12  Hyperbolic Partial Differential Equations

12.1 The One-dimensional Wave Equation

We will now begin to study the second major class of PDEs, \(\,\)hyperbolic equations

  • Vibrating-String Problem

    We consider the small vibrations of a string that is fastened at each end

  • To mathematically describe the vibrations of this string, \(\,\)we consider all the forces acting on a small section \(\Delta x\) of the string

    If the horizontal component of tension is constant \(T\), \(\,\)then the tension acting on each side of the string segment is given by

    \[ \begin{aligned} T_1 \cos\alpha &\approx T \\ T_2 \cos\beta &\approx T \end{aligned}\]

  • In the vertical component of Newton’s second law, \(\,\)the mass of this piece \(\,\rho\Delta x\,\) times its acceleration \(\,u_{tt}\) will be equal to the net force on the piece:

    \[\begin{aligned} \rho\Delta x u_{tt} &= T_2 \sin\beta +T_1 \sin\alpha \\ &\Downarrow \\ \frac{\rho\Delta x}{T} u_{tt} &= \frac{T_2 \sin\beta}{T_2 \cos\beta} +\frac{T_1 \sin\alpha}{T_1 \cos\alpha} = \tan\beta +\tan\alpha \\ &= u_x(x+\Delta x) -u_x(x) \\ &\Downarrow \\ u_{tt} &=\frac{T}{\rho} \frac{u_x(x+\Delta x) -u_x(x)}{\Delta x} \\ &\Downarrow\; c^2 = T/\rho,\;\Delta x \to 0 \\ u_{tt} &= c^2 u_{xx} \end{aligned}\]

    This is the wave equation for \(\,u(x,t)\,\) and \(\,c\) is the speed of propagation of the wave in the string

12.2 \(\,\)D’Alembert Solution of the Wave Equation

  • If the student recalls the parabolic case, \(\,\)we started solving diffusion problems when the space variable was bounded (by separation of variables), \(\,\)and then went on to solve the unbounded case (where \(-\infty < x <\infty\)) by the Fourier transform

  • In the hyperbolic case (wave equation), \(\,\)we will do the opposite. \(\,\)We start by solving the one-dimensional wave equation in free space:

\[ \begin{aligned} u_{tt} &=c^2 u_{xx} && -\infty < x < \infty,\; 0 < t < \infty \\ \begin{array}{r} u(x, 0) \\ u_t(x, 0) \end{array} & \begin{array}{c} = f(x) \\ = g(x) \end{array} && -\infty < x <\infty \end{aligned} \tag{DA}\label{eq:DA}\]

  • We could solve this problem by using the Fourier transform (transforming \(x\)) or the Laplace transform (transforming \(t\)), \(\,\)but we will introduce yet a new technique (canonical coordinate), \(\,\)which will introduce the reader to several new and exciting ideas

STEP 1 \(\,\)Replacing \(\,(x,t)\) by new canonical coordinates \((\xi,\eta)\)

\[\begin{aligned} u_{tt}&=c^2 u_{xx} \\ &\Downarrow\;\color{red}{\xi=x+ct,\;\eta=x-ct} \\ u_t&= u_\xi \xi_t+u_\eta \eta_t =c(u_\xi -u_\eta) \\ u_{tt}&= c(u_{\xi\xi} -u_{\eta\xi})\xi_t +c(u_{\xi\eta} -u_{\eta\eta})\eta_t\\ &=c^2(u_{\xi\xi}-2u_{\xi\eta}+u_{\eta\eta})\\ u_x&=u_\xi \xi_x+u_\eta \eta_x=u_\xi+u_\eta \\ u_{xx}&= u_{\xi\xi}\xi_x+u_{\eta\xi}\xi_x+u_{\xi\eta}\eta_x +u_{\eta\eta}\eta_x\\ &=u_{\xi\xi}+2u_{\xi\eta}+u_{\eta\eta}\\ &\Downarrow \\ u_{\xi\eta}&=0 \end{aligned}\]

STEP 2 \(\,\)Solving the Transformed Equations

\[\begin{aligned} u_{\xi\eta}&= 0 \\ &\Downarrow \\ \text{Integration } &\text{with respect to }\xi \\ &\Downarrow \\ u_{\eta}(\xi,\eta)&=\varphi(\eta) \\ &\Downarrow \\ \text{Integration } &\text{with respect to }\eta \\ &\Downarrow \;\;\phi=\int\varphi\,d\eta \\ u(\xi,\eta)=\phi&(\eta) +\psi(\xi) \\ \end{aligned}\]

STEP 3 \(\,\)Transforming back to the Original Coordinates \(\,x\,\) and \(\,t\)

\[\begin{aligned} u(\xi,\eta)&=\phi(\eta) +\psi(\xi) \\ &\Downarrow\; \xi=x+ct, \;\eta=x-ct \\ u(x,t)=\phi&(x-ct) +\psi(x+ct) \end{aligned}\]

NOTE \(\,\)This is the general solution of the wave equation, \(\,\)and it is interesting in that it physically represents the sum of any two moving waves, \(\,\)each moving in opposite directions with velocity \(\,c\)

STEP 4 \(\,\)Substituting the General Solution into the ICs

\[\begin{aligned} u(x,t)=\phi(x-&ct) +\psi(x+ct) \\ &\Downarrow \;\scriptsize u(x,0)=f(x), \;u_t(x,0)=g(x) \\ \scriptsize\phi(x) +\psi(x) \; & \scriptsize = f(x)\\ \scriptsize-c\phi_x(x) +c\psi_x(x) \; &\scriptsize = g(x) \\ \big\Downarrow \;&{\scriptstyle \text{integrating the 2}^{nd} \text{equation} \text{ from } x_0 \text{ to } x} \\ \scriptsize\phi(x) +\psi(x) = &\scriptsize \, f(x)\\ \scriptsize-c\phi(x) +c\psi(x) = &\scriptsize \, \int_{x_0}^x g(\alpha)\,d\alpha +C \\ &\Downarrow \\ \scriptsize\phi(x)=\frac{1}{2}f(x) -\,& \scriptsize\frac{1}{2c}\int_{x_0}^x g(\alpha)\,d\alpha -\frac{C}{2c} \\ \scriptsize\psi(x)=\frac{1}{2}f(x) +\,& \scriptsize\frac{1}{2c}\int_{x_0}^x g(\alpha)\,d\alpha +\frac{C}{2c} \\ &\Downarrow \\ \end{aligned}\]

\[\color{red}{\begin{aligned} u(x,t) =\frac{1}{2} \left[ f(x -ct)\right. &+ \left. f(x +ct) \right] +\frac{1}{2c} \int_{x-ct}^{x+ct} g(\alpha)\,d\alpha \end{aligned}}\tag{DAS}\label{eq:DAS}\]

This is what we were aiming for, \(\,\)and it is called the D’Alembert solution to \(\eqref{eq:DA}\)

  • Motion of a Simple Square Wave

    \[ \begin{aligned} u(x,0)&= \begin{cases} 1 & \;-1 < x < 1 \\ 0 & \text{everywhere else} \end{cases} \\ u_t(x,0)&=0 \end{aligned}\]

  • Initial Velocity Given

    Suppose now the initial position of the string is at equilibrium and we impose an initial velocity (as in a piano string) of \(\,\sin x\)

    \[\begin{aligned} u(x,0)&= 0\\ u_t(x,0)&=\sin x \end{aligned}\]

    Here, \(\,\)the solution would be

    \[\begin{aligned} u(x,t) &= \frac{1}{2c} \int_{x-ct}^{x+ct} \sin\xi \, d\xi\\ &=\frac{1}{2c} \left[ \cos(x -ct) -\cos(x +ct)\right] \\ &=\frac{1}{c} \sin x \cdot \sin ct \end{aligned}\]

\(~\)

import numpy as np
import matplotlib.pyplot as plt
from matplotlib import animation, rc
from IPython.display import HTML

plt.rcParams['font.size'] = 12
plt.rcParams['xtick.direction'] = 'in'
plt.rcParams['ytick.direction'] = 'in'
fig = plt.figure(figsize=(6, 4))
ax = plt.axes(xlim=(-6.0*np.pi, 6.0*np.pi), ylim=(-1.5, 1.5))

ax.set_xticks([-6.0*np.pi,-3.0*np.pi, 0, 3.0*np.pi, 6.0*np.pi])
ax.set_xticklabels([r'$-6\pi$',r'$-3\pi$','$0$',r'$3\pi$',r'$6\pi$'])
ax.set_yticks([-1.2, -0.6, 0, 0.6, 1.2])
ax.set_xlabel('$x$')
ax.set_ylabel('$u(x,t)$')

plt.close()
time_text = ax.text(-1.5, 1.3, '')
line, = ax.plot([], [], lw=2)
def init():
    time_text.set_text('t = 0.0')
    line.set_data([], [])
    return (line,)

c = 1
def animate(t):
    time_text.set_text(f't = {t:3.1f}')
    x = np.linspace(-6.0*np.pi, 6.0*np.pi, 300)
    u = 1.0/(2.0*c)*(np.cos(x -c*t) -np.cos(x +c*t))
    line.set_data(x, u)
    return (line,)

tt = list(np.linspace(0, 2.0*np.pi/c, 100))
anim = animation.FuncAnimation(fig, animate, 
        init_func=init, frames=tt, interval=200, blit=True) 
HTML('<center>' + anim.to_html5_video() + '</center>')
Initial Velocity Given: \(\;u_t(x,0)=\sin t\)

\(~\)

12.3 \(\,\)More on the D’Alembert Solution

  • We proved that in the last section the solution of the pure initial-value problem

    \[ \begin{aligned} u_{tt} &=c^2 u_{xx} && -\infty < x < \infty,\; 0 < t < \infty \\ \begin{array}{r} u(x, 0) \\ u_t(x, 0) \end{array} & \begin{array}{c} = f(x) \\ = g(x) \end{array} && -\infty < x <\infty \end{aligned}\]

    is given by

    \[ u(x,t) =\frac{1}{2} \left[\, f(x -ct) +f(x +ct) \right] +\frac{1}{2c} \int_{x-ct}^{x+ct} g(\xi)\,d\xi \]

    We now present an interpretation of this solution in the \(\,xt\)-plane at the two specific cases

CASE 1 \(\,\) Initial Position Given; \(\,\) Initial Velocity Zero

  • Let’s consider the following initial condition

    \[\begin{aligned} u(x,0) &=f(x) \\ u_t(x,0) &=0 \end{aligned},\quad -\infty < x < \infty\]

    the D’Alembert solution is

    \[ u(x,t) =\frac{1}{2} \left[ f(x -ct) +f(x +ct) \right] \]

    and the solution \(u(x_0,t_0)\) can be interpreted as being the average of the initial displacement \(\,f(x)\) at the points \((x_0-ct_0,0)\) and \((x_0 +ct_0,0)\) found by backtracking along the lines

    \[\begin{aligned} x -ct&= x_0 -ct_0\\ x +ct&= x_0 +ct_0 \end{aligned}\quad\color{red}{\text{characteristic curves}}\]

  • For example, \(\,\)using this interpretation, \(\,\)the initial condition

    \[\begin{aligned} u(x,0) &= \begin{cases} 1 & \;-1 < x < 1 \\ 0 & \text{everywhere else} \end{cases} \\ u_t(x,0) &=0 \end{aligned}\]

    would give us the solution in the \(\,xt\)-plane shown in figure

CASE 2 \(\,\)Initial Displacement Zero; \(\,\) Velocity Arbitrary

  • Consider now the IC

    \[\begin{aligned} u(x,0) &=0 \\ u_t(x,0) &=g(x) \end{aligned}, \quad -\infty < x < \infty\]

    the solution is

    \[ u(x,t) =\frac{1}{2c} \int_{x -ct}^{x +ct} g(\xi)\,d\xi \]

    and, \(\,\)hence, \(\,\)the solution \(\,u(x_0, t_0)\,\) can be interpreted as integrating the initial velocity between \(x_0 -ct_0\) and \(x_0 +ct_0\) on the initial line \(t=0\)

  • Again, \(\,\)using this interpretation, \(\,\)the solution to the initial-value problem

    \[\begin{aligned} u(x,0) &= 0 \quad -\infty<x<\infty \\ u_t(x,0) &= \begin{cases} 1 & \;-1 < x < 1 \\ 0 & \text{everywhere else} \end{cases} \end{aligned}\]

    has a solution in the \(\,tx\)-plane in figure

  • To find the displacement, \(\,\)we compute the D’Alembert solution

\[\scriptsize \begin{aligned} u(x,t) &=\frac{1}{2c}\int_{x-ct}^{x+ct} g(\xi)\,d\xi & \\ &= \frac{1}{2c} \int_{x-ct}^{x+ct} 0\, d\xi=0, & (x,t) \in \text{Region 1}\\ &= \frac{1}{2c} \int_{-1}^{x+ct} 1\, d\xi=\frac{1+x+ct}{2c}, & (x,t) \in \text{Region 2}\\ &= \frac{1}{2c} \int_{-1}^{1} 1\, d\xi=\frac{1}{c}, & (x,t) \in \text{Region 3}\\ &= \frac{1}{2c} \int_{x-ct}^{1} 1\, d\xi=\frac{1-x+ct}{2c}, & (x,t) \in \text{Region 4}\\ &= \frac{1}{2c} \int_{x-ct}^{x+ct} 0\, d\xi=0, & (x,t) \in \text{Region 5}\\ &= \frac{1}{2c} \int_{x-ct}^{x+ct} 1\, d\xi=t, & (x,t) \in \text{Region 6} \end{aligned}\]

\(~\)

fig = plt.figure(figsize=(6, 4))
ax = plt.axes(xlim=(-15, 15), ylim=(-0.1, 1.5))

ax.set_xticks([-15, -10, -5, 0, 5, 10, 15])
ax.set_yticks([0, 0.5, 1.0, 1.5])
ax.set_xlabel('$x$')
ax.set_ylabel('$u(x,t)$')

plt.close()

time_text = ax.text(-2, 1.3, '')
line, = ax.plot([], [], lw=2)
def init():
    time_text.set_text('t = 0.0')
    line.set_data([], [])
    return (line,)
c = 1
def animate(t):
    time_text.set_text(f't = {t:3.1f}')    
    xx = np.linspace(-15, 15, 300)   
    uu = np.zeros_like(xx)
    
    for i, x in enumerate(xx):
        ch1 = x +c*t
        ch2 = x -c*t
        if ch1 <-1.0 or ch2 > 1.0:
            uu[i] = 0.0
        elif t < 1.0 /c:
            if ch2 <-1.0:
                uu[i] = (1.0 +ch1) /(2.0*c)
            elif ch1 < 1.0:
                uu[i] = t
            else:
                uu[i] = (1.0 -ch2) /(2.0*c)           
        else:
            if ch1 < 1.0:
                uu[i] = (1.0 +ch1) /(2.0*c)
            elif ch2 <-1.0:
                uu[i]=1.0 /c
            else:
                uu[i] = (1.0 -ch2) /(2.0*c) 
                       
    line.set_data(xx, uu)
    return (line,)
tt = list(np.linspace(0, 20, 50))
anim = animation.FuncAnimation(fig, animate, init_func=init, 
    frames=tt, interval=300, blit=True)

HTML('<center>' + anim.to_html5_video() + '</center>')
Initial Velocity Given: \(\;u_t(x,0)=\begin{cases} 1 & \;-1 < x < 1 \\ 0 & \text{everywhere else} \end{cases}\)

\(~\)

  • Solution of the Semi-Infinite String via the D’Alembert Formula

    • In the remainder of the section, \(\,\)we will solve the initial-boundary-value problem for the semi-infinite string

      \[ \begin{aligned} u_{tt} &=c^2 u_{xx} && \color{red}{0 < x < \infty}, \; 0 < t < \infty \\ u(0, t) &= 0 && 0 < t < \infty \\ \begin{array}{r} u(x, 0) \\ u_t(x, 0) \end{array} & \begin{array}{c} = f(x) \\ = g(x) \end{array} && 0 < x <\infty \end{aligned}\]

    • We proceed in a manner similar to that used with the infinite string, \(\,\)which is to find the general solution to the PDE

      \[ u(x,t) = \phi(x -ct) +\psi(x +ct) \tag{GS}\label{eq:GS}\]

      If we now substitute this general solution into the initial conditions, \(\,\)we arrive at

      \[{\begin{aligned} \phi(x - ct)=\frac{1}{2}f(x -ct) \,-\,&\frac{1}{2c}\int_{x_0}^{x -ct} g(\xi)\,d\xi -\frac{C}{2c} \\ \psi(x +ct)=\frac{1}{2}f(x +ct) \,+\,&\frac{1}{2c}\int_{x_0}^{x +ct} g(\xi)\,d\xi +\frac{C}{2c} \\ \end{aligned}} \tag{IM}\label{eq:IM}\]

    • We now have a problem we didn’t encounter when dealing with the infinite string. \(\,\)Since we are looking for the solution \(\,u(x,t)\,\) everywhere in the first quadrant \((x>0,\;t>0)\,\) of the \(\,tx\) plane, \(\,\)it is obvious that we must find

      \[\begin{aligned} \phi(x -ct) &\;\;\; \color{red}{\text{for all } -\infty < x -ct < \infty} \\ \psi(x +ct)&\;\;\; \text{for all } \phantom{xxx} 0 < x +ct < \infty \end{aligned}\]

    • Unfortunately, \(\,\)the first equation only gives us \(\,\phi(x -ct)\,\) for \(\,x -ct \geq 0\), \(\,\)since our initial data \(\,f(x)\) and \(g(x)\) are only known for positive arguments

    • As long as \(x -ct \geq 0\), \(\,\)we have no problem, \(\,\)since we can substitute \(\eqref{eq:IM}\) into the general solution \(\eqref{eq:GS}\) to get

      \[ {u(x,t) =\frac{1}{2} \left[ f(x -ct) +f(x +ct) \right] +\frac{1}{2c} \int_{x-ct}^{x+ct} g(\xi)\,d\xi,\;\;\; x \geq ct} \]

      The question is, \(\,\)what to do when \(x < ct\)?

    • When \(x < ct\), \(\,\)substituting the general solution \(\eqref{eq:GS}\) into the BC \(\,u(0,t)=0\,\) gives

      \[\color{blue}{\phi(-ct)=-\psi(ct)}\]

      and, \(\,\)hence, \(\,\)by functional substitution

      \[ {\phi(x -ct)={\color{red}{-}}\frac{1}{2}f({\color{red}{ct -x}}) {\color{red}{-}}\frac{1}{2c}\int_{x_0}^{{\color{red}{ct -x}}} g(\xi)\,d\xi {\color{red}{-}}\frac{C}{2c}} \]

    • Substituting this value of \(\phi\) into the general solution \(\eqref{eq:GS}\) gives

      \[ {u(x,t) =\frac{1}{2} \left[ f(x +ct) -f(ct -x) \right] +\frac{1}{2c} \int_{ct -x}^{x+ct} g(\xi)\,d\xi,\;\;\;0<x<ct} \]

    • For \(x \geq ct\), \(\,\)the solution is the same as the D’Alembert solution for the infinite wave, while for \(x < ct\), \(\,\)the solution \(u(x,t)\) is modified as a result of the wave reflecting from the boundary (The sign of the wave is changed when it’s reflected)

    • The straight lines

      \[\begin{aligned} x +ct &= \text{constant}\\ x -ct &= \text{constant} \end{aligned}\]

      are known as characteristics, \(\,\)and it is along these lines that disturbances are propagated. \(\,\)Characteristics are generally associated with hyperbolic equations

\(~\)

Example \(\,\)Solve the semi-infinite string problem

\(~\)

\[ \begin{aligned} u_{tt} &=c^2 u_{xx} && 0 < x < \infty,\; 0 < t < \infty \\ u(0,t) &=0 && 0 < t < \infty \\ \begin{array}{r} u(x, 0) \\ u_t(x, 0) \end{array} &\, \begin{array}{l} = f(x) \\ = 0 \end{array} && 0 < x <\infty \end{aligned}\]

\(~\)

fig = plt.figure(figsize=(6, 4))
ax = plt.axes(xlim=(0, 5), ylim=(-1.0, 1.5))

ax.set_xticks([0, 2.5, 5])
ax.set_yticks([-1.0, -0.5, 0.0, 0.5, 1.0, 1.5])
ax.set_xlabel('$x$')
ax.set_ylabel('$u(x,t)$')

plt.close()

time_text = ax.text(2.2, 1.3, '')
line, = ax.plot([], [], lw=2)
def init():
    time_text.set_text('t = 0.0')
    line.set_data([], [])
    return (line,)
c = 1
def f_i(x):
    if x >= 1 and x <= 2:
        u = 1.0
    else:
        u = 0.0
    return u
    
def animate(t):
    time_text.set_text(f't = {t:3.1f}')        
    xx = np.linspace(0, 5, 400)   
    uu = np.zeros_like(xx)
    
    for i, x in enumerate(xx):
        ch1 = x +c*t
        ch2 = x -c*t
        if ch2 >= 0:
            uu[i] = 0.5*(f_i(ch1) +f_i(ch2))
        else:
            uu[i] = 0.5*(f_i(ch1) -f_i(-ch2))
                       
    line.set_data(xx, uu)
    return (line,)
tt = list(np.linspace(0, 7, 100))
anim = animation.FuncAnimation(fig, animate, 
        init_func=init, frames=tt, interval=750, blit=True)
HTML('<center>' + anim.to_html5_video() + '</center>')
Semi-infinite String

\(~\)

Example \(\,\)Solve the semi-infinite string problem

\[ \begin{aligned} u_{tt} &=c^2 u_{xx} && 0 < x < \infty,\; 0 < t < \infty \\ u_x(0,t) & =0 && 0 < t < \infty \\ \begin{array}{r} u(x, 0) \\ u_t(x, 0) \end{array} &\; \begin{array}{l} = f(x) \\ = 0 \end{array} && 0 < x <\infty \end{aligned}\]

in a manner analogous to the way the semi-infinite string problem was solved in the section

Solution \(\,\)For \(x \geq ct\), \(\,\)the solution is the same as the D’Alembert solution for the infinite wave, \(\,\)while for \(x < ct\), \(\,\)the solution \(u(x,t)\) is modified as a result of the wave reflecting from the boundary

\[ {u(x,t) =\frac{1}{2} \left[ f(x +ct) +f(ct -x) \right] +\frac{1}{2c} \left[ \int_0^{ct -x} g(\xi)\,d\xi + \int_0^{x +ct} g(\xi)\,d\xi \right] ,\;\; 0 < x < ct} \]

\(~\)

  • The Nonhomogeneous Wave Equation

    • We consider now the pure nonhomogeneous wave equation:

      \[ \begin{aligned} u_{tt} &\! =c^2 u_{xx} +\color{red}{F(x, t)} && -\infty < x < \infty,\; 0 < t < \infty \\ \begin{array}{r} u(x, 0) \\ u_t(x, 0) \end{array} & \begin{array}{c} = 0 \\ = 0 \end{array} && -\infty < x <\infty \end{aligned}\]

      which includes the forcing term \(F(x, t)\), \(~\)but zero initial conditions

    • We again make the change of variables \(\,\xi=x +ct\,\) and \(\,\eta=x -ct\). \(\,\)The differential equation then becomes

      \[ \begin{aligned} & u_{\xi\eta} = \color{red}{-\frac{1}{4c^2} F(\xi, \eta)} \\ \\ &\color{blue} {\left.\begin{matrix} u = 0\\ u_\xi = u_\eta \end{matrix} \;\;\right\} \;\text{ at } \xi=\eta \;\; \leftarrow \left.\begin{matrix} u = 0 \\ u_t = 0 \end{matrix} \;\;\right\} \text{ at } t=0} \end{aligned}\]

    \[ \begin{aligned} \\[5pt] &\Downarrow {\scriptstyle\text{ integrating with respect to } \xi \text{ and } \eta, \text{ respectively} }\\[5pt] u_\eta \;&{\scriptsize= -\frac{1}{4c^2} \int_\eta^\xi F\left(\bar{\xi}, \eta \right) \,d\bar{\xi} +c_1}, \:\: u_\xi \;{\scriptsize= -\frac{1}{4c^2} \int_\xi^\eta F\left(\xi, \bar{\eta} \right) \,d\bar{\eta} +c_1} \\[5pt] &\Downarrow {\scriptstyle\text{ integrating with respect to } \eta \text{ and } \xi, \text{ respectively}}\\[5pt] u \;&{\scriptsize= -\frac{1}{4c^2} \int_\eta^\xi \left[ \int_{\bar{\eta}}^\xi F\left( \bar{\xi}, \bar{\eta} \right) \,d\bar{\xi} \right ] \,d\bar{\eta} +c_1 \eta +c_2} \\ \;&{\scriptsize= -\frac{1}{4c^2} \int_\xi^\eta \left[ \int_{\bar{\xi}}^\eta F\left( \bar{\xi}, \bar{\eta} \right) \,d\bar{\eta} \right ] \,d\bar{\xi} +c_1 \xi +c_3} \end{aligned}\]

    • We let \(\,\bar{\xi}=\beta +c\tau\;\) and \(\,\bar{\eta}=\beta -c\tau\), \(\,\)(\(t \ge \tau\), \(\,\beta=x\)). \(\,\)The domain of integration \(\,\eta \leq \bar{\eta} \leq \bar{\xi} \leq \xi\;\) becomes

      \[ \begin{aligned} \eta \leq \beta -c\tau &\leq \beta +c\tau \leq \xi \\ &\Downarrow \\ \eta +c\tau \leq &\;\beta \leq\xi -c \tau \\ 0 \leq &\;\tau \leq \frac{1}{2c} (\xi -\eta) \end{aligned}\]

    • The transformation from \(\,(\bar{\xi}, \,\bar{\eta})\,\) to \(\,(\beta, \,\tau)\,\) gives the solution formula

      \[ \begin{aligned} u &= -\frac{1}{4c^2} \int_\eta^\xi \left[ \int_{\bar{\eta}}^\xi F\left( \bar{\xi}, \bar{\eta} \right) \,d\bar{\xi} \right ] \,d\bar{\eta} \\[5pt] &\big\Downarrow {\scriptstyle \;\;d\bar{\xi}\,d\bar{\eta} \,=\, \begin{vmatrix} \frac{\partial \bar{\xi}}{\partial \beta}& \frac{\partial \bar{\xi}}{\partial \tau} \\ \frac{\partial \bar{\eta}}{\partial \beta}& \frac{\partial \bar{\eta}}{\partial \tau} \end{vmatrix} \,d\beta\,d\tau \,=\, -2c \,d\beta\,d\tau}\\[5pt] &=\frac{1}{2c} \int_0^{(\xi -\eta)/2c} \left[ \int_{\eta +c\tau}^{\xi -c\tau} F\left( \beta,\tau \right) \,d\beta \right ] \,d\tau \\[5pt] &\Downarrow {\scriptstyle \;\xi=x +ct, \;\eta=x -ct} \\ {\color{red}{u}}\; &{\color{red}{= \frac{1}{2c} \int_0^t \left[ \int_{x -c(t -\tau)}^{x +c(t -\tau)} F\left( \beta,\tau \right) \,d\beta \right ] \,d\tau}} \end{aligned} \]

\(~\)

12.4 The Wave Equation in Two and Three Dimensions (Free Space)

The problem of this section is to generalize the D’Alembert solution to two and three dimensions

  • Waves in Three Dimensions

    We start by considering spherical waves in three dimensions that have given ICs; \(\,\)that is, \(\,\)we would like to solve the initial value problem

    \[\begin{aligned} u_{tt} &=c^2(u_{xx} +u_{yy} +u_{zz}),\quad \begin{cases} -\infty < x <\infty \\ -\infty < y <\infty \\ -\infty < z <\infty \end{cases} \\ u(&x,y,z,0)=\phi(x,y,z) \\ u_t(&x,y,z,0)=\psi(x,y,z) \end{aligned} \tag{3D}\label{eq:3D}\]

    • To solve this problem, \(\,\)we first solve the simpler one (set \(\phi=0\))

      \[\color{red}{\begin{aligned} u_{tt}&=c^2(u_{xx} +u_{yy} +u_{zz}),\quad (x,y,z) \in \mathbb{R}^3\\ u(&x,y,z,0)=0 \\ u_t(&x,y,z,0)=\psi(x,y,z) \end{aligned}} \tag{3Dv}\label{eq:3Dv}\]

      This problem can be solved by the Fourier transform

      \[\begin{aligned} \mathcal{F}(u) &=\frac{1}{(2\pi)^{3/2}}\iiint_{-\infty}^\infty u(x,y,z,t) e^{-i(\omega_1 x+\omega_2 y+\omega_3 z)}\,dx\,dy\,dz =U(\omega_1,\omega_2,\omega_3,t)\\ \mathcal{F}(\psi)&=\frac{1}{(2\pi)^{3/2}}\iiint_{-\infty}^\infty \psi(x,y,z) e^{-i(\omega_1 x+\omega_2 y+\omega_3 z)}\,dx\,dy\,dz =\Psi(\omega_1,\omega_2,\omega_3) \end{aligned}\]

    \[\scriptsize \begin{aligned} &\Downarrow\;\; \mathcal{F}^{-1} \\ \\ u(x,y,z,t)= \frac{1}{(2\pi)^{3/2}} \lim_{L\to\infty} \underset{\omega_1^2 +\omega_2^2 +\omega_3^2 < L^2}{\iiint} \Psi(\omega_1,\omega_2,\omega_3) & \frac{\sin ct \sqrt{ \omega_1^2 +\omega_2^2 +\omega_3^2 }}{c\sqrt{\omega_1^2 +\omega_2^2 +\omega_3^2 }} \,e^{i(\omega_1 x +\omega_2 y +\omega_3 z)}\,d\omega_1 \,d\omega_2 \,d\omega_3 \\ = \frac{1}{(2\pi)^{3/2}} \iiint_{-\infty}^\infty \psi(\xi,\eta,\zeta)\; {\tiny \left[ \frac{1}{(2\pi)^{3/2}} \lim_{L\to\infty} \underset{\omega_1^2 +\omega_2^2 +\omega_3^2 < L^2}{\iiint} \right. } & {\tiny \left. e^{-i\left[\omega_1 (\xi-x) +\omega_2 (\eta-y) +\omega_3 (\zeta-z) \right]} \; \frac{\sin ct \sqrt{ \omega_1^2 +\omega_2^2 +\omega_3^2 }}{c\sqrt{\omega_1^2 +\omega_2^2 +\omega_3^2 }}\, \,d\omega_1 \,d\omega_2 \,d\omega_3 \,\right] } \,d\xi \,d\eta \,d\zeta \end{aligned}\]

    • We introduce spherical coordinates \((\varrho,\varphi,\vartheta)\) in the \((\omega_1,\omega_2,\omega_3)\) space with the north pole \(\varphi=0\) in the direction of the vector \(\left \langle \xi-x,\eta-y,\zeta-z \right \rangle\). \(\,\)Then the integral in \((\omega_1,\omega_2,\omega_3)\) becomes

      \[\scriptsize\begin{aligned} \frac{1}{(2\pi)^{3/2}} \lim_{L\to\infty} \underset{\omega_1^2 +\omega_2^2 +\omega_3^2 < L^2}{\iiint} &e^{-i\left[\omega_1 (\xi-x) +\omega_2 (\eta-y) +\omega_3 (\zeta-z)\right]}\; \frac{\sin ct \sqrt{ \omega_1^2 +\omega_2^2 +\omega_3^2 }}{c\sqrt{\omega_1^2 +\omega_2^2 +\omega_3^2 }}\, \,d\omega_1 \,d\omega_2 \,d\omega_3 \\ &\;\;\Downarrow \; {\tiny r=|\mathbf{r}|=\sqrt{(\xi-x)^2 +(\eta -y)^2 +(\zeta -z)^2},\;\; \varrho =|\boldsymbol{\varrho}| = \sqrt{\omega_1^2 +\omega_2^2 +\omega_3^2} }\\ &=\frac{1}{(2\pi)^{3/2}} \lim_{L\to\infty} \int_0^L \int_0^{2\pi} \int_0^\pi \;e^{-ir \varrho\cos \varphi} \;\frac{\sin ct \varrho}{c\varrho}\; \varrho^2 \sin\varphi \,d\varphi\,d\vartheta\,d\varrho \\ &\;\;\Downarrow \; {\tiny \alpha=\cos\varphi, \;d\alpha=-\sin\varphi d\varphi } \\ &=\frac{1}{(2\pi)^{3/2}} \lim_{L\to\infty} \frac{2\pi}{c} \int_0^L \; \varrho \sin ct\varrho \underbrace{\left[ \frac{e^{-ir\varrho \alpha}}{-ir\varrho} \right]_{-1}^1 }_{\frac{2}{r\varrho}\sin r\varrho} \,d\varrho \\ &=\frac{1}{(2\pi)^{3/2}} \lim_{L\to\infty} \frac{4\pi}{cr} \int_0^L \;\sin ct \varrho \cdot \sin r\varrho \,d\varrho \end{aligned}\]

    • To treat the integration with respect to \(\xi\), \(\,\eta\), and \(\,\zeta\), \(~\)we introduce spherical coordinates \((r,\phi,\theta)\) in the \((\xi,\eta,\zeta)\) space with their origin at \((x,y,z)\), \(\,\)so that

      \[\begin{aligned} \xi-x &=r\sin\phi \cos\theta \\ \eta-y&=r\sin\phi \sin\theta \\ \zeta-z&=r\cos\phi \end{aligned}\]

      Then the solution becomes

      \[\scriptsize\begin{aligned} {\normalsize\color{red}{u(x,y,z,t)}} &= {\tiny \frac{4\pi}{(2\pi)^3 c} \lim_{L\to \infty} \int_0^\infty \int_0^{2\pi} \int_0^\pi \;\psi(x +r\sin\phi\cos\theta,y +r\sin\phi\sin\theta, z+r\cos\phi)\; \frac{1}{r}\;\left[ \int_0^L \sin ct \varrho \sin r\varrho\, d\varrho \right] r^2\sin\phi \, d\phi \, d\theta \,dr }\\ &= {\tiny \frac{1}{2\pi^2 c} \lim_{L\to \infty} \int_0^L \sin ct\varrho \left[\int_0^\infty \sin r\varrho \underbrace{\left( \int_0^{2\pi} \int_0^\pi \psi(x +r\sin\phi\cos\theta,y +r\sin\phi\sin\theta, z+r\cos\phi) \,r\,\sin\phi \,d\phi\,d\theta \right) }_{g(r)} dr \right] d\varrho } \\ &= {\tiny \frac{1}{2\pi^2 c} \frac{\pi}{2} \int_0^\infty \sin ct \varrho \;\underbrace{\left[\frac{2}{\pi} \int_0^\infty \sin r\varrho \,g(r)\,dr \right]}_{G(\varrho)} d\varrho = \frac{1}{4\pi c} \underbrace{\int_0^\infty \sin ct \varrho \;G(\rho)\,d\varrho}_{g(ct)} } \\ &= {\tiny \frac{t}{4\pi (ct)^2} \int_0^{2\pi}\int_0^\pi \psi(x +ct\sin\phi\cos\theta,y +ct\sin\phi\sin\theta, z+ct\cos\phi) \,(ct)^2\sin\phi \,d\phi \,d\theta } = {\normalsize\color{red}{t\bar{\psi}}} \end{aligned}\]

      where \(\bar{\psi}\) is the average of the initial disturbance \(\psi\) over the surface of the sphere of radius \(\,ct\,\) centered at \((x,y,z)\)

    • The interpretation of this solution is that the initial disturbance \(\,\psi\,\) radiates outward spherically (velocity \(c\)) at each point, \(\,\)so that after so many seconds, \(\,\)the point \(\,(x,y,z)\) will be influenced by those initial disturbances on a sphere (of radius \(\,ct\)) around that point

    • Now, \(\,\)to finish the problem, \(\,\)what about the other half; \(\,\)that is

      \[\color{red}{\begin{aligned} u_{tt}&=c^2(u_{xx} +u_{yy} +u_{zz}),\quad (x,y,z) \in \mathbb{R}^3 \\ u(&x,y,z,0)=\phi(x,y,z) \\ u_t(&x,y,z,0)=0 \end{aligned}} \tag{3Dp}\label{eq:3Dp}\]

      This is easy: \(\,\)A famous theorem developed by Stokes says all we have to do to solve this problem is to change the ICs to \(\,v=0\), \(\,v_t=\phi\), \(\,\)and then differentiate this solution with respect to time:

      \[ v=\int_0^t u\,dt,\;v_t=u, \;u_{tt}=\frac{\partial}{\partial t} v_{tt}, \;u_{xx}=\frac{\partial}{\partial t} v_{xx}\]

    • In other words, \(\,\)we solve

      \[\begin{aligned} v_{tt}&=c^2(v_{xx} +v_{yy} +v_{zz}),\quad (x,y,z) \in \mathbb{R}^3 \\ v(&x,y,z,0)=0 \\ v_t(&x,y,z,0)=\phi(x,y,z) \end{aligned}\]

      to get \(v=t\bar{\phi}\,\) and then differentiate with respect to time. \(\,\)This gives us the solution

      \[\color{red}{u=\frac{\partial}{\partial t} \left[ t\bar{\phi} \right]}\]

      to problem \(\eqref{eq:3Dp}\)

    • We now have the solution to our general three-dimensional problem \(\eqref{eq:3D}\). \(\,\)It’s just

      \[ \color{red}{u(x,y,z,t)=\frac{\partial}{\partial t} \left[ t\bar{\phi} \right]+t\bar{\psi}}\]

      where \(\,\bar{\phi}\,\) and \(\,\bar{\psi}\,\) are the averages of the functions \(\,\phi\,\) and \(\,\psi\,\) over the sphere of radius \(\,ct\,\) centered at \(\,(x,y,z)\)

      \(~\)

    • Suppose now the initial disturbances \(\,\phi\,\) and \(\,\psi\,\) are zero except for a small sphere. \(\,\)As time increases, \(\,\)the radius of the sphere around \((x,y,z)\) increases with velocity \(\,c\)

    • and so after \(\,t_2\) seconds, \(\,\)it will finally intersect the initial disturbance region, \(\,\)and, \(\,\)hence, \(\,u(x,y,z,t)\,\) becomes nonzero

    • For \(\,t_2 < t < t_3\), \(\,\)the solution at \(\,(x,y,z)\,\) will be nonzero, \(\,\)since the sphere intersects the disturbance region

    • but when \(\,t=t_3\), \(\,\)the solution at \(\,(x,y,z)\,\) abruptly becomes zero again. \(\,\)In other words, \(\,\)the wave disturbance originating from the initial-disturbance region has a sharp trailing edge. \(\,\)This general principle is known as Huygen’s principle for the three dimensions, \(\,\)and \(\,\)it is the reason why sound waves in three dimensions stimulate our ears but die off instanteously when the wave has passed

\(~\)

Example \(\,\)Illustrate by picture and words the spherical wave solution of the three-dimensional problem

\[\begin{aligned} u_{tt}&=c^2(u_{xx} +u_{yy} +u_{zz}), \quad \begin{cases} -\infty < x <\infty \\ -\infty < y <\infty \\ -\infty < z <\infty \end{cases} \\ u(&x,y,z,0)=0 \\ u_t(&x,y,z,0)= \begin{cases} \;1 & x^2+y^2+z^2 \leq 1 \\ \;0 & \text{elsewhere} \end{cases} \end{aligned}\]

\(~\)

  • Two-Dimensional Wave Equation

    To solve the two-dimensional problem

    \[\begin{aligned} u_{tt}&=c^2(u_{xx} +u_{yy}),\quad \begin{cases} -\infty < x <\infty \\ -\infty < y <\infty \end{cases} \\ u(&x,y,0)=\phi(x,y) \\ u_t(&x,y,0)=\psi(x,y) \end{aligned} \tag{2D}\label{eq:2D}\]

    we merely let the initial disturbances \(\,\phi\,\) and \(\,\psi\,\) in the three-dimensional problem depend on only the two variables \(\,x\,\) and \(\,y\). \(\,\)Doing this, the three-dimensional formula

    \[ u = \frac{\partial}{\partial t} [t\bar{\phi}] +t\bar{\psi} \]

    for \(\,u\,\) will describe cylindrical waves and, \(\,\)hence, \(\,\)give us the solution for the two-dimensional problem

    • This technique is called the method of descent. \(\,\)Carrying out the computations, \(\,\)we get

      \[\scriptstyle\begin{aligned} \bar{\psi}(x,y,t) &= \frac{1}{4\pi (ct)^2} \color{red}{\int_0^{2\pi} \int_{0}^\pi} \psi(x +ct\sin\phi\cos\theta,y+ct\sin\phi\sin\theta)\, \color{red}{(ct)^2 \sin\phi \, d\phi \, d\theta} \;\;\leftarrow \text{spherical surface}\\ &\Downarrow \;\; {\tiny r = ct \,\sin \phi, \; \,dr =ct \, \cos\phi \,d\phi\;\;\rightarrow\;\; \int_0^{\pi} ct \sin\phi\,d\phi = 2\int_0^{ct}\frac{r}{\sqrt{(ct)^2 -r^2}} dr } \\ &= \frac{1}{2\pi ct} \, \color{red}{\int_0^{2\pi} \int_0^{ct}} \psi(x +r\cos\theta, y+r\sin\theta)\,\frac{1}{\sqrt{(ct)^2 -r^2}}\,\color{red}{r\,dr\,d\theta} \;\;\leftarrow \text{circle interior}\\ \bar{\phi}(x,y,t) &= \frac{1}{2\pi ct} \, \color{red}{\int_0^{2\pi} \int_0^{ct}} \phi(x +r\cos\theta, y+r\sin\theta)\,\frac{1}{\sqrt{(ct)^2 -r^2}}\,\color{red}{r\,dr\,d\theta} \end{aligned}\]

      Note that in this solution, \(\,\)the two integrals of the initial conditions \(\,\phi\,\) and \(\,\psi\,\) are integrated over the interior of a circle with center \(\,(x,y)\,\) and radius \(\,ct\). \(\,\)We see that initial disturbances give rise to sharp leading waves, \(\,\)but not to sharp trailing waves. \(\,\)Thus, Huygen’s principle doesn’t hold in two dimensions

  • One-Dimensional Wave Equation

    Finally, \(\,\)if we assume the initial conditions \(\,\phi\,\) and \(\,\psi\,\) depend only on one variable, \(\,\)this gives rise to plane waves and, \(\,\)hence, \(\,\)the preceding equation descends one more dimension

    \[\begin{aligned} \bar{\psi}(x,t) &= {\tiny \frac{1}{2\pi ct} \, \int_0^{2\pi} \int_0^{ct} \frac{\psi(x +r\cos\theta)}{\sqrt{(ct)^2 -r^2}}\,r\,dr\,d\theta }\\ &\Downarrow {\tiny r^2=x^2 + y^2, \; (ct)^2 -x^2 = \alpha^2 }\\ &={\tiny \frac{1}{2\pi ct}\, \int_{x -ct}^{x +ct} \underbrace{\int_{-\alpha}^{\alpha} \frac{1}{\sqrt{\alpha^2 -y^2}}\,dy}_{ \pi} \;\psi(x) \,dx } = \frac{1}{2ct} \int_{x-ct}^{x+ct} \psi(x)\,dx \\ \bar{\phi}(x,t) &= \frac{1}{2ct} \int_{x-ct}^{x+ct} \phi(x)\,dx \end{aligned}\]

    to the well-known D’Alembert solution. \(\,\)Note in the D’Alembert solution, \(\,\)the initial position \(\,\phi\,\) gives rise to sharp trailing edges, \(\,\)but the initial velocity \(\,\psi\,\) does not. \(\,\)In other words, \(\,\)one dimension is a little unusual in that the initial position satisfies Huygen’s principle, \(\,\)but the initial velocity does not

\(~\)

Example \(\,\)What is the two dimensional solution of the analogous cylindrical-wave problem

\[\begin{aligned} u_{tt}&=c^2(u_{xx} +u_{yy}),\quad \begin{cases} -\infty < x <\infty \\ -\infty < y <\infty \end{cases} \\ u(&x,y,0)=0 \\ u_t(&x,y,0)= \begin{cases} \;1 & x^2+y^2 \leq 1 \\ \;0 & \text{elsewhere} \end{cases} \end{aligned}\]

\(~\)

12.5 Boundary Conditions Associated with the Wave Equation

The purpose of this section is to discuss some of the various types of BCs that are associated with physical problems of wave motions. \(\,\)Here, \(\,\)we will stick to one-dimensional problems where the BCs (linear ones) are generally groupded into one of three kinds

  • First Kind: \(\,\)Controlled End Points

    We are now involved with problems like

    \[ \begin{aligned} u_{tt} &=c^2 u_{xx} && \color{red}{0 < x < 1},\;\; 0 < t < \infty \\ \\ \begin{array} {r} u(0, t) \\ u(1,t) \end{array} & \begin{array} {r} =g_1(t) \\ =g_2(t) \end{array} && 0 < t < \infty \\ \\ \begin{array} {r} u(x, 0) \\ u_t(x, 0) \end{array} & \begin{array} {r} =f(x) \\ =g(x) \end{array} && 0 \leq x \leq L \end{aligned}\]

    where we control the end points so that they move in a given manner

    A typical problem of this kind would involve suddenly twisting (at \(t=1\)) the right end of a fastened rod so many degrees and observing the resulting tortional vibration

  • Second Kind: \(\,\)Force Given on the Boundaries

    In as much as the vertical forces on the string at the left and right ends are given by \(\,Tu_x(0,t)\,\) and \(\,Tu_x(1,t)\), \(\,\)respectively, \(\,\)by allowing the ends of the string to slide vertically on frictionless sleeves, \(\,\)the boundary conditions become

    \[\begin{aligned} u_x(0,t)&=0 \\ u_x(1,t)&=0 \end{aligned}, \quad 0 < t < \infty\]

    If a vertical force \(f(t)\) is applied at the end \(x=1\), \(\,\)then the BC would be

    \[u_x(1,t)=\frac{1}{T}\,f(t)\]

  • Third Kind: \(\,\)Elastic Attachment on the Boundaries

    Consider finally a violin string whose ends are attached to an elastic arrangement like the one shown in figure

    Here, \(\,\)the spring attachments at each end give rise to vertical forces proportional to the displacements \(\,-u(0,t)\,\) and \(\,-u(1,t)\)

    Setting the vertical tensions of the spring at the two ends \(\,-Tu_x(0,t)\,\) and \(\,Tu_x(1,t)\,\) equal to these displacements (multiplied by the spring constant \(h\)) gives us our desired BCs

    \[\begin{aligned} -T u_x(0,t) &= -hu(0,t)\\ Tu_x(1,t) &= -hu(1,t) \end{aligned}\]

    We can rewrite these two homogeneous BCs as

    \[\begin{aligned} u_x(0,t) -\frac{h}{T} u(0,t) &=0\\ u_x(1,t) +\frac{h}{T} u(1,t) &=0 \end{aligned}\]

    If the two spring attachments are displaced according to the functions \(\,\theta_1(t)\,\) and \(\,\theta_2(t)\), \(\,\)we would have the nonhomogeneous BCs

    \[\begin{aligned} u_x(0,t) &=\frac{h}{T} \left[u(0,t) -\theta_1(t) \right]\\ u_x(1,t) &=-\frac{h}{T} \left[u(1,t) -\theta_2(t) \right] \end{aligned}\]

\(~\)

NOTE \(\,\)What is the general nature of BC

\[ u_x(0,t) =\frac{h}{T} \left[u(0,t) -\theta_1(t) \right]\]

when \(~h \to \infty\) and \(~h \to 0\)

\(~\)

12.6 The Finite Vibrating String (Standing Waves)

  • So far, \(\,\)we have studied the wave equation \(\,u_{tt}=c^2 u_{xx}\,\) for the unbounded domain and have found D’Alembert solutions to be certain traveling waves (moving in opposite directions)

  • When we study the same wave equation in a bounded region of space \(0 < x < L\), \(\,\)we find that the waves no longer appear to be moving due to their repeated interaction with the boundaries and, in fact, often appear to be what are known as standing waves

  • Separation of Variables Solution to the Finite Vibrating String

    Consider what happens when a guitar string (fixed at both ends \(x=0,\,L\)) described by the simple hyperbolic IBVP is set in motion

    \[ \begin{aligned} u_{tt} &=c^2 u_{xx} && \color{red}{0 < x < L},\;\; 0 < t < \infty \\ \\ \begin{array} {r} u(0, t) \\ u(1,t) \end{array} &\; \begin{array} {r} =0 \\ =0 \end{array} && 0 < t < \infty \\ \\ \begin{array} {r} u(x, 0) \\ u_t(x, 0) \end{array} &\; \begin{array} {r} =f(x) \\ =g(x) \end{array} && 0 \leq x \leq L \end{aligned} \tag{FS}\label{eq:FS}\]

    • What happens is that the traveling wave solution to the PDE and IC keeps reflecting from the boundaries in such a way that the wave motion does not to be moving, \(\,\)but, \(\,\)in fact, appears to be vibrating in one position:

    • We start by seeking standing-wave solutions to the PDE; \(\,\)that is solutions of the form

      \[u(x,t)=X(x)T(t)\]

      Substituting this expression into the wave equation and separating variables gives us the two ODEs

      \[\begin{aligned} \frac{\;\;T''}{c^2 T} &= \frac{\;\;X''}{X} = - \lambda <0\\ &\Downarrow \\ T'' +c^2\lambda T &=0\\ X'' +\lambda X &=0 \end{aligned}\]

      in which only positive values of \(\lambda\) give feasible (nonzero and bounded) solutions

    • The solutions of these two ODEs yield

      \[\begin{aligned} T(t) &= a \cos c\sqrt{\lambda} t +b\sin c\sqrt{\lambda}t\\ X(x)&=d \cos \sqrt{\lambda}x +e\sin\sqrt{\lambda}x \end{aligned}\]

      Plugging this equation into \(u(0,t)=u(L,t)=0\,\) gives

      \[\scriptsize \begin{aligned} u(0,t) & = d \cdot \,T(t)=0 \;\;\Rightarrow \;\; T(t) \neq 0,\;\;d=0\\ u(L,t) & = e\sin\sqrt{\lambda}L \cdot T(t)=0\;\; \Rightarrow T(t) \neq 0,\;\; e\neq0, \;\sin\sqrt{\lambda}L=0\;\;\\ &\normalsize\Rightarrow \;\;\lambda_n=\left( \frac{n\pi}{L}\right)^2,\;\;n=1,2,3,\cdots \end{aligned}\]

      Hence, \(\,\)we have now found a sequence of simple vibrations (which we subscript with \(n\))

      \[ \begin{aligned} u_n(x,t)&= T_n(t)X_n(x) \\=&\left[ a_n \cos\frac{n\pi c t}{L} +b_n \sin \frac{n\pi c t}{L} \right]\,\sin \frac{n\pi x}{L}, \;\;n=1,2,3,\cdots \end{aligned}\]

    • All of which satisfy the wave equation and the BCs, \(\,\)and constitute a family of standing waves. \(\,\)Since any sum of these vibrations is also a solution to the PDE and BCs (since the PDE and BCs are linear and homogeneous), \(\,\)we add them together in such a way that the resulting sum also agrees with the ICs. \(\,\)Substituting the sum

      \[ u(x,t)= \sum_{n=1}^\infty \left[ a_n \cos\frac{n\pi c t}{L} +b_n \sin \frac{n\pi c t}{L} \right]\,\sin \frac{n\pi x}{L}\]

      into the ICs

      \[u(x,0)=f(x), \;\; u_t(x,0)=g(x)\]

      gives the two equations \[ \sum_{n=1}^\infty a_n \sin \frac{n\pi x}{L}=f(x), \;\; \sum_{n=1}^\infty b_n \frac{n\pi c}{L}\sin\frac{n\pi x}{L}=g(x)\]

    • and using the orthogonality condition

      \[\int_0^L \sin \frac{m\pi x}{L}\, \sin \frac{n\pi x}{L}\,dx=\begin{cases} \;\;0& m \neq n \\ L/2& m = n \end{cases}\]

      we can find the coefficients \(\,a_n\,\) and \(\,b_n\)

      \[\begin{aligned} a_n&= \frac{2}{L} \int_0^L f(x)\, \sin \frac{n\pi x}{L}\,dx\\ b_n&= \frac{2}{n\pi c} \int_0^L g(x)\, \sin \frac{n\pi x}{L}\,dx \end{aligned}\]

NOTE

\[\scriptsize\begin{aligned} \normalsize u(x,t) \;&\normalsize= \sum_{n=1}^\infty \left[ a_n \cos\frac{n\pi c t}{L} +b_n \sin \frac{n\pi c t}{L} \right]\,\sin \frac{n\pi x}{L}\\[5pt] &= \frac{1}{2} \sum_{n=1}^\infty a_n\left[\sin \frac{n\pi}{L} (x+ct) +\sin \frac{n\pi}{L} (x-ct)\right ] \\&\quad-\frac{1}{2} \sum_{n=1}^\infty b_n\left[\cos \frac{n\pi}{L} (x+ct) -\cos \frac{n\pi}{L} (x-ct)\right ]\\[5pt] &\;\;\Big\Downarrow \;\; {\tiny\sum_{n=1}^\infty a_n \sin \frac{n\pi x}{L}=f(x), \;\;\sum_{n=1}^\infty b_n \frac{n\pi c}{L}\sin\frac{n\pi x}{L}=g(x) \;\rightarrow\, -\sum_{n=1}^\infty b_n\, c\cos \frac{n\pi x}{L} +\sum_{n=1}^\infty b_n c=\int^x_0 g(\alpha) d\alpha} \\[5pt] &\normalsize= \frac{1}{2}\left[ f(x+ct) +f(x-ct)\right] +\frac{1}{2c} \int_{x-ct}^{x+ct} g(\alpha)\,d\alpha \end{aligned}\]

\(~\)

Example \(\,\)Find the solution to the vibrating-string problem \(\eqref{eq:FS}\) if the ICs are given by

\[\begin{aligned} u(x,0) &= \sin \frac{\pi x}{L} +0.5 \sin\frac{3\pi x}{L}\\ u_t(x,0) &= 0 \end{aligned}\]

\(~\)

Example \(\,\)What is the solution of the vibrating-string problem \(\eqref{eq:FS}\) if the ICs are

\[\begin{aligned} u(x,0)&= 0\\ u_t(x,0)&= \sin\frac{3\pi x}{L} \end{aligned}\]

\(~\)

Example \(\,\)Solve the damped vibrating-string problem

\[ \begin{aligned} u_{tt} &=c^2 u_{xx} -\beta u_t && \color{red}{0 < x < 1},\;\; 0 < t < \infty \\ \\ \begin{array} {r} u(0, t) \\ u(1,t) \end{array} & \begin{array} {r} =0 \\ =0 \end{array} && 0 < t < \infty \\ \\ \begin{array} {r} u(x, 0) \\ u_t(x, 0) \end{array} & \begin{array} {r} =f(x) \\ =0\phantom{(x)} \end{array} && 0 \leq x \leq 1 \end{aligned}\]

\(~\)

12.7 Solution of Nonhomogeneous Wave Equation via the Finite Transform

  • Consider the nonhomogeneous wave equation

    \[ \begin{aligned} u_{tt} &=c^2 u_{xx} +\sin\pi x && 0 < x < 1,\;\; 0 < t < \infty \\ \\ \begin{array} {r} u(0, t) \\ u(1,t) \end{array} & \begin{array} {r} =0 \\ =0 \end{array} && 0 < t < \infty \\ \\ \begin{array} {r} u(x, 0) \\ u_t(x, 0) \end{array} & \begin{array} {r} =1 \\ =0 \end{array} && 0 \leq x \leq 1 \end{aligned}\]

STEP 1 \(\,\) Determine the transform

Since the \(x\)-variable ranges from \(0\) to \(1\), \(\,\)we use a finite transform. \(\,\)Also, \(\,\)you will see why, \(\,\)in this case, we use the sine transform. \(\,\)We could solve this problem with the Laplace transform by transforming \(\,t\)

STEP 2 \(\,\) Carry out the transformation

\[\begin{aligned} \mathcal{F}_s [u_{tt}]&= c^2\mathcal{F}_s [u_{xx}] +\mathcal{F}_s [\sin\pi x] \\ &\Downarrow \\ \frac{d^2 U_n(t)}{dt^2}&= -(n\pi c)^2 U_n(t) +2n\pi c^2\left[ u(0,t) -(-1)^n u(1,t) \right] +S_n(t) \\ &\Downarrow\; u(0,t)=0, \;\;u(1,t)=0 \\ \frac{d^2 U_n(t)}{dt^2}&= -(n\pi c)^2 U_n(t) +S_n(t) \end{aligned}\]

where

\[ S_n(t) = 2 \int_0^1 \sin \pi x \, \sin n\pi x \,dx = \begin{cases} 1 & n=1 \\ 0 & n=2,3,\cdots \end{cases}\]

If we now transform the initial conditions, \(\,\)we will arrive at the initial conditions for our ODE

\[\begin{aligned} \mathcal{F}_s [u(x,0)]&= U_n(0) = \begin{cases} \displaystyle \frac{4}{n\pi} & n=1,3,\cdots \\ \;\,0 & n=2,4,\cdots \end{cases}\\ \mathcal{F}_s [u_t(x,0)]&= \frac{dU_n(0)}{dt} =0 \end{aligned}\]

So, \(\,\)solving our new initial-value problem, \(\,\)we have

\[\begin{aligned} U_1(t)&= \left( \frac{4}{\pi} -\frac{1}{\pi^2} \right) \,\cos\pi c t +\frac{1}{\pi^2}\\ U_n(t) &= \begin{cases} \quad\;\; 0 & n=2,4,\cdots \\ \displaystyle \frac{4}{n\pi}\cos n\pi c t & n=3,5,\cdots \end{cases} \end{aligned}\]

Hence, \(\,\)the solution \(\,u(x,t)\) of the problem is

\[ u(x,t)=\left[ \left( \frac{4}{\pi} -\frac{1}{\pi^2} \right) \,\cos\pi c t +\frac{1}{\pi^2} \right]\,\sin \pi x +\frac{4}{\pi} \sum_{n=1}^\infty \frac{1}{2n +1} \cos(2n +1)\pi c t \, \sin(2n +1)\pi x\]

12.8 The Vibrating Drumhead (Wave Equation in Polar Coordinates)

  • The problem is to find \(\,u(r,\theta,t)\) (which stands for the height of the drumhead from the plane) that satisfies

    \[\color{red}{\begin{aligned} u_{tt} &= c^2 \left( u_{rr} +\frac{1}{r} u_r +\frac{1}{r^2} u_{\theta\theta} \right) && \\ u(1, \theta, t) &= 0 && 0 < t < \infty \\ u(r, \theta, 0) &= f(r,\theta) && 0 < r < 1\\ u_t(r, \theta, 0) &= g(r,\theta) && \end{aligned}}\]

    We will look for solutions of the form

    \[\color{blue}{u(r,\theta,t)=U(r,\theta)T(t)}\]

    This gives the shape \(\,U(r,\theta)\) of the vibrations times the oscillatory factor \(\,T(t)\)

  • Carrying out this substitution, \(\,\)we arrive at the two equations

    \[\scriptsize\begin{aligned} { u_{tt}} &{ = c^2 \left( u_{rr} +\frac{1}{r} u_r +\frac{1}{r^2} u_{\theta\theta} \right)}\\ &\Downarrow \\ { UT''} &= { c^2 \left( U_{rr} +\frac{1}{r}U_r +\frac{1}{r^2}U_{\theta\theta} \right) T} \\ &\Downarrow \\ { \frac{\;\;T''}{c^2T}} &={\frac{\displaystyle U_{rr} +\frac{1}{r}U_r +\frac{1}{r^2}U_{\theta\theta}}{U} =-\lambda < 0} \\ &\Downarrow \\ \normalsize\color{red}{T''} &\normalsize\color{red}{+ c^2 \lambda T =0} && \text{Simple Harmonic Motion}\\ \normalsize\color{red}{U_{rr}} &\normalsize\color{red}{+\frac{1}{r}U_r +\frac{1}{r^2}U_{\theta\theta} +\lambda U=0} && \text{Helmholtz Equation} \end{aligned}\]

  • For the next step, \(\,\)we want to solve the Helmholtz equation, \(\,\)but, \(\,\)first, \(\,\)it needs a boundary condition. \(\,\)To find it, \(\,\)we substitute \(\,u(r,\theta,t)=U(r,\theta)T(t)\) \(\,\)into the boundary condition of the drumhead to get

    \[\color{red}{u(1,\theta,t)=U(1,\theta)T(t)=0 \;\;\Rightarrow\;\;U(1,\theta)=0}\]

\(~\)

  • Solution of the Helmholtz Eigenvalue Problem

    • To solve

      \[\color{blue}{\begin{aligned} &U_{rr} +\frac{1}{r}U_r +\frac{1}{r^2}U_{\theta\theta} +\lambda U=0\\ &U(1,\theta)=0 \end{aligned}}\]

      we let

      \[\color{red}{U(r,\theta)=R(r)\Theta(\theta)}\]

      and plug it into the Helmholtz BVP

      \[\begin{aligned} U_{rr} +\frac{1}{r}U_r &+\frac{1}{r^2}U_{\theta\theta} +\lambda U=0\\ &\Downarrow U(r,\theta)=R(r)\Theta(\theta)\\ R''\Theta +\frac{1}{r}R' \Theta &+\frac{1}{r^2}R \Theta'' +\lambda R\Theta = 0 \\ &\Downarrow \\ -\frac{r^2 R'' +rR' +\lambda r^2R}{R} &= \frac{\;\Theta''}{\Theta} = -\mu <0 \\ &\Downarrow \\ \color{red}{\Theta'' +\mu \Theta} &\color{red}{=0} \\ \color{red}{r^2 R'' +rR'} &\color{red}{+(\lambda r^2 -\mu) R =0} \end{aligned}\]

      We get a new separation constant \(\,\mu\)

    • It’s obvious that the drumhead is periodic with period \(\,2\pi\,\) in \(\,\theta.\) \(\,\)Thus

      \[\scriptsize\begin{aligned} \Theta'' &+\mu\Theta = 0 \\ &\Downarrow \\ \Theta(\theta) &= a \cos\sqrt{\mu}\theta +b\sin\sqrt{\mu}\theta \\ &\Downarrow \;{\scriptstyle \Theta(0)=\Theta(2\pi),\;\;\Theta'(0)=\Theta'(2\pi)} \\ \begin{bmatrix} 1 -\cos 2\pi\sqrt{\mu} & -\sin 2\pi\sqrt{\mu} \\ \sqrt{\mu} \sin 2\pi\sqrt{\mu} & \sqrt{\mu}\left(1 -\cos 2\pi\sqrt{\mu}\right) \end{bmatrix} &\begin{bmatrix} a \\ b \end{bmatrix} = \begin{bmatrix} 0 \\ 0 \end{bmatrix} \\ \;\;&\overset{\text{non-trivial solution}}{\longrightarrow} \;\; {\tiny \sqrt{\mu}(1 -\cos 2\pi\sqrt{\mu})^2 +\sqrt{\mu} \sin^2 2\pi\sqrt{\mu}=0 \;\rightarrow\;1=\cos 2\pi\sqrt{\mu} } \\ &\Downarrow \\ \color{red}{\mu}\, &\color{red}{= n^2, \;\;n=0,1,2,\cdots} \\ &\Downarrow \\ \color{red}{\Theta_n(\theta)}\, &\color{red}{= a_n \cos n\theta +b_n\sin n\theta} \\ &=\sqrt{a_n^2 +b_n^2} \left(\cos n\theta \cos \phi +\sin n\theta \sin \phi \right),\;\;\tan \phi=\frac{b_n}{a_n} \\ &\color{red}{=c_n \cos (n\theta -\phi)} \end{aligned}\]

    • So, in order to solve the Helmholtz equation, \(~\)we must solve the following differential equation

      \[\color{red}{\begin{aligned} r^2 R_n''\, +\,&rR_n' +(\lambda r^2 -n^2) R_n =0, \;\;\;0<r<1 \\ \left | R_n(0)\right| &< \infty \\ R_n(1)\phantom{|} &= 0 \end{aligned}}\]

      It is the parameterized Bessel’s equation and has two linearly independent solutions \(\color{red}{J_n(\sqrt{\lambda}r)}\) and \(\color{red}{Y_n(\sqrt{\lambda}r)}\)

\(~\)

from scipy.special import jv, yv, jn_zeros

fig = plt.figure(figsize=(6, 8))
ax1 = fig.add_subplot(211)
ax2 = fig.add_subplot(212)

ax1.axis((0, 10, -0.6, 1))    
ax1.set_ylabel('$J_n$')

ax2.axis((0, 10, -3, 1))    
ax2.set_xlabel(r'$\sqrt{\lambda}$')
ax2.set_ylabel('$Y_n$')

ax1.plot([0, 10], [0, 0], 'k:')
ax2.plot([0, 10], [0, 0], 'k:')

x = np.linspace(0, 10, 200)
for n in range(4):
    y1 = jv(n, x)
    y2 = yv(n, x)
    ax1.plot(x, y1, label=rf'$J_{n:d}(\sqrt{{\lambda}})$')
    ax2.plot(x, y2, label=rf'$Y_{n:d}(\sqrt{{\lambda}})$')

ax1.legend(bbox_to_anchor=(1.01, 1), loc='upper left')
ax2.legend(bbox_to_anchor=(1.01, 1), loc='upper left')

plt.show()
Figure 12.1: Bessel’s functions

\(~\)

    • Since the functions \(Y_n(\sqrt{\lambda} r)\) are unbounded at \(r=0\), \(\,\)we choose as our solution \(J_n(\sqrt{\lambda} r)\). \(\,\)The last step in finding \(R_n(r)\) is to use the boundary condition \(R_n(1)=0~\) to find \(\lambda\). \(\,\)We must pick \(\lambda\) to be one of the roots of \(\,\color{red}{J_n(\sqrt{\lambda})=0}\); \(\,\)that is

      \[\color{red}{\sqrt{\lambda_{nm}}=\alpha_{nm}}\]

      where \(\,\alpha_{nm}\) is the \(m\)-th root of \(\,J_n(r)=0\), \(~m=1,2,3,\cdots\). \(\,\)With these roots, \(\,\)we have just solved the Helmholtz eigenvalue problem

    • The eigenvalues \(\,\lambda_{nm}\) are \(\,\alpha_{nm}^2\); \(\,\)and the corresponding eigenfunctions are

      \[\color{red}{U_{nm}(r,\theta) =c_n J_n(\alpha_{nm} r) \cos (n\theta -\phi), \;\;\;n=0,1,2,\cdots, \;\;m=1,2,3,\cdots}\]

      The general shape of \(\,U_{nm}(r,\theta)\,\) is the same for different values of the constants \(\,c_n\,\) and \(\,\phi.\) \(\,\)We plot these functions for the different values of \(\,n\) and \(\,m\) with \(\,c_n=1\,\) and \(\,\phi=0\)

\(~\)

ncols = 4
nrows = 4
    
fig, axs = plt.subplots(ncols=ncols, 
    nrows=ncols, figsize=(7, 7), 
    subplot_kw=dict(projection='polar'))

for ax in axs.flat:
    ax.set(xticks=[], yticks=[])

azimuths = np.radians(np.linspace(0, 360, 100))
zeniths = np.linspace(0, 1, 100)

theta, r = np.meshgrid(azimuths, zeniths)

for n in range(ncols):
    alpha_nm = jn_zeros(n, nrows)
    for m in range(nrows):
        U_nm = jv(n, alpha_nm[m] *r) *np.cos(n *theta)
        axs[n, m].contourf(theta, r, U_nm, levels=[0, 10])
        if m == 0:
            axs[n, 0].set_ylabel(f'n = {n}') 
        if n == 0:
            axs[nrows -1, m].set_xlabel(f'm = {m +1}')
  
plt.show()
Figure 12.2: Helmholtz Eigenfunctions: \(\,U_{nm}(r, \theta)\)

\(~\)

  • Now that we have solved the Helmholtz equation for the basic shapes \(\,U_{nm}(r,\theta)\), \(\,\)the final step is to multiply by the time factor

    \[\color{red}{T_{nm}(t)=A_{nm}\cos c\alpha_{nm}t +B_{nm}\sin c\alpha_{nm} t}\]

    and add the products in such a way that the initial conditions are satisfied. \(\,\)That is, \(\,\)the solution to our problem will be

    \[ \color{red}{u(r,\theta,t) = \sum_{n=0}^\infty \sum_{m=1}^\infty J_n(\alpha_{nm}r) \cos n\theta \left[ A_{nm}\cos c\alpha_{nm}t +B_{nm}\sin c\alpha_{nm}t \right]}\]

    Rather than going through the complicated process of finding \(\,A_{nm}\,\) and \(\,B_{nm}\,\) for the general case, \(\,\)we will find the solution for the situation where \(\,u\,\) is independent of \(\,\theta\) (very common, \(\,n=0\)). \(\,\)In other words, \(\,\)we will assume that the initial position of the drumhead depends only on \(\,r\)

  • In particular, \(\,\)we consider

    \[\begin{aligned} u(r,\theta,0) &= f(\color{red}{r})\\ u_t(r,\theta,0) &= 0 \end{aligned}\]

    (It’s just easy to do the case where \(\,u_t \neq 0\)). \(\,\)With these assumptions, \(\,\)the solution now becomes

    \[ u(r,t) = \sum_{m=1}^\infty A_{\color{red}{0}m} J_{\color{red}{0}}(\alpha_{\color{red}{0}m}r) \cos c\alpha_{\color{red}{0}m}t\]

    and our goal is to find \(\,A_{0m}\,\) so that

    \[ \color{blue}{f(r) = \sum_{m=1}^\infty \color{red}{A_{0m}} J_0(\alpha_{0m}r)} \tag{IC}\label{eq:IC}\]

  • To find the constants \(\,A_{0m}\), \(\,\)we use the orthogonality condition of the Bessel functions:

    \[\int_0^1 rJ_0(\alpha_{0i} r)\,J_0(\alpha_{0j} r) \,dr = \begin{cases} 0 & i \neq j \\ \frac{1}{2}J_1^2(\alpha_{0i}) & i=j \end{cases}\]

    Hence, \(\,\)we multiply each side of equation \(\eqref{eq:IC}\) by \(\,rJ_0(\alpha_{0j}r)\,\) and integrate from \(\,0\,\) to \(\,1\,\), \(\,\)giving

    \[ \int_0^1 rf(r)J_0(\alpha_{0j}r)\,dr = A_{0j} \int_0^1 r J_0^2(\alpha_{0j}r)\,dr \]

    from which we can solve for \(\,A_{0j}\)

    \[ \color{red}{A_{0j}=\frac{2}{J_1^2(\alpha_{0j})} \int_0^1 rf(r)J_0(\alpha_{0j}r)\,dr, \;\;j =1,2,3,\cdots} \]

  • For example \(\,\)the solution to the vibrating drumhead with initial conditions

    \[\begin{aligned} u(r,\theta,0) &= J_0(2.405 r) +0.5 J_0(8.654 r)\\ u_t(r,\theta,0) &= 0 \end{aligned}\]

    would be

    \[ u(r,t) = J_0(2.405 r) \cos(2.405 ct) +0.5 J_0(8.654 r)\cos(8.654 ct) \]

  • We start by drawing \(\,J_0(\sqrt{\lambda})\). \(\,\)Now, \(\,\)in order to graph the functions

    \[J_0(\alpha_{01}r), \,J_0(\alpha_{02}r), \,\cdots,\, J_0(\alpha_{0m}r)\]

    we rescale the \(\,r\)-axis so that \(\,m\)-th root passes through \(\,r=1\)

\(~\)

from mpl_toolkits.axisartist.axislines import SubplotZero

fig = plt.figure(figsize=(7, 8))
ax1 = SubplotZero(fig, 211)
ax2 = SubplotZero(fig, 212)

fig.add_subplot(ax1)
ax1.axis["left", "right", "bottom", "top"].set_visible(False)
ax1.axis["xzero", "yzero"].set_visible(True)
ax1.axis["xzero", "yzero"].set_axisline_style("-|>")

ax1.set_xticks([2, 4, 6, 8, 10])
ax1.set_ylim(-0.5, 1.1)
ax1.set_yticks([-0.5, 0.5, 1.0])
ax1.text(11, -0.03, r'$\sqrt{\lambda}$', fontsize=14)

fig.add_subplot(ax2)
ax2.axis["left", "right", "bottom", "top"].set_visible(False)
ax2.axis["xzero", "yzero"].set_visible(True)
ax2.axis["xzero", "yzero"].set_axisline_style("-|>")

ax2.set_xticks([0.2, 0.4, 0.6, 0.8, 1.0])
ax2.set_ylim(-0.5, 1.1)   
ax2.set_yticks([-0.5, 0.5, 1.0])
ax2.text(1.1, -0.03, '$r$', fontsize=12)

mm = 3
alpha_0m =jn_zeros(0, mm)

r1 = np.linspace(0, 10, 100)
y1 = jv(0, r1)
ax1.plot(r1, y1, label=r'$J_0(\sqrt{\lambda})$')
ax1.plot(alpha_0m, np.zeros(mm), 'ro')

for m in range(mm):
    ax1.text(alpha_0m[m] -0.2, 0.1, rf'$\alpha_{m +1}$', fontsize=12)
ax1.legend()

r2 = np.linspace(0, 1, 100)
for m in range(mm):   
    y2 = jv(0, alpha_0m[m]*r2)
    ax2.plot(r2, y2, label=rf'$J_0(\alpha_{m +1}r)$')
    ax2.plot(alpha_0m[:m +1] /alpha_0m[m], np.zeros(m +1), 'x')
ax2.legend()

plt.show()  

\(~\)

  • In general, \(\,\)what we would do is to expand the initial position \(\,f(r)\,\) into the sum of these basic shapes \(\,J_0(\alpha_{0m} r)\), \(\,\)observe the vibration for each one and then add them up
fig = plt.figure(figsize=(7, 5))
ax = plt.axes(xlim=(0, 1), ylim=(-1.5, 1.5))

ax.set_xticks([0.0, 0.2, 0.4, 0.6, 0.8, 1.0])
ax.set_yticks([-1.5, -1.0, -0.5, 0, 0.5, 1.0, 1.5])
ax.tick_params(pad=10)
ax.set_xlabel('$r$')
ax.set_ylabel('$u(r,t)$')

plt.close()
line1, = ax.plot([], [], lw=1, ls=':')
line2, = ax.plot([], [], lw=1, ls=':')
line_sum, = ax.plot([], [], lw=4)

time_text = ax.text(0.45, 1.3, '')
def init():
    time_text.set_text('t = 0.0')
    line1.set_data([], [])
    line2.set_data([], [])
    line_sum.set_data([], [])
    return (line1, line2, line_sum)

c = 1 
alpha_0m =jn_zeros(0, 3)

def animate(t):
    time_text.set_text(f't = {t:.2f}')        
    r = np.linspace(0, 1, 100)   

    u1 = jv(0, alpha_0m[0] *r) *np.cos(c *alpha_0m[0] *t)
    u2 = 0.5 *jv(0, alpha_0m[2] *r) *np.cos(c *alpha_0m[2] *t)
    u_sum = u1 +u2    
    
    line1.set_data(r, u1)
    line2.set_data(r, u2)
    line_sum.set_data(r, u_sum)
    
    return (line1, line2, line_sum)

tt = np.linspace(0, 10, 100)
anim = animation.FuncAnimation(fig, animate, 
         init_func=init, frames=tt, interval=200, blit=True)

HTML('<center>' + anim.to_html5_video() + '</center>')

\(~\)

Example \(\,\)Solve the vibrating drum head problem

\[ \begin{aligned} u_{tt} &=\nabla^2 u && 0 < r < 1, \;\; 0 < t < \infty \\ \\ u(1,\theta,t) &=0 && 0 < t < \infty \\ \\ \begin{array}{r} u(r,\theta,0) \\ u_t(r,\theta,0) \end{array} & \begin{array}{l} = 1 -r^2 \\ =0 \phantom{-r^2} \end{array} && 0 \leq r \leq 1 \end{aligned}\]

\(~\)

12.9 Dimensionless Problems

  • The basic idea behind dimensionless analysis is that by introducing new (dimensionless) variables in a problem, \(\,\)the problem becomes purely mathematical and contains none of the physical constants that originally characterized it

  • In this way, \(\,\)many different equations in physics, biology, engineering, \(\,\)and chemistry that contain special nuances via physical parameters are all transformed into the same simple form

  • Converting a Diffusion Problem to Dimensionless Form

    • Suppose we start with the initial-boundary-value problem:

      \[ \begin{aligned} u_t &=\alpha u_{xx} && 0 < x < L, \;\; 0 < t < \infty \\[5pt] \begin{array}{r} u(0,t) \\ u(L,t) \end{array} & \begin{array}{l} = T_1 \\ = T_2 \end{array} && 0 < t < \infty \\[5pt] u(x,0) &= \sin \frac{\pi x}{L} && 0 \leq x \leq L \end{aligned}\]

      Our goal is to change this problem to a new equivalent formulation that has the properties

        1. No physical parameters (like \(\alpha\)) in the new equation
        2. The initial and boundary conditions are simpler

      \(~\)

    • Transforming the Dependent Variable \(~u \to U =\displaystyle\frac{u(x,t) -T_1}{T_2 -T_1}\)

      \[ \begin{aligned} U_t &=\alpha U_{xx} && 0 < x < L, \;\; 0 < t < \infty \\[5pt] \begin{array}{r} U(0,t) \\ U(L,t) \end{array} & \begin{array}{l} = 0 \\ = 1 \end{array} && 0 < t < \infty \\[5pt] U(x,0) &= \frac{\displaystyle \sin \frac{\pi x}{L} -T_1}{T_2 -T_1} && 0 \leq x \leq L \end{aligned}\]

      \(~\)

    • Transforming the Space Variable \(~x \to \xi=\;^{\displaystyle x}/_{\displaystyle L}\)

      \[ \begin{aligned} U_t &=\frac{\alpha}{L^2} U_{\xi\xi} && 0 < \xi < 1, \; 0 < t < \infty \\[5pt] \begin{array}{r} U(0,t) \\ U(L,t) \end{array} & \begin{array}{l} = 0 \\ = 1 \end{array} && 0 < t < \infty \\[5pt] U(\xi,0) &= \frac{\sin\pi\xi -T_1}{T_2 -T_1} && 0 \leq \xi \leq 1 \end{aligned}\]

      \(~\)

    • Transforming the Time Variable \(\;\; t \to \tau\)

      Since our goal is to eliminate the constant \(\displaystyle\frac{\alpha}{L^2}\) from the PDE, \(\,\)we proceed as follows:

      1. Try a transformation of the form \(\tau=ct\), \(\,\)where \(c\) is an unknown constant

      2. Compute \(U_t=U_\tau\tau_t=cU_\tau\)

      3. Substitute this derivative into the the PDE to obtain

        \[ cU_\tau =\frac{\alpha}{L^2} U_{\xi\xi}\]

        and, \(\,\)hence, \(\,\)pick \(\displaystyle c=\frac{\alpha}{L^2}\).

      This give us our new time

      \[ \tau=\frac{\alpha}{L^2} t \]

    • We now have the completely dimensionless problem

      \[ \begin{aligned} U_\tau &= U_{\xi\xi} && 0 < \xi < 1, \; 0 < \tau < \infty \\[5pt] \begin{array}{r} U(0,\tau) \\ U(L,\tau) \end{array} & \begin{array}{l} = 0 \\ = 1 \end{array} && 0 < t < \infty \\[5pt] U(\xi,0) &= \frac{\sin\pi\xi -T_1}{T_2 -T_1} && 0 \leq \xi \leq 1 \end{aligned}\]

\(~\)

  • Transforming a Hyperbolic Problem to Dimensionless Form

    Consider the vibrating string

    \[ \begin{aligned} u_{tt} &=c^2 u_{xx} && 0 < x < L,\;\; 0 < t < \infty \\[5pt] \begin{array}{r} u(0,t) \\ u(L,t) \end{array} & \begin{array}{l} = 0 \\ = 0 \end{array} && 0 < t < \infty \\[5pt] \begin{array}{r} u(x,0) \\ u_t(x,0) \end{array} & \begin{array}{l} = \sin \frac{\pi x}{L} +0.5\sin \frac{3\pi x}{L} \\ = 0 \end{array} && 0 < x < L \end{aligned}\]

    • By transforming the independent variables (no need to transform \(u\)) into a new ones

      \[ \xi=\frac{x}{L}\; \text{ and } \;\tau=\frac{c}{L} t \]

      we get the new problem

      \[ \begin{aligned} u_{\tau\tau} & =u_{\xi\xi} && 0 < \xi < 1,\;\; 0 < \tau < \infty \\[5pt] \begin{array}{r} u(0,\tau) \\ u(1,\tau) \end{array} & \begin{array}{l} = 0 \\ = 0 \end{array} && 0 < \tau < \infty \\[5pt] \begin{array}{r} u(\xi,0) \\ u_\tau(\xi,0) \end{array} & \begin{array}{l} = \sin \pi\xi +0.5\sin 3\pi\xi \\ = 0 \end{array} && 0 < \xi < 1 \\ \end{aligned}\]

      which has the solution

      \[ u(\xi,\tau)=\cos \pi \tau \sin\pi\xi +0.5 \cos 3\pi \tau \sin 3\pi \xi \]

\(~\)

Example \(\,\)How could you pick a new space variable \(\,\xi\,\) so that \(\,v\) is eliminated in the equation

\[ u_t +vu_x=0 \]

\(~\)

12.10 \(~\)Classification of PDEs (Cannonical Form of the Hyperbolic Equation)

  • The purpose here is to classify the second-order linear PDE

    \[ \color{red}{A}u_{xx} +\color{red}{B}u_{xy} +\color{red}{C} u_{yy} +Du_x +Eu_y +Fu = G \tag{SL}\label{eq:SL}\]

    in which \(A, B, C, D, E, F,\) and \(\,G\) are, in general, functions of \(x\) and \(y\) in two independent variables and could be constants, as

    • Hyperbolic if \(\,B^2 -4AC > 0~\) at \(\,(x,y)\)

    • Parabolic \(~\,\) if \(\,B^2 -4AC = 0~\) at \(\,(x,y)\)

    • Elliptic \(~~~~~\,\) if \(\,B^2 -4AC < 0~\) at \(\,(x,y)\)

  • and depending on which is true, \(\,\)to transform the equation into a corresponding cannonical (simple) form by introducing new coordinates \(\,\xi=\xi(x,y)\) \(\,\)and \(\,\eta=\eta(x,y)\)

    \[\begin{aligned} u_{\xi\eta} = \Phi(\xi,\eta,u,u_\xi,u_\eta) \; &\text{ or } \; u_{\xi\xi} -u_{\eta\eta} = \Psi(\xi,\eta,u,u_\xi,u_\eta) \\ u_{\eta\eta} &= \Phi(\xi,\eta,u,u_\xi,u_\eta)\\ u_{\xi\xi} +u_{\eta\eta} &= \Phi(\xi,\eta,u,u_\xi,u_\eta) \end{aligned}\]

    The student should note that whether equation \(\eqref{eq:SL}\) is hyperbolic, parabolic, or elliptic depends only on the coefficients of the second derivatives. \(\,\)We now come to the major portion of this section; \(\,\)rewriting hyperbolic equations in their canonical form

  • The Canonical Form of the Hyperbolic Equation

    The objective here is to introduce new coordinates \(\,\xi=\xi(x,y)\,\) and \(\,\eta=\eta(x,y)\,\) so that the general PDE contains only one second derivative \(\,u_{\xi\eta}\,\). \(\,\)First of all, \(\,\)we compute the partial derivatives

    \[\begin{aligned} Au_{xx} +Bu_{xy} +C u_{yy} &+Du_x +Eu_y +Fu = G \\ &\Downarrow\;\; \color{red}{\xi=\xi(x,y), \;\eta=\eta(x,y)} \\ u_x =\; u_\xi &\xi_x +u_\eta \eta_x \\ u_y =\; u_\xi &\xi_y +u_\eta \eta_y \\ u_{xx} =\; u_{\xi\xi}& \xi_x^2 +2u_{\xi\eta}\xi_x\eta_x +u_{\eta\eta}\eta_x^2 +u_\xi\xi_{xx} +u_\eta\eta_{xx}\\ u_{xy} =\;u_{\xi\xi}& \xi_x\xi_y +u_{\xi\eta}(\xi_x\eta_y +\xi_y\eta_x) +u_{\eta\eta}\eta_x\eta_y +u_\xi\xi_{xy} +u_\eta\eta_{xy}\\ u_{yy} =\;u_{\xi\xi}& \xi_y^2 +2u_{\xi\eta}\xi_y\eta_y +u_{\eta\eta}\eta_y^2 +u_\xi\xi_{yy} +u_\eta\eta_{yy}\\[5pt] &\Downarrow \\ \end{aligned}\]

    \[ \begin{aligned} {\color{red}{\bar{A}}}u_{\xi\xi} +{\color{red}{\bar{B}}}u_{\xi\eta} +&{\color{red}{\bar{C}} u_{\eta\eta}} +\bar{D}u_\xi +\bar{E}u_\eta +\bar{F}u = \bar{G} \\[5pt] \text{in which} & \\[5pt] \bar{A}=&\;A\xi_x^2 +B\xi_x \xi_y +C\xi_y^2 \\ \bar{B}=&\;2A\xi_x\eta_x +B(\xi_x \eta_y+\xi_y\eta_x) +2C\xi_y\eta_y \\ \bar{C}=&\;A\eta_x^2 +B\eta_x \eta_y +C\eta_y^2 \\ % \bar{D}=&\;A\xi_{xx} +B\xi_{xy} % +C\xi_{yy} +D\xi_x +E\xi_y \\ % \bar{E}=&\;A\eta_{xx} +B\eta_{xy} % +C\eta_{yy} +D\eta_x +E\eta_y \\ % \bar{F}=&\;F\\ % \bar{G}=&\;G\\ \Downarrow& \\ \begin{pmatrix} 2\bar{A}& \bar{B}\\ \bar{B}& 2\bar{C} \end{pmatrix} =& \; \begin{pmatrix} \xi_x & \xi_y\\ \eta_x & \eta_y \end{pmatrix} \begin{pmatrix} 2A & B\\ B & 2C \end{pmatrix} \begin{pmatrix} \xi_x & \xi_y\\ \eta_x & \eta_y \end{pmatrix}^T \\ \Downarrow& \\ \bar{B}^2 -4\bar{A}\bar{C} =& \; (\xi_x\eta_y-\xi_y\eta_x)^2(B^2 -4AC) \\ =& \; J^2(B^2 -4AC) \end{aligned}\]

    in which \(\,J\,\) is the Jacobian of the transformation and we select the transformation \(\,(\xi,\eta)\,\) such that \(\,J \neq 0\)

  • The next step in our process is to set the coefficients \(\,\bar{A}\,\) and \(\,\bar{C}\,\) to zero and solve for the tranformation \(\,\xi\,\) and \(\,\eta\,\). \(\,\)This will give us the coordinates that reduce the original PDE to canonical form:

    \[\begin{aligned} \bar{A}=&\;A\xi_x^2 +B\xi_x \xi_y +C\xi_y^2 =0\\ \bar{C}=&\;A\eta_x^2 +B\eta_x \eta_y +C\eta_y^2 =0\\ &\Downarrow \\ A\left[ \frac{\xi_x}{\xi_y} \right]^2 &+B\left[ \frac{\xi_x}{\xi_y} \right] +C=0,\;\; A\left[ \frac{\eta_x}{\eta_y} \right]^2 +B\left[ \frac{\eta_x}{\eta_y} \right] +C=0 \\ &\Downarrow \\ \left[ \frac{\xi_x}{\xi_y} \right] =&\frac{-B +\sqrt{\color{blue}{B^2-4AC}}}{2A},\;\; \left[ \frac{\eta_x}{\eta_y} \right] =\frac{-B -\sqrt{\color{blue}{B^2-4AC}}}{2A} \end{aligned}\]

  • We have now reduced the problem to finding the two functions \(\,\xi(x,y)\,\) and \(\,\eta(x,y)\,\) so that their ratio \(\,[\xi_x/\xi_y]\,\) and \(\,[\eta_x/\eta_y]\,\) satisfy the above equation. \(\,\)Finding functions satisfying these conditions is really quite easy once we look for a few moments at figure

    \[\begin{aligned} &\big\Downarrow \;\;{\scriptsize \xi = \text{constant}, \eta = \text{constant}}\\ \frac{dy}{dx} &=-\left[ \frac{\xi_x}{\xi_y} \right]=\frac{B -\sqrt{B^2-4AC}}{2A} =g_1(x,y)\\ \frac{dy}{dx} &=-\left[ \frac{\eta_x}{\eta_y} \right]=\frac{B +\sqrt{B^2-4AC}}{2A} =g_2(x,y)\\ &\Downarrow \;\;{\scriptsize \text{integration and arrangement}}\\ G_1(&x,y) = c_1 \\ G_2(&x,y) = c_2 \\ &\Downarrow \\ \xi &=G_1(x,y) \\ \eta &=G_2(x,y) \end{aligned}\]

\(~\)

Example \(\,\)Consider the simple equation

\[u_{xx} -4u_{yy} +u_x =0,\;\;\;B^2 -4AC=16>0\]

whose characteristic equations are

\[\begin{aligned} \frac{dy}{dx} &=-\left[ \frac{\xi_x}{\xi_y} \right]=\frac{B -\sqrt{B^2-4AC}}{2A} =-2\\ \frac{dy}{dx} &=-\left[ \frac{\eta_x}{\eta_y} \right]=\frac{B +\sqrt{B^2-4AC}}{2A} =2\\ \end{aligned}\]

To find \(\xi\) and \(\eta\), \(\,\)we first integrate for \(x\) and solve for the constants \(c_1\) and \(c_2\), \(\,\)getting

\[\begin{array}{l} y =-2x+c_1 \\ y = 2x+c_2 \\ \end{array} \;\;\Rightarrow\;\; \begin{array}{l} \xi =y+2x=c_1 \\ \eta =y-2x=c_2 \end{array}\]

\(~\)

Example \(\,\)Suppose we start with the equation

\[y^2u_{xx} -x^2 u_{yy}=0,\;\;\;x>0,\;y>0\]

which is a hyperbolic equation in the first quadrant. \(\,\)We consider the problem of finding new coordinates that will change the original equation to canonical form for \(\,x\,\) and \(\,y\,\) in the first quadrant

STEP 1 \(\,\) Solve the two characteristic equations

\[\begin{aligned} \frac{dy}{dx} &=-\left[ \frac{\xi_x}{\xi_y} \right]=\frac{B -\sqrt{B^2-4AC}}{2A} =-\frac{x}{y}\\ \frac{dy}{dx} &=-\left[ \frac{\eta_x}{\eta_y} \right]=\frac{B +\sqrt{B^2-4AC}}{2A} =\frac{x}{y}\\ \end{aligned}\]

Integrating these two equations by the ODE technique of separating variables gives the implicit relationship

\[\begin{aligned} y^2 &-x^2 = c_1 \\ y^2 &+x^2 = c_2 \\ &\Downarrow \\ \xi= \,&y^2 -x^2 \\ \eta = \,&y^2 +x^2 \end{aligned}\]

We substitute this new coordinate into the equation to find the new equation

\[\begin{aligned} y^2 u_{xx} &-x^2 u_{yy}= 0\\ &\Downarrow\;\; \xi=y^2 -x^2,\; \eta=y^2 +x^2 \\ -16x^2y^2 u_{\xi\eta} -2(x^2 +y^2) &u_\xi +2(y^2 -x^2) u_\eta=0 \end{aligned}\]

STEP 2 \(\,\) Finally, \(\,\)solving for \(\,x\,\) and \(\,y\,\) in terms of \(\,\xi\,\) and \(\,\eta\), \(\,\)we get

\[ u_{\xi\eta}=\frac{\eta u_\xi -\xi u_\eta}{2(\xi^2 -\eta^2)}\]

\(~\)

Example \(\,\)Find the new characteristic coordinates for

\[u_{xx} + 4u_{xy} = 0\]

Solve the transformed equation in the new coordinate system and then transform back to the original coordinates to find the solution to the original problem

\(~\)

NOTES

  • The general hyperbolic equation actually has two canonical forms; \(\,\)the other one can be found by making yet another transformation

    \[\begin{aligned} \alpha &=\xi +\eta \\ \beta &=\xi -\eta \end{aligned}\]

    and rewriting the first canonical form in terms of \(\,\alpha\,\) and \(\,\beta\). \(\,\)The partial derivatives are transformed as follows:

    \[\begin{aligned} u_\xi&= u_\alpha \alpha_\xi +u_\beta \beta_\xi=u_\alpha +u_\beta\\ u_\eta&= u_\alpha \alpha_\eta +u_\beta \beta_\eta=u_\alpha -u_\beta\\ u_{\xi\eta}&= u_{\alpha\alpha}\alpha_\eta +u_{\alpha\beta}\beta_\eta +u_{\beta\alpha}\alpha_\eta +u_{\beta\beta} \beta_\eta =u_{\alpha\alpha} -u_{\beta\beta} \end{aligned}\]

  • In the general parabolic equation, \(\,\)we obtain only a single canonical transformation as

    \[ \bar{A}=0, \;\;\frac{dy}{dx}=-\left [ \frac{\xi_x}{\xi_y} \right ],\;\;\;\text{or}\;\;\; \bar{C}=0, \;\;\frac{dy}{dx}=-\left [ \frac{\eta_x}{\eta_y} \right ]\]

    Suppose \(\,\xi\,\) is calculated. \(\,\)What could be the possible choice for \(\,\eta\,\)? \(\,\)We can select \(\,\eta\,\) as any arbitrary function of \(\,x\,\) and \(\,y\,\) such that \(\,\eta\,\) is independent of \(\,\xi\,\) (\(J \neq 0\))

  • In the general elliptic case, \(\,A\lambda^2 +B\lambda +C=0\,\) leads to complex conjugate canonical tramsformation \(\,\xi\,\) and \(\,\eta\). \(\,\)Since \(\,\xi\,\) and \(\,\eta\,\) are complex, \(\,\)we introduce new real variables

    \[\begin{aligned} \alpha &= \frac{1}{2}(\xi +\eta)\\ \beta &= \frac{1}{2i}(\xi -i\eta) \end{aligned}\]

    Under the transformation \(\,(x, y)\to (\alpha, \beta)\), \(\,\)the canonical form is given by

    \[ u_{\alpha\alpha} +u_{\beta\beta}=\Phi(\alpha,\beta,u,u_\alpha,u_\beta)\]

  • The three major classifications of linear PDEs as hyperbolic, parabolic, and elliptic equations essentially classify physical problems into three basic physical types:

    • Wave propagation
    • Diffusion
    • Steady-state problems

    The mathematical solutions of these three types of equations are quite different

\(~\)

12.11 \(~\)First-Order Equations (Method of Characteristics)

  • If the student recalls, \(\,\)when we solved the diffusion equation

    \[ u_t=\alpha u_{xx} - vu_x, \quad -\infty < x < \infty, \;\; 0 < t < \infty\]

    the constant \(\,\alpha\,\) stood for the amount of diffusion, \(\,\)while \(v\) stood for the velocity of the medium

  • Hence if \(\,\alpha=0\) (no diffusion), \(\,\)it is clear (since we only have convection) that the solution will travel along the \(\,x\)-axis with velocity \(\,v\). \(\,\)In other words, \(\,\)if the initial solution is \(\,u(x,0)=\phi(x)\), \(\,\)then the solution to

    \[u_t=-vu_x\]

    would be \(\,u(x,t)=\phi(x-vt)\)

  • So we can think of the first-order equation

    \[a(x,t)u_x +b(x,t)u_t = 0\]

    as the concentration along a stream where the velocity is given by

    \[ v=\frac{a(x,t)}{b(x,t)}\]

  • The purpose of this section is to introduce the notion of first-order partial differential equations and method of characteristics. \(\,\)The problem we will solve is the initial-value problem

    \[ a(x,t) u_x +b(x,t) u_t +c(x,t)u=0 \]

    \[u(x,0)=\phi(x), \;\; {-\infty<x<\infty,\;\; 0< t < \infty}\]

    The solution to this equation is based on a physical fact, \(\,\)that an initial disturbance at some point \(\,x\,\) propagates along a line (curve) in the \(\,tx\)-plane (called a characteristic)

  • With this in mind, \(\,\)the idea is to introduce two new coordinates \(\,s\,\) and \(\,\tau\) (to replace \(\,x\,\) and \(\,t\)) that have the properties

    • \(s\,\) will change along the characteristic curves
    • \(\tau\,\) will change along the initial curve (most likely the line \(\,t=0\))

    \[\begin{aligned} a(x,t) u_x +\, & b(x,t) u_t +c(x,t)u = 0\\ &\big\Downarrow\;\; {\scriptsize\text{characteristic curves } \{ \left( x(s), t(s) \right) : 0<s<\infty \}} \\ \frac{dx}{ds} = a(x,t), &\; \frac{dt}{ds} = b(x,t) \\ &\Downarrow \\ \frac{du}{ds}= u_x \frac{dx}{ds} +u_t\frac{dt}{ds}&=a(x,t) u_x +b(x,t) u_t \\ &\Downarrow \\ \frac{du}{ds} +\, &c(x,t)u =0 \end{aligned}\]

\(~\)

Example \(\,\)Solve the following IVP with constant coefficients:

\[ \begin{aligned} u_x +u_t +2u &=0 && -\infty < x < \infty,\;\; 0 < t < \infty \\ u(x,0) &=\sin x && -\infty < x < \infty \end{aligned}\]

STEP 1 \(\,\) Find the characteristics (along which the initial data propagate)

\[\begin{aligned} \frac{dx}{ds}=1,\;\; \frac{dt} {ds}&=1\quad 0<s<\infty \\ &\Downarrow \\ x(s) = &\;s +c_1 \\ t(s) = &\;s +c_2 \\ &\Downarrow \;\;x(0)= \tau, \;\;t(0)=0 \\ x = &\;s +\tau \\ t = &\;s \\ &\Downarrow \\ x - t = \tau, \;\;\;&-\infty < \tau < \infty \end{aligned}\]

STEP 2 \(\,\) Using this coordinate system, \(\,\)we change the PDE to the ODE

\[\begin{aligned} \frac{du}{ds} +2u&= 0, \;\; u(0)= \sin\tau \\ &\Downarrow \\ u(s,\tau) &=\sin \tau \cdot e^{-2s} \end{aligned}\]

STEP 3 \(\,\) We now solve the transformation for \(\,s\,\) and \(\,\tau\,\) in terms of \(\,x\,\) and \(\,t\):

\[u(x,t) = \sin(x -t) \cdot e^{-2t}\]

\(~\)

Example \(\,\)Solve the following IVP with variable coefficients:

\[ \begin{aligned} x u_x +u_t +tu &=0 && -\infty < x < \infty,\; 0 < t < \infty \\ u(x,0) &=F(x) && -\infty < x < \infty, \;\;\text{an arbitrary initial wave} \end{aligned}\]

STEP 1 \(\,\) Find the characteristics (along which the initial data propagate)

\[\begin{aligned} \frac{dx}{ds} =x,\;\;&\frac{dt}{ds} =1 \quad 0<s<\infty \\ &\Downarrow \\ x(s) &=\; c_1 e^s\\ t(s) &=\; s +c_2 \\ &\Downarrow \;\; x(0)= \tau, \;\;t(0)=0 \\ x &=\;\tau e^s \\ t &=\;s \end{aligned}\]

STEP 2 \(\,\) Using this coordinate system, \(\,\)we change the PDE to the ODE

\[\begin{aligned} \frac{du}{ds} +su&= 0, \;\; u(0)= F(\tau) \\ &\Downarrow \\ u(s,\tau) &=F(\tau) \cdot e^{-s^2/2} \end{aligned}\]

STEP 3 \(\,\) We now solve the transformation for \(\,s\,\) and \(\,\tau\,\) in terms of \(\,x\,\) and \(\,t\):

\[u(x,t) = F(xe^{-t}) \cdot e^{-t^2/2}\]

\(~\)

Example \(\,\) Solve the problem in higher dimensions (surface waves)

\[ \begin{aligned} a u_x +b u_y +c u_t +du &=0 && \scriptsize -\infty < x < \infty, \;-\infty < y < \infty, \; 0< t < \infty \\ u(x,y,0) &=e^{-(x^2 +y^2)} && \scriptsize -\infty < x < \infty, \;-\infty < y < \infty \end{aligned}\]

\(~\)

12.12 \(~\)Nonlinear First-Order Equations (Conservation Equations)

  • One of the most useful partial differential equations in mathematics is the conservation equation

    \[u_t +f_x=0\]

    This equation says that the increase of a physical quantity \(\,u_t\,\) is equal to the change in flux \(\,-f_x\,\) of that quantity across the boundary

  • Derivation of the Conservation Equation

    Suppose we have a stretch of highway on which cars are moving from left to right, \(\,\)and we assume there are no exit and entrance ramps. \(\,\)We let

    \[ \begin{aligned} u(x,t) &=\text{density of cars at } x\;\; (\text{cars per unit length at } x) \\[5pt] f(x,t) &=\text{flux of cars at } x \;\; (\text{cars per minute passing the point } x) \end{aligned}\]

    Then, \(\,\)it is fairly obvious that for a road segment \(\,[a,b]\), \(\,\)the change in the number of cars (with respect to time) is given by both of the following expressions:

    \[\begin{aligned} \text{change in the number of cars in } [a,b] &= \frac{d}{dt} \int_a^b u(x,t)\,dx \\ \text{change in the number of cars in } [a,b] &= f(a,t) -f(b,t) = -\int_a^b \frac{\partial f}{\partial x}(x,t)\,dx \end{aligned}\]

    Setting these two integrals equal to each other, \(\,\)we have

    \[ \int_a^b \frac{\partial u}{\partial t}(x,t)\,dx = -\int_a^b \frac{\partial f}{\partial x}(x,t)\,dx\]

    Finally, \(\,\)since the integral \(\,[a,b]\,\) was arbitrary, \(\,\)we have the one-dimensional conservation equation

    \[u_t +f_x=0\]

    If we carried out a similar analysis in two or three dimensions, \(\,\)we would have

    \[u_t +f_x +f_y=0, \;\; u_t +f_x +f_y +f_z=0\]

\(~\)

  • Conservation Equation Applied to the Traffic Problem

    • We are now ready to see how the conservation equation \(\,u_t +f_x=0\,\) can be applied to the traffic problem

    • To do this, \(\,\)we rewrite the change in flux \(\,f_x\,\) as

      \[f_x = \frac{df}{du} u_x,\;\;\text{chain rule}\]

      and substitute it in the conservation equation. \(\,\)Doing this, \(\,\)we arrive at

      \[ u_t +\frac{df}{du}u_x=u_t+g(u)u_x=0\]

      where we can imagine a car moving upstream or downstream with velocity \(\,g(u)\)

    • Hence, after \(\,t\,\) minutes, the position \(\,x\,\) of the car will be

      \[x = x_0 +g(u)t, \quad\text{characteristic equation}\]

    • Hence, after \(\,t\,\) minutes, the position \(\,x\,\) of the car will be

      \[x = x_0 +g(u)t, \quad\text{characteristic equation}\]

    • Remembering that the concentration \(\,u(x,t)\,\) will not change along each characteristic, \(\,\)if we know the initial density \(\,u(x_0,0)\), then the characteristic equation for \(\,x\,\) can be written

      \[x=x_0+g[u(x_0,0)]t\]

    • It is clear that by knowing these characteristic curves starting from each point (and knowing the solution doesn’t change along these curves), \(\,\)we can piece together the solution \(\,u(x,t)\,\) for all time \(\,t\). \(\,\)We won’t actually find the explicit equation for the solution \(\,u(x,t)\,\) in terms of \(\,x\,\) and \(\,t\,\) but will use our knowledge of the characteristics of the equation to solve some interesting problems

\(~\)

  • Traffic-Flow Problem

    • We will now solve a traffic-flow problem where the flux is given by \(\,f(u)=u^2\) and the initial density of cars is given. \(\,\)In other words, \(\,\)we have

    \[ u_t +2uu_x =0 \quad -\infty < x < \infty, \; 0< t < \infty\]

    \[u(x,0) = \begin{cases} \;\;\; 1 & x \leq 0\\ 1 -x & 0 < x < 1, \;-\infty < x <\infty \\ \;\;\; 0 & 1 \leq x \end{cases}\]

STEP 1

We begin by finding the characteristics starting from each initial point \(\,(x_0,0)\). \(\,\)For \(\,x_0 < 0\), \(\,\)the characteristics are

\[\begin{aligned} x &= x_0 +g[u(x_0,0)] t \\ &= x_0 +g[1] t\\ &= x_0 +2t \\ &\Downarrow \\ t &=\frac{1}{2} (x -x_0) \end{aligned}\]

Now consider those initial points \(\,0 < x_0 < 1\). \(\,\)Here, \(\,\)the characteristic curves are

\[\begin{aligned} x &= x_0 +g[u(x_0,0)] t \\ &= x_0 +g[1 -x_0] t\\ &= x_0 +2(1 -x_0)t \\ &\Downarrow \\ t &=\frac{x -x_0}{2(1 -x_0)} \end{aligned}\]

Finally, \(\,\)for \(\,1 \leq x_0 < \infty\), \(\,\)we have the characteristics

\[\begin{aligned} x &= x_0 +g[u(x_0,0)] t \\ &= x_0 +g[0] t\\ &= x_0 \end{aligned}\]

which represents vertical lines starting at \(\,x_0\)

The solution remains constant along each characteristic line. \(\,\)Now the solution for \(\,x \in \mathbb{R}\), \(\,0 \leq t <\textstyle\frac{1}{2}\,\) is given by

\[u(x,t)= \begin{cases} \quad 1 & \; x \leq 2t,\qquad 0 \leq t < \frac{1}{2}\\ \displaystyle\frac{1 -x}{1 -2t} & 2t \leq x < 1, \;\, 0 \leq t < \frac{1}{2} \\ \quad 0 & \; 1 \leq x, \qquad \; 0 \leq t < \frac{1}{2} \end{cases}\]

The difficulty in this example is that some characteristic lines will intersect each other. \(\,\)In other words, \(\,\)at the intersection point \(\,(1,\textstyle \frac{1}{2})\,\) of the two characteristic lines, \(\,\)the solution becomes multi-valued. \(\,\)This pathology means that the solution in the usual sense fails to exist after \(\,t \geq \textstyle \frac{1}{2}\)

STEP 2

  • Hence, \(\,\)to find the solution past \(\,t=\textstyle \frac{1}{2}\), \(\,\)we must use another type of analysis. \(\,\)When characteristics run together, \(\,\)we have the phenomenon of shock waves (discontinuous solutions), \(\,\)and what we must ask is, \(\,\)how fast does the leading edge of the shork wave propagate along the road?

  • Before proceeding further, \(\,\)it is necessary to introduce the concept of a jump condition: a condition that holds at a discontinuity or abrupt change. \(\,\)Consider one dimensional problem where there is a jump in the conserved quantity \(\,u\):

    \[u_t +f_x=0\]

    Let the solution exhibit a jump (or shork) at \(\,x=x_s(t)\,\) and integrate over the partial domain, \(\,x \in (x_1, x_2)\), \(\,\)where \(\,x_1 < x_s(t) < x_2\),

    \[\scriptsize \begin{aligned} \frac{d}{dt} \left( \int_{x_1}^{x_s(t)}u\,dx +\int_{x_s(t)}^{x_2}u \,dx\right)&= -\int_{x_1}^{x_2} \frac{\partial } {\partial x} f(u)\,dx\\ &\Downarrow \\ \int_{x_1}^{x_s(t)} u_t \,dx +u^-(x_s(t),t)\frac{dx_s}{dt} +\int_{x_s(t)}^{x_2} u_t \,dx &-u^+(x_s(t),t)\frac{dx_s}{dt}= -\left. \underset{}{f(u)}\, \right|_{x_1}^{x_2}\\ x_1 \to x_s(t),\;x_2 \to x_s(t) \;\; &\Downarrow \\ \left[ u^-(x_s(t),t) -u^+(x_s(t),t) \right] \frac{d x_s}{dt} &= f(u^-(x_s(t),t)) -f(u^+(x_s(t),t)) \\ &\Downarrow \\ s=\frac{d x_s}{dt} &=\frac{f(u^-(x_s(t),t)) -f(u^+(x_s(t),t))}{ u^-(x_s(t),t) -u^+(x_s(t),t)} \end{aligned}\]

    in which \(\,s\,\) is the speed of the discontinuity and represents the jump condition for conservation eqaution

  • A shork situation arises in a system where its characteristics intersect, \(\,\)and under these conditions a requirement for a unique single-valued solution is that the solution should satisfy the Lax entropy condition

    \[f'(u^+(x_s(t),t)) < s < f'(u^-(x_s(t),t))\]

    where \(f'(u^+(x_s(t),t))\,\) and \(\,f'(u^-(x_s(t),t))\,\) represent characteristic speeds at upstream and downstream conditions, respectively

  • So, \(\,\)in our example, \(\,\)the speed of the wave front would be(remembering that \(\,f(u)=u^2\))

    \[ s = \frac{1 -0}{1 -0}=1\]

  • Therefore, \(\,\)a shock is formed at \((x,t)=(1,\textstyle\frac{1}{2})\):

    • After that, \(\,\)it moves with constant speed \(1\), \(\,\)along the line \(\,x=t +\textstyle \frac{1}{2}\), \(\,\)for \(\,t \geq \textstyle \frac{1}{2}\).

    • In conclusion, \(\,\)the solution takes different forms in the three different regions in the space-time plane

\[u(x,t)= \begin{cases} \quad 1 & x \leq 2t, \qquad\; 0 \leq t < \frac{1}{2} \;\;\text{or}\;\; x-\frac{1}{2} \leq t, \;\;t \geq \frac{1}{2}\\ \displaystyle\frac{1 -x}{1 -2t} & 2t \leq x < 1, \;\, 0 \leq t < \frac{1}{2} \\ \quad 0 & x \geq 1, \qquad\;\; 0 \leq t < \frac{1}{2} \;\;\text{or}\;\; x-\frac{1}{2} > t, \;\;t \geq \frac{1}{2} \end{cases}\]

\(~\)

Example \(\,\) Solve

\[ \begin{aligned} u_t +uu_x &=0 \qquad -\infty < x < \infty, \; 0 < t < \infty \\ u(x,0) &= \begin{cases} 3(1 -|x|) & |x| \leq 1\\ \quad\; 0 & |x| > 1 \end{cases} \end{aligned}\]

\(~\)

Example \(\,\) Verify that \(\,u=\phi[x-g(u)t]\,\) is an implicit solution of the nonlinear problem

\[ \begin{aligned} u_t +g(u)u_x &=0 \quad -\infty < x < \infty, \; 0 < t < \infty \\ u(x,0) &= \phi(x) \end{aligned}\]

\(~\)

12.13 \(~\)Systems of PDEs

  • In many areas of science, \(\,\)several unknown quantities (and their derivatives) are modeled by a system of interlocking equations. \(\,\)And we seek to find all of these functions simultaneously

  • There are other reasons for studying systems of equations. \(\,\)Although things are not quite so simple in PDE theory, \(\,\)it is often possible to write a single higher-order PDE as a system of first-order equations

  • For example, \(\,\)the telegraph equation

    \[ u_{tt}=c^2u_{xx} +au_t +bu \]

    can be rewritten in equivalent form as three first-order PDEs by introducing the three variables:

    \[\begin{aligned} u_1 &= u \\ u_2 &= u_x \\ u_3 &= u_t \end{aligned}\]

  • It is a simple matter to see that the telegraph equation is equivalent to the three equations:

    \[\begin{aligned} \frac{\partial u_1}{\partial x} &= u_2 \\[2pt] \frac{\partial u_2}{\partial t} &= u_3 \\[2pt] \frac{\partial u_3}{\partial t} &= c^2 \frac{\partial u_2}{\partial x} +au_3 +bu_1 \end{aligned}\]

  • Solution of the Linear System \(\,\mathbf{u}_t +\mathbf{A}\mathbf{u}_x=\mathbf{0}\)

    • Consider the initial-value problem consisting of two PDEs and two ICs

      \[\begin{aligned} \frac{\partial u_1}{\partial t} +8 &\frac{\partial u_2}{\partial x} = 0 \\[2pt] \frac{\partial u_2}{\partial t} +2 &\frac{\partial u_1}{\partial x} = 0 \\[5pt] u_1(x,0)& = f(x) \\ u_2(x,0)& = g(x) \end{aligned}\]

      We start by writing the two PDEs in matrix form

      \[\begin{aligned} \begin{bmatrix} \frac{\partial u_1}{\partial t} \\ \frac{\partial u_2}{\partial t} \end{bmatrix} +&\begin{bmatrix} 0 & 8 \\ 2 & 0 \end{bmatrix} \begin{bmatrix} \frac{\partial u_1}{\partial x} \\ \frac{\partial u_2}{\partial x} \end{bmatrix} = \begin{bmatrix} 0 \\ 0 \end{bmatrix} \\ \end{aligned}\]

      \[\begin{aligned} &\Downarrow \\[2pt] \mathbf{u}_t + &\mathbf{A}\mathbf{u}_x = \mathbf{0} \\[2pt] &\Downarrow \;\; \mathbf{u}=\mathbf{P}\mathbf{v}\\[2pt] \mathbf{v}_t +&\mathbf{P}^{-1} \mathbf{A}\mathbf{P}\mathbf{v}_x = \mathbf{0} \\[2pt] &\Downarrow \\[2pt] \mathbf{v}_t + &\boldsymbol{\Gamma}\mathbf{v}_x = \mathbf{0} \end{aligned}\]

      where

      \[\mathbf{P}= \begin{bmatrix} 2 & -2 \\ 1 & \;\;\,1 \end{bmatrix},\;\; \boldsymbol{\Gamma}= \begin{bmatrix} 4 & \;\;\,0 \\ 0 & -4 \end{bmatrix}\qquad\]

    • These are the two uncoupled equations that we can solve independently

      \[\begin{aligned} &\frac{\partial v_1}{\partial t} +4\frac{\partial v_1}{\partial x} = 0 \\ &\frac{\partial v_2}{\partial t} -4\frac{\partial v_2}{\partial x} = 0 \end{aligned}\]

      These equations have travelling-wave solutions

      \[\begin{aligned} v_1(x,t) &= \phi(x -4t) \\ v_2(x,t) &= \psi(x +4t) \end{aligned}\]

      where \(\,\phi\,\) and \(\,\psi\,\) are arbitrary differentiable functions

    • To obtain our general solution \(\,u\), \(\,\)we merely compute

      \[\mathbf{u}=\mathbf{P}\mathbf{v}= \begin{bmatrix} 2\phi(x -4t) -2\psi(x+4t)\\ \phi(x -4t) +\psi(x+4t) \end{bmatrix}\]

      We now substitute general solution into the ICs to get

      \[\begin{aligned} 2\phi(x) &-2\psi(x) = f(x) \\ \phi(x) &+\psi(x) = g(x) \\ &\Downarrow \\ \phi(x) &= \frac{1}{4} \left[ f(x) +2g(x) \right] \\ \psi(x) &= \frac{1}{4} \left[ 2g(x) -f(x) \right] \end{aligned}\]

      and, \(\,\)hence, \(\,\)the solution to IVP is

      \[\begin{aligned} u_1(x,t) &= \frac{1}{2} \left[ f(x -4t) +2g(x -4t) \right] -\frac{1}{2}\left[ 2g(x +4t) -f(x +4t) \right]\\ u_2(x,t) &= \frac{1}{4} \left[ f(x -4t) +2g(x -4t) \right] +\frac{1}{4}\left[ 2g(x +4t) -f(x +4t) \right] \end{aligned}\]

\(~\)

Example \(\,\) What is the general solution of the system

\[\begin{aligned} &\frac{\partial u_1}{\partial t} +\frac{\partial u_1}{\partial x} +\frac{\partial u_2}{\partial x} = 0\\ &\frac{\partial u_2}{\partial t} +4\frac{\partial u_1}{\partial x} +\frac{\partial u_2}{\partial x} = 0 \end{aligned}\]

\(~\)

Worked Exercises