1:
2: \documentclass[paper, 12pt, letterpaper, epsf]{JHEP}
3: \input{epsf.tex}
4: \usepackage{graphics}
5: \usepackage{epsfig}
6: %\usepackage{amsmath}
7: \usepackage{amssymb}
8: %\renewcommand{\theequation}{\thesubsection.\arabic{equation}}
9:
10: %\NeedsTeXFormat{LaTeX2e}[1995/12/01]
11: %\documentclass[titlepage,12pt]{utarticle}
12:
13: \usepackage{graphics}
14: \usepackage{latexsym,amsmath,amssymb,amscd,cite}
15: \usepackage{pdfsync}
16: \usepackage{epsf}
17: %\usepackage{showkeys}
18: %\usepackage{hyperref}
19: %\hypersetup{colorlinks=true, linkcolor=green, anchorcolor=magenta,
20: %citecolor=blue, filecolor=blue, menucolor=blue, pagecolor=blue, urlcolor=red}
21: \usepackage[all]{xy}
22: %\usepackage{amsmath}
23: \usepackage{amsthm}
24:
25: \newtheorem{Theorem}{Theorem}[section]
26: \newtheorem{Lemma}[Theorem]{Lemma}
27: \newtheorem{Proposition}[Theorem]{Proposition}
28: \theoremstyle{definition}
29: \newtheorem{Definition}[Theorem]{Definition}
30: \theoremstyle{remark}
31: \newtheorem*{Remark}{Remark}
32:
33: \newcommand{\Hom}{{\rm Hom}}
34:
35: \newcommand{\A}{{\Bbb{A}}}
36: \newcommand{\Spec}{{\rm Spec}}
37:
38: \newcommand{\bp}{\begin{Proposition}}
39: \newcommand{\ep}{\end{Proposition}}
40: \newcommand{\bl}{\begin{Lemma}}
41: \newcommand{\el}{\end{Lemma}}
42: \newcommand{\bt}{\begin{Theorem}}
43: \newcommand{\et}{\end{Theorem}}
44: \newcommand{\bd}{\begin{Definition}}
45: \newcommand{\ed}{\end{Definition}}
46: \newcommand{\End}{\rm{End}}
47: \newcommand{\Aut}{\rm{Aut}}
48: %\newcommand{\id}{\rm{id}}
49: \newcommand{\td}{\tilde{}}
50: \newcommand{\Mod}{\rm{Mod}}
51: \newcommand{\Mom}{\rm{Mom}}
52: \newcommand{\Edual}{E[1]^{\rm{v}}}
53: \newcommand{\iEnd}{{\underline \End}}
54: \newcommand{\iHom}{{\underline \Hom}}
55: \newcommand{\Rep}{{\rm Rep}}
56: \newcommand{\g}{{\bf g}}
57: \newcommand{\Mat}{{\rm Mat}}
58: \newcommand{\ev}{{\rm ev}}
59:
60: \DeclareFontFamily{U}{rsf}{}
61: \DeclareFontShape{U}{rsf}{m}{n}{<5> <6> rsfs5 <7> <8> <9> rsfs7 <10-> rsfs10}{}
62: \DeclareMathAlphabet\Scr{U}{rsf}{m}{n}
63:
64: \def\N{{\Bbb N}}
65: \def\Z{{\Bbb Z}}
66: \def\C{{\Bbb C}}
67: \def\R{{\Bbb R}}
68: \def\rk{{\rm rk}}
69: \def\deg{{\rm deg}}
70: \def\Der{{\rm Der}}
71: \def\Sym{{\rm Sym}}
72:
73: \def\lp{\overrightarrow{\partial}}
74: \def\rp{\overleftarrow{\partial}}
75: \def\ld{\overrightarrow{\delta}}
76: \def\rd{\overleftarrow{\delta}}
77: \def\li{\overrightarrow{i}}
78: \def\ri{\overleftarrow{i}}
79: \def\lL{\overrightarrow{L}}
80: \def\rL{\overleftarrow{L}}
81: \def\ltheta{\overrightarrow{\theta}}
82: \def\rtheta{\overleftarrow{\theta}}
83: \def\tf{{\tilde f}}
84: \def\tom{{\tilde \omega}}
85: \def\th{{\tilde \theta}}
86: \def\ta{{\tilde a}}
87: \def\tb{{\tilde b}}
88: \def\lD{\overrightarrow{D}}
89: \def\rD{\overleftarrow{D}}
90: \def\l{\partial^l}
91:
92: %:wlmacros
93: \def\frc#1#2{\frac1{#2}{#1}}
94: %\def\bib#1#2{\bib{#1}{#2}}% for conventional use
95: \def\bib#1#2{\newcommand{#1}{#2}}% for use with refsgen.sh
96:
97: \newcommand{\be}{\begin{equation}}
98: \newcommand{\ee}{\end{equation}}
99: \newcommand{\bea}{\end{eqnarray}}
100: \newcommand{\eea}{\end{eqnarray}}
101: \newcommand{\nn}{\nonumber}
102:
103: \newcommand{\beql}[1]{\begin{eqnarray}\label{eq:#1}}
104:
105: \newcommand{\eeql}{\end{eqnarray}}
106:
107: \long\def\eqn#1#2{\begin{eqnarray}\label{eq:#1}#2\end{eqnarray}}
108: \long\def\begdel#1\enddel{}
109: %\long\def\new#1\endnew{{\bf #1}}
110: \def\rf#1{(\ref{eq:#1})}
111:
112:
113:
114: \newcommand{\Res}{{\rm Res}}
115: \newcommand{\id}{{\rm id}}
116: \newcommand{\Tr}{{\rm Tr}}
117: \newcommand{\tr}{{\rm tr}}
118: \newcommand{\str}{{\rm str}}
119: \newcommand{\diag}{{\rm diag}}
120:
121: \def\lto{\longrightarrow}
122:
123: \def\lab#1{^{(#1)}}
124: \def\Lab#1{^{[#1]}}
125:
126: \def\half{{1\over2}}
127:
128: \def\non{\nonumber}
129:
130: \def\del{\partial}
131: \def\cR{{\mathcal R}}
132: \def\cC{{\mathcal C}}
133:
134: \def\cB{\Scr B}
135: \def\BC{{\mathbb C}}
136:
137: \def\BZ{{\mathbb Z}}
138: \def\BM{{\mathbb M}}
139: \def\BS{{\mathbb S}}
140: \def\BE{{\mathbb E}}
141: \def\BX{{\mathbb X}}
142:
143: \def\BJ{{\mathbb J}}
144: \def\BA{{\mathbb A}}
145: \def\BB{{\mathbb B}}
146: \def\cZ{{\cal Z}}
147: \def\al{\alpha}
148: \def\bfone{{\bf 1}}
149: \def \unity{\bf 1}
150:
151:
152: %paper specific macros
153: \def\bc{boundary condition changing\ }
154:
155: %\def\bp{boundary preserving\ }
156:
157: \def\weff{{\mathcal W}_{\rm eff}}
158:
159: %\def\wlg{{W}_{\rm LG}}
160: \def\wlg{W}
161:
162: \def\hll{\half(\ell_1+\ell_2)}
163: %\def\vp{\varphi}
164: \def\vp{x}
165: \def\del{\partial}
166: \def\JJ{\BJ}
167: \def\EE{\BE}
168:
169:
170:
171: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% conventions
172:
173:
174: \title{On the non-commutative geometry of topological D-branes}
175: \author{~~~Calin Iuliu Lazaroiu\\Department of Mathematics\\Trinity College Dublin\\Dublin 2, Ireland\\calin@maths.tcd.ie\\}
176:
177: \abstract{This is a noncommutative-geometric study of the
178: semiclassical dynamics of finite topological D-brane systems. Starting
179: from the formulation in terms of $A_\infty$ categories, I show that
180: such systems can be described by the noncommutative symplectic
181: supergeometry of $\Z_2$-graded quivers, and give a synthetic formulation of the
182: boundary part of the generalized WDVV equations. In particular, a {\em faithful} generating
183: function for integrated correlators on the disk can be constructed as a
184: linear combination of quiver necklaces, i.e. a function
185: on the noncommutative symplectic superspace defined by the quiver's
186: path algebra. This point of view allows one to construct extended
187: moduli spaces of topological D-brane systems as non-commutative
188: algebraic `superschemes'. They arise by imposing further relations on
189: a $\Z_2$-graded version of the quiver's preprojective algebra, and
190: passing to the subalgebra preserved by a natural group of symmetries. }
191:
192: \preprint{}
193:
194: \begin{document}
195:
196:
197: \tableofcontents
198:
199: \pagebreak
200:
201: \vskip .6in
202:
203:
204: \section{Introduction}
205:
206: The extended moduli spaces \cite{Witten_mirror} of closed topological
207: strings are Frobenius supermanifolds\cite{Dubrovin}, certain types of
208: flat Riemannian supermanifolds whose metric is induced by the
209: generating functional of tree-level closed string amplitudes. This
210: description follows from the consistency constraints on closed string
211: amplitudes on the sphere, known as the WDVV equations
212: \cite{WDVV}. The theory of Frobenius supermanifolds encodes many
213: interesting properties such as the generic lack of
214: obstructions\footnote{A manifestation of this is the
215: Bogomolov-Tian-Todorov theorem \cite{Tian, Todorov} on
216: unobstructedness of deformations of complex structure for Calabi-Yau
217: manifolds, and its generalization to extended deformations due to
218: Barannikov and Kontsevich \cite{BarKon}. The extended moduli space of
219: complex structures for such manifolds coincides with the moduli space
220: of deformations of the associated topological B-type string. } of
221: (not necessarily conformal) topological bulk deformations, and has
222: well-known applications to closed string mirror symmetry.
223:
224:
225: It is natural to ask if a similar description can be given for
226: topological deformations of open strings. A boundary topological
227: string theory admits two types of deformations, which are induced by
228: bulk and boundary observables. As in the closed string case, the
229: open-closed tree-level amplitudes obey consistency conditions (the
230: so-called {\em generalized WDVV equations}), which were derived in
231: \cite{HLL} by worldsheet arguments. While bulk deformations have the
232: same character as in the boundary-free case, boundary deformations
233: behave quite differently. As shown in \cite{HLL}, they are
234: constrained by a {\em homotopy} version of the associativity
235: conditions, leading to an intricate structure known as a cyclic and
236: unital weak $A_\infty$ category (see \cite{CIL5, Stasheff_Kajiura, Kajiura, Gaberdiel, Fukaya_mirror, KS,
237: CIL4, CIL8, CIL9, CIL10, Costello1, Costello2} for related
238: work). The complication arises due to the non-commutativity of
239: boundary insertions on the disk, and leads to difficulties
240: when analyzing the boundary moduli problem. Among these is the observation that the
241: the homotopy associativity constraints for the integrated boundary disk amplitudes
242: $W_{a_1\ldots a_n}$ cannot be `integrated' faithfully to an ordinary
243: generating function of the boundary deformation parameters.
244:
245: In the present paper, we investigate a resolution to this problem, by
246: arguing quite generally that the semiclassical dynamics of
247: open strings in finite topological D-brane systems can be described in
248: the framework of supersymplectic noncommutative geometry. This
249: approach, which is already implicit in the $A_\infty$ constraints of
250: \cite{HLL}, leads us to consider the boundary deformation potential as
251: a function on a {\em noncommutative} space, and allows for a synthetic
252: formulation of the boundary WDVV equations. Moreover, it leads
253: naturally to a construction of boundary moduli spaces of topological
254: D-branes as noncommutative superspaces.
255:
256: Let us explain this for the simple example of a single topological
257: D-brane. In this situation, one could try the following naive proposal
258: for the generating function: \be \nn
259: W_{naive}=\sum_{n}{\frac{1}{n}W_{a_1\ldots a_n}\sigma^{a_1}\ldots
260: \sigma^{a_n}}~~, \ee where one views the boundary deformation
261: parameters as (super)commuting variables $\sigma^a$. However,
262: supercommutativity of the parameters implies that $W_{naive}$ reduces
263: to: \be\nn \sum_{n}{\frac{1}{n}W_{(a_1\ldots a_n)}\sigma^{a_1}\ldots
264: \sigma^{a_n}} \ee where $W_{(a_1\ldots a_n)}$ is the (super)
265: symmetrization of the amplitudes. Thus a generating function based on
266: (super) commuting deformation parameters does not faithfully encode
267: the topological tree-level data of the worldsheet theory, and cannot
268: generally be used to reconstruct the latter.
269:
270: It was suggested in \cite{HLL} (see also \cite{Kajiura}) that this
271: problem might be overcome by viewing the boundary deformation
272: parameters as {\em non-commuting}. While this might seem unusual at
273: first sight, it is in fact quite natural if one recalls that any
274: boundary theory admits Chan-Paton extensions, whose effect is to
275: promote the deformation parameters to (super)matrices $X^a$. As a
276: result, some of the information lost by $W_{naive}$ is preserved by
277: the matrix potential: \be \nn {\hat
278: W}=\sum_{n}{\frac{1}{n}W_{a_1\ldots a_n}{\rm str}(X^{a_1}\ldots
279: X^{a_n})}~~. \ee However, supertraces of matrix monomials of finite
280: dimensions generally obey polynomial constraints. As a consequence,
281: the matrix potential of a fixed Chan-Paton extension can be reduced by
282: such relations, and again fails to faithfully encode the data of the
283: theory. To completely resolve the issue, one has to remove all
284: constraints on $X^a$, which amounts to replacing them by free
285: (and in particular non-commuting) supervariables $s^a$. Hence one is lead to the {\em
286: non-commutative generating function} of \cite{HLL}:
287: \be
288: \nn
289: W=\sum_{n} \frac{1}{n}W_{a_1\ldots a_n}(s^{a_1}\ldots s^{a_n})_c~~,
290: \ee
291: where $(.)_c$ denotes the graded-cyclization operation, which gives an
292: abstract analogue of the supertrace. Notice that
293: $W$ allows one to recover any Chan-Paton extension upon
294: replacing $s^a$ with supermatrices $X^a$, which amounts to considering
295: a finite-dimensional representation of the free associative
296: superalgebra $A=\C\langle\{s^a\} \rangle$. In this way, one can study
297: at once {\em all} Chan Paton extensions of the theory, as well as more general representations
298: obtained by taking morphisms from $A$ to an arbitrary associative superalgebra.
299:
300:
301: Remarkably, the procedure outlined above agrees with a key
302: principle of noncommutative algebraic geometry espoused in
303: \cite{Konts_formal} and developed further in
304: \cite{LB_lectures} (see \cite{Ginzburg_lectures} for an introduction).
305: According to this ideology, `good' notions in
306: affine non-commutative algebraic geometry should induce the
307: corresponding classical notions on the moduli spaces of
308: finite-dimensional representations of the noncommutative coordinate
309: ring. In the example above, the free superalgebra $A$ is the
310: coordinate ring of a noncommutative affine superspace $\A$, while
311: finite-dimensional representations of this algebra
312: (i.e. supermatrix-valued points of the `noncommutative scheme' $\A$)
313: correspond to Chan-Paton extensions of the theory. By insisting
314: that the generating function should faithfully encode the information
315: of integrated amplitudes on the disk, we are lead to consider $W$ as
316: an element of the cyclic subspace $A_c$ of $A$
317: (namely the subspace of $A$ spanned by all graded-cyclic
318: monomials in the generators $s^a$). This matches the interpretation
319: \cite{Konts_formal} of $A_c$ as the space of regular
320: functions on $\A$. Thus $W$ is a function on a
321: non-commutative affine space, and we find that physics
322: reasoning agrees with the approach to non-commutative algebraic
323: geometry advocated in \cite{Konts_formal, LB_lectures}. Moreover, one
324: can show that the boundary topological metric induces an (even or odd)
325: noncommutative symplectic form on $\A$, which makes this affine superspace into
326: a noncommutative symplectic supermanifold. As
327: in the supercommutative case, the symplectic structure
328: determines a bracket $\{.,.\}$ on $A_c$, and one finds that the homotopy associativity
329: constraints of \cite{HLL} are equivalent with the equation:
330: \be
331: \label{BV}
332: \{W,W\}=0~~.
333: \ee
334: Moreover, the unitality condition \cite{HLL, CIL5} on the underlying
335: $A_\infty$ algebra can also be formulated as
336: a constraint on $W$. This gives a
337: non-commutative geometric interpretation of the boundary part of the
338: generalized WDVV equations. As explained in \cite{HLL}, the
339: boundary topological metric of a general worldsheet theory can be even
340: or odd; as a consequence, $\{.,.\}$ is an even or odd Lie bracket. In
341: the latter case, the constraint (\ref{BV}) is a {\em non-commutative}
342: analogue of the classical master equation.
343:
344: Non-commutativity of boundary observable insertions is also
345: responsible for the fact that deformations of topological D-branes
346: are generally obstructed, which is reflected in
347: the typically singular nature of the boundary moduli space. An algebro-geometric approach to
348: boundary deformations was developed in relation to homological mirror
349: symmetry \cite{HMS} in \cite{Fukaya_mirror} and related to the deformation theory of open strings
350: in \cite{CIL4} (see \cite{Polishchuk} for related work); these proposals
351: rely on constructing the moduli space as a commutative algebraic
352: or complex-analytic variety.
353:
354:
355: The observations made above suggest that the deformation theory of
356: topological D-branes can be considered as a problem in {\em
357: noncommutative} geometry. In particular, the possibility of
358: Chan-Paton extensions implies that the boundary moduli space can be
359: viewed as a non-commutative algebraic variety. This is obtained by
360: `extremizing' the noncommutative function $W$ and modding out via
361: appropriate symmetries. More precisely, one can impose the algebraic
362: relations:
363: \be
364: \label{crit}
365: \ld_a W=0~~,
366: \ee
367: where $\ld_a$ is a $\Z_2$-graded version of the so-called {\em cyclic
368: derivatives} of \cite{RSS, Voiculescu} (see also \cite{Konts_formal}).
369: If $J$ is the two-sided ideal generated by $\ld_a W$, then the
370: quotient algebra $\C[{\cal Z}]:=\C[\{s^a\}]/J$ can be viewed as the
371: coordinate ring of a `noncommutative affine scheme' ${\cal Z}$ sitting
372: inside the affine superspace $\A$. Moreover, one can show that the symmetries of
373: the system
374: form a subgroup ${\cal G}$ of the group of noncommutative symplectomorphisms
375: of $\A$. These symmetries also preserve $J$, and thus descend to
376: automorphisms of $\C[{\cal Z}]$. They can be viewed as gauge
377: transformations acting along the `noncommutative vacuum space' ${\cal Z}$.
378: One can thus define a {\em non-commutative extended moduli space}
379: ${\cal M}$ as the affine `noncommutative scheme' whose coordinate ring
380: $\C[{\cal M}]=\C[{\cal Z}]^{\cal G}$ is the ${\cal G}$-invariant part
381: of $\C[{\cal Z}]$. The existence of a unit observable in the boundary
382: sector implies that one of the conditions (\ref{crit}) is a
383: non-commutative moment map constraint in the sense of \cite{BEV,
384: Bergh}. Therefore, the (invariant theory) quotient leading to ${\cal M}$ amounts to
385: modding out a zero-level 'symplectic reduction' of $A$ through the
386: ideal defined by the remaining relations.
387:
388: It turns out that the construction outlined above can be carried out in much greater generality.
389: In fact, as pointed out in \cite{CIL5, HLL}, the
390: homotopy-associativity constraints on disk boundary amplitudes
391: generalize to {\em systems} of D-branes. In
392: this situation, boundary observables are either boundary-preserving or
393: boundary condition-changing, a decomposition which defines the
394: boundary sectors discussed in \cite{CIL1}. The homotopy associativity constraints of \cite{HLL}
395: admit an obvious extension to this case, which can be formulated by
396: saying \cite{CIL5} that the D-brane system defines a (generally weak) cyclic and
397: unital $A_\infty$ category. The objects $u$ of this category are the
398: D-branes themselves, while the morphism spaces $\Hom(u,v)=E_{uv}$ are
399: the spaces of topological observables of strings stretching from $u$
400: to $v$. In this case, the boundary deformation parameters $s^a$ are
401: replaced by $s^i_{uv}$, where $u,v$ run over the
402: topological D-branes, while
403: $i$ indexes a basis $\psi^i_{uv}$ of $E_{uv}$ (in fact, $\{s\}$ can
404: be viewed as a parity-changed dual basis to $\{\psi\}$). Treating
405: $s^i_{uv}$ as non-commuting supervariables leads one to replace the
406: free superalgebra $\C\langle \{s^a\}\rangle$ with the associative
407: superalgebra generated\footnote{Over the subalgebra spanned by the trivial paths.}
408: by $s^i_{uv}$ with the relations:
409: \be
410: s^i_{uv}s^j_{v'w}=0~~{\rm unless~}v=v'~~.
411: \ee
412: This is the the path algebra $A_{\cal Q}$ of a quiver ${\cal Q}$
413: whose vertices are the D-branes $u$, and whose arrows from $u$ to $v$
414: are the index triples $(u,v,i)$ associated to $s^i_{uv}$. This quiver is $\Z_2$-graded since $s^i_{uv}$ can be
415: even or odd; as a consequence, $A_{\cal Q}$ is an
416: associative superalgebra. The case of a single D-brane corresponds to
417: a quiver with a single vertex $u$, whose path algebra coincides with
418: the free superalgebra generated by the $\Z_2$-graded loops at that vertex.
419:
420: It is known that path algebras of quivers are formally smooth in the
421: sense of \cite{CQ}, and in fact they provide good non-commutative
422: generalizations of the coordinate rings of smooth affine varieties
423: \cite{LB_lectures}. One can formulate a non-commutative symplectic
424: geometry for such spaces \cite{Ginzburg, LB_quivers} by extending the
425: construction of \cite{Konts_formal}. For a general topological D-brane system,
426: the space $\A_{\cal Q}$ with coordinate ring $A_{\cal Q}$
427: is a noncommutative {\em super}space, whose symplectic
428: form is induced by the parity change of the boundary topological
429: metrics. As for affine superspaces, one finds an even or odd Lie
430: superbracket $\{.,.\}$ which acts on the space $C^0_R(A_{\cal Q})$ of
431: regular functions on $\A_{\cal Q}$. The latter can be viewed as the
432: vector superspace spanned by necklaces of the quiver (i.e. cycles of
433: the quiver whose marking by the start=endpoint is forgotten). The
434: generating function $W$ of integrated disk boundary amplitudes is an
435: element of this space, i.e. a linear combination of such necklaces;
436: its parity is opposite that of the boundary topological metrics. The
437: categorical $A_\infty$ constraints for the entire collection of
438: boundary operators amount to equation (\ref{BV}). The existence of
439: unit observables in the boundary-preserving sectors can be expressed
440: as a differential constraint on $W$. Once again, a noncommutative
441: moduli space can be constructed as the invariant theory quotient of the
442: noncommutative critical variety ${\cal Z}$ through the symmetries of
443: the system. The existence of boundary unit observables implies that
444: this quotient can be be viewed as a noncommutative 'symplectic
445: reduction', followed by imposing further constraints.
446:
447:
448: The paper is organized as follows. In Section 2, we describe the
449: algebraic structure of finite D-brane systems, starting from the
450: formulation in terms of $A_\infty$ categories found in \cite{HLL,
451: CIL5} (which is summarized in Appendix \ref{data}). Upon introducing a
452: finite-dimensional commutative semisimple algebra $R$, we show that
453: one can express the boundary sector decomposition of \cite{CIL1} as an
454: $R$-superbimodule structure on the total space
455: $E=\oplus_{u,v}{\Hom(u,v)}$ of boundary observables. As a consequence,
456: the cyclic and unital weak $A_\infty$ category determined by
457: integrated string correlators can be described as a cyclic and unital
458: weak $A_\infty$ algebra on the superbimodule $E$. We also summarize
459: the strategy which will be followed in later sections in order to
460: encode this data into a noncommutative generating function. The main
461: step is passing from the superbimodule $E$ to the tensor algebra
462: $A=T_R (E[1]^{\rm v})$ of its parity-changed dual, which is a
463: superalgebra over $R$. The boundary topological metrics of the theory
464: induce an (even or odd) noncommutative symplectic form on $A$. This
465: leads us to study the noncommutative symplectic geometry of
466: $R$-superalgebras, a subject which is addressed in Section 3. In
467: Section 4, we apply this to the tensor algebra $A$, discussing the
468: realization of the abstract objects introduced previously. Upon
469: picking appropriate bases in the space of boundary observables, we show that $A$
470: can be identified with the path algebra of a $\Z_2$-graded quiver, and
471: give coefficient expressions for the various quantities of
472: interest. We also discuss the quiver version of cyclic derivatives and
473: the so-called loop partial derivatives, two types of operators acting
474: on the space of noncommutative functions. In Section 5, we apply the
475: machinery developed in Sections 3 and 4 to finite D-brane
476: systems. Using the results of Section 2, we show that the cyclic and
477: unital weak $A_\infty$ structure on $E$ is encoded faithfully by a
478: noncommutative generating function $W$, which can be expressed as a
479: linear combination of quiver necklaces. After discussing the
480: constraints obeyed by $W$, we say a few words about deformations of
481: the underlying string theory (as opposed to deformations of the
482: boundary data). In Section 6, we discuss symmetries of the topological
483: D-brane system and give the algebraic construction of the
484: noncommutative moduli space. Section 7 gives the realization of our
485: construction in the case of a single D-brane. In Section 8,
486: we discuss the simplest examples with an even and odd boundary topological
487: metric, and a rather general class of examples with odd boundary metrics, some
488: particular cases of which appeared recently in \cite{Katz}. Section 9
489: contains our conclusions.
490:
491:
492: \paragraph{Conventions}
493: Unless specified otherwise, an algebra means an associative and unital algebra over the complex numbers.
494: All morphisms of algebras are assumed to be unital. We will often encounter $\Z\times \Z_2$-graded algebras. To fix the
495: sign convention for such algebras, one must chose a $\Z_2$-valued pairing on the
496: Abelian group $\Gamma:=\Z\times \Z_2$, i.e. a biadditive and
497: symmetric map $\cdot:\Gamma\times \Gamma\rightarrow \Z_2$ specifying
498: how the Koszul rule is applied to bigraded objects. Namely, one agrees
499: that commuting two objects of bidegrees $(n,\alpha)$ and $(m,\beta)$
500: always produces the sign $(n,\alpha)\cdot (m,\beta)$. In the present
501: paper, we work with the following choice of pairing:
502: \be
503: \label{pairing}
504: (n,\alpha)\cdot (m,\beta)=[mn]+\alpha\beta~~,
505: \ee
506: where here and in the rest of the paper the notation $[k]$ for $k\in
507: \Z$ stands for the ${\rm mod}~2$ reduction of $k$. If $\alpha$ is an
508: element of $\Z_2$, we let $(-1)^{\alpha}=+1$ if $\alpha=0$ and $-1$ if
509: $\alpha=1$. We let $[1]$ be the parity change functor on the category
510: of super-vector spaces and $\Sigma_U$ the suspension operator of
511: the supervector space $U$ (i.e. $\Sigma:U\rightarrow U[1]$ is the
512: identity operator of $U$, viewed as an odd map from $U$ to itself).
513: We have $\Sigma_U^2=\id_U$ and $[1]^2\approx {\rm Id}$, where ${\rm Id}$ in the
514: last relation is the identity functor. We will often write $\Sigma$
515: without indicating the vector space on which it acts, since the latter
516: is usually clear from the context. Given a ring $R$, we let ${\rm
517: Mod-R}$, ${\rm R-Mod}$ and ${\rm R-Mod-R}$ and ${\rm
518: Smod-R}$, ${\rm R-Smod}$ and ${\rm R-Smod-R}$
519: be the categories of left-, right- and bi- modules, respectively supermodules over $R$.
520: For a pair $U,V$ of
521: objects of any of these categories, we let $\Hom(U,V)$ be the space of
522: morphisms from $U$ to $V$ in that category. For supermodules, this consists of
523: degree zero $R$-linear maps and is an ordinary (i.e. ungraded) module. For two supermodules,
524: we also let $\iHom(U,V)$ be the space of morphisms in the corresponding
525: category ${\rm Mod-R}$, ${\rm R-Mod}$ or ${\rm R-Mod-R}$ of {\em
526: ungraded} objects, obtained by forgetting the $\Z_2$-grading of
527: $U,V$. The latter consists of all linear maps, without degree
528: conditions; it is known as the {\em inner} morphism space and is
529: $\Z_2$-graded. In fact, we have $\iHom(U,V)=\iHom^0(U,V)\oplus
530: \iHom^1(U,V)$, where the degree zero component coincides with the
531: space of degree zero maps, $\iHom^0(U,V)=\Hom(U,V)$. We will use the same notational
532: convention for the endomorphism and automorphism spaces, for example $\End(U)$
533: is the endomorphism space of $U$ as a supermodule and $\iEnd(U)$ its inner endomorphism
534: space etc. Given an $R$-superbimodule $U$,
535: we set $T_R U:=\oplus_{n\geq 0}{U^{\otimes_R n}}$ (with $U^{\otimes 0}:=R$), the tensor
536: algebra of $U$, viewed as an $\N\times \Z_2$ graded algebra.
537:
538: {\bf Note} Throughout this paper, an $A_\infty$ algebra or category will mean
539: an $A_\infty$ algebra/category whose sequence of defining products $(r_n)$ terminates
540: (i.e. $r_n=0$ for sufficiently large $n$). This condition is purely technical,
541: being needed if one wishes to pass
542: to the dual of a coalgebra in naive manner. The condition can be
543: removed in standard fashion, by considering profinite modules and formal algebras and
544: replacing the relevant maps by continuous maps; then the non-commutative geometric
545: objects mentioned above can be understood in the formal sense. Because most of our
546: considerations generalize straightforwardly to formal case, we adopted the convention of
547: sometimes writing finite sums without indicating the upper bound; in this paper, it is
548: understood that all such sums terminate. When we write a finite sum this manner, it is
549: implied that the sum can be extended to a series in the formal theory.
550:
551:
552: \section{Algebraic description of finite topological D-brane systems}
553: \label{algebraic}
554:
555: In this section, we consider finite D-brane systems in an open
556: topological string theory, i.e. finite collections of D-branes,
557: together with the spaces of boundary observables of all topological
558: strings stretched between them. We will assume that the total space of
559: zero-form boundary observables is finite-dimensional, which is the usual case
560: for topological strings. Using a description derived in \cite{CIL5, HLL}, we
561: will encode the information of all tree-level boundary string amplitudes into
562: an algebraic structure on a certain superbimodule built out of the total space
563: of zero-form boundary observables.
564:
565:
566: The results of \cite{CIL5, HLL} imply that D-brane systems in topological string theory are
567: described by $A_\infty$ categories. The precise statement is
568: as follows (see Appendix \ref{data} for mathematical background):
569:
570: \
571:
572: {\em A topological D-brane system is described by a weak, cyclic and
573: unital $A_\infty$ category ${\cal A}$.}
574:
575: \
576:
577: \noindent
578:
579: In general, the $A_\infty$ category is only $\Z_2$-graded. This
580: grading can be lifted to a $\Z$-grading provided that the relevant
581: $U(1)$ symmetry of the worldsheet theory is non-anomalous. The
582: objects of ${\cal A}$ are the D-branes themselves, while the morphism
583: space $ \Hom_{\cal A}(u,v)$ for two objects $u,v$ is the complex
584: supervector space of boundary zero-form observables for the
585: topological string stretching from $u$ to $v$. The degree of such
586: observables is given by worldsheet Grassmann parity, and will be
587: denoted by $|.|$. With respect to this grading, the $A_\infty$
588: products have degrees $[n]$, where the square bracket indicates mod 2
589: reduction. As in Appendix \ref{data}, it is convenient to work with
590: the parity-changed spaces $\Hom_{\cal A}(u,v)[1]$, the degree of whose
591: elements we denote by a tilde. Hence ${\tilde x}=[|x|+1]$ for all
592: homogeneous morphisms $x$. The unitality condition involves even
593: elements $1_u$ of the endomorphism spaces $\Hom_{\cal A}(u,u)$
594: (equivalently, their odd suspensions $\lambda_u:=\Sigma 1_u\in \Hom_{\cal
595: A}(u,u)[1]$), while the cyclicity constraint involves non-degenerate
596: $\Z_2$-homogeneous pairings (the boundary topological metrics)
597: $\rho_{uv}:\Hom_{\cal A}(u,v)\times \Hom_{\cal A}(v,u)\rightarrow \C$
598: of common degree ${\tilde \omega}\in \Z_2$, obeying the
599: graded-symmetry condition $\rho_{uv}(x,y)=(-1)^{|x||y|}\rho_{vu}(y,x)$
600: for homogeneous elements $x\in \Hom_{\cal A}(u,v)$ and $y\in
601: \Hom_{\cal A}(v,u)$. The latter can also be expressed in
602: terms of the graded-antisymmetric forms $\omega_{uv}:=\rho_{uv}\circ
603: \Sigma^{\otimes 2}:\Hom_{\cal A}(u,v)[1]\times \Hom_{\cal
604: A}(v,u)[1]\rightarrow \C$ obtained from the topological metrics via
605: suspension. The detailed formulation of this structure (which arises
606: by introducing boundary sectors in the derivation of \cite{HLL}) is
607: given in Appendix \ref{data}.
608:
609:
610:
611: \paragraph{Observation}
612:
613: It was argued in \cite{CIL5} (see \cite{CIL2, CIL3} for the dG case) that the $A_\infty$ category
614: ${\cal A}_{full}$ obtained by considering `all' D-branes of a
615: topological string theory must be endowed with a parity-change functor
616: and be equivalent with its category of twisted complexes. This encodes
617: the physical requirement that the collection of all topological
618: D-branes of a given theory is closed under formation of topological
619: D-brane composites. This `quasiunitarity constraint' implies that the
620: cohomology category $H^0({\cal A}_{full})$ is an enhanced triangulated
621: category (with an $A_\infty$ enhancement). In the present paper, we
622: work with a fixed $A_\infty$ sub-category ${\cal A}$ of the full
623: D-brane category ${\cal A}_{full}$, so ${\cal A}$ need not obey this
624: supplementary condition (which is impossible to satisfy with a finite collection of objects).
625:
626: \
627:
628: For the remainder of this paper, we focus on finite D-brane systems,
629: which means that $Ob{\cal A}$ is a finite set and all morphism spaces
630: $\Hom(u,v)$ are finite-dimensional. In this case, we say that the $A_\infty$ category ${\cal A}$ is {\em finite}.
631:
632:
633: \subsection{Encoding the boundary sector decomposition}
634:
635: One can encode the categorical data of ${\cal A}$
636: in an equivalent, but more amenable form. Setting ${\cal Q}_0:=Ob{\cal A}$,
637: consider the $\Z_2$-graded vector space
638: $E:=\oplus_{u,v\in {\cal Q}_0}{\Hom(u,v)}$, which is known as
639: the total boundary space of the D-brane system \cite{CIL1}.
640:
641:
642: \
643:
644: \noindent {\bf Definition}
645: A {\em binary homogeneous decomposition} over the set ${\cal Q}_0$ is
646: a pair $(U, (U_{uv})_{u,v\in {\cal Q}_0})$ such that $U, U_{uv}$ are
647: finite-dimensional complex supervector spaces and $U=\oplus_{(u,v)\in {\cal Q}_0\times {\cal
648: Q}_0}{U_{uv}}$ is a homogeneous decomposition of $U$ indexed by the
649: Cartesian product ${\cal Q}_0\times {\cal Q}_0$. The {\em opposite} of the binary decomposition
650: $(U, (U_{uv})_{(u,v)\in {\cal Q}_0\times {\cal Q}_0})$ is the binary homogeneous decomposition
651: $(U, (U^{opp}_{uv})_{(u,v)\in {\cal Q}_0\times{\cal Q}_0})$, where $U^{opp}_{uv}=U_{vu}$.
652:
653: \
654:
655: A morphism of binary homogeneous decompositions over ${\cal Q}_0$ from
656: $(U, (U_{uv})_{(u,v)\in {\cal Q}_0\times {\cal Q}_0})$ to
657: $(U',(U'_{uv})_{(u,v)\in {\cal Q}_0\times {\cal Q}_0})$
658: is a degree zero linear map $\phi:U\rightarrow U'$ such that
659: $\phi(U_{uv})\subset U'_{uv}$ for all $(u,v)\in {\cal Q}_0\times {\cal Q}_0$.
660: With this definition, binary
661: homogeneous decompositions over ${\cal Q}_0$ form a category. A morphism
662: $\phi$ in this category is an isomorphism iff it is a bijective
663: map. In this case, we have $\phi(U_{uv})=U'_{uv}$ for all $u,v\in
664: {\cal Q}_0$.
665:
666:
667: A topological D-brane system determines two opposite decompositions of its total boundary space,
668: namely $E=\oplus_{(u,v)\in {\cal Q}_0\times {\cal Q}_0}{\Hom_{\cal A}(u,v)}$ and
669: $E=\oplus_{(u,v)\in {\cal Q}_0\times {\cal Q}_0}{\Hom_{\cal A}(v,u)}$. We will view
670: the first of these as fundamental, so we set $E_{uv}:=\Hom_{\cal A}(u,v)$; then
671: $E_{uv}^{opp}:=\Hom_{\cal A}(v,u)$. This is consistent with the convention that morphisms
672: compose forward in the definition of an $A_\infty$ category given in Appendix \ref{data}.
673: The binary homogeneous decomposition $(E, (E_{uv})_{(u,v)\in {\cal Q}_0\times {\cal Q}_0})$
674: is the so-called {\em boundary sector decomposition} of the topological D-brane system \cite{CIL1}.
675:
676: Let us consider the finite-dimensional semisimple commutative algebra $R=\oplus_{u\in {\cal Q}_0}{\C}$,
677: where the right hand side is a direct sum of copies of $\C$, viewed as
678: an algebra over itself. We let $\epsilon_u$ $(u\in {\cal Q}_0$) be
679: the commuting idempotents of $R$ corresponding to the canonical basis
680: elements of the vector space $\C^{{\cal Q}_0}$, so $R=\oplus_{u\in
681: {\cal Q}_0}{\C \epsilon_u}$ with
682: $\epsilon_u\epsilon_v=\delta_{uv}\epsilon_u$. The unit of $R$ is
683: $1_R=\sum_{u\in {\cal Q}_0}{\epsilon_u}$. It is not hard to
684: see that a binary
685: homogeneous decomposition $U=\oplus_{(u,v)\in {\cal Q}_0\times {\cal Q}_0 }{U_{uv}}$
686: amounts to giving an $R$-superbimodule structure on $U$.
687: Indeed, such a decomposition amounts to giving degree
688: zero commuting idempotents $\epsilon_u^l, \epsilon_u^r\in \End_\C(U)$
689: for all $u\in {\cal Q}_0$, namely the projectors on the subspaces
690: $\oplus_{v}{U_{uv}}$ and $\oplus_{v}{U_{vu}}$ respectively (then
691: $U_{uv}=\epsilon_u^l\epsilon_v^r(U)=\epsilon_v^r\epsilon_u^l(U)$). Notice
692: that $U_{uv}$ can be recovered as $U_{uv}=\epsilon_uU\epsilon_v$ from
693: knowledge of the $R$-superbimodule structure.
694: Moreover, a
695: morphism of binary homogeneous decompositions over ${\cal Q}_0$
696: amounts to a morphism of $R$-superbimodules. Hence the category of
697: binary homogeneous decompositions over ${\cal Q}_0$ is equivalent with
698: the category of superbimodules over $R$.
699:
700: Since $R$ is commutative, any $R$-superbimodule $U$ defines another $R$-superbimodule $U^{opp}$
701: whose underlying supervector space coincides with that of $U$ but whose external multiplications
702: are given by:
703: \be
704: \nn
705: \alpha*x*\beta=\beta x\alpha~~~~~~\forall \alpha,\beta\in R~~\forall x\in U~~.
706: \ee
707: The relations $\epsilon_u*E*\epsilon_v=\epsilon_v E \epsilon_u$ show that the binary decomposition
708: determined by $U^{opp}$ is the opposite of the binary decomposition determined by $U$:
709: \be
710: \nn
711: (U^{opp})_{uv}=U_{uv}^{opp}=U_{vu}~~.
712: \ee
713:
714:
715: Applying this to a topological D-brane system, we find that the boundary
716: sector decomposition $E=\oplus_{(u,v)\in {\cal Q}_0\times {\cal Q}_0}{E_{uv}}=
717: \oplus_{(u,v)\in {\cal Q}_0\times {\cal Q}_0}{\Hom_{\cal A}{(u,v)}}$
718: is encoded by an $R$-superbimodule structure on $E$. Moreover, the opposite decomposition
719: $E=\oplus_{(u,v)\in {\cal Q}_0\times {\cal Q}_0}{\Hom_{\cal A}{(v,u)}}$ is encoded by the opposite
720: superbimodule $E^{opp}$. These observations allow one to encode the category-theoretic
721: structure of $A_\infty$ products into compatibility with the
722: $R$-superbimodule structure of $E$. Before stating the relevant
723: result, we need a few more preparations.
724:
725:
726: Given an $R$-superbimodule $U$, its dual $U^{\rm v}:=\Hom_{\rm R-mod} (U,R)=\iHom_{\rm R-Smod}(U,R)$
727: as a left $R$-module becomes an $R$-superbimodule
728: with respect to the external multiplications defined through:
729: \be
730: \label{dual_multiplications}
731: (\alpha f \beta)(x):=f(\alpha x \beta)=\alpha f(x\beta) ~~.
732: \ee
733: We warn the reader that the usual definition of an $R$-superbimodule structure on $U^{\rm v}$ corresponds
734: to the opposite of that given in (\ref{dual_multiplications}). We adopted the convention
735: (\ref{dual_multiplications}) in order to avoid notational morass later on. With this definition, some of
736: the usual isomorphisms involve taking the opposite of certain superbimodules, as we explain in Appendix
737: \ref{isomorphisms}; in return, the formulas of Section 4,5 and 6 simplify considerably.
738:
739:
740: It is not hard to see that the binary homogeneous decomposition
741: $U^{\rm v}=\oplus_{(u,v)\in {\cal Q}_0\times {\cal Q}_0}{(U^{\rm v})_{uv}}$ determined by
742: the $R$-superbimodule structure (\ref{dual_multiplications}) has components:
743: \be
744: \label{dual_spaces}
745: (U^{\rm v})_{uv}=(U_{uv})^*~~,
746: \ee
747: where $(U_{uv})^*:=\iHom_\C(U_{uv},\C)$ is the linear dual of $U_{uv}$ viewed as a supervector space.
748: Notice that there is no reversal of the positions of $u$ and $v$ in relation (\ref{dual_spaces}).
749:
750:
751: With the definition (\ref{dual_multiplications}), a homogeneous
752: $R$-bilinear form\footnote{Recall that a multilinear map $f$ of $R$-superbimodules is required to satisfy
753: $f(\alpha x_1, x_2\ldots x_{n-1}, x_n\beta)=\alpha f(x_1\ldots
754: x_n)\beta$ as well as the balance condition $f(x_1\ldots
755: x_{j-1}\alpha, x_j\ldots x_n)=f(x_1\ldots x_{j-1}, \alpha x_j\ldots
756: x_n)$ for all $j$, where $\alpha,\beta\in R$.
757: In particular, a bilinear map $\sigma:U\times U\rightarrow R$
758: satisfies $\sigma (x\alpha,y)=\sigma (x,\alpha y)$ and $\sigma (\alpha x, y\beta)=\alpha\sigma(x,y)\beta$.}
759: $\sigma:U\times U\rightarrow R$ of degree ${\tilde \sigma}$ induces
760: an $R$-superbimodule map $U^{opp}\stackrel{j_\sigma}{\rightarrow}U[{\tilde \sigma}]^{\rm v}$ given
761: by $x\rightarrow f_x(.):=\sigma(\cdot, x)$.
762: The form is
763: called {\em graded-symmetric} if
764: $\sigma(x,y)=(-1)^{\deg x \deg y}\sigma(y,x)$ for all homogeneous $x,y\in U$
765: and {\em graded-antisymmetric} if
766: $\sigma(x,y)=(-1)^{1+\deg x \deg y}\sigma(y,x)$. In any of these case, it is called {\em
767: nondegenerate} if the map $x\rightarrow f_x$ is an isomorphism of vector spaces.
768:
769:
770: A graded-symmetric form $\rho$ on
771: $U$ induces a graded-antisymmetric form $\omega=\rho\circ
772: \Sigma^{\otimes 2}$ on $U[1]$, given explicitly by:
773: \be
774: \nn
775: \omega(\Sigma x,\Sigma y)=(-1)^{\tilde x}\rho(x,y)~~~~~~\forall x,y \in U~~,
776: \ee
777: where the sign prefactor is due to the Koszul rule (the
778: suspension operator $\Sigma$ is odd). If $\rho$ is homogeneous, then
779: $\omega$ is homogeneous of the same $\Z_2$ degree. Hence giving a
780: graded-symmetric form on $U$ is equivalent to giving a graded-antisymmetric
781: form on $U[1]$. Moreover, $\rho$ is non-degenerate iff
782: $\omega$ is. A non-degenerate, homogeneous and graded-symmetric
783: $R$-bilinear form will be called a {\em metric}, while a
784: non-degenerate, homogeneous and graded-antisymmetric R-bilinear form
785: will be called a {\em symplectic form}. A $R$-superbimodule is called
786: a {\em metric superbimodule} if it is endowed with a metric, and a {\em
787: symplectic superbimodule} if it is endowed with a symplectic form. Metric
788: and symplectic $R$-superbimodules form categories if one defines
789: morphisms in the obvious fashion. Notice that parity change induces
790: an idempotent equivalence between these categories. This reflects the
791: general principle that metric and symplectic superdata are related
792: through parity change. In particular, $(U[1],\omega)$ is a symplectic
793: R-superbimodule iff $(U,\rho)$ (with $\omega=\rho\circ \Sigma^{\otimes
794: 2}$) is a metric R-superbimodule.
795:
796: Giving an $R$-bilinear form $\sigma$ on $U$ amounts to giving $\C$-bilinear forms
797: $\sigma_{uv}:U_{uv}\times U_{vu}\rightarrow \C$ for all $u,v$. Indeed, $R$-bilinearity
798: of $\sigma$ implies:
799: \be
800: \nn
801: \sigma(x\epsilon_v, y)=\sigma(x, \epsilon_vy)~~{\rm and}~~\sigma(\epsilon_u x, y\epsilon_w)=
802: \sigma(x,y)\epsilon_u\epsilon_w~~.
803: \ee
804: Writing $x=\sum_{u,v}{x_{uv}}$ and $y=\sum_{u,v}{y_{uv}}$ with $x_{uv}:=\epsilon_u x \epsilon_v\in U_{uv}$ and
805: $y_{uv}:=\epsilon_u y\epsilon_v\in U_{uv}$, this shows that $\sigma(x_{uv}, y_{u'v'})$ vanishes unless $v=u'$ and
806: $u=v'$. Hence $\sigma$ is completely determined by its restrictions $\sigma'_{uv}$ to the subspaces
807: $U_{uv}\times U_{vu}$. Explicitly, we have $\sigma(x,y)=\sum_{u,v}\sigma'_{uv}(x_{uv},y_{vu})$.
808: Each restriction takes $U_{uv}\times U_{vu}$ into the one-dimensional subspace $\C \epsilon_u$ of $R$, so
809: we can write $\sigma'_{uv}=\sigma_{uv}\epsilon_u$ for some complex-linear maps
810: $\sigma_{uv}:U_{uv}\times U_{vu}\rightarrow \C$. Then:
811: \be
812: \label{sigma_dec}
813: \sigma(x,y)=\sum_{u,v}\sigma_{uv}(x_{uv},y_{vu})\epsilon_u
814: \ee
815: and $\sigma$ is completely determined by $\sigma_{uv}$. Conversely,
816: any family of complex-bilinear maps $\sigma_{uv}:U_{uv}\times
817: U_{vu}\rightarrow \C$ determines an $R$-bilinear map $\sigma$ through
818: relation (\ref{sigma_dec}). The map $U\stackrel{j_\sigma}{\rightarrow} U^{\rm v}$ induced by $\sigma$
819: has the property $j_\sigma(U^{\rm opp}_{uv})\subset (U^{\rm v})_{uv}$ i.e.
820: $j_\sigma(U_{vu})\subset (U_{uv})^*$, and its restrictions to the subspaces $U_{vu}$ can be identified
821: with the maps $U_{vu}\rightarrow (U_{uv})^*$ induced by $\sigma_{uv}$.
822: Thus $\sigma$ is non-degenerate iff $\sigma_{uv}$ are (the later
823: means that all maps $U_{vu}\rightarrow (U_{uv})^*$ determined by
824: $\sigma_{uv}$ are linear isomorphisms). It is also clear that $\sigma$ is
825: graded-symmetric iff $\sigma_{uv}(x,y)=(-1)^{{\tilde x}{\tilde
826: y}}\sigma_{uv}(y,x)$ for all homogeneous $x\in U_{uv}$ and $y\in
827: U_{vu}$; a similar statement holds for graded-antisymmetric forms.
828:
829:
830: For a topological D-brane system, we have $E_{uv}=\Hom_{\cal
831: A}(u,v)$ and the topological metrics $\rho_{uv}:\Hom_{\cal
832: A}(u,v)\times \Hom_{\cal A}(v,u)\rightarrow \C$ are homogeneous of common
833: degree ${\tilde \omega}$, so they determine an $R$-bilinear form
834: $\rho$ on $E$ of the same degree. Moreover, the graded-symmetry and
835: non-degeneracy conditions for $\rho_{uv}$ amount to the condition that
836: $\rho$ is a superbimodule metric on $E$. Similarly, the symplectic
837: forms $\omega_{uv}=\rho_{uv}\circ \Sigma^{\otimes 2}$ determine a
838: superbimodule symplectic form on $E[1]$ of degree ${\tilde \omega}$.
839:
840:
841: A {\em weak $A_\infty$ structure} on a $R$-superbimodule $U$ is a
842: countable family of odd $R$-linear maps $r_n:U[1]^{\otimes
843: n}\rightarrow U[1]$ (equivalently, odd $R$-multilinear maps
844: $r_n:U[1]^n\rightarrow U[1]$, which we denote by the same letters)
845: with $n\geq 0$, which satisfy the conditions\footnote{There is
846: some ambiguity in the sign conventions for various objects
847: related to $A_\infty$ algebras and categories. In this paper, we view the homological
848: derivation $Q$ discussed in Section \ref{geometrization} as the
849: fundamental object, and have defined $r_n$ such that most signs
850: simplify. We refer the reader to \cite{Penkava} for a discussion of
851: other conventions.}:
852: \be
853: \label{ainf}
854: \sum_{0\leq i+j\leq n} (-1)^{{\tilde x}_1+\ldots +{\tilde x}_i} r_{n-j+1}(x_1\ldots x_i, r_j(x_{i+1}
855: \ldots x_{i+j}), x_{i+j+1}\ldots x_n)=0~~
856: \ee
857: for all $n\geq 0$. In these relations, it is understood that $x_1,
858: \ldots, x_n$ are arbitrary homogeneous elements of $U[1]$. The
859: structure is called {\em strong}, if $r_0=0$ and {\em minimal} if
860: $r_0=r_1=0$.
861:
862: We say that a weak $A_\infty$ structure on $U$ is {\em
863: unital} if there exists an even element $1\in U^R$ such
864: that its odd suspension $\lambda =\Sigma 1\in U[1]^R$ satisfies the
865: following conditions:
866: \begin{eqnarray}
867: \label{unitality}
868: r_n(x_1\ldots x_{j-1},\lambda,x_{j+1}\ldots x_n)&=&0~~{\rm for~all}~~~~ n\neq 2 ~{\rm~and~all}~ j~~\nn\\
869: -r_2(\lambda,x)=(-1)^{{\tilde x}} r_2(x,\lambda)&=&x~~,
870: \end{eqnarray}
871: for all homogeneous elements $x,x_j$ of $U$. In this case, $1$ is
872: called the {\em even unit} of the $A_\infty$ structure, while
873: $\lambda$ will be called the {\em odd unit}. It is not hard to
874: see that the unit of a unital $A_\infty$
875: structure is unique. Indeed, given another unit $1'$, set $\lambda':=\Sigma
876: 1'$. Then $r_2(\lambda,\lambda')=-\lambda=-\lambda'$, where we used
877: the second row in (\ref{unitality}) by viewing either $\lambda$ or
878: $\lambda'$ as the unit. This implies $1=1'$.
879:
880: Recall that the {\em center} $U^R$ of an $R$-superbimodule $U$
881: (a.k.a the centralizer of $R$ in $U$) is the
882: homogeneous sub-bimodule consisting of all central elements $x\in U$, i.e. those
883: elements which satisfy $\alpha x=x\alpha$ for all $\alpha\in R$. In
884: terms of the homogeneous binary decomposition, we have
885: $U^R=\oplus_{u\in {\cal Q}_0}{U_{uu}}$. Thus a central element has
886: the form $x=\oplus_{u\in {\cal Q}_0}{x_u}$, with
887: $x_u=\epsilon_ux\epsilon_u=\epsilon_u x=x\epsilon_u\in U_{uu}$. Since
888: $R$ is semisimple, we have a direct sum decomposition of vector spaces:
889: \be
890: \label{CQ_decomp}
891: U=U^R\oplus [R,U]~~,
892: \ee
893: where $[R,U]$ is the complex supervector space generated by the
894: commutators $[\alpha,x]=\alpha x-x\alpha$ with $\alpha\in R$ and $x\in
895: U$. This follows as in \cite{CQ} due to the fact that $\oplus_{u\in
896: Q_0}{\epsilon_u\otimes \epsilon_u}$ is a so-called separability
897: element. It can also be seen directly by using the identities
898: $x-\sum_{u}{\epsilon_u
899: x\epsilon_u}=\frac{1}{2}\sum_{u}{[\epsilon_u,[\epsilon_u, x]]}$ and
900: $\epsilon_u[\epsilon_v, x]\epsilon_u=0$, which
901: hold for any element $x\in U$.
902:
903: Semisimplicity of $R$ implies that the unit of a unital $A_\infty$ structure
904: on an $R$-superbimodule $U$ must be central. To see this, let $\lambda$ be the odd
905: $A_\infty$ unit and consider the central element $\lambda'=\sum_{u}{\epsilon_u \lambda \epsilon_u}$.
906: Then the last row in (\ref{unitality}) implies:
907: \be
908: \nn
909: r_2(x,\lambda')=\sum_{u} r_2(x,\epsilon_u \lambda \epsilon_u)=
910: \sum_{u} r_2(x\epsilon_u, \lambda )\epsilon_u=(-1)^{\tilde x} \sum_u x\epsilon_u =(-1)^{\tilde x}x~~
911: \ee
912: and:
913: \be
914: \nn
915: r_2(\lambda',x)=\sum_{u} r_2(\epsilon_u \lambda \epsilon_u, x)=
916: \sum_{u} \epsilon_u r_2(\lambda, \epsilon_u x)=- \sum_u \epsilon_u x=-x~~,
917: \ee
918: where we used $R$-bilinearity of $r_2$ and the equation $\sum_u \epsilon_u=1$. These two equations
919: imply $\lambda=\lambda'$ by the argument used above to show unicity of the $A_\infty$ unit.
920: This shows that $\lambda$ must be central.
921:
922:
923:
924: A weak $A_\infty$ structure $(r_n)$ on $U$ is
925: called {\em cyclic} if $U[1]$ is endowed with a homogeneous symplectic form
926: $\omega$, such that the following relations are satisfied:
927: \be
928: \label{rrcyc}
929: \omega(x_0,r_n(x_1\ldots x_n))=(-1)^{{{\tilde x}_0+{\tilde x}_1+\tilde x}_0
930: ({\tilde x}_1+\ldots +{\tilde x}_n)}\omega(x_1,r_n(x_2\ldots x_n, x_0))~~.
931: \ee
932: In terms of the metric $\rho$ determined by $\omega=\rho\circ \Sigma^{\otimes 2}$,
933: these relations take the following form, which might be more
934: familiar to some readers: \be
935: \label{rcyc}
936: \rho(x_0,r_n(x_1\ldots x_n))=(-1)^{{\tilde x}_0({\tilde x}_1+\ldots
937: +{\tilde x}_n)}\rho(x_1,r_n(x_2\ldots x_n, x_0))~~. \ee
938:
939: \
940:
941: \noindent We can now state a basic equivalence:
942: \paragraph{\bf Proposition}
943: Giving a finite weak cyclic and unital $A_\infty$ category with object
944: set ${\cal Q}_0$ amounts to giving a weak, cyclic and unital
945: $A_\infty$ structure on a $\Z_2$-graded $R$-superbimodule $E$ of finite
946: complex dimension, over the finite-dimensional semisimple commutative
947: algebra $R=\oplus_{u\in {\cal Q}_0}\C \epsilon_u$.
948:
949:
950: \
951:
952: \noindent
953:
954: In view of this proposition, one can {\em define} finite tree-level
955: topological D-brane systems to be cyclic and unital weak $A_\infty$
956: structures on some $R$-superbimodule of finite dimension over $\C$, where $R$ is a
957: finite-dimensional semisimple commutative $\C$-algebra. The decomposition of the
958: system into constituent D-branes and the decomposition of the total
959: boundary space $E$ into boundary sectors
960: $E_{uv}$ are both encoded by the $R$-superbimodule structure.
961:
962:
963: \paragraph{Sketch of proof}
964: As explained above, the superbimodule structure on $E$ amounts to a
965: homogeneous decomposition $E=\oplus_{(u,v)\in {\cal Q}_0\times {\cal
966: Q}_0}{E_{uv}}$, where $E_{uv}:=\Hom_{\cal A}(u,v)$. The rest of the
967: proof is a lengthy but straightforward check of conditions, showing
968: that compatibility of various maps with the $R$-superbimodule
969: structure of $E$ allows one to translate superbimodule $A_\infty$ data
970: into the categorical $A_\infty$ data listed in Appendix \ref{data}.
971: We already showed above that giving the categorical bilinear forms $\rho_{uv}$
972: amounts to giving a metric $\rho$ on this superbimodule. Similarly,
973: giving an $R$-multilinear map $r_n:E^n\rightarrow E$ amounts to giving
974: $\C$-multilinear maps $r_{u_1\ldots u_{n+1}}: E_{u_1 u_2}\times E_{u_2
975: u_3}\times \ldots \times E_{u_n u_{n+1}}\rightarrow E_{u_1u_{n+1}}$,
976: and it is clear that the weak $A_\infty$ constraints for $r_n$ amount
977: to the categorical weak $A_\infty$ constraints for these maps.
978: Moreover, the cyclicity conditions for $r_n$ with respect to
979: $\omega=\rho\circ\Sigma^{\otimes 2}$ amount to the categorical
980: cyclicity constraints with respect to $\omega_{uv}:=\rho_{uv}\circ
981: \Sigma^{\otimes 2}$. Finally, the $A_\infty$ units $1_u$ of ${\cal
982: A}$ give an even central element $1=\oplus_{u}{1_u}$ in the
983: superbimodule $E$, which is a unit for the superbimodule
984: $A_\infty$ structure. Conversely, giving such a unit amounts
985: to giving elements $1_u$ in each `diagonal' subspace $E_{uu}$, since
986: the unit of an $R$- superbimodule $A_\infty$ structure must be central.
987: The unitality constraints for $r_n$ amount to the categorical unitality constraints
988: for $r_{u_1\ldots u_{n+1}}$.
989:
990: \paragraph{Observation} When the $A_\infty$ structure $(r_n)$ is minimal, the first
991: non-trivial $A_\infty$ constraint implies that the product
992: $\cdot=\Sigma\circ r_2\circ \Sigma^{\otimes 2}$ is associative.
993: Moreover, cyclicity and unitality imply that the triple $(E,\cdot
994: ,\rho)$ is a (non-commutative) Frobenius algebra, whose multiplication and
995: bilinear form are $R$-bilinear. On the other hand, forgetting all higher products of ${\cal A}$ gives a usual
996: (i.e. associative and unital) category endowed with non-degenerate and
997: graded symmetric bilinear pairings $\rho_{uv}$ between opposite spaces
998: of morphisms. As explained in \cite{CIL1}, such a category describes
999: the boundary part of a two-dimensional topological field theory
1000: defined on bordered Riemann surfaces. This gives the following:
1001:
1002:
1003:
1004: \paragraph{\bf Corollary} The boundary sector of a topological field theory in
1005: two-dimensions, in the presence of a finite system of topological D-branes,
1006: is described by a noncommutative Frobenius structure on
1007: an $R$-superbimodule $E$, i.e. a unital noncommutative
1008: Frobenius algebra on the vector superspace $E$, whose associative
1009: product and pairing are $R$-bilinear.
1010:
1011: \
1012:
1013: \subsection{The geometrization strategy}
1014:
1015: The algebraic formulation of the previous subsection
1016: allows one to avoid the notational morass of the category-theoretic description. One
1017: is still left with the rather complicated data of a cyclic and unital
1018: weak $A_\infty$ structure on $E$. To express this synthetically, we will
1019: use a $\Z_2$-graded version of the non-commutative symplectic
1020: geometry of quivers developed in \cite{Ginzburg, LB_quivers}. Much of the content of the next two sections is a
1021: relatively straightforward, though tedious, superextension of the construction
1022: of \cite{Ginzburg, LB_quivers}, so it might be useful to
1023: summarize the main points. Start with a cyclic and unital $A_\infty$
1024: structure on the $R$-superbimodule $E$. To encode this geometrically,
1025: we will proceed as follows:
1026:
1027: (1) Giving a weak
1028: $A_\infty$ structure on the superbimodule $E$ amounts to giving an odd
1029: derivation $Q$ of the tensor algebra $A=T_R E[1]^{\rm v}$, satisfying
1030: the condition $Q^2=0$. Hence $(A,Q)$ can be viewed as a noncommutative version of
1031: the $Q$-manifolds considered in \cite{Konts_Schwarz}.
1032:
1033: (2) The symplectic form on $E[1]$ induces a non-commutative symplectic
1034: form on $A$, and one can develop the non-commutative symplectic
1035: supergeometry of this algebra by extending the approach of
1036: \cite{Ginzburg, LB_quivers}. This gives notions of symplectic and
1037: Hamiltonian superderivations having the classical properties. The
1038: Karoubi complex $C_R(A)$ is acyclic in positive degrees so all symplectic
1039: superderivations are Hamiltonian. There is a $\Z_2$-graded version
1040: $\{.,.\}$ of the Kontsevich bracket, a super-Lie bracket on
1041: $C^0_R(A)[{\tilde \omega}]$, where ${\tilde \omega}$ is the
1042: $\Z_2$-degree of the symplectic form.
1043:
1044: (3) Cyclicity of $(r_n)$ amounts to the condition that $Q$ be a
1045: symplectic derivation, i.e. $L_Q\omega=0$, where $L_Q$ is the Lie
1046: superderivative along $Q$. Thus $(A,Q,\omega)$ is a noncommutative
1047: version of the QP-manifolds of \cite{Konts_Schwarz}.
1048:
1049:
1050: (4) The non-commutative generating function $W$ of the D-brane system
1051: is the Hamiltonian of the homological derivation $Q$; to determine this uniquely, we
1052: require that it `vanishes at zero' in an appropriate sense. Thus $W$ is an
1053: element of degree ${\tilde \omega}+1$ of the supervector space
1054: $C^0_R(A)=A/[A,A]$. The weak $A_\infty$ constraint $Q^2=0\Leftrightarrow
1055: [Q,Q]=0$ is equivalent with the condition $\{W,W\}=0$. The superderivation $Q$ can be
1056: reconstructed as the Hamiltonian derivation determined by $W$, so the $A_\infty$
1057: structure defined by $W$ is automatically cyclic.
1058:
1059:
1060:
1061: (6) Unitality of the weak $A_\infty$ structure is equivalent to the
1062: condition $\frac{1}{2}\ld_\lambda W=\mu$, where $\mu$ is a superized version of
1063: the moment map of \cite{BEV} and $\ld_\lambda$ is the cyclic
1064: derivative of $W$ with respect to the odd $A_\infty$ unit
1065: $\lambda$. The noncommutative generating function determines the
1066: symplectic form through this relation.
1067:
1068:
1069: (7) Since $R=\oplus_{u\in {\cal Q}_0}{\C \epsilon_u}$, the tensor algebra $A$ can be
1070: viewed as the path algebra of a superquiver ${\cal Q}$, a quiver on the
1071: vertex set ${\cal Q}_0$ endowed with a $\Z_2$-valued map on its set of
1072: arrows. The quiver presentation arises by choosing homogeneous bases
1073: of the vector space $E$ which are adapted to its binary homogeneous
1074: decomposition. Using quiver language amounts to working in `special'
1075: coordinates on the non-commutative space determined by $A$, where
1076: `special' means that the coefficients of the non-commutative
1077: symplectic form in such coordinates are complex numbers.
1078:
1079: (8) All constructions have natural quiver interpretations. For
1080: example, $C^0_R(A)$ can be described in terms of necklaces, which in our
1081: case are $\Z_2$-graded. Cyclic derivatives with respect to elements of
1082: the adapted basis translate into cyclic derivatives with respect to
1083: the quiver's arrows $a$. Relative to an appropriate adapted basis,
1084: the unitality constraint amounts to the requirement that $W$ has a certain
1085: dependence on an odd element $\sigma$ of $E[1]^{\rm v}$ determined by
1086: the odd $A_\infty$ unit $\lambda$.
1087:
1088: (9) One can view $A$ as the coordinate ring of a non-commutative
1089: supermanifold $\A_{\cal Q}$. Imposing the relations $\ld_a W=0$ gives the
1090: non-commutative extended vacuum space ${\cal Z}$, a `noncommutative subscheme'
1091: of $\A_{\cal Q}$. The non-commutative extended moduli space ${\cal M}$ is obtained by modding
1092: ${\cal Z}$ (in the GIT sense) through those
1093: symplectomorphisms which correspond to unital and cyclic $A_\infty$ automorphisms
1094: of the underlying D-brane category.
1095:
1096: We now proceed with the detailed discussion of these points.
1097:
1098: \section{Noncommutative symplectic geometry of $R$-superalgebras}
1099: \label{sg}
1100:
1101: In this section, we extend the construction of
1102: \cite{Ginzburg,LB_quivers} to the case of superalgebras. The proofs
1103: of most statements are straightforward adaptations of those given in
1104: \cite{Ginzburg,LB_quivers}, so I will only indicate the points where
1105: our conventions are important or something interesting happens.
1106:
1107: Let $R$ be a unital and commutative algebra over $\C$. An {\em
1108: $R$-superalgebra} is a unital superalgebra $A$ containing $R$ as a
1109: subalgebra in even degrees (notice that this is stronger than
1110: requiring that $A$ be a superalgebra over $R$, since
1111: we require that $R$ sits inside the degree zero subalgebra of $A$).
1112: A morphism of $R$-superalgebras is a morphism of
1113: superalgebras whose restriction to $R$ equals the identity map.
1114:
1115:
1116: \subsection{Noncommutative differential superforms}
1117:
1118: Given an $R$-superalgebra $A$, consider the $R$-superbimodule
1119: $A_R:=A/R$, where $A/R$ stands for the vector space quotient.
1120: We define the {\em space of relative noncommutative
1121: forms of $A$ over $R$} by $\Omega_R A:=\oplus_{n\geq 0}\Omega^n_R A$, where:
1122: \be
1123: \nn
1124: \Omega^n_R A= A\otimes_R T_R^n (A_R)=A\otimes_R A_R^{\otimes_R n}~~.
1125: \ee
1126: We write the elements of this space as $w=a_0da_1\ldots da_n$, where $da$ is the
1127: image of $a\in A$ under the projection $A\stackrel{d}{\rightarrow}
1128: A_R=A/R$ and juxtaposition stands for the tensor product over $R$.
1129: The space $\Omega_R A$ is given a differential algebra structure with product:
1130: \begin{eqnarray}
1131: &&(a_0 da_1\ldots da_n) (b_0 db_1 \ldots db_m):=a_0da_1\ldots da_n d(a_nb_0)db_1\ldots db_m\nn\\
1132: &+&\sum_{i=1}^{n-1}{(-1)^{i}a_0
1133: da_1\ldots d(a_{n-i}a_{n-i+1}) \ldots
1134: da_n db_0}\ldots db_m + (-1)^n a_0 a_1 da_2\ldots da_n db_0\ldots db_m\nn
1135: \end{eqnarray}
1136: and differential $d (a_0da_1\ldots d a_n)=da_0da_1\ldots da_n$.
1137: In this paper, we view $\Omega_R A$ as an $\N\times \Z_2$ graded algebra,
1138: whose $\N$-grading (with components $\Omega^n_R A$) is given by the `rank' of forms, and whose
1139: $\Z_2$ grading is induced from $A$. We denote the bidegree of homogeneous elements by $\deg w=({\bar
1140: w},{\tilde w})\in \N\times \Z_2$. As explained in the introduction, we
1141: always work with the pairing (\ref{pairing}), and will require that $d$
1142: has bidegree $(1,0)\in \Z\times \Z_2$. This means that $d$ has the derivation property:
1143: \be
1144: \nn
1145: d(w_1w_2)=(dw_1)w_2+(-1)^{{\bar w_1}}w_1\cdot d w_2~~{\rm for~all~}~~w_1,w_2\in \Omega_R A~~.
1146: \ee
1147: We stress that in our conventions $d$ has degree {\em zero} with
1148: respect to the $\Z_2$-grading. The detailed construction of $\Omega_R
1149: A$ is given in Appendix \ref{envelope}.
1150: The space $\Omega^1_R A$ has an $A$-superbimodule structure with multiplications:
1151: \be
1152: \nn
1153: \alpha(adb)\beta=(\alpha a)d(b\beta) -(\alpha a b)d\beta~~~~~~~~\forall a,b\in A,~~\alpha,\beta\in R~~.
1154: \ee
1155: As in \cite{CQ}, one has an isomorphism of bigraded
1156: algebras $\Omega_R A\approx T_A (\Omega^1_R A)$, which induces an $A$-superbimodule
1157: structure on $\Omega_R A$. In particular, $\Omega_R A$ is
1158: an $R$-superalgebra (since $R\subset A=\Omega^0_R A \subset \Omega_R A$).
1159:
1160:
1161: As usual, the pair $(\Omega_R A, d)$ has a universality property. To
1162: formulate it, we define an $R$-{\em differential superagebra} to be
1163: an $\N\times \Z_2$-graded unital differential algebra $(\Omega,d)$
1164: such that $\deg(d)=(1,0)$, $R\subset \Omega^{0,0}$ and $d(R)=0$, where
1165: $\Omega^{0,0}$ is the subspace of elements of vanishing bidegree. If
1166: $(\Omega_j,d_j)$ are two such algebras, a map
1167: $\phi:\Omega_1\rightarrow \Omega_2$ is called a morphism of
1168: $R$-differential superalgebras if:
1169:
1170: \noindent (1) $\phi$ is a morphism of unital $\N\times \Z_2$-graded
1171: algebras (in particular, $\phi$ has vanishing bidegree)
1172:
1173: \noindent (2) $\phi|_R=\id_R $
1174:
1175: \noindent (3) $d_2\circ \phi=\phi\circ d_1$~~.
1176:
1177: \noindent With this definition, $R$-differential superalgebras form a
1178: category. The universality property of $(\Omega_R A, d)$ is as follows.
1179: Given any $R$-differential superalgebra $(\Omega, d)$ and a morphism of $R$-superalgebras
1180: $\rho:A\rightarrow \Omega$ such that $\rho(A)\subset \Omega^{0}$, there exists a unique morphism
1181: $u:\Omega_R A\rightarrow \Omega$ of $R$-differential superalgebras such
1182: that $\rho=uj$, where $j:A\rightarrow \Omega_R A$ is the
1183: inclusion. Hence $(\Omega_R A,j)$ is an initial object among the pairs
1184: $(\Omega,\rho)$. In fact, the correspondence $\Omega\rightarrow
1185: \{$ $R$-superalgebra morphisms $\rho:A\rightarrow
1186: \Omega$ with $\rho (A)\subset \Omega^{0}~ \}$ defines a
1187: functor from the category of $R$-differential superalgebras to
1188: the category of sets. The universality property means that $(\Omega_R
1189: A, d)$ represents this functor, so it is the
1190: superdifferential envelope of $A$. In particular, any morphism
1191: $\phi:A_1\rightarrow A_2$ of $R$-superalgebras extends uniquely to a
1192: morphism $\phi^*:\Omega_R A_1\rightarrow \Omega_R A_2$ of $R$-differential
1193: superalgebras.
1194:
1195:
1196: \subsection{Super Lie derivatives and contractions}
1197:
1198: Recall that a left (right) derivation of $A$ is a derivation of $A$
1199: viewed as a left (right) supermodule over $A\otimes A^{op}$. Thus a homogeneous
1200: left derivation $D$ satisfies $D(ab)=(Da)b+(-1)^{{\tilde a}{\tilde D}}a Db$, while
1201: a homogeneous right derivation satisfies $(ab)D=a(bD)+(-1)^{{\tilde b}{\tilde D}}(aD) b$, where ${\tilde D}$ is the
1202: degree of $D$ and we use the convention that left derivations are written to the left, and right derivations
1203: are written to the right. We will sometimes also indicate this by writing arrows above $D$.
1204:
1205:
1206: A relative derivation of $A$ is a derivation which is $R$-linear,
1207: i.e. vanishes on the subalgebra $R$. We let $\Der_l(A) $ and
1208: $\Der_r(A)$ be the complex supervector spaces of {\em relative} left
1209: and right derivations of $A$, viewed as Lie superalgebras with respect to the
1210: supercommutator $[D_1,D_2]=D_1\circ D_2-(-1)^{{\tilde D}_1 {\tilde
1211: D_2}}D_2\circ D_1$, which satisfies $[D_1,D_2]=(-1)^{1+{\tilde D}_1{\tilde D}_2}[D_2,D_1]$.
1212: We let $\Der_{l,r}^\alpha (A)$ be the subspaces
1213: consisting of left and right relative derivations of degree $\alpha$.
1214:
1215: Similarly, let $\Der_{l,r}(\Omega_R A)$ be the $\Z\times \Z_2$-graded
1216: complex vector spaces of {\em relative} left/right derivations of $\Omega_R A$,
1217: i.e. those left/right derivations of $\Omega_R A$ which vanish on $R$. In this
1218: definition, we view $\Omega_R A$ as a $\Z\times \Z_2$ graded algebra
1219: with the degree pairing (\ref{pairing}); thus a bihomogeneous left derivation $D$ of
1220: $\Omega_R A$ satisfies:
1221: \be
1222: \nn
1223: D(w_1w_1)=Dw_1 w_2+(-1)^{\deg D \cdot \deg w_1}w_1 Dw_2~~,
1224: \ee
1225: while a bihomogeneous right derivation obeys:
1226: \be
1227: \nn
1228: (w_1w_2)D=w_1 (w_2D)+(-1)^{\deg D\cdot \deg w_2}(w_1D)w_2~~.
1229: \ee
1230: The spaces $\Der_{l,r}(\Omega_R A)$ are bigraded Lie superalgebras
1231: with respect to the bigraded supercommutator $[D_1,D_2]=D_1\circ
1232: D_2-(-1)^{\deg D_1 \cdot \deg D_2} D_2 \circ D_1$, which satisfies
1233: $[D_1,D_2]=(-1)^{1+\deg D_1\cdot \deg D_2} [D_2,D_1]$. We let
1234: $\Der_{l,r}^\alpha (\Omega_R A)$ be the subspaces consisting of left
1235: relative derivations of bidegree $\alpha$.
1236:
1237: Let $\theta\in \Der_l(A)$ be a homogeneous relative left derivation of
1238: degree ${\tilde \theta}$. The {\em contraction by $\theta$} is the
1239: unique relative left derivation $i_\theta \in \Der_{l}^{-1, {\tilde
1240: \theta}}(\Omega_R A)$ which satisfies $i_\theta a =0$ and $i_\theta (da)=\theta(a)$
1241: for all $a\in A$. The {\em Lie derivative along $\theta$} is
1242: the unique left derivation $L_\theta \in \Der_l^{0, {\tilde
1243: \theta}}(\Omega_R A)$ which satisfies $L_\theta a=\theta(a)$ and $L_\theta(da)=d\theta(a)$
1244: for all $a$. There are obvious versions of these definitions for right derivations.
1245: It is easy to check that $i_\theta$ and $L_\theta$ are well-defined.
1246: Let $\Aut_R(A)$ be the space of automorphisms of $A$ as an
1247: $R$-superalgebra (in particular, all maps $\phi\in \Aut_R(A)$ are even and restrict to the identity on the subalgebra $R$).
1248: For any $\theta,\gamma \in \Der_l(A)$ and $\phi\in \Aut_R(A)$, we have the identities:
1249: \begin{eqnarray}
1250: \label{Cartan}
1251: L_\theta&=&\left[i_\theta, d\right]\nn\\
1252: \left[L_\theta, i_\gamma\right]&=&i_{\left[\theta, \gamma\right]}\nn\\
1253: \left[L_\theta, L_\gamma\right]&=&L_{\left[\theta, \gamma\right]}\nn\\
1254: \left[i_\theta, i_\gamma\right]&=&0\\
1255: \left[L_\theta, d\right]&=&0 \nn\\
1256: L_{\phi \theta \phi^{-1}}&=&\phi^*L_\theta \phi^{* -1}\nn\\
1257: i_{\phi \theta \phi^{-1}}&=&\phi^*i_\theta \phi^{* -1}\nn~~.
1258: \end{eqnarray}
1259: As usual, these follow by noticing that all left and right hand sides
1260: are derivations of the same bi-degree on $\Omega_R A$, and checking agreement on the
1261: generators $a$ and $da$ ($a\in A$). Given $w=a_0da_1\ldots da_n$ with $a_i\in A$, we have:
1262: \begin{eqnarray}
1263: \label{ids}
1264: i_\theta w &=& \sum_{i=1}^n{
1265: (-1)^{i-1+{\tilde \theta}({\tilde
1266: a}_0+\ldots +{\tilde a}_{i-1})}
1267: a_0da_1\ldots da_{i-1} \theta(a_i)da_{i+1}\ldots da_n}\nn\\
1268: L_\theta w&=& \theta(a_0) da_1\ldots da_n+
1269: \sum_{i=1}^n{(-1)^{{\tilde \theta}({\tilde
1270: a}_0+\ldots +{\tilde a}_{i-1})}
1271: a_0da_1\ldots da_{i-1} d\theta(a_i)da_{i+1}\ldots da_n}~~.\nn
1272: \end{eqnarray}
1273:
1274: \subsection{The bigraded Karoubi complex}
1275:
1276: Consider the $\N\times \Z_2$-graded vector space:
1277: \be
1278: C_R(A):=\Omega_R A/[\Omega_R A,\Omega_R A]~~,
1279: \ee
1280: where $[\Omega_R A,\Omega_R A]\subset \Omega_R A$ is the image of the bigraded
1281: commutator map $[.,.]:\Omega_R A\times \Omega_R A\rightarrow \Omega_R A$:
1282: \be
1283: \nn
1284: [w_1,w_2]=w_1w_2-(-1)^{\deg w_1\cdot \deg w_2}w_2 w_1~~.
1285: \ee
1286: Notice that $[\Omega_R A,\Omega_R A]$ is a homogeneous subspace of $\Omega_R A$ but not a
1287: subalgebra. We let $\pi:\Omega_R A\rightarrow C_R(A)$ be the
1288: projection, and use the notation:
1289: \be
1290: \nn
1291: \pi(w):=(w)_c~~{\rm for}~~w\in \Omega_R A~~.
1292: \ee
1293: We also let $C_R^n(A)$ be the $\N$-homogeneous components of $C_R(A)$.
1294:
1295: Any relative derivation of $\Omega_R A$ preserves the subspace $[\Omega_R A, \Omega_R A]$,
1296: so it descends to a well-defined linear operator in $C_R(A)$. In
1297: particular, $d$ induces a differential ${\bar d}$ on $C_R(A)$. The bigraded
1298: vector space $(C_R(A),{\bar d})$ is the {\em relative Karoubi (or
1299: non-commutative de Rham) complex} of $A$ over $R$. This differential space is
1300: $\N\times \Z_2$-graded, and we have $\deg {\bar d}=(1,0)$. The
1301: supervector spaces $H^n_R(A):=H_{\bar d}^n(C_R(A))$ are called the
1302: {\em relative de Rham cohomology groups of $A$}.
1303:
1304: Given a morphism of $R$-superalgebras, the induced map
1305: $\phi^*:(\Omega_R A_1,d_1)\rightarrow (\Omega_R A_2, d_2)$ of $R$-differential superalgebras
1306: satisfies $\phi^*([\Omega_R A_1,\Omega_R A_1])\subset [\Omega_R A_2,\Omega_R
1307: A_2]$, so it descends to a morphism of bigraded complexes
1308: ${\bar \phi}^*:(C_R(A_1),{\bar d}_1)\rightarrow (C_R(A_2),{\bar
1309: d}_2)$. In particular, any endomorphism $\phi$ of $A$ induces an
1310: endomorphism ${\bar \phi}^*$ of $(C_R(A),{\bar d})$, which is an automorphism
1311: if $\phi$ is. We let ${\bar
1312: \phi}$ be the restriction of ${\bar \phi}^*$ to $C^0_R(A)=A/[A,A]$ (here $[A,A]\subset A$ is
1313: the image of the supercommutator $[.,.]:A\times A\rightarrow A$).
1314:
1315: The contraction and Lie operators $i_\theta$, $L_\theta$ also descend
1316: to well-defined $\C$-linear maps on $C_R(A)$, which we denote by ${\bar
1317: i}_\theta$, ${\bar L}_\theta$. It is clear that the induced operators
1318: satisfy all properties listed in eqs. (\ref{Cartan}):
1319: \begin{eqnarray}
1320: \label{Cartan_c}
1321: {\bar L}_\theta &=&\left[{\bar i}_\theta, {\bar d}\right]\nn\\
1322: \left[{\bar L}_\theta, {\bar i}_\gamma\right]&=&{\bar i}_{\left[\theta, \gamma\right]}\nn\\
1323: \left[{\bar L}_\theta, {\bar L}_\gamma\right]&=&{\bar L}_{\left[\theta, \gamma\right]}\nn\\
1324: \left[{\bar i}_\theta, {\bar i}_\gamma\right]&=&0\\
1325: \left[{\bar L}_\theta, {\bar d}\right]&=&0 \nn\\
1326: {\bar L}_{\phi \theta \phi^{-1}}&=&{\bar \phi}^*{\bar L}_\theta {\bar \phi}^{* -1}\nn\\
1327: {\bar i}_{\phi \theta \phi^{-1}}&=&{\bar \phi}^*{\bar i}_\theta {\bar \phi}^{* -1}\nn~~,
1328: \end{eqnarray}
1329: where $\phi\in \Aut_R(A)$~~.
1330:
1331: \subsection{Noncommutative supersymplectic forms}
1332:
1333: An element $\omega\in C^2_R(A)$ is called non-degenerate if the following complex-linear map is bijective:
1334: \be
1335: \nn
1336: \theta\in \Der_l(A)\rightarrow {\bar i}_\theta \omega\in C^1_R(A)~~.
1337: \ee
1338: A {\em relative non-commutative symplectic form} on $A$ is
1339: a $\Z_2$-homogeneous element $\omega\in C^2_R(A)$ which is closed (${\bar d}\omega=0$) and non-degenerate.
1340:
1341:
1342: Given a symplectic form $\omega$ of $\Z_2$-degree ${\tilde \omega}$, a relative derivation $\theta\in \Der_l(A)$
1343: is called {\em symplectic} if ${\bar L}_\theta\omega=0$. Let
1344: $\Der_l^\omega(A)\subset \Der_l(A)$ be the subspace of all symplectic
1345: derivations. By the third property in (\ref{Cartan_c}), this is a (super) Lie subalgebra of $\Der_l(A)$.
1346: As in the even case, it is easy to see that the following map is an isomorphism of supervector spaces:
1347: \be
1348: \label{isom_0}
1349: \theta\in \Der_l^\omega(A)\rightarrow {\bar i}_\theta\omega\in C^1_R(A)_{\rm closed}[{\tilde \omega}]~~.
1350: \ee
1351: Here $C^1_R(A)_{\rm closed}=\{\eta\in C^1_R(A)|{\bar d} \eta=0\}$, a homogeneous subspace of $C^1_R(A)$.
1352:
1353: This implies that any $f\in C^0_R(A)$ defines a unique
1354: element $\theta_f\in \Der_l^\omega(A)$, determined by the equation
1355: ${\bar i}_{\theta_f}\omega={\bar d}f$. Let
1356: $\psi_\omega:C^0_R(A)\rightarrow \Der_l^\omega(A)$ be the complex-linear map
1357: given by $\psi_\omega(f):=\theta_f$. The relation ${\bar
1358: i}_{{\theta}}\omega={\bar d}f$ implies ${\tilde \theta}_f={\tilde
1359: \omega}+{\tilde f}$ for any homogeneous $f$, so the map $\psi_\omega$
1360: is homogeneous of degree ${\tilde \omega}$. It is clear that the
1361: following sequence of supervector spaces is exact:
1362: \be
1363: \label{ex_0}
1364: 0\rightarrow
1365: H^0_R(A)\hookrightarrow C^0_R(A)\stackrel{\psi_\omega}{\rightarrow}\Der_l^\omega(A)[\tom]~~,
1366: \ee
1367: where the map in the middle is the inclusion. Elements $\theta\in {\rm
1368: im}\psi_\omega$ are called {\em Hamiltonian derivations}. Given a
1369: Hamiltonian derivation $\theta$, an element $f\in C^0_R(A)$ such that
1370: $\psi_\omega(f)=\theta\Leftrightarrow \theta=\theta_f$ is called a
1371: {\em Hamiltonian} associated with $\theta$. The sequence (\ref{ex_0})
1372: shows that the Hamiltonian of a Hamiltonian derivation is determined
1373: up to addition of elements of $H^0_R(A)$.
1374:
1375: \subsection{$\Z_2$-graded version of the Kontsevich bracket}
1376: \label{sec:bracket}
1377:
1378: We have the following generalization of an operation introduced in \cite{Konts_formal}.
1379:
1380: \paragraph{\bf Definition}
1381: The {\em Kontsevich bracket} induced by $\omega$ is the $\Z_2$-homogeneous
1382: complex-linear map $\{.,.\}:C^0_R(A)\otimes_\C C^0_R(A)\rightarrow C^0_R(A)$ of degree $\tom$ defined through:
1383: \be
1384: \label{bdef}
1385: \{f,g\}:={\bar i}_{\theta_f}{\bar i}_{\theta_g}\omega~~~~~\forall f,g \in C^0_R(A)~~,\nn
1386: \ee
1387: where $\theta_f=\psi_\omega(f)$ and $\theta_g=\psi_\omega(g)$.
1388:
1389: \
1390:
1391: \noindent Notice the relation:
1392: \be
1393: \label{bL}
1394: \{f,g\}={\bar L}_{\theta_f}g~~.
1395: \ee
1396: The following result gives the basic properties of the bracket in the $\Z_2$-graded case.
1397: \paragraph{\bf Proposition}
1398: The Kontsevich bracket satisfies the identities:
1399: \be
1400: \label{gasym}
1401: \{g,f\}=(-1)^{1+({\tilde f}+{\tilde \omega})({\tilde g}+{\tilde \omega})}\{f,g\}~~.
1402: \ee
1403: and:
1404: \be
1405: \label{gjacobi}
1406: (-1)^{(\tf_1+\tom)(\tf_3+\tom)}
1407: \{f_1,\{f_2,f_3\}\}+(-1)^{(\tf_2+\tom)(\tf_1+\tom)}\{f_2,\{f_3,f_1\}\}+
1408: (-1)^{(\tf_3+\tom)(\tf_2+\tom)}\{f_3,\{f_1,f_2\}\}=0~~.
1409: \ee
1410: Hence $(C^0_R(A)[\tom],\{.,.\})$ is a Lie superalgebra.
1411:
1412: \begin{proof}
1413:
1414: The first property follows immediately from $[{\bar i}_{\theta_f},{\bar i}_{\theta_g}]=0$.
1415: For the second property, let $f_1,f_2,f_3\in C^0_R(A)$ and
1416: $\theta_i:=\phi_\omega(f_i)$. Identities (\ref{Cartan_c}) give:
1417: \begin{eqnarray}
1418: 0={\bar i}_{\theta_2}{\bar i}_{\theta_1}
1419: {\bar i}_{\theta_0}d\omega&=&(-1)^{{\tilde
1420: \theta}_0({\tilde \theta_1}+{\tilde
1421: \theta}_2)}{\bar L}_{\theta_0}{\bar i}_{\theta_2}{\bar i}_{\theta_1}\omega+
1422: (-1)^{1+{\tilde \theta}_1 {\tilde
1423: \theta}_2}{\bar L}_{\theta_1}{\bar i}_{\theta_2}{\bar i}_{\theta_0}\omega+
1424: {\bar L}_{\theta_2}{\bar i}_{\theta_1}{\bar i}_{\theta_0}
1425: \omega\nn\\
1426: &+&
1427: (-1)^{1+\th_0\th_1}{\bar i}_{\theta_2}{\bar i}_{[\theta_0,\theta_1]}\omega+(-1)^{\th_2(\th_0+\th_1)}
1428: {\bar i}_{\theta_1}
1429: {\bar i}_{[\theta_0,\theta_2]}\omega+(-1)^{1+\th_0\th_1+\th_2(\th_0+\th_1)}{\bar i}_{\theta_0}
1430: {\bar i}_{[\theta_1,\theta_2]}\omega~~.\nn
1431: \end{eqnarray}
1432: Using the relations
1433: $[\theta_i,\theta_j]=(-1)^{1+\th_i\th_j}[\theta_j,\theta_i]$ and the
1434: properties of ${\bar L}_{\theta_i}$, the
1435: right hand side can be brought to the form:
1436: \be
1437: 2\left[(-1)^{1+\th_0\th_1+\th_1\th_2+\th_2\th_0}\{f_0,\{f_1,f_2\}\}+(-1)^{1+\th_1\th_2}
1438: \{f_1,\{f_2,f_0\}\}+(-1)^{1+\th_0\th_1}\{f_2,\{f_0,f_1\}\}\right]\nn
1439: \ee
1440: Multiplying with $(-1)^{1+\th_0\th_1+\th_1\th_2}$ leads to the
1441: identity:
1442: \be
1443: (-1)^{\th_0\th_2}\{f_0,\{f_1,f_2\}\}+(-1)^{\th_1\th_0}\{f_1,\{f_2,f_0\}\}+
1444: (-1)^{\th_2\th_1}\{f_2,\{f_0,f_1\}\}=0~~.\nn
1445: \ee
1446: This implies equation (\ref{gjacobi}) upon changing the indices
1447: $0,1,2$ into $1,2,3$ and using $\th_i={\tilde
1448: f}_i+{\tilde \omega}$.
1449:
1450: \end{proof}
1451:
1452:
1453: \noindent The map $\psi_\omega$ has another property which parallels classical behavior.
1454:
1455: \paragraph{\bf Proposition}
1456: {\em We have:
1457: \be
1458: \label{Lie_morphism}
1459: \theta_{\{f,g\}}=[\theta_f, \theta_g]~~~~~~\forall f,g\in C^0_R(A)~~.
1460: \ee
1461: Thus $\psi_\omega:(C^0_R(A)[\tom],\{.,.\}) \rightarrow (\Der_l^\omega(A),[.,.])$ is a morphism of Lie
1462: superalgebras over $\C$. }
1463:
1464: \begin{proof}
1465: Compute:
1466: \be
1467: {\bar d}\{f,g\}={\bar d}{\bar i}_{\theta_f}{\bar i}_{\theta
1468: g}\omega={\bar L}_{\theta_f}{\bar i}_{\theta_g}\omega+{\bar i}_{\theta_f}
1469: {\bar d}({\bar i}_{\theta_g}\omega)={\bar L}_{\theta_f}{\bar i}_{\theta_g}\omega=[{\bar L}_{\theta_f},
1470: {\bar i}_{\theta_g}]\omega={\bar i}_{[\theta_f,\theta_g]}\omega~~.\nn
1471: \ee
1472: In the third equality, we used ${\bar i}_{\theta_g}\omega={\bar d}g$ and ${\bar d}^2=0$,
1473: while in the fourth we used ${\bar L}_{\theta_f}\omega=0$.
1474: \end{proof}
1475:
1476:
1477: An $R$-superalgebra automorphism $\phi\in \Aut_R(A)$ is called a {\em relative
1478: symplectomorphism} if ${\bar \phi}^*(\omega)=\omega$. We let
1479: $\Aut^\omega_R(A)\subset \Aut_R (A)$ be the subgroup of relative
1480: symplectomorphisms of $A$. By the sixth property in (\ref{Cartan_c}),
1481: the obvious action of $\Aut^\omega_{alg}(A)$ on $\Der_l(A)$ preserves
1482: the Lie subalgebra $\Der_l^\omega(A)$. Given a symplectomorphism
1483: $\phi$, the last property in (\ref{Cartan_c}) implies $\phi\circ
1484: \theta_f\circ \phi^{-1}=\theta_{{\bar \phi}(f)}$ for $f\in C^0_R(A)$,
1485: i.e. $\psi_\omega\circ {\bar \phi} ={\rm Ad}_{\phi}\circ \psi_\omega$
1486: for all $\phi\in \Aut^\omega_R(A)$. In turn, this gives $\{{\bar
1487: \phi}(f),{\bar \phi}(g)\}={\bar \phi}(\{f,g\})$. Hence
1488: $\Aut^\omega_R(A)$ acts on $(A[{\tilde \omega}],\{.,.\})$ by Lie
1489: algebra automorphisms.
1490:
1491:
1492:
1493: \section{Noncommutative calculus for finite D-brane systems}
1494:
1495: Consider a finite D-brane system with object set ${\cal Q}_0$. As in
1496: Section \ref{algebraic}, we set $R=\oplus_{u\in {\cal Q}_0}{\C \epsilon_u}$ and let $E$ be the
1497: $R$-superbimodule of boundary sectors. Recall that $E[1]$ carries a
1498: symplectic form $\omega=\rho\circ \Sigma^{\otimes 2}$, the parity change of the topological metric $\rho$.
1499: Setting $V=E[1]^{\rm v}$, we consider the $\N\times \Z_2$ graded tensor
1500: algebra $A=T_R (V)$ whose $\N$-homogeneous subspaces we denote by $A_n=T_R^n (V)$.
1501: With respect to its $\Z_2$-grading, $A$ is an $R$-superalgebra with $R=A_0$.
1502:
1503: \subsection{Generalities}
1504: \label{tensor}
1505:
1506: The symplectic form $\omega$ on $E[1]$ induces a
1507: relative noncommutative symplectic form $\omega_{form}$ on $A$ as
1508: follows. Recall from Section 2 that the symplectic form $\omega$ defines
1509: a map $j_\omega: E[1]^{opp}\rightarrow E[1]^{\rm v}$ which is a morphism of $R$-bimodules, i.e.
1510: an element of $\iHom(E[1]^{opp},E[1]^{\rm v})$.
1511: As explained in Appendix \ref{isomorphisms}, there exists
1512: an isomorphism of $R$-superbimodules between
1513: $\iHom(E[1]^{opp},E[1]^{\rm v})$ and the center of the superbimodule
1514: $E[1]^{\rm v}\otimes_R E[1]^{\rm v}$,
1515: which allows us to view $\omega$ as an element ${\hat \omega}$ of degree ${\tilde \omega}$
1516: of the space $(V\otimes_R V)^R$. If ${\hat \omega}=\sum_{i}{f_i\otimes_R g_i}$ with $f_i,g_i\in V$,
1517: then it is shown in Appendix \ref{isomorphisms} that $\omega$ can be recovered as:
1518: \be
1519: \label{omega_rec}
1520: \omega(x,y)=\sum_i f_i(xg_i(y))~~.
1521: \ee
1522: Using the element ${\hat \omega}$, we define a noncommutative two-form on $A=T_R V$ through the relation:
1523: \be
1524: \label{omega_form_gen}
1525: \omega_{form}=-\frac{1}{2}\sum_{i}{(df_i dg_i)_c}\in C^2_R(A)~~,
1526: \ee
1527: where the minus sign is introduced for later convenience.
1528: It is easy to see that $\omega_{form}$ is well-defined
1529: and of the same $\Z_2$-degree as $\omega$. Moreover, it is not hard to check that this two-form is
1530: symplectic.
1531:
1532:
1533: Since $A=T_R V$, the algebra of noncommutative forms $\Omega_RA$ has a
1534: second $\N$-grading, which is induced from the $\N$-grading of $A$
1535: (with respect to this grading, we have $\deg a =\deg da=1$ for all
1536: $a\in V$, while $R$ sits in degree zero). This induces a similar
1537: $\N$-grading on $C_R(A)$. We let $(\Omega_RA)_n$ and $C_R(A)_n$ be
1538: the homogeneous subspaces determined by this grading. A {\em
1539: constant} noncommutative two-form \footnote{The notion of constant
1540: noncommutative symplectic form depends on the specific realization of
1541: $A$ as a tensor algebra $T_R V$. In particular, this concept is not
1542: invariant under the $R$-superalgebra automorphism group $\Aut_R(A)$
1543: (because a superalgebra automorphism need not preserve the
1544: $\N$-grading of $A$).} on $A$ is an element of $C^2_R(A)_2$. Thus
1545: (\ref{omega_form_gen}) is a constant symplectic form, and any constant
1546: symplectic form on $A$ has such an expansion. We let $CNS^{\tilde \omega}(V)$ be the
1547: vector space of all constant noncommutative symplectic forms on $T_R V$ of degree ${\tilde \omega}$. In
1548: what follows, we often denote $\omega_{form}$ by $\omega$; which of
1549: the two is meant should be clear from the context.
1550:
1551:
1552: Notice that $[A,A]\cap A_0=[A_0,A_0]=[R,R]=0$ since $A_0=R$ and $R$ is commutative. In particular,
1553: we have $C^0_R(A)_0=R$. The following result follows as in \cite{Ginzburg}, by
1554: considering the Euler derivation associated with the $\N$-grading induced from $A$.
1555:
1556: \paragraph{\bf Proposition} {\em We have $H^0_R(A)=C^0_R(A)_0=R$ and $H^n_R(A)=0$ for all $n\geq 1$.}
1557:
1558: \
1559:
1560: \noindent Using this in (\ref{ex_0}) gives a short exact sequence:
1561: \be
1562: \label{ex}
1563: 0\rightarrow
1564: R\rightarrow C^0_R(A)\stackrel{\psi_\omega}{\rightarrow}\Der_l^\omega(A)[\tom]\rightarrow 0~~,
1565: \ee
1566: where surjectivity of $\psi_\omega$ follows by using $H^1_R(A)=0$
1567: in (\ref{isom_0}). In particular, {\em any relative symplectic derivation of $A$
1568: is Hamiltonian}. The sequence (\ref{ex}) also shows that the
1569: Hamiltonian of a symplectic derivation is determined up to addition of
1570: elements of $R$, which can be viewed as the subspace $C^0_R(A)_0$ of $C^0_R(A)$.
1571:
1572: Any element $f$ of $C^0_R(A)$ has a decomposition $f=\sum_{n\geq
1573: 0}{f_n}$, with $f_n\in C^0_R(A)_n$ (in particular, $f_0\in C^0_R(A)_0=R$).
1574: We say that $f$ {\em has order $k$
1575: at zero} if $f_0= f_1= \ldots =f_{k-1}=0$ and $f_k\neq 0$. We say that
1576: $f$ {\em vanishes at zero} if $f_0=0$, i.e. $f$ has order at least one
1577: at zero. We define the {\em canonical Hamiltonian} of a symplectic
1578: derivation $\theta$ to be that Hamiltonian of $\theta$ which vanishes
1579: at zero. The sequence (\ref{ex}) shows that the canonical
1580: Hamiltonian exists and is unique.
1581:
1582:
1583:
1584:
1585:
1586: \subsection{Adapted bases and superquivers}
1587: \label{coord_description}
1588:
1589: Using the binary decomposition $E=\oplus_{(u,v)\in {\cal Q}_0\times {\cal Q}_0}{E_{uv}}$,
1590: consider a homogeneous basis $(\psi_a)$ of the supervector space $E$ having the following properties:
1591:
1592: \
1593:
1594: (1) $a=(u,v, j)$ is a multi-index with $u,v \in {\cal Q}_0$ and $j=1\ldots \dim_\C E_{uv}$
1595:
1596: \
1597:
1598: (2) $\psi_{uvj}$ for $j=1\ldots \dim_\C E_{uv}$ is a
1599: homogeneous basis of $E_{uv}$ for all $u,v\in {\cal Q}_0$.
1600:
1601: \
1602:
1603: \noindent We say that such a homogeneous basis is {\em adapted} to the binary decomposition of $E$.
1604: Setting $e_a:=\Sigma \psi_a$, we let $(s^a)$ be the basis of the super-vector space
1605: $V=E[1]^{\rm v}$ dual to $(e_a)$:
1606: \be
1607: \nn
1608: s^a(e_b)=\delta^a_b~~.
1609: \ee
1610: Relation (\ref{dual_spaces}) shows that
1611: $s^{uvj}$ are bases of $V_{uv}=(E[1]^{\rm v})_{uv}=(E_{uv}[1])^*$, odd dual to the
1612: bases $\psi_{uvj}$ of $E_{uv}$. We set ${\tilde
1613: a}:=\deg s^a=\deg e_a=[|a|+1]$, where $|a|:=|\psi_a|$.
1614:
1615:
1616: It is convenient to keep track of indices by considering a quiver
1617: ${\cal Q}$ determined by the multi-indices $a$. Specifically, the {\em index quiver} ${\cal Q}$ is the quiver
1618: on the vertex set ${\cal Q}_0$ obtained by drawing an arrow from $u$ to $v$ for
1619: each $j=1\ldots \dim_\C V_{uv}$. With this construction, we can identify each
1620: multi-index $a$ with the corresponding arrow of ${\cal Q}$. We let ${\cal Q}_1$ be the set of all
1621: arrows and ${\cal Q}_1(u,v)$ the subset of arrows going from $u$ to $v$. We
1622: also let $h,t:{\cal Q}_1\rightarrow {\cal Q}_0$ be the head and tail maps of ${\cal Q}$.
1623:
1624: The index quiver is in fact a {\em superquiver},
1625: being endowed with a map $\deg: {\cal Q}_1\rightarrow \Z_2$
1626: given by $\deg (a)={\tilde a}$. An arrow $a$
1627: is called even if ${\tilde a}=0$ and odd if ${\tilde a}=1$. The path algebra
1628: $\C {\cal Q}$ becomes a superalgebra by declaring a path $p$ to be even
1629: or odd if it contains an even or odd number of odd arrows. That is, we
1630: define the degree of $p$ by the formula:
1631: \be
1632: \nn
1633: {\tilde p}:=\sum_{j=1}^k{\tilde a_j}~~,
1634: \ee
1635: where $p=a_1\ldots a_k$ is the arrow decomposition. The trivial paths
1636: are taken to be even. The path algebra is in fact $\N\times
1637: \Z_2$-graded, where the $\N$-grading is induced by the length of paths.
1638: We let $(\C {\cal Q})_n$ be the components of degree $n$ with respect to the length grading.
1639:
1640: As usual, the subspace spanned by the trivial paths forms a finite-dimensional semisimple commutative algebra.
1641: We identify this with the boundary sector algebra $R$ by sending the trivial path at $u$
1642: into the idempotent $\epsilon_u$. On the other hand, the subspace spanned by
1643: the arrows is isomorphic with the supervector space $V$:
1644: \be
1645: \nn
1646: (\C {\cal Q})_1\approx \oplus_{u,v\in {\cal Q}_0}\oplus_{a\in {\cal Q}_1(u,v)}\C[{\tilde a}]\approx V~~
1647: \ee
1648: via the identification $a\equiv s^a$. It is also clear that the
1649: $R$-superbimodule structure of $V$ coincides with the
1650: $(\C{\cal Q})_0$-superbimodule structure induced on $(\C{\cal Q})_1$ by multiplication in
1651: the path algebra. In fact, the entire path algebra is isomorphic
1652: with the tensor algebra $T_R (\C {\cal Q})_1$ as an $\N\times
1653: \Z_2$-graded algebra. Combining these observations, we find an isomorphism of bigraded algebras:
1654: \be
1655: \nn
1656: A\approx \C{\cal Q}
1657: \ee
1658: which extends the isomorphism $V \approx (\C {\cal Q})_1$. Hence:
1659:
1660: \
1661:
1662: {\em Choosing an adapted basis of $E$ identifies the tensor algebra
1663: $A=T_R V$ with the path algebra of the index superquiver ${\cal Q}$.}
1664:
1665: \
1666:
1667: \noindent In the next subsections, we explore the consequences of this identification.
1668:
1669: \subsection{Quiver description of the symplectic structure}
1670: \label{symplectic_structure}
1671:
1672: Recall that the symplectic form $\omega$ on $E[1]$ corresponds to bilinear forms $\omega_{uv}:E[1]_{uv}\times
1673: E[1]_{vu}\rightarrow \C$ of common degree ${\tilde \omega}$, such that $\omega_{uv}(x,y)=(-1)^{{\tilde
1674: x}{\tilde y}+1}\omega_{vu}(y,x)$. We define coefficients $(\omega_{ab})_{a,b\in {\cal Q}_1}$
1675: through:
1676: \be
1677: \nn
1678: \omega_{ab}=-\omega_{t(a)h(a)}(e_a,e_b)\in \C~~{\rm~if~}~~t(a)=h(b)~~{\rm~and~}~~h(a)=t(b)
1679: \ee
1680: and zero otherwise. The minus sign in this expression is introduced in
1681: order to simplify certain formulas which will appear in the next sections.
1682: In terms of the superbimodule symplectic form, we have
1683: $\omega(e_a,e_b)=-\omega_{ab}\epsilon_u$ where $u:=t(a)=h(b)$.
1684: The coefficients have the graded-antisymmetry properties:
1685: \be
1686: \label{omega_symm}
1687: \omega_{ab}=(-1)^{{\tilde a}{\tilde b}+1}\omega_{ba}~~
1688: \ee
1689: and satisfy the selection rule:
1690: \be
1691: \label{omega_sel}
1692: \omega_{ab}=0~~{\rm~unless~}~~{\tilde a}+{\tilde b}={\tilde \omega}~~.
1693: \ee
1694: It is not hard to see that the inverse matrix $(\omega^{ab})_{a,b\in {\cal Q}_1}$ also satisfies the selection
1695: rule:
1696: \be
1697: \nn
1698: \omega^{ab}=0~~{\rm unless}~~{\tilde a}+{\tilde b}={\tilde \omega}
1699: \ee
1700: and graded antisymmetry property:
1701: \be
1702: \nn
1703: \omega^{ab}=(-1)^{{\tilde a}{\tilde b}+1}\omega^{ba}~~.
1704: \ee
1705: Moreover, $\omega^{ab}$ vanishes unless $h(a)=t(b)$ and $t(a)=h(b)$.
1706:
1707: Let ${\hat \omega}$ be the central element of $V\otimes_R V$ determined by $\omega$. Since $s^a$ is a vector
1708: space basis of $V$, we can expand ${\hat \omega}=\sum_{a,b}{{\hat \omega}_{ab}s^a\otimes_R s^b}=
1709: \sum_{a}{s^a\otimes_R (s^a)'}$, where we set $(s^a)':=\sum_{b}{\hat \omega}_{ab}s^b$.
1710: Then equation (\ref{omega_rec}) gives $\omega(e_a,e_b)=\sum_{c}s^c(e_a (s^c)'(e_b))={\hat \omega}_{ab}$.
1711: Thus ${\hat \omega}_{ab}=-\omega_{ab}\epsilon_{t(a)}$ and we find that our definition of coefficients
1712: corresponds to the expansion:
1713: \be
1714: \nn
1715: {\hat \omega}=-\sum_{a,b\in {\cal Q}_1}\omega_{ab}s^a\otimes_R s^b\in V\otimes_R V~~,
1716: \ee
1717: where we used the relation $\epsilon_{t(a)} s^a=s^a$, which holds because $s^a\in V_{t(a)h(a)}$.
1718: Using equation (\ref{omega_form_gen}),
1719: it follows that the non-commutative symplectic form induced on $A$ is given by:
1720: \be
1721: \label{omega_form}
1722: \omega_{form}=\frac{1}{2}\sum_{\tiny
1723: \begin{array}{c}a,b\in {\cal Q}_1 \\h(a)=t(b), h(b)=t(a)\end{array}}\omega_{ab}(ds^a ds^b)_c\equiv
1724: \frac{1}{2}\sum_{a,b\in Q_1}\omega_{ab}(da db)_c~~.
1725: \ee
1726: In the last equality, notice that $(dadb)_c$ vanishes unless
1727: $h(a)=t(b)$ and $h(b)=t(a)$, which can be seen immediately by
1728: inserting idempotents $\epsilon_u$ in the appropriate places. As in
1729: \cite{Ginzburg, LB_quivers}, it is easy to check that
1730: non-degeneracy of the constant two-form (\ref{omega_form})
1731: amounts to non-degeneracy of the matrix $(\omega_{ab})_{a,b\in {\cal Q}_1}$. A {\em symplectic
1732: superquiver} is a superquiver whose path algebra is endowed with a
1733: constant symplectic form of type (\ref{omega_form}). Hence picking an
1734: adapted homogeneous basis allows us to encode the information of the
1735: symplectic superbimodule $(E[1],\omega)$ into a symplectic
1736: superquiver. Of course, the inverse correspondence also holds.
1737:
1738:
1739:
1740: Recall that the topological metric $\rho$ on $E$ is given by
1741: $\omega=\rho\circ \Sigma^{\otimes 2}$. We define its coefficients
1742: trough:
1743: \be
1744: \nn
1745: \rho_{ab}=\rho(\psi_a,\psi_b)~~,
1746: \ee
1747: {\em without} a minus sign insertion. Equation $\omega(e_a,e_b)=(-1)^{\tilde a}\rho(\psi_a,\psi_b)$ gives:
1748: \be
1749: \label{omega_rho}
1750: \omega_{ab}=(-1)^{{\tilde a}+1}\rho_{ab}~~.
1751: \ee
1752: The coefficients of $\rho$ have the properties:
1753: \be
1754: \nn
1755: \rho_{ab}=(-1)^{{\tilde a}{\tilde b}+{\tilde \omega}+1}\rho_{ba}=(-1)^{|a||b|}\rho_{ba}
1756: \ee
1757: and:
1758: \be
1759: \nn
1760: \rho_{ab}=0{\rm~unless~}{\tilde a}+{\tilde b}={\tilde \omega}\Longleftrightarrow |a|+|b|={\tilde \omega}~~.\nn
1761: \ee
1762: Relation (\ref{omega_rho}) shows that the inverse of the matrix $(\rho_{ab})$ takes the form:
1763: \be
1764: \rho^{ab}=(-1)^{{\tilde b}+1}\omega ^{ab}~~.\nn
1765: \ee
1766: It is clear that the inverse matrix satisfies the relations:
1767: \be
1768: \label{oinv_symm}
1769: \rho^{ab}=(-1)^{{\tilde a}{\tilde b}+{\tilde \omega}+1}\rho^{ba}=(-1)^{|a||b|}\rho^{ba}~~.
1770: \ee
1771: and
1772: \be
1773: \label{oinv_sel}
1774: \rho^{ab}=0{\rm~unless~}{\tilde a}+{\tilde b}={\tilde \omega}\Longleftrightarrow |a|+|b|={\tilde \omega}~~.
1775: \ee
1776:
1777: The structure theorem for graded antisymmetric matrices implies the following:
1778:
1779: \noindent (1) If ${\tilde \omega}=0$, then we can find an adapted basis and an ordering
1780: $a_1,\ldots, a_{2m},a_{2m+1},\ldots, a_N$ of the arrows ($N={\rm Card} {\cal Q}_1$)
1781: such that $a_1,\ldots, a_{2m}$ are even, $a_{2m+1},\ldots, a_N$ are odd and:
1782: \be
1783: \nn
1784: \omega=\sum_{i=1}^m{(da_id a_{i+m})_c}+\frac{1}{2}\sum_{j=2m+1}^N{(da_j da_j)_c}~~.
1785: \ee
1786: Setting $p_i:=a_i$ and $q_i=a_{i+m}$ for $i=1\ldots m$ and $\xi_\alpha:=a_{2m+\alpha}$ for $\alpha=1\ldots N-2m$,
1787: we can write this in the form:
1788: \be
1789: \label{e_can}
1790: \omega=(dp_idq_i)_c+\frac{1}{2}(d\xi_\alpha d\xi_\alpha)_c
1791: \ee
1792: with even $p_i,q_i$ and odd $\xi_\alpha$. Hence $\omega_{p_iq_j}=-\omega_{q_jp_i}=\delta_{ij}$
1793: and $\omega_{\xi_\alpha\xi_\beta}=\delta_{\alpha\beta}$.
1794:
1795: \noindent (2) If ${\tilde \omega}=1$, then ${\rm Card} {\cal Q}_1=2m$ for some integer $m$ and
1796: we can find an adapted basis and an ordering
1797: $a_1,\ldots,a_{2m}$ of the arrows such that $a_1,\ldots, a_{m}$ are odd, $a_{m+1},\ldots, a_{2m}$ are even and:
1798: \be
1799: \nn
1800: \omega=\sum_{i=1}^m{(da_id a_{i+m})_c}~~.
1801: \ee
1802: Setting $p_i:=a_i$ and $q_i=a_{i+m}$ for $i=1\ldots m$, this becomes:
1803: \be
1804: \label{o_can}
1805: \omega=(dp_idq_i)_c
1806: \ee
1807: with even $p_i$ and odd $q_i$. Hence $\omega_{p_iq_j}=-\omega_{q_jp_i}=\delta_{ij}$.
1808:
1809: In general, one can set $a^*=\sum_{b\in {\cal Q}_1}\omega_{ab}b$,
1810: which brings $\omega$ to the form $\omega=\frac{1}{2}(dada^*)_c$.
1811: For an even $\omega$ in the canonical basis (\ref{e_can}), we have $p_i^*=q_i$, $q_i^*=-p_i$ and
1812: $\xi_\alpha^*=\xi_\alpha$. In this case, $*$ squares to minus the identity on the subspace spanned by $q_i, p_i$
1813: but to plus the identity on the subspace spanned by $\xi_\alpha$.
1814: For odd $\omega$ in the basis (\ref{o_can}), we have $p_i^*=q_i$ and $q_i^*=-p_i$, so
1815: $*$ squares to minus the identity on the entire subspace $A_1=V$.
1816:
1817: It is clear from the above that a D-brane system has different behavior
1818: depending on the parity of $\omega$. We say that the system
1819: is {\em even} or {\em odd} if ${\tilde \omega}=0$, respectively $1$.
1820:
1821:
1822: \subsection{Quiver description of $C^0_R(A)$}
1823: \label{quivc0}
1824:
1825: Any element $f\in A$ has an expansion:
1826: \be
1827: \label{f_path}
1828: f=\sum_{p=path}{f_p p}
1829: \ee
1830: where the sum is over all paths $p$ of ${\cal Q}$ (including the
1831: trivial paths) and where $f_p\in \C$. In this and subsequent
1832: relations, we agree that only a finite number of coefficients
1833: are nonzero, so that all sums are finite. We can also write (\ref{f_path}) as:
1834: \be
1835: \label{f}
1836: f=\sum_{n\geq 0}{f_{a_1\ldots a_n}a_1\ldots a_n}~~,
1837: \ee
1838: where we use implicit summation over the arrows $a_j$ and we agree
1839: that $f_{a_1\ldots a_0}a_1\ldots a_0$ stands for the sum
1840: $c(f):=\sum_{u\in {\cal Q}_0}{f_u\epsilon_u}$ (with $f_u\in \C$).
1841: The product $a_1\ldots a_n$ vanishes unless it is a path. This is seen by
1842: inserting idempotents:
1843: \be
1844: \nn
1845: a_1\ldots a_ia_{i+1}\ldots a_n=a_1\ldots a_i\epsilon_{h(a_i)}\epsilon_{t(a_{i+1})}a_{i+1}\ldots a_n~~,
1846: \ee
1847: and noticing that the right hand side vanishes unless
1848: $h(a_i)=t(a_{i+1})$. Thus only the coefficients $f_{a_1\ldots a_n}$
1849: which correspond to paths $a_1\ldots a_n$ are defined; for convenience, we
1850: define $f_{a_1\ldots a_n}$ to vanish if the word $a_1\ldots a_n$ is not a path.
1851:
1852: The obvious relations
1853: \be
1854: \label{lindep}
1855: ({a_1}\ldots {a_n})_c=(-1)^{({\tilde a}_1+\ldots +
1856: {\tilde a}_i)({\tilde a}_{i+1}+\ldots +{\tilde a}_n)}({a_{i+1}}\ldots
1857: {a_n}{a_1}\ldots {a_i})_c~~.
1858: \ee
1859: show that $({a_1}\ldots {a_n})_c$ vanishes unless $a_1\ldots
1860: a_n$ is a cycle of ${\cal Q}$ (as we will see below, it can still
1861: vanish even for a cycle). For a general element (\ref{f}), relations (\ref{lindep}) give:
1862: \be
1863: \nn
1864: f_c:=\pi(f)=\sum_{n\geq 0}{f_{a_1\ldots a_n}({a_1}\ldots
1865: {a_n}})_c=\sum_{n\geq 0}{f_{(a_1\ldots a_n)}({a_1}\ldots
1866: {a_n})_c}~~,
1867: \ee
1868: where we introduced the `cyclicized coefficients':
1869: \be
1870: \nn
1871: f_{(a_1\ldots a_n)}:=\frac{1}{n}\sum_{i=0}^{n-1}{(-1)^{({\tilde a}_1+\ldots +{\tilde
1872: a}_i)({\tilde a}_{i+1}+\ldots +{\tilde a}_n)}f_{a_{i+1}\ldots
1873: a_n, a_1\ldots a_i}}~~,
1874: \ee
1875: which satisfy $f_{(a_1\ldots a_n)}=(-1)^{({\tilde a}_1+\ldots +
1876: {\tilde a}_i)({\tilde a}_{i+1}+\ldots +{\tilde a}_n)} f_{(a_{i+1}\ldots
1877: a_n, a_1\ldots a_i)}$.
1878: For $n=0$, we set $f_{(u)}=f_{u}$ for all $u\in {\cal Q}_0$.
1879:
1880: The observations made above show that any element $f\in C^0_R(A)$ can be expanded as:
1881: \be
1882: \label{fc_expansion}
1883: f=\sum_{n\geq 0}{f_{a_1\ldots a_n}(a_1\ldots a_n)_c}
1884: \ee
1885: where the coefficients are taken to be graded cyclic:
1886: \be
1887: \nn
1888: f_{a_1\ldots a_n}=(-1)^{({\tilde a}_1+\ldots +
1889: {\tilde a}_i)({\tilde a}_{i+1}+\ldots +{\tilde a}_n)} f_{a_{i+1}\ldots
1890: a_n, a_1\ldots a_i}~~
1891: \ee
1892: and the term $n=0$ in the sum stands for $\sum_{u\in {\cal Q}_0}{f_u\epsilon_u}\in R$.
1893: We also define the {\em strict coefficients} of $f\in C^0_R(A)$ by:
1894: \be
1895: \nn
1896: {\bar f}_{a_1\ldots a_n}:=nf_{a_1\ldots a_n}~~{\rm if}~~ n\neq 0~~
1897: \ee
1898: and ${\bar f}_{u}:=f_u$ for $n=0$. Then the expansion of $f_c$ becomes:
1899: \be
1900: \nn
1901: f_c=c(f)+\sum_{n\geq 1}{\frac{{\bar f}_{a_1\ldots a_n}}{n}({a_1}\ldots {a_n})_c}~~.
1902: \ee
1903:
1904: Consider the set ${\cal C}({\cal Q})$ of cycles of ${\cal Q}$. We say
1905: that two cycles $\gamma_1$ and $\gamma_2$ are {\em equivalent}, and
1906: write $\gamma_1\sim \gamma_2$, if they have the same length and differ
1907: by a cyclic permutation of their arrows (i.e. they differ only in the
1908: choice of the initial=terminal point of the cycle). This is an
1909: equivalence relation on ${\cal C}({\cal Q})$, whose equivalence classes are
1910: known as {\em necklaces}. We let $N({\cal Q}):={\cal C}({\cal Q})/\sim$ denote the
1911: set of necklaces, and write $[\gamma]$ for the equivalence class of a
1912: cycle $\gamma$. We define the length $l([\gamma])$ to be the length of
1913: any representative cycle. The $\Z_2$ degree of a necklace is the
1914: degree of any of its representatives. This gives a well-defined map
1915: from $N({\cal Q})$ to $\Z_2$.
1916:
1917: A cycle $\gamma$ is called primitive if it cannot be
1918: written in the form $\gamma=u^k$ with $u$ a non-trivial cycle and
1919: $k\geq 2$ (with this definition, the trivial paths $\epsilon_u$ are
1920: primitive). Any non-trivial cycle $\gamma$ can be written uniquely in
1921: the form $\gamma=r(\gamma)^{p(\gamma)}$ where $p(\gamma)\in \N^*$ and
1922: $r(\gamma)$ is a primitive cycle. This representation is called the
1923: primitive decomposition of $\gamma$. The integer $p(\gamma)$ is
1924: called the period, while the path $r(\gamma)$ is called the primitive
1925: root of $\gamma$. Given a necklace $\nu$ and representatives
1926: $\gamma_1,\gamma_2\in \nu$, we have $p(\gamma_1)=p(\gamma_2)$ and
1927: $r(\gamma_1)\sim r(\gamma_2)$. This allows us to define the period and
1928: primitive root of necklaces through $p([\gamma])=p(\gamma)$ and
1929: $r([\gamma])=[r(\gamma)]$.
1930:
1931: \paragraph{\bf Definition}
1932: A {\em null necklace} is a necklace $\nu$ such that $p(\nu)$ is
1933: even and $\deg~r(\nu)=1\in \Z_2$.
1934:
1935: \
1936:
1937: \paragraph{\bf Proposition}
1938: Let $\gamma$ be a cycle of the quiver ${\cal Q}$. The vector
1939: $\gamma_c=\pi(\gamma)\in C^0_R(A)$ vanishes if and only if the necklace
1940: $\nu:=[\gamma]$ is null.
1941:
1942: \
1943:
1944: \noindent We let $N_\bullet({\cal Q}):=\{ \nu\in N({\cal Q}) |\nu~{\rm~is~{\rm
1945: not}~null} \}$ be the set of non-null necklaces. Notice that the
1946: trivial paths $\epsilon_u$ are not null, and thus belong to $N_\bullet({\cal Q})$.
1947: \begin{proof}
1948:
1949: Let $\gamma=r^p$ be the primitive decomposition of $\gamma$. Commuting
1950: one copy of $r$ to the right gives:
1951: \be
1952: \gamma_c=\pi(r^p)=(-1)^{(p-1){\tilde r}}\gamma_c,\nn
1953: \ee
1954: where we noticed that $\deg (r)\deg(r^{p-1})={\tilde r}(p-1)$ in
1955: $\Z_2$. Thus $\gamma_c=0$ if $p$ is even and ${\tilde r}$ is odd. The
1956: converse follows from the description of $\pi$ given in Section
1957: \ref{tensor} and the fact that $r$ is the period of $\gamma$.
1958: \end{proof}
1959:
1960: \noindent
1961:
1962: Consider a necklace $\nu=[a_1\ldots
1963: a_n]$. Relations (\ref{lindep}) show that the subspace $V_\nu:=\C({a_1}\ldots
1964: {a_n})_c\subset C^0_R(A)$ depends only on $\nu$, while the proposition implies that $V_\nu=0$ if $\nu$ is null
1965: and $V_\nu\approx \C$ otherwise. This gives a direct sum
1966: decomposition: \be C^0_R(A)=\oplus_{\nu \in N_\bullet({\cal Q})}{V_\nu}~~,\nn
1967: \ee where $V_{u}:=\C\epsilon_u$ for the null paths $\epsilon_u$.
1968: Accordingly, any element $f\in C^0_R(A)$ decomposes as: \be
1969: f=\sum_{\nu \in N_\bullet({\cal Q})}{f(\nu)}~~,\nn \ee where all but a finite number
1970: of the vectors $f(\nu)\in V_\nu$ vanish. If $l(\nu)=n>0$, then: \be
1971: f(\nu)=\sum_{a_1a_2\dots a_n\in \nu}f_{a_1a_2\ldots
1972: a_n}({a_1}\ldots {a_n})_c~~,\nn \ee where the coefficients are
1973: defined as in (\ref{fc_expansion}). Since $f_{a_1\ldots a_n}$ are
1974: cyclic, all vectors $f_{a_1\ldots a_n}({a_1}\ldots {a_n})_c\in
1975: C^0_R(A)$ for which the cycle $a_1\ldots a_n$ belongs to a given
1976: necklace $\nu$ are equal. Thus: \be\nn f(\nu)={\bar f}_{a_1\ldots
1977: a_n}({a_1}\ldots {a_n})_c~~~{\rm for~any~fixed~cycle~}~~a_1\dots a_n
1978: \in \nu~~. \ee (in this relation, no summation over $a_j$ is
1979: implied).
1980:
1981: In view of these observations, the coefficients in the expansion
1982: (\ref{fc_expansion}) associated to null necklaces are not defined.
1983: For simplicity, we will define all such coefficients to be zero. Thus
1984: all coefficients associated with words $a_1 \ldots a_n$ which fail to
1985: correspond to a cycle or correspond to a cycle in a null necklace are
1986: set to zero by definition. We will use this convention repeatedly in
1987: what follows.
1988:
1989:
1990: \paragraph{Observation}
1991: $C^0_R(A)$ can be identified with the vector space $\C^{N_{\cal Q}}=\oplus_{\nu
1992: \in N_{\cal Q}}\C\nu $ generated by the set of non-null necklaces $N_{\cal Q}$. For
1993: this, pick an enumeration of ${\cal Q}_1$, let $c_\nu:=\min \nu$ be the
1994: minimal representative of $\nu$ with respect to the induced
1995: lexicographic order on the set of paths of ${\cal Q}$, and identify $V_\nu$ with $\C$
1996: by sending $\pi(c_\nu)$ into the complex unit. With this
1997: identification, we have $f(\nu)\equiv {\bar
1998: f}_{c_\nu}$ and $f\equiv \sum_{\nu \in N_\bullet({\cal Q})} {\bar f}_{c_\nu}
1999: \nu$. Notice that such an identification requires that we pick an enumeration of
2000: ${\cal Q}_1$.
2001:
2002:
2003: \subsection{Cyclic derivatives and loop partial derivatives}
2004: \label{cycloop}
2005:
2006: The isomorphism $\Omega^1_RA\approx A\otimes V$ of Section \ref{tensor}
2007: shows that any one-form $w\in \Omega_R^1 A$ has well-defined coefficients
2008: $w_a\in A$ determined by:
2009: \be
2010: \nn
2011: w=\sum_{a\in {\cal Q}_1}{(w_a d a)_c}=\sum_{a\in {\cal Q}_1}{ (-1)^{{\tilde a}({\tilde w}+1)}((d a) w_a)_c}~~,
2012: \ee
2013: where we noticed that ${\tilde w}_a={\tilde w}+{\tilde a}$. Inserting
2014: idempotents shows that each $w_a\in A$ must be a linear combination of
2015: paths starting at $h(a)$ and ending at $t(a)$.
2016:
2017: We define the {\em left and right quiver cyclic derivatives} $\ld_a f, f\rd_a\in A$ of
2018: an element $f\in C^0_R(A)$ through:
2019: \be
2020: \label{df}
2021: {\bar d} f=\sum_{a} ((f\rd_a)d a)_c =\sum_{a} (d a (\ld_a f))_c~~.
2022: \ee
2023: This gives linear maps $\ld_a, \rd_a: C^0_R(A)\rightarrow A$, satisfying $f\rd_a=
2024: (-1)^{{\tilde a}({\tilde f}+1)}\ld_a f$.
2025: Equation (\ref{df}) gives:
2026: \begin{eqnarray}
2027: \ld_a f&=&\sum_{n} {\bar f}_{a a_1\ldots a_n}a_1\ldots a_n ~~\nn\\
2028: f\rd_a &=&\sum_{n} {\bar f}_{a_1\ldots a_n a }a_1\ldots a_n~~.\nn
2029: \end{eqnarray}
2030: Notice that our convention (\ref{pairing}) is crucial for these simple formulas.
2031: In particular, for any cycle $\gamma=a_1\ldots a_n$ of ${\cal Q}$:
2032: \begin{eqnarray}
2033: \label{cd_props_general}
2034: \ld_a({a_1}\ldots {a_n})_c &=&\sum_{i=1}^n{(-1)^{({\tilde a}_1+\ldots
2035: +{\tilde a}_{i-1})({\tilde a}+{\tilde
2036: a}_{i+1}+\ldots +{\tilde a}_{n})}\delta_{a,a_i}{a_{i+1}}\ldots
2037: {a_n} {a_1}\ldots {a_{i-1}}}\\
2038: ({a_1}\ldots {a_n})_c\rd_a &=&\sum_{i=1}^n{(-1)^{({\tilde a}+{\tilde a}_1+\ldots +{\tilde a}_{i-1})({\tilde
2039: a}_{i+1}+\ldots +{\tilde a}_{n})}\delta_{a,a_i}{a_{i+1}}\ldots
2040: {a_n} {a_1}\ldots {a_{i-1}}}~~.\nn
2041: \end{eqnarray}
2042: These equations imply:
2043: \begin{eqnarray}
2044: \label{cd_proj}
2045: \pi(\ld_a({a_1}\ldots {a_n})_c) &=&\sum_{i=1}^n(-1)^{{\tilde a}({\tilde a}_1+\ldots
2046: +{\tilde a}_{i-1})}\delta_{a,a_i}(a_1\ldots a_{i-1}a_{i+1}\ldots a_n)_c\\
2047: \pi(({a_1}\ldots {a_n})_c \rd_a) &=&\sum_{i=1}^n (-1)^{{\tilde a}({\tilde a}_{i+1}+\ldots
2048: +{\tilde a}_n)}\delta_{a,a_i}(a_1\ldots a_{i-1}a_{i+1}\ldots a_n)_c~~.\nn
2049: \end{eqnarray}
2050: It is clear that all terms in right hand side vanish
2051: unless $a_i$ is a loop. Hence the
2052: linear operators $\pi\circ \ld_a$ and $\pi\circ \rd_a$ induced
2053: on $C^0_R(A)$ are non-trivial only when $a$ is a loop.
2054:
2055:
2056: We also define {\em loop partial derivatives} $\lp_a,\rp_a$ if $a$ is
2057: a (non-trivial) {\em loop} of the quiver. Given such a loop of
2058: ${\cal Q}$, the loop derivatives $\lp_a\in \Der_l(A)$ and $\rp_a\in
2059: \Der_r(A)$ are the unique $R$-linear left and right derivations of
2060: $A$ of degree ${\tilde a}$ such that:
2061: \be
2062: \label{lp_def}
2063: \lp_a b:=b\rp_a=\delta^b_a \epsilon_u ~~{\rm~for~all}~b\in {\cal Q}_1~~,
2064: \ee
2065: where $u=h(a)=t(a)$.
2066: It is not hard to see that $\lp_a$ and $\rp_a$ are well-defined and $\lp_a f=(-1)^{{\tilde a}({\tilde f}+1)}f\rp_a$.
2067: It is clear from (\ref{lp_def}) that loop partial derivatives supercommute:
2068: \be
2069: \nn
2070: [\lp_a,\lp_b]=[\rp_a,\rp_b]=0~~{\rm for~all~loops}~~a,b\in {\cal Q}_1.
2071: \ee
2072:
2073: Let ${\cal Q}_1(u)$ be the subset of loops at the vertex $u$ and
2074: let $a^*=\sum_{b\in {\cal Q}_1}{\omega_{ab}b}=\sum_{b\in {\cal Q}_1(u)}{\omega_{ab}b} \in A$ be the
2075: conjugate element of the loop $a$ introduced in Subsection \ref{symplectic_structure}.
2076: A simple computation gives the relation:
2077: \be
2078: \nn
2079: {\bar i}_{\lp_a}\omega={\bar d}\pi(a^*)~~,
2080: \ee
2081: which shows that $\lp_a$ is the symplectic relative left derivation
2082: with Hamiltonian $\pi(a^*)\in C^0_R(A)$. Similarly, one can
2083: view $\rp_a$ as the symplectic relative right derivation with Hamiltonian
2084: $\pi(a^*)$. Notice that the necklace of a loop $a$ has only one
2085: representative, so one can identify loops with their projections
2086: through $\pi$; we will sometimes do so in what follows.
2087:
2088: Let $\partial^r_a$ and $\partial_a^l$ be the complex-linear maps
2089: induced by $\lp_a$ and $\rp_a$ on $C_R^0(A)=A/[A,A]$. For any
2090: cycle $\gamma=a_1\ldots a_n$, we have:
2091: \begin{eqnarray}
2092: \partial^l_a\gamma_c&=&\pi(\lp_a \gamma)=\sum_{i=1}^n(-1)^{{\tilde a}({\tilde a}_1+\ldots
2093: +{\tilde a}_{i-1})}\delta_{a,a_i}(a_1\ldots a_{i-1}a_{i+1}\ldots a_n)_c~~\nn\\
2094: \gamma_c\partial^r_a &=&\pi(\gamma\rp_a)=\sum_{i=1}^n(-1)^{{\tilde a}({\tilde
2095: a}_{i+1}+\ldots +{\tilde a}_{n})}\delta_{a,a_i}(a_1\ldots a_{i-1}a_{i+1}\ldots a_n)_c~~.\nn
2096: \end{eqnarray}
2097: Comparing with (\ref{cd_proj}) gives:
2098: \be
2099: \nn
2100: \partial^l_a=\pi\circ \ld_a~~,~~\partial^r_a=\pi\circ \rd_a~~.
2101: \ee
2102: Hence the cyclic and loop partial derivatives induce the same complex-linear operators on $C^0_R(A)$.
2103:
2104: Given an element $x\in E[1]$, we expand $x=\sum_{a\in {\cal Q}_1}{x^a e_a}$ with
2105: $x^a\in \C$ and define the (relative) left cyclic derivative along $x$
2106: via:
2107: \be
2108: \nn
2109: \ld_x f:=\sum_{a\in {\cal Q}_1}{x^a\ld_a f}~~~~\forall f\in C^0_R(A)~~.
2110: \ee
2111: If $x$ is central in $E$, then $x^a$ vanish unless $a$ is a loop. In
2112: this case, we define the (relative) loop left partial derivative along
2113: $x$ via:
2114: \be
2115: \nn
2116: \lp_x f=\sum_{a=loop}{x^a\lp_a f}~~.
2117: \ee
2118: For a central element, these two notions induce the same map on $C^0_R(A)$:
2119: \be
2120: \nn
2121: \partial^l_x f=\sum_{a=loop}{x^a\partial^l_a f}~~.
2122: \ee
2123: These definitions are well-behaved with respect to changes of adapted bases.
2124: It is clear that $\partial_{e_a}^l=\partial_a^l$ etc.
2125:
2126: \paragraph{Observations}
2127:
2128: (1) Loop partial derivatives and quiver cyclic derivatives are related to certain double
2129: derivations introduced in \cite{Bergh}. Consider
2130: the $A$-superbimodule $A\otimes_\C A$, where we use the so-called outer superbimodule structure:
2131: \be
2132: \nn
2133: \alpha (a\otimes b)\beta:=(\alpha a)\otimes (b\beta)~~~~\forall \alpha,\beta, a, b \in A~~.
2134: \ee
2135: Since $R$ is a subalgebra of $A$, this is also an $R$-superbimodule. A {\rm relative double derivation} of $A$
2136: is an $R$-linear derivation of the $A$-superbimodule $A\otimes A$.
2137: As in \cite{Bergh}, consider the double left derivations determined by:
2138: \be
2139: \nn
2140: {\bf D}_a(b)=\epsilon_{t(a)}\otimes \epsilon_{h(a)}\delta^b_a~~~~~\forall a,b\in {\cal Q}_1~~.
2141: \ee
2142: Then the loop partial derivatives can be recovered as:
2143: \be
2144: \nn
2145: \lp_a={\bf m}\circ {\bf D}_a~~{\rm for~~}a{\rm ~~a~loop}
2146: \ee
2147: where ${\bf m}:A\otimes A\rightarrow A$ is the $A$-superbimodule morphism:
2148: \be
2149: \nn
2150: {\bf m}(a\otimes b)=ab~~.
2151: \ee
2152: Similarly, relations (\ref{cd_props_general}) show that the cyclic
2153: derivatives are induced by the maps $m\circ {\bf D}_a$, where
2154: $m:A\otimes A\rightarrow A$ is the linear map ({\em not} a bimodule
2155: morphism !): \be \nn m(a\otimes b)=(-1)^{{\tilde a}{\tilde b}}ba~~.
2156: \ee Indeed, it is not hard to see that $m\circ {\bf D}_a$ vanishes on
2157: $[A,A]$, so it induces a map from $C^0_R(A)$ to $A$ which coincides
2158: with the cyclic derivative $\ld_a$.
2159:
2160: (2) Let ${\cal Q}$ be a quiver with a single vertex $u$. Then all
2161: arrows are loops and $A$ is the free superalgebra $\C\langle
2162: \{a\}\rangle$ generated by these loops. In this case, $\lp_a$ and
2163: $\rp_a$ reduce to the standard left and right partial derivatives of
2164: the free algebra $A$. Moreover, the observations above show that
2165: $\ld_a$ are a superized version of the objects considered in
2166: \cite{RSS, Voiculescu}. This justifies our terminology.
2167:
2168:
2169: \subsection{Description of one-forms and closed two-forms}
2170: \label{12forms}
2171:
2172: Consider the reduced tensor algebra $A_{\geq 1}:=V\otimes_R
2173: A=T_R^{\geq 1} V=\oplus_{n\geq 1}{V^{\otimes_R n}}$, which coincides
2174: with the subspace of $A$ spanned by all paths of length at least one. As in \cite{Ginzburg,LB_quivers}
2175: (see also \cite{Lazarev}), it is easy to see that the super-vector space $C^1_R(A)$ is isomorphic
2176: with the center $A_{\geq 1}^R=(V\otimes_R A)^R$, the space spanned by the non-trivial cycles of the quiver.
2177: The isomorphism $\Xi:A_{\geq 1}^R\rightarrow C^1_R(A)$ has the form:
2178: \be
2179: \label{Xi}
2180: \sum_{i}{x_i\otimes_R f_i}\stackrel{\Xi}{\rightarrow} \sum_{i}{(dx_i f_i)_c}\in C^1_R(A)~~.
2181: \ee
2182: This follows from the observation that any element of $C^1_R(A)$ can be
2183: written uniquely as:
2184: \be
2185: \nn
2186: w=\sum_{n\geq 0}f_{a a_1\ldots a_n}(da a_1\ldots a_n )_c=\sum_{a}(da f_a)_c=\Xi(\sum_a af_a)~~,
2187: \ee
2188: where $f_a:=\sum_{n\geq 0}f_{a a_1\ldots a_n }a_1\ldots a_n$ and
2189: the complex coefficients $f_{a_1\ldots a_n}$ are taken to vanish unless $a_1\ldots a_n$ is
2190: a cycle. Then $\Xi^{-1}(w)=\sum_{n\geq 0} f_{a_1\ldots a_n}a_1\ldots a_n:=f=\sum_a af_a$.
2191: For $g=\sum_{n\geq 0}g_{a_1\ldots a_n} a_1\ldots a_n \in A^R_{\geq 1}$, we have
2192: ${\bar d}\pi(g)={\bar d}g_c=(da\ld_a g_c)_c=
2193: \sum_{n\geq 0} {\bar g}_{(a a_1\ldots a_n)}(da a_1\ldots a_n)_c$, where
2194: ${\bar g}_{(a_1\ldots a_n)}:=n g_{(a_1\ldots a_n)}$ are the strict coefficients of $g_c$.
2195: Thus $w={\bar d}(g)_c$ iff
2196: $f_{a_1\ldots a_n}={\bar g}_{(a_1\ldots a_n)}$. It is clear from these observations that
2197: $w$ is exact iff its coefficients $f_{a_1\ldots a_n}$ are graded-cyclic; in this case, we
2198: can take $g=\frac{1}{n} f_{a_1\ldots a_n}a_1\ldots a_n$ and we have $f_a=\ld_a g$.
2199:
2200: Thus $\Xi(f)\in C^1_R(A)_{\rm closed}= C^1_R(A)_{\rm exact}$. We let $A_{\geq 1}^c$ be the {\em
2201: cyclic subspace} of $A_{\geq 1}^R$, i.e. the subspace
2202: consisting of elements with graded-cyclic coefficients. The space $A_{\geq 1}^c$
2203: consists of linear combinations of non-trivial cycles, such
2204: that cycles belonging to the same necklace of length $n$ appear with
2205: coefficients related by the action of $\Z_n$. This subspace provides
2206: an embedding of $C^0_R(A)/R$ into $A_{\geq 1}$. In fact, the projection
2207: $\pi:A\rightarrow C^0_R(A)$ induces an isomorphism $A_{\geq
2208: 1}^c\approx C^0_R(A)/R=A_{\geq 1}/[A,A]$ (recall that
2209: $[A,A]_0=[R,R]=0$ so $[A,A]\subset A_{\geq 1}$). Thus we have a vector
2210: space decomposition $A_{\geq 1}=[A,A]\oplus A_{\geq 1}^c$, as well as
2211: $A_{\geq 1}^R=[A,A]^R\oplus A_{\geq 1}^c$. In particular, the subspace $A_{\geq 1}^c$ gives a
2212: natural complement of $[A,A]^R$ inside $A_{\geq 1}^R$.
2213:
2214:
2215: Since the Karoubi complex is acyclic in positive degrees, we have $C^2_R (A)_{\rm closed}=
2216: {\bar d}(C^1_R (A))$ and the isomorphism $\Xi$ shows that any closed
2217: two-form can be written as:
2218: \be
2219: \label{dXi}
2220: u=-\sum_{n\geq 0}{(da df_a)_c}={\bar d}\Xi(f)~~.
2221: \ee
2222: where $f=af_a\in A_{\geq 1}^R$ is a combination of cycles of
2223: length at least one. Since $\Xi(A_{\geq 1}^c)=C^1_R(A)_{\rm closed}$, the two-form (\ref{dXi})
2224: vanishes precisely when $f$ belongs to the cyclic subspace $A_{\geq 1}^c$. Thus $\ker(d\Xi)=A_{\geq 1}^c$,
2225: and the map $d\Xi$ induces an isomorphism:
2226: \be
2227: \label{kappa_gen}
2228: \kappa:[A,A]^R\stackrel{\approx}{\rightarrow} C^2_R (A)_{\rm closed}
2229: \ee
2230: between the complement $[A,A]^R$ of this subspace in $A_{\geq 1}^R$ and the space
2231: of closed two forms\footnote{This result is also discussed in \cite{VG}. I thank V. Ginzburg for pointing this
2232: out.}.
2233:
2234:
2235:
2236:
2237: \subsection{Quiver description of the Kontsevich bracket}
2238:
2239: Consider the non-commutative symplectic form (\ref{omega_form}) on $A$.
2240: For $\theta\in \Der_l(A)$, we set $\theta(a):=\theta^a\in A$. $R$-linearity of $\theta$ implies
2241: that $\theta^a$ is a linear combination of paths which start at $t(a)$ and end at $h(a)$.
2242: Equation (\ref{df}) gives:
2243: \be
2244: \nn
2245: {\bar L}_\theta f={\bar i}_\theta {\bar d} f=
2246: \pi(i_\theta\sum_{a\in {\cal Q}_1} (da) \ld_a f)=\sum_{a\in {\cal Q}_1}{(\theta(a )\ld_a f)_c}~~.
2247: \ee
2248: Thus:
2249: \be
2250: \label{dftheta}
2251: {\bar L}_\theta f=\sum_{a\in {\cal Q}_1}(\theta^a \ld_a f)_c~~~~\forall \theta\in \Der_l(A)~~.
2252: \ee
2253: If $\theta$ is homogeneous of degree ${\tilde \theta}$, we set:
2254: \be
2255: \theta_a:=\theta^b\omega_{ba}~~.\nn
2256: \ee
2257: Notice that $\tilde{\theta_a}={\tilde a}+{\tilde \omega}+{\tilde \theta}$. Expanding
2258: $\theta^a=\sum_{n\geq 0}\theta_{a_1\ldots a_n}\,\,^a {a_1}\ldots {a_n}$ with $\theta_{a_1\ldots a_n}\,\,^a\in \C$,
2259: we find $\theta_a=\sum_{n\geq 0}\theta_{a_1\ldots a_n a}{a_1}\ldots {a_n}$, where:
2260: \be
2261: \nn
2262: \theta_{a_1\ldots a_na}=\theta_{a_1\ldots a_n}\,\,^b\omega_{ba}~~.
2263: \ee
2264: As usual, $\theta_{a_1\ldots a_n}\,\,^a$ are taken to vanish unless $a_1\ldots a_n$ is a path. Also notice
2265: that $\theta_{a_1\ldots a_n}\,\,^a$ vanishes automatically unless this path starts at $t(a)$ and ends at $h(a)$.
2266: Similarly, $\theta_{a_1\ldots a_n}$ vanishes unless $a_1\ldots a_n$ is a cycle of ${\cal Q}$.
2267:
2268: An easy computation gives:
2269: \be
2270: \label{hash}
2271: {\bar i}_\theta \omega=(\theta_a da)_c=\sum_{n\geq 0}\theta_{a_1\ldots a_n a}({a_1}\ldots {a_n}da)_c~~.
2272: \ee
2273: Given $f\in C^0_R(A)$, we have ${\bar d}f=(f\rd_a
2274: da)_c$. Comparing with (\ref{hash}) gives $(\theta_f)_a=f\rd_a$, where
2275: $\theta_f$ is the Hamiltonian vector field of $f$. Hence the map
2276: $\psi_\omega:C^0_R(A)\rightarrow \Der_l(A)$ of (\ref{ex}) is
2277: given by:
2278: \be
2279: \label{phiomega}
2280: \theta_f(a):=\theta_f^a=f\rd_b\omega^{ba}=\sum_{n\geq 0}{\bar f}_{a_1\ldots a_n b}\omega^{ba} s^{a_1}\ldots s^{a_n}~~,
2281: \ee
2282: where ${\bar f}_{a_1\ldots a_n}$ are the strict
2283: coefficients of $f$. This allows us to write the Kontsevich bracket in more familiar form.
2284:
2285: \paragraph{\bf Proposition}
2286: \label{coord_bracket}
2287: We have $\{f,g\}= (f\rd_a\omega^{ab}\ld_b g)_c$ for all $f,g\in
2288: C^0_R(A)$.
2289:
2290: \
2291:
2292: \begin{proof}
2293:
2294: Using (\ref{bL}), (\ref{dftheta}) and (\ref{phiomega}), we compute
2295: $\{f,g\}={\bar L}_{\theta_f}(g)=(\theta_f^a\ld_a
2296: g)_c=(f\rd_b\omega^{ba}\ld_a g)_c$.
2297:
2298: \end{proof}
2299:
2300:
2301: \subsection{Some canonical forms and coefficient expressions}
2302: \label{canforms}
2303:
2304: In this subsection, we give some expressions which are useful in applications.
2305: Let $W\in C^0_R(A)$ be an element of degree ${\tilde \omega}+1$. As we will see in
2306: the next section, the boundary generating function of a topological D-brane system
2307: is such an element.
2308:
2309: For ${\tilde \omega}=0$, let us choose an adapted basis as in (\ref{e_can}).
2310: Then $W$ is odd, and one finds:
2311: \be
2312: \nn
2313: \{W, W\}=(W\rd_{p_i}\ld_{q_i}W-W\rd_{q_i}\ld_{p_i}W)_c+(W\rd_{\xi_\alpha}\ld_{\xi_\alpha}W)_c=
2314: 2(W\rd_{p_i}\ld_{q_i}W)_c+(W\rd_{\xi_\alpha}\ld_{\xi_\alpha}W)_c
2315: \ee
2316: since $(W\rd_{p_i}\ld_{q_i}W)_c=-(W\rd_{q_i}\ld_{p_i}W)_c$. Also notice that $W\rd_{\xi_\alpha}=\ld_{\xi_\alpha} W$.
2317:
2318: Now let ${\tilde \omega}=1$ and choose an adapted basis as in (\ref{o_can}).
2319: Then $W$ is even and we have:
2320: \be
2321: \nn
2322: \{W, W\}=(W\rd_{p_i}\ld_{q_i}W-W\rd_{q_i}\ld_{p_i}W)_c=2(W\rd_{p_i}\ld_{q_i}W)_c~~,
2323: \ee
2324: since again $(W\rd_{p_i}\ld_{q_i}W)_c=-(W\rd_{q_i}\ld_{p_i}W)_c$.
2325:
2326: One can also extract the coefficient expression of the cyclic bracket by
2327: direct computation. The case relevant for us is as follows. For $W$ as above, notice that $\rho^{ab}$ can be used to
2328: raise and lower indices `from the left':
2329: \begin{eqnarray}
2330: \label{lifts}
2331: {\bar W}_{a_1\ldots a_{i-1}} {\,\, }^a_{\,\,a_{i+1}\ldots
2332: a_{n}}&:=&\rho^{ab}{\bar W}_{a_1\ldots a_{i-1}b a_{i+1}\ldots
2333: a_n}~~.\nn\\ {\bar W}_{a_1\ldots a_{i-1}b a_{i+1}\ldots
2334: a_n}&:=&\rho_{ab}{\bar W}_{a_1\ldots a_{i-1}} {\,\, }^b_{\,\,
2335: a_{i+1}\ldots a_{n}}\nn~~.
2336: \end{eqnarray}
2337: Then it is shown in Appendix \ref{coeffs} that that bracket of $W$ with itself takes the form:
2338: \be
2339: \frac{1}{2}\{W,W\}=\frac{1}{2}{\bar W}_a{\bar W}^a+
2340: \sum_{n\geq 1}\frac{1}{n}\left(\sum_{0\leq i+j\leq n}(-1)^{{\tilde a_1}+\ldots +{\tilde a}_i}
2341: {\bar W}_{a_1\ldots a_i a a_{i+j+1}\ldots a_n} {\bar W}^a_{\,\,a_{i+1}\ldots
2342: a_{i+j}}\right)({a_1}\ldots {a_n})_c~~,\nn
2343: \ee
2344: which is valid irrespective of the degree of $\omega$.
2345:
2346:
2347: \section{Geometry of finite D-brane systems}
2348: \label{geometrization}
2349:
2350: Consider a finite topological D-brane system with total boundary space
2351: $E$ and boundary algebra $R$. As before, we let $V=E[1]^{\rm v}$ and
2352: consider the tensor algebra $A=T_R V$. As explained in Section
2353: \ref{algebraic}, the data of all integrated boundary correlators on
2354: the disk is encoded by an $R$-superbimodule structure on $E$, together with
2355: a cyclic and unital weak $A_\infty$ structure on this
2356: superbimodule. We will use the machinery developed in the previous two
2357: sections to encode this into a `noncommutative
2358: generating function' $W\in C^0_R(A)$ subject to simple constraints. To
2359: this end, we pick an adapted basis of $E$ and let ${\cal Q}$ be its
2360: index superquiver.
2361:
2362:
2363:
2364:
2365: \subsection{Geometric description of cyclic weak $A_\infty$ structures}
2366:
2367: It turns out that a weak $A_\infty$
2368: structure on the $R$-superbimodule $E$ is the same as an {\em odd}
2369: relative derivation $Q$ of the tensor algebra $A$.
2370: With our conventions, the relation is as follows. Picking adapted
2371: coordinates, we define the coefficients of $Q$ through:
2372: \be
2373: \label{Q_exp}
2374: Q(a)=\sum_{n\geq 0} Q_{a_1\ldots a_n}{\,}^a a_1\ldots a_n
2375: \ee
2376: and construct odd linear maps $r_n:E[1]^{\otimes n}\rightarrow E[1]$ via:
2377: \be
2378: \label{rQ}
2379: r(e_{a_1}\ldots e_{a_n})=Q_{a_1\ldots a_n}{\,}^a e_a~~.
2380: \ee
2381: Thus $Q(s^a)=\sum_{n\geq 0} s^a(r(e_{a_1}\ldots e_{a_n})) s^{a_1}\otimes_R \ldots \otimes_R s^{a_n}$.
2382: Thinking in terms of arrows, it is clear that $r_n$ are $R$-multilinear.
2383: Since $Q$ is odd, we have $[Q,Q]= 2 Q^2$, which implies that $Q^2$ is
2384: a derivation of $A$. Since $a$ generate the algebra, this means that the condition $Q^2=0$ is equivalent with
2385: $Q^2(a)=0$ for all $a$. Using expansion (\ref{Q_exp}), one finds
2386: that this amounts to the relations:
2387: \be
2388: \nn
2389: \sum_{0\leq i+j\leq n}(-1)^{{\tilde a_1}+\ldots +{\tilde a}_i}
2390: Q_{a_1\ldots a_i b a_{i+j+1}\ldots a_n}{\,}^a Q_{\,\,a_{i+1}\ldots
2391: a_{i+j}}{\,}^b=0~~{\rm for~all~}~~ n\geq 0~~,
2392: \ee
2393: which are the $A_\infty$ constraints (\ref{ainf}).
2394:
2395: It is also not hard to check that the nilpotent derivation $Q$ is
2396: symplectic iff the associated $A_\infty$ structure is cyclic. An easy
2397: way to see this is as follows. By the exact sequence (\ref{ex}), we
2398: have that $Q$ is symplectic iff it is Hamiltonian, which via equation
2399: (\ref{phiomega}) amounts to the existence of a $W\in C^0_R(A)$ such
2400: that:
2401: \be
2402: Q_{a_1\ldots a_n}{\,}^a={\bar W}_{a_1\ldots a_nb}\omega^{ba}~~
2403: \ee
2404: or, equivalently:
2405: \be
2406: \label{QW}
2407: Q_{a_1\ldots a_n}={\bar W}_{a_1\ldots a_n}~~.
2408: \ee
2409: As usual, we have set $Q_{a_1\ldots a_n a}:=Q_{a_1\ldots a_n}{\,}^b\omega_{ba}$.
2410: It is clear that a $W$ exists if and only if
2411: the coefficients $Q_{a_1\ldots a_n}$ are cyclic. We have:
2412: \begin{eqnarray}
2413: \label{cyc_arg}
2414: \rho(e_{a_0},r_n(e_{a_1}\ldots e_{a_n}))&=&(-1)^{{\tilde a}_0}\omega(e_{a_0},
2415: r_n(e_{a_1}\ldots e_{a_n}))=(-1)^{1+{\tilde a}_0{\tilde \omega}}
2416: \omega(r_n(e_{a_1}\ldots e_{a_n}), e_{a_0})\nn\\
2417: &=&(-1)^{{\tilde a}_0 {\tilde \omega}}{\bar W}_{a_1\ldots a_n a_0}={\bar W}_{a_0\ldots a_n}~~,
2418: \end{eqnarray}
2419: where we used the superselection rules for $\omega$ and $W$.
2420: Thus:
2421: \be
2422: \label{QW1}
2423: \rho(e_{a_0},r_n(e_{a_1}\ldots e_{a_n}))={\bar W}_{a_0a_1\ldots a_n}~~,
2424: \ee
2425: and we see that $L_Q \omega=0$ implies that the left hand side is
2426: cyclic, which is the cyclicity constraint (\ref{rrcyc}). Conversely,
2427: if the LHS is cyclic then we define $W$ through equation
2428: (\ref{QW1}). Then relations (\ref{cyc_arg}) show that $Q_{a_1\ldots
2429: a_n a_0}={\bar W}_{a_1\ldots a_n a_0}$, i.e. $Q$ is symplectic with
2430: Hamiltonian $W$. Combining everything and noticing that ${\tilde W}={\tilde \omega}+1$ (because
2431: $Q$ is odd), we have:
2432:
2433: \
2434:
2435: {\em Giving a cyclic weak $A_\infty$ structure on $E$ amounts to giving an element $W\in C^0_R(A)$,
2436: of degree ${\tilde \omega}+1$, such that $\{W,W\}=0$.}
2437:
2438: \
2439:
2440:
2441: \paragraph{Observation} The triplet $(A,Q,\omega_{form})$ can be viewed as a noncommutative generalization
2442: of the so-called $QP$-manifolds of \cite{Konts_Schwarz}, while the doublet $(A,Q)$ generalizes the concept of
2443: $Q$-manifold discussed in the same paper (notice, though, that we consider both even and odd symplectic forms,
2444: so we generalize the work of \cite{Konts_Schwarz} in two directions). It was shown in \cite{Konts_Schwarz} that
2445: $QP$-manifolds give the general geometric setting of the classical BV-formalism.
2446: Accordingly, for odd symplectic forms, the
2447: triplet $(A,Q,\omega_{form})$ defines a noncommutative version of that formalism.
2448:
2449:
2450:
2451:
2452: \subsection{The unitality constraint}
2453: \label{sec:unitality}
2454:
2455: We saw that a cyclic weak $A_\infty$ structure on $E$ is the same as
2456: an $R$-linear symplectic derivation $Q\in \Der_l^\omega(A)$ such that
2457: $Q^2=0$. We let $W$ be the {\em canonical} Hamiltonian of $Q$, i.e. that Hamiltonian which vanishes at zero
2458: (see Section \ref{tensor}). Explicitly, equations (\ref{QW}) and (\ref{QW1}) give:
2459: \be
2460: \nn
2461: W=\sum_{n\geq 0}{\frac{1}{n+1}\rho(e_{a_0}, r_n(e_{a_1}\ldots e_{a_n}))(s^{a_0}\ldots s^{a_n}})_c~~,
2462: \ee
2463: which allows us to reconstruct $r_n$ from $W$ provided that we know $\omega=\rho\circ \Sigma^{\otimes 2}$.
2464: The homological derivation $Q$ can be recovered as the Hamiltonian derivation defined by $W$,
2465: which amounts to relations (\ref{QW}).
2466:
2467: For a topological D-brane system, the underlying weak $A_\infty$ structure should be unital.
2468: To formulate this condition in terms of $W$, we write the unitality constraints (\ref{unitality}) as:
2469: \begin{eqnarray}
2470: \label{ueqs}
2471: r_n(e_{a_1}\ldots e_{a_{j-1}},\lambda,e_{a_{j+1}}\ldots e_{a_n})&=&0~~{\rm for~all}~~~~ n\neq 2 ~{\rm~and~all}~
2472: j=1\ldots n~~\nn\\
2473: -r_2(\lambda,e_{a})=(-1)^{{\tilde a}} r_2(e_{a},\lambda)&=&e_a~~,
2474: \end{eqnarray}
2475: where $\lambda$ is an odd central element of $E$.
2476: Given $\lambda=\oplus_{u\in {\cal Q}_0}{\lambda_u}\in E[1]^R$ with $\lambda_u\in E_{uu}[1]$, we
2477: can choose adapted coordinates such that each $\lambda_u$ is one of the odd basis elements $\{e_a\}$.
2478: We then let $\sigma_u\in E[1]^{\rm v}$ be the corresponding elements of the dual basis $\{s^a\}$ of $V$ (those
2479: dual basis elements which satisfy $s^{\sigma_u}(e_a)=\delta_{e_a,\lambda_u}$).
2480: It is clear that each $\sigma_u$ is an odd loop of the quiver starting and ending at the vertex $u$. With such
2481: a choice of adapted basis, we have $\lambda_u=e_{\sigma_u}$ and the first row in (\ref{ueqs}) is equivalent with:
2482: \be
2483: \label{un1}
2484: W_{a_1\ldots a_n}=0 {\rm ~if~}n\neq 3{\rm ~and~any~of~the~arrows~}
2485: a_j{\rm ~coincides~with~any~of~the~loops}~\sigma_u~~
2486: \ee
2487: while the second row amounts to:
2488: \be
2489: \label{un2}
2490: {\bar W}_{\sigma_u ab}=-\omega_{ab}\Longleftrightarrow W_{\sigma_u ab}=-\frac{1}{3}\omega_{ab}
2491: ~~~{\rm~for~}t(a)=h(b)=u~~{\rm~and~~}h(a)=t(b)~~.
2492: \ee
2493: Hence unitality of $(r_n)$ boils down to the requirement that
2494: the adapted basis $\{\psi_a\}$ can be chosen such that the vertices
2495: the index quiver carry distinguished odd loops $\sigma_u$ satisfying
2496: (\ref{un1}) and (\ref{un2}).
2497:
2498: The two conditions above say that $W$ takes the form:
2499: \be
2500: \label{W_decomp}
2501: W=W_g +W_d
2502: \ee
2503: where the `generic' contribution is given by:
2504: \be
2505: \label{Wg}
2506: W_g:=-\omega_{ab}(\sigma a b)_c=
2507: -\sum_{\tiny \begin{array}{c}a,b\in {\cal Q}_1, u\in {\cal Q}_0\\t(a)=h(b)=u\\h(a)=t(b)
2508: \end{array}}{\omega_{ab}(\sigma_u ab)_c}~~
2509: \ee
2510: while the `deformation part'
2511: $W_d$ vanishes at zero and is independent of all $\sigma_u$. In the first form of the last expression,
2512: we used Einstein summation over $a$ and $b$ and have set $\sigma=\sum_{u\in {\cal Q}_0}{\sigma_u}\in A_1$.
2513: Notice the lack of a $1/3$ prefactor in (\ref{Wg});
2514: this is because we brought all terms to a form in which a $\sigma_u$ insertion appears in the first position.
2515:
2516: To describe this more elegantly, notice\footnote{For an
2517: element $f=\sum_{n\geq 0} f_{a_1\ldots a_n}(a_1\ldots a_n)_c\in C^0_R(A)$, the condition $\ld_a f=0$ amounts
2518: to vanishing of all cyclic coefficients $f_{a_1\ldots a_n}$ for which one of the $a_j$ coincides with $a$.
2519: Further, the cyclic derivative $\ld_{\lambda }W$ determines $\ld_{\sigma_u}W=\epsilon_u \ld_{\lambda} W$.
2520: Thus (\ref{W_decomp}) and (\ref{Wg}) amount to $\ld_\lambda\left(W+\omega_{ab}(\sigma a b)_c\right)=0$,
2521: which gives the desired statement.}
2522: that (\ref{W_decomp}) together with (\ref{Wg}) amounts to the condition:
2523: \be
2524: \nn
2525: \ld_\lambda W=-\sum_{a,b\in {\cal Q}_1}\omega_{ab} ab~~
2526: \ee
2527: where $\lambda:=\sum_{u\in {\cal Q}_0}{\lambda_u}$ and
2528: $\ld_\lambda:=\sum_{u\in {\cal Q}_0} \ld_{\sigma_u} $ as in Subsection \ref{cycloop}.
2529: Using the graded antisymmetry of $\omega_{ab}$, the last relation takes the form:
2530: \be
2531: \label{ldcond1}
2532: \ld_\lambda W=-\frac{1}{2}\sum_{a\in {\cal Q}_1}[a,a^*]\in [A,A]_2^R~~,
2533: \ee
2534: where we introduced the conjugate variables
2535: \be
2536: \label{adual_expansion}
2537: a^*:=\sum_{b\in {\cal Q}_1}\omega_{ab}b=
2538: \sum_{\tiny \begin{array}{c}b\in {\cal Q}_1 \\h(a)=t(b), h(b)=t(a)\end{array}}{\omega_{ab}b}~~
2539: \ee
2540: as in Subsection \ref{symplectic_structure}. Here $[A,A]_2$ is the subspace of $[A,A]$
2541: consisting of elements of degree two with respect to the
2542: $\N$-grading, while $[A,A]_2^R\subset [A,A]^R$ is the centralizer of $[A,A]_2$ in $R$.
2543:
2544: Relation (\ref{ldcond1}) allows one reconstruct $\omega$ from $W$. To
2545: formulate this invariantly, remember from Subsection \ref{12forms}
2546: that the space of closed noncommutative two-forms $C^2_R
2547: (A)_{\rm closed}$ is isomorphic with $[A,A]^R$. Restricting the map
2548: (\ref{kappa_gen}) to the subspace $C^2_R (A)_2 \subset C^2_R(A)_{\rm
2549: closed}$ of constant two-forms gives an isomorphism
2550: $[A,A]_2^R\stackrel{\approx}{\rightarrow} C^2_R (A)_2$ whose
2551: explicit form is given by relation (\ref{dXi}):
2552: \be
2553: \label{kappa_0}
2554: \sum_{a,b\in {\cal Q}_1}{f_{ab}ab}\stackrel{\kappa}{\rightarrow} -\sum_{a, b\in {\cal Q}_1}f_{ab}(dadb)_c~~.
2555: \ee
2556: Here $f=\sum_{a,b}f_{ab}ab$ is the general element of $[A,A]_2^R$, with graded-antisymmetric complex coefficients
2557: $f_{ab}=(-1)^{1+{\tilde a}{\tilde b}}f_{ba}$, so we can also write $f=\frac{1}{2}\sum_{a,b}f_{ab}[a,b]$.
2558: Applying this to the noncommutative symplectic form, we find:
2559: \be
2560: \kappa^{-1}(\omega)=-\frac{1}{2}\omega_{ab}ab=-\frac{1}{4}\sum_{a\in {\cal Q}_1}[a,a^*]~~.
2561: \ee
2562: Thus relation (\ref{ldcond1}) can be written as either of the following equivalent conditions:
2563: \be
2564: \label{diff_unitality}
2565: \ld_\lambda W=2\kappa^{-1}(\omega)\Leftrightarrow \omega=\frac{1}{2}\kappa(\ld_\lambda W)~~.
2566: \ee
2567: These observations allow us to write the unitality constraint
2568: (\ref{ldcond1}) as condition (\ref{diff_unitality}). In
2569: particular, the element $\ld_\lambda W\in [A,A]_2^R $ must belong to the
2570: subspace spanned by the $\kappa$-preimages of quiver symplectic forms. To
2571: describe this space, notice that any element $\mu\in [A,A]_2^R$ can
2572: be expanded uniquely as:
2573: \be
2574: \mu=-\frac{1}{4}\sum_{a\in {\cal Q}_1}{[a,a^*]}~~,
2575: \ee
2576: where each $a^*$ is a linear combination of arrows going from $h(a)$
2577: to $t(a)$. We say that $\mu$ is {\em non-degenerate} if the elements
2578: $(a^*)_{a\in {\cal Q}_1}$ form a basis of $V$; in this
2579: case, we can expand $a^*$ as in (\ref{adual_expansion}), with
2580: coefficients $\omega_{ab}$ forming the entries of a
2581: graded-antisymmetric non-degenerate matrix. Moreover, it is clear that $\mu$ has
2582: $\Z_2$-degree ${\tilde \omega}$ iff this matrix satisfies the selection
2583: rules (\ref{omega_sel}). We let $\Mom_V\subset [A,A]_2^R$
2584: be the $\Z_2$-homogeneous subspace of non-degenerate elements in $[A,A]_2^R$ and
2585: let $\Mom^0_V$ and $\Mom^1_V$ be its homogeneous components.
2586: The observations made above show that $\kappa$ induces an isomorphism between
2587: the space $CNS^{\tilde \omega}(V)$ of constant
2588: non commutative symplectic forms on $A$ having degree ${\tilde \omega}$ and the space
2589: $\Mom^{\tilde \omega}_V$. It follows from Proposition 8.1.1 of \cite{BEV} that
2590: ${\rm Mom}_V$ is the space of noncommutative moment maps
2591: associated to quiver symplectic forms on the path algebra $A$.
2592:
2593: \
2594:
2595: \noindent We can now formulate the unitality criterion as follows:
2596:
2597: \paragraph{\bf Proposition}
2598: Let $E[1]$ be a symplectic $R$-superbimodule of finite complex
2599: dimension whose symplectic form has degree ${\tilde \omega}$, let
2600: $A:=T_R E[1]^{\rm v}$ and let $W$ be an element of $C^0_R(A)$
2601: which has degree ${\tilde \omega}+1$ and vanishes at zero. Let $\omega$ be the noncommutative symplectic
2602: form induced on $A$ and assume that $\{W,W\}=0$, where $\{.,.\}$ is the Kontsevich bracket defined by
2603: $\omega$. Then the following statements are equivalent:
2604:
2605: (1) The cyclic weak $A_\infty$ structure determined by $W$ on $E[1]$
2606: is unital
2607:
2608: (2) There exists an odd central element $\lambda\in E^R$ such that
2609: $\frac{1}{2}\ld_\lambda W=\kappa^{-1}(\omega)$.
2610:
2611: \noindent In this case:
2612:
2613: (a) $\Sigma \lambda$ is the unit of the $A_\infty$ structure.
2614:
2615: (b) The element $\mu:=\frac{1}{2}\ld_\lambda W\in A$ belongs to the
2616: subspace $\Mom_V$ of $A$
2617:
2618: (c) The non-commutative symplectic form can be recovered via the
2619: relation $\omega=\kappa (\mu)$.
2620:
2621: (d) Let $\lambda=\sum_{u\in {\cal Q}_0}{\lambda_u}$ ($\lambda_u\in
2622: E_{uu}$) be the decomposition of $\lambda$, and choose an adapted
2623: basis $e_a$ of $E[1]$ containing $\lambda_u$ among the basis
2624: elements. Let $s^a\equiv a$ be the dual basis, and let $\sigma_u$ its
2625: elements associated with $\lambda_u$. Then $\sigma_u$ correspond to odd
2626: loops of the associated superquiver and $W$ takes the form given in
2627: eqs. (\ref{W_decomp}) and (\ref{Wg}), where $W_d$ vanishes at zero and
2628: is independent of all $\sigma_u$.
2629:
2630: \
2631:
2632: \subsection{Noncommutative geometry of D-brane systems}
2633:
2634: Combining the discussion of the previous subsections, we have the following non-commutative geometric description
2635: of finite topological D-brane systems:
2636:
2637: \
2638:
2639: {\em Let $R$ be a finite-dimensional semisimple commutative algebra over $\C$
2640: and $E$ an $R$-superbimodule which is finite-dimensional over $\C$.
2641: Giving a finite topological D-brane system with boundary
2642: decomposition described by ($R$, $E$) and topological metrics of $\Z_2$-degree ${\tilde \omega}$ amounts to giving
2643: a `noncommutative function' $W\in C^0_R(A)$ on the tensor algebra $A=T_R E[1]^{\rm v}$
2644: and an odd central element $\lambda \in E^R$ such that
2645: the following conditions are satisfied:
2646:
2647: \
2648:
2649: \noindent (1) $W$ vanishes at zero and has $\Z_2$-degree ${\tilde \omega}+1$.
2650:
2651: \noindent (2) The element $\mu:=\frac{1}{2}\ld_\lambda W$ belongs to $\Mom_V^{\tilde \omega}$
2652:
2653: \noindent (3) We have $\{W,W\}=0$, where $\{.,.\}$ is the Kontsevich bracket determined on $C^0_R(A)$ by
2654: the constant noncommutative symplectic form $\omega:=\kappa (\mu)$.
2655: }
2656:
2657: \paragraph{Observations}
2658:
2659: (1) The coefficient expressions given in Section
2660: \ref{canforms} show that equation $\{W,W\}=0$ is
2661: equivalent with:
2662: \be
2663: \label{W_inf}
2664: \sum_{0\leq i+j\leq n}(-1)^{{\tilde a_1}+\ldots +{\tilde a}_i}
2665: {\bar W}_{a_1\ldots a_i a a_{i+j+1}\ldots a_n} {\bar W}^a_{\,\,a_{i+1}\ldots
2666: a_{i+j}}=0~~{\rm for~all~}~~ n\geq 0~~,
2667: \ee
2668: where lifting of indices is done from the left with $\rho^{ab}=(-1)^{\tilde
2669: b+1}\omega^{ab}$, and $\omega^{ab}$ is the inverse of the matrix $\omega_{ab}=-{\bar W}_{\sigma ab}$.
2670: The first equation (for $n=0$) is ${\bar W}_a{\bar
2671: W}^a=0$. Notice that $\omega$ is determined by $W$, so the equations are not quadratic.
2672: This countable system of nonlinear algebraic conditions is
2673: a non-commutative analogue of the WDVV equations \cite{WDVV}.
2674:
2675: (2) It was shown in \cite{HLL} that the background satisfies the string equations of motion iff
2676: the underlying $A_\infty$ algebra is minimal. It is clear that
2677: this amounts to the requirement that $W$ has order at least $3$ at the origin.
2678:
2679: (3) The structure given above can be viewed as an 'off-shell extension' of the 'boundary part' of the data described in
2680: \cite{CIL1, Moore_Segal, Moore}. The latter arises in the particular case when $W$ has degree three at the origin
2681: (i.e. the underlying $A_\infty$ structure is minimal), and can be recovered
2682: by forgetting all terms of $W$ of order higher than $3$.
2683: In physics language, the structure of \cite{CIL1, Moore_Segal, Moore} corresponds to keeping only the
2684: boundary three-point functions on the disk, thereby forgetting all {\em integrated} amplitudes. As explained in the
2685: introduction to \cite{CIL1}, this reflects the difference between two-dimensional topological field theory and
2686: topological string theory, namely in the topological field theory one does not consider integration of amplitudes
2687: over the moduli space of the underlying Riemann surface (since by definition
2688: the worldsheet metric is not a dynamical variable).
2689:
2690: \subsection{Deformations of the underlying string theory}
2691: \label{theory_deformations}
2692:
2693: Each solution of the constraints described in the previous subsection represents the tree-level boundary data of
2694: an open topological string theory. We would like to make some basic observations about the space of such
2695: theories.
2696:
2697: It is instructive to consider the trivial approximation $W_d=0$ i.e. $W=W_g$, which --- as we shall see in
2698: a moment --- is appropriate under certain assumptions. Starting from $W_g=-\omega_{ab}(\sigma ab)_c$, we
2699: compute:
2700: \be
2701: \label{rdWg}
2702: -W_g\rd_a=\omega_{\alpha a}\sigma \alpha+(-1)^{{\tilde \beta}{\tilde \omega}} \omega_{a\beta}\beta\sigma+
2703: (-1)^{\tilde \omega}\delta^\sigma_ a\omega_{\alpha\beta}\alpha\beta~~
2704: \ee
2705: and:
2706: \be
2707: \label{ldWg}
2708: -\omega^{ab}\ld_b W_g=(-1)^{\tilde \omega}a\sigma +(-1)^{{\tilde a}+{\tilde \omega}}\sigma a+\omega^{a\sigma}
2709: \omega_{\gamma\delta}\gamma\delta~~.
2710: \ee
2711: Combining these equations and using appropriate cyclic permutations gives:
2712: \be
2713: \nn
2714: \{W_g,W_g\}_c=(W_g\rd_a\omega^{ab}\ld_bW_g)_c=(-1)^{\tilde \omega}\omega^{\sigma\sigma}\omega_{\alpha\beta}
2715: \omega_{\gamma\delta}(\alpha\beta\gamma\delta)_c~~.
2716: \ee
2717: Since $\sigma$ is odd, $\omega^{\sigma\sigma}$ vanishes for degree reasons unless ${\tilde \omega}=0$. When
2718: $\omega$ is even, the term in the right hand side need not vanish.
2719:
2720: Let us assume that ${\tilde \omega}=1$ or that ${\tilde \omega}=0$ but
2721: $\omega^{\sigma\sigma}\omega_{\alpha\beta}\omega_{\gamma\delta}(\alpha\beta\gamma\delta)_c$ vanishes.
2722: In this case, we have $\{W_g,W_g\}=0$ and $W_g$ gives a marked point in the space of open
2723: string theories with underlying supermodule $E$, unit $\lambda$ and symplectic form $\omega$. The cyclic unital
2724: $A_\infty$ algebra corresponding to this solution has a single product $r_2^g$, which is given by:
2725: \be
2726: \nn
2727: r_2^g(e_a,e_b)=(-1)^{{\tilde \omega}+1}\omega_{ab}\omega^{\lambda c}e_c~~{\rm for}~~a,b\neq \sigma
2728: \ee
2729: and by the unitality constraint $-r_2^g(\lambda,e_a)=(-1)^{{\tilde a}} r_2^g(e_a,\lambda)=e_a$ for the
2730: remaining combinations of basis elements. The $A_\infty$ constraints (\ref{ainf}) reduce to:
2731: \be
2732: \nn
2733: r_2^g(r_2^g(x,y),z)+(-1)^{{\tilde x}}r_2^g(x,r_2^g(y,z))=0~~,
2734: \ee
2735: which means that $m_g:=\Sigma\circ r_2\circ \Sigma^{\otimes 2}:E^{\otimes 2}\rightarrow E$ satisfies the associativity condition:
2736: \be
2737: \nn
2738: m_g(m_g(x,y),z)=m_g(x,m_g(y,z))~~.
2739: \ee
2740: Moreover, the unitality constraint for $r_2^g$ amounts to $m_g(\lambda, x)=m_g(x,\lambda)=x$. Hence the distinguished
2741: solution $W=W_g$ corresponds to an associative superalgebra
2742: structure on $E$, and the underlying $A_\infty$ category reduces to
2743: an ordinary (i.e. associative) $\Z_2$-graded category. In the
2744: topological string theory, all integrated boundary correlators on the
2745: disk vanish and the entire information is contained in the boundary
2746: three-point functions. Fixing $\lambda$ and $\omega$, other string
2747: theories with the same units and topological metrics are given by
2748: solutions of the equations $\{W_g+W_d, W_g+W_d\}=0$ and $\ld_\lambda W_d=0$, the first of which reduces to:
2749: \be
2750: \nn
2751: \{W_g, W_d\}+\frac{1}{2}\{W_d,W_d\}=0~~,
2752: \ee
2753: where we used the graded antisymmetry property of the Kontsevich bracket. Letting $Q_g=\theta_{W_g}$
2754: be the (odd) Hamiltonian vector field defined by $W_g$ (i.e. $dW_g=i_{Q_g} \omega$), this equation takes the form:
2755: \be
2756: \label{MC}
2757: L_{Q_g}W_d+\frac{1}{2}\{W_d,W_d\}=0~~.
2758: \ee
2759: Notice that $L_{Q_g}$ squares to zero (since $L_{Q_g}^2=\frac{1}{2}[L_{Q_g}, L_{Q_g}]=\frac{1}{2}L_{[Q_g,Q_g]}=
2760: \frac{1}{2}L_{\theta_{\{W_g,W_g\}}}=0$) and that it acts as an odd derivation of the Lie superalgebra
2761: $(C^0_R(A)[{\tilde \omega}],\{.,\})$ (due to the Jacobi identity for the Kontsevich bracket):
2762: \be
2763: \nn
2764: L_{Q_g}\{f,g\}=\{L_{Q_g}f, g\}+(-1)^{{\tilde f}+{\tilde \omega}}\{f, L_{Q_g} g\}~~.
2765: \ee
2766: Hence $(C^0_R(A)[{\tilde \omega}], L_Q, \{.,.\})$ is a differential Lie superalgebra, and (\ref{MC}) is its
2767: Maurer-Cartan equation. This means that one can study the moduli space of boundary string {\em theories}
2768: with fixed units and topological metrics by using the deformation theory of Lie superalgebras.
2769:
2770:
2771: \section{The noncommutative moduli space}
2772:
2773: In this section, we use the formalism developed above to construct a
2774: noncommutative version of the extended moduli space of finite D-brane
2775: systems (the boundary part of the extended moduli space of topological
2776: strings).
2777:
2778: \subsection{Symmetries}
2779: \label{automf_group}
2780:
2781: Let us fix a D-brane system described by the noncommutative
2782: generating function $W$, with Hamiltonian derivation $Q$ and
2783: symplectic form $\omega$. We assume given adapted coordinates
2784: including odd loops $\sigma_u$ associated with the units of the
2785: underlying $A_\infty$ structure. It is clear from the categorical
2786: formulation of Appendix \ref{data} that a symmetry of the D-brane
2787: system amounts to an automorphism of the underlying cyclic and unital
2788: weak $A_\infty$ category, called a cyclic and unital
2789: $A_\infty$ automorphism (an automorphism is a strict autoequivalence, as
2790: appropriate for a finite category). In this subsection, we describe such symmetries
2791: as symplectomorphisms of $A$ which obey certain supplementary properties.
2792:
2793: Given a relative automorphism $\phi$ of $A$, we set
2794: $a':=\phi(a)$ for all $a\in {\cal Q}_1$, and let $V'\subset A$ be the $R$-sub-bimodule
2795: spanned by the elements $a'$. Then $A$ is isomorphic {\em as a superalgebra} with the tensor
2796: algebra $T_R V'$, and $\phi$ can be viewed as a change of
2797: coordinates from $a$ to $a'$. More precisely, the restriction of $\phi$ to $A_1=V$
2798: gives an isomorphism of $R$-superbimodules $\phi_1:V\rightarrow V'$ and $\phi$ can be identified
2799: with the isomorphism of bigraded $R$-superalgebras
2800: $T_R(\phi)=\oplus_{n\geq 0}{\phi^{\otimes_R n}}:T_RV\rightarrow T_R V'$ induced by $\phi_1$.
2801: Of course, we have $a'=\phi_1(a)=\phi(a)$ and $a'$ is an adapted basis for the $R$-superbimodule $V'$.
2802:
2803: $R$-linearity of $\phi$ implies that
2804: each $a'$ is a linear combination of paths starting at $t(a)$ and
2805: ending at $h(a)$. However, $\phi$ need not be homogeneous with respect
2806: to the $\N$-grading of $A$, so generally each $\phi(a)$ can be a
2807: linear combination of paths of different length. Defining $A'_n$ to be
2808: the subspace spanned by $n$-factor monomials in $a'$, we have
2809: $A'_n\approx T_R^n V'$ and a new decomposition:
2810: \be
2811: \nn
2812: A=\oplus_{n\geq 0}{A'_n}~~
2813: \ee
2814: with $A'_0=R$. In particular, a generic $R$-superalgebra automorphism induces a change of $\N$-grading.
2815:
2816: Equation (\ref{df}) implies:
2817: \be
2818: \nn
2819: {\bar d} {\bar \phi}(f)={\bar\phi}^*({\bar d}f)=\sum_{a} (\phi(f\rd_a)d a')_c~~,
2820: \ee
2821: which shows that the cyclic derivatives with respect to the new coordinates are given by:
2822: \be
2823: \label{ld_tf}
2824: {\bar \phi}(f)\rd_{a'}=\phi(f\rd_a)\Leftrightarrow \ld_{a'} {\bar \phi}(f)=\phi(\ld_a f)~~.
2825: \ee
2826:
2827: Relative superalgebra endomorphisms $\phi$ of $A$ having the property $\phi\circ Q \circ \phi=Q$ correspond to
2828: endomorphisms of the underlying weak $A_\infty$ category. The correspondence is obtained by expanding:
2829: \be
2830: \label{phi_expansion}
2831: \phi(a)=\sum_{n\geq 0}\phi_{a_1\ldots a_n}^a a_1\ldots a_n~~,
2832: \ee
2833: where the complex-valued coefficients $\phi_{a_1\ldots a_n}^a$ vanish unless $a_1\ldots a_n$ is a path starting
2834: at $t(a)$ and ending at $h(a)$. The evenness condition on $\phi$ gives the selection rules:
2835: \be
2836: \nn
2837: \phi^a_{a_1\ldots a_n}=0~~{\rm unless}~{\tilde a}={\tilde a}_1+\ldots +{\tilde a}_n~~.
2838: \ee
2839: The $n=0$ part of (\ref{phi_expansion}) stands for the sum over loops
2840: $\sum_{a\in Q_1(u,u)}\phi^{a}\epsilon_u$, i.e. we use the convention $\phi_{a_1\ldots a_0}^a:=
2841: \phi^a \delta^u_{h(a)}\delta^u_{t(a)}\epsilon_u$. Then the $A_\infty$ morphism associated with $\phi$ is
2842: given by the even $R$-multilinear maps $\phi_n:E[1]^n\rightarrow E[1]$ defined through:
2843: \be
2844: \label{phi_maps}
2845: \phi_n(e_{a_1}\ldots e_{a_n})=\phi^a_{a_1\ldots a_n}e_a~~.
2846: \ee
2847: The conditions $\phi(ab)=\phi(a)\phi(b)$ amount to the complicated relations giving the traditional
2848: definition. The maps (\ref{phi_maps}) define an endomorphism of the
2849: weak $A_\infty$ structure on the superbimodule $E$; as usual,
2850: $R$-multilinearity allows one to decompose them into
2851: complex-multilinear maps describing an endomorphism of the underlying
2852: weak $A_\infty$ category. In particular, the map $\phi_0:R\rightarrow
2853: E[1]$ gives even linear maps $\phi_u:\C\rightarrow E_{uu}[1]$ via the
2854: decomposition $\phi_0(\sum_{u}{\alpha_u
2855: \epsilon_u})=\sum_{u}{\epsilon_u \phi_u(\alpha_u)\epsilon_u}$ for
2856: complex $\alpha_u$; these can also be viewed as the odd elements
2857: $\phi_u(1_\C)=\sum_{h(a)=t(a)=u}\phi^a e_a \in E_{uu}$.
2858:
2859:
2860: An $A_\infty$ endomorphism of $(E,(r_n))$ is called {\em unital} if $\phi_1(\lambda)=\lambda$
2861: and $\phi_n(e_{a_1}\ldots e_{a_n})$ for $n\neq 1$ vanishes when any of
2862: the elements $e_{a_1}\ldots e_{a_n}$ coincides with the odd $A_\infty$ unit $\lambda$. In terms of the
2863: coefficients of $\phi$, this means $\phi^a_\sigma=\delta^a_\sigma$ and
2864: $\phi^{a}_{a_1\ldots a_n}=0$ for all $n\neq 1$, if $\sigma\in \{a_1\ldots a_n\}$. Plugging this into
2865: expansion (\ref{phi_expansion}), we see that the $A_\infty$ endomorphism is unital iff:
2866: \be
2867: \label{phi_unitality}
2868: \phi(\sigma)=\sigma~~~{\rm and}~~~\phi(a)={\rm ~independent~of~}\sigma {\rm~for~all}~~a\neq \sigma~~.
2869: \ee
2870:
2871: The $A_\infty$ endomorphism determined by $\phi$ is {\em cyclic} if $\phi^*(\omega)=\omega$; writing
2872: this condition explicitly gives a series of complicated relations used in the traditional definition.
2873: In particular, a cyclic $A_\infty$ automorphism of $(E,\rho, (r_n)))$ amounts to a
2874: relative symplectomorphism of $A$ preserving the homological derivation $Q$.
2875:
2876: Let $\phi\in \Aut_R^\omega(A)$ be a relative symplectomorphism.
2877: Remember from the end of Subsection \ref{sec:bracket} that
2878: the map $\psi_\omega:C^0_R(A)[{\tilde \omega}]\rightarrow \Der_l^\omega(A)$
2879: is equivariant with respect to the action of $\Aut_R^\omega(A)$. Moreover, the exact sequence (\ref{ex}) shows that
2880: $\psi_\omega$ induces an isomorphism of vector spaces $C^0_R(A)[{\tilde \omega}]/R\approx \Der_l^\omega(A)$.
2881: Thus $Q$ is $\phi$-invariant iff its (canonical) Hamiltonian $W$ is invariant under the action of $\phi$ up to
2882: addition of elements of $R$:
2883: \be
2884: \label{W_variation}
2885: {\bar \phi}(W)=W+\alpha~~{\rm~for~some~}\alpha\in R~~.
2886: \ee
2887: Moreover, the associated $A_\infty$ endomorphism is unital iff $\phi$ satisfies relations
2888: (\ref{phi_unitality}).
2889:
2890: Combining these observations, we see that a cyclic and unital
2891: endomorphism of the underlying $A_\infty$ structure amounts to a
2892: symplectomorphism of $A$ which satisfies (\ref{phi_unitality}) and
2893: (\ref{W_variation}). The {\em symmetry group} ${\cal G}\subset
2894: \Aut^\omega_R(A)$ of the system is the group of all such
2895: symplectomorphisms of $A$. If $\phi$ belongs to ${\cal G}$, equations
2896: (\ref{ld_tf}) and (\ref{phi_unitality}), (\ref{W_variation}) imply: \be \nn \mu=\frac{1}{2}\ld_\sigma
2897: W=\frac{1}{2}\ld_{\sigma'} {\bar \phi}(W)=
2898: \frac{1}{2}\phi(\ld_\sigma W)=\phi(\mu)~~, \ee so $\phi$ preserves
2899: the moment element $\mu$.
2900:
2901:
2902: \subsection{Algebraic construction of the noncommutative moduli space}
2903:
2904: \label{nc_moduli}
2905:
2906: Consider the two-sided ideal $J$ of $A$ generated by the elements:
2907: \be
2908: \nn
2909: \ld_a W\in A ~~(a\in {\cal Q}_1)~~,
2910: \ee
2911: which we shall call the {\em critical ideal} of the noncommutative generating function. Notice that $J$
2912: is also generated by $W\rd_a$, due to the relations
2913: $\ld_a W=(-1)^{{\tilde a}{\tilde \omega}}W\rd_a$. We let
2914: $\C[{\cal Z}]:=A/J$. Since $J$ is $\Z_2$-homogeneous (being generated by
2915: $\Z_2$-homogeneous relations), the associative algebra $\C[{\cal Z}]$ is $\Z_2$
2916: graded. Passage to $\C[{\cal Z}]$ implements the conditions:
2917: \be
2918: \nn
2919: \sum_{n\geq 0} {\bar W}_{aa_1\ldots a_n}a_1\ldots a_n=0~~\forall a\in {\cal Q}_1 \Leftrightarrow
2920: \sum_{n\geq 0} {\bar W}_{a_1\ldots a_n a}a_1\ldots a_n=0~~\forall a \in {\cal Q}_1~~,
2921: \ee
2922: which (in view of the isomorphism $V\otimes_R A\approx C^1_R(A)$) can also be written as:
2923: \be
2924: \nn
2925: {\bar d}W=0~~.
2926: \ee
2927: In particular, the distinguished central element $\lambda=\sum_{u\in
2928: {\cal Q}_0}{\lambda_u}$ gives the relation:
2929: \be
2930: \nn
2931: \ld_\lambda W=0\Leftrightarrow \mu=0\Leftrightarrow \omega_{ab}ab=\frac{1}{2}\omega_{ab}[a,b]=0\Leftrightarrow
2932: \sum_{a\in {\cal Q}_1}[a,a^*]=0~~,
2933: \ee
2934: where, as usual, $[a,b]=ab-(-1)^{{\tilde a}{\tilde b}}ba$ is the supercommutator
2935: in $A$. Hence `extremizing' $W$ automatically imposes the zero-level
2936: constraint for the noncommutative moment map of \cite{BEV, Bergh}.
2937:
2938: For $\phi\in \Aut_R(A)$, relations (\ref{df}) imply $\phi^*({\bar d}W)=(\phi(W\rd_a) da')_c$
2939: and $\phi^*({\bar d}W)={\bar d}({\bar \phi}(W))=({\bar \phi}(W)\rd_a da)_c=({\bar \phi}(W)\rd_a d\phi^{-1}(a'))_c$,
2940: where $a':=\phi(a)$. Expanding $d\phi^{-1}(a')$ in the second expression and comparing with the first,
2941: we find that
2942: $\phi(W\rd_a)$ belongs to the ideal generated by ${\bar
2943: \phi}(W)\rd_a$. This shows that relative automorphisms which preserve
2944: $W$ (i.e. ${\bar \phi}(W)=W$) also preserve the ideal $J$, so they
2945: descend to $R$-linear automorphisms of $\C[{\cal Z}]$ (the $R$-superbimodule
2946: structure on $\C[{\cal Z}]$ is induced from its obvious $A$-superbimodule
2947: structure). In particular, the group ${\cal G}$ preserves $J$, and we obtain
2948: a group morphism $\gamma:{\cal G}\rightarrow \Aut_R(\C[{\cal Z}])$, i.e.
2949: an action of ${\cal G}$ by $R$-linear automorphisms of the superalgebra $\C[{\cal Z}]$.
2950: The canonical epimorphism $\zeta:A\rightarrow \C[{\cal Z}]=A/J$ is ${\cal G}$-equivariant:
2951: \be
2952: \nn
2953: \zeta\circ \phi=\gamma(\phi)\circ \zeta~~~~~~\forall \phi\in {\cal G}~~.
2954: \ee
2955: We set $\C{\cal M}=\C[{\cal Z}]^{\cal G}$, the homogeneous
2956: subalgebra of elements invariant under the action of ${\cal G}$. We will view these algebras
2957: as noncommutative coordinate rings of `noncommutative schemes' ${\cal Z}$, ${\cal M}$, which we
2958: call the {\em noncommutative extended vacuum space} and {\em noncommutative extended moduli space}
2959: respectively.
2960:
2961:
2962: \section{The case of a single D-brane}
2963: \label{single_brane}
2964:
2965: Let us illustrate the general discussion with the simple case of a single D-brane.
2966: Then ${\cal Q}$ consists of $m$ loops at a single vertex, and we
2967: let $m_\pm$ be the numbers of even and odd loops.
2968: The space of boundary observables is
2969: $E=\C^{m_-|m_+}$, with parity-changed dual
2970: $V=E[1]^*=\C^{m_+|m_-}$. The boundary algebra $R$ coincides with $\C$ while
2971: the path algebra is the free superalgebra $A=\C\langle\{a\}\rangle $ generated by all loops. The underlying
2972: $A_\infty$ category has a single object, so it reduces to a weak,
2973: cyclic and unital $A_\infty$ algebra on the supervector space $E$.
2974: This is the structure found in \cite{HLL}.
2975:
2976:
2977: It is easy to see that $C^0(A):=C^0_\C(A)$ can be identified with the cyclic subspace $A_{\rm cyclic}$ of $A$,
2978: defined as the image of the idempotent operator:
2979: \be
2980: \nn
2981: P=\id_{\C}\oplus \oplus_{n\geq 1}\left[\frac{1}{n}\sum_{i=0}^{n-1}{(\gamma_n)^i}\right]\in \End_\C(A)~~.
2982: \ee
2983: Here $\gamma_n\in \End_\C (A_n)$ are the generators of the obvious $\Z_n$ action on $A_n$:
2984: \be
2985: \nn
2986: \gamma_n(x_1\otimes \ldots \otimes x_n)=(-1)^{{\tilde x}_1({\tilde x}_2+\ldots +{\tilde x}_n)}x_2\otimes \ldots
2987: \otimes x_n \otimes x_1~~.
2988: \ee
2989: Thus $A_{\rm cyclic}$ consists of all polynomials $f=\sum_{n \geq 0}{f_{a_1\ldots a_n}a_1\ldots a_n}\in A$
2990: whose complex coefficients satisfy the conditions $f_{a_1\ldots a_n}=(-1)^{{\tilde a}_1({\tilde a}_2+\ldots +
2991: {\tilde a}_n)}f_{a_2\ldots a_n a_1}$. Writing $f=\sum_{n\geq 1}{f_n}$ with
2992: $f_n\in A_n =V^{\otimes n}$, such a polynomial belongs to $A_{\rm cyclic}$ iff $\gamma_n(f_n)=f_n$ for
2993: all $n$.
2994:
2995: Let $\sigma$ be the dual basis element corresponding to the parity changed $A_\infty$ unit $\lambda$.
2996: We assume given a basis of $E$ such that $\sigma$ is one of the odd loops.
2997: The generating function is a constant-free polynomial:
2998: \be
2999: \nn
3000: W= \sum_{n\geq 1}W_{a_1\ldots a_n}a_1\ldots a_n\in A_{\rm cyclic}
3001: \ee
3002: in the non-commuting variables $a$ such that
3003: $W_{a_1\ldots a_n}$ are graded-cyclic and satisfy the conditions
3004: $W_{a_1\ldots a_n}=0$ unless ${\tilde a}_1+\ldots +{\tilde
3005: a}_n={\tilde \omega}+1$, as well as the $A_\infty$ constraints
3006: (\ref{W_inf}). Moreover, we must have $W=W_g +W_d$ with
3007: $W_g=-\frac{1}{3}\omega_{ab}[\sigma ab+(-1)^{{\tilde a}+{\tilde b}}ab\sigma +
3008: (-1)^{{\tilde b}({\tilde a}+1)}b\sigma a]$
3009: and where $W_d$ vanishes at zero and is independent of $\sigma$. The
3010: matrix $(\omega_{ab})$ satisfies properties (\ref{omega_symm}) and
3011: (\ref{omega_sel}) of Section \ref{symplectic_structure} and no further
3012: constraints.
3013:
3014: Any endomorphism $\phi$ of $A$ is determined by its values on the generators:
3015: \be
3016: \nn
3017: \phi(a):=\phi_{a}=\sum_{n\geq 0} \phi^{a}_{a_1\ldots a_n}a_1\ldots a_n
3018: \ee
3019: and can be viewed as an $m$-tuple $(\phi_{a_1}\ldots \phi_{a_m})$ of polynomials
3020: in the non-commuting variables $a$. The degree zero condition on $\phi$ gives the constraints
3021: $\deg\phi(a)={\tilde a}$, so the complex coefficient $\phi^{a}_{a_1\ldots a_n}$ must vanish
3022: unless ${\tilde a}_1+\ldots +{\tilde a}_n={\tilde a}$. The relative automorphism group is the usual
3023: group $\Aut(A)$ of superalgebra automorphisms. An automorphism $\phi$ belongs to the symmetry group ${\cal G}$
3024: if $\phi(\sigma)=\sigma$, $\phi(a)$ are independent of $\sigma$ for $a\neq \sigma$
3025: and $W(\{\phi(a)\})$ equals $W(\{a\})$ as a polynomial in the non-commuting variables $\{a\}$. This imposes
3026: nonlinear algebraic conditions on the coefficients $\phi^a_{a_1\ldots a_n}$.
3027:
3028: \paragraph{Observation} The automorphism group $\Aut(A)$ is a rather exotic object.
3029: In the even case $m_-=0$, it is known\cite{Umirbaev} (see also \cite{DY}) that
3030: $\Aut(A)$ contains wild automorphisms\footnote{ An automorphism is
3031: called wild if it is not a composition of the so-called elementary
3032: automorphisms $(a_1\ldots a_m)\rightarrow (a_1\ldots a_i, \alpha
3033: a_i+f(a_1\ldots a_{i-1}, a_{i+1} \ldots a_m), a_{i+1}\ldots a_m)$ with
3034: $\alpha\in \C^*$ and $f$ a polynomial independent of $x_j$.} as soon
3035: as $m\geq 3$. Even in the commutative case, there are well-known
3036: open problems about automorphisms of polynomial algebras such as the
3037: Jacobian conjecture. The $\Z_2$-graded, noncommutative case does not
3038: seem to have been studied systematically.
3039:
3040:
3041:
3042: \
3043:
3044:
3045:
3046: \section{Examples}
3047: \label{examples}
3048:
3049: \subsection{Even system with a single boundary degree of freedom}
3050:
3051: This is the simplest example relevant for topological sigma models
3052: with target spaces of even complex dimension. In this case, we
3053: have a single D-brane ($R=\C$) with $E=\C$ (concentrated in even degree), $V=\C[1]$ ( a purely odd
3054: supervector space) and ${\tilde
3055: \omega}=0$. The canonical forms of Subsection \ref{symplectic_structure}
3056: show that this is the only possibility when the boundary superspace has dimension one.
3057:
3058: Up to rescaling, the boundary sector contains a single boundary observable,
3059: namely the identity operator $1$, which is the even $A_\infty$ unit.
3060: We set $\lambda=\Sigma 1$ and let $\sigma$ be dual odd element in $V=\C[1]$ (of course, $\sigma$ can be
3061: identified with $\lambda$ since we identify $\C^*$ with $\C$ using the canonical basis of $\C$ given by the unit).
3062: The superquiver consists of the single odd loop
3063: $\sigma$, with path superalgebra $A=\C\langle \sigma\rangle$.
3064: By a change of normalization of the $A_\infty$ products, we can take $\omega_{\sigma \sigma}=1$;
3065: then $\omega_{form}=\frac{1}{2}(d\sigma^2)_c$, and $\sigma$ can also be viewed as the canonical odd coordinate
3066: $\xi$ in equation (\ref{e_can}).
3067:
3068: The non-commutative generating function is an odd
3069: polynomial $W=\sum_{n=odd}{W_n(\sigma^n)_c}\in C^0(A)=A_{\rm cyclic}=A^{\rm odd}$. The
3070: unitality constraint requires the splitting $W=W_d+W_g$ with $W_d$
3071: a constant. Since $W$ must vanish at zero, this gives
3072: $W_d=0$. Thus we must have: \be\nn
3073: W=W_g=-\frac{1}{3} \sigma^3 \ee and the only
3074: non-trivial strict coefficient (see Subsection \ref{quivc0}) is ${\bar W}_{\sigma\sigma\sigma}=-1$. The $A_\infty$
3075: constraint (\ref{W_inf}) is trivially satisfied. Since $W$ is cubic, the $A_\infty$ algebra
3076: contains only the product $r_2$, which is completely determined by the unitality constraint
3077: $r_2(\lambda, \lambda)=-\lambda$.
3078: The associative product $\cdot=\Sigma \circ r_2 \circ \Sigma^{\otimes 2}$ on $E$ is given
3079: by $1\cdot 1=1$, which of course is the unique associative product on $\C$ with unit $1$.
3080:
3081: Since elements
3082: $\phi\in {\cal G}$ must preserve $\sigma$, we have ${\cal
3083: G}=\{\id_A\}$. The critical ideal is generated by $\ld_\sigma W=-\sigma^2$, so
3084: $J=A\sigma^2A$ consists of all polynomials of order at least two at the
3085: origin. Thus $\C[{\cal Z}]=A/(\sigma^2)$ is the Grassmann algebra $\C[\sigma]$
3086: on the odd generator $\sigma$. The noncommutative moduli space coincides
3087: with the noncommutative vacuum space, having coordinate ring: \be
3088: \C[{\cal M}]=\C[{\cal Z}]=\C[\sigma]~~.\nn \ee Hence ${\cal M}={\cal
3089: Z}=\C^{0|1}$, the odd point of usual supergeometry.
3090:
3091: In this extremely simple example, passage to the noncommutative moduli
3092: space gives nothing new, since supercommutativity is imposed as a
3093: consequence of the equations of motion.
3094:
3095:
3096: \subsection{Odd system with two boundary degrees of freedom}
3097: This is the simplest example with an odd topological boundary metric, obtained for
3098: $E=\C^{1|1}$. Choosing canonical coordinates as in (\ref{o_can}), we have:
3099: \be
3100: \nn
3101: \omega_{form}=(dp_0dq_0)_c
3102: \ee
3103: with odd $q_0$ and even $p_0$, and take $\sigma=q_0$. Since
3104: $\sigma$ must be a loop, the only possibility for the boundary algebra
3105: is $R=\C$, i.e. the index superquiver consists of an even and an
3106: odd loop at a single vertex (indeed, $\omega_{form}$ vanishes for any
3107: other boundary structure). Thus $A=\C\langle p_0,q_0\rangle$ is a free
3108: associative superalgebra. In this case, we have $W_g=2(q_0^2p_0)_c$
3109: and $W_d=W_d(p_0)$ must be a constant-free univariate polynomial in
3110: $p_0$. It is not hard to see that the Maurer-Cartan equation
3111: (\ref{MC}) for $W_d$ is trivially satisfied. Indeed, it is clear that
3112: $\{W_d,W_d\}=0$, while direct computation gives:
3113: \be
3114: \nn
3115: \{W_g,W_d\}=-(W_g\rd_{q_0}\ld_{p_0}W_d)_c=
3116: 2([q_0,p_0]\ld_{p_0}W_d)_c=2(q_0[p_0, \ld_{p_0}W_d])_c=0
3117: \ee
3118: where we noticed that the commutator $[p_0, \ld_{p_0}W_d]$ vanishes
3119: because $W_d$ is a polynomial in $p_0$. Hence the general system of
3120: this type is described by $W=W_g+W_d$, where $W_d$ is an arbitrary
3121: constant-free polynomial in $p_0$. The defining equations $\ld_a W=0$
3122: for the noncommutative vacuum space take the form:
3123: \be
3124: \label{def_eqs}
3125: [q_0,p_0]=0~~,~~q_0^2=-\frac{1}{2}\ld_{p_0}W_d(p_0)~~.
3126: \ee
3127: Thus $\C[{\cal Z}]=\C\langle q_0,p_0\rangle/([q_0,p_0],
3128: q_0^2+\frac{1}{2}\ld_{p_0}W_d(p_0))$ and ${\cal Z}$ is a bona-fide
3129: noncommutative superspace. It can be viewed as a `fibration' over the
3130: noncommutative affine line $\A^1$ with coordinate $p_0$, where the
3131: (pure fuzz) fiber is a point-dependent deformation of the usual odd point
3132: $\C^{0|1}$.
3133:
3134: It is known \cite{Cz, ML} that all automorphisms of a free associative
3135: algebra on two generators are tame, i.e. given by iterated composition
3136: of elementary automorphisms of the form $(q_0,p_0)\rightarrow (q_0,
3137: \alpha p_0+f(q_0))$ and $(q_0,p_0)\rightarrow (\beta q_0+g(p_0),p_0)$
3138: with $\alpha,\beta\in \C^*$ and $f,g$ arbitrary univariate
3139: polynomials. Moreover (see \cite{DY} for a more general result),
3140: all algebra automorphisms fixing one variable are triangular. In particular,
3141: automorphisms fixing $q_0$ have the form $(q_0,p_0)\stackrel{\phi}{\rightarrow}(q_0,\alpha p_0+f(q_0))$,
3142: with {\em super}algebra automorphisms obtained by restricting to even polynomials.
3143: Obviously $\phi(p_0)$ is independent of $\sigma=q_0$ iff $f$ is the constant polynomial. In this
3144: case, $\phi$ is a symplectomorphism iff $\alpha=1$. Hence ${\cal G}$ is
3145: a subgroup of the one-dimensional translation group:
3146: \be
3147: \nn
3148: {\cal T}:~~q_0\rightarrow q_0~~,~~p_0\rightarrow p_0+t~~~~(t\in \C)~~.
3149: \ee
3150: Such an automorphism preserves $W_g$ up to addition of constants, and takes $W_d(p_0)$ into $W_d(p_0+t)$.
3151: It follows that $W$ is preserved up to constants iff $W_d$ is linear in $p_0$.
3152: Thus ${\cal G}=\{\id_A\}$ unless $W_d=w(p_0)_c$ for some $w\in \C$, in which case ${\cal G}={\cal T}$.
3153:
3154: Hence the generic case of a nonlinear $W_d$ gives ${\cal M}={\cal Z}$.
3155: When $W_d=w(p_0)_c$, equations (\ref{def_eqs})
3156: reduce to $[q_0,p_0]=0$ and $q_0^2=-\frac{w}{2}$
3157: and we find $\C[{\cal M}]=\C[{\cal Z}]^{\cal T}=\C\langle
3158: q_0\rangle/(q_0^2+\frac{w}{2})$, i.e ${\cal M}$ is the quantum
3159: deformation of the odd point $\C^{0|1}$ given by $q_0^2=-\frac{w}{2}$.
3160:
3161:
3162:
3163: \subsection{A family of odd examples}
3164:
3165: Consider a theory with ${\tilde \omega}=1$, where we choose adapted
3166: coordinates $p_0\ldots p_m, q_0\ldots q_m$ (with $2(m+1)={\rm Card}
3167: {\cal Q}_0$, odd $q_i$ and even $p_i$) such that that $\omega$ has
3168: canonical form (\ref{o_can}). Remember from Subsection
3169: \ref{symplectic_structure} that the coordinates $a^*=\omega^{ab}b$ are
3170: given by:
3171: \be
3172: \nn
3173: p_i^*=q_i~~,~~q_i^*=-p_i~~,
3174: \ee
3175: where $*$ can be viewed as an involution on $A_1=V$. Also remember
3176: that $p_i$ are even and $q_i$ are odd. We assume that $q_0=\sigma$
3177: corresponds to the unit.
3178:
3179:
3180: Since ${\tilde \omega}=1$, we have $\{W_g,W_g\}=0$ and the discussion
3181: of Subsection \ref{theory_deformations} applies. In particular, a
3182: general theory with the given unit and symplectic form is specified by
3183: a solution $W_d$ of the Maurer-Cartan equation (\ref{MC}). For
3184: ${\tilde \omega}=1$, relation (\ref{rdWg}) gives:
3185: \be
3186: \nn
3187: W_g\rd_a=[\sigma,a^*]+\delta_a^\sigma\sum_{\alpha\in Q_1}{\alpha\alpha^*}
3188: \ee
3189: and the conditions $W\rd_a=0$ take the form:
3190: \be
3191: \nn
3192: [\sigma,a^*]+\delta_a^\sigma\sum_{\alpha\in Q_1}{\alpha\alpha^*}=-W_d\rd_a~~,
3193: \ee
3194: where, as usual, $[.,.]$ stands for the graded commutator. In
3195: canonical coordinates, we have $[\sigma,
3196: \sigma^*]=[q_0,q_0^*]=-[q_0,p_0]=[p_0,q_0]$ and $\sum_{\alpha\in
3197: Q_1}{\alpha\alpha^*}=\sum_{i=0}^m{[p_i,q_i]}$. The noncommutative
3198: criticality constraints become:
3199: \be
3200: \label{E1}
3201: 2[p_0,q_0]+\sum_{i=1}^m{[p_i,q_i]}=-W_d\rd_{q_0}~~,
3202: \ee
3203: and:
3204: \be
3205: \label{E2}
3206: [q_0,p_i]=W_d\rd_{q_i}~~\forall i=1\ldots m,~~[q_0, q_i]=-W_d\rd_{p_i}~~~~~\forall i=0\ldots m~~.
3207: \ee
3208:
3209: Let us assume that we are given a particular solution $W=W_g+W_d$ for
3210: which $W_d=W_d(p_1\ldots p_m)$ depends only on $p_1\ldots
3211: p_m$. Special solutions of this type were found in \cite{Katz} for
3212: systems which obey $\Z$-valued selection rules (for the examples of
3213: \cite{Katz}, equation (\ref{MC}) is trivially satisfied by the ansatz
3214: $W_d=W_d(p_1\ldots p_m)$ due to the integer-valued degree condition
3215: obeyed by $W$ in the case of Calabi-Yau compactifications). In this
3216: case, eqs. (\ref{E1}) and (\ref{E2}) reduce to the following defining
3217: relations for the noncommutative vacuum space ${\cal Z}$:
3218: \begin{eqnarray}
3219: \label{Z_eqs}
3220: [q_0,q_i]&=&-\ld_{p_i}W_d~~~~~\forall i=1\ldots m\nn\\
3221: \left[q_0,p_i\right]&=&0~~~~~~~~~~~~~~~~\forall i=1\ldots m\\
3222: \left[q_0,p_0\right] &=&\frac{1}{2}\sum_{j=1}^m{[p_j,q_j]}~~\nn\\
3223: q_0^2&=&0~~\nn~~.
3224: \end{eqnarray}
3225: To arrive at this form, we noticed that $\ld_{p_i} W=W\rd_{p_i}$ since
3226: $p_i$ are even. Note that $[q_0,q_i]=q_0q_i+q_iq_0$ since $[.,.]$ is
3227: the graded commutator.
3228:
3229: For simplicity, let us consider the case when the underlying quiver
3230: has a single vertex. Then we can view ${\cal Z}$ as a `fibration'
3231: over the noncommutative affine plane $\A^{m+1}$ with coordinates
3232: $p_0\ldots p_m$, whose `fiber' is a subspace of the noncommutative
3233: affine space $\A^{m+1}$ (with coordinates $q_0\ldots q_m$) determined
3234: by (\ref{Z_eqs}). The Abelian locus ${\cal Z}_{Ab}$ in ${\cal Z}$ is
3235: obtained by requiring that all variables supercommute. In this case,
3236: eqs. (\ref{Z_eqs}) reduce to the conditions
3237: $\partial_{p_i}W_d^{Ab}(p_1\ldots p_m)=0$ and we find that ${\cal
3238: Z}_{Ab}$ coincides with the critical locus ${\rm
3239: Crit}(W_d^{Ab})\subset \C^{m}$ of $W_d^{Ab}$, which is the usual
3240: vacuum space expected in the supercommutative formulation. The
3241: Abelianization epimorphism $\C\langle q_i, p_i\rangle \rightarrow
3242: \C[q_i,p_i]$ induces an embedding of ${\cal Z}_{Ab}$ into the much
3243: larger noncommutative space ${\cal Z}$. In the noncommutative vacuum
3244: space, one can move away from the critical locus of $W_{Ab}$ at the
3245: price of allowing for a non-vanishing commutator of $q_0$ with $q_i$;
3246: notice that this is possible even along the locus in ${\cal Z}$ where
3247: $p_i$ are required to commute.
3248: Determining the symmetry group ${\cal G}$ and noncommutative moduli
3249: space ${\cal M}$ in this class of examples is rather formidable in
3250: general and will not be attempted here. We only note that the trivial case $m=0$ corresponds
3251: to the limit $W_d=0$ of the example discussed in the previous subsection.
3252:
3253:
3254:
3255:
3256:
3257:
3258: \section{Conclusions}
3259:
3260: We showed that the totality of boundary tree-level data determined by
3261: a topological string theory in a finite D-brane background can be
3262: encoded {\em faithfully} by using the non-commutative algebraic
3263: geometry of a superquiver determined by the boundary decomposition of
3264: the D-brane system. In particular, cyclicity of integrated boundary
3265: amplitudes on the disk and the weak $A_\infty$ constraints on such
3266: amplitudes amount to the condition $\{W,W\}=0$, where the boundary
3267: potential $W$ is a function defined on a {\em noncommutative}
3268: superspace $\A_{\cal Q}$ determined by the quiver. We also found a
3269: differential constraint on $W$ which expresses the presence of unit
3270: boundary observables in the boundary-preserving sector.
3271:
3272: Fixing the bulk worldsheet data, but varying the D-brane background,
3273: gives rise to the (extended) boundary moduli space of such a
3274: system. We argued that this moduli space can be viewed as a
3275: noncommutative superspace ${\cal M}$ constructed as an (invariant
3276: theory) quotient of the `noncommutative critical locus' ${\cal Z}$ of
3277: $W$ by a certain group of symplectomorphisms acting on ${\cal Z}$.
3278:
3279: One upshot of this analysis is that the complicated structure
3280: determined by {\em all} integrated boundary correlators on the disk is
3281: encoded faithfully by a form of noncommutative geometry. According to
3282: this point of view, the theory of topological D-brane deformations is
3283: intrinsically noncommutative. This gives a stringy realization of
3284: non-commutativity at the level of such moduli spaces. It should be
3285: compared with the realization of \cite{SW}, which arises by
3286: translating the effective action of open strings into an action
3287: governing objects (such as connections) defined over a noncommutative
3288: space determined by the {\em closed} strings. In both cases,
3289: non-commutativity originates \cite{Schomerus} from the fact that
3290: insertions of boundary observables on the disk do not commute. Hence
3291: these ostensibly different realizations are related, and it would be
3292: interesting to understand precisely how.
3293:
3294:
3295: \
3296:
3297: \acknowledgments{ I am grateful to V. Ginzburg for pointing out
3298: the relevance of quiver algebras, and to M. Kontsevich for sharing his
3299: ideas. This work originated from discussions during the Third Workshop
3300: on Noncommutative Algebraic Geometry held at the Mittag Leffler
3301: Institute.}
3302:
3303: \appendix
3304:
3305: \section{Topological D-brane systems as cyclic and unital weak $A_\infty$ categories}
3306: \label{data}
3307:
3308:
3309: \subsection{Mathematical background}
3310:
3311: A weak ($\Z_2$-graded) $A_\infty$ category ${\cal A}$ consists of a
3312: collection of objects $Ob{\cal A}$ and complex supervector spaces
3313: $\Hom_{\cal A}(u,v):=\Hom(u,v)$ for $u,v\in Ob {\cal A}$, together with
3314: odd multilinear maps\footnote{Notice that we take morphisms to compose {\em forward} under the \
3315: $A_\infty$ products.} $r_{u_1\ldots
3316: u_{n+1}}:\Hom(u_1, u_2)[1]\times \Hom(u_2, u_3)[1]\times\ldots \times \Hom(u_n, u_{n+1})[1]
3317: \rightarrow \Hom(u_1,u_{n+1})[1]$ for all $n\geq 0$. The
3318: case $n=0$ corresponds to odd maps $r_u:\C\rightarrow \Hom(u,u)[1]$, which
3319: amounts to giving even elements $\theta_u=r_u(1)\in \Hom(u,u)$. The maps $r$ are
3320: required to satisfy certain conditions called the weak $A_\infty$
3321: constraints. To formulate them, let $t(x), h(x)\in Ob {\cal A}$
3322: denote the tail and head of a morphism $x$, the unique objects of
3323: ${\cal A}$ such that $x\in \Hom(t(x), h(x))$. We say that an ordered
3324: collection of morphisms $(x_1,\ldots, x_n)$ is {\em composable} if
3325: $h(x_j)=t(x_{j+1})$ for all $j=1\ldots n-1$. In this
3326: case, we let:
3327: \be
3328: [x_1\ldots x_n]:=t(x_1)t(x_2)\ldots t(x_n)h(x_n)~~,
3329: \ee
3330: viewed as a word on the set $Ob {\cal A}$.
3331: In particular, we set $[x]=t(x) h(x)$ for all morphisms $x$.
3332:
3333: We use $|.|$ to denote the degree of homogeneous elements of
3334: $\Hom(u,v)$ and ${\tilde .}$ for the degree of homogeneous elements of
3335: $\Hom(u,v)[1]$. Then the maps $r_n$ are required to satisfy:
3336: \be
3337: \label{cat_ainf}
3338: \sum_{0\leq i+j\leq n} (-1)^{{\tilde x}_1+\ldots +{\tilde x}_i}
3339: r_{[x_1\ldots x_i][x_{i+j+1}\ldots x_{n}]}
3340: (x_1\ldots x_i, r_{[x_{i+1}\ldots x_{i+j}]}(x_{i+1}\ldots x_{i+j}), x_{i+j+1}\ldots x_n)=0~~
3341: \ee
3342: for any system of $\Z_2$-homogeneous and composable morphisms $(x_1, \ldots, x_n)$.
3343:
3344: A weak $A_\infty$ category is called {\em strong} if $r_u=0\Leftrightarrow \theta_u=0$ for all
3345: $u\in Ob {\cal A}$ and {\em minimal} if it is strong and $r_{uv}=0$
3346: for all $u,v\in Ob{\cal A}$. It is called {\em unital} if one is given
3347: even elements $1_u\in \Hom(u,u)$ for each object $u$, such that the following conditions are
3348: satisfied:
3349: \begin{eqnarray}
3350: \label{cat_unitality}
3351: r_{[x_1\ldots x_{j-1}] [x_{j+1}\ldots x_n]}(x_1\ldots x_{j-1},e_{u_j},x_{j+1}\ldots x_n)&=&0~~
3352: {\rm for~all}~~~~ n\neq 2 ~{\rm~and~all}~ j~~\nn\\
3353: r_{[\lambda_u,x]}(\lambda_u,x)=-x~~,~~r_{[y,\lambda_u]}(y,\lambda_u)&=&(-1)^{{\tilde y}} y~~,
3354: \end{eqnarray}
3355: where $\lambda_u:=\Sigma 1_u\in \Hom(u,u)[1]$ and $u_j:=h(x_{j-1})=t(x_{j+1})$. In these relations, it is understood
3356: that $(x_1,\ldots, x_{j-1},x_{j+1},\ldots, x_n)$ is any composable system consisting of $\Z_2$-homogeneous elements.
3357: The last conditions in (\ref{cat_unitality}) are
3358: imposed for any homogeneous $x\in \Hom(u,v)$ and $y\in \Hom(v,u)$, with arbitrary $u,v$.
3359: The elements $1_u, \lambda_u$ are called {\em units} and {\em odd units} respectively.
3360:
3361: The $A_\infty$ category is called {\em cyclic} if one is given
3362: non-degenerate\footnote{This means that all linear maps $\Hom(v,u)\rightarrow \Hom(u,v)^*$ determined by
3363: $\rho_{uv}$ are bijective.} bilinear forms $\rho_{uv}:\Hom(u,v)\times
3364: \Hom(v,u)\rightarrow \C$, homogeneous of the same $\Z_2$-degree ${\tilde
3365: \omega}$, such that $\rho_{uv}(x,y)=(-1)^{|x||y|}\rho_{vu}(y,x)$ and such that the
3366: following identities are satisfied:
3367: \be
3368: \label{cat_rcyc}
3369: \rho_{t(x_0) h(x_0)}(x_0,r_{[x_1\ldots x_n]}(x_1\ldots x_n))=
3370: (-1)^{{\tilde x}_0({\tilde x}_1+\ldots +{\tilde x}_n)}
3371: \rho_{t(x_1) h(x_1)}(x_1,r_{[x_2\ldots x_0]}(x_2\ldots x_n, x_0))~~,
3372: \ee
3373: whenever $(x_0,x_1 ,\ldots, x_n)$ is a homogeneous composable system
3374: and $[x_0\ldots x_n]$ is a cyclic word, i.e. $h(x_n)=t(x_0)$. Equivalently,
3375: \be
3376: \label{cat_rrcyc}
3377: \omega_{t(x_0) h(x_0)}(x_0,r_{[x_1\ldots x_n]}(x_1\ldots x_n))=
3378: (-1)^{{{\tilde x}_0+{\tilde x}_1+\tilde x}_0({\tilde x}_1+\ldots +{\tilde x}_n)}
3379: \omega_{t(x_1) h(x_1)}(x_1,r_{[x_2\ldots x_0]}(x_2\ldots x_n, x_0))~~.
3380: \ee
3381: where $\omega_{uv}=\rho_{uv}\circ \Sigma^{\otimes 2}$
3382: (i.e. $\omega_{uv}(x,y)=(-1)^{\tilde x}\rho_{uv}(x,y)$) are the
3383: suspended bilinear forms $\omega_{uv}:\Hom(u,v)[1]\times
3384: \Hom(v,u)[1]\rightarrow \C$, which satisfy
3385: $\omega_{uv}(x,y)=(-1)^{{\tilde x}{\tilde y}+1}\omega_{vu}(y,x)$.
3386:
3387:
3388: The concept of $A_\infty$ category was introduced in \cite{Fukaya_inf}
3389: as a generalization of the
3390: notion of $A_\infty$ algebra \cite{Stasheff}. These objects are studied mathematically
3391: in \cite{Keller, Fukaya_mirror, KS, Hasegawa, Lyubashenko1,Lyubashenko2,Lyubashenko3, Lyubashenko4}.
3392: It is by now well-understood that they play an important role in homological algebra, in particular
3393: in giving a natural formulation of derived categories \cite{KS_book}. They also play a crucial
3394: role in the homological mirror symmetry program \cite{HMS}.
3395:
3396: \subsection{Category-theoretic description of topological D-brane systems}
3397:
3398: It was pointed out in \cite{CIL5} (and derived in \cite{HLL} from the
3399: worldsheet perspective, see also \cite{Costello1, Costello2}) that topological D-branes in string theory
3400: form the structure of a weak, cyclic and unital $A_\infty$
3401: category. In this realization, the D-branes of the theory are the
3402: objects of $A$, while each morphism space $\Hom_{\cal A}(u,v)$ is the
3403: space of zero-form topological observables for a string stretching from $u$ to
3404: $v$. The bilinear forms $\rho_{uv}$ are the topological metrics, while
3405: the $A_\infty$ units $1_u$ are the identity observables in the
3406: boundary sectors $\Hom_{\cal A}(u,u)$. The $A_\infty$ products arise
3407: by dualizing the integrated correlators on the disk. These can be
3408: recovered from the former with the aid of the topological metrics:
3409: \be
3410: \nn
3411: \langle \langle x_0\ldots x_n\rangle \rangle =\rho_{h(x_0) t(x_0)}(x_0,r_{[x_1\ldots x_n]}(x_1\ldots x_n))~~,
3412: \ee
3413: for composable $(x_0\ldots x_n)$ such that $[x_0\ldots x_n]$ is a cyclic word.
3414: We refer the reader to \cite{HLL, CIL5} for further details.
3415:
3416: As explained in \cite{HLL, CIL5}, non-vanishing maps $r_u$ are present when
3417: the background of the topological string theory does not satisfy the string equations of motion.
3418: In this case, the elements $\theta_u=r_u(1)\in \Hom_{\cal A}(u,u)$ correspond to
3419: tadpoles. This can usually be corrected by shifting the string vacuum
3420: until all tadpoles are eliminated \cite{CIL5, HLL}. One sometimes has an obstruction to reaching a
3421: solution, a phenomenon which was originally noticed in \cite{Fukaya_book}.
3422:
3423: The structure described above is a homotopy-theoretic generalization
3424: of the boundary data of a topological field theory defined on open
3425: Riemann surfaces, which was studied in \cite{CIL1, Moore_Segal,
3426: Moore}. The latter data arises by forgetting all integrated boundary
3427: correlators, thereby keeping only the information contained in the
3428: boundary units, topological metrics and three-point functions.
3429:
3430: The fact that weak cyclic and unital $A_\infty$ categories describe
3431: topological D-brane systems is implicit in the homological mirror
3432: symmetry conjecture \cite{HMS} and in the work of Fukaya and
3433: collaborators \cite{Fukaya_inf, Fukaya_book, Fukaya_mirror}. The
3434: physical interpretation of this was discussed in \cite{CIL5}, the dG
3435: case having been considered previously in \cite{CIL2, CIL3, CIL6}. A
3436: connection with the D-brane superpotential and Chern-Simons theory was
3437: made in \cite{CIL4, CIL5}, realizing explicitly an observation
3438: originally due to \cite{Witten_CS}. See \cite{CIL7,CIL8,CIL9,CIL10}
3439: for further physics discussion. $A_\infty$ algebras as descriptions of
3440: open string vertices appeared originally in \cite{Gaberdiel} in the
3441: context of bosonic string field theory, which was further studied in
3442: \cite{Kajiura}. As discussed in \cite{CIL4, CIL5, Kajiura}, the
3443: $A_\infty$ structure describing open topological strings can also be
3444: obtained from the results of \cite{Gaberdiel}, by using the well-known
3445: formal analogy of bosonic and topological string theories.
3446:
3447: \section{Some basic isomorphisms}
3448: \label{isomorphisms}
3449:
3450: Convention (\ref{dual_multiplications}) for the superbimodule structure of the dual
3451: affects some standard isomorphisms. Given two $R$-superbimodules $U,V$, the
3452: supervector space $\Hom_{\rm Mod-R}(U^{opp}, V)$ becomes an $R$- superbimodule
3453: with respect to the external multiplications given by $(\alpha\phi\beta)(x)=\alpha\phi(x\beta)$.
3454: There exists an isomorphism of $R$-superbimodules
3455: $\Hom_{\rm Mod-R}(U^{opp}, V) \approx V\otimes_R U^{\rm v}$, whose inverse takes
3456: $y\otimes_R f\in V\otimes_R U^{\rm v}$
3457: into the right $R$-supermodule morphism $\phi:U^{opp}\rightarrow V$ given by
3458: $\phi(x):=yf(x)$. This isomorphism maps the $R$-sub-bimodule $\Hom_{\rm R-Mod-R}(U^{opp},V)$
3459: of $\Hom_{\rm Mod-R}(U^{opp}, V)$ into the center of $V\otimes_R U^{\rm v}$:
3460: \be
3461: \label{basic_isomorphism}
3462: \iHom(U^{opp},V)=\Hom_{\rm R-Mod-R}(U^{opp},V)\approx [V\otimes_R U^{\rm v}]^R~~.
3463: \ee
3464: In particular, we have $\iHom(U^{opp},U)\approx [U^{\rm v}\otimes_R U^{\rm v}]^R$, so
3465: given a bilinear form $\sigma$ on $U$, the map $j_\sigma:U^{opp}\rightarrow U^{\rm v}$ defined in
3466: Section 2 can be identified with a central element ${\hat \sigma}\in [U^{\rm v}\otimes_R U^{\rm v}]^R$.
3467: Tracing through the identifications, one finds that $\sigma$ can be recovered from ${\hat \sigma}$
3468: as follows. If ${\hat \sigma}=\sum_{i} f_i\otimes_R g_i $ with $f_i,g_i\in U^{\rm v}$, then
3469: $\sigma(x,y)=\sum_{i}f_i(xg_i(y))$.
3470:
3471:
3472:
3473: \section{Explicit construction of the differential envelope}
3474: \label{envelope}
3475: The construction below is a slight adaptation of that given \cite{CK}.
3476: Let $A$ be an $R$-superalgebra.
3477: \bd
3478: Consider the $R$-superbimodule $A_R:=A/R$ and let $\nu:A\rightarrow
3479: A_R$ be the natural projection. We let $\nu(a):={\bar a}$ for all
3480: $a\in A$ and let $K_R =A\oplus A_R $, viewed as an $R$-superbimodule.
3481: Define:
3482: \be
3483: \Omega_R A=T_R K_R /J
3484: \ee
3485: where $T_R K$ (endowed with its obvious $\Z\times \Z_2$ grading) is
3486: the tensor algebra on $K$ and $J\subset T_R K_R $ is the $\Z\times
3487: \Z_2$-homogeneous two-sided ideal generated by all elements of the
3488: form ${\overline a}\otimes b+a\otimes {\overline b}-\overline{ab}$ and
3489: $a\otimes b-ab$ with $a,b\in A$. We let $d$ be the unique $R$-linear
3490: derivation of $\Omega_R A$ of bidegree $(1,0)\in \Z\times \Z_2$ (with
3491: respect to the pairing (\ref{pairing})) which satisfies the relations:
3492: \be
3493: da={\bar a}~~,~~d{\bar a}=0~~{\rm for~all~}a\in A~~
3494: \ee
3495: (this derivation is well-defined). The $\Z\times \Z_2$-grading of $\Omega_R A$ is induced by the obvious
3496: $\Z\times \Z_2$-grading of $T_R K_R$.
3497: \ed
3498:
3499: \noindent If $\cdot$ denotes multiplication in $\Omega_R A$, we find
3500: that $\Omega_R A$ is spanned by all finite products of elements of the
3501: form $a$ and $db={\bar b}$ with $a,b\in A$, with the relations
3502: $d(ab)=(da)\cdot b+a\cdot (db)$ and $a\cdot b=ab$
3503: Notice that the unit element of $\Omega_R A$ coincides with $1_A$ due to
3504: the relation $1_A\cdot a=1_A a=a=a1_A=a\cdot 1_A$ for $a\in A$, and
3505: that $d$ squares to zero due to the relations $d{\bar
3506: a}=0\Leftrightarrow d^2a=0$ for $a\in A$. Moreover, any element of
3507: $\Omega_ A$ can be written as a finite sum of elements of the form
3508: $a_0\cdot da_1\cdot \ldots \cdot da_n$ with $a_i\in A$, by applying
3509: the identity:
3510: \begin{eqnarray}
3511: a_0 \cdot{\bar a}_1\cdot\ldots \cdot{\bar a}_n\cdot b_0 \cdot{\bar
3512: b}_1\cdot \ldots \cdot{\bar b}_m&:=&a_0 \cdot{\bar a}_1 \cdot \ldots
3513: {\bar a}_{n-1}\cdot\overline{a_n b_0}\cdot~{\bar b}_1\cdot\ldots
3514: \cdot{\bar b}_m\nn\\ &&+\sum_{i=1}^{n-1}{(-1)^{n-i}a_0 \cdot{\bar
3515: a}_1\cdot\ldots \cdot\overline{a_{n-i}a_{n-i+1}} \cdot\ldots \cdot
3516: {\bar a}_n\cdot {\bar b}_0}\cdot\ldots \cdot{\bar b}_m \nn\\ &&+(-1)^n
3517: (a_0a_1) \cdot{\bar a}_2 \cdot\ldots \cdot{\bar a}_n \cdot{\bar
3518: b}_0\cdot \ldots \cdot{\bar b}_m~~\nn\\ ~~{\rm for~all}~~ a_0\ldots
3519: a_n, b_0\ldots b_m\in A~~, \nn
3520: \end{eqnarray}
3521: which follows from the relations valid in $\Omega_R A$. Furthermore, we
3522: can denote the product in $\Omega_R A$ by juxtaposition.
3523:
3524: \section{Coefficient expression for the cyclic bracket}
3525: \label{coeffs}
3526:
3527: \bp
3528: \label{cycbracket}
3529: Let $f, g\in C^0_R(A)$ have the forms:
3530: \begin{eqnarray}
3531: f&=&\sum_{n\geq 0}{f_{a_1\ldots a_n}({a_1}\ldots {a_n})_c}=c(f)+
3532: \sum_{n\geq 1}{\frac{{\bar f}_{a_1\ldots a_n}}{n}({a_1}\ldots {a_n})_c}
3533: ~~\nn\\
3534: g&=&\sum_{n\geq 0}{g_{a_1\ldots a_n}({a_1}\ldots {a_n})_c}=
3535: c(g)+
3536: \sum_{n\geq 1}{\frac{{\bar g}_{a_1\ldots a_n}}{n}({a_1}\ldots {a_n})_c}~~\nn
3537: \end{eqnarray}
3538: with strict coefficients ${\bar f}_{a_1\ldots a_n}$, ${\bar g}_{a_1\ldots a_n}$
3539: and assume that both $f$ and $g$ are homogeneous of degree
3540: ${\tilde \omega}+1\in \Z_2$. Then the cyclic bracket of $f$ and $g$:
3541: \be
3542: \{f,g\}=c(\{f,g\})+\sum_{n\geq 1}{\frac{{\bar \phi}_{a_1\ldots a_n}}{n} ({a_1}\ldots {a_n})_c}~~,\nn
3543: \ee
3544: has strict coefficients:
3545: \begin{eqnarray}
3546: \label{strict_bracket}
3547: {\bar \phi}_{a_1\ldots a_n}=\sum_{j=0}^n \sum_{i=0}^{n-j} (-1)^{{\tilde a}_1+\ldots + {\tilde a}_i}
3548: &[&{\bar f}_{a_1\ldots a_i a a_{i+j+1}\ldots a_n} {\bar g}^a_{\,\,a_{i+1}\ldots
3549: a_{i+j}}\nn\\
3550: &+&{\bar g}_{a_1\ldots a_i a a_{i+j+1}\ldots a_n}{\bar f}^a_{\,\,
3551: a_{i+1}\ldots a_{i+j}}]~~.
3552: \end{eqnarray}
3553: and:
3554: \be
3555: \label{c}
3556: c(\{f,g\}):={\bar f}_a{\bar g}^a={\bar f}^a{\bar g}_a~~,
3557: \ee
3558: where we lift coefficients with $\rho^{ab}$ as in (\ref{lifts}).
3559: \ep
3560:
3561: \noindent With our conventions, expression (\ref{c}) can be obtained by formally setting
3562: $n=0$ in equation (\ref{strict_bracket})) and dividing by two. The second equality in (\ref{c}) follows upon using the
3563: symmetry property of $\rho^{ab}$ and the selection rule ${\tilde a}={\tilde \omega}+1$.
3564:
3565:
3566: \begin{proof}
3567: By Proposition \ref{coord_bracket}, we have:
3568: \be
3569: \{f,g\}=\sum_{n\geq 0}\sum_{j=0}^n F^{(j)}_{(a_1\ldots a_n)}({a_1}\ldots {a_n})_c\nn
3570: \ee
3571: where:
3572: \be
3573: F^{(j)}_{a_1\ldots a_n}={\bar f}_{a_1\ldots a_jb}\omega^{ba}{\bar g}_{a a_{j+1}\ldots a_n}~~.\nn
3574: \ee
3575: The case $n=0$ can be checked directly, so we discuss only the case $n\geq 1$.
3576:
3577: \noindent If $i+j\leq n$, we compute:
3578: \be
3579: F^{(j)}_{a_{i+1}\ldots a_n, a_1\ldots a_i}={\bar f}_{a_{i+1}\ldots a_{i+j}b}\omega^{ba}
3580: {\bar g}_{aa_{i+j+1}\ldots a_na_1\ldots a_i}=
3581: (-1)^{\sigma_1} {\bar g}_{a_1\ldots a_i a a_{i+j+1}\ldots a_n}{\bar f}^a_{\,\, a_{i+1}\ldots a_{i+j}}~~,\nn
3582: \ee
3583: where we used cyclicity of $f$ and $g$ to permute indices. The sign factor is easily determined
3584: keeping in mind the symmetry property (\ref{oinv_symm}) and the selection rule (\ref{oinv_sel}):
3585: \be
3586: \sigma_1={\tilde \omega}({\tilde a}+1)+({\tilde a}_1+\ldots +{\tilde a}_i)({\tilde a}+
3587: {\tilde a}_{i+j+1}+\ldots +{\tilde a}_n)+({\tilde a}+{\tilde \omega})({\tilde a}_{i+1}+\ldots
3588: +{\tilde a}_{i+j})~~({\rm mod}~2)~~.\nn
3589: \ee
3590: For $i+j>n$, we find:
3591: \be
3592: F^{(j)}_{a_{i+1}\ldots a_n, a_1\ldots a_i}=
3593: (-1)^{{\tilde a}+{\tilde \omega}+1}{\bar f}_{a_{i+1}\ldots a_na_1\ldots a_{j-n+i}a}{\bar g}^a_{\,\,a_{j-n+i+1}\ldots a_i}=
3594: (-1)^{\sigma_2}{\bar f}_{a_1\ldots a_{j-n+i}aa_{i+1}\ldots a_n}{\bar g}^a_{\,\,a_{j-n+i+1}\ldots a_i}~~,\nn
3595: \ee
3596: with:
3597: \be
3598: \sigma_2={\tilde a}+{\tilde \omega}+1+
3599: ({\tilde a}_{i+1}+\ldots +{\tilde a}_n)({\tilde a}+{\tilde a}_1+\ldots +{\tilde a}_{j-n+i})
3600: ~~({\rm mod}~2)~~.\nn
3601: \ee
3602:
3603: \noindent This allows us to write:
3604: \be
3605: \label{two_sums}
3606: \sum_{j=0}^n F^{(j)}_{(a_1\ldots a_n)}=\frac{1}{n}\left[\sum_{j=0}^n\sum_{i=1}^{n-j}
3607: (-1)^{\epsilon_1} F^{(j)}_{a_{i+1}\ldots a_n, a_1\ldots a_i}+ \sum_{j=0}^n\sum_{i=n-j+1}^{n}
3608: (-1)^{\epsilon_2} F^{(j)}_{a_{i+1}\ldots a_n, a_1\ldots a_i}\right]\nn
3609: \ee
3610: where:
3611: \begin{eqnarray}
3612: \epsilon_1&=&{\tilde a}+{\tilde \omega}+
3613: ({\tilde a}+{\tilde \omega}+{\tilde a}_1+\ldots +{\tilde a}_i)({\tilde a}+{\tilde a}_{i+1}+\ldots
3614: +{\tilde a}_{i+j})\nn\\
3615: \epsilon_2&=& 1+{\tilde a}+{\tilde \omega}+
3616: ({\tilde a}_{i+1}+\ldots +{\tilde a}_n)({\tilde a}+{\tilde a}_{j-n+i+1}+\ldots +{\tilde a}_i)~~.\nn
3617: \end{eqnarray}
3618: We now perform the replacement $i\rightarrow i'=i+j-n, j\rightarrow j'=n-j$ in the second
3619: double sum. Denoting the new summation indices $i',j'$ by $i,j$, this gives:
3620: \be
3621: \sum_{j=0}^n F^{(j)}_{(a_1\ldots a_n)}=\frac{1}{n}\sum_{j=0}^n\sum_{i=1}^{n-j}\left[(-1)^{\epsilon_1}
3622: {\bar g}_{a_1\ldots a_i a a_{i+j+1}\ldots a_n}{\bar f}^a_{\,\, a_{i+1}\ldots a_{i+j}}+(-1)^{\tilde \epsilon_2}
3623: {\bar f}_{a_1\ldots a_i a a_{i+j+1}\ldots a_n}{\bar g}^a_{\,\, a_{i+1}\ldots a_{i+j}}\right]~~,\nn
3624: \ee
3625: where:
3626: \be
3627: {\tilde \epsilon}_2=1+{\tilde a}+{\tilde \omega}+({\tilde a}+{\tilde a}_{i+1}+\ldots +{\tilde a}_{i+j})({\tilde a}_{i+j+1}+
3628: \ldots +{\tilde a}_n)~~.\nn
3629: \ee
3630: The next step is to notice that the first term in square brackets vanishes unless:
3631: \be
3632: {\tilde a}+{\tilde a}_1+\ldots +{\tilde a}_i={\tilde g}+{\tilde
3633: a}_{i+j+1}+\ldots +{\tilde a}_n ~~({\rm mod}~2)\nn
3634: \ee
3635: and:
3636: \be
3637: {\tilde a}+{\tilde a}_{i+1}+\ldots +{\tilde a}_{i+j}={\tilde f}+{\tilde \omega}~~
3638: ~~({\rm mod}~2)~~,\nn
3639: \ee
3640: which allows us to replace $\epsilon_1$ by:
3641: \be
3642: \epsilon_1'={\tilde a}+{\tilde \omega}+
3643: ({\tilde f}+{\tilde \omega})({\tilde g}+{\tilde \omega}+{\tilde a}_{i+j+1}+\ldots +
3644: {\tilde a}_n)~~.\nn
3645: \ee
3646: Similarly, the second term in the square brackets vanishes unless:
3647: \be
3648: {\tilde a}+{\tilde a}_{i+1}+\ldots +{\tilde a}_{i+j}={\tilde g}+{\tilde \omega}~~,\nn
3649: \ee
3650: which allows us to replace ${\tilde \epsilon}_2$ by:
3651: \be
3652: \epsilon_2'={\tilde a}+ {\tilde \omega}+1+({\tilde g}+{\tilde \omega})({\tilde a}_{i+j+1}+\ldots +
3653: {\tilde a}_n)~~.\nn
3654: \ee
3655: Finally, using the assumption ${\tilde f}={\tilde g}={\tilde \omega}+1$, we find:
3656: \be
3657: \epsilon'_1=\epsilon_2'=1+{\tilde \omega}+{\tilde a}+{\tilde a}_{i+j+1}+\ldots +{\tilde a}_n={\tilde a}_1+\ldots +
3658: {\tilde a}_i~~({\rm mod}~2)~~,\nn
3659: \ee
3660: where we used the selection rule ${\tilde a}_1+\ldots +{\tilde a}_i+{\tilde a}+
3661: {\tilde a}_{i+j+1}+\ldots
3662: +{\tilde a}_n={\tilde f}={\tilde g}={\tilde \omega}+1$.
3663:
3664: \end{proof}
3665:
3666: \paragraph{Corollary}
3667: \label{maincor}
3668: Let $f=c(f)+\sum_{n\geq 0}{\frac{{\bar f}_{a_1\ldots a_n}}{n}({a_1}\ldots {a_n}})_c$ be a
3669: cyclic element of $A$ of degree
3670: ${\tilde \omega}+1$. Then:
3671: \be
3672: \frac{1}{2}\{f,f\}=\frac{1}{2}{\bar f}_a{\bar f}^a+
3673: \sum_{n\geq 1}\frac{1}{n}\left(\sum_{0\leq i+j\leq n}(-1)^{{\tilde a_1}+\ldots +{\tilde a}_i}
3674: {\bar f}_{a_1\ldots a_i a a_{i+j+1}\ldots a_n} {\bar f}^a_{\,\,a_{i+1}\ldots
3675: a_{i+j}}\right)({a_1}\ldots {a_n})_c~~.\nn
3676: \ee
3677:
3678:
3679:
3680:
3681: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3682: \begin{thebibliography}{100}
3683:
3684: %%%%%%%%%%%%%%%%%%%%%%%%% for bibtex
3685:
3686: \bibitem{Witten_mirror}{E.~Witten,
3687: ``Mirror manifolds and topological field theory,'', in {\em Essays on mirror
3688: manifolds}, pp 120--158, Internat. Press, Hong Kong, 1992 [arXiv:hep-th/9112056].}
3689: %%CITATION = HEP-TH 9112056;%%
3690: \bibitem{Dubrovin}{ B.~Dubrovin,``Geometry of 2-D topological field
3691: theories,'' LNM {\bf 1620}, Springer, 1996 [arXiv:hep-th/9407018].}
3692: %%CITATION = HEP-TH 9407018;%%
3693: \bibitem{WDVV}{R.~Dijkgraaf, H.~Verlinde and E.~Verlinde,
3694: ``Topological Strings In D $<$ 1,'' Nucl.\ Phys.\ B {\bf
3695: 352} (1991) 59.}
3696: %%CITATION = NUPHA,B352,59;%%
3697: \bibitem{BarKon}{ S.~Barannikov, M.~Kontsevich, ``Frobenius
3698: Manifolds and Formality of Lie Algebras of Polyvector Fields'', Internat. Math. Res. Notices {\bf 4}
3699: (1998) 201-215 [alg-geom/9710032]}
3700: \bibitem{Tian}{G. Tian , ``Smoothness of the Universal Deformation Space of
3701: Compact Calabi-Yau Manifolds and its Petersson-Weil Metric,'' in {\it Mathematical
3702: Aspects of String Theory,} 629-646, ed. S.-T. Yau, World Scientific, Singapore,
3703: 1987.}
3704: \bibitem{Todorov}{ A. Todorov, ``The Weil-Petersson Geometry of the Moduli Space of
3705: $SU(n\ge 3)$ (Calabi-Yau) Manifolds I,'' {\it Commun. Math. Phys} {\bf 126} (1989) 325-346.}
3706: \bibitem{HLL}{M.~Herbst, C.~I.~Lazaroiu and W.~Lerche,
3707: ``Superpotentials, A(infinity) relations and WDVV equations for open
3708: topological strings,'' JHEP {\bf 0502} (2005) 071 [arXiv:hep-th/0402110].}
3709: %%CITATION = HEP-TH 0402110;%%
3710: \bibitem{CIL5}{C.~I.~Lazaroiu,
3711: ``D-brane categories,'' Int.\ J.\ Mod.\ Phys.\ A {\bf 18} (2003) 5299
3712: [arXiv:hep-th/0305095].}
3713: \bibitem{Costello1}{K. Costello, ``The A-infinity operad and the moduli space of curves'',
3714: math.AG/0402015.}
3715: \bibitem{Costello2}{K. Costello, ``Topological conformal field theories and Calabi-Yau categories'',
3716: math.QA/0412149.}
3717: %%CITATION = HEP-TH 0305095;%%
3718: \bibitem{Stasheff_Kajiura}{H.~Kajiura, J.~Stasheff,
3719: ``Homotopy algebras inspired by classical open-closed string field theory'',
3720: math.QA/0410291.}
3721: \bibitem{Kajiura}{H. Kajiura, ``Noncommutative homotopy algebras
3722: associated with open strings'', Ph.D. thesis, math.QA/0306332.}
3723: \bibitem{Gaberdiel}{M.~R.~Gaberdiel and B.~Zwiebach,
3724: ``Tensor constructions of open string theories I: Foundations,''
3725: Nucl.\ Phys.\ B {\bf 505} (1997) 569 [arXiv:hep-th/9705038].}
3726: %%CITATION = HEP-TH 9705038;%%
3727: \bibitem{Konts_formal}{M. Kontsevich, ``Formal (non)commutative
3728: symplectic geometry.'' , {\em The Gelfand Mathematical Seminars, 1990--1992},
3729: 173--187, Birkh\"auser Boston, Boston, MA, 1993.}
3730: \bibitem{LB_lectures}{L. LeBruyn, ``Noncommutative~geometry $@n$'',
3731: unpublished book, available at $www.neverendingbooks.org$; see also
3732: ``Three~talks on noncommutative geometry$@n$'', math.RA/0312221.}
3733: \bibitem{Ginzburg_lectures}{V. Ginzburg, ``Lectures on Noncommutative Geometry'',
3734: math.AG/0506603.}
3735: \bibitem{HMS}{M.~Kontsevich, ``Homological algebra of mirror symmetry'', Proc.
3736: Internat. Congr. Math., Z\"urich, Switzerland 1994 (Basel), vol.~1,
3737: Birkh\"auser Verlag, 1995, 120--139.}
3738: \bibitem{Polishchuk}{A. Polishchuk, ``Homological mirror symmetry
3739: with higher products'', {\em Winter School on Mirror Symmetry, Vector Bundles and
3740: Lagrangian Submanifolds} (Cambridge, Massachusetts, 1999) in
3741: AMS/IP Stud. Adv. Math. {\bf 23} pp.~247--259 [math.AG/9901025].}
3742: \bibitem{RSS}{G.-C.~Rota, B.~Sagan, P.~R. Stein, ``A cyclic
3743: derivative in noncommutative algebra'', J. of Algebra {\bf 64}
3744: (1980) 54-75.}
3745: \bibitem{Voiculescu}{D.~Voiculescu, ``A note on cyclic gradients'',
3746: Indiana Univ. Math. J {\bf 49} (2000) 3, 837-841.}
3747: \bibitem{BEV}{ W. Crawley-Boevey, P. Etingof, V. Ginzburg,
3748: ``Noncommutative Geometry and Quiver algebras'', math.AG/0502301}
3749: \bibitem{Bergh}{M. Van den Bergh, ``Double Poisson algebras'', math.QA/0410528.}
3750: \bibitem{CIL1}{C.~I.~Lazaroiu,
3751: ``On the structure of open-closed topological field theory in two dimensions,''
3752: Nucl.\ Phys.\ B {\bf 603} (2001) 497 [arXiv:hep-th/0010269].}
3753: %%CITATION = HEP-TH 0010269;%%
3754: \bibitem{CQ}{J. Cuntz, D. Quillen, ``Algebra extensions and
3755: nonsingularity'', Journal AMS {\bf 8} (1995) 251-289.}
3756: \bibitem{Ginzburg}{ V. Ginzburg, ``Noncommutative symplectic
3757: geometry, quiver varieties and operads'', Math. Res. Lett {\bf 8} (2001) 3, pp
3758: 377-400.}
3759: \bibitem{LB_quivers}{ R. Bocklandt, L. Le Bruyn, ``Necklace Lie
3760: algebras and noncommutative symplectic geometry'', Math. Z. {\bf 240} (2002) 141-167 [math.AG/0010030]}
3761: \bibitem{Katz}{P.~S.~Aspinwall and S.~Katz, ``Computation of superpotentials for D-Branes,''
3762: arXiv:hep-th/0412209.}
3763: %%CITATION = HEP-TH 0412209;%%
3764: \bibitem{CIL2}{C.~I.~Lazaroiu,
3765: ``Generalized complexes and string field theory,''
3766: JHEP {\bf 0106} (2001) 052 [arXiv:hep-th/0102122].}
3767: %%CITATION = HEP-TH 0102122;%%
3768: \bibitem{CIL3}{C.~I.~Lazaroiu,
3769: ``Unitarity, D-brane dynamics and D-brane categories,''
3770: JHEP {\bf 0112} (2001) 031 [arXiv:hep-th/0102183].}
3771: %%CITATION = HEP-TH 0102183;%%
3772: \bibitem{Konts_Schwarz}{ M. Alexandrov, M. Kontsevich, A. Schwarz, O. Zaboronsky,
3773: ``The Geometry of the Master Equation and Topological Quantum Field
3774: Theory'', Int. J. Mod.Phys. {\bf A12} (1997) 1405-1430 [hep-th/9502010].}
3775: \bibitem{Penkava}{M. Penkava,
3776: ``Infinity Algebras, Cohomology and Cyclic Cohomology, and Infinitesimal Deformations'',
3777: math.QA/0111088.}
3778: \bibitem{Lazarev}{ A. Hamilton, A. Lazarev, ``Homotopy algebras and
3779: noncommutative geometry'', math.QA/0410621.}
3780: \bibitem{Umirbaev}{U. U. Umirbaev,
3781: ``Tame and wild automorphisms of polynomial algebras and free associative
3782: algebras'', Max-Planck Institute Bonn, Preprint MPIM2004-108.}
3783: \bibitem{DY}{ V. Drensky, J.-T. Yu, `` The strong Anick conjecture is true'',
3784: math.RA/0507170.}
3785: \bibitem{Cz}{A.J. Czerniakiewicz, ``Automorphisms of a
3786: free associative algebra of rank 2. I, II'',
3787: Trans. Amer. Math. Soc. {\bf 160} (1971) 393-401;
3788: {\bf 171} (1972) 309-315.}
3789: \bibitem{ML}{L.G. Makar-Limanov,
3790: ``On automorphisms of free algebra with two generators'', Functional Anal. Appl. {\bf 4} (1970), 262-263.}
3791: \bibitem{SW}{N.~Seiberg, E.~Witten, ``String theory and
3792: noncommutative geometry,'' JHEP {\bf 9909} (1999) 032 [arXiv:hep-th/9908142].}
3793: %%CITATION = HEP-TH 9908142;%%
3794: \bibitem{Schomerus}{V.~Schomerus,``D-branes and deformation
3795: quantization,'' JHEP {\bf 9906} (1999) 030 [arXiv:hep-th/9903205].}
3796: %%CITATION = HEP-TH 9903205;%%
3797: \bibitem{Moore_Segal}{G. Moore and G. Segal, unpublished; see http://online.kitp.ucsb.edu/online/mp01/}
3798: \bibitem{Moore}{G.~W.~Moore,
3799: ``Some comments on branes, G-flux, and K-theory,''
3800: Int.\ J.\ Mod.\ Phys.\ A {\bf 16} (2001) 936 [arXiv:hep-th/0012007].}
3801: %%CITATION = HEP-TH 0012007;%%
3802: \bibitem{Fukaya_mirror}{
3803: K.~Fukaya, ``Floer homology and mirror symmetry. II'', in {\em Minimal
3804: surfaces, geometric analysis and symplectic geometry} (Baltimore, MD,
3805: 1999), Adv. Stud. Pure Math. {\bf 34}, Math. Soc. Japan, Tokyo,
3806: 2002, pp.~31--127.}
3807: \bibitem{KS}{M.~Kontsevich, Y.~Soibelman,
3808: {\em Homological mirror symmetry and torus fibrations}, math.SG/0011041.}
3809: \bibitem{CIL4}{C.~I.~Lazaroiu, ``String field theory and brane superpotentials,'' JHEP {\bf 0110} (2001) 018
3810: [arXiv:hep-th/0107162].}
3811: %%CITATION = HEP-TH 0107162;%%
3812: \bibitem{CIL8}{C.~I.~Lazaroiu and R.~Roiban,
3813: ``Holomorphic potentials for graded D-branes,''
3814: JHEP {\bf 0202} (2002) 038 [arXiv:hep-th/0110288].}
3815: %%CITATION = HEP-TH 0110288;%%
3816: \bibitem{CIL9}{C.~I.~Lazaroiu,
3817: ``An analytic torsion for graded D-branes,''
3818: JHEP {\bf 0209} (2002) 023 [arXiv:hep-th/0111239].}
3819: %%CITATION = HEP-TH 0111239;%%
3820: \bibitem{CIL10}{C.~I.~Lazaroiu and R.~Roiban,
3821: ``Gauge-fixing, semiclassical approximation and potentials for graded
3822: Chern-Simons theories,'' JHEP {\bf 0203} (2002) 022 [arXiv:hep-th/0112029].}
3823: %%CITATION = HEP-TH 0112029;%%
3824: \bibitem{CK}{R. Coquereaux, D. Kastler, ``Remarks on the
3825: differential envelopes of associative algebras'', Pacific. J. Math {\bf 137} (1989) 2.}
3826: \bibitem{Fukaya_inf}{
3827: K.~Fukaya, ``Morse homotopy, $A_\infty$-category, and Floer
3828: homologies'', Proc. of GARC Workshop on Geometry and Topology '93
3829: (H.~J. Kim, ed.), Lecture Notes {\bf 18}, Seoul Nat. Univ., Seoul,
3830: 1993, pp.~1--102.}
3831: \bibitem{Keller}{B.~Keller, ``Introduction to {A}-infinity algebras and modules'',
3832: Homology, Homotopy and Applications {\bf 3} (2001)1, pp 1--35 [math.RA/9910179]}
3833: \bibitem{Hasegawa}{ K. Lefèvre-Hasegawa ``Sur les A-infini cat\'egories'', Ph.D. Thesis,
3834: math.CT/0310337.}
3835: \bibitem{Lyubashenko1}{V.~V. Lyubashenko, ``Category of $A_\infty$-categories'',
3836: Homology, Homotopy and Applications \textbf{5} (2003) 1,pp. 1--48 [math.CT/0210047] }
3837: \bibitem{Lyubashenko2}{ V. Lyubashenko, S. Ovsienko, ``A construction of an A-infinity category'',
3838: math.CT/0211037.}
3839: \bibitem{Lyubashenko3}{V. Lyubashenko, O. Manzyuk, ``Quotients of unital ${A}_\infty$-categories'',
3840: math.CT/0306018.}
3841: \bibitem{Lyubashenko4}{ V. Lyubashenko, O. Manzyuk, ``Free $A_\infty$-categories'',
3842: math.CT/0312339.}
3843: \bibitem{Stasheff}{J.~D. Stasheff, ``Homotopy associativity of H-spaces, I $\&$ II'',
3844: Trans. Amer. Math. Soc. {\bf 108} (1963), pp. 275--292 and 293--312.}
3845: \bibitem{KS_book}{M. Kontsevich, Y. Soibelman, book in preparation}
3846: \bibitem{Fukaya_book}{K. Fukaya, Y.G. Oh, H. Ohta, K. Ono,
3847: ``Lagrangin intersection Floer theory - anomaly and obstructon'', preprint. }
3848: \bibitem{CIL6}{C.~I.~Lazaroiu,
3849: ``Graded Lagrangians, exotic topological D-branes and enhanced triangulated
3850: categories,''JHEP {\bf 0106} (2001) 064 [arXiv:hep-th/0105063].}
3851: %%CITATION = HEP-TH 0105063;%%
3852: \bibitem{CIL7}{C.~I.~Lazaroiu, R.~Roiban and D.~Vaman,
3853: ``Graded Chern-Simons field theory and graded topological D-branes,''
3854: JHEP {\bf 0204} (2002)023 [arXiv:hep-th/0107063].}
3855: %%CITATION = HEP-TH 0107063;%%
3856: \bibitem{Witten_CS}{
3857: E.~Witten, ``Chern-Simons gauge theory as a string theory'',
3858: The Floer memorial volume, 637--678, Progr. Math. {\bf 133}, Birkhauser, Basel,
3859: 1995 [hep-th/9207094].}
3860: \bibitem{VG}{V. Ginzburg, ``Double derivations and Cyclic homology'', math.KT/0505236.}
3861: \end{thebibliography}
3862: \end{document}
3863:
3864: