math0207036/pra.tex
1: \documentstyle{article}
2: \input amssymb.sty
3: %\input amssym.def
4: %\input amssym
5: \input epsf
6: 
7: \newtheorem{theorem}{Theorem}[section]
8: \newtheorem{lemma}[theorem]{Lemma}
9: \newtheorem{proposition}[theorem]{Proposition}
10: \newtheorem{corollary}[theorem]{Corollary}
11: \newtheorem{conjecture}[theorem]{Conjecture}
12: \newtheorem{definition}[theorem]{Definition}
13: \newtheorem{remark}[theorem]{Remark}
14: \newtheorem{theorem-construction}[theorem]{Theorem--Construction}
15: \newtheorem{lemma-construction}[theorem]{Lemma--Construction}
16: 
17: \begin{document}
18:  
19: \newcommand{\Z}{{\Bbb Z}}
20: \newcommand{\R}{{\Bbb R}}
21: \newcommand{\Q}{{\Bbb Q}}
22: \newcommand{\C}{{\Bbb C}}
23: \newcommand{\lra}{\longrightarrow}
24: \newcommand{\lms}{\longmapsto}
25: \newcommand{\AAA}{{\Bbb A}}
26: \newcommand{\Alt}{{\rm Alt}}
27: \newcommand{\wg}{\wedge}
28: \newcommand{\ol}{\overline}
29: \newcommand{\CP}{{\Bbb C}P}
30: \newcommand{\bwg}{\bigwedge}
31: \newcommand{\caL}{{\cal L}}
32: \newcommand{\PP}{{\Bbb P}}
33: \newcommand{\HH}{{\Bbb H}}
34: \newcommand{\LL}{{\Bbb L}}
35: 
36: 
37: \begin{titlepage}
38: \title{Polylogarithms, regulators, and Arakelov motivic complexes}
39: \author{A. B.  Goncharov}
40: \date{}
41: \end{titlepage}
42: \stepcounter{page}
43: \maketitle
44: \tableofcontents
45: 
46: \section  {Introduction}      
47: 
48: {\bf Abstract}. We construct an explicit regulator map 
49: from the weight $n$ Bloch Higher Chow group  complex to the weight $n$ 
50: Deligne complex 
51: of a regular projective complex algebraic variety $X$.  
52: We define the weight $n$ 
53: Arakelov motivic complex
54:   as the cone of this map 
55: shifted by one.
56:  Its last cohomology group is (a version of) the 
57: Arakelov Chow group defined by H. Gillet and C. Soul\'e ([GS]). 
58:  
59: We 
60: relate the 
61: Grassmannian $n$--logarithms (defined as in [G5])  
62: to the geometry of the symmetric space  
63: $SL_n(\C)/SU(n)$. For  $n=2$  we recover  
64: Lobachevsky's formula expressing the volume of an ideal geodesic 
65: simplex in the hyperbolic space via the dilogarithm. 
66: Using the relationship with symmetric spaces 
67: we construct the Borel regulator 
68: on $K_{2n-1}(\C)$ via the 
69: Grassmannian $n$--logarithms. 
70: 
71: We study the Chow dilogarithm and prove a 
72: reciprocity law 
73: which strengthens Suslin's reciprocity law for Milnor's group $K^M_3$ on curves. 
74: 
75: Our note [G5] can serve as an introduction to this paper.  
76: 
77: {\bf 1. Beilinson's conjectures on special values of L-functions}. 
78: Let $X$ be a regular scheme. 
79: A.A. Beilinson [B1] defined the rational motivic cohomology 
80: of $X$ via Quillen's 
81: algebraic $K$-theory of $X$ by the following formula:
82: \begin{equation} \label{4.29.02.1}
83: H^i_{\cal M}(X, \Q(n)) = K^{(n)}_{2n-i}(X)_{\Q}
84: \end{equation}
85:  where on the right stays the weight $n$ eigenspace for the Adams operations 
86: acting on $K_{2n-i}(X)\otimes \Q$. 
87: For a regular complex algebraic variety $X$  Beilinson defined the regulator map 
88: to the weight $n$ Deligne cohomology of 
89: $X(\C)$: 
90: $$
91: r_B: H_{{\cal M}}^{i}(X, \Q(n)) \lra H_{{\cal D}}^{i}(X(\C), \R(n))
92: $$
93: 
94: 
95: Let $X$ be a regular projective scheme over 
96: $\Q$. Let  $L(h^i(X),s)$  be the  $L$-function related to its i-dimensional
97: cohomology. 
98: Beilinson  conjectured that for any integer $n > 1+i/2$ 
99: its  special value at $s=n$  is described, up to a nonzero rational factor,
100:   by  the regulator map to the weight $n$ {\it real} Deligne cohomology of 
101: $X$ 
102: $$
103: r_B: H_{{\cal M}}^{i+1}(X, \Q(n))_{\Z} \lra H_{{\cal D}}^{i+1}(X\otimes_{\Q}\R_{/\R}, \R(n))
104: $$
105: Here the subscript $\Z$ on the left indicates the subspace of $H_{{\cal M}}^{i+1}(X, \Q(n))$ 
106: coming from a regular model of the scheme $X$ over $\Z$, see [B1] and [RSS] for details. 
107:  
108: 
109: 
110: This conjecture is fully established only when $X= {\rm Spec}(F)$ where
111: $F$ is a number field. In this case the regulator map $r_B$ coincides, up to a 
112: non-zero rational factor, 
113:  with 
114:  the Borel regulator ([B1]), 
115: and the relation with special values of the Dedekind zeta-function of $F$ was given
116: by the Borel theorem [Bo]. 
117: 
118: Although Beilinson's conjectures are far from being
119: proved, it is interesting to see what kind of information about the
120: special values of $L$-functions they 
121: suggest. So we come to the problem of {\it explicit} calculation of
122: Beilinson's regulator. This problem is already very interesting for the
123: Borel regulator.
124: 
125: 
126: 
127: {\bf 2. Regulator maps on motivic complexes and Arakelov motivic cohomology}.
128: Beilinson [B2] and S. Lichtenbaum [L1] conjectured that the weight $n$ 
129: integral 
130: motivic cohomology 
131: of $X$ should appear as  the cohomology of certain complexes of abelian groups 
132: $\Gamma(X;n)$, called the weight $n$ motivic complexes:
133: $$
134: H^i_{\cal M}(X, \Z(n)):= H^i(\Gamma(X;n))
135: $$
136: These  complexes are well defined as objects of the 
137: derived category. They 
138: should appear as the hypercohomology of certain complexes of Zarisky 
139: sheaves. 
140: 
141: The first  motivic complexes satisfying  
142: Beilinson's formula (\ref{4.29.02.1}) were Bloch's  Higher Chow group complexes 
143: ${\cal Z}^{\bullet}(X; n)$ [Bl1]. Later on A.A. Suslin and V.A. Voevodsky 
144: defined several important versions of these complexes. 
145: For another candidates for motivic complexes, called the polylogarithmic motivic 
146: complexes,  see [G1-2]. They are very explicit and the smallest among 
147: possible candidates, however Beilinson's formula (\ref{4.29.02.1}) 
148: is far from being established for them. 
149: 
150: 
151: The real Deligne cohomology arises also as the cohomology  
152: of certain complexes. 
153: It was suggested in [G5] and [G7] that the regulator map should be {\it explicitly} 
154: defined on the level of complexes. 
155: 
156: Let $X$ be a regular projective variety over $\C$. 
157: In Chapter 2 we construct 
158: a  homomorphism of complexes
159: \begin{equation} \label{4.29.02.2}
160: \mbox{Bloch's weight $n$ Higher Chow group complex ${\cal Z}^ {\bullet}(X; n)$ of $X$} 
161: \quad \lra 
162: \end{equation}
163: $$
164: \mbox{the weight $n$ real Deligne complex 
165: ${\cal C}^{\bullet}_{\cal D}(X(\C); n)$ of $X$}
166: $$
167: This construction is a version of the one given in [G5]. 
168: The  complex ${\cal C}^{\bullet}_{\cal D}(X(\C); n)$ is the  
169: truncation $\tau_{\leq 2n}$ of the complex 
170: proposed by 
171: Deligne [Del]. The 
172: weight $n$ Arakelov motivic  complex $\Gamma_{\cal A}^{\bullet}(X; n)$ is the cone of 
173: the map (\ref{4.29.02.2}), shifted by $-1$:
174: \begin{equation} \label{6.11.02.14}
175: \Gamma_{\cal A}^{\bullet}(X; n):= {\rm Cone}\Bigl({\cal Z}^ {\bullet}(X; n)
176: \stackrel{(\ref{4.29.02.2})}{\lra} 
177:  {\cal C}^{\bullet}_{\cal D}(X(\C); n)\Bigr)[-1]
178: \end{equation}
179: For a regular projective variety   $X$ over $\R$ the image of  map 
180: (\ref{4.29.02.2}) lies in the subcomplex
181: $$
182: {\cal C}^{\bullet}_{\cal D}(X_{/\R}; n) := 
183: {\cal C}^{\bullet}_{\cal D}(X(\C); n)^{\overline F_{\infty}}
184: $$
185: where $\overline F_{\infty}$ is the De Rham involution provided by the action of 
186: complex conjugation. 
187: The weight $n$ 
188: {\it real} Arakelov motivic complex 
189: is defined as 
190: \begin{equation} \label{6.11.02.4}
191: \Gamma_{\cal A}^{\bullet}(X_{/\R}; n):= {\rm Cone}\Bigl({\cal Z}^ {\bullet}(X; n)
192: \stackrel{(\ref{4.29.02.2})}{\lra} 
193:  {\cal C}^{\bullet}_{\cal D}(X_{/\R}; n)\Bigr)[-1]
194: \end{equation}
195: Let $X$ be a regular projective variety  $X$ over a number field $F$. We view $X$ over $\Q$: 
196: $X \lra {\rm Spec}(F) \lra {\rm Spec}(\Q)$, and set
197: \begin{equation} \label{6.11.02.4q}
198: \Gamma_{\cal A}^{\bullet}(X_{/F}; n):= {\rm Cone}\Bigl({\cal Z}^ {\bullet}(X; n)
199: \stackrel{}{\lra} 
200:  {\cal C}^{\bullet}_{\cal D}(X\otimes_{\Q} \R_{/\R}; n)\Bigr)[-1]
201: \end{equation}
202: The weight $n$ Arakelov motivic cohomology is the cohomology of this complex. 
203: Our construction 
204: works equally well for the Suslin-Voevodsky versions of the 
205: motivic complexes. 
206: 
207:  Taking the cohomology we get a construction of the 
208: regulator map on motivic cohomology. 
209: For a different construction see [Bl3]. 
210: 
211: 
212: The regulator map on the polylogarithmic motivic complexes 
213: was defined in [G7] explicitly via the classical polylogarithms. 
214: The Arakelov motivic complexes constructed using  regulator maps 
215: on different motivic complexes are supposed to lead to 
216: the same object of the derived category.  
217: However a precise relationship between the construction 
218: given in [G7] and the one 
219: in Chapter 2 is not clear. 
220: 
221: 
222: 
223: {\it Higher Arakelov Chow groups}. The last group of the 
224: complex ${\cal C}^{\bullet}_{\cal D}(X(\C); n)$ consists of 
225: closed distributions of a certain type on $X(\C)$. Replacing it 
226: by the quotient modulo smooth closed 
227: forms of the same type we get the quotient complex 
228: $\widetilde {\cal C}^{\bullet}_{\cal D}(X(\C); n)$. 
229: Changing  ${\cal C}$ to  
230: $\widetilde {\cal C}$ in (\ref{6.11.02.14}) 
231: we define the weight $n$ 
232: Higher Arakelov Chow group complex. Its last cohomology group is isomorphic to 
233: the Arakelov Chow group $\widehat {CH}^n(X(\C))$ as defined by Gillet and Soul\'e 
234: [GS], [S]. 
235: 
236: {\bf Problems}. a) Show that taking
237:  cohomology of
238:  the map (\ref{4.29.02.2}) and using the isomorphism between 
239: the rational Bloch's Higher Chow groups of $X$ and the corresponding part of the rational 
240: $K$-theory of $X$ ([Bl2], [Lev]) 
241: we get a non-zero rational multiple of the Beilinson's regulator map. 
242: 
243: b) To generalize the arithmetic Riemann--Roch theorem 
244: proved by Gillet and Soul\'e to the case of Higher Arakelov Chow groups. 
245:  
246:   {\bf Remark}. The weight $n$ Arakelov motivic complex should be considered as an ingrediant 
247: of a definition of the weight $n$ {\it arithmetic motivic complex}. The latter 
248: is related to the regulator maps on 
249: $H_{{\cal M}}^{\bullet}(X, \Q(n))_{\Z}$, while the former is related  
250: to the ones on $H_{{\cal M}}^{\bullet}(X, \Q(n))$. 
251: Ideally one should 
252: have for every place $p$ of $\Q$ a map from the left hand side of (\ref{4.29.02.2}) 
253: to a certain complex, which for the Archimedian  place should be given by our map. 
254: Then one should take the shifted by $-1$ cone of the sum of these maps. 
255:   
256: {\bf 3. The Chow $n$--logarithm function}. 
257: Let us  describe the  regulator map (\ref{4.29.02.2})
258: in the simplest case when $X = {\rm Spec}(\C)$ is a point. 
259: 
260: Let us choose in $\PP^m$  homogeneous coordinates $(z_0: ... : z_m)$. 
261: The union of the coordinate hyperplanes is a simplex $L$. 
262: Let $\AAA^m$ be the complement to
263:  the hyperplane $z_1 + ... + z_m = z_0$ in $\PP^m$. The abelian group 
264: ${\cal Z}_m({\rm Spec}(\C); n)$ is freely generated by 
265: the codimension $n$ irreducible algebraic cycles in $\AAA^{m}$ intersecting  properly 
266: the faces of the simplex $L$. The intersection with codimension one faces $L_j$ of $L$ provide homomorphisms 
267: $$
268: \partial_j: {\cal Z}_m({\rm Spec}(\C); n) \lra {\cal Z}_{m-1}({\rm Spec}(\C); n); \quad 
269: \partial:= \sum_{j=0}^{m} (-1)^j\partial_j 
270: $$
271: The weight $n$ Higher Chow group complex over ${\rm Spec}(\C)$, where $n>0$, 
272: written as a homological complex,  looks as follows:
273: $$
274: ... \stackrel{\partial}{\lra} {\cal Z}_2({\rm Spec}(\C); n) \stackrel{\partial}{\lra}
275: {\cal Z}_1({\rm Spec}(\C); n) \stackrel{\partial}{\lra} {\cal Z}_0({\rm Spec}(\C); n) 
276: $$
277: 
278: The Deligne complex of a point is the complex $(2\pi i)^{n}\R \lra \C$, with $\C$ is 
279: in the degree $+1$. So it is quasiisomorphic to the group $\R(n-1):= (2\pi i)^{n-1}\R$ placed in degree $+1$. 
280: The
281:  regulator map (\ref{4.29.02.2}) 
282: boils down to a construction  of a homomorphism 
283: $$
284: {\cal Z}_{2n-1}({\rm Spec}(\C); n)  \stackrel{{\cal P}_n}{\lra} \R(n-1), \quad 
285: \mbox{such that 
286: ${\cal P}_n \circ \partial = 0$} 
287: $$
288: It is provided by  
289: a function 
290: ${\cal P}_n$ on the space of 
291: codimension $n$ cycles in $\C\PP^{2n-1}$ intersecting properly 
292:  faces of a simplex $L$. This function, called 
293: the {\it Chow $n$--logarithm function}, was constructed in [G5]. 
294: To recall its construction, observe that 
295: a codimension $n$ cycle given by an irreducible 
296:  subvariety $X$ in $\PP^{2n-1} -L$ provides the 
297: $(n-1)$-dimensional variety $X$ with $2n-1$ rational functions 
298: $f_1, ..., f_{2n-1}$:  These functions are obtained by 
299: restriction of the coordinate functions 
300: $z_i/z_0$ to the cycle $X$. We define a natural $(2n-2)$-form 
301: $r_{2n-2}(f_1, ..., f_{2n-1})$ on $X(\C)$ and set
302: \begin{equation} \label{6.17.02.1}
303: {\cal P}_n(X; f_1, ..., f_{2n-1}):= (2\pi i)^{1-n}\int_{X(\C)}r_{2n-2}(f_1, ..., f_{2n-1})
304: \end{equation}
305: 
306: 
307: {\bf 4. An example: the Chow dilogarithm}. 
308: Let $f_1,f_2,f_3$ be three arbitrary 
309: rational functions on a complex curve $X$. 
310: %$$
311: %d r_2(f_1,f_2,f_3) = {\rm Re}\Bigl(d \log f_1 \wedge d \log f_2 \wedge d \log f_3\Bigr)
312: %$$
313: Set 
314: $$
315: r_2(f_1,f_2,f_3):= 
316: $$
317: $$
318: {\rm Alt}_3 \Bigl(\frac{1}{6}\log|f_1| d \log|f_2| 
319: \wedge d
320: \log|f_3|
321: -\frac{1}{2}
322: \log|f_1| d\arg f_2 \wedge d\arg f_3 \Bigr) 
323: $$
324:  where 
325: ${\rm Alt}_3$ is the alternation of $f_{1}, f_{2}, f_{3}$. 
326:  Consider the space of
327: quadruples $(X;f_1,f_2,f_3)$. It is a union of finite dimensional algebraic varieties. 
328: The Chow dilogarithm is a real function 
329: on its complex points  defined by the formula
330: $$
331: {\cal P}_2(X;f_1,f_2,f_3):= \frac{1}{2\pi i}\int_{X(\C)}r_2(f_1,f_2,f_3)
332: $$
333: The integral converges. 
334: The Chow dilogarithm provides a homomorphism 
335: \begin{equation} \label{chowh}
336: \Lambda^3 \C(X)^* \to \R,\quad  f_1\wedge f_2\wedge f_3 \lms {\cal P}_2(X;f_1,f_2,f_3)
337: \end{equation}
338: 
339: Why does the dilogarithm appear in the name of the function ${\cal P}_2$? 
340: Recall  the classical dilogarithm 
341: $$
342: Li_2(z):= - \int_0^z\log(1-z)d\log z
343: $$ 
344: It has a single-valued cousin, the Bloch-Wigner  function:
345: $$
346: {\cal L}_2(z) := {\rm Im} Li_2(z) + \arg (1-z) \log \vert z \vert 
347: $$
348: 
349: The Chow dilogarithm
350: is defined by a two-dimensional integral
351:  over $X(\C)$, while ${\cal L}_2(z)$ 
352: is given by an 
353: integral over a path in ${\C}{\Bbb P}^1$.  In Chapter 6 we
354:  show that nevertheless the Chow dilogarithm can be expressed by the function ${\cal L}_2(z)$. Here is how it 
355: works when $X = \C\PP^1$. 
356: For $f \in \C(X)$ let $v_x(f)$ be the order of zero of $f$ at $x \in X(\C)$. 
357: Choose a point $\infty$ on ${\Bbb P}^1$. Then 
358: \begin{equation} \label{6.14.02.2}
359: {\cal P}_2(\C\PP^1; f_1, f_2, f_3) =  \sum_{x_i \in {\Bbb P}^1(\C)} 
360: v_{x_1}(f_1)v_{x_2}(f_2)v_{x_3}(f_3){\cal L}_2(r(x_1, x_2, x_3, \infty))
361: \end{equation}where $r(...)$ denotes the cross-ratio of four points on $\PP^1$. 
362: A formula for the Chow dilogarithm on elliptic curves is given 
363: in Chapter 6. 
364: 
365: The function ${\cal L}_2$  satisfies  Abel's five term functional equation: 
366: \begin{equation} \label{6.14.02.1}
367: \sum_{i=1}^{5}(-1)^i {{\cal L}}_2(r(x_1,...,\widehat  x_i,...,x_{5})) =0 
368: \end{equation}
369: The Chow dilogarithm also satisfies functional equations. They appear as a reformulation 
370: of 
371: the fact that the composition 
372: $$
373: {\cal Z}_4(Spec(\C); 2) \stackrel{\partial}{\lra} {\cal Z}_3(Spec(\C); 2) 
374: \stackrel{{\cal P}_2}{\lra} \R(1)
375: $$ 
376: is zero. Namely,  
377: let $Y$ be an algebraic surface with four rational functions 
378: $g_1, ..., g_4$ on it corresponding to an element of ${\cal Z}_4(Spec(\C); 2)$. 
379: To evaluate the composition on this element 
380: we do the following. 
381: Take the divisor ${\rm div} (g_i)$ and 
382: restrict the other functions $g_i$ 
383: to it. Then applying the Chow dilogarithm to the obtained data and 
384: taking the alternating sum over $1 \leq i \leq 4$ we get zero. 
385: In the special case when $Y = \C\PP^2$  and ${\rm div}g_i = (l_i) - (l_{5})$, 
386: where $l_1, ...,  l_{5}$ are five  lines in the plane, this 
387: functional equation  plus (\ref{6.14.02.2}) is equivalent to   Abel's equation 
388: (\ref{6.14.02.1}).
389: 
390:  
391: 
392: {\bf 5. The Grassmannian $n$--logarithm and  symmetric space 
393: $SL_n(\C)/SU(n)$}. 
394: Restricting the Chow $n$-logarithm function to the subvariety of $(n-1)$--planes in 
395: $\C\PP^{2n-1}$ in general position with respect to the simplex $L$ 
396: we get the Grassmannian $n$--logarithm function ${\cal L}^G_n$. 
397: 
398: Let $G$ be a group and $X$ a $G$-set. {\it Configurations} of $n$ points in $X$ are 
399: by definition the points of the quotient $X^n/G$. There is a  natural bijection
400: $$
401: \mbox{ $\{ (n-1)$--planes in 
402:  $\PP^{2n-1}$ in generic position with respect to a simplex $L$\}}/({\Bbb G}_m^*)^{2n-1} \quad 
403: $$
404: $$  <--> \quad \left \{ \mbox{Configurations of} \quad 2n \quad\mbox {generic hyperplanes in } \PP^{n-1}\right \}
405: $$
406: given by intersecting of an $(n-1)$--plane $h$ with the codimension one faces of $L$. 
407: 
408: \begin{figure}[ht]
409: \centerline{\epsfbox{gpol6.eps}}
410: \caption{Toric quotients of Grassmannians and configurations of hyperplanes}
411: \label{gpol6}
412: \end{figure}
413: 
414: 
415: Using it 
416: we can view ${\cal L}^G_n$ as a function on the configurations of $2n$ hyperplanes in $\C \PP^{n-1}$. Applying the projective duality we can consider it as a function 
417: on configurations of $2n$ points in $\C \PP^{n-1}$.
418: 
419: In fact one can define the Grassmannian $n$-logarithm 
420: ${{\cal L}}^G_n(x_1,...,x_{2n})$  as a 
421: function 
422:  on configurations of {\it arbitrary}  $2n$ points in $\C\PP^{n-1}$, see 
423:  Chapter 4. 
424: It is a measurable function which is 
425: real analytic on generic configurations. 
426: It satisfies the two functional equations
427: \begin{equation} \label{6.13.02.100}
428: \sum_{i=0}^{2n}(-1)^i {{\cal L}}^G_n(x_0,...,\widehat  x_i,...,x_{2n}) =0, \quad \sum_{j=0}^{2n}(-1)^j {{\cal L}}^G_n(y_j|y_0,...,\widehat  y_j,...,y_{2n}) = 0
429: \end{equation}
430: In the second formula $(y_0,...,y_{2n})$ is a configuration
431: of $2n+1$ points in $\C\PP^{n}$ and
432: $(y_j|y_0,...,\widehat  y_j,...,y_{2n})$ 
433: is a configuration of $2n$ points in $\C\PP^{n-1}$
434: obtained by projection from  $y_j$. 
435: 
436: \begin{figure}[ht]
437: \centerline{\epsfbox{gpol20.eps}}
438: \caption{The configuration $(y_1|y_2, y_3, y_4, y_5)$ on $\PP^1$}
439: \label{gpol20}
440: \end{figure}
441: 
442: 
443: 
444: It follows from (\ref{6.14.02.2}) that 
445: the Grassmannian dilogarithm is given by the Bloch-Wigner function:
446: \begin{equation} \label{6.13.02.101}
447: {\cal L}^G_2(z_1, ..., z_4) = {\cal L}_2(r(z_1, ..., z_4))
448: \end{equation}
449: Abel's five term equation coincides with 
450: (\ref{6.13.02.100}). (The two functional equations (\ref{6.13.02.100}) 
451: are equivalent when $n=2$). 
452: 
453: Lobachevsky discovered that the dilogarithm 
454: appears in the computation of volumes of geodesic 
455: simplices in the three dimensional hyperbolic space ${\cal H}_3$.  
456: Let $I(z_1,...,z_4)$ be  the ideal geodesic simplex with vertices at the 
457: points $z_1,...,z_4$ on the absolute of ${\cal H}_3$. 
458: The absolute is naturally identified with 
459: $\C\PP^1$. 
460: \begin{figure}[ht]
461: \centerline{\epsfbox{gpol21.eps}}
462: \caption{An ideal simplex in the hyperbolic $3$-space}
463: \label{gpol21}
464: \end{figure}
465: Lobachevsky's formula 
466: relates its volume to the Bloch-Wigner  function:
467: $$
468: {\rm vol}\Bigl(I(z_1,...,z_4)\Bigr) =   {\cal L}_2(r(z_1,...,z_4))
469: $$
470: The volume function ${\rm vol}I(z_1,...,z_4)$ is invariant under the group 
471: $SL_2(\C)$ of isometries of ${\cal H}_3$. So it  
472: depends only on the cross ratio 
473: of the points $z_1,...,z_4$. It   satisfies the five term equation (\ref{6.14.02.1}).
474: Indeed, $\sum(-1)^iI(z_1,..., 
475: \widehat z_i, ..., z_5) = \emptyset$. 
476: By Bloch's theorem  [Bl2] any measurable function $f(z)$ on $\C$
477: satisfying the five term equation  is proportional to ${\cal L}_2(z)$.
478: So we get the formula up to a constant. 
479: 
480: We generalize this picture as follows. 
481:  $\C\PP^{n-1}$ is realized as 
482: the smallest boundary 
483: stratum of the symmetric space ${\Bbb H}_n := SL_n(\C)/SU(n)$. We  define a function 
484: $\psi_n(x_1, ..., x_{2n})$  on configurations of $2n$ points 
485: of the symmetric space. The function $\psi_n$ 
486: is defined by an integral over $\C\PP^{n-1}$ similar to (\ref{6.17.02.1}).  
487: We show that it can be naturally extended to a function $\overline \psi_n$ 
488: on configurations of $2n$ points in a compactification $\overline {\Bbb H}_n$ of 
489: the symmetric space. The Grassmannian $n$--logarithm function 
490: turns out to be the value of the function $\overline \psi_n$ 
491: on configurations of $2n$ points at the smallest boundary strata, 
492: which is identified with $\C\PP^{n-1}$. 
493: 
494: 
495: Now let  $n=2$. Then 
496: $SL_2(\C)/SU(2)$ is identified with the hyperbolic $3$-space. 
497:  We prove in Chapter 7 that  $\psi_2(x_1, x_2, x_3, x_{4})$ is 
498:  the volume of the geodesic simplex with vertices at the points 
499: $x_1, ..., x_4$. Restricting to the ideal geodesic simplices 
500: and using the relation to the Grassmannian dilogarithm 
501: plus  (\ref{6.13.02.101}) we 
502: get a 
503: new  proof of   Lobachevsky's formula. 
504: 
505:  
506: {\bf 6. The Grassmannian $n$--logarithms and the Borel regulator}.
507: For any point $x \in \C\PP^{n-1}$ the function
508: \begin{equation} \label{6.17.02.2}
509: c_{2n-1}^n(g_1,...,g_{2n}):= { {\cal L}}^G_n(g_1x,...,g_{2n}x)
510: \end{equation} 
511: is a measurable $(2n-1)$-cocycle of the Lie group $GL_n(\C)$. 
512: Indeed, it is invariant
513: under the diagonal action of $GL_n(\C)$ and  the cocycle
514: condition is just the first functional equation for the function
515: $ {\cal L}^G_n$. Different points $x$ give canonically cohomologous cocycles. 
516: However {\it a priori} it is not clear that the corresponding   cohomology class 
517:   is non-zero. 
518: 
519: 
520: Let $H^{2n-1}_{m}(GL_n(\C), \R)$ be the space of measurable 
521: cohomology of the Lie group $GL_n(\C)$. It is known that
522: $$
523: H^{\ast}_{m}(GL_n(\C), \R) = \Lambda_{\Bbb
524: R}^{\ast}(b_1,b_3,...,b_{2n-1}) 
525: $$
526: where $b_{2k-1} \in H^{2k-1}_{m}(GL_n(\C), \R)$ are certain
527: canonical generators  called the Borel classes ([Bo1]).
528: \begin {theorem}  
529: \label {0.4}
530: The cohomology class of the Grassmannian cocycle  (\ref{6.17.02.2}) is a non
531: zero rational multiple of the Borel class $b_{2n-1}$. 
532: \end{theorem}
533: 
534: For normalization of the Borel classes and precise relationship between the Grassmannian polylogarithms and the Borel regulator 
535: see Chapter 5, especially Sections 5.4 and 5.5. 
536: 
537: The essential role in  the proof is played by 
538: the fact that the Grassmannian $n$--logarithm function 
539:  ${\cal L}^G_n$ is a boundary value of the function  
540: $\overline \psi_n$. The function 
541: $\overline \psi_n(x_1,...,x_{2n})$ is not continuous at certain boundary points, but 
542: always satisfies the cocycle condition. 
543: So taking any point $x \in \overline {\Bbb H}_n$ we get a cocycle
544: $$
545: c_x(g_1, ..., g_{2n-1}):= \overline \psi_n(g_1x,..., g_{2n}x)
546: $$
547: of the group $GL_n(\C)$. Its cohomology class 
548: does not depend on  $x$. If $x \in {\Bbb H}_n$ 
549: the corresponding  cocycle  is smooth. We can differentiate it, getting   
550: a  cohomology class of the Lie algebra $gl_n$,  and relate it to the 
551: Borel class. On the other hand taking $x$ to be a point on the boundary stratum 
552: $\C\PP^{n-1}$ we recover the Grassmannian cocycle (\ref{6.17.02.2}). 
553: So we get the theorem. 
554: 
555: Combining it with the technique developed in [G1-2] we 
556: get  a simple explicit construction of the Borel regulator 
557: $$
558: K_{2n-1}(\C) \longrightarrow \R
559: $$
560: in terms of the Grassmannian $n$-logarithms.
561:  The second functional equation for ${\cal L}^G_n$ plays an important role in the proof. 
562: Therefore, thanks to the Borel theorem [Bo2], this allows 
563: to  express the  special values of Dedekind $\zeta$--functions at $s=n$ 
564: via the Grassmannian $n$--logarithms. 
565: 
566: The definition of the Higher Chow groups of a variety $X$ 
567: is much simpler than 
568: the definition of algebraic K-groups of $X$. The situation with 
569: the regulator maps is similar. However relating the 
570: special values of the Dedekind $\zeta$-functions to motivic 
571: cohomology of the corresponding  number fields
572: we need to work with the algebraic K-theory (or homology of $GL_n(F)$)  
573: of number fields.  
574: 
575: Chapter 2 is the main core of the paper. In Chapters 3, 4 and 6  
576: the main construction of Chapter 2 is investigated from different points of view. 
577: Chapters 4 and 5 are rather 
578: independent from the other Chapters. 
579: 
580: {\bf Acknowledgement}. 
581: I am grateful to Spencer Bloch for discussions of Arakelov 
582: motivic complexes. 
583: The results of this  paper were discussed in  my course at Brown (Spring 1998), 
584: and   at 
585: the Newton Institute (Cambridge, March 1998). 
586: I am grateful to participants for 
587: their interest and useful comments. I am very grateful to the referees 
588: for many comments which greatly improved the exposition. 
589: 
590: 
591: During the work on this paper I enjoyed  the hospitality and support of the 
592: MPI(Bonn) and 
593: IHES (Bures-sur-Yvette). I am grateful to these 
594: institutions. 
595: This work was supported by the NSF grants DMS-9800998 
596: and  DMS-0099390. 
597: 
598: 
599:  \section { Arakelov motivic complexes}
600: 
601: {\bf 1.  The  Higher Chow group complex}. A (non-degenerate) simplex in 
602: $\PP^m$ is an ordered
603: collection of hyperplanes $L_0,...,L_m$ in generic position, i.e. with empty
604: intersection. 
605: Let us choose in $\PP^m$ a simplex $L$ and a generic hyperplane $H$.
606: We might think about this data as of a simplex in the $m$-dimensional
607: affine space $\AAA^m:= \PP^m - H$.    For any two non-degenerate 
608: simplices in $\AAA^m$ there is a unique affine transformation sending one 
609: simplex to the other.
610: 
611: Let $I=(i_1,...,i_k)$ and $L_I:= L_{i_1} \cap ... \cap L_{i_k}$. 
612: Let $X$ be a regular projective variety over a field $F$. Let 
613: ${\cal Z}_{m}(X; n)$ be
614: the free abelian group generated by irreducible codimension $n$
615: algebraic 
616: subvarieties in $X \times \AAA^m$ which  intersect properly (i.e.  with 
617: the  right codimension) all faces
618: $X \times L_I$.
619: 
620: {\bf Warning}. We use the notation  ${\cal Z}_{m}(X; n)$ for the group 
621: denoted  ${\cal Z}^{n}(X; m)$ by  Bloch. This allows us to use upper and lower 
622: indices to distinguish between the 
623: homological and cohomological notations, see below. 
624: 
625: For a given codimension 1 face $L_i$ of a simplex $L$ in $\AAA^m$ the  other 
626: faces $L_j$ cut a simplex $\widehat  L_i:= \{L_i \cap L_j\}$,  in $L_i$. 
627: So the  intersection with codimension 1 faces $X \times L_i$ provides 
628:  group homomorphisms 
629: $$
630: \partial_i: {\cal Z}_{m}(X; n) \longrightarrow 
631: {\cal Z}_{m-1}(X; n); \qquad \partial:= \sum_{i=0}^m (-1)^i \partial_i
632: $$
633:  Then   $\partial^2 =0$, so 
634: $({\cal Z}_{\bullet}(X; n); \partial)$    
635: is a homological complex. Its homology groups  
636: are Bloch's Higher
637: Chow groups. 
638: By the fundamental theorem of Bloch ([Bl1-2], [Lev])
639: $$
640: H_i({\cal Z}_{\bullet}(X; n)\otimes \Q)
641:  = K^{(n)}_{i}(X)\otimes \Q
642: $$
643: Let us cook up  a cohomological complex setting
644: $$
645: {\cal Z}^{\bullet}(X; n):= {\cal Z}_{2n-\bullet}(X; n)
646: $$ 
647: Its cohomology provides  a definition of the integral motivic cohomology of $X$:
648: $$
649: H_{{\cal M}}^i(X, \Z(n)) := H^i({\cal Z}^{\bullet}(X; n))
650: $$
651: Bloch's theorem guarantees Beilinson's formula (\ref{4.29.02.1}) 
652: for  the rational motivic cohomology. 
653: 
654: 
655: 
656: {\bf 2.  The  Beilinson--Deligne complex}. 
657: Recall that an  $n$-distribution, sometimes also called an $n$-form with generalized function coefficients, or an $n$-current,  on a smooth  oriented manifold $X$ is 
658:  a continuous linear  functional on the space of 
659: $(\dim_{\R}X-n)$--forms with compact support.  Denote by 
660: ${\cal D}^{n}_{X}$ the space of all {\it real} $n$--distributions on $X$.  
661: The  
662: De Rham complex of distributions 
663: $({\cal D}^{\bullet}_{X},d)$ is a resolution of the 
664: constant sheaf $\R$.  The space ${\cal A}^{n}_{X}$ of all smooth $n$--forms 
665: on $X$ is a subspace of 
666: ${\cal D}^{n}_{X}$.
667: 
668: 
669: Let $X$ be a regular projective variety over $\C$.  The 
670: standard weight $n$  Beilinson-Deligne complex 
671: ${\R}^{\bullet}(X; n)_{{\cal D}}$ 
672: is the total complex associated with the following
673: bicomplex: 
674: $$
675: \begin{array}{ccccccccccc} \label{del}
676: \Bigl({\cal D}_{X}^{0}&\stackrel{d}{\longrightarrow}&{\cal
677: D}_{X}^{1}&\stackrel{d}{\longrightarrow}&\ldots&\stackrel{d}{\longrightarrow}&{\cal
678: D}^{n}_{X}&\stackrel{d}{\longrightarrow}&{\cal
679: D}_{X}^{n+1}&\stackrel{d}{\longrightarrow}&\ldots\Bigr) \otimes \R(n-1)\\
680: &&&&&&&&&&\\
681: &&&&&&\uparrow\pi_{n} &&\uparrow\pi_{n}&&\\
682: &&&&&&&&&&\\
683: &&&&&&\Omega^{n}_{X}
684: &\stackrel{\partial}{\longrightarrow}&\Omega_{X}^{n+1}&\stackrel
685: {\partial}{\longrightarrow}&
686: \end{array}
687: $$
688: Here  $\R(n):= (2\pi i)^n\R$ and 
689: $$
690: \pi_n: {\cal D}_{X}^{p}\otimes \C \longrightarrow
691: {\cal D}_{X}^{p}\otimes \R(n-1)
692: $$ is the projection induced by the one  $\C = \R(n-1) \oplus \R(n)  \longrightarrow 
693: \R(n-1)$.  Further,  ${\cal D}^{0}_{X}$ placed in
694: degree 1 and  
695: $(\Omega^{\bullet}_{X}, \partial)$ is the De Rham complex of 
696: holomorphic forms. 
697: 
698: The Beilinson-Deligne complex ${\R}^{\bullet}(X; n)_{{\cal D}}$ 
699: is quasiisomorphic to the complex
700: $$
701: \R(n) \lra {\cal O}_X \lra \Omega^{1}_{X} 
702: \lra \Omega^{2}_{X}\lra ... \lra \Omega^{n-1}_{X}
703: $$
704: 
705: %If $X$ is not projective we should take a 
706: %compactification $\overline X$ of $X$ such that $Y:= \overline X - X$ 
707: %is a normal crossing divisor, and use 
708: %the complex  $F^n\Omega^{\bullet}_{X}<Y>$ of forms 
709: %{\it with logarithmic singularities}
710: % at  $Y$ in the definition above. 
711:  
712: 
713: {\bf 3.  The  truncated Deligne complex}. 
714: Let ${\cal D}_X^{ p,q} = {\cal D}^{ p,q}$ be the abelian group of complex valued distributions of type $(p,q)$ on $X(\C)$. Consider the following cohomological 
715: ``bicomplex'',  where ${\cal D}^{n,n}_{\rm cl} $ 
716: is the subspace of the space ${\cal D}^{n,n}$ of closed currents, and  
717:  ${\cal D}^{0,0}$ is in degree $1$:
718: $$
719: \begin{array}{ccccccccc}
720: &&&&&&&&{\cal D}_{cl}^{n,n}\\
721: &&&&&&&&\\
722: &&&&&&&2 \overline \partial \partial\nearrow&\\
723: &&&&&&&&\\
724: {\cal D}^{0,n-1}&\stackrel{\partial}{\longrightarrow}&{\cal D}^{1,n-1}&\stackrel{\partial}{\longrightarrow}&
725: ...&\stackrel{\partial}{\longrightarrow}&{\cal D}^{n-1,n-1}&&\\
726: &&&&&&&&\\
727: \overline \partial \uparrow &&\overline \partial \uparrow &&&&\overline \partial \uparrow&&\\
728:  ...&...&...&...&...&...&...&&\\
729:  \overline \partial \uparrow&&\overline \partial \uparrow&&&&\overline \partial \uparrow&&\\
730: &&&&&&&&\\
731: {\cal D}^{0,1}&\stackrel{\partial}{\longrightarrow}& {\cal D}^{1,1}&\stackrel{\partial}{\longrightarrow}&... &\stackrel{\partial}{\longrightarrow}& {\cal D}^{n-1,1}&&\\
732: &&&&&&&&\\
733: \overline \partial \uparrow&&\overline \partial \uparrow&&&&\overline \partial \uparrow&&\\
734: &&&&&&&&\\
735: {\cal D}^{0,0}&\stackrel{\partial}{\longrightarrow} &{\cal D}^{1,0}& \stackrel{\partial}{\longrightarrow} &...&\stackrel{\partial}{\longrightarrow} &{\cal D}^{n-1,0}&&
736: \end{array}
737: $$
738: Properly speaking, it is not a bicomplex due to the presence of the operator 
739: $2\overline \partial \partial$, but we can handle it the same way we handle the bicomplexes. Namely, we define its total complex   
740: $Tot^{\bullet}$. It is concentrated 
741: in degrees $[1,2n]$. The   complex $C^{\bullet}_{{\cal D}}(X(\C); n) 
742: = C^{\bullet}_{{\cal D}}(n)$ is a  subcomplex 
743: of the complex 
744: $Tot^{\bullet}$ defined as follows. 
745: Take  the intersection of the part of the complex $Tot^{\bullet}$  
746:   coming from the $n \times n$ square in the diagram (and concentrated in degrees 
747: $[1,2n-1]$) with the complex of distributions with values in $\R(n-1)$. Consider 
748: the   subgroup    ${\cal D}_{\R, cl}^{n, n}(n) \subset {\cal D}_{cl}^{n, n}$  of
749:  the $\R(n)$-valued distributions of type $(n, n)$. They  form a subcomplex in 
750: $Tot^{\bullet}$ because
751:  $\overline \partial \partial$ sends $\R(n-1)$-valued distributions to 
752: $\R(n)$--valued distributions.
753: This is the complex $C^{\bullet}_{{\cal D}}(n)$. It is a truncation of the 
754: complex   considered by Deligne ([Del]). Its cohomology is the absolute Hodge cohomology defined by Beilinson [B3].
755: 
756: 
757: 
758: \begin{proposition} \label{dcom} Let $X$ be a regular complex projective 
759: variety. 
760: Then the    complex $C^{\bullet}_{{\cal D}}(X; n)$  is quasiisomorphic to the 
761: truncated Beilinson-Deligne complex $\tau_{\leq 2n}\R^{\bullet}(X; n)_{{\cal D}}$. 
762: \end{proposition}
763: 
764: {\bf Proof}. We need the following general construction. 
765: Let $f^{\bullet}: X^{\bullet} \lra Y^{\bullet}$ be a morphism of complexes such that
766: the map $f^i $ is injective for $i \leq p$ and surjective for $i \geq p$ (and hence is an isomorphism for $i=p$). Consider a complex
767: $$
768: Z^{\bullet}:= \quad {\rm Coker}f^{<p}[-1] \stackrel{D}{\lra} {\rm Ker}f^{>p}
769: $$
770: where the differential $D: {\rm Coker} f^{p-1} \lra {\rm Ker} f^{p+1}[1]$ is defined via the following diagram (the vertical sequences are exact):
771: $$
772: \begin{array}{ccccccccc}
773: &&&&&& 0 &&0\\
774: &&&&&&\downarrow &&\downarrow \\
775: 0&&0 &&&&{\rm Ker}f^{p+1}&\lra &{\rm Ker}f^{p+2}\\
776: \downarrow &&\downarrow &&&&\downarrow &&\downarrow \\
777: X^{p-2} &\lra &X^{p-1} &\lra &X^{p}&\lra & X^{p+1}&\lra &X^{p+2}\\
778: \downarrow &&\downarrow &&f^p \downarrow = && \downarrow &&\downarrow \\
779: Y^{p-2}&\lra &Y^{p-1} &\lra &Y^{p} &\lra & Y^{p+1}&\lra &Y^{p+2}\\
780: \downarrow &&\downarrow &&&&\downarrow &&\downarrow \\
781: {\rm Coker} f^{p-2}&\lra &{\rm Coker} f^{p-1}&&&& 0&&0\\
782: \downarrow &&\downarrow &&&&&&\\
783: 0&&0 &&&&&&
784: \end{array}
785: $$
786: 
787: \begin{lemma} \label{decom}
788: The complex $Z^{\bullet}$ is canonically quasiisomorphic to ${\rm Cone} (X^{\bullet} \stackrel{f^{\bullet}}{\lra} Y^{\bullet})$.
789: \end{lemma}
790: 
791: {\bf Proof}. Let
792: $$
793: \widetilde  \tau_{<p}X^{\bullet}:= \quad ... \stackrel{d_X}{\lra}  X^{p-2} \stackrel{d_X}{\lra}  X^{p-1} \stackrel{d_X}{\lra}  {\rm Im} d_X
794: $$
795: $$
796: \widetilde  \tau_{\geq p}Y^{\bullet}:= \quad Y^{p}/{\rm Im} d_Y  \stackrel{d_Y}{\lra}  Y^{p+1} 
797: \stackrel{d_Y}{\lra}  Y^{p+2} \stackrel{d_Y}{\lra} ...
798: $$
799: Then there is an exact sequence of complexes $0 \lra \widetilde  \tau_{<p}X^{\bullet} \lra X^{\bullet} \lra \widetilde  \tau_{\geq p}X^{\bullet} \lra 0$. The conditions on the maps $f^{\bullet}$ imply that 
800: $$
801: \widetilde  \tau_{<p}f^{\bullet}: \quad \widetilde  \tau_{<p}X^{\bullet} \lra \widetilde  \tau_{<p}Y^{\bullet} \quad \mbox{is injective}
802: $$
803: $$
804: \widetilde  \tau_{\geq p}f^{\bullet}: \quad \widetilde  \tau_{\geq p}Y^{\bullet} \lra \widetilde  \tau_{\geq p}Y^{\bullet} \quad \mbox{is surjective}
805: $$
806: We get maps of complexes
807: $$
808: {\rm Cone}(\widetilde  \tau_{<p}X^{\bullet} \lra f^{\bullet}(\widetilde  \tau_{<p}X^{\bullet})) \stackrel{\alpha}{\hookrightarrow}
809: {\rm Cone}(X^{\bullet} \lra Y^{\bullet}) \stackrel{\beta}{\longrightarrow} 
810: {\rm Cone}(\widetilde  \tau_{\geq p}Y^{\bullet} \lra \widetilde  \tau_{\geq p}Y^{\bullet}) 
811: $$
812: where $\alpha$ is injective and $\beta$ is surjective.
813: The complex ${\rm Ker} (\beta)/{\rm Im}(\alpha)$ looks as follows:
814: $$
815: \begin{array}{ccccccccc}
816: 0&\lra &0 &\lra & {(f^p)}^{-1}{\rm Im}(d_Y)/{\rm Im}(d_X)&\lra &{\rm Ker} f^{p+1}&\lra &{\rm Ker} f^{p+2}\\
817: \downarrow &&\downarrow &&\downarrow &&\downarrow &&\downarrow \\
818: {\rm Coker} f^{p-2}&\lra &{\rm Coker} f^{p-1}&\lra & {\rm Im}(d_Y)/{f^p}{\rm Im}(d_X)  &\lra &0&\lra &0
819: \end{array}
820: $$
821: Since the map $f^p : {\rm Im}(d_Y)/{\rm Im}(d_X) \to {\rm Im}(d_Y)/{f^p}{\rm Im}(d_X)$ is an isomorphism it is quasiisomorphic to $Z^{\bullet}$. The lemma is proved.
822: 
823: 
824: 
825: Applying the lemma to the morphism of complexes 
826: $$
827: {\rm Tot}({\cal D}^{\geq n, \bullet}) \stackrel{\pi_n}{\longrightarrow} {\cal D}^{n+\bullet}\otimes_{\R}\R(n-1)
828: $$
829: we see that the complex $\R(n)_{{\cal D}}$ is canonically quasiisomorphic to the following complex:
830: 
831: \begin{figure}[ht]
832: \centerline{\epsfbox{gpol1.eps}}
833: \caption{The weight $n$ real Deligne complex }
834: \label{gpol1}
835: \end{figure}
836: To compute the differential ${\cal D}_{\R}^{n-1, n-1}(n-1) \lra {\cal D}_{\R}^{n, n}(n)$ we proceed as follows. Take  $\alpha \in {\cal D}_{\R}^{n-1, n-1}(n-1)$, so $\alpha = (-1)^{n-1}\overline \alpha$. Then 
837: $$
838: d \alpha \quad = \quad \partial \alpha + \overline \partial \alpha \quad = \quad \partial \alpha + (-1)^{n-1}\overline {\partial \alpha } \quad = \quad 2 \pi_n(\partial \alpha )
839: $$
840: Applying $d = \partial + \overline \partial$ again and taking the $(n,n)$-component we get $2\overline \partial \partial (\alpha )$. 
841: Truncating this complex we obtain the proof of the proposition. 
842: (Note that $d^{C}:= 
843: (4\pi i)^{-1}(\partial - \overline \partial)$, 
844: so $dd^{C} = (2\pi i)^{-1}\overline \partial\partial$.)
845: 
846: Now if $X$ is a variety over $\R$, then we set
847: $$
848: C^{\bullet}_{{\cal D}}(X_{/{\R}}; n):= C^{\bullet}_{{\cal D}}(X; n)^{\overline F_{{\infty}}}; \qquad H_{{\cal D}}^i(X_{/{\R}}; \R(n)) := 
849: H^i\Bigl(  C^{\bullet}_{{\cal D}}(X_{/{\R}}; n)\Bigr)
850: $$
851: where $\overline F_{{\infty}}$ is the De Rham involution,  i.e.  the composition  of the involution $F_{{\infty}}$   on $X(\C)$ induced by the complex conjugation  with the complex conjugation of coefficients. 
852: 
853: 
854: \begin{theorem-construction} \label{6.11.02.1} 
855: Let $X$ be a regular complex projective variety. Then 
856: there exists a {\rm canonical} homomorphism of complexes 
857: $$
858: {\cal P}^{\bullet}(n): {\cal Z}^{\bullet}(X; n) 
859: \longrightarrow C^{\bullet}_{{\cal D}}(X; n)
860: $$
861: If $X$ is defined over $\R$ then the image of the map  
862: ${\cal P}^{\bullet}(n)$  lies in  
863: the subcomplex $C^{\bullet}_{{\cal D}}(X_{/\R}; n)$. 
864: \end{theorem-construction}
865: 
866: %The complex $C^{\bullet}_{{\cal D}}(X; n)$ 
867: %is quasiisomorphic to Deligne's complex when $X(\C)$ is compact. 
868: %When $X(\C)$ is non compact 
869: %we define the Arakelov motivic complexes using 
870: %the standard Beilinson--Deligne complexes. 
871: 
872: %\begin{theorem-construction} \label{6.11.02.1s}
873: %If $X$ be an open  regular complex variety then the same construction 
874: %provides a homomorphism of complexes
875: %$$
876: %{\cal P}^{\bullet}(n): {\cal Z}^{\bullet}(X; n) 
877: %\longrightarrow \tau_{\leq 2n}\R^{\bullet}(X; n)_{{\cal D}}
878: %$$
879: %\end{theorem-construction}
880: 
881: To construct this homomorphism we need to define certain 
882: homomorphism $r_{n-1}$ ([G5]). In the next section we 
883: recall its definition  and establish  its basic properties. 
884: Using it we define an $(m-1)$-form 
885: $r_{m-1}(L; H)$ canonically attached to the pair $(\AAA^m; L) = 
886: (\PP^m - H, L)$, and then define the homomorphism ${\cal P}^{\bullet}(n)$.
887: 
888:  
889: {\bf 4. The homomorphism $r_{m-1}$}.    
890: Let $X$ be a variety over $\C$. Let $f_1,...,f_m$ be $m$ rational functions on $X$.
891:  We attach to them the   $(m-1)$-form  
892: \begin {equation} \label{1wq}
893: r_{m-1}(f_1,..., f_m) :=
894: \end {equation}
895: $$
896:  {\rm Alt}_m \sum_{j\geq 0, 2j+1\leq 2m+1} c_{j,m}\log|f_1|d\log|f_2|
897: \wedge ... \wedge d\log|f_{2j+1}|\wedge di\arg f_{2j+2}\wedge ... \wedge
898: di\arg f_{m}
899: $$
900: Here $c_{j,m}:= \frac{1}{(2j+1)!(m-2j-1)!} $ and ${\rm Alt}_m$ is the operation of alternation: 
901: $$
902: {\rm
903: Alt}_m F(x_1,...,x_m):= \sum_{\sigma \in S_m}(-1)^{|\sigma|}F(x_{\sigma
904: (1)},...,x_{\sigma (m)})
905: $$
906: 
907: 
908: 
909: So $r_{m-1}(f_1,..., f_m)$ is an $\R(m-1)$-valued $(m-1)$-form and it
910: is easy to check that 
911: \begin{equation} \label{6.16.04.3}
912: d r_{m-1}(f_1,..., f_m) = \pi_m \Bigl(
913: d \log f_1 \wedge ... \wedge d \log f_m\Bigr) \quad  
914: \end{equation}
915: 
916: The form (\ref{1wq}) is a part of a cocycle representing the product in real Deligne cohomology of
917: 1-cocycles $(\log|f_i|, d\log f_i)$. 
918: 
919: Here is a yet another, a bit more general 
920:  way to look at the homomorphism $r_{m-1}$.  Let 
921: ${\cal A}^i(M)$ be the space of smooth $i$-forms on a real smooth manifold $M$. 
922: Consider the following map
923: \begin{equation} \label{6.16.04.1}
924: \omega_{m-1}: \Lambda^{m}{\cal A}^0(M) \lra {\cal A}^{m-1}(M)
925: \end{equation}
926: $$
927: \omega_{m-1}(\varphi_1 \wedge ... \wedge \varphi_{m}) := 
928: $$
929: $$
930: \frac{1}{m!}{\rm Alt}_{m}
931: \Bigl(\sum_{k=1}^{m}
932: (-1)^{k-1}\varphi_1\partial \varphi_2 \wedge ... \partial \varphi_k \wedge \overline 
933: \partial \varphi_{k+1} \wedge ... \wedge \partial \varphi_{m}\Bigr)
934: $$
935: For example 
936: $$
937: \omega_0(\varphi_1) = \varphi_1;\qquad 
938: \omega_1(\varphi_1 \wedge \varphi_{2}) = \frac{1}{2}\Bigl(\varphi_1\partial \varphi_2  - 
939: \varphi_2\partial \varphi_1 - \varphi_1\overline \partial \varphi_2  +
940: \varphi_2\overline \partial \varphi_1   \Bigr)
941: $$
942: Then one easily checks that 
943: \begin{equation} \label{6.16.04.2}
944: d\omega_{m-1}(\varphi_1 \wedge ... \wedge \varphi_{m}) = 
945: \partial \varphi_1 \wedge ... \wedge \partial \varphi_{m} 
946: +(-1)^m \overline \partial \varphi_1 \wedge ... \wedge \overline \partial \varphi_{m} +
947: \end{equation}
948: $$
949: \sum_{i=1}^{m}(-1)^{i}\overline \partial \partial \varphi_i \wedge 
950: \omega_{m-2}(\varphi_1 \wedge ... \wedge \widehat \varphi_i \wedge ... \wedge \varphi_{m})
951: $$
952: 
953: 
954: 
955: 
956: Now let $f_i$ be rational functions on 
957: a complex algebraic variety $X$. Set $M:= X^0(\C)$, where $X^0$ is the open part 
958: of $X$ where the functions $f_i$ are regular. Then $\varphi_i := \log|f_i|$ 
959: are smooth functions on $M$, and we have an identity 
960: $$
961: \omega_{m-1}(\log |f_1| \wedge ... \wedge \log |f_{m}|) = r_{m-1}(f_1 \wedge ... \wedge f_{m})
962: $$
963: 
964: Observe that $\overline \partial \partial \log|f| = 0$ 
965: on $X^0(\C)$. Therefore the second term in  the formula (\ref{6.16.04.2}) is zero, and so this formula 
966: is consistent with the one (\ref{6.16.04.3}). Notice however that 
967: if we understood
968: $\overline \partial \partial \log|f| $ as a distribution on $X(\C)$, then 
969: by the Poincar\'e-Lelong formula one has 
970: \begin{equation} \label{6.16.04.4}
971: 2\overline \partial \partial \log|f| = 2\pi i \delta(f)
972: \end{equation}  
973: 
974: Our next goal is to interpret the form $r_{m-1}(f_1 \wedge ... \wedge f_{m})$ as a distribution 
975: on $X(\C)$ and calculate the differential of this distribution, taking into account 
976: formula  (\ref{6.16.04.4}). 
977: 
978: 
979: {\bf 5. The distribution $r_{m-1}(f_1 \wedge ... \wedge f_{m})$}. 
980: Recall (see for instance [S]) 
981: that for a subvariety $Y$ of a smooth complex variety $X$ 
982: we define the $\delta$-distribution $\delta_{Y}$ by setting 
983: $$
984: <\delta_{Y}, \omega>:= \int_{Y^0(\C)} \omega
985: $$ 
986: where $Y^0$ is the nonsingular part of $Y$. 
987: 
988: \begin{theorem} \label{8.6.02.2}
989:  Let $Y$ be an arbitrary irreducible subvariety of a smooth complex variety 
990: $X$ and $f_1, ..., f_m \in {\C}^*(Y)$. Then for any smooth differential 
991: form $\omega$
992: with compact support on $X(\C)$ the following integral is convergent:
993: $$
994: \int_{Y^0(\C)}r_{m-1}(f_1, ..., f_m) \wedge i_Y^0\omega
995: $$ 
996: Here $Y^0$ is the nonsingular part of $Y$ and $i_Y^0\omega$ is the
997: restriction of the form $\omega$ to $Y^0(\C)$. 
998: Thus the form $r_{m-1}(f_1,..., f_m)$ defines a distribution 
999: $r_{m-1}(f_1, ...,f_m)\delta_Y$ on 
1000: $X(\C)$ given by 
1001: $$
1002: <r_{m-1}(f_1, ...,f_m)\delta_Y, \omega>:= 
1003: \int_{Y^0(\C)}r_{m-1}(f_1, ..., f_m) \wedge i_Y^0\omega
1004: $$
1005:  It provides  a group homomorphism 
1006: \begin{equation} \label{2}
1007: r_{m-1}: \Lambda^m\C(Y)^{\ast} \longrightarrow {\cal D}^{m-1}_{X(\C)}(m-1) 
1008: \end{equation} 
1009: \end{theorem}
1010: 
1011: {\bf Proof}. We need the following lemma
1012: 
1013: \begin{lemma} \label{8.6.02.112} 
1014: Let $Y$ be a smooth complex projective variety. Then 
1015: for any non zero rational functions $f_1,...,f_m$ on $X$ and 
1016: for any smooth form $\omega$
1017: with compact support on $Y(\C)$ the integral 
1018: $$
1019: \int_{Y(\C)}r_{m-1}(f_1, ..., f_m) \wedge \omega
1020: $$
1021: is convergent. So the form
1022: $r_{m-1}(f_1, ..., f_m)$ defines a distribution 
1023: on $Y(\C)$. 
1024: \end{lemma}
1025: 
1026:  
1027: 
1028: {\bf Basic example}. The integral $\int_{\C}\log|z|d\log(z-a) \wedge
1029: d\log \overline {(z-b)}$ is 
1030: divergent at infinity, where all the functions $z, z-a, z-b$ have a
1031: simple pole, since $\int_{\C}\log|z|\frac{dz\wedge d\overline z}{|z|^2}$ 
1032: is divergent (both near zero and infinity). However 
1033: $$
1034: 4\cdot \int_{\C}\log|z|d\log|z-a| \wedge
1035: d\log |z-b| = 
1036: $$
1037: $$
1038: \int_{\C}\log|z|\Bigl( d\log(z-a) \wedge d\log \overline {(z-b)} \quad + \quad d\log \overline {(z-a)} \wedge d\log  (z-b)\Bigr)
1039: $$ 
1040: is convergent: the divergent parts cancel each other. 
1041: In $r_{m-1}(f_1,...,f_{m})$ such divergences
1042: cancel because of multiplicativity and skew-symmetry of $r_{m-1}$.
1043: 
1044: {\bf Proof of Lemma \ref{8.6.02.112}}.    
1045: Resolving singularities we reduce the statement of the lemma to the case when divisors 
1046: ${\rm div}f_i$ 
1047: have normal crossing. Using the fact that  $r_{m-1}$ is a homomorphism 
1048: to differential forms 
1049: we may suppose that these divisors are 
1050: different.  Our statement is local, so we can assume  that in
1051: local coordinates $z_1,...,z_m$ one has $f_1 = z_1,...,f_k = z_k$ and
1052: ${\rm div}f_j$ for $j>k$ does not intersect  the origin.
1053: After this the statement of lemma is obvious: each term in (\ref{1wq})
1054: defines
1055: a distribution near the origin. For instance the worst possible singularities have 
1056: the term $\log|z_1|d\log|z_2|
1057: \wedge ... \wedge d\log|z_{k}|\wedge \omega$ where $\omega$ is smooth
1058: near the origin. It is clearly integrable with a smooth test form. The lemma is proved.
1059: 
1060: {\bf Remark}. In particular if ${\rm dim}_{\C}X=n$  the integral 
1061: \begin{equation} \label{a5}
1062: \int_{X(\C)}r_{2n}(f_1,...,f_{2n+1})
1063: \end{equation}
1064: is convergent. 
1065: 
1066: 
1067: 
1068: 
1069: Below we use the following form of the resolution of singularities
1070: theorem. Recall that the proper preimage $\widetilde Y$ 
1071: is the closure in $\widetilde X$ 
1072: of the preimage of
1073: the generic part of $Y$.
1074: 
1075: \begin{theorem} \label{8.9.02.4} 
1076: Let $Y$ be an arbitrary 
1077:  subvariety of a regular variety $X$ over a characteristic
1078: zero field, and $Z$ 
1079: a divisor of $Y$. Then there exists a sequence of blow ups 
1080: $\widetilde X \lra X_1 \lra ... \lra  X$ providing a projection 
1081:  $\pi: \widetilde X \to
1082: X$ such that the proper preimage $\widetilde Y$ of $Y$ 
1083: is non singular, 
1084: $\widetilde Z:= p^*Z$, where $p= \pi|_{\widetilde Y}:\widetilde Y \to Y$, 
1085:   is a normal crossing divisor in $\widetilde Y$, 
1086: and restriction of  $\pi$ to the nonsingular part 
1087: of $\widetilde Y- \widetilde Z$ 
1088: is an isomorphism. 
1089: \end{theorem}
1090: 
1091: 
1092: 
1093: 
1094: {\bf Proof of Theorem \ref{8.6.02.2}}. By Theorem \ref{8.9.02.4} 
1095: there exists a sequence of blow ups providing a projection 
1096: $\pi: \widetilde X \lra X$ such that 
1097: the proper preimage  $\widetilde Y$ of $Y \subset X$ is
1098: smooth.  
1099: %Indeed, the sequence of blow ups of $Y$, resolving the singularities
1100: %of $Y$, reinterpreted as a sequence of blow ups of $X$ 
1101: %would do it. 
1102: By the above lemma the integral 
1103: \begin{equation} \label{8.6.02.1}
1104: \int_{\widetilde Y(\C)}r_{m-1}(\pi^*f_1, ..., \pi^*f_m) \wedge \pi^*\omega
1105: \end{equation}
1106: is convergent. Therefore the similar integral
1107: over any Zariski dense subset of $\widetilde Y(\C)$ is also convergent and  
1108: coincides with (\ref{8.6.02.1}). Since 
1109: $\pi$ is an isomorphism on the  nonsingular part of 
1110:  $\widetilde Y- \widetilde Z$, we are done.  
1111: The Theorem \ref{8.6.02.2} is proved. 
1112: 
1113: 
1114: Below we employ notation $r_{m-1}(f_1 \wedge ... \wedge f_m)$ 
1115: for the distribution given by (\ref{2}). 
1116: 
1117: 
1118:  
1119: {\bf 6. Differential of the distribution 
1120: $r_{n-1}(f_1 \wedge ... \wedge f_n)$}. 
1121: Let $X$ be a normal variety. Then there is the residue homomorphism
1122: $$
1123: {\rm Res}: \Lambda^n\C(X)^* \lra \oplus_{Y \subset X^{(1)}}\Lambda^{n-1}\C(Y)^*, \qquad 
1124: $$
1125: where the sum is over all irreducible divisors of $X$. 
1126: 
1127: Here is its definition. Let $K$ be a field with 
1128:  a discrete valuation $v$ and  the residue field $
1129: k_v$.  The group of units $U$ has a natural 
1130: homomorphism $U\longrightarrow k_v^{\ast}\; , \; u\mapsto 
1131: \overline u$.  An element $\pi \in K^{\ast}$ is prime if 
1132: ${\rm ord}_{v}\pi = 1$. There is a homomorphism ${\rm res}_v:\Lambda^{n}K^{\ast} 
1133: \longrightarrow\Lambda^{n-1} k_v^{\ast}$ uniquely defined 
1134: by the  properties $(u_{i}\in U)$: 
1135: $$
1136: {\rm res}_v (\pi\wedge u_{1}\wedge \cdots\wedge u_{n-1}) 
1137: = \overline u_{1}\wedge\cdots \wedge \overline u_{n-1}\quad  \mbox{and}  \quad 
1138: {\rm res}_v
1139: (u_{1}\wedge \cdots \wedge u_{n}) = 0
1140: $$ 
1141: It  does not depend on the choice of $\pi$. 
1142: 
1143: 
1144: 
1145: Observe that if $X$ is normal then 
1146: the local ring of any irreducible divisor of 
1147: $X$ is a discrete valuation ring, so we can apply the above construction. 
1148: We set ${\rm Res}:= \sum {\rm res}_{v}$ where the sum is over 
1149: all valuations of the field $\C(X)$ 
1150: corresponding to the codimension one points of $X$. 
1151: 
1152:  {\bf Remark}. If for any $i$ the restrictions of 
1153: the functions $f_j$ for $j \not = i$ 
1154: to the generic points of all irreducible components of 
1155: the divisor  ${\rm div}f_i$ are non zero, then 
1156: \begin{equation} \label{8.12.02.100}
1157: {\rm Res}(f_1 \wedge ... \wedge f_n) =  \sum_{Y \in X^{(1)}}\sum_{i=1}^n(-1)^{i-1}v_Y(f_i) 
1158: \cdot {f_1}_{|_Y} \wedge ... \wedge \widehat  {f_i}_{|_Y}\wedge ... 
1159: \wedge {f_n}_{|_Y}
1160: \end{equation}
1161: where $v_Y(f)$ is the order of zero of $f$ at the generic point of $Y$.
1162: 
1163: Let $f_1 \wedge ...\wedge  f_n \in \Lambda^n \C(Y)^*$ where $Y$ is a
1164: subvariety of a regular complex variety $X$. We define 
1165: a distribution 
1166: \begin{equation} \label{8.9.02.1}
1167: (r_{n-2}\circ {\rm Res})(f_1 \wedge ...\wedge  f_n)
1168: \end{equation}
1169: on $X(\C)$ as follows. 
1170: 
1171: 
1172: i) If $Y$ is a normal then we have defined 
1173: the residue map ${\rm Res}$ on $\Lambda^n \C(Y)^*$. 
1174: So we define 
1175: (\ref{8.9.02.1}) as 
1176: $$
1177: \sum_{Z \in Y^{(1)}}r_{n-2}{\rm res}_Z(f_1 \wedge ...\wedge  f_n)\delta_Y
1178: $$
1179: i.e. for any smooth form $\omega$ on $X(\C)$ 
1180: $$
1181: <(r_{n-2}\circ {\rm Res})(f_1 \wedge ...\wedge  f_n), \omega> := 
1182: \sum_Z \int_{Z^0(\C)}r_{n-2}{\rm res}_Z(f_1\wedge  ... \wedge f_n)\wedge \omega
1183: $$
1184: 
1185: ii) For an arbitrary $Y$ we take the normalization 
1186: $\pi: Y^{\nu} \to Y$ and define (\ref{8.9.02.1}) as 
1187: $\pi_*((r_{n-2}\circ {\rm Res}) (\pi^*f_1 \wedge ...\wedge \pi^* f_n))$
1188: 
1189: 
1190: 
1191: 
1192: \begin{lemma} \label{Lemma 2.7} Let $Y$ be a subvariety of a regular
1193: complex variety $X$,  $f_1 \wedge ...\wedge  f_n \in \Lambda^n
1194: \C(Y)^*$, and  $Z:= \cup_i {\rm div}f_i$. 
1195:  Let $\pi: \widetilde X \to X$ is a blow up of $X$ 
1196: as in Theorem \ref{8.9.02.4}.  Then 
1197: \begin{equation} \label{8.9.02.3}
1198: \pi_*\Bigl((r_{n-2}\circ {\rm Res})(\pi^* f_1 \wedge
1199: ... \wedge \pi^*  f_n) \Bigr) = 
1200: (r_{n-2}\circ {\rm Res})(f_1 \wedge
1201: ... \wedge f_n) 
1202: \end{equation}
1203: \end{lemma}
1204: 
1205: {\bf Proof}. Thanks to the definition  ii) 
1206: of distribution  (\ref{8.9.02.1}) for non normal varieties 
1207: we may assume without loss 
1208: of generality 
1209: that $Y$ is normal. 
1210: 
1211: 
1212: Let $Z = \cup_{i \in I}Z_i$ be the decomposition of the divisor $Z$ into
1213: irreducible components parametrised by a set $I$. 
1214: By the very definition the right hand side of (\ref{8.9.02.3}) 
1215: is a sum over  $i \in
1216: I$ of 
1217: distributions $\psi_i= r_{n-2}(F_i)$ on
1218: $Z_i$   corresponding to certain elements 
1219: $F_i \in \Lambda^{n-2}\C(Z_i)^*$. 
1220: 
1221: Let $\widetilde Z = \cup_{j \in J}\widetilde Z_j$ 
1222: be the decomposition of  $\widetilde Z$ into
1223: irreducible components parametrised by a set $J$.  
1224: The left hand side of (\ref{8.9.02.3}) 
1225: is a sum over $j \in J$ of distributions $\widetilde \psi_j 
1226:  = r_{n-2}(\widetilde F_j)$  corresponding to certain elements 
1227: $\widetilde F_j \in \Lambda^{n-2}\C(\widetilde Z_j)^*$. 
1228: One has $I \subset
1229: J$ since the  proper preimage $\widetilde Z_i$ 
1230: of $Z_i$ is an irreducible component
1231: of 
1232: $\widetilde Z$.
1233: 
1234: The lemma follows from the following two claims:
1235: $$
1236: \pi_*(\widetilde \psi_i) = \psi_i, \quad i\in I; \qquad 
1237: \pi_*(\widetilde \psi_j) = 0, \quad j\in J - I
1238: $$
1239: The first one is obvious since both distributions 
1240: $\widetilde \psi_i$ and $\psi_i$ are defined  
1241: by their restriction
1242: to any nonsingular Zariski dense open part of the corresponding
1243: divisor, and $\pi$, being  restricted to such a sufficiently small part 
1244: of $\widetilde Z_i$, is an isomorphism. Let us prove the second claim. 
1245: Observe that for $j \in J-I$ the subvariety $\pi(\widetilde Z_j)$ is
1246: of
1247:  codimension 
1248: $2$ in $Y$, and restriction of $\pi$ to a Zariski open part of 
1249: $\widetilde Z_j$ is a
1250: fibration with fibers of positive dimension. We need to show that for
1251: any smooth form $\omega$ on $X(\C)$ 
1252: $$
1253: \int_{\widetilde Z_j(\C)} \widetilde \psi_j \wedge \pi^*\omega 
1254: := \int_{\widetilde Z'_j(\C)} \widetilde \psi_j\wedge \pi^*\omega =0
1255: $$
1256: where $Z_j'$ is a (sufficiently small) 
1257: nonsingular  Zariski dense open part of $Z_j$. 
1258: Observe that $\widetilde \psi_j \wedge \pi^*\omega$ 
1259: is a smooth form on $\widetilde
1260: Z'_j(\C)$. For any vector field
1261: $v$ on $\widetilde Z_j'$ tangent to the fibers of $\pi$  we  have 
1262: $i_v\pi^*\omega =0$. So it suffices to show that 
1263: $i_v \widetilde \psi_j =0$. Since the statement is local, 
1264: we can choose a local equation of an open part of $\widetilde Z'_j$
1265: in the form $\pi^*z =0$. Using it as a local
1266: parameter at the definition of ${\rm res}_{\pi^*z =0}$, we see that 
1267: ${\rm res}_{\pi^*z =0}(\pi^*f_1 \wedge ... \wedge \pi^* f_n)$ is
1268: lifted from $Y$, and hence so is 
1269: $$
1270: \widetilde \psi_j = r_{n-2}
1271:  {\rm
1272:   res}_{\pi^*z =0}
1273: (\pi^*f_1 \wedge ... \wedge \pi^* f_n)
1274: $$ 
1275: Thus  $i_v \widetilde \psi_j =0$. The lemma is proved. 
1276: 
1277: 
1278: \begin{proposition} \label{8.9.02.6}  Let $Y$ be an arbitrary 
1279:  subvariety of a regular
1280: complex variety $X$ and   $f_1 \wedge ...\wedge  f_n \in \Lambda^n
1281: \C(Y)^*$. Then
1282: $$
1283: d r_{n-1}(f_1\wedge ... \wedge f_n) = 
1284: $$
1285: \begin{equation} \label{p106}
1286: \pi_n \Bigl(d \log f_1 
1287: \wedge ... \wedge d \log f_n\Bigr) 
1288: + 2 \pi i \cdot (r_{n-2}\circ {\rm Res})
1289: (f_1\wedge ... \wedge f_n)
1290: \end{equation}
1291: \end{proposition}
1292: 
1293: {\bf Proof}. Let us resolve singularities as in Lemma \ref{Lemma 2.7}. 
1294: Since $\pi$ is a birational isomorphism,
1295:  and the distribution 
1296: $r_{n-1}(f_1 \wedge ... \wedge f_n)$ is determined by its 
1297: restriction to the generic point, 
1298: one has 
1299: $$
1300: \pi_*r_{n-1}( \pi^* f_1 \wedge ... \wedge \pi^*  f_n) = 
1301: r_{n-1}( f_1 \wedge ... \wedge f_n)
1302: $$
1303: So this and 
1304:  Lemma  \ref{Lemma 2.7} imply that we may assume that $\cup_i {\rm div}f_i$ 
1305: is a normal crossing divisor. The proposition 
1306: follows immediately from the Poincar\'e-Lelong formula  (\ref{6.16.04.4})
1307:  $$
1308: d(d i 
1309: \arg f) = 2 \pi i \delta(f) := 2 \pi i \delta_{{\rm div}(f)}
1310: $$
1311: The Proposition \ref{8.9.02.6} is proved. 
1312: 
1313: {\bf Remark}. To prove the  Poincar\'e-Lelong formula one may 
1314: resolve the singularities and argue just as in the proof of Lemma 
1315: \ref{Lemma 2.7} that 
1316: it is sufficient to prove this formula on a blow up. When the divisor of $f$  
1317: is a normal 
1318: crossing divisor the formula follows from 
1319: $$
1320: d(d i 
1321: \arg z) = 2 \pi i \delta(z):= 2 \pi i (\delta_0 - \delta_{\infty}
1322: )$$
1323:  
1324: 
1325: 
1326: 
1327: 
1328: {\bf 7. The distribution $r_{m-1}(L;H)$}. Let $\Omega_L$ be the canonical
1329: $m$-form in ${\Bbb P}^m - L$ with logarithmic singularities
1330: at $L$. It represents a generator of $H^m_{{\rm DR}}({\Bbb P}^m - L)$ defined over $\Z$.  Let us  give its coordinate description. Choose homogeneous coordinates
1331: $(z_0:...:z_m)$ in $\PP^m$ such that 
1332: $L_i $ is given by equation $\{z_i = 0\}$. Then 
1333: $$
1334: \Omega_L =  d\log z_1/z_0 \wedge ... \wedge d\log
1335: z_m/z_0
1336: $$
1337: The form $\Omega_L$ has periods in $\Z(m)$.
1338: So
1339: $\pi_m(\Omega_L)$ is exact. However
1340: there is no 
1341: canonical choice of a primitive $(m-1)$-form for it: the group 
1342: $(\C^{\ast})^m$ acting on $\C \PP^m - L$ leaves the form 
1343: invariant and acts non trivially on the
1344: primitives. 
1345: But if we consider a simplex $L$ in the affine complex space $\AAA^m$
1346: (or, what is the same, choose an additional hyperplane $H$ in  $\C \PP^m$, which should be thought of as the infinite hyperplane) then there
1347: is a {\it canonical}  primitive.
1348: 
1349:   
1350:  
1351: Choose a coordinate system $(z_0:...:z_m)$ in $\PP^m$ as above 
1352: such that $H$ is given by  $\{\sum_{i=1}^{m}
1353: z_i =z_0\}$. Set
1354: \begin{equation} \label{mint4}
1355: r_{m-1}(L;H) := r_{m-1}(z_1/z_0 \wedge ... \wedge  z_m/z_0)
1356: \end{equation}
1357: Here is a more invariant definition. Choose one of the faces 
1358: of the simplex $L$, say 
1359:  $L_0$. Consider the simplex $(H, L_1, ... , L_m)$.
1360: Let $f_i$ be the rational function on $\C \PP^m$ such that 
1361: $(f_i) = L_i - L_0$ normalized by 
1362: $f_i(l_i)= 1$, where $l_i$ is the vertex of the simplex 
1363: $(H, L_1, ... , L_m)$  opposite to the face
1364: $L_i$.  Then $f_i = \frac{z_i}{z_0}$ and 
1365: $$
1366: r_{m-1}(L;H) := r_{m-1}(f_1 \wedge ... \wedge  f_m)
1367: $$
1368: 
1369: \begin{figure}[ht]
1370: \centerline{\epsfbox{gpol5.eps}}
1371: \caption{A simplex $L$ and an infinite hyperplane $H$}
1372: \label{gpol5}
1373: \end{figure}
1374: This  form
1375:  is skewsymmetric with respect to the permutation of the 
1376: hyperplane faces of the simplex $L$. 
1377: One has  
1378: $$
1379: d r_{m-1}(L;H) = \pi_m (\Omega_L) \quad \mbox{in} \quad \C \PP^m - L
1380: $$
1381: So a choice of an ``infinite''  hyperplane $H \subset \C 
1382: \PP^{m}$ provides the form 
1383: $r_{m-1}(L;H)$.
1384: 
1385: {\bf Example}. If $m=1$ then 
1386: $$
1387: \AAA^1 = \PP^1 - \{1\}, \quad L = \{0\} \cup \{\infty \}, \quad  \Omega_L = d\log z, \quad \pi_1 (d\log z) = d\log |z|, \quad 
1388: $$
1389: $$
1390: r_0(\{0\} \cup \{\infty \}; \{1\}) =  \log |z|
1391: $$
1392: 
1393: 
1394: 
1395: 
1396: The $(n-1)$-form  $ r_{n-1}(L;H)$ provides 
1397: an $(n-1)$-distribution on $\C \PP^n$.
1398: Recall   the simplex $\widehat  L_i$ which is cuted out
1399:  by $L$ in the hyperplane $L_i$, and put 
1400: $H_i := L_i \cap H$. Consider the $(n-2)$-form 
1401: $r_{n-2}(\widehat  L_i; H_i )$ on the hyperplane  
1402: $L_i$  as $n$-distribution in $\C \PP^n$. We denote 
1403: it as $r_{n-2}(\widehat  L_i; H_i ) \cdot \delta_{L_i}$.  
1404: 
1405: 
1406: \begin {corollary} \label{dif}
1407: One has
1408: \begin{equation} \label{1p1}
1409: d r_{n-1}(L; H) =  \pi_n ( \Omega_L ) + 2 \pi i \cdot 
1410: \sum_{i=0}^n (-1)^ir_{n-2}(\widehat  L_i; H_i ) \delta_{L_i}
1411: \end{equation}
1412: \end {corollary} 
1413:  
1414: {\bf Proof}. Follows immediately from Proposition \ref{8.9.02.6}. 
1415: 
1416: 
1417: {\bf 8.   A coordinate free description of  the form $r_{m-1}(L;H)$}. 
1418: Let $V_m$ be an $m$-dimensional vector space over a field $F$. 
1419: Choose a volume form ${\rm vol}_m \in {\rm det}V_m^{\ast}$. Set
1420: $\Delta(v_1,...,v_m):= \langle{\rm vol}_m, v_1\wedge ...\wedge v_m\rangle \in F^{\ast}$.
1421: 
1422: \begin {lemma} For a configuration $(l_0,...,l_m)$ of $m+1$ vectors in
1423: generic position 
1424: \begin{equation} \label{4p}
1425: f_m(l_0,...,l_m):= \sum_{i=0}^m(-1)^i \Lambda_{j\not = i}\Delta(l_0,...,\widehat  l_j,..., 
1426: l_i,...,l_m) \in \Lambda^m F^{\ast}
1427: \end{equation}
1428: does not depend on the choice of the volume form ${\rm vol}_m$.
1429: \end {lemma}
1430: 
1431: {\bf Proof}. See proof of Lemma 3.1 in [G3].
1432: 
1433: For a point $z\in \AAA^m- L$ let $l_i(z)$ be the
1434: vector from 
1435: $z$ to the vertex $l_i$, see Figure \ref{gpol4}. 
1436: 
1437: 
1438: \begin{figure}[ht]
1439: \centerline{\epsfbox{gpol4.eps}}
1440: \caption{The vectors $l_i(z)$}
1441: \label{gpol4}
1442: \end{figure}
1443: We get a canonical element 
1444: $$
1445: f_m(l_0(z),...,l_m(z)) \in \Lambda^m\Q(\AAA^m -  L)^{\ast}
1446: $$
1447: If $F=\C$, applying the homomorphism $r_m$ to this 
1448: element we get a canonical $(m-1)$-form in $\C \PP^{m} - L$.
1449: It  coincides with  $ r_{m-1}(L;H)$. 
1450: 
1451: {\bf Example}. If $m=1$ and $(L_0, L_1, H)= (0, \infty, 1)$ then $t = \frac{z}{z-1}$ 
1452: is an affine coordinate on $\PP^1 - \{1\}$ and $$
1453: l_0(t) = \frac{z}{z-1}, 
1454: l_1(t) = \frac{1}{z-1}, \quad \mbox{so} \quad  
1455: f_1(l_0(t), l_1(t)) = \frac{l_0(t)}{l_1(t)} = z
1456: $$ 
1457: 
1458: {\bf Remark}. The map $z\in \AAA^m -  L \lms f_m(l_0(z),...,l_m(z))$ 
1459: provides an isomorphism 
1460: $F_m: CH^m({\rm Spec}(F),0) \lra K_m^M(F)$; 
1461: see [NS] where the isomorphism $F_m$ was presented in a bit different way. 
1462: 
1463: 
1464: 
1465: 
1466: 
1467: 
1468: 
1469: 
1470: 
1471: 
1472: 
1473: {\bf 9. The main construction}. We have to construct a morphism of complexes
1474: $$
1475: \begin{array}{ccccccccc}
1476: ... &\longrightarrow&{\cal Z}^{1}(X; n)&\longrightarrow  &...&\longrightarrow&{\cal Z}^{2n-1}(X; n)&\longrightarrow&{\cal Z}^{2n}(X; n)\\
1477: &&&&&&&&\\
1478: &&\downarrow {\cal P}^{1}(n)&&... &&\downarrow {\cal P}^{2n-1}(n)&&\downarrow 
1479: {\cal P}^{2n}(n)\\
1480: &&&&&&&&\\
1481: 0&\lra &{\cal D}^{0,0}_{\R}(n-1)&  \stackrel{}{\longrightarrow}&...&\stackrel{}{\longrightarrow}&{\cal D}_{\R}^{n-1,n-1}(n-1)&\stackrel{2 \overline \partial \partial}{\longrightarrow}&{\cal D}_{\R}^{n,n}(n)
1482: \end{array}
1483: $$
1484: Let $Y \in {\cal Z}^{2n}(X; n)$ 
1485: be a codimension $n$ cycle in $X$. By definition 
1486: $$
1487: {\cal P}^{2n}(n)(Y):= (2\pi i)^n\delta_Y%, \quad \mbox{where} \quad <\delta_Y,\omega>:= \int_{Y(\C)} \omega
1488: $$
1489: 
1490: Let us construct homomorphisms
1491: $$
1492: {\cal P}^{2n-i}(n):  {\cal Z}^{2n-i}(X; n) 
1493: \longrightarrow {\cal D}^{2n-i-1}_{X(\C)}(n-1), \quad i>0 
1494: $$
1495:  Denote by  $\pi_{{\AAA}^i}$ (resp. $\pi_{X}$) the projection of 
1496: $X \times \AAA^i$ to $\AAA^i$ (resp. $X$), and by 
1497:  $\overline \pi_{{\AAA}^i}$ (resp. $\overline \pi_{X}$) the projection of 
1498: $X \times \C{\Bbb P}^i$ to $\C{\Bbb P}^i$ (resp. $X$). 
1499: 
1500: Recall the element 
1501: \begin{equation} \label{mint3}
1502: \frac{z_1}{z_0} \wedge... \wedge \frac{z_i }{z_0}
1503: \in \Lambda^{i}\C(\AAA^i)^*
1504: \end{equation}
1505: defining the form $r_{i-1}(L;H)$, see (\ref{mint4}). Let 
1506: \begin{equation} \label{mint2}
1507: g_1 \wedge... \wedge g_i \in \Lambda^{i}\C(Y)^*
1508: \end{equation} 
1509:  be the restriction
1510: to $Y$ of the inverse image of element 
1511: (\ref{mint3}) by the projection $\pi^*_{\AAA^i}$. 
1512: The element (\ref{mint2}) provides, by 
1513: Theorem \ref{8.6.02.2}, a distribution on $X(\C) \times \C {\Bbb P}^i$. 
1514: Pushing this distribution down by $(2\pi i)^{n-i}\cdot\pi_{\overline X}$ we get 
1515: the distribution $ {\cal P}^{2n-i}(n)(Y)$:
1516: 
1517: \begin{definition} \label{mint6}
1518: $$
1519: {\cal P}^{2n-i}(n)(Y) := (2\pi i)^{n-i}\cdot
1520: \pi_{\overline X *}r_{i-1}(g_1 \wedge... \wedge g_i) :=
1521: $$
1522: $$
1523: (2\pi i)^{n-i}\cdot \pi_{\overline X *}r_{i-1}\Bigl( i_Y^*\pi^*_{\AAA^i}
1524: (\frac{z_1}{z_0} \wedge... \wedge \frac{z_i }{z_0})\Bigr)
1525: $$
1526: \end{definition}
1527: Here $i_{Y}: Y 
1528: \hookrightarrow X \times  {\Bbb P}^i$
1529: 
1530: {\bf Remark}. This definition works if and only if 
1531: the cycle $Y$ has proper intersection with all codimension one faces 
1532: of $X \times L$. Indeed, if $Y$ does not have proper intersection 
1533: with one of the faces, then the equation of this face restricts to zero 
1534: to $Y$, and so (\ref{mint2}) does not make sence. 
1535: As soon as all equations of the 
1536: codimension one faces restrict to non zero functions on $Y$, 
1537: (\ref{mint2}) makes sense, and we can apply  Theorem \ref{8.6.02.2}. 
1538: 
1539: 
1540: 
1541: 
1542: 
1543: {\bf Remark}. We just proved that the product of distributions $\delta_Y \wedge \pi_{\AAA}^{\ast}r_{i-1}(L;H)$ makes sense and 
1544: $$
1545: {\cal P}^{2n-i}(n)(Y)= \quad (2\pi i)^{n-i}
1546: {\pi_X}_*(\delta_Y \wedge \pi_{\AAA}^{\ast}r_{i-1}(L;H))
1547: $$
1548: 
1549: It is handy to rewrite Definition \ref{mint6} more explicitly 
1550: as an  integral over $Y(\C)$. Namely,  
1551: %let $k= 2 {\rm dim}Y - i +1$. 
1552: %Observe  that 
1553: %$$
1554: %2 {\rm dim}X - k = 2({\rm dim}X +i - {\rm dim}Y) - i - 1 = 2n-i-1
1555: %$$ 
1556: %So $(2n-i-1)$-distributions are functionals on  smooth $k$-forms with 
1557: %compact support on $X(\C)$. 
1558: %Let $\omega$ be such a $k$-form and
1559: let $\omega$ be a smooth form on $X(\C)$ and 
1560: $Y \in  {\cal Z}^{2n-i}(X; n)$. Then 
1561: \begin{equation} \label{mint}
1562: <{\cal P}^{2n-i}(n)(Y),\omega>= \quad (2\pi i)^{n-i}
1563: \int_{Y(\C)}\pi_{\AAA}^{\ast}r_{i-1}(L;H) \wedge \pi_X^{\ast}\omega =  
1564: \end{equation}
1565: \begin{equation} \label{8.9.02.10}
1566: (2\pi i)^{n-i}\int_{Y^0(\C)}  
1567: r_{i-1}(g_1\wedge ... \wedge g_i)\wedge i_{Y(\C)}^*\pi^*_X\omega 
1568: \end{equation}
1569: where $Y^0$ is the nonsingular part of $Y$
1570: 
1571:  
1572: 
1573: 
1574: 
1575:  
1576: Since the form $r_{i-1}(L;H)$ is $\R(i-1)$-valued, for $i>0$ the
1577:  distribution
1578:  ${\cal P}^{2n-i}(n)(Y)$ takes values in $\R(n-1)$. 
1579: Further,  ${\cal P}^{2n}(n)(Y)$ is obviously an $\R(n)$-valued distribution. 
1580: 
1581: Let us show that for $i>0$ the distribution ${\cal P}^{2n-i}(n)(Y)$ lies precisely 
1582: in the left bottom $(n-1) \times (n-1)$ square of  the Dolbeault
1583: bicomplex. The integral (\ref{mint}) 
1584: is non zero only if 
1585: $\pi^*_{X}\omega \wedge \pi^*_{{\Bbb A}}r_{i-1}(L;H)$ is of type 
1586: $$
1587: ({\rm dim}Y, {\rm dim}Y) =  ({\rm dim}X+i-n, {\rm dim}X+i-n) 
1588: $$
1589: Since $r_{i-1}(L;H)$ is an 
1590: $(i-1)$-form we see that the integral vanish if $\omega$ is 
1591: a form of type $(p,q)$ where 
1592: $p$ or $q$ is smaller then ${\rm dim} X+1-n$. This just means that 
1593: the distribution lies in the left bottom $(n-1)\times (n-1)$ square of 
1594: the Dolbeault bicomplex. 
1595: The proposition is proved.
1596: 
1597: Therefore we have constructed the maps  ${\cal P}^{i}(n)$. 
1598: 
1599: \begin{theorem} \label{hhmmoo}
1600: ${\cal P}^{\bullet}(n)$ is a homomorphism of complexes.
1601: \end{theorem}
1602: 
1603: {\bf Proof}. One has
1604: \begin{equation} \label{98}
1605: <d {\cal P}^{2n-i}(n)(Y), \omega> = 
1606: %= - <{\cal P}^{2n-i}(n)(Y), d\omega> = 
1607: %$$
1608: %$$- \int_{\overline Y(\C)}i_{\overline Y}(\overline \pi_X^{\ast}d\omega \wedge
1609: %\overline \pi_{\AAA}^{\ast}r_{i-1}(L;H)) = - \int_{\overline Y(\C)}i_{\overline Y}^*\p%i_X^*d\omega \wedge 
1610: %r_{i-1}(g_1\wedge... \wedge g_i) = 
1611: %$$
1612: %
1613: \int_{\overline Y(\C)}dr_{i-1}(g_1\wedge
1614: ... \wedge g_i)\wedge 
1615: i_{\overline Y}^*\pi_X^*\omega 
1616: \end{equation}
1617: where $g_1\wedge
1618: ... \wedge g_i$ is as in (\ref{mint2}). 
1619: We use Proposition \ref{8.9.02.6} to calculate 
1620: $dr_{i-1}(g_1\wedge... \wedge
1621: g_i)$. 
1622: 
1623: 
1624: To handle the first term in (\ref{p106}) observe that 
1625: by the very definition of the complex 
1626: ${\cal C}_{\cal D}(X;n)$ we need to investigate integral (\ref{98}) 
1627: only for smooth forms $\omega$ of type $(p,q)$ where $|p-q| \leq i-1$. 
1628: Since $\pi_i( \Omega_L ) = \Omega_L \pm \overline 
1629: \Omega_L$ is a sum of forms of type $(i,0)$
1630: and $(0,i)$ 
1631: the form $\pi_{\AAA}^{\ast}\pi_i(\Omega_L)
1632: \wedge \pi_X^{\ast}\omega$ 
1633: can not be of type $(k,k)$.  
1634: Therefore only the second term in (\ref{p106}) contributes. 
1635: 
1636:  
1637: 
1638: 
1639: Since $r_0(f) = \log|f|$, the commutativity of the last square 
1640: follows from the Poincar\'e-Lelong formula 
1641: $
1642: 2 \overline \partial \partial \log |f| = 2 \pi i \cdot \delta_{{\rm div}(f)} 
1643: $. 
1644:  
1645: 
1646: 
1647: {\it Commutativity of the $i$-th square of the diagram, 
1648: $i>1$, counting from the right}. 
1649: If $\overline Y$ is  normal this  
1650: follows from Proposition \ref{8.9.02.6}. Indeed, 
1651: since $Y$ meets the codimension two faces properly all equations of the codimension one
1652: faces of  $X \times L$ but  $z_j$ have non zero restriction 
1653: to the generic point of ${\rm div}z_j$. Therefore we may use formula 
1654: (\ref{8.12.02.100}) to calculate ${\rm Res}$, and then the claim is obvious. 
1655:  
1656: In  the case when $\overline Y$ is not normal 
1657: we face the following subtle problem.
1658: 
1659:  Calculating 
1660: $dr_{i-1}(g_1\wedge... \wedge
1661: g_i)$ and hence $d \circ {\cal P}^{2n-i}(n)(Y)$ when $Y$ is not
1662: normal we need to take 
1663: ${\rm Res}_{\widetilde Z}(\widetilde g_1\wedge... \wedge
1664: \widetilde g_i)$ for all irreducible divisors $\widetilde Z$ in the normalization
1665: $\pi: \widetilde Y \to Y$, where $\widetilde g:= \pi^*g$,
1666: and then take
1667: \begin{equation} \label{8.9.02.7}
1668: \sum_{\widetilde Z} \pi_*r_{i-2}\circ {\rm Res}_{\widetilde Z}
1669: (\widetilde g_1\wedge... \wedge
1670: \widetilde g_i)
1671: \end{equation}
1672: Computation of ${\cal P}^{2n-i+1}(n)(Z)\circ d $ 
1673: does not involve the normalization of $Y$: we intersect
1674: $Y$ with all codimension one faces of $X \times L$. 
1675: So we need to compute (\ref{8.9.02.7}) using the intersection data of $Y$ and
1676: $X \times L$. 
1677: 
1678: 
1679: 
1680: To handle this  
1681:  we use  the condition that $Y$ meets the codimension two faces of
1682: the simplex $X \times L$ properly. It implies that 
1683: \begin{equation} \label{8.9.02.20}
1684: {\rm Res}_{\widetilde Z}(\widetilde g_1\wedge... \wedge
1685: \widetilde g_i) = \pi^*G_{i-1}; \qquad  
1686: G_{i-1}\in \Lambda^{i-1}\C(\pi(\widetilde Z))^*
1687: \end{equation} 
1688: Indeed, since all equations of the codimension one
1689: faces of  $X \times L$ but $z_j$ have non zero restriction 
1690: to the generic point of ${\rm div}z_j$,  the wedge 
1691: product of these restrictions can be taken as $G_{i-1}$. 
1692: 
1693: Having  (\ref{8.9.02.20}) the statement is obvious since 
1694: $$
1695: \pi_* r_{i-2} \pi^*G_{i-1} = [\C(\widetilde Z): \C(\pi (\widetilde Z))]
1696: \cdot r_{i-2}(G_{i-1})
1697: $$ 
1698: Indeed, recall (see [F], page 9) that if $\widetilde Y \to Y$ is the
1699: normalization of  $Y$ and $g \in \C(Y)^* = \C(\widetilde Y)^*$ then 
1700: $$
1701: {\rm ord}_Z(g) = \sum_{\widetilde Z} 
1702: {\rm ord}_{\widetilde Z}(g)[\C(\widetilde Z): \C(Z)]
1703: $$
1704: where the sum is over all irreducible 
1705: divisors projecting onto $Z$. 
1706: 
1707: Theorem \ref{hhmmoo} is proved. Therefore we finished the proof of 
1708:  Theorem-Construction 
1709: \ref{6.11.02.1}. 
1710: 
1711:  
1712: 
1713: {\bf 10. The Higher Arakelov Chow groups}. Let $X$ be a regular complex variety. 
1714: Denote by  $\widetilde  C^{\bullet}_{{\cal D}}(n)$ the quotient of the complex $C^{\bullet}_{{\cal D}}(n)$ along the subgroup ${\cal A}_{cl}^{n,n}(n) \subset 
1715: {\cal D}_{cl}^{n,n}(n) $ of closed smooth form of type $(n,n)$ with values in $\R(n)$. 
1716: 
1717: Consider the cone of the homomorphism ${\cal P}^{\bullet}(n)$ shifted by $-1$:
1718: $$
1719: \widehat  {\cal Z}^\bullet(X; n):= {\rm Cone}
1720: \Bigl( {\cal Z}^\bullet(X; n)
1721:  \stackrel{ }{\longrightarrow} \widetilde  C^{\bullet}_{{\cal D}}(X(\C); n)
1722: \Bigr)[-1]
1723: $$
1724: 
1725: \begin{definition} \label{arackd} The Higher Arakelov Chow groups are 
1726: \begin{equation} \label{6.11.02.11}
1727: \widehat  {CH}^n(X; i):= H^{2n-i}(\widehat  {\cal Z}^{\bullet}(X; n))
1728: \end{equation}
1729: \end{definition}
1730: 
1731: 
1732: 
1733:  Recall the arithmetic Chow groups  defined by Gillet-Soule [GS] as follows:
1734: $$
1735: \widehat  {CH}^n(X):= 
1736: $$
1737: \begin{equation} \label{6.11.02.10}
1738: \frac{\{(Z,g); \frac { \overline \partial \partial}{\pi i} g+\delta_Z 
1739: \in {\cal A}^{n,n}\}}{\{(0,\partial u + \overline \partial v); ({\rm div} 
1740: f, -\log|f|), f \in \C(Y), 
1741: {\rm codim} (Y) = n-1\}}
1742: \end{equation} 
1743: Here $Z$ is a divisor in $X$, $f$ is a rational function on a divisor $Y$ in $X$, 
1744: $$
1745: g \in {\cal D}_{\R}^{n-1,n-1}(n-1), \quad (u,v) \in ({\cal D}^{n-2,n-1} 
1746: \oplus {\cal D}^{n-1,n-2})_{\R}(n-1)
1747: $$
1748:  
1749: 
1750: 
1751: \begin{proposition} \label{ch}
1752:  $\widehat  {CH}^n(X; 0) = \widehat  {CH}^n(X)$.
1753: \end{proposition}
1754: 
1755: 
1756: {\bf Proof}.  Let us look at the very right part of the complex 
1757: $\widehat  {\cal Z}^{\bullet}(X; n)$:
1758: $$
1759: \begin{array}{ccccc}
1760: ...&\longrightarrow&{\cal Z}^{2n-1}(X; n)&\longrightarrow&{\cal Z}^{2n}(X; n) \\
1761: &&&&\\
1762: &&\downarrow {\cal P}^{2n-1}(n)&& \downarrow {\cal P}^{2n}(n)\\
1763: &&&&\\
1764: ({\cal D}^{n-2,n-1} \oplus {\cal D}^{n-1,n-2})_{\R}(n-1)&\stackrel{(\partial , \overline \partial)}{\longrightarrow}&{\cal D}_{\R}^{n-1,n-1}(n-1)&\stackrel{2\overline \partial \partial}{\longrightarrow}&{\cal D}_{\R}^{n,n}(n)/{\cal A}_{\R}^{n,n}(n)
1765: \end{array}
1766: $$
1767: Consider the very  end of the   Gersten complex on $X$:
1768: $$
1769: \prod_{Y \in X_{n-2}}\Lambda^2\C(Y)^* \stackrel{\partial}{\longrightarrow}  \prod_{Y \in X_{n-1}}\C(Y)^* \longrightarrow {\cal Z}_0(X; n) 
1770: $$
1771: where $\partial$ is the tame symbol. It maps to the complex 
1772: $\widehat  {\cal Z}^{\bullet}(X; n)$, i.e. to the top row  of the bicomplex above, 
1773: as follows. Recall that ${\cal Z}_0(X; n) = {\cal Z}^{2n}(X; n)$, so the very right component of our map is provided by this identification. 
1774: Further, a pair $(Y;f)$ where $Y$ is an irreducible codimension $n-1$ subvariety of $X$ maps to the cycle $\{(y,f(y))| y \in Y\} \subset X \times \AAA^1$.
1775:  Similarly  any element in $\Lambda^2\C(Y)^*$ can be represented as a linear combination
1776:  of elements 
1777: $\sum_i (Y;f_i\wedge g_i)$ where $Y$ is an irreducible codimension $n-2$ subvariety of $X$ and $f_i, g_i$ are rational functions on $Y$ such that ${\rm div} f_i$ and 
1778: ${\rm div} g_i$  share no irreducible divisors. Then we send $(Y;f_i\wedge g_i)$ to 
1779: the cycle $(y,f_i(y),g_i(y)) \subset X \times \AAA^2$. 
1780: It is well known that in this way we get  an isomorphism on the 
1781: last two cohomology groups. 
1782: Computing the composition of this map with the homomorphism 
1783: ${\cal P}^{\bullet}(n)$ we end up precisely with the denominator 
1784: in (\ref{6.11.02.10}). The proposition is proved. 
1785: 
1786: 
1787:  
1788: 
1789: \section  { The Chow polylogarithms}                                                              
1790: 
1791: Suppose $X = {\rm Spec}(\C)$. Then ${\cal P}_n(Y): = {\cal P}^{1}(n)(Y)$ is a function on the space of all codimension $n$ cycles in $\PP^{2n-1}$ intersecting properly faces of the simplex $L$. It is called 
1792: the {\it Chow polylogarithm function}. For $i>1$ all the distributions ${\cal P}^{i}(n)(Y)$ are zero. However modifying 
1793: the construction of the previous chapter we get a very interesting object, the Chow polylogarithm,  even when $X$ is a point. The Chow polylogarithm function is the 
1794:  first component of the Chow polylogarithm. One can define the Chow 
1795: polylogarithm for an arbitrary variety $X$, but we spell out the details 
1796: in the most 
1797: interesting case when $X$ is a point. 
1798: 
1799:     
1800: {\bf 1   Chow polylogarithms [G5]}. Let $L = (L_0, ..., L_{p+q})$ 
1801: be  a simplex in $\PP^{p+q}$,
1802:  $H$ is a hyperplane in generic position to $L$, and  $H_i := H \cap L_i$. 
1803: 
1804: Let $ {\cal  Z}^{q}_{p}(L)$ be the variety of all codimension $q$
1805: effective   algebraic cycles  in 
1806: $\C \PP^{p+q}$ which intersect properly, i.e. each irreducible component in the right codimension, 
1807:  all faces of the simplex $L$. 
1808:  It  is a union of an infinite number of finite
1809: dimensional complex algebraic varieties. 
1810: 
1811: {\bf Example}.   
1812: $ \C \PP^n - L = (\C^{\ast})^n$ is an irreducible component of 
1813: $ {\cal  Z}^{n}_{0}(L)$ parametrizing the irreducible subvarieties, i.e. points. 
1814: 
1815: Let $\widehat L_i$ be the simplex in the projective space $L_i$ cut by the hyperplanes $L_j$, $j \not = i$. 
1816: The intersection of a cycle with a codimension
1817: 1 face $L_i$ 
1818: of the simplex $L$ provides a map
1819: $$
1820: a_i: {\cal  Z}^{q}_{p}(L) \longrightarrow  {\cal  Z}^{q}_{p-1}(\widehat L_i), 
1821: \qquad 0 \leq i \leq p+q 
1822: $$
1823: 
1824:  Let $l_j$ be the vertex opposite to the face $L_j$. 
1825: Consider an open part ${\cal  Z}^{q}_{p}(L)^0$ of ${\cal  Z}^{q}_{p}(L)$ 
1826: parametrising the cycles $C$ such that projection with the center $l_j$ sends $C$ to a codimension $q-1$ cycle. Then 
1827: projection with the center at  the vertex $l_j$ of $L$ 
1828:  defines a  map 
1829: $$
1830: b_j:    {\cal  Z}^{q}_{p}(L)^0 \longrightarrow 
1831: {\cal Z}^{q-1}_{p}(\widehat L_j), \qquad 0 \leq i \leq p+q 
1832: $$
1833: 
1834: 
1835: 
1836: \begin{theorem-construction} \label{chw}
1837: For given  $q \geq 0$ there is an          
1838: explicitly constructed chain of     $(q-p-1)$-distributions  $\omega_{p}^q =
1839: \omega_{p}^q(L; H)$ on ${\cal Z}^{q}_{p}(L)$ such that
1840: 
1841: \begin{equation} \label{i}
1842: i)\qquad d \omega^q_{0}(L,H) = \pi_q( \Omega_L)
1843: \end{equation}                                                  
1844: %r({\Bbb  A}^{q-1};\Delta_{q-1})
1845: \begin{equation} \label{ii}                                    
1846: ii)\quad d\omega_{p}^q(L; H) = \sum_{i=0}^{p+q}(-1)^ia_i^{\ast}\omega_{p-1}^q(L; H_i)
1847: \end{equation}
1848: \begin{equation} \label{iii}                                     
1849: iii)\qquad \sum_{j=0}^{p+q+1}(-1)^j b_j^{\ast}\omega_{p}^q(L; H) = 0
1850: \end{equation}
1851: The restriction of  $\omega_{p}^q$ to the subvariety $\widehat 
1852: {\cal Z}^{q}_{p}(L)$ of smooth cycles
1853: in generic position with respect to the simplex $L$ is a real-analytic
1854: differential $(q-p-1)$ form.   
1855: \end{theorem-construction}
1856: 
1857: For a given positive integer 
1858: $q$ the collection  $\{\omega_{p}^q\}$ is called the {\it $q$-th Chow
1859: polylogarithm}.   
1860: 
1861: The varieties $ {\cal  Z}^{q}_{p}(L)$ for $ p\geq 0$ form a truncated simplicial variety 
1862: $ {\cal  Z}^{q}_{\bullet}(L)$. The conditions i) and ii) just mean that 
1863: the  sequence of forms $\omega_{p}^q$ is 
1864:  a $2q$-cocycle in
1865: the  complex computing the Deligne cohomology $H^{2q}(
1866: {\cal Z}^{q}_{\bullet}(L), \R_{\cal D}(q))$.  
1867: 
1868: 
1869: 
1870: 
1871: {\bf Proof}. 
1872: We define 
1873:  $\omega^q_p$  as the Radon transform of the distribution 
1874: $r_{p+q-1}(L;H)$ in $\C \PP^{p+q}$ over the family of cycles $Y_{\xi}$
1875: parametrized by 
1876: $ {\cal Z}^q_p(L)$. This means the following. Consider the incidence variety:
1877: $$
1878:   \Gamma_p := \{(x,\xi) \in \C\PP^{p+q} \times   {\cal Z}^q_p(L)(\C)
1879: \quad  \mbox{such that} \quad x \in Y_{\xi}\}
1880: $$
1881: where $Y_{\xi}$ is the cycle in $\C \PP^{p+q}$ corresponding to $\xi
1882: \in   {\cal  Z}^q_p(L)$. We get a double bundle
1883: $$
1884: \begin{array}{ccccc}
1885: &&  \Gamma_p \subset \C \PP^{p+q} \times {\cal  Z}^q_p(L)(\C)&&\\
1886: &&&&\\
1887: &\pi_1\swarrow&&\searrow \pi_2&\\
1888: &&&&\\
1889: \C \PP^{p+q}&&&&  {\cal  Z}^q_p(L)(\C)
1890: \end{array}
1891: $$
1892: Then
1893: $$
1894: \omega_p^q:= {\pi_2}_{\ast} {\rm Res}_{\Gamma_p} \pi_1^{\ast} (2\pi i)^{-q}r_{p+q}(L;H) 
1895: $$
1896: Observe that $r_{p+q}(L;H)$ is a distribution on $\C \PP^{p+q}$, 
1897: and hence 
1898: $\pi_1^{\ast} r_{p+q}(L;H)$ is a distribution on 
1899: $\C \PP^{p+q} \times {\cal  Z}^q_p(L)(\C)$. The fact that this 
1900: distribution can be restricted to 
1901: $\Gamma_p$ is a version of Theorem \ref{8.6.02.2}, and is proved in the same way.  
1902: The push forward ${\pi_2}_{\ast}$ of this 
1903: distribution is well defined since  $\pi_2$ is a proper map. 
1904: 
1905: The property i)  is true by the very definition.
1906: 
1907: \begin {lemma} \label{2.4}
1908: $\sum_{j=0}^{n+1}(-1)^j b_j^{\ast}\omega^n_0 = 0$.
1909: \end {lemma}
1910: 
1911: {\bf Proof}. Let $s(z_0,...,z_n):= z_1/z_0 \wedge ... 
1912: \wedge z_{n}/z_0$. The lemma follows from  the identity 
1913: $$
1914: \sum_{j=0}^{n+1}(-1)^j s(z_0,...,\widehat  {z_j},..., z_{n+1}) = 0
1915: $$
1916: So we have iii). 
1917: To check  ii) observe that the push forward ${\pi_2}_{\ast}$ of 
1918: distributions commutes with the De Rham differential.  
1919: The theorem is proved.
1920: 
1921: 
1922: 
1923: 
1924: 
1925: {\bf 2. Properties of the Chow polylogarithm function}. 
1926: The  function
1927: ${\cal P}_q:= \omega_{q-1}^q$  on $ {\cal  Z}^{q}_{q-1}(L)$
1928:  is called the {\it Chow $q$--logarithm function}. 
1929: It satisfies
1930: two functional 
1931: equations:
1932: $$
1933: \sum_{i=0}^{2q}(-1)^ia_i^{\ast}{\cal P}_q =0 \qquad \sum_{j=0}^{2q}(-1)^j
1934: b_j^{\ast}{\cal P}_q = 0 
1935: $$
1936: 
1937: \begin{theorem} \label{6.10.04.1q}
1938: The Chow polylogarithm function is invariant under the natural 
1939: action of the torus 
1940: $(\C^*)^{p+q}$ on ${\cal Z}^q_p(\C)$. In particular it does not 
1941: depend on the choice of the hyperplane $H$. 
1942: \end{theorem}
1943: 
1944: {\bf Remark}. The statements of Theorem \ref{6.10.04.1q} are no longer true 
1945: for the  forms $\omega_{p}^q$ for
1946: $p<q-1$. 
1947: 
1948: Here is a reformulation of Theorem \ref{6.10.04.1q}.
1949: \begin {theorem}  
1950: \label {0.2}
1951: Suppose that $dimX=n$ and $f_1,...,f_{2n+1
1952: }$ are rational functions on $X$. Then
1953:  the integral 
1954: $$
1955: (2\pi i)^{1-n}\int_{X(\C)}r_{2n}(f_1,...,f_{2n+1})
1956: $$
1957: does not change if we multiply one of the
1958: functions $f_i$ by a non zero constant.
1959: \end {theorem}
1960:  
1961: {\bf Proof}. Multiplying, say, $f_1$ by $\lambda$ we see that the difference
1962: between the two integrals is
1963: \begin{equation} \label{a6}
1964: \log|\lambda| \sum_k a_k \int_{X(\C)}  {\rm Alt}_{2n} d\log|f_2|\wedge ... \wedge
1965: d\log|f_{2k-1}| \wedge d\arg f_{2k} \wedge ... \wedge d\arg f_{2n+1}
1966: \end{equation}
1967: where the $a_k$ are some rational constants (easily computable from
1968: (\ref{1wq})). We will prove that for each $k$ the corresponding integral in this sum
1969: is already zero. Using the identity
1970: $$
1971: (d\log|f_2| + i d \arg  f_2) \wedge ... \wedge (d\log|f_{2n+1}| + i d \arg
1972:  f_{2n+1}) = 0
1973: $$
1974: we can rewrite the integral
1975: $$
1976: \int_{X(\C)}   d\arg f_2\wedge ... \wedge d\arg f_{2n+1}
1977: $$
1978: as a sum of similar integrals containing $d\log |f_i|$. Our statement
1979: follows from
1980: \begin {proposition}  \label{5.2}
1981: \label {0.2b} Suppose that $dimX=n$. Then
1982: $$ 
1983:  d {\rm Alt}_{2n}\Bigl(\log|f_2|d\log|f_3|\wedge ... \wedge
1984: d\log|f_{2k-1}| \wedge d\arg f_{2k} \wedge ... \wedge d\arg f_{2n+1}\Bigr) = 
1985: $$
1986: \begin{equation} \label{d6}
1987:   {\rm Alt}_{2n} \Bigl( d\log|f_2| \wedge ... \wedge
1988: d\log|f_{2k-1}| \wedge d\arg f_{2k} \wedge ... \wedge d\arg f_{2n+1}\Bigr)
1989: \end{equation}
1990: in the sense of distributions.
1991: \end {proposition}
1992: 
1993: {\bf Proof}. Since $d di \arg f = 2\pi i \delta(f) $, the left hand
1994: side is equal to the right hand side {\it 
1995: plus} the following terms concentrated on the divisors $f_{2k+j} = 0$:
1996: $$
1997: 4\pi\cdot2(n-k+1) {\rm Alt}_{2n} \Bigl( \delta(f_{2n+1})\log|f_2|d\log|f_3| \wedge ... \wedge
1998: d\log|f_{2k-1}| \wedge d\arg f_{2k} \wedge 
1999: ... \wedge d\arg f_{2n}\Bigr)
2000: $$
2001: However all these additional terms  vanish thanks to the following 
2002: proposition. 
2003: 
2004: 
2005: \begin {proposition}  \label{5.3}
2006: \label {0.2a}
2007: Suppose that $dim X = n$. Then for each $0 \leq j \leq n-1$ one has 
2008: \begin{equation} \label{b6}
2009:  {\rm Alt}_{2n}\Bigl(d \log|f_1|
2010: \wedge ... \wedge d\log|f_{2j+1}| \wedge d\arg f_{2j+2}\wedge ... \wedge
2011: d\arg f_{2n} \Bigr) =0
2012: \end{equation}
2013: \begin{equation} \label{c6}
2014:  {\rm Alt}_{2n}\Bigl(d \log|f_1|
2015: \wedge ... \wedge d\log|f_{2j}| \wedge d\arg f_{2j+1}\wedge ... \wedge
2016: d\arg f_{2n} \Bigr) =
2017: \end{equation}
2018: $$
2019:  \frac{(2n)! {n \choose j}}{{2n \choose 2j}}\cdot d \log|f_1|
2020: \wedge ... \wedge d\log|f_{2n}| 
2021: $$
2022: \end {proposition}
2023: 
2024: 
2025: {\bf Proof}. The idea is this. 
2026: One has $n$ equations 
2027: $$
2028: d\log f_{1} \wedge ... \wedge d\log f_{2n} =0
2029: $$
2030: $$
2031: d\log |f_{1}| \wedge d\log f_{2} \wedge ... \wedge d\log f_{2n} =0
2032: $$
2033: \begin{equation} 
2034: ............
2035: \end{equation}
2036: $$
2037: d\log |f_{1}|\wedge ... \wedge d\log |f_{n-1}| \wedge d\log f_{n}
2038: \wedge ... \wedge d\log f_{2n} = 0
2039: $$
2040: Taking imaginary part of each of them and alternating $f_1,...,f_{2n}$ we
2041: get $n$ linear equations. Solving them we get the proposition. 
2042: See the details in the Appendix in [GZ]. 
2043: 
2044: 
2045: 
2046: \section{The Grassmannian polylogarithms}
2047: 
2048:  
2049: 
2050: {\bf 1. Configurations of vectors and Grassmannians: a dictionary}. 
2051: Let $G$ be a group. Let $X$ be a $G$-set. 
2052: We define {\it configurations of $m$ points of} $X$ as $G$-orbits in $X^m$. 
2053: 
2054: {\bf Example 1}. If $X:= V$ is a vector space and $G:= GL(V)$ we get 
2055: configurations  of vectors
2056: in $V$. 
2057: A configuration of vectors $(l_1,...,l_m)$ is  in generic position 
2058: if each $k\leq {\rm dim}V$ of
2059: the vectors are linearly independent.
2060: 
2061: {\bf Example 2}. If $X:= \PP(V)$ is a projective space and $G:= PGL(V)$ we get 
2062: configurations  $(x_1,...,x_m)$ of $m$ points in $\PP(V)$. 
2063: A  configuration of points is in generic position if each $k\leq {\rm dim}V$ of
2064: them generate a plane of dimension $k-1$. 
2065: 
2066: Let $T_{p+q}$ be the quotient of the torus ${\Bbb G}_m^{p+q+1}$ 
2067: by the diagonal subgroup ${\Bbb G}_m = (t, ..., t)$. Below $V_n$ denotes a vector space of dimension $n$. 
2068: 
2069: \begin{lemma-construction} \label{6.12.02.5}
2070: i) There are canonical isomorphisms  between the following sets of geometric
2071: objects: 
2072: 
2073: (a) Configurations of $p+q+1$ vectors in generic position in $V_q$.
2074: 
2075: (b) Isomorphism classes of triples 
2076: $\{$a projective space ${\PP}^{p+q}$ together with a  
2077: simplex $L$, an ``infinite'' hyperplane $H$ in generic position to $L$, and a $p$-dimensional plane in generic position to $L$ (but not necessarily to $H$) $\}$. 
2078: 
2079: (c) Isomorphism classes of triples 
2080: $\{$a vector space $V_{p+q+1}$, a basis $(e_0,...,e_{p+q})$ of 
2081: $V_{p+q+1}$, and a $p+1$-dimensional subspace  of $V_{p+q+1}$ in generic position with respect to the coordinate hyperplanes$\}$.
2082: 
2083: ii) The torus $T_{p+q}$ acts naturally, and without fixed points,  
2084: on each of the objects 
2085: a), b), c), and the isomorphisms above are compatible with this action. 
2086: \end{lemma-construction}
2087: 
2088: {\bf Proof}. i) (a) $ \rightarrow$ (c). 
2089: For the $(p+1)$-dimensional  subspace, take  the kernel of the linear map from
2090: $V_{p+q+1}$ to $V_q$ sending $e_i$ to $l_i$. 
2091: 
2092: (c) $ \rightarrow$ (a). Take the quotient of $V_{p+q+1}/h$ along the given subspace $h$ and consider the images of the vectors $(e_0,...,e_{p+q})$ there. 
2093: 
2094: (c) $ \rightarrow$ (b). Let $\PP^{p+q}:= \PP(V_{p+q+1})$. Let 
2095:  ${\AAA}^{p+q}$ be the affine hyperplane in $V_{p+q+1}$
2096: passing through the ends of the basis vectors $e_i$. 
2097: Then ${\AAA}^{p+q} \subset \PP^{p+q}$.
2098:  The coordinate hyperplanes in $V_{p+q+1}$ provide a simplex 
2099: $L_{p+q} \subset {\PP}^{p+q}$.
2100: The projectivization of a generic $(p+1)$-dimensional 
2101: subspace $h$ in $V_{p+q+1}$ gives a
2102: $p$-plane $\overline h$ in generic position with respect to this simplex. 
2103: (Notice that we do not impose any condition on the mutual 
2104: location of $H$ and $\overline h$. For instance $\overline h$ may be inside of $H$.)
2105: 
2106: (b) $\rightarrow$ (c). The triple $(\PP^{p+q}, H, L)$ provides a unique up to an isomorphism data 
2107: $(V_{p+q+1}, (e_0,...,e_{p+q}))$. Namely, the partial data $(\PP^{p+q}, H)$ 
2108: provides us with $(V_{p+q+1}, \widetilde  H)$ where $\widetilde  H$ is the subspace of 
2109: $V_{p+q+1}$ projecting to $H$. Now the vertices $l_i$ of the simplex $L$ provide coordinate lines $\widetilde  l_i$ in 
2110: $V_{p+q+1}$. Intersecting these coordinate lines with a parallel shift of the subspace $\widetilde  H$ we get a point  on each of the coordinate lines. By definition the endpoints of the basis vectors $e_i$ are these points. Taking the  subspace $\widetilde  h$ in $V_{p+q+1}$ projecting to a given $p$-plane $h$ in 
2111: $\PP^{p+q}$ we get the desired correspondence. 
2112: 
2113: 
2114: 
2115: 
2116:  ii) The torus $T_{p+q}$ acts on the 
2117: configurations of vectors in a) 
2118: as $$ 
2119: (t_1, ..., t_{p+q+1}):  
2120: (l_1, ..., l_{p+q+1}) \lms (t_1l_1, ..., t_{p+q+1} l_{p+q+1})
2121: $$
2122: The torus $T_{p+q}$ is identified with $\PP^{p+q} - L$ in b), 
2123: and so acts naturally on the data in b). The action on the data in c) is similar. 
2124:  The lemma is proved. 
2125: 
2126: If we use the description c)  for  the Grassmannians then 
2127: $b_j$ is obtained by factorization along the coordinate
2128: axis $\langle e_j \rangle$. 
2129: 
2130: 
2131: 
2132: {\bf 2. The Grassmannian and bi-Grassmannian polylogarithms}. 
2133: Let us fix a positive integer $q$. 
2134: The operations $a_i$ and $b_j$ from Section 3.1 transform planes to planes. So we get
2135: the following diagram of varieties called the 
2136:  bi-Grassmannian $\widehat  G(q)$:
2137: $$
2138: \begin{array}{cccccccccccc}
2139: &&&&&&\downarrow...\downarrow&&\downarrow...\downarrow\\
2140: &&&&&&&&&\\
2141: \widehat G(q):=&&&&&\stackrel{\rightarrow}{\stackrel{...}{\rightarrow}}&\widehat G_1^{q+1}&\stackrel{\rightarrow}{\stackrel{...}{\rightarrow}}&\widehat  G_0^{q+1}\\
2142: &&&&&&&&&\\
2143: &&&&\downarrow...\downarrow&&\downarrow...\downarrow&&\downarrow...\downarrow\\
2144: &&&&&&&&&\\
2145: &&...&\stackrel{\rightarrow}{\stackrel{...}{\rightarrow}}&\widehat  G_2^{q}&\stackrel{\rightarrow}{\stackrel{...}{\rightarrow}}&\widehat  G_1^{q}&\stackrel{\rightarrow}{\stackrel{...}{\rightarrow}}&\widehat  G_0^{q}\\
2146: \end{array}
2147: $$
2148: Here the  horizontal arrows are the maps $a_i$ and the vertical ones
2149: are $b_j$.
2150: 
2151: {\bf Remark}. The bi-Grassmannian $\widehat  G(n)$ is not a (semi)bisimplicial
2152: scheme. (It is a truncated semi{\it hypersimplicial} scheme. 
2153: See s.2.6 in [G4]).
2154: 
2155: {\it Configurations of  hyperplanes  and torus quotients of Grassmannians}. 
2156: Let $\widehat  G_p^q$ be the 
2157: Grassmannian of $p$-planes in ${\Bbb  P}^{p+q}$ 
2158: in generic position with respect to a given simplex $L$.
2159: 
2160: Taking the $T_{p+q}$--orbits of the objects 
2161: a) and b) in  the lemma  we arrive at   
2162: 
2163: 
2164: \begin{corollary} \label{6.13.02.1}
2165: There is a 
2166:  bijective correspondence
2167: $$
2168: \widehat G^q_p/T_{p+q} 
2169: <->  \left \{ \mbox{Configurations of $ p+q+1$ generic hyperplanes in } \PP^{p}\right \}
2170: $$
2171: sending a $p$--plane $h$ to the configuration  $
2172: (h \cap L_0,...,h \cap L_{p+q})$      in $h$.
2173:   \end{corollary} 
2174: 
2175: Let $\psi^q_{p}(q)$ be the  restriction of the differential form
2176: $\omega_{p}^q$ to $\widehat  G^{q}_p$. The properties i), ii) from Theorem \ref{chw} 
2177: are exactly the defining conditions for the single--valued 
2178: Grassmannian polylogarithm whose existence was conjectured in 
2179:  [HM], [BMS], see also [GGL]. 
2180: 
2181: Let us extend these forms by zero 
2182: to the other rows of the bi-Grassmannian $\widehat G(q)$, i.e.   
2183: set $\psi^{q+i}_{p}(q) =0$ if $i>0$.   Then the 
2184: property iii) from Theorem \ref{chw} guarantees that the forms $\psi^{q+i}_{p}(q)$ 
2185: form  a  $2q$-cocycle in
2186: the  bicomplex computing the Deligne cohomology $H^{2q}(\widehat 
2187: G(q)_{\bullet}, \R(q)_{\cal D})$. It is called {\it the
2188: bi-Grassmannian $q$-logarithm}. 
2189: ( [G5]). 
2190: 
2191:  
2192: 
2193: A sequence of {\it  
2194: multivalued analytic } forms on Grassmannians satisfying conditions
2195:  similar to  i),  ii) was defined  in [HaM1], [HaM2].
2196: Another construction of the multivalued analytic Grassmannian 
2197: polylogarithms was suggested in [G5] in the more general setting 
2198: of the multivalued Chow polylogarithm. 
2199: 
2200: 
2201: 
2202: 
2203: 
2204: {\bf 3. The Grassmannian $n$-logarithm function}. 
2205: By  Theorem \ref{0.2} the Chow polylogarithm function is invariant under the action 
2206: of the torus $(\C^*)^{2n-1}$. So restricting it  to the open Grassmannian $\widehat  G^n_{n-1} \subset 
2207: \widehat  {\cal Z}^n_{n-1}$ and using  the bijection
2208: $$
2209: \mbox{ $\{ (n-1)$--planes in 
2210:  $\PP^{2n-1}$ in generic position with respect to a simplex $L$\}}/({\Bbb G}_m^*)^{2n-1} \quad 
2211: $$
2212: $$  <--> \quad \left \{ \mbox{Configurations of} \quad 2n \quad\mbox {generic 
2213: hyperplanes in } \PP^{n-1}\right \}
2214: $$
2215: we get a function on the configurations of $2n$ hyperplanes in $\C \PP^{n-1}$, called 
2216: the Grassmannian polylogarithm function ${\cal L}^G_n$. 
2217: 
2218: The Grassmannian polylogarithm function has the following  simple description on 
2219: the language of configurations of hyperplanes. It is intresting that in this description we can  
2220: work with {\it any}  configuration of $2n$ 
2221: hyperplanes,   assuming nothing about their mutual   location. 
2222:  
2223:  Let $h_1,...,h_{2n}$ be $2n$ arbitrary hyperplanes in $\C \PP^{n-1}$. 
2224: Choose an additional hyperplane $h_0$. Let $f_i$ be a
2225: rational function on $\C \PP^{n-1}$ with divisor $(h_i) - (h_0)$. It is 
2226: defined up to a scalar
2227: factor. 
2228: Set
2229: $$
2230: {\cal L}^G_n(h_1,...,h_{2n}):= (2\pi i)^{1-n}\int_{\C \PP^{n-1}}r_{2n-2}
2231: (\sum_{j=1}^{2n}(-1)^j f_1\wedge ...
2232: \wedge \widehat  f_j \wedge ... \wedge f_{2n})
2233: $$
2234: It is skewsymmetric by definition.
2235: Notice that 
2236: $$
2237: \sum_{j=1}^{2n}(-1)^j f_1\wedge ...
2238: \wedge \widehat  f_j \wedge ... \wedge f_{2n} = \frac{f_1}{f_{2n}} \wedge
2239: \frac{f_2}{f_{2n}} \wedge ... \wedge \frac{f_{2n-1}}{f_{2n}} 
2240: $$
2241: So we can define ${\cal L}^G_n(h_1,...,h_{2n})$ as follows: choose rational
2242: functions $g_1,...,g_{2n-1}$ such that ${\rm div} g_i = (h_i) - (h_{2n})$ and put
2243: $$
2244: {\cal L}^G_n(h_1,...,h_{2n}) = (2\pi i)^{1-n}
2245: \int_{\C \PP^{n-1}}r_{2n-2}(g_1,...,g_{2n-1})
2246: $$
2247: 
2248: {\bf Remark}. The function ${\cal L}^G_n$ is defined on the set of all configurations of $2n$ hyperplanes in $\C\PP^{n-1}$. However it is 
2249: not even continuous on this set. It is real analytic on the submanifold of generic 
2250: configurations. Since we put no restrictions on the hyperplanes $h_i$ 
2251: the following theorem is stronger than Theorem \ref{chw} in the case of 
2252: linear subvarieties. 
2253: 
2254: \begin {theorem}  
2255: \label {0.3a} The function ${\cal L}^G_n$ has  the following properties:
2256: 
2257: a) It does not depend on the choice of hyperplane $h_0$.
2258: 
2259: b) For any $2n+1$ hyperplanes $(h_1, ..., h_{2n+1})$ in $\C \PP^{n}$ one has
2260: \begin{equation} \label{4}
2261:  \sum_{j=1}^{2n+1}(-1)^j {\cal L}^G_n(h_j \cap h_1,..., \widehat {h_j \cap h_j}, ..., h_j \cap h_{2n+1}) =0
2262: \end{equation}
2263: 
2264: c) For any $2n+1$ hyperplanes $(h_1, ..., h_{2n+1})$ in $\C \PP^{n-1}$ one has
2265: \begin{equation} \label{5}
2266:  \sum_{j=1}^{2n+1}(-1)^j {\cal L}^G_n(h_1,...,\widehat  h_j,...,h_{2n+1})=0
2267: \end{equation}
2268: \end {theorem}
2269: 
2270:  
2271: 
2272: {\bf Proof}. a) Choose another hyperplane $ h_0'$. Take a rational
2273: function $f_0$ with divisor $(h_0') - (h_0)$. Set $f_i' = \frac{f_i}{f_0}$. Then
2274: $$
2275: \sum_{j=1}^{2n+1}(-1)^j f_1\wedge ...
2276: \wedge \widehat  f_j \wedge ... \wedge f_{2n+1} - \sum_{j=1}^{2n+1}(-1)^j
2277:  f_1' \wedge ...
2278: \wedge \widehat  { f_j'} \wedge ..., \wedge  f_{2n+1}' = 0
2279: $$
2280: 
2281: Indeed, substituting $f_i' = \frac{f_i}{f_0}$ in this formula we find
2282: that the only possible 
2283: nontrivial term $f_0 \wedge f_1 \wedge ... \wedge \widehat  f_i \wedge ...
2284: \wedge \widehat  f_j \wedge ... \wedge f_{2n}$ vanishes because it is symmetric
2285: in $i,j$. 
2286: 
2287: 
2288: b) Let $g_1,...,g_{2n+1}$ be rational functions on $\C \PP^{n}$ with
2289: ${\rm div} g_i = (h_i) - (h_0)$. 
2290: Then
2291: \begin{equation} \label{8}
2292:  dr_{2n-1}\Bigl(\sum_{j=1}^{2n+1}(-1)^j g_1 \wedge ...
2293: \wedge \widehat  g_j \wedge ... \wedge g_{2n+1}\Bigr) = 
2294: \end{equation}
2295: $$
2296: \sum_{j \not = i}(-1)^{j+i-1} 2 \pi i \delta(f_i) \wedge r_{2n-2}\Bigl(
2297: g_1\wedge ... \wedge \widehat  g_i \wedge ...
2298: \wedge \widehat  g_j \wedge ... \wedge g_{2n+1}\Bigr)
2299: $$
2300: (Notice that $d\log g_1 \wedge ... \wedge \widehat  {d\log g_j} \wedge ...
2301: \wedge d\log g_{2n+1} =0$ on $\C \PP^{n}$). 
2302: Integrating (\ref{8}) over $\C \PP^{n}$ we see that the left hand side equals
2303:  zero, while the right hand side equals to the sum of the expressions staying 
2304: on the left of (\ref{4}). So we get b).
2305: 
2306: c) is obvious: we apply $r_{2n-2}$ to the zero element.
2307: Theorem is proved.
2308: 
2309: 
2310: {\bf 4. ${\Bbb P}^1 - \{0, \infty\}$ as a 
2311: special stratum in the configuration space of $2n$ points in $\PP^{n-1}$}. 
2312: A {\it special configuration}  is 
2313: a configuration of $2n$  points  
2314: \begin{equation} \label{--00}
2315: (l_0,...,l_{n-1}, m_0,...,m_{n-1}) 
2316: \end{equation} 
2317: in $\PP^{n-1}$ such that  $l_0,...,l_{n-1}$ are vertices of a 
2318: simplex in $\PP^{n-1}$ and $ m_i$ is a point 
2319: on the edge $l_il_{i+1}$ of the simplex different from $l_i$ and 
2320: $l_{i+1}$,  as on the picture. 
2321: 
2322: \begin{figure}[ht]
2323: \centerline{\epsfbox{fig1coc.eps}}
2324: \caption{A special configuration of $8$ points in $\PP^{3}$.}
2325: \label{fig1coc}
2326: \end{figure}
2327: 
2328: 
2329: 
2330: \begin{proposition} \label{6.12.02.1}
2331: The set of special configurations of $2n$ points in $\PP^{n-1}$
2332: is canonically identified with $\PP^1 \backslash \{0, \infty \}$. 
2333: \end{proposition}
2334: 
2335: \begin{figure}[ht]
2336: \centerline{\epsfbox{gpol12.eps}}
2337: \caption{$\PP^1 - \{0, \infty\}$ is a stratum in the configuration space of $2n$ points in $\PP^{n-1}$.}
2338: \label{gpol12}
2339: \end{figure}
2340: 
2341: 
2342: 
2343: {\bf Proof}. 
2344: We define {\it generalized cross-ratio} 
2345: $$
2346: r(l_0,...,l_{n-1},m_0,...,m_{n-1})  \in  F^*,
2347: $$ 
2348: where $F$ is the common field of definition of the points $l_i, m_j$, as follows. 
2349: Consider the one-dimensional subspaces $L_i, M_j$ in the $n$-dimensional vector space $V$ projecting  to the points $l_i,m_j$ in $\PP^{n-1}$ respectively. The 
2350: subspaces $L_i, M_i, L_{i+1}$ generate a two dimensional subspace. Its  quotient
2351:  along   $M_i$ can be identified with $L_i$ as well as with $L_{i+1}$. 
2352: So we get a canonical linear map 
2353: $
2354: \overline M_i: L_i   \longrightarrow L_{i+1}
2355: $. 
2356: The composition of these maps (the ``linear monodromy'')
2357: $$
2358: \overline M_0 \circ    ... \circ \overline M_{n-1} : L_0 \longrightarrow L_0
2359: $$
2360:  is   multiplication by  an element of $F^*$ 
2361: called the generalized cross-ratio of the special configuration (\ref{--00}).
2362: 
2363: \begin{figure}[ht]
2364: \centerline{\epsfbox{gpol7.eps}}
2365: \caption{The generalized cross-ratio of a special configuration}
2366: \label{gpol7}
2367: \end{figure}
2368: 
2369: 
2370: It is clearly invariant under the cyclic permutation  
2371: $$
2372: l_0 \to l_0 \to ... \to l_{n-1}\to l_0; \quad m_0 \to m_1 \to ... \to m_{n-1}\to m_0
2373: $$
2374:   Notice that $r(l_0,...,l_{n-1},m_0,...,m_{n-1}) =1$
2375: if and only if the points $m_0,...,m_{n-1}$ belong to a hyperplane.
2376: 
2377: 
2378: 
2379: 
2380: Let $\widehat  m_i$
2381: be the point of intersection of the line $l_il_{i+1}$ with the
2382: hyperplane passing through all the points $m_j$ except $m_i$. 
2383: Let $r(x_1, ..., x_4)$ be the cross-ratio of the four points $x_i$ on $\PP^1$. Then 
2384: $$
2385: r(l_0,...,l_{n-1},m_0,...,m_{n-1}) =
2386: r(l_i,l_{i+1},m_i,\widehat  m_{i+1})
2387: $$ 
2388: 
2389: 
2390: {\it The special configurations and classical polylogarithms}. 
2391: Consider the configuration of $2n$  hyperplanes in $\PP^{n-1}$ given by the following 
2392: equations in homogeneous coordinates $z_0: ... :z_{n-1}$
2393: $$
2394: z_0 = 0, \quad ..., \quad z_{n-1} = 0, \quad z_0 = z_1, \quad z_1 + z_2 = z_0,
2395: $$
2396: \begin{equation} \label{ewa99}
2397:  z_2-z_3=0,\quad  ..., \quad z_{n-2} - z_{n-1} = 0, \quad z_{n-1} = az_0
2398: \end{equation}
2399: It admits the following interpretation. Recall that the classical
2400: polylogarithm function 
2401: $Li_{n-1}(z)$ can be defined by an iterated integral:
2402: $$
2403: Li_{n-1}(a) = 
2404: \int_0^a\frac{dt}{1-t} \circ \frac{dt}{t} \circ ... \circ \frac{dt}{t}  = 
2405: \int_{\Delta_a}\frac{dz_1}{z_1} \wedge ... \wedge \frac{dz_{n-1}}{z_{n-1}}
2406: $$
2407: If $a \in (0,1]$, then  the simplex $\Delta_a$ is defined by the equations
2408: $$
2409: \Delta_a:= \quad \{(z_1,...,z_{n-1}) \in \R^{n-1}| \quad 
2410: 0 \leq 1-z_1 \leq z_2 \leq z_3 \leq ... \leq z_{n-1} \leq a\} 
2411: $$
2412: The faces of the simplex $\Delta_a$ can be defined for arbitrary $a$. 
2413: Then the codimension one faces $\{z_i =0\}$ of the coordinate 
2414: simplex 
2415: and the codimension one faces of the simplex $\Delta_a$ 
2416: form the configuration (\ref{ewa99}). 
2417: 
2418: 
2419: 
2420: 
2421: We can  
2422: reorder hyperplanes of this configuration as follows:
2423: $$
2424: z_0 =0, \quad z_1 =0, \quad z_1 =z_0, \quad z_1+z_2 = z_0, \quad z_2 =0, 
2425: $$
2426: $$
2427: z_2 = z_3, \quad z_3 =0, \quad ...\quad , \quad z_{n-2} =0, \quad 
2428: z_{n-2} = z_{n-1}, \quad z_{n-1} = a z_0
2429: $$
2430:  Applying the projective duality to this configuration of hyperplanes  
2431: we get the special configuration of $2n$ points in $\PP^{n-1}$ 
2432: with the generalized cross ratio $a$. 
2433: 
2434: The correspondence between the configuration (\ref{ewa99}) and the 
2435: special configuration of points is illustrated in the case $n=3$ 
2436: on Figure \ref{gpol11.fig}.
2437: 
2438: \begin{figure}[ht]
2439: \centerline{\epsfbox{gpol11.fig.eps}}
2440: \caption{Classical polylogarithm configurations and special configurations}
2441: \label{gpol11.fig}
2442: \end{figure}
2443: 
2444: {\bf Remark}. It is amusing that the special configuration of 
2445: $2n$ points in $\PP^{n-1}$, 
2446: which is related to the classical $n$-logarithm  by Theorem  \ref{clpoly} below,
2447:  is constructed using the geometry of the mixed motive corresponding to $Li_{n-1}(a)$. 
2448: 
2449: 
2450: 
2451: 
2452: 
2453: {\bf 5. Restriction of the Grassmannian $n$--logarithm to the special stratum}. 
2454: The function $Li_n(z)$ has a  remarkable  single-valued
2455: version  ([Z1], [BD]): 
2456: \begin{eqnarray*} 
2457: {\cal L}_{n}(z) &:=& \begin{array}{ll} 
2458: {\rm Re} & (n:\ {\rm odd}) \\ 
2459: {\rm Im} & (n: \ {\rm even}) \end{array} 
2460: \left( \sum^{n-1}_{k=0} \beta_k
2461: \log^{k}\vert z\vert \cdot Li_{n-k}(z)\right)\; , \quad n\geq 
2462: 2 \\ 
2463: \end{eqnarray*}           
2464: It is continuous on $\C \PP^1$. Here   $\frac{2x}{e^{2x} -1}  =
2465: \sum_{k=0}^{\infty}\beta_k x^k $, so $\beta_k = \frac{2^kB_k}{k!}$ 
2466: where the $B_k$ are the Bernoulli numbers. For example ${\cal L}_2(z)$
2467: is the Bloch - Wigner function.
2468: 
2469: Let us consider the following modification of the 
2470:  function ${\cal L}_n(z)$ proposed by A. M. Levin in [Le]:
2471: $$
2472: \widetilde {\cal L}_{n}(x):= 
2473: $$
2474: $$
2475: \frac{(2n-3)}{(2n-2)}
2476: \sum_{\mbox{$k$ even;  $0 \leq k \leq n-2$}}
2477: \frac{2^k (n-2)!(2n-k-3)!}{(2n-3)!(k+1)!(n-k-2)!} {\cal L}_{n-k}(x)\log^k|x|
2478: $$
2479: For example $\widetilde {\cal L}_{n}(x) =  {\cal L}_{n}(x)$ for $n \leq 3$, 
2480: but already $\widetilde {\cal L}_{4}(x)$ is different from ${\cal L}_{4}(x)$. 
2481: A direct integration carried out in Proposition 4.4.1 of [Le] 
2482: shows that  
2483: $$
2484: -(2\pi i)^{n-1}(-1)^{(n-1)(n-2)/2}\widetilde {\cal L}_{n}(x) =
2485: $$
2486: $$
2487: \int_{\C \PP^{n-1}}\log|1-z_1| \prod_{i=1}^{n-1}d\log|z_i| \wedge \prod_{i=1}^{n-2}d\log|z_{i} - z_{i+1}| 
2488: \wedge d\log|z_{n-1} -a| 
2489: $$
2490: This combined with Proposition \ref{0.3} below implies 
2491: 
2492: \begin {theorem}  \label{clpoly}
2493: The value of the function  ${\cal L}^G_n$ on the special configuration
2494: (\ref{--00}) is equal to $$
2495: -(-1)^{n(n-1)/2}4^{n-1}{2n-2\choose n-1}^{-1}\widetilde {\cal L}_{n}(a)
2496: $$ where $a = r(l_0,...,l_{n-1},m_0,...,m_{n-1})  $.
2497: \end {theorem}
2498: 
2499: Another proof in the case $n=2$ is given in Proposition 6.8. 
2500: 
2501: 
2502: \begin {conjecture} \label{conj}
2503: The Chow $n$-logarithm function can be expressed by the
2504: Grassmannian $n$-logarithm function.
2505: \end {conjecture}
2506: 
2507: 
2508: 
2509: 
2510: 
2511:  {\bf Remark}. Suppose that an element $\sum_k\{f^{(k)}_1,...,f^{(k)}_{2n+1}\}
2512: \in K^M_{2n+1}(\C(X))$ has zero residues at all the divisors on an
2513: $n$-dimensional variety $X$ over $\C$. Then it defines an element 
2514: $$
2515: \alpha \in gr^{\gamma}_{2n+1}K_{2n+1}(X) = Ext^{2n+1}_{{\cal M}}(\Q(0)_X,\Q(2n+1)_X) 
2516: $$
2517:  Its direct image to the point is
2518: an element 
2519: $$
2520: \pi_{\ast}(\alpha) \in gr^{\gamma}_{n}K_{2n+1}(Spec \C) = Ext^{1}_{{\cal M}}(\Q(0),\Q(n))
2521: $$
2522:  Applying the regulators we see that the
2523: integral
2524: $\sum_k \int_{X(\C)}r_{2n}(f^{(k)}_1,...,f^{(k)}_{2n+1})$ coincides, up to a factor, 
2525:  with
2526: the value of the Borel regulator map on $\pi_{\ast}(\alpha)$ and so  by
2527: results of the next chapter is 
2528: expressible by the Grassmannian $n$-logarithms. Conjecture  \ref{conj}
2529: tells us that this should be true {\it for any} element in
2530: $K^M_{2n+1}(\C(X))$.
2531: 
2532: 
2533: 
2534: 
2535: 
2536: \section{Grassmannian polylogarithms, symmetric 
2537: spaces and Borel regulators}
2538:  
2539: {\bf 1.  The function $\psi_n$.} Let $V_n$ be a complex 
2540: vector space of dimension $n$. 
2541: Let
2542: $$
2543: \HH_n: = \left \{ \mbox{ positive definite Hermitian forms in  
2544:    $V_n$ }\right \}/\R ^*_+ = SL_n(\C)/SU(n) 
2545: $$
2546: $$
2547: = \left \{ \mbox{ positive definite Hermitian forms in
2548:    $V_n$ with determinant } =1 \right \} 
2549: $$
2550: It is a symmetric space    of rank $n-1$.  For example 
2551: $\HH_2 = {\cal H}_3$ is the hyperbolic 3-space. 
2552: Replacing positive definite by non-negative definite Hermitian forms 
2553: we get a compactification $\overline \HH_n$ of the symmetric space $\HH_n$. 
2554: 
2555:        
2556:   
2557: Let $G_x$ be the subgroup of $SL_N(\C)$ stabilizing the point  $x \in \HH_n$ . 
2558: A point $x$ defines a one-dimensional vector space $M_x$:
2559: $$
2560: x \in \HH_n \longmapsto M_x:= \left \{ \mbox{measures on    } \C \PP^{n-1} %\subset \partial \HH_n   
2561: \mbox{ invariant under }  G_x\right \}
2562: $$
2563:    Namely, a point $x$ corresponds to a hermitian metric in $V_n$.  
2564: This metric provides the Fubini-Studi metric on $\C \PP^{n-1} =P(V_n)$.  
2565: Moreover there is the Fubini-Studi K\"ahler form on $\C \PP^{n-1} = P(V_n)$; 
2566: its imaginary part is a symplectic form. Raising it to 
2567: $(n-1)$-th power we get the Fubini-Studi volume form.
2568: The elements of   $M_x$ are the multiples of  the Fubini-Studi volume form. 
2569: 
2570: 
2571: 
2572: So $\HH_n$  embeds
2573:  into the projectivization of the space of all measures in $\C \PP^{n-1}$. 
2574:  Taking its closure we get a compactification of $\HH_n$. 
2575:   
2576: 
2577:   
2578:  Let us choose for any point $x \in \HH_n$ an invariant measure $\mu_x \in M_x$. 
2579:   Then, for any $y \in {\Bbb H}_n$, the ratio     $ \mu_x/\mu_y$ is a real function on $\C \PP^{n-1}$.
2580: 
2581: Let $x_0,...,x_{2n-1}$ be points  of the symmetric space $SL_n(\C)/SU(n)$.
2582: Consider the following function
2583:   \begin{equation} \label{1221q} 
2584: \psi_n(x_0,...,x_{2n-1}) := \int_{\C \PP^{n-1}}
2585:  \log | \frac{\mu_{x_1}}{\mu_{x_0}}| d\log|\frac{\mu_{x_2}}{\mu_{x_0}} | \wedge ... \wedge d\log| \frac{\mu_{x_{2n-1}}}{\mu_{x_0}} |
2586: \end{equation} 
2587: 
2588:  
2589: 
2590: 
2591: 
2592: 
2593:   {\bf 2. General properties of the function $\psi_n$}.   Let us study the properties of 
2594: integral (\ref{1221q}) in a more general situation. Let $X$ be an $m$-dimensional manifold. For any $m+2$ measures $\mu_0,...,\mu_{m+1}$
2595:  on $ X$ such that $\frac{\mu_{i}}{\mu_j}$ are smooth functions we can construct a differential $m$-form  on $X$:
2596: $$
2597: \overline r_{ m}(\mu_0:...:\mu_{ m+1}) := \log | \frac{\mu_1}{\mu_0}| d\log|\frac{\mu_2}{\mu_0} | \wedge ... \wedge d\log| \frac{\mu_{ m+1}}{\mu_0} |
2598: $$
2599: 
2600: \begin{proposition} \label{1.2.}
2601: The integral 
2602: \begin{equation} \label{i2}
2603: \int_{ X}\overline r_{ m}(\mu_0:...:\mu_{ m+1}) 
2604: \end{equation}
2605: satisfies the following properties:
2606: 
2607: 1) Skew symmetry with respect to the permutations of $\mu_i$.
2608: 
2609: 2) Homogeneity:
2610: $$
2611: \int_{ X}\overline r_{ m}(\lambda_0  \mu_0: ... :\lambda_{ m+1}   \mu_{ m+1}) = \int_{X }\overline r_{ m}(\mu_0:...:\mu_{ m+1}) 
2612: $$
2613: 
2614: 3)Additivity: for any $ m+3$ measures $\mu_i$ on $X $ one has
2615: $$
2616:   \sum_{i=0}^{m+2} (-1)^i \int_{X }\overline r_{ m}(\mu_0: ... :\widehat \mu_i: ... :\mu_{ m+2}) =0
2617: $$
2618: 
2619: 4)  Let $g$ be a diffeomorphism of $X$.  Then   
2620: $$
2621: \int_{X }\overline r_{ m}(g^*\mu_0: ... :g^*\mu_{ m+1})  = \int_{X }\overline r_{ m}( \mu_0: ... : \mu_{ m+1})
2622: $$
2623: \end{proposition}
2624: 
2625: {\bf Proof}. 1). Follows from $\log f \cdot d\log g + \log g \cdot d\log f = 
2626: d (\log f \cdot \log g)$.
2627: 
2628: 2) Using 1) we may assume $\lambda_i =1$ for $i>0$. Then 
2629: $$
2630: \int_{ X} (\overline r_{ m}(\lambda_0\mu_0:\mu_1: ...:\mu_{ m+1}) - \overline r_{ m}(\mu_0:\mu_1 :...:\mu_{ m+1})) = 
2631: $$
2632: $$
2633: -  \log |\lambda| \cdot \int_{ X} 
2634: d(\log| \frac{\mu_2}{\mu_0}| d \log| \frac{\mu_3}{\mu_0}| \wedge ... \wedge d\log| \frac{\mu_{m+1}}{\mu_0}|) =0
2635: $$
2636: 
2637: 3) Taking into account the skewsymmetry of the integral we  have  to prove that 
2638: \begin{equation} \label{12321}
2639: {\rm Alt}_{(0,..., m+2)} \left \{
2640: \log | \frac{\mu_2}{\mu_1}| d\log| \frac{\mu_3}{\mu_1}| \wedge ... \wedge d\log| \frac{\mu_{m+2}}{\mu_1} | \right \} = 0
2641: \end{equation}
2642:     Let us write $ \frac{\mu_i}{\mu_j} = \frac{\mu_i}{\mu_0}/ \frac{\mu_j}{\mu_0}$  and substitute it into  (\ref{12321}). Then the    terms in (\ref{12321})  where
2643: $$
2644: \log | \frac{\mu_2}{\mu_0}| d\log| \frac{\mu_3}{\mu_0}| \wedge ... \wedge d\log| \frac{\mu_{m+2}}{\mu_0} | 
2645: $$
2646:  will appear look as follows:
2647: $$
2648: \log | \frac{\mu_2}{\mu_1}| d\log| \frac{\mu_3}{\mu_1}| \wedge ... \wedge d\log| \frac{\mu_{m+2}}{\mu_1} | - \log | \frac{\mu_1}{\mu_2}| d\log| \frac{\mu_3}{\mu_2}| \wedge ... \wedge d\log| \frac{\mu_{m+2}}{\mu_2} | 
2649: $$
2650: $$
2651: -\log | \frac{\mu_2}{\mu_0}| d\log| \frac{\mu_3}{\mu_0}| \wedge ... \wedge d\log| \frac{\mu_{m+2}}{\mu_0} | + \log | \frac{\mu_0}{\mu_2}| d\log| \frac{\mu_3}{\mu_2}| \wedge ... \wedge d\log| \frac{\mu_{m+2}}{\mu_2} | 
2652: $$
2653: (The first two terms comes from ${\rm Alt}_{(1,..., m+2)}\overline r_m(\mu_1:  ... :\mu_{m+2})$ and the second two from ${\rm Alt}_{(0,2,..., m+2)}\overline r_m(\mu_0: \mu_2: ... :\mu_{m+2})$. The expression ${\rm Alt}_{ m+2}\overline r_m(\mu_0: ... :\widehat  \mu_i: ... :\mu_{m+2})$ provides no such terms if $i>1$). 
2654: 
2655: 4) Clear. The proposition is proved. 
2656: 
2657: 
2658: Recall the following general construction. Let $G$ be a group. 
2659: Let $X$ be a $G$-set and $f$ a function on $X^n$ satisfying 
2660: $$
2661: \sum_{i=1}^{n+1}(-1)^{i}f(x_1, ..., \widehat x_i, ..., x_{n+1}) =0
2662: $$
2663:  Choose a point $x \in X$. Then there is an $(n-1)$-cocycle of the group $G$:
2664: $$
2665: f_x(g_1, ..., g_n):= f(g_1 x, ..., g_n x)
2666: $$  
2667: 
2668: \begin{lemma} \label{point}
2669: The cohomology class of the cocycle $f_x$ does not depend on $x$.
2670: \end{lemma}
2671: 
2672: {\bf Proof}. The difference $f_y - f_x$ is the coboundary of the $(n-2)$-chain 
2673: \begin{equation} \label{1.22.1}
2674: h_{x,y}(g_1, ..., g_{n-1}) = 
2675: \sum_{i=1}^{n-1}(-1)^{k-1} f(g_1 x, g_2 x, ..., g_kx, g_ky, g_{k+1}y, 
2676: ..., g_{n-1} y)
2677: \end{equation}
2678: Here is the geometric picture leading to this formula. Consider the prism 
2679: $\Delta_{g_1, ..., g_{n}}^{(n-1)}\times \Delta_{x,y}^{(1)}$ given by product 
2680: of the $(n-1)$-simplex with vertices 
2681: $g_1, ..., g_n$ by the  $1$-simplex with vertices $(x,y)$. Decomposing its side face  
2682: $\Delta_{g_1, ...,  g_{n-1}}^{(n-2)}\times \Delta_{x,y}^{(1)}$ 
2683: into simplices 
2684:  we come to the right hand side of (\ref{1.22.1}).  Then the terms of the 
2685: formula $f_y - f_x - \delta h_{x,y}$ 
2686: correspond to the boundary faces of the prism. Cutting the prism into
2687:  simplices we see that 
2688: the sum of the terms corresponding to the prism boundary is zero thanks 
2689: to the cocycle relation. 
2690: The lemma is proved. 
2691: 
2692: So, for any $x \in {\Bbb H}_n$, 
2693: $$
2694: (\psi_n)_x(g_0,...,g_{2n-1}):= \psi_n(g_0x,...,g_{2n-1}x) 
2695: $$ 
2696: is a smooth $(2n-1)$-cocycle of $GL_n(\C)$. 
2697: 
2698: {\bf Remark}. This   cocycle   is the restriction to $GL_n(\C)$ 
2699: of  the Bott cocycle for the group of diffeomorphisms of $\C \PP^{n-1}$. 
2700: 
2701: 
2702: 
2703: Let $h_0, \dots, h_{2n-1}$ be any hyperplanes in ${\Bbb CP}^{n-1}$. 
2704: Recall that the Grassmannian $n$-logarithm is defined by
2705: $$
2706: {\cal L}_n^{G}(h_0,\dots,h_{2n-1})=(2\pi i)^{1-n}
2707: \int_{{\Bbb CP}^{n-1}} r_{2n-1}(f_1,\dots, f_{2n-1})$$
2708: where $f_i$ is a rational function on ${\Bbb CP}^{n-1}$ with the divisor $(h_i) -(h_0)$. 
2709: 
2710: 
2711: \begin{proposition} \label{0.3}
2712: One has 
2713: $$
2714: {\cal L}_n^{\rm G}(h_0,\dots,h_{2n-1}) = -\frac {(-4)^{n-1}(n-1)!^2} {(2\pi i)^{n-1} (2n-2)!}
2715: \int_{{\Bbb CP}^{n-1}}\log|f_1|\bigwedge_{j=2}^{2n-1} d\log|f_j| 
2716: $$
2717: \end{proposition}
2718: 
2719: 
2720: {\bf Proof}. See Proposition 6.2 in [GZ]. 
2721: 
2722: 
2723:    {\bf 3. The Grassmannian polylogarithm ${\cal L}_n^G$ as the boundary value
2724:  of the function $\psi_n$.}  
2725: We start from an explicit formula for the Fubini-Studi form. Let $\widehat \PP^{n-1}$ 
2726: be the variety of all hyperplanes in $\PP^{n-1}$. Consider the incidence divisor
2727: $$
2728: D \subset  \widehat \PP^{n-1} \times  \PP^{n-1}  \qquad D := \{(h, x)| x \in h\}
2729: $$
2730:  where $h$ is a hyperplane and $x$ is a point in $\PP^{n-1}$. 
2731: 
2732: Let $(x_0:...:x_{n-1})$ be  the 
2733:  homogeneous coordinates of a point $x$ in $\PP^{n-1}$. 
2734: Let
2735: $$
2736: \sigma_n(x,dx):= \sum_{i=0}^{n-1} (-1)^i x_i dx_0 \wedge ... 
2737: \wedge \widehat dx_i \wedge ... \wedge dx_{n-1}  = i_E {\rm vol}_x
2738: $$
2739: be the Leray from. Here ${\rm vol}_x = dx_0 \wedge ... 
2740: \wedge dx_{n-1}$ and $E =\sum x_i \partial_{x_i}$. 
2741: 
2742:  There is a canonical differential $(n-1,n-1)$-form
2743: $\omega_D$ on $ \widehat \PP^{n-1} \times  \PP^{n-1} - D$ 
2744: with a polar singularity at the divisor $D$. Namely, 
2745: let $x \in V_n$ and $\xi \in V_n^*$. Then 
2746: $$
2747: \omega_D:= \frac{1}{(2\pi i)^{n-1}}\frac{\sigma_n(\xi,d\xi) 
2748: \wedge \sigma_n(x,dx)}{<\xi,x>^{n}}
2749: $$
2750: It is $PGL_{n}$-invariant. 
2751: A hermitian metric $H$ in $V_n$ provides an isomorphism 
2752: $H: V_n \lra \overline V_n^*$, and hence an isomorphism 
2753: ${\Bbb C} \PP^{n-1} \lra \widehat {\overline {{\Bbb C}P}}^{n-1}$. 
2754: The graph $\Gamma_H$ of this map does not intersect the incidence divisor $D$. Thus 
2755: restricting the form $\omega_D$ to $\Gamma_H$ we 
2756: get a volume form on ${{\Bbb C}P}^{n-1}$:
2757: \begin{equation} \label{FUBS}
2758: \omega_{FS}(H) := \frac{1}{(2\pi i)^{n-1}}\frac{\sigma_n(z,dz) \wedge \sigma_n(\overline z, d \overline {z})}{H(z, \overline z)^{n}}
2759: \end{equation} 
2760: It is clearly invariant under the group preserving the Hermitian form $H$. Moreover, 
2761: it is the Fubini-Studi volume form:  
2762: a proof can be obtained by using the explicit formula for the Fubini-Studi
2763:  K\"ahler form given  in [Ar], complement 3. 
2764: 
2765: 
2766: 
2767: One can realize 
2768: $\C \PP^{n-1}$ as the smallest stratum  of the boundary of $\HH_n$ . Namely, for 
2769: a hyperplane $h$ in an $n$-dimensional complex vector space $V_n$ let 
2770: $$
2771: F_h:= \left \{\mbox{nonnegative definite hermitian forms in 
2772:   $V_n$ with  kernel}  h \right \}/ \R _+^*
2773: $$
2774: The set of 
2775:  hermitian forms in $V_n$ with the 
2776: kernel  $h$  is isomorphic to $\R _+^*$,  so  $F_h$ defines a point on 
2777: the boundary of $\overline \HH_n$. 
2778: 
2779:  For any nonzero 
2780: nonnegative definite hermitian form $H$ one can define the corresponding 
2781: Fubini-Studi form by formula (\ref{FUBS}). It is 
2782: a differential form with  singularities along the projectivization of 
2783: the kernel of $H$. In particular if $h$ is a hyperplane then the degenerate hermitian 
2784: form $F_{h}$ provides the Lebesgue measure on the affine space $\C\PP^{n-1} -  h$. 
2785: Indeed, if 
2786: $h_0 = \{z_0 =0\}$ then (\ref{FUBS}) specializes to 
2787:     $$
2788: \frac{1}{(2\pi i)^{n-1}}  d\frac{z_1}{z_0} \wedge ... \wedge d\frac{z_{n-1}}{z_0}  \wedge  d\frac{ \overline z_1}{\overline z_0} \wedge ... \wedge d\frac{ \overline z_{n-1}}{\overline  z_0}  
2789: $$
2790: Denote by $M_{h}$ the one dimensional real vector space generated by this form. 
2791: For any hyperplane $h$ in $\C \PP^{n-1}$ 
2792: let us choose a measure $\mu_{h} \in M_{h}$. 
2793: 
2794: 
2795: 
2796: 
2797: 
2798: \begin{proposition} \label{1/18.1}
2799: For any $2n$ hyperplanes $h_0,...,h_{2n-1}$ in $\C \PP^{n-1}$ the integral 
2800: \begin{equation} \label{1221N} 
2801: \psi_n(h_0,...,h_{2n-1}) := \int_{\C \PP^{n-1}}
2802:  \log | \frac{\mu_{h_1}}{\mu_{h_0}}| d\log|\frac{\mu_{h_2}}{\mu_{h_0}} | \wedge ... \wedge d\log| \frac{\mu_{h_{2n-1}}}{\mu_{h_0}} |
2803: \end{equation} 
2804:  is convergent and equals to 
2805: $$
2806: (-4)^{-n} \cdot (2\pi i)^{n-1} (2n)^{2n-1}{2n-2\choose n-1} \cdot {\cal L}_n^G(h_0, ..., h_{2n-1})
2807: $$
2808: \end{proposition} 
2809: 
2810: {\bf Proof}. Let  $h_1,h_2$  be hyperplanes in $\C \PP^{n-1}$ and $f$ be a rational function   
2811: such that   $(f) = (h_1) -(h_2)$. 
2812: From the explicit description of $M_h$ given above we immediately see that 
2813: \begin{equation} \label{AG362}
2814: \mu_{h_1}/\mu_{h_2} =   \lambda \cdot |f|^{2n}
2815: \end{equation}
2816: Using this and Theorem 2.4 we see that  
2817: integral (\ref{1221N}) is convergent. The second statement follows from 
2818: Proposition \ref{0.3} and (\ref{AG362}). The proposition is proved. 
2819: 
2820: 
2821: More generally, take any $2n$ Hermitian forms 
2822: $H_0, ..., H_{2n-1}$, possibly degenerate. For each of the forms $H_i$ 
2823:  consider the corresponding measure $\mu_{H_i}$ ( a multiple of the Fubini-Studi form related to $H_i$). Using the convergence of the integral (\ref{1221N}) 
2824: we can deduce that the integral 
2825: \begin{equation} \label{1221M} 
2826: \psi_n(H_0,...,H_{2n-1}) := \int_{\C \PP^{n-1}}
2827:  \log | \frac{\mu_{H_1}}{\mu_{H_0}}| 
2828: d\log|\frac{\mu_{H_2}}{\mu_{H_0}} | \wedge ... \wedge 
2829: d\log| \frac{\mu_{H_{2n-1}}}{\mu_{H_0}} | = 
2830: \end{equation}
2831: $$
2832: -n^{2n-1}\cdot \int_{\C \PP^{n-1}}
2833:  \log | \frac{H_1(z, \overline z)}{H_0(z, \overline z)}| d\log|\frac{H_2(z, \overline z)}{H_0(z, \overline z)} | \wedge ... \wedge d\log| \frac{H_{2n-1}(z, \overline z)}{H_{0}(z, \overline z)} |
2834: $$
2835: is also convergent. This enables us to extend  $\psi_n$ to the function 
2836: $\overline \psi_n(x_0,...,x_{2n-1})$ 
2837: on the configuration space of $2n$ points in $\overline \HH_{n-1}$. The function $\overline \psi_n$ is discontinuous. For instance it is discontinuous at the point 
2838: $x_1 = ... = x_{2n-1} = F_h$ for a given hyperplane $h$ in $\C \PP^{n-1}$. 
2839: It is however a smooth function on an open part of any given strata. We will keep the notation
2840: $$
2841: \psi_n(h_0,..., h_{2n-1}) = \overline \psi_n(F_{h_0},...,F_{h_{2n-1}})
2842: $$
2843: 
2844: Applying Lemma \ref{point} to the case when $X$ is $\overline {\Bbb H}_n$ and using only the fact that 
2845: the function $\overline \psi_n(x_0,...,x_{2n-1})$ is well defined for {\it any}
2846:  $2n$ points in $\overline {\Bbb H}_n$ and satisfies the cocycle condition for any $2n+1$ of them we get 
2847: 
2848: \begin{corollary} \label{pointx1}
2849: Let $x \in {\Bbb H}_n$ and let $h$ be a hyperplane in $\C\PP^{n-1}$. 
2850: Then the cohomology classes of the following cocycles coincide:
2851: $$
2852: \psi_n(g_0 x,...,g_{2n-1}x) \quad \mbox{and} \quad \psi_n(g_0 h,...,g_{2n-1}h)
2853: $$ 
2854: \end{corollary}
2855: 
2856: {\bf 4. A normalization of the Borel class $b_n$}. Choose a hermitian metric 
2857: in $V_n$. 
2858: Let $e$ be the corresponding point of the symmetric space ${\Bbb H}_n$; its 
2859: stabilizer is the subgroup $SU(n)$. One has 
2860: $$
2861: \Bigl(\Lambda^{\bullet}T^*_e{\Bbb H}_n\Bigr)^{SU(n)} = 
2862: {\cal A}^{\bullet}(SL_n(\C)/SU(n))^{SL_n(\C)}  
2863: $$
2864: There are well known  
2865: canonical ring isomorphisms (see [B2] and references there):
2866: $$
2867: \Bigl(\Lambda^{\bullet}T^*_e{\Bbb H}_n\Bigr)^{SU(n)}\otimes_{\R} \C = 
2868: \Lambda^{\bullet}(sl_n(\C))^{sl_n(\C)} = 
2869: $$
2870: \begin{equation} \label{1.20.2}
2871: H^{\bullet} (sl_n(\C), \C) \stackrel{\alpha}{=} H_{\rm top}^{\bullet} (SU(n), \C) 
2872: \stackrel{\beta}{=}
2873: H^{\bullet}_m(SL_n(\C), \C)
2874: \end{equation}
2875: where $H^{\bullet}$ is the Lie algebra cohomology, $H^{\bullet}_{\rm top}$ is the topological cohomology, and $H_m(G)$ denotes the measurable cohomology of a Lie group $G$. 
2876: The first isomorphism  is obvious: $T_e{\Bbb H}_n\otimes_{\R} \C = sl_n(\C)$. 
2877: The map 
2878: $$
2879: \alpha_{\rm DR}: \Lambda_{\Q}^{\bullet}(sl_n)^{sl_n} \stackrel{\sim}{\lra} 
2880: H_{\rm DR}^{\bullet} (SL_n(\C), \Q) 
2881: $$ 
2882: sends an $sl_n$--invariant exterior form 
2883: on $sl_n$  to the right--invariant one, and hence 
2884: biinvariant differential form on $SL_n(\C)$. 
2885: Let us describe the map 
2886: $$
2887: \beta_{\rm DR}: H_{\rm DR}^{\bullet} (SL_n(\C), \Q) 
2888: \stackrel{}{\lra}
2889: H^{\bullet}_m(SL_n(\C), \C)
2890: $$ 
2891: 
2892: Let $C$ be a biinvariant, and hence closed, 
2893:  differential $(2n-1)$--form on $SL_n(\C)$. Let us restrict it first 
2894: to the Lie algebra, and then  
2895: to the orthogonal 
2896: complement $su(n)^{\perp}$ to the Lie subalgebra $su(n) \subset  sl_n(\C)$. 
2897: We identify the $\R$--vector spaces $T_e{\Bbb H}_n$ and  $su(n)^{\perp}$. 
2898: The obtained exterior form on $T_e{\Bbb H}_n$ is the restriction of 
2899:  an invariant differential form, denoted $\omega_{C}$,  
2900: on the symmetric space 
2901: ${\Bbb H}_n$. It is a closed differential form. 
2902: 
2903: 
2904: 
2905: For any ordered $2n$ points $x_1,...,x_{2n}$ 
2906: in ${\Bbb H}_n$ there is 
2907: a  geodesic 
2908: simplex $I(x_1,...,x_{2n})$ in ${\Bbb H}_n$. It is constructed inductively 
2909: as follows. Let 
2910: $I(x_1,x_2)$ be the geodesic from $x_1$ to $x_2$. The  
2911: geodesics from $x_3$ to 
2912: the points of  $I(x_1,x_2)$ form a geodesic triangle 
2913: $I(x_1,x_2,x_3)$. All the geodesics from $x_4$ to the points 
2914: of the geodesic triangle $I(x_1,x_2,x_3)$ form a geodesic simplex 
2915: $I(x_1,x_2,x_3,x_4)$, and so on. When the 
2916: rank of the symmetric space is greater than 1 (i.e. $n>2$) 
2917: the geodesic simplex $I(x_1,...,x_k)$ depends on the 
2918: ordering of the vertices $x_1,...,x_k$. 
2919: 
2920: The 
2921:  differential $(2n-1)$-form $\omega_{C}$ on 
2922: $SL_n(\C)/SU(n)$ provides a volume of the geodesic simplex: 
2923: $$
2924: {\rm vol}_C I(x_1,...,x_{2n}):= \int_{I(x_1,...,x_{2n})}\omega_{C}
2925: $$
2926: For every $2n+1$ points $x_1,...,x_{2n+1}$ the boundary of the simplex 
2927: $I(x_1,...,x_{2n+1})$ is the alternating sum of the simplices 
2928: $I(x_1,..., \widehat x_i,..., x_{2n+1})$. Since the form $\omega_{C}$ is 
2929: closed, the Stokes theorem yields 
2930: \begin{equation} \label{6.15.02.1}
2931: \sum_{i=1}^{2n+1}(-1)^i \int_{I(x_1,..., \widehat x_i,..., x_{2n+1})}\omega_{C} = 
2932: \int_{I(x_1,..., x_{2n+1})}d\omega_{C} = 0
2933: \end{equation}
2934: This just means that for a given point $x$ the function
2935: $ {\rm vol_C} I(g_1 x,...,g_{2n} x)$ is a smooth $(2n-1)$-cocycle 
2936: of the Lie group 
2937: $SL_n(\C)$. It was considered by J.Dupont [D]. 
2938: By Lemma \ref{point} cocycles corresponding to different points $x$ 
2939: are canonically cohomologous. 
2940: The obtained cohomology class is the class  $\beta_{\rm DR}([C])$. 
2941: 
2942: 
2943: 
2944: {\bf Remark}. 
2945: ${\rm vol}_CI(x_1,...,x_{2n})$ is independent up to a sign 
2946: of the ordering of its vertices. 
2947: Indeed, 
2948: consider $2n+1$ points $(x_1, x_2, x_1, x_3,...,x_{2n})$ 
2949: and apply relation (\ref{6.15.02.1}).
2950: 
2951: 
2952: {\it The Betti cohomology of $SL_n(\C)$}.  Recall  that $SU(n)$ is a retract of 
2953: $SL_n(\C)$. 
2954: It is well known that 
2955: $$
2956: H_{\rm top}^{\bullet} (SU(n), \Z) = 
2957: H_{\rm top}^{\bullet} (S^3 \times S^5 \times ... \times S^{2n-1}, \Z) = 
2958: \Lambda^*(B_3, 
2959: B_5, ..., B_{2n-1})
2960: $$
2961: The restriction from $SU(n)$ 
2962: to $SU(m)$ kills the classes $B_{2k-1}$ for $k>m$. If $k\leq m$ 
2963: it identifies the class 
2964: $B_{2k-1}$ for $SU(n)$ with the one for $SU(m)$. 
2965: The class $B_{2n-1}$ for $SU(n)$ 
2966: is provided by the fundamental class of the sphere $S^{2n-1} \subset \C^n$. 
2967: Namely, it is the pull back of the 
2968: fundamental class under the map $SU(n) \lra S^{2n-1}$ provided by a choice of a point on 
2969: $S^{2n-1}$.  This sphere has the orientation 
2970: induced by the one of $\C^n$. 
2971: Thus 
2972: \begin{equation} \label{1.20.1}
2973: \Z \cdot B_{n} = {\rm Ker}\Bigl(H_{\rm top}^{2n-1}(SU(n), \Z) \lra 
2974: H_{\rm top}^{2n-1}(SU(n-1), \Z)  \Bigr)
2975: \end{equation} 
2976: The transgression in the Leray spectral sequence for the universal $SU(n)$-bundle 
2977: $EU(n) \lra BU(n)$ provides an isomorphism 
2978: $$
2979: \Z\cdot B_n \lra \frac{H^{2n}(BSU(n), \Z)}{\oplus_{0 < i < 2n}H^i \cdot H^{2n-i}}
2980: $$
2981: and identifies $B_{n}$ with the Chern class 
2982: $c_n \in H^{2n}(BSU(n), \Z)$ of the associated vector bundle. 
2983: 
2984: {\it The De Rham cohomology of $SL_n(\C)$}. Consider the differential form
2985: \begin{equation} \label{7.1.02.2}
2986: \omega_{D_n}:= {\rm tr}(g^{-1}dg)^{2n-1} \in \Omega^{2n-1}(SL_N)
2987: \end{equation} 
2988: Its restriction to the subgroup $SL_m$ is zero for $m<n$. It follows that 
2989: the cohomology class 
2990: $$
2991: [\omega_{D_n}] \in H^{2n-1}_{\rm DR}(SL_n, \C)
2992: $$
2993: is a multiple of $B_n$. 
2994: The Hodge considerations 
2995: show that $[\omega_{D_n}] \in  (2\pi i)^n \Q \cdot 
2996: B_n$. 
2997: 
2998: \begin{lemma} \label{7.1.02.3} The differential form $\omega_{D_n}$ is an 
2999: $\R(n-1)$-valued form. In particular it provides 
3000: a cohomology  class 
3001: $$
3002:  b_n := \beta_{\rm DR}(\omega_{D_n}) \in H^{2n-1}_m(SL_n(\C), \R(n-1)) 
3003: $$
3004: \end{lemma}
3005: 
3006: {\bf Proof}. An easy calculation
3007: shows that the value of the exterior form ${\omega_{D_n}}_{|T_e{\Bbb H}_n}$ 
3008: on 
3009: $$
3010: (e_{1,n} + e_{n,1}) \wedge  i(e_{1,n} - e_{n,1}) \wedge ... \wedge  
3011: (e_{n-1, n} + e_{n, n-1}) \wedge  i(e_{n-1, n} - e_{n, n-1}) \wedge e_{n,n} 
3012: $$
3013: is non zero, and obviously lies in $\Q(n-1)$. 
3014: 
3015: On the other hand the values of the form $\omega_{D_n}$ 
3016: lie in a one dimensional $\R$--vector space. Indeed, the space of 
3017: $su(n)$--invariant real exterior 
3018: $(2n-1)$--forms on the space of all 
3019: hermitian $n \times n$ matrices,
3020:  which have  zero restriction to the subspace of hermitian 
3021: $(n-1) \times (n-1)$ matrices,  
3022: is one-dimensional.  The exterior form ${\omega_{D_n}}_{|T_e{\Bbb H}_n}$ 
3023: belongs to the complexification of this space. The lemma follows from this. 
3024: 
3025: 
3026: 
3027: We call the cohomology class provided by this lemma the Borel class, and 
3028: use it below to construct the Borel regulator. 
3029: 
3030: 
3031:  
3032: 
3033: {\bf 5. Comparison of the Grassmannian and Borel cohomology classes of $GL_n(\C)$}. 
3034: Let $[C_n^G]$ be the cohomology class of the $(2n-1)$--cocycle 
3035: of $GL_n(\C)$ provided by the Grassmannian $n$--logarithm (see Corollary 5.5). 
3036: We want to compare it with the Borel class. 
3037: 
3038: Let us consider the following integral 
3039: $$
3040: \widetilde C_n(H_1, ..., H_{2n-1}):=
3041: $$
3042: \begin{equation} \label{A362} 
3043: -n^{2n-1}\cdot \int_{\C \PP^{n-1}}
3044:  \frac{H_1(z, \overline z)}{(z, \overline z)} d\frac{H_2(z, \overline z)}{(z, \overline z)}  \wedge ... \wedge d\frac{H_{2n-1}(z, \overline z)}{(z, \overline z)} 
3045: \end{equation}
3046: where  the $H_i$ are arbitrary complex matrices and  $H_i(z, \overline z)$ 
3047: are the bilinear form in $z, \overline z$ given by  the matrix $H_i$. 
3048: We claim that it is a $(2n-1)$-cocycle of the Lie algebra $gl_n(\C)$, and it is 
3049: obtained by differentiating  the 
3050: group cocycle provided by the function (\ref{1221M}). 
3051: We put these  facts in the following framework. 
3052: 
3053: If we restrict to the case when $H_i$ are hermitian matrices, 
3054: integral (\ref{A362}) admits the following interpretation. 
3055: Let us construct a  map
3056: $$
3057: {\Bbb M}_e: \C \PP^{n-1} \lra T^*_e{\Bbb H}_n
3058: $$
3059: which is a version of the moment map. 
3060: For a point $z \in \C \PP^{n-1}$ the value of the $\langle M_e(z), v\rangle$ 
3061: of the functional ${\Bbb M}_e(z)$ on a vector $v \in T_e{\Bbb H}_n$ is defined as follows. 
3062: Let $e(t)$ be a path in 
3063: ${\Bbb H}_n$ such that $e(0) =e$ and $\stackrel{{\bf .}}{e}(0) =v$. 
3064: recall the measure $\mu_x$ defined in Section 5.1. Then 
3065: $$
3066: <{\Bbb M}_e(z), v>:= \frac{d}{dt}\log \frac{\mu_{e(t)}(z)}{\mu_{e}(z)}|_{t=0} 
3067: $$ 
3068: Choose coordinates $z_1, ..., z_{n}$ in $V_n$ such that 
3069: $(z, \overline z):= |z_1|^2 + ... + |z_{n}|^2$ corresponds to the point $e$. 
3070:  Then $T_e{\Bbb H}_n$ is identified with the space of hermitian $(n\times n)$ 
3071: matrices $H$. It follows from 
3072: (\ref{FUBS}) that 
3073: \begin{equation} \label{1/18.2} 
3074: <{\Bbb M}_e(z), H>:= n \frac{H(z,\overline z)}{(z,\overline z)}
3075: \end{equation}
3076: The map ${\Bbb M}_e$ is clearly $SU(n)$-invariant. Its image is an  $SU(n)$-orbit 
3077: in $T^*_e{\Bbb H}_n$ isomorphic to $\C \PP^{n-1}$. 
3078: 
3079: We need the following general construction. Let $V$ be a real vector space and $M$  a compact subset of  $V^*$ which is the closure of a $k$-dimensional submanifold. Any element $\omega \in \Lambda^kV$ can be viewed as  a  $k$-form 
3080: $\omega$ on $V^*$. Integrating it over $M$ we get 
3081: an exterior form $C_M \in \Lambda^kV^*$. If $M$ is a cone over $M'$ with the vertex at the origin then $\int_M\omega = \int_{M'}i_E\omega$ where $E$ is the Euler vector field on $V$. 
3082: 
3083: 
3084:  Applying this construction to the cone over the orbit $M_e$ 
3085: based at  the origin we get an  
3086:  $SU(n)$-invariant element 
3087: $\widetilde C_n \in \Lambda^{2n-1}T^*_e{\Bbb H}_n$. 
3088: It follows from (\ref{1/18.2}) that 
3089: it is given by formula (\ref{A362}) multiplied by $2n$.
3090: 
3091: 
3092: Another invariant $(2n-1)$-cocycle 
3093:  $C_n$ of the Lie algebra $gl_n$, considered  by Dynkin [Dy],  is given by
3094: \begin{equation} \label{7.1.02.1}
3095: C_n(X_1, ..., X_{2n-1}) = \frac{1}{n! }{\rm Alt}_{2n-1}{\rm Tr}(X_1 X_2 ... X_{2n-1})
3096: \end{equation}
3097: Let $[C_n]$ be the cohomology class 
3098: of $GL_n(\C)$ 
3099: corresponding to the cocycle $C_n$.  
3100: 
3101: \begin{theorem} \label{1.20.10} One has 
3102: $$
3103: \widetilde C_n \quad = -(-1)^{\frac{(n-1)n}{2}}\frac{(2\pi i)^{n-1}n^{2n-1}(n-1)!}{(2n-1)!} 
3104: \cdot  C_n
3105: $$
3106: and the class $[C_n^G]$ is a non zero rational multiple of  $[C_n]$:
3107: $$
3108: [C_n^G] = - (-1)^{\frac{n(n+1)}{2}}\frac{(n-1)!^3}{(2n-2)!(2n-1)!} [C_n]
3109: $$
3110: \end{theorem}
3111: 
3112: {\bf Proof}. The second claim follows from the first 
3113: using Proposition \ref{1/18.1}. 
3114: 
3115: 
3116: Let us prove the first claim. 
3117: The restriction of the cocycle $C_n$ to the Lie subalgebra $gl_{n-1}(\C)$ equals to zero. 
3118: This follows, for instance,  from the Amitsur-Levitsky theorem: for any $n \times n$ 
3119: matrices $A_1, ..., A_{2n}$ one has 
3120: ${\rm Alt}_{2n} (A_1, ..., A_{2n})=0$. 
3121: 
3122: 
3123: On the other hand the restriction of the cocycle $\widetilde C_n$ 
3124:  to the Lie subalgebra of matrices 
3125: $(a_{ij})$ where $a_{1j} = a_{j1} =0$ 
3126: is zero. 
3127: Indeed, in this case the form we integrate 
3128: in (\ref{A362}) is a differential $(2n-2)$-form in $dz_2, ..., dz_{n-1}, 
3129: d \overline z_2, ..., d\overline z_{n-1}$ and thus it is zero.  
3130: So thanks to (\ref{1.20.2}) and (\ref{1.20.1}) we conclude that  the 
3131: cocycle $C_n $ is proportional to $\widetilde C_n$. 
3132: To determine the proportionality coefficient 
3133: we compute the values of the both cocycle on a special element 
3134: $E_n \in \Lambda^{2n-1}gl_n$. To write it down denote by $e_{i,j}$ the 
3135: elementary $n\times n$ matrix whose only non-zero entry is $1$ on the $(i,j)$ place. Then 
3136: \begin{equation} \label{1/16.2}
3137: E_n:= \bigwedge_{j=1}^{n-1}(e_{j,n}\wedge  e_{n,j} )\wedge e_{n,n} 
3138: \end{equation}
3139:  A direct computation shows that  
3140: $$<C_n, E_n> = 1
3141: $$
3142: Indeed, to get a non zero trace we have to multiply $(n-1)$ blocks 
3143: $e_{n,j} e_{j,n}$, as well as $e_{n,n}$, which can be inserted 
3144: anywhere between these blocks. 
3145: So there are $(n-1)! n = n!$ possibilities. 
3146: 
3147: Let us compute the value of the cocycle $\widetilde C_n$ on $E_n$.  
3148: \begin{lemma}
3149: Integral (\ref{A362}) equals 
3150: \begin{equation} \label{1/18.4} 
3151: \frac{-n^{2n-1}}{(2n-1)!}\cdot {\rm Alt}_{2n-1}\int_{\C \PP^{n-1}}
3152:  \frac{H_1(z, \overline z)d H_2(z, \overline z)  \wedge ... \wedge dH_{2n-1}(z, \overline z)}{(z, \overline z)^{2n-1}} 
3153: \end{equation}
3154: \end{lemma}
3155: 
3156: {\bf Proof}. By Proposition \ref{1.2.} integral (\ref{A362}) equals to 
3157: \begin{equation} \label{1?1} 
3158: \frac{-n^{2n-1}}{(2n-1)!}{\rm Alt}_{2n-1}\cdot \int_{\C \PP^{n-1}}
3159:  \frac{H_1(z, \overline z)}{(z, \overline z)} d\frac{H_2(z, \overline z)}{(z, \overline z)}  \wedge ... \wedge d\frac{H_{2n-1}(z, \overline z)}{(z, \overline z)} 
3160: \end{equation}
3161: One has, for $i=2, ..., 2n-1$, that  
3162: $$
3163: d\frac{H_{i}(z, \overline z)}{(z, \overline z)}  = 
3164: \frac{(z, \overline z) d H_{i}(z, \overline z) - H_{i}(z, \overline z) d (z, \overline z)}{(z, \overline z)^2}
3165: $$
3166: Substituting  
3167: $$
3168: \frac{- H_{i}(z, \overline z) d (z, \overline z)}{(z, \overline z)^2}\quad \mbox{instead of} \quad d\frac{H_i(z, \overline z)}{(z, \overline z)} \quad \mbox{in (\ref{1?1})} 
3169: $$
3170: we get zero since $H_1$ and $H_i$  
3171: appear in a symmetric way and thus disappear after the alternation. 
3172: The lemma follows. 
3173: 
3174: 
3175: Let us calculate integral (\ref{1?1}) in the special case 
3176: $$
3177: H_{2n-1}(z, \overline z) = |z_n|^2, \quad H_{2k-1}(z, \overline z) = z_k \overline z_n, \quad H_{2k}(z, \overline z) = z_n \overline z_k 
3178: $$  
3179: so that   $H_1 \wedge ... \wedge H_{2n-1} =E_n$. 
3180: We will  restrict the integrand to the affine part $\{z_n=1\}$ 
3181: and then perform the integration.  
3182: Since $dH_{2n-1}(z, \overline z) =0$ on $\{z_n=1\}$ and 
3183: $d z_k \wedge d\overline z_k  = -2i dx_k \wedge dy_k$ 
3184: we get
3185: $$
3186: -(-1)^{\frac{(n-1)(n-2)}{2}}\frac{(-2i)^{n-1}n^{2n-1}}{2n-1}\int_{\C ^{n-1}}\frac{
3187:  d^{n-1}x d^{n-1} y}
3188: {(1+ |z_1|^2 + ... + |z_{n-1}|^2)^{2n-1}} =
3189: $$
3190: $$
3191: -(-1)^{\frac{(n-1)n}{2}}\frac{(2i)^{n-1}n^{2n-1}}{2n-1} {\rm vol}(S^{2n-3})
3192: \int_0^{\infty}\frac{r^{2n-3}dr}{(1+r^2)^{2n-1}} =
3193: $$
3194: $$
3195:  -(-1)^{\frac{(n-1)n}{2}}\frac{(2\pi i)^{n-1}n^{2n-1}}{(n-1)!(2n-1)} 
3196: {2n-2 \choose n-1}^{-1}
3197: $$
3198: since the volume of the sphere $S^{2n-3}$ is 
3199: $\frac{2\pi^{n-1}}{(n-2)!}$ and 
3200: \begin{equation} \label{1/18.3}
3201: \int_0^{\infty}\frac{r^{2n-3}dr}{(1+r^2)^{2n-1}} \quad = \quad \frac{1}{2}\int_0^{\infty}\frac{r^{n-2}dr}{(1+r)^{2n-1}} \quad = \quad \frac{1}{2n-2} {2n-2 \choose n-1}^{-1}
3202: \end{equation}
3203: To get the last equality we integrate by parts: 
3204: $$
3205: \int_0^{\infty}\frac{r^{n-2}dr}{(1+r)^{2n-1}} = - \frac{1}{2n-2}\int_0^{\infty}r^{n-2} \left(
3206: \frac{1}{(1+r)^{2n-2}}\right)'dr = 
3207: $$
3208: $$
3209: \frac{n-2}{2n-2}\int_0^{\infty}r^{n-3}\frac{dr}{(1+r)^{2n-2}} = ... = 
3210: \frac{(n-2)!n!}{(2n-2)!}\int_0^{\infty}\frac{dr}{(1+r)^{n+1}} = 
3211: \frac{1}{n-1} {2n-2 \choose n-1}^{-1}
3212: $$
3213: Theorem \ref{1.20.10} is proved. 
3214: 
3215: 
3216: 
3217: 
3218: {\bf 6. Construction of the Borel regulator 
3219: via Grassmannian polylogarithms}. 
3220: Let $G$ be a group. The diagonal map $\Delta: G \lra G \times G$ provides a homomorphism $\Delta_*: H_n(G) \lra H_n(G \times G)$. 
3221: %By the Kunneth formula $H_n(G \times G)_{\Q} = 
3222: %\oplus_{ 0 \leq k \leq n}H_k(G)_{\Q} \otimes H_{n-k}(G)_{\Q}$. 
3223: Recall that 
3224: $$
3225: {\rm Prim}H_n(G) := \{x \in H_n(G)| \Delta_*(x) = x \otimes 1 + 1 \otimes x\}
3226: $$
3227: Set $A_{\Q}:= A \otimes \Q$. One has 
3228: $$
3229: K_n(F)_{\Q} = {\rm Prim}H_n(GL(F))_{\Q} = {\rm Prim}H_nGL_n(F)_{\Q}
3230: $$
3231: where the second isomorphism is provided by Suslin's stabilization theorem. 
3232: 
3233: The Borel regulator is a map 
3234: $$
3235: r_{n}^{\rm Bo}: K_{2n-1}(\C)_{\Q} \lra \R(n-1)
3236: $$
3237: provided by pairing the class $b_n \in H^{2n-1}(GL_{2n-1}(\C), \R(n-1))$ with 
3238: the subspace $K_{2n-1}(\C)_{\Q} \subset H_{2n-1}(GL_{2n-1}(\C), {\Q})$. 
3239: 
3240: Recall the Grassmannian complex $C_*(n)$
3241: $$
3242: ... \stackrel{d}{\lra} C_{2n-1}(n) \stackrel{d}{\lra}  C_{2n-2}(n) 
3243: \stackrel{d}{\lra} ... \stackrel{d}{\lra} C_{0}(n)
3244: $$
3245: where $C_k(n)$ is the free abelian group generated by configurations 
3246: of $k+1$ 
3247: vectors $(l_0, ..., l_{k})$ in generic position in an $n$--dimensional 
3248: vector space over a field $F$, and $d$ is given by the standard 
3249: formula 
3250: (see s. 3.1 in [G2]). The group $C_k(n)$ is in degree $k$. 
3251: Since it is a homological resolution of the trivial $GL_n(F)$--module $\Z$  
3252: (see Lemma 3.1 in [G2]), 
3253: there is a canonical homomorphism 
3254: $$
3255: \varphi_{2n-1}^n: H_{2n-1}(GL_n(F)) \lra H_{2n-1}(C_*(n)) 
3256: $$
3257: Thanks to Lemma \ref{7.1.02.3} 
3258: the Grassmannian $n$--logarithm function provides a homomorphism 
3259: \begin{equation} \label{6.24.02.3}
3260: {\cal L}_n^G: C_{2n-1}(n) \lra \R(n-1); \quad (l_0, ..., l_{2n-1}) \lms 
3261: {\cal L}^G_n(l_0, ..., l_{2n-1})
3262: \end{equation}
3263: Thanks to the first $(2n+1)$--term functional equation for ${\cal L}_n^G$, 
3264: see (\ref{4}),  
3265: it is zero on the subgroup $dC_{2n}(n)$. So it induces a homomorphism
3266: $$
3267: {\cal L}_n^G: H_{2n-1}(C_{*}(n)) \lra \R(n-1);
3268: $$ 
3269: \begin{lemma} \label{6.24.02.1} The composition ${\cal L}_n^G\circ 
3270: \varphi_{2n-1}^n$ coincides with the class $[C_n^G]$.
3271: \end{lemma}
3272: 
3273: {\bf Proof}. Standard, see [G4]. 
3274: 
3275: 
3276: To construct the Borel regulator we extend, as in s. 3.10 of [G2], 
3277: the class $[C_n^G]$ 
3278: to a class of $GL_{2n-1}(\C)$. Let us recall the key steps. 
3279: 
3280: Let $\Z[S]$be the free abelian group generated 
3281: by a set $S$. Let $F$ be a field. 
3282: Applying the covariant functor $\Z \lms \Z[X(F)]$ 
3283: to the bi-Grassmannian 
3284: $\widehat G(n)$ (see s. 4.2), and taking the alternating sum of 
3285: the obtained homomorphisms, 
3286:  we get a bicomplex. 
3287: Using Lemma \ref{6.12.02.5} we see that 
3288: it looks as follows ([G2], s. 3.7):
3289: $$
3290: \begin{array}{cccccccc}
3291: &&&&&... & \stackrel{d}{\lra}&C_{2n-1}(2n-1)\\
3292: &&& &&&&\downarrow \\
3293: &&&...&&...&&...\\
3294: &&& \downarrow &&&&\downarrow \\
3295: & ... & \stackrel{d}{\lra} & C_{2n-1}(n+1) & \stackrel{d}{\lra} & ... & 
3296: \stackrel{d}{\lra} & C_{n+1}(n)\\
3297: &\downarrow && \downarrow &&&&\downarrow \\
3298: ... \stackrel{d}{\lra} & C_{2n-1}(n) & \stackrel{d}{\lra} & C_{2n-2}(n) & 
3299: \stackrel{d}{\lra}& ... &\stackrel{d}{\lra} & C_n(n) 
3300: \end{array}
3301: $$
3302: In particular the bottom row is the stupid truncation of the 
3303: Grassmannian complex at
3304: the group $C_n(n)$. 
3305: The total complex of this bicomplex is a homological complex, 
3306:  called the weight $n$ bi--Grassmannian complex $BC_*(n)$.  
3307: In particular there is a homomorphism 
3308: \begin{equation} \label{6.24.02.2}
3309: H_{2n-1}(C_*(n)) \lra H_{2n-1}(BC_*(n)) 
3310: \end{equation} 
3311: 
3312: In [G1-2] we proved that there are homomorphisms 
3313: $$
3314: \varphi_{2n-1}^m:  H_{2n-1}(GL_m(F)) \lra H_{2n-1}(BC_*(n)), 
3315: \quad m \geq n
3316: $$ 
3317: whose restriction to the subgroup $GL_n(F)$ coincides 
3318: with the composition 
3319: $$
3320: H_{2n-1}(GL_n(F)) \stackrel{{\varphi_{2n-1}^n}}{\lra} H_{2n-1}(C_*(n)) 
3321: \stackrel{(\ref{6.24.02.2})}{\lra} H_{2n-1}(BC_*(n)), 
3322: $$ 
3323: Let us  extend homomorphism 
3324: (\ref{6.24.02.3}) to a homomorphism
3325: $$
3326: {\cal L}_n^G: BC_{2n-1}(n) \lra \R(n-1)
3327: $$
3328: by setting it zero on the groups $C_{2n-1}(n+i)$ for $i>0$. 
3329: The second $(2n-1)$--term functional equation for 
3330: the Grassmannian $n$--logarithm function, see (\ref{5}),  
3331: just means that the composition 
3332: $$
3333: C_{2n}(n+1) \lra C_{2n-1}(n) \stackrel{{\cal L}_n^G}{\lra} \R(n-1),
3334: $$
3335: where the first map is a vertical arrow in $BC_*(n)$, is zero. 
3336: Therefore we get a homomorphism
3337: $$
3338: {\cal L}_n^G: H_{2n-1}(BC_{*}(n)) \lra \R(n-1)
3339: $$ 
3340:  \begin{corollary} \label{1.20.11as}
3341: One has 
3342: $$
3343: [C^G_n] = -(-1)^{n(n+1)/2}  \frac{(n-1)!^2}{(2n-2)!(2n-1)!} 
3344: \cdot \frac{b_n}{n}
3345: $$
3346: \end{corollary}
3347: 
3348: {\bf Proof}. Indeed, 
3349: $$
3350: {\rm Alt}_{2n-1}{\rm Tr}(X_1 \cdot ... \cdot X_{2n-1}) = 
3351: <{\rm tr}(g^{-1}dg)^{2n-1}_{|sl_n}, X_1 \wedge ... \wedge X_{2n-1}>
3352: $$
3353: So the claim  follows from Theorem \ref{1.20.10} since, as it clear from  comparison of formulas 
3354: (\ref{7.1.02.1}) and (\ref{7.1.02.2}), 
3355: $
3356: b_n = n![C_n]
3357: $. 
3358: The corollary is proved. 
3359: 
3360: 
3361: \begin{theorem}  \label{6.24.02.7}
3362: The composition 
3363: \begin{equation} \label{6.24.02.6}
3364:  K_{2n-1}(\C) \stackrel{\sim}{\lra} 
3365: {\rm Prim} H_{2n-1}(GL_{2n-1}(\C), \Q) \stackrel{\varphi_{2n-1}^{2n-1}}{\lra} 
3366: \end{equation} 
3367: $$
3368: H_{2n-1}(BC_{*}(n)_{\Q}) \stackrel{{\cal L}_n^G}{\lra} \R(n-1)
3369: $$
3370: equals to 
3371: \begin{equation} \label{7.02.1.4}
3372: -(-1)^{n(n+1)/2} \cdot \frac{(n-1)!^2}{n(2n-2)!(2n-1)!}r_{n}^{\rm Bo}
3373: \end{equation}
3374: \end{theorem} 
3375: 
3376: {\bf Proof}. Recall that 
3377: restriction to  $GL_n$ of the map $\varphi_{2n-1}^{2n-1}$ 
3378: coincides with the map $\varphi_{2n-1}^n$. Therefore  Lemma \ref{6.24.02.1} 
3379: guarantees that restriction to $GL_n(\C)$ 
3380: of the composition of the last two arrows 
3381:  coincides with the map given by the class $[C_{2n-1}^G]$. 
3382: So Corollary \ref{1.20.11as} implies the theorem. 
3383: 
3384: 
3385: 
3386: {\bf 7. Comparing $[D_n]$ and $B_n$}. 
3387: The following result is not used below. 
3388:  
3389: \begin{theorem} \label{1.20.11}
3390: $$
3391: [D_n] \quad = (2\pi i)^{n}(2n-1)B_n
3392: $$
3393: \end{theorem} 
3394: 
3395: {\bf Proof}. The transgression identifies the class $B_n$ with 
3396: the Chern class of the universal bundle over $BG$, where $G = GL_n(\C)$. 
3397: We will compute explicitly the transgression of the $n$-th component of the 
3398: Chern character of the 
3399: universal vector bundle $p: E \lra BG$. 
3400: Let ${\cal A}$ be a connection 
3401: on $E$. Then the $n$-th Chern class is 
3402: represented by the $2n$-form 
3403: $$
3404: c_n(A):= \frac{tr F_{\cal A}^n}{{(2\pi i)}^n}  
3405: $$
3406: where $F_A := d{\cal A} + {\cal A} \wedge {\cal A}$ is the curvature form. 
3407: 
3408: Let $q: EG \lra BG$ be the principal fibration associated with $E$. 
3409: Then the form $q^*c_n({\cal A})$ is exact. If 
3410: $d\omega = q^*c_n({\cal A})$ and 
3411: $F$ is a fiber of $q$ then $\omega|_F$ 
3412: is closed, its cohomology class is transgressive, and goes to 
3413: $[c_n({\cal A})]$. To do 
3414: the computation we choose a connection ${\cal A}_0$ on $BG$ which is 
3415: flat in a neighbourhood $U$
3416: of a point $x \in BG$. It provides a trivialization of the bundle $E$ 
3417: over $U$ as well as a trivialization  $\varphi: EG|_U \lra G \times U$. 
3418: 
3419: The bundle 
3420: $q^*E$ has  a canonical trivialization. It provides a connection ${\cal A}_1$ 
3421: on $q^*E$. So there are two connections, $q^*{\cal A}_0$ and 
3422: ${\cal A}_1$ on $q^*E$. One has ${\cal A}_1 = q^*{\cal A}_0 + g^{-1}dg$, where 
3423: $(g,u) = \varphi(x)$. Let 
3424: $$
3425: {\cal A}(t):= t {\cal A}_1  +(1-t) q^*{\cal A}_0 = 
3426: tg^{-1}dg + q^*{\cal A}_0
3427: $$ 
3428: It can be thought of as a connection on the lifting of the 
3429: bundle $q^*E$ 
3430: to $EG \times [0,1]$; here $t \in [0,1]$. The curvature $F(t)$ of this 
3431: connection is 
3432: $$
3433: F(t)=  g^{-1}dg dt + t^2g^{-1}dg\wedge  g^{-1}dg
3434: $$
3435: The push forward of the form $trF(t)^n$ down to $EG$ is a primitive for the form 
3436: $trF_{{\cal A}(1)}^n - trF_{{\cal A}(0)}^n$. It is given (in  $q^{-1}U$)  by
3437: $$
3438: \int_0^1tr F(t)^n = \frac{1}{2n-1}tr (g^{-1}dg )^{2n-1}
3439: $$
3440: Theorem \ref{1.20.11} is proved. 
3441: 
3442: 
3443: {\bf 8. On the 
3444: motivic nature of the Grassmannian $n$--logarithm functions}. 
3445: According to the results of the previous section understanding of the Borel regulator, 
3446: and hence special values of the Dedekind $\zeta$--functions, is reduced to 
3447: study of properties of the Grassmannian $n$--logarithm function 
3448: ${\cal L}_n^{G}$. 
3449: 
3450: Recall that a framed mixed Hodge-Tate structure has a natural $\R$-valued invariant ([BD]), called the Lie period. Thus a variation of Hodge-Tate structures ${\Bbb L}$ 
3451: over a base $X$ 
3452: provides  a period function ${\Bbb L}^{\cal M}$ on $X(\C)$. 
3453: 
3454: 
3455: 
3456: \begin{conjecture} \label{6.24.02.10} There exists a variation ${\Bbb L}^{\cal M}_n$ 
3457: of framed 
3458: mixed Tate motives over $\widehat G_{n-1}^n$ such that:
3459: $$
3460: \sum_{i=0}^{2n}(-1)^i a_i^*{\Bbb L}_n^{\cal M} = 0; \quad 
3461: \sum_{j=0}^{2n}(-1)^j b_j^*{\Bbb L}_n^{\cal M} = 0;
3462: $$
3463: and the  Lie period ${\cal L}^{\cal M}_n$ 
3464: of its Hodge realization satisfies 
3465: \begin{equation} \label{6.24.02.11}
3466: {\cal L}^{\cal M}_n - {\cal L}^{G}_n = \sum_{i=0}^{2n-1}(-1)^i 
3467: a_i^*{F}_n
3468: \end{equation}
3469: where $F_n$ is a function on $\widehat G_{n-2}^n(\C)$. 
3470: 
3471: b) The functional equations satisfied by  ${\cal L}^{\cal M}_n$ 
3472:  essentially determine it: 
3473: the space of all smooth/measurable functions satisfying these functional equations is finite dimensional. 
3474: \end{conjecture}
3475: 
3476: {\bf Remark}. The function $F_n$ is obviously 
3477: not determined by 
3478: (\ref{6.24.02.11}) -- add a function coming from $\widehat G^n_{n-3}(\C)$. 
3479: Nevertheless we expect that there exists  a {\it canonical} explicit 
3480: choice for  $F_n$. Then formula (\ref{6.24.02.11}) can be considered 
3481: as an explicit formula for 
3482: ${\cal L}^{\cal M}_n$ in terms of ${\cal L}^{G}$'s. 
3483: 
3484: Moreover we expect that there exists a canonical homotopy 
3485: between the Grassmannian $n$-logarithm (understood as 
3486: a cocycle in the Deligne cohomology of the bi-Grassmannian) 
3487: and its ``motivic'' 
3488: bi-Grassmannian counterpart. 
3489: Observe that the motivic bi-Grassmannian $n$-logarithm should
3490: have  non-trivial components outside of the bottom line of 
3491: the bi-Grassmannian, while the defined above (or in [G5]) 
3492: Grassmannian $n$--logarithm 
3493:  is concentrated entirely 
3494: at the bottom line. 
3495: 
3496: 
3497: A variation of mixed Tate motives over $\widehat G_{n-1}^n$ 
3498: was constructed in [HaM]. However it is not clear how to relate it to 
3499: the function ${\cal L}^G_n$. 
3500: 
3501: 
3502: 
3503: Conjecture \ref{6.24.02.10} is known for $n=2$ and $n=3$. 
3504: 
3505: The $n=2$ case 
3506: follows from (\ref{6.13.02.101}), the well known 
3507: motivic realization of the dilogarithm, and 
3508: Bloch's theorem characterizing the Bloch-Wigner function by 
3509: Abel's $5$--term equation it satisfies. 
3510: 
3511: The $n=3$ case of conjecture \ref{6.24.02.10} follows from the results of 
3512: [G1-2], [GZ] and the motivic realization of the trilogarithm. 
3513: In particular the part b) is given by  
3514:  Theorem 1.10 in [G1]. 
3515: 
3516: {\bf Examples}. 1. $n=2$. Then  ${\cal L}^{\cal M}_2 = {\cal L}^{G}_2$. 
3517: 
3518: 2. $n= 3$. The motivic Grassmannian trilogarithm function 
3519: has been constructed in [G1-2] in terms of the 
3520: classical trilogarithm function. Namely, one has 
3521: $$
3522: {\cal L}^{\cal M}_3(l_0, ..., l_5) = \frac{1}{90} {\rm Alt}_6
3523: {\cal L}_3\Bigl( \frac{\Delta(l_0, l_1, l_3)\Delta(l_1, l_2, l_4)
3524: \Delta( l_2, l_0, l_5)}{\Delta(l_0, l_1, l_4)
3525: \Delta(l_1, l_2, l_5)\Delta(l_2, l_0, l_3)}
3526: \Bigr) 
3527: $$
3528: 
3529: According to Theorem 1.3 of [GZ]  ${\cal L}^{\cal M}_3$ 
3530: is different from ${\cal L}^{G}_3$, and  
3531: $$
3532: F_3(l_0, ..., l_4) = \frac{1}{9}{\rm Alt}_5\Bigl(\log 
3533: |\Delta(l_0, l_1, l_2)| \log |\Delta(l_1, l_2, l_3)| \log |\Delta(l_2, l_3, l_4)| \Bigr) 
3534: $$
3535: 
3536: 3. If $n>3$  then ${\cal L}^{\cal M}_n$ 
3537: is different from ${\cal L}^{G}_n$ since it is already so 
3538: for  the restriction to the special configuration, 
3539: see Theorem \ref{clpoly}. 
3540: 
3541: 
3542: The space of the functional equations 
3543: for the function ${\cal L}^G_3$  is smaller than the one for 
3544: ${\cal L}^{\cal M}_3$, see Chapter 1.5 of [G3].   
3545: A similar situation is expected  
3546: for all $n>3$. 
3547: 
3548: 
3549: The space of the functional equations for the motivic 
3550: $n$--logarithm function ${\cal L}^{\cal M}_n$ should provide an 
3551: explicit construction of the weight $n$ part of the motivic Lie coalgebra of 
3552: an arbitrary field $F$, as explained in s. 4.1 in [G6], taking into 
3553: account the following correction. 
3554: 
3555: {\it Correction}. In s. 4.2 of [G6] the subgroup of the functional equations ${\cal R}_n^G$ is supposed to be defined as the subgroup of all functional equations for the function ${\cal L}^{\cal M}_n$, not ${\cal L}^{G}_n$. 
3556: 
3557: 
3558: 
3559: 
3560:  
3561: \section{The Chow dilogarithm and a reciprocity law}
3562: 
3563: 
3564: The Chow dilogarithm provides a homomorphism $\Lambda^3 \C(X)^* \to \R$ 
3565: given by 
3566: \begin{equation} \label{chowhh}
3567:   f_1\wedge f_2\wedge f_3 \lms {\cal P}_2(X;f_1,f_2,f_3):= \frac{1}{2\pi i}
3568: \int_{X(\C)}r_2(f_1,f_2,f_3)
3569: \end{equation}
3570: 
3571: In this section we show that the Chow dilogarithm can be 
3572: expressed by the function ${\cal L}_2(z)$.  
3573: The precise versions of this claim are discussed below.
3574:  
3575: %The statement  "for any point $(X;f_1,f_2,f_3)$ one has 
3576: %${\cal P}_2(X;f_1,f_2,f_3) = \sum_i {\cal L}_2(z_i)$ for some
3577: % complex numbers $z_i$" 
3578: %is empty since any complex number can be written in this form.  
3579: %However if, for example,  a point 
3580: %$(X;f_1,f_2,f_3) \in {\cal Z}_2^1$ is defined over ${\overline \Q}$, 
3581: %then I claim that one can choose $z_i \in {\overline \Q}$. 
3582: 
3583: {\bf 1. The set up ([G1-2])}. 
3584: For any field $F$ we defined in [G1] the  groups 
3585: $$
3586: {\cal B}_n(F):= \frac{\Z[F^*]}{{\cal R}_n(F)}, \quad n \geq 2
3587: $$
3588: and homomorphisms
3589: $$
3590: \delta_n : {\cal B}_n(F) \lra {\cal B}_{n-1}(F)\otimes F^*; \quad 
3591: \{x\}_n \lms \{x\}_{n-1} \otimes x, \quad n\geq 3
3592: $$
3593: $$
3594: \delta_2: 
3595: {\cal B}_2(F) \lra \Lambda^2F^*, \quad \{x\}_2 \lms (1-x) \wedge x
3596: $$
3597: There is a complex $\Gamma(F;n)$
3598: $$
3599: {\cal B}_{n}(F) \stackrel{\delta_n}{\longrightarrow} {\cal B}_{n- 
3600: 1}(F)\otimes F^{\ast}  \stackrel{\delta_n}{\rightarrow}  \ldots \stackrel{\delta_n}{\rightarrow} {\cal B}_{2}(F)\otimes  \Lambda^{n-2}F^{\ast}
3601: \stackrel{\delta_n}{\rightarrow}
3602:  \Lambda^{n}F^{\ast}
3603: $$
3604: where $\delta_n(\{x\}_k \otimes Y) := \{x\}_{k-1} \otimes x \wedge Y$ for $k>2$, and $(1-x)\wedge x \wedge y$ for $k=2$, 
3605: called the weight $n$ polylogarithmic complex. 
3606: 
3607: If $K$ is a  field with 
3608:  a discrete valuation $v$ and  the residue field $
3609: k_v$, then there is a homomorphism of complexes 
3610: ${\rm res}_v: \Gamma(K, n) \lra \Gamma(k_v, n-1)[-1]$ (see s.\ 14 of \S
3611: 1 in [G1]). For example for $n=3$ we have 
3612: \begin{equation} 
3613: \begin{array}{ccccc}
3614: {\cal B}_3(K)&\stackrel{\delta_3}{\longrightarrow}&{\cal B}_2(K)\otimes
3615: K^{\ast}&
3616: \stackrel{\delta_3}{\longrightarrow}&\Lambda^3K^{\ast}\\
3617:   &&\downarrow {\rm res}_v &&\downarrow  {\rm res}_v \\
3618: &&{\cal B}_2(k_v)&\stackrel{\delta_2}{\longrightarrow}&\Lambda^2k_v^{\ast}
3619: \end{array}
3620: \end{equation}
3621: Here ${\rm res}_v(\{x\}_2 \otimes y)$ is zero unless $v(x) =0$. In the latter case 
3622: it is ${\rm res}_v(\{x\}_2 \otimes y) = v(y) \{\overline x\}_2$, where $\overline x$ 
3623: denotes  projection of $x$ to the residue field of $K$. 
3624: 
3625: 
3626: Let $X$ be a regular  curve over an algebraically closed field
3627: $k$ and $F:= k(X)^{\ast}$.   
3628:  Set ${\rm Res}:= \sum_x {\rm res}_x$ where ${\rm res}_x$ is
3629: the residue homomorphism for the valuation on $F$
3630:  corresponding to a
3631: point $x$ of $X$. For instance for $n=3$ we get a morphism of complexes
3632: $$
3633: \begin{array}{ccccc}
3634: {\cal B}_3(F)&\stackrel{\delta_3}{\longrightarrow}&{\cal B}_2(F )\otimes
3635: F^{\ast}&
3636: \stackrel{\delta_3}{\longrightarrow}&\Lambda^3F^{\ast}\\
3637:   &&\downarrow {\rm Res}&&\downarrow {\rm Res}\\
3638: && {\cal B}_2(k )&\stackrel{\delta_2}{\longrightarrow}& \Lambda^2k^{\ast}
3639: \end{array}
3640: $$
3641:   
3642: We will also need a more explicit version $B_2(F)$ 
3643: of the group ${\cal B}_2(F)$. 
3644: Denote by $R_2(F)$  the subgroup of $\Z[{\Bbb P}^1(F)]$ 
3645: generated by the elements $$
3646: \{0\}, \{\infty\}\quad \mbox{and} \quad \sum_{i=1}^{5}(-1)^i
3647: \{r(x_1,...,\widehat   x_i,...,x_5)\}
3648: $$ 
3649: when $(x_1,...,x_5)$ runs through all $5$-tuples of distinct points in ${\Bbb P}^1(F)$,
3650: and the cross-ratio $r(...)$ is normalized by $r(\infty, 0, 1, x) =x$.  Define the Bloch group $B_2(F)$ as
3651: $$
3652: B_2(F):= \frac{\Z[{\Bbb P}^1(F)]}{R_2(F)}
3653: $$ 
3654: 
3655: 
3656: One can show that $R_2(F)\subset {\cal R}_2(F)$. 
3657: So there is a map 
3658: \begin{equation} \label{6.18.02.10}
3659: i: B_2(k)\lra {\cal B}_2(k)
3660: \end{equation}
3661: induced by the identity map on the generators. 
3662: 
3663: \begin{proposition} \label{mot}
3664: Let $k$ be a number field. Then (\ref{6.18.02.10})  is an isomorphism 
3665: modulo torsion.
3666: \end{proposition}
3667: 
3668: {\bf Proof}. The map $i$ is clearly surjective. The diagram
3669: $$
3670: \begin{array}{ccc}
3671: B_2(k) &\lra&{\cal B}_2(k) \\
3672: &&\\
3673: \downarrow \delta_2&&\downarrow \delta_2\\
3674: &&\\
3675: \Lambda^2k^*&=&\Lambda^2k^*
3676: \end{array}
3677: $$
3678: is commutative. So we need only to show that if $0 \not = x \in B_2(k)_{\Q}$ and $\delta_2(x)=0$, then $i(x)\not =0$. This  follows from the injectivity of the regulator map on $K^{ind}_3(k)_{\Q}$.  
3679: Indeed, by Suslin's theorem  for a field $F$ one has $K^{ind}_3(F)_{\Q} = {\rm Ker} \delta_2\otimes \Q $. Let us identify $K^{ind}_3(\C)_{\Q}$ with this subgroup of $B_2(\C)_{\Q} $. The restriction of the dilogarithm map
3680: $$
3681: B_2(k) \lra {\cal B}_2(k)\lra  (\Z[Hom(k,\C)]\otimes 2 \pi i \R)^+, \qquad \{z\}_2 \lms \{2 \pi i{\cal L}_2(\sigma_i(z))\}
3682: $$
3683:  to the subgroup ${\rm Ker} \delta_2\otimes \Q $  gives the  Borel regulator   $K^{ind}_3(k)_{\Q} \lra \R^{r_2}$ ([G1]) and  thus is injective   by Borel's theorem.  
3684: 
3685: {\bf Remark}. For any field  $k$ the rigidity conjecture for $K_3^{ind}$ 
3686: implies that the map $i$ should be an isomorphism, see [G1].
3687: 
3688: {\bf 2. The strong reciprocity law}. 
3689:  \begin{conjecture} \label{reh}
3690: Let $X$ be a regular projective curve over an algebraically closed field
3691: $k$ and $F:= k(X)^{\ast}$.
3692: Then there exists a canonical homomorphism of groups  
3693: $
3694: h:\Lambda^3F^* \rightarrow
3695: {\cal B}_2(k)
3696: $ 
3697: satisfying the following two conditions:
3698: 
3699: a) $h(k^* \wedge \Lambda^2F^*) =0$ and the diagram 
3700: \begin{equation} \label{6.17.02.3}
3701: \begin{array}{ccccc} \label{reh1}
3702: {\cal B}_3(F)& \stackrel{\delta_3}{\longrightarrow}&{\cal B}_2(F)\otimes F^{\ast} 
3703: & \stackrel{\delta_3}{\longrightarrow} & \Lambda^3F^{\ast}\\
3704: &&&&\\
3705: &&{\rm Res}\downarrow&h \swarrow&\downarrow {\rm Res}\\
3706: &&&&\\
3707: &&{\cal B}_2(k)&\stackrel{\delta_2}{\longrightarrow}&\Lambda^2 k^{\ast} 
3708: \end{array}
3709: \end{equation}
3710: is commutative.
3711: 
3712: b) If $X$ is a  curve over $\C$
3713: then 
3714: \begin{equation} \label{homotq}
3715: \frac{1}{2\pi i}\int_{X(\C)}r_2(f_1\wedge f_2 \wedge f_3) =   {\cal L}_2 \Bigl(h(f_1\wedge
3716: f_2 \wedge f_3)\Bigr) 
3717: \end{equation}
3718: \end{conjecture}
3719: 
3720: 
3721: 
3722: 
3723: 
3724: {\bf Remarks}. 1. b) follows easily if we have a functorial map $h$ such that 
3725: ${\rm Res} = \delta_2 \circ h $, see Theorem \ref{55} below.  
3726: 
3727: 2. According to  Suslin's 
3728:  reciprocity law for the Milnor group $K^M_3(F)$ the projection of 
3729: ${\rm Res}(\Lambda^3F^{\ast}) \subset \Lambda^2k^*$ to $K_2(k)$ is zero. 
3730: Since by Matsumoto's 
3731: theorem  $K_2(k)={\rm Coker}(\delta_2)$,  one has 
3732: ${\rm Res}(\Lambda^3F^{\ast}) \subset {\rm Im}(\delta_2)$. However  
3733: ${\rm Ker}(\delta_2)$ is nontrivial, so  it is {\it a priory} 
3734: unclear that we can lift naturally  the map ${\rm Res}$  to a map $h$.
3735:  One  of the reasons   why we can do   this  is provided by (\ref{homotq}).
3736: 
3737: 
3738: 
3739: 
3740: 
3741: 
3742: 
3743:  
3744: We prove  this conjecture  in the following cases: 
3745: 
3746: a)  $X={\Bbb P}^1$; we construct {\it explicitly} a reciprocity homomorphism  
3747: $h: \Lambda^3F^*\to  B_2(k)$ in Theorem \ref{p4}.
3748: 
3749: b)  $X$ is an elliptic curve over an  algebraically closed  field; we construct {\it explicitly} a reciprocity homomorphism  $h: \Lambda^3F^* \to  {\cal B}_2(k)$ in Theorem \ref{ecur}.
3750: 
3751: c) $k = {\overline \Q}$, $X$ is any curve; see Theorem \ref{rl}.
3752: 
3753: In the cases a)  and b)  the homomorphism $h$  satisfies the following additional property. Let $F = k(X)$ and $k$ is not necessarily algebraically closed. Let $k'$ be the field of definition of the divisors $(f_1), (f_2), (f_3)$. Then $h(f_1\wedge f_2\wedge f_3) \in {\cal B}_2(k')$. 
3754: 
3755: \begin{conjecture}
3756: Let $X$ be a projective regular curve over an algebraically closed field
3757: $k$ and $F:= k(X)$. Then the  homomorphism 
3758: $$
3759: {\rm Res}: \Gamma(F;n) \longrightarrow \Gamma(k;n-1)[-1]
3760: $$  
3761: is homotopic to zero.
3762: \end{conjecture}
3763: 
3764: \begin{lemma} \label{mot1}
3765: Assume that we have a map $h$ such that $h(k^* \wedge \Lambda^2F^*) =0$ and 
3766: ${\rm Res} = \delta_2 \circ h $. Then $ h \circ \delta_3 = {\rm Res}$.
3767: \end{lemma}
3768: 
3769: {\bf Proof}. The image of the group ${\cal B}_2(F)\otimes F^{\ast}$ under the map $h \circ \delta_3$ belongs to the subgroup ${\rm Ker}\delta_2$. Since $h(k^* \wedge \Lambda^2F^*) =0$ one has $h \circ \delta_3 = {\rm Res}$ on ${\cal B}_2(F)\otimes k^{\ast}$. Any element of $k(X)^*$ can be connected via a curve to a constant. This together with the rigidity of ${\rm Ker}\delta_2$ (built into the definition of the group ${\cal B}_2(k)$) implies the result.
3770: 
3771:  
3772: {\bf 3. The $X= \PP^1$ case}. 
3773: Recall that  $v_x(f)$ is the order of zero of $f\in k(X)$ at $x$. 
3774: Choose a point $\infty$ on ${\Bbb P}^1$.
3775: 
3776: \begin{theorem} \label{p4}
3777:  Assume that $k = \overline k$. Then the map $
3778: h:\Lambda^3 k({\Bbb P}^1)^* \rightarrow
3779:  B_2(k)
3780: $   given by the formula 
3781: $$
3782: h(f_1\wedge f_2\wedge f_3):=  \sum_{x_i \in {\Bbb P}^1(k)} 
3783: v_{x_1}(f_1)v_{x_2}(f_2)v_{x_3}(f_3)\{r(x_1, x_2, x_3, \infty)\}_2
3784: $$
3785: satisfies all the conditions of conjecture \ref{reh} modulo $6$-torsion.
3786: \end{theorem}
3787: 
3788: {\bf Proof}. Let us show that  $h$ is independent of the choice of  $\infty$, i.e. 
3789: $$
3790: \sum_{x_i \in {\Bbb P}^1(k)}
3791: v_{x_1}(f_1)v_{x_2}(f_2)v_{x_3}(f_3)\{r(x_1, x_2, x_3, a)\}_2
3792: \in B_2(k)
3793: $$
3794: does not depend on  $a$. Indeed, the 5-term
3795: relation  for the 5-tuple of points $(x_1, x_2, x_3, a, b)$ gives
3796: $$
3797: \sum_{x_i \in {\Bbb P}^1(k)}
3798: v_{x_1}(f_1)v_{x_2}(f_2)v_{x_3}(f_3)\Bigl(\{r(x_1, x_2, x_3, a)\}_2 -
3799: \{r(x_1, x_2, x_3, b)\}_2\Bigr) =
3800: $$
3801: $$
3802: -\sum_{x_i \in {\Bbb P}^1(k)}
3803: v_{x_1}(f_1)v_{x_2}(f_2)v_{x_3}(f_3)\Bigr(\{r(x_1, x_2, a, b )\}_2
3804: -\{r( x_1, x_3, a, b)\}_2 + \{r(x_2, x_3, a, b)\}_2\Bigr)
3805: $$
3806: Each of these 3 terms vanishes because $\sum_{x \in {\Bbb P}^1(k)}v_{x}(f) =0$.
3807: for any $f \in k(\PP^1)^*$. 
3808: 
3809:  \begin{proposition} \label{not}
3810: Let $k= \overline k$. Then modulo  $6$-torsion
3811: $$
3812: h((1-f) \wedge f \wedge g) = \sum_{x \in {\Bbb P}^1(k)}v_{x}(g)\{f(x)\}_2
3813: $$
3814: \end{proposition}
3815: 
3816: {\bf Proof}. Using linearity with respect to $g$ and projective invariance of the cross--ratio we see that it is sufficient to prove the identity for $g=t$. 
3817: Then it boils down to 
3818: \begin{equation} \label{7.3.02.2}
3819: \sum_{x_i \in {\Bbb P}^1(k)}
3820: v_{x_1}(1-f)v_{x_2}(f)\{r(x_1, x_2, 0, \infty)\}_2  = 
3821: \{f(0)\}_2 - \{f(\infty)\}_2
3822: \end{equation}
3823:  
3824: 
3825: \begin{lemma} \label{notxzc}
3826: Applying $\delta_2$ to both parts of 
3827: (\ref{7.3.02.2}) we get the same result  modulo  $6$-torsion.
3828: \end{lemma} 
3829: 
3830: {\bf Proof}. 
3831: Choose a coordinate $t$ on $\PP^1$ such that $f(\infty) =1$. 
3832: Then 
3833: \begin{equation} \label{7.3.02.4}
3834: f(t) = \frac{\prod_i (a_i -t)^{\alpha_i}}{\prod_k (c_k -t)^{\gamma_k}}; \quad 
3835: 1-f(t)= \frac{B \prod_j (b_j -t)^{\beta_j}}{\prod_k (c_k -t)^{\gamma_k}}
3836: \end{equation}
3837: Observe that  $\{f(\infty)\}_2 =0$ modulo $6$-torsion. 
3838: The left hand side equals 
3839: \begin{equation} \label{7.3.02.1}
3840: \sum_{x_i \in {\Bbb P}^1(k)}
3841: v_{x_1}(1-f)v_{x_2}(f)\{x_1/x_2\}_2  =
3842: \end{equation}
3843: $$
3844: \sum \beta_j \alpha_i \{b_j/a_i\}_2 - 
3845: \sum \gamma_k \alpha_i \{c_k/a_i \}_2
3846: - \sum \beta_j\gamma_k \{b_j/c_k\}_2  
3847: $$
3848: Applying $\delta_2$ to it we get 
3849: $$
3850: \sum \beta_j \alpha_i \cdot \frac{a_i-b_j}{a_i} \wedge \frac{b_j}{a_i} - 
3851: \sum \gamma_k \alpha_i \cdot \frac{a_i- c_k}{a_i} \wedge \frac{c_k}{a_i} 
3852: - \sum \gamma_k \beta_j \cdot \frac{c_k - b_j}{c_k} \wedge \frac{b_j}{c_k} = 
3853: $$
3854: $$
3855: \sum \beta_j \alpha_i \cdot b_j \wedge a_i + 
3856: \sum \alpha_i\gamma_k  \cdot a_i \wedge c_k  
3857: + \sum \gamma_k \beta_j \cdot c_k \wedge b_j \quad + 
3858: $$
3859: $$
3860: - \sum_i\frac{\prod_j (a_i-b_j)^{\beta_j }}{\prod_k (a_i-c_k)^{\gamma_k }} 
3861: \wedge   a_i^{\alpha_i}  
3862: +   \sum_j\frac{\prod_i (a_i - b_j)^{\alpha_i}}
3863: {\prod_k (c_k - b_j)^{\gamma_k}} \wedge  b_j^{\beta_j}
3864: -  \sum_k \frac{\prod (a_i- c_k)^{\alpha_i}}{\prod (c_k  - b_j)^{\beta_j}} 
3865: \wedge  c_k^{\gamma_k} 
3866: $$
3867: Using (\ref{7.3.02.4}) we see that the second line equals modulo $2$--torsion 
3868: to
3869: $$ 
3870: - \sum_i (1-f(a_i)) \wedge a_i^{\alpha_i} + \sum_j f(b_j) 
3871: \wedge b_j^{\beta_i}  - \sum_k
3872:  \frac{\prod (a_i- c_k)^{\alpha_i}}{\prod (b_j - c_k)^{\beta_j}} 
3873: \wedge   c_k^{\gamma_k} + 
3874: B \wedge \prod a_i^{\alpha_i}
3875: $$
3876: The first two terms are zero since $f(a_i)=0$ 
3877: and $f(b_j) =1$. The third term equals to 
3878: $-(B \wedge \prod c_k^{\gamma_k})$ since, as it follows from (\ref{7.3.02.4}), 
3879: $f(c_k) = \infty$ and thus 
3880: $$
3881: \frac{\prod (a_i- c_k)^{\alpha_i}}{\prod (b_j - c_k)^{\beta_j}}  = -B 
3882: $$
3883: On the other hand, we have 
3884: $$
3885: \delta_2(\{f(0)\}_2) = (1-f(0)) \wedge f(0) =  
3886: \frac{B\prod_j b_j^{\beta_j}}{\prod_k c_k^{\gamma_k}} \wedge 
3887: \frac{\prod_i a_i^{\alpha_i}}{\prod_k c_k^{\gamma_k}} 
3888: $$
3889: which matches the  expression we got for the left hand side. The lemma is proved.
3890: 
3891: To prove the proposition it remains to use a rigidity argument.
3892: Namely, we need to show that the identity is valid 
3893: for some particular $f$, which is easy, or use proposition \ref{p44} 
3894: plus injectivity of the regulator on $K_3(\overline \Q)_{\Q}$. 
3895: The proposition is proved. 
3896: 
3897: Now let us prove the key fact that  ${\rm Res} = \delta_2\circ h$. We need to show that 
3898:  for any 3 rational functions $f_1, f_2, f_3$ on
3899: ${\Bbb P}^1$  
3900: $$
3901: \sum_{x \in {\Bbb P}^1(k)} {\rm res}_x (f_1\wedge f_2\wedge f_3) = 
3902: $$
3903: \begin{equation} \label {1}
3904: \delta_2  \Bigl( \sum_{x_i \in {\Bbb P}^1(k)}
3905: v_{x_1}(f_1)v_{x_2}(f_2)v_{x_3}(f_3)\{r(x_1, x_2, x_3, \infty)\}_2 \Bigr)
3906: \end{equation}
3907: 
3908: Both sides are obviously homomorphisms
3909: from $\Lambda^3 F^{\ast}$ to $\Lambda^2 k^{\ast}$ which are zero on
3910: $k^{\ast}  \wedge \Lambda^2 F^{\ast}$. (The last property for the map
3911: $\sum {\rm res}_x$ 
3912: is provided by the Weil reciprocity law). We normalize the cross ratio of four points on the projective 
3913: line by  $r(\infty,0,1,z) =z$. 
3914:  So it suffices to
3915: check the formula on elements
3916: $$
3917: \frac{z-a_2}{z-a_1} \wedge \frac{z-b_2}{z-b_1} \wedge \frac{z-c_2}{z-c_1}
3918: $$
3919: In this case it follows from 
3920: $$
3921: \delta_2 \{r(a_2, b_2, c_2, \infty)\}_2 = \delta_2\left\{\frac{a_2-c_2}{b_2-c_2}\right\}_2 = \frac{b_2-a_2}{b_2-c_2}\wedge \frac{a_2-c_2}{b_2-c_2} 
3922: $$
3923: 
3924: 
3925: 
3926: It remains to prove the following proposition.
3927: 
3928: \begin{proposition} \label{p44}
3929: $$
3930: {\cal P}_2({\Bbb P}^1;f_1,f_2,f_3)=   \sum_{x_i \in {\Bbb P}^1(\C)} v_{x_1}(f_1)v_{x_2}(f_2)v_{x_3}(f_3){\cal L}_2(r(x_1, x_2, x_3, \infty))
3931: $$
3932: \end{proposition}
3933: 
3934: {\bf Proof}.  We  immediately reduce the statement  to the situation when $f_1 = 1-z, f_2 = z, f_3 = z-a$ 
3935: which is a particular case of the following lemma  
3936: 
3937: \begin{lemma} \label{l454} Let $X$ be an arbitrary curve over $\C$. Then
3938: $$
3939: \int_{X(\C)}r_2((1-f)\wedge f\wedge g)= - \sum_{x \in X(\C)} v_{x}(g) \cdot
3940: {\cal L}_2(f(x))
3941: $$
3942: \end{lemma}
3943: 
3944: {\bf Proof}. For  functions $f(z)$ and $g(z)$ on $X(\C)$ set
3945: $$
3946: \alpha(f,g): = \log|f|d\log|g| - \log|g|d\log|f|
3947: $$
3948:  Consider the following 1-form on $X(\C)$
3949: \begin{equation} \label{99}
3950:   {\cal L}_{2}(f) d\arg g  -\frac{1}{3} \alpha(1-f,f)\log|g|
3951: \end{equation}
3952: It defines a current on $X(\C)$. We claim that its derivative is equal to:
3953: \begin{equation} \label{11}
3954:   2 \pi \cdot {\cal L}_{2}(f)\delta(g)  + r_2((1-
3955: f) \wedge f \wedge g) 
3956: \end{equation}
3957: Using $d (d \arg g) = 2\pi  \cdot \delta(g)$
3958: %where $\delta(g)$ is a $2$-current given by $<\delta(g), \varphi(x)>:= 
3959: %\sum_{x\in X(\C)} v_x(g)\varphi(x)$ for any smooth test function $\varphi(x)$
3960: and 
3961: \begin{equation} \label{dl2}
3962: d{\cal L}_2(z) = -\log|1-z| d\arg z  + \log|z| d\arg(1-z)
3963: \end{equation}
3964: we see that the  differential of the current (\ref{99}) equals to
3965: $$
3966: 2 \pi  {\cal L}_{2}(f)\delta(g) \quad + \quad \Bigl( -\log|1-f| d\arg f  + \log|f| d\arg(1-f)\Bigr)\wedge d\arg g +
3967: $$
3968: $$
3969: \frac{1}{3}\Bigl(\log|1-f|d\log|f| - \log|f|d\log|1-f|\Bigr)\wedge d\log|g| - \frac{2}{3}\log|g| \cdot d\log|1-f|\wedge d\log|f|
3970: $$
3971: Since $d\log(1-f)\wedge d\log f  = 0$ we have $$
3972: d\log|1-f|\wedge d\log|f| = d\arg(1-f)\wedge d\arg f$$
3973:  Using this and writing  $r_2(f_1 \wedge f_2 \wedge f_3)$ as
3974: $$
3975: \frac{1}{3}(\log|f_1|d\log|f_2|\wedge d\log|f_3| + \quad \mbox{cyclic permutations})
3976: $$
3977: $$
3978:  -  (\log|f_1|d\arg(f_2)\wedge d\arg f_3 + \quad \mbox{cyclic permutations})
3979: $$
3980: we come to (\ref{11}).
3981: Integrating we get the lemma. The theorem is proved. 
3982: 
3983: {\bf 4. Expressing the Chow dilogarithm via  the classical one}. Let $\pi: 
3984: Y \rightarrow S$ be a family of curves over a base $S$ over $\C$ and 
3985: $f_1,f_2,f_3 \in \C(Y)^*$. 
3986: We get a function at the generic point  of $S$.  Its value at 
3987:  $s \in S$ is given by the Chow dilogarithm 
3988: ${\cal P}_2(Y^s;f^s_1,f^s_2,f^s_3)$, where 
3989: $Y^s$ is the fiber of 
3990: $\pi$ at $s$. Denote it by 
3991: ${\cal P}_2(Y \rightarrow S ;f_1,f_2,f_3)$.
3992: 
3993: \begin{theorem} \label{55}
3994: a) Let $\pi: 
3995: Y \rightarrow S$ be a family of curves over a  base $S$ over $\C$. Then 
3996: there are rational functions $\varphi_i$ on $S$ such that  
3997: $$
3998: {\cal P}_2(Y \rightarrow S;f_1,f_2,f_3) = \sum_i {\cal L}_2(\varphi_i(s))
3999: $$
4000: 
4001: b) Let $k = \C(S)$, $X$ is the generic fiber of $\pi$, and $F=k(X)$. Suppose that there exists a map $h:\Lambda^3F^* \rightarrow
4002: {\cal B}_2(k)$ such that ${\rm Res} = \delta_2 \circ h $. Then 
4003: \begin{equation} \label{3211}
4004:  d{\cal P}_2(Y \rightarrow S;f_1,f_2,f_3) \quad = \quad d{\cal L}_2(h(f_1,f_2,f_3))
4005: \end{equation}
4006: \end{theorem}
4007: 
4008: {\bf Proof}. 
4009: a) 
4010: We use  the existence of the transfer map on  $K^M_3$  
4011: to reduce the statement  to the case  $X= {\Bbb P}^1$. 
4012: 
4013: Choose a projection $p: X \rightarrow {\Bbb P}^1$. 
4014: We may suppose without loss of generality that $p$ is a (ramified) 
4015: Galois covering 
4016: with the Galois group $G$. 
4017:  Indeed, let  $p_1:Y \rightarrow X$ 
4018: be such a covering that its composition with $p$ 
4019:  is a Galois covering. Indeed, 
4020: $$
4021: {\cal P}_2(Y\lra S; p_1^*f_1,p_1^*f_2, p_1^*f_3) = \frac{1}{{\rm deg} p_1}{\cal P}_2(X\lra S; f_1, f_2, f_3)
4022: $$
4023: 
4024: Then $\sum_{g \in G} g^*\{f_1, f_2, f_3\} \in p^*K_3^M(k({\Bbb P}^1))$. 
4025: It coincides with $p^*$ of the transfer of the element $\{f_1, f_2, f_3\} \in K_3^M(F)$. 
4026: This means that there exist  $s^{(i)}_1,s^{(i)}_2,s^{(i)}_3  \in k({\Bbb P}^1)$ 
4027: and $g_j,h_j  \in k(X)$ such that
4028: \begin{equation} \label{6.18.02.1}
4029: \sum_{g \in G} g^*(f_1\wedge f_2 \wedge f_3) - p^*\sum_i 
4030: s^{(i)}_1\wedge  s^{(i)}_2\wedge s^{(i)}_3 =  
4031: \sum_j(1-g_j)\wedge g_j \wedge h_j
4032: \end{equation} 
4033: Therefore
4034: $$
4035:  {\cal P}_2(Y \rightarrow S;f_1,f_2,f_3)= 
4036: \frac{1}{|G|}\sum_i {\cal P}_2({\Bbb P}^1 \times S \rightarrow S;
4037:  s^{(i)}_1, s^{(i)}_2, s^{(i)}_3) + 
4038: $$
4039: $$
4040: \sum_j {\cal P}_2({\Bbb P}^1 \times S \rightarrow S; (1-g_j), g_j, h_j)
4041: $$
4042: It remains to use Lemma  \ref{l454}  and Proposition  \ref{p44}. 
4043: The part a) of the theorem is proved. 
4044: 
4045: b) We need the following lemma.
4046: \begin{lemma} \label{2.14}
4047: \begin{equation} \label{dhomot}
4048: d {\cal P}_2(Y \rightarrow S;f_1,f_2,f_3) =  (2 \pi)^{-1} \cdot {\rm Alt}_3 \Bigl(v_x(f_1) \log|f_2(x)| d_s \arg f_3(x) \Bigr)
4049: \end{equation}
4050: \end{lemma}
4051: 
4052: {\bf Proof}. Using $dd \log f  = 2\pi i \delta(f)$  we get  an equality of $3$-currents on $Y$
4053: $$
4054: dr_2(f_1,f_2,f_3) = 
4055: $$
4056: \begin{equation} \label{2233}
4057:  \pi_3\Bigl(\frac{df_1}{f_1}\wedge\frac{df_2}{f_2}\wedge\frac{df_3}{f_3}\Bigr) +  2 \pi  \cdot {\rm Alt}_3 \Bigl(\delta(f_1) \log|f_2(x)| d \arg f_3(x) \Bigr)
4058: \end{equation}
4059: The second term in (\ref{2233}) is a $1$-form on the divisor 
4060: $D:= \cup_{i=1}^3{\rm div}(f_i)$ considered as  a $3$-current on $Y$.  
4061: This $1$-form is the composition of the residue map
4062: $$
4063: {\rm res}: \Lambda^3 \C(Y)^* \longrightarrow \coprod_{X \in Y_1}
4064: \Lambda^2\C(X)^*
4065: $$
4066: with the map
4067: $$
4068: r_1: \Lambda^2 \C(X)^* \longrightarrow  {\cal A}^1({\rm Spec }(\C(X))), \qquad f \wedge g \longmapsto -2\pi (\log|f| d\arg g - \log|g| d\arg f)
4069: $$
4070: 
4071: 
4072: The push forward of the first term in (\ref{2233}) vanishes  
4073: (since the fibers are complex curves the push down  of any $(3,0)$-form 
4074: to $S$ is zero). Integrating the second $3$-current  in (\ref{2233}) 
4075: along the fibers of $Y$ we get (\ref{dhomot}). The lemma is  proved. 
4076: 
4077: 
4078: 
4079:  According to  Lemma \ref{2.14}  and formula (\ref{dl2}) for $d {\cal L}_2$, and using ${\rm Res} = \delta_2 \circ h $ we get the proof of the 
4080: part b) of the theorem. 
4081: 
4082: 
4083: 
4084: 
4085: {\bf Remark}  The  function ${\cal L}_2(z)$ is continuous on $\C {\Bbb P}^1$. Therefore part a) of the theorem 
4086: implies that the function ${\cal P}_2(Y \rightarrow S ;f_1,f_2,f_3)$ can be extended to a continuous 
4087: function on $S$.
4088: 
4089: 
4090: {\bf 5. Conjecture \ref{reh} for $\overline k = \overline \Q$}. 
4091: \begin{theorem} \label{rl}
4092: Let $X$ be a   regular  projective curve over  
4093: ${\overline \Q}$ and $F:= {\overline \Q}(X)$. 
4094: Then there exists a homomorphism   
4095: $
4096: h:\Lambda^3F^* \rightarrow
4097: B_2({\overline \Q})\otimes \Q
4098: $ as in conjecture \ref{reh}  such that for any embedding $\sigma: {\overline \Q} 
4099: \hookrightarrow  \C$ one has 
4100: \begin{equation} \label{homot}
4101: \frac{1}{2\pi i}\int_{X(\C)}r_2(\sigma(f_1\wedge f_2 \wedge f_3)) =  {\cal L}_2 \Bigl(\sigma(h(f_1\wedge
4102: f_2 \wedge f_3))\Bigr) 
4103: \end{equation}
4104: \end{theorem}
4105:  
4106: 
4107: 
4108: {\bf Proof}. It is similar to the proof of theorem \ref{55}. 
4109: Choose a projection $p:X \rightarrow {\Bbb P}^1$. 
4110: We may suppose that $p$ is a Galois covering 
4111: with the Galois group $G$. 
4112:  Indeed, let  $p_1:Y \rightarrow X$ 
4113: be such a covering that its composition with $p$ 
4114:  is a Galois covering. 
4115: Setting  
4116: $$
4117: h(f_1\wedge f_2 \wedge f_3):= h(\frac{1}{{\rm deg}p_1} \cdot p_1^*(f_1\wedge f_2 \wedge f_3))
4118: $$ 
4119: we may suppose that $p$ is Galois.  
4120: 
4121: Then $\sum_{g \in G} g^*\{f_1, f_2, f_3\} \in p^*K_3^M(k({\Bbb P}^1))$. 
4122: So there exist  $s^{(i)}_1,s^{(i)}_2,s^{(i)}_3,  \in k({\Bbb P}^1)$ 
4123: and $g_j,h_j  \in k(X)$ such that (\ref{6.18.02.1}) holds. 
4124: Set
4125: $$
4126: |G| \cdot h(f_1\wedge f_2 \wedge f_3) := \sum_i h(s^{(i)}_1\wedge  s^{(i)}_2\wedge s^{(i)}_3) + \sum_j \sum_{x \in X(\overline \Q)}\{g_j(x)\}_2 \cdot v_x(h_j)
4127: $$
4128:    \begin{lemma} \label{1.8}
4129: Suppose $\sum_i (1-f_i) \wedge f_i \wedge g_i = 0$ in  $\Lambda^3 {\overline \Q}(X)^*$. 
4130: Then $$
4131: \sum_i \sum_{x \in X(\overline \Q)} v_x(g_i) \cdot \{f_i(x)\}_2 =0 \quad \mbox{in the group $ B_2({\overline \Q})$.}
4132: $$ 
4133: \end{lemma}
4134: 
4135:  This lemma implies that $h$ is well defined.
4136: Indeed, suppose that we have a different presentation  
4137: $$
4138: \sum_{g \in G} g^*\{f_1, f_2, f_3\} = p^*\sum_k 
4139: \widetilde  s^{(k)}_1\wedge  \widetilde  s^{(k)}_2\wedge \widetilde  
4140: s^{(k)}_3 +  \sum_j(1-\widetilde  g_j)\wedge \widetilde  g_j \wedge \widetilde  h_j
4141: $$
4142: We need to show that 
4143: \begin{equation} \label{**}
4144: h\Bigl( \sum_k \widetilde  s^{(k)}_1\wedge  \widetilde  s^{k)}_2\wedge 
4145: \widetilde  s^{(k)}_3 - 
4146:  \sum_i  s^{(i)}_1\wedge   s^{(i)}_2\wedge  s^{(i)}_3\Bigr) + 
4147: \end{equation} 
4148: $$
4149: \sum v_x(h_j)\{g_j(x)\}_2 - \sum v_x(\widetilde  h_j)\{\widetilde  g_j(x)\}_2 = 0
4150: $$
4151: There exist $a_j, b_j \in k({\Bbb P}^1)$ such that modulo $k^* \wedge 
4152: \Lambda^2k({\Bbb P}^1)^*$  one has   
4153: $$
4154:  \sum_k \widetilde  s^{(k)}_1\wedge  \widetilde  s^{(k)}_2\wedge \widetilde  s^{(k)}_3 - 
4155:  \sum_i  s^{(i)}_1\wedge   s^{(i)}_2\wedge  s^{(i)}_3 - \sum_j (1-a_j)\wedge a_j 
4156: \wedge b_j =0
4157: $$
4158: According to  Theorem \ref{p4} the homomorphism $h$ for ${\Bbb P}^1$ annihilates 
4159: the left hand side. On the other hand 
4160: $$
4161: p^*\Bigl( \sum_k \widetilde  s^{(k)}_1\wedge  \widetilde  s^{(k)}_2\wedge \widetilde  
4162: s^{(k)}_3 - 
4163:  \sum_i  s^{(i)}_1\wedge   s^{(i)}_2\wedge  s^{(i)}_3\Bigr)  - 
4164: $$
4165: $$
4166: \sum_j(1-\widetilde  g_j)\wedge \widetilde  g_j \wedge \widetilde   h_j +\sum_j(1- g_j)\wedge  g_j \wedge h_j = 0
4167: $$
4168: Using Lemma \ref{1.8} we get (\ref{**}). 
4169: To get (\ref{homot}) we use Theorem \ref{55} and notice that
4170: \begin{equation} \label{***}
4171: {\cal P}_2(Y \rightarrow S;f_1,f_2,f_3) = 1/m \cdot {\cal P}_2(Z \rightarrow S;
4172: p_1^*f_1, p_1^*f_2, p_1^*f_3)
4173: \end{equation} 
4174:  
4175:  
4176: 
4177: {\bf Proof of Lemma \ref{1.8}}.  
4178: For a regular curve $X$ over an algebraically closed field $k$ there is a commutative diagram  
4179: $$
4180: \begin{array}{ccc} \label{reh1a}
4181: B_2(F)\otimes F^{\ast} 
4182: & \stackrel{\delta_3}{\longrightarrow} & \Lambda^3F^{\ast}\\
4183: &&\\
4184: {\rm Res}\downarrow&&\downarrow {\rm Res}\\
4185: &&\\
4186: B_2(k)\otimes \Z[X(k)]&\stackrel{\delta_2}{\longrightarrow}&\Lambda^2 k^{\ast}\otimes \Z[X(k)] 
4187: \end{array}
4188: $$
4189: Thus for any point $x$ of the curve $X$ the element $\sum_{i} v_x(g_i) \cdot 
4190: \{f_i(x)\}_2 $ lies in ${\rm Ker }\delta_2$. Therefore it defines an 
4191: element $\gamma_x \in K_3({\overline \Q})_{\Q}$.
4192: 
4193:  For any  embedding $\sigma: {\overline \Q} 
4194: \hookrightarrow  \C$ the value of the 
4195: Borel regulator on $\sigma(\gamma_x)$  is equal to $\sum_{x \in X} v_x(g_i) \cdot {\cal L}_2(\sigma(f_i(x)))$. 
4196: So by Lemma \ref{l454} the value of the 
4197: Borel regulator on $\sum_x \sigma(\gamma_x)$ is equal to $2 \pi  \cdot\int_{\C {\Bbb P}^1} r_2(\sum_i (1-f_i) \wedge f_i \wedge g_i)$
4198: and hence it is zero by our assumption. So Borel's theorem implies that the element is also zero.
4199: 
4200: A similar argument using Lemma \ref{1.8} shows that the homomorphism $h$ does not depend on the choice of the (finite) Galois extension of $\Q({\Bbb P}^1)$ containing the field $\Q(X)$.
4201: 
4202: 
4203: 
4204: {\bf 6. Explicit formulas for the reciprocity homomorphism $h$ and the Chow dilogarithm in the case of an elliptic curve}. 
4205: Let $E$ be an elliptic curve. We want to calculate the integral 
4206: $\int_{E(\C)} r_2(f_1 \wedge f_2 \wedge f_3)$. 
4207: Let us suppose that $E$ is realized as a
4208: plane curve. Then any rational function $f$ on $E$ can be written as a ratio
4209: of products of {\it linear} homogeneous functions: 
4210: $$
4211: f = \frac{l_1\cdot ... \cdot l_k}{l_{k+1}\cdot ... \cdot l_{2k}}
4212: $$
4213: So it is enough to calculate the integral
4214: $\int_{E(\C)} r_2(l_1/l_0 \wedge l_2/l_0 \wedge l_3/l_0)$  where $l_i$ are 
4215:   linear functions in homogeneous coordinates. 
4216: We will do this in a more general setting. 
4217: 
4218: {\it Notations}. Let $X$ be a plane algebraic curve
4219:  and $l_i$  linear functions in homogeneous coordinates. 
4220: Denote by  $L_i$   the line $l_i =0$ in 
4221: $\PP^2$. 
4222: Let $D_i$ be the divisor $L_i \cap X$. 
4223: Set $l_{ij}:= L_i \cap L_j$. 
4224: For three points $a,b,c$ and a 
4225: divisor $D =\sum n_i (x_i)$ on a 
4226: line we will use  the following notation (see Figure \ref{kj}):
4227: $$
4228: \{r(a,b,c,D)\}_2:= \sum_i n_i \{r(a,b,c,x_i)\}_2
4229: $$ 
4230:  
4231: 
4232: 
4233: \begin{figure}[ht]
4234: \centerline{\epsfbox{kj.eps}}
4235: \caption{Defining $\{r(a,b,c,D)\}_2$ for a plane algebraic curve $X$.}
4236: \label{kj}
4237: \end{figure}
4238: 
4239: \begin{theorem} \label{ecur}
4240:  Let $E$ be an elliptic curve over an algebraically closed field $k$. Then there exists a   homomorphism of groups $h:\Lambda^3F^* \rightarrow
4241: {\cal B}_2(k)$ such that for any linear homogeneous functions $l_0,...,l_3$ one has  
4242: \begin{equation} \label{hrule}
4243: h(l_1/l_0 \wedge l_2/l_0 \wedge l_3/l_0) = -\sum_{i=0}^3(-1)^i\{r(l_{i0},..., \widehat l_{ii}, ..., \l_{i3}, D_i)\}_2
4244: \end{equation}
4245: and which satisfies  all the properties of conjecture \ref{reh}. In particular, if $k =\C$ then
4246: \begin{equation} \label{homot1}
4247: {\cal P}_2(E; f_1\wedge f_2 \wedge f_3) =  {\cal L}_2 \Bigl( h(f_1\wedge
4248: f_2 \wedge f_3 )\Bigr) 
4249: \end{equation}
4250: \end{theorem}
4251: 
4252: 
4253:   
4254: {\bf Proof}. Suppose we have four generic lines $L_0,L_1,L_2,L_3$ in $\PP^2$. Any two of them, say   $L_0$ and $L_1$, provide  a {\it canonical} rational function $(l_0/l_1)$ on $\PP^2$ with the divisor $L_0 - L_1$ normalized by the condition that its value at the point $l_{23}$  is equal to $1$. 
4255: 
4256: 
4257: \begin{lemma} \label{3.9}
4258: a) On the line $L_3$ one has $(l_1/l_0) + (l_2/l_0) =1$.
4259: 
4260: b) $\frac{(l_1/l_0)}{(l_2/l_0)} = -(l_1/l_2)$
4261: \end{lemma} 
4262: 
4263: {\bf Proof}.  Let $m$ be a point on the line $L_3$. Then
4264: \begin{equation} \label{88}
4265: (l_1/l_0)(m) = r(l_{03},l_{13},l_{23},m); \qquad (l_2/l_0)(m) = r(l_{03},l_{23},l_{13},m)
4266: \end{equation}
4267: This gives a).
4268: It follows from this that if the point $m$ approaches the point $l_{03}$ then $\frac{(l_1/l_0)}{(l_2/l_0)}$ tends to $-1$. This implies b).
4269: 
4270: \begin{lemma} \label{lrez}
4271: For any plane curve $X$ one has
4272: $$
4273: \sum_{x \in X} {\rm res}_x \Bigl((l_1/l_0) \wedge (l_2/l_0) \wedge ( l_3/l_0)\Bigr) =
4274: - \delta_2\Bigl(\sum_{i=0}^3(-1)^i    \{r(l_{i0},..., \widehat l_{ii}, ..., \l_{i3}, D_i)\}_2\Bigr)
4275: $$
4276: \end{lemma}
4277: 
4278: 
4279: {\bf Proof}. Let us compute first the   residues at the divisors $D_1,D_2,D_3$  using part a) of Lemma \ref{3.9}. For example 
4280: the residue at $x \in D_1$  is equal to
4281: $$
4282: v_x((l_1/l_0))\cdot (l_2/l_0)(x) \wedge (l_3/l_0)(x) = v_x((l_1/l_0))\cdot (1-(l_3/l_0)(x) ) \wedge (l_3/l_0)(x) 
4283: $$
4284: $$
4285: \stackrel{(\ref{88})}{=} v_x((l_1/l_0))\cdot \{r(l_{13}, l_{10}, l_{12}, x)\}_2 = v_x((l_1/l_0))\cdot \{r(l_{10}, l_{12}, l_{13}, x)\}_2
4286: $$
4287: It remains to compute the residues on the line $L_0$. According to part b) of  Lemma \ref{3.9} one has
4288: $$
4289: (l_1/l_0) \wedge (l_2/l_0) \wedge ( l_3/l_0) = - (l_0/l_1) \wedge (l_2/l_1) \wedge 
4290: (l_3/l_1) 
4291: $$
4292: Using this we reduce the calculation to the previous case. 
4293: 
4294:  \begin{proposition} \label{hrez} Let $E$ be an elliptic curve over an algebraically closed field $k$. Then formula (\ref{hrule}) provides a well defined homomorphism of groups $h:\Lambda^3F^* \rightarrow
4295: {\cal B}_2(k)$.
4296: \end{proposition}
4297: 
4298: 
4299: 
4300: 
4301: {\bf Proof}. Let $D:= \sum_i n_i(x_i)$ be the divisor of a rational function $f$ on $E$. 
4302: To decompose it into a fraction of products of linear homogeneous functions $l_i$ we proceed as follows. Let $l_{x,y}$ (resp. $l_{x }$)  be a linear homogeneous equation of the line in $\PP^2$ through the points $x$ and $y$ on $E$  (resp. $x $ and $-x $).  The divisor of the function $l_{x,y}/l_{x}$ is $(x) +(y) - (x+y) - (0)$. 
4303: If  $D = (x)+(y) + D_1$, we write $f =
4304: l_{x,y}/l_{x+y}\cdot f'$, so $(f') = (0) + (x+y) + D_1$. 
4305:  After a finite number of such steps we get the desired decomposition. 
4306: 
4307: 
4308: 
4309: \begin{figure}[ht]
4310: \centerline{\epsfbox{kj3.eps}}
4311: \caption{}
4312: \label{kj3}
4313: \end{figure}
4314: 
4315: There are the following relations
4316: $$
4317: \frac{l_{x,y}}{l_{x+y}} \cdot \frac{l_{x+y,z}}{l_{x+y+z}} \Big /\frac{l_{y,z}}{l_{y+z}} \cdot \frac{l_{x,y+z}}{l_{x+y+z}} \quad =  \quad \mbox{constant}
4318: $$
4319: 
4320: One can prove that they generate all the relations between the functions $l_{x,y}/l_{x+y}$. 
4321:  So $h$ is well defined if it annihilates the following expression:
4322: $$
4323:  F(x,y,z; l_0,l_2,l_3):= \Bigl(\frac{l_{x,y}}{l_{x+y}} \cdot \frac{l_{x+y,z}}{l_{x+y+z}} \Big / \frac{l_{y,z}}{l_{y+z}} \cdot \frac{l_{x,y+z}}{l_{x+y+z}} \Bigr) \wedge (l_2/l_0) \wedge ( l_3/l_0)  
4324: $$
4325: It follows from   Lemma \ref{lrez} that $\delta_2(F(x,y,z; l_0,l_2,l_3)) =0$. Thus according to the definition of the group ${\cal B}_2(k)$ it is enough to check that $h(F(x',y',z'; l_0,l_2,l_3))=0$ for a certain triple of points $(x',y',z')$. It is easy to see that $h(F(a,a,a; l_0,l_2,l_3))=0$ since then the first factor in $F$ is a constant. The proposition is proved.  
4326: 
4327: 
4328: 
4329: \begin{proposition} \label{hrezza} Let $X$ be an algebraic curve in $\PP^2$ over $\C$ and $l_0,...,l_3$ linear homogeneous functions on $\C^3$. Then one has (using the notations defined above)
4330: \begin{equation} \label{elfo}
4331: \int_{X(\C)} r_2(l_1/l_0 \wedge l_2/l_0 \wedge l_3/l_0) = 2 \pi \cdot \sum_{i=0}^3(-1)^i
4332: {\cal L}_2(r(l_{i0},..., \widehat l_{ii}, ..., \l_{i3}, D_i)
4333: \end{equation}
4334: \end{proposition}
4335: 
4336: {\bf Proof}. It follows from Lemma \ref{lrez} and Theorem \ref{55} that the differentials of the both sides coincide. So their difference is a constant. 
4337: To show that this constant is zero 
4338: we deform $X$ to a   union of lines in $\PP^2$. Using Proposition  \ref{p44} one sees that   formula (\ref{elfo}) is valid when $X$ is a 
4339:  line in $\PP^2$. 
4340: 
4341: 
4342: 
4343: 
4344: \section{Appendix: on volumes of simplices in symmetric spaces}
4345: 
4346:   {\bf 1. Volumes of hyperbolic geodesic simplices as boundary integrals}.
4347:  A point $y$ of  the $n$-dimensional hyperbolic space ${\cal H}_n$ defines a one 
4348: dimensional space $M_y$ of volume forms  on the absolute $\partial{\cal H}_n$. It  
4349: consists of the volume forms  invariant under the 
4350: action of the isotropy group of $y$. We write them as follows. 
4351: Let $x_0, ..., x_{n}$ be the coordinates in a vector space $V_{n+1}$ of dimension $n+1$, 
4352: and 
4353: $
4354: Q(x):= x_0^2 + ...  +  x_{n-1}^2 - x_{n}^2 
4355: $. 
4356: Then ${\cal H}_n$ can be realized as the  
4357: projectivization of the cone $Q(x) < 0$, and its boundary is the projectivization of the cone
4358:  $Q(x) = 0$. Choose a point $y \in {\cal H}_n$. Lifting $y$ to a vector $y' \in V_{n+1}$ 
4359: we have the following 
4360: volume form on the boundary $\partial{\cal H}_n$:
4361: $$
4362: \frac{\delta(Q(x)) \sigma_{n+1}(x,dx)}{(x,y')^{n-1}}
4363: $$
4364: If  $y$ belongs to the boundary $\partial{\cal H}_n$, this formula provides a 
4365: space $M_y$ of singular volume forms   on the absolute; they are invariant under the 
4366:   isotropy group of  $y$. 
4367: 
4368: 
4369: 
4370: Let us choose for any point $y$ such a volume form $\mu_y$. 
4371: For two points $x,y$ the 
4372: ratio $  \mu_x/\mu_y$ is a nonzero  function on the absolute.  
4373: 
4374: 
4375: Let $I(y_0,...,y_{n })$ be the geodesic simplex with vertices at 
4376: $y_0,...,y_{n }$ where the points $y_i$ could be on the absolute. 
4377: Denote by ${\rm vol}I(y_0,...,y_{n})$ the volume of this simplex with respect to the invariant volume form  in ${\cal H}_n$ normalized by the following condition: if we realize the hyperbolic space as the interior of the unit ball $y_1^2 + ... + y_n^2 \leq 1$ then the volume form restricted to the tangent space at the origin $(0, ..., 0)$ is $dy_1 \wedge ... \wedge dy_n$.  
4378:  
4379: \begin{theorem} \label{2112}
4380: For any hyperbolic geodesic simplex 
4381: $I(y_0,...,y_{n})$ one has 
4382: \begin{equation} \label{1221} 
4383: \frac{(n-1)^n{\rm vol}(S^{n-1})}{n} \cdot {\rm vol}I(y_0,...,y_{n}) =   
4384: \int_{\partial {\cal H}_n}
4385: \log | \frac{\mu_{y_1}}{\mu_{y_0}}| d\log|\frac{\mu_{y_2}}{\mu_{y_0}} | \wedge ... \wedge d\log| \frac{\mu_{y_{n}}}{\mu_{y_0}} | 
4386: \end{equation} 
4387:  \end{theorem}
4388: 
4389: 
4390:   
4391:  Let    $\varphi(y_0, ..., y_{n})$ be the function defined by the right hand side of (\ref{1221}). Thanks to property 2) of Proposition \ref{1.2.} it does not depend on the choice of invariant volume forms $\mu_y$.
4392: 
4393: 
4394: 
4395: \begin{proposition} \label{1.21.1}
4396: The function $\varphi(y_0,...,y_{n})$   has  the following properties: It is
4397: 
4398: 1)  A smooth function on the vertices $y_i$.
4399: 
4400: 2) Equal to zero if three of the vertices  belong to the same geodesic. 
4401: 
4402: 3) Additive with respect to cutting of a simplex, i.e. if $y_0,...,y_{n}$ are points 
4403: such  that $y_0,y_1,y_2 $ are on the same geodesic, $y_1$ between $y_0$ and $y_2$, then
4404: $$
4405: \varphi(y_0,y_2,...,y_{n+1}) = \varphi(y_0,y_1,y_3, ...,y_{n+1})  + \varphi(y_1,y_2,...,y_{n+1}) 
4406: $$
4407: 
4408: 4) Invariant under the action of the symmetry group $SO(n,1)$.
4409: \end{proposition}
4410: 
4411: {\bf Proof}. 1) This is clear. 
4412: 
4413: 2) Let us  realize the  hyperbolic space as the interior part of  
4414:  the unit ball in $\R ^{n}$. Consider the geodesic $l$ passing through  the center of the ball  in the vertical  direction. The subgroup $SO(n-1) \subset SO(n,1)$ preserves pointwise this geodesic. So for any point $y$ on the geodesic the invariant volume form $\mu_y$  is invariant under the action of the group $SO(n-1)$. 
4415:  The quotient of $\partial {\cal H}_n$  under the action of $SO(n-1)$ is given by the projection $p: \partial {\cal H}_n \longrightarrow \R $ onto the vertical axis. 
4416: Take three points $y_1,y_2,y_3$ on the geodesic.  Then $\frac{\mu_{y_2}}{\mu_{y_1}}$ and $\frac{\mu_{y_3}}{\mu_{y_1}}$ are  lifted  from the line $\R $. Therefore $d\log|\frac{\mu_{y_2}}{\mu_{y_1}}| \wedge d\log|\frac{\mu_{y_3}}{\mu_{y_1}}|$ = 0.
4417:  So for a degenerate simplex $(y_0,...,y_n)$ one has 
4418: $$ 
4419: \log|\frac{\mu_{y_0}}{\mu_{y_1}}| d\log|\frac{\mu_{y_2}}{\mu_{y_1}}| 
4420: \wedge d\log|\frac{\mu_{y_3}}{\mu_{y_1}}|\wedge ... \wedge d\log|
4421: \frac{\mu_{y_n}}{\mu_{y_1}}| =0
4422: $$
4423: It remains to mention the skewsymmetry  of the integral (\ref{1221}). 
4424: The property 2) is proved.
4425: 
4426: 3) Follows from 2) and the additivity property from  Proposition  
4427: \ref{1.2.}.
4428: 
4429: 4) This is clear from 4) of Proposition  
4430: \ref{1.2.}. The  proposition is proved.
4431: 
4432: 
4433: 
4434: The leading term of the Taylor expansion of the function $\varphi(y_0, y_1, ..., y_n)$ 
4435: when $y_0$ is fixed and $y_1, ..., y_n$ are near $y_0$ provides an exterior $n$-form in 
4436: $T_{y_0}{\cal H}_n$ denoted $\varphi_{y_0}(Y_1, ..., Y_n)$,  $Y_i \in T_{y_0}{\cal H}_n$. 
4437: Let us compare it with the volume form 
4438: ${\rm Vol}_{y_0}(Y_1, ..., Y_n)$ in $T_{y_0}{\cal H}_n$ normalized as 
4439: before Theorem 7.1. 
4440: 
4441: \begin{lemma} \label{1.21.1}
4442: $\varphi_{y_0}(Y_1, ..., Y_n)  = \frac{(n-1)^n}{n}{\rm vol}(S^{n-1}) \cdot
4443: {\rm Vol}_{y_0}(Y_1, ..., Y_n)$. 
4444: \end{lemma}
4445: 
4446: {\bf Proof}.   Below we abuse notations by writing $y$  for $y'$.  
4447: One has $\mu_{y_1}/\mu_{y_2} = \frac{(y_2,x)^{n-1}}{(y_1,x)^{n-1}}$. So 
4448: $$
4449: \varphi(y_0, ..., y_n) = (n-1)^{n}\int_{\partial {\cal H}_n} \log\vert 
4450: \frac{(y_1,x)}{(y_0,x)}\vert  d  \log\vert 
4451: \frac{(y_2,x)}{(y_0,x)}\vert \wedge ... \wedge d  \log\vert 
4452: \frac{(y_n,x)}{(y_0,x)}\vert 
4453: $$
4454: Thus 
4455: $$
4456: \varphi_{y_0}(Y_1, ..., Y_n) = (n-1)^{n}\int_{\partial {\cal H}_n} (Y_1,x) d( Y_2, x) \wedge ... \wedge 
4457: d(Y_n, x)
4458: $$
4459: To do the computation of this integral we may suppose that $y_0 = (0, ..., 0, 1)$, $Y_i = 
4460: \frac{\partial}{\partial y_i}$, so $(Y_i, x) = x_i$. Then the last integral equals to 
4461: $$
4462: (n-1)^{n}\int_{S^{n-1}} x_1 dx_2 \wedge ... \wedge dx_n =  \frac{(n-1)^{n}}{n}{\rm vol}(S^{n-1})
4463: $$
4464: where $S^{n-1}$ is the sphere $x_1^2 + ... + x_n^2 =1$. 
4465: The lemma follows. 
4466: 
4467: 
4468: {\bf Proof of Theorem \ref{2112}}. Let us suppose first that the points $x_i$ 
4469: are inside of the hyperbolic space. 
4470: 
4471: The  function   $\varphi(x_0,...,x_{n})$  defines an 
4472: $n$-density $\widetilde \varphi$ on
4473:   $ {\cal H}_n$. 
4474: Namely,   to define the integral $\widetilde \varphi$  over a simplex $M$ one has to 
4475: subdivide it into small simplices and take the sum of the functions 
4476: $\varphi$ corresponding to their vertices.   
4477: When the simplices are getting smaller the limit exists and is by definition  
4478: $\int_M \widetilde \varphi$. Here we used  the properties 1)  - 3).
4479: More precisely 1) and 2)  imply that $\varphi$ defines an additive volume form on 
4480: ${\cal H}_n$, and 3) (together with 1)) guarantee that this volume form is $\sigma$-additive.
4481: 
4482: 
4483: The skewsymmetry property implies that  $\widetilde \varphi$ is actually a differential $n$-form. It is invariant under the action of 
4484: the group $SO(n,1)$. Therefore it is proportional to the standard volume form. 
4485: 
4486: Now suppose that the vertices $x_i$ can be  on the absolute. Then 
4487: it is easy to see that the corresponding integral 
4488: (\ref{i2})  is still convergent. Moreover, 
4489: if the vertices of the geodesic simplex are in general position 
4490: then it is a continuous function of the vertices. This  implies that  the volume of an 
4491: ideal geodesic simplex is finite (which is, of course, an elementary fact) and 
4492:  coincides with the corresponding integral (\ref{1221}).
4493: 
4494:   Completely similar results are valid for the complex $n$-dimensional   hyperbolic 
4495: space ${{\cal H}_n}^{\C}:= \{|z_1| + ... +|z_n|^2 < 1\}$, $z_i \in \C$, and the  
4496: quaternionic hyperbolic space ${{\cal H}_n}^{\HH}:= \{|q_1| + ... +|q_n|^2 < 1\}$ ($q_i$ are quaternions). Indeed, a point $x$ in each of these spaces defines an invariant volume form $\mu_x$ on the boundary. 
4497: 
4498: 
4499: 
4500: 
4501: 
4502: 
4503: 
4504: 
4505: {\bf 2. Calculation of the volume of a three dimensional ideal geodesic simplex}. 
4506: If $n=3$ the absolute can be identified with $\C \PP^1$, 
4507: and for the ideal simplex with vertices at 
4508: the points $\infty, 0,1,a$ on the absolute we get 
4509: $$
4510: {\rm vol} (I(\infty, 0,1,a)) = 3 c_3 \cdot 
4511: \int_{\C \PP^1} (\log|z| d\log|1-z| - \log|1-z| d\log|z|) \wedge d\log|z-a| =
4512: $$
4513: $$
4514: 3 c_3 \cdot \int_{\C \PP^1} (\log|z| d\arg (1-z) - \log|1-z| d\arg (z))\wedge d\arg (z-a)
4515: $$
4516: because $d \log (z-1) \wedge d \log (z-a) = 0$. 
4517: Here $\log f = \log |f| + i \arg(f)$. 
4518: Using 
4519: $$
4520: d {\cal L}_2(z) =  \log|z| d\arg (1-z)  - \log|1-z| d\arg (z)
4521: $$ 
4522: we rewrite the last integral as 
4523: \begin{equation} \label{1331} 
4524: 3 c_3 \cdot \int_{\C \PP^1} d {\cal L}_2(z) \wedge d\arg (z-a)
4525: \end{equation} 
4526: Computing 
4527: the differential in the sense of distributions we get
4528: $$
4529: d ({\cal L}_2(z) d\arg (z-a)) = 2\pi \cdot  {\cal L}_2(z) \delta(z-a)dx d y + 
4530: d {\cal L}_2(z)\wedge d\arg (z-a)
4531: $$
4532: So the integral of the right hand side over $\C \PP^1$ is zero, i.e. the integral (\ref{1331}) is equal to 
4533: $- 6 \pi c_3 \cdot {\cal L}_2(a)$. ( $c_3 = -1/6\pi$).
4534: 
4535:   
4536: 
4537: 
4538: 
4539: {\bf 3. Volumes of geodesic simplices in $SL_n(\C)/SU(n)$}. Recall 
4540: the invariant differential $(2n-1)$--form $\omega_{D_n}$ in ${\Bbb H}_n$.  
4541: 
4542: {\bf Question}. Is it true that  
4543: \begin{equation} \label{1221?q} 
4544: {\rm vol}_{\omega_{D_n}}(I(x_0,...,x_{2n-1})) = \mbox{\rm constant} 
4545: \times \psi_n(x_0,...,x_{2n-1})?
4546: \end{equation} 
4547: 
4548: 
4549: One can show following the lines of s. 7.1 that the positive answer 
4550: to this question is equivalent to the following statement: if $x_0, x_1, x_2$ 
4551: are on the same geodesic then $\psi_n (x_0,...,x_{2n-1}) =0$. 
4552: 
4553: 
4554: {\bf 4. Another approach to Grassmannian polylogarithms}. 
4555: The following construction was suggested to the author 
4556: during the Fall of 1989, independently, 
4557:  by M. Kontsevich and by J. Nekovar. 
4558: A hyperplane $h$ in an $n$--dimensional 
4559:  complex vector space $V$ determines an arrow in the space of degenerate 
4560: non-negative definite 
4561: hermitian forms in $V$ consisting of the forms with the kernel $h$. 
4562: Let $h_1, ..., h_{2n}$ 
4563: be hyperplanes in $V$. Let $C(h_1, ..., h_{2n})$ be projectivization of 
4564: the convex hull of the arrays corresponding to these
4565: hyperplanes. It is a simplex in $\overline {\Bbb H}_n$. The idea is to 
4566: integrate  
4567: the form $\omega_{D_n}$ over this simplex. 
4568: If $n=2$ this construction provides an ideal geodesic simplex 
4569: in the Cayley realization of the hyperbolic space, given by  
4570: the interior part of a ball in $\R \PP^3$.  However 
4571: the convergence of this integral for $n>2$ 
4572: has not been established yet, 
4573: although it does not seem to be a very difficult problem.  
4574: If the integral is convergent, we get a function on configurations 
4575: of $2n$ hyperplanes in $\CP^{n-1}$. 
4576: It would be very interesting to investigate this construction further and 
4577: compare it with our construction of the Grassmannian polylogarithms. 
4578: 
4579: {\bf 5. A  $(2n-1)$-cocycle  of $GL(\C)$}.
4580: Consider an infinite dimensional $\C$-vector space with a given basis 
4581: $e_1,...,e_m,...$. The group $GL(\C)$ is the 
4582: group of automorphisms of this space moving only finite number of basis
4583: vectors.  
4584: 
4585: Let us describe the restriction of the cocycle to the subgroup
4586: $GL_{n+m}(\C)$ acting on the subspace generated by first $n+m$
4587: vectors. 
4588: Take $2n$ elements $g_1,..., g_{2n}$ of this group. Consider the corresponding
4589: $2n$ $(m+1)$-tuples of vectors:
4590: $$
4591: g_1(e_{n},...,e_{n+m}), \quad  ... , \quad g_{2n}(e_{n},...,e_{n+m})
4592: $$
4593: The set of all  $(m+1)$--tuples of vectors form a vector space. 
4594: Let $\Delta_{2n-1}$ be the standard simplex $\sum^{2n}_{j=1} \lambda_j =
4595: 1$.
4596: Consider the set 
4597: $
4598: C_n^{m} \subset \Delta_{2n-1}
4599: $
4600: consisting of  $(\lambda_1,...,\lambda_{2n})$ such that 
4601: $$
4602: \sum_{i=1}^{2n} \lambda_i \cdot g_i(e_{n},...,e_{n+m})
4603: $$
4604: is an $(m+1)$-tuple of vectors {\it not} in generic position. It is a cycle of
4605: codimension $n$. The Chow polylogarithm function evaluated 
4606: on it provides the desired (measurable) cocycle. 
4607: 
4608: The cocycle property follows from the functional equation for the Chow
4609: polylogarithm and the following general fact. 
4610: The set of those $(\lambda_1,...,\lambda_{2n+1})$ such that 
4611: $
4612: \sum_i \lambda_i \cdot g_i(e_{n},...,e_{n+m})
4613: $
4614: is an $(m+1)$-tuple of vectors {\it not} in generic position is also of
4615: codimension $n$.
4616: 
4617: The construction is consistent with the restriction to $GL_n$ just by definition.
4618: 
4619: {\bf Problem}. Show that the cohomology class of this cocycle is nontrivial and proportional to the Borel class. 
4620: 
4621: 
4622: 
4623: 
4624: \begin{thebibliography}{BGSV}
4625: 
4626: 
4627: 
4628: \bibitem[A]{A} Arnold V.I.: Mathematical methods of classical mechanics.  
4629: Springer Verlag, 1978. 
4630: 
4631: \bibitem[B1]{B1} Beilinson A.A: {\it Higher regulators and values of 
4632: $L$-functions}, VINITI, 24 (1984), 181--238. 
4633: 
4634: \bibitem[B2]{B2} Beilinson A.A: {\it Height pairing between 
4635: algebraic cycles},  Springer Lect. Notes in Math. 1289. 
4636: 
4637: \bibitem[B3]{B3} Beilinson A.A: {\it Notes on absolute Hodge cohomology}. 
4638: Applications of algebraic $K$-theory to algebraic geometry and number theory, 
4639: Part I, II (Boulder, Colo., 1983),
4640:    35--68, Contemp. Math., 55, Amer. Math. Soc., Providence, RI, 1986. 
4641: 
4642: \bibitem[BD]{BD} Beilinson A.A., Deligne P: {\it Interpr\'etation motivique de la conjecture de Zagier} in Symp. in Pure Math., v. 55, part 2, 1994, 23-41.
4643: 
4644: 
4645: \bibitem[BMS]{BMS} Beilinson A.A., MacPherson R., Schechtman, 
4646: V.V: {\it Notes on motivic cohomology}, Duke Math.\ J.\ 54 (1987), 
4647: 679--710. 
4648: 
4649: 
4650: 
4651: \bibitem[Bl]{Bl} Bloch S: {\it Algebraic cycles and higher K-theory}, Adv. in
4652: Math., (1986), v. 61, 267--304.
4653: 
4654: \bibitem[Bl2]{Bl2} Bloch S: {\it The moving lemma for higher Chow groups},
4655:  J. Alg. geometry. 3 (1994).
4656: 
4657: \bibitem[Bl3]{Bl3} Bloch S: {\it Algebraic cycles and the 
4658: Beilinson conjectures}, Cont. Math. 58 (1986) no 1, 65-79.
4659: 
4660: \bibitem[BK]{BK} Bloch S. Kriz I: {\it Mixed Tate motives}, 
4661: Ann. of Math., 140 (1994) vol. 3, 557-605.  
4662: 1992. 
4663: 
4664: \bibitem[Bo1]{Bo1} Borel A: {\it Stable real cohomology of arithmetic groups}, Ann. Sci. Ec. Norm. Super., (4) 7 (1974), 235--272
4665: \
4666: \bibitem[Bo2]{Bo2} Borel A: {\it Cohomologie de $SL_{n}$ et valeurs de 
4667: fonctions z\^{e}ta aux points entiers},  Ann. Sc. Norm. 
4668: Sup. Pisa 4 (1977), 613--636.
4669: 
4670: \bibitem[Bott]{Bott} Bott R.: {\it On the characteristic classes of groups
4671: of diffeomorphisms}, L'ensiegnement Mathematique, T. XXIII, fsc. 3-4
4672: (1977) p. 209-220.
4673: 
4674:  \bibitem[Bu]{Bu} Burgos, J. I.: {\it 
4675: Arithmetic Chow rings and Deligne-Beilinson cohomology},
4676:  J. Algebraic Geom. 6 (1997), no. 2, 335--377. 
4677: 
4678: \bibitem[Con]{Con} Consani C.: {\it Double complexes and Euler's L-factors},  
4679: Compositio Math. 111 (1998) 323-358. 
4680: 
4681: \bibitem[D1]{D1} Dupont J.: {\it Simplicial De Rham cohomolgy and
4682: characteristic classes of flat bundles}, Topology 15, 1976, p. 233-245.
4683: 
4684:  \bibitem[Del]{Del} Deligne, P. {\it Le d\'eterminant de la cohomologie}. 
4685: Current trends in arithmetical algebraic geometry (Arcata, Calif., 1985),
4686:    93--177, Contemp. Math., 67, Amer. Math. Soc., Providence, RI, 1987.
4687: 
4688: \bibitem[F]{F} Fulton W.:. {\it Intersection theorey}. Second edition. Springer 1998. 
4689: 
4690: \bibitem[Dyn]{Dyn} Dynkin E.B.: {\it Topological characteristics of compact Lie groups homomorphisms} Math. Sb. NS 35 (1954), 129-173
4691: 
4692: 
4693: 
4694: \bibitem[GGL]{GGL} Gabrielov A., Gelfand I.M., Losik M.: 
4695: {\it Combinatorial computation of characteristic classes} I, II,
4696: Funct. Anal. i ego pril. 9 (1975) 12-28, ibid 9 (1975) 5-26.
4697: 
4698: %\bibitem[Gel]{Gel}Gelfand I.M.: {\it General theory of hypergeometric
4699: %functions}. Doklady Acad. Nauk USSR, 1986, 288, N 1,   p.14-18.
4700: 
4701: \bibitem[GM]{GM} Gelfand I.M., MacPherson R: {\it Geometry in 
4702: Grassmannians and a generalisation of the dilogarithm}, Adv. 
4703: in Math., 44 (1982) 279--312. 
4704: 
4705:  \bibitem[GS]{GS} Gillet, H.; SoulŽé, Ch.:  
4706: {\it Arithmetic intersection theory}, 
4707: Inst. Hautes ŽÉtudes Sci. Publ. Math. No. 72, (1990), 93--174 (1991). 
4708: 
4709: \bibitem[G1]{G1} Goncharov A.B: {\it Geometry of configurations, 
4710: polylogarithms and motivic cohomology},  
4711: Adv. in Math., 114, (1995), 197--318. 
4712: 
4713: \bibitem[G2]{G2} Goncharov A.B.: {\it Polylogarithms and motivic Galois
4714: groups} in Symp. in Pure Math., v. 55, part 2, 1994, 43 - 97.
4715: 
4716: \bibitem[G3]{G3} Goncharov A.B.: {\it Explicit construction of 
4717: characteristic classes}, I.M. Gelfand seminar 
4718: Adv. Soviet Math., 1993, 
4719: v. 16, p. 169-210. 
4720: 
4721: \bibitem[G4]{G4}
4722: Goncharov, A.B.: {\it Deninger's conjecture on special values of
4723: $L$-functions of elliptic curves at $s=3$},  J. Math. Sci. 81 (1996), N3,  2631-2656,  
4724: alg-geom/9512016. 
4725: 
4726: 
4727: \bibitem[G5]{G5} Goncharov A.B.: {\it Chow polylogarithms and
4728: regulators}. Math. Res. Letters, 2, (1995), 99-114.
4729: 
4730: 
4731: 
4732: 
4733: \bibitem[G6]{G6}
4734: Goncharov, A.B.: {\it Geometry of trilogarithm and the motivic Lie 
4735: algebra of a field}, Regulators in analysis, geometry and number theory, 
4736: 127--165, Progr. Math., 171,
4737:    Birkhauser Boston, Boston, MA, 2000. math.AG/0011168.
4738: 
4739: \bibitem[G7]{G7}
4740: Goncharov, A.B.: {\it Explicit regulator maps on the polylogarithmic 
4741: motivic complexes}, Motives, polylogarithms and Hodge theory, Part I (Irvine, CA, 1998), 245--276, Int.
4742:    Press Lect. Ser., 3, I, Int. Press, Somerville, MA, 2002. 
4743: math.AG/0003086.
4744: 
4745: \bibitem[GZ]{GZ} Goncharov A. B., Zhao J.: {\it The Grassmannian trilogarithms}. 
4746: Compositio Math. 127 (2001), no. 1, 83--108.  math.AG/0011165.
4747: 
4748: 
4749: \bibitem[HM]{HM} Hain R, MacPherson R: {\it Higher Logarithms}, 
4750: Ill. J.\ of Math,, vol. 34, (1990) N2,  392--475. 
4751: 
4752: \bibitem[H]{H} Hain R: {\it The existence of higher logarithms}. Compositio Math. 100 (1996) N3, 247- 276 
4753: 
4754: 
4755: \bibitem[HY]{HY} Hain R, Yang R.: {\it Real Grassmann polylogarithms and Chern classes}. math. Annalen, 304, (1996), 
4756: N1, 157-201. 
4757: 
4758: \bibitem[HM1]{HM1} Hanamura M, MacPherson R: {\it Geometric construction
4759: of polylogarithms}, Duke Math. J. 70 ( 1993)481-516. 
4760: 
4761: 
4762: 
4763: \bibitem[HM2]{HM2} Hanamura M,  MacPherson R.: 
4764: {\it Geometric construction of polylogarithms, II}, 
4765: Progress in Math. vol. 132 (1996) 215-282. 
4766: 
4767: \bibitem[Lei]{Lei} Gerhardt C.I. (ed). G.W.Leibniz Mathematische
4768: Schriften III/1 pp. 336-339 Georg Olms Verlag, Hildesheim. New York
4769: 1971. 
4770: 
4771: \bibitem[Le]{Le} Levin A.M.: {\it Notes on $\R$-Hodge-Tate sheaves}. 
4772: Preprint MPI 2001. 
4773: \bibitem[Lev]{Lev} Levine, M.: {\it Bloch's higher Chow groups revisited. $K$-theory}. Astérisque No. 226 (1994), 10, 235--320.
4774: 
4775: \bibitem[Li]{Li} Lichtenbaum S: {\it Values of zeta functions at  
4776: non-negative integers}, Lect. Notes in Math., 1068, Springer Verlag,
4777: 1984, 127-138. 
4778: 
4779: \bibitem[M]{M} MacPherson R.: {\it Gabrielov, Gelfand, Losic
4780: combinatorial formula for the first Pontryagin class} Seminar Bourbaki 1976
4781: 
4782: \bibitem[NS]{NS} Nesterenko Yu.P., Suslin A.A.: {\it 
4783: Homology of the general linear group over a local ring, and 
4784: Milnor's $K$-theory}. 
4785: Math. USSR-Izv. 34 (1990), no. 1, 121--142.
4786: 
4787: \bibitem[S]{S} Soul\'e C.,  D. Abramovich, J. F. Burnol, J. K. Kramer: 
4788: Lectures on Arakelov geometry.  Cambridge University Press, 
4789: 1992. 
4790: 
4791: \bibitem[Su]{Su} Suslin A.A: {\it Homology of $GL_{n}$,
4792: characteristic classes and Milnor's $K$-theory}.  Lect. Notes in Math.
4793: 1046 (1989), 357--375.
4794: 
4795: \bibitem[Z1]{Z1} Zagier D: {\it Polylogarithms, Dedekind zeta 
4796: functions and the algebraic $K$-theory of fields}, Arithmetic         
4797:  Algebraic Geometry (G.v.d.Geer, F.Oort, J.Steenbrink, eds.),        
4798:   Prog. Math., Vol 89, Birkhauser, Boston, 1991, pp. 391--430. 
4799: 
4800: \end{thebibliography}
4801: 
4802: 
4803:  
4804: 
4805: 
4806: 
4807: 
4808: 
4809: 
4810: 
4811: \bigskip
4812: 
4813: 
4814: 
4815: 
4816: Department of Mathematics, Brown University, Providence, RI 02912, USA. 
4817: 
4818: e-mail sasha@math.brown.edu
4819: 
4820: 
4821: 
4822: 
4823: 
4824: 
4825: 
4826: 
4827: 
4828: 
4829: \end{document}
4830: 
4831: 
4832: 
4833: 
4834: 
4835: 
4836: {\bf 8. Cocycles for continuous cohomology classes of  $GL_N(\C)$}. 
4837: {\it A  $(2n-1)$-cocycle of $GL_n(\C)$}. 
4838: Choose a  hyperplane $h$ in $\C \PP^{n-1}$. Then the
4839: function 
4840: \begin{equation} \label{c10}
4841: c^{n}_{2n-1}(g_1,...,g_{2n}):= {\cal P}_n(g_1 h, ... ,g_{2n} h)
4842: \end{equation}
4843: is a (2n-1)-cocycle of $GL_n(\C)$. The cocycle
4844: condition is just the property c) in Theorem (\ref{0.3a}). Notice that
4845: the cocycle $c^{n}_{2n-1}(g_1,...,g_{2n})$ 
4846: is defined everywhere on 
4847: $GL_n(\C)^{2n}$. 
4848: It is, however,  discontinuous near the identity and only a measurable
4849: function on $GL_n(\C)^{2n}$. 
4850: 
4851: This cocycle essentially coincides with the cocycle
4852: described in the s.5 of the Introduction, see  Remark in s.2 below. 
4853: In particular it represents (a multiple of) the Borel class.
4854: 
4855:  
4856:  
4857: {\it A $(2n-1)$-cocycle for the Borel class of $GL_{n+m}(\C)$}.
4858: The property b) in Theorem (\ref{0.3a}) guarantee that this cocycle can
4859: be extended to the group $GL(\C)$. Namely, to do this for the group
4860: $GL_{n+m}(\C)$ we should proceed as follows.
4861: 
4862: The projective duality provides a bijection
4863: $$
4864: \left \{ \mbox{ Configurations of 2n hyperplanes in } \C \PP^{n-1} \right \} \quad  <--> \quad $$
4865: $$
4866: \left \{ \mbox{ Configurations of 2n  points in } \C \PP^{n-1} \right \}
4867: $$
4868:  So we can consider the dual function $\widehat  {\cal P}^G_n$
4869: on configurations of $2n$ points in $\C \PP^{n-1}$. By definition its
4870: value  value at a 
4871: configuration of points  is equal to the value of ${\cal P}_n$
4872: on the projectively dual configuration of hyperplanes in $\C \PP^{n-1}$
4873: 
4874: {\bf Remark}.  The configuration of $2n$  
4875: vectors corresponding to an $(n-1)$- plane $h$ according to Lemma 1.1 
4876: is different from (dual to) the one constructed above. Nevertheless 
4877: the function $\widetilde  {\cal P}_n$ coincides with the function 
4878: $\widehat  {\cal P}_n$ constructed in s.5 of the Introduction. 
4879: 
4880: 
4881: Let us call by $m$-flag in $\C \PP^{k}$ a sequence of subspaces
4882: $L_{\bullet}:= L_0 \subset L_1 \subset ... \subset L_{m-1}$ where $dim
4883: L_i = i$.  
4884: 
4885: Let  $H_1\ast H_2$ be the joining of planes $H_1, H_2$. In general
4886: $dim(H_1 \ast H_2) = dimH_1 + dimH_2 -1$.  
4887: 
4888: For $2n$ generic $(m+1)$-flags $L_{\bullet}^{(1)}, ... , L_{\bullet}^{(2n)}$ in $\C \PP^{n+m-1}$ set 
4889: $$
4890: \widetilde  {\cal P}_n^{(m)}(L_{\bullet}^{(1)}, ... , L_{\bullet}^{(2n)}):=
4891: \sum_{j_1+...+j_{2n} = m} \widetilde  {\cal P}_n\Bigl((L_{j_1-1}^{(1)}\ast ... \ast L_{j_{2n}-1}^{(2n)}\vert L_{j_1}^{(1)}, ... , L_{j_{2n}}^{(2n)})\Bigr)
4892: $$
4893: Here $(L_{j_1-1}^{(1)}\ast ... \ast L_{j_{2n}-1}^{(2n)}\vert L_{j_1}^{(1)}, ... , L_{j_{2n}}^{(2n)})$
4894: is the configuration of $2n$ points in $\C \PP^{n-1}$ obtained by the
4895: projection of $L_{j_k}^{(k)}$ with the center at $L_{j_1-1}^{(1)}\ast
4896: ... \ast L_{j_{2n}-1}^{(2n)}$ (see fig 2). More precisely, the set of
4897: all planes of 
4898: dimension $j_1+...+j_{2n}$ containing $L_{j_1-1}^{(1)}\ast
4899: ... \ast L_{j_{2n}-1}^{(2n)}$ is a projective space of dimension $n-1$.
4900: Each $L_{j_k}^{(k)}$ provides a point in this space.
4901: 
4902: Choose an $(m+1)$-flag $L_{\bullet}$ in $\C \PP^{n+m-1}$. Set
4903: \begin{equation} \label{8as}
4904:  c_{2n-1}^{n+m}(g_1,...,g_{2n}):= \widetilde  {\cal P}_n^{(m)}(g_1
4905: L_{\bullet}^{(1)}, ... , g_{2n} L_{\bullet}^{(2n)})
4906: \end{equation}
4907: This function is defined only on a Zariski open subset of $GL_{n+m}(\C)^{2n}$  for which the flags in
4908: the right hand side of (\ref{8as}) are in generic position. 
4909: 
4910: 
4911: \begin {Theorem}  
4912: \label {5.5}
4913: a) $c_{2n-1}^{n+m}(g_1,...,g_{2n})$ is a measurable cocycle of the Lie
4914: group $GL_{n+m}(\C)$.
4915: 
4916: b) The   cohomology class of   $(2 \pi i)^n c^{n+m}_{2n-1}$ in
4917:  is a non
4918: zero rational multiple of the Borel class. 
4919: \end{Theorem}
4920:   
4921: (The measurable and continuous cohomology of a Lie group $G$ are
4922:  isomorphic).
4923: 
4924: 
4925: {\bf Proof}. a) follows from the Key Lemma  in s.2.1 in[G3] and  b) from
4926: Theorem 5.12 in [G3]. The existence of the bi-Grassmannian $n$-logarithm
4927: is the main ingredient of the proof. 
4928: 
4929: 
4930: 
4931:  {\bf 9. A sketch of  a direct proof of Theorem (\ref{2112})}. 
4932: Let $Q $ be a nondegenerate quadric in $\PP^n$ and  $L$ 
4933: be a nondegenerate simplex in $\PP^n$. 
4934: Assume that $L$ and $Q$ are in generic position.
4935: 
4936:   
4937: Let $\widetilde Q(x)$ be a quadratic 
4938:  equation of the quadric $Q$ in homogeneous coordinates. Set 
4939: $$
4940: \omega_Q := \frac{{\sqrt det (\widetilde Q)}}{(2\pi i)^n}\frac{\sigma_n(x,dx)}{\widetilde Q(x)^n}
4941: $$
4942: It does not depend on the choice of $\widetilde   Q(x)$.  Let $\omega_L$ be the canonical differential form with logarithmic singularities at
4943: $L$ with integer periods. 
4944: 
4945: Let $\Delta_{L}$ be an $n$-chain (or an $n$-membrane) in ${\Bbb C} \PP^n$ such that $[\Delta_{L}]$ is a generator of $H_n({\Bbb C} \PP^n,L;\Z)$. 
4946: 
4947: Similarly let $\Delta_{Q}$ be an $n$-chain in ${\Bbb C} \PP^n$ such that $[\Delta_{Q}]$ is a generator of  $H_n({\Bbb C} \PP^n,Q;\Z)$ corresponding to the primitive class under the isomorphism
4948: $$
4949: H_{n}({\Bbb C} \PP^{n},Q;\Z) = Ker \Bigl(H_{n-1}(Q;\Z) 
4950: \longrightarrow H_{n-1}({\Bbb C} \PP^{n};\Z)\Bigr)
4951: $$ 
4952:  The choice of a generator corresponds to the choice of the sign in the definition of the form $\omega_Q$.  
4953: 
4954:  
4955:  
4956: Let $\widehat Q$ be the quadric in
4957: $\widehat \PP^n$ dual to $Q$ and   $\widehat L$ be the dual  simplex in $\widehat \PP^n$. The vertices of $\widehat L$ correspond to the codimension one faces of $L$.
4958:  
4959:    
4960: \begin{theorem}
4961:  a)
4962: $
4963: \int_{\Delta_{ \widehat Q}}\omega_D =  \omega_Q,  \qquad \int_{\Delta_{ \widehat L}}\omega_D =  \omega_L
4964: $
4965: 
4966: b)
4967: $
4968: \int_{\Delta_{L \times Q}}\omega_D = \int_{\Delta_{\widehat L}}\omega_Q = \int_{\Delta_{ \widehat  Q}}\omega_L 
4969: $
4970: \end{theorem}
4971:  
4972: 
4973: {\bf Proof}
4974:   b)  follows  from a).  
4975: 
4976: a)Now consider a sphere $S^{n-1}$ in $\R \PP ^n$. It cuts the projective space on two domains, $
4977: {\cal B}_+ \cup {\cal B}_-$ one of which is convex. Consider a simplex
4978:  $L$ inside of the convex ball. Let us think about the dual projective space as of the set of all hyperplanes. Then the hyperplanes which do not intersect the convex ball form a convex domain in $\widehat \R \PP ^n$, which we will denote $\widehat 
4979: {\cal B}_+$. 
4980: 
4981: The simplex $L$ defines a dual simplex $\widehat L$ in $\widehat \R \PP^n$. Namely, each vertex $l_i$ of $L$ produces a face of $\widehat L_i$ of the dual simplex, namely 
4982: $\widehat L_i$ consists of all hyperplanes passing through the vertex $L_i$ which do not intersect the interior part of $L$. Therefore the dual simplex $\widehat L$ lies outside of the dual ball $\widehat {\cal B}_+$
4983: 
4984: The cycle
4985: $$
4986: L \times \widehat {\cal B}_+ \subset \R \PP ^n \times \widehat  {\R \PP^n}
4987: $$
4988: does not intersect the incidence variety $\Delta \subset \R \PP^n \times \widehat  {\R  
4989: \PP^n}$.  In particular the integral of our form 
4990:  over this cycle converges.
4991:  $\int_{L} \omega_D = \omega_{\widehat L}$.  Since  the dual simplex $\widehat L$ does not intersect the interior part of the dual ball, the restriction of the form $\omega_{\widehat L}$ to the dual ball is exact. So we can use the Stokes theorem to reduce the calculation of the integral of $\omega_{\widehat   L}$ over the dual ball to the integral over the sphere.
4992: 
4993: 
4994: {\bf Second proof}. We will suppose that $S$ is rational and connected.
4995: The starting point is the  Lemma \ref{2.14}
4996: 
4997: Let $s$ and $t$ be two (generic) points 
4998: of $S$. Choose a path $\gamma$ from $s$ to $t$. Then
4999: $$
5000: {\cal P}_2(Y^s;f^s_1,f^s_2,f^s_3) - {\cal P}_2(Y^t;f^t_1,f^t_2,f^t_3) = \int_{\gamma}
5001: {\rm Alt}_3 \Bigl(v_x(f_1) \log|f_2(x)| d_s \arg f_3(x) \Bigr)
5002: $$
5003: Let us prove that the right hand side is expressed by ${\cal L}_2$. The integrand on the right is supported on a divisor $D \subset Y$, see proof of Lemma \ref{2.14}. Let us apply 
5004:  the transfer map
5005: $K_2(\C(D))  \longrightarrow K_2(\C(S))$ to the element $Res(f_1 \wedge f_2 \wedge f_3 )$. Let $D_1$ be an irreducible component of $D$. Denote by $\gamma_1$ the preimage of the path $\gamma$ in $D_1$.
5006: The same argumentation as in the first proof  shows that we can assume that $ \C(D_1)$ is a Galois extension of $\C(S) $ and reduce the problem to the calculation of two integrals: $\int_{\gamma'} r_1(1-f,f)$ and $\int_{\gamma} r_1(f,g)$.  Noting that 
5007: $r_1(1-f,f) = d{\cal L}_2(f)$,
5008:  we compute the first integral.  Computing the second   we may suppose that $\gamma$ lies in  $\C  {\Bbb P}^1 $. Then the problem reduces to the calculation of the integral $\int_0^z r_1(z-a,z-b)$. Writing
5009: $$
5010: (z-a) \wedge (z-b) = \left(1-\frac{(z-b)}{(a-b)}\right) \wedge \frac{(z-a)}{(a-b)}\quad \mbox{modulo} \quad \C^* \wedge \C(z)^*
5011: $$ and using that $\int_0^z r_1(c,f(z))=0$ for a constant $c$ we reduce the problem to calculation of the first integral. Theorem is proved. 
5012: 
5013: 
5014: 
5015: c) The special values of $L$-functions 
5016: (Conjectures of Birch-Swinnerton-Dyer, Birch-Tate and Lichtenbaum and Deligne, the theorems of Borel, the regulators of Bloch, Beilinson  and philosophy of mixed motives. 
5017: 
5018: 
5019: