memoire-m2

My M2 Memoire on mapping class groups & their representations

twists.tex (27362B)

  1 \chapter{Dehn Twists}\label{ch:dehn-twists}
  2 
  3 With the goal of studying the linear representations of mapping class groups in
  4 mind, we now start investigating the group structure of \(\Mod(\Sigma)\). We
  5 begin by computing some fundamental examples and then explore how we can use
  6 these examples to understand the structure of the mapping class groups of other
  7 surfaces. Namely, we compute \(\Mod(\mathbb{S}^1 \times [0, 1]) \cong
  8 \mathbb{Z}\), and discuss how its generator gives rise to a convenient
  9 generating set for \(\Mod(\Sigma)\), known as the set of \emph{Dehn twists}.
 10 
 11 The idea here is to reproduce the proof of injectivity in
 12 Observation~\ref{ex:torus-mcg}: by cutting along curves and arcs, we can always
 13 decompose a surface into copies of \(\mathbb{D}^2\) and \(\mathbb{D}^2
 14 \setminus \{0\}\). Observation~\ref{ex:alexander-trick} and
 15 Observation~\ref{ex:mdg-once-punctured-disk} then imply the triviality of
 16 mapping classes fixing such arcs and curves. Formally, this translates to the
 17 following result.
 18 
 19 \begin{proposition}[Alexander method]\label{thm:alexander-method}
 20   Let \(\alpha_1, \ldots, \alpha_n \subset \Sigma\) be essential simple closed
 21   curves or proper arcs satisfying the following conditions.
 22   \begin{enumerate}
 23     \item \([\alpha_i] \ne [\alpha_j]\) for \(i \ne j\).
 24     \item Each pair \((\alpha_i, \alpha_j)\) crosses at most once.
 25     \item Given distinct \(i, j, k\), at least one of \(\alpha_i \cap \alpha_j,
 26       \alpha_i \cap \alpha_k, \alpha_j \cap \alpha_k\) is empty.
 27     \item The surface obtained by cutting \(\Sigma\) along the \(\alpha_i\) is a
 28       disjoint union of disks and once-punctured disks.
 29   \end{enumerate}
 30   Suppose \(f \in \Mod(\Sigma)\) is such that \(f \cdot \vec{[\alpha_i]} =
 31   \vec{[\alpha_i]}\) for all \(i\). Then \(f = 1 \in \Mod(\Sigma)\).
 32 \end{proposition}
 33 
 34 See \cite[Proposition~2.8]{farb-margalit} for a proof of
 35 Proposition~\ref{thm:alexander-method}. We now state some \emph{fundamental}
 36 applications of the Alexander method.
 37 
 38 \begin{example}\label{ex:mcg-annulus}
 39   The mapping class group \(\Mod(\mathbb{S}^1 \times [0, 1])\) is freely
 40   generated by \(f = [\phi]\), where
 41   \begin{align*}
 42     \phi : \mathbb{S}^1 \times [0, 1] & \isoto  \mathbb{S}^1 \times [0, 1] \\
 43                    (e^{2 \pi i t}, s) & \mapsto (e^{2 \pi i (t - s)}, s)
 44   \end{align*}
 45   is the map illustrated in Figure~\ref{fig:dehn-twist-cylinder}. In
 46   particular, \(\Mod(\mathbb{S}^1 \times [0, 1]) \cong \mathbb{Z}\).
 47 \end{example}
 48 
 49 \begin{example}\label{ex:mcg-twice-punctured-disk}
 50   The mapping class group \(\Mod(\mathbb{D}^2 \setminus \{-\sfrac{1}{2},
 51   \sfrac{1}{2}\})\) of the twice punctured unit disk in \(\mathbb{C}\) is
 52   freely generated by \(f = [\phi]\), where
 53   \begin{align*}
 54     \phi : \mathbb{D}^2 \setminus \{-\sfrac{1}{2}, \sfrac{1}{2}\}
 55     & \isoto \mathbb{D}^2 \setminus \{-\sfrac{1}{2}, \sfrac{1}{2}\} \\
 56     z & \mapsto -z
 57   \end{align*}
 58   is the map from Figure~\ref{fig:hald-twist-disk}. In particular,
 59   \(\Mod(\mathbb{D}^2 \setminus \{-\sfrac{1}{2}, \sfrac{1}{2}\}) \cong
 60   \mathbb{Z}\).
 61 \end{example}
 62 
 63 \noindent
 64 \begin{minipage}[b]{.47\linewidth}
 65   \centering
 66   \includegraphics[width=.7\linewidth]{images/dehn-twist-cylinder.eps}
 67   \captionof{figure}{The generator $f$ of $\Mod(\mathbb{S}^1 \times [0, 1])
 68   \cong \mathbb{Z}$ takes the yellow arc on the left-hand side to the arc on
 69   the right-hand side that winds about the curve $\alpha$.}
 70   \label{fig:dehn-twist-cylinder}
 71 \end{minipage}
 72 \hspace{.6cm} %
 73 \begin{minipage}[b]{.47\linewidth}
 74   \centering
 75   \includegraphics[width=.4\linewidth]{images/half-twist-disk.eps}
 76   \captionof{figure}{The generator $f$ of $\Mod(\mathbb{D}^2 \setminus
 77   \{-\sfrac{1}{2}, \sfrac{1}{2}) \cong \mathbb{Z}$ corresponds to the
 78   clockwise rotation by $\pi$ about the origin.}
 79   \label{fig:hald-twist-disk}
 80 \end{minipage}
 81 
 82 Let \(\Sigma\) be an orientable surface, possibly with punctures and non-empty
 83 boundary. Given some closed \(\alpha \subset \Sigma\), we may envision doing
 84 something similar to Example~\ref{ex:mcg-annulus} in \(\Sigma\) by looking at
 85 annular neighborhoods of \(\alpha\). These are precisely the \emph{Dehn twists},
 86 illustrated in Figure~\ref{fig:dehn-twist-bitorus} in the case of the bitorus
 87 \(\Sigma_2\).
 88 
 89 \begin{definition}
 90   Given a simple closed curve \(\alpha \subset \Sigma\), fix a closed annular
 91   neighborhood \(A \subset \Sigma\) of \(\alpha\) -- i.e. \(A \cong
 92   \mathbb{S}^1 \times [0, 1]\). Let \(f \in \Mod(A) \cong \Mod(\mathbb{S}^1
 93   \times [0, 1]) \cong \mathbb{Z}\) be the generator from
 94   Example~\ref{ex:mcg-annulus}. The \emph{Dehn twist \(\tau_\alpha \in
 95   \Mod(\Sigma)\) about \(\alpha\)} is defined as the image of \(f\) under the
 96   inclusion homomorphism \(\Mod(A) \to \Mod(\Sigma)\).
 97 \end{definition}
 98 
 99 \begin{figure}[ht]
100   \centering
101   \includegraphics[width=.6\linewidth]{images/dehn-twist-bitorus.eps}
102   \caption{The Dehn twist about the curve $\alpha$ takes the peanut-shaped curve
103   on the left-hand side to the yellow curve on the right-hand side.}
104   \label{fig:dehn-twist-bitorus}
105 \end{figure}
106 
107 Similarly, using the description of the mapping class group of the
108 twice-puncture disk derived in Example~\ref{ex:mcg-twice-punctured-disk}, the
109 generator of \(\Mod(\mathbb{D}^2 \setminus \{-\sfrac{1}{2}, \sfrac{1}{2}\})\)
110 gives rise the so-called \emph{half-twists}. These are examples of mapping
111 classes that permute the punctures of \(\Sigma\).
112 
113 \begin{definition}
114   Given an arc \(\alpha \subset \Sigma\) joining two punctures in the interior
115   of \(\Sigma\), fix a closed neighborhood \(D \subset \Sigma\) of \(\alpha\)
116   with \(D \cong \mathbb{D}^2 \setminus \{-\sfrac{1}{2}, \sfrac{1}{2}\}\). Let
117   \(f \in \Mod(D) \cong \Mod(\mathbb{D}^2 \setminus \{-\sfrac{1}{2},
118   \sfrac{1}{2}\}) \cong \mathbb{Z}\) be the generator from
119   Example~\ref{ex:mcg-twice-punctured-disk}. The \emph{half-twist \(h_\alpha
120   \in \Mod(\Sigma)\) about \(\alpha\)} is defined as the image of \(f\) under
121   the inclusion homomorphism \(\Mod(D) \to \Mod(\Sigma)\).
122 \end{definition}
123 
124 We can use the Alexander method to describe the kernel of capping and cutting
125 morphisms in terms of Dehn twists.
126 
127 \begin{observation}[Capping exact sequence]\label{ex:capping-seq}
128   Let \(\delta \subset \partial \Sigma\) be a boundary component of \(\Sigma\)
129   and \(\operatorname{cap} : \Mod(\Sigma) \to \Mod(\Sigma \cup_\delta
130   (\mathbb{D}^2 \setminus \{0\}))\) be the corresponding the capping
131   homomorphism from Example~\ref{ex:capping-morphism}. There is an exact
132   sequence
133   \begin{center}
134     \begin{tikzcd}
135       1 \rar &
136       \langle \tau_\delta \rangle \rar &
137       \Mod(\Sigma) \rar{\operatorname{cap}} &
138       \Mod(\Sigma \cup_\delta (\mathbb{D}^2 \setminus \{0\}), 0) \rar &
139       1,
140     \end{tikzcd}
141   \end{center}
142   known as \emph{the capping exact sequence} -- see
143   \cite[Proposition~3.19]{farb-margalit} for a proof.
144 \end{observation}
145 
146 \begin{observation}\label{ex:cutting-morphism-kernel}
147   Let \(\alpha \subset \Sigma\) be a simple closed curve and
148   \(\operatorname{cut} : \Mod(\Sigma)_{\vec{[\alpha]}} \to \Mod(\Sigma
149   \setminus \alpha)\) be the cutting homomorphism from
150   Example~\ref{ex:cutting-morphism}. Then \(\ker \operatorname{cut} = \langle
151   \tau_\alpha \rangle \cong \mathbb{Z}\).
152 \end{observation}
153 
154 It is also interesting to study how the geometry of two curves affects the
155 relationship between their corresponding Dehn twists. For instance,
156 by investigating the geometric intersection number
157 \[
158   \#(\alpha \cap \beta) = \min
159   \left\{
160     |\alpha' \cap \beta'| : [\alpha'] = [\alpha]
161     \text{ and }
162     [\beta'] = [\beta]
163   \right\}
164 \]
165 we can distinguish between powers of Dehn twists
166 \cite[Proposition~3.2]{farb-margalit}.
167 
168 \begin{proposition}\label{thm:twist-intersection-number}
169   Let \(\alpha \subset \Sigma\) be a simple closed curve and \(T_\alpha\) be a
170   representative of \(\tau_\alpha \in \Mod(\Sigma)\). Then \(\#
171   (T_\alpha^k(\beta) \cap \beta) = |k| \cdot \#(\alpha \cap \beta)^2\) for any
172   \(k \in \mathbb{Z}\). In particular, if \(\alpha\) is nontrivial then
173   \(\tau_\alpha\) has infinite order.
174 \end{proposition}
175 
176 \begin{observation}
177   Given \(\alpha, \beta \subset \Sigma\), \(\tau_\alpha = \tau_\beta \iff
178   [\alpha] = [\beta]\). Indeed, if \(\alpha\) and \(\beta\) are non-isotopic,
179   we can find \(\gamma\) with \(\#(\gamma \cap \alpha) > 0\) and \(\#(\gamma
180   \cap \beta) = 0\). It thus follows from
181   Proposition~\ref{thm:twist-intersection-number} that \(\#(T_\alpha(\gamma)
182   \cap \gamma) > \#(T_\beta(\gamma) \cap \gamma)\), so \(\tau_\alpha \ne
183   \tau_\beta\).
184 \end{observation}
185 
186 Many other relations between Dehn twists can derived be in a geometric fashion
187 too.
188 
189 \begin{observation}\label{ex:conjugate-twists}
190   Given \(f = [\phi] \in \Mod(\Sigma)\), \(\tau_{\phi(\alpha)} = f \tau_\alpha
191   f^{-1}\).
192 \end{observation}
193 
194 \begin{observation}[Disjointness relations]
195   Given \(f \in \Mod(\Sigma)\), \([f, \tau_\alpha] = 1 \iff f \cdot [\alpha] =
196   [\alpha]\). In particular, \([\tau_\alpha, \tau_\beta] = 1\) for \(\alpha\)
197   and \(\beta\) disjoint, for we can choose a representative of \(\tau_\beta\)
198   whose support is disjoint from \(\alpha\).
199 \end{observation}
200 
201 \begin{observation}
202   If \(\alpha, \beta \subset \Sigma\) are both nonseparating then \(\tau_\alpha,
203   \tau_\beta \in \Mod(\Sigma)\) are conjugate. Indeed, by the change of
204   coordinates principle we can find \(f \in \Mod(\Sigma)\) with \(f \cdot
205   [\alpha] = [\beta]\) and then apply Observation~\ref{ex:conjugate-twists}.
206 \end{observation}
207 
208 \begin{observation}[Braid relations]\label{ex:braid-relation}
209   Given \(\alpha, \beta \subset \Sigma\) with \(\#(\alpha \cap \beta) = 1\), it
210   is not hard to check that \(\tau_\beta \tau_\alpha \cdot [\beta] =
211   [\alpha]\). From Observation~\ref{ex:conjugate-twists} we then get
212   \((\tau_\alpha \tau_\beta) \tau_\alpha (\tau_\alpha \tau_\beta)^{-1} =
213   \tau_\beta\), from which follows the \emph{braid relation}
214   \[
215     \tau_\alpha \tau_\beta \tau_\alpha = \tau_\beta \tau_\alpha \tau_\beta.
216   \]
217 \end{observation}
218 
219 A perhaps less obvious fact about Dehn twists is the following.
220 
221 \begin{theorem}\label{thm:mcg-is-fg}
222   Let \(\Sigma_{g, r}^b\) be the orientable surface of genus \(g \ge 1\) with
223   \(r\) punctures and \(b\) boundary components. Then the pure mapping class
224   group \(\PMod(\Sigma_{g, r}^b)\) is generated by finitely many Dehn twists
225   about nonseparating curves or boundary components.
226 \end{theorem}
227 
228 The proof of Theorem~\ref{thm:mcg-is-fg} is simple in nature: we proceed by
229 induction in \(g\), \(b\) and \(r\). On the other hand, the induction steps
230 require two ingredients we have not encountered so far, namely the \emph{Birman
231 exact sequence} and the \emph{modified graph of curves}. We now provide a
232 concise account of these ingredients.
233 
234 \section{The Birman Exact Sequence}
235 
236 Having the proof of Theorem~\ref{thm:mcg-is-fg} in mind, it is interesting to
237 consider the relationship between the mapping class group of \(\Sigma_{g,
238 r}^b\) and that of \(\Sigma_{g, r+1}^b = \Sigma_{g, r}^b \setminus \{ x \}\)
239 for some \(x\) in the interior \((\Sigma_{g, r}^b)\degree\) of \(\Sigma_{g,
240 r}^b\). Indeed, this will later allow us to establish the induction step on the
241 number of punctures \(r\).
242 
243 Given an orientable surface \(\Sigma\) and \(x_1, \ldots, x_n \in
244 \Sigma\degree\), denote by \(\Mod(\Sigma \setminus \{x_1, \ldots,
245 x_n\})_{\{x_1, \ldots, x_n\}} \subset \Mod(\Sigma \setminus \{x_1, \ldots,
246 x_n\})\) the subgroup of mapping classes \(f\) that permute \(x_1, \ldots,
247 x_n\) -- i.e. \(f \cdot x_i = x_{\sigma(i)}\) for some permutation \(\sigma \in
248 S_n\). We certainly have a surjective homomorphism
249 \begin{align*}
250   \operatorname{forget} :
251   \Mod(\Sigma \setminus \{x_1, \ldots, x_n\})_{\{x_1, \ldots, x_n\}}
252   & \to \Mod(\Sigma) \\
253   [\phi] & \mapsto [\tilde\phi]
254 \end{align*}
255 which ``\emph{forgets} the additional punctures \(x_1, \ldots,
256 x_n\) of \(\Sigma \setminus \{x_1, \ldots, x_n\}\),'' but what is its kernel?
257 
258 To answer this question, we consider the configuration space \(C(\Sigma, n) =
259 \mfrac{C^{\operatorname{ord}}(\Sigma, n)}{S_n}\) of \(n\) (unordered) points in
260 the interior of \(\Sigma\) -- where \(C^{\operatorname{ord}}(\Sigma, n) = \{
261   (x_1, \ldots, x_n) \in (\Sigma\degree)^n : x_i \ne x_j \ \text{for}\ i \ne j
262 \}\). Denote \(\Homeo^+(\Sigma, \partial \Sigma)_{x_1, \ldots, x_n} = \{\phi
263 \in \Homeo^+(\Sigma, \partial \Sigma) : \phi(x_i) = x_i \}\). From the
264 fibration\footnote{See \cite[Chapter~4]{hatcher} for a reference.}
265 \[
266   \arraycolsep=1.4pt
267   \begin{array}{ccrcl}
268     \Homeo^+(\Sigma, \partial \Sigma)_{x_1, \ldots, x_n}
269     & \hookrightarrow & \Homeo^+(\Sigma, \partial \Sigma)
270     & \relbar\joinrel\twoheadrightarrow & C(\Sigma, n) \\
271     & & \phi & \mapsto & [\phi(x_1), \ldots, \phi(x_n)]
272   \end{array}
273 \]
274 and its long exact sequence in homotopy we then obtain the following
275 fundamental result.
276 
277 \begin{theorem}[Birman exact sequence]\label{thm:birman-exact-seq}
278   Suppose \(\pi_1(\Homeo^+(\Sigma, \partial \Sigma), 1) = 1\). Then there is an
279   exact sequence
280   \begin{center}
281     \begin{tikzcd}[cramped]
282       1 \rar
283       & \pi_1(C(\Sigma, n), [x_1, \ldots, x_n]) \rar{\operatorname{push}}
284       & \Mod(\Sigma \setminus \{x_1, \ldots, x_n\})_{\{x_1, \ldots, x_n\}}
285         \rar{\operatorname{forget}}
286       & \Mod(\Sigma) \rar
287       & 1.
288     \end{tikzcd}
289   \end{center}
290 \end{theorem}
291 
292 \begin{remark}
293   Notice that \(C(\Sigma, 1) = \Sigma\degree \simeq \Sigma\). Hence for \(n =
294   1\) Theorem~\ref{thm:birman-exact-seq} gives us a sequence
295   \begin{center}
296     \begin{tikzcd}
297       1 \rar
298       & \pi_1(\Sigma, x) \rar{\operatorname{push}}
299       & \Mod(\Sigma \setminus \{x\}, x) \rar{\operatorname{forget}}
300       & \Mod(\Sigma) \rar
301       & 1.
302     \end{tikzcd}
303   \end{center}
304 \end{remark}
305 
306 We may regard a simple loop \(\alpha : \mathbb{S}^1 \to C(\Sigma, n)\) based at
307 \([x_1, \ldots, x_n]\) as \(n\) disjoint curves \(\alpha_1, \ldots, \alpha_n :
308 [0, 1] \to \Sigma\) with \(\alpha_i(0) = x_i\) and \(\alpha_i(1) =
309 x_{\sigma(i)}\) for some \(\sigma \in S_n\). The element
310 \(\operatorname{push}([\alpha]) \in \Mod(\Sigma)\) can then be seen as the
311 mapping class that ``\emph{pushes} a neighborhood of \(x_{\sigma(i)}\) towards
312 \(x_i\) along the curve \(\alpha_i^{-1}\),'' as shown in
313 Figure~\ref{fig:push-map} for the case \(n = 1\). Indeed, this goes to
314 show \(\operatorname{push}([\alpha])\) can be descrived as a product of Dehn
315 twists.
316 
317 \begin{fundamental-observation}\label{ex:push-simple-loop}
318   Using the notation of Figure~\ref{fig:push-map},
319   \(\operatorname{push}([\alpha]) = \tau_{\delta_1} \tau_{\delta_2}^{-1} \in
320   \Mod(\Sigma)\).
321 \end{fundamental-observation}
322 
323 \begin{figure}[ht]
324   \centering
325   \includegraphics[width=.35\linewidth]{images/push-map.eps}
326   \caption{The inclusion $\operatorname{push} : \pi_1(\Sigma, x) \to
327   \Mod(\Sigma)$ maps a simple loop $\alpha : \mathbb{S}^1 \to \Sigma$ to the
328   mapping class supported at an annular neighborhood $A$ of $\alpha$. Inside
329   this neighborhood, $\operatorname{push}([\alpha])$ takes the arc joining the
330   boundary components $\delta_i \subset \partial A$ on the left-hand side to
331   the yellow arc on the right-hand side.}
332   \label{fig:push-map}
333 \end{figure}
334 
335 \section{The Modified Graph of Curves}
336 
337 Having established Theorem~\ref{thm:birman-exact-seq}, we now need to address
338 the induction step in the genus \(g\) of \(\Sigma_{g, r}^b\). Our strategy is
339 to apply the following lemma from geometric group theory.
340 
341 \begin{lemma}\label{thm:ggt-lemma}
342   Let \(G\) be a group and \(\Gamma\) be a \emph{connected} graph with \(G
343   \leftaction \Gamma\) via graph automorphisms. Suppose that \(G\) acts
344   transitively on both \(V(\Gamma)\) and \(\{(v, w) \in V(\Gamma)^2 :
345   v \text{ --- } w \text{ in } \Gamma \}\). If \(v, w \in V(\Gamma)\) are
346   connected by an edge and \(g \in G\) is such that \(g \cdot w = v\) then
347   \(G\) is generated by \(g\) and the stabilizer \(G_v\).
348 \end{lemma}
349 
350 We are interested, of course, in the group \(G = \PMod(\Sigma_{g, r}^b)\). As
351 for the the role of \(\Gamma\), we consider the following graph.
352 
353 \begin{definition}
354   The \emph{modified graph of nonseparating curves
355   \(\hat{\mathcal{N}}(\Sigma)\) of a surface \(\Sigma\)} is the graph whose
356   vertices are (unoriented) isotopy classes of nonseparating simple closed
357   curves in \(\Sigma\) and
358   \[
359     \text{\([\alpha]\) --- \([\beta]\) in \(\hat{\mathcal{N}}(\Sigma)\)}
360     \iff \#(\alpha \cap \beta) = 1,
361   \]
362   where \(\#(\alpha \cap \beta)\) is the geometric intersection number of
363   \(\alpha\) and \(\beta\).
364 \end{definition}
365 
366 It is clear from the change of coordinates principle and
367 Observation~\ref{ex:change-of-coordinates-crossing} that the actions of
368 \(\Mod(\Sigma_{g, r}^b)\) on \(V(\hat{\mathcal{N}}(\Sigma_{g, r}^b))\) and
369 \(\{([\alpha], [\beta]) \in V(\hat{\mathcal{N}}(\Sigma_{g, r}^b))^2 : \#(\alpha
370 \cap \beta) = 1 \}\) are both transitive. But why should
371 \(\hat{\mathcal{N}}(\Sigma_{g, r}^b)\) be connected?
372 Historically, the modified graph of nonseparating curves first arose as a
373 \emph{modified} version of another graph, known as \emph{the graph of of
374 curves}.
375 
376 \begin{definition}
377   Given a surface \(\Sigma\), the \emph{graph of curves \(\mathcal{C}(\Sigma)\)
378   of \(\Sigma\)} is the graph whose vertices are (unoriented) isotopy classes
379   of essential simple closed curves in \(\Sigma\) and
380   \[
381     \text{\([\alpha]\) --- \([\beta]\) in \(\mathcal{C}(\Sigma)\)}
382     \iff \#(\alpha \cap \beta) = 0.
383   \]
384   The \emph{graph of nonseparating curves \(\mathcal{N}(\Sigma)\)} is the
385   subgraph of \(\mathcal{C}(\Sigma)\) whose vertices consist of nonseparating
386   curves.
387 \end{definition}
388 
389 Lickorish \cite{lickorish} essentially showed that, apart from a small number
390 of sporadic cases, \(\mathcal{C}(\Sigma_{g, r})\) is connected.
391 
392 \begin{theorem}
393   If \(\Sigma_{g, r}\) is not one \(\Sigma_0 = \mathbb{S}^2, \Sigma_{0, 1},
394   \ldots, \Sigma_{0, 4}, \Sigma_1 = \mathbb{T}^2\) and \(\Sigma_{1, 1}\) then
395   \(\mathcal{C}(\Sigma_{g, r})\) is connected.
396 \end{theorem}
397 
398 In other words, given simple closed curves \(\alpha, \beta \subset \Sigma_{g,
399 r}\), we can find closed \(\alpha = \alpha_1, \alpha_2, \ldots, \alpha_n =
400 \beta\) in \(\Sigma_{g, r}\) with \(\alpha_i\) disjoint from \(\alpha_{i+1}\).
401 Now if \(\alpha\) and \(\beta\) are nonseparating, by inductively adjusting
402 this sequence of curves we obtain the following corollary.
403 
404 \begin{corollary}\label{thm:mofied-graph-is-connected}
405   If \(g \ge 2\) then both \(\mathcal{N}(\Sigma_{g, r})\) and
406   \(\hat{\mathcal{N}}(\Sigma_{g, r})\) are connected.
407 \end{corollary}
408 
409 See \cite[Section~4.1]{farb-margalit} for a proof of
410 Corollary~\ref{thm:mofied-graph-is-connected}. We are now ready to show
411 Theorem~\ref{thm:mcg-is-fg}.
412 
413 \begin{proof}[Proof of Theorem~\ref{thm:mcg-is-fg}]
414   Let \(\Sigma_{g, r}^b\) be the orientable surface of genus \(g \ge 1\) with
415   \(r\) punctures and \(b\) boundary components. We want to establish that
416   \(\PMod(\Sigma_{g, r}^b)\) is generated by a finite number of Dehn twists
417   about nonseparating simple closed curves or boundary components. As promised,
418   we proceed by triple induction on \(r\), \(g\) and \(b\).
419 
420   For the base case, it is clear from Observation~\ref{ex:torus-mcg} and
421   Observation~\ref{ex:punctured-torus-mcg} that \(\Mod(\mathbb{T}^2) \cong
422   \Mod(\Sigma_{1, 1}) \cong \operatorname{SL}_2(\mathbb{Z})\) are generated by
423   the Dehn twists about the curves \(\alpha\) and \(\beta\) from
424   Figure~\ref{fig:torus-mcg-generators}, each corresponding to one of the
425   standard generators
426   \begin{align*}
427     \begin{pmatrix}
428       1 & 1 \\
429       0 & 1
430     \end{pmatrix}
431     &&
432     \begin{pmatrix}
433        1 & 0 \\
434       -1 & 1
435     \end{pmatrix}
436   \end{align*}
437   of \(\operatorname{SL}_2(\mathbb{Z})\).
438 
439   \begin{figure}[ht]
440     \centering
441     \includegraphics[width=.5\linewidth]{images/torus-mcg-generators.eps}
442     \caption{The curves $\alpha$ and $\beta$ whose Dehn twists generate
443     $\Mod(\mathbb{T}^2)$ and $\Mod(\Sigma_{1, 1})$.}
444     \label{fig:torus-mcg-generators}
445   \end{figure}
446 
447   Now suppose \(\PMod(\Sigma_{g, r})\) is finitely-generated by twists about
448   nonseparating curves for \(g \ge 2\) or \(g = 1\) and \(r > 1\). In both
449   case, \(\chi(\Sigma_{g, r}) = 2 - 2g - r < 0\) and thus
450   \(\pi_1(\Homeo^+(\Sigma_{g, r})) = 1\) -- see
451   \cite[Theorem~1.14]{farb-margalit}. The Birman exact sequence from
452   Theorem~\ref{thm:birman-exact-seq} then gives us
453   \begin{center}
454     \begin{tikzcd}
455       1 \rar
456       & \pi_1(\Sigma_{g, r}, x) \rar{\operatorname{push}}
457       & \PMod(\Sigma_{g, r + 1}) \rar{\operatorname{forget}}
458       & \PMod(\Sigma_{g, r}) \rar
459       & 1,
460     \end{tikzcd}
461   \end{center}
462   where \(\Sigma_{g, r + 1} = \Sigma_{g, r} \setminus \{x\}\). Since \(g \ge
463   1\), \(\pi_1(\Sigma_{g, r}, x)\) is generated by finitely many nonseparating
464   loops. We have seen in Observation~\ref{ex:push-simple-loop} that
465   \(\operatorname{push} : \pi_1(\Sigma_{g, r}, x) \to \Mod(\Sigma_{g, r+1},
466   x)\) takes nonseparating simple loops to products of twists about
467   nonseparating simple curves. Furthermore, we may lift the
468   generators of \(\PMod(\Sigma_{g, r})\) to Dehn twists about the corresponding
469   curves in \(\Sigma_{g, r + 1}\). This goes to show that
470   \(\PMod(\Sigma_{g, r + 1})\) is also generated by finitely many twists about
471   simple curves, concluding the induction step on \(r\).
472 
473   As for the induction step on \(g\), fix \(g \ge 1\) and suppose that, for
474   each \(r \ge 0\), \(\PMod(\Sigma_{g, r})\) is finitely generated by twists
475   about nonseparating curves or boundary components. Let us show that the same
476   holds for \(\Mod(\Sigma_{g + 1})\). To that end, we consider the action
477   \(\Mod(\Sigma_{g + 1}) \leftaction \hat{\mathcal{N}}(\Sigma_{g + 1})\). Since
478   \(g + 1 \ge 2\), \(\hat{\mathcal{N}}(\Sigma_{g + 1})\) is connected and the
479   conditions of Lemma~\ref{thm:ggt-lemma} are met. Now recall from
480   Observation~\ref{ex:braid-relation} that, given nonseparating \(\alpha, \beta
481   \subset \Sigma_{g + 1}\) crossing once, \(\tau_\beta \tau_\alpha \cdot
482   [\beta] = [\alpha]\). It thus follows from Lemma~\ref{thm:ggt-lemma} that
483   \(\Mod(\Sigma_{g + 1})\) is generated by \(\tau_\beta \tau_\alpha\) and
484   \(\Mod(\Sigma_{g + 1})_{[\alpha]} = \{ f \in \Mod(\Sigma_{g + 1}) : f \cdot
485   [\alpha] = [\alpha]\}\).
486 
487   In turn, \(\Mod(\Sigma_{g + 1})_{[\alpha]}\) has its index \(2\) subgroup
488   \[
489     \Mod(\Sigma_{g + 1})_{\vec{[\alpha]}}
490     = \{ f \in \Mod(\Sigma_{g + 1}) : f \cdot \vec{[\alpha]} = \vec{[\alpha]}\}
491   \]
492   of mapping classes fixing any given choice of orientation of \(\alpha\). One
493   can check that \(\tau_\beta \tau_\alpha^2 \tau_\beta \in \Mod(\Sigma_{g +
494   1})_{[\alpha]}\) inverts the orientation of \(\alpha\) and is thus a
495   representative of the nontrivial
496   \(\Mod(\Sigma_{g+1})_{\vec{[\alpha]}}\)-coset in
497   \(\Mod(\Sigma_{g+1})_{[\alpha]}\). In particular, \(\Mod(\Sigma_{g+1})\) is
498   generated by \(\Mod(\Sigma_{g+1})_{\vec{[\alpha]}}\), \(\tau_\beta
499   \tau_\alpha\) and \(\tau_\beta \tau_\alpha^2 \tau_\beta\).
500 
501   We now claim \(\Mod(\Sigma_{g+1})_{\vec{[\alpha]}}\) is generated by
502   finitely many twists about nonseparating curves. First observe that
503   \(\Sigma_{g+1} \setminus \alpha \cong \Sigma_{g,2}\), as shown in
504   Figure~\ref{fig:cut-along-nonseparating-adds-two-punctures}.
505   Observation~\ref{ex:cutting-morphism-kernel} then gives us an exact sequence
506   \begin{equation}\label{eq:cutting-seq}
507     \begin{tikzcd}
508       1 \rar &
509       \langle \tau_\alpha \rangle \rar &
510       \Mod(\Sigma_{g+1})_{\vec{[\alpha]}} \rar{\operatorname{cut}} &
511       \PMod(\Sigma_{g,2}) \rar &
512       1.
513     \end{tikzcd}
514   \end{equation}
515   But by the induction hypothesis, \(\PMod(\Sigma_{g, 2})\) is
516   finitely-generated by twists about nonseparating simple closed curves. As
517   before, these generators may be lifted to appropriate twists in
518   \(\Mod(\Sigma_{g+1})_{\vec{[\alpha]}}\). Now by (\ref{eq:cutting-seq}) we get
519   that \(\Mod(\Sigma_{g+1})_{\vec{[\alpha]}}\) is finitely generated by twists
520   about nonseparating curves, as desired. This concludes the induction step in
521   \(g\).
522 
523   \begin{figure}[ht]
524     \centering
525     \includegraphics[width=.75\linewidth]{images/cutting-homeo.eps}
526     \caption{The homeomorphism $\Sigma_{g + 1} \setminus \alpha \cong
527     \Sigma_{g, 2}$: removing the curve $\alpha$ has the same effect as cutting
528     along $\alpha$ and then capping the two resulting boundary components with
529     once-punctured disks, which gives us $\Sigma_{g, 2}$.}
530     \label{fig:cut-along-nonseparating-adds-two-punctures}
531   \end{figure}
532 
533   Finally, we handle the induction in \(b\). The boundaryless case \(b = 0\)
534   was already dealt with before. Now suppose \(\PMod(\Sigma_{g, s}^b)\) is
535   finitely generated by twists about simple closed curves or boundary
536   components for all \(g\) and \(s\). Fix some boundary component \(\delta
537   \subset \partial \Sigma_{g, r}^{b+1}\). From the homeomorphism \(\Sigma_{g,
538   r+1}^b \cong \Sigma_{g, r}^{b+1} \cup_\delta (\mathbb{D}^2 \setminus \{ 0
539   \})\) and the capping exact sequence from Observation~\ref{ex:capping-seq}
540   we obtain a sequence
541   \begin{center}
542     \begin{tikzcd}
543       1 \rar                                              &
544       \langle \tau_\delta \rangle \rar                    &
545       \PMod(\Sigma_{g, r}^{b+1}) \rar{\operatorname{cap}} &
546       \PMod(\Sigma_{g, r+1}^b) \rar                       &
547       1.
548     \end{tikzcd}
549   \end{center}
550   Now by induction hypothesis we may once again lift the generators of
551   \(\PMod(\Sigma_{g, r+1}^b)\) to Dehn twists about the corresponding curves in
552   \(\Sigma_{g, r}^{b+1}\) and add \(\tau_\delta\) to the generating set,
553   concluding the induction in \(b \ge 0\). We are done.
554 \end{proof}
555 
556 There are many possible improvements to this last result. For instance, in
557 \cite[Section~4.4]{farb-margalit} Farb-Margalit exhibit an explicit set of
558 generators of \(\Mod(\Sigma_g^b)\) by adapting the induction steps in the
559 proof of Theorem~\ref{thm:mcg-is-fg}. These are known as the \emph{Lickorish
560 generators}.
561 
562 \begin{theorem}[Lickorish generators]\label{thm:lickorish-gens}
563   If \(g \ge 1\) then \(\Mod(\Sigma_g^b)\) is generated by the Dehn twists
564   about the curves \(\alpha_1, \ldots, \alpha_g, \beta_1, \ldots, \beta_g,
565   \gamma_1, \ldots, \gamma_{g - 1}, \eta_1, \ldots, \eta_{b-1}\) as in
566   Figure~\ref{fig:lickorish-gens}
567 \end{theorem}
568 
569 In the boundaryless case \(b = 0\), we can write \(\tau_{\alpha_3}, \ldots,
570 \tau_{\alpha_g} \in \Mod(\Sigma_g)\) as products of the twists about the
571 remaining curves, from which we get the so-called \emph{Humphreys generators}.
572 
573 \begin{corollary}[Humphreys generators]\label{thm:humphreys-gens}
574   If \(g \ge 2\) then \(\Mod(\Sigma_g)\) is generated by the Dehn twists about the
575   curves \(\alpha_0, \ldots, \alpha_{2g}\) as in
576   Figure~\ref{fig:humphreys-gens}.
577 \end{corollary}
578 
579 \noindent
580 \begin{minipage}[b]{.47\linewidth}
581   \centering
582   \includegraphics[width=\linewidth]{images/lickorish-gens.eps}
583   \captionof{figure}{The curves from Lickorish generators of
584   $\Mod(\Sigma_g^b)$.}
585   \label{fig:lickorish-gens}
586 \end{minipage}
587 \hspace{.6cm} %
588 \begin{minipage}[b]{.47\textwidth}
589   \centering
590   \includegraphics[width=\linewidth]{images/humphreys-gens.eps}
591   \captionof{figure}{The curves from Humphreys generators of $\Mod(\Sigma_g)$.}
592   \label{fig:humphreys-gens}
593 \end{minipage}