You are on page 1of 88

Analytic Solutions of Partial Differential Equations

MATH3414
School of Mathematics, University of Leeds

15 credits
Taught Semester 1,
Year running 2003/04
Pre-requisites MATH2360 or MATH2420 or equivalent.
Co-requisites None.
Objectives: To provide an understanding of, and methods of solution for, the most important
types of partial differential equations that arise in Mathematical Physics. On completion
of this module, students should be able to: a) use the method of characteristics to solve
first-order hyperbolic equations; b) classify a second order PDE as elliptic, parabolic or
hyperbolic; c) use Greens functions to solve elliptic equations; d) have a basic understanding
of diffusion; e) obtain a priori bounds for reaction-diffusion equations.
Syllabus: The majority of physical phenomena can be described by partial differential equations (e.g. the Navier-Stokes equation of fluid dynamics, Maxwells equations of electromagnetism). This module considers the properties of, and analytical methods of solution for
some of the most common first and second order PDEs of Mathematical Physics. In particular, we shall look in detail at elliptic equations (Laplace?s equation), describing steady-state
phenomena and the diffusion / heat conduction equation describing the slow spread of concentration or heat. The topics covered are: First order PDEs. Semilinear and quasilinear
PDEs; method of characteristics. Characteristics crossing. Second order PDEs. Classification and standard forms. Elliptic equations: weak and strong minimum and maximum
principles; Greens functions. Parabolic equations: exemplified by solutions of the diffusion
equation. Bounds on solutions of reaction-diffusion equations.
Form of teaching
Lectures: 26 hours. 7 examples classes.
Form of assessment
One 3 hour examination at end of semester (100%).

ii
Details:
Evy Kersal
e
Office: 9.22e
Phone: 0113 343 5149
E-mail: kersale@maths.leeds.ac.uk
WWW: http://www.maths.leeds.ac.uk/kersale/
Schedule: three lectures every week, for eleven weeks (from 27/09 to 10/12).
Tuesday
13:0014:00 RSLT 03
Wednesday 10:0011:00 RSLT 04
Friday
11:0012:00 RSLT 06
Pre-requisite: elementary differential calculus and several variables calculus (e.g. partial
differentiation with change of variables, parametric curves, integration), elementary algebra (e.g. partial fractions, linear eigenvalue problems), ordinary differential equations (e.g.
change of variable, integrating factor), and vector calculus (e.g. vector identities, Greens
theorem).
Outline of course:
Introduction:
definitions
examples
First order PDEs:
linear & semilinear
characteristics
quasilinear
nonlinear
system of equations
Second order linear PDEs:
classification
elliptic
parabolic
Book list:
P. Prasad & R. Ravindran, Partial Differential Equations, Wiley Eastern, 1985.
W. E. Williams, Partial Differential Equations, Oxford University Press, 1980.
P. R. Garabedian, Partial Differential Equations, Wiley, 1964.
Thanks to Prof. D. W. Hughes, Prof. J. H. Merkin and Dr. R. Sturman for their lecture
notes.

Course Summary
Definitions of different type of PDE (linear, quasilinear, semilinear, nonlinear)
Existence and uniqueness of solutions
Solving PDEs analytically is generally based on finding a change of variable to transform
the equation into something soluble or on finding an integral form of the solution.

First order PDEs


a

Linear equations:

u
u
+b
= c.
x
y

change coordinate using (x, y), defined by the characteristic equation


dy
b
= ,
dx
a

and (x, y) independent (usually = x) to transform the PDE into an ODE.


Quasilinear equations:

change coordinate using the solutions of


dx
= a,
ds

dy
=b
ds

and

du
=c
ds

to get an implicit form of the solution (x, y, u) = F ((x, y, u)).


Nonlinear waves:

region of solution.

System of linear equations:

linear algebra to decouple equations.

Second order PDEs


a

2u
2u
u
2u
u
+
2b
+d
+
c
+e
+ f u = g.
2
2
x
xy
y
x
y
iii

iv

Classification

Type

Canonical form

b2 ac > 0

Hyperbolic

2u + . . . = 0

b2 ac = 0

Parabolic

2u + . . . = 0
2

Elliptic

2u + 2u + . . . = 0
2 2

b2

ac < 0

Characteristics
p
dy
b2 ac
b

=
a
dx
dy
= ab , = x (say)
dx
dy
= b
dx

p

b2 ac , = +
a
= i( )

Elliptic equations: (Laplace equation.) Maximum Principle. Solutions using Greens


functions (uses new variables and the Dirac -function to pick out the solution). Method of
images.
Parabolic equations: (heat conduction, diffusion equation.) Derive a fundamental solution in integral form or make use of the similarity properties of the equation to find the
solution in terms of the diffusion variable
x
= .
2 t
First and Second Maximum Principles and Comparison Theorem give bounds on the solution,
and can then construct invariant sets.

Contents
1 Introduction
1.1 Motivation . . . . . . . . . . . . . . . . . . . . . .
1.2 Reminder . . . . . . . . . . . . . . . . . . . . . . .
1.3 Definitions . . . . . . . . . . . . . . . . . . . . . . .
1.4 Examples . . . . . . . . . . . . . . . . . . . . . . .
1.4.1 Wave Equations . . . . . . . . . . . . . . .
1.4.2 Diffusion or Heat Conduction Equations . .
1.4.3 Laplaces Equation . . . . . . . . . . . . . .
1.4.4 Other Common Second Order Linear PDEs
1.4.5 Nonlinear PDEs . . . . . . . . . . . . . . .
1.4.6 System of PDEs . . . . . . . . . . . . . . .
1.5 Existence and Uniqueness . . . . . . . . . . . . . .
2 First Order Equations
2.1 Linear and Semilinear Equations . . . . . . .
2.1.1 Method of Characteristic . . . . . . .
2.1.2 Equivalent set of ODEs . . . . . . . .
2.1.3 Characteristic Curves . . . . . . . . .
2.2 Quasilinear Equations . . . . . . . . . . . . .
2.2.1 Interpretation of Quasilinear Equation
2.2.2 General solution: . . . . . . . . . . . .
2.3 Wave Equation . . . . . . . . . . . . . . . . .
2.3.1 Linear Waves . . . . . . . . . . . . . .
2.3.2 Nonlinear Waves . . . . . . . . . . . .
2.3.3 Weak Solution . . . . . . . . . . . . .
2.4 Systems of Equations . . . . . . . . . . . . . .
2.4.1 Linear and Semilinear Equations . . .
2.4.2 Quasilinear Equations . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

3 Second Order Linear and Semilinear Equations in Two Variables


3.1 Classification and Standard Form Reduction . . . . . . . . . . . . . .
3.2 Extensions of the Theory . . . . . . . . . . . . . . . . . . . . . . . .
3.2.1 Linear second order equations in n variables . . . . . . . . . .
3.2.2 The Cauchy Problem . . . . . . . . . . . . . . . . . . . . . . .
i

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

1
1
1
2
3
3
4
4
4
5
5
6

.
.
.
.
.
.
.
.
.
.
.
.
.
.

9
9
9
12
14
19
19
20
26
26
27
29
31
31
34

.
.
.
.

37
37
44
44
45

ii

CONTENTS

4 Elliptic Equations
4.1 Definitions . . . . . . . . . . . . . . . . . . . . . . .
4.2 Properties of Laplaces and Poissons Equations . .
4.2.1 Mean Value Property . . . . . . . . . . . .
4.2.2 Maximum-Minimum Principle . . . . . . . .
4.3 Solving Poisson Equation Using Greens Functions
4.3.1 Definition of Greens Functions . . . . . . .
4.3.2 Greens function for Laplace Operator . . .
4.3.3 Free Space Greens Function . . . . . . . .
4.3.4 Method of Images . . . . . . . . . . . . . .
4.4 Extensions of Theory: . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

49
49
50
51
52
54
55
55
60
61
68

5 Parabolic Equations
5.1 Definitions and Properties . . . . . . . . . . . . . . . . . . . . . . .
5.1.1 Well-Posed Cauchy Problem (Initial Value Problem) . . . .
5.1.2 Well-Posed Initial-Boundary Value Problem . . . . . . . . .
5.1.3 Time Irreversibility of the Heat Equation . . . . . . . . . .
5.1.4 Uniqueness of Solution for Cauchy Problem: . . . . . . . . .
5.1.5 Uniqueness of Solution for Initial-Boundary Value Problem:
5.2 Fundamental Solution of the Heat Equation . . . . . . . . . . . . .
5.2.1 Integral Form of the General Solution . . . . . . . . . . . .
5.2.2 Properties of the Fundamental Solution . . . . . . . . . . .
5.2.3 Behaviour at large t . . . . . . . . . . . . . . . . . . . . . .
5.3 Similarity Solution . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3.1 Infinite Region . . . . . . . . . . . . . . . . . . . . . . . . .
5.3.2 Semi-Infinite Region . . . . . . . . . . . . . . . . . . . . . .
5.4 Maximum Principles and Comparison Theorems . . . . . . . . . .
5.4.1 First Maximum Principle . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

69
69
69
70
70
71
71
72
73
74
75
75
76
77
78
79

A Integral of ex in R

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.

81

Chapter 1

Introduction
Contents
1.1
1.2
1.3
1.4
1.5

1.1

Motivation .
Reminder . .
Definitions . .
Examples . .
Existence and

. . . . . . . .
. . . . . . . .
. . . . . . . .
. . . . . . . .
Uniqueness

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

1
1
2
3
6

Motivation

Why do we study partial differential equations (PDEs) and in particular analytic


solutions?
We are interested in PDEs because most of mathematical physics is described by such equations. For example, fluids dynamics (and more generally continuous media dynamics), electromagnetic theory, quantum mechanics, traffic flow. Typically, a given PDE will only be
accessible to numerical solution (with one obvious exception exam questions!) and analytic solutions in a practical or research scenario are often impossible. However, it is vital
to understand the general theory in order to conduct a sensible investigation. For example,
we may need to understand what type of PDE we have to ensure the numerical solution is
valid. Indeed, certain types of equations need appropriate boundary conditions; without a
knowledge of the general theory it is possible that the problem may be ill-posed of that the
method is solution is erroneous.

1.2

Reminder

Partial derivatives: The differential (or differential form) of a function f of n independent


variables, (x1 , x2 , . . . , xn ), is a linear combination of the basis form (dx 1 , dx2 , . . . , dxn )
df =

n
X
f
f
f
f
dxi =
dx1 +
dx2 + . . . +
dxn ,
xi
x1
x2
xn
i=1

where the partial derivatives are defined by


f (x1 , x2 , . . . , xi + h, . . . , xn ) f (x1 , x2 , . . . , xi , . . . , xn )
f
= lim
.
xi h0
h
1

1.3 Definitions

The usual differentiation identities apply to the partial differentiations (sum, product, quotient, chain rules, etc.)
Notations: I shall use interchangeably the notations
f
xi f f xi ,
xi

2f
x2i xj f fxi xj ,
xi xj

for the first order and second order partial derivatives respectively. We shall also use interchangeably the notations
~u u u,
for vectors.
Vector differential operators: in three dimensional Cartesian coordinate system (i, j, k)
we consider f (x, y, z) : R3 R and [ux (x, y, z), uy (x, y, z), uz (x, y, z)] : R3 R3 .
Gradient: f = x f i + y f j + z f k.
Divergence: div u u = x ux + y uy + z uz.
Curl: u = (z uy y uz ) i + (z ux x uz ) j + (x uy y ux ) k.
Laplacian: f 2 f = x2 f + y2 f + z2 f.

Laplacian of a vector: u 2 u = 2 ux i + 2 uy j + 2 uz k.
Note that these operators are different in other systems of coordinate (cylindrical or spherical,
say)

1.3

Definitions

A partial differential equation (PDE) is an equation for some quantity u (dependent variable)
which depends on the independent variables x 1 , x2 , x3 , . . . , xn , n 2, and involves derivatives
of u with respect to at least some of the independent variables.
F (x1 , . . . , xn , x1 u, . . . , xn u, x21 u, x21 x2 u, . . . , xn1 ...xn u) = 0.
Note:
1. In applications xi are often space variables (e.g. x, y, z) and a solution may be required
in some region of space. In this case there will be some conditions to be satisfied on
the boundary ; these are called boundary conditions (BCs).
2. Also in applications, one of the independent variables can be time (t say), then there
will be some initial conditions (ICs) to be satisfied (i.e., u is given at t = 0 everywhere
in )
3. Again in applications, systems of PDEs can arise involving the dependent variables
u1 , u2 , u3 , . . . , um , m 1 with some (at least) of the equations involving more than one
ui .

Chapter 1 Introduction

The order of the PDE is the order of the highest (partial) differential coefficient in the
equation.
As with ordinary differential equations (ODEs) it is important to be able to distinguish
between linear and nonlinear equations.
A linear equation is one in which the equation and any boundary or initial conditions do not
include any product of the dependent variables or their derivatives; an equation that is not
linear is a nonlinear equation.
u
u
+c
= 0,
t
x

first order linear PDE (simplest wave equation),

2u 2u
+ 2 = (x, y),
x2
y

second order linear PDE (Poisson).

A nonlinear equation is semilinear if the coefficients of the highest derivative are functions of
the independent variables only.
(x + 3)

u
u
+ xy 2
= u3 ,
x
y

2
2u
u
u
2 u
+
(xy
+
y
)
+u
+ u2
= u4 .
2
2
x
y
x
y

A nonlinear PDE of order m is quasilinear if it is linear in the derivatives of order m with


coefficients depending only on x, y, . . . and derivatives of order < m.
"

1+

u
y

2 #

"
 2 # 2
u
2u
u
u u 2 u
2
+ 1+
= 0.
2
x
x y xy
x
y 2

Principle of superposition: A linear equation has the useful property that if u 1 and u2
both satisfy the equation then so does u 1 + u2 for any , R. This is often used in
constructing solutions to linear equations (for example, so as to satisfy boundary or initial
conditions; c.f. Fourier series methods). This is not true for nonlinear equations, which helps
to make this sort of equations more interesting, but much more difficult to deal with.

1.4
1.4.1

Examples
Wave Equations

Waves on a string, sound waves, waves on stretch membranes, electromagnetic waves, etc.
1 2u
2u
=
,
x2
c2 t2
or more generally
1 2u
= 2 u
c2 t2
where c is a constant (wave speed).

1.4.2

1.4 Examples

Diffusion or Heat Conduction Equations


u
2u
= 2,
t
x

or more generally
u
= 2 u,
t
or even
u
= (u)
t
where is a constant (diffusion coefficient or thermometric conductivity).
Both those equations (wave and diffusion) are linear equations and involve time (t). They
require some initial conditions (and possibly some boundary conditions) for their solution.

1.4.3

Laplaces Equation

Another example of a second order linear equation is the following.


2u 2u
+ 2 = 0,
x2
y
or more generally
2 u = 0.
This equation usually describes steady processes and is solved subject to some boundary
conditions.
One aspect that we shall consider is: why do the similar looking equations describes essentially
different physical processes? What is there about the equations that make this the cases?

1.4.4

Other Common Second Order Linear PDEs

Poissons equation is just the Lapaces equation (homogeneous) with a known source term
(e.g. electric potential in the presence of a density of charge):
2 u = .
The Helmholtz equation may be regarded as a stationary wave equation:
2 u + k 2 u = 0.
The Schrodinger equation is the fundamental equation of physics for describing quantum mechanical behavior; Schrodinger wave equation is a PDE that describes how the wavefunction
of a physical system evolves over time:
2 u + V u = i

u
.
t

Chapter 1 Introduction

1.4.5

Nonlinear PDEs

An example of a nonlinear equation is the equation for the propagation of reaction-diffusion


waves:
u
2u
=
+ u(1 u) (2nd order),
t
x2
or for nonlinear wave propagation:
u
u
+ (u + c)
= 0;
t
x

(1st order).

The equation
x2 u

u
u
+ (y + u)
= u3
x
y

is an example of quasilinear equation, and


y

u
u
+ (x3 + y)
= u3
x
y

is an example of semilinear equation.

1.4.6

System of PDEs

Maxwell equations constitute a system of linear PDEs:


E=

B = j +

B = 0,

E =

1 E
,
c2 t
B
.
t

In empty space (free of charges and currents) this system can be rearranged to give the
equations of propagation of the electromagnetic field,
2E
= c2 2 E,
t2

2B
= c2 2 B.
t2

Incompressible magnetohydrodynamic (MHD) equations combine Navier-Stokes equation (including the Lorentz force), the induction equation as well as the solenoidal constraints,
U
+ U U = + B B + 2 U + F,
t
B
= (U B) + 2 B,
t
U = 0,

B = 0.

Both systems involve space and time; they require some initial and boundary conditions for
their solution.

1.5 Existence and Uniqueness

1.5

Existence and Uniqueness

Before attempting to solve a problem involving a PDE we would like to know if a solution
exists, and, if it exists, if the solution is unique. Also, in problem involving time, whether
a solution exists t > 0 (global existence) or only up to a given value of t i.e. only for
0 < t < t0 (finite time blow-up, shock formation). As well as the equation there could be
certain boundary and initial conditions. We would also like to know whether the solution of
the problem depends continuously of the prescribed data i.e. small changes in boundary
or initial conditions produce only small changes in the solution.
Illustration from ODEs:
1.

du
= u,
dt

u(0) = 1.

Solution: u = et exists for 0 t <


2.

du
= u2 ,
dt

u(0) = 1.

Solution: u = 1/(1 t) exists for 0 t < 1


3.

du
= u,
dt

u(0) = 0,

has two solutions: u 0 and u = t2 /4 (non uniqueness).


We say that the PDE with boundary or initial condition is well-formed (or well-posed) if its
solution exists (globally), is unique and depends continuously on the assigned data. If any
of these three properties (existence, uniqueness and stability) is not satisfied, the problem
(PDE, BCs and ICs) is said to be ill-posed. Usually problems involving linear systems
are well-formed but this may not be always the case for nonlinear systems (bifurcation of
solutions, etc.)
Example: A simple example of showing uniqueness is provided by:
2 u = F in (Poissons equation).
with u = 0 on , the boundary of , and F is some given function of x.
Suppose u1 and u2 two solutions satisfying the equation and the boundary conditions. Then
consider w = u1 u2 ; 2 w = 0 in and w = 0 on . Now the divergence theorem gives
Z
Z
ww n dS =
(ww) dV,

Z

=
w2 w + (w)2 dV

where n is a unit normal outwards from .

Chapter 1 Introduction

(w)2

(w) dV =

w
dS = 0.
n

Now the integrand


is non-negative in and hence for the equality to hold we must
have w 0; i.e. w = constant in . Since w = 0 on and the solution is smooth, we
must have w 0 in ; i.e. u1 = u2 . The same proof works if u/n is given on or for
mixed conditions.

1.5 Existence and Uniqueness

Chapter 2

First Order Equations


Contents
2.1
2.2
2.3
2.4

2.1
2.1.1

Linear and Semilinear


Quasilinear Equations
Wave Equation . . . .
Systems of Equations

Equations
. . . . . . .
. . . . . . .
. . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

9
19
26
31

Linear and Semilinear Equations


Method of Characteristic

We consider linear first order partial differential equation in two independent variables:
a(x, y)

u
u
+ b(x, y)
+ c(x, y)u = f (x, y),
x
y

(2.1)

where a, b, c and f are continuous in some region of the plane and we assume that a(x, y)
and b(x, y) are not zero for the same (x, y).
In fact, we could consider semilinear first order equation (where the nonlinearity is present
only in the right-hand side) such as
a(x, y)

u
u
+ b(x, y)
= (x, y, u),
x
y

(2.2)

instead of a linear equation as the theory of the former does not require any special treatment
as compared to that of the latter.
The key to the solution of the equation (2.1) is to find a change of variables (or a change of
coordinates)
(x, y), (x, y)
which transforms (2.1) into the simpler equation
w
+ h(, )w = F (, )

where w(, ) = u(x(, ), y(, )).


9

(2.3)

10

2.1 Linear and Semilinear Equations

We shall define this transformation so that it is one-to-one, at least for all (x, y) in some set
D of points in the (x-y) plane. Then, on D we can (in theory) solve for x and y as functions
of , . To ensure that we can do this, we require that the Jacobian of the transformation
does not vanish in D:




x y
J = =

6= {0, }

x y y x
x

for (x, y) in D. We begin looking for a suitable transformation by computing derivatives via
the chain rule
w
w
u
w
w
u
=
+
and
=
+
.
x
x
x
y
y
y

We substitute these into equation (2.1) to obtain






w
w
w
w
+b
+ cw = f.
a
+
+
x
x
y
y
We can rearrange this as





w
a
+b
+ a
+b
+ cw = f.
x
y
x
y

(2.4)

This is close to the form of equation (2.1) if we can choose (x, y) so that
a

+b
=0
x
y

for

(x, y)

in

D.

Provided that /y 6= 0 we can express this required property of as


b
x
= .
y
a
Suppose we can define a new variable (or coordinate) which satisfies this constraint. What
is the equation describing the curves of constant ? Putting (x, y) = k (k an arbitrary
constant), then

d =
dx +
dy = 0
x
y
implies that dy/dx = x /y = b/a. So, the equation (x, y) = k defines solutions of the
ODE
b(x, y)
dy
=
.
(2.5)
dx
a(x, y)
Equation (2.5) is called the characteristic equation of the linear equation (2.1). Its solution
can be written in the form F (x, y, ) = 0 (where is the constant of integration) and defines a
family of curves in the plane called characteristics or characteristic curves of (2.1). (More on
characteristics later.) Characteristics represent curves along which the independent variable
of the new coordinate system (, ) is constant.
So, we have made the coefficient of w/ vanish in the transformed equation (2.4), by
choosing (x, y), with (x, y) = k an equation defining the solution of the characteristic
equation (2.5). We can now choose arbitrarily (or at least to suit our convenience), providing
we still have J 6= 0. An obvious choice is
(x, y) = x.

11

Chapter 2 First Order Equations


Then


1

J =


0
=
,

y

and we have already assumed this on-zero.


Now we see from equation (2.4) that this change of variables,
= x, (x, y),
transforms equation (2.1) to
(x, y)

w
+ c(x, y)w = f (x, y).,

where = a/x + b/y. To complete the transformation to the form of equation (2.3),
we first write (x, y), c(x, y) and f (x, y) in terms of and to obtain
A(, )

w
+ C(, )w = (, ).

Finally, restricting the variables to a set in which A(, ) 6= 0 we have


w C

+ w= ,

A
A
which is in the form of (2.3) with
h(, ) =

C(, )
A(, )

and F (, ) =

(, )
.
A(, )

The characteristic method applies to first order semilinear equation (2.2) as well as linear
equation (2.1); similar change of variables and basic algebra transform equation (2.2) to
K
w
= ,

A
where the nonlinear term K(, , w) = (x, y, u) and restricting again the variables to a set
in which A(, ) = (x, y) 6= 0.
Notation: It is very convenient to use the function u in places where rigorously the function
w should be used. E.g., the equation here above can identically be written as u/ = K/A.
Example: Consider the linear first order equation
x2

u
u
+y
+ xyu = 1.
x
y

This is equation (2.1) with a(x, y) = x 2 , b(x, y) = y, c(x, y) = xy and f (x, y) = 1. The
characteristic equation is
b
y
dy
= = 2.
dx
a
x

12

2.1 Linear and Semilinear Equations

Solve this by separation of variables


Z
Z
1
1
1
dy =
dx ln y + = k,
2
y
x
x

for y > 0,

and x 6= 0.

This is an integral of the characteristic equation describing curves of constant and so we


choose
1
(x, y) = ln y + .
x
Graphs of ln y +1/x are the characteristics of this PDE. Choosing = x we have the Jacobian
J=

1
= 6= 0
y
y

as required.

Since = x,
= ln y +

1
y = e1/ .

Now we apply the transformation = x, = ln y + 1/x with w(, ) = u(x, y) and we have


w
w
w w
1
1 w
w
u
=
+
=
+
2 =
2
,
x
x
x


w
w
w 1
1 w
u
=
+
=0+
= 1/
.
y
y
y
y

e
Then the PDE becomes


1 w
1 w
2 w

+ e1/ 1/
2
+ e1/ w = 1,

e
which simplifies to
2

w
+ e1/ w = 1

then to

w 1 1/
1
+ e
w = 2.

We have transformed the equation into the form of equation (2.3), for any region of (, )space with 6= 0.

2.1.2

Equivalent set of ODEs

The point of this transformation is that we can solve equation (2.3). Think of
w
+ h(, )w = F (, )

as a linear first order ordinary differential equation in , with carried along as a parameter.
Thus we use an integrating factor method
e

h(,) d

R
w
+ h(, ) e

h(,) d

w = F (, ) e

R
 R h(,) d 
e
w = F (, ) e

h(,) d

h(,) d

13

Chapter 2 First Order Equations

Now we integrate with respect to . Since is being carried as a parameter, the constant of
integration may depend on
Z
R
R
h(,) d
w = F (, ) e h(,) d d + g()
e
in which g is an arbitrary differentiable function of one variable. Now the general solution of
the transformed equation is
Z
R
R
R
h(,) d
w(, ) = e
F (, ) e h(,) d d + g() e h(,) d .
We obtain the general form of the original equation by substituting back (x, y) and (x, y)
to get
u(x, y) = e(x,y) [(x, y) + g((x, y))] .
(2.6)
A certain class of first order PDEs (linear and semilinear PDEs) can then be reduced to a
set of ODEs. This makes use of the general philosophy that ODEs are easier to solve than
PDEs.
Example: Consider the constant coefficient equation
a

u
u
+b
+ cu = 0
x
y

where a, b, c R. Assume a 6= 0, the characteristic equation is


dy/dx = b/a
with general solution defined by the equation
bx ay = k, k

constant.

So the characteristics of the PDE are the straight line graphs of bx ay = k and we make
the transformation with
= x, = bx ay.

Using the substitution we find the equation transforms to


c
w
+ w = 0.

a
The integrating factor method gives
 c/a 
e
w =0

and integrating with respect to gives

ec/a w = g(),
where g is any differentiable function of one variable. Then
w = g() ec/a
and in terms of x and y we back transform
u(x, y) = g(bx ay) ecx/a .

14

2.1 Linear and Semilinear Equations

Exercise:

Verify the solution by substituting back into the PDE.

Note: Consider the difference between general solution for linear ODEs and general solution
for linear PDEs. For ODEs, the general solution of
dy
+ q(x)y = p(x)
dx
contains an arbitrary constant of integration. For different constants you get different curves
of solution in (x-y)-plane. To pick out a unique solution you use some initial condition (say
y(x0 ) = y0 ) to specify the constant.
For PDEs, if u is the general solution to equation (2.1), then z = u(x, y) defines a family of
integral surfaces in 3D-space, each surface corresponding to a choice of arbitrary function g
in (2.6). We need some kind of information to pick out a unique solution; i.e., to chose the
arbitrary function g.

2.1.3

Characteristic Curves

We investigate the significance of characteristics which, defined by the ODE


b(x, y)
dy
=
,
dx
a(x, y)
represent a one parameter family of curves whose tangent at each point is in the direction of
the vector e = (a, b). (Note that the left-hand side of equation (2.2) is the derivation of u in
the direction of the vector e, e u.) Their parametric representation is (x = x(s), y = y(s))
where x(s) and y(s) satisfy the pair of ODEs
dx
= a(x, y),
ds

dy
= b(x, y).
ds

(2.7)

The variation of u with respect x = along these characteristic curves is given by


du
u dy u
u b u
=
+
=
+
,
dx
x dx y
x a y
(x, y, u)
from equation (2.2),
=
a(x, y)
such that, in term of the curvilinear coordinate s, the variation of u along the curves becomes
du dx
du
=
= (x, y, u).
ds
dx ds
The one parameter family of characteristic curves is parameterised by (each value of
represents one unique characteristic). The solution of equation (2.2) reduces to the solution
of the family of ODEs


du
du
du
(x, y, u)
(2.8)
= (x, y, u)
or similarly
=
=
ds
dx
d
a(x, y)
along each characteristics (i.e. for each value of ).
Characteristic equations (2.7) have to be solved together with equation (2.8), called the
compatibility equation, to find a solution to semilinear equation (2.2).

15

Chapter 2 First Order Equations

Cauchy Problem: Consider a curve in (x, y)-plane whose parametric form is (x =


x0 (), y = y0 ()). The Cauchy problem is to determine a solution of the equation
F (x, y, u, x u, y u) = 0
in a neighbourhood of such that u takes prescribed values u 0 () called Cauchy data on .
y

=cst
=cst

uo(1)

uo(2)

(xo, yo )

Notes:
1. u can only be found in the region between the characteristics drawn through the endpoint of .
2. Characteristics are curves on which the values of u combined with the equation are not
sufficient to determine the normal derivative of u.
3. A discontinuity in the initial data propagates onto the solution along the characteristics.
These are curves across which the derivatives of u can jump while u itself remains
continuous.
Existence & Uniqueness: Why do some choices of in (x, y)-space give a solution and
other give no solution or an infinite number of solutions? It is due to the fact that the Cauchy
data (initial conditions) may be prescribed on a curve which is a characteristic of the PDE.
To understand the definition of characteristics in the context of existence and uniqueness of
solution, return to the general solution (2.6) of the linear PDE:
u(x, y) = e(x,y) [(x, y) + g((x, y))] .
Consider the Cauchy data, u0 , prescribed along the curve whose parametric form is (x =
x0 (), y = y0 ()) and suppose u0 (x0 (), y0 ()) = q(). If is not a characteristic, the
problem is well-posed and there is a unique function g which satisfies the condition
q() = e(x0 (),y0 ()) [(x0 (), y0 ()) + g(x0 (), y0 ())] .
If on the other hand (x = x0 (), y = y0 ()) is the parametrisation of a characteristic
((x, y) = k, say), the relation between the initial conditions q and g becomes
q() = e(x0 (),y0 ()) [(x0 (), y0 ()) + G] ,

(2.9)

where G = g(k) is a constant; the problem is ill-posed. The functions (x, y) and (x, y)
are determined by the PDE, so equation (2.9) places a constraint on the given data function
q(x). If q() is not of this form for any constant G, then there is no solution taking on these
prescribed values on . On the other hand, if q() is of this form for some G, then there
are infinitely many such solutions, because we can choose for g any differentiable function so
that g(k) = G.

16
Example 1:

2.1 Linear and Semilinear Equations


Consider
2

u
u
+3
+ 8u = 0.
x
y

The characteristic equation is

dy
3
=
dx
2
and the characteristics are the straight line graphs 3x 2y = c. Hence we take = 3x 2y
and = x.
y
=cst
=cst

=cst

(We can see that an and cross only once they are independent, i.e. J 6= 0; and have
been properly chosen.)
This gives the solution
u(x, y) = e4x g(3x 2y)
where g is a differentiable function defined over the real line. Simply specifying the solution
at a given point (as in ODEs) does not uniquely determine g; we need to take a curve of
initial conditions.
Suppose we specify values of u(x, y) along a curve in the plane. For example, lets choose
as the x-axis and gives values of u(x, y) at points on , say
u(x, 0) = sin(x).
Then we need
u(x, 0) = e4x g(3x) = sin(x)

i.e.

g(3x) = sin(x)e4x ,

and putting t = 3x,


g(t) = sin(t/3) e 4t/3 .
This determines g and the solution satisfying the condition u(x, 0) = sin(x) on is
u(x, y) = sin(x 2y/3) e8y/3 .
We have determined the unique solution of the PDE with u specified along the x-axis.
We do not have to choose an axis say, along x = y, u(x, y) = u(x, x) = x 4 . From the
general solution this requires,
u(x, x) = e4x g(x) = x4 ,

so

g(x) = x4 e4x

to give the unique solution


satisfying u(x, x) = x4 .

u(x, y) = (3x 2y)4 e8(xy)

17

Chapter 2 First Order Equations

However, not every curve in the plane can be used to determine g. Suppose we choose to
be the line 3x 2y = 1 and prescribe values of u along this line, say
u(x, y) = u(x, (3x 1)/2) = x2 .
Now we must choose g so that
e4x g(3x (3x 1)) = x2 .
This requires g(1) = x2 e4x (for all x). This is impossible and hence there is no solution
taking the value x2 at points (x, y) on this line.
Last, we consider again to be the line 3x 2y = 1 but choose values of u along this line to
be
u(x, y) = u(x, (3x 1)/2) = e4x .
Now we must choose g so that

e4x g(3x (3x 1)) = e4x .


This requires g(1) = 1, condition satisfied by an infinite number of functions and hence there
is an infinite number of solutions taking the values e 4x on the line 3x 2y = 1.
Depending on the initial conditions, the PDE has one unique solution, no solution at all or
an infinite number or solutions. The difference is that the x-axis and the line y = x are not
the characteristics of the PDE while the line 3x 2y = 1 is a characteristic.
Example 2:
x

u
u
y
=u
x
y

with

u = x2

on

y = x, 1 y 2

Characteristics:

y
dy
= d(xy) = 0 xy = c,
dx
x
So, take = xy and = x. Then the equation becomes
xy

w
w
w
+x
xy
=w

constant.

w
w =0

w
= 0.

Finally the general solution is, w = g() or equivalently u(x, y) = x g(xy). When y = x

with 1 y 2, u = x2 ; so x2 = x g(x2 ) g(x) = x and the solution is

u(x, y) = x xy.

=const

=const

This figure presents the characteristic curves given by xy = constant. The red characteristics
show the domain where the initial conditions permit us to determine the solution.

18

2.1 Linear and Semilinear Equations

Alternative approach to solving example 2:


u
u
x
y
= u with u = x2 on y = x, 1 y 2
x
y
This method is not suitable for finding general solutions but it works for Cauchy problems.
The idea is to integrate directly the characteristic and compatibility equations in curvilinear coordinates. (See also alternative method for solving the characteristic equations for
quasilinear equations hereafter.)
The solution of the characteristic equations
dx
dy
=x
and
= y
ds
ds
gives the parametric form of the characteristic curves, while the integration of the compatibility equation
du
=u
ds
gives the solution u(s) along these characteristics curves.
The solution of the characteristic equations is
x = c 1 es

and y = c2 es ,

where the parametric form of the data curve permits us to find the two constants of
integration c1 & c2 in terms of the curvilinear coordinate along .
The curve is described by
x0 () =

and

y0 () =

with

[2, 1]

and we consider the points on to be the origin of the coordinate s along the characteristics
(i.e. s = 0 on ). So,




x0 = = c 1
x(s, ) = es
, [0, 1].
on (s = 0)

y0 = = c 2
y(s, ) = es

For linear or semilinear problems we can solve the compatibility equation independently of
the characteristic equations. (This property is not true for quasilinear equations.) Along the
characteristics u is determined by
du
= u u = c3 e s .
ds
Now we can make use of the Cauchy data to determine the constant of integration c 3 ,
on , at s = 0,

u0 (x0 (), y0 ()) u0 () = 2 = c3 .

Then, we have the parametric forms of the characteristic curves and the solution
x(s, ) = es ,

y(s, ) = es

and

u(s, ) = 2 es ,

in terms of two parameters, s the curvilinear coordinate along the characteristic curves and
the curvilinear coordinate along the data curve . From the two first ones we get s and
in terms of x and y.
r
x
x

2s
= e s = ln
and xy = 2 = xy ( 0).
y
y

Then, we substitute s and in u(s, ) to find


 r 
r
x
x

= xy
= x xy.
u(x, y) = xy exp ln
y
y

19

Chapter 2 First Order Equations

2.2

Quasilinear Equations

Consider the first order quasilinear PDE


a(x, y, u)

u
u
+ b(x, y, u)
= c(x, y, u)
x
y

(2.10)

where the functions a, b and c can involve u but not its derivatives.

2.2.1

Interpretation of Quasilinear Equation

We can represent the solutions u(x, y) by the integral surfaces of the PDE, z = u(x, y), in
(x, y, z)-space. Define the Monge direction by the vector (a, b, c) and recall that the normal to
the integral surface is (x u, y u, 1). Thus quasilinear equation (2.10) says that the normal
to the integral surface is perpendicular to the Monge direction; i.e. integral surfaces are
surfaces that at each point are tangent to the Monge direction,


x u
a
b y u = a(x, y, u) u + b(x, y, u) u c(x, y, u) = 0.
x
y
1
c

With the field of Monge direction, with direction numbers (a, b, c), we can associate the family
of Monge curves which at each point are tangent to that direction field. These are defined
by

dx
a
c dy b dz
dy
dz
dx
dy b = a dz c dx = 0
=
=
(= ds),
a(x, y, u)
b(x, y, u)
c(x, y, u)
dz
c
b dx a dy
where dl = (dx, dy, dz) is an arbitrary infinitesimal vector parallel to the Monge direction.
In the linear case, characteristics were curves in the (x, y)-plane (see 2.1.3). For the quasilinear equation, we consider Monge curves in (x, y, u)-space defined by
dx
= a(x, y, u),
ds
dy
= b(x, y, u),
ds
du
= c(x, y, u).
ds
Characteristic equations (d{x, y}/ds) and compatibility equation (du/ds) are simultaneous
first order ODEs in terms of a dummy variable s (curvilinear coordinate along the characteristics); we cannot solve the characteristic equations and compatibility equation independently
as it is for a semilinear equation. Note that, in cases where c 0, the solution remains
constant on the characteristics.
The rough idea in solving the PDE is thus to build up the integral surface from the Monge
curves, obtained by solution of the ODEs.
Note that we make the difference between Monge curve or direction in (x, y, z)-space and
characteristic curve or direction, their projections in (x, y)-space.

20

2.2.2

2.2 Quasilinear Equations

General solution:

Suppose that the characteristic and compatibility equations that we have defined have two
independent first integrals (function, f (x, y, u), constant along the Monge curves)
(x, y, u) = c1

and

(x, y, u) = c2 .

Then the solution of equation (2.10) satisfies F (, ) = 0 for some arbitrary function F
(equivalently, = G() for some arbitrary G), where the form of F (or G) depends on the
initial conditions.
Proof:

Since and are first integrals of the equation,


(x, y, u) = (x(s), y(s), u(s)),
= (s) = c1 .

We have the chain rule

dx dy du
d
=
+
+
= 0,
ds
x ds
y ds u ds

and then from the characteristic equations


a
And similarly for
a

+b
+c
= 0.
x
y
u

+b
+c
= 0.
x
y
u

Solving for c gives


a

or

x u
x u

a J[u, x] = b J[y, u]



+b
=0

y u
y u



x
x
where J[x1 , x2 ] = 1 2 .


x1

And similarly, solving for a,

x2

b J[x, y] = c J[u, x].

Thus, we have J[u, x] = J[x, y] b/c and J[y, u] = J[u, x] a/b = J[x, y] a/c.
Now consider F (, ) = 0 remember F ((x, y, u(x, y)), (x, y, u(x, y))) and differentiate
dF =

F
F
dx +
dy = 0
x
y

Then, the derivative with respect to x is zero,






F u
F
F u
=
+
+
+
= 0,
x
x u x
x
u x
as well as the derivative with respect to y




F u
F u
F
+
= 0.
=
+
+
y
y
u y
y
u y

21

Chapter 2 First Order Equations


For a non-trivial solution of this we must have

 



u
u
u
u
+
+
+
+

= 0,
x u x
y
u y
x
u x
y
u y




u

u

,
y u
y u x
u x
u x y
x y
x y
u
u
+ J[u, x]
= J[x, y].
J[y, u]
x
y
Then from the previous expressions for a, b, and c
a

u
u
+b
= c,
x
y

i.e., F (, ) = 0 defines a solution of the original equation.


Example 1:
(y + u)

u
u
+y
=xy
x
y

in

y > 0, < x < ,

with

u = 1+x

on

y = 1.

We first look for the general solution of the PDE before applying the initial conditions.
Combining the characteristic and compatibility equations,
dx
= y + u,
ds
dy
= y,
ds
du
=xy
ds

(2.11)
(2.12)
(2.13)

we seek two independent first integrals. Equations (2.11) and (2.13) give
d
(x + u) = x + u,
ds
and equation (2.12)
1 dy
= 1.
y ds
Now, consider
d
ds

x+u
y

1 d
x + u dy
(x + u)
,
y ds
y 2 ds
x+u x+u

= 0.
=
y
y
=

So, (x + u)/y = c1 is constant. This defines a family of solutions of the PDE; so, we can
choose
x+u
(x, y, u) =
,
y

22

2.2 Quasilinear Equations

such that = c1 determines one particular family of solutions. Also, equations (2.11)
and (2.12) give
d
(x y) = u,
ds
and equation (2.13)
d
du
(x y) (x y) = u .
ds
ds
Now, consider



d 
d
d 
(x y)2 u2 =
(x y)2
u2 ,
ds
ds
ds
d
du
= 2(x y) (x y) 2u
= 0.
ds
ds
Then, (x y)2 u2 = c2 is constant and defines another family of solutions of the PDE. So,
we can take
(x, y, u) = (x y)2 u2 .
The general solution is


x+u
F
, (x y)2 u2 = 0
y

or

(x y)2 u2 = G

x+u
y

for some arbitrary functions F or G.


Now to get a particular solution, apply initial conditions (u = 1 + x when y = 1)
(x 1)2 (x + 1)2 = G(2x + 1) G(2x + 1) = 4x.
Substitute = 2x + 1, i.e. x = ( 1)/2, so G() = 2(1 ). Hence,


x+u
2
2
2
(x y) u = 2 1
= (y x u).
y
y
We can regard this as a quadratic equation for u:


2
xy
2
2
+ (x y) = 0,
u u 2
y
y


2
1 2
1
2
u u xy+
+ 2 = 0.
y
y
y
Then,
1
u =
y





1 2
1
1
1
1
+ xy+
2 = xy+
.
y2
y
y
y
y

Consider again the initial condition u = 1 + x on y = 1


u (x, y = 1) = 1 (x 1 + 1) = 1 x take the positive root.
Hence,

2
u(x, y) = x y + .
y

23

Chapter 2 First Order Equations


Example 2:

using the same procedure solve

x(y u)

u
u
+ y(x + u)
= (x + y)u
x
y

with

u = x2 + 1

on

y = x.

Characteristic equations
dx
= x(y u),
ds
dy
= y(x + u),
ds
du
= (x + y)u.
ds

(2.14)
(2.15)
(2.16)

Again, we seek to independent first integrals. On the one hand, equations (2.14) and (2.15)
give
y

Now, consider

dy
dx
+x
= xy 2 xyu + yx2 + xyu = xy (x + y),
ds
ds
1 du
= xy
from equation (2.16).
u ds
1 du
1 dx 1 dy
+
=
x ds y ds
u ds

Hence, xy/u = c1 is constant and


(x, y, u) =

 xy 
d
= 0.
ln
ds
u
xy
u

is a first integral of the PDE. On the other hand,


dx dy
du

= xy xu xy yu = u(x + y) = ,
ds
ds
ds
d
(x + u y) = 0.
ds
Hence, x + u y = c2 is also a constant on the Monge curves and another first integral is
given by
(x, y, u) = x + u y,

so the general solution is

xy
= G(x + u y).
u
Now, we make use of the initial conditions, u = x 2 + 1 on y = x, to determine G:
x2
= G(x2 + 1);
1 + x2
set = x2 + 1, i.e. x2 = 1, then
G() =

1
,

and finally the solution is


x+uy1
xy
=
.
u
x+uy

Rearrange to finish!

24

2.2 Quasilinear Equations

Alternative approach:
x2

Solving the characteristic equations. Illustration by an example,

u
u
+u
= 1,
x
y

with

u=0

on x + y = 1.

The characteristic equations are


dx
= x2 ,
ds

dy
=u
ds

and

du
= 1,
ds

which we can solve to get


1
,
c1 s
s2
y=
+ c2 s + c 3 ,
2
u = c2 + s, for constants

(2.17)

x=

(2.18)
c 1 , c2 , c3 .

(2.19)

We now parameterise the initial line in terms of :


= x, y = 1 ,
and apply the initial data at s = 0. Hence,
(2.17)

gives

(2.18)

gives

(2.19)

gives

1
1
c1 = ,
c1

1 = c3 c3 = 1 ,

0 = c2 c2 = 0.

Hence, we found the parametric form of the surface integral,


x=
Eliminate s and ,

,
1 s
x=

then

y=

s2
+1
2

and u = s.

=
,
1 s
1 + sx

u2
x
+1
.
2
1 + ux
Invariants, or first integrals, are (from solution (2.17), (2.18) and (2.19), keeping arbitrary
c2 = 0) = u2 /2 y and = x/(1 + ux).
y=

Alternative approach to example 1:


(y + u)

u
u
+y
=xy
x
y

in

y > 0, < x < ,

with

u=1+x

on

y = 1.

Characteristic equations
dx
= y + u,
ds
dy
= y,
ds
du
= x y.
ds

(2.20)
(2.21)
(2.22)

25

Chapter 2 First Order Equations


Solve with respect to the dummy variable s; (2.21) gives,
y = c 1 es ,
(2.20) and (2.22) give
d
(x + u) = x + u x + u = c2 es ,
ds
and (2.20) give
dx
= c1 es + c2 es x,
ds
1
1
x = c3 es + (c1 + c2 ) es and u = c3 es + (c2 c1 ) es .
2
2
Now, at s = 0, y = 1 and x = , u = 1 + (parameterising initial line ),
so,

c1 = 1,

c2 = 1 + 2

and

c3 = 1.

Hence, the parametric form of the surface integral is,


x = es + (1 + ) es ,

y = es

and u = es + es .

Then eliminate and s:


1
1
x = + (1 + ) y =
y
y

1
xy+
y

so
1 1
u= +
y y
Finally,

1
xy+
y

2
u =xy+ ,
y

y.

as before.

To find invariants, return to solved characteristics equations and solve for constants in terms
of x, y and u. We only need two, so put for instance c 1 = 1 and so y = es . Then,
x=

c3 1
+ (1 + c2 ) y
y
2

Solve for c2
c2 =

x+u
,
y

and u =
so

c3 1
+ (c2 1) y.
y
2

x+u
,
y

and solve for c3

1
c3 = (x u y)y, so = (x u y)y.
2
Observe is different from last time, but this does not as we only require two independent
choices for and . In fact we can show that our previous is also constant,
(x y)2 u2 = (x y + u)(x y u),

= (y y) ,
y
= ( 1) which is also constant.

26

2.3 Wave Equation

Summary:

Solving the characteristic equations two approaches.

1. Manipulate the equations to get them in a directly integrable form, e.g.


1 d
(x + u) = 1
x + u ds
and find some combination of the variables which differentiates to zero (first integral),
e.g.


d x+u
= 0.
ds
y
2. Solve the equations with respect to the dummy variable s, and apply the initial data
(parameterised by ) at s = 0. Eliminate and s; find invariants by solving for
constants.

2.3

Wave Equation

We consider the equation


u
u
+ (u + c)
=0
t
x

with u(0, x) = f (x),

where c is some positive constant.

2.3.1

Linear Waves

If u is small (i.e. u2  u), then the equation approximate to the linear wave equation
u
u
+c
=0
t
x

with u(x, 0) = f (x).

The solution of the equation of characteristics, dx/dt = c, gives the first integral of the
PDE, (x, t) = x ct, and then general solution u(x, t) = g(x ct), where the function g is
determined by the initial conditions. Applying u(x, 0) = f (x) we find that the linear wave
equation has the solution u(x, t) = f (x ct), which represents a wave (unchanging shape)
propagating with constant wave speed c.
t

t
cs
t=

=
x
c

=
0

u(x,t)

f(x)

t=0

Note that u is constant where x ct =constant, i.e. on the characteristics.

t=t1

x1=ct1

27

Chapter 2 First Order Equations

2.3.2

Nonlinear Waves

For the nonlinear equation,

u
u
+ (u + c)
= 0,
t
x

the characteristics are defined by


dt
= 1,
ds

dx
=c+u
ds

and

du
= 0,
ds

which we can solve to give two independent first integrals = u

and = x (u + c)t. So,

u = f [x (u + c)t],
according to initial conditions u(x, 0) = f (x). This is similar to the previous result, but now
the wave speed involves u.
However, this form of the solution is not very helpful; it is more instructive to consider
the characteristic curves. (The PDE is homogeneous, so the solution u is constant along
the Monge curves this is not the case in general which can then be reduced to their
projections in the (x, t)-plane.) By definition, = x(c+u)t is constant on the characteristics
(as well as u); differentiate to find that the characteristics are described by
dx
= u + c.
dt
These are straight lines,
x = (f () + c)t + ,
expressed in terms of a parameter . (If we make use of the parametric form of the data curve
: {x = , t = 0, R} and solve directly the Cauchy problem in terms of the coordinate
s = t, we similarly find, u = f () and x = (u + c)t + .) The slope of the characteristics,
1/(c + u), varies from one line to another, and so, two curves can intersect.

()+
c)

t=

{x= ,t=0}

u=f(
) &

x(f

tmin

f()<0

Consider two crossing characteristics expressed in terms of 1 and 2 ,


i.e.

x = (f (1 ) + c)t + 1 ,
x = (f (2 ) + c)t + 2 .

(These correspond to initial values given at x = 1 and x = 2 .) These characteristics


intersect at the time
1 2
,
t=
f (1 ) f (2 )

28

2.3 Wave Equation

and if this is positive it will be in the region of solution. At this point u will not be singlevalued and the solution breaks down. By letting 2 1 we can see that the characteristics
intersect at
1
t= 0 ,
f ()
and the minimum time for which the solution becomes multi-valued is
tmin =

1
,
max[f 0 ()]

i.e. the solution is single valued (i.e. is physical) only for 0 t < t min . Hence, when f 0 () < 0
we can expect the solution to exist only for a finite time. In physical terms, the equation
considered is purely advective; in real waves, such as shock waves in gases, when very large
gradients are formed then diffusive terms (e.g. xx u) become vitally important.
u(x,t)
f(x)

multivalued

breaking

To illustrate finite time solutions of the nonlinear wave equation, consider


f () = (1 ),
0

f () = 1 2.

(0 1),

So, f 0 () < 0 for 1/2 < < 1 and we can expect the solution not to remain single-valued for
all values of t. (max[f 0 ()] = 1 so tmin = 1. Now,

so

u = f (x (u + c)t),

u = [x (u + c)t] [1 x + (u + c)t],

(ct x 1 + ct),

which we can express as


t2 u2 + (1 + t 2xt + 2ct2 )u + (x2 x 2ctx + ct + c2 t2 ) = 0,
and solving for u (we take the positive root from initial data)
u=
Now, at t = 1,


p
1 
2 4t(x ct) .
2t(x

ct)

(1
+
t)
+
(1
+
t)
2t2
u = x (c + 1) +

1 + c x,

so the solution becomes singular as t 1 and x 1 + c.

29

Chapter 2 First Order Equations

2.3.3

Weak Solution

When wave breaking occurs (multi-valued solutions) we must re-think the assumptions in
our model. Consider again the nonlinear wave equation,
u
u
+ (u + c)
= 0,
t
x
and put w(x, t) = u(x, t) + c; hence the PDE becomes the inviscid Burgers equation
w
w
+w
= 0,
t
x
or equivalently in a conservative form
w

+
t
x

w2
2

= 0,

where w2 /2 is the flux function. We now consider its integral form,


 
 
Z x2
Z x2 
Z
w2
w
w2
d x2
+
w(x, t) dx =
dx = 0
dx
t
x 2
dt x1
2
x1 x
x1
where x2 > x1 are real. Then,
Z
w2 (x1 , t) w2 (x2 , t)
d x2
w(x, t) dx =

.
dt x1
2
2
Let us now relax the assumption regarding the differentiability of the our solution; suppose
that w has a discontinuity in x = s(t) with x 1 < s(t) < x2 .
w(x,t)
w(s ,t)

w(s+,t)
x1

s(t)

x2

Thus, splitting the interval [x1 , x2 ] in two parts, we have


Z
Z
d x2
w2 (x1 , t) w2 (x2 , t)
d s(t)
w(x, t) dx +

=
w(x, t) dx,
2
2
dt x1
dt s(t)
Z x2
Z s(t)
w
w
dx w(s+ , t) s(t)
+
dx,
= w(s , t) s(t)
+
t
s(t) t
x1
where w(s (t), t) and w(s+ (t), t) are the values of w as x s from below and above respectively; s = ds/dt.
Now, take the limit x1 s (t) and x2 s+ (t). Since w/t is bounded, the two integrals
tend to zero. We then have

w2 (s , t) w2 (s+ , t)

= s w(s , t) w(s+ , t) .
2
2

30

2.3 Wave Equation

The velocity of the discontinuity of the shock velocity U = s.


If [ ] indicates the jump across
the shock then this condition may be written in the form
 2
w
.
U [w] =
2
The shock velocity for Burgers equation is
U=

1 w2 (s+ ) w2 (s )
w(s+ ) + w(s )
=
.
2 w(s+ ) w(s )
2

The problem then reduces to fitting shock discontinuities into the solution in such a way
that the jump condition is satisfied and multi-valued solution are avoided. A solution that
satisfies the original equation in regions and which satisfies the integral form of the equation
is called a weak solution or generalised solution.
Example: Consider the inviscid Burgers equation
w
w
+w
= 0,
t
x
with initial conditions

1
1
w(x = , t = 0) = f () =

for 0,
for 0 1,
for 1.

As seen before, the characteristics are curves on which w = f () as well as x f () t = are


constant, where is the parameter of the parametric form of the curve of initial data, .
For all (0, 1), f 0 () = 1 is negative (f 0 = 0 elsewhere), so we can expect that all the
characteristics corresponding to these values of intersect at the same point; the solution of
the inviscid Burgers equation becomes multi-valued at the time
tmin = 1/ max[f 0 ()] = 1, (0, 1).
Then, the position where the singularity develops at t = 1 is
x = f () t + = 1 + = 1.

st

=c

w=

1&

t
x

w=0 & x=cst

t=

1&

w=

x=1

multivalued
solution

t>1
x

w(x,t) 1
t=0

weak solution
(shock wave)

w(x,t) 1
t=0

U
x

t=1
x

31

Chapter 2 First Order Equations


As time increases, the slope of the solution,

1x
w(x, t) =

1t
0

for
for
for

x t,

t x 1,

x 1,

with 0 t < 1,

becomes steeper and steeper until it becomes vertical at t = 1; then the solution is multivalued. Nevertheless, we can define a generalised solution, valid for all positive time, by
introducting a shock wave.
Suppose shock at s(t) = U t + , with w(s , t) = 1 and w(s+ , t) = 0. The jump condition
gives the shock velocity,
w(s+ ) + w(s )
1
U=
= ;
2
2
furthermore, the shock starts at x = 1, t = 1, so = 1 1/2 = 1/2. Hence, the weak solution
of the problem is, for t 1,

1
0 for x < s(t),
w(x, t) =
where s(t) = (t + 1).
1 for x > s(t),
2

2.4
2.4.1

Systems of Equations
Linear and Semilinear Equations

These are equations of the form


n 
X

aij ux(j)

bij uy(j)

j=1

= ci ,

i = 1, 2, . . . , n


u
= ux ,
x

for the unknowns u(1) , u(j) , . . . , u(n) and when the coefficients aij and bij are functions only
of x and y. (Though the ci could also involve u(k) .)
In matrix notation
Aux + Buy = c,
where

. . . a1n
b11 . . . b1n

.. ,
.. ,
..
B = (bij ) = ... . . .
.
.
.
an1 . . . ann
bn1 . . . bnn
(1)

u
c1
u(2)
c2


c = . and u = . .
.
.
.
.
cn
u(n)

a11
..
A = (aij ) = .

E.g.,

(1)
(2)
(2)
u(1)
x 2ux + 3uy uy = x + y,

(2)
(1)
(2)
2
2
u(1)
x + ux 5uy + 2uy = x + y ,

32

2.4 Systems of Equations

can be expressed as

 " (1) # 
 " (1) # 

1 2 ux
3 1 uy
x+y
(2) + 5
(2) = x2 + y 2 ,
1 1
2
ux
uy
or Aux + Buy = c where




3 1
1 2
, B=
A=
5 2
1 1

 
1/3 2/3
If we multiply by A1 =
1/3 1/3


x+y
.
and c = 2
x + y2


A1 A ux + A1 B uy = A1 c,

we obtain ux + Duy = d,

 

7/3 1
3 1
1/3 2/3
1
and d = A1 c.
=
where D = A B =
8/3 1
1/3 1/3 5 2
We now assume that the matrix A is non-singular (i.e., the inverse A 1 exists ) at least
there is some region of the (x, y)-plane where it is non-singular. Hence, we need only to
consider systems of the form
ux + Duy = d.


We also limit our attention to totally hyperbolic systems, i.e. systems where the matrix D
has n distinct real eigenvalues (or at least there is some region of the plane where this holds).
D has the n distinct eigenvalues 1 , 2 , . . . , n where det(i I D) = 0 (i = 1, . . . , n), with
i 6= j (i 6= j) and the n corresponding eigenvectors e 1 , e2 , . . . , en so that
Dei = i ei .
The matrix P = [e1 , e2 , . . . , en ] diagonalises

1 0
0 2

= 0 ...

0 ...
0 ...
We now put u = P v,
then
and

D via P 1 DP = ,

... ...
0
0
...
0

..
.
. ...
0

0 n1 0
...
0
n

P vx + Px v + DP vy + DPy v = d,
P 1 P vx + P 1 Px v + P 1 DP vy + P 1 DPy v = P 1 d,

which is of the form


vx + vy = q,
where
q = P 1 d P 1 Px v P 1 DPy v.

The system is now of the form

vx(i) + i vy(i) = qi

(i = 1, . . . , n),

where qi can involve {v (1) , v (2) , . . . , v (n) } and with n characteristics given by

dy
= i .
dx
This is the canonical form of the equations.

33

Chapter 2 First Order Equations


Example 1:

Consider the linear system


(1)

(2)

(2)
ux

(1)
9 uy

ux + 4 uy = 0,
+

= 0,

with initial conditions u = [2x, 3x] T on y = 0.


0 4
.
Here, ux + Duy = 0 with D =
9 0
Eigenvalues:
det(D I) = 0 2 36 = 0 = 6.


Eigenvectors:

   
   
2
x
0
x
6 4
for = 6,
=

=
3
y
0
9 6 y

   
   
6 4 x
0
x
2
=

=
for = 6.
9 6 y
0
y
3

Then,







1 3 2
6 0
2 2
1
1
.
and P DP =
, P =
P =
0 6
3 3
12 3 2




2 2
6 0
So we put u =
v and vx +
v = 0, which has general solution
3 3
0 6 y
v (1) = f (6x y)

and

v (2) = g(6x + y),

i.e.
u(1) = 2v (1) + 2v (2)

and u(1) = 3v (1) 3v (2) .

Initial conditions give


2x = 2f (6x) + 2g(6x),
3x = 3f (6x) 3g(6x),
so, f (x) = x/6 and g(x) = 0; then
1
(6x y),
3
1
= (6x y).
2

u(1) =
u(2)
Example 2:

Reduce the linear system




4y x 2x 2y
ux +
ux = 0
2y 2x 4x y

to canonical form in the region of the (x, y)-space where it is totally hyperbolic.
Eigenvalues:


4y x
2x 2y
det
= 0 {3x, 3y}.
2y 2x
4x y
The system is totally hyperbolic everywhere expect where x = y.

34

2.4 Systems of Equations

Eigenvalues:
1 = 3x e1 = [1, 2]T ,
2 = 3y e2 = [2, 1]T .

So,




1 1 2
1 2
1
, P =
P =
2 1
3 2 1


1 2
Then, with u =
v we obtain
2 1


vx +

2.4.2

and


3x 0
.
DP =
0 3y



3x 0
v = 0.
0 3y y

Quasilinear Equations

We consider systems of n equations, involving n functions u (i) (x, y) (i = 1, . . . , n), of the form
ux + Duy = d,
where D as well as d may now depend on u. (We have already shown how to reduce a more
general system Aux + Buy = c to that simpler form.) Again, we limit our attention to totally
hyperbolic systems; then
= P 1 DP D = P P 1 ,

using the same definition of P , P 1 and the diagonal matrix , as for the linear and semilinear
cases. So, we can transform the system in its normal form as,
P 1 ux + P 1 uy = P 1 d,
such that it can be written in component form as
n
X
j=1

Pij1

(j)
(j)
u + i
u
x
y

n
X

Pij1 dj ,

(i = 1, . . . , n)

j=1

where i is the ith eigenvalue of the matrix D and wherein the ith equation involves differentiation only in a single direction the direction dy/dx = i . We define the ith characteristic,
with curvilinear coordinate si , as the curve in the (x, y)-plane along which
dx
= 1,
dsi

dy
= i
dsi

or equivalently

dy
= i .
dx

Hence, the directional derivative parallel to the characteristic is


(j)
(j)
d (j)
u =
u + i
u ,
dsi
x
y
and the system in normal form reduces to n ODEs involving different components of u
n
X
j=1

Pij1

d (j) X 1
u =
Pij dj
dsi
j=1

(i = 1, . . . , n).

35

Chapter 2 First Order Equations


Example: Unsteady, one-dimensional motion of an inviscid compressible adiabatic gas.
Consider the equation of motion (Euler equation)
u
1 P
u
+u
=
,
t
x
x
and the continuity equation

+
+u
= 0.
t
x
x
If the entropy is the same everywhere in the motion then P = constant, and the motion
equation becomes
u
u c2
+u
+
= 0,
t
x
x
where c2 = dP/d = P/ is the sound speed. We have then a system of two first order
quasilinear PDEs; we can write these as
w
w
+D
= 0,
t
x

with

 
u
w=

and


u c2 /
D=
.

The two characteristics of this hyperbolic system are given by dx/dt = where are the
eigenvalues of D;


u c2 /
= 0 (u )2 = c2 and = u c.
det(D I) =

u
The eigenvectors are [c, ]T for and [c, ]T for + , such that the usual matrices are






1 c
uc
0
c c
1
1
, such that = T DT =
.
T =
, T =
0
u+c

2c c

Put and the curvilinear coordinates along the characteristics dx/dt = u c and dx/dt =
u + c respectively; then the system transforms to the canonical form
dt
dt
=
= 1,
d
d

dx
= u c,
d

dx
= u + c,
d

du
d
c
=0
d
d

and

du
d
+c
= 0.
d
d

36

2.4 Systems of Equations

Chapter 3

Second Order Linear and


Semilinear Equations in Two
Variables
Contents
3.1
3.2

3.1

Classification and Standard Form Reduction . . . . . . . . . . . .


Extensions of the Theory . . . . . . . . . . . . . . . . . . . . . . . .

37
44

Classification and Standard Form Reduction

Consider a general second order linear equation in two independent variables


a(x, y)

2u
u
2u
u
2u
+ 2b(x, y)
+ c(x, y) 2 + d(x, y)
+ e(x, y)
+ f (x, y)u = g(x, y);
2
x
xy
y
x
y

in the case of a semilinear equation, the coefficients d, e, f and g could be functions of x u, y u


and u as well.
Recall, for a first order linear and semilinear equation, a u/x+b u/y = c, we could define
new independent variables, (x, y) and (x, y) with J = (, )/(x, y) 6= {0, }, to reduce
the equation to the simpler form, u/ = (, ).
For the second order equation, can we also transform the variables from (x, y) to (, ) to put
the equation into a simpler form?
So, consider the coordinate transform (x, y) (, ) where and are such that the Jacobian,



(, )
x y
J=
=
6= {0, }.
(x, y)
x

Then by inverse theorem there is an open neighbourhood of (x, y) and another neighbourhood
of (, ) such that the transformation is invertible and one-to-one on these neighbourhoods.
As before we compute chain rule derivations
u
u
u
=
+
,
x
x x

u
u
u
=
+
,
y
y y
37

38

3.1 Classification and Standard Form Reduction



 
2 u 2 u 2
u 2
2 u
2
+ 2
+
,
+2
+
x
x x
x
x2
x2
 
 
2 u 2 u 2 u 2
2 u 2
u 2
2u
+
2
+
=
+
+
,
y 2
2 y
y y
2 y
y 2
y 2


2 u
2 u

u 2
2 u u 2
2u
= 2
+
+
+
+
.
+ 2
xy
x y x y y x
x y
xy
xy
2u
2u
=
x2
2

The equation becomes

2u
2u
2u
+
2B
+
C
+ F (u , u , u, , ) = 0,
2

(3.1)

where

 2

2

A=a
+ 2b
,
+c
x
x y
y






+c
+b
+
,
B=a
x x
x y y x
y y
 2
 2

+ 2b
.
C=a
+c
x
x y
y


We write explicitly only the principal part of the PDE, involving the highest-order derivatives
of u (terms of second order).
It is easy to verify that
2

(B AC) = (b ac)

x y y x

2

where (x y y x )2 is just the Jacobian squared. So, provided J 6= 0 we see that the
sign of the discriminant b2 ac is invariant under coordinate transformations. We can use
this invariance properties to classify the equation.
Equation (3.1) can be simplified if we can choose and so that some of the coefficients A,
B or C are zero. Let us define,
D =

/x
/y

and D =

/x
;
/y

then we can write


A=

aD2

+ 2bD + c

2


B = (aD D + b (D + D ) + c)
,
y y
 
 2
2
C = aD + 2bD + c
.
y

Now consider the quadratic equation

aD 2 + 2bD + c = 0,

(3.2)

Chapter 3 Second Order Linear and Semilinear Equations in Two Variables


whose solution is given by
D=

39

b2 ac
.
a

If the discriminant b2 ac 6= 0, equation (3.2) has two distinct roots; so, we can make both
coefficients A and C zero if we arbitrarily take the root with the negative sign for D and
the one with the positive sign for D ,

b b2 ac
/x
=
A = 0,
(3.3)
D =
/y
a

/x
b + b2 ac
D =
=
C = 0.
/y
a
Then, using D D = c/a and D + D = 2b/a we have
B=


2
ac b2
B 6= 0.
a
y y

Furthermore, if the discriminant b 2 ac > 0 then D and D as well as and are real. So,
we can define two families of one-parameter characteristics of the PDE as the curves described
by the equation (x, y) = constant and the equation (x, y) = constant. Differentiate along
the characteristic curves given by = constant,
d =

dx +
dy = 0,
x
y

and make use of (3.3) to find that this characteristics satisfy

b + b2 ac
dy
=
.
dx
a

(3.4)

Similarly we find that the characteristic curves described by (x, y) = constant satisfy

dy
b b2 ac
=
.
(3.5)
dx
a
If the discriminant b2 ac = 0, equation (3.2) has one unique root and if we take this root
for D say, we can make the coefficient A zero,
D =

b
/x
= A = 0.
/y
a

To get independent of , D has to be different from D , so C =


6 0 in this case, but B is
now given by




 2

b


b
b
B = a D + b + D + c
= +c
,
a
a
y y
a
y y
so that B = 0. When b2 ac = 0 the PDE has only one family of characteristic curves, for
(x, y) = constant, whose equation is now
dy
b
= .
dx
a
Thus we have to consider three different cases.

(3.6)

40

3.1 Classification and Standard Form Reduction


1. If b2 > ac we can apply the change of variable (x, y) (, ) to transform the original
PDE to
2u
+ (lower order terms) = 0.

In this case the equation is said to be hyperbolic and has two families of characteristics
given by equation (3.4) and equation (3.5).
2. If b2 = ac, a suitable choice for still simplifies the PDE, but now we can choose
arbitrarily provided and are independent and the equation reduces to the
form
2u
+ (lower order terms) = 0.
2
The equation is said to be parabolic and has only one family of characteristics given by
equation (3.6).
3. If b2 < ac we can again apply the change of variables (x, y) (, ) to simplify the
equation but now this functions will be complex conjugate. To keep the transformation
real, we apply a further change of variables (, ) (, ) via
= + = 2 <(),

= i( ) = 2 =(),

2u
2u
2u
=
+

2 2
so, the equation can be reduced to
i.e.,

(via the chain rule);

2u
2u
+
+ (lower order terms) = 0.
2 2
In this case the equation is said to be elliptic and has no real characteristics.
The above forms are called the canonical (or standard) forms of the second order linear or
semilinear equations (in two variables).
Summary:
b2 ac
Canonical form
Type

>0
2u

+ ... = 0

Hyperbolic

=0
2u
2

+ ... = 0

Parabolic

<0
2u
2

2u
2

+ ... = 0

Elliptic

E.g.
The wave equation,

2
2u
2 u

c
= 0,
w
t2
x2
is hyperbolic (b2 ac = c2w > 0) and the two families of characteristics are described
by dx/dt = cw i.e. = x cw t and = x + cw t. So, the equation transforms into its
canonical form 2 u/ = 0 whose solutions are waves travelling in opposite direction
at speed cw .

Chapter 3 Second Order Linear and Semilinear Equations in Two Variables

41

The diffusion (heat conduction) equation,


2 u 1 u

= 0,
x2 t
is parabolic (b2 ac = 0). The characteristics are given by dt/dx = 0 i.e. = t =
constant.
Laplaces equation,

2u 2u
+ 2 = 0,
x2
y

is elliptic (b2 ac = 1 < 0).


The type of a PDE is a local property. So, an equation can change its form in different regions
of the plane or as a parameter is changed. E.g. Tricomis equation
y

2u 2u
+ 2 = 0,
x2
y

(b2 ac = 0 y = y)

is elliptic in y > 0, parabolic for y = 0 and hyperbolic in y < 0, and for small disturbances
in incompressible (inviscid) flow
1 2u 2u
+ 2 = 0,
1 m2 x2
y

(b2 ac =

1
)
1 m2

is elliptic if m < 1 and hyperbolic if m > 1.


Example 1:

Reduce to the canonical form


2
2u
1
2 u

2xy
=
y
+
x
2
2
x
xy
y
xy
2

2u

3 u

+x

3 u

a = y2
Here b = xy
so b2 ac = (xy)2 x2 y 2 = 0 parabolic equation.

2
c=x
On = constant,

dy
b + b2 ac
b
x
=
= = = x2 + y 2 .
dx
a
a
y
We can choose arbitrarily provided and are independent. We choose = y. (Exercise,
try it with = x.) Then
u
u
= 2x ,
x

u
u u
= 2y
+
,
y

2u
2u
2u
= 4xy 2 + 2x
,
xy

and the equation becomes


2y 2

2
u
2u
2 u
=
2
,
+
4x
x2

2
u
2u
2u
2u
2 u
=
2
+
4y
,
+
4y
+
y 2

2u
u
2u
2u
u
2u
+ 4x2 y 2 2 8x2 y 2 2 4x2 y
+ 2x2
+ 4x2 y 2 2



2
2
1
u
u
u
u
u
2xy 3
,
+ x2 2 =
+ 2x3 y
+ x3
+ 4x2 y

xy

42

3.1 Classification and Standard Form Reduction


2 u 1 u

= 0.
2

i.e.

(canonical form)

This has solution


u = f () + 2 g(),
where f and g are arbitrary functions (via integrating factor method), i.e.
u = f (x2 + y 2 ) + y 2 g(x2 + y 2 ).
We need to impose two conditions on u or its partial derivatives to determine the functions
f and g i.e. to find a particular solution.
Example 2:

Reduce to canonical form and then solve

2u
2u
u
2u
+

2
+ 1 = 0 in 0 x 1, y > 0, with u =
= x on y = 0.
2
2
x
xy
y
y

a=1
so b2 ac = 9/4 (> 0) equation is hyperbolic.
Here b = 1/2

c = 2
Characteristics:


dy
1 3
/x
/x
.
= = 1 or 2
=
or
dx
2 2
/y
/y
Two methods of solving:
1. directly:
1
dy
= 2 x y = constant
dx
2

and

dy
= 1 x + y = constant.
dx

2. simultaneous equations:

y )
=2
x=
=x
x
y
2

=x+y
y=

x
y

So,

u
1 u u
=
+
,
y
2

u u
u
=
+
,
x

1 2u 1 2u
2u
2u
=
+
+
,
xy
2 2
2
2

1
( + 2)
3
2
( )
3

2u
2u
u
2u
=
+
2
+
x2
2

2
2u
1 2u
2u
2u
=

+
,
y 2
4 2

and the equation becomes


u
2u
2u
2u 1 2u 1 2u
2u 1 2u
2u
+
2

+
2
+ 1 = 0,
+
+

2
2

2
2 2
2
2
2 2

9 2u
+ 1 = 0,
2

canonical form.

Chapter 3 Second Order Linear and Semilinear Equations in Two Variables

43

So 2 u/ = 2/9 and general solution is given by


2
u(, ) = + f () + g(),
9
where f and g are arbitrary functions; now, we need to apply two conditions to determine
these functions.
When, y = 0, = = x so the condition u = x at y = 0 gives
2
2
u( = x, = x) = x2 + f (x) + g(x) = x f (x) + g(x) = x + x2 .
9
9

(3.7)

Also, using the relation


u
1 u u
1
1
2
=
+
= f 0 () + g 0 (),
y
2

9
2
9
the condition u/y = x at y = 0 gives
u
1
1
2
1
10
( = x, = x) = x f 0 (x) x + g 0 (x) = x g 0 (x) f 0 (x) =
x,
y
9
2
9
2
9
1
5
g(x) f (x) = x2 + k,
(3.8)
2
9
where k is a constant. Solving equation (3.7) and equation (3.8) simultaneously gives,
and after integration,

2
2
2
f (x) = x x2 k
3
9
3
or, in terms of and

f () =

and

g(x) =

2
2
2
2 k
3
9
3

1
4
2
x + x2 + k,
3
9
3

1
4
2
and g() = + 2 + k.
3
9
3

So, full solution is


2
2
1
4
2
u(, ) = + 2 + + 2 ,
9
3
9
3
9
1
2
= (2 + ) + ( )(2 + ).
3
9
u(x, y) = x + xy +
Example 3:

y2
.
2

(check this solution.)

Reduce to canonical form


2u
2u
2u
+
= 0.
+
x2 xy
y 2

a=1
Here b = 1/2
so b2 ac = 3/4 (< 0) equation is elliptic.

c=1
Find and via

)
)

= y 12 (1 + i 3)x
dy/dx = (1 + i 3)/2

= constant on dy/dx = (1 i 3)/2


= y 12 (1 i 3)x
= constant on

44

3.2 Extensions of the Theory

To obtain a real transformation, put


= + = 2y x
So,

and = i( ) = x 3.

u u
u
=
+ 3
,
x

u
u
2u
=2
,
=
y

x2
2u
2u
2u
3
= 2
+
2
,
xy
2

and the equation transforms to

2u
2u
2u
3

2
+
3
,
2

2
2u
2u
=
4
,
y 2
2

2u
2u
2u
2u
2u
2u

2
+
3

2
+
2
+
4
= 0.
3
3
2

2
2

3.2
3.2.1

2u
2u
+
= 0,
2
2

canonical form.

Extensions of the Theory


Linear second order equations in n variables

There are two obvious ways in which we might wish to extend the theory.
To consider quasilinear second order equations (still in two independent variables.) Such
equations can be classified in an invariant way according to rule analogous to those developed
above for linear equations. However, since a, b and c are now functions of x u, y u and u its
type turns out to depend in general on the particular solution searched and not just on the
values of the independent variables.
To consider linear second order equations in more than two independent variables. In such
cases it is not usually possible to reduce the equation to a simple canonical form. However,
for the case of an equation with constant coefficients such a reduction is possible. Although
this seems a rather restrictive class of equations, we can regard the classification obtained as
a local one, at a particular point.
Consider the linear PDE
n
X

aij

i,j=1

X u
2u
+
+ cu = d.
bi
xi xj
xi
i=1

Without loss of generality we can take the matrix A = (a ij ), i, j = 1 n, to be symmetric


(assuming derivatives commute). For any real symmetric matrix A, there is an associate
orthogonal matrix P such that P T AP = where is a diagonal matrix whose element are
the eigenvalues, i , of A and the columns of P the linearly independent eigenvectors of A,
ei = (e1i , e2i , , eni ). So
P = (eij )

and = (i ij ), i, j = 1, , n.

Now consider the transformation x = P , i.e. = P 1 x = P T x (P orthogonal) where


x = (x1 , x2 , , xn ) and = (1 , 2 , , n ); this can be written as
xi =

n
X
j=1

eij j

and

j =

n
X
i=1

eij xi .

Chapter 3 Second Order Linear and Semilinear Equations in Two Variables


So,

X u
u
=
eik
xi
k

45

n
X
2u
u
=
eik ejr .
xi xj
k r

and

k=1

k,r=1

The original equation becomes,


n
n
X
X

i,j=1 k,r=1

u
aij eik ejr + (lower order terms) = 0.
k r

But by definition of the eigenvectors of A,


eTk

n
X

A er =

i,j=1

eik aij ejr r rk .

Then equation simplifies to


n
X

k=1

u
+ (lower order terms) = 0.
k2

We are now in a position to classify the equation.


Equation is elliptic if and only if all k are non-zero and have the same sign. E.g.
Laplaces equation
2u 2u 2u
+ 2 + 2 = 0.
x2
y
z
When all the k are non-zero and have the same sign except for precisely one of them,
the equation is hyperbolic. E.g. the wave equation
2u
c2
t2

2u 2u 2u
+ 2+ 2
x2
y
z

= 0.

When all the k are non-zero and there are at least two of each sign, the equation
ultra-hyperbolic. E.g.
2u 2u
2u 2u
+
=
+
;
x21 x22
x23 x24
such equation do not often arise in mathematical physics.
If any of the k vanish the equation is parabolic. E.g. heat equation
u

3.2.2

2u 2u 2u
+ 2 + 2
x2
y
z

= 0.

The Cauchy Problem

Consider the problem of finding the solution of the equation


a

2u
2u
2u
+
2b
+ F (x u, y u, u, x, y) = 0
+
c
x2
xy
y 2

46

3.2 Extensions of the Theory

which takes prescribed values on a given curve which we assume is represented parametrically in the form
x = (), y = (),
for I, where I is an interval, 0 1 , say. (Usually consider piecewise smooth
curves.)
We specify Cauchy data on : u, u/x and u/y are given I, but note that
we cannot specify all these quantities arbitrarily. To show this, suppose u is given on by
u = f (); then the derivative tangent to , du/d, can be calculated from du/d = f 0 ()
but also
u dx u dy
du
=
+
,
d
x d y d
u
u
= 0 ()
+ 0 ()
= f 0 (),
x
y
so, on , the partial derivatives u/x, u/y and u are connected by the above relation.
Only derivatives normal to and u can be prescribed independently.
So, the Cauchy problem consists in finding the solution u(x, y) which satisfies the following
conditions

u((), ()) = f ()

,
u

and
((), ()) = g()
n

where I and /n = n denotes a normal derivative to (e.g. n = [ 0 , 0 ]T ); the


partial derivatives u/x and u/y are uniquely determined on by these conditions.
Set, p = u/x and q = u/y so that on , p and q are known; then
dp
2 u dx
2 u dy
=
+
ds
x2 ds xy ds

and

dq
2 u dx 2 u dy
=
+ 2 .
ds
xy ds
y ds

Combining these two equations with the original PDE gives the following system of equations
for 2 u/x2 , 2 u/xy and 2 u/y 2 on (in matrix form),

2u
x2
2u
xy
2u
y 2

F
dp
ds
dq
ds

2b

dx
where M =
ds

dy
ds
dx
ds

0 .

dy
ds

So, if det(M) 6= 0 we can solve the equations uniquely and find 2 u/x2 , 2 u/xy and
2 u/y 2 on . By successive differentiations of these equations it can be shown that the
derivatives of u of all orders are uniquely determined at each point on for which det(M) 6= 0.
The values of u at neighbouring points can be obtained using Taylors theorem.
So, we conclude that the equation can be solved uniquely in the vicinity of provided
det(M) 6= 0 (Cauchy-Kowaleski theorem provides a majorant series ensuring convergence of
Taylors expansion).

Chapter 3 Second Order Linear and Semilinear Equations in Two Variables

47

Consider what happens when det(M) = 0, so that M is singular and we cannot solve uniquely
for the second order derivatives on . In this case the determinant det(M) = 0 gives,
a

dy
ds

2

dx dy
+c
2b
ds ds

dx
ds

2

= 0.

But,
dy
dy/ds
=
dx
dx/ds
and so (dividing through by dx/ds), dy/dx satisfies the equation,

 2
dy
dy
b b2 ac
dy
2b
+ c = 0,
i.e.
=
a
dx
dx
dx
a

or

dy
b
= .
dx
a

The exceptional curves , on which, if u and its normal derivative are prescribed, no unique
solution can be found satisfying these conditions, are the characteristics curves.

48

3.2 Extensions of the Theory

Chapter 4

Elliptic Equations
Contents
4.1
4.2
4.3
4.4

4.1

Definitions . . . . . . . . . . . . . . . . . . . . . . . . .
Properties of Laplaces and Poissons Equations . .
Solving Poisson Equation Using Greens Functions
Extensions of Theory: . . . . . . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

49
50
54
68

Definitions

Elliptic equations are typically associated with steady-state behavior. The archetypal elliptic
equation is Laplaces equation
2 u = 0,

e.g.

2u 2u
+ 2 = 0 in 2-D,
x2
y

and describes
steady, irrotational flows,
electrostatic potential in the absence of charge,
equilibrium temperature distribution in a medium.
Because of their physical origin, elliptic equations typically arise as boundary value problems
(BVPs). Solving a BVP for the general elliptic equation
L[u] =

n
X

i,j=1

aij

X u
2u
+
+ cu = F
bi
xi xj
xi
i=1

(recall: all the eigenvalues of the matrix A = (a ij ), i, j = 1 n, are non-zero and have the
same sign) is to find a solution u in some open region of space, with conditions imposed
on (the boundary of ) or at infinity. E.g. inviscid flow past a sphere is determined by
boundary conditions on the sphere (u n = 0) and at infinity (u = Const).
There are three types of boundary conditions for well-posed BVPs,
1. Dirichlet condition u takes prescribed values on the boundary (first BVP).
49

50

4.2 Properties of Laplaces and Poissons Equations


2. Neumann conditions the normal derivative, u/n = n u is prescribed on the
boundary (second BVP).
In this case we have compatibility conditions (i.e. global constraints):
E.g., suppose u satisfies 2 u = F on and n u = n u = f on . Then,
Z
Z
Z
Z
u
2 u dV =
dS
(divergence theorem),
u dV =
u n dS =
n

Z
Z

F dV =
f dS
for the problem to be well-defined.

3. Robin conditions a combination of u and its normal derivative such as u/n + u


is prescribed on the boundary (third BVP).

y
n

Sometimes we may have a mixed problem, in which u is given on part of and u/n given
on the rest of .
If encloses a finite region, we have an interior problem; if, however, is unbounded, we
have an exterior problem, and we must impose conditions at infinity.
Note that initial conditions are irrelevant for these BVPs and the Cauchy problem for elliptic
equations is not always well-posed (even if Cauchy-Kowaleski theorem states that the solution
exist and is unique).
As a general rule, it is hard to deal with elliptic equations since the solution is global, affected
by all parts of the domain. (Hyperbolic equations, posed as initial value or Cauchy problem,
are more localised.)
From now, we shall deal mainly with the Helmholtz equation 2 u + P u = F , where P and
F are functions of x, and particularly with the special one if P = 0, Poissons equation, or
Laplaces equation, if F = 0 too. This is not too severe restriction; recall that any linear
elliptic equation can be put into the canonical form
n
X
2u
k=1

x2k

+ = 0

and that the lower order derivatives do not alter the overall properties of the solution.

4.2

Properties of Laplaces and Poissons Equations

Definition: A continuous function satisfying Laplaces equation in an open region , with


continuous first and second order derivatives, is called an harmonic function. Functions u

51

Chapter 4 Elliptic Equations

in C 2 () with 2 u 0 (respectively 2 u 0) are call subharmonic (respectively superharmonic).

4.2.1

Mean Value Property

Definition: Let x0 be a point in and let BR (x0 ) denote the open ball having centre x 0
and radius R. Let R (x0 ) denote the boundary of BR (x0 ) and let A(R) be the surface area
of R (x0 ). Then a function u has the mean value property at a point x 0 if
Z
1
u(x) dS
u(x0 ) =
A(R) R
for every R > 0 such that BR (x0 ) is contained in . If instead u(x0 ) satisfies
Z
1
u(x0 ) =
u(x) dV,
V (R) BR
where V (R) is the volume of the open ball B R (x0 ), we say that u(x0 ) has the second
mean value property at a point x0 . The two mean value properties are equivalent.

BR
x0
x

Theorem:
on .
Proof:

If u is harmonic in an open region of R n , then u has the mean value property

We need to make use of Greens theorem which says,



Z 
Z

v
u
u
v
dS =
v 2 u u 2 v dV.
n
n
S
V

(4.1)

(Recall: Apply divergence theorem to the function vu uv to state Greens theorem.)


Since u is harmonic, it follows from equation (4.1), with v = 1, that
Z
u
dS = 0.
S n
Now, take v = 1/r, where r = |x x0 |, and the domain V to be Br (x0 ) B (x0 ), 0 < < R.
Then, in Rn x0 ,

 
1
1
2
2
r
=0
v= 2
r r
r r

52

4.2 Properties of Laplaces and Poissons Equations

so v is harmonic too and equation (4.1) becomes


Z
Z
Z
Z
v
v
v
v
u
dS +
dS =
dS
dS = 0
u
u
u
n
n
r
r
r

v
dS =
r

u
r

v
dS
r

1
2

i.e

u dS =

1
r2

u dS.

Since u is continuous, then as 0 the LHS converges to 4 u(x 0 , y0 , z0 ) (with n = 3, say),


so
Z
1
u dS.
u(x0 ) =
A(r) r
Recovering the second mean value property (with n = 3, say) is straightforward
Z

1
r3
u(x0 ) =
u(x0 ) d =
3
4
2

1
u dS d =
4

u dV.

Br

The inverse of this theorem holds too, but is harder to prove. If u has the mean value property
then u is harmonic.

4.2.2

Maximum-Minimum Principle

One of the most important features of elliptic equations is that it is possible to prove theorems
concerning the boundedness of the solutions.
Theorem:

Suppose that the subharmonic function u satisfies


2 u = F

in , with F > 0 in .

Then u(x, y) attains his maximum on .


Proof: (Theorem stated in 2-D but holds in higher dimensions.) Suppose for a contradiction
that u attains its maximum at an interior point (x 0 , y0 ) of . Then at (x0 , y0 ),
u
= 0,
x

u
= 0,
y

2u
0
x2

and

2u
0,
y 2

since it is a maximum. So,


2u 2u
+ 2 0,
x2
y

which contradicts F > 0 in .

Hence u must attain its maximum on , i.e. if u M on , u < M in .


Theorem: The weak Maximum-Minimum Principle for Laplaces equation.
Suppose that u satisfies
2 u = 0 in a bounded region ;
if m u M on , then m u M in .

53

Chapter 4 Elliptic Equations

Proof: (Theorem stated in 2-D but holds in higher dimensions.) Consider the function
v = u + (x2 + y 2 ), for any > 0. Then 2 v = 4 > 0 in (since 2 (x2 + y 2 ) = 4), and
using the previous theorem,
v M + R2 in ,
where u M on and R is the radius of the circle containing . As this holds for any ,
let 0 to obtain
u M in ,
i.e., if u satisfies 2 u = 0 in , then u cannot exceed M , the maximum value of u on .
Also, if u is a solution of 2 u = 0, so is u. Thus, we can apply all of the above to u to
get a minimum principle: if u m on , then u m in .
This theorem does not say that harmonic function cannot also attain m and M inside
though. We shall now progress into the strong Maximum-Minimum Principle.
Theorem: Suppose that u has the mean value property in a bounded region and that u
= . If u is not constant in then u attains its maximum value on
is continuous in
the boundary of , not in the interior of .
then it attains its maximum
Proof: Since u is continuous in the closed, bounded domain

M somewhere in .
Our aim is to show that, if u attains its max. at an interior point of , then u is constant in

.
Suppose u(x0 ) = M and let x? be some other point of . Join these points with a path
covered by a sequence of overlapping balls, B r .

x?
x

x0

x1

Consider the ball with x0 at its center. Since u has the mean value property then
Z
1
M = u(x0 ) =
u dS M.
A(r) r
This equality must hold throughout this statement and u = M throughout the sphere surrounding x0 . Since the balls overlap, there is x 1 , centre of the next ball such that u(x 1 ) = M ;
the mean value property implies that u = M in this sphere also. Continuing like this gives
u(x? ) = M . Since x? is arbitrary, we conclude that u = M throughout , and by continuity
Thus if u is not a constant in it can attain its maximum value only on the
throughout .
boundary .

54

4.3 Solving Poisson Equation Using Greens Functions

Corollary: Applying the above theorem to u establishes that if u is non constant it can
attain its minimum only on .
Also as a simple corollary, we can state the following theorem. (The proof follows immediately
the previous theorem and the weak Maximum-Minimum Principle.)
Theorem: The strong Maximum-Minimum Principle for Laplaces equation.
i.e. solution of 2 u = 0 in and continuous in ,
with M and
Let u be harmonic in ,
m the maximum and minimum values respectively of u on the boundary . Then, either

m < u < M in or else m = u = M in .


Note that it is important that be bounded for the theorem to hold. E.g., consider u(x, y) =
ex sin y with = {(x, y)| < x < +, 0 < y < 2}. Then 2 u = 0 and on the boundary
of we have u = 0, so that m = M = 0. But of course u is not identically zero in .
Corollary: If u = C is constant on , then u = C is constant in . Armed with the
above theorems we are in position to prove the uniqueness and the stability of the solution
of Dirichlet problem for Poissons equation.
Consider the Dirichlet BVP
2 u = F in with

u = f on

and suppose u1 , u2 two solutions to the problem. Then v = u 1 u2 satisfies


2 v = 2 (u1 u2 ) = 0

in , with v = 0 on .

Thus, v 0 in , i.e. u1 = u2 ; the solution is unique.


To establish the continuous dependence of the solution on the prescribed data (i.e. the
stability of the solution) let u1 and u2 satisfy
2 u{1,2} = F in

with

u{1,2} = f{1,2} on ,

with max |f1 f2 | = . Then v = u1 u2 is harmonic with v = f1 f2 on . As before,


So, the
v must have its maximum and minimum values on ; hence |u 1 u2 | in .
solution is stable small changes in the boundary data lead to small changes in the solution.
We may use the Maximum-Minimum Principle to put bounds on the solution of an equation
without solving it.
The strong Maximum-Minimum Principle may be extended to more general linear elliptic
equations
n
n
X
X
2u
u
L[u] =
aij
+
+ cu = F,
bi
xi xj
xi
i,j=1

i=1

and, as for Poissons equation it is possible then to prove that the solution to the Dirichlet
BVP is unique and stable.

4.3

Solving Poisson Equation Using Greens Functions

We shall develop a formal representation for solutions to boundary value problems for Poissons equation.

55

Chapter 4 Elliptic Equations

4.3.1

Definition of Greens Functions

Consider a general linear PDE in the form


L(x)u(x) = F (x) in ,
where L(x) is a linear (self-adjoint) differential operator, u(x) is the unknown and F (x) is
the known homogeneous term.
(Recall: R L is self-adjoint if L = L? , where L? is defined by hv|Lui = hL? v|ui and where
hv|ui = v(x)w(x)u(x)dx ((w(x) is the weight function).)

The solution to the equation can be written formally

u(x) = L1 F (x),
where L1 , the inverse of L, is some integral operator. (We can expect to have LL 1 =
LL1 = I, identity.) We define the inverse L 1 using a Greens function: let
u(x) = L

F (x) =

G(x, )F ()d,

(4.2)

where G(x, ) is the Greens function associated with L (G is the kernel). Note that G
depends on both the independent variables x and the new independent variables , over
which we integrate.
Recall the Dirac -function (more precisely distribution or generalised function) (x) which
has the properties,
Z

(x) dx = 1

and

Rn

Rn

(x ) h() d = h(x).

Now, applying L to equation (4.2) we get


Lu(x) = F (x) =

LG(x, )F () d;

hence, the Greens function G(x, ) satisfies


u(x) =

4.3.2

G(x, ) F () d

with L G(x, ) = (x ) and x, .

Greens function for Laplace Operator

Consider Poissons equation in the open bounded region V with boundary S,


2 u = F in V .

(4.3)

56

4.3 Solving Poisson Equation Using Greens Functions

y
n

V
x

Then, Greens theorem (n is normal to S outward from V ), which states


Z

u v v u dV =

Z 
S

u
v
v
u
n
n

dS,

for any functions u and v, with h/n = n h, becomes


Z

u v dV =
V

vF dV +

Z 
S

u
v
v
u
n
n

dS;

so, if we choose v v(x, ), singular at x = , such that 2 v = (x ), then u is solution


of the equation

Z 
Z
v
u
vF dV
u
u() =
dS
(4.4)
v
n
n
S
V
which is an integral equation since u appears in the integrand. To address this we consider
another function, w w(x, ), regular at x = , such that 2 w = 0 in V . Hence, apply
Greens theorem to the function u and w

Z 
Z
Z

w
u
2
2
dS =
u
w
u w w u dV =
wF dV.
n
n
S
V
V
Combining this equation with equation (4.4) we find
u() =

(v + w)F dV

Z 
S

u (v + w) (v + w)
n
n

dS,

so, if we consider the fundamental solution of Laplaces equation, G = v + w, such that


2 G = (x ) in V ,
u() =

GF dV

Z 
S

G
u
G
n
n

dS.

(4.5)

Note that if, F , f and the solution u are sufficiently well-behaved at infinity this integral
equation is also valid for unbounded regions (i.e. for exterior BVP for Poissons equation).
The way to remove u or u/n from the RHS of the above equation depends on the choice
of boundary conditions.

57

Chapter 4 Elliptic Equations


Dirichlet Boundary Conditions

Here, the solution to equation (4.3) satisfies the condition u = f on S. So, we choose w
such that w = v on S, i.e. G = 0 on S, in order to eliminate u/n form the RHS of
equation (4.5).
Then, the solution of the Dirichlet BVP for Poissons equation
2 u = F in V with u = f on S
is
u() =

GF dV
2 v

f
S

G
dS,
n

where G = v +w (w regular at x = ) with


= (x) and 2 w = 0 in V and v +w = 0
on S. So, the Greens function G is solution of the Dirichlet BVP
2 G = (x ) in V,

with G = 0 on S.
Neumann Boundary Conditions

Here, the solution to equation (4.3) satisfies the condition u/n = f on S. So, we choose
w such that w/n = v/n on S, i.e. G/n = 0 on S, in order to eliminate u from the
RHS of equation (4.5).
However, the Neumann BVP
2 G = (x ) in V,
G
= 0 on S,
with
n
which does not satisfy a compatibility equation, has no solution. Recall that the Neumann
BVP 2 u = F in V , with u/n = f on S, is ill-posed if
Z
Z
f dS.
F dV 6=
S

We need to alter the Greens function a little to satisfy the compatibility equation; put
2 G = + C, where C is a constant, then the compatibility equation for the Neumann
BVP for G is
Z
Z
1
0 dS = 0 C = ,
( + C) dV =
V
S
V

where V is the volume of V . Now, applying Greens theorem to G and u:



Z
Z 

u
G
2
2
G u u G dV =
G
dS
u
n
n
V
S
we get

u() =

GF dV +
V

Gf dS +
S

1
V
|

u dV .
{z }

This shows that, whereas the solution of Poissons equation with Dirichlet boundary conditions is unique, the solution of the Neumann problem is unique up to an additive constant u

which is the mean value of u over .

58

4.3 Solving Poisson Equation Using Greens Functions

Thus, the solution of the Neumann BVP for Poissons equation


u
2 u = F in V with
= f on S
n
is
Z
Z
u() = u

GF dV +
Gf dS,
V

2 v

where G = v + w (w regular at x = ) with


= (x ), 2 w = 1/V in V and
w/n = v/n on S. So, the Greens function G is solution of the Neumann BVP
1
2 G = (x ) + in V,
V
G
= 0 on S.
with
n

Robin Boundary Conditions


Here, the solution to equation (4.3) satisfies the condition u/n + u = f on S. So, we
choose w such that w/n + w = v/n v on S, i.e. G/n + G = 0 on S. Then,


Z
Z 
Z 
u
G
G
G
+ G(u f ) dS = Gf dS.
dS =
u
u
n
n
n
S
S
S
Hence, the solution of the Robin BVP for Poissons equation
u
2 u = F in V with
+ u = f on S
n
is
Z
Z
Gf dS,
GF dV +
u() =
S

where G = v + w (w regular at x = ) with 2 v = (x ) and 2 w = 0 in V and


w/n + w = v/n v on S. So, the Greens function G is solution of the Robin BVP
with

2 G = (x ) in V,
G
+ G = 0 on S.
n

Symmetry of Greens Functions


The Greens function is symmetric (i.e., G(x, ) = G(, x)). To show this, consider two
Greens functions, G1 (x) G(x, 1 ) and G2 (x) G(x, 2 ), and apply Greens theorem to
these,

Z
Z 

G2
G1
G1 2 G2 G2 2 G1 dV =
G1
dS.
G2
n
n
V
S
Now, since, G1 and G2 are by definition Greens functions, G 1 = G2 = 0 on S for Dirichlet
boundary conditions, G1 /n = G2 /n = 0 on S for Neumann boundary conditions or
G2 G1 /n = G1 G2 /n on S for Robin boundary conditions, so in any case the right-hand
side is equal to zero. Also, 2 G1 = (x 1 ), 2 G2 = (x 2 ) and the equation becomes
Z
Z
G(x, 2 ) (x 1 ) dV,
G(x, 1 ) (x 2 ) dV =
V

G( 2 , 1 ) = G( 1 , 2 ).

Nevertheless, note that for Neumann BVPs, the term 1/V which provides the additive constant to the solution to Poissons equation breaks the symmetry of G.

59

Chapter 4 Elliptic Equations


Example:
Consider the 2-dimensional Dirichlet problem for Laplaces equation,
2 u = 0 in V , with u = f on S (boundary of V ).
Since u is harmonic in V (i.e. 2 u = 0) and u = f on S, then Greens theorem gives

Z
Z 
u
v
2
dS.
u v dV =
f
v
n
n
V
S
Note that we have no information about u/n on S or u in V . Suppose we choose,
v=


1
ln (x )2 + (y )2 ,
4

then 2 v = 0 on V for all points except P (x = , y = ), where it is undefined.


To eliminate
pthis singularity, we cut this point P out i.e, surround P by a small circle of
radius = (x )2 + (y )2 and denote the circle by , whose parametric form in polar
coordinates is
: {x = cos , y = sin with > 0 and (0, 2)}.

Hence, v = 1/2 ln and dv/d = 1/2 and applying Greens theorem to u and v in
this new region V ? (with boundaries S and ), we get


Z 
Z 
v
v
u
u
f
dS +
u
dS = 0.
(4.6)
v
v
n
n
n
n
S

since 2 u = 2 v = 0 for all point in V ? . By transforming to polar coordinates, dS = d


and u/n = u/ (unit normal is in the direction ) onto ; then
Z
Z
u
ln 2 u
v
dS =
d 0 as 0,
2 0
n
and also
Z 2
Z 2
Z
Z 2
1
1
1
v
v
d =
u d =
u d u(, ) as 0,
u dS =
u

2 0

2 0
n
0
and so, in the limit 0, equation (4.6) gives

Z 
v
u
dS, where
f
u(, ) =
v
n
n
S

v=


1
ln (x )2 + (y )2 .
4

60

4.3 Solving Poisson Equation Using Greens Functions

now, consider w, such that 2 w = 0 in V but with w regular at (x = , y = ), and with


w = v on S. Then Greens theorem gives


Z 
Z
Z 

w
w
u
u
u2 w w2 u dV =
u
dS
f
dS = 0
w
+v
n
n
n
n
S
V
S
since 2 u = 2 w = 0 in V and w = v on S. Then, subtract this equation from equation
above to get


Z 
Z
Z 
w
u
v
u

dS
f
dS = f
u(, ) =
v
f
+v
(v + w) dS.
n
n
n
n
n
S
S
S
Setting G(x, y; , ) = v + w, then
u(, ) =

G
dS.
n

Such a function G then has the properties,


2 G = (x ) in V,

4.3.3

with G = 0 on S.

Free Space Greens Function

We seek a Greens function G such that,


G(x, ) = v(x, ) + w(x, )

where

2 v = (x ) in V.

How do we find the free space Greens function v defined such that 2 v = (x ) in V ?
Note that it does not depend on the form of the boundary. (The function v is a source term
and for Laplaces equation is the potential due to a point source at the point x = .)
As an illustration of the method, we can derive that, in two dimensions,
v=


1
ln (x )2 + (y )2 ,
4

as we have already seen. We move to polar coordinate around (, ),


x = r cos

&

y = r sin ,

and look for a solution of Laplaces equation which is independent of and which is singular
as r 0.

Dr r

Cr

61

Chapter 4 Elliptic Equations


Laplaces equation in polar coordinates is


v
2 v 1 v
1
r
= 2+
=0
r r
r
r
r r

which has solution v = B ln r + A with A and B constant. Put A = 0 and, to determine


the constant B, apply Greens theorem to v and 1 in a small disc D r (with boundary Cr ), of
radius r around the origin (, ),
Z
Z
Z
v
2
(x ) dV = 1,
v dV =
dS =
Dr
Dr
Cr n
so we choose B to make

Cr

v
dS = 1.
n

Now, in polar coordinates, v/n = v/r = B/r and dS = rd (going around circle C r ).
So,
Z 2
Z 2
B
1
rd = B
d = 1 B = .
r
2
0
0

Hence,


1
1
1
ln r = ln r 2 = ln (x )2 + (y )2 .
2
4
4
(We do not use the boundary condition in finding v.)
v=

Similar (but more complicated) methods lead to the free-space Greens function v for the
Laplace equation in n dimensions. In particular,

n = 1,
|x |,


1
n = 2,
ln |x |2 ,
v(x, ) =
4

|x |2n , n 3,

(2 n)An (1)

where x and are distinct points and A n (1) denotes the area of the unit n-sphere. We shall
restrict ourselves to two dimensions for this course.
Note that Poissons equation, 2 u = F , is solved in unbounded Rn by
Z
v(x, ) F () d
u(x) =
Rn

where from equation (4.2) the free space Greens function v, defined above, serves as Greens
function for the differential operator 2 when no boundaries are present.

4.3.4

Method of Images

In order to solve BVPs for Poissons equation, such as 2 u = F in an open region V with
some conditions on the boundary S, we seek a Greens function G such that, in V
G(x, ) = v(x, ) + w(x, )

where 2 v = (x )

and 2 w = 0 or 1/V(V ).

62

4.3 Solving Poisson Equation Using Greens Functions

Having found the free space Greens function v which does not depend on the boundary
conditions, and so is the same for all problems we still need to find the function w, solution
of Laplaces equation and regular in x = , which fixes the boundary conditions (v does not
satisfies the boundary conditions required for G by itself). So, we look for the function which
satisfies
2 w = 0 or 1/V(V ) in V,

(ensuring w is regular at (, )),

with

w = v (i.e. G = 0) on S for Dirichlet boundary conditions,


w
v
G
or
=
(i.e.
= 0) on S for Neumann boundary conditions.
n
n
n
To obtain such a function we superpose functions with singularities at the image points of
(, )). (This may be regarded as adding appropriate point sources and sinks to satisfy the
boundary conditions.) Note also that, since G and v are symmetric then w must be symmetric
too (i.e. w(x, ) = w(, x)).
Example 1
Suppose we wish to solve the Dirichlet BVP for Laplaces equation
2 u =

2u 2u
+ 2 = 0 in y > 0
x2
y

with

u = f (x) on y = 0.

We know that in 2-D the free space function is


v=
If we superpose to v the function
w=+


1
ln (x )2 + (y )2 .
4

1
ln (x )2 + (y + )2 ,
4

solution of 2 w = 0 in V and regular at (x = , y = ), then




1
(x )2 + (y )2
G(x, y, , ) = v + w =
.
ln
4
(x )2 + (y + )2
G=v+w

y
x=

(, )

y =

y=

+ (, )

Note that, setting y = 0 in this gives,


1
ln
G(x, 0, , ) =
4

y
v

(x )2 + 2
(x )2 + 2

= 0, as required.

63

Chapter 4 Elliptic Equations


The solution is then given by
u(, ) =

G
dS.
n

Now, we want G/n for the boundary y = 0, which is



G
G
1
=
=


n S
y y=0
(x )2 + 2

Thus,

u(, ) =

(exercise, check this).

f (x)
dx,
(x )2 + 2

and we can relabel to get in the original variables


y
u(x, y) =

f ()
d.
( x)2 + y 2

Example 2
Find Greens function for the Dirichlet BVP
2 u =

2u 2u
+ 2 = F in the quadrant x > 0, y > 0.
x2
y

We use the same technique but now we have three images.

(, ) +

(, )
S
x

(, )

+ (, )

Then, the Greens function G is


G(x, y, , ) =

So,



1
1
ln (x )2 + (y )2 +
ln (x )2 + (y + )2
4
4


1
1
2

ln (x + ) + (y + )2 +
ln (x + )2 + (y )2 .
4
4

1
G(x, y, , ) =
ln
4

"

#

(x )2 + (y )2 (x + )2 + (y + )2
,
((x )2 + (y + )2 ) ((x + )2 + (y )2 )

and again we can check that G(0, y, , ) = G(x, 0, , ) = 0 as required for Dirichlet BVP.

64

4.3 Solving Poisson Equation Using Greens Functions

Example 3
Consider the Neumann BVP for Laplaces equation in the upper half-plane,
2 u =

2u 2u
+ 2 = 0 in y > 0
x2
y

with

u
u
=
= f (x) on y = 0.
n
y

(, )

x=

S
y

y=

y =

(, )
G=v+w

Add an image to make G/y = 0 on the boundary:


G(x, y, , ) =
Note that,



1
1
ln (x )2 + (y )2
ln (x )2 + (y + )2 .
4
4

G
1
=
y
4

2(y )
2(y + )
+
2
2
(x ) + (y )
(x )2 + (y + )2

and as required for Neumann BVP,






G
G
1
2
2
=
=
+
= 0.
n S
y y=0 4 (x )2 + 2 (x )2 + 2

Then, since G(x, 0, , ) = 1/2 ln (x )2 + 2 ,
Z +

1
u(, ) =
f (x) ln (x )2 + 2 dx,
2
Z +

1
i.e. u(x, y) =
f () ln (x )2 + y 2 d,
2
Remind that all the theory on Greens function has been developed in the case when the
equation is given in a bounded open domain. In an infinite domain (i.e. for external problems)
we have to be a bit careful since we have not given conditions on G and G/n at infinity.
For instance, we can think of the boundary of the upper half-plane as a semi-circle with
R +.
y

S2

S1

+R
x

65

Chapter 4 Elliptic Equations


Greens theorem in the half-disc, for u and G, is

Z 
Z

G
u
2
2
u
dS.
G u u G dV =
G
n
n
S
V

Split S into S1 , the portion along the x-axis and S2 , the semi-circular arc. Then, in the above
equation we have to consider the behaviour of the integrals
Z
Z
Z
Z
u
G
u
G
G
u
(1)
dS
R d and (2)
dS
R d
G
u
n
R
R
0
0
S2
S2 n

as R +. Greens function G is O(ln R) on S 2 , so from integral (1) we need u/R to


fall off sufficiently rapidly with the distance: faster than 1/(R ln R) i.e. u must fall off faster
than ln(ln(R)). In integral (2), G/R = O(1/R) on S 2 provides a more stringent constraint
since u must fall off more rapidly that O(1) at large R. If both integrals over S 2 vanish as
R + then we recover the previously stated results on Greens function.
Example 4
Solve the Dirichlet problem for Laplaces equation in a disc of radius a,


u
1 2u
1
2
r
+ 2 2 = 0 in r < a with u = f () on r = a.
u=
r r
r
r

+
y

(x, y)

S
r

(, )

Consider image of point P at inverse point Q


P = ( cos , sin ),
Q = (q cos , q sin ),
with q = a2 (i.e. OP OQ = a2 ).

1
ln (x )2 + (y )2
4


1
a2
a2
2
2
+
ln (x
cos ) + (y
sin ) + h(x, y, , )
4

G(x, y, , ) =

(with 2 + 2 = 2 ).

We need to consider the function h(x, y, , ) to make G symmetric and zero on the boundary.
We can express this in polar coordinates, x = r cos , y = r sin ,


1
(r cos a2 / cos )2 + (r sin a2 / sin )2
G(r, , , ) =
+ h,
ln
4
(r cos cos )2 + (r sin sin )2

 2
r + a4 /2 2a2 r/ cos( )
1
+ h.
ln
=
4
r 2 + 2 2r cos( )

66

4.3 Solving Poisson Equation Using Greens Functions

Choose h such that G = 0 on r = a,



 2
1
a + a4 /2 2a3 / cos( )
+ h,
G|r=a =
ln
4
a2 + 2 2a cos( )

 2 2
a + a2 2a cos( )
1
=
+h=0
ln
4
2 2 + a2 2a cos( )

1
h=
ln
4

2
a2

Note that,


 2

a4
a2 r
1
1
ln r 2 + 2 2
cos( ) +
ln
4

4
a2


1
r 2 2
2
=
ln a + 2 2r cos( )
4
a

w(r, , , ) =

is symmetric, regular and solution of 2 w = 0 in V . So,


 2

a + r 2 2 /a2 2r cos( )
1
ln
G(r, , , ) = v + w =
,
4
r 2 + 2 2r cos( )

G is symmetric and zero on the boundary. This enable us to get the result for Dirichlet
problem for a circle,

Z 2
G
a d,
u(, ) =
f ()
r r=a
0
where

1
G
=
r
4

2r2 /a2 2 cos( )


2r 2 cos( )
2
2
2
2
2
a + r /a 2r cos( ) r + 2 2r cos( )

so




G
1
2 /a cos( )
G
a cos( )
=
=
,

n S
r r=a 2 a2 + 2 2a cos( ) a2 + 2 2a cos( )
=

Then

2 a 2
1
.
2 a a2 + 2 2a cos( )

1
u(, ) =
2

2
0

and relabelling,
a2 r 2
u(r, ) =
2

a2 2
f () d,
a2 + 2 2a cos( )

a2

r2

f ()
d.
2a r cos( )

Note that, from the integral form of u(r, ) above, we can recover the Mean Value Theorem.
If we put r = 0 (centre of the circle) then,
1
u(0) =
2

f () d,
0

i.e. the average of an harmonic function of two variables over a circle is equal to its value at
the centre.

67

Chapter 4 Elliptic Equations

Furthermore we may introduce more subtle inequalities within the class of positive harmonic
functions u 0. Since 1 cos( ) 1 then (a r) 2 a2 2ar cos( ) + r 2 (a + r)2 .
Thus, the kernel of the integrand in the integral form of the solution u(r, ) can be bounded
1 ar
a2 r 2
1
1 a+r

.
2
2
2 a + r
2 (a r) a 2ar cos( ) + r 2
2 a r
For positive harmonic functions u, we may use these inequalities to bound the solution of
Dirichlet problem for Laplaces equation in a disc
Z
Z
1 a + r 2
1 a r 2
f () d u(r, )
f () d,
2 a + r 0
2 a r 0
i.e. using the Mean Value Theorem we obtain Harnacks inequalities
a+r
ar
u(0) u(r, )
u(0).
a+r
ar
Example 5
Interior Neumann problem for Laplaces equation in a disc,


1
u
1 2u
2
u=
r
+ 2 2 = 0 in r < a,
r r
r
r
u
= f () on r = a.
n
Here, we need
1
G = (x )(y ) +
V
2

with


G
= 0,
r r=a

where V = a2 is the surface area of the disc. In order to deal with this term we solve the
equation



1
r2
1
2
r
=

(r)
=
+ c1 ln r + c2 ,
(r) =
r r
r
a2
4a2
and take the particular solution with c 1 = c2 = 0. Then, add in source at inverse point and
an arbitrary function h to fix the symmetry and boundary condition of G
G(r, , , ) =


1
ln r 2 + 2 2r cos( )
4
 2

a
r2
2 r 2
1
2
ln 2 a + 2 2r cos( ) +
+ h.

a
4a2

So,
G
h
1
2r 2 cos( )
1
2r 2a2 / cos( )
r
+
=

+
,
2
2
2
4
2
2
2
r
4 r + 2r cos( ) 4 r + a / 2a r/ cos( ) 2a
r




1
1
a cos( )
a a2 / cos( )
h
G
=

+
,
+
+
r r=a
2 a2 + 2 2a cos( ) a2 + a4 /2 2a3 / cos( )
2a
r r=a

1
h
1 a cos( ) + 2 /a cos( )
,
+
+
=
2
2 + a2 2a cos( )
2a
r
r=a

68

4.4 Extensions of Theory:



1
h
G
1
+
+
=
r r=a
2 a 2a r r=a

and


h
=0
r r=a

implies

G
= 0 on the boundary.
r

Then, put h 1/2 ln(a/) ; so,





 2 2 r 2
r2
1
2
2
+
.
G(r, , , ) = ln r + 2r cos( ) a + 2 2r cos( )
4
a
4 a2

On r = a,

h
2 i
1
1
ln a2 + 2 2a cos( )
+
,
4 
4
 1
1
=
ln a2 + 2 2a cos( )
.
2
2

G|r=a =

Then,

f () G|r=a a d,

Z 2 
 1
a
2
2
=u

ln a + 2a cos( )
f () d.
2 0
2

u(, ) = u
+

Now, recall the Neumann problem compatibility condition,


Z 2
f () d = 0.
0

Indeed,

u dV =

So the term involving

u
dS
n

from divergence theorem

f () d = 0.

f ()d in the solution u(, ) vanishes; hence

Z 2

a
ln a2 + 2 2a cos( ) f () d,
u(, ) = u

2 0
Z 2

a
or u(r, ) = u

ln a2 + r 2 2ar cos( ) f () d.
2 0
Exercise:

4.4

Exterior Neumann problem for Laplaces equation in a disc,


Z 2

a
ln a2 + r 2 2ar cos( ) f () d.
u(r, ) =
2 0

Extensions of Theory:

Alternative to the method of images to determine the Greens function G: (a) eigenfunction method when G is expended on the basis of the eigenfunction of the Laplacian
operator; conformal mapping of the complex plane for solving 2-D problems.
Greens function for more general operators.

Chapter 5

Parabolic Equations
Contents
5.1
5.2
5.3
5.4

5.1

Definitions and Properties . . . . . . . . . . . . . .


Fundamental Solution of the Heat Equation . . .
Similarity Solution . . . . . . . . . . . . . . . . . . .
Maximum Principles and Comparison Theorems

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

69
72
75
78

Definitions and Properties

Unlike elliptic equations, which describes a steady state, parabolic (and hyperbolic) evolution
equations describe processes that are evolving in time. For such an equation the initial state
of the system is part of the auxiliary data for a well-posed problem.
The archetypal parabolic evolution equation is the heat conduction or diffusion equation:
2u
u
=
t
x2

(1-dimensional),

or more generally, for > 0,


u
= ( u)
t
= 2 u ( constant),

2u
(1-D).
x2
Problems which are well-posed for the heat equation will be well-posed for more general
parabolic equation.
=

5.1.1

Well-Posed Cauchy Problem (Initial Value Problem)

Consider > 0,
u
= 2 u in Rn , t > 0,
t
with u = f (x) in Rn at t = 0,
and

|u| < in Rn , t > 0.


69

70

5.1 Definitions and Properties

Note that we require the solution u(x, t) bounded in R n for all t. In particular we assume
that the boundedness of the smooth function u at infinity gives u| = 0. We also impose
conditions on f ,
Z
Rn

|f (x)|2 dx < f (x) 0 as |x| .

Sometimes f (x) has compact support, i.e. f (x) = 0 outside some finite region.(E.g., in 1-D,
see graph hereafter.)
u

+ x

5.1.2

Well-Posed Initial-Boundary Value Problem

Consider an open bounded region of R n and > 0;


u
= 2 u in , t > 0,
t
with u = f (x) at t = 0 in ,
u
(x, t) = g(x, t) on the boundary .
and u(x, t) +
n
Then, = 0 gives the Dirichlet problem, = 0 gives the Neumann problem (u/n = 0
on the boundary is the zero-flux condition) and 6= 0, 6= 0 gives the Robin or radiation
problem. (The problem can also have mixed boundary conditions.)
If is not bounded (e.g. half-plane), then additional behavior-at-infinity condition may be
needed.

5.1.3

Time Irreversibility of the Heat Equation

If the initial conditions in a well-posed initial value or initial-boundary value problem for an
evolution equation are replaced by conditions on the solution at other than initial time, the
resulting problem may not be well-posed (even when the total number of auxiliary conditions
is unchanged). E.g. the backward heat equation in 1-D is ill-posed; this problem,
u
2u
=
in 0 < x < l, 0 < t < T,
t
x2
with u = f (x) at t = T, x (0, l),

and u(0, t) = u(l, t) = 0

for t (0, T ),

which is to find previous states u(x, t), (t < T ) which will have evolved into the state f (x),
has no solution for arbitrary f (x). Even when a the solution exists, it does not depend
continuously on the data.
The heat equation is irreversible in the mathematical sense that forward time is distinguishable from backward time (i.e. it models physical processes irreversible in the sense of the
Second Law of Thermodynamics).

71

Chapter 5 Parabolic Equations

5.1.4

Uniqueness of Solution for Cauchy Problem:

The 1-D initial value problem


2u
u
=
, x R, t > 0,
t
x2
with

u = f (x) at t = 0 (x R), such that

|f (x)|2 dx < .

has a unique solution.


Proof:
We can prove the uniqueness of the solution of Cauchy problem using the energy method.
Suppose that u1 and u2 are two bounded solutions. Consider w = u 1 u2 ; then w satisfies
w
2w
=
t
x2
with

w=0

( < x < , t > 0),

at t = 0

Consider the function of time


Z
1 2
I(t) =
w (x, t) dx,
2

( < x < )

such that

I(0) = 0

and

and


w
= 0, t.
x
I(t) 0 t (as w 2 0),

which represents the energy of the function w. Then,


Z
Z
Z
dI
1 w2
w
2w
dx (from the heat equation),
=
dx =
dx =
w
w
dt
2 t
t
x2




Z 
w
w 2
= w

dx (integration by parts),
x
x



Z 
w
w 2
=
dx 0 since
= 0.
x
x

Then,

0 I(t) I(0) = 0, t > 0,

since dI/dt < 0. So, I(t) = 0 and w 0 i.e. u 1 = u2 , t > 0.

5.1.5

Uniqueness of Solution for Initial-Boundary Value Problem:

Similarly we can make use of the energy method to prove the uniqueness of the solution of
the 1-D Dirichlet or Neumann problem
u
2u
=
in 0 < x < l, t > 0,
t
x2
with u = f (x) at t = 0, x (0, l),
or

u(0, t) = g0 (t)
u
(0, t) = g0 (t)
x

and

and

u(l, t) = gl (t), t > 0 (Dirichlet),


u
(l, t) = gl (t), t > 0 (Neumann).
x

72

5.2 Fundamental Solution of the Heat Equation

Suppose that u1 and u2 are two solutions and consider w = u1 u2 ; then w satisfies
w
2w
=
(0 < x < l, t > 0),
t
x2
with w = 0 at t = 0 (0 < x < l),
and
or

w(0, t) = w(l, t) = 0, t > 0 (Dirichlet),


w
w
(0, t) =
(l, t) = 0, t > 0 (Neumann).
x
x

Consider the function of time


I(t) =

1
2

w2 (x, t) dx,

such that

I(0) = 0

and I(t) 0 t

(as w 2 0),

which represents the energy of the function w. Then,


Z l
Z
dI
1 l w2
2w
=
dx =
,
w
dt
2 0 t
x2
0




Z l
Z l
w 2
w 2
w l

dx =
dx 0.
= w
x 0
x
x
0
0
Then,
0 I(t) I(0) = 0, t > 0,
since dI/dt < 0. So I(t) = 0 t > and w 0 and u 1 = u2 .

5.2

Fundamental Solution of the Heat Equation

Consider the 1-D Cauchy problem,


u
2u
on < x < , t > 0,
=
t
x2
with u = f (x) at t = 0 ( < x < ),
Z
such that
|f (x)|2 dx < .

Example: To illustrate the typical behaviour of the solution of this Cauchy problem, consider the specific case where u(x, 0) = f (x) = exp(x 2 ); the solution is


x2
1
exp
u(x, t) =
1 + 4t
(1 + 4t)1/2

(exercise: check this).

2
t exp(x
Starting with u(x, 0) = exp(x2 ) at t = 0, the solution becomes
u(x,
t)

1/2

/4t),
for t large, i.e. the amplitude of the solution scales as 1/ t and its width scales as t.

73

Chapter 5 Parabolic Equations

t=0
u

t=1
t = 10

Spreading of the Solution: The solution of the Cauchy problem for the heat equation
spreads such that its integral remains constant:
Z
Q(t) =
u dx = constant.

Proof:

Consider
dQ
=
dt

2u
dx (from equation),
2
x
 
u
=
= 0 (from conditions on u).
x

u
dx =
t

So, Q = constant.

5.2.1

Integral Form of the General Solution

To find the general solution of the Cauchy problem we define the Fourier transform of u(x, t)
and its inverse by
Z +
1
U (k, t) =
u(x, t) eikx dx,
2
Z +
1
u(x, t) =
U (k, t) eikx dk.
2
So, the heat equation gives,

Z + 
1
U (k, t)
2

+ k U (k, t) eikx dk = 0
t
2

x,

which implies that the Fourier transform U (k, t) satisfies the equation
U (k, t)
+ k 2 U (k, t) = 0.
t
The solution of this linear equation is
2

U (k, t) = F (k) ek t ,

74

5.2 Fundamental Solution of the Heat Equation

where F (k) is the Fourier transform of the initial data, u(x, t = 0),
Z +
1
f (x) eikx dx.
F (k) =
2
R +
(This requires |f (x)|2 dx < .) Then, we back substitute U (k, t) in the integral form
of u(x, t) to find,

Z + Z +
Z +
1
1
2
ik
k 2 t ikx
f () e
d ek t eikx dk,
u(x, t) =
F (k) e
e dk =
2
2

Z +
Z +
1
2
=
ek t eik(x) dk d.
f ()
2

Now consider
H(x, t, ) =

k 2 t ik(x)

dk =

"

x
exp t k i
2t

2

(x )2

4t

dk,

since the exponent satisfies




x
k t + ik(x ) = t k ik
t
2

= t

"

x
ki
2t

2

#
(x )2
,
+
4t2

and set k i(x )/2t = s/ t, with dk = ds, such that


r


Z +
(x)2 /4t
(x )2 ds
s2
=
H(x, t, ) =
e
,
e
exp
4t
t
t

Z +

2
since
es ds = (see appendix A).
So,

1
u(x, t) =
4 t
Where the function



Z +
(x )2
f () exp
d =
K(x , t) f () d
4t

 2
1
x
K(x, t) =
exp
4t
4 t
is called the fundamental solution or source function, Greens function, propagator, diffusion kernel of the heat equation.

5.2.2

Properties of the Fundamental Solution

The function K(x, t) is solution (positive) of the heat equation for t > 0 (check this) and has
a singularity only at x = 0, t = 0:

1. K(x, t) 0 as t 0+ with x 6= 0 (K O(1/ t exp[1/t])),

2. K(x, t) + as t 0+ with x = 0 (K O(1/ t)),

3. K(x, t) 0 as t + (K O(1/ t)),


Z
K(x , t) d = 1
4.

75

Chapter 5 Parabolic Equations

At any time t > 0 (no matter how small), the solution to the initial value problem for the heat
equation at an arbitrary point x depends on all of the initial data, i.e. the data propagate
with an infinite speed. (As a consequence, the problem is well posed only if behaviour-atinfinity conditions are imposed.) However, the influence of the initial state dies out very
rapidly with the distance (as exp(r 2 )).

5.2.3

Behaviour at large t

Suppose that the initial data have a compact support or decays to zero sufficiently quickly
as |x| and that we look at the solution of the heat equation on spatial scales, x, large
compared to the spatial scale of the data and at t large. Thus, we assume the ordering
x2 /t O(1) and 2 /t O() where  1 (so that, x/t O( 1/2 )). Then, the solution
Z
2
ex /4t +
2
f () e
d =
f () e /4t ex/2t d,
4 t

 2
2 /4t Z +
x
x
F (0)
e
,
f () d ' exp
'
4t
4 t
2t

1
u(x, t) =
4 t

(x)2 /4t

where F (0) is the Fourier transform of f at k = 0, i.e.


Z +
Z +

1
ikx
F (k) =
f (x) dx = 2 F (0).
f (x) e
dx
2

So, at large t, on large spatial


scales x the solution evolves as u ' u 0 / t exp( 2 ) where u0

is a constant and = x/ 2t is the diffusion variable. This solution spreads and decreases as
t increases.

5.3

Similarity Solution

For some equations, like the heat equation, the solution depends on a certain grouping of the
independent variables rather than depending on each of the independent variables independently. Consider the heat equation in 1-D
2u
u
D 2 = 0,
t
x
and introduce the dilatation transformation
= a x,

= b t

and w(, ) = c u(a , b ), R.

This change of variables gives


u
w
w
= c
= bc
,
t
t

and

u
w
w
= c
= ac
x
x

2
2
2u
ac w
2ac w
=

.
=

x2
2 x
2

So, the heat equation transforms into

bc

2w
w
2ac D
=0

i.e.

bc

w
2w
2ab D

= 0,

76

5.3 Similarity Solution

and is invariant under the dilatation transformation (i.e. ) if b = 2a. Thus, if u solves the
equation at x, t then w = c u solve the equation at x = a , t = b .
Note also that we can build some groupings of independent variables which are invariant
under this transformation, such as

a/b

a x
(b

t)

a/b

x
ta/b

which defines the dimensionless similarity variable (x, t) = x/ 2Dt, since b = 2a. (
if x or t 0 and = 0 if x = 0.) Also,
w
c/b

c u

(b t)

c/b

u
tc/b

= v()

suggests that we look for a solution of the heat equation of the form u = t c/2a v(). Indeed,
since the heat equation is invariant under the dilatation transformation, then we also expect
the solution to be invariant under that transformation. Hence, the partial derivatives become,
c

u
c c/2a1

1
=
t
v() + tc/2a v 0 ()
= tc/2a1
v() v 0 () ,
t
2a
t
2
a

since /t = x/(2t 2Dt) = /2t, and


u

tc/2a1/2 0
= tc/2a v 0 ()
=
v (),
x
x
2D

tc/2a1 00
2u
=
v ().
x2
2D

Then, the heat equation reduces to an ODE



(5.1)
t/21 v 00 () + v 0 () v() = 0.

with = c/a, such that u = t/2 v and = x/ 2Dt. So, we may be able to solve the heat
equation through (5.1) if we can write the auxiliary conditions on u, x and t as conditions
on v and . Note that, in general, the integral transform method is able to deal with more
general boundary conditions; on the other hand, looking for similarity solution permits to
solve other types of problems (e.g. weak solutions).

5.3.1

Infinite Region

Consider the problem


u
2u
= D 2 on < x < , t > 0,
t
x
with u = u0 at t = 0, x R , u = 0 at t = 0, x R+ ,
and u u0 as x ,

u 0 as x , t > 0.

u0

t=0

77

Chapter 5 Parabolic Equations

We look for a solution of the form u = t /2 v(), where (x, t) = x/ 2Dt, such that v() is
solution of equation (5.1). Moreover, since u = t /2 v() u0 as , where u0 does
not depend on t, must be zero. Hence, v is solution of the linear second order ODE
v 00 () + v 0 () = 0

with v u0 as

and v 0 as +.

Making use of the integrating factor method,


 2

 2 /2 0 
2
2 /2 00
e
v () + exp
v 0 () =
e
v () = 0 e /2 v 0 () = 0 ,
2

v () = 0 e

2 /2

v() = 0

h2 /2

dh + 1 = 2

/ 2

es ds + 1 .

Now, apply the initial conditions to determine the constants 2 and 1 . As , we

have v = 1 = u0 and as , v = 2 + u0 = 0, so 2 = u0 / . Hence, the solution


to this Cauchy problem in the infinite region is
!
!
Z 2
Z
/ 2

v() = u0

5.3.2

es ds

x / 4Dt

i.e.

u(x, t) = u0

es ds .

Semi-Infinite Region

Consider the problem


2u
u
= D 2 on 0 < x < , t > 0,
t
x
with u = 0 at t = 0, x R+ ,
u
and
= q at x = 0, t > 0, u 0 as x , t > 0.
x

Again, we look for a solution of the form u = t /2 v(), where (x, t) = x/ 2Dt, such that
v() is solution of equation (5.1). However, the boundary conditions are now different

t(1)/2 0

t(1)/2 0
u

/2 0

=
=t
v ()
=
v ()
v (0) = q,
x
x
x x=0
2D
2D

since q does notdepend on t, 1 must be zero. Hence from equation (5.1), the function v,
such that u = v t, is solution of the linear second order ODE

v 00 () + v 0 () v() = 0 with v 0 (0) = q 2D and v 0 as +.


Since the function v = is solution of the above ODE, we seek for solutions of the form
v() = () such that
v 0 = + 0

and v 00 = 20 + 00 .

Then, back-substitute in the ODE


00 + 20 + 2 0 + = 0

i.e.

2 + 2
2
00
=

= .
0

78

5.4 Maximum Principles and Comparison Theorems

After integration (integrating factor method or another), we get


2

2
1
2
e /2
ln | | = 2 ln
+ k = ln 2
+ k 0 = 0
= 0
2

2
2
0

An integration by part gives


"
# Z
2
es /2
() = 0

s2 /2

ds

e /2
+

+ 1 . = 2

es /2
ds + 1 .
s2

s2 /2

ds

+ 3 .

Hence, the solution becomes,



Z
2
v() = 2 e /2 +

es

2 /2


ds + 3 ,

where the constants of integration 2 and 3 are determined by the initial conditions:


Z
Z
2
0
s2 /2
2 /2
2 /2
v = 2 e
e
ds + e
+
+ 3 = 2
es /2 ds + 3 ,
0

so that v 0 (0) = 3 = q 2D. Also


 Z
as +, v 2

s2 /2

ds + 3

since

s2 /2

Z
ds = 2

h2

The solution of the equation becomes


v() = 2
= 2

5.4

e
e

2 /2

2 /2

=q

4D

u(x, y) = q

4Dt

Z
+ 2
Z
2

2 /2

/ 2

e
+
e
/ 2

h2

dh

dh

+
e
/ 2

Dt

= 0 2 = 3

r
2

=
.
dh =
2
2

Z
2

x2 /4Dt

h2

h2

2
,

+ 3 ,


+ 2
!

+ 3 ,
2

dh ,

x/ 4Dt

h2

dh .

Maximum Principles and Comparison Theorems

Like the elliptic PDEs, the heat equation or parabolic equations of most general form satisfy
a maximum-minimum principle.
Consider the Cauchy problem,
u
2u
=
t
x2

in

< x < , 0 t T.

and define the two sets V and VT as


and

V = {(x, t) (, +) (0, T )},

VT = {(x, t) (, +) (0, T ]}.

79

Chapter 5 Parabolic Equations


Lemma: Suppose

u 2 u
2 <0
t
x

in V

and u(x, 0) M,

then u(x, t) < M in VT .


Proof:

Suppose u(x, t) achieves a maximum in V , at the point (x 0 , t0 ). Then, at this point,


u
= 0,
t

u
=0
x

and

2u
0.
x2

But, 2 u/x2 0 is contradictory with the hypothesis 2 u/x2 > u/t = 0 at (x0 , t0 ).
Moreover, if we now suppose that the maximum occurs in t = T then, at this point
u
0,
t

u
=0
x

and

2u
0,
x2

which again leads to a contradiction.

5.4.1

First Maximum Principle

Suppose

u 2 u
2 0
t
x

in V

and u(x, 0) M,

then u(x, t) M in VT .
Proof: Suppose there is some point (x 0 , t0 ) in VT (0 < t T ) at which u(x0 , t0 ) = M1 > M .
Put w(x, t) = u(x, t) (t t0 ) where = (M1 M )/t0 < 0. Then,
u 2 u
w 2 w

=
2 |{z}
<0
t
x2
|t {z x }
>0

(in form of lemma),

and by lemma,

w(x, t) < max{w(x, 0)}

in VT ,

< M + t0 ,
M1 M
t0 ,
<M+
t0
w(x, t) < M1

in VT .

But, w(x0 , t0 ) = u(x0 , t0 ) (t0 t0 ) = u(x0 , t0 ) = M1 ; since (x0 , t0 ) VT we have a


contradiction.

80

5.4 Maximum Principles and Comparison Theorems

Appendix A
2
x
in R
Integral of e
Consider the integrals
I(R) =

s2

ds and

I=

such that I(R) I as R +. Then,


Z R
Z R
Z RZ
2
x2
y 2
I (R) =
e
dx
e
dy =
0

es ds

(x2 +y 2 )

dx dy =

e(x

2 +y 2 )

dx dy.

Since its integrand is positive, I 2 (R) is bounded by the following integrals


Z
Z
Z
2
2
(x2 +y 2 )
(x2 +y 2 )
e
dx dy <
e
dx dy <
e(x +y ) dx dy,

where : {x R+ , y R+ |x2 + y 2 = R2 } and + : {x R+ , y R+ |x2 + y 2 = 2R2 }.

R 2

R 2

R
R
+

R 2 x

Hence, after polar coordinates transformation, (x = cos , y = sin ), with dx dy = d d


and x2 + y 2 = 2 , this relation becomes
Z /2 Z R
Z /2 Z R2
2
2
e d d < I 2 (R) <
e d d.
0

Put, s =
so that ds = 2 d, to get
Z R2
Z 2
2R s

s
2
e ds < I (R) <
e ds i.e.
4 0
4 0
81





2
2
1 eR < I 2 (R) <
1 e2R .
4
4

82
Take the limit R + to state that

I2
4
4
So,

e
0

since exp(x2 ) is even on R.

s2

i.e.

I =
4
2

and I =
(I > 0).
2

Z +

es ds = ,

ds =
2

You might also like