memoire-m2
My M2 Memoire on mapping class groups & their representations
Name | Size | Mode | |
.. | |||
sections/introduction.tex | 23839B | -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
\chapter{Introduction}\label{ch:introduction} Ever since ancestral humans first stepped foot on the surface of Earth, Mankind has pondered the shape of the planet we inhabit. More recently, mathematicians have spent the past centuries trying to understand the topology of manifolds and, in particular, surfaces. Orientable compact surfaces were perhaps first classified by Gauss in the early 19th century. The proof of the following formulation of the classification, often attributed to Möbius, was completed in the 1920s with the work of Radò and others. \begin{theorem}[Classification of surfaces]\label{thm:classification-of-surfaces} Any closed connected orientable surface is homeomorphic to the connected sum \(\Sigma_g\) of the sphere \(\mathbb{S}^2\) with \(g \ge 0\) copies of the torus \(\mathbb{T}^2 = \mfrac{\mathbb{R}^2}{\mathbb{Z}^2}\). Any compact connected orientable surface \(\Sigma\) is homeomorphic to the surface \(\Sigma_g^b\) obtained from \(\Sigma_g\) by removing \(b \ge 0\) open disks with disjoint closures. \end{theorem} The integer \(g \ge 0\) in Theorem~\ref{thm:classification-of-surfaces} is called \emph{the genus of \(\Sigma\)}. We also have the noncompact surface \(\Sigma_{g, r}^b = \Sigma_g^b \setminus \{x_1, \ldots, x_r\}\), where \(x_1, \ldots, x_r\) lie in the interior of \(\Sigma_g^b\). The points \(x_1, \ldots, x_r\) are called the \emph{punctures} of \(\Sigma_{g, r}^b\). Throughout these notes, all surfaces considered will be of the form \(\Sigma = \Sigma_{g, r}^b\). Any such \(\Sigma\) admits a natural compactification \(\widebar\Sigma\) obtained by filling its punctures. We denote \(\Sigma_{g, r} = \Sigma_{g, r}^0\). All closed curves \(\alpha, \beta \subset \Sigma\) we consider lie in the interior of \(\Sigma\) and intersect transversely. Unless explicitly stated otherwise, the curves \(\alpha, \beta\) are assumed to be \emph{unoriented} -- i.e. we regard them as subsets of \(\Sigma\). Despite the apparent clarity of the picture painted by Theorem~\ref{thm:classification-of-surfaces}, there are still plenty of interesting, sometimes unanswered, questions about surfaces and their homeomorphisms. For instance, we can use the classification of surfaces to deduce information about how different curves in \(\Sigma\) are related by its homeomorphisms. \begin{observation}[Change of coordinates principle] Given oriented nonseparating simple closed curves \(\alpha, \beta : \mathbb{S}^1 \to \Sigma = \Sigma_{g, r}^b\), we can find an orientation-preserving homeomorphism \(\phi : \Sigma \isoto \Sigma\) fixing \(\partial \Sigma\) pointwise such that \(\phi(\alpha) = \beta\) with orientation. To see this, we consider the surface \(\Sigma_\alpha\) obtained by cutting \(\Sigma\) along \(\alpha\): we subtract the curve \(\alpha\) from \(\Sigma\) and then add one additional boundary component \(\delta_i\) in each side of \(\alpha\), as shown in Figure~\ref{fig:change-of-coordinates}. By identifying \(\delta_1\) with \(\delta_2\) we can see \(\Sigma\) as a quotient of \(\Sigma_\alpha\). Since \(\alpha\) is nonseparating, \(\Sigma_\alpha\) is a connected surface of genus \(g - 1\). In other words, \(\Sigma_\alpha \cong \Sigma_{g-1,r}^{b+2}\). Similarly, \(\Sigma_\beta \cong \Sigma_{g-1, r}^{b+2}\) also has two additional boundary components \(\delta_1', \delta_2' \subset \partial \Sigma_\beta\). Now by the classification of surfaces we can find an orientation-preserving homeomorphism \(\tilde\phi : \Sigma_\alpha \isoto \Sigma_\beta\). Even more so, we can choose \(\tilde\phi\) taking \(\delta_i\) to \(\delta_i'\). The homeomorphism \(\tilde\phi\) then descends to a self-homeomorphism \(\phi\) of the quotient surface \(\Sigma \cong \mfrac{\Sigma_\alpha}{\sim} \cong \mfrac{\Sigma_\beta}{\sim}\) with \(\phi(\alpha) = \beta\), as desired. \end{observation} \begin{figure}[ht] \centering \includegraphics[width=.5\linewidth]{images/change-of-coords-cut.eps} \caption{The surface $\Sigma_\alpha \cong \Sigma_{g-1, r}^{b+2}$ for a certain $\alpha \subset \Sigma$.} \label{fig:change-of-coordinates} \end{figure} By cutting \(\Sigma\) along curves \(\alpha, \alpha' \subset \Sigma\) crossing once, we can also show the following result. \begin{observation}\label{ex:change-of-coordinates-crossing} Let \(\alpha, \beta, \alpha', \beta' \subset \Sigma\) be nonseparating curves such that each pair \((\alpha, \alpha'), (\beta, \beta')\) crosses exactly once. Then we can find an orientation-preserving \(\phi : \Sigma \isoto \Sigma\) fixing \(\partial \Sigma\) pointwise such that \(\phi(\alpha) = \beta\) and \(\phi(\alpha') = \beta'\) -- without orientation. \end{observation} Given a surface \(\Sigma\), the group \(\Homeo^+(\Sigma, \partial \Sigma)\) of orientation-preserving homeomorphisms of \(\Sigma\) fixing each point in \(\partial \Sigma\) is a topological group\footnote{Here we endow \(\Homeo^+(\Sigma, \partial \Sigma)\) with the compact-open topology.} with a rich geometry, but its algebraic structure is often regarded as too complex to tackle. More importantly, all of this complexity is arguably unnecessary for most topological applications, in the sense that usually we are only really interested in considering \emph{homeomorphisms up to isotopy}. For example, \begin{enumerate} \item Isotopic \(\phi \simeq \psi \in \Homeo^+(\Sigma, \partial \Sigma)\) determine the same application \(\phi_* = \psi_* : \pi_1(\Sigma, x) \to \pi_1(\Sigma, x)\) and \(\phi_* = \psi_* : H_1(\Sigma, \mathbb{Z}) \to H_1(\Sigma, \mathbb{Z})\) at the levels of homotopy and homology. \item The diffeomorphism class of the mapping torus \(M_\phi = \mfrac{\Sigma \times [0, 1]}{(x, 0) \sim (\phi(x), 1)}\) -- a fundamental construction in low-dimensional topology -- is invariant under isotopy. \end{enumerate} It is thus more natural to consider the group of connected components of \(\Homeo^+(\Sigma, \partial \Sigma)\), a countable discrete group known as \emph{the mapping class group}. This will be the focus of the dissertation at hand. \begin{definition}\label{def:mcg} The \emph{mapping class group \(\Mod(\Sigma)\) of an orientable surface \(\Sigma\)} is the group of isotopy classes of orientation-preserving homeomorphisms \(\Sigma \isoto \Sigma\), where both the homeomorphisms and the isotopies are assumed to fix \(\partial \Sigma\) pointwise. \[ \Mod(\Sigma) = \mfrac{\Homeo^+(\Sigma, \partial \Sigma)}{\simeq} \] \end{definition} There are many variations of Definition~\ref{def:mcg}. \begin{observation}\label{ex:action-on-punctures} Any \(\phi \in \Homeo^+(\Sigma, \partial \Sigma)\) extends uniquely to a homeomorphism \(\tilde\phi\) of \(\widebar\Sigma\) that permutes the set \(\{x_1, \ldots, x_r\} = \widebar\Sigma \setminus \Sigma\) of punctures of \(\Sigma\). We may thus define an action \(\Mod(\Sigma) \leftaction \{x_1, \ldots, x_r\}\) via \(f \cdot x_i = \tilde\phi(x_i)\) for \(f = [\phi] \in \Mod(\Sigma)\) -- which is independent of the choice of representative \(\phi\) of \(f\). \end{observation} \begin{definition} Given an orientable surface \(\Sigma\) and a puncture \(x \in \widebar\Sigma\) of \(\Sigma\), denote by \(\Mod(\Sigma, x) \subset \Mod(\Sigma)\) the subgroup of mapping classes that fix \(x\). The \emph{pure mapping class group \(\PMod(\Sigma) \subset \Mod(\Sigma)\) of \(\Sigma\)} is the subgroup of mapping classes that fix every puncture of \(\Sigma\). \end{definition} \begin{observation}\label{ex:action-on-curves} Given an oriented simple closed curve \(\alpha : \mathbb{S}^1 \to \Sigma\), denote by \(\vec{[\alpha]}\) and \([\alpha]\) the isotopy classes of \(\alpha\) with and without orientation, respectively -- i.e \(\vec{[\alpha]} = \vec{[\beta]}\) if \(\alpha \simeq \beta\) as functions and \([\alpha] = [\beta]\) if \(\vec{[\alpha]} = \vec{[\beta]}\) or \(\vec{[\alpha]} = \vec{[\beta^{-1}]}\). There are natural actions \(\Mod(\Sigma) \leftaction \{ \vec{[\alpha]} \, | \, \alpha : \mathbb{S}^1 \to \Sigma \}\) and \(\Mod(\Sigma) \leftaction \{ [\alpha] \, | \, \alpha \subset \Sigma \}\) given by \begin{align*} f \cdot \vec{[\alpha]} & = \vec{[\phi(\alpha)]} & f \cdot [\alpha] & = [\phi(\alpha)] \end{align*} for \(f = [\phi] \in \Mod(\Sigma)\). \end{observation} \begin{definition} Given a simple closed curve \(\alpha \subset \Sigma\), we denote by \(\Mod(\Sigma)_{\vec{[\alpha]}}\) and \(\Mod(\Sigma)_{[\alpha]}\) the subgroups of mapping classes that fix \(\vec{[\alpha]}\) -- for any given choice of orientation of \(\alpha\) -- and \([\alpha]\), respectively. \end{definition} While trying to understand the mapping class group of some surface \(\Sigma\), it is interesting to consider how the geometric relationship between \(\Sigma\) and other surfaces affects \(\Mod(\Sigma)\). Indeed, different embeddings \(\Sigma' \hookrightarrow \Sigma\) translate to homomorphisms at the level of mapping class groups. \begin{example}[Inclusion homomorphism]\label{ex:inclusion-morphism} Let \(\Sigma' \subset \Sigma\) be a closed subsurface. Given \(\phi \in \Homeo^+(\Sigma', \partial \Sigma')\), we may extend \(\phi\) to \(\tilde{\phi} \in \Homeo^+(\Sigma, \partial \Sigma)\) by setting \(\tilde{\phi}(x) = x\) for \(x \in \Sigma\) outside of \(\Sigma'\) -- which is well defined since \(\phi\) fixes every point in \(\partial \Sigma'\). This construction yields a group homomorphism \begin{align*} \Mod(\Sigma') & \to \Mod(\Sigma) \\ [\phi] & \mapsto [\tilde\phi], \end{align*} known as \emph{the inclusion homomorphism}. \end{example} \begin{example}[Capping homomorphism]\label{ex:capping-morphism} Let \(\delta \subset \partial \Sigma\) be a boundary component of \(\Sigma\). We refer to the inclusion homomorphism \(\operatorname{cap} : \Mod(\Sigma) \to \Mod(\Sigma \cup_\delta (\mathbb{D}^2 \setminus \{0\}))\) as \emph{the capping homomorphism}. \end{example} \begin{example}[Cutting homomorphism]\label{ex:cutting-morphism} Given a simple closed curve \(\alpha \subset \Sigma\), any \(f \in \Mod(\Sigma_{g+1})_{\vec{[\alpha]}}\) has a representative \(\phi \in \Homeo^+(\Sigma, \partial \Sigma)\) fixing \(\alpha\) point-wise -- so that \(\phi\) restricts to a homeomorphism of \(\Sigma \setminus \alpha\). Furthermore, if \(\phi\!\restriction_{\Sigma \setminus \alpha} \simeq 1\) in \(\Sigma \setminus \alpha\) then \(\phi \simeq 1 \in \Homeo^+(\Sigma, \partial \Sigma)\) -- see \cite[Proposition~3.20]{farb-margalit}. There is thus a group homomorphism \begin{align*} \operatorname{cut} : \Mod(\Sigma)_{\vec{[\alpha]}} & \to \Mod(\Sigma\setminus\alpha) \\ [\phi] & \mapsto [\phi\!\restriction_{\Sigma \setminus \alpha}], \end{align*} known as \emph{the cutting homomorphism}. \end{example} As goes for most groups, another approach to understanding the mapping class group of a given surface \(\Sigma\) is to study its actions. We have already seen simple examples of such actions in Observation~\ref{ex:action-on-punctures} and Observation~\ref{ex:action-on-curves}. An important class of actions of \(\Mod(\Sigma)\) are its \emph{linear representations} -- i.e. the group homomorphisms \(\Mod(\Sigma) \to \GL_n(\mathbb{C})\). These may be seen as actions \(\Mod(\Sigma) \leftaction \mathbb{C}^n\) where each \(f \in \Mod(\Sigma)\) acts via some linear isomorphism \(\mathbb{C}^n \isoto \mathbb{C}^n\). \section{Representations} Here we collect a few fundamental examples of linear representations of \(\Mod(\Sigma)\). \begin{observation} Recall \(H_1(\Sigma_g, \mathbb{Z}) \cong \mathbb{Z}^{2g}\), with standard basis given by \([\alpha_1], [\beta_1], \ldots, [\alpha_g], [\beta_g] \in H_1(\Sigma_g, \mathbb{Z})\) for \(\alpha_1, \ldots, \alpha_g, \beta_1, \ldots, \beta_g\) as in Figure~\ref{fig:homology-basis}. The Abelian group \(H_1(\Sigma_g, \mathbb{Z})\) is endowed with a natural \(\mathbb{Z}\)-bilinear alternating form given by the \emph{algebraic intersection number} \([\alpha] \cdot [\beta] = \sum_{x \in \alpha \cap \beta} \operatorname{ind}\,x\) -- where the index \(\operatorname{ind}\,x = \pm 1\) of an intersection point is given by Figure~\ref{fig:intersection-index}. In terms of the standard basis of \(H_1(\Sigma_g, \mathbb{Z})\), this form is given by \begin{align}\label{eq:symplectic-form} [\alpha_i] \cdot [\beta_j] & = \delta_{i j} & [\alpha_i] \cdot [\alpha_j] & = 0 & [\beta_i] \cdot [\beta_j] & = 0 \end{align} and thus coincides with the pullback of the standard \(\mathbb{Z}\)-bilinear symplectic form in \(\mathbb{Z}^{2g}\). \end{observation} \begin{example}[Symplectic representation]\label{ex:symplectic-rep} Given \(f = [\phi] \in \Mod(\Sigma_g)\), we may consider the map \(\phi_* : H_1(\Sigma_g, \mathbb{Z}) \to H_1(\Sigma_g, \mathbb{Z})\) induced at the level of singular homology. By homotopy invariance, the map \(\phi_*\) is independent of the choice of representative \(\phi\) of \(f\). By the functoriality of homology groups we then get a \(\mathbb{Z}\)-linear action \(\Mod(\Sigma_g) \leftaction H_1(\Sigma_g, \mathbb{Z}) \cong \mathbb{Z}^{2g}\) given by \(f \cdot [\alpha] = \phi_*([\alpha]) = [\phi(\alpha)]\). Since pushforwards by orientation-preserving homeomorphisms preserve the indices of intersection points, \((f \cdot [\alpha]) \cdot (f \cdot [\beta]) = [\alpha] \cdot [\beta]\) for all \(\alpha, \beta \subset \Sigma_g\) and \(f \in \Mod(\Sigma_g)\). In light of (\ref{eq:symplectic-form}), this implies \(\Mod(\Sigma_g)\) acts on \(\mathbb{Z}^{2g}\) via symplectomorphisms. We thus obtain a group homomorphism \(\psi : \Mod(\Sigma_g) \to \operatorname{Sp}_{2g}(\mathbb{Z}) \subset \GL_{2g}(\mathbb{C})\), known as \emph{the symplectic representation of \(\Mod(\Sigma_g)\)}. \end{example} \noindent \begin{minipage}[b]{.47\linewidth} \centering \includegraphics[width=.9\linewidth]{images/homology-generators.eps} \captionof{figure}{The curves $\alpha_1, \beta_1, \ldots, \alpha_g, \beta_g \subset \Sigma_g$ that generate its first homology group.} \label{fig:homology-basis} \end{minipage} \hspace{.6cm} % \begin{minipage}[b]{.47\linewidth} \centering \includegraphics[width=.9\linewidth]{images/intersection-index.eps} \vspace*{.75cm} \captionof{figure}{The index of an intersection point $x \in \alpha \cap \beta$.} \label{fig:intersection-index} \end{minipage} The symplectic representation already allows us to compute some important examples of mapping class groups: namely, that of the torus \(\mathbb{T}^2 = \Sigma_1\) and the once-punctured torus \(\Sigma_{1, 1}\). \begin{observation}[Alexander trick]\label{ex:alexander-trick} The group \(\Homeo^+(\mathbb{D}^2, \mathbb{S}^1)\) of homeomorphisms of the unit disk \(\mathbb{D}^2 \subset \mathbb{C}\) is contractible. In particular, \(\Mod(\mathbb{D}^2) = 1\). Indeed, for any \(\phi \in \Homeo^+(\mathbb{D}^2, \mathbb{S}^1)\) the isotopy \begin{align*} \phi_t : \mathbb{D}^2 & \to \mathbb{D}^2 \\ z & \mapsto \begin{cases} (1 - t) \phi(\sfrac{z}{1 - t}) & \text{if } 0 \le |z| \le 1 - t \\ z & \text{otherwise} \end{cases} \end{align*} that ``fixes the band \(\{ z \in \mathbb{D}^2 : |z| \ge 1 - t \}\) and does \(\phi\) inside the sub-disk \(\{ z \in \mathbb{D}^2 : |z| \le 1 - t\}\)'' joins \(\phi = \phi_0\) and \(1 = \phi_1\). \end{observation} \begin{observation}\label{ex:mdg-once-punctured-disk} By the same token, \(\Mod(\mathbb{D}^2 \setminus \{0\}) = 1\). \end{observation} \begin{observation}[Linearity of $\Mod(\mathbb{T}^2)$]\label{ex:torus-mcg} The symplectic representation \(\psi : \Mod(\mathbb{T}^2) \to \operatorname{Sp}_2(\mathbb{Z}) = \operatorname{SL}_2(\mathbb{Z})\) is a group isomorphism. In particular, \(\Mod(\mathbb{T}^2) \cong \operatorname{SL}_2(\mathbb{Z})\). To see \(\psi\) is surjective, first observe \(\mathbb{Z}^2 \subset \mathbb{R}^2\) is \(\operatorname{SL}_2(\mathbb{Z})\)-invariant. Hence any matrix \(g \in \operatorname{SL}_2(\mathbb{Z})\) descends to an orientation-preserving homeomorphism \(\phi_g\) of the quotient \(\mathbb{T}^2 = \mfrac{\mathbb{R}^2}{\mathbb{Z}^2}\), which satisfies \(\psi([\phi_g]) = g\). To see \(\psi\) is injective we consider the curves \(\alpha_1\) and \(\beta_1\) from Figure~\ref{fig:homology-basis}. Given \(f = [\phi] \in \Mod(\mathbb{T}^2)\) with \(\psi(f) = 1\), \(f \cdot \vec{[\alpha_1]} = \vec{[\alpha_1]}\) and \(f \cdot \vec{[\beta_1]} = \vec{[\beta_1]}\), so we may choose a representative \(\phi\) of \(f\) fixing \(\alpha_1 \cup \beta_1\) pointwise. Such \(\phi\) determines a homeomorphism \(\tilde \phi\) of the surface \(\mathbb{T}_{\alpha_1 \beta_1}^2 \cong \mathbb{D}^2\) obtained by cutting \(\mathbb{T}^2\) along \(\alpha_1\) and \(\beta_1\), as in Figure~\ref{fig:cut-torus-along}. Now by the Alexander trick from Observation~\ref{ex:alexander-trick}, \(\tilde\phi\) must be isotopic to the identity. The isotopy \(\tilde\phi \simeq 1 \in \Homeo^+(\mathbb{D}^2, \mathbb{S}^1)\) then descends to an isotopy \(\phi \simeq 1 \in \Homeo^+(\mathbb{T}^2)\), so \(f = 1 \in \Mod(\mathbb{T}^2)\) as desired. \end{observation} \begin{figure}[ht] \centering \includegraphics[width=.55\linewidth]{images/torus-cut.eps} \caption{By cutting $\mathbb{T}^2$ along $\alpha_1$ we obtain a cylinder, where $\beta_1$ determines a yellow arc joining the two boundary components. Now by cutting along this yellow arc we obtain a disk.} \label{fig:cut-torus-along} \end{figure} \begin{observation}\label{ex:punctured-torus-mcg} By the same token, \(\Mod(\Sigma_{1, 1}) \cong \operatorname{SL}_2(\mathbb{Z})\). \end{observation} \begin{remark} Despite the fact \(\psi : \Mod(\mathbb{T}^2) \to \operatorname{SL}_2(\mathbb{Z})\) is an isomorphism, the symplectic representation is \emph{not} injective for surfaces of genus \(g \ge 2\) -- see \cite[Section~6.5]{farb-margalit} for a description of its kernel. Korkmaz and Bigelow-Budney \cite{korkmaz-linearity, bigelow-budney} showed there exist injective linear representations of \(\Mod(\Sigma_2)\), but the question of linearity of \(\Mod(\Sigma_g)\) remains wide-open for \(g \ge 3\). Recently, Korkmaz \cite[Theorem~3]{korkmaz} established the lower bound of \(3 g - 3\) for the dimension of an injective representation of \(\Mod(\Sigma_g)\) in the \(g \ge 3\) case -- if one such representation exists. \end{remark} Another fundamental class of examples of representations are the so-called \emph{TQFT representations}. \begin{definition} A \emph{cobordism} between closed oriented surfaces \(\Sigma\) and \(\Sigma'\) is a triple \((W, \phi_+, \phi_-)\) where \(W\) is a smooth oriented compact \(3\)-manifold with \(\partial W = \partial_+ W \amalg \partial_- W\), \(\phi_+ : \Sigma \isoto \partial_+ W\) is an orientation preserving diffeomorphism and \(\phi_- : \Sigma' \isoto \partial_- W\) is an orientation-reversing diffeomorphism. We may abuse the notation and denote \(W = (W, \phi_+, \phi_-)\). \end{definition} \begin{definition} We denote by \(\Cob\) the category whose objects are (possibly disconnected) closed oriented surfaces and whose morphisms \(\Sigma \to \Sigma'\) are diffeomorphism classes\footnote{Here we only consider orientation-preserving diffeomorphisms $\varphi : W \isoto W'$ that are compatible with the boundary identifications in the sense that $\varphi(\partial_\pm W) = \partial_\pm W'$ and $\psi_\pm = \varphi \circ \phi_\pm$.} of cobordisms between \(\Sigma\) and \(\Sigma'\), with composition given by \[ [W, \phi_-, \phi_+] \circ [W', \psi_-, \psi_+] = [W \cup_{\psi_- \circ \phi_+^{-1}} W', \phi_-, \psi_+] \] for \([W, \phi_-, \phi_+] : \Sigma \to \Sigma'\) and \([W', \psi_-, \phi_+] : \Sigma' \to \Sigma''\). We endow \(\Cob\) with the monoidal structure given by \begin{align*} \Sigma \otimes \Sigma' & = \Sigma \amalg \Sigma' & [W,\phi_+,\phi_-] \otimes [W',\psi_+,\psi_-] & = [W \amalg W', \phi_+ \amalg \psi_+, \phi_- \amalg \psi_-]. \end{align*} \end{definition} \begin{definition}[TQFT]\label{def:tqft} A \emph{topological quantum field theory} (abbreviated by \emph{TQFT}) is a functor \(\mathcal{F} : \Cob \to \Vect\) satisfying \begin{align*} \mathcal{F}(\emptyset) & = \mathbb{C} & \mathcal{F}(\Sigma \otimes \Sigma') & = \mathcal{F}(\Sigma) \otimes \mathcal{F}(\Sigma') & \mathcal{F}([W] \otimes [W']) & = \mathcal{F}([W]) \otimes \mathcal{F}([W']), \end{align*} where \(\Vect\) denotes the category of finite-dimensional complex vector spaces. \end{definition} \begin{observation} Given \(\phi \in \Homeo^+(\Sigma_g)\), we may consider the so-called \emph{mapping cylinder} \(C_\phi = (\Sigma_g \times [0, 1], \phi, 1)\), a cobordism between \(\Sigma_g\) and itself -- where \(\partial_+ (\Sigma_g \times [0, 1]) = \Sigma_g \times 0\) and \(\partial_- (\Sigma_g \times [0, 1]) = \Sigma_g \times 1\). The diffeomorphism class of \(C_\phi\) is independent of the choice of representative of \(f = [\phi] \in \Mod(\Sigma_g)\), so \(C_f = [C_\phi] : \Sigma_g \to \Sigma_g\) is a well defined morphism in \(\Cob\). \end{observation} \begin{example}[TQFT representations]\label{ex:tqft-reps} It is clear that \(C_1\) is the identity morphism \(\Sigma_g \to \Sigma_g\) in \(\Cob\). In addition, \(C_{f \cdot g} = C_f \circ C_g\) for all \(f, g \in \Mod(\Sigma_g)\) -- see \cite[Lemma~2.5]{costantino}. Now given a TQFT \(\mathcal{F} : \Cob \to \Vect\), by functoriality we obtain a linear representation \begin{align*} \rho_{\mathcal{F}} : \Mod(\Sigma_g) & \to \GL(\mathcal{F}(\Sigma_g)) \\ f & \mapsto \mathcal{F}(C_f). \end{align*} \end{example} As simple as the construction in Example~\ref{ex:tqft-reps} is, in practice it is not that easy to come across functors as the ones in Definition~\ref{def:tqft}. This is because, in most interesting examples, we are required to attach some extra data to our surfaces to get a well defined association \(\Sigma_g \mapsto \mathcal{F}(\Sigma_g)\). Moreover, the condition \(\mathcal{F}([W] \circ [W']) = \mathcal{F}([W]) \circ \mathcal{F}([W'])\) may only hold up to multiplication by scalars. Hence constructing an actual functor typically requires \emph{extending} \(\Cob\) and \emph{tweaking} \(\Vect\). Such functors give rise to linear and projective representations of the \emph{extended mapping class groups} \(\Mod(\Sigma_g) \times \mathbb{Z}\). We refer the reader to \cite{costantino, julien} for constructions of one such TQFT and its corresponding representations: the so-called \emph{\(\operatorname{SU}_2\) TQFT of level \(r\)}, first introduced by Witten and Reshetikhin-Tuarev \cite{witten, reshetikhin-turaev} in their foundational papers on quantum topology. Besides Example~\ref{ex:symplectic-rep} and Example~\ref{ex:tqft-reps}, not a lot of other linear representations of \(\Mod(\Sigma_g)\) are known. Indeed, the representation theory of mapping class groups remains a mystery at large. In Chapter~\ref{ch:representations} we provide a brief overview of the field, as well as some recent developments. More specifically, we highlight Korkmaz' \cite{korkmaz} proof of the triviality of low-dimensional representations and comment on his classification of \(2g\)-dimensional representations. To that end, in Chapter~\ref{ch:dehn-twists} and Chapter~\ref{ch:relations} we survey the group structure of mapping class groups: its relations and known presentations.