--- title: "Holomorphic function" sort_title: "Holomorphic function" date: 2021-02-25 categories: - Mathematics - Complex analysis layout: "concept" --- In complex analysis, a complex function $$f(z)$$ of a complex variable $$z$$ is called **holomorphic** or **analytic** if it is complex differentiable in the neighbourhood of every point of its domain. This is a very strong condition. As a result, holomorphic functions are infinitely differentiable and equal their Taylor expansion at every point. In physicists' terms, they are extremely "well-behaved" throughout their domain. More formally, a given function $$f(z)$$ is holomorphic in a certain region if the following limit exists for all $$z$$ in that region, and for all directions of $$\Delta z$$: $$\begin{aligned} \boxed{ f'(z) = \lim_{\Delta z \to 0} \frac{f(z + \Delta z) - f(z)}{\Delta z} } \end{aligned}$$ We decompose $$f$$ into the real functions $$u$$ and $$v$$ of real variables $$x$$ and $$y$$: $$\begin{aligned} f(z) = f(x + i y) = u(x, y) + i v(x, y) \end{aligned}$$ Since we are free to choose the direction of $$\Delta z$$, we choose $$\Delta x$$ and $$\Delta y$$: $$\begin{aligned} f'(z) &= \lim_{\Delta x \to 0} \frac{f(z + \Delta x) - f(z)}{\Delta x} = \pdv{u}{x} + i \pdv{v}{x} \\ &= \lim_{\Delta y \to 0} \frac{f(z + i \Delta y) - f(z)}{i \Delta y} = \pdv{v}{y} - i \pdv{u}{y} \end{aligned}$$ For $$f(z)$$ to be holomorphic, these two results must be equivalent. Because $$u$$ and $$v$$ are real by definition, we thus arrive at the **Cauchy-Riemann equations**: $$\begin{aligned} \boxed{ \pdv{u}{x} = \pdv{v}{y} \qquad \pdv{v}{x} = - \pdv{u}{y} } \end{aligned}$$ Therefore, a given function $$f(z)$$ is holomorphic if and only if its real and imaginary parts satisfy these equations. This gives an idea of how strict the criteria are to qualify as holomorphic. ## Integration formulas Holomorphic functions satisfy **Cauchy's integral theorem**, which states that the integral of $$f(z)$$ over any closed curve $$C$$ in the complex plane is zero, provided that $$f(z)$$ is holomorphic for all $$z$$ in the area enclosed by $$C$$: $$\begin{aligned} \boxed{ \oint_C f(z) \dd{z} = 0 } \end{aligned}$$ {% include proof/start.html id="proof-int-theorem" -%} Just like before, we decompose $$f(z)$$ into its real and imaginary parts: $$\begin{aligned} \oint_C f(z) \dd{z} &= \oint_C (u + i v) \dd{(x + i y)} = \oint_C (u + i v) \:(\dd{x} + i \dd{y}) \\ &= \oint_C u \dd{x} - v \dd{y} + i \oint_C v \dd{x} + u \dd{y} \end{aligned}$$ Using Green's theorem, we integrate over the area $$A$$ enclosed by $$C$$: $$\begin{aligned} \oint_C f(z) \dd{z} &= - \iint_A \pdv{v}{x} + \pdv{u}{y} \dd{x} \dd{y} + i \iint_A \pdv{u}{x} - \pdv{v}{y} \dd{x} \dd{y} \end{aligned}$$ Since $$f(z)$$ is holomorphic, $$u$$ and $$v$$ satisfy the Cauchy-Riemann equations, such that the integrands disappear and the final result is zero. {% include proof/end.html id="proof-int-theorem" %} An interesting consequence is **Cauchy's integral formula**, which states that the value of $$f(z)$$ at an arbitrary point $$z_0$$ is determined by its values on an arbitrary contour $$C$$ around $$z_0$$: $$\begin{aligned} \boxed{ f(z_0) = \frac{1}{2 \pi i} \oint_C \frac{f(z)}{z - z_0} \dd{z} } \end{aligned}$$ {% include proof/start.html id="proof-int-formula" -%} Thanks to the integral theorem, we know that the shape and size of $$C$$ is irrelevant. Therefore we choose it to be a circle with radius $$r$$, such that the integration variable becomes $$z = z_0 + r e^{i \theta}$$. Then we integrate by substitution: $$\begin{aligned} \frac{1}{2 \pi i} \oint_C \frac{f(z)}{z - z_0} \dd{z} &= \frac{1}{2 \pi i} \int_0^{2 \pi} f(z) \frac{i r e^{i \theta}}{r e^{i \theta}} \dd{\theta} = \frac{1}{2 \pi} \int_0^{2 \pi} f(z_0 + r e^{i \theta}) \dd{\theta} \end{aligned}$$ We may choose an arbitrarily small radius $$r$$, such that the contour approaches $$z_0$$: $$\begin{aligned} \lim_{r \to 0}\:\: \frac{1}{2 \pi} \int_0^{2 \pi} f(z_0 + r e^{i \theta}) \dd{\theta} &= \frac{f(z_0)}{2 \pi} \int_0^{2 \pi} \dd{\theta} = f(z_0) \end{aligned}$$ {% include proof/end.html id="proof-int-formula" %} Similarly, **Cauchy's differentiation formula**, or **Cauchy's integral formula for derivatives** gives all derivatives of a holomorphic function as follows, and also guarantees their existence: $$\begin{aligned} \boxed{ f^{(n)}(z_0) = \frac{n!}{2 \pi i} \oint_C \frac{f(z)}{(z - z_0)^{n + 1}} \dd{z} } \end{aligned}$$ {% include proof/start.html id="proof-dv-formula" -%} By definition, the first derivative $$f'(z)$$ of a holomorphic function exists and is: $$\begin{aligned} f'(z_0) = \lim_{z \to z_0} \frac{f(z) - f(z_0)}{z - z_0} \end{aligned}$$ We evaluate the numerator using Cauchy's integral theorem as follows: $$\begin{aligned} f'(z_0) &= \lim_{z \to z_0} \frac{1}{z - z_0} \bigg( \frac{1}{2 \pi i} \oint_C \frac{f(\zeta)}{\zeta - z} \dd{\zeta} - \frac{1}{2 \pi i} \oint_C \frac{f(\zeta)}{\zeta - z_0} \dd{\zeta} \bigg) \\ &= \frac{1}{2 \pi i} \lim_{z \to z_0} \frac{1}{z - z_0} \oint_C \frac{f(\zeta)}{\zeta - z} - \frac{f(\zeta)}{\zeta - z_0} \dd{\zeta} \\ &= \frac{1}{2 \pi i} \lim_{z \to z_0} \frac{1}{z - z_0} \oint_C \frac{f(\zeta) (z - z_0)}{(\zeta - z)(\zeta - z_0)} \dd{\zeta} \end{aligned}$$ This contour integral converges uniformly, so we may apply the limit on the inside: $$\begin{aligned} f'(z_0) &= \frac{1}{2 \pi i} \oint_C \Big( \lim_{z \to z_0} \frac{f(\zeta)}{(\zeta - z)(\zeta - z_0)} \Big) \dd{\zeta} = \frac{1}{2 \pi i} \oint_C \frac{f(\zeta)}{(\zeta - z_0)^2} \dd{\zeta} \end{aligned}$$ Since the second-order derivative $$f''(z)$$ is simply the derivative of $$f'(z)$$, this proof works inductively for all higher orders $$n$$. {% include proof/end.html id="proof-dv-formula" %}