[Commits] [svn:einsteintoolkit] Workshop_Spring_2012/numerical_relativity/ (Rev. 29)

diener at cct.lsu.edu diener at cct.lsu.edu
Tue Apr 3 16:52:38 CDT 2012


User: diener
Date: 2012/04/03 04:52 PM

Modified:
 /numerical_relativity/
  numerical_relativity.tex

Log:
 Add some slides on finite differencing.

File Changes:

Directory: /numerical_relativity/
=================================

File [modified]: numerical_relativity.tex
Delta lines: +89 -27
===================================================================
--- numerical_relativity/numerical_relativity.tex	2012-04-03 18:30:44 UTC (rev 28)
+++ numerical_relativity/numerical_relativity.tex	2012-04-03 21:52:37 UTC (rev 29)
@@ -15,7 +15,7 @@
 %  \item 3+1 spacetime decomposition.
   \item The BSSN formulation.
   \item Puncture data.
-%  \item Finite Differencing.
+  \item Finite Differencing.
 %  \item The method of lines.
  \end{itemize}
 }
@@ -31,33 +31,33 @@
 \begin{align}
 {\cal H} & \equiv  R + K^2 - K_{ij} K^{ij} = 0, \\
 {\cal M}^i & \equiv  D_j (K^{ij} - \gamma^{ij} K) = 0.
-\end{align}
-These equations were used a lot in the early years of numerical relativity.
+\end{align}\pause
+These equations were used a lot in the early years of numerical relativity.\pause
 
 However, it was later discovered that the ADM evolution equations are only
-weakly hyperbolic.
+weakly hyperbolic.\pause
 
 In the mid to late 90's a new formulation was introduced that proved to be
 much more robust and stable: The BSSN formulation.
 }
 
 \frame{\frametitle{The BSSN formulation (continued)}
-\setlength{\vs}{-1.5ex}
+\setlength{\vs}{-0.5ex}
 Introduce a conformal rescaling of the three metric\vspace{\vs}
 \begin{equation}
 \gamma_{ij} = \psi^4 \tg_{ij}.\vspace{\vs}
-\end{equation}
-We choose $\psi = \gamma^{1/12}$ such that the determinant of $\tg_{ij}$ is 1\vspace{\vs}
+\end{equation}\pause
+We choose $\psi = \gamma^{1/12}$ such that the determinant of $\tg_{ij}$ is 1%\vspace{\vs}
 \begin{equation}
 \mathrm{det}(\tg_{ij}) = \mathrm{det}(\psi^{-4}\gamma_{ij}) = 
 \mathrm{det}(\gamma^{-1/3}\gamma_{ij}) = 
-\gamma^{-1} \mathrm{det}(\gamma_{ij}) = 1. \vspace{\vs}
-\end{equation}
+\gamma^{-1} \mathrm{det}(\gamma_{ij}) = 1.\vspace{\vs}
+\end{equation}\pause
 In addition we introduce a trace decomposition of the extrinsic curvature.\vspace{\vs}
 \begin{align}
 K & = \gamma^{ij} K_{ij}, \\
 A_{ij} & = K_{ij} - \frac{1}{3} \gamma_{ij} K.\vspace{\vs}
-\end{align}
+\end{align}\pause
 We then promote the following variables to evolution variables\vspace{\vs}
 \begin{align}
   \phi &= \ln \psi = \frac{1}{12} \ln \gamma,
@@ -75,16 +75,17 @@
 \begin{equation}
   \tG^i = \tg^{jk} \tG^i{}_{jk} = - \partial_j \tg^{ij},
 \end{equation}
-to evolved variables as well.
+to evolved variables as well.\pause
+
 The final set of evolution variables are  $\phi$, $K$, $\tg_{ij}$, $\tA_{ij}$
-and $\tG^i$.
+and $\tG^i$.\pause
 
 The BSSN evolution equations can be derived from the ADM equations. As an
 example take the equation for $\phi$.
 \begin{equation}
 \dt \phi  = \partial_0\phi = \partial_0 \left (\frac{1}{12}\ln\gamma\right)
 = \frac{1}{12}\frac{1}{\gamma}\partial_0\gamma.
-\end{equation}
+\end{equation}\pause
 Using the expression for the derivative of the determinant of the metric
 in terms of the derivatives of the metric ($\partial_0\gamma = \gamma 
 \gamma^{ij} \partial_0\gamma_{ij}$) we find
@@ -114,13 +115,11 @@
 & - 2 \tilde{A}^{ij} \partial_j\alpha
 + 2 \alpha ( \tilde{\Gamma}^i{}_{jk} \tilde{A}^{jk} + 6 \tilde{A}^{ij}
 \partial_j \phi - \frac{2}{3} \tg^{ij} \partial_j K).
-\end{align}%\vspace{\vs}
+\end{align}
 }
 
 \frame{\frametitle{The BSSN formulation (continued)}
-%\raisebox{-1.2\vs}[0pt][0pt]{
 Here $R_{ij} =  \tilde{R}_{ij} + R^{\phi}_{ij}$, where\vspace{1.0\vs}
-%}\vspace{\vs}
 \begin{align}
 R^{\phi}_{ij} = {} & - 2 \tilde{D}_i
 \tilde{D}_j \phi - 2 \tilde{\gamma}_{ij}
@@ -142,18 +141,18 @@
 \tilde{{\cal G}} & \equiv \tg-1 = 0, \\
 \tilde{{\cal A}} & \equiv \tg^{ij}\tA_{ij} = 0, \\
 \tilde{{\cal L}}^{i} & \equiv \tG^{i}+\partial_j\tg^{ij} = 0.
-\end{align}
+\end{align}\pause
 The constraints $\tilde{{\cal G}}$ and $\tilde{{\cal A}}$ are enforced actively
-at each timestep.
+at each timestep.\pause
 
 The other constraints ($\tilde{{\cal H}}$, $\tilde{{\cal M}}^i$ and
-$\tilde{{\cal L}}^{i}$) are not enforced.
+$\tilde{{\cal L}}^{i}$) are not enforced.\pause
 
 To improve stability and to help maintain $\tilde{{\cal L}}^{i}$ at a low level
 the following rule is employed in an implementation
 \begin{itemize}
 \item Where derivatives of $\tG^{i}$ are needed the evolved $\tG^{i}$ are used
-      directly.
+      directly.\pause
 \item Where $\tG^{i}$ are needed without taking derivatives
       $\tg^{jk}\tG^{i}{}_{jk}$ are used instead.
 \end{itemize}
@@ -161,7 +160,7 @@
 
 \frame{\frametitle{The BSSN formulation (continued)}
 In order to evolve a spacetime with the BSSN equations, you have to specify
-the gauges $\alpha$ and $\beta^{i}$.
+the gauges $\alpha$ and $\beta^{i}$.\pause
 
 Most codes use the ``moving puncture'' gauges
 \begin{align}
@@ -179,12 +178,12 @@
 \gamma^{\mathrm{ph}}_{ij} = \psi^4\gamma_{ab}, \mbox{\hspace{2em}} K^{\mathrm{ph}}_{ij} = \psi^{-2} K_{ij},
 \end{equation}
 where $\gamma_{ij}$ is chosen to be the flat metric and $K_{ij}$ is assumed
-to be tracefree. \\
+to be tracefree.\pause \\
 The constraint equations become\vspace{\vs}
 \begin{align}
 0 & = \Delta\psi +\frac{1}{8}K^{ij}K_{ij}\psi^{-7} \\
 0 & = D_j K^{ij}.
-\end{align}
+\end{align}\pause
 The momentum constraint has an analytic solution\vspace{\vs}
 \begin{align}
 K^{ij}_{\mathrm{BY}} = {} & \frac{3}{2r^2}\left ( P^i n^j + P^j n^i -
@@ -197,21 +196,84 @@
 
 \frame{\frametitle{Puncture data (continued)}
 \setlength{\vs}{-0.8ex}
-For $K_{ij} = 0$ the Hamiltonian constraint has a simple solution for $N$ black holes.
+For $K_{ij} = 0$ the Hamiltonian constraint has a simple solution for $N$
+black holes.
 \begin{equation}
 \psi = 1+\sum_{i=1}^N\frac{m_i}{2|\vec{r}-\vec{r}_i|}.
-\end{equation}
+\end{equation}\pause
 Inspired by this, for $K_{ij}\ne 0$ we make the ansatz
 \begin{equation}
 \psi=\frac{1}{\alpha}+u, \mbox{\hspace{2em}} 
 \frac{1}{\alpha} = \sum_{i=1}^N\frac{m_i}{2|\vec{r}-\vec{r}_i|},
-\end{equation}
+\end{equation}\pause
 In which case the Hamiltonian constraint becomes an equation for $u$
 \begin{equation}
 \Delta u+\frac{1}{8}\alpha^7 K^{ij}K_{ij}(1+\alpha u)^{-7} = 0.
-\end{equation}
+\end{equation}\pause
 It can be shown that $u$ will be $C^2$ at the location of the `punctures' and
 that $\psi$ has a unique solution.
 }
 
+\frame{\frametitle{Finite Differencing}
+With finite differencing we discretize a function by sampling it at a 
+collection of grid points.\pause
+
+The grid points are usually (but not necessarily) equally spaced. \pause
+
+We can then approximate derivatives of a function at a grid point by a
+weigthed sum of function values at grid points in the neighbourhood (the
+stencil) of the grid point. \pause
+
+As an example consider a stencil containing the grid point ($f_i$) and it's two 
+nearest neighbors ($f_{i-1}$ and $f_{i+1}$) with 
+$\Delta x=x_{i+1}-x_i=x_i-x_{i-1}$.
+\begin{equation}
+  \left. \frac{df}{dx} \right |_{x_i}\approx \frac{1}{\Delta x}\sum_{j=-1}^{j=1} a_j f_{i+j}.
+\end{equation}
+}
+\frame{\frametitle{Finite Differencing (continued)}
+The coefficients $a_j$ can be found be expanding $f$ in a Taylor series around
+$x_i$ for the grid points in the stencil
+\begin{eqnarray}
+f(x_{i-1}) & = & f(x_i) - \left. \frac{df}{dx}\right |_{x_i} \Delta x
+             + \frac{1}{2} \left. \frac{d^2 f}{dx^2}\right |_{x_i} \Delta x^2
+             + O((\Delta x)^3) \nonumber \\
+f(x_i) & = & f(x_i) \nonumber \\
+f(x_{i+1}) & = & f(x_i) + \left. \frac{df}{dx}\right |_{x_i} \Delta x
+             + \frac{1}{2} \left. \frac{d^2 f}{dx^2}\right |_{x_i} \Delta x^2
+             + O((\Delta x)^3) \nonumber
+\end{eqnarray} \pause
+Requiring that the weighted sum approximates the derivative yields
+the following equations for $a_{-1}$, $a_0$ and $a_1$
+\begin{eqnarray}
+0  & = & a_{-1}+a_0+a_1 \nonumber \\
+1 & = & -a_{-1}+a_1 \nonumber \\
+0 & = & a_{-1}+a_1 \nonumber
+\end{eqnarray}
+with the solution $a_{-1}=-1/2, a_0=0, a_1=1/2$.
+}
+
+\frame{\frametitle{Finite Differencing (continued)}
+Thus we find that
+\begin{equation}
+\left. \frac{df}{dx} \right |_{x_i} = \frac{f_{i+1}-f_{i-1}}{2\Delta x}
+ + O((\Delta x)^2).
+\end{equation}\pause
+Similarly we find for the second derivative that
+\begin{equation}
+\left. \frac{d^2f}{dx^2} \right |_{x_i} = \frac{f_{i+1}-2 f_i+f_{i-1}}{(\Delta x)^2}
+ + O((\Delta x)^2).
+\end{equation}\pause
+These finite difference operators are second order accurate.\pause
+
+Higher order accuracy or higher order derivatives require larger stencils.\pause
+
+Another way of looking at finite differencing operators is through 
+interpolating polynomials.\pause
+
+Either approach gives the same coefficients for the same stencil.\pause
+
+It is also clear from either approach that the error estimates are only
+correct if $f$ is smooth enough.
+}
 \end{document}



More information about the Commits mailing list