memoire-m2

My M2 Memoire on mapping class groups & their representations

introduction.tex (23839B)

  1 \chapter{Introduction}\label{ch:introduction}
  2 
  3 Ever since ancestral humans first stepped foot on the surface of Earth, Mankind
  4 has pondered the shape of the planet we inhabit. More recently, mathematicians
  5 have spent the past centuries trying to understand the topology of manifolds
  6 and, in particular, surfaces. Orientable compact surfaces were perhaps first
  7 classified by Gauss in the early 19th century. The proof of the following
  8 formulation of the classification, often attributed to Möbius, was completed in
  9 the 1920s with the work of Radò and others.
 10 
 11 \begin{theorem}[Classification of surfaces]\label{thm:classification-of-surfaces}
 12   Any closed connected orientable surface is homeomorphic to the connected sum
 13   \(\Sigma_g\) of the sphere \(\mathbb{S}^2\) with \(g \ge 0\) copies of the
 14   torus \(\mathbb{T}^2 = \mfrac{\mathbb{R}^2}{\mathbb{Z}^2}\). Any compact
 15   connected orientable surface \(\Sigma\) is homeomorphic to the surface
 16   \(\Sigma_g^b\) obtained from \(\Sigma_g\) by removing \(b \ge 0\) open disks
 17   with disjoint closures.
 18 \end{theorem}
 19 
 20 The integer  \(g \ge 0\) in Theorem~\ref{thm:classification-of-surfaces} is
 21 called \emph{the genus of \(\Sigma\)}. We also have the noncompact surface
 22 \(\Sigma_{g, r}^b = \Sigma_g^b \setminus \{x_1, \ldots, x_r\}\), where \(x_1,
 23 \ldots, x_r\) lie in the interior of \(\Sigma_g^b\). The points \(x_1, \ldots,
 24 x_r\) are called the \emph{punctures} of \(\Sigma_{g, r}^b\). Throughout these
 25 notes, all surfaces considered will be of the form \(\Sigma = \Sigma_{g,
 26 r}^b\). Any such \(\Sigma\) admits a natural compactification
 27 \(\widebar\Sigma\) obtained by filling its punctures. We denote \(\Sigma_{g, r}
 28 = \Sigma_{g, r}^0\). All closed curves \(\alpha, \beta \subset \Sigma\) we
 29 consider lie in the interior of \(\Sigma\) and intersect transversely. Unless
 30 explicitly stated otherwise, the curves \(\alpha, \beta\) are assumed to
 31 be \emph{unoriented} -- i.e. we regard them as subsets of \(\Sigma\).
 32 
 33 Despite the apparent clarity of the picture painted by
 34 Theorem~\ref{thm:classification-of-surfaces}, there are still plenty of
 35 interesting, sometimes unanswered, questions about surfaces and their
 36 homeomorphisms. For instance, we can use the classification of surfaces to
 37 deduce information about how different curves in \(\Sigma\) are related by its
 38 homeomorphisms.
 39 
 40 \begin{observation}[Change of coordinates principle]
 41   Given oriented nonseparating simple closed curves \(\alpha, \beta :
 42   \mathbb{S}^1 \to \Sigma = \Sigma_{g, r}^b\), we can find an
 43   orientation-preserving homeomorphism \(\phi : \Sigma \isoto \Sigma\) fixing
 44   \(\partial \Sigma\) pointwise such that \(\phi(\alpha) = \beta\) with
 45   orientation. To see this, we consider the surface \(\Sigma_\alpha\) obtained
 46   by cutting \(\Sigma\) along \(\alpha\): we subtract the curve \(\alpha\)
 47   from \(\Sigma\) and then add one additional boundary component \(\delta_i\)
 48   in each side of \(\alpha\), as shown in
 49   Figure~\ref{fig:change-of-coordinates}. By identifying \(\delta_1\) with
 50   \(\delta_2\) we can see \(\Sigma\) as a quotient of \(\Sigma_\alpha\). 
 51 
 52   Since \(\alpha\) is nonseparating, \(\Sigma_\alpha\) is a connected surface
 53   of genus \(g - 1\). In other words, \(\Sigma_\alpha \cong
 54   \Sigma_{g-1,r}^{b+2}\). Similarly, \(\Sigma_\beta \cong \Sigma_{g-1,
 55   r}^{b+2}\) also has two additional boundary components \(\delta_1', \delta_2'
 56   \subset \partial \Sigma_\beta\). Now by the classification of surfaces we can
 57   find an orientation-preserving homeomorphism \(\tilde\phi : \Sigma_\alpha
 58   \isoto \Sigma_\beta\). Even more so, we can choose \(\tilde\phi\) taking
 59   \(\delta_i\) to \(\delta_i'\). The homeomorphism \(\tilde\phi\) then descends
 60   to a self-homeomorphism \(\phi\) of the quotient surface \(\Sigma \cong
 61   \mfrac{\Sigma_\alpha}{\sim} \cong \mfrac{\Sigma_\beta}{\sim}\) with
 62   \(\phi(\alpha) = \beta\), as desired.
 63 \end{observation}
 64 
 65 \begin{figure}[ht]
 66   \centering
 67   \includegraphics[width=.5\linewidth]{images/change-of-coords-cut.eps}
 68   \caption{The surface $\Sigma_\alpha \cong \Sigma_{g-1, r}^{b+2}$ for a
 69   certain $\alpha \subset \Sigma$.}
 70   \label{fig:change-of-coordinates}
 71 \end{figure}
 72 
 73 By cutting \(\Sigma\) along curves \(\alpha, \alpha' \subset \Sigma\) crossing
 74 once, we can also show the following result.
 75 
 76 \begin{observation}\label{ex:change-of-coordinates-crossing}
 77   Let \(\alpha, \beta, \alpha', \beta' \subset \Sigma\) be nonseparating curves
 78   such that each pair \((\alpha, \alpha'), (\beta, \beta')\) crosses exactly
 79   once. Then we can find an orientation-preserving \(\phi : \Sigma \isoto
 80   \Sigma\) fixing \(\partial \Sigma\) pointwise such that \(\phi(\alpha) =
 81   \beta\) and \(\phi(\alpha') = \beta'\) -- without orientation.
 82 \end{observation}
 83 
 84 Given a surface \(\Sigma\), the group \(\Homeo^+(\Sigma, \partial \Sigma)\) of
 85 orientation-preserving homeomorphisms of \(\Sigma\) fixing each point in
 86 \(\partial \Sigma\) is a topological group\footnote{Here we endow
 87 \(\Homeo^+(\Sigma, \partial \Sigma)\) with the compact-open topology.} with a
 88 rich geometry, but its algebraic structure is often regarded as too complex to
 89 tackle. More importantly, all of this complexity is arguably unnecessary for
 90 most topological applications, in the sense that usually we are only really
 91 interested in considering \emph{homeomorphisms up to isotopy}. For example,
 92 \begin{enumerate}
 93   \item Isotopic \(\phi \simeq \psi \in \Homeo^+(\Sigma, \partial \Sigma)\)
 94     determine the same application \(\phi_* = \psi_* : \pi_1(\Sigma, x) \to
 95     \pi_1(\Sigma, x)\) and \(\phi_* = \psi_* : H_1(\Sigma, \mathbb{Z}) \to
 96     H_1(\Sigma, \mathbb{Z})\)
 97     at the levels of homotopy and homology.
 98 
 99   \item The diffeomorphism class of the mapping torus \(M_\phi = \mfrac{\Sigma
100     \times [0, 1]}{(x, 0) \sim (\phi(x), 1)}\) -- a fundamental construction in
101     low-dimensional topology -- is invariant under isotopy.
102 \end{enumerate}
103 
104 It is thus more natural to consider the group of connected components of
105 \(\Homeo^+(\Sigma, \partial \Sigma)\), a countable discrete group known as
106 \emph{the mapping class group}. This will be the focus of the dissertation at
107 hand.
108 
109 \begin{definition}\label{def:mcg}
110   The \emph{mapping class group \(\Mod(\Sigma)\) of an orientable surface
111   \(\Sigma\)} is the group of isotopy classes of orientation-preserving
112   homeomorphisms \(\Sigma \isoto \Sigma\), where both the homeomorphisms and
113   the isotopies are assumed to fix \(\partial \Sigma\) pointwise.
114   \[
115     \Mod(\Sigma) = \mfrac{\Homeo^+(\Sigma, \partial \Sigma)}{\simeq}
116   \]
117 \end{definition}
118 
119 There are many variations of Definition~\ref{def:mcg}.
120 
121 \begin{observation}\label{ex:action-on-punctures}
122   Any \(\phi \in \Homeo^+(\Sigma, \partial \Sigma)\) extends uniquely to a
123   homeomorphism \(\tilde\phi\) of \(\widebar\Sigma\) that permutes the set
124   \(\{x_1, \ldots, x_r\} = \widebar\Sigma \setminus \Sigma\) of punctures of
125   \(\Sigma\). We may thus define an action \(\Mod(\Sigma) \leftaction \{x_1,
126   \ldots, x_r\}\) via \(f \cdot x_i = \tilde\phi(x_i)\) for \(f = [\phi] \in
127   \Mod(\Sigma)\) -- which is independent of the choice of representative
128   \(\phi\) of \(f\).
129 \end{observation}
130 
131 \begin{definition}
132   Given an orientable surface \(\Sigma\) and a puncture \(x \in
133   \widebar\Sigma\) of \(\Sigma\), denote by \(\Mod(\Sigma, x) \subset
134   \Mod(\Sigma)\) the subgroup of mapping classes that fix \(x\). The \emph{pure
135   mapping class group \(\PMod(\Sigma) \subset \Mod(\Sigma)\) of \(\Sigma\)} is
136   the subgroup of mapping classes that fix every puncture of \(\Sigma\).
137 \end{definition}
138 
139 \begin{observation}\label{ex:action-on-curves}
140   Given an oriented simple closed curve \(\alpha : \mathbb{S}^1 \to \Sigma\),
141   denote by \(\vec{[\alpha]}\) and \([\alpha]\) the isotopy classes of
142   \(\alpha\) with and without orientation, respectively -- i.e \(\vec{[\alpha]}
143   = \vec{[\beta]}\) if \(\alpha \simeq \beta\) as functions and \([\alpha] =
144   [\beta]\) if \(\vec{[\alpha]} = \vec{[\beta]}\) or \(\vec{[\alpha]} =
145   \vec{[\beta^{-1}]}\). There are natural actions \(\Mod(\Sigma) \leftaction \{
146     \vec{[\alpha]} \, | \, \alpha : \mathbb{S}^1 \to \Sigma \}\) and
147   \(\Mod(\Sigma) \leftaction \{ [\alpha] \, | \, \alpha \subset \Sigma \}\)
148   given by
149   \begin{align*}
150     f \cdot \vec{[\alpha]} & = \vec{[\phi(\alpha)]} &
151     f \cdot [\alpha]       & = [\phi(\alpha)]
152   \end{align*}
153   for \(f = [\phi] \in \Mod(\Sigma)\).
154 \end{observation}
155 
156 \begin{definition}
157   Given a simple closed curve \(\alpha \subset \Sigma\), we denote by
158   \(\Mod(\Sigma)_{\vec{[\alpha]}}\) and \(\Mod(\Sigma)_{[\alpha]}\) the
159   subgroups of mapping classes that fix \(\vec{[\alpha]}\) -- for any given
160   choice of orientation of \(\alpha\) -- and \([\alpha]\), respectively.
161 \end{definition}
162 
163 While trying to understand the mapping class group of some surface \(\Sigma\),
164 it is interesting to consider how the geometric relationship between \(\Sigma\)
165 and other surfaces affects \(\Mod(\Sigma)\). Indeed, different embeddings
166 \(\Sigma' \hookrightarrow \Sigma\) translate to homomorphisms at the level of
167 mapping class groups.
168 
169 \begin{example}[Inclusion homomorphism]\label{ex:inclusion-morphism}
170   Let \(\Sigma' \subset \Sigma\) be a closed subsurface. Given \(\phi \in
171   \Homeo^+(\Sigma', \partial \Sigma')\), we may extend \(\phi\) to
172   \(\tilde{\phi} \in \Homeo^+(\Sigma, \partial \Sigma)\) by setting
173   \(\tilde{\phi}(x) = x\) for \(x \in \Sigma\) outside of \(\Sigma'\) -- which
174   is well defined since \(\phi\) fixes every point in \(\partial \Sigma'\).
175   This construction yields a group homomorphism
176   \begin{align*}
177     \Mod(\Sigma') & \to \Mod(\Sigma) \\
178      [\phi] & \mapsto [\tilde\phi],
179   \end{align*}
180   known as \emph{the inclusion homomorphism}.
181 \end{example}
182 
183 \begin{example}[Capping homomorphism]\label{ex:capping-morphism}
184   Let \(\delta \subset \partial \Sigma\) be a boundary component of \(\Sigma\).
185   We refer to the inclusion homomorphism \(\operatorname{cap} : \Mod(\Sigma)
186   \to \Mod(\Sigma \cup_\delta (\mathbb{D}^2 \setminus \{0\}))\) as \emph{the
187   capping homomorphism}.
188 \end{example}
189 
190 \begin{example}[Cutting homomorphism]\label{ex:cutting-morphism}
191   Given a simple closed curve \(\alpha \subset \Sigma\), any \(f \in
192   \Mod(\Sigma_{g+1})_{\vec{[\alpha]}}\) has a representative \(\phi \in
193   \Homeo^+(\Sigma, \partial \Sigma)\) fixing \(\alpha\) point-wise -- so that
194   \(\phi\) restricts to a homeomorphism of \(\Sigma \setminus \alpha\).
195   Furthermore, if \(\phi\!\restriction_{\Sigma \setminus \alpha} \simeq 1\) in
196   \(\Sigma \setminus \alpha\) then \(\phi \simeq 1 \in \Homeo^+(\Sigma,
197   \partial \Sigma)\) -- see \cite[Proposition~3.20]{farb-margalit}. There is
198   thus a group homomorphism
199   \begin{align*}
200     \operatorname{cut} : \Mod(\Sigma)_{\vec{[\alpha]}}
201     & \to \Mod(\Sigma\setminus\alpha) \\
202     [\phi] & \mapsto [\phi\!\restriction_{\Sigma \setminus \alpha}],
203   \end{align*}
204   known as \emph{the cutting homomorphism}.
205 \end{example}
206 
207 As goes for most groups, another approach to understanding the mapping class
208 group of a given surface \(\Sigma\) is to study its actions. We have already
209 seen simple examples of such actions in
210 Observation~\ref{ex:action-on-punctures} and
211 Observation~\ref{ex:action-on-curves}. An important class of actions of
212 \(\Mod(\Sigma)\) are its \emph{linear representations} -- i.e. the group
213 homomorphisms \(\Mod(\Sigma) \to \GL_n(\mathbb{C})\). These may be seen as
214 actions \(\Mod(\Sigma) \leftaction \mathbb{C}^n\) where each \(f \in
215 \Mod(\Sigma)\) acts via some linear isomorphism \(\mathbb{C}^n \isoto
216 \mathbb{C}^n\).
217 
218 \section{Representations}
219 
220 Here we collect a few fundamental examples of linear representations of
221 \(\Mod(\Sigma)\).
222 
223 \begin{observation}
224   Recall \(H_1(\Sigma_g, \mathbb{Z}) \cong \mathbb{Z}^{2g}\), with standard
225   basis given by \([\alpha_1], [\beta_1], \ldots, [\alpha_g], [\beta_g] \in
226   H_1(\Sigma_g, \mathbb{Z})\) for \(\alpha_1, \ldots, \alpha_g, \beta_1,
227   \ldots, \beta_g\) as in Figure~\ref{fig:homology-basis}. The Abelian group
228   \(H_1(\Sigma_g, \mathbb{Z})\) is endowed with a natural
229   \(\mathbb{Z}\)-bilinear alternating form given by the \emph{algebraic
230   intersection number} \([\alpha] \cdot [\beta] = \sum_{x \in \alpha \cap
231   \beta} \operatorname{ind}\,x\) -- where the index \(\operatorname{ind}\,x =
232   \pm 1\) of an intersection point is given by
233   Figure~\ref{fig:intersection-index}. In terms of the standard basis of
234   \(H_1(\Sigma_g, \mathbb{Z})\), this form is given by
235   \begin{align}\label{eq:symplectic-form}
236     [\alpha_i] \cdot [\beta_j]  & = \delta_{i j} &
237     [\alpha_i] \cdot [\alpha_j] & = 0            &
238     [\beta_i]  \cdot [\beta_j]  & = 0
239   \end{align}
240   and thus coincides with the pullback of the standard \(\mathbb{Z}\)-bilinear
241   symplectic form in \(\mathbb{Z}^{2g}\).
242 \end{observation}
243 
244 \begin{example}[Symplectic representation]\label{ex:symplectic-rep}
245   Given \(f = [\phi] \in \Mod(\Sigma_g)\), we may consider the map \(\phi_* :
246   H_1(\Sigma_g, \mathbb{Z}) \to H_1(\Sigma_g, \mathbb{Z})\) induced at the
247   level of singular homology. By homotopy invariance, the map \(\phi_*\) is
248   independent of the choice of representative \(\phi\) of \(f\). By the
249   functoriality of homology groups we then get a \(\mathbb{Z}\)-linear action
250   \(\Mod(\Sigma_g) \leftaction H_1(\Sigma_g, \mathbb{Z}) \cong
251   \mathbb{Z}^{2g}\) given by \(f \cdot [\alpha] = \phi_*([\alpha]) =
252   [\phi(\alpha)]\). Since pushforwards by orientation-preserving homeomorphisms
253   preserve the indices of intersection points, \((f \cdot [\alpha]) \cdot (f
254   \cdot [\beta]) = [\alpha] \cdot [\beta]\) for all \(\alpha, \beta \subset
255   \Sigma_g\) and \(f \in \Mod(\Sigma_g)\). In light of
256   (\ref{eq:symplectic-form}), this implies \(\Mod(\Sigma_g)\) acts on
257   \(\mathbb{Z}^{2g}\) via symplectomorphisms. We thus obtain a group
258   homomorphism \(\psi : \Mod(\Sigma_g) \to \operatorname{Sp}_{2g}(\mathbb{Z})
259   \subset \GL_{2g}(\mathbb{C})\), known as \emph{the symplectic representation
260   of \(\Mod(\Sigma_g)\)}.
261 \end{example}
262 
263 \noindent
264 \begin{minipage}[b]{.47\linewidth}
265   \centering
266   \includegraphics[width=.9\linewidth]{images/homology-generators.eps}
267   \captionof{figure}{The curves $\alpha_1, \beta_1, \ldots, \alpha_g, \beta_g
268   \subset \Sigma_g$ that generate its first homology group.}
269   \label{fig:homology-basis}
270 \end{minipage}
271 \hspace{.6cm} %
272 \begin{minipage}[b]{.47\linewidth}
273   \centering
274   \includegraphics[width=.9\linewidth]{images/intersection-index.eps}
275   \vspace*{.75cm}
276   \captionof{figure}{The index of an intersection point $x \in \alpha \cap
277   \beta$.}
278   \label{fig:intersection-index}
279 \end{minipage}
280 
281 The symplectic representation already allows us to compute some important
282 examples of mapping class groups: namely, that of the torus \(\mathbb{T}^2 =
283 \Sigma_1\) and the once-punctured torus \(\Sigma_{1, 1}\).
284 
285 \begin{observation}[Alexander trick]\label{ex:alexander-trick}
286   The group \(\Homeo^+(\mathbb{D}^2, \mathbb{S}^1)\) of homeomorphisms of the
287   unit disk \(\mathbb{D}^2 \subset \mathbb{C}\) is contractible. In particular,
288   \(\Mod(\mathbb{D}^2) = 1\). Indeed, for any \(\phi \in
289   \Homeo^+(\mathbb{D}^2, \mathbb{S}^1)\) the isotopy
290   \begin{align*}
291     \phi_t : \mathbb{D}^2 & \to     \mathbb{D}^2 \\
292                         z & \mapsto
293     \begin{cases}
294       (1 - t) \phi(\sfrac{z}{1 - t}) & \text{if } 0 \le |z| \le 1 - t \\
295       z                              & \text{otherwise}
296     \end{cases}
297   \end{align*}
298   that ``fixes the band \(\{ z \in \mathbb{D}^2 : |z| \ge 1 - t \}\) and does
299   \(\phi\) inside the sub-disk \(\{ z \in \mathbb{D}^2 : |z| \le 1 - t\}\)''
300   joins \(\phi = \phi_0\) and \(1 = \phi_1\).
301 \end{observation}
302 
303 \begin{observation}\label{ex:mdg-once-punctured-disk}
304   By the same token, \(\Mod(\mathbb{D}^2 \setminus \{0\}) = 1\).
305 \end{observation}
306 
307 \begin{observation}[Linearity of $\Mod(\mathbb{T}^2)$]\label{ex:torus-mcg}
308   The symplectic representation \(\psi : \Mod(\mathbb{T}^2) \to
309   \operatorname{Sp}_2(\mathbb{Z}) = \operatorname{SL}_2(\mathbb{Z})\) is a
310   group isomorphism. In particular, \(\Mod(\mathbb{T}^2) \cong
311   \operatorname{SL}_2(\mathbb{Z})\). To see \(\psi\) is surjective, first
312   observe \(\mathbb{Z}^2 \subset \mathbb{R}^2\) is
313   \(\operatorname{SL}_2(\mathbb{Z})\)-invariant. Hence any matrix \(g \in
314   \operatorname{SL}_2(\mathbb{Z})\) descends to an orientation-preserving
315   homeomorphism \(\phi_g\) of the quotient \(\mathbb{T}^2 =
316   \mfrac{\mathbb{R}^2}{\mathbb{Z}^2}\), which satisfies \(\psi([\phi_g]) = g\).
317   To see \(\psi\) is injective we consider the curves \(\alpha_1\) and
318   \(\beta_1\) from Figure~\ref{fig:homology-basis}. Given \(f = [\phi] \in
319   \Mod(\mathbb{T}^2)\) with \(\psi(f) = 1\), \(f \cdot \vec{[\alpha_1]} =
320   \vec{[\alpha_1]}\) and \(f \cdot \vec{[\beta_1]} = \vec{[\beta_1]}\), so we
321   may choose a representative \(\phi\) of \(f\) fixing \(\alpha_1 \cup
322   \beta_1\) pointwise. Such \(\phi\) determines a homeomorphism \(\tilde \phi\)
323   of the surface \(\mathbb{T}_{\alpha_1 \beta_1}^2 \cong \mathbb{D}^2\)
324   obtained by cutting \(\mathbb{T}^2\) along \(\alpha_1\) and \(\beta_1\), as
325   in Figure~\ref{fig:cut-torus-along}. Now by the Alexander trick from
326   Observation~\ref{ex:alexander-trick}, \(\tilde\phi\) must be isotopic to the
327   identity. The isotopy \(\tilde\phi \simeq 1 \in \Homeo^+(\mathbb{D}^2,
328   \mathbb{S}^1)\) then descends to an isotopy \(\phi \simeq 1 \in
329   \Homeo^+(\mathbb{T}^2)\), so \(f = 1 \in \Mod(\mathbb{T}^2)\) as desired.
330 \end{observation}
331 
332 \begin{figure}[ht]
333   \centering
334   \includegraphics[width=.55\linewidth]{images/torus-cut.eps}
335   \caption{By cutting $\mathbb{T}^2$ along $\alpha_1$ we obtain a cylinder,
336   where $\beta_1$ determines a yellow arc joining the two boundary components.
337   Now by cutting along this yellow arc we obtain a disk.}
338   \label{fig:cut-torus-along}
339 \end{figure}
340 
341 \begin{observation}\label{ex:punctured-torus-mcg}
342   By the same token, \(\Mod(\Sigma_{1, 1}) \cong
343   \operatorname{SL}_2(\mathbb{Z})\).
344 \end{observation}
345 
346 \begin{remark}
347   Despite the fact \(\psi : \Mod(\mathbb{T}^2) \to
348   \operatorname{SL}_2(\mathbb{Z})\) is an isomorphism, the symplectic
349   representation is \emph{not} injective for surfaces of genus \(g \ge 2\) --
350   see \cite[Section~6.5]{farb-margalit} for a description of its kernel.
351   Korkmaz and Bigelow-Budney \cite{korkmaz-linearity, bigelow-budney} showed
352   there exist injective linear representations of \(\Mod(\Sigma_2)\), but the
353   question of linearity of \(\Mod(\Sigma_g)\) remains wide-open for \(g \ge
354   3\). Recently, Korkmaz \cite[Theorem~3]{korkmaz} established the lower bound
355   of \(3 g - 3\) for the dimension of an injective representation of
356   \(\Mod(\Sigma_g)\) in the \(g \ge 3\) case -- if one such representation
357   exists.
358 \end{remark}
359 
360 Another fundamental class of examples of representations are the so-called
361 \emph{TQFT representations}.
362 
363 \begin{definition}
364   A \emph{cobordism} between closed oriented surfaces \(\Sigma\) and
365   \(\Sigma'\) is a triple \((W, \phi_+, \phi_-)\) where \(W\) is a smooth
366   oriented compact \(3\)-manifold with \(\partial W = \partial_+ W \amalg
367   \partial_- W\), \(\phi_+ : \Sigma \isoto \partial_+ W\) is an orientation
368   preserving diffeomorphism and \(\phi_- : \Sigma' \isoto \partial_- W\) is an
369   orientation-reversing diffeomorphism. We may abuse the notation and denote
370   \(W = (W, \phi_+, \phi_-)\).
371 \end{definition}
372 
373 \begin{definition}
374   We denote by \(\Cob\) the category whose objects are (possibly disconnected)
375   closed oriented surfaces and whose morphisms \(\Sigma \to \Sigma'\) are
376   diffeomorphism classes\footnote{Here we only consider orientation-preserving
377   diffeomorphisms $\varphi : W \isoto W'$ that are compatible with the boundary
378   identifications in the sense that $\varphi(\partial_\pm W) = \partial_\pm W'$
379   and $\psi_\pm = \varphi \circ \phi_\pm$.} of cobordisms between \(\Sigma\)
380   and \(\Sigma'\), with composition given by
381   \[
382     [W, \phi_-, \phi_+] \circ [W', \psi_-, \psi_+]
383     = [W \cup_{\psi_- \circ \phi_+^{-1}} W', \phi_-, \psi_+]
384   \]
385   for \([W, \phi_-, \phi_+] : \Sigma \to \Sigma'\) and \([W', \psi_-, \phi_+] :
386   \Sigma' \to \Sigma''\). We endow \(\Cob\) with the monoidal structure given
387   by
388   \begin{align*}
389     \Sigma \otimes \Sigma'
390     & = \Sigma \amalg \Sigma' &
391     [W,\phi_+,\phi_-] \otimes [W',\psi_+,\psi_-]
392     & = [W \amalg W', \phi_+ \amalg \psi_+, \phi_- \amalg \psi_-].
393   \end{align*}
394 \end{definition}
395 
396 \begin{definition}[TQFT]\label{def:tqft}
397   A \emph{topological quantum field theory} (abbreviated by \emph{TQFT})
398   is a functor \(\mathcal{F} : \Cob \to \Vect\) satisfying
399   \begin{align*}
400     \mathcal{F}(\emptyset) & = \mathbb{C} &
401     \mathcal{F}(\Sigma \otimes \Sigma')
402     & = \mathcal{F}(\Sigma) \otimes \mathcal{F}(\Sigma') &
403     \mathcal{F}([W] \otimes [W'])
404     & = \mathcal{F}([W]) \otimes \mathcal{F}([W']),
405   \end{align*}
406   where \(\Vect\) denotes the category of finite-dimensional complex vector
407   spaces.
408 \end{definition}
409 
410 \begin{observation}
411   Given \(\phi \in \Homeo^+(\Sigma_g)\), we may consider the so-called
412   \emph{mapping cylinder} \(C_\phi = (\Sigma_g \times [0, 1], \phi, 1)\), a
413   cobordism between \(\Sigma_g\) and itself -- where \(\partial_+ (\Sigma_g
414   \times [0, 1]) = \Sigma_g \times 0\) and \(\partial_- (\Sigma_g \times [0,
415   1]) = \Sigma_g \times 1\). The diffeomorphism class of \(C_\phi\) is
416   independent of the choice of representative of \(f = [\phi] \in
417   \Mod(\Sigma_g)\), so \(C_f = [C_\phi] : \Sigma_g \to \Sigma_g\) is a well
418   defined morphism in \(\Cob\).
419 \end{observation}
420 
421 \begin{example}[TQFT representations]\label{ex:tqft-reps}
422   It is clear that \(C_1\) is the identity morphism \(\Sigma_g \to \Sigma_g\)
423   in \(\Cob\). In addition, \(C_{f \cdot g} = C_f \circ C_g\) for all \(f, g
424   \in \Mod(\Sigma_g)\) -- see \cite[Lemma~2.5]{costantino}. Now given a TQFT
425   \(\mathcal{F} : \Cob \to \Vect\), by functoriality we obtain a linear
426   representation
427   \begin{align*}
428     \rho_{\mathcal{F}} : \Mod(\Sigma_g) & \to \GL(\mathcal{F}(\Sigma_g)) \\
429                                       f & \mapsto \mathcal{F}(C_f).
430   \end{align*}
431 \end{example}
432 
433 As simple as the construction in Example~\ref{ex:tqft-reps} is, in practice it
434 is not that easy to come across functors as the ones in
435 Definition~\ref{def:tqft}. This is because, in most interesting examples, we
436 are required to attach some extra data to our surfaces to get a well defined
437 association \(\Sigma_g \mapsto \mathcal{F}(\Sigma_g)\). Moreover, the condition
438 \(\mathcal{F}([W] \circ [W']) = \mathcal{F}([W]) \circ \mathcal{F}([W'])\) may
439 only hold up to multiplication by scalars.
440 
441 Hence constructing an actual functor typically requires \emph{extending}
442 \(\Cob\) and \emph{tweaking} \(\Vect\). Such functors give rise to linear and
443 projective representations of the \emph{extended mapping class groups}
444 \(\Mod(\Sigma_g) \times \mathbb{Z}\). We refer the reader to \cite{costantino,
445 julien} for constructions of one such TQFT and its corresponding
446 representations: the so-called \emph{\(\operatorname{SU}_2\) TQFT of level
447 \(r\)}, first introduced by Witten and Reshetikhin-Tuarev \cite{witten,
448 reshetikhin-turaev} in their foundational papers on quantum topology.
449 
450 Besides Example~\ref{ex:symplectic-rep} and Example~\ref{ex:tqft-reps}, not a
451 lot of other linear representations of \(\Mod(\Sigma_g)\) are known. Indeed,
452 the representation theory of mapping class groups remains a mystery at large.
453 In Chapter~\ref{ch:representations} we provide a brief overview of the field,
454 as well as some recent developments. More specifically, we highlight Korkmaz'
455 \cite{korkmaz} proof of the triviality of low-dimensional representations and
456 comment on his classification of \(2g\)-dimensional representations. To that
457 end, in Chapter~\ref{ch:dehn-twists} and Chapter~\ref{ch:relations} we survey
458 the group structure of mapping class groups: its relations and known
459 presentations.
460