Differential equation

 Lecture 1

What is an Ordinary differential equation?

It is of the form $\phi(x,y,y',y'',\ldots y^{(n)})=0$

We are primarily interested in knowing if the ODE can be solved, what the solution is, and if it has a unique solution. 

The order of a differential equation is the highest derivative of $y$ in the equation and the degree of the DE is the power of the highest derivative of $y$.

First-Order Differential Equation:

The equations of the form $a(x)y'+b(x)y+c(x)=0$. If $c(x)=0$, we get a homogeneous differential equation.

Homogeneous first-order differential equation

These equations are of the form $a(x)y'+b(x)y=0$. We can then have,
$y'=-\dfrac{b(x)}{a(x)}y=-p(x)y$
$\implies \displaystyle{\int}\dfrac{dy}{y}=\displaystyle{\int}-p(x)dx$
$\implies |y|=e^{c-\int p(x)dx}=\kappa e^{P(x)}$ where $P(x)=-\int p(x)dx$

Inhomogeneous first-order differential equation

They are of the form $y'=-p(x)y-q(x)$. The solution we get is,
$y=I(\int q(x)I dx)$ where $I=e^{-\int p(x)dx}$.


Lecture 2

First-order differential equation

1. Case 1: y'=g(x)

This case is trivial using the fundamental theorem calculus, which gives us 

$y=\int\limits_{x_0}^xg(x)dx+c$

 

2. Case 2: Linear equations

A) Homogeneous: $y’=-p(x)y$

For this case we will have $y=e^{-\int p(t)dt}$. (Say $P(x)=-\int\$)

This brings us to the initial value theorem which wants $y(x)$ which satisfies the DE and $y_0=y(x_0)$

It is worth noting that for a $p:I\to X$, there exists a unique function $y=e^{-\int\limits_{x_0}^x}p(t)dt$ such that $y(x_0)=y_0$ and satisfies the DE.

(Proof later)

B) Non Homogeneous: $y’=-p(x)y-q(x)$

In this case we will have $\dfrac{d}{dx}ye^{P(x)}=-q(x)e^{P(x)}$ as it will give us

$-ye^{P(x)}p(x)+y’e^{P(x)}=-q(x)e^{P(x)}$

Now from the above equation, we can say

$ye^{P(x)}=-\int\limits_{x_0}^x q(t)e^{P(t)}dt$

 

3. Case 3: Non linear DE

A) Separable: y’=f(x,y)=g(x)h(y)

Now $\dfrac{dy}{dx}=g(x)h(y)\Rightarrow \int\dfrac{dy}{h(y)}=\int\dfrac{g(x)}{dx}$ and $h(y)$ must be non vanishing.

Theorem: Suppose $I$ is a closed interval and $U$ an open interval and we have $g:I\to \mathbb{R}$ and $h:U\to \mathbb{R}$. Then there exists $J\subseteq I$ such that $x\in J$ and there exists a unique $y(x)$ such that it satisfies the differential equation and $y(x_0)=y_0$.

$G(x)=\int\limits_{x_0}^x g(t)dt$ and $H(y)=\int\limits_{y_0}^y\dfrac{1}{h(t)}dt$ then we will have $y=H^{-1}(G(x))$

Proof: (Later XP)

B) Exact DE

Suppose we have a differential equation of the form

$P(x,y)dx+Q(x,y)dy=0$

We may consider the above equation as $\phi(x,y)=c$ which means $\dfrac{d\phi(x,y)}{dx}=\dfrac{\partial \phi}{\partial x}+\dfrac{\partial \phi}{\partial y}\dfrac{dy}{dx}$

On comparison we can see that $\dfrac{\partial \phi}{\partial x}=P(x,y)$ and $\dfrac{\partial \phi}{\partial y}=Q(x,y)$. 

Lecture 3

A system of linear differential equations

Say $\overline{x}=(x_1,x_2,\ldots x_n)$ and we are given the system of equations $\dot{\overline{x}}=\overline{v}(t,\overline{x})$ where $\overline{v}(t, \overline{x})=(v_1(t,\overline{x}),\ldots v_n(t, \overline{x}))$. This would like the following

$\dfrac{dx_1}{dt}=v_1(t, \overline{x})$

$\dfrac{dx_2}{dt}=v_2(t, \overline{x})$

$\dfrac{dx_3}{dt}=v_3(t, \overline{x})$

$\ldots$

$\dfrac{dx_n}{dt}=v_n(t, \overline{x})$

A solution to these family of equations, say is  $\overline{\phi}$, and it satisfies $\dot{\overline{\phi}}=\overline{v}(t,\overline{\phi})$

Theorem: Any $n$th order differential equation can be converted into a system of 1st order DEs.

$x^{(n)}=F(t,x,x’,x’’\ldots x^{(n-1)})$

Consider $\overline{y}=\overline{v}(t,\overline{y})$ where $\overline{y}=(y_1,y_2\ldots y_n)$. Here we will consider $y_i=\dfrac{dy_{i-1}}{dt}$ where $1\leq i \leq n$ and $y^{(n)}=F(t,y’,y’’,\ldots y^{n-1})$. Now this gives us $\overline{v}(t,\overline{y})=\dot{\overline{y}}=(y_2,\ldots F(t,\overline{y}))$.  

For a differential equation which is of the form $\dot{\overline{x}}=\overline{v}(t,\overline{x})$, we call it Non-autonomous and for types of the form $\dot{\overline{x}}=\overline{y(\overline{x})}$, we call it Autonomous

Geometric interpretation

Let’s focus on the Autonomous equations $\dot{\overline{x}}=\overline{v}(\overline{x})$. The set of all possible $x$s forms the phase space or the state phase, $U\subseteq \mathbb{R}^n$.

Now let’s consider a deterministic, finite dimensionafinite-dimensionalle process. Say the phase space is $M$. Therefore the process is $M\to M$. Say the state $\overline{x}\to g^t(\overline{x})$ under this process. Here the variable $t$ represents the time elapsed to reach that state. Note that

$g^{t+s}(\overline{x})=g^{t}(g^s(\overline{x}))$. It is interesting to note that $\{g^t:M\to M | t\in \mathbb{R} \}$ can be treated as a group. Say $g^{t}(\overline{x})=f(t)$. We will have $f(s+t)=f(s)\oplus f(t)$.

This brings us to the idea of Phase curves $\{M,\{g^t\}\}$ which connects all the different states of $\overline{x}$, forming a curve. (Note that $g^t$ is differentiable).

Say $\overline{v}(\overline{x})=\dfrac{d}{dt}g^t(\overline{x})\vert_{t=0}$. The right-hand side gives us a tangent vector at every state. We get a vector field from the family of differential equations, $\dot{\overline{x}}=\overline{v}(\overline{x})$. Below is a representation of a vector field with the phase curve.



Lecture 4

Consider the phase space $U\subseteq \mathbb{R}^n$. Recall that a vector field is a system of differential equations $\dot{\overline{x}}=\overline{v}(\overline{x})$. Now if the map $\overline{v}$ is continuous/differentiable/continuously differentiable, then the vector field is continuous/differentiable/continuously differentiable.

Aside: The vector field is a section of  tangent bundles

Using the phase field, we can get solutions for $\dot{\overline{x}}=\overline{w}(\overline{x})$.

Have a look at the picture below. (The phase space is in yellow)


For given $U,\{g^t\}$ where $g^t:U\to U$ is a diffeomorphism, we say that the phase velocity is $\dfrac{d}{dt}g^t(x)\vert_{t=0}:=\overline{v}(\overline{x})$

 Let’s have a talk on some stuffs.

Definition: We say a point, an equilibrium in the phase space if $g^t(x)=x$. It simply does not change with time ($t$).

Theorem: There exists one and only one phase curve passing through a point on the phase space. (Proof?)

Extended Phase Space: We call the space $\mathbb{R}\times U$, where $U$ is the phase space. We will get to know soon on its importance.

Integral curves: Here we will consider the product space $\mathbb{R}\times U$ (which we have defines as the extended phases space). The curve $(t,\phi (t))$ where $\phi=g^t$, is called the Integral curve. (They are just normal freaking graphs!)

Now under which conditions on $\overline{v}$ and $t_0,x_0$, can we guarantee a n unique integral curve through $(t_0,x_0)$. This brings us to the following theorem.

Theorem: Before stating the main statement of the theorem, have a look on the 2 equations below

1. $\dot{x}=v(x)$

2. $x=\phi(t)$ where $t\in \mathbb{R}$ and $x\in U$ (Here $U$ is the phase space).

The Statement: For given $v:U\to \mathbb{R}$, and given $t_0\in \mathbb{R}$ and $x_0\in U$, we will have a unique solution $\phi$ to equation 1 which satisfies equation 2.

Also, if solutions $\phi_1$ and $\phi_2$ satisfy 2, then they will coincide in a common neighborhood around $t=t_0$.

We will then have

$t-t_0=\displaystyle{\int\limits_{x_0}^{x=\phi(t)}\dfrac{1}{v(\xi)}d \xi}$

Proof:

Take the function $\psi(x)=t_0+\int\limits_{x_0}^{x}\dfrac{d\xi}{v(\xi)}$. We will then have $\dfrac{d\psi (x)}{dx}\vert_{x=x_0}=\dfrac{1}{\dfrac{1}{v(x_0)}}=v(x_0)\neq 0$.

As $\psi(x)$ is differentiable in a neighborhood of $x_0$ as $\dfrac{1}{v(\xi)}$ is continuous for $v(\psi)\neq 0$, by inverse function theorem, we can find $\phi$ such that $\psi(\phi(t))=t$ and $\phi(\psi(x))=x$. Now we will have

$\dfrac{d\psi (x)}{dx}\vert_{x=\phi(t)}\cdot\dfrac{d\phi (t)}{t}=1\Rightarrow \dfrac{d\phi }{dt}=\dfrac{1}{\dfrac{d\psi}{dx}}=v(x)$ 




Comments