1: \documentclass[a4paper]{article}
2: \usepackage{amssymb,amsbsy,amsmath,amsfonts,amssymb,amscd}
3: \usepackage{latexsym}
4: \usepackage{geometry}
5: %\usepackage{showkeys}
6: % \usepackage{pdfsync}
7: \input{macros.tex}
8:
9: \input xy
10: \xyoption{all}
11:
12:
13: \newcommand{\bysame}{\hspace{1.5cm}}
14: \newcommand{\ra}{\rightarrow}
15: \renewcommand{\H}{{\mathrm H}}
16: \newcommand{\cD}{{\cal D}}
17: \newcommand{\cE}{{\cal E}}
18: \newcommand{\cF}{{\cal F}}
19: \newcommand{\cG}{{\cal G}}
20: \newcommand{\cK}{{\cal K}}
21: \newcommand{\cS}{{\cal S}}
22: \newcommand{\cT}{{\cal T}}
23: \newcommand{\cN}{{\cal N}}
24: \newcommand{\cH}{{\cal H}}
25: \newcommand{\cR}{{\cal R}}
26:
27: \newcommand{\tGam}{{\tilde{\Gamma}}}
28: \newcommand{\tA}{{\tilde{A}}}
29: \newcommand{\tF}{{\tilde{F}}}
30:
31: \newcommand{\Ga}{{\bb G}_a}
32: \newcommand{\Gm}{{\bb G}_m}
33:
34:
35: % Fritz compatibility macros
36:
37: \def\normalbaselines{\baselineskip20pt\lineskip3pt\lineskiplimit3pt }
38: \def\mapright#1{\smash{\mathop{\longrightarrow}\limits^{#1}}}
39: \def\mapdown#1{\Big\downarrow\rlap{$\vcenter{\hbox{$\scriptstyle#1$}}$}}
40:
41:
42: \def\g{{\Gamma}}
43: \def\a{{\Aut}}
44: \hsize= 15cm
45: \vsize= 20.5cm
46: \font\Bbb=msbm10
47: \font\Bbbb=msbm7
48: \def\I{{\bf I}}
49: \def\BZ{{\hbox{\Bbb Z}}}
50: \def\BL{{\hbox{\Bbb L}}}
51: \def\BH{{\hbox{\Bbb H}}}
52: \def\BK{{\hbox{\Bbb K}}}
53: \def\BP{{\hbox{\Bbb P}}}
54: \def\BF{{\hbox{\Bbb F}}}
55: \def\BG{{\hbox{\Bbb G}}}
56: \def\BN{\hbox{\Bbb N}}
57: \def\BR{{\hbox{\Bbb R}}}
58: \def\BC{{\hbox{\Bbb C}}}
59: \def\BQ{{\hbox{\Bbb Q}}}
60: \def\BT{\hbox{\Bbb T}}
61: \def\A{{\hbox{\rm Aut}}}
62: \def\Ro{{\bar R_{\rm old}}}
63: \def\A{{\rm Aut}}
64:
65: \def\op{{\rm op}}
66: \def\R{{\bf R}}
67: \def\Ad{{\rm Ad}}
68: \def\g{{\Gamma}}
69: \def\u{{{\mathsf R}_{\rm u}(G)}}
70: \def\s{{{\mathsf R}_{\rm u}(G_1)}}
71: \def\t{{{\mathsf R}_{\rm u}(G_2)}}
72:
73: \baselineskip= 13pt
74: \parindent= 0pt
75: \parskip= 3pt plus 1pt minus 0.5pt
76:
77: \def \Box{\lower .1 em
78: \vbox{\hrule \hbox{\vrule \hskip .6 em \vrule height .6 em} \hrule}}
79: \def \Mid{\quad \vrule \quad}
80: \font \msbm= msbm10
81: \def \rfish{\mathbin {\hbox {\msbm \char'157}}}
82:
83: % End of Fritz compatibility
84:
85:
86:
87: \title{ Linear Representations of the Automorphism Group of a
88: Free Group}
89:
90: \author{Fritz Grunewald\\
91: Mathematisches Institut \\
92: Heinrich-Heine-Universit\"at\\
93: D-40225 D\"usseldorf\\
94: e-mail: fritz@math.uni-duesseldorf.de
95: \and
96: Alexander Lubotzky\\
97: Einstein Institute of Mathematics\\
98: The Hebrew University of Jerusalem\\
99: Jerusalem, 91904, Israel\\
100: e-mail: alexlub@math.huji.ac.il}
101:
102:
103:
104: \begin{document}
105: \maketitle
106:
107:
108: \begin{abstract}
109: Let $F_n$ be the free group on $n\ge 2$ elements and $\A(F_n)$ its group of
110: automorphisms. In this paper we present a rich collection of
111: linear representations of
112: $\A(F_n)$ arising through the action of finite index subgroups of it on
113: relation modules of finite quotient groups of $F_n$. We show
114: (under certain conditions) that the images
115: of our representations are arithmetic groups.
116: \newline
117:
118: \noindent
119: 2000 Mathematics Subject Classification: Primary 20F28, 20E05; \\
120: Secondary 20E36, 20F34, 20G05
121: \end{abstract}
122:
123: \medskip
124:
125:
126: \bigskip
127: \bigskip
128:
129:
130: \tableofcontents
131: \medskip
132:
133: \bigskip
134: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
135: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
136:
137:
138:
139: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
140: \section{Introduction}
141: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
142:
143: Let $F_n$ be the free group on $n\ge 2$ elements and $\A(F_n)$ its group of
144: automorphisms. The latter is a much studied group, but very little seems to be
145: known about its (finite dimensional) complex representation theory
146: (see for example \cite{FP}, \cite{DJ}, \cite{PORA}, \cite{RA}, \cite{LP}).
147: In fact,
148: as far as we know the only representations studied (at least when $n\ge 3$)
149: are:
150: \begin{itemize}
151: \item The representation
152: \begin{equation}\label{furep}
153: \rho_1 : \A(F_n)\to \A(F_n/F'_n)\cong \GL(n,\BZ)
154: \end{equation}
155: where $F_n'$ is the commutator subgroup of $F_n$.
156: \item The representations factoring through the homomorphisms
157: $$\rho_i : \A(F_n)\to \A(F_n/F^{(i+1)}_n)$$
158: where $F^{(i+1)}_n$ ($i=0,1,\ldots$) stands for the lower
159: central series of $F_n$.
160: \item Representations with finite image.
161: \end{itemize}
162: What is common to all of the above mentioned reprentations $\rho$ is that if we
163: denote by ${\cal H}:=\overline{\rho(\A(F_n))}$ the Zariski closure of
164: $\rho(\A(F_n))$, then the semi-simple part of the connected component
165: ${\cal H}^\circ$ of ${\cal H}$ is either trivial or $\SL(n,\BC)$. Moreover,
166: $\rho({\rm IA}(F_n))$ is always virtually solvable for these representations,
167: where ${\rm IA}(F_n)$ is the kernel of $\rho_1$.
168:
169: In this paper we will show that the representation theory of $\A(F_n)$ is in
170: fact much richer. Our main Theorem (Theorem \ref{teo} below)
171: implies for example
172: \begin{theorem}\label{teo4}
173: Let $n\ge 2$, $k\ge 1$, $h_1<\ldots <h_k$, $m_1,\ldots ,m_k$ be
174: natural numbers. Let $\BQ(\zeta_{m_i})$ be the field of
175: $m_i$-th roots of unity and
176: $\BZ(\zeta_{m_i})$ its ring of integers. There is a
177: subgroup $\Gamma\le \A(F_n)$ of finite index and a representation
178: $$
179: \rho : \Gamma \to \prod_{i=1}^k\SL((n-1)h_i,\BQ(\zeta_{m_i}))^{m_i}
180: $$
181: such that $\rho(\Gamma)$ is commensurable with
182: $\prod_{i=1}^k\SL((n-1)h_i,\BZ(\zeta_{m_i}))^{m_i}$.
183: \end{theorem}
184: Theorem \ref{teo4} shows that $\A(F_n)$ has a representation $\rho$ such that
185: the connected component of $\overline{\rho(\A(F_n))}$ is isomorphic to
186: $$\prod_{i=1}^k\SL((n-1)h_i,\BC)^{m_i}$$
187: as above. In fact the proof shows that this is even true for
188: $\overline{\rho({\rm IA}(F_n))}$ and in particular the latter
189: is very far from being virtually solvable (see Section \ref{Concl1} for more).
190:
191: Specifying $m_1=\ldots =m_k=1$ in Theorem \ref{teo4}, we deduce
192: \begin{corollary}\label{coro4}
193: Let $n\ge 2$, $k\ge 1$ be natural numbers. There is a
194: subgroup $\Gamma\le \A(F_n)$ of finite index and a representation
195: $$
196: \rho : \Gamma \to \prod_{i=1}^k\SL((n-1)i,\BZ)
197: $$
198: such that $\rho(\Gamma)$ is of finite index in $\prod_{i=1}^k\SL((n-1)i,\BZ)$.
199: \end{corollary}
200: Specialising Theorem \ref{teo4} even further to
201: $n=3$, $k=1$, $h_1=m_1=1$ we get that
202: $\A(F_3)$ has a subgroup of finite index which can be mapped onto a subgroup
203: of finite index in $\SL(2,\BZ)$. This implies
204: \begin{corollary}\label{kaz}
205: The automorphism group $\A(F_3)$ is large, that is it has a subgroup of finite
206: index which can be mapped onto a free nonabelian group. In particular
207: $\A(F_3)$ does not have Kazdhan's property (T).
208: \end{corollary}
209: The corollary is easy for $\A(F_2)$, see the discussion in Section
210: \ref{Proofs}, but it is a
211: well known open problem for larger $n$. Our solution for
212: $n=3$ does not indicate what should be the answer for $n>3$.
213:
214: We shall describe now how the representations to be considered here arise.
215: Let $G$ be a finite group and $\pi : F_n\to G$ a surjective homomorphism
216: of the free group $F_n$ onto $G$. Let $R$ be the kernel of
217: $\pi$. We define
218: \begin{equation}
219: \Gamma(R):=\{\, \varphi\in \A(F_n)\Mid \varphi(R)=R \, \}
220: \end{equation}
221: and
222: \begin{equation}\label{I1}
223: \Gamma(G,\pi):=\{\, \varphi\in \Gamma(R) \Mid
224: \varphi\ {\rm induces\ the\ identity\ on}\ F_n/R \, \}.
225: \end{equation}
226: These are subgroups of finite index in $\A(F_n)$. We let
227: \begin{equation}
228: \bar R:=R/R'
229: \end{equation}
230: be the relation module of the presentation $\pi : F_n\to G$.
231: The action of $F_n$ on $R$ by conjugation leads
232: to an action of $G$ on $\bar R$.
233:
234: The structure of $\bar R$ as a $G$-module is described by
235: Gasch\"utz' theory (see Section \ref{suga} and the references therein).
236: It says in particular that
237: \begin{equation}\label{G1}
238: \BQ\otimes_{\bb Z} \bar R\cong \BQ\oplus \BQ[G]^{n-1}
239: \end{equation}
240: as a module over the rational group ring $\BQ[G]$ (see also (\ref{gasch})).
241: Our work can be described as an equivariant Gasch\"utz' theory where we try to
242: describe the action of $\Gamma(G,\pi)$ on $\bar R$: Every $\varphi\in
243: \Gamma(R)$ induces a linear automorphism $\bar \varphi$
244: of $\bar R$ and (as an application of the theorem of Gasch\"utz)
245: $\Gamma(G,\pi)$ consists exactly of those elements $\varphi\in\Gamma(R)$
246: for which $\bar\varphi$ is $G$-equivariant.
247: Already now we can see from (\ref{G1}) why the number $n-1$ plays such an
248: important role in Theorem \ref{teo4}.
249:
250: To be more precise:
251: The relation module $\bar R$ is a finitely generated
252: free abelian group, let $t$ be its $\BZ$-rank. Then
253: \begin{equation}\label{I4}
254: {\cal G}_{G,\pi}:=\Aut_G(\BC\otimes_{\bb Z} \bar R)\le \GL(t,\BC)
255: \end{equation}
256: is a $\BQ$-defined algebraic subgroup of $\GL(t,\BC)$, in fact
257: ${\cal G}_{G,\pi}$ is the centraliser of the group $G$ acting on
258: $\BC\otimes_\BZ \bar R$ through matrices with rational entries. Let
259: \begin{equation}
260: {\cal G}_{G,\pi}^1 \le \SL(t,\BC)
261: \end{equation}
262: be the kernel of all $\BQ$-defined homomorphisms from the complex algebraic
263: group $\BC\otimes{\cal G}_{G,\pi}$ to the multiplicative group. This is a
264: $\BQ$-defined subgroup of ${\cal G}_{G,\pi}$. We shall describe it in more
265: detail in Section \ref{suga}.
266: We define
267: \begin{equation}
268: {\cal G}_{G,\pi}(\BZ):=\{\, \phi\in {\cal G}_{G,\pi}\Mid \phi(\bar
269: R)=\bar R\,\}.
270: \end{equation}
271: This is an arithmetic subgroup of the $\BQ$-defined algebraic group
272: ${\cal G}_{G,\pi}$ which contains the arithmetic subgroup
273: \begin{equation}
274: {\cal G}^1_{G,\pi}(\BZ):=\{\, \phi\in {\cal G}^1_{G,\pi}\Mid \phi(\bar
275: R)=\bar R\,\}
276: \end{equation}
277: of ${\cal G}^1_{G,\pi}$.
278: Following the definitions we obtain an
279: integral linear representation
280: \begin{equation}\label{I5}
281: \rho_{G,\pi} : \Gamma(G,\pi)\to {\cal G}_{G,\pi}(\BZ),\qquad
282: \rho_{G,\pi}(\varphi)=\bar\varphi\quad (\varphi\in \Gamma(G,\pi)).
283: \end{equation}
284: Examples show (see Section \ref{Choose}) that in general $\rho_{G,\pi}$ is not
285: onto, neither is its image of finite index in ${\cal G}_{R,\pi}(\BZ)$.
286: However, our main result shows that, at least for redundant presentations, the
287: image of $\rho_{G,\pi}$ captures the whole semi-simple part of
288: ${\cal G}_{R,\pi}(\BZ)$. More precisely:
289: \begin{theorem}\label{teo} Assume $n$ is a natural number with $n\ge 4$.
290: Let $\pi : F_n\to G$ be a redundant presentation of the finite group $G$. Then
291: $\rho_{G,\pi}(\Gamma(G,\pi))\cap {\cal G}^1_{G,\pi}$ is
292: of finite index in the arithmetic group ${\cal G}_{G,\pi}^1(\BZ)$.
293: \end{theorem}
294: A presentation $\pi : F_n\to G$ is called {\it redundant} if there is a basis
295: $x_1,\ldots ,x_n$ of $F_n$ such that $\pi(x_n)=1$. This is equivalent to the
296: kernel of $\pi$ containing at least one primitive element
297: (that is a member of a basis) of $F_n$.
298:
299: In general we do not know whether
300: $\rho_{G,\pi}(\Gamma(G,\pi))\cap {\cal G}_{G,\pi}^1(\BZ)$ is of finite index
301: in $\rho_{G,\pi}(\Gamma(G,\pi))$. Section \ref{SL} contains criteria under
302: which this is going to happen.
303:
304: Choosing the finite group in various ways we obtain a rich variety of
305: linear representations of subgroups of finite index in $\A(F_n)$ with
306: arithmetic image groups. Let us start with some simple examples.
307:
308: {\it Example 1:} Let $G=C_2$ be the group of order $2$ and $\pi :F_n\to G$
309: be any surjective homomorphism. Since we generally have assumed $n\ge 2$, the
310: representation is redundant. As an abelian group we have
311: $\bar R\cong \BZ^{2(n-1)+1}$. Also $\BQ\otimes_{\bb Z} \bar R$
312: can be decomposed into
313: the $+1$-eigenspace of a generator of $G$ and the corresponding
314: $-1$-eigenspace. We obtain a decomposition
315: $\BQ\otimes_{\bb Z} \bar R=\BQ^n\oplus \BQ^{n-1}$
316: which has to be respected by
317: $\rho_{G,\pi}(\Gamma(G,\pi))$. Thus we obtain a representation
318: $$
319: \rho_{G,\pi} : \Gamma(G,\pi) \to \GL(n,\BQ)\times \GL(n-1,\BQ)
320: $$
321: which we show to have an image commensurable with $\SL(n,\BZ)\times
322: \SL(n-1,\BZ)$.
323:
324: The first factor is the familiar one arising from the representation
325: (\ref{furep}). But the second is somewhat less expected. It already gives rise
326: to a representation which is suitable for proving Corollary \ref{kaz}.
327:
328: {\it Example 2:} Let $p$ be a prime number, $\zeta_p\in\BC$ a non trivial
329: $p$-th root of unity. Let $K:=\BQ(\zeta_p)$ be the $p$-th cyclotomic field
330: and $\BZ(\zeta_p)$ its ring of integers.
331: Taking $G=C_p$ the cyclic group of order
332: $p$ we have $\bar R\cong \BZ^{p(n-1)+1}$. Now $p(n-1)+1=(p-1)(n-1)+n$ and we
333: get a representation
334: $$
335: \rho_{G,\pi} : \Gamma(G,\pi) \to \GL(n,\BQ)\times \GL(n-1,K)
336: $$
337: which we show to have an image commensurable with $SL(n,\BZ)\times
338: SL(n-1,\BZ(\zeta_p))$. This example is treated in Section \ref{Choose} along
339: with the case of a general cyclic group.
340:
341: Let $G$ be now an arbitrary finite group and $\pi: F_n\to G$ a redundant
342: epimorphism. Decompose the group algebra $\BQ[G]$
343: as
344: $$\BQ[G]=\BQ\oplus \prod_{i=2}^\ell M(h_i,D_i)$$
345: where the $D_i$ are division algebras. Theorem \ref{teo} implies that there is
346: a homomorphism $\rho$ from a finite index subgroup $\Gamma$ of $\A(F_n)$ into
347: a subgroup of $\prod_{i=2}^\ell\GL((n-1)h_i,D_i)$ where the image contains an
348: arithmetic subgroup of
349: $$\Lambda:=\prod_{i=2}^\ell\SL((n-1)h_i,{\cal R}_i)$$ where
350: ${\cal R}_i$ is an order in $D_i$. In general we do not know whether
351: $\rho(\Gamma)\cap\Lambda$ is of finite index in $\rho(\Gamma)$. It will be
352: especially interesting if it is not! In this case
353: $\rho(\Gamma)/(\rho(\Gamma)\cap\Lambda)$ would be an infinite abelian group and
354: this would imply that $\A(F_n)$ does not have Kazdhan property (T). We have
355: some results in the opposite direction showing, for example, that if $G$ is
356: metabelian then $\rho(\Gamma)$ is in fact commensurable to $\Lambda$.
357:
358: In any event, whenever $\rho(\Gamma)\cap\Lambda$ is of finite index or not, it
359: becomes of finite index after we divide by the center of
360: $\prod_{i=2}^\ell\GL((n-1)h_i,D_i)$. This shows that in all cases an
361: arithmetic group commensurable to $\Lambda$ is a quotient of some finite index
362: subgroup of $\A(F_n)$. Thus Theorem \ref{teo} provides a very rich class of
363: arithmetic subgroups which are quotients of (finite index subgroups) of
364: $\A(F_n)$. Theorem \ref{teo4} is obtained from the main result by some special
365: choices of the finite group $G$ and by an application of Margulis
366: super-rigidity (see Section \ref{Proofs}).
367:
368: The present paper opens up a large number of interesting problems. Among
369: others it shows that the classical Torelli subgroup (i.e. ${\rm IA}(F_n)$)
370: is just a first example in a series of infinitely many.
371: The kernel of $\rho_{G,\pi}$, for any (redundant) homomorphism
372: $\pi:F_n\to G$ of $F_n$ to a finite group $G$,
373: can be considered as a natural generalisation of
374: ${\rm IA}(F_n)$ which corresponds to the case where $G$ is the trivial group.
375: Just as ${\rm IA}(F_n)$, the kernel of $\rho_{G,\pi}$ is residually
376: torsion-free nilpotent and, at least when $\pi$ is redundant,
377: the image of $\rho_{G,\pi}$ is an arithmetic group hence
378: is finitely presented, which implies that
379: ${\rm ker}(\rho_{G,\pi})$ is a finitely generated normal subgroup.
380: Is ${\rm ker}(\rho_{G,\pi})$ a finitely generated group? In case of the
381: classical Torelli group this is a result of Magnus (see \cite{MKS}).
382:
383: Let us now describe the method of the proof of the main result and the layout
384: of the paper: In Section \ref{Two}, we describe in detail the rational
385: relation module $\BQ\otimes_{\bb Z}\bar R$ together
386: with ${\cal G}_{G,\pi}$ which
387: is the centraliser of $G$ when acting on the relation module. In Section
388: \ref{Drei}, we describe the homomorphism from $\Gamma (G,\pi)$ to
389: ${\cal G}_{G,\pi}$, which is given by the action of $\Gamma (G,\pi)$
390: on $\BQ\otimes_\BZ\bar R$. We compute the action of various elements of
391: $\Gamma (G,\pi)$ on $\BQ\otimes_{\bb Z}\bar R$ in Section \ref{Vier}
392: using results
393: on Fox calculus from Section \ref{Drei}. A careful choice of elements from
394: $\Gamma (G,\pi)$ allows us to produce sufficiently many unipotent elements
395: of ${\cal G}_{G,\pi}(\BZ)$ to appeal to results of Vaserstein \cite{Vas}
396: (see also Venkataramana \cite{Venk} for generalisations) to deduce
397: in Section \ref{Fuenf} that these
398: elements generate a finite index subgroup. Section \ref{Choose} is devoted to
399: a detailed study of the case when $G$ is a cyclic group. In this case we give
400: fairly complete results. Some of them are used in Section \ref{Proofs} for the
401: proof of Theorem \ref{teo4}.
402: In Section \ref{SL} we discuss the question
403: whether the image of $\Gamma (G,\pi)$ is commensurable with
404: ${\cal G}_{G,\pi}^1(\BZ)$ or an infinite extension of it. We solve it
405: for cyclic groups and also for metabelian groups.
406: Some remarks, computational
407: results and open problems are given in Section \ref{Concl}.
408:
409: In a subsequent paper \cite{GLu} we apply similar methods to study linear
410: representations of the mapping class group.
411:
412: {\it Acknowledgement:} The authors are grateful to A. Rapinchuk for a useful
413: conversation which led to the results about abelian groups in Section
414: \ref{SL}. They thank the Minerva-Landau Center for its support and especially the
415: Institute for Advanced Study in Princeton, where this
416: work was done, for its hospitality in the academic year 2005/2006. The second
417: author thanks the NSF, the Monell Foundation
418: and the Ellentuck Fund for support.
419:
420: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
421: \section{Relation modules and representations of subgroups of finite index in
422: $\A(F_n)$ }\label{Two}
423: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
424:
425: In this section we collect some results on the structure of the relation
426: modules of finite groups. We also consider
427: integrality properties of the linear reprentations of
428: subgroups of finite index in $\A(F_n)$ introduced in the introduction.
429:
430:
431: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
432: \subsection{The rational group ring}\label{sura}
433: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
434:
435: Here $G$ is a finite group and $\BQ[G]$ is its rational group ring.
436: Our basic reference for the structural results we need is the book of Curtis
437: and Reiner \cite{CR} (see also \cite{S}), in particular
438: Sections IV and XI therein.
439:
440: We start to set up some notation.
441: Let $N$ be an irreducible left $\BQ[G]$-module. We write
442: $$
443: {\bf I}_N(M)
444: $$
445: for the $N$-isotypic component of the left $\BQ[G]$-module $M$. We also choose
446: representatives $N_1,\ldots , N_\ell$ for the isomorphism classes of
447: irreducible left $\BQ[G]$-modules and abreviate
448: $$
449: {\bf I}_i(M):={\bf I}_{N_i}(M)\qquad (i=1,\ldots,\ell)
450: $$
451: for a left $\BQ[G]$-module $M$. We take $N_1$ to be the one dimensional
452: trivial module.
453:
454: Viewing $\BQ[G]$ as a left $\BQ[G]$-module we obtain a direct product
455: decomposition
456: \begin{equation}\label{ri}
457: \BQ[G]=\prod_{i=1}^\ell\, {\bf I}_i(\BQ[G]).
458: \end{equation}
459: In this case every ${\bf I}_i(\BQ[G])$ is also a right submodule of
460: $\BQ[G]$. Furthermore every ${\bf I}_i(\BQ[G])$ is an ideal of $\BQ[G]$ having
461: its own unity.
462: The decomposition
463: (\ref{ri}) is a decomposition of $\BQ[G]$ as a direct product of
464: $\BQ$-algebras.
465:
466: We let $P_i\in \BQ[G]$ ($i=1,\ldots,\ell$) be the right projector of
467: $\BQ[G]$ onto ${\bf I}_i(\BQ[G])$, that is we have
468: \begin{itemize}
469: \item[{\bf RP}:]
470: $AP_i=A$ for all $A\in {\bf I}_i(\BQ[G])$
471: and $AP_i=0$
472: if $A$ is in the direct product of the subrings
473: ${\bf I}_j(\BQ[G])$ with ${\bf I}_i(\BQ[G])$ excluded.
474: \end{itemize}
475: Notice that we also have
476: \begin{itemize}
477: \item[{\bf LP}:]
478: $P_iA=A$ for all $A\in {\bf I}_i(\BQ[G])$ and $P_iA=0$
479: if $A$ is in the direct product of the subrings
480: ${\bf I}_j(\BQ[G])$ with ${\bf I}_i(\BQ[G])$ excluded.
481: \end{itemize}
482: In fact we have $P_i\in {\bf I}_i(\BQ[G])$ for $i=1,\ldots ,\ell$
483: and $P_i$ can be thought of as the
484: unit element in ${\bf I}_i(\BQ[G])$.
485:
486: Given two left $\BQ[G]$-modules $M_1$, $M_2$ we define
487: ${\rm Hom}_G(M_1,M_2)$ to be the vector space of $\BQ[G]$-module homomorphisms
488: from $M_1$ to $M_2$. The $\BQ$-algebra of $\BQ[G]$-module endomorphisms of
489: $M_1$ is denoted by ${\rm End}_G(M_1)$.
490:
491: Let us write $S^{\rm op}$ for the {\it opposite ring}
492: of an associative ring $S$.
493: Mapping an element $A\in {\bf I}_i(\BQ[G])$ to the
494: multiplication from the right by $A$ defines $\BQ$-algebra isomorphisms
495: \begin{equation}\label{isoppp}
496: {\bf I}_i(\BQ[G])^{\rm op}\to
497: {\rm End}_G({\bf I}_i(\BQ[G])) \qquad (i=1,\ldots ,\ell).
498: \end{equation}
499: The existence of the isomorphisms (\ref{isoppp}) implies that
500: \begin{equation}\label{dim17}
501: {\rm dim}_{\bb Q}({\rm Hom}_G({\bf I}_i(\BQ[G])^f,{\bf I}_i(\BQ[G]))
502: =f{\rm dim}_{\bb Q}({\bf I}_i(\BQ[G]))
503: \end{equation}
504: for all $f\in\BN$ and $i=1,\ldots,\ell$.
505:
506: The algebras
507: $$
508: D_i:={\rm End}_G(N_i)\qquad (i=1,\ldots,\ell)
509: $$
510: of left $\BQ[G]$-module endomorphism of the irreducible modules $N_i$ are
511: finite dimensional division algebras with cyclotomic number fields as center.
512: We set
513: $$h_i:={\rm dim}_{D_i}(N_i)\qquad (i=1,\ldots,\ell).$$
514: Then the algebra ${\bf I}_i(\BQ[G])$
515: is isomorphic to a $h_i\times h_i$ matrix algebra $M(h_i,D_i)$ over $D_i$,
516: that is
517: \begin{equation}\label{ra1}
518: {\bf I}_i(\BQ[G])\cong M(h_i,D_i)\qquad (i=1,\ldots ,\ell).
519: \end{equation}
520: Using (\ref{isoppp}) we get isomorphisms of algebras
521: \begin{equation}\label{ra1oppp}
522: {\rm End}_G({\bf I}_i(\BQ[G])) \cong M(h_i,D_i^{\rm op})
523: \qquad (i=1,\ldots ,\ell).
524: \end{equation}
525: We shall further discuss integral versions of the isomorphisms (\ref{ri}) and
526: (\ref{ra1}). Our reference here is \cite{CR} Section XI or \cite{D}
527: Section XX.
528:
529: Let $K$ be a number field with ring of integers ${\cal O}$
530: and $S$ a finite dimensional
531: $K$-algebra. A subring ${\cal R}$ of $S$ is called an {\it order} if
532: it contains a $K$-basis of $S$ and if there is a subring ${\cal O}_0$ of
533: finite index in ${\cal O}$ such that ${\cal R}$ is a finitely generated
534: ${\cal O}_0$-module
535: (subrings are always assumed to contain the identity element of the bigger
536: ring). Let $h$ be a natural number and $S=M(h,D)$ where $D$ is a
537: finite dimensional division algebra over $K$ with center $K$. We write
538: ${\rm End}_S(S^m)$ ($m\in\BN$) for the algebra of $S$-left module
539: endomorphisms of the $S$-left module $S^m$. We obtain an algebra isomorphism
540: \begin{equation}\label{ord}
541: \Theta: {\rm End}_S(S^m)\to M(m,S)^\op=M(m,S^\op)=M(m,M(h,D^\op)).
542: \end{equation}
543: Here we identify $M(m,S)^\op$ in the usual way with $M(m,S^\op)$ and then with
544: $M(m,M(h,D^\op))$.
545: We call a subgroup $\Lambda$ of the additive group of $S^m$ a {\it lattice} if
546: $\Lambda$ contains a $K$-basis of $S^m$ and if
547: there is a subring ${\cal O}_0$ of finite index in ${\cal O}$ such that
548: $\Lambda$ is a finitely generated ${\cal O}_0$-module.
549: We write
550: ${\rm End}_S(S^m)_{\Lambda}$ for the ring of those endomorphisms stabilising
551: $\Lambda$. The following is well known (see \cite{CR}, \cite{D}).
552: \begin{lemma}\label{repkon} Let $\Lambda\le S^m$ ($m\in \BN$) be a lattice.
553: There is a sublattice $\tilde\Lambda\le \Lambda$ (of finite index in
554: $\Lambda$) and an order
555: ${\cal R}$ in $D$ such that
556: $$\Theta({\rm End}_S(S^m)_{\tilde\Lambda})\subset M(m,M(h,{\cal R}^\op)). $$
557: \end{lemma}
558: Let ${\rm Aut}_S(S^m)_\Lambda$ be the group of invertible elements in
559: ${\rm End}_S(S^m)_\Lambda$. Note that
560: ${\rm Aut}_S(S^m)_{\tilde\Lambda}$ is commensurable with
561: ${\rm Aut}_S(S^m)_{\Lambda}$ in particular
562: ${\rm Aut}_S(S^m)_{\Lambda}\cap {\rm Aut}_S(S^m)_{\tilde\Lambda}$
563: has finite index in ${\rm Aut}_S(S^m)_{\Lambda}$ and we obtain a
564: representation
565: \begin{equation}\label{ord1}
566: \Theta: {\rm Aut}_S(S^m)_{\Lambda}\cap {\rm Aut}_S(S^m)_{\tilde\Lambda}
567: \to \GL(m,M(h,{\cal R}^\op))=\GL(mh,{\cal R}^\op).
568: \end{equation}
569:
570:
571: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
572: \subsection{A result of Gasch\"utz}\label{suga}
573: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
574:
575: Next we need to report on a result of Gasch\"utz (see \cite{G}, \cite{Gr} and
576: \cite{JR}).
577: For this, let $G$ be a finite group and $\pi : F_n\to G$ a
578: surjective homomorphism
579: of a rank $n$ free group $F_n$ onto $G$. Let $R$ be the kernel of
580: $\pi$ and $\bar R:=R/R'$ the corresponding relation module with its action
581: (from the left) of $G$. The result of Gasch\"utz says that there is
582: a $\BQ[G]$-module isomorphism
583: \begin{equation}\label{gasch}
584: \kappa:\BQ\oplus \BQ[G]^{n-1}\to \BQ\otimes_{\bb Z}\bar R.
585: \end{equation}
586: where $\BQ$ is to be the trivial one dimensional $\BQ[G]$-module.
587: We fix a $\BQ[G]$-module isomorphism $\kappa$ once for all and identify
588: $\BQ\otimes_\BZ\bar R$ with $\BQ\oplus \BQ[G]^{n-1}$. In
589: particular we obtain ${\rm dim}_\BQ(\BQ\otimes_\BZ\bar{R})=1+|G|(n-1)$,
590: which is also clear from the formula of Reidemeister and Schreier.
591: We deduce
592: \begin{equation}\label{gaschi1}
593: {\bf I}_1(\BQ\otimes_{\bb Z}\bar R)={\bf I}_1(\BQ[G])^{n}=\BQ^n,
594: \end{equation}
595: \begin{equation}\label{gaschi}
596: {\bf I}_i(\BQ\otimes_{\bb Z}\bar R)={\bf I}_i(\BQ[G])^{n-1}
597: \qquad (i=2,\ldots,\ell).
598: \end{equation}
599:
600: From (\ref{dim17}) we get
601: \begin{lemma}\label{dim19}
602: Let $G$ be a finite group and $\pi : F_n\to G$ a
603: surjective homomorphism. Let $\bar R:=R/R'$ be the corresponding
604: relation module. Then we have:
605: $${\rm dim}_{\bb Q}({\rm Hom}_G({\bf I}_1(\BQ\otimes_{\bb Z}\bar{R}),
606: {\bf I}_1(\BQ[G])))=n$$
607: $${\rm dim}_{\bb Q}(
608: {\rm Hom}_G({\bf I}_i(\BQ\otimes_{\bb Z}\bar{R}), {\bf I}_i(\BQ[G])))=
609: (n-1){\rm dim}_{\bb Q}({\bf I}_i(\BQ[G])$$
610: \end{lemma}
611:
612: The result of Gasch\"utz enables us further to decompose the algebraic group
613: ${\cal G}_{G,\pi}$ and to describe the components.
614: We define for $i=1,\ldots,\ell$:
615: \begin{equation}\label{stabi}
616: {\cal G}_{G,\pi,i}:=\left\{\, \phi\in {\cal G}_{G,\pi}
617: \Mid \phi |_{{\bb C}\otimes{\bf I}_j({\bb Q}\otimes_{\bb Z}\bar R)}=
618: {\rm Id}_{{\bb C}\otimes{\bf I}_j({\bb Q}\otimes_{\bb Z}\bar R)}\
619: {\rm for}\ j\ne i\, \right\}.
620: \end{equation}
621: Note that the groups ${\cal G}_{G,\pi,i}$ centralise each other and
622: have pairwise trivial intersection. It is also easy to see that we have an
623: internal product decomposition:
624: $$
625: {\cal G}_{G,\pi}=\prod_{i=1}^\ell\, {\cal G}_{G,\pi,i}
626: $$
627: of $\BQ$-defined algebraic groups.
628:
629: We write $\GL(d,S)$ ($d\in \BN$) for the group of
630: invertible $d\times d$-matrices with entries in an associative ring $S$.
631: Following the discussion of the previous section we infer that the
632: identification made using
633: (\ref{gasch}) defines isomorphisms
634: $$
635: \Sigma_i : {\cal G}_{G,\pi,i}(\BQ)\to \GL(n-1,{\bf I}_i(\BQ[G])^{\rm op})
636: =\GL((n-1)h_i,D_i^{\rm op})
637: $$
638: for $i=2,\dots \ell$ and also an isomorphism
639: $$
640: \Sigma_1 : {\cal G}_{G,\pi,1}(\BQ)\to \GL(n,{\bf I}_1(\BQ[G])^{\rm op})=
641: \GL(n,\BQ).
642: $$
643: Together they define an isomorphism
644: $$
645: \Sigma : {\cal G}_{G,\pi}(\BQ)\to \GL(n,\BQ)\times \prod_{i=2}^\ell
646: \GL((n-1)h_i,D_i^{\rm op}).
647: $$
648:
649: Suppose that $i$ is one of the numbers $1,\ldots,\ell$ and let
650: $K_i$ be the center of ${\bf I}_i(\BQ[G])^{\rm op}$. As already
651: mentioned, this is a cyclotomic number field. The group
652: $\GL((n-1)h_i,D_i^{\rm op})$ is the group of $K_i$-rational points of a
653: $K_i$-defined linear algebraic group ${\cal H}_i$. The above discussion shows
654: that the $\BQ$-defined algebraic group ${\cal G}_{G,\pi,i}$
655: is equal to the base
656: field restriction of ${\cal H}_i$ from $K_i$ to $\BQ$. For this concept see
657: \cite{PLR}. We shall add a description of the
658: groups ${\cal G}_{G,\pi}^1$ introduced in the introduction.
659: Every element $g$ of $\GL((n-1)h_i,D_i^{\rm op})$ ($i=1,\ldots,\ell$)
660: acts by multiplication on $D_i^{(n-1)h_i}$. Taking a $K_i$-vector space
661: identification $D_i=K_i^{d_i}$ we associate to $g$ a square matrix of size
662: $(n-1)h_id_i$. The determinant of this matrix is called the reduced norm
663: ${\rm Nrd}(g)$ of $g$. The group of elements of reduced norm $1$ is
664: conventionally denoted by $\SL((n-1)h_i,D_i^{\rm op})$.
665: We have
666: $$
667: {\cal G}^1_{G,\pi,i}(\BQ)=\{\, g\in {\cal G}_{G,\pi,i}(\BQ)\Mid
668: {\rm Nrd}(\Sigma_i(g))=1\,\}.
669: $$
670:
671: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
672: \subsection{Representations of subgroups of finite index in
673: $\A(F_n)$}\label{repsfini}
674: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
675:
676: In this section we describe a matrix version of the representation (\ref{I5}).
677: To do this we use Lemma \ref{repkon} which constructs an integral
678: matrix representation of a subgroup of finite index in $\Gamma(G,\pi)$ which
679: is compatible with the representations
680: $\Sigma_i$ ($i=1,\ldots,\ell$) and $\Sigma$
681: from the previous section.
682:
683: Let $G$ be a finite group and $\pi : F_n\to G$ a surjective homomorphism
684: of the free group $F_n$ onto $G$. Let $R$ be the kernel of
685: $\pi$ and $\bar R$ the corresponding relation module. Let further
686: $\Gamma(G,\pi)$ be the subgroup of $\A(F_n)$ defined in
687: (\ref{I1}). In the sequel we shall often use the following fact without
688: further notice.
689: \begin{lemma} An element $\varphi\in\Gamma(R)$ is in $\Gamma(G,\pi)$ if and
690: only if the map $\bar\varphi :\bar R\to \bar R$ induced by $\varphi$ is
691: $G$-equivariant.
692: \end{lemma}
693: \begin{proof} If $\varphi$ is in $\Gamma(G,\pi)$ then
694: for every $w\in F_n$ we have a $r_w\in R$ with $\varphi(w)=wr_w$. Hence
695: $\varphi(wrw^{-1})=wr_w\varphi(r)r_w^{-1}w^{-1}$ holds for every $r\in R$.
696: The latter is equal to $ w\varphi(r)w^{-1}$ modulo $R'$ which
697: implies that $\bar\varphi :\bar R\to \bar R$ is
698: $G$-equivariant.
699:
700: If $\bar\varphi :\bar R\to \bar R$ is $G$-equivariant then for every
701: $w\in F_n$ the congruence
702: $\varphi(w)\varphi(r)\varphi(w)^{-1}\equiv w \varphi(r) w^{-1}$
703: holds modulo
704: $R'$ for every $r\in R$. This means that $w$ and $\varphi(w)$ act the same on
705: $\bar R$. The result of Gasch\"utz implies that the action of $G$ on $\bar R$
706: is faithful, that is no element except the identity of $G$ acts as the
707: identity. We conclude that $\varphi(w)w^{-1}$ is in $R$.
708: \end{proof}
709:
710: Keeping the notations set up in subsections \ref{sura}, \ref{suga} we define
711: $$
712: \Lambda_i:=\bar R\cap {\bf I}_i(\BQ\otimes_{\bb Z} \bar R)
713: \quad (i=1,\ldots ,\ell).
714: $$
715: Each $\Lambda_i$ is a $\BZ[G]$-submodule of
716: $\bar R\cap {\bf I}_i(\BQ\otimes_{\bb Z} \bar R)$ and
717: $\prod_{i=1}^\ell \Lambda_i\le \bar R$
718: is a finite index $\BZ[G]$-submodule.
719: Notice that the $\Lambda_i$ is a sublattice of
720: ${\bf I}_i(\BQ\otimes_{\bb Z} \bar R)=M(h_i,D_i)^{n-1}$
721: ($i=2,\ldots,\ell$) in the
722: sense of the previous section, so is
723: $\Lambda_1$ in ${\bf I}_1(\BQ\otimes_{\bb Z} \bar R)=\BQ^{n}$. Applying Lemma
724: \ref{repkon} we choose orders ${\cal R}_i\subset D_i$ and finite index
725: additive subgroups $\tilde\Lambda_i\le \Lambda_i$ ($i=1,\ldots,\ell$).
726: We define
727: \begin{equation}\label{hut}
728: \hat\Gamma(G,\pi)\le \Gamma(G,\pi)
729: \end{equation}
730: to be the simultaneous stabiliser of all the
731: $\tilde\Lambda_i\le \Lambda_i$ ($i=1,\ldots,\ell$).
732: Notice further that $\hat\Gamma(G,\pi)$
733: is of finite index in $\Gamma(G,\pi)$. Using the various representations
734: $\Theta$ from Lemma \ref{repkon} we obtain representations
735: $$
736: \sigma_{G,\pi,1}: \hat\Gamma (G,\pi) \to \GL(n,\BZ),\qquad
737: \sigma_{G,\pi,i}: \hat\Gamma (G,\pi) \to
738: \GL((n-1)h_i,{\cal R}_i^{\rm op})
739: $$
740: ($i=2,\ldots, \ell$),
741: which come by projection from the analogously defined representation
742: $$
743: \sigma_{G,\pi}: \hat\Gamma(G,\pi) \to
744: \GL(n,\BZ)\times \prod_{i=2}^\ell \GL(n-1,M(h_i,{\cal R}_i^{\rm op})).
745: $$
746: Notice that the diagram
747: $$
748: \xymatrix{
749: \hat\Gamma(G,\pi) \ar[r]^{} \ar[d]^{\sigma_{G,\pi}}
750: & {\cal G}_{G,\pi}(\BQ) \ar[ld]^{\Sigma} \\
751: \GL(n,\BQ)\times \prod_{i=2}^\ell \GL(n-1,M(h_i,D_i^{\rm op})) &
752: }
753: $$
754: is commutative.
755:
756:
757: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
758: \section{Derivations and the relation module}\label{Drei}
759: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
760:
761: This section contains a brief discussion of the
762: correspondences between Fox-derivatives, the
763: relation module of a finite group and cohomology. We apply these results to
764: obtain a certain canonical system of right module generators of
765: ${\rm Hom}_G(\BQ\otimes_{\bb Z} \bar R,\BQ[G])$
766: where $G$ is a finite group and
767: $\bar R$ is the relation module coming from a presentaion $\pi :F_n\to G$.
768:
769:
770: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
771: \subsection{Fox calculus and cohomology}
772: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
773:
774: For $n\in\BN$, let $F_n$ be the free group with basis $x_1,\ldots ,x_n$. We
775: write $\BZ[F_n]$, $\BQ[F_n]$ for the group rings of $F_n$ over the integers or
776: the rational numbers, respectively.
777: Let
778: $$
779: \frac{\partial}{\partial x_i}:F_n\to \BZ[F_n]\qquad (i=1,\ldots n)
780: $$
781: be the Fox-derivatives (see \cite{Bi} for a more elaborate exposition).
782: They are defined by the rules
783: \begin{equation}\label{F1}
784: \frac{\partial x_j}{\partial x_i}=\delta_{ij},\qquad \
785: \frac{\partial w_1w_2}{\partial x_i}=\frac{\partial w_1}{\partial x_i}
786: +w_1\frac{\partial w_2}{\partial x_i}
787: \end{equation}
788: for $i,\, j=1,\ldots n$ and $\ w_1,\, w_2\in F_n$. As usual,
789: $\delta_{ij}$ denoting the Kronecker symbol ($\delta_{ij}=1$ if $i=j$ and
790: $\delta_{ij}=0$ if $i\ne j$). For later use we note some consequences of
791: (\ref{F1}). For $i=1,\ldots , n$ and $e\in \BN$ we have
792: \begin{equation}\label{F2}
793: \frac{\partial x_i^e}{\partial x_i}=\sum_{k=0}^{e-1}x_i^k,\qquad\quad
794: \frac{\partial x_i^{-e}}{\partial x_i}=-\sum_{k=1}^{e}x_i^{-k}.
795: \end{equation}
796: Let further $w:=v_1x_i^{e_1} v_2x_i^{e_2}\cdot
797: \ldots \cdot v_kx_i^{e_k}v_{k+1}$ ($e_1,\ldots ,e_k\in\BZ$)
798: be an element of $F_n$ with the subwords
799: $v_j$ ($j=1,\ldots ,k+1$) not involving $x_i$, then
800: \begin{equation}\label{F3}
801: \frac{\partial w}{\partial x_i}=v_1\frac{\partial x_i^{e_1}}{\partial x_i}+
802: v_1x_i^{e_1}v_2 \frac{\partial x_i^{e_2}}{\partial x_i}+\ldots +
803: v_1x_i^{e_1}\cdot v_2x_i^{e_2}\cdot \ldots \cdot v_k
804: \frac{\partial x_i^{e_k}}{\partial x_i}.
805: \end{equation}
806: Let $G$ be a finite group generated by its elements
807: $g_1,\ldots , g_n$ ($n\in \BN$) and let
808: $$
809: \pi: F_n\to G\qquad \pi(x_1)=g_1,\ldots ,\pi(x_n)=g_n
810: $$
811: be the corresponding surjective group homomorphism.
812: The homomorphism $\pi:F_n\to G$ extends by linearity to
813: surjective ring homomorphisms:
814: $$
815: \pi: \BZ [F_n]\to \BZ[G]\qquad \pi: \BQ [F_n]\to \BQ[G].
816: $$
817: Let $R\le F_n$ be the kernel of $\pi$ and set $\bar{R}:=R/R'$
818: to be the commutator factor group of $R$. Let $\bar{} :R\to R/R'$ stand for
819: the quotient homomorphism.
820: The free group $F_n$ acts by
821: conjugation on $R$. Since its subgroup $R$ acts trivially on
822: $\bar{R}$ we get an induced action of $G$ on $\bar{R}$ which
823: we denote by $g\cdot u$ ($g\in G$, $u\in \bar{R}$). We generally write the
824: composition in $\bar{R}$ additively.
825: The free abelian group $\bar{R}$ then obtains the structure of a left
826: $\BZ[G]$-module which extends to a left $\BQ[G]$-module structure on
827: $\BQ\otimes_\BZ\bar{R}$. We have
828: \begin{lemma}\label{leF0}
829: Let $G$ be a group and $\pi :F_n\to G$ a homomorphism.
830: Let $R$ be the kernel of $\pi$. Then
831: $$\pi\left( \frac{\partial r}{\partial x_i}\right)=0$$
832: for all $r\in R'$ and $i=1,\ldots n$.
833: \end{lemma}
834: \begin{proof}
835: Note first that (\ref{F1}) implies
836: $\frac{\partial w^{-1}}{\partial x_i}=-w^{-1}
837: \frac{\partial w}{\partial x_i}$ for all $w\in F_n$ and $i=1,\ldots n$.
838: It is enough to show the result for single commutators
839: $r=r_1r_2r_1^{-1}r_2^{-1}$. A repeated application of (\ref{F1}) implies that
840: $$\frac{\partial r}{\partial x_i}=
841: \frac{\partial r_1}{\partial x_i}+
842: r_1\frac{\partial r_2}{\partial x_i}-
843: r_1r_2r_1^{-1}\frac{\partial r_1}{\partial x_i}-
844: r_1r_2r_1^{-1}r_2^{-1}\frac{\partial r_2}{\partial x_i}$$
845: for all $i=1,\ldots n$. As $\pi(r_1)=\pi(r_2)=1$ the result follows.
846: \end{proof}
847:
848: The Fox-derivatives
849: $\frac{\partial}{\partial x_i}$ now induce maps
850: $$
851: \partial_i: \bar{R}\to \BZ[G]\qquad (i=1,\ldots ,n)
852: $$
853: defined as follows.
854: For every $u\in\bar{R}$ choose $w_u\in R$ with $ \bar w_u=u$ and
855: define
856: \begin{equation}\label{deldef}
857: \partial_i(u):=\pi\left( \frac{\partial w_u}{\partial x_i}\right)\qquad
858: (i=1,\ldots ,n).
859: \end{equation}
860: We have
861: \begin{lemma}\label{leF}
862: The maps $\partial_1,\ldots \partial_n$ do not depend on the choices of the
863: preimages $w_u$ ($u\in\bar{R}$). They are $\BZ[G]$-module
864: homomorphisms when $\BZ[G]$ acts on itself from the left.
865: \end{lemma}
866: \begin{proof}
867: The first statement follows from Lemma \ref{leF0}. For the second note that
868: $$\frac{\partial grg^{-1}}{\partial x_i}=
869: \frac{\partial g}{\partial x_i}+g \frac{\partial r}{\partial x_i}-
870: grg^{-1}\frac{\partial g}{\partial x_i}$$
871: for all $g,\, r\in F_n$ and all $i=1,\ldots n$.
872: When $r$ is in $R$, $\pi(r)=1$ and the projection of the right hand side into
873: $\BZ[G]$ is equal to $\pi(g) \pi(\frac{\partial r}{\partial x_i})$.
874: \end{proof}
875: We also write
876: $\partial_i: \BQ\otimes_{\bb Z}\bar{R}\to \BQ[G]$
877: for the induced $\BQ[G]$-module homomorphisms.
878:
879: Let $t$ be the $\BZ$-rank of the free abelian group $\bar{R}$.
880: Note that $t=1+|G|(n-1)$ and also
881: $t={\rm dim}_{\bb Q}(\BQ\otimes_{\bb Z}\bar{R})$.
882: Let
883: $B:=(u_1,\ldots, u_t)$ be a $\BQ$-basis of $\BQ\otimes_{\bb Z}\bar{R}$.
884: We consider here
885: \begin{equation}\label{pm}
886: J_B:= \left( \begin{array}{ccc}
887: \partial_1(u_1) & \ldots & \partial_n(u_1)\\
888: \vdots & & \vdots \\
889: \partial_1(u_t) & \ldots & \partial_n(u_t)
890: \end{array} \right),\qquad
891: \end{equation}
892: which is a $t\times n$ matrix with entries in $\BQ[G]$.
893:
894: Given a left $\BQ[G]$-module $M$ and $s\in\BN$ we write
895: $$M^s:=\left\{
896: \left(
897: \begin{array}{c}
898: m_1 \\ \vdots \\ m_s
899: \end{array} \right) \Mid m_1,\ldots, m_s\in M\right\}
900: $$
901: for the corresponding $\BQ[G]$-module of column vectors.
902: The matrix $J_B$ induces a $\BQ$-linear map
903: $J_B:M^n\to M^t$. Notice for later use that given a $\BQ[G]$-submodule
904: $N\le M$ we have $J_B(N^n)\subseteq N^t$.
905:
906: Given the left $\BQ[G]$-module $M$ we write
907: $$
908: {\rm Der}(G,M):=\{\, d:G\to M\Mid d(gh)=gd(h)+d(g) \ {\rm for\ all}\ g,h\in
909: G\,\}
910: $$
911: for the vector space of derivations from $G$ to $M$.
912: Using the generators $g_1,\ldots ,g_n$ of our finite group $G$ we define
913: for $d\in {\rm Der}(G,M)$
914: $$
915: L(d):=
916: \left(\begin{array}{c}
917: d(g_1) \\ \vdots \\ d(g_n)
918: \end{array}\right)\in M^n.
919: $$
920: We have
921: \begin{lemma}\label{coholemm}
922: The map $L: {\rm Der}(G,M)\to M^n$ is injective and its image is equal to
923: ${\rm Ker}(J_B)$. Hence $L$ defines a $\BQ$-vector space isomorphism
924: between ${\rm Der}(G,M)$ and ${\rm Ker}(J_B)$.
925: \end{lemma}
926: \begin{proof}
927: Let $d$ be a derivation in the kernel of $L$. We infer that
928: $d(g_1)=\ldots=d(g_n)=0$. The derivation rule and the fact that the
929: $g_1,\ldots ,g_n$ generate $G$ implies that $d$ is identically zero.
930: This shows that $L$ is injective.
931:
932: Let $D:F_n\to M$ be a derivation where $F_n$ acts on $M$ through the
933: homomorphism $\pi : F_n\to G$. Since both sides are derivations and agree on
934: the $x_i$'s we find
935: \begin{equation}\label{deriv}
936: D(w)=\frac{\partial w}{\partial x_1}D(x_1)+\ldots
937: +\frac{\partial w}{\partial x_n}D(x_n)
938: \end{equation}
939: for all $w\in F_n$.
940:
941: Given a derivation $d: G\to M$, define an induced derivation
942: $D: F_n\to M$ by setting $D(w):=d(\pi(w))$. Applying (\ref{deriv}) we find
943: for every $r\in R$
944: $$0=d(1)=D(r)=\sum_{i=1}^n \frac{\partial r}{\partial x_i}D(x_i)=
945: \sum_{i=1}^n \partial_i (r) d(g_i).$$
946: Applying this to $r=u_1,\ldots, u_t$ shows that the image of $L$ is contained
947: in the kernel of $J_B$.
948:
949: Let now
950: $$m:=\left(\begin{array}{c}
951: m_1 \\ \vdots \\ m_n
952: \end{array}\right)\in M^n
953: $$
954: be an $n$-tuple of elements of $M$. It is well known that there is a
955: derivation $D :F_n\to M$ with $D(x_1)=m_1,\ldots, D(x_n)=m_n$.
956: If $m$ is an element of the kernel of $J_B$ then (\ref{deriv}) implies that
957: $D(r)=0$ for all $r\in R$. The derivation $D$ then induces a derivation
958: $d: G\to M$ with the property $d(g_1)=m_1,\ldots, d(g_n)=m_n$. This shows that
959: every element of the kernel of $J_B$ is in the image of $L$.
960: \end{proof}
961: \begin{corollary}\label{cohocor}
962: Let $M$ be a $\BQ[G]$-module then the kernel of the map
963: $J_B :M^n\to M^t$ is zero if $M$ is a trivial module and has $\BQ$-vector
964: space dimension equal to ${\rm dim}_{\bb Q}(M)$ if M does not have a trivial
965: submodule.
966: \end{corollary}
967: \begin{proof}
968: Given an element $m\in M$ we let $d_m :G\to M$ be the corresponding inner
969: derivation $d_m(g)=(g-1)m$. Let
970: $$
971: {\rm IDer}(G,M):=\{\, d_m \Mid m\in M\,\}\subseteq {\rm Der}(G,M)
972: $$
973: be the space of inner derivations. It is well known that
974: the first cohomology group $H^1(G,M)={\rm Der}(G,M)/{\rm IDer}(G,M)$
975: is zero ($G$ is a finite group). Note also that the map
976: $\mu : M\to {\rm IDer}(G,M)$,
977: defined by $\mu(m):=d_m$ ($m\in M$) is surjective.
978:
979: If $M$ is a trivial module then $\mu$ is identically zero, that implies
980: ${\rm IDer}(G,M)=\{0\}$ and hence also ${\rm Der}(G,M)=\{0\}$.
981: We find from Lemma \ref{coholemm} that
982: ${\rm dim}_{\bb Q}({\rm ker}(J_B))=0.$
983:
984: If $M$ does not contain the trivial module then
985: $\mu$ is injective, that implies
986: ${\rm dim}_{\bb Q}({\rm IDer}(G,M))={\rm dim}_{\bb Q}(M)$. We find from
987: Lemma \ref{coholemm} that
988: ${\rm dim}_{\bb Q}({\rm ker}(J_B))={\rm dim}_{\bb Q}({\rm Der}(G,M))={\rm
989: dim}_{\bb Q}(M).$
990: \end{proof}
991:
992: The results of this section will be applied in the next section in case of the
993: relation modules.
994:
995:
996: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
997: \subsection{${\rm Hom}_G(\BQ\otimes_{\bb Z} \bar R,\BQ[G])$
998: as right $\BQ[G]$-module}\label{Hom}
999: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1000:
1001: We keep here our notation from before, that is $G$ is a finite group
1002: generated by its elements $g_1,\ldots , g_n$ ($n\in \BN$) and
1003: $\pi: F_n\to G$ is the homomorphism defined by
1004: $\pi(x_1)=g_1,\ldots ,\pi(x_n)=g_n$. Furthermore $R$ is the kernel of $\pi$
1005: and $\bar R$ is the corresponding relation module.
1006: As the main result we shall show that the
1007: $\partial_1,\ldots,\partial_n\in {\rm Hom}_G(\BQ\otimes_\BZ \bar R,\BQ[G])$,
1008: which were constructed in the previous subsection generate
1009: ${\rm Hom}_G(\BQ\otimes_{\bb Z} \bar R,\BQ[G])$ if this
1010: space is considered as a
1011: $\BQ[G]$-module from the right in the following way.
1012:
1013: Let $M$ be a left $\BQ[G]$-module. For
1014: $f\in {\rm Hom}_G(M ,\BQ[G])$
1015: and $A\in \BQ[G]$ we define $f^A\in {\rm Hom}_G(M ,\BQ[G])$ by
1016: \begin{equation}
1017: f^A(x):=f(x)A\qquad (x\in M).
1018: \end{equation}
1019: Thus the $\BQ$-vector space ${\rm Hom}_G(M ,\BQ[G])$
1020: becomes a right $\BQ[G]$-module.
1021:
1022: We shall prove here:
1023: \begin{proposition}\label{wich} The maps $\partial_1,\ldots, \partial_n$
1024: (defined in (\ref{deldef})) generate
1025: ${\rm Hom}_G(\BQ\otimes_{\bb Z} \bar{R},\BQ[G])$
1026: as right $\BQ[G]$-module.
1027: \end{proposition}
1028:
1029: To prepare for the proof we decompose the matrix $J_B$ into submatrices.
1030: To do this we choose
1031: the basis $B$ adapted to the decomposition
1032: $$
1033: \BQ\otimes_{\bb Z}\bar{R}=\prod_{i=1}^\ell\,
1034: {\bf I}_i(\BQ\otimes_{\bb Z}\bar{R}).
1035: $$
1036: That is we choose bases
1037: $B_i=(u_{i1},\ldots u_{it_i})$
1038: $(t_i:={\rm dim}_{\bb Q}({\bf I}_i(\BQ\otimes_{\bb Z}\bar{R}))
1039: $
1040: for $i=1,\ldots,\ell$ and put them together to obtain a
1041: $\BQ$-vector space basis
1042: $$
1043: B=(u_1=u_{11},\ldots,u_t=u_{\ell t_\ell})
1044: $$
1045: of $\BQ\otimes_{\bb Z}\bar{R}$.
1046:
1047: Consider the matrices
1048: \begin{equation}\label{pmi}
1049: J_B^{P_i}:= \left( \begin{array}{ccc}
1050: \partial_1(u_1)P_i & \ldots & \partial_n(u_1)P_i\\
1051: \vdots & & \vdots \\
1052: \partial_1(u_t)P_i & \ldots & \partial_n(u_t)P_i
1053: \end{array} \right)=
1054: \left( \begin{array}{ccc}
1055: \partial_1^{P_i}(u_1) & \ldots & \partial_n^{P_i}(u_1)\\
1056: \vdots & & \vdots \\
1057: \partial_1^{P_i}(u_t) & \ldots & \partial_n^{P_i}(u_t)
1058: \end{array} \right)
1059: \end{equation}
1060: for $i=1,\ldots,\ell$. Notice that according to
1061: property {\bf LP} of the right projector $P_i$ (see section \ref{sura})
1062: the maps
1063: $$J_B,\, J_B^{P_i}: {\bf I}_i(\BQ[G])^n\to {\bf I}_i(\BQ[G])^t$$
1064: are the same. Furthermore we have
1065: $\partial_j(u)\in {\bf I}_i(\BQ[G])$ ($j=1,\ldots ,n$)
1066: for $u\in {\bf I}_i(\BQ\otimes_{\bb Z}\bar{R})$.
1067: This follows since each $\partial_j$ is a $\BQ[G]$-module homomorphism which
1068: has to respect isotypic components.
1069: Thus property {\bf RP} of the right projector $P_i$
1070: implies that all entries of $J_B^{P_i}$ are equal to zero except
1071: those in the submatrix
1072: \begin{equation}\label{pmis}
1073: J_{B_i}:= \left(\begin{array}{ccc}
1074: \partial_1(u_{i1})P_i & \ldots & \partial_n(u_{i1})P_i\\
1075: \vdots & & \vdots \\
1076: \partial_1(u_{i t_i})P_i & \ldots & \partial_n(u_{i t_i})P_i
1077: \end{array} \right)
1078: \end{equation}
1079: hence the kernels of the two maps
1080: $J_B : {\bf I}_i(\BQ[G])^n\to {\bf I}_i(\BQ[G])^t$ and
1081: $J_{B_i}:{\bf I}_i(\BQ[G])^n\to {\bf I}_i(\BQ[G])^{t_i}$
1082: are the same. Note finally that we have
1083: $$
1084: J_{B_i}=\left(\begin{array}{ccc}
1085: \partial_1^{P_i}(u_{i1}) & \ldots & \partial_n^{P_i}(u_{i1})\\
1086: \vdots & & \vdots \\
1087: \partial_1^{P_i}(u_{i t_i}) & \ldots & \partial_n^{P_i}(u_{i t_i})
1088: \end{array} \right)
1089: $$
1090: for all $i=1,\ldots,\ell$.
1091:
1092: \medskip
1093: \begin{prof} {\it of Proposition \ref{wich}:}
1094: We have the decomposition
1095: $$
1096: {\rm Hom}_G(\BQ\otimes_{\bb Z}\bar{R}, \BQ[G])=
1097: \prod_{i=1}^\ell
1098: {\rm Hom}_G({\bf I}_i(\BQ\otimes_{\bb Z}\bar{R}), {\bf I}_i(\BQ[G]))
1099: $$
1100: of ${\rm Hom}_G(\BQ\otimes_{\bb Z}\bar{R}, \BQ[G])$ as
1101: a right $\BQ[G]$-module.
1102: Since the $\partial_1^{P_i},\ldots ,\partial_n^{P_i}$ are zero on
1103: ${\bf I}_j(\BQ\otimes_{\bb Z}\bar{R})$ for $j\ne i$ it is enough to show that
1104: they generate
1105: ${\rm Hom}_G({\bf I}_i(\BQ\otimes_{\bb Z}\bar{R}), {\bf I}_i(\BQ[G]))$
1106: as a right $\BQ[G]$-module.
1107: Consider the $\BQ$-linear map
1108: $\Phi_i : {\bf I}_i(\BQ[G])^n \to
1109: {\rm Hom}_G({\bf I}_i(\BQ\otimes_{\bb Z}\bar{R}), {\bf I}_i(\BQ[G]))$
1110: given by
1111: $$
1112: \Phi_i\left(\left(
1113: \begin{array}{c}
1114: m_1 \\ \vdots \\ m_n
1115: \end{array} \right)\right):=
1116: \partial_1^{P_i m_1}+ \ldots +\partial_n^{P_i m_n}.
1117: $$
1118: Notice that ${\bf I}_i(\BQ[G])\cdot \BQ[G]={\bf I}_i(\BQ[G])$ and hence
1119: the image of $\Phi_i$ is the right $\BQ[G]$-submodule of
1120: ${\rm Hom}_G({\bf I}_i(\BQ\otimes_{\bb Z}\bar{R}), {\bf I}_i(\BQ[G]))$
1121: generated by
1122: the $\partial_1^{P_i},\ldots ,\partial_n^{P_i}$.
1123: The key observation now is
1124: \begin{equation}\label{dime}
1125: {\rm kernel}(J_{B_i})={\rm kernel}(\Phi_i)
1126: \end{equation}
1127: which is obvious by what is explained above. Notice that in fact
1128: $(m_1,\ldots,m_n)^t$ is in the kernel of $\Phi_i$ if and only if
1129: $\partial_1^{P_im_1}(u_{ik})+\ldots +\partial_n^{P_im_n}(u_{ik})=0$
1130: holds for all $k=1,\ldots,t_i$.
1131:
1132: We can treat now the case $i=1$ (the case of the trivial module).
1133: Here we have by (\ref{gaschi1}) that
1134: ${\rm dim}_{\bb Q}({\bf I}_1(\BQ\otimes_{\bb Z}\bar{R}))=n$ and the kernel
1135: of $\Phi_1$ is trivial (by Corollary \ref{cohocor}). Since we have
1136: ${\rm dim}_{\bb Q}({\rm Hom}_G({\bf I}_1(\BQ\otimes_{\bb Z}\bar{R}),
1137: {\bf I}_1(\BQ[G])))=n$ (by Lemma \ref{dim19}) the map $\Phi_1$ has to be
1138: surjective.
1139:
1140: In the cases $i >1$ we have
1141: $ {\rm dim}_{\bb Q}({\rm kernel}(\Phi_i))={\rm dim}_{\bb Q}({\bf I}_i(\BQ[G]))$
1142: by (\ref{dime}) and Corollary \ref{cohocor}. By Lemma \ref{dim19} and
1143: (\ref{dim17}) we have
1144: ${\rm dim}_{\bb Q}(
1145: {\rm Hom}_G({\bf I}_i(\BQ\otimes_{\bb Z}\bar{R}), {\bf I}_i(\BQ[G]))=
1146: (n-1){\rm dim}_{\bb Q}({\bf I}_i(\BQ[G]))$
1147: which implies that $\Phi_i$ has to be surjective.
1148: \end{prof}
1149:
1150: As a corollary to Proposition \ref{wich} we note
1151: \begin{corollary}\label{wichcor}
1152: The maps
1153: $\partial_1^{P_i},\ldots ,\partial_n^{P_i}$
1154: generate
1155: ${\rm Hom}_G({\bf I}_i(\BQ\otimes_{\bb Z}\bar{R}), {\bf I}_i(\BQ[G]))$
1156: as right ${\bf I}_i(\BQ[G])$-module for every $i=1,\ldots ,\ell$.
1157: \end{corollary}
1158: We shall give now a reformulation of the last corollary which will be needed
1159: later. Let
1160: ${\rm Id}_{i,k}\in {\bf I}_i(\BQ\otimes_{\bb Z}\bar{R})$
1161: ($k=1,\ldots,n-1$) be the images
1162: under $\kappa_i$ (see (\ref{gaschi}), (\ref{gaschi1}))
1163: of the unit elements of of the components of
1164: ${\bf I}_i(\BQ\otimes_{\bb Z}\bar{R})^{n-1}$ respectively of
1165: ${\bf I}_1(\BQ\otimes_{\bb Z}\bar{R})^{n}$ with $k=1,\ldots,n$.
1166: Using the identification (\ref{ra1oppp}) we think of
1167: $f\in{\rm Hom}_G({\bf I}_i(\BQ\otimes_{\bb Z}\bar{R}), {\bf I}_i(\BQ[G]))$
1168: being identified with the collection of matrices
1169: $f({\rm Id}_{i,1})=A_{i,1},\ldots ,
1170: f({\rm Id}_{i,n-1})=A_{i,n-1}\in M(h_i,D_i^\op)$ or
1171: $f({\rm Id}_{1,1})=A_{1,1},\ldots ,
1172: f({\rm Id}_{i,n})=A_{1,n}\in M(h_1,D_1^\op)=\BQ$.
1173:
1174: \begin{corollary}\label{wichcor1}
1175: Let $i$ be one of the numbers $2,\ldots,\ell$.
1176: Given $A_{1},\ldots ,A_{n-1}\in M(h_i,D_i^\op)$
1177: there are $B_1,\ldots,B_n\in \BQ[G]$ such that
1178: $$\partial_1^{P_iB_1}+\ldots +\partial_n^{P_iB_n}({\rm Id}_{i,k})=A_k
1179: \qquad (k=1,\ldots,n-1).$$
1180: Given $A_{1},\ldots ,A_{n}\in M(h_1,D_1^\op)$
1181: there are $B_1,\ldots,B_n\in \BQ[G]$ such that
1182: $$\partial_1^{P_1B_1}+\ldots +\partial_n^{P_1B_n}({\rm Id}_{1,k})=A_k
1183: \qquad (k=1,\ldots,n).$$
1184: \end{corollary}
1185:
1186: {\it Remark:} We sketch here another (more conceptual)
1187: proof of Proposition \ref{wich}. Features of the proof given above
1188: will play a role in the sequel.
1189:
1190: Using $H^2(G,\BQ\otimes_{\bb Z}\bar R)=0$ we
1191: obtain an injective homomorphism
1192: $$\epsilon : F_n/R'\to (\BQ\otimes_{\bb Z}\bar R)\rfish G$$
1193: from $F_n/R'$ to the semidirect product on the right hand side.
1194: On $R/R'\le F_n/R'$ the homomorphism $\epsilon$ induces the inclusion of
1195: $R/R'$ into $\BQ\otimes_{\bb Z}\bar R\le (\BQ\otimes_{\bb Z}\bar R)\rfish G$.
1196: The second component of $\epsilon$ coincides with $\pi$.
1197: Associate to $f\in{\rm Hom}_G(\BQ\otimes_{\bb Z}\bar R,\BQ[G])$ the
1198: homomorphism $\tau_f: (\BQ\otimes_{\bb Z}\bar R)\rfish G\to \BQ[G]\rfish G$
1199: ($\tau_f((u,g))=(f(u),g)$). Composition with $\epsilon$ creates a homomorphism
1200: $$\lambda_f: F_n\to F_n/R'\to \BQ[G]\rfish G, \qquad
1201: \lambda_f(w)=(d_f(w),\pi(w)),\quad (w\in F_n)$$
1202: The map $d_f:F_n \to \BQ[G]$ is a derivation. Hence we have for $w\in F_n$
1203: $$d_f(w)=\frac{\partial w}{\partial x_1}d_f(x_1)+\ldots
1204: +\frac{\partial w}{\partial x_n}d_f(x_n).
1205: $$
1206: Restricting this formula to $R/R'$ we have expressed $f$ in the desired way.
1207:
1208: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1209: \section{Redundant presentations}\label{Vier}
1210: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1211:
1212: In this section, we present a special decomposition of the relation module
1213: $\bar R$ for redundant presentations of finite groups. We then describe the
1214: action of various Nielsen automorphisms of $F_n$ on $\bar R$ with respect to
1215: this decomposition.
1216:
1217: Let $F_{n}$ ($n\in\BN, n\ge 2$)
1218: be the free group with basis $x_1,\ldots ,x_{n-1},\, y$.
1219: Let $G$ be a finite group generated by its elements
1220: $g_1,\ldots , g_{n-1}$ and let
1221: \begin{equation}\label{redu1}
1222: \pi: F_{n}\to G\qquad \pi(x_1)=g_1,\ldots ,\pi(x_{n-1})=g_{n-1},\ \pi(y)=1
1223: \end{equation}
1224: be the surjective group homomorphism corresponding to these data.
1225: We keep the notation introduced above of objects related to the presentation
1226: $\pi$ of $G$. That is $R$ is the kernel of $\pi$ and $\bar R=R/R'$ is the
1227: relation module, etc..
1228: We shall construct certain Nielsen-like elements in $\hat\Gamma(G,\pi)$
1229: (see \ref{hut}) which
1230: will allow us to find many unipotent elements in the image of the
1231: representation
1232: \begin{equation}\label{rep12}
1233: \sigma_{G,\pi}: \hat\Gamma(G,\pi) \to
1234: \GL(n,\BZ)\times \prod_{i=2}^\ell \GL(n-1,M(h_i,{\cal R}_i^\op))
1235: \end{equation}
1236: defined in Section \ref{repsfini}.
1237:
1238: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1239: \subsection{Decomposition of the relation module}
1240: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1241:
1242: With the notation fixed as indicated above we define
1243: $$
1244: R_{\rm old}:=\{\, x\in \langle x_1,\ldots ,x_{n-1}\rangle\Mid \pi(x)=1\,\}\
1245: {\rm and}\ \Ro:=R_{\rm old}/R'_{\rm old}.
1246: $$
1247: The injection $R_{\rm old}\to R$ defines a $\BZ[G]$-module homomorphism
1248: $\Ro\to \bar R$. Since $y\in R$, we can consider the $\BZ[G]$-submodule
1249: of $\bar R$ generated by $\bar y\in \bar R$.
1250:
1251: \begin{lemma} The homomorphism $\Ro\to\bar R$ is injective and so is the
1252: homomorphism
1253: $\BZ[G]\to \bar R$ defined by $A\mapsto A\cdot \bar y$.
1254: Identifying $\Ro$ with its image in $\bar R$ we obtain a direct
1255: sum decomposition
1256: $$
1257: \bar R=\bar R_{\rm old}\oplus \BZ[G]\cdot \bar y.
1258: $$
1259: as a $\BZ[G]$-module.
1260: \end{lemma}
1261: \begin{proof}
1262: Let us temporarily write $\hat R_{\rm old}$ for the image of $R_{\rm old}$ in
1263: $\bar R$. We first show that $\bar R=\hat R_{\rm old}+ \BZ[G]\cdot \bar y$.
1264: Given any $w\in F_n$, by looking at the image of $w$ modulo the normal closure
1265: of $y$ in $F_n$, we see that $w$ can be written in a unique way as
1266: $w=w_1w_2$ with $w_1,\, w_2\in F_n$ such
1267: that $w_1$ does not involve the letter $y$ whereas $w_2$ is a product of
1268: conjugates of powers of $y$. If $w$ is an element of $R$ then $w_1$ has to be
1269: in $R_{\rm old}$. This implies $\bar R=\hat R_{\rm old}+ \BZ[G]\cdot \bar y$.
1270:
1271: We infer that $\BQ\otimes_{\bb Z}\bar R=\BQ\otimes_{\bb Z}\hat R_{\rm old}
1272: + \BQ[G]\cdot \bar y$. We now count dimensions using the result of
1273: Gasch\"utz. As $\bar R$ and $\bar R_{\rm old}$ are free over $\BZ$,
1274: the lemma follows.
1275: \end{proof}
1276:
1277: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1278: \subsection{Nielsen Automorphisms}
1279: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1280:
1281: We study here the effect of certain Nielsen automorphisms on the relation
1282: module $\bar R$.
1283: We define elements $\alpha_i\in \A(F_n)$ ($i=1,\ldots ,n-1$) by
1284: \begin{equation}\label{al}
1285: \alpha_i(x_i):=yx_i.
1286: \end{equation}
1287: Our convention here is that values not given are identical to the argument,
1288: that is we assume
1289: $\alpha_i(x_j):=x_j,\ {\rm for}\ j\ne i\ {\rm and}\ \alpha_i(y)=y.$
1290:
1291: For $X\in \langle x_1,\ldots ,x_{n-1}\rangle$ we define elements
1292: $\beta_X\in \A(F_n)$ by
1293: \begin{equation}\label{be}
1294: \beta_X(y):=XyX^{-1}
1295: \end{equation}
1296: and finally for $U\in R_{\rm old}$ we define $\gamma_U\in \A(F_n)$ by
1297: \begin{equation}\label{ge}
1298: \gamma_U(y):=Uy.
1299: \end{equation}
1300: The $\alpha_i$, $\beta_X$, $\gamma_U$ extend to automorphisms of $F_{n}$.
1301: We have
1302: \begin{lemma}\label{NLe}
1303: The $\alpha_i$ ($i=1,\ldots,n-1$), $\beta_X$
1304: ($X\in \langle x_1,\ldots ,x_{n-1}\rangle$) and $\gamma_U$
1305: ($U\in R_{\rm old}$) are in
1306: $\Gamma(G,\pi)$ and the following formulas hold for the
1307: corresponding automorphisms induced on $\bar R$:
1308: \begin{equation}\label{alf}
1309: \bar\alpha_i(u)=u+\partial_i(u)\cdot \bar y\quad {\it for}\
1310: u\in \bar R_{\rm old}\ {\it and}\qquad
1311: \bar\alpha_i(\bar y)=\bar y.
1312: \end{equation}
1313: \begin{equation}\label{bef}
1314: \bar\beta_X(u)=u,\quad {\it for}\ u\in \bar R_{\rm old}\ {\it and}\qquad
1315: \bar\beta_X(\bar y)=\pi(X)\cdot \bar y,
1316: \end{equation}
1317: \begin{equation}\label{gef}
1318: \bar\gamma_U(u)=u,\quad {\it for}\ u\in \bar R_{\rm old}\ {\it and}\qquad
1319: \bar\gamma_U(\bar y)=\bar U+\bar y,
1320: \end{equation}
1321: \end{lemma}
1322: \begin{proof}
1323: It is first of all clear from our definitions that all the $\alpha_i$,
1324: $\beta_X$, $\gamma_U$ are elements of\
1325: $\Gamma(R):=\{\, \varphi\in \A(F_n)\Mid \varphi(R)=R \, \}$. Moreover all of
1326: them induce the identity on $F_n/R$ and hence are even in $\Gamma(G,\pi)$. It
1327: follows (see Section \ref{repsfini}) that they act
1328: as $\BZ[G]$-automorphisms on $\bar R$.
1329:
1330: Next we prove formula (\ref{alf}). Fix $i\in\{1,\ldots ,n-1\}$
1331: and define for $e\in\BN$
1332: $$
1333: W_e:=y\cdot x_iyx_i^{-1}\cdot\ldots \cdot x_i^{e-1}yx_i^{1-e},\quad
1334: W_{-e}:=x_i^{-1}y^{-1}x_i\cdot\ldots \cdot x_i^{-e}y^{-1}x_i^{e}\in F_n
1335: $$
1336: and add $W_0:=1$. We then have
1337: $\alpha_i(x_i^e)=W_e\cdot x_i^e$
1338: for all $e\in \BZ$. For $u\in\Ro$ choose $w\in R_{\rm old}$ with $\bar w=u$.
1339: We decompose $w\in \langle x_1,\ldots ,x_n\rangle$ as
1340: $w:=v_1x_i^{e_1} v_2x_i^{e_2}\cdot
1341: \ldots \cdot v_kx_i^{e_k}v_{k+1}$
1342: ($e_1,\ldots ,e_k\in\BZ$) such that the $v_j$ do not involve $x_i$. Define
1343: further
1344: $$
1345: V_1:=v_1,\ V_2:=v_1x_i^{e_1}v_2,\ldots ,V_k:=v_1x_i^{e_1}v_2\ldots
1346: x_i^{e_k}v_k.
1347: $$
1348: We then have
1349: $$
1350: \alpha_i(w)=V_1W_{e_1}V_1^{-1}\cdot V_2W_{e_2}V_2^{-1}
1351: \cdot\ldots\cdot
1352: V_k W_{e_2} V_k^{-1}\cdot w.
1353: $$
1354: Switching to the additive notation and using formulas (\ref{F2}) we find
1355: $$
1356: \bar\alpha_i(u)=u+\pi\left(v_1\frac{\partial x_i^{e_1}}{\partial x_i}+
1357: v_1x_i^{e_1}v_2 \frac{\partial x_i^{e_2}}{\partial x_i}+\ldots +
1358: v_1x_i^{e_1} v_2x_i^{e_2}\cdot \ldots \cdot v_k
1359: \frac{\partial x_i^{e_k}}{\partial x_i}\right)\cdot \bar y
1360: $$
1361: and formula (\ref{alf}) follows from (\ref{F3}) and (\ref{deldef}).
1362:
1363: To prove formula (\ref{bef}) notice that $\beta_X$ is the identity on $\Ro$,
1364: subsequently (\ref{bef}) is clear.
1365:
1366: All statements made about the automorphisms $\gamma_U$ ($U\in R_{\rm old}$)
1367: are evident. Notice that the map $\bar\gamma_U$ depends only on the image
1368: $\bar U$ of $U$ in $\bar R_{old}$.
1369: \end{proof}
1370: For later use we introduce now a notation for some elements in
1371: $\Gamma(G,\pi)$ obtained by composing the $\alpha_i$ and $\beta_X$. For
1372: every $g\in G$ we choose $X_g\in \langle x_1,\ldots ,x_{n-1}\rangle$
1373: with $\pi(X_g)=g$ and
1374: define $\beta_g:=\beta_{X_g}$. We infer from formulas (\ref{alf}) and
1375: (\ref{bef})
1376: $$\bar\beta_g\bar\alpha_i^m\bar\beta_{g^{-1}}(u+A\cdot\bar y)=
1377: u+(\partial_i(u)mg+A)\cdot\bar y\qquad (u\in \Ro,\, A\in\BZ[G])
1378: $$
1379: for all $g\in G$, $m\in \BZ$.
1380: Given an element $B=m_1g_1+\ldots +m_kg_k\in \BZ[G]$
1381: ($m_1,\ldots ,m_k\in\BZ$, $g_1,\ldots ,g_k\in G$) we define
1382: $\eta_{B,i}\in\A(F_n)$ by
1383: \begin{equation}\label{eta}
1384: \eta_{B,i}:=\beta_{g_1}\alpha_i^{m_1}\beta_{g_1^{-1}}\circ\ldots \circ
1385: \beta_{g_k}\alpha_i^{m_k}\beta_{g_k^{-1}}.
1386: \end{equation}
1387: for $i=1,\ldots ,n-1$.
1388: By Lemma \ref{NLe} it follows that $\eta_{B,i}\in \Gamma(G,\pi)$ for
1389: $i=1,\ldots ,n$ and $B\in \BZ[G]$.
1390: Using the notation of Section \ref{Hom} we have
1391: $$
1392: \bar\eta_{B,i}(u+A\cdot\bar y)=u+(\partial_i^B(u)+A)\cdot\bar y
1393: \qquad (u\in \Ro,\, A\in\BZ[G])
1394: $$
1395: for $i=1,\ldots ,n-1$ and $B\in\BZ[G]$. This formula makes it clear that
1396: $$
1397: \bar\eta_{m_1B_1+m_2B_2,i}=\bar\eta_{B_1,i}^{m_1}\circ \bar\eta_{B_2,i}^{m_2}
1398: $$
1399: holds for all $i=1,\ldots ,n-1$ and $B_1, \, B_2\in\BZ[G]$ and $m_1,\, m_2\in
1400: \BZ$.
1401:
1402: Using the representation (\ref{furep}) we set
1403: $$
1404: \A^+(F_n):=\{\, \phi\in \A(F)\Mid {\rm det}(\rho_1(\phi))=1\, \}.
1405: $$
1406: The subgroup $\A^+(F_n)$ is of index $2$ in $\A(F_n)$. For $n\ge 3$ it is
1407: a perfect group, that is $\A^+(F_n)$ is equal to its own commutator subgroup.
1408: We note that all elements $\bar\eta_{B,i}$ are in $\A^+(F_n)$.
1409:
1410:
1411: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1412: \section{Arithmeticity in case of a redundant presentation}\label{Fuenf}
1413: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1414:
1415: In this section we prove Theorem \ref{teo}: The decomposition of $\bar R$ to
1416: its isotypic components shows that our
1417: main goal is to prove that the image of $\hat\Gamma(G,\pi)$ in (\ref{rep12})
1418: contains a finite index subgroup of each component
1419: $$\SL(n-1,M(h_i,{\cal R}_i^\op))=\SL((n-1)h_i,{\cal R}_i^\op).$$
1420: To prove this we will show that this image contains sufficiently
1421: many unipotent elements with nonzero entries in the bottom strip of $h_i$ rows
1422: and also in the most right hand strip of $h_i$ columns of this
1423: $\SL((n-1)h_i,{\cal R}_i^\op)$. Then we can appeal to a theorem of Vaserstein
1424: which ensures that the elementary matrices with entries from a non-zero
1425: two-sided ideal of ${\cal R}_i$ generate a finite index subgroup of the
1426: arithmetic group $\SL((n-1)h_i,{\cal R}_i^\op)$.
1427:
1428: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1429: \subsection{Elementary matrices}\label{Fuenf1}
1430: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1431:
1432: This section sets up certain notations and facts about elementary matrices.
1433:
1434: Let $S$ be an associative ring with unity.
1435: We denote by $\GL(n,S)$
1436: ($n\in\BN$) the group
1437: of invertible $n\times n$ matrices with entries in $S$. We write $I_n$ for the
1438: identity matrix. Given $1\le i,\, j\le n$ with $i\ne j$ and $s\in S$ we
1439: define $E_{ij}(s)\in \GL(n,S)$ to be the corresponding elementary matrix, that
1440: is $E_{ij}(s)$ is equal to $I_n$ except for the $(i,\, j)$ position, where we
1441: put $s$. For $1\le i,\, j\le n$ with $i\ne j$, $1\le k,\, l \le n$
1442: with $k\ne l$ and $s_1,\, s_2\in S$
1443: the Steinberg relations between the elementary matrices are
1444: \begin{equation}\label{stein}
1445: [E_{ij}(s_1),E_{k\ell}(s_2)]:=\left\{\begin{array}{ccc}
1446: 1 & {\rm for}\ j\ne k, & i\ne l, \\
1447: E_{il}(s_1s_2) & {\rm for}\ j= k, & i\ne l, \\
1448: E_{kj}(-s_2s_1) & {\rm for}\ j\ne k, & i= l, \\
1449: \end{array} \right.
1450: \end{equation}
1451: where
1452: $[E_{ij}(s_1),E_{k l}(s_2)]=E_{ij}(s_1)E_{kl}(s_2)
1453: E_{ij}(s_1)^{-1}E_{kl}(s_2)^{-1}$. For every admissible pair $i,\, j$ we have
1454: $E_{ij}(s_1)E_{ij}(s_2)=E_{ij}(s_1+s_2)$ for all $s_1,\, s_2\in S$.
1455:
1456: Let ${\lie a}\subset S$ be an ideal of $S$. Given natural numbers $m_1,\,
1457: m_2\in\BN$ we denote by $M(m_1,m_2;{\lie a})$ the space of
1458: $m_1\times m_2$-matrices with entries in ${\lie a}$.
1459: Suppose now that $m_1+m_2=n$. For
1460: $A\in M(m_1,m_2;{\lie a})$ we define
1461: $$
1462: H(A):=\left(\begin{array}{cc}
1463: I_{m_2} & 0\\
1464: A & I_{m_1}
1465: \end{array} \right),\qquad
1466: V(A):=\left(\begin{array}{cc}
1467: I_{m_2} & A^t\\
1468: 0 & I_{m_1}
1469: \end{array} \right)
1470: $$
1471: Here $A^t$ stands for the transpose of the matrix $A$. Suppose that
1472: $A:=(a_{ij})$ where $i=1,\ldots,m_1$, $j=1,\ldots,m_2$ with
1473: $a_{ij}\in{\lie a}$, then we have
1474: $$
1475: H(A)=\prod_{i=1}^{m_1} \prod_{j=1}^{m_2}\, E_{m_2+i\, j}(a_{ij}),
1476: \qquad
1477: V(A)=\prod_{i=1}^{m_1} \prod_{j=1}^{m_2}\, E_{j\, m_2+i}(a_{ij}).
1478: $$
1479: Given $m_1,\, m_2\in\BN$ with $m_1+m_2=n$ and an
1480: ideal $\lie{a}\le S$ we define
1481: $$
1482: H(m_1,m_2;{\lie a}):=\{\, H(A)\Mid A\in M(m_1,m_2; {\lie a})\,\},
1483: $$
1484: $$
1485: V(m_1,m_2;{\lie a}):=\{\, V(A)\Mid A\in M(m_1,m_2; {\lie a})\,\}.
1486: $$
1487: Both $H(m_1,m_2;{\lie a})$ and $V(m_1,m_2;{\lie a})$
1488: are abelian subgroups of $\GL(n,S)$.
1489:
1490: We apply this notation in case of the orders $M(h_i,{\cal R}_i)$
1491: of the simple
1492: factor rings ${\bf I}_i(\BQ[G])$ of the rational group ring $\BQ[G]$.
1493: So let $K$ be a number field and $D$ a
1494: finite dimensional division algebra over $K$ with center $K$.
1495: Let ${\cal R}$ be a order in $D$. It is well
1496: known that every two-sided ideal $\lie{a}\le {\cal R}$
1497: is additively finitely
1498: generated and has finite index in ${\cal R}$
1499: if it is non-zero. We have
1500:
1501: \begin{proposition}\label{wasser}
1502: Let $n\in \BN$ with $n\ge 3$ and let $m_1,\, m_2\in\BN$ satisfy $m_1+m_2=n$.
1503: Let ${\lie a}$ be a non-zero two-sided ideal of ${\cal R}$
1504: Then $H(m_1,m_2;{\lie a})$ and $V(m_1,m_2;{\lie a})$ generate a subgroup of
1505: finite index in $\SL(n,{\cal R})$.
1506: \end{proposition}
1507: \begin{proof} Let $U$ be the sugroup generated by
1508: $H(m_1,m_2;{\lie a})$ and $V(m_1,m_2;{\lie a})$. Using
1509: the Steinberg relations (\ref{stein}) we conclude
1510: that there is a non-zero two-sided ideal ${\lie b}$ ($={\lie a}^2$) in ${\cal
1511: R}$ such that every $E_{ij}(b)$ ($b\in {\lie b}$, $i\ne j$) is in $U$. The
1512: main result of \cite{Vas} implies that $U$ has finite index in
1513: $\SL(n,{\cal R})$.
1514: \end{proof}
1515:
1516:
1517: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1518: \subsection{Constructing elements of ${\cal G}_{G,\pi}^1(\BZ)$}
1519: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1520:
1521:
1522: Let $F_{n}$ ($n\in\BN,\, n\ge 3$)
1523: be the free group with basis $x_1,\ldots ,x_{n-1},\, y$, $G$ a finite group
1524: with generators
1525: $g_1,\ldots , g_{n-1}$ and let
1526: $$
1527: \pi: F_{n}\to G\qquad \pi(x_1)=g_1,\ldots ,\pi(x_{n-1})=g_{n-1},\ \pi(y)=1
1528: $$
1529: be the redundant presentation resulting from the data. Notice the assumption
1530: $n\ge 3$, if $n=2$ then $G$ is cyclic. We enclude this case into the next
1531: section.
1532:
1533: Given $i=2,\ldots ,\ell$ and an ideal ${\lie a}$ of the order
1534: ${\cal R}_i^\op$,
1535: let us define
1536: \begin{equation}\label{pde1}
1537: H_i({\lie a}):=H_i(h_i,(n-2)h_i;{\lie a})\le \GL(n,\BZ)
1538: \times \prod_{i=2}^\ell \GL(n-1,M(h_i,{\cal R}_i^\op))
1539: \end{equation}
1540: to be
1541: $H(h_i,(n-2)h_i;{\lie a})\le \GL((n-1)h_i,{\cal R}_i^\op)=
1542: \GL(n-1,M(h_i,{\cal R}_i^\op))$
1543: but considered as a subgroup of
1544: the big direct product on the right hand side of (\ref{pde1})
1545: (with all components of the factors different from the $i$-th factor equal to
1546: the identity matrix). In case $i=1$ we have ${\cal O}_1=\BZ$
1547: and we define
1548: $H_1({\lie a})=H_1(1,n-1;{\lie a})$
1549: with $n-1$ replaced by $n$. Let us also
1550: define
1551: \begin{equation}\label{pde2}
1552: V_i({\lie a}):=V_i(h_i,(n-2)h_i;{\lie a})\le \GL(n,\BZ)
1553: \times \prod_{i=2}^\ell \GL(n-1,M(h_i,{\cal R}_i^\op))
1554: \end{equation}
1555: in a similar way. We have
1556: \begin{lemma}\label{eli1}
1557: Let $i$ be one of the $1,\ldots ,\ell$ and $g\in H_i({\cal R}_i^\op)$.
1558: There is an
1559: $e \in\BN$ such that $g^e$ is in $\sigma_{G,\pi}(\hat\Gamma(G,\pi))$
1560: where $\sigma_{G,\pi}$ is as in (\ref{rep12}).
1561: \end{lemma}
1562: \begin{proof} We prove the statement for $i\ge 2$, for $i=1$ the proof is only
1563: different in notation.
1564:
1565: Let $i\ge 2$ be fixed and consider $g\in H_i({\cal R}_i^\op)$.
1566: There are matrices $A_1,\ldots,A_{n-2}\in M(h_i,{\cal R}_i^\op)$ such that
1567: the $i$-th component of $g_i$ of $g$ is equal to
1568: $$g_i=\left(\begin{array}{cccccc}
1569: I_{h_i} & 0 & 0 & \ldots & 0 & 0\\
1570: 0 & I_{h_i} & 0 & \ldots & 0 & 0\\
1571: \vdots & \vdots & \vdots & \vdots & \vdots & \vdots \\
1572: 0 & 0 & 0 & \ldots & I_{h_i} & 0\\
1573: A_1 & A_2 & A_3 & \ldots & A_{n-2} & I_{h_i}
1574: \end{array} \right),
1575: $$
1576: all other components being equal to the identity matrix in the corresponding
1577: component group.
1578: By Corollary \ref{wichcor1}
1579: there are $B_1,\ldots,B_{n-1}\in \BQ[G]$ such that
1580: $$\partial_1^{P_iB_1}+\ldots +\partial_{n-1}^{P_iB_{n-1}}({\rm Id}_{i,k})=A_k$$
1581: for $k=1,\ldots, n-2$. Choose $e \in\BN$ to have sufficiently many divisors so
1582: that $eP_iB_1,\ldots ,eP_iB_{n-2}$ are in $\BZ[G]$. Put
1583: $$\gamma:=\eta_{eP_iB_1,1}\circ \ldots \circ\eta_{eP_iB_{n-1},n-1} $$
1584: with the $\eta$ defined in (\ref{eta}).
1585: Increase now (if
1586: necessary) $e$ to ensure $\gamma\in \hat\Gamma(G,\pi)$ (see (\ref{hut})).
1587: We have $\sigma_{G,\pi}(\gamma)=g^e$ and the lemma is proved.
1588: \end{proof}
1589: \begin{lemma}\label{eli2}
1590: Let $i$ be one of the $1,\ldots ,\ell$ and $g\in V_i({\cal R}_i^\op)$.
1591: There is an
1592: $e \in\BN$ such that $g^e$ is in $\sigma_{G,\pi}(\hat\Gamma(G,\pi))$.
1593: \end{lemma}
1594: \begin{proof}
1595: We prove the statement for $i\ge 2$, again for $i=1$ the proof is only
1596: different in notation.
1597:
1598: Let $i\ge 2$ be fixed and consider $g\in V_i({\cal R}_i^\op)$.
1599: There are matrices $A_1,\ldots,A_{n-2}\in M(h_i,{\cal R}_i^\op)$ such that
1600: the $i$-th component of $g_i$ of $g$ is equal to
1601: $$g_i=\left(\begin{array}{cccccc}
1602: I_{h_i} & 0 & 0 & \ldots & 0 & A_1\\
1603: 0 & I_{h_i} & 0 & \ldots & 0 & A_2\\
1604: \vdots & \vdots & \vdots & \vdots & \vdots & \vdots \\
1605: 0 & 0 & 0 & \ldots & I_{h_i} & A_{n-2}\\
1606: 0 & 0 & 0 & \ldots & 0 & I_{h_i}
1607: \end{array} \right),
1608: $$
1609: The matrices $A_1,\ldots,A_{n-2}$ can be considered as elements of
1610: $M(h_i,{\cal R}_i)$, after all $M(h_i,{\cal R}_i^\op)$ and
1611: $M(h_i,{\cal R}_i)$ are the same sets. Using the convention of Section
1612: \ref{repsfini} we consider $u:=(A_1,\ldots,A_{n-2})$ as an element of
1613: $M(h_i,D_i)^{n-2}\subseteq \BQ\otimes_{\bb Z}\bar R_{\rm old}$
1614: (see (\ref{gaschi})).
1615: Then there is $e \in\BN$ such that $eu\in \bar R_{\rm old}$.
1616: We choose $U\in R_{\rm old} $ with $\bar U=eu$. We
1617: take $e \in\BN$ to have sufficiently many divisors so
1618: that $\bar\gamma_U$ (see (\ref{ge})) is in $\hat\Gamma(G,\pi)$. We have
1619: $\bar\gamma_U=g^e$ and the lemma is proved.
1620: \end{proof}
1621:
1622: Since every $H_i({\cal R}_i^\op)$ or $V_i({\cal R}_i^\op)$ ($i=1,\ldots,\ell$)
1623: is a finitely generated abelian group the two previous lemmas immediately
1624: imply
1625: \begin{proposition}\label{prfi}
1626: There is $e\in \BN$ such that
1627: \begin{equation}\label{pde3}
1628: \prod_{i=1}^\ell H_i(e{\cal R}_i)\le \sigma_{G,\pi}(\hat\Gamma(G,\pi))
1629: \quad {and}\quad \prod_{i=1}^\ell V_i(e{\cal R}_i)\le
1630: \sigma_{G,\pi}(\hat\Gamma(G,\pi)).
1631: \end{equation}
1632: \end{proposition}
1633:
1634: Finally we can state
1635: \begin{theorem}\label{prfit} Let $n\in \BN$ satisfy $n\ge 4$, then
1636: $$\sigma_{G,\pi}(\hat\Gamma(G,\pi))\cap \SL(n,\BZ)
1637: \times \prod_{i=2}^\ell \SL((n-1)h_i,{\cal R}_i^\op))$$
1638: is of finite index in $\SL(n,\BZ)
1639: \times \prod_{i=2}^\ell \SL((n-1)h_i,{\cal R}_i^\op))$.
1640: \end{theorem}
1641: \begin{proof} Just use Proposition \ref{prfi} together with Proposition
1642: \ref{wasser}.
1643: \end{proof}
1644:
1645: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1646: \section{Cyclic groups}\label{Choose}
1647: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1648:
1649:
1650: In this section we present a detailed study of the case when
1651: $G$ is a cyclic group. We restrict ourselves to a particular
1652: type of (redundant) presentation, but following the discussion in Section
1653: \ref{Concl3} this is in fact the general case. An important feature
1654: is that we are able to present a set of
1655: generators of $\Gamma(G,\pi)$, in the case $n\ge 3$. For a general
1656: group $G$ we are not
1657: able to do this. As a consequence we can determine the precise image of
1658: $\Gamma(G,\pi)$ (not only up to commensurability)
1659: acting on its invariant pieces of the relation module.
1660: Some of the results here will be needed in Section
1661: \ref{Proofs}, but we believe that this section is of independent interest.
1662: In fact, our method can be observed here in a very simple, almost elementary
1663: situation.
1664:
1665: Let us call the attention of the reader to the fact that our results
1666: concerning cyclic groups, in fact, cover the general case of $G$ being
1667: abelian. This follows since every irreducible representation of an abelian
1668: group factors through a cyclic quotient, see the discussion in Section
1669: \ref{SL3}.
1670:
1671: Let the free group $F_n$ ($n\ge 2$) be generated
1672: by $x,\, y_1,\ldots, y_{n-1}$. Let us introduce the following elements of
1673: $\A(F_n)$. Our convention here is again that values
1674: not given are identical to the argument.
1675: \begin{itemize}
1676: \item $\delta_i(x):=y_ix$ and $\epsilon_i(x):=xy_i$ for $i=1,\ldots ,n-1$,
1677: \item $\varphi_i(y_i):=xy_i$ and
1678: $\psi_i(y_i):=y_ix$ for $i=1,\ldots ,n-1$,
1679: \item $\lambda_{ij}(y_i):=y_jy_i$ and $\nu_{ij}(y_i):=y_iy_j$
1680: for $i,\, j=1,\ldots ,n-1$ with $i\ne j$.
1681: \end{itemize}
1682: A theorem of Nielsen (see \cite{MKS}) asserts that these elements generate
1683: $\A^+(F_n)$ where $\A^+(F_n)$ is the kernel of the homomorphism
1684: $\rho_1: \A(F_n) \to \GL(n,\BZ)$ followed by the determinant. If
1685: $\Gamma \le \A(F_n)$ is a subgroup, we define $\Gamma^+:=\Gamma\cap
1686: \A^+(F_n)$. Let us further introduce
1687: \begin{itemize}
1688: \item $\kappa_{jk}(x):= x [y_j,y_k]$, $\kappa_{ijk}(y_i):= y_i [y_j,y_k]$
1689: for $1\le i,\, j,\, k\le n-1$ with $i\ne j$, $i\ne k$,
1690: \item $\tau_{ij}(y_i):= y_i [x,y_j]$ for $1\le i,\, j\le n-1$
1691: with $i\ne j$.
1692: \end{itemize}
1693: The set $T_n$ consisting of the $\kappa_{jk}$, $\kappa_{ijk}$,
1694: $\tau_{ij}$ (with indices as above), the
1695: $\delta_i^{-1}\circ \epsilon_i$, $\varphi_i^{-1}\circ \psi_i$ for
1696: $i=1,\ldots ,n-1$ and the $\lambda_{ij}\circ\nu_{ij}^{-1}$ for
1697: $1\le i,\, j\le n-1$ with $i\ne j$
1698: generates the group ${\rm IA}(F_n)$ by
1699: another theorem of Nielsen (see \cite{MKS}). Notice that $T_n$ is contained in
1700: $\Gamma(G,\pi)$ for every presentation $\pi$ of an abelian group $G$.
1701:
1702: Let us also introduce the following notation concerning matrix groups. We
1703: write
1704: $$\Gamma^1(n,m)\le \SL(n,\BZ)$$
1705: for the subgroup of $\SL(n,\BZ)$ consisting of those elements having a first
1706: row which is congruent to $(1,0,\ldots,0)$ modulo $m\in\BN$. We need
1707: \begin{lemma}\label{eli11} Let $n$ and $m$ be natural numbers with $n\ge
1708: 3$. The group $\Gamma^1(n,m)$ is generated by the elementary matrices
1709: $$E_{ij}(1)\quad (i,\, j=1,\ldots,n, i\ne j, i\ne 1),\qquad E_{1j}(m) \quad
1710: (j=2,\ldots,n).$$
1711: \end{lemma}
1712: This lemma is an elementary exercise using the euclidean algorithm in $\BZ$.
1713:
1714:
1715: We first treat the cyclic group of order $2$ which we call $C_2$. We have
1716: singled out this case since it is completely elementary and also plays a
1717: special role in the proof of Corollary \ref{kaz}.
1718: Let $g$ be a generator of $C_2$. We choose
1719: the presentation
1720: $$
1721: \pi:F_n\to C_2,\quad \pi(x)=g,\ \pi(y_1)=\ldots=\pi(y_{n-1})=1.
1722: $$
1723: The corresponding relation module $\bar R$ has the following elements as a
1724: $\BZ$-basis.
1725: $$\overline{x^2},\ \bar y_1,\ldots,\bar y_{n-1},\
1726: \overline{xy_1x^{-1}}=g\bar y_1,\ldots,
1727: \overline{xy_{n-1}x^{-1}}=g\bar y_{n-1}.$$
1728: The $\BQ$-vector space $\BQ\otimes_{\bb Z} \bar R$ decomposes as
1729: $\BQ\otimes_{\bb Z} \bar R=V_1\oplus V_{-1}$ where
1730: $V_1$, $V_{-1}$ are the $\pm 1$
1731: eigenspaces of $g$ respectively.
1732: Set $\bar R_1:=\bar R\cap V_1$ and $\bar R_{-1}:=\bar R\cap V_{-1}$.
1733: Notice that $\bar R_1+\bar R_{-1}$ is of finite index in $\bar R$. Introduce
1734: \begin{equation}\label{basi}
1735: v_i:=\bar y_i+\overline{xy_{i}x^{-1}},\quad
1736: w_i:=\bar y_i-\overline{xy_{i}x^{-1}}\qquad (i=1,\ldots ,n-1).
1737: \end{equation}
1738: Then $\overline{x^2},v_1,\ldots,v_{n-1}$ is a $\BZ$-basis of $\bar R_1$ and
1739: $w_1,\ldots,w_{n-1}$ is a $\BZ$-basis of $\bar R_{-1}$.
1740: Since $\Gamma(C_2,\pi)$
1741: leaves $\bar R_1$ and $\bar R_{-1}$ invariant we obtain,
1742: with the above $\BZ$-bases being chosen, representations
1743: \begin{equation}\label{rep2}
1744: \sigma_1:\Gamma^+(C_2,\pi)\to \GL(n,\BZ),\qquad
1745: \sigma_{-1}:\Gamma^+(C_2,\pi)\to \GL(n-1,\BZ).
1746: \end{equation}
1747: With all these data given we have:
1748: \begin{proposition}\label{cyc2} Let $n\ge 2$ be a natural number. The
1749: following hold.
1750:
1751: {\rm (i)} The group $\Gamma^+(C_2,\pi)$ is generated by the
1752: automorphisms $\delta_i$, $\epsilon_i$
1753: ($i=1,\ldots,n-1$), the $\lambda_{ij}$ and $\nu_{ij}$
1754: ($i,\, j=1,\ldots ,n-1$, $i\ne j$), the $\varphi_i^2$ ($i=1,\ldots,n-1$) and
1755: the elements of the set $T_n$ introduced above.
1756:
1757: {\rm (ii)} The image of $\sigma_1$ is equal to $\Gamma^1(n,2)$.
1758:
1759: {\rm (iii)} The index of $\Gamma^+(C_2,\pi)$ in $\A^+(F_n)$ is $2^{n}-1$.
1760:
1761: {\rm (iv)} The representation $\sigma_{-1}$ is surjective onto $\GL(n-1,\BZ)$.
1762:
1763: {\rm (v)} The group $\sigma_{-1}({\rm IA}(F_3))\le \GL(2,\BZ)$ is generated by
1764: the matrices
1765: $$\left(\begin{array}{cc}
1766: -1 & 0\\
1767: 0 & 1
1768: \end{array} \right),\qquad
1769: \left(\begin{array}{cc}
1770: 1 & 0\\
1771: 0 & -1
1772: \end{array} \right),\qquad
1773: \left(\begin{array}{cc}
1774: 1 & 2\\
1775: 0 & 1
1776: \end{array} \right),\qquad
1777: \left(\begin{array}{cc}
1778: 1 & 0\\
1779: 2 & 1
1780: \end{array} \right).
1781: $$
1782: \end{proposition}
1783:
1784: The proof of this proposition is elementary but repeated below
1785: in the more general case, so we skip it here.
1786: Items (ii), (iv) and (v) follow
1787: from formulas describing the action of the given
1788: generators of $\Gamma^+(C_2,\pi)$ on the bases in $\bar R_1$ and $\bar R_{-1}$.
1789: An important ingredient is Lemma \ref{eli11}. Note that $\Gamma^1(2,2)$ is
1790: generated by the elementary matrices
1791: $$ E_{21}(1)=\left(\begin{array}{cc}
1792: 1 & 0\\
1793: 1 & 1
1794: \end{array} \right),\qquad
1795: E_{12}(2)=\left(\begin{array}{cc}
1796: 1 & 2\\
1797: 0 & 1
1798: \end{array} \right).$$
1799: For $\Gamma^1(2,m)$ ($m\ge 3$) the analogous statement is not true.
1800: Proposition \ref{cyc2} already
1801: implies Corollary \ref{kaz}.
1802:
1803: We turn now to the case of a general cyclic group $C_m=\langle g\rangle$
1804: ($m\in \BN$).
1805: We choose the presentation
1806: $$
1807: \pi:F_n\to C_m,\quad \pi(x)=g,\ \pi(y_1)=\ldots=\pi(y_{n-1})=1.
1808: $$
1809: The corresponding relation module $\bar R$ has the following
1810: $1+m(n-1)$ elements as a
1811: $\BZ$-basis
1812: $$\overline{x^m},\
1813: \overline{x^{k}y_1x^{-k}},
1814: \ldots, \overline{x^{k}y_{n-1}x^{-k}}\qquad (k=0,\ldots,m-1).$$
1815: The rational group ring of $C_m$ decomposes as
1816: \begin{equation}\label{grouri}
1817: \BQ[C_m]=\prod_{d|m}\BQ(\zeta_d).
1818: \end{equation}
1819: Here $\zeta_d\in\BC$ is a primitive $d$-th root of unity. We consider
1820: each of the $\BQ(\zeta_d)$ as a $C_m$-module letting $g$ act by multiplication
1821: with $\zeta_d$. The ring of integers $\BZ(\zeta_d)$ is then a
1822: $\BZ[C_m]$-submodule of $\BQ(\zeta_d)$.
1823: We mimic the decomposition (\ref{grouri}) inside the relation module
1824: by introducing
1825: \begin{equation}
1826: v_i(d,m):=\sum_{k=0}^{m-1}{\rm Tr}(\zeta_d^k)\, \overline{x^ky_ix^{-k}}
1827: \end{equation}
1828: for $i=1,\ldots,n-1$.
1829: Here ${\rm Tr}(\zeta_d^k)$ is the trace of $\zeta_d^k$ taken from
1830: $\BQ(\zeta_d)$ to $\BQ$. Notice that the $v_i(d,m)$ are generalisations of the
1831: generators in (\ref{basi}).
1832: Let $\bar R_d$ be the $C_m$-submodule of $\bar R$ generated by the
1833: $v_i(d,m)$ for $i=1,\ldots,n-1$ if $d>1$ and define
1834: $\bar R_1$ to be the $C_m$-submodule of $\bar R$ generated by the
1835: $v_i(1,m)$ for $i=1,\ldots,n-1$ together with $\overline{x^m}$. By setting
1836: $$\Lambda_{d}(v_i(d,m)):=\left(0,\ldots,0,
1837: \sum_{k=0}^{m-1}{\rm Tr}(\zeta_d^k)\zeta_d^k,0,\ldots,0\right)
1838: \qquad (i=1,\ldots,n-1)$$
1839: (with the non-zero entry in the $i$-th component) for $d>1$ and extending
1840: $C_m$-linearly we obtain well defined additive homomorphisms
1841: $$\Lambda_d:\bar R_d\to \BZ(\zeta_d)^{n-1}\qquad (d|m,\, d>1). $$
1842: The following are true
1843: \begin{itemize}
1844: \item the homomorphism $\Lambda_d$ is $C_m$-equivariant and injective (onto an
1845: ideal in $\BZ(\zeta_d)$),
1846: \item the $\BZ$-rank of $\bar R_d$ is $(n-1)\varphi(d)$
1847: for $d>1$ and $n$ for $d=1$,
1848: \item $\bar R_d\cap \bar R_{d'}=\langle 0\rangle$ for distinct divisors $d,\,
1849: d'$ of $m$,
1850: \item the (direct) sum of all $\bar R_d$ ($d|m$) has finite index in $\bar R$,
1851: \item each $\bar R_d$ ($d|m$) is left invariant by $\Gamma(C_m,\pi)$.
1852: \end{itemize}
1853: As above we transport the action of $\Gamma(C_m,\pi)$ on $\bar R_d$
1854: to the image of $\lambda_d$ and obtain our representations
1855: $$\sigma_{d}: \Gamma(G,\pi)\to \GL(n-1,\BQ(\zeta_d))\qquad (d|m,\, d>1).$$
1856: To control the image of $\sigma_{d}$ we first show:
1857: \begin{proposition}\label{genem}
1858: Let $n$ and $m$ be natural numbers with $n\ge 3$.
1859: The group $\Gamma^+(C_m,\pi)$ is generated by the
1860: automorphisms $\delta_i$, $\epsilon_i$
1861: ($i=1,\ldots,n-1$), the $\lambda_{ij}$ and $\nu_{ij}$
1862: ($i,\, j=1,\ldots ,n-1$, $i\ne j$) and the $\varphi_i^m$,
1863: ($i=1,\ldots,n-1$) and
1864: the elements of the set $T_n$ introduced above.
1865: \end{proposition}
1866: \begin{proof}
1867: Since every element of $\Gamma(C_m,\pi)$ has
1868: to leave the normal closure of
1869: $x^m$ in $F_n$ invariant
1870: and has to fix $x$ modulo $R$
1871: we see that $\rho_1(\Gamma^+(C_m,\pi))$ has to be
1872: contained in $\Gamma^1(n,m)$ (for $\rho_1$ see (\ref{furep})).
1873: Examining the action of elements in the
1874: statement of the proposition on $F_n/F_n'$ and using Lemma \ref{eli11} we find
1875: that $\rho_1(\Gamma^+(C_m,\pi))=\Gamma^1(n,m)$.
1876:
1877: Now our set of generators contains enough elements to generate ${\rm IA}(F_n)$
1878: by the theorem of Nielsen and also enough elements to generate the image of
1879: $\Gamma^+(G,\pi)$ in $\SL(n,\BZ)$ by Lemma \ref{eli11}.
1880: The proposition follows.
1881: \end{proof}
1882: Finally we can state:
1883: \begin{proposition}\label{cycfin}
1884: Let $n$ and $m$ be natural numbers with $n\ge 2$.
1885: The image of the representation $\sigma_{d}$ ($d|m,\, d>1$)
1886: is $\GL^+(n-1,\BZ(\zeta_d))$
1887: where $\GL^+(n-1,\BZ(\zeta_d))$ is the subgroup consisting of those elements in
1888: $\GL(n-1,\BZ(\zeta_d))$ which have a power of $\zeta_d$
1889: as determinant.
1890: \end{proposition}
1891: \begin{proof}
1892: Assume first that $n\ge 3$.
1893: Evaluating $\sigma_{d}$ on the generators from Proposition
1894: \ref{genem} we see that $\sigma_{d}(\Gamma^+(C_m,\pi))$ is contained in
1895: $\GL^+(n-1,\BZ(\zeta_d))$.
1896: We now show that all elementary matrices $E_{ij}(\zeta_d^k)$
1897: ($i,\, j=1,\ldots n-1,\, k\in \BZ$) are in the image of $\sigma_{d}$. This is
1898: done by evaluating $\sigma_{d}$ on some of the generators from Proposition
1899: \ref{genem}. We report only a special case: Take $\tau_{ij}$ for a pair
1900: $(i,j)$ with $i\ne j$. We have
1901: $\tau_{ij} (y_i)=y_i[x,y_j]$. This leads to
1902: $$\bar\tau_{ij}(v_i(d,m))=v_i(d,m)+gv_j(d,m)-v_j(d,m)$$
1903: which in turn shows that $E_{ij}(\zeta_d-1)$ is in the image of
1904: $\sigma_{d}$. Since $E_{ij}(1)$ is also present (use $\lambda_{ij}$) we have
1905: shown that $E_{ij}(\zeta_d)$ is contained in the image of
1906: $\sigma_{d}$.
1907:
1908: After more elementary considerations like that we can show that all
1909: the elementary matrices with entries in $\BZ(\zeta_d)$ are in the image of
1910: $\sigma_{d}$. These matrices generate $\SL(n-1,\BZ(\zeta_d))$: For $n\ge 4$
1911: this follows from results in \cite{Vas}, for $n=3$ and $d\ge 5$ we apply the
1912: main result of \cite{Vas1}, in case $n=3$ and $d\le 4$ the ring $\BZ(\zeta_d)$
1913: is euclidean, in case $n=2$ nothing has to be proved.
1914:
1915: As a last step we use the $\varphi_i^{-1}\circ \psi_i$ ($i=1,\ldots, n-1$) to
1916: show that the image of $\sigma_d$ contains a matrix with determinant equal to
1917: $\zeta_d$. This finishes the proof for $n\ge 3$.
1918:
1919: Let us finally discuss the case $n=2$ which seems trivial, but in fact is
1920: not. Note that $\GL(1,\BZ(\zeta_d))$ contains elements of infinite order
1921: whenever $d\ge 5$, $d\ne 6$ holds. Since we do not have generators for
1922: $\Gamma^+(2,m)$ ($m\ge 3$) at hand we cannot directly rule out the
1923: possibility that such an element lies in the image of $\sigma_d$. We argue as
1924: follows. Let $\varphi\in \Gamma^+(2,m)$ be such that for some divisor $d$ of
1925: $m$ the image $\sigma_d(\varphi)$ has a determinant of infinite order. We
1926: extend $\varphi$ to an element of $\psi$ of $\Gamma^+(3,m)$ by setting
1927: $\psi(x):=\varphi(x)$, $\psi(y_1):=\varphi(y_1)$, $\psi(y_2):=y_2$. Since the
1928: determinants of $\sigma_d(\varphi)$ and $\sigma_d(\psi)$ coincide we have
1929: finished also the case $n=2$.
1930: \end{proof}
1931: Proposition \ref{cycfin} allows us to conclude the following result concerning
1932: the abelianisations of the images of $\Gamma(C_m,\pi)$ under the
1933: representations $\sigma_d$ ($d|m,\, d>1$).
1934: \begin{corollary}\label{cycfincor}
1935: Let $n$ and $m$ be natural numbers with $n\ge 2$.
1936: Then $\sigma_{d}(\Gamma(C_m,\pi))$ ($d|m,\, d>1$) has finite abelianisation.
1937: \end{corollary}
1938: \begin{proof}
1939: It is well known that $\GL^+(n-1,\BZ(\zeta_d))$ has finite abelianisation for
1940: all $d\in\BN$: For $n=2$ this fact is obvious,
1941: for $n=3$ and $d=2,\, 3,\, 4,\, 6$ a presentation of
1942: $\SL(2,\BZ(\zeta_d))$ (see \cite{EGM}) can be used, for $d=5$ or $d\ge 7$ this
1943: fact is contained in \cite{Serresl}.
1944: For other values of $n,\, d$ the group $\GL^+(n-1,\BZ(\zeta_d))$ has Kazdhan's
1945: property (T).
1946: Since $\sigma_{d}(\Gamma(C_m,\pi))$ contains
1947: $\sigma_{d}(\Gamma^+(C_m,\pi))$ as a subgroup of finte index the result
1948: follows. \end{proof}
1949:
1950: We finally remark that
1951: $$\sigma_{d} ({\rm IA}(F_n))\le \GL^+(n-1,\BZ(\zeta_d))\qquad (d|m,\, d>1)
1952: $$
1953: is equal to the full congruence subgroup of $\GL^+(n-1,\BZ(\zeta_d))$ with
1954: respect to the ideal of $\BZ(\zeta_d)$ generated by $\zeta_d-1$. This can be
1955: shown by a more detailed analysis of the arguments involved in the proof of
1956: Proposition \ref{cycfin}.
1957:
1958: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1959: \section{Completion of proofs}\label{Proofs}
1960: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1961:
1962: In this section we complete the proofs of all the results promised in the
1963: introduction.
1964:
1965:
1966: Theorem \ref{prfit} proves Theorem \ref{teo}. Both theorems were formulated
1967: only for $n\ge 4$ but our method applies also to the
1968: case $n=3$. The only obstacle for $n=3$ is that congruence elementary matrices
1969: in $\SL(2)=\SL(n-1)$ do not always generate a finite index subgroup, for
1970: example the matrices
1971: $$ \left(\begin{array}{cc}
1972: 1 & 3\\
1973: 0 & 1
1974: \end{array} \right),\qquad
1975: \left(\begin{array}{cc}
1976: 1 & 0\\
1977: 3 & 1
1978: \end{array} \right)$$
1979: do not generate a subgroup of finite index in $\SL(2,\BZ)$. By the result of
1980: Vaserstein \cite{Vas1} this is a real obstacle (in the case of commutative
1981: rings of endomorphisms) only for $\SL(2,\BZ)$ and $\SL(2)$ of a ring of
1982: integers in an imaginary quadratic number field. Thus our results still work
1983: for $n=3$ in many cases. For example
1984: if the
1985: group algebra $\BQ[G]$ of the finite group $G$ has
1986: (except for the trivial module) no irreducible module $N$ with
1987: ${\rm dim}_D(N)=1$ ($D={\rm End}_G(N)$). If there are such modules $N$ but
1988: if in each case $D$ is a number field which is not $\BQ$
1989: or imaginary quadratic we are also fine. The other cases remain open.
1990:
1991: \smallskip
1992: \begin{prof} {\it of Theorem \ref{teo4}:} $n=2$: In this case we
1993: just consider the (surjective) representation
1994: $$\rho_1 : \A(F_2)\to \A(F_2/F'_2)\cong \GL(2,\BZ).$$
1995: It is well known that $\GL(2,\BZ)$ contains a free subgroup
1996: of any rank ($\ge 2$) of finite index. Hence any finitely generated group
1997: is the image of a subgroup of finite index in $\A(F_2)$. Note that the
1998: groups appearing in Theorem \ref{teo4} as image groups are finitely generated
1999: since they are arithmetic groups.
2000:
2001: $n=3$: Here we use the representation
2002: $\sigma_{-1}:\Gamma^+(C_2,\pi)\to \GL(2,\BZ)$
2003: from Section \ref{Choose} to find
2004: for every $r\in \BN$ with $r\ge 2$ a subgroup of finite index in $\A(F_3)$
2005: which can be mapped onto the free group of rank $r$. We then argue as in the
2006: case $n=2$.
2007:
2008: $n\ge 4$: We first consider the case $k=1$. For $h,\, m\in\BN$ we have to find
2009: a subgroup $\Gamma\le \A(F_n)$ of finite index and a representation
2010: $$
2011: \rho : \Gamma \to \SL((n-1)h,\BQ(\zeta_{m}))^{m}
2012: $$
2013: such that $\rho(\Gamma)$ is commensurable with
2014: $\SL((n-1)h,\BZ(\zeta_{m}))^{m}$. To do this we take the group
2015: $$G:=S_{h+1}\times C_m\times C_m.$$
2016: As an input for Theorem \ref{prfit} we need suitable $\BQ[G]$-modules. We let
2017: $PM_{h+1}$ be the standard $h+1$ dimensional $\BQ[S_{h+1}]$-permutation
2018: module and let $M_h$ be the kernel of the augmentation map from
2019: $PM_{h+1}$ to $\BQ$. The $\BQ[S_{h+1}]$-module $M_h$ is
2020: irreducible and has $\BQ$-dimension
2021: $h$. We put $N:=\BQ(\zeta_m)\otimes_\BQ M_h$ and consider $N$ as a
2022: $\BQ$-vector space. Here, as before, $\zeta_m\in\BC$ is a primitive $m$-th
2023: root of unity. Let $\epsilon : C_m\times C_m\to \langle \zeta_m\rangle$ be a
2024: surjective homomorphism. We turn the $\BQ$-vector space
2025: $N=\BQ(\zeta_m)\otimes_\BQ M_h$ into a $\BQ[G]$-module $N_\epsilon$ by
2026: letting $S_{h+1}$ act on $M_h$ and letting $g\in C_m\times C_m$ act by
2027: multiplication by $\epsilon(g)$ on $\BQ(\zeta_m)$. The following are clear
2028: for every surjective homomorphism
2029: $\epsilon : C_m\times C_m\to \langle \zeta_m\rangle$.
2030: \begin{itemize}
2031: \item ${\rm dim}_\BQ(N_\epsilon)=\varphi(m)h$,
2032: \item ${\rm End}_G(N_\epsilon)=\BQ(\zeta_m)$,
2033: \item ${\rm dim}_{\BQ(\zeta_m)}(N_\epsilon)=h$,
2034: \item if $\epsilon_1,\,\epsilon_2: C_m\times C_m\to \langle \zeta_m\rangle$
2035: are two surjective homomorphisms with distinct kernels then $N_{\epsilon_1}$
2036: and $N_{\epsilon_2}$ are not isomorphic as $\BQ[G]$-modules.
2037: \end{itemize}
2038: Our group $G$ can be generated by $3$ elements: take $g_1,\, g_2,\, g_3\in
2039: S_{h+1}$ such that $g_3$ and the commutator of $g_1,\, g_2$ generate
2040: $S_{h+1}$ and consider the elements
2041: $(g_1,1,0)$, $(g_2,0,1)$, $(g_3,0,0)$ in $G$. Let $\pi :F_n\to G$ be a
2042: redundant presentation of $G$ (remember the assumption is $n\ge 4$).
2043: Since the number of surjective homomorphisms
2044: $\epsilon : C_m\times C_m\to \langle \zeta_m\rangle$ with pairwise distinct
2045: kernels is bigger or equal to $m$ we may apply Theorem \ref{prfit} to obtain
2046: a subgroup $\Gamma \le \A(F_n)$ and a representation
2047: $$\sigma :\Gamma\to \GL((n-1)h,\BQ(\zeta_m))^m$$
2048: such that $\sigma(\Gamma)\cap \SL((n-1)h,\BZ(\zeta_m))^m$ is of finite index
2049: in $\SL((n-1)h,\BZ(\zeta_m))^m$. Here we do not know whether
2050: $\sigma(\Gamma)\cap \SL((n-1)h,\BZ(\zeta_m))^m$ is of finite index in
2051: $\sigma(\Gamma)$. However,
2052: there is a subgroup $\Delta\le \GL((n-1)h,\BZ(\zeta_m))^m$ of finite index
2053: such that $ \SL((n-1)h,\BZ(\zeta_m))^m\le \Delta$ and such that
2054: $\Delta\le Z\cdot \SL((n-1)h,\BZ(\zeta_m))^m$ with a subgroup
2055: $Z\le \GL((n-1)h,\BZ(\zeta_m))^m$ which is central and satisfies
2056: $Z\cap \SL((n-1)h,\BZ(\zeta_m))^m=\langle 1\rangle$. Let $\Gamma_0\le \Gamma$
2057: be the inverse image of $\Delta$ under $\sigma$. The representation
2058: $$\sigma_0: \Gamma_0\to \Delta/Z=\SL((n-1)h,\BZ(\zeta_m))^m$$
2059: satisfies the requirements for $k=1$. Another way of looking at the last
2060: construction is to project a finite index subgroup of the image of
2061: $\sigma(\Gamma)$ into ${\rm PGL}(n-1,\BZ(\zeta_m))$ which contains
2062: ${\rm PSL}(n-1,\BZ(\zeta_m))$ as a finite index subgroup. We then intersect
2063: this subgroup in ${\rm PGL}(n-1,\BZ(\zeta_m))$ with
2064: ${\rm PSL}(n-1,\BZ(\zeta_m))$ and pull the resulting subgroup of
2065: ${\rm PSL}(n-1,\BZ(\zeta_m))$ back to obtain a finite index subgroup of
2066: $\Gamma$ which is mapped into ${\rm SL}(n-1,\BZ(\zeta_m))$.
2067:
2068: We turn now to the general case. Let $\Gamma_1,\ldots,\Gamma_k\le \A(F_n)$
2069: be the subgroups of finite index and
2070: $$\sigma_i: \Gamma_i\to \SL((n-1)h_i,\BZ(\zeta_{m_i}))^{m_i}\qquad
2071: (i=1,\ldots,k)$$
2072: be the representations constructed above. Set
2073: $\Gamma:=\Gamma_1\cap\ldots\cap\Gamma_k$ and consider the representation
2074: $$\sigma: \Gamma\to \prod_{i=1}^k\SL((n-1)h_i,\BZ(\zeta_{m_i}))^{m_i}$$
2075: which maps an element $\gamma\in\Gamma$ to the tuple
2076: $(\sigma_1(\gamma),\ldots,\sigma_k(\gamma))$. The image of $\Gamma$ projects
2077: onto an arithmetic subgroup in each of the factors of the direct product.
2078: We finish the proof by applying the following lemma.
2079: \begin{lemma}\label{Margu}
2080: Let ${\cal H}_1,\ldots ,{\cal H}_k$ be pairwise non-isogeneous
2081: $\BQ$-defined simple linear algebraic groups which all have real rank greater
2082: or equal to $2$. Let
2083: $$\Gamma\le \prod_{i=1}^k{\cal H}_i^{m_i}(\BQ)$$
2084: be a subgroup such that all projections into the factors ${\cal
2085: H}_i^{m_i}(\BQ)$ are arithmetic groups. Then $\Gamma$ is an arithmetic
2086: subgroup of $\prod_{i=1}^k{\cal H}_i^{m_i}$, that is $\Gamma$ is commensurable
2087: with $\prod_{i=1}^k{\cal H}_i^{m_i}(\BZ)$.
2088: \end{lemma}
2089: This lemma is proved using Margulis-superrigidity. See \cite{L} item (iv) on
2090: pages 328-329 or
2091: \cite{GP} Section 2.
2092: \end{prof}
2093:
2094: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2095: \section{(SL) or not (SL)?}\label{SL}
2096: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2097:
2098: In this section we show some results concerning
2099: the question whether $\rho_{G,\pi}(\Gamma(G,\pi))\cap{\cal G}^1_{G,\pi}$ is of
2100: finite index in $\rho_{G,\pi}(\Gamma(G,\pi))$ or not (for the notation see the
2101: introduction). The relevance of this
2102: question and the notation are explained in the introduction. We change
2103: here from
2104: the consideration of the rational relation module to its complex version.
2105: This brings certain technical advantages and still
2106: reflects the problem.
2107:
2108: Let us set up some notation for this section. Let $G$ be a finite group and
2109: $\pi:F_n\to G$ an epimorphism with kernel $R$. We define $\Gamma(G,\pi)$,
2110: ${\cal G}_{G,\pi}$ and ${\cal G}^1_{G,\pi}$ as in the introduction and
2111: consider the action of $\Gamma(G,\pi)$ now on the complex relation module
2112: $\BC\otimes_\BZ\bar R$. The result of Gasch\"utz now says that
2113: $$\BC\otimes_{\bb Z}\bar R=\BC\oplus \BC[G]^{n-1}$$
2114: as $\BC[G]$-modules.
2115:
2116: If $Q$ is an irreducible $\BC[G]$-module and
2117: $M$ is an arbitrary $\BC[G]$-module we keep the notation ${\bf I}_Q(M)$
2118: for the $Q$-isotypic component inside $M$. We hope that this creates no
2119: confusion with the ${\bf I}_N(M)$ of Section \ref{sura} where $N,\, M$ are
2120: $\BQ[G]$-modules.
2121:
2122: Let $Q$ be an non-trivial irreducible $\BC[G]$-module of complex dimension
2123: $h_Q$. Let ${\cal G}_{G,\pi,Q}$ be the stabiliser of
2124: $I_Q(\BC\otimes_\BZ\bar R)$ inside ${\cal G}_{G,\pi}$ (see \ref{stabi}).
2125: By our usual procedure we obtain a representation
2126: \begin{equation}
2127: \rho_{G,\pi,Q}:\Gamma(G,\pi)\to {\cal G}_{G,\pi,Q}=\GL(n-1,M(h_Q,\BC))=
2128: \GL((n-1)h_Q,\BC).
2129: \end{equation}
2130: We call the pair $(G,Q)$ of {\it type (SL)} if
2131: $\rho_{G,\pi,Q}(\Gamma(G,\pi))\cap \SL((n-1)h_Q,\BC)$ is of finite index in
2132: $\rho_{G,\pi,Q}(\Gamma(G,\pi))$ for every $n\in \BN$ and every epimorphism
2133: $\pi:F_n\to G$. We say the finite group $G$ is of {\it type (SL)}
2134: if for all non-trivial irreducible $\BC[G]$-module $Q$ the pair
2135: $(G,Q)$ is of type (SL).
2136:
2137: Note that, since $\Gamma(G,\pi)$ is finitely generated,
2138: $\rho_{G,\pi,Q}(\Gamma(G,\pi))\cap \SL((n-1)h_Q,\BC)$ is of finite index in
2139: $\rho_{G,\pi,Q}(\Gamma(G,\pi))$ if and only if $\rho_{G,\pi,Q}(\Gamma(G,\pi))$
2140: contains no element with a determinant of infinite order.
2141:
2142: Note that the above implies that once we find a finite group $G$
2143: which is not of type (SL), we have found a subgroup $\Gamma$ of finite
2144: index in some $\A(F_n)$ which has an epimorphism
2145: onto $\BZ$.
2146:
2147: Let $G$ be a finite group $G$ and $N_2,\ldots,N_\ell$ its non-trivial
2148: irreducible $\BQ[G]$-modules. Let $\pi:F_n\to G$ be a surjective homomorphism.
2149: The following are equivalent:
2150: \begin{itemize}
2151: \item For all $i=2,\ldots,\ell$ the intersection
2152: $$\sigma_{G,\pi,i}(\Gamma(G,\pi))\cap \SL((n-1)h_i,{\cal R}_i^{\rm op})$$
2153: has finite index in
2154: $\sigma_{G,\pi,i}(\Gamma(G,\pi))\le \GL((n-1)h_i,{\cal R}_i^{\rm op})$.
2155: \item For all non-trivial
2156: irreducible $\BC[G]$-modules $Q$ the intersection
2157: $$\rho_{G,\pi,Q}(\Gamma(G,\pi))\cap \SL((n-1)h_Q,\BC)$$
2158: is of finite index in $\rho_{G,\pi,Q}(\Gamma(G,\pi))$.
2159: \end{itemize}
2160: This is seen by introducing the decomposition of the
2161: complexifications $\BC\otimes_\BQ N_i$ into irreducible $\BC[G]$-modules.
2162:
2163: There are some non-trivial irreducible $\BC[G]$-modules $Q$ which are
2164: obviously of type (SL):
2165: \begin{proposition}\label{aal}
2166: Let $G$ be a finite group and $Q$ a non-trivial irreducible $\BC[G]$-module.
2167: If there is an (irreducible) $\BQ[G]$-module $N$ with $Q=\BC\otimes_\BQ N$
2168: then $(G,Q)$ is of type (SL).
2169: \end{proposition}
2170: \begin{proof}
2171: We note that under the above assumptions the endomorphism algebra
2172: $D:={\rm End}_G(N)$ is a division algebra
2173: over the rational numbers with $\BQ$ as
2174: center. In this case $\SL(m,{\cal R})$ (i.e. the group of elements of
2175: reduced norm $1$ in $M(m,{\cal R})$) is of finite index in
2176: $\SL(m,{\cal R})$ for any order ${\cal R}\le D$ and any $m\in \BN$.
2177: \end{proof}
2178: In case of the symmetric groups $S_m$ ($m\in \BN$) it is known, that every
2179: $\BC[S_m]$-module satisfies the assumptions of Proposition \ref{aal}, see
2180: \cite{S}, Section 13. This implies:
2181: \begin{corollary}
2182: All the symmetric groups $S_m$ are of type (SL).
2183: \end{corollary}
2184: In case of the alternating group $A_5$ Proposition \ref{aal} applies to
2185: the irreducible $\BC[A_5]$-modules of dimensions $4$ and $6$ but not to the
2186: two $3$-dimensional irreducible $\BC[A_5]$-modules. They fit together to form
2187: a $6$-dimensional irreducible $\BQ[A_5]$-module $N$ with
2188: ${\rm End}_G(N)=\BQ(\sqrt{5})$.
2189:
2190:
2191: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2192: \subsection{Reduction Steps}\label{SL2}
2193: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2194:
2195: After the definitions above we discuss some reduction steps which will allow
2196: us to recognise pairs $(G,Q)$ of type (SL).
2197:
2198: \begin{lemma}\label{redu01}
2199: Let $G$ be a finite group with normal subgroup $H_0$ and quotient $H=G/H_0$.
2200: Let $Q$ be a
2201: non-trivial irreducible $\BC[G]$-module on which $H_0$ acts trivially. If
2202: $(H,Q)$ is of type (SL) then $(G,Q)$ also has this property.
2203: \end{lemma}
2204: \begin{proof} Let $\pi:F_n\to G$ ($n\ge 2$) be any epimorphism
2205: and let $R_G$ be the kernel of $\pi$. Let $\pi_0 :F_n\to G\to H$ be the
2206: resulting presentation of $H$ and $R_H$ its kernel. Obviously $R_G$ is a
2207: subgroup of finite index in $R_H$. It is also easily seen that $\Gamma(G,\pi)$
2208: is a subgroup of finite index of $\Gamma(H,\pi_0)$. The inclusion of
2209: $R_G$ into $R_H$ gives rise to a surjective $\BC$-linear map
2210: $$\Theta: \BC\otimes_{\bb Z} \bar R_G\to \BC\otimes_{\bb Z} \bar R_H.$$
2211: This linear map is equivariant for the induced actions of $G$ and
2212: $\Gamma(G,\pi)$.
2213: By restriction we obtain a map
2214: \begin{equation}\label{fro1}
2215: \Theta_Q: {\bf I}_Q(\BC\otimes_{\bb Z} R_G)\to{\bf I}_Q(\BC\otimes_{\bb Z} R_H)
2216: \end{equation}
2217: which is surjective and $H$-equivariant.
2218: Gasch\"utz' result implies that $\Theta_Q$ is in
2219: fact an ($\Gamma(G,\pi)$-equivariant) isomorphism.
2220:
2221: We have to show that ${\rm det}(\rho_{G,\pi,Q}(\varphi))$
2222: is of finite order
2223: for any $\varphi\in\Gamma(G,\pi)$. We compute this determinant on the right
2224: hand side of (\ref{fro1}) where our assumption implies this property.
2225: \end{proof}
2226:
2227: \begin{lemma}\label{redu2}
2228: Let $G$ be a finite group with normal subgroup $H$. Let $Q$ be a
2229: non-trivial irreducible $\BC[G]$-module which is induced from $H$, that is,
2230: $Q$ is of the form $Q={\rm Ind}_H^G(Q_0)$ where $Q_0$ is an irreducible
2231: $\BC[H]$-module. If
2232: $(H,Q_0)$ is of type (SL) then $(G,Q)$ also has this property.
2233: \end{lemma}
2234: \begin{proof} Let $\pi:F_n\to G$ ($n\ge 2$) be any epimorphism
2235: and let $R$ be the kernel of $\pi$. The subgroup
2236: $\pi^{-1}(H)\le F_n$ is a free group on $m=1+[G:H](n-1)$ generators,
2237: we denote it
2238: by $F_m$. We then have a commutative diagram
2239: $$\xymatrix{
2240: 1 \ar[r] & R \ar[r] & F_n \ar[r]^{\pi} & G \ar[r] & 1\\
2241: 1 \ar[r] & R \ar[u]^{=}\ar[r] & F_m \ar[u]\ar[r]^{\pi_0} & H\ar[u]\ar[r] & 1\\
2242: }
2243: $$
2244: where $\pi_0$ is the restriction of $\pi$ to $F_m$. Note that
2245: $R={\rm ker}(\pi)={\rm ker}(\pi_0)$.
2246:
2247: An element $\varphi\in\Gamma(G,\pi)$ stabilises the subgroup $F_m\le F_n$ and
2248: as an automorphism of $F_m$ it lies in $\Gamma(H,\pi_0)$. On the vector space
2249: $\BC\otimes_\BZ\bar R$ the linear maps induced by $\varphi$ as an element of
2250: $\Gamma(G,\pi)$ and by $\varphi$ as an element of $\Gamma(H,\pi_0)$ coincide.
2251: We denote both by $\bar\varphi$.
2252:
2253: The complex relation module $\BC\times_\BZ\bar R$ now has a structure as
2254: a $\BC[G]$-module and a structure of a $\BC[H]$-module. In fact the second is
2255: just the restriction of the first.
2256:
2257: Let $Q=Q_0\oplus Q_1\oplus\ldots \oplus Q_r$ be the decomposition of $Q$ into
2258: irreducible $\BC[H]$-modules
2259: We now prove that
2260: \begin{equation}\label{frob}
2261: {\bf I}_Q(\BC\otimes_{\bb Z}\bar R)=
2262: {\bf I}_{Q_0}(\BC\otimes_{\bb Z}\bar R)\oplus
2263: {\bf I}_{Q_1}(\BC\otimes_{\bb Z}\bar R)\oplus\ldots \oplus
2264: {\bf I}_{Q_r}(\BC\otimes_{\bb Z}\bar R).
2265: \end{equation}
2266: On the left hand side the isotypic component of the $\BC[G]$-module $Q$ is
2267: taken whereas on the right hand side we see the the isotypic component of the
2268: corresponding $\BC[H]$-modules. Clearly the left hand side of (\ref{frob})
2269: is contained in the right hand side.
2270: Note that, since $H$ is normal in $G$ the module $Q$ is also induced from any
2271: of the $Q_0,\ldots, Q_r$.
2272: The Frobenius reciprocity law (\cite{S}, chapter 7)
2273: implies that none of the $Q_0,\ldots, Q_r$ can occur as an irreducible
2274: constituent of the restriction of an irreducible $\BC[G]$-module different
2275: from $Q$.
2276:
2277: We have to show that ${\rm det}(\rho_{G,\pi,Q}(\varphi))$ is of finite order
2278: for any $\varphi\in\Gamma(G,\pi)$. We compute this determinant on the right
2279: hand side of (\ref{frob}) where our assumption implies this property.
2280: \end{proof}
2281:
2282:
2283: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2284: \subsection{Abelian and metabelian groups}\label{SL3}
2285: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2286:
2287: This section contains results which describe the image of $\rho_{G,\pi}$ for
2288: abelian and then for metabelian groups $G$ up to commensurability. In Section
2289: \ref{Choose} we have proved a very precise result in this direction for cyclic
2290: groups $G$. This allowed us to conclude
2291: for $n\ge 2$ that the image of $\Gamma(G,\pi)$
2292: (under certain linear representations) has finite abelianisation.
2293: We present here the following much stronger result for $n\ge 3$. This result
2294: is due to A. Rapinchuk (it appears in a somewhat hidden form in \cite{RA},
2295: page 150), we thank him for the permission to include it.
2296: \begin{proposition}\label{rapi}
2297: Let $n\ge 3$ be a natural number, $G$ a finite abelian group $G$ and let
2298: $\pi:F_n\to G$ be any surjective homomorphism. Then $\Gamma(G,\pi)$ has finite
2299: abelianisation.
2300: \end{proposition}
2301: \begin{proof}
2302: Since $G$ is abelian we have $F_n'\le R$ where $R$ is, as always, the kernel
2303: of $\pi$. Consequently we have ${\rm IA}(F_n)\le \Gamma(G,\pi)$. Writing
2304: $\tilde\Gamma(G,\pi):=\rho_1(\Gamma(G,\pi))\le \GL(n,\BZ)$ for the image of
2305: $\rho_1(\Gamma(G,\pi))$ in $\GL(n,\BZ)$ we have the exact sequence of groups
2306: $$1\to {\rm IA}(F_n)\to \Gamma(G,\pi)\to \tilde\Gamma(G,\pi)\to 1$$
2307: Note that $\tilde\Gamma(G,\pi)$ is a subgroup of finite index in
2308: $\GL(n,\BZ)$. Since $\Gamma(G,\pi)$ is a finitely generated group we need only
2309: show that
2310: $$H^1(\Gamma(G,\pi),\BQ)={\rm Hom}(\Gamma(G,\pi),\BQ)=0$$
2311: We shall do that by applying the edge-term sequence of the Hochschild-Serre
2312: spectral sequence for the above short exact sequence of groups. We shall use
2313: that there is an exact sequence of $\BQ$-vector spaces
2314: \begin{equation}\label{Hose}
2315: H^1(\tilde\Gamma(G,\pi),\BQ)\to H^1(\Gamma(G,\pi),\BQ)\to
2316: H^1({\rm IA}(F_n),\BQ)^{\tilde\Gamma(G,\pi)}
2317: \end{equation}
2318: First of all the left hand side of (\ref{Hose}) is $0$ since $n\ge 3$ and
2319: $\GL(n,\BZ)$ has Kazdhan's property (T).
2320:
2321: Next we evaluate the right hand side of (\ref{Hose}). By a theorem of
2322: Bachmuth (\cite{Ba} and also \cite{F}) there is a $\A(F_n)$-equivariant
2323: isomorphism
2324: $$H^1({\rm IA}(F_n),\BZ)\to {\rm Hom}_{\bb Z}(F_n/F_n',F_n'/\gamma_2(F_n))$$
2325: where $\gamma_2(F_n):=[[F_n,F_n],F_n]$ is the third term of the lower central
2326: series of $F_n$. Let $\GL^\pm (n,\BC)$ be the subgroup of
2327: $\GL (n,\BC)$ consisting of those elements having determinant $\pm 1$. Let
2328: $\BC^n$ be the standard $\GL^\pm (n,\BC)$-module of dimension $n$ and
2329: $\Lambda^2(\BC^n)$ be its outer square. Consider $\GL^\pm (n,\BZ)$ as a
2330: subgroup of $\GL^\pm (n,\BC)$. By identifying $F_n'/\gamma_2(F_n))$ with
2331: $\Lambda^2(F_n/F_n')=\Lambda^2(\BZ^n)$ we obtain an
2332: $\GL^\pm (n,\BZ)$-equivariant
2333: isomorphism
2334: $$\BC\otimes_{\bb Z} {\rm Hom}_{\bb Z}(F_n/F_n',F_n'/\gamma_2(F_n))\to
2335: {\rm Hom}(\BC^n,\Lambda^2(\BC^n)).$$
2336: It is well known (see \cite{F}, page 427)
2337: that the right hand side splits as a $\GL^\pm (n,\BC)$-module
2338: as
2339: $${\rm Hom}(\BC^n,\Lambda^2(\BC^n))=\BC^n\oplus V_n$$
2340: where $V_n$ is a $\GL^\pm (n,\BC)$-module of dimension
2341: ${\rm dim}_\BC(V_n)=n(n+1)(n-2)/2$ which is irreducible even as
2342: $\SL(n,\BC)$-module. Hence the space of $\SL(n,\BC)$-invariants in
2343: ${\rm Hom}(\BC^n,\Lambda^2(\BC^n))$ is $0$. Since the Zariski closure of
2344: $\tilde\Gamma(G,\pi)$ contains $\SL(n,\BC)$ the group
2345: ${\rm Hom}_\BZ(F_n/F_n',F_n'/\gamma_2(F_n))^{\tilde\Gamma(G,\pi)}$ is $0$.
2346: This shows that the right hand side of (\ref{Hose}) is $0$. Together with the
2347: above the proposition is proved.
2348: \end{proof}
2349: Proposition \ref{rapi} is clearly not true for $n=2$, since the left-most term
2350: in (\ref{Hose}) contributes to the $H^1(\Gamma (G,\pi),\BQ)$. See also the
2351: tables in Section \ref{Comp}.
2352: As an immediate corollary from Proposition \ref{rapi} we get:
2353: \begin{corollary}
2354: All finite abelian groups are of type (SL).
2355: \end{corollary}
2356: This corollary can also be deduced from Corollary \ref{cycfincor} (in fact
2357: this corollary is needed in the case $n=2$) and Lemma \ref{redu01}.
2358: But the statement
2359: incorporated in Proposition \ref{rapi} (for $n\ge 3$) is in fact stronger.
2360:
2361:
2362: A metabelian group is a group $G$ with abelian commutator subgroup $G'$.
2363: We prove:
2364: \begin{proposition}
2365: All metabelian finite groups $G$ are of type (SL)
2366: \end{proposition}
2367: \begin{proof}
2368: We use induction on the order of $G$ and the fact proved above that all finite
2369: abelian groups have property (SL).
2370:
2371: Let $Q$ be a non-trivial irreducible $\BC[G]$-module.
2372: We may assume by induction and Lemma \ref{redu01} that $G$ acts faithfully on
2373: $Q$.
2374:
2375: {Case 1:} the commutator subgroup is not contained in the center ${\rm Z}(G)$
2376: of $G$. The restriction of the module $Q$ to $G'$ cannot be
2377: isotypic. Otherwise the abelian subgroup $G'$ would act by scalar matrices on
2378: $Q$ and $G'$ would have to be contained in the center of $G$.
2379: Now Proposition 24 of \cite{S} says that $Q$ is either isotypic or induced.
2380: It follows that
2381: $Q$ is induced from a proper subgroup $H$ of $G$ containing $G'$.
2382:
2383: By our construction $H$ is normal in $G$.
2384: Hence this case is finished by induction and
2385: application of Lemma \ref{redu2}.
2386:
2387: {Case 2:} the commutator subgroup is contained in the center ${\rm Z}(G)$
2388: of $G$. In this case $G$ is nilpotent of class $2$. Unless $G$ is abelian (the
2389: case treated above) we find $g\in G$ with $g\notin {\rm Z}(G)$. The subgroup
2390: $A:=\langle G',g\rangle$ generated by $G'$ and $G$ is abelian, normal in $G$
2391: and not contained in ${\rm Z}(G)$. We then proceed as in case 1.
2392: \end{proof}
2393:
2394:
2395:
2396: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2397: \section{Concluding remarks and suggestions for further research}\label{Concl}
2398: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2399:
2400: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2401: \subsection{Dependence on the presentation}\label{Concl3}
2402: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2403:
2404: The results of the current paper can be considered as a first step toward a
2405: systematic study of the relation module of a finite group as a
2406: $\Gamma(G,\pi)$-module, i.e. an equivariant Gasch\"utz's theory. Here we gave
2407: a quite satisfactory answer in the case of redundant presentations. It is not
2408: clear to what extend this represents the general case.
2409:
2410: For a fixed finite group $G$ let us look at all the possible epimorphisms from
2411: the free group $F_n$ to $G$ and ${\bf R}(n,G)$ the set of their kernels. The
2412: automorphism group $\A(F_n)$ acts on this set. It is easy to see that the
2413: equivariant Gasch\"utz theory that we are trying to develop here depends only
2414: on the orbits of $\A(F_n)$ on ${\bf R}(n,G)$ and not on the actual
2415: presentation. Various authors studied the transitivity properties
2416: of the action of $\A(F_n)$ on ${\bf R}(n,G)$. It is not transitive in general
2417: (a first example was given by B. H. Neumann \cite{Neu})
2418: but it is in some interesting cases. For example, it is transitive if $G$ is a
2419: cyclic group and $n\ge 2$. Thus our results in Section \ref{Choose} give the
2420: full picture for all relation modules of cyclic groups. It is also
2421: transitive if $n>2{\rm log}_2(|G|)$ or if $G$ is solvable and $n>d(G)$, where
2422: $d(G)$ denotes the minimal number of generators of $G$ (see \cite{DU}). An
2423: old conjecture of Wiegold predicts transitivity for finite simple groups if
2424: $n\ge 3$, in which case again $n>d(G)$ holds. For some partial results see
2425: \cite{Gi}, \cite{Ev} and the references in \cite{Pak}.
2426: It was even suggested in \cite{Pak}
2427: that for every finite group $G$ it is transitive if $n>d(G)$.
2428:
2429: In \cite{Ev} it is shown that if a finite group $G$
2430: has spread 2 (i.e. for every
2431: $g_1,\, g_2\in G$ with $g_1\ne 1 \ne g_2$
2432: there is $h\in G$ such that $\langle g_1,\, h\rangle=
2433: \langle g_2,\, h\rangle=G$) then $\A(F_n)$ acts transitively on the set of
2434: kernels of all redundant presentations. Guralnick and Shalev
2435: \cite{GS} proved that almost all
2436: finite simple groups have spread 2.
2437:
2438: Theorem 1.4 is not true in general if the presentation is not redundant, at
2439: least when $n=2$, see the $A_5$-example in Section \ref{Comp}.
2440: The story for $n\ge 3$ may be different as $\A(F_2)$ behaves
2441: in many ways differently from $\A(F_n)$ for $n\ge 3$.
2442:
2443:
2444:
2445: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2446: \subsection{Representations of IA($F_n$)}\label{Concl1}
2447: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2448:
2449: A well known theorem of Formanek and Procesi (see \cite{FP}) asserts
2450: that the automorphism group $\A(F_n)$ has, for $n\ge 3$, no faithful
2451: linear representation over any field. Their proof suggests a stronger
2452:
2453: {\it Conjecture: For every linear representation $\rho$ of $\A(F_n)$ the image
2454: $\rho({\rm Inn}(F_n))$ is virtually solvable.}
2455:
2456: As far as we know, in all previously constructed cases of linear
2457: representations even
2458: $\rho({\rm IA}(F_n))$ had this property.
2459: It should be mentioned that in all the representations $\rho$ studied in this
2460: paper, $\rho({\rm Inn}(F_n))$ is finite but $\rho({\rm IA}(F_n))$ is far from
2461: being virtually solvable and, in fact, our results show that
2462: $\rho({\rm IA}(F_n))$ can be Zariski dense in very large semi-simple groups.
2463: Our reprentations seem to be the first known with this property.
2464: We have:
2465: \begin{proposition}\label{teo3}
2466: Let $n\ge 3$, $k\ge 1$, $h_1<\ldots <h_k$, $m_1,\ldots ,m_k$ be
2467: natural numbers. Let $\BQ(\zeta_{m_i})$ be the field of
2468: $m_i$-th roots of unity and
2469: $\BZ(\zeta_{m_i})$ its ring of integers. There is a
2470: subgroup $\Gamma\le {\rm IA}(F_n)$ of finite index and a representation
2471: $$
2472: \rho : \Gamma \to \prod_{i=1}^k\SL((n-1)h_i,\BQ(\zeta_{m_i}))^{m_i}
2473: $$
2474: such that $\rho(\Gamma)$ is commensurable with
2475: $\prod_{i=1}^k\SL((n-1)h_i,\BZ(\zeta_{m_i}))^{m_i}$.
2476: \end{proposition}
2477: \begin{proof}
2478: The proof of Theorem \ref{teo4} shows that there is a subgroup of finite index
2479: $\Gamma_0\le\A(F_n)$ and a representation
2480: \begin{equation}\label{Mech}
2481: \rho_0 : \Gamma_0 \to \SL(n,\BZ)\times
2482: \prod_{i=1}^k\SL((n-1)h_i,\BQ(\zeta_{m_i}))^{m_i}
2483: \end{equation}
2484: such that the image $\rho_0(\Gamma_0)$ is the internal direct product of its
2485: intersections with the factors in (\ref{Mech}).
2486: These intersections in turn have
2487: finite index in the corresponding factor. Using the fact that a normal
2488: subgroup of a subgroup of finite index in $\SL(n,\BZ)$ ($n\ge 3$) is either
2489: finite or has finite index we conclude the proof of the prposition.
2490: \end{proof}
2491:
2492: Another easy consequence of Theorem \ref{teo4} concerns representations of the
2493: the outer automorphism group ${\rm Out}(F_n)$. We have:
2494: \begin{proposition}\label{teoo}
2495: Let $n\ge 2$, $k\ge 1$, $h_1<\ldots <h_k$, $m_1,\ldots ,m_k$ be
2496: natural numbers. Let $\BQ(\zeta_{m_i})$ be the field of
2497: $m_i$-th roots of unity and
2498: $\BZ(\zeta_{m_i})$ its ring of integers. There is a
2499: subgroup $\Gamma\le {\rm Out}(F_n)$ of finite index and a representation
2500: $$
2501: \rho : \Gamma \to \prod_{i=1}^k\SL((n-1)h_i,\BQ(\zeta_{m_i}))^{m_i}
2502: $$
2503: such that $\rho(\Gamma)$ is commensurable with
2504: $\prod_{i=1}^k\SL((n-1)h_i,\BZ(\zeta_{m_i}))^{m_i}$.
2505: \end{proposition}
2506: Here we only have to note that the image of the inner automorphisms under
2507: any of the representations from Theorem \ref{teo4} is finite. This finite
2508: subgroup of the image can be avoided by going to a subgroup of finite index in
2509: $\Gamma$.
2510:
2511: As a consequence of Propositions \ref{teo3}, \ref{teoo} we get:
2512: \begin{corollary}\label{coro8}
2513: The groups ${\rm IA}(F_3)$ and ${\rm Out}(F_3)$ are large.
2514: \end{corollary}
2515:
2516: Our result from Theorem \ref{teo4} can
2517: conviniently be summarised in the language
2518: of the pro-algebraic completion ${\cal A}(\A(F_n))$ of $\A(F_n)$.
2519: For a definition and
2520: discussion of properties of the pro-algebraic completion (also called the
2521: Hochschild-Mostow group) of a group see
2522: \cite{LM} and \cite {BLMM}. We have
2523: \begin{proposition}\label{teo9}
2524: Let $n\ge 2$ be a natural number and ${\cal S}(\A(F_n))$ be
2525: the semisimple part of the connected component of the identity of the
2526: pro-algebraic completion ${\cal A}(\A(F_n))$. Then for every $h\in \BN$, the
2527: group $\SL((n-1)h,\BC)$ appears infinitely many times as a factor
2528: in ${\cal S}(\A(F_n))$. That is, there is a surjective homomorphism
2529: $$
2530: {\cal S}(\A(F_n))\to \prod_{h=1}^\infty \prod_{i=1}^\infty \SL((n-1)h,\BC).
2531: $$
2532: \end{proposition}
2533: Propositions \ref{teo3}, \ref{teoo} imply similar results for the groups
2534: ${\rm IA}(F_n)$, ${\rm Out}(F_n)$ ($n\ge 3$).
2535:
2536: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2537: \subsection{Representations of $\Gamma(G,\pi)$ into affine groups}\label{Lini}
2538: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2539:
2540: In this section we give another construction of a linear representation
2541: of subgroups of finite index in $\A(F_n)$. This representation takes values in
2542: affine groups (which have a non-trivial unipotent radical). The construction
2543: originates from the proof of Gasch\"utz' result as presented in \cite{JR}.
2544: Besides giving new types of image groups, it is of importance for the
2545: computer algorithms which we have used to create the computational results
2546: described in the next subsection.
2547:
2548: Let $G$ be a finite group and $\pi : F_n\to G$ a surjective homomorphism
2549: of the free group $F_n$ onto $G$. Let $R$ be the kernel of
2550: $\pi$ and $\bar R$ the corresponding relation module. Let further
2551: $\Gamma(G,\pi)$ be the subgroup of $\A(F_n)$ defined in
2552: (\ref{I1}). Given a group $H$ and a commutative ring $S$ we write $S[H]$ for
2553: the corresponding group ring and ${\cal I}(S[H])$ for its augmentation ideal.
2554: If $H_0\le H$ is a subgroup we define ${\cal I}(S[H],H_0)$ to be the two-sided
2555: ideal of $S[H]$ generated by the $h-1$ for $h\in H_0$.
2556:
2557:
2558:
2559: Gasch\"utz' result (\ref{gasch}) is proved (in \cite{JR})
2560: by considering the exact sequence
2561: \begin{equation}\label{gashpr}
2562: 0\to \BQ\otimes_\BZ \bar R\to {\cal I}(\BQ[F_n])/({\cal I}(\BQ[F_n],R)
2563: \cdot{\cal I}(\BQ[F_n]))
2564: \to {\cal I}(\BQ[G])\to 0
2565: \end{equation}
2566: of $G$-equivariant homomorphisms.
2567: The right hand map is induced by $\pi : F_n\to G$ while the left hand map
2568: comes from the map from $R$ to ${\cal I}(\BQ[F_n])$ which sends $r\in R$ to
2569: $r-1$. The free group $F_n$ acts on the middle term in (\ref{gashpr}) by
2570: multiplication from the left. This leads to an action of the
2571: finite group $G$ on this term.
2572:
2573: We now let $\Gamma(G,\pi)$ act on the left hand term as before,
2574: on the middle term
2575: by its action on $F_n$ and trivially on $\BQ[G]$.
2576: It is straightforward to see that the sequence
2577: (\ref{gashpr}) is then $\Gamma(G,\pi)$-equivariant. We obtain a
2578: representation
2579: \begin{equation}\label{gashprep}
2580: \eta_{G,\pi}:\Gamma(G,\pi)\to
2581: {\rm Hom}_G({\cal I}(\BQ[G]),\BQ\otimes_{\bb Z} \bar R)\rfish {\cal
2582: G}_{G,\pi}(\BQ).
2583: \end{equation}
2584: The semi-direct product in (\ref{gashprep}) is formed with respect to the
2585: action of ${\cal G}_{G,\pi}(\BQ)$ on $\BQ\otimes_\BZ \bar R$.
2586:
2587: Our methods show:
2588: \begin{theorem}\label{teo16} Assume $n$ is a natural number with $n\ge 4$.
2589: Let $\pi : F_n\to G$ be a redundant presentation of the finite group $G$. Then
2590: $$\eta_{G,\pi}(\Gamma(G,\pi))\cap
2591: {\rm Hom}_G({\cal I}(\BQ[G]),\BQ\otimes_\BZ \bar R)\rfish
2592: {\cal G}^1_{G,\pi}(\BQ)$$ is
2593: of finite index in the arithmetic group
2594: ${\rm Hom}_G({\cal I}(\BZ[G]),\bar R)\rfish{\cal G}_{G,\pi}^1(\BZ)$.
2595: \end{theorem}
2596: Theorem \ref{teo16} can then be used to prove
2597: \begin{theorem}\label{teo17}
2598: Let $n\ge 2$, $h$, $m$ be
2599: natural numbers. Let $\BQ(\zeta_{m})$ be the field of
2600: $m$-th roots of unity and
2601: $\BZ(\zeta_{m})$ its ring of integers. There is a
2602: subgroup $\Gamma\le \A(F_n)$ of finite index and a representation
2603: $$
2604: \eta : \Gamma \to \left(\BQ(\zeta_m)^{(n-1)h}\rfish
2605: \SL((n-1)h,\BQ(\zeta_{m}))\right)^{m}
2606: $$
2607: such that $\eta(\Gamma)$ is commensurable with
2608: $(\BZ(\zeta_m)^{(n-1)h}\rfish \SL((n-1)h,\BZ(\zeta_{m})))^{m}$.
2609: \end{theorem}
2610:
2611:
2612: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2613: \subsection{Computational results}\label{Comp}
2614: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2615:
2616: An important feature of our methods is that subgroups of quite high indices in
2617: $\A(F_n)$ may be computationally located. We describe here some of the results
2618: of our computer calculations. The computations where done using the computer
2619: algebra system MAGMA.
2620:
2621: We shall describe our approach now in detail for $n=2$. Let
2622: $F_2:=\langle x,\, y\rangle$ be the free group. The automorphism group
2623: $\A^+(F_2)$ is generated by $\alpha,\, \beta$ which are given by
2624: $$\alpha(x)=y^{-1},\quad \alpha(y)=x\qquad \beta(x)=x^{-1}y^{-1},\quad
2625: \beta(y)=x.$$
2626: The group $\A^+(F_2)$ is finitely presented, a presentation is (see
2627: \cite{Neu1})
2628: \begin{equation}\label{neuma}
2629: \A^+(F_2)=\langle\, \alpha,\, \beta \Mid \alpha^4=\beta^3=
2630: \alpha^2\beta^2\alpha^2\beta\alpha\beta\alpha^2\beta^2\alpha=1\,\rangle.
2631: \end{equation}
2632: Let $\pi: F_2\to G$ be a presentation of a finite group. If $\varphi\in
2633: \A^+(F_2)$ is given by its values on $x,\, y$ it is easy to check whether
2634: $\varphi$ is in $\Gamma(G,\pi)$ or not. Using a random process we generate
2635: a list of words $\varphi_1,\ldots, \varphi_k\in \Gamma(G,\pi)$ in the
2636: automorphisms $\alpha,\, \beta$. We then wait until
2637: $$\Delta:=\langle \, \varphi_1,\ldots, \varphi_k\,\rangle$$
2638: has finite index in $\A^+(F_2)$. This is checked using a Todd-Coxeter algorithm
2639: using the presentation (\ref{neuma}). By this approach we first of all find the
2640: following table. The first column contains the finite group $G$, which
2641: we always think of being given as a permutation group. The third and fourth
2642: column contain the images $x,\, y$ under the
2643: surjective homomorphism $\pi:F_2\to G$.
2644: The fifth column gives the index of
2645: $\Delta\le \Gamma(G,\pi)$ in $\A^+(F_2)$. We give
2646: the abelianised group $\Delta^{\rm ab}$ in the last column.
2647: \bigskip
2648: \begin{center}
2649: \begin{tabular}{|c|c|c|c|c|c|}
2650: \hline
2651: $G$ & $|G|$ & $\pi(x)$ & $\pi(y)$ & $[\A^+(F_2):\Delta]$ & $\Delta^{\rm ab}$\\
2652: \hline
2653: $C_2$ & 2 & (1,2) & (1) & 3 & $\BZ^2\times C_2\times C_4$\\
2654: $C_3$ & 3 & (1,2,3) & (1) & 8 & $\BZ\times C_3^2$\\
2655: $C_4$ & 4 & (1,2,3,4) & (1) & 12 & $\BZ^2\times C_4$\\
2656: $C_2\times C_2$ & 4 & (1,2) & (3,4) & 6 & $\BZ^2\times C_2^3$\\
2657: $C_5$ & 5 & (1,2,3,4,5) & (1) & 24 & $\BZ^3\times C_5$\\
2658: $C_6$ & 6 & (1,2,3,4,5,6) & (1) & 24 & $\BZ^3\times C_6$\\
2659: $S_3$ & 6 & (1,2,3) & (1,2) & 18 & $\BZ^2\times C_2$\\
2660: $C_7$ & 7 & (1,2,3,4,5,6,7) & (1) & 48 & $\BZ^5\times C_7$\\
2661: $D_4$ & 8 & (1,2,3,4) & (1,4)(2,3) & 24 & $\BZ^3\times C_2$\\
2662: $Q_8$ & 8 & (1,7,2,8)(3,6,4,5) & (1,4,2,3)(5,7,6,8) & 24 & $\BZ^2\times C_4$\\
2663: $D_5$ & 10 & (1,2,3,4,5) & (1,5)(2,4) & 30 & $\BZ^2\times C_2$\\
2664: $A_4$ & 12 & (1,2,3) & (1,2)(3,4) & 96 & $\BZ^3$\\
2665: $S_3\times C_2$ & 12 & (1,3)(4,5) & (1,2) & 36 & $\BZ^3\times C_2$\\
2666: Sm($12,1$) & 12 & $\sigma_1$ & $\sigma_2$ & 72 & $\BZ^3\times C_2$\\
2667: $A_5$ & 60 & (1,2,3,4,5) & (1,2,3) & 1080 & $\BZ^{17}$\\
2668: \hline
2669: \end{tabular}
2670: \end{center}
2671: The notation for the finite groups is: $C_n$ is the cyclic group of order $n$,
2672: $D_n$ is the dihedral group of order $2n$,
2673: $Q_8$ is the quaternion group of order $8$
2674: and Sm($12,1$) is the non-abelian group of order $12$ not isomorphic to
2675: $A_4$ or $S_3\times C_2$. The two permutations $\sigma_1,\, \sigma_2$ are:
2676: $$\sigma_1:=(1, 8, 4, 11)(2, 9, 5, 12)(3, 7, 6, 10),\qquad
2677: \sigma_2:=(1, 9, 4, 12)(2, 7, 5, 10)(3, 8, 6, 11).$$
2678:
2679: Proceeding with our computer calculations, we then evaluated
2680: the $\varphi_1,\ldots, \varphi_k$ (which now generate a
2681: subgroup of finite index in $\Gamma(G,\pi)$) on the middle term of the sequence
2682: (\ref{gashpr}). Locating the isotypic component of an irreducible
2683: $\BQ[G]$-module $N$ inside all three terms of (\ref{gashpr}) we were able
2684: to compute $\rho_{G,\pi,N}(\varphi_i)$ ($i=1,\ldots,k$) as matrices. Computing
2685: determinants we could decide whether
2686: $\rho_{G,\pi}(\Gamma(G,\pi))\cap {\cal G}_{G,\pi}^1(\BZ)$ is of finite index
2687: in $\rho_{G,\pi}(\Gamma(G,\pi))$ or not.
2688:
2689: We have considered at least one presentation $\pi: F_2\to G$ for
2690: every group $G$ with $|G|\le 60$ which can be generated by 2 elements and
2691: have found that $\rho_{G,\pi}(\Gamma(G,\pi))\cap {\cal G}_{G,\pi}^1(\BZ)$ is
2692: of finite index in $\rho_{G,\pi}(\Gamma(G,\pi))$.
2693: Of course we have run experiments on many more groups
2694: (like $\PSL(2,7)$) but always found a
2695: similar result.
2696:
2697: Let us report some more on the particularly interesting case $G=A_5$.
2698: Let $N$ be the irreducible $\BQ[G]$-module of dimension 6. We have
2699: ${\rm End}_G(N)=\BQ(\sqrt{5})$. Its ring of integers is
2700: ${\cal O}=\BZ((1+\sqrt{5})/2)$.
2701: Identifying the isotypic component of $N$ in $\BQ[G]$ with $M(3,\BQ(\sqrt{5}))$
2702: we obtain a representation
2703: $$\rho: \Gamma(G,\pi)\to \GL(1,M(3,\BQ(\sqrt{5})))=\GL(3,\BQ(\sqrt{5})).$$
2704: Running the programm described above we have found
2705: \begin{proposition}\label{compu18}
2706: For every presentation $\pi: F_2\to G=A_5$ there is a subgroup
2707: $\Delta\le \Gamma(G,\pi)$
2708: of finite index such that $\rho_{A_5,\pi}(\Delta)$
2709: is contained up to change of bases in the subgroup of
2710: $\SL(3,{\cal O})$ consisting of elements which have $(1,0,0)$ as a first row.
2711: \end{proposition}
2712: Proposition \ref{compu18} shows that Theorem \ref{teo} is not true for $n=2$
2713: and
2714: non-redundant presentations. It is not clear what Proposition \ref{compu18}
2715: suggests toward the general case. It may suggest that Theorem \ref{teo} is not
2716: true for non-redundant
2717: presentations but it may also be that $n=2$ is exceptional.
2718:
2719: We have developed similar programms also for the cases $n=3$ and $n=4$.
2720: They lead to the following tables. The notation is the same as in the above
2721: table for $n=2$. The presentations used for $n=4$ take the value $1$ on the
2722: fourth generator of the free group.
2723:
2724: \medskip
2725: \centerline{\it Subgroups in $\A(F_3)$}
2726: \begin{center}
2727: \begin{tabular}{|c|c|c|c|c|c|c|}
2728: \hline
2729: $G$ & $|G|$ & $\pi(x)$ & $\pi(y)$ & $\pi(z)$ & $[\A(F_3):\Delta]$ &
2730: $\Delta^{\rm ab}$ \\
2731: \hline
2732: $C_2$ & 2 & (1,2) & (1) & (1) & 7 & $C_2^3$\\
2733: $C_3$ & 3 & (1,2,3) & (1) & (1) & 26 & $C_6$ \\
2734: $S_3$ & 6 & (1,2,3) & (1,2) & (1) & 168 & $C_2^3$\\
2735: $D_4$ & 8 & (1,3) & (1,4,3,2) & (1) & 336 & $C_2^8$ \\
2736: $Q_8$ & 8 & (1,7,2,8)(3,6,4,5) & (1,4,2,3)(5,7,6,8) & (1) & 336 & $C_2^7$\\
2737: $D_5$ & 10 & (1,2,3,4,5) & (1,5)(2,4) & (1) & 840 & $C_2^5$ \\
2738: $A_4$ & 12 & (1,2,3) & (1,2)( 3,4) & (1) & 1560 & $C_6$ \\
2739: $S_3\times C_2$ & 12 & (1,3)(4,5) & (1,2) & (1) & 1008 & $C_2^9$\\
2740: Sm($12,1$) & 12 & $\sigma_1$ & $\sigma_2$ & (1) & 1344 & $C_2^2\times C_4$\\
2741: $A_5$ & 60 & (1,2,3,4,5) & (1,2,3) & (1) & 200160 & ? \\
2742: \hline
2743: \end{tabular}
2744: \end{center}
2745:
2746: \medskip
2747: \centerline{\it Subgroups in $\A(F_4)$}
2748: \begin{center}
2749: \begin{tabular}{|c|c|c|c|}
2750: \hline
2751: $G$ & $|G|$ & $[\A(F_4):\Delta]$ & $\Delta^{\rm ab}$ \\
2752: \hline
2753: $C_2$ & 2 & 15 & $C_2^2$\\
2754: $C_3$ & 3 & 80 & $C_6$\\
2755: $S_3$ & 6 & 80 & $C_6$ \\
2756: $D_4$ & 8 & 3360 & $C_2^4$ \\
2757: $Q_8$ & 8 & 840 & $C_2^6$ \\
2758: $D_5$ & 10 & 930 & $C_2^3$ \\
2759: $A_4$ & 12 & 1680 & $C_3\times C_6$ \\
2760: $S_3\times C_2$ & 12 & 2730 & $C_2^4$ \\
2761: Sm($12,1$) & 12 & 3120 & $C_2^2\times C_4$\\
2762: $A_5$ & 60 & 213098 & ? \\
2763: \hline
2764: \end{tabular}
2765: \end{center}
2766: We have run experiments on many finite groups $G$ and
2767: presentations $\pi: F_n\to G$, we have always found
2768: that $\rho_{G,\pi}(\Gamma(G,\pi))\cap {\cal G}_{G,\pi}^1(\BZ)$ is
2769: of finite index in $\rho_{G,\pi}(\Gamma(G,\pi))$.
2770:
2771:
2772:
2773:
2774:
2775: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2776: \begin{thebibliography}{99}
2777: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2778:
2779: \bibitem{Ba} S.\, Bachmuth,
2780: {\it Induced automorphisms of free groups and free metabelian groups.}
2781: Trans. Amer. Math. Soc. {\bf 122}, (1966), 1--17
2782:
2783: \bibitem{BLMM} H. Bass, A. Lubotzky, A.R. Magid, S. Mozes,
2784: {\it The proalgebraic completion of rigid groups.}
2785: Geometriae Dedicata {\bf 95}, (2002), 19--58
2786:
2787: \bibitem{Bi} J.\, Birman,
2788: {\it Braids, links and mapping class groups. \/},
2789: Ann. of Math. Stud. {\bf 82}, Princeton Univ. Press, Princeton (1974)
2790:
2791: \bibitem{CR} C.W.\, Curtis, I. Reiner,
2792: {\it Representation Theory of finite Groups and associative Algebras.}
2793: Interscience Publishers (1962)
2794:
2795: \bibitem{D} M. Deuring,
2796: {\it Algebren.}
2797: Ergebnisse {\bf 41}, Springer Verlag, New York, Berlin, Heidelberg (1968)
2798:
2799: \bibitem{DJ} D. Djokovi\'c, V. Platonov,
2800: {\it Low dimensional representations of $\A(F_2)$.}
2801: Manuscripta Math. {\bf 89}, (1996), 475--509
2802:
2803: \bibitem{DU} M. Dunwoody,
2804: {\it Nielsen Transformations}.
2805: In: Computational Problems in Abstract Algebra (1970), Pergamon, Oxford,
2806: 45--46
2807:
2808: \bibitem{EGM} J. Elstrodt, F. Grunewald, J. Mennicke,
2809: {\it Groups acting on Hyperbolic Space: Harmonic Analysis
2810: and Number Theory.}
2811: Springer Monographs in Math., (1998), 524 pp
2812:
2813: \bibitem{Ev} M. Evans,
2814: {\it Presentations of groups involving more generators than are necessary.}
2815: Proc London Math. Soc. {\bf 67}, (1993), 106--126
2816:
2817: \bibitem{Ev1} M. Evans,
2818: {\it T-systems of certain finite simple groups.}
2819: Math. Proc. Cambridge Philos. Soc. {\bf 113}, (1993), 9--22
2820:
2821: \bibitem{F} E. Formanek,
2822: {\it Characterizing a free group in its automorphism group.}
2823: J. Algebra, {\bf 133}, (1990), 424--432
2824:
2825: \bibitem{FP} E. Formanek, C. Procesi,
2826: {\it The automorphism group of a free group is not linear.}
2827: J. Algebra, {\bf 149}, (1992), 494--499
2828:
2829: \bibitem{G} W. Gasch\"utz,
2830: {\it \"Uber modulare Darstellungen endlicher Gruppen, die von freien Gruppen
2831: induziert werden.}
2832: Math. Z. {\bf 60}, (1954), 274--286
2833:
2834: \bibitem{Gi} R. Gilman,
2835: {\it Finite quotients of the automorphism group of a free group.}
2836: Canad. J. Math. {\bf 29}, (1977), 541--551
2837:
2838: \bibitem{Gr} K. Gruenberg,
2839: {\it Relation modules of finite groups.}
2840: CBMS {\bf 25}, Amer. Math. Soc. (1976)
2841:
2842: \bibitem{GLu} F. Grunewald,, A. Lubotzky,
2843: {\it Linear representations of the mapping class group.}
2844: In preparation
2845:
2846: \bibitem{GP} F. Grunewald, V. Platonov,
2847: {\it Rigidity results for groups with radical, cohomology of finite groups and
2848: arithmeticity problems.}
2849: Duke Math. J. {\bf 100}, (1999), 321--358
2850:
2851: \bibitem{GS} R. Guralnick, A. Shalev,
2852: {\it On the spread of finite simple groups.}
2853: Combinatorica {\bf 23}, (2003), 73--87
2854:
2855: \bibitem{JR} M. Jarden, J. Ritter,
2856: {\it Normal Automorphisms of absolute Galois Groups of $\wp$-adic Fields.}
2857: Duke Math. J. {\bf 47}, (1980), 47--56
2858:
2859: \bibitem{L} A. Lubotzky,
2860: {\it Torsion in profinite completions of torsion-free groups.}
2861: Quart. J. Math. Oxford Series {\bf 44}, (1993), 327--332
2862:
2863: \bibitem{LM} A. Lubotzky, A.R. Magid,
2864: {\it Varieties of representations of finitely generated groups.}
2865: Mem. Amer. Math. Soc. {\bf 58}, (1985)
2866:
2867: \bibitem{LP} A. Lubotzky, I. Pak,
2868: {\it The product replacement algorithm and Kazhdan's property (T).}
2869: J. Amer. Math. Soc. {\bf 14}, (2001), 347--363
2870:
2871: \bibitem{MKS} W.\, Magnus, A. Karass, D. Solitar,
2872: {Combinatorial Group Theory.}
2873: Interscience Publishers (1966)
2874:
2875: \bibitem{Neu1} B.H. Neumann,
2876: {\it Die Automorphismengruppe der freien Gruppen.}
2877: Math. Ann. {\bf 107}, (1933), 367--386
2878:
2879: \bibitem{Neu} B.H. Neumann,
2880: {\it On a question of Gasch\"utz.}
2881: Archiv der Math. {\bf 7}, (1956), 87--90
2882:
2883: \bibitem{N} J. Nielsen,
2884: {\it Die Isomorphismengruppe der freien Gruppen.}
2885: Math. Ann. {\bf 91}, (1924), 169--208
2886:
2887: \bibitem{Pak} I. Pak,
2888: {\it What do we know about the product replacement algorithm?}
2889: Groups and Computation III (Columbus, Ohio, 1999), Ohio State
2890: Univ. Math. Res. Inst. Publ. {\bf 8}, 301--347, de Gruyter, Berlin, (2001)
2891:
2892: \bibitem{PLR} V. Platonov, A. Rapinchuk,
2893: {\it Algebraic Groups and Number Theory.} Pure and Applied Mathematics {\bf
2894: 139}, Academic Press, (1994)
2895:
2896: \bibitem{PORA} A. Potapchik, A. Rapinchuk,
2897: {\it Low-dimensional linear representations of $\A(F_n)$, $n\ge 3$.}
2898: Transactions of the Amer. Math. Soc. {\bf 352}, (1999), 1437--1451
2899:
2900: \bibitem{RA} A. Rapinchuk,
2901: {\it $n$-dimensional linear representations of $\A(F_n)$, and more.}
2902: J. of Algebra {\bf 197}, (1997), 146--152
2903:
2904: \bibitem{Serresl} J.-P., Serre,
2905: {\it Le probl\'eme des groupes des congruence pour $\SL_2$.}
2906: Ann. of Math., {\bf 92}, (1970), 489--527
2907:
2908: \bibitem{S} J.-P., Serre,
2909: {\it Linear Representations of Finite Groups.}
2910: Graduate Text in Math. {\bf 42},
2911: Springer Verlag, New York, Berlin, Heidelberg (1993)
2912:
2913: \bibitem{Vas} L.\, Vaserstein,
2914: {\it The structure of classical arithmetic groups of rank greater than one.}
2915: Math. USSR Sbornik (English translation), {\bf 20}, (1973), 465--492
2916:
2917: \bibitem{Vas1} L.\, Vaserstein,
2918: {\it The group $\SL_2$ over Dedekind rings of arithmetic type.}
2919: Mat. Sb. {\bf 89}, (1972), 313--322
2920:
2921: \bibitem{Venk} T.N.\, Venkataramana,
2922: {\it On systems of generators of arithmetic subgroups of higher rank groups.}
2923: Pacific Journal of Math., {\bf 166, 1}, (1994), 193--212
2924:
2925: \end{thebibliography}
2926:
2927:
2928: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2929: \end{document}
2930:
2931:
2932:
2933:
2934:
2935:
2936:
2937: