1: \documentclass[10pt]{amsart}
2: \input{epsf.tex}
3:
4: \usepackage{latexsym,amsfonts,amssymb,epsfig,verbatim}
5: \usepackage{amsmath, latexsym, graphics, textcomp,psfrag}
6: \usepackage{float}
7: %\usepackage{showlabels}
8: \usepackage{amsthm}
9:
10: \usepackage[all,2cell,dvips]{xy}
11: %\usepackage[dvips]{graphicx}
12:
13: %\renewcommand{\topfraction}{0.75}
14: %\renewcommand{\textfraction}{0.1}
15:
16: \setlength{\parskip}{2ex}
17: \setlength{\parindent}{0in}
18:
19: \newcommand{\nc}{\newcommand}
20: \nc{\dmo}{\DeclareMathOperator}
21: \nc{\nt}{\newtheorem}
22:
23: \nc{\ds}{\displaystyle}
24: \nc{\ens}{\ensuremath}
25:
26:
27: \theoremstyle{plain}
28: \nt{thm}{Theorem}[section]
29: \nt*{main}{Main Theorem}
30: \nt{lem}[thm]{Lemma}
31: \nt{cor}[thm]{Corollary}
32: \nt*{cor*}{Corollary}
33: \nt{prop}[thm]{Proposition}
34: \nt*{q}{Question}
35: \nt{fact}{Fact}
36: \nt*{claim}{Claim}
37: \nt{conjecture}[thm]{Conjecture}
38:
39: \theoremstyle{remark}
40: \newtheorem{example}[thm]{Example}
41: \newtheorem{exercise}[thm]{Exercise}
42: \newtheorem{definition}[thm]{Definition}
43: \newtheorem{remark}[thm]{Remark}
44: \newtheorem{definition-remark}[thm]{Definition-Remark}
45: \newtheorem{examples}[thm]{Examples}
46: \newtheorem*{definition*}{Definition}
47: \newtheorem*{remark*}{Remark}
48: \newtheorem{question}[thm]{Question}
49: \newtheorem{obs}[thm]{Observation}
50: \newtheorem{obsdef}[thm]{Observation/Definition}
51:
52:
53: \def\pmf{{\cal PMF}}
54: \def\mf{{\cal MF}}
55: \def\r{{\mathbb R_+}}
56: \def\H{{\mathbb H}}
57: \def\G{{\Gamma}}
58: \def\S{{\Sigma}}
59:
60: \def\C{{\mathcal C}}
61:
62: \nc{\N}{\mathcal{N}}
63:
64: \def\S{{\mathcal S}}
65:
66:
67: \dmo{\out}{Out}
68: \dmo{\aut}{Aut}
69: \dmo{\gl}{GL}
70: \dmo{\SL}{SL}
71: \dmo{\sy}{Sp}
72: \dmo{\mcg}{Mod}
73:
74: \dmo{\stu}{\tilde{St}}
75: \dmo{\st}{St}
76: \dmo{\lku}{\tilde{Lk}}
77: \dmo{\lk}{Lk}
78: \dmo{\dlk}{Lk_<}
79: \dmo{\cdlk}{Lk_\ll}
80:
81: \nc{\T}{\mathcal{T}}
82: \nc{\tm}{\mathcal{M}}
83: \nc{\X}{\mathcal{X}}
84: \nc{\Y}{\mathcal{Y}}
85: \nc{\K}{\mathcal{K}}
86: \nc{\IA}{\textrm{IA}}
87:
88: \nc{\sg}{\Sigma_g}
89: \nc{\modg}{\mcg(\sg)}
90: \nc{\mods}{\mcg(S)}
91: \nc{\spz}{\sy_{2g}(\Z)}
92:
93: \nc{\Z}{\mathbb{Z}}
94: \nc{\R}{\mathbb{R}}
95: \nc{\Q}{\mathbb{Q}}
96: \nc{\I}{\mathcal{I}}
97: \nc{\W}{W}
98:
99: \nc{\outf}{\out(F_n)}
100: \nc{\glz}{\gl_n(\Z)}
101: \nc{\outfn}{\out(F_n)}
102: \nc{\glnz}{\gl_n(\Z)}
103: \nc{\glzw}{W \backslash \! \glz}
104:
105: %\nc{\kpo}{K(\pi,1)}
106: \nc{\kpo}{K(\T_n,1)}
107:
108: \nc{\rt}{\tilde{\rho}}
109: \nc{\strt}{\st(\rt)}
110: \nc{\strto}{\st(\rt_1)}
111: \nc{\strtt}{\st(\rt_2)}
112: \nc{\strtop}{\st(\rt_1')}
113: \nc{\strttp}{\st(\rt_2')}
114: \nc{\lkrt}{\lk(\rt)}
115: \nc{\str}{\st(\rho)}
116: \nc{\lkr}{\lk(\rho)}
117: \nc{\dlkr}{\dlk(\rho)}
118: \nc{\cdlkr}{\cdlk(\rho)}
119:
120: \nc{\li}{L_i}
121: \nc{\lit}{\tilde\li}
122:
123: \nc{\seg}{\mathcal{S}}
124: \nc{\lseg}{\underline{\mathcal{S}}}
125:
126: \nc{\slice}{\Sigma_I}
127: \nc{\bzt}{\ens{{\mathbb Z}^2}}
128:
129: \nc{\norm}[1]{\|#1\|}
130: \nc{\bignorm}[1]{\left\|#1\right\|}
131:
132: \nc{\tbt}[4]{\ensuremath{\left(\begin{array}{rr} #1 & #2 \\ #3 & #4\end{array}\right)}}
133:
134: \nc{\twbo}[2]{\ensuremath{\left(\begin{array}{c} #1 \\ #2 \\ \end{array}\right)}}
135:
136: \nc{\tbo}[3]{\ensuremath{\left(\begin{array}{c} #1 \\ #2 \\ #3 \end{array}\right)}}
137:
138: \nc{\ttt}[9]{\ensuremath{\left(\begin{array}{rrr} #1 & #2 & #3 \\ #4 & #5 & #6 \\ #7 & #8 & #9 \end{array}\right)}}
139:
140: \nc{\p}[1]{\paragraph{{\bf #1}}}
141:
142: \nc{\pics}[3]{\epsfysize=#3 cm \begin{figure}[htb]
143: \center{\leavevmode \epsfbox{#1.eps}} \caption{#2} \label{#1pic}
144: \end{figure}}
145:
146: \nc{\beqn}{\begin{equation}}
147: \nc{\eeqn}{\end{equation}}
148:
149: \nc{\bl}{ \begin{list}{$\cdot$}{
150: \setlength{\leftmargin}{.5in}
151: \setlength{\rightmargin}{.5in}
152: \setlength{\parsep}{0.5ex plus .2ex minus 0ex}
153: \setlength{\itemsep}{0.2ex plus 0.2ex minus 0ex}
154: }
155: }
156: \nc{\blwide}{ \begin{list}{$\cdot$}{
157: \setlength{\leftmargin}{.25in}
158: \setlength{\rightmargin}{.25in}
159: \setlength{\parsep}{0.5ex plus .2ex minus 0ex}
160: \setlength{\itemsep}{0.2ex plus 0.2ex minus 0ex}
161: }
162: }
163:
164: \nc{\el}{\end{list}}
165:
166: \nc{\bpf}{\begin{proof}}
167: \nc{\epf}{\end{proof}}
168:
169: \nc{\si}{\sigma}
170: \nc{\ep}{\epsilon}
171: \nc{\al}{\alpha}
172: \nc{\be}{\beta}
173: \nc{\Ga}{\Gamma}
174: \nc{\la}{\lambda}
175:
176: \nc{\margin}[1]{\marginpar{\scriptsize #1}}
177:
178: \nc{\set}[1]{\{#1\}}
179:
180: \nc{\bc}{\begin{comment}}
181: \nc{\ec}{\end{comment}}
182:
183:
184: \begin{document}
185:
186: \title[Dimension of the Torelli group for $\out(F_n)$]{Dimension of the Torelli group for \boldmath$\out(F_n)$}
187:
188: \author{Mladen Bestvina}
189: \author{Kai-Uwe Bux}
190: \author{Dan Margalit}
191:
192: \address{Mladen Bestvina: Department of Mathematics\\ University of Utah\\ 155 S 1440 East \\ Salt Lake City, UT 84112-0090}
193: \email{bestvina@math.utah.edu}
194:
195: \address{Kai-Uwe Bux: Department of Mathematics\\ University of Virginia\\ Kerchof
196: Hall 229\\ Charlottesville, VA 22903-4137}
197: \email{kb2ue@virginia.edu}
198:
199: \address{Dan Margalit: Department of Mathematics\\ University of Utah\\ 155 S 1440 East \\ Salt Lake City, UT 84112-0090}
200: \email{margalit@math.utah.edu}
201:
202: \thanks{The first and third authors gratefully acknowledge support by
203: the National Science Foundation.}
204:
205:
206: \keywords{$Out(F_n)$, Torelli group, cohomological dimension}
207:
208: \subjclass[2000]{Primary: 20F36; Secondary: 20F28}
209:
210: \maketitle
211:
212: \begin{center}\today\end{center}
213:
214: \begin{abstract}
215: Let $\T_n$ be the kernel of the natural map $\outf \to \glz$. We
216: use combinatorial Morse theory to prove that $\T_n$ has an Eilenberg--MacLane space which is $(2n-4)$-dimensional and that
217: $H_{2n-4}(\T_n,\Z)$ is not finitely generated ($n \geq 3$). In particular, this
218: recovers the result of Krsti\'c--McCool that $\T_3$ is not finitely
219: presented. We also give a new proof of the fact, due to Magnus, that
220: $\T_n$ is finitely generated.
221: \end{abstract}
222:
223: %%%
224: %%%
225: %%%
226:
227: \section{Introduction}
228:
229: There is a natural homomorphism from $\out(F_n)$, the group of outer
230: automorphisms of the free group on $n$ generators, to $\gl_n(\Z)$,
231: given by abelianizing the free group $F_n$. It is a theorem of
232: Nielsen that this map is surjective \cite{jn}. We call its kernel the
233: \emph{Torelli subgroup} of $\out(F_n)$, and we denote it by $\T_n$:
234:
235: \[ 1 \to \T_n \to \out(F_n) \to \gl_n(\Z) \to 1 \]
236:
237: \begin{main} \label{m1}
238: For $n \geq 3$, we have:
239: \begin{enumerate}
240: \item $\T_n$ has a $(2n-4)$-dimensional Eilenberg--MacLane space.
241: \item $H_{2n-4}(\T_n,\Z)$ is infinitely generated.
242: \item \label{fg} $\T_n$ is finitely generated.
243: \end{enumerate}
244: \end{main}
245:
246: Part~(\ref{fg}) of the main theorem is due to Magnus; we give our own
247: proof in Section~\ref{section:finite generation}. We remark that
248: $\T_1$ is obviously trivial and $\T_2$ is trivial by a classical
249: result of Nielsen \cite{jn} (we give a new proof of the latter fact in
250: Section~\ref{section:finite generation}).
251:
252: The group $\T_n$, like any torsion free subgroup of $\out(F_n)$, acts
253: freely on the spine for outer space (see Section~\ref{section:kpo}),
254: and therefore has an Eilenberg--MacLane space of dimension $2n-3$, the dimension of this
255: spine. Our theorem improves this upper bound on the dimension and
256: shows that $2n-4$ is sharp.
257:
258: When $n=3$, we obtain that $H_2(\T_3,\Z)$ is not finitely generated,
259: and this immediately implies the result of Krsti\'c--McCool that
260: $\T_3$ is not finitely presented \cite{km}.
261:
262: \p{Historical background} The question of whether $H_k(\T_n,\Z)$ is finitely generated, for
263: various values of $k$ and $n$, is a long standing problem with few
264: solutions. This question was explicitly asked by Vogtmann in her
265: survey article \cite{kv}. We now give a brief history of related
266: results, all of which are recovered by our main theorem.
267:
268: Nielsen proved in 1924 that $\T_3$ is finitely generated \cite{jn}. Ten years later, Magnus proved
269: that $\T_n$ is finitely generated for every $n$ \cite{wm}.
270:
271: Smillie--Vogtmann proved in 1987 that, if $2 < n < 100$ or $n > 2$ is
272: even, then $H_\star(\T_n,\Z)$ is not finitely
273: generated \cite{sv1} \cite{sv2}. Their method is to consider the
274: rational Euler characteristics of the groups in the short exact
275: sequence defining $\T_n$ (see \cite{kv}).
276:
277: The Krsti\'c--McCool result that $\T_3$
278: is not finitely presented was proven in 1997, via completely algebraic
279: methods \cite{km}.
280: It is a general fact that if the second homology of a group is not
281: finitely generated, then the group is not finitely presented.
282:
283: \p{Large abelian subgroups} It follows from the second part of the
284: main theorem that the first part is sharp; i.e., $\T_n$ does not have
285: an Eilenberg--MacLane space of dimension less than $2n-4$. A simpler proof that the
286: cohomological dimension of $\T_n$ is at least $2n-4$ is to simply
287: exhibit an embedding of $\Z^{2n-4}$ into $\T_n$. There is a subgroup
288: $\Z^{2n-4} \cong G < \T_n$ consisting of elements with representative automorphisms
289: given by:
290: \[
291: \begin{array}{rcl}
292: x_1 & \mapsto & x_1 \\
293: x_2 & \mapsto & x_2 \\
294: x_3 & \mapsto & [x_1,x_2]^{p_3}x_3[x_1,x_2]^{q_3} \\
295: & \vdots & \\
296: x_n & \mapsto & [x_1,x_2]^{p_n}x_n[x_1,x_2]^{q_n}
297: \end{array}
298: \]
299: for varying $p_i$ and $q_i$ (the $x_i$ are generators for $F_n$).
300:
301: In Section~\ref{section:infinite generation}, we prove that specific
302: conjugates of $G$ represent independent classes in
303: $H_{2n-4}(\T_n,\Z)$, thus proving the second part of the main theorem.
304: These conjugates are exactly the generators of $H_{2n-4}(\tm_n,\Z)$,
305: where $\tm_n$, called the ``toy model'', is a particularly simple
306: subcomplex of the Eilenberg--MacLane space $\Y_n$ defined in
307: Section~\ref{section:kpo}. In Section~\ref{section:infinite
308: generation}, we prove that $H_{2n-4}(\tm_n,\Z)$ injects into
309: $H_{2n-4}(\Y_n,\Z)$. Since the homology of $\tm_n$ is not finitely
310: generated in any dimension greater than 1, we are led to the following
311: question:
312:
313: \begin{q}
314: \label{question:middle homology}
315: Does $H_\ast(\tm_n,\Z)$ inject into $H_\ast(\Y_n,\Z)$?
316: \end{q}
317:
318: \p{Mapping class groups.} The term ``Torelli group'' comes from the
319: theory of mapping class groups. Let $\sg$ be a closed surface of
320: genus $g \geq 1$. The \emph{mapping class group} of $\sg$, denoted
321: $\modg$, is the group of isotopy classes of orientation preserving
322: homeomorphisms of $\sg$. The \emph{Torelli group}, $\I_g$, is the
323: subgroup of $\modg$ acting trivially on the homology of $\sg$. As
324: $\modg$ acts on $H_1(\sg,\Z)$ by symplectic automorphisms, $\I_g$ is
325: defined by:
326: \[ 1 \to \I_g \to \modg \to \spz \to 1 \]
327: It is a classical theorem of Dehn, Nielsen, and Baer that the natural
328: map $\modg \to \out(\pi_1(\sg))$ is an isomorphism. In this sense
329: $\T_n$ is the direct analog of $\I_g$.
330:
331: Our (lack of) knowledge of the finiteness properties of $\I_g$ mirrors
332: that for $\T_n$. Using the fact that $\mcg(\Sigma_1) \cong \SL_2(\Z) =
333: \sy_2(\Z)$, it is
334: obvious that $\I_1$ is trivial. Johnson showed in 1983 that $\I_g$ is
335: finitely generated for $g \geq 3$ \cite{dj1}. In 1986,
336: McCullough--Miller showed that $\I_2$ is not finitely generated
337: \cite{mcm}, and Mess improved on this in 1992 by showing that $\I_2$
338: is a free group of infinite rank \cite{gm}. At the same time, Mess
339: further showed that $H_3(\I_3,\Z)$ is not finitely generated. In
340: Kirby's problem list, Mess asked about finiteness properties in higher
341: genus \cite{rk}.
342:
343:
344: \p{Automorphisms vs. outer automorphisms.} Strictly speaking, Magnus
345: and Krsti\'c--McCool study the group $\K_n$, by which we mean the
346: kernel of $\aut(F_n) \to \gl_n(\Z)$, where $\aut(F_n)$ is the
347: automorphism group of the free group. By considering the short
348: exact sequence
349: \[ 1 \to F_n \to \K_n \to \T_n \to 1 \]
350: we see that $\K_n$ is finitely generated if and only if $\T_n$ is
351: finitely generated. Moreover, it follows from our main theorem and the spectral
352: sequence associated to this short exact sequence
353: that $H_{2n-3}(\K_n,\Z)$ is not finitely generated and if $k$ is the
354: smallest index so that $H_k(\T_n,\Z)$ is not finitely generated, then
355: $H_k(\K_n,\Z)$ is not finitely generated.
356:
357: From a topological point of view, $\T_n$ is the more natural group to
358: study.
359:
360: In the literature, $\T_n$ is sometimes denoted by
361: $\IA_n$ for ``identity on abelianization'' (see, e.g. \cite{kv}).
362: However, since Krsti\'c--McCool use $\IA_n$ to denote the kernel of $\aut(F_n)
363: \to \gl_n(\Z)$, we avoid this notation to eliminate the confusion.
364: The notation $\K_n$ comes from Magnus \cite{wm}.
365:
366: \p{Acknowledgements.}
367: We would like to thank Bob Bell, Mikhail Gromov, Jon McCammond, and Kevin
368: Wortman for helpful conversations. We are especially grateful to Karen Vogtmann for
369: explaining her unpublished work.
370:
371:
372: %%%
373: %%%
374: %%%
375:
376: \section{An Eilenberg--MacLane space}
377: \label{section:kpo}
378:
379: In Section~\ref{section:spine}, we recall
380: the definition of Culler--Vogtmann's spine for Outer space. Then, in
381: Section~\ref{section:quotient}, we describe the quotient of this space by
382: $\T_n$. This quotient is a $(2n-3)$-dimensional Eilenberg--MacLane space for $\T_n$.
383:
384: A \emph{rose} is a graph with one vertex. The \emph{standard rose} in
385: rank $n$, denoted $R_n$, is a particular rose which is fixed once and
386: for all. We denote the standard generators of $F_n \cong \pi_1(R_n)$
387: by $x_1, \dots, x_n$.
388:
389: %%%
390:
391: \subsection{Spine for Outer space} \label{section:spine}
392: Culler--Vogtmann introduced the \emph{spine for Outer space}, which we
393: denote by $\X_n$, as a tool for studying $\out(F_n)$ \cite{cv}. This
394: is a simplicial complex defined in terms of marked graphs.
395:
396: A \emph{marked graph} is a pair $(\Ga,g)$, where $\Ga$ is a finite
397: metric graph (1-dimensional cell complex with a metric) with no
398: separating edges and no vertices of valence less than 3 and $g:R_n \to
399: \Ga$ is a homotopy equivalence ($g$ is called the \emph{marking}). We
400: say that two marked graphs $(\Ga,g)$ and $(\Ga',g')$ are
401: \emph{equivalent} if $g' \circ g^{-1}$ is homotopic to an isometry,
402: where $g^{-1}$ is any homotopy inverse of $g$. We will denote the
403: equivalence class $[(\Ga,g)]$ by $(\Ga,g)$.
404:
405: The vertices of $\X_n$ are equivalence classes of marked graphs where
406: all edges have length 1. A set of vertices
407: \[ \set{(\Ga_1,g_1), \dots, (\Ga_k,g_k)} \]
408: is said to span a simplex if $\Ga_{i+1}$ is obtained from $\Ga_i$ by
409: collapsing a forest in $\Ga_i$, and $g_{i+1}$ is the marking obtained
410: from $g_i$ via this operation.
411:
412: We can think of arbitrary points of $\X_n$ as marked metric graphs:
413: for instance, as we move along an edge between two vertices in $\X_n$,
414: the length of some edge in the corresponding graphs (more generally,
415: the lengths of the edges in a forest) varies between 0 and 1.
416:
417: There is a natural right action of $\out(F_n)$ on $\X_n$. Namely,
418: given $\phi \in \out(F_n)$ and $(\Ga,g) \in \X_n$, the action is
419: given by:
420: \[ (\Ga,g) \cdot \phi = (\Ga,g \circ \phi) \]
421: (here we are using the fact that every element $\phi$ of $\out(F_n)$
422: can be realized by a homotopy equivalence $R_n \to R_n$, also denoted
423: $\phi$, uniquely up to homotopy).
424:
425: Culler--Vogtmann proved the following result \cite{cv}:
426:
427: \begin{thm}
428: \label{cv thm}
429: For $n \geq 2$, the space $\X_n$ is contractible.
430: \end{thm}
431:
432: This theorem has the consequence that the virtual cohomological
433: dimension of $\out(F_n)$ is equal to $2n-3$, the dimension of
434: $\X_n$.
435:
436: The \emph{star of a rose} in $\X_n$ is the union of the closed
437: simplices containing the vertex corresponding to a rose. The key idea
438: for Theorem~\ref{cv thm} is to think of $\X_n$ as the union of stars of
439: vertices corresponding to marked roses. We take an analogous approach
440: in this paper.
441:
442: %%%
443:
444: \subsection{The quotient}\label{section:quotient}
445:
446: Baumslag--Taylor proved that $\T_n$ is torsion free \cite{bt}. We
447: also know that the action of $\T_n$ on $\X_n$ is free: by the
448: definition of the action, point stabilizers correspond to graph
449: isometries, and isometries act nontrivially on homology.
450: Finally, the action is simplicial, and so it follows
451: that the quotient of $\X_n$ by $\T_n$ is an Eilenberg--MacLane space for $\T_n$:
452: \[ \Y_n = \X_n/\T_n \]
453:
454:
455: \p{Homology markings.} Since $\out(F_n)$ identifies every pair of
456: isometric graphs of $\X_n$, points of $\Y_n$ can be thought of as
457: equivalence classes of pairs $(\Ga,g)$, where $\Ga$ is a metric graph
458: (as before), and $g$ is a \emph{homology marking}; that is, $g$ is an
459: equivalence class of homotopy equivalences $R_n \to \Ga$, where two
460: homotopy equivalences are equivalent if (up to isometries of $\Ga$)
461: they induce the same map $H_1(R_n,\Z) \to H_1(\Ga,\Z)$.
462:
463: Via the marking $g$, we can think of the (oriented) edges of $\Ga$ as
464: elements of $H^1(R_n) \cong \Z^n$ (if $e$ is an edge and $x$ is a
465: simplicial 1-chain, then $e(x)$ is the number of times $e$
466: appears in $x$). As such, if we think of the generators $x_1,
467: \dots, x_n$ of $\pi_1(R_n)$ as elements of $H_1(R_n,\Z)$, then we can
468: \emph{label} each oriented edge $a$ of $\Ga$ by the corresponding row
469: vector:
470: \[ (a(g(x_1)), \dots, a(g(x_n))) \]
471: where $a(g(x_i))$ is the number of times
472: $g(x_i)$ runs over $a$ homologically, with sign.
473:
474: In this way, a point of $\Y_n$ is given by a labelled graph, and two
475: such graphs represent the same point in $\Y_n$ if and only if there is
476: a label preserving graph isomorphism between them (i.e. if the map
477: induces the identity on cohomology). See Figure~\ref{figure:not
478: frontier} for an example of a labelled graph. We remark that this
479: example exhibits the fact that $\Y_n$ is not a simplicial
480: complex---there are two edge collapses (and hence two edges in $\Y_n$)
481: taking this point to the rose with the identity marking.
482:
483: \begin{figure}[htb]
484: \psfrag{x1}{$(1,0,0)$}
485: \psfrag{x2}{$(0,1,0)$}
486: \psfrag{x3}{$(0,0,1)$}
487: \centerline{\includegraphics[scale=.35]{label}}
488: \caption{An example of a labelled graph.}
489: \label{figure:not frontier}
490: \end{figure}
491:
492: When convenient, we will confuse the points of $\Y_n$ with the
493: corresponding marked graphs.
494:
495: We will make use of the following generalities about marked graphs in $\Y_n$:
496:
497: \begin{prop}
498: \label{lemma1}
499: Let $(\Ga,g)$ be a marked graph.
500: \begin{enumerate}
501: \item \label{labels dont change} If an edge of $\Ga$ is collapsed, the labels of the
502: remaining edges do not change.
503: \item \label{parallel}Any two edges of $\Ga$ with the same label (up
504: to sign) are \emph{parallel}
505: in the sense that the union of their interiors disconnects $\Ga$.
506: \item \label{switch condition} The sum of the labels of the (oriented) edges coming into a
507: vertex of $\Ga$ is equal to the sum of the labels of the edges leaving the
508: vertex.
509: \end{enumerate}
510: \end{prop}
511:
512: We leave the proofs to the reader.
513:
514:
515: \p{Roses.} Let $\Ga$ be a rose with edges $a_1, \dots, a_n$, and let
516: $g:R_n \to \Ga$ be a homology marking. Up to isometries of $\Ga$, the marking $g$ gives an element of
517: $\gl_n(\Z)$, called the \emph{marking matrix}; the rows are exactly the labels of the edges.
518:
519: Since all edges have length 1, the isometry group of $\Ga$ is generated
520: by swapping edges and by reversing the orientations of edges; the
521: former operation has the effect of switching rows of the matrix, and
522: the latter corresponds to changing signs of rows. Thus, in this case,
523: $(\Ga,g)$ gives rise to an element of $\W\backslash\gl_n(\Z)$, where
524: $\W=\W_n$ is the signed permutation subgroup of $\gl_n(\Z)$, acting on
525: the left. In fact, this gives a bijection between roses in $\Y_n$ and
526: elements of $\W\backslash\gl_n(\Z)$, as $\out(F_n)$ acts transitively
527: on the roses of $\X_n$.
528:
529: \medskip
530:
531: The right action
532: of $\out(F_n)$ on $\X_n$ descends to a right action of $\gl_n(\Z)$ on
533: $\Y_n$. In particular, the action on roses is given by the
534: right action of $\gl_n(\Z)$ on $\W\backslash\gl_n(\Z)$.
535:
536: %%%
537: %%%
538: %%%
539:
540: \section{Stars of roses}
541: \label{section:star of a rose}
542: \label{section:frontier}
543:
544: As with $\X_n$, we like to
545: think of the quotient $\Y_n$ as the union of stars of roses. By
546: definition, the \emph{star of a rose} in $\Y_n$ is the image of the
547: star of a rose in $\X_n$. Thus, it consists of graphs which can be
548: collapsed to a particular rose.
549: We now discuss some of the basic properties of the star of a rose.
550:
551: %%%
552:
553: \subsection{Labels in the star of a rose} We will need several observations
554: about the behavior of labels in the star of a rose. The proofs of the
555: various parts of the propositions are straightforward and are left to
556: the reader. In each of the statements, let $\rho$ be a rose in $\Y_n$
557: represented by a marked graph $(\Ga,g)$. Say that its edges $a_1,
558: \dots, a_n$ are labelled by $v_1, \dots, v_n \in \Z^n$.
559:
560: \begin{prop}
561: \label{lemma2}
562: If $(\Ga',g')$ is a marked graph in $\str$, we have:
563: \begin{enumerate}
564: \item \label{vi} For each $i$, there is an edge of $\Ga'$ labelled $\pm v_i$.
565: \item \label{forest}The edges not labelled $\pm v_i$ form a forest.
566: \item \label{circles}If each edge of $\Ga'$ is labelled $\pm v_i$,
567: then the union of edges labelled $\pm v_i$ (for any particular $i$)
568: is a topological circle (see, e.g., Figure~\ref{figure:not frontier}).
569: \item \label{coefficients} The label of any edge of $\Ga'$ is of the form
570: \[ \sum_{i=1}^n k_i v_i \]
571: where $k_i \in \{-1,0,1\}$.
572: \end{enumerate}
573: \end{prop}
574:
575: We have the following converse to the first two parts of the previous proposition:
576:
577: \begin{prop}
578: \label{converse}
579: If $(\Ga',g')$ is a marked graph which has, for each $i$, at least one
580: edge of length 1 labelled $\pm v_i$, then $(\Ga',g')$ is in
581: $\str$.
582: \end{prop}
583:
584: We also have a criterion for when a marked graph is in the frontier of
585: the star of a rose:
586:
587: \begin{prop}
588: \label{lemma:frontier condition}
589: A marked graph $(\Ga',g')$ in $\str$ is in the frontier
590: of $\str$ if and only if it has at least one edge of length 1 whose
591: label is not $\pm v_i$ for any $i$. In this case, the given label is
592: a label for some rose whose star contains $(\Ga',g')$.
593: \end{prop}
594:
595: %%%
596:
597: \subsection{Ideal edges}
598: \label{section:ideal edges}
599:
600: Let $\rho=(\Ga,g)$ be a rose whose edges $a_1, \dots, a_n$ are labelled $v_1, \dots, v_n$,
601: as above. An \emph{ideal edge} is any formal sum:
602: \[ \sum_{i=1}^n k_i a_i \]
603: where $k_i \in \{-1,0,1\}$, and at least two of the $k_i$ are nonzero.
604: An ideal edge is a ``direction'' in $\str$ in the following sense:
605: for any ideal edge, we can find a marked graph $(\Ga',g')$ in the
606: frontier of $\str$ where one of the edges of $(\Ga',g')$ has the label
607: $\sum k_i v_i$. If a marked graph in $\str$ has an edge of length 1
608: with label $\sum k_i v_i$, we say that the marked graph
609: \emph{realizes} the ideal edge $\sum k_i a_i$.
610:
611: \begin{lem}
612: \label{lem:1 edge}
613: Given any ideal edge for a particular rose, there is a 1-edge blowup
614: of $\rho$ in the frontier of $\str$ which realizes that ideal edge.
615: \end{lem}
616:
617: The lemma is proven by example. See Figure~\ref{figure:ideal edge}
618: for a picture of a 1-edge blowup realizing the ideal edge
619: $a_1-a_3+a_4$ in rank 5 (apply Proposition~\ref{lemma1}(\ref{switch
620: condition})). Also, we see that there are many graphs satisfying the
621: conclusion of the lemma---the other graphs are obtained by moving the
622: loops labelled $v_2$ and $v_5$ arbitrarily around the graph.
623:
624: \begin{figure}[H]
625: \psfrag{v1}{$v_1$}
626: \psfrag{v2}{$v_2$}
627: \psfrag{v3}{$v_3$}
628: \psfrag{v4}{$v_4$}
629: \psfrag{v5}{$v_5$}
630: \psfrag{v1v3v4}{$v_1-v_3+v_4$}
631: \centerline{\includegraphics[scale=.5]{1edge}}
632: \caption{A 1-edge blowup realizing the ideal edge $a_1-a_3+a_4$.}
633: \label{figure:ideal edge}
634: \end{figure}
635:
636: Our notion of an ideal edge is simply the homological version of the
637: ideal edges of Culler--Vogtmann \cite{cv}.
638:
639: We say that an ideal edge $\iota'$ is \emph{subordinate} to the ideal
640: edge $\iota = \sum k_i a_i$ if $\iota'$ is obtained by
641: changing some of the $k_i$ to zero.
642: A \emph{2-letter ideal edge} is an ideal edge of
643: the form $k_{i}a_{i} + k_{j}a_{j}$. Two ideal edges are said
644: to be \emph{opposite} if one can be obtained from the other by
645: changing the sign of exactly one coefficient. The following facts are
646: used in Section~\ref{section:finite generation}:
647:
648: \begin{lem}
649: \label{lemma:subordinate}
650: Let $\rho=(\Ga,g)$ be a rose whose edges $a_1, \dots, a_n$ are
651: labelled $v_1, \dots, v_n$. Suppose that $\iota$ and $\iota'$ are
652: ideal edges and that either:
653: \begin{enumerate}
654: \item $\iota'$ is subordinate to $\iota$, or
655: \item $\iota$ and $\iota'$ are 2-letter ideal edges which are not opposite.
656: \end{enumerate}
657: In either case, there is a marked graph $(\Ga',g')$ in $\str$ which
658: simultaneously realizes $\iota$ and $\iota'$.
659: \end{lem}
660:
661: \bpf
662:
663: In each case, we can explicitly describe the desired graph. If
664: $\iota'$ is subordinate to the ideal edge $\iota = \sum k_i
665: a_i$, we start with a 1-edge blowup realizing $\iota$
666: (Lemma~\ref{lem:1 edge}), and then blow
667: up another edge to separate the edges which appear in $\iota'$ from
668: those which do not. Figure~\ref{figure:simultaneous} (left hand side)
669: demonstrates this for $\iota=v_1+v_2+v_3+v_4$ and $\iota' = v_1+v_2$
670: in rank 4.
671:
672: For the case of two 2-letter ideal edges which are not opposite,
673: without loss of generality it suffices to demonstrate marked graphs
674: which simultaneously realize $v_1+v_2$ with $-v_1-v_2$, $v_2+v_3$, or
675: $v_3+v_4$ (the arbitrary case is obtained by renaming/reorienting
676: edges and by attaching extra 1-cells to any vertex). See
677: Figure~\ref{figure:simultaneous} (right hand side) for a
678: demonstration. One can use Proposition~\ref{lemma1}(\ref{switch
679: condition}) to verify the labels.
680: %
681: \epf
682:
683: \begin{figure}[H]
684: \psfrag{v1}{$v_1$}
685: \psfrag{v2}{$v_2$}
686: \psfrag{v3}{$v_3$}
687: \psfrag{v4}{$v_4$}
688: \psfrag{v1v2}{$v_1+v_2$}
689: \psfrag{v2v3}{$v_2+v_3$}
690: \psfrag{v3v4}{$v_3+v_4$}
691: \psfrag{v1v2v3v4}{$v_1+v_2+v_3+v_4$}
692: \psfrag{-v1v2}{$-v_1-v_2$}
693: \centerline{\includegraphics[scale=.5]{compatible}}
694: \caption{Marked graphs simultaneously realizing subordinate ideal
695: edges (left) and 2-letter ideal edges which are not opposite
696: (right).}
697: \label{figure:simultaneous}
698: \end{figure}
699:
700: The reader may verify that opposite ideal edges are never
701: simultaneously realized.
702:
703: We remark that, in the framework established by Culler--Vogtmann, one
704: can think of this lemma in terms of compatibility of partitions, in
705: which case the proof is immediate; see
706: \cite{cv}.
707:
708: %%%
709:
710: \subsection{Homotopy type} In the remainder of this section, we
711: prove that the star of any rose retracts onto the subcomplex consisting
712: of ``cactus graphs'', and this subcomplex is homeomorphic to a union
713: of $(n-2)$-tori.
714:
715: We define a \emph{rank $n$ cactus graph} inductively as follows. A
716: rank 1 cactus graph is a graph with 1 vertex and 1 edge (i.e. a circle
717: with a distinguished point). In general, a rank $n$ cactus graph is
718: obtained by gluing a rank 1 cactus graph to a rank $n-1$ cactus graph
719: along the vertex of the rank 1 cactus graph. The set of vertices of
720: the new graph is the union of the sets of vertices of the original two
721: graphs.
722: We note that a rank $n$ cactus graph has exactly $n$ embedded
723: circles, and every edge belongs to exactly one embedded circle
724: (Figure~\ref{figure:not frontier} is an example).
725:
726: Let $\C(\rho)$ denote the space of cactus graphs in $\str$.
727: Given any $\rho'$, there is a canonical homeomorphism $\C(\rho) \to
728: \C(\rho')$, once we choose orderings of the edges of $\rho$ and
729: $\rho'$. Thus, we can unambiguously use $\C_n$ to denote the space of
730: cactus graphs in the star of a rose in rank $n$.
731:
732: In the remainder, assume that $\rho=(\Ga,g)$ is a rose in $\Y_n$ with
733: edges $a_1, \dots, a_n$ labelled $v_1, \dots, v_n$.
734:
735: \begin{lem}
736: \label{lemma:retraction to cactus graphs}
737: $\str$ strongly deformation retracts
738: onto $\C(\rho)$.
739: \end{lem}
740:
741: \bpf
742:
743: For every marked graph in $\str$, the set of edges
744: whose label is not $\pm v_i$ is a forest
745: (Proposition~\ref{lemma2}(\ref{forest})). We perform a strong deformation
746: retraction of $\str$ by shrinking the edges of each such forest in
747: each marked graph in $\str$.
748:
749: Consider any marked graph $(\Ga',g')$ in the image of the retraction.
750: By Propositions~\ref{lemma2}(\ref{vi}) and~\ref{lemma2}(\ref{circles}),
751: there is a circle of edges labelled $\pm v_i$ for each $i$. We
752: consider the ``dual graph'' obtained by assigning a vertex to each
753: such circle (the \emph{circle vertices}) and each intersection point
754: (the \emph{point vertices}) and we connect a point vertex to a circle
755: vertex if the point is contained in the circle. It follows from
756: Proposition~\ref{lemma1}(\ref{parallel}) that this graph is a tree,
757: and hence $(\Ga',g')$ is a cactus graph.
758: %
759: \epf
760:
761: \begin{cor}
762: \label{star dim}
763: For $n \geq 2$, the star of any rose $\str$ in $\Y_n$ is homotopy equivalent to a complex of dimension $n-2$.
764: \end{cor}
765:
766: \bpf
767:
768: By the definition of cactus graphs, we can see that the dimension
769: increases with slope 1 with respect to dimension, starting at $n=2$.
770: Since $\C_2$ is a point, $\C_n$ is a complex of dimension
771: $n-2$. An application of Lemma~\ref{lemma:retraction to cactus
772: graphs} completes the proof.
773: %
774: \epf
775:
776: We can filter $\C(\rho)$ by subsets according to the
777: number of vertices in the cactus graphs:
778: \[ \set{\rho} = V_0 \subset V_1 \subset \cdots \subset V_{n-2} = \C(\rho) \]
779: Each $V_i$ consists of cactus graphs with $i-1$ vertices.
780:
781: Our goal now is to give a generating set for $\pi_1(\C(\rho))$.
782: Since $V_2$ is simple to understand, the following proposition will
783: make it easy to do this.
784:
785: \begin{prop}
786: \label{prop:cactus space}
787: There is a cell structure on $\C(\rho)$ so that the $i$-skeleton is
788: exactly $V_i$.
789: \end{prop}
790:
791: \bpf
792:
793: We proceed inductively. The 0-skeleton is one point $V_0 = \set{\rho}$.
794:
795: Let $i > 0$. Any marked graph $(\Ga',g')$ in $V_i - V_{i-1}$ lies in
796: a unique $i$-cell $C$. For each $i$, let $k_i$ be the number of
797: edges of $\Ga'$ labelled $\pm v_i$. If we reparameterize so that the sum of the
798: lengths of the edges of $\Ga'$ labelled $\pm v_i$ is 1, then we get a
799: $(k_i-1)$-simplex for each $i$, and $C$ is the product of these simplices.
800:
801: The boundary of $C$ is the set of points where some edge is
802: assigned length 0. Clearly, $\partial C \subset V_{i-1}$. and so
803: the proposition follows.
804: %
805: \epf
806:
807: For the remainder of this section, we use the cell structure given by
808: Proposition~\ref{prop:cactus space}, which is different from the cell
809: structure inherited from $\Y_n$.
810:
811: Let $V_1^1$ be the subset of $V_1$ consisting of graphs with a vertex
812: of valence 4 and a vertex of valence $2n-2$ (i.e. only a single loop
813: is ``travelling'' around another loop). We will see in
814: Section~\ref{section:finite generation} that the obvious generators
815: for $\pi_1(V_1^1)$ correspond to one of the two types of Magnus
816: generators for $\T_n$.
817:
818: \begin{prop}
819: \label{cor:gens for cactus space}
820: The subcomplex $V_1^1$ contains a generating set for $\pi_1(\C(\rho))$.
821: \end{prop}
822:
823: \bpf
824:
825: First, each of $\C_1$ and $\C_2$ is a single point. In
826: rank 3, $V_1^1=V_1$. Thus, in all of these cases, the proposition is
827: vacuously true. For the remainder, assume $n \geq 4$.
828:
829: As per Proposition~\ref{prop:cactus space}, $V_1$ can be thought of as
830: the 1-skeleton of the cell complex $\C(\rho)$. This subcomplex has
831: 1 vertex (the rose $\rho$) and an edge for each combinatorial type of labelled
832: graph with 2 vertices. We now need to show that any such \emph{standard loop} $\alpha$ in
833: $V_1$ can be written in $\pi_1(\C(\rho))$ as a product of loops in
834: $V_1^1$. Our strategy is to show that $V_2$ is a union of 2-tori and
835: that $\pi_1(V_1^1)$ surjects onto $\pi_1(V_2)$.
836:
837: If $(\Ga',g')$ is a point of $V_2-V_1$, then there are two
838: possibilities: the three vertices of $\Ga'$ either lie on the same
839: circle or they do not; see Figure~\ref{figure:v3}. If they do all lie
840: on some ``central circle'', then we obtain a 2-torus by fixing one
841: vertex and letting the other two vertices ``move around'' the central
842: circle (really we are changing lengths so as to give the appearance of
843: this motion). In the other case, there are two central circles. By
844: fixing the middle intersection point and letting the other two
845: intersection points move around the respective circles, we again see a
846: torus.
847:
848: Consider a standard loop $\al$ of $V_1$. At an interior point of $\alpha$,
849: there is a central circle with two vertices, and the two vertices have
850: valence, say, $p=p(\al)$ and $q=q(\al)$. By definition of $V_1$, we
851: have that $p$ and $q$ are even and at least $4$; say $p \leq q$. We thus
852: have a filtration of $V_1$: $\al$ is in $V_1^k$ if $(p-2)/2
853: \leq k$. The number $(p-2)/2$ is the number of loops glued to that vertex,
854: other than the central circle.
855:
856: Now, suppose that $\al$ is a standard loop of $V_1^k$ for some $k
857: \geq 2$. At any interior point of $\al$, we perform a blowup so that
858: we end up with a graph in $V_2-V_1$ of the first type (left side of
859: Figure~\ref{figure:v3}). Moreover, we choose the blowup so that (at
860: least) one of the vertices has valence 4. The fundamental group of the
861: corresponding torus is generated by a standard loop from $V_1^{k-1}$ and a standard loop
862: from $V_1^1$, and so $\al$ can be written as a product of such loops.
863: By induction, $\al$ can be written as a product of loops from $V_1^1$.
864: %
865: \epf
866: \begin{figure}[ht]
867: \includegraphics[scale=.35]{v3.eps}
868: \caption{Two types of graphs in $V_2$.}\label{figure:v3}
869: \end{figure}
870:
871: \p{Remark} For completeness, we mention that the entire space
872: $\C_n$ can be thought of as a union of $(n-2)$-tori, and the
873: intersection between any two of these tori is a lower dimensional
874: torus which is a product of diagonals of coordinate subtori. It is
875: straightforward to prove this, given what we have already done.
876: However, we will not need this fact.
877:
878: %%%
879: %%%
880: %%%
881:
882: \section{Cohomological Dimension}
883: \label{section:morse function}
884:
885: We now give the argument for the first part of the main theorem, that
886: $\Y_n$ is homotopically $(2n-4)$-dimensional. The basic strategy is to
887: put an ordering on the stars of roses of $\Y_n$ (we think of the
888: ordering as a Morse function) and then to glue the stars of roses
889: together in the prescribed order. This is in the same spirit as the
890: proof of Culler--Vogtmann that $\X_n$ is contractible.
891:
892: %%%
893:
894: \subsection{Morse function} The ordering on roses will come from an
895: ordering on matrices. We start with vectors.
896: By the \emph{norm} of an element $v=(a_1,\cdots,a_n)$ of $\Z^n$, we mean:
897: \[ |v|=(|a_1|,\cdots,|a_n|)\in \Z_+^n \]
898: where the elements of $\Z_+^n$ are ordered lexicographically.
899: Consider the matrix:
900: \[ M = \left(\begin{array}{c} v_1 \\ v_2 \\ \vdots \\ v_n
901: \end{array}\right) \]
902: The \emph{norm} of $M$ is:
903: \[ |M| = (|v_n|,\dots,|v_1|) \in (\Z_+^n)^n \]
904: where $(\Z_+^n)^n$ has the lexicographic ordering on the $n$ factors.
905: We say that $M$ is a \emph{standard representative} for an element of
906: $\W\backslash\gl_n(\Z)$ if $|v_n| < \cdots < |v_1|$ (i.e. if it is a
907: representative with smallest norm). Note that two rows of a matrix in
908: $\gl_n(\Z)$ cannot have the same norm, for otherwise these
909: two rows would be equal after reducing modulo 2, and the resulting
910: matrix would not be invertible.
911:
912: We declare the norm of an element of $\W\backslash\gl_n(\Z)$ to be the
913: norm of a standard representative, and the norm of a
914: rose in $\Y_n$ to be the norm of the corresponding element of
915: $\W\backslash\gl_n(\Z)$.
916:
917: In what follows, the following fact will be important:
918:
919: \begin{lem}
920: \label{lemma:neighbors}
921: If the stars of two roses
922: intersect, then the roses have different norms.
923: \end{lem}
924:
925: \bpf
926:
927: If $M$ and $M'$ are marking matrices for neighboring roses, then
928: $M'=NM$, where each entry of $N$ is either $-1$, $0$, or $+1$ (apply Proposition~\ref{lemma2}(\ref{coefficients})).
929: Then, if $|M|=|M'|$, it follows that $N$ is the identity modulo 2, and
930: so $N \in \W$.
931: %
932: \epf
933:
934: The norm on roses turns the set of roses into a well-ordered set. We
935: use this fact without mention in the transfinite induction arguments for Theorem~\ref{thm:dim} and Proposition~\ref{proposition:normal generation}.
936:
937: %%%
938:
939: \subsection{The induction}
940: \label{subsection:the induction}
941:
942: We define an \emph{initial segment} of $\Y_n$ to be a union of stars of
943: a set of roses that is closed under taking smaller roses (i.e. a
944: sublevel set of the ``Morse function'' given on stars of roses). Note
945: that, in general, an initial segment consists of infinitely many
946: roses. If we show that each initial segment is $(2n-4)$-dimensional,
947: it will follow by transfinite induction that $\Y_n$ has the same
948: property.
949:
950: To this end, we define the \emph{descending link} of a rose in $\Y_n$
951: to be the intersection of its star with the union of all stars of
952: roses of strictly smaller norm (by Lemma~\ref{lemma:neighbors}, we need not worry about roses of equal norm). The descending link of a rose $\rho$, denoted $\dlkr$, is a subset of the frontier of its star. We will prove the following in
953: Section~\ref{section:descending links}:
954:
955: \begin{prop}
956: \label{desc link dim}
957: For $n \geq 3$, descending links are homotopically $(2n-5)$-dimensional.
958: \end{prop}
959:
960: Given this, we can prove the first part of the main theorem:
961:
962: \begin{thm}
963: \label{thm:dim}
964: For $n \geq 3$, the complex $\Y_n$ is homotopy equivalent to a complex of dimension at most $2n-4$.
965: \end{thm}
966:
967: \bpf
968:
969: We proceed by transfinite induction on initial segments. The base step is Corollary~\ref{star dim}.
970:
971: Whenever we
972: glue the star of a rose $\str$ to an initial segment $\S$ in order to
973: make a new initial segment $\hat \S$, we can think of this as a
974: diagram of spaces:
975: \[ \S \leftarrow \dlkr \rightarrow \str \]
976: By the inductive
977: hypothesis, $\S$ is homotopy equivalent to a $(2n-4)$-dimensional
978: space $\S'$. Denote by $\str'$ the $(n-2)$-complex homotopy equivalent to $\str$ given by Proposition~\ref{star dim}. By
979: Proposition~\ref{desc link dim}, the descending link $\dlkr$ is homotopy equivalent to a $(2n-5)$-dimensional space
980: $\dlkr'$. We choose maps $\dlkr' \to \S'$ and $\dlkr' \to \str'$ so
981: that the following diagram commutes up to homotopy:
982:
983: \begin{figure}[H]
984: \begin{center}
985: \scalebox{1}{ \xymatrix{\str' \ar@{<->}[r] & \str \\
986: \dlkr' \ar@{->}[d] \ar@{->}[u] \ar@{<->}[r] & \dlkr \ar@{->}[d] \ar@{->}[u] \\
987: \S' \ar@{<->}[r] & \S}}
988: \end{center}
989: \end{figure}
990: It follows that the colimit of the diagram of spaces in the left column is homotopy equivalent to the colimit of the diagram of spaces
991: in the right column (see e.g. \cite[Proposition 4G.1]{ah}). The
992: former, call it $\hat \S'$, is $(2n-4)$-dimensional (consider the double mapping cylinder), and
993: the latter is $\hat \S$. By construction, the homotopy equivalence $\hat S' \to \hat S$ extends the homotopy equivalence $S' \to S$.
994:
995: By transfinite induction, we thus build a homotopy model $Z_n$ for
996: $\Y_n$. By the inductive construction given above, $Z_n$ has a
997: filtration by subcomplexes $\set{Z_n^\al}$, each equipped with a
998: homotopy equivalence $h_\al:Z_n^\al \to \Y_n^\al$ for some some initial
999: segment $\Y_n^\al$. What is more, the induced map $h:Z_n \to
1000: \Y_n$, when restricted to $Z_n^\al$, is precisely $h_\al$. It follows
1001: that $h$ is a homotopy equivalence (see, e.g., the discussion
1002: following \cite[Proposition 4G.1]{ah}). Since $Z_n$ has dimension at
1003: most $2n-4$ (by construction), we are done.
1004: %
1005: \epf
1006:
1007: \p{Remark.} If one wants to avoid transfinite induction, it is
1008: possible to alter the Morse function so that it is the same locally
1009: (i.e. Proposition~\ref{desc link dim} and its proof do not change)
1010: but the image of the Morse function is order isomorphic to the
1011: positive integers.
1012:
1013:
1014: %%%
1015: %%%
1016: %%%
1017:
1018: \section{Finite generation}
1019: \label{section:finite generation}
1020:
1021: In this section, we recall the definition of the Magnus generating set
1022: for $\T_n$, and explain how our point of view recovers the result that
1023: these elements do indeed generate $\T_n$
1024: (Theorem~\ref{proposition:finite generation} below).
1025:
1026: Throughout the section (and the appendix), we denote an element $\phi$ of $\out(F_n)$
1027: by:
1028: \[ [\Phi(x_1), \dots, \Phi(x_n)] \]
1029: where $x_1,\dots,x_n$ are the generators of $F_n$, and $\Phi$ is a
1030: representative automorphism for $\phi$.
1031:
1032: %%%
1033:
1034: \subsection{Magnus generators}
1035:
1036: Magnus proved that $\T_n$ is generated by:
1037: \[ \begin{array}{rcl}
1038: K_{ik} & = & [x_1, \dots, x_kx_ix_k^{-1}, \dots, x_n] \\
1039: K_{ikl} & = & [x_1, \dots, x_i[x_k,x_l], \dots, x_n]
1040: \end{array}\]
1041: for distinct $i$, $k$, and $l$.
1042:
1043: We can see the $K_{ik}$ as loops in the star of a rose in $\Y_n$.
1044: Consider the picture in Figure~\ref{figure:not frontier}. As
1045: mentioned in Section~\ref{section:quotient}, shrinking
1046: either of the parallel edges gives a path leading to the rose with the
1047: identity marking, and so this is a loop in the star of that rose in
1048: $\Y_n$. By considering what is happening on the level of homotopy (as
1049: opposed to homology), we see that this loop is exactly $K_{23}$ (see
1050: \cite{cv}). By attaching more loops at one of the vertices, and
1051: renaming the edges, we see that we can obtain any $K_{ik}$ in the star
1052: of the identity rose. In the stars of other roses, the
1053: analogously defined loops are conjugates of the $K_{ik}$. What is
1054: more, we have:
1055:
1056: \begin{prop}
1057: \label{proposition:generators of star of a rose}
1058: The fundamental group of the star of the rose with the identity
1059: marking is generated by the $K_{ik}$.
1060: \end{prop}
1061:
1062: The proposition follows immediately from the fact that the loops in
1063: the above discussion corresponding to the $K_{ik}$ are exactly the
1064: standard generators for $\pi_1(V_1)$ from Proposition~\ref{cor:gens
1065: for cactus space}.
1066:
1067: %%%
1068:
1069: \subsection{Proof of finite generation}
1070:
1071: Our proof that the Magnus generators generate $\pi_1(Y_n) \cong \T_n$
1072: rests on the following two topological facts about descending links
1073: which we prove in Section~\ref{section:descending links}:
1074:
1075: \begin{prop}
1076: \label{proposition:descending links are nonempty}
1077: Descending links are nonempty, except for that of the rose with the identity marking.
1078: \end{prop}
1079:
1080: \begin{prop}
1081: \label{proposition:descending links are connected}
1082: Descending links are connected.
1083: \end{prop}
1084:
1085: Combining Propositions~\ref{proposition:generators of star of a rose},
1086: \ref{proposition:descending links are nonempty},
1087: and~\ref{proposition:descending links are connected} with Van
1088: Kampen's theorem and the transitivity of the action of $\out(F_n)$ on
1089: stars of roses, we see that the fundamental group of any initial segment of $\Y_n$ is normally generated by the $K_{ik}$. By transfinite induction, we have:
1090:
1091: \begin{prop}
1092: \label{proposition:normal generation}
1093: $\T_n$ is normally generated by the $K_{ik}$.
1094: \end{prop}
1095:
1096: The group generated by the $K_{ik}$
1097: is not normal in $\out(F_n)$, as any element
1098: of this subgroup is of the form:
1099: \[ [g_1x_1g_1^{-1},g_2x_2g_2^{-1},\dots,g_nx_ng_n^{-1}] \]
1100: Thus, to find a generating set for $\T_n$, we need to add more
1101: elements.
1102:
1103: We have the following result of Magnus:
1104:
1105: \begin{prop}
1106: \label{magnus}
1107: For any $n$, the group generated by
1108: \[ \set{K_{ik},K_{ikl}: i \neq k < l \neq i} \]
1109: is normal in $\out(F_n)$.
1110: \end{prop}
1111:
1112: It is now easy to prove the following, which is the third part of our
1113: main theorem:
1114:
1115: \begin{thm}
1116: \label{proposition:finite generation}
1117: $\T_n$ is finitely generated. In particular, it is generated by $\set{K_{ik},K_{ikl}}$.
1118: \end{thm}
1119:
1120: Proposition~\ref{magnus} is also one of the steps in
1121: Magnus's proof that the $K_{ik}$ and $K_{ikl}$ generate $\K_n$
1122: \cite{wm}. For completeness, we give Magnus's proof of
1123: Proposition~\ref{magnus} in the appendix.
1124:
1125: %%%
1126:
1127: \subsection{Proof that \boldmath$\T_2$ is trivial} Since there are two
1128: ways to blow up a rank 2 rose, it follows that the star of a rose in
1129: $\Y_2$ is homeomorphic to an interval and that the frontier is
1130: homeomorphic to $S^0$. If we glue the stars of roses together
1131: inductively according to our Morse function as in
1132: Section~\ref{section:morse function}, then at each stage we are gluing
1133: a contractible space (the star of the new rose) to a contractible
1134: space (the previous initial segment is contractible by induction)
1135: along a contractible space (Propositions~\ref{proposition:descending links
1136: are nonempty} and~\ref{proposition:descending links are connected} and
1137: the fact that the frontier is $S^0$). It follows that each initial
1138: segment, and hence all of $\Y_2$, is contractible; hence, $\T_2=1$.
1139:
1140: It is more illuminating to draw a diagram of $\X_2=\Y_2$. It is a
1141: tree, with edges representing stars of roses. This tree is naturally
1142: dual to the classical Farey graph, with the matrix
1143: $\left(\begin{smallmatrix} a&b\\ c&d\end{smallmatrix}\right)$
1144: corresponding to the unordered pair $\{\frac ba,\frac dc\}$. See
1145: Figure \ref{id18}.
1146:
1147: \begin{figure}[H]
1148: \centerline{\includegraphics{id-det.18}}
1149: \caption{A part of the Farey graph and the dual tree $\Y_2$.}
1150: \label{id18}
1151: \end{figure}
1152:
1153:
1154: %%%
1155: %%%
1156: %%%
1157:
1158: \section{Descending links}
1159: \label{section:descending links}
1160:
1161: Recall that the descending link $\dlkr$ of a rose $\rho$ is the
1162: intersection of its star with the union of stars of roses of strictly
1163: smaller norm. The goal of this section is to prove
1164: Propositions~\ref{proposition:descending links are nonempty},
1165: \ref{proposition:descending links are connected}, and~\ref{desc link
1166: dim}, that descending links are nonempty, connected, and homotopically
1167: $(2n-5)$-dimensional.
1168:
1169: As in Section~\ref{section:frontier}, let $\rho$ be a rose represented by a marked graph $(\Ga,g)$ whose edges $a_i$ are labelled $v_i$. We assume the $a_i$ are ordered so that the marking matrix
1170: \[ M = \left(\begin{array}{c} v_1 \\ \vdots \\ v_n \end{array}\right) \]
1171: is a standard representative.
1172:
1173: %%%
1174:
1175: \subsection{Descending ideal edges}
1176:
1177: An ideal edge for $\rho$ is called \emph{descending} if any of the corresponding
1178: 1-edge blowups (Lemma~\ref{lem:1 edge}) lies in $\dlkr$. Every edge of a marked graph in $\str$ which is not labelled $\pm v_i$ corresponds to some ideal edge; if the corresponding ideal edge is descending, we may say that the edge is descending.
1179:
1180: We now give a criterion for checking whether or not a particular ideal
1181: edge is descending.
1182:
1183: \begin{lem}
1184: \label{descending criterion}
1185: Let $\iota = a_1 + a_{i_1} + \cdots + a_{i_m}$ be an ideal edge. The
1186: following are equivalent:
1187: \begin{enumerate}
1188: \item $\iota$ is descending
1189: \item any of the corresponding 1-edge blowups lies in $\dlkr$
1190: \item all of the corresponding 1-edge blowups lie in $\dlkr$
1191: \item $|v_1 + v_{i_1} + \cdots + v_{i_m}| < |v_1|$
1192: \end{enumerate}
1193:
1194: Similarly, $\bar \iota = -a_1 + a_{i_1} + \cdots + a_{i_m}$ is descending if and only if
1195: $|v_1 - (v_{i_1} + \cdots + v_{i_m})| < |v_1|$.
1196: \end{lem}
1197:
1198: \bpf
1199:
1200: A 1-edge blowup which realizes the ideal edge $\iota$ lies in $m+1$
1201: stars of roses (Proposition~\ref{converse}). Namely, for each of $v_1, v_{i_1}, \dots v_{i_m}$,
1202: we get a new marking matrix by replacing that vector with:
1203: \[ v_1 + v_{i_1} + \cdots + v_{i_m} \]
1204: and leaving all other row vectors the same. To see if $\iota$ is
1205: descending, we look at the smallest of these matrices.
1206: We claim that the smallest is:
1207: \[ N = \left(\begin{array}{c} v_1 + v_{i_1} + \cdots + v_{i_m} \\ v_2 \\ \vdots \\ v_n \end{array}\right) \]
1208: Indeed, suppose we had replaced some other row vector, say $v_i$, with
1209: $v_1 + v_{i_1} + \cdots + v_{i_m}$, obtaining a matrix $N'$. Now,
1210: forgetting the order of the rows, $N$
1211: and $N'$ share $n-1$ rows, and $N$ has the row vector $v_i$ whereas
1212: $N'$ has the row vector $v_1$. By the assumption that $M$ is a
1213: standard representative, we have $|v_i| < |v_1|$. Now, if we put
1214: $|N'|$ in standard form, it is easy to find a representative for the
1215: $N$-coset with smaller norm than the standard representative for
1216: $N'$---simply replace the row of $N'$ consisting of $v_1$ with the
1217: vector $v_i$. The norm of $N$ is less than or equal to the norm of
1218: this representative, so the claim is proven.
1219:
1220: Now both directions are easy: if $|v_1 + v_{i_1} + \cdots + v_{i_m}| <
1221: |v_1|$ then $|N|$ is obviously strictly less than $|M|$ (the given
1222: representative has smaller norm) and so $\iota$ is
1223: descending; conversely, if $|v_1 + v_{i_1} + \cdots + v_{i_m}| \geq
1224: |v_1|$, then the given representative is in standard form and
1225: obviously has norm at least $|M|$. (We remark that the last
1226: inequality must be strict by Lemma~\ref{lemma:neighbors}.)
1227:
1228: The second statement follows by symmetry.
1229: %
1230: \epf
1231:
1232: It is not hard to prove a stronger statement than the one given here. However, the relatively simple result given suffices for our purposes, and the generalities are notationally unpleasant.
1233:
1234: %%%
1235:
1236: \begin{cor}
1237: \label{lemma:description of descending}
1238: A marked graph in $\str$ is in $\dlkr$ if and only if it realizes a descending ideal edge.
1239: \end{cor}
1240:
1241: As a consequence of Lemma~\ref{descending criterion}, we see that
1242: there exist pairs of marked graph which can never be simultaneously
1243: descending.
1244:
1245: \begin{lem}
1246: \label{forbidden pairs}
1247: If the ideal edge $\iota = a_1 + a_{i_1} + \cdots + a_{i_m}$ is descending
1248: then $\bar \iota = -a_1 +
1249: a_{i_1} + \cdots + a_{i_m}$ is not descending.
1250: \end{lem}
1251:
1252: Generalizations of Lemma~\ref{descending criterion} lead to analogous generalizations of the current lemma.
1253:
1254: \bpf
1255:
1256: To simplify notation, let $w_0=v_1$, $w_1 = v_{i_1}$,
1257: $w_2=v_{i_2}$, etc. We will denote particular entries in each of
1258: these row vectors by using double indices; i.e., $w_{jk}$ is the
1259: $k^{\mbox{\tiny th}}$ entry of the row vector $w_j$.
1260:
1261: Let $k$ be the smallest number so that
1262: \[ |w_{0k} + w_{1k} + w_{2k} + \cdots + w_{mk}| \neq |w_{0k}| \]
1263: Note that there is such a $k$, for otherwise, the
1264: original matrix $M$ would not be invertible (reduce modulo 2).
1265:
1266: Applying Lemma~\ref{descending criterion}, we see that $\iota$ is descending if and only if
1267: \begin{equation}\tag{1}
1268: |w_{0k} + w_{1k} + w_{2k} + \cdots + w_{mk}| < |w_{0k}|
1269: \end{equation}
1270: (we are using the minimality of $k$). It follows that $w_{1k} +
1271: w_{2k} + \cdots + w_{mk}\neq 0$ and that the sign of this sum differs
1272: from that of $w_{0k}$. Thus, we have:
1273: \begin{equation}\tag{2}
1274: |w_{0k} - (w_{1k} + w_{2k} + \cdots + w_{mk})| > |w_{0k}|
1275: \end{equation}
1276: and so $\bar \iota$ is not descending. By symmetry, we are done.
1277: %
1278: \epf
1279:
1280:
1281: %%%
1282:
1283: \subsection{Proof of Propositions~\ref{proposition:descending links are nonempty} and~\ref{proposition:descending links are connected}}
1284:
1285: As usual, let $\rho$ be a rose represented by
1286: a marked graph $(\Ga,g)$ with edges $a_1, \dots, a_n$ labelled by
1287: $v_1, \dots, v_n$, and assume that the edges are ordered so that the
1288: marking matrix
1289: \[ M = \left(\begin{array}{c} v_1 \\ \vdots \\ v_n \end{array}\right) \]
1290: is a standard representative.
1291:
1292: We first give the proof that descending links are nonempty:
1293:
1294: \begin{proof}[Proof of Proposition~\ref{proposition:descending
1295: links are nonempty}]
1296:
1297: Let $k$ be the first column of $M$ which is not a coordinate vector
1298: (since $M$ is a standard representative, it follows that the entries
1299: in the first $k-1$ column vectors agree with the identity matrix up
1300: to sign). If we denote the $j^{\mbox{\tiny th}}$ entry of $v_i$ by $v_{ij}$,
1301: then $v_{kk}$ is nonzero. This follows from the fact
1302: that $M$ is a standard representative and the fact that $M$ is
1303: invertible.
1304:
1305: Since the $k^{\mbox{\tiny th}}$ column is not a coordinate vector (and since $M$ is
1306: invertible), there is a $j$, different from $k$, so that $v_{jk}$ is
1307: nonzero. If there is a $j > k$ such that $v_{jk} \neq 0$, then, since
1308: $M$ is a standard representative, $|v_{jk}| \leq |v_{kk}|$, and $a_k +
1309: \epsilon_j a_j$ is a descending ideal edge for some $\epsilon_j = \pm 1$. If
1310: $v_{jk}=0$ for all $j > k$, it follows that $v_{kk} = \pm 1$ (since
1311: $M$ is invertible) and there is some $j < k$ so that $v_{jk} \neq 0$.
1312: But then, again, $a_j + \epsilon a_k$ is descending for some $\epsilon
1313: = \pm 1$.
1314: %
1315: \epf
1316:
1317: Here is the proof that descending links are connected:
1318:
1319: \begin{proof}[Proof of Proposition~\ref{proposition:descending
1320: links are connected}]
1321:
1322: We first claim that if $\iota$ is any descending ideal edge, then
1323: there is a subordinate 2-letter ideal edge $\iota'$ which is also
1324: descending; see Section~\ref{section:ideal edges} for definitions. It
1325: will then follow from Lemma~\ref{lemma:subordinate} and
1326: Corollary~\ref{lemma:description of descending} that there is a path in
1327: $\dlkr$ between the 1-edge blowup realizing $\iota$ to the 1-edge
1328: blowup realizing $\iota'$ (the graph simultaneously realizing $\iota$
1329: and $\iota'$ is the midpoint of the path).
1330:
1331: To prove the claim, we need some notation. First, recall the
1332: notations $\rho$, $a_i$, $v_i$, and $M$ from above. Also, say
1333: (without loss of generality) that $\iota = a_{i_1} + a_{i_2} + \cdots +
1334: a_{i_m}$, and denote $v_{i_j}$ by $w_j$. Starting with the matrix
1335: with the $w_i$ as rows, we obtain a matrix $M'$ by deleting all
1336: columns without a nonzero entry.
1337: The $ij^{\mbox{\tiny th}}$ entry
1338: of $M'$ is denoted $w_{ij}$.
1339:
1340: We proceed in two cases. If the first column of $M'$ is not a
1341: coordinate vector, then at least two of the $w_{i1}$ are nonzero, in
1342: particular, $w_{11} \neq 0$. Without loss of generality, say $w_{11}
1343: > 0$. Since $\iota$ is descending, there must be a $k$ so that
1344: $w_{k1} < 0$, and since $M$ is a standard representative, we have
1345: $|w_{k1}| \leq |w_{11}|$. It follows that $a_{i_1} + a_{i_k}$ is
1346: descending, and this completes the proof of the first case.
1347:
1348: If the first column of $M'$ is a coordinate vector (i.e. $w_{11} = \pm
1349: 1$ and $w_{k1} = 0$ for $k > 1$), then we look at the second column of
1350: $M'$. Without loss of generality, assume $w_{22} > 0$. At this point
1351: there are three subcases. If $w_{k2}=0$ for all $k > 2$, then
1352: $a_{i_1}+a_{i_2}$ is descending, since $\iota$ is descending. If there is a $k > 2$ so that
1353: $w_{k2} < 0$ then $a_{i_2}+a_{i_k}$ is descending (since $M$ is a
1354: standard representative). If $w_{k2} \geq 0$ for all $k > 2$ and
1355: $w_{k2} \neq 0$ for at least one $k > 2$, then, since $\iota$ is descending, it follows that $w_{12}
1356: < 0$ and so $a_{i_1}+a_{i_k}$ is
1357: descending for any $k > 2$ with $w_k > 0$.
1358:
1359: We now claim that given any two descending 2-letter ideal edges, there
1360: is a path between the corresponding points in $\dlkr$. This follows, as above, from
1361: Lemma~\ref{lemma:subordinate} and Corollary~\ref{lemma:description of
1362: descending}, in addition to the fact that opposite 2-letter ideal
1363: edges cannot both be descending (Lemma~\ref{forbidden pairs}).
1364: This completes the proof.
1365: %
1366: \epf
1367:
1368:
1369: %%%
1370:
1371: \subsection{Completely descending link}
1372:
1373: We now shift our attention to Proposition~\ref{desc link dim}. Let
1374: $\rho$ be a rose represented by a marked graph $(\Ga,g)$, and say that
1375: $\Ga$ has edges $a_1, \dots, a_n$ labelled by $v_1, \dots, v_n$.
1376:
1377: The main argument for the proof (Section~\ref{section:dimension of
1378: descending links} below) is purely combinatorial, referring only to
1379: isomorphism types of labelled graphs. As things stand, however, we
1380: cannot describe $\dlkr$ in terms of combinatorial graphs
1381: without metrics. Indeed, given a marked graph in $\dlkr$,
1382: if we shrink the descending edges to have length less than 1 (while
1383: staying in the frontier by enlarging a nondescending edge), then the
1384: resulting marked graph is not in $\dlkr$
1385: (Corollary~\ref{lemma:description of descending}).
1386:
1387: To remedy this problem we perform a deformation retraction of $\dlkr$
1388: onto the \emph{completely descending link}, which we define to be the
1389: subset of $\dlkr$ consisting of marked graphs where each edge not
1390: labelled $\pm v_i$ is descending. The deformation retraction is
1391: achieved by simply shrinking all edges which correspond to
1392: nondescending ideal edges. Recall that these edges form a forest (Proposition~\ref{lemma2}(\ref{forest})),
1393: so there is no obstruction.
1394: We denote the completely descending link of $\rho$ by $\cdlkr$.
1395:
1396: \begin{lem}
1397: \label{def ret}
1398: For any given rose $\rho$, the completely descending link $\cdlkr$ is
1399: a strong deformation retract of the descending link $\dlkr$. In particular, the two
1400: are homotopy equivalent.
1401: \end{lem}
1402:
1403: We see that $\cdlkr$ has the desired cell structure: a cell
1404: is given by a combinatorial type of labelled graph and the cells are
1405: parameterized by the lengths of the edges in the graph. To be more
1406: precise, let $(\Ga'g')$ be a marked graph in $\cdlkr$, and for each
1407: $i$, let $k_i$ be the number of edges of $\Ga'$ labelled $\pm v_i$.
1408: For each $i$ we thus get a $(k_i-1)$-simplex by projecting
1409: \[ \set{(t_1, \dots, t_{k_i}) \in [0,1]^{k_i}: t_j=1 \mbox{ for some
1410: } j} \]
1411: to the simplex $\Delta_i = \set{\sum t_i=1}$. This projection is a
1412: homeomorphism. For each edge not labelled $\pm v_i$, we allow its
1413: length to vary arbitrarily within $[0,1]$, as long as one such edge
1414: has length 1. If $k_0$ is the
1415: number of such edges, then, as above, we get a $(k_0-1)$-simplex
1416: $\Delta_0$. Thus, the cell corresponding to $(\Ga',g')$ has a cell
1417: structure given by the product:
1418: \[ \Delta_0 \times \cdots \times \Delta_n \]
1419: We now summarize some of the important features of this cell
1420: structure:
1421: \begin{prop}
1422: \label{prop:cdlk}
1423: Consider a cell $C$ of $\cdlkr$ as above.
1424: \begin{enumerate}
1425: \item Passing to faces of $C$ corresponds to collapsing forests in $\Ga'$.
1426: \item $C$ is top-dimensional if and only if all vertices of $\Ga'$
1427: have valence 3.
1428: \item If $\Ga'$ has $v$ vertices, then $C$ has dimension $v-2$.
1429: \end{enumerate}
1430: \end{prop}
1431:
1432: %%%
1433:
1434: \subsection{Proof of Proposition~\ref{desc link dim}}
1435: \label{section:dimension of descending links}
1436:
1437: In this section we show that the completely descending link for any
1438: rose is homotopy equivalent to a complex of dimension $2n-5$ (Proposition~\ref{cdl dim}). Since
1439: the completely descending link is a deformation retract of the
1440: descending link (Lemma~\ref{def ret}), Proposition~\ref{desc link dim} follows as a corollary.
1441:
1442: As usual, let $\rho=(\Ga,g)$ be a rose in $\Y_n$, with edges $a_1,
1443: \dots, a_n$ labelled by $v_1, \dots, v_n$. If $(\Ga',g')$ is any
1444: marked graph in $\str$, we define the \emph{$v_i$-loop} as the image of $a_i$
1445: under a homotopy inverse of the collapsing map $\Ga' \to \Ga$.
1446:
1447:
1448: \begin{prop}
1449: \label{cdl dim}
1450: Let $n \geq 3$. For any rose $\rho$ in $\Y_n$, there is a strong deformation retraction of $\cdlkr$ onto a
1451: complex of dimension $2n-5$.
1452: \end{prop}
1453:
1454: \bpf
1455:
1456: If any top-dimensional cell of $\cdlkr$ has a free face in
1457: $\cdlkr$, then there is a homotopy equivalence (deformation
1458: retraction) of $\cdlkr$ which collapses away this cell. We perform
1459: this process inductively until we arrive at a subcomplex $L$ where no top-dimensional cell
1460: has a free face.
1461:
1462: We now suppose that $L$ is $(2n-4)$-dimensional,
1463: i.e., it has at least one top-dimensional cell. Among
1464: these, choose a cell $C$ where the total number of edges
1465: $\ell$ of a $v_1$-loop is minimal. Call the loop $P$ and choose one
1466: of its edges labelled $\pm v_1$ and call it $e$; see the leftmost
1467: diagram in Figure~\ref{fig:generic}. Say that $C$ is given
1468: by a marked graph $(\Ga',g')$.
1469:
1470: Firstly, note that $\ell$ is not 1, since there are no graphs with
1471: separating edges in $\Y_n$.
1472:
1473: \begin{figure}[ht]
1474: \psfrag{v1}{$v_1$}
1475: \includegraphics[scale=.5]{induction.eps}
1476: \caption{The top edge is $e$. The horizontal path plus $e$ is $P$.}\label{fig:generic}
1477: \end{figure}
1478:
1479: If we collapse any edge of $P-e$ (middle of Figure~\ref{fig:generic}),
1480: we move to a codimension 1 face of $C$. There are two ways to move to
1481: a new top-dimensional cell, since there are two other blowups of the
1482: resulting valence 4 vertex. One way reduces the length of $P$ (top
1483: right of Figure~\ref{fig:generic}), so by the minimality assumption for
1484: $C$, this is not a cell of $L$. Since we are assuming $C$ does not
1485: have any free faces, the other top-dimensional cell (bottom right of
1486: Figure~\ref{fig:generic}), call it $C'$, must be in $L$. The marked
1487: graphs in $C$ and $C'$ have the same labels outside of $P$; the
1488: difference is that the order of the edges leaving $P$ has changed
1489: (Proposition~\ref{lemma1}(\ref{labels dont change}) is applied twice).
1490:
1491: Continuing in this way, we see that if we permute the edges leaving
1492: $P$ in any way, we arrive at cells which are necessarily part of $L$.
1493: In particular, the graph obtained by taking the edge which leaves $P$
1494: at one endpoint of $e$ and moving it to the other endpoint of $e$
1495: gives a descending cell $\bar{C}$.
1496:
1497: We now argue that $C$ and $\bar{C}$ are opposite in the sense of
1498: Lemma~\ref{forbidden pairs}. Consider either endpoint of $e$ in
1499: $\Ga'$. This is a valence 3 vertex, as shown in Figure~\ref{fig:end
1500: of e}. By Proposition~\ref{lemma1}(\ref{switch condition}) and
1501: Proposition~\ref{lemma2}(\ref{coefficients}), the labels must be as in
1502: the left hand side of the figure. When we move the edge labelled
1503: $\sum k_iv_i$ to the other end of $e$ (as above), the labels must be
1504: as shown in the right hand side of the figure; the key point is that
1505: the labels and orientations do not change for $e$ and the edge being
1506: moved. It is then possible for us to determine the label for the
1507: third edge leaving the vertex where these edges meet. By
1508: Lemma~\ref{forbidden pairs} and Corollary~\ref{lemma:description of
1509: descending}, we have a contradiction.
1510: %
1511: \epf
1512:
1513:
1514: \begin{figure}[ht]
1515: \psfrag{1}{$v_1$}
1516: \psfrag{2}{$\displaystyle \sum_{i\geq 2} k_iv_i$}
1517: \psfrag{3}{$v_1+\displaystyle \sum_{i\geq 2} k_iv_i$}
1518: \psfrag{4}{$v_1-\displaystyle \sum_{i\geq 2} k_iv_i$}
1519: \includegraphics[scale=.75]{endofe.eps}
1520: \caption{Labels at endpoints of $e$.}
1521: \label{fig:end of e}
1522: \end{figure}
1523:
1524: %%%
1525: %%%
1526: %%%
1527:
1528: \section{The toy model and infinite generation of top homology}
1529: \label{section:infinite generation}
1530: \label{section:toy model}
1531:
1532: In this section we prove the second part of the main theorem, that
1533: $H_{2n-4}(\T_n,\Z)$ is not finitely generated when $n \geq 3$. In order to do this,
1534: we define a subcomplex $\tm_n$ of $\Y_n$, called the ``toy model'', we
1535: find an explicit infinite basis for $H_{2n-4}(\tm_n,\Z)$, and then we
1536: show that the inclusion $\tm_n \to \Y_n$ induces a monomorphism
1537: $H_{2n-4}(\tm_n,\Z) \to H_{2n-4}(\Y_n,\Z)$
1538: (Theorem~\ref{prop:independent tori}).
1539:
1540: %%%
1541:
1542: \subsection{Description of the toy model}
1543: \label{section:description of the toy model}
1544:
1545: Let $\rho=(\Ga,g)$ be the rose in $\Y_n$ with the identity marking,
1546: let $x_i$ denote the edges of the standard rose $R_n$, and let $a_i$ denote the corresponding edges of $\Ga$.
1547:
1548: Consider the set of points $\tm_n^0=\set{(\Gamma',g')}$ in $\str$ where $g'(x_1) \cup
1549: g'(x_2)$ is a rank 2 rose. We define the \emph{toy model} to be the
1550: subset $\tm_n$ of $\Y_n$ given by: \[ \tm_n = \bigcup_{p_i,q_i\in \Z} \tm_n^0 \cdot
1551: \left[
1552: \begin{array}{cccccc}
1553: 1 & 0 & p_3 & p_4 & \cdots & p_n \\
1554: 0 & 1 & q_3 & q_4 & \cdots & q_n \\
1555: 0 & 0 & 1& 0 & \cdots & 0 \\
1556: 0 & 0 & 0& 1 & \cdots & 0 \\
1557: \vdots & & & & \ddots & \vdots \\
1558: 0 & 0 & 0& 0 & \cdots & 1 \\
1559: \end{array}
1560: \right ]
1561: \]
1562:
1563: \p{Another point of view} We now give a different description of
1564: $\tm_n$, which will make it easier to find its homotopy type.
1565:
1566: For each marked graph $(\Ga',g')$ of $\tm_n$, the union $g'(x_1) \cup g'(x_2)$
1567: is a rank 2 rose in $\Ga'$, and $\Ga'$ has $n-2$ edges $a_3, \dots,
1568: a_n$ labelled $v_3, \dots, v_n$, where $v_i$ is a coordinate vector
1569: with $+1$ in the $i^{\mbox{\tiny th}}$ spot.
1570: By considering the starting and ending points of $a_3, \dots, a_n$ as
1571: points in $g'(x_1) \cup g'(x_2)$, a path in $\tm_n$ can be thought of
1572: as a path in the configuration space of $n-2$ pairs of points in the
1573: universal abelian cover of $g'(x_1) \cup g'(x_2)$, which is $U = (\R
1574: \times \Z) \cup (\Z \times \R)$. To make this precise, for each
1575: metric graph $(\Ga',g')$ in $\tm_n$, we rescale the metric so that
1576: $g'(x_1)$ and $g'(x_2)$ both have length 1. After doing this, the
1577: endpoints of the $a_i$ give a well-defined subset of the metric cover
1578: $U$.
1579:
1580: If, in the configuration space, we move the two points
1581: corresponding to the endpoints of some $a_i$ by the same integral
1582: vector, then the corresponding point in $\tm_n$ does not change.
1583:
1584: \begin{prop}
1585: \label{proposition:toy model as config space}
1586: The above construction defines a homeomorphism:
1587: \[ (U^2)^{n-2}/(\Z^2)^{n-2} \to \tm_n \]
1588: \end{prop}
1589:
1590: At this point, the proof is straightforward and is left to the reader.
1591:
1592: A typical graph in $\tm_7$ is shown in Figure~\ref{fig:torus
1593: graph}. That graph is ``maximally blown up'' in the sense that it has
1594: the greatest number of valence 3 vertices possible in $\tm_7$.
1595:
1596: \begin{figure}[ht]
1597: \includegraphics[scale=.5]{torus.eps}
1598: \caption{A maximally blown up graph in $\tm_7$.}
1599: \label{fig:torus graph}
1600: \end{figure}
1601:
1602: %%%
1603:
1604: \subsection{Homotopy type of the toy model}
1605: \label{section:rank 3 toy model}
1606:
1607: We start by focusing our attention on the rank 3 toy model
1608: $\tm_3$. In general, we have $\tm_n \cong (\tm_3)^{n-2}$, and so we will be able to deduce the finiteness properties of $\tm_n$ from those of $\tm_3$.
1609:
1610: Via Proposition~\ref{proposition:toy model as config space}, we can
1611: think of $\tm_3$ as pairs of points in $U$. However, it will simplify
1612: our analysis if we thicken $U$ to a space $V$, which we now
1613: define. First, for any integers $p$ and $q$, denote by $D_{p,q}$ the
1614: open disk of radius $r$ around $(p,q) \in \R^2$, for some fixed $r$
1615: close to zero. Then, define $V = \R^2 - \cup D_{p,q}$.
1616:
1617: The straight line retraction of $V$ onto $U$ gives a homotopy
1618: equivalence from $V^2/\Z^2$ to $U^2/\Z^2 \cong \tm_3$. Thinking of
1619: $V^2/\Z^2$ as pairs of points in $V$, we immediately see the following
1620: features:
1621:
1622: \begin{enumerate}
1623: \item The diagonal of $V^2/\Z^2$ is a torus with one boundary
1624: component.
1625: \item For each $(p,q)$, there is a
1626: 2-torus $Z_{p,q} = \partial D_{0,0} \times \partial D_{p,q}$.
1627: \end{enumerate}
1628:
1629: We will now use Morse theory to argue that these features capture
1630: the homotopy type of $\tm_3$. We consider the Morse function
1631: $d: V^2/\Z^2 \to \R$ which assigns to a point in $V^2/\Z^2$ the
1632: Euclidean distance between the pair of points in $V^2$.
1633:
1634: \begin{figure}[ht]
1635: \psfrag{minset}{minset}
1636: \psfrag{index 1}{index 1}
1637: \psfrag{index 2}{index 2}
1638: \includegraphics[scale=.4]{critpoints.eps}
1639: \caption{Critical points for the toy model in rank 3.\label{critpointspic}}
1640: \end{figure}
1641:
1642: We see that $d$ has the following features, depicted in Figure~\ref{critpointspic}:
1643: \blwide
1644:
1645: \item The minset is the torus with one boundary component
1646: corresponding to the diagonal of $V^2/\Z^2$.
1647:
1648: \item There is a horizontal 1-cell of index 1 critical points
1649: corresponding to pairs of points lying diametrically
1650: opposite from each other on $\partial D_{0,0}$.
1651:
1652: \item For every $(p,q)\neq (0,0)$, there is an index 2 critical point,
1653: corresponding to the two points of tangency of $\partial
1654: D_{0,0}$ and $\partial D_{p,q}$ with the unique circle tangent to both.
1655:
1656: \item At all other points, there is a well-defined gradient flow, and
1657: so there are no other critical points.
1658: \el
1659:
1660: We notice that each index two critical point is the maximum point of
1661: the corresponding $Z_{p,q}$. Thus, at each critical point of index 2,
1662: a 2-cycle is added. One can also see that, at the horizontal 1-cell,
1663: there is another torus being added. Since there are no critical
1664: points of index greater than 2, these classes are nontrivial in $H_2(\tm_3,\Z)$.
1665:
1666: In higher rank, we define the torus $Z_{p,q}$ to be the $(n-2)$-fold
1667: product $(\partial D_{0,0} \times \partial D_{p,q})^{n-2}$.
1668: As $\tm_n \cong \tm_3^{n-2}$, we have:
1669:
1670: \begin{prop}
1671: \label{proposition:generators for top homology of toy model}
1672: The $Z_{p,q}$ freely generate $H_{2n-4}(\tm_n,\Z)$.
1673: \end{prop}
1674:
1675: To formalize the above argument, one can use Morse theory for
1676: manifolds with corners (see \cite{db}). There is one technicality:
1677: the critical values of $d$ are not isolated; this is easily overcome
1678: by replacing each $D_{p,q}$ with an ellipse (or proving a more
1679: general Morse theory).
1680:
1681: We remark that the image of $\pi_1(Z_{p,q}) \cong \Z^{2n-4}$ in
1682: $\pi_1(\Y_n) \cong \T_n$ is a conjugate of the the subgroup $G$
1683: of $\T_n$ described in the introduction. To see this, one simply
1684: needs to understand the effect of blowups and blowdowns on the
1685: \emph{homotopy} classes of marked graphs; see~\cite{cv}.
1686:
1687: %%%
1688:
1689: \subsection{Independence in homology} We now set out to prove
1690: that the $Z_{p,q}$ represent independent classes in $\Y_n$
1691: (Theorem~\ref{prop:independent tori}). In particular, this will
1692: prove the second part of the main theorem.
1693:
1694: By understanding the homotopy equivalences
1695: \[ (V^2)^{n-2}/(\Z^2)^{n-2} \to (U^2)^{n-2}/(\Z^2)^{n-2}
1696: \to \tm_n \]
1697: we can give a concrete description of the $Z_{p,q}$ in terms of marked
1698: graphs. First of all, we have:
1699:
1700: \begin{lem}
1701: \label{seg has tori}
1702: Each $Z_{p,q}$ is contained in the union of stars of roses with marking matrices of the form:
1703: \[ \left [
1704: \begin{array}{cccccc}
1705: 1 & 0 & p_3 & p_4 & \cdots & p_n \\
1706: 0 & 1 & q_3 & q_4 & \cdots & q_n \\
1707: 0 & 0 & 1& 0 & \cdots & 0 \\
1708: 0 & 0 & 0& 1 & \cdots & 0 \\
1709: \vdots & & & & \ddots & \vdots \\
1710: 0 & 0 & 0& 0 & \cdots & 1 \\
1711: \end{array}
1712: \right ]
1713: \]
1714: where each $p_i \in [p-1,p+1]$ and $q_i \in [q-1,q+1]$.
1715: \end{lem}
1716:
1717: The main fact we will need about the stars of roses listed in
1718: Lemma~\ref{seg has tori} is that any ideal edge of the form
1719: \[ \pm v_1 \pm v_2 + \sum_{i \geq 3} k_i v_i \]
1720: is \emph{ascending}.
1721: We will also need the following observation about the $Z_{p,q}$:
1722:
1723: \begin{lem}
1724: \label{lemma:weakly realized}
1725: For any $Z_{p,q}$ and any rose $\rho$, we can give $Z_{p,q}$ the same cell structure as $\cdlkr$. In particular, if a marked graph in some $Z_{p,q}$ has an edge of length less than 1 corresponding to an ideal
1726: edge $\iota$, then there is a point in that $Z_{p,q}$ which realizes
1727: $\iota$.
1728: \end{lem}
1729:
1730: For the remainder, fix a $Z_{p,q}$, and let $\rho$ be a rose of
1731: greatest norm which intersects $Z_{p,q}$. By Lemma~\ref{seg has
1732: tori}, if $p$ and $q$ are both nonzero, then $\rho$ is unique; if
1733: one of them is zero, then there are $2^{n-2}$ choices for $\rho$; and if
1734: $p=q=0$, then there are $2^{2n-4}$ choices.
1735: From Lemma~\ref{seg has tori}, we deduce the following key fact:
1736:
1737: \begin{lem}
1738: \label{big torus}
1739: Any map from $\set{Z_{p,q}}$ to roses, sending $Z_{p,q}$ to any rose $\rho$
1740: of maximal norm with $Z_{p,q} \cap \str \neq \emptyset$, is injective.
1741: In particular, given any finite subset of $\set{Z_{p,q}}$, the rose of
1742: maximal norm intersecting this set has nonempty intersection with
1743: exactly one torus in this set.
1744: \end{lem}
1745:
1746: Let $\dlkr$ denote the descending link for $\rho$ as
1747: defined in Section~\ref{subsection:the induction}.
1748:
1749: \begin{lem}
1750: \label{torus sphere}
1751: The intersection $Z_{p,q} \cap \dlkr$ is homeomorphic to $S^{2n-5}$.
1752: \end{lem}
1753:
1754: \bpf
1755:
1756: We assume $p,q > 0$, with the other cases handled similarly.
1757:
1758: Under the homotopy equivalence $(V^2)^{n-2}/(\Z^2)^{n-2} \to
1759: (U^2)^{n-2}/(\Z^2)^{n-2}$, we can identify $Z_{p,q}$ with the
1760: configuration space of $n-2$ pairs of points in $U \subset \R^2$
1761: where the first point $z_i$ in each pair lies on the coordinate square
1762: with vertices at $(0,0)$ and $(1,1)$ and the second point $z_i'$ in
1763: each pair lies on the square with vertices $(p,q)$ and $(p+1,q+1)$.
1764:
1765: The rose $\rho$ is realized when each $z_i$ is at $(0,0)$ and each
1766: $z_i'$ is at the point $(p+1,q+1)$. The points of $Z_{p,q} \cap
1767: \dlkr$ are exactly the set of points where each $z_i$ is within a
1768: distance of $1/2$ from the origin, each $z_i'$ is within $1/2$ of
1769: $(p+1,q+1)$, and at least one $z_i$ or $z_i'$ has distance exactly $1/2$.
1770:
1771: In other words, each $z_i$ and $z_i'$ is allowed to move within a
1772: closed interval, and such a configuration is in $Z_{p,q} \cap \dlkr$
1773: if at least one of the points is on the boundary of its interval.
1774: Thus, $Z_{p,q} \cap \dlkr$ is homeomorphic to $\partial I^{2n-4} \cong
1775: S^{2n-5}$.
1776: %
1777: \epf
1778:
1779: The following completes the proof of the main theorem:
1780:
1781: \begin{thm}
1782: \label{prop:independent tori}
1783: Let $n\geq 3$. The $Z_{p,q}$ form an infinite set of independent
1784: classes in $H_{2n-4}(\Y_n,\Z) \cong H_{2n-4}(\T_n,\Z)$. In other
1785: words, $H_{2n-4}(\tm_n,\Z)$ injects into $H_{2n-4}(\Y_n,\Z)$.
1786: \end{thm}
1787:
1788: \bpf
1789:
1790: Given a finite subset $A$ of $\set{Z_{p,q}}$, let $Z_{p,q}$ be an element
1791: which intersects a rose $\rho$ of highest norm (marking matrix as in
1792: Lemma~\ref{seg has tori}, the $i^{\mbox{\tiny th}}$ row corresponds to
1793: $v_i$). We know that there is a strong deformation retraction of
1794: $\dlkr$ onto a complex of dimension $2n-5$ (Lemma~\ref{def ret} plus
1795: Proposition~\ref{cdl dim}). The goal is to show that the image of the
1796: sphere $Z_{p,q} \cap \dlkr$ is embedded in this complex. Even better,
1797: we will show that the deformation retractions of Lemma~\ref{def ret}
1798: and Proposition~\ref{cdl dim} do not move the points of $Z_{p,q} \cap
1799: \dlkr$.
1800:
1801: To see why this proves this proposition, we consider the long exact
1802: sequence associated to the pair $(\str,\dlkr)$:
1803: \[
1804: \begin{array}{l} \cdots \to H_{2n-4}(\str) \to H_{2n-4}(\str,\dlkr)
1805: \\ \qquad
1806: \to H_{2n-5}(\dlkr) \to H_{2n-5}(\str) \to \cdots
1807: \end{array}
1808: \]
1809: By excision, $Z_{p,q}$ corresponds to a class in
1810: $H_{2n-4}(\str,\dlkr)$. The image in $H_{2n-5}(\dlkr)$ is the class
1811: $Z_{p,q} \cap \dlkr$, which is nontrivial once we show $Z_{p,q} \cap
1812: \dlkr$ is embedded in the
1813: $(2n-5)$-dimensional deformation retract of $\dlkr$ (Lemma~\ref{torus sphere}). It follows
1814: that $Z_{p,q}$ is nontrivial in $H_{2n-4}(\str,\dlkr)$ and hence, via
1815: excision, in $H_{2n-4}(\seg',\seg)$ where $\seg$ is the largest initial segment
1816: not containing $\rho$ and $\seg'=\seg\cup\str$. By Lemma~\ref{big
1817: torus}, each element of $A$ other than $Z_{p,q}$ is trivial in
1818: $H_{2n-4}(\seg',\seg)$, and so $Z_{p,q}$ is linearly independent from
1819: these, which is what we wanted to show.
1820:
1821: \begin{comment}
1822: Here is why this proves the proposition: by excision, $Z_{p,q}$
1823: corresponds to a class in $H_{2n-4}(\str,\dlkr)$
1824:
1825: since $Z_{p,q} \cap \dlkr$ is homeomorphic to $S^{2n-5}$ (Lemma~\ref{torus sphere}) and the deformation retraction of $\dlkr$ is $(2n-5)$-dimensional, it will follow that $Z_{p,q} \cap \dlkr$ represents a nontrivial class in $H_{2n-5}(\dlkr)$. Then, we consider the long exact sequence associated to the pair $(\str,\dlkr)$:
1826: \[
1827: \begin{array}{l} \cdots \to H_{2n-4}(\str) \to H_{2n-4}(\str,\dlkr)
1828: \\ \qquad \to
1829: H_{2n-5}(\dlkr) \to H_{2n-5}(\str) \to \cdots
1830: \end{array}
1831: \]
1832: The first term shown is trivial (Proposition~\ref{star
1833: dim}), and so (by excision) $Z_{p,q}$ corresponds to a nontrivial class in
1834: $H_{2n-4}(\str,\dlkr)$. But by excision, this is a nontrivial class
1835: in $H_{2n-4}(\seg',\seg)$ where $\seg$ is the largest initial segment
1836: not containing $\rho$ and $\seg'=\seg\cup\str$. It follows that
1837: $Z_{p,q}$ is not in the space spanned by the tori contained in $\seg$,
1838: and this gives the proposition.
1839:
1840: \end{comment}
1841:
1842: Thus, we are reduced to showing that the two deformation retractions do not move the sphere $Z_{p,q} \cap \dlkr$. We handle each in turn.
1843:
1844: For the deformation retraction of $\dlkr$ onto $\cdlkr$, we need to
1845: show that $Z_{p,q} \cap \dlkr$ is already contained in $\cdlkr$.
1846: Suppose that there were a point of $Z_{p,q} \cap \dlkr$ which were not
1847: contained in $\cdlkr$. By
1848: Lemma~\ref{lemma:weakly realized}, there is a point which realizes an
1849: \emph{ascending} ideal edge (Lemma~\ref{lemma:neighbors}), and this implies that $Z_{p,q}$ intersects some rose
1850: of higher norm, contradicting the choice of $\rho$.
1851:
1852: We now focus on the deformation retraction of Proposition~\ref{cdl
1853: dim}. A marked graph representing a maximal cell of $\cdlkr$ must
1854: have disjoint $v_1$- and $v_2$-loops. Indeed, there are no valence 4 vertices, so the overlap would have to contain an ascending edge by
1855: Proposition~\ref{lemma1}(\ref{switch condition}), the statement after
1856: Lemma~\ref{seg has tori}, and Lemma~\ref{descending criterion}. The
1857: codimension 1 cells which get collapsed are obtained from these
1858: maximal cells by collapsing an edge of the $v_1$-loop. Thus, the
1859: corresponding graphs still have disjoint $v_1$- and $v_2$-loops. On
1860: the other hand, in any graph of $Z_{p,q}$, the $v_1$-loop and the
1861: $v_2$-loop intersect in exactly 1 point. Thus, no points of $Z_{p,q}$
1862: are moved during this retraction, so we are done.
1863: %
1864: \epf
1865:
1866:
1867: %%%
1868: %%%
1869: %%%
1870:
1871: \appendix
1872:
1873: \section*{Appendix: Proof of Proposition~\ref{magnus}}
1874: \label{section:magnus}
1875:
1876: This appendix contains Magnus's proof of Proposition~\ref{magnus}. In
1877: this section, we freely use the notation of
1878: Section~\ref{section:finite generation}.
1879:
1880: Let $K$ be the subgroup of $\out(F_n)$ generated by the $K_{ik}$ and
1881: $K_{ikl}$ for distinct $i$, $k$, and $l$. We now prove
1882: Proposition~\ref{magnus}, that $K$ is normal in $\out(F_n)$.
1883:
1884: \bpf[Proof of Proposition~\ref{magnus}]
1885:
1886: We choose the following generating set for $\out(F_n)$:
1887: \[ \begin{array}{rcl}
1888: \delta_{12} & = & [x_1 x_2, x_2, \dots, x_n] \\
1889: \Omega_1 & = & [x_1^{-1}, x_2, \dots, x_n] \\
1890: \Pi_{i-1} & = & [x_1, \dots, x_{i-2}, x_i, x_{i-1}, x_{i+1},
1891: \dots, x_n]
1892: \end{array}\]
1893:
1894: It suffices to show that the conjugates of the $K_{ik}$ and
1895: $K_{ikl}$ by the chosen generators of $\out(F_n)$ (and their
1896: inverses) are elements of $K$.
1897:
1898: We have the following simplifications:
1899:
1900: \begin{enumerate}
1901:
1902: \item Operations on disjoint sets of elements commute.
1903:
1904: \item Since $\Omega_1$ and $\Pi_{i-1}$ have order 2, we don't need to
1905: conjugate by their inverses.
1906:
1907: \item We don't need to conjugate by $\delta_{12}^{-1}$ since \[ (\Pi_1
1908: \Omega_1 \Pi_1) \delta_{12} (\Pi_1 \Omega_1 \Pi_1)^{-1} =
1909: \delta_{12}^{-1} \]
1910:
1911: \item Since $K_{ikl} = K_{ilk}^{-1}$, we may assume $k < l$.
1912:
1913: \item Any outer automorphism $\psi$ of the form
1914: \[ [x_1, \dots, g' x_i g, \dots, x_n] \]
1915: where $gg'$ is an element of the commutator subgroup of the subgroup $H$
1916: of $F_n$ generated by $\set{x_k:k \neq i}$ is an element of $K$.
1917:
1918: \bigskip
1919:
1920: To see that $\psi \in K$, first note that, by postcomposing with a product of $K_{i\star}^{\pm 1}$,
1921: we may assume that $g'=1$. Now, we know that the commutator subgroup of $H$ is normally generated
1922: by the $[x_k,x_l]$, where $k$ and $l$ are both different from $i$.
1923: Therefore, it suffices to handle the case of
1924: \[ g = h [x_k,x_l] h^{-1} = [hx_kh^{-1},hx_lh^{-1}] \]
1925: where $h = x_{i_1} \cdots x_{i_p}$ is an arbitrary element of $H$.
1926: It is elementary to check that
1927: \[ \psi = P^{-1} K_{ikl} P \]
1928: where
1929: \[ P = \prod_{j \neq i} K_{ji_p} \cdots \prod_{j \neq i} K_{ji_2}
1930: \prod_{j \neq i} K_{ji_1} \]
1931: \end{enumerate}
1932:
1933: \bigskip
1934:
1935: Given these simplifications, it is straightforward to check (case by case) that the
1936: conjugates by $\Pi_{i-1}$, $\Omega_1$, and $\delta_{12}$ of each
1937: $K_{ik}$ and $K_{ikl}$ are elements of $K$. There is one exception; we give Magnus's
1938: computation for this difficult case here:
1939: \[ \delta_{12} K_{2l1} \delta_{12}^{-1} =
1940: K_{l2}K_{l1}^{-1}K_{1l}K_{l1}K_{2l1}K_{12l}K_{l2}^{-1}K_{2l}^{-1} \]
1941: %
1942: \epf
1943:
1944:
1945: \bibliographystyle{plain}
1946: \bibliography{tn}
1947:
1948: \end{document}