Home Blog Page 6

Instability of solutions to nonlinear systems

0

In this post, we propose two results in the instability of solutions to nonlinear systems. Here, we study ordinary differential equations. Well-know stability theorems are due to Liapunov, based on the linearization of the vector field.

Consider a continuous function $f:\Omega\subset \mathbb{R}^d\to\mathbb{R}^d$, and $x_0\in \Omega$. Let the Cauchy problem\begin{align*}\tag{Eq}\dot{u}(t)=f(u(t)),\quad u(0)=x_0.\end{align*}

The flow of an autonomous system

According to Peano’s theorem, the maximal solution of the differential equation $({\rm Eq})$ exists. We denote by $J_{x_0}$ the interval in which the maximal solution is well defined. We denote \begin{align*}D(f)=\bigcup_{x\in\Omega}\left(J_x\times\{x\}\right).\end{align*} The flow associated to the equation $({\rm Eq})$ is the following application \begin{align*} \Phi: D(f)\to \Omega,\quad (t,x)\mapsto \Phi(t,x)=u(t),\end{align*}where $u$ is the maximal solution associated to the initial condition $u(0)=x$. Also, we denote $\Phi_t(x)=\Phi(t,x)$.

We mention that $J_{\Phi_t(x)}=J_x-t$, and if $t_1+t_2\in J_x$, we have\begin{align*}\Phi_{t_2}\circ \Phi_{t_1} (x)= \Phi_{t_1+t_2}(x).\end{align*}

Assume that $f$ is locally Lipschitz on $\Omega$. Then $D(f)$ is an open set of $\mathbb{R}\times \Omega$. Moreover, the flow $\Phi$ is locally Lipschitz, in particular, it is continuous, on $D(f)$.

 Instability of solutions to nonlinear systems

An equilibrium point of $f:\Omega\to \mathbb{R}^d$ is an element $x_0\in \Omega$ such that $f(x_0)=0$. An immediate consequence is that the constant function equal to $x_0$ satisfies the differential equation $({\rm Eq})$.

The equilibrium point $x_0$ is stable if and only if for any neighborhood $W$ of $x_0$ in $\Omega,$ there exists a neighborhood $U$ of $x_0$ in $\Omega$ such that for all $x\in U,$ $\Phi_t(x)$ is defined for any $t\ge 0,$ i.e. global solution, and takes value in $W$.

The equilibrium $x_0$ is unstable if and only if it is not stable.

The equilibrium $x_0$ is asymptotically stable if and only if it is stable and there exists a neighborhood $W$ of $x_0$ in $\Omega$ such that for any $x\in W,$ $\Phi_t(x)$ is defined for any $t\ge 0$ and $\Phi_t(x)\to x_0$ as $t\to +\infty$.

Direct sum of subspaces examples

0

The philosophy behind the direct sum of subspaces is the decomposition of vector spaces as a sum of disjoint spaces. In fact, this is very important for defining the projections; so restricting the work only on the subspaces instead of working on the enter vector space.

Definition of the direct sum of subspaces

Let $E$ be a vector space and let $F$ and $H$ be subspaces of $E$. We define the sum of the two spaces $F$ and $H$ by \begin{align*}F+H:=\{x+y:x\in F,\;y\in H\}.\end{align*} We say that this sum is direct if $F\cap H=\{0\}$. In this case, we write $F\oplus H$.

On the other hand, we say that $F$ and $H$ supplementary vector spaces in $E$ is the sun $F+H$ is direct and $E=F+H$. We then write $E=F\oplus H$.

An example of supplementary vector spaces

Let $\mathcal{E}$ the vector space of all functions from $[-1,1]$ to $\mathbb{R}$. On this space, we consider three subspaces \begin{align*}F_1&:=\{f\in \mathcal{E}: f\;\text{is constant}\},\cr F_2&:=\{f\in \mathcal{E}: f_{|[-1,0]}=0\},\cr F_3&:=\{f\in \mathcal{E}: f_{|[0,1]}=0\}.\end{align*}Prove the following direct sum \begin{align*}\mathcal{E}=F_1\oplus F_2\oplus F_3.\end{align*}

First of all let us prove that each function in $\mathcal{E}$ can be decomposed as the sum of three functions belonging to $F_1,F_2$ and $F_3$. In fact, let $f\in \mathcal{E}$. Let define the following functions\begin{align*}f_1(t):=f(0),\qquad \forall t\in [-1,1],\cr f_2(t):=\begin{cases} 0,& t\in [-1,0],\cr f(t)-f(0),& t\in (0,1],\end{cases}\cr f_3(t):=\begin{cases} f(t)-f(0),& t\in [-1,0),\cr 0,& t\in [0,1].\end{cases}\end{align*}Denote $g=f_1+f_2+f_3$ and prove that $g=f$. As $f_2(0)=f_3(0)=0,$ then $g(0)=f_(0)=f(0)$. On the other hand, let $t\in [-1,0),$ by definition of $f_2$ and $f_2$ we have $f_2(t)=0$ and $f_3(t)=f(t)-f(0)$. Hence for any $t\in [-1,0),$\begin{align*}g(t)=f(0)+0+(f(t)-f(0))=f(t).\end{align*}Now let $t\in (0,1]$. We know that $f_2(t)=f(t)-f(0)$ and $f_3(t)=0$. hence, for any $t\in (0,1],$\begin{align*}g(t)=f(0)+(f(t)-f(0))+0=f(t).\end{align*}Finally, $g(t)=f(t)$ for any $t\in [-1,1]$, so that $\mathcal{E}=F_1+F_2+F_3$.

Let us now $h_1\in F_1,\;h_2\in F_2$ and $h_3\in F_3$ such that $h_1+h_2+h_3=0$. We shall prove that all these functions are null. In fact, as $h_2(0)=h_3(0)=0$, then for all $t\in [-1,1]$ we have \begin{align*}0=h_1(0)+h_2(0)+h_3(0)=h_1(0).\end{align*}As $h_1$ is the constant function $[-1,0]$ then for any $t\in [-1,1],$ $h_1(t)=h_1(0)=0$. So that\begin{align*}f_2+f_3=0\quad\text{on}\; [-1,0].\end{align*}As $f_2\in F_2,$ then $f_2=$ on $[-1,0]$. This implies that $f_3=0$ on $[-1,0]$. But we know that $f_3=0$ on $[0,1]$. Then $f_3=0$ on $[-1,1]$. Hence $f_2=0$ on $[-1,1]$. this ends the proof.

Open mapping theorem in functional analysis

0

In this article, we give an application of the open mapping theorem in functional analysis. This fundamental theorem in functional analysis plays a key role in the study of evolution equations.

In the sequel, we propose a nice functional analysis worksheet on some deep applications in different domains of mathematical analysis.

An application of the open mapping theorem

Exercise: Let $f$ be an application from $\mathbb{R}^n$ to $\mathbb{R}^n$. Suppose that there exists $C>0$ such that \begin{align*}\forall (x,y)\in \mathbb{R}^n\times \mathbb{R}^n ,\quad \langle f(x)-f(y),x-y\rangle\ge C\|x-y\|^2.\end{align*}We propose to demonstrate that $ f $ is a homeomorphism from $ \mathbb{R}^n $ on $ \mathbb{R}^n $. To do so, consider the differential equation \begin{align*}\tag{Eq} \dot{x}(t)=-f(x(t)).\end{align*}

  • Let $x$ and $y$ two solutions of $(Eq)$ defined on $[0,T]$ with $T>0$. If we put $\varphi=\|x-y\|^2,$ prove that \begin{align*} \varphi(t)\le \varphi(0)e^{-2Ct},\quad \forall t\in [0,T].\end{align*} Deduce that \begin{align*} \|\dot{x}(t)\|\le \|\dot{x}(0)\| e^{-Ct}.\end{align*}
  • Let $x$ be a maximal solution of the Cauchy problem $\dot{x}=-f(x)$ and $x(0)=0$. prove that $x$ is defined on $[0,+\infty)$, global solution, and $x$ admits a finite limit $\ell$ as $t\to +\infty$ such that $f(\ell)=0$.
  • Give a conclusion

Solution: 1) As the application $x-y$ is differentiable, then $\varphi$ is differentiable and \begin{align*}\varphi'(t)&=2\langle x(t)-y(t),x'(t)-y'(t)\rangle\cr & \le -2C\varphi(t).\end{align*} This implies that \begin{align*} \varphi(t)\le \varphi(0)e^{-2Ct}.\end{align*}

On the other hand, let we fix $s\in [0,T[$. It is not difficult to see that the application $x_s:[0,T-s]\to \mathbb{R}^n$ such that $t\mapsto x_s(t)=x(t+s)$ is a solution of the equation $(Eq)$. Then by taking $y=x_s$ in the above estimate, we get \begin{align*}\forall t\in [0,T-s],\quad \|x(t+s)-x(t)\|\le \|x(t)-x(0)\| e^{-C t}.\end{align*} This inequality is true for any $(s,t)$ with $0\le s<T$ and $0\le t\le T-s$. Taking $s\in (0,T-t)$, so \begin{align*}\left\|\frac{x(t+s)-x(t)}{s}\right\|\le \left\|\frac{x(t)-x(0)}{s}\right\| e^{-C t}.\end{align*} Now by letting $s\to 0,$ we obten the result.

2) Let $x:[0,T[\to \mathbb{R}^n$ the maximal solution of the Cauchy problem with initial condition $x(0)=0$. If $T<+\infty$, then $\|x(t)\|$ explodes if $ t $ is close to $T$. This means that there exists $\delta>0$ such that $\|x(t)\|$ is not bounded on $[T-\delta,T[$. On the other hand, for $t\in [0,T]$, \begin{align*} \|x(t)\|&\le \int^t_0 \|\dot{x}(\sigma)\|d\sigma\cr & \le \int^t_0 \|\dot{x}(0)\| e^{-C \sigma}d\sigma\cr & \le \frac{\|\dot{x}(0)\|}{C}.\end{align*} This is a contradiction. Thus $T=+\infty$. Then \begin{align*} \|\dot{x}(t)\|\le \|\dot{x}(0)\| e^{-C t},\qquad\forall t\in [0,+\infty).\end{align*} This implies that the improper integral \begin{align*}\int^{+\infty}_0 \|\dot{x}(t)\|dt\end{align*} converges. Thus $x$ has a limit $\ell$ at $+\infty$. On the other hand, as $\dot{x}(t)\to 0$ as $t\to 0,$ then $f(\ell)=0$.

3) If we replace $ f $ by $ f-y $ which satisfies the same hypotheses, we see that $ f $ is a surjective from $\mathbb{R}^n$ to $\mathbb{R}^n$. According to the first hypothesis on $f,$ we deduce that $f$ is also injective. Then it realizes a bijection from $\mathbb{R}^n$ to $\mathbb{R}^n$ . By using Cauchy-Schwarz inequality we have \begin{align*} \|f(x)-f(y)\|\ge C \|x-y\|,\qquad \forall (x,y)\in \mathbb{R}^n\times \mathbb{R}^n.\end{align*} From this we deduce that \begin{align*} \|f^{-1}(x)-f^{-1}(y)\|\le \frac{1}{C} \|x-y\|,\qquad \forall (x,y)\in \mathbb{R}^n\times \mathbb{R}^n.\end{align*} This implies that $f^{-1}$ is continuous.

Differential calculus in Banach spaces

0

Differential calculus in Banach spaces is a very important part of mathematics. In fact, the treatment of partial differential equations strongly depends on this theory. Think about the heat equation. We mention that such equations use function spaces as a workspace. The latter is a Banach space, such as the Lebesgue space, the space of continuous and bounded functions. Hence the importance of differentiation in Banach spaces.

A Banach space $E$ is a vector space endowed with a norm $\|\cdot\|$ for which Cauchy sequences converge in $E$. On the other hand, we say that $E$ is complete with respect to the norm $\|\cdot\|$.

We shall denote by $\mathcal{L}(E)$ the space of linear continuous applications from $E$ into $E$, bounded operators. If $T\in \mathcal{L}(E) $, there exists $M\ge 1$ such that $$\|Tx\|\le M \|x\|,\quad \forall  x\in E.$$ As we will see in the sequel, the differentiability of an map is a linear continuous applications. Contrary to the one dimension case, where a derivative is just a real number.

Differentiation in the real number set

Let us start by defining the differentiation of functions of one variable. We say that a function $f:D\subset \mathbb{R}\to \mathbb{R}$ is differentiable at $c\in D$ if the following limit\begin{align*}\ell=\lim_{x\to c\,x\neq c}\frac{f(x)-f(c)}{x-c}\end{align*} exists. In this case, we set $\ell:f'(c)$; called the derivative of $f$ at the point $c$.

Also, we can write \begin{align*} f(c+h)=f(c)+f'(c)h+\varepsilon(h),\end{align*} where $ \varepsilon(h)/h \to 0$ as $h\to 0$.

We say that $f$ is differentiable on $D$ if $f$ is differentiable at all points of $D$. In this case, we define the differential function of $f$ by $f’:D\to \mathbb{R}$.

The important thing to remember is that the derivative at a point $ c $ is a real number, that is $f'(c)\in\mathbb{R}$.

Differential calculus in Banach spaces

In this part, we shall work with an infinite-dimensional Banach space $(E,\|\cdot\|)$. Let $f: E\to E$ be an application, not necessarily linear, and $x_0,h\in E$. We say that $f$ is differentiable at $x_0$ if there exists a linear continuous application $L\in \mathbb{L}(E)$ such that \begin{align*}f(x_0+h)=f(x_0)+Lh+\varepsilon(h)\end{align*}such that\begin{align*}\lim_{h\to 0}\frac{ \varepsilon(h)}{\|h\|}=0.\end{align*}

In this case we say that $f $ is differentiable at $x_0$ and $L$ the differential of $f$ at $x_0$, denoted by $L:=Df(x_0)$.

Contrary to $\mathbb{R}$, the differential of $f$ at a point $x_0$ is a linear continuous application from $E$ to $E$; that is $Df(x_0)\in\mathcal{L}(E)$.

Illustrate this definition: differentiability in the matrix spaces 

To that purpose, denote by $E=\mathcal{M}_{m,n}(\mathbb{R})$ the Euclidian spaces of matrices of order $m\times n$ endowed with the inner product\begin{align*}(A,B)\in E\times E\mapsto \langle A,B \rangle:={\rm Tr}(A^T B),\end{align*}where ${\rm Tr}$ is the trace. For a function $f:U\subset E\to \mathbb{R}$ defined on a open set $U$ of $E$ and differentiable in $X\in U,$ we denote by $\nabla f(X)$ the gradient of $f$ in $X$.

General matrices

Fix $A\in E$ and let $f_A$ the application\begin{align*}f_A: E\to \mathbb{R},\quad X\mapsto f_A(X)={\rm Tr}(A^T X).\end{align*}Let us show that $f_A$ is differentiable on $E$ and determine $\nabla f_A(X)$. In fact, we know that ${\rm Tr}(\cdot)$ is linear. Then $f_A$ is a linear application, which is continuous because it takes values in $\mathbb{R}$. Thus $f_A$ is differentiable on $E$ and the differential application of $f_A$ is $Df_A(X)=f_A$ for any $X\in E$. We then have\begin{align*}Df_A(X)H=f_A(X)={\rm Tr}(A^T H)=\langle A,H \rangle.\end{align*}Hence, we obtain\begin{align*}\nabla f_A(X)=A,\qquad \forall X\in E.\end{align*}

Differentiability of the space of invertible matrices 

Assume that $m=n$, then we will work with square matrices of order $n$. Let $A\in E:=\mathcal{M}_n(\mathbb{R})$. Denote by  $U$ the set of invertible matrices of $E$. Let $g:U\subset E\to \mathbb{R}$ defined by\begin{align*}g_A(X)={\rm Tr}(X^{-1} A).\end{align*}Now we prove that $g_A$ is differentiable on all $X\in U$ and calculate $\nabla g_A(X)$. We recall that the application $\varphi(X)=X^{-1}$ is differentiable on all $X\in U$. In addition, the differential $D\varphi(X)H=-X^{-1}HX^{-1}$ for all $X\in U$ and $H\in E$. We can write \begin{align*}g_A=\psi\circ \varphi,\end{align*}where $\varphi: U\to E$ and $\psi:E\to \mathbb{R}$ are such that $\varphi(X)=X^{-1}$ and $\psi(Y)={\rm Tr}(YA)$. The application $psi$ is differentiable on each $Y\in E$ with $D\psi(Y)K={\rm Tr}(KA)$ for $K\in E$. On the other hand, the application $\varphi$ is differentiable on each $X\in U$ and\begin{align*}D\varphi(X)H=-X^{-1}HX^{-1}.\end{align*}Thus $g_A=\psi\circ\varphi$ is differentiable on each $X\in U$ and\begin{align*}Dg_A(X)H&=D\psi(\varphi(X))(D\varphi(X)H)\cr &= {\rm Tr}(-X^{-1}HX^{-1} A)\cr &= -{\rm Tr}(-X^{-1}AX^{-1} H).\end{align*}This implies that \begin{align*}\nabla g_A(X)=-(X^{-1}AX^{-1})^T.\end{align*}

Using a canonical inner product 

Let $a\in\mathbb{R}^n,$ $U$ and $E$ as in the previous question.  We define an application $h_a:U\to \mathbb{R}$ by\begin{align*}h_a(X):=\langle X^{-1}a,a\rangle.\end{align*} Here $\langle\cdot,\cdot\rangle$ is the canonical inner product in $\mathbb{R}^n$. Prove that $h_a$ is differentiable on all $X\in U,$ and determine $\nabla h_a(X)$. In fact, we write $h_a=\xi\circ \varphi$ where \begin{align*}xi: E\to \mathbb{R},\quad \xi(Y)=\langle Ya,a\rangle.\end{align*}Remark that $\xi$ is linear and continuous. Then it is differentiable on $E$ and $D\xi(Y)=\xi$. Thus $h_a$ is differentiable in all $X\in U$ and\begin{align*}Dh_a(X)H&=\langle -X^{-1}HX^{-1}a,a\rangle\cr &= -a^T\left(X^{-1}AX^{-1}\right)a\cr &= -{\rm Tr}(X^{-1}aa^T X^{-1}H).\end{align*}Hence\begin{align*}\nabla h_a(X)=-(X^{-1}aa^T X^{-1})^T.\end{align*}

Heat equation using the Fourier series

0

In this article, we learn how to solve the heat equation using the Fourier series. The heat equation belongs to the class of partial differential equations widely used in physics. It describes the diffusion of heat over a region of space. Of course, there are many approaches to studying this equation. Here we mainly use the simplest approach.

The heat equation model

The Fourier series was introduced by the French mathematician and politician Fourier to solve the heat equation. The latter is modeled as follows: let us consider a metal bar. Knowing, at the initial instant, the temperature at each point of the bar and, at all times; the temperature at both ends; can we determine; at any time and at any point; the temperature of the bar?

This problem has been modeled and the temperature $ u $ which is a function of time $ t $ and the point $ x $ is the solution of a partial differential equation called heat equation.

We can assume that the bar is the segment $ [0, L] $. Denote $ Q =(0, L)\times(0, + \infty) $, and $ \overline{Q} = [0, L] \times [0, + \infty) $. The problem to solve is as follows: find a function $ u $ such that \begin{align*}\tag{H1}u\in \mathcal{C}^0(\overline{Q}),\quad u\in \mathcal{C}^2(Q),\end{align*}\begin{align*}\tag{H2}\frac{\partial u}{\partial t}=\frac{\partial^2 u}{\partial x^2}\quad\text{on}\quad Q,\end{align*} \begin{align*}\tag{H3} u(0,t)=u(L,t)=0,\qquad t\in [0,+\infty),\end{align*}\begin{align*}\tag{H4} u(x,0)=h(x),\quad x\in [0,L],\end{align*} where $h$ is a $C^1$ function on the closed interval $[0,L]$ satisfying $h(0)=h(L)=0,$ this is needed to give more regularity to the solution $u$.

The heat equation using the Fourier series approach

We write the solution of (H2) as $$ u (x, t) = f (x) g (t).$$ Then equation $(H2)$ is equivalent to $f(x)g'(t)=f”(x)g(t)$. Let us now find a solution $u$ correponding to the case $f(x)\neq 0$ pour any $x\in ]0,L[$, and $g(t)\neq 0$ for all $t>0$. We deduce that\begin{align*}\frac{f”(x)}{f(x)}=\frac{g'(t)}{g(t)},\quad \forall x\in (0,L),\;t>0.\end{align*}This is true only if the two sides of this equality are constants. This means that, there exists a constant $\lambda\in \mathbb{R}$ such that for any $xin (0,L)$ and $t>0,$ \begin{align*} f”(x)=\lambda f(x)\quad\text{and}\quad g'(t)=\lambda g(t).\end{align*}

In the case of $\lambda>,$ the solution of the first equation is given by \begin{align*} f(x)=a e^{\sqrt{ \lambda }x}+b e^{-\sqrt{ \lambda }x}.\end{align*} The condition $(H3)$ implies that $a+b=0$ and $ a e^{\sqrt{ \lambda }L}+b e^{-\sqrt{ \lambda }L} =0$ . Hence $a=b=0,$ and then $u(x,t)=0,$ this is a contradiction with the condition $(H4)$. On the other hand, if $\lambda=0,$ we have $f”(x)=0$. Then $f(x)=ax+b$. But, due to $(H3)$, we get $u=0,$ and this is a contacdition. The only case that remains is $\lambda <0$. We can wrire $\lambda=-\xi^2<0$. In this case, \begin{align*}f(x)=a\cos \xi x+b\sin \xi x,\quad g(t)=ce^{-\xi^2 t}.\end{align*}

Now using the condition $(H3)$, we deduce that $a=0$ and $\xi=\frac{n\pi}{L}$ for $n\in\mathbb{Z}$. We then have a family of solutions of the form\begin{align*} u_n(x,t)=b_n \sin\left( \frac{n\pi}{L}x\right) e^{- \frac{n^2\pi^2}{L^2}t }. \end{align*}

As the equation $(H2)$ is linear then any sum of $u_n$ is also a solution. We then introduce \begin{align*} u(x,t)=\sum_{n=1}^{+\infty} b_n \sin\left( \frac{n\pi}{L}x\right) e^{- \frac{n^2\pi^2}{L^2}t } .\end{align*}

Remark: Let us mention the isometry of the Fourier transform on the Hilbert space $L^2$ can help in solving the PDE. In fact, using the Fourier transform can help transform the heat equation as PDE into an ordinary differential equation. Then we can use elementary materials to solve the equation. On the other hand, one can use semigroup theory to solve the heat equation.

Gronwall lemma

0

In this article, we state and prove Gronwall lemma and give some of its applications. In fact, we will use this lemma for the stability of the solution to the differential equations. In particular, the Lyapunov stability of nonlinear systems.

The proof of Gronwall’s lemma

Many proofs of known theorems in mathematics are based on Gronwall’s lemma. In fact, this lemma is used to prove the uniqueness of the solution of differential equations, as well as the stability of equilibrium points.

Lemma: Let $\varphi:[a,b]\to [0,+\infty)$ be a continuous function and assume that there exist two real constants $\alpha,\beta\ge 0$ such that for any $c\in [a,b]$, \begin{align*}\varphi(t)\le \alpha+\beta\left|\int^t_c \varphi(s)ds\right|,\qquad \forall t\in [a,b].\end{align*}Then \begin{align*}\varphi(t)\le \alpha e^{\beta |t-c|},\quad\forall t\in [a,b].\end{align*}

Proof: For simplicity, we suppose that $t\ge c$. We select\begin{align*}F(t)= \alpha e^{\beta |t-c|},\quad\forall t\in [a,b]. \end{align*}Then $F$ is a $C^1$ function and $F'(t)=\beta \varphi(t)\le \beta F(t)$. By multiplying $e^{-\beta t}$ on the both sides of this inequality, we obtain \begin{align*} e^{-\beta t} F'(t)- e^{-\beta t} \beta F(t)\le 0.\end{align*} Hence, \begin{align*} \frac{d}{ds} \left(e^{-s\beta}F(s)\right)\le 0.\end{align*} By integrating between $c$ and $t$, the result then follows.

Another version of the lemma: Let $\varphi,\psi:[a,b]\to [0,+\infty)$ be a continuous function and assume that there exist two real constants $\alpha,\beta\ge 0$ such that for any $c\in [a,b]$, \begin{align*}\varphi(t)\le \alpha+\beta\left|\int^t_c \psi(s)\varphi(s)ds\right|,\qquad \forall t\in [a,b].\end{align*}Then \begin{align*}\varphi(t)\le \alpha \exp\left(\beta\left|\int^t_c \psi(s)ds\right|\right),\quad\forall t\in [a,b].\end{align*}

Applications to the stability of the solution of differential equations

In this paragraph, we give an application of the Gronwall lemma to the stability of nonlinear differential equations. In fact, let $\lambda>0$ be a real number, and consider the following semilinear Cauchy problem \begin{align*}\tag{CP} \begin{cases}\dot{u}(t)=-\lambda u(t)+\arctan(t^2 u(t)^5) \frac{\sin{u(t)}}{1+2 u(t)^2},&t\ge 0,\cr u(0)=\frac{1}{2}.\end{cases}\end{align*} If we select \begin{align*}f(t,x)=\lambda x+\arctan(t^2 x^5)\frac{\sin(x)}{1+2x^2},\quad (t,x)\in \mathbb{R}^+\times \mathbb{R},\end{align*} then our equation has the following standard form \begin{align*}\tag{CP} \begin{cases}\dot{u}(t)=f(t,u(t)),& t\ge 0,\cr u(0)=\frac{1}{2}.\end{cases}\end{align*} Clearly $f$ is a $C^1$ function, then locally Lipschitz function. Thus by the Cauchy-Lipschitz theorem, the Cauchy problem (CP) admits a unique maximal solution $u:[0,T)\to \mathbb{R}$. This solution is global, i.e. defined on all $[0,+\infty)$ because $|f(t,x)|\le \lambda |x|+1$ for any $(t,x)\in\mathbb{R}^+\times \mathbb{R}$.

By computing the derivative $(e^{\lambda t}u(t))’$ and using the expression of $\dot{u}(t)$ in (CP), we can prove that \begin{align*}u(t)=\frac{1}{2}e^{-\lambda t}+\int^t_0 e^{-\lambda (t-s)}\arctan(t^2 u(s)^5) \frac{\sin{u(s)}}{1+2 u(s)^2}.\end{align*} Now using the facts that $|\arctan(y)|\le \frac{\pi}{2},$ $|\sin(y)|\le |y|$ and $\frac{1}{1+y^2}\le 1,$ for any $y\in\mathbb{R}$, we obtain \begin{align*} |u(t)|\le \frac{1}{2}e^{-\lambda t}+\frac{\pi}{2} e^{-\lambda t}\int^t_0 e^{\lambda s}|u(s)|ds,\quad t\ge 0.\end{align*} It follows that \begin{align*} e^{\lambda t}|u(t)|\le \frac{1}{2}+\frac{\pi}{2}\int^t_0 e^{\lambda s}|u(s)|ds,\quad t\ge 0.\end{align*} Now by using the Gronwall lemma we have \begin{align*} e^{\lambda t}|u(t)|\le \frac{1}{2}e^{\frac{\pi}{2}t}.\end{align*} Hence $$ |u(t)|\le \frac{1}{2}e^{-\left(\lambda-\frac{\pi}{2}\right)t},\quad \forall t\ge 0.$$ We deduce that the solution, of the equilibrium point $0$, is exponentially stable if $\lambda>\frac{\pi}{2}$.

Stability analysis of differential equations

0

The stability analysis of solutions of differential equations is one of the most important axes in ODE. Here w gives a concise course on the stability of equilibrium (critical) points of differential equations.

The stability analysis of solutions of differential equations

The existence of solutions is guaranteed by two fundamental theorems, Peano’s theorem which gives on the existence of solutions, and the Cauchy-Lipschitz theorem which gives the existence and the uniqueness of solutions. We also mention that several types of solutions are introduced for nonlinear Cauchy problems. In fact, the local solution; the maximal solution, and the global solution. For stability, the solution must be global.

Critical points of functions

A critical (or an equilibrium) point of a continuous function $F:\mathbb{R}^d\to \mathbb{R}^d$ is an element $x^\ast\in \mathbb{R}^d$ such that $F(x^\ast)=0$.

Let us consider the differential equation \begin{align*}\tag{Eq}\dot{u}(t)=F(u(t)),\quad t\ge t_0,\quad u(t_0)=x_0.\end{align*} As $F$ is continuous, by Peano’s theorem we know that there exists a maximal solution $u:[t_0,T)\to \mathbb{R}^d$.

If $x^\ast\in\mathbb{R}^d$ is a critical point of $F,$ then the function $r(t)=x^\ast$ for any $t\ge 0$ satisfies $\dot{r}(t)=F(r(t))$. The function $t\mapsto r(t)$ is called the reference or the stationary solution of the differential equation.

Having the stationary solution, one wonders if the other solutions to the differential equations are somehow close to this reference solution. This is the starting point of the theory of stability, which we study below.

Stability analysis of solutions to differential equations

The critical point $x^\ast$ is called exponentially stable if there exist constants $\delta,\omega,M\in (0,\infty),$ with $\delta$ small enough, such that if the initial state $u(t_0)=x_0$ is closed to the critical point $x^\ast$, i.e. $\|x_0-x^\ast\| \le \delta,$ then

  • the solution $u(t)$ is well defined for any $t\ge t_0$. i.e. $u$ is a global solution, and
  • $\|u(t)-x^\ast\|\le M e^{-\omega t}\|x_0-x^\ast\|$ for any $t\ge 0$.

The linear case: We assume that $F(x)=Ax,$ for $x\in\mathbb{R}^p,$ where $A$ is a square matrix of order $p$. As $F$ is linear, then $F(0)=0,$ so that $0$ is an equilibrium point for $F$. On the other hand, the solution of the differential equation $\dot{u}(t)=A u(t)$ is given by the following exponential of the matrix $A,$ \begin{align*}u(t)=e^{t A}u(0):=\sum_{n=0}^\infty \frac{t^n}{n!} A^n u(0),\qquad \forall t\ge 0.\end{align*}

If $A$ is a diagonal matrix with complex nombers $\lambda_1,\lambda_2,\cdots,\lambda_p$ at the diagonal, then it is clear that the matrix $e^{t A}$ is diagonal with $e^{\lambda_i}$ for $i=1,cdots,p$ at the diagonal. Thus \begin{align*}\|u(t)\|\le e^{(\sup_{1\le i\le p} {\rm Re}\lambda_i)t}.\end{align*}

In this case, the equilibrium point $0$ is exponentially stable if and only if the spectral bound $s(A):=\sup_{1\le i\le p} {\rm Re}\lambda_i < 0$. More generally, we have the following result

Lyapunov stability theorem: linear differential equation

Let $A$ be a square matrix with spectrum $\sigma(A)$, the set of eigenvalues of $A$. Then there exist constants $\omega >0$ a $M>0$ such that \begin{align*}\|e^{tA}\|\le M e^{-\omega t},\qquad \forall t\ge 0,\end{align*} if and only if the spectral bounded satisfies \begin{align*}s(A)=\sup\{ {\rm Re}\lambda: \lambda\in \sigma(A)\}.\end{align*}

Sketch of the proof: If $A$ is diagonalizable, then there exists an invertible matrix $P$ such that $A=P^{-1}D P,$ where $P$ is a diagonal matrix. Remark that for all $n\in \mathbb{N},$ $A^n=P^{-1}D^n P$. This implies that \begin{align*}e^{tA}=P^{-1} e^{tD} P,\quad \forall t\ge 0.\end{align*}Now, assume that $s(A) < 0. $ Using the previous result, for any $\omega\in (0,-s(A))$, we have \begin{align*}\left\| e^{t A}\right\|\le M e^{-\omega t},\quad\forall t\ge 0,\end{align*} where $M=\|P^{-1}\|\|P\|$. Conversely, by contraposition assume that $s(A)\ge 0,$ then there exists $\lambda\in\sigma(A)$ such that ${\rm Re}\lambda\ge 0$. Let $x\in \mathbb{R}^p$ the eigenvector associated with $\lambda,$ so that $Ax=\lambda x$. Then $A^n=\lambda^n x$. This implies that $e^{t A}x=e^{\lambda t}x$. Thus $\|e^{t A}x\|=e^{{\rm \lambda} t}\|x\|$ dont converge to zero as $t\to +\infty$. This ends the proof.

If $A$ is not diagonalizable, i.e. when $\sigma(A)=\{\lambda_1,\cdots,\lambda_r\}$ with $r < p$ and each $\lambda_i$ has a multiplicity $n_i$ such that $n_1+\cdots+n_r=p$. In this case, there exists a square invertible matrix $S$ such $A=S^{-1}JS$, where $J$ is a matrix of Jordan, $J$ is diagonalizable by block $J=diag(J_{\lambda_1},\cdots,J_{\lambda_r})$ with each block $J_{\lambda_i}$ is an upper triangular matrix with $\lambda_i$ repeated in the diagonal. We also have \begin{align*}e^{tA}=S^{-1} diag\left(e^{t J_{\lambda_1}},\cdots,e^{t J_{\lambda_r}}\right)S.\end{align*}

Proof of peano existence theorem

0

We propose a nice proof of Peano existence theorem. This theorem shows that the continuity of the vector field suffices for the existence of solutions to the ODE; ordinary nonlinear differential equations. We notice that this theorem does not guarantee the uniqueness of the solution.

Local solutions

Let $I$ be an interval of $\mathbb{R}$ and $\Omega$ an open set of $\mathbb{R}^d$ with $d\in\mathbb{N}^\ast$. Let $f:I\times \Omega\to \mathbb{R}^d$ be a continuous function. Let $(t_0,x_0)\in I\times\Omega$ and consider the Cauchy problem \begin{align*}(CP)\quad\begin{cases} \dot{u}(t)=f(t,u(t)),& t\in I,\cr u(t_0)=x_0.\end{cases}\end{align*}

A local solution to the Cauchy problem $(CP)$ is a couple $(J,\varphi)$, where $J$ is a subinterval of $I$ containing $t_0$, $\varphi:J\to \Omega$ is a function of class $C^1$ with $u(t_0)=x_0$ and satisfying the differential equation in $(CP)$.

The proof of Peano existence theorem

Peano’s theorem says that if $f$ is continuous then the Cauchy problem $(CP)$ admits a least a local solution $(J,u)$.

The idea of the proof of this theorem: Let $r>0$ such that the closed ball $\overline{B}(x_0,r)\subset \Omega$. By reducing $I$ to a compact, we can assume that $f$ is bounded by a constant $M>0$ on $I\times \overline{B}(x_0,r) $. We set \begin{align*}J:=I\cap \left[t_0-\frac{r}{M}, t_0+\frac{r}{M}\right].\end{align*}

Let $(f_k)_k$ be a sequence of functions of class $C^1$ from $\mathbb{R}\times \mathbb{R}^d$ to $\mathbb{R}^d$ be converging to $f$ on $J\times \overline{B}(x_0,r)$, bounded by $M$. Each $f_k$ is locally Lipschitz, we can then use a version of Cauchy-Lipschitz theorem which gives the existence of a solution $u_k:J\to \overline{B}(x_0,r)$ such that\begin{align*} \forall t\in J,\quad \dot{u}_k(t)=f_k(t,u_k(t)),\quad u_k(t_0)=x_0.\end{align*}

By taking the integral between $t_0$ and $t$, we obtain \begin{align*}\forall t\in J,\quad u_k(t)=x_0+\int^t_{t_0} f_k(s,u_k(s))ds.\end{align*}On the other hand, a simple argument shows that the sequence of function $(u_k)_k$ is equicontinuous and equi-bounded. Then by using Ascoli theorem, there exists a subsequence $(u_{n_k})_k$ of $(u_k)_k$ that converge on $J$ to an application $u$ continue on $J$ into $ \overline{B}(x_0,r) $. This is the local solution to the Cauchy problem $(CP)$.

The uniqueness of the solution is not assured by Peano’s theorem

The continuity of $f$ in the proof of the Peano theorem is not sufficient for the uniqueness of the solution.

The following example strengthens this remark: the Cauchy problem \begin{align*} x'(t)=\sqrt{x(t)},\quad x(0)=0.\end{align*}The problem admits two solution; the null solution and the following solution \begin{align*}t\mapsto \begin{cases}\frac{t^2}{4},& t\ge 0,\cr -\frac{t^2}{4},& t\le 0.\end{cases}\end{align*}

Even and odd numbers

0

Welcome to the fascinating world of mathematics, where even and odd numbers reign supreme! These two categories of numbers are the foundation upon which countless mathematical concepts and theorems are built. Whether you’re a mathematician or simply someone who wants to deepen their understanding of the mathematical universe, understanding even and odd numbers is essential.

In this article, we’ll take a deep dive into the mesmerizing world of even and odd numbers, uncovering their unique properties, exploring their practical applications, and discovering the hidden beauty within their numerical patterns. Get ready to be amazed by the wonders of mathematics

Defining Even and Odd Numbers

First, let’s get clear on the two unique groups of integers we’re dealing with here:

Even numbers

Let’s talk about even numbers! These are the integers that can be divided by 2 without leaving any leftovers. So, if you can express a number “$n$” as $2$ multiplied by another integer “$k$”, then it’s considered even: $n=2k$. Easy peasy, right? Some examples of even numbers are $-2, 0, 4, 8,$ and even the ultimate answer to life, the universe, and everything $- 42$!

Odd Numbers

Did you know that odd numbers are integers that just can’t be divided by $2$ evenly? It’s true! If you have an integer “$n$” and it can be expressed as $n=2k+1$ with “k” being an integer, then it’s considered odd. Some examples of odd numbers include $-3, 1, 7, 15 $, and $33$. So next time you come across a number that can’t be divided by $2$ evenly, you’ll know it’s an odd one!

Properties of Even and Odd Numbers

Even numbers possess distinctive characteristics when subjected to arithmetic operations. The addition or subtraction of two even numbers invariably results in an even number. Conversely, the addition or subtraction of two odd numbers yields an even number. However, the multiplication of even numbers always yields an even product.

The arithmetic operations of addition, subtraction, and multiplication applied to a pair of even numbers result in an even number.
Let $a$ and $b$ be even integers. It follows that there exist two integers $p$ and $q$ such that $a=2p$ and $b=2q$. Consequently, we have that $$ a+b=2(p+q),\quad a-b=2(p-q),\quad ab=2(2pq).$$ Hence, $a+b$, $a-b$, and $ab$ are all even integers.
An odd number is obtained when an even number is added to or subtracted from another odd number.
The proof is concluded by demonstrating that the following equations hold true for any integers $k$ and $k’$: $$ 2k+(2k’+1)=2(k+k’)+1,\quad 2k-(2k’+1)=2(k-k’)+1.$$
When an even number is multiplied by an odd number, the resulting product is an even number.
In fact, this follows for the equation $$ (2k)(2k’+1)=2(2kk’+k’).$$

Worksheets on even and odd numbers

The objective of this inquiry is to ascertain whether the value of $c$, which is given by the expression $(2n+1)^2+2n-1$, is an even or odd number.
We can write \begin{align*}c&=(2n+1)^2+2n-1\cr &= 4n^2+4n+1+2n-1\cr &= 4n^2+6n\cr &= 2(2n^2+3n).\end{align*} This shows that $c$ is an even number.
From the following list of numbers, determine which are even: $$ 5^2,\quad \frac{378}{3},\quad 15^2-8.$$
We have
  1. $5^2=5\times 5=25$, it is an odd number.
  2. $\frac{378}{3}=126=2\times 63,$ it is an even number.
  3. $15^2-8=225-8=217=2\times 108+1$, it is an odd number.
Demonstrate that the square of an even integer is also an even integer.
Assuming $N$ to be an even integer, it can be expressed as $N=2n$, where $n$ is a natural number. By squaring both sides of the aforementioned equation, we obtain $N^2=(2n)^2=4n^2=2(2n^2)$. Consequently, $N^2$ is an even integer
Demonstrate that the product of two consecutive integers is an even number.

Let $k$ be an integer and consider the selection of $N=k(k+1)$, which is the product of two consecutive integers. There are two possible cases for $k$: when $k$ is even or odd.

First, let us assume that $k$ is even. In this case, we can express $k$ as $k=2n$, where $n$ is an integer. Substituting this value of $k$ into the expression for $N$, we obtain $N=2n(2n+1)=2p$, where $p=2n^2+n$ is an integer. Therefore, $N$ is an even number.

Second, let us assume that $k$ is odd. In this case, we can express $k$ as $k=2n+1$, where $n$ is an integer. Substituting this value of $k$ into the expression for $N$, we obtain $N=(2n+1)(2n+1+1)=(2n+1)(2n+2)$. This can be rewritten as $N=2q$, where $q=(2n+1)(n+1)$ is an integer. Therefore, $N$ is also an even number.

Fourier Transform properties and applications

0

In this post, we study Fourier transform properties and give some applications. In fact, this transformation helps in converting partial differential equations to ODE. A simple method to solve the heat equation is the Fourier transform technique.

The Riemann-Lebesgue Lemma

We denote by $L^1(\mathbb{R})$ the Lebesgue space of measurable functions $f:\mathbb{R}\to \mathbb{R} $ such that\begin{align*} \|f\|_1^2:=\int_{\mathbb{R}} |f(x)|dx<\infty.\end{align*}

Also, we recall that $\|\cdot\|_1$ is a norm on $ L^1(\mathbb{R})$ and that $( L^1(\mathbb{R}), \|\cdot\|_1 ) $ is a Banach space, no reflexive. On the other hand, the space of continuously differentiable functions with compact support $\mathcal{C}_c^1(\mathbb{R})$ is dense in the space $ L^1(\mathbb{R}) $.

As for any $t,x\in\mathbb{R}$, we have $|e^{-itx}f(x)|=|f(x)|$, then for $f\in L^1(\mathbb{R})$, the function $(x\mapsto e^{-itx}f(x) )\in L^1(\mathbb{R}) $. We then have the following definition:

Definition: For $f\in L^1(\mathbb{R}) $ we define the Fourier transform of $f$ by \begin{align*}\hat{f}(t)=\int_{\mathbb{R}} e^{-itx}f(x) dx,\qquad t\in \mathbb{R}.\end{align*}

Clearly $f\mapsto\hat{f}$ is a linear application on $ L^1(\mathbb{R})$. We have the following Riemann-Lebesgue result.

Theorem: The Fourier transform is an application from $ L^1(\mathbb{R})$ to $\mathcal{C}_0(\mathbb{R})$, where \begin{align*} \mathcal{C}_0(\mathbb{R}) :=\{f\in \mathcal{C}(\mathbb{R}) : \lim_{|t|\to +\infty}f(t)=0\}.\end{align*}

Proof: Here we use the density of $\mathcal{C}_c^1(\mathbb{R})$ is dense in the space $ L^1(\mathbb{R}) $. Hence, in the first step, we will prove the result for $f\in \mathcal{C}_c^1(\mathbb{R}) $. In this case, and by using integration by part, we get for any $t\neq 0,$ \begin{align*}\hat{f}(t)=\frac{1}{it}\int_{\mathbb{R}}e^{-itx}f'(x)dx.\end{align*}This implies that\begin{align*} | \hat{f}(t) |\le \frac{1}{|t|}|f’|_1.\end{align*}

If $f\in L^1(\mathbb{R}) ,$ by density there exists a sequence $(f_n)_n\subset \mathcal{C}_c^1(\mathbb{R}) $ that converges to $f$ in $ L^1(\mathbb{R}) $. Thus, for any $\varepsilon>0,$ there exists $m\in\mathbb{N}$ such that $\|f_m-f\|_1\le \frac{\varepsilon}{2}$. Moreover, remark that\begin{align*} \hat{f}(t) = \hat{f_m}(t) + \hat{f}(t) – \hat{f_m}(t),\end{align*} so that \begin{align*} | \hat{f}(t) |&\le \frac{1}{|t|}\|f_m’\|_1+\|f-f_m\|_1\cr & \le \frac{1}{|t|}\|f_m’\|_1 + \frac{\varepsilon}{2}. \end{align*} This ends the proof.

Fourier Transform in the Schwartz space

We define the Schwartz space by\begin{align*}\mathcal{S}(\mathbb{R})=\{f\in \mathcal{C}^\infty(\mathbb{R}):\forall p,q\in\mathbb{N}&,\;\exists C_{p,q}>0:\cr & |x^pf^{(q)}(x)|\le C_{pq},\; \forall x\in\mathbb{R}\}.\end{align*} The first remark is that $ \mathcal{S}(\mathbb{R}) \subset L^1(\mathbb{R})$ continuously and densely. Hence we can prove all properties of the Fourier transform in $\mathcal{S}(\mathbb{R})$. These properties are translated to $L^1(\mathbb{R})$ by density.

As example we have $(x\mapsto e^{-x^2})\in \mathcal{S}(\mathbb{R}) $. On the other hand, we can easily prove that if $f\in \mathcal{S}(\mathbb{R}) ,$ then also $f’$ and $x\mapsto xf(x)$ are in $ \mathcal{S}(\mathbb{R}) $.

Theorem: If $f\in \mathcal{S}(\mathbb{R}) $ then the Fourier transform $\hat{f}\in \mathcal{S}(\mathbb{R}) $.

Fear of mathematics, solution

0

This article discusses the fear of mathematics. The latter alone crystallizes many of the nervous breakdowns and tears in students during their journey in public and private education. Mathematics is involved in all school programs. From elementary school to engineering school: one day, you may find yourself blocked by an exercise and lose your nerves.

Math allows students to understand the world around them and which above all determines many professional careers! Success in math is a guarantee of winning a good degree. The fear of mathematics is therefore due to its difficulties and to the fact that it is thanks to it that we select brilliant students.

What is the most terrifying thing in mathematics?

It is one of the only disciplines taught in school that can cause such blockage in children, preventing them from speaking in public or giving them apprehension before each lesson. Why? Because in mathematics, we are TRUE or FALSE. In contrast, in each class, only three to five students feel good about math. As in athletics, after the first round, all players go into two parts, a very small set in front and another large set behind.

In order to solve the fear of math problems, it is more convenient to focus on the origin of the blockage:

  • Is it the idea of ​​doing math that keeps you from being aware?
  • Is it the fear of failing in general?
  • You do not know how to succeed in a problem and it bothers you?
  • Is it a specific dimension of mathematics that is causing you problems?

The fear of mathematics is often caused by a lack of self-esteem or a negative thought: what will my teacher say if I am wrong? Why can’t I fix this simple problem?

In addition, there is real social pressure regarding mathematics, which is considered “the future” rather than a literary subject for example. Failure to master them would mean failure. We also use mathematics on a daily basis, whether at the supermarket or the restaurant, and not knowing how to use its multiplication tables can quickly make us laugh. At least that’s how a person can feel.

Thus, the nature of your math block will determine how you can manage it, and thus you can overcome your fear without a problem. Choose the statement that seems most specific to describe your situation:

  • You find it hard to calm down when your math teacher asks you to solve an equation,
  • You don’t like solving a math problem on the blackboard,
  • You hate learning math or doing exercises,
  • You don’t know how to handle your math difficulties,
  • Sometimes it is enough to be aware of the blockage to resolve it.

How do avoid blockages in mathematics?

  • Take private math lessons with a private math teacher. This very specific type of teaching mathematics makes it possible to unlock students’ problems.
  • Use concrete supports: being afraid of mathematics can be solved thanks to an adapted pedagogy and concrete supports.
  • Look for the psychological blockage source: in some cases, the blockage is caused by family problems. The math teacher can then become a psychologist.
  • Go back to the basics of math. In particular, classical formulas and mathematical identities.

All students, from elementary to high school know how important mathematics is; not only in the school curriculum but also in professional and daily life. Hence the need to be attentive in mathematics and to listen to your math teacher. Realizing this importance, mathematics causes panic, anxiety, anxiety, and panic attack … In short, this is called “fear of mathematics”. To the point of even talking about psychological blockages, anxiety disorders, panic attacks, and anxiety attacks. Here are some tips to avoid fear in math lessons:

Tips for becoming a good math teacher

0

You love math and want to be a teacher? here we propose some tips for becoming a good math teacher. You need also to know how to motivate your students in mathematics. This question deserves a detailed answer because many students fear mathematics. Becoming a mathematics teacher attracts more and more people. But to teach this type of program takes a lot of skills and patience.

Here comes tips for becoming a good math teacher

Teaching math and getting your students interested throughout the cycle requires finding various ploys. Above all, becoming a good math teacher requires a solid background in basic and advanced math

Motivate students to love math

As a math teacher, it is not always easy to get students interested in a math problem. In general, being able to capture students’ attention by teaching them math is certainly one of the main concerns of a teacher. However, whether you are an associate professor or a certified professor, or you are registered on a platform to give online courses, you often want to offer your students effective and lasting learning.Faced with the blocking that some may have in mathematics, this task is far from easy.

Many students believe that mathematics is just an abstraction of studying algebra and geometry and is similar to a mind sport or an intelligence test. The challenge of a good math teacher is therefore to show his students that it is prejudiced and that one can use mathematics on a daily basis, and every day.

Helping students understand mathematics, acquiring the vocabulary necessary for success in mathematics, and mastering the different theorems, requires passion but also a certain methodology.

How to be a good math teacher?

Without a doubt, in every school or university, there are good teachers and bad teachers. It is understood that there are teachers who have a very high level of mathematics but have problems transmitting their know-how. So a bad teacher is one who cannot communicate well with these students.

A good math teacher is, above all, in our view, someone who can transmit the taste – or even the passion – of his discipline to his students. In theory, a good teacher is one who manages to:

  • Enjoy teaching.
  • Make you want to work and create a climate conducive to work.
  • Use diagrams and designs to simplify complex math problems
  • Encourage your audience
  • Highlight the qualities of his student

Speaking of good math teachers, are there bad students?

This question depends on certain situations. From primary to high school, there are normally no bad students in math. There are only students who have not yet managed to understand mathematics! Indeed, at this level, we only rely on elementary mathematics, or on a basis, sometimes without proof. In most cases, the student can understand mathematics using diagrams and designs. For example, the initial definition of an integral of a continuous function is only a surface, the teacher avoids talking about the constriction of the integral.

We do not use epsilon to give the definition of a sequence limit. In secondary to prove that the function $ t \mapsto \sin(t) $ has no limit to infinity, as $t$ goes to $+ \infty$, we only communicate to the student that the sine function is $ 2 \pi $ periodic and has a graph between -1 and 1 and repeated in the same way on each interval of $ 2 pi $. As a math teacher, your main mission will be to introduce your student to your mathematical universe! Whatever his course, primary, middle school, or high school, you must support him in his success by showing seriousness and pedagogy.

Now, after high school, a student can also study maths at university. In this case, we select two types of students. Students only use mathematics to study other sciences such as physics, and students who wish to have a full mathematics program. For the latter, we can speak of bad students. Because the program somehow requires an ability to absorb math. To have a career in mathematics, you need a minimum of intuition.