Finished chapter 10

This commit is contained in:
Robert 2021-07-29 00:05:23 +02:00
parent 489afda40e
commit 2a8e0ef278
4 changed files with 14997 additions and 14036 deletions

View file

@ -11,4 +11,5 @@
\subfile{sections/identity_continuation.tex}
\subfile{sections/laurent_series.tex}
\subfile{sections/residual_calculus.tex}
\subfile{sections/potential_theory.tex}
\end{document}

View file

@ -0,0 +1,286 @@
% !TeX root = ../../script.tex
\documentclass[../../script.tex]{subfiles}
\begin{document}
\section{Application: Potential Theory}
\begin{defi}[Harmonic Function]
A function $\phi: U \rightarrow \realn$ with $U \subset \realn^d$ open, $d \in \natn$ is said to be harmonic if
\[
\laplacian\phi(x) = \sum_{j=1}^d \partial_j^2 \phi(x) = 0, \quad x \in \realn^d
\]
\end{defi}
\begin{thm}\label{thm:10.37}
If $f: U \rightarrow \cmpln$, $U \subset \cmpln$ is holomorphic, then $\Re(f)$ and $\Im(f)$ are harmonic on $U$.
\end{thm}
\begin{proof}
If $f$ is holomorphic on $U$, then $f$ is also analytic, which means it is infinitely differentiable on $U$.
Using the Cauchy-Riemann equations the desired statement can be shown.
\end{proof}
\begin{eg}[Potential problems in $\realn^2$]
Let $U \subset \realn^2$ be a domain with smooth boundary. A typical problem from electrostatics is
\begin{align*}
\laplacian \phi(x, y) &= 0 & (x, y) &\in U \\
\phi(x, y) &= \phi_0(x, y) & (x, y) &\in \boundary{U}
\end{align*}
where $\phi: \closure{U} \rightarrow \realn$ is the desired function with given boundary values $\phi_0 \in C(\boundary{}{U}, \realn)$.
Such a boundary value problem is known as the Dirichlet problem.
An important example is the Dirichlet problem for the upper half plane
\[
\set[y > 0]{(x, y) \in \realn^2}
\]
with $\phi_0 \in C(\realn)$. We want to assume that $\phi_0$ decreases like
\[
\abs{\phi_0(x)} \le \frac{c}{1 + \abs{x}}, \quad x \in \realn, c > 0
\]
near infinity.
\end{eg}
\begin{thm}[Poisson integral formula for the upper half plane]\label{thm:10.38}
The function
\[
\phi(x, y) = \rec{\pi} \int_{-\infty}^{\infty} \phi_0(t) \frac{y}{(x-t)^2 + y^2} \dd{t}
\]
solves the Dirichlet problem for the upper half plane
\begin{align*}
\laplacian\phi(x, y) &= 0 & (x, y) &\in \set[y > 0]{(x, y) \in \realn^2} \\
\phi(x, 0) &= \phi_0(x) & x &\in \realn
\end{align*}
\end{thm}
\begin{proof}
The function
\begin{equation}
f(z) := \rec{\pi i} \int_{-\infty}^{\infty} \phi_0(t) \rec{t-z} \dd{t}
\end{equation}
is holomorphic on $\cmpln_+$ with
\begin{equation}
\Re(f(x + iy)) = \rec{\pi} \int_{-\infty}^{\infty} \phi_0(t) \Im\left(\rec{t-z}\right) \dd{t}
\end{equation}
We find that
\begin{equation}
\Im\left(\rec{t - x - iy}\right) = \Im\left(\rec{(t - x) - iy}\right) = \Im\left(\frac{(t - x) + iy}{(t - x)^2 + y^2} \right) = \frac{y}{(t-x)^2 + y^2}
\end{equation}
and thus
\begin{equation}
\Re(f(x + iy)) = \phi(x, y)
\end{equation}
for $y > 0$. However according to \Cref{thm:10.37} this means that $\phi$ is harmonic for $y > 0$:
\begin{equation}
\laplacian \phi(x, y) = 0, \quad \forall(x, y) \in \set[y > 0]{(x, y) \in \realn^2}
\end{equation}
It remains to be shown that the boundary values are being accounted for. For that we rewrite the Poisson integral formula as
\begin{equation}
\phi(x, y) = (\phi_0 * P_y)(x) := \int_{\realn} \phi(t) P_y(x - t) \dd{t}
\end{equation}
with the Poisson kernel $P_y(x) = \rec{\pi} \frac{y}{x^2 + y^2}$. We will finish this proof later.
\end{proof}
\begin{defi}[Convolution]
We define
\[
L^1(\realn^d) := \set[\int_{\realn^d} \abs{f(x)} \dd{x} < \infty]{f: \realn^d \rightarrow \realn \text{ measurable}}
\]
the Lebesgue space of absolutely integrable functions. It is a complete, normed space with
\[
\norm{f}_{L^1} = \int_{\realn^d} \abs{f} \dd{x}
\]
It induces a metric space with the metric
\[
d(f, g) := \norm{f - g}_{L^1}
\]
Let $\anyseqdef[f]{L^1(\realn^d)}$. This sequence converges to $f \in L^1(\realn^d)$ if
\[
\norm{f_n - f}_{L^1} \conv{n \rightarrow \infty} 0
\]
Since $L^1$ is complete, every Cauchy sequence converges.
For $f, g \in L^1(\realn^d)$
\[
(f * g)(x) := \int_{\realn^d} f(y) g(x - y) \dd{y}
\]
is said to be the convolution of $f$ and $g$.
\end{defi}
\begin{thm}
The convolution is well defined as a mapping
\[
*: L^1(\realn^d) \times L^1(\realn^d) \longrightarrow L^1(\realn^d)
\]
with $\norm{f * g}_{L^1} \le \norm{f}_{L^1} \norm{g}_{L^1}$.
The space $(L^1(\realn^d), *)$ is a commutative and associative algebra, i.e. $\forall f, g, h \in L^1(\realn^d)$:
\begin{enumerate}[(i)]
\item $f * g = g * f$
\item $(f * g) * h = f * (g * h)$
\item $f * (g + h) = f * g + f * h$
\item $\forall \lambda \in \cmpln: \quad \lambda(f * g) = (\lambda f) * g = f * (\lambda g)$
\end{enumerate}
\end{thm}
\begin{proof}
We will only be proving that the mapping is well defined and that the inequality holds.
First, let
\begin{equation}
f, g \in L^1 \cap L^{\infty} := \set[\esssup_{x \in \realn^d} \abs{f(x)} < \infty]{f: \realn^d \rightarrow \realn \text{ measurable}}
\end{equation}
Then the convolution $f * g$ is well defined (pointwise almost everwhere), because
\begin{equation}
\abs{f(y) g(x - y)} \leexpl{$g \in L^{\infty}$} C \abs{f(y)} \implies f \in L^1 \implies \text{integrable}
\end{equation}
We then find that
\begin{equation}
\begin{split}
\norm{f * g}_{L^1} &= \int_{\realn^d} \abs{(f * g)(x)} \dd{x} = \int_{\realn^d} \abs{\int_{\realn^d} f(y) g(x - y) \dd{y}} \dd{x} \\
&\le \int_{\realn^d} \int_{\realn^d} \abs{f(y)} \abs{g(x - y)} \dd{y}\dd{x} \\
&= \int_{\realn^d} \abs{f(y)} \int_{\realn^d} \abs{g(x - y)} \dd{x}\dd{y}
\end{split}
\end{equation}
By substituting $z = x - y$ we get
\begin{equation}
= \int_{\realn^d} \abs{f(y)} \int_{\realn^d} \abs{g(z)} \dd{z} \dd{y} = \norm{f}_{L^1} \norm{g}_{L^1}
\end{equation}
For more general $f, g \in L^1$ we can approximate $f$ via a function sequence
\begin{equation}
f_n := \min \set{f, n} \in L^1 \cap L^{\infty}
\end{equation}
Then $f_n \conv{n \rightarrow \infty} f$ in $L^1$, since
\begin{equation}
\abs{f_n(x)} \le \abs{f(x)} \quad \forall x \in \realn^d
\end{equation}
By using the previous results we can conclude
\begin{equation}
\norm{f_n * g - f_m * g}_{L^1} = \norm{(f_n - f_m) * g}_{L^1} \le \underbrace{\norm{f_n - f_m}_{L^1}}_{\conv{n,m \rightarrow \infty} 0} \norm{g}_{L^1}
\end{equation}
So $(f_n * g)_{n \in \natn}$ is a Cauchy sequence in $L^1$ and thus
\begin{equation}
f * g := \lim_{n \rightarrow \infty} f_n * g
\end{equation}
\end{proof}
\begin{rem}
One can show that $(L^1(\realn^d), *)$ does not have a neutral element, i.e.
\[
\nexists \delta \in L^1(\realn^d): \quad f * \delta = f \quad \forall f \in L^1(\realn^d)
\]
\end{rem}
\begin{defi}[Good kernels, Approximative identity]
A sequence of convolution kernels $\anyseqdef[K]{L^1(\realn^d)}$ is said to be a class of good kernels if
\begin{align*}
\forall n \in \natn: &\quad \int_{\realn^d} K_n(x) \dd{x} = 1 \\
\exists M > 0 ~\forall n \in \natn: &\quad \int_{\realn^d}\abs{K_n(x)} \dd{x} \le M \\
\forall \delta > 0: &\quad \limn \int_{\abs{x} > \delta} \abs{K_n(x)} \dd{x} = 0
\end{align*}
A sequence of good kernels with $K_n \ge 0$ for all $n \in \natn$ is called Dirac sequence.
\end{defi}
\begin{thm}[Smoothing by convolution with good kernels]\label{thm:10.42}
Let $\anyseqdef[K]{L^1(\realn^d)}$ be a class of good kernels. Then:
\begin{enumerate}[(i)]
\item If $f \in L^1(\realn^d)$ then
\[
\norm{f * K_n - f}_{L^1(\realn^d)} = 0
\]
\item If $K_n \subset C^m(\realn^d) ~\forall n \in \natn$ and if the partial derivatives $\partial^{\alpha} K_n, ~\abs{\alpha} \le m$ are bounded, then
\[
f * K_n \in C^m(\realn^d) \text{ and } \partial^{\alpha} (f * K_n) = f * \partial^{\alpha} K_n
\]
\item If $f \in C(\realn^d)$ is bounded, then
\[
\limn (f * K_n)(x) = f(x), \quad \forall x \in \realn^d
\]
\end{enumerate}
\end{thm}
\begin{eg}
Let $\anyseqdef[\epsilon]{(0, \infty)}$ be a null sequence. Then
\begin{align*}
\text{Poisson kernels}& & P_{k}(x) &:= \rec{\pi} \frac{\epsilon_k}{x^2 + \epsilon_k^2} \\
\text{Gauß kernels}& & \delta_k(x) &:= (2\pi \epsilon_k^2)^{-\frac{d}{2}} e^{-\frac{\abs{x}^2}{2 \epsilon_k^2}}, \quad x \in \realn^d
\end{align*}
are classes of good kernels. Now let $0 \le \phi \in L^1(\realn^d)$ with $\norm{\phi}_{L^1} = 1$. Then
\[
\phi_k(x) = \rec{\epsilon_k^d} \phi\left(\frac{x}{\epsilon}\right)
\]
is a class of good kernels. We can show that
\[
P_{\epsilon_k} (x) = \rec{\pi} \rec{\epsilon_k^2} \frac{\epsilon_k}{\left(\frac{x}{\epsilon_k}\right)^2 + 1} = \rec{\epsilon_k} \rec{\pi} \rec{\left(\frac{x}{\epsilon_k}\right)^2 + 1} = \rec{\epsilon_k} P\left(\frac{x}{\epsilon_k}\right)
\]
One has to show that
\[
P(x) = \rec{\pi} \rec{1 + x^2} \in L^1
\]
with
\[
\int_{\realn} P(x) \dd{x} = 1
\]
To do that we can calculate
\[
\int_{\realn} \underbrace{P(x)}_{\ge 0} \dd{x} = \rec{\pi} \int_{\realn} \rec{1 + x^2} \dd{x} = \rec{\pi} (\arctan(\infty) - \arctan(-\infty)) = \rec{\pi} \left(\frac{\pi}{2} - \left(-\frac{\pi}{2}\right)\right) = 1
\]
\end{eg}
\begin{proof}[Proof of \Cref{thm:10.42} (iii)]
Consider $f * K_n(x) - f(x)$. We can calculate
\begin{equation}
\begin{split}
f * K_n(x) - f(x) &= \int_{\realn} f(x - y) K_n(y) \dd{y} - f(x) \underbrace{\int_{\realn^d} K_n(y) \dd{y}}_{=1} \\
&= \int_{\realn^d} \left(f(x - y) - f(x) \right) K_n(y) \dd{y}
\end{split}
\end{equation}
Since $f$ is continuous in $x$,
\begin{equation}
\forall \epsilon > 0 ~\exists \delta > 0: \quad \abs{f(x - y) - f(x)} < \epsilon \quad \forall \abs{y} < \delta
\end{equation}
From the definition of good kernels follows that
\begin{equation}
\exists M > 0: \quad \int_{\realn^d} \abs{K_n(y)} \dd{y} \le M
\end{equation}
which lets us conclude
\begin{equation}
\int_{\abs{y} < \delta} \underbrace{\abs{f(x - y) - f(x)}}_{< \epsilon} \abs{K_n(y)} \dd{y} < \epsilon \int_{\abs{y} < \delta} \abs{K_n(x)} \dd{y} \le \epsilon M
\end{equation}
By utilising the boundedness of $f$ and (iii) from the definition of good kernels we can show
\begin{equation}
\int_{\abs{y} \ge \delta} \underbrace{\abs{f(x - y) - f(x)}}_{\le 2c} \abs{K_n(y)} \dd{y} \le 2c \underbrace{\int_{\abs{y} \ge \delta} \abs{K_n(y)} \dd{y}}_{\epsilon} \le 2c\epsilon
\end{equation}
We can now use the previous results to show that
\begin{equation}
\begin{split}
\abs{f * K_n(x) - f(x)} &\le \int_{\realn^d} \abs{f(x - y) - f(x)} \abs{K_n(y)} \dd{y} \\
&= \underbrace{\int_{\abs{y} < \delta} \abs{f(x - y) - f(x)} \abs{K_n(y)} \dd{y}}_{\le M\epsilon} + \underbrace{\int_{\abs{y} \ge \delta} \abs{f(x - y) - f(x)} \abs{K_n(y)} \dd{y}}_{\le 2c\epsilon} \\
&\le \epsilon(M + 2c)
\end{split}
\end{equation}
Since $\epsilon$ can be chosen arbitrarily it follows that
\begin{equation}
f * K_n(x) \conv{n \rightarrow \infty} f(x) \quad \forall x \in \realn^d
\end{equation}
With this it is now easy to finish \textbf{the proof for \Cref{thm:10.38}}. We had seen that
\begin{equation}
\phi(x, y) = \int_{\realn} \phi_0(t) P_y(x - t) \dd{t} = (\phi_0 * P_y)(x)
\end{equation}
Now let $\anyseqdef[\epsilon]{(0, \infty)}$ be a null sequence. Since $\phi_0$ is continuous and bounded,
and since $\left(P_{\epsilon_n}\right) \subset {L^1(\realn^d)}$ is a class of good kernels, it follows from what we have just proven that
\begin{equation}
\lim_{n \rightarrow \infty} (\phi_0 * P_{\epsilon_n})(x) = \phi_0(x)
\end{equation}
All in all it follows that
\begin{align}
\laplacian \phi(x, y) &= 0 & x &\in \realn, ~y > 0 \\
\phi(x, 0) &\conv{y \rightarrow 0} \phi_0(x) & \forall x &\in \realn \nonumber
\end{align}
\end{proof}
\begin{rem}
If $\psi: U \rightarrow \cmpln$ is holomorphic, $U \subset \cmpln$ open and $V \subset U$ a domain, then $\psi(V)$ is also a domain with $\psi(\boundary{V}) = \boundary{\psi(V)}$.
(Open mapping principle). Then the solution to the Dirichlet problem on $V$
\begin{align*}
\laplacian \phi &= 0 & \text{on } &V \\
\phi &= \phi_0 & \text{on } &\boundary{V}
\end{align*}
can be obtained through a holonomic transformation $\psi: V \rightarrow \cmpln_+$ with $\phi(V) = \cmpln_+$ of the Dirichlet problem on the upper half plane.
\end{rem}
\end{document}

28745
script.pdf

File diff suppressed because it is too large Load diff

View file

@ -40,6 +40,7 @@
\DeclareMathOperator{\diag}{diag}
\DeclareMathOperator{\const}{const}
\DeclareMathOperator{\Log}{Log}
\DeclareMathOperator*{\esssup}{ess\,sup}
\newcommand{\natn}{\mathbb{N}}
\newcommand{\intn}{\mathbb{Z}}