C2999_PDF_Intro.pdf

(1005 KB) Pobierz
Introduction
Some Denitions,
Formulas, Methods, and Solutions
0.1. Classication of Second-Order Partial Differential
Equations
0.1.1. Equations with Two Independent Variables
0.1.1-1. Examples of equations encountered in applications.
Three basic types of partial differential equations are distinguished—parabolic,
hyperbolic,
and
elliptic.
The solutions of the equations pertaining to each of the types have their own characteristic
qualitative differences.
The simplest example of a
parabolic
equation is the heat equation
£  
¡
where the variables and play the role of time and the spatial coordinate, respectively. Note that
equation (1) contains only one highest derivative term.
The simplest example of a
hyperbolic
equation is the wave equation
£  
2
where the variables and play the role of time and the spatial coordinate, respectively. Note that
the highest derivative terms in equation (2) differ in sign.
The simplest example of an
elliptic
equation is the Laplace equation
¤  
2
+
where and play the role of the spatial coordinates. Note that the highest derivative terms in
equation (3) have like signs.
Any linear partial differential equation of the second-order with two independent variables can
be reduced, by appropriate manipulations, to a simpler equation which has one of the three highest
derivative combinations specied above in examples (1), (2), and (3).
0.1.1-2. Types of equations. Characteristic equations.
Consider a second-order partial differential equation with two independent variables which has the
general form
¤  
£  
¨
¤  
§
¤   £  
¦
£  
( , )
¥
2
2
, , ,
© 2002 by Chapman & Hall/CRC
¡
¤
£
¤
£
+2
( , )
¤
£
+
( , )
=
,
¡  
¡  
©
¡
 
¡
 
¡
2
 
2
 
¡
2
£  
 
2
¡
 
¡
2
¢  
 
2
¡
2
 
¢  
£
£
¢
¢
¤
£
¤
£
¡  
2
= 0,
(1)
2
= 0,
(2)
2
= 0,
(3)
2
,
(4)
where , , are some functions of and that have continuous derivatives up to the second-order
inclusive.*
Given a point ( , ), equation (4) is said to be
2
¦
§
¥
¦
¤
£
§
¦
¥
parabolic
hyperbolic
elliptic
at this point.
In order to reduce equation (4) to a canonical form, one should rst write out the characteristic
equation
2
2
−2
+
= 0,
which splits into two equations
£
§
¥
¦
¦
£
§
¤
£
¦
¤
¥
+
2
¥
and
and nd their general integrals.
¤
¥
In this case, equations (5) and (6) coincide and have a common general integral,
( , )
=
¤
£
.
in accordance with the relations
¤
£
By passing from ,
to new independent variables ,
=
( , ),
¤
£
=
( , ),
where
=
( , ) is any twice differentiable function that satises the condition of nondegeneracy
(
of the Jacobian
( ,, ))
in the given domain, we reduce equation (4) to the canonical form
 
 
¨
As , one can take
=
or
=
.
It is apparent that, just as the heat equation (1), the transformed equation (7) has only one
highest-derivative term.
In the degenerate case where the function
1
does not depend on the derivative
,
equation (7) is an ordinary differential equation for the variable , in which serves as a parameter.
¥
The general integrals
of equations (5) and (6) are real and different. These integrals determine two different families of
real characteristics.
* The right-hand side of equation (4) may be nonlinear. The classication and the procedure of reducing such equations
to a canonical form are only determined by the left-hand side of the equation.
© 2002 by Chapman & Hall/CRC
4
1
,
¤
£
( , )
=
¤
£
( , )
=
§
¦
0.1.1-4. Canonical form of hyperbolic equations (case
2
>
0).
2
¡
!
 
¨
2
1
, , ,
¡
=
,
¡  
¡  
©
¡
2
 
 
§
¥
¦
0.1.1-3. Canonical form of parabolic equations (case
2
£
§
2
¦
§
¥
¦
if
2
§
¥
¦
¤
¤
£
if
if
2
= 0,
>
0,
<
0
= 0,
= 0,
(5)
(6)
¥
¤
¤
£
£
$ #
" !
¤
£
3 120 )
(&
'
%
= 0).
.
(7)
This is the so-called rst canonical form of a hyperbolic equation.
The transformation
= +
,
= −
brings the above equation to another canonical form,
 
¢  
5
¨
 
©
where
3
= 4
2
. This is the so-called second canonical form of a hyperbolic equation. Apart from
notation, the left-hand side of the last equation coincides with that of the wave equation (2).
¥
In this case the general integrals of equations (5) and (6) are complex conjugate; these determine
two families of complex characteristics.
Let the general integral of equation (5) have the form
6
¤
£
46
( , )
+
¤
£
( , )
=
,
we reduce equation (4) to the canonical form
 
 
¨
 
Apart from notation, the left-hand side of the last equation coincides with that of the Laplace
equation (3).
0.1.2. Equations with Many Independent Variables
¡  
¡  
©
¡
where
are some functions that have continuous derivatives with respect to all variables to the
second-order inclusive, and
x
= {
1
, ,
}.
[The right-hand side of equation (8) may be nonlinear.
The left-hand side only is required for the classication of this equation.]
At a point
x
=
x
0
, the following quadratic form is assigned to equation (8):
C
B
C
B ¥
=
B
D
,
=1
C
© 2002 by Chapman & Hall/CRC
@
A
C
,
=1
C
1
(x
0
)
.
@ £  
998
8 8
£  
¨
@ £
£  
B
9898
8
£  
(x)
C
x,
,
¡
2
=
,
,
,
8 8
998
7
Consider a second-order partial differential equation with
has the form
 
B ¥
@
A
B
independent variables
1
,
,
@ £
£
2
2
4
, , ,
¡
+
=
,
¡  
¡  
©
¡
 
¡
2
 
 
2
¤
£
4
=
( , ),
¤
£
=
( , ),
where ( , ) and ( , ) are real-valued functions.
By passing from , to new independent variables ,
¤
£
4
§
¦
0.1.1-5. Canonical form of elliptic equations (case
2
<
0).
2
= −1,
in accordance with the relations
5
5
2
2
3
, , ,
¡
¢
=
,
¡  
¡  
¡
 
¡
2
¢  
 
2
 
5
¢
 
2
¨
, , ,
¡
=
,
¡  
¡  
©
5
¢
¡
 
 
we reduce equation (4) to
 
2
¤
£
4
=
( , ),
¤
£
=
( , ),
By passing from ,
£
¤
¤
£
£
to new independent variables ,
in accordance with the relations
.
,
¨
¤
£
C
B ¥
¨
.
that
(8)
(9)
TABLE 1
Classication of equations with many independent variables
Type of equation (8) at a point
x
=
x
0
Parabolic (in the broad sense)
Hyperbolic (in the broad sense)
Elliptic
Coefcients of the canonical form (11)
B §
G
@
At least one coefcient of the
B §
B §
is zero
All
are nonzero and some
B §
differ in sign
All
are nonzero and have like signs
By an appropriate linear nondegenerate transformation
7
8 8
998
6
B
E
B §
¤
¨
=
the quadratic form (9) can be reduced to the canonical form
B B §
=
D
=1
where the coefcients assume the values
1, −1,
and
0.
The number of negative and zero coefcients
in (11) does not depend on the way in which the quadratic form is reduced to the canonical form.
Table 1
presents the basic criteria according to which the equations with many independent
variables are classied.
Suppose all coefcients of the highest derivatives in (8) are constant,
=
const. By introducing
the
form
B
G
are the coefcients of the linear transformation (10), we reduce equation (8) to the canonical
@ ¤  
8 8
998
¤  
=1
Here, the coefcients are the same as in the quadratic form (11), and
y
= {
1
, ,
}.
Among the parabolic equations, it is conventional to distinguish the parabolic
equations in the narrow sense, i.e., the equations for which only one of the coefcients, , is zero,
while the other is the same, and in this case the right-hand side of equation (12) must contain the
rst-order partial derivative with respect to .
In turn, the hyperbolic equations are divided into normal hyperbolic equations—
for which all but one have like signs—and ultrahyperbolic equations—for which there are two or
more positive and two or more negative .
Specic equations of parabolic, elliptic, and hyperbolic types will be discussed further in
Subsection 0.2.
References for Section
0.1: V. M. Babich, M. B. Kapilevich, S. G. Mikhlin, et al. (1964), S. J. Farlow (1982), D. Colton
(1988), E. Zauderer (1989), A. N. Tikhonov and A. A. Samarskii (1990), I. G. Petrovsky (1991), W. A. Strauss (1992),
R. B. Guenther and J. W. Lee (1996), D. Zwillinger (1998).
B §
B §
SQ
R
E
§
@ ¤
8
9898
¤
B §
B §
3 I
3 P
120 )
120 )
(&
'
(&
'
%
%
=1
2
B
1
y,
,
0.2. Basic Problems of Mathematical Physics
0.2.1. Initial and Boundary Conditions. Cauchy Problem.
Boundary Value Problems
Every equation of mathematical physics governs innitely many qualitatively similar phenomena
or processes. This follows from the fact that differential equations have innitely many particular
© 2002 by Chapman & Hall/CRC
¡
=
1
,
,
¡  
¡  
©
¡
2
 
.
B
B
8 8
998
the new independent variables
@
E
1
,
,
in accordance with the formulas
£
H
=
E
E
E
¤
C
B ¥
@
B
A
G
=1
E
E
@ ¤
@
FEA
(
= 1,
, )
(10)
¤   B §
B
¤
B
A
B §
2
,
(11)
, where
(12)
solutions. The specic solution that describes the physical phenomenon under study is separated
from the set of particular solutions of the given differential equation by means of the initial and
boundary conditions.
Throughout this section, we consider linear equations in the -dimensional Euclidean space
or in an open domain
(exclusive of the boundary) with a sufciently smooth boundary
=
.
0.2.1-1. Parabolic equations. Initial and boundary conditions.
In general, a linear second-order partial differential equation of the parabolic type with independent
variables can be written as
x,
[ ]
=
(x, ),
(1)
¡  
¡
7
¡  
7
@
T
V
U
U
 
@
W
T
,
=1
C
=1
Parabolic equations govern unsteady thermal, diffusion, and other phenomena dependent on time .
Equation (1) is called homogeneous if (x, )
≡ 0.
Cauchy problem
(
≥ 0,
x
). Find a function that satises equation (1) for >
0
and the
initial condition
=
(x) at
= 0.
(3)
¢
¡
Boundary value problem*
(
≥ 0,
x
). Find a function
the initial condition (3), and the boundary condition
V
d
Y
that satises equation (1) for >
0,
( >
0).
¢
x,
In general,
x,
is a rst-order linear differential operator in the space variables
x
with coefcient de-
pendent on
x
and . The basic types of boundary conditions are described below in Subsection 0.2.2.
The initial condition (3) is called homogeneous if (x)
≡ 0.
The boundary condition (4) is called
homogeneous if (x, )
≡ 0.
0.2.1-2. Hyperbolic equations. Initial and boundary conditions.
Consider a second-order linear partial differential equation of the hyperbolic type with independent
variables of the general form
`
Y
X
¢  
7
b
¢
¢
d
Y
c
2
x,
where the linear differential operator
x,
is dened by (2). Hyperbolic equations govern unsteady
wave processes, which depend on time .
Equation (5) is said to be homogeneous if (x, )
≡ 0.
). Find a function that satises equation (5) for >
0
and the
Cauchy problem
(
≥ 0,
x
initial conditions
=
0
(x) at
= 0,
(6)
=
1
(x) at
= 0.
*
Boundary value problems
for parabolic and hyperbolic equations are sometimes called
mixed
or
initial-boundary value
problems.
¢
¢
¢
¡
¢
`
b
b
Y ¢
¡
¡
Y
X
 
@
T
V
¢
© 2002 by Chapman & Hall/CRC
¢
¡
+
(x, )
¢
¡  
¡
2
¢  
 
[ ]
=
(x, ),
W
[ ]
=
(x, ) at
x
¢
¡
(4)
(5)
¢
¢
a
B
B
a
C
B
¢
¡
C
¢
B ¥
,
(x, )
¢
b
`
U
B
V
¡
@ £
@
T
8998
8
x
= {
1
,
£
},
@
A
C
@
,
=1
C
=1
2
,
§
B
£  
B
£  
B
£  
A
C
¢
V
B
c
Y
x,
X
B ¦
(x, )
(x, )
>
0.
¡ ¢
¢
¢
B ¥
[ ]
¡
+
A
2
 
@
@
where
A
¢
`
¡
Y
X
¢  
+
(x, ) ,
(2)
¢
Zgłoś jeśli naruszono regulamin