1 |
jgs |
102 |
% $Id$ |
2 |
|
|
|
3 |
|
|
\chapter{The First Steps} |
4 |
|
|
\label{FirstSteps} |
5 |
|
|
|
6 |
|
|
\begin{figure} |
7 |
|
|
\centerline{\includegraphics[width=\figwidth]{FirstStepDomain}} |
8 |
|
|
\caption{Domain $\Omega=[0,1]^2$ with outer normal field $n$.} |
9 |
|
|
\label{fig:FirstSteps.1} |
10 |
|
|
\end{figure} |
11 |
|
|
|
12 |
jgs |
107 |
In this chapter we will give an introduction how to use \escript to solve |
13 |
|
|
a partial differential equation \index{partial differential equation} (PDE \index{partial differential equation!PDE}). The reader should be familiar with Python. The knowledge presented at the Python tutorial at \url{http://docs.python.org/tut/tut.html} |
14 |
|
|
is sufficient. It is helpful if the reader has some basic knowledge of PDEs \index{partial differential equation}. |
15 |
jgs |
102 |
|
16 |
jgs |
107 |
The PDE \index{partial differential equation} we wish to solve is the Poisson equation \index{Poisson equation} |
17 |
jgs |
102 |
\begin{equation} |
18 |
|
|
-\Delta u =f |
19 |
|
|
\label{eq:FirstSteps.1} |
20 |
|
|
\end{equation} |
21 |
jgs |
107 |
for the solution $u$. The function $f$ is the given right hand side. The domain of interest, denoted by $\Omega$ |
22 |
jgs |
102 |
is the unit square |
23 |
|
|
\begin{equation} |
24 |
|
|
\Omega=[0,1]^2=\{ (x\hackscore 0;x\hackscore 1) | 0\le x\hackscore{0} \le 1 \mbox{ and } 0\le x\hackscore{1} \le 1 \} |
25 |
|
|
\label{eq:FirstSteps.1b} |
26 |
|
|
\end{equation} |
27 |
jgs |
107 |
The domain is shown in \fig{fig:FirstSteps.1}. |
28 |
jgs |
102 |
|
29 |
|
|
$\Delta$ denotes the Laplace operator\index{Laplace operator} which is defined by |
30 |
|
|
\begin{equation} |
31 |
|
|
\Delta u = (u\hackscore {,0})\hackscore{,0}+(u\hackscore{,1})\hackscore{,1} |
32 |
|
|
\label{eq:FirstSteps.1.1} |
33 |
|
|
\end{equation} |
34 |
jgs |
107 |
where, for any function $w$ and any direction $i$, $u\hackscore{,i}$ |
35 |
|
|
denotes the partial derivative \index{partial derivative} of $u$ with respect to $i$. |
36 |
jgs |
102 |
\footnote{Some readers |
37 |
|
|
may be more familiar with the Laplace operator\index{Laplace operator} being written |
38 |
|
|
as $\nabla^2$, and written in the form |
39 |
|
|
\begin{equation*} |
40 |
jgs |
107 |
\nabla^2 u = \nabla^t \cot \nabla u = \frac{\partial^2 u}{\partial x\hackscore 0^2} |
41 |
jgs |
102 |
+ \frac{\partial^2 u}{\partial x\hackscore 1^2} |
42 |
|
|
\end{equation*} |
43 |
|
|
and \eqn{eq:FirstSteps.1} as |
44 |
|
|
\begin{equation*} |
45 |
|
|
-\nabla^2 u = f |
46 |
|
|
\end{equation*} |
47 |
|
|
} |
48 |
jgs |
107 |
Basically, in the subindex of a function, any index to the left of the comma denotes a spatial derivative with respect |
49 |
jgs |
102 |
to the index. To get a more compact form we will write $w\hackscore{,ij}=(w\hackscore {,i})\hackscore{,j}$ |
50 |
|
|
which leads to |
51 |
|
|
\begin{equation} |
52 |
|
|
\Delta u = u\hackscore{,00}+u\hackscore{,11}=\sum\hackscore{i=0}^2 u\hackscore{,ii} |
53 |
|
|
\label{eq:FirstSteps.1.1b} |
54 |
|
|
\end{equation} |
55 |
|
|
In some cases, and we will see examples for this in the next chapter, |
56 |
|
|
the usage of the nested $\sum$ symbols blows up the formulas and therefore |
57 |
jgs |
107 |
it is convenient to use the Einstein summation convention \index{summation convention}. This |
58 |
|
|
drops the $\sum$ sign and assumes that a summation over a repeated index is performed |
59 |
jgs |
102 |
("repeated index means summation"). For instance we write |
60 |
|
|
\begin{eqnarray} |
61 |
|
|
x\hackscore{i}y\hackscore{i}=\sum\hackscore{i=0}^2 x\hackscore{i}y\hackscore{i} \\ |
62 |
|
|
x\hackscore{i}u\hackscore{,i}=\sum\hackscore{i=0}^2 x\hackscore{i}u\hackscore{,i} \\ |
63 |
|
|
u\hackscore{,ii}=\sum\hackscore{i=0}^2 u\hackscore{,ii} \\ |
64 |
jgs |
107 |
x\hackscore{ij}u\hackscore{i,j}=\sum\hackscore{j=0}^2\sum\hackscore{i=0}^2 x\hackscore{ij}u\hackscore{i,j} \\ |
65 |
jgs |
102 |
\label{eq:FirstSteps.1.1c} |
66 |
|
|
\end{eqnarray} |
67 |
|
|
With the summation convention we can write the Poisson equation \index{Poisson equation} as |
68 |
|
|
\begin{equation} |
69 |
|
|
- u\hackscore{,ii} =1 |
70 |
|
|
\label{eq:FirstSteps.1.sum} |
71 |
|
|
\end{equation} |
72 |
|
|
On the boundary of the domain $\Omega$ the normal derivative $n\hackscore{i} u\hackscore{,i}$ |
73 |
|
|
of the solution $u$ shall be zero, ie. $u$ shall fulfill |
74 |
|
|
the homogeneous Neumann boundary condition\index{Neumann |
75 |
|
|
boundary condition!homogeneous} |
76 |
|
|
\begin{equation} |
77 |
|
|
n\hackscore{i} u\hackscore{,i}= 0 \;. |
78 |
|
|
\label{eq:FirstSteps.2} |
79 |
|
|
\end{equation} |
80 |
|
|
$n=(n\hackscore{i})$ denotes the outer normal field |
81 |
|
|
of the domain, see \fig{fig:FirstSteps.1}. Remember that we |
82 |
|
|
are applying the Einstein summation convention \index{summation convention}, i.e |
83 |
jgs |
107 |
$n\hackscore{i} u\hackscore{,i}= n\hackscore{0} u\hackscore{,0} + |
84 |
|
|
n\hackscore{1} u\hackscore{,1}$. |
85 |
jgs |
102 |
\footnote{Some readers may familiar with the notation |
86 |
|
|
\begin{equation*} |
87 |
|
|
\frac{\partial u}{\partial n} = n\hackscore{i} u\hackscore{,i} |
88 |
|
|
\end{equation*} |
89 |
|
|
for the normal derivative.} |
90 |
|
|
The Neumann boundary condition of \eqn{eq:FirstSteps.2} should be fulfilled on the |
91 |
|
|
set $\Gamma^N$ which is the top and right edge of the domain: |
92 |
|
|
\begin{equation} |
93 |
|
|
\Gamma^N=\{(x\hackscore 0;x\hackscore 1) \in \Omega | x\hackscore{0}=1 \mbox{ or } x\hackscore{1}=1 \} |
94 |
|
|
\label{eq:FirstSteps.2b} |
95 |
|
|
\end{equation} |
96 |
jgs |
107 |
On the bottom and the left edge of the domain which is defined |
97 |
jgs |
102 |
as |
98 |
|
|
\begin{equation} |
99 |
|
|
\Gamma^D=\{(x\hackscore 0;x\hackscore 1) \in \Omega | x\hackscore{0}=0 \mbox{ or } x\hackscore{1}=0 \} |
100 |
|
|
\label{eq:FirstSteps.2c} |
101 |
|
|
\end{equation} |
102 |
|
|
the solution shall be identically zero: |
103 |
|
|
\begin{equation} |
104 |
|
|
u=0 \; . |
105 |
|
|
\label{eq:FirstSteps.2d} |
106 |
|
|
\end{equation} |
107 |
jgs |
107 |
This kind of boundary condition is called a homogeneous Dirichlet boundary condition |
108 |
jgs |
102 |
\index{Dirichlet boundary condition!homogeneous}. The partial differential equation in \eqn{eq:FirstSteps.1.sum} together |
109 |
|
|
with the Neumann boundary condition \eqn{eq:FirstSteps.2} and |
110 |
|
|
Dirichlet boundary condition in \eqn{eq:FirstSteps.2d} form a so |
111 |
|
|
called boundary value |
112 |
jgs |
107 |
problem\index{boundary value problem} (BVP\index{boundary value problem!BVP}) for |
113 |
|
|
the unknown |
114 |
jgs |
102 |
function $u$. |
115 |
|
|
|
116 |
|
|
|
117 |
|
|
\begin{figure} |
118 |
|
|
\centerline{\includegraphics[width=\figwidth]{FirstStepMesh}} |
119 |
|
|
\caption{Mesh of $4 \time 4$ elements on a rectangular domain. Here |
120 |
|
|
each element is a quadrilateral and described by four nodes, namely |
121 |
|
|
the corner points. The solution is interpolated by a bi-linear |
122 |
|
|
polynomial.} |
123 |
|
|
\label{fig:FirstSteps.2} |
124 |
|
|
\end{figure} |
125 |
|
|
|
126 |
|
|
In general the BVP\index{boundary value problem!BVP} cannot be solved analytically and numerical |
127 |
|
|
methods have to be used construct an approximation of the solution |
128 |
|
|
$u$. Here we will use the finite element method\index{finite element |
129 |
|
|
method} (FEM\index{finite element |
130 |
|
|
method!FEM}). The basic idea is to fill the domain with a |
131 |
jgs |
107 |
set of points called nodes. The solution is approximated by its |
132 |
jgs |
102 |
values on the nodes\index{finite element |
133 |
jgs |
107 |
method!nodes}. Moreover, the domain is subdivided into small, |
134 |
|
|
sub-domain called elements \index{finite element |
135 |
jgs |
102 |
method!element}. On each element the solution is |
136 |
|
|
represented by a polynomial of a certain degree through its values at |
137 |
|
|
the nodes located in the element. The nodes and its connection through |
138 |
|
|
elements is called a mesh\index{finite element |
139 |
jgs |
107 |
method!mesh}. \fig{fig:FirstSteps.2} shows an |
140 |
jgs |
102 |
example of a FEM mesh with four elements in the $x_0$ and four elements |
141 |
|
|
in the $x_1$ direction over the unit square. |
142 |
|
|
For more details we refer the reader to the literature, for instance |
143 |
|
|
\Ref{Zienc,NumHand}. |
144 |
|
|
|
145 |
|
|
\escript provides the class \Poisson to define a Poisson equation \index{Poisson equation}. |
146 |
|
|
(We will discuss a more general form of a PDE \index{partial differential equation!PDE} |
147 |
|
|
that can be defined through the \LinearPDE class later). The instantiation of |
148 |
|
|
a \Poisson class object requires the specification of the domain $\Omega$. In \escript |
149 |
|
|
the \Domain class objects are used to describe the geometry of a domain but it also |
150 |
|
|
contains information about the discretization methods and the actual solver which is used |
151 |
|
|
to solve the PDE. Here we are using the FEM library \finley \index{finite element |
152 |
|
|
method}. The following statements create the \Domain object \var{mydomain} from the |
153 |
|
|
\finley method \method{Rectangle} |
154 |
|
|
\begin{python} |
155 |
jgs |
107 |
from esys.finley import Rectangle |
156 |
|
|
mydomain = Rectangle(l0=1.,l1=1.,n0=40, n1=20) |
157 |
jgs |
102 |
\end{python} |
158 |
|
|
In this case the domain is a rectangle with the lower, left corner at point $(0,0)$ and |
159 |
|
|
the right, upper corner at $(\var{l0},\var{l1})=(1,1)$. |
160 |
jgs |
107 |
The arguments \var{n0} and \var{n1} define the number of elements in $x\hackscore{0}$ and |
161 |
jgs |
102 |
$x\hackscore{1}$-direction respectively. For more details on \method{Rectangle} and |
162 |
|
|
other \Domain generators within the \finley module, |
163 |
|
|
see \Chap{CHAPTER ON FINLEY}. |
164 |
|
|
|
165 |
jgs |
107 |
The following statements define the \Poisson class object \var{mypde} with domain \var{mydomain} and |
166 |
jgs |
102 |
the right hand side $f$ of the PDE to constant $1$: |
167 |
|
|
\begin{python} |
168 |
jgs |
107 |
from esys.escript import Poisson |
169 |
|
|
mypde = Poisson(mydomain) |
170 |
|
|
mypde.setValue(f=1) |
171 |
jgs |
102 |
\end{python} |
172 |
|
|
We have not specified any boundary condition but the |
173 |
|
|
\Poisson class implicitly assumes homogeneous Neuman boundary conditions \index{Neumann |
174 |
|
|
boundary condition!homogeneous} defined by \eqn{eq:FirstSteps.2}. With this boundary |
175 |
|
|
condition the BVP\index{boundary value problem!BVP} we have defined has no unique solution. In fact, with any solution $u$ |
176 |
|
|
and any constant $C$ the function $u+C$ becomes a solution as well. We have to add |
177 |
|
|
a Dirichlet boundary condition \index{Dirichlet boundary condition}. This is done |
178 |
jgs |
107 |
by defining a characteristic function \index{characteristic function} |
179 |
|
|
which has positive values at locations $x=(x\hackscore{0},x\hackscore{1})$ where Dirichlet boundary condition is set |
180 |
jgs |
102 |
and $0$ elsewhere. In our case of $\Gamma^D$ defined by \eqn{eq:FirstSteps.2c}, |
181 |
jgs |
107 |
we need a function which is positive for the cases $x\hackscore{0}=0$ or $x\hackscore{1}=0$: |
182 |
jgs |
102 |
\begin{python} |
183 |
|
|
x=mydomain.getX() |
184 |
|
|
gammaD=x[0].whereZero()+x[1].whereZero() |
185 |
|
|
\end{python} |
186 |
jgs |
107 |
In the first statement, the method \method{getX} of the \Domain \var{mydomain} |
187 |
|
|
gives access to locations |
188 |
|
|
in the domain defined by \var{mydomain}. The object \var{x} is actually a \Data object |
189 |
|
|
which we will learn more about later. \code{x[0]} returns the $x\hackscore{0}$ coordinates of the locations and |
190 |
jgs |
102 |
\code{x[0].whereZero()} creates function which equals $1$ where \code{x[0]} is (nearly) equal to zero |
191 |
jgs |
107 |
and $0$ elsewhere. |
192 |
|
|
Similarly, \code{x[1].whereZero()} creates function which equals $1$ where \code{x[1]} is |
193 |
|
|
equal to zero and $0$ elsewhere. |
194 |
|
|
The sum of the results of \code{x[0].whereZero()} and \code{x[1].whereZero()} gives a function on the domain \var{mydomain} which is exactly positive where $x\hackscore{0}$ or $x\hackscore{1}$ is equal to zero. |
195 |
jgs |
102 |
|
196 |
jgs |
107 |
The additional parameter \var{q} of the \code{setValue} method of the \Poisson class defines the |
197 |
jgs |
102 |
characteristic function \index{characteristic function} of the locations |
198 |
|
|
of the domain where homogeneous Dirichlet boundary condition \index{Dirichlet boundary condition!homogeneous} |
199 |
|
|
are set. The complete definition of our example is now: |
200 |
|
|
\begin{python} |
201 |
jgs |
107 |
from esys.linearPDEs import Poisson |
202 |
jgs |
102 |
x = mydomain.getX() |
203 |
|
|
gammaD = x[0].whereZero()+x[1].whereZero() |
204 |
jgs |
107 |
mypde = Poisson(domain=mydomain) |
205 |
|
|
mypde = setValue(f=1,q=gammaD) |
206 |
jgs |
102 |
\end{python} |
207 |
jgs |
107 |
The first statement imports the \Poisson class definition form the \linearPDEsPack module which is part of the \ESyS package. |
208 |
|
|
To get the solution of the Poisson equation defined by \var{mypde} we just have to call its |
209 |
jgs |
102 |
\method{getSolution}. |
210 |
|
|
|
211 |
|
|
Now we can write the script to solve our test problem (Remember that |
212 |
jgs |
107 |
lines starting with '\#' are comment lines in Python) (available as \file{mypoisson.py} |
213 |
jgs |
102 |
in the \ExampleDirectory): |
214 |
|
|
\begin{python} |
215 |
jgs |
107 |
from esys.finley import Rectangle |
216 |
jgs |
102 |
from esys.linearPDEs import Poisson |
217 |
|
|
# generate domain: |
218 |
jgs |
107 |
mydomain = Rectangle(l0=1.,l1=1.,n0=40, n1=20) |
219 |
jgs |
102 |
# define characteristic function of Gamma^D |
220 |
|
|
x = mydomain.getX() |
221 |
|
|
gammaD = x[0].whereZero()+x[1].whereZero() |
222 |
|
|
# define PDE and get its solution u |
223 |
|
|
mypde = Poisson(domain=mydomain,f=1,q=gammaD) |
224 |
|
|
u = mypde.getSolution() |
225 |
|
|
# write u to an external file |
226 |
|
|
u.saveDX("u.dx") |
227 |
|
|
\end{python} |
228 |
|
|
The last statement writes the solution to the external file \file{u.dx} in |
229 |
|
|
\OpenDX file format. \OpenDX is a software package |
230 |
|
|
for the visualization of scientific, engineering and analytical data and is freely available |
231 |
|
|
from \url{http://www.opendx.org}. |
232 |
|
|
|
233 |
|
|
\begin{figure} |
234 |
|
|
\centerline{\includegraphics[width=\figwidth]{FirstStepResult.eps}} |
235 |
jgs |
107 |
\caption{\OpenDX Visualization of the Possion Equation Solution for $f=1$} |
236 |
jgs |
102 |
\label{fig:FirstSteps.3} |
237 |
|
|
\end{figure} |
238 |
|
|
|
239 |
jgs |
107 |
You can edit the script file using your favourite text editor (or the Integrated DeveLopment Environment IDLE |
240 |
|
|
for Python, see \url{http://idlefork.sourceforge.net}). If the script file has the name \file{mypoisson.py} \index{scripts!\file{mypoisson.py}} you can run the |
241 |
jgs |
102 |
script from any shell using the command: |
242 |
|
|
\begin{verbatim} |
243 |
|
|
python mypoisson.py |
244 |
|
|
\end{verbatim} |
245 |
|
|
After the script has (hopefully successfully) been completed you will find the file \file{u.dx} in the current |
246 |
|
|
directory. An easy way to visualize the results is the command |
247 |
|
|
\begin{verbatim} |
248 |
jgs |
107 |
dx -prompter & |
249 |
jgs |
102 |
\end{verbatim} |
250 |
jgs |
107 |
to start the generic data visualization interface of \OpenDX. \fig{fig:FirstSteps.3} shows the result. |