Riemann Mapping Theorem

The Riemann mapping theorem is one of the most central results in complex analysis. However, the standard proof [1] lacks intuition. We will see an intuitive proof using ideas from topology and partial differential equations (PDEs). No background in topology or PDEs is assumed. However, I will state and use (without proof) some theorems from topology that are intuitively obvious, even though their proofs are nontrivial. I will refer to these theorems as i.o., and in the final section I will provide references for their proofs.

To the hesitant student of mathematics: it is a standard practice among mathematicians to use established i.o. theorems, without having seen the proof. This is because they want to use their time to focus on the more conceptual and creative aspects of their work.

Prerequisites. The reader is expected to be familiar with undergraduate complex analysis up to the Cauchy integral formula (e.g., Chapters 1 and 2.1–2.4 from [1]).

We begin by introducing some definitions.

Paths and loops. Let \(a>0\). A continuous function \( p:[0,a] \to \mathbb{C} \) is called a continuous path. If \(p\) is also injective, we call it simple. A continuous path \( p:[0,a] \to \mathbb{C} \) is called a loop if \( p(0) = p(a). \) A simple loop is also called a Jordan curve. Finally, regarding paths that go to infinity, we will only need the following type: we say that \(p:(0,a) \to \mathbb{C}\) is a simple continuous path that starts and ends at infinity if \(p\) is continuous and injective, and \( p(t) \to \infty \) as \( t \to 0 \) or \( t \to a \).

We will need the following i.o. theorem about Jordan curves. If you are unaware of the term connected component, you can see [6].

Theorem 1. Let \( \ell:[0,a] \rightarrow \mathbb{C} \) be a Jordan curve, and let \( L := \ell([0,a]) \) denote its image. Then the complement \( \mathbb{C} \setminus L \) consists of exactly two connected components, both of which are open. One of the components is bounded (called the interior, denoted \( A_\ell \)), and the other is unbounded (called the exterior, denoted \( E_\ell \)). The curve \( L \) is the boundary of both components. Furthermore, the interior \( A_\ell \) and its closure \(A_\ell \cup L\) are both simply connected.

Before stating the main theorem, we introduce one more concept.

Biholomorphisms. Let \(A, B\subseteq \mathbb{C}\) be open. If a holomorphism \(f:A\rightarrow B\) is a bijection, then we say that \(f\) is a biholomorphism. We will prove momentarily that if \(f\) is a biholomorphism, then its inverse is holomorphic. If there exists a biholomorphism between two open sets \(A, B\), then we say that \(A\) and \(B\) are biholomorphic.

Theorem 2 (Riemann mapping theorem). Let \( A \subseteq \mathbb{C} \) be non-empty, simply connected, and open. Suppose that \( A \) is not all of \( \mathbb{C} \). Then there exists a biholomorphism from \( A \) onto the open unit disc \( D := \{ z \in \mathbb{C} : |z| < 1 \} \).
Illustration of the Jordan curve and indicated directions

Example of biholomorphism, as described by the Riemann mapping theorem.

We will prove a slightly simplified version of Theorem 2. This is in order to focus on the main ideas, without being distracted by the technical details needed for the fully general case. In [2], the authors generalize the approach we take here, to prove Theorem 2 in its full generality.

We prove Theorem 2 under the following assumption:

Assumption. The set \(A\) is either the interior of a Jordan curve (bounded case), or it has a boundary that is a simple continuous path that starts and ends at infinity (unbounded case).
Illustration of the Jordan curve and indicated directions

A bounded and an unbounded set satisfying our assumption.

Note that not all sets \(A\) in Theorem 2 satisfy our assumption, e.g., \(\mathbb{C}\setminus \{z\in \mathbb{C}: Re(z)\ge 0\}\).

We will first do the proof for the bounded case. Then, we will prove the unbounded case by a simple reduction to the bounded one.

The bounded case

We will need some tools to attack the bounded case. The first tool is topological and complex-analytic. It is called the argument principle, and it is covered in this article of mine. Even if you are familiar with it, you will most probably benefit from that article, since the proof is fully visual. We will also need two applications of the argument principle, covered here. Make sure you have understood the material from these articles, before you continue.

The third tool is the following theorem:

Theorem 3. Let \( \ell:[0,a] \rightarrow \mathbb{C} \) be a Jordan curve, and let \( L := \ell([0,a]) \) denote its image and \(A\) its interior. Let \(g:L\rightarrow \mathbb{R}\) be a continuous function. Then, there exists a unique function \(u: A \cup L \rightarrow \ \mathbb{R}\) that is continuous on \(A \cup L\), harmonic on \(A\), and equal to \(g\) on \(L\).

In PDE terms, this theorem says that the Dirichlet problem for the Laplace equation, on a domain bounded by a Jordan curve, has a unique solution. Theorem 3 has a very intuitive proof using probability theory. Even though the full proof requires an advanced concept called Brownian motion, the core idea can be described in elementary terms — and the solution \( u \) can actually be constructed in a fully elementary way! I suggest you accept Theorem 3 as a fact for now and continue with the rest of the article. When you're done, if you're interested in understanding how Theorem 3 is proved, then:

Now, let's move on with the proof of the bounded case.

Without loss of generality, we assume that \(0\in A\) (otherwise we can just translate \(A\)). We will construct a biholomorphism that sends zero to zero. First, we need a definition. Let \(\bar{S}\) denote the closure of a set \(S\).

Definition. Let \( A, B \subseteq \mathbb{C} \) open sets. A function \( f : \bar{A} \to \bar{B} \) is a strong biholomorphism if
  1. It is a continuous bijection.
  2. The restriction \( f|_A \) is a biholomorphism sending \( A \) onto \( B \).

The first thing we will do is deduce some properties satisfied by all strong biholomorphisms \(f:\bar{A}\rightarrow \bar{D}\) with \(f(0)=0\). These properties will give us insight into how to actually construct a biholomorphism from \(A\) onto \(D\) satisfying \(f(0)=0\). This function will in fact be a strong biholomorphism, but we will not prove this here, as it is not needed for our theorem.

Claim 1. Let \(f:\bar{A}\rightarrow \bar{D}\) be a strong biholomorphism with \(f(0)=0\). Then, there exists a holomorphic function \(H:A\rightarrow \mathbb{C}\) such that \(f(z)=ze^{H(z)}\), for all \(z\in A\). Furthermore, the real part \(\text{Re}(H)\) can be continuously extended to the closure \(\bar{A}\).

Again, it is possible to show that \(\text{Im}(H)\) can also be continuously extended to \(\bar{A}\), which implies that the identity \(f(z)=ze^{H(z)}\) holds on \(\bar{A}\). However, we won't need this fact and so we won't prove it here.

Proof

Let \( g : \bar{A} \to \mathbb{C} \) be defined as

\[ g(z) := \begin{cases} f(z)/z, & z \in \bar{A} \setminus \{0\} \\ f'(0), & z = 0 \end{cases} \]

I claim that \( g \) is holomorphic on \( A \) and continuous on \( \bar{A} \). Indeed, locally around zero,

\[ f(z) = f'(0)z + \frac{f''(0)}{2}z^2 + \dots \]

Thus, locally around zero,

\[ g(z) = f'(0) + \frac{f''(0)}{2}z + \dots \]

So, \( g \) is holomorphic at zero, and clearly it is also holomorphic everywhere else on \( A \). The continuity of \( g \) at points of \( \partial A \) follows from the continuity of \( f \).

Now, from Theorem 2 in this article, we know that \( f' \neq 0 \) on \( A \). Furthermore, since \( f \) is injective, we have \( g \neq 0 \) on \( \bar{A} \). This implies that there exists a holomorphic function \( H : A \to \mathbb{C} \) such that \( g(z) = e^{H(z)}, \) for all \( z \in A \). For a proof of this implication, you can see [4].

Let \( u(z) \) and \( v(z) \) be the real and imaginary parts of \( H \). Since \( g \) is continuous on \( \bar{A} \) and \( u(z) = \ln(|g(z)|) \), the function \( u(z) \) can be continuously extended on \( \bar{A} \). \(\ \square\)

Continuing on the setting of Claim 1, we have that on the set \(A\), \(f(z)=ze^{u(z)+iv(z)}\), and so

\[ |f(z)|=|z|e^{u(z)} \]

for all \(z\in A\). Since \(f\) and \(u\) are continuous on \(\bar{A}\), the above equality extends on \(\partial A\). Now, the crucial moment: since \(f\) is a strong biholomorphism, we have \(f(\partial A)=\partial D\). This property actually tells us how to construct the function \(u\). Indeed, since \(|f(z)|=1\) on \(\partial A\), we have that \(u(z)=-\log |z| \), for all \(z\in \partial A\). But \(u\) is the real part of a holomorphic function on \(A\), and thus it is harmonic on \(A\). This means that we can construct \(u\) by solving this Dirichlet problem!

Construction of the imaginary part

What about \(v\)? We can construct it by using \(u\)! Remember that these are coupled by the Cauchy-Riemann equations: \(u_x=v_y\) and \(u_y=-v_x\). Geometrically, this means that \(\nabla v\) is \(\nabla u\) rotated by \(90^\circ\) counter-clockwise. Since we have \(\nabla v\), we can get \(v\) up to an additive constant:

\[ v(x,y)=\int_\gamma \nabla v(s)\cdot ds + v(0) \]
where the above integral is the line integral of \(\nabla v\) along any piecewise-smooth curve \(\gamma\) that starts at zero, ends at \((x,y)\), and lies entirely inside \(A\).

Thus, as long as there exists a strong biholomorphism \(f:\bar{A}\rightarrow \bar{D}\) with \(f(0)=0\), this function \(f\) is uniquely determined up to a rotation (expressed by the constant \(v(0)\)).

The biholomorphism

We now use these insights to construct a biholomorphism from \( A \) onto \( D \). Let \( u \) be the unique function that is harmonic on \( A \), continuous on \( \bar{A} \), and equals \( -\log |z| \) on \( \partial A \). Let

\[ v(x, y) := \int_\gamma (-u_y(s), u_x(s)) \cdot ds \]

where \( \gamma \) is any piecewise-smooth curve that starts at the origin, ends at \( (x, y) \), and lies entirely inside \( A \). We will prove that the function \( f : A \to \mathbb{C} \) defined as

\[ f(z) = z e^{u(z) + i v(z)} \]

is holomorphic, sends \( A \) onto \(D \), and is injective. The fact that \( f \) is holomorphic is immediate, since \( u \) and \( v \) satisfy the Cauchy-Riemann equations (why?). We now prove the rest.

The boundary has all the information

To highlight the main idea, we first do the proof under the assumption that \( f \) can be continuously extended on \( \bar{A} \). We then drop this assumption by tweaking our argument. As I mentioned in this article, whenever I say "argument principle", I mean its general version, given in Corollary 1 of that article.

Let \( \ell : [0,a] \to \mathbb{C} \) be a positively oriented Jordan curve that parametrizes \( \partial A \). By the way we constructed \( u \), we have \( f(\partial A) \subseteq \partial D \), and so the loop \( f \circ \ell \) lies entirely on the unit circle. This implies that \( |f(z)| \leq 1 \) for all \( z \in A \). Indeed, if \( |f(z_0)| > 1 \) for some \( z_0 \in A \), then by the argument principle \( W(f \circ \ell, f(z_0)) \geq 1 \), a contradiction. Since \( f \) is an open map , we have \( f(A) \subseteq D \).

We need to show that \( f \) is onto and injective. We will see that these are also consequences of the argument principle. What is \( W(f \circ \ell, 0) \)? By the argument principle, it equals the number of roots of \( f \) in \( A \), counting multiplicities. Clearly, zero is the only root of \( f \). Combining this with

\[ f'(0) = \lim_{z \to 0} \frac{f(z)}{z} = e^{u(0)} > 0 \]

we have \( W(f \circ \ell, 0) = 1 \). By Proposition 1 in this article, we have

\[ W(f \circ \ell, q) = 1 \]

for all \( q \in D \). By the argument principle, for each \( q \in D \), there exists exactly one \( z \in A \) such that \( f(z) = q \). \( \ \square\)

Dropping the continuity assumption

Even if we don't assume that \(f\) can be continuously extended on \( \bar{A} \), we can still use the continuity of \(u\) on the boundary! We first prove the following intuitively obvious proposition.
Proposition 1. For all \(\delta>0\), there exists a Jordan curve \(\ell_\delta:[0,a]\rightarrow \mathbb{C}\) that lies inside \(A,\) and is such that \(|\ell_\delta (t)-\ell(t)|<\delta\), for all \(t\in [0,a]\).

Let \( z_0 \in A \), and consider a positively oriented circle \( C_{z_0, \varepsilon}(t) = \varepsilon e^{i 2\pi t} + z_0 \), \( t \in [0,1] \), that lies inside \(A\). By Theorem 2 from this article, there exists a function \( \gamma : [0,1] \times [0,1] \rightarrow \mathbb{C} \) such that for all \(t\in[0,1]\), \(\gamma(0,t) = C_{z_0, \varepsilon}(t),\ \gamma(1,t) = \ell(at),\) and for all \( s \in (0,1) \), the curve \( \gamma(s, \cdot) \) is a Jordan curve lying inside \( A \setminus \bar{D}_{z_0, \varepsilon} \). Remember that \( \bar{D}_{z_0, \varepsilon} \) denotes the closed disc centered at \( z_0 \) with radius \( \varepsilon \).

Since \( [0,1] \times [0,1] \) is compact, the continuous function \( \gamma \) is in fact uniformly continuous. Thus, for any \( \delta > 0 \), there exists an \( \eta \in (0,1) \) such that

\[ |\gamma(1 - \eta, t) - \gamma(1, t)| < \delta \]
for all \( t \in [0,1] \). By setting \( \ell_\delta(t) := \gamma(1 - \eta, t/a) \) for \( t \in [0,a] \), we are done. \(\ \square \)

We now explain why for all small enough \(\delta>0\), the loop \(f\circ \ell_\delta\) is very close to the unit circle. Since \(u\) is continuous on the boundary, we can continuously extend \(|f|\) on \(\bar{A}\). Also, since \(\bar{A}\) is compact, \(|f|\) is uniformly continuous on \(\bar{A}\). This implies that for any \(\varepsilon>0\), there is a \(\delta_0>0\), such that for all \(\delta\in (0,\delta_0)\),

\[ 1-\varepsilon < |f\circ \ell_\delta (t)|<1+\varepsilon \]

Using this fact, we can immediately adapt the proof we did under the continuity assumption. This adaptation is left as an exercise for the reader.

The unbounded case

Let \( A \) be an unbounded, simply connected, and open subset of \( \mathbb{C} \) that is not the whole \( \mathbb{C} \). Suppose the boundary of \( A \) is a simple continuous path that starts and ends at infinity. We will construct a simple biholomorphism from \( A \) to a bounded, simply connected, and open subset of \( \mathbb{C} \), whose boundary is a Jordan curve. We proved that such a set is biholomorphic to \( D \). Thus, we will be able to conclude that \( A \) is biholomorphic to \( D \).

We will use the following i.o. theorem.

Theorem 4. Let \( p : (0, a) \to \mathbb{C} \) be a simple continuous path that starts and ends at infinity. Let \( A \) be an unbounded, simply connected, and open set with boundary \( P := p((0, a)) \). Then, \( \mathbb{C} \setminus P \) has exactly two connected components: \( A \) and another set \(B\). The set \( B \) is also unbounded, simply connected, and open.

We apply this theorem to our setting. Without loss of generality, we assume that \(0\in B\) (otherwise, we can translate \(A\)). Let \(g(z):=1/z\). By applying \(g\) to the path \(p\), we get a Jordan curve. This is formally expressed by defining \( \tilde{p} : [0, a] \to \mathbb{C} \) as follows:

\[ \tilde{p}(t) := \begin{cases} g(p(t)), & t \in (0, a) \\ 0, & t = 0 \text{ or } t = a \end{cases} \]

Observe that \(\tilde{p}\) is indeed a Jordan curve. By proving the following claim, we complete the proof of the unbounded case.

Claim 2. The set \(g(A)\) is the interior of the Jordan curve \(\tilde{p}\).
Remember that continuous functions send connected sets to connected sets. Also, holomorphic functions send open sets to open sets. Thus, \(g(A)\) and \(g(B\setminus\{0\})\) are both open and connected. They are also disjoint. Furthermore, observe that their union is \(\mathbb{C}\setminus \tilde{P}\), where \( \tilde{P}:=\tilde{p}([0,a]) \). Thus, \(g(A)\) and \(g(B\setminus\{0\})\) are the two connected components of \(\mathbb{C}\setminus \tilde{P}\) (why?). Using Theorem 1 and the fact that \(g(B\setminus\{0\})\) is unbounded, we are done. \(\ \square\)

References for the i.o. theorems

For completeness, I provide references for proving the i.o. theorems used in this article, and also in the one about the argument principle. As I said in the beginning, even though these theorems are intuitively obvious, their proofs are nontrivial. Several of them follow from a central theorem in topology called Jordan–Schoenflies (JS) theorem [7].

This article. Theorem 1 follows from JS (even though the first part is the simplest Jordan curve theorem). Theorem 4 follows from Theorem 1, by applying the inversion idea that we used for the unbounded case.

Argument principle article. Theorem 1 is proved using a topological concept called covering maps (see Chapter 9 of [8]). Theorem 4 also follows immediately from the techniques of that chapter, and it is the simplest example of a topological fact called Hopf’s Degree Theorem that is valid in any number of dimensions. Now, about Theorem 2, its first part is proven in Section 65 of [8], and the second part follows from the Annulus theorem [9], which is a consequence of JS. Finally, Theorem 5 follows from JS.

References

  1. Complex Analysis, Elias M. Stein and Rami Shakarchi.
  2. The Riemann Mapping Theorem from Riemann's Viewpoint, by Robert Greene and Kang-Tae Kim. Note: the authors use Perron's method for the Dirichlet problem. However, the probabilistic solution that we employ here works equally well, as it is shown in [3].
  3. Probabilistic techniques in Analysis, by Richard F. Bass. Chapter 2.1.
  4. Complex Analysis, Elias M. Stein and Rami Shakarchi. Theorem 6.2, Chapter 3. Even though the theorem is in Chapter 3, a reader having the prerequisites for this article can understand it.
  5. Probability: Theory and Examples, Rick Durrett. Chapter 9.5. You will also need a proof that all the points of a Jordan curve are regular. This is proved in [3], Proposition 1.14.
  6. Complex Analysis, Elias M. Stein and Rami Shakarchi. Chapter 1, exercise 6.a.
  7. An Elementary Proof of the Jordan-Schoenflies Theorem, Stewart S. Cairns.
  8. Topology, James Munkres (second edition).
  9. Knots and Links, Dale Rolfsen, p.11.