Partial Differential Equations III & V, Exercise Sheet 6: Solutions
Lecturer: Amit Einav

  1. The Fourier transform: The heat equation with source term.

  2. The Fourier transform: The transport equation.

  3. The Fourier transform: Schrödinger’s equation.

  4. The Fourier transform: The wave equation. Use the Fourier transform to derive the solution \[u(x,t)=\frac 12 [g(x-ct)+g(x+ct)]\] of the wave equation \[\begin{aligned} u_{tt} = c^2 u_{xx} \quad & \textrm{for } (x,t) \in \mathbb{R} \times (0,\infty), \\ u(x,0) = g(x) \quad & \textrm{for } x \in \mathbb{R}, \\ u_t(x,0) = 0 \quad & \textrm{for } x \in \mathbb{R}, \end{aligned}\] where the constant \(c>0\) is the wave speed. This is known as D’Alembert’s solution. Hint: Use Q2(i) and the fact that \(\cos(c \xi t) = [\exp(i c \xi t) + \exp(- i c \xi t)]/2\).

    Taking the Fourier transform of \(u_{tt} = c^2 u_{xx}\) with respect to the \(x\) variable gives \[\widehat{u_{tt}} = \widehat{c^2 u_{xx}} \quad \Longleftrightarrow \quad \hat{u}_{tt}(\xi,t) = c^2 (i \xi)^2 \hat{u}(\xi,t) = - c^2 \xi^2 \hat{u}(\xi,t).\] Taking the Fourier transform of the initial condition \(u(x,0)=g(x)\) gives \[\hat{u}(\xi,0)=\hat{g}(\xi).\] Taking the Fourier transform of the initial condition \(u_t(x,0)=0\) gives \[\hat{u}_t(\xi,0)=0.\] We have reduced the PDE to a one-parameter family of uncoupled ODEs, indexed by \(\xi\): \[\hat{u}_{tt} = -c^2 \xi^2 \hat{u}, \qquad \hat{u}(\xi,0)=\hat{g}(\xi), \qquad \hat{u}_t(\xi,0) = 0.\] Recall that the ODE \(\ddot{x} = -\lambda^2 x\) has solution of the form \(x(t) = A \cos(\lambda t) + B \sin(\lambda t)\). Applying this with \(x = \hat{u}\), \(\lambda = c \xi\) gives \[\hat{u}(\xi,t) = A \cos(c \xi t) + B \sin (c \xi t).\] The initial conditions \(\hat{u}(\xi,0)=\hat{g}(\xi)\), \(\hat{u}_t(\xi,0) = 0\) imply that \(A = \hat{g}(\xi)\) and \(B=0\). Therefore \[\begin{aligned} \hat{u}(\xi,t) & = \hat{g}(\xi) \cos(c \xi t) \\ & = \hat{g}(\xi) \left[ \frac{\exp(i c \xi t) + \exp(- i c \xi t)}{2} \right] \\ & = \frac 12 \exp(i c \xi t) \hat{g}(\xi) + \frac 12 \exp(-i c \xi t) \hat{g}(\xi) \\ & = \frac 12 \, \widehat{\tau_{-a} g}(\xi) + \frac 12 \, \widehat{\tau_{a} g}(\xi) \end{aligned}\] where \(a=ct\), by Q2(i). Taking the inverse Fourier transform gives \[\begin{aligned} u(x,t)& =\frac 12 \, \tau_{-a} \, g(x) + \frac12 \, \tau_{a} \, g(x) \\ &= \frac 12 \, g(x+a) + \frac 12 \, g(x-a) \\ & = \frac 12 [g(x+ct)+g(x-ct)] \end{aligned}\] as required.

  5. The Fourier transform of a derivative. By definition \[\begin{aligned} \widehat{u'}(\xi) & = \frac{1}{\sqrt{2 \pi}} \int_{-\infty}^\infty u'(x) e^{-i \xi x} \, dx \\ & = - \frac{1}{\sqrt{2 \pi}} \int_{-\infty}^\infty u(x) \frac{d}{dx} e^{-i \xi x} \, dx & \textrm{(integration by parts)} \\ & = - \frac{1}{\sqrt{2 \pi}} \int_{-\infty}^\infty u(x) (-i \xi) e^{-i \xi x} \, dx \\ & = i \xi \frac{1}{\sqrt{2 \pi}} \int_{-\infty}^\infty u(x) e^{-i \xi x} \, dx \\ & = i \xi \hat{u}(\xi) \end{aligned}\] as required.

  6. The Fourier transform of a convolution. By definition \[\begin{aligned} \widehat{u*v}(\xi) & = \frac{1}{\sqrt{2 \pi}} \int_{-\infty}^\infty (u*v)(x) e^{-i \xi x} \, d x \\ & = \frac{1}{\sqrt{2 \pi}} \int_{-\infty}^\infty \left( \int_{-\infty}^\infty u(z) v(x-z) \, dz \right) e^{-i \xi x} \, d x \\ & = \frac{1}{\sqrt{2 \pi}} \int_{-\infty}^\infty \left( \int_{-\infty}^\infty u(z) v(x-z) \, dz \right) e^{- i \xi z} e^{-i \xi(x-z)} \, d x \\ & = \frac{1}{\sqrt{2 \pi}} \int_{-\infty}^\infty \left( \int_{-\infty}^\infty v(x-z) e^{-i \xi(x-z)} \, d x \right) u(z) e^{- i \xi z} \, d z \\ & = \int_{-\infty}^\infty \left( \frac{1}{\sqrt{2 \pi}} \int_{-\infty}^\infty v(\tilde{x}) e^{-i \xi \tilde{x}} \, d \tilde{x} \right) u(z) e^{- i \xi z} \, d z & (\tilde{x} = x-z) \\ & = \int_{-\infty}^\infty \hat{v}(\xi) u(z) e^{- i \xi z} \, d z \\ & = \sqrt{2 \pi} \, \hat{v}(\xi) \frac{1}{\sqrt{2 \pi}} \int_{-\infty}^\infty u(z) e^{- i \xi z} \, d z \\ & = \sqrt{2 \pi} \, \hat{v}(\xi) \hat{u}(\xi) \end{aligned}\] as desired.

  7. Proof of the Sobolev embedding using the Fourier transform.

  8. Fundamental Solution of the Heat Equation. We compute \[\begin{aligned} \Phi_t(\boldsymbol{x},t) & = \frac{1}{(4 \pi k)^{\frac n2}} e^{-\frac{|\boldsymbol{x}|^2}{4kt}} \left( - \frac n2 t^{-\frac n2 - 1} + \frac{|\boldsymbol{x}|^2}{4 k t^2} \right) \\ \frac{\partial \Phi}{\partial x_i}(\boldsymbol{x},t) & = \frac{1}{(4 \pi k t)^{\frac n2}} e^{-\frac{|\boldsymbol{x}|^2}{4kt}} \left( - \frac{2 x_i}{4kt} \right), \\ \frac{\partial^2 \Phi}{\partial x_i^2}(\boldsymbol{x},t) & = \frac{1}{(4 \pi k t)^{\frac n2}} e^{-\frac{|\boldsymbol{x}|^2}{4kt}} \left( - \frac{2}{4kt} + \left( \frac{2 x_i}{4kt} \right)^2 \right) = \frac{1}{(4 \pi k t)^{\frac n2}} e^{-\frac{|\boldsymbol{x}|^2}{4kt}} \left( - \frac{1}{2kt} + \frac{ x_i^2}{4k^2t^2} \right), \\ \Delta \Phi (\boldsymbol{x},t) & = \sum_{i=1}^n \frac{\partial^2 \Phi}{\partial x_i^2}(\boldsymbol{x},t) = \frac{1}{(4 \pi k t)^{\frac n2}} e^{-\frac{|\boldsymbol{x}|^2}{4kt}} \left( - \frac{n}{2kt} + \frac{ |\boldsymbol{x}|^2}{4k^2t^2} \right). \end{aligned}\] Therefore \(\Phi\) satisfies \(\Phi_t(\boldsymbol{x},t) = k \Delta \Phi(\boldsymbol{x},t)\) for all \(\boldsymbol{x}\in \mathbb{R}^n\), \(t>0\), as required.

  9. Finite speed of propagation for a degenerate diffusion equation.

  10. The mathematical equation that caused the banks to crash. Let \(t(x,\tau)\) satisfy \[\tau = T-t(x,\tau).\] Differentiating this expression with respect to \(\tau\) and \(x\) gives \[\begin{gathered} t_\tau = -1, \qquad t_x = 0. \end{gathered}\] Let \(S(x,\tau)\) satisfy \[\begin{equation} \label{eq:BS} x = \ln \left( \frac{S(x,\tau)}{K} \right) + \left( r - \frac 12 \sigma^2 \right) \tau. \end{equation}\] Differentiating this expressions with respect to \(\tau\) gives \[\begin{aligned} 0 = \frac{K}{S} \frac{S_\tau}{K} + r - \frac 12 \sigma^2 \quad \Longleftrightarrow \quad S_\tau = \left( \frac12 \sigma^2 - r \right) S. \end{aligned}\] Differentiating equation \(\eqref{eq:BS}\) with respect to \(x\) gives \[1 = \frac{K}{S} \frac{S_x}{K} \quad \Longleftrightarrow \quad S_x = S.\] Therefore \[\begin{aligned} u_\tau & = C e^{r \tau} \left[ rV + V_S S_\tau + V_t t_\tau \right] = C e^{r \tau} \left[ rV + \left( \frac12 \sigma^2 - r \right) S V_s - V_t \right], \\ u_x & = Ce^{r \tau} \left[ V_S S_x + V_t t_x \right] = Ce^{r \tau} S V_S, \\ u_{xx} & = Ce^{r \tau} \left[ S_x V_S + S V_{SS} S_x \right] = Ce^{r \tau} \left[ S V_S + S^2 V_{SS} \right]. \end{aligned}\] Therefore \[\begin{aligned} u_\tau - \frac 12 \sigma^2 u_{xx} & = C e^{r \tau} \left[ rV + \left( \frac12 \sigma^2 - r \right) S V_s - V_t - \frac 12 \sigma^2 (S V_S + S^2 V_{SS} ) \right] \\ & = - C e^{r \tau} \left[ V_t + \frac12 \sigma^2 S^2 V_{SS} + rS V_S - rV \right] \\ & = 0 \end{aligned}\] since \(V\) satisfies the Black-Scholes PDE. This completes the proof.

  11. The energy method: Uniqueness for the heat equation in a time dependent domain. Let \(u\) and \(v\) be solutions and let \(w=u-v\). Then \(w\) satisfies \[\begin{aligned} w_{t} - k w_{xx} =0 \quad & \textrm{for } (x,t) \in U, \\ w(a(t),t)=0 \quad & \textrm{for } t \in [0,T], \\ w(b(t),t)=0 \quad & \textrm{for } t \in [0,T], \\ w(x,0)=0 \quad & \textrm{for } x \in (a(0),b(0)). \end{aligned}\] Multiply the equation \(w_t = k w_{xx}\) by \(w\) and integrate over \((a(t),b(t))\) to obtain \[\begin{equation} \label{Q6:1} \int_{a(t)}^{b(t)} w w_t \, dx = k \int_{a(t)}^{b(t)} w w_{xx} \, dx. \end{equation}\] Recall the Fundamental Theorem of Calculus: \[\frac{d}{dt} \int_{a(t)}^{b(t)} f(x,t) \, dx = \int_{a(t)}^{b(t)} f_t(x,t) \, dx + \dot{b}(t) f(b(t),t) - \dot{a}(t) f(a(t),t).\] We can use this to rewrite the left-hand side of equation \(\eqref{Q6:1}\) as follows: \[\begin{align} \nonumber \int_{a(t)}^{b(t)} w w_t \, dx & = \frac{d}{dt} \frac{1}{2} \int_{a(t)}^{b(t)} w^2(x,t) \, dx - \frac12 \dot{b}(t) w^2(b(t),t) + \frac 12 \dot{a}(t) w^2(a(t),t) \\ \label{Q6:2} & = \frac{d}{dt} \frac{1}{2} \int_{a(t)}^{b(t)} w^2(x,t) \, dx \end{align}\] since \(w(a(t),t)=w(b(t),t)=0\). We rewrite the right-hand side of equation \(\eqref{Q6:1}\) using integration by parts: \[\begin{equation} \label{Q6:3} k \int_{a(t)}^{b(t)} w w_{xx} \, dx = k w w_x \Big|_{a(t)}^{b(t)} - k \int_{a(t)}^{b(t)} w_x^2 \, dx = - k \int_{a(t)}^{b(t)} w_x^2 \, dx, \end{equation}\] again using the fact that \(w(a(t),t)=w(b(t),t)=0\). Substituting \(\eqref{Q6:2}\) and \(\eqref{Q6:3}\) into \(\eqref{Q6:1}\) gives \[\frac{d}{dt} \frac{1}{2} \int_{a(t)}^{b(t)} w^2(x,t) \, dx = - k \int_{a(t)}^{b(t)} w_x^2 \, dx \le 0.\] Let \[E(t) = \frac{1}{2} \int_{a(t)}^{b(t)} w^2(x,t) \, dx.\] We have shown that \(\dot{E}(t) \le 0\). Hence \[0 \le E(t) \le E(0) = 0\] since \(w=0\) for \(t=0\). Consequently \(E(t)=0\) for all \(t\). Therefore \(w=0\) in \(U\) and so \(u=v\), as required.

  12. The energy method: Uniqueness for a 4th-order heat equation. Let \(u\) and \(v\) be solutions and let \(w=u-v\). Then \(w\) satisfies \[\begin{align} \label{Q7:1} w_t + k w_{xxxx} = 0 \quad & \textrm{for } (x,t) \in (a,b) \times (0,T], \\ \label{Q7:2} w(a,t) = w(b,t)=0 \quad & \textrm{for } t \in [0,T], \\ \label{Q7:3} w_x(a,t) = w_x(b,t)=0 \quad & \textrm{for } t \in [0,T], \\ \label{Q7:4} w(x,0) = 0 \quad & \textrm{for } x \in (a,b). \end{align}\] Multiplying the equation \(w_t = - k w_{xxxx}\) by \(w\) and integrating over \((a,b)\) gives \[\begin{aligned} \int_a^b \underbrace{w w_t}_{\frac12 \frac{\partial}{\partial t} w^2} \, dx & = - k \int_a^b w w_{xxxx} \, dx \\ & = -k w w_{xxx} \Big|_a^b + k \int_a^b w_x w_{xxx} \, dx \\ & = k \int_a^b w_x w_{xxx} \, dx & \textrm{(by \eqref{Q7:2})} \\ & = k w_x w_{xx} \Big|_a^b - k\int_a^b w_{xx} w_{xx} \, dx \\ & = - k\int_a^b w^2_{xx} \, dx & \textrm{(by \eqref{Q7:3}).} %\\ %& = - k \int_a^b w_{xx}^2 \, dx. \end{aligned}\] Therefore \[\frac{d}{dt} \frac12 \int_a^b w^2 \, dx = - k \int_a^b w_{xx}^2 \, dx \le 0\] and so \[0 \le \int_a^b w^2(x,t) \, dx \le \int_a^b w^2(x,0) \, dx = 0\] by \(\eqref{Q7:4}\). We conclude that \(w=0\) and hence \(u=v\), as required.

    We consider \(u_t + k u_{xxxx}=0\) to be the 4th-order version of the heat equation \(u_t - k u_{xx} = 0\) since it has the same energy-decay property: \[\frac{d}{dt} \| u \|^2_{L^2([a,b])} \le 0\] (provided that \(u\) and \(u_x\) vanish at \(x=a\) and \(x=b\)). The equation \(u_t - k u_{xxxx}=0\) looks more similar to the heat equation \(u_t - k u_{xx}=0\) (because of the minus sign), but its \(L^2\)–energy grows with time: \[\frac{d}{dt} \| u \|^2_{L^2([a,b])} \ge 0.\]

  13. Asymptotic behaviour of the heat equation with time independent data. Let \(w(\boldsymbol{x},t)=u(\boldsymbol{x},t)-v(\boldsymbol{x})\). We need to prove that \(\lim_{t \to \infty} \| w \|_{L^2(\Omega)} = 0\). By subtracting the PDEs for \(u\) and \(v\) we find that \(w\) satisfies \[\begin{aligned} w_t(\boldsymbol{x},t) - k \Delta w(\boldsymbol{x},t) = 0 \quad & \textrm{for } (\boldsymbol{x},t) \in \Omega \times (0,\infty), \\ w(\boldsymbol{x},t) = 0 \quad & \textrm{for } (\boldsymbol{x},t) \in \partial \Omega \times [0,\infty), \\ w(\boldsymbol{x},0) = u_0(\boldsymbol{x}) - v(\boldsymbol{x}) \quad & \textrm{for } \boldsymbol{x}\in \Omega. \end{aligned}\] Multiplying the equation \(w_t = k \Delta w\) by \(w\) and integrating by parts over \(\Omega\) gives \[\begin{align} \nonumber \int_{\Omega} w w_t \, d \boldsymbol{x}= k \int_\Omega w \Delta w \, d \boldsymbol{x} \quad & \Longleftrightarrow \quad \frac{d}{dt} \frac12 \int_{\Omega} w^2 \, d \boldsymbol{x}= k \int_{\partial \Omega} w \, \nabla w \cdot \boldsymbol{n}\, dS - k \int_{\Omega} \nabla w \cdot \nabla w \, d \boldsymbol{x} \\ \label{Q12:1} & \Longleftrightarrow \quad \frac{d}{dt} \frac12 \int_{\Omega} w^2 \, d \boldsymbol{x}= - k \int_{\Omega} |\nabla w|^2 \, d \boldsymbol{x} \end{align}\] since \(w=0\) on \(\partial \Omega\). By the Poincaré inequality, there exists a constant \(C_p>0\) such that \[\int_\Omega |w|^2 \, d \boldsymbol{x}\le C_p \int_\Omega |\nabla w|^2 \, d \boldsymbol{x}.\] Multiplying this by \(-k/C_p\) gives \[\begin{equation} \label{Q12:2} - \frac{k}{C_p} \int_\Omega |w|^2 \, d \boldsymbol{x}\ge -k \int_\Omega |\nabla w|^2 \, d \boldsymbol{x}. \end{equation}\] Combining equations \(\eqref{Q12:1}\), \(\eqref{Q12:2}\) yields \[\frac{d}{dt} \frac12 \int_{\Omega} w^2 \, d \boldsymbol{x}\le - \frac{k}{C_p} \int_\Omega |w|^2 \, d \boldsymbol{x}.\] Define \[E(t) = \int_{\Omega} w^2(\boldsymbol{x},t) \, d \boldsymbol{x}= \| w \|^2_{L^2(\Omega)}\] and \(\lambda = \frac{2k}{C_p}\). We have shown that \[\dot{E} \le - \lambda E.\] By the Gr̈onwall inequality, \[E(t) \le e^{-\lambda t} E(0).\] Since \(\lambda > 0\), we conclude that \(E(t) \to 0\) as \(t \to \infty\). Therefore \(w \to 0\) in \(L^2(\Omega)\) as \(t \to \infty\), as required.

  14. Asymptotic behaviour of the heat equation with time independent data in the \(L^\infty\)–norm.

  15. Applications of the maximum principle: Uniqueness and bounds on solutions.

  16. Application of the maximum principle: Comparison Principle. Define \(v = u_1 - u_2\). Then \(v\) satisfies \[\begin{aligned} \frac{\partial v}{\partial t}(\boldsymbol{x},t) - k \Delta v(\boldsymbol{x},t) = f_1(\boldsymbol{x}) - f_2(\boldsymbol{x}) \quad & \textrm{for } (\boldsymbol{x},t) \in \Omega \times (0,T], \\ v(\boldsymbol{x},t) = g_1(\boldsymbol{x})-g_2(\boldsymbol{x}) \quad & \textrm{for } (\boldsymbol{x},t) \in \partial \Omega \times [0,T], \\ v(\boldsymbol{x},0) = u^0_1(\boldsymbol{x}) - u^0_2(\boldsymbol{x}) \quad & \textrm{for } \boldsymbol{x}\in \Omega. \end{aligned}\] Since \(f_1 \le f_2\), then \[v_t - k \Delta v = f_1 - f_2 \le 0 \quad \textrm{in } \Omega_T.\] Therefore the weak maximum principle implies that \[\max_{\overline{\Omega_T}} v = \max_{\Gamma_T} v.\] For \((\boldsymbol{x},t) \in \Gamma_T\), \[v(\boldsymbol{x},t) = \left\{ \begin{array}{cl} g_1(\boldsymbol{x}) - g_2(\boldsymbol{x}) & \textrm{if } (\boldsymbol{x},t) \in \partial \Omega \times [0,T], \\ u^0_1(\boldsymbol{x}) - u^0_2(\boldsymbol{x}) & \textrm{if } t=0, \, \boldsymbol{x}\in \Omega. \end{array} \right.\] But \[g_1 - g_2 \le 0, \qquad u^0_1 - u^0_2 \le 0.\] Therefore \(v \le 0\) on \(\Gamma_T\) and hence \[\max_{\overline{\Omega_T}} v = \max_{\Gamma_T} v \le 0.\] Hence \(v \le 0\) in \(\Omega_T\) and so \(u_1 \le u_2\) in \(\Omega_T\), as required.

  17. Eigenfunctions of the Laplacian and an application to the heat equation. Formally (not worrying about interchanging limits and infinite sums), \[\begin{aligned} 0 & = v_t - k \Delta v \\ & = \sum_{n=1}^\infty \dot{c}_n(t) u_n(\boldsymbol{x}) - k \sum_{n=1}^\infty c_n(t) \Delta u_n(\boldsymbol{x}) \\ & = \sum_{n=1}^\infty \dot{c}_n(t) u_n(\boldsymbol{x}) + k \sum_{n=1}^\infty c_n(t) \lambda_n u_n(\boldsymbol{x}) \\ & = \sum_{n=1}^\infty (\dot{c}_n(t) + k \lambda_n c_n(t) ) \, u_n(\boldsymbol{x}). \end{aligned}\] Since \(\{u_n\}_{n \in \mathbb{N}}\) forms an orthogonal basis, it follows that \[\dot{c}_n(t) + k \lambda_n c_n(t) = 0\] for all \(n\). We also have \[v(\boldsymbol{x},0) = g(\boldsymbol{x}) \quad \Longleftrightarrow \quad \sum_{n=1}^\infty c_n(0) u_n(\boldsymbol{x}) = \sum_{n=1}^\infty g_n u_n(\boldsymbol{x}) \quad \Longleftrightarrow \quad \sum_{n=1}^\infty (c_n(0)-g_n) u_n(\boldsymbol{x}) = 0.\] Again, since \(\{u_n\}_{n \in \mathbb{N}}\) forms an orthogonal basis, it follows that \[c_n(0)=g_n\] for all \(n\). We have reduced the PDE for \(v\) to a one-parameter family of uncoupled ODEs, indexed by \(n\): \[\dot{c}_n(t) = - k \lambda_n c_n(t), \qquad c_n(0)=g_n.\] These ODEs have solutions \[c_n(t) = g_n e^{- k \lambda_n t}.\] Therefore \[v(\boldsymbol{x},t) = \sum_{n=1}^\infty g_n e^{- k \lambda_n t} u_n(\boldsymbol{x})\] as required.