Skip to main content
Log in

Uniformization of Riemann surfaces revisited

  • Published:
Annals of Global Analysis and Geometry Aims and scope Submit manuscript

Abstract

We give an elementary and self-contained proof of the uniformization theorem for noncompact simply connected Riemann surfaces.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Data availability

Not applicable.

References

  1. Albin, P., Rochon, F.: A local families index formula for \(\overline{\partial }\)-operators on punctured Riemann surfaces. Commun. Math. Phys. 289(2), 483–527 (2009)

    Article  MathSciNet  Google Scholar 

  2. Albin, P., Aldana, C.L., Rochon, F.: Ricci flow and the determinant of the Laplacian on non-compact surfaces. Commun. Partial Differ. Equ. 38(4), 711–749 (2013)

    Article  MathSciNet  Google Scholar 

  3. Carathéodory, C.: Über die gegenseitige Beziehung der Ränder bei der konformen Abbildung des Inneren einer Jordanschen Kurve auf einen Kreis. Math. Ann. 73, 305–320 (1913)

    Article  MathSciNet  Google Scholar 

  4. Cartan, H.: Théorie élémentaire des fonctions analytiques d’une ou plusieurs variables complexes. Hermann, Paris (1985)

  5. Donaldson, S.: Riemann Surfaces. Oxford University Press, Oxford (2011)

    Book  Google Scholar 

  6. de Saint-Gervais, H.P.: Uniformization of Riemann Surfaces: Revisiting a Hundred-Year-old Theorem. European Mathematical Society, Zürich (2016)

    Book  Google Scholar 

  7. Farkas, H.M., Kra, I.: Riemann Surfaces. Springer, Berlin (1980)

    Book  Google Scholar 

  8. Forster, O.: Lectures on Riemann Surfaces. Springer, Berlin (1999)

    Google Scholar 

  9. Hubbard, J.H.: Teichmüller theory and Applications to Geometry, Topology and Dynamics, vol. 1, Matrix Editions, Ithaca NY (2006)

  10. Koebe, P.: Über die Uniformisierung beliebiger analytischer Kurven, Nachr. Ges. Wiss. Göttingen (1907), 191–210 and 633–649

  11. McCleary, J.: A first course in topology: continuity and dimension. American Mathematical Society, Providence (2006)

    Book  Google Scholar 

  12. Milnor, J.W.: Topology from a Differentiable Viewpoint. The University Press of Virginia, Virginia (1965)

    MATH  Google Scholar 

  13. Osgood, W.F.: On the existence of the Green’s function for the most general simply connected plane region. Trans. AMS 1, 310–314 (1900)

    MathSciNet  MATH  Google Scholar 

  14. Popescu-Pampu,P.: What is the genus?, Lecture Notes in Mathematics 2162, Springer (2016)

  15. Perron, O.: Eine neue Behandlung der ersten Randwertaufgabe für \(\Delta \)u=0. Math. Z. 18, 42–54 (1923)

    Article  MathSciNet  Google Scholar 

  16. Poincaré, H.: Sur l’uniformisation des fonctions analytiques. Acta Math. 31, 1–64 (1907)

    Article  MathSciNet  Google Scholar 

  17. Radó, T.: Über den Begriff der Riemannschen Fläche. Acta Szeged 2, 101–121 (1925)

    MATH  Google Scholar 

  18. Riemann, B.: Grundlagen für eine allgemeine Theorie der Functionen einer verändlicher komplexer Grösse, Inauguraldissertation, Göttingen (1851)

  19. Stein, E., Shakarchi, R.: Complex Analysis. Princeton University Press, New Jersey (2003)

    MATH  Google Scholar 

  20. Tu, L.W.: An Introduction to Manifolds. Springer, Berlin (2011)

    Book  Google Scholar 

Download references

Acknowledgements

We would like to thank Sergiu Moroianu for useful discussions. The authors were partially supported from the project PN-III-P4-ID-PCE-2020-0794 (funding organization: Unitatea Executivă pentru Finanţarea Învăţământului Superior, a Cercetării, Dezvoltării şi Inovării) and from a Bitdefender Junior Researcher Fellowship.

Funding

Both authors were partially supported from the project: PN-III-P4-ID-PCE-2020-0794 and from a Bitdefender Junior Researcher Fellowship.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Rareş Stan.

Ethics declarations

Conflicts of interest

Not applicable.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A Proof of Perron’s principle

Since the family \(\mathcal {F}\) is locally bounded above, we can define \(u:X \longrightarrow \mathbb {R}\), \(u=\sup _{f\in \mathcal {F}} f.\) Our aim is to show that u is harmonic. Since harmonicity is a local property, it suffices to prove that u is harmonic on a disk \(D\subset X\). Let \(A=\{ z_0, z_1,... \} \subset D\) be a dense subset. For each j, there exists a sequence \((v_{jk})_{k \in \mathbb N} \subset \mathcal F\) such that \(u(z_j)=\lim _{k\rightarrow \infty } v_{jk}(z_j)\). Since \(\mathcal F\) is a Perron family, the map \(h_1=v_{11}^{(D)}\) belongs to \(\mathcal F\), is harmonic on D, and \(h_1 \ge v_{11}\). Suppose we constructed the functions \(\{h_1,...,h_n\} \subset \mathcal F\) harmonic on D such that \(h_k \ge h_{k-1}\) and \(h_k \ge v_{ij}\) for all \(k\in \{1,...,n\}\), \(i,j\in \{1,...,k\}\). Pick \(h_{n+1}=\max (v_{11},v_{12},...,h_n)^{(D)} \in \mathcal F\) harmonic on D, with \(h_{n+1} \ge h_{n}\) and \(h_{n+1} \ge v_{ij}\) for any \(i,j \in \{1,...,n+1\}\). Then \(h_n(z_j) \ge v_{jk}(z_j)\) for all \(n \ge k \ge j\). Letting \(k\rightarrow \infty \) in the previous inequality, we get:

$$\begin{aligned}\lim _{n\rightarrow \infty }h_n(z_j)=u(z_j).\end{aligned}$$

Since \((h_n)_{n \in \mathbb N}\) is an increasing sequence, we can consider the Lebesgue measurable function

$$\begin{aligned} w:D\longrightarrow \mathbb {R},&w=\lim _{n\rightarrow \infty } h_n\le \sup _{g\in \mathcal {F}}g_{\vert _{D}}=u_{\vert _D}. \end{aligned}$$

By the dominated convergence theorem, for every disk \(D'\) centered at a point \(x\in D\) we have

$$\begin{aligned} w(x)=\lim _{n\rightarrow \infty } h_n(x)=\lim _{n\rightarrow \infty } \frac{1}{\pi }\int _{D'} h_n(z)\mathrm{d}z=\frac{1}{\pi }\int _{D'} \lim _{n\rightarrow \infty } h_n(z)\mathrm{d}z=\frac{1}{\pi }\int _{D'} w(z)\mathrm{d}z. \end{aligned}$$

Here we used the mean property on disks for harmonic functions, which is an immediate consequence of the definition. This implies easily that w is \(C^0\). Again by dominated convergence,

$$\begin{aligned} \frac{1}{2\pi } \int _{0}^{2\pi } w(e^{it}) \mathrm{d}t= & {} \frac{1}{2\pi } \int _{0}^{2\pi } \lim _{n\rightarrow \infty } h_n(e^{it}) \mathrm{d}t= \lim _{n\rightarrow \infty } \frac{1}{2\pi }\int _{0}^{2\pi } h_n(e^{it}) \mathrm{d}t\\= & {} \lim _{n\rightarrow \infty } h_n(x)=w(x) \end{aligned}$$

thus showing that w is harmonic. We claim that \(w=u\).

Since \(h_n \in \mathcal F\) for any \(n \in \mathbb N\), then \(h_n \le u\), so \(w \le u\). But \(w(z_j)=u(z_j)\) for all \(j \in \mathbb N\), hence \(w \ge v\) on the dense subset \(A \subset D\) for every \(v \in \mathcal F\). Since \(\mathcal F\) is a family of continuous functions, it follows that \(w \ge v\) for any \(v \in \mathcal F\), therefore \(w \ge u\), which ends the proof.

Appendix B Proof of Dirichlet’s principle

If f is constant, the conclusion is clear, otherwise let \(m<M\) be the infimum, respectively, the supremum of f on \(\partial Y\). Consider the family \(\mathcal F\) of continuous functions \(g:Y \longrightarrow [m,M]\) which are subharmonic on the interior of Y with \(g_{\vert _{\partial Y}} \le f\). The first condition for \(\mathcal F\) to be a Perron family is evident. Let \(g \in \mathcal F\) and \(D'\) a disk in Y. Using Remark 4, \(g^{(D')}\) is subharmonic. By the maximum principle, \(g^{(D')}\) still takes values in [mM]. Thus \(g^{(D')} \in \mathcal F\). Now \(\mathcal {F}\ne \emptyset \) since it contains the constant function m. Let \(F=\sup _{g \in \mathcal F} g\). By Perron’s principle, F is harmonic on .

We must prove that for every \(x\in \partial Y\), \(\lim _{\xi \rightarrow x}F(\xi )=f(x)\). Let us first show that \(\liminf _{\xi \rightarrow x} F(\xi )\ge f(x)\). If \(f(x)=m\), there is nothing to prove, hence let us suppose that \(f(x)>m\). We work in a disk \(D\subset X\) centered at x with local coordinate \(\xi \).

Fix \(\epsilon >0\) such that \(f(0)-\epsilon >m\). There exists \(1>\delta >0\) such that for every \(\xi \in \partial Y\) with \(|\xi |<\delta \) we have \(f(0)-\epsilon <f(\xi )\). Consider the exterior normal to \(\partial Y\) at \(x=0\). Take a circle centered at a point p on this normal, of radius r small enough such that it touches Y only in 0. Choose \(t>1\) such that \(\log \frac{1}{t-1}+f(0)-\epsilon <m\) and set \(R=rt\). By decreasing r if needed, we can assume that \(R<\delta \). Clearly, \(|p|\le |\xi -p|\) for every \(\xi \in Y\cap D\). Define

$$\begin{aligned} u:D\rightarrow [m, \infty ),&u(\xi )=\max \left( m,\log \tfrac{|p|}{|\xi -p|} +f(0)-\epsilon \right) . \end{aligned}$$

Notice that u is the maximum of two harmonic functions, hence subharmonic. When \(|\xi |\) is close to 0, \(u(\xi )=\log \tfrac{|p|}{|\xi -p|}+f(0)-\epsilon \), while for \(|\xi |\ge R\), \(u(\xi )=m\). Hence u can be continuously extended (by the constant value m) to \(Y\cup D\), and is subharmonic on . It is straightforward to check that \(u_{\vert _{Y}}\) belongs to the family \(\mathcal {F}\). Thus, for small \(|\xi |\), we get:

$$\begin{aligned} F(\xi ) \ge u(\xi )=\log \tfrac{|p|}{|\xi -p|}+f(0)-\epsilon . \end{aligned}$$

Since \(\epsilon \) was arbitrarily small, we obtain \(\liminf _{\xi \rightarrow 0} F(\xi ) \ge f(0)\). We now prove that \(\limsup _{\xi \rightarrow 0} F(\xi ) \le f(0)\). If \(f(0)=M\), this is clear. Otherwise, take \(\epsilon >0\) such that \(f(0)+\epsilon < M\). As above, consider \(t>1\) and \(R=rt\) such that \(-\log \frac{1}{t-1}+f(0)+ \epsilon >M\), and \(f(\xi )<f(0)+ \epsilon \) for \(|\xi | \le R\). Define

$$\begin{aligned} U:D\rightarrow (-\infty ,M],&U(\xi )=\min \left( M,-\log \tfrac{|p|}{|\xi -p|} + f(0)+\epsilon \right) . \end{aligned}$$

Notice that U is the minimum of two harmonic functions, thus \(-U\) is subharmonic. Also \(U(\xi )=M\) for \(|\xi |>R \), \(U(\xi )=-\log \tfrac{|p|}{|\xi -p|} + f(0)+\epsilon \) for small \(|\xi |\), \(U \ge m\), and \(U(\xi )\ge f(\xi )\) on \(\partial Y\). For every \(g\in \mathcal {F}\), the continuous function \(g-U\) is nonpositive on the boundary of the compact domain \(\{|\xi |\le R \}\cap Y\) and subharmonic in the interior. By the maximum principle it follows that on this compact set \(g\le U\), hence \(F\le U\). Thus for small \(|\xi |\),

$$\begin{aligned} F(\xi )\le -\log \tfrac{|p|}{|\xi -p|}+f(0)+\epsilon . \end{aligned}$$

Therefore \(\limsup _{\xi \rightarrow 0} F(\xi ) \le f(0)\), showing that \(\lim _{\xi \rightarrow 0} F(\xi )\) exists and equals f(0).

Appendix C Second countability in the presence of a holomorphic function

Let us prove that if there exists a nonconstant holomorphic function \(f:X\longrightarrow {\mathbb {C}}\), then the Riemann surface X is second countable.

Let \({\mathcal {B}}\) be a countable basis of topology for \({\mathbb {C}}\). Let \({\mathcal {A}}\) be the set of those connected components of each \(f^{-1}(U)\), \(U \in {\mathcal {B}}\), which are second countable. We claim that A is a basis of topology for X. Indeed, let \(D \subset X\) be an open set and \(x \in D\). By the identity theorem for holomorphic functions, \(f^{-1}\left( f(x) \right) \) is discrete, hence there exists an open neighborhood W of x, relatively compact in D, such that \(W \cap f^{-1} \left( f(x) \right) = \lbrace x \rbrace \). We have that \(f \left( \partial \overline{W} \right) \) is compact in \({\mathbb {C}}\). Since \(f(x) \notin f \left( \partial \overline{W} \right) \), there exists \(U \in {\mathcal {B}}\) which contains f(x) such that \(U \cap f \left( \partial \overline{W} \right) =\emptyset \). Let V be the connected component of \(f^{-1}(U)\) which contains x. Since \(V \cap \partial \overline{W}=\emptyset \), we get that \(V \subset W\), hence V is second countable. We found \(x \in V \subset D\), with \(V \in {\mathcal {A}}\), therefore the claim is proved. We next prove that \({\mathcal {A}}\) is countable.

Each \(V \in {\mathcal {A}}\) intersects countably many other elements in \({\mathcal {A}}\). Otherwise, there would exist \(U \in {\mathcal {B}}\) such that V intersects uncountably many connected components of \(f^{-1}(U)\). It would follow that V contains uncountably many disjoint open subsets, which is a contradiction.

Consider \(V_0 \in {\mathcal {A}}\). There exists a countable number of open sets from \({\mathcal {A}}\) which intersect \(V_0\), denote their union with \(V_0\) by \(V_1\). For \(k \ge 1\), define \(V_{k+1}\) as the union of \(V_k\) with those open sets in \({\mathcal {A}}\) which intersect \(V_k\). By induction, \(V_k\) is countable. It is clear that \(\cup _{k \ge 1} V_k\) is both open and closed, hence \(\cup _{k \ge 1} V_k=X\). Therefore all the open sets from \({\mathcal {A}}\) appear in this process, so \({\mathcal {A}}\) is a countable union of countable sets, hence countable.

Appendix D Compactness of families of injective holomorphic functions

Proof of Lemma 13

Denote by \(r_n=\sup \{ r \in \mathbb {R}: r\mathbb {D}\subset f_n({\mathbb {D}}) \}\). Let \(a_n \in \partial (r_n \overline{ \mathbb {D}}) {\setminus } f_n({\mathbb {D}})\). From the Schwarz lemma for the function \({f_n^{-1}}_{\vert _{r_n \mathbb {D}}}\), it follows that \(|a_n|=r_n \le 1\). By extracting a subsequence, we can assume that \((a_n)_{n \in \mathbb {N}}\) is convergent. Consider the function \(g_n: {\mathbb {D}} \longrightarrow \mathbb {C}\), \(g_n=a_n^{-1} f_n\). It is easy to see that \({\mathbb {D}} \subset g_n({\mathbb {D}})\) and also \(1 \notin g_n({\mathbb {D}})\). Since \(g_n\) is holomorphic and injective, it follows that \(g_n({\mathbb {D}})\) is simply connected; hence, we can construct the square root

$$\begin{aligned}&\psi _n:g_n({\mathbb {D}})\longrightarrow {\mathbb {C}}^{*}&\psi _n(z)=ie^{\frac{1}{2}\int _0^z \frac{dw}{w-1}}. \end{aligned}$$

Then \(\psi _n(0)=i\) and \(\psi _n^2(z)=z-1\) for \(z \in g_n({\mathbb {D}})\). Since \(\psi _n^2\) is injective, the image of \(\psi _n\) cannot contain pairs of the form \((w,-w)\). Thus, there exists a disk D centered at \(-i\) disjoint from the image of \(\psi _n\) for every n. Define \(h_n=\psi _n \circ g_n: {\mathbb {D}} \longrightarrow \mathbb {C}{\setminus } D\). There exists \(0<r<\infty \) so that the homography \(\alpha (z)=\frac{1}{z+i}\) maps \(\mathbb {C}{\setminus } D\) into \(r\mathbb {D}\). Using Montel’s Theorem and relabeling the sequence, we can assume that \( \alpha \circ h_{n} \) converges to a holomorphic map \(h:\mathbb {D}\longrightarrow r \overline{\mathbb {D}}\). If h is nonconstant, it is open by Sect. 2; hence, it maps \(\mathbb {D}\) to \(r\mathbb {D}\). Thus we can compose it with the inverse homography \(\alpha ^{-1}\). Otherwise, h equals the constant \(\frac{1}{2i}\). In both cases, \(h_n\) converges to \(\alpha ^{-1} \circ h\), thus \(f_{n}=a_{n}\left( 1+ h_n^2 \right) \) converges in \(\mathcal O (\mathbb {D})\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Anghel, C., Stan, R. Uniformization of Riemann surfaces revisited. Ann Glob Anal Geom 62, 603–615 (2022). https://doi.org/10.1007/s10455-022-09860-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10455-022-09860-2

Keywords

Navigation