1: \documentclass[letterpaper]{article}
2: \usepackage[margin=2cm]{geometry}
3: \usepackage{array}
4: %%\usepackage{rotating}
5: \usepackage[leqno]{amsmath}
6: \usepackage{amsfonts,amssymb,amsthm}
7: \usepackage{amscd,amsxtra}
8: %%\usepackage{diagrams}
9: %%\diagramstyle[amstex]
10:
11:
12: %%%%%%%%%%%%%%%%%%%%%%%
13: %% euler script for mathcal
14: \usepackage{eucal}
15: %%\usepackage[mathscr]{eucal}
16:
17: %%\usepackage{stmaryrd}
18: %%\usepackage{mathbbol}
19: \usepackage{mathrsfs} %provides \mathscr
20:
21: %%%%%%%%%%%%%%%%%%%%%%% concrete fonts and math
22: %%\usepackage{beton}
23: %%\usepackage{euler}
24: %%\usepackage{concmath}
25: %%\usepackage{concrete}
26:
27: %%%%%%%%%%%%%%%%%%%%%%% postscript fonts
28: %%\usepackage{times}
29: %%\usepackage{mathtime}
30: %%\usepackage{pslatex}
31: %%\usepackage{mathptm}
32: %%\usepackage[scaled=.92]{helvet}
33: %%\renewcommand{\familydefault}{\sfdefault}
34:
35: %%%%%%%%%%%%%%%%%%%%%%% other
36: %%\usepackage{comment}
37:
38: %%XyPic
39: \usepackage[all]{xy}
40: %%\CompileMatrices %xypic: precompile diagrams
41: \SelectTips{cm}{} %xypic: arrowheads with cm fonts (like AmsCD)
42:
43: %%\usepackage{color,graphicx}
44:
45: %%\usepackage[colorlinks=true]{hyperref}
46: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
47: \newtheorem{theorem}{Theorem}[section]
48: \newtheorem{proposition}[theorem]{Proposition}
49: \newtheorem{lemma}[theorem]{Lemma}
50: \newtheorem{corollary}[theorem]{Corollary}
51: %%
52: %%\newtheorem{alphthm}{Theorem}
53: %%\renewcommand{\thealphthm}{\Alph{alphthm}}
54: %%
55: \theoremstyle{definition}
56: \newtheorem{definition}[theorem]{Definition}
57: \newtheorem{example}[theorem]{Example}
58: %%\newtheorem*{assumption}{Assumption}
59: %%
60: \theoremstyle{remark} \newtheorem{remark}[theorem]{Remark}
61:
62: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
63: \numberwithin{equation}{section}
64: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
65:
66: %% Curly (=cal in the sense of mathrsfs) symbol
67: \newcommand{\curly}[1]{\mathscr{#1}}
68: \newcommand{\cD}{\curly{D}}
69: \newcommand{\cL}{\curly{L}}
70:
71:
72: %% fields and other notable objects requiring a blackboard
73: %% boldface symbol
74: \newcommand{\field}[1]{\ensuremath{\mathbb{#1}}}
75: \newcommand{\ZZ}{\field{Z}}
76: \newcommand{\RR}{\field{R}}
77: \newcommand{\CC}{\field{C}}
78: \newcommand{\HH}{\field{H}}
79: \newcommand{\PP}{\field{P}}
80: \newcommand{\TT}{\field{T}}
81:
82: %% objects related to complex analysis and alg. curves
83: \newcommand{\projspace}[1]{\PP^{#1}}
84: \newcommand{\pione}{\projspace{1}}
85:
86: %% cross-ratio
87: \newcommand{\CR}[4]{[#1\colon #2\colon #3\colon #4]}
88:
89: %% differential complexes
90: \newcommand{\complex}[1]{\mathsf{#1}} %%
91: \newcommand{\CCC}{\complex{C}}
92: %% \newcommand{\complex}[2][\bullet]{{#2}^{#1}}
93: %% \newcommand{\CCC}{{\complex[{}]{C}}}
94: %% Cech complexes
95: \newcommand{\sheafcech}{\Check{\curly{C}}}
96:
97: %% Categories
98: \newcommand{\cat}[1]{\mathsf{#1}}
99:
100: %% Sheaves
101: \newcommand{\sheaf}[1]{\underline{\mathnormal{#1}}}
102: %%\newcommand{\locsys}[1]{\mathbb{#1}}
103: %%\newcommand{\locsys}[1]{\mathrm{#1}}
104: \newcommand{\sha}[2][\bullet]{\sheaf{A}_{#2}^{#1}}
105: \newcommand{\shomega}[2][\bullet]{\sheaf{\Omega}_{#2}^{#1}}
106: \newcommand{\she}[2][\bullet]{\sheaf{\mathcal{E}}_{#2}^{#1}}
107: \newcommand{\sho}[1]{\mathcal{O}_{#1}}
108: \newcommand{\deligne}[3][\bullet]{#2(#3)^{#1}_\mathcal{D}}
109: \newcommand{\deltilde}[3][\bullet]{%
110: \smash[t]{\widetilde{#2(#3)}}^{#1}_\mathcal{D}}
111: \newcommand{\delub}[3][\bullet]{%
112: \smash[b]{\underline{#2(#3)}}^{#1}_\mathcal{D}}
113: \newcommand{\dhh}[2][\bullet]{D(#2)_\mathit{h.h.}^{#1}}
114: \newcommand{\delH}[4][\bullet]{H^{#1}_\mathcal{D}(#2, #3(#4))}
115: \newcommand{\dhhH}[3][\bullet]{%
116: H^{#1}_{\mathcal{D}_\mathit{h.h.}}(#2,#3)}
117:
118: %% quasi-isomorphism
119: \newcommand{\qi}{\xrightarrow{\simeq}}
120: \newcommand{\lqi}{\overset{\simeq}{\longrightarrow}}
121:
122: %% hypercohomology
123: \newcommand{\hyper}[1]{\mathbf{#1}}
124: \newcommand{\HHH}{\hyper{H}}
125: \newcommand{\RRR}{\hyper{R}}
126: \newcommand{\LLL}{\hyper{L}}
127:
128: %% Cech coverings
129: \newcommand{\cover}[1]{\mathcal{#1}}
130:
131: %% differentials
132: \newcommand{\del}{\partial} %usual partial
133: \newcommand{\delb}{\Bar\partial} %same with bar
134: \newcommand{\deltacheck}{\Check\delta} %for Cech complexes
135: %% double complexes
136: \newcommand{\delp}{\partial^\prime}
137: \newcommand{\delpp}{\partial^{\prime\prime}}
138: \newcommand{\deltap}{\d}%
139: \newcommand{\deltapp}{\delta}
140:
141: %% operator names
142: %% \renewcommand{\d}{\operatorname{d}}
143: \DeclareMathOperator{\dd}{d} \DeclareMathOperator{\id}{id}
144: \DeclareMathOperator{\I}{I} \DeclareMathOperator{\Tot}{Tot}
145: \DeclareMathOperator{\Hom}{Hom} \DeclareMathOperator{\Tor}{Tor}
146: \DeclareMathOperator{\Ext}{Ext} \DeclareMathOperator{\Ker}{Ker}
147: \DeclareMathOperator{\cone}{Cone} \DeclareMathOperator{\Coker}{Coker}
148: \DeclareMathOperator{\im}{Im} \DeclareMathOperator{\SL}{SL}
149: \DeclareMathOperator{\PSL}{PSL} \DeclareMathOperator{\SO}{SO}
150: \DeclareMathOperator{\Ad}{Ad} \DeclareMathOperator{\ad}{ad}
151: \DeclareMathOperator{\deck}{Deck}
152:
153: %% miscellaneous
154: \newcommand{\tame}[2]{\bigl(#1,#2\bigr]}
155: \newcommand{\tate}{2\pi\sqrt{-1}}
156: \newcommand{\bil}[2]{\bigl(#1\,\vert\,#2\bigr)}
157: \newcommand{\dual}[2]{\langle#1\,,\,#2\rangle}
158: \newcommand{\abs}[1]{\lvert#1\rvert}
159: \newcommand{\norm}[1]{\lVert#1\rVert}
160: \newcommand{\interior}{\lrcorner}
161: \newcommand{\onehalf}{\frac{1}{2}}
162: \newcommand{\onefourth}{\frac{1}{4}}
163: \newcommand{\ihalf}{\frac{\sqrt{-1}}{2}}
164: \newcommand{\eqdef}{\overset{\mathrm{def}}{=}}
165: \newcommand{\pic}[1]{\mathrm{Pic}(#1)}
166: \newcommand{\cm}[1]{\mathscr{CM}(#1)}
167: \newcommand{\li}{\mathit{Li}_2}
168: \newcommand{\bwli}{{\cL}_2}
169: \newcommand{\bwd}{{\cD}_2}
170:
171: %% other
172: \newcommand{\bei}{Be\u\i{}linson}
173: \newcommand{\cech}{\v{C}ech}
174:
175: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
176: \setcounter{tocdepth}{2}
177: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
178:
179: \title{On hermitian-holomorphic classes related to uniformization, the
180: dilogarithm, and the Liouville Action}
181:
182:
183: \author{Ettore Aldrovandi\\
184: Department of Mathematics\\
185: Florida State University\\
186: Tallahassee, FL 32306-4510, USA\\
187: \texttt{aldrovandi@math.fsu.edu} }
188:
189: %%\author{Ettore Aldrovandi}
190: %%\address{Department of Mathematics\\
191: %% Florida State University\\
192: %% Tallahassee, FL 32306-4510, USA}
193: %%\email{aldrovandi@math.fsu.edu}
194:
195: \date{}
196:
197: \begin{document}
198:
199: \maketitle
200:
201: \begin{abstract}
202: Metrics of constant negative curvature on a compact Riemann surface
203: are critical points of the Liouville action functional, which in
204: recent constructions is rigorously defined as a class in a
205: \cech-de~Rham complex with respect to a suitable covering of the
206: surface.
207:
208: We show that this class is the square of the metrized holomorphic
209: tangent bundle in hermitian-holomorphic Deligne cohomology. We achieve
210: this by introducing a different version of the hermitian-holomorphic
211: Deligne complex which is nevertheless quasi-isomorphic to the one
212: introduced by Brylinski in his construction of Quillen line bundles. We
213: reprove the relation with the determinant of cohomology construction.
214:
215: Furthermore, if we specialize the covering to the one provided by a
216: Kleinian uniformization (thereby allowing possibly disconnected
217: surfaces) the same class can be reinterpreted as the transgression of
218: the regulator class expressed by the Bloch-Wigner dilogarithm.
219: \end{abstract}
220:
221: \tableofcontents
222:
223: \section{Introduction}
224: \label{sec:introduction}
225:
226: Metrics of constant negative curvature play a very important role in
227: uniformization problems for compact Riemann surfaces of genus $g>1$. The
228: condition that the scalar curvature associated to a conformal metric on
229: a Riemann surface $X$ be equal to $-1$ is equivalent to the fact that
230: the associated conformal factor satisfies a nonlinear partial
231: differential equation known as the Liouville equation.
232:
233: The Liouville equation appears as early as in one of the approaches
234: considered by Poincar\'e to attack the uniformization theorem
235: \cite{poincare:liou}. In relatively recent times, it has received
236: considerable attention in Theoretical and Mathematical Physics due to
237: the key role it plays in Polyakov's approach to String Theory
238: \cite{polyakov:bosonic}, especially from the point of view of
239: non-critical strings and two-dimensional quantum gravity. In this
240: context one refers to the conformal factor of the metric as the
241: Liouville ``field.''
242:
243: As usual in the context of differential equations with a physical
244: motivation, one would normally like to formulate a variational principle
245: to express the Liouville equation as an extremum condition. Namely,
246: given a Riemann surface $X$ and the space $\cm{X}$ of all conformal
247: metrics on it, the metric of constant negative curvature should be a
248: critical point of a functional defined over $\cm{X}$. This functional is
249: the Liouville action. As it happens, action functionals may turn out to
250: be even more relevant than the equations they are associated to. The
251: Liouville action is no exception in this sense: it has deep connection
252: with the geometry of Teichm\"uller spaces
253: \cite{zogtak1987-1,zogtak1987-2}, and in Physics it describes the
254: conformal anomaly in String Theory.
255:
256: Providing a rigorous mathematical definition of the Liouville action
257: functional is however far from trivial. The very geometric properties of
258: the Liouville field itself prevent expressing the corresponding
259: functional as a plain integral of a $2$-form on a Riemann surface.
260: Correction terms are required, typically in the form of integration of
261: lower degree forms over the $1$-skeleton of an appropriate simplicial
262: realization of $X$. (One should notice that this behavior is not
263: specific to the Liouville equation, and it is by now possible to give a
264: characterization, in terms of homological algebra, of these type of
265: functionals, see ref.\ \cite{MR1908413}.)
266:
267: It is possible to directly determine the necessary correction terms by
268: requiring that the variational problem be well defined. This, however,
269: is not completely satisfactory from the point of view of certain
270: applications to deformation theory, where a consistent definition across
271: a \emph{family} of surfaces is required. Quite recently, a more
272: systematic construction, based on the homological algebra techniques
273: developed by the author and L.\ A.\ Takhtajan in \cite{aldtak1997}, was
274: carried out by L.\ A.\ Takhtajan and L.-P.\ Teo in ref.\
275: \cite{math.CV/0204318}, generalizing the earlier results of
276: \cite{zogtak1987-1,zogtak1987-2}. The authors of ref.\
277: \cite{math.CV/0204318} constructed a \cech\ cocycle with respect to the
278: \'etale cover of $X$ associated to a quasi-Fuchsian (and more generally
279: Kleinian) uniformization. Since their construction works across
280: (Kleinian) deformations, it could be exploited to obtain results of
281: global nature on the analytic geometry of Kleinian deformation spaces.
282: As a further result, the authors of loc.\ cit.\ were able to rigorously
283: prove the validity of the ``holography principle'' for the Liouville
284: action corresponding to a large class of Kleinian (in particular
285: Fuchsian and quasi-Fuchsian) uniformizations. Specifically, they proved
286: that given a second kind Kleinian group, the corresponding Liouville
287: action can be obtained as the regularized limit of the hyperbolic volume
288: of the corresponding associated $3$-manifold. This extends to the
289: general Kleinian case a previous formula obtained by Krasnov
290: \cite{MR2002k:81230} for classical Schottky groups.
291:
292: Our interest in this matter is two-fold. From the perspective of the
293: newer methods adopted in \cite{aldtak2000}, the covering map $U\to X$
294: associated to the uniformization by a discrete group $\Gamma=\deck
295: (U/X)$ is but one of the many possible covers comprising an appropriate
296: category $\cat{C}$ of, say, local diffeomorphisms $U\to X$---the most
297: obvious choice being that of standard open cover $\cover{U}=\{ U_i\}$
298: with associated space $U=\coprod_i U_i$. In particular one expects to be
299: able to apply the methods of \cite{aldtak1997} and \cite{aldtak2000}
300: uniformly on a class of reasonably behaved covers of $X$.
301:
302: Second, the focus of ref.\ \cite{aldtak2000} was on the rigorous
303: definition of a functional for quasi-conformal deformation of the
304: Riemann surface $X$ and its application to the study of projective
305: structures. A main result is that the construction of the action is
306: possible thanks to the vanishing of the ``tame symbol'' (see refs.\
307: \cite{del:symbole} and \cite{brymcl:deg4_II} for the relevant
308: definitions) $\tame{T_X}{T_X}$, where $T_X$ is the holomorphic tangent
309: line bundle of $X$. The vanishing determines local choices (with respect
310: to a cover) of a Bloch-type dilogarithm which then allow for a
311: cohomological construction of the action. There are many indication that
312: the Liouville action ought to be the hermitian square of a functional of
313: the type studied in \cite{aldtak2000}.\footnote{From a physical point of
314: view this originates in the modular geometry approach to Conformal
315: Field Theory advocated by Friedan and Shenker in \cite{MR88b:81146}.
316: Mathematically speaking, it is one of the many proposed forms of the
317: holomorphic factorization property for determinant line bundles.} Thus
318: it is natural to ask whether there is an analogous mechanism as the one
319: in loc.\ cit.\ to obtain a general construction of the Liouville action
320: by replacing the holomorphic symbol maps and dilogarithms with
321: corresponding real objects.
322:
323: In this paper we answer this question in the affirmative. More
324: precisely, we show that the Liouville action (up to the area term
325: which is given by an ordinary $2$-form) can be computed as a
326: symbol map taking values in hermitian holomorphic Deligne
327: cohomology, first introduced by Brylinski and McLaughlin in their
328: study of degree four characteristic classes
329: \cite{brymcl:deg4_II}. (By way of comparison, the tame symbols
330: used in ref.\ \cite{aldtak2000} used holomorphic and smooth
331: Deligne cohomology.) In particular we show that the dilogarithm
332: type terms are replaced here by the Bloch-Wigner function, the
333: real valued counterpart of the dilogarithm (see refs.
334: \cite{MR2001i:11082}, and \cite{MR94k:19002,MR2002g:52013} for a
335: review.)
336:
337: The appearance of the Bloch-Wigner function ties very well with
338: the holography property of the Liouville function proved in
339: \cite{math.CV/0204318} in the following sense. As mentioned
340: before, the Liouville action (up to the area term) relative to a
341: Kleinian uniformization\footnote{Note that $X$ is allowed to be
342: disconnected.} $U\to X$ can also be computed as the
343: ``regularized volume'' of the associated $3$-manifold $N=\Gamma
344: \backslash (U \cup \HH^3)$, where $\Gamma =\deck (U/X)$ as
345: before, $U\subset \PP^1$ is the domain of discontinuity for
346: $\Gamma$, and $\HH^3$ is the standard hyperbolic $3$-space. (To
347: define the regularized volume would lead us too far afield. It
348: suffices to mention that the conformal factor of a metric on
349: $X=\partial N$ can be used to select a compact submanifold
350: $N_\epsilon$ whose volume is finite. One then subtracts from the
351: volume of $N_\epsilon$ the areas of the boundary components and
352: other carefully chosen constants independent of the metric
353: structure, so that the resulting quantity will have a finite
354: limit as $\epsilon \to 0$.) On the other hand, the hyperbolic
355: volume in three dimensions corresponds to a three dimensional
356: (purely imaginary) class on $\PSL_2(\CC)$ expressible through the
357: Bloch-Wigner dilogarithm, the so-called \emph{regulator class.}
358: We show that the regulator is precisely the class that needs to
359: be killed in order to close the cohomological descent conditions
360: required to calculate the Liouville action for a covering map
361: $U\to X$ with covering group a Kleinian group $\Gamma$. This is
362: possible, since for a second kind Kleinian group the quotient
363: $\HH^3/\Gamma$ is non-compact, hence it carries no cohomology in
364: dimension three, so the class represented by the Bloch-Wigner
365: function, pulled back to $\Gamma$ via the imbedding $\Gamma
366: \hookrightarrow \PSL_2(\CC)$, vanishes.
367:
368: Returning to the cohomological interpretation of the construction
369: of the Liouville action, it should also be noted that leaving
370: aside the area term, our results show that the cohomologically
371: non trivial part is indeed a square. Namely, for a conformal
372: metric $\rho \in \cm{X}$ we consider the pair $(T_X,\rho)$ as a
373: holomorphic line bundle equipped with an hermitian metric. Then,
374: using that hermitian holomorphic Deligne cohomology has a cup
375: product, we show that the Liouville action is just the square of
376: the class of $(T_X,\rho)$. In fact this identification holds at
377: the level of cocycles, rather than only for the corresponding
378: classes.
379:
380: Again leaving aside the area term, it immediately follows from
381: the properties of hermitian holomorphic Deligne cohomology that
382: most of the story carries over to the case of a \emph{pair} of
383: holomorphic line bundles $L$ and $L'$ equipped with hermitian
384: metrics $\rho$ and $\rho'$, respectively. Furthermore, Brylinski
385: shows in \cite{bry:quillen} that the pairing of two such
386: holomorphic line bundles with metrics corresponds to the pairing
387: defined by Deligne on the determinant of cohomology in
388: \cite{MR89b:32038}. Without introducing the machinery of
389: $2$-gerbes, we reobtain this result in our setting. Specifically,
390: we directly obtain Gabber's formula for the hermitian metric on
391: the determinant line from the explicit cocycle for the cup
392: product of two metrized line bundles. In turn this shows that the
393: Liouville action is a multiple of the determinant of cohomology,
394: thereby generalizing earlier results (cf.\
395: \cite{zograf1990})---without assuming criticality.
396:
397:
398: \subsection{Organization of the paper}
399: \label{sec:Organization-paper}
400:
401: This paper is organized as follows.
402: Sections~\ref{sec:deligne-complexes}
403: and~\ref{sec:herm-holom-deligne} are devoted to expounding some
404: background material for the sake of keeping this paper
405: self-contained and to put the reader in position of reproducing
406: the necessary calculations. Section~\ref{sec:deligne-complexes}
407: contains background facts on Deligne cohomology, paying special
408: attention to the product structures and the cone constructions.
409: We provide some examples and collect some facts about the
410: dilogarithm from the point of view of Deligne cohomology. The
411: particular model of hermitian holomorphic Deligne cohomology we
412: use later in the paper requires certain constructions available
413: in the literature, and recalled in
414: section~\ref{sec:deligne-complexes}, to be slightly modified in
415: order to obtain a (graded) commutative product. The necessary
416: arguments, being somewhat outside the line of development of the
417: paper are presented in Appendix~\ref{sec:Cones}. Hermitian
418: holomorphic Deligne cohomology is introduced in
419: section~\ref{sec:herm-holom-deligne}. We give the definition as
420: in refs.\ \cite{brymcl:deg4_II} and \cite{bry:quillen}, and then
421: introduce another model which, albeit more complex, has the
422: advantage for us of keeping the metric structure explicit. We
423: have explicitly proved the isomorphism in
424: Lemma~\ref{lemma:qi-dhh-bry}. For the sake of completeness, we
425: give an explicit description of the cocycle determined by a
426: holomorphic line bundle with hermitian metric, and in
427: sect.~\ref{sec:Cup-prod-herm} we explicitly compute the cup
428: product of two metrized line bundles for later usage. Results
429: about the existence of a fiber integration map are mentioned in
430: the paper, so some background material is provided in
431: sect.\ \ref{sec:integr-along-fiber}.
432:
433: Sections~\ref{sec:Conf-metr-Liouv} and~\ref{sec:conf-metr-herm}
434: form the core of the paper. The direct construction of the
435: Liouville action according to the techniques of refs.\
436: \cite{aldtak1997,aldtak2000,math.CV/0204318} is presented in
437: section~\ref{sec:Conf-metr-Liouv}. Since explicit calculations
438: have been presented in great detail in ref.\
439: \cite{math.CV/0204318}, and the calculations we need are quite
440: straightforward, we keep details to a minimum. In
441: Proposition~\ref{prop:2} and Corollary~\ref{cor:1} we show that
442: the Liouville action functional computed via descent theory does
443: solve the variational problem. These results have appeared also
444: in ref.\ \cite{math.CV/0204318} and are presented for here
445: completeness, although our choice for the cover is different. The
446: framework of loc.\ cit.\ is that of a Kleinian cover $U\to X\cong
447: U/\Gamma$, where $\Gamma$ is a second kind purely loxodromic
448: geometrically finite Kleinian group, which then we treat in some
449: detail in sect.\ \ref{sec:liouv-funct-bloch}. We illustrate how
450: the genuine Bloch-Wigner function of sect.\
451: \ref{sec:remarks-dilogarithm} appears in the descent equations
452: relative to the Kleinian cover. Having observed descent equations
453: close on general cohomological grounds in sect.\
454: \ref{sec:solut-vari-probl}, we now point out that for the case of
455: a Kleinian cover this is due to the vanishing of the regulator
456: class for the non-compact $3$-manifold $\HH^3/\Gamma$.
457:
458: A true geometric construction of the Liouville action, which does
459: not rely on the arguments of sect.\ \ref{sec:Conf-metr-Liouv} to
460: close the descent equations, is carried out in
461: section~\ref{sec:conf-metr-herm}, which contains our main result:
462: We compare the descent calculations with the cup products
463: computed in section~\ref{sec:herm-holom-deligne} and conclude
464: that the quadratic part of the Liouville action is in fact
465: (modulo an area term) the cup square of the metrized holomorphic
466: tangent bundle in Hermitian holomorphic Deligne cohomology, see
467: Theorem~\ref{thm:2}, Corollary~\ref{corollary:1}, and
468: Proposition~\ref{prop:3}. It follows that from this point of view
469: the descent equations close automatically, without the need for
470: special arguments. We then prove that the cup square is
471: identified with the determinant of cohomology construction in
472: Theorem~\ref{thm:3} and Corollary~\ref{corollary:2}. Auxiliary
473: facts about the homological algebra of cones and conventions on
474: Kleinian groups are stored in the appendices.
475:
476: Finally, we draw our conclusions in sect.\
477: \ref{sec:conclusions-outlook}.
478:
479: \subsection{Notation and conventions}
480: \label{sec:notation}
481:
482: If $z$ is a complex number, then $\pi\sb{p} (z) \eqdef \onehalf ( z +
483: (-1)\sp p \Bar z)$, and similarly for any other complex quantity, e.g.
484: complex valued differential forms. If $A$ is a subring of $\RR$, we will
485: use the ``twist'' $A(j) = (2\pi \sqrt{-1})^j\,A$.
486:
487: If $X$ is a complex manifold, $\sha{X}$ and $\shomega{X}$ denote
488: the de~Rham complexes of smooth $\CC$-valued and holomorphic
489: forms, respectively. We denote by $\she{X}$ the de~Rham complex
490: of sheaves of real valued differential forms and by $\she{X} (j)$
491: the twist $\she{X} \otimes_\RR \RR(j)$. We set $\sho{X} \equiv
492: \shomega[0]{X}$ as usual. When needed, $\sha[{p,q}]{X}$ will
493: denote the sheaf of smooth $(p,q)$-forms. We use the standard
494: decomposition $d=\del + \delb$ according to types. Furthermore,
495: we introduce the differential operator $d^c = \del -\delb$
496: (contrary to the convention, see, e.g. \cite{lang:arakelov}, we
497: omit the factor $1/(4\pi \sqrt{-1})$). We have $2\del\delb =
498: d^cd$. The operator $d^c$ is an imaginary one, and accordingly,
499: we have the rules
500: \begin{gather*}
501: d\pi_p(\omega) = \pi_p(d\omega)\\
502: d^c\pi_p(\omega) = \pi_{p+1}(d^c\omega)
503: \end{gather*}
504: for any complex form $\omega$.
505:
506: An open cover of $X$ will be denoted by $\cover{U}_X$. If $\{U_i\}_{i\in
507: I}$ is the corresponding collection of open sets, we denote $U_{ij} =
508: U_i\cap U_j$, $U_{ijk} = U_i\cap U_j\cap U_k$, and so on. We can also
509: consider more general covers $\cover{U}_X = \{ U_i \to X\}_{i\in I}$
510: where the maps are regular coverings in the appropriate category. In
511: this case intersections are replaced by $(n+1)$-fold fibered products
512: \begin{math}
513: U_{i_0 i_1\dots i_n}= U_{i_0} \times_X\dots \times_X U_{i_n}\,.
514: \end{math}
515: Open coverings fit this more general description, since if $U_i$ and
516: $U_j$ are two open sets, then $U_i\cap U_j = U_i\times_X U_j$, where the
517: fiber product is taken with respect to the inclusion maps. As another
518: example, one can consider regular covering maps $U\to X$ with $\Gamma =
519: \mathrm{Deck}(U/X)$, and in this case, taking the fiber product over $X$
520: $(n+1)$-times, one gets
521: \begin{math}
522: U\times_X\dots\times_X U = U\times \Gamma \times \dots \times \Gamma\,,
523: \end{math}
524: where the group factor is repeated $n$-times. This includes the cases of
525: Kleinian (and in particular Fuchsian) covers of Riemann surfaces.
526:
527: The \emph{nerve} of the cover $\cover{U}_X$ is
528: the simplicial object
529: \begin{math}
530: n\mapsto N_n(\cover{U}_X) = \coprod U_{i_0} \times_X\dots \times_X
531: U_{i_n}
532: \end{math}
533: where $N_n(\cover{U}_X)$ maps into $N_{n-1}(\cover{U}_X)$ in $(n+1)$
534: ways by forgetting in turn each factor. For open covers this just yields
535: the expected inclusion maps.
536:
537: If $\sheaf{F}^\bullet$ is a complex of sheaves on $X$, its \cech\
538: resolution with respect to a covering $\cover{U}_X\to X$ is the double
539: complex
540: \begin{math}
541: \CCC^{p,q} (\sheaf{F}) \eqdef \check{C}^q(\cover{U}_X,\sheaf{F}^p)\,,
542: \end{math}
543: where the $q$-cochains with values in $\sheaf{F}^p$ are given by
544: \begin{math}
545: \prod \sheaf{F}^p (U_{i_0} \times_X\dots \times_X U_{i_n})\,.
546: \end{math}
547: The \cech\ coboundary operator is denoted $\deltacheck$. The sign
548: convention we are going to use is that the index along the \cech\
549: resolution is the \emph{second} one, so if we denote by $d$ the
550: differential in the complex $\sheaf{F}^\bullet$, the total differential
551: in the total simple complex of $\check{C}^q(\cover{U}_X,\sheaf{F}^p)$
552: will be $D=d\pm \deltacheck$. For open covers we just get the familiar
553: \cech\ (hyper)cohomology. The other interesting example is that of a
554: regular covering map $U\to X$: \cech\ cohomology with respect to this
555: cover is the same as group cohomology for $\Gamma = \mathrm{Deck}(U/X)$
556: with coefficients in the $\Gamma$-module $\sheaf{F}^p(U)$.
557:
558: The Koszul sign rule that results in a sign being picked whenever two
559: degree indices are formally exchanged is applied. In particular, for
560: \cech\ resolutions of complexes of sheaves, it leads to the following
561: conventions. If $\sheaf{G}^\bullet$ is a second complex of sheaves on
562: $X$, then one defines the cup product
563: \begin{equation*}
564: \cup : \CCC^{p,q}(\sheaf{F}) \otimes \CCC^{r,s}(\sheaf{G})
565: \longrightarrow \Check{C}^{q+s}(\cover{U}_X,\sheaf{F}^p\otimes
566: \sheaf{G}^r) \subset \CCC^{p+r,q+s}(\sheaf{F}\otimes\sheaf{G})
567: \end{equation*}
568: of two elements $\{f_{i_0,\dots,i_q}\}\in \CCC^{p,q}(\sheaf{F})$
569: and $\{g_{j_0,\dots,j_s}\} \in \CCC^{r,s}(\sheaf{G})$ by
570: \begin{equation*}
571: (-1)^{qr}\,f_{i_0,\dots,i_q}\otimes
572: g_{i_q,i_{q+1},\dots,i_{q+s}} \,.
573: \end{equation*}
574:
575: \section*{Acknowledgments}
576: Parts of this work were completed during visits at the
577: International School for Advanced Studies (SISSA) in Trieste,
578: Italy, and at the Department of Mathematics, Instituto Superior
579: T\'ecnico in Lisbon, Portugal. I would like to thank both
580: institutions for support and for creating a friendly and
581: stimulating research environment. I would also like to thank
582: Paolo Aluffi, Phil Bowers, Ugo Bruzzo, Johan Dupont, Leon
583: Takhtajan for illuminating discussions and/or patiently answering
584: my many questions. Also, special thanks are due to the referee
585: for his or her thoroughness and for providing very detailed
586: comments.
587:
588: \section{Deligne complexes}
589: \label{sec:deligne-complexes}
590:
591: \subsection{Cup products on cones}
592: \label{sec:cup-products-cones}
593:
594: Recall that the cone of a map $f: X^\bullet \to Y^\bullet$
595: between two complexes is the complex $C^\bullet (f) =
596: X^\bullet[1]\oplus Y^\bullet$ with differential $d_f(x,y) =
597: (-d\,x, f(x) + d\,y)$, where $[k]$ denotes the shift functor. The
598: cone fits into the exact sequence
599: \begin{equation*}
600: 0 \longrightarrow
601: Y^\bullet \longrightarrow C^\bullet (f) \longrightarrow
602: X^\bullet[1] \longrightarrow 0\,.
603: \end{equation*}
604: The following constructions are a special case of those considered by
605: \bei\ in ref.\ \cite{bei:hodge_coho}. Suppose we are given complexes
606: $X\sp\bullet\sb i$, $Y\sp\bullet\sb i$, and $Z\sp\bullet\sb i$ and maps
607: $f\sb i : X\sp\bullet\sb i \rightarrow Z\sp\bullet \sb i$, $g\sb i :
608: Y\sp\bullet\sb i \rightarrow Z\sp\bullet \sb i$, for $i=1,2,3$. Suppose
609: also that we have product maps $X^\bullet_1\otimes X^\bullet_2
610: \xrightarrow{\cup} X^\bullet_3$, and similarly for $Y^\bullet_i$, and
611: $Z^\bullet_i$, strictly compatible with the $f_i$, $g_i$ in the obvious
612: sense. Then we can consider the cones
613: \begin{equation*}
614: \cone \big( f_i -g_i \big)[-1] \equiv
615: \cone \big( X^\bullet_i\oplus Y^\bullet_i \xrightarrow{f_i -
616: g_i} Z^\bullet_i \big)[-1]\,.
617: \end{equation*}
618: For a real parameter $\alpha$, there is a family of products
619: \begin{equation}
620: \label{eq:1}
621: \cone \big( f_1 - g_1 \big)[-1] \otimes
622: \cone \big( f_2 - g_2 \big)[-1] \xrightarrow{\cup_\alpha}
623: \cone \big( f_3 - g_3 \big)[-1]
624: \end{equation}
625: determined as follows. For
626: \begin{math}
627: (x_i,y_i,z_i) \in X^\bullet_i \oplus Y^\bullet_i \oplus
628: Z^{\bullet -1}_i\,,\;i=1,2\,,
629: \end{math}
630: one defines
631: \begin{equation}
632: \label{eq:2}
633: \begin{split}
634: (x_1,y_1,z_1) \cup_\alpha (x_2,y_2,z_2) =
635: \Big(&x_1\cup x_2, y_1\cup y_2, \\
636: &(-1)^{\deg (x_1)}
637: \big((1-\alpha )f_1(x_1) + \alpha g_1(y_1) \big) \cup z_2 \\
638: &\quad +z_1\cup \big( \alpha f_2(x_2) +
639: (1-\alpha)g_2(y_2)\big) \Big)\,.
640: \end{split}
641: \end{equation}
642: Note that $\deg (x_1)=\deg (x_1,y_1,z_1)$. Checking that
643: $\cup_\alpha$ is a map of complexes is a straightforward routine
644: calculation. Different products for different values $\alpha$ and
645: $\beta$ of the real parameter are homotopic. Explicitly, we have
646: \begin{equation*}
647: (x_1,y_1,z_1) \cup_\alpha (x_2,y_2,z_2) -
648: (x_1,y_1,z_1) \cup_\beta (x_2,y_2,z_2) =
649: \big( d \, h_{\alpha,\beta} + h_{\alpha,\beta} d \big)
650: ((x_1,y_1,z_1) \otimes (x_2,y_2,z_2)) \,,
651: \end{equation*}
652: where the homotopy
653: \begin{equation*}
654: h_{\alpha,\beta} : \Tot \Big( \cone \big( f_1 - g_1\big)[-1]
655: \otimes \cone \big( f_2 - g_2\big)[-1] \Big)^\bullet
656: \longrightarrow \cone \big( f_3 - g_3\big)[-1]^{\bullet -1}
657: \end{equation*}
658: is given by the formula
659: \begin{equation}
660: \label{eq:3}
661: h_{\alpha,\beta} ((x_1,y_1,z_1) \otimes (x_2,y_2,z_2)) =
662: (\alpha -\beta) \, (-1)^{\deg (x_1) -1} (0,0,z_1\cup
663: z_2) \,.
664: \end{equation}
665: If the products
666: \begin{math}
667: X^\bullet_1 \otimes X^\bullet_2
668: \xrightarrow{\cup} X^\bullet_3\,,
669: \end{math}
670: etc., are graded commutative, then the swap functor on the tensor
671: product maps the $\cup_\alpha$ product structure on the cones into the
672: $\cup_{1-\alpha}$ structure. Using the homotopy~\eqref{eq:3} it follows
673: at once that there is a well defined graded commutative product in
674: cohomology.
675:
676: If we do not assume the product structures $X^\bullet_1\otimes
677: X^\bullet_2 \xrightarrow{\cup} X^\bullet_3$, etc., are strictly
678: compatible with the maps $f_i$, $g_i$, some of the preceding
679: constructions must be slightly modified. With an eye toward
680: certain constructions to be carried out later in this paper, let
681: us assume we have compatibility up to homotopy, namely there
682: exist maps
683: \begin{gather*}
684: h \colon \bigl( X_1\otimes X_2 \bigr)^\bullet
685: \longrightarrow Z_3^{\bullet -1}\\
686: k \colon \bigl( Y_1\otimes Y_2 \bigr)^\bullet
687: \longrightarrow Z_3^{\bullet -1}
688: \end{gather*}
689: such that
690: \begin{equation}
691: \label{eq:4}
692: \begin{gathered}
693: f_3\circ \cup - \cup \circ (f_1\otimes f_2)
694: = d\, h + h\,d \\
695: g_3\circ \cup - \cup \circ (g_1\otimes g_2)
696: = d\, k + k\,d\,,
697: \end{gathered}
698: \end{equation}
699: with obvious meaning of the symbols.
700: \begin{lemma}
701: \label{lemma:mod-cup-prod}
702: Let $X_i$, $Y_i$, $Z_i$ and the maps $f_i$, $g_i$ be as above. Let
703: $\alpha$ be a real parameter. We have a product of type~\eqref{eq:1}
704: for the cones $\cone(f_i-g_i)[-1]$ defined by the following
705: modification of formula~\eqref{eq:2}:
706: \begin{equation}
707: \label{eq:5}
708: \begin{split}
709: (x_1,y_1,z_1) \cup_\alpha (x_2,y_2,z_2) =
710: \Big(&x_1\cup x_2, y_1\cup y_2, \\
711: &(-1)^{\deg (x_1)}
712: \big((1-\alpha )f_1(x_1) + \alpha g_1(y_1) \big) \cup z_2 \\
713: &\quad +z_1\cup \big( \alpha f_2(x_2) +
714: (1-\alpha)g_2(y_2)\big)\\
715: &\qquad -h(x_1\otimes x_2) +k(y_1\otimes y_2)
716: \Big)\,.
717: \end{split}
718: \end{equation}
719: The product~\eqref{eq:5} is a map of complexes and two products
720: $\cup_\alpha$ and $\cup_\beta$ are related by the same homotopy
721: formula~\eqref{eq:3}.
722: \end{lemma}
723: \begin{proof}
724: Direct verification.
725: \end{proof}
726: This modified framework carries over to the full structure considered by
727: \bei\ in ref.\ \cite{bei:hodge_coho}. We will still refer to this
728: modified product as the \bei\ product. It is also necessary to relax the
729: assumption that the products
730: \begin{math}
731: X^\bullet_1 \otimes X^\bullet_2
732: \xrightarrow{\cup} X^\bullet_3\,,
733: \end{math}
734: etc., be graded commutative. It is possible to complete all the diagrams
735: so that the permutation of factors in the tensor products still yields a
736: homotopy commutative product~\eqref{eq:5} for the cones. As a
737: consequence the induced product in cohomology will still be graded
738: commutative. Explicit formulas are not needed except to ensure this
739: latter fact, therefore we shall not discuss this matter any further and
740: refer the reader to the appendix, where a brief but explicit treatment
741: can be found.
742:
743: \subsection{Deligne complexes}
744: \label{sec:deligne-compl}
745:
746: Let $X$ be a complex manifold. Recall the standard Hodge
747: filtration of $\shomega{X}$:
748: \begin{equation}
749: \label{eq:6}
750: F^p\shomega{X} : 0 \longrightarrow \cdots \longrightarrow
751: \shomega[p]{X} \longrightarrow \cdots \longrightarrow
752: \shomega[n]{X}\,,
753: \end{equation}
754: where $n=\dim_\CC X$.
755:
756: The corresponding filtration for the complex of smooth
757: $\CC$-valued forms is defined as follows: denote by $F^p\sha{X}$
758: the subcomplex of $\sha{X}$ comprising forms of type $(r,s)$
759: where $r$ is at least $p$, so that $F^p\sha[k]{X} = \oplus_{r\geq
760: p} \sha[{r,k-r}]{X}$. Then (cf. \cite{del:hodge_II}) the
761: inclusion $\shomega{X} \hookrightarrow \sha{X}$ is a
762: quasi-isomorphism respecting the filtrations, namely
763: $F^p\shomega{X} \hookrightarrow F^p\sha{X}$, and the latter
764: inclusion induces an isomorphism in cohomology.
765:
766: If $A$ is a subring of $\RR$, and $\imath$ and $\jmath$ denote
767: the inclusions of $A(p)$ and $F^p\shomega{X}$ into $\shomega{X}$
768: respectively, the $p$-th Deligne complex of sheaves is
769: defined by
770: \begin{equation}
771: \label{eq:7}
772: \deligne{A}{p} = \cone \bigl( A(p)_X \oplus F^p\shomega{X}
773: \xrightarrow{\imath -\jmath} \shomega{X} \bigr)[-1]\,.
774: \end{equation}
775: It is quasi-isomorphic to the complex:
776: \begin{equation}
777: \label{eq:8}
778: \cone \bigl( A(p)_X \oplus F^p\sha{X}
779: \xrightarrow{\imath -\jmath} \sha{X} \bigr)[-1]\,,
780: \end{equation}
781: where $\imath$ and $\jmath$ have the same meaning. We also notice
782: the quasi-isomorphism
783: \begin{equation}
784: \label{eq:9}
785: \deligne{A}{p}\lqi
786: \bigl( A(p)_X
787: \overset{\imath}{\longrightarrow} \sho{X}
788: \overset{d}{\longrightarrow} \shomega[1]{X}
789: \overset{d}{\longrightarrow} \dotsm
790: \overset{d}{\longrightarrow} \shomega[{p-1}]{X}\bigr)\,.
791: \end{equation}
792: When $A=\RR$ there are further quasi-isomorphisms, namely
793: \begin{equation*}
794: \deligne{\RR}{p} \xrightarrow{\simeq} \cone \big(
795: F^p\shomega{X} \rightarrow \she{X} (p-1)\big) [-1]
796: \xrightarrow{\simeq} \cone \big( F^p\sha{X} \rightarrow \she{X}
797: (p-1)\big) [-1]
798: \end{equation*}
799: since the maps
800: \begin{equation*}
801: \bigl( \RR(p) \rightarrow \shomega{X}\bigr)
802: \xrightarrow{\simeq} \bigl( \RR(p)
803: \rightarrow \CC \bigr) \xrightarrow{\simeq} \RR(p-1)
804: \xrightarrow{\simeq} \she{X}(p-1)
805: \end{equation*}
806: are all quasi-isomorphisms in the derived category, cf.
807: \cite{esn-vie:del}. Here we have used $\CC \cong \RR(p) \oplus
808: \RR(p-1)$. Following op.~cit., we set:
809: \begin{equation}
810: \label{eq:11}
811: \deltilde{\RR}{p} \eqdef
812: \cone \big( F^p\sha{X} \xrightarrow{-\pi\sb{p-1}} \she{X} (p-1)\big)
813: [-1] \, .
814: \end{equation}
815: Again, there is an explicit quasi-isomorphism
816: (\cite{MR86h:11103,esn-vie:del}):
817: \begin{equation}
818: \label{eq:12}
819: \begin{gathered}
820: \rho_p : \deligne{\RR}{p} \overset{\simeq}{\longrightarrow}
821: \deltilde{\RR}{p} \\
822: \rho_p \lvert_{\RR(p)} = 0\,,\quad \rho_p
823: \lvert_{F^p\shomega{X}} = \mathit{incl} \,,\quad \rho_p
824: \lvert_{\shomega{X}} = \pi_{p-1}
825: \end{gathered}
826: \end{equation}
827:
828: The \emph{Deligne cohomology groups} of $X$ with coefficients in
829: $A(p)$ are the hypercohomology groups
830: \begin{equation*}
831: \delH{X}{A}{p} = \HHH^\bullet (X, \deligne{A}{p})\,.
832: \end{equation*}
833: and clearly, any complex quasi-isomorphic to $\deligne{A}{p}$
834: would do. In order to perform calculations with these cohomology
835: groups we shall normally resort to a \cech\ resolution with
836: respect to an open cover $\cover{U}_X$ of $X$ or an \'etale map
837: $\cover{U}_X \to X$, e.g. a regular cover with deck group
838: $\Gamma$.
839:
840: One of the important properties of Deligne cohomology is the existence
841: of a graded commutative cup product
842: \begin{equation}
843: \label{eq:13}
844: \delH[i]{X}{A}{p} \otimes \delH[j]{X}{A}{q}
845: \xrightarrow{\cup} \delH[i+j]{X}{A}{p+q}\,,
846: \end{equation}
847: which follows from the existence of the \bei{} cup product at the
848: level of Deligne complexes whose construction was recalled above.
849: There are products $A(p) \otimes A(q) \to A(p+q)$ and
850: $F^p\shomega{X} \otimes F^q\shomega{X} \to F^{p+q}\shomega{X}$,
851: plus the obvious (wedge) product on $\shomega{X}$, thus it
852: follows from the cone version~\eqref{eq:7} that the Deligne
853: complexes come equipped with the \bei{} product, and therefore
854: the Deligne cohomology groups inherit the graded commutative cup
855: product~\eqref{eq:13}. The explicit form, that is, the
856: translation of~\eqref{eq:2} to the case at hand can be found
857: in~\cite{esn-vie:del}. The explicit form of the cup product for
858: the complex~\eqref{eq:11} as computed in \cite{MR86h:11103} (see
859: also \cite{esn-vie:del}) will be needed in the sequel. Let
860: $(\omega\sb 1, \eta\sb 1)$ be an element of degree $n$ in
861: $\deltilde{\RR}{p}$---this means that $\omega\sb 1\in
862: F^p\shomega[n]{X}$ and $\eta\sb 1\in \she[n-1]{X}(p-1)$---and
863: $(\omega\sb 2, \eta\sb 2)$ any element in $\deltilde{\RR}{q}$.
864: The product is defined by the formula:
865: \begin{equation}
866: \label{eq:15}
867: (\omega_1,\eta_1) \,\Tilde\cup\, (\omega_2,\eta_2) = \bigl(
868: \omega\sb 1 \wedge \omega\sb 2 ,
869: (-1)^n\, \pi\sb p\omega\sb 1 \wedge \eta\sb 2
870: +\eta\sb 1\wedge \pi\sb q \omega\sb 2 \bigr)\,.
871: \end{equation}
872: The product $\Tilde\cup$ is a morphism of complexes and (modulo
873: the quasi-isomorphisms $\rho_p$) is homotopic to the \bei\
874: products $\cup_\alpha$ on the complexes $\deligne{\RR}{p}$.
875: Specifically, if we denote an element of $\deligne{\RR}{p}$ by
876: the triple $(r,f,\omega)$, where $r\in \RR(p)_X$, $f\in
877: F^{p}\!\shomega{X}$, and $\omega \in \shomega{X}$, the homotopy
878: between $\Tilde\cup \circ (\rho_p\otimes \rho_q)$ and
879: $\rho_{p+q}\circ \cup_\alpha$ is given by
880: \begin{equation}
881: \label{eq:16}
882: \Tilde h ((r,f,\omega)\otimes (r',f',\omega'))
883: = (-1)^{\deg \omega} \bigl(0,
884: (1-\alpha)\, \pi_p\omega\wedge\pi_{q-1}\omega'
885: -\alpha\, \pi_{p-1}\omega\wedge\pi_{q}\omega'
886: \bigr)\,.
887: \end{equation}
888:
889:
890: \subsection{Examples}
891: \label{sec:examples}
892:
893: \subsubsection{}
894: Let $A=\ZZ$. From~\eqref{eq:9} we have $\deligne{\ZZ}{1} \qi
895: \sho{X}^\times[-1]$ via the standard exponential sequence, so
896: that
897: \begin{math}
898: \delH[k]{X}{\ZZ}{1} \cong H^{k -1}(X, \sho{X}^\times)\,.
899: \end{math}
900: In particular $\delH[1]{X}{\ZZ}{1}\cong H^0(X,\sho{X}^\times)$,
901: the global invertibles on $X$, and $\delH[2]{X}{\ZZ}{1} \cong
902: \pic{X}$, the Picard group of line bundles over $X$.
903:
904: If an open cover $\{U_i\}_{i\in I}$ of $X$ is chosen, the class
905: of a line bundle $L$ in $\delH[2]{X}{\ZZ}{1}$ can be represented
906: via a \cech\ resolution by the cocycle $(f_{ij},c_{ijk})$, where
907: $f_{ij}\in \sho{X}(U_{ij})$, and $c_{ijk}\in \ZZ(1)_X(U_{ijk})$.
908: Thus the functions $f_{ij}$ should be interpreted as
909: \emph{logarithms} of the corresponding transition functions for
910: $L$. Then, the collection $c_{ijk}=(\deltacheck f)_{ijk}$
911: provides a representative for the first Chern Class $c_1(L)$.
912: Similarly, an invertible function $f$ would be described by the
913: collection $f_i$ of its logarithms on each open $U_i$, plus the
914: ``integers'' $m_{ij} = f_j -f_i\in \ZZ(1)$ on each $U_{ij}$.
915:
916: \subsubsection{}
917: Still using the exponential sequence, $\deligne{\ZZ}{2} \qi
918: \bigl( \sho{X}^\times \xrightarrow{d\log}
919: \shomega[1]{X}\bigr)[-1]$. Thus $\delH[2]{X}{\ZZ}{2}$ is the
920: group of isomorphism classes of holomorphic line bundles with
921: (holomorphic) connection. Using the (in fact, any) product
922: $\deligne{\ZZ}{1} \otimes \deligne{\ZZ}{1} \to \deligne{\ZZ}{2}$,
923: the cup product of two global invertible holomorphic functions
924: $f$ and $g$ on $X$ yields a line bundle with connection---the
925: tame symbol---denoted by $\tame{f}{g}$ whose class is in
926: $\delH[2]{X}{\ZZ}{2}$,
927: see~\cite{del:symbole,bloch:dilog_lie,rama:reg_hei}. Higher cup
928: products in this spirit have been studied
929: in~\cite{brymcl:deg4_II}.
930:
931: \subsubsection{}
932: If $A=\RR$, we have
933: \begin{math}
934: \delH[2p]{X}{\RR}{p} = H^{2p}(X,\RR(p)) \cap H^{p,p}(X)\,.
935: \end{math}
936: The \cech\ representative $(c_{ijk},f_{ij})$ of a class in
937: $\delH[2]{X}{\ZZ}{1}$ mentioned above maps to the cocycle $(-d{}
938: f_{ij}, -\abs{f_{ij}})$ under~\eqref{eq:12}. Taking into account
939: that the $f_{ij}$ are the logarithms of the transition functions,
940: the corresponding $(1,1)$ class would be given by the associated
941: canonical connection, see sec.~\ref{sec:line-bun}.
942:
943: \subsubsection{}\label{sec:ex_4}
944: $\delH[1]{X}{\RR}{1}$ is the group of real valued functions $f$ on $X$
945: such that there exists a holomorphic one-form $\omega$ such that $\pi\sb
946: 0\omega = df$. In other words it is the group of those smooth functions
947: $f$ such that $\del f$ is holomorphic, which amounts to say that such an
948: $f$ itself is harmonic.
949:
950:
951: \subsection{Remarks on the cup product $f\cup g$}
952: \label{sec:remarks-dilogarithm}
953:
954: It is convenient to consider the case of the cup product of two
955: invertible functions $f$ and $g$ in various complexes in more detail,
956: and to introduce some related notions we shall need later.
957:
958: As observed, $\deligne{\ZZ}{1} \qi \sho{X}^\times[-1]$ and an
959: invertible function $f$ can be considered as an element of
960: $\delH[1]{X}{\ZZ}{1}$. Therefore, via~\eqref{eq:12}, it induces
961: $\rho_1(f) \in \delH[1]{X}{\RR}{1}$ represented by $(d\log f,
962: \log\,\abs{f})$. (Note that $\log \abs{f}$ fits the description
963: of $\delH[1]{X}{\RR}{1}$ in \ref{sec:ex_4}.) Given two such $f$
964: and $g$, the expression for the cup product~\eqref{eq:15} gives
965: the following element of $\delH[2]{X}{\RR}{2}$:
966: \begin{equation}
967: \label{eq:17}
968: \rho_1(f)\,\Tilde\cup\, \rho_1(g) =
969: \bigl( d\log f \wedge d\log g,
970: -\pi_1(d\log f)\, \log\,\abs{g}
971: +\log\,\abs{f}\, \pi_1(d\log g) \bigr)\,.
972: \end{equation}
973: The first term is obviously zero when $X$ is a curve. Given $f$
974: and $g$, invertible on $X$, let us define the imaginary $1$-form:
975: \begin{equation}
976: \label{eq:18}
977: r_2(f,g) = \pi_1(d\log f)\, \log\,\abs{g}
978: -\log\,\abs{f}\, \pi_1(d\log g)\,.
979: \end{equation}
980: On the other hand, the cup product of $f$ and $g$ as elements of
981: $\delH[1]{X}{\ZZ}{1}$ yields an element $f\cup g$ of
982: $\delH[2]{X}{\ZZ}{2}$ represented by
983: \begin{math}
984: \bigl( d\log f \wedge d\log g, \log f\; d\log g \bigr)
985: \end{math}
986: (if we use the $\cup_0$ product) and this maps via $\rho_2$ to
987: the element
988: \begin{equation*}
989: \bigl( d\log f \wedge d\log g, -\pi_1(\log f\; d\log g)
990: \bigr)\,.
991: \end{equation*}
992: This is equal to~\eqref{eq:18} only up to homotopy. Indeed, using
993: \begin{math}
994: \pi_{p+q-1}(a\wedge b) = \pi_p(a)\wedge \pi_{q-1}(b) +
995: \pi_{p-1}(a)\wedge \pi_q(b)\,,
996: \end{math}
997: we find
998: \begin{equation}
999: \label{eq:19}
1000: r_2(f,g) = d \bigl( \pi_1(\log f)\, \log\,\abs{g} \bigr)
1001: -\pi_1 \bigl( \log f\; d\log g \bigr)\,,
1002: \end{equation}
1003: where the first term is just the explicit homotopy as computed
1004: from~\eqref{eq:16}.
1005:
1006: Recall that the \emph{tame symbol}
1007: (\cite{del:symbole,bloch:dilog_lie}) $\tame{f}{g}$ associated to
1008: $f$ and $g$ is the line bundle with connection determined (up to
1009: isomorphism) by the class $f\cup g$. A ``Bloch dilogarithm''
1010: (\cite{esn-vie:del}) is (the logarithm of) a horizontal
1011: trivializing section, namely a function $L$ on $U\subset X$
1012: satisfying the equation
1013: \begin{equation*}
1014: dL = -\log f\; d\log g\,.
1015: \end{equation*}
1016: Thus $L$ realizes the isomorphism $\tame{f}{g}\cong \sho{X}$ over
1017: $U\subset X$.
1018:
1019: Thus a Bloch dilogarithm will only locally be available, in
1020: general. However, when $g=1-f$ then $\tame{1-f}{f}$ is
1021: \emph{globally} trivial~\cite{bloch:dilog_lie,esn-vie:del}, i.e.
1022: $\tame{1-f}{f}\cong \sho{X}$. This is the the \emph{Steinberg
1023: relation} satisfied by the Tame symbol. It can be deduced from
1024: the following universal case. Set $f=z$, $X=\PP^1 \setminus
1025: \{0,1,\infty\}$, consequently $g=1-z$. Then $L$ is identified
1026: with the classical Euler dilogarithm $\li$, namely
1027: \begin{equation*}
1028: \li (z) = -\int_0^z \log (1-t) \frac{dt}{t}\,,
1029: \end{equation*}
1030: see~\cite{del:symbole} and~\cite{MR94k:19002} for details.
1031:
1032: On $\PP^1 \setminus \lbrace 0,1,\infty\rbrace$ the classical
1033: dilogarithm has a single valued parter, denoted $\bwd$,
1034: introduced by Bloch and Wigner:
1035: \begin{equation}
1036: \label{eq:20}
1037: \bwd (z) = \arg (1-z) \log\,\abs{z} + \im\li (z)\,.
1038: \end{equation}
1039: $\bwd$ is real-analytic on $\PP^1 \setminus \{0,1,\infty\}$ and
1040: extends continuously to $\PP^1$. That it is single-valued can be
1041: verified as follows. Choose a branch of the logarithm, say the
1042: principal one, to define $\li$ (and $\bwd$) on the cut plane $\CC
1043: \setminus (-\infty,0] \cup [1,\infty)$. Then one shows that the
1044: expression~\eqref{eq:20} is in fact single-valued by analytic
1045: continuation along paths based, say, at $1/2\in \CC$, and winding
1046: around the points $0$ and $1\in\CC$. Explicit computations can be
1047: found in ref.~\cite{MR2001i:11082}.
1048:
1049: It is convenient to introduce
1050: \begin{equation*}
1051: \bwli (z) = \sqrt{-1}\, \bwd (z) \,,
1052: \end{equation*}
1053: so that
1054: \begin{equation*}
1055: d\bwli = r_2(1-z,z)\,.
1056: \end{equation*}
1057: More generally, if $L$ trivializes $\tame{f}{g}$ over $U$ in the
1058: sense explained above, we can associate a function $\bwli (f,g)$
1059: over $U$ such that
1060: \begin{equation*}
1061: d \bwli(f,g) = r_2(f,g)
1062: \end{equation*}
1063: via the position
1064: \begin{equation}
1065: \label{eq:21}
1066: \bwli (f,g) = \pi_1(\log f)\, \log\,\abs{g} +\im L\,.
1067: \end{equation}
1068:
1069:
1070: \section{Constructions in hermitian holomorphic Deligne cohomology}
1071: \label{sec:herm-holom-deligne}
1072:
1073: In this section we recall the definition of hermitian holomorphic
1074: Deligne cohomology. In ref.~\cite{bry:quillen} Brylinski
1075: introduced certain complexes $C(l)^\bullet$, for a positive
1076: integer $l$, in order to compare the \bei{}-Chern classes of a
1077: holomorphic vector bundle $E$ with the Cheeger-Chern-Simons
1078: classes determined by $(E,\nabla)$, where $\nabla$ is the
1079: \emph{canonical connection,} namely the unique connection
1080: compatible with both the holomorphic and hermitian structures.
1081: The cohomology groups determined by these complexes are aptly
1082: named \emph{Hermitian Holomorphic Deligne cohomology groups.} For
1083: a holomorphic \emph{line} bundle equipped with the canonical
1084: connection, the complex $C(1)^\bullet$ encodes the reduction of
1085: the structure group from $\CC$ to $\TT$ afforded by the hermitian
1086: fiber metric.
1087:
1088: In the following we will need to compute Hermitian Holomorphic
1089: Deligne cohomology by means of different---but
1090: quasi-isomorphic---sheaf complexes we denote $\dhh{l}$. These
1091: complexes are tailored to a direct description of a metrized line
1092: bundle in terms of local representatives of the hermitian fiber
1093: metric. Since the two constructions are related by a
1094: quasi-isomorphism, the resulting cohomology groups are the same.
1095:
1096: \subsection{Hermitian holomorphic Deligne cohomology}
1097: \label{sec:herm-holom-deligne-1}
1098:
1099: In ref.~\cite{bry:quillen}, where Brylinski introduces the
1100: complexes:
1101: \begin{equation}
1102: \label{eq:24}
1103: C(l)^\bullet = \cone \bigl(
1104: \ZZ(l)_X \oplus (F^l\!\sha{X}\cap \sigma^{2l}\she{X}(l))
1105: \longrightarrow \she{X}(l)
1106: \bigr)[-1]\,,
1107: \end{equation}
1108: where $\sigma^p$ denotes the (sharp) truncation in degree $p$,
1109: namely for a complex $\sheaf{F}^\bullet$, $\sigma^p\sheaf{F}^k$
1110: is zero for $k<p$ and equal to $\sheaf{F}^k$ for $k\geq p$.
1111: In~\eqref{eq:24} we take the cone of the difference between the
1112: two inclusions. We have the following:
1113: \begin{definition}[\cite{bry:quillen}]
1114: \label{def:1}
1115: The hermitian holomorphic Deligne cohomology groups are the
1116: hypercohomology groups of the complexes~\eqref{eq:24}:
1117: \begin{equation}
1118: \label{eq:23}
1119: \dhhH[p]{X}{l} \eqdef \HHH^{\,p}(X,C(l)^\bullet)\,.
1120: \end{equation}
1121: \end{definition}
1122: The complexes~\eqref{eq:24} are expressed as cones, and therefore
1123: admit a (standard) \bei\ product~\eqref{eq:2}. (The wedge product
1124: induces cup products on both $\she{X}(l)$ and $F^l\!\sha{X}\cap
1125: \sigma^{2l}\she{X}(l)$.)\footnote{It appears signs should be
1126: adjusted in the product formula quoted in
1127: ref.~\cite{bry:quillen}, and that using~\eqref{eq:2} is more
1128: appropriate.} It follows there is a graded commutative product
1129: on cohomology:
1130: \begin{equation}
1131: \label{eq:27}
1132: \dhhH[i]{X}{l} \otimes \dhhH[j]{X}{k}
1133: \overset{\cup}{\longrightarrow}
1134: \dhhH[i+j]{X}{l+k}\,.
1135: \end{equation}
1136: Also, from the standard cone exact sequences we get
1137: (cf.\ ref.~\cite{bry:quillen}):
1138: \begin{equation}
1139: \label{eq:84}
1140: \dotsm \longrightarrow H^{2l-1}(X,\RR (l))
1141: \longrightarrow \dhhH[2l]{X}{l} \longrightarrow
1142: H^{2l}(X,\ZZ (l)) \oplus A^{(l,l)}(X)_{\RR(l)}
1143: \longrightarrow H^{2l}(X,\RR (l))
1144: \longrightarrow \dotsm
1145: \end{equation}
1146: where $A^{(l,l)}(X)_{\RR(l)}$ denotes the space of smooth
1147: $\RR(l)$-valued global $(l,l)$-forms on $X$. Thus we see
1148: hermitian holormorphic classes are $\ZZ(l)$-valued classes
1149: represented by (necessarily) closed $\RR(l)$-valued $2l$-forms of
1150: pure type $(l,l)$. For a line bundle this corresponds to a
1151: structure group reduction from $\sho{X}$ to $\she[0]{X}$, namely
1152: $\TT$-valued sections, at the same time controlling the Hodge
1153: type of the resulting class, cf.\ ref.~\cite{bry:quillen}
1154:
1155: Later (cf. sec.~\ref{sec:line-bun}) we will want to work with the
1156: hermitian structure on a holomorphic line bundle, together with
1157: the imaginary $(1,1)$-form built from the canonical connection,
1158: directly in the holomorphic frame. To carry this out in general
1159: for $(l,l)$ classes, we introduce the complex
1160: \begin{equation}
1161: \label{eq:22}
1162: \dhh{l} = \cone \bigl(
1163: \deligne{\ZZ}{l}\oplus (F^l\!\sha{X}\cap \sigma^{2l}\she{X}(l))
1164: \longrightarrow \deltilde{\RR}{l}
1165: \bigr)[-1]\,.
1166: \end{equation}
1167: The map
1168: \begin{math}
1169: \deligne{\ZZ}{l} \to \deltilde{\RR}{l}
1170: \end{math}
1171: is the composite of the obvious map
1172: \begin{math}
1173: \deligne{\ZZ}{l} \to \deligne{\RR}{l}
1174: \end{math}
1175: with the quasi-isomorphism $\rho_{\,l}$ defined by~\eqref{eq:12}. We
1176: will simply denote it by $\rho_l$ in the sequel, suppressing the first
1177: morphism in the notation. The map
1178: \begin{math}
1179: (F^l\!\sha{X}\cap \sigma^{2l}\she{X}(l)) \to \deltilde{\RR}{l}
1180: = \cone \bigl( F^l\!\sha{X} \to \she{X}(l-1) \bigr)[-1]
1181: \end{math}
1182: is induced by the inclusion of $(F^l\!\sha{X}\cap
1183: \sigma^{2l}\she{X}(l))$ into $F^l\!\sha{X}$. In~\eqref{eq:22} we
1184: take the cone of the difference between these two maps. The
1185: complex~\eqref{eq:22} offers another model for Hermitian
1186: holomorphic Deligne cohomology. Indeed we have:
1187: \begin{lemma}
1188: \label{lemma:qi-dhh-bry}
1189: The complexes $C(l)^\bullet$ and $\dhh{l}$ are
1190: quasi-isomorphic.
1191: \end{lemma}
1192: \begin{proof}
1193: By elementary manipulation of cones
1194: \begin{equation*}
1195: \dhh{l}= \cone \Bigl(
1196: F^l\!\sha{X}\cap \sigma^{2l}\she{X}(l) \to
1197: \cone \bigl( \deligne{\ZZ}{l} \to
1198: \deltilde{\RR}{l} \bigr) \Bigr)[-1]
1199: \end{equation*}
1200: and clearly:
1201: \begin{math}
1202: \cone \bigl( \deligne{\ZZ}{l} \to \deltilde{\RR}{l} \bigr)
1203: \qi \cone \bigl( \ZZ(l)_X \to \RR(l)_X \bigr) \qi \cone
1204: \bigl( \ZZ(l)_X \to \she{X}(l) \bigr)\,,
1205: \end{math}
1206: where all arrows are quasi-isomorphisms. Thus
1207: \begin{align*}
1208: \dhh{l} &\lqi \cone \Bigl( F^l\!\sha{X}\cap
1209: \sigma^{2l}\she{X}(l) \to
1210: \cone \bigl( \ZZ(l)_X \to \she{X}(l) \bigr) \Bigr)[-1] \\
1211: &= \cone \bigl( \ZZ(l)_X \oplus (F^l\!\sha{X}\cap
1212: \sigma^{2l}\she{X}(l)) \to \she{X}(l) \bigr)[-1] \\ &\equiv
1213: C(l)\,,
1214: \end{align*}
1215: as wanted.
1216: \end{proof}
1217: It follows from Lemma~\ref{lemma:qi-dhh-bry} that
1218: \begin{math}
1219: \HHH^p(X,\dhh{l}) \cong \dhhH[p]{X}{l}\,,
1220: \end{math}
1221: so we can use either complex to compute the Hermitian-holomorphic
1222: Deligne cohomology groups.
1223:
1224: Again from the cone exact sequence applied to~\eqref{eq:22}, we
1225: see the groups $\dhhH{X}{l}$ also satisfy the exact sequence
1226: \begin{equation}
1227: \label{eq:25}
1228: \dotsm \longrightarrow \delH[2l-1]{X}{\RR}{l}
1229: \longrightarrow
1230: \dhhH[2l]{X}{l} \longrightarrow
1231: \delH[{2l}]{X}{\ZZ}{l} \oplus A^{(l,l)}(X)_{\RR(l)}
1232: \longrightarrow \delH[{2l}]{X}{\RR}{l}
1233: \longrightarrow \dotsm
1234: \end{equation}
1235: which we can rewrite as
1236: \begin{equation*}
1237: \dotsm \longrightarrow \delH[2l-1]{X}{\RR}{l}
1238: \longrightarrow \dhhH[2l]{X}{l}
1239: \longrightarrow \delH[{2l}]{X}{\ZZ}{l} \oplus
1240: A^{(l,l)}(X)_{\RR(l)} \longrightarrow H^{2l}(X,\RR(l))\cap
1241: H^{l,l}(X) \longrightarrow \dotsm
1242: \end{equation*}
1243: Thus the elements of $\dhhH[2l]{X}{l}$ map onto those
1244: $(l,l)$-forms representing the Hodge classes corresponding to
1245: $\delH[2l]{X}{\ZZ}{l}$.
1246:
1247: The complexes $\deligne{\ZZ}{l}$, $\deltilde{\RR}{l}$, and
1248: $F^l\!\sha{X}\cap \sigma^{2l}\she{X}(l)$ appearing in the
1249: cone~\eqref{eq:22}, all have cup products. It follows that we
1250: have the \bei{} family of products
1251: \begin{equation}
1252: \label{eq:26}
1253: \dhh{l} \otimes \dhh{k} \overset{\cup_\alpha}{\longrightarrow}
1254: \dhh{l+k} \,.
1255: \end{equation}
1256: \begin{remark}
1257: The product $\cup_\alpha$ in eq.\eqref{eq:26} should be
1258: intended in the modified version provided by~\eqref{eq:5},
1259: since the complexes $\deligne{\ZZ}{l}$ and $\deltilde{\RR}{l}$
1260: have product structures that are compatible with the map
1261: $\rho_l$ only up to homotopy (given by formula~\eqref{eq:16}).
1262: Moreover, these complexes have product structures that are
1263: themselves graded commutative up to homotopy: that the
1264: product~\eqref{eq:26} is graded commutative up to homotopy
1265: follows from Proposition~\ref{prop:4} in the appendix.
1266: \end{remark}
1267:
1268: \subsection{Hermitian holomorphic line bundles}
1269: \label{sec:line-bun}
1270:
1271: A hermitian holomorphic line bundle or, equivalently, a metrized
1272: line bundle, cf. \cite{lang:arakelov}, is a holomorphic line
1273: bundle $L$ over $X$ together with a hermitian fiber metric $\rho
1274: : L \rightarrow \RR_{\geq 0}$. As a rule, we will not distinguish
1275: $L$ and its sheaf of holomorphic sections. We will also use the
1276: alternate notation $\abs{s}_\rho$ to denote the length of a local
1277: section $s$ of $L$ with respect to $\rho$. Metrized line bundles
1278: can be tensor multiplied and an inverse is defined, see. op. cit.
1279: An isomorphism of metrized line bundles $(L,\rho)$ and
1280: $(L^\prime,\rho^\prime)$ is defined in the obvious way, namely it
1281: is a map $\phi : L \to L^\prime$ such that $\abs{s}_\rho =
1282: \abs{\phi(s)}_{\rho^\prime}$ for some local section $s$ of $L$.
1283: We denote by $\widehat{\pic{X}}$ the group of isomorphism classes
1284: of metrized line bundles.
1285:
1286: If $L$ is trivialized over a \cech\ cover $\cover{U}_X =
1287: \{U_i\}_{i\in I}$ by sections $s_i$, then as usual we obtain the
1288: cocycle of transition functions $g_{ij}\in \sho{X}^\times
1289: (U_i\cap U_j)$ by writing $s_j = s_i\, g_{ij}$. Then if
1290: $(L,\rho)$ is a metrized line bundle, we can define the positive
1291: function $\rho_i = \abs{s_i}^2_\rho$, namely the local
1292: representative of the hermitian structure with respect to the
1293: given trivialization. It follows that the various local
1294: representatives satisfy
1295: \begin{equation}
1296: \label{eq:29}
1297: \rho_j = \rho_i \, \abs{g_{ij}}^2\,.
1298: \end{equation}
1299: Let us work out the local version of the isomorphism introduced
1300: above. Let $s'_i$ be a local section of $L'$ over $U_i$.
1301: Introduce analogous (primed) quantities for $L'$ as we just did
1302: for $L$. Given the isomorphism $\phi: L\to L'$ we have $\phi(s_i)
1303: = s'_i\, f_i$, for some $f_i\in \sho{X}^\times(U_i)$. Then we
1304: find $f_i\, g_{ij} = g'_{ij}\, f_j$ and $\rho_i = \rho'_i\,
1305: \abs{f_i}^2$.
1306:
1307: Still working with respect to the chosen cover $\cover{U}_X$, a
1308: \emph{connection} compatible with the holomorphic structure is
1309: the datum of a collection of $(1,0)$-forms $\xi_i \in
1310: \sha[{1,0}]{X}(U_i)$ satisfying
1311: \begin{equation}
1312: \label{eq:30}
1313: \xi_j - \xi_i = d\log g_{ij}\,.
1314: \end{equation}
1315: Note for future reference that $\sha[{1,0}]{X} =
1316: F^1\!\sha[1]{X}$. The connection is compatible with the hermitian
1317: metric if
1318: \begin{equation}
1319: \label{eq:31}
1320: \pi_0(\xi_i) = \onehalf d\log \rho_i\,.
1321: \end{equation}
1322: Using $d=\del +\delb$ and decomposition with respect to
1323: $(p,q)$-types, we find the familiar relation
1324: \begin{equation}
1325: \label{eq:32}
1326: \xi_i = \del \log \rho_i
1327: \end{equation}
1328: for the \emph{unique} connection compatible with both the complex
1329: and hermitian structures \cite{gh:alg_geom}.
1330:
1331: The global $2$-form
1332: \begin{equation}
1333: \label{eq:33}
1334: c_1(\rho) = \delb\del\log \rho_i
1335: \end{equation}
1336: represents the first Chern class of $L$ in $H^2(X,\RR(1))$.
1337: Actually, the class of $c_1(\rho)$ is a pure Hodge class in
1338: $H^{1,1}(X)$ and, according to the examples, it is the image of
1339: the first Chern class of $L$ under the map $\delH[2]{X}{\ZZ}{1}
1340: \to H^2_\mathcal{D}(X,\RR(1))$ induced by $\ZZ(1) \to \RR(1)$.
1341: Observe that $c_1(\rho) = c_1(\rho')$ under the isomorphism
1342: considered above.
1343:
1344: The following proposition can be found, for example, in
1345: refs.~\cite{brymcl:deg4_II,bry:quillen}. (It apparently was
1346: observed first by Deligne, cf.~\cite{esn:char}.) The proof is
1347: based on writing out an explicit cocycle in a \cech\ resolution
1348: of $C(1)^\bullet$ or $\dhh{1}$. This calculation will be needed
1349: later on with the complex $\dhh{1}$, so we provide a proof here.
1350: \begin{proposition}
1351: \label{prop:1}
1352: \begin{equation*}
1353: \widehat{\pic{X}} \cong \dhhH[2]{X}{1}
1354: \end{equation*}
1355: \end{proposition}
1356: \begin{proof}
1357: Recall that $\dhh{1}$ is the cone of the map $\rho_1 -\jmath$,
1358: where $\rho_1: \deligne{\ZZ}{1} \to \deltilde{\RR}{1}$, and
1359: \begin{math}
1360: \jmath : F^1\!\sha{X}\cap \sigma^{2}\she{X}(1) \to
1361: \deltilde{\RR}{1}\,.
1362: \end{math}
1363: By unraveling the structure of all the cones involved we have:
1364: \begin{equation*}
1365: \begin{CD}
1366: \ZZ(1)_X @>>> \shomega[1]{X}\oplus\sho{X} @>>>
1367: \shomega[2]{X}\oplus \shomega[1]{X} @>>>
1368: \shomega[3]{X}\oplus \shomega[2]{X} @>>> \dotsm \\
1369: & & @VV{\imath\oplus\pi_0}V @VV{\imath\oplus\pi_0}V
1370: @VV{\imath\oplus\pi_0}V \\
1371: & & F^1\!\sha[1]{X} \oplus \she[0]{X} @>>> F^1\!\sha[2]{X}
1372: \oplus \she[1]{X} @>>>
1373: F^1\!\sha[3]{X} \oplus \she[2]{X} @>>> \dotsm \\
1374: & & & & @AA{\jmath\oplus 0}A @AA{\jmath\oplus 0}A \\
1375: & & & & F^1\!\sha[2]{X}\cap \she[2]{X}(1) @>>>
1376: F^1\!\sha[3]{X}\cap \she[3]{X}(1) @>>> \dotsm
1377: \end{CD}
1378: \end{equation*}
1379: With respect to this diagram, an element of total degree $2$
1380: can be written in the form:
1381: \begin{equation}
1382: \label{eq:37}
1383: \setlength{\extrarowheight}{4pt}
1384: \begin{array}{c|c|c}
1385: c_{ijk} & -\;d \log\,g_{ij}\oplus \log\,g_{ij} & 0 \\
1386: \hline
1387: 0 & \xi_i \oplus \sigma_i & \mathsf{X} \\
1388: \hline
1389: 0 & 0 & \eta_i
1390: \end{array}
1391: \end{equation}
1392: for $\xi_i\in F^1\!\sha[1]{X}(U_i)$, $\sigma_i \in
1393: \she[0]{X}(U_i)$ and $\eta_i\in (F^1\!\sha[2]{X}\cap
1394: \she[2]{X}(1))(U_i)$. To make sense out of~\eqref{eq:37}, note
1395: that each entry is an element of the object in the
1396: corresponding position in the left $3\times 3$ part of the
1397: previous diagram. Then since the total degree is $2$, the
1398: degree of each element \emph{in the complex} $\dhh{1}$ is $2$
1399: minus the \cech\ degree as found in~\eqref{eq:37}. The top line
1400: is the class of $L$ in $\deligne[2]{\ZZ}{1}$. Finally, the
1401: entry marked $\mathsf{X}$ means there is no applicable
1402: element---it would have degree $3$.
1403:
1404: A totally routine calculation shows that~\eqref{eq:37} is a
1405: degree $2$ cocycle if and only if
1406: relations~\eqref{eq:30},~\eqref{eq:31} and $\eta_i = d\xi_i =
1407: 2\,\delb\del \sigma_i$ are satisfied (with $\sigma_i = \onehalf
1408: \log \rho_i$). Thus $\eta_i = c_1(\rho)\vert_{U_i}$ and all the
1409: relations defining a metrized line bundle with its canonical
1410: connection are satisfied. The verification that adding an
1411: appropriate coboundary to the cocycle leads to an isomorphic
1412: metrized bundle in the sense explained above is also routine.
1413: Finally, the correspondence between component-wise addition of
1414: cocycles modulo coboundaries and the group operation in
1415: $\widehat{\pic{X}}$ is again a direct verification.
1416: \end{proof}
1417: From the proof of Proposition~\ref{prop:1} we have that the
1418: cocycle representing $(L,\rho)$ in the \cech\ resolution of
1419: $\dhh{1}$ has the expression:
1420: \begin{equation}
1421: \label{eq:38}
1422: \setlength{\extrarowheight}{4pt}
1423: \begin{array}{c|c|c}
1424: c_{ijk} & -\;d \log\,g_{ij}\oplus \log\,g_{ij} & 0 \\
1425: \hline
1426: 0 & \del \log\rho_i \oplus \onehalf \log \rho_i & \mathsf{X} \\
1427: \hline
1428: 0 & 0 & \delb\del\,\log \rho_i
1429: \end{array}
1430: \end{equation}
1431: This cocycle is expressed purely in terms of holomorphic and
1432: metric data (the local expression for the fiber metric $\rho$),
1433: and it also explicitly encodes the canonical connection and its
1434: curvature.
1435:
1436: By way of comparison, a degree two cocycle in the \cech\
1437: resolution of $C(1)^\bullet$ with respect to the same cover would
1438: be given by a triple
1439: \begin{equation*}
1440: \bigl( \eta_i\oplus \sigma_i, \log u_{ij}, c_{ijk} \bigr)\,,
1441: \end{equation*}
1442: where $\eta_i\in \sha[(1,1)]{X}(U_i)_{\RR(1)}$, $\sigma_i \in
1443: \she[1]{X}(1)(U_i)$, $\log u_{ij}\in \she[0]{X}(1)(U_{ij})$ and
1444: $c_{ijk}\in \ZZ(1)$, satisfying the following relations: aside
1445: from the standard cocycle condition for the $\TT$-valued
1446: functions $u_{ij}$, we must have
1447: \begin{equation}
1448: \label{eq:86}
1449: \begin{gathered}
1450: d\eta_i = 0\,,\\
1451: d\sigma_i=\eta_i\,,\\
1452: \sigma_j -\sigma_i = d \log u_{ij}\,.
1453: \end{gathered}
1454: \end{equation}
1455: Given a holomorphic trivialization of $(L,\rho)$ as above, the
1456: position
1457: \begin{equation}
1458: \label{eq:87}
1459: \eta_i = \delb\del\log \rho_i\,, \quad
1460: \sigma_i = \onehalf d^c \log \rho_i\,, \quad
1461: u_{ij} = g_{ij}/\abs{g_{ij}}
1462: \end{equation}
1463: satisfies relations~\eqref{eq:86}. From here, the canonical
1464: connection is recovered. It is also easily seen that another
1465: solution of~\eqref{eq:86} is equivalent to the one above. The
1466: following is evident:
1467: \begin{lemma}
1468: \label{lem:1}
1469: Relations~\eqref{eq:87} provide an explicit quasi-isomorphism
1470: between the \cech\ resolutions of $\dhh{1}$ and $C(1)^\bullet$.
1471: \end{lemma}
1472:
1473: On the
1474: other hand, in~\eqref{eq:38} the relations defining the canonical
1475: connection are already enforced at the level of the complex.
1476:
1477: Finally, if $[L,\rho]$ and $[L',\rho']$ are the classes
1478: corresponding to the metrized bundles $(L,\rho)$ and
1479: $(L',\rho')$, then we write
1480: \begin{math}
1481: [ L\otimes L', \rho\rho'] = [L,\rho] + [L',\rho']\,.
1482: \end{math}
1483:
1484: \subsection{Cup product of hermitian holomorphic line bundles}
1485: \label{sec:Cup-prod-herm}
1486:
1487: If $L$ and $L'$ are two line bundles on $X$, their cup product in
1488: Deligne cohomology would produce a class in $\delH[4]{X}{\ZZ}{2}$
1489: denoted $\tame{L}{L'}$. Calculations were carried out
1490: in~\cite{brymcl:deg4_II} where a geometric interpretation of
1491: $\tame{L}{L'}$ as a 2-gerbe and its connection with the
1492: determinant of cohomology (when $X$ is the total space of a
1493: family of Riemann surfaces) were put forward. The structure
1494: ensuing from the generalization to line bundles equipped with a
1495: hermitian metric was further analyzed in ref.~\cite{bry:quillen}
1496: by means of hermitian holomorphic Deligne cohomology.
1497:
1498: The cup product of two metrized line bundles $(L,\rho)$ and
1499: $(L',\rho')$ in hermitian holomorphic Deligne cohomology produces
1500: a class in $\dhhH[4]{X}{2}$. Despite the more involved definition
1501: of $\dhh{l}$ as opposed to that of $C(l)^\bullet$, it will
1502: nonetheless be more advantageous from the perspective of
1503: sect.~\ref{sec:conf-metr-herm} to use the former to calculate the
1504: desired class. Thus let us explicitly compute a representative of
1505: this class in terms of the expression~\eqref{eq:38} for the class
1506: of a metrized line bundle and the modified \bei\
1507: product~\eqref{eq:5} for
1508: \begin{equation*}
1509: \dhh{1} \otimes \dhh{1} \xrightarrow{\cup_{\alpha}}
1510: \dhh{2} \,.
1511: \end{equation*}
1512: A diagram for the complex $\dhh{2}$ analogous to that for
1513: $\dhh{1}$ displayed in the proof of Proposition~\ref{prop:1} is:
1514: \begin{equation}
1515: \label{eq:39}
1516: \begin{CD}
1517: \ZZ(2)_X @>>> \sho{X} @>>> \shomega[2]{X}\oplus
1518: \shomega[1]{X} @>>> \shomega[3]{X}\oplus \shomega[2]{X} @>>>
1519: \shomega[4]{X}\oplus \shomega[3]{X} @>>>
1520: \dotsm \\
1521: & & @VV{\pi_1}V @VV{\imath\oplus\pi_1}V
1522: @VV{\imath\oplus\pi_1}V @VV{\imath\oplus\pi_1}V \\
1523: & & \she[0]{X}(1) @>>> F^2\!\sha[2]{X} \oplus \she[1]{X}(1)
1524: @>>> F^2\!\sha[3]{X} \oplus \she[2]{X}(1) @>>>
1525: F^2\!\sha[4]{X} \oplus \she[3]{X}(1) @>>>
1526: \dotsm \\
1527: & & & & & & & & @AA{\jmath\oplus 0}A \\
1528: & & & & & & & & F^1\!\sha[4]{X}\cap \she[4]{X}(2) @>>> \dotsm
1529: \end{CD}
1530: \end{equation}
1531: The cup product of two elements of the form~\eqref{eq:38} in the
1532: \cech\ resolution of $\dhh{1}$ would result in an analogous
1533: $5\times 3$ table. The degrees (in the cone) start from $0$ in
1534: the leftmost entry in the top and bottom rows, and from $1$ in
1535: the center one. The overall signs are determined by those in the
1536: \bei\ product~\eqref{eq:5} plus those arising from the \cech\
1537: resolution as explained in sect.~\ref{sec:notation}. Actually, it
1538: is visually less cumbersome to display the resulting expression
1539: as in~\eqref{tab:big-cup} below, where the corresponding
1540: bidegrees are explicitly indicated: the first degree is the
1541: overall degree in the cone and the second is the \cech\ degree.
1542: (We have explicitly written only the nonzero terms.)
1543: \begin{small}
1544: \begin{equation}
1545: \renewcommand{\onehalf}{\tfrac{1}{2}}
1546: \renewcommand{\arraystretch}{1.3}
1547: %%\setlength{\extrarowheight}{2pt}
1548: \begin{array}{c|c}
1549: (0,4) & c_{ijk}c'_{klm} \\ \hline
1550: (1,3) & -c_{ijk} \log \, g'_{kl} \\ \hline
1551: (2,2) & \bigl( -d \log \, g_{ij} \wedge d \log\, g'_{jk} \oplus
1552: -\log \, g_{ij} \, d\log\,g'_{jk} \bigr) \bigoplus
1553: -\pi_1\log\,g_{ij}\, \log\,\abs{g'_{jk}} \\ \hline
1554: (3,1) &
1555: \begin{aligned}
1556: \bigl( -(1-\alpha) \, d\log g_{ij} \wedge \del \log
1557: \rho'_j +\alpha \, \del \log \rho_i \wedge d \log g'_{ij}
1558: \bigr)
1559: \oplus
1560: \Bigl\{ (1-\alpha) \bigl( d^c\log\,\abs{g_{ij}}\;
1561: \onehalf \log\rho'_j &-\log\,\abs{g_{ij}}\; \onehalf d^c
1562: \log\,\rho'_j
1563: \bigr) \\
1564: +\alpha \bigl( -\onehalf d^c
1565: \log\rho_i\;\log\,\abs{g'_{ij}} &+\onehalf \log\rho_i\;
1566: d^c\log\,\abs{g'_{ij}} \bigr) \Bigr\}
1567: \end{aligned} \\ \hline
1568: (4,0) &
1569: \begin{aligned}
1570: \delb\del \log\rho_i \wedge
1571: \delb\del \log \rho'_i
1572: \bigoplus
1573: \bigl( & \alpha\, \delb\del\log\rho_i \wedge \del \log
1574: \rho'_i +(1-\alpha)\, \del\log \rho_i \wedge
1575: \delb\del\log\rho'_i
1576: \bigr) \\
1577: & \oplus \bigl( \alpha\, \delb\del \log\rho_i\;
1578: \onehalf\log\rho'_i +(1-\alpha)\, \onehalf \log\rho_i \;
1579: \delb\del\log\rho'_i \bigr)
1580: \end{aligned}
1581: \end{array}
1582: \label{tab:big-cup}
1583: \end{equation}
1584: \end{small}
1585:
1586:
1587: \subsection{Integration along the fiber}
1588: \label{sec:integr-along-fiber}
1589:
1590: To conclude this introduction, let us quickly mention that
1591: Hermitian holomorphic Deligne cohomology has an integration along
1592: the fiber map. Namely, if $\pi\colon X\rightarrow S$ is a proper
1593: submersion of complex manifolds, it follows from
1594: \cite{MR91j:14017} that Deligne cohomology has an integration
1595: along the fiber $\int_\pi$, hence by
1596: \cite{brymcl:deg4_II,bry:quillen} there is a map
1597: \begin{equation}
1598: \label{eq:28}
1599: \dhhH[i]{X}{l} \longrightarrow \dhhH[i-2d]{S}{l-d}
1600: \end{equation}
1601: where $d$ is the complex dimension of the fiber, and commutative
1602: diagrams analogous to \cite[Theorem 5.1]{bry:quillen}.
1603:
1604: We will be interested in the case of complex relative dimension
1605: $1$, namely $\pi \colon X\rightarrow S$ is a holomorphic
1606: fibration with compact connected Riemann surfaces as fibers. From
1607: \eqref{eq:28} in degree $4$ we have the map:
1608: \begin{math}
1609: \dhhH[4]{X}{2} \rightarrow \dhhH[2]{S}{1}
1610: \end{math}
1611: into the group of complex hermitian line bundles on $S$ (cf.\
1612: Proposition \ref{prop:1}).
1613:
1614: We refer to refs.\ \cite{brymcl:deg4_II,bry:quillen} for
1615: a complete treatment of the map~\eqref{eq:28}, in particular for
1616: the ``trace'' map
1617: \begin{equation*}
1618: \RRR^\bullet\pi_* (\dhh{l}) \longrightarrow
1619: \dhh[\bullet-2d]{l-d}\,,
1620: \end{equation*}
1621: which induces~\eqref{eq:28} at the level of cohomology. (Clearly
1622: any quasi-isomorphic model for Hermitian-holomorphic cohomology
1623: will do.) We will limit ourselves to observe the following: if
1624: $\cover{U}_S$ is a good cover of $S$, and $\cover{U}_X$ a good
1625: cover of $X$ which refines $\pi^{-1}\cover{U}_S$, the $i$-{th}
1626: direct image $\RRR^i\pi_* \sheaf{F}^\bullet$ of any complex
1627: $\sheaf{F}^\bullet$ on $X$ is the sheaf on $S$ associated to the
1628: presheaf
1629: \begin{equation*}
1630: V \longmapsto H^i \bigl(\Tot
1631: \Check{C}^\bullet ( \cover{U}_X\cap \pi^{-1}(V),\sheaf{F}^\bullet)
1632: \bigr)\,.
1633: \end{equation*}
1634: Then unraveling the cone structure of ${\dhh{l}}_X$ (or of
1635: $C(l)^\bullet_X$) reduces to computing the direct images of
1636: $\ZZ(l)_X$ and $\sha{X}$, the de~Rham complex. The trace map is
1637: then obtained by capping a total cocycle, say in
1638: \begin{equation}
1639: \label{eq:89}
1640: \bigoplus_{p+q=i}
1641: \Check{C}^q ( \cover{U}_X\cap \pi^{-1}(V),\sha[p]{X})\,,
1642: \end{equation}
1643: with a representative of the fundamental class $[M]$ of the
1644: smooth model $M$ of the fiber of $\pi^{-1}(V)\cong V\times M$ (as
1645: smooth manifolds). Explicit representatives for $[M]$ for a
1646: triangulation subordinated to the nerve of $\cover{U}_X\cap
1647: \pi^{-1}(V)$ are computed in ref.\ \cite{aldtak2000}. If
1648: $\omega^p_{i_0,\dots,i_q}$ are the components of a cocycle
1649: $\omega^i_V$ in~\eqref{eq:89}, then
1650: \begin{equation*}
1651: \omega^i_V \cap [M]
1652: =\sum \omega^p_{i_0,\dots,i_q} \cap \Delta^{2d-q}_{i_0,\dots,i_q}
1653: \end{equation*}
1654: where the $\Delta^{2d-q}_{i_0,\dots,i_q}$ are \emph{signed}
1655: generators of the nerve of dimension $2d -q$. The exact
1656: combinatorics can be found in ref.\ \cite{aldtak2000}, so we will
1657: not further pursue the matter here.
1658:
1659: Let us conclude by noticing that for $\pi\colon X\to S$ the cup
1660: product of $(L,\rho)$ and $(L',\rho')$ induces a metrized line
1661: bundle on $S$. {F}rom the preceding discussion, the connection
1662: component, for example, will be obtained by following the (two)
1663: components $F^2\!\sha{X}\cap \she[\bullet \geq 4]{X}(2)$
1664: in~\eqref{tab:big-cup}. The one in degree $(4,0)$ will be capped
1665: with generators $\Delta^{2}$ of dimension $2$ (hence integrated
1666: over $2$-simplices), those in degree $(3,1)$ with generators
1667: $\Delta^{1}$ of dimension $1$.
1668:
1669:
1670: From \cite[Theorem 6.1]{brymcl:deg4_II} there is a
1671: corresponding map
1672: \begin{math}
1673: A^{(2,2)}(X)_{\RR(2)} \rightarrow A^{(1,1)}(S)_{\RR(1)}\,.
1674: \end{math}
1675: {F}rom the explicit cocycles we have computed the corresponding
1676: representative in $H^2(S,\RR(1))\cap H^{1,1}(S)$ of the first
1677: Chern class of the resulting metrized line bundle is
1678: \begin{equation*}
1679: \int_\pi c_1(\rho) \wedge c_1(\rho')\,,
1680: \end{equation*}
1681: as in \cite[Proposition 6.6.1]{MR89b:32038}.
1682:
1683:
1684: \section{Conformal metrics and the Liouville functional}
1685: \label{sec:Conf-metr-Liouv}
1686:
1687: Let $X$ be a compact Riemann surface of genus $g\geq 2.$ For simplicity
1688: we can assume $X$ to be connected, although this is not necessary, and
1689: in fact this assumption will be dropped when dealing with Kleinian
1690: groups.
1691:
1692: Let $\cm{X}$ be the space of conformal metrics on $X$. Locally on $X$,
1693: if $z$ is a local analytic coordinate defined on an open set $U$, any
1694: metric $ds^2$ can be represented as
1695: \begin{equation*}
1696: ds^2 = \rho \, \abs{dz}^2
1697: \end{equation*}
1698: for a positive function $\rho : U \to \RR_{>0}$. According to
1699: sec.~\ref{sec:line-bun}, a conformal metric corresponds to
1700: considering the metrized line bundle $(T_X,\rho)$, where $T_X$ is
1701: the holomorphic tangent bundle of $X$. With respect to a cover
1702: $\cover{U}_X$ of $X$, the conformal factors $\rho_i$ and
1703: $\rho_j$, associated to $U_i$ and $U_j$ respectively, satisfy the
1704: relation
1705: \begin{equation}
1706: \label{eq:41}
1707: \rho_j = \rho_i\, \abs{z'_{ij}}^2\,,
1708: \end{equation}
1709: where $z'_{ij} = dz_i/dz_j$ and $z_i$ is a local coordinate
1710: defined over $U_i$.
1711:
1712: It follows from the uniformization theorem that there exists a
1713: unique conformal metric of scalar curvature equal to $-1$, the
1714: Poincar\'e metric. Locally on $U_i\subset X$, the condition for
1715: the metric to have curvature $-1$ is equivalent to the nonlinear
1716: PDE
1717: \begin{equation}
1718: \label{eq:42}
1719: \frac{\del^2}{\del z_i\del\Bar z_i} \phi_i = \onehalf
1720: \exp {\phi_i} \,,
1721: \end{equation}
1722: known as the Liouville equation, for the smooth function $\phi_i
1723: = \log \rho_i$. Observe that equation~\eqref{eq:42} can be
1724: written in the form
1725: \begin{equation*}
1726: c_1(\rho) = \sqrt{-1}\,\omega_\rho\,,
1727: \end{equation*}
1728: where we have used the K\"ahler form associated to the metric:
1729: \begin{equation*}
1730: \omega_\rho\vert_{U_i} = \ihalf \rho_i \,
1731: dz_i\wedge d\Bar z_i\,.
1732: \end{equation*}
1733: This representation makes it apparent that the Liouville equation
1734: is independent of the choice of the coordinate system. On the
1735: other hand, a direct verification of this fact is immediate using
1736: $\phi_j -\phi_i = \log \abs{z'_{ij}}^2$.
1737:
1738:
1739: \subsection{Variational problem for conformal metrics}
1740: \label{sec:vari-probl-conf}
1741:
1742: It is well known that the Liouville equation has a local
1743: variational principle in the following sense. Let $D$ be a region
1744: in the complex plane. Then equation~\eqref{eq:42} is the
1745: Euler-Lagrange equation for the variational problem defined by
1746: the action functional
1747: \begin{equation}
1748: \label{eq:43}
1749: S[\phi] = \ihalf \int_D \bigl( \del \phi \wedge \delb \phi
1750: + e^\phi\,dz\wedge d\Bar z \,\bigr) \,,
1751: \end{equation}
1752: defined on smooth functions $\phi :D \to \RR$, with the condition
1753: that variations
1754: \begin{math}
1755: \left.\frac{d}{d\alpha} (\phi_\alpha) \right\vert_{\alpha=0}
1756: \end{math}
1757: of $\phi$ be zero on $\del D$. However, it is easily seen that
1758: the functional~\eqref{eq:43} cannot be defined globally on $X$,
1759: since, as a consequence of~\eqref{eq:41}, the first term under
1760: the integral sign would not yield a well-defined $2$-form on $X$.
1761: (The second one would of course present no problems, it would
1762: just give the area $A_X(\rho)$ of $X$ with respect to the given
1763: metric $\rho\,\abs{dz}^2 = e^\phi\,\abs{dz}^2$.)
1764: Accordingly, it is convenient to write the integrand in~\eqref{eq:43} as
1765: \begin{equation*}
1766: \sqrt{-1}\,\omega[\phi] + \omega_\rho
1767: \end{equation*}
1768: where we have defined the $2$-form\footnote{Note that
1769: equation~\eqref{eq:44} defines an imaginary form. The reason for this
1770: choice will be apparent later.}
1771: \begin{equation}
1772: \label{eq:44}
1773: \omega^{0}[\phi] = \onehalf \del \phi \wedge \delb \phi\,,
1774: \end{equation}
1775: and restrict our considerations
1776: to the first term of~\eqref{eq:43} which we denote
1777: \begin{equation}
1778: \label{eq:45}
1779: \Check{S}[\phi] = \sqrt{-1}\,\int_D \omega^{0}[\phi]\,.
1780: \end{equation}
1781:
1782: There is by now an established procedure on how to address the problem
1783: caused by the fact that~\eqref{eq:44} is not globally defined. In
1784: general terms, given the choice of a conformal metric $ds^2\in\cm{X}$
1785: and a cover $\cover{U}_X$, one works with the full \cech-de~Rham complex
1786: $\Check{C}^\bullet(\cover{U}_X,\she{X}(1))$ with respect to
1787: $\cover{U}_X$, rather than with just differential forms. The
1788: $2$-form~\eqref{eq:44} is then completed to a total degree
1789: $2$-cocycle---to be denoted $\Omega[\phi]$. This results in a class in
1790: $H^2(X,\RR(1))$ after taking cohomology. (Whether or not there also is a
1791: de~Rham type theorem will depend on the acyclicity properties of
1792: $\cover{U}_X$.)
1793:
1794: This scheme has been previously carried out not quite for covers
1795: of $X$ by open sets, but actually for different choices of planar
1796: coverings. For the covering associated to a Schottky
1797: uniformization of $X$ a generalization of eq.~\eqref{eq:43} was
1798: written in ref.\ \cite{zogtak1987-2}. More recently, a detailed
1799: calculation of the cocycle for the general case of a covering
1800: associated to a Kleinian uniformization was carried out in ref.\
1801: \cite{math.CV/0204318} by exploiting the homological methods
1802: developed in ref.\ \cite{aldtak1997}. Note, however, that from the
1803: point of view of ref.\ \cite{aldtak2000} these planar coverings
1804: are ``\'etale'' coordinates on $X$, so the group cohomology
1805: constructions required to work with the various kinds of
1806: uniformization coverings just follow from specializing the \cech\
1807: formalism to the coverings at hand.
1808:
1809: Finally, the integration in eqs.~\eqref{eq:43} or~\eqref{eq:45} should
1810: be replaced by the evaluation of $\Omega[\phi]$ against an appropriate
1811: representative $\Sigma$ of the fundamental class of $X$. The
1812: ``appropriate'' form for both $\Sigma$ and the evaluation will be
1813: dictated by the chosen cover $\cover{U}_X$ \emph{and} the cohomology
1814: theory being used. Typically $\Sigma$ will be a cycle in a double
1815: complex of $\cover{U}_X$-small simplices, where the differentials are
1816: the singular one and the one determined by the face maps induced by the
1817: cover. Thus, in the case of a \cech\ cover, it will be the complex
1818: determined by a triangulation of $X$ subordinated to the open
1819: $\cover{U}_X$, and in the same way, for a planar cover the singular
1820: complex of the planar domain $\cover{U}_X$ tensored with an appropriate
1821: bar resolution of the group of deck transformations. These issues have
1822: been discussed at length in refs.\
1823: \cite{aldtak1997,aldtak2000,math.CV/0204318}, so we will not repeat the
1824: discussion here. Whenever we have a cocycle extending~\eqref{eq:44} and
1825: a cycle $\Sigma$ representing $X$ we state
1826: \begin{definition}
1827: \label{def:2}
1828: The Liouville functional (without the area term) is given by
1829: the evaluation:
1830: \begin{equation}
1831: \label{eq:46}
1832: \Check{S}[\phi] = -\frac{1}{\tate}
1833: \dual{\Omega[\phi]}{\Sigma}\,.
1834: \end{equation}
1835: For the complete functional we add the area term
1836: \begin{equation}
1837: \label{eq:47}
1838: S[\phi] = \Check{S}[\phi] +\frac{1}{2\pi}
1839: \int_X\omega_\rho\,.
1840: \end{equation}
1841: \end{definition}
1842: \begin{remark}
1843: $\dual{\Omega[\phi]}{\Sigma}\in \RR(1)$, and $\Check{S}[\phi]$ (or
1844: $S[\phi]$) is real. Division by $\tate$ is conventional, but note that
1845: $\RR(1) \qi \RR_+$ via $\exp(\cdot / \tate )$. In the sequel it will
1846: be more convenient to work with imaginary classes. (See also
1847: section~\ref{sec:conf-metr-herm}.)
1848: \end{remark}
1849: In next two subsections we examine these constructions in some
1850: detail. For definiteness, we initially make use of an ordinary
1851: open cover. First, we recall the direct construction of a cocycle
1852: generalizing~\eqref{eq:44}, and we show that this way
1853: eq.~\eqref{eq:42} indeed is the resulting extremum condition.
1854: Then we emphasize the role played by Deligne cohomology and the
1855: tame symbol. These aspects will become important when introducing
1856: a Kleinian uniformization later in the paper, when we discuss
1857: connections with the dilogarithm function. In a later section we
1858: shall tackle the question of its geometrical significance by
1859: making full use of the hermitian-holomorphic version of Deligne
1860: cohomology presented in section~\ref{sec:herm-holom-deligne}, and
1861: we show that the cocycle constructed following refs.
1862: \cite{zogtak1987-2,math.CV/0204318} corresponds to the square of
1863: $(T_X,\rho)$ in hermitian holomorphic Deligne cohomology.
1864:
1865:
1866: \subsection{Direct construction of the Liouville cocycle}
1867: \label{sec:Direct-constr-Liouv}
1868:
1869: \subsubsection{Initial setup}
1870: \label{sec:Initial-setup}
1871:
1872: Let $X$ a compact Riemann surface of genus greater than $2$. We
1873: shall not include the area term in our explicit calculations,
1874: therefore it makes sense to extend our considerations to a
1875: general metrized line bundle $(L,\rho)$. Of course, whenever
1876: referring to a conformal metric or to the variational problem for
1877: the Liouville equation, it will be be assumed that $L=T_X$ and
1878: that $g_{ij}= z'_{ij}$. Upon choosing a cover $\cover{U}_X$,
1879: which for now we assume to be a \cech\ cover by open sets, the
1880: pair $(L,\rho)$ is described in terms of the data expounded in
1881: section~\ref{sec:line-bun}. Our starting point will be the
1882: $0$-cochain
1883: \begin{equation}
1884: \label{eq:48}
1885: \omega^{0}_i[\log\rho_i]
1886: =\onehalf\,\del \log \rho_i \wedge \delb \log \rho_i
1887: = -\onehalf d \log \rho_i \wedge \onehalf d^c \log \rho_i
1888: \end{equation}
1889: with values in $\she[2]{X}(1)(U_i)$.
1890: \begin{remark}
1891: \label{rem:1}
1892: A generalization for eq.~\eqref{eq:48} would be to consider a
1893: \emph{pair} of metrized line bundles $(L,\rho)$ and $(L',\rho')$, and
1894: then the analog of~\eqref{eq:48} would be
1895: \begin{equation}
1896: \label{eq:49}
1897: \omega^{0}_i[\log\rho_i,\log\rho'_i] = \onehalf \Bigl(
1898: -\onehalf d \log \rho_i \wedge \onehalf d^c \log \rho'_i
1899: +\onehalf d^c \log \rho_i \wedge \onehalf d \log \rho'_i
1900: \Bigr)
1901: \end{equation}
1902: Note, however, that the expressions are quadratic. Moreover,
1903: $L\otimes L'$ has metric $\rho\rho'$, so that there is a
1904: natural ``polarization identity''
1905: \begin{equation}
1906: \label{eq:50}
1907: \omega[\log\rho,\log\rho'] =
1908: \onefourth \omega[\log \rho\rho']
1909: - \onefourth \omega[\log \frac{\rho}{\rho'}]
1910: \end{equation}
1911: where we have omitted the indexes for simplicity of notation. We
1912: shall comment later on the significance of eq.~\eqref{eq:50}.
1913: \end{remark}
1914:
1915:
1916: \subsubsection{Computation}
1917: \label{sec:computation}
1918:
1919: Let us extend~\eqref{eq:48} to a degree $2$ cocycle in the total
1920: simple complex associated to the double complex
1921: \begin{math}
1922: \Check{C}^\bullet(\cover{U}_X,\she{X}(1))
1923: \end{math}
1924: of \cech\ cochains with values in the de~Rham complex of imaginary
1925: smooth forms. This is accomplished in the usual fashion (see
1926: e.g.~\cite{bott1982}) by finding a $1$-cochain of $1$-forms
1927: $\omega^{1}_{ij}[\log \rho]$ on $U_{ij}$ and a $2$-cochain of $0$-forms
1928: $\omega^{2}_{ijk}[\log \rho]$ on $U_{ijk}$ such that the relations
1929: \begin{equation*}
1930: \deltacheck \omega^{0} = -d \omega^{1}\,, \qquad
1931: \deltacheck \omega^{1} = d \omega^{2}\,, \qquad
1932: \deltacheck \omega^{2} = 0
1933: \end{equation*}
1934: are satisfied. Of course, the remaining one, namely $d\omega =0$
1935: is automatically satisfied for dimensional reasons. It turns out
1936: that to a great extent these relations are explicitly computable
1937: without further assumptions, such as that the cover $\cover{U}_X$
1938: be good. The needed calculations are fairly standard, and they
1939: are presented in great detail in ref.~\cite{math.CV/0204318}, so
1940: we shall be brief. (The observation in \cite{lang:arakelov} that
1941: on a Riemann surface for two smooth functions $f$ and $g$ one has
1942: \begin{math}
1943: df \wedge d^c g = dg\wedge d^c f
1944: \end{math}
1945: is useful in carrying out the calculations.) The first two steps
1946: are as follows.
1947:
1948: First, one has:
1949: \begin{equation}
1950: \label{eq:51}
1951: \begin{gathered}
1952: \omega^{0}_j[\rho] - \omega^{0}_i[\rho]
1953: = -d\,\omega^{1}_{ij} [\rho] \\
1954: \omega^{1}_{ij} [\rho]
1955: = \onehalf\log\rho_i\;d^c \log\,\abs{g_{ij}}
1956: +d^c \log\,\abs{g_{ij}}\; \onehalf \log \rho_j
1957: \end{gathered}
1958: \end{equation}
1959: The next step yields:
1960: \begin{equation*}
1961: \begin{aligned}
1962: \deltacheck\bigl( \omega^{1}[\log\rho] \bigr)_{ijk} &=
1963: \omega^{1}_{jk} [\rho] -\omega^{1}_{ik} [\rho]
1964: +\omega^{1}_{ij} [\rho] \\
1965: &= \log\,\abs{g_{ij}} \; d^c \log\,\abs{g_{jk}}
1966: -d^c \log\,\abs{g_{ij}} \; \log\,\abs{g_{jk}}
1967: \end{aligned}
1968: \end{equation*}
1969: and notice that
1970: \begin{math}
1971: d^c \log\,\abs{g_{ij}} = \pi_1 d\,\log \,g_{ij}\,,
1972: \end{math}
1973: and
1974: \begin{math}
1975: \pi_{p+q-1}(a\wedge b) = \pi_p(a)\wedge \pi_{q-1}(b) +
1976: \pi_{p-1}(a)\wedge \pi_q(b)\,,
1977: \end{math}
1978: so we have
1979: \begin{equation}
1980: \label{eq:52}
1981: \deltacheck\bigl( \omega^{1}[\log\rho] \bigr)_{ijk} =
1982: \pi_1\bigl( \log\,g_{ij} \; d\,\log\, g_{jk} \bigr)
1983: -d \bigl( \pi_1(\log\,g_{ij})\; \log\,\abs{g_{jk}} \bigr) \,.
1984: \end{equation}
1985: Observe that now the problem of continuing the descent becomes
1986: independent of the chosen metric $\rho$. The most direct way of
1987: proceeding is the following. If we assume the cover $\cover{U}_X$ to be
1988: acyclic for the de~Rham complex $\she{X}(1)$, then there exists
1989: $\omega^{2}_{ijk} \in \she[0]{X}(1)(U_{ijk})$ such that
1990: $d\omega^{2}_{ijk} = \deltacheck (\omega^{1}[\log\rho])_{ijk}$.
1991: Furthermore, consistency on a quadruple intersection requires that
1992: $\deltacheck\omega^{2}$ be a $3$-cocycle with values in $\RR(1)_X$.
1993: Since $H^3(X,\RR(1))=0$, this cocycle must be a coboundary, therefore,
1994: up to readjusting the constants, there exists a choice of
1995: $\omega^{2}_{ijk}$ such that $\deltacheck\omega^{2} = 0$, and
1996: furthermore, $\omega^{2}_{ijk}$ does not depend on the metric structure.
1997:
1998:
1999: \subsubsection{Solution to the variational problem}
2000: \label{sec:solut-vari-probl}
2001:
2002: The previous preliminary calculation is sufficient from the point of
2003: view of finding the extrema. To this effect, we set $L = T_X$, for
2004: $\rho$ a conformal metric in $\cm{X}$. Notice that the space of
2005: conformal metrics on $X$ is affine over $C^\infty(X,\RR) \equiv
2006: \mathcal{E}^0(X)$: if $ds^2$ and ${ds'}^2$ are two conformal metrics
2007: with local expressions $\rho_i\,\abs{dz_i}^2$ and
2008: $\rho'_i\,\abs{dz_i}^2$ respectively, then there exists $\sigma \in
2009: C^\infty(X,\RR)$ such that $\log\rho'_i =\sigma\vert_{U_i} +
2010: \log\rho_i$. The change from $\Omega[\log\rho]$ to $\Omega[\log\rho
2011: +\sigma]$ can be exactly computed thanks to the fact that the last step
2012: in the determination of $\Omega[\log\rho]$ is independent of $\rho$ and
2013: the quadratic character of~\eqref{eq:44}. Indeed we have:
2014: \begin{proposition}
2015: \label{prop:2}
2016: \begin{equation*}
2017: S[\log\rho +\sigma] -S[\log\rho] =
2018: \frac{1}{\tate} \int_X \bigl(
2019: \onehalf \del\sigma \wedge \delb\sigma
2020: +\sigma\, c_1(\rho) -\sqrt{-1}(e^\sigma -1)
2021: \omega_\rho\bigr)
2022: \end{equation*}
2023: \end{proposition}
2024: \begin{proof}
2025: The change in $\omega^{0}_i[\log \rho]$ is computed as
2026: \begin{equation*}
2027: \omega^{0}_i[\log \rho + \sigma] -\omega^{0}_i[\log \rho] =
2028: -\onehalf d\sigma_i \wedge \onehalf d^c\sigma_i
2029: + \sigma_i\, \onehalf dd^c \log \rho_i
2030: - \onehalf d(\sigma_i \, d^c \log\rho_i)\,,
2031: \end{equation*}
2032: where we set $\sigma_i \equiv \sigma\rvert_{U_i}$. Note that
2033: the first two terms on the right hand side are globally
2034: well-defined $2$-forms. On the other hand
2035: \begin{equation*}
2036: \omega^{1}_{ij}[\log \rho +\sigma]
2037: -\omega^{1}_{ij}[\log\rho] =
2038: \sigma_i\,d^c \log\,\abs{g_{ij}}\,.
2039: \end{equation*}
2040: Letting $\chi_i = \sigma_i \, d^c \log\rho_i$, we see that
2041: \begin{equation*}
2042: \Omega[\log \rho +\sigma] - \Omega[\log \rho] =
2043: -\onehalf d\sigma \wedge \onehalf d^c\sigma
2044: + \sigma\, \onehalf dd^c \log \rho -D\chi\,,
2045: \end{equation*}
2046: and taking the area terms into account, establishes the
2047: formula.
2048: \end{proof}
2049: As a consequence, we obtain
2050: \begin{corollary}
2051: \label{cor:1}
2052: The Liouville equation~\eqref{eq:42} is the Euler-Lagrange
2053: equation for the Liouville functional~\eqref{eq:47} introduced
2054: in definition~\ref{def:2}. The critical point is
2055: non-degenerate.
2056: \end{corollary}
2057: \begin{proof}
2058: Replacing $\sigma$ with $t\,\sigma$, $t\in\RR$, in the previous
2059: proposition we find the infinitesimal change in $\Omega[\log
2060: \rho]$ to be
2061: \begin{equation*}
2062: \left.\frac{d}{dt}\right\rvert_{t=0}
2063: S[\log \rho + t\sigma] = -\frac{1}{\tate}\, \int_X \sigma
2064: \bigl( c_1(\rho) - \sqrt{-1}\,\omega_\rho \bigr)\,,
2065: \end{equation*}
2066: and it follows that $S[\log\rho]$ has an extremum if and only
2067: if the Liouville equation is satisfied. Non degeneracy follows
2068: from the quadratic part in the exact change formula in
2069: proposition~\ref{prop:2}.
2070: \end{proof}
2071: \begin{remark}
2072: The fact that the change in the cocycle is given by a pure $2$-form
2073: term up to total coboundary can also be analyzed in terms of gluing
2074: properties of variational bicomplexes, cf. Theorem 1.2 in ref.\
2075: \cite{MR1908413} and the proof of Theorem 1 in ref.\ \cite{aldtak2000}.
2076: From this perspective, Corollary~\ref{cor:1} is a direct consequence
2077: of the affine structure of the space $\cm{X}$ of conformal metrics.
2078: \end{remark}
2079:
2080:
2081: \subsection{A cup product}
2082: \label{sec:cup-product}
2083: Formula~\eqref{eq:52} can be handled in a more geometric fashion
2084: as follows. From sec.~\ref{sec:remarks-dilogarithm} we can
2085: rewrite~\eqref{eq:52} as
2086: \begin{equation}
2087: \label{eq:53}
2088: \deltacheck\bigl( \omega^{1}[\log\rho] \bigr)_{ijk} =
2089: -r_2(g_{ij}, g_{jk})\,,
2090: \end{equation}
2091: and we have the collection of tame symbols
2092: $\tame{g_{ij}}{g_{jk}}$ associated any triple intersection
2093: $U_{ijk}$ in the cover $\cover{U}_X$, \cite{brymcl:deg4_II}.
2094: These symbols glue to form a global symbol $\tame{L}{L}$. As a
2095: cohomology class on a curve $X$, however, $\tame{L}{L}$ will be
2096: zero (that is, there is a global object in the associated
2097: $2$-stack, cf. \cite{bry:quillen}), so that it will be possible
2098: to choose local functions $L_{ijk}$ such that
2099: \begin{equation*}
2100: d L_{ijk} = -\log\,g_{ij} \; d\,\log\, g_{jk}\,,
2101: \end{equation*}
2102: as explained in detail in \cite{aldtak2000}. Note that we still need the
2103: cover $\cover{U}_X$ to be fine enough. Moreover, cf. loc.~cit., the
2104: collection $L_{ijk}$ can be chosen in a way that
2105: \begin{equation*}
2106: \deltacheck L _{ijkl} = -c_{ijk} \log \, g_{kl} + n_{ijkl}\,,
2107: \end{equation*}
2108: where $n_{ijkl} \in \ZZ(2)$. Therefore
2109: from~\eqref{eq:21},~\eqref{eq:52} and~\eqref{eq:53}, we can set
2110: \begin{equation}
2111: \label{eq:54}
2112: \omega^{2}_{ijk} =
2113: -\bigl( \pi_1(\log\,g_{ij})\; \log\,\abs{g_{jk}} \bigr)
2114: -\pi_1 L_{ijk}\,,
2115: \end{equation}
2116: namely according to sec.~\ref{sec:remarks-dilogarithm} we have set
2117: $\omega^{2}_{ijk} = -\bwli (g_{ij},g_{jk})$. Now, the last compatibility
2118: condition is satisfied, indeed we have:
2119: \begin{align*}
2120: \deltacheck \omega^{2}_{ijkl} &=
2121: -c_{ijk}\, \log \,\abs{g_{kl}} -\pi_1 (\deltacheck L_{ijkl}) \\
2122: &= -c_{ijk}\, \log \,\abs{g_{kl}}
2123: +\pi_1 (c_{ijk} \log g_{kl} ) \\
2124: &= 0\,.
2125: \end{align*}
2126: As a result, we obtain a $2$-cocycle in the the total complex associated
2127: to $\Check{C}^\bullet (\cover{U}_X,\she{X}(1))$ as before, with a more
2128: geometric interpretation of $\omega^{2}_{ijk}$ in terms of the
2129: trivialization of the symbol $\tame{L}{L}$. Moreover, notice that if $X$
2130: is a curve then $\she{X}(1)[1] \qi \deltilde{\RR}{2}$ thus we may
2131: interpret the class so determined by $\Omega[\log \rho]=
2132: \omega^{0}[\log\rho] +\omega^{1}[\log\rho] +\omega^{2}$ as a (degree
2133: $3$) class in $\delH[3]{X}{\RR}{2}$.
2134:
2135:
2136: \subsection{Two line bundles}
2137: \label{sec:two-line-bundles}
2138:
2139: For a pair $(L,\rho)$, $(L',\rho')$ of metrized line bundles we can
2140: complete~\eqref{eq:49} to a cocycle $\Omega[\log\rho,\log\rho']$ via an
2141: analogous procedure to the one presented in
2142: sects.~\ref{sec:Direct-constr-Liouv} and~\ref{sec:cup-product}. The
2143: relevant calculations being entirely similar, we limit ourselves to
2144: quoting the relevant expressions. Starting from~\eqref{eq:49}, which we
2145: rewrite in the form
2146: \begin{equation}
2147: \label{eq:55}
2148: \omega^{0}_i[\log\rho_i,\log\rho'_i] =
2149: \onehalf d^c \log \rho_i \wedge \onehalf d \log \rho'_i\,,
2150: \end{equation}
2151: the corresponding expression for the degree $(1,1)$ term is:
2152: \begin{equation}
2153: \label{eq:56}
2154: \omega^{1}_{ij}[\log\rho,\log\rho'] =
2155: \onehalf \log\rho_i\;d^c \log\,\abs{g'_{ij}}
2156: +\onehalf d^c \log\,\abs{g_{ij}}\; \log \rho'_j\,.
2157: \end{equation}
2158: Computing the \cech\ coboundary we find:
2159: \begin{equation}
2160: \label{eq:57}
2161: \begin{aligned}
2162: \bigl(\deltacheck \omega^{1}[\log\rho,\log\rho']\bigr)_{ijk} &=
2163: \log\,\abs{g_{ij}} \; d^c \log\,\abs{g'_{jk}}
2164: -d^c \log\,\abs{g_{ij}} \; \log\,\abs{g'_{jk}} \\
2165: &= -r_2(g_{ij},g'_{jk})\,,
2166: \end{aligned}
2167: \end{equation}
2168: from which $\omega^{2}_{ijk}$ (now independent of $\rho$ and $\rho'$)
2169: can be obtained as $-\bwli (g_{ij},g'_{jk})$ by looking at a collection
2170: $L_{ijk}$ such that
2171: \begin{equation*}
2172: d L_{ijk} = -\log\,g_{ij} \; d\,\log\, g'_{jk}\,,
2173: \end{equation*}
2174: from the triviality of the symbol $\tame{L}{L'}$.
2175:
2176: \subsection{Additional remarks on the Liouville functional and the
2177: Bloch-Wigner dilogarithm}
2178: \label{sec:liouv-funct-bloch}
2179:
2180: We wish to compare the previous constructions to those of ref.\
2181: \cite{math.CV/0204318}. To this end we need to specifically
2182: consider the case of a cover of $X$ provided by a Kleinian
2183: uniformization. (The reader should consult loc.\ cit.\ for
2184: reference and complete details.) The comparison offers a better
2185: perspective on the absence of cohomological obstructions in the
2186: calculations of ref.\ \cite{math.CV/0204318}, and on the
2187: relations with three-dimensional hyperbolic geometry.
2188:
2189: Let $\Gamma$ be a purely loxodromic second kind Kleinian group
2190: satisfying all the conventions spelled out in
2191: Appendix~\ref{sec:kleinian-groups}, to which we refer for the
2192: notation. Let $U_\Gamma\subset \pione$ be the region of
2193: discontinuity, and $X=U_\Gamma/\Gamma$ the resulting (possibly
2194: disconnected) Riemann surface.
2195:
2196: A conformal metric $\rho$ on $X=X_1\sqcup \dots \sqcup X_n$
2197: appears as an automorphic function on $U_\Gamma$:
2198: \begin{equation}
2199: \label{eq:62}
2200: \onehalf \log \rho - \onehalf \log \rho\circ \gamma
2201: = \log\, \abs{\gamma'}\,,\quad \gamma\in \Gamma\,.
2202: \end{equation}
2203: Eq.~\eqref{eq:62} is the direct translation of \eqref{eq:29}
2204: following the principles of \cite{aldtak2000}. Accordingly, the
2205: first two terms of the Liouville cocycle computed by applying the
2206: procedure explained in sect.~\ref{sec:Conf-metr-Liouv} are:
2207: \begin{gather}
2208: \omega^{0}[\log\rho]
2209: = -\onehalf d \log \rho \wedge \onehalf d^c \log \rho \label{eq:63} \\
2210: \omega^{1}_{\gamma} [\log\rho]
2211: = \bigl(\onehalf\log\rho +\onehalf \log \rho
2212: \circ \gamma \bigr)\; d^c \log\,\abs{\gamma'}\label{eq:64}
2213: \end{gather}
2214: and computing the coboundary of~\eqref{eq:64} according to the
2215: prescription in Appendix~\ref{sec:kleinian-groups} yields
2216: \begin{equation}
2217: \label{eq:65}
2218: \deltacheck\bigl( \omega^{1}[\log\rho]
2219: \bigr)_{\gamma_1,\gamma_2}
2220: =-r_2((\gamma_1\gamma_2)', \gamma'_2)\,,
2221: \end{equation}
2222: where $r_2$ has been introduced in eqs.~\eqref{eq:18}
2223: and~\eqref{eq:19}. Note that, as in sect.~\ref{sec:cup-product},
2224: the coboundary of $\omega^{1}$ is a cup product in real Deligne
2225: cohomology:
2226: \begin{equation*}
2227: r_2((\gamma_1\gamma_2)', \gamma'_2)
2228: = (d\log (\gamma_1\gamma_2)', \log\, \abs{(\gamma_1\gamma_2)'})
2229: \cup (d\log \gamma'_2, \log\, \abs{\gamma'_2})\,.
2230: \end{equation*}
2231: where the two classes are associated to the rational functions
2232: $(\gamma_1\gamma_2)'$ and $\gamma'_2$, respectively (cf.\
2233: sect.~\ref{sec:remarks-dilogarithm}). Hence, we can work with the
2234: double complex of group cochains on $\Gamma$ with values in the
2235: real Deligne complex $\deltilde{\RR}{2}\qi \she{U_\Gamma}(1)$ on
2236: the region of discontinuity $U_\Gamma$. Since from this point on
2237: only rational functions with singularities at certain prescribed
2238: points will appear, following ref.\ \cite{math.AG/0003086} we
2239: will consider the Deligne complex on the generic point
2240: $\eta_{\pione}$ of $\pione$.
2241:
2242: For any two elements $\gamma_i, \gamma_j \in \Gamma$ define
2243: $T_{ij}\in \PSL_2(\CC)$ by
2244: \begin{equation}
2245: \label{eq:66}
2246: z \longmapsto T_{ij} (z) = \CR{z}{z_{ij}}{z_j}{\infty}
2247: = \frac{z-z_{ij}}{z-z_j}\,.
2248: \end{equation}
2249: where $z_j = \gamma_j^{-1}(\infty)$ and $z_{ij} =
2250: \gamma_j^{-1}(z_i)$. Following ref.\ \cite{math.CV/0204318}, we
2251: introduce the $1$-cochain on $\Gamma$ with values in
2252: $\she[1]{\pione}(1)(\eta_{\pione})$:
2253: \begin{equation}
2254: \label{eq:68}
2255: \kappa_\gamma = \log\,\abs{c_\gamma} \, \pi_1 d\log \gamma'\,.
2256: \end{equation}
2257: We have:
2258: \begin{lemma}
2259: \label{lemma:kappa}
2260: \begin{math}
2261: r_2((\gamma_1\gamma_2)', \gamma'_2) =
2262: 4\, d \bigl( \bwli \circ T_{12} \bigr)
2263: +\deltacheck\kappa_{\gamma_1,\gamma_2}\,,
2264: \end{math}
2265: where $\bwli = \sqrt{-1} \bwd$, and $\bwd$ is the standard
2266: Bloch-Wigner dilogarithm, cf.\ \ref{sec:remarks-dilogarithm}.
2267: \end{lemma}
2268: \begin{proof}
2269: A straightforward calculation exploiting relation~\eqref{eq:60}.
2270: \end{proof}
2271: Since obviously $d\kappa_\gamma=0$, we can redefine
2272: \begin{math}
2273: \omega^{1}_\gamma \to \omega^{1}_\gamma + \kappa_\gamma\,,
2274: \end{math}
2275: so that using the lemma, from eq.~\eqref{eq:65} we have:
2276: \begin{equation}
2277: \label{eq:69}
2278: \deltacheck\bigl( \omega^{1}[\log\rho]
2279: \bigr)_{\gamma_1,\gamma_2}
2280: = -4\, d\,\bigl(\bwli \circ T_{12} \bigr)\,.
2281: \end{equation}
2282: For convenience of notation, let us temporarily set
2283: \begin{math}
2284: \Hat\omega^{2}_{\gamma_1,\gamma_2} := -4\, \bigl(\bwli \circ
2285: T_{12} \bigr)\,.
2286: \end{math}
2287: We then have the following
2288: \begin{lemma}
2289: \label{lem:3}
2290: \begin{equation}
2291: \label{eq:71}
2292: D(\omega^{0}+\omega^{1}+\Hat\omega^{2}) =
2293: -4 \,\bwli (
2294: \CR{\infty}%
2295: {\gamma_1(\infty)}{\gamma_1\gamma_2(\infty)}%
2296: {\gamma_1\gamma_2\gamma_3(\infty)}) \,.
2297: \end{equation}
2298: \end{lemma}
2299: \begin{proof}
2300: By construction,
2301: \begin{math}
2302: D(\omega^{0}+\omega^{1}+\Hat\omega^{2}) =
2303: \deltacheck\Hat\omega^{2}\,,
2304: \end{math}
2305: and the latter \cech\ coboundary is computed as
2306: \begin{equation*}
2307: \begin{split}
2308: \deltacheck\Hat\omega^{2}_{\gamma_1,\gamma_2, \gamma_3}
2309: &= -4 \Bigl(\bwli (\CR{z}{z_{23}}{z_3}{\infty})
2310: -\bwli (\CR{z}{z_{123}}{z_3}{\infty})\\
2311: &\qquad +\bwli (\CR{z}{z_{123}}{z_{23}}{\infty})
2312: -\bwli (\CR{\gamma_3(z)}{z_{12}}{z_2}{\infty}) \Bigr)
2313: \end{split}
2314: \end{equation*}
2315: In the last term in the previous relation we have
2316: \begin{math}
2317: \CR{\gamma_3(z)}{z_{12}}{z_2}{\infty}
2318: =\CR{z}{z_{123}}{z_{23}}{z_3}\,.
2319: \end{math}
2320: The Bloch-Wigner dilogarithm satisfies the $5$-term
2321: relation~\cite{MR2002g:52013,MR94k:19002}:
2322: \begin{equation}
2323: \label{eq:70}
2324: \sum_{i=0}^4 (-1)^i \bwd ([a_0:\dots :\Hat{a_i}:\dots :a_4]) = 0\,,
2325: \end{equation}
2326: where $a_0,\dots, a_4 \in \pione$ and the hat sign denotes
2327: omission. As a consequence we have:
2328: \begin{equation*}
2329: \deltacheck\Hat\omega^{2}_{\gamma_1,\gamma_2, \gamma_3} =
2330: -4 \,\bwli (\CR{z_{123}}{z_{23}}{z_3}{\infty}) \,,
2331: \end{equation*}
2332: and again by the invariance of the cross ratio, we
2333: obtain~\eqref{eq:71}.
2334: \end{proof}
2335: \begin{remark}
2336: \begin{math}
2337: \bwd (\CR{a}{b}{c}{d} )
2338: \end{math}
2339: is the hyperbolic volume of the ideal hyperbolic tetrahedron
2340: with vertices at the points
2341: \begin{math}
2342: a,b,c,d \in \pione\,,
2343: \end{math}
2344: see, e.g.\ refs.\ \cite{MR84b:53062b,MR88k:57032,MR94k:19002}.
2345: \end{remark}
2346: It follows from the five-term relation~\eqref{eq:70} that the
2347: right hand side of eq.~\eqref{eq:71} defines an $\RR(1)$-valued
2348: $3$-cocycle on $\Gamma$. Moreover, this cocycle is already defined
2349: on $\PSL_2(\CC)$, where its class is known to generate the
2350: Eilenber-Mac~Lane cohomology group
2351: \begin{math}
2352: H^3(\PSL_2(\CC),\RR(1))\,,
2353: \end{math}
2354: \cite{MR2001i:11082,MR88k:57032,MR94k:19002}. It is also known
2355: that up to a factor $2$ it agrees with the imaginary part of the
2356: second Cheeger-Simons universal secondary class $\Hat C_2$.
2357:
2358: Thus the complete Liouville cocycle
2359: \begin{math}
2360: \Omega = \omega^{0}+\omega^{1}+\omega^{2}
2361: \end{math}
2362: subordinated to the cover $U_\Gamma \to X$ is found as follows.
2363: The pullback of $\Hat C_2$ along the inclusion map
2364: \begin{math}
2365: \Gamma \hookrightarrow \PSL_2(\CC)
2366: \end{math}
2367: is zero, since ($\Gamma$ being of second kind) the
2368: $3$-manifold $M_\Gamma = \HH^3/\Gamma$ is non-compact, and
2369: $H^\bullet(\Gamma,\RR(1)) \cong H^\bullet(M_\Gamma,\RR(1))$. It
2370: follows that the restriction of the cocycle given by the
2371: Bloch-Wigner dilogarithm to $\Gamma$ must be a coboundary, hence
2372: there exists a group $2$-cochain $c$ on $\Gamma$ with values in
2373: $\RR(1)$ such that
2374: \begin{equation}\label{eq:67}
2375: 4 \,\bwli (
2376: \CR{\infty}%
2377: {\gamma_1(\infty)}{\gamma_1\gamma_2(\infty)}%
2378: {\gamma_1\gamma_2\gamma_3(\infty)})
2379: = (\deltacheck c)_{\gamma_1,\gamma_2,\gamma_3}\,.
2380: \end{equation}
2381: It follows that the cochain $c$ provides the necessary
2382: ``integration constants,'' namely the required $2$-cochain on
2383: $\Gamma$ with values in $\she[0]{U_\Gamma}$ to complete the
2384: Liouville cocycle is
2385: \begin{math}
2386: \omega^{2}_{\gamma_1,\gamma_2}
2387: = -4\, \bigl(\bwli \circ
2388: T_{12} \bigr)
2389: + c_{\gamma_1,\gamma_2}\,.
2390: \end{math}
2391:
2392:
2393: \section{Conformal metrics and hermitian holomorphic cohomology}
2394: \label{sec:conf-metr-herm}
2395:
2396: In sect.~\ref{sec:Conf-metr-Liouv} we presented a construction of a
2397: degree $2$, $\RR(1)$-valued class corresponding to a conformal metric
2398: $\rho\in \cm{X}$, represented by the cocycle $\Omega[\log\rho]$.
2399: Supplemented by the area of $X$ computed with respect to $\rho$, it
2400: provides a global functional for the variational problem associated to
2401: the Liouville equation~\eqref{eq:42}. We now show that it coincides
2402: with the square of the class of $(T_X,\rho)$ in hermitian holomorphic
2403: Deligne cohomology introduced in sect.~\ref{sec:herm-holom-deligne}.
2404: Moreover, we show that this equality holds at the cocycle level. More
2405: generally, without considering the area term, we show that the cup
2406: product of $(L,\rho)$ and $(L',\rho')$ in hermitian holomorphic Deligne
2407: cohomology coincides with the class of $\Omega[\log\rho,\log\rho']$, and
2408: again the equality in fact holds at the cocycle level.
2409:
2410: \subsection{Comparison on a curve}
2411: \label{sec:comparison-curve}
2412:
2413: Let $X$ be a compact Riemann surface. From the results of
2414: sect.~\ref{sec:Cup-prod-herm}, the cup product of the classes of
2415: $(L,\rho)$ and $(L',\rho')$ in hermitian holomorphic Deligne cohomology
2416: yields a class in $\dhhH[4]{X}{2}$, and on a curve we only capture the
2417: $2$-dimensional part of this class. Indeed, in the exact
2418: sequence~\eqref{eq:25}, the cohomology class corresponding to the symbol
2419: $\tame{L}{L'}\in \delH[4]{X}{\ZZ}{2}$ is zero, and
2420: $A^{(2,2)}(X)_{\RR(2)}$ is also zero for obvious dimensional reasons, so
2421: we have:
2422: \begin{equation*}
2423: \dotsm \longrightarrow \delH[3]{X}{\RR}{2}
2424: \longrightarrow
2425: \dhhH[4]{X}{2} \longrightarrow 0\,.
2426: \end{equation*}
2427: It follows that the class $[L,\rho]\cup [L',\rho']$ must come from an
2428: element in $\delH[3]{X}{\RR}{2}$. As already remarked, on a curve we
2429: have $\deltilde{\RR}{2} \qi \she{X}(1)[-1]$, thus
2430: $\delH[3]{X}{\RR}{2}\cong \HHH^3(X,\she{X}(1)[-1]) \cong
2431: \HHH^2(X,\she{X}(1)) \cong H^2(X,\RR(1))$, in agreement with the
2432: calculations performed in sect.~\ref{sec:Conf-metr-Liouv}.
2433:
2434: More in detail, in complex dimension $1$ the second hermitian
2435: holomorphic Deligne complex $\dhh{1}$ simplifies considerably and
2436: diagram~\eqref{eq:39} reduces to
2437: \begin{equation}
2438: \label{eq:72}
2439: \begin{CD}
2440: \ZZ(2) @>-\imath>> \sho{X} @>-d>> \shomega[1]{X} \\
2441: & & @VV\pi_1V @VV\pi_1V \\
2442: & & \she[0]{X}(1) @>-d>> \she[1]{X}(1) @>-d>> \she[2]{X}(1)
2443: \end{CD}
2444: \end{equation}
2445: so that $\dhh{2}$ becomes just the cone of the morphism
2446: $\deligne{\ZZ}{2} \overset{\pi_1}{\longrightarrow}
2447: \she{X}(1)[-1]$. In other words, on a curve $X$ we have that
2448: $\dhh{2}$ is given by the complex
2449: \begin{equation}
2450: \label{eq:73}
2451: \begin{CD}
2452: \ZZ(2)_X @>{-\imath}>>
2453: \sho{X} @>{(-d,-\pi_1)}>>
2454: \shomega[1]{X}\oplus \she[0]{X}(1) @>{-\pi_1+d}>>
2455: \she[1]{X}(1) @>{d}>>
2456: \she[2]{X}(1)\,,
2457: \end{CD}
2458: \end{equation}
2459: where the differentials have been written explicitly. We can see
2460: the complex $\she{X}$ appears as a subcomplex in~\eqref{eq:73}
2461: and the shift of two positions to the right clearly accounts for
2462: the cohomology degree shift from $2$ to $4$.
2463:
2464: Our main result is the following comparison
2465: \begin{theorem}
2466: \label{thm:2}
2467: Let $X$ be a compact Riemann surface of genus $g>1$. Let $(L,\rho)$
2468: and $(L',\rho')$ be two hermitian holomorphic line bundles. The class
2469: of $[L,\rho]\cup [L',\rho']$ in $\dhhH[4]{X}{2} \cong H^2(X,\RR(1))$
2470: coincides with the one represented by the cocycle $\Omega[\log
2471: \rho,\log \rho']$ constructed in section~\ref{sec:Conf-metr-Liouv}.
2472: \end{theorem}
2473: \begin{proof}
2474: We have observed above that $\dhhH[4]{X}{2} \cong
2475: H^2(X,\RR(1))$, and by construction the class of
2476: \begin{math}
2477: \Omega[\log\rho,\log\rho']
2478: \end{math}
2479: is in $\HHH^2(X,\she{X}(1)) \cong H^2(X,\RR(1))$. Note that for $X$
2480: connected they must coincide up to a proportionality factor, since
2481: $H^2(X,\RR(1))\cong \RR(1)$ in this case. In general, we compute the
2482: proportionality factor from the explicit cocycles from
2483: sects.~\ref{sec:Cup-prod-herm} and~\ref{sec:Direct-constr-Liouv}
2484: to~\ref{sec:two-line-bundles}. (Since $\Omega[\log\rho,\log\rho']$ is
2485: computed under suitable acyclicity assumptions on the cover, so we
2486: will use such a cover to establish the comparison.)
2487:
2488: Let us assume $L$ and $L'$ and their respective hermitian metric
2489: structures are represented by cocycles of type~\eqref{eq:38} with
2490: respect to the chosen cover $\cover{U}_X$. Specializing the general
2491: expression in table~\ref{tab:big-cup} in sect.~\ref{sec:Cup-prod-herm}
2492: to the case at hand we obtain, with reference to~\eqref{eq:73}:
2493: \begin{equation}
2494: \label{eq:74}
2495: \renewcommand{\arraystretch}{1.3}
2496: \begin{array}{cc}
2497: (0,4) & c_{ijk}c'_{klm} \\ \hline
2498: (1,3) & -c_{ijk} \log \, g'_{kl} \\ \hline
2499: (2,2) & -\log \, g_{ij} \, d\log\,g'_{jk} \oplus
2500: -\pi_1\log\,g_{ij}\, \log\,\abs{g'_{jk}} \\ \hline
2501: (3,1) &
2502: (1-\alpha) \bigl( d^c\log\,\abs{g_{ij}}\; \onehalf \log\rho'_j
2503: -\log\,\abs{g_{ij}}\; \onehalf d^c \log\,\rho'_j
2504: \bigr)
2505: +\alpha \bigl( -\onehalf d^c \log\rho_i\;\log\,\abs{g'_{ij}}
2506: +\onehalf \log\rho_i\; d^c\log\,\abs{g'_{ij}} \bigr) \\ \hline
2507: (4,0) &
2508: \alpha\, \delb\del \log\rho_i\;
2509: \onehalf\log\rho'_i +(1-\alpha)\, \onehalf \log\rho_i \;
2510: \delb\del\log\rho'_i
2511: \end{array}
2512: \end{equation}
2513: where we have followed the convention explained in the introduction
2514: for the bidegrees. Let us denote by $\theta^{i}$ the term of bidegree
2515: $(4-i,i)$ in~\eqref{eq:74} and by $\Theta$ the total cocycle. (For
2516: simplicity, we suppress $\rho$ and $\rho'$ from the notation.) A
2517: direct calculation shows that
2518: \begin{equation*}
2519: \theta^{0}_i = \omega^{0}_i + d\lambda^{0}_i\,,\qquad
2520: \theta^{1}_{ij} = \omega^{1}_{ij} -\deltacheck\lambda^{0}_{ij}\,,
2521: \end{equation*}
2522: where $\omega^{0}_i$ is given by eq.~\eqref{eq:55}, $\omega^{1}$
2523: is given by eq.~\eqref{eq:56}, and
2524: \begin{equation*}
2525: \lambda^0_i = \alpha\,
2526: \onehalf d^c \log\rho_i \, \log\rho'_i
2527: +(1-\alpha)\, \onehalf \log\rho_i \, d^c \log\rho'_i\,.
2528: \end{equation*}
2529: Note that at this point we could simply \emph{define} $\Omega = \Theta
2530: -D\lambda^0$. Furthermore, note that $\Omega$ does not explicitly
2531: depend on the parameter $\alpha$ from the \bei\ product~\eqref{eq:5}.
2532: To finish the comparison, let us assume the cover $\cover{U}_X$ allows
2533: us to find a collection $L_{ijk}\in \sho{X}(U_{ijk})$ such that
2534: \begin{equation}
2535: \label{eq:75}
2536: \begin{aligned}
2537: d L_{ijk} &= -\log\,g_{ij} \; d\,\log\, g'_{jk}\\
2538: \deltacheck L _{ijkl} &= -c_{ijk} \log \, g'_{kl} + n_{ijkl}\,,
2539: \end{aligned}
2540: \end{equation}
2541: as in sect.~\ref{sec:cup-product}. In this way we have
2542: \begin{math}
2543: \Omega =
2544: (\omega^{0},\omega^{1}, \omega^{2})
2545: \end{math}
2546: with $\omega^{2}_{ijk} = -\bwli (g_{ij},g'_{jk})$. $\Omega$ is a
2547: cocycle of total degree $2$ in $\Tot \Check{C}^\bullet
2548: (\cover{U}_X,\she{X})$, and it injects (via the exact sequence of the
2549: cone) into $\Tot\Check{C}^\bullet (\cover{U}_X,\dhh{2})$ as
2550: \begin{equation*}
2551: (\omega^{0},\omega^{1}, 0\oplus\omega^{2})\,.
2552: \end{equation*}
2553: Then via equations~\eqref{eq:75} it is easily seen that
2554: \begin{equation*}
2555: \theta^{2}_{ijk}=d_{\dhh{2}} (-L_{ijk})
2556: + 0\oplus \omega^{2}_{ijk}\,,
2557: \end{equation*}
2558: where $d_{\dhh{2}}$ is the differential in $\dhh{2}$, and therefore
2559: \begin{equation*}
2560: \Theta = (\omega^{0},\omega^{1}, 0\oplus \omega^{2})
2561: +D\lambda^{0} +D(-L,n)\,,
2562: \end{equation*}
2563: where we have put $D=d_{\dhh{2}}\pm \deltacheck$ for the total
2564: differential. Thus the two cocycles constructed via the direct method
2565: of sect.~\ref{sec:Direct-constr-Liouv} and the cup product of metrized
2566: bundles define the same class. By direct comparison, the
2567: proportionality factor is $1$.
2568: \end{proof}
2569: In light of the previous theorem, the polarization identity in
2570: Remark~\ref{rem:1} is now easily explained. Using
2571: \begin{math}
2572: [L\otimes L',\rho\rho'] = [L,\rho] + [L',\rho']
2573: \end{math}
2574: and
2575: \begin{math}
2576: [L\otimes {L'}^\vee,\rho/\rho'] = [L,\rho] - [L',\rho'] \,,
2577: \end{math}
2578: and the (graded) commutativity of the cup product
2579: \begin{equation*}
2580: \dhhH[2]{X}{1} \otimes \dhhH[2]{X}{1}
2581: \overset{\cup}{\longrightarrow}
2582: \dhhH[4]{X}{2}\,,
2583: \end{equation*}
2584: we obtain the polarization identity
2585: \begin{equation*}
2586: 4 [L,\rho]\cup [L',\rho'] = [L\otimes L',\rho\rho']^2
2587: -[L\otimes {L'}^\vee,\rho/\rho']^2\,,
2588: \end{equation*}
2589: where the squares in the right hand side refer to $\cup$. A polarization
2590: identity at the level of representative cocycles, and hence the one in
2591: Remark~\ref{rem:1}, follow by applying Thm.~\ref{thm:2} to the latter
2592: identity.
2593:
2594: By choosing $L=L'=T_X$, the holomorphic tangent
2595: bundle of $X$, we immediately obtain:
2596: \begin{corollary}
2597: \label{corollary:1}
2598: Let $\rho\in\cm{X}$ be a conformal metric. The Liouville functional
2599: without area term~\eqref{eq:46} is given by the (evaluation of) the
2600: square $[T_X,\rho]\cup [T_X,\rho]$ with respect to the cup product in
2601: hermitian holomorphic Deligne cohomology. The full-fledged Liouville
2602: functional is obtained by adding the area term
2603: \begin{math}
2604: \frac{1}{2\pi}\int_X \omega_\rho
2605: \end{math}
2606: to~\eqref{eq:46}.
2607: \end{corollary}
2608: \begin{remark}
2609: Due to the specific form of the differential in the complex $\dhh{2}$
2610: the descent equations are explicit and close automatically. Therefore
2611: the cocycle resulting from the calculation of the cup product
2612: sidesteps the problem of the explicit calculation of the last term,
2613: unlike the more direct version from sect.~\ref{sec:Conf-metr-Liouv}.
2614: Thus, thanks to the explicit character of the calculation, specific
2615: assumptions on the nature of the cover $\cover{U}_X$ are not required.
2616: \end{remark}
2617: It follows from Thm~\ref{thm:2}, corollary~\ref{corollary:1} and
2618: the previous remark that definition~\ref{def:2} applies to any
2619: (\'etale) cover $\cover{U}_X \to X$. Indeed, proposition~\ref{prop:2}
2620: from sect.~\ref{sec:Direct-constr-Liouv} can be reformulated at the
2621: cocycle level as follows:
2622: \begin{proposition}
2623: \label{prop:3}
2624: Let $X$ be a compact, genus $g>1$ Riemann surface and let
2625: $\cover{U}_X\to X$ be a cover. For a conformal metric $\rho \in
2626: \cm{X}$ and $\sigma \in C^\infty(X,\RR)$, there is a cocycle
2627: $\Hat{\Omega}_{\cover{U}_X} [\log\rho]$ solving the variational
2628: problem for the Liouville equation.
2629: \end{proposition}
2630: \begin{proof}
2631: If $\rho$ is a conformal metric, let the pair $(T_X,\rho)$ be
2632: represented, as an hermitian line bundle, by a cocycle $c(T_X,\rho)$
2633: with respect to the cover $\cover{U}_X$. We set
2634: \begin{equation*}
2635: \Omega[\log\rho] = c(T_X,\rho)\cup c(T_X,\rho)\,,
2636: \end{equation*}
2637: and a simple calculation starting from eq.~\eqref{eq:74} yields
2638: \begin{equation*}
2639: \Omega[\log\rho +\sigma] - \Omega[\log\rho] =
2640: \sigma\,c_1(\rho) + \onehalf\,\sigma\, \del\delb\sigma +
2641: D\chi\,,
2642: \end{equation*}
2643: where
2644: \begin{math}
2645: \chi_i = \onehalf \sigma \, \onehalf d^c\log \rho_i
2646: -\onehalf d^c\sigma\, \onehalf \log\rho_i\,.
2647: \end{math}
2648: Now define
2649: \begin{equation*}
2650: \Hat{\Omega}_{\cover{U}_X}[\log\rho] = \Omega[\log\rho]
2651: -\sqrt{-1}\omega_\rho\,.
2652: \end{equation*}
2653: We see that it yields the formula in Proposition~\ref{prop:2}. In
2654: particular we have that
2655: \begin{equation*}
2656: \left.\frac{d}{dt}\right\rvert_{t=0}
2657: \Hat{\Omega}_{\cover{U}_X}[\log\rho +t\sigma]
2658: \equiv \sigma ( c_1(\rho) -\sqrt{-1}\omega_\rho )\,,
2659: \end{equation*}
2660: where $\equiv$ means ``up to total coboundary.''
2661: \end{proof}
2662:
2663: \subsection{Determinant of cohomology}
2664: \label{sec:determ-cohom}
2665:
2666: Let again $L$ and $L'$ be two holomorphic line bundles with
2667: hermitian metrics $\rho$ and $\rho'$, respectively, on the
2668: compact Riemann surface $X$. Brylinski proves in
2669: \cite{bry:quillen} that the cup product of $L$ and $L'$ in
2670: hermitian holomorphic Deligne cohomology yields the (logarithm)
2671: of the metric $\norm{\cdot}$ on the Deligne pairing $\dual{L}{L'}$
2672: defined in \cite{MR89b:32038}.
2673:
2674: It follows, via Thm~\ref{thm:2} and the
2675: isomorphism~\ref{lemma:qi-dhh-bry} between our version of hermitian
2676: holomorphic cohomology and Brylinski's, that the class of
2677: $\Omega[\log\rho,\log\rho']$ is also equal to $\log
2678: \norm{\dual{L}{L'}}$. It is worthwhile to provide a direct proof of this
2679: fact starting from the explicit cocycle given in \eqref{eq:74}.
2680:
2681: First, we need to recall a few definitions from \cite{MR89b:32038}. A
2682: complex line $\dual{L}{L'}$ is assigned to the pair $(L, L')$ as
2683: follows. Let $D$ and $D'$ be divisors on $X$ corresponding to $L$ and
2684: $L'$, and assume they have disjoint supports. Consider two rational
2685: sections, $s$ and $s'$ such that $(s)=D$ and $(s')=D'$. To this datum
2686: one assigns a copy of the complex line generated by the symbol
2687: $\dual{s}{s'}$ subject to the relations:
2688: \begin{equation}
2689: \label{eq:76}
2690: \begin{aligned}
2691: \dual{fs}{s'} &= f(D') \dual{s}{s'}\\
2692: \dual{s}{gs'} &= g(D) \dual{s}{s'}
2693: \end{aligned}
2694: \end{equation}
2695: whenever $f$ is a rational function with divisor $(f)$ disjoint from
2696: $D'$, and similarly for $g$. The Weil reciprocity relation
2697: $f(\mathit{div}(g)) = g(\mathit{div}(f))$ (cf. ref.\ \cite{gh:alg_geom})
2698: for two rational functions $f$ and $g$ with disjoint divisors implies
2699: that the relations~\eqref{eq:76} are consistent and the complex line
2700: depends only on the pair $L$, $L'$. When the line bundles are equipped
2701: with hermitian metrics, generically denoted by $\norm{\cdot}$, the
2702: assignment\footnote{We write the square explicitly, whereas the symbol
2703: $\norm{\cdot}$ used in ref.~\cite{MR89b:32038} denotes the
2704: \emph{square} of the norm.}
2705: \begin{equation}
2706: \label{eq:77}
2707: \log\, \norm{\dual{s}{s'}}^2 = \frac{1}{\tate}
2708: \int_X \del\delb\log \norm{s}^2 \,\log\norm{s'}^2
2709: + \log\norm{s}^2[D'] +\log\norm{s'}^2[D]
2710: \end{equation}
2711: is compatible with the relations~\eqref{eq:76} and defines an hermitian
2712: metric on the complex line $\dual{L}{L'}$. In formula~\eqref{eq:77} the
2713: operator $\del\delb$ is to be computed in the sense of distributions.
2714:
2715: Having covered the main definitions, we can now state
2716: \begin{theorem}
2717: \label{thm:3}
2718: The cup product of $(L,\rho)$ with $(L',\rho')$ in hermitian
2719: holomorphic Deligne cohomology corresponds to the logarithm of the
2720: norm \eqref{eq:77} on the Deligne pairing $\dual{L}{L'}$. The
2721: proportionality factor is $-\pi\,\sqrt{-1}$.
2722: \end{theorem}
2723: \begin{proof}
2724: Let $D$ and $D'$ be divisors with disjoint support on $X$ corresponding
2725: to $L$ and $L'$, respectively. Using the same technique as in refs.
2726: \cite{bry:quillen,MR89b:32038}, consider two $C^\infty$ positive real
2727: functions $f_1$ and $f_2$ such that $f_1+f_2=1$ and $f_1$ (resp. $f_2$)
2728: vanishes in a neighborhood of the support of $D'$ (resp. $D$). Also, set
2729: $U_1 = X\setminus \mathrm{supp}(D')$ and $U_2 = X\setminus
2730: \mathrm{supp}(D)$. Thus $\{f_1,f_2\}$ is just a partition of unity
2731: subordinated to the cover $\cover{U}_X = \{ U_1, U_2\}$.
2732:
2733: The only two terms different from zero in the cocycle $\Theta$
2734: in~\eqref{eq:74} representing the class $[L,\rho]\cup [L',\rho']$ with
2735: respect to this cover are $\theta^{0}_i$ and $\theta^{1}_i$, with the
2736: \cech\ index $i \in \{ 1,2 \}$. Thus the class we are after is
2737: equivalently given by the integral
2738: \begin{equation}
2739: \label{eq:78}
2740: \int_X f_1\theta^{0}_1 +f_2\theta^{1}_2
2741: + df_2 \wedge \theta^{1}_{21} \,,
2742: \end{equation}
2743: which is arrived at by applying in the standard homotopy operator based
2744: on the partition of unity $\{f_1,f_2\}$: from
2745: $\deltacheck\theta^{1}_{12}=0$ we obtain that $\theta^{1}_{12}$ is the
2746: coboundary of the cochain $j\to \sum_{i=1,2} f_i\theta^{1}_{ij}$ and
2747: then we use $\theta^{0}_2-\theta^{0}_2=-d\theta^{1}_{12}$. Observe that
2748: the $2$-form in~\eqref{eq:78} is globally well defined over $X$:
2749: $\theta^{1}_{21}$ is defined only on $U_1\cap U_2$, but $df_2$ has
2750: support on $U_1\cap U_2$, so their wedge product is defined everywhere;
2751: similarly, $f_i\,\theta^{0}_i$ is everywhere defined thanks to the fact
2752: that $f_i$ has support in $U_i$, $i=1,2$.
2753:
2754: Consider rational sections $s$ and $s'$ of $L$ and $L'$ such that
2755: $\mathit{div}(s)=D$ and $\mathit{div}(s')=D'$ as above. With respect to
2756: the two-element cover $\cover{U}_X=\{U_1,U_2\}$, the section $s$
2757: corresponds to the pair $\{s_1,s_2\}$, and similarly for
2758: $s'=\{s'_1,s'_2\}$. Since $\mathrm{supp}(D)$ is contained in $U_1$ but
2759: not in $U_2$, and the other way around for $D'$, it follows that $s_2$
2760: and $s'_1$ are actually invertible functions over their respective
2761: domains. Following \cite{lang:arakelov}, we can assume that $s$ and $s'$
2762: are in fact the rational section $\mathbf{1}$, so that $s_2=1$ and
2763: $s'_1=1$, and therefore:
2764: \begin{gather*}
2765: \abs{s}^2_\rho =
2766: \begin{cases}
2767: \log \rho_1 + \log \abs{s_1}^2 & \text{on } U_1\,,\\
2768: \log \rho_2 & \text{on } U_2\,,
2769: \end{cases}\\ \intertext{and}
2770: \abs{s'}^2_{\rho'} =
2771: \begin{cases}
2772: \log\, \rho'_1 & \text{on } U_1\,,\\
2773: \log \rho'_2 + \log\, \abs{s'_2}^2 & \text{on } U_2\,.
2774: \end{cases}\\
2775: \end{gather*}
2776: Furthermore,
2777: \begin{math}
2778: \log \, \abs{g'_{21}} = \log \, \abs{s'_2}
2779: \end{math}
2780: on $U_1\cap U_2$.
2781: Let us denote by $\norm{\cdot}=\abs{\cdot}_\rho= \abs{\cdot}_{\rho'}$
2782: for simplicity. Using the relevant entries from~\eqref{eq:74} we have:
2783: \begin{equation*}
2784: \begin{gathered}
2785: \theta^{0}_1 = c_1(\rho)\,\log \norm{s'}\,,\\
2786: \theta^{0}_2 = \delb\del \log \rho_2 \; \onehalf \log \rho'_2
2787: = dd^c \,\log\,\norm{s} \; \onehalf \log\rho'_2\,.
2788: \end{gathered}
2789: \end{equation*}
2790: An elementary integration by parts leads to:
2791: \begin{equation*}
2792: \begin{split}
2793: \int_X f_2 \theta^{0}_2
2794: &= \int_X f_2 \log\,\norm{s}\, c_1(\rho') \\
2795: &+ \int_X \log\,\norm{s} \, df_2 \wedge \onehalf d^c \log\rho'_2
2796: - \int_X \onehalf \log\rho'_2 \, df_2 \wedge d^c \log\,\norm{s}\,.
2797: \end{split}
2798: \end{equation*}
2799: On the other hand, we have
2800: \begin{equation*}
2801: \int_X df_2 \wedge \theta^{1}_{12} =
2802: \int_X \log\,\norm{s}\; df_2\wedge d^c \log\,\abs{s'_2}
2803: -\int_X \log\,\abs{s'_2} \; df_2\wedge d^c \log\,\norm{s}\,,
2804: \end{equation*}
2805: and putting all terms together we obtain
2806: \begin{equation}
2807: \label{eq:79}
2808: \begin{split}
2809: \int_X f_1\theta^{0}_1 +f_2\theta^{1}_2
2810: + df_2 \wedge \theta^{1}_{21}
2811: &=\int_X f_1\,c_1(\rho)\,\log \norm{s'}
2812: +\int_X f_2 \log\,\norm{s}\, c_1(\rho') \\
2813: &+ \int_X \log\,\norm{s}\; df_2\wedge d^c \log\,\norm{s'}
2814: +\int_X \log\,\norm{s'} \; df_1\wedge d^c \log\,\norm{s}
2815: \end{split}
2816: \end{equation}
2817: which, if expressed in terms of the squares of the norms, is (up
2818: to a factor) the logarithm of $\norm{\dual{s}{s'}}$, as it is
2819: found in \cite[formula 6.5.1]{MR89b:32038}. This version is due
2820: to O. Gabber. Via the Poincar\'e-Lelong lemma (see,
2821: e.g.\ \cite{gh:alg_geom})
2822: \begin{equation*}
2823: \delb\del \log\,\norm{s} = c_1(\rho) + \tate\,[D]\,,
2824: \end{equation*}
2825: where $[D]$ is the delta-current supported at the divisor of $s$,
2826: and similarly for $s'$, formula~\eqref{eq:79} can be recast into:
2827: \begin{equation}
2828: \label{eq:80}
2829: \begin{split}
2830: \int_X f_1\theta^{0}_1 +f_2\theta^{1}_2
2831: + df_2 \wedge \theta^{1}_{21} &=
2832: \int_X dd^c \log\,\norm{s}\; \log\,\norm{s'}
2833: -\tate\,\log\,\norm{s} [D']
2834: -\tate\,\log\,\norm{s'} [D] \\
2835: &= -\tate\; \log\,\norm{\dual{s}{s'}}
2836: \end{split}
2837: \end{equation}
2838: which is what we wanted to show.
2839: \end{proof}
2840: This allows us to recast the Liouville functional for conformal metrics
2841: on $X$ in the following form.
2842: \begin{corollary}
2843: \label{corollary:2}
2844: The exponential of the Liouville functional defines an hermitian
2845: metric on the complex line $\dual{T_X}{T_X}$, namely for a conformal
2846: metric $\rho\in \cm{X}$ we have:
2847: \begin{equation*}
2848: \exp S[\log\rho] = \norm{\dual{T_X}{T_X}}\, \exp
2849: \tfrac{1}{2\pi} A_X(\rho)\,,
2850: \end{equation*}
2851: where $A_X(\rho)$ is the area of $X$ with respect to $\rho$.
2852: \end{corollary}
2853: \begin{remark}
2854: The above corollary justifies the choice made in
2855: Definition~\ref{def:2} for the various factors $\tate$.
2856: \end{remark}
2857: A similar result has been obtained in ref.\ \cite{zograf1990} by
2858: considering the Liouville action functional defined on the Schottky
2859: space, and in fact the results in loc.\ cit. are formulated in terms of
2860: a Schottky \emph{family.}
2861:
2862: Indeed, the statement in Corollary~\ref{corollary:2} can be immediately
2863: reformulated for a family $\pi \colon X\rightarrow S$ with base
2864: parameter space $S$ by considering the relative holomorphic tangent line
2865: bundle $T_{X/S}$ with an hermitian fiber metric $\rho$. (Thus $\rho_s\in
2866: \cm{X_s}$ for every fiber $X_s\,, s\in S$.) Notice that the fiber metric
2867: $\rho$ needs not be critical (i.e.\ satisfying the fiberwise constant
2868: negative curvature condition).
2869:
2870:
2871: \section{Conclusions and outlook}
2872: \label{sec:conclusions-outlook}
2873:
2874: In this paper we have provided a geometric description of the
2875: construction of the Liouville action functional for constant
2876: negative curvature metrics on compact Riemann surfaces of genus
2877: $g\geq 2$.
2878:
2879: Our approach was to construct a Hermitian holomorphic Deligne
2880: cohomology class as the cup square of the metrized holomorphic
2881: tangent bundle $T_X$, and then show at the level of cocycles that
2882: (modulo an area term) this construction agrees with those
2883: in~\cite{math.CV/0204318} and earlier works.
2884:
2885: Furthermore, our construction leads to the identification of the
2886: class corresponding to the Liouville action with the determinant
2887: of cohomology construction. Hence it could serve as an
2888: alternative construction of the latter in terms of different
2889: choices of cocycles.
2890:
2891: One of the most important properties of the Liouville functional
2892: from the works~\cite{zogtak1987-1,zogtak1987-2,zograf1990} is the
2893: link with the Weil-Petersson geometry of the Teichm\"uller space.
2894: From this point of view, an analysis of the behavior of the
2895: Liouville action for families of Riemann surfaces is crucial. In
2896: particular, a delicate analytic computation of the variation of
2897: the Action with respect to the moduli was carried out in
2898: the afore mentioned works to establish the link with the
2899: Weil-Petersson metric.
2900:
2901: We have not pursued these matters in the present work limiting
2902: ourselves to establish the existence of a class for a metrized
2903: bundle on the base of a family $\pi\colon X\to S$. A more precise
2904: analysis would require, not only a more explicit description of
2905: the map $\int_\pi$ (also advocated in~\cite{brymcl:deg4_II}), but
2906: a full computation of the Leray sequence associated to the family
2907: $\pi$. This is instrumental in defining \emph{relative} Hermitian
2908: holomorphic classes, and in analyzing their variation with
2909: respect to base parameters.
2910:
2911: We hope to pursue this and other problems related to the
2912: extension of the present work to singular metrics in a different
2913: publication.
2914:
2915:
2916: \appendix
2917:
2918: \section{Cones}
2919: \label{sec:Cones}
2920:
2921: In the main body of the paper we have used iterated cones to
2922: define the hermitian holomorphic Deligne complexes. One technical
2923: problem one has to face concerns the homotopy (graded)
2924: commutativity of the modified \bei\ product defined in
2925: eq.~\eqref{eq:5}. A problem arises because the factors in the
2926: cones are cones themselves and therefore they have multiplication
2927: structures which are graded commutative up to homotopy to begin
2928: with. We want to show that even in this situation the final
2929: resulting product on cones is again homotopy graded commutative.
2930: This ensures that on cohomology the product will be genuinely
2931: graded commutative, so that in particular hermitian holomorphic
2932: Deligne cohomology as defined in
2933: section~\ref{sec:herm-holom-deligne} has the correct product
2934: structure.
2935:
2936:
2937: \subsection{Cones and homotopies}
2938: \label{sec:cones-homotopies}
2939:
2940: We consider the following situation. For $i=1,2,3$ we have maps of
2941: complexes:
2942: \begin{math}
2943: f_i \colon A^\bullet_i \rightarrow B^\bullet_i\,,
2944: \end{math}
2945: and for $i<j$ maps
2946: \begin{math}
2947: a_{ji} \colon A^\bullet_{i} \rightarrow A^\bullet_j
2948: \end{math}
2949: and
2950: \begin{math}
2951: b_{ji} \colon B^\bullet_{i} \rightarrow B^\bullet_j\,.
2952: \end{math}
2953: Also, let
2954: \begin{math}
2955: C^\bullet(f_i) =\cone (f_i:A^\bullet_i\rightarrow B^\bullet_i) \,,
2956: \end{math}
2957: for $i=1,2,3$.
2958:
2959: First, consider the homotopy commutative diagram:
2960: \begin{equation}
2961: \label{eq:81}\vcenter{%
2962: \xymatrix{%
2963: A^\bullet_j \ar[r]^{f_j} \ar[d]_{a_{ij}} & B^\bullet_j
2964: \ar[d]^{b_{ij}} \ar@{=>}[dl]^{s_{ij}} \\
2965: A^\bullet_i \ar[r]_{f_i} & B^\bullet_i}}
2966: \end{equation}
2967: where
2968: \begin{math}
2969: s_{ij} \colon A^\bullet_j \rightarrow B^{\bullet -1}_i
2970: \end{math}
2971: is the homotopy map of complexes:
2972: \begin{equation*}
2973: f_ia_{ij} -b_{ij}f_j = d\,s_{ij} + s_{ij}d\,.
2974: \end{equation*}
2975: An immediate verification yields
2976: \begin{lemma}
2977: \label{lemma:cone-map}
2978: The diagram~\eqref{eq:81} can be extended to
2979: \begin{equation*}
2980: \xymatrix{%
2981: A^\bullet_j \ar[r]^{f_j} \ar[d]_{a_{ij}} &
2982: B^\bullet_j \ar[r] \ar[d]^{b_{ij}} \ar@{=>}[dl]^{s_{ij}} &
2983: C^\bullet (f_j) \ar[d]^{c_{ij}} \ar[r]^{[1]} &
2984: A^\bullet_j \ar[d]^{a_{ij}}\\
2985: A^\bullet_i \ar[r]_{f_i} & B^\bullet_i \ar[r] &
2986: C^\bullet (f_i) \ar[r]_{[1]} & A^\bullet_i
2987: }
2988: \end{equation*}
2989: where the map $c_{ij}$ is given by
2990: \begin{math}
2991: \bigl(
2992: \begin{smallmatrix}
2993: a_{ij} & 0 \\ -s_{ij} & b_{ij}
2994: \end{smallmatrix}
2995: \bigr)
2996: \end{math}
2997: and the squares containing the cones are in fact strictly commutative.
2998: \end{lemma}
2999: \begin{remark}
3000: This lemma is nothing other than the statement that any
3001: homotopy commutative diagram of the form~\eqref{eq:81} in the
3002: category of complexes in an abelian category can be extended to
3003: a (homotopy) commutative diagram of distinguished triangles,
3004: that is, one of the axioms defining a triangulated category,
3005: see, e.g. \cite{weibel_hom_alg}.
3006: \end{remark}
3007: For $k<j<i$ consider the homotopy commutative triangle
3008: \begin{equation*}
3009: \xymatrix{%
3010: & A^\bullet_j \ar[dr]^{a_{ij}} \ar@{=>}[d]^(.6){\alpha_{ijk}}& \\
3011: A^\bullet_k \ar[0,2]_{a_{ik}} \ar[ur]^{a_{jk}} & &
3012: A^\bullet_i}
3013: \end{equation*}
3014: where
3015: \begin{math}
3016: a_{ik} -a_{ij}a_{jk} = d\alpha_{ijk} + \alpha_{ijk} d\,,
3017: \end{math}
3018: and similarly for the complexes $B^\bullet_i$, with a
3019: corresponding homotopy $\beta_{ijk}$. Thus
3020: \begin{math}
3021: \alpha_{ijk} \colon
3022: A^\bullet_k \rightarrow A^{\bullet -1}_i
3023: \end{math}
3024: and
3025: \begin{math}
3026: \beta_{ijk} \colon
3027: B^\bullet_k \rightarrow B^{\bullet -1}_i\,.
3028: \end{math}
3029: Now consider the diagram:
3030: \begin{equation}
3031: \label{eq:82}\vcenter{%
3032: \xymatrix@!{%
3033: A^\bullet_k \ar[0,2]^{a_{jk}} \ar[dr]_(.7){a_{ik}} \ar[2,0]_{f_k} & &
3034: A^\bullet_j \ar[dl]^{a_{ij}} \ar[2,0]^{f_j}
3035: \ar@{=>}[];[dl]+/ul 5ex/\\
3036: & A^\bullet_i \ar[2,0]^(.4){f_i} & \\
3037: B^\bullet_k \ar'[r]^(.7){b_{jk}}[rr] \ar[dr]_{b_{jk}}
3038: \ar@{=>}[ur] \ar@2{-->}[-1,0];[-2,1] & &
3039: B^\bullet_j \ar[dl]^{b_{ij}} \ar@{=>}[ul]
3040: \ar@{=>}[];[dl]+/ul 5ex/ \\
3041: & B^\bullet_i &
3042: }}
3043: \end{equation}
3044: The faces in~\eqref{eq:82} are homotopy commutative, however we assume
3045: that composing the \emph{faces} is strictly commutative, namely the two
3046: possible homotopies
3047: \begin{math}
3048: b_{ij}\,b_{jk}\,f_k \Longrightarrow f_i\,a_{ik}
3049: \end{math}
3050: must be equal. Concretely, this corresponds to the relation
3051: \begin{equation}
3052: \label{eq:83}
3053: s_{ik} + \beta_{ijk}\,f_{k} = f_i\,\alpha_{ijk}
3054: +s_{ij}\,a_{jk} +b_{ij}\,s_{jk}\,.
3055: \end{equation}
3056: We have:
3057: \begin{lemma}
3058: \label{lemma:cone-homotopy}
3059: The map
3060: \begin{equation*}
3061: \bigl(
3062: \begin{smallmatrix}
3063: -\alpha_{ijk} & 0 \\ 0 & \beta_{ijk}
3064: \end{smallmatrix}
3065: \bigr) \colon C^\bullet (f_k) \longrightarrow
3066: C^{\bullet -1}(f_i)
3067: \end{equation*}
3068: realizes the homotopy
3069: \begin{equation*}
3070: \xymatrix{%
3071: & C^\bullet (f_j) \ar[dr]^{c_{ij}} \ar@{=>}[d]& \\
3072: C^\bullet (f_k) \ar[ur]^{c_{jk}} \ar[rr]_{c_{ik}} & &
3073: C^\bullet (f_i)}
3074: \end{equation*}
3075: \end{lemma}
3076: \begin{proof}
3077: It is an elementary calculation based on writing
3078: \begin{math}
3079: c_{ik} - c_{ik}\,c_{jk}
3080: \end{math}
3081: explicitly via the matrix representation given in
3082: Lemma~\ref{lemma:cone-map} and using eq.~\eqref{eq:83}.
3083: \end{proof}
3084:
3085:
3086: \subsection{Applications}
3087: \label{sec:applications}
3088:
3089: Consider the same setup as in section~\ref{sec:cup-products-cones}, with
3090: the same complexes $X^\bullet_i$, etc., and diagrams:
3091: \begin{equation*}
3092: \mathcal{D}_i \eqdef
3093: X^\bullet_i \overset{f_i}{\longrightarrow} Z^\bullet_i
3094: \overset{g_i}{\longleftarrow} Y^\bullet_i
3095: \end{equation*}
3096: from which we construct the cones
3097: \begin{equation*}
3098: C(\mathcal{D}_i) =
3099: \cone (X^\bullet_i\oplus Y^\bullet_i
3100: \xrightarrow{f_i-g_i} Z^\bullet_i)[-1]\,,
3101: \quad i=1,2,3\,.
3102: \end{equation*}
3103: Moreover, following ref.\ \cite{bei:hodge_coho}, define
3104: $\mathcal{D}_i\otimes \mathcal{D}_j$ by taking the tensor product
3105: component-wise. Thus
3106: \begin{equation*}
3107: \mathcal{D}_1\otimes \mathcal{D}_2 =
3108: X^\bullet_1 \otimes X^\bullet_2
3109: \xrightarrow{f_1\otimes f_2}
3110: Z^\bullet_1\otimes Z^\bullet_2
3111: \xleftarrow{g_1\otimes g_2}
3112: Y^\bullet_1\otimes Y^\bullet_2\,.
3113: \end{equation*}
3114: Assuming as in section~\ref{sec:cup-products-cones} that the product
3115: maps are compatible with the $f_i$, etc., the diagram
3116: \begin{equation*}
3117: \mathcal{D}_1\otimes \mathcal{D}_2 \rightarrow \mathcal{D}_3
3118: \end{equation*}
3119: is of the same type as~\eqref{eq:81}, and therefore
3120: lemma~\ref{lemma:cone-map} implies lemma~\ref{lemma:mod-cup-prod}.
3121:
3122: Now, let the multiplication maps
3123: \begin{math}
3124: X^\bullet_1 \otimes X^\bullet_2 \rightarrow X^\bullet_3
3125: \end{math}
3126: be graded commutative up to homotopy and similarly for the
3127: $Y^\bullet_i$ and the $Z^\bullet_i$. We are interested in the
3128: commutativity properties of multiplication map given by the \bei\
3129: product~\eqref{eq:5}.
3130: \begin{proposition}
3131: \label{prop:4}
3132: The multiplication map
3133: \begin{math}
3134: \cup_\alpha\colon C(\mathcal{D}_1)\otimes C(\mathcal{D}_2)
3135: \longrightarrow C(\mathcal{D}_3)
3136: \end{math}
3137: given by~\eqref{eq:5} is homotopy graded commutative.
3138: \end{proposition}
3139: \begin{proof}
3140: The permutation operation on tensor products induces the diagram
3141: \begin{equation*}
3142: \xymatrix{%
3143: {\mathcal{D}_1}\otimes {\mathcal{D}_2} \ar[dr] \ar[rr] & &
3144: {\mathcal{D}_2}\otimes {\mathcal{D}_1} \ar[dl] \\
3145: & {\mathcal{D}}_3 & }
3146: \end{equation*}
3147: which is of type~\eqref{eq:82} and we can apply
3148: lemma~\ref{lemma:cone-homotopy}.
3149: \end{proof}
3150: It follows from the proposition that the cohomology inherits a well
3151: defined graded commutative product. This in particular applies to the
3152: definition of hermitian holomorphic Deligne cohomology that uses the
3153: cone~\eqref{eq:22}. Therefore we conclude that the cup
3154: product~\eqref{eq:27} is graded commutative, as wanted.
3155:
3156: \section{Conventions on Kleinian groups and fractional linear
3157: transformations}
3158: \label{sec:kleinian-groups}
3159:
3160: As a reference the reader can consult, among many others, the
3161: book \cite{MR99g:30055}. Let $\Gamma$ be a finitely generated
3162: purely loxodromic non-elementary Kleinian group of the second
3163: kind, so there is a nonempty discontinuity region
3164: $U_\Gamma\subset \pione$. The limit set is $L_\Gamma = \pione
3165: \setminus U_\Gamma$. According to Ahlfors' finiteness theorem the
3166: quotient $X=U_\Gamma/\Gamma$ is a finite union of analytically
3167: finite Riemann surfaces. Thus:
3168: \begin{equation*}
3169: U_\Gamma/\Gamma = U_1/\Gamma_1 \sqcup \dots \sqcup U_n/\Gamma_n\,,
3170: \end{equation*}
3171: where $U_1,\dots,U_n$ are the inequivalent components of
3172: $U_\Gamma$ and $\Gamma_1, \dots, \Gamma_n$ their stabilizers. By
3173: way of example, a Schottky group has just one component, whereas
3174: a Fuchsian or quasi-Fuchsian group will have exactly two
3175: components.
3176:
3177: We consider the map $U_\Gamma \to X$ as an \'etale cover, and
3178: \cech\ cohomology with respect to it translates into group
3179: cohomology for the group $\Gamma$, where the coefficient modules
3180: are sections over $U_\Gamma$ of the relevant sheaves. The group
3181: action is by pull-back. According to the conventions of ref.\
3182: \cite{aldtak2000} we will write the coboundary operation as:
3183: \begin{equation}
3184: \label{eq:58}
3185: (\deltacheck c)_{\gamma_1,\dots,\gamma_n} =
3186: c_{\gamma_2,\dots,\gamma_n}
3187: +\sum_{i=1}^{n-1}(-1)^i
3188: c_{\gamma_1,\dots,\gamma_i\gamma_{i+1},\dots,\gamma_n}
3189: +(-1)^n (c_{\gamma_1,\dots,\gamma_{n-1}})\cdot {\gamma_n}\,,
3190: \end{equation}
3191: where $c$ is an $n$-cochain with values in some \emph{right}
3192: $\Gamma$-module $A$. The expression~\eqref{eq:58} is the \cech\
3193: coboundary applied to the nerve of the cover $U_\Gamma \to X$.
3194:
3195: We assume that $\Gamma$ is normalized, namely the point $\infty$
3196: belongs to the limit set $L_\Gamma$. If $\gamma\in \Gamma$
3197: corresponds to the fractional linear transformation:
3198: \begin{equation*}
3199: \pione \ni z \longmapsto \gamma (z) = \frac{az +b}{cz +d}\,,
3200: \end{equation*}
3201: we have
3202: \begin{equation}
3203: \label{eq:59}
3204: \gamma' (z) = \frac{\det \gamma}{c^2(z -z_\gamma)^2}\,,
3205: \end{equation}
3206: where $z_\gamma = -\frac{d}{c} \equiv \gamma^{-1}(\infty)$. Set
3207: \begin{equation*}
3208: c(\gamma) \equiv c_\gamma \eqdef \frac{\det \gamma}{c^2}\,.
3209: \end{equation*}
3210: The following properties are easily verified. If $z_0$ and $z_\infty$
3211: are the attracting and repelling fixed points, respectively, then
3212: \begin{equation*}
3213: c_\gamma =
3214: \frac{(z_0 -z_\infty)^2\lambda_\gamma}{(1-\lambda_\gamma)^2} \,,
3215: \end{equation*}
3216: where $\lambda_\gamma$ is the dilating factor. For $\gamma_i$ and
3217: $\gamma_j$ two elements of $\Gamma$, denote:
3218: \begin{equation*}
3219: z_i = \gamma^{-1}(\infty)\,, \quad
3220: z_{ij} = (\gamma_i\gamma_j)^{-1}(\infty) =
3221: \gamma^{-1}_j (z_i)\,, \quad c_i = c_{\gamma_i}\,, \quad
3222: c_{ij} = c_{\gamma_i\gamma_j}\,.
3223: \end{equation*}
3224: We have the following relation:
3225: \begin{equation}
3226: \label{eq:60}
3227: c_{\gamma_1\gamma_2}
3228: = \frac{c_{\gamma_1}}{c_{\gamma_2}}\, (z_{12} -z_2)^2\,.
3229: \end{equation}
3230: Finally, given four points $z_1,z_2, z_3, z_4\in \pione$, we define
3231: their cross-ratio by:
3232: \begin{equation}
3233: \label{eq:61}
3234: \CR{z_1}{z_2}{z_3}{z_4}
3235: = \frac{(z_1-z_2)(z_3-z_4)}{(z_1-z_4)(z_3-z_2)}\,.
3236: \end{equation}
3237:
3238: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3239: \bibliography{general}
3240: \bibliographystyle{hamsplain}
3241: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3242:
3243: \end{document}
3244:
3245:
3246: % LocalWords: holomorphic uniformization Liouville Ettore Aldrovandi Deligne
3247: % LocalWords: Brylinski Quillen Kleinian Wigner Poincar Polyakov's variational
3248: % LocalWords: Teichm uller homological ref Takhtajan cocycle etale Fuchsian
3249: % LocalWords: loc Krasnov diffeomorphisms refs cohomological Friedan Shenker
3250: % LocalWords: dilogarithms regularized submanifold imbedding
3251: