Pde
Pde
Pde
Part I
Introduction
For the equation to be of second order, a, b, and c cannot all be zero. Define
its discriminant to be b2 – 4ac. The properties and behavior of its solution
are largely dependent of its type, as classified below.
α2 uxx = ut
Where the constant coefficient α2 is the thermo diffusivity of the bar, given
by α2 = k / ρs. (k = thermal conductivity, ρ = density, s = specific heat, of the
material of the bar.)
Further, let us assume that both ends of the bar are kept constantly at 0
degree temperature (abstractly, by connecting them both to a heat reservoir
Those two conditions are called the boundary conditions of this problem.
They literally specify the conditions present at the boundaries between the
bar and the outside. Think them as the “environmental factors” of the given
problem.
But before any of those boundary and initial conditions could be applied, we
will first need to process the given partial differential equation. What can
we do with it? There are other tools (by Laplace transforms, for example),
but the most accessible method to us is called the method of Separation of
Variables. The idea is to somehow de-couple the independent variables,
therefore rewrite the single partial differential equation into 2 ordinary
differential equations of one independent variable each (which we already
know how to solve). We will solve the 2 equations individually, and then
combine their results to find the general solution of the given partial
differential equation. For a reason that should become clear very shortly, the
method of Separation of Variables is sometimes called the method of
Eigenfunction Expansion.
u=XT uxx = X ″ T
ux = X ′ T utt = X T ″
ut = X T ′ uxt = utx = X ′ T ′
α2 X ″ T = X T ′.
2
Dividing both sides by α X T :
X ′′ T ′
=
X α2T
2
(α is a constant, so it could go to either side of the equation, but it is usually,
and more conveniently, moved to the t side.) The equation is now
“separated”, as all the x-terms are on the left and t-terms are on the right.
Note: The above step would not have been possible if either X = 0 or T = 0.
However, if either part is zero, then u = XT = 0, which will trivially satisfy
the given equation α2 uxx = ut . This constant zero solution is called the
trivial solution of the equation. We know this is going to be the case.
Therefore, we will assume from this point onward that X ≠ 0 and T ≠ 0. We
look for only the nonzero solutions.
X ′′ T ′
= =−λ .
X α2T
Why must the two sides be the same constant? Well, think what would
happen if one of the sides isn’t a constant. The equation would not be true
for all x and t if that were the case, because then one side/variable could be
held at a fixed value while the other side/variable changes.
Next, equate first the x-term and then the t-term with −λ. We have
X ′′
=−λ → X ″ = −λX → X ″ + λX = 0,
X
and,
T′
=−λ → T ′ = −α2 λ T → T ′ + α2 λ T = 0.
α2T
Consequently, the single partial differential equation has now been separated
into a simultaneous system of 2 ordinary differential equations. They are a
second order homogeneous linear equation in terms of x, and a first order
linear equation (it is also a separable equation) in terms of t. Both of them
can be solved easily using what we have already learned in this class.
T ′ + α2 λ T = 0 .
The general solution (that satisfies the boundary conditions) shall be solved
from this system of simultaneous differential equations. Then the initial
condition u(x, 0) = f (x) could be applied to find the particular solution.
t 3 X ″ T + x3 X T ″ = 0 ,
t3 X ″ T = − x3 X T ″.
t 3 T − x3 X
=
T ′′ X ′′ .
Now insert a constant of separation:
t 3 T − x3 X
= = −λ .
T ′
′ X ′′
t3 T = −λ T ″ → λ T ″ + t3 T = 0,
− x3 X = −λ X ″ → λ X ″ − x3 X = 0.
X ′ T + 2 X ′ T ′ − 10 X T ″ = 0,
X ′ T + 2 X ′ T ′ = 10 X T ″.
T + 2 T ′ = −λ T ″ → λ T ″ + 2 T ′ + T = 0,
10 X = −λ X ′ → λ X ′ + 10 X = 0.
What we have done thus far is to separate the heat conduction equation, with
2 independent variables, into 2 equations of one variable each. Meanwhile
we have also rewritten the boundary conditions, so that they now associate
with the spatial variable x only.
Hence, the next goal is to find the eigenvalues λ such that the boundary
value problem (1)
will have a nonzero solution satisfying both boundary conditions. Since the
form of the general solution of the second order linear equation is dependent
on the type of roots that its characteristic equation has. In this example, the
characteristic equation is r 2 + λ = 0. The type of roots it has is dependent on
its discriminant, which is simply −4λ. We will attempt to find λ by
separately considering the 3 possible types of the solution arise from the
different roots of the characteristic equation.
X(0) = 0 = C1 + C2 → C2 = − C1
X(x) = C1 + C2 x.
X(0) = 0 = C1 → C1 = 0
X(L) = 0 = C1 + C2 L = C2 L → C2 = 0 (because L > 0)
The (positive) values of λ for which the equation will have a solution
satisfying the specified boundary conditions, i.e. the eigenvalues of this BVP,
are
2 n2 π 2
λ =σ = 2 , n = 1, 2, 3, …
L
What are the eigenfunctions, then? Retracing our work above, we see that
they occur only when λ are positive (therefore, the general solution in the
form of X(x) = C1 cos(σx) + C2 sin(σx)), that C1 must be zero, and C2 could
2 2 2
be any nonzero constant. Lastly, λ = n π /L , where n = 1, 2, 3, … are all
the positive integers.
nπ x
X n = sin
L , n = 1, 2, 3, …
Case 1: If λ < 0:
X(0) = 0 = C1 + C2 → C2 = − C1
σL −σL σL −σL
X ′(L) = 0 = C1 σ e − C2 σ e = C1 σ ( e +e )
Case 2: If λ = 0:
X(0) = 0 = C1 → C1 = 0
X ′(L) = 0 = C2 → C2 = 0
2(2n − 1) 2 π 2
λ =σ = , n = 1, 2, 3, …
4 L2
(2n − 1) π x
X n = sin , n = 1, 2, 3, …
2L
T ′ + α2 λ T = 0. (2)
−α 2 n 2π 2 t / L2
Tn (t ) = Cn e , n = 1, 2, 3, …
We can now assemble the result from the two ordinary differential equations
to find the solutions of the partial differential equation. Recall that the
assumption in the separation of variables method is that the PDE has
solutions in the form u(x, t) = X(x)T(t). Therefore, we see that the solutions
of the one-dimensional heat conduction equation, with the boundary
conditions u(0, t) = 0 and u(L, t) = 0, are in the form
−α 2 n 2π 2 t / L2 nπ x
u n ( x, t ) = X n ( x) Tn (t ) = Cn e sin
L ,
n = 1, 2, 3, …
∞
−α 2 n 2π 2 t / L2 nπ x
u ( x, t ) = ∑ Cn e sin
n =1 L .
Once the general solution has been found, we can now apply the initial
condition in order to find the particular solution. Set t = 0 in the general
solution above, and equate it with the initial condition u(x, 0) = f (x):
∞
nπ x
u ( x,0) = ∑ Cn sin = f ( x) .
n =1 L
Take a few seconds to grasp what this equation says, and we quickly realize
that we are not out of the woods yet. The equation above specifies that the
(arbitrary) initial condition must be equal to an infinite series of sine terms,
and there are infinitely many coefficients cn that need to be solved. Can this
equation even be solved, in general? The answer is a reassuring “yes”. But
to get there we still need to know several things. Namely, what kind of
functions could be expressed as a series of sines (and, more generally, sines
and/or cosines)? Even if such functions exist, how does the arbitrary initial
condition f (x) fit in? Lastly, how could we possibly solve for the infinitely
many coefficients with just this one equation given?
4. Multiply the results from steps (2) and (3), and sum up all the
products to find the general solution.
7. X ″ + λX = 0, X(0) = 0 , X ′(2π) = 0.
8. X ″ + λX = 0, X ′(0) = 0 , X(2π) = 0.
9. X ″ + λX = 0, X ′(0) = 0 , X ′(2π) = 0.
12. (a) Show that any positive eigenvalue of the boundary value problem
X ″ + λX = 0, X(0) + X ′(0) = 0, X(L) = 0,
2
must be in the form λ = σ , where σ satisfies the equation σ = tan(σL). (b) Is
0 an eigenvalue of this problem?
13. Show that 0 is not an eigenvalue, and that any positive eigenvalue of the
boundary value problem
X ″ + λX = 0, X(0) − X ′(0) = 0, X(L) + 2X ′(L) = 0,
2 2σ 2 − 1
must be in the form λ = σ , where σ satisfies the equation cot(σL ) = .
3σ
Part II
Fourier Series
a0 ∞ nπ x nπ x
f ( x ) = + ∑ a n cos + bn sin
2 n =1 L L
mπ x
L
1
am = ∫ f ( x) cos dx , m = 0, 1, 2, 3, …
L −L L
nπ x
L
1
bn = ∫ f ( x) sin dx , n = 1, 2, 3, …
L −L L
The coefficients a’s are called the Fourier cosine coefficients (including a0,
the constant term, which is in reality the 0-th cosine term), and b’s are called
the Fourier sine coefficients.
Note 3: Even though that the “=” sign is usually used to equate a periodic
function and its Fourier series, we need to be a little careful. The function f
and its Fourier series “representation” are only equal to each other if, and
whenever, f is continuous. Hence, if f is continuous for −∞ < x < ∞, then f is
exactly equal to its Fourier series; but if f is piecewise continuous, then it
disagrees with its Fourier series at every discontinuity. (See the Fourier
Convergence Theorem below for what happens to the Fourier series at a
discontinuity of f .)
Note 7: The constant term in the Fourier series, which has expression
L L
a0 1 1 1
= ⋅ ∫ f ( x) cos(0) dx =
2 L −∫L
f ( x) dx ,
2 2 L −L
is just the average or mean value of f (x) on the interval [−L, L]. Since f is
periodic, this average value is the same for every period of f. Therefore, the
constant term in a Fourier series represents the average value of the function
f over its entire domain.
2 2
mπ x 1 x2
L
1 1 1
a0 = ∫ f ( x) cos dx = ∫−2 x dx = 2 2 = ( 2 − 2) = 0
L −L L 2 −2
2
2
nπ x nπx
L
1 1
an = ∫ f ( x) cos dx = ∫ x cos dx
L −L L 2 −2
2
1 2 x nπx
2 2
nπx 2
=
2 nπ
sin −
2 − 2 nπ ∫−2sin 2 dx
1 2x
2
nπx 4 nπx
= sin + 2 2 cos
2 nπ 2 nπ 2 −2
1 4 4
= 0 + 2 2 cos( nπ ) − 0 + 2 2 cos( − nπ ) = 0
2 nπ nπ
Hence, there is no nonzero cosine coefficient for this function. That is,
its Fourier series contains no cosine terms at all. (We shall see the
significance of this fact a little later.)
2
nπ x nπ x
L
1 1
bn = ∫ f ( x) sin dx = ∫ x sin dx
L −L L 2 −2
2
1 − 2 x nπ x
2 2
nπ x −2
2 −2 nπ −∫2
= cos − cos dx
2 nπ 2
2
1 − 2x nπ x 4 nπ x
= cos + 2 2 sin
2 nπ 2 nπ 2 −2
1 − 4 4
= cos( nπ ) − 0 − cos( − nπ ) − 0
2 nπ nπ
−2
= (cos(nπ ) + cos(nπ ) ) = − 4 cos( nπ )
nπ nπ
4
nπ , n = odd (−1) n +1 4
= =
−4
, n = even nπ .
nπ
4 ∞
(−1) n + 1 nπ x
Therefore, f ( x ) = ∑
π n =1 n
sin
2
.
4 ∞
(−1) n + 1 nπ x
f ( x) = ∑ sin .
π n =1 n 2
4 4
1 1 x2 1
a0 = ∫ x dx = = (8 − 0) = 4
20 2 2 0
2
For n = 1, 2, 3, … :
4
1 nπx
an = ∫ x cos dx
20 2
1 2x
4
nπx 4 nπx
= sin + 2 2 cos
2 nπ 2 nπ 2 0
1 4 4
= 0 + 2 2 cos( 2nπ ) − 0 + 2 2 cos(0) = 0
2 nπ nπ
4
1 nπx
bn = ∫ x sin dx
20 2
1 − 2x
4
nπx 4 nπx
= cos + 2 2 sin
2 nπ 2 nπ 2 0
1 − 8 −4
= cos( 2nπ ) − 0 − (0 − 0 ) =
2 nπ nπ
Consequently,
a0 ∞ nπ x nπ x −4 ∞ 1 nπx
f ( x) = + ∑ an cos + bn sin = 2+ ∑ sin
2 n =1 L L π n =1 n 2 .
8 ∞
1 (2n − 1)π x
Answer : f ( x) = 1 −
π 2 ∑ (2n − 1)
n =1
2
cos
2
∞
8 1
Answer : f ( x) =
π
∑ (2n − 1) sin((2n − 1)π x)
n =1
Example: The Fourier series (period 2π) representing f (x) = 6 cos(x) sin(x) is
not exactly itself as given, since the product cos(x) sin(x) is not a term in a
Fourier series representation. However, we can use the double-angle
formula of sine to obtain the result: 6 cos(x) sin(x) = 3 sin(2x).
Consequently, the Fourier series is f (x) = 3 sin(2x).
A consequence of this theorem is that the Fourier series of f will “fill in” any
removable discontinuity the original function might have. A Fourier series
will not have any removable-type discontinuity.
4 ∞
(−1) n + 1 nπ x
f ( x) = ∑ sin .
π n =1 n 2
The following figures are the graphs of various finite n-th partial sums of the
series above.
n=3
n = 10
n = 30
n = 50
This behavior is known as the Gibbs Phenomenon. It further states that the
partial sums of a Fourier series will overshoot a jump discontinuity by an
amount approximately equal to 9% of the jump. That is, near each jump
discontinuity, the overshoot amounts to about
for large n. Further, this overshoot does not go away for any finitely large n.
We have seen a few graphs of its partial sums. But what will the graph of
the actual Fourier series look like?
Examples: cos(x), sec(x), any constant function, x2, x4, x6, … , x −2, x −4, …
Most functions, however, are neither even nor odd. There is one function
that is both even and odd. (What is it?)
The table below summaries the result of performing the common arithmetic
operations on a pair of even and/or odd functions:
The result above can be extended to arbitrarily many terms. For example, a
sum of three or more even functions will again be even. (Care needs to be
taken in the cases where 3 or more odd functions forming a product/quotient.
For example, a product of 3 odd functions will be odd, but a product of 4
odd functions is even.)
L L
∫
−L
f ( x ) dx = 2 ∫ f ( x) dx .
0
∫
−L
f ( x ) dx = 0 .
Suppose f is an even periodic function of period 2L, then its Fourier series
contains only cosine (include, possibly, the constant term) terms. It will not
have any sine term. That is, its Fourier series is of the form
a0 ∞ nπ x
f ( x) = + ∑ an cos
2 n =1 L .
Conversely, any periodic function whose Fourier series has the form of a
cosine series as shown must be an even periodic function. Computationally,
this means that the Fourier coefficients of an even periodic function are
given by
mπ x mπ x
L L
1 2
am = ∫ f ( x) cos dx = ∫ f ( x) cos dx ,
L −L L L0 L
m = 0, 1, 2, 3, …
bn = 0, n = 1, 2, 3, …
Notice that the integrand in the definite integral used to find the cosine
coefficients a’s is an even function (it is a product of two even functions, f (x)
and cos x). Therefore, we can use the symmetric property of even functions
to simplify the integral.
If f is an odd periodic function of period 2L, then its Fourier series contains
only sine terms. It will not have any cosine term. That is, its Fourier series
is of the form
∞
nπ x
f ( x) = ∑ bn sin
n =1 L
Conversely, any periodic function whose Fourier series has the form of a
sine series as shown must be an odd periodic function. Therefore, the
Fourier coefficients of an odd periodic function are given by
am = 0, m = 0, 1, 2, 3, …
nπ x
L
2
bn =
L ∫0
f ( x) sin
L
dx , n = 1, 2, 3, …
Example: We have calculated earlier that the function f (x) = x, −2 < x < 2,
f (x + 4) = f (x), has as its Fourier series consists of purely sine terms:
4 ∞ (−1) n + 1 nπ x
f ( x) = ∑ sin .
π n =1 n 2
We now see that this sine series signifies that the function is odd periodic.
It is perhaps not very obvious, but the integrand in the integral for the
Fourier sine coefficients is another even function. It is a product of two odd
functions, f (x) and sin x, which makes it even. Therefore, we can again take
advantage of the symmetric property of even functions to simplify the
integral.
Here is an outline of how this can be done. Start with a function that is
defined only on an interval of finite length, from 0 to L. First expand the
function to be defined on the interval from −L to L such that the function is
an even or an odd function as required. Then define the function to be
periodic with a period of T = 2L by requiring F(x + 2L) = F(x). This process
is actually much easier than it sounds. Mathematically, the process can be
achieved rather simply, as described below.
Given f (x) defined on [0, L]. Its even extension of period 2L is:
f ( x), 0 ≤ x ≤ L
F ( x) = , F(x + 2L) = F(x).
f (− x), − L < x < 0
a0 ∞ nπ x
Where F ( x) = + ∑ an cos , such that
2 n =1 L
mπ x
L
2
am =
L ∫
0
f ( x) cos
L
dx , m = 0, 1, 2, 3, …
bn = 0, n = 1, 2, 3, …
Given f (x) defined on (0, L). Its odd extension of period 2L is:
∞
nπ x
Where F ( x ) = ∑ bn sin , such that
n =1 L
am = 0, m = 0, 1, 2, 3, …
nπ x
L
2
bn =
L ∫
0
f ( x) sin
L
dx , n = 1, 2, 3, …
Example: Let f (x) = x, 0 ≤ x < 2. Find its cosine and sine series
extensions of period 4.
8 ∞
1 (2n − 1)π x
Answers: Cosine series: f ( x) = 1 − ∑
π 2 n =1 (2n − 1) 2
cos
2
4 (−1) n + 1
∞
nπ x
Sine series: f ( x) = ∑ sin
π n =1 n 2
We now know that the above equation says that the initial condition needs to
be an odd periodic function of period 2L. Since the initial condition could
be an arbitrary function, it usually means that we would need to “force the
issue” and expand it into an odd periodic function of period 2L. That is
∞
nπ x
f ( x ) = ∑ bn sin
n =1 L .
nπ x
L
2
Cn = bn = ∫ f ( x) sin dx .
L0 L
2
Since the standard form of the heat conduction equation is α uxx = ut,
2
we see that α = 8; and we also note that L = 5. Therefore, the general
solution is
∞
nπ x
u ( x, t ) = ∑ Cn e −α 2 n 2π 2 t / L2
sin
n =1 L
∞
nπ x
= ∑ Cn e −8 n 2π 2 t / 25
sin
n =1 5
C5 = b5 = 2,
C10 = b10 = −4,
C25 = b25 = 1,
Cn = bn = 0, for all other n, n ≠ 5, 10, or 25.
Hence,
2
)π 2 t / 25 2
)π 2 t / 25
u ( x, t ) = 2e−8(5 sin(πx) − 4e−8(10 sin(2πx)
2
)π 2 t / 25
+ e−8( 25 sin(5πx)
5
nπ x nπ x
L
2 2
bn =
L ∫0
f ( x ) sin
L
dx = ∫ x sin
50 5
dx
2 − 5 x nπ x
5 5
nπ x −5
5 0 nπ ∫0
= cos − cos dx
5 nπ 5
5
2 − 5x nπ x 25 nπ x
= cos + 2 2 sin
5 nπ 5 nπ 5 0
2 − 25
= cos( nπ ) − 0 − (0 − 0 )
5 nπ
− 10
= cos( nπ )
nπ
10
nπ , n = odd (−1) n+110
= =
− 10 nπ
, n = even
nπ
10 (−1) n +1
∞
nπ x
f ( x) = ∑ sin .
π n =1 n 5
10 ∞
(−1)n +1 −8n 2π 2t / 25 nπ x
u( x, t ) =
π
∑n =1 n
e sin
5 .
9 – 12 Expand each function into its cosine series and sine series
representations of the indicated period. Determine the values to which each
series converge to at x = 0, x =2, and x = −2.
9. f (x) = 3 − x, T = 6. 10. f (x) = e, T = 2π.
x, 0 ≤ x < 2
11. f (x) = sin x, T = 2π. 12. f ( x) = , T = 6.
2, 2 ≤ x < 3
14. Solve the heat conduction problem of the given initial conditions.
1. f (x) = 5, f (0) = 5.
2. f (x) = 1 + sin(2x), f (0) = 1.
∞
6 1
3. f ( x) = 3 + ∑ sin( nπx) , f (0) = 3.
π n =1 n
π2 1 ∞
π π2
4. f ( x) = + ∑ 2 cos(2nx ) − sin( 2nx ) , f ( 0) = .
3 n =1 n n 2
π2 (−1) n ∞
5. f ( x) = + 4 ∑ 2 cos( nx ) , f (0) = 0.
3 n =1 n
4 ∞
1 nπ nπx 1 nπ nπx
6. f ( x) = 1 +
π
∑ n sin cos + 1 − cos
2 2 n 2
sin
2
,
n =1
f (0) = 2.
2 4 ∞ 1
7. f ( x) = − ∑ cos( 2nx ) , f (0) = 0.
π π n =1 (2n − 1)(2n + 1)
1 1 ∞
8. f ( x) = + ∑ cos( n( x − c )) , f (0) = 0.
2π π n =1
3 ∞ 12 nπx
9. Cosine series: f ( x ) = + ∑ 2 2
cos ,
2 n =1 ( 2n − 1) π 3
∞
9 nπx
Sine series: f ( x ) = ∑ sin ;
n = 1 nπ 3
The cosine series converges to 3, 1, and 1; the sine series converges to 0, 1,
and −1, respectively, at x = 0, x =2, and x = −2.
10. Cosine series: f (x) = e,
∞
4e
Sine series: f ( x ) = ∑ sin( nx) ;
n =1 ( 2 n − 1)π
The cosine series converges to e at all 3 points. The sine series converges
to 0, e, and − e, respectively, at x = 0, x =2, and x = −2.
2 4 ∞ 1
11. Cosine series: f ( x ) = − ∑ cos(2nx ) ,
π π n =1 ( 2n − 1)(2n + 1)
Sine series: f (x) = sin x;
Both series converge to 0 at x = 0 and to sin(2) at x = 2. At x = −2, the
cosine series converges to sin(2), the sine series to − sin(2).
Part III
Keep in mind that, throughout this section, we will be solving the same
partial differential equation, the homogeneous one-dimensional heat
conduction equation:
α2 uxx = ut
where u(x, t) is the temperature distribution function of a thin bar, which has
2
length L, and the positive constant α is the thermo diffusivity constant of
the bar. The equation will now be paired up with new sets of boundary
conditions.
Now let us consider the situation where, instead of them being kept at
constant 0 degree temperature, the two ends of the bar are also sealed with
perfect insulation so that no heat could escape to the outside environment
(recall that the side of the bar is always perfectly insulated in the one-
dimensional assumption), or vice versa. The new boundary conditions are
ux(0, t) = 0 and ux(L, t) = 0*, reflecting the fact that there will be no heat
transferring, spatially, across the points x = 0 and x = L. (Hence, this is a
Neumann type problem.) The heat conduction problem becomes the initial-
boundary value problem below.
The first step is the separation of variables. The equation is the same as
before. Therefore, it will separate into the exact same two ordinary
differential equations as in the first heat conduction problem seen earlier.
The new boundary conditions separate into
As before, we cannot choose T(t) = 0. Else we could only get the trivial
solution u(x,t) = 0, rather than the general solution. Hence, the new
boundary conditions should be X ′(0) = 0 and X ′(L) = 0.
*
The conditions say that the instantaneous rate of change with respect to x, the spatial variable (i.e., the rate
of point-to-point heat transfer), is zero at each end. They do not suggest that the temperature is constant
(that is, there is no change in temperature through time, which would require ut = 0) at each end.
T ′ + α2 λ T = 0 .
n2 π 2
λ= 2 , n = 1, 2, 3, …
L
The third step is to substitute the positive eigenvalues found above into the
equation of t and solve:
n 2π 2
T′+α 2
T =0.
L2
As a result, the solutions of the second equation are just the ones we have
gotten the last time
Tn (t ) = Cn e −α n π 2 t / L2
2 2
, n = 1, 2, 3, …
u 0( x , t ) = C 0,
−α 2 n 2π 2 t / L2 nπ x
u n ( x, t ) = X n (t ) Tn (t ) = Cn e cos
L ,
n = 1, 2, 3, …
The general solution is their linear combination. Hence, for a bar with both
ends insulated, the heat conduction problem has general solution:
∞
nπ x
u ( x, t ) = C 0 + ∑ C n e −α 2 n 2π 2 t / L2
cos
n =1 L .
∞
nπ x
u ( x,0) = C0 + ∑ Cn cos = f ( x) .
n =1 L
We see that the requirement is that the initial temperature distribution f (x)
must be a Fourier cosine series. That is, it needs to be an even periodic
function of period 2L. If f (x) is not already an even periodic function, then
we will need to expand it into one and use the resulting even periodic
extension of f (x) in its place in the above equation. Once this is done, the
coefficients C’s in the particular solution are just the corresponding Fourier
cosine coefficients of the initial condition f (x). (Except for the constant term,
where the relation C0 = a0 / 2 holds, instead.)
nπ x
L
2
C n = an =
L ∫
0
f ( x ) cos
L
dx , n = 1, 2, 3, …
C0 = a 0 / 2
2
First note that α = 3 and L = 8, and the fact that the boundary
conditions indicating this is a bar with both ends perfectly insulated.
Substitute them into the formula we have just derived to obtain the
general solution for this problem:
∞
nπ x
u ( x, t ) = C0 + ∑ Cn e −3n 2π 2 t / 64
cos
n =1 8 .
Check the initial condition f (x), and we see that it is already in the
require form of a Fourier cosine series of period 16. Therefore, there
is no need to find its even periodic extension. Instead, we just need to
extract the correct Fourier cosine coefficients from f (x):
C0 = a0 / 2 = 9,
C2 = a2 = −3,
C16 = a16 = −6,
Cn = an = 0, for all other n, n ≠ 0, 2, or 16.
2
)π 2 t / 64 πx 2
)π 2 t / 64
u ( x, t ) = 9 − 3e−3( 2 cos( ) − 6e−3(16 cos(2π x)
4
u(x, 0) = f (x).
α2 vxx = 0.
Divide both sides by α2 and integrate twice with respect to x, we find that
v(x) must be in the form of a degree 1 polynomial:
v(x) = Ax + B.
Then, rewrite the boundary conditions in terms of v: u(0, t) = v(0) = T1, and
u(L, t) = v(L) = T2. Apply those 2 conditions to find that:
v(0) = T1 = A(0) + B = B → B = T1
Therefore,
T2 − T1
v( x ) = x + T1 .
L
That is, just rename the function u as v, ignore the time variable t, and put
whatever x-coordinate specified directly into v(x).
Once the steady-state solution has been found, we can set it aside for the
time being and proceed to find the transient part of solution, w(x, t). First we
will need to rewrite the given initial-boundary value problem slightly. Keep
in mind that the initial and boundary conditions as originally given were
meant for the temperature distribution function u(x, t) = v(x) + w(x, t). Since
we have already found v(x), we shall now subtract out the contribution of v(x)
from the initial and boundary values. The results will be the conditions that
the transient solution w(x, t) alone must satisfy.
Note: Recall that u(0, t) = v(0) = T1, and u(L, t) = v(L) = T2.
∞
nπ x
w( x, t ) = ∑ Cn e−α n π 2 t / L2
2 2
sin
n =1 L .
Where the coefficients Cn are equal to the corresponding Fourier sine
coefficients bn of the (newly rewritten) initial condition w(x, 0) = f (x) − v(x).
(Or those of w(x, 0)’s odd periodic extension, of period 2L, if it is not already
an odd periodic function of the correct period.) Explicitly, they are given by
nπ x
L
2
Cn = bn =
L ∫ ( f ( x) − v( x) )sin
0
L
dx , n = 1, 2, 3, …
u ( x, t ) = v( x) + w( x, t )
T2 − T1 ∞ nπ x
x + T1 + ∑ Cn e−α n π t / L sin
2 2 2 2
= .
L n =1 L
2
First we note that α = 8 and L = 5. Since T1 = 10 and T2 = 90, the
steady-state solution is v(x) = (90 − 10)x / 5 + 10 = 16x + 10. We then
subtract v(0) = 10 from u(0, t), v(5) = 90 from u(5, t), and v(x) = 16x +
10 from u(x, 0) to obtain a new set of initial-boundary values that the
transient solution w(x,t) alone must satisfy:
C5 = b5 = 2,
C10 = b10 = −4,
C30 = b30 = 1,
Cn = bn = 0, for all other n, n ≠ 5, 10, or 30.
2
)π 2 t / 25
u ( x, t ) = v( x) + w( x, t ) = 16 x + 10 + 2e−8(5 sin(π x)
−8(10 2 )π 2 t / 25 −8 ( 30 2 )π 2 t / 25
− 4e sin(2π x) + e sin(6π x)
lim u ( x, t ) = v( x) .
t →∞
This relation is true only for the solution of heat conduction equation
(modeling diffusion-like processes that are thermodynamically irreversible).
Physically speaking, v(x) describes the eventual state of maximum entropy
as dictated by the second law of Thermodynamics.
Caution: The above relation is not true, in general, for solutions of the wave
equation in the next section. This difference is due to the fact that the wave
equation models wave-like motions which are thermodynamically reversible
processes.
v(0) = 50 = A(0) + B = B → B = 50
v′(6) = 0 = A → A=0
Hence,
v′(10) = 25 = A → A = 25
Hence,
v(0) = 35 = A(0) + B = B → B = 35
0 = 7A + 35
A = −5
4. Multiply the results from steps (2) and (3), and sum up all the
products to find the general solution respect to the given
homogeneous boundary conditions. Add to it the steady-state solution
(from step 0, if applicable) to find the overall general solution.
Here is a list of heat conduction problems and their solutions. All solutions
obey the homogeneous one-dimensional heat conduction equation
α2 uxx = ut .
They only differ in boundary conditions (which are given below for each
problem). The initial condition is always arbitrary, u(x, 0) = f (x).
∞
nπ x
u ( x, t ) = ∑ Cn e −α 2 n 2π 2 t / L2
sin
n =1 L .
Expand f (x) to be a Fourier sine series, then Cn = bn.
Steady-state solution is v(x) = 0.
∞
nπ x
u ( x, t ) = C 0 + ∑ C n e −α 2 n 2π 2 t / L2
cos
n =1 L .
Expand f (x) to be a Fourier cosine series, then C0 = a0 / 2, and Cn = an,
n = 1, 2, 3, …
Steady-state solution is v(x) = C0.
3. Bar with T1 degrees at the left end, and T2 degrees at the right end
(boundary conditions: u(0, t) = T1, u(L, t) = T2)
T2 − T1 ∞ nπ x
u ( x, t ) = x + T1 + ∑ Cn e −α 2 n 2π 2 t / L2
sin
L n =1 L .
Expand (f (x) − v(x)) to be a Fourier sine series, then Cn = bn.
T2 − T1
Steady-state solution is v( x) = x + T1 .
L
13. Consider the heat conduction equation below, subject to each of the 4
sets of boundary conditions.
Given the common initial condition u(x, 0) = 300, determine the temperature
at the midpoint of the bar (at x = 5) after a very long time has elapsed.
Which set of boundary conditions will give the highest temperature at that
point?
(d) lim u ( x, t ) = 0 .
t →∞
−8π t
sin(2πx) − 2e −25π
2 2
10. (a) u ( x, t ) = −5 x + 40 + 5e
t/2
sin(5πx / 2) .
11. (a) and (b) lim u (4, t ) = 20 .
t →∞
∞
nπ x
12. (a) u ( x, t ) = 32 + ∑ Cn e
−9 n π 2 2
t / 16
sin
n =1 4
nπ x
4
1
(b) C n = ∫ ( f ( x ) − 32 ) sin dx , n = 1, 2, 3, …
2 0
4
(c) lim u ( x, t ) = 32
t →∞
Part IV
a2 uxx = utt
The two boundary conditions reflect that the two ends of the string are
clamped in fixed positions. Therefore, they are held motionless at all time.
The equation comes with 2 initial conditions, due to the fact that it contains
the second partial derivative of time, utt. The two initial conditions are the
initial (vertical) displacement u(x, 0), and the initial (vertical) velocity
ut(x, 0)*, both are arbitrary functions of x alone. (Note that the string is
merely the medium for the wave, it does not itself move horizontally, it only
vibrates, vertically, in place. The resulting undulation, or the wave-like
“shape” of the string, is what moves horizontally.)
*
Velocity = rate of change of displacement with respect to time. The other first partial derivative ux
represents the slope of the string at a point x and time t.
We first let u(x, t) = X(x)T(t) and separate the wave equation into two
ordinary differential equations. Substituting uxx = X ″ T and utt = X T ″ into
the wave equation, it becomes
a2 X ″ T = X T ″.
X ′′ T ′′
=
X a2 T
X ′′ T ′′
= =−λ
X a2 T
X ′′
=−λ → X ″ = −λX → X ″ + λX = 0,
X
T ′′
=−λ → T ″ = −a2 λ T → T ″ + a2 λ T = 0.
a2 T
The boundary conditions also separate:
T ″ + a2 λ T = 0.
X ″ + λX = 0, X(0) = 0, X(L) = 0.
We have already solved this eigenvalue problem, recall. The solutions are
n2 π 2
Eigenvalues: λ= 2 , n = 1, 2, 3, …
L
nπ x
Eigenfunctions: X n = sin , n = 1, 2, 3, …
L
Next, substitute the eigenvalues found above into the second equation to find
T(t). After putting eigenvalues λ into it, the equation of T becomes
n 2π 2
T ′′ + a
2
T =0.
L2
anπ
r=± i.
L
Thus, the solutions are simple harmonic:
anπ t anπ t
Tn (t ) = An cos + Bn sin
L L , n = 1, 2, 3, …
Multiplying each pair of Xn and Tn together and sum them up, we find the
general solution of the one-dimensional wave equation, with both ends fixed,
to be
There are two sets of (infinitely many) arbitrary coefficients. We can solve
for them using the two initial conditions.
Set t = 0 and apply the first initial condition, the initial (vertical)
displacement of the string u(x, 0) = f (x), we have
∞
nπ x
u ( x,0) = ∑ ( An cos(0) + Bn sin(0) ) sin
n =1 L
∞
nπ x
= ∑ An sin = f ( x)
n =1 L
Therefore, we see that the initial displacement f (x) needs to be a Fourier sine
series. Since f (x) can be an arbitrary function, this usually means that we
need to expand it into its odd periodic extension (of period 2L). The
coefficients An are then found by the relation An = bn, where bn are the
corresponding Fourier sine coefficients of f (x). That is
nπ x
L
2
An = bn = ∫ f ( x ) sin dx .
L0 L
Notice that the entire sequence of the coefficients An are determined exactly
by the initial displacement. They are completely independent of the other
sequence of coefficients Bn, which are determined solely by the second
initial condition, the initial (vertical) velocity of the string. To find Bn, we
differentiate u(x, t) with respect to t and apply the initial velocity,
ut(x, 0) = g(x).
∞
anπ nπ x
ut ( x,0) = ∑ Bn sin = g ( x) .
n =1 L L
We see that g(x) needs also be a Fourier sine series. Expand it into its odd
periodic extension (period 2L), if necessary. Once g(x) is written into a sine
series, the previous equation becomes
∞
anπ nπ x ∞
nπ x
ut ( x,0) = ∑ Bn sin = g ( x) = ∑ bn sin
n =1 L L n =1 L
Therefore,
nπ x
L
L 2
Bn =
anπ
bn =
anπ ∫ g ( x) sin
0
L
dx .
nπ x
L
2
An = ∫ f ( x ) sin dx , n = 1, 2, 3, …
L0 L
Therefore,
∞
anπ t nπ x
u ( x, t ) = ∑ An cos sin
n =1 L L .
Use the product formula sin(A) cos(B) = [sin(A − B) + sin(A + B)] / 2, the
solution above can be rewritten as
1 ∞ nπ ( x − at ) nπ ( x + at )
u ( x, t ) = ∑ An sin + sin
2 n =1 L L
In which F(x) is the odd periodic extension (period 2L) of the initial
displacement f (x).
†
Jean le Rond d”Alembert (1717 – 1783) was a French mathematician and physicist. He is perhaps best
known to calculus students as the inventor of the Ratio Test for convergence.
Special case II: Zero initial displacement, nonzero initial velocity: f (x) = 0,
g(x) ≠ 0.
nπ x
L
2
Bn =
anπ ∫
0
g ( x) sin
L
dx , n = 1, 2, 3, …
Therefore,
∞
anπ t nπ x
u ( x, t ) = ∑ Bn sin sin
n =1 L L .
2
First note that a = 9 (so a = 3), and L = 5.
∞
3nπ t 3nπ t nπ x
u ( x, t ) = ∑ An cos + Bn sin sin
n =1 5 5 5 .
A5 = b5 = 4,
A10 = b10 = −1,
A25 =b25 = −3,
An = bn = 0, for all other n, n ≠ 5, 10, or 25.
Indeed, you could easily verify (do this as an exercise) that the solution
obtained this way is identical to our previous answer. Just apply the addition
formula of sine function ( sin(α ± β) = sin(α)cos(β) ± cos(α)sin(β) ) to each
term in the above solution and simplify.
2
As in the previous example, a = 9 (so a = 3), and L = 5.
Therefore, the general solution remains
∞
3nπ t 3nπ t nπ x
u ( x, t ) = ∑ An cos + Bn sin sin
n =1 5 5 5 .
nπ x nπ x
L 5
2 2
Bn =
anπ ∫
0
g ( x) sin
L
dx =
3nπ ∫0
4 sin
5
dx
80
, n = odd
= 3n 2π 2
0 , n = even
Therefore,
∞
80 3( 2n − 1)π t ( 2n − 1)π x
u ( x, t ) = ∑ sin sin
n =1 3( 2n − 1) 2 π 2 5 5 .
In addition to the fact that the constant a is the standing wave’s propagation
speed, several other observations can be readily made from the solution of
the wave equation that give insights to the nature of the solution.
To reduce the clutter, let us look at the form of the solution when there is no
initial velocity (when g(x) = 0). The solution is
∞
anπ t nπ x
u ( x, t ) = ∑ An cos sin
n =1 L L .
The sine terms are functions of x. They described the spatial wave patterns
(the wavy “shape” of the string that we could visually observe), called the
normal modes, or natural modes. The frequencies of those sine waves that
we could see, nπ / L, are called the spatial frequencies of the wave. They are
also known as the wave numbers. It measures the angular motion, in radians,
per unit distance that the wave travels. The “period” of each spatial (sine)
function, 2/( nπ / L) = 2L / n, is the wave length of each term. Meanwhile, the
cosine terms are functions of t, they give the vertical displacement of the
string relative to its equilibrium position (which is just the horizontal, or the
x-axis). They describe the up-and-down vibrating motion of the string at
each point of the string. These temporal frequencies (the frequencies of
functions of t; in this case, the cosines’) are the actual frequencies of
oscillating motion of vertical displacement. Since this is the undamped
wave equation, the motion of the string is simple harmonic. The frequencies
of the cosine terms, anπ / L (measured in radians per second), are called the
natural frequencies of the string. In a string instrument, they are the
frequencies of the sound that we could hear. The corresponding natural
periods (= 2π / natural frequency) are, therefore, T = 2L / an.
For n = 1, the observable spatial wave pattern is that of sin(πx / L). The wave
length is 2L, meaning the length L string carries only a half period of the
sinusoidal motion. It is the string’s first natural mode. The first natural
For n = 2, the spatial wave pattern is sin(2πx / L) is the second natural mode.
Its wavelength is L, which is the length of the string itself. The second
natural frequency of oscillation, 2aπ / L, is also called the second harmonic,
or the first overtone, of the string. It is exactly twice of the string’s
fundamental frequency; hence its wavelength (= L) is only half as long.
Acoustically, it produces a tone that is exactly one octave higher than the
first harmonic. For n = 3, the third natural frequency, 3aπ / L, is also called
the third harmonic, or the second overtone. It is 3 times larger than the
fundamental frequency and, at a 3:2 ratio over the second harmonic, is
situated exactly halfway between the adjacent octaves (at the second and the
fourth harmonics). The fourth natural frequency (fourth harmonic/ third
overtone), 4aπ / L, is four times larger than the fundamental frequency and
twice of that the second natural frequency. The tone it produces is, therefore,
exactly 2 octaves and 1 octave higher than those generated by the first and
second harmonics, respectively. Together, the sequence of all positive
integer multiples of the fundamental frequency is called a harmonic series
(not to be confused with that other harmonic series that you have studied in
calculus).
The motion of the string is the combination of all its natural modes, as
indicated by the terms of the infinite series of the general solution. The
presence, and magnitude, of the nature modes are solely determined by the
(Fourier sine series expansion of) initial conditions.
Lastly, notice that the “wavelike” behavior of the solution of the undamped
wave equation, quite unlike the solution of the heat conduction equation
discussed earlier, does not decrease in amplitude/intensity with time. It
never reaches a steady state (unless the solution is trivial, u(x, t) = 0, which
occurs when f (x) = g(x) = 0). This is a consequence of the fact that the
undamped wave motion is a thermodynamically reversible process that
needs not obey the second law of Thermodynamics.
Second natural mode (oscillates at the 2nd natural frequency / 2nd harmonic):
Third natural mode (oscillates at the 3rd natural frequency / 3rd harmonic):
∞
anπ t anπ t nπ x
u ( x, t ) = ∑ An cos + Bn sin sin
n =1 L L L .
nπ x
L
2
An = ∫ f ( x ) sin dx , and
L0 L
nπ x
L
2
Bn =
anπ ∫ g ( x) sin
0
L
dx .
(b) u(x, 0) = 0,
ut(x, 0) = 6.
(c) u(x, 0) = 0,
ut(x, 0) = 12sin(2x) − 16sin(5x) + 24sin(6x).
4. Verify that the D’Alembert solution, u(x, t) = [F(x − at) + F(x + at)] / 2,
where F(x) is an odd periodic function of period 2L such that F(x) = f (x) on
the interval 0 < x < L, indeed satisfies the given initial-boundary value
problem by checking that it satisfies the wave equation, boundary conditions,
and initial conditions.
a2 uxx = utt + γ ut + k u.
a2 uxx = utt + γ ut , γ ≠ 0.
Suppose boundary conditions remain as the same (both ends fixed): (0, t) = 0,
and u(L, t) = 0.
a2X ″ T = X T ″ + γ X T ′,
2
Divide both sides by a X T , and insert a constant of separation:
X ′′ T ′′ + γ T ′
= 2
= −λ .
X a T
X ″ = −λ X → X ″ + λ X = 0,
T ″ + γ T ′ = − a2 λ T → T ″ + γ T ′ + a2 λ T = 0.
T ″ + γ T ′ + α2 λ T = 0 .
n2 π 2
Eigenvalues: λ= 2 , n = 1, 2, 3, …
L
nπ x
Eigenfunctions: X n = sin
L , n = 1, 2, 3, …
Example: A flexible string of length L has its two ends firmly secured.
Assume there is no damping. Suppose the string has a weight density of 1
Newton per meter. That is, it is subject to, uniformly across its length, a
constant force of F(x, t) = 1 unit per unit length due to its own weight.
Let u(x, t) be the vertical displacement of the string, 0 < x < L, and at any
time t > 0. It satisfies the nonhomogeneous one-dimensional undamped
wave equation:
a2 uxx + 1 = utt.
The usual boundary conditions u(0, t) = 0, and u(L, t) = 0, apply. Plus the
initial conditions u(x, 0) = f (x) and ut(x, 0) = g(x).
−1 2
v( x) = x + C1 x + C 2 .
2a 2
−1 2 L
v( x) = x + x.
2a 2 2a 2
Comment: Thus, the sag of a wire or cable due to its own weight can be
seen as a manifestation of the steady-solution of the wave equation. The sag
is also parabolic, rather than sinusoidal, as one might have reasonably
assumed, in nature.
We can then subtract out v(x) from the equation, boundary conditions,
and the initial conditions (try this as an exercise), the transient
solution w(x, t) must satisfy:
∞
anπ t anπ t nπ x
w( x, t ) = ∑ An cos + Bn sin sin
n =1 L L L .
u ( x, t ) = v( x) + w( x, t )
−1 2 L ∞
anπ t anπ t nπ x
= 2 x + 2 x + ∑ An cos + Bn sin sin
2a 2a n =1 L L L
nπ x
L
2
An = ∫ ( f ( x) − v( x) )sin dx , and
L0 L
nπ x
L
2
Bn =
anπ ∫
0
g ( x) sin
L
dx .
Note: Since the velocity ut(x, t) = vt(x) + wt(x, t) = 0 + wt(x, t) = wt(x, t). The
initial velocity does not need any adjustment, as ut(x, 0) = wt(x, 0) = g(x).
The last type of the second order linear partial differential equation in 2
independent variables is the two-dimensional Laplace equation, also called
the potential equation. Unlike the other equations we have seen, a solution
of the Laplace equation is always a steady-state (i.e. time-independent)
solution. Indeed, the variable t is not even present in the Laplace equation.
The Laplace equation describes systems that are in a state of equilibrium
whose behavior does not change with time. Some applications of the
Laplace equation are finding the potential function of an object acted upon
by a gravitational / electric / magnetic field, finding the steady-state
temperature distribution of the (2- or 3-dimensional) heat conduction
equation, and the steady-state flow of an ideal fluid (where the flow velocity
forms a vector field that has zero curl and zero divergence).
Since the time variable is not present in the Laplace equation, any problem
of the Laplace equation will not, therefore, have any initial condition. A
Laplace equation problem has only boundary conditions.
Let u(x, y) be the potential function at a point (x, y), then it is governed by
the two-dimensional Laplace equation
uxx + uyy = 0.
Similarly, suppose u(x, y, z) is the potential function at a point (x, y, z), then it
is governed by the three-dimensional Laplace equation
Now let us take a step back and see the bigger picture: how the
homogeneous heat conduction and wave equations are structured, and how
they are related to the Laplace equation of the same spatial dimension.
∇2 u = uxx + uyy.
(As we have just noted, in the one-variable case, the Laplaian of u(x),
degenerates into ∇2 u = u″.)
α2 ∇2 u = ut,
α2 uxx = ut
α2 (uxx + uyy) = ut
a2 ∇2 u = utt,
where the constant a is the propagation velocity of the wave motion. Thus,
the homogeneous wave equations of 1-, 2-, and 3-dimension are,
respectively,
a2 uxx = utt
Now let us consider the steady-state solutions of these heat conduction and
wave equations. In each case, the steady-state solution, being independent
of time, must have all zero as its partial derivatives with respect to t.
Therefore, in every instance, the steady-state solution can be found by
setting, respectively, ut or utt to zero in the heat conduction or the wave
equations and solve the resulting equation. That is, the steady-state solution
of a heat conduction equation satisfies
α2 ∇2 u = 0,
‡
Even the electromagnetic waves are described by this equation. It can be easily shown by vector calculus
that any electric field E and magnetic field B satisfying the Maxwell’s Equations will also satisfy the 3-
dimensional wave equation, with propagation speed a = c ≈ 299792 km/s, the speed of light in vacuum.
∇2 u = 0.
This universal equation that all the steady-state solutions of heat conduction
and wave equations have to satisfy is the Laplace / potential equation!
uxx = 0,
uxx + uyy = 0,
X ″ Y + X Y ″ = 0,
X ″ Y = −X Y ″.
X ′′ Y ′′
=−
X Y
Now that the independent variables are separated to the two sides, we can
insert the constant of separation. Unlike the previous instances, it is more
convenient to denote the constant as positive λ instead.
X ′′ Y ′′
=− =λ
X Y
Y ′′
− =λ → Y ″ = −λ Y → Y ″ + λ Y = 0.
Y
X ″ − λX = 0, X(0) = 0,
The next step is to solve the eigenvalue problem. Notice that there is
another slight difference. Namely that this time it is the equation of Y that
gives rise to the two-point boundary value problem which we need to solve.
However, except for the fact that the variable is y and the function is Y,
rather than x and X, respectively, we have already seen this problem before
(more than once, as a matter of fact; here the constant L = b). The
eigenvalues of this problem are
n2 π 2
λ =σ = 2 ,
2
n = 1, 2, 3, …
b
nπ y
Yn = sin , n = 1, 2, 3, …
b
n 2π 2
X ′′ − X =0, X(0) = 0.
b2
n 2π 2 nπ
Its characteristic equation, r −
2
= 0 , has real roots r = ± .
b2 b
nπ − nπ
x x
X = C1 e b
+ C2 e b
.
X(0) = 0 = C1 + C2 → C2 = −C1.
nbπ x −nπ
x
X n = Cn e − e b .
eθ − e −θ
sinh θ = ,
2
nπ x
X n = K n sinh , n = 1, 2, 3, …
b
Combining the solutions of the two equations, we get the set of solutions
that satisfies the two-dimensional Laplace equation, given the specified
boundary conditions:
nπ x nπ y
u n ( x, y ) = X n ( x) Yn ( y ) = K n sinh sin ,
b b
n = 1, 2, 3, …
The general solution, as usual, is just the linear combination of all the above,
linearly independent, functions un(x, y). That is,
∞
nπ x nπ y
u ( x, y ) = ∑ K n sinh sin
n =1 b b .
To find the particular solution, we will use the fourth boundary condition,
namely, u(a, y) = f (y).
∞
anπ nπ y
u (a, y ) = ∑ K n sinh sin = f ( y)
n =1 b b
We have seen this story before. There is nothing really new here. The
summation above is a sine series whose Fourier sine coefficients are
bn = Kn sinh(anπ / b). Therefore, the above relation says that the last
boundary condition, f (y), must either be an odd periodic function (period =
2b), or it needs to be expanded into one. Once we have f (y) as a Fourier sine
series, the coefficients Kn of the particular solution can then be computed:
anπ nπ y
b
2
K n sinh = bn = ∫ f ( y ) sin dy
b b0 b
Therefore,
nπ y
b
bn 2
Kn =
anπ
=
anπ ∫ f ( y ) sin
b
dy
.
sinh b sinh 0
b b
1 1
u rr + u r + 2 uθθ = 0 .
r r
1 1
R′′Θ + R′Θ + 2 RΘ′′ = 0 .
r r
This equation above can be rewritten into two ordinary differential equations:
r2 R″ + rR′ − λR = 0,
Θ″ + λΘ = 0.
A0 ∞
u (r , θ ) = + ∑ ( An cos nθ + Bn sin nθ ) r n .
2 n =1
A0 ∞
u (k , θ ) = ( )
+ ∑ An k n cos nθ + Bn k n sin nθ = f (θ ) .
2 n =1
a0 ∞
f (θ ) = + ∑ (an cos nθ + bn sin nθ ) .
2 n =1
For a problem on the unit circle, whose radius k = 1, the coefficients An and
Bn are exactly identical to, respectively, the Fourier coefficients an and bn of
the boundary condition f (θ).