Chapter 3
Linear Second-Order Equations
In this chapter, we derive the heat equation in general dimension, the equation of wave propa-
gation in an elastic membrane, and the Poisson's equation. We present the properties of each
equation and then classify the linear second-order partial differential equations into elliptic,
parabolic, and hyperbolic equations. The main reason for t his classification is that ea ch class
of equation has distinct solutions that behave differently. For instance, elliptic equations
have s o luti o ns that are smooth and have no singularities, whereas hyperbolic equations have
solutions that have s ingularities and can exhibit wave-like b eh avior. Understanding these
differences is crucial for solving and interpreting the solutions of partial differential equations
in various fields of science and engineering.
3.1 Heat equation
The heat equation is a fundamental equation in physics, engineering, and applied mathe-
matics that is used to model a wide range of phenomena, inclu ding heat transfer, thermal
diffusion, and mass diffusion. In this section, we will first de rive the heat equation for solid
conductive continua, and then i ntroduce the advection-diffusion equation for fluids.
3.1.1 Derivation of heat eq u ation
In the first chapter, we demonstrated the derivation of the equation for heat flow in a
conductive rod. In this chapter, we pres ent the derivation in three-dimensional space (R
3
)
to gain a deeper understanding of it s physics.
Consider a solid conductive continuum R
n
, which can be a bounded or unbounded set,
with or without a boundary. If is a bounded domain, we suppose that the only possibility
that exchange heat with the ambient space is through its boundary bnd(Ω). Let u(x; t)
denote th e temperature of an arbitrary point x2 at time t. For any bounded region D ,
the thermal energy inside D at time t is defined by the integral:
Q
D
(t) =
ZZZ
D
q(x; t) dV ;
1
where q is the density function of the heat energy. Therefore, the increme nt in energy can
be expressed as:
Q
D
(t + δt) ¡Q
D
(t) =
ZZZ
D
fq(x; t + δt) ¡ q(x; t)gdV :
In the limiting case a s δt!0, we obtain the fo llowing relation:
d
dt
Q
D
(t) =
ZZZ
D
ρc@
t
u(x; t) dV ;
where ρ is the density and c is the specific heat capacity of the continu um.
On the other hand, we can relate the rate o f change of thermal energy inside D to the heat
exchange through its boundary bnd(D). The rate of heat exch a nge through the boundary
of D is proportional to the temperature gradient, according to Fourier's law , which states
that th e heat flux J(x; t) at a point x in D is given by J(x; t)=¡ α(x)ru(x; t), where α(x)
is the thermal conductivity at x. Note that the negative sign arises because the direction
of the tempe rature gradient is towards increasing temperature, whereas the direction of the
heat flux is from hotter to colder regions. T he ne t ux leaving D through bnd(D) is equal
to the surface integral
ZZ
bnd(D)
J(z; t) ·ν(z) dS = ¡
ZZ
bnd(D)
α(z) ru(z; t) ·ν(z) dS ;
where ν(z) is the outward-pointing unit normal vector at z 2 bnd(D). By the conservation
of energy principle, we have
d
dt
Q
D
(t) = ¡
ZZ
bnd(D)
J(z; t) ·ν(z) dS ;
where the negative sign accounts for the fact that the integral measures the flux leaving D
at time t. Substituting
dQ
D
dt
by the triple integral and rearranging, we obtain
ZZZ
D
ρc@
t
u(x; t) dV =
ZZ
bnd(D)
α(z) ru(z; t) ·ν(z) dS:
Using Gauss's theorem, we can convert the surface integral in the right-hand side of the
previous relation to a volume integral. Applying the theorem yields:
ZZZ
D
div (α (x)ru(x; t))dV =
ZZ
bnd(D)
α(x)ru(x; t) ·ν(x)dS;
and y rearranging the terms, we arrive at the integral form of the heat equation:
ZZZ
[ρc@
t
u(x; t) ¡div (α(x) ru(x; t))] dV = 0:
Since this equation holds for arbitrary , we obtain the follow ing differential form of the
heat equation
u
t
(x; t) =
1
ρc
div (α(x) ru(x; t)):
2 Linear Second-Order Equations
If α is constant, then we have
u
t
(x; t) = k u;
where k =
α
ρc
> 0.
Remark 3.1. The he a t equation that we derived invo lves a first-order partial derivative of u
with respect to time. Therefore, to fully describe the heat dynamics, we need to specify the
initial condition u(x; 0) for x 2 . This is called the initial condition o f the heat equation.
In fact, the initial condition provides the initial thermal energy distribution for the sy stem,
which drives its subsequent behav ior. Additi o nally, in some physical systems, there may be
an internal source or sink of thermal energy that is not accounted for in our de rivation. In
the presence of such a source or sink, the heat equation takes the fo rm:
u
t
=
1
ρc
div (αru) + h;
where the function h represents the rate at which the internal source or sink produces or
absorbs energy.
3.1.2 Convection-diffusion equation
The convection-diffusion equation is a partial differential equation that is widely used to
model heat and mass transfer in fluids, such as in chemical engineering and environmental
science. The equation takes into account both convection, which is the transport of heat
or mass by the fluid flow, and diffusion, which is the transpo rt of heat or mass due to a
concentration gradient.
In the modeling of the heat problem, we assumed th a t heat is transferred only by conduc-
tivity. This assumption make s sense if the medium is solid. If the medium is a fluid or gas,
the heat can be transferred in the form of convection i n addition to diffusion. In this scenario,
the heat flux J consists of two terms, the conductivity term and the convection term:
J = J
cond
+ J
conv
:
If the fluid moves with velo city V , then J
conv
= uV , and thus
ZZ
bnd(D)
J ·ν dS = ¡
ZZ
bnd(D)
αru ·ν dS +
ZZ
bnd(D)
uV ·ν dS:
Therefore, the heat equation reads
ρcu
t
+div (uV ) =div (αru):
The above equation is called the convection-diffusion equation. The first term on the left-
hand side of the equation represents the change in temp erature with respect to time. The
second term represents the convective transport of heat by the uid flow, while the third
term represents the diffusive transpo rt due to a co nc entration gradient. In the presence of
a source or sink, the heat equation takes the form:
ρcu
t
+div (uV ) =div (αru) + h:
3.1 Heat equation 3
The convection-diffusion is a fundamental equation in many areas of science and engineering
and has many applications, such as modeling air and water pollution, heat transfer in pipes,
and drug delivery in biological systems.
3.1.3 Boundary conditions for bounded domains
Continuing with o ur discussion of the heat equation on bounded domain, it is worth noting
that the solution to a partial d ifferential equation, and especially to the heat equation,
depends heavily on the boundary conditions imposed on the bounded domain .
Recall that the Newton cooling law relates the te mperature inside and outside a region
. Speci fically, if T denotes the tempe rature outside , the law states that the heat flux
through bnd(Ω), is proportional to the difference T ¡u, where u is the temperature function
inside . This relationship is expressed mathematically by the Robin boundary condition
@u
@n
= κ(T ¡u);
where ν denotes the outward normal to the boundary, and κ is the conductivity factor of
the boundary. When κ = 0, the bounda ry is perfectly insulated, and the condition reduces
to the hom o g en eous Neumann boundary condition
@u
@n
= 0. More generally, the Neumann
boundary condition takes the form
@u
@n
= g on bnd(Ω), where g is a given function. In the
limiting case where κ is very large, the Ro bin condition can be rewritten as
T ¡u =
1
κ
@u
@n
;
and taking the limit κ!1 yields the condition u =T on bnd(Ω), which is called the Dirichlet
boundary condition.
Dirichlet boundary condition. The Dirichlet boundary condition is used when the
temperature at the boundary of the domain is known, such that u(z; t)= g(z; t) for
z 2bnd(Ω). This condition is named after the German mathematician Peter Gustav
Lejeune Dirichlet. For a heat equation, this means that the temperature at bnd(Ω) is
kept a t the known value g(z; t). If g is identically zero, then the condition is call ed
the homogeneous Dirichlet condition.
u = g
PDE inside
In one-dimensional problems, such as a conductive rod extending from x = 0 to
x = L, the condition becomes u(0; t) = g
0
and u(L; t) = g
L
, where g
0
and g
L
can be
constants or functions of tim e.
4 Linear Second-Order Equations
Neumann boundary condition. This condition is named after the German math-
ematician Carl Neumann. Mathematically, the Neumann bou ndary c ondition is
expressed as
@u
@n
:= ru ·ν = g;
where
@u
@n
is the direct ional derivative of u along the unit normal direction ν on
bnd(Ω). In the context of a heat problem, this condition specifies the heat flux through
the boundary bnd(Ω) in terms of function g. In part icular, if g is identically zero,
the boundary is perfectly insulated.
@u
@n
= g
ν
PDE inside
For a one-dimensional problem where the co nductive rod extends from x = 0 to
x = L, the condition reads:
u
x
(0; t ) = g
0
; u
x
(L; t) = g
L
:
Robin's or mixed boundary condition. As we observed above, the Robin boundary
condition is a straightforward derivation of the Newton cooling law in the context of
a heat problem. Mathematically, it can be expressed as
αu (z; t) + β
@u
@n
(z; t) = g(z; t):
where α and β are some coefficients, and g is a given function. In particular, if g
is identically zero, the condition is called the homogeneous Robin boundary condi-
tion. The Robin boundary condition is commonly used in problems that involve heat
transfer between a solid and a fluid, where the heat transfer coefficient (i.e., α / β)
depends on the properties of the fluid and the surface of the solid. It is also used
to model heat tran sfer in biological tissues, where the heat transfer coefficient can
depend on the blood flow rate a nd other physiological factors. The Robin boundary
condition is named after the French mathematician Victor Gustave Robin.
αu + β
@u
@n
= g
PDE inside
For a one-dimensional problem, the condition reads
α
1
u(0; t) + β
1
u
x
(0; t ) = g
0
α
2
u(L; t) + β
2
u
x
(L; t) = g
L
:
3.1 Heat equation 5
3.1.4 Fundamental solution of heat equatio n
Consider a con ductive rod of diffusivity factor k > 0, extended over the entire real line (¡1;
1). Suppose that an infinitely small portion of the rod at x = 0 has a temperature of u = 1,
and the temperature i s zero everywhere else. As time progresses, the heat diffuses across the
rod, but the total thermal energy remains constant as it was at t = 0. It can be shown that
the heat profile evolves according to the function:
Φ(x; t) =
1
4πkt
p
e
¡
x
2
4kt
;
for t > 0. In fact, one can verify that Φ
tt
= k Φ
xx
through di rect calculation. The graph of
Φ(x;t) demon strates the evolution of heat profile for t > 0. The following figure shows Φ(x;t)
for t = 0.1; 0.5 a nd 1 where k = 1:
-4 -2 0 2 4
0
0.2
0.4
0.6
0.8
That the total thermal energy remains constant over time follows from the fact that the
integral of Φ for a ny fixed value o f t is constant:
Z
¡1
1
Φ(x; t)dx = 1:
In fact, using the change of variable v =
x
4kt
p
, we have:
Z
¡1
1
Φ(x; t)dx =
1
π
p
Z
¡1
1
e
¡v
2
dv = 1:
This confirms the conservation of thermal energy since the initial heat sou rce is a unit
pointwise source at x = 0. The function Φ is called the Gaussian or heat kernel of a heat
equation.
Exercise 3.1. By direct calculation, verify that Φ(x; t) satisfies the heat equation u
t
= ku
xx
.
Now, consider the heat problem given by the partial differential equation
u
t
= ku
xx
; x 2(¡1; 1); t > 0;
with initial condition u(x; 0) = f(x), where f(x) is a continuous and integrable fun ction in
(¡1; 1). The follow ing theorem gives the solution to this problem:
6 Linear Second-Order Equations
Theorem 3.1. The solution to the given problem is given by the convolution
u(x; t) = f Φ :=
Z
¡1
1
Φ(x ¡ y; t) f(y) dy:
Proof. We rst show that u defined as above satisfies the partial differential equation. We
have
u
t
(x; t) =
Z
¡1
1
Φ
t
(x ¡ y; t) f (y) dy;
and u
xx
is
u
x x
(x; t) =
Z
¡1
1
Φ
xx
(x ¡ y; t) f(y) dy:
The above representations of the partial derivatives are justified by the assumption of inte-
grability of f, th a t is ,
Z
¡1
1
jf(x)jdx < 1;
and the Dominated Convergence Theorem which allows us to take the limits from outsi de
the integrals to inside the integrals. Accordingly, we have
u
t
¡ku
xx
=
Z
¡1
1
t
(x ¡ y; t) ¡kΦ
x x
(x ¡ y; t)] f(y) dy:
By the above exercise, we have Φ
t
¡k Φ
x x
= 0, and this completes the claim. It remains to
show that the convolution satisfies the initial condition for t !0. We have
lim
t!0
u(x; t) = lim
t!0
1
4πkt
p
Z
¡1
1
e
¡
(x¡y)
2
4kt
f(y)dy = lim
t!0
1
4πkt
p
Z
¡1
1
e
¡
y
2
4kt
f(x ¡ y)dy:
Using the substitution v =
y
4kt
p
, we can write
lim
t!0
u(x; t) = lim
t!0
1
π
p
Z
¡1
1
e
¡v
2
f(x ¡ 4kt
p
v)dv:
If we pass the limit t!0 inside the integral (which is justified by the Dominant Convergence
Theorem), we obtain
lim
t!0
u(x; t) =
1
π
p
Z
¡1
1
e
¡v
2
lim
t!0
f(x ¡ 4kt
p
v)dv:
Since f is continuous, we ob tain
lim
t!0
u(x; t) =
1
π
p
Z
¡1
1
e
¡v
2
f(x)dv =
f(x)
π
p
Z
¡1
1
e
¡v
2
dv = f (x);
which completes the proof.
Example 3.1. Consider the following problem:
8
<
:
u
t
= u
x x
u(x; 0) =
1 ¡1 < x < 1
0 otherwise
:
3.1 Heat equation 7
The initial heat p rofile, denoted by f(x), is 1 for x in the interval (¡1; 1), and zero elsewhere.
The figure below show s the solution at time t =0.1. Note that the solution is smooth, despite
the initial condition having jumps a t x = ±1. This is a characteristic property of the heat
equation, w he re it immediately smooths out any discontinuities after t = 0.
-4 -2 0 2 4
0
0.2
0.4
0.6
0.8
1
In this equation, we have taken k = 1. Remember that k is proportional to the diffusivity
of the rod. Therefore, for larger values of k, the solution u(x; t) spreads out faster, as show n
in the figure below for k = 4.
-4 -2 0 2 4
0
0.2
0.4
0.6
0.8
1
Exercise 3.2. Another interesting property of the convolution solution of the heat equation is that
the initial temperature of distant points can have an i mmediate impact on the heat evolution of a fixed
po int. This is because the solution invo lves an integral over the entire domain, and the contribution
from faraway points is not necess arily negligible. This is in contrast to so me other types of differential
equations, where the evolution of a point is only influenced by its immediate neighbors. This property
of the h eat equation m akes it particularly useful for modeling systems where heat can quickly spread
throughout a medium, such as in heat transfer pr ob lems in engineering or physics.
For the simple heat equation u
t
= u
xx
, suppose the initial condition is given by f (x) = δ(x) + δ(x ¡3),
that is, two point heat sources at x = 0 and x =3. Find the convolution solution and then use the following
cod e in Matlab to draw the solution at t = 0.5:
x=-10:0.01:10;
k=1;t=0.5;
Fi=@(z) exp(-z.^2/(4*k*t))/sqrt(4*pi*k*t);
plot(x,Fi(x)+Fi(x-3))
8 Linear Second-Order Equations
Exercise 3 .3. From the ab ove exercise, we learned that the heat equation is a diffusion equation,
which means that heat is gradually spread throughout the rod. This spreading process is affected by the
diffusivity factor k and the initial heat distribution, but it also depends on the position of the fixed p oint
relative to all other points on the rod. In other words, the heat at distant points has a n immediate impact
on the heat evolution of a fixed point. In addition to its smoothing-out property, another important
feature of the heat equation is its time-invariance. This means that the solutions to the system:
u
t
= ku
xx
t > 0
u(x; 0) = f (x)
;
and
u
t
= ku
xx
t > t
0
u(x; t
0
) = Φ(x; t
0
) f(x)
;
are the same. In other words, we can define a flow as
φ
t
(f (x)) = Φ(x; t) f(x);
that satisfies the property
φ
t+s
(f(x)) = φ
t
(φ
s
(f (x))):
This means that the heat evolution of a fixed point depends only on the initial heat distribution over
the entire rod a nd the diffusivity factor k, and not on the specific time at which we consider the system.
The ow φ
t
thus provides a way to study the time-evolution of the heat equation in a systematic way, by
tracking how initial heat distributions evolve over time. We will prove this fact in subsequent chapters
using the Fourier transform. However, we can verify t he fact for the Dirac delta function, f(x)=δ(x),
here. Consider the heat equation u
t
= u
xx
with initial condition u(x; 0)=δ(x). Show that the solution
of this problem at time t + s is equal to the solution of the system
u
t
= u
xx
u(x; s) = Φ(x; s)
;
at time t > 0. In other words, the following relation holds:
Φ(x; t) Φ(x; s) = Φ(x; t + s);
for t; s > 0.
Exercise 3.4. The above exercise helps us to gain a better understanding of the solution to a heat
equation as a convolution integral. Specifically, a heat problem can be viewed as a system with the
response φ
t
(f (x)) for the initial condition f . When f (x) = δ(x), the imp ulse response of the system is
Φ(x; t). Recall that any continuous function f(x) can be expressed as the convolution
f(x) =
Z
¡1
1
f(y) δ(x ¡ y) dy:
We can rewrite this integral as a Riemann sum
f(x)
X
n
f(y
i
) δ(x ¡ y
i
) δy
i
:
The response of the heat system to this function is
φ
t
(f(x))
X
n
f(y
i
) φ
t
[δ(x ¡y
i
)] δy
i
=
X
n
f(y
i
) Φ(x ¡ y
i
; t) δy
i
;
Returning to the integral formulation, we obtain
φ
t
(f(x)) =
Z
¡1
1
f(y) Φ(x ¡y; t) dy:
In the above proof, we have used the following properties
a) The solution of the system
u
t
= u
xx
u(x; 0) = αδ(x)
;
for a constant α is u = αΦ(x; t).
3.1 Heat equation 9
b) The solution of the system
u
t
= u
xx
u(x; 0) = δ(x ¡y)
;
is u = Φ(x ¡ y; t).
Justify both properties through direct calculation.
Exercise 3.5. We can compare the heat kernel Φ(x; t) to the Gaussian or normal probability density
function of a random variable X with expected value µ = 0 and variance σ
2
, given by
N(x; σ) =
1
σ 2π
p
e
¡
x
2
2σ
2
:
Upon comparison, we see that the standard deviation of the Gaussian distribution is related to time as
σ = 2t
p
.
a) Use a Z-tabl e and find a > 0 in terms of σ such that
P (¡a < X < a) = 0.95:
b) Since 95 percent energy of Φ is concentrated in (¡a; a), we can use this value to approximate the
convolution integral Φ(x; t) f(x) as
Φ(x; t) f (x)
Z
¡a
a
Φ(x ¡ y; t) f(y) dy:
c) Suppose jf (x)jM, show the inequality
Φ(x; t ) f (x) ¡
Z
¡a
a
Φ(x ¡ y; t) f (y) dy
0.05M :
Exercise 3.6. Verify that the function Φ
n
given by the function
Φ
n
(x
1
; :::; x
n
; t) = (4πkt)
¡n/2
e
¡(x
1
2
+···+x
n
2
)/4kt
;
solves the heat equation
u
t
= ku;
on R
n
, and furthermore the convolution
u(x; t) = Φ
n
(x; t) f (x);
solve the heat problem
u
t
= ku
u(x; 0) = f (x)
;
where f is a continuous and integrable function in R
n
. For the proof, we need the result
Z
R
n
Φ
n
(x
1
; :::x
n
; t)dV = 1;
which we accept without prof.
Exercise 3.7. Assume th at f is a piecewise continuous function and integrable in (¡1; 1). Suppose
that lim
x!0
+
f(x) = a and lim
x!0
¡
f(x) = ¡a. Show the following relation
lim
t!0
1
4πkt
p
Z
¡1
1
f(x) e
¡
x
2
4kt
dx = 0:
Exercise 3.8. Consider the equation u
t
= u
xx
for x 2 (¡1; 1), and t > 0. If u(x; t) and u
x
(x; t) are
bo unded and furthermore lim
x!±1
u(t; x) = 0, then show the following integral is decreasing
P (t) =
Z
¡1
1
ju(x; t)j
2
dx:
Conclude that the following problem
u
t
= u
xx
u(0; x) = f(x)
;
10 Linear Second-Order Equations
has a unique solution under the above as sumptions.
Exercise 3.9. Consider the equation u
t
=div ( α ru) on R
+
× R
n
. Suppose u(x; t) and ru(x; t) are
bo unded, and furthermore lim
jxj!±1
u(x; t) = 0. Show that the following integral is decreasin g
P (t) =
Z
¡1
1
ju(x; t)j
2
dV :
Conclude that the following problem
u
x
=div (αru)
u(x; 0) = f (x)
;
has a unique solution under the above as sumptions.
3.1.5 Solution for problems in bounded domains
The convolution solution is only applicable for heat problems defined on doma ins without
boundary. To see t his, consider the simple heat problem on the interval [0; 1]:
8
<
:
u
t
= ku
x x
x 2(0; 1); t > 0
u(0; t) = u(1; t) = 0
u(x; 0) = f (x)
:
Note that the convolution integral does not satisfy the prescribed boundary conditions at
x = 0; 1. Therefore, the heat ke rnel Φ (x; t) cannot be used to obtain the solution of heat
problems defined on domains with boundaries. In subsequent chapters, we will study partial
differential equations defined on bounded domains, and we will need to develop new tools
that allow us to express the solution o f PDEs as infinite s eries o r improper integrals .
However, we can still obtain some impo rtant properties o f the heat equati on on bounded
domains by solving boundary value problems. T he following exercises illustrate some of these
properties.
Exercise 3.10. Consider the equation u
t
= k u
xx
in (0; L) where k > 0 is a constant. Show th at the
integral
P (t) =
Z
0
L
ju(x; t)j
2
dx;
is decreasing for the following boundary conditions
i. u(0; t) = u(L; t) = 0
ii. u(0; t) ¡u
x
(0; t) = 0, u(L; t) + u
x
(L; t) = 0.
Exercise 3.11. Let be a bounded open set with smooth boundary bnd(Ω). Consider the following
Dirichlet problem
(
u
t
=div (αru)
uj
bnd(Ω)
=0
:
For the power function
P (t) =
Z
ju(x; t)j
2
dV ;
show P (t) is decreasing.
Exercise 3.12. Consider the following heat problem
8
<
:
u
t
= ku
xx
¡ru
u
x
(0; t) = u
x
(L; t) = 0
u(x; 0) = f (x)
:
3.1 Heat equation 11
where r is a constant. Let E(t) be the following energy function
E(t) =
Z
0
L
u(x; t) dx:
Prove the following relation
E(t) =
Z
0
L
f(x) dx
e
¡rt
:
Exercise 3.13. Repeat the above problem for the following equation
8
>
>
>
>
<
>
>
>
>
:
u
t
=div (αru) ¡ru
@u
@n
bnd(Ω)
= 0
u(x; 0) = f (x)
:
Exercise 3.14. Let be an open bounded set with smooth boundary bnd(Ω). Consider the following
heat problem on
8
>
>
>
>
<
>
>
>
>
:
u
t
= ku
@u
@n
bnd(Ω)
= 0
u(x; 0) = f (x)
;
a) Show the following relation
Z
u(x; t) dV =
Z
f(x) dV :
b) Now for the problem
8
>
>
>
>
<
>
>
>
>
:
u
t
= ku ¡ru
@u
@n
bnd(Ω)
= 0
u(x; 0) = f (x)
;
where r is a constant, show the following relation
Z
u(x; t) dV =
Z
f(x) dV
e
¡rt
:
3.2 Wave equation
3.2.1 Derivation of the eq u ation
In the first chapter, we obtained the equation of motion for a vibrating string by modeling it
as an infinite number of small mass-spring systems. We now extend this approach to derive
the wave equation for an elastic membrane in two-dimensional Euclidean space. The method
we use can be generalized to any dimensi o n R
n
, where n is greater than or equal to one.
Let be an elastic membrane in the two-dimensional space, and let D be an infinitesimal
element of area δA =δx δy. For the displacement function u(x; y; t), the tension on the
boundary of D is given by T =τ ru, where τ is called the stress and depend s on the phy sical
characteristics of the elastic material. The vector force exerted on D can be de fined by the
integral
F =
I
D
τ ru ·νdl;
12 Linear Second-Order Equations
where ν is the unit exterior normal vector t o the boundary of D. Applying Newton's second
law to this small portion of the membrane, we obtain
ρδAu
tt
= F ;
where ρ is the d en sity of the membrane. By applying the divergence theorem, we can write
u
tt
=
1
ρ
lim
δA!0
1
δA
ZZ
D
div (τ ru) dS =
1
ρ
div (τ ru):
If the medium is homogeneous with co nstant τ and ρ, we can write the derived equation
as u
tt
= c
2
u, where c=
τ
ρ
q
is the wave speed of the elastic medium. This equation is
known as the wave equation f o r elastic membranes. This method can b e extended to higher-
dimensional spaces, where the wave equation takes a similar form.
The derived formulation is incomplete without specifying the initial conditions for the
equation. Note that the equation c o ntains a second-order partial derivative of u with respect
to time. Therefore, the equation should be accompanied by two initial conditions: one to
specify the initial displacement u(x; y; 0) and the seco nd to specify the initial velocity u
t
(x;
y; 0). These initial conditions correspond, respectively, to the initial stretch potential energy
of the membrane and the initial kinetic energy. These energies cause a dynamic for the
membrane that forces it to oscillate like a wave function. Evidently, if u(x; y; 0) = 0 and
u
t
(x; y; 0) = 0, then u = 0 for all t. Therefore, we can write the wave equation in general
dimension as follows
8
>
>
<
>
>
:
u
tt
= c
2
u
u(x; y; 0) = f(x; y)
u
t
(x; y; 0) = g(x; y)
;
for x 2 R
n
, where f and g are initial displacement and velocity respectively.
If is a bounded domain, such as a string, then the natural boundary condition for the
equation is the homogeneo us Dirichlet condition. Physically, this conditio n means that the
endpoints of th e string are fastened and fixed. However, other possible boundary conditions
include the Neumann an d Robin conditions, which we have seen previously for the heat
equation. The wave equation with homogeneous Robin's boundary condition on a bounded
domain is
8
>
>
>
>
>
>
>
>
<
>
>
>
>
>
>
>
>
:
u
tt
= c
2
u on
αu + β
@u
@n
= 0 on bnd(Ω)
u(x; y; 0) = f (x; y)
u
t
(x; y; 0) = g(x; y)
:
3.2.2 Damped wave problem
In the formulation of the wave equation, we neglected the drag or resistance force that the
membrane may experience. To include such a possible forces, we assume that the drag forces
is proportional to the velocity of the displacement, that is,
f
drag
/u
t
;
3.2 Wave equation 13
and write the damped wave equation as
u
tt
+ ξu
t
= c
2
u;
where ξ > 0 is a constant. If in addition, the membrane is under an external force h, the
equation reads
u
tt
+ ξu
t
= c
2
u + h:
3.2.3 Solution of 1D wave equation
The one-dimensional wave equation u
tt
= c
2
u
x x
has a simple solution. By denoting the
time derivative with the operator @
t
and the space derivative with @
x
, we can represent
the equation as (@
t
¡c@
x
)(@
t
+ c@
x
) [u] = 0 or equivalently (@
t
+ c@
x
)(@
t
¡c@
x
) [u] = 0. This
representation al lows us to decompose the wave equation into two uncoupled first-order
PDEs:
u
t
+ cu
x
= 0
u
t
¡cu
x
= 0
;
which can be solved for arbitrary functions F (x ¡ ct) and G(x + c t). Thus, the general
solution to the equation is u(x; t) = F (x ¡ct) + G(x + ct).
Now, consider the wave equation with initial conditions u(x;0)= f (x), and u
t
(x;0)= g(x).
It is possible to determine F ;G such that the general solution satisfies these initial conditions.
To find F and G, we can u se the initial conditions to set up a system of equations. From u(x;
0)= f(x), we have F (x)+ G(x)= f(x). Taking the derivative with respect to t a nd evaluating
at t = 0, we get u
t
(x; 0) = F
0
(x) ¡G
0
(x) =
1
c
g(x), and finally we obtain the following system
(
F (x) + G(x) = f(x)
F
0
(x) ¡G
0
(x) =
1
c
g(x)
:
A simple manipulation gives the following solution
u(x; t) =
1
2
[f(x ¡ct) + f(x + ct)] +
1
2c
Z
x¡ct
x+ct
g(s) ds:
The above formula is common ly known a s the D'Alembert formula for solving 1D wave equa-
tions, named after the French mathematician Jean-Baptiste Rond d'Alembert. According to
the formula, the value of u at an arbitrary point (x; t) is determined by the values of f and
g in the segment [x ¡ct; x + ct]. This segment is referred to as the wave equation's influen ce
domain, as illustrated in the figure below. It is worth noting that if g 0, then the value o f
u(x; t) represents the average of the right wave f(x ¡ct) and the left wave f(x + ct), bo th
propagating with the same speed c.
In the wave equation, the value of u at an arbitrary point (x; t) is determined not only by
the initial conditions but also by the values of u a nd its derivatives in the segment [x ¡ct;
x + ct]. This means that the influence of far distant po ints on a fixed point is not immediate,
but rather it takes time for the values to propagate and reach the fixed point, and this time
is determined by the velocity c. So the behavior of the wave equation is very different from
that of the heat equation, where the value of u a t a point is determined by the co nvolution
over entire axis x.
14 Linear Second-Order Equations
x + ctx ¡ct x
(x; t)
x
η = x + ct
t
ξ = x ¡ct
Example 3.2. Consider the following problem
8
<
:
u
tt
= 9u
x x
u(x; 0) = f(x)
u
t
(x; 0) = 0
;
where f(x) is the function
f(x) =
1 ¡1 x 1
0 otherwise
The figure below illustrates the solution at time t = 0 and t = 1 for c = 3.
-6 -4 -2 0 2 4 6
0
0.2
0.4
0.6
0.8
1
As it is observed, the initial condition is split into two branches, one moving to the right
and one moving to the left with the same speed c = 3. T his is because the solution to the
wave equation is a superposition of two waves propagating in opposite directions. Each
wave carries half of the energy of the initial function, which is why we see two identical
waves moving away from the origin in the figure. While the heat equation can sm ooth out
discontinuities in the initial condition, wave equations are unable to do so, which can lead
to issues with properly defining the partial differential equation at these points.
Exercise 3.15. Consider the equation
u
tt
¡c
2
u
xx
= 0:
Take ξ = x ¡ct and η = x + ct. Show that the equation under the transformation reduces to
@
ξη
u = 0:
Solve the obtained equation and conclude that the solution of the equation is
u(x; t) = F (x ¡ct) + G(x + ct);
for arbitrary functions F ; G.
3.2 Wave equation 15
Exercise 3.16. Consider the non-homogeneous equation
(
u
tt
= c
2
u
xx
+ h(t; x)
u(x; 0) = u
t
(x; 0) = 0
:
a) Use the characteristic change of variable ξ = x ¡ct, η = x + ct to ch an ge the equation u
tt
= c
2
u
xx
to the form u
ξη
= 0.
b) Use this method to show that the solution to the above non-homogeneous problem is
u(x; t) =
Z Z
h(t
0
; x
0
) dt
0
dx
0
;
where is the inside of the triangle constructed on the characteristic ξ = x ¡ct, and η = x + ct
3.2.4 Solution on nite string
We will now derive the solution to a 1D wave problem on a string o f length L w hos e endpoints
at x = 0 and x = L are fixed. The problem is given by the wave equation:
u
tt
=c
2
u
xx
with the initial and bo undary conditions:
8
<
:
u(t; 0)=u(t; L)=0
u(x; 0)= f(x)
u
t
(x; 0)= g(x)
:
To solve this equation, we can use the following formula:
u(t; x) =
1
2
[f
odd
(x ¡ct) + f
odd
(x + ct)] +
1
2c
Z
x¡ct
x+ct
g
odd
(s) ds: (3.1)
Here, f
odd
and g
odd
are the odd extensions of f and g, respectively, and are 2L-periodic
functions. Recall that the odd extension of a f unction f defined on [0; L] is
f
odd
(x) =
f(x) 0 x L
¡f(¡x) ¡L x 0
:
The solution given by the above formula is obtained by applying the D'Alembert formula to
the odd extension of the initial condition f(x) and the odd extension of the initial velocity
g(x). The first term in solution represents the contrib ution of the initial displacement to
the solution, and the second term repres ents the contribution of the initial velocity to the
solution.
The boundary conditions u(t; 0)=u(t; L)=0 are automatically satisfied by this solution
since fodd(0)= f
odd
(L)=0.
Example 3.3. Consider the following problem
8
>
>
>
>
>
>
<
>
>
>
>
>
>
:
u
tt
= u
xx
u(0; t) = u(4; t) = 0
u(x; 0) = f (x)
u
t
(x; 0) = 0
;
16 Linear Second-Order Equations
where
f(x) =
8
<
:
x ¡1 1 x 2
3 ¡x 2 x 3
0 otherwise
:
The figure below illustrates the initial condition f(x), and u(x; 1):
0 1 2 3 4
0
0.2
0.4
0.6
0.8
1
For t > 1, two branches of waves hit the boundary points x = 0 and x = 4. According to
formula (3.1), the o dd extensio n of the wave moves back after it hits the end point. The
figure below de picts the solution u(x; t) fo r t = 0; 1; 3, and 4.
0 2 4
-1
-0.5
0
0.5
1
0 2 4
-1
-0.5
0
0.5
1
0 2 4
-1
-0.5
0
0.5
1
0 2 4
-1
-0.5
0
0.5
1
Exercise 3.17. Verify the formula ( 3.1).
3.2 Wave equation 17
Exercise 3.18. Consider the following equation
8
>
>
>
>
>
>
>
>
<
>
>
>
>
>
>
>
>
:
u
tt
= u
xx
¡1; 0 < x < 1
u(0; t) = u(1; t) = 0
u(x; 0) =
1
2
x(x ¡1);
u
t
(x; 0) = sin(πx)
: (3.2)
Verify that the following function is the solution of the above problem:
u(x; t) =
1
2
x(x ¡1) +
1
π
sin(πx) sin(πt):
Exercise 3.19. Consider the following wave problem
8
>
>
>
>
>
>
<
>
>
>
>
>
>
:
u
tt
= c
2
u
xx
u(0; t) = u(L; t) = 0
u(x; 0) = f (x)
u
t
(x; 0) = g(x)
:
The energy of the wave function is defined by the following formula
E(t) :=
Z
0
L
[ju
t
(x; t)j
2
+c
2
ju
x
(x; t)j
2
] dx:
a) Suppose that f
0
(x) is integrable. Verify the following formula which is known as the conservation
of energy of the wave function
E(t) =
Z
0
L
fjg(x)j
2
+c
2
jf
0
(x)j
2
gdx:
b) Conclude that the the given problem has a unique solution.
Exercise 3.20. Let R
2
be a bounded set with smooth boundary bnd(Ω). Consider the following
wave problem on
8
>
>
>
>
>
>
<
>
>
>
>
>
>
:
u
tt
= c
2
u
uj
bnd(Ω)
=0
u(x; y; 0) = f (x; y )
u
t
(x; y; 0) = g(x; y)
;
and assume that
ZZ
jg j
2
+c
2
jrf j
2
dS < 1:
a) The energy of the vibrating membrane is defined through the following integral
E(t): =
ZZ
[ju
t
(x; y; t)j
2
+c
2
jru(x; y; t)j
2
] dS
Prove that E(t) is constant and conclude the following relation
E(t) =
ZZ
[jg j
2
+c
2
jrf j
2
] dS:
b) Prove that the above problem can not have multiple solutions.
18 Linear Second-Order Equations
Exercise 3.21. Consider the following wave problem
8
>
>
>
>
>
>
<
>
>
>
>
>
>
:
u
tt
= u
xx
¡u
u(0; t) = u(L; t) = 0
u(x; 0) = f (x)
u
t
(x; 0) = g(x)
:
i. Show the following conservation of energy for the above problem
Z
0
L
(juj
2
+ ju
t
j
2
+ ju
x
j
2
) dx =
Z
0
L
(jf j
2
+ jf
0
j
2
) dx:
ii. If f (0) = f(L), show that
Z
0
L
(juj
2
+ ju
t
j
2
+ ju
x
j
2
) dx =
Z
0
L
jf + f
0
j
2
dx:
Exercise 3.22. Consider the following damped equation
8
>
>
>
>
>
>
<
>
>
>
>
>
>
:
u
tt
+ ξu
t
= u
xx
u(0; t) = u(L; t) = 0
u(x; 0) = f (x)
u
t
(x; 0) = g(x)
;
where ξk 0 is a constant. For E(t), the energy defined as
E(t): =
Z
0
L
(ju
t
j
2
+ ju
x
j
2
) dx;
show the following relation
e
¡2ξt
E(0) E(t) E(0):
Exercise 3.23. Consider the following wave problem
8
>
>
>
>
>
>
<
>
>
>
>
>
>
:
u
tt
= c
2
u
uj
bnd(Ω)
=0
u(x; 0) = f (x)
u
t
(x; 0) = g(x)
;
where x 2 R
n
, and is a bounded domain wit h smooth boundary bnd(Ω). Verify the following
conservation of energy formula for the problem
Z
[ju
t
(x; t)j
2
+c
2
jru(x; t)j
2
] dV =
Z
[jg(x)j
2
+c
2
jrf(x)j
2
] dV :
Exercise 3.24. Consider the following wave problem
8
<
:
u
tt
= u ¡u
u(t; bnd(Ω)) = 0;
u(0; p) = f(p); @
t
u(0; p) = 0
;
for x 2 R
n
, and is a bounded domain with the smooth boundary b nd( Ω). Show the following
conservation of energy for the above problem
Z
(juj
2
+ ju
t
j
2
+ jruj
2
) dV =
Z
(jf j
2
+ jrf j
2
) dV :
3.2 Wave equation 19
Exercise 3.25. Consider the following damped equation
8
>
>
>
>
>
>
<
>
>
>
>
>
>
:
u
tt
+ ξu
t
= u
uj
bnd(Ω)
=0
u(x; 0) = f (x)
u
t
(x; 0) = g(x)
;
where ξ > 0 is a constant, and is the unit cube [0; 1]
n
in R
n
.
i. For E(t), the energy defined by
E(t): =
Z
ju
t
j
2
+
Z
jruj
2
;
show that E(t) E(0) for t > 0.
ii. Show the following relation
E(t) E(0) e
¡ξt
3.3 Poisson and Lapl ace equations
The Po isson and Laplace equations are two important partial differential equations that
arise in many areas of science and engineering. The Poisson equation describes the steady-
state behavior of a field, such as the electric potential or temperature, in the presence
of a given source or charge distribution. The Laplace equation is a special case of the
Poi sson equation where the source is zero, and it describes the equilibrium state of the field.
These equations have many applications in physics, engineering, and mathematics, and their
solutions can provide valuable i nsights into the behavior of physical systems. In this chapter,
we will discuss the Poisson and Laplace equations and their solutions in one, two, and three
dimensions.
3.3.1 Derivation of Poisson an d Laplace equations
Let be an open subset of R
n
. The Poisson's equation has the general form ¡u = f . If
is bounded with boundary bnd(Ω), the associated boundary condition can be
αu + β
@u
@n
bnd(Ω)
= g;
where α and β a re constants. I n particular, if f is identically zero, the partial differential
equation is called a Laplace equation.
The Poisson equation c a n be considered as the steady-state solution of a heat problem.
For example, the steady-state solution of the non-homogeneous heat equation
8
>
>
<
>
>
:
u
t
= u + f on
αu + β
@u
@n
= g on bnd(Ω)
u(x; 0) = u
0
(x)
;
is expressed as the equation
(
¡u = f on
αu + β
@u
@n
= g on bnd(Ω)
;
20 Linear Second-Order Equations
which is independent of th e initial condition u
0
.
In the electrostatic context, the Poisson and Laplace equations arise as the potential field
of electric charges. Consider a distribution of electric cha rge with density ρ(r) in R
3
. The
electric field at r 2R
3
generated by this distribution is determined by the following integral:
E(r) =
1
4π"
0
Z
R
3
r~ ¡r~
0
jr~ ¡r~
0
j
3
ρ(r
0
) dr
0
; (3.3)
where "
0
> 0 is the permittivity constant. We can apply Gauss's law, which relates the flux of
electric field through a surface to the charge enclosed by that surface, to a bounded domain
R
3
:
ZZ
bnd(Ω)
E(r) ·ν dS =
Q
"
0
;
where Q i s the total charge enclosed by bn d(Ω):
Q =
ZZZ
ρdV :
Using the divergence theorem, this can be rewritten in terms of the dive rgen ce of the electric
field:
ZZZ
W
div (E) dV =
1
"
0
ZZZ
ρdV :
It turns out that the electric field E is potential in the sense that there is a scalar function
φ such that E = ¡rφ. Accordingly, we obtain the following differential equation for φ:
¡φ(r) =
ρ(r)
"
0
:
Exercise 3.26. Recall that if we p ut a pointwise electric charge q at the origin, it pro duces an electric
field E that bu the Columb law is
E(r) =
q
4π"
0
jrj
2
r^;
for any r 2R
3
¡f0g, where r^ is the unit vector in the direction o f r.
a) Ver ify that for φ = ¡
q
4π"
0
jrj
, we have E = ¡rφ.
b) For any r 2R
3
¡f0g, verify the relation φ(r) = 0.
Exercise 3.27. f the electrical field E(r ) is known, its associated potential φ with the relation E =¡rφ
is derived by the relation
φ(r) =
Z
1
r
E(r
0
) ·T
^
dl;
where T
^
is the unit tangent vector to the straight line connecting 1 to r.
Exercise 3.28. The most well know conservative field in physics is the gravitation field. If a particle
with the mass m is located at the origin in R
3
, then its potential U (x; y; x) is
U = ¡
m
jrj
;
for r 2R
3
¡f0g. Now, suppose that in R
n
; n 3, t he potential function U has the form
U(x
1
; :::; x
n
) = ¡
m
(x
1
2
+ ···+ x
n
2
)
α/2
;
3.3 Poisson and Laplace equations 21
for some constant α. Find α such that U = 0.
Exercise 3.29. Despite the heat and wave equations, the Poisson and Laplace equations with Neumann
bo undary conditions can be tricky and the existence of a solution may fail. Consider the following
equation:
(
¡u = f on
@u
@n
= g on bnd(Ω)
:
a) Use the divergence theorem to show that the necessary condition for the existence of a solution
to the equation is:
I
bnd(Ω)
g =
Z
f:
b) Even if a solution u exists for the equation, show that the solution is not unique and there are
infinitely many other solutions.
c) Let B R
2
be the unit disk centered at the origin. Find α such that the following equation is
solvable.
(
u = 1 + x
2
y
@u
@n
S
= α
;
where S = bnd(B) is the unit circle.
Exercise 3.30. The Maxwell's equations are as follows
r·E =
ρ
"
0
; r·B = 0; r×E = ¡B
t
; r×B = µσE + µ"
0
E
t
;
where E ; B are respectively the electri c and magnetic fields in space, ρ is t he density of the electric
charge, µ is the magnetic inductive capacity and " is the electric inductive capacity and σ is the electrical
conductivity. Use the above equations and derive the following equation
E = µ"
0
E
tt
+ µσE
t
+
1
"
0
rρ
3.3.2 Fundamental solution of Poisson's equation
Consider the equation ¡u = f(r), where r 2R
3
. We can write f as a convolution integral
using the Dirac delta function δ:
f(r) =
Z
R
3
f(r
0
) δ(r ¡r
0
) dV :
This suggests that the solution to the above equation is related to the solution of th e equation
¡u = δ(r). Recall that the potential function φ = ¡
1
4π"
0
jrj
describes the electric field of a
pointwise unit charge a t the origin. We can take u(r) = Φ(r) :=
1
4π jrj
for r 2 R
3
¡f0g, and
although a rigorous proof is beyond the scope of this book, we c a n verify the following:
¡∆Φ = 0 = δ(r);
for r 2R
3
¡f0g. Moreover, for any ball B
a
of radius a centered at 0, we have:
ZZZ
B
a
δ(r) dV = 1:
22 Linear Second-Order Equations
Using the divergence theorem, we can rewrite the left-hand side of the above equation:
¡
ZZZ
B
a
∆ΦdV = ¡
ZZ
bnd(B
a
)
rΦ ·νdS = ¡
ZZ
bnd(B
a
)
r
1
4π j rj
·r^dS ;
where we used the fact that ν = r^ on the surface of the ball B
a
. Using the relation
r
1
4π j rj
= ¡
r^
4π jrj
2
;
and the equality jrj
2
= a
2
on bnd(B
a
), we obtain
¡
ZZ
bnd(B
a
)
r
1
4π j rj
·r^dS =
1
4πa
2
ZZ
bnd(B
a
)
dS = 1:
Therefore,
¡
ZZZ
B
a
∆Φ(r) dV =
ZZZ
B
a
δ(r) dV ;
for any a > 0. We can now obtain the solution to the equation ¡u = f using the convolution
integral:
u(r) =
ZZZ
R
3
f(r
0
) Φ(r ¡r
0
) dV :
Example 3.4. Consider the equation ¡u(r) = jrje
¡jrj
. The solution of the equation is
u(r) =
1
4π
ZZZ
R
3
jr
0
je
¡jr
0
j
jr ¡r
0
j
dV :
The value of u a t the origin r = 0 is equal to
u(0) =
1
4π
ZZZ
R
3
e
¡jr
0
j
dV =
1
4π
lim
R!1
Z
¡π
π
Z
0
π
Z
0
R
e
¡r
0
r
0
2
sin(θ) dr
0
= 2:
For r =/ 0, we can use a numerical integrator to find the value of u(r) by the above integral.
Remark 3.2. While the convolution solution for the Poisson equation works well on R
n
, it
does not always satisfy the desired bou ndary conditions when applied to bounded domains .
In fact, finding a closed-form solution for the Poisson equation on a general b o unded domain
with arbitrary boundary conditions is typically impossible. However, in later chapters, we
will develop tools to solve this equation on specific types of domains such as rectangles, disks,
cylinders, and spheres.
Exercise 3.31. Let f = e
¡jrj
2
. Use a numerical integrator and find the solution of equation ¡u = f
on the sphere jrj= 1.
Exercise 3.32. Let us find the fundamental solut ion o f the Poisson equation in R
2
¡∆Φ = δ(r ):
3.3 Poisson and Laplace equations 23
Assume Φ = Φ(jrj), where jrj= x
2
+ y
2
p
. Calculate Φ and obtain th e following equation for r =/ 0:
d
2
Φ
djrj
2
+
1
jrj
dΦ
djrj
= 0;
Solve the equation and obtain Φ(r) = C lnjpj for some constant C. Find C.
3.3.3 Harmonic functions
Harmonic functions play an important role in many areas of mathematics and physics. A
smooth function u defined on an open set R
n
is c a lle d harmonic if it satisfies the Laplace
equation u(x) = 0 for all x 2 . There are many examples of harmonic functions in R
n
,
including some simple ones that can be obtained using basic calculus. The following exercise
gives a few such examples.
Exercise 3.33. Verify that the following functions are solutions t o the Laplace equation in R
2
:
i. f (x; y) = constant
ii. f (x; y) = ax + by
iii. f(x; y) = x
2
¡ y
2
iv. f (x; y ) = cos (nx) cosh (ny).
Harmonic fu nc tions po ssess several interesting properties. One of these properties is that
the net flux of the gradient vector field F = ru through an arbitrary closed surface S is zero.
Specifically, for any closed surface S that encloses the open se t , we have
0 =
Z
u =
I
S
ru ·ν; (3.4)
where ν is the unit outward norma l vector to S. This property also explains why the Laplace
equation u = 0 corresponds to the steady state of a heat equation u
t
= u: if we interpret
u as the temperature f unction, then the gradient ru represents the heat flux, and since
u is s teady, the net heat flux through any closed surface S must be zero, otherwise the
temperature o f points inside S would change in time. This observation leads to the following
important theorem
Theorem 3.2. A harmonic function u in an open set can not have a proper maximum
and minimum inside .
The theorem expresses a fundamental result in the theory of harmo nic functions. It
states that a smooth function u, which is harmonic in an open set , cannot have a proper
maximum or minimum inside . By proper maximum or minimum, we mean a point x
0
and
a neighborhood D containing x
0
such that u(x
0
) > u(x) or u(x
0
) < u(x) for all x 2D.
To understand why this is true, consider the associated vector field F = r u. As we saw
in t he first chapter, proper maximum points behave like sinks for the vector field, while
minimum points behave like so urces. In either case, the net flux passing through the level
curves arou nd the m a ximum or minimum points is not zero, which violates the relation (3.4).
24 Linear Second-Order Equations
For example, the function u = x y e
¡x
2
¡y
2
can not be a harmonic function. The figure
below illustrates th e associated vector field of the function. Observe that the net flux passing
through the level curves around the maximum and minimum points are non-zero.
The following fact is an immediate result of the above argument:
Theorem 3.3. (Maximum principle) If u is harmonic function in an bounded open
set R
n
and if u is continuous on cl(Ω), then u attains its maximum and minimum on
bnd(Ω).
Exercise 3.34. Let R
n
be a bounded open set with smooth boundary bnd(Ω). Use the maximum
principle and prove that the unique solution of the Laplace equation
u = 0 on
u = a on bnd(Ω)
;
for a a constant, is the constant solution u = a.
Exercise 3.35. Let R
n
be a bounded open set with smooth boundary bnd(Ω). Use the maximum
principle and show that if there is a s olution to the following problem
u = f on
u = g on bnd(Ω)
;
then it is unique.
Exercise 3.36. Let R
n
be a bounded open set with smooth boundary bnd(Ω). Use the maximum
principle and show that if there is a s olution to the following problem
(
¡u = f on
u +
@u
@n
= g on bnd(Ω)
;
then it is unique.
Exercise 3.37. Let be an open bounded set with smooth boundary bnd(Ω). Prove that the necessary
condition that the following problem has a non-trivial sol ution is λ > 0
(
u = ¡λu on
u +
@u
@n
= 0 on bnd(Ω)
:
Exercise 3.38. Let us write the wave equation on a bounded domain as follows
1
c
2
u
tt
= u:
3.3 Poisson and Laplace equations 25
When c !1, the equation approached to the Laplace equation u = 0. Therefore, we can consider the
Laplace equation as a wave pr opagation equation with the wave speed infinity. This justifies t ha t if the
condition on bnd(Ω) changes, this change affects immediately on the value of u inside .
a) Ver ify that the functio n u = 1 + "r
2
sin(2θ) satisfies the Laplace equation
u = 0 on B
u = 1 + " sin(2θ) on bnd(B)
;
where B is the unit disk in R
2
.
b) For " = 0, the unique solution is u = 1. For the boundary condition uj
bnd(B)
=1 + "sin(2θ), the
solution inside B changes immediately to u = 1 + "r
2
sin(2θ). Run the following code in Matlab
and draw the solut ion for " = 0; 0.5; 1.
[th,r]=meshgrid(-pi:pi/20:pi,0:0.2:1);
up=@(e) 1+e*r.^2.*sin(2*th);
[x,y]=pol2cart(th,r,up(0.5));
axis([-1 1 -1 1 0 2]);
subplot(2,2,1)
surf(x,y,up(0));
axis([-1 1 -1 1 0 2]);
title('$\epsilon=0$','interpreter','latex','fontsize',12);
subplot(2,2,2)
surf(x,y,up(0.5));
axis([-1 1 -1 1 0 2]);
title('$\epsilon=0.5$','interpreter','latex','fontsize',12);
subplot(2,2,3)
surf(x,y,up(1));
axis([-1 1 -1 1 0 2]);
title('$\epsilon=1$','interpreter','latex','fontsize',12)
Another consequence of the relation (3.4) is the Mean Value Property of harmonic f unc-
tion. In its rigorous form, it states that i f u is a harmonic function on an open set in R
n
,
and B
r
(x) denotes the open ball of radius r centered at x, and S
r
(x) denotes the sphere
bnd(B
r
(x)), then for any x 2 and r > 0 such that B
r
(x) , we have:
u(x) =
1
jS
r
j
I
S
r
(x)
u(z) dS;
where jS
r
j denotes the volume of the ball sphere S
r
. In other words, the value of a harmonic
function at any point is the average of its values over any sphere cente red at that point.
In the case of harmonic functions on R
2
, the Mean Va lue Property can be expressed
as follows: if u is a harmonic function on an open set , and B
r
(z
0
) denotes the open disk
of radius r centered at z
0
= (x
0
; y
0
) and contained in , then
u(z
0
) =
1
2πr
I
C
r
(z
0
)
u(z) dz;
where C
r
(z
0
) denotes the circle of radius r centered at z
0
, oriented counterclock wise. In other
words, the value of a harmonic function at any point is the average of its values over any
circle centered at that point.
To see why this is true, suppose u is harmonic in an open set R
2
and consider two
circles centered at z
0
=(x
0
; y
0
) with radii r and r + " inside . For an arbit rary point z on
C
r
, the ci rcle of radius r, consider the a ssociated point z + " r^ on C
r+ "
.
26 Linear Second-Order Equations
z
z + "r^
r + "
r
We can calculate the limit:
lim
"!0
1
"
1
r + "
I
C
r+"
(z
0
)
u(z) dl ¡
1
r
I
C
r
(z
0
)
u(z) dl
:
Using polar coordinates for both integrals, we ob tain the expression in the bracket as:
1
r + "
I
C
r+"
(z
0
)
u(z) dl ¡
1
r
I
C
r(z
0
)
u(z) dl = "
Z
¡π
π
ru(jξ j; θ) ·r^;
where ξ is a point in the line connecting z to z + " r^. T herefore, we obtain:
lim
"!0
1
"
1
r + "
I
C
r+"
(z
0
)
u(z) dl ¡
1
r
I
C
r
(z
0
)
u(z) dl
= lim
"!0
Z
¡π
π
ru(jξj; θ) ·r^;
and if we pass the limit inside the integral, we obtain
lim
"!0
Z
¡π
π
ru(jξj; θ) ·r^ =
Z
¡π
π
ru(r; θ) ·r^ = 0
where the last equality is justified by the relation (3.4). This calculation justifies the fact that:
d
dr
1
r
I
C
r
(z
0
)
u(z) dl = 0;
and then the integral
1
r
I
C
r
(z
0
)
u(z) dl
is a constant. In particular, if r !0, we obtain
lim
r!0
1
r
I
C
r
(z
0
)
u(z) dl = u(z
0
) lim
r!0
1
r
I
C
r
(z
0
)
dl = 2πu(z
0
);
and then
u(x
0
; y
0
) =
1
2πr
I
C
r
(z
0
)
u(z)dl:
This is the first version of the mean value property of ha rmonic functions in 2D.
Example 3.5. Let us verify the mean value property for the harmonic function u = x
2
¡y
2
.
We have
1
2πr
Z
¡π
π
δ
2
(cos
2
θ ¡sin
2
θ) =
r
2π
Z
¡π
π
cos(2θ) = 0;
3.3 Poisson and Laplace equations 27
and the obtained value is equal to u(0; 0). Now, consider the arbitrary point z
0
: (x
0
; y
0
). The
circle C of radius r centered at z
0
has the following representation in the polar coordinate
x = x
0
+ r cosθ; y = y
0
+ r sinθ:
Hence,
1
2πr
I
C
r
(x
2
¡y
2
)dl =
1
2πr
Z
¡π
π
[(x
0
+ r cosθ)
2
¡(y
0
+ r sinθ)
2
] = x
0
2
¡y
0
2
= u(z
0
):
Exercise 3 .39. Verify that the function u = x
3
+ y
3
¡3x
2
y ¡3xy
2
is harmonic and calculate the following
integral
I
C
u(x; y) dl;
where C is the circle of radius 1 centered at the point (1; 1).
Exercise 3.40. We prove another version of Mean Value Property for harmonic functions in 2D
u(x
0
; y
0
) =
1
πr
2
ZZ
B
r
(x
0
;y
0
)
u(x; y) dxdy:
For this, use the polar coordinate a nd write
ZZ
B
r
(x
0
;y
0
)
u(x; y) dxdy =
Z
0
r
Z
¡π
π
u(r; θ) rdr:
Then, use the first version of the mean value property and conclude the claim.
Exercise 3.41. Use the spherical coordinate and prove the first and second versions of the Mean Value
Proper ty for harmonic functions in 3D.
3.4 Cl as sification of linear second-order PD Es
The differential equations we have explored thus far are instances of linear second-order
equations. Understanding the behav ior of their solutions is a significant aspect of studying
these equations, both qualitatively and quantitatively. In order to analyze them effectively,
it is helpful to classify linear second-order partial differential equations into three main
categories: elliptic, parabolic, and hyperbolic equations. This classification is based on the
properties of the principal part of the d ifferential operator near a specific point, which in
turn determines the characteristic beh avior of the solution in the vicinity of that point.
3.4.1 Operator form, li nearity and superposition
A useful perspective on linear differential equations is to view them as operator equations.
When dealing with a real-valued function u(x) defined on a domain R
n
, the partial deriv-
ative
@u
@x
j
can be seen as a mapping @
j
that takes functions from the domain of continuously
differentiable functions and returns continuous functions:
@
j
: C
1
(Ω) !C(Ω):
28 Linear Second-Order Equations
Here, C
1
represents the set of continuously differentiable functions, and C represents the set
of continuous functions.
By considering
@u
@x
j
as an o perator @
j
, we can analyze how it acts on functions and how
it re lates to other operators in the differential equation. This viewpoint allows us to study
the properties of the operator, such a s linearity, compo sition, and its effect on the solutions
of the differential equation. Similarly, we can extend the notion of differenti a l operators to
higher order derivatives. For instance, the se cond partial derivative operator can be denoted
as @
ij
and maps functions from the domain of twice continuously differentiable functions:
@
ij
: C
2
(Ω) !C(Ω):
Furthermore, we can define a complete linear second-order differential operator as:
L :=
X
i;j =1
n
a
ij
(x) @
ij
+
X
j=1
n
b
j
(x) @
j
+ c(x); (3.5)
where a
i; j
(x); b
j
(x) and c(x) are some continuous or smooth functions defined on the d omain
. This general form allows us to represent a wide range of second-order linear differential
equations.
Recall from linear algebra that a mapping T from R
n
to R
m
is called linear if it satisfies
the following relation for any α
1
; α
2
2R and u
1
; u
2
2U:
T (α
1
v~
1
+ α
2
v~
2
) = α
1
T (v~
1
) + α
2
T (v~
2
):
Similarly, we can apply this concept to the differential operator L. The operator L is linear
because it satisfies the following property:
L(α
1
u
1
(x) + α
2
u
2
(x)) = α
1
T (u
1
(x)) + α
2
T (u
2
(x));
where u
1
(x) and u
2
(x) are functions in the domain of the operator L, and α
1
and α
2
are
constants.
Although we wo n't delve into the full vector space formulation of partial differential
operators in this book, representing partial de rivatives in o perator notation can help us
understand the classification of second-order linear PDEs. In general, a linear second-order
partial differential equation can be written as:
L[u]= f ;
where L is the ope rator defined in (3.5). Here, L represents a linear mapping that takes a
function u as input and produces the corresponding function f as output. The goal of such
an equation is to determine the function u, given a known function f.
Interpreting a linear partial differential equation as an operator equation has several
advantages. I t allows us to make a connection between linear algebra, which studies the
properties of linear mappings in finite-dimensional vector spaces, and differential equations.
One benefit of this approach is the superposition principle, which applies to homogeneous
equations of the form L[u] = 0.
3.4 Classification of linear second-order PD Es 29
Proposition 3.1. If u
1
; :::; u
N
are homogeneous solutions of linear equation L[u] = 0, then
the function
u =
X
j=1
N
c
j
u
j
;
for arbitrary c
j
2 R solves the equation. Moreover, if u
p
is a particular solution for the
equation L[u] = f, then the function
u = u
p
+
X
j=1
N
c
j
u
j
;
solves the equation if u
1
; ::: ; u
N
are in the kernel or null space of L.
Exercise 3.42. Let L be the linear operator L = @
x
+ @
y
. Consider the equation L[u] = x + y.
a) Find the general homogeneous solution of the equation.
b) Verify that u
p
= xy is a particular solution of the equation. Therefore, u = u
h
+ xy is the general
solution of the equation. It is simply verified that u
p
=
1
2
x
2
+
1
2
y
2
is also a particular solution.
Show that the latter particular solution can be derived from the general solution u = u
h
+ xy.
Exercise 3.43. Find the gener al solution of the linear equation L[u] = x for u = u(x; y), where
L = @
t
¡c@
x
for a constant c.
Exercise 3.44. Consider the following equation
u
tt
¡c
2
u
xx
= 0;
and let L be the oper ator L : = @
tt
¡c
2
@
xx
. Show that
L = L
1
L
2
;
where L
1
:= @
t
¡c@
x
, L
2
: =@
t
+ c@
x
. With the aid of the the decomposition, find two linearly independent
solution of the given equation.
By taking some precautions, we can also extend the superposition principle to the infinite
or continuous versions. For example, if u
1
; u
2
; ::: are in the null space of L, then L[u] = 0 for
the function
u =
X
j=1
1
c
j
u
j
:
Similarly, if u(x; k) solves the equation L[u] = 0 for k in some interval I, then the following
function also solves the equation:
u =
Z
I
c(k) u(x; k) dk:
Example 3.6. Consider the partial differential equation u =0 defined on the half-plane x >
0. An example of a solution to this equation is give n by the functions u
k
(x; y) = e
¡kx
sin(ky)
for any k 2(¡1; 1). Now, suppose we define c(k) as follows:
c(k) =
0 k < 0
1 k > 0
:
Then, the function
u(x; y) :=
Z
¡1
1
c(k) u
k
(x; y) dk =
y
x
2
+ y
2
;
30 Linear Second-Order Equations
solves the equation u =0 on x > 0. However, as x approach es 0, the function u(x; y) becomes
unbounded and cannot be extended to the boundary of the domain. To remedy this, we can
choose c(k) as
c(k) =
(
0 k < 0
e
¡k
k > 0
:
This transforms the superposition integral to
u(x; y) =
y
(x + 1)
2
+ y
2
:
This solution can be extended to the boundary x = 0 in a proper manner.
Remark 3.3. The version o f infinite superposition should be used with caution. For instance,
let us consider the equation L =
d
2
dx
2
+ 1 and the functions u
n
(x) =
1
n
2
cos(nx) for any
o dd n, which solve the equation. The f unction
u =
X
n=1;n:odd
1
u
n
(x) = cos(x) +
1
4
cos(2x) +
1
9
cos(3x) + ···+
converges tof(x)=
π
2
8
¡
π
4
jxj for x 2(¡π; π). However, this function is discontinuous and does
not have a derivative at x = 0. Therefore, we cannot define L [u ] as equal to zero at x = 0,
and we must be careful when using the infinite version of superposition.
Exercise 3.45. Find a solution to equation u = 0 on the half-plane x > 0 such that u on the boundary
x = 0 is equal to
y
4 + y
2
.
Exercise 3.46. Consider linear equation @
xx
u = xy, for u = u(x; y).
a) Show that u
p
=
1
6
x
3
y solves the equation.
b) Find the homogeneous solution of the equation and write down the general solution.
Exercise 3.47. Find the gener al s olution of the following equation
L[u] = xy;
for a function u = u(x; y), where L = @
xy
.
Exercise 3.48. We know that the linear equations enjoy superposition principle. This property does
not necessarily hold for non-linear equations. Verify that functions u
1
= x + c
1
and u
2
= y + c
2
are the
solutions to the nonlinear equation
ju
x
j
2
+ ju
y
j
2
= 1;
however, u
1
+ u
2
is not a solution.
3.4.2 Classification: c onstant coefficient case
Linear seco nd-order partial differential equations can be classified based on their principal
part, which is defined by the sum:
L
p
:=
X
i;j
a
ij
(x) @
ij
:
3.4 Classification of linear second-order PD Es 31
This classification is based on the idea that the behavior of a PDE near a given poi nt is
determined by the highest-order derivatives app earing in the equation at that point. To
illustrate the logic behind the classification, we first consider the case where a
ij
are constants,
and the domain of the problem is open subsets of the plane.
It is simply seen that L
p
can be rewritten in the matrix form as follows:
L
p
:= (
@
1
@
2
) A
@
1
@
2
;
where A is the symmetric coefficient matrix given by
A =
0
@
a
11
a
12
+ a
21
2
a
12
+ a
21
2
a
22
1
A
:
From linear algebra, the eigenvalues of A are real. Let λ
1
and λ
2
be the eigenvalues of A.
Based on these eigenvalues, we can classify the PDE as follows:
Elliptic: λ
1
λ
2
>0.
Solutions to elliptic PDEs are smooth and have no s ingu larities. This type of
PDE typically arises in problems involving steady-state phenomena, such as
in electrostatics or flu id mechanics. The most familiar example is the Poisson
equation ¡u = f, where operator can be written as
= (
@
1
@
2
)
1 0
0 1
@
1
@
2
;
with the coefficient matrix A =
1 0
0 1
having eigenvalues λ
1
= λ
2
= 1.
Parabolic: λ
1
λ
2
=0.
Solutions to parabolic PDEs have a smooth ing effect that can regularize initial
data. This type o f PDE often arises in problems involving diffusion, heat
flow, or time-dependent phenomena that evolve in a certain direction. The
most fa miliar equation is the heat equation u
t
= ku
xx
in the plane (x; t). The
coefficient matrix is
A =
0 0
0 k
;
withe eigenvalues λ
1
= 0, λ
2
= k.
Hyperbolic: λ
1
λ
2
<0.
Solutions to hyperbolic PDEs have wave-like behavior and can exhibit shock
waves and other d iscontinuities. This type of PDE typically arises in problems
involving wave propagation or vibrations. The most familiar example is a wave
equation u
tt
= c
2
u
xx
in the (x; t)-plane. The coefficient matrix of this equation
is
A =
¡1 0
0 c
2
!
;
32 Linear Second-Order Equations
with eigenvalues λ
1
= ¡1, λ
2
= c
2
.
Remark 3.4. It is known from linear algebra that λ
1
λ
2
= det(A), and thus the classification
can be done based on the sign of det(A) too. Therefore, we can classify the P DE based on
the sign of det(A), as follows:
Elliptic: det(A)>0
Parabolic: det(A)=0
Hyperbolic: det(A)<0
Example 3.7. Consider the following equation
2u
x x
¡αu
xy
+ 2u
yy
+ 3u
x
= u + x:
We classify the type of the equation based on the value of α. The coefficient matrix of the
principal part of the equation is
A =
2
α
2
α
2
2
!
:
We have det(A) = 4 ¡
α
2
4
, and thus the equation is elliptic if α
2
< 16 or ¡4 < α < 4. The
equation is hyperbolic if α > 4 or α < ¡4. The equation is parabolic if α = 4 or ¡4.
Exercise 3.49. Classify following equations
a) u
xx
+ u
yy
+ 2u
xy
+ u
y
= 1 ¡u
b) u
xx
+ 2u
yy
¡3u
xy
= u.
Exercise 3.50. Determine the type of the following equation based on the values of α
u
xx
+ u
yy
+ 2αu
xy
= 0:
Definition 3.1. Let L
p
be the principal part of a linear second-order partial differential
equation (PDE), given by
L
p
=
X
i;j=1
n
a
ij
@
ij
;
where a
ij
are constant coefficients. The associated coefficient matrix of the operator is defined
by A = [a~
ij
] where a~
ij
=
a
ij
+ a
ji
2
. Let λ
1
; :::; λ
n
be the eigenvalues of A.
The PDE is called elliptic if all eigenvalues λ
1
; :::; λ
n
have the same sign.
The PDE is called hyperbolic if one eigenvalue has the opposite sign of the other
eigenvalues.
The PDE is called parabolic if one eigenvalue is zero, and the other eigenvalues have
the same sign.
Exercise 3.51. For function u = u(x; y; t), classify the equations
a) u
t
= ku,
3.4 Classification of linear second-order PD Es 33
b) u
tt
= c
2
u.
c) u
xt
+ u
yy
+ u
xx
+ 3u
x
= 0.
Exercise 3.52. Find α such that the following equation is parabolic in the coordinate (t; x; y):
u
t
= 2u
xx
+ αu
xy
+ 2u
yy
:
3.4.3 Canonic al form
For the principal part L
p
, the terms @
ij
for i =/ j are called multiplicative terms. We can
remove these terms by performing a lin ear transformation of the co ordinate. Before intro-
ducing the transformation, let's explain the terminology of ellipticity, parabolicity, and
hyberbolicity by comparing them to the types of conic sections.
Recall that the equation of a conic section is given by:
ax
2
+ 2bxy + cy
2
+ dx + ey = const:
The principal part of this section is ax
2
+ 2bxy + cy
2
, which ca n be represented in matrix
form as
ax
2
+ 2bxy + cy
2
= (
x y
) A
x
y
;
where th e c oefficient matrix A is symmetric and has real eigenvalues λ
1
; λ
2
, and two orthog-
onal eigenvectors. Let Q = [v~
1
jv~
2
] be the eigenvector matrix of A with jv~
1
j= jv~
2
j= 1.
By linear algebra, we know that:
Q
¡1
AQ =
λ
1
0
0 λ
2
:
Thus, the transformation
x
y
= Q
X
Y
;
maps the old coordinate system to the new coordinate system (X ; Y ), and transforms the
conic section to the following form:
λ
1
X
2
+ λ
2
Y
2
= first and zero order terms:
If λ
1
and λ
2
have the same sign, the conic section is an ellipse. If one of λ
1
or λ
2
is zero,
the conic section is a parabola. If λ
1
and λ
2
have opposite signs, then the conic section is a
hyperbola.
This argument shows how we can remove the multiplicative terms and why the termi-
nology is justified based on geometrical observation.
Example 3.8. Consider the equation
2x
2
+ 2y
2
¡2xy = 1:
In the matrix form, we can represent the equation as
(
x y
)
2 ¡1
¡1 2
x
y
= 1:
34 Linear Second-Order Equations
The coeffi cient matrix A =
2 ¡1
¡1 2
has two eigenvalues λ = 1; 3 with associated eigenvectors
v~
1
=
0
B
B
@
2
p
2
2
p
2
1
C
C
A
, v~
2
=
0
B
B
@
¡
2
p
2
2
p
2
1
C
C
A
. Consider the coordinate transformation
Q
X
Y
=
x
y
; (3.6)
where Q is the matrix of eigenvectors. By performing this transformation to the equation,
we obtain
(
x y
)
2 ¡1
¡1 2
x
y
= (
X Y
)Q
t
2 ¡1
¡1 2
Q
X
Y
= (
X Y
)
1 0
0 3
X
Y
:
Thus, th e given equation in the new coo rdinate system has the representation X
2
+ 3Y
2
= 1
which is a ellipse. Note that Q is a 45 degree rotation m a trix, and thus the coord inate change
is just a rotation of the old coordinate system.
-0.5 0.5
-0.5
0.5
-1 -0.5 0.5 1
We can employ the above technique for linear partial differential equation.
Example 3.9. Let us rewrite the following equation in the canonical form
2u
xx
+ 2u
yy
¡2u
xy
+ 3u
x
= u + x:
The coefficient matrix of the principal part is A =
2 ¡1
¡1 2
, with the eigenvalues λ
1
= 1,
and λ
2
= 3. The eigenvectors of the coefficient matrix A are v~
1
=
1
2
p
1
1
, v~
2
=
1
2
p
¡1
1
, and
thus Q =
1
2
p
1 ¡1
1 1
is the transformation matrix. Define the new coordinate system @
X
; @
Y
through the following transformation
@
x
@
y
= Q
@
X
@
Y
: (3.7)
By performing this transformation to the principal part of the equation, we reach
2u
x x
+ 2u
yy
¡2u
x y
= u
XX
+ 3u
YY
:
We need also to transform the first and zero terms of the equation as well. From (3.7), we
have
@
x
=
2
p
2
@
X
¡
2
p
2
@
Y
;
3.4 Classification of linear second-order PD Es 35
and x =
2
p
2
X ¡
2
p
2
Y . Finally, the given PDE reduces to the following one in the new
coordinate system
u
XX
+ 3u
YY
+
3 2
p
2
u
X
¡
3 2
p
2
u
Y
= u +
2
p
2
X ¡
2
p
2
Y :
Exercise 3.53. Wr ite down the canonical form of the following equation
u
xx
+ u
yy
+ 4u
xy
+ 2
p
u
y
= u
Exercise 3.54. Find values of α s uch that the equation
u
xx
+ 3u
yy
+ 2αu
xy
+ 2u
x
¡3u
y
= 0;
is elliptic, parabo lic or hyperbolic. For α = 1 write down the equation in the new coordinate and
determine the associated linear transformation.
Exercise 3.55. Write the following equations in the canonical form and classify them into elliptic,
parabolic and hyperbolic equations
a) u
xy
= 0
b) u
xx
+ 2u
xy
= u
c) u
xx
¡u
yy
¡2u
xy
= u
x
+ u
d) 3u
xx
+ 4u
xy
+ 6u
yy
= u
x
¡u
e) u
xx
¡4u
xy
+ 4u
yy
¡u
x
= 1
f) ¡2u
xx
¡6u
xy
+ 6u
yy
¡u
x
= 0
g) u
xx
+ 4u
xy
+ u
yy
+ 2
p
u
y
= u
Exercise 3.56. Consider the following equation
u
tt
+ u
xx
+ 4u
tx
= 2
p
x:
a) Determine the type of equation in terms of elliptic, parabolic o r hyperbolic equation. Your answer
should be based on a calculation.
b) Use a change of coordinate from (t; x) to ( T ; X) and rewrite the equation in the normal form
without any multiplicative partial differentiation.
c) For the new equation in the coordinate (T ; X), find a solution in the form u(T ; X) = αX
a
+ βT
b
for suitable constants α; a; β; b.
Exercise 3.57. Consider the following equation
2u
xx
+ 3u
yy
+ 2 2
p
u
xy
= 0
a) Determine the type of the equation in terms of ellipticity, parabolicity and hyberbolicity.
b) Write the canonical form of the equation with the aid of a transformation of the type
Q
@
x
@
y
=
@
X
@
Y
;
for an appropriate matrix Q.
c) In the new coordinate (X ; Y ), use the transformation ξ = Y + i2X, η = Y ¡i2X, and derive the
following equation in (ξ; η)
u
ξη
= 0:
Find the general solution of the equation in ( X ; Y ).
36 Linear Second-Order Equations
3.4.4 Classification: variable coefficients
The classification of linear second-order PDEs with variable coefficients is similar to that of
equations with constant coefficients. However, for variable c oefficient equations, the classi-
fication is loca l, meaning it holds only in an open neighborhood of the focal point, and not
for the entire domain. The following example illustrates this fact.
Example 3.10. Consider the differential operator
L := (1 + x
2
y)@
xx
¡2xy@
xy
+ y@
yy
:
The coefficient matrix of L
p
is
A(x; y) =
1 + x
2
y ¡xy
¡xy y
!
:
Note that det(A)= y. Thu s, L is elliptic in the half plane y >0 and hype rbolic in the half
plane y <0. For a ny point on th e x-axis, the e quation exhibits degenerate behavior.
Exercise 3.58. Consider the differ ential operator
L = x@
xx
+ xy @
xy
+ y@
yy
:
Rewrite L in the form
L =div ([a
ij
(x; y)] r) + fir st term derivatives
where [a
ij
] is a symmetric matri x.
Problem 3. 1 . Find domains in which the following operators can be classified into elliptic, parabolic
or hyperbolic typ e
i. L := y@
xx
¡y@
yy
+ 2 (x ¡1)@
xy
ii. L := cos(xy)@
xx
+ cos(xy)@
yy
+ 2 sin(xy)@
xy
:
iii. L := x@
xx
+ y@
yy
+ 2xy@
xy
iv. L := x
2
+ y
2
p
@
xx
+ x
2
+ y
2
p
@
yy
+ 2@
xy
.
3.4 Classification of linear second-order PD Es 37