hep-th0111239/rs.tex
1: \newif\iffigs\figstrue
2: %  Uncomment the next line if you don't want the figures:
3: %\figsfalse
4: 
5: \documentclass[paper, 12pt, letterpaper, epsf]{JHEP}
6: \input{epsf.tex}
7: \usepackage{graphics}
8: \def\Bbb{\bf}
9: \def\C{{\Bbb C}}
10: \def\R{{\Bbb R}}
11: \def\Z{{\Bbb Z}}
12: \def\H{{\Bbb H}}
13: 
14: 
15: \def\Hom{\operatorname{Hom}}
16: \def\Tors{\operatorname{Tors}}
17: \def\Ker{\operatorname{Ker}}
18: \def\Spec{\operatorname{Spec}}
19: \def\Area{\operatorname{Area}}
20: \def\Vol{\operatorname{Vol}}
21: \def\ad{\operatorname{ad}}
22: \def\tr{\operatorname{tr}}
23: \def\Pic{\operatorname{Pic}}
24: \def\disc{\operatorname{disc}}
25: \def\cpl{\operatorname{cpl}}
26: \def\Img{\operatorname{Im}}
27: \def\Rea{\operatorname{Re}}
28: \def\Gr{\operatorname{Gr}}
29: \def\SO{\operatorname{SO}}
30: \def\Sl{\operatorname{SL}}
31: \def\GO{\operatorname{O{}}}
32: \def\SU{\operatorname{SU}}
33: \def\GU{\operatorname{U{}}}
34: \def\Sp{\operatorname{Sp}}
35: \def\Spin{\operatorname{Spin}}
36: \def\rank{\operatorname{rank}}
37: \def\bearray{\begin{eqnarray}}
38: \def\eearray{\end{eqnarray}}
39: \def\bearraynn{\begin{eqnarray*}}
40: \def\eearraynn{\end{eqnarray*}}
41: \def\bfig{\begin{figure}}
42: \def\efig{\end{figure}}
43: \def\Aff{\operatorname{Aff}}
44: \def\diag{\operatorname{diag}}
45: \def\opeq#1{\advance\lineskip#1 \advance\baselineskip#1
46:         \advance\lineskiplimit#1}
47: \def\eqalignsq#1{\null\,\vcenter{\opeq{2.5\jot}\mathsurround=0pt
48:         \everycr={}\tabskip=0pt\offinterlineskip
49:         \halign{\strut\hfil$\displaystyle{##}$&$\displaystyle{{}##}$\hfil
50:         \crcr#1\crcr}}\,\null}
51: \def\eqalign#1{\null\,\vcenter{\opeq{2.5\jot}\mathsurround=0pt
52:         \everycr={}\tabskip=0pt
53:         \halign{\strut\hfil$\displaystyle{##}$&$\displaystyle{{}##}$\hfil
54:         \crcr#1\crcr}}\,\null}
55: 
56: \def\sm{$\sigma$-model}
57: \def\nlsm{non-linear \sm}
58: \def\smm{\sm\ measure}
59: \def\CY{Calabi--Yau}
60: \def\LG{Landau-Ginzburg}
61: \def\cR{{\Scr R}}
62: \def\cM{{\Scr M}}
63: \def\cA{{\Scr A}}
64: \def\cB{{\Scr B}}
65: \def\cK{{\Scr K}}
66: \def\cD{{\Scr D}}
67: \def\cH{{\Scr H}}
68: \def\cT{{\Scr T}}
69: \def\cL{{\Scr L}}
70: \def\cF{{\Scr F}}
71: 
72: 
73: \def\spnh{\Spin(32)/\Z_2}
74: 
75: \def\Pf{{\em Proof: }}
76: 
77: \def\ker{{\rm ker}}
78: \def\coker{{\rm coker}}
79: \def\rank{{\rm rank}}
80: \def\im{{\rm im}}
81: \def\dim{{\rm dim}}
82: \def\codim{{\rm codim}}
83: \def\Card{{\rm Card}}
84: \def\li{{\rm linearly independent}}
85: \def\ld{{\rm linearly dependent}}
86: \def\deg{{\rm deg}}
87: \def\det{{\rm det}}
88: \def\Div{{\rm Div}}
89: \def\supp{{\rm supp}}
90: \def\Gr{{\rm Gr}}
91: \def\End{{\rm End}}
92: \def\Aut{{\rm Aut}}
93: 
94: 
95: \newtheorem{Proposition}{Proposition}[section]
96: \newtheorem{Definition}{Definition}[section]
97: \newtheorem{Theorem}{Theorem}[section]
98: \newtheorem{Lemma}{Lemma}[section]
99: \newtheorem{Corrolary}{Corrolary}[section]
100: 
101: 
102: 
103: \newcommand{\be}{\begin{equation}}
104: \newcommand{\ee}{\end{equation}}
105: \newcommand{\bea}{\begin{eqnarray}}
106: \newcommand{\eea}{\end{eqnarray}}
107: 
108: \newcommand{\bp}{\begin{Proposition}}
109: \newcommand{\ep}{\end{Proposition}}
110: \newcommand{\bt}{\begin{Theorem}}
111: \newcommand{\et}{\end{Theorem}}
112: \newcommand{\bl}{\begin{Lemma}}
113: \newcommand{\el}{\end{Lemma}}
114: \newcommand{\bc}{\begin{Corrolary}}
115: \newcommand{\ec}{\end{Corrolary}}
116: \newcommand{\nn}{\nonumber}
117: 
118: 
119: 
120: \font\mybbb=msbm10 at 8pt
121: \font\mybb=msbm10 at 12pt
122: \def\bbb#1{\hbox{\mybbb#1}}
123: \def\bb#1{\hbox{\mybb#1}}
124: \def\pRe{\bbb{R}}
125: \def\Re {\bb{R}}
126: \def\P{\bb{P}}
127: \def\Z {\bb{Z}}
128: \def\q{\bb{a}}
129: \def\id{\protect{{1 \kern-.28em {\rm l}}}}
130: \def\cn{{\cal N}}
131: 
132: \def\p#1{{\phi{}^{(#1)}}}
133: \def\hp#1{{{\phi'}{}^{(#1)}}}
134: \def\tp#1{{{\tilde\phi}{}^{(#1)}}}
135: \def\c#1{{c{}^{(#1)}}}
136: \def\hc#1{{{c'}{}^{(#1)}}}
137: \def\tc#1{{{\tilde c}{}^{(#1)}}}
138: 
139: %\def\boldphi{{\phi\hspace{-6.62pt}\phi\hspace{-6.62pt}\phi}}
140: \def\boldphi{\mbox{\boldmath $\phi$}}
141: \def\boldpsi{\mbox{\boldmath $\psi$}}
142: 
143: %%\boldwe is used in the definition of \bwe. It can also be used separately. 
144: \def\boldwe{\mbox{\boldmath $\wedge$}}
145: \def\bwe{{{{\boldwe\hspace{-8.9pt}\boldwe}\hspace{-8.8pt}\boldwe}
146: \hspace{-8.8pt}\boldwe}}
147: \def\bweft{{{{\boldwe\hspace{-8.1pt}\boldwe}\hspace{-8pt}\boldwe}
148: \hspace{-8pt}\boldwe}}
149: 
150: 
151: \def\pb#1{{{\boldphi}{}^{(#1)}}}
152: \def\hpb#1{{{\bf\boldphi'}{}^{(#1)}}}
153: \def\tpb#1{{{\bf\tilde\boldphi}{}^{(#1)}}}
154: \def\cb#1{{{\bf c}{}^{(#1)}}}
155: \def\hcb#1{{{\bf c'}{}^{(#1)}}}
156: \def\tcb#1{{{\bf\tilde c}{}^{(#1)}}}
157: 
158: 
159: \font\gothics=ygoth at 10pt
160: \font\gothicl=ygoth at 12pt
161: \def\gg#1{\hbox{\gothics#1}}
162: 
163: \def\gG{\hbox{\gothicl G}}
164: 
165: \font\frakl=yfrak at 12pt
166: \font\fraks=yfrak at 10pt
167: 
168: \def\fG{\hbox{\frakl G}}
169: 
170: 
171: 
172: 
173: %EOF
174: 
175: 
176: 
177: 
178: \newcommand{\rf}[1]{(\ref{#1})}
179: \renewcommand{\theequation}{\thesection.\arabic{equation}}
180: %\renewcommand{\thefootnote}{\fnsymbol{footnote}}
181: 
182: %\def\appendix#1{
183:  % \addtocounter{section}{1} \setcounter{equation}{0}
184:   %\renewcommand{\thesection}{\Alph{section}} \section*{Appendix}
185:   %\addcontentsline{toc}{section}{Appendix \thesection\ \ \ #1} }
186: 
187: \newcommand{\newsection}{    % Numeration of eqs. is automatic
188: \setcounter{equation}{0}
189: \section}
190: \newcommand{\non}{\nonumber \\*}
191: 
192: \newcommand{\eq}[1]{Eq.~(\ref{#1})}
193: \newcommand{\nonu}{\nonumber}
194: \def \ov {\over }
195: \def\bea{\begin{eqnarray}}
196: \def\eea{\end{eqnarray}}
197: \def \ci {\cite}
198: \def \de{\partial}
199: \def \x{{\bf x}}
200: \def\y{{\bf y}}
201: \def\k{{\bf k}}
202: \def\LB{\left(}
203: \def\RB{\right)}
204: \def\be{\begin{equation}}
205: \def\ee{\end{equation}}
206: \def\ba{\begin{eqnarray}}
207: \def\ea{\end{eqnarray}}
208: \def\la{\label}
209: \def \bi{\bibitem}
210: \def \Tr {{\rm tr}}
211: \def\O{{\cal O}}
212: \def\LA{\langle}
213: \def\RA{\rangle}
214: \def\X{{\bar X}}
215: \def\a{\alpha}
216: \def\b{\beta}
217: \def\e{\epsilon}
218: \def\A{{\cal A}}
219: \def\n{{\nabla}}
220: \def\G{{\Gamma}}
221: \def\1{{{(1)}}}\def\2{{{(2)}}}\def\3{{{(3)}}}  
222: 
223: 
224: 
225: \font\mybbb=msbm10 at 8pt
226: \font\mybb=msbm10 at 12pt
227: \def\bbb#1{\hbox{\mybbb#1}}
228: \def\bb#1{\hbox{\mybb#1}}
229: \def\pRe{\bbb{R}}
230: \def\Re {\bb{R}}
231: \def\Z {\bb{Z}}
232: \def\q{\bb{a}}
233: \def\id{\protect{{1 \kern-.28em {\rm l}}}}
234: \def\cn{{\cal N}}
235: 
236: \def\p#1{{\phi{}^{(#1)}}}
237: \def\hp#1{{{\phi'}{}^{(#1)}}}
238: \def\tp#1{{{\tilde\phi}{}^{(#1)}}}
239: \def\c#1{{c{}^{(#1)}}}
240: \def\hc#1{{{c'}{}^{(#1)}}}
241: \def\tc#1{{{\tilde c}{}^{(#1)}}}
242: 
243: %\def\boldphi{{\phi\hspace{-6.62pt}\phi\hspace{-6.62pt}\phi}}
244: \def\boldphi{\mbox{\boldmath $\phi$}}
245: 
246: 
247: %%\boldwe is used in the definition of \bwe. It can also be used separately. 
248: \def\boldwe{\mbox{\boldmath $\wedge$}}
249: \def\bwe{{{{\boldwe\hspace{-8.9pt}\boldwe}\hspace{-8.8pt}\boldwe}
250: \hspace{-8.8pt}\boldwe}}
251: \def\bweft{{{{\boldwe\hspace{-8.1pt}\boldwe}\hspace{-8pt}\boldwe}
252: \hspace{-8pt}\boldwe}}
253: 
254: 
255: \def\pb#1{{{\boldphi}{}^{(#1)}}}
256: \def\hpb#1{{{\bf\boldphi'}{}^{(#1)}}}
257: \def\tpb#1{{{\bf\tilde\boldphi}{}^{(#1)}}}
258: \def\cb#1{{{\bf c}{}^{(#1)}}}
259: \def\hcb#1{{{\bf c'}{}^{(#1)}}}
260: \def\tcb#1{{{\bf\tilde c}{}^{(#1)}}}
261: 
262: 
263: \def \bs{\star^\dagger}
264: 
265: 
266: 
267: \usepackage{graphics}
268: 
269: 
270: \title{An analytic torsion for graded D-branes}
271: 
272: \author{C.~I.~Lazaroiu
273: \\C.~N.~Yang Institute for Theoretical Physics\\
274: SUNY at Stony BrookNY11794-3840,
275: U.S.A.\\calin@insti.physics.sunysb.edu}
276: 
277: 
278: \abstract{I consider the semiclassical approximation of the 
279: graded Chern-Simons field theories describing certain systems of topological A
280: type branes in the large radius limit of Calabi-Yau compactifications.
281: I show that the semiclassical 
282: partition function can be expressed in terms of a 
283: certain (differential) numerical invariant which is 
284: a version of the analytic torsion of Ray and Singer, 
285: but associated with flat graded superbundles. I also discuss a `twisted' version 
286: of the Ray-Singer norm, and show its independence of metric data.
287: As illustration, I consider
288: graded D-brane pairs of unit relative grade with a scalar condensate in the 
289: boundary condition changing sector. 
290: For the particularly simple case when the reference flat connections
291: are trivial, I show that the generalized torsion reduces to a 
292: power of the classical Ray-Singer invariant of the base 3-manifold. }
293: 
294: \preprint{YITP-SB 01-62}
295: 
296: \begin{document}
297: 
298: 
299: \tableofcontents
300: 
301: \pagebreak
302: \vskip .6in
303: 
304: 
305: \section{Introduction}
306: 
307: D-brane composite formation is a subject of central importance for 
308: a deeper understanding of open string theory dynamics. Of particular 
309: interest is the incarnation of such processes for the case of superstring 
310: compactifications on Calabi-Yau backgrounds, which provide a rich source 
311: of potential phenomenological applications. 
312: 
313: A systematic study of D-brane composites 
314: is best performed with the tools of string field theory. Indeed, such 
315: processes involve {\em off-shell} string dynamics, which is best captured
316: in string field language. 
317: Given the complexity superstring models, 
318: a satisfactory formulation of superstring field theory is 
319: lacking Calabi-Yau backgrounds. However, part of the dynamics of 
320: spacetime fields originating from the chiral primary sector allows for an 
321: explicit  description in terms of topological strings\cite{Witten_nlsm,
322: Witten_mirror, Witten_CS}. 
323: 
324: A fundamental observation made in \cite{Douglas_Kontsevich}(see also 
325: \cite{Zaslow_Polishchuk} and \cite{Seidel}) is that 
326: D-branes in Calabi-Yau backgrounds are graded objects. In our context,
327: this means that topological D-branes will 
328: carry certain integral data which specify a branch of their 
329: `BPS grade', which was discussed at length in \cite{Douglas_Kontsevich}.
330: It was recently proposed \cite{com1, com3, Diaconescu, sc, bv} that the 
331: dynamics of graded topological D-branes is described
332: by certain target space theories which have the form of  
333: `graded Chern-Simons models'\footnote{For the A-model, these 
334: theories do not take into account worldsheet instanton corrections
335: (which can be formally 
336: incorporated along the lines of \cite{Fukaya, Fukaya2}).
337: In the present paper, we always work in the large radius limit
338: of a given Calabi-Yau compactification, where such corrections 
339: can be neglected.}. These are versions of Chern-Simons field 
340: theories associated with graded superbundles, and whose fields assemble into a 
341: `graded superconnection of total degree one'.
342: This approach, which is intimately connected with the derived category 
343: program of \cite{Kontsevich} and \cite{Douglas_Kontsevich, Aspinwall, 
344: Douglas_Aspinwall}, allows  for 
345: a description of topological D-brane physics in standard field theoretic terms.
346: It also leads to a physical representation of the extended moduli space of 
347: open strings \cite{bv, gauge}. 
348: 
349: One benefit of this description is that the formal analogy between graded 
350: and ungraded Chern-Simons theories suggests a wealth of generalizations 
351: of classical connections between physics and certain mathematical invariants. 
352: In this note, we take a first step in this direction for the graded 
353: Chern-Simons theory of A-type branes.
354: 
355: It is well-known that the semiclassical approximation of 
356: usual Chern-Simons theory (which forms the string 
357: field theory of single, ungraded topological A-branes) is 
358: related to the analytic torsion of Ray and Singer. 
359: In fact, the Chern -Simons approach 
360: allows one to `discover' the torsion from physical considerations. 
361: It is natural to ask what is the analogue of this relation for 
362: graded Chern-Simons models. Proceeding in physical manner, we  
363: consider the semiclassical approximation of graded Chern-Simons theories, 
364: from which we extract a generalized version of Ray-Singer torsion, 
365: which should prove useful for a topological characterization of 
366: D-brane composites in Calabi-Yau backgrounds.
367: 
368: 
369: The note is organized as follows. In Section 2, we give a short review 
370: of the graded Chern-Simons theory of \cite{sc, bv} and 
371: recall the D-brane interpretation of its vacuum configurations. 
372: In Section 3, we consider the semiclassical approximation 
373: to these models around a general background. 
374: Upon using the methods of \cite{ST, Schwarz_resolvent, Adams_Sen} and 
375: related work, we evaluate the partition function in this limit 
376: up to a real prefactor related to the isotropy subgroup 
377: of the background. This allows us to express the result in terms of 
378: a certain generalization of the analytic torsion of Ray and Singer 
379: \cite{Ray, RS1, RS_symp}. This numerical invariant 
380: is related to (but does not seem to coincide with) 
381: a quantity considered in \cite{Bismut_Lott}. Upon writing the semiclassical 
382: partition function in terms of a `twisted Ray-Singer metric', 
383: we use general results of \cite{Schwarz_resolvent, Adams_Sen}, 
384: in order to show that the extended Ray-Singer norm 
385: is independent of the metric data employed in the gauge-fixing 
386: procedure.
387: We also discuss the argument of the semiclassical partition function, 
388: which displays a metric dependence reminiscent of that 
389: known from usual Chern-Simons theory. 
390: Section 4 considers the generalized 
391: torsion of  graded D-brane pairs
392: (of unit relative grade) in the presence of a scalar condensate. 
393: For the case of trivial underlying flat connections, we show that 
394: the generalized torsion reduces to a power of the classical 
395: Ray-Singer invariant of the three-cycle. 
396: 
397: \
398: 
399: \noindent{\bf Conventions} There exist a few different conventions 
400: for the Ray-Singer torsion in the literature. In this note, we
401: define the Ray-Singer torsion
402: of a closed three-manifold $L$ (with respect to the trivial flat line bundle) 
403: by:
404: \be
405: \label{conventions}
406: T(L)=\prod_q{det^{',reg}(\Delta_q)^{\frac{q(-1)^q}{2}}}~~,
407: \ee
408: where $\Delta_q$ is the $q$-form Laplacian and $det^{',reg}(\Delta_q)$ is 
409: the regularized determinant of the restriction to the orthocomplement of its 
410: kernel.
411: Other papers define the Ray-Singer torsion to be $T(L)^{-1}, T(L)^{2}$ or 
412: $T(L)^{-2}$, where $T(L)$ is the quantity in (\ref{conventions}).
413: Our conventions for the generalized torsion will be an extension 
414: of (\ref{conventions}). 
415: 
416: 
417: 
418: \section{Graded Chern-Simons theories and topological D-branes}
419: 
420: \subsection{The string field theory of graded topological A-branes}
421: 
422: Consider a special Lagrangian 3-cycle $L$ in a a Calabi-Yau threefold $X$, 
423: endowed with the `fundamental' orientation discussed in \cite{sc, gauge}. 
424: Given a collection of graded topological branes 
425: (of different grades $n$) wrapping $L$, 
426: we form the total bundle ${\bf E}=\oplus_{n}{E_n}$, endowed with the 
427: $\Z$-grading induced by $n$. Here $E_n$ are flat bundles which describe the 
428: worldvolume backgrounds of the various topological D-branes. 
429: Throughout this paper, we shall assume that the system contains a finite 
430: number of graded D-branes, i.e. $n$ takes values in a finite set of integers.
431: We also make the convention that a form on $L$ of rank lying outside 
432: the interval $0..3$ is defined to be zero. Note that we consider 
433: {\em complex} flat vector bundles $E_n$, which are not required to be unitary 
434: (i.e. there need not exist metrics on $E_n$ which are covariantly-constant 
435: with respect to the flat connections).
436: 
437: The graded Chern-Simons theory of \cite{sc, bv} describes sections  
438: $u$ of the bundle:
439: \be
440: {\cal V}=\Lambda^*(T^*L)\otimes End({\bf E})~~,~~
441: \ee
442: which we endow with the total grading 
443: ${\cal V}=\oplus_{t}{{\cal V}^t}$, where:
444: \be
445: {\cal V}^t=\oplus_{\tiny \begin{array}{c}k, m, n\\k+n-m=t\end{array}}
446: {\Lambda^k(T^*L)\otimes Hom(E_m, E_n)}~~.
447: \ee
448: The degree of $u$ with respect to this grading is:
449: \be
450: |u|=rk u + \Delta(u)~~,
451: \ee
452: where $\Delta(u)=n-m$ if $u\in \Omega^*(L, Hom(E_m,E_n))$. The grading $|.|$ 
453: is related (after localization) 
454: to the worldsheet $U(1)$ charge of topological string states.
455: The space ${\cal H}=\Gamma({\cal V})$ of sections of ${\cal V}$ is
456: {\em total boundary space} of \cite{sc}, and can be viewed as the 
457: collection of open string states of the system. 
458: It is endowed with the grading ${\cal H}^k=\Gamma({\cal V}^k)$.
459: 
460: Triple string interactions are described by the so-called 
461: {\em total boundary product}, which is defined through:
462: \be
463: \label{bullet}
464: u\bullet v= (-1)^{\Delta(u)rk v}u\wedge v~~,
465: \ee
466: where the wedge product on the right hand side includes composition 
467: of morphisms in $End({\bf E})$. This associative product is compatible 
468: with the grading and admits the identity endomorphism of 
469: ${\bf E}$ as a neutral element:
470: \be
471: \label{bullet_props}
472: |u\bullet v|=|u|+|v|~~, ~~1\bullet u=u\bullet 1=u~~.
473: \ee
474: (note that $|1|=0$). 
475: 
476: The direct sum $A^{(0)}=\oplus_n{A_n}$ of the flat connections on $E_n$
477: induces a flat structure on $End({\bf E})$. If $d^{(0)}$ 
478: is de Rham differential twisted by this flat connection,  
479: then it acts as a degree one derivation of the boundary product
480: (since the connection induced on $End({\bf E})$ 
481: is in the `adjoint representation').
482: To allow for more general backgrounds, we shift by elements 
483: $\phi\in {\cal H}^1$, which allows us to build the object 
484: $d=d^{(0)}+[\phi,.]_\bullet$, where $[.,.]_\bullet$ stands for the graded 
485: commutator\footnote{This is given by
486: $[u,v]_\bullet:=u\bullet v -(-1)^{|u||v|}v\bullet u$.}
487:  in the associative algebra $({\cal H}, \bullet)$. 
488: It is clear that $d$ is a degree one derivation of $({\cal H},\bullet)$
489: (since so are both terms in its definition): 
490: \be
491: \label{d_props}
492: |du|=|u|+1~~,~~d(u\bullet v)=(du)\bullet v+(-1)^{|u|}u\bullet (dv)~~.
493: \ee
494: In the language of 
495: \cite{Bismut_Lott}, $d$ is a `graded superconnection of total degree one'.
496: We refer the reader to \cite{Quillen} for basic facts regarding 
497: superconnections. 
498: For what follows, we shall pick a reference {\em flat}\footnote{Flatness
499: means that $d^2=0$, and will be required by the equations of motion.} 
500: graded superconnection $d$ (of degree one)
501: which need not coincide with the original flat connection $A^{(0)}$; 
502: the formalism is independent of this choice.
503: 
504: One also has a graded trace on $\Omega^*(L, End({\bf E}))$, which is defined through:
505: \be
506: \label{str}
507: str(u)=\sum_{n}{(-1)^n tr(u_{nn})}~~,~~{\rm for}~~u=\oplus_{m,n}{u_{mn}}~~,
508: \ee
509: with $u_{mn}\in \Omega^*(L, Hom(E_m, E_n))$. 
510: The bilinear form:
511: \be
512: \label{bf_A}
513: \langle u, v \rangle:=\int_{L}{str(u\bullet v)}~~ 
514: \ee
515: is non-degenerate and has the properties:
516: \bea
517: \label{form_invariance}
518: \langle u, v\rangle =(-1)^{|u||v|}\langle v, u \rangle~~,~~
519: \langle du, v\rangle +(-1)^{|u|}\langle u, dv\rangle=0~~,~~
520: \langle u\bullet v, w\rangle =\langle u, v\bullet w\rangle~~ 
521: \eea
522: It also obeys the selection rule $\langle u, v\rangle=0$ unless $|u|+|v|=3$,
523: so that the sign prefactor 
524: in the first equation of (\ref{form_invariance}) is always 
525: one (though it is more convenient to write it out explicitly, as we did). 
526: 
527: 
528: The string field theory of \cite{sc,bv} is described by the action
529: \footnote{This is the real part of the complex action considered in 
530: \cite{gauge}. In that paper, we were interested in the moduli space, 
531: which can be understood without worrying about the real and imaginary parts.
532: In the present note, we shall consider the path integral, which requires that 
533: we work with the physical action, an object which must always be real.}:
534: \be
535: \label{action}
536: S(\phi)=\int_{L}{str\left[\frac{1}{2}\phi\bullet d\phi+\frac{1}{3}
537:     \phi\bullet\phi\bullet\phi\right]} +cc=
538: \frac{1}{2}\langle \phi, d\phi\rangle +\frac{1}{3}\langle
539: \phi,\phi\bullet \phi\rangle +cc~~~~,  
540: \ee which is defined on the degree one
541: component 
542: \be
543: {\cal H}^1=\{\phi\in {\cal H}||\phi|=1\}=
544: \Gamma(\oplus_{k+n-m=1}{\Lambda^k(T^*L)\otimes Hom(E_m,E_n)})
545: \ee 
546: of the total boundary space. 
547: This defines a `graded Chern-Simons field theory', which governs the 
548: dynamics of background shifts $\phi$. The equations of motion:
549: \be
550: \label{mc}
551: d\phi+\frac{1}{2}[\phi,\phi]_\bullet =0 \Leftrightarrow d\phi+
552: \phi\bullet \phi=0~~(\phi\in {\cal H}^1)~~
553: \ee
554: are equivalent with the requirement that the shifted 
555: superconnection $d_\phi=d+[\phi,.]_\bullet$ satisfies $(d_\phi)^2=0$. 
556: Hence the extrema of (\ref{action}) describe flat superconnections 
557: of total degree one. 
558: 
559: The equations of motion (\ref{mc}) are invariant
560: \footnote{One has 
561: $d(\phi^g)+\phi^g\bullet \phi^g=g\bullet(d\phi+\phi\bullet \phi)\bullet 
562: g^{-1}$. The object $d\phi+\phi\bullet \phi$ is the curvature of the 
563: associated superconnection.} 
564: with respect to gauge transformations of the form: 
565: \be
566: \label{Gauge}
567: \phi\rightarrow \phi^g=g\bullet \phi \bullet g^{-1}+g\bullet dg^{-1}~~,
568: \ee
569: where $g$ is an invertible element of the subalgebra $({\cal H}^0, \bullet)$.
570: This means that $g$ belongs to the group of units:
571: \be
572: \label{Gauge_group}
573: {\cal G}=\{g\in {\cal H}^0|{\rm~exists~}~g^{-1}\in {\cal H}^0~{\rm~such~that~}
574: ~g\bullet g^{-1}=g^{-1}\bullet g=1\}~~.
575: \ee
576: One has the standard transformation rule for the `covariant differential'
577: $d_\phi$:
578: \be
579: d_\phi(g^{-1}\bullet u\bullet g)=g^{-1}\bullet d_{\phi^g}\bullet g
580: \Leftrightarrow d_\phi(Ad_{g^-1}u)=Ad_{g^{-1}}(d_{\phi^g}u)~~,
581: \ee
582: for $u\in {\cal H}$, where $Ad_{g}(u):=g\bullet u \bullet g^{-1}$.
583: The moduli space results upon dividing the space ${\cal S}$ of solutions to 
584: (\ref{mc}) through the gauge symmetries (\ref{Gauge}). 
585: This is a standard Maurer-Cartan problem. 
586: 
587: 
588: In general, the 
589: the action (\ref{action}) is invariant only under `small' gauge 
590: transformations \cite{gauge}, i.e. those transformations for which 
591: $g$ can be written in the form:
592: \be
593: \label{exp}
594: g=e^\alpha_\bullet :=\sum_{k\geq 0}{\frac{1}{k!}\alpha^{\bullet k}}~~,
595: \ee
596: where $\alpha^{\bullet k}$ stands for $k$-th iteration of 
597: the $\bullet$-product of $\alpha$ with itself (and we define 
598: $\alpha^{\bullet 0}:=1$). 
599: 
600: 
601: 
602: For infinitesimal $\alpha$, the gauge 
603: transformations (\ref{Gauge}) become:
604: \be
605: \label{gauge}
606: \phi\rightarrow\phi+\delta_\alpha\phi~~, 
607: \ee 
608: with $\delta_\alpha\phi=-d\alpha-[\phi,\alpha]_\bullet$. 
609: For small $\phi$ and $\alpha$, 
610: the moduli problem  reduces to its linearized version:
611: \be
612: \label{mc_lin}
613: d\phi=0~~,~~\phi\equiv \phi-d\alpha~~,
614: \ee
615: which describes infinitesimal deformations of the background. 
616: 
617: 
618: 
619: \paragraph{Observation} Since graded Chern-Simons theories contain higher rank 
620: forms, a complete analysis of their gauge symmetries (with a view toward 
621: quantization) requires the Batalin-Vilkovisky formalism. The classical part 
622: of this analysis was carried out in \cite{bv}, while gauge fixing and 
623: propagators are discussed in \cite{gbv}. In the present paper, we 
624: will be able to avoid the BV formalism by employing the 
625: `resolvent method' of \cite{Schwarz_resolvent, Schwarz_resolvent2}. 
626: The BV analysis gives the same results (as expected 
627: from the general observations of \cite{Blau_Thompson}). 
628: 
629: \subsection{D-brane interpretation of the background}
630: 
631: Our string field theory is formulated around a specific 
632: background described by certain collections of graded D-branes. 
633: A shift of the background (which can be achieved by use of a solution $\phi$ 
634: to (\ref{mc})) deforms the differential $d$ into $d_\phi u=du+[\phi,u]$. 
635: It is easy to check (see \cite{com1, com3, gauge}) 
636: that such shifts preserve the relevant  properties of $d$. 
637: When expanded around the shifted background, the 
638: string field action has the same form (\ref{action}) up to the addition 
639: of an irrelevant constant \cite{com1}.  
640: 
641: As explained in \cite{com1, com3}, a general background admits a D-brane 
642: interpretation which allows for a systematic description in the language of 
643: category theory. This is related to the basic fact \cite{Bismut_Lott} 
644: that degree one superconnections can be decomposed as the sum of 
645: a usual connection and a collection of bundle-valued forms, which can be 
646: interpreted as condensates of spacetime fields associated with 
647: boundary condition changing operators. The decomposition properties of 
648: the boundary product and bilinear form with respect to these fields
649: can then be used to extract the D-brane content of the background
650: \cite{com1, com3}. The D-brane interpretation is different at various 
651: points in the moduli space. This identifies our moduli space of vacua 
652: with the moduli space of topological D-brane composites \cite{com1, com3}.
653: We refer the reader to the papers just cited for a detailed 
654: presentation of this analysis. It was shown on general grounds in \cite{com1}
655: (and, for the models discussed here, in \cite{sc}) that the collection of 
656: D-brane composites describing a background $\phi$ defines a `differential 
657: graded category', part of 
658: which can be related to certain enhanced triangulated 
659: categories \cite{BK}\footnote{We warn the reader that the description 
660: of `D-brane category dynamics' in \cite{com1,com3} is only local, 
661: i.e. based on the linearized version (\ref{mc_lin}) of the equations 
662: of motion; similar limitations apply to the triangulated category picture, 
663: which is a consequence of this. This is related to 
664: the issue of finding a good global definition for the 
665: `moduli space of a triangulated category', which seems to be best approached 
666: at the string field theory level.}. For the B-model
667: version of the theory (which was considered in \cite{Diaconescu}), 
668: the relevant triangulated category essentially coincides with 
669: the derived category of coherent sheaves. 
670: 
671: From many points of view, the string field theory description must come 
672: {\em before} any discussion in terms of triangulated categories. For example, 
673: a proper definition of deformations seems to require 
674: string field data, and possesses a standard (i.e. Maurer-Cartan) 
675: formulation only at the string field level \cite{gauge}. 
676: 
677: 
678: 
679: \section{The semiclassical approximation and generalized Ray-Singer torsion}
680: 
681: We are interested in the partition function $Z$ of (\ref{action}). 
682: Fixing a flat superconnection $A$, this can be determined by performing a 
683: background perturbation expansion around $A$. It is a standard 
684: fact that the semiclassical contribution to this expansion is given (up to the 
685: trivial factor $e^{-i\lambda S(A)}$, which we shall ignore) by certain
686: functional determinants, which encode the contribution of the kinetic term.
687: The purpose of this section is to evaluate this approximation in order 
688: to extract a certain numerical invariant associated 
689: with our models. 
690: 
691: \subsection{Hermitian structure}
692: 
693: Our gauge-fixing procedure will require that we pick a Riemannian 
694: metric $g$ on $L$ and Hermitian metrics on the bundles
695: $E_n$. The last determine a Hermitian metric $g_{\bf E}$ on ${\bf E}$, 
696: and thus a Hermitian metric on $End({\bf E})$, given by:
697: \be
698: (\alpha,\beta)=tr(\alpha^\dagger\circ \beta)~~{\rm~for~}~~\alpha,\beta\in 
699: End({\bf E}_p)~~,p\in L~~,
700: \ee
701: where $\alpha^\dagger$ is the Hermitian conjugate of $\alpha$ with respect to 
702: $g_{\bf E}$. On the other hand, $g$ induces a Hermitian 
703: metric $(.,.)$ on $\Lambda^*(T^*L)$
704: in the standard fashion. This is given by: 
705: \be
706: (*{\overline \omega})\wedge \eta=(\omega, \eta)vol_g~~,{\rm~for~}~~
707: \omega, \eta\in \Lambda^*(T^*_pL)~~,
708: \ee
709: where $vol_g$ is the volume form induced by $g$ on $L$ (with respect to the 
710: fundamental orientation on $L$), while 
711: $*$ is the {\em complex linear} Hodge operator, 
712: which is involutive and satisfies:
713: \be
714: rk(*\omega)=3-rk\omega~~.
715: \ee
716: Combining these, we obtain a Hermitian metric $(.,.)_{\cal V}$ on the bundle 
717: ${\cal V}=\Lambda^*(T^*L)\otimes End({\bf E})$, which can be described through
718: the relation:
719: \be
720: \label{Vmetric}
721: tr(*u^\dagger\wedge v)=(u,v)_{\cal V}
722: vol_g~~{\rm~for~}u,v\in {\cal V}_p~~,~~p\in L~~.
723: \ee
724: It is clear that $(u,v)_{\cal V}$ vanishes (on bi-homogeneous elements) 
725: unless $\Delta(u)=\Delta(v)$ and $rk u= rk v$ (and thus unless $|u|=|v|$). 
726: In particular, we have induced scalar products on each of the subbundles 
727: ${\cal V}^k$, which are mutually orthogonal with respect to (\ref{Vmetric}).
728: Integration over $L$ gives a Hermitian scalar product on the 
729: space ${\cal H}=\Gamma({\cal V})$:
730: \be
731: \label{hmetric}
732: h(u,v)=\int_{L}{(u,v)vol_g}=
733: \int_{L}{tr(*u^\dagger\wedge v)}~~,~~{\rm for}~~u,v\in {\cal H}~~.
734: \ee
735: 
736: 
737: 
738: \subsubsection{The conjugation operator} 
739: 
740: This was discussed in detail \cite{gauge}, so I will be short. 
741: Since the bilinear form (\ref{bf_A}) is non-degenerate, there exists a unique 
742: antilinear operator $c$ on ${\cal H}$ with the property:
743: \be
744: \label{h}
745: h(u,v)=\langle cu,v\rangle=\int_{L}{str[(cu)\bullet v]}~~.
746: \ee
747: This has the following form on decomposable elements:
748: \be
749: \label{c_A}
750: c(\omega\otimes f)=(-1)^{n+\Delta(f)(1+rk \omega)}
751: (*{\overline \omega})\otimes f^\dagger~~,~~{\rm~for~}~~\omega\in \Omega^*(L)~~
752: {\rm~and~}~~f\in Hom(E_m,E_n)~~.
753: \ee
754: The conjugation operator satisfies:
755: \be
756: \label{c_props}
757: |cu|=3-|u|~~, ~~c^2=id~~,
758: \ee
759: and thus obeys the axiomatic framework of \cite{superpot}.
760: Note that $c$ is anti-unitary:
761: \be
762: h(u,v)=h(cv,cu)~~.
763: \ee
764: In the ungraded case (${\bf E}=E_0$), 
765: $c$ reduces to the {\em antilinear} Hodge operator 
766: ${\overline *}$, coupled to the bundle $End(E)$.
767: 
768: 
769: \subsubsection{The adjoint of $d$ and the Laplacian}
770: 
771: As mentioned above, the 
772: scalar product $h$ satisfies the selection rule:
773: \be
774: \label{h_sel}
775: h(u,v)=0 ~~,~~{\rm~unless~}~~|u|=|v|~~.
776: \ee
777: Considering the Hermitian conjugate $d^\dagger$ of $d$ (with respect to $h$), we have:
778: \be
779: d^\dagger u= (-1)^{|u|}cdcu~~,~~|d^\dagger u|=|u|-1~~,~~(d^\dagger)^2=0~~.
780: \ee
781: Using this operator, one constructs the `deformed Laplacian' 
782: $\Delta=dd^\dagger+d^\dagger d$, which is an order two elliptic operator
783: on ${\cal H}$. We also note the relations:
784: \bea
785: \label{cdd}
786: cd^\dagger u= (-1)^{|u|}dcu~~&,&~~d^\dagger cu=(-1)^{|u|+1}cdu~~\nn\\
787: d^\dagger dc=cdd^\dagger~~&,&~~dd^\dagger c=cd^\dagger d~~.
788: \eea
789: 
790: \subsection{The spacetime ghost grading}
791: 
792: According to (\ref{h_sel}), we have induced 
793: Hermitian scalar products on each of 
794: the subspaces ${\cal H}^k$, which are mutually orthogonal with respect 
795: to $h$. For what follows, it will be convenient to use the grading 
796: $s(u)=1-|u|$, which in the BV formalism corresponds to the ghost number 
797: of the string field theory \cite{bv}. 
798: We shall use the notation ${\cal H}(\sigma)={\cal H}^{1-\sigma}$ 
799: for the homogeneous subspaces of ${\cal H}$ with respect to this grading, 
800: and the notation $H_\sigma({\cal H})=H^{1-\sigma}({\cal H})$ for the 
801: associated components of $H^*({\cal H})$ (since the ghost grading 
802: is decreased by $d$, this takes the original cochain complex $({\cal H},d)$ 
803: into a chain complex, which is why we use homological notation).
804: With this convention, the string field lies in the subspace ${\cal H}(0)$ 
805: of vanishing ghost number. 
806: Note that:
807: \be
808: \label{s_props}
809: s(du)=s(u)-1~~,~~s(cu)=-1-s(u)~~,~~s(cdu)=-s(u)~~.
810: \ee
811: In particular, $cd$ induces a well-defined antilinear operator on the 
812: physical subspace ${\cal H}(0)$. 
813: 
814: It is clear from the definition of $c$ that a nonzero element $u\in {\cal H}$ 
815: always has a nonzero conjugate partner $cu$. Since the number of graded 
816: components $E_n$ is assumed to be finite, there exists 
817: a maximal value of the degree $s$, which we shall denote by $N$.
818: The second relation in (\ref{s_props}) then shows that the minimal value of 
819: $s$ is $-1-N$. Accordingly, the degree $|.|=1-s(.)$ lies in the interval
820: $1-N\dots 2+N$. 
821: 
822: We also recall the standard decompositions:
823: \be
824: {\cal H}=K\oplus im d \oplus im d^\dagger~~,~~ker d=imd \oplus K~~,~~ker d^\dagger= im d^\dagger \oplus K~~,
825: \ee
826: where $K=ker \Delta=ker d\cap ker d^\dagger$. Hodge theory leads 
827: to the identification $K\approx H_d^*({\cal H})$.
828: 
829: \subsection{Complex and real determinants and induced Euclidean scalar 
830: products}
831: 
832: In the next subsection, it will be convenient to `decompose our fields
833: into real components' when performing the path integral. For that purpose, 
834: we collect some useful facts and definitions. This is entirely trivial and 
835: well-known, but I shall state it explicitly nonetheless. 
836: 
837: The complex vector space ${\cal H}$ 
838: can be viewed as a real vector space upon restriction 
839: of the field of scalars. Then the Hermitian product $h$
840: induces an Euclidean scalar product $(.,.):=Re h$ on ${\cal H}$. 
841: Given a complex vector subspace $W$ of ${\cal H}$, it is easy to check 
842: that for any vector $u$ in ${\cal H}$, one has the equivalence:
843: \be
844: \label{EH}
845: (u,v)=0{\rm ~for~all~}v\in W\Leftrightarrow <u,v>=0{\rm~for~all~}v\in W.
846: \ee
847: Indeed, if $W$ contains a vector $v$, then it also contains the vector 
848: $iv$, and 
849: one has $(u,iv)=Reh(u,iv)=Re [ih(u,v)]= -Im h(u,v)$. Relation (\ref{EH}) 
850: shows that the orthogonal complements of $W$ with respect to $h$ and 
851: $(.,.)$ coincide, so we shall use the symbol $W^\perp$ to denote either.
852: 
853: A complex-linear operator $O$ on a complex vector space $V$ 
854: can also be viewed as a real-linear operator on the underlying 
855: real vector space. 
856: Therefore, one has two notions for the determinant of $O$, namely the
857: complex and real determinant.
858: The first is the determinant of $O$ viewed as a complex-linear map, 
859: i.e. the determinant of its matrix in a complex-linear basis
860: $e_\alpha$ of $V$:
861: \be
862: \label{real_matrix}
863: Oe_\alpha=a_{\beta\alpha}e_\beta\Rightarrow det O=det (a_{\alpha\beta})~~.
864: \ee
865: It can be described more geometrically by considering the maximal exterior 
866: power $\Lambda^d V$ of $V$ viewed as a complex vector space
867: (here $d$ is the complex dimension of $V$). Indeed, $O$ induces a linear 
868: endomorphism of $\Lambda^d V$. Since the latter space is complex 
869: one-dimensional, this endomorphism is canonically identified with a complex 
870: number, which is the determinant of $O$. In fact, $\Lambda^d V$
871: has a basis given by $e_1\wedge\dots \wedge e_d$, and 
872: $O(e_1)\wedge\dots \wedge O(e_d)=det O e_1\wedge\dots \wedge e_d$.
873: 
874: The real determinant arises from a similar construction, where one views 
875: $O$ as a real linear operator on the real vector space $V$ (whose 
876: real dimension is $2d$). The latter admits the basis 
877: $e_1\dots e_d, ie_1\dots ie_d$, which gives:
878: \bea
879: Oe_\alpha&=&x_{\beta\alpha}e_\alpha+y_{\beta\alpha}(ie_\alpha)~~\nn\\
880: O(ie_\alpha)&=&-y_{\beta\alpha}e_\alpha+x_{\beta\alpha}(ie_\alpha)~~,
881: \eea
882: where $x_{\alpha\beta}=Re a_{\alpha\beta}$ and 
883: $y_{\alpha\beta}=Im a_{\alpha\beta}$. The real determinant $det_\R O$ 
884: is the determinant of the resulting real $2d\times 2d$ matrix. This can again be 
885: identified with the operator induced by $O$ on the real one-dimensional 
886: vector space $\Lambda_\R^{2d} V$, where $\Lambda_\R$ denotes the real 
887: exterior product. One has $O(e_1)\wedge_\R\dots \wedge_\R O(e_d)
888: \wedge_\R O(ie_1)\wedge_\R \dots \wedge_\R O(ie_d)=det_R(O) 
889: e_1\wedge_\R \dots \wedge_\R e_d\wedge (ie_1)\wedge_\R \dots \wedge_\R(ie_d)$.
890: One can show that:
891: \be
892: det_\R O= |det O|^2~~,
893: \ee
894: a formula which will be used repeatedly in later subsections. 
895: 
896: 
897: 
898: \subsection{The semiclassical partition function in the 
899: resolvent formalism}
900: 
901: \subsubsection{The computation} 
902: 
903: To compute the semiclassical partition function, 
904: we use the resolvent formalism of \cite{Schwarz_resolvent,Schwarz_resolvent2}, 
905: in its generalized version discussed in \cite{Adams_Sen0, Adams_Sen}\footnote{
906: A streamlined presentation of this material can be found in the thesis of 
907: D. H. Adams. The author thanks him for bringing this thesis to his attention.}
908: (see \cite{Adams_scl, Adams_FP, Adams_cpct} and 
909: \cite{GK, NC} for related issues). For this, we focus on the kinetic term: 
910: \be
911: \label{Skin}
912: S_{kin}(\phi)=Re \langle \phi, d\phi\rangle=
913: Re \int_{L}{str(\phi \bullet d\phi)}~~,
914: \ee
915: where $d:=d_A$ is the differential on ${\cal H}$ associated with the
916: background $A$. The semiclassical partition function is formally given by:
917: \be
918: \label{Zformal}
919: Z_{scl}=\int{{\cal D}[\phi]e^{-i\lambda S_{kin}(\phi)}}~~,
920: \ee
921: where $\lambda>0$ is the coupling constant.
922: This path integral is of course ill-defined. The most severe problem 
923: is the presence of zero modes, which are related to the gauge invariance 
924: $\phi\rightarrow \phi+d\omega$, with $\omega$ an element of 
925: ${\cal H}(1)={\cal H}^0=
926: \oplus_{m,n}{\Omega^{m-n}(L, Hom(E_m,E_n))}$. A precise 
927: characterization of zero modes is as follows. Using the metric 
928: induced on ${\cal H}(0)$, we write $S_{kin}$ in the form:
929: \be
930: S_{kin}(\phi)=Re~h(\phi, cd\phi)=(\phi, cd\phi)~~,
931: \ee
932: where $(.,):=Reh(.,.)$ is the Euclidean scalar product induced by 
933: $h$, if we view ${\cal H}^1={\cal H}(0)$ as a real vector 
934: space. 
935: The zero modes lie in the kernel of this quadratic functional.
936: Since $(.,.)$ is nondegenerate, this coincides with the kernel of the 
937: antilinear operator $T_0:=cd:{\cal H}(0)\rightarrow {\cal H}(0)$. 
938: In fact, $T_0$ becomes a (real-linear) 
939: self-adjoint operator, as can be checked by using invariance of 
940: $\langle.,.\rangle$ with respect to $d$:
941: \be
942: (T_0u, v)=(u,T_0v)~~{~~\rm~for~~}~u, v \in {\cal H}(0)\Rightarrow 
943: T_0^t=T_0~~,
944: \ee
945: where $^t$ stands for the adjoint with respect to $(.,.)$.
946: It is also easy to check that:
947: \be
948: (\phi, T_0\phi)=(\phi_M, T_0\phi_M)~~,
949: \ee 
950: where we decomposed $\phi:=\phi_M\oplus \phi_K$ with 
951: $\phi_K\in ker d=ker T_0$ and $\phi_M \in (ker d)^\perp=(ker T_0)^\perp$. 
952: If we let $T_0'$ denote the restriction of $T_0$ to the orthogonal 
953: complement of its kernel, then formal integration in (\ref{Zformal}) leads to:
954: \be
955: \label{Zformal1}
956: Z_{scl}=vol(ker(T_0))\int{{\cal D}[\phi_M]e^{-i\lambda (\phi_M, T_0'\phi_M)}}~~.
957: \ee
958: 
959: Both factors in (\ref{Zformal1}) are ill-defined, 
960: since the first involves the `volume' of a vector space
961: while the second is a Gaussian integral for a generally indefinite 
962: quadratic form. The second problem is solved in standard manner by 
963: replacing $T_0 '$ with the strictly positive operator 
964: $|T_0'|=\sqrt{T_0'T_0'}=
965: \sqrt{(d^\dagger d)'_{s=0}}$, 
966: where the subscript $s=0$ indicates that 
967: $d^\dagger d$ is restricted to the subspace ${\cal H}(0)$.
968: This leads to the expression:
969: \be
970: \int{{\cal D}[\phi_M]e^{-i\lambda (\phi_M, T_0'\phi_M)}}=
971: ct\times (det_\R|T_0'|)^{-1/2}~~
972: \ee
973: where $det_\R|T_0'|$ denotes the real determinant of $|T_0'|$. 
974: Using $det_\R |T'_0|^{-1/2}=det |T_0'|^{-1}= (det'_{s=0}(d^\dagger d))^{-1/2}$,
975: we obtain:
976: \be
977: \label{Zformal2}
978: Z_{scl}=ct\times vol(ker(T_0)) (det'_{s=0}d^\dagger d)^{-1/2}~~,
979: \ee
980: where, for a Hermitian operator $O$, 
981: $det' O$ denotes the complex determinant of $O|_{(ker O)^\perp}$.
982: 
983: The last factor of (\ref{Zformal2}) is still ill-defined, since it involves 
984: an infinite product of eigenvalues. This is a standard problem, which 
985: is solved by zeta-function regularization.
986: We remind the reader that, given a positive elliptic operator $O$, 
987: acting on sections of a vector bundle over $L$, 
988: its zeta function $\zeta_O(z)$ $(z\in \C)$ is defined as follows.
989: For sufficiently large $Rez$, $\zeta_O(z)$ is given by the expansion:
990: \be
991: \zeta_O(z)=\sum_{\lambda}{\frac{n_\lambda}{\lambda^z}}~~,
992: \ee
993: where the sum is over the (strictly positive) 
994: eigenvalues $\lambda$ of $O$, whose multiplicities we denote by 
995: $n_\lambda$. 
996: The analytic continuation of this series to the 
997: complex plane is meromorphic and regular at the origin. 
998: This allows one to define the regularized determinant through the expression:
999: $det^{reg}(O):= e^{-\zeta'_{O}(0)}$, where 
1000: $\zeta'_{O}(z):=\frac{d}{dz}\zeta_{O}(z)$. 
1001: Applying this to the operator $|T_0'|^2$, we obtain:
1002: \be
1003: \label{detprime}
1004: det'^{,reg}_{s=0}(d^\dagger d)= e^{-\zeta'_{(d^\dagger d)'_{s=0}}(0)}~~.
1005: \ee 
1006: 
1007: 
1008: To regularize the volume factor in (\ref{Zformal2}), 
1009: we use the so-called {\em method of resolvents} 
1010: \cite{Schwarz_resolvent, Adams_Sen0, Adams_Sen},
1011: which proceeds as follows\footnote{Since we work with complex fields, 
1012: (i.e. $\phi$ varies in a complex vector space)
1013: we will have to make some minor modifications in order to adapt the work of 
1014: \cite{Adams_Sen}.}. First, we notice that the kinetic 
1015: operator $T_0=cd$ admits an {\em elliptic resolvent}.
1016: More precisely, one has the complex of {\em real} vector spaces:
1017: \be
1018: \label{res}
1019: ({\cal R})~~:~~0\rightarrow{\cal H}(N)
1020: \stackrel{T_N=d}{\longrightarrow}\dots 
1021: \stackrel{T_2=d}{\longrightarrow} 
1022: {\cal H}(1)\stackrel{T_1=d}{\longrightarrow}{\cal H}(0) 
1023: \stackrel{T_0=cd}{\longrightarrow}{\cal H}(0)~~,
1024: \ee
1025: (where $N$ denotes the maximal value of the spacetime ghost degree, 
1026: as discussed in Subsection 3.2), and the associated Laplace operators:
1027: \be
1028: \label{Laplacians}
1029: \Delta_\sigma:=T_\sigma^t T_\sigma+
1030: T_{\sigma+1} T_{\sigma+1}^t
1031: = d^\dagger d+dd^\dagger:{\cal H}(\sigma)\rightarrow {\cal H}(\sigma)
1032: \ee
1033: (with $T_{N+1}:=0$) are elliptic for all $\sigma=0\dots N$. 
1034: To arrive at (\ref{Laplacians}), 
1035: we used the fact that $T_\sigma^t=T_\sigma^\dagger$ for 
1036: $\sigma>0$ and $T_0^t T_0=d^\dagger d$ for $\sigma=0$.
1037: Notice that ${\cal H}(\sigma)={\cal H}^{1-\sigma}=\Gamma({\cal V}^{1-\sigma})$
1038: are in fact complex vector spaces, but the last map in (\ref{res}) 
1039: is complex antilinear, rather than complex linear.  This is why we 
1040: can only view (\ref{res}) as a complex of real-linear maps. 
1041: 
1042: To regularize 
1043: $vol(ker T_0)$, one picks auxiliary Hermitian metrics 
1044: $h_\sigma^H$ on the homology spaces 
1045: $H_\sigma({\cal R})=ker T_\sigma/im T_{\sigma+1}=H^{1-\sigma}_d({\cal H})$
1046: of the resolvent, and considers the Hodge isomorphisms:
1047: \be
1048: \label{fsigma}
1049: f_\sigma:K(\sigma)=K^{1-\sigma}=
1050: ker T_\sigma\cap ker T_\sigma^t
1051: \stackrel{\approx}{\longrightarrow} 
1052: H_\sigma({\cal R})=H^{1-\sigma}_d({\cal H})~~
1053: \ee 
1054: induced by the obvious 
1055: projections $ker T_\sigma\rightarrow H_\sigma({\cal R})$. 
1056: Given this data, one can formally compute \cite{Adams_Sen}:
1057: \bea
1058: \label{vol_formal0}
1059: vol(ker T_0)=\prod_{\sigma=0}^N{
1060: \left[\left(\frac{det'_\R(T_{\sigma+1}^t T_{\sigma+1})}
1061: {det_\R(f_\sigma^t f_\sigma)}\right)^{1/2}vol(H_\sigma({\cal R}))
1062: vol({\cal H}(\sigma+1))\right]^{(-1)^{\sigma}}}~~,
1063: \eea
1064: where $det_\R$ denotes the determinant of real-linear maps and
1065: we defined $det'_\R(T_{N+1}^t T_{N+1}):=1$ and $vol({\cal H}(N+1)):=1$. 
1066: Relation (\ref{vol_formal0}) gives:
1067: \bea
1068: \label{vol_formal}
1069: vol(ker T_0)=\prod_{\sigma=0}^N{
1070: \left[\frac{det'_{s=\sigma+1}(d^\dagger d)}
1071: {det(f_\sigma^\dagger f_\sigma)}vol(H_\sigma({\cal R}))vol({\cal H}(\sigma+1))
1072: \right]^{(-1)^{\sigma}}}~~
1073: \eea
1074: (where again we define $det'_{s=N+1}(d^\dagger d):=1$).
1075: This expression is given a meaning by dropping 
1076: the ill-defined factors
1077: $vol(H_\sigma({\cal R}))$ and $vol({\cal H}(\sigma))$ 
1078: and using zeta-function regularization for
1079: the determinants of the positive operators 
1080: $(d^\dagger d)'_{s=\sigma}=
1081: [(d^\dagger d)|_{ker (d)^\perp}]_{s=\sigma}$: 
1082: \be
1083: vol^{reg}(ker T_0)=\prod_{\sigma=0}^N{
1084: \left[\frac{det^{',reg}_{s=\sigma+1}(d^\dagger d)}
1085: {det(f_\sigma^\dagger f_\sigma)}\right]^{(-1)^{\sigma}}}~~.
1086: \ee
1087: Combining with (\ref{Zformal2}), we conclude:
1088: \bea
1089: \label{Zf}
1090: Z_{scl}=Cdet'^{,reg}_{s=0}(d^\dagger d)^{-1/2}
1091: \prod_{\sigma=0}^N{
1092: \left[\frac{det'^{,reg}_{s=\sigma+1}(d^\dagger d)}
1093: {det(f_\sigma^\dagger f_\sigma)}\right]^{(-1)^{\sigma}}}=\\
1094: C e^{\frac{1}{2}\zeta'_{(d^\dagger d)'_{s=0}}(0)
1095: +\sum_{\sigma=1}^N{(-1)^\sigma \zeta'_{(d^\dagger d)'_{s=\sigma}}(0)}}
1096: \prod_{\sigma=0}^N{
1097: \left[det(f_\sigma^\dagger f_\sigma)^{(-1)^{\sigma+1}}\right]}~~,\nn
1098: \eea
1099: where $C$ is a complex constant. If one uses the normalization conventions
1100: of \cite{Adams_Sen}, then $C$ can be written as:
1101: \be
1102: \label{C}
1103: C=\left(\frac{\pi}{\lambda}\right)^{\zeta_{|T_0|}(0)/2}e^{-\frac{i\pi}{4}
1104: \eta_{T_0}(0)}~~,
1105: \ee
1106: where $\zeta_{|T_0|}:=\zeta_{|T_0'|}$
1107: is the zeta-function of the positive operator $|T_0'|$ and 
1108: $\eta_{T_0}:=\eta_{cd_{s=0}}$ is the eta-function of $T_0$. 
1109: The latter is defined through analytic continuation of 
1110: the following expression\footnote{Remember that $T_0$ is 
1111: self-adjoint only as a real-linear operator. To find its eigenvalues, 
1112: one must of course consider the complexification $W$ of the underlying
1113: real vector space of ${\cal H}(0)$. The complex dimension of $W$ 
1114: equals $2dim_\C {\cal H}(0)$. 
1115: The selfadjoint operator $T_0$ induces a 
1116: {\em complex-linear} Hermitian operator $T_0^\C$
1117: on $W$. It is the real 
1118: eigenvalues of this operator which appear in (\ref{eta})  
1119: for our case.}
1120: (which is valid for $Re z>>0$):
1121: \be
1122: \label{eta}
1123: \eta_{T_0}(z)=\sum_{\lambda>0}{\frac{n_\lambda}{\lambda^z}}-
1124: \sum_{\lambda<0}{\frac{n_\lambda}{|\lambda|^z}}~~,
1125: \ee
1126: the sums being over the strictly positive and strictly negative 
1127: eigenvalues of $T_0$ $(n_\lambda$ are the multiplicities). We refer the reader 
1128: to \cite{Adams_Sen} for the justification of (\ref{C}) (in particular, 
1129: it is shown there that $\eta_{T_0}(0)$ is well-defined). Here we only note 
1130: that the absolute value of $C$ produced by a putative non-perturbative 
1131: solution of the full graded Chern-Simons theory 
1132: need not strictly agree with (\ref{C}), 
1133: due to corrections from a possibly nontrivial gauge stabilizer
1134: of the background superconnection $A$ (this is 
1135: similar to what happens for the ungraded case 
1136: \cite{Adams_Sen, Adams_FP, Adams_scl, Adams_cpct}).
1137: Therefore, the modulus of expression (\ref{C}) should {\em not} 
1138: be taken at face value.
1139: 
1140: \subsubsection{Generalized Ray-Singer torsion}
1141: 
1142: Let us write the partition function (\ref{Zf}) in an alternate form. 
1143: First we introduce the notation 
1144: $\zeta_{\sigma}(z):=\zeta_{(d^\dagger d)'_{s=\sigma}}(z)$.
1145: Lemma 3.1 of \cite{Adams_Sen} establishes the relation:
1146: \be
1147: \label{zeta_rel}
1148: \zeta_\sigma(z)+\zeta_{\sigma+1}(z)=\zeta_{\Delta_\sigma}(z)
1149: ~~(\sigma \geq 0)~~.
1150: \ee
1151: where $\zeta_{\Delta_\sigma}$ is a shorthand for $\zeta_{\Delta'_\sigma}$.
1152: This follows from the fact that 
1153: $\Delta'_\sigma=(dd^\dagger)'\oplus (d^\dagger d)'$ (since 
1154: $(ker \Delta)^\perp=im d^\dagger \oplus im d=(ker d)^\perp\oplus 
1155: (ker d^\dagger)^\perp=ker (d^\dagger d)^\perp\oplus (ker d d^\dagger)^\perp$),
1156: which implies
1157: $\zeta_{\Delta'_\sigma}(z)=\zeta_{(d^\dagger d)'_\sigma}(z)+
1158: \zeta_{(dd^\dagger)'_\sigma}(z)$.
1159: Then one notices that $d$ gives an isomorphism between 
1160: $ker(d^\dagger d)^\perp=ker (d)^\perp$ and 
1161: $ker(dd^\dagger)^\perp=ker(d^\dagger)^\perp=im(d)$, thereby inducing 
1162: a bijection between the nonzero eigenvalues of $(d^\dagger d)_{\sigma+1}$ and 
1163: $(dd^\dagger)_\sigma$ (if $d^\dagger du =\lambda u$ then 
1164: $dd^\dagger du=\lambda du$ etc.). Therefore, 
1165: $\zeta_{(dd^\dagger)'_\sigma}(z)=\zeta_{\sigma+1}(z)$, which leads to
1166: (\ref{zeta_rel}).
1167: 
1168: Equation (\ref{zeta_rel}) implies:
1169: \be
1170: \label{s1}
1171: \zeta_0+2\sum_{\sigma=1}^N{(-1)^\sigma\zeta_\sigma}=
1172: \sum_{\sigma=0}^N{(-1)^\sigma(1+2\sigma)\zeta_{\Delta_\sigma}}~~.
1173: \ee
1174: The next step is to notice that the last two equations in 
1175: (\ref{cdd}) imply $\Delta c=c\Delta$ and thus:
1176: \be
1177: c\Delta_\sigma=\Delta_{-1-\sigma}c~~,~~\Delta_\sigma c=c\Delta_{-1-\sigma}~~,
1178: \ee
1179: because $s(cu)=-1-s(u)$. 
1180: Since $c$ is a bijection, this shows that the operators
1181: $\Delta_\sigma$ and $\Delta_{-1-\sigma}$ are iso-spectral (i.e. have the 
1182: same eigenvalues, with the same multiplicities). Therefore:
1183: \be
1184: \zeta_{\Delta_\sigma}(z)=\zeta_{\Delta_{-1-\sigma}}(z)~~{\rm~~for~~}
1185: \sigma\in \{ -1-N \dots N\}~~.
1186: \ee
1187: Thus:
1188: \be
1189: \label{s2}
1190: \sum_{\sigma=0}^N{(-1)^\sigma(1+2\sigma)\zeta_{\Delta_\sigma}}=
1191: \sum_{\sigma=-N-1}^N{(-1)^\sigma\sigma\zeta_{\Delta_\sigma}}=
1192: \sum_{q=1-N}^{N+2}{(-1)^q q\zeta_{\Delta^{(q)}}}~~,
1193: \ee
1194: where $\Delta^{(q)}=\Delta_{1-q}$ is the Laplacian on the space 
1195: ${\cal H}^q={\cal H}(1-q)$. 
1196: Combining (\ref{s1}) and (\ref{s2}), we obtain:
1197: \be
1198: \label{sums_final}
1199: \frac{1}{2}\zeta_0+\sum_{\sigma=1}^N{(-1)^\sigma\zeta_\sigma}=
1200: \frac{1}{2}\sum_{\sigma=-N-1}^{N}{(-1)^\sigma\sigma\zeta_{\Delta_\sigma}}=
1201: \frac{1}{2}\sum_{q=1-N}^{N+2}{(-1)^q q\zeta_{\Delta^{(q)}}}~~,
1202: \ee
1203: where $\sigma$ and $q$ in the right hand side 
1204: run over their maximal domains of definition 
1205: $\{-1-N\dots N\}$ and $\{1-N\dots N+2\}$. 
1206: Let us define the quantity:
1207: \be
1208: \label{T}
1209: T(L, A):=\prod_{q=1-N}^{N+2}{[det^{',reg}\Delta^{(q)}]^{\frac{(-1)^qq}{2}}}=
1210: e^{-\frac{1}{2}\sum_{q=1-N}^{N+2}{(-1)^q q\zeta'_{\Delta^{(q)}}(0)}}=
1211: e^{-\frac{1}{2}
1212: \sum_{\sigma=-N-1}^N{(-1)^\sigma\sigma \zeta'_{\Delta_\sigma}(0)}}~~,
1213: \ee
1214: which is a generalized version of the analytic 
1215: torsion of Ray and Singer 
1216: \cite{Ray, RS1, RS_symp}. Then:
1217: \be
1218: Z_{scl}=C T(L, A)^{-1}
1219: \prod_{\sigma\geq 0}{
1220: \left[det(f_\sigma^\dagger f_\sigma)^{(-1)^{\sigma+1}}\right]}~~.
1221: \ee
1222: It is convenient to write $Z_{scl}=C{\tilde Z}_{scl}$, where:
1223: \be
1224: {\tilde Z}_{scl}=T(L, A)^{-1}
1225: \prod_{\sigma\geq 0}{
1226: \left[det(f_\sigma^\dagger f_\sigma)^{(-1)^{\sigma+1}}\right]}~~.
1227: \ee
1228: 
1229: {\bf Observation} Relations (\ref{cdd}) also imply that the 
1230: operators $(d^\dagger d)_\sigma|_{ker(d)^\perp}$ and 
1231: $(dd^\dagger)_{-1-\sigma}|_{ker (d^\dagger)^\perp}$ are iso-spectral.
1232: Since the latter is iso-spectral with 
1233: $(d^\dagger d)_{-\sigma}|_{(ker d)^\perp}$, this gives the relation:
1234: \be
1235: \zeta_\sigma(z)=\zeta_{-\sigma}(z)~~,
1236: \ee
1237: which implies $\frac{1}{2}\zeta_0+\sum_{\sigma=1}^N{(-1)^\sigma\zeta_\sigma}=
1238: \frac{1}{2}\sum_{\sigma=-N}^N{(-1)^\sigma\zeta_\sigma}$.
1239: Together with (\ref{sums_final}), this gives:
1240: \be
1241: \frac{1}{2}\sum_{q=1-N}^{N+2}{(-1)^q q\zeta_{\Delta^{(q)}}}=
1242: \frac{1}{2}\sum_{\sigma=-N}^N{(-1)^\sigma\zeta_\sigma}~~
1243: \ee
1244: and leads to the following expression for the torsion:
1245: \be
1246: T(L,A)=e^{-\frac{1}{2}
1247: \sum_{\sigma=-N}^N{(-1)^\sigma\zeta'_{\sigma}(0)}}~~.
1248: \ee
1249: Note that the sum in the exponent does not contain a term with 
1250: $\sigma=-N-1$.
1251: 
1252: 
1253: \subsubsection{A Hermitian analogue of the Ray-Singer metric}
1254: 
1255: The factor $J:=\prod_{\sigma\geq 0}{
1256: \left[det(f_\sigma^\dagger f_\sigma)^{(-1)^{\sigma+1}}\right]}$
1257: can be described geometrically as follows. 
1258: We first notice that there exists a unique antilinear involution
1259: $c_*:H_{\sigma}({\cal R})\rightarrow H_{-1-\sigma}({\cal R})$ with the 
1260: property:
1261: \be
1262: f_\sigma c=c_*f_{-1-\sigma}~~,
1263: \ee
1264: where we consider the Hodge isomorphisms 
1265: (\ref{fsigma}) for all $\sigma=-N-1\dots N$. Given this involution, we 
1266: use the metrics $h^H_\sigma$ for $\sigma \geq 0$ to introduce
1267: metrics on the spaces $H_\sigma({\cal R})=
1268: H^{1-\sigma}({\cal H})$ with $\sigma<0$ through the relations:
1269: \be
1270: h^H_\sigma(\alpha,\beta)=h^H_{-1-\sigma}(c_*\beta,c_*\alpha)
1271: ~~{\rm~for~all~}~~\sigma=-N-1 \dots -1~~.
1272: \ee
1273: With respect to these metrics, the operators $c_*$ are anti-unitary.
1274: Use of particular extension of the auxiliary data $h^H_\sigma (\sigma\geq 0)$ 
1275: to the entire cohomology $H^*_d({\cal H})$ is crucial for the validity 
1276: of certain arguments below. 
1277: 
1278: We have $f_\sigma^\dagger f_\sigma=cf_{-1-\sigma}^\dagger f_{-1-\sigma} c$, 
1279: which implies:
1280: \be
1281: det(f^\dagger_\sigma f_\sigma)=det (f^\dagger_{-1-\sigma}f_{-1-\sigma})~~{\rm~for~all~}\sigma~~,
1282: \ee 
1283: where we used $c^2=id$.
1284: This allows us to write:
1285: \be
1286: \label{J}
1287: J=\prod_{\sigma\geq 0}{
1288: \left[det(f_\sigma^\dagger f_\sigma)^{(-1)^{1-\sigma}}\right]}=
1289: \prod_{\sigma=-N-1}^{N}{(detf^\dagger_\sigma f_\sigma)^{\frac{(-1)^{1-\sigma}}{2}}}=
1290: \prod_{q=1-N}^{2+N}{
1291: det(f^{(q)\dagger} f^{(q)})^{\frac{(-1)^{q}}{2}}}~~,
1292: \ee
1293: where we defined $f^{(q)}:=f_{1-q}:K^q\rightarrow H^{q}_d({\cal H})$.
1294: 
1295: Defining $b_\sigma:=dim_\C H_\sigma({\cal R})$ 
1296: and $b^q:=dim_\C H^{q}({\cal H})$ (so that $b_\sigma:=b^{1-\sigma}$), let
1297: us consider the `complex determinant line': 
1298: \be
1299: {\cal L}=det H^*_d({\cal H})=
1300: \otimes_{\sigma=-N-1}^N{
1301: \Lambda^{b_\sigma}[H_\sigma({\cal R})^{*(1-\sigma)}]}=\otimes_{q=1-N}^{2+N}
1302: {\Lambda^{b^q}[H^q_d({\cal H})^{*q}]}~~,
1303: \ee
1304: which is a one-dimensional 
1305: complex vector space. This space carries two Hermitian metrics. 
1306: The first metric is induced by $g$ and $g_{\bf E}$ 
1307: and arises upon transporting
1308: the restriction of $h$ to the harmonic subspaces $K(\sigma)$ to metrics 
1309: $h_\sigma=h^{(1-\sigma)}$ on $H_\sigma({\cal H})=H_d^{1-\sigma}({\cal H})$ 
1310: through the Hodge isomorphisms 
1311: $f_\sigma:K(\sigma)\rightarrow H_d^{1-\sigma}({\cal H})$: 
1312: \be
1313: \label{hsigma}
1314: h_\sigma(u,v)=h(f_\sigma^{-1}u,f_\sigma^{-1}v)
1315: =h^H_\sigma((f_\sigma^{-1})^\dagger f_\sigma^{-1}u,v)=
1316: h^H_\sigma((f_\sigma f_\sigma^\dagger)^{-1}u, v) {\rm~~for~~} u,v\in 
1317: H_d^{1-\sigma}({\cal H})~~.
1318: \ee
1319: The metrics (\ref{hsigma}) 
1320: induce a metric $h_{\cal L}$ 
1321: on ${\cal L}$ known as the {\em Hermitian} $L_2$-{\em metric}.
1322: If $e^{(q)}_\alpha$ are bases for the complex vector spaces $H^q_d({\cal H})$, 
1323: then $e:=\otimes_{q}{(e^{(q)}_1\wedge \dots \wedge e^{(q)}_{b^q})^{*q}}$ 
1324: is a complex basis of ${\cal L}$, 
1325: whose squared norm in the metric $h_{\cal L}$ is given by:
1326: \be
1327: ||e||^2_{\cal L}=h_{\cal L}(e,e)=\prod_{q}{[detG^{(q)}]^{(-1)^q}}~~,
1328: \ee
1329: where $G^{(q)}_{\alpha\beta}=h^{(q)}(e^{(q)}_\alpha, e^{(q)}_\beta)$ are the (positive-definite) 
1330: Hermitian Gramm matrices.
1331: On the other hand, the auxiliary metrics $h_\sigma^H=h^{(1-\sigma)}_H$ on 
1332: $H_\sigma({\cal H})=H{1-\sigma}_d({\cal H})$ 
1333: induce a metric $h^H_{\cal L}$ on ${\cal L}$, for which:
1334: \be
1335: (||e||^H_{\cal L})^2=h^H_{\cal L}(e,e)=\prod_{q}{[detG_H^{(q)}]^{(-1)^q}}~~,
1336: \ee
1337: with $(G_H^{(q)})_{\alpha\beta}=h_H^{(q)}(e^{(q)}_\alpha, e^{(q)}_\beta)$.
1338: It is clear from this description and from (\ref{J}), (\ref{hsigma}) that the two norms  
1339: on ${\cal L}$ are related through:
1340: \be
1341: \label{norm_relation}
1342: ||.||^H_{\cal L}=J||.||_{\cal L}~~.
1343: \ee
1344: Let us consider the quantity:
1345: \be
1346: ||.||_{RS}=T(L,A)||.||_{\cal L}~~,
1347: \ee
1348: which is a norm on ${\cal L}$ generalizing the standard Ray-Singer norm
1349: \cite{BZ}. This norm is independent of $h^H_\sigma$.
1350: With this definition, one can write ${\tilde Z}_{scl}$ as:
1351: \be
1352: {\tilde Z}_{scl}=\frac{||.||^H_{\cal L}}{~||.||_{RS}}~~,
1353: \ee
1354: which displays the dependence of ${\tilde Z}_{scl}$ on 
1355: the auxiliary data $h^H_\sigma$.
1356: 
1357: \subsubsection{Metric induced on the real determinant line}
1358: 
1359: The traditional formulation of Ray-Singer torsion in the ungraded case involves connections 
1360: defined on {\em real} vector bundles. As a consequence, 
1361: one obtains a norm defined not on the complex determinant line
1362: but on its real analogue. Since we work with complex vector bundles
1363: and connections, it is more natural in our case 
1364: to use the metric of the previous subsection, which 
1365: is defined on the complex determinant line.  Here we explain 
1366: the relation between this formulation and the traditional description.
1367: 
1368: Let us define the {\em real determinant line}: 
1369: \be
1370: \Lambda:=\otimes_{\sigma=-N-1}^N{
1371: \Lambda^{2b_\sigma}_\R[H_\sigma({\cal R})^{*(1-\sigma)}]}=\otimes_{q=1-N}^{2+N}
1372: {\Lambda_\R^{2b^q}[H^q_d({\cal H})^{*q}]}~~,
1373: \ee
1374: where all antisymmetrized and tensor products are now taken over the 
1375: field of real numbers. 
1376: There exists a natural isomorphism between $\Lambda$ and the 
1377: second exterior power $\Lambda^2_\R {\cal L}$ of the complex line, 
1378: where ${\cal L}$ is viewed as a real two-dimensional vector space 
1379: by restriction of the field of scalars.
1380: This isomorphism can be described as follows. Given bases 
1381: $e^{(q)}_1\dots e^{(q)}_{b^q}$ of the complex vector spaces $H^q({\cal H})$, 
1382: consider the bases $e^{(q)}_1\dots e^{(q)}_{b^q}, ie^{(q)}_1\dots ie^{(q)}_{b^q}$ of the 
1383: underlying real vector spaces, as well as the associated bases 
1384: $e:=\otimes_{q}{(e_1\wedge \dots\wedge e_{b^q})^{*q}}$ and 
1385: $\epsilon:=\otimes_{q}{(e_1\wedge_\R \dots \wedge_\R e_{b^q}
1386: \wedge_\R (ie_1)\wedge_\R \dots \wedge_\R (ie_{b^q}))^{*q}}$ of ${\cal L}$ and
1387: $\Lambda$. Under a change $e^{(q)}_i\rightarrow e'^{(q)}_i=M_qe_i$ 
1388: of the bases $e^{(q)}_1\dots e^{(q)}_{b^q}$ (where $M_q$ are invertible complex-
1389: linear operators in $H^q({\cal H})$), we have:
1390: \bea
1391: e'&:=&\otimes_{q}{(e'_q\wedge \dots \wedge e'_q)^{*q}} =\prod_{q}{(det M_q)^{(-1)^q}}e~~\nn\\
1392: \epsilon'&:=&\otimes_{q}{(e'_1\wedge_\R \dots \wedge_R e'_{b^q}
1393: \wedge'_\R (ie'_1)\wedge_\R \dots \wedge_\R (ie'_{b^q}))^{*q}}=
1394: |\prod_q{det M_q^{(-1)^q}}|^2\epsilon~~.
1395: \eea
1396: Considering the basis $e,ie$ of the real vector space ${\cal L}$, 
1397: we obtain a basis $\epsilon_0:=e\wedge_\R (ie)$ of $\Lambda^2_\R{\cal L}$, 
1398: which transforms as follows:
1399: \be
1400: \epsilon'_0:=e'\wedge_\R (ie')=|\prod_{q}{(det M_q)^{(-1)^q}}|^2 \epsilon_0~~.
1401: \ee
1402: The isomorphism $\phi:\Lambda^2_\R{\cal L}\rightarrow \Lambda$ 
1403: is obtained by taking $\epsilon_0$ into $\epsilon$.
1404: This is well-defined because $\epsilon$ and $\epsilon_0$ have the same 
1405: transformation rules. Note that both of the real lines $\Lambda$ and $\Lambda^2_\R{\cal L}$ 
1406: are equipped with orientations induced from the complex structure of 
1407: $H^q({\cal H})$, and the vectors $\epsilon_0$ and $\epsilon$ always agree 
1408: with these orientations.
1409: 
1410: Given the metrics $g$ on $L$ and $g_{\bf E}$ on 
1411: ${\bf E}$, we consider these constructions for 
1412: bases $e_1\dots e_{b^q}$  which are orthonormal 
1413: with respect to the Hermitian metrics $h^q=h_{1-q}$ which induced (through $f_\sigma$) on 
1414: $H^q({\cal H})$.
1415: In this case, only unitary transformations $M_q$ are allowed, hence 
1416: $\epsilon_0$ and $\epsilon$ are uniquely determined by the metric data (since $|det M_q|=1$), 
1417: while $e$ is determined up to multiplication by the phase factor $\prod_{q}{detM_q^{(-1)^q}}$. 
1418: By construction, the Hermitian $L_2$ metric on ${\cal L}$ is uniquely determined by the 
1419: constraint\footnote{Indeed, any element $v$ of ${\cal L}$ has the form 
1420: $u=\alpha e$ with $\alpha$ a complex constant, so relation (\ref{norm0}) 
1421: determines $||v||=|\alpha|$. Knowledge of the norm on ${\cal L}$ then 
1422: determines the metric in standard fashion.}:
1423: \be
1424: \label{norm0}
1425: ||e||_{\cal L}=1~~.
1426: \ee
1427: This induces a Euclidean metric $(.,.)_{\cal L}:=Re h_{\cal L}( .,.)$ on 
1428: the underlying real vector space, which makes $(e, ie)$ into a real 
1429: orthonormal basis of ${\cal L}$. Upon taking the second exterior power, 
1430: we have an induced Euclidean metric $(.,.)$ on $\Lambda_\R^2{\cal L}$, 
1431: which is uniquely determined by the constraint:
1432: \be
1433: \label{norm1}
1434: ||\epsilon_0||=1~~.
1435: \ee
1436: In fact, given $u=\alpha e$ and $v=\beta e$ in ${\cal L}$, 
1437: with $\alpha, \beta$ some complex constants, we have 
1438: $u\wedge_\R v=Im({\overline \alpha}\beta) \epsilon_0$ and 
1439: $||u\wedge_R v||=|Im({\overline \alpha}\beta)|$.
1440: 
1441: 
1442: On the other hand, one has an {\em Euclidean $L_2$-metric} $(.,.)_\Lambda$
1443: on $\Lambda$, which is uniquely determined by the condition:
1444: \be
1445: \label{norm2}
1446: ||\epsilon||_\Lambda=1~~.
1447: \ee
1448: This is the metric induced by the Euclidean scalar products
1449: $(.,.)_\sigma:=Re h_\sigma(.,.)$ 
1450: associated with the Hermitian metrics
1451: $h_\sigma$ on $H_\sigma({\cal H})$. Since $\phi$ maps $\epsilon_0$ into 
1452: $\epsilon$, it is clear from (\ref{norm1}) and (\ref{norm2}) that 
1453: $||\phi(v)||=||v||_\Lambda$. Hence the Euclidean $L_2$ metric is naturally 
1454: induced by the Hermitian $L_2$ metric via the (metric-independent) 
1455: isomorphism $\phi$. 
1456: 
1457: Conversely, the Euclidean $L_2$ metric determines the 
1458: vector $\epsilon$ (and thus the vector $\epsilon_0$), via relation 
1459: (\ref{norm2}). (The apparent sign ambiguity is fixed by 
1460: the condition that $\epsilon$ must agree with the orientation of $\Lambda$).
1461: Then $e$ is determined up to a phase factor by the relation 
1462: $e\wedge_\R (ie)=\epsilon_0$. Since multiplying $e$ by a 
1463: phase factor does not 
1464: affect the metric determined by relation (\ref{norm0}), it follows 
1465: that $||.||_{\cal L}$ is uniquely determined by the 
1466: Euclidean $L_2$ metric.
1467: 
1468: We conclude that the Hermitian and Euclidean $L_2$ metrics completely 
1469: determine each other. Which one we use is simply a matter of convention. 
1470: In this note, we use the metric $||.||_{\cal L}$, which is better adapted to 
1471: our complex formalism. In the ungraded case $({\bf E}=E_0$), 
1472: the Euclidean $L_2$ metric arises when considering the complex bundle
1473: as a real bundle by forgetting its complex structure. 
1474: 
1475: 
1476: 
1477: \subsubsection{Metric-independence of the Ray-Singer norm}
1478: 
1479: It turns out that the generalized Ray-Singer norm is {\em independent} 
1480: of $g$ and $g_{\bf E}$, and therefore completely independent of metric
1481: data. This statement, which parallels a well-known result for the ungraded 
1482: case due to Ray and Singer,  follows from the general discussion given in 
1483: \cite{Schwarz_resolvent} and \cite{Adams_Sen}. 
1484: The results of \cite{Adams_Sen} assure independence of $||.||_{RS}$
1485: of all metric data due to the following facts:
1486: 
1487: \
1488: 
1489: (1) The resolvent (\ref{res}) is {\em topological}, i.e. the complex 
1490: \be
1491: ~~0\rightarrow{\cal H}(N)
1492: \stackrel{T_N=d}{\longrightarrow}\dots 
1493: \stackrel{T_2=d}{\longrightarrow} 
1494: {\cal H}(1)\stackrel{T_1=d}{\longrightarrow}ker(d_{s=0})=ker(T_0)~~
1495: \ee
1496: is defined without reference to any metric.
1497: 
1498: 
1499: \
1500: 
1501: (2) The resolvent is also {\em elliptic}, i.e. the deformed Laplacians
1502: $\Delta_\sigma=T^t_\sigma T_\sigma+T_{\sigma+1}T_{\sigma +1}^t=
1503: (dd^\dagger +d^\dagger d)_{s=\sigma}$
1504: are elliptic operators for all $\sigma=0\dots N$.
1505: 
1506: \
1507: 
1508: (3) The base manifold $L$ is odd dimensional. 
1509: 
1510: \
1511: 
1512: The proof (which can be found in \cite{Adams_Sen}) 
1513: involves a combination of results in elliptic 
1514: operator theory with arguments in linear algebra.
1515: 
1516: 
1517: A particularly simple case arises when the background superconnection $A$
1518: is {\em acyclic}, i.e. the cohomology $H_d^*({\cal H})$ 
1519: vanishes in all degrees. In this situation, the quantity 
1520: ${\tilde Z}_{scl}$ coincides with $T(L, A)^{-1}$ and
1521:  is independent of all metric data. Examples of acyclic backgrounds are 
1522: provided by condensation of scalars in boundary condition changing sectors 
1523: of `topological brane-antibrane pairs', as discussed in detail in 
1524: \cite{bv} and \cite{gauge}. 
1525: 
1526: \subsubsection{The complex prefactor}
1527: 
1528: Let us turn to the complex prefactor $C$ given in (\ref{C}).
1529: A result of \cite{Adams_Sen} implies that:
1530: \be
1531: \zeta_{|T_0|}(0)=-\sum_{\sigma=0}^N{(-1)^\sigma dim H_\sigma({\cal R})}=
1532: \sum_{q=1-N}^1{(-1)^q dim H^q_d({\cal H})}~~,
1533: \ee
1534: which shows that the absolute value of $C$ is independent of metric data. \
1535: This shows that $|Z_{scl}|$ depends only on the metrics $h^H_\sigma$.
1536: 
1537: On the other hand, the value $\eta_{T_0}(0)$ may depend on $g$ and 
1538: $g_{\bf E}$, 
1539: which means that the phase factor of the semiclassical partition function 
1540: will generally carry a metric dependence. This parallels well-known results 
1541: valid for the ungraded case (standard Chern-Simons field theory)
1542: \cite{Witten_tcs}.
1543: 
1544: 
1545: \section{Example: D-brane pairs of unit relative grade 
1546: in a scalar background}
1547: 
1548: Consider a D-brane pair $(a,b)$ such that $grade(a)=0$ and $grade(b)=1$. 
1549: In this case, the underlying graded bundle is ${\bf E}=E_a\oplus E_b$, where 
1550: $E_a$ and $E_b$ are the flat bundles describing the reference 
1551: D-brane background. We let $A_a$ and $A_b$ be associated flat connections.
1552: The space ${\cal H}^q$ consists of sections of the bundle
1553: ${\cal V}^q=\Lambda^q(T^*L)\otimes End(E_a)\oplus \Lambda^q(T^*L)
1554: \otimes End(E_b)\oplus \Lambda^{q-1}(T^*L)
1555: \otimes Hom(E_a, E_b)\oplus \Lambda^{q+1}(T^*L)\otimes Hom(E_b, E_a)$. 
1556: Such elements can be viewed as matrices of bundle-valued forms:
1557: \be
1558: \label{u}
1559: u=\left[\begin{array}{cc}
1560: u_q&u_{q+1}\\
1561: u_{q-1}&{\hat u}_q
1562: \end{array}\right]~~,~~{\rm for}~~|u|=q~~,
1563: \ee
1564: where the subscript denotes form rank and the bundle 
1565: components of $u_q=u_{aa}, {\hat u}_q=u_{bb}, u_{q+1}=u_{ba}$ and 
1566: $u_{q-1}=u_{ab}$ satisfy $u_{\alpha\beta}\in 
1567: \Omega^{q+grade(\alpha)-grade(\beta)}(L, Hom(E_\alpha, E_\beta))$. 
1568: This corresponds to presenting $u$ as the direct sum:
1569: \be
1570: u=\oplus_{\alpha,\beta\in \{a,b\}}{u_{\alpha\beta}}~~.
1571: \ee
1572: 
1573: We shall consider backgrounds of the form:
1574: \be
1575: \label{phi}
1576: \phi=\left[\begin{array}{cc} 0&0\\\phi_0&0\end{array}\right]~~,~~|\phi|=1~~,
1577: \ee
1578: where $\phi_0$ is a
1579: zero-form valued in the bundle $Hom(E_a, E_b)$. In this case, the 
1580: equations of motion $d\phi+\phi\bullet \phi=0$ reduce to $d\phi_0=0$, which 
1581: means that $\phi_0$ is a covariantly-constant section of $Hom(E_a,E_b)$. 
1582: The shifted background is the flat superconnection 
1583: $A=(A_a\oplus A_b)+\phi$. We shall assume that $\phi_0$ is a 
1584: bundle morphism in the restricted sense, i.e. we require\footnote{This assumption 
1585: allows us to treat the kernel and cokernel of $\phi_0$ as subbundles of 
1586: $E_a$ and $E_b$. Allowing maps $\phi_0$ of non-constant rank leads to situations
1587: which are better described in terms of sheaf theory. This is very similar to the case of holomorphic 
1588: bundles vs. holomorphic sheaves. Note that constancy of $rk \phi_0(p)$ 
1589: is a purely technical assumption -- there is no {\em physical} 
1590: reason to restrict to scalar backgrounds of constant rank.} that
1591: the rank of the fiber maps $\phi_0(p):(E_a)_p\rightarrow (E_b)_p$ 
1592: is independent of the point $p\in L$. 
1593: 
1594: Let us assume that the reference flat bundles $(E_a, A_a)$ and $(E_b, A_b)$ 
1595: are unitary, i.e. they admit covariantly-constant Hermitian metrics
1596: \footnote{In 
1597: physical language, this means that we are dealing with unitary connections, 
1598: i.e. connections whose matrices are anti-Hermitian in appropriate 
1599: local frames.}.
1600: We shall pick the metric $g_{\bf E}$
1601: on ${\bf E}$ to be induced by two such metrics. 
1602: With this hypothesis, 
1603: it was showed in \cite{gauge} that the deformed Laplacian 
1604: $\Delta_\phi=d_\phi d_\phi^\dagger+
1605: d_\phi^\dagger d_\phi$ in the  background $\phi_0$ has the form:
1606: \be
1607: \label{DD}
1608: \Delta_\phi u=
1609: \left[\begin{array}{cc}
1610: \Delta_\phi^{(aa)}u_q&\Delta_\phi^{(ba)}u_{q+1}\\
1611: \Delta_\phi^{(ab)}u_{q-1}&\Delta_\phi^{(bb)}{\hat u}_q
1612: \end{array}\right]~~,
1613: \ee
1614: with:
1615: \bea
1616: \Delta_\phi^{(\alpha\beta)}u_{\alpha\beta}&=&\Delta u_{\alpha\beta} +
1617: u_{\alpha\beta}\circ t^{(\alpha)}+
1618: t^{(\beta)}\circ u_{\alpha\beta}~~,
1619: \eea
1620: where we defined:
1621: \be
1622: t^{(a)}=\phi_0^\dagger \phi_0\in End(E_a)~~{\rm~and~}~~t^{(b)}=\phi_0\phi_0^\dagger \in End(E_b)~~
1623: \ee
1624: and where $\Delta u_{\alpha\beta}$ stands for the form 
1625: Laplacian coupled to the flat connection induced by $A_\alpha$ and $A_\beta$ 
1626: on the bundle $Hom(E_\alpha,E_\beta)$. In direct sum notation, we have:
1627: \be
1628: \Delta_\phi=\oplus_{\alpha\beta}{\Delta_\phi^{(\alpha\beta)}}~~.
1629: \ee
1630: The fact that the deformed Laplacian decomposes in this manner is a characteristic of 
1631: scalar background of the particular form (\ref{phi}).
1632: 
1633: 
1634: Since the subspaces $\Omega^*(L, Hom(E_\alpha,E_\beta))\subset {\cal H}$  
1635: are mutually orthogonal with respect to the scalar product $h$, we have:
1636: \be
1637: p_q:=det^{',reg}_{|.|=q}\Delta_\phi=
1638: det^{',reg}_{rk =q}\Delta_\phi^{(aa)}
1639: det^{',reg}_{rk =q}\Delta_\phi^{(bb)}
1640: det^{',reg}_{rk =q-1}\Delta_\phi^{(ab)}
1641: det^{',reg}_{rk =q+1}\Delta_\phi^{(ba)}~~
1642: \ee
1643: and:
1644: \be
1645: T(L,A)^2=\prod_{q=1-N}^{N+2}{(det^{',reg}_{|.|=q}\Delta_\phi)^{q(-1)^q}}=
1646: \prod_{q=1-N}^{N+2}{p_q^{q (-1)^q}}~~.
1647: \ee
1648: 
1649: It was showed in \cite{gauge} that the kernel $K$ of $\Delta_\phi$ 
1650: coincides (up to a twist of gradings) 
1651: with the space of harmonic forms valued in the 
1652: flat bundle $End(R\oplus I^\perp)$, where $R=ker \phi$ and 
1653: $I=im \phi$ are (flat) subbundles of $E_a$ and $E_b$:
1654: {\scriptsize \be
1655: K^q=\Omega^q_{harm}(L, End(R))\oplus \Omega^q_{harm}(L, End(I^\perp))\oplus 
1656: \Omega^{q-1}_{harm}(L, Hom (R, I^\perp))\oplus \Omega^{q+1}_{harm}
1657: (L, Hom(I^\perp, K))~~.
1658: \ee}
1659: Using the covariantly-constant metrics 
1660: on $E_a$ and $E_b$, we identify $I^\perp$ with 
1661: the cokernel $Q$ of the flat bundle map $\phi_0:E_a\rightarrow E_b$:
1662: \be
1663: I^\perp\approx Q:=coker \phi_0=E_b/im \phi_0~~.
1664: \ee
1665: Hodge theory leads to identifications $H^q_{d_\phi}({\cal H})=K^q$ and 
1666: $\Omega^k_{harm}(V)\approx H^k(V)$, where $V$ is any of the 
1667: flat bundles $End(R)$, $End(I^\perp)$, $Hom(R, I^\perp)$ and 
1668: $Hom(I^\perp, R)$, while $H^*(V)$ stands for the usual cohomology with 
1669: coefficients in the local system determined by $V$. Combining everything, 
1670: we obtain:
1671: {\scriptsize \bea
1672: \label{Hd}
1673: H^q_{d_\phi}({\cal H})=H^q(L, End(R))\oplus H^q(L, End(Q))\oplus 
1674: H^{q-1}(L, Hom (R, Q))\oplus H^{q+1}(L, Hom(Q, R))~~.
1675: \eea}
1676: Using (\ref{Hd}), we find that the 
1677: Ray-Singer metric is defined on the determinant line:
1678: {\footnotesize \bea
1679: {\cal L}&=&det H^*_{d_\phi}({\cal H})=
1680: \otimes_q{\Lambda^{max}H^q_{d_\phi}({\cal H})^{* q}}=\\
1681: &=&det H^*(L, End(R))\otimes det H^*(L,End(Q))
1682: \otimes [det H^*(L,Hom(R, Q))]^*\otimes [det H^*(L, Hom(Q, R))]^*~~.\nn
1683: \eea}
1684: 
1685: \subsection{The case of trivial flat bundles}
1686: 
1687: Let us consider the particularly simple 
1688: case when $E_a$ and $E_b$ (of ranks $r_a$ and $r_b$) 
1689: are trivial as flat vector 
1690: bundles, so that all components $u_k, {\hat u}_k, u_{k+1}, u_{k-1}$ can be 
1691: viewed as operator-valued forms, and $d$ coincides with the de Rham 
1692: differential acting on such forms. In this situation, the condition 
1693: $d\phi_0=0$ means that $\phi_0$ is a {\em constant} linear operator from 
1694: $\C^{r_a}$ to $\C^{r_b}$. Since the flat bundle structure is trivial, we 
1695: expect to obtain a particularly simple expression for $T(L,A)$. 
1696: We show below that this expectation is fulfilled. 
1697: 
1698: For this, we first notice that the non-negative operators
1699: $t^{(a)}=\phi_0^\dagger \phi_0\in End(\C^{r_a})$ 
1700: and $t^{(b)}=\phi_0\phi_0^\dagger\in End(\C^{r_b})$ have the same nonzero eigenvalues. 
1701: This follows form the following variant of the argument used in Subsection 
1702: 3.4.2. Noting that $ker (\phi_0^\dagger \phi_0)^\perp=ker (\phi_0)^\perp$ 
1703: and $ker (\phi_0\phi_0^\dagger)^\perp=ker (\phi_0^\dagger)^\perp
1704: =im(\phi_0) $, the restriction of $\phi_0$ gives an isomorphism:
1705: \be
1706: \phi_0:ker(\phi_0^\dagger \phi_0)^\perp\stackrel{\approx}{\rightarrow}
1707: ker (\phi_0\phi_0^\dagger)^\perp~~.
1708: \ee
1709: This induces a bijection between the non vanishing eigenvalues of 
1710: $\phi_0^\dagger\phi_0$ and $\phi_0\phi_0^\dagger$, since 
1711: $\phi_0^\dagger \phi_0 v=\lambda v$ for some $\lambda\in \R_+^*$ 
1712: and $v\in  \C^{r_a}$
1713: implies $\phi_0\phi_0^\dagger(\phi v)=\lambda \phi_0 v$. 
1714: In particular, the ranks of the two operators coincide:
1715: \be
1716: rk(\phi_0^\dagger \phi_0)=rk(\phi_0 \phi_0^\dagger):=\rho~~,
1717: \ee
1718: while their defects are given by:
1719: \be
1720: dim~ker (\phi_0^\dagger \phi_0)=rk R=r_a-\rho~~,~~
1721: dim~ker (\phi_0 \phi_0^\dagger)=rk Q=r_b-\rho~~.
1722: \ee
1723: 
1724: 
1725: Consider unitary transformations $S_a$ and $S_b$ of $\C^{r_a}$ and $\C^{r_b}$ 
1726: which diagonalize $\phi_0^\dagger \phi_0$ and $\phi_0\phi_0^\dagger$:
1727: \bea
1728: \label{S_tf}
1729: &&S_a\phi_0^\dagger \phi_0 S_a^{-1}=D^{(a)}:=diag(\lambda_1^{(a)}\dots 
1730: \lambda^{(a)}_{r_a})~~\nn\\
1731: &&S_b\phi_0 \phi_0^\dagger S_b^{-1}=D^{(b)}:=
1732: diag(\lambda^{(b)}_1\dots \lambda^{(b)}_{r_b})~~,
1733: \eea
1734: with $\lambda_i^{(a)}, \lambda^{(b)}_j\geq 0$.
1735: In view of the above, we can always pick $S_a$ and $S_b$ such that:
1736: \bea
1737: \lambda^{(a)}_i&=&\lambda^{(b)}_i:=\lambda_i~{\rm~for~}~i=1\dots 
1738: \rho~~{\rm and}\nn\\
1739: \lambda^{(a)}_i&=&0~{\rm~for~}~i=\rho+1 \dots r_a~~\\
1740: \lambda^{(b)}_j&=&0~{\rm~for~}~j=\rho+1 \dots r_b~~.\nn
1741: \eea
1742: 
1743: It is easy to see that the unitary transformation 
1744: $S=S_a\oplus S_b\equiv \left[\begin{array}{cc}S_a&0\\0&S_b\end{array}\right]$
1745: preserves the decomposition (\ref{DD}) of $\Delta_\phi$ 
1746: while bringing $\Delta_\phi^{(\alpha\beta)}$ to the form:
1747: \be
1748: \Delta_\phi^{(\alpha\beta)}u_{\alpha\beta} = 
1749: \Delta u_{\alpha\beta}+D^{(\beta)}\circ u_{\alpha\beta} +
1750: u_{\alpha\beta} \circ D^{(\alpha)}~~,
1751: \ee
1752: which act on the matrix components\footnote{Picking orthonormal 
1753: bases $e_i^{(a)}$ and $e_j^{(b)}$ of $\C^{r_a}$ and $\C^{r_b}$, we
1754: define $u_{\alpha\beta}^{ji}$ through $u_{\alpha\beta}(e_i^{(\alpha)})=
1755: u_{\alpha\beta}^{ji}e_j^{(\beta)}$.} of $u_{\alpha\beta}$ as:
1756: \be
1757: \Delta_\phi^{(\alpha\beta)}u^{ji}_{\alpha\beta} = \Delta u^{ji}_{\alpha\beta}+
1758: (\lambda^{(\alpha)}_i+\lambda^{(\beta)}_j)u^{ji}_{\alpha\beta}~~.
1759: \ee
1760: This leads to the expression:
1761: \be
1762: p_q=p_q^{(1)}p_q^{(2)}p_q^{(3)}~~,
1763: \ee
1764: where:
1765: \bea
1766: p_q^{(1)}&=&det^{',reg}_{rk=q}(\Delta)^{(r_a-\rho)^2+(r_b-\rho)^2}
1767: \left[det^{',reg}_{rk=q-1}(\Delta)
1768: det^{',reg}_{rk=q+1}(\Delta)\right]^{(r_a-\rho)(r_b-\rho)}~~\\
1769: p_q^{(2)}&=&
1770: \prod_{i,j=1}^\rho{\left[
1771: det^{',reg}_{rk=q}(\Delta +\lambda_i+\lambda_j)\right]^2
1772: det^{',reg}_{rk=q-1}(\Delta +\lambda_i+\lambda_j)
1773: det^{',reg}_{rk=q+1}(\Delta +\lambda_i+\lambda_j)}~~\nn\\
1774: p_q^{(3)}&=&\prod_{i=1}^{\rho}{
1775: det^{',reg}_{rk=q}(\Delta +\lambda_i)^{2(r_a+r_b-2\rho)}
1776: \left[det^{',reg}_{rk=q-1}(\Delta +\lambda_i)
1777: det^{',reg}_{rk=q+1}(\Delta +\lambda_i)\right]^{r_a+r_b-2\rho}}~~.\nn
1778: \eea
1779: It is easy to check that: 
1780: \be
1781: \prod_q{\left[det^{',reg}_{rk=q-1}(\Delta +\mu)
1782: det^{',reg}_{rk=q+1}(\Delta +\mu)\right]^{q(-1)^q}}=
1783: \left(\prod_q{\left[det^{',reg}_{rk=q}(\Delta +\mu)\right]^{q(-1)^q}}\right)
1784: ^{-2}~~.
1785: \ee
1786: Therefore, one has
1787: $\prod_{q}{(p_q^{(2)})^{q(-1)^q}}=
1788: \prod_{q}{(p_q^{(3)})^{q(-1)^q}}=1$ and we obtain:
1789: \be
1790: \label{final}
1791: T(L,A)=\left[\prod_q{(p_q)^{q(-1)^q}}\right]^{1/2}=
1792: \left[\prod_q{(p_q^{(1)})^{q(-1)^q}}\right]^{1/2}=
1793: T(L)^{(r_a-r_b)^2}
1794: ~~.
1795: \ee
1796: where $T(L)$ is the usual Ray-Singer torsion of $L$
1797: (i.e. the standard analytic torsion for the trivial flat complex line bundle 
1798: ${\cal O}_L$ over $L$):
1799: \be
1800: T(L)=\prod_q{det^{',reg}_{rk=q}(\Delta)^{\frac{q(-1)^q}{2}}}=
1801: e^{-\frac{1}{2}\sum_{q=0}^3{(-1)^{q} q \zeta'_{\Delta_q'}(0)}}~~,
1802: \ee
1803: with $\Delta_q$ the Laplacian on $\Omega^q(L)$.  
1804: The determinant line can be found by noticing that 
1805: $R={\cal O}_L^{\oplus (r_a-\rho)}$ and $Q={\cal O}_L^{\oplus (r_b-\rho)}$,
1806: where ${\cal O}_L$ is the trivial complex flat line bundle over $L$. 
1807: This implies 
1808: $End(R)={\cal O}_L^{\oplus (r_a-\rho)^2}$, 
1809: $End(Q)={\cal O}_L^{\oplus (r_b-\rho)^2}$
1810: and 
1811: $Hom(R,Q)=Hom(Q,R)={\cal O}_L^{\oplus (r_a-\rho)(r_b-\rho)}$
1812: (as flat line bundles), thereby giving the determinant line:
1813: \be
1814: {\cal L}=det H^*(L, {\cal O}_L)^{\otimes (r_a-r_b)^2}=
1815: det H^*(L)^{\otimes (r_a-r_b)^2}~~.
1816: \ee
1817: Combining with (\ref{final}), we see that $T(L,A)$ is the norm induced 
1818: on $det H^*(L, {\cal O}_L)^{\otimes (r_a-r_b)^2}$ by the 
1819: usual Ray-Singer norm on $det H^*(L, {\cal O}_L)=det H^*(L)$ (considered
1820: in the `complex' formalism discussed in Subsection 3.4.3).
1821: 
1822: 
1823: {\bf Observation} 
1824: If $r_a=r_b$, then it was showed in 
1825: \cite{bv, gauge} that a background $\phi_0$ which is a flat isomorphism 
1826: leads to an acyclic composite, i.e. the complex 
1827: of the shifted differential $d_\phi$ is acyclic in all degrees. 
1828: In this case (and with trivial flat connections $A_a$ and $A_b$), 
1829: the generalized Ray-Singer torsion is equal to one. 
1830: 
1831: \acknowledgments{I thank R. Roiban for comments on the manuscript and 
1832: Martin Rocek for interest and support.
1833: The present work was supported by the Research Foundation under NSF 
1834: grant PHY-9722101.}
1835: 
1836: \
1837: 
1838: \begin{thebibliography}{100}
1839: \bibitem{BZ}{J.~M.~Bismut, W.~Zhang, {\em An extension of a theorem of 
1840: Cheeger and Muller}, Asterisque {\bf 205} (1992).}
1841: \bibitem{gauge}{ C. I. Lazaroiu, R. Roiban, {\em 
1842: Holomorphic potentials for graded D-branes},  hep-th/0110288.} 
1843: \bibitem{gbv}{C.~I.~Lazaroiu, R.~Roiban, {\em 
1844: Holomorphic potentials for graded D-branes}, 
1845: JHEP 02(2002) 038, hep-th/0110288. }
1846: \bibitem{Witten_tcs}{E.~Witten, {\em Quantum field theory and the 
1847: Jones polynomial}, Commun Math Phys {\bf 121} (1989) 351.}
1848: \bibitem{Ray}{D.~B.~Ray, {\em Reidemeister torsion and the Laplacian 
1849: on lens spaces}, Adv. in Math. {\bf 4} (1970), 109--126.}
1850: \bibitem{RS1}{D.~B.~Ray, I.~M.~Singer, 
1851: {\em R-torsion and the Laplacian on Riemannian manifolds}, 
1852: Adv. Math {\bf 7} (1971), 145-210.}
1853: \bibitem{RS_symp}{D.~B.~Ray, I.~M.~Singer, {\em Analytic torsion}, 
1854: Proceedings of Symposia in Pure Mathematics, vol. {\bf 23}, p 167, 
1855: American Mathematical Society 1973.}
1856: \bibitem{ST}{A. S. Schwarz, I.S. Tyupkin, {\em Quantization of antisymmetric
1857:       tensors and Ray-Singer Torsion}, Nucl. Phys. {\bf B 242} (1984) 436.}
1858: \bibitem{Schwarz_resolvent}{A.~Schwarz, 
1859: {\em The partition function of a degenerate quadratic functional
1860: and Ray-Singer invariants}, Lett. Math. Phys, {\bf 2} (1978), 247--252.} 
1861: \bibitem{Schwarz_resolvent2}{A.~Schwarz, 
1862: {\em The partition function of a degenerate functional}, 
1863: Commun. Math. Phys. {\bf 67} (1979), 1--16.} 
1864: \bibitem{Adams_Sen0}{D.~H. Adams, S.~Sen, 
1865: {\em Phase and scaling properties of determinants
1866: arising in topological field theories},
1867: Phys. Lett. {\bf B 353} (1995) no. 4, 495--500, hep-th/9506079.}
1868: \bibitem{Adams_Sen}{D.~H. Adams, S.~Sen, 
1869: {\em Partition Function of a Quadratic Functional and Semiclassical 
1870: Approximation for Witten's 3-Manifold Invariant}, hep-th/9503095.}
1871: \bibitem{Adams_FP}{D.~H.~Adams, {\em 
1872: A note on the Faddeev-Popov determinant and Chern-Simons perturbation theory},
1873: Lett.Math.Phys. {\bf 42} (1997) 205-214, hep-th/9704159.} 
1874: \bibitem{Adams_scl}{D.~H.~Adams, 
1875: {\em The semiclassical approximation for the Chern-Simons partition function}, 
1876: Phys.Lett. B417 (1998) 53-60, hep-th/9709147.}
1877: \bibitem{Adams_cpct}{D.~H.~Adams, 
1878: {\em Perturbative expansion in gauge theories on compact manifolds}, 
1879: hep-th/9602078.} 
1880: \bibitem{NC}{C.~Nash, D.~O' Connor, 
1881: {\em BRST Quantization and the Product Formula for the Ray-Singer Torsion},
1882: Int.J.Mod.Phys. {\bf A10} (1995) 1779-1806, hep-th/9310038.}
1883: \bibitem{GK}{ J.~Gegenberg, G.~Kunstatter, 
1884: {\em The Partition Function for Topological Field Theories}, 
1885: Ann. Phys. {\bf 231} (1994) 270-289, hep-th/9304016.}
1886: \bibitem{Blau_Thompson}{M.~Blau, G.~Thompson, 
1887: Ann. Phys {\bf 205}(1991) 130.}
1888: \bibitem{bv}{C.~I.~Lazaroiu, R.~Roiban and D.~Vaman, 
1889: {\em Graded Chern-Simons field theory and graded topological
1890: D-branes}, hep-th/0107063.}
1891: \bibitem{BK}{A.~Bondal, M.~M. Kapranov, {\em Enhanced triangulated categories},
1892: Math. USSR Sbornik, vol 70 (1991) no 1, 93.}
1893: \bibitem{Douglas_Diaconescu}{D.~E.~Diaconescu, M.~Douglas, 
1894: {\em D-branes on Stringy Calabi-Yau Manifolds}, hep-th/0006224.}
1895: \bibitem{Douglas_Kontsevich}{M.~R.~Douglas, {\em D-branes, Categories
1896:       and N=1 Supersymmetry}, hep-th/0011017.}  
1897: \bibitem{com1}{
1898:     C.~I.~Lazaroiu, {\em Generalized complexes and string field
1899:     theory}, JHEP 06 (2001) 052.}  
1900: \bibitem{com3}{C.~I.~Lazaroiu, {\em
1901:     Unitarity, D-brane dynamics and D-brane categories},
1902:     hep-th/0102183. } 
1903: \bibitem{Aspinwall}{P.~S.~Aspinwall,
1904:     A.~Lawrence, {\em Derived Categories and Zero-Brane Stability},
1905:     hep-th/0104147.}  
1906: \bibitem{Diaconescu}{D.~E.~Diaconescu, {\em
1907:     Enhanced D-Brane Categories from String Field Theory},
1908:     hep-th/0104200.}  
1909: \bibitem{Douglas_Aspinwall}{P.~S.~Aspinwall, M.~R.~Douglas, 
1910: {\em D-Brane Stability and Monodromy}, hep-th/0110071.}
1911: \bibitem{Fukaya}{K.~Fukaya, {\em Morse homotopy,
1912:     $A^\infty$-category and Floer homologies}, in {\em Proceedings of
1913:     the GARC Workshop on Geometry and Topology}, ed. by H.~J.~Kim,
1914:     Seoul national University (1994), 1-102; {\em Floer homology,
1915:     $A^\infty$-categories and topological field theory}, in {\em
1916:     Geometry and Physics}, Lecture notes in pure and applied
1917:     mathematics, {\bf 184}, pp 9-32, Dekker, New York, 1997; {\em
1918:     Floer homology and Mirror symmetry, I}, preprint available at
1919:     $http://www.kusm.kyoto-u.ac.jp/~\tilde{}~fukaya/fukaya.html.$}
1920: \bibitem{Fukaya2}{K. Fukaya, Y.-G. Oh, H.~Ohta, K.~Ono, {\em
1921:       Lagrangian intersection Floer theory - anomaly and obstruction},
1922:     preprint available at
1923:     $http://www.kusm.kyoto-u.ac.jp/~\tilde{}~fukaya/fukaya.html.$}
1924: \bibitem{Zaslow_Polishchuk}{A.~Polishchuk, E.~Zaslow, 
1925: {\em Categorical mirror symmetry: the elliptic curve}, 
1926: math.AG/9801119.}
1927: \bibitem{Kontsevich}{M.~Kontsevich,
1928:     {\em Homological algebra of mirror symmetry}, Proceedings of the
1929:     International Congress of Mathematicians, (Zurich, 1994),
1930:     120--139, Birkhauser, alg-geom/9411018.}
1931: \bibitem{Witten_CS}{ E.~Witten,{\em
1932:     Chern-Simons gauge theory as a string theory}, The Floer memorial
1933:     volume, 637--678, Progr. Math., 133, Birkhauser, Basel, 1995,
1934:     hep-th/9207094.}  
1935: \bibitem{Seidel}{P.~Seidel, {\em Graded
1936:     Lagrangian submanifolds}, Bull. Soc. Math. France {\bf 128}
1937:     (2000), 103-149, math.SG/9903049.}
1938: \bibitem{Bismut_Lott}{J.~M.~Bismut and J.~Lott, {\em Flat vector
1939:       bundles, direct images and higher analytic torsion}, J. Amer.
1940:     Math Soc {\bf 8} (1992) 291.}  
1941: \bibitem{Quillen}{D.~Quillen, {\em
1942:     Superconnections and the Chern character}, Topology, {\bf 24},
1943:     No.1.(1085), 89-95.}  
1944: \bibitem{superpot}{C. I. Lazaroiu, {\em String field theory and brane 
1945: superpotentials}, hep-th/0107162.}
1946: \bibitem{Witten_nlsm}{ E.~Witten, {\em Topological sigma models},
1947:     Commun. Math. Phys.  {\bf 118} (1988),411.}
1948: \bibitem{Witten_mirror}{E.~Witten, {\em Mirror manifolds and
1949:       topological field theory}, Essays on mirror manifolds, 120--158,
1950:     Internat. Press, Hong Kong, 1992, hep-th/9112056.}
1951: \bibitem{sc}{C.~I.~Lazaroiu, {\em Graded
1952:     Lagrangians, exotic topological D-branes and enhanced triangulated
1953:     categories}, JHEP 0106 (2001) 064.}
1954: \end{thebibliography}
1955: \end{document}
1956: 
1957: \bibitem{Camporesi}{R.~Camporesi, 
1958: {\em Harmonic analysis and propagators on homogeneous spaces}, 
1959: Phys. Rept. {\bf 196} (1990):1-134.}
1960: \bibitem{BCVZ}{A.~A.~Bytsenko, G.~Cognola, L.~Vanzo, S.~Zerbini, 
1961: {\em Quantum fields and extended objects in Space-Times with constant 
1962: curvature spatial section}, Phys.Rept. {\bf 266} 
1963: (1996) 1--126, hep-th/9905061.}
1964: 
1965: 
1966: 
1967: 
1968: 
1969: 
1970: 
1971: 
1972: 
1973: 
1974: 
1975: 
1976: 
1977: 
1978: 
1979: