memoire-m2
My M2 Memoire on mapping class groups & their representations
presentation.tex (24869B)
1 \chapter{Relations Between Twists}\label{ch:relations} 2 3 Having found a convenient set of generators for \(\Mod(\Sigma_g)\), it is now 4 natural to ask what the relations between such generators are. In this chapter, 5 we highlight some additional relations between Dehn twists and the geometric 6 intuition behind them, culminating in the statement of a presentation for 7 \(\Mod(\Sigma_g)\) whose relations can be entirely explained in terms of the 8 geometry of curves in \(\Sigma_g\) -- see 9 Theorem~\ref{thm:wajnryb-presentation}. 10 11 \begin{fundamental-observation}[Lantern relation] 12 Let \(\Sigma_0^4\) be the surface of genus \(0\) with \(4\) boundary 13 components and \(\alpha, \beta, \gamma, \delta_1, \ldots, \delta_4 \subset 14 \Sigma_0^4\) be as in Figure~\ref{fig:latern-relation}. Consider the surfaces 15 \(\Sigma_0^3 = \Sigma_0^4 \cup_{\delta_1} \mathbb{D}^2\) and \(\Sigma_{0,1}^3 16 = \Sigma_0^4 \cup_{\delta_1} (\mathbb{D}^2 \setminus \{ 0 \})\), as well as 17 the map \(\operatorname{push} : \pi_1(\Sigma_0^3, 0) \to 18 \Mod(\Sigma_{0,1}^3)\). Let \(\eta_1, \eta_2, \eta_3 : \mathbb{S}^1 \to 19 \Sigma_0^3\) be the loops from Figure~\ref{fig:lantern-relation-capped}, so 20 that \([\eta_1] \cdot [\eta_2] = [\eta_3]\) in \(\pi_1(\Sigma_0^3, 0)\). From 21 Observation~\ref{ex:push-simple-loop} we obtain 22 \[ 23 (\tau_{\delta_2} \tau_\alpha^{-1}) (\tau_{\delta_3} \tau_\gamma^{-1}) 24 = \operatorname{push}([\eta_1]) \cdot \operatorname{push}([\eta_2]) 25 = \operatorname{push}([\eta_3]) 26 = \tau_\beta \tau_{\delta_4}^{-1} 27 \in \Mod(\Sigma_{0, 1}^3). 28 \] 29 Using the capping exact sequence from Observation~\ref{ex:capping-seq}, we 30 can then see \(\tau_{\delta_2} \tau_\alpha^{-1} \tau_{\delta_3} 31 \tau_\gamma^{-1}, \tau_\beta \tau_{\delta_4}^{-1} \in \Mod(\Sigma_0^4)\) 32 differ by a power of \(\tau_{\delta_1}\). In fact, one can show 33 \((\tau_{\delta_2} \tau_\alpha^{-1} \tau_{\delta_3} \tau_\gamma^{-1}) 34 (\tau_\beta \tau_{\delta_4}^{-1})^{-1} = \tau_{\delta_1}^{-1} \in 35 \Mod(\Sigma_0^4)\). Now the disjointness relations \([\tau_{\delta_i}, 36 \tau_\alpha] = [\tau_{\delta_i}, \tau_\beta] = [\tau_{\delta_i}, \tau_\gamma] 37 = 1\) give us the \emph{lantern relation} (\ref{eq:lantern-relation}) in 38 \(\Mod(\Sigma_0^4)\). 39 \begin{equation}\label{eq:lantern-relation} 40 \tau_\alpha \tau_\beta \tau_\gamma 41 = \tau_{\delta_1} \tau_{\delta_2} \tau_{\delta_3} \tau_{\delta_4} 42 \end{equation} 43 \end{fundamental-observation} 44 45 \noindent 46 \begin{minipage}[t]{.5\linewidth} 47 \centering 48 \includegraphics[width=\linewidth]{images/lantern-relation.eps} 49 \captionof{figure}{Two views of $\Sigma_0^4$: on the left-hand side we see 50 the \emph{lantern-like} surface we get by subtracting \(4\) disjoint open 51 disks from \(\mathbb{S}^2\), and on the right-hand side we see the disk with 52 three open disks subtracted from its interior.} 53 \label{fig:latern-relation} 54 \end{minipage} 55 \hspace{.6cm} % 56 \begin{minipage}[t]{.44\linewidth} 57 \centering 58 \includegraphics[width=.45\linewidth]{images/lantern-relation-capped.eps} 59 \captionof{figure}{The curves $\eta_1, \eta_2, \eta_3 \subset \Sigma_0^3$ 60 from the proof of the lantern relation.} 61 \label{fig:lantern-relation-capped} 62 \end{minipage} 63 64 We may exploit different embeddings \(\Sigma_0^4 \hookrightarrow \Sigma\) and 65 their corresponding inclusion homomorphisms \(\Mod(\Sigma_0^4) \to 66 \Mod(\Sigma)\) to obtain interesting relations between the corresponding Dehn 67 twists in \(\Mod(\Sigma)\). For example, the lantern relation can be used to 68 compute \(\Mod(\Sigma_g^b)^\ab\) for \(g \ge 3\). 69 70 \begin{proposition}\label{thm:trivial-abelianization} 71 The Abelianization \(\Mod(\Sigma_g^b)^\ab = 72 \mfrac{\Mod(\Sigma_g^b)}{[\Mod(\Sigma_g), \Mod(\Sigma_g)]}\) is cyclic. 73 Moreover, if \(g \ge 3\) then \(\Mod(\Sigma_g^b)^\ab = 0\). In other words, 74 \(\Mod(\Sigma_g)\) is a perfect group for \(g \ge 3\). 75 \end{proposition} 76 77 \begin{proof} 78 By Theorem~\ref{thm:lickorish-gens}, \(\Mod(\Sigma_g^b)^\ab\) is generated by 79 the image of the Lickorish generators, which are all conjugate and thus 80 represent the same class in the Abelianization. In fact, any nonseparating 81 \(\alpha \subset \Sigma_g^b\) is conjugate to the Lickorish generators too, 82 so \(\Mod(\Sigma_g^b)^\ab = \langle [\alpha] \rangle\). 83 84 Now for \(g \ge 3\) we can embed \(\Sigma_0^4\) in \(\Sigma_g^b\) in such a 85 way that all the corresponding curves \(\alpha, \beta, \gamma, \delta_1, 86 \ldots, \delta_4 \subset \Sigma_g^b\) are nonseparating, as shown in 87 Figure~\ref{fig:latern-relation-trivial-abelianization}. The lantern relation 88 (\ref{eq:lantern-relation}) then becomes 89 \[ 90 3 \cdot [\tau_\alpha] 91 = [\tau_\alpha] + [\tau_\beta] + [\tau_\gamma] 92 = [\tau_{\delta_1}] + [\tau_{\delta_2}] 93 + [\tau_{\delta_3}] + [\tau_{\delta_4}] 94 = 4 \cdot [\tau_\alpha] 95 \] 96 in \(\Mod(\Sigma_g^b)^\ab\). In other words, \([\tau_\alpha] = 0\) and thus 97 \(\Mod(\Sigma_g^b)^\ab = 0\). 98 \end{proof} 99 100 \begin{figure}[ht] 101 \centering 102 \includegraphics[width=.5\linewidth]{images/lantern-relation-trivial-abelianization.eps} 103 \caption{The embedding of $\Sigma_0^4$ in $\Sigma_g^b$ for $g \ge 3$.} 104 \label{fig:latern-relation-trivial-abelianization} 105 \end{figure} 106 107 To get extra relations we need to investigate certain branched covers \(\Sigma 108 \to \mathbb{D}^2 \setminus \{x_1, \ldots, x_r\}\), as well as the relationship 109 between \(\Mod(\Sigma)\) and \(\Mod(\mathbb{D}^2 \setminus \{x_1, \ldots, 110 x_r\})\). This is what is known as \emph{the Birman-Hilden theorem}. 111 112 \section{The Birman-Hilden Theorem}\label{birman-hilden} 113 114 Let \(\Sigma_{0, r}^1 = \mathbb{D}^2 \setminus \{x_1, \ldots, x_r\}\) be the 115 surface of genus \(0\) with \(r\) punctures and one boundary component. We 116 begin our investigation by providing an alternative description of its mapping 117 class group. Namely, we show that \(\Mod(\Sigma_{0, r}^1)\) is the braid group 118 on \(r\) strands. 119 120 \begin{definition} 121 The \emph{braid group on \(n\) strands} \(B_n\) is the fundamental group 122 \(\pi_1(C(\mathbb{D}^2, n), *)\) of the configuration space \(C(\mathbb{D}^2, 123 n) = \mfrac{C^{\operatorname{ord}}(\mathbb{D}^2, n)}{S_n}\) of \(n\) points 124 in the interior of the disk. The elements of \(B_n\) are referred to as 125 \emph{braids}. 126 \end{definition} 127 128 \begin{example} 129 Given \(i = 1, \ldots, n-1\), we define \(\sigma_i \in B_n\) as in 130 Figure~\ref{fig:braid-group-generator}. 131 \end{example} 132 133 \begin{figure}[ht] 134 \centering 135 \includegraphics[width=.25\linewidth]{images/braid-group-generator.eps} 136 \caption{The braid $\sigma_i$.} 137 \label{fig:braid-group-generator} 138 \end{figure} 139 140 The third Reidemeister move translates to the so-called \emph{braid 141 relations} 142 \[ 143 \sigma_i \sigma_{i+1} \sigma_i = \sigma_{i+1} \sigma_1 \sigma_i 144 \] 145 in \(B_n\), which motivates the name used in 146 Observation~\ref{ex:braid-relation}. In his seminal paper on braid groups, 147 Artin \cite{artin} gave the following finite presentation of \(B_n\). 148 149 \begin{theorem}[Artin] 150 \[ 151 \arraycolsep=1.2pt 152 B_n = 153 \left\langle 154 \sigma_1, \ldots, \sigma_{n - 1} : 155 \begin{array}{rll} 156 \sigma_i \sigma_{i+1} \sigma_i & = \sigma_{i+1} \sigma_i \sigma_{i+1} & 157 \quad \text{for all} \ i, \\ 158 \sigma_i \sigma_j & = \sigma_j \sigma_i & 159 \quad \text{for} \ j \ne i + 1 \ \text{and} \ j \ne i - 1 160 \end{array} 161 \right\rangle. 162 \] 163 \end{theorem} 164 165 As promised, we now show that \(B_n\) coincides with \(\Mod(\Sigma_{0, n}^1)\). 166 Recall from Theorem~\ref{thm:birman-exact-seq} that there is an exact 167 sequence 168 \begin{center} 169 \begin{tikzcd} 170 1 \rar 171 & B_n \rar{\operatorname{push}} 172 & \Mod(\Sigma_{0, n}^1) \rar 173 & \cancelto{1}{\Mod(\mathbb{D}^2)} \rar 174 & 1, 175 \end{tikzcd} 176 \end{center} 177 for \(\Homeo^+(\mathbb{D}^2, \mathbb{S}^1)\) is contractible by 178 Observation~\ref{ex:alexander-trick}. We thus obtain the following result. 179 180 \begin{proposition} 181 The map \(\operatorname{push} : B_n \to \Mod(\Sigma_{0, n}^1)\) is a group 182 isomorphism. 183 \end{proposition} 184 185 \noindent 186 \begin{minipage}[b]{.47\linewidth} 187 \begin{observation}\label{ex:braid-group-center} 188 Using the capping exact sequence from Observation~\ref{ex:capping-seq} and 189 the Alexander method, one can check that the center \(Z(\Mod(\Sigma_{0, 190 n}^1))\) of \(\Mod(\Sigma_{0, n}^1)\) is freely generated by the Dehn twist 191 \(\tau_\delta\) about the boundary \(\delta = \partial \Sigma_{0, n}^1\). It 192 is also not very difficult to see that \(\operatorname{push} : B_n \to 193 \Mod(\Sigma_{0, n}^1)\) takes \(\sigma_1 \cdots \sigma_{n-1}\) to the 194 rotation by \(\sfrac{2\pi}{n}\) as in Figure~\ref{fig:braid-group-center}, 195 which is an \(n\)-th root of \(\tau_\delta\). Hence the center \(Z(B_n)\) is 196 freely generated by \(z = (\sigma_1 \cdots \sigma_{n - 1})^n\). 197 \end{observation} 198 \end{minipage} 199 \hspace{.6cm} % 200 \begin{minipage}[b]{.47\textwidth} 201 \centering 202 \includegraphics[width=.4\linewidth]{images/braid-group-center.eps} 203 \captionof{figure}{The clockwise rotation by $\sfrac{2\pi}{n}$ about an axis 204 centered around the punctures $x_1, \ldots, x_n$ of $\Sigma_{0, n}^1$.} 205 \label{fig:braid-group-center} 206 \end{minipage} 207 \smallskip 208 209 To get from \(\Sigma_{0, n}^1\) to surfaces of genus \(g > 0\) we may consider 210 the \emph{hyperelliptic involution} \(\iota : \Sigma_g \isoto \Sigma_g\), which 211 rotates \(\Sigma_g\) by \(\pi\) around some axis as in 212 Figure~\ref{fig:hyperelliptic-involution}. Given \(\ell < g\) and \(b = 1, 2\), 213 we can also embed \(\Sigma_\ell^b\) in \(\Sigma_g\) in such way that \(\iota\) 214 restricts to an involution\footnote{This involution does not fix $\partial 215 \Sigma_\ell^b$ point-wise.} \(\Sigma_\ell^b \isoto \Sigma_\ell^b\). 216 217 \begin{figure}[ht] 218 \centering 219 \includegraphics[width=\linewidth]{images/hyperelliptic-involution.eps} 220 \caption{The hyperelliptic involution $\iota$.} 221 \label{fig:hyperelliptic-involution} 222 \end{figure} 223 224 It is clear from Figure~\ref{fig:hyperelliptic-involution} that the quotients 225 \(\mfrac{\Sigma_\ell^1}{\iota}\) and \(\mfrac{\Sigma_\ell^2}{\iota}\) are both 226 disks, with boundary corresponding to the projection of the boundaries of 227 \(\Sigma_\ell^1\) and \(\Sigma_\ell^2\), respectively. Given \(b = 1, 2\), the 228 quotient map \(\Sigma_\ell^b \to \mfrac{\Sigma_\ell^b}{\iota} \cong 229 \mathbb{D}^2\) is a double cover with \(2\ell + b\) branch points corresponding 230 to the fixed points of \(\iota\). We may thus regard 231 \(\mfrac{\Sigma_\ell^b}{\iota}\) as the disk \(\Sigma_{0, 2\ell + b}^1\) with 232 \(2\ell + b\) punctures in its interior, as shown in 233 Figure~\ref{fig:hyperelliptic-covering}. We also draw the curves \(\alpha_1, 234 \ldots, \alpha_{2\ell} \subset \Sigma_\ell^b\) of the Humphreys generators of 235 \(\Mod(\Sigma_g)\). Since these curves are invariant under the action of 236 \(\iota\), they descend to arcs \(\bar{\alpha}_1, \ldots, \bar{\alpha}_{2\ell + b} \subset \Sigma_{0, 2\ell + b}^1\) joining the punctures of the quotient 237 surface. 238 239 \begin{figure}[ht] 240 \centering 241 \includegraphics[width=.77\linewidth]{images/hyperelliptic-covering.eps} 242 \caption{The double branched covers given by $\iota$.} 243 \label{fig:hyperelliptic-covering} 244 \end{figure} 245 246 \begin{observation}\label{ex:push-generators-description} 247 The map \(\operatorname{push} : B_{2\ell + b} \to \Mod(\Sigma_{0, 2\ell + b}^1)\) takes \(\sigma_i\) to the half-twist \(h_{\bar{\alpha}_i}\) about 248 the arc \(\bar{\alpha}_i \subset \Sigma_{0, 2\ell + b}^1\). 249 \end{observation} 250 251 We now study the homeomorphisms of \(\Sigma_\ell^1\) and \(\Sigma_\ell^2\) that 252 descend to the quotient surfaces and their mapping classes, known as \emph{the 253 symmetric mapping classes}. 254 255 \begin{definition} 256 Let \(\ell \ge 0\) and \(b = 1, 2\). The \emph{group of symmetric 257 homeomorphisms of \(\Sigma_\ell^b\)} is \(\SHomeo^+(\Sigma_\ell^b, \partial \Sigma_\ell^b) = 258 \{\phi \in \Homeo^+(\Sigma_\ell^b, \partial \Sigma_\ell^b) : [\phi, \iota] = 1\}\). The 259 \emph{symmetric mapping class group of \(\Sigma_\ell^b\)} is the subgroup 260 \(\SMod(\Sigma_\ell^1) = \{ [\phi] \in \Mod(\Sigma_\ell^b) : \phi \in 261 \SHomeo^+(\Sigma_\ell^b, \partial \Sigma_\ell^b) \}\). 262 \end{definition} 263 264 Fix \(b = 1\) or \(2\). It follows from the universal property of quotients 265 that any \(\phi \in \SHomeo^+(\Sigma_\ell^b, \partial \Sigma_\ell^b)\) defines 266 a homeomorphism \(\bar \phi : \Sigma_{0, 2\ell+b}^1 \isoto \Sigma_{0, 267 2\ell+b}^1\). This yields a homomorphism of topological groups 268 \begin{align*} 269 \SHomeo^+(\Sigma_\ell^b, \partial \Sigma_\ell^b) 270 & \to \Homeo^+(\Sigma_{0, 2\ell + b}^1, \partial \Sigma_{0, 2\ell + b}^1) \\ 271 \phi 272 & \mapsto \bar \phi, 273 \end{align*} 274 which is surjective because any \(\psi \in \Homeo^+(\Sigma_{0, 2\ell + b}^1, 275 \partial \Sigma_{0, 2\ell + b}^1)\) lifts to \(\Sigma_\ell^b\). 276 277 It is also not difficult to see \(\SHomeo^+(\Sigma_\ell^b, \partial 278 \Sigma_\ell^b) \to \Homeo^+(\Sigma_{0, 2\ell + b}^1, \partial \Sigma_{0, 2\ell 279 + b}^1)\) is injective: the only candidates for elements of its kernel are 280 \(1\) and \(\iota\), but \(\iota\) is not an element of 281 \(\SHomeo^+(\Sigma_\ell^b, \partial \Sigma_\ell^b)\) since it does not fix 282 \(\partial \Sigma_\ell^b\) point-wise. Now since we have a continuous bijective 283 homomorphism we find 284 \[ 285 \begin{split} 286 \pi_0(\SHomeo^+(\Sigma_\ell^b, \partial \Sigma_\ell^b)) 287 & \cong \pi_0(\Homeo^+(\Sigma_{0, 2\ell+b}^1, \partial \Sigma_{0, 2\ell+b}^1)) \\ 288 & = \mfrac{\Homeo^+(\Sigma_{0,2\ell+b}^1, \partial \Sigma_{0, 2\ell+b}^1)}{\simeq} \\ 289 & = \Mod(\Sigma_{0, 2\ell+b}^1) \\ 290 & \cong B_{2\ell + b}. 291 \end{split} 292 \] 293 294 We would like to say \(\pi_0(\SHomeo^+(\Sigma_\ell^b, \partial \Sigma_\ell^b)) = 295 \SMod(\Sigma_\ell^b)\), but a priori the story looks a little more complicated: 296 \(\phi, \psi \in \SHomeo^+(\Sigma_\ell^b, \partial \Sigma_\ell^b)\) define the same class 297 in \(\SMod(\Sigma_\ell^b)\) if they are isotopic, but they may not lie in same 298 connected component of \(\SHomeo^+(\Sigma_\ell^b, \partial \Sigma_\ell^b)\) if they are 299 not isotopic \emph{through symmetric homeomorphisms}. Birman-Hilden 300 \cite{birman-hilden} showed that this is never the case. 301 302 \begin{theorem}[Birman-Hilden] 303 If \(\phi, \psi \in \SHomeo^+(\Sigma_\ell^b, \partial \Sigma_\ell^b)\) are isotopic 304 then \(\phi\) and \(\psi\) are isotopic through symmetric homeomorphisms. In 305 particular, there is an isomorphism 306 \begin{align*} 307 \SMod(\Sigma_\ell^b) & \isoto \Mod(\Sigma_{0, 2\ell + b}) \\ 308 [\phi] & \mapsto [\bar \phi]. 309 \end{align*} 310 \end{theorem} 311 312 \begin{observation} 313 Using the notation of Figure~\ref{fig:hyperelliptic-covering}, the 314 Birman-Hilden isomorphism \(\SMod(\Sigma_\ell^b) \isoto \Mod(\Sigma_{0, 2g + b})\) takes \(\tau_{\alpha_i}\) to the half twist \(h_{\bar{\alpha}_i} \in 315 \Mod(\Sigma_{0, 2g + b})\). This can be checked by looking at 316 \(\iota\)-invariant annular neighborhoods of the curves \(\alpha_i\) -- 317 \cite[Section~9.4]{farb-margalit}. 318 \end{observation} 319 320 \begin{fundamental-observation}[$k$-chain relations] 321 The Birman-Hilden isomorphism \(\SMod(\Sigma_\ell^1) \isoto \Mod(\Sigma_{0, 322 2\ell+1}^1)\) takes the twists \(\tau_\delta \in \SMod(\Sigma_\ell^1)\) about 323 the boundary \(\delta = \partial \Sigma_\ell^1\) to \(\tau_{\bar\delta}^2 \in 324 \Mod(\Sigma_{0, 2\ell+1}^1)\). Similarly, \(\SMod(\Sigma_\ell^2) \isoto 325 \Mod(\Sigma_{0, 2\ell+2})\) takes \(\tau_{\delta_1} \tau_{\delta_2} \in 326 \SMod(\Sigma_\ell^2)\) to \(\tau_{\bar\delta_1} = \tau_{\bar\delta_2}\). In 327 light of Observation~\ref{ex:push-generators-description}, 328 Observation~\ref{ex:braid-group-center} translates into the so-called 329 \emph{\(k\)-chain relations} in \(\SMod(\Sigma_\ell^b) \subset 330 \Mod(\Sigma_g)\). 331 \[ 332 \arraycolsep=1.4pt 333 \begin{array}{rlcrll} 334 (\sigma_1 \cdots \sigma_k)^{2k+2} & = z^2 \in B_{k + 1} & 335 \; \rightsquigarrow & 336 \; (\tau_{\alpha_1} \cdots \tau_{\alpha_k})^{2k + 2} & = \tau_\delta & 337 \; \text{for } k = 2 \ell \text{ even} \\ 338 (\sigma_1 \cdots \sigma_k)^{k+1} & = z \in B_{k + 1} & 339 \; \rightsquigarrow & 340 \; (\tau_{\alpha_1} \cdots \tau_{\alpha_k})^{k + 1} 341 & = \tau_{\delta_1} \tau_{\delta_2} & 342 \; \text{for } k = 2 \ell + 1 \text{ odd} 343 \end{array} 344 \] 345 \end{fundamental-observation} 346 347 We may also exploit the quotient \(\mfrac{\Sigma_g}{\iota} \cong \mathbb{S}^2\) 348 to obtain other relations. Since \(\iota\) has \(2g + 2\) fixed points in 349 \(\Sigma_g\), we get branched double cover \(\Sigma_g \to \Sigma_{0, 2g+2}\). 350 351 \begin{theorem}[Birman-Hilden without boundary]\label{thm:boundaryless-birman-hilden} 352 If \(g \ge 2\) then we have an exact sequence 353 \begin{center} 354 \begin{tikzcd} 355 1 \rar 356 & \langle [\iota] \rangle \rar 357 & C_{\Mod(\Sigma_g)}([\iota]) \rar 358 & \Mod(\Sigma_{0, 2g + 2}) \rar 359 & 1, 360 \end{tikzcd} 361 \end{center} 362 where \(C_{\Mod(\Sigma_g)}([\iota]) \subset \Mod(\Sigma_g)\) is the 363 commutator subgroup of \([\iota]\) and the right map takes \([\phi] \in 364 C_{\Mod(\Sigma_g)}([\iota])\) to \([\bar \phi] \in \Mod(\Sigma_{0, 2g + 365 2})\). 366 \end{theorem} 367 368 \begin{fundamental-observation}[Hyperelliptic relations] 369 Let \(\alpha_1, \ldots, \alpha_{2g}, \delta \subset \Sigma_g\) be as in 370 Figure~\ref{fig:hyperellipitic-relations}. Then 371 \begin{equation}\label{eq:hyperelliptic-eq} 372 [\iota] 373 = \tau_\delta \tau_{\alpha_{2g}} \cdots \tau_{\alpha_1} 374 \tau_{\alpha_1} \cdots \tau_{\alpha_{2g}} \tau_\delta. 375 \end{equation} 376 Indeed, \(C_{\Mod(\Sigma_g)}([\iota]) \to \Mod(\Sigma_{0, 2g+2})\) takes 377 \(\tau_\delta \tau_{\alpha_{2g}} \cdots \tau_{\alpha_1}\) to the rotation 378 from Figure~\ref{fig:hyperelliptic-relation-rotation}, while 379 \(\tau_{\alpha_1} \cdots \tau_{\alpha_{2g}} \tau_\delta\) is taken to its 380 inverse. By Theorem~\ref{thm:boundaryless-birman-hilden}, 381 \[ 382 \tau_\delta \tau_{\alpha_{2g}} \cdots \tau_{\alpha_1} 383 \tau_{\alpha_1} \cdots \tau_{\alpha_{2g}} \tau_\delta 384 \in \ker (C_{\Mod(\Sigma_g)}([\iota]) \to \Mod(\Sigma_{0, 2g+2})) 385 = \langle [\iota] \rangle \cong \mathbb{Z}/2. 386 \] 387 One can then show \(\tau_\delta \tau_{\alpha_{2g}} \cdots \tau_{\alpha_1} 388 \tau_{\alpha_1} \cdots \tau_{\alpha_{2g}} \tau_\delta\) inverts the 389 orientation of \(\alpha_1\), so \(\tau_\delta \tau_{\alpha_{2g}} \cdots 390 \tau_{\alpha_1} \tau_{\alpha_1} \cdots \tau_{\alpha_{2g}} \tau_\delta \ne 1\) 391 and (\ref{eq:hyperelliptic-eq}) follows. In particular, we obtain the 392 so-called \emph{hyperelliptic relations} (\ref{eq:hyperelliptic-rel-1}) and 393 (\ref{eq:hyperelliptic-rel-2}) in \(\Mod(\Sigma_g)\). 394 \begin{align}\label{eq:hyperelliptic-rel-1} 395 (\tau_\delta \tau_{\alpha_{2g}} \cdots \tau_{\alpha_1} 396 \tau_{\alpha_1} \cdots \tau_{\alpha_{2g}} \tau_\delta)^2 397 & = 1 \\ 398 \label{eq:hyperelliptic-rel-2} 399 [\tau_\delta \tau_{\alpha_{2g}} \cdots \tau_{\alpha_1} 400 \tau_{\alpha_1} \cdots \tau_{\alpha_{2g}} \tau_\delta, \tau_\delta] 401 & = 1 402 \end{align} 403 \end{fundamental-observation} 404 405 \noindent 406 \begin{minipage}[b]{.47\textwidth} 407 \centering 408 \includegraphics[width=.7\linewidth]{images/hyperelliptic-relation.eps} 409 \vspace*{.5cm} 410 \captionof{figure}{The curves from the Humphreys generators of 411 $\Mod(\Sigma_g)$ and the curve $\delta$ from the hyperelliptic relations.} 412 \label{fig:hyperellipitic-relations} 413 \end{minipage} 414 \hspace{.6cm} % 415 \begin{minipage}[b]{.47\textwidth} 416 \centering 417 \includegraphics[width=.33\linewidth]{images/sphere-rotation.eps} 418 \captionof{figure}{The clockwise rotation by $\sfrac{\pi}{g + 1}$ about an 419 axis centered around the punctures of $\Sigma_{0, 2g + 1}$.} 420 \label{fig:hyperelliptic-relation-rotation} 421 \end{minipage} 422 \medskip 423 424 \section{Presentations of Mapping Class Groups} 425 426 Having explored some of the relations in \(\Mod(\Sigma)\), it is natural to ask 427 if these relations are enough to completely describe the structure of 428 \(\Mod(\Sigma)\). Different presentations of mapping class groups are due to 429 the work of Birman-Hilden \cite{birman-hilden}, Gervais \cite{gervais} and many 430 others. Wajnryb \cite{wajnryb} derived a presentation of \(\Mod(\Sigma_g)\) 431 only using the relations discussed in Chapter~\ref{ch:dehn-twists} and 432 Section~\ref{birman-hilden}. This is quite a satisfactory result, for we have 433 seen that all of these relations can be explained in terms of the topology of 434 \(\Sigma_g\). 435 436 \begin{theorem}[Wajnryb]\label{thm:wajnryb-presentation} 437 Suppose \(g \ge 3\). If \(\alpha_0, \ldots, \alpha_g\) are as in 438 Figure~\ref{fig:humphreys-gens} and \(a_i = \tau_{\alpha_i} \in 439 \Mod(\Sigma_g)\) are the Humphreys generators, then there is a presentation 440 of \(\Mod(\Sigma_g)\) with generators \(a_0, \ldots a_{2g}\) subject to the 441 following relations. 442 \begin{enumerate} 443 \item The \emph{disjointness relations} \([a_i, a_j] = 1\) for \(\alpha_i\) 444 and \(\alpha_j\) disjoint. 445 446 \item The \emph{braid relations} \(a_i a_j a_i = a_j a_i a_j\) for 447 \(\alpha_i\) and \(\alpha_j\) crossing once. 448 449 \item The \emph{\(3\)-chain relation} \((a_1 a_2 a_3)^4 = a_0 b_0\), where 450 \[ 451 b_0 = (a_4 a_3 a_2 a_1 a_1 a_2 a_3 a_4) 452 a_0 (a_4 a_3 a_2 a_1 a_1 a_2 a_3 a_4)^{-1}. 453 \] 454 455 \item The \emph{lantern relation} \(a_0 b_2 b_1 = a_1 a_3 a_5 b_3\), where 456 \begin{align*} 457 b_1 & = (a_4 a_5 a_3 a_4)^{-1} a_0 (a_4 a_5 a_3 a_4) \\ 458 b_2 & = (a_2 a_3 a_1 a_2)^{-1} b_1 (a_2 a_3 a_1 a_2) \\ 459 b_2 & = u b_1 u^{-1} \\ 460 u & = (a_6 a_5) (a_4 a_3 a_2) (a_6 a_5)^{-1} b_1 (a_6 a_5) a_1^{-1} 461 (a_4 a_3 a_2)^{-1}. 462 \end{align*} 463 464 \item The \emph{hyperelliptic relation} \([a_{2g} \cdots a_1 a_1 \cdots 465 a_{2g}, d] = 1\), where \(d = n_g\) for \(n_1 = a_1\), \(n_2 = b_0\) and 466 \begin{align*} 467 n_{i + 2} & = w_i n_i w_i^{-1} \\ 468 w_i & = (a_{2i + 4} a_{2i + 3} a_{2i + 2} n_{i + 1}) 469 (a_{2i + 1} a_{2i}^2 a_{2i + 1}) 470 (a_{2i + 3} a_{2i + 2} a_{2i + 4} a_{2i + 3}) 471 (n_1 a_{2i + 2} a_{2i + 1} a_{2i}). 472 \end{align*} 473 \end{enumerate} 474 \end{theorem} 475 476 \begin{remark} 477 The mapping classes \(b_0, \ldots, b_3, d\) in the statement of 478 Theorem~\ref{thm:wajnryb-presentation} correspond to the Dehn twists about 479 the curves \(\beta_0, \ldots, \beta_3, \delta \subset \Sigma_g\) highlighted 480 in Figure~\ref{fig:wajnryb-presentation-curves}, so Wajnryb's presentation is 481 not as intractable as it might look at first glance. 482 \end{remark} 483 484 \begin{figure}[ht] 485 \centering 486 \includegraphics[width=.7\linewidth]{images/wajnryb-presentation-curves.eps} 487 \caption{The curves from Wajnryb's presentation.} 488 \label{fig:wajnryb-presentation-curves} 489 \end{figure} 490 491 Different presentations can be used to compute the Abelianization of 492 \(\Mod(\Sigma_g)\) for \(g \le 2\). Indeed, if \(G = \langle g_1, \ldots, g_n : 493 R \rangle\) is a finitely-presented group, then \(G^\ab = \langle g_1, \ldots, 494 g_n : R, [g_i, g_j] \text{ for all } i, j \rangle\). Using this approach, 495 Farb-Margalit \cite[Section~5.1.3]{farb-margalit} show the Abelianization is 496 given by 497 \begin{center} 498 \begin{tabular}{r|c|l} 499 \(g\) & \(\Sigma_g\) & \(\Mod(\Sigma_g)^\ab\) \\[1pt] 500 \hline 501 & & \\[-10pt] 502 \(0\) & \(\mathbb{S}^2\) & \(0\) \\ 503 \(1\) & \(\mathbb{T}^2\) & \(\mathbb{Z}/12\) \\ 504 \(2\) & \(\Sigma_2\) & \(\mathbb{Z}/10\) \\ 505 \end{tabular} 506 \end{center} 507 for closed surfaces of small genus. In \cite{korkmaz-mccarthy} Korkmaz-McCarthy 508 showed that even though \(\Mod(\Sigma_2^b)\) is not perfect, its commutator 509 subgroup is. In addition, they also show \([\Mod(\Sigma_g^b), 510 \Mod(\Sigma_g^b)]\) is normally generated by a single mapping class. 511 512 \begin{proposition}\label{thm:commutator-is-perfect} 513 The commutator subgroup \(\Mod(\Sigma_2^b)' = [\Mod(\Sigma_2^b), 514 \Mod(\Sigma_2^b)]\) is perfect -- i.e. \(\Mod(\Sigma_2^b)^{(2)} = 515 [\Mod(\Sigma_2^b)', \Mod(\Sigma_2^b)']\) is the whole of 516 \(\Mod(\Sigma_2^b)'\). 517 \end{proposition} 518 519 \begin{proposition}\label{thm:commutator-normal-gen} 520 If \(g \ge 2\) and \(\alpha, \beta \subset \Sigma_g\) are simple closed 521 crossing only once, then \(\Mod(\Sigma_g)'\) is \emph{normally generated} by 522 \(\tau_\alpha \tau_\beta^{-1}\) -- i.e. if \(\tau_\alpha \tau_\beta^{-1} \in 523 N \normal \Mod(\Sigma_g)'\) then \(\Mod(\Sigma_g)' \subset N\). 524 \end{proposition} 525 526 The different presentations of \(\Mod(\Sigma_g)\) may also be used to study its 527 representations. Indeed, in light of Theorem~\ref{thm:wajnryb-presentation}, a 528 representation \(\rho : \Mod(\Sigma_g) \to \GL_n(\mathbb{C})\) is nothing other 529 than a choice of \(2g + 1\) matrices \(\rho(\tau_{\alpha_0}), \ldots, 530 \rho(\tau_{\alpha_{2g}}) \in \GL_n(\mathbb{C})\) satisfying the relations 531 \strong{(i)} to \strong{(v)} as above. In the next chapter, we will discuss how 532 these relations may be used to derive obstructions to the existence of 533 nontrivial representations of certain dimensions.