PDE Lecture Notes
PDE Lecture Notes
1.1 Definitions
The simplest differential equation is perhaps the Malthusian equation modelling population growth:
du
= ku, (1.1.1)
dt
where k is a constant, u is the unknown function (representing population size) and t is the independent
variable. A differential equation involving only one independent variable is called ordinary differential
equation(ODE). So (1.1.1) is such. An example of a different type of differential equations is the heat
equation
∂u ∂2u ∂2u ∂2u
= k( 2 + 2 + 2 ) (1.1.2)
∂t ∂x ∂y ∂z
where k is the heat conductivity constant, u is the unknown function (representing temperature) and x,
y, z and t are the independent variables. This is an example of a partial differential equation.
∂u ∂u ∂u ∂ 2 u ∂ 2 u ∂2u
F (x, u, , ,···, , 2 , 2 , · · · , 2 , · · ·) = 0, (1.1.3)
∂x1 ∂x2 ∂xn ∂x1 ∂x2 ∂xn
or in the operator form:
L[u](x) = f (x), (1.1.4)
where x = (x1 , x2 , · · · , xn ) are the independent variables, u = u(x1 , x2 , · · · , xn ) the unknown function,
f = f (x) = f (x1 , x2 , · · · , xn ) the given function. It is agreed that L[u] includes all the terms in F that
involve u.
i.e. the linear combination of two (hence any number of ) solutions is also a solution of the PDE. This is
called the superposition principle for linear PDEs, and does not hold for non-linear ones.
More examples
∂u
where ux denotes the partial derivative
, etc.
∂x
General solutions, initial and boundary conditions
For ODEs, we talked about their general solutions; for example, the general solution, i.e. the set of all
solutions of (1.1.1) is
u(t) = const. Cekt (1.1.6).
For PDEs, we can do the same. For example, consider the transport equation
ut + V ux = 0, (1.1.7)
where u represents the density of a substance (e.g. a pollutant) in a one dimensional medium (e.g. a river)
that moves with constant velocity V . Let
where φ is an arbitrary smooth function. We check that this is always a solution of (1.1.7):
ut + V ux = −V φ0 (x − V t) + V φ0 (x − V t) = 0.
On the other hand, we will show in the next chapter that given any solution u of (1.1.7), there exists a
function φ such that (1.1.7) holds. Thus the general solution of (1.1.7) is given by (1.1.8). Observe that
in the ODE general solution (1.1.6) there is an arbitrary constant C, while in the PDE general solution
(1.1.8) there is an arbitrary function. This is typical for PDEs.
As in the case of ODEs, to pick out a unique solution from a family of solutions, we need initial and
boundary conditions that are mostly physically motivated/demanded. For (1.1.7), if the initial condition
1
u(x, 0) = , x ∈ (−∞, ∞) (1.1.9)
1 + x2
is given (physically the initial density is given), then
1
φ(x) = , x ∈ (−∞, ∞) (1.1.10)
1 + x2
and so u can be uniquely determined
1
u(x, t) = , (x, t) ∈ (−∞, ∞) × [0, ∞).
1 + (x − V t)2
If the “river”(medium) runs from x = 0 to x = ∞, then to determine a unique solution of (1.1.7) we need
not only the initial condition (1.1.9) for x ∈ [0, ∞), but also a boundary condition at x = 0 (the origin
of the river) such as
ξ 2 + η 2 = 1, (1.2.2)
which is the equation of an ellipse. ((1.2.2) can be obtained by taking Fourier transform of both sides
with f being a δ−function, and then dropping the Fourier transform of u.) In this connection, we say
that the Laplace equation (1.2.1) is elliptic. In the same fashion, the wave equation
utt − uxx = f
ξ2 − η2 = 1
hence we say the wave equation is hyperbolic. For the heat equation
ut − uxx = f,
if we take Laplace transform in the t-variable and then Fourier transform in the x-variable, we then obtain
the equation of a parabola
ξ = −η 2 .
Thus the heat equation is said to be parabolic.
It turns out that all 2nd-order linear PDEs
can be classified into three types: elliptic, parabolic and hyperbolic. Here the coefficients a11 , a12 , a22 ,
b1 , b2 , c and the righthand side f are functions of the variables x1 and x2 in general. (Note ux1 x2 = ux2 x1
for smooth u and so there is no need to include a ux2 x1 -term in (1.2.3).) Before discussing how to do the
4
classification, we mention the significance of classifying PDEs: it can be proved by using advanced PDE
theory that solutions of an elliptic equation are as smooth as allowed by the coefficients and the righthand
side of the equation; solutions of a parabolic equation are smoothed right after the initial time, while the
smoothness of solutions of a hyperbolic equation is neither improved nor destroyed after the initial time.
Other differences are: second order elliptic and parabolic equations satisfy the maximum principle; the
speed of propagation of disturbance is finite for hyperbolic equations, infinite for parabolic equations.
Now let us come back to the question of classifying (1.2.3). The type of this equation is a local property
so we freeze the coefficients at a fixed point on the x1 x2 −plane (this is related to the “freezing coefficient
method”in the regularity theory for elliptic and parabolic equations). So we now assume the coefficients in
(1.2.3) are constants. The type of this PDE does not depend on the lower order terms and the righthand
side. Thus we focus on the second order terms.
We determine the type of (1.2.3) by checking to see if there exists a linear change of variables
2
X
yi = bij xj , i = 1, 2,
j=1
so that in y1 , y2 variables, the highest order terms of (1.2.3) are the same as that of the Laplace, or the
wave or the heat equation. By chain rule, we have
2
X
uxk xl = uyi yj bik bjl .
i,j=1
Since ux1 x2 = ux2 x1 , we can let a21 = a12 and rewrite the leading terms of (1.2.3) as
X 2 2
X X2
uxk xl akl = akl bik bjl uyi yj . (1.2.4)
k,l=1 i,j=1 k,l=1
The coefficient of uyi yj on the righthand side is just the ij−element of matrix BAB T , where A = (aij )
and B = (bij ). Since A is symmetric, there exists an orthogonal matrix B such that BAB T is diagonal:
BAB T = diag(λ1 , λ2 )
where the λ’s are the eigenvalues of the matrix A. Then in y1 , y2 variables, the leading terms of (1.2.3)
are written as
λ1 uy1 y1 + λ2 uy2 y2 .
Case 1. det A > 0, i.e, a11 a22 > a212 . In this case the λ’s have the same sign (recall det A = λ1 λ2 ); without
√ √
loss of the generality, assume that both are positive. If we let z1 = y1 / λ1 and z2 = y2 / λ2 , then the
leading terms of (1.2.3) can be further reduced to
uz1 z1 + uz2 z2 .
Thus in Case 1, we say (1.2.3) is elliptic, the same type as the Laplace equation.
Case 2. det A < 0, i.e, a11 a22 < a212 . In this case the λ’s have the opposite signs. We say that (1.2.3) is
hyperbolic, the same type as the wave equation.
Case 3. det A = 0, i.e, a11 a22 = a212 . In this case one of the λ’s is 0, the other is not (if both are 0, then
A = 0 and (1.2.3) is not a second order PDE). We say that (1.2.3) is parabolic, the same type as the
heat equation.
If the coefficients of the leading order terms in (1.2.3) are not constant, then we define the type of (1.2.3)
at a point P on x1 x2 −plane, as above.
5
Assignment 1
1. For each of the PDEs below, find its order, linearity and homogeneity.
(1) ut + uux = 0 (Burger’s equation).
(2) xut − uxx + 2x + sin t = 0 (Degenerate heat equation).
(3) utt − (uxx + uyy + uzz ) = −u + u3 (Klein-Gordon equation).
(4) (1 + u2y )uxx − 2ux uy uxy + (1 + u2x )uyy = 0 (Minimal surface equation).
Show that the only ones that are unchanged under all axis-rotations (rotation invariant) have the form
a · (uxx + uyy ) + b u = 0,
where a and b are constants. Hint: Use the discussion in the text, especially (1.2.4)
4. Classify the following equations as hyperbolic, parabolic or elliptic. If the type changes in the xy−plane,
find the region for each type.
(1) (1 + x2 ) uxx + (1 + y 2 ) uyy + x ux + y uy = 0
(2) uxx + (1 + y)2 uyy = 0
(3) e2x uxx + 2ex+y uxy + e2y uyy = 0
(4) uxx + y uyy = 0
(5) uxx + xy uyy = 0
(6) y uxx − x uyy + ux + y uy = 0
6
5. Show by direct substitution that u(x, t) = f (x + 2t) + g(x − 2t) is a solution of the PDE
utt − 4uxx = 0
where A, B and C are constants. Find the general solution of the above equation when
(1) equation (∗) is hyperbolic;
(2) equation (∗) is parabolic.
Perhaps the most important first order PDE is the transport equation, also called continuity equation.
It can be used to describe the evolution of distribution of mass, energy, electric charges and biological
population, etc. Thus this equation is fundamental to applied sciences.
Here is the general idea in the derivation of the transport equation. Let x represent location in Rn
(n = 1, 2, 3 are the most relevant cases) and let t represent time. Consider a substance that moves in
Rn ; this substance may be a pollutant in a river, population of a biological species, even thermal energy.
Let u(x, t) be the density function of the substance, measured in mass/volume (we emphasize that here
“mass”may mean energy); suppose the velocity vector of the particle of the substance which is at location
x at time t is V(x, t). In multi-variate Calculus we learned that given a surface S with unit normal vector
field n, the rate at which mass crosses the surface in the direction of normal n is given by the “flux integral”
Z
u(x, t)V · n(x, t)dS. (2.1.1)
S
Note that the mass may flow across one part of the surface in the direction of n, and across the other
part in the opposite direction. Thus (2.1.1) is really the net rate at which mass crosses the surface in
the direction of normal n. The substance may be created or degraded so we assume that the creation-
degradation rate is f (x, t). f (x, t) is calculated by taking a small neighborhood of point x with volume
∆V , then dividing the rate of change of mass in the neighborhood by ∆V and sending ∆V to zero. Thus
the unit of f is mass/time/volume.
Now take an arbitrary bounded region Ω in Rn with smooth boundary ∂Ω. By conservation of mass, the
net rate of change of mass in Ω = net rate at which mass crosses ∂Ω in the direction of the inner normal
+ total creation-degradation rate in Ω. Thus we have
Z Z Z
d
u(x, t)dx = u(x, t)V · (−n)(x, t)dS + f (x, t)dx, (2.1.2)
dt Ω ∂Ω Ω
where n is the unit outer normal vector field of ∂Ω. By Divergence Theorem, the surface integral in the
above equation is equal to Z
− ∇ · (uV)dx,
Ω
and so (2.1.2) can be rewritten as
Z
(ut + ∇ · (uV) − f )(x, t)dx = 0.
Ω
7
If we assume the integrand is continuous, then it must be identically equal to 0: otherwise, there exists a
small ball B on which the integrand is positive or negative, so if we take Ω to be B then the integral is
non-zero; a contradiction!
Now we have derived the transport equation
ut + ∇ · (uV) = f. (2.1.3)
We introduce
J = uV
which is called flux. Then (2.1.3) is rewritten as
ut + ∇ · J = f. (2.1.4)
Traffic equation
This is a specific example of the transport equation. Of concern is the traffic problem on one straight lane:
- x
direction of traffic
We use the continuum hypothesis to describe the traffic flow. Let V (x, t) be the velocity of traffic at
position x and time t ( unit : length/time ). The traffic density at time t is defined by
(In (2.1.1), we take the “surface”to be the point x and the normal n to be 1 (one dimensional vector), so
the flux integral is simply the flux J(x, t).) The flux J(x, t) is the rate at which cars cross point x (in the
positive direction) at time t. Now by the general transport equation (2.1.4), we have
∂u ∂J
+ =0
∂t ∂x
or,
∂u ∂
+ (u · V ) = 0,
∂t ∂x
which is called the conservation law or conservation equation for the traffic problem.
Example 2.2.1 Let us start with the simplest transport/traffic equation, already discussed in Section 1.1:
∂u ∂u
+V =0
∂t ∂x (2.2.1)
u(x, 0) = φ(x),
where V is a constant. Recall that in Section 1.1, we announced that the solution is
The idea used to solve this problem is to reduce the PDE to an ODE, by restricting the solution u(x, t)
to a curve on the xt−plane so that u now is a function of one variable and satisfies an ODE. The curve
is called characteristic curve of the PDE. This is called the method of characteristics. Let us carry
out the plan now. Consider a curve x = x(t); after restricting u on the curve it becomes a function of t
only (the function u(x(t), t)). Then by chain rule, we have
du dx
= ut + ux . (2.2.3)
dt dt
We wish to relate the righthand side to the PDE, so we demand that
dx
= V.
dt
Thus
x = x(t) = V t + C, (2.2.4)
where C is an arbitrary constant. (Note that all these characteristics are parallel lines and fill the entire
x-t plane.) Now by our choice of the curve, (2.2.3) and the PDE, we have
du
= ut + V ux = 0,
dt
hence u = M where the constant M depends on the characteristic curve and hence on C, i.e, u is a
function C, which is just x − V t according to (2.2.4). Thus we write
u(x, t) = f (x − V t) (2.2.5)
for some function f . On the other hand, as has already been verified in Section 1.1, (2.2.5) is always
a solution of the PDE (2.2.1) for any smooth function f . Thus the set of all solutions, or the general
solution of the PDE is given by (2.2.5).
We invoke the initial condition in (2.2.1) as has been done in Section 1.1 and so the solution of (2.2.1) is
Before we discuss the method of characteristics further, let us get some feel of this solution. If we take the
snapshot of the graph of u as a function of x at time t, it is just the graph of the initial value φ(x) shifted
to the right by V (to the left if V < 0). The graph is travelling with velocity V without changing its
shape. For this reason, u is called a travelling wave solution. This solution can be interpreted as modelling
a lump of (non-diffusive) pollutant in a river with water moving with constant speed V .
Example 2.2.2
∂u ∂u
+y =u+y
∂x ∂y (2.2.6)
u(x, 1) = φ(x),
hence by
y = Cex .
So u, when restricted on a fixed characteristic curve (so constant C is fixed), satisfies
du
= ux + yuy = u + Cex .
dx
Solving this ODE, we obtain
u = Cxex + M ex ,
where M is a constant that depends on the characteristic curve and hence on C, and thus M = f (C) for
a function f . But on the curve, C = ye−x . We then have the general solution of the PDE:
So f (0) = 1 and nothing else we can say about f . Then we cannot uniquely determine u(x, y) when y 6= 0.
The moral of this story is that if an initial condition is imposed on a characteristic curve, then there may
not exist a solution; if there exists a solution, the solution is not unique.
(We use the differentials because we have the advantage of re-writing the equation in the form of either
dx dt
dt = · · · or dx = · · ·.) In the linear case (a and b independent of u), (2.2.7) is an ODE that may be solved
as in the previous examples. But in the nonlinear case, (2.2.7) appears to be useless because it involves
the unknown solution u. This is not so as we can see from the following.
Example 2.2.3 Consider the following quasilinear traffic/transport equation
∂u ∂u
+ c(u) =0
∂t ∂x (2.2.8)
u(x, 0) = φ(x).
This formula for u is not given in explicit form; this is a nonlinear phenomenon.
There is a more striking phenomenon due to nonlinearity: the initial value can be smooth, yet the solution
develops a singularity in finite time. This occurs when c0 (u) > 0 for all u and when there exist two points
x1 < x2 such that φ(x1 ) > φ(x2 ). In this case, the two characteristic lines issued from x1 and x2 on the
x−axis intersect at time
x2 − x1
t=− , (2.2.11)
c(φ(x2 )) − c(φ(x1 ))
which is positive. But along each characteristic line, the solution u must be constant. This shows that at
and after the time given by (2.2.11), a smooth solution is impossible to exist; something that determines
the smoothness of the solution must have broken down before or at this time. The question is: exactly
what has broken down? To answer this question, (2.2.10) is useful: differentiating both sides of it with
respect to x, we have
ux = φ0 (x − c(u)t)(1 − tc0 (u)ux ),
φ0 (x − c(u)t)
ux = .
1 + tc0 (u)φ0 (x − c(u)t)
Suppose there exists x0 such that
φ0 (x0 ) < 0.
Then along the characteristic line issued from the point x0 on the x−axis,
φ0 (x0 )
ux = .
1 + tc0 (φ(x0 ))φ0 (x0 )
Thus along the characteristic line as t tends to time
−1
> 0,
c0 (φ(x 0
0 ))φ (x0 )
the slope of u becomes unbounded. This is called the steepening effect. At or before the time given
above, the classical solution ceases to exist; the breakdown time or life-span of the classical solution is given
by
−1
ts = min .
x0 ∈R: φ (x0 )<0 c (φ(x0 ))φ0 (x0 )
0 0
When a classical solution stops to exist, we say a shock occurs. The subscript s in ts refers to the word
“shock”. Now the natural question is: how do we define a solution after the shock occurs? The resolution
is to introduce the notion of “weak solution”. In this course, we will not cover this topic.
We remark that if c0 (u) > 0 for all u and if the initial value φ is non-decreasing, then no two characteristic
lines intersect and the smooth solution exist for all time t ≥ 0.
It is helpful to interpret physically the appearance of shock: consider the case where c(u) = u and think
of (2.2.8) as a traffic/transport equation modelling the density of a substance flowing in a pipe. Then
the velocity function V is u/2 and so the more dense the substance is distributed, the faster it moves. If
the initial value φ is non-decreasing, particles move no slower than the ones behind (to the left), so we
11
do not expect the steepening effect. On the other hand, if the initial value is not non-decreasing, then
some particles will catch up some others that are initially ahead of them, causing “collision”, and thus the
steepening effect.
Example 2.2.4 (Discontinuous initial value.) Find the solution of the following Burger’s equation with
discontinuous initial condition:
ut + uux = 0
−1, x < 0,
u(x, 0) = φ(x) =
1, x > 0.
Solution
We study the following problem:
ut + uux = 0
−1, x < −,
u(x, 0) = φ(x) = x/, − ≤ x ≤ ,
1, x > .
By (2.2.10), we have
−1, x − ut < −,
u(x, t) = φ(x − u(x, t)t) = (x − ut)/, − ≤ x − ut ≤ ,
x − ut > ,
1,
i.e.
−1, x < −(t + ),
u(x, t) = x/(t + ), −(t + ) ≤ x ≤ (t + ),
1, x > t + .
Sending → 0, we get the solution of the original problem
−1, x < −t,
u(x, t) = x/t, −t ≤ x ≤ t,
1, x > t.
2
It is interesting to observe that the initial discontinuity is smoothed as illustrated in following figure. We
can say Burger’s equation favors non-decreasing initial values and dislikes other ones.
u(x, 0) 6 u(x, t0 ) 6
1 1
-x -x
x = t0
−1 −1
determined. If a curve is not non-characteristic at every point on it, then it is said to be a characteristic
curve of the PDE.
Now let us show that in the case of
this new definition of characteristic curves is equivalent to (2.2.7). Consider the following equations
(
aut + bux = f (x, t, u) (PDE)
(2.2.13)
ut dt + ux dx = du|Γ (du is known on Γ)
i.e.
" #( ) ( )
a b ut f (x, t, u)
=
dt dx ux du Γ
which in turn is equivalent to the fact that ut and ux cannot be uniquely determined. Thus (2.2.7) implies
that the curve Γ is characteristic in the sense of the new definition. Conversely, if (2.2.7) does not hold
at a point P (and for prescribed values of u on Γ), then both ut and ux at P = (x0 , t0 ) can be uniquely
determined. Moreover all higher order partial derivatives of u can be determined uniquely: differentiating
(2.2.13) with respect to t and then plugging in (x, t) = (x0 , t0 ), we have
(
autt + buxt = known quantity
utt + uxt dx/dt = known quantity,
thus (2.2.7) implies that utt and uxt at point P can be uniquely determined. Other higher-order derivatives
can be determined uniquely in the same fashion. Therefore if (2.2.7) fails at P , then Γ is non-characteristic
there. We have completed the proof of the equivalence of (2.2.7) and the new definition of characteristic
curves in the case of (2.2.12).
The advantage of the new definition of characteristic curves is that it applies to systems of PDEs where
we cannot foresee the likes of (2.2.7) easily. To make this point, we supply the following examples
Example 2.2.5 Consider the following first-order system of partial differential equations:
(
a1 ux + b1 uy + c1 vx + d1 vy = f1 (x, y)
a2 ux + b2 uy + c2 vx + d2 vy = f2 (x, y)
i.e.
ux
f1 (x, y)
a1 b1 c1 d1
f2 (x, y)
a2 uy
b2 c2 d2
dx
=
dy 0 0 du
vx
0 0 dx dy
vy
dv
Γ
a1 b1 c1 d 1
a2 b2 c2 d 2
det
dx dy
= 0.
0 0
0 0 dx dy
As in the previous discussion regarding (2.2.12), if the ODE system fails, then Γ is non-characteristic.
Thus this ODE system defines the characteristic equation.
2
Unlike in the case of single equation (2.2.12), the system of PDEs in the above example can no longer be
reduced to a systems of ODEs along a characteristic curve. Then what is the point of finding these curves?
The answer lies partially in the fact that if the “initial value”is prescribed on a characteristic curve, then
generally either the existence or the uniqueness of solution fails, as we have seen in Example 2.2.2. To have
the existence and uniqueness, we should prescribe the initial value on a non-characteristic curve: indeed, a
general local existence theorem called Cauchy-Kovalevski Theorem says that if everything (coefficients
and given functions) in the system is analytic, if the non-characteristic curve and the initial value are also
analytic, then in a neighborhood of the curve, the initial value problem has one and only one analytic
solution, which is obtained from the Taylor series mentioned in the definition of non-characteristic curves.
We close this section by one more example.
Example 2.2.6∗ (Advanced problem)
The rotation-free 2-D gas equations are given by
1
−ρ(ux + vy ) = ρx u + ρy v = (upx + vpy ).
c2
14
which is √
dy uv ± c u2 + v 2 − c2
= .
dx u2 − c2
2
Assignment 2
1. Solve the following initial value problem: 3ut + 5ux = 0, u(x, 0) = exp(−x2 ).
4. This exercise makes the point that the boundary condition for transport equations has to be given
carefully: show that the PDE ut + ux = 0, x ∈ [0, 1], t ∈ R, has no smooth solutions satisfying the
boundary condition u(0, t) = 1, u(1, t) = 2. Explain this physically. Hint: Draw several characteristic
curves.
(a) Find the time ts when shock first occurs; (b) Solve the initial value problem before time ts .
Prove that (a) for each fixed t ≥ 0, u is non-decreasing in x; (b) for each fixed x, u is non-increasing in
t ≥ 0. Hint: Argue by contradiction to prove (a).
The unit for J is energy/time/length (length2 should be replaced in by lengthn−1 if n = 1, 2). Now by
2
where the gradient is taken in the x−variable, and k > 0 (so J points in the direction in which the
temperature function decreases most rapidly). Moreover an empirical law says that thermal energy density
is proportional to the temperature; more precisely,
E = cρu, (3.1.3)
where ρ is the density function of the medium, c the specific heat which is positive. In principle, k, c, ρ
depend on location and time, however, for simplicity we assume in this course that they all are constant.
k is called the thermal conductivity of the medium. Combining (3.1.1-3), we have the heat equation
where a2 = k/(cρ) which is called the heat diffusion coefficient of the medium, and f = F/(cρ). The
unit of heat diffusion coefficient is length2 /time.
Reaction-diffusion equation
Let u(x, t) be the density function of a diffusive substance. Diffusion refers to the tendency of particles of
the substance moving from the region of higher concentration to the region of lower concentration. Then
Fick’s law, an analog of Fourier’s law, states that the flux of the substance is given by
where k is positive. Again, for simplicity in this course we assume that k is constant; it is called diffusion
coefficient. Now by the transport equation (2.1.4), we have the reaction-diffusion equation
If heat conduction occurs in a region Ω with boundary ∂Ω, obviously what happens on the boundary to
the temperature function affects the temperature function inside the region. Thus it is necessary to give
a boundary condition before solving the heat equation. (This also holds for reaction-diffusion equations.)
16
The physical meaning of the first kind B.C. is that the temperature at boundary is specified. For example,
if a rod of length l with the left end being put in ice-water and the right end being put in boiling water,
then the boundary condition is u(0, t) = 0 (◦ C) and u(l, t) = 100 (◦ C).
∂u
J · n = −k = 0,
∂n
where n is the unit normal vector field on the boundary. In general, we can specify the normal component
of the flux on the boundary:
∂u
−k (x, t) = µ(x, t), x ∈ ∂Ω.
∂n
This is the general Neumann B.C..
∂u
−k (x, t) = H(u(x, t) − µ(x, t)),
∂n
where H is a positive constant, called coefficient of surface heat transfer. It can be re-written as
∂u
(x, t) + h(u(x, t) − µ(x, t)) = 0, x ∈ ∂Ω
∂n
where h = H/k is positive. This is the general Robin B.C. It is also called the radiation B.C.. Observe
in the limit h → 0, Robin B.C. becomes the homogeneous Neumann B.C.; in the limit h → ∞, Robin B.C.
becomes the Dirichlet B.C..
All these boundary conditions make sense for the reaction-diffusion equation.
ut = a2 ∆u + f (x, t)
x ∈ Ω, t > 0
u(x, 0) = φ(x), x∈Ω (3.3.1)
Boundary conditions on ∂Ω,
17
where φ(x) is the initial temperature distribution. Since the PDE and the boundary condition are lin-
ear, the uniqueness of the above problem is satisfied if and only if the homogeneous heat equation with
homogeneous boundary condition has only the trivial solution u = 0.
ut = a2 ∆u
x ∈ Ω, t > 0
u(x, 0) = 0, x∈Ω (3.3.2)
Homogeneous boundary conditions on ∂Ω,
Here by homogeneity in the boundary condition, we mean that µ is zero. Multiplying both sides of the
homogeneous heat equation by u and integrating on Ω, we have
1 d
R 2 R
2 dt Ω u (x, t)dx = uu (x, t)dx
ΩR t
2
= a Ω u∆u(x, t)dx
= a2 Ω (∇ · (u∇u) − |∇u|2 )(x, t)dx
R
where we have used the divergence theorem. The boundary integral is zero in the cases of Dirichlet
and Neumann boundary conditions; it is non-positive in the Robin case because h is positive. Thus the
following function of t is nonincreasing: Z
u2 (x, t)dx. (3.3.3)
Ω
This together with the facts that this function is non-negative and is equal to zero initially implies that
Z
u2 (x, t)dx ≡ 0, t > 0.
Ω
Model problem
Consider the following initial-boundary value problem for the heat equation:
ut = a2 uxx
(0 < x < l, t > 0)
u(0, t) = 0, u(l, t) = 0 (3.4.1)
u(x, 0) = φ(x).
In the above model problem, the PDE and boundary condition are homogeneous.
Separation of variables
The strategy to solve (3.4.1) is to first find solutions that satisfy the PDE and the boundary condition,
and then take care of the initial condition. The solutions to start with have the following form
We say that the variables x and t are separated. Substituting this into the PDE, we have
i.e.
X 00 (x) T 0 (t)
= 2 .
X(x) a T (t)
6 6
function of t
function of x
Notice that the left hand side of the above equation is a function only of x, while the right hand side is a
pure function of t. The only possibility is that both sides of the above equation are constants, i.e.
X 00 (x) T 0 (t)
= 2 ≡ −λ (constant),
X(x) a T (t)
where the negative sign is selected because we will show that λ > 0. Thus we have two separated equations:
( 00
X (x) + λ X(x) = 0 (0 < x < l),
(3.4.3)
T 0 (t) + a2 λ T (t) = 0 (t > 0).
Eigenvalue problem
The boundary conditions in (3.4.4) leads to that X(0) = 0 and X(l) = 0. Thus we get
X 00 (x) + λ X(x) = 0
(
(0 < x < l),
(3.4.5)
X(0) = 0, X(l) = 0
X(x) = 0 is obviously a solution of the above problem. But it gives only a trivial solution. Thus we demand
that X 6= 0. (3.4.5) is analogous to the eigenvalue problem of a linear transformation and so (3.4.5) is said
to be an eigenvalue problem, the unknown constant λ is called eigenvalue, any nontrivial function X
is called eigenfunction, the set of all eigenfunctions corresponding to a fixed eigenvalue λ, plus the zero
function, is called the eigenspace. Any eigenspace is linear in the sense that if u1 and u2 belong to the
eigenspace, then so does c1 u1 + c2 u2 for any constants c1 and c2 .
It is easy to check that λ = 0 leads to a trivial solution and thus 0 is not an eigenvalue. For λ 6= 0, the
general solution of the ODE in (3.4.5) is clearly given by
√ √
−λ x
X(x) = C1 e + C2 e− −λ x
.
This leads to
√
2 −λ l = i 2nπ (n = 1, 2, · · ·).
19
Principle of superposition
For each n, the second equation in (3.4.3) becomes
nπa 2
Tn0 (t) + Tn (t) = 0 ∀ n.
l
It is easy to show that
nπa 2
Tn (t) = φn exp − t , (n = 1, 2, · · ·) (3.4.8)
l
where φn is an arbitrary constant. Thus we have infinitely many functions
nπa 2 nπx
un (x, t) = Xn (x) · Tn (t) = φn exp − t sin , (n = 1, 2, · · ·), (3.4.9)
l l
that satisfy the PDE and the B.C. (3.4.1). According to the principle of superposition for homogeneous
linear equations, a finite sum of solutions is still a solution; this still holds for an infinite sum if the
convergence of the sum (series) is guaranteed. To take care of the initial condition, we boldly form the
sum of all solutions un to obtain
∞
X nπa 2 nπx
u(x, t) = φn exp − t sin . (3.4.10)
n=1
l l
(The first person to do so was Fourier.) Formally, that is, non-rigorously, u satisfies the PDE and the B.C.
Indeed, because of the exponential decay of un and its partial derivatives as n → ∞ if t > 0, it is possible
to prove that if all the φn are bounded, then u is well-defined (meaning: the series converges) and all its
partial derivatives exist and are continuous in the upper plane t > 0; moreover u satisfies the PDE and
the B.C. for t > 0. The rigorous proofs of these statements are not required for this course, unless your
instructor insists otherwise.
It is our hope that by choosing the coefficients φn suitably, u satisfies the initial condition. We wish
∞
X nπx
u(x, 0) = φn sin = φ(x). (3.4.11)
n=1
l
That means the coefficients φn in (3.4.10) are determined by the Fourier expansion of initial temperature
φ(x). Recall the formula
(
2 l
Z
nπx mπx 0 m 6= n
sin sin dx = (3.4.12)
l 0 l l 1 m = n.
nπx
Then multiplying (3.4.11) by sin and integrating term-by-term on [0, l], we have
l
2 l
Z
nπx
φn = φ(x) sin dx (3.4.13)
l 0 l
20
Remark 1
Notice that
∞
X nπa 2 nπx
u(x, t) = φn exp − t sin −→ 0 as t → + ∞.
n=1
l l
This result coincides with our physical intuition. The temperature will be zero eventually, since the
terminals of the rod are put in ice-water mixture and there is neither creation nor degradation of thermal
energy inside of the rod (f (x, t) = 0).
Remark 2
Notice that the convergence rate of the term
nπa 2
Tn (t) = exp − t −→ 0 as t → + ∞
l
is faster, if the integer n is larger. That means the higher frequency components, which is related to
the larger n in sin nπxl , will vanish faster than the lower frequency components. Thus we expect the
temperature will be smoothed as time increases. Indeed, as we mentioned before u(x, t) as a function of
x becomes infinitely smooth right after t = 0, even if the initial φ is very rough (e.g. nowhere continuous
but is square-integrable in the sense of Lesbegue). This phenomenon is called the smoothing effect of
heat equation, which also holds for general parabolic equations.
Remark 3
According to Remark 2, a solution u of the heat equation (with a B.C.) at time t = 1 must be infinitely
smooth in x. So if we want to solve the heat equation backward from time t = 1, the initial value u(x, 1)
must be at least this smooth. So in general, solving the heat equation backward in time is an ill-posed
problem. It is ill-posed because of another reason: the solutions
1 nπa 2 nπx
un (x, t) = exp − t sin , (n = 1, 2, · · ·),
n l l
at time t = 0 are all bounded between −1/n and 1/n and hence are small if n is large, but at a negative
time, say, t = −1, un becomes unbounded as n → ∞. Thus small input leads to unbounded output; this
makes no physical sense and is highly unstable. This discussion of ill-posedness also applies to general
parabolic equations. The most famous backward parabolic equation is the Black-Scholes PDE that arises
in Finance
∂V 1 ∂2V ∂V
+ σ 2 S 2 2 + rS − rV = 0. (3.4.14)
∂t 2 ∂S ∂S
2
Note that the sign of the coefficient of ∂∂SV2 is positive, unlike the heat equation. The coefficient becomes
negative if we change variable by replacing t by T − t (T is an arbitrary constant). Black-Scholes equation
in its original form (3.4.14) can only be solved if a terminal condition V (S, T ) = φ(S) is given, that is,
V (S, t) can be determined by φ for t ≤ T . Fortunately, solving Black-Scholes equation with a terminal
condition at the “expiration day”(of the call option) is helpful for stock investors. See Exercise 9.
In Chapter 5, we will see that the backward problem is well-posed for wave propagation problems (hyper-
bolic equation), e.g. sonar, electro-magnetic wave.
Remark 4
The methodology of separation of variables applies to general linear parabolic equations in higher spatial
dimensions. The technical difficulty in handling the higher dimensional case is that X 00 in the eigenvalue
problem (3.4.5) now becomes the Laplacian ∆X, so we can no longer use the ODE method to solve the
eigenvalue problem easily. In this course, we will not go into this in detail.
21
Theorem 3.5.1
Suppose that λm and λn (λm 6= λn ) are two eigenvalues of the following general eigenvalue problem:
00
X (x) + λ X(x) = 0 (a < x < b),
α1 X 0 (a) + α2 X(a) = 0, (3.5.1)
β1 X 0 (b) + β2 X(b) = 0,
where {α1 , α2 } are not all zeroes, {β1 , β2 } are not all zeroes, and Xm (x), Xn (x) are eigenfunctions
corresponding to λm , λn respectively. Then the eigenfunctions Xn (x) and Xm (x) are orthogonal in the
following sense:
Z b
Xn (x)Xm (x) dx = 0. (3.5.2)
a
Proof
Since Xm (x), Xn (x) are eigenfunctions w.r.t. eigenvalues λm , λn , we have
( 00
Xm (x) + λm Xm (x) = 0, (1)
Xn00 (x) + λn Xn (x) = 0. (2)
Notice that
0
(
α1 Xm (a) + α2 Xm (a) = 0,
α1 Xn0 (a) + α2 Xn (a) = 0.
{α1 , α2 } is a non-trivial solution of the above system, we must have
0
Xm (a) Xm (a)
det 0 = 0,
Xn (a) Xn (a)
i.e.
0
Xm (a)Xn (a) − Xm (a)Xn0 (a) = 0.
22
Similarly we have
0
Xm (b)Xn (b) − Xm (b)Xn0 (b) = 0.
Then (3.5.3) becomes
Z b
(λm − λn ) Xn Xm dx = 0.
a
Since λm 6= λn , one must have
Z b
Xn Xm dx = 0.
a
2
In the previous section, we have seen that the eigenvalue problem (3.4.5) with Dirichlet B.C. has infinitely
many eigenvalues (3.4.6) that diverge to infinity, each of which has a one dimensional eigenspace. We also
note that all the eigenfunctions corresponding to λn change sign exactly n − 1 times. All these can be
generalized for the more general eigenvalue problem (3.5.1) as follows.
Theorem 3.5.2
(i) The eigenvalues of (3.5.1) are real and form an increasing and diverging sequence
(ii) For each n, the eigenspace Vn corresponding to the eigenvalue λn is one dimensional.
(iii) Any eigenfunction corresponding to λn changes sign exactly n − 1 times; in particular, every eigen-
function corresponding to λ1 is either positive or negative on (a, b).
We point out that if the B.Cs are not given as in (3.5.1), then eigenspaces may not be one-dimensional
(e.g. in the case of the periodic B.C, all except the first eigenspaces are two dimensional - see Exercise
3, part (2)). Theorems 3.5.1 and 3.5.2 form the Sturm-Liouville theory that actually works when X 00 in
(3.5.1) is replaced by the more general operator (a(x)X 0 )0 + c(x)X. We will not go into details in this
course.
Suppose we have obtained all eigenvalues {λn , n = 1, 2, · · ·}, and their corresponding eigenfunctions
{Xn (x), n = 1, 2, · · ·}. To solve the heat equation ut = a2 uxx with the B.C. in (3.5.1), we form
The right hand side of (3.5.5) with φn given by (3.5.6) is called the generalized Fourier series or the
eigenexpansion of φ. There is a subtle point in the above discussion: what we have actually shown is
that if the right hand side of (3.5.5) converges to φ, then φn is given by (3.5.6), namely, only the generalized
Fourier series can converge to φ. But then logically, we must answer the following question: what’s the
requirement on φ so that the the generalized Fourier series/eigenexpansion of φ converges to φ in some
sense? The following theorem answers the question.
Theorem 3.5.3
(i) If φ is square integrable, i.e, φ2 has a finite integral on [a,b], then the eigenexpansion of φ converges to
φ in the L2 sense:
Z b Xk
|φ(x) − φn Xn (x)|2 dx → 0, as k → ∞.
a n=1
(ii) If φ, φ0 and φ00 all are continuous on [a, b] and φ satisfies the B.C. in (3.5.1), then the eigenexpansion
of φ converges to φ uniformly on [a, b].
Remark In some special cases, such as the case when the B.C. at both x = 0 and x = l is Dirichlet or
when the B.C. at both boundary points is Neumann (so the eigen-expansion is either the classical sine
series or the cosine series), to have the uniform convergence in (ii), we actually do not even need the
existence of φ00 , and in the Neumann case we do not need φ to satisfy the B.C.. In these two special
cases, weaker assumptions yield weaker conclusions: the Fourier series converges to φ(x) at every x in the
open interval (a, b) if φ is continuous and φ0 is piecewise continuous on [a, b] (piecewise continuity means
continuity everywhere except at finitely many points where the function still has finite one-sided limits);
In the more general case when φ and φ0 are piecewise continuous on [a, b], the Fourier series converges to
the average of the right and left limits of φ
1
(φ(x − 0) + φ(x + 0))
2
at every x in (a, b); the sine series converges to 0 at x = 0, l (because all the sine terms in the series are
equal to 0 at these two end points); the cosine series at x = 0 converges to φ(0 + 0), at x = l to φ(l − 0).
1◦
X 00 (x) + λ X(x) = 0
(
(0 < x < l),
X(0) = 0, X(l) = 0
nπ 2
Eigenvalues: λn = l , n = 1, 2, · · ·
Eigenfunctions: Xn (x) = sin nπx
l , n = 1, 2, · · ·.
Rl l
Integral formula: 0
(Xn (x))2 dx = , n = 1, 2, · · · .
2
2◦
X 00 (x) + λ X(x) = 0
(
(0 < x < l),
X 0 (0) = 0, X 0 (l) = 0
nπ 2
Eigenvalues: λn = l , n = 0, 1, 2, · · ·
Eigenfunctions: Xn (x) = cos nπx
l , n = 0, 1, 2, · · ·.
24
Rl Rl l
Integral formula: 0
(X0 (x))2 dx = l; 0
(Xn (x))2 dx = , n = 1, 2, · · · .
2
3◦
X 00 (x) + λ X(x) = 0
(
(0 < x < l),
X(0) = 0, X 0 (l) = 0
2
(n+ 21 )π
Eigenvalues: λn = l , n = 0, 1, 2, · · ·
(n+ 12 )πx
Eigenfunctions: Xn (x) = sin l , n = 0, 1, 2, · · ·.
Rl l
Integral formula: 0
(Xn (x))2 dx = , n = 0, 1, 2, · · · .
2
4◦
X 00 (x) + λ X(x) = 0
(
(0 < x < l),
X 0 (0) = 0, X(l) = 0
2
(n+ 12 )π
Eigenvalues: λn = l , n = 0, 1, 2, · · ·
(n+ 12 )πx
Eigenfunctions: Xn (x) = cos l , n = 0, 1, 2, · · ·.
Rl l
Integral formula: 0
(Xn (x))2 dx = , n = 0, 1, 2, · · · .
2
Example 3.5.1
X 00 (x) + λ X(x) = 0
(
(−π < x < π),
X(−π) = 0, X 0 (π) = 0
Solution
Let ξ = x + π and X(x) = X(ξ − π) ≡ y(ξ). Then the above eigenvalue problem becomes
( 00
y (ξ) + λ y(ξ) = 0 (0 < ξ < 2π),
y(0) = 0, y 0 (2π) = 0
According to the result 3◦ , the eigenvalues and eigenfunctions are given by
2 2
(n + 21 )π
n 1
λn = = + , n = 0, 1, 2, · · ·
2π 2 4
and
n 1
yn (ξ) = sin + ξ, n = 0, 1, 2, · · · .
2 4
Therefore the solution of the original problem is given by
2
n 1
λn = + , n = 0, 1, 2, · · ·
2 4
and
n 1
Xn (x) = sin + (x + π), n = 0, 1, 2, · · · .
2 4
2
Example 3.5.2
X 00 (x) + λ X(x) = 0
(
(0 < x < l),
X(0) = 0, X 0 (l) + hX(l) = 0
25
Hence λ ≥ 0. If λ = 0, then the right hand of the last equation is zero, implying X 0 (x) ≡ 0 on [0, l]. So
X(x) ≡ Const. on [0, l]. But X(0) = 0 and so X(x) ≡ 0, which contradicts the early agreement that an
eigenfunction is not identically equal to zero. We have shown that any eigenvalue must be positive.
Now we can write the general solution of the ODE in the eigenvalue problem as :
√ √
X(x) = C1 sin( λx) + C2 cos( λx).
i.e.
µ √
= − tan µ (µ = λ l).
hl
This equation has infinitely many nonzero roots as illustrated below.
µ
(
(((
((( (( h l
(((
( (((((
(((
((((
((( • π • 2π • 3π -µ
µ1 µ2 µ3
− tan µ
In section 3.4, the method of separation of variables is applied to solve the homogeneous heat conduction
problem with homogeneous boundary conditions. For non-homogeneous boundary conditions, the method-
ology is to transfer the non-homogeneous BCs to homogeneous ones. For non-homogeneous equations, the
methodology is to first expand the inhomogeneous term by the eigenfunctions and then use the technique
for solving non-homogeneous ODE.
Model problem
ut = a2 uxx + f (x, t)
(0 < x < l, t > 0),
u(0, t) = µ1 (t), u(l, t) = µ2 (t) (t > 0), (3.6.1)
u(x, 0) = φ(x) (0 < x < l).
Thus the function w(x, t) must satisfy the non-homogeneous boundary condition, i.e.
Ut = a2 Uxx + F (x, t)
(0 < x < l, t > 0),
U (0, t) = 0, U (l, t) = 0 (t > 0), (3.6.5)
U (x, 0) = Φ(x) (0 < x < l),
where
∂w
F (x, t) = f (x, t) −
∂t (3.6.6)
Φ(x) = φ(x) − w(x, 0).
The original problem (3.6.1) for u(x, t) with non-homogeneous BC is transformed to the problem (3.6.5)
for U (x, t) with homogeneous BC.
Now we expand F (x, t) and initial condition Φ(x) in {Xn (x), n = 1, 2, · · ·}, i.e.
∞
X
F (x, t) = Fn (t) Xn (x) (3.6.7)
n=1
and
∞
X
Φ(x) = Φn Xn (x), (3.6.8)
n=1
where Rl
0
F (x, t)Xn (x) dx
Fn (t) = Rl
0
(Xn (x))2 dx
and Rl
0
Φ(x)Xn (x) dx
Φn = Rl .
0
(Xn (x))2 dx
We also expand the solution
∞
X
U (x, t) = Tn (t) Xn (x). (3.6.9)
n=1
Multiplying both sides of each of the above equation by Xk (x) and integrating on [0, l], then using the
orthogonality proved in Theorem 3.5.1, we have
Solving this initial value problem for the linear ODE, we obtain
Z t
2 2
Tn (t) = e−a λn (t−τ )
Fn (τ ) dτ + Φn e−a λn t
. (3.6.11)
0
In this fashion, the solution U (x, t) of (3.6.5) and the solution u(x, t) = U (x, t) + w(x, t) of (3.6.1) are
determined.
Summary
The following is a step-by-step description for the method of separation of variables:
Step 1: (Nonhomogeneous BCs)
Let u(x, t) = U (x, t) + w(x, t). Select w(x, t) to transform the non-homogeneous BC to a homogeneous
one.
Step 2: (Eigenvalue problem)
Let U (x, t) = X(x) · T (t). Derive the eigenvalue problem for X(x) from the homogeneous PDE and BC.
Step 3: (Eigenvalues and eigenfunctions)
Find the eigenvalues λn and eigenfunctions Xn (x) from the eigenvalue problem.
28
Step 4: (Eigenexpansions)
Find Fourier/eigenexpansions of the non-homogeneous term in the equation, F (x, t), and of the initial
value Φ(x).
Step 5: (Solution of non-homogeneous equation)
P
Set U (x, t) = Tn (t) Xn (x). Find Tn (t) from the ODE derived from the nonhomogeneous PDE and the
initial condition.
Example 3.6.1
ut = a2 uxx + x − π
(0 < x < π, t > 0),
u(0, t) = π, u(π, t) = 0 (t > 0), (3.6.12)
u(x, 0) = π − x
(0 < x < π).
Solution
Step 1: (Nonhomogeneous BC)
Let u(x, t) = U (x, t) + w(x, t) and w(x, t) = π − x. The problem (3.6.12) becomes
Ut = a2 Uxx + x − π
(0 < x < π, t > 0),
U (0, t) = 0, U (π, t) = 0 (t > 0), (3.6.13)
U (x, 0) = 0 (0 < x < π).
λn = n2 ; Xn (x) = sin(nx), n = 1, 2, · · · .
Step 4: ( Eigen-expansions)
Let
∞
X
F (x, t) = x − π = Fn sin(nx).
n=1
2 π
Z
2
Fn = (x − π) sin(nx) dx = − n = 1, 2, · · · .
π 0 n
The general solution of the homogeneous ODE is C exp(−(na)2 t), and the particular solution of the above
non-homogeneous ODE is (by inspection) −2/(n3 a2 ). Thus the general solution is given by
2 2 2
Tn (t) = C e−n a t
− .
n3 a 2
The initial condition Tn (0) = 0 leads to
2 2 2
Tn (t) = e−n a t
−1 .
n3 a2
Therefore we get that
∞ ∞
X X 2 −n2 a2 t
U (x, t) = Tn (t) Xn (x) = e − 1 sin(nx),
n=1 n=1
n3 a2
and
∞
X 2 −n2 a2 t
u(x, t) = U (x, t) + w(x, t) = π − x + e − 1 sin(nx).
n=1
n3 a2
2
(x − ξ)2
1
G(x, t; ξ) = √ exp − (t > 0), (3.7.1)
2a πt 4a2 t
where x and t are variables, ξ ∈ R a parameter. It can be directly verified that G ∈ C ∞ w.r.t. x or t, and
satisfies the homogeneous heat conduction equation for any ξ, i.e.
Remark
There are several approaches to the derivation of the fundamental solution G(x, t; ξ) . In Appendix 3.1,
G(x, t; ξ) is derived by the method of separation of variables. Other approaches, e.g. similarity analysis,
Fourier transformation, may be found in other textbooks.
Property 1◦
G(x, t; ξ)
1 6
√
2a πt
-x
ξ
30
Property 2◦
∂m
G(x, t; ξ) −→ 0 as x → ∞, f or t > 0, ξ ∈ (−∞, + ∞). (3.7.3)
∂xm
Property 3◦
Z +∞
G(x, t; ξ)dx = 1, f or t > 0, ξ ∈ (−∞, + ∞). (3.7.4)
−∞
Property 4◦
0 x 6= ξ
lim+ G(x, t; ξ) = (3.7.5)
t→0 +∞ x = ξ.
Property 5◦
Z +∞
lim f (ξ)G(x, t; ξ)dξ = f (x), (3.7.6)
t→0+ −∞
Theorem 3.7.1
Consider Cauchy problem for the heat equation:
ut = a2 uxx −∞ < x < +∞, t>0
(3.7.7)
u(x, 0) = φ(x) −∞ < x < +∞,
where φ(x) is bounded and continuous on R. The following function gives a bounded solution
Z +∞
u(x, t) = φ(ξ)G(x, t; ξ)dξ. (3.7.8)
−∞
Proof
Z +∞
ut − a2 uxx = φ(ξ)(Gt − a2 Gxx )dξ = 0.
−∞
Z +∞
lim+ u(x, t) = lim+ φ(ξ)G(x, t; ξ)dξ = φ(x).
t→0 t→0 −∞
We shall prove in the future the uniqueness of Cauchy problem, i.e. it has at most one bounded solution.
Thus (3.7.8) is the only bounded solution of the Cauchy problem. We emphasize that the Cauchy problem
has no boundary condition. In applications, the initial values are often piecewise continuous, then at every
continuous point x of φ, we still have limt→0+ u(x, t) = φ(x); at a discontinuous point x, we have
1
lim u(x, t) = (φ(x − 0) + φ(x + 0)),
t→0+ 2
that is, the limit is the average of the left and right limits of the initial value. Since there are just discretely
many discontinuous points of φ where the initial condition is not satisfied exactly, we still say (3.7.8) is
the solution of the Cauchy problem.
31
u(x, t) = ∆m G(x, t; x0 ).
Thus the fundamental solution G(x, t; x0 ) is the mass distribution function when a unit point mass is put
at x0 at time t = 0.
Critique. The above interpretation of the fundamental solution actually reveals a flaw of the heat/diffusion
equation: if we put one unit of thermal density or point mass at 0 at time t = 0 and if there is no cre-
ation of thermal energy or mass, then even if the thermal energy/mass travels at the speed of light - the
supposedly highest speed at which any object can travel according to Einstein - we still need to wait for a
while to feel a fragment of the thermal energy/mass 10 miles away; but the temperature or mass density
10 miles away from the origin at any time t is raised from 0 to G(10, t; 0) > 0 ! This is certainly absurd,
and is enough for the purist to immediately abandon the heat equation. But wait: if a2 = 1 and t = 0.1,
then G(10, t; 0) = 2.3811 × 10−109 , so tiny that the flaw of the heat/diffusion equation is practically harm-
less! Moreover, the heat/diffusion equation has been proved to match experimental data so well that it is
accepted as the standard model in the science and engineering community.
Example 3.7.1
32
ut = a2 uxx
0 < x < +∞, t > 0,
u(x, 0) = 0 0 < x < +∞, (∗)
u(0, t) = N0 , t > 0.
Solution
Consider the standard Cauchy problem on the whole line:
ut = a2 uxx
(
−∞ < x < +∞, t > 0,
(∗)0
u(x, 0) = φ(x) −∞ < x < +∞,
The solution of (∗)0 satisfies problem (*), except the boundary condition u(0, t) = N0 . We will select a
suitable Φ(x) to satisfy this boundary condition. According to Theorem 3.7.1, the solution of (∗)0 is given
by
Z +∞ Z 0
u(x, t) = φ(ξ)G(x, t; ξ)dξ = Φ(ξ)G(x, t; ξ)dξ.
−∞ −∞
Φ(x) = 2N0 .
Z +∞
2 2
u(x, t) = √ N0 e−y dy
π x√
2a t
Z +∞ Z x√ !
2 −y 2 2a t
−y 2
= N0 √ e dy − e dy
π 0 0
x√
= N0 1 − erf( ) ,
2a t
where the so-called error function is a standard special function in most of mathematical software and
function tables, defined by
Z x
2 2
erf(x) = √ e−u du
π 0
ut = a2 ∆u, x ∈ Rn , t > 0,
1 −|x − ξ|2
G(x, t; ξ) = exp( ).
(4πa2 t)n/2 4a2 t
Properties 1◦ -5◦ , and Theorem 3.7.1 hold with the obvious modifications.
This section is devoted to qualitative study of the heat equation. Consider a refrigerator occupying region
Ω; and think about the maximum of the temperature function u(x, t) inside the refrigerator during time
interval [0, T ]. The maximum of u(x, t) must be achieved either at a boundary point at some time between
t = 0 and t = T , or inside Ω at time t = 0 (as in the case of a refrigerator which is turned on at time
t = 0). Thus
max u = max u, (3.8.1)
DT ΓT
where DT = Ω × (0, T ] and ΓT = (∂Ω × [0, T ]) ∪ (Ω × {0}). DT is called the parabolic interior , and ΓT
the parabolic boundary of the cylinder Ω × [0, T ].
This is the physical interpretation of the weak maximum principle for the heat equation that we
will formulate mathematically now. First let us think about how to use the PDE language to model a
refrigerator. We have a refrigerator if and if the creation-degradation rate of thermal energy is non-positive,
i.e. f in (3.1.4) is non-positive in DT . Now we are ready to state
ut − a2 ∆u ≤ 0. (3.8.2)
ut (x0 , t0 ) = 0, or ≥ 0. (3.8.3)
On the other hand, consider the function u(x, t0 ), which assumes its maximum value at x0 . By the Second
Derivative Test, the Hessian matrix
∂2u
( (x0 , t0 ))1≤i, j≤n
∂xi ∂xj
is non-positive-definite. So the trace of this matrix is non-positive. But the trace is just the Laplace of u
and hence
∆u(x0 , t0 ) ≤ 0.
Combining this with (3.8.3) we reach a contradiction to (3.8.2) at point (x0 , t0 ). We have proved (3.8.1)
when the strict inequality in (3.8.2) holds.
Now suppose we do not have the strict inequality. Define a new function
where > 0 is a constant. The new function satisfies the strict inequality in (3.8.2) and so it satisfies
max v = max v.
DT ΓT
Comparison principle
Let u and v satisfy
ut − a2 ∆u ≥ vt − a2 ∆v
x ∈ Ω, T ≥ t > 0
u(x, t) ≥ v(x, t), x ∈ ∂Ω, T ≥ t > 0 (3.8.4)
u(x, 0) ≥ v(x, 0), x ∈ Ω.
Then
u(x, t) ≥ v(x, t), x ∈ Ω × [0, T ].
35
Moreover, if there exists (x0 , t0 ) ∈ Ω × (0, T ] where u and v touch each other, then u and v are identical,
at least before time t0 .
Proof
Let w = v − u. Then w satisfies (3.8.2) and hence by the weak maximum principle, we have
max w = max w.
DT ΓT
But on the vertical component of ΓT , w ≤ 0 by the B.C.; the same is true at time t = 0 by the initial
condition. Thus w ≤ 0 and u ≤ v on DT .
If u and v touch at (x0 , t0 ), then w achieves its maximum in the parabolic interior DT . It follows from
the strong maximum principle that w is identically equal to zero before time t0 . This completes the proof.
ut − a2 ∆u = f (x, t)
x ∈ Ω, T ≥ t > 0
u(x, t) = g(x, t), x ∈ ∂Ω, T ≥ t > 0
u(x, 0) = φ(x), x ∈ Ω.
Then
max |u1 − u2 | ≤ max |φ1 − φ2 | + max |g1 − g2 | + T max |f1 − f2 |. (3.8.5)
DT Ω ∂Ω×[0,T ] DT
Proof
Let v = u1 − u2 and
Then
wt − a2 ∆w ≥ vt − a2 ∆v
x ∈ Ω, T ≥ t > 0
w(x, t) ≥ v(x, t), x ∈ ∂Ω, T ≥ t > 0
w(x, 0) ≥ v(x, 0), x ∈ Ω.
By the comparison principle, we have v ≤ w on DT . Similarly, −v ≤ w on DT . These and the fact that
w is less than or equal to the right hand side of (3.8.5) imply (3.8.5).
36
In applications, the structure components, i.e. the source term f , the boundary value g and the initial
value φ are measured experimentally and hence are not given precisely. Equation (3.8.5) says that small
errors in measuring these data result in a small error in the solution. Thus we have structural stability.
ut − a2 ∆u ≤ 0, x ∈ Rn , 0 < t ≤ T,
(3.8.6)
u(x, 0) = φ(x), x ∈ Rn .
Then
sup u = sup φ. (3.8.7)
Rn ×[0,T ] Rn
Proof
The idea is the same as in the case of bounded region. The new difficulty is that the sup of u may not be
assumed at a finite point. To overcome this, we use a function Γ(x, t) that satisfies the heat equation, but
has (exponential) growth as |x| → ∞, so when we subtract u by Γ, the sup is achieved at a finite point.
This Γ is a modification of the fundamental solution:
1 2 2
Γ(x, t) = √ e|x| /(4a (T +1−t)) .
T +1−t
We check by direct computation that
Γt − a2 ∆Γ = 0.
Take a small constant and define
Then v satisfies the strict inequality in (3.8.6). The sup of v on Rn × [0, T ] is assumed somewhere at
(x0 , t0 ). If t0 > 0, then we reach a contradiction as in the case of bounded region. So t0 = 0 and hence
u(x, 0) = φ(x), x ∈ Rn .
We mention that without the boundedness condition, there is no uniqueness: the Cauchy problem has
infinitely many unbounded solutions that have the 0 initial value.
∞
X nπa 2 nπx
u(x, t) = φn exp − t sin
n=1
l l
∞
!
2 l
Z
X nπξ nπa 2 nπx
= φ(ξ) sin dξ exp − t sin
n=1
l 0 l l l
∞
Z l X !
2 nπξ nπx nπa 2
= sin sin exp − t φ(ξ) dξ
0 n=1
l l l l
Z lX
≡ ·φ(ξ) dξ
0
nπ P
Let ωn = . The in the above formula becomes
l
∞
X X 2 2 2
= sin(ωn ξ) sin(ωn x)e−ωn a t
n=1
l
Notice that
nπ π π
ωn = = n · ≡ n · ∆ω (∆ω = )
l l l
-ω
ω0 ω1 ω2 ωn
Thus
dI(c) c
= − I(c).
dc 2b
Solving this (linear and also separable) ODE, we have
c2
I(c) = I(0)e− 4b ,
38
(x − ξ)2
1
G(x, t; ξ) = √ exp −
2a πt 4a2 t
Notice that
Z lX Z +∞
u(x, t) = ·φ(ξ) dξ −→ [G(x, t; ξ) − G(−x, t; ξ)]φ(ξ) dξ as l → +∞.
0 0
where f (·) is bounded on R and is continuous at x. By (3.7.4) and the the formula for G, we have
Z +∞ Z +∞
f (x) = f (x) G(x, t; ξ)dξ = f (x)G(x, t; ξ)dξ
−∞ −∞
1
Z +∞
2 √
√ e−η (f (x + 4taη) − f (x))dη. (∗ ∗ ∗)
π −∞
39
Then
1
Z
2 √
√ | e−η (f (x + 4taη) − f (x))dη| ≤ 2M · = .
π |η|≥L 2M
Since f is continuous at x, there exists a small τ > 0 such that
√
|f (x + 4taη) − f (x)| ≤ for any 0 < t < τ, |η| ≤ L,
and hence 2 √ 2
√1 | e−η (f (x √1 e−η dη
R R
π |η|≤L
+ 4taη) − f (x))dη| ≤ π |η|≤L
R +∞ −η2
≤ √1 e dη ·
π −∞
= .
Then the absolute value of (***) is less than or equal to
Z Z
| |+| | ≤ + for any 0 < t < τ.
|η|≥L |η|≤L
Assignment 3
1. (Transmission conditions)
Consider a surface S that separates two media with different thermal conductivities k1 and k2 . Let u1
and u2 be the temperature in the media. Suppose the media are in intimate contact along the surface S
so we have
u1 = u2 on S. (1)
Prove that on S,
∂u1 ∂u2
−k1 = −k2 , (2)
∂n ∂n
where n is the unit normal vector field of the surface S. ( (1) and (2) are called transmission conditions.)
Hint: Take an arbitrary patch ∆S of S, and think about the rate at which thermal energy crosses the
patch in the direction of the normal.
of the body and the coating be k1 and k2 , respectively. Prove that on the boundary ∂Ω1 of the body, we
have approximately Robin boundary condition
∂u1 k2
k1 + (u1 − H) = 0, (3)
∂n δ
where n is the unit outer normal vector field of ∂Ω1 . (Equation (3) is called the effective boundary
condition; its significance is that with it we do not need to solve, analytically or numerically, the heat
equation inside the coating–we just need to solve it inside the body with (3) as the B.C.) To insulate the
body well, what should be the scaling relationship of k2 and δ? Hint: start with (2); fix a point x on ∂Ω1 ,
and define f (τ ) = u2 (x + τ n). Then perform a Taylor expansion of f at 0.
(2)
X 00 (x) + λX(x) = 0,
(
0<x<l
X(x) is a periodic function with period l
where h is a nonzero constant that may not be positive. Note that negative eigenvalues may appear. In
this case, what is the behavior of the solution for the corresponding initial-boundary value problem for
the homogeneous heat equation?
u(x, 0, t) = u(x, b, t) = 0
u(x, y, 0) = xy
41
Reference answer:
∞ X
∞ 2
m2
X
m+n 4ab n 2 2 nπx mπy
u(x, y, t) = (−1) 2
exp − 2
+ 2 π k t · sin sin .
m=1 n=1
mnπ a b a b
Find the limit of u(x, t) as t → ∞ by inspecting the general solution formula obtained by separation of
variables. (You do not need to compute all the Fourier coefficients.) Interpret your result physically;
generalize it, without proof, to the case of general initial value and higher spatial dimensions.
9. (Black-Scholes equation)
Consider the terminal value problem for the Black-Scholes equation
(
∂V 1 2 2 ∂2V ∂V
∂t + 2 σ S ∂S 2 + rS ∂S − rV = 0, S > 0, 0 < t < T,
(5)
V (S, T ) = φ(S), S > 0,
where S is the price of a stock (as independent variable), V the call option value (as the dependent
variable), σ the volatility of the stock, r the risk-free interest rate, T the expiration day of the option.
This homework is designed to show that this terminal value problem can be transformed to the Cauchy
problem for the heat equation and therefore (5) can be solved explicitly.
(a) Introduce new variables
S = Kex , t = T − τ /(σ 2 /2),
where the constant K is the striking price. Let v(x, τ ) = V (S, t). Show that
∂v ∂2v 2r ∂v 2r
= + ( 2 − 1) − v.
∂τ ∂x2 σ ∂x σ 2
where φ ≥ 0 but is not identically equal to zero on [0, l], satisfying φ00 < 0 on (0, l).
(a) Prove that that u(x, t) > 0 for (x, t) ∈ (0, l) × (0, ∞). Hint: first use the weak minimum principle and
then the strong minimum principle.
(b) Prove that ut (x, t) < 0 for (x, t) ∈ (0, l) × (0, ∞). Hint: Let w = ut ; first find the initial-boundary
problem that w solves, then apply the maximum principles to w.
(c) Draw the graph of u vs x and put arrows on the graph to indicate the behavior of the graph as t
increases. Can you predict the behavior without proving (b) rigorously? What if the initial value changes
its concavity?
11. Consider the solution (3.7.8) of the Cauchy problem (3.7.7). If φ is bounded on R and has a jump
discontinuity at point x, prove that
1
lim u(x, t) = (φ(x − 0) + φ(x + 0)).
t→0+ 2
Prove that u is identically equal to zero on [0,l] for all t ≤ 0. To this end, recall the energy that we have
defined before
Z l
E(t) = u2 (x, t)dx.
0
We just need to prove that E(t) is identically equal to zero for all t < 0. We argue by contradiction by
assuming that there exists t0 < 0 such that E(t0 ) > 0. By continuity, there exists a t1 ∈ (t0 , 0] such that
E is positive on [t0 , t1 ) and is equal to zero at t1 . Without loss of generality, assume t1 = 0. Now proceed
as follows
(a) Prove
Z l
E 00 (t) = 4a4 u2xx (x, t)dx.
0
43
(E 0 )2 ≤ E E 00 , t ∈ [t0 , 0).
(e) Prove that (d) contradicts the assumption that u(x, 0) ≡ 0 (E(0) = 0).
∆u = 0 ,
∂2u ∂2u
+ 2 =0 ( in 2 − D ),
∂x2 ∂y
(4.1.1)
∂2u ∂2u ∂2u
+ 2 + 2 =0 ( in 3 − D ).
∂x2 ∂y ∂z
A solution of the Laplace equation is called a harmonic function. The inhomogeneous version of Laplace
equation
−∆u = f , (4.1.2)
where u(x, y) and v(x, y) are real-valued functions. An analytic function is one that is expressible as a
power series in z, i.e.
+∞
X
f (z) = an z n ,
n=0
which are called the Cauchy-Riemann equations. If we differentiate them, we find that
So that ∆u = 0. Similarly we have ∆v = 0. Thus the real and imaginary parts of an analytic function are
harmonic.
Irrotational and incompressible fluid
Consider a fluid that is irrotational, that is, its velocity vector field v(x) (x ∈ R3 ) satisfies
curl v = 0.
By multivariate calculus, if the region occupied by the fluid has no holes, then the velocity vector field has
a potential φ: ∇φ = v. Suppose the fluid is also incompressible, that is,
∇ · v = 0.
∆φ = 0.
Electrostatics
Suppose at point x0 there is a point charge qx0 . By Coulomb’s law, the electric force exerted by qx0 on a
positive unit point charge located at point x is
qx0 (x − x0 )
Ex0 (x) = ,
4π|x − x0 |3
where we define the unit of charges so that the Coulomb constant is just 1/(4π). The electric potential
induced by the charge qx0 is a scalar function Ux0 of x such that
It can also be verified that the “electric field ”Ex0 is divergence-free, hence the electric potential function
satisfies the Laplace equation
∆Ux0 (x) = 0, x 6= x0 .
Now consider the scenario of a continuous distribution of charges inside a dielectric material Ω. Sup-
pose the density function of charges is f (x) (unit: charges/volume). Then the amount of charge in an
infinitesimal volume element dVx0 at x0 is qx0 = f (x0 )dVx0 . Treating these volume elements as point
charges, by the superposition principle we have that the induced electric potential is
Z
f (x0 )
V (x) = dVx0
Ω 4π|x − x0 |
where the integral is with respect to x0 . If f is smooth, then one can prove (very hard!) that V satisfies
Poisson equation
−∆V (x) = f (x), x ∈ Ω. (4.1.3)
This is also true outside Ω if we interpret that the density function f is zero outside Ω.
Solution
We use a trial solution in the form of
u(x, y) = X(x)Y (y),
and take care of the PDE, and the B.C. at the lateral boundary and at the bottom of the square first
(because there, the B.C. is homogeneous). By the PDE, we have
and so
X 00 (x) Y 00 (y)
=− = Const. = −λ.
X(x) Y (y)
X satisfies the Neumann B.C. X 0 (0) = 0 = X 0 (1). Thus
λn = n2 π 2 , Xn (x) = cos(nπx), n = 0, 1, · · · .
On the other hand, Y satisfies B.C. Y (0) = 0 (again, do not worry about the B.C. at y = 1 at this
moment). The general solution for the Y −ODE is given by Yn (y) = aenπy + be−nπy for n = 1, 2, · · · ; and
Y0 (y) = ay + b for n = 0. By B.C. Y (0) = 0, we have
Now we form
∞
X ∞
X
u(x, y) = un (x, y) = a0 y + an (enπy − e−nπy ) cos(nπx),
n=0 n=1
46
which satisfies the PDE, and the B.C. at the lateral boundary and the bottom of the square. To satisfy
the B.C. at the top of the square, we demand
∞
X
a0 + an (enπ − e−nπ ) cos(nπx) = x,
n=1
Z 1
2
an (enπ − e−nπ ) = 2 x cos(nπx) dx = (cos(nπ) − 1), n = 1, 2, · · · .
0 n2 π 2
Thus
∞
1 X 2((−1)n − 1) enπy − e−nπy
u(x, y) = y+ cos(nπx).
2 n=1
n2 π 2 enπ − e−nπ
x = r cos θ
y = r sin θ
The Laplace operator has the form (See Appendix 3.1 for proof.) :
∂2 1 ∂ 1 ∂2
∆= + + (4.2.1)
∂r2 r ∂r r2 ∂θ2
z
6
* (x, y, z)
r
z
θ
P -y
ψ P
P
47
x = r sin θ cos ψ
y = r sin θ sin ψ
z = r cos θ.
The Laplace operator has the form (See Appendix 3.1 for proof.) :
∂2 ∂2
2 ∂ 1 ∂ ∂ 1
∆= 2 + + 2 sin θ + 2 2
∂r r ∂r r sin θ ∂θ ∂θ r sin θ ∂ψ 2
Separation of variables
We assume that
The solutions of the ODE (4.2.2) are exponential functions (for λ < 0), linear functions (for λ = 0) and
trigonometric functions (for λ > 0). Because of the periodicity, one must have
eigenvalues : λ n = n2 n = 0, 1, 2, · · ·
eigenfunctions : Θn = An cos(nθ) + Bn sin(nθ) n = 0, 1, 2, · · · .
R0 (r) = C0 + D0 ln r.
When λ = λn = n2 > 0, we assume that R(r) = rk . Substituting this into equation (4.2.3), we have
k(k − 1) + k − n2 = 0, or k = ±n.
That means
R(r) = Cn rn + Dn r−n .
48
Now we see that the general solution of the Laplace equation in polar coordinates has the following series
form:
+∞
X
Cn rn + Dn r−n (An cos(nθ) + Bn sin(nθ)) .
u(r, θ) = C0 + D0 ln r + (4.2.4)
n=1
Remark
From formula (4.2.4) it follows that the following functions are special harmonic functions:
1, ln r
rn cos(nθ), rn sin(nθ) n = 1, 2, · · ·
r−n cos(nθ), r−n sin(nθ) n = 1, 2, · · · .
The first two harmonic functions, which are functions of r only, may be derived directly from the Laplace
equation. Let
u(r, θ) = R(r).
Then ∆u = 0 leads to ODE
1
R00 + R0 = 0.
r
The solution is clearly
u(r, θ) = R(r) = a + b ln r.
1
The constant function is a trivial solution of the Laplace equation. The singular function − ln r, in fact,
2π
plays an important role in solving the Poisson equation, which will be studied in Sections 4.3 and 4.4.
The solution u(r, θ) must be continuous, and then bounded, at r = 0. Therefore the coefficients Dn ,
including D0 , in (4.2.4) must vanish. The solution of (4.2.5) may be written as
+∞
a0 X n
u(r, θ) = + r (an cos(nθ) + bn sin(nθ)) , (4.2.6)
2 n=1
That means the coefficients {an , bn } must be the Fourier coefficients of given function φ(θ) i.e.
1 2π
Z
an = φ(ξ) cos(nξ) dξ n = 0, 1, 2, · · · (4.2.7a)
π 0
1 2π
Z
bn = φ(ξ) sin(nξ) dξ n = 1, 2, 3, · · · . (4.2.7b)
π 0
49
Example 4.2.2
(
∆u = 0 a < r < b, 0 ≤ θ < 2π
u(a, θ) = 0, u(b, θ) = f (θ).
Solution
This is a problem in the interior of an annulus. The general solution (4.2.4) and boundary condition at
r=a:
+∞
X
Cn an + Dn a−n (An cos(nθ) + Bn sin(nθ)) = 0
u(a, θ) = C0 + D0 ln a +
n=1
lead to
C0 + D0 ln a = 0
Cn an + Dn a−n = 0.
Let C0 = −D0 ln a and Cn an = −Dn a−n . The solution is written as
+∞
r X r n r −n
u(r, θ) = D0 ln + − (an cos(nθ) + bn sin(nθ)) .
a n=1 a a
Thus we have
Z 2π
1 b
D0 = f (ξ) dξ (µ = )
2π ln µ 0 a
Z 2π
1
an = f (ξ) cos(nξ) dξ
π(µn − µ−n ) 0
Z 2π
1
bn = f (ξ) sin(nξ) dξ.
π(µ − µ−n )
n
0
2
Example 4.2.3
∆u = 0
r > a, 0 ≤ θ < 2π
∂u
= 0, lim (u(r, θ) − V0 r cos θ) = 0.
∂r r=a r→+∞
Solution
Notice that the boundary condition at infinity is inhomogeneous. Let
According to the homogeneous boundary condition at infinity, the coefficients D0 and Cn , including C0 ,
in (4.2.4) are zero. Then the solution of this problem should have the form
+∞
X
w(r, θ) = C0 + r−n (An cos(nθ) + Bn sin(nθ)) ,
n=1
and
+∞
∂w X
= − n r−(n+1) (An cos(nθ) + Bn sin(nθ)) .
∂r n=1
Comparing both sides of this equation we see that all of the coefficients An and Bn are zero except A1 .
The coefficient A1 is determined by −a−2 A1 = −V0 , i.e. A1 = a2 V0 . Thus we have
Example 4.2.4
Let us solve the following boundary value problem for a Poisson equation
(
∆u = cos θ 1 ≤ r ≤ 2, 0 ≤ θ < 2π
u|r=1 = 0, u|r=2 = 2.
Solution
Unlike in the previous two examples, u is no longer harmonic on the region and hence the formula (4.2.4)
for the generation solution does not work for this problem. What we can do now is to use the idea used
before for solving inhomogeneous heat equations: for each fixed r ∈ [1, 2], think of u(r, θ) as an 2π-periodic
function of θ, and expand it into a Fourier series involving An cos nθ + Bn sin nθ (n = 0, 1, · · ·):
∞
X
u(r, θ) = A0 (r) + (An (r) cos nθ + Bn (r) sin nθ).
n=1
(The coefficients An and Bn may vary when r varies and hence in general, they are functions of r. ) Now
the formula (4.2.1), the PDE and the above expansion for u imply
∞
A00 (r) X A0n (r) n2 An (r) Bn0 (r) n2 Bn (r)
A000 (r) + + 00
(An (r) + − 00
) cos nθ + (Bn (r) + − ) sin nθ = cos θ.
r n=1
r r2 r r2
Comparing both sides, we have that all the An = 0 = Bn , except A0 and A1 which must satisfy
A00 (r) A0 (r) A1 (r)
A000 (r) + = 0; A001 (r) + 1 − = 1.
r r r2
Multiplying the A0 -equation by r, we have
For the A1 -equation, by guessing, we have a particular solution r2 /3; the general solution for the corre-
sponding homogeneous equation
A0 (r) A1 (r)
A001 (r) + 1 − =0
r r2
is given by C1 r + C2 r−1 . Thus for the original inhomogeneous equation
the general solution is given by C1 r + C2 r−1 + r2 /3. Now we see that u must be in the form of
and
C1 + C2 + 1/3 = 0
C1 + C2 /2 + 4/3 = 0.
Poisson’s formula
Substituting (4.2.7) into (4.2.6), we have
+∞
a0 X n
u(r, θ) = + r (an cos(nθ) + bn sin(nθ))
2 n=1
+∞
1 2π
Z Z 2π
X 1
= φ(ξ)dξ + rn φ(ξ) cos(nξ)dξ cos(nθ)
2π 0 n=1
π 0
Z 2π
1
+ φ(ξ) sin(nξ)dξ sin(nθ)
π 0
+∞
!
1 2π 1 2π
Z Z X
= φ(ξ)dξ + φ(ξ) rn cos n(ξ − θ) dξ
2π 0 π 0 n=1
Z 2π +∞
!
1 X
= φ(ξ) 1 + 2 rn cos n(ξ − θ) dξ
2π 0 n=1
+∞
X
1+2 rn cos n(ξ − θ)
n=1
+∞
X +∞
X
= 1+ rn ein(ξ−θ) + rn e−in(ξ−θ) ( i2 = −1)
n=1 n=1
+∞ +∞
X n X n
= 1+ rei(ξ−θ) + re−i(ξ−θ)
n=1 n=1
e in(ξ−θ)
e−in(ξ−θ)
= 1+ +
1 − rein(ξ−θ) 1 − re−in(ξ−θ)
1 − r2
= .
1 − 2r cos(ξ − θ) + r2
Finally we get the so-called Poisson’s formula or Poisson’s integral for the solution of Laplace equation
in the unit circle:
Z 2π
1 1 − r2
u(r, θ) = φ(ξ) dξ. (4.2.8)
2π 0 1 − 2r cos(ξ − θ) + r2
Let ρ = Rr (a > 0). The above formula leads to the Poisson’s formula for the solution of Laplace equation
in the circle with radius R:
Z 2π
1 R 2 − ρ2
u(ρ, θ) = φ(ξ) 2 dξ, (4.2.9)
2π 0 R − 2Rρ cos(ξ − θ) + ρ2
where 0 ≤ ρ < R, 0 ≤ θ < 2π. This formula has theoretical value and is beautiful; but you will have a
very hard time computing it when, say, φ(θ) = cos θ, while by the separation of variables method, you can
easily get the solution u = r cos θ = x of (4.2.5) with B.C u(1, θ) = cos θ. Thus when solving boundary
value problem (4.2.5), the first thing to try is still the separation of variables method!
1◦ δ(M ; M0 ) = 0 ∀ M ∈ Rn and M 6= Mo
2◦ δ(M0 ; M0 ) = +∞
Z
3◦ φ(M )δ(M ; M0 )dM = φ(M0 ), ∀ bounded and continuous function φ defined on Rn .
Rn
Observe that δ(M ; M0 ) = δ(M − M0 ; 0). So it makes sense to write δ(M ; M0 ) = δ(M − M0 ).
If we take φ ≡ 1, then Z
δ(M − M0 )dM = 1.
Rn
But since δ is everywhere equal to zero except at M0 , the integral must be zero (recall the value of a
function at one point does not influence the value of its integral). We have a contradiction! Thus, strictly
speaking the δ-function is not a function. In fact, it is a “functional”: it acts on a function φ (as the input)
and the output is a number given by φ(M0 ). This is the point of view taken in advanced mathematics
53
courses. Here in this course, for the sake of simplicity, we still treat δ-function as a function. This practice,
as was intended by its inventor the physicist Dirac, produces correct results as long as we restrain ourselves
from doing wild things such as squaring the δ-function.
The mysterious δ-function can be approximated by “earthly”, i.e. ordinary functions, if they are concen-
trated at a point to form spikes:
Theorem 4.3.1
Let {fm (x)} be a sequence of functions satisfying that
(i) there exists a constant K such that for every m,
Z
|fm (x)|dx ≤ K;
Rn
Then the sequence {fm (x)} converges to Aδ(x − x0 ) in the following weak sense:
Z Z
lim fm (x)φ(x)dx = Aφ(x0 ) = Aφ(x)δ(x − x0 )dx,
m→∞ Rn Rn
for any φ which is bounded and continuous on Rn . In fact, we only need the boundedness φ on Rn and
its continuity at point x0 .
Proof
We just need to show Z
lim fm (x)(φ(x) − φ(x0 ))dx = 0. (4.3.1)
m→∞ Rn
Since φ is continuous at x0 , for any > 0, there exists r > 0 such that |φ(x) − φ(x0 )| < if |x − x0 | ≤ r.
Then R R
| |x−x0 |≤r fm (x)(φ(x) − φ(x0 ))dx| ≤ |x−x0 |≤r |fm (x)||φ(x) − φ(x0 )|dx
R
≤ |x−x0 |≤r |fm (x)|dx
R (4.3.2)
≤ Rn |fm (x)|dx
= K.
On the other hand, let L be an upper bound of |φ|. Because of the concentration assumption, there exists
M such that if m ≥ M , we have Z
|fm (x)|dx < .
|x−x0 |>r
Example 4.3.1
In Section 3.7, the fundamental solution of heat equation
(ξ − ξ0 )2
1
G(ξ, t; ξ0 ) = √ exp − (t > 0)
2a πt 4a2 t
is introduced. Property 3◦ there implies conditions (i) and (iii) with x0 = ξ0 . Now for any fixed r > 0,
Z Z
1 −|η|2
G(ξ, t; ξ0 )dξ = √ √ e dη,
|ξ−ξ0 |>r π |η|>r/(2a t)
which converges to zero as t → 0+ . Thus condition (ii) is also satisfied and Theorem 4.33.1 implies that
G(ξ, t; ξ0 ) converges to δ(ξ − ξ0 ) weakly as t → 0+ . Recall this is proved in Appendix 3.2 where the
arguments are similar to the ones in the proof of Theorem 4.3.1.
G(ξ, t; ξ0 )
1 6
√
2a πt
-ξ
ξ0
Fundamental solution
The solution G0 (x) of the following equation:
The rigorous treatment is a nice application of Theorem 4.3.1 and will make the future proofs (such as
Property 2◦ in Section 4.3) involving the fundamental solution rigorous. Let us consider the 2D case first.
Notice that δ(x) is radially symmetric about the origin. Then it is reasonable to seek radially symmetric
solution:
G0 (x) = F (r), r = |x|.
Then we have
d2 F
1 dF 1 d dF
∆G0 (x) = + = r .
dr2 r dr r dr dr
(4.3.3) leads to
d dF
r = −r δ(x),
dr dr
Z r
dF
r = − r δ(x)dr
dr 0 Z
1 2π r
Z
=− δ(x)r · drdθ
2π Z0 Z 0
1
=− δ(x) dx
2π Z Z|x|<r
1
=− δ(x) dx
2π R2
1
=− .
2π
1
Therefore we have the ODE dF/dr = −1/2πr and its solution F (r) = − ln r +C, where C is a constant.
2π
Since constant C is a trivial solution of Laplace equation ∆u = 0, we simply take C = 0. Finally we get
that
1
G0 (x) = − ln |x|. (4.3.5).
2π
∂ 2 G 1 ∂G 22 C
−∆G = −( 2
+ )=− 2 .
∂r r ∂r (r + 2 )2
We want to choose C so that the righthand side converges to δ(x) weakly and (4.3.4) holds. The integral
of the righthand side on R2 is equal to
Z 2π Z ∞
22 C
− 2 2 2
rdrdθ = −2πC,
0 0 (r + )
so conditions (i) and (iii) in Theorem 4.3.1 are satisfied with A = −2πC. To verify condition (ii), take a
fixed r0 > 0, we estimate Z Z ∞
|22 C| 4π2 |C|
(r 2 +2 )2 dx = (r 2 +2 )2 rdr
|x|>r0 r0 Z ∞
2 1
≤ 4π |C| r 4 rdr
r0
56
Thus C = −1/(2π). By the last sentence in the statement of Theorem 4.3.1, we actually have (4.3.4).
where ωn is the surface area of the unit sphere in Rn . We can use either of the two methods presented
above to obtain these formulas; (4.3.4) holds if the regularization is defined by
1
G (x) = . (4.3.8)
(n − 2)ωn (|x|2 + 2 )(n−2)/2
G0 (x − x0 ).
Formally, by the chain rule we have
Rigorously, in (4.3.4) we replace φ(x) by φ(x + x0 ) and then change variable y = x + x0 , we are led to
Z
lim+ (−∆G )(y − x0 )φ(y)dy = φ(x0 ) for any φ that is bounded on Rn and continuous at x0 .
→0 Rn
(4.3.10)
Notice that in the 3D case, the fundamental solution with singularity at x0 is exactly the electric potential
induced by a unit point charge located at x0 (see Section 4.1).
5 · (v 5 u) = 5v · 5u + v∆u.
Integrating this equation on a bounded region Ω with piecewise smooth boundary ∂Ω, we have
57
Z Z Z Z Z Z
5 ·(v 5 u)dA = 5 v · 5udA + v∆udA (4.4.1)
Ω Ω Ω
where F is a smooth vector field in Ω and n is the unit outward normal vector field on ∂Ω. Using this in
(4.4.1) we obtain
Z Z Z Z I I
∂u
5 v · 5udA + v∆udA = (v 5 u) · ndS = v dS.
Ω Ω ∂Ω ∂Ω ∂n
Z Z Z Z I
∂u
v∆udA = − 5 v · 5udA + v dS , (4.4.2)
Ω Ω ∂Ω ∂n
which is called the Green’s first identity or first Green’s formula. What we are doing here is to shift
the “burden”(the derivative) on u to v; in doing so, the boundary integral is a necessary evil. The reader
may benefit from thinking about the simple case where Ω is an interval [a, b]: by integration by parts, we
have Z b Z Z b b
vu00 dx = vdu0 = − v 0 u0 dx + vu0 (b) − vu0 (a),
a a a
where the last two terms can be regarded as an integral on two boundary points a, b. Because of this, in
PDE literature, (4.4.2) is called integration by parts.
From the Green’s first identity, one can easily get that
Z Z I
∂v ∂u
(u∆v − v∆u)dA = (u −v )dS , (4.4.3)
Ω ∂Ω ∂n ∂n
∆u = 0 in a bounded region Ω.
Property 1◦
I
∂u
dS = 0. (4.4.4)
∂Ω ∂n
It can be easily proved by taking v(x, y) = 1 in either Green’s first identity or Green’s second identity.
This property is a necessary condition for a harmonic function and often used to validate the boundary
conditions for Laplace equation.
Property 2◦
58
I
∂u ∂G0
u(M0 ) = (G0 (M ) − u(M ) )dSM , (4.4.5)
∂Ω ∂n ∂n
I
∂u ∂G0
(G0 (M ) − u(M ) )dSM
∂Ω ∂n ∂n
Z Z
= (G0 ∆u(M ) − u(M )∆G0 )dAM ( take v = G0 )
Ω
Z Z
= u δ(M − M0 )dAM ( − ∆G0 = δ(M − M0 ) and ∆u = 0 )
Ω
Sending to zero, the left hand side becomes the one with replaced by 0 because the the singularity is
away from the boundary; the right hand side can be re-written as the same integral on the entire space
Rn by defining u ≡ 0 outside Ω, then by (4.3.10), the right hand side of (4.4.6) converges to u(M0 ). 2
Remark 1. (4.4.5) is called the boundary representation of harmonic functions. Since for M ∈ ∂Ω,
G0 (M − M0 ) and ∂G∂n (M − M0 ) are infinite smooth functions of M0 ∈ interior of Ω, by the boundary
0
−∆u(M ) = f (M ), M ∈ Ω,
Property 3◦
I
1
u(M0 ) = u dS (2 − D),
2πR ∂B(M0 ,R)
I (4.4.7)
1
u(M0 ) = u dS (3 − D),
4πR2 ∂B(M0 ,R)
59
n
M b
*
ρ
b
R
M0
We start with (4.4.5) with Ω replaced by B(M0 , R). In the 2D case, the fundamental solution is given by
(4.3.5). Let ρ = |M − M0 |. We compute
∂ ∂ 1 1
(ln ρ) = (ln ρ) = = .
∂n ∂B(M0 ,R) ∂ρ ∂B(M0 ,R) ρ ∂B(M0 ,R) R
I I I
1 1 ∂u 1 ln R ∂u
u(M0 ) = u − (ln R) dS = u dS − dS
2π ∂B(M0 ,R) R ∂n 2πR ∂B(M0 ,R) 2π ∂B(M0 ,R) ∂n
where the right hand sides are the averages of u on the disk/ball. This can be proved, in the 2D case for
example, by replacing every R by r < R, by multiplying (4.4.7) by 2πr, and then integrating in r:
Z R Z RI
u(M0 )2πr dr = u dS dr.
0 0 ∂B(M0 ,r)
The left hand side is equal to πR2 u(M0 ); the right hand side is equal to
RR
B(M0 ,R)
u dA because the
length element dS on ∂B(M0 , r) times dr is just the area element dA on B(M0 , R).
Remark 3. If u is not harmonic but satisfies the Poisson equation with a source term
−∆u(M ) = f (M ) ≥ 0, M ∈ Ω,
60
then by modifying the proof of (4.4.7) and using (4.4.5’) (the details are supplied at the end of this remark),
we have I
1
u(M0 ) ≥ u dS (2 − D),
2πR ∂B(M0 ,R)
I (4.4.70 )
1
u(M0 ) ≥ u dS (3 − D),
4πR2 ∂B(M0 ,R)
and Z
1
u(M0 ) ≥ u dA (2 − D),
πR2 B(M0 ,R)
Z (4.4.80 )
1
u(M0 ) ≥ u dV (3 − D).
4πR3 /3 B(M0 ,R)
Of course, if f ≤ 0, then we just reverse the direction of the inequality sign “≥”.
Now we give the details of the proof of (4.4.70 ) in the 2D case. By (4.4.50 ) and the proof of Property 3,
we see
I I Z
1 ln R ∂u 1
u(M0 ) = u dS − dS − ln |M − M0 |f (M ) dAM .
2πR ∂B(M0 ,R) 2π ∂B(M0 ,R) ∂n 2π B(M0 ,R)
To rewrite the middle integral on the right hand side, we take in Green’s first identity v = 1 and then use
the PDE for u. Then
I Z Z Z Z
∂u
dS = ∆u(M ) dAM = − f (M ) dAM .
∂B(M0 ,R) ∂n B(M0 ,R) B(M0 ,R)
Thus I Z
1 1
u(M0 ) = u dS + (ln R − ln |M − M0 |)f (M ) dAM .
2πR ∂B(M0 ,R) 2π B(M0 ,R)
Since the integrand in the last integral is nonnegative, we complete the proof of (4.4.70 ) in the 2D case.
As in the case of the heat equation, three kinds of boundary conditions can be prescribed:
∂u
(x) = ψ(x), ∀ x ∈ ∂Ω − − − − Neumann boundary condition,
∂n
∂u
(x) + hu(x) = µ(x), ∀ x ∈ ∂Ω − − − − Robin boundary condition, (h>0)
∂n
Theorem 4.4.1 The solution of Poisson equation (4.4.9) with Dirichlet or Robin boundary condition
is unique; with Neumann boundary condition, the solution is unique up to an additive constant.
Proof.
Since the Poisson equation and the three boundary conditions are all linear, it is only required to prove
that the following problem
∆u = 0 in Ω
homogenuous B.C. on ∂Ω
61
has unique solution u(x) = 0 for Dirichlet or Robin boundary condition, and u(x) = constant for Neumann
boundary condition.
The PDE ∆u = 0 and Green’s first identity lead to
Z I
∂u
5 u · 5u dV = u dS.
Ω ∂Ω ∂n
∂u
For the homogeneous Dirichlet B.C. (u = 0) or Neumann B.C. ( = 0), we have
∂n
Z I
∂u
5 u · 5u dV = u dS = 0.
Ω ∂Ω ∂n
∂u
For the homogeneous Robin boundary condition ( + hu = 0), we have
∂n
Z I I
∂u
5 u · 5u dΩ = u dS = − hu2 dS ≤ 0.
Ω ∂Ω ∂n ∂Ω
Since h > 0, one must have hu2 ≥ 0. So for all three cases, we have
Z
5 u · 5u dV = 0.
Ω
5u ≡ 0, i.e. u = constant.
In the Dirichlet case, u is zero on the boundary and so the constant function must be identically equal
to zero; in the Robin case, since u is constant, the B.C. also implies that u is zero on the boundary and
hence in the interior. 2
Dirichlet Principle
Again, let Ω be bounded region in Rn with piecewise smooth boundary. For any smooth function v defined
on Ω, motivated by electrostatic considerations, we define the energy of v as
Z
1
E(v) = |∇v|2 dV.
2 Ω
Consider the Dirichlet boundary value problem for the Laplace equation
(
∆u = 0 in Ω
(4.4.10)
u=φ on ∂Ω.
Theorem 4.4.2 (Dirichlet Principle) Let u be the unique smooth solution of (4.4.10). Let Γφ be the
set of all smooth functions v defined on Ω satisfying the boundary condition i.e.
v=φ on ∂Ω.
E(v) ≥ E(u) ∀ v ∈ Γφ .
Conversely, any minimizer u of E in Γφ is a solution of the Dirichlet boundary value problem (4.4.10).
Proof
62
Suppose u is a solution of (4.4.10). Let w = v − u. Then w vanishes on the boundary, i.e. w|∂Ω = 0. We
compute
Z Z
1
E(v) = E(u + w) = | 5 (u + w)|2 dV = E(u) + 5 u · 5w dV + E(w).
2 Ω Ω
If there exists a point x0 ∈ Ω where ∆u 6= 0, say, ∆u(x0 ) > 0, then by continuity, ∆u > 0 in a small ball
B(x0 , ). Pick a smooth function w that is positive inside the ball, equal to zero outside it. Then
Z Z
w∆u dV = w∆u dV > 0,
Ω B(x0 ,)
Remark
The proof of Theorem 4.4.2 is simple, but this is an important mathematical theorem based on the
physical idea of energy. It is a general principle in physics that any system prefers to approach the state
of lowest energy, called the ground state. Theorem 4.4.2 is a mathematical manifestation of this physical
principle.
Equation (4.4.11) characterizes harmonic functions, that is, it is a necessary and sufficient condition for
u to be harmonic. In advanced PDE courses, it is called the weak formulation of harmonic functions;
any function that satisfies (4.4.11), even those that we do not know their smoothness, is called a weak
solution of the Laplace equation. It can be proved a weak solution is necessarily a smooth solution.
where f is a smooth function with bounded support (meaning that f ≡ 0 outside a large ball). Motivated
by the electrostatic consideration in Section 4.1, we conjecture that
Z
u(x) = G0 (x − x0 )f (x0 )dVx0 (4.4.13)
Rn
is a solution.
Formal proof
Recall from (4.3.9)
−∆G0 (x − x0 ) = δ(x − x0 ).
We compute R
−∆u(x) = Rn (−∆G0 )(x − x0 )f (x0 )dVx0
R
= Rn δ(x − x0 )f (x0 )dVx0
R
= Rn δ(x0 − x)f (x0 )dVx0
= f (x).
Rigorous proof
Choose a large L such that outside the ball B(0, L), f = 0. Change variable in (4.4.13) by letting
y = x − x0 . Then Z
u(x) = G0 (y)f (x − y)dVy ,
Rn
R
−∆u(x) = Rn G0 (y)(−∆x ) (f (x − y)) dVy
R
= Rn G0 (y)(−∆y ) (f (x − y)) dVy (by chain rule)
R
= B(x,L+1) G0 (y)(−∆y ) (f (x − y)) dVy (by assumption on f )
R
= lim→0+ B(x,L+1) G (y)(−∆y ) (f (x − y)) dVy
R
= lim→0+ B(x,L+1) (−∆y )G (y)f (x − y)dVy +
+ lim→0+ ∂B(x,L+1) ∂G∂n (y)
f (x − y) − G (y) ∂f (x−y)
H
∂n dS ( Green second formula)
R
= lim→0+ B(x,L+1) (−∆y )G (y)f (x − y)dVy (f terms in boundary integral = 0)
R
= lim→0+ Rn (−∆y )G (y)f (x − y)dVy
= f (x) (by (4.3.10)).
where f is the density function of charges continuously distributed in the space, and u is the electric
potential thus induced.
We start with the simplest case: suppose −u00 ≤ 0 on an interval I. Then u is concave up. If u achieves
its maximum value at an interior point of I, then u must be a constant function. In general, we have the
following
Theorem 4.5.1 (Strong maximum principle) Suppose in a connected region Ω function u satisfies
−∆u ≤ 0, (4.5.1)
64
then the maximum function value is taken only on the boundary unless u is a constant function.
Proof.
We just need to show that if u(M0 ) reaches the maximum function value umax and M0 is an interior point
of Ω, one must have
# Γ -+
# c c c c
c
c M∗
"! M M3 Mk−1
"! M1 2
M0
Since Ω is a connected region, one can find a continuous curve Γ ⊂ Ω connecting M0 and M ∗ . Since M0
is an interior point in Ω, one can find a disk B(M0 , R0 ) ⊂ Ω. According to (4.4.8’), in the 2D case (the
higher dimensional case can be handled in the same way) one has
Z
1
u(M0 ) ≤ u dA.
πR02 B(M0 ,R0 )
Denote the intersection of curve Γ and circle ∂B(M0 , R0 ) along the positive direction by M1 . Then we
have M1 ∈ Ω and u(M1 ) = umax . The above argument can be iteratively repeated, leading to a series of
interior points {M0 , M1 , M2 , · · ·} such that
Eventually we reach a finite number k such that M ∗ falls into the disk B(Mk , Rk ). This completes the
proof. 2
Remark 1. There is another proof of this result: observe that u also satisfies (3.8.2) so the parabolic
strong maximum principle applies, which yields the elliptic strong maximum principle above.
Remark 2. If the inequality in (4.5.1) is reversed, then we have the strong minimum principle. If u
is harmonic, then we have both principles.
Remark 3. The strong maximum principle implies the weak maximum principle:
max u = max u.
Ω ∂Ω
The goal of this section is to solve the Dirichlet boundary value problem
∆u = 0, x ∈ Ω,
(4.6.1)
u(x) = φ(x), x ∈ ∂Ω,
where for the time being, we assume that Ω is a bounded region in Rn with piecewise smooth boundary.
Recall we have obtained (4.4.5) from which we have:
I
∂u ∂G0
u(x0 ) = (G0 (x − x0 ) (x) − φ(x) (x − x0 ))dSx .
∂Ω ∂n ∂n
∂u
This almost achieves the goal of solving (4.6.1), except that we do not know ∂n on the boundary. To get
rid of this term on the righthand side, for each fixed x0 in the interior of Ω, we find a function ψ(x; x0 )
which is smooth in x on the closure of Ω, satisfying
∆x ψ(x; x0 ) = 0, x ∈ Ω,
ψ(x; x0 ) = −G0 (x − x0 ), x ∈ ∂Ω.
It satisfies
−∆x G(x; x0 ) = δ(x − x0 ), x ∈ Ω,
(4.6.3)
G(x; x0 ) = 0, x ∈ ∂Ω.
The physical interpretation of Green’s function in the 3−dimensional case is that it is the electric
potential function induced by a unit point-charge located at x0 with ∂Ω grounded (so the potential on
the boundary is zero). ψ in (4.6.2) is called the regular part of Green’s function.
Now repeating the proof of (4.4.5) with G0 replaced by G, we have the solution formula for (4.6.1)
I
∂G
u(x0 ) = − φ(x) (x; x0 )dSx . (4.6.4)
∂Ω ∂n
Thus solving the Dirichlet problem (4.6.1) boils down to merely finding the Green’s function, which
sounds very promising. But finding Green’s function is equivalent to finding its regular part; and finding
the regular part ψ is still a Dirichlet boundary problem, though with a special boundary value. This is
the reason why for general regions, it is impossible to find explicit formulas for the Green’s functions; in
fact, only for balls and half-spaces, will we be able to find these in this course. The Green’s function for a
rectangle is, surprisingly, not easy to find and has to be given by a series (thus in this case the preferred
method is still the method of separation of variables).
at x0 in the upper space with the plane x3 = 0 grounded (so the potential is 0 on the plane). To
ground the plane, we simply put a negative unit point charge at the mirror image x∗0 of x0 (so x∗0 =
(x01 , x02 , · · · , x0n−1 , −x0n )). In this fashion, we get the Green’s function of the upper half-space
Indeed, G(x; x0 ) is identically equal to zero on the plane xn = 0 and is harmonic on the upper half-space
(note the singularity of G(x; x∗0 ) occurs in the lower half-space). And these hold in case of any spatial
dimensions (recall that G0 is a radial function and note that x0 and x∗0 are equidistance from any point
x on the plane xn = 0); the 2D case is illustrated Figure 4.6.1.
x2
6
xb0 = (x01 , x02 )
@
@
@
@
@
@ x
@b - x1
b
x∗0 = (x01 , − x02 )
Figure 4.6.1 Mirror reflected point x∗0 of x0 for the upper half-plane
∂G
Now let us come back to (4.6.4) by first computing on the plane xn = 0. In the 2D case,
∂n
∂G(x; x0 ) ∂G −x02 1
=− = · . (4.6.7)
∂n x2 =0 ∂x2 x2 =0 π (x1 − x01 )2 + x202
In the 3D case,
∂G(x; x0 ) ∂G x03 1
=− =− ·p 3. (4.6.8)
∂n x3 =0 ∂x3 x3 =0 2π (x1 − x01 )2 + (x2 − x02 )2 + x203
Combining (4.6.4) and (4.6.7) and (4.6.8), for the 2D case we have
y0 +∞
Z
φ(x)
u(x0 , y0 ) = dx, (4.6.9)
π −∞ (x − x0 )2 + y02
for 3D case we have
Z +∞ Z +∞
z0 φ(x, y)
u(x0 , y0 , z0 ) = 3 dxdy. (4.6.10)
2π
p
−∞ −∞ (x − x0 )2 + (y − y0 )2 + z02
67
These two formulas for the solution of the boundary problem (4.6.5) are called Poisson integrals for the
half-spaces.
We now check rigorously that if φ is bounded and continuous, then (4.6.9) and (4.6.10) indeed are solutions
of the boundary value problem (4.6.5) in 2D and 3D cases, respectively. Since the 3D case is similar to
the 2D case, we shall focus on the latter case. By inspecting the explicit formula for the Green’s function,
we note the following symmetry
Because of this and the fact that G(x; x0 ) is harmonic in x-variable, it is also harmonic in x0 -variable,
except at x0 = x. Thus for any fixed x = (x, y) on the x-axis,
∂G(x; x0 ) y0
− =
∂y π((x − x0 )2 + y02 )
is harmonic in x0 = (x0 , y0 )-variable in the upper half-plane {y0 > 0}. Now it follows that
Z +∞
y0
∆x0 u(x0 ) = ∆x 0 [ ]φ(x)dx = 0.
−∞ π((x − x0 )2 + y02 )
Now we check that that u given by (4.6.9) satisfies the boundary condition in (4.6.5) in the sense of
y0
Γ(x; (x0 , y0 )) = ,
π((x − x0 )2 + y02 )
Example 4.6.1
∆u = 0
in Ω = {M = (x, y); − ∞ < x < +∞, y > 0 }
u0 x>0
u(x, 0) =
.
0 x<0
Solution
Using the Poisson’s integral, we find
68
Z +∞
y f (ξ)
u(x, y) = dξ
π −∞ (ξ − x)2 + y 2
Z +∞
y u0
= dξ
π 0 (ξ − x)2 + y 2
+∞
ξ−x
Z
u0 dζ
= set ζ =
π −x/y 1 + ζ2 y
u0 +∞
= tan−1 ζ −x/y
π
u0 2 −1 x
= 1 + tan .
2 π y
2
R R 1 |x0 |
G0 (x − x∗0 ) = = G0 ( (x − x∗0 )). (4.6.14)
|x0 | |x0 | 4π|x − x∗0 | R
So we take the regular part of the Green’s function to be ψ(x; x0 ) = −G0 ( |xR0 | (x − x∗0 )). It is harmonic in
the ball and it is equal to − |xR0 | G0 (x − x∗0 ) = −G0 (x − x0 ) for x ∈ ∂BR (0). Thus the Green’s function is
given by
|x0 |
G(x; x0 ) = G0 (x − x0 ) − G0 ( (x − x∗0 )). (4.6.15)
R
x
bH
H
HH
H
HH
ψ b Hb
H -
O x0 x∗0
BR (0)
69
Figure 4.6.2 Reflected point x∗0 of x0 for Dirichlet problem in the ball BR (0)
It turns out that (4.6.15) is also the formula for the Green’s function for the ball in all spatial dimensions
n ≥ 2: By the chain rule, it is easy to check that the regular part in (4.6.15) is harmonic in the ball; it
remains to check that the right-hand side is 0 for x on the boundary of the ball. By (4.6.12) and R = |x|,
we have
|x0 | |x|
= ∗ .
|x| |x0 |
Thus we have two similar triangles ∆Oxx0 ' ∆Ox∗0 x. This in turn implies that
|x − x0 | |x0 |
= , (4.6.16)
|x − x∗0 | |x|
|x0 |
|x − x0 | = |x − x∗0 |.
R
Then because G0 is a radial function, we obtain
|x0 |
G0 (x − x0 ) = G0 ( (x − x∗0 )).
R
This completes the checking that (4.6.15) indeed is the Green’s function for the ball in any spatial dimen-
sions bigger than 1.
Now we use the Green’s function to solve the Dirichlet boundary value problem:
∆u = 0, x ∈ BR (0),
(4.6.17)
u(x) = φ(x), x ∈ ∂BR (0),
∂G(x; x0 )
− = −∇G(x; x0 ) · n = (−G00 (r)∇r + G00 (ρ)∇ρ) · n (by chain rule)
∂n 1−n
ρ1−n
r x
= ∇r − ∇ρ · (by (4.3.8))
ωn ωn R
r1−n x
= (∇r − ∇ρ) · (for x ∈ ∂BR (0), r = ρ)
ωn R
1−n
|x0 | x − x∗0
r x − x0 x
= − ·
ωn |x − x0 | R |x − x∗0 | R
r1−n x − x0 |x0 |2 x − x∗0 x
= − · (we use (4.6.16))
ωn |x − x0 | R2 |x − x0 | R
r1−n R2 (x − x0 ) · x − |x0 |2 (x − x∗0 ) · x
=
ωn |x − x0 |R3
1−n 2 2
r R − |x0 |
= (use (4.6.13))
ωn |x − x0 |R
2 2
R − |x0 | 1
= .
ωn R |x − x0 |n
R2 − |x0 |2 φ(x)
Z
u(x0 ) = dS (4.6.18)
∂BR (0) ωn R |x − x0 |n
This is called the Poisson formula for the ball BR (0). In the 2D case (n = 2), this formula is just
(4.2.9) if we express every item in polar coordinates.
70
Then
∂ ∂ ∂
= cos θ + sin θ
∂r ∂x ∂y
1 ∂ ∂ ∂
= − sin θ + cos θ
r ∂θ ∂x ∂y
i.e.
∂ ∂ 1 ∂
= cos θ − sin θ
∂x ∂r r ∂θ
(2)
∂ ∂ 1 ∂
= sin θ + cos θ
∂y ∂r r ∂θ
Notice that
∂2
∂ 1 ∂ ∂ 1 ∂
= cos θ − sin θ cos θ − sin θ
∂x2 ∂r r ∂θ ∂r r ∂θ
∂2 1 ∂2
1 ∂
= cos2 θ − cos θ sin θ − +
∂r2 r2 ∂θ r ∂r∂θ
2
∂2
1 ∂ ∂ 1 ∂
− sin θ − sin θ + cos θ + 2 sin θ cos θ + sin θ 2 .
r ∂r ∂r∂θ r ∂θ ∂θ
∂2
∂ 1 ∂ ∂ 1 ∂
= sin θ + cos θ sin θ + cos θ
∂y 2 ∂r r ∂θ ∂r r ∂θ
∂2 1 ∂2
1 ∂
= sin2 θ + sin θ cos θ − +
∂r2 r2 ∂θ r ∂r∂θ
∂2 ∂2
1 ∂ 1 ∂
+ cos θ cos θ + sin θ + 2 cos θ − sin θ + cos θ 2 .
r ∂r ∂r∂θ r ∂θ ∂θ
Finally
∂2 ∂2 ∂2 1 ∂ 1 ∂2
∆= + = + + . (3)
∂x2 ∂y 2 ∂r2 r ∂r r2 ∂θ2
For spherical coordinates, let s = r sin θ. Then the 3-D coordinate transformation becomes a pair of 2-D
transformations:
x = s cos ψ z = r cos θ
and (4)
y = s sin ψ s = r sin θ .
Thus we have
∂2 ∂2 ∂2 1 ∂ 1 ∂2
+ = + + . (5)
∂x2 ∂y 2 ∂s2 s ∂s s2 ∂ψ 2
∂2 ∂2 ∂2 1 ∂ 1 ∂2
+ = + + . (6)
∂z 2 ∂s2 ∂r2 r ∂r r2 ∂θ2
71
∂2 ∂2 ∂2 ∂2 1 ∂ 1 ∂2 1 ∂ 1 ∂2
2
+ 2+ 2 = 2+ + 2 2+ + 2 . (7)
∂x ∂y ∂z ∂r r ∂r r ∂θ s ∂s s ∂ψ 2
∂2 ∂2 ∂2 ∂2 ∂2 ∂2
2 ∂ 1 ∂ 1
+ + = + + 2 sin θ 2 + cos θ + .
∂x2 ∂y 2 ∂z 2 ∂r 2 r ∂r r sin θ ∂θ ∂θ r2 sin2 θ ∂ψ 2
or
∂2 ∂2 ∂2 ∂2 ∂2
2 ∂ 1 ∂ ∂ 1
∆= + + = + + 2 sin θ + . (8)
∂x2 ∂y 2 ∂z 2 ∂r 2 r ∂r r sin θ ∂θ ∂θ r2 sin2 θ ∂ψ 2
Assignment 4
2. A spherical shell with inner radius 1 and outer radius 2 has s steady-state temperature distribution,
∂u
u. Its inner boundary is held at 100◦ C. Its outer boundary satisfies = −γ, where γ is a constant.
∂n
(a) Find the temperature u. Hint: everything is radial and hence so is u.
(b) What are the hottest and coldest temperatures?
(c) Can you choose γ so that the temperature on the outer boundary is 20◦ C?
3. Suppose that u is a harmonic function in disk D = {r < 2} and that u = 3 sin(2θ) + 1 for r = 2.
Without finding the solution, answer the following questions:
(a) Find the maximum value of u in D;
(b) Calculate the value of u at the origin.
4. Find the Green’s function G(M ; M0 ) for Dirichlet problem in the first quadrant of plane:
(
−∆G = δ(M − M0 ) in Ω = {M = (x, y); x > 0, y > 0}
G|x=0 = 0, G|y=0 = 0
72
(2)
(
∆u = 1 in Ω = {M = (r, θ); 0 ≤ r < R, 0 ≤ θ < 2π}
u(R, θ) = 0
Hint: the region and B.C. are radially symmetric and hence the solution should be radially symmetric.
(3)
∆u = A r2 sin(2θ) in Ω = {M = (r, θ); 0 ≤ r < R, 0 ≤ θ < 2π}
2
u(R, θ) = 0
Hint: For each fixed r, u(r, θ) is 2π−periodic function of θ which can be expanded by the eigenfunctions
with 2π-period B.C.. Thus u takes the form of
∞
X
u(r, θ) = A0 (r) + (An (r) cos(nθ) + Bn (r) sin(nθ)).
n=1
(4)
∆u = 0 in Ω = {M = (x, y); 0 < x < π, 0 < y < π}
u(0, y) = 0 u(π, y) = cos2 y
uy (x, 0) = 0 uy (x, π) = 0
(5)
∆u = 0 in Ω = {M = (x, y); 0 < x < a, 0 < y < b}
u(0, y) = 0 u(a, y) = 0
∂u
+u = 0 u(x, b) = g(x)
∂y
y=0
6. Find the solutions that depend only on r of the Helmholtz equation −∆u = λ2 u in 3-D, where
λ > 0 is a constant.
in Ω ⊂ R3
∆u = f
∂u = g on ∂Ω
∂n
unless
Z I
f dx = g dS.
Ω ∂Ω
8. Let Ω be a bounded domain in Rn with smooth boundary. Consider Poisson equation with Neumann
boundary condition
−∆GN (x; x0 ) = δ(x − x0 ), x ∈ Ω,
∂GN
∂n (x; x0 ) = const. C, x ∈ ∂Ω,
73
where n is the unit outer normal of ∂Ω, x0 is a fixed point in Ω. Do this problem formally.
(i) Find the value of const. C such that the above BVP has a solution.
(ii) By using GN , find a formula for u(x0 ), where u is a solution of
−∆u = f (x), x ∈ Ω,
∂u
∂n = g(x), x ∈ ∂Ω.
−∆u = f (x), x ∈ R3 ,
lim u = 0,
|x|→∞
where
1, if |x| ≤ 1,
f (x) =
0, if |x| > 1.
sup u ≤ 2n infn u.
Rn R
Hint: take an arbitrary pair of points P and Q. Let R = |P −Q|. Use the mean value property of harmonic
functions on the balls BR (P ) and B2R (Q) (balls centered at P and Q with radius R and 2R, respectively).
11. (Liouville Theorem) Prove that any harmonic function u in the whole Rn that is either bounded
from below or above must be a constant function. Hint: Consider either u − inf u or sup u − u.
12. (Decay rate of harmonic functions) Suppose u is harmonic in the exterior of the ball BR (0) in
R3 such that it decays at infinity:
lim u(x) = 0.
|x|→∞
(i) Define
v(x) = M G0 (x) − u(x),
where G0 is the fundamental solution of Laplace equation, and the constant M is taken large enough such
that v > 0 on ∂BR (0). Prove that v is positive in the exterior of BR (0). Hint: argue by contradiction and
use the strong minimum principle.
(ii) Prove that u decays at infinity at least as fast as the fundamental solution.
b b - x
0 l
Z lp
1 + u2x (x, t) dx
0
may change as t changes (while the total mass does not change), strictly speaking ρ is a function of t.
But because of the idealization that ∂u
∂x is small, the total length of the string is approximately constant
in t. On the other hand, the string may be inhomogeneous and thus we assume ρ depends on x only. We
introduce the tension T (x) of the string at x, which is the magnitude of the tensile force exerted on the
part of the string on one side of the particle x by the part on the other side. For a string that is “aging
”quickly, T should vary with respect to t appreciably. Thus T may depend on both x and t (however,
as we will see, the transversality of the motion of the string implies that T is independent of x). We
assume that an external force (such as damping force) is exerted on the string in the transverse direction;
to describe this force, we let
T (x1 , t)
x1 x2
Now we take an arbitrary piece of the string between particles x1 and x2 with x1 < x2 . We apply Newton’s
Second Law to this piece. The tensile force exerted to the piece at x2 (by the part of string on the right-
hand side) is tangent to the string and has magnitude T (x2 , t), and so the force is T (x2 , t) times the unit
tangent vector. To get the tangent vector, think of the string at time t being parameterized by the function
75
x 7→ (x, u(x, t)). Then the unit tangent vector that we are seeking for is
(1, u (x , t))
p x 2 ,
1 + u2x (x2 , t)
and so the tensile force at particle x2 at time t is given by
(1, u (x , t))
p x 2 T (x2 , t).
1 + u2x (x2 , t)
Similarly, the tensile force at particle x1 at time t is given by
(1, ux (x1 , t))
−p T (x1 , t).
1 + u2x (x1 , t)
Notice that the external (non-tensile) force exerted on the piece of the string is given by
Z x2
F (x, t) dx(0, 1).
x1
Since the string moves only transversely, the horizontal components of the tensile forces cancel:
T (x2 , t) T (x1 , t)
p =p .
2
1 + ux (x2 , t) 1 + u2x (x1 , t)
Because of our assumption that ux is small, this leads to, approximately,
Dividing both sides by ρ, we have the wave equation in the standard form
where a2 = T /ρ, f (x, t) = F (x, t)/ρ(x). In the rest of this chapter, for simplicity we assume that T and ρ
are constants and hence a is also a constant. The dimension of a is
s
p mass × length/time2
f orce/density = = length/time.
mass/length
76
So the unit of a is that of speed! Indeed, we will see in the future that a is the speed of wave propagation
along the string.
where f (x) is the initial displacement (i.e. the initial shape of string) and g(x) the initial velocity (i.e.
the initial motion of the string).
The boundary conditions for wave equation may be classified into three types:
First kind B.C. (Dirichlet B.C.)
u(0, t) = µ1 (t); u(l, t) = µ2 (t). (5.1.3)
The physical meaning of the first kind B.C. is that the terminal points are forced to move vertically
according to certain formulas µ1 (t) and µ2 (t), respectively. For the simplest case i.e. u(0, t) = 0, the
terminal point x = 0 is fixed.
?
u(0, t)
- x
6
The physical meaning of the Robin B.C. is that the left-end point of the string is put on an elastic
foundation, which is modelled by an elastic spring, and so the external force f1 (t) in the discussion of
77
Neumann B.C. is now given by Hooke’s law: f1 (t) = −ku(0, t), where k > 0 is the spring constant, and
we assume that u = 0 is the equilibrium position of the spring. Now
k
ux (0, t) − u(0, t) = 0,
T (t)
which is the Robin B.C at x = 0. The Robin B.C at x = l can be interpreted in the same fashion.
(x0 (s), y 0 (s), ux (x(s), y(s), t)x0 (s) + uy (x(s), y(s), t)y 0 (s)) ~t + (~t · ∇u)~k
T~ = p =q
0 2 0 2 0
(x ) (s) + (y ) (s) + (ux (x(s), y(s), t)x (s) + uy (x(s), y(s), t)y (s)) 0 2
1 + (∇u · ~t)2
~ = (−∇u(x(s),
N p
y(s), t), 1)
.
1 + |∇u|2
Now we have
0 ~ ~ − (x0 (s) − ux ∇u · ~t)~j + (y 0 (s)ux − x0 (s)uy )~k
~
D = T~ × N ~ = (y (s) − uy ∇u · t)iq
p
1 + (∇u · ~t)2 1 + |∇u|2
≈ (y 0 (s), −x0 (s), y 0 (s)ux − x0 (s)uy ),
where in the last step we drop all the quadratic terms of ∇u (which are much smaller than the linear
terms of ∇u, by our assumption that the motion of the membrane is mild). So the horizontal component
of the tensile force is
Z Z q Z
0 0 0 ~ 2
(y (s), −x (s))T ds = nT 1 + (∇u · t) ds ≈ nT ds
Γ Z ∂D ∂D
= (Tx , Ty )dxdy (Divergence Theorem) = 0
D
because there is no horizontal motion. Since region D is arbitrary, we have (Tx , Ty ) ≡ 0 and so T depends
78
only on t: T = T (t). On the other hand, the vertical component of the tensile force on Γt is
Z Z q
~ · ~kT ds0 =
D (y 0 (s)ux − x0 (s)uy ))T 1 + (∇u · ~t)2 ds
Γt Z ∂D
≈ ∇u · nT (t)ds
Z ∂D
= T (t)∆u dxdy.
D
Suppose the density function of the membrane is ρ(x, y) and the external (non-tensile) force on the mem-
brane is modelled by F (x, y, t) with unit force/area. Then by Newton Second Law, we obtain
Z Z Z
utt ρ dxdy = T (t)∆u dxdy + F dxdy.
D D D
with K(t) and P (t) being kinetic energy and potential energy of the dynamic system, respectively.
For a vibrating string, at the beginning we do not know how to define its potential energy; but we can
easily define its kinetic energy: for a moving particle with mass m, the kinetic energy of the particle is
1
defined as mv 2 , where v is the velocity of the particle. Thus naturally the kinetic energy of a vibrating
2
string should be given by
Z l Z l
1 1
K(t) = (ρdx)(ut )2 = ρu2t dx. (5.2.2)
0 2 2 0
dE dP dK
= + = 0.
dt dt dt
dK
Then the potential energy P (t) may be derived from − mathematically. Suppose that there is
dt
no external force applied to the string, that tension T is constant, and that we have the free boundary
condition at the both ends of the string. Observe
Z l Z l Z l Z l
dK d 1 2
= ρut utt dx = ρut (a2 uxx ) dx = − T utx ux dx = − T u dx.
dt 0 0 0 dt 0 2 x
79
1 l
Z
P (t) = T u2x dx. (5.2.3)
2 0
The total energy is given by
1 l
Z
E(t) = (ρu2t + T u2x ) dx. (5.2.4)
2 0
Since the multiplication of the energy E by a constant will not change the conservation property, the
mathematical version of the total energy is written in the following form
1 l 2
Z
E(t) = (ut + a2 u2x ) dx. (5.2.5)
2 0
Theorem 5.2.1 Consider the wave equation (5.1.1) with initial condition (5.1.2) and one of the bound-
ary conditions (5.1.3-5). The initial-boundary value problem has at most one solution.
Proof.
Because of the linearity of the wave equation and boundary conditions, it suffices to prove that the following
problem
utt = a2 uxx
x ∈ (0, l), t ∈ (−∞, ∞)
u(x, 0) = 0, ut (x, 0) = 0 x ∈ (0, l)
t ∈ (−∞, ∞)
homogenuous B.C. x = 0, l,
dE l
− a2 ut · ux
0
= 0. (5.2.6)
dt
dE
In the case of either homogeneous Dirichlet or Neumann B.C., the boundary term is 0. Thus = 0,
dt
i.e. the total energy E(t) is a constant (conservation of energy). Now the zero initial conditions lead to
E(t) = E(0) = 0. Hence ux ≡ 0 ≡ ut and u ≡ constant. Now by the initial condition u(x, 0) = 0, we have
u ≡ 0.
In the case of homogeneous Robin boundary condition (5.1.5), one has
l ha2 d 2
a2 ut · ux = −ha2 [ut (l, t) · u(l, t) + ut (0, t) · u(0, t)] = − [u (l, t) + u2 (0, t)]
0 2 dt
80
!
Z l
1 d k 2 k
u2t a2 u2x 2
(ha2 =
+ dx + [u (l, t) + u (0, t)] = 0. )
2 dt 0 ρ ρ
which leads to Z l
e =1 u2t + a2 u2x
1k 2
[u (l, t) + u2 (0, t)] = constant.
E(t) dx +
2 0 2ρ
The integral above is the total energy of spring and the boundary part is the energy stored in the elastic
foundation. E(t),
e in fact, is the total energy of the whole system: the vibrating string and the elastic
foundation. By the initial conditions, E(0)
e = 0; hence E(t) ≡ 0, from which it follows that u ≡ constant.
Finally the initial condition u(x, 0) = 0 implies u ≡ 0.
2
The method of separation of variables introduced in Chapters 3 and 4 can be also applied to the hyperbolic
equations. We will not repeat the description for the method, since it is almost the same as that discussed in
Section 3.4-3.6. We select several typical examples and pay more attentions to the physical understanding
of the Fourier series form of the solutions.
Example 5.3.1
A string with length l is fixed at two ends, x = 0 and x = l. The middle point of the string is lifted to
height h and released at time t = 0. Find the subsequent vibration of the string.
Solution
The problem is to solve the following wave equation with boundary and initial conditions:
utt = a2 uxx
0 < x < l, t ∈ (−∞, ∞)
u(0, t) = 0, u(l, t) = 0 t ∈ (−∞, ∞) (5.3.1)
u(x, 0) = φ(x), ut (x, 0) = 0 0 < x < l,
where
2h l
l x
0≤x≤
2
φ(x) =
2h (l − x) l
≤ x ≤ l.
l 2
We start with a trial solution of the form
Based on the result in Section 3.5, the eigenvalues and eigenfunctions are given by
81
nπ 2
λn = n = 1, 2, 3, · · ·
l
nπx
Xn = sin n = 1, 2, 3, · · · .
l
Now the ODE for the T −part is
T 00 (t) + a2 λn T (t) = 0.
Now let
+∞
X
u(x, t) = Tn (t)Xn (x).
n=1
One should notice that the PDE and the boundary condition are already satisfied. So we only need to
take care of the initial conditions.
By the first initial condition, we have
+∞
X
φ(x) = an Xn (x),
n=1
and so Z l
2 nπx
an = φ(x) sin dx
l 0 l
l
!
Z Z l
2 2 2h nπx 2h nπx
= x sin dx + (l − x) sin dx
l 0 l l l
2
l l
8h nπ
= sin
n2 π 2 2
8h(−1)k
n = 2k − 1, k = 1, 2, · · ·
= (2k − 1)2 π 2
0 n = 2k, k = 1, 2, · · ·
from which we have that all the bn are zero. Finally the solution u(x, t) is given by
+∞
X nπa nπ
u(x, t) = an cos t sin x .
n=1
l l
i.e.
+∞
8h(−1)k
X (2k − 1)πa (2k − 1)π
u(x, t) = cos t sin x .
(2k − 1)2 π 2 l l
k=1
2
Harmonic vibrations
If the initial conditions in (5.3.1) are changed to general ones, the solution of Tn (t) is still in the form of
(5.3.2). Then the string vibration can be expressed as
82
+∞
X nπa nπa nπ
u(x, t) = an cos t + bn sin t sin x .
n=1
l l l
Now let us consider a fixed point, x = x0 . The motion at this position is clearly given by
+∞
X nπa
u(x0 , t) = An cos t + θn ,
n=1
l
nπ
where An = αn sin x0 . That means the vibration at every particle of the string is a combination of
l
a series of harmonic vibrations with frequency:
nπa
ωn = n = 1, 2, 3, · · · ,
l
with a phase shift θn . The fundamental tone of the string vibration is determined by the lowest frequency
,e.g. ω1 = πa/l in Example 5.3.1. If the length of the string is cut to be one-half, e.g. the mid-point of
the string is fixed by the finger of a violin player, the frequency of fundamental tone is doubled.
• The heat equation, as we have noticed, cannot be solved backward in time; however, for the wave
equation the solution formula (5.3.3) makes as much sense for t > 0 as for t < 0. So the wave
equation can be solved both forward and backward.
• The heat equation enjoys the maximum principle, while the wave equation does not: consider the
initial-boundary value problem again, but with φ(x) = sin(πx/l); then the unique solution is u(x, t) =
cos(πat/l) sin(πx/l) which obviously changes sign, while u(x, 0) ≥ 0 on the interval [0, l].
• The solution of the heat equation with homogeneous Dirichlet boundary condition decays exponen-
tially as t → ∞, while the solution of the wave equation with the same boundary condition oscillates
with its total energy remaining constant.
To close this section, we supply an example which has Neumann boundary condition:
Example 5.3.2
utt − 4uxx = 0, x ∈ (0, 1), t ∈ R,
ux (0, t) = 0 = ux (1, t), t ∈ R,
u(x, 0) = 0, ut (x, 0) = x, x ∈ (0, 1).
Solution
This initial-boundary value problem models the situation that the two ends of string move freely in the
vertical direction and the string initially lies on the x−axis and is given an upward velocity x at point x.
Our intuition tells us that the string should “float”to infinity and at the meantime vibrate as t ↑ ∞. Let’s
compute and then see if this is the case from the formula for the solution.
83
Similarly as in Example 5.3.1, we start with the trial solution u(x, t) = X(x)T (t). Then again the PDE
leads to
X 00 (x) T 00 (t)
= = −λ ( constant ).
X(x) 4T (t)
The corresponding eigenvalue problem is
X 00 (x) + λX(x) = 0
(
X 0 (0) = 0, X 0 (1) = 0.
λ n = n2 π 2 n = 0, 1, 2, · · ·
Xn = cos nπx n = 0, 1, 2, · · · .
and
T0 (t) = a0 + b0 t.
Now we form
+∞
X
u(x, t) = Tn (t)Xn (x),
n=0
which satisfies the PDE and the boundary condition.
By the initial condition u(x, 0) = 0, we have
+∞
X
0= an Xn (x),
n=0
and so all an , including a0 , are zero. By the second initial condition ut (x, 0) = x, we have
+∞
X
x = b0 + bn 2nπ cos nπx.
n=1
Thus Z 1
1
b0 = xdx = ,
0 2
1
(−1)n − 1
Z
bn 2nπ = 2 x cos nπxdx = .
0 n2 π 2
Now
+∞
t X (−1)n − 1
u(x, t) = + cos 2nπt cos nπx.
2 n=1 2n3 π 3
Now we check to see if this match our physical intuition about the string floating to infinity and oscillating
as t ↑ ∞: observe that the absolute value of the general term in the series is dominated by 1/(n3 π 3 ), and
so the absolute value of the series is dominated by
+∞
X 1
n 3 π3
n=1
84
which, by the “p-test”, is finite. Thus u(x, t) grows in the order of t/2 (the first term in the formula for
u) as t ↑ ∞. On the other hand, the cosine functions in t−variable in the formula make u oscillate when
it ascends to infinity.
utt = a2 uxx
(
x ∈ (−∞, +∞), t ∈ (−∞, ∞)
(5.4.1)
u(x, 0) = φ(x), ut (x, 0) = ψ(x).
Such kind of problems, in which only initial conditions are imposed, are called Cauchy problems. In
Section 3.7, we use the fundamental solution to study the Cauchy problem for the heat equation. Now we
use the so-called d’Alembert formula to study this problem.
d’Alembert’s formula
We re-write the wave equation in the operator form
∂ 2 ∂ 2
( ) − (a ) u(x, t) = 0.
∂t ∂x
The operator reminds us the algebraic expression A2 − B 2 which can be factored as (A + B)(A − B).
Factoring the operator in the same fashion we have
∂ ∂ ∂ ∂
( + a )( − a )u(x, t) = 0.
∂t ∂x ∂t ∂x
We are tempted to introduce new independent variables η and ξ such that
∂ ∂ ∂ ∂ ∂ ∂
−a = , +a = .
∂t ∂x ∂η ∂t ∂x ∂ξ
This and the chain rule imply that
∂x ∂t ∂x ∂t
= −a, = 1, = a, = 1.
∂η ∂η ∂ξ ∂ξ
Then it makes sense to choose
x = −aη + aξ, t = η + ξ,
or equivalently
x − at x + at
η=− , ξ= .
2a 2a
Now in the new independent variables, the wave equation becomes
uηξ = 0.
The functions f (·) and g(·) are determined by the initial conditions. Using the initial conditions, we have
(
f (x) + g(x) = φ(x)
−af 0 (x) + ag 0 (x) = ψ(x)
i.e.
f (x) + g(x)
= φ(x)
1 x
Z
−f (x) + g(x) =
ψ(s)ds + C,
a 0
which is called d’Alembert’s formula. Since the derivation of d’Alembert’s formula is constructive, the
uniqueness of the solution of Cauchy problem (5.4.1) is also proved.
This formula does not apply to the initial-boundary value problem such as (5.3.1) when the interval
(x − at, x + at) is not included in the interval (0, l).
Travelling waves
We recall that the solution u(x, t) of the wave equation is decomposed into two parts
T ravelling-
-x
x0 x = x0 + at
Obviously the travelling speed is a. Thus f (x − at) is called a right-travellingpwave with speed a.
Similarly g(x + at) is called a left-travelling wave with speed a. Notice that a = T /ρ. Therefore the
wave propagation is faster, if the tension T is larger or the density ρ is smaller. One can get a better
understanding of the travelling waves from the following example.
l 3l l 2l
Sketch the solution of problem (5.4.1) at times: t = 0, , , , and .
2a 4a a a
(This is a “three-finger” pluck, with all three fingers removed from the string at once.)
Solution
u
6
h
@
t = 0 @
@ h/2
@
@ - x
−2l −l 0 l 2l
6
t = l/2a
H
HH h/2
HH
- x
−2l −l 0 l 2l
6
t = 3l/4a
HH h/2
H H
H
HH
H - x
−2l −l 0 l 2l
6
t = l/a
H
H H
HH h/2
HH HH
H - x
−2l −l 0 l 2l
6
t = 2l/a
H
H H H h/2
HH HHH -
H - x
−2l −l 0 l 2l
Figure 5.4.2 Separation, propagation and interaction of travelling waves in plucked string
At time t = 0, the original wave with amplitude h separates to a pair of waves which start to travel in the
two directions at speed a and at half the original amplitude. These two travelling waves interact within
the period t ∈ (0, l/a). They are completely separated after time t = l/a, since the base length of two
travelling waves is 2l and the relative travelling speed is 2a.
Reflection of waves
d’Alembert’s formula cannot be applied directly to the situation when there is a boundary point. Indeed,
when the infinitely long string is clamped at a point, and when an incident wave reaches this boundary
point, then the string exerts an, say, upward tensile force to the fixed boundary point, which in turn
exerts a downward force to the string with equal magnitude (Newton’s third law), resulting in a reflected
wave. This effect is not included in (5.4.4). However, as we will see in the following example, after some
preparation d’Alembert’s formula can still be used to solve the problem.
87
utt = a2 uxx
x ∈ (0, +∞), t ∈ (−∞, ∞)
u(x, 0) = φ(x), ut (x, 0) = ψ(x), x ∈ (0, +∞) (5.4.5)
t ∈ (−∞, ∞).
u(0, t) = 0,
We use the method of reflection, which was used in Section 3.7 to solve the heat conduction problem
on the half-line. Consider the standard Cauchy problem (5.4.1) with the following initial conditions:
utt = a2 uxx
x ∈ (−∞, +∞), t>0
φ(x) x>0 ψ(x) x>0 (5.4.6)
u(x, 0) = φ̃(x) = 0 x=0 , ut (x, 0) = ψ̃(x) = 0 x=0
−φ(−x) −ψ(−x)
x<0 x<0
where φ̃(x) and ψ̃(x) defined in (−∞, +∞) are the odd extension of functions φ(x) and ψ(x) which were
originally defined on the half-line (0, +∞). By Exercise 6, the solution of (5.4.6) is odd in x and hence it
is also the solution of (5.4.5). Then the solution of (5.4.5) is given by
1 x+at
Z
1
u(x, t) = [φ̃(x − at) + φ̃(x + at)] + ψ̃(s)ds.
2 2a x−at
We will compute the solution only for t > 0.
Case I x − at > 0, i.e. x > at
We have φ̃(x − at) = φ(x − at) and ψ̃(s) = ψ(s) for s ≥ x − at. Then the solution keeps its original form
in (5.4.4), i.e.
1 x+at
Z
1
u(x, t) = [φ(x − at) + φ(x + at)] + ψ(s)ds (x > at).
2 2a x−at
Substituting φ̃(ξ) = −φ(−ξ) and ψ̃(ξ) = −ψ(−ξ), where ξ = x − at < 0, into the above formula, we get
that
Z 0 Z x+at
1 1 1
u(x, t) = [−φ(−x + at) + φ(x + at)] + − ψ(−s)ds + ψ(s)ds.
2 2a x−at 2a 0
Notice that
Z 0 Z at−x
− ψ(−s)ds = − ψ(w)dw (set w = −s).
x−at 0
So
at−x x+at
−1
Z Z
1 1 1
u(x, t) = φ(−x + at) − ψ(s)ds + φ(x + at) + ψ(s)ds, (5.4.7)
2 2a 0 2 2a 0
that is,
Z at+x
1 1
u(x, t) = [−φ(at − x) + φ(at + x)] + ψ(s)ds (x < at). (5.4.8)
2 2a at−x
88
If both φ and ψ are positive (on the interval (0, ∞)), then the third and the fourth terms in (5.4.7) form
a positive, left-travelling wave, and the first and the second terms form a negative, right-travelling wave
which is the reflected wave. 2
Domain of dependence
Now let us investigate the solution of (5.4.1) at (x0 , t0 ), which is given by
1 x0 +at0
Z
1
u(x0 , t0 ) = [φ(x0 − at0 ) + φ(x0 + at0 )] + ψ(s)ds,
2 2a x0 −at0
It is clear that the solution at (x0 , t0 ) is determined by the initial information in interval [x0 −at0 , x0 +at0 ].
t
6
(x0 , t0 )
t0 b
• • • - x
x0 − at0 6
x0 x0 + at0 x∗
In daily life, the phenomenon of sound propagation is the simplest example. A person who is standing at
position x0 can only hear at time t0 > 0 the sound produced at the distance at0 at time t = 0, where a
is the speed of sound propagation (a = 344 M/Sec.); the solution value at (x0 , t0 ) has no relation to the
initial “information”at position x∗ (x∗ > x0 + at0 or x∗ < x0 − at0 ), since the initial information is too
far away to reach the position x0 .
We further investigate the line segment M1 M2 as illustrated below.
t
6
(x0 , t0 )
t0 b
M1 = (x1 , t∗ )
t∗ b b • M2 = (x2 , t∗ )
M1 M2 x∗
• • - x
x0 − at0 6
x0 x0 + at0
The solution at any point on M1 M2 , u(x, t∗ ) (x1 < x < x2 ) can be determined by the initial conditions
in (x0 − at0 , x0 + at0 ) and is well-defined. We denote
1 x0 +aτ0 ∗
Z
1
u(x0 , t0 ) = [φ∗ (x0 − aτ0 ) + φ∗ (x0 + aτ0 )] + ψ (s)ds,
2 2a x0 −aτ0
Since t∗ < t0 is arbitrary, the solution u(x0 , t0 ) can be determined by the information on any line segment
89
in the triangle. Meanwhile the information in the exterior of the triangle has no effect to the solution
value u(x0 , t0 ). Thus the triangle is called the domain of dependence of point (x0 , t0 ).
t
6
(x0 , t0 )
- x
x0 − at0 x0 + at0
Now let us consider the domain of dependence for the problem (5.4.5) which takes into account the
reflection of waves from boundary x = 0. For Case I, since x0 − at0 > 0, i.e. t0 < x0 /a, the influence
of boundary reflection has not yet reached x0 . Thus the domain of dependence at (x0 , t0 ) is the same as
illustrated in Figure 5.4.2. For Case II, x0 − at0 < 0. By (5.4.8) the domain of dependence of (x0 , t0 ) is
as illustrated below.
6
(x0 , t0 )
- x
x0 − at0 at0 − x0 x0 + at0
Figure 5.4.3 Domain of dependence at (x0 , t0 ) due to the reflection of waves at boundary x = 0
Domain of influence
Now let us consider the influence of the initial conditions at point x0 . Pick a point (x, t) in upper-half of
the x-t plane and draw its domain of dependence as illustrated below.
t
6
(x, t)
• - x
6 x0 6
x − at x + at
It is clear that the initial conditions at x0 will affect the solution at (x, t), if and only if (x0 , 0) is an interior
point of the bottom of the domain of dependence, i.e. x0 > x − at and x0 < x + at. In other words, the
initial conditions at x0 have influence on the domain
t
6
x + at = x0 x − at = x0
@
@
@
@
@
@• - x
x0
Figure 5.4.4a Domain of influence for initial conditions at x0
For the initial conditions in the interval (x1 , x2 ), the domain of influence is shown as below.
t
6
x + at = x1 x − at = x2
@
@
@
@
@
@• • - x
x1 x2
Figure 5.4.4b Domain of influence for initial conditions in interval (x1 , x2 )
Characteristic lines
We have seen that the two sets of parallel straight lines in x-t plane
t
x − at = constant
6
- x
and
t
6
x + at = constant
@ @ @ @ @ @ @ @ @ @
@ @ @ @ @ @ @ @ @ @
@ @ @ @ @ @ @ @ @ @
@ @ @ @ @ @ @ @ @ @
@ @ @ @ @ @ @ @ @ @
@ @ @ @ @ @ @ @ @ @ - x
91
are extremely important. The domain of dependence and domain of influence are bounded by these lines.
These lines are called characteristic lines of the wave equation.
Comparison of heat and wave equations, part 2. (This is a continuation of the discussion at the
end of Section 5.3.)
• As we have seen in Section 3.7, the heat equation predicts infinite speed of propagation of heat; on
the other hand, we have just found that the speed of propagation vibration waves modelled by the
wave equation is finite.
• Heat equations have a smoothing effect, but wave equations do not enhance the smoothness of initial
conditions (see Example 5.4.1).
Assignment 5
1. Derive the wave equation for a string that moves in a medium in which the resistance force between
the string and the medium is proportional to the velocity of the string.
2. Solve the following initial-boundary value problems for the wave equation
(i)
utt − 4uxx = 0, x ∈ (0, 1), t ∈ R,
u(0, t) = 0 = u(1, t), t ∈ R,
u(x, 0) = sin(πx), ut (x, 0) = sin(4πx), x ∈ (0, 1).
(ii)
utt − a2 uxx = 0, x ∈ (0, 1), t ∈ R,
ux (0, t) = 0 = ux (1, t), t ∈ R,
u(x, 0) = x, ut (x, 0) = 0, x ∈ (0, 1).
(iii)
utt − a2 uxx = 0, x ∈ (0, 1), t ∈ R,
u(0, t) = 0, ux (1, t) = 1, t ∈ R,
u(x, 0) = 0, ut (x, 0) = cos(πx), x ∈ (0, 1).
utt = uxx , x ∈ R, t ∈ R.
Is it possible that u(x, 1) is smoother than u(x, 0)? Is it possible to have a maximum principle for the
wave equation?
utt − a2 uxx = 0, x ∈ R, t ∈ R,
utt = uxx , x ∈ R, t ≥ 0.
Ft − Gx = 0.
Integrate this equation and apply Green’s Theorem to the trapezoid bounded by the x-axis, the charac-
teristic lines passing through (x0 , t0 ), and the horizontal line t = t.