Você está na página 1de 199

Partial Dierential Equations

T. Muthukumar
tmk@iitk.ac.in
April 18, 2014
ii
Contents
Notations vii
1 PDE: An Introduction 1
1.1 Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Well-Posedness of PDE . . . . . . . . . . . . . . . . . . . . . . 9
1.3 Three Basic PDE: History . . . . . . . . . . . . . . . . . . . . 11
1.4 Continuity Equation . . . . . . . . . . . . . . . . . . . . . . . 13
2 First Order PDE 17
2.1 Family Of Curves . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2 Linear Transport Equation . . . . . . . . . . . . . . . . . . . . 19
2.2.1 One Space Dimension . . . . . . . . . . . . . . . . . . . 19
2.2.2 Higher Dimension . . . . . . . . . . . . . . . . . . . . . 20
2.2.3 Inhomogeneous Transport Equation . . . . . . . . . . . 21
2.3 Integral Surfaces and Monge Cone . . . . . . . . . . . . . . . . 22
2.3.1 Quasi-linear Equations . . . . . . . . . . . . . . . . . . 22
2.3.2 Nonlinear Equations . . . . . . . . . . . . . . . . . . . 23
2.4 Method of Characteristics . . . . . . . . . . . . . . . . . . . . 24
2.5 Complete Integrals . . . . . . . . . . . . . . . . . . . . . . . . 32
2.5.1 Envelopes and General Integrals . . . . . . . . . . . . . 33
2.5.2 Method Of Characteristics . . . . . . . . . . . . . . . . 34
2.6 Cauchy Problem . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.6.1 Quasilinear . . . . . . . . . . . . . . . . . . . . . . . . 38
2.6.2 Nonlinear . . . . . . . . . . . . . . . . . . . . . . . . . 48
2.7 Non-smooth Solutions and Shocks . . . . . . . . . . . . . . . . 57
3 Classication by Characteristics 61
3.1 Semilinear . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
iii
CONTENTS iv
3.1.1 Semilinear: Two Dimension . . . . . . . . . . . . . . . 63
3.2 Quasilinear . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
3.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.4 System of First Order PDE . . . . . . . . . . . . . . . . . . . 70
3.5 System of Second Order PDE . . . . . . . . . . . . . . . . . . 72
3.6 Invariance of Discriminant . . . . . . . . . . . . . . . . . . . . 73
3.7 Standard or Canonical Forms . . . . . . . . . . . . . . . . . . 74
3.8 Reduction to Standard Form . . . . . . . . . . . . . . . . . . . 75
4 Wave Equation 81
4.1 One Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . 81
4.1.1 The Vibrating String: Derivation . . . . . . . . . . . . 81
4.1.2 Travelling Waves . . . . . . . . . . . . . . . . . . . . . 83
4.1.3 Domain of Dependence and Inuence . . . . . . . . . . 88
4.1.4 Standing Waves: Separation of Variable . . . . . . . . 91
4.2 Higher Dimensions . . . . . . . . . . . . . . . . . . . . . . . . 95
4.2.1 Spherical Means . . . . . . . . . . . . . . . . . . . . . . 96
4.2.2 Odd Dimension . . . . . . . . . . . . . . . . . . . . . . 100
4.2.3 Inhomogeneous Wave equation . . . . . . . . . . . . . . 101
4.3 Eigenvalue Problem of Laplacian . . . . . . . . . . . . . . . . 103
4.3.1 In Rectangle . . . . . . . . . . . . . . . . . . . . . . . . 104
4.3.2 In Disk . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
5 Heat Equation 107
5.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
5.2 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . 108
5.3 Heat Flow on a Bar . . . . . . . . . . . . . . . . . . . . . . . . 108
5.4 On a Circular Wire . . . . . . . . . . . . . . . . . . . . . . . . 110
5.5 Inhomogeneous Equation . . . . . . . . . . . . . . . . . . . . . 112
5.6 Steady State Equation . . . . . . . . . . . . . . . . . . . . . . 113
5.7 Fundamental Solution of Heat Equation . . . . . . . . . . . . 114
6 The Laplacian 117
6.1 Properties of Laplacian . . . . . . . . . . . . . . . . . . . . . . 118
6.2 Ill-Posedness of Cauchy Problem . . . . . . . . . . . . . . . . . 121
6.3 Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . 122
6.4 Harmonic Functions . . . . . . . . . . . . . . . . . . . . . . . 124
6.4.1 Spherical Harmonics . . . . . . . . . . . . . . . . . . . 125
CONTENTS v
6.4.2 Properties of Harmonic Functions . . . . . . . . . . . . 130
6.5 Existence and Uniqueness . . . . . . . . . . . . . . . . . . . . 138
6.6 Perrons Method . . . . . . . . . . . . . . . . . . . . . . . . . 140
6.6.1 Non-existence of Solutions . . . . . . . . . . . . . . . . 141
6.6.2 Characterizing regular points . . . . . . . . . . . . . . 141
6.7 with Simple Geometry . . . . . . . . . . . . . . . . . . . . . 142
6.8 Poisson Equation . . . . . . . . . . . . . . . . . . . . . . . . . 150
6.8.1 Fundamental Solution of Laplacian . . . . . . . . . . . 151
6.8.2 Existence and Uniqueness . . . . . . . . . . . . . . . . 153
6.8.3 Greens Function . . . . . . . . . . . . . . . . . . . . . 157
6.8.4 Greens Function for half-space . . . . . . . . . . . . . 163
6.8.5 Greens Function for a disk . . . . . . . . . . . . . . . . 164
6.8.6 Conformal Mapping and Greens Function . . . . . . . 167
6.8.7 Dirichlet Principle . . . . . . . . . . . . . . . . . . . . 167
6.9 Neumann Boundary Condition . . . . . . . . . . . . . . . . . . 169
Appendices 171
A The Gamma Function 173
B Normal Vector of a Surface 175
C Implicit Function Theorem 177
D Divergence Theorem 179
E Surface Area and Volume of Disk in R
n
181
F Molliers and Convolution 185
G Duhamels Principle 187
Bibliography 189
Index 191
CONTENTS vi
Notations
Symbols

n
i=1

2
x
2
i

_

x
1
,

x
2
, . . . ,

xn
_
denotes an open subset of R
n
, not necessarily bounded
denotes the boundary of
D

1
x
1

1
. . .

n
xn
n
and = (
1
, . . . ,
n
)
Function Spaces
C
k
(X) is the class of all k-times (k 1) continuously dierentiable functions
on X
C

(X) is the class of all innitely dierentiable functions on X


C

c
(X) is the class of all innitely dierentiable functions on X with com-
pact support
General Conventions
B
r
(x) denotes the open disk with centre at x and radius r
S
r
(x) denotes the circle or sphere with centre at x and radius r
w
n
denotes the surface area of a n-dimensional sphere of radius 1.
vii
NOTATIONS viii
Chapter 1
PDE: An Introduction
A partial dierential equation (PDE) is an equation involving an unknown
function u of two or more variables and some or all of its partial derivatives.
The partial dierential equation is usually a mathematical representation of
problems arising in nature, around us. The process of understanding physical
systems can be divided in to three stages:
(i) Modelling the problem or deriving the mathematical equation (in our
case it would be formulating PDE). The derivation process is usually a
result of conservation laws or balancing forces.
(ii) Solving the equation (PDE). What do we mean by a solution of the
PDE?
(iii) Studying properties of the solution. Usually, we do not end up with a
denite formula for the solution. Thus, how much information about
the solution can one extract without any knowledge of the formula?
1.1 Denitions
Recall that the ordinary dierential equations (ODE) dealt with functions
of one variable, u : R R. The subset could have the interval form
(a, b). The derivative of u at x is dened as
u

(x) := lim
h0
u(x + h) u(x)
h
,
1
CHAPTER 1. PDE: AN INTRODUCTION 2
provided the limit exists. The derivative gives the slope of the tangent line
at x . How to generalise this notion of derivative to a function u :
R
n
R? These concepts are introduced in a course on multi-variable
calculus. However, we shall jump directly to concepts necessary for us to
begin this course.
Let be an open subset of R
n
and let u : R be a given function.
We denote the directional derivative of u at x , along a vector R
n
, as
u

(x) = lim
h0
u(x + h) u(x)
h
,
provided the limit exists. The directional derivative of u at x , along
the standard basis vectors e
i
= (0, 0, . . . , 1, 0, . . . , 0) is called the i-th partial
derivative of u at x and is given as
u
x
i
=
u
x
i
(x) = lim
h0
u(x +he
i
) u(x)
h
.
Similarly, one can consider higher order derivatives, as well. We now in-
troduce Schwartzs multi-index notation for derivative, which will be used
to denote a PDE in a concise form. A multi-index Z
n
+
is a n-tuple
= (
1
, . . . ,
n
) of non-negative integers and let [[ =
1
+ . . . +
n
. If
and are two multi-indices, then means
i

i
for all 1 i n.
Also, = (
1

1
, . . . ,
n

n
), ! =
1
! . . .
n
! and x

= x

1
1
. . . x
n
n
for any x R
n
. A k-degree polynomial in n variables can be represented as

||k
a

.
The partial dierential operator of order is denoted as,
D

1
x
1

1
. . .

n
x
n
n
=

||
x
1

1
. . . x
n
n
.
For k N, we dene D
k
u(x) := D

u(x) [ [[ = k. Thus, for k = 1, we


regard Du as being arranged in a vector,
=
_
D
(1,0,...,0)
, D
(0,1,0,...,0)
, . . . , D
(0,0,...,0,1)
_
=
_

x
1
,

x
2
, . . . ,

x
n
_
.
We call this the gradient vector. The dot product of the gradient vector with
itself := is called the Laplacian and :=

n
i=1

2
x
2
i
. Similarly, for
CHAPTER 1. PDE: AN INTRODUCTION 3
k = 2, we regard D
2
as being arranged in a matrix form (called the Hessian
matrix),
D
2
=
_
_
_
_
_
_

2
x
2
1
. . .

2
x
1
xn

2
x
2
x
1
. . .

2
x
2
xn
.
.
.

2
xnx
1
. . .

2
x
2
n
_
_
_
_
_
_
nn
.
The trace of the Hessian matrix is called the Laplace operator, denoted as
:=

n
i=1

2
x
2
i
. Note that under some prescribed order on multi-indices ,
D
k
u(x) can be regarded as a vector in R
n
k
. Then [D
k
u[ := (

||=k
[D

u[
2
)
1/2
.
In particular, [u[ = (

n
i=1
u
2
x
i
)
1/2
and [D
2
u[ = (

n
i,j=1
u
2
x
i
x
j
)
1/2
.
Example 1.1. Let u(x, y) : R
2
R be u(x, y) = ax
2
+ by
2
. Then
u = (u
x
, u
y
) = (2ax, 2by).
D
2
u =
_
u
xx
u
yx
u
xy
u
yy
_
=
_
2a 0
0 2b
_
Note that, for convenience, we can view u : R
2
R
2
and D
2
u : R
2

R
4
= R
2
2
, by assigning some ordering to the partial derivatives .
Denition 1.1.1. Let be an open subset of R
n
. A k-th order PDE F is a
given map F : R
n
k
R
n
k1
. . . R
n
R R having the form
F
_
D
k
u(x), D
k1
u(x), . . . Du(x), u(x), x
_
= 0, (1.1.1)
for each x and u : R is the unknown.
A general rst order PDE is of the form F(Du(x), u(x), x) = 0 and, in
particular, for a two variable function u(x, y) the PDE will be of the form
F(u
x
, u
y
, u, x, y) = 0. If u(x, y, z) is a three variable function, then the PDE
is of the form F(u
x
, u
y
, u
z
, u, x, y, z) = 0. A general second order PDE is of
the form F(D
2
u(x), Du(x), u(x), x) = 0.
As we know the PDE is a mathematical description of the behaviour of
the associated system. Thus, our foremost aim is to solve the PDEs for
the unknown function u, usually called the solution of the PDE. The rst
expected notion of solution is as follows:
Denition 1.1.2. We say u : R is a solution (in the classical sense)
to the k-th order PDE (1.1.1),
CHAPTER 1. PDE: AN INTRODUCTION 4
if u is k-times dierentiable with the k-th derivative being continuous
and u satises the equation (1.1.1).
Example 1.2. Consider the equation u
x
(x, y) = 0. If we freeze the variable y,
the equation is very much like an ODE. Integrating both sides, we would get
u(x, y) = f(y) as a solution, where f is any arbitrary function of y. Thus, the
family of solution depends on the choice of f C
1
. Similarly, the solution
of u
y
(x, y) = 0 is u(x, y) = g(x) for any choice of g C
1
.
Example 1.3. Consider the equation u
t
(x, t) = u(x, t). If we freeze the vari-
able x, the equation is very much like an ODE. Integrating both sides, we
would get u(x, t) = f(x)e
t
as a solution, where f is any arbitrary function of
x. Thus, the family of solution depends on the choice of f.
Example 1.4. Let us solve for u in the equation u
xy
= 4x
2
y. Unlike previous
example, the PDE here involves derivatives in both the variable. Still one
can solve this PDE for a general solution. We rst integrate w.r.t x both
sides to get u
y
= (4/3)x
3
y + f(y). Then, integrating again w.r.t y, we get
u(x, y) = (2/3)x
3
y
2
+ F(y) + g(x), where F(y) =
_
f(y) dy.
Example 1.5. Consider the equation u
x
(x, y) = u
y
(x, y). At rst look this
doesnt look simple for solving directly. But a change of coordinates rewrites
the equation in a simpler form. If we choose the coordinates w = x + y and
z = xy, we have u
x
= u
w
+u
z
and u
y
= u
w
u
z
. Substituting this, we get
u
z
(w, z) = 0 which is the form considered in Example 1.2. Hence its solution
is u(w, z) = g(w) for g C
1
and, hence, u(x, y) = g(x + y).
Observe in the above examples that, in contrast to ODE, the family of
solutions of a PDE may be indexed by a function rather than a constant.
Exercise 1. Determine a and b so that u(x, y) = e
ax+by
is a solution to
u
xxxx
+u
yyyy
+ 2u
xxyy
= 0.
Exercise 2. Determine the relation between a and b if u(x, y) = f(ax + by)
is a solution to 3u
x
7u
y
= 0 for any dierentiable function f such that
f

(z) ,= 0 for all real z. (Answer: a = 7b/3).


It is not necessary that the general form a solution has unique represe-
nation. Note that in the example below we have three dierent family of
solutions for the same PDE and we may have more!
Example 1.6. Consider the PDE u
t
(x, t) = u
xx
(x, t).
CHAPTER 1. PDE: AN INTRODUCTION 5
(i) Note that u(x, t) = c is a solution of the PDE, for any constant c R.
Thus, we have a family of solutions depending on c.
(ii) The function u : R
2
R dened as u(x, t) = (1/2)x
2
+t +c, for a given
constant c R, is a solution of the PDE u
t
= u
xx
. Because u
t
= 1,
u
x
= x and u
xx
= 1. We have another family of solutions for the same
PDE.
(iii) Note that u(x, t) = e
ax+bt
is a solution to the PDE u
t
= u
xx
if b = a
2
.
Note that u
t
= bu, u
x
= au and u
xx
= a
2
u.
As of today, there is no universal way of solving a given PDE. Thus,
the PDEs have to be categorised based on some common properties, all of
which may be expected to have a common technique to solve. One such
classication is given below.
Denition 1.1.3. We say F is linear if (1.1.1) has the form

||k
a

(x)D

u(x) = f(x)
for given functions f and a

([[ k). If f 0, we say F is homogeneous.


F is said to be semilinear, if it is linear only in the highest order, i.e., F has
the form

||=k
a

(x)D

u(x) + a
0
(D
k1
u, . . . , Du, u, x) = 0.
We say F is quasilinear if it has the form

||=k
a

(D
k1
u(x), . . . , Du(x), u(x), x)D

u + a
0
(D
k1
u, . . . , Du, u, x) = 0,
i.e., the highest order derivative coecient contains derivative only upto the
previous order. Finally, we say F is fully nonlinear if it depends nonlinearly
on the highest order derivatives.
Note that the classication is heirarchical, i.e., we have the inclusion
linear semilinear quasilinear fully nonlinear.
Thus, common sense tells that we classify a PDE based on the smallest class
it sits in the heirarchy.
CHAPTER 1. PDE: AN INTRODUCTION 6
Example 1.7. (i) a
1
(x)u
xx
+ a
2
(x)u
xy
+ a
3
(x)u
yy
+ a
4
(x)u
x
+ a
5
(x)u
y
=
a
6
(x)u is linear.
(ii) xu
y
yu
x
= u is linear.
(iii) xu
x
+ yu
y
= x
2
+y
2
is linear.
(iv) u
tt
c
2
u
xx
= f(x, t) is linear.
(v) y
2
u
xx
+ xu
yy
= 0 is linear.
(vi) u
x
+u
y
u
2
= 0 is semilinear.
(vii) u
t
+ uu
x
+ u
xxx
= 0 is semilinear.
(viii) u
2
tt
+u
xxxx
= 0 is semilinear.
(ix) u
x
+uu
y
u
2
= 0 is quasilinear.
(x) uu
x
+ u
y
= 2 is quasilinear.
(xi) u
x
u
y
u = 0 is nonlinear.
Example 1.8 (Examples of Linear PDE). Transport Equation u
t
(x, t)+b

x
u(x, t) = 0 for some given b R
n
assuming that x R
n
.
Laplace Equation u = 0.
Poisson Equation u(x) = f(x).
Poisson Equation u(x) = f(u).
Helmholtz Equation u +k
2
u = 0, for a given constant k.
Heat Equation u
t
u = 0.
Kolmogorovs Equation u
t
A D
2
u +b u = 0, for given nn matrix
A = (a
ij
) and b R
n
. The rst scalar product is in R
n
2
and the second
is in R
n
.
Wave Equation u
tt
u = 0
General Wave Equation u
tt
A D
2
u+b u = 0, for given nn matrix
A = (a
ij
) and b R
n
. The rst scalar product is in R
n
2
and the second
is in R
n
.
CHAPTER 1. PDE: AN INTRODUCTION 7
Schrodinger Equation iu
t
+ u = 0.
Airys Equation u
t
+ u
xxx
= 0.
Beam Equation u
t
+ u
xxxx
= 0.
Example 1.9 (Examples of Nonlinear PDE). Inviscid Burgers Equation
u
t
+ uu
x
= 0, for x R
Eikonal Equation [u(x)[ = f(x) is a rst order nonlinear equation. This
equation arises in geometrical optics, optimal control and computer
vision etc. In fact, the name eikon is a greek word for image.
Hamilton-Jacobi Equation u
t
+ H(u, x) = 0.
Minimal Surface Equation

_
u
_
1 +[u[
2
_
= f(x)
arises in geometry. The graph of the solution u dened on the domain
(say convex domain, for simplicity) has the given mean curvature
f : R. When f 0 the equation is called minimal surface
equation. This is a second order elliptic type PDE.
Image Processing A degenerate elliptic equation
u
_
u
[u[
_
= f (x)
is a system of three equations where the solution u : R
2
R
3
,
which measures the intensity of red, green and blue pixels iin a coloured
image. The problem is given a noisy image f : R
3
, we seek a de-
noised image u by diusing the noise in directions parallel to the image
edges. The > 0 is a diusive scaling.
Monge-Ampere Equation det(D
2
u) = f(x, u, u) is a fully nonlinear
PDE encountered in optimal transport problems. The gradient of the
solution u, u, maps optimal transportation path.
Schrodinger Equation iu
t
+ u V (u)u = 0. The solution u : R
n

[t
0
, ) C, a wavefunction, is associated with a particle of mass m
and driven by a potential V (u).
CHAPTER 1. PDE: AN INTRODUCTION 8
Korteweg de Vries (KdV) Equation u
t
+ u
x
+ uu
x
+u
xxx
= 0.
Exercise 3. Classify all the important PDE listed in Example 1.8.
Exercise 4. Classify the PDEs in the hierarchy of linearity:
(i) (y u)u
x
+ xu
y
= xy + u
2
.
(ii) uu
2
x
xu
y
=
2
x
u
3
.
(iii) x
2
u
x
+ (y x)u
y
= y sin u.
(iv) (sin y)u
x
e
x
u
y
= e
y
u.
(v) u
x
+ sin(u
y
) = u.
(vi) uu
x
+ x
2
u
yyy
+ sin x = 0.
(vii) u
x
+e
x
2
u
y
= 0.
(viii) u
tt
+ (sin y)u
yy
e
t
cos y = 0.
(ix) x
2
u
xx
+ e
x
u = xu
xyy
.
(x) e
y
u
xxx
+ e
x
u = sin y + 10xu
y
.
(xi) y
2
u
xx
+ e
x
uu
x
= 2xu
y
+ u.
(xii) u
x
u
xxy
+ e
x
uu
y
= 5x
2
u
x
.
(xiii) u
t
= k
2
(u
xx
+ u
yy
) + f(x, y, t).
(xiv) x
2
u
xxy
+ y
2
u
yy
log(1 + y
2
)u = 0.
(xv) u
x
+u
3
= 1.
(xvi) u
xxyy
+ e
x
u
x
= y.
(xvii) uu
xx
+u
yy
u = 0.
(xviii) u
xx
+ u
t
= 3u.
Exercise 5. Rewrite the following PDE in the new coordinates v and w.
(i) u
x
+u
y
= 1 for v = x + y and w = x y.
CHAPTER 1. PDE: AN INTRODUCTION 9
(ii) au
t
+ bu
x
= u for v = ax bt and w = t/a where a, b ,= 0.
(iii) au
x
+ bu
y
= 0 for v = ax + by and w = bx ay, where a
2
+ b
2
> 0.
(iv) u
tt
= c
2
u
xx
for v = x + ct and w = x ct.
(v) u
xx
+ 2u
xy
+ u
yy
= 0 for v = x and w = x y.
(vi) u
xx
2u
xy
+ 5u
yy
= 0 for v = x + y and w = 2x.
(vii) u
xx
+4u
xy
+4u
yy
= 0 for v = y 2x and w = x. (should get u
ww
= 0).
(viii) u
xx
+ 2u
xy
3u
yy
= 0 for v = y 3x and w = x + y.
1.2 Well-Posedness of PDE
We have seen through examples that a given PDE may one, many or no solu-
tion(s). The choice of our solution depends on the motivation for the study.
One needs to enforce extra side conditions to a particular solution. These
extra conditions are called boundary conditions (boundary value problem)
or whose value at initial time is known(initial value problem). It is usually
desirable to solve a well-posed problem, in the sense of Hadamard . By well-
posedness we mean that the PDE along with the boundary condition (or
initial condition)
(a) has a solution (existence)
(b) the solution is unique (uniqueness)
(c) the solution depends continuously on the data given (stability).
Any PDE not meeting the above criteria is said to be ill-posed. If too many
initial/boundary conditions are specied, then the PDE will have no solu-
tion. If too few initial/boundary conditions are specied, then the PDE will
not have unique solution. However, if the right amount of initial/boundary
conditions are specied, but at wrong places, then the solution may exist and
be unique except that that it may not depend continuously on the initial or
boundary data.
CHAPTER 1. PDE: AN INTRODUCTION 10
Example 1.10. The initial value problem
_
u
tt
= u
xx
in R (0, )
u(x, 0) = u
t
(x, 0) = 0
has the trivial solution u(x, t) = 0. We consider the same problem with a
small change in data, i.e.,
_
_
_
u
tt
= u
xx
in R (0, )
u(x, 0) = 0
u
t
(x, 0) = sin
_
x

_
which has the solution u

(x, t) =
2
sin(x/) sin(t/). Since
sup
(x,t)
[u

(x, t) u(x, t)[ =


2
sup
(x,t)
[sin(x/) sin(t/)[ =
2
the problem is well-posed.
Example 1.11 (Ill-posed). The initial value problem
_
u
tt
= u
xx
in R (0, )
u(x, 0) = u
t
(x, 0) = 0
has the trivial solution u(x, t) = 0. We consider the same problem with a
small change in data, i.e.,
_
_
_
u
tt
= u
xx
in R (0, )
u(x, 0) = 0
u
t
(x, 0) = sin
_
x

_
which has the solution u

(x, t) =
2
sin(x/) sinh(t/). Since
sup
(x,t)
[u

t
(x, t) u
t
(x, t)[ = sup
(x,t)
[sin(x/)[ =
and
lim
t
sup
x
[u

(x, t) u(x, t)[ = lim


t
sup
x
[sin(x/)[
= lim
t

2
[sinh(t/)[ = ,
the problem is ill-posed. Because a small change in initial data leads to a
large change in the solution.
CHAPTER 1. PDE: AN INTRODUCTION 11
Exercise 6. Consider the initial value problem
u
t
+ u
xx
= 0 (x, t) R (0, )
with u(x, 0) = 1.
(a) Show that u(x, t) 1 is a solution to this problem.
(b) Show that u
n
(x, t) = 1 +
e
n
2
t
n
sin(nx) is a solution to the problem with
initial value
u(x, 0) = 1 +
sin(nx)
n
.
(c) Find sup
x
[u
n
(x, 0) 1[.
(d) Find sup
x
[u
n
(x, t) 1[.
(e) Show that the problem is ill-posed.
The fundamental question is, given a PDE, nd these extra conditions
that make a PDE well-posedness.
1.3 Three Basic PDE: History
The study of partial dierential equations started as a tool to analyse the
models of physical science. The PDEs usually arise from the physical laws
such as balancing forces (Newtons law), momentum, conservation laws etc.
The rst PDE was introduced in 1752 by dAlembert as a model to study
vibrating strings. He introduced the one dimensional wave equation

2
u(x, t)
t
2
=

2
u(x, t)
x
2
.
This was then generalised to two and three dimensions by Euler (1759) and
D. Bernoulli (1762), i.e.,

2
u(x, t)
t
2
= u(x, t),
where =

3
i=1

2
x
2
i
.
CHAPTER 1. PDE: AN INTRODUCTION 12
In physics, a eld is a physical quantity associated to each point of space-
time. A eld can be classied as a scalar eld or a vector eld according to
whether the value of the eld at each point is a scalar or a vector, respec-
tively. Some examples of eld are Newtons gravitational eld, Coulombs
electrostatic eld and Maxwells electromagnetic eld.
Given a vector eld V , it may be possible to associate a scalar eld u,
called potential, such that u = V . Moreover, the gradient of any function
u, u is a vector eld. In gravitation theory, the gravity potential is the
potential energy per unit mass. Thus, if E is the potential energy of an
object with mass m, then u = E/m and the potential associated with a mass
distribution is the superposition of potentials of point masses.
The Newtonian gravitation potential can be computed to be
u(x) =
1
4
_

(y)
[x y[
dy
where (y) is the density of the mass distribution, occupying R
3
, at y. In
1782, Laplace discovered that the Newtons gravitational potential satises
the equation:
u = 0 on R
3
.
Thus, the operator = is called the Laplacian and any function whose
Laplacian is zero (as above) is said to be a harmonic function.
Later, in 1813, Poisson discovered that on the Newtonian potential
satises the equation:
u = on .
Such equations are called the Poisson equation. The identity obtained by
Laplace was, in fact, a consequence of the conservation laws and can be gener-
alised to any scalar potential. Green (1828) and Gauss (1839) observed that
the Laplacian and Poisson equations can be applied to any scalar potential
including electric and magnetic potentials. Suppose there is a scalar potential
u such that V = u for a vector eld V and V is such that
_

V d = 0
for all closed surfaces . Then, by Gauss divergence theorem
1
(cf.
Appendix D), we have
_

V dx = 0 .
1
a mathematical formulation of conservation laws
CHAPTER 1. PDE: AN INTRODUCTION 13
Thus, V = divV = 0 on and hence u = (u) = V = 0
on . Thus, the scalar potential is a harmonic function. The study of
potentials in physics is called Potential Theory and, in mathematics, it is
called Harmonic Analysis. Note that, for any potential u, its vector eld
V = u is irrotational, i.e., curl(V ) = V = 0.
Later, in 1822 J. Fourier on his work on heat ow in Theorie analytique
de la chaleur introduced the heat equation
u(x, t)
t
= u(x, t),
where =

3
i=1

2
x
2
i
. The heat ow model was based on Newtons law of
cooling.
Thus, by the beginning of 19th century, the three most important PDEs
were identied.
1.4 Continuity Equation
Let us consider an ideal compressible uid (viz. gas) occupying a bounded
region R
n
(in practice, we take n = 3, but the derivation is true for
all dimensions). For mathematical precision, we assume to be a bounded
open subset of R
n
. Let (x, t) denote the density of the uid for x at
time t I R, for some open interval I. Mathematically, we presume that
C
1
(I). We cut a region
t
and follow
t
, the position at time t,
as t varies in I. For mathematical precision, we will assume that
t
have C
1
boundaries (cf. Appendix D). Now, the law of conservation of mass states
that during motion the mass is conserved and mass is the product of density
and volume. Thus, the mass of the region as a function of t is constant and
hence its derivative should vanish. Therefore,
d
dt
_
t
(x, t) dx = 0.
We regard the points of
t
, say x
t
, following the trajectory x(t) with
velocity v(x, t). We also assume that the deformation of
t
is smooth, i.e.,
CHAPTER 1. PDE: AN INTRODUCTION 14
v(x, t) is continuous in a neighbourhood of I. Consider
d
dt
_
t
(x, t) dx = lim
h0
1
h
_
_

t+h
(x, t + h) dx
_
t
(x, t) dx
_
= lim
h0
_
t
(x, t +h) (x, t)
h
dx
+ lim
h0
1
h
_
_

t+h
(x, t + h) dx
_
t
(x, t + h) dx
_
The rst integral becomes
lim
h0
_
t
(x, t +h) (x, t)
h
dx =
_
t

t
(x, t) dx.
The second integral reduces as,
_

t+h
(x, t + h) dx
_
t
(x, t + h) dx =
_

(x, t + h)
_

t+h

t
_
=
_

t+h
\t
(x, t + h) dx

_
t\
t+h
(x, t + h) dx.
We now evaluate the above integral in the sense of Riemann. We x
t. Our aim is to partition the set (
t+h

t
) (
t

t+h
) with cylinders
and evaluate the integral by letting the cylinders as small as possible. To
do so, we choose 0 < s 1 and a polygon that covers
t
from outside
such that the area of each of the face of the polygon is less than s and the
faces are tangent to some point x
i

t
. Let the polygon have m faces.
Then, we have x
1
, x
2
, . . . x
m
at which the faces F
1
, F
2
, . . . , F
m
are a tangent
to
t
. Since
t+h
is the position of
t
after time h, any point x(t) moves to
x(t +h) = v(x, t)h. Hence, the cylinders with base F
i
and height v(x
i
, t)h is
expected to cover our annular region depending on whether we move inward
or outward. Thus, v(x
i
, t)(x
i
) is positive or negative depending on whether

t+h
moves outward or inward, where (x
i
) is the unit outward normal at
x
i

t
.
_

t+h
\t
(x, t+h) dx
_
t\
t+h
(x, t+h) dx = lim
s0
m

i=1
(x
i
, t)v(x
i
, t)(x
i
)hs.
CHAPTER 1. PDE: AN INTRODUCTION 15
Thus,
1
h
_
_

t+h
(x, t + h) dx
_
t
(x, t + h) dx
_
h0

_
t
(x, t)v(x, t) (x) d.
By Greens theorem (cf. Appendix D), we have
d
dt
_
t
(x, t) dx =
_
t
_

t
+ div(v)
_
dx.
Now, using conservation of mass, we get

t
+ div(v) = 0 in R. (1.4.1)
Equation (1.4.1) is called the equation of continuity. In fact, any quantity
that is conserved as it moves in an open set satises the equation of con-
tinuity (1.4.1).
CHAPTER 1. PDE: AN INTRODUCTION 16
Chapter 2
First Order PDE
In this chapter, we try to nd the general solutions and particular of rst
order PDE. A general rst PDE has the form
F(u(x), u(x), x) = 0.
2.1 Family Of Curves
We observed in the previous chapters that solutions of PDE occur as family of
curves given by a constant or arbitrary function. In fact, by eliminating the
constant or function, via dierentiation may lead to the dierential equation
it solves. We now look at some family of curves which arise as a solution to
rst order PDEs.
Let A R
2
be an open subset that represents a parameter set and con-
sider
u : R
2
A R
a two parameter family of smooth surfaces in R
3
, u(x, y, a, b), where (a, b)
A. For instance, u(x, y, a, b) = (x a)
2
+ (y b)
2
is a family of circles
with centre at (a, b). Dierentiate w.r.t x and y, we get u
x
(x, y, a, b) and
u
y
(x, y, a, b), respectively. Eliminating a and b from the two equations, we
get a rst order PDE
F(u
x
, u
y
, u, x, y) = 0
whose solutions are the given surfaces u.
Example 2.1. Consider the family of circles
u(x, y, a, b) = (x a)
2
+ (y b)
2
.
17
CHAPTER 2. FIRST ORDER PDE 18
Thus, u
x
= 2(x a) and u
y
= 2(y b) and eliminating a and b, we get
u
2
x
+ u
2
y
4u = 0
is a rst order PDE.
Example 2.2. Find the rst order PDE, by eliminating the arbitrary function
f, satised by u.
(i) u(x, y) = xy + f(x
2
+ y
2
)
(ii) u(x, y) = f(x/y)
Proof. (i) Dierentiating the given equation w.r.t x and y, we get
u
x
= y + 2xf

, u
y
= x + 2yf

,
respectively. Eliminating f

, by multiplying y and x respectively, we


get
yu
x
xu
y
= y
2
x
2
.
(ii) Dierentiating the given equation w.r.t x and y, we get
u
x
=
1
y
f

, , u
y
=
x
y
2
f

,
respectively. Eliminating f

, by multiplying x and y respectively, we


get
xu
x
+ yu
y
= 0.
Example 2.3. Find the rst order PDE, by eliminating the arbitrary constants
a and b, satised by u
(i) u(x, y) = (x + a)(y +b)
(ii) u(x, y) = ax + by
Proof. (i) Dierentiating the given equation w.r.t x and y, we get
u
x
= y + b, u
y
= x +a,
respectively. Eliminating a and b, we get
u
x
u
y
= u.
CHAPTER 2. FIRST ORDER PDE 19
(ii) Dierentiating the given equation w.r.t x and y, we get
u
x
= a, u
y
= b
respectively. Eliminating a and b, we get
xu
x
+ yu
y
= u.
Exercise 7. Find the rst order PDE, by eliminating the arbitrary function
f, satised by u(x, y) = e
x
f(2x y). (Answer: u
x
+ 2u
y
u = 0).
Exercise 8. Find the rst order PDE, by eliminating the arbitrary function
f, satised by u(x, y) = e
4x
f(2x 3y). (Answer: 3u
x
+ 2u
y
+ 12u = 0).
2.2 Linear Transport Equation
2.2.1 One Space Dimension
Derivation
Imagine a object (say, a wave) moving on the surface (of a water) with
constant speed b. At any time instant t, every point on the wave would have
travelled a distance of bt from its initial position. At any given instant t,
let u(x, t) denote the shape of the wave in space. Let us x a point (x
0
, 0)
on the wave. Now note that the value of u(x, t) is constant along the line
x = bt + x
0
or x bt = x
0
. Therefore the directional derivative of u in the
direction of (b, 1) is zero. Therefore
0 = u(x, t) (b, 1) = u
t
+ bu
x
.
This is a simple rst order linear equation called the transport equation.
Solving
We wish to solve the transport equation u
t
+ bu
x
= 0 which describes the
motion of an object moving with constant speed b, as seen by a xed observer
A.
Let us imagine that another observer B (say, on a skateboard) moving
with speed b observes the same object, moving in the direction of the object.
CHAPTER 2. FIRST ORDER PDE 20
For B the wave would appear stationary while for A, the xed observer,
the wave would appear to travel with speed b. What is the equation of the
motion of the object for the moving observer B? To understand this we need
to identify the coordinate system for B relative to A. Let us x a point x
at time t = 0. After time t, the point x remains as x for the xed observer
A, while for the moving observer B, the point x is now x bt. Therefore
the coordinate system for B is (, s) where = x bt and s = t. Let v(, s)
describe the motion of the object from Bs view. Then the PDE describing
the motion of object, as seen by B, is v
s
(, s) = 0. Therefore, v(, s) = f(),
for some arbitrary function f (suciently dierentiable), is the solution from
Bs perspective. To solve the problem from As perspective, we observe the
relations
u
t
= v

t
+ v
s
s
t
= bv

+ v
s
and
u
x
= v

x
+ v
s
s
x
= v

.
Therefore, u
t
+ bu
x
= bv

+ v
s
+ bv

= v
s
and, hence, u(x, t) = v(, s) =
f() = f(x bt).
2.2.2 Higher Dimension
We consider the homogeneous initial-value transport problem in higher di-
mensions. Given a vector b R
n
, we need to nd u : R
n
[0, ) R
satisfying
u
t
(x, t) + b u(x, t) = 0 in R
n
(0, ) (2.2.1)
By setting a new variable y = (x, t) in R
n
(0, ), (2.2.1) can be rewritten
as
(b, 1)
y
u(y) = 0 in R
n
(0, ).
This means that the directional derivative of u(y) along the direction (b, 1)
is zero. Thus, u must be constant along all lines in the direction of (b, 1).
The parametric representation of a line passing through a given point (x, t)
R
n
[0, ) and in the direction of (b, 1) is given by s (x + sb, t + s), for
all s t. Thus, u is constant on the line (x +sb, t +s) for all s t and,
in particular, the value of u at s = 0 and s = t are same. Hence,
u(x, t) = u(x tb, 0).
The procedure explained above can be formalised as below.
CHAPTER 2. FIRST ORDER PDE 21
The equation of a line passing through (x, t) and parallel to (b, 1) is (x, t)+
s(b, 1), for all s (t, ), i.e., (x + sb, t + s). Thus, for a xed (x, t)
R
n
(0, ), we set v(s) := u(x+sb, t +s) for all s (t, ). Consequently,
dv(s)
ds
= u(x + sb, t + s)
d(x + sb)
ds
+
u
t
(x +sb, t +s)
d(t + s)
ds
= u(x + sb, t + s) b +
u
t
(x +sb, t +s)
and from (2.2.1) we have
dv
ds
= 0 and hence v(s) constant for all s R.
Thus, in particular, v(0) = v(t) which implies that u(x, t) = u(x tb, 0).
If the value of u is known at time t = 0, for instance, u(x, 0) = g(x) on
R
n
t = 0 for a function g : R
n
R, then
u(x, t) = u(x tb, 0) = g(x tb).
Since (x, t) was arbitrary in R
n
(0, ), we have u(x, t) = g(x tb) for
all x R
n
and t 0. Thus, g(x tb) is a classical solution to (2.2.1) if
g C
1
(R
n
). If g , C
1
(R
n
), we shall call g(x tb) to be a weak solution of
(2.2.1).
2.2.3 Inhomogeneous Transport Equation
We now consider the inhomogeneous transport problem. Given a vector
b R
n
and a function f : R
n
(0, ) R, nd u : R
n
[0, ) R
satisfying
u
t
(x, t) + b u(x, t) = f(x, t) in R
n
(0, ) (2.2.2)
As before, we set v(s) := u(x + sb, t + s), for all s R, and for any given
point (x, t) R
n
(0, ). Thus,
dv(s)
ds
= b u(x + sb, t + s) +
u
t
(x + sb, t + s) = f(x + sb, t + s).
CHAPTER 2. FIRST ORDER PDE 22
Consider,
u(x, t) u(x tb, 0) = v(0) v(t)
=
_
0
t
dv
ds
ds
=
_
0
t
f(x + sb, t + s) ds
=
_
t
0
f(x + (s t)b, s) ds.
Thus, u(x, t) = u(x tb, 0) +
_
t
0
f(x + (s t)b, s) ds solves (2.2.2).
2.3 Integral Surfaces and Monge Cone
2.3.1 Quasi-linear Equations
We begin by describing the method for rst order quasi-linear PDE,
F(u, u, x) := b(x, u(x)) u(x) c(x, u(x)) = 0 for x ,
where R
n
open subset, b(x, u(x)) R
n
and c(x, u(x)) R. Thus,
we have (b(x, u(x)), c(x, u(x))) (u(x), 1) = 0. Finding the solution u is
equivalent to nding the surface S in R
n+1
which is graph of the solution u
of the given quasi-linear PDE, i.e.,
S = (x, z) R [ u(x) z = 0.
The equation of the surface S is given by G(x, z) = u(x) z. The normal
to S, at any point, is given by the gradient of G. Hence, G = (u(x), 1)
(cf. Appendix B). Therefore, for every point (x
0
, u(x
0
)) S, the coecient
vector (b(x
0
, u(x
0
)), c(x
0
, u(x
0
))) R
n+1
is perpendicular to the normal vec-
tor (u(x
0
), 1). Thus, the coecient vector must lie on the tangent plane
at (x
0
, u(x
0
)) of S. Dene the vector eld V (x, z) = (b(x, z), c(x, z)) formed
by the coecients of the quasi-linear PDE. Then, we note from the above
discussion that S must be such that the coecient vector eld V is tangential
to S at every point of S.
Denition 2.3.1. A curve in R
n
is said to be an integral curve for a given
vector eld, if the curve is tangent to the vector eld at each of its point.
CHAPTER 2. FIRST ORDER PDE 23
Similarly, a surface in R
n
is said to be an integral surface for a given vector
eld, if the surface is tangent to the vector eld at each of its point.
In the spirit of the above denition and arguments, nding a solution
to the quasi-linear PDE is equivalent to nding an integral surface S corre-
sponding to the coecient vector eld V . We view an integral surface w.r.t
V as an union of integral curves w.r.t V .
2.3.2 Nonlinear Equations
The arguments in previous section can be carried over to a general nonlinear
rst order PDE. Consider the rst order nonlinear PDE, F : R
n
R R
such that (cf. (1.1.1))
F(u(x), u(x), x) = 0 in ,
where R
n
, F is given and u is the unknown to be found. We want
nd a surface described by the solution u in R
n+1
. In the quasi-linear case,
using the equation, we obtained a unique direction (b, c) which is tangential,
at each point, to our desired surface S! In the non-linear case, however, we
have no such unique direction cropping out of the equation. For any point
x , the point (x, z), where z = u(x) is in the solution surface. Further,
if p = u(x) then we have the relation F(p, z, x) = 0. Thus, for a xed
(x, z) R, we consider the more general equation F(p, z, x) = 0 and
denote the solution set as
V (x, z) := p R
n
[ F(p, z, x) = 0.
Therefore, solving (2.5.1) is equivalent to nding a u C
1
() such that, for
all x , there is a pair (x, z) for which z = u(x) and p = u(x). Every
choice of p is a possible normal vector candidate (p, 1) at (x
0
, z
0
) on S.
In general, these family of normals envelope a cone with vertex at (x
0
, z
0
)
perpendicular to S. As p varies in V (x, z), we have a possible family of
(tangent) planes through (x
0
, z
0
) given by the equation
(z z
0
) = p (x x
0
),
where one of the planes is tangential to the surface S. The envelope of this
family is a cone C(x
0
, z
0
), called Monge cone, with vertex at (x
0
, z
0
). The
envelope of the family of planes is that surface which is tangent at each of
its point to some plane from the family.
CHAPTER 2. FIRST ORDER PDE 24
Denition 2.3.2. A surface S in R
n+1
is said to be an integral surface if
at each point (x
0
, z
0
) S R
n
R it is tangential to the Monge cone with
vertex at (x
0
, z
0
).
2.4 Method of Characteristics
We have already noted that solving a rst order PDE is equivalent to nding
an integral surface corresponding to the given PDE. The integral surfaces
are usually the union of integral curves, also called the characteristic curves.
Thus, nding an integral surface boils down to nding a family of character-
istic curves. The method of characteristics gives the equation to nd these
curves in the form of a system of ODE. The method of characteristics is a
technique to reduce a given rst order PDE into a system of ODE and then
solve the ODE using known methods, to obtain the solution of the rst order
PDE.
Let us consider the rst order nonlinear PDE (2.5.1) in new independent
variables p R
n
, z R and x . Consequently,
F(p, z, x) = F(p
1
, p
2
, . . . , p
n
, z, x
1
, x
2
, . . . , x
n
)
is a map of 2n + 1 variable. We now introduce the derivatives (assume it
exists) of F corresponding to each variable,
_

p
F = (F
p
1
, . . . , F
pn
)

x
F = (F
x
1
, . . . , F
xn
).
The method of characteristics reduces a given rst order PDE to a system of
ODE. The present idea is a generalisation of the idea employed in the study
of linear transport equation (cf. (2.2.1)). We must choose a curve x(s) in
such that we can compute u and u along this curve. In fact, we would
want the curve to intersect the boundary.
We begin by dierentiating F w.r.t x
i
in (2.5.1), we get
n

j=1
F
p
j
u
x
j
x
i
+ F
z
u
x
i
+ F
x
i
= 0.
Thus, we seek to nd x(s) such that
n

j=1
F
p
j
(p(s), z(s), x(s))u
x
j
x
i
(x(s)) + F
z
(p(s), z(s), x(s))p
i
(s)
+ F
x
i
(p(s), z(s), x(s)) = 0.
CHAPTER 2. FIRST ORDER PDE 25
To free the above equation of second order derivatives, we dierentiate p
i
(s)
w.r.t s,
dp
i
(s)
ds
=
n

j=1
u
x
i
x
j
(x(s))
dx
j
(s)
ds
and set
dx
j
(s)
ds
= F
p
j
(p(s), z(s), x(s)).
Thus,
dx(s)
ds
=
p
F(p(s), z(s), x(s)). (2.4.1)
Now substituting this in the rst order equation, we get
dp
i
(s)
ds
= F
z
(p(s), z(s), x(s))p
i
(s) F
x
i
(p(s), z(s), x(s)).
Thus,
dp(s)
ds
= F
z
(p(s), z(s), x(s))p(s)
x
F(p(s), z(s), x(s)). (2.4.2)
Similarly, we dierentiate z(s) w.r.t s,
dz(s)
ds
=
n

j=i
u
x
j
(x(s))
dx
j
(s)
ds
=
n

j=i
u
x
j
(x(s))F
p
j
(p(s), z(s), x(s))
Thus,
dz(s)
ds
= p(s)
p
F(p(s), z(s), x(s)). (2.4.3)
We have 2n+1 rst order ODE called the characteristic equations of (2.5.1).
The steps derived above can be summarised in the following theorem:
Theorem 2.4.1. Let u C
2
() solve (2.5.1) and x(s) solve (2.4.1), where
p(s) = u(x(s)) and z(s) = u(x(s)). Then p(s) and z(s) solve (2.4.2) and
(2.4.3), respectively, for all x(s) .
CHAPTER 2. FIRST ORDER PDE 26
We end this section with remark that for linear, semi-linear and quasi-
linear PDE one can do away with (2.4.2), the ODE corresponding to p,
because for these problems (2.4.3) and (2.4.1) form a determined system.
However, for a fully nonlinear PDE one needs to solve all the 3 ODEs to
compute the characteristic curve. The method of characteristics may be
generalised to higher order hyperbolic PDEs.
Remark 2.4.2. If the PDE is linear, i.e., a and b are independent of u,
then the characteristic curves are lying in the xy-plane. If the a and b are
constants (independent of both x and u) then the characteristic curves are
straight lines. In the linear case the characteristics curves will not intersect.
This is easily seen from the fact that if they intersect then, at the point of
intersection, they have the same tangent. This is not possible.
(a) In rst order linear problem, the ODE reduces to, as follows: Let
F(u, u, x) := b(x) u(x) +c(x)u(x) = 0 x .
Then, in the new variable, F(p, z, x) = b(x) p + c(x)z. Therefore, by
(2.4.1), we have
dx(s)
ds
=
p
F = b(x(s)).
Also,
dz(s)
ds
= b(x(s)) p(s) = b(x(s)) u(x(s)) = c(x(s))z(s).
(b) For a semi-linear PDE
F(u, u, x) := b(x) u(x) + c(x, u(x)) = 0 for x ,
we have
dx(s)
ds
=
p
F = b(x(s)).
Also,
dz(s)
ds
= b(x(s), z(s)) p(s) = b(x(s), z(s)) u(x(s)) = c(x(s), z(s)).
CHAPTER 2. FIRST ORDER PDE 27
(c) A quasi-linear PDE has the form
F(u, u, x) := b(x, u(x)) u(x) + c(x, u(x)) = 0 for x .
Then, in the new variable, F(p, z, x) = b(x, z) p +c(x, z). Therefore, by
(2.4.1), we have
dx(s)
ds
=
p
F = b(x(s), z(s)).
Also,
dz(s)
ds
= b(x(s), z(s)) p(s) = b(x(s), z(s)) u(x(s)) = c(x(s), z(s)).
Example 2.4. To understand the derivation of the characteristic equation and
the geometry involved, explained in this section, let us see what is happening
two varibale rst order quasi-linear equation:
a(x, y, u)u
x
+b(x, y, u)u
y
= c(x, y, u). (2.4.4)
Solving for u(x, y) in the above equation is equivalent to nding the surface
S (x, y, u(x, y)) generated by u in R
3
. If u is a solution of (2.4.4), at
each (x, y) in the domain of u,
a(x, y, u)u
x
+ b(x, y, u)u
y
= c(x, y, u)
a(x, y, u)u
x
+ b(x, y, u)u
y
c(x, y, u) = 0
(a(x, y, u), b(x, y, u), c(x, y, u)) (u
x
, u
y
, 1) = 0
(a(x, y, u), b(x, y, u), c(x, y, u)) (u(x, y), 1) = 0.
But (u(x, y), 1) is normal to S at the point (x, y) (cf. Appendix B).
Hence, the coecients (a(x, y, u), b(x, y, u), c(x, y, u)) are perpendicular to
the normal. Thus, the coecients (a(x, y, u), b(x, y, u), c(x, y, u)) lie on the
tangent plane to S at (x, y, u(x, y)). Hence, nding u is equivalent to nd-
ing the integral surface corresponding to the coecient vector eld V =
(a(x, y, u), b(x, y, u), c(x, y, u)).
The surface is the union of curves which satisfy the property of S. Thus,
for any curve S such that at each point of , the vector V (x, y) =
(a(x, y, u), b(x, y, u), c(x, y, u) is tangent to the curve. Parametrizing the
curve by the variable s, we see that we are looking for the curve =
x(s), y(s), z(s) R
3
such that
dx
ds
= a(x(s), y(s), u(x(s), y(s))),
dy
ds
= b(x(s), y(s), u(x(s), y(s))),
CHAPTER 2. FIRST ORDER PDE 28
and
dz
ds
= c(x(s), y(s), u(x(s), y(s))).
The three ODEs obtained are called characteristic equations. The union
of these characteristic (integral) curves give us the integral surface.
We shall now illustrate the method of characteristics for various examples
of rst order PDE.
Example 2.5 (Linear Transport Equation). We have already solved the linear
transport equation by elementary method. We solve the same now using
method of characteristics. Consider the linear transport equation in two
variable,
u
t
+ bu
x
= 0, x R and t (0, ),
where the constant b R is given. Thus, the given vector eld V (x, t) =
(b, 1, 0). The characteristic equations are
dx
ds
= b,
dt
ds
= 1, and
dz
ds
= 0.
Solving the 3 ODEs, we get
x(s) = bs + c
1
, t(s) = s + c
2
, and z(s) = c
3
.
Eliminating the parameter s, we get the curves (lines) x bt = a constant
and z = a constant. Therefore, u(x, t) = f(x bt) is the general solution,
for an arbitrary function f.
In the inhomogeneous equation case, where the ODE corresponding to z
is not zero, we intend to seek a function (x, u) such that the solution u is
dened implicitly by (x, u) = a constant. Suppose there is such a function
then, by setting z := u,

x
i
+
z
u
x
i
= 0 i = 1, 2, . . . , n.
Assuming
z
,= 0, we get
u
x
i
=

x
i

z
i = 1, 2, . . . , n.
If u solves b(x, u) u(x) = c(x, u) then (b(w), c(w))
w
= 0 is a homoge-
neous rst order equation of in n + 1 variables. In this case,
dw(s)
ds
= (b, c) and
d(w(s))
ds
= 0.
CHAPTER 2. FIRST ORDER PDE 29
Example 2.6 (Inhomogeneous Transport Equation). Given a constant b R
and a function f(x, t) , we wish to solve the inhomogeneous linear transport
equation,
u
t
(x, t) + bu
x
(x, t) = f(x, t), x R and t (0, ).
As before, the rst two ODE will give the projection of characteristic curve
in the xt plane, x bt = a constant, and the third ODE becomes
dz(s)
ds
= f(x(s), t(s)).
Lets say we need to nd the value of u at the point (x, t). The parametri-
sation of the line passing through (x, t) and (x bt, 0) in the rst variable
is x(s) = x bt + bs and in the second variable is t(s) = s, as s varies from
0 to t. If z has to be on the integral curve, then z(s) = u(x bt + bs, s).
Therefore, the third ODE becomes,
dz(s)
ds
= f(x(s), s) = f(x bt +bs, s).
Integrating both sides from 0 to t, we get
_
t
0
f(x b(t s), s) ds = z(t) z(0)
= u(x, t) u(x bt, 0).
Thus,
u(x, t) = u(x bt, 0) +
_
t
0
f(x b(t s), s) ds
is the general solution.
Example 2.7. Let us compute the general solution (in terms of arbitrary
functions) of the rst order PDE xu
x
(x, y) + yu
y
(x, y) = u(x, y). The char-
acteristic equations (ODEs) are
dx
ds
= x(s)
dy
ds
= y(s) and
dz
ds
= z(s).
Thus, x(s) = c
1
e
s
, y(s) = c
2
e
s
and z(s) = c
3
e
s
. Eliminating the parameter s,
we get y/x = c
4
and z/x = c
5
. Thus, the general solution is F(y/x, z/x) = 0
for an arbitrary function F. Explicitly,
u(x, y) = xg(y/x) or u(x, y) = yf(x/y),
CHAPTER 2. FIRST ORDER PDE 30
for some arbitrary smooth functions f and g. Compare this answer with
Example 2.3(ii).
Example 2.8. Let us compute the general solution (in terms of arbitrary
functions) of the rst order PDE yu(x, y)u
x
(x, y) + xu(x, y)u
y
(x, y) = xy.
The characteristic equations are
dx
ds
= yz,
dy
ds
= xz and
dz
ds
= xy.
Hence,
0 = x
dx
ds
y
dy
ds
=
d(x
2
)
ds

d(y
2
)
ds
=
d(x
2
y
2
)
ds
.
Thus, x
2
y
2
= c
1
and, similarly, x
2
z
2
= c
2
. Hence, the general solution
is F(x
2
y
2
, x
2
z
2
) = 0 for some arbitrary function F. Explicitly, for some
f and g,
u
2
(x, y) = x
2
+ f(x
2
y
2
) or u
2
(x, y) = y
2
+g(x
2
y
2
).
Example 2.9. Let us compute the general solution (in terms of arbitrary
functions) of the rst order PDE 2yu
x
+ uu
y
= 2yu
2
. The characteristic
equations (ODEs) are
dx
ds
= 2y(s)
dy
ds
= z(s) and
dz
ds
= 2y(s)z
2
(s).
Solving in the parametric form is quite cumbersome, because we will have a
second order nonlinear ODE of y, y

(s) = 2y(y

)
2
. However, for u ,= 0, we get
dz
dy
= 2y(s)z(s) solving which we get ln [z[ = y
2
+ c
1
or z = c
2
e
y
2
. Similarly,
dy
dx
=
z
2y
=
c
2
e
y
2
2y
solving which we get c
2
x + e
y
2
= c
3
or, equivalently,
xze
y
2
+ e
y
2
= c
3
. Thus, the general solution is
F(e
y
2
(1 +xz), e
y
2
z) = 0.
or
u(x, y) = f
_
e
y
2
(1 + xu)
_
e
y
2
CHAPTER 2. FIRST ORDER PDE 31
for some arbitrary smooth functions f. Note that u 0 is a solution if we
choose f = 0. The characteristic curves are (1 + xu)e
y
2
= a constant and
along these curves ue
y
2
is constant.
Example 2.10. Let us compute the general solution (in terms of arbitrary
functions) of the rst order PDE u
x
+2xu
y
= u
2
. The characteristic equations
(ODEs) are
dx
ds
= 1
dy
ds
= 2x(s) and
dz
ds
= z
2
(s).
Solving which we get x(s) = s + c
1
, y(s) = s
2
+ 2c
1
s + c
2
and z(s) =
1/(s +c
3
). Eliminating s between x and y, we get the characteristic curves
to be y x
2
= a constant and x + 1/z = a constant. Thus, the general
solution is F(y x
2
, x + 1/z) = 0. Explicitly,
u(x, y) =
1
x + f(y x
2
)
for some arbitrary smooth functions f.
Example 2.11. Let us compute the general solution (in terms of arbitrary
functions) of the rst order PDE yu
x
xu
y
= 2xyu. The characteristic
equations (ODEs) are
dx
ds
= y(s)
dy
ds
= x(s) and
dz
ds
= 2xyz.
To avoid cumbersome ODE, let us begin by assuming y ,= 0, then dividing
the entire equation by y, we get
dx
ds
= 1
dy
ds
= x(s)/y(s) and
dz
ds
= 2xz.
Solving which we get x(s) = s + c
1
, y(s) = s
2
2c
1
s + 2c
2
and [z(s)[ =
c
3
e
s
2
+2c
1
s
. Eliminating s between x and y, we get the characteristic curves
to be y
2
+ x
2
= a constant and z = c
4
e
x
2
. Thus, the general solution is
F(y
2
+ x
2
, e
x
2
z) = 0. Explicitly,
u(x, y) = f(y
2
+x
2
)e
x
2
for some arbitrary smooth functions f.
CHAPTER 2. FIRST ORDER PDE 32
Example 2.12. Let us compute the general solution (in terms of arbitrary
functions) of the rst order PDE u
x
1
+e
x
1
u
x
2
+e
x
3
u
x
3
= (2x
1
+e
x
1
)e
u
. The
characteristic equations (ODEs) are
dx
1
ds
= 1
dx
2
ds
= e
x
1
dx
3
ds
= e
x
3
and
dz
ds
= (2x
1
+ e
x
1
)e
z(s)
.
Solving which we get x
1
(s) = s +c
1
, x
2
(s) = e
c
1
e
s
+c
2
, e
x
3
(s)
= s +c
3
and
e
z(s)
= s
2
2c
1
s e
s+c
1
+ c
4
.
Eliminating s between x
1
and x
2
, we get x
2
e
x
1
= a constant, e
x
3
+x
1
= a
constant and the general solution is F(x
2
e
x
1
, e
x
3
+x
1
, e
u
+x
2
1
+e
x
1
) = 0.
Explcitly,
e
u
= f(x
2
e
x
1
, e
x
3
+ x
1
) x
2
1
e
x
1
for some arbitrary smooth functions f.
2.5 Complete Integrals
In this section, we study the form of general solutions of a rst order PDE,
i.e.,
F(u(x), u(x), x) = 0 in . (2.5.1)
Let A R
n
be an open set which is the parameter set. Let us introduce the
n (n + 1) matrix
(D
a
u, D
2
xa
u) :=
_
_
_
u
a
1
u
x
1
a
1
. . . u
xna
1
.
.
.
.
.
.
.
.
.
.
.
.
u
an
u
x
1
an
. . . u
xnan
_
_
_
.
Denition 2.5.1. A C
2
function u = u(x; a) is said to be a complete integral
in A if u(x; a) solves (2.5.1) for each a A and the rank of the matrix
(D
a
u, D
2
xa
u) is n.
The condition on the rank of the matrix means that u(x; a) strictly de-
pends on all the n components of a.
Example 2.13. The nonlinear equation u
x
u
y
= u has a complete integral
u(x, y; a, b) = xy +ab + (a, b) (x, y). We shall derive this solution in Exam-
ple 2.22.
CHAPTER 2. FIRST ORDER PDE 33
Example 2.14. The nonlinear equation u
2
x
+ u
y
= 0 has a complete integral
u(x, y; a, b) = ax ya
2
+ b.
Example 2.15. For any given f : R
n
R, the complete integral of the
Clairauts equation
x u + f(u) = u
is
u(x; a) = a x + f(a) a R
n
.
Example 2.16. The complete integral of the eikonal equation [u[ = 1 is
u(x; a, b) = a x + b for all a S(0, 1) and b R.
Example 2.17. The Hamilton-Jacobi is a special case of the nonlinear equa-
tion where F(x, z, p) = p
n
+ H(x, p
1
, . . . , p
n1
) where H is independent of z
and p
n
. For any given H : R
n
R, the complete integral of the Hamilton-
Jacobi equation
u
t
+ H(u) = 0
is
u(x, t; a, b) = a x tH(a) +b a R
n
, b R.
2.5.1 Envelopes and General Integrals
Denition 2.5.2. Let R
n
and A R
m
be open subsets and let u =
u(x; a) be a C
1
function of both x and a. Suppose the equation D
a
u(x; a) = 0
is solvable for a, as a C
1
function of x, say a = (x), i.e., D
a
u(x; (x)) = 0,
then v(x) := u(x; (x)) is the envelope of the functions u(; a)
aA
.
The idea is that for each x , the graph of v is tangent to the graph of
u(; a) for a = (x).
Example 2.18. The complete integral of the nonlinear PDE u
2
(1+[u[
2
) = 1
is u(x; a) = (1 [x a[
2
)
1/2
with [x a[ < 1. Now, solving D
a
u =

xa
(1|xa|
2
)
1/2
= 0 for a, we get a = (x) := x. Thus, the envelope is v(x) =
1.
Theorem 2.5.3. Suppose for each a A, u(; a) is a solution to (2.5.1)
and the envelope v of u, given as v(x) = u(x, (x)), exists then v also solves
(2.5.1).
CHAPTER 2. FIRST ORDER PDE 34
Proof. Since v(x) = u(x; (x)), for each i = 1, 2, . . . , n,
v
x
i
(x) = u
x
i
+
m

j=1
u
a
j

j
x
i
(x) = u
x
i
because D
a
u(x; (x)) = 0. Therefore,
F(v(x), v(x), x) = F(u(x; (x)), u(x; (x)), x) = 0.
Denition 2.5.4. The general integral is the envelope w of the functions
u(x; a

, h(a

)) where a

:= (a
1
, a
2
, . . . , a
n1
) and h : A

R
n1
R, provided
it exists.
Example 2.19. The envelope of the nonlinear equation u
x
u
y
= u, considered
in Example 2.13, is u(x, y; a, b) = xy + ab + (a, b) (x, y). Let h : R R be
dened as h(a) = a, then u(x, y; a, h(a)) = u(x, y; a, a) = xy +a
2
+a(x +y).
Solve D
a
u = 2a +x +y = 0 which yields a = (x) :=
(x+y)
2
. Therefore, the
envelope w(x) = u(x, y; (x), h((x))) = (x y)
2
/4.
Example 2.20. Consider the eikonal equation [u(x, y)[ = 1 in two di-
mension, i.e.,
_
u
2
x
+ u
2
y
= 1 whose complete integral is u(x, y; a
1
, a
2
) =
(x, y) (cos a
1
, sin a
1
) + a
2
. Consider h 0, then u(x, y; a
1
, h(a
1
)) = (x, y)
(cos a
1
, sin a
1
). Thus, solving D
a
1
u = xsin a
1
+ y cos a
1
= 0, we get
a
1
= arctan(y/x). Since
cos(arctan(z)) =
1

1 +z
2
and sin(arctan(z)) =
z

1 + z
2
,
we have the envelope w(x) =
_
x
2
+ y
2
for non-zero vectors.
Example 2.21. Consider the Hamilton-Jacobi Equation u
t
+[u[
2
= 0 whose
complete integral is u(x, t; a, b) = x a t[a[
2
+ b, where a R
n
and b R.
Consider h 0, then u(x, t; a, h(a)) = x a t[a[
2
. Thus, solving D
a
u =
x 2ta = 0, we get a =
x
2t
. We get the envelope w(x) =
|x|
2
4t
.
2.5.2 Method Of Characteristics
We have already derived the characteristic equation for a rst order PDE.
We illustrate through examples its application for nonlinear PDE.
CHAPTER 2. FIRST ORDER PDE 35
Example 2.22. Let us compute a complete integral of the rst order PDE
u
x
u
y
= u(x, y). The equation is of the form F(p, z, x) = p
1
p
2
z. The
characteristic equations are (using (2.4.1))
_
dx(s)
ds
,
dy(s)
ds
_
= (p
2
(s), p
1
(s)),
(using (2.4.2))
dp(s)
ds
= p(s)
and (using (2.4.3))
dz(s)
ds
= (p
1
(s), p
2
(s)) (p
2
(s), p
1
(s)) = 2p
1
(s)p
2
(s)
Thus, on integrating, we get p
1
(s) = c
1
e
s
and p
2
(s) = c
2
e
s
. Solving for z, we
get
z(s) = c
1
c
2
e
2s
+ c
3
.
Using p, we solve for x to get x(s) = c
2
e
s
+b and y(s) = c
1
e
s
+a.Therefore,
u(x, y) = (y a)(x b) + c
3
is a complete integral for arbitrary constants a and b, if we choose c
3
= 0.
We have already seen in Example 2.13 that u(x, y) = xy +ab + (a, b) (x, y)
is a complete integral.
Example 2.23. Let us nd a dierent complete integral of the nonlinear PDE
u
x
u
y
= u. Note that F(p, z, x) = p
1
p
2
z. Then the ODE
_
dx(s)
ds
,
dy(s)
ds
_
= (p
2
(s), p
1
(s)),
dp(s)
ds
= p(s)
Thus, on integrating, we get p
1
(s) = c
1
e
s
and p
2
(s) = c
2
e
s
. Therefore,
p
1
/p
2
= a. Using this equation with p
1
p
2
= z, we get p
1
=

az and
CHAPTER 2. FIRST ORDER PDE 36
p
2
=
_
z/a. Now,
dz(s)
ds
= (p
1
(s), p
2
(s))
_
dx(s)
ds
,
dy(s)
ds
_
=

az
dx(s)
ds

_
z/a
dy(s)
ds
1

z
dz(s)
ds
=
_

a
dx(s)
ds
+ 1/

a
dy(s)
ds
_
2

z =
_
ax + y/

a
_
+c
3
.
Thus,
u(x, y) =
_
b +
1
2
_
ax + y/

a
_
_
2
is a complete integral, if we had chosen a > 0.
Note that previous two examples compute two dierent complete integral
for same equation. However, in both examples, no choice of a and b will give
the zero solution u 0. Thus, u 0 is called singular solution.
Example 2.24. Let us nd the complete integral, general solution and singular
solution of the fully non-linear PDE u
2
x
+ u
2
y
= 1 + 2u. Since F(p, z, x) =
p
2
1
+ p
2
2
1 2z, the ODEs are
_
dx(s)
ds
,
dy(s)
ds
_
= (2p
1
(s), 2p
2
(s)),
dp(s)
ds
= (2p
1
, 2p
2
)
Thus, on dividing and integrating, we get p
1
/p
2
= a. Using the PDE, we get
(1 + a
2
)p
2
2
= 1 + 2z. Thus,
p
2
=
_
1 + 2z
1 + a
2
p
1
= a
_
1 + 2z
1 +a
2
.
CHAPTER 2. FIRST ORDER PDE 37
Now,
dz(s)
ds
= (p
1
(s), p
2
(s))
_
dx(s)
ds
,
dy(s)
ds
_
=
_
1 + 2z
1 + a
2
_
a
dx(s)
ds
+
dy(s)
ds
_
1

1 + 2z
dz(s)
ds
=
1

1 + a
2
_
a
dx(s)
ds
+
dy(s)
ds
_

1 + 2z =
ax +y

1 + a
2
b.
Thus,
u(x, y) =
1
2
_
ax + y

1 + a
2
+b
_
2

1
2
is a complete integral. Note that no choice of a and b will give the constant
solution u = 1/2. Thus, u 1/2 is called singular solution.
Exercise 9. Find the general solution (in terms of arbitrary functions) of the
given rst order PDE
(i) xu
x
+ yu
y
= xe
u
with x > 0. (Answer: u(x, y) = f(y/x) for some
arbitrary f).
(ii) u
x
+ u
y
= y + u. (Answer: u(x, y) = (1 + y) + f(y x)e
x
).
(iii) x
2
u
x
+ y
2
u
y
= (x + y)u. (Answer: u(x, y) = f((1/x) (1/y))(x y)).
(iv) x(y
2
u
2
)u
x
y(u
2
+x
2
)u
y
= (x
2
+y
2
)u. (Answer: u(x, y) =
x
y
f(x
2
+
y
2
+ u
2
)).
(v) (ln(y + u))u
x
+u
y
= 1.
(vi) x(y u)u
x
+ y(u x)u
y
= (x y)u.
(vii) u(u
2
+ xy)(xu
x
yu
y
) = x
4
.
(viii) (y + xu)u
x
(x +yu)u
y
= x
2
y
2
.
(ix) (y
2
+u
2
)u
x
xyu
y
+xu = 0.
(x) (y u)u
x
+ (u x)u
y
= x y.
CHAPTER 2. FIRST ORDER PDE 38
(xi) x(y
2
+ u)u
x
y(x
2
+ u)u
y
= (x
2
y
2
)u.
(xii)

1 x
2
u
x
+ u
y
= 0.
(xiii) (x + y)uu
x
+ (x y)uu
y
= x
2
+ y
2
.
(xiv) Find a complete integral of uu
x
u
y
= u
x
+ u
y
.
(xv) Find a complete integral of u
2
x
+ u
2
y
= xu.
2.6 Cauchy Problem
Recall that the general solution of the transport equation depends on the
value of u at time t = 0, i.e., the value of u on the curve (x, 0) in the xt-
plane. Thus, the problem of nding a function u satisfying the rst order
PDE (2.4.4) such that u is known on a curve in the xy-plane is called the
Cauchy problem.
Denition 2.6.1. A Cauchy problem states that: given a hypersurface
R
n
, can we nd a solution u of F(x, u(x), u(x)) = 0 whose graph contains
?
The question that arises at this moment is that: Does the knowledge of u
on any hypersurface R
n
lead to solving the rst order PDE? The answer
is a no. For instance, in the transport problem, if we choose the curve
= (x, t) [ x bt = 0, then we had no information to conclude u o the
line x bt = 0.
2.6.1 Quasilinear
Consider the general rst order qausilinear PDE with n independent variable
F(x, u, Du, D
2
u) := b(x, u(x)) Du(x) c(u(x), x), (2.6.1)
where b = b
i
is a vector with n components. Let be an hypersurface
dened implicitly as S(x) = 0 and S ,= 0. Given u on , can we calculate
all rst derivatives of u on ? To answer this, we map to a hyperplane

0
by the map :
0
with its components = (
1
,
2
, . . . ,
n
) such
that
n
(x) = S(x) and J

(x) ,= 0, non-zero Jacobian of , for all x . For


suciently regular hypersurfaces such a mapping exists. Note that
0
is a
CHAPTER 2. FIRST ORDER PDE 39
subset of the hyperplane whose nal coordinate is zero. Let y = (x) and
v(y) := u(x). Then
u
x
i
=
n

k=1
v
y
k

k
x
i
.
Using the above equation, the rst order linear PDE becomes
n

i,k=1
b
i

k
x
i
v
y
k
= c(v, y), (2.6.2)
where RHS is all known on
0
. To understand LHS on
0
, note that
v
y
k
(y) = lim
h0
1
h
[v(y
1
, . . . , y
k
+ h, . . . , y
n1
, 0) v(y
1
, . . . , y
k
, . . . , y
n1
, 0)].
Therefore, we know v
y
k
, for all k = 1, 2, . . . , n 1, on
0
. We only do not
know v
yn
on
0
. Thus, (2.6.2) can be rewritten as
n

i=1
b
i

n
x
i
v
yn
= terms known on
0
.
Since
n
(x) = S(x), we can compute v
yn
if
n

i=1
b
i
(x)S
x
i
,= 0
on . Note that S is the normal to the hypersurface S(x) = 0.
Denition 2.6.2. We say a hypersurface R
n
is non-characteristic
w.r.t (2.6.1) if
n

i=1
b
i
(u(x), x)
i
(x) ,= 0 x ,
where (x) is the normal vector of at x.
For instance, in the two dimension case, =
1
(r),
2
(r) R
2
is
non-characteristic for the quasilinear Cauchy problem
_
a(x, y, u)u
x
+ b(x, y, u)u
y
= c(x, y, u) (x, y)
u = g on
(2.6.3)
if is nowhere tangent to (a(
1
,
2
, g), b(
1
,
2
, g)), i.e.,
(a(
1
,
2
, g), b(
1
,
2
, g)) (

2
,

1
) ,= 0.
CHAPTER 2. FIRST ORDER PDE 40
Example 2.25. Consider the equation
2u
x
(x, y) + 3u
y
(x, y) = 1 in R
2
.
Let be a straight line y = mx +c in R
2
. The parametrisation of the line is
(r) := (r, mr + c) for r R. Therefore,
(a(
1
(r),
2
(r)), b(
1
(r),
2
(r))) (

2
(r),

1
(r)) = (2, 3) (m, 1) = 3 2m.
Thus, the line is not a non-characteristic for m = 3/2, i.e., all lines with slope
3/2 is not a non-characteristic.
Theorem 2.6.3. Let a, b and c, the coecients of (2.6.3), have continuous
partial derivatives w.r.t x, y, u. Let (r) := (
1
(r),
2
(r)) be the parametriza-
tion of an initial curve on which u(
1
(r),
2
(r)) = (r) such that
1
,
2
and
are continuously dierentiable and the initial curve is non-characteristic,
i.e.,
(a(
1
,
2
, ), b(
1
,
2
, )) (

2
,

1
) ,= 0.
Then there exists a unique solution u(x, y) in some neighbourhood of which
satises (2.6.3).
Proof. The characteristic curves are solution to the ODEs
dx
ds
(r, s) = a(x, y, u);
dy
ds
(r, s) = b(x, y, u) and
du
ds
(r, s) = c(x, y, u)
such that on , x(r, s) =
1
(r), y(r, s) =
2
(r) and z(r, s) = (r). The
non-characteristic property of implies that the following Jacobian on :
_
x
r
x
s
y
r
y
s
_
= (a(
1
,
2
, ), b(
1
,
2
, )) (

2
,

1
) ,= 0.
By implicit function theorem (cf. Theorem C.0.2), one can solve for r and
s in terms of x and y, locally, in the neighbourhood of . Set v(x, y) =
u(r(x, y), s(x, y)). We will show that v is a unique solution to (2.6.3). Note
that, by implicit function theorem, the value of v on is
v(x, y) = u(
1
(r),
2
(r)) = (r).
CHAPTER 2. FIRST ORDER PDE 41
Moreover,
av
x
+ bv
y
= a(u
r
r
x
+u
s
s
x
) + b(u
r
r
y
+ u
s
s
y
)
= u
r
(ar
x
+br
y
) +u
s
(as
x
+bs
y
)
= u
r
(r
x
x
s
+ r
y
y
s
) + u
s
(s
x
x
s
+ s
y
y
s
)
= u
r
dr
ds
+u
s
ds
ds
=
du
ds
= c
Theorem 2.6.4. If the initial is non-characteristic and there exists func-
tions and , as above, then there exists a unique (depending on and )
solution u(x, y) in some neighbourhood of which satises (2.6.5).
Recall that we already introduced the notion of well-posedness of a PDE
in Chapter 1. We see the existence issues in a example which highlights the
importance of well-posedness of Cauchy problem. In particular, if is not
non-characteristic, then the Cauchy problem is not well-posed.
Example 2.26. (i) Find the general solution (in terms of arbitrary func-
tions) of the rst order PDE 2u
x
(x, y) + 3u
y
(x, y) + 8u(x, y) = 0.
(ii) For the PDE given above, check for the characteristic property of the
following curves
(a) y = x in the xy-plane
(b) y =
3x1
2
.
(iii) Discuss the particular solutions of the above PDE, corresponding to
(a) u(x, x) = x
4
on y = x
(b) u(x, (3x 1)/2) = x
2
on y = (3x 1)/2
(c) u(x, (3x 1)/2) = e
4x
.
Observe the nature of solutions for the same PDE on a characteristic
curve and on non-characteristic curve.
(i) The characteristic equations are
dx
ds
= 2,
dy
ds
= 3 and
dz
ds
= 8z.
CHAPTER 2. FIRST ORDER PDE 42
Hence,
x(s) = 2s + c
1
y(s) = 3s + c
2
and z(s) = c
3
e
8s
.
Thus, 3x 2y = c
4
and z = c
4
e
4x
or z = c
5
e
8y/3
. Hence, the general
solution is F(3x 2y, e
4x
z) = 0. Explicitly, for some f and g,
u(x, y) = f(3x 2y)e
4x
or u(x, y) = g(3x 2y)e
8y/3
.
(ii) (a) Parametrise the curve y = x as (r) : r (r, r). Thus
1
(r) =

2
(r) = r. Since the coecients of the PDE are a(r) = 2 and
b(r) = 3, we have
(a, b) (

2
(r),

1
(r)) = (2, 3) (1, 1) = 2 + 3 = 1 ,= 0.
Hence is non-characteristic.
(b) Parametrise the curve y = (3x 1)/2 as (r) : r (r, (3r 1)/2).
Hence
1
(r) = r and
2
(r) = (3r 1)/2 and
(a, b) (

2
(r),

1
(r)) = (2, 3) (3/2, 1) = 3 + 3 = 0.
Hence is a characteristic curve.
(iii) Recall that the general solution is F(3x 2y, e
4x
z) = 0 or
u(x, y) = f(3x 2y)e
4x
or u(x, y) = g(3x 2y)e
8y/3
.
(a) Now, u(x, x) = x
4
implies F(x, e
4x
x
4
) = 0. Thus,
e
4x
z = e
12x8y
(3x 2y)
4
and
u(x, y) = (3x 2y)
4
e
8(xy)
.
Thus, we have a unique solution u.
(b) Using the given condition, we have F(1, x
2
e
4x
) = 0. Either f(1) =
x
2
e
4x
or f(x
2
e
4x
) = 1. The rst case is not valid (multi-valued
function). The second case corresponds to z = e
4x
which will not
satisfy the Cauchy data. Hence there is no solution u solving the
given PDE with the given data.
CHAPTER 2. FIRST ORDER PDE 43
(c) Once again using the given condition, we have F(1, x
2
e
4x
) = 0.
Either f(1) = x
2
e
4x
or f(x
2
e
4x
) = 1. The rst case is not valid
(multi-valued function). The second case corresponds to z = e
4x
which will satisfy the Cauchy data. Since there many choices of f
that satises f(x
2
e
4x
) = 1, we have innite number of solutions (or
choices for) u that solves the PDE.
Example 2.27. (i) Find the general solution (in terms of arbitrary func-
tions) of the rst order PDE u
x
(x, y) + u
y
(x, y) = 1.
(ii) For the PDE given above, check for the characteristic property of the
following curves
(a) the x-axis, (x, 0), in the xy-plane
(b) y = x.
(iii) Discuss the particular solutions of the above PDE, corresponding to
(a) u(x, 0) = (x) on x-axis.
(b) u(x, x) = x on y = x.
(c) u(x, x) = 1 on y = x.
Observe the nature of solutions for the same PDE on a characteristic
curve and on non-characteristic curve.
(i) The characteristic equations are
dx
ds
= 1,
dy
ds
= 1 and
dz
ds
= 1.
Hence,
x(s) = s + c
1
y(s) = s + c
2
and z(s) = s + c
3
.
Thus, y x = c
4
and z x = c
5
or x y = c
4
and z y = c
5
. Hence,
for some f and g,
u(x, y) = x + f(y x) or u(x, y) = y + g(x y).
CHAPTER 2. FIRST ORDER PDE 44
(ii) (a) Parametrise the curve x-axis as (r) : r (r, 0). Thus
1
(r) = r
and
2
(r) = 0. Since the coecients of the PDE are a(r) = 1 and
b(r) = 1, we have
(a, b) (

2
(r),

1
(r)) = (1, 1) (0, 1) = 1 ,= 0.
Hence is non-characteristic.
(b) Parametrise the curve y = x as (r) : r (r, r). Hence
1
(r) =
r =
2
(r) and
(a, b) (

2
(r),

1
(r)) = (1, 1) (1, 1) = 1 + 1 = 0.
Hence is a characteristic curve.
(iii) Recall that the general solution is
u(x, y) = x + f(y x) or u(x, y) = y + g(x y).
(a) Now, u(x, 0) = (x) implies f(x) = (x) +x or g(x) = (x), and
u(x, y) = y +(x y).
Thus, we have a unique solution u.
(b) Using the given condition, we have f(0) = 0 or g(0) = 0. One has
many choices of function satisying these conditions. Thus, we have
innite number of solutions (or choices for) u that solves the PDE.
(c) Once again using the given condition, we have f(0) = 1 x or
g(0) = 1 x for all x R. This implies f and g are not well
dened. We have no function f and g, hence there is no solution u
solving the given PDE with the given data.
For any given (smooth enough) function : R R, consider the linear
transport equation
_
u
t
+ bu
x
= 0 x R and t (0, )
u(x, 0) = (x) x R.
(2.6.4)
We know that the general solution of the transport equation is u(x, t) = u(x
bt, 0). Thus, u(x, t) = (x bt) is the unique solution of (2.6.4). We derive
the particular solution of the Cauchy problem (2.6.4) using parametrisation
CHAPTER 2. FIRST ORDER PDE 45
of the data curve . The example also shows how the information on data
curve is reduced as initial condition for the characteristic ODEs. Note that
in the example below the data curve is parametrised using the variable r
and the characteristic curves is parametrised using the variable s.
Example 2.28. We shall compute the solution of the Cauchy problem (2.6.4).
We rst check for non-characteristic property of . Note that (x, 0),
the x-axis of xt-plane, is the (boundary) curve on which the value of u is given.
Thus, (, ) = (x, 0, (x)) is the known curve on the solution surface of u.
We parametrize the curve with r-variable, i.e., =
1
(r),
2
(r) =
(r, 0). is non-characteristic, because (b, 1) (0, 1) = 1 ,= 0. The charac-
teristic equations are:
dx(r, s)
ds
= b,
dt(r, s)
ds
= 1, and
dz(r, s)
ds
= 0
with initial conditions,
x(r, 0) = r, t(r, 0) = 0, and z(r, 0) = (r).
Solving the ODEs, we get
x(r, s) = bs + c
1
(r), t(r, s) = s +c
2
(r)
and z(r, s) = c
3
(r) with initial conditions
x(r, 0) = c
1
(r) = r
t(r, 0) = c
2
(r) = 0, and z(r, 0) = c
3
(r) = (r).
Therefore,
x(r, s) = bs + r, t(r, s) = s, and z(r, s) = (r).
We solve for r, s in terms of x, t and set u(x, t) = z(r(x, t), s(x, t)).
r(x, t) = x bt and s(x, t) = t.
Therefore, u(x, t) = z(r, s) = (r) = (x bt).
CHAPTER 2. FIRST ORDER PDE 46
Example 2.29. Let := (x, y) R
2
[ x > 0, y > 0. Let := (x, 0) [ x >
0. Consider the linear PDE
_
xu
y
(x, y) yu
x
(x, y) = u(x, y) in
u(x, 0) = (x) on .
The parametrisation of the initial curve is (r) := (r, 0) for r > 0. Therefore,
(a(
1
(r),
2
(r)), b(
1
(r),
2
(r))) (

2
(r),

1
(r)) = (0, r) (0, 1) = r ,= 0.
Hence, the given initial curve is non-characteristic. The characteristic equa-
tions are
dx(r, s)
ds
= y;
dy(r, s)
ds
= x and
dz(r, s)
ds
= z(s)
with initial conditions
x(r, 0) = r, y(r, 0) = 0, and z(r, 0) = (r).
Note that
d
2
x(r, s)
ds
= x(r, s) and
d
2
y(r, s)
ds
= y(r, s).
Then, x(r, s) = c
1
(r) cos s + c
2
(r) sin s and y(r, s) = c
3
(r) cos s + c
4
(r) sin s.
Using the initial condition, we get c
1
(r) = r and c
3
(r) = 0. Also,
0 = y(r, 0) =
dx(r, s)
ds
[
s=0
= c
1
(r) sin 0 + c
2
(r) cos 0 = c
2
(r).
and, similarly, c
4
(r) = r. Also, z(r, s) = c
5
(r)e
s
where c
5
(r) = (r). Thus,
we have (x(r, s), y(r, s)) = (r cos s, r sin s), where r > 0 and 0 s /2.
Hence, r = (x
2
+y
2
)
1/2
and s = arctan(y/x). Therefore, for any given (x, y),
we have
u(x, y) = z(r, s) = (r)e
s
= (
_
x
2
+ y
2
)e
arctan(y/x)
.
Example 2.30. Let := (x, y) R
2
[ y > 0. Let := (x, 0) [ x R.
Consider the semi-linear PDE
_
u
x
(x, y) + u
y
(x, y) = u
2
(x, y) in
u(x, 0) = (x) on .
The parametrisation of the initial curve is (r) := (r, 0) for all r R. There-
fore,
(a(
1
(r),
2
(r)), b(
1
(r),
2
(r))) (

2
(r),

1
(r)) = (1, 1) (0, 1) = 1 ,= 0.
CHAPTER 2. FIRST ORDER PDE 47
Hence, the given initial curve is non-characteristic. The characteristic equa-
tions are
dx(r, s)
ds
= 1,
dy(r, s)
ds
= 1 and
dz(r, s)
ds
= z
2
(r, s)
with initial conditions
x(r, 0) = r, y(r, 0) = 0, and z(r, 0) = (r).
Therefore, x(r, s) = s + c
1
(r), y(r, s) = s + c
2
(r) and z(r, s) =
1
s+c
3
(r)
. Using
the initial conditions, we get c
1
(r) = r, c
2
(r) = 0 and c
3
(r) = ((r))
1
.
Note that this makes sense only if (r) ,= 0 for all r. To overcome this
situation, we write z(r, s) =
(r)
1(r)s
. Also, we have (x(r, s), y(r, s)) = (s+r, s),
where s 0 and r R. Moreover, r = x y and s = y. Therefore,
u(x, y) = z(r, s) =
(r)
1(r)s
=
(xy)
1(xy)y
. Note that the non-linearity in the
z-variable, even though the equation is linear, may cause a possible blow-up
(or singularity) in the solution. For instance, even if we assume is bounded,
a very large value of y may induce a singularity.
Example 2.31. Consider the quasi-linear PDE
_
u
t
(x, t) + u(x, t)u
x
(x, t) = x in
u(x, 0) = 1 on .
The parametrisation of the initial curve is (r) := (r, 0) for all r R. There-
fore,
(a(
1
(r),
2
(r)), b(
1
(r),
2
(r))) (

2
(r),

1
(r)) = (u, 1) (0, 1) = 1 ,= 0.
Hence, the given initial curve is non-characteristic. The characteristic equa-
tions are
dx(r, s)
ds
= z(r, s),
dt(r, s)
ds
= 1 and
dz(r, s)
ds
= x(r, s)
with initial conditions
x(r, 0) = r, t(r, 0) = 0, and z(r, 0) = 1.
Therefore, t(r, s) = s + c
1
(r), x(r, s) = c
2
(r)e
s
+ c
3
(r)e
s
and z(r, s) =
c
4
(r)e
s
+ c
5
(r)e
s
. Using the initial conditions, we get c
1
(r) = 0, c
2
(r) =
(r +1)/2 = c
4
(r), c
3
(r) = (r 1)/2 and c
5
(r) = (1 r)/2. Solving for r and
s, in terms of x, t and z, we get s = t and
r =
2x e
t
+e
t
e
t
+ e
t
.
Therefore, u(x, y) = x
e
2t
1
e
2t
+1
+
2e
t
e
2t+1
.
CHAPTER 2. FIRST ORDER PDE 48
2.6.2 Nonlinear
Denition 2.6.5. We say R
n
is non-characteristic for the nonlin-
ear Cauchy problem
_
F(u(x), u(x), x) = 0 x
u = g on
(2.6.5)
if is nowhere tangent to the Monge cone, i.e., there exists function v such
that F(v(r), g(r),
1
(r), . . . ,
n
(r)) = 0 and v
i
= g
x
i
, for i = 1, 2, . . . , n 1
on , and satises
n

i=1
F
p
i
(v, g,
1
, . . . ,
n
) (x) ,= 0
where (x) is normal to at x.
In particular, in the two dimension case, =
1
(r),
2
(r) R
2
is
non-characteristic for the nonlinear Cauchy problem (2.6.5) if is nowhere
tangent to the Monge cone, i.e., there exists function (r) and (r) such that
F((r), (r), g(r),
1
(r),
2
(r)) = 0 and g

(r) = (r)

1
(r) + (r)

2
(r) and
satises
(F
p
2
(
1
,
2
, g, , ), F
p
1
(
1
,
2
, g, , )) (

2
,

1
) ,= 0.
Example 2.32. For a nonlinear PDE the concept of characteristics also depend
on initial values. Consider the nonlinear PDE

n
i=1
u
2
x
i
= 1 with u(x

, 0) =
(x

). Note that u = x
i
is a solution with (x

) = x
i
. However, any choice
of such that [
x
(x

)[
2
> 1 has no solution.
Suppose v is a solution of F(x, u, u) = 0 in a neighbourhood of some
point x
0
.
Example 2.33. (i) Find the general solution (in terms of arbitrary func-
tions) of the rst order PDE xu
x
(x, y) + 2xuu
y
(x, y) = u(x, y).
(ii) For the PDE given above, check if the following curves in xy-plane are
non-characteristic and discuss the particular solutions of the PDE
(a) y = 2x
2
+ 1 and u(x, 2x
2
+ 1) = 2x.
(b) y = 3x
3
and u(x, 3x
3
) = 2x
2
.
CHAPTER 2. FIRST ORDER PDE 49
(c) y = x
3
1 and u(x, x
3
1) = x
2
.
Observe the nature of solutions for the same PDE on a characteristic
curve and on non-characteristic curve.
(i) The characteristic equations are
dx
ds
= x,
dy
ds
= 2xz and
dz
ds
= z.
Hence,
x(s) = c
1
e
s
z(s) = c
2
e
s
and y(s) = c
1
c
2
e
2s
+ c
3
.
Thus, y = c
2
/c
1
x
2
+c
3
and z = c
2
/c
1
x. Therefore, y zx = c
3
and, for
some f,
u(x, y) = xf(y xu).
The characteristic curves are y xu = a constant which depends on u.
(ii) (a) Parametrise the curve y = 2x
2
+ 1 as (r) : r (r, 2r
2
+ 1). Thus

1
(r) = r and
2
(r) = 2r
2
+ 1. Since the coecients of the PDE
are a(r) = r and b(r, u) = 4r
2
, we have
(a, b) (

2
(r),

1
(r)) = (r, 4r
2
) (4r, 1) = 4r
2
+ 4r
2
= 0.
Hence is not non-characteristic. But on the characteristic curves
y 2x
2
= 1 the function u = 2x solves the PDE. Elsewhere
the solution is non-unique and there are many choices because
u(x, 2x
2
+ 1) = 2x implies f(1) = 2. Thus, we have innite num-
ber of solutions (or choices for) u that solves the PDE on other
characterisitc curves.
(b) Parametrise the curve y = 3x
3
as (r) : r (r, 3r
3
). Hence

1
(r) = r and
2
(r) = 3r
3
and
(a, b) (

2
(r),

1
(r)) = (r, 4r
3
) (9r
2
, 1) = 9r
3
+4r
3
= 5r
3
,= 0
for r ,= 0. Hence is a non-characteristic curve. Using (u(x, 3x
3
) =
2x
2
we get 2x
2
= xf(3x
3
2x
3
) which implies 2x = f(x
3
). Thus,
f(x) = 2x
1/3
and u(x, y) = 2x(y xu)
1/3
or u
3
(x, y) +8x
4
u = 8x
3
y.
Thus, we have a unique solution. The characteristic curves are
y 2x
3
= a constant and the data u = 2x
2
is given on y 3x
3
= 0.
CHAPTER 2. FIRST ORDER PDE 50
(c) Parametrise the curve y = x
3
1 as (r) : r (r, r
3
1). Hence

1
(r) = r and
2
(r) = r
3
1 and
(a, b) (

2
(r),

1
(r)) = (r, 2r
3
) (3r
2
, 1) = 3r
3
+2r
3
= r
3
,= 0
for r ,= 0. Hence is a non-characteristic curve. Using (u(x, x
3

1) = x
2
we get x = f(1). Thus f is not well dened and, hence
there is no solution. The characteristic curves are y x
3
= a con-
stant and u = x
2
given on y x
3
= 1 is not a solution.
Example 2.34. We now give an example of a fully non linear PDE. Let :=
(x, y) R
2
[ x > 0. Let := (0, y) [ y R. Consider the fully
non-linear PDE
_
u
x
u
y
= u(x, y) in
u(0, y) = y
2
on .
The parametrisation of the initial curve is (r) := (0, r) for all r R. We
rst look for the functions and such that
(r)(r) = r
2
and 2r = (r).
Solving them, we get (r) = 2r and (r) = r/2. Since F(p, z, x) = p
1
p
2
z,
we have
F
p
2

2
(r) = p
1
= (r) = r/2 ,= 0, for r ,= 0.
Hence, the given initial curve is non-characteristic. The characteristic equa-
tions are (using (2.4.1))
_
dx(r, s)
ds
,
dy(r, s)
ds
_
= (p
2
(r, s), p
1
(r, s)),
(using (2.4.2))
dp(r, s)
ds
= p(r, s)
and (using (2.4.3))
dz(r, s)
ds
= (p
1
(r, s), p
2
(r, s)) (p
2
(r, s), p
1
(r, s)) = 2p
1
(r, s)p
2
(r, s)
with initial conditions
x(r, 0) = 0, y(r, 0) = r, z(r, 0) = r
2
, p
1
(r, 0) = (r) =
r
2
CHAPTER 2. FIRST ORDER PDE 51
and
p
2
(r, 0) = (r) = 2r.
Thus, on integrating, we get p
1
(r, s) = (r/2)e
s
and p
2
(r, s) = 2re
s
, for all
s R. Using p, we solve for x to get x(r, s) = 2r(e
s
1) and y(r, s) =
(r/2)(e
s
+ 1). Solving for z, we get
z(r, s) =
r
2
2
(e
2s
+ 1).
Solving r and s in terms of x and y, we get
r =
4y x
4
and e
s
=
x + 4y
4y x
.
Hence u(x, y) = z(r(x, y), s(x, y)) =
(x+4y)
2
16
.
Example 2.35. Consider the fully non-linear PDE
_
u
x
u
y
= u(x, y) in R
2
u(x, 1 + x) = x
2
on .
The parametrisation of the initial curve is (r) := (r, 1 + r) for all r R.
We rst look for the functions and such that
(r)(r) = r
2
and 2r = (r) + (r).
Solving them, we get (r) = r = (r). Since F(p, z, x) = p
1
p
2
z, we have
F
p
1

1
F
p
2

2
(r) = p
2
p
1
= (r) (r) = r r = 0, r.
Hence, the given initial curve is not non-characteristic. The characteristic
equations are (using (2.4.1))
_
dx(r, s)
ds
,
dy(r, s)
ds
_
= (p
2
(r, s), p
1
(r, s)),
(using (2.4.2))
dp(r, s)
ds
= p(r, s)
and (using (2.4.3))
dz(r, s)
ds
= (p
1
(r, s), p
2
(r, s)) (p
2
(r, s), p
1
(r, s)) = 2p
1
(r, s)p
2
(r, s)
CHAPTER 2. FIRST ORDER PDE 52
with initial conditions
x(r, 0) = r, y(r, 0) = 1 + r, z(r, 0) = r
2
, p
1
(r, 0) = (r) = r
and
p
2
(r, 0) = (r) = r.
Thus, on integrating, we get p
1
(r, s) = re
s
and p
2
(r, s) = re
s
, for all s R.
Using p, we solve for x to get x(r, s) = re
s
and y(r, s) = re
s
+ 1. Solving for
z, we get z(r, s) = r
2
e
2s
. Note that there is no unique way of solving r and s
in terms of x and y. In fact, we have three possible represenation of u, viz.,
u = x
2
, u = (y 1)
2
and u = x(y 1). This is because the Jacobian is zero.
Of these three possibilities, only u = x(y 1) satises the equation.
Example 2.36. For any given R, consider the fully non-linear PDE
_
u
2
x
+ u
2
y
= 1 in R
2
u(x, x) = x on .
The parametrisation of the initial curve is (r) := (r, r) for all r R. We
rst look for the functions and such that

2
(r) +
2
(r) = 1 and = (r) + (r).
We can view (r) = cos and (r) = sin where is such that cos +sin =
. Since F(p, z, x) = p
2
1
+ p
2
2
1, we have
F
p
1

1
F
p
2

2
(r) = 2(p
1
p
2
) = 2(cos sin ) = 0, = /4 +k,
where k = 1, 2, . . .. Hence, the given initial curve is non-characteristic for
,= /2 +k for all k and cos +sin = . The characteristic equations are
(using (2.4.1))
_
dx(r, s)
ds
,
dy(r, s)
ds
_
= (2p
1
(r, s), 2p
2
(r, s)),
(using (2.4.2))
dp(r, s)
ds
= 0
and (using (2.4.3))
dz(r, s)
ds
= (p
1
(r, s), p
2
(r, s)) (2p
1
(r, s), 2p
2
(r, s)) = 2(p
2
1
+ p
2
2
) = 2
CHAPTER 2. FIRST ORDER PDE 53
with initial conditions
x(r, 0) = r, y(r, 0) = r, z(r, 0) = r, p
1
(r, 0) = (r) = cos
and
p
2
(r, 0) = (r) = sin .
Thus, on integrating, we get p
1
(r, s) = cos and p
2
(r, s) = sin . Using p, we
solve for x to get x(r, s) = 2s cos + r and y(r, s) = 2s sin + r. Solving for
z, we get z(r, s) = 2s + r. Solving r and s in terms of x and y, we get
s =
x y
2(cos sin )
and r =
xsin y cos
sin cos
.
Therefore, the general solution is
u(x, y) =
x y + (y cos xsin )
cos sin
=
x y + (cos + sin )(y cos xsin )
cos sin
=
x(1 sin cos sin
2
) + y(cos
2
+ sin cos 1)
cos sin
= xcos +y sin .
Example 2.37. Consider the equation xu
x
u
y
+ yu
2
y
= 1 with u(2r, 0) = r.
Note that F(p, z, x, y) = xp
1
p
2
+ yp
2
2
1 = 0. Thus the ODE
_
dx(s)
ds
,
dy(s)
ds
_
= (xp
2
, xp
1
+ 2yp
2
),
dp(s)
ds
= (p
1
p
2
, p
2
2
)
Thus, on integrating, we get p
1
/p
2
= a. Using the PDE, we get (xa+y)p
2
2
= 1.
Thus,
p
2
=
1

xa + y
p
1
=
a

xa + y
.
Now,
dz(s)
ds
= (p
1
(s), p
2
(s))
_
dx(s)
ds
,
dy(s)
ds
_
=
1

xa + y
_
a
dx(s)
ds
+
dy(s)
ds
_
u(x, y) = 2

ax + y +b.
CHAPTER 2. FIRST ORDER PDE 54
Thus, (u(x, y) b)
2
= 4(ax + y) is a complete integral. Using the initial
data, we get (r b)
2
= 8ar. Dierentiating w.r.t r, we get r = 4a + b and
eliminating r, we get b = h(a) := 2a. Hence, (u + 2a)
2
= 4(ax + y). Now
solving for a in D
a
u = 0, we get 4(u+2a) = 4x. Then a = (x) := (xu)/2
and the solution is
x
2
= 4
_
x u
2
_
x + 4y
which yields u(x, y) =
x
2
+4y
2x
.
Example 2.38. Consider the equation y(u
2
x
u
2
y
)+uu
y
= 0 with u(2y, y) = 3y.
Note that F(p, z, x, y) = y(p
2
1
p
2
2
) +zp
2
= 0. Thus the ODEs become
_
dx(s)
ds
,
dy(s)
ds
_
= (2yp
1
, z 2yp
2
),
dp(s)
ds
= (p
1
p
2
, p
2
1
)
Thus, on integrating, we get p
2
1
p
2
2
= a. Using the PDE, we get ya+up
2
= 0.
Thus,
p
2
=
ay
z
p
1
=
_
a +
a
2
y
2
z
2
.
Now,
dz(s)
ds
= (p
1
(s), p
2
(s))
_
dx(s)
ds
,
dy(s)
ds
_

a
dx(s)
ds
=
1
_
z
2
+ ay
2
_
z
dz(s)
ds
+ ay
dy(s)
ds
_

ax b =
_
z
2
+ ay
2
u
2
(x, y) = (b + x

a)
2
ay
2
.
Using the initial data, we get 9y
2
= (b +2y

a)
2
ay
2
which is satised with
b = 0 and a = 3. Thus, u(x, y)
2
= 3(x
2
y
2
).
Example 2.39. Consider the equation y(u
2
x
u
2
y
) + uu
y
= 0 with u(r
2
, 0) =
2r. Note that F(p, z, x, y) = y(p
2
1
p
2
2
) + zp
2
= 0. As in above example,
we get u
2
(x, y) = (b + x

a)
2
ay
2
. Using the initial conditions, we get
4r
2
= (b + r
2

a)
2
. Dierentiating this w.r.t to r and solving for r, we get
r
2
=
1

a
_
2

a
b
_
.
CHAPTER 2. FIRST ORDER PDE 55
Substituting this in the equation of r, we get b = 1/

a. This gives
u
2
=
_
1

a
+ x

a
_
2
ay
2
=
1
a
(1 + xa)
2
ay
2
.
Now solving for a in D
a
u = 0, we get
0 = 2
_
1

a
+x

a
__

1
2a

a
+
x
2

a
_
y
2
y
2
=
_
x
2

1
a
2
_
a = (x) :=
1
_
x
2
y
2
.
We choose the positive root above to keep a > 0 so that all roots above made
sense. Therefore,
u
2
(x, y) =
_
x
2
y
2
_
1 +
x
_
x
2
y
2
_
2

y
2
_
x
2
y
2
.
Exercise 10. Find the general solution of the following PDE. Check if the
given data curve is non-characteristic or not. Also nd the solution(s) (if it
exists) given the value of u on the prescribed curves.
(i) 2u
t
+ 3u
x
= 0 with u(x, 0) = sin x.
(ii) u
x
u
y
= 1 with u(x, 0) = x
2
.
(iii) u
x
+ u
y
= u with u(x, 0) = cos x.
(iv) u
x
u
y
= u with u(x, x) = sin x.
(v) 4u
x
+u
y
= u
2
with u(x, 0) =
1
1+x
2
.
(vi) au
x
+ u
y
= u
2
with u(x, 0) = cos x.
(vii) u
x
+ 4u
y
= x(u + 1) with u(x, 5x) = 1.
(viii) (1 xu)u
x
+ y(2x
2
+ u)u
y
= 2x(1 xu). Also, when u(0, y) = e
y
on
x = 0.
CHAPTER 2. FIRST ORDER PDE 56
(ix) e
2y
u
x
+ xu
y
= xu
2
. Also, when u(x, 0) = e
x
2
on y = 0.
(x) u
x
2xuu
y
= 0. Also, when u(x, 2x) = x
1
on y = 2x and when
u(x, x
3
) = x on y = x
3
.
(xi) 3u
x
+ u
y
= 0 with u(x, 0) = e
x
2
. (Answer: u(x, y) = e
(x+3y)
2
).
(xii) yu
x
+xu
y
= x
2
+y
2
with u(x, 0) = 1+x
2
and u(0, y) = 1+y
2
. (Answer:
u(x, y) = xy +[x
2
y
2
[).
(xiii) yu
x
+ xu
y
= 4xy
3
with u(x, 0) = x
4
and u(0, y) = 0.
(xiv) yu
x
+ xu
y
= u with u(x, 0) = x
3
.
(xv) u
x
+yu
y
= y
2
with u(0, y) = sin y.
(xvi) u
x
+yu
y
= u
2
with u(0, y) = sin y.
(xvii) u
x
+yu
y
= u with u(x, 3e
x
) = 2.
(xviii) u
x
+yu
y
= u with u(x, e
x
) = e
x
.
(xix) u
x
+xu
y
= u with u(1, y) = (y).
(xx) xu
x
+ u
y
= 3x u with u(x, 0) = arctan x.
(xxi) xu
x
+ u
y
= 0 with u(x, 0) = (x).
(xxii) xu
x
+ yu
y
= u with u(x, 1) = 2 + e
|x|
.
(xxiii) xu
x
+ yu
y
= xe
u
with u(x, x
2
) = 0.
(xxiv) xu
x
yu
y
= 0 with u(x, x) = x
4
.
(xxv) e
2y
u
x
+ xu
y
= xu
2
with u(x, 0) = e
x
2
.
(xxvi) uu
x
+ u
y
= 1 with u(2r
2
, 2r) = 0 for r > 0. (Answer: No solution for
y
2
> 4x).
(xxvii) (y u)u
x
+ (u x)u
y
= x y with u(x, 1/x) = 0.
(xxviii) x(y
2
+ u)u
x
y(x
2
+ u)u
y
= (x
2
y
2
)u with u(x, x) = 1.
(xxix)

1 x
2
u
x
+ u
y
= 0 with u(0, y) = y.
CHAPTER 2. FIRST ORDER PDE 57
Exercise 11. Solve the equation xu
x
+ 2yu
y
= 0 with u(1, y) = e
y
. Does a
solutions exist with data on u(0, y) = g(y) or u(x, 0) = h(x)? What happens
to characteristic curves at (0, 0)?
Exercise 12. Solve the equation yu
x
+xu
y
= 0 with u(0, y) = e
y
2
. In which
region of the plane is the solution uniquely determined?
Exercise 13. Solve the equation u
x
+ yu
y
= 0 with u(x, 0) = 1. Also, solve
the equation with u(x, 0) = x. If there is no solution, give reasons for non-
existence.
Exercise 14. (i) Find a complete integral of x(u
2
x
+u
2
y
) = uu
x
with u given
on the curve
(a) u(2y, y) = 5y
(b) u(0, r
2
) = 2r
(ii) Find a complete integral of 4uu
x
u
3
y
= 0 with u given on the curve
u(0, r) = 4r.
2.7 Non-smooth Solutions and Shocks
Observe that in the Cauchy problem of Transport equation (2.6.4), suppose
we choose a discontinuous , for instance,
(x) =
_
1 x > 0
0 x 0
then u(x, t) inherits this jump continuity. In applications it is often necessary
to consider such solutions which, by our denition, is not even a dierentiable,
hence, not a solution. This demand motivates us to redene the notion of
solution to accommodate non-smooth solutions. Further, in some quasilinear
situations, eventhough we start with a smooth initial data , jump disconti-
nuity might occur at some positive time t. For instance, consider the Burgers
equation given as
_
u
t
+ uu
x
= 0 x R and t (0, )
u(x, 0) = (x) x R.
CHAPTER 2. FIRST ORDER PDE 58
We parametrize the curve with r-variable, i.e., =
1
(r),
2
(r) =
(r, 0). is non-characteristic, because (u, 1) (0, 1) = 1 ,= 0. The charac-
teristic equations are:
dx(r, s)
ds
= z,
dt(r, s)
ds
= 1, and
dz(r, s)
ds
= 0
with initial conditions,
x(r, 0) = r, t(r, 0) = 0, and z(r, 0) = (r).
Solving the ODE corresponding to z, we get z(r, s) = c
3
(r) with initial
conditions z(r, 0) = c
3
(r) = (r). Thus, z(r, s) = (r). Using this in the
ODE of x, we get
dx(r, s)
ds
= (r).
Solving the ODEs, we get
x(r, s) = (r)s + c
1
(r), t(r, s) = s + c
2
(r)
with initial conditions
x(r, 0) = c
1
(r) = r and t(r, 0) = c
2
(r) = 0.
Therefore,
x(r, s) = (r)s + r, and t(r, s) = s.
Solving r and s, in terms of x, t and z, we get s = t and r = xzt. Therefore,
u(x, t) = (x tu) is the solution in the implicit form.
Example 2.40. If the data (x) = c, some constant, then u(x, t) = c and the
characteristic curves are t = x/c.
Example 2.41. In the Burgers equation, suppose we choose the data (x) =
x, then
u(x, t) =
x
1 + t
.
Note that u has a singularity at (0, 1). This can be observed in the following
way: u takes the constant value c along the line t = x/c 1 and all these
curves intersect at (0, 1) which means u is multiply dened at (0, 1) or,
rather undened at (0, 1).
CHAPTER 2. FIRST ORDER PDE 59
Example 2.42. In the Burgers equation, suppose we choose the data to be
the function
(x) =
_

_
1 x 0
1 x 0 x 1
0 x 1.
Then the characteristic curves are
x =
_

_
t +c c 0
(1 c)t 0 c 1
c c 1.
Therefore,
u(x, t) =
_

_
1 x t
1x
1t
t x 1
0 x 1.
Note that for t 1 the solution behaves well, but for t 1, the characteristics
start crossing each other on the line t = x and u(x, t) takes both 0 and 1 on
the line t = x, for t 1. This situation is called the shock.
Example 2.43. In the Burgers equation, suppose we choose to be the
function
(x) =
_

_
1 x < 1
x 1 x 1
1 1 < x.
Then the characteristic curves are
x =
_

_
t + c c < 1
c(t + 1) 1 c 1
t + c 1 < c.
Therefore,
u(x, t) =
_

_
1 x +t < 1
x
t+1
(t + 1) x (t + 1)
1 1 < x t.
CHAPTER 2. FIRST ORDER PDE 60
Example 2.44. In the Burgers equation, suppose we choose to be the
function
(x) =
_
0 x < 0
1 x > 0.
Then note that
u(x, t) =
_
0 x < 0
1 x > t,
but there is no information of u on the wedge 0 < x < t.
Chapter 3
Classication by
Characteristics
A general second order PDE is of the form F (D
2
u(x), Du(x), u(x), x) = 0,
for each x R
n
and u : R is the unknown. A Cauchy problem
poses the following: Given the knowledge of u on a smooth hypersurface
can one nd the solution u of the PDE? The knowledge of u on is
said to be the Cauchy data.
What should be the minimum required Cauchy data for the Cauchy prob-
lem to be solved? Viewing the Cauchy problem as an initial value problem
corresponding to ODE, we know that a unique solution exists to the second
order ODE
_
_
_
y

(x) +P(x)y

(x) +Q(x)y(x) = 0 x I
y(x
0
) = y
0
y

(x
0
) = y

0
.
where P and Q are continuous on I (assume I closed interval of R) and for
any point x
0
I. This motivates us to dene the Cauchy problem for second
order PDE as:
_
_
_
F (D
2
u(x), Du(x), u(x), x) = 0 x
u(x) = g(x) x
Du(x) (x) = h(x) x
(3.0.1)
where is the outward unit normal vector on the hypersurface and g, h
are known functions on .
61
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 62
3.1 Semilinear
Consider the general second order qausilinear PDE with n independent vari-
able
F(x, u, Du, D
2
u) := A(x) D
2
u D(u, u, x), (3.1.1)
where A = A
ij
is an n n matrix with entries A
ij
(x, u, u), D
2
u is the
Hessian matrix. The dot product in LHS is in R
n
2
. Since we demand the
solution to be in C
2
, the mixed derivatives are equal and we can assume,
without loss generality that, A is symmetric. In fact if A is not symmetric,
we can replace A with A
s
:=
1
2
(A
0
+A
t
0
), which is symmetric since A D
2
u =
A
s
D
2
u.
Let be an hypersurface dened implicitly as S(x) = 0 and S ,= 0.
Given u and u on , can we calculate all other derivatives of u on ?
To answer this, we map to a hyperplane
0
by the map :
0
with
its components = (
1
,
2
, . . . ,
n
) such that
n
(x) = S(x) and J

(x) ,= 0,
non-zero Jacobian of , for all x . For suciently regular hypersurfaces
such a mapping exists. Note that
0
is a subset of the hyperplane whose
nal coordinate is zero. Let y = (x) and v(y) := u(x). Then
u
x
i
=
n

k=1
v
y
k

k
x
i
u
x
i
x
j
=
n

k,l=1
v
y
k
y
l

k
x
i

l
x
j
+
n

k=1
v
y
k

k
x
i
x
j
.
Using the second equation the second order linear PDE becomes
n

i,j,k,l=1
A
ij

k
x
i

l
x
j
v
y
k
y
l
=

D(v, v, y)
n

i,k=1
A
ij
v
y
k

k
x
i
x
j
, (3.1.2)
where RHS is all known on
0
. To understand LHS on
0
, note that using
the rst equation, v
y
k
(y
1
, . . . , y
n1
, 0), for all k = 1, 2, . . . , n, are known in

0
. Therefore, we know v
y
k
y
l
, for all l = 1, 2, . . . , n 1, on
0
because
v
y
k
y
l
(y) = lim
h0
1
h
[v
y
k
(y
1
, . . . , y
l
+ h, . . . , y
n1
, 0) v
y
k
(y
1
, . . . , y
l
, . . . , y
n1
, 0)].
We only do not know v
ynyn
on
0
. Therefore the (3.1.2) can be rewritten as
n

i,j=1
A
ij

n
x
i

n
x
j
v
ynyn
= terms known on
0
.
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 63
Since
n
(x) = S(x), we can compute v
ynyn
if
n

i,j=1
A
ij
(x)S
x
i
S
x
j
,= 0
on . Note that S is the normal to the hypersurface S(x) = 0.
Denition 3.1.1. We say a hypersurface R
n
is non-characteristic
w.r.t (3.1.1) if
n

i,j=1
A
ij
(x)
i
(x)
j
(x) ,= 0.
where (x) is the normal vector of at x.
Since any real symmetric matrix can always be diagonalised, there is
a coordinate transformation T(x) such that the matrix T(x)A(x)T
t
(x) is
diagonal with diagonal entries, say
1
,
2
, . . . ,
n
, for each x . Since A(x)
is real symmetric all
i
R, for all i. Thus, we classify PDE at a point
x based on the eigenvalues of the matrix A(x). Let p denote the number
of eigenvalues that are strictly positive and z denote the number of zero
eigenvalues.
Denition 3.1.2. We say a PDE is hyperbolic at a point x , if z = 0
and either p = 1 or p = n 1. We say it is parabolic if z > 0. We say it is
elliptic, if z = 0 and either p = n or p = 0. If z = 0 and 1 < p < n 1 then
the PDE is said to be ultra hyperbolic.
Note that the elliptic case corresponds to the situation that every hy-
persurface S(x) = 0 with S ,= 0 is non-characteristic corresponding the
elliptic operator, i.e., there are no real characteristics curves. Thus, one can
equivalently dene a linear second order PDE to be elliptic at x if
n

i,j=1
A
ij
(x)
i

j
,= 0 R
n
0.
3.1.1 Semilinear: Two Dimension
Consider the Cauchy problem (3.0.1) in two variables and set x = (x, y). Let
denote the unit tangent vector on . Then, the directional derivative along
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 64
the tangent vector, Du(x, y) (x, y) = g

(x, y) is known because g is known.


Thus, we may compute directional derivative of u in any direction, along ,
as a linear combination of Du and Du . Using this we may reformulate
(3.0.1) as
_

_
F (D
2
u, Du, u, x, y) = 0 (x, y)
u(x, y) = g(x, y) (x, y)
u
x
(x, y) = h
1
(x, y) (x, y)
u
y
(x, y) = h
2
(x, y) (x, y)
with the compatibility condition that g(s) = h
1

1
(s) + h
2

2
(s), where s
(
1
(s),
2
(s)) is the parametrisation
1
of the hypersurface . The compatibil-
ity condition is an outcome of the fact that
u(s) = u
x

1
(s) + u
y

2
(s).
The above condition implies that among g, h
1
, h
2
only two can be assigned
independently.
Consider the Cauchy problem for the second order semi-linear PDE in
two variables (x, y) R
2
,
_

_
A(x, y)u
xx
+ 2B(x, y)u
xy
+C(x, y)u
yy
= D (x, y)
u(x, y) = g(x, y) (x, y)
u
x
(x, y) = h
1
(x, y) (x, y)
u
y
(x, y) = h
2
(x, y) (x, y) .
(3.1.3)
where D(x, y, u, u
x
, u
y
) may be non-linear and is a smooth
2
curve in .
Also, one of the coecients A, B or C is identically non-zero (else the PDE
is not of second order). Let s (
1
(s),
2
(s)) be a parametrisation of the
curve . Then we have the compatibility condition that
g(s) = h
1

1
(s) +h
2

2
(s).
By computing the second derivatives of u on and considering u
xx
, u
yy
and u
xy
as unknowns, we have the linear system of three equations in three
unknowns on ,
Au
xx
+2Bu
xy
+Cu
yy
= D

1
(s)u
xx
+
2
(s)u
xy
=

h
1
(s)

1
(s)u
xy
+
2
(s)u
yy
=

h
2
(s).
1
() denotes the derivative with respect to space variable and () denotes the derivative
with respect to parameter
2
twice dierentiable
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 65
This system of equation is solvable if the determinant of the coecients are
non-zero, i.e.,

A 2B C

1

2
0
0
1

2

,= 0.
Denition 3.1.3. We say a curve R
2
is characteristic (w.r.t
(3.1.3)) if
A
2
2
2B
1

2
+C
2
1
= 0.
where (
1
(s),
2
(s)) is a parametrisation of .
Note that the geometry hidden in the above denition is very similar
to that we encountered in rst order equation. Since = (
2
,
1
) is the
normal to at each point, the above denition says that the curve is non-
characteristic if
2

i,j=1
A
ij

j
= A
2
2
2B
1

2
+C
2
1
,= 0
where A
11
= A, A
12
= A
21
= B and A
22
= C. If y = y(x) is a representation
of the curve (locally, if necessary), we have
1
(s) = s and
2
(s) = y(s).
Then the characteristic equation reduces as
A
_
dy
dx
_
2
2B
dy
dx
+ C = 0.
Therefore, the characteristic curves of (3.1.3) are given by the graphs whose
equation is
dy
dx
=
B

B
2
AC
A
.
Thus, we have three situations arising depending on the sign of the dis-
criminant, B
2
AC. This classies the given second order PDE based on
the sign of its discriminant d = B
2
AC.
Denition 3.1.4. We say a second order PDE is of
(a) hyperbolic type if d > 0,
(b) parabolic type if d = 0 and
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 66
(c) elliptic type if d < 0.
The hyperbolic PDE have two families of characteristics, parabolic PDE
has one family of characteristic and elliptic PDE have no characteristic. We
caution here that these names are no indication of the shape of the graph
of the solution of the PDE. The classication tells us the right amount of
initial/boundary condition to be imposed for a PDE to be well-posed. For
hyperbolic, which has two real characteristics, requires as many initial con-
dition as the number of characteristics emanating from initial time and as
many boundary conditions as the number of characteristics that pass into the
spatial boundary. For parabolic, which has exactly one real characteristic,
we need one boundary condition at each point of the spatial boundary and
one initial condition at initial time. For elliptic, which has no real charac-
teristic curves, we need one boundary condition at each point of the spatial
boundary.
Note that the classication depends on the determinant of the coecient
matrix
_
A B
B C
_
For every (x, y) , the matrix is symmetric and hence diagonalisable.
If
1
,
2
are the diagonal entries, then d =
1

2
. Thus, a equation is
hyperbolic at a point (x, y) if the eigen values have opposite sign. It is ellipic
if the eigenvalues have same sign and is parabolic if, at least, one of the
eigenvalue is zero.
3.2 Quasilinear
All the arguments of semilinear PDE can be carried over to a quasilinear
PDE A(x, u(x), Du(x)). For each specic point x
0
, u(x
0
) = u
0
and A
0
=
A(x
0
, u
0
, u(x
0
)). The solutions to characteristic equation for a quasilinear
equation depends on the solution considered. Set
U := (x, z, p) [ x , z R, p R
n
.
Denition 3.2.1. A quasilinear equation
A(x, u(x), Du(x)) D
2
u(x) = D(x, u(x), Du(x))
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 67
is said to be elliptic if the matrix A
ij
(x, z, p) is positive denite for each
(x, z, p) U. Further,
0 < (x, z, p)[[
2

i,j=1
A
ij
(x, z, p)
i

j
(x, z, p)[[
2
R
n
.
The bounds (x, z, p) and (x, z, p) are minimum and maximum eigen-
values, respectively. If / is uniformly bounded in U then PDE is uniformly
elliptic. The interesting thing about uniformly elliptic equation is that they
behave very similar to linear elliptic equations.
Example 3.1. Consider the minimal surface equation

_
u
_
1 +[u[
2
_
= f(x)
where the second order coecients are
A
ij
(x, z, p) = (1 +[p[
2
)
1/2
_

ij

p
i
p
j
1 +[p[
2
_
and
(x, z, p) =
1
(1 +[p[
2
)
3/2
(x, z, p) =
1
(1 +[p[
2
)
1/2
.
Thus, the equation is not uniformly elliptic.
The minimal surface equation and the capillary equation are not uni-
formly elliptic.
3.3 Examples
Example 3.2 (Wave Equation). For a given c R, u
yy
c
2
u
xx
= 0 is hyper-
bolic. Since A = c
2
, B = 0 and C = 1, we have d = B
2
AC = c
2
> 0. The
eigen values of the coecient matrix are 1, c
2
which have opposite sign.
Example 3.3 (Heat Equation). For a given c R, u
y
cu
xx
= 0 is parabolic.
Since A = c, B = 0 and C = 0, thus d = B
2
AC = 0. The eigen values
of the coecient matrix are 0, c has a zero eigenvalue.
Example 3.4 (Laplace equation). u
xx
+u
yy
= 0 is elliptic. Since A = 1, B = 0
and C = 1, thus d = B
2
AC = 1 < 0. The eigen values of the coecient
matrix are 1, 1 which have same sign.
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 68
Example 3.5 (Velocity Potential Equation). In the equation (1 M
2
)u
xx
+
u
yy
= 0, A = (1 M
2
), B = 0 and C = 1. Then d = B
2
AC = (1 M
2
).
The eigen values of the coecient matrix are 1 M
2
, 1. Thus, for M > 1
(opposite sign), the equation is hyperbolic (supersonic ow), for M = 1 (zero
eigenvalue) it is parabolic (sonic ow) and for M < 1 (same sign) it is elliptic
(subsonic ow).
Note that the classication of PDE is dependent on its coecients. Thus,
for constant coecients the type of PDE remains unchanged throughout
the region . However, for variable coecients, the PDE may change its
classication from region to region.
Example 3.6. An example is the Tricomi equation , u
xx
+ xu
yy
= 0. The
discriminant of the Tricomi equation is d = x. The eigenvalues are 1, x.
Thus, tricomi equation is hyperbolic when x < 0, elliptic when x > 0 and
degenerately parabolic when x = 0, i.e., y-axis. Such equations are called
mixed type.
The notion of classication of second order semi-linear PDE, discussed in
this section, could be generalised to quasi-linear, non-linear PDE and system
of ODE. However, in these cases the classication may also depend on the
solution u, as seen in the examples below.
Example 3.7. Consider the quasi-linear PDE u
xx
uu
yy
= 0. The discrimi-
nant is d = u. The eigenvalues are 1, u(x). It is hyperbolic for u > 0
3
,
elliptic when u < 0 and parabolic when u = 0.
Example 3.8. Consider the quasi-linear PDE
(c
2
u
2
x
)u
xx
2u
x
u
y
u
xy
+ (c
2
u
2
y
)u
yy
= 0
where c > 0. Then d = B
2
AC = c
2
(u
2
x
+ u
2
y
c
2
) = c
2
([u[
2
c
2
). It is
hyperbolic if [u[ > c, parabolic if [u[ = c and elliptic if [u[ < c.
Example 3.9. Find the family of characteristic curves for the following second
order PDE, whenever they exist.
(i) For a given c R, u
yy
c
2
u
xx
= 0.
(ii) For a given c R, u
y
cu
xx
= 0.
(iii) u
xx
+ u
yy
= 0.
3
The notation u > 0 means x [ u(x) > 0
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 69
(iv) u
xx
+ xu
yy
= 0.
Solution. (i) We have already seen the equation is hyperbolic and hence
it should have two characteristic curves. The characteristic curves are
given by the equation
dy
dx
=
B

B
2
AC
A
=

c
2
c
2
=
1
c
.
Thus, cy x = a constant is the equation of the two characteristic
curves. Note that the characteristic curves y = x/c+y
0
are boundary
of two cones in R
2
with vertex at (0, y
0
).
(ii) We have already seen the equation is parabolic and hence it should
have one characteristic curve. The characteristic curve are given by the
equation
dy
dx
=
B

B
2
AC
A
= 0.
Thus, y = a constant is the equation of the characteristic curve. i.e.,
any horizontal line in R
2
is a charateristic curve.
(iii) We have already seen the equation is elliptic and hence has no real
characteristics.
(iv) The equation is of mixed type. In the region x > 0, the characteristic
curves are y 2x
3/2
/3 = a constant.
Exercise 15. Classify the following second order PDE in terms of the number
of characteristics:
(a) 3u
xx
+u
xy
+ 2u
yy
= 0.
(b) u
zz
+ u
z
+ u
rr
+
1
r
u

+ c = 0.
(c) u
t
+ u
x
+ u
xx
= 0.
(d) 4u
xx
+y
2
u
x
+xu
x
+ u
yy
+ 4u
xy
4xy = 0.
(e) xu
xx
+ xu
xy
+yu
yy
= 0.
(f) xu
xx
+ yu
xy
+ c = 0.
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 70
(g) x
2
yu
xx
+ xyu
xy
y
2
u
yy
= 0.
(h) sin xu
xx
+ 2 cos xu
xy
+ sin xu
yy
= 0.
(i) u
xx
+ 4u
xy
+ 5u
yy
+u
x
+ 2u
y
= 0.
(j) u
xx
4u
xy
+ 4u
yy
+ 3u
x
+ 4u = 0.
(k) u
xx
+ 2u
xy
3u
yy
+ 2u
x
+ 6u
y
= 0.
(l) (1 + x)u
xx
+ 2xyu
xy
y
2
u
yy
= 0.
(m) 2u
xx
4u
xy
+ 7u
yy
u = 0.
(n) u
xx
2 cos xu
xy
sin
2
xu
yy
= 0.
(o) yu
xx
+ 2(x 1)u
xy
(y + 2)u
yy
= 0.
(p) yu
xx
+ u
xy
x
2
u
yy
u
x
u = 0.
Exercise 16. Classify the following second order PDE, in terms of the number
of characteristics, and nd their characteristics, when it exists:
(a) u
xx
+ (5 + 2y
2
)u
xy
+ (1 + y
2
)(4 + y
2
)u
yy
= 0.
(b) yu
xx
+ u
yy
= 0.
(c) yu
xx
= xu
yy
.
(d) u
yy
xu
xy
+ yu
x
+ xu
y
= 0.
(e) y
2
u
xx
+ 2xyu
xy
+x
2
u
yy
= 0.
(f) u
xx
+ 2xu
xy
+ (1 y
2
)u
yy
= 0.
3.4 System of First Order PDE
Recall that any second order ODE y

= f(x, y, y

) can be equivalently written


as a system of ODE, by using u
1
= y and u
2
= y

, as follows:
_
u

1
= u
2
u

2
= f(x, u
1
, u
2
).
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 71
A similar procedure also makes a second order PDE into a system of rst
order PDEs. Thus, we expect that our classication of second order PDE to
induce a classication for a system of rst order PDE.
A general system of m rst order linear PDE in n variables will be of the
form
n

j=1
A
j
(x, u)u
x
j
= f (x, u) in , (3.4.1)
where each A
j
(x, u) is a m m matrix, u(x) = (u
1
, . . . , u
m
) and f =
(f
1
, . . . , f
m
) has m components. Following the arguments for semilinear PDE
and using the map (x) and v(y) between and
0
, we get
n

j=1
A
j
(x, u)v
yn
S
x
j
= terms known on
0
.
This system is solvable for v
yn
if
det
_
n

j=1
A
j
(x, u)S
x
j
_
,= 0
on .
Denition 3.4.1. We say a hyperspace in R
n
is non-characteristic if
det
_
n

j=1
A
j
(x, u)
j
(x)
_
,= 0
on , where (x) = (
1
, . . . ,
n
) is the normal at x of .
Denition 3.4.2. The system of rst order PDE (3.4.1) is called hyperbolic
if the mm matrix
n

j=1

j
A
j
(x, u)
is diagonalisable for every = (
j
) R
n
and every x . If the eigen-
values are all distinct, for non-zero R
n
, then the PDE is called strictly
hyperbolic.
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 72
Denition 3.4.3. The system of rst order PDE (3.4.1) is called elliptic if
det
_
n

j=1
A
j
(x, u)
j
_
= 0
only for = 0.
Example 3.10 (Beltrami Equations). Consider the system of rst order equa-
tions
_
W(x, y)u
x
(x, y) b(x, y)v
x
(x, y) c(x, y)v
y
(x, y) = 0
W(x, y)u
y
(x, y) + a(x, y)v
x
(x, y) + b(x, y)v
y
(x, y) = 0
where W, a, b, c are given such that W ,= 0 and the matrix
_
a b
b c
_
is positive denite. Set
A
1
=
_
W b
0 a
_
A
2
=
_
0 c
W b
_
.
Then the system can be rewritten as
A
1
_
u
x
v
x
_
+A
2
_
u
y
v
y
_
=
_
0
0
_
.
Then
det
_
n

j=1
A
j
(x, u)
j
_
=

W
1
b
1
c
2
W
2
a
1
+ b
2

= W(a
2
1
+ 2b
1

2
+ c
2
2
) ,= 0
if ,= 0. Therefore, the Beltrami equation is elliptic. The Beltrami system
is a generalization of the Cauchy-Riemann equations.
3.5 System of Second Order PDE
A general system of m second order linear PDE in n variables will be of the
form
n

i,j=1
A
ij
(x, u, u)u
x
i
x
j
+ lower order terms = 0 in , (3.5.1)
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 73
where each A
ij
is a m m matrix and u(x) = (u
1
, . . . , u
m
). Following the
arguments for semilinear PDE and using the map (x) and v(y) between
and
0
, we get
n

i,j=1
A
ij
(x, u, u)v
ynyn
S
x
i
S
x
j
= terms known on
0
.
Denition 3.5.1. The system is called elliptic if
det
_
n

i,j=1
A
ij

j
_
= 0
only for = 0.
3.6 Invariance of Discriminant
The classication of second order semi-linear PDE is based on the discrimi-
nant B
2
AC. In this section, we note that the classication is independent
of the choice of coordinate system (to represent a PDE). Consider the two-
variable semilinear PDE
A(x, y)u
xx
+2B(x, y)u
xy
+C(x, y)u
yy
= D(x, y, u, u
x
, u
y
) (x, y) (3.6.1)
where the variables (x, y, u, u
x
, u
y
) may appear non-linearly in D and R
2
.
Also, one of the coecients A, B or C is identically non-zero (else the PDE is
not of second order). We shall observe how (3.6.1) changes under coordinate
transformation.
Denition 3.6.1. For any PDE of the form (3.6.1) we dene its discrimi-
nant as B
2
AC.
Let T : R
2
R
2
be the coordinate transformation with components
T = (w, z), where w, z : R
2
R. We assume that w(x, y), z(x, y) are such
that w, z are both continuous and twice dierentiable w.r.t (x, y), and the
Jacobian J of T is non-zero,
J =

w
x
w
y
z
x
z
y

,= 0.
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 74
We compute the derivatives of u in the new variable,
u
x
= u
w
w
x
+ u
z
z
x
,
u
y
= u
w
w
y
+u
z
z
y
,
u
xx
= u
ww
w
2
x
+ 2u
wz
w
x
z
x
+u
zz
z
2
x
+ u
w
w
xx
+u
z
z
xx
u
yy
= u
ww
w
2
y
+ 2u
wz
w
y
z
y
+ u
zz
z
2
y
+ u
w
w
yy
+ u
z
z
yy
u
xy
= u
ww
w
x
w
y
+ u
wz
(w
x
z
y
+ w
y
z
x
) + u
zz
z
x
z
y
+u
w
w
xy
+ u
z
z
xy
Substituting above equations in (3.6.1), we get
a(w, z)u
ww
+ 2b(w, z)u
wz
+ c(w, z)u
zz
= d(w, z, u, u
w
, u
z
).
where D transforms in to d and
a(w, z) = Aw
2
x
+ 2Bw
x
w
y
+Cw
2
y
(3.6.2)
b(w, z) = Aw
x
z
x
+B(w
x
z
y
+ w
y
z
x
) + Cw
y
z
y
(3.6.3)
c(w, z) = Az
2
x
+ 2Bz
x
z
y
+ Cz
2
y
. (3.6.4)
Note that the coecients in the new coordinate system satisfy
b
2
ac = (B
2
AC)J
2
.
Since J ,= 0, we have J
2
> 0. Thus, both b
2
ac and B
2
AC have the
same sign. Thus, the sign of the discriminant is invariant under coordinate
transformation. All the above arguments can be carried over to quasi-linear
and non-linear PDE.
3.7 Standard or Canonical Forms
The advantage of above classication helps us in reducing a given PDE into
simple forms. Given a PDE, one can compute the sign of the discriminant
and depending on its clasication we can choose a coordinate transformation
(w, z) such that
(i) For hyperbolic, a = c = 0 or b = 0 and a = c.
(ii) For parabolic, c = b = 0 or a = b = 0. We conveniently choose
c = b = 0 situation so that a ,= 0 (so that division by zero is avoided in
the equation for characteristic curves).
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 75
(iii) For elliptic, b = 0 and a = c.
If the given second order PDE (3.6.1) is such that A = C = 0, then
(3.6.1) is of hyperbolic type and a division by 2B (since B ,= 0) gives
u
xy
=

D(x, y, u, u
x
, u
y
)
where

D = D/2B. The above form is the rst standard form of second order
hyperbolic equation. If we introduce the linear change of variable X = x +y
and Y = x y in the rst standard form, we get the second standard form
of hyperbolic PDE
u
XX
u
Y Y
=

D(X, Y, u, u
X
, u
Y
).
If the given second order PDE (3.6.1) is such that A = B = 0, then
(3.6.1) is of parabolic type and a division by C (since C ,= 0) gives
u
yy
=

D(x, y, u, u
x
, u
y
)
where

D = D/C. The above form is the standard form of second order
parabolic equation.
If the given second order PDE (3.6.1) is such that A = C and B = 0,
then (3.6.1) is of elliptic type and a division by A (since A ,= 0) gives
u
xx
+ u
yy
=

D(x, y, u, u
x
, u
y
)
where

D = D/A. The above form is the standard form of second order
elliptic equation.
Note that the standard forms of the PDE is an expression with no mixed
derivatives.
3.8 Reduction to Standard Form
Consider the second order semi-linear PDE (3.6.1) not in standard form. We
look for transformation w = w(x, y) and z = z(x, y), with non-vanishing
Jacobian, such that the reduced form is the standard form.
If B
2
AC > 0, we have two characteristics. We are looking for the
coordinate system w and z such that a = c = 0. This implies from equation
(3.6.2) and (3.6.4) that we need to nd w and z such that
w
x
w
y
=
B

B
2
AC
A
=
z
x
z
y
.
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 76
Therefore, we need to nd w and z such that along the slopes of the charac-
teristic curves,
dy
dx
=
B

B
2
AC
A
=
w
x
w
y
.
This means that, using the parametrisation of the characteristic curves,
w
x

1
(s) + w
y

2
(s) = 0 and

w(s) = 0. Similarly for z. Thus, w and z
are chosen such that they are constant on the characteristic curves.
The characteristic curves are found by solving
dy
dx
=
B

B
2
AC
A
and the coordinates are then chosen such that along the characteristic curve
w(x, y) = a constant and z(x, y) = a constant. Note that w
x
z
y
w
y
z
x
=
w
y
z
y
_
2
A

B
2
AC
_
,= 0.
Example 3.11. Let us reduce the PDE u
xx
c
2
u
yy
= 0 to its canonical form.
Note that A = 1, B = 0, C = c
2
and B
2
AC = c
2
and the equation is
hyperbolic. The characteristic curves are given by the equation
dy
dx
=
B

B
2
AC
A
= c.
Solving we get y cx = a constant. Thus, w = y +cx and z = y cx. Now
writing
u
xx
= u
ww
w
2
x
+ 2u
wz
w
x
z
x
+ u
zz
z
2
x
+ u
w
w
xx
+ u
z
z
xx
= c
2
(u
ww
2u
wz
+u
zz
)
u
yy
= u
ww
w
2
y
+ 2u
wz
w
y
z
y
+ u
zz
z
2
y
+ u
w
w
yy
+ u
z
z
yy
= u
ww
+ 2u
wz
+u
zz
c
2
u
yy
= c
2
(u
ww
+ 2u
wz
+ u
zz
)
Substituting into the given PDE, we get
0 = 4c
2
u
wz
= u
wz
.
Example 3.12. Let us reduce the PDE u
xx
x
2
yu
yy
= 0 given in the region
(x, y) [ x R, x ,= 0, y > 0 to its canonical form. Note that A = 1, B = 0,
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 77
C = x
2
y and B
2
AC = x
2
y. In the given region x
2
y > 0, hence the
equation is hyperbolic. The characteristic curves are given by the equation
dy
dx
=
B

B
2
AC
A
= x

y.
Solving we get x
2
/2 2

y = a constant. Thus, w = x
2
/2 + 2

y and
z = x
2
/2 2

y. Now writing
u
x
= u
w
w
x
+u
z
z
x
= x(u
w
+u
z
)
u
y
= u
w
w
y
+ u
z
z
y
=
1

y
(u
w
u
z
)
u
xx
= u
ww
w
2
x
+ 2u
wz
w
x
z
x
+u
zz
z
2
x
+u
w
w
xx
+u
z
z
xx
= x
2
(u
ww
+ 2u
wz
+u
zz
) + u
w
+ u
z
u
yy
= u
ww
w
2
y
+ 2u
wz
w
y
z
y
+u
zz
z
2
y
+u
w
w
yy
+ u
z
z
yy
=
1
y
(u
ww
2u
wz
+ u
zz
)
1
2y

y
(u
w
u
z
)
x
2
yu
yy
= x
2
(u
ww
2u
wz
+ u
zz
) +
x
2
2

y
(u
w
u
z
)
Substituting into the given PDE, we get
0 = 4x
2
u
wz
+
2

y + x
2
2

y
u
w
+
2

y x
2
2

y
u
z
= 8x
2

yu
wz
+ (x
2
+ 2

y)u
w
+ (2

y x
2
)u
z
.
Note that w + z = x
2
and w z = 4

y. Now, substituting x, y in terms of


w, z, we get
0 = 2(w
2
z
2
)u
wz
+
_
w +z +
w z
2
_
u
w
+
_
w z
2
w z
_
u
z
= u
wz
+
_
3w + z
4(w
2
z
2
)
_
u
w

_
w + 3z
4(w
2
z
2
)
_
u
z
.
Example 3.13. Let us reduce the PDE u
xx
+u
xy
2u
yy
+ 1 = 0 given in the
region (x, y) [ 0 x 1, y > 0 to its canonical form. Note that A = 1,
B = 1/2, C = 2 and B
2
AC = 9/4 > 0. Hence the equation is hyperbolic.
The characteristic curves are given by the equation
dy
dx
=
B

B
2
AC
A
=
1
2

3
2
= 2 or 1.
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 78
Solving we get y 2x = a constant and y +x = a constant. Thus, w = y 2x
and z = y + x.
In the parabolic case, B
2
AC = 0, we have a single characteristic. We
are looking for a coordinate system such that either b = c = 0.
Example 3.14. Let us reduce the PDE e
2x
u
xx
+ 2e
x+y
u
xy
+ e
2y
u
yy
= 0 to its
canonical form. Note that A = e
2x
, B = e
x+y
, C = e
2y
and B
2
AC = 0.
The PDE is parabolic. The characteristic curves are given by the equation
dy
dx
=
B
A
=
e
y
e
x
.
Solving, we get e
y
e
x
= a constant. Thus, w = e
y
e
x
. Now, we
choose z such that w
x
z
y
w
y
z
x
,= 0. For instance, z = x is one such choice.
Then
u
x
= e
x
u
w
+ u
z
u
y
= e
y
u
w
u
xx
= e
2x
u
ww
+ 2e
x
u
wz
+ u
zz
e
x
u
w
u
yy
= e
2y
u
ww
+ e
y
u
w
u
xy
= e
y
(e
x
u
ww
u
wz
)
Substituting into the given PDE, we get
e
x
e
y
u
zz
= (e
y
e
x
)u
w
Replacing x, y in terms of w, z gives
u
zz
=
w
1 +we
z
u
w
.
Example 3.15. Let us reduce the PDE y
2
u
xx
2xyu
xy
+ x
2
u
yy
=
1
xy
(y
3
u
x
+
x
3
u
y
) to its canonical form. Note that A = y
2
, B = xy, C = x
2
and
B
2
AC = 0. The PDE is parabolic. The characteristic curves are given by
the equation
dy
dx
=
B
A
=
x
y
.
Solving, we get x
2
+ y
2
= a constant. Thus, w = x
2
+y
2
. Now, we choose z
such that w
x
z
y
w
y
z
x
,= 0. For instance, z = x is one such choice. Then
u
x
= 2xu
w
+ u
z
u
y
= 2yu
w
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 79
In the elliptic case, B
2
AC < 0, we have no real characteristics. Thus,
we choose w, z to be the real and imaginary part of the solution of the
characteristic equation.
Example 3.16. Let us reduce the PDE x
2
u
xx
+y
2
u
yy
= 0 given in the region
(x, y) R
2
[ x > 0, y > 0 to its canonical form. Note that A = x
2
,
B = 0, C = y
2
and B
2
AC = x
2
y
2
< 0. The PDE is elliptic. Solving the
characteristic equation
dy
dx
=
iy
x
we get ln x i ln y = c. Let w = ln x and z = ln y. Then
u
x
= u
w
/x
u
y
= u
z
/y
u
xx
= u
w
/x
2
+ u
ww
/x
2
u
yy
= u
z
/y
2
+ u
zz
/y
2
Substituting into the PDE, we get
u
ww
+u
zz
= u
w
+u
z
.
Example 3.17. Let us reduce the PDE u
xx
+2u
xy
+5u
yy
= xu
x
to its canonical
form. Note that A = 1, B = 1, C = 5 and B
2
AC = 4 < 0. The PDE is
elliptic. The characteristic equation is
dy
dx
= 1 2i.
Solving we get x y i2x = c. Let w = x y and z = 2x. Then
u
x
= u
w
+ 2u
z
u
y
= u
w
u
xx
= u
ww
+ 4u
wz
+ 4u
zz
u
yy
= u
ww
u
xy
= (u
ww
+ 2u
wz
)
Substituting into the PDE, we get
u
ww
+ u
zz
= x(u
w
+ 2u
z
)/4.
Replacing x, y in terms of w, z gives
u
ww
+ u
zz
=
z
8
(u
w
+ 2u
z
).
CHAPTER 3. CLASSIFICATION BY CHARACTERISTICS 80
Example 3.18. Let us reduce the PDE u
xx
+ u
xy
+ u
yy
= 0 to its canonical
form. Note that A = 1, B = 1/2, C = 1 and B
2
AC = 3/4 < 0. The
PDE is elliptic. Solving the characteristic equation
dy
dx
=
1
2
i

3
2
we get 2y = x i

3x +c. Let w = 2y x and z =

3x.
Exercise 17. Rewrite the PDE in their canonical forms and solve them.
(a) u
xx
+ 2

3u
xy
+ u
yy
= 0
(b) x
2
u
xx
2xyu
xy
+ y
2
u
yy
+ xu
x
+ yu
y
= 0
(c) u
xx
(2 sin x)u
xy
(cos
2
x)u
yy
(cos x)u
y
= 0
(d) u
xx
+ 4u
xy
+ 4u
yy
= 0
Chapter 4
Wave Equation
4.1 One Dimension
The one dimensional wave equation is the rst ever partial dierential equa-
tion (PDE) to be studied by mankind, introduced in 1752 by dAlembert as
a model to study vibrating strings. He introduced the one dimensional wave
equation

2
u(x, t)
t
2
=

2
u(x, t)
x
2
.
dAlembert used the travelling wave technique to solve the wave equation.
In this chapter we shall explain this technique of dAlembert and also give
the standing wave technique which motivates the idea of separation of vari-
able and in turn the evolution of Fourier series. The wave equations was
generalised to two and three dimensions by Euler (1759) and D. Bernoulli
(1762), respectively. Note the invariance of the wave equation under the
transformation t t, i.e., if u(x, t) is a solution to the wave equation for
t 0, then u(x, t) := u(x, ) is a solution of the wave equation for t < 0 and
:= t > 0, because d/dt = 1, u
t
(x, t) = u

(x, ). This means that


wave equation is reversible in time and do not distinguish between past and
future.
4.1.1 The Vibrating String: Derivation
Let us consider a homogeneous string of length L, stretched along the x-axis,
with one end xed at x = 0 and the other end xed at x = L. We assume
that the string is free to move only in the vertical direction. Let > 0 denote
81
CHAPTER 4. WAVE EQUATION 82
the density of the string and T > 0 denote the coecient of tension of the
string. Let u(x, t) denote the vertical displacement of the string at the point
x and time t.
We shall imagine the string of length L as system of N objects, for N
suciently large. Think of N objects sitting on the string L at equidistant
(uniformly distributed). The position of the n-th object on the string is
given by x
n
= nL/N. One can think of the vibrating string as the harmonic
oscillator of N objects governed by the tension on the string (which behaves
like the spring). Let y
n
(t) = u(x
n
, t) denote the displacement of the object x
n
at time t. The distance between any two successive objects is h = x
n+1
x
n
=
L/N. Then mass of each of the N object is mass of the string divided by N.
Since mass of the string is L, mass of each of the object x
n
, n = 1, 2, . . . , N,
is h. Thus, by Newtons second law, hy

n
(t) is same as the force acting on
the n-th object. The force on x
n
is coming both from left (x
n1
) and right
(x
n+1
) side. The force from left and right is given as T(y
n1
y
n
)/h and
T(y
n+1
y
n
)/h, respectively. Therefore,
hy

n
(t) =
T
h
y
n+1
(t) + y
n1
(t) 2y
n
(t)
=
T
h
u(x
n
+h, t) +u(x
n
h, t) 2u(x
n
, t)
y

n
(t) =
T

_
u(x
n
+ h, t) + u(x
n
h, t) 2u(x
n
, t)
h
2
_
Note that assuming u is twice dierentiable w.r.t the x variable, the term
on RHS is same as
T

1
h
_
u(x
n
+h, t) u(x
n
, t)
h
+
u(x
n
h, t) u(x
n
, t)
h
_
which converges to the second partial derivative of u w.r.t x as h 0. The
h 0 is the limit case of the N objects we started with. Therefore the
vibrating string system is governed by the equation

2
u
t
2
=
T

2
u
x
2
where T is the tension and is the density of the string. Equivalently,

2
u
t
2
= c
2

2
u
x
2
(4.1.1)
where c
2
= T/, c > 0, on x (0, L) and t > 0.
CHAPTER 4. WAVE EQUATION 83
Exercise 18. Show that the wave equation (4.1.1), derived above can be
written as
u
zz
= u
ww
in (w, z) (0, L) (0, ).
under a new coordinate system (w, z). One may, in fact choose coordinate
such that the string is xed between (0, ).
Proof. Set w = x/a and z = t/b, where a and b will be chosen appropriately.
Then, w
x
= 1/a and z
t
= 1/b. Therefore, u
x
= u
w
/a, u
t
= u
z
/b, a
2
u
xx
= u
ww
and b
2
u
tt
= u
zz
. Choosing a = 1 and b = 1/c. Choosing a = L/ and
b = L/c makes the domain (0, ).
4.1.2 Travelling Waves
Consider the wave equation u
tt
= c
2
u
xx
on R(0, ), describing the vibra-
tion of an innite string. We have already seen in Chapter 3 that the equation
is hyperbolic and has the two characteristics x ct= a constant. Introduce
the new coordinates w = x + ct, z = x ct and set u(w, z) = u(x, t). Thus,
we have the following relations, using chain rule:
u
x
= u
w
w
x
+ u
z
z
x
= u
w
+ u
z
u
t
= u
w
w
t
+ u
z
z
t
= c(u
w
u
z
)
u
xx
= u
ww
+ 2u
zw
+ u
zz
u
tt
= c
2
(u
ww
2u
zw
+ u
zz
)
In the new coordinates, the wave equation satises u
wz
= 0. Integrating
1
this twice, we have u(w, z) = F(w) + G(z), for some arbitrary functions F
and G. Thus, u(x, t) = F(x + ct) + G(x ct) is a general solution of the
wave equation.
Consider the case where G is chosen to be zero function. Then u(x, t) =
F(x+ct) solves the wave equation. At t = 0, the solution is simply the graph
of F and at t = t
0
the solution is the graph of F with origin translated to the
left by ct
0
. Similarly, choosing F = 0 and G = F, we have u(x, t) = F(xct)
also solves wave equation and at time t is the translation to the right of the
graph of F by ct. This motivates the name travelling waves and wave
equation. The graph of F is shifted to right or left with a speed of c.
1
We are assuming the function is integrable, which may be false
CHAPTER 4. WAVE EQUATION 84
Now that we have derived the general form of the solution of wave equa-
tion, we return to understand the physical system of a vibrating innite
string. The initial shape (position at initial time t = 0) of the string is given
as u(x, 0) = g(x), where the graph of g on R
2
describes the shape of the
string. Since we need one more data to identify the arbitrary functions, we
also prescribe the initial velocity of the string, u
t
(x, 0) = h(x).
Another interesting property that follows from the general solution is
that for any four points A, B, C and D that form a rectangle bounded by
characteristic curves in R R
+
then u(A) + u(C) = u(B) + u(D) because
u(A) = F() +G(), u(C) = F() +G(), u(B) = F() +G() and u(D) =
F() +G().
Theorem 4.1.1. Given g C
2
(R) and h C
1
(R), there is a unique C
2
solution u of the Cauchy initial value problem (IVP) of the wave equation,
_
_
_
u
tt
(x, t) c
2
u
xx
(x, t) = 0 in R (0, )
u(x, 0) = g(x) in R
u
t
(x, 0) = h(x) in R,
(4.1.2)
which is given by the dAlemberts formula
u(x, t) =
1
2
(g(x + ct) + g(x ct)) +
1
2c
_
x+ct
xct
h(y) dy. (4.1.3)
Proof. The general solution is u(x, t) = F(x + ct) + G(x ct) with F, G
C
2
(R). Using the initial position we get
F(x) + G(x) = g(x).
Thus, g should be C
2
(R). Now, u
t
(x, t) = c (F

(w) G

(z)) and putting


t = 0, we get
F

(x) G

(x) =
1
c
h(x).
Thus, h should be C
1
(R). Now solving for F

and G

, we get 2F

(x) =
g

(x) + h(x)/c. Similarly, 2G

(x) = g

(x) h(x)/c. Integrating


2
both these
equations, we get
F(x) =
1
2
_
g(x) +
1
c
_
x
0
h(y) dy
_
+c
1
2
assuming they are integrable and the integral of their derivatives is itself
CHAPTER 4. WAVE EQUATION 85
and
G(x) =
1
2
_
g(x)
1
c
_
x
0
h(y) dy
_
+ c
2
.
Since F(x) +G(x) = g(x), we get c
1
+c
2
= 0. Therefore, the solution to the
wave equation is given by (4.1.3).
Aliter. Let us derive the dAlemberts formula in an alternate way. Note
that the wave equation can be factored as
_

t
+ c

x
__

t
c

x
_
u = u
tt
c
2
u
xx
= 0.
We set v(x, t) =
_

t
c

x
_
u(x, t) and hence
v
t
(x, t) +cv
x
(x, t) = 0 in R (0, ).
Notice that the above rst order PDE obtained is in the form of homogeneous
linear transport equation (cf. (2.2.1)), which we have already solved. Hence,
for some smooth function ,
v(x, t) = (x ct)
and (x) := v(x, 0). Using v in the original equation, we get the inhomoge-
neous transport equation,
u
t
(x, t) cu
x
(x, t) = (x ct).
Recall the formula for inhomogenoeus transport equation (cf. (2.2.2))
u(x, t) = g(x at) +
_
t
0
(x a(t s), s) ds.
Since u(x, 0) = g(x) and a = c, in our case the solution reduces to,
u(x, t) = g(x + ct) +
_
t
0
(x + c(t s) cs) ds
= g(x + ct) +
_
t
0
(x + ct 2cs) ds
= g(x + ct) +
1
2c
_
xct
x+ct
(y) dy
= g(x + ct) +
1
2c
_
x+ct
xct
(y) dy.
CHAPTER 4. WAVE EQUATION 86
But (x) = v(x, 0) = u
t
(x, 0) cu
x
(x, 0) = h(x) cg

(x) and substituting


this in the formula for u, we get
u(x, t) = g(x + ct) +
1
2c
_
x+ct
xct
(h(y) cg

(y)) dy
= g(x + ct) +
1
2
(g(x ct) g(x + ct))
+
1
2c
_
x+ct
xct
h(y) dy
=
1
2
(g(x ct) + g(x + ct)) +
1
2c
_
x+ct
xct
h(y) dy
For c = 1, the dAlemberts formula takes the form
u(x, t) =
1
2
(g(x t) + g(x + t)) +
1
2
_
x+t
xt
h(y) dy.
A useful observation from the dAlemberts formula is that the regularity of
u is same as the regularity of its initial value g.
Theorem 4.1.2 (Dirichlet Condition). Given g C
2
[0, ), h C
1
[0, )
and C
2
(0, ), there is a unique C
2
solution u of the homogeneous Cauchy
initial value problem (IVP) of the wave equation,
_

_
u
tt
(x, t) c
2
u
xx
(x, t) = 0 in (0, ) (0, )
u(x, 0) = g(x) in [0, )
u
t
(x, 0) = h(x) in [0, )
u(0, t) = (t) in (0, ),
(4.1.4)
where , g, h satises the compatibility condition
g(0) = (0), g

(0) =

(0), h(0) =

(0).
Proof. We rst consider case 0, homogeneous Dirichlet conditions. We
extend g as an odd function on (, ) by setting g(x) = g(x) for
x [0, ). Therefore, g is an odd function in R. Then, we have a unique
solution u C
2
solving the Cauchy problem in R (0, ). But note that
v(x, t) = u(x, t) is also a solution to the same Cauchy problem. Thus,
u(x, t) = u(x, t) for all (x, t) R R
+
and u(0, t) = 0.
CHAPTER 4. WAVE EQUATION 87
Now, we consider the case when is a non-zero function. The line ct = x
divides the quadrant in to two regions
1
and
2
dened as

1
= (x, t) [ x > ct
and

2
= (x, t) [ x < ct.
For (x, t)
1
, the solution
u
1
(x, t) =
1
2
(g(x ct) + g(x + ct)) +
1
2c
_
x+ct
xct
h(y) dy.
On the line x = ct, we get
(x) := u
1
(x, x/c) =
1
2
(g(0) + g(2x)) +
1
2c
_
2x
0
h(y) dy.
Let u
2
be the solution in
2
of
_
_
_
u
tt
(x, t) c
2
u
xx
(x, t) = 0 in
2
u(x, x/c) = (x) in x = ct
u(0, t) = (t) in (0, ).
Fix A := (x, t)
2
. One of the characteristic curve through A intersects
t-axis at B := (0, t x/c). The other characteristic curve intersects the
line ct = x at C := (1/2(ct + x, t + x/c). The characteristic curve through
B intersects ct = x at D := 1/2(ct x, t x/c). The four points form a
parallelogram in
2
. Therefore, we know that
u
2
(x, t) + u
2
(1/2(ct x, t x/c)) = u
2
(0, t x/c) + u
2
(1/2(ct + x, t + x/c))
and, hence,
u
2
(x, t) = (t x/c) + (1/2(ct + x)) (1/2(ct x))
= (t x/c) +
1
2
(g(0) + g(ct + x)) +
1
2c
_
ct+x
0
h(y) dy

1
2
(g(0) + g(ct x))
1
2c
_
ctx
0
h(y) dy
= (t x/c) +
1
2
(g(ct + x) g(ct x)) +
1
2c
_
ct+x
ctx
h(y) dy.
CHAPTER 4. WAVE EQUATION 88
By setting
u(x, t) =
_
u
1
(x, t) (x, t)
1
u
2
(x, t) (x, t)
2
and the fact that alll derivatives of u are continuous along ct = x line, due
to the compatibility condition implies that u is a solution (4.1.4).
Corollary 4.1.3 (Dirichlet Condition). Given g C
2
[0, L], h C
1
[0, L] and
, C
2
(0, ), there is a unique C
2
solution u of the homogeneous Cauchy
initial value problem (IVP) of the wave equation,
_

_
u
tt
(x, t) c
2
u
xx
(x, t) = 0 in (0, L) (0, )
u(x, 0) = g(x) in [0, L]
u
t
(x, 0) = h(x) in [0, L]
u(0, t) = (t) in (0, )
u(L, t) = (t) in (0, ),
(4.1.5)
where , , g, h satises the compatibility condition
g(0) = (0), g

(0) =

(0), h(0) =

(0)
and
g(L) = (0), g

(L) =

(0), h(L) =

(0).
Proof. We rst consider case = 0, homogeneous Dirichlet conditions.
We extend g as an odd function on [L, L] by setting g(x) = g(x) for
x [0, L]. Therefore, g is an odd 2L-periodic function in R. Similarly,
we extend h and we have g C
2
(R) and h C
1
(R). Then, we have a
unique solution u C
2
solving the Cauchy problem in R (0, ). But
note that v(x, t) = u(x, t) is also a solution to the same Cauchy problem.
Thus, u(x, t) = u(x, t) for all (x, t) R R
+
and u(0, t) = 0. Similarly,
w(x, t) = u(2Lx, t) is also a solution and, hence, u(x, t) = u(2Lx, t)
which implies u(L, t) = 0.
Consider the lines ct = x and ct = x +cL, then we will obtain u in the
four regions as u
1
, u
2
, u
3
and u
4
. Then follow the idea similar to proof in
above theorem.
4.1.3 Domain of Dependence and Inuence
Note that the solution u(x, t) depends only on the interval [x ct, x + ct]
because g takes values only on the end-points of this interval and h takes
CHAPTER 4. WAVE EQUATION 89
values between this interval. The interval [xct, x+ct] is called the domain
of dependence. Thus, the region of R (0, ) on which the value of u(x, t)
depends forms a triangle with base [x ct, x + ct] and vertex at (x, t). The
domain of dependence of (x, t) is marked in x-axis by the characteristic curves
passing through (x, t).
Given a point p on the x-axis what values of u on (x, t) will depend on
the value of g(p) and h(p). This region turns out to be a cone with vertex
at p and is called the domain of inuence. The domain of inuence is the
region bounded by the characteristic curves passing through p.
If the initial data g and h are supported in the interval B
x
0
(R) then the
solution u at (x, t) is supported in the region B
x
0
(R+ct). Consequently, if g
and h have compact support then the solution u has compact support in R
for all time t > 0. This phenomenon is called the nite speed of propagation.
Theorem 4.1.4 (Inhomogeneous Wave Equation). Given g C
2
(R), h
C
1
(R) and f C
1
(R [0, ), there is a unique C
2
solution u of the inho-
mogeneous Cauchy initial value problem (IVP) of the wave equation,
_
_
_
u
tt
(x, t) c
2
u
xx
(x, t) = f(x, t) in R (0, )
u(x, 0) = g(x) in R
u
t
(x, 0) = h(x) in R,
(4.1.6)
given by the formula
1
2
[g(x + ct) + g(x ct)] +
1
2c
_
_
x+ct
xct
h(y) dy +
_
t
0
_
x+c(ts)
xc(ts)
f(y, s) dy ds
_
.
Proof. Fix (x, t) R R
+
. Consider the open triangle in R R
+
with
vertices (x, t), (x ct, 0) and (x + ct, 0), and denote it by T(x, t). Thus,
T(x, t) := (y, s) R R
+
[ [y x[ < c(t s).
The boundary of the triangle T(x, t) consists of three parts
T
0
:= (y, 0) [ x ct < y < x + ct,
T
+
:= (y, s) R (0, t) [ cs + y = x + ct
and
T

:= (y, s) R (0, t) [ cs y = x +ct.


CHAPTER 4. WAVE EQUATION 90
The unit outward normal at each point of the boundary T(x, t) is given by
= (
1
,
2
) dened as
(y, s) =
_

_
(0, 1) (y, s) T
0
1

1+c
2
(1, c) (y, s) T
+
1

1+c
2
(1, c) (y, s) T

.
Therefore, by Gauss divergence theorem (cf. (D.0.1))
_
T(x,t)
f((y, s) dy ds =
_
T(x,t)
_
u
tt
(y, s) c
2
u
xx
(y, s)

dy ds
=
_
T(x,t)
_
u
t

2
c
2
u
x

d
=
_
T
0
_
u
t

2
c
2
u
x

d +
_
T
+
_
u
t

2
c
2
u
x

d
+
_
T

_
u
t

2
c
2
u
x

d
=
_
x+ct
xct
u
t
(y, 0) dy +
c

1 + c
2
_
T
+
[u
t
cu
x
] d
+
c

1 + c
2
_
T

[u
t
+ cu
x
] d
Note that the second and third integral are just the direction derivatives of
u along the direction (1, c) and (1, c) in the line T
+
and T

, respectively.
Therefore, we have
_
T(x,t)
f((y, s) dy ds =
_
x+ct
xct
h(y) dy + cu(x, t) cu(x + ct, 0)
+c[u(x, t) u(x ct, 0)]
u(x, t) =
1
2
[g(x + ct) + g(x ct)]
+
1
2c
_
_
x+ct
xct
h(y) dy +
_
t
0
_
x+c(ts)
xc(ts)
f(y, s) dy ds
_
.
CHAPTER 4. WAVE EQUATION 91
Aliter. We introduce a new function v dened as v(x, t) = u
t
(x, t) and rewrite
(4.1.6) as
U

(x, t) + AU(x, t) = F(x, t)


where U = (u, v), F = (0, f) and
A =
_
0 1
c
2
2
x
2
0
_
with the initial condition G(x) := U(x, 0) = (g(x), h(x)). The solution
U(x, t) is given as (cf. Appendix G)
U(x, t) = S(t)G(x) +
_
t
0
S(t s)F(s) ds
where S(t) is a solution operator of the homoegeneous system of rst order
PDE. Therefore, by dAlemberts formula,
S(t)(g, h) =
_
1
2
[g(x +ct) + g(x ct)] +
1
2c
_
x+ct
xct
h(y) dy
c
2
[g

(x + ct) g

(x ct)] +
1
2
[h(x + ct) + h(x ct)]
_
and, hence, u(x, t) has the required represenation.
4.1.4 Standing Waves: Separation of Variable
Recall the set-up of the vibrating string given by the equation u
tt
= u
xx
,
we have normalised the constant c. Initially at time t, let us say the string
has the shape of the graph of v, i.e., u(x, 0) = v(x). The snapshot of the
vibrating string at each time are called the standing waves. The shape of
the string at time t
0
can be thought of as some factor (depending on time)
of v. This observation motivates the idea of separation of variable, i.e.,
u(x, t) = v(x)w(t), where w(t) is the factor depending on time, which scales
v at time t to t with the shape of u(x, t).
The fact that endpoints are xed is given by the boundary condition
u(0, t) = u(L, t) = 0.
We are also given the initial position u(x, 0) = g(x) (at time t = 0) and initial
velocity of the string at time t = 0, u
t
(x, 0) = h(x). Given g, h : [0, L] R
such that g(0) = g(L) = 0 and h(0) = h(L), we need to solve the initial value
problem (4.1.5) with = 0.
CHAPTER 4. WAVE EQUATION 92
Let us seek for solutions u(x, t) whose variables can be separated. Let
u(x, t) = v(x)w(t). Dierentiating and substituting in the wave equation, we
get
v(x)w

(t) = c
2
v

(x)w(t)
Hence
w

(t)
c
2
w(t)
=
v

(x)
v(x)
.
Since RHS is a function of x and LHS is a function t, they must equal a
constant, say . Thus,
v

(x)
v(x)
=
w

(t)
c
2
w(t)
= .
Using the boundary condition u(0, t) = u(L, t) = 0, we get
v(0)w(t) = v(L)w(t) = 0.
If w 0, then u 0 and this cannot be a solution to (4.1.5). Hence, w , 0
and v(0) = v(L) = 0. Thus, we need to solve the eigen value problem for the
second order dierential operator.
_
v

(x) = v(x), x (0, L)


v(0) = v(L) = 0,
Note that the can be either zero, positive or negative. If = 0, then
v

= 0 and the general solution is v(x) = x + , for some constants and


. Since v(0) = 0, we get = 0, and v(L) = 0 and L ,= 0 implies that = 0.
Thus, v 0 and hence u 0. But, this cannot be a solution to (4.1.5).
If > 0, then v(x) = e

x
+ e

x
. Equivalently,
v(x) = c
1
cosh(

x) + c
2
sinh(

x)
such that = (c
1
+c
2
)/2 and = (c
1
c
2
)/2. Using the boundary condition
v(0) = 0, we get c
1
= 0 and hence
v(x) = c
2
sinh(

x).
Now using v(L) = 0, we have c
2
sinh

L = 0. Thus, c
2
= 0 and v(x) = 0.
We have seen this cannot be a solution.
CHAPTER 4. WAVE EQUATION 93
Finally, if < 0, then set =

. We need to solve the simple


harmonic oscillator problem
_
v

(x) +
2
v(x) = 0 x (0, L)
v(0) = v(L) = 0.
The general solution is
v(x) = cos(x) + sin(x).
Using v(0) = 0, we get = 0 and hence v(x) = sin(x). Now using
v(L) = 0, we have sin L = 0. Thus, either = 0 or sin L = 0. But
= 0 does not yield a solution. Hence L = k or = k/L, for all non-
zero k Z. Since > 0, we can consider only k N. Hence, for each k N,
there is a solution (v
k
,
k
) for the eigen value problem with
v
k
(x) =
k
sin
_
kx
L
_
,
for some constant b
k
and
k
= (k/L)
2
. It now remains to solve w for each
of these
k
. For each k N, we solve for w
k
in the ODE
w

k
(t) + (ck/L)
2
w
k
(t) = 0.
The general solution is
w
k
(t) = a
k
cos
_
ckt
L
_
+ b
k
sin
_
ckt
L
_
.
For each k N, we have
u
k
(x, t) =
_
a
k
cos
_
ckt
L
_
+ b
k
sin
_
ckt
L
__
sin
_
kx
L
_
for some constants a
k
and b
k
. The situation corresponding to k = 1 is called
the fundamental mode and the frequency of the fundamental mode is
1
2
c
L
=
c
2L
=
_
T/
2L
.
The frequency of higher modes are integer multiples of the fundamental fre-
quency.
CHAPTER 4. WAVE EQUATION 94
The general solution of (4.1.5), by principle of superposition, is
u(x, t) =

k=1
_
a
k
cos
_
ckt
L
_
+ b
k
sin
_
ckt
L
__
sin
_
kx
L
_
.
Note that the solution is expressed as series, which raises the question of
convergence of the series. Another concern is whether all solutions of (4.1.5)
have this form. We ignore these two concerns at this moment.
Since we know the initial position of the string as the graph of g, we get
g(x) = u(x, 0) =

k=1
a
k
sin
_
kx
L
_
.
This expression is again troubling and rises the question: Can any arbitrary
function g be expressed as an innite sum of trigonometric functions? An-
swering this question led to the study of Fourier series. Let us also, as
usual, ignore this concern for time being. Then, can we nd the the con-
stants a
k
with knowledge of g. By multiplying sin
_
lx
L
_
both sides of the
expression of g and integrating from 0 to L, we get
_
L
0
g(x) sin
_
lx
L
_
dx =
_
L
0
_

k=1
a
k
sin
_
kx
L
_
_
sin
_
lx
L
_
dx
=

k=1
a
k
_
L
0
sin
_
kx
L
_
sin
_
lx
L
_
dx
Therefore, the constants a
k
are given as
a
k
=
2
L
_
L
0
g(x) sin
_
kx
L
_
.
Finally, by dierentiating u w.r.t t, we get
u
t
(x, t) =

k=1
ck
L
_
b
k
cos
ckt
L
a
k
sin
ckt
L
_
sin
_
kx
L
_
.
Employing similar arguments and using u
t
(x, 0) = h(x), we get
h(x) = u
t
(x, 0) =

k=1
b
k
kc
L
sin
_
kx
L
_
and hence
b
k
=
2
kc
_
L
0
h(x) sin
_
kx
L
_
.
CHAPTER 4. WAVE EQUATION 95
4.2 Higher Dimensions
We shall denote as the Laplacian w.r.t. the space variable and the wave
equation is denoted as u := u
tt
c
2
u. The Cauchy initial value problem
in higher dimensions is
_
_
_
u = f(x, t) in R
n
(0, )
u(x, 0) = g(x) in R
n
u
t
(x, 0) = h(x) in R
n
.
(4.2.1)
Due to the linearity of the wave operator , any solution u = v+w+z where
v, w and z are, respectively, solutions of
_
_
_
v = 0 in R
n
(0, )
v(x, 0) = g(x) in R
n
v
t
(x, 0) = 0 in R
n
,
(4.2.2)
_
_
_
w = 0 in R
n
(0, )
w(x, 0) = 0 in R
n
w
t
(x, 0) = h(x) in R
n
(4.2.3)
and
_
_
_
z = f(x, t) in R
n
(0, )
z(x, 0) = 0 in R
n
z
t
(x, 0) = 0 in R
n
.
(4.2.4)
Theorem 4.2.1 (Duhamels Principle). Let w
h
be a solution to (4.2.3). Then
v(x, t) = w
g
t
(x, t)
and
z(x, t) =
_
t
0
w
fs
(x, t s) ds,
where f
s
= f(, s), are solutions to (4.2.2) and (4.2.4).
Proof. Since w
g
solves (4.2.3) with h = g, we have
v = w
g
t
(x, t) =

t
(w
g
) = 0
on R
n
(0, ). Further, v(x, 0) = w
g
t
(x, 0) = g(x) in R
n
and v
t
(x, 0) =
w
g
tt
(x, 0) = c
2
w
g
(x, 0) = 0. Thus, v solves (4.2.2). Now, let us denote
CHAPTER 4. WAVE EQUATION 96
w(x, t) = w
fs
(x, ts), for 0 < s < t, which satises (4.2.3) with h() = f(c,s).
Then,
z
t
(x, t) = w
fs
(x, 0) +
_
t
0
w
fs
t
(x, t s) ds =
_
t
0
w
fs
t
(x, t s) ds
and
z
tt
(x, t) = w
fs
t
(x, 0) +
_
t
0
w
fs
tt
(x, t s) ds
= f(x, t) + c
2

_
t
0
w
fs
(x, t s) ds
= f(x, t) + c
2
z.
Therefore, z solves (4.2.4).
The Duhamels principle can be viewed as a generalization of the method
of variations of constants in ODE (cf. Appendix G). Owing to the above
theorem it is enough solve for w in (4.2.3). To do so, we shall employ the
method of spherical means which reduces such problem to one dimensional
framework.
4.2.1 Spherical Means
More generally, in this section we solve for v +w which is the solution to the
wave equation
_
_
_
u = 0 in R
n
(0, )
u(x, 0) = g(x) in R
n
u
t
(x, 0) = h(x) in R
n
.
(4.2.5)
For a xed x R
n
and t (0, ), the spherical mean of a C
2
function
u(x, t) is given as
M(x; r, t) :=
1

n
r
n1
_
Sr(x)
u(y, t) d
y
where
n
is the surface area of the unit sphere in R
n
. Equivalently, after
setting z = (y x)/r,
M(x; r, t) :=
1

n
_
S
1
(0)
u(x + rz, t) d
z
.
CHAPTER 4. WAVE EQUATION 97
If we take the above form of M(x; r, t) as a denition, we observe that it is
valid for all r R and M(x; 0, t) = u(x, t). Then
M
r
(x; r, t) =
1

n
_
S
1
(0)
n

i=1
u
y
i
(x +rz, t)z
i
d
z
=
1

n
r
n1
_
Sr(x)

y
u(y, t) z d
y
=
1

n
r
n1
_
Br(x)

y
u(y, t) dy.
If u(x, t) is a solution to (4.2.5) then
r
n1
M
r
(x; r, t) =
1
c
2

n
_
Br(x)
u
tt
(y, t) dy
=
1
c
2

n
_
r
0
_
Ss(x)
u
tt
(y, t) d
y
ds
d
dr
(r
n1
M
r
(x; r, t)) =
1
c
2

n
_
Sr(x)
u
tt
(y, t) d
y
=
r
n1
c
2

2
t
2
_
1

n
r
n1
_
Sr(x)

y
u(y, t) d
y
_
=
r
n1
c
2
M
tt
.
Thus, the spherical means satises the one space variable PDE
r
1n
d
dr
(r
n1
M
r
) = c
2
M
tt
or
M
rr
+
n 1
r
M
r
= c
2
M
tt
,
which is called the Euler-Poisson-Darboux equation. Also, if u(x, t) is a
solution to (4.2.5) then using the initial condition, we get
G(x; r) := M(x; r, 0) =
1

n
_
S
1
(0)
g(x + rz) d
z
and
H(x; r) := M
t
(x; r, 0) =
1

n
_
S
1
(0)
h(x + rz) d
z
.
In the following section we shall illustrate solving the Euler-Poisson-
Darboux equation in three and two dimensions. For higher dimensions solv-
ing E-P-D equation is tedious.
CHAPTER 4. WAVE EQUATION 98
Theorem 4.2.2 (Three dimensions). Given g C
3
(R
3
) and h C
2
(R
3
)
there exists a unique solution u C
2
(R
3
[0, ) of (4.2.5) given by the
Poissons formula
u(x, t) =
1
4c
2
_

t
_
1
t
_
Sct(x)
g(y) d
y
_
+
1
t
_
Sct(x)
h(y) d
y
_
.
Proof. For the Euclidean dimension three, the Euler-Poisson-Darboux equa-
tion becomes
c
2
(rM)
rr
= (rM)
tt
.
Thus, rM is a solution to the one dimensional Cauchy problem
_
_
_
c
2
(rM)
rr
(x; r, t) = (rM)
tt
(x; r, t) in (0, ) (0, )
rM(x; r, 0) = rG(x; r) in (0, )
rM
t
(x; r, 0) = rH(x; r) in (0, )
and is given by the dAlembert formula
M(x; r, t) =
1
2r
((r ct)G(r ct) + (r + ct)G(r + ct))+
1
2rc
_
r+ct
rct
yH(y) dy,
as long as, the domain of dependence [r ct, r +ct] is in (0, ). Otherwise,
we extend G and H to entire R as

G(x; r) =
_

_
G(x; r) r > 0
g(x) r = 0
G(x; r) r < 0
and

H(x; r) =
_

_
H(x; r) r > 0
h(x) r = 0
H(x; r) r < 0.
Invoking Lemma 4.2.3, we get

M(x; r, t) as
1
2r
_
(r ct)

G(r ct) + (r + ct)

G(r + ct)
_
+
1
2rc
_
r+ct
rct
y

H(y) dy
Since

G and

H are even functions, we have
_
(rct)
rct
yH(y) dy = 0
CHAPTER 4. WAVE EQUATION 99
and, therefore

M(x; r, t)
1
2r
_
(r + ct)

G(ct + r) (ct r)

G(ct r)
_
+
1
2rc
_
r+ct
ctr
y

H(y) dy.
Since
u(x, t) = lim
r0
+

M(x; r, t),
by LHospitals rule, we get
u(x, t) =
d
dt
(tG(ct)) + tH(ct) = ctG

(ct) + G(ct) + tH(ct).


Lemma 4.2.3. r

G, r

H C
2
(R).
Proof. It follows from the denition that
lim
r0+
G(x; r) = g(x) and lim
r0+
H(x; r) = h(x).
Thus, r

G, r

H C(R). Since G, H C
1
(R), r

G, r

H C
1
(R). Further, since
G

and H

are bounded as r 0+, r



G, r

H C
2
(R).
The domain of dependence of (x, t
0
) for the three dimensional wave equa-
tion is the boundary of the three dimensional sphere with radius ct.
The Hadamards method of descent is the technique of nding a solu-
tion of the two dimensional wave equation using the three dimensional wave
equation.
Theorem 4.2.4 (Method of Descent). Given g C
3
(R
2
) and h C
2
(R
2
)
there exists a unique solution u C
2
(R
2
[0, ) of (4.2.5) given by the
Poissons formula
1
2c
_

t
_
_
Bct(x,y)
g(, )
_
c
2
t
2

2
d d
_
+
_
Bct(x,y)
h(, )
c
2
t
2

2
d d
_
where
=
_
( x)
2
+ ( y)
2
.
CHAPTER 4. WAVE EQUATION 100
Proof. Let v be a solution of (4.2.5) in two dimensions with g C
3
(R
2
) and
h C
2
(R
2
). Then
u(x, y, z, t) := v(x, y, t)
is solution to (4.2.5) in three dimensions where g and h are given independent
of z. Since u(x, y, z, t) = u(x, y, 0, t)+zu
z
(x, y, z, t) for 0 < < 1 and u
z
= 0,
we have v(x, y, t) = u(x, y, 0, t). Therefore, using the poisson formula in three
dimensions, we get
v(x, y, t) =
1
4c
2
_

t
_
1
t
_
Sct(x,y,0)
g(, , ) d
_
+
1
t
_
Sct(x,y,0)
h(, , ) d
_
.
Recall that g and h are independent of , therefore g(, , ) = g(, ) and
h(, , ) = h(, ). The sphere has the equation (x)
2
+(y)
2
+
2
= c
2
t
2
.
The exterior normal is given as
=
S
[S[
=
_
x
ct
,
y
ct
,

ct
_
and the surface element d is given as d = dd
1
|
3
|
where

3
=
_
c
2
t
2
( x)
2
( y)
2
ct
with the positive sign applying when > 0 and negative sign applying when
< 0. Set
=
_
( x)
2
+ ( y)
2
.
In the two dimensions, the domain of dependence is the entire disk
B
ct
0
(x
0
, y
0
) in contrast to three dimensions which had only the boundary
of the sphere as domain of dependence.
4.2.2 Odd Dimension
One can copy the idea of three dimension to any odd dimension, if we rewrite
the Euler-Poisson-Darboux equation in approriate form.
CHAPTER 4. WAVE EQUATION 101
Exercise 19. If n is odd, show that the correct form of M that satises the
one dimensional wave equation is
_
1
r

r
_n3
2
(r
n2
M(r, t)).
For instance, when n = 5, r
2
M
r
+ 3rM satises the one dimensional wave
equation.
We have already noted that the solution at a given point is determined by
the value of intial data in a subset of the initial hypersurface. Consequently,
changing initial value outside the domain of dependence does not change
values of solutions.
Also, it takes time for the initial data to make inuence. Suppose g and
h have their support in B
r
(x
0
). Then the support of u(, t) is contained in

yBr(x
0
)
B
t
(y) = B
r+ct
(x
0
). The support of u spreads at a nite speed and
is called the nite speed propagation.
4.2.3 Inhomogeneous Wave equation
We have already derived in Theorem 4.2.1 the formula for inhomogeneous
equation (4.2.4).
Theorem 4.2.5. For any f C
1
, the solution u(x, t) of (4.2.4) is given as
u(x, t) =
_

_
1
2c
_
t
0
_
_
x+c(ts)
xc(ts)
f(y, s)dy
_
ds n = 1
1
4c
_
t
0
_
_
B
c(ts)
(x)
f(y,s)

c
2
(ts)
2
r
2
dy
_
ds n = 2
1
4c
2
_
Bct(x)
f(y,tr/c)
r
dy n = 3.
Proof. The Poissons formula corresponding to the three dimension case gives
the formula for
w
fs
(x, t s) =
1
4c
2
(t s)
_
S
c(ts)
(x)
f(y, s) d
y
.
CHAPTER 4. WAVE EQUATION 102
and
u(x, t) =
_
t
0
w
fs
(x, t s) ds =
1
4c
2
_
t
0
_
S
c(ts)
(x)
f(y, s)
t s
d
y
ds
=
1
4c
2
_
ct
0
_
S (x)
f(y, t /c)

d
y
d [setting = c(t s)]
=
1
4c
2
_
Bct(x)
f(y, t r/c)
r
dy
where r = [x y[. Similarly, one can derive the formulae for one and two
dimensions.
Note that in the three dimensional case the integrand is not taken at time
t, but at an earlier time t r/c. Thus, the integrand in this case is called
retarded potential.
Example 4.1. Consider the wave equation
_
_
_
u
tt
(x, t) c
2
u
xx
(x, t) = sin 3x in (0, ) (0, )
u(0, t) = u(, t) = 0 in (0, )
u(x, 0) = u
t
(x, 0) = 0 in (0, ).
We look for the solution of the homogeneous wave equation
_

_
w
tt
(x, t) c
2
w
xx
(x, t) = 0 in (0, ) (0, )
w(0, t) = w(, t) = 0 in (0, )
w(x, 0) = 0 in (0, )
w
t
(x, 0) = sin 3x in (0, ).
By separation of variable technique, we know that the general solution of w
is
w(x, t) =

k=1
[a
k
cos(kct) + b
k
sin(kct)] sin(kx)
and
w(x, 0) =

k=1
a
k
sin(kx) = 0.
Thus, a
k
= 0, for all k N. Also,
w
t
(x, 0) =

k=1
b
k
ck sin(kx) = sin 3x.
CHAPTER 4. WAVE EQUATION 103
Hence, b
k
s are all zeroes, except for k = 3 and b
3
= 1/3c. Thus,
w(x, t) =
1
3c
sin(3ct) sin(3x)
and
u(x, t) =
_
t
0
w(x, t s) ds =
1
3c
_
t
0
sin(3c(t s)) sin 3xds
=
sin 3x
3c
_
t
0
sin(3c(t s)) ds =
sin 3x
3c
cos(3c(t s))
3c

t
0
=
sin 3x
9c
2
(1 cos 3ct) .
4.3 Eigenvalue Problem of Laplacian
The separation of variable technique can be used for studying wave equation
on 2D Rectangle and 2D Disk etc. This leads to studying the eigen value
problem of the Laplacian. For a given open bounded subset R
2
, the
Dirichlet eigenvalue problem,
_
u(x, y) = u(x, y) (x, y)
u(x, y) = 0 (x, y) .
Note that, for all R, zero is a trivial solution of the Laplacian. Thus,
we are interested in non-zero s for which the Laplacian has non-trivial
solutions. Such an is called the eigenvalue and corresponding solution u

is called the eigen function.


Note that if u

is an eigen function corresponding to , then u

, for all
R, is also an eigen function corresponding to . Let W be the real vector
space of all u : R continuous (smooth, as required) functions such that
u(x, y) = 0 on . For each eigenvalue of the Laplacian, we dene the
subspace of W as
W

= u W [ u solves Dirichlet EVP for given .


Theorem 4.3.1. There exists an increasing sequence of positive numbers
0 <
1
<
2
<
3
< . . . <
n
< . . . with
n
which are eigenvalues of
the Laplacian and W
n
= W
n
is nite dimensional. Conversely, any solution
u of the Laplacian is in W
n
, for some n.
CHAPTER 4. WAVE EQUATION 104
Though the above theorem assures the existence of eigenvalues for Lapla-
cian, it is usually dicult to compute them for a given . In this course, we
shall compute the eigenvalues when is a 2D-rectangle and a 2D-disk.
4.3.1 In Rectangle
Let the rectangle be = (x, y) R
2
[ 0 < x < a, 0 < y < b. We wish to
solve the Dirichlet EVP in the rectangle
_
u(x, y) = u(x, y) (x, y)
u(x, y) = 0 (x, y) .
The boundary condition amounts to saying
u(x, 0) = u(a, y) = u(x, b) = u(0, y) = 0.
We look for solutions of the form u(x, y) = v(x)w(y) (variable separated).
Substituting u in separated form in the equation, we get
v

(x)w(y) v(x)w

(y) = v(x)w(y).
Hence

(x)
v(x)
= +
w

(y)
w(y)
.
Since LHS is function of x and RHS is function y and are equal they must
be some constant, say . We need to solve the EVPs
v

(x) = v(x) and w

(y) = ( )w(y)
under the boundary conditions v(0) = v(a) = 0 and w(0) = w(b) = 0.
As seen before, while solving for v, we have trivial solutions for 0. If
> 0, then v(x) = c
1
cos(

x)+c
2
sin(

x). Using the boundary condition


v(0) = 0, we get c
1
= 0. Now using v(a) = 0, we have c
2
sin

a = 0. Thus,
either c
2
= 0 or sin

a = 0. We have non-trivial solution, if c


2
,= 0,
then

a = k or

= k/a, for k Z. For each k N, we have
v
k
(x) = sin(kx/a) and
k
= (k/a)
2
. We solve for w for each
k
. For each
k, l N, we have w
kl
(y) = sin(ly/b) and
kl
= (k/a)
2
+ (l/b)
2
. For each
k, l N, we have
u
kl
(x, y) = sin(kx/a) sin(ly/b)
and
kl
= (k/a)
2
+ (l/b)
2
.
CHAPTER 4. WAVE EQUATION 105
4.3.2 In Disk
Let the disk of radius a be = (x, y) R
2
[ x
2
+ y
2
< a
2
. We wish to
solve the Dirichlet EVP in the disk
_
_
_
1
r

r
_
r
u
r
_

1
r
2

2
u

2
= u(r, ) (r, )
u() = u( + 2) R
u(a, ) = 0 R.
We look for solutions of the form u(r, ) = v(r)w() (variable separated).
Substituting u in separated form in the equation, we get

w
r
d
dr
_
r
dv
dr
_

v
r
2
w

() = v(r)w().
Hence dividing by vw and multiplying by r
2
, we get

r
v
d
dr
_
r
dv
dr
_

1
w
w

() = r
2
.
r
v
d
dr
_
r
dv
dr
_
+ r
2
=
1
w
w

() = .
Solving for non-trivial w, using the periodicity of w, we get for
0
= 0,
w
0
() =
a
0
2
and for each k N,
k
= k
2
and
w
k
() = a
k
cos k + b
k
sin k.
For each k N 0, we have the equation,
r
d
dr
_
r
dv
dr
_
+ (r
2
k
2
)v = 0.
Introduce change of variable x =

r and x
2
= r
2
. Then
r
d
dr
= x
d
dx
.
rewriting the equation in new variable y(x)) = v(r)
x
d
dx
_
x
dy(x)
dx
_
+ (x
2
k
2
)y(x) = 0.
CHAPTER 4. WAVE EQUATION 106
Note that this none other than the Bessels equation. We already know that
for each k N 0, we have the Bessels function J
k
as a solution to the
Bessels equation. Recall the boundary condition on v, v(a) = 0. Thus,
y(

a) = 0. Hence

a should be a zero of the Bessels function.


For each k N0, let z
kl
be the l-th zero of J
k
, l N. Hence

a = z
kl
and so
kl
= z
2
kl
/a
2
and y(x) = J
k
(x). Therefore, v(r) = J
k
(z
kl
r/a). For each
k N 0 and l N, we have
u
kl
(r, ) = J
k
(z
kl
r/a) sin(k) or J
k
(z
kl
r/a) cos(k)
and
kl
= z
2
kl
/a
2
.
Chapter 5
Heat Equation
5.1 Derivation
Let a homogeneous material occupy a region R
n
with C
1
boundary. Let
k denote the thermal conductivity (dimensionless quantity) and c be the heat
capacity of the material. Let u(x, t) be a function plotting the temperature
of the material at x in time t. The thermal energy stored at x , at
time t, is cu(x, t). If v(x, t) denotes the velocity of (x, t), by Fourier law, the
thermal energy changes following the gradients of temperature, i.e.,
cu(x, t)v(x, t) = ku.
The thermal energy is the quantity that is conserved (conservation law) and
satises the continuity equation (1.4.1). Thus, we have
u
t

k
c
u = 0.
If the material occupying the region is non-homogeneous, anisotropic,
the temperature gradient may generate heat in preferred directions, which
themselves may depend on x . Thus, the conductivity of such a material
at x , at time t, is given by a nn matrix K(x, t) = (k
ij
(x, t)). Thus, in
this case, the heat equation becomes,
u
t
div
_
1
c
Ku
_
= 0.
The heat equation is an example of a second order equation in divergence
form. The heat equation gives the temperature distribution u(x, t) of the
107
CHAPTER 5. HEAT EQUATION 108
material with conductivity k and capacity c. In general, we may choose
k/c = 1, since, for any k and c, we may rescale our time scale t (k/c)t.
The Cauchy initial value problem (IVP) of the heat equation,
_
u
t
(x, t)
x
u(x, t) = 0 in R
n
(0, )
u(x, 0) = g(x) in R
n
,
(5.1.1)
where g C(). We end this section with a remark that under the transfor-
mation t t, in contrast to the wave equation, the heat equation changes
to a background equation. This signies that the heat equation describes
irreversible process, i.e., it is not possible to nd the distribution of temper-
ature at an earlier time t < t
0
, if the distribution is given at t
0
.
5.2 Boundary Conditions
To make the heat equation u
t
u = f in (0, T) well-posed, where is
a bounded open subset of R
n
, we choose to specify the boundary condition
u(x, 0) = g(x) on t = 0 and one of the following conditions on
(0, T):
(i) (Dirichlet condition) u(x, t) = h(x, t);
(ii) (Neumann condition)
x
u(x, t) (x) = h(x, t), where nu(x) is the unit
outward normal of (x, t) (0, T);
(iii) (Robin condition)
x
u(x, t) + cu(x, t) = h(x, t) for any c > 0.
(iv) (Mixed condition) u(x, t) = h(x, t) on
1
and
x
u(x, t) = h(x, t) on

2
, where
1

2
= (0, T) and
1

2
= .
5.3 Heat Flow on a Bar
The equation governing the heat propagation in a bar of length L is
u
t
=
1
(x)(x)

x
_
(x)
u
x
_
where (x) is the specic heat at x, (x) is density of bar at x and (x) is
the thermal conductivity of the bar at x. If the bar is homogeneous, i.e, its
CHAPTER 5. HEAT EQUATION 109
properties are same at every point, then
u
t
=

2
u
x
2
with , , being constants.
Let L be the length of a homogeneous rod insulated along sides insulated
along sides and its ends are kept at zero temperature. Then the temperature
u(x, t) at every point of the rod, 0 x L and time t 0 is given by the
equation
u
t
= c
2

2
u
x
2
where c is a constant.
The temperature zero at the end points is given by the Dirichlet boundary
condition
u(0, t) = u(L, t) = 0.
Also, given is the initial temperature of the rod at time t = 0, u(x, 0) = g(x),
where g is given (or known) such that g(0) = g(L) = 0. Given g : [0, L] R
such that g(0) = g(L) = 0, we look for all the solutions of the Dirichlet
problem
_
_
_
u
t
(x, t) c
2
u
xx
(x, t) = 0 in (0, L) (0, )
u(0, t) = u(L, t) = 0 in (0, )
u(x, 0) = g(x) on [0, L].
We look for u(x, t) = v(x)w(t) (variable separated). Substituting u in sepa-
rated form in the equation, we get
v(x)w

(t) = c
2
v

(x)w(t)
w

(t)
c
2
w(t)
=
v

(x)
v(x)
.
Since LHS is function of t and RHS is function x and are equal they must be
some constant, say . Thus,
w

(t)
c
2
w(t)
=
v

(x)
v(x)
= .
Thus we need to solve two ODE to get v and w,
w

(t) = c
2
w(t)
CHAPTER 5. HEAT EQUATION 110
and
v

(x) = v(x).
But we already know how to solve the eigenvalue problem involving v. For
each k N, we have the pair (
k
, v
k
) as solutions to the EVP involving v,
where
k
= (k)
2
/L
2
and v
k
(x) = sin
_
kx
L
_
some constants b
k
. For each
k N, we solve for w
k
to get
ln w
k
(t) =
k
c
2
t + ln
where is integration constant. Thus, w
k
(t) = e
(kc/L)
2
t
. Hence,
u
k
(x, t) = v
k
(x)w
k
(t) =
k
sin
_
kx
L
_
e
(kc/L)
2
t
,
for some constants
k
, is a solution to the heat equation. By superposition
principle, the general solution is
u(x, t) =

k=1
u
k
(x, t) =

k=1

k
sin
_
kx
L
_
e
(kc/L)
2
t
.
We now use the initial temperature of the rod, given as g : [0, L] R to
nd the particular solution of the heat equation. We are given u(x, 0) = g(x).
Thus,
g(x) = u(x, 0) =

k=1

k
sin
_
kx
L
_
Since g(0) = g(L) = 0, we know that g admits a Fourier Sine expansion and
hence its coecients
k
are given as

k
=
2
L
_
L
0
g(x) sin
_
kx
L
_
.
5.4 On a Circular Wire
We intend solve the heat equation in a circle (circular wire) of radius one
which is insulated along its sides. Then the temperature u(, t) at every
point of the circle, R and time t 0 is given by the equation
u
t
= c
2

2
u

2
CHAPTER 5. HEAT EQUATION 111
where c is a constant. We note that now u(, t) is 2-periodic in the variable
. Thus,
u( + 2, t) = u(, t) R, t 0.
Let the initial temperature of the wire at time t = 0, be u(, 0) = g(), where
g is a given 2-periodic function. Given a 2-periodic function g : R R,
we look for all solutions of
_
_
_
u
t
(, t) c
2
u

(, t) = 0 in R (0, )
u( + 2, t) = u(, t) in R (0, )
u(, 0) = g() on R t = 0.
We look for u(, t) = v()w(t) with variables separated. Substituting for u
in the equation, we get
w

(t)
c
2
w(t)
=
v

()
v()
= .
For each k N 0, the pair (
k
, v
k
) is a solution to the EVP where

k
= k
2
and
v
k
() = a
k
cos(k) + b
k
sin(k).
For each k N 0, we get w
k
(t) = e
(kc)
2
t
. For k = 0
u
0
(, t) = a
0
/2 (To maintain consistency with Fourier series)
and for each k N, we have
u
k
(, t) = [a
k
cos(k) + b
k
sin(k)] e
k
2
c
2
t
Therefore, the general solution is
u(, t) =
a
0
2
+

k=1
[a
k
cos(k) + b
k
sin(k)] e
k
2
c
2
t
.
We now use the initial temperature on the circle to nd the particular solu-
tion. We are given u(, 0) = g(). Thus,
g() = u(, 0) =
a
0
2
+

k=1
[a
k
cos(k) + b
k
sin(k)]
Since g is 2-periodic it admits a Fourier series expansion and hence
a
k
=
1

g() cos(k) d,
CHAPTER 5. HEAT EQUATION 112
b
k
=
1

g() sin(k) d.
Note that as t the temperature of the wire approaches a constant a
0
/2.
Exercise 20. Solve the heat equation for 2D Rectangle and 2D Disk
5.5 Inhomogeneous Equation
In this section we solve the inhomogeneous heat equation, using Duhamels
principle. The Duhamels principle states that one can obtain a solution of
the inhomogeneous IVP for heat from its homogeneous IVP.
For a given f, let u(x, t) be the solution of the inhomogeneous heat equa-
tion,
_
_
_
u
t
(x, t) c
2
u(x, t) = f(x, t) in (0, )
u(x, t) = 0 in (0, )
u(x, 0) = 0 in .
As a rst step, for each s (0, ), consider w(x, t; s) as the solution of the
homogeneous problem (auxiliary)
_
_
_
w
s
t
(x, t) c
2
w
s
(x, t) = 0 in (s, )
w
s
(x, t) = 0 in (s, )
w
s
(x, s) = f(x, s) on s.
Since t (s, ), introducing a change of variable r = ts, we have w
s
(x, t) =
w(x, t s) which solves
_
_
_
w
t
(x, r) c
2
w(x, r) = 0 in (0, )
w(x, r) = 0 in (0, )
w(x, 0) = f(x, s) on .
Duhamels principle states that
u(x, t) =
_
t
0
w
s
(x, t) ds =
_
t
0
w(x, t s) ds
CHAPTER 5. HEAT EQUATION 113
and u solves the inhomogenous heat equation. Suppose w is C
2
, we get
u
t
(x, t) =

t
_
t
0
w(x, t s) ds
=
_
t
0
w
t
(x, t s) ds + w(x, t t)
d(t)
dt
w(x, t 0)
d(0)
dt
=
_
t
0
w
t
(x, t s) ds + w(x, 0)
=
_
t
0
w
t
(x, t s) ds + f(x, t).
Similarly,
u(x, t) =
_
t
0
w(x, t s) ds.
Thus,
u
t
c
2
u = f(x, t) +
_
t
0
_
w
t
(x, t s) c
2
w(x, t s)
_
ds
= f(x, t).
5.6 Steady State Equation
Consider the Cauchy problem with inhomogeneous Dirichlet boundary con-
ditions
_
_
_
u
t
(x, t) c
2
u(x, t) = 0 in (0, )
u(x, t) = = (x, t) in [0, )
u(x, 0) = g(x) on
such that, for all x ,
g(x) = (x, 0), g

(x) =

(x, 0).
The steady-state solution of the heat equation is dened as
v(x) = limt u(x, t).
Note that v satises the equation v = 0, since v
t
= 0. Further v satises
the boundary condition on as
v(x) = limt (x, t).
CHAPTER 5. HEAT EQUATION 114
5.7 Fundamental Solution of Heat Equation
We shall now derive the fundamental solution of the heat equation (5.1.1).
Taking Fourier transform both sides of the equation, we get

u
t
u =

f
u
t
(, t)

u(, t) =

f(, t)
u
t
(, t)
n

j=1
i
2

2
j
u(, t) =

f(, t)
u
t
(, t) +[[
2
u(, t) =

f(, t).
The solution of the above ODE is given by (cf. Appendix G)
u(, t) = g()e
||
2
t
+
_
t
0
e
||
2
(ts)

f(, s) ds
where we have used the initial condition u(, 0) = g() of the ODE. Therefore,
by inverse Fourier formula,
u(x, t) = (2)
n/2
_
R
n
_
g()e
||
2
t
e
ix
+
_
t
0
e
ix||
2
(ts)

f(, s) ds
_
d
= (2)
n
_
R
n
__
R
n
g(y)e
iy
dy
_
e
||
2
t
e
ix
d
+(2)
n
_
R
n
_
t
0
__
R
n
f(y, s)e
iy
dy
_
e
ix||
2
(ts)
ds d
= (2)
n
_
R
n
g(y)
__
R
n
e
i(xy)||
2
t
d
_
dy
+(2)
n
_
R
n
_
t
0
f(y, s)
__
R
n
e
i(xy)||
2
(ts)
d
_
ds dy
=
_
R
n
g(y)K(x, y, t) dy +
_
R
n
_
t
0
f(y, s)K(x, y, t s) ds dy
where
K(x, y, t) = (2)
n
_
R
n
e
i(xy)||
2
t
d.
Note that
i (x y) [[
2
t =
_

t i
(x y)
2

t
_

t i
(x y)
2

t
_

[x y[
2
4t
CHAPTER 5. HEAT EQUATION 115
and, set =
_

t i
(xy)
2

t
_
. Therefore, d =

td. Using this substituion


in K and simplifying, we get
K(x, y, t) = (4t)
n/2
e
|xy|
2
/4t
called the heat kernel or the fundamental solution of heat equation. The
function K can be motivated in another way. Note that if u(x, t) is a solu-
tion of the heat equation, then (u T

)(x, t) is also a solution of the heat


equation, where T

(x, t) = (x,
2
t) is a linear transformation for any ,= 0.
This scaling or dilation is called the parabolic scaling. Thus, we look for a
solution u(x, t) = v(t)w(r
2
/t), where r = [x[. Substituting this separation
of variable in the heat equation, we derive v(t) = t
n/2
and w(t) = e
r
2
/4t
.
This motivates us to dene the fundamental solution as
K(x, t) :=
_

1
4t
n/2
e
r
2
/4t
x R
n
, t > 0
0 x R
n
, t < 0.
CHAPTER 5. HEAT EQUATION 116
Chapter 6
The Laplacian
A general second order linear elliptic equation is of the form
n

i,j=1
a
ij
(x)u
x
i
x
j
+
n

i=1
b
i
(x)u
x
i
+ c(x)u(x) = f(x)
where A(x) = a
ij
(x) is real, symmetric and positive denite. If A is a
constant matrix then with a suitable transformation one can rewrite
n

i,j=1
a
ij
(x)u
x
i
x
j
= v
where v(x) := u(Tx). We introduced (cf. Chapter 1) Laplacian to be the
trace of the Hessain matrix, :=

n
i=1

2
x
2
i
. The Laplace operator usually
appears in physical models associated with dissipative eects (except wave
equation). The importance of Laplace operator can be realised by its ap-
pearance in various physical models. For instance, the heat equation

t
,
the wave equation

2
t
2
,
or the Schr odingers equation
i

t
+ .
The Laplacian is a linear operator, i.e., (u+v) = u+v and (u) =
u for any constant R.
117
CHAPTER 6. THE LAPLACIAN 118
6.1 Properties of Laplacian
The Laplace operator commutes with the translation and rotation opera-
tor. For any a R
n
, we dene the translation map T
a
: C() C() as
(T
a
u)(x) = u(x + a). The invariance of Laplacian under translation means
that T
a
= T
a
. For any u C
2
(), (T
a
u)
x
i
(x) = u
x
i
(x + a) and
(T
a
u)
x
i
x
i
(x) = u
x
i
x
i
(x + a). Thus, (T
a
u)(x) = u(x + a).
For any orthogonal n n matrix O (O
1
= O
t
), we dene R : C()
C() as Ru(x) = u(Ox). The invariance of Laplacian under rotation means
that R = R . Let y = Ox. Then, y
j
=

n
i=1
O
ji
x
i
and, by chain rule,
(Ru)
x
i
=
n

j=1
u
y
j
y
j
x
i
=
n

j=1
u
y
j
O
ji
.
Therefore,
x
Ru = O
t

y
u and
( R)u(x) =
x
Ru
x
(Ru) = O
t

y
u O
t

y
u = OO
t

y
u
y
u =
y
u.
But
y
u = (u)(Ox) = (R )u(x).
A radial function is constant on every sphere about the origin. Since
Laplacian commutes with rotations, it should map the class of all radial
functions to itself.
In cartesian coordiantes, the n-dimensional Laplacian is given as
:=
n

i=1

2
x
2
i
.
In polar coordinates (2 dimensions), the Laplacian is given as
:=
1
r

r
_
r

r
_
+
1
r
2

2
where r is the magnitude component (0 r < ) and is the direction
component (0 < 2). The direction component is also called the azimuth
angle or polar angle.
Exercise 21. Show that the two dimensional Laplacian has the representation
:=
1
r

r
_
r

r
_
+
1
r
2

2
in polar coordinates.
CHAPTER 6. THE LAPLACIAN 119
Proof. Using the fact that x = r cos and y = r sin , we have
x
r
= cos ,
y
r
= sin and
u
r
= cos
u
x
+ sin
u
y
.
Also,

2
u
r
2
= cos
2

2
u
x
2
+ sin
2

2
u
y
2
+ 2 cos sin

2
u
xy
.
Similarly,
x

= r sin ,
y

= r cos and
u

= r cos
u
y
r sin
u
x
.
Also,
1
r
2

2
u

2
= sin
2

2
u
x
2
+ cos
2

2
u
y
2
2 cos sin

2
u
xy

1
r
u
r
.
Therefore,

2
u
r
2
+
1
r
2

2
u

2
=

2
u
x
2
+

2
u
y
2

1
r
u
r
.
and hence
u =

2
u
r
2
+
1
r
2

2
u

2
+
1
r
u
r
.
In cylindrical coordinates (3 dimensions), the Laplacian is given as
:=
1
r

r
_
r

r
_
+
1
r
2

2
+

2
z
2
where r [0, ), [0, 2) and z R. In spherical coordinates (3 dimen-
sions), the Laplacian is given as
:=
1
r
2

r
_
r
2

r
_
+
1
r
2
sin

_
sin

_
+
1
r
2
sin
2

2
where r [0, ), [0, ] (zenith angle or inclination) and [0, 2)
(azimuth angle).
Note that in one dimension, n = 1, =
d
2
dx
2
.
CHAPTER 6. THE LAPLACIAN 120
Proposition 6.1.1. Let n 2 and u be a radial function, i.e., u(x) = v(r)
where x R
n
and r = [x[, then
u(x) =
d
2
v(r)
dr
2
+
(n 1)
r
dv(r)
dr
.
Proof. Note that
r
x
i
=
[x[
x
i
=
(
_
x
2
1
+ . . . +x
2
n
)
x
i
=
1
2
(x
2
1
+ . . . +x
2
n
)
1/2
(2x
i
)
=
x
i
r
.
Thus,
u(x) =
n

i=1

x
i
_
u(x)
x
i
_
=
n

i=1

x
i
_
dv(r)
dr
x
i
r
_
=
n

i=1
x
i

x
i
_
1
r
dv(r)
dr
_
+
n
r
dv(r)
dr
=
n

i=1
x
2
i
r
d
dr
_
dv(r)
dr
1
r
_
+
n
r
dv(r)
dr
=
n

i=1
x
2
i
r
_
1
r
d
2
v(r)
dr
2

1
r
2
dv(r)
dr
_
+
n
r
dv(r)
dr
=
r
2
r
_
1
r
d
2
v(r)
dr
2

1
r
2
dv(r)
dr
_
+
n
r
dv(r)
dr
=
d
2
v(r)
dr
2

1
r
dv(r)
dr
+
n
r
dv(r)
dr
=
d
2
v(r)
dr
2
+
(n 1)
r
dv(r)
dr
.
Hence the result proved.
More generally, the Laplacian in R
n
may be written in polar coordinates
as
:=

2
r
2
+
n 1
r

r
+
1
r
2

S
n1
CHAPTER 6. THE LAPLACIAN 121
where
S
n1 is a second order dierential operator in angular variables only.
The above forms of Laplacian gives an insight into how the Laplacian treats
radial function and angular functions. The angular part of Laplacian is
called the Laplace-Beltrami operator acting on S
n1
(unit sphere of R
n
) with
Riemannian metric induced by the standard Euclidean metric in R
n
.
6.2 Ill-Posedness of Cauchy Problem
Recall that for a second order Cauchy problem we need to know both u
and its normal derivative on a data curve contained in . However, the
Cauchy problem for Laplacian (more generally for elliptic equations) is not
well-posed. In fact, the Cauchy problem for Laplace equation on a bounded
domain is over-determined.
Example 6.1 (Hadamard). Consider the Cauchy problem for Laplace equa-
tion
_
_
_
u
xx
+ u
yy
= 0
u(0, y) =
cos ky
k
2
u
x
(0, y) = 0,
where k > 0 is an integer. It is easy to verify that there is a unique solution
u
k
(x, y) =
cosh(kx) cos(ky)
k
2
of the Cauchy problem. Note that for any x
0
> 0,
[u
k
(x
0
, n/k)[ =
cosh(kx
0
)
k
2
.
Since, as k , n/k 0 and [u
k
(x
0
, n/k)[ the Cauchy problem is
not stable, and hence not well-posed.
Exercise 22. Show that the Cauchy problem for Laplace equation
_
_
_
u
xx
+ u
yy
= 0
u(x, 0) = 0
u
y
(x, 0) = k
1
sin kx,
where k > 0, is not well-posed. (Hint: Compute explicit solution using sep-
aration of variable. Note that, as k , the Cauchy data tends to zero
uniformly, but the solution does not converge to zero for any y ,= 0. There-
fore, a small change from zero Cauchy data (with corresponding solution
being zero) may induce bigger change in the solution.)
CHAPTER 6. THE LAPLACIAN 122
This issue of ill-posedness of the Cauchy problem is very special to second
order elliptic equations. In general, any hyperbolic equation Cauchy problem
is well-posed, as long as the hyperbolicity is valid in the full neighbourhood
of the data curve.
Example 6.2. Consider the Cauchy problem for the second order hyperbolic
equation
_
_
_
y
2
u
xx
yu
yy
+
1
2
u
y
= 0 y > 0
u(x, 0) = f(x)
u
y
(x, 0) = g(x).
The general solution to this problem can be computed as
u(x, y) = F
_
x +
2
3
y
3/2
_
+ G
_
x
2
3
y
3/2
_
.
On y = 0 u(x, 0) = F(x) + G(x) = f(x). Further,
u
y
(x, y) = y
1/2
F

_
x +
2
3
y
3/2
_
y
1/2
G

_
x
2
3
y
3/2
_
and u
y
(x, 0) = 0. Thus, the Cauchy problem has no solution unless g(x) = 0.
If g 0 then the solution is
u(x, y) = F
_
x +
2
3
y
3/2
_
F
_
x
2
3
y
3/2
_
+ f
_
x
2
3
y
3/2
_
for arbitrary F C
2
. Therefore, when g 0 the solution is not unique.
The Cauchy problem is not well-posed because the equation is hyperbolic
(B
2
AC = y
3
) not in the full neighbourhood of the data curve y = 0.
6.3 Boundary Conditions
To make the Poisson equation u = f in well-posed, where is a
bounded open subset of R
n
, we choose to specify one of the following condi-
tions on the boundary, , of :
(i) (Dirichlet condition) u = g;
(ii) (Neumann condition) u = g, where (x) is the unit outward normal
of x ;
CHAPTER 6. THE LAPLACIAN 123
(iii) (Robin condition) u + cu = g for any c > 0.
(iv) (Mixed condition) u = g on
1
and u = h on
2
, where
1

2
=
and
1

2
= .
The Poisson equation with Neumann boundary condition comes with a
compatibility condition. Note that, by Guass divergence theorem (cf. Corol-
lary D.0.7), if u is a solution of the Neumann problem then u satises, for
every connected component of ,
_

u =
_

u (Using GDT)

f =
_

g.
The second equality is called the compatibility condition. Thus, for a Neu-
mann problem the given data f, g must necessarily satisfy the compatibility
condition. Otherwise, the Neumann problem does not make any sense.
The aim of this chapter is solve
_
u(x) = f(x) in
one of the above inhomogeneous boudary condition on ,
for any open subset R
n
. Note that, by linearity of Laplacian, u = v +w
where v is a harmonic function
1
that solves
_
v(x) = 0 in
one of the above inhomogeneous boudary condition on ,
and w solves the Poisson equation
2
_
w(x) = f(x) in
one of the above homogeneous boudary condition on .
Therefore, we shall restrict our attention to solving only for v and w. We
begin our analysis with v called harmonic functions.
1
dened later
2
dened later
CHAPTER 6. THE LAPLACIAN 124
6.4 Harmonic Functions
The one dimensional Laplace equation is a ODE and is solvable with solu-
tions u(x) = ax + b for some constants a and b. But in higher dimensions
solving Laplace equation is not so simple. For instance, a two dimensional
Laplace equation
u
xx
+ u
yy
= 0
has the trivial solution, u(x, y) = ax + by + c, all one degree polynomials of
two variables. In addition, xy, x
2
y
2
, x
3
3xy
2
, 3x
2
y y
3
, e
x
sin y and
e
x
cos y are all solutions to the two variable Laplace equation. In R
n
, it is
trivial to check that all polynomials up to degree one, i.e.

||1
a

is a solution to u = 0 in R
n
. But we also have functions of higher degree
and functions not expressible in terms of elementary functions as solutions
to Laplace equation. For instance, note that u(x) =

n
i=1
x
i
is a solution to
u = 0 in R
n
.
Denition 6.4.1. Let be an open subset of R
n
. A function u C
2
() is
said to be harmonic on if u(x) = 0 in .
We already remarked that every scalar potential is a harmonic function.
Gauss was the rst to deduce some important properties of harmonic func-
tions and thus laid the foundation for Potential theory and Harmonic Anal-
ysis.
Due to the linearity of , sum of any nite number of harmonic func-
tions is harmonic and a scalar multiple of a harmonic function is harmonic.
Moreover, harmonic functions can be viewed as the null-space of the Laplace
operator, say from C
2
() to C(), the space of continuous functions.
In two dimension, one can associate with a harmonic function u(x, y), a
conjugate harmonic function, v(x, y) which satisfy the rst order system of
PDE called the Cauchy-Riemann equations,
u
x
= v
y
and u
y
= v
x
.
Harmonic functions and holomorphic functions (dierentiable complex func-
tions) are related in the sense that, for any pair (u, v), harmonic and its conju-
gate, gives a holomorphic function f(z) = u(x, y) +iv(x, y) where z = x+iy.
CHAPTER 6. THE LAPLACIAN 125
Conversely, for any holomorphic function f, its real part and imaginary part
are conjugate harmonic functions. This observation gives us more examples
of harmonic functions, for instance, since all complex polynomials f(z) = z
m
are holomorphic we have (using the polar coordinates) u(r, ) = r
m
cos m
and v(r, ) = r
m
sin m are harmonic functions in R
2
for all m N. Simi-
larly, since f(z) = log z = ln r +i is holomorphic in certain region, we have
u(r, ) = ln r and v(r, ) = are harmonic in R
2
(0, 0) and R
2
= 0,
respectively.
Exercise 23. Show that there are innitely many linearly independent har-
monic functions in the vector space C
2
(R
2
).
6.4.1 Spherical Harmonics
A polynomial of degree k in n-variables is of the form
P
k
(x) :=

||k
a

.
A polynomial P is said to be homogeneous of degree k if P(x) =
k
P(x)
for any real ,= 0. Note that a homogeneous polynomial of degree k is of
the form

||=k
a

.
The number of possible n-tuples such that [[ = k is given by
_
n+k1
k
_
. Let
1
k
(R
n
) denote the set of all homogeneous harmonic polynomial of degree k
in n variables. Note that 1
k
(R
n
) forms a vector space. Recall that the class
of harmonic functions, call it N(), is a (vector) subspace of C
2
(R
n
) and,
hence, 1
k
(R
n
) N() C
2
(R
n
).
Two Dimensions
Consider a general homogeneous polynomial
P
k
(x, y) :=
k

i=0
a
i
x
i
y
ki
CHAPTER 6. THE LAPLACIAN 126
of degree k in R
2
(two variables). Note that P
k
contains k + 1 coecients
3
.
Then
P
k
(x, y) =
k

i=2
a
i
i(i 1)x
i2
y
ki
+
k2

i=0
a
i
(k i)(k i 1)x
i
y
ki2
is a homogeneous polynomial of degree k 2 and, hence, contains k 1
coecients. If P
k
1
k
(R
2
), i.e. P
k
(x, y) = 0, then all the k 1 coecients
should vanish. Thus, we have k 1 equations relating the k + 1 coecients
of P
k
and, hence, 1
k
(R
2
) is of dimension two (since k + 1 (k 1) = 2).
Let us now nd the basis of the two dimensional space 1
k
(R
2
). In polar
coordinates, P
k
(r, ) = r
k
Q
k
() where
Q
k
() =
k

i=0
a
i
(cos )
i
(sin )
ki
.
Note that Q
k
is the restriction of P
k
to S
1
and are called spherical harmonics.
If P
k
1
k
(R
2
) then, using the polar form of Laplacian, we get
r
k2
_
Q

k
() + k
2
Q
k
()

= 0.
Therefore, for all r > 0, Q
k
() is a solution to the ODE
Q

k
() + k
2
Q
k
() = 0.
Therefore, Q
k
() = cos k + sin k and P
k
(r, ) = r
k
(cos k + sin k).
Thus, P
k
is a linear combination r
k
cos k and r
k
sin k. In fact, if we identify
each vector (x, y) R
2
with the complex number z = x + iy, then we have
shown that Re(z
k
), Im(z
k
) are the basis of 1
k
(R
2
). If we choose
1
and

1
such that =
1
sin
1
and =
1
cos
1
, then we can rewrite the
polynomial as
P
k
(r, ) =
1
r
k
cos(k +
1
).
Thus, we immediately see that the zero set of P
k
(r, ) in R
2
will be a family of
k straight lines passing through origin such that between any two consecutive
lines the angle is same.
3
_
k+1
k
_
= k + 1
CHAPTER 6. THE LAPLACIAN 127
Three Dimensions
Consider a general homogeneous polynomial
P
k
(x) :=

||=k
a

of degree k in R
3
(three variables). Note that P
k
contains
_
k+2
k
_
=
(k+2)(k+1)
2
coecients. Then P
k
(x) is a homogeneous polynomial of degree k 2 and,
hence, contains
k(k1)
2
coecients. If P
k
1
k
(R
3
), i.e. P
k
(x) = 0, then all
the
k(k1)
2
coecients should vanish. Thus, we have
k(k1)
2
equations relating
the
(k+2)(k+1)
2
coecients of P
k
and, hence, 1
k
(R
3
) is of dimension
(k + 2)(k + 1) k(k 1)
2
= 2k + 1.
The basis of the 2k+1 dimensional space 1
k
(R
3
) is given in terms of the Leg-
endre functions which we shall describe now. In spherical coordinates, x =
r sin cos , y = r sin sin and z = r cos . Thus, P
k
(r, , ) = r
k
R()Q()
where
R()Q() =

||=k
a

(sin )

1
+
2
(cos )

3
(cos )

1
(sin )

2
.
The separated variable assumption above is not a issue because dierential
operator is linear. Note that RQ is the restriction of P
k
to S
2
and are
called spherical harmonics. If P
k
1
k
(R
2
) then, using the spherical form of
Laplacian, we get
r
k2
_
k(k + 1) sin
2
+ sin
2

()
R()
+ sin cos
R

()
R()
+
Q

()
Q()
_
= 0.
Therefore, for all r > 0, we have equality
k(k + 1) sin
2
+ sin
2

()
R()
+ sin cos
R

()
R()
=
Q

()
Q()
.
Since LHS is a function of and RHS is a function of they must be equal
to some constant . Then, we have to solve for the eigenvalue problem
Q

() = Q()
CHAPTER 6. THE LAPLACIAN 128
where Q is 2-periodic. This has the solution, for all m N 0, = m
2
and Q
m
() =
m
cos m +
m
sin m. For = m
2
we solve for R() in
R

() +
cos
sin
R

() = R()
_
m
2
sin
2

k(k + 1)
_
(0, ).
Set w = cos . Then
dw
d
= sin .
R

() = sin
dR
dw
and R

() = sin
2

d
2
R
dw
2
cos
dR
dw
In the new variable w, we get the Legendre equation
(1 w
2
)R

(w) 2wR

(w) =
_
m
2
1 w
2
k(k + 1)
_
R(w) w [1, 1].
For each k N 0, this has the Legendre polynomials, R
k,m
(cos ), as its
solutions. Therefore, in general,
P
k
(r, , ) = r
k
(cos m + sin m)R
k,m
(cos ).
However, we are interested only those R
k,m
which gives a polynomial of degree
k in R
3
. Thus, for m = 0, 1, . . . , k,
R
k,m
(w) = (1 w
2
)
m/2
d
k+m
dw
k+m
(1 w
2
)
k
.
Note that, for each xed k and all 1 m k, the collection
R
k,0
(cos ), cos mR
k,m
(cos ), sin mR
k,m
(cos ) 1
k
(R
3
)
is 2k +1 linearly independent homogeneous harmonic polynomials of degree
k and forms a basis. Thus, each P
k
is a linear combination of these basis
elements.
The zero sets of P
k
exhibit properties depending on m. For m = 0 the
harmonic polynomial P
k
is a constant multiple of R
k,0
(cos ). Since R
k,0
(w)
has k distinct zeros in [1, 1] arranged symmetrically about w = 0, there are
k distince zeros of R
k,0
(cos ) in (0, ) arranged symmetrically about /2.
Thus on S
2
, the unit sphere, the function R
k,0
(cos ) vanishes on k circles
circumscribed in the latitudinal direction. For k odd the circle along equator
CHAPTER 6. THE LAPLACIAN 129
is also a zero set. The function R
k,0
(cos ) and its constant multiples are
called zonal harmonics.
If 0 < m < k, then the spherical harmonics is of the form
(cos m + sin m) sin
m

d
k+m
dw
k+m
(1 w
2
)
k
.
If the rst term is zero then tan m = /. This corresponds to great
circle through the north pole and south pole of S
2
and the angle between
the planes containing two consecutive great circle is /m. The second term
vanishes on = 0 and = corresponding to the north and south pole,
respectively. The third term vanishes on km latitude circle. Thus, we have
orthogonally intersecting family of circles which form the zero set which are
called tesseral harmonics.
If m = k then the spherical harmonics is of the form
(cos k + sin k) sin
k

and it vanishes for = 0, = or tan k = /. The rst two cases


corresponds to the north and south pole, respectively, and the last case cor-
responds to great circles through the north pole and south pole of S
2
and
the angle between the planes containing two consecutive great circle is /k.
Thus, the great circles divide the S
2
in to 2k sectors and are called sectorial
harmoics.
Higher Dimensions
Consider a general homogeneous polynomial
P
k
(x) :=

||=k
a

of degree k in R
n
(n variables). Note that P
k
contains
_
n+k1
k
_
coecients.
Then P
k
(x) is a homogeneous polynomial of degree k 2 and, hence, con-
tains
_
n+k3
k2
_
coecients. If P
k
1
k
(R
n
), i.e. P
k
(x) = 0, then all the
_
n+k3
k2
_
coecients should vanish. Thus, we have
_
n+k3
k2
_
equations relating
_
n+k1
k
_
coecients of P
k
and, hence, 1
k
(R
n
) is of dimension
:=
_
n + k 1
k
_

_
n + k 3
k 2
_
.
CHAPTER 6. THE LAPLACIAN 130
In polar form, P
k
(r, ) = r
k
Q() where S
n1
and if P
k
(r, ) 1
k
(R
n
)
then
P
k
=

2
P
k
r
2
+
n 1
r
P
k
r
+
1
r
2

S
n1P
k
= 0
where
S
n1 is a second order dierential operator in angular variables only
called the Laplace-Beltrami operator. Therefore, we have
r
k2
[
S
n1Q() + k(n + k 2)Q()] = 0
and for r > 0,

S
n1Q() + k(n + k 2)Q() = 0.
6.4.2 Properties of Harmonic Functions
In this section we shall study properties of harmonic functions. We shall as-
sume the divergence theorems from multivariable calculus (cf. Appendix D).
Also, note that if u is a harmonic function on then, by Gauss divergence
theorem (cf. Theorem D.0.6),
_

d = 0.
Denition 6.4.2. Let be an open subset of R
n
and w
n
=
2
n/2
(n/2)
(cf. Ap-
pendix E) be the surface area of the unit sphere S
1
(0) of R
n
.
(a) A function u C() is said to satisfy the rst mean value property
(I-MVP) in if
u(x) =
1

n
r
n1
_
Sr(x)
u(y) d
y
for any B
r
(x) .
(b) A function u C() is said to satisfy the second mean value property
(II-MVP) in if
u(x) =
n

n
r
n
_
Br(x)
u(y) dy for any B
r
(x) .
CHAPTER 6. THE LAPLACIAN 131
Exercise 24. Show that u satises the I-MVP i
u(x) =
1

n
_
S
1
(0)
u(x +rz) d
z
.
Similarly, u satises II-MVP i
u(x) =
n

n
_
B
1
(0)
u(x +rz) dz.
Exercise 25. Show that the rst MVP and second MVP are equivalent. That
is show that u satises (a) i u satises (b).
Owing to the above exercise we shall, henceforth, refer to the I-MVP and
II-MVP as just mean value property (MVP).
We shall now prove a result on the smoothness of a function satisfying
MVP.
Theorem 6.4.3. If u C() satises the MVP in , then u C

().
Proof. We rst consider u

:=

u, the convolution of u with molliers, as


introduced in Theorem F.0.10. where

:= x [ dist(x, ) > .
We shall now show that u = u

for all > 0, due to the MVP of u and the


radial nature of . Let x

. Consider
u

(x) =
_

(x y)u(y) dy
=
_
B(x)

(x y)u(y) dy (Since supp(

) is in B

(x))
=
_

0

(r)
__
Sr(x)
u(y) d
y
_
dr (cf. Theorem E.0.8)
= u(x)
n
_

0

(r)r
n1
dr (Using MVP of u)
= u(x)
_

0

(r)
__
Sr(0)
d
y
_
dr
= u(x)
_
B(0)

(y) dy = u(x).
Thus u

(x) = u(x) for all x

and for all > 0. Since u

) for
all > 0 (cf. Theorem F.0.10), we have u C

) for all > 0.


CHAPTER 6. THE LAPLACIAN 132
Theorem 6.4.4. Let u be a harmonic function on . Then u satises the
MVP in .
Proof. Let B
r
(x) be any ball with centre at x and for some r > 0.
For the given harmonic function u, we set
v(r) :=
1

n
r
n1
_
Sr(x)
u(y) d
y
.
Note that v is not dened at 0, since r > 0. We have from Exercise 24 that
v(r) =
1

n
_
S
1
(0)
u(x +rz) d
z
.
Now, dierentiating both sides w.r.t r, we get
dv(r)
dr
=
1

n
_
S
1
(0)
u(x + rz) z d
z
=
1

n
r
n1
_
Sr(x)
u(y)
(y x)
r
d
y
Since [x y[ = r, by setting := (y x)/r as the unit vector, and applying
the Gauss divergence theorem along with the fact that u is harmonic, we get
dv(r)
dr
=
1

n
r
n1
_
Sr(x)
u(y) d
y
=
1

n
r
n1
_
Br(x)
u(y) dy = 0.
Thus, v is a constant function of r > 0 and hence
v(r) = v() > 0.
Moreover, since v is continuous (constant function), we have
v(r) = lim
0
v()
= lim
0
1

n
_
S
1
(0)
u(x + z) d
z
=
1

n
_
S
1
(0)
lim
0
u(x + z) d
z
(u is continuous on S
1
(0))
=
1

n
_
S
1
(0)
u(x) d
z
= u(x) (Since
n
is the surface area of S
1
(0)).
Thus, u satises I-MVP and hence the II-MVP.
CHAPTER 6. THE LAPLACIAN 133
Corollary 6.4.5. If u is harmonic on , then u C

().
The above corollary is a easy consequence of Theorem 6.4.4 and Theo-
rem 6.4.3. We shall now prove that any function satisfying MVP is harmonic.
Theorem 6.4.6. If u C() satises the MVP in , then u is harmonic
in .
Proof. Since u satises MVP, by Theorem 6.4.3, u C

(). Thus, u
makes sense. Now, suppose u is not harmonic in , then there is a x
such that u(x) ,= 0. Without loss of generality, lets say u(x) > 0.
Moreover, since u is continuous there is a s > 0 such that, for all y B
s
(x),
u(y) > 0. As done previously, we set for r > 0,
v(r) :=
1

n
r
n1
_
Sr(x)
u(y) d
y
.
Thus, v(r) = u(x) for all r > 0 and hence v is a constant function of r and
v

(s) = 0. But
0 =
dv(s)
dr
=
1

n
r
n1
_
Bs(x)
u(y) dy > 0
is a contradiction. Therefore, u is harmonic in .
Above results leads us to conclude that a function is harmonic i it sat-
ises the MVP.
Exercise 26. If u
m
is a sequence of harmonic functions in converging to u
uniformly on compact subsets of , then show that u is harmonic in .
Theorem 6.4.7 (Strong Maximum Principle). Let be an open, connected
(domain) subset of R
n
. Let u be harmonic in and M := max
y
u(y).
Then
u(x) < M x
or u M is constant in .
Proof. We dene a subset S of as follows,
S := x [ u(x) = M.
CHAPTER 6. THE LAPLACIAN 134
If S = , we have u(x) < M for all x . Suppose S ,= . Then S is closed
subset of , since u is continuous. Now, for any x S, by MVP
u(x) =
n

n
r
n
_
Br(x)
u(y) dy for every r such that B
r
(x) .
Thus, we have
M = u(x) =
n

n
r
n
_
Br(x)
u(y) dy M
Hence equality will hold above only when u(y) = M for all y B
r
(x). Thus,
we have shown that for any x S, we have B
r
(x) S. Therefore, S is open.
Since is connected, the only open and closed subsets are or . Since
S was assumed to be non-empty, we should have S = . Thus, u M is
constant in .
Corollary 6.4.8 (Weak maximum Principle). Let be an open, bounded
subset of R
n
. Let u C() be harmonic in . Then
max
y
u(y) = max
y
u(y).
Proof. Let M := max
y
u(y). If there is a x such that u(x) = M, then
u M is constant on the connected component of containing x. Thus,
u = M on the boundary of the connected component which is a part of
.
Aliter. Since , we have max

u max

u. It only remains to
prove the other equality. For the given harmonic function u and for a xed
> 0, we set v

(x) = u(x) + [x[


2
, for each x . For each x ,
v

= u + 2n > 0. Recall that


4
if a function v attains local maximum
at a point x , then in each direction its second order partial derivative
v
x
i
x
i
(x) 0, for all i = 1, 2, . . . , n. Therefore v(x) 0. Thus, we argue
that v

does not attain (even a local) maximum in . But v

has to have a
maximum in , hence it should be attained at some point x

, on the
boundary. For all x ,
u(x) v

(x) v

(x

) = u(x

) + [x

[
2
max
x
u(x) + max
x
[x[
2
.
The above inequality is true for all > 0. Thus, u(x) max
x
u(x), for all
x . Therefore, max

u max
x
u(x). and hence we have equality.
4
v C
2
(a, b) has a local maximum at x (a, b) then v

(x) = 0 and v

(x) 0
CHAPTER 6. THE LAPLACIAN 135
Theorem 6.4.9 (Estimates on derivatives). If u is harmonic in , then
[D

u(x)[
C
k
r
n+k
|u|
1,Br(x)
B
r
(x) and each [[ = k
where the constants C
0
=
n
n
and C
k
= C
0
(2
n+1
nk)
k
for k = 1, 2, . . ..
Proof. We prove the result by induction on k. Let k = 0. Since u is harmonic,
by II-MVP we have, for any B
r
(x) ,
[u(x)[ =
n

n
r
n

_
Br(x)
u(y) dy

n
r
n
_
Br(x)
[u(y)[ dy
=
n

n
r
n
|u|
1,Br(x)
=
C
0
r
n
|u|
1,Br(x)
.
Now, let k = 1. Observe that if u is harmonic then by dierentiating the
Laplace equation and using the equality of mixed derivatives, we have u
x
i
:=
u
x
i
is harmoic, for all i = 1, 2, . . . , n. Now, by the II-MVP of u
x
i
, we have
[u
x
i
(x)[ =
n2
n

n
r
n

_
B
r/2
(x)
u
x
i
(y) dy

=
n2
n

n
r
n

_
S
r/2
(x)
u
i
d
y

(by Gauss-Green theorem)

2n
r
|u|
,S
r/2
(x)
.
Thus, it now remains to estimate |u|
,S
r/2
(x)
. Let z S
r/2
(x), then
B
r/2
(z) B
r
(x) .
But, using k = 0 result, we have
[u(z)[
C
0
2
n
r
n
|u|
1,B
r/2
(z)

C
0
2
n
r
n
|u|
1,Br(x)
.
Therefore, |u|
,S
r/2
(x)

C
0
2
n
r
n
|u|
1,Br(x)
and using this in the estimate of u
x
i
,
we get
[u
x
i
(x)[
C
0
n2
n+1
r
n+1
|u|
1,Br(x)
.
CHAPTER 6. THE LAPLACIAN 136
Hence
[D

u(x)[
C
1
r
n+1
|u|
1,Br(x)
for [[ = 1.
Let now k 2 and be a multi-index such that [[ = k. We assume the
induction hypothesis that the estimate to be proved is true for k 1. Note
that D

u =
D

u
x
i
for some i 1, 2, . . . , n and [[ = k 1. Moreover,
if u is harmonic then by dierentiating the Laplace equation and using the
equality of mixed derivatives, we have
D

u
x
i
is harmoic for i = 1, 2, . . . , n.
Thus, following an earlier argument, we have
[D

u(x)[ =

u(x)
x
i

=
nk
n

n
r
n

_
B
r/k
(x)
D

u(y)
x
i
dy

=
nk
n

n
r
n

_
S
r/k
(x)
D

u
i
d
y

nk
r
|D

u|
,S
r/k
(x)
.
It now only remains to estimate |D

u|
,S
r/k
(x)
. Let z S
r/k
(x), then
B
(k1)r/k
(z) B
r
(x) . But, using induction hypothesis for k 1, we
have
[D

u(z)[
C
k1
k
n+k1
((k 1)r)
n+k1
|u|
1,B
(k1)r/k
(z)

C
k1
k
n+k1
((k 1)r)
n+k1
|u|
1,Br(x)
.
Therefore, using the above estimate for D

u, we get
[D

u(x)[
C
k1
nk
n+k
(k 1)
n+k1
r
n+k
|u|
1,Br(x)
=
C
0
2
(n+1)(k1)
n
k
(k 1)
k1
k
n+k
(k 1)
n+k1
r
n+k
|u|
1,Br(x)
=
C
0
(2
n+1
nk)
k
r
n+k
_
k
k 1
_
n
_
1
2
n+1
_
|u|
1,Br(x)
=
C
0
(2
n+1
nk)
k
r
n+k
_
k
2(k 1)
_
n
_
1
2
_
|u|
1,Br(x)

C
k
r
n+k
|u|
1,Br(x)
since
_
k
2(k 1)
_
n
_
1
2
_
1.
CHAPTER 6. THE LAPLACIAN 137
Hence
[D

u(x)[
C
k
r
n+k
|u|
1,Br(x)
for [[ = k, k 2.
Theorem 6.4.10 (Liouvilles Theorem). If u is bounded and harmonic on
R
n
, then u is constant.
Proof. For any x R
n
and r > 0, we have the estimate on the rst derivative
as,
[u(x)[
C
1
r
n+1
|u|
1,Br(x)
=
2
n+1
n

n
r
n+1
|u|
1,Br(x)

2
n+1
n

n
r
n+1
|u|
,R
n
n
r
n
=
2
n+1
r
|u|
,R
n 0 as r .
Thus, u 0 in R
n
and hence u is constant.
Exercise 27. Show that if u is harmonic in , then u is analytic in . (Hint:
Use the estimates on derivatives with Stirlings formula and Taylor expan-
sion).
We end our discussion on the properties of harmonic function with Har-
nack inequality. The Harnack inequality states that non-negative harmonic
functions cannot be very large or very small at any point without being so
everywhere in a compact set containing that point.
Theorem 6.4.11 (Harnacks Inequality). Let u be harmonic in and u 0
in , then for each connected open subset there is a constant C > 0
(depending only on ) such that
sup
x
u(x) C inf
x
u(x).
In particular,
1
C
u(y) u(x) Cu(y) x, y .
CHAPTER 6. THE LAPLACIAN 138
Proof. Set r :=
1
4
dist(, ). Let x, y such that [xy[ < r. By II-MVP,
u(x) =
n

n
2
n
r
n
_
B
2r
(x)
u(z) dz

n
2
n
r
n
_
Br(y)
u(z) dz =
1
2
n
u(y).
Thus, 1/2
n
u(y) u(x). Interchanging the role of x and y, we get 1/2
n
u(x)
u(y). Thus, 1/2
n
u(y) u(x) 2
n
u(y) for all x, y such that [x y[ r.
Now, let x, y . Since is compact and connected in , we can pick
points x = x
1
, x
2
, . . . , x
m
= y such that
m
i=1
B
i
, where B
i
:= B
r/2
(x
i
)
and are sorted such that B
i
B
i+1
,= , for i = 1, 2, . . . , m1. Hence, note
that [x
i+1
x
i
[ r. Therefore,
u(x) = u(x
1
)
1
2
n
u(x
2
)
1
2
2n
u(x
3
) . . .
1
2
(m1)n
u(x
m
) =
1
2
(m1)n
u(y).
Thus, C can be chosen to be
1
2
(m1)n
.
6.5 Existence and Uniqueness
A consequence of the maximum principle is the uniqueness of the harmonic
functions.
Theorem 6.5.1 (Uniqueness of Harmonic Functions). Let be an open,
bounded subset of R
n
. Let u
1
, u
2
C
2
() C() be harmonic in such that
u
1
= u
2
on , then u
1
= u
2
in .
Proof. Note that u
1
u
2
is a harmonic function and hence, by weak maximum
principle, should attain its maximum on . But u
1
u
2
= 0 on . Thus
u
1
u
2
0 in . Now, repeat the argument for u
2
u
1
, we get u
2
u
1
0
in . Thus, we get u
1
u
2
= 0 in .
Let u C
2
() C() be a solution of the Dirichlet problem
_
u(x) = 0 x
u(x) = g(x) x .
(6.5.1)
By the strong maximum principle (cf. Theorem 6.4.7), if is connected and
g 0 and g(x) > 0 for some x then u(x) > 0 for all x .
CHAPTER 6. THE LAPLACIAN 139
Theorem 6.5.2. Let be an open bounded connected subset of R
n
and
g C(). Then the Dirichlet problem (6.5.1) has atmost one solution
u C
2
() C(). Moreover, if u
1
and u
2
are solution to the Dirichlet
problem corresponding to g
1
and g
2
in C(), respectively, then
(a) (Comparison) g
1
g
2
on and g
1
(x
0
) > g
2
(x
0
) for some x
implies that u
1
> u
2
in .
(b) (Stability) [u
1
(x) u
2
(x)[ max
y
[g
1
(y) g
2
(y)[ for all x .
Proof. The fact that there is atmost one solution to the Dirichlet problem
follows from the Theorem 6.5.1. Let w = u
1
u
2
. Then w is harmonic.
(a) Note that w = g
1
g
2
0 on . Since g
1
(x
0
) > g
2
(x
0
) for some
x
0
, then w(x) > 0 for all x . This proves the comparison
result.
(b) Again, by maximum principle, we have
w(x) max
y
[g
1
(y) g
2
(y)[x .
This proves the stability result.
We remark that the uniqueness result is not true for unbounded domains.
Example 6.3. Consider the problem (6.5.1) with g 0 in the domain =
x R
n
[ [x[ > 1. Obviously, u = 0 is a solution. But we also have a
non-trivial solution
u(x) =
_
ln [x[ n = 2
[x[
2n
1 n 3.
Example 6.4. Consider the problem (6.5.1) with g 0 in the domain =
x R
n
[ x
n
> 0. Obviously, u = 0 is a solution. But we also have a
non-trivial solution u(x) = x
n
.
We have shown above that if a solution exists for (6.5.1) then it is unique
(cf. Theorem 6.5.1). So the question that remains to be answered is on
the existence of solution of (6.5.1), for any given domain . In the modern
theory, there are three dierent methods to address this question of existence,
viz., Perrons Method, Layer Potential (Integral Equations) and L
2
methods.
CHAPTER 6. THE LAPLACIAN 140
6.6 Perrons Method
Denition 6.6.1. We say a function w C() is a barrier at x
0
if
there is a neighbourhood U of x
0
such that
1. w is superharmonic in U
2. w > 0 in ( U) x
0
and w(x
0
) = 0.
Denition 6.6.2. Any point on is said to be regular (w.r.t Laplacian)
if there exists a barrier at that point.
A necessary and sucient condition for the existence of solution to (6.5.1)
is given by the following result:
Theorem 6.6.3. The Dirichlet problem (6.5.1) is solvable for any arbitrary
bounded domain and for any arbitrary g on i all the points in are
regular.
Proof. One way is obvious. If (6.5.1) is solvable and x
0
. Then, the
solution to
_
w = 0 in
w = g on ,
where g(x) = [x x
0
[, is a barrier function at x
0
. Thus, any x
0

is a regular point. The converse is proved using the Perrons method for
subharmonic functions.
Denition 6.6.4. A bounded domain R
n
is said to satisfy the exterior
sphere condition if for every point x
0
there is a ball B := B
R
(y) such
that B = x
0
.
Lemma 6.6.5. If satises the exterior sphere condition then all boundary
points of are regular.
Proof. For any x
0
, we dene the barrier function at x
0
as
w(x) =
_
R
2n
[x y[
2n
for n 3
ln
_
|xy|
R
_
for n = 2.
CHAPTER 6. THE LAPLACIAN 141
Theorem 6.6.6. Any bounded domain with C
2
boundary satises the exte-
rior sphere condition.
Denition 6.6.7. A bounded domain R
n
is said to satisfy the exterior
cone condition if for every point x
0
there is a nite right circular cone
K with vertex at x
0
such that K = x
0
.
Exercise 28. Any domain satisfying the exterior sphere condition also satises
the exterior cone condition.
Exercise 29. Every bounded Lipschitz domain satises the exterior cone con-
dition.
Lemma 6.6.8. If satises the exterior cone condition then all boundary
points of are regular.
6.6.1 Non-existence of Solutions
In 1912, Lebesgue gave an example of a domain on which the classical Dirich-
let problem is not solvable. Let
:= (x, y, z) R
3
[ r
2
+ z
2
< 1; r > e
1/2z
for z > 0.
Note that is the unit ball in R
3
with a sharp inward cusp, called Lebesgue
spine, at (0, 0, 0). The origin is a not regular point of .
However, there do exist domains with inward cusps for which the classical
problem is solvable, for instance, consider
:= (x, y, z) R
3
[ r
2
+z
2
< 1; r > z
2k
for z > 0,
for any positive integer k. The proof of this fact involves the theory of
capacities.
6.6.2 Characterizing regular points
The Wieners criterion gives the necessary and sucient condition for the
regularity of the boundary points. For n 3 and a xed (0, 1), the
Wieners criterion states that a point x
0
is regular i the series

i=0
C
i

i(n2)
diverges, where C
i
:= cap
2
x / [ [x x
0
[
i
.
CHAPTER 6. THE LAPLACIAN 142
6.7 with Simple Geometry
The method of separation of variables was introduced by dAlembert (1747)
and Euler (1748) for the wave equation. This technique was also employed by
Laplace (1782) and Legendre (1782) while studying the Laplace equation and
also by Fourier while studying the heat equation. The motivation behind the
separation of variable technique will be highlighted while studying wave
equation.
Theorem 6.7.1 (2D Rectangle). Let = (x, y) R
2
[ 0 < x < a and 0 <
y < b be a rectangle in R
2
. Let g : R which vanishes on three sides
of the rectangle, i.e., g(0, y) = g(x, 0) = g(a, y) = 0 and g(x, b) = h(x) where
h is a continuous function h(0) = h(a) = 0. Then there is a unique solution
to (6.5.1) on this rectangle with given boundary value g.
Proof. We begin by looking for solution u(x, y) whose variables are separated,
i.e., u(x, y) = v(x)w(y). Substituting this form of u in the Laplace equation,
we get
v

(x)w(y) + v(x)w

(y) = 0.
Hence
v

(x)
v(x)
=
w

(y)
w(y)
.
Since LHS is function of x and RHS is function y, they must equal a constant,
say . Thus,
v

(x)
v(x)
=
w

(y)
w(y)
= .
Using the boundary condition on u, u(0, y) = g(0, y) = g(a, y) = u(a, y) =
0, we get v(0)w(y) = v(a)w(y) = 0. If w 0, then u 0 which is not a
solution to (6.5.1). Hence, w , 0 and v(0) = v(a) = 0. Thus, we need to
solve,
_
v

(x) = v(x), x (0, a)


v(0) = v(a) = 0,
the eigen value problem for the second order dierential operator. Note that
the can be either zero, positive or negative.
If = 0, then v

= 0 and the general solution is v(x) = x +, for some


constants and . Since v(0) = 0, we get = 0, and v(a) = 0 and a ,= 0
implies that = 0. Thus, v 0 and hence u 0. But, this can not be a
solution to (6.5.1).
CHAPTER 6. THE LAPLACIAN 143
If > 0, then v(x) = e

x
+ e

x
. Equivalently,
v(x) = c
1
cosh(

x) + c
2
sinh(

x)
such that = (c
1
+c
2
)/2 and = (c
1
c
2
)/2. Using the boundary condition
v(0) = 0, we get c
1
= 0 and hence
v(x) = c
2
sinh(

x).
Now using v(a) = 0, we have c
2
sinh

a = 0. Thus, c
2
= 0 and v(x) = 0.
We have seen this cannot be a solution.
If < 0, then set =

. We need to solve
_
v

(x) +
2
v(x) = 0 x (0, a)
v(0) = v(a) = 0.
(6.7.1)
The general solution is
v(x) = cos(x) + sin(x).
Using the boundary condition v(0) = 0, we get = 0 and hence v(x) =
sin(x). Now using v(a) = 0, we have sin a = 0. Thus, either = 0
or sin a = 0. But = 0 does not yield a solution. Hence a = k or
= k/a, for all non-zero k Z. Hence, for each k N, there is a solution
(v
k
,
k
) for (6.7.1), with
v
k
(x) =
k
sin
_
kx
a
_
,
for some constant
k
and
k
= (k/a)
2
. We now solve w corresponding to
each
k
. For each k N, we solve for w
k
in the ODE
_
w

k
(y) =
_
k
a
_
2
w
k
(y), y (0, b)
w(0) = 0.
Thus, w
k
(y) = c
k
sinh(ky/a). Therefore, for each k N,
u
k
=
k
sin
_
kx
a
_
sinh
_
ky
a
_
CHAPTER 6. THE LAPLACIAN 144
is a solution to (6.5.1). The general solution is of the form (principle of
superposition) (convergence?)
u(x, y) =

k=1

k
sin
_
kx
a
_
sinh
_
ky
a
_
.
The constant
k
are obtained by using the boundary condition u(x, b) = h(x)
which yields
h(x) = u(x, b) =

k=1

k
sinh
_
kb
a
_
sin
_
kx
a
_
.
Since h(0) = h(a) = 0, we know that h admits a Fourier Sine series. Thus

k
sinh
_
kb
a
_
is the k-th Fourier sine coecient of h, i.e.,

k
=
_
sinh
_
kb
a
__
1
2
a
_
a
0
h(x) sin
_
kx
a
_
.
Theorem 6.7.2 (2D Disk). Let = (x, y) R
2
[ x
2
+y
2
< R
2
be the disk
of radius R in R
2
. Let g : R is a continuous function. Then there is
a unique solution to (6.5.1) on the unit disk with given boundary value g.
Proof. Given the nature of the domain, we shall use the Laplace operator in
polar coordinates,
:=
1
r

r
_
r

r
_
+
1
r
2

2
where r is the magnitude component and is the direction component. Then
is the circle of radius one. Then, solving for u(x, y) in the Dirichlet
problem is to equivalent to nding U(r, ) : R such that
_
_
_
1
r

r
_
r
U
r
_
+
1
r
2

2
U

2
= 0 in
U(r, + 2) = U(r, ) in
U(R, ) = G() on
(6.7.2)
where U(r, ) = u(r cos , r sin ), G : [0, 2) R is G() = g(cos , sin ).
Note that both U and G are 2 periodic w.r.t . We will look for solution
CHAPTER 6. THE LAPLACIAN 145
U(r, ) whose variables can be separated, i.e., U(r, ) = v(r)w() with both
v and w non-zero. Substituting it in the polar form of Laplacian, we get
w
r
d
dr
_
r
dv
dr
_
+
v
r
2
d
2
w
d
2
= 0
and hence
r
v
d
dr
_
r
dv
dr
_
=
1
w
_
d
2
w
d
2
_
.
Since LHS is a function of r and RHS is a function of , they must equal a
constant, say . We need to solve the eigen value problem,
_
w

() w() = 0 R
w( + 2) = w() .
Note that the can be either zero, positive or negative. If = 0, then
w

= 0 and the general solution is w() = +, for some constants and


. Using the periodicity of w,
+ = w() = w( + 2) = + 2 +
implies that = 0. Thus, the pair = 0 and w() = is a solution. If
> 0, then
w() = e

+e

.
If either of and is non-zero, then w() as , which contra-
dicts the periodicity of w. Thus, = = 0 and w 0, which cannot be a
solution. If < 0, then set =

and the equation becomes


_
w

() +
2
w() = 0 R
w( + 2) = w()
Its general solution is
w() = cos() + sin().
Using the periodicity of w, we get = k where k is an integer. For each
k N, we have the solution (w
k
,
k
) where

k
= k
2
and w
k
() =
k
cos(k) +
k
sin(k).
CHAPTER 6. THE LAPLACIAN 146
For the
k
s, we solve for v
k
, for each k = 0, 1, 2, . . .,
r
d
dr
_
r
dv
k
dr
_
= k
2
v
k
.
For k = 0, we get v
0
(r) = ln r + . But ln r blows up as r 0, but any
solution U and, hence v, on the closed unit disk (compact subset) has to be
bounded. Thus, we must have the = 0. Hence v
0
. For k N, we need
to solve for v
k
in
r
d
dr
_
r
dv
k
dr
_
= k
2
v
k
.
Use the change of variable r = e
s
. Then e
s ds
dr
= 1 and
d
dr
=
d
ds
ds
dr
=
1
e
s
d
ds
.
Hence r
d
dr
=
d
ds
. v
k
(e
s
) = e
ks
+e
ks
. v
k
(r) = r
k
+r
k
. Since r
k
blows
up as r 0, we must have = 0. Thus, v
k
= r
k
. Therefore, for each
k = 0, 1, 2, . . .,
U
k
(r, ) = a
k
r
k
cos(k) + b
k
r
k
sin(k).
The general solution is
U(r, ) =
a
0
2
+

k=1
_
a
k
r
k
cos(k) + b
k
r
k
sin(k)
_
.
To nd the constants, we must use U(R, ) = G(). If G C
1
[0, 2], then G
admits Fourier series expansion. Therefore,
G() =
a
0
2
+

k=1
_
R
k
a
k
cos(k) + R
k
b
k
sin(k)

where
a
k
=
1
R
k

G() cos(k) d,
b
k
=
1
R
k

G() sin(k) d.
Using this in the formula for U and the uniform convergence of Fourier series,
we get
U(r, ) =
1

G()
_
1
2
+

k=1
_
r
R
_
k
(cos k cos k + sin k sin k)
_
d
=
1

G()
_
1
2
+

k=1
_
r
R
_
k
cos k( )
_
d.
CHAPTER 6. THE LAPLACIAN 147
Using the relation

k=1
_
r
R
_
k
cos k( ) = Re
_

k=1
_
r
R
e
i()
_
k
_
= Re
_
r
R
e
i()
1
r
R
e
i()
_
=
R
2
rRcos( )
R
2
+ r
2
2rRcos( )
1
=
rRcos( ) r
2
R
2
+ r
2
2rRcos( )
in U(r, ) we get
U(r, ) =
R
2
r
2
2
_

G()
R
2
+r
2
2rRcos( )
d.
Note that the formula derived above for U(r, ) can be rewritten in Carte-
sian coordinates and will have the form
u(x) =
R
2
[x[
2
2R
_
S
R
(0)
g(y)
[x y[
2
dy.
This can be easily seen, by setting y = R(x
1
0
cos +x
2
0
sin ), we get dy = Rd
and [x y[
2
= R
2
+r
2
2rRcos( ). This is called the Poisson formula.
More generally, the unique solution to the Dirichlet problem on a ball of
radius R centred at x
0
in R
n
is given by Poisson formula
u(x) =
R
2
[x x
0
[
2

n
R
_
S
R
(x
0
)
g(y)
[x y[
n
dy.
We will derive this general form later (cf. (6.8.7)).
Theorem 6.7.3 (3D Sphere). Let = (x, y, z) R
3
[ x
2
+y
2
+z
2
< 1 be
the unit sphere in R
3
. Let g : R is a continuous function. Then there
is a unique solution to (6.5.1) on the unit sphere with given boundary value
g.
Proof. Given the nature of domain, the Laplace operator in spherical coor-
dinates,
:=
1
r
2

r
_
r
2

r
_
+
1
r
2
sin

_
sin

_
+
1
r
2
sin
2

2
.
CHAPTER 6. THE LAPLACIAN 148
where r is the magnitude component, is the inclination (zenith or elevation)
in the vertical plane and is the azimuth angle (in the direction in horizontal
plane). Solving for u in (6.5.1) is equivalent to nding U(r, , ) : R
such that
_

_
1
r
2

r
_
r
2 U
r
_
+
1
r
2
sin

_
sin
U

_
+
1
r
2
sin
2

2
U

2
= 0 in
U(1, , ) = G(, ) on
(6.7.3)
where U(r, , ) and G(, ) are appropriate spherical coordinate function
corresponding to u and g. We will look for solution U(r, , ) whose variables
can be separated, i.e., U(r, , ) = v(r)w()z() with v, w and z non-zero.
Substituting it in the spherical form of Laplacian, we get
wz
r
2
d
dr
_
r
2
dv
dr
_
+
vz
r
2
sin
d
d
_
sin
dw
d
_
+
vw
r
2
sin
2

d
2
z
d
2
= 0
and hence
1
v
d
dr
_
r
2
dv
dr
_
=
1
wsin
d
d
_
sin
dw
d
_

1
z sin
2

d
2
z
d
2
.
Since LHS is a function of r and RHS is a function of (, ), they must equal
a constant, say . If Azimuthal symmetry is present then z() is constant
and hence
dz
d
= 0. We need to solve for w,
sin w

() + cos w

() +sin w() = 0, (0, )


Set x = cos . Then
dx
d
= sin .
w

() = sin
dw
dx
and w

() = sin
2

d
2
w
dx
2
cos
dw
dx
In the new variable x, we get the Legendre equation
(1 x
2
)w

(x) 2xw

(x) + w(x) = 0 x [1, 1].


We have already seen that this is a singular problem (while studying S-L
problems). For each k N 0, we have the solution (w
k
,
k
) where

k
= k(k + 1) and w
k
() = P
k
(cos ).
CHAPTER 6. THE LAPLACIAN 149
For the
k
s, we solve for v
k
, for each k = 0, 1, 2, . . .,
d
dr
_
r
2
dv
k
dr
_
= k(k + 1)v
k
.
For k = 0, we get v
0
(r) = /r +. But 1/r blows up as r 0 and U must
be bounded in the closed sphere. Thus, we must have the = 0. Hence
v
0
. For k N, we need to solve for v
k
in
d
dr
_
r
2
dv
k
dr
_
= k(k + 1)v
k
.
Use the change of variable r = e
s
. Then e
s ds
dr
= 1 and
d
dr
=
d
ds
ds
dr
=
1
e
s
d
ds
.
Hence r
d
dr
=
d
ds
. Solving for m in the quadratic equation m
2
+m = k(k +1).
m
1
= k and m
2
= k 1. v
k
(e
s
) = e
ks
+ e
(k1)s
. v
k
(r) = r
k
+ r
k1
.
Since r
k1
blows up as r 0, we must have = 0. Thus, v
k
= r
k
.
Therefore, for each k = 0, 1, 2, . . .,
U
k
(r, , ) = a
k
r
k
P
k
(cos ).
The general solution is
U(r, , ) =

k=0
a
k
r
k
P
k
(cos ).
Since we have azimuthal symmetry, G(, ) = G(). To nd the constants,
we use U(1, , ) = G(), hence
G() =

k=0
a
k
P
k
(cos ).
Using the orthogonality of P
k
, we have
a
k
=
2k + 1
2
_

0
G()P
k
(cos ) sin d.
Now that we have sucient understanding of harmonic functions, solution
of homogeneous Laplace equation, with Dirichlet boundary conditions we
next attempt to solve the inhomogeneous Laplace equation (called Poisson
equation) with homogeneous boundary conditions.
CHAPTER 6. THE LAPLACIAN 150
6.8 Poisson Equation
We now wish to solve the Poisson equation, for any given f (under some
hypothesis) nd u such that
u = f in R
n
. (6.8.1)
Recall that we have already introduced the notion convolution of functions
(cf. Appendix F) while discussing C

properties of harmonic functions. We


also observed that the dierential operator can be accumulated on either side
of the convolution operation. Suppose there is a function K with the property
that K is the identity of the convolution operation, i.e., f K = f, then
we know that u := f K is a solution of (6.8.1).
Denition 6.8.1. We shall say a function K to be the fundamental solution
of the Laplacian, , if K is the identity with respect to the convolution
operation.
We caution that the above denition is not mathematically precise be-
cause we made no mention on what the function K could be and its dif-
ferentiability, even its existence is under question. We shall just take it as a
informal denition.
We note the necessry condition for any K to be a fundamental solution.
Observe that K is such that f K for all f in the given space of functions
in R
n
. In particular, one can choose f 1. Thus, the necessary condition
for a fundamental solution is 1 K = 1, i.e.,
_
R
n
K(x) dx = 1.
Equivalently, the necessary condition for K is
lim
r
_
Br(0)
K(x) dx = 1,
which by Gauss divergence theorem (all informally) means
lim
r
_
Sr(0)
K(y) (y) d
y
= 1.
CHAPTER 6. THE LAPLACIAN 151
6.8.1 Fundamental Solution of Laplacian
The invariance of Laplacian under rotation motivates us to look for a ra-
dial fundamental solution. Recall how Laplacian treats radial functions (cf.
Proposition 6.1.1) and, consequently, we have
Corollary 6.8.2. The function u(x) = ax + b solves u = 0 in R. For
n 2, if u is a radial function on R
n
then u = 0 on R
n
0 i
u(x) =
_
a + b ln [x[ if n = 2,
a +
b
2n
[x[
2n
if n 3
where a, b are some constants.
Proof. For radial functions u(x) = v(r) where r = [x[. Observe that u(x) =
0 i v

(r) +
(n1)
r
v

(r) = 0. Now, integrating both sides w.r.t r, we get


v

(r)
v

(r)
=
(1 n)
r
ln v

(r) = (1 n) ln r + ln b
v

(r) = br
(1n)
Integration both sides, once again, yields
v(r) =
_
b ln r + a if n = 2
b
2n
r
2n
+a if n ,= 2.
The reason to choose the domain of the Laplacian as R
n
0 is because
the operator involves a r in the denominator. However, for one dimensional
case we can let zero to be on the domain of Laplacian, since for n = 1, the
Laplace operator is unchanged. Thus, for n = 1, u(x) = a+bx is a harmonic
function in R
n
.
Note that as r 0, v(r) . Thus, u has a singularity at 0. In fact,
for any given vector x
0
R
n
, u(x x
0
) = 0 for all x R
n
x
0
. We shall
choose a, b such that for every sphere S
r
(0) about the origin, we have
_
Sr(0)
v

(r) d = 1.
CHAPTER 6. THE LAPLACIAN 152
Thus,
1 =
_
Sr(0)
v

(r) d =
_
b
r
(2r) for n = 2
br
1n
(r
n1

n
) for n 3.
This is possible only for the choice
b =
_
1
2
for n = 2
1
n
for n 3.
The constant a can be chosen arbitrarly, but to keep things simple, we choose
a 0 for n 2. For convention sake, we shall add minus () sign (notice
the minus sign in (6.8.1)).
Denition 6.8.3. For any xed x
0
R
n
We say K(x
0
, x), dened as
K(x
0
, x) :=
_

1
2
ln [x x
0
[ (n = 2)
|xx
0
|
2n
n(n2)
(n 3),
is the fundamental solution of at any given x
0
R
n
.
We end this section by emphasising that the notion of fundamental so-
lution has a precise denition in terms of the Dirac measure. The Dirac
measure, at a point x R
n
, is dened as,

x
(E) =
_
1 if x E
0 if x / E
for all measurable subsets E of the measure space R
n
. The Dirac measure
has the property that
_
E
d
x
= 1
if x E and zero if x / E. Also, for any integrable function f,
_
R
n
f(y) d
x
= f(x).
In this new set-up a fundamental solution K(x
0
, ) can be dened as the
solution corresponding to
x
0
, i.e.,
K(x
0
, x) =
x
0
in R
n
.
CHAPTER 6. THE LAPLACIAN 153
Note that the above equation, as such, makes no sense because the RHS is a
set-function taking subsets of R
n
as arguments, whereas K is a function on
R
n
. To give meaning to above equation, one needs to view
x
as a distribution
(introduced by L. Schwartz) and the equation should be interpreted in the
distributional derivative sense. The Dirac measure is the distributional limit
of the sequence of molliers,

, in the space of distributions.


6.8.2 Existence and Uniqueness
In this section, we shall give a formula for the solution of the Poisson equation
(6.8.1) in R
n
in terms of the fundamental solution.
Theorem 6.8.4. For any given f C
2
c
(R
n
), u := K f is a solution to the
Poisson equation (6.8.1).
Proof. By the property of convolution (cf. proof of Theorem F.0.10), we
know that D

u(x) = (K D

f)(x) for all [[ 2. Since f C


2
c
(R
n
), we
have u C
2
(R
n
). The diculty arises due to the singularity of K at the
CHAPTER 6. THE LAPLACIAN 154
origin. Consider, for any xed m > 0,
u(x) =
_
R
n
K(y)
x
f(x y) dy
=
_
Bm(0)
K(y)
x
f(x y) dy +
_
R
n
\Bm(0)
K(y)
x
f(x y) dy
=
_
Bm(0)
K(y)
x
f(x y) dy +
_
R
n
\Bm(0)
K(y)
y
f(x y) dy
=
_
Bm(0)
K(y)
x
f(x y) dy +
_
Sm(0)
K(y)
y
f(x y) d
y

_
R
n
\Bm(0)

y
K(y)
y
f(x y) dy (cf. Corollary D.0.7)
=
_
Bm(0)
K(y)
x
f(x y) dy +
_
Sm(0)
K(y)
y
f(x y) d
y
+
_
R
n
\Bm(0)

y
K(y)f(x y) dy

_
Sm(0)
f(x y)
y
K(y) d
y
(cf. Corollary D.0.7)
=
_
Bm(0)
K(y)
x
f(x y) dy +
_
Sm(0)
K(y)
y
f(x y) d
y

_
Sm(0)
f(x y)
y
K(y) d
y
:= I
m
(x) + J
m
(x) +K
m
(x).
But, due to the compact support of f, we have
[I
m
(x)[ |D
2
f|
,R
n
_
Bm(0)
[K(y)[ dy.
Thus, for n = 2,
[I
m
(x)[
m
2
2
_
1
2
+[ ln m[
_
|D
2
f|
,R
n
and for n 3, we have
[I
m
(x)[
m
2
2(n 2)
|D
2
f|
,R
n.
CHAPTER 6. THE LAPLACIAN 155
Hence, as m 0, [I
m
(x)[ 0. Similarly,
[J
m
(x)[
_
Sm(0)
[K(y)
y
f(x y) [ d
y
|f|
,R
n
_
Sm(0)
[K(y)[ d
y
.
Thus, for n = 2,
[J
m
(x)[ m[ ln m[|f|
,R
n
and for n 3, we have
[J
m
(x)[
m
(n 2)
|f|
,R
n.
Hence, as m 0, [J
m
(x)[ 0. Now, to tackle the last term K
m
(x), we note
that a simple computation yields that
y
K(y) =
1
n|y|
n
y. Since we are in
the m radius sphere [y[ = m. Also the unit vector outside of S
m
(0), as a
boundary of R
n
B
m
(0), is given by y/[y[ = y/m. Therefore,

y
K(y) =
1

n
m
n+1
y y =
1

n
m
n1
.
K
m
(x) =
_
Sm(0)
f(x y)
y
K(y) d
y
=
1

n
m
n1
_
Sm(0)
f(x y) d
y
=
1

n
m
n1
_
Sm(x)
f(y) d
y
Since f is continuous, for every > 0, there is a > 0 such that [f(x)
f(y)[ < whenever [x y[ < . When m 0, we can choose m such that
m < and for this m, we see that Now, consider
[K
m
(x) (f(x))[ =

f(x)
1

n
m
n1
_
Sm(x)
f(y) d
y

=
1

n
m
n1
_
Sm(x)
[f(x) f(y)[ d
y
< .
Thus, as m 0, K
m
(x) f(x). Hence, u solves (6.8.1).
CHAPTER 6. THE LAPLACIAN 156
Remark 6.8.5. Notice that in the proof above, we have used the Greens
identity eventhough our domain is not bounded (which is a hypothesis for
Greens identity). This can be justied by taking a ball bigger than B
m
(0)
and working in the annular region, and later letting the bigger ball approach
all of R
n
.
A natural question at this juncture is: Is every solution of the Poisson
equation (6.8.1) of the form K f. We answer this question in the following
theorem.
Theorem 6.8.6. Let f C
2
c
(R
n
) and n 3. If u is a solution of (6.8.1)
and u is bounded, then u has the form u(x) = (Kf)(x)+C, for any x R
n
,
where C is some constant.
Proof. We know that (cf. Theorem 6.8.4) u

(x) := (K f)(x) solves (6.8.1),


the Poisson equation in R
n
. Moreover, u

is bounded for n 3, since K(x)


0 as [x[ and f has compact support in R
n
. Also, since u is given to be
a bounded solution of (6.8.1), v := u u

is a bounded harmonic function.


Hence, by Liouvilles theorem, v is constant. Therefore u = u

+C, for some


constant C.
We turn our attention to studying Poisson equation in proper subsets of
R
n
. Let be an open bounded subset of R
n
with C
1
boundary .
Theorem 6.8.7 (Uniqueness). Let be an open bounded subset of R
n
. For
the Poisson equation u = f with one of Dirichlet, Robin or Mixed condi-
tions on , there exists at most one solution u C
2
() C
1
(). In the
Neumann problem two solutions dier by a constant.
Proof. Let u and v be solutions of the Poisson equation with same boundary
conditions on . Then w := u v is a harmonic function, w = 0, with
homogeneous boundary condition on . By Greens identity D.0.7, we have
_

[w[
2
dx =
_

w(w ) d.
For the Drichlet, Neumann and Mixed case, the RHS is zero. For the Robin
condition the RHS is negative,
_

w(w ) d = c
_

w
2
d 0.
CHAPTER 6. THE LAPLACIAN 157
Thus, in all the four boundary conditions
_

[w[
2
dx 0
and w = 0. Therefore, w = u v is constant in the connected components
of . In the case of Dirichlet, mixed and Robin the constant has to be zero,
by Maximum principle
5
. Thus, u = v in these three cases.
To begin with we shall focus on the study of Dirichlet problem. The
Dirichlet problem is stated as follows: Given f : R and g : R,
nd u : R such that
_
u = f in
u = g on
(6.8.2)
Lemma 6.8.8. Let f be bounded and locally Holder continuous
6
with expo-
nent 1 in . Then u := K f C
2
(), u = f in .
Theorem 6.8.9 (Existence). Let be a bounded domain with all boundary
points being regular w.r.t Laplacian. The classical Dirichlet problem (6.8.2) is
solvable (hence uniquely) for any bounded, locally Holder continuous function
f in and continuous function g on .
Proof. Recall that K is a fundamental solution of . Set w(x) := f K in
R
n
then w = f. Set v = u w. Then (6.8.2) is solvable i
_
v = 0 in
v = g w on
is solvable. The equation for v is solvable by Theorem 6.6.3.
6.8.3 Greens Function
We shall now attempt to solve the Poisson equation on a proper open subset
of R
n
. This is done via the Greens function. We begin by motivating the
Greens function. For any x , choose m > 0 such that B
m
(x) . Set

m
:= B
m
(x). Now applying the second identity of Corollary D.0.7 for
5
or, simply, from the fact that a non-zero c will contradict the continuous extension of
w to boundary.
6
H older continuous in each compact subset of
CHAPTER 6. THE LAPLACIAN 158
any u C
2
() and v
x
(y) = K(y x), the fundamental solution on R
n
x,
on the domain
m
, we get
_
m
u(y)
y
v
x
(y) dy

_
m
v
x
(y)
y
u(y) dy =
_
m
_
u(y)
v
x

(y) v
x
(y)
u(y)

_
d
y

_
m
v
x
(y)
y
u(y) dy =
_
m
_
u(y)
v
x

(y) v
x
(y)
u(y)

_
d
y

+
_
Bm(x)
=
_

+
_
Sm(x)
_
Bm(x)
v
x
(y)
y
u(y) dy

_
Sm(x)
u(y)
v
x

(y) d
y
+
_
Sm(x)
v
x
(y)
u(y)

d
y
=
_

_
u(y)
v
x

(y) v
x
(y)
u(y)

_
d
y
+
_

v
x
(y)
y
u(y) dy
I
m
(x) + K
m
(x) + J
m
(x) =
_

_
u(y)
K

(y x) K(y x)
u(y)

_
d
y
+
_

K(y x)
y
u(y) dy
The LHS is handled exactly as in the proof of Theorem 6.8.4, since u is
a continuous function on the compact set and is bounded. We repeat the
arguments below for completeness sake. Consider the term I
m
.
[I
m
(x)[ |D
2
u|
,
_
Bm(x)
[K(y x)[ dy.
Thus,
[I
m
(x)[
_
m
2
2
_
1
2
+[ ln m[
_
|D
2
u|
,
for n = 2
m
2
2(n2)
|D
2
u|
,
for n 3.
Hence, as m 0, [I
m
(x)[ 0. Next, consider the term K
m
(x). Note that

y
K(yx) =
1
n|yx|
n
(yx). Since we are in the m radius sphere [yx[ = m.
CHAPTER 6. THE LAPLACIAN 159
Also the unit vector inside of S
m
(x), as a boundary of B
m
(x), is given
by (y x)/[y x[ = (y x)/m. Therefore,

y
K(y x) =
1

n
m
n+1
(y x) (y x) =
1

n
m
n1
.
Thus,
K
m
(x) =
_
Sm(x)
u(y)
y
K(y x) d
y
=
1

n
m
n1
_
Sm(x)
u(y) d
y
Since u is continuous, for every > 0, there is a > 0 such that [u(x)u(y)[ <
whenever [x y[ < . When m 0, we can choose m such that m <
and for this m, we see that Now, consider
[K
m
(x) (u(x))[ =

u(x)
1

n
m
n1
_
Sm(x)
u(y) d
y

=
1

n
m
n1
_
Sm(x)
[u(x) u(y)[ d
y
< .
Thus, as m 0, K
m
(x) u(x). Finally, we consider the term J
m
(x),
[J
m
(x)[
_
Sm(x)
[K(y x)
y
u(y) [ d
y
|
y
u|
,
_
Sm(x)
[K(y x)[ d
y
.
Thus, for n = 2,
[J
m
(x)[
_
m[ ln m[|
y
u|
,
for n = 2
[J
m
(x)[
m
(n2)
|
y
u|
,
for n 3.
Hence, as m 0, [J
m
(x)[ 0. Therefore, letting m 0, we have the
identity
u(x) =
_

_
K(y x)
u(y)

u(y)
K

(y x)
_
d
y

K(yx)
y
u(y) dy
(6.8.3)
CHAPTER 6. THE LAPLACIAN 160
For the Dirichlet problem, u is known in and u is known on .
Thus, (6.8.3) gives an expression for the solution u, provided we know the
normal derivative
u(y)

along . But this quantity is usually an unknown for


Dirichlet problem. Thus, we wish to rewrite (6.8.3) such that the knowledge
of the normal derivative is not necessary. To do so, we introduce a function

x
(y), for a xed x , as the solution of the boundary-value problem,
_

x
(y) = 0 in

x
(y) = K(y x) on .
(6.8.4)
Now applying the second identity of Corollary D.0.7 for any u C
2
()
and v(y) =
x
(y), we get
_

_
u


x
u

_
d
y
=
_

(u
y

y
u) dy.
Therefore, substituting the following identity
_

K(y x)
u(y)

d
y
=
_

x
(y)
y
u(y) dy +
_

u(y)

x
(y)

d
y
in (6.8.3), we get
u(x) =
_

(
x
(y) K(y x))
y
udy +
_

u(
x
(y) K(y x)) d
y
.
The identity above motivates the denition of what is called the Greens
function.
Denition 6.8.10. For any given open subset R
n
and x, y such
that x ,= y, we dene the Greens function as
G(x, y) :=
x
(y) K(y x).
Rewriting (6.8.3) in terms of Greens function,we get
u(x) =
_

G(x, y)
y
u(y) dy +
_

u(y)
G(x, y)

d
y
.
Thus, in the arguments above we have proved the following theorem.
CHAPTER 6. THE LAPLACIAN 161
Theorem 6.8.11. Let be a bounded open subset of R
n
with C
1
boundary.
Also, given f C() and g C(). If u C
2
() solves the Dirichlet
problem (6.8.2), then u has the representation
u(x) =
_

G(x, y)f(y) dy +
_

g(y)
G(x, y)

d
y
. (6.8.5)
Observe that we have solved the Dirichlet problem (6.8.2) provided we
know the Greens function. The construction of Greens function depends
on the construction of
x
for every x . In other words, (6.8.2) is solved
if we can solve (6.8.4). Ironically, computing
x
is usually possible when
has simple geometry. We shall identify two simple cases of , half-space and
balll, where we can explicitly compute G.
The Greens function is the analogue of the fundamental solution K for
the boundary value problem. This is clear by observing that, for a xed
x , G satises (informally) the equation,
_
G(x, ) =
x
in
G(x, ) = 0 on ,
where
x
is the Dirac measure at x.
Theorem 6.8.12. For all x, y such that x ,= y, we have G(x, y) =
G(y, x), i.e., G is symmetric in x and y.
Proof. Let us x x, y . For a xed m > 0, set
m
= (B
m
(x) B
m
(y))
and applying Greens identity for v() := G(x, ) and w() := G(y, ), we get
_
m
_
v(z)
w(z)

w(z)
v(z)

_
d
z
=
_
m
v(z)
z
w(z) dz

_
m
w(z)
z
v(z) dz
_
m
_
v(z)
w(z)

w(z)
v(z)

_
d
z
= 0
_
Sm(x)
_
v
w

w
v

_
d
z
=
_
Sm(y)
_
w
v

v
w

_
d
z
J
m
(x) + K
m
(x) = J
m
(y) + K
m
(y).
CHAPTER 6. THE LAPLACIAN 162
[J
m
(x)[
_
Sm(x)
[v(z)
z
w(z) [ d
z
|w|
,
_
Sm(x)
[v(z)[ d
z
= |w|
,
_
Sm(x)
[
x
(z) K(z x)[ d
z
.
Thus, for n = 2,
[J
m
(x)[ (2m|
x
|
,
+ m[ ln m[) |w|
,
and for n 3, we have
[J
m
(x)[
_

n
m
n1
|
x
|
,
+
m
(n 2)
_
|w|
,
.
Hence, as m 0, [J
m
(x)[ 0. Now, consider the term K
m
(x),
K
m
(x) =
_
Sm(x)
w(z)
v(z)

d
z
=
_
Sm(x)
w(z)
K

(z x) d
z

_
Sm(x)
w(z)

x
(z)

d
z
.
The second term goes to zero by taking the sup-norm outside the integral.
To tackle the rst term, we note that
z
K(z x) =
1
n|zx|
n
(z x). Since
we are in the m radius sphere [z x[ = m. Also the unit vector outside of
S
m
(x), as a boundary of B
m
(x), is given by (zx)/[zx[ = (zx)/m.
Therefore,

z
K(z x) =
1

n
m
n+1
(z x) (z x) =
1

n
m
n1
.
_
Sm(x)
w(z)
z
K(z x) d
z
=
1

n
m
n1
_
Sm(x)
w(z) d
z
Since w is continuous in y, for every > 0, there is a > 0 such that
[w(z) w(x)[ < whenever [x z[ < . When m 0, we can choose m
such that m < and for this m, we see that Now, consider

n
m
n1
_
Sm(x)
w(z) d
z
w(x)

=
1

n
m
n1
_
Sm(x)
[w(z) w(x)[ d
z
< .
CHAPTER 6. THE LAPLACIAN 163
Thus, as m 0, K
m
(x) w(x). Arguing similarly, for J
m
(y) and K
m
(y),
we get G(y, x) = G(x, y).
6.8.4 Greens Function for half-space
In this section, we shall compute explicitly the Greens function for positive
half-space. Thus, we shall have
R
n
+
= x = (x
1
, x
2
, . . . , x
n
) R
n
[ x
n
> 0
and
R
n
+
= x = (x
1
, x
2
, . . . , x
n
R
n
[ x
n
= 0.
To compute the Greens function, we shall use the method of reection.
The reection technique ensures that the points on the boundary (along
which the reection is done) remains unchanged to respect the imposed
Dirichlet condition.
Denition 6.8.13. For any x = (x
1
, x
2
, . . . , x
n
) R
n
+
, we dene its reec-
tion along its boundary R
n1
as x

= (x
1
, x
2
, . . . , x
n
).
It is obvious from the above denition that, for any y R
n
+
, [y x

[ =
[y x[. Given a xed x R
n
+
, we need to nd a harmonic function
x
in
R
n
+
, as in (6.8.4). But K( x) is harmonic in R
n
+
x. Thus, we use the
method of reection to shift the singularity of K from R
n
+
to the negative
half-space and dene

x
(y) = K(y x

).
By denition,
x
is harmonic in R
n
+
and on the boundary
x
(y) = K(y x).
Therefore, we dene the Greens function to be G(x, y) = K(y x

) K(y
x), for all x, y R
n
+
and x ,= y. It now only remains to compute the normal
derivative of G. Recall that
y
K(y x) =
1
n|yx|
n
(y x). Thus,

y
G(x, y) =
1

n
_
y x

[y x

[
n

y x
[y x[
n
_
Therefore, when y R
n
+
, we have

y
G(x, y) =
1

n
[y x[
n
(x x

).
CHAPTER 6. THE LAPLACIAN 164
Since the outward unit normal of R
n
+
is = (0, 0, . . . , 0, 1), we get

y
G(x, y) =
2x
n

n
[y x[
n
.
Denition 6.8.14. For all x R
n
+
and y R
n
+
, the map
P(x, y) :=
2x
n

n
[y x[
n
is called the Poisson kernel for R
n
+
.
Now substituing for G in (6.8.5), we get the Poisson formula for u,
u(x) =
_
R
n
+
[K(y x) K(y x

)]f(y) dy +
2x
n

n
_
R
n
+
g(y)
[y x[
n
d
y
. (6.8.6)
It now remains to show that the u as dened above is, indeed, a solution of
(6.8.2) for R
n
+
.
Exercise 30. Let f C(R
n
+
) be given. Let g C(R
n1
) be bounded. Then
u as given in (6.8.6) is in C
2
(R
n
+
) and solves (6.8.2).
6.8.5 Greens Function for a disk
In this section, we shall compute explicitly the Greens function for a ball of
radius r > 0 and centred at a R
n
, B
r
(a). As usual, we denote the surface
of the disk as S
r
(a), the circle of radius r centred at a. We, once again, use
the method of reection but, this time reected along the boundary of the
disk.
Denition 6.8.15. For any x R
n
a, we dene its reection along the
circle S
r
(a) as x

=
r
2
(xa)
|xa|
2
+ a.
The idea behind reection is clear for the unit disk, i.e., when a = 0 and
r = 1, as x

=
x
|x|
2
. The above denition is just the shift of origin to a and
dilating the unit disk by r.
CHAPTER 6. THE LAPLACIAN 165
Now, for any y S
r
(a) and x ,= a, consider
[y x

[
2
= [y a[
2
2(y a) (x

a) +[x

a[
2
= r
2
2r
2
(y a)
_
x a
[x a[
2
_
+

r
2
(x a)
[x a[
2

2
=
r
2
[x a[
2
([x a[
2
2(y a) (x a) +r
2
)
=
r
2
[x a[
2
([x a[
2
2(y a) (x a) +[y a[
2
)
=
r
2
[x a[
2
[y x[
2
Therefore,
|xa|
r
[yx

[ = [yx[ for all y S


r
(a). For each xed x B
r
(a), we
need to nd a harmonic function
x
in B
r
(a) solving (6.8.4). Since K( x)
is harmonic in B
r
(a) x, we use the method of reection to shift the
singularity of K at x to the complement of B
r
(a). Thus, we dene

x
(y) = K
_
[x a[
r
(y x

)
_
x ,= a.
For n 3, K
_
|xa|
r
(y x

)
_
=
|xa|
2n
r
2n
K(y x

). Thus, for n 3,
x
solves
(6.8.4), for x ,= a. For n = 2,
K
_
[x a[
r
(y x

)
_
=
1
2
ln
_
[x a[
r
_
+ K(y x

).
Hence
x
solves (6.8.4) for n = 2. Note that we are yet to identify a harmonic
function
a
corresponding to x = a. We do this by setting
a
to be the
constant function

a
(y) :=
_

1
2
ln r (n = 2)
r
2n
n(n2)
(n 3).
Thus,
a
is harmonic and solves (6.8.4) for x = a. Therefore, we dene the
Greens function to be
G(x, y) := K
_
[x a[
r
(y x

)
_
K(yx) x, y B
r
(a), x ,= a and x ,= y
CHAPTER 6. THE LAPLACIAN 166
and
G(a, y) :=
_

1
2
ln
_
r
|ya|
_
(n = 2)
1
n(n2)
(r
2n
[y a[
2n
) (n 3).
We shall now compute the normal derivative of G. Recall that

y
K(y x) =
1

n
[y x[
n
(y x)
and one can compute
y
K
_
|xa|
r
(y x

)
_
=
|xa|
2n
r
2n
n|yx

|
n
(yx

). Therefore,

y
G(x, y) =
1

n
_
[x a[
2n
(y x

)
r
2n
[y x

[
n

y x
[y x[
n
_
.
If y S
r
(a), we have

y
G(x, y) =
1

n
[y x[
n
_
[x a[
2
r
2
(y x

) (y x)
_
=
1

n
[y x[
n
_
[x a[
2
r
2
1
_
(y a)
Since the outward unit normal at any point y S
r
(a) is
1
r
(y a), we have

y
G(x, y) =
1

n
[y x[
n
_
[x a[
2
r
2
1
_
n

i=1
1
r
(y
i
a
i
)
2
=
r

n
[y x[
n
_
[x a[
2
r
2
1
_
.
Denition 6.8.16. For all x B
r
(a) and y S
r
(a), the map
P(x, y) :=
r
2
[x a[
2
r
n
[y x[
n
is called the Poisson kernel for B
r
(a).
Now substituing for G in (6.8.5), we get the Poisson formula for u,
u(x) =
_
Br(a)
G(x, y)f(y) dy +
r
2
[x a[
2
r
n
_
Sr(a)
g(y)
[y x[
n
d
y
. (6.8.7)
It now remains to show that the u as dened above is, indeed, a solution of
(6.8.2) for B
r
(a).
Exercise 31. Let f C(B
r
(a)) be given. Let g C(S
r
(a)) be bounded.
Then u as given in (6.8.7) is in C
2
(B
r
(a)) and solves (6.8.2).
CHAPTER 6. THE LAPLACIAN 167
6.8.6 Conformal Mapping and Greens Function
In two dimensions, the Greens function has a nice connection with conformal
mapping. Let w = f(z) be a conformal mapping from an open domain
(connected) R
2
onto the interior of the unit circle. The Greens function
of is
G(z, z
0
) =
1
2
ln

1 f(z)f(z
0
)
f(z) f(z
0
)

where z = x
1
+ ix
2
and z
0
= y
1
+ iy
2
.
6.8.7 Dirichlet Principle
The Dirichlet principle (formulated, independently by Gauss, Lord Kelvin
and Dirichlet) states that the solution of the Dirichlet problem minimizes
the corresponding energy functional.
Let be an open bounded subset of R
n
with C
1
boundary and let
f : R and g : R be given. For convenience, recall the Dirichlet
problem ((6.8.2)),
_
u = f in
u = g on .
Any solution u of (6.8.2) is in V = v C
2
() [ v = g on . The
energy functional J : V R is dened as
J(v) :=
1
2
_

[v[
2
dx
_

fv dx
Theorem 6.8.17 (Dirichlets principle). A C
2
() function u solves (6.8.2)
i u minimises the functional J on V , i.e.,
J(u) J(v) v V.
Proof. Let u C
2
() be a solution of (6.8.2). For any v V , we multiply
CHAPTER 6. THE LAPLACIAN 168
both sides of (6.8.2) by u v and integrating we get,
_

(u)(u v) dx =
_

f(u v) dx
_

u (u v) dx =
_

f(u v) dx
_

_
[u[
2
fu
_
dx =
_

(u v fv) dx

[u v[
_

fv dx

1
2
_

_
[u[
2
+[v[
2
_
dx
_

fv dx
(since 2ab a
2
+ b
2
)
J(u) J(v).
Thus, u minimises J in V . Conversely, let u minimise J in V . Thus,
J(u) J(v) v V
J(u) J(u + t) (for any C
2
() such that = 0 on )
0
1
t
(J(u + t) J(u))
0
1
t
_
1
2
_

_
t
2
[[
2
+ 2t u
_
dx t
_

fdx
_
Taking limit t 0 both sides, we get
0
_

udx
_

fdx C
2
() s.t. = 0 on .
Choosing in place of we get the reverse inequality, and we have equality
in the above. Thus,
_

u dx =
_

fdx C
2
() s.t. = 0 on
_

(u f)dx = 0 C
2
() s.t. = 0 on .
Thus u solves (6.8.2).
CHAPTER 6. THE LAPLACIAN 169
6.9 Neumann Boundary Condition
The Neumann problem is stated as follows: Given f : R and g :
R, nd u : R such that
_
u = f in
u

= g on
(6.9.1)
where
u

:= u and = (
1
, . . . ,
n
) is the outward pointing unit normal
vector eld of . Thus, the boundary imposed is called the Neumann
boundary condition. The solution of a Neumann problem is not necessarily
unique. If u is any solution of (6.9.1), then u +c for any constant c is also a
solution of (6.9.1). More generally, for any v such that v is constant on the
connected components of , u + v is a solution of (6.9.1).
CHAPTER 6. THE LAPLACIAN 170
Appendices
171
Appendix A
The Gamma Function
The gamma function : (0, ) R is dened as,
(x) =
_

0
e
t
t
x1
dt x (0, ).
Note that the gamma function is dened as an improper integral and its
existence has to be justied. Observe that for a xed x > 0,
[e
t
t
x1
[ = e
t
t
x1
t
x1
t > 0,
since for t > 0, [e
t
[ 1. Now, since
_
1
0
t
x1
dt exists, we have by comparison
test the existence of the integral
_
1
0
e
t
t
x1
dt. Now, for t , e
t
t
x1
0
and hence the is a constant C > 0 such that
t
x1
e
t
C/t
2
t 1.
Since
_

1
1/t
2
dt exists, we again have using comparison test the existence of
the integral
_

1
e
t
t
x1
dt. In fact, the gamma function can be dened for
any complex number z C such that Re(z) > 0.
It is worth noting that the gamma function generalises the notion of
factorial of positive integers. This would be the rst property we shall prove.
Exercise 32. Show that (x + 1) = x(x). In particular, for any positive
integer n, (n + 1) = n!. Also, show that (1) = 1 and (1/2) =

.
Further, for any positive integer n,
(n + 1/2) = (n 1/2)(n 3/2) . . . (1/2)

.
(Hint: Use integration by parts and change of variable).
173
APPENDIX A. THE GAMMA FUNCTION 174
Exercise 33. Show that is continuous on (0, ).
Exercise 34. Show that the logarithm of is convex on (0, ).
We shall now show that is the only possible generalisation of the notion
of factorial of positive integers satisfying the above properties.
Theorem A.0.1. Let f be positive and continuous on (0, ) and let log f
be convex on (0, ). Also, let f satisfy the recursive equation
f(x + 1) = xf(x) x > 0
and f(1) = 1, then f(x) = (x) for all x > 0.
Appendix B
Normal Vector of a Surface
Let S(x, y, z) = 0 be the equation of a surface S in R
3
. Let us a x a
point p
0
= (x
0
, y
0
, z
0
) S. We need to nd the normal vector at p
0
for
the surface S. Let us x an arbitrary curve C lying on the surface passing
through the point p
0
. Let the parametrized form of the curve C be given as
r(t) = (x(t), y(t), z(t)) such that r(t
0
) = p
0
. Since the curve C r(t) lies
on the surface for all t, we have S(r(t)) = 0. Thus, S(x(t), y(t), z(t)) = 0.
Dierentiating w.r.t t (using chain rule), we get
S
x
dx(t)
dt
+
S
y
dy(t)
dt
+
S
z
dz(t)
dt
= 0
(S
x
, S
y
, S
z
) (x

(t), y

(t), z

(t)) = 0
S(r(t)) r

(t) = 0.
In particular, the above computation is true for the point p
0
. Since r

(t
0
)
is the slope of the tangent at t
0
to the curve C, we see that the vector S(p
0
)
is perpendicular to the tangent vector at p
0
. Since this argument is true for
any curve that passes through p
0
. We have that S(p
0
) is normal vector to
the tangent plane at p
0
. If, in particular, the equation of the surface is given
as S(x, y, z) = u(x, y) z, for some u : R
2
R, then
S(p
0
) = (S
x
(p
0
), S
y
(p
0
), S
z
(p
0
))
= (u
x
(x
0
, y
0
), u
y
(x
0
, y
0
), 1) = (u(x
0
, y
0
), 1).
175
APPENDIX B. NORMAL VECTOR OF A SURFACE 176
Appendix C
Implicit Function Theorem
Theorem C.0.2 (Implicit Function Theorem). Let R
m
R
n
be an open
subset and f : R
n
such that f is continuously dierentiable (C
1
) in .
Let (x
0
, y
0
) be such that f(x
0
, y
0
) = 0 and the n n matrix
D
y
f(x
0
, y
0
) :=
_
_
_
f
1
y
1
(x
0
, y
0
)
f
1
yn
(x
0
, y
0
)
.
.
.
.
.
.
.
.
.
fn
y
1
(x
0
, y
0
)
fn
yn
(x
0
, y
0
)
_
_
_
is non-singular, then there is a neighbourhood U R
m
of x
0
and a unique
function g : U R
n
such that g(x
0
) = y
0
and, for all x U, f(x, g(x)) = 0.
Further g is continuously dierentiable in U.
177
APPENDIX C. IMPLICIT FUNCTION THEOREM 178
Appendix D
Divergence Theorem
Denition D.0.3. For an open set R
n
we say that its boundary is
C
k
(k 1), if for every point x , there is a r > 0 and a C
k
dieomor-
phism : B
r
(x) B
1
(0) ( i.e.
1
exists and both and
1
are k-times
continuously dierentiable) such that
1. ( B
r
(x)) B
1
(0) x R
n
[ x
n
= 0 and
2. ( B
r
(x)) B
1
(0) x R
n
[ x
n
> 0
We say is C

if is C
k
for all k = 1, 2, . . . and is analytic if is
analytic.
Equivalently, a workable denition of C
k
boundary would be the follow-
ing: if for every point x , there exists a neighbourhood B
x
of x and a
C
k
function : R
n1
R such that
B
x
= x B
x
[ x
n
> (x
1
, x
2
, . . . , x
n1
).
The divergence of a vector eld is the measure of the magnitude (outgoing
nature) of all source (of the vector eld) and absorption in the region. The
divergence theorem was discovered by C. F. Gauss in 1813
1
which relates
the outward ow (ux) of a vector eld through a closed surface to the
behaviour of the vector eld inside the surface (sum of all its source and
sink). The divergence theorem is, in fact, the mathematical formulation
of the conservation law.
1
J. L. Lagrange seems to have discovered this, before Gauss, in 1762
179
APPENDIX D. DIVERGENCE THEOREM 180
Theorem D.0.4. Let be an open bounded subset of R
n
with C
1
boundary.
If v C
1
() then
_

v
x
i
dx =
_

v
i
d
where = (
1
, . . . ,
n
) is the outward pointing unit normal vector eld and
d is the surface measure of .
The domain need not be bounded provided [v[ and

v
x
i

decays as
[x[ . The eld of geometric measure theory attempts to identify the
precise condition on and v for which divergence theorem or integration
by parts hold.
Corollary D.0.5 (Integration by parts). Let be an open bounded subset
of R
n
with C
1
boundary. If u, v C
1
() then
_

u
v
x
i
dx +
_

v
u
x
i
dx =
_

uv
i
d.
Theorem D.0.6 (Gauss). Let be an open bounded subset of R
n
with C
1
boundary. Given a vector eld V = (v
1
, . . . , v
n
) on such that v
i
C
1
()
for all 1 i n, then
_

V dx =
_

V d. (D.0.1)
Corollary D.0.7 (Greens Identities). Let be an open bounded subset of
R
n
with C
1
boundary. Let u, v C
2
(), then
(i)
_

(vu +v u) dx =
_

v
u

d,
where
u

:= u and := .
(ii)
_

(vu uv) dx =
_

_
v
u

u
v

_
d.
Proof. Apply divergence theorem to V = vu to get the rst formula. To get
second formula apply divergence theorem for both V = vu and V = uv
and subtract one from the other.
Appendix E
Surface Area and Volume of
Disk in R
n
Theorem E.0.8 (Polar coordinates). Let f : R
n
R be continuous and
integrable. Then
_
R
n
f(x) dx =
_

0
__
Sr(a)
f(y) d
y
_
dr
for each a R
n
. In particular, for each r > 0,
d
dr
__
Br(a)
f(x) dx
_
=
_
Sr(a)
f(y) d
y
.
Theorem E.0.9. Prove that
_
R
n
e
|x|
2
dx = 1.
Further, prove that the surface area
n
of S
1
(0) in R
n
is
2
n/2
(n/2)
and the volume of the ball B
1
(0) in R
n
is
n
/n. Consequently, for any x R
n
and the r > 0, the surface area of S
r
(x) is r
n1

n
and the volume of B
r
(x)
is r
n

n
/n.
181
APPENDIX E. SURFACE AREA AND VOLUME OF DISK IN R
N
182
Proof. We rst observe that
e
|x|
2
= e
(

n
i=1
x
2
i
)
=
n

i=1
e
x
2
i
.
Therefore,
I
n
:=
_
R
n
e
|x|
2
dx =
_
R
n
n

i=1
e
x
2
i
dx
=
n

i=1
_
R
e
t
2
dt
=
__
R
e
t
2
dt
_
n
= (I
1
)
n
_
R
n
e
|x|
2
dx =
__
R
e
t
2
dt
_
2(n/2)
=
_
(I
1
)
2
_
n/2
= (I
2
)
n/2
=
__
R
2
e
|y|
2
dy
_
n/2
=
__
2
0
_

0
e
|y|
2
dy
_
n/2
=
__
2
0
_

0
e
r
2
r dr d
_
n/2
(since jacobian is r)
=
_
2
_

0
e
r
2
r dr
_
n/2
=
_

_

0
e
s
ds
_
n/2
(by setting r
2
= s)
=
__

0
e
q
dq
_
n/2
(by setting s = q)
= ((1))
n/2
= 1.
Let
n
denote the surface area of the unit sphere S
1
(0) in R
n
, i.e.,

n
=
_
S
1
(0)
d,
APPENDIX E. SURFACE AREA AND VOLUME OF DISK IN R
N
183
where d is the n 1-dimensional surface measure. Now, consider
1 =
_
R
n
e
|x|
2
dx
=
_
S
1
(0)
_

0
e
r
2
r
n1
dr d
=
n
_

0
e
r
2
r
n1
dr
=

n
2
n/2
_

0
e
s
s
(n/2)1
ds (by setting s = r
2
)
=

n
(n/2)
2
n/2
.
Thus,
n
=
2
n/2
(n/2)
. We shall now compute the volume of the disk B
1
(0).
Consider,
_
B
1
(0)
dx =
n
_
1
0
r
n1
dr =

n
n
.
For any x R
n
and r > 0, we observe by the shifting of origin that the
surface area of S
r
(x) is same as the surface area of S
r
(0). Let S
r
(0) = s
R
n
[ [s[ = r. Now
_
Sr(0)
d
s
=
_
S
1
(0)
r
n1
d
t
= r
n1

n
,
where t = s/r. Thus, the surface area of S
r
(x) is r
n1

n
. Similarly, volume
of a disk B
r
(x) is r
n

n
/n.
APPENDIX E. SURFACE AREA AND VOLUME OF DISK IN R
N
184
Appendix F
Molliers and Convolution
Exercise 35. Show that the Cauchys exponential function, f : R R dened
as
f(x) =
_
exp(x
2
) if x > 0
0 if x 0,
is innitely dierentiable, i.e., is in C

(R).
Using the above Cauchys exponential function, one can construct func-
tions in C

c
(R
n
).
Exercise 36. For x = (x
1
, . . . , x
n
) R
n
, show that : R
n
R dened as
(x) =
_
cexp(
1
1|x|
2
) if [x[ < 1
0 if [x[ 1,
is in C

c
(R
n
) with supp() = B
1
(0), ball with centre 0 and radius 1, where
c
1
=
_
|x|1
exp(
1
1|x|
2
) dx.
Thus, one can introduce a sequence of functions in C

c
(R
n
), called mol-
liers. For > 0, we set

(x) =
_
c
n
exp(

2

2
|x|
2
) if [x[ <
0 if [x[ ,
(F.0.1)
Equivalently,

(x) =
n
(x/).
Exercise 37. Show that

0 and
_
R
n

(x) dx = 1 and is in C

c
(R
n
) with
support in B

(0).
185
APPENDIX F. MOLLIFIERS AND CONVOLUTION 186
Let f, g L
1
(R
n
). Their convolution f g is dened as, for x R
n
,
(f g)(x) =
_
R
n
f(x y)g(y) dy.
The integral on RHS is well-dened, since by Fubinis Theorem and the
translation invariance of the Lebesgue measure, we have
_
R
n
_
R
n
[f(xy)g(y)[ dxdy =
_
R
n
[g(y)[ dy
_
R
n
[f(xy)[ dx = |g|
1
|f|
1
< .
Thus, for a xed x, f(x y)g(y) L
1
(R
n
).
Theorem F.0.10. Let R
n
be an open subset of R
n
and let

:= x [ dist(x, ) > .
If u : R is locally integrable, i.e., for every compact subset K ,
_
K
[u[ < +, then u

:=

u is in C

).
Proof. Fix x

. Consider
u

(x + he
i
) u

(x)
h
=
1
h
_

(x +he
i
y)

(x y)] u(y) dy
=
_
B(x)
1
h
[

(x + he
i
y)

(x y)]u(y) dy.
Now, taking lim
h0
both sides, we get
u

(x)
x
i
= lim
h0
_
B(x)
1
h
[

(x + he
i
y)

(x y)]u(y) dy
=
_
B(x)

(x y)
x
i
u(y) dy
(interchange of limits is due to the uniform convergence)
=
_

(x y)
x
i
u(y) dy =

x
i
u.
Similarly, one can show that, for any tuple , D

(x) = (D

u)(x).
Thus, u

).
Appendix G
Duhamels Principle
Consider the rst order inhomogeneous ODE
_
x

(t) + ax(t) = f(t) in (0, )


x(0) = x
0
.
(G.0.1)
Multiplying the integration factor e
at
both sides, we get
[e
at
x(t)]

= e
at
f(t)
and
x(t) = e
at
_
t
0
e
as
f(s) ds + ce
at
.
Using the initial condition x(0) = x
0
, we get
x(t) = x
0
e
at
+
_
t
0
e
a(st)
f(s) ds.
Notice that x
0
e
at
is a solution of the homogeneous ODE. Thus, the solution
x(t) can be given as
x(t) = S(t)x
0
+
_
t
0
S(t s)f(s) ds
where S(t) is a solution operator of the linear equation, given as S(t) = e
at
.
Consider the second order inhomogeneous ODE
_
_
_
x

(t) + a
2
x(t) = f(t) in (0, )
x(0) = x
0
x

(0) = x
1
.
(G.0.2)
187
APPENDIX G. DUHAMELS PRINCIPLE 188
We introduce a new function y such that
x

(t) = ay(t).
Then
y

(t) =
f(t)
a
ax(t)
and the second order ODE can be rewritten as a system of rst order ODE
X

(t) + AX(t) = F(t)


where X = (x, y), F = (0, f/a) and
A =
_
0 a
a 0
_
with the initial condition X
0
:= X(0) = (x
0
, x
1
/a). We introduce the matrix
exponential e
At
=

n=1
(At)
n
n!
. Then, multiplying the integration factor e
At
both sides, we get
[e
At
X(t)]

= e
At
F(t)
and
X(t) = X
0
e
At
+
_
t
0
e
A(st)
F(s) ds.
Notice that X
0
e
At
is a solution of the homogeneous ODE. Thus, the solution
X(t) can be given as
X(t) = S(t)X
0
+
_
t
0
S(t s)F(s) ds
where S(t) is a solution operator of the linear equation, given as S(t) = e
At
.
Bibliography
189
BIBLIOGRAPHY 190
Index
ball
surface area of, 181
volume of, 181
characteristic equations, 25
convolution, 131, 186
Dirac measure, 152, 161
directional derivative, 2
Dirichlet Problem, 157
elliptic PDE, 63, 66
energy functional, 167
function
Cauchys exponential, 185
fundamental solution, 150, 152
gamma function, 173
Gauss divergence result, 180
gradient, 2
Greens function, 160
Greens identities, 180
Hadamard, 9
Hadamards wellposedness, 9
harmonic function, 12, 124
Harnack inequality, 137
Hessian matrix, 3, 62
hyperbolic PDE, 63, 65
integral curve, 22
integral surface, 23
Laplace operator, 3, 12, 117
Liouvilles theorem, 137
maximum principle
strong, 133
weak, 134
mean value property, 130
method of characteristics, 24
method of reection, 163, 164
molliers, 185
Monge cone, 23
Neumann boundary condition, 169
Neumann Problem, 169
parabolic PDE, 63, 65
Poisson equation, 12, 150
Poisson formula, 164, 166
Poisson kernel, 164, 166
radial function, 118
transport equation, 20
tricomi equation, 68
191

Você também pode gostar