math0409214/KM2.tex
1: %Last changed by DK on 8 September 2004; minor changes on 14 January 2006
2: 
3: \documentclass[12pt]{amsart}
4: \usepackage{epsfig,amscd,amssymb,times}
5: %\usepackage{backref}
6: %\usepackage{graphics}
7: \setlength{\headheight}{6.15pt}
8: 
9: \newtheorem{theorem}{Theorem}%[section]
10: \newtheorem{lemma}[theorem]{Lemma}
11: \newtheorem{proposition}[theorem]{Proposition}
12: \newtheorem{corollary}[theorem]{Corollary}
13: \theoremstyle{definition}
14: \newtheorem{definition}[theorem]{Definition}
15: \newtheorem{problem}[theorem]{Problem}
16: \newtheorem{example}[theorem]{Example}
17: \newtheorem{conjecture}[theorem]{Conjecture}
18: \theoremstyle{remark}
19: \newtheorem{remark}[theorem]{Remark}
20: %\numberwithin{figure}{section}
21: %\numberwithin{table}{section}
22: 
23: \def\varph{{\varphi}}
24: \def\ra{{\rightarrow}}
25: \def\lra{{\longrightarrow}}
26: \def\om{{\omega}}
27: \def\Om{{\Omega}}
28: \def\la{{\lambda}}
29: \def\al{{\alpha}}
30: 
31: \def\La{{\Lambda}}
32: \def\ga{{\gamma}}
33: \def\Ga{{\Gamma}}
34: \def\vGa{{\varGamma}}
35: \def\bC{{\mathbb C}}
36: \def\bZ{{\mathbb Z}}
37: \def\bQ{{\mathbb Q}}
38: \def\bR{{\mathbb R}}
39: \def\inv{{^{-1}}}
40: \def\Sg{{\Sigma_g}}
41: \def\Mg{{\mathcal M}_g}
42: \def\Msg{{\mathcal M}_{g,*}}
43: \def\M1g{{\mathcal M}_{g,1}}
44: \def\I1g{{\mathcal I}_{g,1}}
45: 
46: \newcommand\Hom{\operatorname{Hom}}
47: \newcommand\End{\operatorname{End}}
48: \newcommand\Aut{\operatorname{Aut}}
49: \newcommand\Der{\operatorname{Der}}
50: \newcommand\Out{\operatorname{Out}}
51: \newcommand\vol{\operatorname{vol}}
52: \newcommand\sgn{\operatorname{sgn}}
53: \newcommand\Flux{\operatorname{Flux}}
54: \newcommand\Symp{\operatorname{Symp}}
55: \newcommand\Ham{\operatorname{Ham}}
56: \newcommand\eFlux{\operatorname{\widetilde{Flux}}}
57: \newcommand\Cal{\operatorname{Cal}}
58: \newcommand\Ker{\operatorname{Ker}}
59: \newcommand\Cok{\operatorname{Cok}}
60: \newcommand\im{\operatorname{Im}}
61: \newcommand\Int{\operatorname{Int}}
62: \newcommand\sign{\operatorname{sign}}
63: \newcommand\BS{\operatorname{BSymp}}
64: \newcommand\ES{\operatorname{ESymp}}
65: \newcommand\BD{\operatorname{BDiff}}
66: \newcommand\BG{\operatorname{B\bar\Gamma}}
67: \newcommand\Diff{\operatorname{Diff}}
68: \newcommand\id{\operatorname{id}}
69: \catcode`\@=\active % make @ other character ( is default)
70: 
71: \begin{document}
72: \title[Characteristic classes of foliated surface bundles]
73: {Characteristic classes of foliated surface bundles with area-preserving 
74: holonomy}
75: \author{D.~Kotschick}
76: \address{Mathematisches Institut, Ludwig-Maxi\-mi\-lians-Universit\"at 
77: M\"unchen,
78: Theresienstr.~39, 80333 M\"unchen, Germany}
79: \email{dieter{\char'100}member.ams.org}
80: \author{S.~Morita}
81: \address{Department of Mathematical Sciences\\
82: University of Tokyo \\Komaba, Tokyo 153-8914\\
83: Japan}
84: \email{morita{\char'100}ms.u-tokyo.ac.jp}
85: 
86: \keywords{symplectomorphism, area-preserving diffeomorphism,
87: foliated surface bundle, Hamiltonian symplectomorphism}
88: \thanks{The first author is grateful to the {\sl Deutsche 
89: Forschungsgemeinschaft}
90: for support of this work. The second author is partially supported by JSPS 
91: Grant
92: 16204005}
93: \subjclass{Primary 57R17, 57R50, 57M99; secondary 57R50, 58H10}
94: 
95: \begin{abstract}
96: Making use of the {\it extended} flux homomorphism defined in~\cite{KM03}
97: on the group $\Symp\Sg$ of symplectomorphisms of a closed oriented surface
98: $\Sg$ of genus $g\geq 2$, we introduce new characteristic classes of foliated
99: surface bundles with symplectic, equivalently area-preserving, total holonomy.
100: These characteristic classes are stable with respect to $g$ and we show that
101: they are highly non-trivial. We also prove that the second homology of
102: the group $\Ham\Sg$ of Hamiltonian symplectomorphisms of $\Sg$, equipped with
103: the discrete topology, is very large for all $g\geq 2$.
104: \end{abstract}
105: 
106: \maketitle
107: 
108: \section{Introduction}\label{s:intro}
109: 
110: In this paper we study the homology of symplectomorphism groups of
111: surfaces considered as discrete groups. We shall prove that certain
112: homology groups are highly non-trivial by constructing characteristic
113: classes of foliated surface bundles with area-preserving holonomy, and
114: proving non-vanishing results for them.
115: 
116: Let $\Sg$ be a closed oriented surface of genus $g\geq 2$, and $\Diff_+\Sg$
117: its group of orientation preserving selfdiffeomorphisms. We fix an area
118: form $\omega$ on $\Sg$, which, for dimension reasons, we can also think of
119: as a symplectic form. We denote by $\Symp\Sg$ the subgroup of $\Diff_+\Sg$
120: preserving the form $\om$. The classifying space $\BS ^\delta\Sg$ for the group
121: $\Symp\Sg$ with the {\it discrete} topology is an Eilenberg-MacLane space
122: $K(\Symp^\delta\Sg,1)$ which classifies foliated $\Sg$-bundles with
123: area-preserving total holonomy groups.
124: 
125: Our construction of characteristic classes proceeds as follows.
126: Let $\Symp_0\Sg$ be the identity component of $\Symp\Sg$.
127: A well-known theorem of Moser~\cite{Moser} concerning volume-preserving
128: diffeomorphisms implies that the quotient $\Symp\Sg/\Symp_0\Sg$ can be
129: naturally identified with the mapping class group $\Mg$, so that
130: we have an extension
131: $$
132: 1\lra\Symp_0\Sg\lra\Symp\Sg\overset{p}{\lra}\Mg\lra 1 \ .
133: $$
134: There is a surjective homomorphism $\Flux\colon\Symp_0\Sg\ra H^1(\Sg;\bR)$,
135: called the flux homomorphism. In~\cite{KM03} we proved that this homomorphism
136: can be extended to a crossed homomorphism
137: $$
138: \widetilde\Flux\colon\Symp\Sg\lra H^1(\Sg;\bR) \ ,
139: $$
140: which we call the {\it extended} flux homomorphism. This extension is 
141: essentially
142: unique in the sense that the associated cohomology class
143: $$
144: [\widetilde\Flux]\in H^1(\Symp^\delta\Sg;H^1(\Sg;\bR))
145: $$
146: with twisted coefficients is uniquely defined. Now we consider the powers
147: $$
148: [\widetilde\Flux]^k\in H^k(\Symp^\delta\Sg;H^1(\Sg;\bR)^{\otimes k})\quad
149: (k=2,3,\cdots) \ ,
150: $$
151: and apply $\Mg$-invariant homomorphisms
152: $$
153: \la\colon H^1(\Sg;\bR)^{\otimes k}\lra \bR
154: $$
155: to obtain cohomology classes
156: $$
157: \la([\widetilde\Flux]^k)\in H^{k}(\BS^\delta\Sg;\bR)
158: $$
159: with constant coefficients.
160: The usual cup-product pairing $H^1(\Sg;\bR)^{\otimes 2}\ra\bR$
161: is the main example of such a homomorphism $\la$.
162: 
163: This method of constructing constant cohomology classes out of twisted ones
164: was already used in~\cite{Morita96} in the case of the mapping class group,
165: where the Torelli group (respectively the Johnson homomorphism) played the
166: role of $\Symp_0\Sg$ (respectively of the flux homomorphism) here. In that
167: case, it was proved in {\it loc.~cit.}~that all the Mumford--Morita--Miller
168: classes can be obtained in this way. The precise formulae were given
169: in~\cite{KM96,KM01}, with the important implication that no other classes
170: appear. In our context here, we can go further by enhancing the coefficients
171: $\bR$ to associated $\bQ$-vector spaces which appear as the targets of various
172: multiples of the {\it discontinuous} cup-product pairing
173: $$
174: H^1(\Sg;\bR)\otimes_\bZ H^1(\Sg;\bR)\lra S^2_\bQ\bR \ ,
175: $$
176: where $S^2_\bQ\bR$ denotes the second symmetric power of $\bR$ over
177: $\bQ$, see Section~\ref{s:main} for the details. In this way, we obtain
178: many new characteristic classes in
179: $$
180: H^*(\Symp^\delta\Sg;S^*(S^2_\bQ\bR)) \ ,
181: $$
182: where
183: $$
184: S^*(S^2_\bQ\bR)=\bigoplus_{k=1}^\infty S^k(S^2_\bQ\bR)
185: $$
186: denotes the symmetric algebra of $S^2_\bQ\bR$.
187: On the other hand, we proved in~\cite{KM03} that any power
188: $$
189: e_1^k\in H^{2k}(\Symp^\delta\Sg;\bQ)
190: $$
191: of the first Mumford--Morita--Miller class $e_1$ is non-trivial for
192: $k\leq \frac{g}{3}$. Now we can consider the cup products of $e_1^k$ with
193: the new characteristic classes defined above. The main purpose of the present
194: paper is to prove that these characteristic classes are all non-trivial in a
195: suitable stable range.
196: 
197: The contents of this paper is as follows. In Section~\ref{s:main} precise
198: statements of the main results are given. In Section~\ref{s:trans} we study,
199: in detail, the transverse symplectic class of foliated $\Sg$-bundles with
200: area-preserving total holonomy groups. In Section~\ref{s:open} we construct
201: two kinds of {\it extended} flux homomorphisms for open surfaces
202: $\Sigma_g^0=\Sg\setminus D^2$. We compare these extended flux homomorphisms
203: with the one obtained in the case of closed surfaces. This is used in
204: Section~\ref{s:secondopen} to generalize our result on the second homology of
205: the symplectomorphism group to the case of open surfaces. Sections~\ref{s:open}
206: and~\ref{s:secondopen} are the heart of this paper. Then in Section~\ref{s:proofs} we use the results of the previous sections to show the non-t 
207: riviality of cup
208: products of various characteristic classes, thus yielding proofs of the main
209: results about the homology of symplectomorphism groups as discrete groups. In
210: the final Section~\ref{s:misc} we give definitions of yet more characteristic
211: classes, other than the ones given in Section~\ref{s:main}. We propose several
212: conjectures and problems about them.
213: 
214: \section{Statement of the main results}\label{s:main}
215: 
216: Consider the usual cup-product pairing
217: $$
218: \iota\colon H^1(\Sg;\bR)\otimes H^1(\Sg;\bR)\lra \bR
219: $$
220: in cohomology, dual to the intersection pairing in homology. For simplicity,
221: we denote $\iota(u,v)$ by $u\cdot v$, where $u,v\in H^1(\Sg;\bR)$.
222: We first lift this pairing as follows.
223: 
224: As in Section~\ref{s:intro}, let $S^2_{\bQ}\bR$ denote the second symmetric
225: power of $\bR$ over $\bQ$. In other words, this is a vector space over $\bQ$,
226: consisting of the homogeneous polynomials of degree two generated by the
227: elements of $\bR$ considered as a vector space over $\bQ$. For each element
228: $a\in\bR$, we denote by $\hat a$ the corresponding element in $S^1_\bQ\bR$.
229: Thus any element in $S^2_\bQ\bR$ can be expressed as a finite sum
230: $$
231: \hat a_1\hat b_1+\cdots +\hat a_k\hat b_k
232: $$
233: with $a_i, b_i\in\bR$. We have a natural projection
234: $$
235: S^2_\bQ\bR\lra \bR
236: $$
237: given by the correspondence $\hat a\mapsto a\ (a\in\bR)$.
238: 
239: With this terminology we make the following definition:
240: \begin{definition}[{\it Discontinuous intersection pairing}]
241: Define a pairing
242: $$
243: \tilde\iota\colon H^1(\Sg;\bR)\times H^1(\Sg;\bR)\lra S^2_\bQ\bR
244: $$
245: as follows. Choose a basis $x_1,\cdots , x_{2g}$ of $H^1(\Sg;\bQ)$.
246: For any two elements, $u,v\in H^1(\Sg;\bR)$, write
247: $$
248: u=\sum_i a_i x_i,\quad v=\sum_i b_i x_i\quad (a_i, b_i\in\bR) \ .
249: $$
250: Then we set
251: $$
252: \tilde\iota (u,v)=\sum_{i,j} \iota(x_i,x_j) \hat a_i \hat b_j \in S^2_\bQ
253: \bR \ .
254: $$
255: \end{definition}
256: Clearly $\tilde\iota$ followed by the projection $S^2_\bQ\bR\lra \bR$ is
257: nothing but the usual intersection or cup-product pairing $\iota$.
258: Henceforth we simply write $u\odot v$ for $\tilde\iota (u,v)$.
259: 
260: It is easy to see that $\tilde\iota$ is well defined independently of the
261: choice of basis in $H^1(\Sg;\bQ)$. We can see from the following proposition
262: that $\tilde\iota$ enumerates all $\bZ$-multilinear skew-symmetric pairings
263: on $H^1(\Sg;\bR)$ which are $\Mg$-invariant. Let $\La^2_{\bZ} H^1(\Sg;\bR)$
264: denote the second exterior power, over $\bZ$, of $H^1(\Sg;\bR)$ considered as
265: an abelian group, rather than as a vector space over $\bR$. Also let
266: $\left(\La^2_{\bZ} H^1(\Sg;\bR)\right)_{\Mg}$ denote the abelian group of
267: coinvariants of $\La^2_{\bZ} H^1(\Sg;\bR)$ with respect to the natural action
268: of $\Mg$.
269: \begin{proposition}\label{prop:la2}
270: There exists a canonical isomorphism
271: $$
272: \left(\La^2_{\bZ} H^1(\Sg;\bR)\right)_{\Mg}\cong S^2_{\bQ} \bR
273: $$
274: given by the correspondence
275: $$
276: \left(\sum_i a_i u_i\right)\land
277: \left(\sum_j b_j v_j\right)\longmapsto
278: \sum_{i,j} \iota(u_i,v_j)\ \hat a_i \hat b_j\ \ ,
279: $$
280: where $a_i, b_j\in\bR$, $u_i,v_j\in H^1(\Sg;\bQ)$.
281: \end{proposition}
282: In order not to digress, we refer the reader to the appendix for a proof.
283: 
284: Before we can define some new cocycles on the group $\Symp\Sg$, we have to
285: recall some facts from~\cite{KM03}. The symplectomorphism group $\Symp\Sg$
286: acts on its identity component by conjugation, and acts on $H^{1}(\Sg;\bR)$
287: from the left via $\varph(w)=(\varph^{-1})^*(w)$.
288: The flux homomorphism $\Flux\colon\Symp_{0}\Sg\rightarrow H^1(\Sg;\bR)$
289: is equivariant with respect to these actions by Lemma~6 of~\cite{KM03}.
290: Its extension $\widetilde{\Flux}\colon\Symp\Sg\rightarrow H^1(\Sg;\bR)$
291: is a crossed homomorphism for the above action in the sense that
292: \begin{equation}\label{eq:eflux}
293: \widetilde{\Flux}(\varph\psi)=\widetilde{\Flux}(\varph)+
294: (\varph\inv)^*\widetilde{\Flux}(\psi) \ .
295: \end{equation}
296: \begin{definition}\label{def:al}
297:      Let $\varph_1,\ldots,\varph_{2k}\in\Symp\Sg$, and
298:      $$
299: \xi_i=((\varph_1 \ldots\varph_{i-1})^{-1})^*\widetilde{\Flux}(\varph_i) \ .
300: $$
301: Define a $2k$-cocycle $\tilde\al^{(k)}$ with values in $S^{k}(S^2_\bQ\bR)$ by
302: \begin{align*}
303: \tilde\al^{(k)}&(\varph_1,\ldots,\varph_{2k})\\
304: &=
305: \frac{1}{(2k)!}\sum_{\sigma\in\mathfrak{S}_{2k}}
306: \sgn\sigma\ (\xi_{\sigma(1)}\odot\xi_{\sigma(2)})\ldots
307: (\xi_{\sigma(2k-1)}\odot\xi_{\sigma(2k)})\in S^{k}(S^2_\bQ\bR) \ ,
308: \end{align*}
309: where the sum is over permutations in the symmetric group $\mathfrak{S}_{2k}$.
310: \end{definition}
311: 
312: That $\tilde\al^{(k)}$ is indeed a cocycle is easy to check by a standard
313: argument in the theory of cohomology of groups using~\eqref{eq:eflux}. Thus
314: we have the corresponding cohomology classes
315: $$
316: \tilde\al^{(k)}\in H^{2k}(\Symp^\delta\Sg;S^{k}(S^2_\bQ\bR)) \ ,
317: $$
318: denoted by the same letters. If we apply the canonical projection
319: $S^{k}(S^2_\bQ\bR)\ra\bR$ to these classes, we obtain real cohomology classes
320: $$
321: \al^k\in H^{2k}(\Symp^\delta\Sg;\bR) \ ,
322: $$
323: which are the usual cup products of the first one $\al\in 
324: H^2(\Symp^\delta\Sg;\bR)$.
325: The refined classes $\tilde\al^{(k)}$ can be considered as a twisted version of
326: {\it discontinuous invariants} in the sense of~\cite{Morita85} arising from the
327: flux homomorphism.
328: 
329: Now we can state our first main result.
330: \begin{theorem}\label{th:hev}
331: For any $k\geq 1$ and $g\geq 3k$, the characteristic classes
332: $$
333: e_1^k, e_1^{k-1}\tilde\al, \ldots, e_1\tilde\al^{(k-1)}, \tilde\al^{(k)}
334: $$
335: induce a surjective homomorphism
336: $$
337: H_{2k}(\Symp^{\delta}\Sg;\bZ)\lra
338: \bZ\oplus S^{2}_{\bQ}\bR\oplus\cdots\oplus S^{k}(S^2_\bQ\bR) \ .
339: $$
340: \end{theorem}
341: For $k=1$ this is not hard to see, so we give the proof right away.
342: For $k>1$ the proof is given in Section~\ref{s:proofs} below and requires
343: the technical results developed in the body of this paper.
344: 
345: Consider the subgroup $\Ham\Sg$ of $\Symp_0\Sg$ consisting of all
346: Hamiltonian symplectomorphisms of $\Sg$. As is well known
347: (see~\cite{Banyaga,MS}), we have an extension
348: \begin{equation}\label{Ham}
349: 1\lra\Ham\Sg\lra\Symp_0\Sg\ \overset{\Flux}{\lra}\
350: H^1(\Sg;\bR)\lra 1 \ .
351: \end{equation}
352: This gives rise to a $5$-term exact sequence in cohomology:
353: \begin{align*}
354: 0 \lra H^{1}(H^1(\Sg;\bR)^\delta;\bZ) &\stackrel{\Flux^{*}}{\lra}
355: H^{1}(\Symp_0^\delta\Sg;\bZ)\lra
356: H^{1}(\Ham^\delta\Sg;\bZ)^{H^1_{\bR}}\\
357: &\lra H^{2}(H^1(\Sg;\bR)^\delta;\bZ)\stackrel{\Flux^{*}}{\lra}
358: H^{2}(\Symp_0^\delta\Sg;\bZ) \ ,
359: \end{align*}
360: where we have written $H^1_{\bR}$ for $H^1(\Sg;\bR)$. Now $\Ham\Sg$ is
361: a perfect group by a result of Thurston~\cite{Thurston0}, see also
362: Banyaga~\cite{Banyaga}. Therefore, $\Flux^{*}$ injects the second
363: cohomology of $H^1(\Sg;\bR)$ as a discrete group into that of
364: $\Symp_0^\delta\Sg$. By definition, the class $\tilde\alpha$ is the image
365: of the class of $\tilde\iota$ under $\Flux^{*}$. So $\tilde\alpha$ is
366: nontrivial on $\Symp_{0}^{\delta}\Sg$, and is defined on the whole
367: $\Symp^{\delta}\Sg$. We conclude that $\tilde\alpha$ defines a
368: surjective homomorphism
369: $$
370: H_{2}(\Symp^{\delta}\Sg;\bZ)\rightarrow S^{2}_{\bQ}\bR \ ,
371: $$
372: for any $g\geq 2$. We already proved in~\cite{KM03} that the first
373: Mumford--Morita--Miller class $e_{1}$ defines a surjection
374: $H_{2}(\Symp^{\delta}\Sg;\bZ)\longrightarrow \bZ$ for all $g\geq 3$.
375: Clearly the two classes are linearly independent because $\tilde\alpha$
376: is nonzero on $\Symp_{0}^{\delta}\Sg$, to which $e_{1}$ restricts trivially.
377: This proves Theorem~\ref{th:hev} in the easy case when $k=1$.
378: 
379: We can restrict the homomorphism
380: $$
381: \Flux^*\colon H^*(H^1(\Sg;\bR)^\delta;\bR)\lra
382: H^*(\Symp_0^\delta\Sg;\bR)
383: $$
384: to the {\it continuous cohomology}
385: $$
386: H^*_{ct}(H^1(\Sg;\bR)^\delta;\bR)\cong H_*(T^{2g};\bR)
387: \subset H^*(H^1(\Sg;\bR)^\delta;\bR) \ ,
388: $$
389: see Section~\ref{s:trans} for the precise definition. Thereby we obtain a ring
390: homomorphism
391: $$
392: \Flux^*\colon H_*(T^{2g};\bR)\longrightarrow H^*(\Symp_0^\delta\Sg;\bR) \ ,
393: $$
394: where $T^{2g}=K(\pi_1\Sg,1)$ is the Jacobian manifold of $\Sg$, and the ring
395: structure on the homology of $T^{2g}$ is induced by the Pontrjagin product.
396: Let $\om_0\in H_2(T^{2g};\bR)$ be the homology class represented by the dual
397: of the standard symplectic form on $T^{2g}$. We decompose the
398: $Sp(2g,\bR)$-module $H_k(T^{2g};\bR)$ into irreducible components.
399: For this, consider the homomorphism
400: $$
401: \om_0\wedge \ \colon H_{k-2}(T^{2g};\bR)\longrightarrow H_k(T^{2g};\bR)
402: $$
403: induced by the wedge product with $\om_0$. On the one hand, it is easy
404: to see using Poincar\'e duality on $T^{2g}$, that the above homomorphism is
405: surjective for any $k\geq g+1$. On the other hand, it is well-known
406: (see~\cite{FH}), that the kernel of the contraction homomorphism
407: $$
408: C\colon H_k(T^{2g};\bQ)\longrightarrow H_{k-2}(T^{2g};\bQ)
409: $$
410: induced by the intersection pairing
411: $H_2(T^{2g};\bQ)\cong \La^2 H_1(\Sg;\bQ)\ra\bQ$
412: is the irreducible representation of the algebraic group $Sp(2g,\bQ)$
413: corresponding to the Young diagram $[1^{k}]$ for any $k\leq g$.
414: Let $[1^k]_\bR=[1^k]\otimes\bR$ denote the real form of this representation.
415: Then we have a direct sum decomposition
416: \begin{equation}\label{eq:irrep}
417: H_k(T^{2g};\bR)= [1^k]_\bR\oplus \om_0\land H_{k-2}(T^{2g};\bR)
418: \quad (k\leq g) \ .
419: \end{equation}
420: \begin{theorem}\label{th:om2}
421: The kernel of the homomorphism
422: $$
423: \Flux^*\colon H_*(T^{2g};\bR)\lra H^*(\Symp_0^\delta\Sg;\bR)
424: $$
425: induced by the flux homomorphism is the ideal generated by the subspace
426: $\om_0\land H_1(T^{2g};\bR)\subset H_3(T^{2g};\bR)$, and the image of this
427: homomorphism can be described as
428: $$
429: \im \Flux^*\cong \bR\oplus \bigoplus_{k=1}^{g} [1^k]_\bR \ ,
430: $$
431: where $\bR$ denotes the image of the subspace of $H_2(T^{2g};\bR)$ spanned
432: by $\om_0$.
433: \end{theorem}
434: %\begin{theorem}\label{th:om2}
435: %Let $\om_0\in H_2(T^{2g};\bR)$ be the homology class represented by the dual
436: %of the standard symplectic form on $T^{2g}$. The kernel of the homomorphism
437: %$$
438: %\Flux^*\colon H_*(T^{2g};\bR)\lra H^*(\Symp_0^\delta\Sg;\bR)
439: %$$
440: %induced by the flux homomorphism is the ideal generated by the subspace
441: %$\om_0\land H_1(T^{2g};\bR)\subset H_3(T^{2g};\bR)$.
442: %\end{theorem}
443: 
444: The group $H^1(\Sg;\bR)$ acts on $\Ham\Sg$ by outer automorphisms.
445: In particular, it acts on the homology $H_*(\Ham^{\delta}\Sg;\bZ)$
446: of the discrete group $\Ham^{\delta}\Sg$ so that we can consider the
447: coinvariants $H_*(\Ham^{\delta}\Sg;\bZ)_{H^1_{\bR}}$, where for simplicity
448: we have written $H^1_{\bR}$ instead of $H^1(\Sg;\bR)$.
449: The following result shows that the second homology group
450: $H_2(\Ham^{\delta}\Sg;\bZ)$ is highly non-trivial.
451: \begin{theorem}\label{th:ham}
452: For any $g\geq 2$, there exists a natural injection
453: $$
454: H^1(\Sg;\bR)\subset H_2(\Ham^{\delta}\Sg;\bZ)_{H^1_{\bR}} \ .
455: $$
456: \end{theorem}
457: 
458: \section{The transverse symplectic class}\label{s:trans}
459: 
460: Since $\Sg$ is an Eilenberg-MacLane space, the total space
461: of the universal foliated $\Sg$-bundle over the classifying
462: space $\BS^\delta\Sg$ is again a $K(\pi,1)$ space.
463: Hence if we denote by $\ES^{\delta}\Sg$ the fundamental group of
464: this total space, then we obtain a short
465: exact sequence
466: \begin{equation}\label{eq:esymp}
467: 1\lra\pi_1\Sg\lra\ES^{\delta}\Sg\lra\Symp^{\delta}\Sg\lra 1
468: \end{equation}
469: and any cohomology class of the total space can be considered as
470: an element in the group cohomology of $\ES^{\delta}\Sg$.
471: Now on the total space of any foliated $\Sg$-bundle with total holonomy
472: group contained in $\Symp\Sg$ there is a closed $2$-form $\tilde\om$
473: which restricts to the symplectic form $\om$ on each fiber. At the
474: universal space level, the de Rham cohomology class of $\tilde\om$
475: defines a class $v\in H^2(\ES^\delta;\bR)$ which we call the transverse
476: symplectic class. We normalize the symplectic form $\om$ on $\Sg$
477: so that its total area is equal to $2g-2$. It follows that the
478: restriction of $v$ to a fiber is the same as the negative of the Euler
479: class $e\in H^2(\ES^\delta;\bR)$ of the vertical tangent bundle.
480: 
481: Let $\ES_0\Sg$ denote the subgroup of $\ES^{\delta}\Sg$ obtained by
482: restricting the extension~\eqref{eq:esymp} to $\Symp_0\Sg\subset\Symp\Sg$.
483: Since any foliated $\Sg$-bundle with total holonomy in $\Symp_0\Sg$ is
484: trivial as a differentiable $\Sg$-bundle, there exists an isomorphism
485: $$
486: \ES_0\Sg\cong \pi_1\Sg\times\Symp_0\Sg \ .
487: $$
488: Henceforth we identify the above two groups.
489: By the K\"unneth decomposition, we have an isomorphism
490: \begin{equation}\label{eq:kunneth}
491: \begin{split}
492: H^2(&\ES_0^\delta\Sg;\bR)\cong H^2(\Sg;\bR)\oplus\\
493: &\oplus \left(H^1(\Sg;\bR)\otimes
494: H^1(\Symp_0^\delta\Sg;\bR)\right)\oplus H^2(\Symp_0^\delta\Sg;\bR) \ ,
495: \end{split}
496: \end{equation}
497: where we identify $H^*(\pi_1\Sg;\bR)$ with $H^*(\Sg;\bR)$.
498: %which is finite dimensional.
499: Let $\mu\in H^2(\Sg;\bZ)$ be the fundamental cohomology class
500: of $\Sg$. Clearly the Euler class $e\in H^2(\ES_0^\delta\Sg;\bR)$
501: is equal to $(2-2g)\mu$.
502: The flux homomorphism %$\Flux\colon\Symp_0\Sg\ra H^1(\Sg;\bR)$
503: gives rise to an element
504: \begin{align*}
505: [\Flux]\in \hspace{3mm} &\Hom_\bZ (H_1(\Symp_0^\delta\Sg;\bZ),H^1(\Sg;\bR))\\
506: \cong & \Hom_\bR(H_1(\Symp_0^\delta\Sg;\bR),H^1(\Sg;\bR))\\
507: \cong & H^1(\Sg;\bR)\otimes H^1(\Symp_0^\delta\Sg;\bR) \ ,
508: \end{align*}
509: where the last isomorphism exists because $H^1(\Sg;\bR)$
510: is finite dimensional. Choose a symplectic basis
511: $x_1,\ldots,x_g$, $y_1,\ldots,y_g$
512: of $H_1(\Sg;\bR)$ and denote by $x_1^*,\ldots,x_g^*$, $y_1^*,\ldots,y_g^*$
513: the dual basis of $H^1(\Sg;\bR)$. Then Poincar\'e duality
514: $H_1(\Sg;\bR)\cong H^1(\Sg;\bR)$ is given by the
515: correspondence $x_i\mapsto -y_i^*, y_i\mapsto x_i^*$.
516: The element $[\Flux]$ can be described explicitly as
517: \begin{equation}\label{eqn:flux}
518: [\Flux]=\sum_{i=1}^{g} (x_i^*\otimes \tilde x _i+y_i^*\otimes \tilde y_i)
519: \in H^1(\Sg;\bR)\otimes H^1(\Symp_0^\delta\Sg;\bR)
520: \end{equation}
521: where $\tilde x_i,\tilde y_i\in H^1(\Symp_0^\delta\Sg;\bR)\cong
522: \Hom (H_1(\Symp_0\Sg;\bZ);\bR)$ is defined by the
523: equality
524: $$
525: \Flux(\varph)=\sum_{i=1}^{g}
526: (\tilde x_i(\varph) x_i+\tilde y_i(\varph) y_i)\quad (\varph\in \Symp_0\Sg).
527: $$
528: The elements $\tilde x_i,\tilde y_i$ can be also interpreted
529: as follows.
530: The flux homomorphism induces a homomorphism in cohomology
531: \begin{equation}\label{eq:flux*}
532: \Flux^*\colon H^*(H^1(\Sg;\bR)^\delta;\bR)\lra H^*(\Symp_0^\delta\Sg;\bR)
533: \end{equation}
534: where the domain
535: $$
536: H^*(H^1(\Sg;\bR)^\delta;\bR)\cong \Hom_{\bZ}(\La_{\bZ}^*(H^1(\Sg;\bR)),\bR)
537: $$
538: is the cohomology group of $H^1(\Sg;\bR)$ considered as a {\it discrete
539: abelian group}, rather than as a vector space over $\bR$, so that it is
540: a very large group. Its {\it continuous part} is defined as
541: \begin{align*}
542: H^*_{ct}(H^1(\Sg;\bR)^\delta;\bR)=&\Hom_{\bR}(\La_{\bR}^*(H^1(\Sg;\bR)),\bR)\\
543: \subset &\Hom_{\bZ}(\La_{\bR}^*(H^1(\Sg;\bR)),\bR)\\
544: \subset &\Hom_{\bZ}(\La_{\bZ}^*(H^1(\Sg;\bR)),\bR)\cong
545: H^*(H^1(\Sg;\bR)^\delta;\bR) \ ,
546: \end{align*}
547: where the second inclusion is induced by the natural projection
548: $$
549: \La_{\bZ}^*(H^1(\Sg;\bR))\longrightarrow \La_{\bR}^*(H^1(\Sg;\bR)) \ .
550: $$
551: Denoting by $T^{2g}=K(\pi_1\Sg,1)$ the Jacobian torus of $\Sg$, there is
552: a canonical isomorphism
553: $$
554: \La^*_{\bR}H^1(\Sg;\bR)\cong H^*(T^{2g};\bR) \ ,
555: $$
556: so that we can identify
557: $$
558: H^*_{ct}(H^1(\Sg;\bR)^\delta;\bR)\cong \Hom_\bR(H^*(T^{2g};\bR);\bR)
559: \cong H_*(T^{2g};\bR) \ .
560: $$
561: Thus, by restricting the homomorphism $\Flux^{*}$ in~\eqref{eq:flux*}
562: to the continuous cohomology, we obtain a homomorphism
563: $$
564: \Flux^*\colon H_*(T^{2g};\bR)\longrightarrow H^*(\Symp_0^\delta\Sg;\bR) \ .
565: $$
566: It is easy to see that under this homomorphism we have
567: $$
568: \tilde x_i=\Flux^*(x_i) \ ,\quad \tilde y_i=\Flux^*(y_i) \ .
569: $$
570: Let $\om_0\in \La^2_{\bR}H_1(\Sg;\bR)$ be the symplectic class
571: defined by
572: $$
573: \om_0=\sum_{i=1}^g x_i\land y_i
574: $$
575: and set
576: $$
577: \tilde\om_0=\Flux^*(\om_0)
578: =\sum_{i=1}^{g} \tilde x_i \tilde y_i\in H^2(\Symp_0^\delta\Sg;\bR) \ .
579: $$
580: 
581: \begin{lemma}\label{lem:flux2}
582: We have the equality
583: $$
584: [\Flux]^2=-2\mu\otimes \tilde\om_0\in
585: H^2(\Sg;\bR)\otimes H^2(\Symp_0^\delta\Sg;\bR) \ .
586: $$
587: \end{lemma}
588: \begin{proof}
589: A direct calculation using the expression~\eqref{eqn:flux} yields
590: $$
591: [\Flux]^2=-\sum_{i=1}^{g} (x_i^* y_i^*\otimes\tilde x_i\tilde y_i
592: +y_i^* x_i^*\otimes \tilde y_i\tilde x_i) \ .
593: $$
594: Since $x_i^*y_i^*=-y_i^*x_i^*=\mu$, we obtain
595: %\begin{align*}
596: $$
597: [\Flux]^2=-2\mu\otimes\sum_{i=1}^{g} \tilde x_i\tilde y_i
598: =-2\mu\otimes\tilde\om_0
599: $$
600: %\end{align*}
601: as required.
602: \end{proof}
603: 
604: Now we can completely determine the transverse symplectic class $v$ of
605: foliated $\Sg$-bundles whose total holonomy groups are contained
606: in the identity component $\Symp_0\Sg$ of $\Symp\Sg$ as follows.
607: \begin{proposition}\label{prop:v}
608: On the subgroup $\ES_0^\delta\Sg$ the transverse symplectic class
609: $v\in H^2(\ES_0^\delta\Sg;\bR)$ is given by
610: $$
611: v=(2g-2)\mu+[\Flux]+\frac{1}{2g-2}\tilde\om_0
612: $$
613: under the isomorphism \eqref{eq:kunneth}.
614: Furthermore, the homomorphism
615: $$
616: \tilde\om_0\otimes H^1_{ct}(\Symp_0^\delta\Sg;\bR)\lra
617: H^3(\Symp_0^\delta\Sg;\bR)
618: $$
619: induced by the cup product is trivial, where
620: $H^1_{ct}(\Symp_0^\delta\Sg;\bR)\cong H_1(\Sg;\bR)$
621: denotes the subgroup of $H^1(\Symp_0^\delta\Sg;\bR)$
622: generated by the continuous cohomology classes
623: $\tilde x_i, \tilde y_i$.
624: In particular, $\tilde\om_0^2=0$.
625: \end{proposition}
626: \begin{proof}
627: Since the restriction of $v$ to each fiber is equal to the negative of
628: that of the Euler class $e$ by our normalization, $v$ restricts to
629: $(2g-2)\mu$ on the fiber. This gives the first component of the formula.
630: The second component follows from Lemma~8 of~\cite{KM03}.
631: Thus we can write
632: $$
633: v=(2g-2)\mu+[\Flux]+ \ga \in H^*(\Sg;\bR)\otimes H^*(\Symp_0^\delta\Sg;\bR)
634: $$
635: for some $\ga\in H^2(\Symp_0^\delta\Sg;\bR)$.
636: Now observe that $v^2=0$ because $\tilde\om^2=0$.
637: Also, because we have restricted to $\Symp_{0}\Sg$, we have
638: $\mu^2=0$, $\mu [\Flux]=0$. Hence we obtain
639: $$
640: [\Flux]^2+\ga^2+ 2(2g-2)\mu \ga+2 [\Flux] \ga=0 \ .
641: $$
642: It follows that
643: \begin{align*}
644: [\Flux]^2+2(2g-2)\mu \ga&=0\\
645: [\Flux] \ga&=0\\
646: \ga^2&=0
647: \end{align*}
648: because these three elements belong to different summands in the
649: K\"unneth decomposition
650: of $H^*(\Sg;\bR)\otimes H^*(\Symp_0^\delta;\bR)$.
651: If we combine the first equality above and Lemma~\ref{lem:flux2},
652: then we can conclude that
653: \begin{equation}\label{eqn:ga}
654: \ga=\frac{1}{2g-2}\tilde\om_0 \ .
655: \end{equation}
656: This proves the first claim of the proposition.
657: If we substitute~\eqref{eqn:flux} and~\eqref{eqn:ga} in
658: the second equality above, then we see that
659: $\tilde\om_0\tilde x_i=\tilde\om_0\tilde y_i=0$ for any $i$,
660: whence the second claim.
661: Observe that the third equality $\ga^2=0$, which is equivalent to
662: $\tilde\om_0^2=0$ by the above, is a consequence of the second claim.
663: \end{proof}
664: 
665: Now we can calculate the restriction of the cocycle $\alpha$ defined
666: in Section~\ref{s:main} to the identity component $\Symp_{0}\Sg$.
667: \begin{proposition}\label{prop:alom}
668: Let $i\colon\Symp_0^\delta\Sg\ra\Symp^\delta\Sg$ be the inclusion.
669: Then %we have
670: $$
671: i^*\al = 2 \tilde\om_0\in H^2(\Symp_0^\delta\Sg;\bR)
672: $$
673: and
674: $$
675: i^*\al^2=0 \ .
676: $$
677: \end{proposition}
678: \begin{proof}
679: Let $\varph, \psi\in \Symp_0\Sg$ be any two elements. Then, by the
680: definition of $\al$, see Definition~\ref{def:al} and the subsequent
681: discussion, we have
682: $$
683: \al(\varph,\psi)=\iota(\Flux(\varph), \Flux(\psi)) \ .
684: $$
685: By the definition of the cohomology classes $\tilde x_i, \tilde y_i$, we have
686: $$
687: \Flux(\varph)=\sum_{i=1}^{g}
688: (\tilde x_i(\varph) x_i+\tilde y_i(\varph) y_i)\ ,\quad
689: \Flux(\psi)=\sum_{i=1}^{g}
690: (\tilde x_i(\psi) x_i+\tilde y_i(\psi) y_i)\ .
691: $$
692: Hence we obtain
693: $$
694: \al(\varph,\psi)=\sum_{i=1}^g \left\{\tilde x_i(\varph) \tilde y_i(\psi)
695: -\tilde y_i(\varph) \tilde x_i(\psi)\right\}\ .
696: $$
697: Using the Alexander--Whitney cup product, we also have
698: $$
699: \tilde\om_0(\varph,\psi)=\sum_{i=1}^g \tilde x_i(\varph) \tilde y_i(\psi)
700: =-\sum_{i=1}^g \tilde x_i(\psi) \tilde y_i(\varph) \ .
701: $$
702: Thus $\al=2 \tilde\om_0$ as required. The last statement follows from
703: Proposition~\ref{prop:v}.
704: \end{proof}
705: We see from this proposition that $i^*\al^2=0$ in
706: $H^4(\Symp_0^\delta\Sg;\bR)$, while we will prove the
707: non-triviality of $\al^2\in H^4(\Symp^\delta\Sg;\bR)$, which is a special
708: case of Theorem~\ref{th:hev}.
709: One could say that the non-triviality of $\al^2$ is realized
710: by an interaction of the two groups $\Symp_0\Sg$ and $\Mg$.
711: 
712: %\section{The second homology group of $\Symp^\delta\Sg$}\label{s:second}
713: 
714: %In this section, we prove Theorem \ref{th:hev} for the special
715: %case $k=1$. In fact we prove the following slightly stronger statement.
716: %\begin{theorem}\label{th:h2}
717: %For any $g\geq 3$ the characteristic classes $e_1$ and $\tilde \al$ induce
718: %a surjective homomorphism
719: %$$
720: %H_2(\Symp^\delta\Sg;\bZ)\lra \bZ\oplus S^2_\bQ\bR \ .
721: %$$
722: %For $g=2$, $\tilde \al$ induces a surjection
723: %$$
724: %H_2(\Symp^\delta\Sigma_2;\bZ)\ra S^2_\bQ\bR \ .
725: %$$
726: %\end{theorem}
727: 
728: %Recall that we have the following two group extensions
729: %\begin{equation}\label{eq:hss}
730: %1\lra\Ham\Sg\lra\Symp_0\Sg\ \overset{\Flux}{\lra}\ H^1(\Sg;\bR)\lra 1 \ ,
731: %\end{equation}
732: %\begin{equation}\label{eq:ssm}
733: %1\lra\Symp_0\Sg\lra\Symp\Sg\lra\Mg\lra 1 \ .
734: %\end{equation}
735: %\begin{proposition}\label{prop:fl2}
736: %The homomorphism
737: %$$
738: %\Flux_*\colon H_2(\Symp_0^\delta\Sg;\bZ)\lra H_2(H^1(\Sg;\bR)^\delta;\bZ)
739: %\cong \La^2_{\bZ}H^1(\Sg;\bR)
740: %$$
741: %induced by the flux homomorphism is surjective for any $g\geq 2$.
742: %\end{proposition}
743: %\begin{proof}
744: %We consider the Hochschild-Serre exact sequence
745: %\begin{align*}
746: %H_2&(\Symp_0^\delta\Sg;\bZ)\ \overset{\Flux_*}{\lra}\
747: %H_2(H^1(\Sg;\bR)^\delta;\bZ)\lra H_1(\Ham^\delta\Sg;\bZ)_{H^1_{\bR}}\\
748: %&\lra H_1(\Symp_0^\delta\Sg;\bZ)\lra
749: %H_1(H^1(\Sg;\bR)^\delta;\bZ)\lra 0
750: %\end{align*}
751: %of the group extension~\eqref{eq:hss}. It is known that the group $\Ham\Sg$
752: %is perfect (see Banyaga~\cite{Banyaga}). Hence $H_1(\Ham^\delta\Sg;\bZ)=0$.
753: %The result follows.
754: %\end{proof}
755: 
756: %\begin{proof}[Proof of Theorem~\ref{th:h2}]
757: %First of all, we proved in~\cite{KM03} that for $g\geq 3$ the
758: %first Mumford-Morita-Miller class $e_1$, more precisely $\frac{1}{12}
759: %e_1$, induces a surjection $H_2(\Symp^\delta\Sg;\bZ)\ra \bZ$.
760: 
761: %Next we consider the group extension~\eqref{eq:ssm} and the spectral
762: %sequence $\{E^r_{p,q},d^r\}$ for its integral homology. Then we have
763: %$$
764: %E^2_{0,2}\cong H_0(\Mg;H_2(\Symp_0^\delta\Sg;\bZ))
765: %\cong H_2(\Symp_0^\delta\Sg;\bZ)_{\Mg}  \ .
766: %$$
767: %By Proposition~\ref{prop:fl2}, the homomorphism between the
768: %spaces of $\Mg$-coinvariants
769: %$$
770: %H_2(\Symp_0^\delta\Sg;\bZ)_{\Mg}\lra H_2(H^1(\Sg;\bR)^\delta;\bZ)_{\Mg}
771: %$$
772: %induced by the flux homomorphism is surjective.
773: %Since by Proposition~\ref{prop:la2} the target of this surjection can be
774: %identified with $S^2_\bQ\bR$, we obtain a surjection
775: %$$
776: %E^2_{0,2}\lra S^2_\bQ\bR \ .
777: %$$
778: %The restriction of the characteristic class $\tilde\al$ to the subgroup
779: %$\Symp_0\Sg\subset\Symp\Sg$ clearly detects this surjection.
780: %Hence it survives to the $E^\infty$-term and
781: %there exists a surjection
782: %$$
783: %H_2(\Symp^\delta\Sg;\bZ)\supset E^\infty_{0,2}\lra S^2_\bQ\bR \ .
784: %$$
785: %Again because $\tilde\al$ is defined on the whole group $\Symp\Sg$
786: %and not just on $\Symp_0\Sg$, we conclude that the homomorphism
787: %$H_2(\Symp^\delta\Sg;\bZ)\ra S^2_\bQ\bR$ is surjective.
788: 
789: %Finally, we have to prove that the two characteristic classes $e_1$
790: %and $\tilde \al$ are linearly independent. But this follows easily
791: %from the fact that the non-triviality of the $S^2_\bQ\bR$ factor
792: %can be realized on cycles whose supports lie in the subgroup $\Symp_0\Sg$,
793: %while the characteristic class $e_1$ is trivial on these cycles because
794: %topologically they are trivial $\Sg$-bundles. This completes the proof.
795: %\end{proof}
796: 
797: \section{Extended flux homomorphisms for open surfaces}\label{s:open}
798: 
799: We consider the open surface $\Sigma_g^0=\Sigma_g\setminus D$ obtained
800: from $\Sg$ by removing a closed embedded disk $D\subset\Sg$. Let
801: $j\colon\Sigma_g^0\ra \Sg$ be the inclusion. We denote by $\Symp^c\Sigma_g^0$
802: the symplectomorphism group of $(\Sigma_g^0,j^*\om)$ with {\it compact 
803: supports}.
804: Hence the group $\Symp^c\Sigma_g^0$ can be considered as a subgroup of 
805: $\Symp\Sg$
806: with inclusion $j\colon\Symp^c\Sigma_g^0\ra\Symp\Sg$. Let $\Symp_0^c\Sigma_g^0$
807: be the identity component of $\Symp^c\Sigma_g^0$. Clearly $\Symp_0^c\Sigma_g^0$
808: is a subgroup of $\Symp_0\Sg$.
809: 
810: Let $\M1g$ denote the mapping class group of $\Sg$ {\it relative to}
811: the embedded disk $D^2\subset\Sg$, equivalently, the mapping class
812: group of the compact surface $\overline{\Sigma_g^0}$ with boundary.
813: We have a natural homomorphism $p\colon\Symp^c\Sigma_g^0\ra\M1g$
814: which is easily seen to be surjective. Moreover, Moser's theorem~\cite{Moser}
815: adapted to the present case (see~\cite{Tsuboi} for a general statement),
816: implies that the kernel of this surjection is precisely the group
817: $\Symp_0^c\Sigma_g^0$. We summarize the situation in the following
818: diagram:
819: \begin{equation}
820: \begin{CD}
821: 1@>>> \Symp^{c}_{0}\Sigma_g^{0} @>{i^c}>>
822: \Symp^c\Sigma_g^{0}@>{p}>> \M1g @>>> 1\\
823: @. @V{j_0}VV @V{j}VV @V{q}VV  @.\\
824: 1@>>> \Symp_{0}\Sigma_g @>{i}>> \Symp\Sigma_g @>>> \Mg @>>> 1
825: \end{CD}
826: \label{eqn:GG}
827: \end{equation}
828: where $q\colon\M1g\ra\Mg$ denotes the natural projection.
829: 
830: The restriction of the flux homomorphism
831: \begin{equation}\label{eqn:Flux}
832: \Flux\colon\Symp_0\Sg\lra H^1(\Sg;\bR)
833: \end{equation}
834: to the subgroup $\Symp_0^c\Sigma_g^0$, denoted $j^*\Flux$, can be described
835: as follows. The restriction $j^{*}\omega$ of the area form $\omega$
836: to the open surface $\Sigma_{g}^{0}$ is exact. Choose a $1$-form $\la$ %on 
837: $\Sigma_g^0$
838: such that $d\la=-j^*\om$. Then, for any element $\varph\in\Symp_0^c\Sigma_g^0$,
839: the $1$-form $\la-\varph^*\la$ is a closed form with compact support.
840: Hence the corresponding de Rham cohomology class $[\la-\varph^*\la]$,
841: which can be shown to be independent of the choice of $\la$,
842: is an element of the first cohomology group $H^1_c(\Sigma_g^0;\bR)$
843: of $\Sigma_g^0$ with compact support. It is easy to see that
844: $H^1_c(\Sigma_g^0;\bR)$ is canonically isomorphic to $H^1(\Sg;\bR)$,
845: and that under this isomorphism
846: \begin{equation}\label{eq:Fluxc}
847: (j^*\Flux)(\varph)=[\la-\varph^*\la]\in H^1_c(\Sigma_g^0;\bR)
848: \cong H^1(\Sg;\bR) %\quad (\varph\in\Symp_0^c\Sigma_g^0)
849: \end{equation}
850: for all $\varph\in\Symp_0^c\Sigma_g^0$, see Lemma~10.14 of~\cite{MS}.
851: We obtain the following commutative diagram:
852: \begin{equation}
853: \begin{CD}
854: \Symp^{c}_{0}\Sigma_g^{0} @>{j^*\Flux}>> H^1_{c}(\Sigma_g^{0};\bR)\\
855: @VVV @VV{\cong}V\\
856: \Symp_{0}\Sigma_g @>>{\Flux}> H^1(\Sg;\bR).
857: \end{CD}
858: \label{eqn:FFlux}
859: \end{equation}
860: From now on we identify $H^1_{c}(\Sigma_g^{0};\bR)$ with $H^1(\Sg;\bR)$.
861: 
862: As was already mentioned in the Introduction, we proved in~\cite{KM03} that
863: the flux homomorphism~\eqref{eqn:Flux} can be extended to a crossed 
864: homomorphism
865: \begin{equation}\label{eqn:TFlux}
866: \widetilde{\Flux}\colon\Symp\Sg\lra H^1(\Sg;\bR) \ ,
867: \end{equation}
868: and that the extension is unique up to the addition of coboundaries.
869: The restriction of such a crossed homomorphism to the subgroup
870: $\Symp^c\Sigma_g^0\subset \Symp\Sg$, denoted $j^*\widetilde{\Flux}$,
871: is of course an extension of the flux homomorphism $j^*\Flux$.
872: However, we also have another extension of the same flux homomorphism
873: $j^*\Flux$ to the group $\Symp^c\Sigma_g^0$ as follows.
874: \begin{proposition}\label{prop:Fluxb}
875: The map
876: $$
877: \widetilde{\Flux}_c\colon\Symp^c\Sigma_g^0\lra H^1_c(\Sigma_g^0;\bR)
878: \cong H^1(\Sg;\bR) %\quad (\varph\in\Symp^c\Sigma_g^0)
879: $$
880: defined by
881: $\widetilde{\Flux}_c(\varph)=[(\varph\inv)^*\la-\la]\in H^1_c(\Sigma_g^0;\bR)$
882: is a crossed homomorphism which extends the flux homomorphism $j^*\Flux$.
883: Its cohomology class
884: $[\widetilde{\Flux}_c]\in H^1(\Symp^c\Sigma_g^0;H^1(\Sg;\bR))$
885: is uniquely determined independently of the choice of the
886: $1$-form $\la$ such that $d\la=-j^*\om$.
887: \end{proposition}
888: \begin{proof}
889: First observe that, for any $\varph\in \Symp_0^c\Sigma_g^0$, we have
890: $$
891: [\la-\varph^*\la]=(\varph\inv)^*[\la-\varph^*\la]=[(\varph\inv)^*\la-\la]
892: $$
893: because $\varph$ acts trivially on $H^1(\Sg;\bR)$. Hence
894: by~\eqref{eq:Fluxc} we have
895: $$
896: \widetilde{\Flux}_c(\varph)=(j^*\Flux)(\varph) \ .
897: $$
898: Next, for any two elements $\varph, \psi\in \Symp^c\Sigma_g^0$, we have
899: \begin{align*}
900: \widetilde{\Flux}_c(\varph\psi)&=[((\varph\psi)\inv)^*\la-\la]\\
901: &=[(\varph\inv)^*\la-\la +(\varph\inv)^*(\psi\inv)^*\la-(\varph\inv)^*\la]\\
902: &=[(\varph\inv)^*\la-\la]+[(\varph\inv)^*((\psi\inv)^*\la-\la)]\\
903: &=\widetilde{\Flux}_c(\varph)+(\varph\inv)^*\widetilde{\Flux}_c(\psi).
904: \end{align*}
905: Therefore $\widetilde{\Flux}_c$ is a crossed homomorphism which extends 
906: $j^*\Flux$.
907: 
908: Finally, let $\la'$ be another $1$-form on $\Sigma_g^0$
909: satisfying $d\la'=-j^*\om$, and let $\widetilde{\Flux}'_c$ be the corresponding
910: crossed homomorphism. Then $a=\la'-\la$ is a closed $1$-form defining a de
911: Rham cohomology class $[a]\in H^1(\Sigma_g^0;\bR)\cong H^1(\Sg;\bR)$.
912: Now
913: \begin{align*}
914: \widetilde{\Flux}'_c(\varph)&=[(\varph\inv)^*\la'-\la']\\
915: &=[(\varph\inv)^*(\la+a)-(\la+a)]\\
916: &=[(\varph\inv)^*\la-\la]+[(\varph\inv)^*a-a]\\
917: &=\widetilde{\Flux}_c(\varph)+(\varph\inv)^*[a]-[a] \in H^1(\Sg;\bR)
918: \ .
919: \end{align*}
920: This shows that the difference $\widetilde{\Flux}'_c-\widetilde{\Flux}_c$
921: is a coboundary, completing the proof of the proposition.
922: \end{proof}
923: 
924: We have proved that the restriction $j^*\Flux$ of the flux
925: homomorphism~\eqref{eqn:Flux} to the subgroup $\Symp_0^c\Sigma_g^0$ has
926: two extensions to the group $\Symp^c\Sigma_g^0$ as a crossed homomorphism.
927: One is the restriction $j^*\widetilde{\Flux}$ of $\widetilde{\Flux}$, and
928: the other is $\widetilde{\Flux}_c$. We will show that these two crossed
929: homomorphisms are essentially different. More precisely, we will show that
930: the difference of these two crossed homomorphisms can be expressed by an
931: element of the cohomology group
932: \begin{equation}\label{eqn:hm1g}
933: H^1(\M1g;H^1(\Sg;\bR)) \ .
934: \end{equation}
935: It was proved in~\cite{Morita89} that $H^1(\M1g;H^1(\Sg;\bZ))$ is
936: isomorphic to $\bZ$ for all $g\geq 2$. It follows that the above
937: group~\eqref{eqn:hm1g} is isomorphic to $\bR$ and if
938: $$
939: k\colon \M1g \lra  H^1(\Sg;\bZ)
940: $$
941: is any crossed homomorphism whose cohomology class is a
942: generator of $H^1(\M1g;H^1(\Sg;\bZ))$, then the associated
943: crossed homomorphism
944: $$
945: k_\bR\colon\M1g \lra  H^1(\Sg;\bR)
946: $$
947: represents the element $1\in H^1(\M1g;H^1(\Sg;\bR))\cong\bR$.
948: Let $p^*k_\bR\in H^1(\Symp^c\Sg;H^1(\Sg;\bR))$ be the
949: class induced from $k_\bR$ by the projection $p\colon\Symp^c\Sg\ra\M1g$.
950: 
951: \begin{theorem}\label{th:h1sc}
952: For any $g\geq 2$, there exists an isomorphism
953: $$
954: H^1(\Symp^c\Sigma_g^0;H^1(\Sg;\bR))\cong \bR\oplus\Hom_{\bZ}(\bR,\bR)
955: $$
956: such that $p^*k_\bR$ and $[j^*\widetilde{\Flux}]$ represent
957: the classes $1\in \bR$ and $\id\in \Hom_{\bZ}(\bR,\bR)$ respectively.
958: \end{theorem}
959: \begin{proof}
960: The top extension in~\eqref{eqn:GG} gives rise to an exact sequence
961: \begin{align*}
962: & 0\ \lra \ H^1(\M1g;H^1(\Sg;\bR))\ \cong\bR\ \overset{p^*}{\lra} \\
963:    H^1(&\Symp^c\Sg;H^1(\Sg;\bR))\lra
964: H^1(\Symp^c_0\Sg;H^1(\Sg;\bR))^{\M1g}\lra\cdots .
965: \end{align*}
966: It was proved in~\cite{KM03} that $j^*\Flux$ induces an isomorphism
967: $$
968: j^*\Flux\colon H_1(\Symp^c_0\Sigma_g^0;\bZ)\cong
969: H^1_c(\Sigma_g^0;\bR)=H^1(\Sg;\bR) \ .
970: $$
971: Hence, by an argument similar to  the one given for $\Symp_0\Sg$ 
972: in~\cite{KM03},
973: we have an isomorphism
974: $$
975: H^1(\Symp^c_0\Sigma_g^0;H^1(\Sg;\bR))^{\M1g}\cong \Hom_\bZ(\bR,\bR) \ ,
976: $$
977: and clearly $j^*[\widetilde{\Flux}]$ corresponds to $\id\in \Hom_\bZ(\bR,\bR)$.
978: The result follows from this.
979: \end{proof}
980: 
981: \begin{theorem}\label{th:ffk}
982: We have the identity
983: $$
984: [\widetilde{\Flux}_c]=j^*[\widetilde{\Flux}]- p^*k_\bR
985: $$
986: in $H^1(\Symp^c\Sigma_g^0;H^1(\Sg;\bR))$.
987: \end{theorem}
988: \begin{proof}
989: Since both crossed homomorphisms $\widetilde{\Flux}_c$ and 
990: $j^*\widetilde{\Flux}$
991: are extensions of the flux homomorphism $j^*\Flux$, the proof of
992: Theorem~\ref{th:h1sc} implies that
993: $$
994: [\widetilde{\Flux}_c]=j^*[\widetilde{\Flux}]+a\ p^*k_\bR
995: $$
996: for some constant $a\in\bR$. Let $\I1g\subset\M1g$ denote the Torelli
997: subgroup consisting of mapping classes which act trivially on homology.
998: We set ${\mathcal I}\Symp^c\Sigma_g^0=p^{-1}(\I1g)\subset\Symp^c\Sigma_g^0$.
999: If we restrict the crossed homomorphims $\widetilde{\Flux}_c$ and
1000: $\widetilde{\Flux}$ to this subgroup ${\mathcal I}\Symp^c\Sigma_g^0$, then
1001: they become homomorphisms which depend only on the cohomology classes
1002: $[\widetilde{\Flux}_c]$ and $j^*[\widetilde{\Flux}]$, and not on the
1003: particular crossed homomorphisms representing these cohomology classes.
1004: This is because any crossed homomorphism which is a coboundary is trivial
1005: on ${\mathcal I}\Symp^c\Sigma_g^0$. It was proved in~\cite{Morita89}
1006: that a generator of the group $H^1(\M1g;H^1(\Sg;\bZ))\cong \bZ$ is
1007: characterized by the fact that the {\it Poincar\'e dual} of its value on
1008: a single non-trivial element $\varph \in \I1g$ is equal to $\pm C 
1009: \tau(\varph)$,
1010: where $\tau\colon\I1g\ra \La^3 H_1(\Sg;\bZ)$ denotes the
1011: (first) Johnson homomorphism and $C\colon\La^3 H_1(\Sg;\bZ)\ra H_1(\Sg;\bZ)$
1012: denotes the contraction. Hence we only have to compute
1013: the values of $\widetilde{\Flux}_c$ and $\widetilde{\Flux}$
1014: on some particular element $\tilde\varph\in {\mathcal I}\Symp^c\Sigma_g^0$.
1015: We choose such element as follows.
1016: 
1017: \begin{figure}
1018: \epsfig{file=figure1.eps,width=12cm}
1019: %\epsfig{file=figure1.pdf,width=12cm}
1020: \caption{ }\label{fig:torelli}
1021: \end{figure}
1022: 
1023: Consider an embedded disk $D'$ in $\Sigma_g^0=\Sg\setminus D$ as depicted
1024: in Figure~\ref{fig:torelli} and set $\Sigma_g^{00}=\Sg\setminus (D\cup D')$.
1025: We also consider two disjoint simple closed curves $C_1, C_2\subset 
1026: \Sigma_g^{00}$
1027: as shown in Figure~\ref{fig:torelli}.
1028: Let $\tau_i\ (i=1,2)$ denote the Dehn twist along $C_i$
1029: and choose lifts $\tilde\tau_i\in\Symp^c\Sigma_g^0$ of $\tau_i$ with
1030: supports in small neighborhoods of the $C_{i}$.
1031: Now we set
1032: $$
1033: \tilde\varph=\tilde\tau_1 \tilde\tau_2^{-1} \ .
1034: $$
1035: Since $\varph=\tau_1\tau_2^{-1}$ belongs to the Torelli group $\I1g\subset 
1036: \M1g$,
1037: its lift $\tilde\varph$ is an element of the subgroup
1038: ${\mathcal I}\Symp^c\Sigma_g^0$. We would like to compute the difference
1039: \begin{equation}\label{eq:difference}
1040: \begin{split}
1041: \widetilde{\Flux}_c(\tilde\varph)-\widetilde{\Flux}(\tilde\varph)
1042: &= a\ p^*k_\bR(\tilde\varph)\\
1043: &=a\ k(\varph)\in H^1_c(\Sigma_g^0;\bR)\cong H^1(\Sg;\bR) \ .
1044: \end{split}
1045: \end{equation}
1046: 
1047: Now we consider the long exact sequence
1048: \begin{equation}\label{eq:s0}
1049: \begin{split}
1050: 0\lra H^1(\Sg;\bR)&\overset{i^*}{\lra}H^1(\Sigma_g^{00};\bR)
1051: \lra \\
1052: & H^2(\Sg,\Sigma_g^{00};\bR)\cong\bR^2\lra
1053: H^2(\Sg;\bR)\lra 0
1054: \end{split}
1055: \end{equation}
1056: of the pair $(\Sg, \Sigma_g^{00})$.
1057: Also consider the following short exact sequences
1058: \begin{equation}\label{eq:s0'}
1059: \begin{split}
1060: 0\lra H^1_c(\Sigma_g^0;\bR)\cong H^1(\Sg;\bR)&\overset{j^*}{\lra}
1061: H^1(\Sigma_g^{00};\bR)\lra \bR\lra 0\\
1062: 0\lra H^1_c(\Sigma_g\setminus D';\bR)\cong H^1(\Sg;\bR)&\overset{(j')^*}{\lra}
1063: H^1(\Sigma_g^{00};\bR)\lra \bR\lra 0 \ .
1064: \end{split}
1065: \end{equation}
1066: In view of the above exact sequences~\eqref{eq:s0} and~\eqref{eq:s0'},
1067: we can determine the value of~\eqref{eq:difference} in the
1068: group $H^1(\Sigma_g^{00};\bR)$ because both groups
1069: $H^1_c(\Sigma_g^0;\bR)$ and $H^1(\Sg;\bR)$ are embedded in it.
1070: Choose $1$-forms
1071: $\la$ and $\la'$ on $\Sigma_g^0=\Sg\setminus D$ and
1072: $\Sg\setminus D'$ respectively, such that
1073: $$
1074: d\la=-j^*\om,\quad d\la'=-(j')^*\om
1075: $$
1076: where $j\colon\Sigma_g^0\ra\Sg$ and $j'\colon\Sg\setminus D'\ra\Sg$ are
1077: the inclusions. Then $d(\la-\la')=0$ so that $\nu=\la-\la'$ is a
1078: closed $1$-form on $\Sigma_g^{00}$. We have the identity
1079: \begin{align*}
1080: (\tilde\varph\inv)^*\la-\la&=(\tilde\varph\inv)^*(\la'+\nu)-(\la'+\nu)\\
1081: &=(\tilde\varph\inv)^*\la'-\la'+(\tilde\varph\inv)^*\nu-\nu \ .
1082: \end{align*}
1083: Hence we have the equality
1084: \begin{equation}\label{eq:[]}
1085: [(\tilde\varph\inv)^*\la-\la]=[(\tilde\varph\inv)^*\la'-\la']+
1086: [(\tilde\varph\inv)^*\nu-\nu]
1087: \end{equation}
1088: in the group $H^1(\Sigma_g^{00};\bR)$.
1089: By the definition of $\widetilde{\Flux}_c$, we have
1090: \begin{equation}\label{eq:fla}
1091: \widetilde\Flux_c(\tilde\varph)=[(\tilde\varph\inv)^*\la-\la]\in
1092: H^1_c(\Sigma_g^0;\bR)\subset H^1(\Sigma_g^{00};\bR) \ .
1093: \end{equation}
1094: Next we compute $\widetilde{\Flux}(\tilde\varph)$. For this, observe
1095: that $\tilde\varph$ is isotopic to the identity as an element of
1096: $\Symp^c(\Sg\setminus D')\subset \Symp\Sg$, although it is {\it not}
1097: isotopic to the identity as an element of $\Symp^c(\Sg\setminus D)$.
1098: Hence $\tilde\varph\in \Symp^c_0(\Sg\setminus D')$ and
1099: $\widetilde{\Flux}(\tilde\varph)=((j')^*\Flux)(\tilde\varph)$.
1100: By replacing $D$ with $D'$ in equation~\eqref{eq:Fluxc}, we obtain
1101: \begin{equation}\label{eq:Fluxc'}
1102: \begin{split}
1103: ((j')^*\Flux) &(\tilde\varph)=[\la'-\tilde\varph^*\la']\\
1104: = &[(\tilde\varph\inv)^*\la'-\la']\in
1105: H^1_c(\Sigma_g\setminus D';\bR)\subset H^1(\Sigma_g^{00};\bR) \ .
1106: \end{split}
1107: \end{equation}
1108: By combining the equations~\eqref{eq:difference}, \eqref{eq:[]},
1109: \eqref{eq:fla} and~\eqref{eq:Fluxc'}, we shall prove
1110: \begin{equation}\label{eq:nutau}
1111: [(\tilde\varph\inv)^*\nu-\nu]=
1112: a k(\tilde\varph)=a k(\varph)=a PD\circ C \tau(\varph) \ ,
1113: \end{equation}
1114: where $PD$ denotes the Poincar\'e duality isomorphism.
1115: 
1116: Clearly $(\tilde\varph\inv)^*\nu-\nu$ is a closed $1$-form on $\Sg$ whose
1117: support is contained in a neighborhood of $C_1\cup C_2$.
1118: Hence the Poincar\'e dual of the de Rham cohomology class
1119: $[(\tilde\varph\inv)^*\nu-\nu]\in H^1(\Sg;\bR)$ is a
1120: multiple of the homology class
1121: $[C_1]\in H_1(\Sg;\bZ)$ represented by the simple closed curve
1122: $C_1$ with a fixed orientation depicted in Figure~\ref{fig:torelli}.
1123: However, according to Johnson~\cite{Johnson80}, we have
1124: $$
1125: \tau(\varph)=(x_1\land y_1+\ldots +x_{g-1}\land y_{g-1})\land [C_1]
1126: $$
1127: where $x_1. y_1,\ldots ,x_{g-1}, y_{g-1}$ is a symplectic basis
1128: of the homology group of the left hand subsurface of $\Sg$
1129: obtained by cutting $\Sg$ along the simple closed curve $F_2$
1130: depicted in Figure~\ref{fig:torelli}. It follows that
1131: $$
1132: C \tau(\varph)=2(g-1)[C_1] \ .
1133: $$
1134: This checks the equality~\eqref{eq:nutau}, for some constant $a$.
1135: 
1136: To determine this constant, it is enough to compute the
1137: value of the cohomology class $[(\tilde\varph\inv)^*\nu-\nu]$
1138: on the homology class represented by the oriented simple
1139: closed curve $E$ depicted in Figure~\ref{fig:torelli}.
1140: Observe that the homology class
1141: $\tilde\varph\inv_*[E]-[E]\in H_1(\Sigma_g^{00};\bZ)$
1142: can be represented by the oriented simple closed curve
1143: $F_1$ also depicted in Figure~\ref{fig:torelli}.
1144: Let $\widetilde D$ denote the right hand compact subsurface
1145: of $\Sg$ obtained by cutting along $F_1$.
1146: Thus $\widetilde D$ is diffeomorphic to a disk which contains
1147: the original embedded disk $D$ in its interior.
1148: Also let $\Sigma =\Sg\setminus\textrm{Int}\tilde D$.
1149: 
1150: Now we compute
1151: \begin{align*}
1152: [(\tilde\varph\inv)_*\nu-\nu]([E])&=\nu((\tilde\varph\inv)_*[E])-\nu([E])\\
1153: &=\nu([F_1])
1154: =\int_{F_1} \la-\la'\\
1155: &=\int_{\partial \Sigma}\la-\int_{-\partial \widetilde D}\la'\\
1156: &=\int_{\Sigma} d\la +\int_{\widetilde D}d\la'\\
1157: &=\int_{\Sg} -\om =2-2g \ .
1158: \end{align*}
1159: Thus we can conclude that $a=-1$, completing the proof.
1160: \end{proof}
1161: 
1162: \section{The second homology group of
1163: $\Symp^{c,\delta}\Sigma_g^0$}\label{s:secondopen}
1164: 
1165: In this section, we generalize the claim of Theorem~\ref{th:hev}
1166: about the second homology, i.~e.~for $k=1$, to the case of the
1167: open surface $\Sigma_g^0$. We have the cohomology class
1168: $$
1169: j^*\tilde \al\in H^2(\Symp^{c,\delta}\Sigma_g^0;S^2_\bQ\bR)
1170: $$
1171: induced by the inclusion $j\colon\Sigma_g^0\ra\Sg$.
1172: \begin{theorem}\label{th:h20}
1173: The characteristic classes $e_1$ and $j^*\tilde \al$ induce a surjective
1174: homomorphism
1175: $$
1176: H_2(\Symp^{c,\delta}\Sigma_g^0;\bZ)\lra \bZ\oplus S^2_\bQ\bR
1177: $$
1178: for any $g\geq 3$. For $g=2$, the class $j^*\tilde \al$ induces
1179: a surjection
1180: $$
1181: H_2(\Symp^{c,\delta}\Sigma_2^0;\bZ)\lra S^2_\bQ\bR \ .
1182: $$
1183: \end{theorem}
1184: The proof of this theorem, which occupies the rest of this section,
1185: consists of a rather long and delicate argument. We first describe the
1186: reason why the easy proof of Theorem~\ref{th:hev} in the case $k=1$,
1187: which treated the case of
1188: closed surfaces, does not work for Theorem~\ref{th:h20}, thereby
1189: making the difficulty in the case of open surfaces explicit.
1190: 
1191: Consider the top extension
1192: \begin{equation}\label{eq:scm}
1193: 1 \lra \Symp^{c}_{0}\Sigma_g^{0} \overset{i^c}{\lra}
1194: \Symp^c\Sigma_g^{0} \overset{p}{\lra} \M1g \lra 1
1195: \end{equation}
1196: in the commutative diagram~\eqref{eqn:GG}. We have the following proposition,
1197: contrasting with our discussion of the closed case in Section~\ref{s:main}.
1198: \begin{proposition}\label{prop:fl2c}
1199: For any $g\geq 2$ the image of the homomorphism
1200: $$
1201: (j^*\Flux)_*\colon H_2(\Symp_0^{c,\delta}\Sigma_g^0;\bZ)\lra
1202: H_2(H^1(\Sg;\bR)^\delta;\bZ)\cong \La^2_{\bZ}H^1(\Sg;\bR)
1203: $$
1204: induced by the restriction $j^*\Flux$ of the flux homomorphism
1205: to the subgroup $\Symp_0^{c}\Sigma_g^0$
1206: is equal to the kernel of the natural intersection pairing
1207: $$
1208: \La^2_{\bZ}H^1(\Sg;\bR)\lra \bR \ .
1209: $$
1210: \end{proposition}
1211: \begin{proof}
1212: Recall that we have an extension
1213: \begin{equation}\label{eq:hcss}
1214: 1\lra\Ham^c\Sigma_g^0\lra\Symp_0^c\Sigma_g^0\ \overset{j^*\Flux}{\lra}\
1215: H^1(\Sg;\bR)\lra 1 \ ,
1216: \end{equation}
1217: where $\Ham^c\Sigma_g^0$ denotes the subgroup consisting of
1218: Hamiltonian symplectomorphisms with compact supports
1219: (see~\cite{MS} as well as~\cite{KM03}).
1220: Also recall that there is a surjective homomorphism
1221: $$
1222: {\rm Cal}\colon \Ham^c \Sigma_g^0 \lra\bR
1223: $$
1224: called the (second) Calabi homomorphism (see~\cite{Calabi}).
1225: Banyaga~\cite{Banyaga} proved that the kernel of this
1226: homomorphism is perfect. Hence we have an isomorphism
1227: $$
1228: H_1(\Ham^{c,\delta}\Sigma_g^0;\bZ)\cong \bR \ .
1229: $$
1230: Now we consider the Hochschild--Serre exact sequence
1231: \begin{align*}
1232: H_2(\Symp_0^{c,\delta}\Sigma_g^0;\bZ)&\ \overset{j^*\Flux_*}{\lra}\
1233: H_2(H^1(\Sg;\bR)^\delta;\bZ)\overset{\partial}{\lra}
1234: H_1(\Ham^{c,\delta}\Sigma_g^0;\bZ)_{H^1_{\bR}}\cong\bR\\
1235: &\lra H_1(\Symp_0^{c,\delta}\Sigma_g^0;\bZ)\
1236: \overset{j^*\Flux_*}{\lra}\
1237: H_1(H^1(\Sg;\bR)^\delta;\bZ)\lra 0
1238: \end{align*}
1239: of the group extension~\eqref{eq:hcss}. We proved in~\cite{KM03}
1240: (Proposition~11 and Corollary~12) that the last homomorphism
1241: $j^*\Flux_*$ in the above sequence is an isomorphism, and that
1242: the boundary operator $\partial$ coincides with the intersection
1243: pairing. The result follows.
1244: \end{proof}
1245: 
1246: \begin{corollary}\label{cor:trivial}
1247: The restriction of $\al\in H^2(\Symp^\delta\Sg;\bR)$ to the subgroup
1248: $\Symp_0^{c,\delta}\Sigma_g^0$ is trivial.
1249: \end{corollary}
1250: \begin{proof}
1251: This follows from Proposition~\ref{prop:fl2c} and the definition of
1252: the cohomology class $\al$.
1253: \end{proof}
1254: Thus, in order to prove the non-triviality of $\al$ on the group
1255: $\Symp^c\Sigma_g^0$, we must combine the roles of the two groups $\M1g$
1256: and $\Symp_0^{c,\delta}\Sigma_g^0$. This contrasts sharply
1257: with the case of closed surfaces treated in Section~\ref{s:main}.
1258: 
1259: Let $\{E_{p,q}^r\}$ be the Hochschild--Serre spectral sequence for the
1260: integral homology of the extension~\eqref{eq:scm}.
1261: This gives rise to two short exact sequences
1262: \begin{equation}\label{eq:ss}
1263: \begin{split}
1264: 0\lra \Ker& \lra H_2(\Symp^{c,\delta}\Sigma_g^0;\bZ)\lra
1265: E^\infty_{2,0} \lra 0\\
1266: & 0\lra E^\infty_{0,2}\lra\Ker \lra E^\infty_{1,1}\lra 0 \ ,
1267: \end{split}
1268: \end{equation}
1269: where $E^\infty_{2,0} \subset H_2(\M1g;\bZ)$ concerns the first
1270: Mumford-Morita-Miller class already discussed in~\cite{KM03}.
1271: Proposition~\ref{prop:fl2c} shows that the image of the map
1272: $$
1273: E^\infty_{0,2}=\im\left(H_2(\Symp_0^{c,\delta}\Sigma_g^0;\bZ)
1274: \ra H_2(\Symp^{c,\delta}\Sigma_g^0;\bZ)\right)\lra S^2_\bQ\bR
1275: $$
1276: defined by the Kronecker product with $\tilde\al$ is
1277: precisely the kernel of the natural map $S^2_\bQ\bR\ra\bR$.
1278: It remains to determine the $E^\infty_{1,1}$-term.
1279: We have
1280: \begin{equation}\label{eq:E1}
1281: E^2_{1,1}=H_1(\M1g;H_1(\Symp_0^{c,\delta}\Sigma_g^0;\bZ))
1282: \cong H_1(\M1g;H_1(\Sg;\bR))
1283: \end{equation}
1284: because, as was already mentioned above, it was proved in~\cite{KM03}
1285: that $j^*\Flux$ induces an isomorphism
1286: $H_1(\Symp_0^{c,\delta}\Sigma_g^0;\bZ)\cong H^1(\Sg;\bR)$.
1287: 
1288: We now recall a result which was essentially proved in~\cite{Morita89}.
1289: %Here we describe it because we need a precise form of the result.
1290: Without repeating everything done in~\cite{Morita89}, we  want to give
1291: a precise statement and proof of what is needed in the sequel.
1292: We use the Lickorish generators for the mapping class group $\M1g$,
1293: denoted $\la_i, \mu_i, \nu_i$ as in Figure~1 of~\cite{Morita89}, and a
1294: symplectic basis $x_1,\cdots,x_g,y_1,\cdots,y_g$ of $H_1(\Sg;\bZ)$.
1295: In particular
1296: \begin{equation}\label{eq:action}
1297: \nu_i(x_i)=x_i-y_i+y_{i+1},\quad \mu_i(x_i)=x_i-y_i
1298: \end{equation}
1299: which we record here for later use.
1300: \begin{proposition}\label{prop:h1m1}
1301: For any $g\geq 2$, we have an isomorphim
1302: $$
1303: H_1(\M1g;H_1(\Sg;\bZ))\cong \bZ \ .
1304: $$
1305: Furthermore, for any $i=1,2,\cdots,g-1$, the element
1306: $$
1307: c=\nu_i\otimes x_i-\mu_i\otimes x_i+\mu_{i+1}\otimes x_{i+1}
1308: $$
1309: is a $1$-cycle of $\M1g$ with twisted coefficients in $H_1(\Sg;\bZ)$
1310: and it represents a generator of the above infinite cyclic group.
1311: \end{proposition}
1312: \begin{proof}
1313: The group extension
1314: $$
1315: 1\lra \pi_1\Sg\lra \Msg\lra \Mg\lra 1 \ ,
1316: $$
1317: where $\Msg$ denotes the mapping class group of $\Sg$ {\it relative}
1318: to a basis point, yields the Hochschild--Serre exact sequence
1319: \begin{equation}\label{eq:msg}
1320: H_2(\Mg;H)\lra (H\otimes H)_{\Mg}\lra H_1(\Msg;H)
1321: \lra H_1(\Mg;H)\lra 0 \ .
1322: \end{equation}
1323: Here and henceforth $H$ is a shorthand for $H_1(\Sg;\bZ)$.
1324: It is easy to see that the intersection pairing induces an isomorphism
1325: $(H\otimes H)_{\Mg}\cong\bZ$ and the element $x_1\otimes y_1$, for
1326: example, represents a generator. It was proved in~\cite{Morita89}
1327: that
1328: $$
1329: H^1(\Msg;H^1(\Sg;\bZ))\cong Z,\quad
1330: H_1(\Mg;H^1(\Sg;\bZ))\cong \bZ/(2g-2)\bZ \ .
1331: $$
1332: If we apply the crossed homomorphism $f\colon\Msg\ra H^1(\Sg;\bZ)$
1333: given in the above cited paper, which detects a generator of the
1334: above infinite cyclic group, to the element $x_1\otimes y_1$
1335: considered as a cycle of $\Msg$ with coefficients in $H$, we obtain
1336: $$
1337: f(x_1)(y_1)=2-2g \ .
1338: $$
1339: On the other hand, $c$ is a cycle because
1340: $$
1341: \partial c=x_i-\nu_i(x_i)-x_i+\mu_i(x_i)+x_{i+1}-\mu_{i+1}(x_{i+1})
1342: =0
1343: $$
1344: by~\eqref{eq:action}, and it was shown that
1345: \begin{equation}\label{eq:k}
1346: f(c)=f(\nu_i)(x_i)-f(\mu_i)(x_i)+f(\mu_{i+1})(x_{i+1})=1 \ ,
1347: \end{equation}
1348: see~\cite{Morita89} for details. In view of the exact sequence~\eqref{eq:msg},
1349: we can conclude that $H_1(\Msg;H)\cong \bZ$. Finally, it is easy to deduce
1350: from the central group extension $0\ra \bZ\ra \M1g\ra\Msg\ra 1$ that we have an
1351: isomorphism $H_1(\M1g;H)\cong H_1(\Msg;H)$. This finishes the proof.
1352: \end{proof}
1353: 
1354: Going back to the $E^2_{1,1}$-term in~\eqref{eq:E1}, we have
1355: $$
1356: E^2_{1,1}\cong H_1(\M1g;H\otimes\bR)\cong\bR
1357: $$
1358: by Proposition~\ref{prop:h1m1}. Now consider the exact sequence
1359: $$
1360: E^2_{3,0}\cong H_3(\M1g;\bZ)\ \overset{d^2}{\lra} \
1361: E^2_{1,1}\cong \bR\lra E^3_{1,1}=E^\infty_{1,1}\lra 0 \ .
1362: $$
1363: Harer~\cite{Harer92} determined the third stable
1364: rational cohomology group
1365: $$
1366: \lim_{g\to\infty} H^3(\M1g;\bQ)
1367: $$
1368: of the mapping class group to be trivial\footnote{There is now a final
1369: result on the stable cohomology of $\Mg$ due to Madsen and Weiss~\cite{MW}.}.
1370: It follows that $E^2_{3,0}$ is a finite group for all sufficiently large $g$.
1371: Hence we can conclude that $E^\infty_{1,1}\cong \bR$ for such $g$.
1372: It is natural to expect that this $\bR$ will recover the missing $\bR$ in
1373: $E^\infty_{0,2}$ so that we obtain the surjectivity of the $\tilde\al$-factor
1374: in Theorem~\ref{th:h20}. It turns out that this is indeed the case, and below
1375: we shall give a proof of this fact which does not use Harer's result mentioned
1376: above. Before doing so, we have to prepare some general facts concerning
1377: group (co)homology in small degrees.
1378: 
1379: Consider a group extension
1380: \begin{equation}\label{eq:ext}
1381: 1\lra K\lra G\lra Q\lra 1 \ ,
1382: \end{equation}
1383: and
1384: %let $\{E^r_{p,q}\}$ be the Hochschild--Serre spectral
1385: %sequence for the integral homology of $G$.
1386: suppose we are given a $1$-cycle $c=\sum_i q_i\otimes u_i\in Z_1(Q;H_1(K))$
1387: of the group $Q$ with coefficients in the abelianization $H_1(K)$ of $K$
1388: considered as a natural $Q$-module, where $q_i\in Q$ and $u_i\in H_1(K)$.
1389: \begin{lemma}\label{lem:cc}
1390: For any choices of lifts $\tilde q_i\in G$ of the $q_i$
1391: and representatives $k_i\in K$ with $[k_i]=u_i$, the element
1392: $$
1393: \tilde c=\sum_{i} \left\{(\tilde q_i, k_i)+(\tilde q_i k_i,\tilde q_i^{-1})
1394: -(\tilde q_i,\tilde q_i^{-1})-(\id,\id)\right\}+ d
1395: $$
1396: is a $2$-cycle of $G$, where $d$ is a $2$-chain of the group $K$ such that
1397: $$
1398: \partial d=\sum_i\left\{(\tilde q_i k_i\tilde q_i^{-1})-(k_i)\right\}
1399: \ .
1400: $$
1401: Furthermore, $p_*([\tilde c])=0\in H_2(Q;\bZ)$, where $p\colon G\ra Q$ denotes
1402: the projection, and in the short exact sequence
1403: \begin{align*}
1404: 0\lra &E^\infty_{0,2}\lra
1405: \Ker\left(H_2(G;\bZ)\overset{p_*}{\lra} H_2(Q;\bZ)\right)\\
1406: & \lra E^\infty_{1,1}\
1407: \left(\cong H_1(Q;H_1(K))/d^2(E^2_{3,0})\right)\lra 0
1408: \end{align*}
1409: arising from the Hochschild--Serre spectral sequence of~\eqref{eq:ext} the
1410: class $[\tilde c]\in \Ker$ is a lift of $[c]\in H_1(Q;H_1(K))$.
1411: \end{lemma}
1412: \begin{proof}
1413: Since $c$ is a cycle by the assumption, we have
1414: $$
1415: \partial c=\sum_i (q_i(u_i)-u_i)=0 \in H_1(K) \ .
1416: $$
1417: It follows that there exists a $2$-chain $d\in C_2(K;\bZ)$
1418: with the property described in the statement of the lemma.
1419: Then a direct computation shows that $\partial \tilde c=0$.
1420: Clearly $p_*(\tilde c)=0$ and it is easy to check
1421: the rest of the required assertions.
1422: \end{proof}
1423: 
1424: The following can be proved by a standard argument in the cohomology theory
1425: of groups, see~\cite{Brown}.
1426: \begin{lemma}\label{lem:ch}
1427: Let $G$ be a group and $M$ a $G$-module. Assume we have a
1428: $G$-invariant skew-symmetric bilinear pairing
1429: $$
1430: \iota\colon M\times M\lra A \ ,
1431: $$
1432: where $A$ is an abelian group with trivial $G$-action, and
1433: we are given two crossed homomorphisms
1434: $$
1435: f_i\colon G\lra M\quad (i=1,2)
1436: $$
1437: so that $f_i(gh)=f_i(g)+ g_* f_i(h)$ for all $g,h\in G$.
1438: Then the assignment
1439: $$
1440: G\times G\ni (g,h)\longmapsto\iota(f_1(g),g_*f_2(h))\in A \ ,
1441: $$
1442: which we denote by $f_1\cdot f_2$, is a $2$-cocycle of $G$ with values
1443: in $A$ and its cohomology class in $H^2(G;A)$ depends only on the cohomology
1444: classes $[f_i]\in H^1(G;M)$ of the crossed homomorphisms $f_i$.
1445: Furthermore, $f_2\cdot f_1$ is cohomologous to $f_1\cdot f_2$,
1446: so that $[f_2\cdot f_1]=[f_1\cdot f_2]\in H^2(G;A)$.
1447: \end{lemma}
1448: %\begin{proof}
1449: %In fact, the assignment
1450: %$(g,h)\mapsto f_1(g)\otimes g_*f_2(h)\in M\otimes M$ is nothing
1451: %other than the Alexander-Whitney cup product $f_1\cup f_2$ of the
1452: %two $1$-cocycles $f_i$ of $G$ with values in $M$. The cocycle
1453: %$f_1\cdot f_2$ is obtained from this by applying the
1454: %$G$-invariant skew-symmetric pairing $M\otimes M\ra A$
1455: %to the usual cup product $[f_1][f_2]$,
1456: %namely $[f_1\cdot f_2]=\iota_*([f_1][f_2])$, so that
1457: %its cohomology class depends only on those of $f_i$.
1458: %As is well-known, $[f_2][f_1]=- [f_1][f_2]$ under the
1459: %isomorphism $t$ of $M\otimes M$ given by $t(m,n)=(n,m)$.
1460: %Since $\iota$ is skew-symmetric by assumption, we obtain
1461: %the equality $[f_2\cdot f_1]=[f_1\cdot f_2]$ as required.
1462: %This completes the proof.
1463: %\end{proof}
1464: 
1465: Now in the situation of Lemma~\ref{lem:ch}, we consider the case where
1466: $G=\Symp^c\Sigma_g^0$, $M=H^1(\Sg;\bR)$ and $\iota$ is the intersection
1467: pairing. Then we have three crossed homomorphisms
1468: $$
1469: j^*\widetilde{\Flux},\ \widetilde{\Flux}_c,\ p^*k_\bR:
1470: \Symp^c\Sigma_g^0\lra H^1(\Sg;\bR)
1471: $$
1472: and, by the definition of $\al$, we have
1473: $j^*\al=[j^*\widetilde{\Flux}\cdot j^*\widetilde{\Flux}]$.
1474: \begin{proposition}\label{prop:fluxc}
1475: In the above notation, we have
1476: $[\widetilde{\Flux}_c\cdot\widetilde{\Flux}_c]=0$
1477: and
1478: $$
1479: j^*\al=2 [p^*k_\bR\cdot\widetilde{\Flux}_c]-p^* e_1 \ .
1480: $$
1481: \end{proposition}
1482: \begin{proof}
1483: Define a map
1484: $$
1485: \widetilde{\Cal}\colon\Symp^c\Sigma_g^0\lra \bR
1486: $$
1487: by the formula
1488: $$
1489: \widetilde{\Cal}(\varph)=\int_{\Sigma_g^0}(\varph\inv)^*\la\land \la
1490: %\quad (\varph\in \Symp^c\Sigma_g^0).
1491: \ .
1492: $$
1493: The restriction of this map to the subgroup $\Ham^c\Sigma_g^0$ is a
1494: homomorphism, which, suitably normalized, is called the (second) Calabi
1495: homomorphism (see~\cite{Calabi}). In our previous paper~\cite{KM03}, we
1496: examined how the restriction of $\widetilde{\Cal}$ to the subgroup
1497: $\Symp^c_0\Sigma_g^0$ fails to be a homomorphism. Here we extend this
1498: discussion to the whole group $\Symp^c\Sigma_g^0$. For any two elements
1499: $\varph, \psi\in \Symp^c\Sigma_g^0$, we claim that
1500: \begin{equation}\label{eq:cal}
1501: \widetilde{\Cal}(\varph\psi)
1502: = \widetilde{\Cal}(\varph)+\widetilde{\Cal}(\psi)+
1503: \widetilde{\Flux}_c(\varph)\cdot (\varph\inv)^*\widetilde{\Flux}_c(\psi)
1504: \ .
1505: \end{equation}
1506: This is because
1507: \begin{align*}
1508: \widetilde{\Cal}(\varph\psi)=&\int_{\Sigma_g^0}
1509: ((\varph\psi)\inv)^*\la\land\la\\
1510: =& \int_{\Sigma_g^0} (\varph\inv)^*((\psi\inv)^*\la-\la)\land\la
1511: +(\varph\inv)^*\la\land\la\\
1512: =& \int_{\Sigma_g^0}(\varph\inv)^*\widetilde{\Flux}_c(\psi)\land
1513: ((\varph\inv)^*\la-\widetilde{\Flux}_c(\varph))+\widetilde{\Cal}(\varph)\\
1514: =& \widetilde{\Flux}_c(\varph)\cdot (\varph\inv)^*\widetilde{\Flux}_c(\psi)
1515: +\int_{\Sigma_g^0} ((\psi\inv)^*\la-\la)\land\la +\widetilde{\Cal}(\varph)\\
1516: =& \widetilde{\Flux}_c(\varph)\cdot (\varph\inv)^*\widetilde{\Flux}_c(\psi)
1517: +\widetilde{\Cal}(\psi)+\widetilde{\Cal}(\varph) \ .
1518: \end{align*}
1519: Equation~\eqref{eq:cal} implies that the $2$-cocycle
1520: $\widetilde{\Flux}_c\cdot \widetilde{\Flux}_c$ of the group $\Symp^c\Sigma_g^0$
1521: is a coboundary. Hence $[\widetilde{\Flux}_c\cdot
1522: \widetilde{\Flux}_c]=0$ as claimed.
1523: 
1524: The definition of $\al$ implies
1525: $$
1526: j^*\al= [j^*\widetilde{\Flux}\cdot j^*\widetilde{\Flux}]
1527: $$
1528: so that, by Theorem~\ref{th:ffk}, we can write
1529: \begin{equation}\label{eq:all}
1530: j^*\al=
1531: [\widetilde{\Flux}_c\cdot\widetilde{\Flux}_c]+[\widetilde{\Flux}_c\cdot
1532: p^*k_\bR]+[p^*k_\bR\cdot \widetilde{\Flux}_c]+p^*[k_\bR\cdot k_\bR] \ .
1533: \end{equation}
1534: By Lemma~\ref{lem:ch} we have
1535: $[\widetilde{\Flux}_c\cdot p^*k_\bR]=[p^*k_\bR\cdot
1536: \widetilde{\Flux}_c]$,
1537: and it was proved in~\cite{Morita89b} that $[k_\bR\cdot k_\bR]=-e_1$.
1538: If we substitute these relations in \eqref{eq:all}, we obtain the desired
1539: identity.
1540: \end{proof}
1541: 
1542: \begin{proof}[Proof of Theorem~\ref{th:h20}]
1543: In view of Proposition~\ref{prop:fl2c} and the discussion following
1544: Corollary~\ref{cor:trivial}, it suffices to show that there exist $2$-cycles
1545: of the group $\Symp^c\Sigma_g^0$ such that the evaluations of $j^*\al$ on
1546: them have as values any real number. To construct such $2$-cycles, we use
1547: the $1$-cycle $c$ of $\M1g$ with coefficients in $H$ described in
1548: Proposition~\ref{prop:h1m1}, which represents a generator of 
1549: $H_1(\M1g;H)\cong\bZ$.
1550: In order to adapt to the situation here, we consider the dual cycle
1551: $$
1552: c^*=\nu_i\otimes y^*_i-\mu_i\otimes y^*_i+\mu_{i+1}\otimes y^*_{i+1}
1553: $$
1554: with coefficients in $H^1(\Sg;\bZ)$, where the
1555: $x^*_1,\cdots,x^*_g,y^*_1,\cdots,y^*_g$ denote the dual basis of 
1556: $H^1(\Sg;\bZ)$.
1557: If $k\colon\M1g\ra H^1(\Sg;\bZ)$ is a crossed homomorphism which represents
1558: a generator of $H^1(\M1g;H^1(\Sg;\bZ))$, then we have
1559: \begin{equation}\label{eq:kc}
1560: k(c^*)=k(\nu_i)\cdot y^*_i-k(\mu_i)\cdot y^*_i+k(\mu_{i+1})\cdot
1561: y^*_{i+1}=1 \ .
1562: \end{equation}
1563: For any real number $r\in\bR$, we consider the $1$-cycle
1564: $$
1565: c^*_r=\nu_i\otimes r y^*_i-\mu_i\otimes r y^*_i+\mu_{i+1}\otimes r y^*_{i+1}
1566: $$
1567: of $\M1g$ with coefficients in $H^1(\Sg;\bR)$.
1568: Now we apply Lemma~\ref{lem:cc} in the case where
1569: $G=\Symp^c\Sigma_g^0$, $K=\Symp^c_0\Sigma_g^0$, and $Q=\M1g$.
1570: We proved in~\cite{KM03} that the flux homomorphism induces an isomorphism
1571: $$
1572: \Flux\colon H_1(\Symp^c_0\Sigma_g^0;\bZ)\cong H^1(\Sg;\bR).
1573: $$
1574: Therefore,
1575: $c^*_r$ can be considered as a $1$-cycle of $\M1g$ with coefficients
1576: in the abelianization of $\Symp^c_0\Sigma_g^0$.
1577: Hence, if we choose elements
1578: $$
1579: \tilde{\nu}_i, \tilde{\mu}_i\in \Symp^c\Sigma_g^0,
1580: \quad \varph^r_i\in \Symp^c_0\Sigma_g^0
1581: $$
1582: such that
1583: \begin{equation}\label{eq:kf}
1584: p(\tilde{\nu}_i)=\nu_i,\quad p(\tilde{\mu}_i)=\mu_i,
1585: \quad \Flux(\varph^r_i)= \Flux_c(\varph^r_i)=r y^*_i \ ,
1586: \end{equation}
1587: where $p\colon\Symp^c\Sigma_g^0\ra\M1g$ denotes the projection as before,
1588: then
1589: \begin{align*}
1590: \tilde{c}^*_r=& (\tilde{\nu}_i, \varph^r_i)+
1591: (\tilde{\nu}_i\varph^r_i,\tilde{\nu}^{-1}_i)-
1592: (\tilde{\nu}_i,\tilde{\nu}^{-1}_i)-(\id,\id)+\\
1593: & (\tilde{\mu}_i, \varph^r_i)+
1594: (\tilde{\mu}_i\varph^r_i,\tilde{\mu}^{-1}_i)-
1595: (\tilde{\mu}_i,\tilde{\mu}^{-1}_i)-(\id,\id)+\\
1596: &  (\tilde{\mu}_{i+1}, \varph^r_{i+1})+
1597: (\tilde{\mu}_{i+1}\varph^r_{i+1},\tilde{\mu}^{-1}_{i+1})-
1598: (\tilde{\mu}_{i+1},\tilde{\mu}^{-1}_{i+1})-(\id,\id)+d
1599: \end{align*}
1600: is a $2$-cycle of $\Symp^c\Sigma_g^0$, where $d$ is a $2$-chain of the
1601: group $\Symp^c_0\Sigma_g^0$ such that
1602: $$
1603: \partial d=(\tilde{\nu}_i\varph^r_i\tilde{\nu}^{-1}_i)-(\varph^r_i)
1604: +(\tilde{\mu}_i\varph^r_i\tilde{\mu}^{-1}_i)-(\varph^r_i)
1605: +(\tilde{\mu}_{i+1}\varph^r_{i+1}\tilde{\mu}^{-1}_{i+1})-(\varph^r_{i+1}) \ .
1606: $$
1607: Now we claim that
1608: \begin{equation}\label{eq:cr}
1609: j^*\al(\tilde{c}^*_r)=2r \ ,
1610: \end{equation}
1611: which will finish the proof of the theorem.
1612: To show this, observe first that $p^*e_1(\tilde{c}^*_r)=0$ because
1613: clearly $p_*(\tilde{c}^*_r)=0$. Hence, by Proposition~\ref{prop:fluxc},
1614: $$
1615: j^*\al(\tilde{c}^*_r)=2 [p^*k\cdot\widetilde{\Flux}_c](\tilde{c}^*_r)
1616: \ .
1617: $$
1618: Observe that
1619: $$
1620: [p^*k\cdot\widetilde{\Flux}_c]\left((\tilde{\nu}_i\varph^r_i,
1621: \tilde{\nu}^{-1}_i)-(\tilde{\nu}_i,\tilde{\nu}^{-1}_i)\right)=0 \ ,
1622: $$
1623: because $p^*k(\tilde{\nu}_i\varph^r_i)=p^*k(\tilde{\nu}_i)$ and
1624: $\varph^r_i$ acts trivially on the homology of $\Sigma_g^0$.
1625: The same is true for two other similar terms. Since $d$ is a $2$-chain
1626: of $\Symp^c_0\Sigma_g^0$, the evaluation of $[p^*k\cdot\widetilde{\Flux}_c]$
1627: on it vanishes. Keeping in mind equations~\eqref{eq:kf} and~\eqref{eq:kc},
1628: we can now conclude that
1629: $$
1630: j^*\al(\tilde{c}^*_r)=2 [p^*k\cdot\widetilde{\Flux}_c](\tilde{c}^*_r)
1631: =2r p^*k(\tilde{c}^*)=2r \ .
1632: $$
1633: This proves~\eqref{eq:cr} and hence the theorem.
1634: \end{proof}
1635: 
1636: \section{Proof of the main results}\label{s:proofs}
1637: 
1638: In this section we give the proofs of the main results described
1639: in Section~\ref{s:main}.
1640: 
1641: \begin{proof}[Proof of Theorem~\ref{th:hev}]
1642: Here we follow the argument of~\cite{Morita2} and of our previous
1643: paper~\cite{KM03} to prove the non-triviality of the cup products
1644: $e_1^k\tilde\al^{(\ell)}$. For this we first observe that, similar
1645: to the class $e_1$, the class $\tilde\al$ is stable, with respect to
1646: $g$, and also that it is primitive in the following sense. For each
1647: $k$, consider the genus $kg$ surface
1648: $\Sigma_{kg,1}=\Sigma_{kg}\setminus \Int D^2$ with one boundary component
1649: as the boundary connected sum
1650: $$
1651: \Sigma_{kg,1}=\Sigma_{g,1}\ \natural \cdots\natural\ \Sigma_{g,1}
1652: $$
1653: of $k$ copies of $\Sigma_{g,1}=\Sg\setminus \Int D^2$.
1654: This induces a homomorphism
1655: \begin{equation}\label{eqn:fk}
1656: f_k\colon \Symp^c \Sigma_g^0\times\cdots\times \Symp^c \Sigma_g^0
1657: \lra \Symp^c \Sigma_{kg}^0
1658: \end{equation}
1659: from the direct product of $k$ copies of the group $\Symp^c \Sigma_g^0$ to
1660: $\Symp^c \Sigma_{kg}^0$.
1661: %Namely we set the supports of the images of the $k$-components of the
1662: %former group to be disjoint in the target group.
1663: Under this homomorphism we have the equality
1664: $$
1665: f_{k}^*(\tilde\al)=\tilde\al\times 1\times\cdots\times 1+\cdots
1666: +1\times\cdots\times 1\times \tilde\al \ ,
1667: $$
1668: which follows easily from the definition of $\tilde\al$. Now we can combine
1669: Theorem~\ref{th:h20} with the above property to show the required assertion
1670: in the theorem. This finishes the proof.
1671: \end{proof}
1672: 
1673: \begin{proof}[Proof of Theorem~\ref{th:om2}]
1674: The fact that the ideal generated by $\om_0\land H_1(T^{2g};\bR)$
1675: is contained in the kernel of $\Flux^*$ has already been proved in
1676: Proposition~\ref{prop:v}. To show that $\Ker \Flux^*$ is precisely
1677: this ideal, we use the decomposition~\eqref{eq:irrep} of $H^k(T^{2g};\bR)$
1678: into irreducible summands given in Section~\ref{s:main}. It is easy to
1679: see that the quotient of this module divided by the ideal generated by
1680: $\om_0\land H_1(T^{2g};\bR)\subset H_3(T^{2g};\bR)$ is precisely
1681: $$
1682: \bR\oplus \bigoplus_{k=1}^{g} [1^k]_\bR \ .
1683: $$
1684: The $\bR$-summand in degree $2$ corresponds to the class $\alpha$
1685: and its non-triviality has already been shown in Theorem~\ref{th:hev}.
1686: Hence, to prove the assertion, it remains to show that $\Flux^*([1^k])$
1687: is non-trivial for any $k\leq g$. The highest weight vector of the
1688: irreducible representation $[1^k]$ is $x_1\land\cdots\land x_k$, where
1689: $x_1,\cdots,x_g,y_1,\cdots,y_g$ is a symplectic basis of $H_1(\Sg;\bZ)$
1690: as before. Now the definition of the flux homomorphism implies that, for
1691: any $i$, there exists an element $\varph_i\in\Symp_0\Sg$ such that
1692: $\Flux(\varph_i)=x_i^*$. Here we can choose the support of $\varph_i$ to
1693: be contained in an arbitrarily small neighbourhood of a simple closed curve
1694: which represents the homology class $x_i$. Then the $k$ elements
1695: $\varph_1,\cdots,\varph_k$ mutually commute because their supports are
1696: disjoint. Hence they form a cycle of $\Symp_0\Sg$ supported on a
1697: $k$-dimensional torus, and the cohomology class
1698: $\Flux^*(x_1\land\cdots\land x_k)\in H^{k}(\Symp_0^\delta\Sg;\bR)$
1699: takes a non-zero value (namely $1$) on this cycle. This completes the proof.
1700: \end{proof}
1701: 
1702: \begin{proof}[Proof of Theorem~\ref{th:ham}]
1703: We consider the extension
1704: $$
1705: 1\lra \Ham\Sg\lra \Symp_0\Sg\overset{\Flux}{\lra}H^1(\Sg;\bR)\lra 1 \ .
1706: $$
1707: Let $\{E^r_{p,q}\}$ be the Hochschild--Serre spectral sequence for its
1708: homology. Since $\Ham\Sg$ is perfect by Thurston~\cite{Thurston0}, see
1709: also~\cite{Banyaga}, we have
1710: $$
1711: E^2_{1,1}\cong H_1(H^1(\Sg;\bR)^\delta;H_1(\Ham^\delta\Sg;\bZ))=0 \ .
1712: $$
1713: Hence the differential $d^2\colon E^2_{3,0}\ra E^2_{1,1}$ vanishes, so that
1714: $$
1715: E^3_{3,0}\cong E^2_{3,0}\cong H_3(H^1(\Sg;\bR)^\delta;\bZ)
1716: \cong \La_\bZ^3 H^1(\Sg;\bR) \ .
1717: $$
1718: Similarly $E^2_{2,1}=0$, so that the differential $d^2\colon E^2_{2,1}\ra 
1719: E^2_{0,2}$
1720: vanishes and
1721: $$
1722: E^3_{0,2}\cong E^2_{0,2}\cong
1723: H_0(H^1(\Sg;\bR)^\delta;H_2(\Ham^\delta\Sg;\bZ))
1724: \cong H_2(\Ham^\delta\Sg;\bZ)_{H^1_\bR} \ .
1725: $$
1726: However, $E^\infty_{3,0}=E^4_{3,0}$ is equal to the image of the homomorphism
1727: $$
1728: \Flux_*\colon H_3(\Symp_0^\delta\Sg;\bZ)\lra
1729: H_3(H^1(\Sg;\bR)^\delta;\bZ) \ .
1730: $$
1731: Now we can conclude that the exact sequence
1732: $$
1733: E^\infty_{3,0}=E^4_{3,0}\lra E^3_{3,0}\lra E^3_{0,2}=E^2_{0,2}
1734: $$
1735: yields an exact sequence
1736: \begin{equation}
1737: H_3(\Symp_0^\delta\Sg;\bZ)\lra H_3(H^1(\Sg;\bR)^\delta;\bZ)\lra
1738: H_2(\Ham^\delta\Sg;\bZ)_{H^1_\bR} \ .
1739: \label{eq:eh}
1740: \end{equation}
1741: By the same computation as above in the dual context of cohomology,
1742: we obtain an exact sequence
1743: \begin{equation}
1744: H^2(\Ham^\delta\Sg;\bR)^{H^1_\bR}\lra H^3(H^1(\Sg:\bR)^\delta;\bR)
1745: \overset{\Flux^*}{\lra} H^3(\Symp_0^\delta\Sg;\bR) \ .
1746: \label{eq:ech}
1747: \end{equation}
1748: Proposition~\ref{prop:v} (see also Theorem~\ref{th:om2})
1749: implies that the continuous cohomology classes
1750: $$
1751: \tilde\om_0\land H_1(\Sg;\bR)\subset H^3(H^1(\Sg:\bR)^\delta;\bR)
1752: $$
1753: vanish under the homomorphism $\Flux^*$ so that they can be
1754: lifted to elements of $H^2(\Ham^\delta\Sg;\bR)^{H^1_\bR}$
1755: by~\eqref{eq:ech}. Now consider the cycles
1756: $$
1757: \om_0 \land H^1(\Sg;\bR)\subset \La^3_\bZ H^1(\Sg;\bR)\cong
1758: H_3(H^1(\Sg;\bR)^\delta;\bZ) \ ,
1759: $$
1760: where $\om_0\in\La^2_\bZ H^1(\Sg;\bR)$, and also their images
1761: in $H_2(\Ham^\delta\Sg;\bZ)_{H^1_\bR}$ in the exact sequence~\eqref{eq:eh}.
1762: If we consider the Kronecker products of these cycles with the above lifted
1763: cohomology classes, we can conclude that $\om_0 \land H^1(\Sg;\bR)$ maps
1764: injectively into $H_2(\Ham^\delta\Sg;\bZ)_{H^1_\bR}$. This completes the proof.
1765: \end{proof}
1766: 
1767: \begin{remark}
1768: It would be interesting to obtain explicit group cocycles of $\Ham\Sg$ which
1769: represent the above degree two cohomology classes. There should also be a
1770: relation to the work of Ismagilov~\cite{Is}. We shall pursue this elsewhere.
1771: \end{remark}
1772: 
1773: \section{Further discussion}\label{s:misc}
1774: 
1775: As in Section~\ref{s:trans}, let $e, v\in H^2(\ES^{\delta}\Sg;\bR)$ be
1776: the Euler class and the transverse symplectic class, respectively.
1777: By analogy with the definition $e_1=\pi_*(e^2)$, where
1778: $$
1779: \pi_*\colon H^4(\ES^{\delta}\Sg;\bR)\lra H^2(\Symp^\delta\Sg;\bR)
1780: $$
1781: denotes the integration over the fiber, we can define a cohomology class
1782: $$
1783: v_1 \in H^2(\Symp^\delta\Sg;\bR)
1784: $$
1785: by setting $v_1=\pi_*(ev)$. After we conjectured that $v_1$ is a linear
1786: combination of the two classes $\al$ and $e_1$, Kawazumi~\cite{Kawazumi}
1787: kindly provided a proof. More precisely, he pointed out that the
1788: contraction formula, Theorem~6.2 of~\cite{KM01}, can be adapted to the
1789: case of the cohomology class $\al$ of the group $\ES^{\delta}\Sg$, and
1790: that the following equality holds:
1791: \begin{equation}
1792: \al=-\pi_*\left((e+v)^2\right)=-e_1-2 v_1 \ .
1793: \label{v1}
1794: \end{equation}
1795: %We now interpret this relation on the subgroup
1796: %$\Symp_0^\delta\Sg\subset \Symp^\delta\Sg$ from the point of view of the
1797: %results of Section~\ref{s:trans}.
1798: %Consider the following commutative diagram, where $\Msg$
1799: %denotes the mapping class group of $\Sg$ relative to a base point:
1800: %\begin{equation}
1801: %\begin{CD}
1802: %1@>>> \ES^\delta_0\Sg  @>{\tilde i}>> \ES^\delta\Sg @>>> \Msg @>>> 1 \\
1803: %@. @V{\pi}VV @V{\pi}VV @VVV  @.\\
1804: %1@>>> \Symp^\delta_{0}\Sg @>{i}>> \Symp^\delta\Sg @>>> \Mg @>>> 1 \ .
1805: %\end{CD}
1806: %\label{eqn:ES}
1807: %\end{equation}
1808: 
1809: %If we restrict $ev$ to the
1810: %subgroup $\ES^\delta_0\Sg$, then Proposition~\ref{prop:v} implies that
1811: %$$
1812: %\tilde i^*(ev)=\{(2-2g)\mu\}\{(2g-2)\mu+[\Flux]+\frac{1}{2g-2}\tilde\om_0\}
1813: %=-\mu \tilde\om_0 \ .
1814: %$$
1815: %Hence
1816: %\begin{equation}
1817: %\pi_*(\tilde i^*(ev))=-\tilde\om_0=-\frac{1}{2}i^*\al \ ,
1818: %\label{eq:ev}
1819: %\end{equation}
1820: %where the second equality follows from Proposition~\ref{prop:alom}.
1821: %By the naturality of the integration over the fiber, we also have
1822: %$$
1823: %\pi_*(\tilde i^*(ev))=i^*\pi_*(ev)=i^* v_1 \ ,
1824: %$$
1825: %so that we can conclude
1826: %$$
1827: %i^* v_1=-\frac{1}{2}i^*\al \ .
1828: %$$
1829: %Now the Hochschild--Serre exact sequence of the bottom extension
1830: %of~\eqref{eqn:ES} implies that
1831: %$
1832: %v_1=-\frac{1}{2} \al + c e_1
1833: %$
1834: %for some constant $c\in\bR$. Kawazumi's argument mentioned above
1835: %implies that the constant $c$ equals $-\frac{1}{2}$.
1836: 
1837: Since we know by~\cite{KM03} that $e^2\not=0$, we could also apply
1838: integration over the fiber to the cohomology class $e^2v$ in order to
1839: obtain some more cohomology. However, $e^2 v$ vanishes in
1840: $H^6(\ES^\delta\Sg;\bR)$ by the Bott vanishing theorem.
1841: %The class $e^2$ is equal to the first Pontrjagin class $p_1$
1842: %of the normal bundle of the transverse foliation on the foliated
1843: %surface bundle so that, at the level of a suitable associated
1844: %jet bundle, it is contained in the ideal generated by the canonical
1845: %forms $\theta^1, \theta^2$ with respect to the fiber $\Sg$.
1846: %On the other hand, $v$ is represented by a non-zero constant times
1847: %the closed $2$ form $\theta^1\theta^2$ which is the (pull back of
1848: %the) area from of $\Sg$. Hence the form representing $p_1 v$ vanishes
1849: %identically.
1850: 
1851: 
1852: %\subsection{Further classes}
1853: 
1854: %In this subsection, we introduce some further characteristic classes
1855: %for foliated surface bundles with area-preserving holonomy.
1856: 
1857: A more promising approach to find more cohomology for the
1858: symplectomorphism groups is the following.
1859: Consider the extension
1860: \begin{equation}
1861: 1\lra\Symp_0\Sg\lra\Symp\Sg\overset{p}{\lra}\Mg\lra 1 \ .
1862: \label{eq:ssm}
1863: \end{equation}
1864: On the one hand, Theorem~\ref{th:om2} shows that we have an injection
1865: \begin{equation}
1866: \bigoplus_{k=1}^g [1^k]_\bR \subset H^*(\Symp_0^\delta\Sg;\bR) \ .
1867: \label{eq:1k}
1868: \end{equation}
1869: On the other hand, Looijenga~\cite{Looijenga} determined the stable
1870: cohomology $H^*(\Mg;V)$ of the mapping class group with coefficients
1871: in any irreducible representation $V$ of the algebraic group $Sp(2g;\bQ)$.
1872: In particular, the cohomology groups $H^*(\Mg;[1^k])$ are highly non-trivial.
1873: In the spectral sequence for the cohomology of the extension~\eqref{eq:ssm},
1874: there are many non-trivial $E_2$-terms
1875: $$
1876: E^{p,q}_2=H^p(\Mg;H^q(\Symp_0^\delta\Sg;\bR)) \ .
1877: $$
1878: For example, it was proved in~\cite{Morita93} that 
1879: $H^1(\Mg;[1^3]_\bR)\cong\bR$,
1880: and an explicit computation using Looijenga's formula shows
1881: $H^2(\Mg;[1^2]_\bR)\cong\bR$. Hence we have injections
1882: \begin{align*}
1883: &\bR\subset E^{1,3}_2=H^1(\Mg;H^3(\Symp_0^\delta\Sg;\bR)) \\
1884: &\bR\subset E^{2,2}_2=H^2(\Mg;H^2(\Symp_0^\delta\Sg;\bR))
1885: \end{align*}
1886: for all sufficiently large $g$. It seems likely that these
1887: two copies of $\bR$ survive to the $E_\infty$ term, so that they
1888: define certain cohomology classes in $H^4(\Symp^\delta\Sg;\bR)$.
1889: More generally, the summands~\eqref{eq:1k} and the non-trivial
1890: cohomology groups $H^*(\Mg;[1^k]_\bR)$ should give rise to
1891: infinitely many cohomology classes in $H^*(\Symp^\delta\Sg;\bR)$.
1892: \begin{problem}
1893: Prove that these cohomology classes are non-trivial.
1894: \end{problem}
1895: 
1896: Next, there are completely different candidates for possible classes
1897: in $H^*(\Symp^\delta\Sg;\bR)$ coming from the cohomology of the Lie
1898: algebra $V_n$ of formal Hamiltonian vector fields on $\bR^{2n}$
1899: first studied by Gelfand, Kalinin and Fuks in~\cite{GKF}. This Lie
1900: algebra $V_n$ contains ${\mathfrak sp}(2n,\bR)$ as a subalgebra
1901: consisting of vector fields corresponding to linear symplectomorphisms.
1902: Let $\BG_2^\om$ be the Haefliger classifying space for the pseudogroup
1903: of local symplectomorphisms of $\bR^2$ with respect to the standard
1904: symplectic form. Then there is a natural homomorphism
1905: \begin{equation}
1906: H^*_c(V_1;{\frak sp}(2,\bR))\lra H^*(\BG_2^\om;\bR)
1907: \label{eq:vbg}
1908: \end{equation}
1909: from the continuous cohomology of $V_1$ relative to the subalgebra
1910: ${\frak sp}(2,\bR)$ to the real cohomology group of $\BG_2^\om$.
1911: 
1912: There is also an obvious continuous mapping
1913: $$
1914: K(\ES^\delta\Sg,1)\lra \BG_2^\om
1915: $$
1916: which classifies the transversely symplectic codimension $2$
1917: foliation on the classifying space for the group $\ES^\delta\Sg$,
1918: that is the total space of the universal foliated $\Sg$-bundle over
1919: $\BS^\delta\Sg$. This induces homomorphisms
1920: \begin{equation}
1921: H^*(\BG_2^\om;\bR)\lra H^*(\ES^\delta\Sg;\bR)
1922: \overset{\pi_*}{\lra} H^{*-2}(\Symp^\delta\Sg;\bR) \ ,
1923: \label{eq:bgs}
1924: \end{equation}
1925: where the last homomorphism is the integration along the fibre.
1926: Combining~\eqref{eq:vbg} and~\eqref{eq:bgs} we obtain a homomorphism
1927: \begin{equation}
1928: H^*_c(V_1;{\frak sp}(2,\bR))\lra H^{*-2}(\Symp^\delta\Sg;\bR) \ .
1929: \label{eq:}
1930: \end{equation}
1931: 
1932: Now, Gelfand, Kalinin and Fuks~\cite{GKF} found a new cohomology class
1933: in $H^7_c(V_1;{\frak sp}(2,\bR))$.
1934: %besides standard elements, e.g. the transverse
1935: %symplectic class as well as the first Pontrjagin class.
1936: Later, Metoki~\cite{Metoki} extended their computation and found another
1937: exotic class in $H^9_c(V_1;{\frak sp}(2,\bR))$. It seems to be widely
1938: believed that there should exist infinitely many such exotic classes.
1939: \begin{problem}
1940: Study the cohomology classes in $H^*(\Symp^\delta\Sg;\bR)$ induced from
1941: exotic classes in $H^{*+2}_c(V_1;{\frak sp}(2,\bR))$. In particular, prove
1942: that the two elements in $H^5(\Symp^\delta\Sg;\bR)$ and 
1943: $H^7(\Symp^\delta\Sg;\bR)$
1944: induced from the classes found by Gelfand, Kalinin and Fuks and by Metoki,
1945: are non-trivial.
1946: \end{problem}
1947: 
1948: Recall that Harer~\cite{Harer85} proved that the homology groups of the mapping
1949: class groups $\Mg$ stabilize with respect to the genus $g$ in a certain stable
1950: range. In view of the fact that all the characteristic classes we introduced in
1951: this paper are stable with respect to $g$, we would like to propose the 
1952: following
1953: problem, although it appears to be beyond the range of available techniques 
1954: at the
1955: moment:
1956: \begin{problem}
1957: Determine whether the homology groups of $\Symp^\delta \Sg$ stabilize
1958: with respect to $g$, or not. In particular, is it true that
1959: $$
1960: H_2(\Symp^\delta\Sg;\bZ)\cong \bZ\oplus S^2_\bQ\bR
1961: $$
1962: for all $g\geq 3$ ?
1963: \end{problem}
1964: 
1965: \section*{Appendix: Proof of Proposition~\ref{prop:la2}}
1966: 
1967: To prove Proposition~\ref{prop:la2}, observe first that the second exterior
1968: power $\La^2_{\bZ}H^1(\Sg;\bR)$ over $\bZ$ is naturally isomorphic to the same
1969: power $\La^2_{\bQ}H^1(\Sg;\bR)$ over $\bQ$ because $H^1(\Sg;\bR)$ is a uniquely
1970: divisible abelian group. Choose a Hamel basis $a_\la\ (\la\in A)$ of $\bR$ as a
1971: vector space over $\bQ$. Then we can write
1972: $$
1973: H^1(\Sg;\bR)=\sum_{\la} a_\la H^1(\Sg;\bQ) \ .
1974: $$
1975: Hence
1976: $$
1977: \La^2_{\bQ} H^1(\Sg;\bR)=\sum_{\la} a_\la \La^2_{\bQ}H^1(\Sg;\bQ)\oplus
1978: \sum_{\la<\mu} a_\la H^1(\Sg;\bQ)\otimes a_\mu H^1(\Sg;\bQ) \ ,
1979: $$
1980: where we choose a total order in the index set $A$. Clearly, this is a
1981: decomposition of $\Mg$-modules. It is easy to see that the intersection
1982: pairing gives rise to an isomorphism
1983: $$
1984: \left(\La^2_{\bQ}H^1(\Sg;\bQ)\right)_{\Mg}\cong \bQ \ .
1985: $$
1986: We also have
1987: %\begin{align*}
1988: $$
1989: \left(H^1(\Sg;\bQ)\otimes H^1(\Sg;\bQ)\right)_{\Mg}
1990: \cong
1991: \left(S^2H^1(\Sg;\bQ)\oplus \La^2H^1(\Sg;\bQ)\right)_{\Mg}\\
1992: \cong \bQ \ .
1993: $$
1994: %\end{align*}
1995: Here we have used the fact that
1996: $$
1997: \left(S^2H^1(\Sg;\bQ)\right)_{\Mg}=0 \ ,
1998: $$
1999: which is true because the action of $\Mg$ on $S^2H^1(\Sg;\bQ)$ factors
2000: through that of the algebraic group $Sp(2g,\bQ)$, and $S^2H^1(\Sg;\bQ)$
2001: is a non-trivial irreducible $Sp(2g,\bQ)$-module.
2002: 
2003: Thus we obtain an isomorphism
2004: $$
2005: \left(\La^2_{\bZ} H^1(\Sg;\bR)\right)_{\Mg}\cong
2006: \left(\sum_{\la} a_\la\otimes a_\la \bQ\right)\oplus
2007: \left(\sum_{\la<\mu} a_\la\otimes a_\mu\bQ\right) \ .
2008: $$
2009: It is easy to see that the right-hand side of the above expression
2010: can be naturally identified with $S^2_\bQ\bR$. This completes the proof.
2011: 
2012: 
2013: 
2014: \bibliographystyle{amsplain}
2015: 
2016: \begin{thebibliography}{999}
2017: 
2018: \bibitem{Banyaga}
2019: A.~Banyaga,
2020: {\it Sur la structure du groupe des diff\'eomorphismes qui
2021: pr\`eservent une forme symplectique},
2022: Comment.~Math.~Helv.~{\bf 53}
2023: (1978), 174--227.
2024: 
2025: \bibitem{Brown}
2026: K. S.  Brown,
2027: {\sl Cohomology of Groups},
2028: Graduate Texts in Mathematics vol.~87,
2029: Springer Verlag,
2030: 1982.
2031: 
2032: \bibitem{Calabi}
2033: E.~Calabi,
2034: {\it On the group of automorphisms of a symplectic manifold},
2035: in {\sl Problems in Analysis},
2036: ed.~R.~Gunning, Princeton University Press, 1970, 1--26.
2037: 
2038: \bibitem{FH}
2039: W. Fulton, J. Harris,
2040: {\sl Representation Theory},
2041: Graduate Texts in Mathematics vol.~129,
2042: Springer Verlag 1991.
2043: 
2044: \bibitem{GKF}
2045: I.~M.~Gelfand, D.~I.~Kalinin, D.~B.~Fuks,
2046: {\it The cohomology of the Lie algebra of Hamiltonian formal vector fields},
2047: (Russian) Funkcional.~Anal.~i Prilozen.~{\bf 6} (1972), 25--29;
2048: English trans.~in Functional Anal.~Appl.~{\bf 6} (1972), 193--196.
2049: 
2050: \bibitem{Harer85}
2051: J.~Harer, {\it Stability of the homology of the mapping class group of
2052: an orientable surface},
2053: Ann.~of Math.~{\bf 121} (1985), 215--249.
2054: 
2055: \bibitem{Harer92}
2056: J.~Harer,
2057: {\it The third homology group of the moduli space of curves},
2058: Duke Math.~J.~{\bf 63}
2059: (1992), 25--55.
2060: 
2061: \bibitem{Is}
2062: R.~S.~Ismagilov,
2063: {\it Inductive limits of area-preserving diffeomorphism groups},
2064: (Russian) Funktsional.~Anal.~i Prilozhen.~{\bf 37} (2003), 36--50, 95;
2065: Engl.~trans.~in Funct.~Anal.~Appl.~{\bf 37} (2003), 191--202.
2066:  
2067: \bibitem{Johnson80}
2068: D.~Johnson,
2069: {\it An abelian quotient of the mapping class group ${\mathcal I}_g$},
2070: Math.~Ann.~{\bf 249}
2071: (1980), 225--242.
2072: 
2073: \bibitem{Kawazumi}
2074: N.~Kawazumi,
2075: private communication.
2076: 
2077: \bibitem{KM96}
2078: N.~Kawazumi, S.~Morita,
2079: {\it The primary approximation to the cohomology of the moduli
2080: space of curves and cocycles for the stable characteristic classes},
2081: Math.~Research Letters {\bf 3}
2082: (1996), 629--641.
2083: 
2084: \bibitem{KM01}
2085: N.~Kawazumi, S.~Morita,
2086: {\it The primary approximation to the cohomology of the moduli
2087: space of curves and cocycles for the Mumford-Morita-Miller classes},
2088: preprint.
2089: 
2090: \bibitem{KM03}
2091: D.~Kotschick, S.~Morita,
2092: {\it Signatures of foliated surface bundles and the symplectomorphism
2093: groups of surfaces}, Topology {\bf 44} (2005), 131--149.
2094: 
2095: \bibitem{Looijenga}
2096: E.~Looijenga,
2097: {\it Stable cohomology of the mapping class group with symplectic
2098: coefficients and of the universal Abel-Jacobi map},
2099: J.~Algebraic Geometry {\bf 5}
2100: (1996), 135--150.
2101: 
2102: \bibitem{MW}
2103: I.~Madsen, M.~Weiss,
2104: {\it The stable moduli space of Riemann surfaces: Mumford's conjecture},
2105: preprint math.AT/0212321 v3 14Jul2004.
2106: 
2107: \bibitem{MS}
2108: D.~McDuff, D.~Salamon,
2109: {\sl Introduction to Symplectic Topology},
2110: second edition, Oxford University Press,
2111: 1998.
2112: 
2113: \bibitem{Metoki}
2114: S.~Metoki,
2115: {\sl Non-trivial cohomology classes of Lie algebras of volume
2116: preserving formal vector fields},
2117: thesis, University of Tokyo, 2000.
2118: 
2119: \bibitem{Morita85}
2120: S.~Morita,
2121: {\it Discontinuous invariants of foliations},
2122: Adv.~Stud.~Pure Math.~{\bf 5}
2123: (1985), 169--193.
2124:  
2125: \bibitem{Morita2}
2126: S.~Morita,
2127: {\it Characteristic classes of surface bundles},
2128: Invent.~Math.~{\bf 90}
2129: (1987), 551--577.
2130: 
2131: \bibitem{Morita89}
2132: S.~Morita,
2133: {\it Families of Jacobian manifolds and characteristic classes of surface 
2134: bundles I},
2135: Ann.~Inst.~Fourier {\bf 39}
2136: (1989), 777--810.
2137: 
2138: \bibitem{Morita89b}
2139: S.~Morita,
2140: {\it Families of Jacobian manifolds and characteristic classes of surface 
2141: bundles II},
2142: Math.~Proc.~Camb.~Phil.~Soc.~{\bf 105}
2143: (1989), 79--101.
2144: 
2145: \bibitem{Morita93}
2146: S.~Morita, {\it The extension of Johnson's homomorphism from the Torelli
2147: group to the mapping class group}, Invent.~Math.~{\bf 111} (1993), 197--224.
2148: 
2149: \bibitem{Morita96}
2150: S.~Morita,
2151: {\it A linear representation of the mapping class group
2152: of orientable surfaces and characteristic classes of surface bundles},
2153: in {\sl Proceedings of the Taniguchi Symposium on Topology and Teichm\"uller
2154: Spaces} held in Finland, July 1995,
2155: World Scientific
2156: 1996, 159--186.
2157: 
2158: \bibitem{Moser}
2159: J.~Moser,
2160: {\it On the volume elements on a manifold},
2161: Trans.~Amer.~Math.~Soc.~{\bf 120}
2162: (1965), 286--294.
2163: 
2164: \bibitem{Thurston0}
2165: W. Thurston,
2166: {\it On the structure of volume-preserving diffeomorphisms},
2167: unpublished manuscript.
2168: 
2169: \bibitem{Tsuboi}
2170: T. Tsuboi,
2171: {\it The Calabi invariant and the Euler class},
2172: Trans.~Amer.~Math.~Soc.~{\bf 352} (2000), 515--524.
2173: 
2174: 
2175: \end{thebibliography}
2176: 
2177: \end{document}
2178: