14. The Riemann mapping theorem  PDF TEX

The Riemann mapping theorem

The Riemann mapping theorem

  • \(D=\{z\in \mathbb{C}: |z|<1\}\) is the open unit disc centered at the origin.

Theorem.

Let \(\Omega \neq \mathbb{C}\) be a simply connected region. Then \(\Omega\) is conformally equivalent to \(D\). Moreover, the assumption \(\Omega \neq \mathbb{C}\) is necessary.

Remark.

  • In view of the Liouville theorem, the assumption \(\Omega \neq \mathbb{C}\) is necessary. Indeed, if \(f:\mathbb C\to D\) is a conformal map, then \(f\) is bounded, since \(|f(z)|<1\) for all \(z\in \mathbb C\). Hence, by the Liouville theorem \(f\) must be constant, but then it cannot be injective.

  • However, \(\mathbb C\) and \(D\) are homeomorphic. The mapping \[\mathbb C\ni z\mapsto \frac{z}{1+|z|}\in D\] is the desired homeomorphism.

Simple topological lemma

Lemma.

Let \(\Omega\) be an open subset of \(\mathbb C\). Then there exists a sequence \((K_{n})_{n \in\mathbb N}\subseteq \Omega\) of compact sets such that \[\Omega=\bigcup_{n=1}^{\infty} K_{n}\quad \text{ and } \quad K_{n} \subseteq \operatorname{int} K_{n+1} \quad \text{ for } \quad n\in\mathbb N,\] where \(\operatorname{int}K_{n+1}\) denotes the interior of \(K_{n+1}\). Further for a compact set \(K\subseteq \Omega\), we have \(K \subseteq K_{n}\) for some \(n \in\mathbb N\).

Proof: For \(n \in\mathbb N\), let \[K_n=\overline{D}(0, n)\cap\{z \in\Omega: d(z, \mathbb{C}\setminus\Omega) \geq 1/n\}.\]

  • It is clear that \(K_{n} \subseteq \Omega\) and it is bounded.

  • Each \(K_n\) is closed and consequently compact.

  • Moreover, \(K_{n} \subseteq K_{n+1}\) for \(n\in\mathbb N\).

  • We prove that \(K_{n} \subseteq \operatorname{int} K_{n+1}\). Taking \(r=\frac{1}{n}-\frac{1}{n+1}\), it suffices to show \[D(z, r) \subseteq K_{n+1} \quad \text { for } \quad z \in K_{n}.\]

  • Let \(z \in K_{n}\) and \(\zeta \in D(z, r)\). Then \[|\zeta-z|<\frac{1}{n}-\frac{1}{n+1} \quad \text { and } \quad |z-a| \geq \frac{1}{n} \quad \text { for }\quad a \notin \Omega.\]

  • Therefore for \(a \notin \Omega\), we have \[|\zeta-a| \geq|z-a|-|\zeta-z| \geq \frac{1}{n}-\left(\frac{1}{n}-\frac{1}{n+1}\right)=\frac{1}{n+1}.\]

  • This implies \(\zeta \in K_{n+1}\), since \(|\zeta|<|z|+\frac{1}{n} \leq n+\frac{1}{n} \leq n+1\).

  • Hence \(K_{n} \subseteq \operatorname{int}{K}_{n+1}\) for \(n \in\mathbb N\) and \(K_{1} \subseteq K_{2} \subseteq K_{3} \subseteq \ldots\).

  • We observe that \(\bigcup_{n=1}^{\infty} K_{n} \subseteq \Omega\), since each \(K_{n} \subseteq \Omega\).

  • For the reverse inclusion, let \(z \in \Omega\) and \(M\in\mathbb N\) be so that \(|z| \leq M\). Since \(\Omega\) is open, there exists \(\rho>0\) such that \(D(z, \rho) \subseteq \Omega\). Let \(N\in\mathbb N\) be the least integer greater than or equal to \(\max (M, \rho^{-1})\).

  • Then \(|z| \leq M \leq N\) and for \(a \notin \Omega\) we have \(|z-a| \geq \rho \geq \frac{1}{N}\). Therefore \(z \in K_{N}\) and thus \(\Omega \subseteq \bigcup_{n=1}^{\infty} K_{n}\), hence \(\Omega=\bigcup_{n=1}^{\infty} K_{n}\).

  • We also have a stronger result \(\Omega=\bigcup_{n=1}^{\infty} \operatorname{int}{K}_{n}\).

  • Indeed, since \(K_{n} \subseteq \operatorname{int}{K}_{n+1}\), we have \[\Omega \subseteq \bigcup_{n=1}^{\infty} K_{n} \subseteq \bigcup_{n=2}^{\infty} \operatorname{int}{K}_{n} \subseteq \bigcup_{n=1}^{\infty} \operatorname{int}{K}_{n}\subseteq \Omega.\]

  • Let \(K\) be a compact subset of \(\Omega\). Then \[K \subseteq \bigcup_{n=1}^{\infty} \operatorname{int}{K}_{n}\] is an open cover of \(K\). Therefore, \(K \subseteq \bigcup_{n=1}^{P} \operatorname{int}{K}_{n}\) for some \(P\in \mathbb N\).

  • Thus \[K \subseteq K_{1} \cup K_{2}\cup \ldots \cup K_{P}=K_{P}. \qquad \tag*{$\blacksquare$}\]

Hurwitz lemma

Lemma.

Let \(\Omega\subseteq \mathbb C\) be a region and \((f_{n})_{n \in\mathbb N}\) be a sequence such that \(f_{n} \in H(\Omega)\) converges uniformly on compact subsets of \(\Omega\). Assume that \(f_{n}\) is one-to-one for any \(n \in\mathbb N\). Then the limit function \(f=\lim_{n\to\infty}f_n\) is either constant or one-to-one on \(\Omega\).

Proof: Note that \(f \in H(\Omega)\). ( Why?)

  • Assume that \(f\) is not one-to-one, then there exist \(z_{1}, z_{2} \in \Omega\) such that \(z_{1} \neq z_{2}\) with \(f\left(z_{1}\right)=f\left(z_{2}\right)\).

  • For \(n \in\mathbb N\), we define \(g_{n} \in H(\Omega)\), and \(g \in H(\Omega)\) by setting \[g_{n}(z)=f_{n}(z)-f_{n}\left(z_{1}\right), \quad \text { and }\quad g(z)=f(z)-f\left(z_{1}\right).\] Then \(g\left(z_{2}\right)=0\) and also \(g_{n}\left(z_{2}\right) \neq 0\), since \(f_{n}\) is one-to-one.

  • Further, we may suppose that \(g\) is not constant in \(\Omega\) otherwise \(f\) is constant in \(\Omega\) and the assertion follows.

  • Since the zeros of \(g\) are isolated, there is \(r>0\) so that \(\left|z_{1}-z_{2}\right|>r\) and \(g(z) \neq 0\) whenever \(z\in\Omega\) satisfies \(0<\left|z-z_{2}\right| \leq r\).

  • Let \(\gamma\) be a circle centered at \(z_{2}\) with radius \(r\). Then there exists \(\delta>0\) such that \(|g(z)|>\delta\) for \(z \in \gamma^{*}\).

  • By the uniform convergence there exists \(n_0\in\mathbb N\) such that for every \(n \geq n_{0}\), we have \[\left|g_{n}(z)-g(z)\right|<\frac{\delta}{2} \quad \text { uniformly for any }\quad z \in \gamma^{*}.\]

  • Therefore for \(z \in \gamma^{*}\), we have \[\left|g_{n}(z)\right| \geq|g(z)|-\left|g_{n}(z)-g(z)\right| \geq \delta-\frac{\delta}{2}=\frac{\delta}{2}.\]

  • Now we see that \(\frac{1}{g_{n}(z)}\) converges uniformly to \(\frac{1}{g(z)}\) on \(\gamma^*\). Also \(g_{n}^{\prime}(z)\) converges uniformly to \(g^{\prime}(z)\) on \(\gamma^*\).

  • Hence \[\lim _{n \rightarrow \infty} \frac{g_{n}^{\prime}(z)}{g_{n}(z)}=\frac{g^{\prime}(z)}{g(z)}\] uniformly on \(\gamma^*\).

  • We observe that \(\frac{g_{n}^{\prime}(z)}{g_{n}(z)} \in H\left(\overline{D}\left(z_{2}, r\right)\right)\). Now by the Cauchy theorem we obtain that \[0=\frac{1}{2 \pi i} \int_{\gamma} \frac{g_{n}^{\prime}(z)}{g_{n}(z)} d z \quad \text { for }\quad n \in\mathbb N.\]

  • Consequently, by the argument principle \[0=\lim _{n \rightarrow \infty}\frac{1}{2 \pi i} \int_{\gamma} \frac{g_{n}^{\prime}(z)}{g_{n}(z)} d z=\frac{1}{2 \pi i} \int_{\gamma} \frac{g^{\prime}(z)}{g(z)} d z=N_g \geq 1.\]

  • This is a contradiction completing the proof of Hurwitz lemma.$$\tag*{$\blacksquare$}$$

Normal families of analytic functions

Definition.
  • Let \(\Omega\subseteq \mathbb C\) be a region and \(\mathcal{F}\) be a family of analytic functions in \(\Omega\). Thus \(\mathcal{F}\) is a sub-family of \(H(\Omega)\). Further \(\mathcal{F}\) is called a normal family if every sequence of elements of \(\mathcal{F}\) contains a subsequence which converges uniformly on compact subsets of \(\Omega\).

  • \(\mathcal{F}\) is uniformly bounded on compact subsets of \(\Omega\) if for every compact subset \(K\) of \(\Omega\) there exists \(M=M(K)\) such that \[|f(z)| \leq M \quad \text { for } \quad f \in \mathcal{F},\ z \in K.\]

  • The family \(\mathcal{F}\) is called equicontinuous on compact subsets of \(\Omega\) if for \(\varepsilon>0\) and compact subset \(K \subseteq \Omega\) there exists \(\delta>0\) depending only on \(\varepsilon\) and \(K\) such that \(\left|f\left(z_{1}\right)-f\left(z_{2}\right)\right|<\varepsilon\) for \(f \in \mathcal{F}\) and \(z_{1}, z_{2} \in K\) satisfying \(\left|z_{1}-z_{2}\right|<\delta\).

Montel lemma

Remark.

  • The limit of the subsequence in the above definition (i) belongs to \(H(\Omega)\) but it need not belong to \(\mathcal{F}\).

Lemma.

Let \(\mathcal{F}\) be a family of \(H(\Omega)\) with \(\Omega\subseteq \mathbb C\) region. Assume that \(\mathcal{F}\) is uniformly bounded on compact subsets of \(\Omega\). Then \(\mathcal{F}\) is a normal family.

Proof: We show that \(\mathcal{F}\) is equicontinuous on compact subsets of \(\Omega\).

  • By the topological lemma, there exists a sequence of compact sets \((K_{n})_{n\in\mathbb N}\subseteq \Omega\) such that \(K_{n} \subseteq \operatorname{int}K_{n+1}\) for \(n\in\mathbb N\) that satisfies \[\Omega=\bigcup_{n=1}^{\infty} K_{n}.\] Also, every compact subset of \(\Omega\) is contained in \(K_{n}\) for some \(n\in\mathbb N\).

  • Let \(n \in\mathbb N\). Since \(K_{n}\) is compact, \(\left(\operatorname{int}K_{n+1}\right)^{c}\) is closed and \[K_{n} \cap\left(\operatorname{int}K_{n+1}\right)^{c}=\varnothing,\]

  • We can also find \(\delta_{n}>0\) such that \[\left|z_{1}-z_{2}\right|>2 \delta_{n}\quad \text { for } \quad z_{1} \in K_{n}, \quad z_{2} \notin \operatorname{int}K_{n+1}.\]

  • Thus \[\overline{D}\left(z, 2 \delta_{n}\right) \subseteq \operatorname{int}K_{n+1} \subseteq K_{n+1} \quad \text { for } \quad z \in K_{n}.\]

  • Let \(z^{\prime} \in K_{n}\), and \(z^{\prime \prime} \in K_{n}\) with \(\left|z^{\prime}-z^{\prime \prime}\right|<\delta_{n}\).

  • Let \(\gamma\) be a circle centered at \(z^{\prime}\) and with radius \(2 \delta_{n}\).

  • Thus \(z^{\prime \prime}\) lies inside the circle \(\gamma\) and \(\gamma^{*} \subseteq K_{n+1}\).

  • Let \(f \in \mathcal{F}\), then by the Cauchy integral formula \[\begin{aligned} f\left(z^{\prime}\right)-f\left(z^{\prime \prime}\right) & =\frac{1}{2 \pi i} \int_{\gamma} f(\zeta)\left(\frac{1}{\zeta-z^{\prime}}-\frac{1}{\zeta-z^{\prime \prime}}\right) d \zeta \\ & =\frac{z^{\prime}-z^{\prime \prime}}{2 \pi i} \int_{\gamma} \frac{f(\zeta)}{\left(\zeta-z^{\prime}\right)\left(\zeta-z^{\prime \prime}\right)} d \zeta. \end{aligned}\]

  • If \(\left|z^{\prime}-z^{\prime \prime}\right|<\delta_{n}\), we observe that \[\left|\zeta-z^{\prime}\right|=2 \delta_{n} \quad \text { and } \quad \left|\zeta-z^{\prime \prime}\right|=\left|\zeta-z^{\prime}+z^{\prime}-z^{\prime \prime}\right|>2 \delta_{n}-\delta_{n}=\delta_{n}.\]

  • Since \(\mathcal{F}\) is uniformly bounded on compact subsets of \(\Omega\), there exists a constant \(M\left(K_{n+1}\right)\) depending only on \(K_{n+1}\) such that \[\left|f\left(z^{\prime}\right)-f\left(z^{\prime \prime}\right)\right|<\frac{\left|z^{\prime}-z^{\prime \prime}\right|}{2 \pi} \frac{4\pi \delta_n}{2 \delta_{n}^2 } M\left(K_{n+1}\right)=\frac{M\left(K_{n+1}\right)}{\delta_{n}}\left|z^{\prime}-z^{\prime \prime}\right|.\]

  • The above inequality holds for all \(f \in \mathcal{F}\) and \(z^{\prime}, z^{\prime \prime} \in K_{n}\) whenever \(\left|z^{\prime}-z^{\prime \prime}\right|<\delta_{n}\).

  • Let \(\varepsilon>0\) and set \[\delta=\delta(n)=\frac{\varepsilon \delta_{n}}{\varepsilon+M\left(K_{n+1}\right)}<\delta_{n}.\]

  • Then for \(\left|z^{\prime}-z^{\prime \prime}\right|<\delta\), we have \[\frac{M\left(K_{n+1}\right)}{\delta_{n}}\left|z^{\prime}-z^{\prime \prime}\right|<\frac{M\left(K_{n+1}\right)}{\delta_{n}} \delta=\frac{\varepsilon M\left(K_{n+1}\right)}{\varepsilon+M\left(K_{n+1}\right)}<\varepsilon.\]

  • Therefore we have \[\left|f\left(z^{\prime}\right)-f\left(z^{\prime \prime}\right)\right|<\varepsilon \tag{*}\] for \(f \in \mathcal{F}\) and \(z^{\prime}, z^{\prime \prime} \in K_{n}\) and \(\left|z^{\prime}-z^{\prime \prime}\right|<\delta\).

  • Thus the family \(\mathcal{F}\) is equicontinuous on compact subsets of \(\Omega\), since every compact subset \(K\) of \(\Omega\) is contained in \(K_{n}\) for some \(n\in\mathbb N\), and therefore (*) holds for all \(f \in \mathcal{F}\) and \(z^{\prime}, z^{\prime \prime} \in K\) so that \(\left|z^{\prime}-z^{\prime \prime}\right|<\delta\).

  • Let \((f_{m})_{m\in\mathbb N}\subseteq\mathcal{F}\). We show that it has a subsequence which converges uniformly on compact subsets of \(\Omega\).

  • Let \(E\) be a countable dense set of \(\Omega\).

  • For example, we take \(E\) to be the set of all points of \(\Omega\) with rational coordinates. We arrange the elements of \(E\) as \(w_{1}, w_{2}, w_3, w_4, \ldots\).

  • Since \((f_{m}\left(w_{1}\right))_{m\in\mathbb N}\) is a bounded sequence by assumption on \(\mathcal F\), the sequence \((f_{m})_{m\in\mathbb N}\) has a subsequence \((f_{m1})_{m\in\mathbb N}\) so that \((f_{m 1}\left(w_{1}\right))_{m\in\mathbb N}\) converges. Here, we have used the Bolzano–Weierstrass theorem.

  • Similarly, the sequence \((f_{m1})_{m\in\mathbb N}\) has a subsequence \((f_{m2})_{m\in\mathbb N}\) such that \((f_{m 2}\left(w_{2}\right))_{m\in\mathbb N}\) converges.

  • Proceeding recursively, we see that for \(i \in\mathbb N\) there is a subsequence \((f_{m i})_{i\in\mathbb N}\) of \((f_{m, i-1})_{i\in\mathbb N}\) such that \((f_{m i}\left(w_{i}\right))_{i\in\mathbb N}\) converges.

  • Here we write \(f_{m 0}\) for \(f_{m}\).

  • Now the diagonal sequence \((f_{mm})_{m\in\mathbb N}\) converges at all \(w\in E\).

  • We show that \((f_{mm})_{m\in\mathbb N}\) converges uniformly on \(K_{n}\) for any \(n \in\mathbb N\).

  • Then the diagonal sequence \((f_{mm})_{m\in\mathbb N}\) converges uniformly on all compact subsets \(K\) of \(\Omega\) since \(K \subseteq K_{n}\) for some \(n\in\mathbb N\).

  • For \(n \in \mathbb N\) and \(\delta=\delta(n)\) as above, we have \[K_{n} \subseteq \bigcup_{z \in K_{n}} D(z, \delta).\]

  • Then \[K_{n} \subseteq \bigcup_{z \in E \cap K_{n}} D(z, \delta),\] since \(E \cap K_{n}\) is dense in \(K_{n}\). Since \(K_{n}\) is compact, we observe that the above open cover admits a finite subcover.

  • Thus there exist \(z_{1}, z_{2}, \ldots, z_{p} \in E \cap K_{n}\) such that \[K_{n} \subseteq D\left(z_{1}, \delta\right) \cup D\left(z_{2}, \delta\right)\cup \cdots \cup D\left(z_{p}, \delta\right).\]

  • For \(\varepsilon>0\), there exists \(M\) depending only on \(\varepsilon\) and \(K_{n}\) such that \[\left|f_{r r}\left(z_{i}\right)-f_{s s}\left(z_{i}\right)\right| \leq \varepsilon\] for \(r \geq M, s \geq M\) and \(1 \leq i \leq p\).

  • Let \(z \in K_{n}\). Then \(z \in D\left(z_{i}, \delta\right)\) for some \(i\) with \(1 \leq i \leq p\).

  • Further, by the equicontinuity, we have \[\begin{aligned} |f_{r r}(z)&-f_{s s}(z)| \\ & \leq\left|f_{r r}(z)-f_{r r}\left(z_{i}\right)\right|+\left|f_{r r}\left(z_{i}\right)-f_{s s}\left(z_{i}\right)\right|+\left|f_{s s}(z)-f_{s s}\left(z_{i}\right)\right| \\ & <\varepsilon+\varepsilon+\varepsilon=3 \varepsilon \end{aligned}\] wherever \(r \geq M, s \geq M\). Thus \((f_{mm})_{m\in\mathbb N}\) converges uniformly on \(K_{n}\).

  • Since every compact subset of \(\Omega\) is contained in \(K_{n}\) for some \(n\in\mathbb N\), we conclude that \((f_{mm})_{m\in\mathbb N}\) converges uniformly on compact subsets of \(\Omega\). This completes the proof of Montels’ lemma. $$\tag*{$\blacksquare$}$$

Proof: Assume that \(\Omega\neq \mathbb C\) is a simply connected region. Let \(z_{0} \in \Omega\) and define \[\Sigma=\{\psi \in H(\Omega): \psi \text{ is one-to-one on } \Omega \text{ and } \psi(\Omega) \subseteq D\}.\] Our aim is to prove that \(\Sigma\) contains an element which is onto.

  • In fact, we will show that for \(\psi \in \Sigma\), which is not onto \(D\), there exists \(\psi_{1} \in \Sigma\) such that \[\left|\psi_{1}^{\prime}\left(z_{0}\right)\right|>\left|\psi^{\prime}\left(z_{0}\right)\right|. \tag{*}\]

  • Next we consider \[\eta=\sup \left\{\left|\psi^{\prime}\left(z_{0}\right)\right|: \psi \in \Sigma\right\},\tag{**}\] and we will prove that the supremum is assumed for some \(\psi_{0} \in \Sigma\).

  • Then it will be clear that \(\psi_{0}\) has to be an onto function. Otherwise, if \(\psi_{0}\) is not onto \(D\) then there is \(\psi_1\in\Sigma\) satisfying (*). Hence, by (**), we have \(\eta=\left|\psi_0^{\prime}\left(z_{0}\right)\right|<\left|\psi_{1}^{\prime}\left(z_{0}\right)\right|\le \eta\), which is impossible.

The proof will consist of a few steps.

Step 1. We first prove that \(\Sigma\neq\varnothing\).

  • From Lecture 8 we know that if \(\Omega\subseteq \mathbb C\) is a simply connected region, then for every closed curve \(\Gamma\) in \(\Omega\) we have \[\operatorname{Ind}_{\Gamma}(\alpha)=0 \quad \text{ whenever }\quad \alpha\not\in \Omega.\]

  • From Lecture 9, we know that the latter is equivalent to the statement that for every \(f\in H(\Omega)\) satisfying \(1/f\in H(\Omega)\) there exists \(g\in H(\Omega)\) such that \(f=g^2\).

  • Since \(\Omega \neq \mathbb{C}\), let \(w_{0} \in \mathbb{C}\) such that \(w_{0} \notin \Omega\). Consider \[f(z)=z-w_{0}.\]

  • We observe that \(f \in H(\Omega)\) and \(f\) has no zero in \(\Omega\) since \(w_{0} \notin \Omega\).

  • Thus \(1 / f \in H(\Omega)\). Therefore, there exists \(g \in H(\Omega)\) such that \[f(z)=g^{2}(z) \quad \text { for }\quad z \in \Omega.\]

  • Let \(a \in \Omega\). By open mapping theorem, we observe that \(g(\Omega)\) is open containing \(g(a)\).

  • Therefore there exists \(r>0\) such that \[D(g(a), r) \subseteq g(\Omega).\]

  • Now we show that \(D(-g(a), r) \cap g(\Omega)=\varnothing\). Suppose there exists \(z_{1} \in \Omega\) such that \(g\left(z_{1}\right) \in D(-g(a), r)\). Thus \[\left|g\left(z_{1}\right)+g(a)\right|<r \quad \iff \quad \left|-g\left(z_{1}\right)-g(a)\right|<r\]

  • Then \[-g\left(z_{1}\right) \in D(g(a), r) \subseteq g(\Omega) .\]

  • Therefore there exists \(z_{2} \in \Omega\) such that \(-g\left(z_{1}\right)=g\left(z_{2}\right)\).

  • By squaring both sides of this equation, we derive that \[z_{1}-w_{0}=f(z_1)=\left(-g\left(z_{1}\right)\right)^{2}=\left(g\left(z_{2}\right)\right)^{2}=f(z_2)=z_{2}-w_{0}\] implying \(z_{1}=z_{2}\). Thus \(g\left(z_{1}\right)=0\), and consequently \(z_{1}-w_{0}=0\).

  • This is a contradiction since \(z_{1} \in \Omega\) and \(w_{0} \notin \Omega\). Hence, \[D(-g(a), r) \cap g(\Omega)=\varnothing.\]

  • Now we consider \[\psi(z)=\frac{r}{g(z)+g(a)} \quad \text { for }\quad z \in \Omega.\]

  • We observe that \(\psi(z) \in H(\Omega)\) and \(|\psi(z)| \leq 1\) for \(z \in \Omega\), since \[|g(z)+g(a)| \geq r\quad \text{ for } \quad z \in \Omega.\]

  • In fact \(|\psi(z)|<1\) for \(z \in \Omega\) by the maximum modulus principle.

  • Further \(\psi \in H(\Omega)\) and is one-to-one on \(\Omega\), since \(\psi\left(z^{\prime}\right)=\psi\left(z^{\prime \prime}\right)\) implies \(g^{2}\left(z^{\prime}\right)=g^{2}\left(z^{\prime \prime}\right)\), and therefore \(z^{\prime}=z^{\prime \prime}\).

  • Hence \(\psi \in \Sigma\) as desired.

Step 2. We will show that if \(\psi \in \Sigma\) and \(\psi(\Omega)\) is a proper subset of \(D\), then there exists \(\psi_{1} \in \Sigma\) satisfying \[\left|\psi_{1}^{\prime}\left(z_{0}\right)\right|>\left|\psi^{\prime}\left(z_{0}\right)\right|. \tag{*}\]

  • Fix \(\psi \in \Sigma\). Since \(\psi(\Omega)\) is a proper subset of \(D\), there exists \(\alpha \in D\) such that \(\alpha \notin \psi(\Omega)\). We consider \(\phi_{\alpha} \circ \psi\), where \[\phi_{\alpha}(w)=\frac{w-\alpha}{1-\overline{\alpha} w}.\]

  • We recall that \(\phi_{\alpha}\) is an automorphism of \(D\). For \(z \in \Omega\), observe that \[\phi_{\alpha} \circ \psi(z)=\phi_{\alpha}(\psi(z))=\frac{\psi(z)-\alpha}{1-\overline{\alpha} \psi(z)}=0\] only when \(\psi(z)=\alpha\) which is not the case since \(\alpha \notin \psi(\Omega)\).

  • Therefore \(\phi_{\alpha} \circ \psi \in H(\Omega)\) has no zero in \(\Omega\). Then, there exists \(g \in H(\Omega)\) such that \[g^{2}(z)=\phi_{\alpha} \circ \psi(z) \quad \text { for }\quad z \in \Omega.\]

  • By writing \(s(w)=w^{2}\) for \(w \in D\), we rewrite the last equation as \[s \circ g=\phi_{\alpha} \circ \psi\quad \text { in } \quad \Omega.\]

  • If \(g\left(z_{1}\right)=g\left(z_{2}\right)\) for \(z_{1},z_{2}\in\Omega\), then \(\psi\left(z_{1}\right)=\psi\left(z_{2}\right)\), since \(\phi_{\alpha}\) is one-to-one. Therefore, \(g\) is one-to-one, and \(g \in \Sigma\).

  • Let \[\psi_{1}=\phi_{\beta} \circ g \quad \text { with }\quad g\left(z_{0}\right)=\beta.\]

  • Observe that \[\psi_{1}\left(z_{0}\right)=\phi_{\beta}\left(g\left(z_{0}\right)\right)=\phi_{\beta}(\beta)=0.\]

  • Hence, \[\psi=\phi_{-\alpha} \circ s \circ g=\phi_{-\alpha} \circ s \circ \phi_{-\beta} \circ \psi_{1}=F \circ \psi_{1},\] where \[F=\phi_{-\alpha} \circ s \circ \phi_{-\beta}.\]

  • By the chain rule and \(\psi_{1}\left(z_{0}\right)=0\), we have \[\psi^{\prime}\left(z_{0}\right)=F^{\prime}\left(\psi_{1}\left(z_{0}\right)\right) \psi_{1}^{\prime}\left(z_{0}\right)=F^{\prime}(0) \psi_{1}^{\prime}\left(z_{0}\right).\]

  • Inequality (*) will follow if we show that \(\left|F^{\prime}(0)\right|<1\), since \[\left|\psi^{\prime}\left(z_{0}\right)\right|=\left|F^{\prime}(0)\right|\left|\psi_{1}^{\prime}\left(z_{0}\right)\right|.\]

  • Recall the following lemma from the previous lecture.

Lemma.

Let \(f\) be non-constant and analytic in \(D\), and satisfy \(|f(z)|<1\) for \(z\in D\). Let \(w \in D\) with \(f(w)=a\). Then \[\left|f^{\prime}(w)\right| \leq \frac{1-|a|^{2}}{1-|w|^{2}} .\] Moreover equality occurs only when \[f=\phi_{-a} \circ\left(c \phi_{w}\right) \text { in } D,\] for some constant \(c\) whose absolute value is \(1\).

  • We observe that \(F(D) \subseteq D\), and let \(F(0)=a\). Then by the lemma with \(w=0\), we have \[\left|F^{\prime}(0)\right| \leq 1-|a|^{2}.\]

  • Suppose that \(\left|F^{\prime}(0)\right|=1\). Then \(a=0\) and by the second part of the previous lemma, we conclude with \(w=0\) that \(F(z)=\lambda z\) for \(z \in D\) where \(\lambda\) is a constant of absolute value \(1\).

  • This is not possible since \(F\) is not one-to-one as \(s(w)=w^{2}\) is not one-to-one. Hence \(\left|F^{\prime}(0)\right|<1\) and the proof of (*) is complete.

Step 3. In this step we finish the proof. In Step 1, we have proved that \(\Sigma\neq\varnothing\), hence we can define \[\eta=\sup \left\{\left|\psi^{\prime}\left(z_{0}\right)\right|: \psi \in \Sigma\right\},\tag{**}\]

  • By the inverse mapping theorem we have \(\left|\psi^{\prime}\left(z_{0}\right)\right|>0\) for every \(\psi \in \Sigma\). Hence \(\eta>0\).

  • There exists a sequence \((\psi_{n})_{n\in\mathbb N}\subseteq\Sigma\) such that \[\lim _{n \rightarrow \infty}\left|\psi_{n}^{\prime}\left(z_{0}\right)\right|=\eta>0.\]

  • We observe that \(|\psi(z)|<1\) for \(\psi \in \Sigma\).

  • In particular, \(\Sigma\) is uniformly bounded on compact subsets of \(\Omega\).

  • Therefore \(\Sigma\) is a normal family by Montel’s lemma.

  • Hence the above sequence \((\psi_{n})_{n\in\mathbb N}\subseteq\Sigma\) has a subsequence which we denote again by \((\psi_{n})_{n\in\mathbb N}\subseteq\Sigma\), and which converges uniformly on compact subsets of \(\Omega\) satisfying \[\lim _{n \rightarrow \infty}\left|\psi_{n}^{\prime}\left(z_{0}\right)\right|=\eta>0.\]

  • Let \[\lim _{n \rightarrow \infty} \psi_{n}(z)=h(z)\quad \text { for } \quad z \in \Omega\] converge uniformly on compact subsets of \(\Omega\). Then \(h \in H(\Omega)\).

  • Further \[\lim _{n \rightarrow \infty} \psi_{n}^{\prime}(z)=h^{\prime}(z)\quad \text { for }\quad z \in \Omega\] converges uniformly on compact subsets \(\Omega\).

  • Therefore \[\lim _{n \rightarrow \infty}\left|\psi_{n}^{\prime}\left(z_{0}\right)\right|=\left|h^{\prime}\left(z_{0}\right)\right|,\] which implies \[\left|h^{\prime}\left(z_{0}\right)\right|=\eta>0.\]

  • Hence \(h\) cannot be constant, otherwise we would have \(\eta=0\).

  • Therefore \(|h(z)|<1\) for \(z \in \Omega\) by the maximum modulus principle, in other words \[h(\Omega)\subseteq D.\]

  • By the Hurwitz lemma \(h\) must be also one-to-one, thus \(h\in\Sigma\) and is the maximizer for (**) as desired.

  • Hence \(h(\Omega)=D\) as it was explained by combining (*) and (**) at the beginning of the proof.

  • This completes the proof of the Riemann mapping theorem. $$\tag*{$\blacksquare$}$$

Top