finished metric spaces

This commit is contained in:
Robert 2021-03-28 01:33:45 +01:00
parent bdcb40e561
commit c74c37a2c5
7 changed files with 344 additions and 1 deletions

View file

@ -0,0 +1,8 @@
\documentclass[../script.tex]{subfiles}
%! TEX root = ../../script.tex
\begin{document}
\chapter{Multivariable Calculus}
\subfile{sections/partial_total_diff.tex}
\end{document}

View file

@ -0,0 +1,323 @@
\documentclass[../../script.tex]{subfiles}
% !TEX root = ../../script.tex
\begin{document}
\section{Convergence of Function sequences}
\begin{defi}[Pointwise convergence]
Let $M$ be a set, $f_n: M \rightarrow \field ~~\forall n \in \natn$ and $f: M \rightarrow \field$.
The sequence $\seq{f}$ is said to be pointwise convergent to $f$ if
\[
\limn f_n(x) = f(x) ~~\forall x \in M
\]
\end{defi}
\begin{eg}
Consider
\begin{align*}
f_n: [0, 1] &\longrightarrow \realn \\
x &\longmapsto \begin{cases}
1 - nx, & x \in [0, \frac{1}{n}] \\
0, & \text{else}
\end{cases}
\end{align*}
\begin{center}
\begin{tikzpicture}[domain=0:1]
\begin{axis}[xmin=0, ymin=0, xmax=1, ymax=1, samples=50]
\addplot[black, thick] {1-x} node[above right, pos=0.6] {$f_1$};
\addplot[black, thick] {1-2*x} node[above right, pos=0.4] {$f_2$};
\addplot[black, thick] {1-3*x} node[right, pos=0.3] {$f_3$};
\addplot[black, thick] {1-4*x} node[left, pos=0.23] {$f_4$};
\end{axis}
\end{tikzpicture}
\end{center}
The $f_n$ are continuous for all $n \in \natn$ and converge pointwise to
\begin{align*}
f: [0, 1] &\longrightarrow \realn \\
x &\longmapsto \begin{cases}
1, & x = 0 \\
0, & x \ne 0
\end{cases}
\end{align*}
$f$ is not continuous.
\end{eg}
\begin{rem}
Let $M$ be a set. Then
\[
B(M) = \set[\exists K \in \realn: ~~|f(x)| < K ~~\forall x \in M]{f_n: M \longrightarrow \field}
\]
is a linear subspace of the space of all functions $M \rightarrow \field$. We can define the supremum norm
\begin{align*}
\supnorm{\cdot}: B(M) &\longrightarrow \realn \\
f &\longmapsto \sup_{x \in M}\set{\abs{f(x)}}
\end{align*}
\end{rem}
\begin{proof}
We will now proof that $\supnorm{\cdot}$ is a norm.
It is defined, because
\begin{equation}
\supnorm{f} = 0 \implies \abs{f(x)} = 0 ~~\forall x \in M
\end{equation}
This implies
\begin{equation}
f(x) = 0 ~~\forall x \in M \implies f = 0
\end{equation}
The triangle inequality is proven by first considering
\begin{equation}
\abs{f(x)} \le \supnorm{f} ~~\forall f \in B(M) ~\forall x \in M
\end{equation}
Let $f, g \in B(M)$, then
\begin{equation}
\abs{f(x) + g(x)} \le \abs{f(x)} + \abs{g(x)} \le \supnorm{f} + \supnorm{g} ~~\forall x \in M
\end{equation}
Which implies
\begin{equation}
\supnorm{f + g} = \sup_{x \in M} \abs{f(x) + g(x)} \le \supnorm{f} + \supnorm{g}
\end{equation}
\end{proof}
\begin{defi}[Uniform convergence]
A sequence of bounded functions $\seq{f}$,
\[
f_n: M \longrightarrow \field
\]
is said to be uniformly convergent to $f: M \rightarrow \field$ if its norm converges.
\[
\supnorm{f_n - f} \conv{n \rightarrow \infty} 0
\]
\end{defi}
\begin{rem}
Formally, pointwise convergence means
\[
\forall \epsilon > 0 ~\forall x \in M ~\exists N \in \natn ~\forall n \ge N: ~~\abs{f_n(x) - f(x)} < \epsilon
\]
and uniform convergence means
\[
\forall \epsilon > 0 ~\exists N \in \natn ~\forall x \in M ~\forall n \ge N: ~~\abs{f_n(x) - f(x)} < \epsilon
\]
\end{rem}
\begin{thm}
The function space $B(M)$ is complete.
\end{thm}
\begin{proof}
Let $\anyseqdef[f]{B(M)}$ be a Cauchy sequence in terms of $\supnorm{\cdot}$. Firstly, we have for some fixed $x \in M$
\begin{equation}
\abs{f_n(x) - f_m(x)} \le \supnorm{f_n - f_m}
\end{equation}
Since $\seq{f}$ is a Cauchy sequence, $(f_n(x))$ is also a Cauchy sequence in $\field_0$. Because $\field$ is complete,
$(f_n(x))$ converges, and we define
\begin{equation}
f(x) = \limn f_n(x)
\end{equation}
thus $\seq{f}$ converges pointwise to $f$. Let $\epsilon > 0$. Then
\begin{equation}
\exists N \in \natn: ~~\supnorm{f_n \cdot f_m} < \epsilon ~~\forall n, m \ge N
\end{equation}
Then $\forall x \in M, ~\forall n, m \ge N$ we have
\begin{equation}
\abs{f_n(x) - f_m(x)} \le \supnorm{f_n - f_m} < \epsilon
\end{equation}
We can find the limit for $m \rightarrow \infty$
\begin{equation}
\abs{f(x) - f_n(x)} \le \epsilon
\end{equation}
and
\begin{equation}
\supnorm{f} = \sup_{x \in M}\abs{f} \le \sup_{x \in M} \abs{f(x) - f_n(x)} + \sup_{x \in M} \abs{f_n(x)} = \epsilon + \supnorm{f_n}
\end{equation}
Thus, $f$ is bounded. Furthermore
\begin{equation}
\supnorm{f - f_n} = \sup_{x \in M} \abs{f(x) - f_n(x)} \le \epsilon
\end{equation}
which in turn implies
\begin{equation}
\supnorm{f - f_n} \conv{n \rightarrow \infty} 0
\end{equation}
\end{proof}
\begin{defi}
Let $\metric$ be a metric space, then $C_b(X)$ is said to be the space of all continuous bounded functions.
\end{defi}
\begin{rem}
If $X$ is compact (e.g. a bounded, closed subset of $\realn^n$) then all continuous functions are bounded. We then write $C(X)$ for $C_b(X)$.
\end{rem}
\begin{thm}
Let $\metric$ be a metric space. $C_b(X)$ is closed in $B(X)$. In other words, every uniformly convergent sequence of continuous functions converges to a continuous function.
\end{thm}
\begin{proof}
Let $\anyseqdef[f]{C_b(X)}$ be a sequence that uniformly converges to $f \in B(X)$.
Let $x \in X$ and $\epsilon > 0$, then
\begin{equation}
\exists N \in \natn: ~~\supnorm{f - f_n} y \frac{\epsilon}{3} ~~\forall n \ge N
\end{equation}
Choose a fixed $n \ge N$. Since $f_n$ is continuous, this means that
\begin{equation}
\exists \delta > 0: ~~\abs{f_n(x) - f_n(y)} < \frac{\epsilon}{3} ~~\forall y \in \oball[\delta](x)
\end{equation}
Then we have for all such $y$
\begin{equation}
\begin{split}
\abs{f(x) - f(y)} &\le \abs{f(x) - f_n(x)} + \abs{f_n(x) - f_n(y)} + \abs{f_n(y) - f(y)} \\
&\le 2 \cdot \supnorm{f - f_n} + f_n(x) - f_n(y) < \epsilon
\end{split}
\end{equation}
This proves the continuity of $f$ in $x$. Since $x \in X$ was chosen arbitrarily, $f$ is continuous everywhere.
\end{proof}
\begin{defi}
Let $x_0 \in \field$ and $\anyseqdef[a]{\field}$. Then
\[
\series{n} a_n (x - x_0)^n
\]
is called a power series around $x_0$. The number
\[
\rho := \sup\set[\series{n} a_n (x - x_0)^n \text{ converges}]{\abs{x - x_0}}
\]
is the convergence radius.
\begin{center}
\begin{tikzpicture}
\draw[->, thick] (0, 0) -- (0, 4) node [above] {};
\draw[->, thick] (0, 0) -- (4, 0) node [right] {};
\draw[fill] (2, 2) circle [radius=1pt] node [below right] {$x_0$};
\draw[dashed] (2, 2) circle [radius=1.2cm];
\draw[<->, rotate around={45:(2, 2)}] (2, 2) -- node[above] {$\rho$} (3.2, 2);
\draw[fill, rotate around={100:(2, 2)}] (3.2, 2) circle [radius=1pt] node [above] {$x$};
\end{tikzpicture}
\end{center}
\end{defi}
\begin{rem}
All results so far (including proofs) can be extended to $\realn^n$-valued functions, or functions with values in a Banach space in general.
\end{rem}
\begin{thm}
Let $\series{n} a_n (x - x_0)^n$ be a power series with convergence radius $\rho \in [0, \infty) \cup \set{\infty}$.
If $\abs{x - x_0} < \rho$ then the series converges absolutely, for $\abs{x - x_0} > \rho$ it diverges.
\[
\frac{1}{\rho} = \limsupn \sqrt[n]{\abs{a_n}}
\]
\end{thm}
\begin{proof}
W.l.o.g. choose $x_0 = 0$:
For $\abs{x} > \rho$ the series diverges by definition. If $\abs{x} < \rho$ then there exists $y \in \field$ such that
$\abs{x} < \abs{y} \le \rho$ and $\series{n} a_n y^n$ convergent. Especially, $(a_n y^n)$ is a null sequence.
This means $\exists C > 0$ such that $\abs{a_n y^n} \le C ~~\forall n \in \natn$
\begin{equation}
\series{n} \abs{a_n x^n} = \series{n} \abs{a_n y^n} \abs{\frac{x}{y}}^n \le C \cdot \series{n} \abs{\frac{x}{y}}^n < \infty
\end{equation}
This statement only holds for $\rho > 0$.
\end{proof}
\begin{rem}
\begin{enumerate}[(i)]
\item We have
\[
\rho = \sup \set[\series{n}\abs{a_n}a^n \text{ converges}]{a \in [0, \infty)}
\]
\item If the following limit exists, then
\[
\rho = \limn \frac{\abs{a_n}}{\abs{a_{n+1}}}
\]
\end{enumerate}
\end{rem}
\begin{eg}
The series
\[
\series{n} x^n
\]
is convergent on $(-1, 1)$, so $\rho = 1$. The limit function is
\[
x \longmapsto \frac{1}{1 - x}
\]
\end{eg}
\begin{thm}
Let $\series{n} a_n (x - x_0)^n$ be a power series with convergence radius $\rho > 0$.
Let $0 < a < \rho$. Then this power series converges uniformly on $\cball[a](x_0)$. Especially
\begin{align*}
f: \oball[\rho](x_0) &\longrightarrow \realn \\
x &\longmapsto \series{n} a_n (x_n - x_0)^n
\end{align*}
\end{thm}
\begin{proof}
W.l.o.g. choose $x_0 = 0$. Let $0 < a < \rho$. We know that $\series{n} a_nx^n$ converges on $\cball[a](0)$.
\begin{center}
\begin{tikzpicture}
\draw[->, thick] (0, -3) -- (0, 3);
\draw[->, thick] (-3, 0) -- (3, 0);
\draw[fill] (0, 0) circle [radius=1pt];
\draw[solid] (0, 0) circle [radius=1.8cm];
\draw[dashed] (0, 0) circle [radius= 2.4cm] node[above right=2.4cm] {$\rho$};
\node at (2, 0.2) {$a$};
\end{tikzpicture}
\end{center}
Define
\begin{equation}
\begin{split}
f_n: \cball[a](0) &\longrightarrow \field \\
x &\longmapsto x^n ~~\forall n \in \natn
\end{split}
\end{equation}
We can see that
\begin{equation}
\supnorm{f} = \sup_{x \in \cball[a](0)} \abs{f_n} = \sup_{x \in \cball[a](0)} = a^n
\end{equation}
and thus
\begin{equation}
\series{n} a_n f_n \implies \series{n} \supnorm{a_n f_n} = \series{n} \abs{a_n}^n < \infty
\end{equation}
because $a < \rho$. The series $\series{n} a_n f_n$ is absolutely convergent in $C(\cball[a](0))$.
Since $C(\cball[a](0))$ is complete, $\series{n} a_n f_n$ is convergent because the partial sums $\series[N]{n} a_n f_n$ are continuous $\forall N \in \natn$.
Therefore $f$ is also continuous on $\cball[a](0)$. Let $x \in \oball[\rho](0)$. Then there exists some $a > 0$ such that $\abs{x} < a < \rho$.
Thus, $f$ is continuous on $\cball[a](0)$. Since $\cball[a](0)$ contains a neighbourhood of $x$, and continuity is a local property,
$f$ is also continuous in $x$. Because $x \in \oball[\rho](0)$ was chosen arbitrarily, $f$ is continuous.
\end{proof}
\begin{rem}
$\exp$, $\sin$, $\cos$ are continuous.
\end{rem}
\begin{eg}
The statements above can be extended to Banach space-valued power series (e.g. matrix-valued functions).
The norm on $\realn^{n \times n}$ is
\[
\norm{A} = \sup\set[{\forall x \in \oball[1](0)}]{\norm{Ax}}
\]
Define
\[
\exp(A) := \series{0} \frac{A^n}{n!}
\]
This converges $\forall A \in \realn^{n \times n}$, because
\begin{align*}
\series{n} \norm{\frac{A^n}{n!}} &= \series{n} \frac{1}{n!} \norm{A^n} \le \series{n} \frac{1}{n!} \norm{A}^n \\
&= \exp(\norm{A}) < \infty
\end{align*}
Thus, $\series{n} \frac{A^n}{n!}$ converges absolutely. Now consider the function
\begin{align*}
\realn &\longrightarrow \realn^{n \times n} \\
t &\longmapsto \exp(At)
\end{align*}
This is a matrix-valued power series
\[
\exp(At) = \series{n} \frac{(At)^n}{n!} = \series{n} \frac{A^n}{n!} t^n
\]
with a convergence radius of $\rho = \infty$. In this case $\exp(A + B)$ doesn't necessarily have to equal $\exp(A) \cdot \exp(B)$.
\end{eg}
\end{document}

View file

@ -0,0 +1,10 @@
\documentclass[../../script.tex]{subfiles}
%! TEX root = ../../script.tex
\begin{document}
\section{Partial and Total Differentiability}
\begin{defi}
\end{defi}
\end{document}

View file

@ -304,7 +304,7 @@ Let $V$ be a vector space. A non-empty set $W \subset V$ is called a vector subs
\begin{eg}
Consider
\[
W = \set[chi \in \realn]{(\chi, \chi) \in \realn^2}
W = \set[\chi \in \realn]{(\chi, \chi) \in \realn^2}
\]
This is a subspace, because
\[

View file

@ -8,4 +8,5 @@
\subfile{sections/seq_ser_limits.tex}
\subfile{sections/open_closed_sets.tex}
\subfile{sections/continuity.tex}
\subfile{sections/conv_of_funcseq.tex}
\end{document}

Binary file not shown.

View file

@ -156,5 +156,6 @@
\subfile{chapters/linear_algebra.tex}
\subfile{chapters/real_analysis_2.tex}
\subfile{chapters/topo_of_metr_spaces.tex}
\subfile{chapters/multivar_calc.tex}
\end{document}