notes

uni notes
git clone git://popovic.xyz/notes.git
Log | Files | Refs

main.tex (141355B)


      1 \documentclass{article}
      2 \usepackage[utf8]{inputenc}
      3 \usepackage{amssymb}
      4 \usepackage{amsmath}
      5 \usepackage{amsthm}
      6 %\usepackage[ngerman]{babel}
      7 \usepackage[english]{babel}
      8 \usepackage{geometry}
      9 \usepackage{cases}
     10 \usepackage{wasysym}
     11 \usepackage{enumitem}
     12 \usepackage{thmtools}
     13 \usepackage{tikz}
     14 \usepackage{pgfplots}
     15 \usepgfplotslibrary{polar}
     16 \pgfplotsset{compat = newest}
     17 
     18 \usetikzlibrary{angles,quotes} % for pic (angle labels)
     19 \usetikzlibrary{arrows.meta}
     20 \usetikzlibrary{calc}
     21 \usetikzlibrary{decorations.markings}
     22 \usetikzlibrary{bending}
     23 
     24 \tikzstyle{xline}=[black,thick,smooth]
     25 \tikzstyle{width}=[{Latex[length=5,width=3]}-{Latex[length=5,width=3]},thick]
     26 \tikzstyle{mydashed}=[dash pattern=on 1.7pt off 1.7pt]
     27 
     28 \tikzset{>=latex}
     29 \tikzset{
     30   traj/.style 2 args={xline,postaction={decorate},decoration={markings,
     31     mark=at position #1 with {\arrow{<}},
     32     mark=at position #2 with {\arrow{<}}}
     33   }
     34 }
     35 
     36 \geometry{
     37   left=3cm,
     38   right=3cm,
     39   top=2cm,
     40   bottom=4cm,
     41   bindingoffset=5mm
     42 }
     43 
     44 
     45 \newcommand*{\N}{\mathbb{N}}
     46 \newcommand*{\Z}{\mathbb{Z}}
     47 \newcommand*{\Q}{\mathbb{Q}}
     48 \newcommand*{\R}{\mathbb{R}}
     49 \newcommand*{\C}{\mathbb{C}}
     50 \newcommand*{\I}{\mathbb{I}}
     51 \newcommand*{\T}{\mathbb{T}}
     52 
     53 \newcommand*{\Ns}{\N^*}
     54 
     55 \newcommand*{\Hs}{{H^s}}
     56 \newcommand*{\dHs}{{\dot H^s}}
     57 \newcommand*{\Lp}{{L^p}}
     58 \newcommand*{\Lq}{{L^q}}
     59 \newcommand*{\Rd}{{\mathbb{R}^d}}
     60 \newcommand*{\Rn}{{\mathbb{R}^n}}
     61 \newcommand*{\half}[1]{\frac{#1}{2}}
     62 \newcommand*{\jbr}[1]{{\langle #1 \rangle}}
     63 \newcommand*{\ag}{\alpha}
     64 \newcommand*{\bg}{\beta}
     65 \newcommand*{\og}{\omega}
     66 \newcommand*{\Og}{\Omega}
     67 \newcommand*{\supp}{\text{supp}}
     68 \newcommand*{\xj}{{x_j}}
     69 \newcommand*{\reci}[1]{{\frac{1}{#1}}}
     70 \newcommand*{\tr}{\text{tr}}
     71 \newcommand*{\grad}{\text{grad}}
     72 \newcommand*{\spanl}{\text{span}}
     73 \newcommand*{\sgn}{\text{sgn}}
     74 \newcommand*{\limn}{\lim_{n\to\infty}}
     75 \newcommand*{\Per}{\text{Per}}
     76 \newcommand*{\diam}{\text{diam}}
     77 \newcommand*{\divg}{\text{div}}
     78 
     79 \newcommand{\vectwo}[2]{\begin{pmatrix} #1\\#2\end{pmatrix}}
     80 \newcommand{\vecthree}[3]{\begin{pmatrix} #1\\#2\\#3\end{pmatrix}}
     81 
     82 \newcommand*{\mattwo}[4]{\begin{pmatrix}
     83     #1&#2\\#3&#4
     84 \end{pmatrix}}
     85 \newcommand*{\matthree}[9]{\begin{pmatrix}
     86     #1&#2&#3\\#4&#5&#6\\#7&#8&#9
     87 \end{pmatrix}}
     88 %\newcommand*{\matfour}[16]{\begin{pmatrix}
     89  %   #1&#2&#3&#4\\#5&#6&#7&#8\\#9&#10&#11&#12\\#13&#14&#15&#16
     90 %\end{pmatrix}}
     91 
     92 \newcommand*{\matutwo}[3]{\begin{pmatrix}
     93     #1&#2\\ &#3\\
     94 \end{pmatrix}}
     95 \newcommand*{\matuthree}[6]{\begin{pmatrix}
     96     #1&#2&#3\\ &#4&#5\\ &&#6
     97 \end{pmatrix}}
     98 
     99 \newcommand*{\matdiagtwo}[2]{\mattwo{#1}{\,}{\,}{#2}}
    100 \newcommand*{\matdiagthree}[3]{\matthree{#1}{\,}{\,}{\,}{#2}{\,}{\,}{\,}{#3}}
    101 
    102 \newcommand{\matdoublediagthree}[5]{\matthree{#1}{#2}{\,}{\,}{#3}{#4}{\,}{\,}{#5}}
    103 
    104 \newcommand{\quests}
    105 {
    106 Modelling
    107 Autonomous ODE in R^n
    108 Invariant subspaces
    109 Stability of equilibria
    110 Polar coordinates
    111 Asymptotic behavior
    112 LaSalle's invariance principle
    113 Hamiltonian systems in 2D
    114 Special Hamiltonian systems: Newtonian systems
    115 Gradient systems in R^n
    116 First integral (or constant of motion)
    117 How to find centers
    118 Stable and unstable manifolds
    119 Center manifold
    120 Andronov bifurcation
    121 
    122 Ideas from the General theory of dynamical systems
    123 * (Def 2.29) Circle rotations
    124 Maps with complicated orbit structure
    125 Outlook: Coding for other systems
    126 }
    127 
    128 \declaretheorem[
    129 	name=Theorem,
    130 	numberwithin=section
    131 	]{thm}
    132 \declaretheorem[
    133 	name=Lemma,
    134 	sibling=thm,
    135 	]{lem}
    136 \declaretheorem[
    137 	name=Proposition,
    138 	sibling=thm,
    139 	]{prop}
    140 \declaretheorem[
    141 	name=Corollary,
    142 	sibling=thm,
    143 	]{cor}
    144 
    145 \declaretheorem[
    146 	name=Definition,
    147 	style=definition,
    148 	sibling=thm,
    149 	numbered=yes,
    150 	]{defin}
    151 
    152 \declaretheorem[
    153 	name=Remark,
    154 	style=remark,
    155 	numbered=no
    156 	]{rem}
    157 \declaretheorem[
    158 	name=Recall,
    159 	style=remark,
    160 	numbered=no
    161 	]{rec}
    162 \declaretheorem[
    163 	name=Example,
    164 	style=remark,
    165 	numbered=no
    166 	]{exam}
    167 \declaretheorem[
    168 	name=Notation,
    169 	style=remark,
    170 	numbered=no
    171 	]{notation}
    172 \declaretheorem[
    173 	name=Homework,
    174 	style=remark,
    175 	numbered=no
    176 	]{hw}
    177 
    178 
    179 \title{Dynamical Systems and Nonlinear Differential Equations 2023S \\
    180 Lecturers: Roland Zweimüller, Balázs Boros}
    181 \author{Nikolas Hauschka}
    182 \date{}
    183 
    184 \begin{document}
    185 
    186 \maketitle
    187 \tableofcontents
    188 
    189 \setcounter{section}{-1}
    190 
    191 \section{Notations and conventions}
    192 
    193 For simplicity, we use some notations and conventions, which aren't always specified. If not otherwise specified, we use the following things:
    194 
    195 \begin{enumerate}
    196     \item $U$ and $V$ are open sets.
    197 
    198     \item $U$ and $V$ are subsets of $\Rn$.
    199 
    200     \item For $A \subseteq \Rn$ we define $C^0(U,A) := C(U,A) := \{f:U \to A, f \text{ is continuous}\}$.
    201 
    202     \item $C^{d+1}(U,A) := \{f:U\to A, f \text{ is differetiable and each first partial derivative lies in } C^d(U,A)\}$.
    203 
    204     \item If $A = \Rn$ then we might write $C^d(U)$ instead of $C^d(U,\Rn)$.
    205 
    206     \item For $m,n\in\Z$ with $m\leq n$, we define $\jbr{m,n}:=\{m,\dots,n\}=[m,n]\cap\Z$.
    207 
    208     \item We also might write $\jbr{n}$ instead of $\jbr{1,n}$.
    209 
    210     \item We use $\N := \{0,1,\dots\}$ and $\Ns:=\{1,2,\dots\}$.
    211 
    212     \item We define $\T := \R / \Z$, which can be viewed as the interval $[0,1]$ where the ends are identified.
    213 
    214     \item If an element of $\T$ is added/subtracted/multplied with an element of $\R$, the result is the corresponding element of $\T \pmod1$. We won't always state that it's calculated $\pmod1$.
    215 
    216     \item For $a,b \in \T$ with $b<a$ the interval we define the interval $[a,b] := [a,1) \cup [0,b]$ if not otherwise stated.
    217 \end{enumerate}
    218 
    219 \section{Part 1}
    220 
    221 \subsection{Modelling}
    222 
    223 When modeling we have to make certain decisions. Some types of models are listed below.
    224 
    225 \begin{itemize}
    226     \item Continuous time or discrete time.
    227 
    228     \item Deterministic or stochastic.
    229 
    230     \item State space:
    231     \begin{itemize}
    232         \item finite, discrete
    233 
    234         \item metric space
    235 
    236         \item Euclidean space (finite dimensional)
    237 
    238         \item Hilbert space
    239     \end{itemize}
    240 
    241     \item Homogeneous (ODE) or inhomogeneous (PDE)
    242 
    243     \item Future depends only on the present or also on the past.
    244 
    245     \item autonomous or non-autonomous
    246 \end{itemize}
    247 
    248 In this lecture we choose: Continuous time, deterministic, Euclidean space, only on the present, autonomous.
    249 \newline
    250 \newline
    251 \subsection{Autonomous ODE in $\R^n$}
    252 \begin{thm}
    253     Given
    254     $$\dot x(t) = f(x(t))$$
    255     where $f:U \to \R^n$ is locally Lipschitz continuous. Then for all $p \in U$ there exists a unique  solution $t\mapsto \phi(t,p)$ on a maximal interval such that $\phi(0,p) = p$.
    256 \end{thm}
    257 
    258 Details can be found in the book "Differential Equations and dynamical systems" by Perko.
    259 
    260 \begin{exam}[Lotka reactions]
    261     $$\begin{array}{rcl}
    262     X &\stackrel{\kappa_1}{\to}& 2X\\
    263     X+Y &\stackrel{\kappa_2}{\to}& 2Y\\
    264     Y &\stackrel{\kappa_3}{\to}& 0
    265     \end{array}$$
    266 
    267     $$\begin{array}{rcccl}\dot x &=&\kappa_1x-\kappa_2xy &=& x(\kappa_1-\kappa_2y)\\
    268     \dot y&=&\kappa_2xy-\kappa_3y &=& y(\kappa_2x-\kappa_3)\\
    269     &&\kappa_1,\kappa_2, \kappa_3 &>& 0
    270     \end{array}$$
    271     Let's go through the details. The line $X \stackrel{\kappa_1}{\to} 2X$ that the species of type $X$ doubles at the rate $\kappa_1$. So it makes sense to get an equation like $\dot x =\kappa_1x$. The third line is similar except that the species of type $Y$ decreases with a rate of $\kappa_3$. So we get an equation like $\dot y=-\kappa_3y$. The second line is more complicated. Here the transformation occurs at a rate of $\kappa_2$ when two species of the different types meet. If the $X$ count or the $Y$ is doubled then there are twice as many meetings between $X$ and $Y$ so the expression $\kappa_2xy$ makes sense. By this transformation the $Y$ count increases while the $X$ count decreases, so we have to add $\kappa_2xy$ to the $\dot y$-equation and subtract it from the $\dot x$-equation.
    272 \end{exam}
    273 
    274 \begin{exam}[Ivanova reactions]
    275     $$\begin{array}{rcl}Z+X &\stackrel{\kappa_1}{\to}& 2X\\
    276     X+Y &\stackrel{\kappa_2}{\to}& 2Y\\
    277     Y+Z &\stackrel{\kappa_3}{\to}& 2Y
    278     \end{array}$$
    279 
    280 
    281 
    282     $$\begin{array}{rcl}\dot x &=& x(\kappa_1z - \kappa_2y)\\
    283     \dot y &=& y(\kappa_2x - \kappa_3z)\\
    284     \dot z &=& z(\kappa_3y - \kappa_1x)\end{array}$$
    285 \end{exam}
    286 
    287 \begin{exam}[Competitive Lotka-Volterra systems (two competing species)]
    288     $$\dot x=x(r_1+b_{11}x+b_{12}y)$$
    289     $$\dot y=y(r_2+b_{21}x+b_{22}y)$$
    290     With the parameters $r_j > 0, b_{ij} < 0$ for $i,j = 1,2$.
    291 \end{exam}
    292 
    293 \begin{exam}[Cyclic competition of three species.]
    294     $$\begin{array}{rcl}\dot x &=& x(1-x-\alpha y-\beta z)\\
    295     \dot y &=& y(1-\beta x-y-\alpha z)\\
    296     \dot z &=& x(1-\alpha x-\beta y-z)
    297     \end{array}$$
    298     With the parameters $\alpha,\beta > 0$ and the restriction $x,y,z \geq 0$.
    299 \end{exam}
    300 
    301 \begin{exam}[Lotka-Volterra equation]
    302     $$\dot x_i = x_i\left(r_i+\sum_{l = 1}^n b_{i,l}x_l\right), i\in\jbr{n}$$
    303     With the parameters $r \in \R^n, B \in \R^{n\times n}$ and the restriction $x \in \R^n_{\geq 0}$.
    304 \end{exam}
    305 
    306 \begin{exam}[Replicator dynamics on the simplex (e.g. rock-paper-scissors)]
    307     $$\dot x_i  = x_i((Ax)_i-x^TAx), i \in\jbr{n}$$
    308     With $A \in \R^{n\times n}$ and the restriction $x\in\R_{\geq0}^n$ such that $\sum_{j=1}^n x_j = 1$.
    309 \end{exam}
    310 
    311 \begin{exam}[Pendulum]
    312     Newton's second law of motion.
    313     $$\begin{aligned}F &= ma\\
    314     -mg\sin(x)&=ml\ddot x\\
    315     \ddot x + \frac{g}{l}\sin(x)&=0\end{aligned}$$
    316 \end{exam}
    317 
    318 \begin{exam}[Van der Pol oscilator (electrical engineering)]
    319     $$\ddot x-\mu(1-x^2)\dot x+x=0$$
    320     With the parameter $\mu \in \R$ and the restriction $x\in\R_{\geq0}$.
    321 \end{exam}
    322 
    323 \begin{exam}[SIR (epidemiology): $S \to I \to R$]
    324     The population can be divided into three groups:\\
    325     $S:$ susceptible\\
    326     $I:$ infected\\
    327     $R:$ recovered
    328     $$\begin{array}{rcl}\dot S&=&-\beta SI\\
    329     \dot I&=&\beta SI - \gamma I\\
    330     \dot R&=&\gamma I
    331     \end{array}$$
    332     Where $\beta > 0$ is the rate of transmission, $\gamma > 0$ is the rate of recovery and $S+I+R$ is constant.
    333 \end{exam}
    334 
    335 \begin{exam}[Two body problem]
    336     Here $r_1(t),r_2(t) \in \R^3$ describe the positions, $\dot r_1(t), \dot r_2(t) \in \R^3$ describe the velocities and $\ddot r_1(t),\ddot r_2(t) \in \R^3$ describe the accelerations. Also $m_1$ and $m_2$ are the masses and $\gamma > 0$ is the gravitational constant.
    337     $$m_1\ddot r_1 = - \frac{\gamma m_1 m_2}{|r_2-r_1|^3}(r_1-r_2)$$
    338     $$m_2\ddot r_2 = - \frac{\gamma m_1 m_2}{|r_2-r_1|^3}(r_2-r_1)$$
    339 \end{exam}
    340 
    341 \begin{exam}[Lorenz equation]
    342     $$\begin{array}{rcl}\dot x &=& \sigma(y-x)\\
    343     \dot y &=& \rho x-y-xz\\
    344     \dot z &=& xy-\beta z
    345     \end{array}$$
    346     Where $\sigma,\rho, \beta > 0$ and $x,y,z \in \R$. It shows chaos for $\sigma = 10, \rho = 28, \beta=\frac83$.
    347 \end{exam}
    348 
    349 \begin{exam}[Linear ODEs]
    350     $$\dot x(t) = Ax(t)$$
    351     Where $x(t)\in \R^n$ and $A \in \R^{n \times n}$. The solution is the function $x: \R \to \R^{n}$ with $x(t)=e^{At}x(0)$ for all $t$, where
    352     $$e^{At}:=\sum_{k=0}^\infty\frac{t^k}{k!}A^k.$$
    353 \end{exam}
    354 
    355 \begin{thm}[Real Jordan normalform]
    356     For any $A \in \R^{n\times n} $ there is an invertible $P \in \R^{n \times n}$, such that $B = P^{-1}AP$ is a block diagonal matrix with blocks of the form\\
    357     $$\matdoublediagthree\lambda1\ddots1\lambda \text{ or } \matdoublediagthree{D}{I_2}{\ddots}{I_2}{D},$$
    358     where $\lambda \in \sigma(A)\cap \R$ or $\mu \pm i\omega \in \sigma(A)\setminus\R$ ($\sigma(A)$ is the set of the generalized eigenvalues of $A$) and
    359     $$D = \mattwo\mu{-\omega}\omega\mu, \quad I_2=\mattwo1001.$$
    360 \end{thm}
    361 
    362 \begin{cor}
    363     Suppose that $P,B$ are as above. Then
    364     $$x(t)=Pe^{Bt}P^{-1}x(0).$$
    365 \end{cor}
    366 
    367 \begin{prop}
    368     If
    369     $$B = \matdoublediagthree{\lambda}{1}{\ddots}{1}{\lambda} \in \mathbb{R}^{m \times m}$$
    370     then
    371     $$e^{Bt} = e^{\lambda t}\matuthree1\dots{\frac{t^{m-1}}{(m-1)!}}\ddots\vdots1.$$
    372     If
    373     $$B = \matdoublediagthree{D}{I_2}{\ddots}{I_2}{D}  \in \mathbb{R}^{2m \times 2m}$$
    374     then
    375     $$e^{Bt}=e^{\mu t}\matuthree R\dots{\frac{Rt^{m-1}}{(m-1)!}}\ddots\vdots R,$$
    376     where
    377     $$R = \mattwo{\cos(\omega t)}{-\sin(\omega t)}{\sin(\omega t)}{\cos(\omega t)}.$$
    378 \end{prop}
    379 
    380 \begin{cor}
    381     Each coordinate of each solution of $\dot x = Ax$ is a linear combination of $e^{\mu t}t^k\cos(\omega t)$ or $e^{\mu t}t^k\sin(\omega t)$, where $\mu\pm i \omega \in \sigma(A)$ and $k$ is smaller than the multiplicity of $\lambda$ in the minimal polynomial of $A$.
    382 \end{cor}
    383 
    384 Let $n = 2$. Then we have the following cases:
    385 \begin{itemize}
    386     \item If $B = \matdiagtwo\lambda\mu$ then $e^{Bt} = \matdiagtwo{e^{\lambda t}}{e^{\mu t}}$.
    387 
    388     \item If $B = \matutwo\lambda1\lambda$ then $e^{Bt}=e^{\lambda t}\matutwo1t1$.
    389 
    390     \item If $B = \mattwo\mu{-\omega}\omega\mu$ then $e^{Bt}=e^{\mu t}\mattwo{\cos(\omega t)}{-\sin(\omega t)}{\sin(\omega t)}{\cos(\omega t)}$.
    391 \end{itemize}
    392 
    393 Depending on the case, the diagram looks different.
    394 
    395 \begin{enumerate}
    396     \item Saddle: $\matdiagtwo\lambda\mu$ with $\lambda < 0 < \mu$.
    397     The flow moves towards the origin on the horizontal axis and moves away on the vertical axis.
    398 
    399     \begin{tikzpicture}
    400         \begin{axis}[
    401             xmin = -4, xmax = 4,
    402             ymin = -4, ymax = 4,
    403             zmin = 0, zmax = 1,
    404             axis equal image,
    405             view = {0}{90},
    406             samples = 9,
    407             samples y = 9,
    408         ]
    409 
    410         \addplot3[
    411             quiver = {
    412                 u = {-x},
    413 		        v = {y},
    414                 scale arrows = 0.25,
    415             },
    416             -stealth,
    417         ] {0};
    418         \end{axis}
    419     \end{tikzpicture}
    420 
    421     \item Stable nodes:
    422 
    423     \begin{itemize}
    424         \item $B=\begin{pmatrix}
    425         \lambda & 0\\
    426         0 & \mu
    427         \end{pmatrix}$ with $\lambda = \mu < 0$.
    428 
    429         \begin{tikzpicture}
    430             \begin{axis}[
    431                 xmin = -4, xmax = 4,
    432                 ymin = -4, ymax = 4,
    433                 zmin = 0, zmax = 1,
    434                 axis equal image,
    435                 view = {0}{90},
    436                 samples = 9,
    437                 samples y = 9,
    438             ]
    439 
    440             \addplot3[
    441                 quiver = {
    442                     u = {-x},
    443     		        v = {-y},
    444                     scale arrows = 0.25,
    445                 },
    446                 -stealth,
    447             ] {0};
    448             \end{axis}
    449         \end{tikzpicture}
    450 
    451         \item $B=\begin{pmatrix}
    452         \lambda & 0\\
    453         0 & \mu
    454         \end{pmatrix}$ with $\lambda < \mu < 0$.
    455 
    456         \begin{tikzpicture}
    457             \begin{axis}[
    458                 xmin = -4, xmax = 4,
    459                 ymin = -4, ymax = 4,
    460                 zmin = 0, zmax = 1,
    461                 axis equal image,
    462                 view = {0}{90},
    463                 samples = 9,
    464                 samples y = 9,
    465             ]
    466 
    467             \addplot3[
    468                 quiver = {
    469                     u = {-2*x},
    470     		        v = {-y},
    471                     scale arrows = 0.25,
    472                 },
    473                 -stealth,
    474             ] {0};
    475             \end{axis}
    476         \end{tikzpicture}
    477 
    478         \item $B=\begin{pmatrix}
    479         \lambda & 1\\
    480         0 & \lambda
    481         \end{pmatrix}$ with $\lambda < 0$.
    482 
    483         \begin{tikzpicture}
    484             \begin{axis}[
    485                 xmin = -4, xmax = 4,
    486                 ymin = -4, ymax = 4,
    487                 zmin = 0, zmax = 1,
    488                 axis equal image,
    489                 view = {0}{90},
    490                 samples = 9,
    491                 samples y = 9,
    492             ]
    493 
    494             \addplot3[
    495                 quiver = {
    496                     u = {-x+y},
    497     		        v = {-y},
    498                     scale arrows = 0.25,
    499                 },
    500                 -stealth,
    501             ] {0};
    502             \end{axis}
    503         \end{tikzpicture}
    504     \end{itemize}
    505 
    506     The flow moves towards the origin from all sides.
    507 
    508     \item Focus: $\begin{pmatrix}
    509         \mu & -\omega\\
    510         \omega & \mu
    511     \end{pmatrix}$ with $\omega \neq 0, \mu \neq 0$.
    512 
    513     \begin{tikzpicture}
    514         \begin{axis}[
    515             xmin = -4, xmax = 4,
    516             ymin = -4, ymax = 4,
    517             zmin = 0, zmax = 1,
    518             axis equal image,
    519             view = {0}{90},
    520             samples = 9,
    521                 samples y = 9,
    522         ]
    523 
    524         \addplot3[
    525             quiver = {
    526                 u = {x-y},
    527 		        v = {x+y},
    528                 scale arrows = 0.25,
    529             },
    530             -stealth,
    531         ] {0};
    532         \end{axis}
    533     \end{tikzpicture}
    534 
    535     The flow moves around the origin and moves closer or further away from it (depending on the sign of $\mu$).
    536 
    537     \item Center: $\begin{pmatrix}
    538         \mu & -\omega\\
    539         \omega & \mu
    540     \end{pmatrix}$ with $\omega \neq 0, \mu = 0$.
    541 
    542     \begin{tikzpicture}
    543         \begin{axis}[
    544             xmin = -4, xmax = 4,
    545             ymin = -4, ymax = 4,
    546             zmin = 0, zmax = 1,
    547             axis equal image,
    548             view = {0}{90},
    549             samples = 9,
    550             samples y = 9,
    551         ]
    552 
    553         \addplot3[
    554             quiver = {
    555                 u = {-y},
    556 		        v = {x},
    557                 scale arrows = 0.25,
    558             },
    559             -stealth,
    560         ] {0};
    561         \end{axis}
    562     \end{tikzpicture}
    563 
    564     The flow moves periodically around the origin in circles.
    565 \end{enumerate}
    566 
    567 If $0 \in \sigma(A)$ then the origin is not an isolated equilibrium.\\
    568 %Bifurcation diagram.\\
    569 \newline
    570 Let $\delta = \det(A)$ and $\tau = \tr(A)$. Then the characteristic polynomial of $A$ is: $x^2-\tau x + \delta$.
    571 Given
    572 $$\dot x = Ax$$
    573 we define $B := P^{-1}AP$ and $y := P^{-1}x$. Then we get
    574 $$\dot y = P^{-1}\dot x = P^{-1}Ax = P^{-1}APy = By.$$
    575 
    576 \subsection{Invariant subspaces}
    577 
    578 \begin{defin}
    579     Let $A\in\C^{n\times n}$ and $\lambda \in \sigma(A), v \in \C^n$ is a generalized eigenvector, if
    580     $$(\lambda I-A)^kv=0$$
    581     for some $k\in\Ns$.
    582 \end{defin}
    583 
    584 \begin{thm}
    585     Let $A\in \R^{n\times n}$ with eigenvalues as follows:
    586     \begin{itemize}
    587         \item $\lambda_1,\dots,\lambda_k\in\R$
    588 
    589         \item $\lambda_j=\mu_j+i\omega_j, \overline{\lambda_j}=\mu_j-i\omega_j, \quad j\in\jbr{k+1,m}\quad (2m-k=n)$
    590     \end{itemize}
    591 
    592     Then the set $\{u_1,\dots,u_k,u_{k+1},v_{k+1},\dots,u_m,v_m\}$ is a basis of $\R^n$, where $u_1, \dots,u_k$ are generalized eigenvectors corresponding to $\lambda_1, \dots, \lambda_k$ and the $u_j\pm i v_j$ are generalized eigenvectors corresponding to the $\mu_j \pm i\omega_j$ where $j\in\jbr{k+1,m}$.
    593 \end{thm}
    594 
    595 \begin{defin}
    596     We define the following subspaces:
    597     \begin{itemize}
    598         \item Stable subspace: $E^s=\spanl\{u_j,\nu_j:\Re(\lambda_j)<0\}$.
    599 
    600         \item Center subspace: $E^c=\spanl\{u_j,\nu_j:\Re(\lambda_j)=0\}$.
    601 
    602         \item Unstable subspace: $E^u=\spanl\{u_j,\nu_j:\Re(\lambda_j)>0\}$.
    603     \end{itemize}
    604     We also define:
    605     \begin{itemize}
    606         \item $s(A) = \dim(E^s)$
    607 
    608         \item $c(A) = \dim(E^c)$
    609 
    610         \item $u(A) = \dim(E^u)$
    611     \end{itemize}
    612 \end{defin}
    613 
    614 \begin{exam}
    615     Given
    616     $$\begin{aligned}
    617     \dot x&=-2x-y\\
    618     \dot y &= x-2y\\
    619     \dot z &= 3z,
    620     \end{aligned}$$
    621     we can read the matrix
    622     $$A=\matthree{-2}{-1}{0}{1}{-2}{0}{0}{0}{3}.$$
    623     We get the following eigenpairs:
    624     \begin{itemize}
    625         \item $\lambda_1 = 3, \quad u_1 =\vecthree001$
    626 
    627         \item $\lambda_2 = -2+i, \quad u_2+iv_2 = \vecthree010+i\vecthree100$
    628     \end{itemize}
    629     Also
    630     $$\begin{array}{ll}E^s=(x,y)-\text{plane}, &\quad s(A) = 2,\\
    631     E^c=\{0\}, &\quad c(A)=0,\\
    632     E^u=z-\text{axis}, &\quad u(A)=1.
    633     \end{array}$$
    634 \end{exam}
    635 
    636 \begin{exam}
    637     Given
    638     $$\begin{array}{rcl}
    639     \dot x &=& -y\\
    640     \dot y &=& x\\
    641     \dot z &=& 2z,
    642     \end{array}$$
    643      we can read the matrix
    644     $$A=\matthree{0}{-1}{0}{1}{0}{0}{0}{0}{2}$$
    645     We get the following eigenpairs:
    646     \begin{itemize}
    647         \item $\lambda_1=2, \quad u_1=\vecthree001$
    648 
    649         \item $\lambda_2=i, \quad u_2+iv_2=\vecthree010+i\vecthree100$
    650     \end{itemize}
    651     $\begin{array}{ll}E^s=\{0\},\\E^c=(x,y)-\text{plane},\\
    652     E^u=z-\text{axis}.
    653     \end{array}$
    654 \end{exam}
    655 
    656 \begin{exam}
    657     Given
    658     $$\dot x=0$$
    659     $$\dot y = x,$$
    660     we can read the matrix
    661     $$A=\mattwo0010$$
    662     Since $(A-0)^2 = 0$, we get the following eigenpairs:
    663     \begin{itemize}
    664         \item $\lambda_1=0, \quad u_1=\vectwo01$
    665 
    666         \item $\lambda_2=0, \quad u_2\vectwo10$.
    667     \end{itemize}
    668     $\begin{array}{ll}E^s=\{0\},\\
    669     E^c=(x,y)-\text{plane},\\
    670     E^u=\{0\}.
    671     \end{array}$
    672 \end{exam}
    673 
    674 \begin{thm}
    675     The whole space is a direct sum of the three subspaces, meaning $\R^n=E^s\oplus E^c \oplus E^u$. Also the following statements hold:
    676     \begin{itemize}
    677         \item $E^i$ is invariant ($i = s,c,u)$.\\
    678         If $x(0)\in E^i$ then $x(t)=e^{At}x(0)\in E^i$ for all $t \in \R$.
    679 
    680         \item If $x(0) \in E^s$ then $x(t) \to 0$ as $t\to \infty$. Moreover there exist $K,\alpha > 0$, such that
    681         $$\|e^{At}\| \leq  Ke^{-\alpha t}$$
    682         for all $ t \geq 0$.
    683 
    684         \item If $x(0)\in E^u$ then $x(t) \to 0$ as $t \to -\infty$. Moreover there exist $L,\beta > 0$, such that
    685         $$\|e^{At}\|\leq Le^{\beta t}$$
    686         for all $t\leq 0$.\\
    687         %$|x(t)|\leq Le^{\beta t}|x(0)| \forall t \leq 0$
    688         \item $E^s = \{ p \in \R^n:e^{At}p \to 0$ as $t \to \infty\}$
    689         \item $E^u = \{ p \in \R^n:e^{At}p \to 0$ as $t \to -\infty\}$
    690     \end{itemize}
    691 \end{thm}
    692 
    693 \begin{defin}
    694     We say that a matrix $A$ is stable if $\Re(\lambda) < 0$ for all $\lambda \in\sigma(A)$.
    695 \end{defin}
    696 
    697 \begin{itemize}
    698     \item For $n=2$ the matrix $A$ is stable if and only if $\det(A) > 0$ and $\tr(A) < 0$. The characteristic polynomial is $\lambda^2-\tr(A)\lambda+\det(A)$.
    699 
    700     %\det(A)>M\tr(A)
    701     \item For $n=3$ the matrix $A$ is stable if and only if $\det(A) < 0, \tr(A)< 0,M>0$, where $M$ is the sum of the $2\times2$ principal minors
    702     $$M =\det\mattwo{a_{11}}{a_{12}}{a_{21}}{a_{22}} + \det\mattwo{a_{11}}{a_{13}}{a_{31}}{a_{33}}+\det\mattwo{a_{22}}{a_{23}}{a_{32}}{a_{33}}.$$
    703     The characteristic polynomial is (up to the faktor $(-1)$) $\lambda^3 -\tr(A)\lambda^2+M\lambda-\det(A)$.
    704 \end{itemize}
    705 
    706 
    707 \begin{thm}[Routh-Hurwitz]
    708     Each root of $x^n+b_{n-1}x^{n-1}+\dots +b_1x+b_0$ has a negative real part if and only if all the leading principal minors of $H$ are positive, where
    709     $$H := \begin{pmatrix}
    710     b_{n-1} & 1 & 0 & 0 & 0 & 0 & 0 & \dots & 0\\
    711     b_{n-3} & b_{n-2} & b_{n-1} & 1 & 0 & 0 & 0 & \dots & 0\\
    712     b_{n-5} & b_{n-4} & b_{n-3} & b_{n-2} & b_{n-1} & 1 & 0 & \dots & 0\\
    713     \vdots & \ddots & \ddots & \ddots & \ddots & \ddots & \ddots & \ddots & \vdots\\
    714     0 & \dots & 0 & 0 & b_0 & b_1 & b_2 & b_3 & b_4\\
    715     0 & \dots & 0 & 0 & 0 & 0 & b_0 & b_1 & b_2\\
    716     0 & \dots & 0 & 0 & 0 & 0 & 0 & 0 & b_0
    717     \end{pmatrix}.$$
    718 \end{thm}
    719 Applying this to $\det(xI-A)$, we get
    720 \begin{itemize}
    721     \item $b_0 = (-1)^n\det(A)$
    722 
    723     \item $b_{n-1} = -\tr(A)$
    724 
    725     \item $b_j=(-1)^{n-j}\sum_{\alpha\subseteq \jbr{n}, |\alpha|=n-j} A[\alpha|\alpha]$\\
    726     The sum of principal minors corresponding to the rows/columns in $\alpha$.
    727 \end{itemize}
    728 
    729 \begin{thm}[Lyapunov]
    730     For $A\in \R^{n\times n}$. The following are equivalent:
    731     \begin{enumerate}
    732         \item $A$ is stable.
    733 
    734         \item There is a symmetric $Q>0$, such that $QA+A^TQ < 0$.
    735 
    736         \item For every symmetric $C <0$ there is a symmetric $Q>0$ such that $QA+A^TQ=C$.
    737     \end{enumerate}
    738 \end{thm}
    739 
    740 \begin{proof}
    741     We prove $(2) \Rightarrow (1) \Rightarrow (3) \Rightarrow (2)$.
    742     \begin{itemize}
    743         \item $(2) \Rightarrow (1)$: Let $Q$ be as in the statement. Now we take the time derivative of $x^TQx$. For better readability, we write $x$ instead of $x(t)$.
    744         $$\frac{d}{dt}(x^TQx) = \dot x^TQx+x^TQ\dot x = x^TA^TQx+x^TQAx = x^T(QA+A^TQ)x.$$
    745         The last expression is negative if $x(t) \neq 0$. The level sets $\{x \in \R^n: x^TQx = \gamma\}$ for $\gamma > 0$ are ellipsoids and solutions are going inwards, because if the time increases then $x^TQx$ decreases. Therefore $x(t) = e^{At}p\to 0$ as $t\to\infty$ for all $p \in \R^n$. So we get $E^s = \R^n$ and $A$ is stable.
    746 
    747         \item $(1)\Rightarrow (3):$ Let $C < 0$ be a symmetric matrix. Let
    748         $$Q = -\int_0^\infty e^{A^Tt}Ce^{At}dt.$$ It exists and is finite, because $\|e^{At}\|\leq Ke^{-\alpha t}$ for $t\geq 0$.
    749         \begin{itemize}
    750             \item Clearly $Q=Q^T>0$, because
    751             $$Q^T = \left(-\int_0^\infty e^{A^Tt}Ce^{At}dt\right)^T = -\int_0^\infty\left(e^{A^Tt}Ce^{At}\right)^Tdt= -\int_0^\infty\left(e^{At}\right)^TC^T\left(e^{A^Tt}\right)^Tdt = Q$$
    752             and
    753             $$x^TQx = -\int_0^\infty x^Te^{A^Tt}Ce^{At}xdt = -\int_0^\infty \left(e^{A^Tt}x\right)^TCe^{At}xdt = -\int_0^\infty y^TCydt > 0.$$
    754 
    755             \item We also note that
    756             $$QA+A^TQ = -\int_0^\infty e^{A^Tt}Ce^{At}A+A^Te^{A^Tt}Ce^{At}dt = -\int_0^\infty \frac{d}{dt}(e^{A^Tt}Ce^{At})dt =$$
    757             $$=-[e^{A^Tt}Ce^{At}]_{t=0}^{t=\infty} = -(0-C)=C.$$
    758         \end{itemize}
    759 
    760         \item $(3)\Rightarrow (2)$: Trivial.
    761     \end{itemize}
    762 \end{proof}
    763 
    764 \begin{defin}
    765     For $k \in\Ns$, we say that $h:U\to V$ is a homeomorphism (or $C^k$-diffeomorphism) if it is bijective and $h,h^{-1}$ are continuous (or $C^k$).\\
    766 \end{defin}
    767 
    768 \begin{defin}[Flow]
    769     Let  $f \in C^1(U)$. Denote the solution of
    770     $$\begin{aligned}\dot x(t) &= f(x(t))\\
    771     x(0) &= p\end{aligned}$$
    772     by $t \mapsto \varphi(t,p)$.\\
    773     The flow of $\dot x=f(x)$ is
    774     $$\varphi_t:\{p\in U: t\in I(p)\} \to U$$
    775     $$p \mapsto \varphi(t,p)$$
    776     For fixed $t$, the set $\varphi_t(K)$ is where a set $K$ is moved after time $t$ (as opposed to $\{\varphi(t,p)\colon t \in I(p)\}$, which is the trajectory of a single point $p$).
    777 \end{defin}
    778 
    779 For $\dot x = Ax$, the flow is $\varphi_t = e^{At}:\R^n\to\R^n$.
    780 
    781 \begin{prop}
    782     The flow has the following properties (where these make sense):
    783     \begin{enumerate}
    784         \item $\varphi_0=id$
    785 
    786         \item $\varphi_s\circ\varphi_t = \varphi_{t+s}$
    787 
    788         \item $\varphi_t\circ\varphi_{-t}=id$
    789     \end{enumerate}
    790 \end{prop}
    791 
    792 \begin{defin}[Conjugate flow]
    793     We have the following ODEs
    794     $$\dot x=f(x)$$
    795     in $U$ with the flow $\varphi_t$ and
    796     $$\dot y=g(y)$$
    797     in $V$ with the flow $\psi_t$.\\
    798     We say that the flows $\varphi_t$ and $\psi_t$ are topologically conjugate if there is a homeomorphism $h:U\to V$ such that $h\circ \varphi_t=\psi_t\circ h$. If $h$ is a $C^k$-diffeomorphism, we say $\varphi_t$ and $\psi_t$ are $C^k$ conjugate.
    799 \end{defin}
    800 
    801 \begin{thm}
    802     If the matrices $A,B \in\R^{n\times n}$ are stable then the flows $e^{At}, e^{Bt}$ are topologically conjugate. (There is a homeomorphism $\R^n\to\R^n$ such that $h(e^{At}p)=e^{Bt}h(p)$ for all $p\in\R^n$ and $t\in\R$.)
    803 \end{thm}
    804 
    805 \begin{proof}[Sketch of the proof.]
    806     Without loss of generality, assume $B = -I$ (because being topological conjugate is an equivalence relation). Because $A$ is stable, we can find a symmetric $Q > 0$ satisfying $QA+A^TQ < 0$. Let
    807     $$S = \{p\in\R^n:p^TQp=1\}.$$
    808     Then for all $p \in \R^n\setminus \{0\}$ there is a unique $\tau(p)$ such that $e^{A\tau(p)}p\in S$. It exists, because $e^{At}p$ approaches $0$ at $\infty$ and $\infty$ at $-\infty$ and the flow is continuous. It is unique because the time derivative of $x(t)^TQx(t)$ is negative (see proof of the Lyapunov theorem). Then $\tau:\R^n \setminus\{0\} \to \R$ is continuous (because solutions depend on the initial conditions continuously).\\
    809     %$$e^{A\tau(e^{At}p)}\left(e^{At}p\right) = e^{A(\tau(e^{At}p)+t)}p = e^{A\tau(p)}p,$$
    810     Since $p$ and $e^{At}p$ are on the same orbit with a time difference of $t$, we know that
    811     $$\tau(e^{At}p)=\tau(p)-t.$$
    812     Define
    813     $$h(p) := \begin{cases}
    814         e^{(A+I)\tau(p)}p&p \neq 0\\
    815         0&p=0.
    816     \end{cases}$$
    817     The orbits are mapped to orbits: Indeed, for $p\neq 0$ we get
    818     $$h(e^{At}p)=e^{(A+I)\tau(e^{At}p)}e^{At}p = e^{(A+I)(\tau(p)-t)}e^{At}p = e^{-It}e^{(A+I)\tau(p)}p  = e^{-It}h(p).$$
    819     We omit the proof that $h$ is a homeomorphism.
    820 \end{proof}
    821 
    822 Hence, a stable focus and a stable node are topologically conjugate.
    823 
    824 \begin{defin}
    825     We say that $A\in \R^{n\times n}$ is hyperbolic if $\Re(\lambda) \neq 0$ for all $\lambda \in \sigma(A)$.
    826 \end{defin}
    827 
    828 \begin{thm}
    829     For $A, B \in \R^{n\times n}$ hyperbolic matrices, the following statements are equivalent:
    830     \begin{itemize}
    831         \item $s(A) = s(B)$ and $u(A) = u(B)$
    832 
    833         \item $e^{At}$ and $e^{Bt}$ are topologically conjugate.
    834     \end{itemize}
    835 \end{thm}
    836 
    837 \begin{exam}
    838     The matrices
    839     $$A = \mattwo{0}{-2}{2}{0}, \quad B=\mattwo{0}{-1}{1}{0}$$
    840     are not topologically conjugate. To show this, we compute
    841     $$e^{At} = \mattwo{\cos(2t)}{-\sin(2t)}{\sin(2t)}{\cos(2t)}, \quad e^{Bt} = \mattwo{\cos(t)}{-\sin(t)}{\sin(t)}{\cos(t)}.$$
    842     Suppose that there is a homeomorphism $h$ such that
    843     $$h(e^{At}p) = e^{Bt}h(p)$$
    844     for all $p\in\R^2$ and for all $t\in \R$. Setting $t = \pi$ yields $h(p)= -h(p)$ and then $h(p) = 0$ for all $p \in \R^2$, which is a contradiction.
    845 \end{exam}
    846 
    847 \begin{thm}
    848     If $c(A) = c(B) =n$ then $e^{At}$ and $e^{Bt}$ are topologically conjugate if and only if $A$ and $B$ are similar matrices.
    849 \end{thm}
    850 
    851 \begin{thm}
    852     The flows $e^{At}, e^{Bt}$ are topologically conjugate if and only if $s(A) = s(B), u(A) = u(B), c(A) = c(B)$ and $A|_{E^c(A)}, B|_{E^c(B)}$ are similar.
    853 \end{thm}
    854 
    855 \begin{prop}
    856     $A$ and $B$ are $C^1$-conjugate if and only if $A$ and $B$ are similar.
    857 \end{prop}
    858 
    859 \begin{proof}
    860     We need to show both directions.
    861     \begin{itemize}
    862         \item $\Leftarrow$: We just use the transformation matrix as a homeomorphism.
    863 
    864         \item $\Rightarrow$: By definition $h(e^{At}p)=e^{Bt}h(p)$ for all $t,p$. Differentiate with respect to $p$ and set $p = 0$.
    865         $$h'(e^{At}0)e^{At}=e^{Bt}h'(0)$$
    866         $$h'(0)e^{At}=e^{Bt}h'(0),$$
    867         where $h'$ is the Jacobian matrix of $h$. Differentiate with respect to $t$ and set $t=0$.
    868         $$h'(0)Ae^{A0}=Be^{B0}h'(0)$$
    869         Let $P := (h'(0))^{-1}$. Then
    870         $$B=P^{-1}AP.$$
    871     \end{itemize}
    872 \end{proof}
    873 
    874 Hence $C^1$-conjugacy is too strong: e.g.
    875 
    876 $$\dot x = -x$$
    877 and
    878 $$\dot y = -2y$$
    879 are not $C^1$-conjugate.
    880 
    881 \begin{rem}[Related to Routh-Hurwitz]
    882     Given the real polynomial $x^n+b_{n-1}x^{n-1} +\dots + b_0$, where all roots have negative real parts. Then $b_{n-1} > 0, \dots, b_0 > 0$.
    883 \end{rem}
    884 
    885 \begin{proof}
    886     We first consider the simplest cases.
    887     \begin{itemize}
    888         \item $n=1$: The polynomial $x + b_0$ has the root $-b_0$. By assumtion, it has to have a negative real part, so $b_0 > 0$.
    889 
    890         \item $n=2$: The polynomial $x^2+b_1x+b_0=(x-\lambda_1)(x-\lambda_2)$ has the roots $\lambda_1$ and $\lambda_2$, which have a negative real part. Then $b_0 = \lambda_1\lambda_2$ and $b_1 = -\lambda_1-\lambda_2$. Obviously $b_1 > 0$. If $\lambda_1$ and $\lambda_2$ are real, then they are negative and $b_0>0$. Otherwise, they are of the form $\mu\pm i\omega$. In this case, $b_0 = (\mu+i\omega)(\mu-i\omega) = \mu^2+\omega^2 > 0$.
    891 
    892         \item For higher degree, find the decomposition of the polynomial to irreducible factors, which leads to linear and quadratic polynomials. Then each coefficient of the factors is positive. Multiplying those factors together, we get positive coefficients of the whole polynomial.
    893     \end{itemize}
    894 \end{proof}
    895 
    896 \subsection{Stability of equilibria}
    897 
    898 Let the following general differential equation be given:
    899 $$\dot x(t)=f(x(t)),$$
    900 where $f \in C^1(U)$. For $p\in U$ and $ t\mapsto \phi(t,p)$ denote the solution for which $x(0)=p$ holds (on a maximal time interval $I(p)$).
    901 
    902 \begin{defin}
    903     We say that $p\in U$ is an equilibrium if $f(p) = 0$. In this case, $\phi(t,p)$ is constant over time. So $\phi_t(p)=p$, meaning that $p$ is a fixed point of the flow for all $t \in \R$.
    904 \end{defin}
    905 
    906 \begin{defin}
    907     An equilibrium $p$ is Lyapunov stable if for all $\epsilon > 0$ there is a $\delta > 0$ such that $|q-p| < \delta$ implies $|\phi(t,q)-p| < \epsilon$ for all $t\geq 0$. Otherwise $p$ is called unstable.
    908 \end{defin}
    909 
    910 \begin{defin}
    911     An equilibrium is called attracting if there is a neighborhood $U_0 \subseteq U$ of $p$ such that $\lim_{t\to\infty} \phi(t,q)=p$ for all $q\in U_0$.
    912 \end{defin}
    913 
    914 \begin{defin}
    915     An equilibrium is said to be asymptotically stable if it is Lyapunov stable and attracting.
    916 \end{defin}
    917 
    918 \begin{exam}[An attracting equilibrium that is not Lyapunov stable]
    919     Let the following system be given in polar coordinates
    920     $$\begin{array}{ll}\dot r=r(1-r), &\quad r\geq 0,\\
    921     \dot \theta=\sin^2\left(\half\theta\right), &\quad (0\leq \theta < 2\pi).\end{array}$$
    922 
    923     \begin{tikzpicture}
    924             \begin{polaraxis}[
    925                 ymax = 1.5,
    926             ]
    927 
    928             \addplot3[
    929                 samples=12,
    930                 samples y=12,
    931                 quiver = {
    932                     u = {deg(sin(atan2(y,x)/2)^2)},%{veclen(y,x)*(1-veclen(y,x))},
    933     		        v = {veclen(y,x)*(1-veclen(y,x))},%{sin(atan2(y,x)/2)^2},
    934                     scale arrows = 0.5,
    935                 },
    936                 -latex,
    937                 domain=-1.5:1.5,
    938                 domain y=-1.5:1.5,
    939                 data cs=cart
    940             ] (x, y, 0);
    941             \end{polaraxis}
    942         \end{tikzpicture}
    943 
    944     The point $(r^*,\theta^*) = (1,0)$ is not Lyaponov stable, but attracts $\R^2\setminus \{0\}$. In the polar form it attracts all points with $r > 0$ and $0\leq \theta < 2\pi$.
    945 \end{exam}
    946 
    947 \begin{exam}
    948     Given
    949     $$\begin{array}{rcl}\dot x &=& -y-x^3\\
    950     \dot y &=& x-y^3.\end{array}$$
    951     Let $V(x,y):= x^2+y^2, \R^2\to\R$. Taking the derivative with respect to $t$ gives us
    952     $$\frac{d}{dt}V(x,y) = \nabla V(x,y)\cdot\vectwo{\dot x}{\dot y} = 2(x\dot x+y\dot y) =  2(x(-y-x^3)+ y(x-y^3)) = -2(x^4+y^4) < 0.$$
    953 \end{exam}
    954 
    955 In general, $V \in C^1(\Rn,\R)$. When differentiating with respect to $t$ and setting $t=0$ we get
    956 $$\frac{d}{dt} V(x)|_{t=0} = \nabla V(x)\cdot \dot x|_{t=0} = \nabla V(x)\cdot f(x)|_{t=0}.$$
    957 So
    958 $$\dot V(x) := \nabla Vf(x) := \nabla V(x)\cdot f(x), \dot V: U\to \R$$
    959 is the derivative of $V$ with respect to $f$.
    960 
    961 \begin{defin}
    962     Let $p$ be an equilibrium and $U_0 \subseteq U$ an open neighbourhood of $p$. A function $V: U_0\to \R$ is called a Lyapunov function for $p$ if:
    963     \begin{enumerate}
    964         \item $V(p) = 0$ and $V(x) > 0$ for $x \in U_0\setminus\{p\}$.
    965 
    966         \item $V$ is continuous in $U_0$, $C^1$ in $U_0\setminus \{p\}$ and $\dot V(x)\leq 0$ for $x \in U_0\setminus \{p\}$.
    967     \end{enumerate}
    968     If $\dot V(x) < 0$ for all $x\in U_0\setminus\{p\}$, then we call it a strict Lyapunov function for $p$.
    969 \end{defin}
    970 
    971 We can imagine that each orbit moves along $V$ such that the value of $V$ doesn't increase. So as long as the orbit starts close at $p$, it stays close to it.
    972 
    973 \begin{hw}
    974     Consider
    975     $$\dot x = f(x)$$
    976     in $\R$ with an equilibrium $f(p)=0$. Show that $p$ is asymptotically stable if and only if $V(x) = |x-p|$ is a strict Lyapunov function for $p$ in a neighborhood.
    977 \end{hw}
    978 
    979 \begin{hw}
    980     Pendulum with friction: Given
    981     $$\begin{aligned}\dot x&=y\\
    982     \dot y &= \sin x-\delta y\end{aligned}$$
    983     where $\delta \geq 0$. Show that
    984     $$V(x,y) := \half{y^2}-\cos x$$
    985     is a Lyapunov function at the origin.
    986 \end{hw}
    987 
    988 \begin{hw}
    989     For which $A\in\R^{n\times n}$ is the origin of $\dot x = Ax$ Lyapunov stable?
    990 \end{hw}
    991 
    992 \begin{thm}[Lyapunov, 1892]
    993     The following statements hold:
    994     \begin{enumerate}
    995         \item If there is a Lyapunov function for the equilibrium $p$ then $p$ is Lyapunov stable.
    996 
    997         \item If there is a strict Lyaponov function for the equilibrium $p$ then $p$ is asymptotically stable.
    998     \end{enumerate}
    999 \end{thm}
   1000 
   1001 \begin{proof}
   1002     For the proof, we use some $\epsilon$-$\delta$-tricks.
   1003     \begin{enumerate}
   1004         \item Let $\epsilon > 0$ such that
   1005         $$\overline{B_\epsilon(p)} \subseteq U_0.$$
   1006         Let
   1007         $$\alpha := \min_{|x-p|=\epsilon}V(x) > 0.$$
   1008         Clearly, $V(p)<\alpha$. Let $\delta > 0$ such that $|x-p|<\delta$ implies $V(x) < \alpha$ (recall $V$ is continuous).\\
   1009         Then for $q$ with $|q-p| < \delta$ we have $V(\varphi(t,q))< \alpha$ for $t \geq 0$. This is, because $V(\varphi(0,q)) = V(q) < \alpha$ and $\dot V \leq 0$. So the flow will never intersect the sphere $\{x\in\Rn:|x-p| = \epsilon\}$. Therefore we get $|\varphi(t,q)-p|<\epsilon$ for $t\geq 0$.
   1010 
   1011         \item We show that $\lim_{t\to\infty}\varphi(t,q) = p$ if $|q-p|<\delta$. Since $t\mapsto V(\varphi(t,q))$ is monotone decreasing, the following limit exists:
   1012         $$\lim_{t\to\infty}V(\varphi(t,q))=:c \geq 0.$$
   1013         \begin{itemize}
   1014             \item If $c = 0$, then the orbit stays in a compact neighborhood of $p$, since $V$ is Lyapunov stable. Therefore the orbit accumulates at a certain point $r$. Since $c = 0$, the only accumulation point possible is $p$. So $\lim_{t\to\infty} \varphi(t,q) = p$.
   1015 
   1016             \item If $c > 0$ then there is a $\gamma > 0$ such that $|\varphi(t,q)-p|\geq \gamma$ for $t\geq 0$. On the compact set
   1017             $$\{x\in U_0: \gamma \leq |x-p|\leq \epsilon\},$$
   1018             we get that
   1019             $$\dot V(x)\leq -\beta < 0$$
   1020             for every $x$. Therefore $V(\varphi(t,q))\to-\infty$ as $t\to\infty$, which is a contradiction.
   1021         \end{itemize}
   1022     \end{enumerate}
   1023 \end{proof}
   1024 
   1025 \begin{hw}
   1026     Given the following scalar ODE:
   1027     $$\dot x = x^3 \sin^2\left(\reci x\right)$$
   1028     Is the origin Lyapunov stable? Can you find a Lyapunov function?
   1029 \end{hw}
   1030 
   1031 \begin{hw}
   1032     Prove the following statements about the pendulum with friction:
   1033     \begin{itemize}
   1034         \item If $\delta = 0$ then the origin is Lyapunov stable.
   1035 
   1036         \item If $\delta > 0$ then the origin is asymptotically stable.
   1037     \end{itemize}
   1038 \end{hw}
   1039 
   1040 \begin{hw}
   1041     Is the following statement true or false? Give a proof or a counterexample.
   1042     \begin{itemize}
   1043         \item If $V:\R^n\to\R$ is a strict Lyapunov function for the equilibrium $p$, then $\lim_{t\to\infty} \varphi(t,q) = p$ for all $q \in\R^n$.\\
   1044     \end{itemize}
   1045 \end{hw}
   1046 
   1047 \begin{hw}
   1048     Lorenz:\\
   1049     Given the following system of equations:
   1050     $$\begin{array}{rcl}\dot x&=&\sigma(y-x)\\
   1051     \dot y &=& \rho x-y-xz\\
   1052     z &=& xy-\beta z\end{array}$$
   1053     Show that the origin is globally asymptotically stable if $0<\rho < 1$.\\
   1054     Hint: $V(x,y,z) = \rho x^2+\sigma y^2+\sigma z^2$
   1055 \end{hw}
   1056 
   1057 Given
   1058 $$\dot x(t)=f(x(t))$$
   1059 where $f \in C^1(U)$ and $f(p) = 0$. Then
   1060 $$f(x)=f'(p)(x-p)+r(x)$$
   1061 where $f'$ is the Jacobin matrix of $f$.
   1062 $$f'=\matthree{\partial_1f_1}{\dots}{\partial_nf_1}{\vdots}{}{\vdots}{\partial_1f_n}{\dots}{\partial_nf_n}=(\partial_jf_i)_{i,j=1}^n$$
   1063 Also $r(p) = 0$ and $\lim_{x\to\infty}\frac{r(x)}{|x-p|}=0$. Let $A := f'(p)$.
   1064 \begin{thm}
   1065     If $A$ is stable then $p$ is asymptotically stable for $\dot x=f(x)$.
   1066 \end{thm}
   1067 
   1068 \begin{proof}
   1069     Assume without loss of generality that $p=0$. Let $Q = Q^T > 0$ such that $QA+A^TQ<0$.\\
   1070     We claim that $V(x)=x^TQx$ is a strict Lyapunov function for $p$. So we need to show that $\dot V(x) < 0$ for $x \neq p$.\\
   1071     $$\dot V(x) = \nabla V(x)\cdot f(x)=2x^TQ (Ax+r(x))$$
   1072     Because $x^TQAx$ is a scalar, we can transpose it, to get to $x^TA^TQx$. Then we get
   1073     $$\dot V(x) = x^T(QA+A^TQ)x + 2x^T Q r(x).$$
   1074     We define the inner product
   1075     $$\jbr{x,y}:=x^TQy$$
   1076     with the induced norm
   1077     $$\|x\|^2 := \jbr{x,x}=x^TQx.$$
   1078     Now let's calculate the following quotient:
   1079     $$\frac{\dot V(x)}{V(x)}=\frac{x^T(QA+A^TQ)x}{x^TQx}+2\frac{\jbr{x,r(x)}}{\|x\|^2}$$
   1080     The first term is smaller than or equal $-c$ for $c := \min_{\|x\|=1}\left(\frac{-x^T(QA+A^TQ)x}{x^TQx}\right)>0$. For the second term:
   1081     $$\frac{\jbr{x,r(x)}}{\|x\|^2}\leq \frac{\|x\|\|r(x)\|}{\|x\|^2} = \frac{\|r(x)\|}{\|x\|} \to 0$$
   1082     as $x \to 0$. Therefore, $\frac{\dot{V}}{V}$ is negative in a neighborhood of the origin. So for $x \neq 0$ close to $p$, $\dot V(x) < 0$ so $V$ is a Lyapunov function.
   1083 \end{proof}
   1084 
   1085 \begin{hw}
   1086     Show that the origin is asymptotically stable for the pendulum with friction $(\delta > 0)$ (linearize).
   1087 \end{hw}
   1088 
   1089 \begin{defin}
   1090     An equilibrium $p$ is hyperbolic if $f'(p) \in \R^{n\times n}$ has no eigenvalue with zero real part.
   1091 \end{defin}
   1092 
   1093 \begin{thm}[Hartman-Grobman]
   1094     If $f$ is $C^1$ and $p$ is a hyperbolic equilibrium then the local flow $\varphi_t$ is topologically conjugate to the local flow of
   1095     $$\dot y = f'(p)y$$
   1096     at the origin. (There is an open neighborhood $U_0$ of $p$, an open neighborhood $U_1$ of $0 \in \R^n$ and a homeomorphism $h:U_0 \to U_1$ such that
   1097     $$h\circ \varphi_t = e^{At}\circ h$$
   1098     (here $A := f'(p)$) where this is defined ($\varphi_t(x) \in U_0$).)\\
   1099     (Simon, Perko, Teschl)
   1100 \end{thm}
   1101 
   1102 \begin{defin}
   1103     Let $p$ be an equilibrium and $A = f'(p)$.
   1104 
   1105     \begin{itemize}
   1106         \item We say that $p$ is a sink if $\Re(\lambda) < 0$ for all $\lambda \in \sigma(A)$.
   1107 
   1108         \item We say that $p$ is a source if $\Re(\lambda)>0$ for all $\lambda \in \sigma(A)$.
   1109 
   1110         \item We say that $p$ is a saddle if it is hyperbolic, meaning th there is a $\lambda_1 \in \sigma(A)$ with $\Re(\lambda_1)<0$ and $\lambda_2 \in \sigma(A)$ with $\Re(\lambda_2)> 0$.
   1111     \end{itemize}
   1112 \end{defin}
   1113 
   1114 \begin{hw}
   1115     Classify all of the equilibria of
   1116     $$\begin{aligned}\dot x&=x^2-y^2-1\\
   1117     \dot y &= 2y.\end{aligned}$$
   1118 \end{hw}
   1119 
   1120 \begin{exam}
   1121     Show that the flows of
   1122     $$\begin{aligned}\dot x &= -x\\
   1123     \dot y &= y+x^2\end{aligned}$$
   1124     and
   1125     $$\begin{aligned}\dot u &= -u\\
   1126     \dot v &= v\end{aligned}$$
   1127     are topologically conjugate (not only locally, but even globally).\\
   1128     Let
   1129     $$h:(x,y)\mapsto (x,y+\frac{x^2}{3}).$$
   1130     Then
   1131     $$h^{-1}: (u,v)\mapsto (u,v-\frac{u^2}{3})$$ and $h$ is a homeomorphism. Let
   1132     $$(u,v) = h(x,y)$$
   1133     or
   1134     $$\begin{aligned}u&=x\\
   1135     v&=y+\frac{x^2}{3}\end{aligned}.$$
   1136     Then
   1137     $$\begin{aligned}\dot u &= \dot x=-x = -u\\
   1138     \dot v &= \dot y + \frac{2}{3}x\dot x=y+x^2+\frac{2}{3}(-x^2)= y+\frac{x^2}{3}=v.\end{aligned}$$
   1139 \end{exam}
   1140 
   1141 \subsection{Polar coordinates}
   1142 
   1143 If there is a system in Cartesian form like
   1144 $$\begin{aligned}\dot x &= f(x,y)\\
   1145 \dot y &= g(x,y)\end{aligned}$$
   1146 we can get to a system in polar form like
   1147 $$\begin{aligned}\dot r = F(r,\theta)\\
   1148 \dot \theta = G(r,\theta)\end{aligned}$$
   1149 where
   1150 $$\begin{aligned}&x = r\cos(\theta), &y = r\sin(\theta)\\
   1151 &r = \sqrt{x^2+y^2}, &\theta = \arctan(y/x).\end{aligned}$$
   1152 Using some derivative tricks, we get
   1153 $$\dot x = \dot r \cos(\theta) +r(-\sin(\theta))\dot \theta, \quad \dot y = \dot r \sin(\theta) + r \cos(\theta) \dot \theta,$$
   1154 $$\dot r = \frac{2x\dot x +2y \dot y}{2\sqrt{x^2+y^2}}, \quad \dot \theta = \frac{\frac{\dot yx - y\dot x}{x^2}}{1+\frac{y^2}{x^2}} = \frac{\dot yx-y\dot x}{x^2+y^2},$$
   1155 
   1156 $$\dot x=-y\dot \theta +x\frac{\dot r}{r}, \quad
   1157 \dot y = x\dot \theta +y\frac{\dot r}{r},$$
   1158 $$\dot r = \frac{x\dot x + y\dot y}{r}, \quad
   1159 \dot \theta = \frac{x\dot y-\dot xy}{r^2}.$$
   1160 
   1161 Linearization at a nonhyperbolic equilibrium (these are situations, where the Hartman-Grobman theorem does not apply):
   1162 
   1163 \begin{exam}
   1164     1D
   1165     \begin{itemize}
   1166         \item Given
   1167         $$\dot x = -x^3,$$
   1168         the equilibrium $0$ attracts from both sides.
   1169 
   1170         \item Given
   1171         $$\dot x = x^3,$$
   1172         the equilibrium $0$ repells.
   1173 
   1174         \item Given
   1175         $$\dot x = x^2,$$
   1176         the flow goes to the right.
   1177 
   1178         \item Given
   1179         $$\dot x = 0,$$
   1180         everything is stationary. (This is the linearization of all three equations above.)
   1181     \end{itemize}
   1182 
   1183 \end{exam}
   1184 
   1185 \begin{exam}
   1186     2D
   1187     \begin{itemize}
   1188         \item Given
   1189         $$\begin{aligned}\dot r &= -r^3\\
   1190         \dot \theta &= 1\end{aligned}$$
   1191         in polar coordinates or
   1192         $$\begin{aligned}\dot x &= -y-xy^2-x^3\\
   1193         \dot y &= x-y^3-x^2y\end{aligned}$$
   1194         in Cartesian coordinates, the flow spirals inwards. So the origin is asymptotically stable.
   1195 
   1196         \item Given
   1197         $$\begin{aligned}\dot r &= r^3\\
   1198         \dot \theta &= 1\end{aligned}$$
   1199         in polar coordinates or
   1200         $$\begin{aligned}\dot x &= -y+xy^2-x^3\\
   1201         \dot y &= x+y^3+x^2y\end{aligned}$$
   1202         in Cartesian coordinates, the flow spirals outwards. So the origin is repelling.
   1203 
   1204         \item Given
   1205         $$\begin{aligned}\dot r &= 0\\
   1206         \dot \theta &= 1+r\cos(\theta)\end{aligned}$$
   1207         in polar coordinates or
   1208         $$\begin{aligned}\dot x &= -y-xy\\
   1209         \dot y &= x+x^2\end{aligned}$$
   1210         in Cartesian coordinates, we get a center at the origin.
   1211 
   1212         \item All three examples above have the same linearization:
   1213         $$\begin{aligned}\dot x &= -y\\
   1214         \dot y &= x\end{aligned}.$$
   1215         Here the flow moves in circles around the origin.
   1216     \end{itemize}
   1217 \end{exam}
   1218 
   1219 \subsection{Asymptotic behavior}
   1220 
   1221 What do solutions do as $t \to \infty$? Given
   1222 $$\dot x(t) = f(x(t))$$
   1223 where $f \in C^1(U, \Rn)$ and $t \mapsto \varphi(t,p)$ is the solution with $x(0) = p$.
   1224 
   1225 \begin{defin}
   1226     For $p \in U$, the $\omega$-limit of $p$ is
   1227     $$\omega(p)=\{q\in U: \exists (t_n)_{n\in\N} \subseteq \R \text{ such that } \lim_{n\to\infty}t_n = \infty \text{ and } \lim_{n\to\infty}\varphi(t_n,p)=q\}.$$
   1228     For $p\in U$, the $\alpha$-limit of $p$ is
   1229     $$\alpha(p) = \{q\in U:\exists (t_n)_{n\in\N} \subseteq \R \text{ such that }\lim_{n\to\infty}t_n=-\infty \text{ and } \lim_{n\to\infty}\varphi(t_n,p)=q\}.$$
   1230 \end{defin}
   1231 
   1232 \begin{exam}
   1233 Let's do some examples.
   1234 \begin{enumerate}
   1235     \item Given:
   1236     $$\begin{aligned}\dot x&=x\\
   1237     \dot y &= -y.\end{aligned}$$
   1238     \begin{itemize}
   1239         \item If $p = 0$ then $\alpha(p) = \{0\}$ and $\omega(p)=\{0\}.$
   1240 
   1241         \item If $p \neq 0$ is on the $x$-axis then $\alpha(p)=\{0\}$ and $\omega(p)=\emptyset$.
   1242 
   1243         \item If $p \neq 0$ is on the $y$-axis then $\alpha(p) = \emptyset$ and $\omega(p) = \{0\}$.
   1244 
   1245         \item If $p$ is on no axis then $\alpha(p) = \emptyset$ and $\omega(p) = \emptyset$.
   1246     \end{itemize}
   1247 
   1248     \item Given:
   1249     $$\begin{aligned}\dot x &= -y+x(1-x^2-y^2)\\
   1250     \dot y &= x+y(1-x^2-y^2)\end{aligned}$$
   1251     or
   1252     $$\begin{aligned}\dot r &= r(1-r^2)\\
   1253     \dot \theta &= 1\end{aligned}$$
   1254 
   1255     \begin{itemize}
   1256         \item If $p = 0$ then $\alpha(p) = \{0\}$ and $\omega(p) = \{0\}$.
   1257 
   1258         \item If $p \neq 0$ is inside the unit circle then $\alpha(p) = \{0\}$ and $\omega(p)$ is the boundary of the unit circle.
   1259 
   1260         \item If $|p| = 1$ then $\alpha(p)$ and $\omega(p)$ are the boundary of the unit circle.
   1261 
   1262         \item If $p$ is outside the unit circle then $\alpha(p)=\emptyset$ and $\omega(p)$ is the boundary of the unit circle.
   1263     \end{itemize}
   1264 
   1265 
   1266 
   1267     \item Given:
   1268     $$\begin{aligned}\dot x &= -y+x(1-x^2-y^2)\\
   1269     \dot y &= x+y(1-x^2-y^2)\\
   1270     \dot z &= \beta > 0 \text{ (constant)} \end{aligned}$$
   1271     Then for every $p \in \R^3$ we get $\alpha(p) = \omega(p) = \emptyset$.
   1272 
   1273     \item Same as 3, but identify $0$ and $2\pi$ on the $z$-axis. Then the invariant cylinder becomes an invariant torus. If $\beta\in\Q$ then the torus is filled with periodic orbits (and the $\omega$-limit is one of these periodic orbits, provided $p$ is not on the $z$-axis).\\
   1274     If $\beta \notin \Q$ then there are everywhere dense orbits on the torus and $\omega(p)$ is the torus for all $p$ not on the $z$-axis.
   1275 
   1276     \item Lorenz equation: Strange attractor.
   1277 \end{enumerate}
   1278 \end{exam}
   1279 
   1280 \begin{thm}
   1281     The following statements are true:
   1282     \begin{enumerate}
   1283         \item $\omega(p)$ is closed.
   1284 
   1285         \item $\omega(p)$ is invariant (if $q \in \omega(p)$ then $\varphi(t,q) \in \omega(p)$ for all $t \in \mathbb{R}$).
   1286 
   1287         \item If $\{\varphi(t,p):t\geq 0\}$ is bounded then $\omega(p)$ is nonempty an connected.
   1288     \end{enumerate}
   1289 \end{thm}
   1290 
   1291 \begin{proof}
   1292     Let's go through the proof.
   1293     \begin{enumerate}
   1294         \item Let $q\notin \omega(p)$. Then there is an open neighborhood $U_0\subseteq U$ of $q$ and $T > 0$, such that $\varphi(t,p)\notin U_0$ for all $t \geq T$. So after time $T$, the solution does not enter $U_0$. Therefore $U_0 \subseteq \omega(p)^c$ implying that $\omega(p)^c$ is open and $\omega(p)$ is closed.
   1295 
   1296         \item Fix $q \in \omega(p)$. Then there is a sequence $(t_n)_{n\in\N} \subseteq \R$, such that $\lim_{n\to\infty}t_n = \infty$ and $\lim_{n\to \infty}\varphi(t_n,p) = q$.\\
   1297         Fix $t\in\R$ and let $\tau_n = t_n+t$. Then $\lim_{n\to\infty}\tau_n=\infty$ and $$\lim_{n\to\infty}\varphi(\tau_n,p) = \lim_{n\to\infty}\varphi(t_n+t,p) = \lim_{n\to\infty}\varphi(t,\varphi(t_n,p))=\varphi(t,q)$$
   1298         (because of the continuous dependence of solutions on the initial conditions). Therefore, we get $\varphi(t,q) \in \omega(p)$.
   1299 
   1300         \item The solution lies in a compact set $K$. Take any sequence $(t_n)_{n\in\N}\subseteq\R$ such that $\lim_{n\to\infty}t_n=\infty$. Then $(\varphi(t_n,p))_{n\in\N} \subseteq K$ and there is a subsequence $(\varphi(t_{n_k},p))_k$ such that $\lim_{k\to\infty}\varphi(t_{n_k},p) \in K$ exists. Therefore $\omega(p) \neq \emptyset$.\\
   1301         Now we show that $\omega(p)$ is connected. By way of contradiction, assume there exist $G_1, G_2$ open sets such that:
   1302         \begin{itemize}
   1303             \item $G_1$ and $G_2$ are disjoint.
   1304 
   1305             \item $G_1 \cup G_2 \supseteq \omega(p)$.
   1306 
   1307             \item $G_1 \cap \omega(p) \neq \emptyset$ and $G_2 \cap \omega(p) \neq \emptyset$.
   1308         \end{itemize}
   1309 
   1310         Then there are sequences $(t_n)_{n\in\N}, (\tau_n)_{n\in\N} \subseteq \R$ such that
   1311         $\lim_{n\to\infty} t_n=\infty$, $\lim_{n\to\infty}\tau_n=\infty$,
   1312         $$\lim_{n\to\infty}\varphi(t_n,p)=q_1 \in G_1 \cap \omega(p)$$
   1313         and
   1314         $$\lim_{n\to\infty}\varphi(\tau_n,p)=q_2  \in G_2 \cap \omega(p),$$
   1315         where $t_1 < \tau_1 < t_2 < \tau_2 < \dots$.\\
   1316         Let $\theta_n$ be in the interval $(t_n,\tau_n)$ such that $\varphi(\theta_n,p)\notin G_1 \cup G_2$. Then as above $(\varphi(\theta_n,p))_{n\in\N}$ has a convergent subsequence, thus we found a point in $\omega(p) \setminus(G_1 \cup G_2)$ which is a contradiction.
   1317     \end{enumerate}
   1318 \end{proof}
   1319 
   1320 \begin{hw}
   1321     Sketch a 2D example with nonempty and disconnected $\omega$-limit set.
   1322 \end{hw}
   1323 
   1324 In a one-dimensional system the $\omega$-limit of a point $p$ is either $\emptyset$ or $\{q\}$, where $q$ is an equilibrium. What happens in the two-dimensional case?
   1325 
   1326 \begin{thm}[Poincar\'e-Bendixson]
   1327     Let $U \subseteq \R^2$ be open. Let
   1328     $$\dot x(t)=f(x(t))$$
   1329     be given where $f\in C^1(U,\R^2)$. Let $K \subseteq U$ be a forward invariant ($p \in K$ implies $\varphi(t,p)\in K$ for all $t \geq 0$), compact set with finitely many equilibria. Then for any $p \in K$ the set $\omega(p)$ is one of the following:
   1330     \begin{itemize}
   1331         \item $\omega(p) = \{q\}$, where $q$ is an equilibrium.
   1332 
   1333         \item $\omega(p) = \gamma$, where $\gamma$ is a closed orbit (the image of a periodic solution).
   1334 
   1335         \item $\omega(p)$ is the union of $\{q_1, \dots ,q_n\}$ and countably many non closed orbits $\gamma$ such that $\alpha(\gamma) = q_i$ and $\omega(\gamma)=q_j$.
   1336     \end{itemize}
   1337 \end{thm}
   1338 
   1339 \begin{defin}
   1340     If $q$ is an equilibrium and $\gamma$ is an orbit with $\lim_{t \to \pm\infty}\gamma(t) = q$, then $\gamma$ is called a homoclinic orbit.
   1341 \end{defin}
   1342 
   1343 \begin{defin}
   1344     If $q_1, q_2$ are two different equilibria and $\gamma$ is an orbit with $\lim_{t\to-\infty}\gamma(t) = q_1$ and $\lim_{t\to\infty}\gamma(t) = q_2$, then $\gamma$ is called a heteroclinic orbit.
   1345 \end{defin}
   1346 
   1347 \begin{exam}
   1348     The system
   1349     $$\begin{aligned}\dot x &= x(1-x)(x-y)\\
   1350     \dot y &= y(1-y)(2x-1)\end{aligned}$$
   1351     has an attractin rectangle.
   1352 \end{exam}
   1353 
   1354 \begin{rem}
   1355     The Poincar\'e-Bendixson-Theorem does not apply to 2D manifolds other than subsets of $\R^2$ or the $2$-sphere.
   1356 \end{rem}
   1357 
   1358 For instance, let
   1359 $$\dot x = v$$
   1360 for $v \in \R^2$ such that $\frac{v_1}{v_2}\notin \Q$ in the $2$-torus $\mathbb{T}^2$.
   1361 $$x(t)=tv+p$$
   1362 is dense in $\mathbb{T}^2$ and $\omega(p) = \mathbb{T}^2$.
   1363 
   1364 \begin{defin}
   1365     Given
   1366     $$\begin{aligned}
   1367         \dot x_1 = f_1(x)\\
   1368         \dot x_2 = f_2(x)\\
   1369         \vdots\\
   1370         \dot x_n = f_n(x)
   1371     \end{aligned}$$
   1372     for $f = (f_1, \dots, f_n) \in C(U)$. For $i \in \jbr{n}$, we define $\{x \in U: f_i(x) = 0\}$ the $x_i$-nullcline.
   1373 \end{defin}
   1374 
   1375 \paragraph{Chlorine dioxide-Iodine-Malonic-Acid reaction: $(X = I, Y = ClO_2^-)$}
   1376 Given
   1377 $$\begin{aligned} \dot x &= a-x-\frac{4xy}{x^2+1}\\
   1378 \dot y &= bx\left(1-\frac{y}{x^2+1}\right),\end{aligned}$$
   1379 where $a,b > 0$ and $x,y\geq 0$. We first compute the nullclines.
   1380 $$\begin{aligned}\dot x = 0 &\Leftrightarrow y = \frac{(a-x)(x^2+1)}{4x}\\
   1381 \dot y = 0 &\Leftrightarrow y = x^2+1\end{aligned}$$
   1382 Solving this system of equation gives
   1383 $$x^* = \frac a5,\quad y^* = 1+(x^*)^2.$$
   1384 Now we compute the Jacobian matrix $J$ at $(x^*,y^*)$.
   1385 $$J (x^*,y^*) = \reci{(x^*)^2+1}\mattwo{3(x^*)^2-5}{-4x^*}{2b(x^*)^2}{-bx^*}$$
   1386 To get information about the eigenvalues, we calculate the determinant and the trace of it.
   1387 $$\det J(x^*,y^*)=\frac{5bx^*}{(x^*)^2+1}>0$$
   1388 $$\tr J(x^*,y^*)=\frac{3(x^*)^2-bx^*-5}{(x^*)^2+1}$$
   1389 $$\sgn(\tr J(x^*,y^*)) = \sgn(3a^2-5ab-125)$$
   1390 \begin{itemize}
   1391     \item For $b > \frac35a-\frac{25}a$ it is asymptotically stable.
   1392 
   1393     \item For $b < \frac35a-\frac{25}a$ it is repelling, and, by the Poincar\'e-Bendixson Theorem, there is a closed orbit.
   1394 
   1395     \item At $b = \frac35a-\frac{25}a$ a supercritical Andronov--Hopf bifurcation occurs (such a bifurcation will be discussed on May 8th).
   1396 \end{itemize}
   1397 
   1398 \begin{defin}
   1399     A limit cycle is a closed orbit $\gamma$ which is the $\omega$-limit or $\alpha$-limit of at least one point outside $\gamma$.
   1400 \end{defin}
   1401 
   1402 \begin{defin}
   1403     A periodic attractor is a closed orbit such that for all $p$ in a neighborhood of $\gamma$, we get
   1404     $$\omega(p)=\gamma.$$
   1405     (Note: a neighborhood of $\gamma$ in $\R^2$ is an annulus, in $\R^n$ it is a torus.)
   1406 \end{defin}
   1407 
   1408 \begin{hw}
   1409     Given
   1410     $$\begin{aligned}
   1411         \dot x &= y+xh(r)\\
   1412         \dot y &= x+yh(r)\\
   1413         r &= \sqrt{x^2+y^2},
   1414     \end{aligned}$$
   1415     write in polar coordinates and find $h$ such that there are infinitely many limit cycles.
   1416 \end{hw}
   1417 
   1418 \begin{thm}[Green's theorem]
   1419     Let $F:\R^2\to\R^2$ be smooth, $D\subseteq \R^2$ be a simply connected region, $\gamma$ be its boundary and
   1420     $n: \gamma \to\R^2$ is the outward facing unit normal vector field. Then
   1421     $$\int_\gamma F\cdot n=\iint_D\divg F.$$
   1422     With $F=(P,Q)$ this reads as
   1423     $$\oint_\gamma Pdy-Qdx = \iint_D\partial_x P+\partial_y Q.$$
   1424 
   1425 \end{thm}
   1426 
   1427 \begin{thm}[Bendixson-Dulac criterion]
   1428     Given
   1429     $$\begin{aligned}
   1430         \dot x = f(x,y)\\
   1431         \dot y = g(x,y)
   1432     \end{aligned}$$
   1433     where $(f,g):D \to\R^2$ is $C^1$ and $D \subseteq \R^2$ is a simply connected region.\\
   1434     If there is an $h:\R^2\to\R$ which is $C^1$ such that $\text{div}(hf,hg) > 0$ in $D$ (or $\text{div}(hf,hg)<0$ in $D$) then there is no closed orbit that lies entirely in $D$.
   1435 \end{thm}
   1436 
   1437 \begin{proof}
   1438     Suppose by contradiction that there is a closed orbit $\gamma$ (with period $T$) with $\gamma(t) = (x(t),y(t))$. Then
   1439     $$0 < \iint_{\text{int} \gamma} \text{div}(hf,hg) \stackrel{\text{Green}}{=} \oint_\gamma hfdy-hgdx = \int_0^T (hf\dot y-hg\dot x)dt  = \int_0^T (hfg-hgf)dt = 0$$
   1440     which is a contradiction.
   1441 \end{proof}
   1442 
   1443 \begin{hw}
   1444     Rule out closed orbits.
   1445     \begin{enumerate}
   1446         \item Given
   1447         $$\begin{aligned}
   1448             \dot x &= y\\
   1449             \dot y &= -\sin(x)-\delta y
   1450         \end{aligned}$$
   1451         where $\delta > 0$ and $(x,y)\in\R^2$.
   1452 
   1453         \item Given
   1454         $$\begin{aligned}
   1455             \dot x &= x(1-x)(x-y)\\
   1456             \dot y &= y(1-y)(2x-1)
   1457         \end{aligned}$$
   1458         where $(x,y) \in (0,1)^2$.
   1459 
   1460         \item The intraspecific competition is decribed by
   1461         $$\begin{aligned}
   1462             \dot x &= x(n-ax+by)\\
   1463             \dot y &= y(s+cx-dy)
   1464         \end{aligned}$$
   1465         where $a,d > 0$ and $(x,y)\in\R_+^2$.
   1466 
   1467         \item Given
   1468         $$\ddot x + p(x)\dot x+q(x)=0$$
   1469         where $p(x) > 0$ for all $x \in \R$.
   1470     \end{enumerate}
   1471 \end{hw}
   1472 
   1473 \paragraph{Hilbert's 16th problem:}
   1474 Given
   1475 $$\begin{aligned}
   1476     \dot x = f(x,y)\\
   1477     \dot y = g(x,y)
   1478 \end{aligned}$$
   1479 where $f,g$ are polynomials of degree $n$. For a fixed $n$, find the maximum number of limit cycles, and denote this as $H(n)$. It is known that any fixed planar polynomial ODE has only finitely many limit cycles. However, it still not known whether $H(2)$ is finite. It is known that $H(2) \geq 4$, and in fact, people conjecture that $H(2)=4$. Further, there is a cubic example with 13 limit cycles, and hence we know that $H(3)\geq 13$.
   1480 
   1481 \subsection{LaSalle's invariance principle}
   1482 
   1483 Given
   1484 $$\dot x = f(x),$$
   1485 where $f \in C^1(U, \Rn)$.
   1486 
   1487 \begin{lem}
   1488     Let $V \in C(U,\R)$ such that $V(\varphi(\cdot, p))$ is monotone increasing (or decreasing). Then $V$ is constant on $\omega(p)$.
   1489     \label{Constant on orbit}
   1490 \end{lem}
   1491 
   1492 \begin{proof}
   1493     Let $q,r \in \omega(p)$. Then there are sequences $(t_n)_{n\in\N}\subseteq \R$ and $(\tau_n)_{n\in\N} \subseteq \R$ such that $\lim_{n\to\infty} t_n = \infty, \lim_{n\to\infty}\tau_n = \infty, \limn \varphi(t_n,p)=q$ and $ \limn \varphi(\tau_n,p)=r$. By refining the sequences, we may assume that $t_n \leq \tau_n \leq t_{n+1}$ for every $n$. Hence,
   1494     $$V(\varphi(t_n,p)) \leq V(\varphi(\tau_n,p)) \leq V(\varphi(t_{n+1},p))$$
   1495     for every $n$. Letting $n$ approach $\infty$, we get $V(q) \leq V(r) \leq V(q)$. So $V(r)=V(q)$, showing that $V$ is constant on $\omega(p)$.
   1496 \end{proof}
   1497 
   1498 \begin{thm}
   1499     Let $V \in C^1(U,\R)$ such that $\nabla Vf(x)\leq 0$ for all $x \in U$ (or $\geq 0$). Then
   1500     $$\omega(p)\subseteq \{x\in U:\nabla Vf(x)=0\}$$
   1501     for all $p \in U$. Moreover, $\omega(p)$ is contained in the maximal invariant subset of $\{x\in U:\nabla Vf(x)=0\}$.
   1502 \end{thm}
   1503 
   1504 \begin{proof}
   1505     Suppose $q \in \omega(p)$. Then $\varphi(t,q) \in \omega(p)$ for all $t \geq 0$. By assumtion, the map $t \mapsto V(\varphi(t,q))$ monotone decreasing. By lemma \ref{Constant on orbit} it is constant. By differentiating with respect to time, we get
   1506     $$\nabla V(\varphi(t,q))\partial_t\varphi(t,q) = 0.$$
   1507     Because $\dot x = f(x)$, we get
   1508     $$\nabla V(\varphi(t,q))f(\varphi(t,q)) = 0$$
   1509     and setting $t=0$, we get $\nabla Vf(q)=0$.
   1510 \end{proof}
   1511 
   1512 \begin{rem}
   1513     Such a function is called a Lyapunov function. ($\nabla Vf \leq 0$ or $\nabla Vf \geq 0$.)
   1514 \end{rem}
   1515 
   1516 \begin{thm}
   1517     If $V:\Rn \to \R$ is a strict Lyapunov function for the equilibrium $p$ and $V$ is radially unbounded (meaning $V(x) \to \infty$ if $|x| \to \infty$; hence, the sublevelsets of $V$ are bounded) then $\limn\varphi(t,q) = p$ for all $q \in \Rn$ (i.e., $p$ is globally asymptotically stable).
   1518 \end{thm}
   1519 
   1520 \begin{exam}[Pendulum with friction]
   1521     Given
   1522     $$\begin{aligned}
   1523         \dot x &= y\\
   1524         \dot y &= -\sin(x)-\delta y
   1525     \end{aligned}$$
   1526     where $\delta > 0$. We define $V(x,y) = \half {y^2}-\cos(x)$ to be our candidate. Then
   1527     $$\nabla Vf(x,y) = \sin(x)\cdot y + y(-\sin(x)-\delta y)=-\delta y^2 \leq 0.$$
   1528     So
   1529     $$\{(x,y)\in\R^2:\nabla Vf(x,y)=0\} = x-\text{axis}.$$
   1530     The maximal invariant subset of it is $\{(\xi \pi,0), \xi \in \Z\}$. Therefore every bounded solution converges to a point here.
   1531 \end{exam}
   1532 
   1533 \begin{exam}
   1534     Given
   1535     $$\begin{aligned}
   1536         \dot x&= x\\
   1537         \dot y &= -y,
   1538     \end{aligned}$$
   1539     we choose $V(x,y) = x^2-y^2$ as our candidate. Then
   1540     $$\dot V(x,y) = 2x\dot x-2y\dot y = 2(x^2+y^2) \geq 0.$$ So $\dot V(x,y) = 0$ if and only if $(x,y)=0$. Hence, $\omega(p) = \{(0,0)\}$ or $\omega = \emptyset$. The former holds if $p$ is on the $y-$axis, while the latter holds when $p$ is outside the $y$-axis.
   1541 \end{exam}
   1542 
   1543 \subsection{Hamiltonian systems in 2D}
   1544 
   1545 Let $U \subseteq R^2$ be open, $H \in C^2(U, \R)$. We analyze systems of the form
   1546 $$\begin{aligned}
   1547     \dot x &= \partial_y H\\
   1548     \dot y &= -\partial_x H.
   1549 \end{aligned}\quad (H)$$
   1550 
   1551 \begin{thm}[Conservation of energy]
   1552     $H$ is constant along trajectories.
   1553 \end{thm}
   1554 
   1555 \begin{proof}
   1556     We just need to show that the time derivative vanishes.
   1557     $$\frac{d}{dt} H(x(t),y(t)) = \partial_x H\dot x +\partial_y H \dot y = \partial_x H \partial_y H +\partial_y H (-\partial_x H)=0$$
   1558 \end{proof}
   1559 
   1560 \begin{rem}
   1561     Equilibria of $H$ are exactly the critical points of $H$ (where the gradient of $H$ vanishes).
   1562 \end{rem}
   1563 
   1564 \begin{defin}
   1565     An equilibrium $p$ of $\dot x = f(x)$ in $\Rn$ is said to be degenerate if $0 \in \sigma(f'(p))$. Otherwise, it is called non-degenerate.
   1566 \end{defin}
   1567 
   1568 \begin{rem}
   1569     For $n = 2$, a non-degenerate equilibrium is either hyperbolic or a center of the linearized system.
   1570 \end{rem}
   1571 
   1572 \begin{defin}
   1573     An equilibrium $p$ of $\dot x = f(x)$ on $\R^2$ is called
   1574     \begin{itemize}
   1575         \item a center if all nearby orbits are closed.
   1576 
   1577         \item a saddle if there are trajectories $\Gamma_1, \Gamma_2$ that approach $p$ as $t\to\infty$, $\Gamma_3, \Gamma_4$ that approach $p$ as $t\to-\infty$ and there is a $\delta>0$ such that all other trajectories in $B_\delta(p)\setminus\{p\}$ leave $B_\delta(p)$ as $t \to \pm\infty$.
   1578     \end{itemize}
   1579 \end{defin}
   1580 
   1581 \begin{thm}
   1582     Any non-degenerate equilibrium of $(H)$ is either a saddle (when $p$ is saddle of $H$) or a center (if $p$ is a local extremum of $H$).
   1583 \end{thm}
   1584 
   1585 \begin{proof}
   1586     Without loss of generality the equilibrium is at the origin. The linearization is $\vectwo{\dot x}{\dot y}=A\vectwo xy$, where
   1587     $$A = \mattwo{\partial_x\partial_y H}{\partial_y^2H}{-\partial_x^2H}{-\partial_y\partial_x H}.$$
   1588     Then $\tr A = 0$ and $\det A= -(\partial_x\partial_yH)^2+\partial_y^2 H\partial_x^2 H$. The Hessian $\mathcal H$ of $H$ is
   1589     $$\mathcal H = \mattwo{\partial_x^2H}{\partial_x\partial_y H}{\partial_x\partial_yH}{\partial_y^2H}$$
   1590     Then $\det \mathcal H = \det A$. Since the equilibrium is non-degenerate, we have $\det\mathcal H \neq 0$. Then we get two cases.
   1591     \begin{enumerate}
   1592         \item In the first case, we have $\det \mathcal H < 0$ ($\mathcal H$ is indefinite). Therefore it's a saddle of $H$ and a saddle of the linearized system. So the equilibrium is a saddle.
   1593 
   1594         \item In the other case, we have $\det \mathcal H > 0$ ($\mathcal H$ is positive or negative definite). Therefore it's a strict local extremum of $H$. Then the level sets are closed curves. Since $H$ is constant along solutions, the equilibrium is a center.
   1595     \end{enumerate}
   1596 
   1597 \end{proof}
   1598 
   1599 \subsection{Special Hamiltonian systems: Newtonian systems}
   1600 
   1601 Given
   1602 $$\ddot x = F(x)$$
   1603 in $\R$ (scalar ODE, second order). We can use the substitution $y:= \dot x$ to get two ODEs of order one
   1604 
   1605 
   1606 $$\begin{aligned}
   1607     \dot x &= y\\
   1608     \dot y &= F(x).
   1609 \end{aligned}\quad (N)$$
   1610 (planar ODE, first order). Now let
   1611 $$H(x,y) = \half{y^2}-\int_{x_0}^x F(s)ds$$
   1612 In this case, $H(x,y)$ is called the total energy and $\half{y^2}$ is the kinetic energy. The rest is the potential energy, so
   1613 $$H(x,y) = T(y)+U(x).$$
   1614 
   1615 \begin{thm}
   1616     The following statements hold:
   1617     \begin{enumerate}
   1618         \item The equilibria of $(N)$ all lie on the $x$-axis.
   1619 
   1620         \item The point $(x^*,0)$ is an equilibrium of $(N)$ if and only if $x^*$ is a critical point of $U$ ($\Leftrightarrow$ $F(x^*) = 0$).
   1621 
   1622         \item If $x^*$ is a strict local max/min of $U$ then $(x^*,0)$ is a saddle/center of $(N)$.
   1623 
   1624         \item The phase portrait of $(N)$ is symmetric to the $x$-axis.
   1625     \end{enumerate}
   1626 \end{thm}
   1627 
   1628 \begin{exam}
   1629     Given
   1630     $$\ddot x +\sin(x) = 0,$$
   1631     so $F(x) = -\sin(x)$. By substitution, we get
   1632     $$\begin{aligned}
   1633         \dot x &= y\\
   1634         \dot y &= -\sin(x).
   1635     \end{aligned}$$
   1636     The potential energy is
   1637     $$U(x) = -\int_0^x-\sin(s)ds = -[\cos(s)]_{s=0}^{s=x} = 1-\cos(x).$$
   1638     Then the total energy is
   1639     $$H(x,y) = \half {y^2}+1-\cos(x).$$
   1640     So $H(x,y) = 2$ if and only if $y = \pm\sqrt{2(\cos x+1)}$. We can conclude that $(-\pi,0)$ and $(\pi,0)$ are connected by heteroclinic orbits.\\
   1641     \\
   1642     Equilibria: The point $(0,0)$ is a center (Lyapunov stable), while $(\pm \pi,0)$ are saddles (not Lyapunov stable).
   1643 \end{exam}
   1644 
   1645 \subsection{Gradient systems in $\Rn$}
   1646 
   1647 Let  $V \in C^2(U,\R)$ and
   1648 $$\dot x = -\nabla V(x)^T,$$
   1649 where $\nabla V(x) = (\partial_{x_1} V, \dots, \partial_{x_n} V)$.
   1650 So
   1651 $$\dot x_i = -\partial_{x_i} V, \quad i \in\jbr{n} \text{ for all }$$
   1652 Sometimes the system
   1653 $$\dot x = \nabla V(x)^T$$
   1654 is analyzed (but not here). Critical points of $V$ are the points, where $\nabla V(x) = 0$ (and these are exactly the equilibria of the ODE).\\
   1655 Regular points of $V$ are the points, where $\nabla V(x) \neq 0$. In this case $\nabla V(x)$ is orthogonal to the level sets $\{V = c\}$ and the vector field points to the steepest descent.
   1656 
   1657 \begin{exam}
   1658     Let $V(x,y) = \half1(2x^2+y^2)$.
   1659     $$\begin{aligned}
   1660         \dot x &= -2x\\
   1661         \dot y &= -y.
   1662     \end{aligned}$$
   1663     The level sets look like ellipses. The flows are orthogonal to the level sets and converge to the origin.
   1664 \end{exam}
   1665 
   1666 \begin{lem}
   1667     For a gradient system, the following statements hold:
   1668     \begin{enumerate}
   1669         \item $V$ is a Lyapunov function ($\dot V \leq 0$).
   1670 
   1671         \item $\dot V(p)=0$ if and only if $p$ is an equilibrium.
   1672 
   1673         \item Strict local minima of $V$ are asymptotically stable equilibria.
   1674 
   1675         \item The $\omega$-limit set consists of equilibria only.
   1676 
   1677         \item If equilibria are isolated then every bounded trajectory converges to an equilibrium.
   1678     \end{enumerate}
   1679 \end{lem}
   1680 
   1681 \begin{proof}
   1682     The statements are simple to prove.
   1683     \begin{enumerate}
   1684         \item $\dot V = \nabla V\cdot(-\nabla V)^T = -\|\nabla V\|^2 \leq 0$.
   1685 
   1686         \item We know that $p$ is an equilibrium if and only if $\nabla V(p) = 0$. This is the case if and only if $\dot V(p) = 0$.
   1687 
   1688         \item Follows from the Lyapunov stability theorem.
   1689 
   1690         \item Follows from LaSalle.
   1691 
   1692         \item trivial.
   1693     \end{enumerate}
   1694 \end{proof}
   1695 
   1696 \begin{exam}
   1697     Let $V(x,y) = (x+1)^2(x-1)^2+y^2$. So
   1698     $$\begin{aligned}
   1699         \dot x &= -4(x+1)x(x-1)\\
   1700         \dot y &= -2y.
   1701     \end{aligned}$$
   1702     Equilibria are:
   1703     \begin{itemize}
   1704         \item $(-1,0)$ asymptotically stable
   1705 
   1706         \item $(0,0)$ saddle
   1707 
   1708         \item $(1,0)$ asymptotically stable
   1709     \end{itemize}
   1710 \end{exam}
   1711 
   1712 Linearization at an equilibrium $-(\partial_{x_i}\partial_{x_j}V)_{i,j}^n$ is symmetric. Therefore eigenvalues are real.
   1713 
   1714 \begin{thm}
   1715     A non-degenerate equilibrium of a planar gradient system is either a saddle (when it is a saddle of $V$), a stable node (when it is a strict local minimum of $V$) or an unstable node (when it is a strict local maximum of $V$).
   1716 \end{thm}
   1717 
   1718 \begin{rem}
   1719     Surprisingly $\omega(p)$ can be a continuum of equilibria (for example, a circle).
   1720 \end{rem}
   1721 
   1722 How to recognize a gradient system?
   1723 The system $\dot x = f(x)$ is a gradient system in $\Rn$ if and only if $\partial_{x_j}f_i = \partial_{x_i}f_j$ for $i,j\in\jbr{n}$.
   1724 
   1725 \begin{hw}
   1726     Given $\dot x=Ax,$ where $ A\in\R^{n\times n}$. Which linear systems are gradient systems? Find $V$.
   1727 \end{hw}
   1728 
   1729 \begin{thm}
   1730     Given
   1731     $$\begin{aligned}
   1732         \dot x &= f(x,y)\\
   1733         \dot y &= g(x,y).
   1734     \end{aligned}$$
   1735     Rotate the vector field by $-90$ degree.
   1736     $$\begin{aligned}
   1737         \dot x &= g(x,y)\\
   1738         \dot y &= -f(x,y)
   1739     \end{aligned}$$
   1740     A center of the first system is a node in the second system. A saddle of the first system is a saddle of the second system. Focus refers to focus. At non-equilibrium points, the trajectories are orthogonal. The first one is a Hamiltonian system if and only if the second one is a gradient system. (Flows and level sets switch roles.)
   1741 \end{thm}
   1742 
   1743 \begin{hw}
   1744     Given
   1745     $$\begin{aligned}
   1746         \dot x &= y\\
   1747         \dot y &= -x+x^2.
   1748     \end{aligned}$$
   1749     Find the Hamiltonian $H$ and sketch the phase portrait.
   1750 \end{hw}
   1751 
   1752 \subsection{First integral (or constant of motion)}
   1753 
   1754 Given
   1755 $$\dot x(t)=f(x(t)),$$
   1756 where $f \in C^1(U, \Rn)$.
   1757 
   1758 \begin{defin}
   1759     A function $V \in C^1(U,\Rn)$ is called a first integral or a constant of motion if $\dot V = 0$ in $U$.
   1760 \end{defin}
   1761 
   1762 \begin{exam}[Lotka reactions]
   1763     Given
   1764     $$\begin{array}{rcl}
   1765     X &\stackrel{\kappa_1}{\to}& 2X\\
   1766     X+Y &\stackrel{\kappa_2}{\to}& 2Y\\
   1767     Y &\stackrel{\kappa_3}{\to}& 0
   1768     \end{array}$$
   1769     we can model it as a system of differential equations.
   1770     $$\begin{array}{rcccl}\dot x &=&\kappa_1x-\kappa_2xy &=& x(\kappa_1-\kappa_2y)\\
   1771     \dot y&=&\kappa_2xy-\kappa_3y &=& y(\kappa_2x-\kappa_3)\\
   1772     &&\kappa_1,\kappa_2, \kappa_3 &>& 0
   1773     \end{array}$$
   1774     The ositive equilibrium is $(x^*,y^*)=\left(\frac{\kappa_3}{\kappa_2},\frac{\kappa_1}{\kappa_2}\right)$. Calculating the Jacobian matrix at $(x^*, y^*)$ gives
   1775     $$J = \mattwo0{-\kappa_3}{\kappa_1}0.$$
   1776     So $\sigma(J)=\{\pm\omega i\}$ with $\omega = \sqrt{\kappa_1\kappa_3}$. Now we define
   1777     $$V(x,y)=\kappa_3\log x + \kappa_1\log y-\kappa_2(x+y).$$
   1778     Then
   1779     $$\dot V(x,y) = \partial_xV\dot x+\partial_y V\dot y = \left(\frac{\kappa_3}{x}-\kappa_2\right)x(\kappa_1-\kappa_2y)+\left(\frac{\kappa_1}{y}-\kappa_2\right)y(\kappa_2x-\kappa_3) = 0,$$
   1780     so $V$ is a first integral.
   1781 \end{exam}
   1782 
   1783 \begin{exam}[Ivanova reactions]
   1784     Given
   1785     $$\begin{array}{rcl}Z+X &\stackrel{\kappa_1}{\to}& 2X\\
   1786     X+Y &\stackrel{\kappa_2}{\to}& 2Y\\
   1787     Y+Z &\stackrel{\kappa_3}{\to}& 2Y
   1788     \end{array}$$
   1789     or
   1790     $$\begin{array}{rcl}\dot x &=& x(\kappa_1z - \kappa_2y)\\
   1791     \dot y &=& y(\kappa_2x - \kappa_3z)\\
   1792     \dot z &=& z(\kappa_3y - \kappa_1x),\end{array}$$
   1793     define
   1794     $$V_1(x,y,z) = x+y+z.$$
   1795     Then
   1796     $$\dot V_1=\dot x+\dot y+\dot z = 0,$$
   1797     so $V_1$ is a first integral. We can get a different first integral defining
   1798     $$V_2(x,y,z) = x^{\kappa_3}y^{\kappa_1}z^{\kappa_2} \text{ or } \log x^{\kappa_3}y^{\kappa_1}z^{\kappa_2}.$$
   1799     In those cases
   1800     $$\dot V_2=0$$
   1801     and we can conclude that $x(t)^{\kappa_3}y(t)^{\kappa_1}z(t)^{\kappa_2}$ is constant.
   1802 \end{exam}
   1803 
   1804 \begin{defin}[Lotka-Volterra equation]
   1805     A Lotka-Volterra equation is a system of equations like
   1806     $$\dot x_i = x_i\left(r_i + \sum_{k=1}^na_{ik}x_k\right)$$
   1807     where $i \in \jbr{n}$.
   1808 \end{defin}
   1809 
   1810 \begin{exam}[A Newtonian system]
   1811     Given
   1812     $$\begin{aligned}
   1813         \dot x &= y\\
   1814         \dot y &= -x+x^3
   1815     \end{aligned}$$
   1816     we can find the first integral
   1817     $$V(x,y)=\half{y^2}+\half{x^2}-\frac{x^4}{4}$$
   1818     since
   1819     $$\dot V(x,y) = (x-x^3)y +y(-x+x^3)=0.$$
   1820 \end{exam}
   1821 
   1822 \begin{rem}
   1823     For a Hamiltonian system
   1824     $$\begin{aligned}
   1825         \dot x&=\partial_y H\\
   1826         \dot y &= -\partial_x H
   1827     \end{aligned}$$
   1828     $H$ is a constant of motion.
   1829 \end{rem}
   1830 
   1831 \begin{rem}
   1832     A planar vector field $(f,g)$ is divergence-free if and only if
   1833     $$\begin{aligned}
   1834         \dot x &= f(x,y)\\
   1835         \dot y &= g(x,y)
   1836     \end{aligned}$$
   1837     is Hamiltonian system. Why?
   1838     Let $H$ be such that
   1839     $$\begin{aligned}
   1840         \partial_y H &= f\\
   1841         \partial_x H &= -g.
   1842     \end{aligned}$$
   1843     This holds if and only if
   1844     $$\partial_y\partial_xH = \partial_x\partial_yH.$$
   1845     Now we rearrange it a bit
   1846     $$\partial_xf = -\partial_yg$$
   1847     or
   1848     $$\divg(f,g) = \partial_xf+\partial_y g = 0.$$
   1849     How to find $H$?
   1850     $$H(x,y) = \int f(x,y)dy+c(x)$$
   1851     where $c(x)$ is found by making sure $-\partial_xH=g$ holds.
   1852 \end{rem}
   1853 
   1854 \begin{hw}
   1855     The system
   1856     $$\begin{aligned}
   1857         \dot x &= -2x-2y-2\\
   1858         \dot y &= -2x+2y-2
   1859     \end{aligned}$$
   1860     is divergence-free. Find the Hamiltonian function $H$.
   1861 \end{hw}
   1862 
   1863 \begin{lem}
   1864     Let $f:U\to \Rn$ locally Lipschitz, $v:U\to \R_{>0}$ locally Lipschitz. Then the orbits of $\dot x(t)=f(x(t))$ and $\dot y(t)=v(y(t))f(y(t))$ coincide and the directions of the flows are the same.
   1865 \end{lem}
   1866 
   1867 \begin{exam}
   1868     The systems
   1869     $$\begin{aligned}
   1870         \dot x &= y\\
   1871         \dot y &= -x+x^3
   1872     \end{aligned}
   1873     \leftrightarrow
   1874     \begin{aligned}
   1875         \dot x &= 2y\\
   1876         \dot y &= -2x+2x^3
   1877     \end{aligned}
   1878     \leftrightarrow
   1879     \begin{aligned}
   1880         \dot x &= (x^2+y^2+1)y\\
   1881         \dot y &= (x^2+y^2+1)(-2x+2x^3)
   1882     \end{aligned}$$
   1883     have the same orbits.
   1884 \end{exam}
   1885 
   1886 \begin{exam}[Lotka ODE in $\R^2_+$]
   1887     For $v(x,y)=\reci{xy}$ let
   1888     $$\begin{aligned}
   1889         \dot x &= \frac{\kappa_1}{y}-\kappa_2\\
   1890         \dot y &= \kappa_2-\frac{\kappa_3}{x}
   1891     \end{aligned}$$
   1892     be given.
   1893 \end{exam}
   1894 
   1895 \begin{hw}
   1896     Find the Hamiltonian function $H$.
   1897 \end{hw}
   1898 
   1899 \begin{defin}
   1900     For
   1901     $$\begin{aligned}
   1902         \dot x=f(x,y)\\
   1903         \dot y = g(x,y)
   1904     \end{aligned}$$
   1905     the function $v(x,y) > 0$ is called a Dulac function if $\divg(vf,vg) = 0$.
   1906 \end{defin}
   1907 
   1908 \subsection{How to find centers?}
   1909 
   1910 Given
   1911 $$\begin{aligned}
   1912     \dot x &= f(x,y)\\
   1913     \dot y &= g(x,y)
   1914 \end{aligned}$$
   1915 where $(x^*,y^*)$ is an equilibrium and $J \in \R^{2\times 2}$ is the Jacobian matrix at $(x^*,y^*)$. If $\tr J = 0$ and $\det J > 0$ then $\sigma(J) = \{\pm \omega i\}$ with $\omega = \sqrt{\det J}$. In this case the equilibrium of the linearization is a center. Try to find a Dulac function.
   1916 
   1917 \subsubsection{Planar S-systems}
   1918 
   1919 Given
   1920 $$\begin{aligned}
   1921     \dot x_1 &= \alpha_1 x_1^{g_{11}}x_2^{g_{12}}-\beta_1x_1^{h_{11}}x_2^{h_{12}}\\
   1922     \dot x_2 &= \alpha_2 x_1^{g_{21}}x_2^{g_{22}}-\beta_2x_1^{h_{21}}x_2^{h_{22}}
   1923 \end{aligned}$$
   1924 where $\alpha_1, \alpha_2, \beta_1, \beta_2 > 0$ and $g_{ij}, h_{ij} \in \R$ for $i,j = 1,2$. With some nonlinear transformation:
   1925 $$\begin{aligned}
   1926     \dot u &= e^{a_1u+b_1v}-e^{a_2u+b_2v}\\
   1927     \dot v &= e^{a_3u+b_3v}-e^{a_4u+b_4v}
   1928 \end{aligned}$$
   1929 For which $a_i,b_i$ is the origin a center?
   1930 
   1931 $$J=\mattwo{a_1-a_2}{b_1-b_2}{a_3-a_4}{b_3-b_4}$$
   1932 is the Jacobian at the origin. Assume $\tr J=0$ and $\det J > 0$. Assume further $a_1=a_2$ and $b_3=b_4$. Use the Dulac function $e^{-a_1u-b_4v}$. Then
   1933 $$\begin{aligned}
   1934     \dot u &= e^{(b_1-b_4)v}-e^{(b_2-b_4)v}\\
   1935     \dot v &= e^{(a_3-a_1)u}-e^{(a_4-a_1)u}
   1936 \end{aligned}$$
   1937 is divergence-free and the origin is a center.
   1938 
   1939 \begin{hw}
   1940     Find $H(u,v)$, such that
   1941     $$\begin{aligned}
   1942         \dot u &= \partial_v H\\
   1943         \dot v &= -\partial_u H.
   1944     \end{aligned}$$
   1945 \end{hw}
   1946 
   1947 \subsubsection{Reversible systems}
   1948 
   1949 Given
   1950 $$\begin{aligned}
   1951     \dot x &= f(x,y)\\
   1952     \dot y &= g(x,y).
   1953 \end{aligned}$$
   1954 Assume $f$ is odd in $y$ ($f(x,-y) = -f(x,y)$) and $g$ is even in $y$ ($g(x,-y) = g(x,y)$). For example, Newtonian systems. If $(x(t), y(t))$ is a solution, then $(\tilde x(t), \tilde y(t)):=(x(-t),-y(-t))$ is also a solution. We can show this directly.
   1955 $$\dot{\tilde x}(t) = -\dot x(-t)=-f(x(-t),y(-t))=-f(\tilde x(t),-\tilde y(t)) = f(\tilde x(t),\tilde y(t))$$
   1956 $$\dot{\tilde y}(t) = \dot y(-t)=g(x(-t),y(-t))=g(\tilde x(t),-\tilde y(t)) = g(\tilde x(t),\tilde y(t))$$
   1957 
   1958 \begin{thm}
   1959     An equilibrium on the $x$-axis of a reversible system with purely imaginary eigenvalues is a center.
   1960 \end{thm}
   1961 
   1962 More generally:
   1963 \begin{itemize}
   1964     \item Let $R: \R^2\to \R^2$ be a reflection along a line.
   1965 
   1966     \item A vector field $F:\R^2\to \R^2$ is said to be reversible with respect to $R$ if $-R^{-1}\circ F\circ R = F$.
   1967 
   1968     \item A similar theorem holds.
   1969 \end{itemize}
   1970 
   1971 \paragraph{Reversibility with respect to $x = y$ line:}
   1972 
   1973 Generally, a system of the form
   1974 $$\begin{aligned}
   1975     \dot x &= f(x,y)\\
   1976     \dot y &= -f(y,x)
   1977 \end{aligned}$$
   1978 is reversible with respect to the $x=y$ line.
   1979 
   1980 \begin{exam}
   1981     Given
   1982     $$\begin{aligned}
   1983         \dot u &= e^{(a_1-a)u+(b_1-b)v}-e^{(a_2-a)u+(b_2-b)v}\\
   1984         \dot v &= e^{(a_3-a)u+(b_3-b)v}-e^{(a_4-a)u+(b_4-b)v}
   1985     \end{aligned}$$
   1986     where we assume that
   1987     $$\begin{aligned}
   1988         a_1-a = b_4-b\\
   1989         a_2-a = b_3-b\\
   1990         a_3-a = b_2-b\\
   1991         a_4-a = b_1-b,
   1992     \end{aligned}$$
   1993     the system is reversible with respect to the $u=v$ line. There exist such $a,b$ if and only if $$a_1-b_4=a_2-b_3 = a_3-b_2=a_4-b_1 \quad (*).$$
   1994     Then the origin is thus a center if $(*)$ holds and $\det J > 0$ (notice that $\tr J=0$  follows automatically from $(*)$).
   1995 \end{exam}
   1996 
   1997 
   1998 
   1999 \begin{hw}
   2000     Figure out when
   2001     $$\begin{aligned}
   2002         \dot u &= f(u,v)\\
   2003         \dot v &= g(u,v)
   2004     \end{aligned}$$
   2005     is reversible with respect to
   2006     \begin{enumerate}
   2007         \item the $v$-axis
   2008 
   2009         \item the $v=-u$ line.
   2010     \end{enumerate}
   2011     And try to find planar $S$-systems with a reversible center (especially in the latter case, i.e., reversible with respect to the line $v=-u$).
   2012 \end{hw}
   2013 
   2014 \subsection{Stable and unstable manifolds}
   2015 
   2016 \begin{exam}
   2017     For the system
   2018     $$\begin{aligned}
   2019         \dot x &= -x\\
   2020         \dot y &= -y+x^2\\
   2021         \dot z &= z+x^2
   2022     \end{aligned}$$
   2023     the only equilibrium is at the origin. The Jacobian matrix at the origin is
   2024     $$J = \matdiagthree{-1}{-1}{1}.$$
   2025     $$\begin{aligned}
   2026         E^s &= (x,y)-\text{axis}\\
   2027         E^c &= \{0\}\\
   2028         E^u &= z-\text{axis}
   2029     \end{aligned}$$
   2030     Solution:
   2031     $$\begin{aligned}
   2032         x(t) &= x(0)e^{-t}\\
   2033         y(t) &= y(0)e^{-t}+x(0)^2(e^{-t}-e^{-2t})\\
   2034         z(t) &= z(0)e^t+\frac{x(0)^2}{3}(e^t-e^{-2t})
   2035     \end{aligned}$$
   2036     Obeserve that if $z(0) = -\frac{x(0)^2}{3}$ then
   2037     $$(x(t),y(t),z(t)) \stackrel{t\to\infty}{\to} (0,0,0)$$
   2038     and if $x(0) = y(0) = 0$ then
   2039     $$(x(t),y(t),z(t)) \stackrel{t\to-\infty}{\to} (0,0,0).$$
   2040     Define
   2041     $$W^s=\{(x,y,z)\in\R^3:z=-\frac{x^2}{3}\}$$
   2042     the stable manifold (tangent to $E^s$) and
   2043     $$W^u=\{(x,y,z)\in\R^3:x=y=0\}$$
   2044     the unstable manifold ($W^u = E^u$).
   2045 \end{exam}
   2046 
   2047 Setting
   2048 $$\dot x(t) = f(x(t))$$
   2049 where $f\in C^1(U,\Rn)$. Assume $0\in U$ is a hyperbolic equilibrium ($f(0) = 0$ and $\Re \lambda \neq 0$ for all $\lambda \in \sigma(f'(0))$. Then $\Rn = E^s \oplus E^u$.
   2050 
   2051 \begin{thm}[Stable and unstable manifolds]
   2052     There exist a neighborhood $\tilde U$ of the origin, $\psi\in C^1(E^s\cap\tilde U, E^u)$ and $\phi\in C^1(E^u\cap \tilde U, E^s)$ for which $W^s=\{(x_s,\psi(x_s)): x_s\in E^s\cap \tilde U\}$ is the stable manifold and $W^u=\{(\psi(x_u),x_u),x_u\in E^u \cap \tilde U\}$ is the unstable manifold with
   2053     \begin{enumerate}
   2054         \item $W^s$ is positively invariant ($\varphi_t(W^s)\subseteq W^s$ for $t \geq 0$).
   2055 
   2056         $W^u$ is negatively invariant ($\varphi_t(W^u)\subseteq W^u$ for $t \leq 0$).
   2057 
   2058         \item $W^s$ is tangential to $E^s$ at the origin. $W^u$ is tangential to $E^u$ at the origin.
   2059 
   2060         \item $\lim_{t\to\infty} \varphi(t,p) = 0$ for $p\in W^s$ and $\lim_{t\to-\infty} \varphi(t,p)=0$ for $p\in W^u$.
   2061     \end{enumerate}
   2062 \end{thm}
   2063 
   2064 \begin{rem}
   2065     If $f$ is $C^r$ for $r \geq 1$ (or analytic) then $\psi$ and $\phi$ are also $C^r$ (or analytic).
   2066 \end{rem}
   2067 
   2068 \begin{exam}
   2069     Given
   2070     $$\begin{aligned}
   2071         \dot x &= -x\\
   2072         \dot y &= y-x^2.
   2073     \end{aligned}$$
   2074     Draw the phase portrait.
   2075     We first come to the obvious conclusions
   2076     $$\begin{aligned}
   2077         E^s = x\text{-axis}\\
   2078         E^u = y\text{-axis}
   2079     \end{aligned}$$
   2080     $$\begin{aligned}
   2081         W^s &= ?\\
   2082         W^u &= y\text{-axis}.
   2083     \end{aligned}$$
   2084     So how do we find $W^s$. We choose the following Ansatz for the stable manifold:
   2085     $$y = \psi(x)=a_2x^2+a_3x^3+\dots$$
   2086     It must be tangential to the $x$-axis. So we drop the first two terms $a_0$ and $a_1x$.
   2087     $$\dot y = \psi'(x)\dot x$$
   2088     Pluggin it in and using the system from before, we get
   2089     $$(a_2-1)x^2+\sum_{k=3}^\infty a_kx^k=\left(\sum_{k=2}^\infty ka_kx^{k-1}\right)(-x).$$
   2090     Comparing coefficients gives us:
   2091     $$\begin{aligned}
   2092         a_2-1 &=-2a_2\\
   2093         a_k &= -ka_k \text{ for } k \geq 3
   2094     \end{aligned}$$
   2095     Therefore $a_2 = \reci3$ and $a_k = 0$ for $k\geq 3$.
   2096 \end{exam}
   2097 
   2098 \begin{defin}[Global mainfolds]
   2099     Global stable manifold: $S = \bigcup_{t \leq 0} \varphi_t(W^s)$\\
   2100     Global unstable manifold: $U=\bigcup_{t\geq 0}\varphi_t(W^u)$.
   2101 \end{defin}
   2102 
   2103 \begin{prop}
   2104     The global stable and unstable manifolds are invariant (meaning $\varphi_t(S)\subseteq S$ for all $t \in \mathbb{R}$ and $\varphi_t(U)\subseteq U$ for all $t \in \mathbb{R}$) and
   2105     $$\lim_{t\to \infty}\varphi(t,p)=0$$
   2106     for $p \in S$ and
   2107     $$\lim_{t\to-\infty}\varphi(t,p) = 0$$
   2108     for $p\in U$.
   2109 \end{prop}
   2110 
   2111 \begin{exam}
   2112     The system
   2113     $$\begin{aligned}
   2114         \dot x &=-x-y^2\\
   2115         \dot y &= y+x^2
   2116     \end{aligned}$$
   2117     is reversible with respect to the line where $x = y$ (since we can swap the roles of $x$ and $y$ without changing the system).
   2118     $$J(x,y) = \mattwo{-1}{-2y}{2x}{1}$$
   2119     $$J(0,0)=\matdiagtwo{-1}1$$
   2120     $$J(-1,-1)=\mattwo{-1}{2}{-2}{1}$$
   2121     $S$ and $U$ intersect in a homoclinic loop.
   2122 \end{exam}
   2123 
   2124 \subsection{Center manifold}
   2125 
   2126 Given
   2127 $$\begin{aligned}
   2128     \dot x &= xy+x^3\\
   2129     \dot y &= -y-2x^2
   2130 \end{aligned}$$
   2131 we get the Jacobian matrix at the origin
   2132 $$J = \matdiagtwo{0}{-1}.$$
   2133 Since $\sigma(J) = \{0,-1\}$ the Jacobian matrix is not hyperbolic and
   2134 $$\begin{aligned}
   2135     E^s&= y-\text{axis}\\
   2136     E^c&= x-\text{axis}\\
   2137     E^u&= \{0\}.
   2138 \end{aligned}$$
   2139 
   2140 Setting $x \in \R^c, y \in \R^{s+u}$ where $c+s+u=n$, we get
   2141 $$\begin{aligned}
   2142     \dot x &= Ax+f(x,y) \text{ with } f(0,0) = 0, f'(0,0)=0\\
   2143     \dot y &= By+g(x,y) \text{ with } g(0,0) = 0, g'(0,0)=0,
   2144 \end{aligned}$$
   2145 where $(0,0) \in U$ and $(f,g)\in C^r(U,\Rn)$. The Jacobian matrix at $(0,0)$ is
   2146 $$J=\matdiagtwo{A}{B}$$
   2147 Assume $\Re \lambda = 0$ for all $\lambda \in \sigma(A)$ and $\Re \lambda \neq 0$ for all $\lambda \in \sigma(B)$.
   2148 
   2149 \begin{thm}[Center manifold]
   2150     There exists a neighborhood $\tilde U$ of the origin and a function $\psi \in C^r(E^c\cap \tilde U,E^s\oplus E^u)$ for which
   2151     \begin{enumerate}
   2152         \item $W^c = \{(x,\psi(x)): x\in E^c\cap \tilde U\}$ is locally invariant.
   2153 
   2154         \item $\psi(0) = 0, \psi'(0) = 0$ (meaning $W^c$ is tangential to $E^c$ at the origin.
   2155 
   2156         \item $B\psi(x)+g(x,\psi(x)) = \psi'(x)(Ax+f(x,\psi(x)))$ (this helps finding $\psi$)
   2157 
   2158         \item The system
   2159         $$\begin{aligned}
   2160             \dot x &= Ax +f(x,\psi(x))\\
   2161             \dot y &= By
   2162         \end{aligned}$$
   2163         is locally topologically equivalent to the original system at the origin. (Generalization of the Hartman-Grobman Theorem to the nonhyperbolic case.)
   2164     \end{enumerate}
   2165 \end{thm}
   2166 
   2167 \begin{rem}
   2168     Differentiating $y(t)=\psi(x(t))$ with respect to the time, one gets 3).
   2169 \end{rem}
   2170 
   2171 \paragraph{Bad news:}
   2172 The center manifold is not unique, and need not be as smooth as $(f,g)$. (E.g.\ for analytic $(f,g)$ there can be a continuum of center manifolds, and only one of them is analytic. Or even none of them is analytic!)
   2173 
   2174 \paragraph{Good news:}
   2175 The Taylor series expansion of $\psi$ is unique (as far as it exists), and can usually be computed. All orbits near the origin that stay near the origin for all time $t\in\R$ are contained in every center manifold.
   2176 
   2177 \begin{exam}
   2178     Given
   2179     $$\begin{aligned}
   2180         \dot x &= xy+x^3\\
   2181         \dot y &= -y-2x^2
   2182     \end{aligned}$$
   2183     approximate the center manifold with the ansatz
   2184     $$y=\psi(x)=a_2x^2+a_3x^3+\dots$$
   2185     The linear term is missing, because the center manifold must be tangential to the center subspace. By the invariance of $W^c$, $y=\psi'(x)x$.
   2186     $$-\sum_{k=2}^\infty a_kx^k-2x^2=\left(\sum_{k=2}^\infty ka_kx^{k-1}\right)\left(x\sum_{k=2}^\infty a_kx^k+x^3\right)$$
   2187     Comparing the coefficients of $x^2$, we get $$-a_2-2=0$$
   2188     or
   2189     $$a_2=-2.$$
   2190     So we conclude that
   2191     $$\psi(x)=-2x^2+O(x^3)$$
   2192     and
   2193     $$\dot x = x(-2x^2+O(x^3))+x^3=-x^3+O(x^4).$$
   2194     So the trajectories on the center manifold tend to the origin.
   2195 \end{exam}
   2196 
   2197 \begin{exam}
   2198     Given
   2199     $$\begin{aligned}
   2200         \dot x &= x^2y-x^5\\
   2201         \dot y &= -y+x^2,
   2202     \end{aligned}$$
   2203     a similar analysis as above reveals
   2204     $$y=\psi(x)=x^2+O(x^5)$$
   2205     so
   2206     $$\dot x = x^4+O(x^5).$$
   2207     Hence, on any center manifold, the origin attracts for $x<0$ and repels for $x>0$.
   2208 
   2209 \end{exam}
   2210 
   2211 \subsection{Andronov--Hopf bifurcation}
   2212 
   2213 \begin{exam}
   2214     Given
   2215     $$\begin{aligned}
   2216         \dot x &= \alpha x-y-x(x^2+y^2)\\
   2217         \dot y &= x+\alpha y - y(x^2+y^2),
   2218     \end{aligned}$$
   2219     where $\alpha \in \R$ is a parameter.
   2220     The Jacobian at the origin is
   2221     $$\mattwo\alpha{-1}1\alpha$$
   2222     with the eigenvalues $\alpha \pm i$. In polar coordinates, the system is
   2223     $$\begin{aligned}
   2224         \dot r &= r(\alpha-r^2)\\
   2225         \dot \theta &= 1.
   2226     \end{aligned}$$
   2227     \begin{itemize}
   2228         \item If $\alpha < 0$: The solutions spiral counter clockwise to the origin. It is linearly stable. It approaches the origin at an exponential speed.
   2229 
   2230         \item If $\alpha = 0$: The solutions spiral counter clockwise to the origin much slower.
   2231 
   2232         \item If $\alpha > 0$: We get a stable limit cycle of radius $\sqrt{\alpha}$.
   2233     \end{itemize}
   2234     Supercritical Andronov--Hopf bifurcation.
   2235 \end{exam}
   2236 
   2237 \begin{exam}
   2238     Given
   2239     $$\begin{aligned}
   2240         \dot x &= \alpha x-y+x(x^2+y^2)\\
   2241         \dot y &= y + \alpha y + y(x^2+y^2)
   2242     \end{aligned}$$
   2243     or
   2244     $$\begin{aligned}
   2245         \dot r &= r(\alpha + r^2)\\
   2246         \dot \theta &= 1.
   2247     \end{aligned}$$
   2248     \begin{itemize}
   2249         \item If $\alpha < 0$: We get an unstable limit cycle of radius $\sqrt{-\alpha}$.
   2250 
   2251         \item If $\alpha = 0$: The solutions spiral counter clockwise away from the origin much slower than for $\alpha>0$.
   2252 
   2253         \item If $\alpha > 0$: The solutions spiral counter clockwise away from the origin. It is linearly stable. It leaves the origin at an exponential speed.
   2254     \end{itemize}
   2255     Subcritical Andronov--Hopf bifurcation.
   2256 \end{exam}
   2257 
   2258 \begin{exam}
   2259     Given
   2260     $$\begin{aligned}
   2261         \dot x &= \alpha x - y\\
   2262         \dot y &= x+\alpha y
   2263     \end{aligned} \quad\text{ or in polar form }\quad\begin{aligned}
   2264         \dot r &= \alpha r\\
   2265         \dot \theta &= 1.
   2266     \end{aligned}$$
   2267     \begin{itemize}
   2268         \item If $\alpha < 0$: The solutions spiral counter clockwise to the origin. It is linearly stable. It approaches the origin at an exponential speed.
   2269 
   2270         \item If $\alpha = 0$: The origin is a center.
   2271 
   2272         \item If $\alpha > 0$: The origin repells.
   2273     \end{itemize}
   2274     Vertical Andronov--Hopf bifurcation.
   2275 \end{exam}
   2276 
   2277 \begin{thm}
   2278     Let $U \subseteq \R^2$ be open and
   2279     $$\dot x(t) = f_\alpha(x(t))$$
   2280     be a familiy of ODEs, where $\alpha \in (-\epsilon, \epsilon)$ for some $\epsilon > 0$. We assume that
   2281     $$(\alpha,x) \mapsto f_\alpha(x)$$
   2282     is $C^1$ with respect to $\alpha$ and  $C^3$ with respect to $x$. Assume $f_\alpha(0) = 0$ for all $\alpha \in (-\epsilon, \epsilon)$. Let
   2283     $$\mu(\alpha)\pm i\omega(\alpha)$$
   2284     be the eigenvalues of the Jacobian matrix at the origin,
   2285     where $\mu(0)=0$ and $\omega(0) > 0$.
   2286     Assume $\mu'(0)\neq 0$ (transverality) and $L_1\neq 0$ (nondegeneracy), where $L_1$ is the first focal value (the computation is explained in the next lemma). Then there are invertible coordinate and parameter changes and also time reparametrisation transforming the
   2287     $$\dot x(t)=f_\alpha(x(t))$$
   2288     to
   2289     $$d_t y(t)=\mattwo{\beta}{-1}{1}{\beta}y\pm|y|^2y+O(|y|^4),$$ where $\pm$ is the sign of $L_1$. Furthermore, omitting $O(|y|^4)$ leads to a family that is locally topologically equivalent near the origin.
   2290     \begin{itemize}
   2291         \item $L_1<0$: supercritical Andronov--Hopf (stable limit cycle)
   2292 
   2293         \item $L_1>0$: subcritical Andronov--Hopf (unstable limit cycle)
   2294     \end{itemize}
   2295 \end{thm}
   2296 
   2297 \begin{proof}
   2298     e.g. Kuznetsov: Elements of bifurcation theory (Section 3.5)
   2299 \end{proof}
   2300 
   2301 \begin{lem}
   2302     Given
   2303     $$\begin{aligned}
   2304         \dot x &= f(x,y)=-\omega y+\sum_{i+j=2} f_{ij}x^iy^j\\
   2305         \dot y &= g(x,y)=\omega x+\sum_{i+j=2} g_{ij}yx^iy^j.
   2306     \end{aligned}$$
   2307     We define, $f_{ij}=\frac{1}{i!j!}\frac{\partial^{i+j}f}{\partial_x^i \partial_y^j}\Big|_{(x,y)=(0,0)}$ and $g_{ij}=\frac{1}{i!j!}\frac{\partial^{i+j}g}{\partial_x^i \partial_y^j}\Big|_{(x,y)=(0,0)}$.
   2308     Then
   2309     $$L_1 = 3f_{30}+f_{12}+3g_{03}+g_{21}+\reci{\omega}[f_{11}(f_{20}+f_{02})-g_{11}(g_{20}+g_{02})+2f_{02}g_{02}-2f_{20}g_{20}]$$
   2310     (Bautin 1949).
   2311 \end{lem}
   2312 
   2313 Other names for the focal value
   2314 \begin{itemize}
   2315     \item Lyapunov value
   2316 
   2317     \item Lyapunov coefficient
   2318 
   2319     \item Lyapunov constant
   2320 
   2321     \item Lyapunov quantity
   2322 
   2323     \item Poincar\'e--Lyapunov coefficient
   2324 
   2325     \item Poincar\'e constant
   2326 
   2327     \item Bautin constant
   2328 
   2329     \item Fokusgröße
   2330 
   2331     \item Strudelgröße
   2332 \end{itemize}
   2333 
   2334 \begin{exam}[Brusselator]
   2335     Given
   2336     $$0 \stackrel{\stackrel{\kappa_1}{\rightarrow}}{\stackrel{\leftarrow}{\kappa_2}} X \stackrel{\kappa_3}{\rightarrow} Y$$
   2337     $$2X+Y\rightarrow^{\kappa_4}3X$$
   2338     where $\kappa_1,\kappa_2,\kappa_2,\kappa_2 > 0$ are parameters. We get
   2339     $$\begin{aligned}
   2340         \dot x&= \kappa_1-(\kappa_2+\kappa_3)x+\kappa_4x^2y\\
   2341         \dot y &= \kappa_3x-\kappa_4x^2y
   2342     \end{aligned}$$
   2343     where $x,y\geq 0$. Then the
   2344     positive equilibrium is $(x^*,y^*) = \left(\frac{\kappa_1}{\kappa_2},\frac{\kappa_2\kappa_3}{\kappa_1\kappa_4}\right)$. The Jacobian at this point is
   2345     $$J=\mattwo{-\kappa_2+\kappa_3}{\frac{\kappa_1^2\kappa_4}{\kappa_2^2}}{-\kappa_3}{-\frac{\kappa_1^2\kappa_4}{\kappa_2^2}}$$ and also
   2346     $$\tr J=-\kappa_2+\kappa_3-\frac{\kappa_1^2\kappa_4}{\kappa_2^2}$$
   2347     $$\det J=\frac{\kappa_1^2\kappa_4}{\kappa_2} > 0.$$
   2348     Fix $\kappa_1,\kappa_2,\kappa_4$ and keep only $\kappa_3$ as a parameter
   2349     $$\mu(\kappa_3) = \Re \lambda_{12}(\kappa_3) = \half1\left(\kappa_3-\left(\kappa_2+\frac{\kappa_1^2\kappa_4}{\kappa_2^2}\right)\right)$$
   2350     $$\mu'(\kappa_3) = \half1$$
   2351     so the transversality holds.
   2352     \begin{enumerate}
   2353         \item Shift the equilibrium to the origin
   2354         $$\tilde x(t) = x(t)-x^*$$
   2355         $$\tilde y(t) = y(t)-y^*$$
   2356         $$\begin{aligned}
   2357             &\dot{\tilde x} = \dot x = \kappa_1-(\kappa_2+\kappa_3)(\tilde x+x^*)+\kappa_4(\tilde x+x^*)^2(\tilde y+y^*)\\
   2358             &\dot{\tilde y} = \dot y = \kappa_3(\tilde x+x^*)-\kappa_4(\tilde x+x^*)^2(\tilde y+y^*).
   2359         \end{aligned}$$
   2360 
   2361         \item Eliminate $\kappa_3$ by $\tr J = 0$.
   2362         $$J = \mattwo abc{-a}.$$
   2363         We define
   2364         $$T=\mattwo10{-\frac a\omega}{-\frac b\omega}$$
   2365         so
   2366         $$T^{-1} = \mattwo10{-\frac ab}{-\frac\omega b}.$$
   2367         Therefore
   2368         $$TJT^{-1} = \mattwo0{-\omega}\omega0.$$
   2369         In general $\dot x = f(x), f(0)=0$. We define $u = Tx$ where $T \in \R^{n\times n}$ is invertible. So
   2370         $$\dot u = T\dot x = Tf(x) = Tf(T^{-1}u).$$
   2371         In this case the Jacobian at $u = 0$ is $Tf'(0)T^{-1}$.\\
   2372         Back to the Brusselator:
   2373         $$\vectwo uv = T\vectwo{\tilde{x}}{\tilde{y}}$$
   2374         $$\dot u = -\omega v+\left(\frac{\kappa_2^2}{\kappa_1}-\frac{\kappa_1\kappa_4}{\kappa_2}\right)u^2-2\sqrt{\kappa_2\kappa_4}uv-\kappa_4u^3-\frac{\kappa_2^2}{\kappa_1^2}\omega u^2v$$
   2375         $$\dot v = \omega u.$$
   2376         So we get
   2377         $$L_1 = -3\kappa_4+\reci\omega(-2\sqrt{\kappa_2\kappa_4})\left(\frac{\kappa_2^2}{\kappa_1}-\frac{\kappa_1\kappa_4}{\kappa_2}\right)=-\frac{1}{\kappa_1^2}(2\kappa_2^3+\kappa_1^2\kappa_4),$$ where we used
   2378         $\omega = \sqrt{\det J}=\kappa_1\sqrt{\frac{\kappa_4}{\kappa_2}}$. Hence, $L_1<0$, and the Andronov--Hopf bifurcation is supercritical. Therefore for $\kappa_3$ slightly larger than $\kappa_2+\frac{\kappa_1^2\kappa_4}{\kappa_2^2}$, the repelling positive equilibrium is surrounded by a stable limit cycle.
   2379     \end{enumerate}
   2380 \end{exam}
   2381 In general, we have
   2382 $$\dot x(t) = f_\alpha(x(t))$$
   2383 where $x(t) \in \R^n, \alpha \in (-\epsilon, \epsilon), f_\alpha(0)=0$ and
   2384 $$\sigma(f_\alpha'(0))=\{\mu(\alpha)\pm \omega(\alpha)i, \lambda_3(\alpha),\dots, \lambda_n(\alpha)\},$$
   2385 where $\mu(0)=0, \omega(0)>0, \mu'(0)\neq 0, \Re \lambda_j(0)\neq 0 \, \forall j \in \jbr{3,n}$. At $\alpha = 0$ there is a 2-dim center manifold. Figure out the behavior there. A similar theorem holds as in 2d, and a recipe for computing the first focal value is available (but not easy).
   2386 
   2387 \paragraph{Back to 2d:}
   2388 What if $L_1 = 0$? Then further work is needed to answer the stability of the equilibrium at the critical parameter value. One has to compute $L_2$ and the formula is roughly one page long and uses up to fifth order derivatives.
   2389 
   2390 \begin{exam}
   2391     Given
   2392     $$\begin{aligned}
   2393         \dot r &=r(\alpha_1+\alpha_2r^2-r^4)\\
   2394         \dot \theta &= 1.
   2395     \end{aligned}$$
   2396     \begin{itemize}
   2397         \item If $\alpha_1 < 0$ and $\alpha_2 > 2\sqrt{-\alpha_1}$ then there are 2 limit cycles (the outer one is stable, the inner one is unstable).
   2398 
   2399         \item If $\alpha_1<0$ and $\alpha_2 = 2\sqrt{-\alpha_1}$, then there is a semistable limit cycle (it attracts solutions from the outside, but repels in the inside).
   2400 
   2401         \item If $L_2\neq 0$ then we talk about a Bautin bifurcation, where, among other things, the above-described fold bifurcation of limit cycles occurs (with a semistable limit cycle at the critical value).
   2402     \end{itemize}
   2403 \end{exam}
   2404 
   2405 \section{Part 2}
   2406 
   2407 \subsection{Ideas from the General theory of dynamical systems}
   2408 
   2409 Dynamical systems theory is understanding the long-term behavior of a dynamical system. We will focus on (autonomous) deterministic system. For the whole chapter $X$ is the statespace of a system and all $x\in X$ are the possible states of the system.
   2410 
   2411 \subsubsection{Continuous vs. discrete time DS}
   2412 
   2413 In the continuous case a flow is defined like the following:
   2414 
   2415 \begin{defin}
   2416     \item A flow $\phi =(\phi_t)_{t\in\R}$ in $X$ is a family of maps $\phi_t:X\to X$ with $t\in \R$, satisfying $\phi_{s+t}=\phi_s\circ\phi_t$. This is a group of transformations. The elements are invertible, because $\phi_t\circ\phi_{-t}=id_X$. We say, $\phi_t(x)$ is the state at time $t$ where $x$ is the start state.
   2417 \end{defin}
   2418 
   2419 For simplicity, we use the notation $(\phi_t)_t$ for $(\phi_t)_{t\in \R}$.
   2420 
   2421 \begin{defin}
   2422     A semiflow on $X$ is a family $\phi=(\phi_t)_{t\geq0}$ of maps which satisfies $\phi_{s+t}=\phi_s\circ\phi_t$ for $s,t\geq 0$ and $\phi_0 = id_X$. The maps need not be invertible.
   2423 \end{defin}
   2424 
   2425 We can define the same for discrete time systems.
   2426 
   2427 \begin{defin}
   2428     A flow is a family $(T_n)_{n\in\N}$ of maps $T_n:X\to X$ such that $T_{n+m}=T_n\circ T_m$ for all $m,n\in \N$. Then $T_n = \underbrace{T\circ\dots\circ T}_{n-\text{times}} =T^n$, where $T := T_1$. $T^0 = id_X$.
   2429 \end{defin}
   2430 
   2431 For simplicity, we use the notation $(\phi_n)_n$ for $(\phi_n)_{n\in \N}$.
   2432 
   2433 \begin{exam}
   2434     Maps on the interval/circle:
   2435     \begin{enumerate}
   2436         \item Circle rotation: Fix $\alpha\in\T$. Start at $e^{2\pi ix}$, where $x \in \T$. Rotate by $\alpha$ to get $e^{2\pi i (x+\alpha)}$.
   2437         $$T:\T\to\T,\quad Tx = x+\alpha$$
   2438 
   2439         \item Angular doubling map:
   2440         $$T:\T\to\T, \quad Tx = 2x$$
   2441 
   2442         \item Full logistic map:
   2443         $$T:[0,1]\to[0,1], \quad Tx = 4x(1-x)$$
   2444 
   2445         \item Tent map: Linear  interpolation of the points $(0,0),(0.5,1),(1,0)$.
   2446 
   2447         \item Maps of the square/torus:
   2448     \end{enumerate}
   2449 \end{exam}
   2450 
   2451 \subsubsection{Continuous time systems can also give you discrete systems: $(\phi_t)_{t\geq0}$ semiflows of $X$}
   2452 
   2453 There are different possibilities to convert a continuous semiflow to a discrete flow. We show two of them.
   2454 
   2455 \begin{enumerate}
   2456     \item Fix a constant $\tau>0$ and consider the system at times $0,\tau, 2\tau, \dots$.
   2457     $$x, \phi_\tau(x)=:Tx, \phi_{2\tau}(x)=:T^2x,\dots$$
   2458 
   2459     \item Poincare section: Let $Y\subseteq X$ be a submanifold of lower dimension such that $\forall y \in Y$ there is a $\tau(y)>0$ (minimal) such that $\phi_{\tau(y)}\in Y$. Then we define the Poincare map of $Y$ as
   2460     $$T:Y\to Y, \quad Ty:=\phi_{\tau(y)}\in Y.$$
   2461 \end{enumerate}
   2462 
   2463 Discrete time systems are Poincare maps $T: Y\to Y$. Let $\tau:Y\to[\epsilon,\infty)$. Define a semiflow on:
   2464 $$X=\{(y,s):y\in Y,s < \tau(y)\}$$
   2465 $$\phi_t(y,s):=\begin{cases}
   2466     (y,s+t) & s+t < \tau(y)\\
   2467     (Ty,t-(\tau(y)-s)) & \tau(y) \leq t+s < \tau(Ty)\\
   2468     (T^2y,t-(\tau(Ty)-s)) & \tau(Ty) \leq t+s < \tau(T^2y)\\
   2469     \vdots
   2470 \end{cases}$$
   2471 In this case, the semiflow starts at $(y,s)$. Then it moves upwards right before it reaches $(y,\tau(y))$. Then it jumps to $(Ty,0)$ and moved upwards. Right before it reaches $(Ty, \tau(Ty))$, it jumps to $(T^2y,0)$ and so on.
   2472 
   2473 \subsubsection{Relations between systems}
   2474 
   2475 \begin{defin}
   2476     Two maps $T:X\to X$ and $S:Y\to Y$ are conjugate, if there is a bijective $\eta:X\to Y$, such that
   2477     $$\eta \circ T=S\circ \eta.$$
   2478     This means that $\eta$ is just a change of coordinates and $S$ represents $T$ in these new coordinates.\\
   2479     \\
   2480     If $\eta$ is just onto, then $\eta$ is a semi-conjugacy (factor map).
   2481 \end{defin}
   2482 
   2483 %\begin{exam}
   2484 %    $T(a,b) := (Sa,H(a,b))$
   2485 %    $$\eta(a,b) := a$$
   2486 %\end{exam}
   2487 
   2488 
   2489 \begin{exam}
   2490     Let $X := \R^2, Y := \R$. Now we define maps
   2491     $$\begin{array}{ll}
   2492         T:X \to X & T(a,b) := (2a,3b)\\
   2493         S:Y \to Y & S(a) := 2a\\
   2494         \eta:X\to Y & \eta(a,b) := a.
   2495     \end{array}$$
   2496     Here we can see that
   2497     $$\eta(T(a,b)) = \eta(2a,3b) = 2a$$
   2498     while
   2499     $$S(\eta(a,b)) = S(a) = 2a$$
   2500     so we conclude that
   2501     $$\eta \circ T = S \circ \eta.$$
   2502     Since $\eta$ is surjective, we know that it is a semi-conjugacy, but due to the lack of injectivity, it is not a conjugacy.
   2503 \end{exam}
   2504 
   2505 \subsubsection{Example: Mathematical Billiards "table" $Q \subseteq \R^2$, open}
   2506 
   2507 Let $Q$ be bounded by $C^3$ curves. We assume that there is no friction and the reflections are elastic. The state of a moving billiard sphere can be represented by an element of $Q\times S^1$ (position and direction). To make sure that the flow is continuous with respect to the time even during a reflection, we have to identify the incoming and outgoing directions on the boundary. It can have very complicated dynamics.
   2508 
   2509 \paragraph{Poincare sections:}
   2510 Let $Y = \partial Q\times S^1$ be the set of the states at the boundaries. Then we can define the collision map $T:Y\to Y$. So if the system starts at $x \in Y$ then $Tx$ is the first point where it hits the boundary.
   2511 
   2512 \begin{rem}
   2513     Such a system is usually not continuous with respect to the starting condition, because a small change of the starting position might change the fact that the flow hits a certain obstacle.
   2514 \end{rem}
   2515 
   2516 \subsubsection{Questions and structure}
   2517 
   2518 \begin{defin}
   2519     For a given $x\in X$ we define the following:
   2520     \begin{itemize}
   2521         \item Discrete case: Let $T:X\to X$ be a map. Then we define $(T^nx)_n$ to be the forward orbit.
   2522 
   2523         \item Continuous case: Let $(\phi_t)_{t\geq 0}$ be a semiflow. Then we define $(\phi_tx)_{t\geq0}$ to be the forward orbit.
   2524     \end{itemize}
   2525 \end{defin}
   2526 
   2527 \begin{defin}
   2528     For $A\subseteq X$ and $T:X\to X$ we use the notation $T^{-n}(A) := (T^n)^{-1}(A)$ for the preimage.
   2529 \end{defin}
   2530 
   2531 Using these definitions, we can ask some questions for $A\subseteq X$:
   2532 
   2533 \begin{itemize}
   2534     \item For given $x\in X$ and $n\in\N$ is $T^nx \in A$ or equivalently, $x \in T^{-n}A$?
   2535 
   2536     \item Is there a $k\in\N$ such that $T^kx\in A$ or equivalently, $x \in \bigcup_{k\geq 1}T^{-k}A$?
   2537 
   2538     \item For $x \in A$ are there infinitely many $k\geq1$ such that $T^kx\in A$? (Recurrence)
   2539 \end{itemize}
   2540 
   2541 \paragraph{Coarse structure:}
   2542 
   2543 Sometime subsets have useful properties.
   2544 
   2545 \begin{defin}
   2546     A set $A\subseteq X$ is called forward invariant, if $TA\subseteq A$.
   2547 \end{defin}
   2548 
   2549 In such a case, we can just study $T|_A:A\to A$.
   2550 
   2551 \begin{defin}
   2552     A set $A\subseteq X$ is called completely invariant, if $TA\subseteq A$ and $TA^c\subseteq A^c$.
   2553 \end{defin}
   2554 
   2555 \paragraph{Topological dynamics:}
   2556 Let $X$ be a (often compact) topological space. We can ask some questions.
   2557 
   2558 \begin{itemize}
   2559     \item Does $T^nx$ converge to $y$? Does $y\in \omega(x):=\{\text{Limit points of } (T^nx)_n\}$?
   2560 
   2561     \item Does $d_X(T^nx,T^ny)$ approach $0$?
   2562 
   2563     \item Is $x\in\omega(x)$? ($x$ recurrent point)
   2564 \end{itemize}
   2565 Usually $A$ should be open or closed. We want forward invariant sets to be closed.
   2566 In case of a conjugacy, we prefer a topological conjugacy which is additionally a homeomorphism. In the case of differentiable dynamics the set $X$ is a $C^r$-manifold. In the case of measurable dynamics $X$ is a measure space (Ergodic Theory).
   2567 
   2568 \subsection{Circle rotations}
   2569 
   2570 Let $X = \T$. We define $A\subseteq \T$ to be an arc (connected subset). We use $\lambda$ as the length of the arc. Let $\alpha \in [0,1]$, and
   2571 $$T:\T\to \T, \quad Tx:=x+\alpha.$$
   2572 Obviously, $T$ is invertible and isometric.
   2573 
   2574 \subsubsection{Rational rotation}
   2575 
   2576 We assume that $\alpha = \frac pq \in \Q$. Then
   2577 $$T^q x=x+q\frac pq=x.$$
   2578 
   2579 \begin{prop}
   2580     Every orbit is $q$-periodic. So it is recurrent.
   2581 \end{prop}
   2582 
   2583 \begin{rem}
   2584     We can predict that if $y$ is in a neighborhood of $x$ then $T^ny$ is close to $T^nx$.
   2585 \end{rem}
   2586 
   2587 \subsubsection{Irrational Rotations}
   2588 
   2589 We assume that $\alpha \notin \Q$. In this case, every orbit $(T^nx)_{n\geq 0}$ is an infinite set.
   2590 
   2591 \begin{prop}
   2592     Every orbit is dense.
   2593 \end{prop}
   2594 
   2595 \begin{proof}
   2596     We show that for all $\epsilon > 0$ the orbit $(T^nx)_n$ is $\epsilon$-dense, meaning that for every $a \in X$ there is a $n\in N$ such that $d(a,T^nx) < \epsilon$.\\
   2597     \\
   2598     Take $\epsilon > 0$ then by Bolzano Weierstraß there exist $m,n$, such that $0<d(T^mx,T^{m+n}x) < \epsilon$. Let $y := T^mx$ and $S:=T^n$. Then $d(y, Sy) < \epsilon$ and $(S^jy, S^{j+1}y) < \epsilon$ for every $j$. So $(S^jy)_j$ is $\epsilon$-dense. Since $(S^jy)_j\subseteq (T^nx)_n$, we know that $(T^nx)_n$ is also $\epsilon$-dense.
   2599 \end{proof}
   2600 
   2601 \begin{cor}
   2602     Every $x$ is recurrent.
   2603 \end{cor}
   2604 
   2605 \begin{rem}
   2606     We predict that there is no small open set containing the orbit of $y$.
   2607 \end{rem}
   2608 
   2609 \subsubsection{Linear flows on the $2$-torus $\T^2$}
   2610 Given
   2611 $$\begin{aligned}
   2612     \dot x&=\omega_1\\
   2613     \dot y&=\omega_2,
   2614 \end{aligned}$$
   2615 where $\omega_1,\omega_2 > 0$. Then
   2616 $$(\phi_t(x,y))=(x+t\omega_1,y+t\omega_2).$$
   2617 Let $Y:= \{0\}\times \T$. This is a global Poincare section, because every flow line meets $Y$ infinitely often. Let $T:Y\to Y$ be the Poincare map. So $T(0,y) = (0,y+\alpha)$, where $\alpha=\frac{\omega_2}{\omega_1}$. So $T$ corresponds to the circle rotation by $\alpha$.
   2618 
   2619 \begin{prop}
   2620     The following statements hold.
   2621     \begin{enumerate}
   2622         \item If $\frac{\omega_2}{\omega_1} \in \Q$ then every $\phi$-orbit is periodic.
   2623 
   2624         \item If $\frac{\omega_2}{\omega_1} \notin \Q$ then every $\phi$-orbit is dense in the torus.
   2625     \end{enumerate}
   2626 \end{prop}
   2627 
   2628 \begin{cor}
   2629     Every $\phi$-orbit is recurring.
   2630 \end{cor}
   2631 
   2632 \subsubsection{Some notions of topological dynamics}
   2633 
   2634 \begin{defin}
   2635     A map $T$ on a topological space $X$ is called topologically transitive if it has a dense orbit, meaning that there is an $x\in X$ such that $(T^nx)_n$ is dense.
   2636 \end{defin}
   2637 This means that under certain conditions, we can go from any open set to any other open set. So all orbits are dense. In this case there is no closed forward invariant set other than $\emptyset$ and $X$
   2638 
   2639 \begin{defin}
   2640     IThere is no closed forward invariant set other than $\emptyset$ and $X$ then the system $T$ is called minimal.
   2641 \end{defin}
   2642 
   2643 \subsubsection{Distribution of orbits}
   2644 
   2645 Let $X = \T$, $\alpha \in \R\setminus\Q$ and $Tx=x+\alpha$. Let $A\subseteq \T$ be an arc, consisting of more than one point. Then each orbit intersects $A$. The question is, how many times does the state land in $A$? Let $S_n:\T\to \R$ with
   2646 $$S_n(A):=\sum_{k=0}^{n-1}1_A\circ T^k.$$
   2647 Then
   2648 $$S_n(A)(x)=1_A(x)+1_A(Tx)+\dots+1_A(T^{n-1}x)$$
   2649 is the number of visits (occupation time) of $(T^kx)_{k=0}^{n-1}$. The relative frequency of visits can be calculated, using
   2650 $$\reci nS_n(A).$$
   2651 
   2652 \begin{prop}[Equidistribution of orbits]
   2653     For all arcs $A\subseteq \T$, we get
   2654     $$\limn\frac{S_n(A)(x)}n = \lambda(A),$$
   2655     where $\lambda(A)$ is the Lebesgue measure of $A$. This even converges uniformly in $x\in X$.
   2656 \end{prop}
   2657 
   2658 \begin{defin}
   2659     We call the limit the asymptotic frequency of visits to $A$.
   2660 \end{defin}
   2661 
   2662 \begin{lem}
   2663     If $A,B\subseteq X$ are arcs with $\lambda(B) < \lambda(A)$, then there is an $N_0\in\Ns$ such that
   2664     $$S_{n+N}(A) > S_n(B)$$
   2665     for each $N\geq N_0$ and $n\in\N$.
   2666     \label{Monotonicity}
   2667 \end{lem}
   2668 
   2669 \begin{proof}
   2670     Since $\lambda(B)<\lambda(A)$ and all orbits are dense, there is $N_0 \in\Ns$ such that $T^{N_0}B\subseteq A$, hence $T^nx\in B$ implies $T^{n+N_0}x\in A$. Therefore
   2671     $$S_n(B)\leq S_{n+N_0}(A) \leq S_{n+N}(A)$$
   2672     for all $N \geq N_0$.
   2673 \end{proof}
   2674 
   2675 Let
   2676 $$\overline S(A) := \limsup_{n \to \infty} \reci nS_n(A)$$
   2677 and
   2678 $$\underline S(A) := \liminf_{n \to \infty} \reci nS_n(A).$$
   2679 
   2680 \begin{lem}
   2681     The following statements hold:
   2682     \begin{enumerate}
   2683         \item If $A,B$ are arcs with $\lambda(B) < \lambda(A)$, then
   2684         $$\overline S(A) \geq \overline S(B)$$
   2685         on $\T$.
   2686 
   2687         \item If $A_1,\dots,A_m$ are pairwise disjoint arcs, then
   2688         $$\overline S\left(\bigcup_{j=1}^mA_j\right) \leq \sum_{j=1}^m \overline S(A_j).$$
   2689 
   2690         \item $\overline S(A) = 1-\underline S(A^c)$.
   2691     \end{enumerate}
   2692 \end{lem}
   2693 
   2694 \begin{proof}
   2695     We only provide proof for the first two statements.
   2696     \begin{enumerate}
   2697         \item We use lemma \ref{Monotonicity}.
   2698         $$\overline S(B)=\limsup_{n\to \infty} \reci nS_n(B) \leq \limsup_{n\to\infty}\reci n S_{n+N_0}(A) = \limsup_{n\to\infty}\underbrace{\frac{n+N_0}n}_{\to1}\reci{n+N_0} S_{n+N_0}(A) = \overline S(A)$$
   2699 
   2700         \item Obviously we get
   2701         $$S_n\left(\bigcup_{j=1}^mA_j\right) = \sum_{j=1}^mS_n(A_j)$$
   2702         for a disjoint union. So
   2703         $$\limsup_{n\to\infty}\reci n\underbrace{S_n\left(\bigcup_{j=1}^mA_j\right)}_{=\sum_{j=1}^mS_n(A_j)} = \limsup_{n\to\infty}\sum_{j=1}^m\reci nS_n(A_j)\leq \sum_{j=1}^m\limsup_{n\to\infty}\reci nS_n(A_j) = \sum_{j=1}^m \overline S(A_j).$$
   2704 
   2705         \item Exercise
   2706     \end{enumerate}
   2707 \end{proof}
   2708 
   2709 \begin{lem}
   2710     If $A$ is an arc with $\lambda(A) = \reci k$, then $\overline S(A) \leq \reci {k-1}$.
   2711 \end{lem}
   2712 
   2713 \begin{proof}
   2714     Take $k-1$ pairwise disjoint arcs $A_1,\dots,A_{k-1}$ of length $\reci{k-1}$ such that $X = A_1 \cup \dots \cup A_{k-1}$. By Lemma \ref{Monotonicity}, there is $N_j \in \Ns$ such that $S_{n+N}(A_j)\geq S_n(A)$ for all $N \geq N_j$.
   2715     $$\sum_{j=1}^{k-1}S_{n+N}(A_j)\geq(k-1)S_n(A)$$
   2716     for all $N \geq \max_{j\in \jbr{k-1}} N_j =: \overline N$. The left side is
   2717     $$S_{n+N}\left(\bigcup_{j=1}^{k-1}A_j\right)=S_{n+N}(\T)=n+N,$$
   2718     so we get
   2719     $$\reci{k-1}\geq \frac{n}{n+N}\reci nS_n(A).$$
   2720     Letting $n$ approach infinity, we get
   2721     $$\reci{k-1}\geq \overline S(A).$$
   2722 \end{proof}
   2723 
   2724 \begin{proof}[Proof of proposition]
   2725     Take any arc $A\subseteq X$, $\epsilon > 0$. Choose another arc $B\supseteq A$ such that
   2726     $$\lambda(A)\leq\lambda(B)=\frac lk<\lambda(A)+\epsilon$$
   2727     with $k = k(\epsilon)>\reci\epsilon$. Then
   2728     $$\overline S(A)\leq\overline S(B)=\overline S\left(\bigcup_{j=1}^lB_j\right)\leq \sum_{j=1}^l\overline S(B_j) \leq \frac l{k-1}=\frac lk\frac{k}{k-1} < (\lambda(A)+\epsilon)\frac k{k-1}$$
   2729     with $B=\bigcup_{j=1}^l B_j$, a disjoint union with $\lambda(B_j)=\reci k$ for $j \in\jbr{l}$. Letting $\epsilon$ approach $0$ and $k$ approach $\infty$, we get
   2730     $$\overline S(A)\leq \lambda(A)$$
   2731     Also, $\underline S(A)\geq \lambda(A)$, because of $\overline S(A)=1-\underline S(A^c)$. We conclude that
   2732     $$\underline S(A) = \overline S(A) = \lambda(A).$$
   2733 \end{proof}
   2734 
   2735 \paragraph{An application:}
   2736 Take $k\in\Ns$ which is not a power of $10$. Let $p\in\jbr{9}$. Consider $k^j$ where $j\in\Ns$. Let
   2737 $$\sigma_n:=\#\{j\in\jbr{0,n-1}: \text{ the decimal expansion of } k^j \text{ starts with } p\}.$$
   2738 Then
   2739 $$\limn\frac{\sigma_n}n =\log_{10}(p+1)-\log_{10}(p).$$
   2740 
   2741 \begin{exam}
   2742     Let $k=2, n=10$ and $p = 1$. Then we take a look at the powers of $2$ up to $2^9$.
   2743     $$1,2,4,8,16,32,64,128,256,512$$
   2744     The numbers $1,16$ and $128$ start with $1$, so
   2745     $$\sigma_{10}=3.$$
   2746 \end{exam}
   2747 
   2748 \begin{proof}
   2749     The number $k^j$ starts with the digit $p$. This means there exists $l\in\N$ such that $k^j=p\cdot10^l+q$, where $0\leq q < 10^l$. We can write it like this:
   2750     $$p = \frac{k^j}{10^l}$$
   2751     Doing some algebra, we get
   2752     $$0<\log_{10}(p)=\log_{10}(k^j)-\log_{10}(10^l)= j\cdot\log_{10}k-l<\log_{10}(p+1)\leq1.$$
   2753     Obviously we can choose $l$ such that $j\cdot\log_{10}k-l$ lands in the interval $(0,1]$. Let $\alpha:=\log_{10}k\notin\Q$ and $T:\T\to\T$ with $Tx:=x+\alpha$. Then $T^jx=x+j\alpha$. By the proposition, we get
   2754     $$\frac{\sigma_n}{n}=\frac{S_n(A_p)(0)}{n}\to \lambda(A_p).$$
   2755 \end{proof}
   2756 
   2757 \subsubsection{More general circle maps}
   2758 
   2759 Let $X = \T$, $\pi:\R\to \T$, $\pi(x) = x \pmod 1$. Let $T:\T\to\T$ be continuous (often a homeomorphism). Recall that rotations are defined such that $Tx = x+\alpha$. If $\alpha \in \Q$ then all orbits are periodic with the same periodicity. Otherwise all orbits are dense.
   2760 \newline
   2761 \newline
   2762 Question: Can we classify rotations?\\
   2763 Recall $S$ and $T$ are topologically conjugate ($T\cong S$) if there is a homeomorphism $\eta:\T\to\T$ such that $\eta\circ T=S\circ \eta$. Equivalently $S = \eta \circ T\circ \eta^{-1}$.
   2764 
   2765 \begin{defin}
   2766     If $T$ has a periodic point, we define $\Per(T)$ to be the smallest $k\in\N$ such that there is a $x\in \T$ which is $k$-periodic.
   2767 \end{defin}
   2768 
   2769 \begin{hw}
   2770     Show that if $S\cong T$ then $\Per(S)=\Per(T)$. This means that $\Per(T)$ is invariant under topologically conjugacy. Therfore if $\Per(S) \neq \Per(T)$, we can conclude that $S$ and $T$ are not topologically conjugate.
   2771 \end{hw}
   2772 
   2773 The question is: What happens if $T$ and $S$ are non-periodic?
   2774 \newline
   2775 \newline
   2776 General $T$ can display more complicated dynamics. For example let
   2777 $$\Delta_L(x):= Lx-x$$
   2778 be a displacement function where $L:\R\to\R$ satisfies $\pi\circ L = T \circ \pi$. In the case of the rotation, the displacement function is just $\alpha$ which is then $1$-periodic.
   2779 
   2780 \begin{prop}
   2781     If $T:\T\to\T$ is continuous, then there exists a continuous $L:\R\to\R$ with
   2782     $$\pi\circ L=T\circ \pi.$$
   2783 \end{prop}
   2784 
   2785 \begin{defin}
   2786     Any such $L$ is a lift of $T$ to $\R$.
   2787 \end{defin}
   2788 
   2789 \begin{proof}
   2790     The idea is to keep track of how many rounds the point has done. This number is then added to the position.
   2791 \end{proof}
   2792 
   2793 \begin{prop}
   2794     If $T: \T\to\T$, then:
   2795     \begin{enumerate}
   2796         \item If $L$ is a lift of $T$ then $\{L+m:m\in\Z\}$ is the collections of all lifts.
   2797 
   2798         \item $L(x+1)-L(x)$ is an integer independent of $x$ and $L$.
   2799     \end{enumerate}
   2800 \end{prop}
   2801 
   2802 \begin{defin}
   2803     For a given map $T$ we define
   2804     $$\deg(T):=L(x+1)-L(x)$$
   2805     to be the degree of $T$. This does not depend on $x$.
   2806 \end{defin}
   2807 
   2808 \begin{rem}
   2809     $\deg(T)$ counts how often $T$ maps around $\T$.
   2810 \end{rem}
   2811 
   2812 \begin{rem}
   2813     $L(x+m)=L(x)+m\deg(T)$
   2814 \end{rem}
   2815 
   2816 \begin{prop}
   2817     Additionally:
   2818     \begin{enumerate}
   2819         \setcounter{enumi}{2}
   2820         \item If $T$ is a homeomorphism of $\T$, then $L$ is a homeomorphism of $\R$ and $\deg(T)=\pm1$.
   2821     \end{enumerate}
   2822 \end{prop}
   2823 
   2824 \begin{rem}
   2825     The sign of the degree gives information about the orientation.\\
   2826     $\deg(T)=1$ means $T$ is orientation preserving (o.p.).\\
   2827     $\deg(T)=-1$ means $T$ is orientations reversing (o.r.).
   2828 \end{rem}
   2829 
   2830 \begin{prop}
   2831     In addition:
   2832     \begin{enumerate}
   2833         \setcounter{enumi}{3}
   2834         \item $L_\Delta(x) := L(x)-\deg(T)x$ is continuous, bounded, $1$-periodic on $\R$.
   2835 
   2836         \item $L^m$ is a lift of $T^m$ for all $m\in\Z$.
   2837 
   2838         \item If $S,T$ are orientation preserving homeomorphisms of $X$ with $S=\eta\circ T\circ \eta^{-1}$ and $L,H$ lifts of $T,\eta$, then $H\circ L\circ H^{-1}$ is a lift of $S$.
   2839     \end{enumerate}
   2840 \end{prop}
   2841 
   2842 \begin{proof}
   2843     Again, we only prove the first two statements, since the rest is easy to show.
   2844     \begin{enumerate}
   2845         \item If $L$ is a lift of $T$ then $\tilde L:=L+m$ ist also a lift. If $L,\tilde L$ are lifts then
   2846         $$\pi\circ \tilde L = T\circ \pi=\pi\circ L.$$ Therefore $\tilde L(x)-L(x)\in\Z$ for all $x$. Since $\tilde L-L$ is continuous, $\tilde L-L$ is constant.
   2847 
   2848         \item We easily see that
   2849         $$\pi(L(x+1))=T(\underbrace{\pi(x+1)}_{=\pi(x)})=\pi(L(x)).$$ Therefore $L(x+1)-L(x)\in\Z$ is constant.
   2850     \end{enumerate}
   2851 \end{proof}
   2852 
   2853 \begin{lem}
   2854     Let $T$ be an orientation preserving homeomorphism of $\T$ with lift $L$. Then for $\Delta_L(x) := Lx-x$ we get
   2855     $$\Delta_L(x)\leq \Delta_L(y)+1$$
   2856     for all $x,y\in\R$.
   2857 \end{lem}
   2858 
   2859 \begin{proof}
   2860     Set $k:=\lfloor y-x\rfloor\in\Z$ and $\theta\in[0,1)$ such that $y=x+k+\theta$. Then
   2861     $$\Delta_L(y)=L(y)-L(x+k)+L(x+k)-(x+k)+\underbrace{(x+k)-y}_{=-\theta\leq0} \leq \underbrace{L(y)-L(x+k)}_{\leq 1 \text{ since } y\in[x+k,x+k+1]}+\Delta_L(x+k)$$
   2862 \end{proof}
   2863 
   2864 \begin{rem}[Recall:]
   2865     If $T$ is a orientation preserving homeomorphism of $X$ with lift $L$, then
   2866     $$\rho(L)(x):=\limn\frac{L^nx-x}n$$
   2867     exists in $\R$, is an integer and doesn't depend on $x$. So we define
   2868     $$\rho(L):=\rho(L)(x)$$
   2869     and we get
   2870     $$\rho(L+m)=\rho(L)+m$$
   2871     for all $m\in\Z$.
   2872 \end{rem}
   2873 
   2874 \begin{exam}
   2875     Let $T:\T\to\T$ with $Tx=x+\alpha$ and $L:\R\to\R$ with $Lx=x+\alpha$. Then
   2876     $$\rho(L)=\limn\frac{x+n\alpha-x}n = \alpha.$$
   2877 \end{exam}
   2878 
   2879 So we can say that $\rho(L)$ describes the average rotation speed.
   2880 
   2881 \begin{defin}
   2882     We say that $\rho(L)$ is the rotation number of $L$. Also $\rho(T):=\pi(\rho(L))$ is the rotation number of $T$.
   2883 \end{defin}
   2884 
   2885 \begin{rem}
   2886     $\rho(T^q)=q\rho(T)\pmod1$
   2887 \end{rem}
   2888 
   2889 \begin{lem}
   2890     If $(a_n)_{n\in\Ns}\subseteq\R$ are such that there exist $k\in\N,\kappa\in\R$ with
   2891     $$a_{m+n}\leq a_m+a_{n+k}+\kappa\quad (*)$$
   2892     for every $m,n\in\Ns$, then the following limit exists:
   2893     $$\limn \frac{a_n}n\in[-\infty,\infty)$$
   2894 \end{lem}
   2895 
   2896 \begin{proof}
   2897     We split this proof into three parts.
   2898     \begin{enumerate}
   2899         \item We show that we can assume wlog that $k=0$. Using $(*)$ with the switched roles $m \mapsto n, n\mapsto k, k \mapsto k$, we get
   2900         $$a_{n+k}\leq a_n+a_{2k}+\kappa $$
   2901         for all $n\in\Ns$. Therefore
   2902         $$a_{m+n}\leq a_m+a_{n+k}+\kappa \leq a_m+a_n+a_{2k}+\kappa$$
   2903         for all $n,m\in\Ns$. Setting $k=0$ and replacing $a_0+\kappa$ by $\kappa$, we get the result
   2904         $$a_{m+n} \leq a_m+a_n\kappa.$$
   2905 
   2906         \item Now we show that
   2907         $$\liminf_{n\to\infty}\frac{a_n}n\in\R.$$
   2908         By induction, we get
   2909         $$a_n = a_{1+\dots+1} \leq na_1+(n-1)\kappa.$$
   2910         Therefore
   2911         $$\left|\frac{a_n}{n}\right| \leq |a_1|+\frac{n-1}{n}|\kappa| \leq |a_1|+|\kappa|.$$
   2912         We showed that $\left(\frac{a_n}{n}\right)_n$ is bounded, so the limes inferior is real.
   2913 
   2914         \item In the case $k=0$, let
   2915         $$a:=\liminf_{n\to\infty}\frac{a_n}n\in\R.$$
   2916         Take $\epsilon > 0$, we show that
   2917         $$\limsup_{n\to\infty}\frac{a_n}n\leq a+\epsilon.$$
   2918         Fix $n\in\N$ such that $\frac{a_n}n < a+\frac\epsilon3$ and $\frac\kappa n<\frac\epsilon3$. Any $l\geq n$ can be written as
   2919         $$l=jn+r$$
   2920         with $j\in\N$, $r\in\jbr{0,n-1}$, then
   2921         $$\frac{a_l}l=\frac{a_{n+\dots+n+r}}l\leq\underbrace{\frac{ja_n}l}_{j \leq \frac ln}+\frac{a_r}l+\frac{j\kappa}l\leq\frac{a_n}n+\underbrace{\frac{\max(a_1,\dots,a_{n-1})}l}_{l \text{ large}}+\frac\kappa n < a+\frac{\epsilon}{3}+\frac{\epsilon}3+\frac\epsilon3<a+\epsilon$$
   2922         if $l \geq l_0$.
   2923     \end{enumerate}
   2924 \end{proof}
   2925 
   2926 \begin{proof}[Proof of Remark.]
   2927     We want to show that $\rho(L)(x)=\limn\frac{L^nx-x}{n}$ exists and doesn't depend on $x$.
   2928     \begin{enumerate}
   2929         \item We claim that $\rho(L)(x)$ is independent of $x$. Since $L$ is a homeomorphism of $\R$ and a lift of a orientation preserving homeomorhpism, we know $L(x+1)=L(x)+1$. So
   2930         $$|x-y|<1$$
   2931         implies
   2932         $$|L^nx-L^ny|<1.$$
   2933         Hence, for $x,y\in(0,1]:$
   2934         $$\left|\frac{L^nx-x}n-\frac{L^ny-y}n\right|\leq\reci n(|L^nx-L^ny|+|x-y|)<\frac2n\to0,$$
   2935         since $\Delta_{L^n}(x)=L^nx-x$ is $1$ periodic on $\R$. Therefore, if the limit exists then it doesn't depend on the input.
   2936 
   2937         \item Fix $x\in\R$, let $a_n:=L^nx-x$. We want to show that $\limn\frac{a_n}n\in\R$. Then
   2938         $$a_{m+n}=L^{m+n}x-x=\underbrace{L^m(L^nx)-L^nx}_{\Delta_{L^m(y)}}+\underbrace{L^nx-x}_{=a_n},$$
   2939         where $y:=L^nx$. By the lemma, this is less than or equal to
   2940         $$\Delta_{L^m}(x)+1+a_n=a_m+a_n+1$$
   2941         for all $m,n\in\Ns$. By the subadditivity lemma we have the limit
   2942         $$\limn\frac{a_n}n\in[-\infty,\infty)$$
   2943         and it's also larger than $-\infty$. So
   2944         $$\frac{a_n}n = \reci n\sum_{k=0}^{n-1}\underbrace{\Delta_L(T^kx)}_{\geq c > -\infty}\geq c > -\infty.$$
   2945 
   2946         \item Rest: Do it yourself.
   2947     \end{enumerate}
   2948 \end{proof}
   2949 
   2950 \begin{rec}
   2951     $T\cong S$ if they are topologically conjugate. That means that there is a homeomorphism $\eta:\T\to\T$ such that $\eta\circ T=S\circ \eta$.
   2952 \end{rec}
   2953 
   2954 \begin{rec}
   2955     For periodic rotations, $\Per(T)$ is invariant: If $S\cong T$ then $\Per(S) = \Per(T)$. The converse is not true.
   2956 \end{rec}
   2957 
   2958 \begin{prop}
   2959     Let $T,S: \T\to\T$ be orientation preserving homeomorphisms of $\T$, with $S=\eta\circ T\circ\eta^{-1}$, then $\rho(T)=\rho(S)$. So $\rho(T)$ is an invariant for topologically conjugacy.
   2960 \end{prop}
   2961 
   2962 \begin{exam}
   2963     For all $\alpha,\beta \in (0,1]\setminus\Q$ with $\alpha \neq \beta$ and $\alpha \neq 1-\beta$ the corresponding rotations $T,S$ are not conjugate.
   2964 \end{exam}
   2965 
   2966 \begin{proof}
   2967     Let $L$ and $H$ be lifts of $T$ and $\eta$. Then $H\circ L\circ H^{-1}$ is a lift of $S$. For $x\in\R$
   2968     $$\frac{(H\circ L\circ H^{-1})^nx-x}n=\frac{(H\circ L^n\circ H^{-1})x-x}n=\frac{H(L^n(H^{-1}x))-L^n(H^{-1}x)}n+\frac{L^n(H^{-1}x)-H^{-1}x}n+\frac{H^{-1}x-x}n$$
   2969     $$=\underbrace{\frac{\Delta_H(L^n(H^{-1}x))}n}_{\to0}+\frac{L^ny-y}n+\underbrace{\frac{\Delta_{H^{-1}}(x)}n}_{\to0} \to \rho(L).$$
   2970     So the rotation numbers coincide.
   2971 \end{proof}
   2972 
   2973 \begin{prop}
   2974     If $T:\T\to\T$ is an orienation preserving homeomorphism, then $\rho(T) \in\Q$ if and only if $T$ has a periodic point.
   2975 \end{prop}
   2976 
   2977 \begin{proof}
   2978     We show both directions.
   2979     \begin{enumerate}
   2980         \item $\Leftarrow:$ Suppose $T^qx=x$ for some $q\in\Ns$. Take a lift of $T$, then $L^qx=x+p$ for some $p\in\Z$. Then, $L^{mq}x = x+mp$ for all $m \in \N$. Hence
   2981         $$\rho(L)=\lim_{m\to\infty} \frac{L^{mq}x-x}{mq}=\lim_{m\to\infty}\reci{mq}\sum_{k=0}^{m-1}\underbrace{(L^{(k+1)q}x-L^{kq}x)}_{=p}=\frac{mp}{mq}=\frac pq \in\Q.$$
   2982 
   2983         \item $\Rightarrow:$ Assume that $\rho(L)=\frac pq$ for some lift $L$ of $T$. Then
   2984         $$\rho(L^q)=p\rho(L) = p\cdot\frac pq = p = 0 \pmod1$$
   2985         Therefore $\rho(T^q) = 0$. We show that $S:=T^q$ has a fixed point. We know that $\rho(S)=0$. Take a lift $L$ of $S$ with $L(0) \in [0,1)$. We have
   2986         $$\Delta_L(x)=Lx-x\notin \Z$$
   2987         for $x \in \R$ (otherwise: $S(\pi(x))=\pi(L(x))=\pi(x+m)=\pi(x)$, so $S = id$, which has fixed points). Since
   2988         $$\Delta_L(0)=L(0)-0=L(0)\in(0,1)$$
   2989         and $\Delta_L$ is continuous on $\R$, $0<\Delta_L<1$ on $\R$. By continuity on $[0,1]$ there is a
   2990         $$0<\delta\leq \Delta_L(x)\leq (1-\delta)<1 \quad\forall x \in\R.$$
   2991         Then
   2992         $$L^n0=L^n0-0 = \sum_{k=0}^{n-1}\underbrace{(L(L^k0)-L^k0)}_{\Delta_L(L^k0)}$$
   2993         for $n\in\Ns$. Then
   2994         $$\frac{n\delta}n \leq \underbrace{\frac{L^n(0)-0}n}_{\to\rho(L)}\leq \frac{n(1-\delta)}n = 1-\delta<1.$$
   2995         Therefore $\rho(L) \neq 0 \pmod1$. This is a contradiction.
   2996     \end{enumerate}
   2997 \end{proof}
   2998 
   2999 \subsubsection{Circle homeomorphisms with periodic points}
   3000 
   3001 \begin{prop}
   3002     Let $T:\T\to\T$ be an orientation preserving homeomorphism, $\rho(T)=\frac pq$ with $p,q$ relatively prime. Then, for every periodic point $x$ of $T$:
   3003     \begin{enumerate}
   3004         \item $x$ has minimal period $q$.
   3005 
   3006         \item The ordering of $(x,Tx,\dots, T^{q-1}x)$ on $\T$ is the same as the ordering of $(x,Sx,\dots,S^{q-1}x)$ where $Sx:=x+\frac pq \pmod1$.
   3007     \end{enumerate}
   3008 \end{prop}
   3009 
   3010 \begin{exam}
   3011     Let $T$ be an orientation preserving homeomorphism with $\rho(T)=\frac23$. Let $z$ be a $3$-periodic point. Let $S:\T\to\T$ with $Sx=x+\frac23$. We observe the ordering $(z, Sz, S^2z)$ with $z\in \T$. Those are oriented clockwise. For  the points $(z,Tz,T^2z)$ also have the same orientation.
   3012 \end{exam}
   3013 
   3014 \begin{exam}
   3015     Suppose that $T$ has exactly one $2$-cycle: $\{x_1,x_2\}$, $Tx_1=x_2$ and $Tx_2=x_1$. So both are fixed points of $S:=T^2$. Without loss of generality we assume that $x_1 = 0$. Then for all $y$ there is an $\overline m\in\{1,2\}$ such that
   3016     $$d(T^{2n}y,T^{2n}x_{\overline m}) \to 0$$
   3017     since $T$ is continuous:
   3018     $$d(T^{2n+1}y,T^{2n+1}x_{\overline m}) \to 0.$$
   3019     So $d(T^ny,T^nx_{\overline m}) \to 0$ as $n\to\infty$.
   3020 \end{exam}
   3021 
   3022 \begin{thm}
   3023     Let $T:\T\to\T$ be a orientation preserving homeomorphism with $\rho(T) = \frac pq$ with $\{x_1,\dots,x_M\}$ of $q$-periodic points, where $x_1,\dots,x_M$ are ordered on a circle. Then for all $m\in\jbr{1,M}$ there is $\overline m\in \{m,m+1\}\pmod M$ such that for all $y \in (x_m,x_{m+1})$ we get
   3024     $$d(T^ny,T^nx_{\overline m}) \to 0.$$
   3025 \end{thm}
   3026 
   3027 Let $C:=\{x:T^qx\}$ be closed. In the case of a rational rotation, $C = \T$. Sometimes $C$ consists of finitely many points. Sometimes $C$ is an interval.
   3028 
   3029 \begin{lem}
   3030     Let $J=[a,b]\subseteq\T$ a closed interval, $f:J\to J$ a homeomorphism, increasing with $f(y)\neq y$ for all $y\in(a,b)$ then for all $y \in (a,b)$
   3031     $$\limn f^n(y)=\begin{cases}
   3032         a & \text{if } f(x)<x \text{ on } $(a,b)$\\
   3033         b & \text{ otherwise }
   3034     \end{cases}$$
   3035 \end{lem}
   3036 
   3037 \begin{proof}
   3038     Case $f(x) < x$: Then $y>f(y)>f^2(y)>\dots$. This sequence is monotone and bounded, so it has a limit $z$. By continuity, we get $f(z) = z$ and $z=a$. The other case works similarly.
   3039 \end{proof}
   3040 
   3041 If we apply this to $T^q$, then for any $y\in\T\setminus C$ there is an $x\in C$ such that
   3042 $$d(T^{mq}y,T^{mq}x)\to0.$$
   3043 By continuoity, we get
   3044 $$d(T^{mq+r}x,T^{mq+r}y)\to0$$
   3045 for $r\in\jbr{1,q-1}$ and $m\to\infty$.
   3046 
   3047 \begin{thm}
   3048     For $T$ as above with $\rho(T)\in\Q$, there is a nonempty closed $C\subseteq \T$ of $q$-periodic points (same ordering as for rotation $S$). If $I$ is a maximal interval $\T\setminus C$, then there exists a $x\in \partial I\subseteq C$ such that for all $y\in I$
   3049     $$d(T^ny,T^nx)\to0.$$
   3050 \end{thm}
   3051 
   3052 \begin{rec}
   3053     The $\omega$-limit of $x$ is
   3054     $$\omega(x):=\{z\in X: (k_n)_n: k_n\to\infty \text{ and } T^{k_n}x\to z\}.$$
   3055 \end{rec}
   3056 
   3057 \begin{hw}
   3058     If $T$ is continuous:
   3059     \begin{enumerate}
   3060         \item $\omega(x)$ is closed
   3061 
   3062         \item $T\omega(x) \subseteq \omega(x)$
   3063     \end{enumerate}
   3064 \end{hw}
   3065 
   3066 
   3067 \begin{exam}
   3068     In the theorem $\omega(x)=\{x,Tx,\dots,T^{q-1}x\}$ and $\omega(y)=\omega(x)$.
   3069 \end{exam}
   3070 
   3071 Now we take a look at orientation preserving homeomorphisms with no periodic points. So $\rho(T) = \alpha\notin\Q$.
   3072 
   3073 \begin{rec}
   3074     For $Sx:=x+\alpha$, every orbit is dense ($\omega(x)=\T$).
   3075 \end{rec}
   3076 
   3077 \begin{lem}
   3078     Let $T$ be as above, $m \in \jbr{n-1}$ and $I$ be a closed arc with endpoints $T^nx, T^mx$ for $x\in\T$. Then for any $y\in\T$ the orbit $(T^ky)_{k\in\N}$ meets $I$.
   3079 \end{lem}
   3080 
   3081 \begin{proof}
   3082     We need to show that $X\subseteq \bigcup_{k\in\N}T^{-k}I$. Consider $I_k:=T^{-k(n-m)}I$, where $k\in\N$. Then $I_k,I_{k+1}$ have a common endpoint. Suppose $\bigcup I_k \neq \T$, then $J_m := (I_1\cup\dots\cup I_m)_m$ is growing. The right ends of $J_m$ calling it $(z_m)_m$ points is a monotone sequence. But then, $(z_m)_m$ has a finite limit. So define
   3083     $$z=\lim_{k\to\infty} T^{-k(n-m)}(T^mx) = \lim_{k\to\infty}T^{-(k-1)(n-m)}(T^mx) = \lim_{k\to\infty}T^{n-m}(T^{-k(n-m)}(T^mx)) = T^{n-m}z$$
   3084     Therefore $z$ is a periodic point. So we have a contradiction. So $\T = \bigcup_{k\in\N}I_k$.
   3085 \end{proof}
   3086 
   3087 \begin{prop}
   3088     Let $T$ be as above, then
   3089     \begin{enumerate}
   3090         \item $\omega(x)=\omega(y)$ for all $x,y\in\T$.
   3091 
   3092         \item $E:=\omega(x)$ is a perfect set (it has no isolated points).
   3093 
   3094         \item $E=\T$ or $E$ is nowhere dense (the interior of the closure is empty).
   3095 
   3096         \item Cantor set: It's perfect and nowhere dense.
   3097     \end{enumerate}
   3098 \end{prop}
   3099 
   3100 \begin{proof}
   3101     We only proof the first statement. Maybe we'll do the others in the future.
   3102     \begin{enumerate}
   3103         \item Take $z\in\omega(x)$, then there exists $(l_n)_n$ going to $\infty$ such that $T^{l_n}x\to z$. Consider $J_n:=[T^{l_n}x,T^{l_{n+1}}x]$ (the shorter interval). By the lemma there is a sequence $k_n\to\infty$: $T^{k_n}y\in J_n$ (fill in the details).
   3104 
   3105         Since $T^{l_n}x\to z$, we get $d(J_n,z)\to0$. So $T^{k_n}y\to z$ which proves that $z\in \omega(y)$.
   3106 
   3107         \item Not now
   3108 
   3109         \item Not now
   3110     \end{enumerate}
   3111 \end{proof}
   3112 
   3113 \begin{thm}
   3114     Let $T:\T\to\T$ be an orientation preserving homeomorphism of a circle with $\alpha := \rho(T)\notin\Q$. Then there exists a monotone topological semiconjugacy $\eta:\T\to\T$ with $\eta\circ T=S\circ \eta$, where $S$ is a rotation.
   3115 
   3116     \begin{enumerate}
   3117         \item If $T$ is topologically transitive (so $E = \T$), then $\eta$ is a homeomorphism, so a topological conjugate.
   3118 
   3119         \item Otherwise: $\eta$ is not injective ($S$ is just a factor of $T$).
   3120     \end{enumerate}
   3121 \end{thm}
   3122 
   3123 \begin{thm}
   3124     Depending on the smoothness of $\eta$ there are different possibilities.
   3125     \begin{enumerate}
   3126         \item For $C^1$ diffeomorphisms both (1) and (2) are possible.
   3127 
   3128         \item For $C^2$ diffeomorphisms only (1) is possible.
   3129     \end{enumerate}
   3130 \end{thm}
   3131 
   3132 \subsection{Maps with complicated orbit structure}
   3133 
   3134 \subsubsection{Warmup}
   3135 
   3136 Very simple maps on the interval/circle can have very complicated dynamics.\\
   3137 Limitations for predictions, often due to:
   3138 
   3139 \begin{defin}
   3140     Let $(X,d)$ be a metric space. The map $T:X\to X$ has sensitive dependence (on initial conditions) if there is a $\delta > 0$ (the sensitivity constant) such that for all $x\in X$ and $\epsilon > 0$ there exists $y\in Y$ such that
   3141     $$d(x,y)< \epsilon,$$
   3142     but
   3143     $$d(T^nx,T^ny) \geq \delta$$
   3144     for some $n\in\Ns$.
   3145 \end{defin}
   3146 
   3147 \begin{exam}[Prototypical]
   3148     Let $X = \T$.
   3149     \begin{itemize}
   3150         \item Angle doubling map: Let $T:\T\to\T$ with $Tx:=2x$. We take a look at the lift $L:\R\to\R$, with $Lx:=2x$. More generally, we can replace $2$ by an integer $m\geq 2$ where $m=\deg(T)$. In this case, we have $m$ cylinder sets (maximal subintervals, such that $T$ is injective on it). That means that $T|_z$ is injective, where $z$ is a branch. $T^n$ has $m^n$ branches.
   3151     \end{itemize}
   3152 \end{exam}
   3153 
   3154 \begin{defin}
   3155     A uniformly expanding map on $\T$ is a map $T\in C^1(\T,\T)$ with $|T'|\geq \delta > 1$. So there is a lift $L$.
   3156 \end{defin}
   3157 
   3158 It can be shown that $T^n$ has $m^n$ cylinders and $\diam(Z)\leq \rho^{-n}$ for each cylinder $Z$. Check that we can assume that $T0=0$ (by shifting).
   3159 
   3160 \subsubsection{Basic properties}
   3161 
   3162 \paragraph{Periodic orbits:}
   3163 We know that $T^n$ has $m^n$ branches. Each branch meets the diagonal at exactly one point. Therefore, there are $m^n-1$ fixed points of $T^n$. So the set of the periodic points is dense (since the $\diam(Z)\to0$ for every cylinder $Z$ as $n\to\infty$).
   3164 
   3165 \paragraph{Question:}
   3166 What about non-periodic points?
   3167 
   3168 \begin{prop}
   3169     Every uniformly expanding map $T$ on $\T$ has sensitive dependence.
   3170 \end{prop}
   3171 
   3172 \begin{proof}[Proof for $Tx=2x$]
   3173     If $x\neq y$ with $d(x,y) < \reci4$. Then $d(Tx,Ty)=2d(x,y)$. So we can take $\delta = \reci4$.
   3174 \end{proof}
   3175 
   3176 \begin{itemize}
   3177     \item There are always countably infinitely many periodic points.
   3178 
   3179     \item $T$ has sensitive dependence.
   3180 
   3181     \item What about other types of orbits? Dense orbits? Recurrent orbits?
   3182 \end{itemize}
   3183 
   3184 \begin{rec}
   3185     Continuous map $T$ on $\T$ is topologically transitive if and only if for all open $U,V\neq \emptyset$ there is an $n\in\Ns$ such that $U\cap T^{-n}V\neq \emptyset$.
   3186 \end{rec}
   3187 
   3188 \begin{defin}
   3189     $T$ on $X$ is topologically mixing, if for all nonempty open $U,V\subseteq X$ there is a $N\in\N$ such that $U\cap T^{-n}V\neq \emptyset$ for all $n\geq N$.
   3190 \end{defin}
   3191 
   3192 \begin{exam}
   3193     No rotation is topologically mixing, since:
   3194 \end{exam}
   3195 
   3196 \begin{prop}
   3197     If $X$ has at least three different points and $T$ is an isometry. Then $T$ is not topologically mixing.
   3198 \end{prop}
   3199 
   3200 \begin{proof}
   3201     Exercise.
   3202 \end{proof}
   3203 
   3204 \begin{exam}
   3205     The doubling map is topologically mixing. Without loss of generality $U,V$ are open intervals. So $\diam(U) > 0$. Take $N$ so large that $\diam(U)>2\cdot 2^{-N}=\diam(Z)$ where $Z$ is a cylinder of $T^n$. In this case there is a cylinder $Z_0$ lying completely in $U$. So $\T = T^nZ_0\subseteq T^nU$.
   3206 \end{exam}
   3207 
   3208 \begin{prop}
   3209     Every uniformly expanding circle map $T$ is topologically mixing.
   3210 \end{prop}
   3211 
   3212 \begin{proof}
   3213     Same argument: $T^N$ has $m^n$ branches/cylinders $Z$ with $TZ = \T$. We know that $\diam(Z)\leq \rho^{-N}$ and do the same.
   3214 \end{proof}
   3215 
   3216 \begin{prop}
   3217     $T$ as above has dense orbits.
   3218 \end{prop}
   3219 
   3220 \begin{prop}
   3221     If $T\in C(X,X)$ and $X$ has at least $2$ points, then topologically mixing implies sensitive dependence.
   3222 \end{prop}
   3223 
   3224 \begin{proof}
   3225     Exercise.
   3226 \end{proof}
   3227 
   3228 \subsubsection{Symbolic dynamics and coding}
   3229 
   3230 The case of the doubling map $T:\T\to\T$ with $Tx = 2x$.
   3231 \newline
   3232 \newline
   3233 Binary/dyadic expansions:
   3234 $$x = 0.\omega_0\omega_1\omega_2\dots = \sum_{j\in\N}\frac{\omega_j}{2^{j+1}},\quad \omega_j\in\{0,1\}.$$
   3235 We can imagine that $\omega_0$ splits the interval into two pieces. If $\omega_0=0$ we use the left piece. Otherwise we use the right one. Then we keep doing it with $\omega_1,\omega_2,\dots,\omega_{l-1}$ to get a smaller interval. We get the interval
   3236 $$\left[\sum_{j=0}^{n-1}\frac{\omega_j}{2^{j+1}},\sum_{j=0}^{n-1}\frac{\omega_j}{2^{j+1}}+\reci{2^r}\right] =: Z_{(\omega_0,\omega_1,\dots,\omega_{r-1})}.$$
   3237 Now our map $T$ can simply be written as
   3238 $$Tx=\omega_0+\sum_{j\in\N}\frac{\omega_{j+1}}{2^{j+1}} = \sum_{j\in\N}\frac{\omega_{j+1}}{2^{j+1}} = 0.\omega_1\omega_2\dots$$
   3239 Also
   3240 $$T^nx=0.\omega_n\omega_{n+1}\dots$$
   3241 So $T^nx\in Z_{(\omega_n,\dots,\omega_{n+r-1})}$ for all $r,n \in \N$. For every $\omega = (\omega_j)_{j\in\N}\in\Omega_2:=\{0,1\}^{\N}$ there exists an $x$ such that $x=0.\omega_0\omega_1\dots$
   3242 
   3243 \paragraph{Application:}
   3244 \begin{enumerate}
   3245     \item $T$ has $2^p-1$ points of period $\leq p$: So there exists $2^p$ tuples $(\omega_0,\dots,\omega_{p-1})$. Take $\omega=\overline{(\omega_0,\dots,\omega_{p-1},\dots)}$
   3246 
   3247     \item $T$ has dense orbits: Take $\omega$ which contains every finite block:
   3248     $$\omega=0,1,00,01,10,11,000,\dots, \hat \omega_0,\dots,\tilde \omega_{r-1},\dots$$
   3249     $\eta:\Omega_2\to\T$
   3250     $x := \eta(\omega)$, $T^nx=0.\tilde\omega_0,\dots,\tilde\omega_{r-1}\in Z_{(\omega_0,\dots,\omega_{r-1})}$. So the orbit lands in every cylinder and is therefore dense.
   3251 
   3252     \item There exists $x\in Z_{(0,0,1)}$ such that $T^nx\in Z_{(0,0,1)}$ for all $n\geq 2$. Take $\omega=0,0,1,0,1,0,1,0,1,0,1,\dots$. Such that only one time there are zwo consecutive zeros. In this case, the orbit will never return to the first quater.
   3253 
   3254     \item For every left-/right sequence there is a 0-/1 sequence. So there is a $x$ such that $T^nX$ on left/right of presented times.
   3255 
   3256     \item The shift means: $T$ can be represented on $\Omega_2$ by
   3257     $$\sigma:\Omega_2\to \Omega_2$$
   3258     the shift map with $\sigma((\omega_j)_j) := (\omega_{j+1})$. So we get a commuting diagram with $\sigma$, $\eta$ and $T$. In this case $\eta$ is a semi conjugacy (but not injective because of different expansions for the same number).
   3259 \end{enumerate}
   3260 
   3261 \paragraph{Metric on $\Omega_2$:}
   3262 Let $\omega:=(\omega_j)_j, \tilde \omega:=(\tilde \omega_j)_j$. Let $s(\omega,\tilde\omega):=\inf\{j\in\N:\omega_j\neq\tilde\omega_j\}\in[0,\infty]$. Now let $d(\omega,\tilde\omega):=2^{-s(\omega,\tilde\omega)}$. Now let
   3263 $$[\omega_0,\dots,\omega_{r-1}]:=\{\tilde\omega:\tilde\omega_j=\omega_j\quad j \in \jbr{0,r-1}\}.$$
   3264 This is a $2^{-r}$ neighborhood of $\omega$, a cylinder set in $\Omega_2$.
   3265 
   3266 \begin{prop}
   3267     The following statements hold:
   3268     \begin{enumerate}
   3269         \item $d$ is a metric.
   3270 
   3271         \item $\eta$ is continuous.
   3272     \end{enumerate}
   3273 \end{prop}
   3274 
   3275 However $\eta$ is not injective in this situation:
   3276 $$\omega_0\dots\omega_{r-1}10000000\dots = \omega_0\dots\omega_{r-1}01111111\dots$$
   3277 Such $x$ have exactly two preimages. These $x$ are exactly the endpoints of the diadic intervals. All other $x$ have exactly one preimage $\omega$ such that $x=\eta(\omega)$.
   3278 \newline
   3279 \newline
   3280 We imagine that we half the interval and shrink both parts. Then we repeat it and get something similar to the cantor set.
   3281 
   3282 \begin{prop}
   3283     $\Omega_2$ has no isolated points and is totally disconnected and compact. So this is a cantor set.
   3284 \end{prop}
   3285 
   3286 \subsubsection{The general uniformly expanding circle maps $T:\T\to\T$ of degree 2}
   3287 
   3288 Let $Z_0,Z_1$ be the two cylinders. So the natural semipartition is $S = \{Z_0,Z_1\}$ with overlapping endpoints. Previously $x\in Z_{(0,1,0)}$ meant $x \in Z_0$, $Tx\in Z_1$ and $T^2x\in Z_0$. The same holds for this general case.
   3289 $$Z_{(\omega_0,\dots,\omega_{r-1},\omega_r)} := \text{The non degenerate interval in } Z_{(\omega_0,\dots,\omega_{r-1})}\cap T^{-r}Z_{\omega_r}$$
   3290 Then $x\in Z_{(\omega_0,\dots,\omega_{r-1})}$ implies $T^jx\in Z_{\omega_j}$ for $j\in\jbr{0,r-1}$. Moreover for every $\omega=(\omega_j)_j\in\Omega_2$ there exists exactly one $x = \eta_T(\omega)\in\T$ such that $x \in Z_{(\omega_0,\dots,\omega_{r-1})}$ for all $r\in\Ns$. Indeed
   3291 $$\bigcap_{r\in\Ns} Z_{(\omega_0,\dots,\omega_{r-1})}$$
   3292 is a singleton because it's a nested sequence of non empty compact intervals and $\diam(Z_{(\dots)}) \leq \rho^{-r} \to 0$. :)\\
   3293 So we define $\eta_T:\Omega_2\to\T$ like that. For $x = \eta_T(\omega)$, consider $y := Tx$. Then $T^jy=T^{j+1}x\in Z_{\omega_{j+1}}$ for all $j\in\N$. So
   3294 $$\eta_T\circ \sigma = T\circ \eta_T.$$
   3295 Check the following:
   3296 \begin{enumerate}
   3297     \item $\eta_T$ is continuous.
   3298 
   3299     \item $x$ has two preimages if and only if $x$ is the endpoint of some cylinder. This looks like
   3300     $$\omega = \omega_0\dots\omega_{r-1}10000000\dots,\tilde \omega= \omega_0\dots\omega_{r-1}01111111\dots.$$
   3301     with $\eta(\omega)$, define $\eta_T^{-1}(x) := \omega$ (not $\tilde \omega$).
   3302 \end{enumerate}
   3303 
   3304 \begin{thm}
   3305     Any two uniformly expanding circle maps $T, \tilde T$ of degree $2$ are topologically conjugate. Hence they are isomorphic to the doubling map.
   3306 \end{thm}
   3307 
   3308 \begin{proof}
   3309     Let $\psi:= \eta_{\tilde T}\circ\eta_T^{-1}:\T\to\T$. Then
   3310     $$\psi\circ T=\tilde T\circ \psi.$$
   3311     Also $\psi$ is a bijection. \\
   3312     Claim: $\psi$ is a homeomorphism. It suffices to show that it's continuous at every $x\in\T$ (by changing the roles of $T, \tilde T$).\\
   3313     Let $\epsilon > 0$. Let $Z_{(\omega_0,\dots,\omega_{r-1})}$ be a cylinder of $T$ and $\tilde Z_{(\omega_0,\dots,\omega_{r-1})}$ be a cylinder of $\tilde T$. Take $r\in\Ns$ so large that choose four adjacent cylinders $\tilde Z(l) = \tilde Z_{(\omega_0^{(l)},\dots,\omega_{r-1}^{(l)})}$ where $l\in\jbr{1,4}$ such that $y$ is between $Z(1)$ and $Z(4)$ and all of them are contained in the $\epsilon$ neighborhood. Consider $Z(l) = Z_{(\omega_0^{(l)},\dots,\omega_{r-1}^{(l)})}$ where $l\in\jbr{1,4}$. By construction $Z(l)$ is mapped to $\tilde Z(l)$. Choose $\delta > 0$ such that $B_\delta(x)\subseteq \bigcup_{l=1}^4Z(l)$. Then
   3314     $$\psi(B_j(x))\subseteq \psi\left(\bigcup_{l=1}^4Z(l)\right)=\bigcup_{l=1}^4\tilde Z(l)\subseteq B_\epsilon(y).$$
   3315 \end{proof}
   3316 
   3317 \begin{rem}
   3318     Analogous construction works for degree $m\geq 2$ maps with the sequence space $\Omega_m:=\jbr{0,m-1}^\N=\{(\omega_j)_{j\geq0}:\omega_j\in\jbr{0,m-1}\}$.
   3319 \end{rem}
   3320 
   3321 \begin{thm}
   3322     Two uniformly expanding circle maps $T, \tilde T$ of degree $m$ and $\tilde m\geq 2$ respectively are topologically conjugate if and only if $m=\tilde m$.
   3323 \end{thm}
   3324 
   3325 \begin{proof}
   3326     We look at the two cases.
   3327     \begin{itemize}
   3328         \item $m=\tilde m$: The same as seen before.
   3329 
   3330         \item $m\neq \tilde m$: The number of preimages of a point $(m,\tilde m)$ is invariant under conjugaticy.
   3331     \end{itemize}
   3332 \end{proof}
   3333 
   3334 \subsection{Outlook: Coding for other systems}
   3335 
   3336 \begin{defin}
   3337     For arbitrary $X$ and $T:X\to X$ and a partition $S=\{Z_i\}_{i\in I}$, where $I$ is finite/countable. For any $x\in X$ the $S$-itineray ($S$-name) of $x$ is the sequence
   3338     $$\gamma_T(x):=\omega=(\omega_j)_{j\in\N}\in\Omega_I=I^\N,$$
   3339     with $T^jx \in Z_{\omega_j}$ for all $j\in\N$. So $\gamma_T:X\to \Omega_I$.
   3340 \end{defin}
   3341 
   3342 For $y = Tx$, we get $T^jy=T^{j+1}x \in Z_{\omega_{j+1}}$ that means that $\gamma \circ T = \sigma \circ \gamma$.
   3343 
   3344 \paragraph{Warning:}
   3345 In general, $\gamma$ is not injective. For example, $S = \{X\}$ doesn't tell anything. We want that
   3346 $$Z_{(\omega_0,\dots,\omega_{r-1})}:=\bigcap_{j=0}^{r-1}T^{-j}Z_{\omega_j}$$
   3347 shrinks to at most one point.
   3348 
   3349 \begin{rem}
   3350     Sometimes, it is useful to use partitions with overlaps.
   3351 \end{rem}
   3352 
   3353 \begin{rem}
   3354     In general, even for nicest $S$, $\gamma_T$ need not be surjective. And $\gamma_T(X) \subseteq \Omega_I$ can be very complicated.
   3355 \end{rem}
   3356 
   3357 \begin{exam}
   3358     Let $T:\T\to\T$ a surjective piecewise not continuous linear map. We assume that the graph doesn't contain $[0.5,0.5]^2$. Then $\gamma_T(X)=\{\omega: (\omega_j,\omega_{j+1})\neq(1,1)\quad  \forall j\in\N\}$.
   3359 \end{exam}
   3360 
   3361 It $T$ is invertible, use $2$-sided sequences
   3362 $$\hat\Omega_I:=\{\og=(\og_j)_{j\in\Z}:\og_j\in\Z\}=I^\Z$$
   3363 $$\hat\Omega_I\to\hat\Omega_I, \quad (\omega_j)_{j\in\Z}\mapsto (\og_{j+1})_{j\in\Z}$$
   3364 and we define $\gamma_I(x)=(i_n)_{n\in\Z}$, where $T^nx\in Z_{i_n}$ for all $n\in\Z$.
   3365 
   3366 \begin{exam}
   3367     Baker map on $\T^2$. Let $Z_0$ be the closed left half and $Z_1$ be the closed right half. Then $\{Z_0, Z_1\}$ is almost a partition. Let $T:\T^2\to\T^2$ with
   3368     $$T(s,t) = \left(2s,\half t+\half{1_{Z_1}(s,t)}\right).$$
   3369     Then for all $\og\in\hat\Og_I=\{0,1\}^\Z$ there exists an $x$ such that $\gamma(x)=\og$. If we restrict ourselves to the first component of $T$, we get
   3370     $$S(s) = 2s.$$
   3371     This is an example of a general hyperbolic map with expanding (horizontal) and contracting (vertical) directions.
   3372 \end{exam}
   3373 
   3374 \subsubsection{Outlook: Measurable dynamics (Ergotic theory)}
   3375 
   3376 We have already seen chaotic maps with sensitive dependence transitivity (coexistence of diffent types of orbits). What do most/many orbits do?
   3377 
   3378 \begin{exam}
   3379     Doubling map $T:\T\to\T$, $Tx = 2x$. Let $f:= 1_{Z_1}$ and $X_k \in f\circ T^k:\T\to\{0,1\}$. Then
   3380     $$x \in Z_{[\og_0,\dots,\og_{n-1}]}$$
   3381     if and only if
   3382     $$T^kx\in Z_{\og_k}$$
   3383     for all $k \in\jbr{0,n-1}$. Also $\lambda(Z_{[\omega_0,\dots,\og_{n-1}]})=2^{-n}$. Suppose we split $x$ at random, using $P = \lambda$ (the Lebesgue measure on $(X, B_X)$). Then
   3384     $$P[(X_0,\dots,X_{n-1})=(\og_0,\dots,\omega_{n-1})]=\lambda(Z_{(\og_0,\dots,\og_{n-1})})=2^{-n}$$
   3385     This deterministic dynamival system generates the most random process possible if the initial condition is regarded random.\\
   3386     Recall that the strong law of large numbers implies
   3387     $$\reci n\sum_{k=0}^{n-1}X_k \to \half1$$
   3388     almost sure. This is a special case of the Ergotic theorem (for the doublong map):
   3389     $$\forall f \in L^1(\lambda): \reci n\sum_{k=0}^{n-1}f\circ T^k \to \int fd\lambda \text{ almost everywhere}.$$
   3390 \end{exam}
   3391 
   3392 \begin{thm}
   3393     Let $A \in B_X$ be any measurable set and $f:=1_A$. Then
   3394     $$\reci n\sum_{k=0}^{n-1}1_A\circ T^k \to\lambda(A) \text{ almost everywhere}$$
   3395 \end{thm}
   3396 
   3397 What is behind this?
   3398 \begin{thm}
   3399     Suppose $T$ is measurable on a measurable space $(X,B_X)$ such that for all $A\in B_X$ the exists $\mu(A) \in[0,1]$ such that
   3400     $$\reci n\int_{k=0}^{n-1}1_A\circ T^k \to \mu(A)$$
   3401     then
   3402     \begin{enumerate}
   3403         \item $\mu$ is a measure on $B_X$ and $\mu(X)=1$ (showing the finite additivity is easy, but not the $\sigma$-additivity)
   3404 
   3405         \item $\mu$ is $T$-invariant: $\mu\circ T^{-1}=\mu$, meaning that $\mu(T^{-1}A)=\mu(A)$ for all $A\in B_X$.
   3406     \end{enumerate}
   3407 \end{thm}
   3408 
   3409 \begin{proof}
   3410     We only prove the second property. Since $1_{T^{-1}A}=1_a\circ T$, we get
   3411     $$\underbrace{\reci n\sum_{k=0}^{n-1}1_{T^{-1}A}\circ T^k}_{\to\mu(T^{-1}A)}=\reci n\sum_{k=1}^{n} 1_A\circ T^k= \underbrace{\reci n\sum_{k=0}^{n-1}1_A\circ T^k}_{\to \mu(A)} \underbrace{-\reci n1_A+\reci n1_A\circ T^k}_{\to0}.$$
   3412     So we conclude that $\mu(T^{-1}A) = \mu(A)$.
   3413 \end{proof}
   3414 
   3415 The existence of $T$-invariant probability measure $\mu$ enables quantitative analysis of $T$, for example, the ergodic theorem implies that
   3416 $$\reci n \sum_{k=0}^{n-1}1_A\circ T^k \text{ converges } \lambda-\text{almost everywhere}.$$
   3417 This is in particular useful if $\mu$ is equivalent to $\lambda$ (meaning they have the same $0$-sets)
   3418 
   3419 \begin{exam}
   3420     If $T$ is a uniformly expanding circle map and $C^2$ (or $C^{1+\epsilon}$) then there is a $\mu\circ T^{-1} \cong \lambda$.
   3421 \end{exam}
   3422 
   3423 \begin{exam}
   3424     Rotations preserve $\lambda$.
   3425 \end{exam}
   3426 
   3427 \begin{exam}
   3428     Any Hamiltonian system has an invariant measure $\cong \lambda^d$ (Liouville measure), which is finite on compact subsetes.
   3429 \end{exam}
   3430 
   3431 \begin{exam}
   3432     Math billiards.
   3433 \end{exam}
   3434 
   3435 Sample result: Poincare-Carathedory recurrence theorem.
   3436 
   3437 \begin{thm}
   3438     Let $(X, B_X, \mu)$ be a finite measure space and $T$ be measurable with $\mu = \mu \circ T^{-1}$. Then for all $B \in B_X$ and $\mu$-almost all $x\in B$ will return to $B$ infinitely often. That means that there is an increasing, diverging sequence $(n_k)_k$ such that $T^{n_k}x \in B$.
   3439 \end{thm}
   3440 
   3441 \begin{proof}
   3442     Let's go through the proofs.
   3443     \begin{enumerate}
   3444         \item We say that $A\in B_X$ is a wandering set if no $x\in A$ returns. So $A$ and $T^{-k}A$ are disjoint for all $k\in\Ns$. Applying $T^{-m}$ to both sets, we get that $T^{-m}A$ and $T^{-m-k}A$ and so
   3445         $$A, T^{-1}A, T^{-2}A, \dots$$
   3446         are pairwise disjoint. Then we get,
   3447         $$\infty > \mu(x) \geq \mu\left(\bigcup_{k\in\N}T^{-k}A\right)=\sum_{k\in\N}\mu(T^{-k}A) = \sum_{k\in\N}\mu(A) = \infty\mu(A).$$
   3448         Therefore $\mu(A) = 0$.
   3449 
   3450         \item Take any $B\in B_X$. Let
   3451         $$A:= B\setminus\bigcup_{k\in\Ns}T^{-k}B,$$
   3452         the set of returning points. Then $A$ is wandering. By 1. we know that $\mu(A)=0$. But
   3453         $$\mu(\{x\in B: x \text{ is not infinitely recurrent to } B\}) = \mu\left(\bigcup_{n\in\N}T^{-n}A\right)\leq\sum_{n\in\N}\underbrace{\mu(T^{-n}A)}_{=\mu(A) = 0} = 0.$$
   3454     \end{enumerate}
   3455 \end{proof}
   3456 
   3457 \end{document}