memoire-m2

My M2 Memoire on mapping class groups & their representations

NameSizeMode
..
sections/twists.tex 27362B -rw-r--r--
001
002
003
004
005
006
007
008
009
010
011
012
013
014
015
016
017
018
019
020
021
022
023
024
025
026
027
028
029
030
031
032
033
034
035
036
037
038
039
040
041
042
043
044
045
046
047
048
049
050
051
052
053
054
055
056
057
058
059
060
061
062
063
064
065
066
067
068
069
070
071
072
073
074
075
076
077
078
079
080
081
082
083
084
085
086
087
088
089
090
091
092
093
094
095
096
097
098
099
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
\chapter{Dehn Twists}\label{ch:dehn-twists}

With the goal of studying the linear representations of mapping class groups in
mind, we now start investigating the group structure of \(\Mod(\Sigma)\). We
begin by computing some fundamental examples and then explore how we can use
these examples to understand the structure of the mapping class groups of other
surfaces. Namely, we compute \(\Mod(\mathbb{S}^1 \times [0, 1]) \cong
\mathbb{Z}\), and discuss how its generator gives rise to a convenient
generating set for \(\Mod(\Sigma)\), known as the set of \emph{Dehn twists}.

The idea here is to reproduce the proof of injectivity in
Observation~\ref{ex:torus-mcg}: by cutting along curves and arcs, we can always
decompose a surface into copies of \(\mathbb{D}^2\) and \(\mathbb{D}^2
\setminus \{0\}\). Observation~\ref{ex:alexander-trick} and
Observation~\ref{ex:mdg-once-punctured-disk} then imply the triviality of
mapping classes fixing such arcs and curves. Formally, this translates to the
following result.

\begin{proposition}[Alexander method]\label{thm:alexander-method}
  Let \(\alpha_1, \ldots, \alpha_n \subset \Sigma\) be essential simple closed
  curves or proper arcs satisfying the following conditions.
  \begin{enumerate}
    \item \([\alpha_i] \ne [\alpha_j]\) for \(i \ne j\).
    \item Each pair \((\alpha_i, \alpha_j)\) crosses at most once.
    \item Given distinct \(i, j, k\), at least one of \(\alpha_i \cap \alpha_j,
      \alpha_i \cap \alpha_k, \alpha_j \cap \alpha_k\) is empty.
    \item The surface obtained by cutting \(\Sigma\) along the \(\alpha_i\) is a
      disjoint union of disks and once-punctured disks.
  \end{enumerate}
  Suppose \(f \in \Mod(\Sigma)\) is such that \(f \cdot \vec{[\alpha_i]} =
  \vec{[\alpha_i]}\) for all \(i\). Then \(f = 1 \in \Mod(\Sigma)\).
\end{proposition}

See \cite[Proposition~2.8]{farb-margalit} for a proof of
Proposition~\ref{thm:alexander-method}. We now state some \emph{fundamental}
applications of the Alexander method.

\begin{example}\label{ex:mcg-annulus}
  The mapping class group \(\Mod(\mathbb{S}^1 \times [0, 1])\) is freely
  generated by \(f = [\phi]\), where
  \begin{align*}
    \phi : \mathbb{S}^1 \times [0, 1] & \isoto  \mathbb{S}^1 \times [0, 1] \\
                   (e^{2 \pi i t}, s) & \mapsto (e^{2 \pi i (t - s)}, s)
  \end{align*}
  is the map illustrated in Figure~\ref{fig:dehn-twist-cylinder}. In
  particular, \(\Mod(\mathbb{S}^1 \times [0, 1]) \cong \mathbb{Z}\).
\end{example}

\begin{example}\label{ex:mcg-twice-punctured-disk}
  The mapping class group \(\Mod(\mathbb{D}^2 \setminus \{-\sfrac{1}{2},
  \sfrac{1}{2}\})\) of the twice punctured unit disk in \(\mathbb{C}\) is
  freely generated by \(f = [\phi]\), where
  \begin{align*}
    \phi : \mathbb{D}^2 \setminus \{-\sfrac{1}{2}, \sfrac{1}{2}\}
    & \isoto \mathbb{D}^2 \setminus \{-\sfrac{1}{2}, \sfrac{1}{2}\} \\
    z & \mapsto -z
  \end{align*}
  is the map from Figure~\ref{fig:hald-twist-disk}. In particular,
  \(\Mod(\mathbb{D}^2 \setminus \{-\sfrac{1}{2}, \sfrac{1}{2}\}) \cong
  \mathbb{Z}\).
\end{example}

\noindent
\begin{minipage}[b]{.47\linewidth}
  \centering
  \includegraphics[width=.7\linewidth]{images/dehn-twist-cylinder.eps}
  \captionof{figure}{The generator $f$ of $\Mod(\mathbb{S}^1 \times [0, 1])
  \cong \mathbb{Z}$ takes the yellow arc on the left-hand side to the arc on
  the right-hand side that winds about the curve $\alpha$.}
  \label{fig:dehn-twist-cylinder}
\end{minipage}
\hspace{.6cm} %
\begin{minipage}[b]{.47\linewidth}
  \centering
  \includegraphics[width=.4\linewidth]{images/half-twist-disk.eps}
  \captionof{figure}{The generator $f$ of $\Mod(\mathbb{D}^2 \setminus
  \{-\sfrac{1}{2}, \sfrac{1}{2}) \cong \mathbb{Z}$ corresponds to the
  clockwise rotation by $\pi$ about the origin.}
  \label{fig:hald-twist-disk}
\end{minipage}

Let \(\Sigma\) be an orientable surface, possibly with punctures and non-empty
boundary. Given some closed \(\alpha \subset \Sigma\), we may envision doing
something similar to Example~\ref{ex:mcg-annulus} in \(\Sigma\) by looking at
annular neighborhoods of \(\alpha\). These are precisely the \emph{Dehn twists},
illustrated in Figure~\ref{fig:dehn-twist-bitorus} in the case of the bitorus
\(\Sigma_2\).

\begin{definition}
  Given a simple closed curve \(\alpha \subset \Sigma\), fix a closed annular
  neighborhood \(A \subset \Sigma\) of \(\alpha\) -- i.e. \(A \cong
  \mathbb{S}^1 \times [0, 1]\). Let \(f \in \Mod(A) \cong \Mod(\mathbb{S}^1
  \times [0, 1]) \cong \mathbb{Z}\) be the generator from
  Example~\ref{ex:mcg-annulus}. The \emph{Dehn twist \(\tau_\alpha \in
  \Mod(\Sigma)\) about \(\alpha\)} is defined as the image of \(f\) under the
  inclusion homomorphism \(\Mod(A) \to \Mod(\Sigma)\).
\end{definition}

\begin{figure}[ht]
  \centering
  \includegraphics[width=.6\linewidth]{images/dehn-twist-bitorus.eps}
  \caption{The Dehn twist about the curve $\alpha$ takes the peanut-shaped curve
  on the left-hand side to the yellow curve on the right-hand side.}
  \label{fig:dehn-twist-bitorus}
\end{figure}

Similarly, using the description of the mapping class group of the
twice-puncture disk derived in Example~\ref{ex:mcg-twice-punctured-disk}, the
generator of \(\Mod(\mathbb{D}^2 \setminus \{-\sfrac{1}{2}, \sfrac{1}{2}\})\)
gives rise the so-called \emph{half-twists}. These are examples of mapping
classes that permute the punctures of \(\Sigma\).

\begin{definition}
  Given an arc \(\alpha \subset \Sigma\) joining two punctures in the interior
  of \(\Sigma\), fix a closed neighborhood \(D \subset \Sigma\) of \(\alpha\)
  with \(D \cong \mathbb{D}^2 \setminus \{-\sfrac{1}{2}, \sfrac{1}{2}\}\). Let
  \(f \in \Mod(D) \cong \Mod(\mathbb{D}^2 \setminus \{-\sfrac{1}{2},
  \sfrac{1}{2}\}) \cong \mathbb{Z}\) be the generator from
  Example~\ref{ex:mcg-twice-punctured-disk}. The \emph{half-twist \(h_\alpha
  \in \Mod(\Sigma)\) about \(\alpha\)} is defined as the image of \(f\) under
  the inclusion homomorphism \(\Mod(D) \to \Mod(\Sigma)\).
\end{definition}

We can use the Alexander method to describe the kernel of capping and cutting
morphisms in terms of Dehn twists.

\begin{observation}[Capping exact sequence]\label{ex:capping-seq}
  Let \(\delta \subset \partial \Sigma\) be a boundary component of \(\Sigma\)
  and \(\operatorname{cap} : \Mod(\Sigma) \to \Mod(\Sigma \cup_\delta
  (\mathbb{D}^2 \setminus \{0\}))\) be the corresponding the capping
  homomorphism from Example~\ref{ex:capping-morphism}. There is an exact
  sequence
  \begin{center}
    \begin{tikzcd}
      1 \rar &
      \langle \tau_\delta \rangle \rar &
      \Mod(\Sigma) \rar{\operatorname{cap}} &
      \Mod(\Sigma \cup_\delta (\mathbb{D}^2 \setminus \{0\}), 0) \rar &
      1,
    \end{tikzcd}
  \end{center}
  known as \emph{the capping exact sequence} -- see
  \cite[Proposition~3.19]{farb-margalit} for a proof.
\end{observation}

\begin{observation}\label{ex:cutting-morphism-kernel}
  Let \(\alpha \subset \Sigma\) be a simple closed curve and
  \(\operatorname{cut} : \Mod(\Sigma)_{\vec{[\alpha]}} \to \Mod(\Sigma
  \setminus \alpha)\) be the cutting homomorphism from
  Example~\ref{ex:cutting-morphism}. Then \(\ker \operatorname{cut} = \langle
  \tau_\alpha \rangle \cong \mathbb{Z}\).
\end{observation}

It is also interesting to study how the geometry of two curves affects the
relationship between their corresponding Dehn twists. For instance,
by investigating the geometric intersection number
\[
  \#(\alpha \cap \beta) = \min
  \left\{
    |\alpha' \cap \beta'| : [\alpha'] = [\alpha]
    \text{ and }
    [\beta'] = [\beta]
  \right\}
\]
we can distinguish between powers of Dehn twists
\cite[Proposition~3.2]{farb-margalit}.

\begin{proposition}\label{thm:twist-intersection-number}
  Let \(\alpha \subset \Sigma\) be a simple closed curve and \(T_\alpha\) be a
  representative of \(\tau_\alpha \in \Mod(\Sigma)\). Then \(\#
  (T_\alpha^k(\beta) \cap \beta) = |k| \cdot \#(\alpha \cap \beta)^2\) for any
  \(k \in \mathbb{Z}\). In particular, if \(\alpha\) is nontrivial then
  \(\tau_\alpha\) has infinite order.
\end{proposition}

\begin{observation}
  Given \(\alpha, \beta \subset \Sigma\), \(\tau_\alpha = \tau_\beta \iff
  [\alpha] = [\beta]\). Indeed, if \(\alpha\) and \(\beta\) are non-isotopic,
  we can find \(\gamma\) with \(\#(\gamma \cap \alpha) > 0\) and \(\#(\gamma
  \cap \beta) = 0\). It thus follows from
  Proposition~\ref{thm:twist-intersection-number} that \(\#(T_\alpha(\gamma)
  \cap \gamma) > \#(T_\beta(\gamma) \cap \gamma)\), so \(\tau_\alpha \ne
  \tau_\beta\).
\end{observation}

Many other relations between Dehn twists can derived be in a geometric fashion
too.

\begin{observation}\label{ex:conjugate-twists}
  Given \(f = [\phi] \in \Mod(\Sigma)\), \(\tau_{\phi(\alpha)} = f \tau_\alpha
  f^{-1}\).
\end{observation}

\begin{observation}[Disjointness relations]
  Given \(f \in \Mod(\Sigma)\), \([f, \tau_\alpha] = 1 \iff f \cdot [\alpha] =
  [\alpha]\). In particular, \([\tau_\alpha, \tau_\beta] = 1\) for \(\alpha\)
  and \(\beta\) disjoint, for we can choose a representative of \(\tau_\beta\)
  whose support is disjoint from \(\alpha\).
\end{observation}

\begin{observation}
  If \(\alpha, \beta \subset \Sigma\) are both nonseparating then \(\tau_\alpha,
  \tau_\beta \in \Mod(\Sigma)\) are conjugate. Indeed, by the change of
  coordinates principle we can find \(f \in \Mod(\Sigma)\) with \(f \cdot
  [\alpha] = [\beta]\) and then apply Observation~\ref{ex:conjugate-twists}.
\end{observation}

\begin{observation}[Braid relations]\label{ex:braid-relation}
  Given \(\alpha, \beta \subset \Sigma\) with \(\#(\alpha \cap \beta) = 1\), it
  is not hard to check that \(\tau_\beta \tau_\alpha \cdot [\beta] =
  [\alpha]\). From Observation~\ref{ex:conjugate-twists} we then get
  \((\tau_\alpha \tau_\beta) \tau_\alpha (\tau_\alpha \tau_\beta)^{-1} =
  \tau_\beta\), from which follows the \emph{braid relation}
  \[
    \tau_\alpha \tau_\beta \tau_\alpha = \tau_\beta \tau_\alpha \tau_\beta.
  \]
\end{observation}

A perhaps less obvious fact about Dehn twists is the following.

\begin{theorem}\label{thm:mcg-is-fg}
  Let \(\Sigma_{g, r}^b\) be the orientable surface of genus \(g \ge 1\) with
  \(r\) punctures and \(b\) boundary components. Then the pure mapping class
  group \(\PMod(\Sigma_{g, r}^b)\) is generated by finitely many Dehn twists
  about nonseparating curves or boundary components.
\end{theorem}

The proof of Theorem~\ref{thm:mcg-is-fg} is simple in nature: we proceed by
induction in \(g\), \(b\) and \(r\). On the other hand, the induction steps
require two ingredients we have not encountered so far, namely the \emph{Birman
exact sequence} and the \emph{modified graph of curves}. We now provide a
concise account of these ingredients.

\section{The Birman Exact Sequence}

Having the proof of Theorem~\ref{thm:mcg-is-fg} in mind, it is interesting to
consider the relationship between the mapping class group of \(\Sigma_{g,
r}^b\) and that of \(\Sigma_{g, r+1}^b = \Sigma_{g, r}^b \setminus \{ x \}\)
for some \(x\) in the interior \((\Sigma_{g, r}^b)\degree\) of \(\Sigma_{g,
r}^b\). Indeed, this will later allow us to establish the induction step on the
number of punctures \(r\).

Given an orientable surface \(\Sigma\) and \(x_1, \ldots, x_n \in
\Sigma\degree\), denote by \(\Mod(\Sigma \setminus \{x_1, \ldots,
x_n\})_{\{x_1, \ldots, x_n\}} \subset \Mod(\Sigma \setminus \{x_1, \ldots,
x_n\})\) the subgroup of mapping classes \(f\) that permute \(x_1, \ldots,
x_n\) -- i.e. \(f \cdot x_i = x_{\sigma(i)}\) for some permutation \(\sigma \in
S_n\). We certainly have a surjective homomorphism
\begin{align*}
  \operatorname{forget} :
  \Mod(\Sigma \setminus \{x_1, \ldots, x_n\})_{\{x_1, \ldots, x_n\}}
  & \to \Mod(\Sigma) \\
  [\phi] & \mapsto [\tilde\phi]
\end{align*}
which ``\emph{forgets} the additional punctures \(x_1, \ldots,
x_n\) of \(\Sigma \setminus \{x_1, \ldots, x_n\}\),'' but what is its kernel?

To answer this question, we consider the configuration space \(C(\Sigma, n) =
\mfrac{C^{\operatorname{ord}}(\Sigma, n)}{S_n}\) of \(n\) (unordered) points in
the interior of \(\Sigma\) -- where \(C^{\operatorname{ord}}(\Sigma, n) = \{
  (x_1, \ldots, x_n) \in (\Sigma\degree)^n : x_i \ne x_j \ \text{for}\ i \ne j
\}\). Denote \(\Homeo^+(\Sigma, \partial \Sigma)_{x_1, \ldots, x_n} = \{\phi
\in \Homeo^+(\Sigma, \partial \Sigma) : \phi(x_i) = x_i \}\). From the
fibration\footnote{See \cite[Chapter~4]{hatcher} for a reference.}
\[
  \arraycolsep=1.4pt
  \begin{array}{ccrcl}
    \Homeo^+(\Sigma, \partial \Sigma)_{x_1, \ldots, x_n}
    & \hookrightarrow & \Homeo^+(\Sigma, \partial \Sigma)
    & \relbar\joinrel\twoheadrightarrow & C(\Sigma, n) \\
    & & \phi & \mapsto & [\phi(x_1), \ldots, \phi(x_n)]
  \end{array}
\]
and its long exact sequence in homotopy we then obtain the following
fundamental result.

\begin{theorem}[Birman exact sequence]\label{thm:birman-exact-seq}
  Suppose \(\pi_1(\Homeo^+(\Sigma, \partial \Sigma), 1) = 1\). Then there is an
  exact sequence
  \begin{center}
    \begin{tikzcd}[cramped]
      1 \rar
      & \pi_1(C(\Sigma, n), [x_1, \ldots, x_n]) \rar{\operatorname{push}}
      & \Mod(\Sigma \setminus \{x_1, \ldots, x_n\})_{\{x_1, \ldots, x_n\}}
        \rar{\operatorname{forget}}
      & \Mod(\Sigma) \rar
      & 1.
    \end{tikzcd}
  \end{center}
\end{theorem}

\begin{remark}
  Notice that \(C(\Sigma, 1) = \Sigma\degree \simeq \Sigma\). Hence for \(n =
  1\) Theorem~\ref{thm:birman-exact-seq} gives us a sequence
  \begin{center}
    \begin{tikzcd}
      1 \rar
      & \pi_1(\Sigma, x) \rar{\operatorname{push}}
      & \Mod(\Sigma \setminus \{x\}, x) \rar{\operatorname{forget}}
      & \Mod(\Sigma) \rar
      & 1.
    \end{tikzcd}
  \end{center}
\end{remark}

We may regard a simple loop \(\alpha : \mathbb{S}^1 \to C(\Sigma, n)\) based at
\([x_1, \ldots, x_n]\) as \(n\) disjoint curves \(\alpha_1, \ldots, \alpha_n :
[0, 1] \to \Sigma\) with \(\alpha_i(0) = x_i\) and \(\alpha_i(1) =
x_{\sigma(i)}\) for some \(\sigma \in S_n\). The element
\(\operatorname{push}([\alpha]) \in \Mod(\Sigma)\) can then be seen as the
mapping class that ``\emph{pushes} a neighborhood of \(x_{\sigma(i)}\) towards
\(x_i\) along the curve \(\alpha_i^{-1}\),'' as shown in
Figure~\ref{fig:push-map} for the case \(n = 1\). Indeed, this goes to
show \(\operatorname{push}([\alpha])\) can be descrived as a product of Dehn
twists.

\begin{fundamental-observation}\label{ex:push-simple-loop}
  Using the notation of Figure~\ref{fig:push-map},
  \(\operatorname{push}([\alpha]) = \tau_{\delta_1} \tau_{\delta_2}^{-1} \in
  \Mod(\Sigma)\).
\end{fundamental-observation}

\begin{figure}[ht]
  \centering
  \includegraphics[width=.35\linewidth]{images/push-map.eps}
  \caption{The inclusion $\operatorname{push} : \pi_1(\Sigma, x) \to
  \Mod(\Sigma)$ maps a simple loop $\alpha : \mathbb{S}^1 \to \Sigma$ to the
  mapping class supported at an annular neighborhood $A$ of $\alpha$. Inside
  this neighborhood, $\operatorname{push}([\alpha])$ takes the arc joining the
  boundary components $\delta_i \subset \partial A$ on the left-hand side to
  the yellow arc on the right-hand side.}
  \label{fig:push-map}
\end{figure}

\section{The Modified Graph of Curves}

Having established Theorem~\ref{thm:birman-exact-seq}, we now need to address
the induction step in the genus \(g\) of \(\Sigma_{g, r}^b\). Our strategy is
to apply the following lemma from geometric group theory.

\begin{lemma}\label{thm:ggt-lemma}
  Let \(G\) be a group and \(\Gamma\) be a \emph{connected} graph with \(G
  \leftaction \Gamma\) via graph automorphisms. Suppose that \(G\) acts
  transitively on both \(V(\Gamma)\) and \(\{(v, w) \in V(\Gamma)^2 :
  v \text{ --- } w \text{ in } \Gamma \}\). If \(v, w \in V(\Gamma)\) are
  connected by an edge and \(g \in G\) is such that \(g \cdot w = v\) then
  \(G\) is generated by \(g\) and the stabilizer \(G_v\).
\end{lemma}

We are interested, of course, in the group \(G = \PMod(\Sigma_{g, r}^b)\). As
for the the role of \(\Gamma\), we consider the following graph.

\begin{definition}
  The \emph{modified graph of nonseparating curves
  \(\hat{\mathcal{N}}(\Sigma)\) of a surface \(\Sigma\)} is the graph whose
  vertices are (unoriented) isotopy classes of nonseparating simple closed
  curves in \(\Sigma\) and
  \[
    \text{\([\alpha]\) --- \([\beta]\) in \(\hat{\mathcal{N}}(\Sigma)\)}
    \iff \#(\alpha \cap \beta) = 1,
  \]
  where \(\#(\alpha \cap \beta)\) is the geometric intersection number of
  \(\alpha\) and \(\beta\).
\end{definition}

It is clear from the change of coordinates principle and
Observation~\ref{ex:change-of-coordinates-crossing} that the actions of
\(\Mod(\Sigma_{g, r}^b)\) on \(V(\hat{\mathcal{N}}(\Sigma_{g, r}^b))\) and
\(\{([\alpha], [\beta]) \in V(\hat{\mathcal{N}}(\Sigma_{g, r}^b))^2 : \#(\alpha
\cap \beta) = 1 \}\) are both transitive. But why should
\(\hat{\mathcal{N}}(\Sigma_{g, r}^b)\) be connected?
Historically, the modified graph of nonseparating curves first arose as a
\emph{modified} version of another graph, known as \emph{the graph of of
curves}.

\begin{definition}
  Given a surface \(\Sigma\), the \emph{graph of curves \(\mathcal{C}(\Sigma)\)
  of \(\Sigma\)} is the graph whose vertices are (unoriented) isotopy classes
  of essential simple closed curves in \(\Sigma\) and
  \[
    \text{\([\alpha]\) --- \([\beta]\) in \(\mathcal{C}(\Sigma)\)}
    \iff \#(\alpha \cap \beta) = 0.
  \]
  The \emph{graph of nonseparating curves \(\mathcal{N}(\Sigma)\)} is the
  subgraph of \(\mathcal{C}(\Sigma)\) whose vertices consist of nonseparating
  curves.
\end{definition}

Lickorish \cite{lickorish} essentially showed that, apart from a small number
of sporadic cases, \(\mathcal{C}(\Sigma_{g, r})\) is connected.

\begin{theorem}
  If \(\Sigma_{g, r}\) is not one \(\Sigma_0 = \mathbb{S}^2, \Sigma_{0, 1},
  \ldots, \Sigma_{0, 4}, \Sigma_1 = \mathbb{T}^2\) and \(\Sigma_{1, 1}\) then
  \(\mathcal{C}(\Sigma_{g, r})\) is connected.
\end{theorem}

In other words, given simple closed curves \(\alpha, \beta \subset \Sigma_{g,
r}\), we can find closed \(\alpha = \alpha_1, \alpha_2, \ldots, \alpha_n =
\beta\) in \(\Sigma_{g, r}\) with \(\alpha_i\) disjoint from \(\alpha_{i+1}\).
Now if \(\alpha\) and \(\beta\) are nonseparating, by inductively adjusting
this sequence of curves we obtain the following corollary.

\begin{corollary}\label{thm:mofied-graph-is-connected}
  If \(g \ge 2\) then both \(\mathcal{N}(\Sigma_{g, r})\) and
  \(\hat{\mathcal{N}}(\Sigma_{g, r})\) are connected.
\end{corollary}

See \cite[Section~4.1]{farb-margalit} for a proof of
Corollary~\ref{thm:mofied-graph-is-connected}. We are now ready to show
Theorem~\ref{thm:mcg-is-fg}.

\begin{proof}[Proof of Theorem~\ref{thm:mcg-is-fg}]
  Let \(\Sigma_{g, r}^b\) be the orientable surface of genus \(g \ge 1\) with
  \(r\) punctures and \(b\) boundary components. We want to establish that
  \(\PMod(\Sigma_{g, r}^b)\) is generated by a finite number of Dehn twists
  about nonseparating simple closed curves or boundary components. As promised,
  we proceed by triple induction on \(r\), \(g\) and \(b\).

  For the base case, it is clear from Observation~\ref{ex:torus-mcg} and
  Observation~\ref{ex:punctured-torus-mcg} that \(\Mod(\mathbb{T}^2) \cong
  \Mod(\Sigma_{1, 1}) \cong \operatorname{SL}_2(\mathbb{Z})\) are generated by
  the Dehn twists about the curves \(\alpha\) and \(\beta\) from
  Figure~\ref{fig:torus-mcg-generators}, each corresponding to one of the
  standard generators
  \begin{align*}
    \begin{pmatrix}
      1 & 1 \\
      0 & 1
    \end{pmatrix}
    &&
    \begin{pmatrix}
       1 & 0 \\
      -1 & 1
    \end{pmatrix}
  \end{align*}
  of \(\operatorname{SL}_2(\mathbb{Z})\).

  \begin{figure}[ht]
    \centering
    \includegraphics[width=.5\linewidth]{images/torus-mcg-generators.eps}
    \caption{The curves $\alpha$ and $\beta$ whose Dehn twists generate
    $\Mod(\mathbb{T}^2)$ and $\Mod(\Sigma_{1, 1})$.}
    \label{fig:torus-mcg-generators}
  \end{figure}

  Now suppose \(\PMod(\Sigma_{g, r})\) is finitely-generated by twists about
  nonseparating curves for \(g \ge 2\) or \(g = 1\) and \(r > 1\). In both
  case, \(\chi(\Sigma_{g, r}) = 2 - 2g - r < 0\) and thus
  \(\pi_1(\Homeo^+(\Sigma_{g, r})) = 1\) -- see
  \cite[Theorem~1.14]{farb-margalit}. The Birman exact sequence from
  Theorem~\ref{thm:birman-exact-seq} then gives us
  \begin{center}
    \begin{tikzcd}
      1 \rar
      & \pi_1(\Sigma_{g, r}, x) \rar{\operatorname{push}}
      & \PMod(\Sigma_{g, r + 1}) \rar{\operatorname{forget}}
      & \PMod(\Sigma_{g, r}) \rar
      & 1,
    \end{tikzcd}
  \end{center}
  where \(\Sigma_{g, r + 1} = \Sigma_{g, r} \setminus \{x\}\). Since \(g \ge
  1\), \(\pi_1(\Sigma_{g, r}, x)\) is generated by finitely many nonseparating
  loops. We have seen in Observation~\ref{ex:push-simple-loop} that
  \(\operatorname{push} : \pi_1(\Sigma_{g, r}, x) \to \Mod(\Sigma_{g, r+1},
  x)\) takes nonseparating simple loops to products of twists about
  nonseparating simple curves. Furthermore, we may lift the
  generators of \(\PMod(\Sigma_{g, r})\) to Dehn twists about the corresponding
  curves in \(\Sigma_{g, r + 1}\). This goes to show that
  \(\PMod(\Sigma_{g, r + 1})\) is also generated by finitely many twists about
  simple curves, concluding the induction step on \(r\).

  As for the induction step on \(g\), fix \(g \ge 1\) and suppose that, for
  each \(r \ge 0\), \(\PMod(\Sigma_{g, r})\) is finitely generated by twists
  about nonseparating curves or boundary components. Let us show that the same
  holds for \(\Mod(\Sigma_{g + 1})\). To that end, we consider the action
  \(\Mod(\Sigma_{g + 1}) \leftaction \hat{\mathcal{N}}(\Sigma_{g + 1})\). Since
  \(g + 1 \ge 2\), \(\hat{\mathcal{N}}(\Sigma_{g + 1})\) is connected and the
  conditions of Lemma~\ref{thm:ggt-lemma} are met. Now recall from
  Observation~\ref{ex:braid-relation} that, given nonseparating \(\alpha, \beta
  \subset \Sigma_{g + 1}\) crossing once, \(\tau_\beta \tau_\alpha \cdot
  [\beta] = [\alpha]\). It thus follows from Lemma~\ref{thm:ggt-lemma} that
  \(\Mod(\Sigma_{g + 1})\) is generated by \(\tau_\beta \tau_\alpha\) and
  \(\Mod(\Sigma_{g + 1})_{[\alpha]} = \{ f \in \Mod(\Sigma_{g + 1}) : f \cdot
  [\alpha] = [\alpha]\}\).

  In turn, \(\Mod(\Sigma_{g + 1})_{[\alpha]}\) has its index \(2\) subgroup
  \[
    \Mod(\Sigma_{g + 1})_{\vec{[\alpha]}}
    = \{ f \in \Mod(\Sigma_{g + 1}) : f \cdot \vec{[\alpha]} = \vec{[\alpha]}\}
  \]
  of mapping classes fixing any given choice of orientation of \(\alpha\). One
  can check that \(\tau_\beta \tau_\alpha^2 \tau_\beta \in \Mod(\Sigma_{g +
  1})_{[\alpha]}\) inverts the orientation of \(\alpha\) and is thus a
  representative of the nontrivial
  \(\Mod(\Sigma_{g+1})_{\vec{[\alpha]}}\)-coset in
  \(\Mod(\Sigma_{g+1})_{[\alpha]}\). In particular, \(\Mod(\Sigma_{g+1})\) is
  generated by \(\Mod(\Sigma_{g+1})_{\vec{[\alpha]}}\), \(\tau_\beta
  \tau_\alpha\) and \(\tau_\beta \tau_\alpha^2 \tau_\beta\).

  We now claim \(\Mod(\Sigma_{g+1})_{\vec{[\alpha]}}\) is generated by
  finitely many twists about nonseparating curves. First observe that
  \(\Sigma_{g+1} \setminus \alpha \cong \Sigma_{g,2}\), as shown in
  Figure~\ref{fig:cut-along-nonseparating-adds-two-punctures}.
  Observation~\ref{ex:cutting-morphism-kernel} then gives us an exact sequence
  \begin{equation}\label{eq:cutting-seq}
    \begin{tikzcd}
      1 \rar &
      \langle \tau_\alpha \rangle \rar &
      \Mod(\Sigma_{g+1})_{\vec{[\alpha]}} \rar{\operatorname{cut}} &
      \PMod(\Sigma_{g,2}) \rar &
      1.
    \end{tikzcd}
  \end{equation}
  But by the induction hypothesis, \(\PMod(\Sigma_{g, 2})\) is
  finitely-generated by twists about nonseparating simple closed curves. As
  before, these generators may be lifted to appropriate twists in
  \(\Mod(\Sigma_{g+1})_{\vec{[\alpha]}}\). Now by (\ref{eq:cutting-seq}) we get
  that \(\Mod(\Sigma_{g+1})_{\vec{[\alpha]}}\) is finitely generated by twists
  about nonseparating curves, as desired. This concludes the induction step in
  \(g\).

  \begin{figure}[ht]
    \centering
    \includegraphics[width=.75\linewidth]{images/cutting-homeo.eps}
    \caption{The homeomorphism $\Sigma_{g + 1} \setminus \alpha \cong
    \Sigma_{g, 2}$: removing the curve $\alpha$ has the same effect as cutting
    along $\alpha$ and then capping the two resulting boundary components with
    once-punctured disks, which gives us $\Sigma_{g, 2}$.}
    \label{fig:cut-along-nonseparating-adds-two-punctures}
  \end{figure}

  Finally, we handle the induction in \(b\). The boundaryless case \(b = 0\)
  was already dealt with before. Now suppose \(\PMod(\Sigma_{g, s}^b)\) is
  finitely generated by twists about simple closed curves or boundary
  components for all \(g\) and \(s\). Fix some boundary component \(\delta
  \subset \partial \Sigma_{g, r}^{b+1}\). From the homeomorphism \(\Sigma_{g,
  r+1}^b \cong \Sigma_{g, r}^{b+1} \cup_\delta (\mathbb{D}^2 \setminus \{ 0
  \})\) and the capping exact sequence from Observation~\ref{ex:capping-seq}
  we obtain a sequence
  \begin{center}
    \begin{tikzcd}
      1 \rar                                              &
      \langle \tau_\delta \rangle \rar                    &
      \PMod(\Sigma_{g, r}^{b+1}) \rar{\operatorname{cap}} &
      \PMod(\Sigma_{g, r+1}^b) \rar                       &
      1.
    \end{tikzcd}
  \end{center}
  Now by induction hypothesis we may once again lift the generators of
  \(\PMod(\Sigma_{g, r+1}^b)\) to Dehn twists about the corresponding curves in
  \(\Sigma_{g, r}^{b+1}\) and add \(\tau_\delta\) to the generating set,
  concluding the induction in \(b \ge 0\). We are done.
\end{proof}

There are many possible improvements to this last result. For instance, in
\cite[Section~4.4]{farb-margalit} Farb-Margalit exhibit an explicit set of
generators of \(\Mod(\Sigma_g^b)\) by adapting the induction steps in the
proof of Theorem~\ref{thm:mcg-is-fg}. These are known as the \emph{Lickorish
generators}.

\begin{theorem}[Lickorish generators]\label{thm:lickorish-gens}
  If \(g \ge 1\) then \(\Mod(\Sigma_g^b)\) is generated by the Dehn twists
  about the curves \(\alpha_1, \ldots, \alpha_g, \beta_1, \ldots, \beta_g,
  \gamma_1, \ldots, \gamma_{g - 1}, \eta_1, \ldots, \eta_{b-1}\) as in
  Figure~\ref{fig:lickorish-gens}
\end{theorem}

In the boundaryless case \(b = 0\), we can write \(\tau_{\alpha_3}, \ldots,
\tau_{\alpha_g} \in \Mod(\Sigma_g)\) as products of the twists about the
remaining curves, from which we get the so-called \emph{Humphreys generators}.

\begin{corollary}[Humphreys generators]\label{thm:humphreys-gens}
  If \(g \ge 2\) then \(\Mod(\Sigma_g)\) is generated by the Dehn twists about the
  curves \(\alpha_0, \ldots, \alpha_{2g}\) as in
  Figure~\ref{fig:humphreys-gens}.
\end{corollary}

\noindent
\begin{minipage}[b]{.47\linewidth}
  \centering
  \includegraphics[width=\linewidth]{images/lickorish-gens.eps}
  \captionof{figure}{The curves from Lickorish generators of
  $\Mod(\Sigma_g^b)$.}
  \label{fig:lickorish-gens}
\end{minipage}
\hspace{.6cm} %
\begin{minipage}[b]{.47\textwidth}
  \centering
  \includegraphics[width=\linewidth]{images/humphreys-gens.eps}
  \captionof{figure}{The curves from Humphreys generators of $\Mod(\Sigma_g)$.}
  \label{fig:humphreys-gens}
\end{minipage}