PDE notes: Wave Equation

2.1. Introduction

The one-dimensional wave equation for small displacement of a perfectly elastic string of length {l} with no frictional forces and no restoring forces is

\displaystyle \rho_0 u_{tt} = T_0 u_{xx} \ \ \ \ \ (88)

or

\displaystyle u_{tt} = c^2 u_{xx} \ \ \ \ \ (89)

for {t\geq0}, {0\leq x \leq l}, where {c^2=\frac{T_0}{\rho_0}}. We assume both tension {T_0} and linear density {\rho_0} to be constant. Tension {T_0} has units of {\frac{mass}{time^2}} and linear density has units of {\frac{mass}{length^2}}, so {c=\sqrt{\frac{T_0}{\rho_0}}} has the units of {\frac{length}{time}}, which is also the unit of speed. Thus, {c} turns out to be the velocity of wave propagation along the string. The one-dimensional wave equation models sound waves, water waves, vibrations in solids, and longitudinal or torsional vibrations in a rod, among other things.

Since the PDE in (89) contains the second time derivative, two initial conditions are required. The initial conditions usually take the follwing form:

\displaystyle \mathrm{Initial \hspace{0.2cm} displacement:} \hspace{0.5cm} u(x,0)= f(x) \ \ \ \ \ (90)

\displaystyle \mathrm{Initial \hspace{0.2cm} velocity:} \hspace{0.5cm} v(x,0)=u_t(x,0)= g(x) \ \ \ \ \ (91)

Typical boundary conditions are of the same form as thus given in the discussion of one-dimensional heat equation. For homogeneous Dirichlet conditions,

\displaystyle u(0,t)=0=u(l,t) \ \ \ \ \ (92)

for {t \geq 0}. The end ends of the vibrating strings are fixed.

For homogeneous Neumann conditions,

\displaystyle u_x(0,t)=0=u_x(l,t) \ \ \ \ \ (93)

for {t \geq 0}. These conditions are be achieved, for example, by attaching the ends of the string to a frictionless sleeve that moves vertically.

2.2. Solution by separation of variables

We now solve the one-dimensional wave equation with homogeneous Dirichlet boundary conditions. The following problem is defined for {0 \leq x \leq L} and {t \geq 0}:

\displaystyle \mathrm{PDE: } \hspace{0.5cm} u_{tt}=c^2u_{xx} \ \ \ \ \ (94)

\displaystyle \mathrm{BC1:} \hspace{0.5cm} u(0,t)=0 \ \ \ \ \ (95)

\displaystyle \mathrm{BC2:}\hspace{0.5cm} u(L,t)=0 \ \ \ \ \ (96)

\displaystyle \mathrm{IC1:}\hspace{0.5cm} u(x,0) = f(x) \ \ \ \ \ (97)

\displaystyle \mathrm{IC2:}\hspace{0.5cm} u_x(x,0) = g(x) \ \ \ \ \ (98)

Solution Since both the PDE and the boundary conditions are linear and homogeneous, the method of separation of variables is attempted. We look for special product solutions of the form:

\displaystyle u(x,t)=\phi(x)h(t) \ \ \ \ \ (99)

Substitute (99) into (94) yields

\displaystyle \phi(x)h''(t)=c^2h(t)\phi''(x) \ \ \ \ \ (100)

Divide both sides by {\phi(x)h(t)c^2} separates the variables:

\displaystyle \frac{1}{h(t)c^2}h''(t)= \frac{1}{\phi(x)}\phi''(x)=-\lambda \ \ \ \ \ (101)

where {\lambda} is the separation constant. This implies that

\displaystyle h''=-\lambda hc^2 \ \ \ \ \ (102)

\displaystyle \phi''=-\lambda \phi \ \ \ \ \ (103)

We consider (103) first since it has a complete set of boundary conditions. Letting {\phi=e^{rx}}, then {\phi''=r^2e^{rx}}. So {r=\pm\sqrt{-\lambda}}. Using results from the heat equation, we know that {\lambda < 0} and {\lambda=0} yields trivial solutions. For {\lambda > 0}, the general solution is

\displaystyle \phi(x) = c_1\mathrm{cos}(\sqrt{\lambda}x)+c_2\mathrm{sin}(\sqrt{\lambda}x) \ \ \ \ \ (104)

Apply {\phi(0)=0} and we arrive at {c_1=0}. Consider {\phi(L)=0}, we arrive at {c_2\mathrm{sin}(\sqrt{\lambda}L)=0}. If {c_2=0}, we get a trivial solution. Therefore, we get a nontrivial solution if and only if

\displaystyle \mathrm{sin}(\sqrt{\lambda}L)=0 \ \ \ \ \ (105)

This means that

\displaystyle \sqrt{\lambda}L= n\pi \ \ \ \ \ (106)

The eigenvalues are

\displaystyle \lambda= (\frac{n\pi}{L})^2, n=1,2,3... \ \ \ \ \ (107)

The eigenfunctions corresponding to the eigenvalues are

\displaystyle \phi(x)=c_2 \mathrm{sin}(\frac{n\pi x}{L}), n=1,2,3... \ \ \ \ \ (108)

The time-dependent part of the solution is

\displaystyle h(t) = c_3\mathrm{cos}(\frac{cn\pi t}{L})+c_4\mathrm{sin}(\frac{cn\pi t}{L}) \ \ \ \ \ (109)

for {n=1,2,3...}. Thus, for {n\geq 0}, each of the functions

\displaystyle u_n(x,t)=\phi_n(x)h_n(t) = \mathrm{sin}(\frac{n\pi x}{L})(A_n\mathrm{cos}(\frac{cn\pi t}{L})+B_n\mathrm{sin}(\frac{cn\pi t}{L})) \ \ \ \ \ (110)

is a solution to the PDE and satisfies the boundary condition. By superposition principle, we can solve the initial value problem by considering a linear combinations of all product solutions:

\displaystyle u(x,t) = \sum^{\infty}_{0}\mathrm{sin}(\frac{n\pi x}{L})(A_n\mathrm{cos}(\frac{cn\pi t}{L})+B_n\mathrm{sin}(\frac{cn\pi t}{L})) \ \ \ \ \ (111)

The initial conditions in (97) and (98) are satisfied if,

\displaystyle f(x)=\sum^{\infty}_{0}A_n\mathrm{sin}(\frac{n\pi x}{L}) \ \ \ \ \ (112)

\displaystyle g(x)=\sum^{\infty}_{0}B_n\frac{cn\pi}{L}\mathrm{sin}(\frac{n\pi x}{L}) \ \ \ \ \ (113)

We can consider the fact that {\mathrm{sin}(\frac{n\pi x}{L})} satisfies the following orthogonality relation:

\displaystyle \int_{0}^{L} \mathrm{sin}(\frac{n \pi x}{L})\mathrm{sin}(\frac{m \pi x}{L}) \,dx = \left\{ \begin{array}{lr} 0 & n \neq m \\ \frac{L}{2} & n=m \\ \end{array} \right. \ \ \ \ \ (114)

Multiply (112) by {\mathrm{sin}(\frac{m \pi x}{L})} and integrating from {0} to {L} yields

\displaystyle \int_{0}^{L} f(x)\mathrm{sin}(\frac{m \pi x}{L})\, dx = A_{n}\int_{0}^{L}\mathrm{sin^2}(\frac{m\pi x}{L})\, dx \ \ \ \ \ (115)

Solving for {A_n} yields,

\displaystyle A_n=\frac{2}{L} \int_{0}^{L}f(x)\mathrm{sin}(\frac{n \pi x}{L})\, dx \ \ \ \ \ (116)

Multiply (113) by {\mathrm{sin}(\frac{m \pi x}{L})} and integrating from {0} to {L} yields

\displaystyle \int_{0}^{L} f(x)\mathrm{sin}(\frac{m \pi x}{L})\, dx = B_n\frac{cn\pi}{L}\int_{0}^{L}\mathrm{sin^2}(\frac{m\pi x}{L})\, dx \ \ \ \ \ (117)

Solving for {B_n} yields,

\displaystyle B_n\frac{cn\pi}{L}=\frac{2}{L} \int_{0}^{L}f(x)\mathrm{sin}(\frac{n \pi x}{L})\, dx \ \ \ \ \ (118)

Therefore, the PDE with homogeneous Dirichlet boundary conditions has a simple explicit solution.

{\square}

The product solutions are also called the normal modes of vibration. The coefficients of {t} inside the product solutions, namely {\frac{n\pi c}{L}, n=1,2,3...}, are called the frequencies. The fundamental mode of the string has a frequency of {\frac{\pi c}{L}}. The {n}th overtone is just the {n}th integral multiple of the fundamental.

We now consider the motion of a vibrating string governed by the homogeneous Neumann boundary conditions for the wave equation:

\displaystyle \mathrm{PDE: } \hspace{0.5cm} u_{tt}=c^2u_{xx} \ \ \ \ \ (119)

\displaystyle \mathrm{BC1:} \hspace{0.5cm} u_x(0,t)=0 \ \ \ \ \ (120)

\displaystyle \mathrm{BC2:}\hspace{0.5cm} u_x(L,t)=0 \ \ \ \ \ (121)

\displaystyle \mathrm{IC1:}\hspace{0.5cm} u(x,0) = f(x) \ \ \ \ \ (122)

\displaystyle \mathrm{IC2:}\hspace{0.5cm} u_x(x,0) = g(x) \ \ \ \ \ (123)

Solution Again, we use the method of separation of variables. We look for production solutions of the form:

\displaystyle u(x,t)=\phi(x)h(t) \ \ \ \ \ (124)

with

\displaystyle h''=-\lambda hc^2 \ \ \ \ \ (125)

\displaystyle \phi''=-\lambda \phi \ \ \ \ \ (126)

The general solution for (126) is

\displaystyle \phi(x) = c_1\mathrm{cos}(\sqrt{\lambda}x)+c_2\mathrm{sin}(\sqrt{\lambda}x) \ \ \ \ \ (127)

We also need

\displaystyle \phi'(x) = \sqrt{\lambda}(c_2\mathrm{cos}(\sqrt{\lambda}x)-c_1\mathrm{sin}(\sqrt{\lambda}x)) \ \ \ \ \ (128)

Apply {\phi'(0)=0} and we arrive at {c_2=0}. Apply {\phi'(L)=0} and we arrive at {\sqrt{\lambda} c_1\mathrm{sin}(\sqrt{\lambda}x)=0}. Since {c_1=0} yields a trivial solution, we get a nontrivial solution if and only if

\displaystyle \mathrm{sin}(\sqrt{\lambda}L)=0 \ \ \ \ \ (129)

This means that

\displaystyle \sqrt{\lambda}L= n\pi \ \ \ \ \ (130)

The eigenvalues are

\displaystyle \lambda= (\frac{n\pi}{L})^2, n=1,2,3... \ \ \ \ \ (131)

The eigenfunctions corresponding to the eigenvalues are

\displaystyle \phi(x)=c_1 \mathrm{cos}(\frac{n\pi x}{L}), n=1,2,3... \ \ \ \ \ (132)

The time-dependent part of the solution is

\displaystyle h(t) = c_3\mathrm{cos}(\frac{cn\pi t}{L})+c_4\mathrm{sin}(\frac{cn\pi t}{L}) \ \ \ \ \ (133)

for {n=0,1,2,3...}. Thus, for {n\geq 0}, each of the functions

\displaystyle u_n(x,t)=\phi_n(x)h_n(t) = \mathrm{cos}(\frac{n\pi x}{L})(A_n\mathrm{cos}(\frac{cn\pi t}{L})+B_n\mathrm{sin}(\frac{cn\pi t}{L})) \ \ \ \ \ (134)

is a solution to the PDE and satisfies the boundary condition. By superposition principle, we can solve the initial value problem by considering a linear combinations of all product solutions:

\displaystyle u(x,t) = A_0+\sum^{\infty}_{1}\mathrm{cos}(\frac{n\pi x}{L})(A_n\mathrm{cos}(\frac{cn\pi t}{L})+B_n\mathrm{sin}(\frac{cn\pi t}{L})) \ \ \ \ \ (135)

The initial conditions in (97) and (98) are satisfied if,

\displaystyle f(x)=A_0+\sum^{\infty}_{1}A_n\mathrm{cos}(\frac{n\pi x}{L}) \ \ \ \ \ (136)

\displaystyle g(x)=A_0+\sum^{\infty}_{1}B_n\frac{cn\pi}{L}\mathrm{cos}(\frac{n\pi x}{L}) \ \ \ \ \ (137)

for {0<x<L}. We use the fact that {\mathrm{cos}(\frac{n\pi x}{L})} satisfies the following orthogonality relation:

\displaystyle \int_{0}^{L} \mathrm{cos}(\frac{n \pi x}{L})\mathrm{cos}(\frac{m \pi x}{L}) \,dx = \left\{ \begin{array}{lr} 0 & n \neq m \\ \frac{L}{2} & n=m\neq 0 \\ L & n=m=0 \end{array} \right. \ \ \ \ \ (138)

Multiply (136) by {\mathrm{cos}(\frac{m \pi x}{L})} and integrating from {0} to {L} yields

\displaystyle \int_{0}^{L} f(x)\mathrm{cos}(\frac{m \pi x}{L})\mathrm{dx} = A_{n}\int_{0}^{L}\mathrm{cos^2}(\frac{m\pi x}{L})\mathrm{dx} \ \ \ \ \ (139)

Solving for {A_n} yields,

\displaystyle A_0= \frac{1}{L} \int_{0}^{L}f(x) \mathrm{dx} \ \ \ \ \ (140)

\displaystyle A_n=\frac{2}{L} \int_{0}^{L}f(x)\mathrm{cos}(\frac{n \pi x}{L})\mathrm{dx}, n\geq 1 \ \ \ \ \ (141)

Multiply (137) by {\mathrm{cos}(\frac{m \pi x}{L})} and integrating from {0} to {L} yields

\displaystyle B_n\frac{cn\pi}{L}=\frac{2}{L} \int_{0}^{L}f(x)\mathrm{cos}(\frac{n \pi x}{L})\mathrm{dx}, n\geq 1 \ \ \ \ \ (142)

{\square}

Leave a Reply