Holomorphic Functions

The power of complex analysis doesn’t come from complex-differentiability at a point (though that is certainly not a trivial property for a function). Its power comes from complex-differentiability in a neighborhood around that point.

Definition 1. Let f : \mathbb C \to \mathbb C be differentiable at some point z_0 \in \mathbb C.

  • f is holomorphic at z_0 if there exists a neighborhood U of z_0 such that f is complex-differentiable on U.
  • f is holomorphic on an open set U if it is holomorphic at every point z_0 \in U.
  • f is entire if it is holomorphic on \mathbb C.

For simplicity, call a nonempty open subset D \subseteq \mathbb C \cong \mathbb R^2 a domain if and only if it is (path-)connected. Here, a path from z_1 to z_2 is the image \gamma([0,1]) of a continuous map \gamma : [0, 1] \to \mathbb C such that \gamma(0) = z_1 and \gamma(1) = z_1.

Let f : \mathbb C \to \mathbb C is analytic on a domain D containing z_0.

Theorem 1. We have f' = 0 on D if and only if f = f(z_0) on D.

Proof. For the direction (\Rightarrow), z_0 = x_0 + i y_0 and fix z = x + iy \in D such that there exists a con line

L_{z_0, z}:= \{ (1-t) \cdot z_0 + t \cdot z : t \in [0, 1]\}.

belong to D. If f' = 0, then

u_x = u_y = v_x = v_y = 0.

Observe that the function \gamma : [0, 1] \to \mathbb C defined by

\gamma(t) := f((1-t) \cdot z_0 + t \cdot z)

is differentiable on L. By the mean value theorem, there exists c \in [0, 1] such that

\gamma(1) = \gamma(0) + \gamma'(c) \cdot (z - z_0).

However,

\gamma'(c) = f'((1-c) \cdot z_0 + c \cdot z) \cdot (z - z_0) = 0 \cdot (z-z_0) = 0,

so that f(z) = \gamma(1) = \gamma(0) = f(z_0). For arbitrary z \in D, connect it to z_0 using the polygonal lines L_{z_0,z_1}, L_{z_1,z_2}, \dots, L_{z_{n-1},z_n}, L_{z_n,z}. By the previous argument,

\gamma(z) = \gamma(z_n) = \cdots = \gamma(z_1) = \gamma(z_0).

The direction (\Leftarrow) is obvious.

Analytic functions also play a crucial role in harmonic analysis. But first, define second-order partial derivatives by u_{x_i x_j} := (u_{x_i})_{x_j} whenever they exist.

Theorem 2. If the functions u = \mathbf f_1, v = \mathbf f_2 have continuous first- and second-order partial derivatives, then they are harmonic in D in that they satisfy the following Laplace equations:

\left\{ \begin{aligned} u_{xx} + u_{yy} &= 0, \\ v_{xx} + v_{yy} &= 0. \end{aligned} \right.

Proof. Since f is analytic, its component functions satisfy the Cauchy-Riemann equations:

u_x = v_y,\quad u_y = -v_x.

Taking necessary derivatives,

u_{xx} = v_{yx},\quad u_{yy} = -v_{xy},\quad v_{xx} = -u_{yx},\quad v_{yy} = u_{xy}.

Thus, we prove the theorem if we can prove that u_{xy} = u_{yx}. This we claim holds by Clairaut’s theorem below.

Theorem 3 (Clairaut’s Theorem). If u has continuous first- and second-order partial derivatives on a domain D, then u_{xy} = u_{yx}.

Proof. Suppose without loss of generality that D is an open square containing some (x_0, y_0) and (x_0 \pm \epsilon, y_0 \pm \epsilon). Define the Fréchet-differentiable functions f,g, h : (-\epsilon, \epsilon)^2 \to D by

\begin{aligned}f(s, t) &:= u(x_0 + s, y_0 + t) - u(x_0 + s, y_0), \\ g(s, t) &:= u(x_0 + s, y_0 + t) - u(x_0 , y_0+t), \\ h(s, t) &:= u(x_0 + s, y_0 + t) - u(x_0 + s, y_0) - u(x_0 , y_0+t) + u(x_0, y_0). \end{aligned}

Now fix (s, t) \in (-\epsilon, \epsilon)^2 such that s \neq 0, t \neq 0. Use the mean value theorem to obtain \xi_1,\xi_2 \in (0, 1) such that

\begin{aligned} h(s,t) &= f(s,t) - f(0, t) = f_x(\xi_1 s, t) \cdot s  \\ &=  u_{xy}(x_0 + \xi_1 s, y_0 + \xi_2 t) \cdot s \cdot t\end{aligned}

Similarly, use the mean value theorem to obtain \xi_3,\xi_4 \in (0, 1) such that

\begin{aligned} h(s,t) &= g_y(s, \xi_3 t) \cdot t = u_{yx}(x_0 + \xi_4 s, y_0 + \xi_3 t) \cdot s \cdot t\end{aligned}

Hence,

u_{xy}(x_0 + \xi_1 s, y_0 + \xi_2 t) = u_{yx}(x_0 + \xi_4 s, y_0 + \xi_3 t) .

Taking (s, t) \to (0, 0), and applying continuity, u_{xy}(x_0, y_0) = u_{yx}(x_0, y_0). Since (x_0, y_0) is arbitary, we obtain u_{xy} = u_{yx}, as required.

The converse of Theorem 2 holds as well.

Corollary 1. Let f : \mathbb C \to \mathbb C. If u,v are harmonic functions on D that satisfy the Cauchy-Riemann equations, then f is holomorphic on D.

And just like with complex-differentiability, we obtain a family of holomorphic functions.

Theorem 4. Suppose f, g : \mathbb C \to \mathbb C are holomorphic on a domain D. Then f \pm g, f \cdot g are holomorphic on D. Furthermore, if g \neq 0 on D, then f/g is holomorphic on D. Finally, if g is holomorphic on f(D), then g \circ f is holomorphic on D, yielding the usual product, quotient, and chain rules in real-variable calculus.

Particularising to D = \mathbb C, we obtain families of entire functions. Polynomials are our canonical examples, but the exponential is really the star of complex analysis.

Theorem 5. The exponential function \exp : \mathbb C \to \mathbb C characterised by

\exp(x +iy ) = e^{x} \cdot (\cos(y) + i \sin(y))

is entire.

Proof. Writing u = e^x \cdot \cos(y) and v = e^x \cdot \sin(y), it is clear that

u_x = v_y,\quad u_y = -v_x.

Furthermore,

u_{xx} + u_{yy} = u + (-u) = 0,\quad v_{xx} + v_{yy} = v + (-v) = 0.

By Corollary 1, \exp is holomorphic on \mathbb C. Therefore, \exp is entire.

Corollary 2. The complex trigonometric functions \cos, \sin defined by

\displaystyle \cos(z) := \frac{e^{iz} + e^{-iz}}{2},\quad \sin(z) := \frac{e^{iz} - e^{-iz}}{2i}

are entire.

What about the logarithm? This is where we will need to take a pitstop and re-evaluate our life decisions. Or rather, approach it with a log more caution than the other functions so far.

—Joel Kindiak, 11 Aug 25, 2346H

,

Published by


Leave a comment