Math 180 Spring 2018 Notes
Math 180 Spring 2018 Notes
Math 180 Spring 2018 Notes
Adolfo J. Rumbos
c Draft date April 10, 2018
2
Contents
1 Preface 5
3 Classification of PDEs 43
3.1 Linearity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.2 Second Order PDEs . . . . . . . . . . . . . . . . . . . . . . . . . 46
4 Solving PDEs 51
4.1 Using Characteristic Curves . . . . . . . . . . . . . . . . . . . . . 51
4.1.1 Solving the One–Dimensional Wave Equation . . . . . . . 52
4.1.2 Solving First–Order PDEs . . . . . . . . . . . . . . . . . . 57
4.2 Using Symmetry to Solve PDEs . . . . . . . . . . . . . . . . . . . 64
4.2.1 Radially Symmetric Solutions . . . . . . . . . . . . . . . . 64
4.2.2 Dilation-Invariant Solutions . . . . . . . . . . . . . . . . . 69
4.2.3 Solving the Diffusion Equation . . . . . . . . . . . . . . . 73
3
4 CONTENTS
C Kernels 197
C.1 The Dirichlet Kernel . . . . . . . . . . . . . . . . . . . . . . . . . 197
Chapter 1
Preface
5
6 CHAPTER 1. PREFACE
Chapter 2
ut − kuxx = 0, (2.2)
7
8 CHAPTER 2. MODELING WITH PDES
has to be accounted by how much of the substance goes into the system and
how much of it goes out of the system. For the case in which Q is also assumed
to be differentiable (again, this is a mathematical assumption that would need
some justification), the conservation principle can be succinctly stated as
dQ
= Rate of Q in − Rate of Q out. (2.3)
dt
In the cases to be considered in this section, the conservation principle in (2.3)
might lead to a differential equation, or a system of differential equations, and
so the theory of differential equations will be used to help in the analysis of the
model.
In the derivation of the equations governing fluid motion, we will have the
opportunity to apply the conservation principle in (2.3) several times.
Suppose we are following the motion of a fluid in some region R in three–
dimensional space; see Figure 2.1.1. We assume that the fluid is a continuum
(x, y, z)
B
with density function ρ(x, y, x, t), in units of mass per unit volume, so that the
mass of a fluid element of volume dV = dxdydz around a point (x, y, z) at time
t is, approximately,
ρ(x, y, x, t)dV,
where dV denotes the volume of the fluid element. It then follows that the mass
of fluid contained in a subregion B of R (see Figure 2.1.1) at time t is given by
ZZZ
M (B, t) = ρ(x, y, x, t) dV. (2.4)
B
If we assume that the components of the velocity field ~u are differentiable with
continuous derivatives throughout the region R and for all times t (i.e., ~u is a
C 1 vector field), then a solution to the system of ordinary differential equations
in (2.5) subject to the initial conditions in (2.6) is guaranteed to exist over
some maximal interval of time containing 0. The solution (x(t), y(t), y(t)) of
the system in (2.5) subject to the initial conditions in (2.6) defines a path in
space,
t 7→ (x(t), y(t), y(t)),
for t in the maximal interval of existence, which describes the motion of a fluid
element located at (x, y, z) at time t = 0. The path traced by the fluid element
as it moves in time is called a pathline; Figure 2.1.1 shows what a pathline
through (x, y, z) might look like. If we knew the velocity field at any point
in space and at any time, we could compute the pathline through (x, y, z) by
integrating the equations in (2.5) and imposing the initial conditions in (2.6):
Z t
x(t) = x + u1 (x(τ ), y(τ ), z(τ ), τ ) dτ ;
0
Z t
y(t) = y+ u2 (x(τ ), y(τ ), z(τ ), τ ) dτ ; (2.7)
0
Z t
z(t) = z+ u3 (x(τ ), y(τ ), z(τ ), τ ) dτ.
0
However, the velocity field is usually not known, and we need to do more mod-
eling to find equations involving u1 , u2 and u3 that we hope we can solve.
By the principle of conservation of mass, the rate of change in the mass of fluid
contained in B has to be accounted for by how much fluid is entering the region
and how much is leaving per unit of time:
dMB
= Rate of fluid into B − Rate of fluid out of B. (2.9)
dt
The equation in (2.9) is an instance of the conservation principle in (2.3).
If we assume that ρ is a C 1 function in R, we can compute the left–hand
side of the equation by differentiating under the integral in (2.8):
ZZZ
dMB ∂ρ
= (x, y, x, t) dV. (2.10)
dt B ∂t
Next, we compute the right–hand side of the expression in (2.9). Let ~n denote
the unit vector normal to the boundary, ∂B, of the region B pointing outward.
The outward unit normal, ~n(x, y, z), to the boundary of B is guaranteed to exist
at every point (x, y, z) ∈ ∂B if we assume that ∂B is a smooth surface. Then,
the rate of fluid passing through an element of area, dA, on the surface ∂B can
be expressed, approximately, as
ρ ~u · ~n dA, (2.11)
where ~u · ~n denotes the dot product of ~u and ~n. Note that the expression in
(2.11) is in units of mass per unit of time. Integrating the expression in (2.11)
over the boundary of B yields the net flux of mass across the surface ∂B,
ZZ
ρ ~u · ~n dA. (2.12)
∂B
Since the outward unit normal, ~n, points away from the region B, the expression
in (2.12) measures the flux of fluid away from the region B, if it is positive; if
the expression in (2.12) is negative, it measures the net amount of fluid per unit
time that enters B. We can therefore write the conservation principle in (2.9)
as ZZ
dMB
=− ρ ~u · ~n dA. (2.13)
dt ∂B
To understand the reason for the minus sign on the right–hand side of the
expression in (2.13), observe that a net increase in the amount of fluid in the
region B, which yields a positive sign for the derivative in the left–hand side of
(2.13), corresponds to a net amount of fluid flowing into the region B across the
boundary ∂B.
2.1. MODELING FLUID FLOW 11
Since we are assuming that the boundary of B is smooth, we can apply the
Divergence Theorem to rewrite the integral in the right–hand side of (2.13) as
follows: ZZ ZZZ
ρ ~u · ~n dA = ∇ · (ρ~u) dV, (2.14)
∂B B
where ∇ · (ρ~u) denotes the divergence of the vector field ρ~u; that is,
∂ ∂ ∂
∇ · (ρ~u) = (ρu1 ) + (ρu2 ) + (ρu3 ). (2.15)
∂x ∂y ∂z
In view of (2.10) and (2.14), we see that we can rewrite the conservation
equation in (2.13) as
ZZZ ZZZ
∂ρ
dV = − ∇ · (ρ~u) dV,
B ∂t B
or ZZZ
∂ρ
+ ∇ · (ρ~u) dV = 0. (2.16)
B ∂t
If we assume that the vector field ~u and the scalar field ρ are C 1 functions over
R and for all times t, then the fact that (2.16) holds true for any subregion B of
R with smooth boundary implies that integrand on the left–hand side of (2.16)
must be 0 over R and for all t; that is,
∂ρ
+ ∇ · (ρ~u) = 0, in R and for all t. (2.17)
∂t
The equation in (2.17) is an example of a partial differential equation (PDE)
involving the functions ρ, u1 , u2 and u3 ; in fact, using the definition of divergence
(see (2.15)), the PDE in (2.17) can be rewritten as
∂ρ ∂ ∂ ∂
+ (ρu1 ) + (ρu2 ) + (ρu3 ) = 0. (2.18)
∂t ∂x ∂y ∂z
The PDE in (2.17) is called the continuity equation and it expresses the
conservation principle for a quantity of density ρ that flows according to a veloc-
ity field ~u in some region in space. For one–dimensional flow with linear density
ρ(x, t) and scalar velocity field u(x, t), for x ∈ R and t ∈ R, the continuity
equation reads
∂ρ ∂
+ (ρu) = 0; (2.19)
∂t ∂x
see (2.18). The equation in (2.19) is an example of a first order PDE because
the first derivatives of the functions ρ and u are involved. As it stands, the PDE
in (2.19) involves two unknown functions, the density, ρ, and the velocity, u.
Thus, we will need one more relation or equations in order for us to even begin
to solve the problem posed by the modeling that led to the PDE in (2.19). An
interesting example is provided by the following application to modeling traffic
flow.
12 CHAPTER 2. MODELING WITH PDES
u1r - u2 r -
postulate a traffic density, ρ(x, t), measured in units of number of cars per unit
length of road at location x and time t. We interpret ρ(x, t) as follows: Consider
a section of the road from x to x + ∆x at time t. Let ∆N ([x, x + ∆x], t) denote
the number of cars in the section [x, x + ∆x] at time t. We define ρ(x, t) by the
expression
∆N ([x, x + ∆x], t)
ρ(x, t) = lim , (2.20)
∆x→0 ∆x
provided that the limit on the right–hand side of (2.20) exists. It follows from
(2.20) that, if a continuous traffic density, ρ(x, t), is known for all x and t, then
the number of cars in a section of the road from x = a to x = b, where a < b,
at time t is given by
Z b
∆N ([a, b], t) = ρ(x, t) dx.
a
We assume that at each point x along the road and at each time t the velocity
of vehicle at that location and time is dictated by a function u(x, t), which we
also assume to be a C 1 function. It follows from these assumptions and the
derivations in this section that the one–dimensional equation of continuity in
(2.19) applies to this situation.
Ideally, we would like to find a solution, ρ, to (2.19) subject to some initial
condition
ρ(x, 0) = ρo (x), (2.21)
for some initial traffic density profile, ρo , along the road. In order to solve this
problem, we postulate that u is a function of traffic density—the higher the
density, the lower the traffic speed, for example. We may therefore write
the parameters ρmax and vmax , the simplest model for the relationship between
v and ρ is the constitutive equation
ρ
u = vmax 1 − . (2.23)
ρmax
We therefore arrive at the initial value problem (IVP):
∂ρ ∂ ρ
+ vmax ρ 1− = 0 for x ∈ R, t > 0;
∂t ∂x ρmax
(2.24)
ρ(x, 0) = ρo (x), for x ∈ R,
where we have incorporated the continuity equation in (2.19), the initial con-
dition in (2.21), and the constitutive relation in (2.23), which is an instance of
(2.22).
The partial differential equation model for traffic flow (2.24) presented in this
section, based on the equation of continuity in (2.19) and a constitutive relation
for the traffic velocity, u, and the traffic density ρ (of which (2.23) is just an
example), was first introduced by Lighthill and Whitman in 1955 (see [LW55]);
it was also treated by Richards in 1956, [Ric56]. In a subsequent section in
these notes we will present an analysis of this model based on the method of
characteristics.
We end this section with an alternate derivation of the conservation of mass
equation in (2.17). In this approach we focus on the amount of fluid contained
in a region B as the fluid in this region moves according to flow dictated by
the velocity field ~u. Suppose we begin to observe a portion of fluid in B at
time t = 0. We assume that B is bounded and has smooth boundary ∂B. At
some time t > 0, the portion of fluid in B has moved as a consequence of the
fluid motion. We denote by Bt the portion of the fluid that we are following at
time t (see Figure 2.1.3). To see how Bt comes about, consider a fluid element
located at (x, y, z) at time t = 0. At time t > 0, the fluid element will be located
at (x(t), y(t), z(t)), where the functions x(t), y(t) and z(t) are solutions to the
system of ordinary differential equations in (2.5) subject to the initial conditions
in (2.6). We denote the point (x(t), y(t), z(t)) by ϕt (x, y, z), and note that the
map
(x, y, z) 7→ ϕt (x, y, z), for all (x, y, z) ∈ R,
yields a C 1 map from R to R. Furthermore, ϕt is an invertible map for each t
in the interval of existence for the initial value problem in (2.5) and (2.6). We
shall refer to ϕt as the fluid flow map; it gives the location of a fluid element
initially at (x, y, y) at time t as a result of fluid motion. It then follows that Bt
is the image of B under the flow map ϕt ; that is,
Bt = ϕt (B). (2.25)
The total mass of the fluid in Bt is a function of time that we compute as follows
ZZZ
m(t) = ρ(ϕt (x, y, z), t) dV. (2.26)
Bt
14 CHAPTER 2. MODELING WITH PDES
Bt
which is the mass of the portion of fluid in the region B. As the flow of the fluid
moves the region B, its shape might change. However, because of conservation
of mass, the mass of fluid contained in Bt must the same as that contained in
the region B at time t = 0; that is,
dm
= 0, for all t. (2.29)
dt
dm
Before we compute , we first rewrite the integral defining m(t) in (2.26)
dt
by means of the change of variables provided by the flow map ϕt (see (2.25).
We have ZZZ
m(t) = ρ(x, y, z, t)J(x, y, z, t) dxdydz
B
where J(x, y, z, t) the Jacobian of the map ϕt ; that is, J(x, y, z, t) is the deter-
minant of the derivative map of ϕt . We then have that
ZZZ
dm ∂
= [ρJ] dxdydz,
dt B ∂t
or ZZZ
dm ∂J ∂ρ
= ρ + J dxdydz. (2.30)
dt B ∂t ∂t
2.1. MODELING FLUID FLOW 15
Making the change of variables provided by the flow map in the integral in (2.30)
we obtain that
ZZZ
dm 1 ∂
= ρ(ϕt (x, y, z), t) J(ϕt (x, y, z), t)
dt Bt J(ϕt (x, y, z), t) ∂t
(2.31)
∂
+ [ρ(ϕt (x, y, z), t)] dV.
∂t
It can be shown that
∂
[J(ϕt (x, y, z), t) = J(ϕt (x, y, z), t) ∇ · ~u(ϕt (x, y, z), t), (2.32)
∂t
see page 8 in [CM93]. Thus, substituting (2.32) into (2.31), we get
ZZZ
dm ∂
= (∇ · ~u(ϕt (x, y, z), t))ρ(ϕt (x, y, z), t) + [ρ(ϕt (x, y, z), t)] dV,
dt Bt ∂t
which we can write as
ZZZ
dm Dρ
= (∇ · ~u)ρ + dV, (2.33)
dt Bt Dt
where we have set
Dρ ∂
= [ρ(ϕt (x, y, z), t)]
Dt ∂t
∂
= [ρ(x(t), y(t), z(t), t)] (2.34)
∂t
∂ρ dx ∂ρ dy ∂ρ dz ∂ρ
= + + + ,
∂x dt ∂y dt ∂z dt ∂t
where we have used the Chain Rule in the last step of the calculations in (2.34)
and assumed that the density ρ is a C 1 field. We therefore have that
∂ ∂ρ ∂ρ ∂ρ ∂ρ
[ρ(x(t), y(t), z(t), t)] = + u1 + u2 + u3 , (2.35)
∂t ∂t ∂x ∂y ∂z
where we have used the fact that (x(t), y(t), z(t)) solves the system of ordinary
differential equations in (2.5). Writing (2.35) in vector notation we obtain
∂ ∂ρ
[ρ(x(t), y(t), z(t), t)] = + ~u · ∇ρ, (2.36)
∂t ∂t
∂ρ ∂ρ ∂ρ
where ∇ρ = , , is the gradient of ρ. The expression in (2.36) is
∂x ∂y ∂z
called the material derivative of the field ρ. It is also referred to as the
Dρ
convective derivative of ρ and is usually denoted by , so that
Dt
Dρ ∂ρ
= + ~u · ∇ρ. (2.37)
Dt ∂t
16 CHAPTER 2. MODELING WITH PDES
Dg ∂g
= + ~u · ∇g. (2.38)
Dt ∂t
The material derivative of g in (2.38) expresses the rate of change of g along the
pathlines as a result of the fact that the field g might change in time as well as
a result of the motion of the fluid. The material derivative of a C 1 vector field,
→
−
G = (g1 , g2 , g3 ), is
→
−
DG Dg1 Dg2 Dg3
= , , ,
Dt Dt Dt Dt
which can be written as
→
− →
−
DG ∂G →
−
= + (~u · ∇) G . (2.39)
Dt ∂t
Dρ
+ (∇ · ~u)ρ = 0, in R, for all t, (2.41)
Dt
Dρ
where the material derivative, , of ρ is given in (2.37); that is,
Dt
Dρ ∂ρ
= + ~u · ∇ρ
Dt ∂t
(2.42)
∂ρ
= + ∇ · (ρ~u) − (∇ · ~u)ρ
∂t
substituting the result of the calculations in (2.42) into (2.41) then yields
∂ρ
+ ∇ · (ρ~u) = 0, in R, for all t,
∂t
which is the continuity equation in (2.17). We have also shown that the equation
in (2.41) is an equivalent form of the continuity equation. We shall rewrite it
here as
Dρ
= −(∇ · ~u)ρ, in R, for all t. (2.43)
Dt
2.1. MODELING FLUID FLOW 17
→
−
ZZZ
Π B (t) = ρ(x(t), y(t), z(t), t)~u(x(t), y(t), z(t), t) dV,
Bt
or
→
−
ZZZ
Π B (t) = ρ(ϕt (x, y, x), t)~u(ϕt (x, y, x), t) dV,
Bt
→
−
ZZZ
Π B (t) = ρ~u dV, (2.44)
Bt
(see Figure 2.1.3). The principle of conservation of momentum states that the
rate of change of the total momentum of the fluid in Bt has to be accounted for
by the balance of forces acting on Bt :
→
−
dΠB
= Balance of Forces on Bt ; (2.45)
dt
this is, in fact, Newton’s second law of motion.
There are two types of forces acting on the portion of fluid in Bt that con-
tribute to the balance of forces in the right–hand side of the equation in (2.45).
There are forces of stress due to the fluid surrounding the region Bt , and there
are external, or body forces, such as gravity or electromagnetic forces. We can
then rewrite the conservation of momentum equation in (2.45) as
→
−
dΠB →
− →
−
= S B (t) + F B (t), (2.46)
dt
→
−
where S B (t) denotes the total vector sum of the stress forces acting on Bt , and
→
−
F B (t) the total vector sum of body forces acting on Bt .
We assume that
→
−
ZZZ
F B (t) = f~ dV, (2.47)
Bt
where the vector field f~(x, y, z, t) gives the total forces per unit volume acting
on an element of fluid around the point (x, y, z) at time t.
In this section we shall make a special assumption when modeling the stress
forces acting on the fluid. We assume that the fluid under consideration is an
ideal fluid. This means that at any point, (x, y, y), on a surface in the fluid,
the stress force per unit area exerted across the surface is given by
p(x, y, z, t)~n
18 CHAPTER 2. MODELING WITH PDES
where ~n is a unit vector perpendicular to the surface at (x, y, z) and time t, and
p(x, y, z, t) is a scalar field called the pressure. It then follows that
→
−
ZZ
S B (t) = − p~n dA, (2.48)
∂Bt
Writing the unit vector ~n in Cartesian coordinates, (n1 , n2 , n3 ), we see that the
stress forces term in (2.49) has components
ZZ ZZ ZZ
− pn1 dA, − pn2 dA, and − pn3 dA.
∂Bt ∂Bt ∂Bt
and ZZ ZZZ
∂p
− pn3 dA = − dV.
∂Bt Bt ∂z
→
−
Substituting these expressions into the definition of S B (t) in (2.48) we obtain
ZZZ
~B (t) = −
S ∇p dV, (2.50)
Bt
where
∂p ∂p ∂p
∇p = , , (2.51)
∂x ∂y ∂z
is the gradient of p. Combining (2.50), (2.48) and (2.46), we can rewrite the
conservation of momentum equation in (2.49) as
→
− ZZZ ZZZ
dΠB
=− ∇p dV + f~ dV, (2.52)
dt Bt Bt
where J(x, y, z, t) the Jacobian of the map ϕt ; that is, J(x, y, z, t) is the deter-
minant of the derivative map of ϕt . We then have that
ZZZ ZZZ
d ∂
ρ~u dV = [Jρ~u] dxdydz,
dt Bt B ∂t
or ZZZ ZZZ
d ∂J ∂
ρ~u dV = ρ~u + [ρ~u]J dxdydz. (2.54)
dt Bt B ∂t ∂t
Substituting (2.32) into (2.54) yields
ZZZ ZZZ
d ∂
ρ~u dV = (∇ · ~u)ρ~u + [ρ~u] J dxdydz,
dt Bt B ∂t
Using the expression for the material derivative of a vector field in (2.39), we
can rewrite (2.55) as
ZZZ ZZZ
d D
ρ~u dV = (ρ~u) + (∇ · ~u)ρ~u dV. (2.56)
dt Bt Bt Dt
Using the definition of the convective derivative for a vector field in (2.39) we
have that
D D~u Dρ
(ρ~u) = ρ + ~u, (2.57)
Dt Dt Dt
where
Dρ ∂ρ
= + ~u · ∇ρ
Dt ∂t
∂ρ
+ ∇ · (ρ~u) − ρ∇ · ~u;
=
∂t
it then follows from the conservation mass equation in (2.17) that
Dρ
= −ρ∇ · ~u, (2.58)
Dt
20 CHAPTER 2. MODELING WITH PDES
D D~u
(ρ~u) = ρ − (∇ · ~u)ρ~u. (2.59)
Dt Dt
D
Substituting the expression for (ρ~u) in (2.59) into the expression for the
Dt
rate of change of momentum in (2.56) yields
ZZZ ZZZ
d D~u
ρ~u dV = ρ dV. (2.60)
dt Bt Bt Dt
or ZZZ
D~u
ρ + ∇p − f~ dV = 0, for all t. (2.61)
Bt Dt
Assuming that the fields ρ, ~u and p are C 1 over R and for all times t, and that
the field f~ is continuous over R and for all times t, we see that the integrand in
the left–hand side of (2.61) is continuous over R and for all times t. Thus, since
(2.61) holds true for all bounded subregions, Bt , of R with smooth boundary,
we conclude that
D~u
ρ + ∇p − f~ = 0, in R, for all t,
Dt
or
D~u
ρ = −∇p + f~, in R, for all t, (2.62)
Dt
which is the differential form of the conservation of momentum principle.
Observe that the PDE in (2.62) is a vector differential equation in three
dimensions. As such, it is really a system of three first–order PDEs:
Du1 ∂p
ρ = − + f1 ;
Dt ∂x
Du
2 ∂p
ρ = − + f2 ; (2.63)
Dt ∂y
ρ Du3 = − ∂p + f .
3
Dt ∂z
The equations in (2.18) and (2.63) constitute a system of four first–order
PDEs in the (possibly) unknown scalar fields u1 , u2 , u3 , ρ and p (the body
forces field f~ can usually be determined from the outset). Thus, in order to
2.1. MODELING FLUID FLOW 21
have any hope for solving the system of conservation equations in (2.17) and
(2.62), we need to have at least one more relation, or equation, involving the
velocity field, ~u, the density, ρ, and the pressure, p. Another relation will be
provided by the principle of conservation of energy to be discussed in the next
section.
→
−
The expression in (2.60) holds true for any C 1 vector field G in R,
→
−
→
−
ZZZ ZZZ
d DG
ρ G dV = ρ dV,
dt Bt Bt Dt
As the shape of the region Bt changes with the flow, the volume of Bt might
change also. We compute the rate at which the volume changes by first rewriting
the expression for v(t) in (2.65) as
ZZZ
v(t) = J(ϕt (x, y, z), t) dV. (2.66)
B
In an incompressible flow the volume of any portion of the fluid does not
change with time. We therefore obtain from (2.68) that
ZZZ
∇ · ~u dV = 0, for all t. (2.69)
Bt
Since the expression in (2.69) holds true for any Bt in R, it follows that, for
case in which the velocity field, ~u, is C 1 in R, the condition for the flow to be
incompressible is
∇ · ~u = 0, in R for all t. (2.70)
We show in this section that, in an ideal incompressible fluid, the kinetic energy
in the portion of the fluid in Bt is conserved.
The kinetic energy of the portion of the fluid Bt at time t is given by
ZZZ
1
E(t) = ρk~uk2 dV, (2.71)
2 Bt
where kuk2 = ~u · ~u is the square of the Euclidean norm of the velocity field ~u.
The rate of change of E in (2.71) is given by the Transport Theorem in
(2.64) to be ZZZ
dE 1 D
= ρ [k~uk2 ] dV, (2.72)
dt 2 Bt Dt
where
D D~u
[k~uk2 ] = 2~u · ,
Dt Dt
so that, in view of (2.72),
ZZZ
dE D~u
= ρ~u · dV,
dt Bt Dt
or ZZZ
dE D~u
= ~u · ρ dV. (2.73)
dt Bt Dt
Substituting the law of conservation of momentum expression for an ideal fluid
in (2.62) into the right–hand side of (2.73) then yields
ZZZ
dE
= ~u · −∇p + f~ dV,
dt Bt
The right–most integral in (2.74) measures the rate at which body forces do
work in the portion of fluid in Bt at time t. In order to understate the other
2.1. MODELING FLUID FLOW 23
integral in (2.74) we use the assumption that the fluid is incompressible, stated
as the PDE in (2.70), to obtain
∇ · (p~u) = ∇p · ~u + p∇ · ~u = ∇p · ~u,
Applying the divergence theorem to the integral on the right–hand side of (2.75)
yields ZZZ ZZ
∇p · ~u dV = p~u · ~n dA, (2.76)
Bt ∂Bt
where ~n denotes the outward unit normal vector the boundary of Bt . Substi-
tuting the expression in (2.76) into the right–hand side of (2.74) then yields
ZZ ZZZ
dE
=− p~u · ~n dA + f~ · ~u dV. (2.77)
dt ∂Bt Bt
ZZ
Observe that − p~u · ~n dA gives the rate at which the stress forces are
∂Bt
doing work on the portion of fluid in Bt . Hence, the equation in (2.77) is a
statement of the conservation of kinetic energy.
where ρo , ~uo and po are given functions defined in R. Since, we want the flow
to remain within the region R, we also impose the boundary condition
where we are assuming that R has a smooth boundary ∂R. The condition in
(2.80) forbids fluid to cross in or out of the boundary.
gives, approximately, the number of particles that cross a small section of the
surface of area dA, per unit time, in a direction perpendicular to the surface at
that point. It then follows that the number of particles per unit time crossing
the smooth boundary of a region B ⊂ R into that region (see Figure 2.2.4) is
given by
→
−
ZZ
− J (x, y, z, t) · ~n dA, (2.81)
∂B
where the minus sign in (2.81) takes into account that we are taking ~n to be
the outward unit normal to ∂B. The expression in (2.81) is called the flux of
particles across the boundary of B.
Assume that the concentration of particles in the region R at any time t is
given by a C 1 scalar field, u, so that number of particles contained in the region
B is given at time t is given by
ZZZ
NB (t) = u(x, y, x, t) dxdydx, for all t. (2.82)
B
Assuming that particles are not being created or destroyed, we get the conser-
vation principle
→
−
ZZ
dNB
=− J (x, y, z, t) · ~n dA (2.83)
dt ∂B
Since we are assuming that u is a C 1 field, we can differentiate under the integral
sign in (2.82) to rewrite (2.83) as
→
−
ZZZ ZZ
∂u
dxdydx = − J (x, y, z, t) · ~n dA (2.84)
B ∂t ∂B
→
−
If we also assume that the vector field J is a C 1 function, we can use the
Divergence Theorem to rewrite the right–hand side of (2.84) to obtain
→
−
ZZZ ZZZ
∂u
dV = − ∇ · J dV
B ∂t B
or ZZZ
∂u →
−
+∇· J dV = 0. (2.85)
B ∂t
Since (2.85) holds true for all bounded subsets, B, of R, and all times t, we
obtain the PDE
∂u →
−
+ ∇ · J = 0, in R, for all t. (2.86)
∂t
The PDE in (2.86) has two unknown functions: the concentration u and the
→
−
flux field J . Thus, in order to complete the modeling, we need a constitutive
26 CHAPTER 2. MODELING WITH PDES
→
−
equation relating u and J . This is provided by Fick’s First Law of Diffusion
(see [Ber83, pg. 18]):
→
−
J = −D∇ · u, in R, for all t, (2.87)
∂u
for some given function f : R → R, and some integrability conditions on u,
∂x
and f .
The equation in (2.92) also describes the flow of heat in a cylindrical metal
rod of constant cross–sectional area whose cylindrical boundary is insulated so
that heat can only flow in or out of the rod through the cross sections a the ends
of the rod (see Assignment #4). In this case u(x, t) denotes the temperature in
x
0 L
the cross–section of the rod located at x and at time t, and the constant D is
given by
κ
D= ,
cρ
where ρ is the density, c is the specific heat, and κ is the heat conductivity of
the material the of the rod (see Assignment #4). Thus, (2.92) ia also called the
heat equation. In this case D is called the thermal diffusivity.
In these notes we will see how to solve the heat equation in (2.92) subject
the initial and boundary conditions
u(x, 0) = f (x), for 0 < x < L;
u(0, t) = To (t), for t > 0;
u(L, t) = TL (t), for t > 0,
where f , To and TL are given functions of single variable. We will also solve the
problem with the boundary conditions
∂u (0, t) = 0,
for t > 0;
∂x
∂u
(L, t) = 0, for t > 0.
∂x
These conditions imply that heat cannot flow through the end cross–sections
either; so that the rod is totally insulated.
x R
dimensional space. Specifically, the shape of the soap film spanning the wire
loop, can be modeled by the graph of a smooth function, u : R → R, defined on
the closure of a bounded region, R, in the xy–plane with smooth boundary ∂R.
The physical explanation for the shape of the soap film relies on the variational
principle that states that, at equilibrium, the configuration of the film must be
such that the energy associated with the surface tension in the film must be the
lowest possible. Since the energy associated with surface tension in the film is
proportional to the area of the surface, it follows from the least–energy principle
that a soap film must minimize the area; in other words, the soap film spanning
the wire loop must have the shape of a smooth surface in space containing
the wire loop with the property that it has the smallest possible area among
all smooth surfaces that span the wire loop. In this section we will develop a
mathematical formulation of this variational problem.
The wire loop can be modeled by the curve determined by the set of points:
Φ : R → R3
given by
Φ(x, y) = (x, y, u(x, u)), for all x ∈ R, (2.93)
where R = R ∪ ∂R is the closure of R, and
u: R → R
u ∈ C 2 (R) ∩ C(R).
that is,
Ag = {u ∈ C 2 (R) ∩ C(R) | u = g on ∂R}. (2.94)
Next, we see how to compute the area of the surface Su = Φ(R), where Φ is
the map given in (2.93) for u ∈ Ag , where Ag is the class of functions defined
in (2.94).
The grid lines x = c and y = d, for arbitrary constants c and d, are mapped
by the parametrization Φ into curves in the surface Su given by
y 7→ Φ(c, y)
and
x 7→ Φ(x, d),
respectively. The tangent vectors to these paths are given by
∂u
Φy = 0, 1, (2.95)
∂y
and
∂u
Φx = 1, 0, , (2.96)
∂x
respectively. The quantity
kΦx × Φy k∆x∆y (2.97)
gives an approximation to the area of portion of the surface Su that results
from mapping the rectangle [x, x + ∆x] × [y, y + ∆y] in the region R to the
surface Su by means of the parametrization Φ given in (2.93). Adding up all
the contributions in (2.97), while refining the grid, yields the following formula
for the area Su : ZZ
area(Su ) = kΦx × Φy k dxdy. (2.98)
R
30 CHAPTER 2. MODELING WITH PDES
Using the definitions of the tangent vectors Φx and Φy in (2.95) and (2.96),
respectively, we obtain that
∂u ∂u
Φx × Φ y = − , − , 1 ,
∂x ∂y
so that s 2 2
∂u ∂u
kΦx × Φy k = 1+ + ,
∂x ∂y
or p
kΦx × Φy k = 1 + |∇u|2 ,
where |∇u| denotes the Euclidean norm of ∇u. We can therefore write (2.98)
as ZZ p
area(Su ) = 1 + |∇u|2 dxdy. (2.99)
R
A : Ag → R
by ZZ p
A(u) = 1 + |∇u|2 dxdy, for all u ∈ Ag , (2.100)
R
which gives the area of the surface parametrized by the map Φ : R → R3 given
in (2.93) for u ∈ Ag . We will refer to the map A : Ag → R defined in (2.100)
as the area functional. With the new notation we can restate the variational
problem of this section as follows:
Problem 2.3.1 (Variational Problem 1). Out of all functions in Ag , find one
such that
A(u) 6 A(v), for all v ∈ Ag . (2.101)
That is, find a function in Ag that minimizes the area functional in the class
Ag .
Problem 2.3.1 is an instance of what has been known as Plateau’s problem
in the Calculus of Variations. The mathematical question surrounding Pateau’s
problem was first formulated by Euler and Lagrange around 1760. In the middle
of the 19th century, the Belgian physicist Joseph Plateu conducted experiments
with soap films that led him to the conjecture that soap films that form around
wire loops are of minimal surface area. It was not until 1931 that the American
mathematician Jesse Douglas and the Hungarian mathematician Tibor Radó,
independently, came up with the first mathematical proofs for the existence of
minimal surfaces. In this section we will derive a necessary condition for the
existence of a solution to Problem 2.3.1, which is expressed in terms of a PDE
that u ∈ Ag must satisfy, the minimal surface equation.
2.3. VARIATIONAL PROBLEMS 31
f 0 (0) = 0. (2.105)
We will see next that, since we are assuming that u ∈ C 2 (R) ∩ C(R) and
ϕ ∈ Cc∞ (R), f is indeed differentiable. To see why this is the case, use (2.104)
and (2.100) to compute
ZZ p
f (t) = 1 + |∇(u + tϕ)|2 dxdy, for all t ∈ R, (2.106)
R
where
∇(u + tϕ) = ∇u + t∇ϕ, for all t ∈ R,
by the linearity of the differential operator ∇. It then follows that
= ∇u · ∇u + t∇u · ∇ϕ + t∇ϕ · ∇u + t2 ∇ϕ · ∇ϕ
Since the integrand in (2.107) is C 1 , we can differentiate under the integral sign
to get
∇u · ∇ϕ + t|∇ϕ|2
ZZ
0
f (t) = p dxdy, (2.108)
R 1 + |∇u|2 + 2t∇u · ∇ϕ + t2 |∇ϕ|2
32 CHAPTER 2. MODELING WITH PDES
for all ϕ ∈ Cc∞ (R), where the second integral in (2.111) is a path integral
around the boundary of R. Since ϕ ∈ Cc∞ (R) vanishes in a neighborhood of the
boundary of R, it follows from (2.111) that
!
∇u
ZZ
∇· p ϕ dxdy = 0, for all ϕ ∈ Cc∞ (R). (2.112)
1 + |∇u| 2
R
The BVP in (2.114) is called the Dirichlet problem for the minimal surface
equation.
2.3. VARIATIONAL PROBLEMS 33
x R
|∇u| is very small throughout R. We can therefore use the linear approximation
√ 1
1 + t ≈ 1 + t, for small |t|, (2.117)
2
34 CHAPTER 2. MODELING WITH PDES
so that
ZZ
1
A(u) ≈ area(R) + |∇u|2 dxdy, for all u ∈ Ag . (2.118)
2 R
Thus, according to (2.120), for wire loops close to a horizontal plane, mini-
mal surfaces spanning the wire loop can be approximated by solutions to the
following variational problem,
Problem 2.3.2 (Variational Problem 2). Out of all functions in Ag , find one
such that
D(u) 6 D(v), for all v ∈ Ag . (2.121)
It can be shown that a necessary condition for u ∈ Ag to be a solution to
the Variational Problem 2.3.2 is that u solves the boundary value problem
∆u = 0 in R;
(2.122)
u = g on ∂R,
where
∆u = uxx + uyy ,
the two–dimensional Laplacian. The BVP in (2.122) is called the Dirichlet
Problem for Laplace’s equation.
x
0 L
x
0 L
after it is plucked to a configuration like that pictured in Figure 2.3.9 and then
released. We assume that the shape of the plucked string is described by a
continuous function, f , of x, for x ∈ [0, L]. At any time t > 0, the shape of
the string is described by a function, u, of x and t; so that u(x, t) gives the
vertical displacement of a point in the string located at x when the string is in
the equilibrium position pictured in Figure 2.3.8, and at time t > 0. We then
have that
u(x, 0) = f (x), for all x ∈ [0, L]. (2.123)
In addition to the initial condition in (2.123), we will also prescribe the initial
speed of the string,
∂u
(x, 0) = g(x), for all x ∈ [0, L], (2.124)
∂t
where g is a continuous function of x; for instance, if the plucked string is
released from rest, then g(x) = 0 for all x ∈ [0, L]. We also have the boundary
conditions,
u(0, t) = u(L, t) = 0, for all t, (2.125)
which model the assumption that the ends of the string do not move.
The question we would like to answer is: Given the initial conditions in
(2.123) and (2.124), and the boundary conditions in (2.125), can we determine
the shape of the string, u(x, t), for all x ∈ [0, L] and all times t > 0. We will
answer this questions in a subsequent chapter in these notes. In this section,
though, we will derive a necessary condition in the form of a PDE that u must
satisfy in order for it to describe the motion of the vibrating string.
In order to find the PDE governing the motion of the string, we will formulate
the problem as a variational problem. We will use Hamilton’s Principle in
Mechanics, or the Principle of Least Action. This principle states that the
the path that configurations of a mechanical system take from time t = 0 to
t = T is such that a quantity called the action is minimized (or optimized)
along the path. The action is defined by
Z T
A= [K(t) − V (t)] dt, (2.126)
0
36 CHAPTER 2. MODELING WITH PDES
where K(t) denotes the kinetic energy of the system at time t, and V (t) its
potential energy at time t. For the case of a string whose motion is described by
small vertical displacements u(x, t), for all x ∈ [0, L] and all times t, the kinetic
energy is given by
Z L 2
1 ∂u
K(t) = ρ(x) (x, t) dx. (2.127)
2 0 ∂t
To see how (2.127) comes about, note that the kinetic energy of a particle of
mass m is
1
K = mv 2 ,
2
where v is the speed of the particle. Thus, for a small element of the string whose
projection on the x–axis is the interval [x, x+∆x], so that its approximate length
is ∆x, the kinetic energy is, approximately,
2
1 ∂u
∆K ≈ ρ(x) (x, t) . (2.128)
2 ∂t
Thus, adding up the kinetic energies in (2.128) over all elements of the string,
and letting ∆x → 0, yields the expression in (2.127), which we rewrite as
Z L
1
K(t) = ρu2t dx, for all t, (2.129)
2 0
so that "Z #
L p
V (t) = τ 1 + u2x dx − L , for all t. (2.131)
0
Since we are considering small vertical displacements of the string, we can lin-
earize the expression in in (2.130) by means of the linear approximation in
(2.117) to get
Z L Z L Z L
p 1 1
1 + u2x dx ≈ [1 + u2x ] dx = L + u2x dx,
0 0 2 2 0
where we have substitute the expressions for K(t) and V (t) in (2.129) and
(2.132), respectively, into the expression for the action in (2.126).
We will use the expression for the action in (2.133) define a functional in the
class of functions A defined as follows: Let R = (0, L) × (0, T ), the cartesian
product of the open intervals (0, L) and (0, T ). Then, R is an open rectangle
in the xt–plane. We say that u ∈ A if u ∈ C 2 (R) ∩ C(R), and u satisfies the
initial conditions in (2.123) and (2.124), and the boundary conditions in (2.125).
Then, the action functional,
A : A → R,
is defined by the expression in (2.133), so that
ZZ
1 2
ρut − τ u2x dxdt,
A(u) = for u ∈ A. (2.134)
2 R
Next, for ϕ ∈ Cc∞ (R), note that u + sϕ ∈ A, since ϕ has compact support
in R, and therefore ϕ and all its derivatives are 0 on ∂R. We can then define a
real valued function h : R → R by
Using the definition of the functional A in (2.134), we can rewrite h(s) in (2.135)
as ZZ
1
ρ[(u + sϕ)t ]2 − τ [(u + sϕ)x ]2 dxdt
h(s) =
2 R
ZZ
1
ρ[ut + sϕt ]2 − τ [ux + sϕx ]2 dxdt,
=
2 R
so that
ZZ
h(s) = A(u) + s [ρut ϕt − τ ux ϕx ] dxdt + s2 A(ϕ), (2.136)
R
for s ∈ R, where we have used the definition of the action functional in (2.134).
It follows from (2.136) that h is differentiable and
ZZ
h0 (s) = [ρut ϕt − τ ux ϕx ] dxdt + 2sA(ϕ), for s ∈ R. (2.137)
R
The principle of least action implies that, if u describes the shape of the string,
then s = 0 must be ac critical point of h. Hence, h0 (0) = 0 and (2.137) implies
that ZZ
[ρut ϕt − τ ux ϕx ] dxdt = 0, for ϕ ∈ Cc∞ (R), (2.138)
R
38 CHAPTER 2. MODELING WITH PDES
for C 1 functions ψ and ϕ, where n1 is the first component of the outward unit
normal, ~n, on ∂R (wherever this vector is defined), and
ZZ Z ZZ
∂ϕ ∂ψ
ψ dxdt = ψϕn2 ds − ϕ dxdt,
R ∂t ∂R R ∂t
where n2 is the second component of the outward unit normal, ~n, to obtain
ZZ Z ZZ
∂
ρut ϕt dxdt = ρut ϕn2 ds − [ρut ]ϕ dxdt,
R ∂R R ∂t
so that ZZ ZZ
∂
ρut ϕt dxdt = − [ρut ]ϕ dxdt, (2.139)
R R ∂t
since ϕ has compact support in R.
Similarly, ZZ ZZ
∂
τ ux ϕx dxdt = − [τ ux ]ϕ dxdt. (2.140)
R R ∂x
Next, substitute the results in (2.139) and (2.140) into (2.138) to get
ZZ
∂ ∂
[ρut ] − [τ ux ] ϕ dxdt = 0, for ϕ ∈ Cc∞ (R). (2.141)
R ∂t ∂x
Thus, we obtain from (2.141) that
∂2u ∂2u
ρ − τ = 0, in R, (2.142)
∂t2 ∂x2
since we area assuming that u is C 2 , ρ is a continuous function of x, and τ is
constant.
The PDE in (2.142) is called the one–dimensional wave equation. It is
sometimes written as
∂2u τ ∂2u
= ,
∂t2 ρ ∂x2
or
∂2u ∂2u
2
= c2 2 , (2.143)
∂t ∂x
where
τ
c2 = ,
ρ
for the case in which ρ and τ are assumed to be constant.
The wave equation in (2.142) or (2.143) is a second order, linear, hyperbolic
PDE.
2.4. MODELING SMALL AMPLITUDE VIBRATIONS 39
x
0 x x + ∆x L
Figure 2.4.10 shows a snapshot of the vibrating string at some time t. The
figure also highlights a small section of the string that is above an interval
[x, x + ∆x], for some x ∈ (0, L). As in the derivation in Section 2.3.3, u(x, t) is
the vertical displacement of a point on the string located at x when the string is
at equilibrium, and time t (we are assuming here that the vibrations have small
amplitude; so that, we may assume that each particle along the string moves
along a vertical direction); ρ is the density of the material making up the string
(in units of mass per unit of length); and τ is the tension along the string in
units of force. We assume that u has continuous partial derivatives,
tan(θ) = ux . (2.144)
1
cos(θ) = p . (2.146)
1 + u2x
40 CHAPTER 2. MODELING WITH PDES
If we assume that ∆x is very small, we may treat the section of the string
above the interval [x, x + ∆x] as a particle of mass given approximately by
ρ(x)∆s,
where ∆s is the arc–length of the section of the string, which is given, approxi-
mately, by p
∆s ≈ 1 + u2x ∆x.
Thus, the mass of the section of the string above the interval [x, x + ∆x] is,
approximately, p
∆m ≈ ρ(x) 1 + u2x ∆x. (2.147)
Newton’s Second Law of Motion applied to a particle of mass ∆m given in
(2.147) reads
∂2u
(∆m) = forces acting on the particle in vertical dirction. (2.148)
∂t2
The forces acting on the particle in the vertical direction are the force of
gravity and the vertical components of the tension, which acts tangentially to
the string and away from the section above the interval [x, x+∆x]. There might
be other forces acting on the string as well; for instance, a violin string might be
acted on by the bow when playing. We will put together the forces other than
tension into a function, F (x, t), in units of force per length. Thus, the forces,
other than tension, acting on the particle are
The vertical component of the tension acting on the section of the string above
the interval [x, x + ∆x] is
The PDE in (2.152), as it stands, is a nonlinear equation (we will see why this
is the case in the next section on classification of PDEs). We can approximate
the equation by a linear PDE by using the assumption that the vibrations of
the string are on very small amplitude. In the case of very small amplitude
vibrations, ux , which is related to the tangent of the angle the string makes
with a horizontal direction, is very small in magnitude. Thus, we can make the
approximation
ux ≈ 0
in (2.152) to obtain
ρutt = τ uxx + F (x, t). (2.153)
Observe that the equation in (2.153) leads to the wave equation in (2.142) for
the case F (x, t) = 0 for all x ∈ [0, L] and t > 0. The equation in (2.142) will be
called in this notes the one–dimensional, linear, homogeneous wave equation.
We will refer to the equation in (2.153) with F 6≡ 0 as the one–dimensional,
linear, non–homogeneous wave equation. In these notes we will show how to
construct solutions of those equations.
42 CHAPTER 2. MODELING WITH PDES
Chapter 3
43
44 CHAPTER 3. CLASSIFICATION OF PDES
equations, while the first two equations in the system in (3.1) and the PDE in
(3.5) are not linear. In the next section we will discuss properties of linear
equations, and how those properties can be very helpful in the construction of
solutions, and proofs of uniqueness for some initial/boundary value problems.
The PDE in (3.5) is in a general class of equations of the form
Duxx − ut = 0, (3.11)
3.1 Linearity
Laplace’s equation (3.9), the one–dimensional wave equation (3.10), and the
one–dimensional diffusion equations (3.11) are linear equations. To understand
the use of this terminology in the context of PDEs, note that Laplace’s equation
(3.9) can also be written as
∆u = 0,
where ∆ : C 2 (R) → C(R) defines a linear operator between the spaces of func-
tions C 2 (R) and C(R) given by
L(c1 u + c2 v) = c1 Lu + c2 Lv.
L(c1 u + c2 v) = c1 0 + c2 0 = 0,
(x(t), y(t))
second order derivatives, uxx , uxy and uyy , of u (and higher order derivatives)
on Γ. Is this can be done, we can attempt to construct a solution of the PDE in
(3.17) by building Taylor series expansions around every point on Γ using the
values of u and its derivatives based on the conditions in (3.18), (3.19) and (3.20)
and derivatives of the expressions in (3.19) and (3.20). The first step in this
construction consists of taking the derivatives of derivatives of the expressions
in (3.19) and (3.20) and combining these with the information provided by the
PDE (3.17) to obtain the linear system
= f˙
ẋ uxx + ẏ uxy
ẋ uxy + ẏ uyy = ġ (3.21)
a uxx + b uxy + c uyy = d,
for the unknowns uxx , uxy and uyy on Γ, where a dot on top of a variable
denotes derivative with respect to t:
dx dy ˙ df dg
ẋ = , ẏ = , f= and ġ = .
dt dt dt dt
Note that the system in (3.21) can be written in matrix form as
f˙
ẋ ẏ 0 uxx
0 ẋ ẏ uxy = ġ . (3.22)
a b c uyy d
The matrix equation in (3.22) can be solved for the second derivatives of u, in
terms of the data (f˙, ġ, d) on Γ, provided that the determinant of the matrix
ẋ ẏ 0
0 ẋ ẏ (3.23)
a b c
48 CHAPTER 3. CLASSIFICATION OF PDES
is not zero. The case in which the determinant of the matrix in (3.23) yields
the equation
a(ẏ)2 − bẋẏ + c(ẋ)2 = 0. (3.24)
Dividing the equation in (3.24) by ẋ2 and using the fact that
ẏ dy
= ,
ẋ dx
by the Chain Rule, we obtain the ordinary differential equation (ODE)
2
dy dy
a −b + c = 0. (3.25)
dx dx
We shall refer to the ODE in (3.25) as the characteristic equation to the PDE
in (3.17). Solutions to the ODE in (3.25) are called characteristic curves of
dy
the PDE in (3.17). Assuming that a 6= 0 in R, we can solve for in (3.25) to
dx
get √
dy b ± b2 − 4ac
= . (3.26)
dx 2a
We have three possibilities, depending on whether b2 − 4ac is positive, zero, or
negative.
If b2 − 4ac > 0, then the PDE in (3.17) has two families of characteristic
curves given by the solutions to the two ODEs in (3.26). In this case we say
that the PDE in (3.17) is hyperbolic.
If b2 −4ac = 0, then the PDE in (3.17) has one family of characteristic curves
given by the solution to the ODE
dy b
= .
dx 2a
In this case we say that the PDE in (3.17) is parabolic.
If b2 − 4ac < 0, then the ODE in (3.25) has no real solutions. Thus, the
PDE in (3.17) has no (real) characteristic curves. In this case we say that the
PDE in (3.17) is elliptic.
Example 3.2.1 (The One–dimensional Wave Equation). In the case of the
linear second order equation
or
dt 1
=± ,
dx c
which we can rewrite as
dx
= ±c. (3.28)
dt
The equations in (3.28) is a pair of ODEs that can be solved to yield the two
families of curves
x = ct + ξ, (3.29)
and
x = −ct + η, (3.30)
where ξ and η are the parameters for each of the families of characteristic
curves in (3.29) and (3.30). The family of characteristic curves described by
the equations in (3.29) consists of parallel lines of (positive) slope 1/c in the
xt–plane with x–intercept ξ ∈ R. Some of those characteristic curves are shown
in Figure 3.2.2.
There is really no general theory for solving any given PDE of the form in (2.1),
In Section 3.2 we defined characteristic curves for second order PDEs in two
variables, and saw how characteristic curves can be used to come up with a
classification scheme for those equations. In this section we see how to use
characteristic curves to construct solutions to certain types of PDEs in two
variables. We begin with the example of the one–dimensional wave equation.
51
52 CHAPTER 4. SOLVING PDES
x = ct + ξ, (4.2)
and
x = −ct + η. (4.3)
These families of curves consist of parallel straight lines in the xt–plane of
slope 1/c and of slope −1/c, respectively. We will see in the next section that
characteristic curves carry information about solutions of the equation from one
point on the curve to another point on the same curve. This suggests that we
we consider the PDE in (4.1) along the curves given in (4.2) and (4.3). We can
do this my considering the parameters ξ and η in (4.2) and (4.3) as a new set
of variables,
ξ = x − ct, (4.4)
and
η = x + ct. (4.5)
If we are give a solution, u, to the PDE in (4.1), we can use the change of
variables provided by (4.4) and (4.5) to define a function, v, of ξ and η in terms
of u by means of
v(ξ, η) = u(x, t), (4.6)
where x and t can be obtained in terms of ξ and θ by solving the linear system
x − ct = ξ;
x + ct = η,
so that
1 1
x = η + ξ;
2 2
(4.7)
1 1
t = η − ξ.
2c 2c
Alternatively, we can rewrite (4.6) as
or
utt = c2 [vξξ − 2vξη + vηη ], (4.16)
where we have used the equality of the mixed second partial derivatives.
Since we are assuming that u solves the PDE
1
uxx − utt = 0, (4.17)
c2
it follows from (4.12), (4.16) and (4.17) that v solves the second order PDE
vξη = 0. (4.18)
Note that the PDE in (4.18) is also a hyperbolic, second order, linear PDE
(in this case a = c = 0 and b = 1, so that b2 − 4ac = 1 > 0). In contrasts
with the hyperbolic PDE in (4.17), the PDE in (4.18) can be solved directly by
integration. Indeed, writing (4.18) as
∂
[vξ ] = 0,
∂η
we see that
vξ = h(ξ), (4.19)
1
where h is an arbitrary C function of a single variable; and writing (4.19) as
∂
[v(ξ, η)] = 0,
∂ξ
we see that
v(ξ, η) = F (ξ) + G(η), (4.20)
where F is an antiderivative of h (i.e., F 0 = h), and G is an arbitrary C 2 function
of a single variable.
The function v defined by the expression in (4.20), where F and G are
arbitrary C 2 functions of a single variable, is the general solution to the PDE
in (4.18). We can use it, along with (4.8), (4.4) and (4.5) to obtain the general
solution to the one–dimensional wave equation
namely,
u(x, t) = F (x − ct) + G(x + ct), (4.22)
where F and G are arbitrary C 2 functions of a single variable. The expression in
(4.22) is known as d’Alembert’s solutions to the one–dimensional wave equation.
4.1. USING CHARACTERISTIC CURVES 55
We now use the general solution (4.22) to the one–dimensional wave equation
in (4.21) to construct a solution to the initial value problem in (4.1). In this
construction we will need to assume that f is C 2 and g is C 1 .
Differentiate the expression for u in (4.22) with respect to t to obtain
where we have applied the Chain Rule. Next, apply the initial conditions in
(4.1) to obtain the equations
F (x) + G(x) = f (x);
(4.24)
−cF 0 (x) + cG0 (x) = g(x),
(x, t)
ξ η x
solves the initial value problem (4.1) for the one–dimensional wave equation.
In order to understand what the solution to the IVP in (4.1) displayed in
(4.30) is saying, refer to Figure 4.1.1. Suppose we want to compute the value
of u at x and at time t > 0; that is, u(x, t), where (x, t) is a point in the
xt–plane. Two characteristic curves cross at that point: one with x–intercept
labeled ξ in Figure 4.1.1, and the other with x–intercept labeled η in the figure.
These correspond to the values x − ct and x + ct, respectively. According to the
expression for u in (4.29), the value of u at (x, t) is the average the values of the
initial data f at those two points, plus t times the average of all the values of
the initial speed, g, over the interval [ξ, η].
For the special case in which the initial speed is zero throughout R, we obtain
from (4.30) the special form
1
u(x, t) = [f (x − ct) + f (x + ct)], for x ∈ R and t ∈ R. (4.31)
2
1
The function u in (4.31) is made up of two traveling wave forms: f (x−ct),
2
1
which moves to the right with speed c, and f (x + ct), which moves to the left
2
with speed c. We illustrate this for the spacial case in which the initial data
f is in Cc∞ (R), with supp(f ) = [−1, 1]; see Figure 4.1.2. Figure 4.1.3 shows
the supports of the initial data and two of the traveling waves at some time
t > 0 later with ct > 2. Figure 4.1.4 shows the two pulses traveling in opposite
directions at that instant of time. Note that the two pulses in Figure 4.1.4 have
half of the amplitude of the initial pulse in Figure 4.1.2.
4.1. USING CHARACTERISTIC CURVES 57
−1 x
1
−1 x
1
We will first define the concept of characteristic curves for the PDE in (4.33).
The discussion here is analogous to the discussion on characteristic curves for
the second order equation in (3.17) on page 46. As in that discussion, the
58 CHAPTER 4. SOLVING PDES
−1 x
1
Observe that the ODE in (4.38) is equivalent to the system of first order ODEs:
dx
dt = a(x, y);
(4.39)
dy
= b(x, y).
dt
The system of ODEs in (4.39) defines the characteristic curves for the first–order
linear PDE in (4.33). Since, we are assuming that a and b ar C ∞ functions,
solutions to the system of first–order ODEs in (4.39) is guaranteed to have a
unique solution around to ∈ R subject to the the initial condition (x(to ), y(to )) =
(xo , yo ). Thus, in theory, characteristic curves for the PDE in (4.33) can always
be computed.
Suppose that we have computed the characteristic curves for the PDE in
(4.33) according to the system of ODEs in (4.39). Let one of those characteristics
be given by the parametrization
= ut + cux
= 0,
where we have used the Chain Rule, (4.45), and the assumption that u solves
the PDE in (4.44).
We can solve the ODE in (4.47) to obtain that
u(x, t) = constant along characteristic curves (4.48)
Since the characteristic curves in (4.46) are indexed by ξ, we can rewrite (4.48)
as
u(x, t) = F (ξ), (4.49)
where F is an arbitrary C 1 function of a real variable. Next, solve for ξ in (4.46)
and substitute into (4.49) to obtain the general solutions to the PDE in (4.44),
u(x, t) = F (x − ct), for x ∈ R and t ∈ R. (4.50)
For the case in which c > 0, (4.50) describes a traveling wave moving to the
right with speed c.
Finally, using the initial condition in (4.44), we get that
F (x) = f (x), for all x ∈ R,
so that
u(x, t) = f (x − ct), for x ∈ R and t ∈ R,
is a solution to the IVP in (4.44).
The method of characteristic curves illustrated thus far also applied to the
quasi–linear equation in (4.32). In this case, the equations to the characteristic
curves read
dx
dt = a(x, y, u);
(4.51)
dy
= b(x, y, u).
dt
Along characteristic curves u solves the ODE
du
= c(x, y, u). (4.52)
dt
In general, we might not be able to obtain an explicit formula for a solution of
the PDE in (4.32) based on the system (4.50)–(4.52). But, in some cases, we
might be able to obtain an expression that gives u(x, y) implicitly. We illustrate
this in the following example.
62 CHAPTER 4. SOLVING PDES
Here t is playing the role of y in the general discussion. In this case, the equation
for the characteristic curves is
dx
= u. (4.54)
dt
In order to solve the ODE in (4.54) we need information on the function u, which
is what ultimately we are trying too determine. The information is provided by
the differential equation that u satisfies along characteristic curves; namely,
du
= 0,
dt
which implies that u must be constant along characteristic curves. Thus, we
can set
u = F (ξ), (4.55)
where ξ is a parameter indexing the characteristic curves, and F is an arbitrary
C 1 function of a single variable. Substituting the expression for u in (4.55) into
the equation for the characteristic curves in (4.54) yields
dx
= F (ξ),
dt
which can be solved to yield the equation for the characteristic curves of the
PDE in (4.53):
x = F (ξ)t + ξ. (4.56)
Observe that in this case the characteristic curves are straight lines in the xt–
plane with x–intercept ξ and slope 1/F (ξ). Note that the slopes of the charac-
teristic curves depend on the value of the solution on the characteristic curves,
according to (4.55).
We can solve for ξ in (4.56) and use (4.55) to get
ξ = x − u(x, t)t
and then substitute this value into (4.55) to obtain an implicit formula for
u(x, t):
u(x, t) = F (x − u(x, t)t), for x ∈ R and t > 0. (4.57)
Next, use the initial condition in (4.53) to obtain from (4.57) that
so that
u(x, t) = f (x − u(x, t)t), for x ∈ R and t > 0,
provides an expression that defines u(x, t) implicitly.
4.1. USING CHARACTERISTIC CURVES 63
y = ξx, (4.65)
where ξ is a real parameter. Thus, the characteristic curves for the PDE in
(4.62) is a pencil of straight lines through the origin in R2 .
Now, along the characteristic curves for the PDE in (4.62), u solves the ODE
du
= 2u. (4.66)
dt
Next, combine the ODE in (4.66) with the first ODE in (4.63) to obtain the
ODE
du 2u
= , for x 6= 0. (4.67)
dx x
The ODE in (4.67) can be solved by separation of variables to yield
u = F (ξ)x2 , (4.68)
so that
x = ξ cos φ + η sin φ;
(4.73)
y = −ξ sin φ + η cos φ.
In view of the equations in (4.73), we can think of u as a function of ξ and η,
which we will denote by v(ξ, η); so that
where x and y on the right–hand side of (4.74) are given in terms of ξ and η in
(4.73).
Applying the Chain Rule, we obtain from (4.74) that
∂ξ ∂η
ux = vξ + vη ,
∂x ∂x
where
∂ξ ∂η
= cos φ and = sin φ, (4.75)
∂x ∂x
in view of the equations in (4.72). Thus,
Next, differentiate on both sides of (4.76) with respect to x and apply the Chain
Rule to get
∂ξ ∂η ∂ξ ∂η
uxx = cos φ vξξ + vξη + sin φ vηξ + vηη ,
∂x ∂x ∂x ∂x
so that, using (4.75) and the fact that the mixed second partial derivatives of
C 2 functions are equal,
Similarly, taking the partial derivative with respect to y on both sides of (4.77),
and using
∂ξ ∂η
= − sin φ and = cos φ,
∂y ∂y
which follow from (4.72), we obtain from (4.77) that
vξξ + vηη = 0,
which has the same form as Laplace’s equation. We therefore conclude that
Laplace’s equation in (4.69) is invariant under rotations. This suggests that we
look for solutions of (4.69) that are functions of a combination of the independent
variables that is independent of the rotation parameter φ. To obtain such a
combination, use (4.72) to compute
= x2 + y 2 ,
p
so that x2 +y 2 or x2 + y 2 are combinations of the independent variables, x and
y, that do not depend on φ, the rotation parameter; that is, they are rotationally
invariant. We will therefore look for solutions of the Laplace’s equation in (4.69)
that are of the form
p
u(x, y) = f ( x2 + y 2 ), for (x, y) ∈ R2 , (4.80)
of the form
u(x, y) = f (r), for (x, y) ∈ R2 , (4.82)
4.2. USING SYMMETRY TO SOLVE PDES 67
where p
r= x2 + y 2 , (4.83)
2
and f : (0, ∞) → R is a C function.
Write the expression in (4.83) r2 = x2 + y 2 and differentiate on both sides
with respect to x, applying the Chain Rule to get
∂r
2r = 2x,
∂x
from which we get that
∂r x
= , for r > 0. (4.84)
∂x r
Similar calculations show that
∂r y
= , for r > 0, (4.85)
∂y r
Next, use the Chain Rule to obtain from (4.82) that
∂r
ux = f 0 (r) ,
∂x
so that, by virtue of (4.84),
x 0
ux = f (r), for r > 0. (4.86)
r
Similar calculations using (4.82) and (4.85) yield
y 0
uy = f (r), for r > 0. (4.87)
r
Next, use the Product Rule, the Quotient Rule, and the Chain Rule to obtain
from (4.86) that
∂r ∂r
1 0 rf 00 (r) − f 0 (r)
uxx = f (r) + x ∂x ∂x ;
r r2
thus, using (4.84),
1 0 x2 x2
uxx = f (r) + 2 f 00 (r) − 3 f 0 (r), for r > 0. (4.88)
r r r
Similar calculations, using (4.85) and (4.87) yield
1 0 y2 y2
uyy = f (r) + 2 f 00 (r) − 3 f 0 (r), for r > 0. (4.89)
r r r
Next, add the expressions in (4.88) and (4.89) to obtain
2 0 x2 + y 2 00 x2 + y 2 0
uxx + uyy = f (r) + f (r) − f (r), for r > 0,
r r2 r3
68 CHAPTER 4. SOLVING PDES
Solution: Since the annulus Ω in (4.95) has radial symmetry, and the
boundary conditions in (4.96) are also radially symmetric, it makes sense to
look for radially symmetric solutions of problem (4.96). According to the result
of Example 4.2.1, these are of the form given in (4.94); namely
p
u(x, y) = c1 ln x2 + y 2 + c2 , for (x, y) ∈ Ω, (4.97)
c1 ln r1 + c2 = a (4.98)
and
c1 ln r2 + c2 = b, (4.99)
in view of (4.97). Solving the system of equations in (4.98) and (4.99) for c1
and c2 yields
b−a
c1 = ,
ln(r2 /r1 )
and
a ln r2 − b ln r1
c2 = .
ln(r2 /r1 )
Substituting these values for c1 and c2 into (4.97) yields a solution,
b−a p a ln r2 − b ln r1
u(x, y) = ln x2 + y 2 + , for (x, y) ∈ Ω, (4.100)
ln(r2 /r1 ) ln(r2 /r1 )
Setting
v(ξ, η) = u(x, y), (4.105)
where x and y are given in terms of ξ and η by the equations in (4.104), we
compute, using the Chain Rule, we obtain from (4.105) that
∂ξ ∂η
ux = vξ + vη ,
∂x ∂x
where, by virtue of the equations in (4.103),
∂ξ ∂η
=α and = 0,
∂x ∂x
so that
ux = αvξ . (4.106)
Similarly,
uy = βvη . (4.107)
Next, differentiate on both sides of (4.106) and apply the Chain Rules as in the
previous calculations to obtain
2s 0 s2
uxx = 2
f (s) + 2 f 00 (s), for x 6= 0. (4.121)
x x
Next, apply the Chain Rule to obtain from (4.116) that
∂s
uy = f 0 (s) , (4.122)
∂y
72 CHAPTER 4. SOLVING PDES
where
∂s 1
= , for x 6= 0. (4.123)
∂y x
It then follows from (4.122) and (4.123) that
1 0
uy = f (s), for x 6= 0. (4.124)
x
Differentiate on both sides of (4.124) with respect to y, apply the Chain Rule,
and use (4.123) to obtain
1 00
uyy = f (s), for x 6= 0. (4.125)
x2
Next, add the expressions in (4.121) and (4.125) to get
2s 0 1 + s2 00
uxx + uyy = 2
f (s) + f (s), for x 6= 0. (4.126)
x x2
It follows from (4.126) that, if u solves Laplace’s equation in R2 , then f solves
the second order ODE
2s 0 1 + s2 00
f (s) + f (s) = 0, for x 6= 0,
x2 x2
or
(1 + s2 )f 00 (s) + 2sf 0 (s) = 0. (4.127)
In order to solve the ODE in (4.127), set
so that
dv
(1 + s2 ) + 2sv = 0. (4.129)
ds
The first order ODE in (4.129) can be solved by separating variables to yield
Z Z
1 2s
dv = − ds,
v 1 + s2
or
1
ln |v| = ln + co , (4.130)
1 + s2
for some constant co .
Exponentiating on both sides of (4.130) and using the continuity of v we
obtain
c1
v(s) = , for s ∈ R, (4.131)
1 + s2
and some constant c1 . It follows from (4.128) and (4.131) that
c1
f 0 (s) = , for s ∈ R,
1 + s2
4.2. USING SYMMETRY TO SOLVE PDES 73
u = c1 θ + c2 ,
∂u ∂2u
= D 2, for x ∈ R and t > 0, (4.134)
∂t ∂x
where D > 0 is the diffusivity constant. We proceed as in Section 4.2.2 by finding
conditions on parameters α and β so that the diffusion equation in (4.134) is
invariant under the change of variables
ξ = αx;
(4.135)
τ = βt,
where αβ 6= 0.
Write
v(ξ, τ ) = u(x, t), (4.136)
where x and t are given in terms of ξ and τ by inverting the system in (4.136),
x = ξ/α;
t = τ /β,
so that
ux = αvξ . (4.137)
Similarly,
ut = βvτ . (4.138)
Next, differentiate on both sides of (4.137) and apply the Chain Rules as in the
previous calculations to obtain
Hence, the diffusion equation in (4.134) is invariant under the change of variables
in (4.136) provided that
β = α2 . (4.141)
It follows from (4.140) and (4.141) that the diffusion equation in (4.134) is
invariant under the dilation
ξ = αx;
(4.142)
τ = α2 t.
It follows from (4.142) that combinations of the variables that are independent
of the dilation parameter, α, are
ξ2 x2 ξ x
= or √ =√ , for τ > 0 and t > 0.
τ t τ t
Thus, in order to find dilation–invariant solutions of the one–dimensional diffu-
sion equation, we look for solutions of the form
x
u(x, t) = f √ , for t > 0, (4.143)
t
so that
dv s
+ v = 0. (4.154)
ds 2D
76 CHAPTER 4. SOLVING PDES
The first order ODE in (4.154) can be solved by separating variables to yield
Z Z
1 s
dv = − ds,
v 2D
or
s2
ln |v| = − + co , (4.155)
4D
for some constant co .
Exponentiating on both sides of (4.155) and using the continuity of v we
obtain 2
v(s) = c1 e−s /4D , for s ∈ R, (4.156)
and some constant c1 . It follows from (4.156) and (4.153) that
2
f 0 (s) = c1 e−s /4D
, for s ∈ R,
and constants c1 and c2 . It follows from (4.143) and (4.157) that dilation–
invariant solutions of one–dimensional diffusion equation in (4.134) are of the
form
Z x/√t
2
u(x, t) = c1 e−z /4D dz + c2 , for x ∈ R and t > 0, (4.158)
0
Lu = 0,
77
78 CHAPTER 5. SOLVING LINEAR PDES
and
u(L, t) = 0, for t > 0, (5.3)
and the initial conditions
and
ut (x, 0) = 0, for x ∈ [0, L], (5.5)
where f : [0, L] → R is a given, real valued, continuous function defined on the
interval [0, L].
We recall that the constant c in (5.1) is given by
τ
c2 = , (5.6)
ρ
where ρ is the constant density of the string in units of mass per length, and τ
is the constant tension of the string in units of force.
The PDE in (5.1), together with the boundary conditions in (5.2) and (5.3),
and the initial conditions in (5.4) and (5.5), models the small amplitude vibra-
tions of a taut string of length L that is fixed at the end–points 0 and L of the
interval [0, L], that is plucked from rest from an initial shape given by the graph
of the function f . In this section we construct a solution of this initial–boundary
value problem, which we write in concise form as
utt − c2 uxx = 0, for x ∈ (0, L) and t > 0;
u(0, t) = u(L, t) = 0, for t > 0;
(5.7)
u(x, 0) = f (x), for x ∈ [0, L];
ut (x, 0) = 0, for x ∈ [0, L].
extended to a continuous function on [0, ∞). We will also require that these
solutions satisfy the boundary conditions in (5.2) and (5.3).
Once we find solutions of (5.1) of the form given in (5.8) that satisfy the
boundary conditions in (5.2) and (5.3), we take advantage of the fact that the
PDE in (5.1) is linear to construct linear combinations of these solutions, which
will also be solutions, to approximate solutions that satisfy the initial conditions
in (5.4) and (5.5). This is the general outline of the procedure to be used in
this section, which is know in the literature as the method of separation of
variables.
and
utt (x, t) = y(x)h00 (t), for x ∈ (0, L), and t > 0. (5.10)
If we require that u be a solution of the PDE in (5.1), substituting the expression
for uxx in (5.9) and for utt in (5.10) into (5.1) then yields
and
h00 (t) + λc2 h(t) = 0, for t > 0. (5.14)
Next, if u is to satisfy the boundary conditions in (5.2) and (5.3), it must
be the case that
y(0)h(t) = 0 and y(L)h(t) = 0, for all t > 0. (5.15)
At this point we make the observation that the function defined by
u(x, t) = 0, for all x ∈ [0, L] and t > 0, (5.16)
is a solution of the PDE in (5.1). We call it the trivial solution of the equation.
We also note that the function in (5.16) satisfies the boundary conditions in (5.2)
and (5.3) and the initial condition in (5.5). However, for a general function
f : [0, L] → R, the trivial function in (5.16) does not satisfy the initial condition
in (5.4). Thus, if we are to construct a solution of the initial–boundary problem
in (5.7), it cannot be the trivial solution. Hence, we look for nontrivial solutions
of the PDE in (5.1) of the form in (5.8).
Consequently, if u ∈ C 2 ([0, L] × [0, ∞), R) of the form given in (5.8) is a
nontrivial solution of the PDE in (5.1) that also satisfies the boundary conditions
in (5.15), then it must be the case that
y(0) = 0 and y(L) = 0. (5.17)
To see why this is the case, use the assumption that u is nontrivial to conclude
that there exists to > 0 such that h(to ) 6= 0 and, according to (5.15),
y(0)h(to ) = 0 and y(L)h(to ) = 0. (5.18)
We can see now that (5.17) follows from (5.18) and the fact that h(to ) 6= 0.
We therefore look for nontrivial solutions of the the two–point, boundary
value problem (BVP)
00
y + λy = 0, for x ∈ (0, L);
y(0) = 0; (5.19)
y(L) = 0.
b = 0,
By the same token, since y also has to satisfy the boundary condition
y(L) = 0
aL = 0,
we get that
a = 0,
since L > 0. Consequently, in view of (5.21),
y(x) = 0, for x ∈ R,
y 00 − µ2 y = 0. (5.22)
m2 − µ2 = 0,
which has roots µ and −µ; hence, according to (A.5) in Appendix A.1, the
general solution of the ODE in (5.22) is
The homogeneous system in (5.24) has only the trivial solution if and only if
1 1
det µL 6= 0, (5.25)
e e−µL
where
1 1
det = e−µL − eµL .
eµL e−µL
1 1
det =0
eµL e−µL
if and only if
eµL = e−µL
if and only if
e2µL = 1
if and only if
2µL = 0,
from which we conclude that µ = 0, which contradicts the assumption that
µ > 0. Hence, (5.25) must hold true, from which we conclude that the system
in (5.24) has only the trivial solution
c1 = c2 = 0;
which shows that, if λ < 0, then the BVP in (5.19) has only the trivial solution.
It remains to consider the case λ > 0. Thus, assume that λ = ω 2 , where
ω > 0. In this case, the ODE in BVP (5.19) becomes
y 00 + ω 2 y = 0, (5.26)
m2 + ω 2 = 0,
The boundary condition y(0) = 0 in BVP (5.19) implies from (5.27) that c1 = 0;
so that, in view of (5.27),
c2 sin ωL = 0. (5.29)
ωL = nπ, for n ∈ Z,
m2 + λn c2 , for n = 1, 2, 3, . . . ,
√ √
which has complex conjugate roots λn ci and − λn ci. Consequently, by (A.6)
in Appendix A, the general solutions of the ODEs in (5.35) are given by
nπct nπct
hn (t) = an cos + bn sin , (5.36)
L L
84 CHAPTER 5. SOLVING LINEAR PDES
The assumption that the series in the integrand on the left–hand side of (5.47)
justifies the term–by–term integration of the series to get
X∞ Z L nπx mπx Z L mπx
cn sin sin dx = f (x) sin dx. (5.48)
n=1 0 L L 0 L
To evaluate the integrals on the left–hand side on (5.48), we use the trigono-
metric identity
1
sin A sin B = [cos(A − B) − cos(A + B)] , for A, B ∈ R. (5.49)
2
Applying (5.49) to the integrands on the left–hand side of (5.48), we get that
nπx mπx 1
(n − m)πx
(n + m)πx
sin sin = cos − cos , (5.50)
L L 2 L L
for m, n ∈ N. We integrate each of the terms on the right–hand side of (5.50)
from 0 to L separately. First, integrate the right–most term in (5.50) to get
Z L L
(n + m)πx L (n + m)πx
cos dx = sin = 0,
0 L (n + m)π L 0
2 L
Z nπx
cn = f (x) sin dx, for n ∈ N, (5.55)
L 0 L
for computing the coefficients in the Fourier series expansion for f in (5.46).
Given f ∈ C([0, L], R), define
N
X nπx
SN (x) = cn sin , for x ∈ [0, L], (5.56)
n=1
L
where SN is given in (5.56), and for which the convergence in (5.57) is uniform,
so that the calculations leading to (5.55) are justified.
If the statement in (5.57) is true, we write
∞
X nπx
cn sin = f (x), for x ∈ [0, L], (5.58)
n=1
L
where the integral in (5.59) denotes the Riemann integral. If this is the case,
we can the define Fourier coefficients of f as follows:
Z L
1
ao = f (x) dx; (5.60)
2L −L
Z L
1 nπx
an = f (x) cos dx, for n = 1, 2, 3, . . . ; (5.61)
L −L L
and
Z L
1 nπx
bn = f (x) sin dx, for n = 1, 2, 3, . . . . (5.62)
L −L L
Example 5.1.1. Let f : [0, L] → R be the function giving the first initial con-
dition in the problem (5.7). Since we are assuming that f (0) = f (L) = 0, f can
be extended to a continuous, odd, periodic function of period 2L be defining
and
f (x + 2L) = f (x), for all x ∈ R.
In this case, the formula for the Fourier coefficients of f in (5.60), (5.61) and
(5.62 yield
an = 0, for n = 0, 1, 2, 3, . . .
and
Z L
2 nπx
bn = f (x) sin dx, for n = 1, 2, 3, . . . , (5.64)
L 0 L
since f is odd and therefore the integrand in (5.62) is even.
Thus, in this case, the Fourier series expansion of f yields
∞
X nπx
bn sin , for x ∈ [0, L],
n=1
L
where the coefficients bn , for n ∈ N are given in (5.64). This is the same
expansion given in (5.46) with cn in place of bn .
5.1. SOLVING THE VIBRATING STRING EQUATION 89
where SN is given by (5.65). If (5.66) holds true, we say that the Fourier series
expansion for f given in (5.63) converges to f pointwisely, and write
X∞ h nπx nπx i
ao + an cos + bn sin = f (x), for x ∈ [−L, L]. (5.67)
n=1
L L
N
!
Z L
X 1 nπy nπx
+ f (y) cos dy cos (5.68)
n=1
L −L L L
N
!
Z L
X 1 nπy nπx
+ f (y) sin dy sin ,
n=1
L −L L L
for x ∈ [0, L], where we have used y as the variable of integration in the definition
of the Fourier coefficients because the variable x is being used as the argument
of SN .
Next, interchange summation and integration and use the distributive prop-
erty to in the last two terms on the right–hand side of (5.68) to get
Z L N h
1 X nπx nπy nπx nπy i
f (y) cos cos + sin sin dy. (5.69)
L −L n=1
L L L L
The term in the summation in (5.69) can be simplifies using the trigonometric
identity
to yield
L N
nπ(x − y)
Z
1 X
f (y) cos dy. (5.70)
L −L n=1
L
Consequently, the expression in (5.68) can be rewritten as
N !
1 L
nπ(x − y)
Z
1 X
SN (x) = f (y) + cos dy, for x ∈ [0, L], (5.71)
L −L 2 n=1 L
as Z L
1
SN (x) = f (y)DN (x − y) dy, for x ∈ [0, L]. (5.73)
L −L
The function defined in (5.72) is called the Dirichlet kernel. Observe that
DN is even and periodic of period 2L. Furthermore, it can be shown that
1 πθ
sin N +
2 L
DN (θ) = , for θ 6= 0. (5.74)
πθ
2 sin
2L
z =x−y
or Z x+L
1
SN (x) = f (x − z)DN (z) dz, for x ∈ [−L, L];
L x−L
5.1. SOLVING THE VIBRATING STRING EQUATION 91
so that,
Z L
1
SN (x) = f (x − z)DN (z) dz, for x ∈ [−L, L], (5.77)
L −L
In the proof of Theorem 5.1.2 we will use the following result known as the
Riemann–Lebesgue Lemma. We state here a spacial form of the result in the
context of 2L–periodic functions and the Fourier coefficients.
Theorem 5.1.3. Assume that F : R → R is bounded, 2L–periodic, and inte-
grable over [−L, L]. Let an , for n = 0, 1, 2, 3, . . ., and bn , for n = 1, 2, 3, . . .,
denote the Fourier coefficients of F given in (5.60)–(5.62), with F in place of F .
Then,
lim an = 0 and lim bn = 0. (5.80)
n→∞ n→∞
and Z L
N πy
lim F (y) sin dy = 0. (5.82)
N →∞ −L L
A more general version of the Riemann–Lebesgue Lemma is the following.
Theorem 5.1.4. Assume that F : [a, b] → R is absolutely integrable on [a, b];
so that,
Z b
|F (x)| dx < ∞.
a
92 CHAPTER 5. SOLVING LINEAR PDES
Then,
Z b
lim F (x) cos(λx) dx = 0, (5.83)
|λ|→∞ a
and Z b
lim F (x) sin(λx) dx = 0. (5.84)
|λ|→∞ a
For a proof of this result, see [Tol62]. We presents two examples that the
the result in Theorem 5.1.4 plausible.
b
1
= sin(λx) ;
λ a
so that
Z b
1
F (x) sin(λx) dx = [sin(λb) − sin(λa)], for λ 6= 0. (5.85)
a λ
Take absolute value of both sides of (5.85) and apply the triangle inequality to
get
Z b
2
F (x) cos(λx) dx 6 , for λ 6= 0,
a |λ|
from which (5.83) follows for the case F (x) = 1 for all x.
Similar calculations show that (5.84) is also true in this case.
that is, F 0 is absolutely integrable over [a, b]. Then, (5.83) and (5.84) hold true.
or Z b
1
F (x) cos(λx) dx = [F (b) sin(λb) − F (a) sin(λa)]
a λ
(5.87)
Z b
1 0
− F (x) sin(λx) dx.
λ a
Next, take absolute value of both sides of (5.87) and apply the triangle inequality
to get
Z b Z b
2M 1
F (x) cos(λx) dx 6 + |F 0 (x)| dx, for λ 6= 0, (5.88)
a |λ| |λ| a
where M is an upper bound of F on [a, b]. In view on (5.86), we can see that
(5.83) follows from (5.88).
Similar calculations show that (5.84) is also true in this case.
1 L
Z
f (x) = f (x)DN (z) dz, for x ∈ [−L, L]. (5.89)
L −L
1 L
Z
SN (x) − f (x) = [f (x − z) − f (x)]DN (z) dz, for x ∈ [−L, L], (5.90)
L −L
Next, we use the formula for the Dirichlet kernel in (5.74) to write (5.90) as
1 L f (x − z) − f (x)
Z
1 πz
SN (x) − f (x) = sin N + dz, (5.91)
L −L 2 sin πz
2 L
2L
for x ∈ [−L, L].
Write
f (x − z) − f (x)
G(x, z) = πz , for z 6= 0 and x, z ∈ [−L, L], (5.92)
2 sin
2L
and observe that, using L’Hospital’s Rule,
f (x − z) − f (x) L
lim πz = − f 0 (x),
z→0
2 sin π
2L
where we have used the assumption that f is a C 1 function. Consequently,
defining
L
G(x, 0) = − f 0 (x), for x ∈ [−L, L], (5.93)
π
94 CHAPTER 5. SOLVING LINEAR PDES
We then see that the function G defined in (5.92) and (5.93), for x ∈ [−L, L]
and z ∈ [−L, L], is continuous on [−L, L] × [−L, L]. (Observe that denominator
in (5.92) is zero only at z = 0 in [−L, L].) Thus, the function G is bounded and
continuous on [−L, L] × [−L, L], and therefore the map
is absolutely integrable for each x ∈ [−L, L]. (We are aiming here at applying
the Riemann–Lebesgue Lemma as stated in (5.81) and (5.82).)
Using G(x, z) as defined in (5.92) and (5.93), we can rewrite (5.91) as
Z L
1 1 πz
SN (x) − f (x) = G(x, z) sin N+ dz,
L −L 2 L
since G(x, z) is absolutely integrable over [−L, L] for each x ∈ [−L, L]. Thus,
we can apply the result of the Riemann–Lebesgue Lemma in (5.82) to deduce
that
Z L πz
N πz
lim G(x, z) cos sin dz = 0, for each x ∈ [−L, L]. (5.95)
N →∞ −L 2L L
from which (5.79) follows. We have therefore completed the proof of Theorem
5.1.2.
5.1. SOLVING THE VIBRATING STRING EQUATION 95
N > Nε ⇒ |SN (x) − f (x)| < ε, for all x ∈ [−L, L]. (5.97)
We will prove this as a consequence of the fact that f ∈ C 1 (R, R); so that, f 0 is
bounded on [−L, L]; that is, there exists a constant M > 0 such that
Our goal is to show that, given any ε > 0, we can find Nε ∈ N such that the
absolute value of the integral on the right–hand side of (5.99) is less than εL for
all N > Nε and all x ∈ [−L, L].
Before we proceed any further, let’s make the change of variables y = −z in
the integral on the right–hand side of (5.99) and use the fact that the Dirichlet
kernel DN is an even function to write
Z L
1
SN (x) − f (x) = [f (x + y) − f (x)]DN (y) dy, for x ∈ [−L, L]. (5.100)
L −L
Let ε > 0 be given and let δ > 0 (to be chosen later) be such that δ < L.
Write the expression on the right–hand side of (5.100) as the sum of the
three terms
1 −δ
Z
I1 (x) = [f (x + y) − f (x)]DN (y) dy, for x ∈ [−L, L], (5.101)
L −L
Z δ
1
I2 (x) = [f (x + y) − f (x)]DN (y) dy, for x ∈ [−L, L], (5.102)
L −δ
and
Z L
1
I3 (x) = [f (x + y) − f (x)]DN (y) dy, for x ∈ [−L, L]; (5.103)
L δ
so that,
L πy/L
Z
1 1
VN (y) = sin N + r dr,
π 0 2 2 sin(r/2)
L πy/L
Z
1 1 1
VN (y) = sin N + r − dr
π 0 2 2 sin(r/2) r
(5.108)
1
Z πy/L sin N+ r
L 2
+ dr.
π 0 r
5.1. SOLVING THE VIBRATING STRING EQUATION 97
Define
1 1
− , if r 6= 0 and r ∈ [−π, π];
2 sin(r/2) r
g(r) = (5.109)
0, if r = 0.
L πy/L
Z
1
VN (y) = g(r) sin N + r dr
π 0 2
(5.110)
L (N +1/2)πy/L sin t
Z
+ dt,
π 0 t
1
where we have used the change of variables t = N + r in the second
2
integral on the right–hand side of (5.108).
Next, since
Z (N +1/2)πy/L Z ∞
sin t sin t π
lim dt = dt = ,
N →∞ 0 t 0 t 2
the second integral on the right–hand side of (5.110) is bounded.
We can bound the first integral on the right hand side of (5.110) as follows:
Z πy/L
1
g(r) sin N + r dr 6 π max |g(r)|, for y ∈ [−L, L].
0 2 y∈[−π,π]
The first term in (5.112) can be estimated using the Mean Value Theorem and
the estimates in (5.107) and (5.98) as follows:
where we have used (5.111), (5.112), (5.113), (5.114) and the triangle inequality.
Consequently, the integral I2 in (5.102) can be estimated using (5.115) as
4M M1 δ
|I2 (x)| 6 , for all x ∈ [−L, L]. (5.116)
L
We choose δ > 0 such that
4M M1 δ ε
< ,
L 3
or
εL
δ< .
12M M1
We then obtain from (5.116) that
ε
|I2 (x)| < , for all x ∈ [−L, L]. (5.117)
3
Next, we estimate I3 in (5.103).
Use the formula for the Dirichlet kernel in (5.74) to write (5.103) as
1 L f (x + y) − f (x)
Z
1 πy
I3 (x) = sin N + dy, (5.118)
L δ 2 sin(πy/2L) 2 L
f (x + y) − f (x)
F (x, y) = , for δ 6 y 6 L and − L 6 x 6 L. (5.119)
2 sin(πy/2L)
Observe that the function F defined in (5.119) is a C 1 function over the rectangle
[−L, L] × [δ, L]. Hence, there exist positive constants M2 and M3 such that
and
∂
[F (x, y)] 6 M3 , for all δ 6 y 6 L and − L 6 x 6 L. (5.121)
∂y
With the definition of F given in (5.119), we can rewrite (5.118) as
1 L
Z
1 πy
I3 (x) = F (x, y) sin N + dy, for x ∈ [−L, L]. (5.122)
L δ 2 L
Next, let
1 π
N+ =λ (5.123)
2 L
in (5.122) to get
Z L
1
I3 (x) = F (x, y) sin λy dy, for x ∈ [−L, L]. (5.124)
L δ
and for λ 6= 0, where we have used the estimates in (5.120) and (5.121).
Combining (5.124), (5.123) and the estimate in (5.125) we obtain
2M2 + M3 L
|I3 (x)| 6 , for all x ∈ [−L, L]. (5.126)
π(N + 1/2)
It follows from (5.126) that, if
6M2 + 3M3 L
N> ,
πε
then
ε
|I3 (x)| <
, for all x ∈ [−L, L].
3
Thus, there exists N1 ∈ N such that
ε
N > N1 ⇒ |I3 (x)| < , for all x ∈ [−L, L]. (5.127)
3
Similar calculations show that there exists N2 ∈ N such that
ε
N > N2 ⇒ |I1 (x)| < , for all x ∈ [−L, L]. (5.128)
3
Letting Nε = max{N1 , N2 } and combining (5.104), (5.117), (5.127) and (5.128),
we obtain that
N > Nε ⇒ |SN (x) − f (x)| < ε, for all x ∈ [−L, L],
which proves that (SN ) converges uniformly to f as N → ∞.
100 CHAPTER 5. SOLVING LINEAR PDES
In this section we will see an alternate proof of the limit facts in (5.131 based
in the fact that
∞
X
(|an | + |bn |) < ∞, (5.132)
n=1
N h
X nπx nπx i
SN (x) = ao + an cos + bn sin , (5.135)
n=1
L L
Z L nπx
sin dx = 0, for all n ∈ N, (5.138)
−L L
Z L nπx mπx L,
if n = m;
cos cos dx = (5.139)
−L L L
0, if n 6= m,
Z L nπx mπx L,
if n = m;
sin sin dx = (5.140)
−L L L
0, if n 6= m,
and
Z L nπx mπx
sin cos dx = 0, for all m, n ∈ N, (5.141)
−L L L
We begin with
N h
!
X nπx nπx i
2
(SN (x)) = ao + an cos + bn sin
n=1
L L
N h
!
X mπx mπx i
× ao + am cos + bm sin ,
m=1
L L
N
X mπx N
X nπx
+ ao bm sin + ao an cos
m=1
L m=1
L
X N
N X nπx mπx
+ an am cos cos
n=1 m=1
L L
X N
N X nπx mπx
+ an bm cos sin (5.142)
n=1 m=1
L L
N
X nπx
+ ao bn sin
m=1
L
X N
N X nπx mπx
+ am bn sin cos
n=1 m=1
L L
X N
N X nπx mπx
+ bn bm sin sin .
n=1 m=1
L L
Next, integrate on both sides of (5.142) from −L to L and use the integration
facts in (5.137)–(5.141) to obtain that
Z L N
X N
X
(SN (x))2 dx = 2La2o + La2n + Lb2n ,
−L n=1 n=1
or
Z L N
X
2
2La2o a2n + b2n ,
(SN (x)) dx = +L for N = 1, 2, 3, . . . (5.143)
−L n=1
5.1. SOLVING THE VIBRATING STRING EQUATION 103
Next, we compute
N h
!
Z L Z L X nπx nπx i
f (x)SN (x) dx = f (x) ao + an cos + bn sin dx,
−L −L n=1
L L
" N
#
Z L X nπx
+ f (x) an cos dx
−L n=1
L
" N
#
Z L X nπx
+ f (x) bn sin dx,
−L n=1
L
or
Z L Z L
f (x)SN (x) dx = ao f (x) dx
−L −L
N
X Z L nπx
+ an f (x) cos dx (5.144)
n=1 −L L
N
X Z L nπx
+ bn f (x) sin dx.
n=1 −L L
or
Z L N
X
f (x)SN (x) dx = 2La2o + L a2n + b2n .
(5.145)
−L n=1
to rewrite (5.136) as
N
!
Z L Z L X
2 2
2La2o a2n b2n
(SN (x) − f (x)) dx = (f (x)) dx − +L + , (5.146)
−L −L n=1
for N = 1, 2, 3, . . .
104 CHAPTER 5. SOLVING LINEAR PDES
Note that, if f is square–integrable over [−L, L], the left–hand side of (5.147) is
monotone, increasing sequence that is bounded above. Hence, it has a limit as
N → ∞. We therefore have that
∞
1 L
X Z
2 2 2
(f (x))2 dx.
2ao + an + bn 6 (5.148)
n=1
L −L
for the case in which f is square integrable over [−L, L]. We have therefore
proved Proposition 5.1.8.
Next, assume that f : R → R is differentiable, 2L–periodic and with square–
integrable derivative over [−L, L]. Thus, letting a0n and b0n , for n ∈ N, denote
the Fourier coefficients of f 0 , we obtain from Proposition 5.1.8 applied to f 0 that
∞
X
(a0n )2 + (b0n )2 < ∞.
(5.149)
n=1
L 1 0
|an | = · |b |, for n = 1, 2, 3, . . . , (5.152)
π n n
5.1. SOLVING THE VIBRATING STRING EQUATION 105
and
L 1 0
|bn | = · |a |, for n = 1, 2, 3, . . . (5.153)
π n n
Then, using the inequality
1 2 1 2
|x||y| 6 x + y , for all x, y ∈ R,
2 2
and
L 1 0 2
|bn | 6 + (an ) , for n = 1, 2, 3, . . . (5.155)
2π n2
Adding the estimates in (5.154) and (5.155) we obtain
L 2 0 2 0 2
|an | + |bn | 6 + (an ) + (b n ) , for n = 1, 2, 3, . . .
2π n2
Consequently,
N
" N N
#
X L X 1 X
0 2 0 2
(|an | + |bn |) 6 2 + (an ) + (bn ) , (5.156)
n=1
2π n=1
n2 n=1
for N = 1, 2, 3, . . .
∞
X 1
Now, since the series converges, it follows from Proposition 5.1.8
n=1
n2
that, in the case that f 0 is square–integrable over [−L, L], the sequences on the
right–hand side of (5.156) converge and
N
" ∞ ∞
#
X L X 1 X
0 2 0 2
(|an | + |bn |) 6 2 + (an ) + (bn ) , (5.157)
n=1
2π n=1
n2 n=1
which is the claim mane in (5.132). We summarize this result in the following
proposition.
106 CHAPTER 5. SOLVING LINEAR PDES
Next, we will show that the summability condition in (5.159) will imply that
the sequence (SN ) converges uniformly to f over [−L, L] as N → ∞. This
result will be attained as a consequence of the Weierstrass Majorant Theorem
or Weierstrass M–Test.
and
∞
X
Mn < ∞.
n=1
(See, for example, [Rud53, Theorem 7.10, pg. 119] for a proof of the Weier-
strass M–Test).
We will also need the following lemma.
and
g(x) 6 1, for − π 6 x 6 xo − δ or xo + δ 6 x 6 π. (5.167)
Furthermore, g(x) is a linear combination of the set {1, cos x, sin x}. To see this,
use the trigonometric identity
to write
1 1
sin2 x = − cos(2x), for all x ∈ R,
2 2
and
2 sin x cos x = sin(2x), for all x ∈ R,
2
we can see from (5.166) that (g(x)) is a linear combination of elements from
the set
{1, cos x, sin x, cos(2x), sin(2x)}.
Similarly, noting that (g(x))2 = g(x)(g(x))2 and using the trigonometric iden-
tities
1 1
cos x cos(2x) = cos x + cos(3x), for all x ∈ R,
2 2
1 1
sin x sin(2x) = cos x1 cos(3x), for all x ∈ R,
2 2
1 1
cos x sin(2x) = sin x + sin(3x), for all x ∈ R,
2 2
and
1 1
sin x cos(2x) = − sin x + sin(3x), for all x ∈ R,
2 2
we can see that (g(x))3 is a linear combination of of elements from the set
{1, cos x, sin x, cos(2x), sin(2x), cos(3x), sin(3x), cos(4x), sin(4x), . . .},
for all n ∈ N. (Recall that the span of a set is the collection of all finite a linear
combination of of elements from the set).
It then follows from the assumptions in (5.160), (5.161) and (5.162) that
Z π
h(x)(g(x))n dx = 0, for all n ∈ N. (5.169)
−π
Z xo +δ
+ h(x)(g(x))n dx (5.170)
xo −δ
Z π
+ h(x)(g(x))n dx,
xo +δ
for all n ∈ N, and estimating the first and last integral in the right–hand side
of (5.170) as follows.
5.1. SOLVING THE VIBRATING STRING EQUATION 109
so that Z π
h(x)(g(x))n dx 6 2M π, for all n ∈ N, (5.171)
xo +δ
where
M = max |h(x)| dx. (5.172)
x∈[−π,π]
Similarly,
Z xo −δ
h(x)(g(x))n dx 6 2M π, for all n ∈ N, (5.173)
−π
Next, let
r= min g(x). (5.176)
xo −δ/26x6xo +δ/2
We can now see that (5.180) is in direct contradiction with (5.169). Conse-
quently, we conclude that h(x) = 0 for all x ∈ [−π, π], if it is continuous and all
its Fourier coefficients are 0.
and Z π
h(x) sin(nx) dx = 0, for all n ∈ N,
−π
5.1. SOLVING THE VIBRATING STRING EQUATION 111
respectively.
We have therefore shown that the hypotheses in Lemma 5.1.12 are satisfied
for h given in (5.184), from which we obtain that f (y) = 0 for all y ∈ R.
where
nπx nπx
gn (x) = an cos + bn sin , for x ∈ [−L, L] and n ∈ N. (5.187)
L L
Then, by the triangle inequality, we obtain from (5.187) that
where the series on the right–hand side of (5.190) converges absolutely and
uniformly on [−L, L]. Consequently, the Fourier coefficients of g are an and bn ,
112 CHAPTER 5. SOLVING LINEAR PDES
for n ∈ N, which are the same as the Fourier coefficients of f for n ∈ N. Observe
also that Z L
g(x) dx = 0. (5.191)
−L
where Z L
2 nπx
bn = f (x) sin dx, for n ∈ N. (5.198)
L 0 L
We can then use the Weierstrass M-Test (Theorem 5.1.11) to deduce that the
series
∞ nπx
X nπct
bn sin cos , (5.199)
n=1
L L
where bn , for n ∈ N are given in (5.198), converges absolutely and uniformly for
x ∈ [0, L] and t > 0. Indeed, setting
nπx
nπct
gn (x, t) = bn sin cos , for x ∈ [0, L], t > 0 (5.200)
L L
and n ∈ N, we see that
|gn (x, t)| 6 |bn |, for x ∈ [0, L], t > 0 and n ∈ N.
Consequently, in view of (5.197), the Weierstrass M–Test applies. We therefore
deduce that
∞
X
gn (x, t) (5.201)
n=1
converges absolutely and uniformly for x ∈ [0, L] and t > 0. Thus, in view of
(5.200) and (5.201), we see that the series in (5.199) converges absolutely and
uniformly for x ∈ [0, L] and t > 0.
Define
∞ nπx
X nπct
u(x, t) = bn sin cos , for x ∈ [0, L], t > 0. (5.202)
n=1
L L
We have therefore shown that the function u : [0, L]×[0, ∞) → R given in (5.202)
is well–defined. In particular, we get from (5.202) that
∞
X nπx
u(x, 0) = bn sin , for x ∈ [0, L]. (5.203)
n=1
L
114 CHAPTER 5. SOLVING LINEAR PDES
It follows from (5.198) and Theorem 5.1.14 that the series on the right–hand
side of (5.203) converges to f (x), for x ∈ [0, L]. Hence,
This is the first of the initial conditions in the vibrating string problem in (5.7).
We note also that the function u defined in (5.202) also satisfies the boundary
conditions in (5.7):
To see whether or not u in (5.202) satisfies the second initial condition in problem
(5.7), we first need to see that u is differentiable with respect to t and that
∞
∂ X πc nπx nπct
[u(x, t)] = − nbn sin sin , (5.205)
∂t n=1
L L L
for x ∈ [0, L], t > 0. This will require to determine conditions on f for which
that series
∞ nπx
X nπct
nbn sin sin (5.206)
n=1
L L
converges for x ∈ [0, L] and t > 0. We will answer these questions in a subse-
quent example.
the Weierstrass M-Test would imply the absolute and uniform convergence of
the series in (5.206)
Consider the general situation of a 2L–periodic function, f : R → R. Assume
now that f is twice differentiable, with second derivative, f 00 , that is square–
integrable on [−L, L]; so, that
Z L
|f 00 (x)|2 dx < ∞. (5.208)
−L
Let a0n and b0n , for n ∈ N, denote the Fourier coefficients of f 0 . Applying the
result of Proposition 5.1.9 to f 0 , instead of f , we obtain that
∞
X
(|a0n | + |b0n |) < ∞, (5.209)
n=1
in view of (5.208).
5.1. SOLVING THE VIBRATING STRING EQUATION 115
2 L
Z nπx
bn = f (x) sin dx, for n ∈ N, (5.217)
L 0 L
116 CHAPTER 5. SOLVING LINEAR PDES
and
∞
∂ X π nπx nπct
[u(x, t)] = nbn cos cos , (5.219)
∂x n=1
L L L
respectively.
The series in (5.218) and (5.219) converge absolutely and uniformly, by the
Weierstrass M–Test, provided that we can show that
∞
X
n|bn | < ∞. (5.220)
n=1
However, (5.220) follows from (5.215) in Proposition 5.1.17 because we are as-
suming that f 00 is square integrable over [−L, L]. We therefore conclude that
the series in (5.218) and (5.219) converge absolutely and uniformly for x ∈ [0, L]
and t > 0. In particular, we obtain from (5.218) that
so that the function u defined in (5.216) satisfies the second of the initial con-
ditions in the vibrating string problem in (5.7).
the the function u defined in (5.216) has second partial derivatives, utt and uxx ,
given by
∞
π 2 c2 2 nπx
X nπct
utt (x, t) = − n bn sin cos , (5.221)
n=1
L2 L L
in view of (5.224).
Next, use the identities in (5.210) and (5.211), applied to a00n and b00n , to get
that
nπ 0
a00n = b , for n = 1, 2, 3, . . . (5.227)
L n
and
nπ
b00n = − a0n , for n = 1, 2, 3, . . . (5.228)
L
118 CHAPTER 5. SOLVING LINEAR PDES
Using the identities in (5.210) and (5.211) again, we obtain from (5.227) and
(5.228) that
π2
a00n = − 2 n2 an , for n = 1, 2, 3, . . .
L
and
π2
b00n = − 2 n2 bn , for n = 1, 2, 3, . . . ,
L
from which we get
π2 2
|a00n | = n |an |, for n = 1, 2, 3, . . . (5.229)
L2
and
π2 2
|b00n | =
n |bn |, for n = 1, 2, 3, . . . . (5.230)
L2
It follows from (5.229) and (5.230) that
L2
n2 (|an | + |bn |) = (|a00n | + |b00n |) , for n = 1, 2, 3, . . . (5.231)
π2
In view of (5.231) we see that the assertion in (5.225) follows from (5.226), and
the proof of the proposition is now complete.
has a solution.
Proof: Define u : [0, L] → R by
∞ nπx
X nπct
u(x, t) = bn sin cos , for x ∈ [0, L], t > 0, (5.233)
n=1
L L
2 L
Z nπx
bn = f (x) sin dx, for n ∈ N, (5.234)
L 0 L
the Fourier coefficients the odd, 2L–periodic extension of f : [0, L] → R.
5.2. FUNDAMENTAL SOLUTIONS 119
or Z ∞ √
v(x, t) dx = c1 4Dπ, for all t > 0. (5.238)
−∞
that is,
1
c1 = √ . (5.239)
4Dπ
Substituting the value of c1 in (5.239) into the definition of v(x, t) in (5.237),
we obtain
1 2
v(x, t) = √ e−x /4Dt , for x ∈ R and t > 0. (5.240)
4πDt
5.2. FUNDAMENTAL SOLUTIONS 121
We shall denote the expression for v(x, t) defined in (5.240) by p(x, t), so that
1 2
p(x, t) = √ e−x /4Dt , for x ∈ R and t > 0. (5.241)
4πDt
It then follows from what we have shown thus far that the function p defined in
(5.241) is a C ∞ function defined in R × (0, ∞) that solves the one–dimensional
diffusion equation in (5.235); that is,
∂p ∂2p
= D 2, for x ∈ R and t > 0. (5.242)
∂t ∂x
Also, it follows from (5.238) and (5.239) that
Z ∞
p(x, t) dx = 1, for all t > 0. (5.243)
−∞
In addition to (5.244), the function p defined in (5.241) has the following prop-
erties:
Proposition 5.2.1 (Properties of p). Let p(x, t) be as defined in (5.241) for
x ∈ R and t > 0.
(i) p(x − y, t) > 0 for all x, y ∈ R and t > 0
(ii) If x 6= y, then lim p(x − y, t) = 0.
t→0+
∂u ∂2u
= D 2, x ∈ R, t > 0;
∂t ∂x (5.245)
u(x, 0) = f (x), x ∈ R,
exist and
lim f (y) 6= lim− f (y).
y→x+ y→x
Z ∞
u(x, t) = p(x − y, t)f (y) dy, for x ∈ R and t > 0, (5.246)
−∞
We will see in this section that (5.247) holds true foe values of x at which f is
continuous. For values of x at which f has a jump discontinuity
f (x+ ) + f (x− )
lim u(x, t) = ,
t→0+ 2
respectively.
We state the main result of this section as the following proposition:
5.2. FUNDAMENTAL SOLUTIONS 123
∂u ∂2u
(x, t) = D 2 (x, t), for x ∈ R and t > 0. (5.247)
∂t ∂x
Furthermore,
lim u(x, t) = f (x). (5.248)
t→0+
if f is continuous at x, and
f (x+ ) + f (x− )
lim u(x, t) = , (5.249)
t→0+ 2
if f has a jump discontinuity at x.
Once we have proved Proposition 5.2.3, we will have constructed a solution
Z ∞
u(x, t) = p(x − y, t)f (y) dy, for x ∈ R and t > 0, (5.250)
−∞
to the initial value problem of the initial value for the one–dimensional diffusion
for the case of continuous initial data f , where p is defined in (5.241). Thus,
a solution of the initial value problem in (5.245) is obtained by integrating
f (y)p(x − y, t) over y in the entire real line. The map
or
1 2
(x, y, t) 7→ √e−(x−y) /4Dt , for all x, y ∈ R and t > 0,
4πDt
is usually called the heat kernel; we shall also call it the fundamental solu-
tion to the one–dimensional diffusion equation. We will denote it by K(x, y, t),
so that K : R2 × (0, ∞) → R and
1 2
K(x, y, t) = √ e−(x−y) /4Dt , for all x, y ∈ R and t > 0. (5.251)
4πDt
We shall reiterate the properties of the heat kernel that we have discussed for
future reference in the following proposition, we will add the additional obser-
vation that K is symmetric in x and y; that is K(x, y, t) = K(y, x, t) for all
x, y ∈ R and t > 0.
Proposition 5.2.4 (Properties of the Heat Kernel). Let K(x, y, t) be as defined
in (5.251) for x, y ∈ R and t > 0.
(i) K(x, y, t) = K(y, x, t) for all x, y ∈ R and t > 0.
(ii) K(x, y, t) > 0 for all x, y ∈ R and t > 0.
1 The function u is C 2 in the first variable, and C 1 in the second variable
124 CHAPTER 5. SOLVING LINEAR PDES
Z ∞
(iii) K(y, x, t) dy = 1 for all x ∈ R and t > 0.
−∞
Before we prove Proposition 5.2.3, we will establish two Lemmas; the first
one involves the error function,
Erf : R → R,
defined by Z x
2 2
Erf(x) = √ e−r dr, for x ∈ R, (5.252)
π 0
and its properties:
Proposition 5.2.5. Let Erf : R → R be as given in (5.252). Then,
(i) Erf(0) = 0;
(ii) lim Erf(x) = 1;
x→∞
Lemma 5.2.6. Let p(x, t) be as defined in (5.241) for x ∈ R and t > 0. For
δ > 0, Z ∞
lim p(x, t) dx = 0. (5.253)
t→0+ δ
5.2. FUNDAMENTAL SOLUTIONS 125
and Z −δ
lim+ p(x, t) dx = 0. (5.254)
t→0 −∞
x
Proof: Make the change of variables y = √ to write
4Dt
Z ∞ Z ∞
1 2
p(x, t) dx = √ · e−x /4Dt dx
δ δ 4πDt
Z ∞
1 2
= √ √ e−y dy
π δ/ 4Dt
1 δ
= 1 − Erf √ ,
2 4Dt
where we have used the definition of the error function in (5.252) and the fact
that Z ∞ √
2 π
e−y dy = .
0 2
We then have that
Z ∞
1 δ
p(x, t) dx = 1 − Erf √ , for t > 0. (5.255)
δ 2 4Dt
Lemma 5.2.7. Let p(x, t) be as defined in (5.241) for x ∈ R and t > 0. Then, we
have the following estimates on integrals of the absolute values of the derivatives
of p:
Z ∞
∂p 1
(x − y, t) dy 6 , for all x ∈ R and t > 0, (5.256)
−∞ ∂t t
and
Z ∞
∂p 1
(x − y, t) dy = √ , for all x ∈ R and t > 0, (5.257)
−∞ ∂x πDt
1 2
p(x − y, t) = √ e−(x−y) /4Dt , for all x, y ∈ R and t > 0, (5.258)
4πDt
126 CHAPTER 5. SOLVING LINEAR PDES
∂ 1 (x − y)2
[p(x − y, t)] = − p(x − y, t) + p(x − y, t), (5.259)
∂t 2t 4Dt2
for all x, y ∈ R and t > 0. Next, take absolute value on both sides of (5.259),
apply the triangle inequality, and use the positivity of the heat kernel (see (ii)
in Proposition 5.2.4) to get
∂ 1 (x − y)2
[p(x − y, t)] 6 p(x − y, t) + p(x − y, t), (5.260)
∂t 2t 4Dt2
for all x, y ∈ R and t > 0. Integrating on both sides of (5.260) and using (5.244)
(see (iii) in Proposition 5.2.4) yields
Z ∞ Z ∞
∂ 1 (x − y)2
[p(x − y, t)] dy 6 + 2
p(x − y, t) dy, (5.261)
−∞ ∂t 2t −∞ 4Dt
for all x ∈ R and t > 0. The right–most integral in (5.262) can be evaluated
using integration by parts to yield
Z ∞ Z ∞
2 −ξ 2 2
ξ e dξ = 2 ξ 2 e−ξ dξ
−∞ 0
∞
Z ∞
−ξ 2 2
= −ξe + e−ξ dξ,
0 0
so that Z ∞ √
2 π
ξ 2 e−ξ dξ = . (5.263)
−∞ 2
Combining (5.263), (5.262) and (5.261) yields the estimate
Z ∞
∂ 1
[p(x − y, t)] dy 6 , for x ∈ R and t > 0,
−∞ ∂t t
5.2. FUNDAMENTAL SOLUTIONS 127
which is (5.256).
In order to establish (5.257), first take the partial derivative with respect to
x on both side of (5.258) to get
∂ x−y
[p(x − y, t)] = − p(x − y, t), for x ∈ R and t > 0, (5.264)
∂x 2Dt
so that, taking absolute value on both sides of (5.264) and integrating,
Z ∞ Z ∞
∂ |x − y|
[p(x − y, t)] dy = p(x − y, t) dy, (5.265)
−∞ ∂x −∞ 2Dt
so that 2
∞
|x − y| e−(x−y) /4Dt
Z
1
· √ dy = √ . (5.267)
−∞ 2Dt 4πDt πDt
The statement in (5.257) now follows by putting together the results in (5.267),
(5.266) and (5.265).
where p(x − y, t) denotes the heat kernel given in (5.258). We will show that
u solves the one–dimensional diffusion equation in (5.247). Before we do that,
though, we need to verify that the expression in (5.269) does indeed define a
128 CHAPTER 5. SOLVING LINEAR PDES
In order to derive the estimate in (5.270), use the positivity of the heat kernel
(see (ii) in Proposition 5.2.4), (5.244) and (5.268) to compute
Z ∞ Z ∞
|p(x − y, t)f (y)| dy 6 M p(x − y, t) dy,
−∞ −∞
so that
Z ∞
|p(x − y, t)f (y)| dy 6 M, for x ∈ R and t > 0, (5.271)
−∞
which implies (5.270). Observe that the estimate in (5.271) also implies that
by virtue of (5.269).
The fact that u defined in (5.269) solves the one–dimensional diffusion equa-
tion in (5.247) will follow from the fact that the heat kernel itself solves the
one–dimensional heat equation,
∂ ∂2
[p(x − y, t)] = D 2 [p(x − y, t)], for x, y ∈ R and t > 0; (5.272)
∂t ∂x
(see also (5.242). Indeed, suppose for the moment that we can interchange
differentiation and integration in the definition of u in (5.269), so that
Z ∞
∂u ∂p
(x, t) = (x − y, t)f (y) dy, for x ∈ R and t > 0, (5.273)
∂t −∞ ∂t
and
∞
∂2u ∂2p
Z
(x, t) = (x − y, t)f (y) dy, for x ∈ R and t > 0. (5.274)
∂x2 −∞ ∂x2
Thus, combining (5.273) and (5.274),
Z ∞
∂2u ∂2p
∂u ∂p
(x, t) − D 2 (x, t) = (x − y, t) − D 2 (x − y, t) f (y) dy,
∂t ∂x −∞ ∂t ∂x
which shows that (5.247) holds true by virtue of (5.272)
The expressions in (5.273) and (5.274) are justified by the assumption that
f is bounded (see (5.268) and the estimates (5.256) and (5.257) in Lemma 5.2.7;
namely, Z ∞
∂p 1
(x − y, t) dy 6 , for all x ∈ R and t > 0.
−∞ ∂t t
5.2. FUNDAMENTAL SOLUTIONS 129
and Z ∞
∂p 1
(x − y, t) dy = √ , for all x ∈ R and t > 0.
−∞ ∂x πDt
Observe that, (5.274) and (5.272) imply the estimate
Z ∞ 2
∂ p 1
2
(x − y, t) dy 6 , for all x ∈ R and t > 0.
−∞ ∂x Dt
where we have used the definition of u(x, t) in (5.269) and (5.244) (see also the
fact (iii) in Proposition 5.2.4). We then have that
Z ∞
u(x, t) − f (x) = p(x − y, t)(f (y) − f (x)) dy,
−∞
so that Z ∞
|u(x, t) − f (x)| 6 p(x − y, t)|f (y) − f (x)| dy, (5.276)
−∞
where we have used the fact that p(x, t) is positive for all x ∈ R and all t > 0.
Next, re-write the integral on the right–hand side of (5.276) as a sum of
three integrals,
Z ∞
p(x − y, t)|f (y) − f (x)| dy =
−∞
Z x−δ
p(x − y, t)|f (y) − f (x)| dy
−∞
(5.277)
Z x+δ
+ p(x − y, t)|f (y) − f (x)| dy
x−δ
Z ∞
+ p(x − y, t)|f (y) − f (x)| dy.
x+δ
130 CHAPTER 5. SOLVING LINEAR PDES
We first estimate the middle integral on the right–hand side of (5.277), using
(5.275) and (5.244) to get
Z x+δ
ε
p(x − y, t)|f (y) − f (x)| dy < . (5.278)
x−δ 3
Next, use (5.268) and the triangle inequality to obtain the following estimate
for the last integral on the right–hand side of (5.277),
Z ∞ Z ∞
p(x − y, t)|f (y) − f (x)| dy 6 2M p(x − y, t) dy. (5.279)
x+δ x+δ
where we have also used the symmetry of the heat kernel (see (i) in Proposition
5.2.4). It follows from (5.280) and (5.253) in Lemma 5.2.6 that
Z ∞
lim+ p(x − y, t)|f (y) − f (x)| dy = 0;
t→0 x+δ
which yields (5.248) and assertion (i) of Proposition 5.2.3 has been proved.
(ii) Assume that f has a jump discontinuity at x and put
Let ε > 0 be given. It follows from (5.284) that there exists δ > 0 such that
ε
x < y < x + δ ⇒ |f (y) − f (x+ )| < , (5.285)
3
and
ε
x − δ < y < x ⇒ |f (y) − f (x− )| < . (5.286)
3
Use the definition of u(x, t) in (5.269) to write
∞
f (x+ ) + f (x− )
Z
1 1
u(x, t) − = p(x − y, t)f (y) dy − f (x+ ) − f (x− ),
2 −∞ 2 2
by virtue of (5.244), (5.243) and the symmetry of the heat kernel (see (i) in
Proposition 5.2.4). We therefore have that
f (x+ ) + f (x− )
u(x, t) −
2
Z x
= p(x − y, t)(f (y) − f (x− )) dy
−∞
Z ∞
+ p(x − y, t)(f (y) − f (x+ )) dy,
xo
so that
f (x+ ) + f (x− )
u(x, t) −
2
Z x
6 p(x − y, t)|f (y) − f (x− )| dy (5.288)
−∞
Z ∞
+ p(x − y, t)|f (y) − f (x+ )| dy,
x
We re-write the last integral on the right–hand side of (5.288) as a sum of two
132 CHAPTER 5. SOLVING LINEAR PDES
integrals,
Z ∞
p(x − y, t)|f (y) − f (x+ )| dy
x
Z x+δ
= p(x − y, t)|f (y) − f (x+ )| dy (5.289)
x
Z ∞
+ p(x − y, t)|f (y) − f (x+ )| dy,
x+δ
where
Z x+δ Z x+δ
+ ε ε
p(x − y, t)|f (y) − f (x )| dy < p(x − y, t) dy < , (5.290)
x 3 x 6
Thus, letting δ3 = min{δ1 , δ2 } we see that the conjunction of (5.292) and (5.293),
together with (5.288), implies that
f (x+ ) + f (x− )
0 < t < δ3 ⇒ u(x, t) − < ε.
2
We have therefore established (5.249) and the proof of part (ii) of Proposition
5.2.3 is now complete.
Example 5.2.8. Solve the initial value problem for the diffusion equation in
(5.245), where (
1, if − 1 < x 6 1;
f (x) = (5.294)
0, elsewhere.
5.2. FUNDAMENTAL SOLUTIONS 133
or
Z 1
1 2
u(x, t) = √ e−(x−y) /4Dt
dy, for x ∈ R and t > 0, (5.295)
4πDt −1
x−y
Make the change variables r = √ in (5.295) to obtain
4Dt
Z x−1
√
1 4Dt 2
u(x, t) = − √ e−r dr, for x ∈ R and t > 0,
π x+1
√
4Dt
or x+1 x−1
Z √ Z √
1 4Dt 2 1 4Dt 2
u(x, t) = √ e−r dr − √ e−r dr, (5.296)
π 0 π 0
for x ∈ R and t > 0.
Making use of the error function defined in (5.252), we can rewrite (5.296)
as
1 x+1 x−1
u(x, t) = Erf √ − Erf √ , (5.297)
2 4Dt 4Dt
for x ∈ R and t > 0. Figure 5.2.4 shows plots of the graph of y = u(x, t),
where u(x, t) is as given in (5.297), for various values of t in the case 4D = 1.
A few interesting properties of the function u given in (5.297) are apparent by
examining the pictures in Figure 5.2.4. First, the graph of y = u(x, t) is smooth
for all t > 0. Even though the initial temperature distribution, f , in (5.294)
is not even continuous, the solution to the initial value problem (5.245) given
in (5.297) is in fact infinitely differentiable as soon as the process gets going
for t > 0. Secondly, the values, u(x, t), of the function u given in (5.297)) are
positive at all values of x ∈ R and t > 0. In particular, for values of x with
|x| > 1, where the initial temperature is zero, the temperature rises instantly
134 CHAPTER 5. SOLVING LINEAR PDES
1.0
0.8
0.6
x+1 x-1
0.5 erf - erf
-3 -2 -1 1 2 3 10 10
Computed by WolframÈAlpha
for t > 0. Thus, the diffusion model for heat propagation predicts that heat
propagates with infinite speed. Thirdly, we see from the pictures in Figure 5.2.4
that
lim u(x, t) = 0, for all x ∈ R. (5.298)
t→∞
so that the integrability condition in (5.299) holds true for the function in
(5.294).
Before we establish that (5.298) is true for any solution of the initial value
problem (5.245) in which the initial condition satisfies (5.299), we will first need
to derive other properties of the function u given in (5.246).
Proposition 5.2.9. Let f : R → R be continuous and satisfying (5.299); that
is, Z ∞
|f (x)|2 dx < ∞.
−∞
Put
Z ∞
u(x, t) = p(x − y, t)f (y) dy, for x ∈ R and t > 0. (5.300)
−∞
5.2. FUNDAMENTAL SOLUTIONS 135
Then, Z ∞
|u(x, t)|2 dx < ∞, for all t > 0, (5.301)
−∞
and Z ∞ 2
∂u
(x, t) dx < ∞, for all t > 0. (5.302)
−∞ ∂x
where we have also used (5.244) and the positivity of the heat kernel (see (ii)
and (iii) in Proposition 5.2.4).
Integrate with respect to x on both sides of (5.303) to get
Z ∞ Z ∞Z ∞
|u(x, t)|2 dx 6 p(x − y, t)|f (y)|2 dydx, (5.304)
−∞ −∞ −∞
for t > 0. Interchanging the order of integration in the integral on the right–
hand side of (5.304) we obtain
Z ∞ Z ∞ Z ∞
|u(x, t)|2 dx 6 |f (y)|2 p(x − y, t) dx dy, (5.305)
−∞ −∞ −∞
so that ∞
(x − y)
Z
∂u
(x, t) = − p(x − y, t) f (y) dy, (5.308)
∂x −∞ 2Dt
for x ∈ R and t > 0.
136 CHAPTER 5. SOLVING LINEAR PDES
Proceeding as in the first part of this proof, use the Cauchy–Schwarz in-
equality (or Jensen’s inequality) to obtain from (5.308) that
2 ∞
(x − y)2
Z
∂u
(x, t) 6 p(x − y, t) |f (y)|2 dy, (5.309)
∂x −∞ 4D2 t2
for t > 0.
Observe that the inner integral in the right–hand side of (5.310) is simply
the variance, 2Dt, of the probability density function p(x, t), so that
Z ∞
(x − y)2 p(x − y, t) dx = 2Dt, for all y ∈ R and t > 0. (5.311)
−∞
∂2u
∂u
= D 2, for x ∈ R, t > 0;
∂t ∂x
u(x, 0) = f (x), for x ∈ R;
Z ∞ (5.312)
|u(x, t)|2 dx < ∞, for all t > 0;
−∞
Z ∞ 2
∂u
(x, t) dx < ∞, for all t > 0.
−∞ ∂x
5.2. FUNDAMENTAL SOLUTIONS 137
Then, Z ∞ Z ∞
|v(x, t)|2 dx 6 |f (x)|2 dx, for t > 0. (5.313)
−∞ −∞
Proof: Let v denote any solution to the problem (5.312), where f satisfies the
integrability condition in (5.299).
In order to establish (5.313), set
Z ∞
E(t) = |v(x, t)|2 dx, for all t > 0. (5.314)
−∞
It follows from the integrability condition in (5.312) that E(t) in (5.314) is well
defined for all t > 0 as a real valued function, E : [0, ∞) → R. Note also that
Z ∞
E(0) = |f (x)|2 dx, (5.315)
−∞
for t > 0.
We note that the integrability conditions in (5.312) imply that
lim v(x, t) = 0 and lim v(x, t) = 0, for t > 0, (5.317)
x→∞ x→−∞
and
lim vx (x, t) = 0 and lim vx (x, t) = 0. for t > 0, (5.318)
x→∞ x→−∞
so that Z ∞ 2
∂v
E 0 (t) = − (x, t) dx, for t > 0, (5.319)
−∞ ∂x
by virtue of (5.317), (5.318) and the last integrability condition in (5.312).
Now, it follows from (5.319) that
E 0 (t) 6 0, for all t > 0,
so that E is nondecreasing in t and therefore
E(t) 6 E(0), for all t > 0. (5.320)
The estimate in (5.313) follows from (5.320) in view of (5.314) and (5.315).
138 CHAPTER 5. SOLVING LINEAR PDES
It follows from the linearity of the differential equation in (5.321) that w also
solves the diffusion equation; indeed,
so that
|w(x, t)|2 6 |v(x, t)|2 + 2|v(x, t)| · |u(x, t)| + |u(x, t)|2 , (5.323)
2ab 6 a2 + b2 , for a, b ∈ R,
in (5.323) to get
so that Z ∞
|w(x, t)|2 dx < ∞, for t > 0,
−∞
5.2. FUNDAMENTAL SOLUTIONS 139
since both u and v satisfy the integrability conditions in problem (5.321). Sim-
ilarly, we can show that
Z ∞
|wx (x, t)|2 dx < ∞, for t > 0.
−∞
Now, observe that, since both v and u satisfy the initial condition in problem
(5.321,
so that
v(x, t) = u(x, t), for all x ∈ R and t > 0,
in view of the definition of w in (5.322). Hence, any solution to the problem in
(5.321) must be that given by (5.300).
then
lim u(x, t) = 0, for all x ∈ R. (5.327)
t→∞
for all x ∈ R and t > 0. Next, square on both sides of (5.328) and apply the
Cauchy–Schwarz inequality to get
Z ∞ −(x−y)2 /2Dt Z ∞
2 1 e
|u(x, t)| 6 √ √ dy |f (y)|2 dy, (5.329)
8πDt −∞ 2πDt −∞
where 2
∞
e−(x−y) /2Dt
Z
√ dy = 1. (5.330)
−∞ 2πDt
Combining (5.329) and (5.330), we then get
Z ∞
1
|u(x, t)|2 6 √ |f (y)|2 dy, (5.331)
8πDt −∞
for x ∈ R and t > 0.
It follows from (5.326) and (5.331) that
• First, in view of the radial symmetry of the domain, we will express prob-
lem (5.3) in polar coordinate r and θ.
5.3. DIRICHLET PROBLEM FOR THE UNIT DISC 141
∂D1
D1
x
• Next, we look for a special type of solutions that are products of a function
of r and a function of θ. In other words, we look for solutions in which
the variables separate; this is where we use the method of separation
of variables.
• When looking for solutions that are nonzero over the domain by means of
separation of variables, we are invariable led to an eigenvalue problem.
Solution of the eigenvalue problem leads to a family of solutions in one
(or both of the variables), called eigenfunctions. These eigenfunctions
generate a special family of solutions.
We will also exploit the linearity of the PDE and the boundary condition in
(5.332) and use the principle of superposition to construct a solution of the
problem by superposing simple solutions of the problem. The strategy then is
to, first, find a special class of functions of r and θ that solve Laplace’s equation,
and then use sums of those solutions to construct a solution that also satisfies
the boundary condition.
142 CHAPTER 5. SOLVING LINEAR PDES
We start out by looking for special solutions of the PDE in (5.334) of the
form
v(r, θ) = h(r)z(θ), for r > 0 and − π < θ < π, (5.335)
where h : [0, ∞) → R is a continuous functions that is C 2 in (0, ∞), and z : R →
R is a C 2 , periodic function of period 2π. We can therefore compute the partial
derivatives,
∂v
(r, θ) = h0 (r)z(θ), r > 0, −π < θ 6 π;
∂r
∂2v
(r, θ) = h00 (r)z(θ), r > 0, −π < θ 6 π;
∂r2
∂2v
(r, θ) = h(r)z 00 (θ), r > 0, −π < θ 6 π,
∂θ2
and substitute them into the PDE in (5.334) to obtain
1 1
h00 (r)z(θ) + h0 (r)z(θ) + 2 h(r)z 00 (θ) = 0, for r > 0, −π < θ 6 π. (5.336)
r r
Assuming that v(r, θ) is not zero for all values of r and θ, and dividing on both
sides of (5.336) by v(r, θ) as given in (5.335), we obtain
Since (5.338) holds true for all values of r and θ in (0, 1) and (−π, π], respectively,
it follows from (5.338) that each side of the equation in (5.338) must be equal
to a constant.2 Call that constant λ so that
h00 (r) h0 (r) z 00 (θ)
r2 +r =− = λ, for 0 < r < 1, −π < θ 6 π. (5.339)
f (r) h(r) z(θ)
and
r2 h00 (r) + rh0 (r) = λh(r), for r > 0. (5.341)
The requirement that the function g in (5.334) be periodic of period 2π yields
the following conditions for z:
−z 00 (θ) = λz(θ),
for − π < θ < π;
z(−π) = z(π); (5.343)
0
z (−π) = z 0 (π).
Substituting the result in (5.345) into the left–hand side of (5.344) then yields
Z π Z π
[z 0 (θ)]2 dθ = λ [z(θ)]2 dθ. (5.346)
−π −π
Since
Z π z is a nontrivial solution of the two–point BVP in (5.343), it follows that
[z(θ)]2 dθ > 0. We can therefore solve (5.346) for λ to obtain
−π
Z π
[z 0 (θ)]2 dθ
λ = Z−π
π ,
[z(θ)]2 dθ
−π
z 00 (θ) = 0,
5.3. DIRICHLET PROBLEM FOR THE UNIT DISC 145
−πc1 + c2 = πc2 + c2 ,
from which we get that 2πc1 = 0, so that c1 = 0. It then follows from (5.347)
any solution of the BVP in (5.343) with λ = 0 must be constant:
λo = 0, (5.349)
and note that any solution of the BVP in (5.343) for λo is a constant multiple
of ϕo given in (5.350); so that
where ao denotes a real constant, represents all solutions of the two–point BVP
in (5.350) corresponding to the eigenvalue λo = 0.
Next, we look for positive eigenvalues of the BVP in (5.343). For the case
in which λ > 0 in (5.343), the general solution of the ODE in (5.343) is
√ √
z(θ) = c1 cos( λθ) + c2 sin( λθ), for all θ, (5.352)
Imposing the the boundary conditions in (5.343) to the functions given in (5.352)
and (5.353) yields the system of equations
√ √ √ √
√ c1 cos(−√ λπ) +√c2 sin(−√λπ) = c1 cos( √ λπ)√+ c2 sin( √λπ); √
−c1 λ sin(− λπ) + c2 λ cos(− λπ) = −c1 λ sin( λπ) + c2 λ cos( λπ);
√ (5.354)
thus, dividing the second equation in (5.354) by λ, since λ > 0, and using the
fact that cos is even and sin is odd,
√ √ √ √
c1 cos( √λπ) − c2 sin(−√λπ) = c1 cos( √ λπ) + c2 sin( √λπ);
c1 sin( λπ) + c2 cos( λπ) = −c1 sin( λπ) + c2 cos( λπ),
146 CHAPTER 5. SOLVING LINEAR PDES
With the values for λ given in (5.361), we now proceed to solve the ODE in
(5.341) for the radial component of the special solutions of the BVP in (5.334)
of the form given in (5.335); namely,
We shall first solve (5.362) for the case n = 0. In this case the equation becomes
where we have divided by r > 0. Observe that the equation in (5.363) can be
written as
d
[rh0 (r)] = 0, for r > 0,
dr
which can be integrated to yield
Hence, it follows from (5.365) and (5.366) that, for n = 0, a solution of (5.362)
is given by
h(r) = c, for all r, (5.367)
is a solution, for some constant c. Taking c = 1 in (5.367) we get the solution
of (5.362) corresponding to n = 0:
Next, consider the case n > 1 in (5.362). In this case the differential equation
in (5.362) is an ODE of Euler type:
The ODE in (5.369) can be solved by looking for solutions of the form
or
q(q − 1)rq + qrq − n2 rq = 0, for r > 0,
or
[q(q − 1) + q − n2 ]rq = 0, for r > 0. (5.371)
It follows from (5.371) that
q(q − 1) + q − n2 = 0,
or
q 2 − n2 = 0,
or
(q + n)(q − n) = 0,
from which we get that
q = ±n. (5.372)
It follows from (5.371) and (5.372) that
Let’s assume for the moment that the series in (5.377) converges also for
r = 1, so that we can apply the boundary condition in (5.334) to get
∞
X
ao + [an cos(nθ) + bn sin(nθ)] = f (θ), for − π < θ 6 π. (5.378)
n=0
Assuming for the moment that the series on the left–hand side of (5.378) con-
verges in such a way that it can be integrated term by term, we can compute
the values of the coefficients an , for n = 0, 1, 2, . . ., and bn , for n = 1, 2, 3, . . ., in
terms of the function f by means of the following integration facts:
Z π
sin(nθ) cos(mθ) dθ = 0, for all m, n = 1, 2, 3, . . . ; (5.379)
−π
(
π
if m 6= n;
Z
0,
cos(nθ) cos(mθ) dθ = (5.380)
−π π, if m = n;
and (
π
if m 6= n;
Z
0,
sin(nθ) sin(mθ) dθ = (5.381)
−π π, if m = n.
Similar calculations (this time multiplying the equation in (5.378) on both sides
by sin(mθ), integrating from −π to π, and using the integral identities in (5.379)
and (5.381)) lead to
1 π
Z
bm = f (θ) sin(mθ) dθ, for m = 1, 2, 3, . . . . (5.385)
π −π
The numbers defined in (5.382), (5.384) and (5.385) are the Fourier coef-
ficients of the 2π–periodic function f . Note that the Fourier coefficients of f
are defined whenever f is absolutely integrable over the interval [−π, π].
Definition 5.3.4 (Absolute Integrability). A function f : [−π, π] → R is said
to be absolutely integrable over [−π, π] whenever
Z π
|f (θ)| dθ < ∞. (5.386)
−π
Note that f doesn’t have to be continuous for (5.386) to hold true. For
instance, if f is bounded and piecewise continuous then (5.386) holds true;
indeed, suppose that f piecewise continuous and
If the integral in (5.387) ia understood as the Lebesgue integral, and kf kL1 < ∞,
we will say that f is an L1 function and write f ∈ L1 (−π, π). We shall refer to
kf kL1 as the L1 norm of f ∈ L1 (−π, π).
The existence of the Fourier coefficients of f in (5.382), (5.384) and (5.385)
is guaranteed for absolutely integrable 2π–periodic functions, f , or for f ∈
L1 (−π, π). This is the content of the following proposition.
Proposition 5.3.6 (Existence of the Fourier Coefficients). Let an , for n =
0, 1, 2, . . ., be as given in (5.382) and (5.384), and bn , for n = 1, 2, 3, . . ., be as
in (5.385), where f ∈ L1 (−π, π). Then,
1
|an | 6 kf kL1 , for n = 0, 1, 2, 3, . . . ; (5.388)
π
and
1
|bn | 6 kf kL1 , for n = 1, 2, 3, . . . . (5.389)
π
5.3. DIRICHLET PROBLEM FOR THE UNIT DISC 151
Proof: The estimates in (5.388) and (5.389) follow from properties of the inte-
gral. For ao , we get from (5.382) that
Z π
1
|ao | 6 |f (θ)| dθ,
2π −π
1 1
|ao | 6 kf kL1 6 kf kL1 .
2π π
For n = 1, 2, 3, . . . we obtain from (5.384) that
1 π
Z
|an | 6 |f (θ)| | cos(nθ)| dθ
π −π
Z π
1
6 |f (θ)| dθ,
π −π
since | cos(nθ)| 6 1 for all θ and all n, which yields (5.388). Similar calculations
lead to (5.389).
Proof: The conclusion will follow by comparing with the geometric series since
0 6 r < 1 and
kf kL1 n
|rn an cos(nθ)| 6 rn |an | 6 r , for all n,
π
152 CHAPTER 5. SOLVING LINEAR PDES
where we have used the estimate (5.388) in Proposition 5.3.6. Similarly, using
(5.389) in Proposition 5.3.6,
kf kL1 n
|rn bn sin(nθ)| 6 r ,
π
for all n.
∞
X
− n2 rn [an cos(nθ) + bn sin(nθ)],
n=0
∞
X
nrn−1 [an cos(nθ) + bn sin(nθ)],
n=0
and
∞
X
n(n − 1)rn−2 [an cos(nθ) + bn sin(nθ)],
n=0
or
∞ ∞ Z
!
Z b X X b
uk (x) dx = uk (x) dx,
a k=1 k=1 a
Proof: The assumptions that f is piecewise C 1 and 2π–periodic imply that the
Fourier coefficients of f satisfy the estimate
∞
X
(|ak | + |bk |) < ∞. (5.393)
k=0
and all r ∈ [0, 1] and θ ∈ [−π, π]. Thus, the absolute values of the terms of
the series in (5.392) are “majorized” by the terms of the convergent series in
(5.393). It then follows by the the Weierstrass M–Test for uniform convergence
([Rud53, Theorem 7.10, pg. 119]) that the series in (5.392) converges uniformly
for r ∈ [0, 1] and θ ∈ [−π, π].
We will get a chance to use the Weierstrass M–Test for uniform convergence
once again to justify the following calculations based on the trigonometric series
representation for v(r, θ) in (5.392) and the assumption that f is a piecewise
C 1 , 2π–periodic function.
First, substitute the formulas defining the Fourier coefficients of f in (5.382),
5.3. DIRICHLET PROBLEM FOR THE UNIT DISC 155
Z π
1
+ f (ξ) sin(nξ) dξ sin(nθ)
π −π
Z π ∞ Z π
1 1X n
= f (ξ) dξ + r cos(nθ) cos(nξ)f (ξ) dξ
2π −π π n=0 −π
Z π
+ sin(nθ) sin(nξ)f (ξ) dξ
−π
Z π
1
= f (ξ) dξ
2π −π
∞ Z
1X π n
+ r [cos(nθ) cos(nξ) + sin(nθ) sin(nξ)]f (ξ) dξ,
π n=0 −π
Thus, the assertion follows by the Weierstrass M–Test for uniform convergence,
for 0 6 r < 1.
Hence, interchanging the order of summation and integration in (5.394), we
can write
Z π " ∞
#
1 X
v(r, θ) = 1+ 2rn cos[n(θ − ξ)] f (ξ) dξ, (5.395)
2π −π n=1
156 CHAPTER 5. SOLVING LINEAR PDES
and for n ∈ N. It follows from (5.398) that the absolute values of the terms of
the series in (5.396) are “majorized” by the terms of the convergent geometric
series
X∞
2rn ,
n=1
for 0 6 r < 1. Hence, the Weierstrass M–Test applies, and we conclude that
the series defining P (r, θ) in (5.396) converges uniformly in θ for 0 6 r < 1.
This arguments can be carried out further to prove that, for any 0 < R < 1,
the series defining P (r, θ) in (5.396) converges uniformly for θ ∈ [−π, π] and
r ∈ [0, R]. Hence, P (r, θ) defines a continuous function in the open unit disc,
D1 , in R2 . This follows from the following important consequence of the uniform
convergence of a sequence of continuous functions:
For a proof of this proposition see Rudin [Rud53, Theorem 7.12, pg. 20].
We will next show that, in fact, the Poisson kernel is C 2 in D1 and that it
solves Laplace’s equation in D1 . In order to show that the partial derivatives of
5.3. DIRICHLET PROBLEM FOR THE UNIT DISC 157
converge uniformly. This assertion will follow from the following proposition
Proposition 5.3.14 (Term–by–Term Differentiation). Let (uk ) be a sequence
of functions that are differentiable over a closed and bounded interval, [a, b].
Assume that the series
X∞
u0k
k=1
converges absolutely and uniformly over [a, b]. Assume also that the series
∞
X
uk (xo )
k=1
converges absolutely at some point xo in [a, b]. Then, the series converges
∞
X
uk
k=1
or "∞ ∞
#
d X X
uk (x) = u0k (x) ,
dx
k=1 k=1
for all θ ∈ [−π, π]; so that the series in (5.399) are “majorized” by the series
∞
X ∞
X
2nrn−1 and 2nrn , (5.400)
n=1 n=1
respectively; both of the series in (5.400) converge by the Ratio Test (or the
Root Test), since 0 6 r < 1. It then follows by the Weierstrass M–Test for
158 CHAPTER 5. SOLVING LINEAR PDES
respectively. It then follows from Proposition 5.3.14 that the partial derivatives
of the Poisson kernel in (5.396) have partial derivatives in D1 given by
∞
∂ 1 X
[P (r, θ)] = 2nrn−1 cos(nθ), 0 6 r < 1 and θ ∈ [−π, π], (5.401)
∂r 2π n=1
and
∞
∂ 1 X
[P (r, θ)] = − 2nrn sin(nθ), 0 6 r < 1 and θ ∈ [−π, π], (5.402)
∂θ 2π n=1
and
∞
∂2 1 X 2 n
[P (r, θ)] = − 2n r cos(nθ), 0 6 r < 1 and θ ∈ [−π, π], (5.404)
∂θ2 2π n=1
where the series in (5.401) and (5.402) have been differentiated term–by–term.
Next, substitute the partial derivatives in (5.401), (5.403) and (5.404) into
the expression for the Laplacian of P in polar coordinates to get
∞
∂2P 1 ∂P 1 ∂2P 1 X
+ + = 2n(n − 1)rn−2 cos(nθ)
∂r2 r ∂r r2 ∂θ2 2π n=1
∞
1 X
+ 2nrn−1 cos(nθ)
2πr n=1
∞
1 X 2 n
− 2n r cos(nθ) (5.405)
2πr2 n=1
∞
1X
= [n(n − 1) + n − n2 ]rn−2 cos(nθ)
π n=1
= 0,
5.3. DIRICHLET PROBLEM FOR THE UNIT DISC 159
for all θ ∈ [−π, π] and 0 6 r < 1. We have therefore shown that the Poisson
kernel solves Laplace’s equation in the open unit disc.
Next, integrate the series in (5.396) over the interval [−π, π], which is justi-
fied by the uniform convergence of the series, to obtain
Z π
P (r, θ) dθ = 1, for all 0 6 r < 1. (5.406)
−π
The series defining the Poisson kernel in (5.396) can actually be evaluated
by using the identity
2 cos(nθ) = einθ + e−inθ ,
and then adding geometric series. Indeed,
∞
X ∞
X
2rn cos(nθ) = rn [einθ + e−inθ ]
n=1 n=1
∞
X ∞
X
= rn [eiθ ]n + rn [e−iθ ]n
n=1 n=1
∞
X ∞
X
= [reiθ ]n + [re−iθ ]n ,
n=1 n=1
which simplifies to
∞
X reiθ − r2 + re−iθ − r2
2rn cos(nθ) = ,
n=1
1 − reiθ − re−iθ + r2
or
∞
X r[eiθ + e−iθ ] − 2r2
2rn cos(nθ) = ,
n=1
1 − r[eiθ + e−iθ ] + r2
or
∞
X 2r cos(θ) − 2r2
2rn cos(nθ) = , 0 6 r < 1, θ ∈ [−π, π]. (5.407)
n=1
1 − 2r cos(θ) + r2
Substituting the value of the series in (5.407) into (5.396) then yields the formula
1 1 − r2
P (r, θ) = , for 0 6 r < 1 and θ ∈ [−π, π], (5.408)
2π 1 − 2r cos(θ) + r2
for the Poisson kernel.
160 CHAPTER 5. SOLVING LINEAR PDES
We will next use the formula in (5.408) for the Poisson kernel for the unit disk
to derive further properties of the Poisson kernel. We summarize these proper-
ties, as well as the ones we have already established using the representation in
(5.396), in the following proposition.
(iii) P is harmonic in D1 ;
Z π
(iv) P (r, θ − ξ) dξ = 1, for all ξ ∈ R and all 0 6 r < 1.
−π
Proof: In order to prove (i) and (ii), first note that, for all θ ∈ R and r > 0,
2r cos θ 6 2r,
so that
1 − 2r cos θ + r2 > 1 − 2r + r2 ,
or
1 − 2r cos(θ) + r2 > (1 − r)2 , for 0 6 r < 1 and θ ∈ R. (5.409)
It follows from (5.409) and the formula for P (r, θ) in (5.408) that P (r, θ) is
defined for all r ∈ [0, r) and all θ ∈ R, and P (r, θ) > 0 for r ∈ [0, 1) and all
θ ∈ R; we have therefore establlished (i).
From (5.409) we also obtain that
Thus, the denominator in the formula for P (r, θ) in (5.408) is not zero for
0 6 r < 1 and θ ∈ R; hence, since the numerator and denominator of the
expression defining P (r, θ) in (5.408) are C ∞ functions, (ii) also follows.
We have already established that P satisfies Laplace’s equation in D1 (see
the calculations leading up to (5.405) on page 158) using the definition of P in
(5.396). Thus, P is harmonic in D1 and so we have established (iii).
The integral identity in (iv) will follow from (5.406) and the 2π–periodicity
of P (r, θ) in θ. Indeed, making the change of variables ζ = θ − ξ in the integral
5.3. DIRICHLET PROBLEM FOR THE UNIT DISC 161
in (iv) we have
Z π Z θ−π
P (r, θ − ξ) dξ = − P (r, ζ) dζ
−π θ+π
Z θ+π
= P (r, ζ) dζ
θ−π
Z π
= P (r, ζ) dζ
−π
= 1,
for all θ ∈ R.
Next, use the formula for P (r, θ) in (5.408) to obtain that
1 1 − r2
P (r, θ − ξ) = ,
2π 1 − 2r + r2
for θ = ξ, from which we get that
1 1+r
P (r, θ − ξ) = , for 0 6 r < 1 and θ = ξ. (5.410)
2π 1 − r
The assertion in (vi) follows from (5.410).
To prove (v), first note that
= sin2 (θ − ξ)
so that
The assertion in (v) then follows from (5.411) and the expression for the Poisson
kernel in (5.408).
and some positive constant M . The goal of this section is to use the properties
of the Poisson kernel listed in Proposition 5.3.15 to prove that the function
162 CHAPTER 5. SOLVING LINEAR PDES
v : D1 → R defined by
Z π
P (r, θ − ξ)f (ξ) dξ, for 0 6 r < 1, θ ∈ [−π, π];
−π
v(r, θ) = (5.412)
f (θ), for r = 1, θ ∈ [−π, π],
where P (r, θ) denotes the Poisson kernel for the unit disc in R2 given in (5.396)
or (5.408), solves the Dirichlet problem (5.334) for the unit disc in R2 .
We first show that v ∈ C 2 (D1 , R) and that it solves Laplace’s equation in
D1 ; in polar coordinates,
∂ 2 v 1 ∂v 1 ∂2v
+ + = 0, for 0 < r < 1, θ ∈ [−π, π]. (5.413)
∂r2 r ∂r r2 ∂θ2
This will follow from (iii) in Proposition 5.3.15, provided we can show that
differentiation under the integral sign in the first part of the definition of v in
(5.412) is valid. Indeed, property (iii) in Proposition 5.413 says that
∂2P 1 ∂P 1 ∂2P
2
+ + 2 = 0, for 0 6 r < 1, θ ∈ [−π, π]. (5.414)
∂r r ∂r r ∂θ2
Thus, assuming for the moment that differentiation under the integral sign in
(5.412) is valid, we have that, for 0 6 r < 1 and θ ∈ [−π, π],
Z π 2
∂ 2 v 1 ∂v 1 ∂2v ∂
2
+ + 2 2 = 2
[P (r, θ − ξ)]f (ξ) dξ
∂r r ∂r r ∂θ −π ∂r
Z π
1 ∂
+ [P (r, θ − ξ)]f (ξ) dξ
−π r ∂r
π
1 ∂2
Z
+ [P (r, θ − ξ)]f (ξ) dξ,
−π r2 ∂θ2
which can be written as
Z π 2
1 ∂2
∂ 1 ∂
∆v = 2
[P (r, θ − ξ)] + [P (r, θ − ξ)] + 2 2 [P (r, θ − ξ)] f (ξ) dξ,
−π ∂r r ∂r r ∂θ
where we have used the short–hand notation, ∆v, for the Laplacian of v. The
fact that v is harmonic in D1 then follows from the previous identity and (5.414).
We will next see that differentiation under the integral sign is justified. In
order to do this, we first note that the continuity of f implies that there exists
a positive constant, M , such that
|f (θ)| 6 M, for all θ ∈ [−π, π]. (5.415)
In view of (5.415) and (5.414), in order to justify the differentiation under the
integral sign in the first part of the definition of u in (5.412), it suffices to prove
that
∂ ∂ ∂2
[P (r, θ − ξ)], [P (r, θ − ξ)] and [P (r, θ − ξ)]
∂θ ∂r ∂θ2
5.3. DIRICHLET PROBLEM FOR THE UNIT DISC 163
∂ 1 (1 − r2 )2r sin(θ − ξ)
[P (r, θ − ξ)] = − ,
∂θ 2π (1 − 2r cos(θ − ξ) + r2 )2
by virtue of the expression for the Poisson kernel in (5.408). Next, take absolute
values on both sides of (5.416) and use the estimate in (5.409) to get
∂ 2r
[P (r, θ − ξ)] 6 P (r, θ − ξ), (5.417)
∂θ (1 − r)2
where we have used the positivity of the Poisson kernel in (i) of Proposition
5.3.15. Integrating on both sides of the inequality in (5.417) form −π to π and
using property (iv) in Proposition 5.3.15 we obtain that
Z π
∂ 2r
[P (r, θ − ξ)] dξ 6 , for 0 6 r < 1 and θ ∈ [−π, π],
−π ∂θ (1 − r)2
∂
which shows that [P (r, θ − ξ)] is absolutely integrable over [−π, π] for 0 6
∂θ
r < 1 and θ ∈ [−π, π].
Next, take partial derivative with respect to θ on both sides of (5.416) to get
∂2 2r cos(θ − ξ)
[P (r, θ − ξ)] = − P (r, θ − ξ)
∂θ2 1 − 2r cos(θ − ξ) + r2
4r2 sin2 (θ − ξ)
+ P (r, θ − ξ)
(1 − 2r cos(θ − ξ) + r2 )2
2r sin(θ − ξ) ∂
− 2
[P (r, θ − ξ)],
1 − 2r cos(θ − ξ) + r ∂θ
∂2 2r cos(θ − ξ)
[P (r, θ − ξ)] = − P (r, θ − ξ)
∂θ2 1 − 2r cos(θ − ξ) + r2
4r2 sin2 (θ − ξ)
+ P (r, θ − ξ)
(1 − 2r cos(θ − ξ) + r2 )2
4r2 sin2 (θ − ξ)
+ P (r, θ − ξ),
(1 − 2r cos(θ − ξ) + r2 )2
164 CHAPTER 5. SOLVING LINEAR PDES
or
∂2 2r cos(θ − ξ)
[P (r, θ − ξ)] = − P (r, θ − ξ)
∂θ2 1 − 2r cos(θ − ξ) + r2
(5.418)
8r2 sin2 (θ − ξ)
+ P (r, θ − ξ).
(1 − 2r cos(θ − ξ) + r2 )2
Taking absolute values on both sides of (5.418) and applying the triangle in-
equality, we obtain
∂2 2r 8r2
2
[P (r, θ − ξ)] 6 2
P (r, θ − ξ) + P (r, θ − ξ), (5.419)
∂θ (1 − r) (1 − r)4
where we have also used the estimate in (5.409) and the positivity of the Poisson
kernel (see property (i) in Proposition 5.3.15). Integrating from −π to π on both
sides of (5.419) then yields
Z π
∂2 2r 8r2
2
[P (r, θ − ξ)] dξ 6 + , for 0 6 r < 1, θ ∈ R,
−π ∂θ (1 − r)2 (1 − r)4
where we have also used property (iv) in Proposition 5.3.15; thus, we have shown
∂2
that [P (r, θ − ξ)] is absolutely integrable over [−π, π] for 0 6 r < 1 and
∂θ2
θ ∈ R.
Next, differentiate the Poisson kernel in (5.408) with respect to r, for 0 6
r < 1, to obtain
∂ 1 −2r 1 (1 − r2 )(2r − 2 cos(θ − ξ))
[P (r, θ − ξ)] = 2
− ,
∂r 2π 1 − 2r cos(θ − ξ) + r 2π (1 − 2r cos(θ − ξ) + r2 )2
where we have applied the Product Rule; so that, in view of the expression for
the Poison kernel in (5.408),
∂ 2r 2r − 2 cos(θ − ξ)
[P (r, θ − ξ)] = − 2
P (r, θ − ξ) − P (r, θ − ξ),
∂r 1−r 1 − 2r cos(θ − ξ) + r2
or
∂ −2r 2r − cos(θ − ξ)
[P (r, θ − ξ)] = + P (r, θ − ξ), (5.420)
∂r 1 − r2 1 − 2r cos(θ − ξ) + r2
for 0 6 r < 1 and all ξ and θ in R.
Next, take absolute values on both sides of (5.420), applying the triangle
inequality, and use the estimate in (5.409) to obtain
∂ 2r 2r + 1
[P (r, θ − ξ)] 6 + P (r, θ − ξ), (5.421)
∂r 1 − r2 (1 − r)2
for 0 6 r < 1 and θ, ξ ∈ R, where we have also used the positivity of the Poisson
kernel (see property (i) in Proposition 5.3.15).
5.3. DIRICHLET PROBLEM FOR THE UNIT DISC 165
∂
which shows that [P (r, θ − ξ)] is absolutely integrable over [−π, π] for 0 6
∂r
r < 1 and θ ∈ [−π, π].
Hence, differentiation under the integral sign in the first part of the definition
of u in (5.412) is justified. (See the results in Appendix B.1). We have therefore
established that the function v defined in (5.412) is in C 2 (D1 ) and satisfies
Laplace’s equation. It remains to prove that v ∈ C(D1 ) and that it satisfies
the boundary conditions in problem (5.334). This will be accomplished once we
prove the following lemma:
Lemma 5.3.16 (Boundary Limits of the Poisson Integral Representation). Let
v be as given in (5.412) where f is continuous on [−π, π]. Then, for every
ζ ∈ [−π, π],
lim |v(r, θ) − f (ζ)| = 0. (5.422)
(r,θ)→(1,ζ)
so that
Z π
v(r, θ) − f (ζ) = P (r, θ − ξ)(f (ξ) − f (ζ)) dξ. (5.424)
−π
Next, take absolute values on both sides of (5.424) and use the positivity of the
Poisson kernel (see property (i) in Proposition 5.3.15) to obtain that
Z π
|v(r, θ) − f (ζ)| 6 P (r, θ − ξ)|f (ξ) − f (ζ)| dξ. (5.425)
−π
We’ll next divide the integral on the right–hand side of (5.425) into three inte-
grals over the domains [−π, ζ − δ1 ], [ζ − δ1 , ζ + δ1 ] and [ζ + δ1 , π], respectively.
166 CHAPTER 5. SOLVING LINEAR PDES
Next, we estimate the integral over [ζ + δ1 , π]. Using the estimate in (5.415)
and the triangle inequality we obtain
Z π Z π
P (r, θ − ξ)|f (ξ) − f (ζ)| dξ < 2M P (r, θ − ξ) dξ, (5.427)
ζ+δ1 ζ+δ1
P (r, ω) < P (r, δ1 /2), for all ω ∈ [δ1 /2, π]. (5.431)
Calculations similar to those leading to (5.434) show that there exists δ3 > 0
such that
Z ζ−δ1
δ1 ε
|r − 1| < δ3 and |θ − ζ| < ⇒ P (r, θ − ξ)|f (ξ) − f (ζ)| dξ < . (5.435)
2 −π 3
δ1
Letting δ = min , δ2 , δ3 , we see that, in view of (5.425), (5.426), (5.434)
2
and (5.435),
This completes the proof of the boundary limits lemma for the case ζ ∈ (−π, π).
The case in which ζ is one of the end–points of the interval [−π, π] can be treated
in an analogous manner to the interior point case using one–sided limits at those
points.
where P (r, θ) denotes the Poisson kernel for the unit disc in R2 given in (5.396)
or (5.408); that is,
1 1 − r2
P (r, θ) = , for 0 6 r < 1 and θ ∈ [−π, π], (5.438)
2π 1 − 2r cos(θ) + r2
We showed in Section 5.3.5 that
Z π
v(r, θ) = P (r, θ − ξ)f (ξ) dξ, for 0 6 r < 1, θ ∈ [−π, π], (5.439)
−π
which shows that v is continuous at every point (1, ζ) ∈ ∂D1 . This completes
the proof of the following result.
Theorem 5.3.17 (Existence for the Dirichlet Problem for the Disc, I). Let
f : R → R be a continuous, 2π–periodic function. There exists a function v ∈
C 2 (D1 , R) ∩ C(D1 , R) that solves the BVP in (5.436).
By converting back to Cartesian coordinates from polar coordinates, we
obtain a solution of the Dirichlet problem in (5.332); namely,
uxx + uyy = 0 in D1 ;
(5.440)
u(x, y) = g(x, y), for (x, y) ∈ ∂D1 ,
so that, (ω1 (ξ), ω2 (ξ)) ∈ ∂D1 for all ξ ∈ (−π, π]. Consequently, denoting the
Euclidean norm of vector in R2 by | · |,
|(x, y) − (ω1 , ω2 )|2 = r2 − 2(x, y) · (ω1 , ω2 ) + 1, (5.441)
where the middle–term on the right–hand side of (5.441) is the dot–product of
the the vectors (x, y) and (ω1 , ω2 ). Thus,
|(x, y) − (ω1 , ω2 )|2 = r2 − 2r cos(θ − ξ) + 1. (5.442)
Hence, in view of (5.438), we see from (5.442) that
1 1 − |(x, y)|2
P (r, θ − ξ) = , (5.443)
2π |(x, y) − (ω1 , ω2 )|2
for (x, y) ∈ D1 and (ω1 , ω2 ) ∈ ∂D1 .
The right–hand side of the equation in (5.443) gives the Poisson kernel in
Cartesian coordinates. We write
1 1 − |(x, y)|2
P ((x, y), (ω1 , ω2 )) = , (5.444)
2π |(x, y) − (ω1 , ω2 )|2
for (x, y) ∈ D1 and (ω1 , ω2 ) ∈ ∂D1 . Then, in view of (5.439) and (5.444), we
define
I
u(x, y) = P ((x, y), ω)g(ω) dsω , for (x, y) ∈ D1 , (5.445)
∂D1
where ω = (ω1 , ω2 ) ∈ ∂D1 and dsω is the element of arc–length along the circle
∂D1 .
The boundary limit Lemma 5.3.16 implies that
I
lim P ((x, y), ω)g(ω) dsω = g(z1 , z2 ), (5.446)
(x,y)→(z1 ,z2 ) ∂D1
(x,y)∈D1
for all (z1 , z2 ) ∈ ∂D1 . Thus, in view of (5.437), (5.445) and (5.446), we see that
the function u : D1 → R given by
I
P ((x, y), ω)g(ω) dsω , for (x, y) ∈ D1 ;
∂D1
u(x, y) = (5.447)
for (x, y) ∈ ∂D1 ,
g(x, y),
where P ((x, y), ω), for (x, y) ∈ D1 and ω ∈ ∂D1 , is the Poisson kernel for the
unit disc, D1 , given in (5.444), gives a solution of the boundary value problem
in (5.440). We state this fact as the following existence theorem, which is,
essentially, the Cartesian coordinates version of Theorem 5.3.17.
Theorem 5.3.18 (Existence for the Dirichlet Problem for the Disc, II). For
any given g ∈ C(∂D1 , R), there exists a function u ∈ C 2 (D1 , R) ∩ C(D1 , R)
that solves the BVP in (5.440). Indeed, u is given by the Poisson integral
representation in (5.447).
170 CHAPTER 5. SOLVING LINEAR PDES
|f (x, y) − f (ξ, η)| 6 M |(x, y) − (ξ, η)|α , for all (x, y), (ξ, η) ∈ Ω. (5.450)
where C is a constant, are harmonic in the punctured plane, R2 \{(0, 0)}; that
is, W solves
uxx + uyy = 0 in R2 \{(0, 0)}. (5.452)
We also saw that the functions in (5.451) are radially symmetric solutions of
the problem in (5.452).
5.4. GREEN’S FUNCTIONS 171
where we have written W for W(xo ,yo ) and Dε for Dε (xo , yo ), and the minus
sign before the second integral in (5.459) indicates the fact that the outward,
unit normal of ∂Ωε at a point on ∂Dε (xo , yo ), points toward the point (xo , yo ).
We can therefore rewrite (5.458) as
ZZ I I
∂u ∂W ∂u ∂W
W ∆u dxdy = W −u ds − W −u ds,
Ωε ∂Ω ∂n ∂n ∂Dε ∂n ∂n
so that,
ZZ ZZ ZZ
W ∆u dxdy − W ∆u dxdy 6 |W ||∆u| dxdy. (5.462)
Ω Ωε Dε
We can use integration by parts and L’Hospital’s Rule to evaluate the integral
on the right–hand side of (5.464) to get
2
ε2
ZZ
ε
|W (x, y)| dxdy = −2π|C| ln ε − . (5.465)
Dε 2 4
for 0 < ε < 1. Letting ε → 0+ in (5.466) and using the fact that
lim ε ln ε = 0, (5.467)
ε→0+
which follows from L’Hospital’s Rule, we obtain from (5.466) and (5.467) that
ZZ ZZ
lim+ W ∆u dxdy − W ∆u dxdy = 0,
ε→0 Ωε Ω
where M1 is given by
M1 = max |∇u(x, y)|,
(x,y)∈Ω
It remains to evaluate
I
∂W
lim u ds,
ε→0+ ∂Dε ∂n
∂W
where the outward, normal derivative, , of W given in (5.454) is the deriva-
∂n
tive of W with respect to the outward, radial direction from (xo , yo ) on the
boundary of Dε (xo , yo ). Wring
p
r = (x − xo )2 + (y − yo )2 ,
Next, let ε → 0+ on both sides of (5.460), and use the limits in (5.461),
(5.470) and (5.473), to get that
ZZ I
∂u ∂W
−2πCu(xo , yo ) = − W ∆u dxdy + W −u ds. (5.474)
Ω ∂Ω ∂n ∂n
From this point on we set
−2πC = 1,
or
1
C=− ; (5.475)
2π
so that, the equation in (5.474) becomes
ZZ I
∂u ∂W
u(xo , yo ) = − W ∆u dxdy + W −u ds, (5.476)
Ω ∂Ω ∂n ∂n
5.4. GREEN’S FUNCTIONS 175
defined by
1 p
W(xo ,yo ) (x, y) = − ln (x − xo )2 + (y − yo )2 , if (x, y) 6= (xo , yo ). (5.477)
2π
It what follows we will show how to use the formula in (5.476) to obtain a
representation of harmonic function u ∈ C 2 (Ω, R)∩C(Ω, R) in terms of its values
on the boundary of Ω. If we succeed, we will be able to obtain a representation
for a solution of the Dirichlet problem in (5.448); namely,
uxx + uyy = 0 in Ω;
(5.478)
u(x, y) = g(x, y), for (x, y) ∈ ∂Ω,
∆H = 0, in Ω. (5.479)
1 p
W(x,y) (ξ, η) = − ln (ξ − x)2 + (η − y)2 , if (ξ, η) 6= (x, y), (5.484)
2π
and (x, y) ∈ Ω, we obtain from (5.483) the integral representation formula
ZZ
u(x, y) = − (W(x,y) (ξ, η) + H(ξ, η))∆u(ξ, η) dξdη
Ω
I (5.485)
∂
− u(ξ, η) [W(x,y) (ξ, η) + H(ξ, η)] ds,
∂Ω ∂n
for all (x, y) ∈ Ω. Thus, according to (5.488), if we are able to find a Green’s
function for a domain Ω ⊂ R2 , we will be able to represent any function u ∈
C 2 (Ω, R) ∩ C(Ω, R) in terms of its Laplacian, ∆u, in Ω, and its boundary values
on ∂Ω. In other words, we will be able to construct a candidate for a solution
of the Dirichlet problem for Poisson’s equation in (5.449); namely,
ZZ
u(x, y) = G((x, y), (ξ, η))f (ξ, η) dξdη
Ω
I (5.489)
∂
− g(ξ, η) [G((x, y), (ξ, η))] ds,
∂Ω ∂n
5.4. GREEN’S FUNCTIONS 177
Thus, setting
∂
P ((x, y), (ξ, η)) = − [G((x, y), (ξ, η)) (5.491)
∂n
in (5.490), we recover the Poisson integral representation formula for a harmonic
function in Ω in terms of of its boundary values,
I
u(x, y) = g(ξ, η)P ((x, y), (ξ, η)) ds, for (x, y) ∈ Ω, (5.492)
∂Ω
(ξ, η) 7→ G((x, y), (ξ, η)) − W(x,y) (ξ, η), for (ξ, η) ∈ Ω,
is harmonic in Ω and
We note that a given domain Ω can have at most one Green function. For
suppose that there are two functions, G1 ((x, y), (ξ, η)) and G2 ((x, y), (ξ, η)),
with the properties that the maps
(ξ, η) 7→ Gi ((x, y), (ξ, η)) − W(x,y) (ξ, η), for (ξ, η) ∈ Ω, and i = 1, 2,
Hi (ξ, η) = Gi ((x, y), (ξ, η)) − W(x,y) (ξ, η), for (ξ, η) ∈ Ω, and i = 1, 2,
Consequently,
G1 ((x, y), (ξ, η)) = G2 ((x, y), (ξ, η)), for all (ξ, η) ∈ Ω with (ξ, η) 6= (x, y).
In the following example, we compute the Green’s function for the disc of
radius R around the origin, DR , in R2 .
H(x,y) (ξ, η) = −W(x,y) (ξ, η), for all (ξ, η) ∈ ∂DR . (5.494)
In the construction, we will use the fact that, if (x∗ , y ∗ ) 6∈ DR , then the function
p
(ξ, η) 7→ ln (ξ − x∗ )2 + (η − y ∗ )2 , for (ξ, η) ∈ DR ,
or
(ξ, η) 7→ ln |(ξ, η) − (x∗ , y ∗ )|, for (ξ, η) ∈ DR ,
is harmonic in DR .
We first consider the case in which (x, y) ∈ DR and (x, y) 6= (0, 0); the case
(x, y) = (0, 0) with be dealt with separately.
Assume that (x, y) ∈ DR and (x, y) 6= (0, 0). We pick (x∗ , y ∗ ) to be a point
on the ray emanating from the origin, (0, 0), and going through the point (x, y)
at a distance λ > R from the origin; so that,
λ
(x∗ , y ∗ ) = (x, y), (5.495)
|(x, y)|
1
H(x,y) (ξ, η) = ln (Q(x, y)|(ξ, η) − (x∗ , y ∗ )|) , for (ξ, η) ∈ DR , (5.496)
2π
5.4. GREEN’S FUNCTIONS 179
Where Q(x, y) > 0, along with λ > R in (5.495), be chosen so that the boundary
condition in (5.494) is satisfied.
We note that H(x,y) is harmonic in DR , since |(x∗ , y ∗ )| = λ > R, according
to (5.495).
Next, use the definition of (x∗ , y ∗ ) in (5.495) to compute, for (ξ, η) ∈ ∂DR ,
λ
|(ξ, η) − (x∗ , y ∗ )|2 = R2 − 2 (x, y) · (ξ, η) + λ2 ,
|(x, y)|
or
R2 |(x, y)|
λ
|(ξ, η) − (x∗ , y ∗ )|2 = − 2(x, y) · (ξ, η) + λ|(x, y)| . (5.497)
|(x, y)| λ
R2
λ= , (5.498)
|(x, y)|
we obtain from (5.497) that
R2
|(ξ, η) − (x∗ , y ∗ )|2 = |(x, y)|2 − 2(x, y) · (ξ, η) + R2 ,
|(x, y)|2
or
R2
|(ξ, η) − (x∗ , y ∗ )|2 = |(x, y) − (ξ, η)|2 ,
|(x, y)|2
from which we get that
|(x, y)|
|(ξ, η) − (x∗ , y ∗ )| = |(ξ, η) − (x, y)|, for (ξ, η) ∈ ∂DR . (5.499)
R
It follows from (5.499) that
1 |(x, y)| 1
ln |(ξ, η) − (x∗ , y ∗ )| = ln (|(ξ, η) − (x, y)|) , (5.500)
2π R 2π
for all (ξ, η) ∈ ∂DR . Thus, comparing (5.496) with the left–hand side of (5.500)
we see that, we see that we can take
|(x, y)|
Q(x, y) =
R
in (5.496) to obtain
1 |(x, y)|
H(x,y) (ξ, η) = ln |(ξ, η) − (x∗ , y ∗ )| , for (ξ, η) ∈ DR , (5.501)
2π R
where, according to (5.495) and (5.498),
R2
(x∗ , y ∗ ) = (x, y), for (x, y) ∈ DR with (x, y) 6= (0, 0). (5.502)
|(x, y)|2
180 CHAPTER 5. SOLVING LINEAR PDES
which is the assertion in (5.494). It then follows from the definition of the Green
function in (5.486) and (5.487) that
for (x, y), (ξ, η) ∈ DR with (x, y) 6= (ξ, η), and (x, y) 6= (0, 0), or
1 |(x, y)| ∗ ∗
G((x, y), (ξ, η)) = ln |(ξ, η) − (x , y )|
2π R
(5.503)
1
− ln (|(ξ, η) − (x, y)|) ,
2π
for (x, y), (ξ, η) ∈ DR with (x, y) 6= (ξ, η), and (x, y) 6= (0, 0).
Next, for the case (x, y) = (0, 0), observe that the constant function
1
H(ξ, η) = ln R, for (ξ, η) ∈ R2 ,
2π
is harmonic everywhere. Furthermore, setting
1
W (ξ, η) = − ln |(ξ, η)|, for (ξ, η) 6= (0, 0),
2π
we see that
H(ξ, η) = −W (ξ, η), for all (ξ, η) ∈ ∂DR .
Thus, we define
1 1
G((0, 0), (ξ, η)) = ln R − ln |(ξ, η)|, for (ξ, η) 6= (0, 0). (5.504)
2π 2π
The function, G((x, y), (ξ, η)), defined in (5.503) and (5.504) is the Green
function of the disc of radius R around the origin in R2 . We can use it to solve
the following Dirichlet problem for Poisson’s equation in DR :
−∆u(x, y) = f (x, y), for (x, y) ∈ DR ;
(5.505)
u(x, y) = g(x, y), for (x, y) ∈ ∂DR ,
1 |(x, y)| 1
G((x, y), (ξ, η)) = ln + ln |(ξ, η) − (x∗ , y ∗ )|2
2π R 4π
(5.508)
1
− ln |(ξ, η) − (x, y)|2 ;
4π
so that, taking the partial derivative with respect to ξ,
∂ 1 ξ − x∗
G((x, y), (ξ, η)) =
∂ξ 2π |(ξ, η) − (x∗ , y ∗ )|2
(5.509)
1 ξ−x
− ,
2π |(ξ, η) − (x, y)|2
∂ 1 ξ 2 + η 2 − x∗ ξ − y ∗ η
G((x, y), (ξ, η)) =
∂n 2πR |(ξ, η) − (x∗ , y ∗ )|2
1 ξ 2 + η 2 − xξ − yη
− ,
2πR |(ξ, η) − (x, y)|2
∂ 1 R2 − (x∗ , y ∗ ) · (ξ, η)
G((x, y), (ξ, η)) =
∂n 2πR |(ξ, η) − (x∗ , y ∗ )|2
(5.511)
1 R2 − (x, y) · (ξ, η)
− ,
2πR |(ξ, η) − (x, y)|2
182 CHAPTER 5. SOLVING LINEAR PDES
1 R2 − (x, y) · (ξ, η)
− ,
2πR |(ξ, η) − (x, y)|2
for (ξ, η) ∈ ∂DR , which can be rewritten as
∂ 1 |(x, y)|2 − R2
G((x, y), (ξ, η)) = , for (ξ, η) ∈ ∂DR ,
∂n 2πR |(ξ, η) − (x, y)|2
from which we get that
∂ 1 R2 − |(x, y)|2
− G((x, y), (ξ, η)) = , for (ξ, η) ∈ ∂DR . (5.512)
∂n 2πR |(ξ, η) − (x, y)|2
Observe that the expression in (5.512) also works for (x, y) = (0, 0), according
to the definition of G((0, 0), (ξ, η)) in (5.504).
The expression on the right–hand side of (5.512) is the Poisson kernel,
P ((x, y), (ξ, η)), for the disc DR (compare with the result of Problem 2 in As-
signment #6). We therefore get from (5.506) and (5.512) that the solution of
the Dirichlet problem for Poisson’s equation in (5.505) should have the repre-
sentation
ZZ
u(x, y) = G((x, y), (ξ, η))f (ξ, η) dξdη
DR
I (5.513)
+ P ((x, y), (ξ, η))g(ξ, η) ds,
∂DR
where G((x, y), (ξ, η)) is the Green’s function of DR defined in (5.503) and
(5.504), is a candidate for a solution of the Dirichlet problem for Poisson’s
equation in DR :
−∆u(x, y) = f (x, y), for (x, y) ∈ DR ;
(5.517)
u(x, y) = 0, for (x, y) ∈ ∂DR ,
G((x, y), (ξ, η)) = H((x, y), (ξ, η)) + W ((x, y), (ξ, η)), (5.520)
184 CHAPTER 5. SOLVING LINEAR PDES
is harmonic if Ω; so that,
and
1
W ((x, y), (ξ, η)) = − ln |(ξ, η) − (x, y)|, for (x, y) 6= (ξ, η). (5.522)
2π
It is also the case that
G((x, y), (ξ, η)) = 0, for all (x, y) ∈ Ω and (ξ, η) ∈ ∂Ω. (5.523)
It then follows from the definition of the Green’s function in (5.520), (5.521)
and (5.522) that v1 is harmonic in Ω\{(x1 , y1 )}; indeed, there exists a harmonic
function h1 : Ω → R such that
v1 (ξ, η) = h1 (ξ, η) + W ((x1 , y1 ), (ξ, η)) for (ξ, η) ∈ Ω\{(x1 , y1 )}. (5.527)
Furthermore,
v1 (ξ, η) = 0, for (ξ, η) ∈ ∂Ω, (5.528)
by virtue of (5.523).
Similarly, defining v2 : Ω\{(x2 , y2 )} → R by
v2 (ξ, η) = h2 (ξ, η) + W ((x2 , y2 ), (ξ, η)), for (ξ, η) ∈ Ω\{(x2 , y2 )}; (5.530)
|∇v1 (ξ, η)| 6 M10 , for all (ξ, η) ∈ Dεo (x2 , y2 ). (5.537)
Consequently, using the estimates in (5.537) and (5.538), we obtain the estimate
I
∂v1
h2 (ξ, η) ds 6 M10 M2 2πε, for 0 < ε < εo . (5.539)
∂D2 ∂n
Combining the results in (5.540) and (5.542) with (5.536) we conclude that
I
∂v1
lim+ v2 ds = 0. (5.543)
ε→0 ∂Dε (x2 ,y2 ) ∂n
5.4. GREEN’S FUNCTIONS 187
and
M20 = max |∇h2 (ξ, η)|.
(ξ,η)∈D εo (x2 ,y2 )
the average value of v1 over the circle ∂Dε (x2 , y2 ). Thus, since v1 is continuous
on Dεo (x2 , y2 ), it follows from (5.549) that
I
1 ∂
lim+ v1 ln |(ξ, η) − (x2 , y2 )| ds = v1 (x2 , y2 ). (5.550)
ε→0 2π ∂D2 ∂n
188 CHAPTER 5. SOLVING LINEAR PDES
Hence, letting ε → 0+ in (5.546), and using the limits in (5.548) and (5.550),
we get that I
∂v2
lim+ v1 ds = −v1 (x2 , y2 ). (5.551)
ε→0 ∂Dε (x2 ,y2 ) ∂n
Similar calculations to those leading to (5.551) can be used to show that
I
∂v1
lim+ v2 ds = −v2 (x1 , y1 ). (5.552)
ε→0 ∂Dε (x1 ,y1 ) ∂n
v1 (x2 , y2 ) = v2 (x1 , y1 ),
is harmonic as a function of (x, y), if (x, y) 6= (ξ, η). Using the definition of G
in (5.520), we then get that
H((x, y), (ξ, η)) = G((x, y), (ξ, η)) − W ((x, y), (ξ, η)),
|f (x, y) − f (ξ, η)| 6 M |(x, y) − (ξ, η)|α , for all (x, y), (ξ, η) ∈ Ω. (5.555)
It follows from (5.555) that f is continuous on Ω and that there exists a constant
Mo > 0 such that
−∆u = f, in Ω.
Furthermore, it follows from the definition of u in (5.557) and from (5.554) that
Thus, the function u given in (5.557) solves the Dirichlet problem for Poisson’s
equation given in (5.519); namely,
−∆u(x, y) = f (x, y), for (x, y) ∈ Ω;
(5.558)
u(x, y) = 0, for (x, y) ∈ ∂Ω.
190 CHAPTER 5. SOLVING LINEAR PDES
Appendix A
In this appendix we present some of the facts from the theory of ordinary dif-
ferential equations (ODEs) that are used in these notes. We begin with linear
second order ODEs with constant coefficients.
ay 00 + by 0 + cy = 0, (A.1)
am2 + bm + c = 0. (A.3)
191
192 APPENDIX A. FACTS FROM THE THEORY OF ODES
If the roots, r1 and r2 , of the characteristic equation in (A.3) are real and
distinct, the general solution of the ODE in (A.1) is given by
Theorems About
Integration
∂H
Assume that the functions H and are absolutely integrable over [a, b].
∂t
1
Then, h is C and its derivative is given by
Z b
0 ∂
h (t) = [H(x, t)] dx.
a ∂t
∂ ∂
Assume that the functions H, [H(x, y, t)] and [H(x, y, t)] are absolutely
∂y ∂t
193
194 APPENDIX B. THEOREMS ABOUT INTEGRATION
integrable over [a, b]. Then, h is C 1 and its partial derivatives are given by
Z t
∂ ∂
[h(y, t)] = [H(x, y, t)] dx
∂y a ∂y
and Z t
∂ ∂
[h(x, t)] = H(t, y, t) + [H(x, y, t)] dx.
∂t a ∂t
→
− →
−
Then F has units of mass per unit length, per unit time. The vector field F in
(B.5) is called the flow field and it measures the amount of fluid per unit time
that goes through a cross section of unit length perpendicular to the direction of
→
−
V . Thus, to get a measure of the amount of fluid per unit time that crosses the
boundary ∂Ω in direction away from the region Ω, we compute the line integral
→
−
I
F ·nb ds, (B.6)
∂Ω
where ds is the element of arc–length along ∂Ω, and n b is unit vector that is
perpendicular to the curve ∂Ω and points away from Ω. The, expression in (B.6)
→
−
is called the flux of the flow field F across ∂Ω and it measures the amount of
fluid per unit time that crosses the boundary ∂Ω.
→
− →
−
On the other hand, the divergence, div F , of the flow field F in (B.5) has
units of mass/time × length2 , and it measures the amount of fluid that diverges
from a point per unit time per unit area. Thus, the integral
→
−
ZZ
div F dxdy (B.7)
Ω
the total amount of fluid leaving the reagin Ω per unit time. In the case where
there are not sinks or sources of fluid inside the region Ω, the integrals in (B.6)
and (B.7) must be equal; so that,
→
− →
−
ZZ I
div F dxdy = F ·nb ds. (B.8)
Ω ∂Ω
where n
b is the outward, unit, normal vector to ∂Ω that exists everywhere on
∂Ω, except possibly at finitely many points.
→
−
that, for the C 1 vector field F given in (B.3), the line integral on the right–hand
side of (B.9) can be written as
1
→
−
I Z
1
F ·n
b ds = (P (σ(t)), Q(σ(t))) · (ẏ(t), −ẋ(t)) |σ 0 (t)| dt,
∂Ω 0 |σ 0 (t)|
or
1
→
−
I Z
F ·n
b ds = [P (σ(t))ẏ(t) − Q(σ(t))ẋ(t)] dt,
∂Ω 0
which we can write, using differentials, as
→
−
I I
F ·nb ds = (P dy − Qdx). (B.11)
∂Ω ∂Ω
→
−
Thus, using the definition of the divergence of F in (B.4) and (B.11), we can
rewrite (B.9) as
ZZ I
∂P ∂Q
+ dxdy = (P dy − Qdx), (B.12)
Ω ∂x ∂y ∂Ω
for the area of the region Ω enclosed by a simple closed curve ∂Ω.
Appendix C
Kernels
We present two derivations of (C.2). The first one involves the use of Euler’s
formula
eiy = cos y + i sin y, for all y ∈ R. (C.3)
The second involves the use of some trigonometric identities.
πθ
Denote by x and compute
L
N
X 1 − ei(N +1)x
eix = , for x 6= 0. (C.4)
n=0
1 − eix
Observe that the real part of the expression on the left–hand side of (C.4) is
"N # N
X X
ix
Re e =1+ cos(nx), for x ∈ R, (C.5)
n=0 n=1
197
198 APPENDIX C. KERNELS
or
1 − cos x − cos[(N + 1)x](1 − cos x) + sin[(N + 1)x]) sin x
It then follows that the real part of the expression on the right–hand side of
(C.4) is
or
1
sin N + x
1 − ei(N +1)x
1 2
Re = + x , (C.9)
1 − eix 2 2 sin
2
where we have used the trigonometric identity
sin(A − B) = sin A cos B − cos A sin B, for A, B ∈ R.
In view of (C.4), (C.5) and (C.9), we can then write
1
N sin N + x
X 1 2
1+ cos(nx) = + x , for x 6= 0. (C.10)
n=1
2 2 sin
2
1
Finally, subtract from both sides of (C.10) to get
2
1
N sin N + x
1 X 2
+ cos(nx) = x , for x 6= 0,
2 n=1 2 sin
2
from which we derive (C.2), in view of (C.1).
Alternatively, use (C.1) to compute
X N
πθ 1 πθ πθ nπθ
sin DN (θ) = sin + sin cos , (C.11)
2L 2 2L n=1
2L L
and use the trigonometric identity
2 sin A cos B = sin(A + B) + sin(A − B)
to each of the terms in the sum in (C.11) to get
N X N
X πθ nπθ 1 1 πθ 1 1 πθ
sin cos = sin n + + sin −n ,
n=1
2L L n=1
2 2 L 2 2 L
or
N N
X πθ nπθ X 1 1 πθ 1 1 πθ
sin cos = sin n+ − sin n − ,
n=1
2L L n=1
2 2 L 2 2 L
since sin is an odd function. Note that the sum above telescopes to
N
X πθ nπθ 1 1 πθ 1 πθ
sin cos = sin N + − sin . (C.12)
n=1
2L L 2 2 L 2 2L
Combining (C.11) and (C.12) yields
πθ 1 1 πθ
sin DN (θ) = sin N + ,
2L 2 2 L
from which (C.2) follows.
200 APPENDIX C. KERNELS
Bibliography
201