theory on $E$. The
2334: associated moduli space ${\cal M}$ of string vacua is the moduli space
2335: of flat connections on $E$, which can be described through rank one
2336: solutions $\phi\in \Omega^1(End(E))$ to the Maurer-Cartan equation for
2337: the commutator algebra of the graded associative algebra
2338: $(\Omega^*(L),\wedge)$, which is the boundary algebra of topological
2339: A-type strings stretching from $a$ to $a$: \be
2340: \label{mc0}
2341: d\phi+\frac{1}{2}[\phi,\phi]_\wedge=0 \Longleftrightarrow
2342: d\phi+\phi\wedge \phi=0~~, \ee divided through the gauge group
2343: generated by transformations of the form: \be
2344: \label{gauge_pair0}
2345: \phi\rightarrow \phi-d\alpha-[\phi,\alpha]_\wedge~~. \ee In these
2346: equations, $[u,v]_\wedge=u\wedge v -(-1)^{rk u~rk v} v \wedge u$ is
2347: the graded commutator built from the usual wedge product of
2348: $End(E)$-valued forms and taken with respect to the grading given by
2349: form rank. Since $rk\alpha=rk\phi -1$ it follows that in equation
2350: (\ref{gauge_pair0}) the graded commutator is a genuine commutator.
2351: The tangent space to this moduli space is given by $H^1(End(E))$, as
2352: can be seen by considering the linearized form of (\ref{mc}) and
2353: (\ref{gauge_pair0}).
2354:
2355: An important problem in open topological string theory and mirror
2356: symmetry is to understand the significance of the so-called {\em
2357: extended moduli space} ${\cal M}_e$, obtained by considering
2358: deformations along directions in the other cohomology groups, namely
2359: $H^0(End(E))$, $H^2(End(E))$ and $H^3(End(E))$. The extended moduli
2360: space can be defined formally through the so-called `extended
2361: deformation theory' of \cite{Kontsevich_Felder}, which gives a precise
2362: technique for constructing a (formal) supermanifold ${\cal M}_e$ whose
2363: tangent space at the origin is given by the full cohomology group
2364: $H^*(End(E))$. Since $H^1(End(E))$ is a subspace of $H^*(End(E))$, the
2365: unextended moduli space ${\cal M}$ can be identified with a
2366: submanifold of ${\cal M}_e$. Points in the complement ${\cal
2367: M}_e-{\cal M}$ are understood as `generalized' vacua of the open
2368: string theory in the given boundary sector, i.e. topological string
2369: backgrounds which do not have an immediate geometric interpretation.
2370:
2371: The formal approach of extended deformation theory tells us nothing
2372: about the physical significance of the generalized backgrounds
2373: described by points lying in ${\cal M}_e-{\cal M}$. This problem has
2374: obstructed efforts to fulfill the program outlined in
2375: \cite{Witten_nlsm, Witten_mirror} of gaining a better understanding of
2376: mirror symmetry through a study of extended moduli spaces. It is a
2377: remarkable fact (pointed out in \cite{sc}) that the theory of graded
2378: D-branes allows us to give an {\em explicit} physical description of
2379: such points.
2380:
2381: To understand this, let us consider the case of graded D-branes $a$
2382: and $b=a[n]$ on $L$, whose underlying flat bundles coincide
2383: ($E_a=E_b:=E$) and whose grades differ by $n$. In this case, the
2384: string field $\phi$ belongs to the space ${\cal
2385: H}^1=\Omega^1(End(E))^{\oplus 2}\oplus \Omega^{1-n}(End(E))\oplus
2386: \Omega^{1+n}(End(E))$, and the classical equation of motion
2387: (\ref{eom}) is replaced by the Maurer-Cartan equation for the
2388: commutator algebra of the boundary algebra ${\cal H}$, which gives the
2389: equation of motion (\ref{eom}) for the string field theory
2390: (\ref{action_pair}): \be
2391: \label{mc}
2392: d\phi+\frac{1}{2}[\phi,\phi]_\bullet=0~~, \ee with the gauge group
2393: generated by transformations of the form (\ref{gauge}), which in our
2394: case become: \bea
2395: \label{gauge_pair}
2396: \delta\phi^{(1)}_{aa}~~~&=&-d\alpha^{(0)}_{aa}-
2397: [\phi^{(1)}_{aa},\alpha^{(0)}_{aa}]_{\bullet}-
2398: \phi^{(1+n)}_{ba}\bullet\alpha^{(-n)}_{ab}-
2399: \alpha^{(n)}_{ba}\bullet\phi^{(1-n)}_{ab}~~\nn\\
2400: \delta\phi^{(1)}_{bb}~~~&=&-d\alpha^{(0)}_{bb}-
2401: [\phi^{(1)}_{bb},\alpha^{(0)}_{bb}]_{\bullet}-
2402: \phi^{(1-n)}_{ab}\bullet
2403: \alpha^{(n)}_{ba}-\alpha^{(-n)}_{ab}\bullet\phi^{(1+n)}_{ba}~~\nn\\
2404: \delta\phi^{(1+n)}_{ba}&=&-d\alpha^{(n)}_{ba}-
2405: \phi^{(1+n)}_{ba}\bullet\alpha^{(0)}_{bb}-\alpha^{(0)}_{aa}\bullet
2406: \phi^{(1+n)}_{ba}-
2407: \phi^{(1)}_{aa}\bullet\alpha^{(n)}_{ba}-\alpha^{(n)}_{ba}\bullet\phi^{(1)}_{bb}~~\\
2408: \delta\phi^{(1-n)}_{ab}&=&-d\alpha^{(-n)}_{ab}-
2409: \phi^{(1-n)}_{ab}\bullet\alpha^{(0)}_{aa}-\alpha^{(0)}_{bb}
2410: \bullet\phi^{(1-n)}_{ab}-
2411: \phi^{(1)}_{bb}\bullet\alpha^{(-n)}_{ab}-\alpha^{(-n)}_{ab}
2412: \bullet\phi^{(1)}_{aa}~~,\nn \eea with a parameter
2413: $\alpha=\left[\begin{array}{cc}\alpha^{(0)}_{aa}& \alpha^{(n)}_{ba}\\
2414: \alpha^{(-n)}_{ab}&\alpha^{(0)}_{bb}\end{array}\right]$, where
2415: forms of rank outside of the range $0..3$ are of course vanishing.
2416:
2417: As in (\ref{d'}), the graded commutator is now taken with respect to
2418: the boundary product $\bullet$ and worldsheet charge $|~.~|$. For
2419: $n=0$, this equation describes deformations of the block-diagonal flat
2420: connection $A=A_a\oplus A_a$ to a flat connection on the bundle
2421: $E\oplus E$ (which need not have block-diagonal form). However, the
2422: moduli space of solutions of (\ref{mc}) for $0<|n|<3$ contains slices
2423: of the {\em extended} moduli space of flat connections on $E$. For
2424: example a graded D-brane system with $n=1$ contains deformations
2425: generated by $H^0(End(E))$ and $H^2(End(E))$, while for $n=2$ we
2426: obtain deformations generated by $H^3(End(E))$. Hence {\em one
2427: represents extended deformations of $E$ through condensation of
2428: boundary condition changing operators in a pair of graded
2429: topological D-branes based on $E$} (figure 2).
2430:
2431: \hskip 1.0 in
2432: \begin{center}
2433: \scalebox{0.5}{\input{moduli.pstex_t}}
2434: \end{center}
2435: \begin{center}
2436: Figure 2. {\footnotesize Points in the extended moduli space of an
2437: ungraded D-brane $a$ can be represented through condensation of
2438: boundary condition changing operators between $a$ and one of its
2439: higher shifts $b=a[n]$. For example, deformations along
2440: $H^0(End(E))$, $H^1(End(E))$ and $H^2(End(E))$ can be represented
2441: by condensing operators in the sectors $Hom(a,a[1])$, $Hom(a,a)$
2442: and $Hom(a[1],a)$ respectively.}
2443: \end{center}
2444:
2445: \
2446:
2447:
2448: A complete description of the extended moduli space requires
2449: consideration of all graded branes $a[n]$ with $n\in \Z$, and is
2450: discussed at that level of generality in \cite{sc}; the condensates
2451: resulting from a given graded D-brane pair only describe various
2452: slices through ${\cal M}_e$. One approach to the physical significance
2453: of points in ${\cal M}_e-{\cal M}$ is to study the quantum dynamics of
2454: the associated string field theories. A first step in this direction
2455: is the BV analysis carried out in this paper. This interpretation of
2456: graded D-brane condensates is the main reason for our interest in
2457: graded Chern-Simons theory.
2458:
2459: \subsection{Acyclic composites}
2460:
2461: The composite resulting from a condensation process can happen to be
2462: {\em acyclic}, i.e. the cohomology of the shifted BRST operator $d'$
2463: may vanish in all degrees. In this case, the theory expanded around
2464: the composite contains only BRST trivial states, and the resulting
2465: point in the (extended) moduli space can be interpreted as a closed
2466: string vacuum.
2467:
2468:
2469: \paragraph{Observation} The axioms of string field theory imply that the
2470: bilinear form $\langle .,. \rangle$ induces a perfect pairing between
2471: the BRST cohomology groups $H^k_{d'}({\cal H})$ and
2472: $H^{3-k}_{d'}({\cal H})$ for any worldsheet $U(1)$ charge $k$. In
2473: particular, acyclicity of a composite can be established by checking
2474: vanishing of only half of the cohomology groups $H^k_{d'}({\cal H})$.
2475:
2476: \subsection{Acyclic composites for $n=1$}
2477:
2478: \subsubsection{Assumptions}
2479:
2480: In this subsection we discuss a particular class of condensation
2481: processes which is somewhat similar to tachyon condensation in bosonic
2482: string theory in that it produces acyclic composites. Condensation
2483: processes of the type discussed below can take place for a graded
2484: D-brane pair satisfying the conditions:
2485:
2486: \
2487:
2488: (1) the relative grade of the pair is $n=1$
2489:
2490: \
2491:
2492: (2) $E_a$ and $E_b$ are isomorphic {\em as flat bundles}.
2493:
2494: \
2495:
2496: (3)
2497: $H^1(End(E_a))=H^1(End(E_b))=H^1(Hom(E_a,E_b))=H^2(Hom(E_a,E_b))=0$.
2498:
2499: \
2500:
2501: The argument presented below is closely related to ideas of
2502: \cite{Vafa_cs}, though our physical realization seems to be different.
2503:
2504: It is not hard to see that condition (2) amounts to the existence of a
2505: {\em covariantly constant} isomorphism between $E_a$ and $E_b$. This
2506: means a bundle isomorphism $f:E_a\rightarrow E_b$ which satisfies
2507: $df=0$ \footnote{It is not hard to see that such a map gives an
2508: isomorphism of the underlying flat structures. In the language of
2509: flat connections, the condition $df=0$ implies that $f$ induces a
2510: gauge transformation which takes $A_a$ to $A_b$ (one obtains
2511: $A_b=fA_af^{-1}-(df)f^{-1}$ upon choosing local frames for $E_a$ and
2512: $E_b$). Alternately, $f$ gives an isomorphism between flat
2513: trivializations of $E_a$ and $E_b$.} (remember that the
2514: differential $d$ on $Hom(E_a, E_b)$ is coupled to the background flat
2515: connections $A_a$ and $A_b$). The constraints on the first cohomology
2516: of $E_a$ and $E_b$ mean that these flat connections possess no
2517: deformations.
2518:
2519: \subsubsection{Construction of acyclic composites}
2520:
2521: With the assumptions discussed above, one can construct a solution
2522: $\phi$ of the string field equations of motion which leads to an
2523: acyclic D-brane composite. Indeed, let us look for a solution of the
2524: form $\phi=\left[\begin{array}{cc}0&0\\\phi_{ab}^{(0)} &0
2525: \end{array}\right]$. In this case, the equations of motion
2526: $d\phi+\phi\bullet\phi=0$ reduce to $d\phi_{ab}^{(0)}=0$, which says
2527: that $\phi_{ab}^{(0)}$ is covariantly constant section of
2528: $Hom(E_a,E_b)$. We shall further require that $\phi^{(0)}_{ab}$ is a
2529: bundle isomorphism, which is possible in view of our second
2530: assumption. Condensation of the boundary condition changing state
2531: $\phi^{(0)}_{ab}$ leads to a new D-brane background which can be
2532: viewed as composite of the D-branes $a$ and $b$. The theory expanded
2533: around this background is endowed with the shifted BRST operator given
2534: by $d'_\phi~u=d\phi+[\phi,u]_\bullet$. We wish to show that $d'_\phi$
2535: has trivial cohomology. Since $n=1$, the total boundary space of our
2536: D-brane system has non-vanishing components of charge $k=-1,0,1,2,3$
2537: and $4$. Moreover, Poincar\'e duality identifies the worldsheet BRST
2538: cohomologies in degrees $k$ and $3-k$, so that it suffices to show
2539: vanishing of $H^k_{d'}({\cal H})$ for $k=-1, 0$ and $1$.
2540:
2541: \subsubsection{Proof of acyclicity}
2542:
2543: \paragraph{Vanishing of $H^{-1}_{d'_\phi}({\cal H})$}
2544:
2545: A state of charge $-1$ has the form $u=\left[\begin{array}{cc} 0 &
2546: u_{ba}^{(0)}\\0 & 0 \end{array}\right]$. In this case, the BRST
2547: closure condition $d'_\phi u=0\Leftrightarrow du+\phi\bullet
2548: u+u\bullet \phi=0$ is equivalent with: \be
2549: u^{(0)}_{ba}\bullet\phi^{(0)}_{ab}=0~~,~~ du^{(0)}_{ba}=0~~,~~
2550: \phi_{ab}^{(0)}\bullet u_{ba}^{(0)}=0~~. \ee Since $\phi^{(0)}_{ab}$
2551: is a bundle isomorphism, this is equivalent with $u_{ba}^{(0)}=0$ i.e.
2552: $u=0$, which shows that $H^{-1}_{d'_\phi}({\cal H})=0$.
2553:
2554: \paragraph{Vanishing of $H^0_{d'_\phi}({\cal H})$}
2555:
2556: Given a charge zero state $u=\left[\begin{array}{cc} u_{aa}^{(0)}&
2557: u_{ba}^{(1)}\\0 & u_{bb}^{(0)} \end{array}\right]$, the BRST
2558: closure condition $d'_\phi u=0\Leftrightarrow du+\phi\bullet
2559: u-u\bullet \phi =0$ reads: \be
2560: \label{closed0}
2561: du_{aa}^{(0)}=u^{(1)}_{ba}\bullet\phi^{(0)}_{ab}~~,~~
2562: du^{(1)}_{ba}=0~~,~~ \phi_{ab}^{(0)}\bullet
2563: u^{(0)}_{aa}=u^{(0)}_{bb}\bullet \phi^{(0)}_{ab}~~,~~
2564: du_{bb}^{(0)}=-\phi_{ab}^{(0)}\bullet u_{ba}^{(1)}~~. \ee On the
2565: other hand, the exactness condition
2566: $u=d'_\phi\alpha=d\alpha+\phi\bullet \alpha+\alpha\bullet\phi$ (for an
2567: element $\alpha=\left[\begin{array}{cc} 0 & \alpha_{ba}^{(0)}\\0 & 0
2568: \end{array}\right]$ of charge $-1$) is equivalent with: \be
2569: \label{exact0}
2570: u^{(0)}_{aa}=\alpha^{(0)}_{ba}\bullet \phi^{(0)}_{ab}~~,~~
2571: u^{(1)}_{ba}=d\alpha^{(0)}_{ba}~~,~~
2572: u^{(0)}_{bb}=\phi^{(0)}_{ab}\bullet \alpha^{(0)}_{ba}~~. \ee
2573:
2574: The assumption $H^2(Hom(E_a,E_b))=0\Leftrightarrow
2575: H^1(Hom(E_b,E_a))=0$ allows us to solve the second equation of
2576: (\ref{closed0}) in the form: \be
2577: \label{ex1}
2578: u^{(1)}_{ba}=d\alpha^{(0)}_{ba}~~, \ee with $\alpha^{(0)}_{ba}$ a
2579: section of $Hom(E_b,E_a)$, determined up to \be
2580: \label{amb}
2581: \alpha^{(0)}_{ba}\rightarrow \alpha^{(0)}_{ba}+f^{(0)}_{ba}~~, \ee
2582: where $f^{(0)}_{ba}$ is a {\em covariantly constant} section of
2583: $Hom(E_b, E_a)$. Upon substituting (\ref{ex1}) into the first and
2584: fourth equations of (\ref{closed0}), these can be solved as: \be
2585: \label{ex2}
2586: u^{(0)}_{aa}=\alpha^{(0)}_{ba}\bullet \phi^{(0)}_{ab}+f^{(0)}_{aa}~~
2587: \ee and \be
2588: \label{ex3}
2589: u^{(0)}_{bb}=\phi^{(0)}_{ab}\bullet \alpha^{(0)}_{ba}+f^{(0)}_{bb}~~,
2590: \ee where $f^{(0)}_{aa}$ and $f^{(0)}_{bb}$ are covariantly constant
2591: sections of $End(E_a)$ and $End(E_b)$ and we used
2592: $d\phi^{(0)}_{ab}=0$. Using these expressions in the third equation
2593: of (\ref{closed0}) gives: \be \label{ex20} \phi^{(0)}_{ab}\bullet
2594: f^{(0)}_{aa}=f^{(0)}_{bb}\bullet \phi^{(0)}_{ab}~~, \ee a condition
2595: which can be satisfied upon choosing
2596: $f^{(0)}_{bb}=\phi^{(0)}_{ab}f^{(0)}_{aa}(\phi^{(0)}_{ab})^{-1}$ (this
2597: is covariantly constant since each of the factors is). With this
2598: choice, both $f^{(0)}_{aa}$ and $f^{(0)}_{bb}$ can be eliminated from
2599: (\ref{ex2}) and (\ref{ex3}) upon using the freedom (\ref{amb}) to
2600: redefine: \be\label{shift0} \alpha^{(0)}_{ba}\rightarrow \alpha^{'(0)}_{ba}:=
2601: \alpha^{(0)}_{ba}+ f^{(0)}_{aa}\bullet (\phi^{(0)}_{ab})^{-1}~~. \ee
2602: This shows that equations (\ref{ex1}), (\ref{ex2}) and (\ref{ex3}) can
2603: be brought to the form (\ref{exact0}). We conclude that
2604: $H^0_{d'_\phi}({\cal H})$ vanishes as well.
2605:
2606:
2607: \paragraph{Vanishing of $H^1_{d'_\phi}({\cal H})$}
2608:
2609: If $u=\left[\begin{array}{cc} u_{aa}^{(1)}& u_{ba}^{(2)}\\u_{ab}^{(0)}
2610: & u_{bb}^{(1)} \end{array}\right]$ is a state with $|u|=1$, then
2611: the BRST closure condition $d'_\phi u=0\Leftrightarrow
2612: du+[\phi,u]_\bullet =0$ is equivalent with the system: \bea
2613: \label{closed}
2614: du_{aa}^{(1)}+u^{(2)}_{ba}\bullet\phi^{(0)}_{ab}&=&0~~\nn\\
2615: du^{(2)}_{ba}&=&0~~\nn\\
2616: du_{ab}^{(0)}+\phi_{ab}^{(0)}\bullet u^{(1)}_{aa}+u^{(1)}_{bb}\bullet
2617: \phi^{(0)}_{ab}&=&0~~\\
2618: du_{bb}^{(1)}+\phi_{ab}^{(0)}\bullet u_{ba}^{(2)}&=&0~~.\nn \eea
2619:
2620:
2621: Using the assumption $H^2(Hom(E_b,E_a))=0$, the second equation can be
2622: solved as: \be
2623: \label{exact1}
2624: u_{ba}^{(2)}=d\alpha_{ba}^{(1)}~~, \ee for some one form
2625: $\alpha_{ba}^{(1)}$ with coefficients in $Hom(E_b,E_a)$. Upon
2626: substituting this into the first equation and using
2627: $d\phi^{(0)}_{ab}=0$, one obtains: \be
2628: \label{exact2}
2629: u_{aa}^{(1)}=d\alpha_{aa}^{(0)}-\alpha^{(1)}_{ba}\bullet\phi^{(0)}_{ab}~~,
2630: \ee for some section $\alpha_{aa}^{(0)}$ of $End(E_a)$. To arrive at
2631: this equation, we made use of the assumption $H^1(End(E_a))=0$. The
2632: section $\alpha^{(0)}_{aa}$ is determined up to transformations of the
2633: form: \be
2634: \label{ambiguity}
2635: \alpha_{aa}^{(0)}\rightarrow \alpha_{aa}^{(0)}+f^{(0)}_{aa}~~, \ee
2636: with $f_{aa}^{(0)}$ a {\em covariantly constant} (flat) section of
2637: $End(E_a)$. Using the solution for $u^{(2)}_{ba}$ and the condition
2638: $d\phi^{(0)}_{ab}=0$, we can similarly solve the last equation of
2639: (\ref{closed}): \be
2640: \label{exact3}
2641: u_{bb}^{(1)}=d\alpha_{bb}^{(0)}+\phi_{ab}^{(0)}\bullet
2642: \alpha^{(1)}_{ba}~~, \ee where we used the assumption
2643: $H^1(End(E_b))=0$. Finally, combining (\ref{exact1}) and
2644: (\ref{exact2}) allows us to solve the third equation in
2645: (\ref{closed}): \be
2646: \label{exact4_intmd}
2647: u_{ab}^{(0)}=\phi_{ab}^{(0)}\bullet
2648: \alpha_{aa}^{(0)}-\alpha^{(0)}_{bb} \bullet
2649: \phi^{(0)}_{ab}+f^{(0)}_{ab}~~, \ee with $f_{ab}^{(0)}$ a covariantly
2650: constant section of $Hom(E_a,E_b)$.
2651:
2652: Since both sections $f^{(0)}_{ab}$ and $\phi_{ab}^{(0)}$ of
2653: $Hom(E_a,E_b)$ are covariantly constant, the section
2654: $g^{(0)}_{aa}:=(\phi_{ab}^{(0)})^{-1}\circ f_{ab}^{(0)}$ of $End(E_a)$
2655: satisfies $dg^{(0)}_{aa}=0$. Upon using the ambiguity
2656: (\ref{ambiguity}), we can therefore absorb $f_{aa}^{(0)}$ in
2657: $\alpha_{aa}^{(0)}$ by defining: \be
2658: \alpha_{aa}^{'(0)}=\alpha_{aa}^{(0)}+g_{aa}^{(0)}~~. \ee This allows
2659: us to re-write (\ref{exact4_intmd}) in the
2660: form: \be u_{ab}^{(0)}=\phi_{ab}^{(0)}\bullet
2661: \alpha_{aa}^{'(0)}-\alpha_{bb}^{(0)} \bullet \phi^{(0)}_{ab}~~. \ee
2662:
2663: Hence we can assume without loss of generality that
2664: $\alpha_{aa}^{(0)}$ has been chosen such that: \be
2665: \label{exact4}
2666: u_{ab}^{(0)}=\phi_{ab}^{(0)}\bullet
2667: \alpha_{aa}^{(0)}-\alpha^{(0)}_{bb} \bullet \phi^{(0)}_{ab}~~. \ee
2668: The last step is to define the state $\alpha=\left[\begin{array}{cc}
2669: \alpha_{aa}^{(0)}&\alpha_{ba}^{(1)}\\0 & \alpha_{bb}^{(0)}
2670: \end{array}\right]$, which has charge zero. It is then easy to
2671: check that equations (\ref{exact1}), (\ref{exact2}) and (\ref{exact3})
2672: and (\ref{exact4}) are equivalent with the condition: \be
2673: u=d'_\phi\alpha=d\alpha+[\phi,\alpha]_\bullet~~. \ee Thus every
2674: $d'_\phi$-closed degree one state is $d'_\phi$-exact, which means that
2675: the BRST complex of $d'_\phi$ is also trivial in degree one. This
2676: shows that the composite obtained by condensation of $\phi_{ab}^{(0)}$
2677: is acyclic.
2678:
2679: Note that the boundary condition changing state $\phi^{(0)}_{ab}$ is
2680: physical, since it is BRST-closed and has charge $1$. In fact, the
2681: (`unshifted') BRST cohomology $H^1_d({\cal H}_{ab})$ in degree one
2682: coincides with the space $H^0_d(Hom(E_a,E_b))$ of covariantly constant
2683: sections of $Hom(E_a,E_b)$.
2684:
2685:
2686: \paragraph{Observations}
2687:
2688: (1) We wish to stress that one does {\em not} need to consider the
2689: extended action (\ref{extended_action}), nor the extended boundary
2690: space ${\cal H}_e$, in order to understand condensation processes.
2691:
2692: (2) We mention that the assumption $E_a\approx E_b$ is necessary only
2693: for establishing acyclicity in the sector with $U(1)$ charge $-1$.
2694: In the sector with vanishing $U(1)$ charge,
2695: this condition can be avoided by choosing
2696: $f_{aa}^{(0)}=g_{ba}^{(0)}\bullet \phi^{(0)}_{ab}$ and
2697: $f_{bb}^{(0)}=\phi^{(0)}_{ab}\bullet g_{ba}^{(0)}$
2698: in equations (\ref{ex2}) and (\ref{ex3}), where
2699: $g_{ba}$ is a covariantly constant section of
2700: $Hom(E_b,~E_a)$. Then equation (\ref{ex20}) is automatically satisfied
2701: and one shifts $\alpha_{ba}^{(0)}$ to $\alpha_{ba}^{'(0)}=\alpha_{ba}^{(0)}+
2702: g_{ba}^{(0)}$ instead of equation (\ref{shift0}).
2703: Hence vanishing of $H_{d'}^0({\cal H})$
2704: can be established without requiring
2705: invertibility of $\phi^{(0)}_{ab}$. A similar
2706: modification of the argument is possible when showing
2707: vanishing of $H^1_{d'}({\cal H})$; namely one chooses $f_{ab}^{(0)}$
2708: to be of the form $f_{ab}^{(0)}=\phi_{ab}^{(0)}\bullet g_{aa}^{(0)}$,
2709: with $g_{aa}^{(0)}$ a covariantly constant section of $End(E_a)$.
2710: However, the condition that $\phi_{ab}$ is invertible is crucial
2711: for assuring vanishing of $H^{-1}_{d'}({\cal H})$.
2712:
2713: (3) The fact that a D-brane pair satisfying our conditions can
2714: condense to an acyclic composite gives some justification for viewing
2715: these systems as `topological brane-antibrane pairs'. However, we
2716: were able to establish the existence of such processes only under
2717: restrictive topological assumptions on the original background. To
2718: understand just how restrictive our hypotheses are, consider the case
2719: of two singly-wrapped graded D-branes, for which $E_a=E_b={\cal O}_L$,
2720: the trivial line bundle over $L$ (endowed with the trivial flat
2721: connection $A_a=A_b=0$). In this situation, our argument requires
2722: vanishing of the cohomology group $H^1(L)\approx H^2(L)$, which in
2723: particular means that the cycle $L$ must be a rational homology sphere.
2724: It is well-known that Calabi-Yau threefolds contain a wealth of
2725: special Lagrangian cycles which do not satisfy this
2726: condition (for example, special Lagrangian
2727: 3-tori, which according to the conjecture of \cite{SYZ} should give a
2728: fibration of $X$ if $X$ admits a geometric mirror). Our argument does
2729: not apply to such cycles, even in the singly-wrapped case. It is an
2730: interesting problem to study composite formation in this more general
2731: situation.
2732:
2733:
2734: \section{Direct construction of the master action for D-brane pairs}
2735:
2736: In this section we give a direct construction of the classical BV
2737: actions for the string field theories of graded D-brane pairs. This
2738: will allow us to recover the extended theory (\ref{extended_action})
2739: by applying the more familiar cohomological formalism. Moreover, we
2740: shall give a BV level proof that there are precisely six inequivalent
2741: classical BV actions, namely those associated to the relative gradings
2742: $n=0,\,\pm1,\,\pm2$ and $(n\ge 3,\,\,n\le-3)$, thus leading to a
2743: $\Z_6$ multiplicity of such systems. This gives the string field
2744: theory realization of an observation made in
2745: \cite{Douglas_Kontsevich}.
2746:
2747: As already mentioned at the end of Section 4, the BV algorithm
2748: prolongs a basis for the space of classical fields to a basis ${\cal
2749: B}= (e^{*\alpha}, e_{\alpha})$ for the space of extended fields,
2750: which is Darboux for the underlying odd symplectic form. Since the
2751: coefficients of the latter are particularly simple in this basis, the
2752: BV bracket takes the form (\ref{Darboux_bracket}).
2753:
2754: In this section we apply the BV algorithm in a component approach, and
2755: carry out the computations in terms of the usual wedge product of
2756: forms (which we shall write as juxtaposition in order to simplify
2757: notation), rather than in terms of the extended boundary product $*$
2758: used in the previous section. This makes for an explicit
2759: presentation, at the cost of introducing somewhat complicated
2760: formulae. A more synthetic description (which uses the results of the
2761: geometric formalism in order to simplify certain steps) is given in
2762: the next section. As discussed in Section 5 \ref{sec:gradedpairs}, it
2763: is enough to consider the cases with relative grading $n\equiv
2764: n_b-n_a\ge 0$, since the rest can be obtained by reversing the roles
2765: of $a$ and $b$ \footnote{Two D-brane pairs with relative gradings $n$
2766: and $-n$ are related by the conjugation operation discussed in
2767: \cite{sc}. While this is a `symmetry' at the classical level, two
2768: systems related in this manner should not be identified. The
2769: conjugation symmetry of \cite{sc} is akin to the charge conjugation
2770: of particle physics, and conjugated configurations must be
2771: considered physically distinct.}.
2772:
2773:
2774: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2775: \subsection{The BV action for $n=1$ }
2776:
2777: $${}$$
2778:
2779:
2780:
2781: For the case $n=1$, the classical fields are two one-forms, a two-form
2782: and a zero-form: \be {\phi}=\pmatrix{ \phi^{(1)} & \phi^{(2)}\cr
2783: \phi^{(0)} & \phi^{'(1)} \cr}~~.
2784: \label{eq:n=1}
2785: \ee In terms of the usual wedge product (\ref{wedge}), the classical
2786: action (\ref{action_pair}) reads: \bea S(\phi)=\int_L &&tr_a\left[
2787: {1\ov 2}\left(\p{1}d\p{1}-\p{2}d\p{0}\right)+{1\ov
2788: 3}\left(\p{1}\p{1}\p{1}+
2789: \p{1}\p{2}\p{0}\right)\right]~~ \\
2790: -\int_{L}&&tr_b\left[{1\ov
2791: 2}\left(\hp{1}d\hp{1}-\p{0}d\p{2}\right)+{1\ov 3}\left(
2792: \hp{1}\hp{1}\hp{1}+\hp{1}\p{0}\p{2}\right)\right]~~.\nn \eea
2793:
2794: Upon substituting $\alpha=-C_1\lambda$ into the gauge transformations
2795: (\ref{gauge_pair0}), where $\lambda$ is a Grassmann-odd constant
2796: parameter and: \be C_1=\pmatrix{ c_1^{(0)} & c_1^{(1)}\cr 0&
2797: {c'}_1^{(0)} \cr}~~,
2798: \label{eq:n1gen1}
2799: \ee is the corresponding matrix of ghosts, we find the BRST
2800: transformations of the physical fields: \bea &&\delta^{(1)}\p{1}=
2801: \left[dc_1^{(0)}+[\p{1},\,c_1^{(0)}]- c_1^{(1)}\p{0}\right]\lambda
2802: \nonumber\\
2803: &&\delta^{(1)}\p{2}=
2804: \left[dc_1^{(1)}+(\p{1}c_1^{(1)}+\p{2}{c'}_1^{(0)} -c_1^{(0)}\p{2}+
2805: c_1^{(1)}\hp{1})\right]\lambda
2806: \nonumber\\
2807: &&\delta^{(1)}\p{0}= (\p{0}c_1^{(0)}-{c'}_1^{(0)}\p{0})\lambda\nonumber\\
2808: &&\delta^{(1)} \hp{1}=\left[d{c'}_1^{(0)}
2809: +[\hp{1},\,{c'}_1^{(0)}]-\p{0}c_1^{(1)}\right] \lambda
2810: \label{brst}~~.
2811: \eea Requiring nilpotence of $\delta^{(1)}$ leads to the following
2812: ghost transformations: \be \delta^{(1)} c_1^{(0)}= c_1^{(0)}
2813: c_1^{(0)}\lambda~~, ~~\delta^{(1)} {c'}_1^{(0)}= {c'}_1^{(0)}
2814: {c'}_1^{(0)}\lambda~~, ~~\delta^{(1)}
2815: c_1^{(1)}=(c_1^{(0)}c_1^{(1)}+c_1^{(1)}{c'}_1^{(0)})\lambda~~.
2816: \label{ghn1brst0}
2817: \ee
2818:
2819: It is easy to see that the gauge transformations (\ref{gauge_pair})
2820: vanish on-shell for a particular choice of parameters.
2821: Oversimplifying, consider the transformation\footnote{We remind the
2822: reader that the cohomological (BRST) approach extends the gauge
2823: algebra ${\bf g}$ of Subsection 2.2.2 to the jet bundle associated
2824: with our fields. Thus, one views the various partial derivatives of
2825: a generator as being independent quantities. This is why we can
2826: consider `differential zero modes', i.e. expressions involving the
2827: exterior derivative of some generator. The jet bundle extension is a
2828: standard device for describing the on-shell algebraic structure of
2829: gauge transformations and observables, and forms the heart of the
2830: modern approach to BRST quantization. We refer the reader to
2831: \cite{Stasheff_bv} for an elegant review of the jet bundle formalism
2832: and further references.} $\delta\p2 = -d\alpha^{(1)} -\p1
2833: \alpha^{(1)}+ \alpha^{(1)}\hp1$ and take $\alpha^{(1)}$ to be of the
2834: form $\alpha^{(1)}
2835: =-(d\beta^{(0)}+\p{1}\beta^{(0)}+\beta^{(0)}\hp{1})$, with
2836: $\beta^{(0)}$ a section of $Hom(E_b,E_a)$. Substituting this into the
2837: gauge transformation, one finds that $\delta\p2$ vanishes when $\p1$
2838: and $\hp1$ satisfy their equations of motion (in this exemplification
2839: we ignored the role of $\alpha^{(0)}$ and $\alpha^{'(0)}$ in the BRST
2840: transformation). It follows that our gauge algebra is {\em
2841: reducible}. In such a situation, one must introduce ghosts for
2842: ghosts. Accordingly, we consider a second generation ghost: \be
2843: C_2=\pmatrix{ 0 & c_2^{(0)}\cr 0 & 0\cr}~~,
2844: \label{eq:n1gen2}
2845: \ee which allows us to extend the BRST transformation of $c_1^{(1)}$
2846: by adding: \bea
2847: &&{\delta}_2c_1^{(1)}=\left(dc_2^{(0)}+\p{1}c_2^{(0)}-c_2^{(0)}\hp{1}
2848: \right) \lambda~~. \eea Allowing $\delta_2$ to act trivially on the
2849: physical fields $\phi$ and requiring nilpotence of $\delta^{(2)}\equiv
2850: \delta^{(1)}+\delta_2$ leads to the following corrections to the
2851: transformations of $c_1^{(0)}$ and ${c'}_1^{(0)}$: \be
2852: {\delta}_2c_1^{(0)}=c_2^{(0)}\p{0}\lambda~~,~~
2853: {\delta}_2{c'}_1^{(0)}=\p{0}c_2^{(0)}\lambda~~ \ee and specifies the
2854: BRST transformation of the second generation ghost: \bea \delta_2
2855: c_2^{(0)}=(-c_1^{(0)}c_2^{(0)}+c_2^{(0)}{c'}_1^{(0)})\lambda~~.
2856: \label{ghn1brst1}
2857: \eea Since there are no further zero modes, the gauge algebra is first
2858: order reducible and the full BRST transformations are given by: \bea
2859: \delta\equiv \delta^{(2)}=\delta^{(1)}+\delta_2
2860: \label{eq:fullBRSTn1}
2861: \eea where we let $\delta^{(1)}$ act trivially on second generation
2862: ghosts: $\delta^{(1)} c_2^{(0)}=0$.
2863:
2864:
2865: The BV construction now introduces antifields $\Phi^*, C_1^*$ and
2866: $C_2^*$ for each of the fields $\phi, C_1$ and $C_2$: \be
2867: \Phi^*=\pmatrix{ \phi^{*(2)}&\phi^{*(3)}\cr
2868: \phi^{*(1)}&\phi'^{*(2)}\cr}~~,~~ C_1{}^*=\pmatrix{ c_1^{*(3)} & 0
2869: \cr c_1^{*(2)} & {c'}_1^{*(3)}\cr}~~,~~ C_2{}^*=\pmatrix{ 0 & 0\cr
2870: c_2^{*(3)} & 0\cr}~~. \ee
2871: \noindent
2872: Note that the antifield of each matrix block sits in the transposed
2873: position in the full matrix, for example ($\phi^{(1)}$,
2874: $\phi^{*(2)}$), ($\phi^{'(1)}$, $\phi^{'*(2)}$), ($\phi^{(2)}$,
2875: $\phi^{*(1)}$) and ($\phi^{(0)}$, $\phi^{*(3)}$) form field-antifield
2876: pairs. The classical action $S$ is extended by adding a term of the
2877: form $S_1=tr(\Phi^* \delta\phi/\lambda+ C_1^* \delta
2878: C_1/\lambda+\dots)$~. The Grassmann parities of antifields are chosen
2879: such that $S_1$ is even: $g({\rm field})+g({\rm antifield})={\hat
2880: 1}~(mod~2)$.
2881:
2882: Following this procedure we write the first order action: \bea S_1&&
2883: =\int_L\,tr_a\Big[~
2884: \phi^{*(2)}\left(dc_1^{(0)}+[\p{1},\,c_1^{(0)}]-c_1^{(1)}\p{0}\right)
2885: \nonumber\\
2886: &&~~~~~~~~ +c_1^{*(3)}\left(c_1^{(0)}c_1^{(0)}+c_2^{(0)}\p{0}\right)
2887: +\phi^{*(3)}\left( \p{0}c_1^{(0)}-{c'}_1^{(0)}\p{0} \right)\Big]\nn\\
2888: &&+\int_L\,tr_b\Big[~ \phi^{*(1)}\left(
2889: dc_1^{(1)}+(\p{1}c_1^{(1)}+\p{2}{c'}_1^{(0)}-c_1^{(0)}\p{2}+
2890: c_1^{(1)}\hp{1}) \right)\nn\\
2891: &&~~~~~~~~+
2892: \phi'^{*(2)}\left(d{c'}_1^{(0)}+[\hp{1},\,{c'}_1^{(0)}]-\p{0}c_1^{(1)}\right)
2893: \nonumber\\
2894: &&~~~~~~~~+
2895: c_1^{*(2)}\left(c_1^{(0)}c_1^{(1)}+c_1^{(1)}{c'}_1^{(0)}+dc_2^{(0)}+\p{1}c_2^{(0)}-
2896: c_2^{(0)}\hp{1}\right)
2897: \nonumber\\
2898: && {}~~~~~~~+
2899: {c'}^{*(3)}_1\left({c'}_1^{(0)}{c'}_1^{(0)}+\p{0}c_2^{(0)}\right) +
2900: c_2^{*(3)}\left( -c_1^{(0)}c_2^{(0)}+c_2^{(0)}{c'}_1^{(0)} \right)
2901: \Big]~~. \eea
2902:
2903: Higher order terms in the antifield expansion $S_{BV}=S_0+S_1+S_2+...$
2904: are constructed from the requirement that $S_{BV}$ satisfies the
2905: master equation. Since vanishing of $\left\{S,\,S_1\right\}$ is
2906: guaranteed by gauge invariance of the classical action, the next step
2907: is to compute the BV bracket $\left\{S_1,\,S_1\right\}$ and find a
2908: non-vanishing result, linear in antifields\footnote{In fact, the BRST
2909: transformation of $c_2^{(0)}$ can also be inferred from the
2910: requirement $\left\{S_1,\,S_1\right\}$ contains no terms which are
2911: independent of antifields. }. Thus, we have to supplement $S_1$ by
2912: a further term $S_2$ quadratic in antifields, such that $\{ S_1,
2913: S_1\}+2\{S ,S_2\}=0$. It is not very hard to check that the choice:
2914: \be S_2=\int_L\,tr_a\left(c_2^{(0)}\phi^{*(1)} \phi^{*(2)}-
2915: c_2^{(0)}\phi'^{*(2)}\phi^{*(1)}\right) ~~, \ee assures that the BV
2916: action: \bea
2917: \label{bv1}
2918: S_{BV}=S+S_1+S_2 \eea obeys the master equation. Some details of the
2919: relevant computation are given in Appendix C.
2920:
2921: To show that the BV action coincides with the coordinate-free
2922: expression (\ref{extended_action}), we set: \bea
2923: \label{coords1}
2924: {\hat \phi}&=&\pmatrix{ c_1^{(0)}+\phi^{(1)}-\phi^{*(2)} +c_1^{*(3)}&
2925: c_2^{(0)}+c_1^{(1)}+\p{2} -\phi^{*(3)}\cr
2926: \phi^{(0)}+\phi^{*(1)}+c_1^{*(2)}+c_2^{*(3)}& {c'}_1^{(0)}
2927: +\hp{1}+{\phi'}^{*(2)}-{c'}_1^{*(3)}
2928: \cr}\nn\\
2929: &=&c_2+c_1+\phi+\phi^*+c_1^*+c_2^*~~,
2930: \label{eq:superfieldn1}
2931: \eea where we defined: \be \phi^*=\pmatrix{
2932: -\phi^{*(2)}&-\phi^{*(3)}\cr \phi^{*(1)}&\phi'^{*(2)}\cr}~~,~~
2933: c_1{}^*=\pmatrix{ c_1^{*(3)} & 0 \cr c_1^{*(2)} &
2934: -{c'}_1^{*(3)}\cr}~~,~~ \ee and: \be
2935: c_1=C_1~~,~~c_2=C_2~~,~~c_2{}^*=C_2^*~~,
2936: \label{notghn1}
2937: \ee notations which will be useful in the next section. Counting
2938: worldsheet $U(1)$ charges and Grassmann parities shows that ${\hat
2939: \phi}$ is an element of $M={\cal H}_e^1$. Relation (\ref{coords1})
2940: can be viewed as the expression of the vector ${\hat \phi}$ in the
2941: particular linear coordinates on $M$ built by the BV algorithm.
2942:
2943: One can check by direct computation that substitution of
2944: (\ref{coords1}) into the extended action (\ref{extended_action})
2945: recovers the expanded form (\ref{bv1}) upon expressing everything in
2946: terms of the usual wedge product. This shows that the BV action
2947: (\ref{bv1}) is simply the form of the extended action in the
2948: particular linear coordinates (\ref{coords1}). To show equivalence
2949: with the geometric formalism of Section 4, we must also check that
2950: (\ref{coords1}) gives Darboux coordinates for the odd symplectic form
2951: (\ref{omega}). As mentioned in Subsection 4.8.1, the odd symplectic
2952: form is completely determined by its restriction to {\em even} vector
2953: fields, which can be identified with the form $\omega_0$ of equation
2954: (\ref{omega0_wedge}). It thus suffices to check that
2955: (\ref{omega0_wedge}) reduces to Darboux form in the coordinates
2956: (\ref{coords1}). This follows by direct computation upon substituting
2957: (\ref{coords1}) into (\ref{omega0_wedge}) \footnote{To obtain Darboux
2958: coordinates per se, one must choose bases for the spaces of
2959: bundle-valued forms corresponding to each block in the decomposition
2960: of our matrices of morphisms. We leave the details to the reader.}:
2961: \bea \omega_0=tr_a\Big[ &&(\delta c_1^{*(3)}\we \delta
2962: c_1^{(0)}-\delta c_1^{(0)}\we\delta c_1^{*(3)})+
2963: (\delta\phi^{*(2)}\we \delta\p1 - \delta\p1\we\delta\phi^{*(2)})\nn\\
2964: +&&(\delta c_2^{*(3)}\we\delta c_2^{(0)}-\delta c_2^{(0)}\we \delta
2965: c_2^{*(3)})
2966: +(\delta c_1^{*(2)}\we\delta c_1^{(1)}-\delta c_1^{(1)}\we \delta c_1^{*(2)})\nn\\
2967: +&&(\delta\phi^{*(1)}\we \delta\p2-\delta\p2\we\delta\phi^{*(1)})\Big]\nn\\
2968: +tr_b\Big[&& (\delta {c'}_1^{*(3)}\we\delta {c'}_1^{(0)}-\delta
2969: {c'}_1^{(0)}\we\delta {c'}_1^{*(3)})
2970: +(\delta\phi'^{*(2)}\we \delta\hp1-\delta\hp1\we\phi'^{*(2)})\nn\\
2971: +&&(\delta\phi^{*(3)}\we
2972: \delta\p0-\delta\p0\we\delta\phi^{*(3)})\Big]~~. \eea We conclude
2973: that the homological construction for a D-brane pair with relative
2974: grading $n=1$ agrees with the BV system discussed in Section 4.
2975:
2976: \subsection{The BV action for $n=2$}
2977:
2978: We next consider D-brane pairs with relative grading $n=2$. For such
2979: systems, the classical fields are two one-forms and a three form: \be
2980: \phi =\pmatrix{ \phi^{(1)} & \phi^{(3)}\cr 0 & {\phi'}^{(1)} \cr}~~.
2981: \ee Since $n$ is even, the relative grading plays no role in the
2982: boundary product $\bullet$, which reduces to the usual wedge product
2983: (\ref{wedge}). The classical action is: \bea
2984: \label{action_2}
2985: \!\!\!\!\!\!\!\!\!\!\!\!\!\!\!\! S(\phi)&&=\int_L {tr_a\left[\frac 12
2986: \p1 d\p1+\frac 13 \p1\p1\p1\right]} +\int_{L}{tr_b\left[\frac 12
2987: \hp1 d\hp1+\frac 13 \hp1\hp1\hp1 \right]}~~. \eea Note that the
2988: field $\phi^{(3)}$ does not appear in the action, and thus its
2989: components define classical flat directions in the moduli space of the
2990: theory. We stress that, even though the action (\ref{action_2}) is
2991: invariant with respect to the shift symmetry $\phi^{(3)}\rightarrow
2992: \phi^{(3)}+b^{(3)}$ (with $b^{(3)}$ an arbitrary 3-form valued in the
2993: bundle $Hom(E_b,E_a))$, we do {\em not} include such symmetries in our
2994: gauge algebra. The reason is that a nonzero background value of
2995: $\phi^{(3)}$ corresponds to condensation of boundary condition
2996: changing states between the graded branes $a$ and $b$, and thus leads
2997: to a {\em physically relevant} shift of the string vacuum. This
2998: forbids us to identify $\phi^{(3)}$ and $\phi^{(3)}+b^{(3)}$. This
2999: means that we cannot treat shifts of $\phi^{(3)}$ as {\em gauge}
3000: symmetries.
3001:
3002:
3003: Since one of the classical fields is a massless 3-form, we expect
3004: first, second and third generation ghosts: \be C_1=\pmatrix{
3005: c_1^{(0)}&c_1^{(2)}\cr 0&{c'}_1^{(0)}\cr}~~,~~ C_2=\pmatrix{
3006: 0&c_2^{(1)}\cr 0&0\cr}~~,~~ C_3=\pmatrix {0&c_3^{(0)}\cr 0&0\cr}~~.
3007: \ee The BRST transformations of the classical fields read: \bea
3008: \delta^{(1)}\p1=\left(dc_1^{(0)}+[\p{1},\,c_1^{(0)}]\right)\lambda~~,~~
3009: \delta^{(1)}\hp1=\left(d{c'}_1^{(0)}+[\hp{1},\,{c'}_1^{(0)}]\right)\lambda\\
3010: \delta^{(1)}\p3=\left(dc_1^{(2)}-c_1^{(2)}\hp{1}+\p{1}c_1^{(2)}-c_1^{(0)}\p{3}+
3011: \p{3}{c'}_1^{(0)}\right)\lambda~~.
3012: \label{brst2o}~~\nn
3013: \eea The transformations of first generation ghosts are derived by
3014: requiring nilpotence of $\delta^{(1)}$ on $\phi$: \be \delta^{(1)}
3015: c_1^{(2)}=\left(c_1^{(0)}c_1^{(2)}+c_1^{(2)}{c'}_1^{(0)}\right)\lambda~~,~~
3016: \delta^{(1)} c_1^{(0)}=c_1^{(0)}c_1^{(0)}\lambda~~,~~ \delta^{(1)}
3017: c_1^{'(0)}=c_1^{'(0)}{c'}_1^{(0)}\lambda~~. \ee Once again,
3018: (\ref{brst2o}) has on-shell zero modes, which requires us to extend
3019: the BRST transformation of $c_1^{(2)}$ by adding the variation: \be
3020: \delta_2 c_1^{(2)}
3021: =(dc_2^{(1)}+\p{1}c_2^{(1)}+c_2^{(1)}\hp{1})\lambda~~.
3022: \label{eq:2ndgengh}
3023: \ee Allowing $\delta_2$ to act trivially on the physical fields
3024: ($\delta_2\phi=0$) and requiring nilpotence of
3025: $\delta^{(2)}=\delta^{(1)}+\delta_2$ implies that the second
3026: generation ghosts must transform as \be {\delta_2
3027: c_2^{(1)}}=(-c_1^{(0)}c_2^{(1)}+ c_2^{(1)}{c'}_1^{(0)})\lambda~~.
3028: \ee Equation (\ref{eq:2ndgengh}) has further zero modes which lead to
3029: the third generation ghosts and to the transformations: \be {\delta_3
3030: c_2^{(1)}}=\left(dc_3^{(0)}-c_3^{(0)}\hp{1}+\p{1}c_3^{(0)}\right)\lambda~~,~~
3031: \delta_3
3032: c_3^{(0)}=\left(c_1^{(0)}c_3^{(0)}+c_3^{(0)}{c'}_1^{(0)}\right)\lambda~~.
3033: \label{eq:3rdgengh}
3034: \ee Since there are no further zero modes, the theory is second order
3035: reducible and we conclude that the full BRST transformations are given
3036: by: \be
3037: \delta\equiv\delta^{(3)}=\delta^{(2)}+\delta_3=\delta^{(1)}+\delta_2+\delta_3
3038: \label{eq:fullBRSTn2}
3039: \ee where $\delta^{(1)}$ acts trivially on the second and third
3040: generation ghosts, $\delta_2$ acts trivially on the physical fields
3041: and the third generation ghosts and $\delta_3$ acts trivially on the
3042: physical fields and the first generation ghosts.
3043:
3044: We next add the antifields: \be \Phi^*=\pmatrix{ \phi^{*(2)}& 0\cr
3045: \phi^{*(0)}&{\phi'}^{*(2)}\cr}~~,~~ C_1^*=\pmatrix{ c_1^{*(3)}&0\cr
3046: c_1^{*(1)} &{c'}_1^{*(3)}\cr}~~,~~ C_2^*=\pmatrix{ 0&0\cr c_2^{*(2)}
3047: &0\cr}~~,~~ C_3^*=\pmatrix{ 0&0\cr c_3^{*(3)}&0\cr}~~, \ee and build
3048: the first order action $S_1$ as in the previous section.
3049:
3050: Since $\{S_1, S_1\}$ does not vanish, the last step is to add a term
3051: $S_2$ quadratic in antifields, such that the classical master equation
3052: is satisfied. This leads to the full BV action: \bea
3053: \label{bv2}
3054: S_{BV}=\int_L\, tr_a&&\Big[{1\ov 2}\p{1}d\p{1}+{1\ov 3}\p{1}\p{1}\p{1}
3055: +\phi^{*(2)}\left(dc_1^{(0)}+[\p{1},\,c_1^{(0)}]\right)\nn\\
3056: &&+c_1^{*(3)}\,c_1^{(0)}c_1^{(0)}
3057: -c_3^{(0)}\left({\phi'}^{*(2)}c_1^{*(1)}-c_1^{*(1)}\phi^{*(2)}
3058: +\phi^{*(0)}
3059: c_1^{*(3)}-{c'}^{*(3)}_1\phi^{*(0)} \right)\nn\\
3060: &&+c_2^{(1)}\left({\phi'}^{*(2)}\phi^{*(0)}+\phi^{*(0)}\phi^{*(2)}\right)\Big]
3061: +\nn\\
3062: +\int_L\,tr_b&&\Big[ {1\ov 2}\hp{1}d\hp{1}+{1\ov 3}\hp{1}\hp{1}\hp{1}
3063: +{\phi'}^{*(2)}\left(d{c'}_1^{(0)}+[\hp{1},{c'}_1^{(0)}]\right)\nn\\
3064: &&+\phi^{*(0)}\left(dc_1^{(2)}-c_1^{(2)}\hp{1}+\p{1}c_1^{(2)}-c_1^{(0)}
3065: \p{3}+\p{3}c_1^{'(0)}\right)\nn\\
3066: &&
3067: +c_1^{*(1)}\left(c_1^{(0)}c_1^{(2)}+c_1^{(2)}{c'}_1^{(0)}+dc_2^{(1)}+\p{1}c_2^{(1)}
3068: +c_2^{(1)}\hp{1}\right)
3069: \nonumber\\
3070: &&
3071: +c_2^{*(2)}\left(dc_3^{(0)}-c_3^{(0)}\hp{1}+\p{1}c_3^{(0)}-c_1^{(0)}c_2^{(1)}
3072: +c_2^{(1)}{c'}_1^{(0)}\right)
3073: \nonumber\\
3074: &&+{c'}^{*(3)}_1{c'}_1^{(0)}{c'}_1^{(0)}-
3075: c_3^{*(3)}\,\left(c_1^{(0)}c_3^{(0)}+c_3^{(0)}{c'}_1^{(0)}\right)
3076: \Big]~~. \eea A brief discussion on the master equation can be found
3077: in Appendix C.
3078:
3079: This action can be cast into the form (\ref{extended_action}) provided
3080: that we identify the extended string field with: \bea
3081: \label{coords2}
3082: {\hat \phi}&=&\pmatrix{ c_1^{(0)}+\phi^{(1)}-\phi^{*(2)} +c_1^{*(3)}&
3083: c_3^{(0)}+c_2^{(1)}+c_1^{(2)}+ \phi^{(3)}\cr -\phi^{*(0)}+c_1^{*(1)}
3084: -c_2^{*(2)}+c_3^{*(3)}&
3085: {c'}_1^{(0)} +\hp{1}-{\phi'}^{*(2)}+ {c'}^{*(3)}_1\cr}~~\nn\\
3086: \nn\\
3087: &=&c_3+c_2+c_1+\phi+\phi^*+c_1^*+c_2^* +c_3^*~~,
3088: \label{eq:superfieldn2}
3089: \eea where we defined: \be \phi^*=\pmatrix{-\phi^{*(2)}& 0\cr
3090: -\phi^{*(0)}&-{\phi'}^{*(2)}\cr}~~,~~ c_2^*=\pmatrix{ 0&0\cr
3091: -c_2^{*(2)} &0\cr}~~ \ee and \be
3092: c_1=C_1~~,~~c_2=C_2~~,~~c_3=C_3~~,~~c_1^*=C_1^*~~,~~c_3^*=C_3^*~~.
3093: \label{ghn2}
3094: \ee We have $deg{\hat \phi}={\hat 1}$, as required. Note that for
3095: $n=2$, the graded trace $str_e$ in (\ref{extended_action}) coincides
3096: with the ordinary trace. One checks again by direct computation that
3097: (\ref{coords2}) gives Darboux coordinates for the odd symplectic form
3098: and that the extended action (\ref{extended_action}) reduces to
3099: (\ref{bv2}) in this basis.
3100:
3101: \subsection{The BV action for $n=0$ }
3102:
3103: We next consider D-brane pairs with relative grading $n=0$. In this
3104: case, the gauge algebra is irreducible and one needs only first
3105: generation ghosts. The physical fields and their ghosts are given by:
3106: \be \phi=\pmatrix{ \phi^{(1)} & {\check\phi}^{(1)}\cr
3107: {\tilde\phi}^{(1)} & {\phi'}^{(1)} \cr} ~~~~~~~~~,~~~~~~~~~
3108: c_1=\pmatrix{ c_1^{(0)} & {\check c}_1^{(0)}\cr {\tilde c}_1^{(0)}&
3109: {c'}_1^{(0)} \cr}
3110: \label{fields0}~~.
3111: \ee Because the relative shift vanishes, the product $\bullet$ is the
3112: usual wedge product of forms and the classical action is the CS action
3113: for $E_a\oplus E_b$. Axelrod and Singer \cite{AS1} quantized the same
3114: classical theory, in the Faddeev-Popov approach. They expressed the
3115: gauged fixed action as an `extended CS action', where the top form in
3116: the extended string field $\hat\phi$ is constrained to vanish. Here we
3117: show that the BV construction yields the same extended action, but for
3118: us the top form is unconstrained \footnote{This BV action is of course
3119: well-known (see, for example, \cite{Ikemori}).}.
3120:
3121: The BV action for this system is: \be
3122: \label{bv0}
3123: S_{BV}=\int_L~tr_{E_a\oplus E_b}\left[{1\ov 2}\phi d\phi+{1\ov
3124: 3}\phi\phi\phi +\phi{}^*\left(dc_1+\phi
3125: c_1-c_1\phi\right)+c_1^*c_1c_1 \right]~~, \ee where \be
3126: \Phi^*=\pmatrix{ \phi^{*(2)} & \phi^{*(2)}\cr \phi^{*(2)}&
3127: {\phi'}^{*(2)} \cr} ~~{\rm~and~}~~ c_1^*=\pmatrix{ c_1^{*(3)} &
3128: c_1^{*(3)}\cr c_1^{*(3)} & c^{'*(3)}_1 \cr}
3129: \label{eq:antifn=0}
3130: \ee are the antifields associated with (\ref{fields0}).
3131:
3132: To present $S_{BV}$ in the form (\ref{extended_action}), one defines
3133: the extended string field: \be
3134: \label{coords0}
3135: \hat \phi=c_1+ \phi + \phi^* + c_1^*~~, \ee where $\phi^*:=-\Phi^*$.
3136: Upon substituting this into (\ref{extended_action}), we recover the BV
3137: action (\ref{bv0}). It is also easy to check that (\ref{coords0})
3138: gives Darboux coordinates for the odd symplectic form.
3139:
3140:
3141: \subsection{The BV action for $n\geq 3$}
3142:
3143: For relative grading $n\ge 3$, the physical field contains two
3144: one-forms: \be {\phi}=\pmatrix{ \phi^{(1)} & 0 \cr 0 & {\phi'}^{(1)}
3145: \cr} \ee and the action appears as the direct sum or difference
3146: (according to the supertrace prescription) of two Chern-Simons terms:
3147: \bea S=\int_{L}{tr_a\left[
3148: \frac{1}{2}{\phi}{}^{(1)}d{\phi}{}^{(1)}+\frac{1}{3}
3149: ({\phi}{}^{(1)})^3\right]}+(-1)^n \int_{L}{tr_b\left[
3150: \frac{1}{2}{\phi'}{}^{(1)}d{\phi'}{}^{(1)}+\frac{1}{3}
3151: ({\phi'}{}^{(1)})^3\right]}~~.\label{truncation} \eea
3152:
3153: Based on this form, it would naively seem that we can obtain the BV
3154: action simply as a sum or difference of the classical master actions
3155: associated with the two Chern-Simons theories,
3156: $S_{BVnaive}=S_{BV}^a+S_{BV}^b~$, where, according to the discussion
3157: of the previous subsection, the BV actions for the two sectors have
3158: the extended Chern-Simons form. However, this conclusion does not
3159: take into account the physically correct form of the gauge algebra.
3160: Indeed, a direct sum construction of the BV action is only justified
3161: if {\em both} the action and the gauge algebra are of direct sum form.
3162:
3163:
3164: A systematic approach to the analysis of pairs with $n\geq 3$ is
3165: afforded by the device of `gauging zero' \cite{warren}, which in our
3166: case amounts to formally applying the BRST procedure to forms of ranks
3167: higher than $3$. For this, we shall pretend that an $(n+1)$-form is
3168: not identically vanishing in three dimensions\footnote{For the
3169: rigorous reader, we note that this procedure can be justified in the
3170: jet bundle approach \cite{Stasheff_bv, Barnich}. Consider the
3171: bundle ${\cal W}=End({\bf E})= End(E_a\oplus E_b)$, whose typical
3172: fiber we denote by $W$. One defines local coordinates $x^i$ on the
3173: base manifold and $u^a$ on the fiber. The infinite jet bundle
3174: $J^\infty {\cal W}$ is a prolongation of ${\cal W}$ to an
3175: infinite-dimensional vector bundle with coordinates $u_I^a=(u^a,
3176: u_{i_1}^a, u_{i_1 i_2}^a, \dots)$ where $i_k\in \{1, 2, 3\}$ and the
3177: indices are symmetrized; the coordinates $u_I=u_{i_1..i_k}$, for a
3178: symmetric multi-index $I=(i_1\dots i_k)$, are viewed as formal
3179: partial derivatives of $u_a$ and regarded as functionally
3180: independent. A section $\sigma$ of the jet bundle decomposes as
3181: $\sigma =\sum_{k=0}^\infty{\sigma_k}$, according to the number $k$
3182: of formal derivatives; the component $\sigma_0$ defines a section of
3183: the unextended bundle ${\cal W}$. A section $s$ of ${\cal W}$ (with
3184: components $u^a\circ s =s^a$) induces a section of $J^\infty {\cal
3185: W}$, the {\em associated jet} $j^\infty s$, which has the property
3186: $u_I^a\circ j^\infty s = \partial_{i_1}\partial_{i_2}\dots
3187: \partial_{i_r}s^a$. This decomposes in a sum $j^\infty s=(j^\infty
3188: s)_0+(j^\infty s)_1+(j^\infty s)_2+\dots~$. The zeroth component of
3189: the jet is the original section, $(j^\infty s)_0=s$. One can also
3190: define forms on the infinite jet bundle: an element $\omega^{(r,s)}$
3191: of the {\em variational bicomplex} $\Omega^{*,*}(J^\infty {\cal W})$
3192: can be locally expressed in the basis $\theta_{I_1}^{a_1} \wedge
3193: \dots\theta_{I_r}^{a_r}\wedge dx^{i_1}\wedge\dots \wedge dx^{i_s}$
3194: where the {\em contact forms} $\theta_I^a=du_I^a-\sum_{i=1}^3
3195: u_{Ii}^a dx^i$ span the vertical directions, and have the property
3196: $(j^\infty s)^*(\theta^a_I)=d(\partial_I\varphi^a)-
3197: (\partial_{Ii}s^a)dx^i =0$. The space ${\cal H}$ of classical
3198: fields is the space of sections of ${\cal W}\otimes \Lambda^*T^*L$,
3199: which prolongs to the space $\Omega^{0,*}(J^\infty{\cal W})$ of {\em
3200: horizontal forms}. Hence given a classical field
3201: $\phi=\sum_{s=1}^3{\phi_{i_1\dots i_s} dx^{i_1}\wedge \dots \wedge
3202: dx^{i_s}}$, we can define its jet $j^\infty(\phi)= \sum_{s=1}^3{
3203: j^\infty(\phi_{i_1\dots i_s})dx^{i_1}\wedge \dots \wedge
3204: dx^{i_s}}\in \Omega^{0,*}(J^\infty{\cal W})$, whose zeroth
3205: component $j(\phi)_0$ coincides with $\phi$. The trick of
3206: introducing forms of rank higher than three can be understood as
3207: working with elements of $\Omega^{*,*}(J^\infty {\cal W})$. For our
3208: purpose, a `formal $k$-form' $\omega$ is an element of the
3209: variational bicomplex having total degree $p+q=k$. This can be
3210: decomposed as $\omega=\omega^{(k,0)}\oplus \omega^{(k-1,1)} \oplus
3211: \omega^{(k-2,2)} \oplus \omega^{(k-3,3)}$ and has a horizontal
3212: component $\omega^{(0,k)}$ if and only if the formal rank $k$ lies
3213: between $0$ and $3$. A similar construction can be given for the
3214: various ghost generations, upon considering Grassmann-valued forms.
3215: At the end, we shall `evaluate' all such forms, i.e. take the zeroth
3216: jet component of their horizontal projection, which is a
3217: (Grassmann-valued) section of the physical bundle ${\cal W}\otimes
3218: \Lambda^*T^*L$. This defines a BV field, i.e. an element of the
3219: extended boundary space ${\cal H}_e$. It is clear that `evaluation'
3220: in this sense applied on $\omega$ will produce zero unless the
3221: formal rank $k$ belongs to $0..3$, in which case the result of
3222: evaluation will be the zeroth jet component of $\omega^{(0,k)}$,
3223: i.e. a `true' $k$-form field. This recovers the procedure of
3224: \cite{warren}.}, and declare that the physical fields are \be
3225: {\phi}=\pmatrix{ \phi^{(1)} & \p{n+1} \cr 0 & {\phi'}^{(1)} \cr}~~.
3226: \ee
3227:
3228: We next perform the BV construction taking this as a starting point.
3229: Forms of ranks higher than 3 will be set to zero only in the final
3230: result.
3231:
3232: Correspondingly, we have the BRST variations: \bea
3233: \delta^{(1)}\p1=\left(dc^{(0)}_1+[\p{1},\,c_1^{(0)}]\right)\lambda~~,~~
3234: \delta^{(1)}\hp1=\left(d c'{}^{(0)}_1+[\hp{1},\, c'{}_1^{(0)}]\right)
3235: \lambda~~~~~~~\nn\\
3236: \delta^{(1)}\p{n+1}=\left(d c^{(n)}_1 +(-1)^{1+n}
3237: c^{(n)}_1\hp{1}+\p{1} c^{(n)}_1+\p{n+1} c'{}^{(0)}_1-
3238: c^{(0)}_1\p{n+1}\right)\lambda~~.~~~~~~~ \eea
3239:
3240: These extend to nilpotent transformations provided that we use the
3241: following variations of the ghosts: \bea \delta^{(1)}
3242: c^{(0)}_1&=&c^{(0)}_1c^{(0)}_1\lambda ~~;~~~ \delta^{(1)}
3243: c'{}^{(0)}_1=c'{}^{(0)}_1c'{}^{(0)}_1\lambda ~~;~~~ \delta^{(1)}
3244: c'{}^{(n)}_1=(c{}^{(0)}_1c{}^{(n)}_1+c{}^{(n)}_1c'{}^{(0)}_1)\lambda
3245: \label{eq:lev1}~~.~~~~~~~
3246: \eea One can write the following more synthetic form of the first
3247: level BRST transformations: \be \delta^{(1)}
3248: \phi=\left(dC_1+[\phi,\,C_1]_\bullet\right)\lambda~~,~{\rm with}~~~~
3249: C_1=\pmatrix{ c^{(0)}_1& c^{(n)}_1\cr 0& c'{}_1^{(0)}\cr}~~.
3250: \label{eq:lev1mat}
3251: \ee
3252:
3253: On the classical shell the transformation $\delta^{(1)}\p{n+1}$ has a
3254: residual local invariance given by \be \delta_2 c{}^{(n)}_1=\left(d
3255: c{}^{(n-1)}_2 +\p{1} c^{(n-1)}_2+ (-1)^n
3256: c^{(n-1)}_2\hp{1}\right)\lambda~~, \ee where $c^{(n-1)}_2$ is the
3257: second generation ghost. With the notations introduced in Sections
3258: 7.1 and 7.2, the full BRST transformations
3259: $\delta^{(2)}\equiv\delta^{(1)}+\delta_2$ at the second level coincide
3260: with $\delta^{(1)}$ when acting on $\phi^{(1)}$, ${\phi'}{}^{(1)}$,
3261: $c_1^{(0)}$, ${c'}{}_1^{(0)}$, and act on $c_1^{(n)}$ as: \bea
3262: \label{unu}
3263: \delta^{(2)}c^{(n)}_1 =(d c{}^{(n-1)}_2 +\p{1} c^{(n-1)}_2+
3264: (-1)^nc^{(n-1)}_2\hp{1} +
3265: c{}^{(0)}_1c{}^{(n)}_1+c{}^{(n)}_1c'{}^{(0)}_1)\lambda~~. \eea The
3266: transformations of second generation ghosts are determined by
3267: requiring nilpotence of (\ref{unu}). On the other hand, on-shell zero
3268: modes of (\ref{unu}) lead to ghosts of the next generation.
3269:
3270: Since we want to discuss all cases with $n\geq 3$, we give an
3271: inductive proof that the BRST transformations of the $k^{th}$
3272: generation ghosts have the form: \bea
3273: \label{k1}
3274: \delta^{(k)}c{}^{(n-k+1)}_{k}&=&\left(d c{}^{(n-k)}_{k+1} +\p{1}
3275: c^{(n-k)}_{k+1}+ (-1)^{n-k+2}
3276: c^{(n-k)}_{k+1}\hp{1}\right.\nonumber\\
3277: &{}&~~~~+(-1)^{k+1}\left.c{}^{(0)}_1 c{}^{(n-k+1)}_k+c{}^{(n-k+1)}_k
3278: c'{}^{(0)}_1\right)\lambda~~. \eea Assume that (\ref{k1}) holds for
3279: the $k-1$ ghosts: \bea \delta^{(k-1)}c{}^{(n-k+2)}_{k-1}&=&\left(d
3280: c{}^{(n-k+1)}_{k} +\p{1} c^{(n-k+1)}_{k}+ (-1)^{n-k+1}
3281: c^{(n-k+1)}_{k}\hp{1}\right.\nonumber\\
3282: &+&(-1)^{k}\left.c{}^{(0)}_1 c{}^{(n-k+2)}_{k-1}+c{}^{(n-k+2)}_{k-1}
3283: c'{}^{(0)}_1\label{k2}\right)\lambda\label{doi1}~~. \eea Then
3284: nilpotence of the BRST transformation of the $(k-1)^{th}$ generation
3285: ghost implies: \be \delta_{k-1} c_k{}^{(n-k+1)}=\left((-1)^{k+1}
3286: c_1^{(0)}c_k{}^{(n-k+1)} +c_k{}^{(n-k+1)} c'{}_1^{(0)}
3287: \right)\lambda
3288: \label{doi2}~~.
3289: \ee We finally note that the BRST transformations (\ref{doi2}) have
3290: the following on-shell zero modes: \be \delta_k
3291: c_{k}^{(n-k+1)}=\left(d c{}^{(n-k)}_{k+1} +\p{1} c^{(n-k)}_{k+1}+
3292: (-1)^{n-k+2} c^{(n-k)}_{k+1}\hp{1} \right)\lambda\label{trei}~~.
3293: \ee We recover the full BRST transformation (\ref{k1}) as the sum of
3294: (\ref{doi2}) and (\ref{trei}),
3295: $\delta^{(k)}c_{k}^{(n-k+1)}=\sum_{j=1}^k \delta_j c_{k}^{(n-k+1)}
3296: =(\delta_{k-1} +\delta_k) c_{k}^{(n-k+1)}$, where we used the fact
3297: that all transformations with the exception those of levels $k-1$ and
3298: $k$ act trivially on $c_{k}^{(n-k+1)}$.
3299:
3300: Note that the Grassmann parity of the ghosts depends on the generation
3301: number \be
3302: \label{gparity}
3303: g(c_k^{(n-k+1)})=(-)^k~~. \ee
3304:
3305: We now set to zero all forms of ranks higher than $3$, which
3306: eliminates all ghosts except for those of ranks between zero and
3307: three. The ansatz: \bea
3308: \label{coords_general}
3309: \!\!\!\!\!\!\!\!\! \!\!\!\!\!\!\!\!\! &&{\hat\phi}=
3310: \nonu\\
3311: \!\!\!\!\!\!\!\!\!\!\! \!\!\!\!\!\!\!\!\! &&\!\!\!\!\! \pmatrix{
3312: c_1^{(0)}+\phi^{(1)} - \phi^{*(2)}+c_1^{*(3)}&
3313: c_{n+1}^{(0)}+c_{n}^{(1)}+c_{n-1}^{(2)}+c_{n-2}^{(3)} \cr
3314: (-)^{n+1}c_{n-2}^{*(0)}+c_{n-1}^{*(1)}-(-)^{n}c_{n}^{*(2)}+
3315: c_{n+1}^{*(3)} & c_1^{'(0)}+\phi^{'(1)}-(-)^{n}\phi^{'*(2)} +(-)^n
3316: c_1^{'*(3)}
3317: \cr}\nonumber\\
3318: \eea shows that the resulting BV action coincides with the extended
3319: action (\ref{extended_action}). It is also easy to check that
3320: (\ref{coords_general}) define Darboux coordinates for the odd
3321: symplectic form (\ref{omega}). Moreover, because of equation
3322: (\ref{gparity}), the degree of the extended string field is~$1$.
3323:
3324:
3325:
3326: \subsection{On the equivalence of the extended actions with $n\ge 3$}
3327:
3328: Given the fact that the classical actions and physical fields of the
3329: theories with relative grading $n\ge 3$ are identical, we could
3330: attempt to relate the extended actions by a canonical transformation.
3331: As shown in the previous subsection, in these cases the extended
3332: action can be obtained by adding a vanishing $(n+1)$-form in the
3333: sector $Hom(E_a,\,E_b)$. Without this trick, one would conclude that
3334: the BV action is the graded sum of the two master actions
3335: corresponding to $E_a$ and $E_b$. Therefore, it seems natural to
3336: inquire whether the extended action can be expressed in this form. We
3337: will first isolate from the extended action the graded sum part and
3338: then search for a canonical transformation which annihilates the
3339: remnant.
3340:
3341:
3342: Let the extended field ${\hat\phi}$ for some grade difference $n\ge 3$
3343: be \be {\hat \phi}={\hat\phi_0}+{\hat c}+{\hat c}{}^*~~, \ee where:
3344: \bea
3345: {\hat\phi}_0&=&\pmatrix{c_1^{(0)}+\phi^{(1)}+\phi^{*(2)}+c_1^{*(3)}&0\cr
3346: 0& c'{}_1^{(0)}+\phi'{}^{(1)}+\phi'{}^{*(2)}+c'{}_1^{*(3)} \cr}
3347: \equiv
3348: \pmatrix{{\phi}&0\cr 0&{\phi}'\cr} \nonumber \\
3349: \nn\\
3350: \nn\\
3351: {\hat
3352: c}&=&\pmatrix{0&c_{k+3}^{(0)}+c_{k+2}^{(1)}+c_{k+1}^{(2)}+c_{k}^{(3)}
3353: \cr 0&0\cr}\equiv \pmatrix{0&c\cr 0&0\cr} \nn\\
3354: \nn\\
3355: {\hat c}{}^*&=&\pmatrix{0&0\cr
3356: c_{k+3}^{*(0)}+c_{k+2}^{*(1)}+c_{k+1}^{*(2)}+c_{k}^{*(3)}
3357: &0\cr}\equiv \pmatrix{0& 0 \cr c^*&0\cr}~~. \eea Then the extended
3358: action is: \be S_e =S_e ({\hat \phi}_0)+\int_L\,str_e\Big[ {\hat c}^*
3359: *\big(d{\hat c}+{\hat \phi}_0*{\hat c}+{\hat c}*{\hat
3360: \phi}_0\big)\Big]~~, \ee where $S_e({\hat \phi}_0)$ is the graded
3361: sum of the two BV actions corresponding to $E_a$ and $E_b$. We now
3362: look for a canonical transformation which annihilates the last term of
3363: the previous equation: \bea
3364: \label{Delta_S}
3365: \Delta S_e \equiv S_e({\hat\phi})-S_e({\hat\phi}_0)&=&(-1)^n
3366: \int_L\,tr_b\Big[ c{}^**\left(dc+\phi * c+c *\phi'\right)
3367: \Big]\nn\\
3368: &=&\int_L\,tr_a\Big[ c*\left(-dc{}^*+\phi' * c{}^*+c{}^* * \phi\right)
3369: \Big]~~. \eea
3370:
3371: Let $\Psi$ be the generator of this canonical transformation, a
3372: Grassmann-odd function of fields and antifields whose ghost number
3373: equals $-1$. If $ad_\Psi(F):=\{\Psi,F\}$ is the left adjoint action
3374: of $\Psi$ on Grassmann-valued functions $F$ of fields and antifields,
3375: then the associated one-parameter group of canonical transformations
3376: acts as $e^{t ad_\Psi}F$. Notice that the BV bracket of two monomials
3377: of degrees $p$ and $q$ in fields and antifields gives either zero or a
3378: homogeneous polynomial of degree $p+q-2$. Since (\ref{Delta_S})
3379: contains terms of degrees $2$ and $3$ in fields and antifields, it
3380: follows that the adjoint action of a term of order $p$ in the Taylor
3381: expansion of $\Psi$ will produce zero or a sum of polynomials of
3382: degrees $p$ and $p+1$. This implies that the only source of quadratic
3383: terms in $e^{t ad_\Psi}\Delta S_e$ is the quadratic term in the
3384: expansion of $\Psi$ acting an arbitrary number of times on the
3385: quadratic term in the expansion of $\Delta S_e$. Therefore, a crucial
3386: property of the desired generator $\Psi$ is that the action of $e^{t
3387: ad_\Psi}$ on the quadratic part of $\Delta S_e$ vanishes.
3388:
3389: If we also also require that $S_e({\hat \phi}_0)$ be invariant under
3390: the canonical transformation we are looking for, it follows that the
3391: only choice (up to a factor) for the quadratic term in $\Psi$ is \be
3392: \Psi_2=\int_L str_e[{\hat c}*{\hat c}{}^*]=
3393: \int_Ltr_a[c*c{}^*]=-(-)^n\int_Ltr_b[c{}^* * c]~~, \ee since
3394: $S_e({\hat \phi}_0)$ is independent of the pair $({\hat c}, {\hat
3395: c}^*)$. The fact that $\Delta S_e$ is linear in fields and
3396: antifields, \be tr_a[c{\partial \ov \partial c}\Delta S_e]=\Delta
3397: S_e~~~~~~~ tr_b[c{}^*{\partial \ov \partial c{}^*}\Delta S_E]=\Delta
3398: S_e \ee implies that: \be ad_{\Psi_2}\Delta S_e\equiv \{\Psi_2,\,
3399: \Delta S_e\}=-(1+(-1)^n)\Delta S_e~~. \ee It follows that a finite
3400: transformation of $\Delta S_e$ has the form: \be e^{t
3401: ~ad_{\Psi_2}}\Delta S_e = {1\ov 2} \left(1-(-1)^n\right)\Delta S_e
3402: +{1\ov 2} \left(1+(-1)^n\right)e^{-2t}\Delta S_e~~.
3403: \label{eq:cannytransf}
3404: \ee We therefore find that for odd $n$ the canonical transformation
3405: generated by $\Psi_2$ maps $\Delta S_e$ to itself. In particular,
3406: since $\Psi_2$ was unique up to a factor, this implies that $\Delta
3407: S_e$ cannot be removed. Hence $S_e$ cannot be canonically transformed
3408: into a difference of CS actions for $n=2k+1\geq 3$.
3409:
3410: For even $n$, the transformation (\ref{eq:cannytransf}) multiplies
3411: $\Delta S_e$ by the factor $e^{-2t}$ which is non-vanishing for any
3412: finite $t$. In the limit $t\rightarrow +\infty$, $\Delta S_e$ is
3413: mapped to zero. Thus, by choosing $\Psi=\Psi_2$ it seems that we can
3414: remove the full $\Delta S_e$. Unfortunately, the limit $t\rightarrow
3415: +\infty$ describes a point in the closure of the space of canonical
3416: transformations and it is not clear whether such a transformation is
3417: indeed allowed.
3418:
3419: Thus, the extended actions with relative grading $n\ge 3$ (with $n$ of
3420: a given parity) do not seem to be canonically equivalent, even though
3421: their classical counterparts are. However, this does not necessarily
3422: mean that the physical phenomena described by them are different,
3423: since the differences between the actions arise, as shown in the
3424: previous section, by gauging the symmetries associated to a vanishing
3425: field. In \cite{warren} a rather similar situation was analyzed (the
3426: classical action was taken to vanish and the BV action was constructed
3427: using the gauge symmetries of a 5-form field strength coupled to
3428: gravity, in four dimensions). The conclusion of the analysis was that
3429: the vanishing of physical observables survives quantization and the
3430: fields introduced by the BV construction contribute only to gauge
3431: dependent correlation functions. The analogy with the situation in
3432: that paper suggests that once quantization is performed, the
3433: off-diagonal sector $Hom(E_a,\,E_b)$ does not contribute to any
3434: observable, and the equivalence of classical actions with $n\geq 3$
3435: extends to an equivalence between the algebras of observables at the
3436: quantum level.
3437:
3438:
3439:
3440: \section{Relation between the constructive and geometric approach}
3441:
3442:
3443: As discussed in Sections 4 and 7, the role of $S_e$ as a tree-level BV
3444: action for graded D-brane pairs can be understood from two quite
3445: different perspectives, a `top-down' approach which makes use of the
3446: geometric formalism and a `bottom-up' approach which uses the standard
3447: homological approach. While the geometric formalism has the advantage
3448: of extreme generality (for example, it applies to an arbitrary number
3449: of graded D-branes, for which direct computation in components is
3450: impractical) and conceptual elegance, the discussion of Section 7 is
3451: more explicit and constructive (providing, in particular, a deeper
3452: understanding of the origin of various ghosts). In this section we
3453: explain the relation between the two descriptions, and give a more
3454: synthetic formulation of the recursive computations of the previous
3455: section.
3456:
3457: It is clear from the direct analysis of the previous section that the
3458: gauge algebra of our models is generally reducible. Therefore, the
3459: BRST procedure extends $S$ by adding the first order action $S_1$,
3460: which is built recursively from a tower of ghosts of various
3461: generations, accompanied by the corresponding antifields. This leads
3462: to successive extension of the string field $\phi$ to enlarged
3463: collections of fields $\boldphi^{(k)}$ and antifields
3464: $\boldphi^{*(k)}$. At each step $k$, one builds the $k^{th}$
3465: approximation $S_1^{(k)}$ (defined on ${\cal L}^{(k)}\oplus {\cal
3466: N}^{(k)}$) to the first order action: \be
3467: \label{Sk}
3468: S^{(k)}_1(\boldphi^{(k)}, \boldphi^{*(k)})= -\omega(\boldphi^{*(k)},
3469: {\bf q}^{(k)}(\boldphi^{(k)}))~~, \ee where ${\bf q}^{(k)}$ is the
3470: BRST generator at order $k$ (identified with a nonlinear operator
3471: acting on the linear space of truncated fields). The procedure stops
3472: at the step $k_m=\sigma+1$ dictated by order of reducibility $\sigma$
3473: of the gauge algebra. Then ${\bf q}^{(k_m)}={\bf q}$ and
3474: $S_1=S^{(k_m)}_1=-\omega(\boldphi^*, {\bf q}(\boldphi))$. In view of
3475: (\ref{Sexp}), the full master action $S_{BV}=S_e$ is obtained from
3476: $S+S_1$ by adding the following term quadratic in in antifields: \be
3477: S_2(\boldphi^*,\boldphi)=\omega(\boldphi, \boldphi^**\boldphi^*)~~.
3478: \ee To describe this in the geometric language of Section 4, we
3479: consider the expansion $\boldphi=\sum_{s\geq 0}{\phi_s}$, with
3480: $\phi_s\in M_s$. The BRST operator (\ref{BRST}) has the form: \be
3481: {\bf q}(\boldphi)=\oplus_{s\geq 0}{{\bf q}_s(\boldphi)}~~, \ee with
3482: ${\bf q}_s(\boldphi)\in \Pi M_s=P_s$. Expanding (\ref{BRST}) gives:
3483: \be {\bf q}_s(\boldphi)=-d\phi_{s+1}-\sum_{i+j=s+1}{\phi_i*\phi_j}~~.
3484: \ee Consider the subspaces: \be {\cal L}^{(k)}:=\oplus_{0\leq s\leq
3485: k}{M_{s}}~~,~~ {\cal R}^{(k)}:=\oplus_{s>k}{M_{s}}~~, \ee which give
3486: complementary ascending and descending sequences of approximations for
3487: ${\cal L}$: \bea
3488: \label{chains}
3489: & &{\cal L}^{(0)}~\subset {\cal L}^{(1)}~\subset {\cal
3490: L}^{(2)}~\subset ...
3491: ~\subset{\cal L}^{(k_m)}={\cal L}~~\nn\\
3492: & &0={\cal R}^{(k_m)}\subset {\cal R}^{(k_{m-1})}\subset
3493: ...\subset {\cal R}^{(1)}\subset{\cal R}^{(0)}~~,\\
3494: & &{\cal L}^{(k)}\oplus {\cal R}^{(k)}={\cal L}~~.\nn \eea We also
3495: define ${\cal N}^{(k)}= \oplus_{-k\leq s<0}{M_{s}}$, which give an
3496: ascending sequence of approximations for ${\cal N}$. Then the $k$-th
3497: approximation to ${\bf q}$ is obtained as: \be {\bf
3498: q}^{(k)}=P^{(k)}{\bf q}|_{{\cal L}^{(k)}}~~, \ee where $P^{(k)}$ is
3499: the projector of $\Pi{\cal L}$ onto $\Pi {\cal L}^{(k)}$, parallel
3500: with the subspace $\Pi{\cal R}^{(k)}$. Defining: \be
3501: \boldphi^{(k)}:=\sum_{0\leq s\leq k}{\phi_s}\in {\cal L}^{(k)}~~,~~
3502: \boldphi^{*(k)}:=\sum_{0\leq s\leq k}{\phi^*_s}\in {\cal N}^{(k)}~~,
3503: \ee we can write: \be {\bf q}^{(k)}(\boldphi^{(k)})= \oplus_{0\leq
3504: s\leq k}{{\bf q}^{(k)}_s(\boldphi^{(k)})}~~, \ee with: \be
3505: \label{qk}
3506: {\bf q}^{(k)}_s(\boldphi^{(k)})= -d\phi_{s+1}-\sum_{\tiny
3507: \begin{array}{c}i+j=s+1\\0\leq i,j \leq k
3508: \end{array}}{\phi_i*\phi_j}~~.
3509: \ee This explicit expression for ${\bf q}^{(k)}$ allows us to recover
3510: the truncation $S_1^{(k)}$ of the first order action upon applying the
3511: prescription (\ref{Sk}). For comparison with the case of D-brane
3512: pairs, we list the first three approximations $k=1,2,3$.
3513:
3514: \paragraph{First approximation}
3515:
3516: The first order extended field and antifield are: \bea
3517: {\boldphi}^{(1)}=\phi+c_1~~,~~{\boldphi}^{*(1)}=\phi^*+c_1^*~~. \eea
3518: Expanding (\ref{qk}) gives: \bea
3519: \label{brst1}
3520: -{\bf q}_0^{(1)}(\boldphi^{(1)})&=&dc_1+\phi*c_1+c_1*\phi~~\nn\\
3521: -{\bf q}_1^{(1)}(\boldphi^{(1)})&=&c_1*c_1~~. \eea \footnote{In the
3522: previous section we used the infinitesimal BRST transformation
3523: denoted by $\delta$. As usual, the infinitesimal transformations are
3524: given by the action of the generator multiplied by the parameter of
3525: the transformation. In our case, at the first step in the iterative
3526: procedure, we have: \bea \delta\boldphi^{(1)}\equiv -{\bf
3527: q}^{(1)}({\boldphi}^{(1)})\lambda~~, \nn \eea where $\lambda$ is a
3528: Grassmann-odd constant which plays the role of parameter for the
3529: BRST transformations. Projecting this relation onto $M_0({\hat 0})$
3530: and $M_1({\hat 1})$ we find the infinitesimal transformations used
3531: previously: \bea
3532: \label{brst1lambda}
3533: \delta^{(1)}\phi=(dc_1+\phi*c_1+c_1*\phi)\lambda~~,
3534: ~~\delta^{(1)}c_1=(c_1*c_1)\lambda~~.\nn \eea }. For D-brane pairs
3535: with relative grading $n=0$ the gauge algebra is irreducible ($o=0$)
3536: and $S_1=S_1^{(1)}$.
3537:
3538:
3539: \paragraph{Second approximation}
3540:
3541: For this, one introduces second generation ghosts $c_2$ and antighost
3542: $c^*_2$. The BRST transformations of the extended field
3543: ${\boldphi}^{(2)}=\phi+c_1+c_2$ have the expanded form: \bea
3544: \label{brst2}
3545: -{\bf q}^{(2)}_0(\boldphi^{(2)})&=&dc_1+\phi*c_1+c_1*\phi~~\nn\\
3546: -{\bf q}^{(2)}_1(\boldphi^{(2)})&=&dc_2+\phi*c_2+c_2*\phi+c_1*c_1~~\\
3547: -{\bf q}^{(2)}_2(\boldphi^{(2)})&=&c_2*c_1+c_1*c_2~~.\nn \eea The
3548: second approximation $S_1^{(2)}$ to $S_1$ is given by (\ref{Sk}) with
3549: the antifields ${\boldphi}^{*(2)}=\phi^*+c_1^*+c_2^*$. For D-brane
3550: pairs of relative grading $n=1$, the gauge algebra is first stage
3551: reducible ($o=1$) and $S_1=S_1^{(2)}$.
3552:
3553:
3554: \paragraph{Third approximation}
3555:
3556: We continue by adding a third generation ghost $c_3$ and antifield
3557: $c_3^*$. This gives the extended field
3558: ${\boldphi}^{(3)}=\phi+c_1+c_2+c_3$, whose components has the BRST
3559: transformations (\ref{qk}): \bea
3560: \label{brst3}
3561: -{\bf q}^{(3)}_0(\boldphi^{(3)})&=&dc_1+\phi*c_1+c_1*\phi~~\nn\\
3562: -{\bf q}^{(3)}_1(\boldphi^{(3)})&=&dc_2+\phi*c_2+c_2*\phi+c_1*c_1~~\nn\\
3563: -{\bf q}^{(3)}_2(\boldphi^{(3)})&=&dc_3+\phi*c_3+c_3*\phi+c_2*c_1+c_1*c_2\\
3564: -{\bf q}^{(3)}_3(\boldphi^{(3)})&=&c_2*c_2+c_1*c_3+c_3*c_1~~.\nn \eea
3565: The third approximation $S_1^{(3)}$ is constructed from (\ref{Sk})
3566: with the antifields ${\boldphi}^{*(3)}=\phi^*+c_1^*+c_2^*+c_3^*$. This
3567: coincides with the full first order action for the case of relative
3568: grading $n=3$. For D-brane pairs with relative grading $n=2$, the
3569: gauge algebra is second order reducible ($o=2$) and $S_1=S_1^{(3)}$.
3570:
3571:
3572: To check agreement with previous computations, let us first consider
3573: the case $n=1$. It is not hard to check that upon inserting
3574: (\ref{coords1}) into (\ref{brst2}) one recovers equation
3575: (\ref{eq:fullBRSTn1}) of the previous section. For $n=2$ one recovers
3576: the BRST transformations (\ref{eq:fullBRSTn2}) upon inserting
3577: (\ref{coords2}) into (\ref{brst3}). It is also easy to check that the
3578: difference between $\delta^{(k)}=-{\bf q}^{(k)}\lambda$ and
3579: $\delta^{(k-1)}=-{\bf q}^{(k-1)}\lambda$ of this section produces
3580: $\delta_{k}$ of the previous section. It is clear that the covariant
3581: formulation given above can be generalized to systems containing more
3582: than two graded D-branes.
3583:
3584: \section{Conclusions and directions for further research}
3585:
3586: We studied graded D-brane systems along the lines proposed in
3587: \cite{sc}, showing that the extended action written down in that paper
3588: plays the role of classical master action for the string field theory
3589: of such backgrounds. We gave a completely general proof of the master
3590: equation by making use of a certain $\Z$-graded version of the
3591: geometric BV formalism, which is based on the concept of graded
3592: supermanifolds recently introduced by T. Voronov. We argued that
3593: graded supermanifolds are the correct framework for a covariant
3594: description of BV systems, and discussed the basics of the geometric
3595: approach within this context. These results are of independent
3596: interest for foundational studies of BV quantization.
3597:
3598: We also performed a direct construction of the master action for the
3599: case of graded D-brane pairs. This allowed us to identify the various
3600: components of the extended string field as ghosts and antifields
3601: produced by the BV procedure, and explain their origin in the
3602: reducibility of the gauge algebra. Upon using the formalism of
3603: \cite{com1} and \cite{com3}, we analyzed formation of D-brane
3604: composites in systems of two graded D-branes, and in particular gave a
3605: rigorous construction of acyclic condensates for the case of unit
3606: relative grading. This gives a detailed implementation and
3607: generalization of ideas proposed in \cite{Vafa_cs}, though in a
3608: somewhat different context.
3609:
3610: For the case of graded D-brane pairs, we showed that the six classical
3611: theories corresponding to various relative grades arise through
3612: different choices of ghost grading and classical gauge for two
3613: underlying master actions, the extended Chern-Simons and extended
3614: super-Chern-Simons functionals. We also showed that these theories
3615: are inequivalent in the case when the underlying special Lagrangian
3616: three-cycle is topologically nontrivial. This sheds new light on the
3617: `mod~6~periodicity' of the D-brane grade discussed from a worldsheet
3618: perspective in \cite{Douglas_Kontsevich}.
3619:
3620: Our results can be viewed as a starting point for the quantization of
3621: such systems. As already pointed out in \cite{sc}, the string field
3622: theory of graded D-branes gives a concrete representation of points of
3623: the {\em extended} boundary moduli space, thereby holding the promise
3624: for a better understanding of its physical significance. This should
3625: be of direct relevance for the homological mirror symmetry programme.
3626: A quantum analysis of our theories around such backgrounds should lead
3627: to new physical information, as well as to certain extensions of
3628: various geometric invariants. This and related matters are the subject
3629: of ongoing research. Here we only note that a thorough study away from
3630: the large radius limit must take into account destabilization and
3631: other instanton effects \cite{Fukaya, Fukaya2, boundary}.
3632:
3633: Let us finally mention that a similar analysis can be carried out for
3634: the open B-model, in which case one deals with (graded) D6-branes
3635: wrapping the entire Calabi-Yau manifold. In that case, the relevant
3636: string field theory is a graded version of holomorphic Chern-Simons
3637: theory \cite{com3, Diaconescu}. It is clear that all of our
3638: constructions have a parallel in such situations, provided that one
3639: makes the appropriate modifications. Since the B model does not
3640: receive instanton corrections, the BV systems associated to graded
3641: B-type branes should be viewed as a description of the associated
3642: deformation theory; for example, they allow for a classification of
3643: the local string field observables, which can be used to deform such
3644: models. Some of these issues are currently under investigation.
3645:
3646:
3647:
3648: \acknowledgments{ We thank W.~Siegel, M.~Ro\v{c}ek , S.~Vandoren, and
3649: D.~E.~Diaconescu for interesting conversations and interest in our
3650: work. The authors are supported by the Research Foundation under
3651: NSF grant PHY-9722101.
3652:
3653: \appendix
3654:
3655:
3656: \section{Graded supermodules and super-bimodules}
3657:
3658: Consider an associative superalgebra $G$ ($\Z_2$-graded associative
3659: algebra with a unit). We remind the reader that a {\em right/left
3660: supermodule} $U$ over $G$ is a $\Z_2$-graded right/left G-module
3661: such that the scalar multiplication $U\times G\rightarrow U$
3662: (respectively $G\times U\rightarrow U$) is homogeneous of degree zero:
3663: \be deg(u\alpha)=degu+deg\alpha~~{\rm~respectively~}deg(\alpha
3664: u)=degu+deg\alpha \ee where $deg$ denotes the grading on $U$ or $G$. A
3665: {\em super-bimodule} $U$ is simultaneously a left and right $G$-module
3666: such that the left and right scalar multiplications are compatible:
3667: \be \alpha u=(-1)^{deg\alpha~deg u}u\alpha~~. \ee It is clear that
3668: this relation determines one scalar multiplication given the other, so
3669: a left or right supermodule can be made into a super-bimodule in a
3670: unique manner.
3671:
3672: \section{Left and right vector fields}
3673:
3674: Consider a DeWitt-Rogers supermanifold $M$ modeled over the algebra of
3675: constants $G$. The space ${\cal F}(M,G)$ of G-valued (smooth)
3676: functions is then a super-bimodule over $G$ with respect to pointwise
3677: scalar multiplication: \be (F\alpha)(p)=F(p)\alpha~~,~~(\alpha
3678: F)(p)=\alpha F(p)~~. \ee It is also a ring with respect to pointwise
3679: multiplication of functions: \be (FG)(p)=F(p)G(p)~~. \ee Combining
3680: the two structures, we obtain a superalgebra over $G$; this
3681: statement means that one has relations such as $\alpha FG=
3682: (-1)^{\epsilon_\alpha\epsilon_F}F\alpha G=
3683: (-1)^{\epsilon_\alpha(\epsilon_F+\epsilon_G)}FG\alpha$.
3684:
3685: A left (right) vector field on $M$ is a left (right) graded derivation
3686: of this algebra. For left vector fields, this means: \be
3687: X_l(F\alpha)=X_l(F)\alpha~~,~~X_l(FG)=X_l(F)G+
3688: (-1)^{\epsilon_F\epsilon_{X_l}}FX_l(G)~~, \ee while for right vector
3689: fields one requires: \be X_r(\alpha F)=\alpha
3690: X_r(F)~~,~~X_r(FG)=FX_r(G)+ (-1)^{\epsilon_G \epsilon_{X_r}}X_r(F)G~~.
3691: \ee In these relations, $\epsilon_{X_l}$ and $\epsilon_{X_r}$ are the
3692: vector field parities. One also defines scalar multiplications: \be
3693: (\alpha X_l)(F):=\alpha X_l(F)~~,~~(Y_l\alpha)(F)=Y_l(F)\alpha~~. \ee
3694: With these operations and grading (and the obvious definition of
3695: addition), the space ${\cal X}_l(M)$ of left derivations of ${\cal
3696: F}(M)$ is a left $G$-supermodule, while the space ${\cal X}_r(M)$ of
3697: right derivations is a right $G$-supermodule.
3698:
3699: It is common procedure to identify left and right vector fields on $M$
3700: in the following manner. If $X_l$ is a left vector field, one defines
3701: $\phi(X_l)$ by: \be \phi(X_l)(F):=(-1)^{\epsilon_X\epsilon_F}X_l(F)~~.
3702: \ee It is the easy to check that $\phi$ is a degree zero isomorphism
3703: between the graded Abelian groups $(X_l(M),+)$ and $(X_r(M),+)$.
3704: Moreover, one has the property: \be
3705: \label{phi_mult}
3706: \phi(\alpha~X_l)=(-1)^{\epsilon_\alpha\epsilon_X}\phi(X_l)\alpha~~.
3707: \ee This allows one to identify ${\cal X}_l(M)$ and ${\cal X}_r(M)$ to
3708: a single abstract space ${\cal X}(M)$, the space of vector fields on
3709: $M$. This space inherits left and right module structures from
3710: $X_l(M)$ and $X_r(M)$, and relation (\ref{phi_mult}) shows that the
3711: two structures are compatible. Thus ${\cal X}(M)$ is a super-bimodule
3712: over $G$. An element $X$ of ${\cal X}(M)$ can be viewed as a pair
3713: $(X_l, X_r)$ with $X_l\in {\cal X}_l(G)$ and $X_r\in {\cal X}_r(G)$,
3714: where $X_l$ and $X_r$ are $\phi$-related, i.e. $X_r=\phi(X_l)$. In
3715: this situation, one uses the notation: \be
3716: X_l:=\stackrel{\rightarrow}{X}~~,~~X_r:=\stackrel{\leftarrow}{X}~~,
3717: \ee and writes $\stackrel{\rightarrow}{X}F$ for $X_l(F)$ and
3718: $F\stackrel{\leftarrow}{X}$ for $X_r(F)$. With these notations, the
3719: $\phi$-relatedness condition becomes: \be
3720: \stackrel{\rightarrow}{X}F=(-1)^{\epsilon_X\epsilon_F}
3721: F\stackrel{\leftarrow}{X}~~. \ee This is the rigorous construction
3722: behind the conventions used in Section 4.1. We warn the reader that
3723: the vector fields $\partial^l_a$ and $\partial^r_a$ defined in that
3724: section do {\em not} form a $\phi$-related pair, in spite of the
3725: similarity with the notation used in this appendix. In the body of
3726: this paper, we use exclusively the convention that left and right
3727: vector fields are identified to abstract objects $X$, and the left and
3728: right components of $X$ are denoted by superscript arrows as explained
3729: above.
3730:
3731:
3732:
3733: \section{Direct check of the master equation for D-brane pairs}
3734:
3735: In this appendix we exemplify the steps needed for the proof that the
3736: BV action satisfies the master equation. We choose to take here an
3737: easier (but equivalent) path, namely we prove that the BRST
3738: transformations are nilpotent. More precisely, we show that: \be
3739: \delta({\lambda_2})\delta({\lambda_1}) \hat\phi=\{\{\hat\phi,
3740: S_{BV}\lambda_1\},S_{BV}\lambda_2\}=0~~, \ee where $\lambda_1,
3741: \lambda_2$ are anti-commuting constants \footnote{Each block component
3742: in the matrix $\hat\phi$ can be viewed locally as a matrix of forms.
3743: For example, $(\phi^{(1)})^{ji}$ and its antifield
3744: $(\phi^{*(2)})^{ij}$ have indices $i,j=1,\dots rk E_a $ while
3745: $(\phi^{(0)})^{ik}$ and its antifield $(\phi^{(3)})^{ki}$ have
3746: indices $i=1,\dots rk E_a$, $k=1,\dots rk E_b$. To avoid complicated
3747: notation we suppress all such indices.} . For the case with
3748: relative grading $n=1$, consider the BRST transformations of $\p1$:
3749: \bea
3750: \label{sss}
3751: \delta({\lambda_2})\delta({\lambda_1}) \p1&&=\{\left(dc_1^{(0)}
3752: -c_1^{(0)}\p1+\p1c_1^{(0)}-c_1^{(1)}\p0-c_2^{(0)}\phi^{*(1)}\right)\lambda_1,
3753: S_{BV}\lambda_2\}\nonu\\
3754: &&=\left[ -d((c_1^{(0)})^2+c_2^{(0)}\p0) +
3755: ((c_1^{(0)})^2+c_2^{(0)}\p0)\p1
3756: -\p1((c_1^{(0)})^2+c_2^{(0)}\p0)\right.
3757: \nonu\\
3758: && +(dc_1^{(0)} -c_1^{(0)}\p1+\p1c_1^{(0)}-c_1^{(1)}\p0-c_2^{(0)}\phi^{*(1)})c_1^{(0)}\nonu\\
3759: &&+ c_1^{(0)}(dc_1^{(0)} -c_1^{(0)}\p1+\p1c_1^{(0)}-c_1^{(1)}\p0-c_2^{(0)}\phi^{*(1)})\nonu\\
3760: && +(c_1^{(0)}c_1^{(1)}+c_1^{(1)}c_1^{'(0)}+dc_2^{(0)}+\p1c_2^{(0)}-c_2^{(0)}\hp1)\p0\nonu\\
3761: && +(c_1^{(0)}c_2^{(0)}-c_2^{(0)}c_1^{'(0)})\phi^{*(1)} +c_1^{(1)}(\p0c_1^{(0)}-c_1^{'(0)}\p0)\nonu\\
3762: &&
3763: \left.-c_2^{(0)}(-d\p0+\p0\p1-\hp1\p0-\phi^{*(1)}c_1^{(0)}-c_1^{'(0)}\phi^{*(1)})
3764: \right]\lambda_1\lambda_2=0~~. \eea We similarly have: \bea
3765: \delta({\lambda_2})\delta({\lambda_1})
3766: \p0&&=\{\left( \p0c_1^{(0)}+c_1^{'(0)}\p0\right)\lambda_1, S_{BV} \lambda_2\}\nonu\\
3767: &&=\left[-(\p0c_1^{(0)}-c_1^{'(0)}\p0)c_1^{(0)}+\p0(c_2^{(0)}\p0+(c_1^{(0)})^2)\right.\nonu\\
3768: &&\left.-((c_1^{'(0)})^2+\p0c_2^{(0)})\p0-
3769: c_1^{'(0)}(\p0c_1^{(0)}-c_1^{'(0)}\p0)\right]\lambda_1\lambda_2=0~~,
3770: \eea as well as: \bea \delta({\lambda_2})\delta({\lambda_1})
3771: \p2&&=\{\left(dc_1^{(1)}+\p2c_1^{'(0)}-c_1^{(0)}\p2+\p1c_1^{(1)}+c_1^{(1)}\hp1
3772: +\phi^{*(2)}c_2^{(0)}+c_2^{(0)}\phi^{'*(2)}\right)\lambda_1, S_{BV} \lambda_2\}\nonu\\
3773: &&=\left[-(d(c_1^{(0)}c_1^{(1)}+c_1^{(1)}c_1^{'(0)}+dc_2^{(0)}+\p1c_2^{(0)}-c_2^{(0)}\hp1)\right.\nonu\\
3774: &&-(\p1(c_1^{(0)}c_1^{(1)}+c_1^{(1)}c_1^{'(0)}+dc_2^{(0)}+\p1c_2^{(0)}-c_2^{(0)}\hp1)\nonu\\
3775: &&-(c_1^{(0)}c_1^{(1)}+c_1^{(1)}c_1^{'(0)}+dc_2^{(0)}+\p1c_2^{(0)}-c_2^{(0)}\hp1)\hp1\nonu\\
3776: &&+(dc_1^{(1)}+\p2c_1^{'(0)}-c_1^{(0)}+c_1^{(1)}\hp1+\p1c_1^{(1)}+\phi^{*(2)}c_2^{(0)}+c_2^{(0)}\phi^{'*(2)})c_1^{'(0)}\nonu\\
3777: &&+c_1^{(0)}(dc_1^{(1)}+\p2c_1^{'(0)}-c_1^{(0)}+c_1^{(1)}\hp1+\p1c_1^{(1)}+\phi^{*(2)}c_2^{(0)}+c_2^{(0)}\phi^{'*(2)})\nonu\\
3778: &&+((c_1^{(0)})^2+c_2^{(0)}\p0)\p2-\p2((c_1^{'(0)})^2+\p0c_2^{(0)})\nonu\\
3779: &&-c_1^{(1)}(dc_1^{'(0)}+\hp1c_1^{'(0)}-c_1^{'(0)}\hp1-\p0c_1^{(1)}-\phi^{*(1)}c_2^{(0)})\nonu\\
3780: &&(dc_1^{(0)}+\p1c_1^{(0)}-c_1^{(0)}\p1+c_1^{(1)}\p0-c_2^{(0)}\phi^{*(1)})c_1^{(1)}\nonu\\
3781: &&+\phi^{*(2)}(c_1^{(0)}c_2^{(0)}-c_2^{(0)}c_1^{'(0)})-(c_1^{(0)}c_2^{(0)}-c_2^{(0)}c_1^{'(0)})\phi^{'*(2)}\nonu\\
3782: &&+(d\p1+\p1{}^2+\p2\p0-\phi^{*(2)}c_1^{(0)}-c_1^{(0)}\phi^{*(2)}-c_1^{(1)}
3783: \phi^{*(1)}+c_2^{(0)}
3784: c_1^{*(2)})c_2^{(0)}\nonu\\
3785: &&+c_2^{(0)}(-d\hp1-\hp1{}^2-\p0\p2-c_1^{'(0)}\phi^{'*(2)}
3786: -\phi^{'*(2)}c_1^{'(0)}+\phi^{*(1)}c_1^{(1)}\nonu\\
3787: &&\left.-c_1^{*(2)}c_2^{(0)})\right]\lambda_1\lambda_2=0~~. \eea
3788:
3789: Nilpotence of the BRST transformations in the ghost sector follows
3790: from similar calculations. For instance: \bea
3791: \delta_{\lambda_2}\delta_{\lambda_1}c_2^{(0)}&=&\{(c_1^{(0)}c_2^{(0)}-c_2^{(0)}c_1^{'(0)})\lambda_1, S_{BV} \lambda_2\}\nonu\\
3792: &=&\left[c_1^{(0)}(c_1^{(0)}c_2^{(0)}-c_2^{(0)}c_1^{'(0)})-((c_1^{(0)})^2+c_2^{(0)}\p0)c_2^{(0)}\right.\nonu\\
3793: &+&\left.c_2^{(0)}((c_1^{'(0)})^2+\p0c_2^{(0)})+((c_1^{(0)}c_2^{(0)}-c_2^{(0)}c_1^{'(0)})c_1^{'(0)})\right]
3794: \lambda_1\lambda_2=0~~. \eea To prove the master equation, one must
3795: also analyze the transformation of antifields.
3796:
3797:
3798: Let us now exemplify nilpotence of the bracket for the case of
3799: relative grading $n=2$. For $\phi^{(1)}$, one has: \bea
3800: \delta_{\lambda_2}\delta_{\lambda_1}\p1&&=\{\left(dc_1^{(0)}
3801: -c_1^{(0)}\p1+\p1c_1^{(0)}-c_2^{(0)}c_1^{*(1)}-c_2^{(1)}
3802: \phi^{*(0)}\right)\lambda_1,
3803: S_{BV}\lambda_2\}\nonu\\
3804: &&=\left[-d((c_1^{(0)})^2-c_2^{(0)}\phi^{*(0)})-\p1 ((c_1^{(0)})^2-c_2^{(0)}\phi^{*(0)})\right.\nonu\\
3805: &&+((c_1^{(0)})^2-c_2^{(0)}\phi^{*(0)})\p1+c_1^{(0)}( dc_1^{(0)}
3806: -c_1^{(0)}\p1+\p1c_1^{(0)}-c_2^{(0)}c_1^{*(1)}-c_2^{(1)}\phi^{*(0)})
3807: \nonu\\
3808: &&+(dc_1^{(0)}
3809: -c_1^{(0)}\p1+\p1c_1^{(0)}-c_2^{(0)}c_1^{*(1)}-c_2^{(1)}\phi^{*(0)})c_1^{(0)}+(c_1^{(0)}c_2^{(0)}+c_2^{(0)}c_1^{'(0)})c_1^{*(1)}
3810: \nonu\\
3811: &&-c_2^{(0)}(d\phi^{*(0)}+\hp1 \phi^{*(0)}-\phi^{*(0)}\p1-c_1^{*(1)}c_1^{(0)}+c_1^{'(0)}c_1^{*(1)})\nonu\\
3812: &&-(dc_2^{(0)}-c_2^{(0)}\hp1+\p1c_2^{(0)}-c_1^{(0)}c_2^{(1)}+c_2^{(1)}c_1^{'(0)})\phi^{*(0)}\nonu\\
3813: &&+\left.c_2^{(1)}(\phi^{*(0)}c_1^{(0)}+
3814: c_1^{'(0)}\phi^{*(0)})\right]\lambda_1\lambda_2=0~~. \eea
3815:
3816:
3817:
3818:
3819: \begin{thebibliography}{100}
3820:
3821: \bibitem{Douglas_Kontsevich}{M.~R.~Douglas, {\em D-branes, Categories
3822: and N=1 Supersymmetry}, hep-th/0011017.} \bibitem{com1}{
3823: C.~I.~Lazaroiu, {\em Generalized complexes and string field
3824: theory}, hep-th/0102122.} \bibitem{com3}{C.~I.~Lazaroiu, {\em
3825: Unitarity, D-brane dynamics and D-brane categories},
3826: hep-th/0102183. } \bibitem{Aspinwall}{P.~S.~Aspinwall,
3827: A.~Lawrence, {\em Derived Categories and Zero-Brane Stability},
3828: hep-th/0104147.} \bibitem{Diaconescu}{D.~E.~Diaconescu, {\em
3829: Enhanced D-Brane Categories from String Field Theory},
3830: hep-th/0104200.} \bibitem{sc}{C.~I.~Lazaroiu, {\em Graded
3831: Lagrangians, exotic topological D-branes and enhanced
3832: triangulated categories}, hep-th/0105063}
3833: \bibitem{Oz_superconn}{ M.~Alishahiha, H.~Ita, Y.~Oz, {\em On
3834: Superconnections and the Tachyon Effective Action},
3835: hep-th/0012222.} \bibitem{Oz_triples}{Y.~Oz, T.~Pantev,
3836: D.~Waldram, {\em Brane-Antibrane Systems on Calabi-Yau Spaces},
3837: hep-th/0009112.} \bibitem{Fukaya}{K.~Fukaya, {\em Morse homotopy,
3838: $A^\infty$-category and Floer homologies}, in {\em Proceedings
3839: of the GARC Workshop on Geometry and Topology}, ed. by
3840: H.~J.~Kim, Seoul national University (1994), 1-102; {\em Floer
3841: homology, $A^\infty$-categories and topological field theory},
3842: in {\em Geometry and Physics}, Lecture notes in pure and applied
3843: mathematics, {\bf 184}, pp 9-32, Dekker, New York, 1997; {\em
3844: Floer homology and Mirror symmetry, I}, preprint available at
3845: $http://www.kusm.kyoto-u.ac.jp/~\tilde{}~fukaya/fukaya.html.$}
3846: \bibitem{Fukaya2}{K. Fukaya, Y.-G. Oh, H.~Ohta, K.~Ono, {\em
3847: Lagrangian intersection Floer theory - anomaly and obstruction},
3848: preprint available at
3849: $http://www.kusm.kyoto-u.ac.jp/~\tilde{}~fukaya/fukaya.html.$}
3850: \bibitem{Kontsevich_recent}{M.~Kontsevich, Y.~Soibelman, {\em
3851: Homological mirror symmetry and torus fibrations},
3852: math.SG/0011041.} \bibitem{Ooguri}{ H.~Ooguri, Y.~Oz, Z.~Yin,
3853: {\em D-branes on Calabi-Yau spaces and their mirrors}, Nucl.Phys.
3854: {\bf B477} (1996) 407-430.} \bibitem{Kontsevich}{M.~Kontsevich,
3855: {\em Homological algebra of mirror symmetry}, Proceedings of the
3856: International Congress of Mathematicians, (Zurich, 1994),
3857: 120--139, Birkhauser, alg-geom/9411018.}
3858: \bibitem{Vafa_cs}{C.~Vafa, {\em Brane/anti-Brane Systems and $U(N|M)$
3859: Supergroup}, hep-th/0101218.} \bibitem{Kontsevich_Schwarz}{ M.
3860: Alexandrov, M. Kontsevich, A. Schwarz, O. Zaboronsky, {\em The
3861: Geometry of the Master Equation and Topological Quantum Field
3862: Theory}, Int.J.Mod.Phys. {\bf A12} (1997) 1405-1430,
3863: hep-th/9502010.} \bibitem{Voronov}{T.~Voronov, {\em Graded
3864: manifolds and Drinfeld doubles for Lie bialgebroids},
3865: math.DG/0105237.} \bibitem{Witten_CS}{ E.~Witten,{\em
3866: Chern-Simons gauge theory as a string theory}, The Floer
3867: memorial volume, 637--678, Progr. Math., 133, Birkhauser, Basel,
3868: 1995, hep-th/9207094.} \bibitem{Seidel}{P.~Seidel, {\em Graded
3869: Lagrangian submanifolds}, Bull. Soc. Math. France {\bf 128}
3870: (2000), 103-149, math.SG/9903049.}
3871: \bibitem{Bismut_Lott}{J.~M.~Bismut and J.~Lott, {\em Flat vector
3872: bundles, direct images and higher analytic torsion}, J. Amer.
3873: Math Soc {\bf 8} (1992) 291.} \bibitem{Quillen}{D.~Quillen, {\em
3874: Superconnections and the Chern character}, Topology, {\bf 24},
3875: No.1.(1085), 89-95.} \bibitem{Gaberdiel}{ M.~Gaberdiel,
3876: B.~Zwiebach, {\em Tensor constructions of open string theories
3877: I:Foundations}, Nucl. Phys {\bf B505} (1997) 569,
3878: hep-th/9705038.} \bibitem{Zwiebach_open}{ B.~Zwiebach, {\em
3879: Oriented open-closed string theory revisited}, Annals. Phys.
3880: {\bf 267} (1988), 193, hep-th/9705241.} \bibitem{supergroupCS}{
3881: J.~H.~Horne, {\em Skein relations and Wilson loops in Chern-Simons
3882: gauge theory}, Nucl. Phys. {\bf B334} (1990) 669; Bourdeau, E.J.
3883: Mlawer, H. Riggs and H.J. Schnitzer, {\em The quasirational fusion
3884: structure of $SU(M|N)$ Chern-Simons and W-Z-W theories}, Nucl.
3885: Phys. {\bf B372} (1992) 303; L. Rozansky and H. Saleur, {\em
3886: Reidemeister torsion, the Alexander polynomial and $U(1,1)$
3887: Chern-Simons theory}, J. Geom. Phys. {\bf 13} (1994) 105.}
3888: \bibitem{Witten_antibracket}{E.~Witten, {\em A note on the antibracket
3889: formalism}, Mod. Phys. Lett. {\bf A5} (1990) 487.}
3890: \bibitem{Henneaux_geom}{M.~Henneaux, {\em Geometric Interpretation of
3891: the Quantum Master Equation in the BRST--anti-BRST Formalism},
3892: Phys. Lett. {\bf B282} (1992) 372, hep-th/9205018.}
3893: \bibitem{Khudaverdian}{O.~M.~Khudaverdian, A.~P.~Nersessian, {\em On
3894: the Geometry of the Batalin-Vilkovisky Formalism}, Mod. Phys.
3895: Lett. {\bf A8} (1993) 2377-2386, hep-th/9303136.}
3896: \bibitem{Schwarz_geom}{A.~Schwarz, {\em Geometry of Batalin-Vilkovisky
3897: quantization}, Commun.Math.Phys. {\bf 155} (1993) 249-260.}
3898: \bibitem{Schwarz_semiclassical}{A.~Schwarz, {\em Semiclassical
3899: approximation in Batalin-Vilkovisky formalism},
3900: Commun.Math.Phys. {\bf 158} (1993) 373-396.}
3901: \bibitem{Schwarz_symms}{A.~Schwarz, {\em Symmetry transformations in
3902: Batalin-Vilkovisky formalism}, Lett.Math.Phys. {\bf 31} (1994)
3903: 299-302.} \bibitem{Schwarz_superanalogues}{A.~Schwarz, {\em
3904: Superanalogs of symplectic and contact geometry and their
3905: applications to quantum field theory}, hep-th/9406120.}
3906: \bibitem{FHST}{J.~M.~L.~Fisch, M.~Henneaux, J.~Stasheff,
3907: C.~Teitelboim, {\em Existence, Uniqueness And Cohomology Of The
3908: Classical BRST Charge With Ghosts Of Ghosts},
3909: Commun.~Math.~Phys. {\bf 120} (1989) 379.}
3910: \bibitem{FH}{J.~M.~Fisch and M.~Henneaux, {\em Homological
3911: Perturbation Theory And The Algebraic Structure Of The Antifield
3912: - Antibracket Formalism For Gauge Theories}, Commun.~Math.~Phys.
3913: {\bf 128} (1990) 627.} \bibitem{Stasheff_bv}{ J.~Stasheff, {\em
3914: The (secret ?) homological algebra of the Batalin-Vilkovisky
3915: approach}, Secondary calculus and cohomological physics (Moscow,
3916: 1997), 195--210, Contemp. Math. {\bf 219} (1998), hep-th/9712157.}
3917: \bibitem{Henneaux_lectures}{M.~Henneaux, {\em Lectures On The
3918: Antifield - BRST Formalism For Gauge Theories}, Nuclear Physics
3919: B (Proc. Suppl.) {\bf 18A} (1990).} \bibitem{Gomis}{J.~Gomis,
3920: J.~Paris, S.~Samuel, {\em Antibracket, Antifields and Gauge-Theory
3921: Quantization}, Phys.Rept. {\bf 259} (1995) 1-145.}
3922: \bibitem{Berezin}{F.~A.~Berezin, {\em The Method of Second
3923: Quantization}, Pure and Appl. Phys. {\bf 24}, Academic Press,
3924: New -York, 1966.} \bibitem{DeWitt}{B.~DeWitt, {\em
3925: Supermanifolds}, Cambridge Univ. Press, Cambridge, 1984.}
3926: \bibitem{Rogers}{A.~Rogers, {\em A global theory of supermanifolds},
3927: J.~Math.~Phys {\bf 21} (1980) 1352-1365.}
3928: \bibitem{Schmitt}{T.~Schmitt: {\em Supergeometry and Quantum Field
3929: Theory, or: What is a Classical Configuration?}, Rev.Math.Phys.
3930: {\bf 9} (1997) 993-1052, hep-th/9607132.}
3931: \bibitem{Manin}{Yu.~I.~Manin {\em Three constructions of Frobenius
3932: manifolds: a comparative study}, Asian J. Math. {\bf 3} (1999),
3933: no. 1, 179--220, math.QA/9801006. }
3934: \bibitem{Kontsevich_Felder}{M.~Kontsevich, {\em Deformation
3935: quantization of Poisson Manifolds}, I, mat/9709010.}
3936: \bibitem{Witten_nlsm}{ E.~Witten, {\em Topological sigma models},
3937: Commun. Math. Phys. {\bf 118} (1988),411.}
3938: \bibitem{Witten_mirror}{E.~Witten, {\em Mirror manifolds and
3939: topological field theory}, Essays on mirror manifolds, 120--158,
3940: Internat. Press, Hong Kong, 1992, hep-th/9112056.}
3941: \bibitem{SYZ}{A.~Strominger, S.-T.~Yau, E.~Zaslow, {\em Mirror
3942: Symmetry is T-Duality} Nucl.Phys. {\bf B479} (1996) 243}
3943: \bibitem{AS1}{S.~Axelrod, I.~M.`Singer, {\em Chern-Simons perturbation
3944: theory}, Proceedings of the XXth International Conference on
3945: Differential Geometric Methods in Theoretical Physics, Vol. 1, 2
3946: (New York, 1991), 3--45, World Sci. Publishing, River Edge, NJ,
3947: 1992, hep-th/9110056.} \bibitem{Ikemori}{H.~Ikemori, {\em
3948: Extended form method of antifield-BRST formalism for topological
3949: quantum field theories}, Class. Quant. Grav. {\bf 10} (1993)
3950: 233, hep-th/9206061.} \bibitem{Barnich}{G.~Barnich, R.~Fulp,
3951: T.~Lada, J. Stasheff, The sh Lie structure of Poisson brackets in
3952: Field Theory, Commun.~Math.~Phys.{\bf 191} (1998) 585.}
3953: \bibitem{warren}{W. Siegel, {\em Quantum Equivalence of Different
3954: Field Representations}, Phys. Lett. {\bf B103} (1981) 107.}
3955: \bibitem{boundary}{C.~I.~Lazaroiu, {\em Instanton amplitudes in
3956: open-closed topological string theory}, hep-th/0011257.}
3957:
3958:
3959:
3960:
3961:
3962: \end{thebibliography}
3963:
3964: \end{document}
3965:
3966: