Martins Job
Martins Job
A Thesis
Presented to
In Partial Fulfillment
Master of Science
in
Mathematics
By
June 2016
c 2016
ii
The thesis of Jorge Dimas Granados Del Cid is approved.
June 2016
iii
ABSTRACT
By
cluding their applications, classifications, and methods of solving them. We show the
tion (with lower order terms), as well as other cases. We derive the solutions of some
partial differential equations of 2nd order using the method of separation of variables.
mixed, periodic and Robin. A discussion of the eigenvalues related to various bound-
the coefficients of the series solutions, is included. The thesis concludes with a pre-
iv
TABLE OF CONTENTS
Abstract................................................................................................................. iv
Chapter
1. Introduction .............................................................................................. 1
4. Eigenvalues................................................................................................ 46
5. Coefficients ................................................................................................ 53
v
5.1. Coefficients in the Case of Dirichlet Boundary Conditions............. 53
vi
LIST OF FIGURES
Figure
3. cosh γl 6= 0................................................................................................ 39
11. for a0 < 0, a0 + a1 < −a0 al l, one negative eigenvalue where the func-
vii
CHAPTER 1
Introduction
entities that include an unidentified, multivariable function and its partial derivatives.
Definition 1.1. A partial derivative is the derivative with respect to one variable of
u(x + h, y, z, . . .) − u(x, y, z, . . .)
lim
h→0 h
These partial derivatives of u, with respect to independent variables such as x, y,
.
t. . . , are written as
or
∂ ∂2 ∂2 ∂3
u, u, u, u, . . .
∂x ∂x2 ∂y∂x ∂x3
we write
∂2
uxy = (ux )y = u,
∂y∂x
If we have a function of one variable, say x, then the only partial derivative of
∂f
f (x), is just the derivative f 0 (x) and equations involving functions of one variable
∂x
and their derivatives are called ordinary differential equations (ODEs).
ux + tutt = t2 (1.0.2a)
1
ut − k 2 uxx = cos t (1.0.2b)
ut (x, t) + x2 + t2 ux (x, t) = 0,
(1.0.2g)
Definition 1.2. The order of an ODE, or a PDE equation is the maximal number of
derivatives (or partial derivatives, respectively) taken with respect to the independent
variable(s).
third order, and in (1.0.2g), the first u has been differentiated just once, and so has
Definition 1.3. L is linear if for any fucntions u and v and constant c we have
slightly different way; we examine the factorization of the differential operator L and
ux + uy = 0 transport (1.0.3a)
2
ux − yuy = 0 transport (1.0.3b)
L, then L is linear. We show this with equations (1.0.3a), (1.0.3b) and (1.0.3c)
∂ ∂
Lu = ux + uy = + u (1.0.4a)
∂x ∂y
∂ ∂
Lu = ux − yuy = −y u (1.0.4b)
∂x ∂y
∂ ∂
Lu = ux + uuy = +u u (1.0.4c)
∂x ∂y
The differential operator in (1.0.4a) is linear because there is no u in it; that is, no u
∂ ∂
in the differential operator ∂x + ∂y ; in (1.0.4b), the y in the differential operator
makes no difference, so ux − yuy is linear. And equation (1.0.4c) is not linear because
= ux + vx + uy + vy
= (ux + uy ) + (vx + vy )
= Lu + Lv,
3
and
functions u and v,
L(u + v) = (u + v)x + (u + v)y + 1
= ux + vx + uy + vy + 1
but
Lu + Lv = ux + uy + 1 + vx + vy + 1
= ux + vx + uy + vy + 2
this means
L(u + v) 6= Lu + Lv,
However, given
ut − uxx + 1 = 0
ut − uxx = −1,
and think of it as Lu = −1; the left side is linear (u can be factored from the
4
g 6= 0 and Lu = g is homogeneous if g = 0.
cos xy 2 ux − y 2 uy = tan x2 + y 2
(1.0.5a)
is a homogeneous equation.
5
CHAPTER 2
possible forms, called canonical form. The most general case of second-order linear,
where the coefficients A, B, and C are functions of x and y and do not vanish
simultaneously...[1, p 57].
B 2 − 4AC
parabolic, if B 2 − 4AC = 0
6
Example 2.2.
ut − γuxx = 0 (2.0.4)
therefore, it is of the elliptic PDE type; while (2.0.3) has the discriminant of the form
B 2 − 4AC > 0, connoting a hyperbolic PDE type; and (2.0.4) is a parabolic type,
be of one type at a set of points, and another type at some other points; this may
uξξ − uηη + · · · = 0,
uξξ + · · · = 0,
uξξ + uηη + · · · = 0.
The dots represent the terms involving u and its first partial derivatives ux and uy
7
only. We will use the change of variables
and the chain rule to transform the general equation Auxx + Buxy + Cuyy + Dux +
uξξ − uηη + · · · = 0,
uξξ + · · · = 0,
uξξ + uηη + · · · = 0
[1, p. 58].
form starts by stating the partial derivatives of u with respect to x and y in terms of
partials of ξ and η. v will be used on the right side when applying the chain rule to
u(x, y) = v(ξ(x, y), η(x, y)) to compute the second-order functions uxx , uxy , and uyy :
ux = vξ ξx + vη ηx
8
uy = vξ ξy + vη ηy
Buxy = vξξ Bξx ξy + vξη Bξx ηy + vξη Bξy ηx + vηη Bηx ηy + · · · , (2.1.5)
= avξξ
9
= bvξη
= cvηη
consequently we have from the transformation ξ = ξ(x, y), η = η(x, y), and chain
∂ (ξ, η) ξx ξy
= = ξx ηy − ξy ηx 6= 0
∂ (x, y) ηx ηy
tions whose Jacobians are different than zero....we conclude that the type
10
The following formula multiplies the discriminant of the general PDE formula Auxx +
transformation of coordinates.
(b)2 − (4ac)
B 2 − 4AC (ξx ηy − ξy ηx )2
the canonical forms by taking coefficients a and c as polynomials and then complete
11
the square to find the roots, this will make coefficients a and c zero and the reduction
will result in a hyperbolic type PDE. We proceed in the fallowing manner; take both
2 2
ξx ξx ηx ηx
A +B + C = 0, A +B +C =0
ξy ξy ηy ηy
ξx B 1 p 2
+ =± (B − 4AC)
ξy 2A 2A
ξx B 1 p 2
=− ± (B − 4AC)
ξy 2A 2A
12
p
ξx −B ± (B 2 − 4AC)
we have the two roots =
ξy 2A
ξx ηx
pick one root for ξy
, and one root for ηy
p p
ξx −B − (B 2 − 4AC) ηx −B + (B 2 − 4AC)
= = .
ξy 2A ηy 2A
dξ = ξx dx + ξy dy = 0,
along the coordinate line ξ(x, y) = constant. The total derivative re-introduces orig-
dy ξx
dξ = ξx dx + ξy dy = 0 into =− .
dx ξy
dy ηx
=− .
dx ηy
ξx ηx
Replacing and above, we get
ξy ηy
√ √
dy −B − B 2 − 4AC dy −B + B 2 − 4AC
− = , − = . (2.2.1)
dx 2A dx 2A
13
We solve for c1 and c2
√ √
B+ B 2 − 4AC B− B 2 − 4AC
c1 = y − x, c2 = y − x
2A 2A
ξy = 1, ηy = 1.
∂(ξ, η) ξx ξy 1√ 2
= =− B − 4AC 6= 0.
∂(x, y) ηx ηy A
to get:
14
=0
1
B 2 − 4AC vξη + 0 · vηη + · · ·
0 · vξξ −
A
1
B 2 − 4AC vξη + 0 + · · ·
=0−
A
1
B 2 − 4AC vξη + · · ·
=− as needed
A
15
Therefore, for A 6= 0,
vξη + · · · = 0 (2.2.2)
vξξ − vηη + · · · = 0.
∂ ∂ ∂ ∂
= − v
∂ξ ∂ξ ∂η ∂η
∂ ∂ ∂ ∂ ∂ ∂ ∂ ∂
= − + − v
∂ξ ∂ξ ∂ξ ∂η ∂ξ ∂η ∂η ∂η
∂ ∂ ∂ ∂
= − + v;
∂ξ ∂η ∂ξ ∂η
ξ+η = α
ξ − η = β,
then we have
β+α α−β
ξ= , η= ,
2 2
16
take partial derivatives of η and ξ with respect to α and β
1 1
ξα = and ξβ =
2 2
1 1
ηα = and ηβ = −
2 2
with respect to ξ and η of u(α, β) = v(ξ(α, η), η(α, β)). By the chain rule
uα = vξ ξα + vη ηα
1 1 1 1 1 1 1 1
uαβ = vξξ · + vξη · − + vηξ · + vηη − + ···
2 2 2 2 2 2 2 2
1 1
= vξ2 + [0] − vη2 + · · ·
4 4
= vξξ − vηη + · · ·
= 0.
17
2.4 Parabolic Reduction
form; this implies a and b or c and b must be zero, and B 2 − 4AC = 0. We set
2
ξx ξx
A +B +C =0. (2.4.1)
ξy ξy
dξ = ξx dx + ξy dy = 0 ,
and gives
ξx dy
−ξx dx = ξy dy =⇒ − =
ξy dx
2 2
ξx ξx dy dy
A +B +C =0 becomes A −B +C =0
ξy ξy dx dx
and the quadratic formula gives one number since we must have B 2 − 4AC = 0; so,
let
√
dy B± 0 B
− =− =− .
dx 2A 2A
dy B
So = gives
dx 2A
B
dy = dx
2A
B
y = x + c1
2A
18
B
c1 = y − x
2A
B
⇒ξ = y− x
2A
and b gives
B 2 − 4AC ηy ,
=
B
ξ=y− and η = x.
2A
B
ξx = − and ξy = 1
2A
ηx = 1 and ηy = 0
19
and the Jacobian is not zero:
∂(ξ, η) ξ ξ
= x y = −1 6= 0. (2.4.2)
∂(x, y) ηx ηy
holds and “shows that the sign of the discriminant B 2 − AC remains invariant”
[1, p. 60].
So, we reduce avξξ + bvξη + cvηη · · · = 0 to a canonical form of the parabolic type; we
take the partial derivatives of η and ξ with respect to x and y and plug them into
then we have
2
B B
a = A − +B − (1) + C (1)2
2A 2A
1 2
= C− B
4A
1
B 2 − 4AC
= −
4A
= 0
1
= − (0)
4A
B B
b = 2A − (1) + B − (0) + (1) (1)
2A 2A
+2C (1) (0)
= −B + B
20
= 0
=⇒ vηη · · · = 0.
The discriminant in this case should be b2 − 4ac < 0; so we can set b = 0 and c = a,
or a − c = 0.
we have that
21
this gives a “coupled” system; that is, both coordinates ξ and η, show up in both
equations,
giving
2
ξx + iηx ξx + iηx
A +B + C = 0,
ξy + iηy ξy + iηy
ξx + iηx
a quadratic equation. Let = φ; dividing by A, we get
ξy + iηy
B C
φ2 + φ + = 0,
A A
2 2
2 B B C B
φ + φ+ =− +
A 2A A 2A
2 2
B B C
φ+ = −
2A 2A A
22
s
2
B B C
φ+ =± −
2A 2A A
s
2
B B C
φ+ =± −
2A 2A A
r
B B2 C
φ+ =± 2
−
2A s 4A A
B 4A2 B 2 C
φ+ =± −
2A 4A2 4A2 A
s 2
B 1 B 4A2 C
φ+ =± −
2A 4A2 4A2 A
s
B 1 4A2 B 2 4A2 C
φ+ =± −
2A 2A 4A2 A
B 1 p 2
φ+ =± (B − 4AC)
2A 2A
B 1 p 2
φ=− ± (B − 4AC)
2A p2A
−B ± (B 2 − 4AC)
φ= .
2A
αx βx
we have complex roots. Let φ = for one root, and φ = for the other root; so,
αy βy
p p
αx −B + i (4AC − B 2 ) βx −B − i (4AC − B 2 )
= , = ,
αy 2A βy 2A
αx βx
these are complex conjugates. The total derivative replaces and :
αy βy
dα = αx dx + αy dy = 0
23
dy αx
gives =− ;
dx αy
dy βx
=− ,
dx βy
then we have
p p
B − i (4AC − B 2 ) B + i (4AC − B 2 )
dy = dx , dy = dx,
2A 2A
Let c1 = α, and c2 = β
p p
B−i (4AC − B 2 ) B + i (4AC − B 2 )
α=y− x, β=y− x, (2.5.2)
2A 2A
∂2 ∂2
vξξ + vηη = + v
∂ξ 2 ∂η 2
∂ ∂ ∂ ∂
= + v
∂ξ ∂ξ ∂η ∂η
∂ ∂ ∂ ∂ ∂ ∂ ∂ ∂
= + − + v
∂ξ ∂ξ ∂ξ ∂η ∂ξ ∂η ∂η ∂η
∂ ∂ ∂ ∂ ∂ ∂
= + − + v
∂ξ ∂ξ ∂η ∂η ∂ξ ∂η
24
∂ ∂ ∂ ∂
= + − v,
∂ξ ∂η ∂ξ ∂η
1 1 1 1
ξ = α+ β and η = iβ − iα.
2 2 2 2
1 1 √
give ξ=y− Bx and η= x 4AC − B 2 .
2A 2A
Taking partial derivatives of ξ and η we have
1 √
ηy = 0, ‘ ηx = AC − B 2 ,
2A
1
ξx = − B, ξy = 1.
2A
∂(ξ, η) ξx ξy 1 √
= =− AC − B 2 6= 0.
∂(x, y) ηx ηy 2A
25
We have a = c, and need b equal to zero,
= a (vξξ + vηη ) · · · ,
26
CHAPTER 3
Separation Of Variables
where each represent a different type: the wave equation represents ‘hyperbolic’ equa-
tions; while the Laplace equation is an elliptic type of equation, and the heat equa-
tion’s type is parabolic. Some conditions have to be given for the equations above
Because PDEs typically have so many solutions,. . . we single out one so-
27
ux − a0 u = f3 (t) at x = 0
ux + al u = f4 (t) at x = l
∂u
is used when and u is applied at the boundary.
∂n
Solving a linear, homogeneous, second order ODEs using the characteristic polynomial
method.
where X = X(x), with a, b and c constants. We look for a solution X(x) in the form
X = erx , then
0
X = rerx ,
00
X = r2 erx ,
00 0
use this to insert into aX + bX + cX = 0, this gives
erx ar2 + br + c = 0,
ar2 + br + c = 0. (3.2.2)
28
√ √
−b + b2 − 4ac −b − b2 − 4ac
r1 = , r2 = . (3.2.3)
2a 2a
These roots, (3.2.3), give solution cases for the differential equation
00 0
aX + bX + cX = 0 where X(x) = erx .
if r1 6= r2 & b2 − 4ac < 0 (these are complex roots) the solutions have the form
= eκx [(c1 cos γx + ic1 sin γx) + (c2 cos γx − ic2 sin γx)]
29
3.3 Dirichlet boundary conditions
00
next is to have XT = c2 X 00 T be divided by c2 XT
00
X(x)T (t) c2 X 00 T
= = −λ,
c2 XT c2 XT
X 00 + λX = 0 and T 00 + c2 λT = 0,
X 00 + β 2 X = 0, and T 00 + c2 β 2 T = 0.
c = β 2 , this implies that since r1 6= r2 & 02 − 4β 2 < 0, the solution has the form
30
where κ = 0 and γ = β in eκx [C1 cos γx + C2 sin γx] .
= A(1) + B(0)
this gives A = 0.
So X (x) = A cos (βx) + B sin (βx) becomes X(x) = B sin(βx) and the right
X(l) = B sin(βl) = 0.
If B = 0, then u(x, y) = X(x)T (t) = X(0)T (t) = 0; this is trivial. Instead, we let
nπ
βl = nπ, where nπ are zeros (roots) of the sine function; so, β = , and
l
nπ 2 πx
βn2 = = λn , Xn (x) = sin (n = 1, 2, 3, . . .)
l l
For T, there are no boundary conditions so we have that u(x, t) = X(x)T (t) becomes
nπct nπct nπx
un (x, t) = Cn cos + Dn sin sin , (3.3.5)
l l l
n is a set of integers from one to infinity and Cn and Dn are constants. So, we
have infinitely many solutions because for each n, we have a different un and we
31
get a sum of solutions as another solution of utt = c2 uxx for 0 < x < l and
u(0, t) = 0 = u(l, t) :
X nπct nπct
nπx
u(x, t) = Cn cos + Dn sin sin , (3.3.6)
n
l l l
X nπx
and (3.3.6) “solves equations (3.3.1), (3.3.2) and (3.3.3) provided that φ(x) = An sin
l
X nπc nπx
and ψ(x) = Bn sin ” [4, p. 89].
l l
0 00
T X
= = −λ = constant,
kT X
32
00
and, as before, the boundary conditions give for −X = λX
nπ 2 nπx
λn = , Xn (x) = sin (n = 1, 2, 3, . . . ). (3.4.4)
l l
0
T d
For T (with = ln |T |), we have
T dt
0
T
0= = −β 2 k
T
d
ln |T | = −β 2 k
dt
nπ 2
ln |Tn | = − kt + c
l
nπ 2
− kt + c
Tn = e l
nπ 2
− kt
≡ An e l .
∞
2 nπx
An e−(nπ/l) kt sin
X
u(x, t) = (3.4.5)
n=1
l
Note that Tn is an exponential function; also note that we have just one initial condi-
tion because the diffusion equation contains only one partial derivative with respect
∞
X nπx
φ(x) = An sin . (3.4.6)
n=1
l
Example 3.2. Consider waves in a resistant medium that satisfy the problem
33
2πc
where r is a constant and 0 < r < . Write down the series expansion of the solu-
l
tion.
write
then
T 00 (t) X 00 (x) T 0 (t)
− + r = 0,
c2 T (t) X (x) c2 T (t)
For T we have
1 00 r
2
T (t) + 2 T 0 (t) + λT (t) = 0,
c c
34
and multiplying c2 on both sides gives:
We use the characteristic equation µ2 + rµ + λc2 = 0, to find the roots for T ; by the
√
q
nπ 2 2
−r ± r2 − 4c2 λn −r ± r2 − 4 l
c
µn = =
2 2
2πc
Given that 0 < r < , we have
l
2
4π 2 c2 4π 2 c2 n2
2πc 2πc
r< =⇒ r2 < = ≤ , n = 1, 2, 3, . . . ,
l l l2 l2
so q
nπ 2 2
r±i 4 − r2
r
l
c r nπ 2 2 r2
µn = − =− ±i c − .
2 2 l 4
Then,
r r
− ± iwn t nπ 2 2 r2
Tn (t) = e 2 , where wn = 4 c −
l 4
and
r
− ± iwn t
Tn (t) = e 2
r
− t ±iw t
= e 2 e n
r
− t
= e 2 (cos wn t ± i sin wn t) .
Hence,
r ∞ r
− tX nπx nπ 2 2 r2
un (x, t) = e 2 (cos wn t + i sin wn t) sin , where wn = 4 c − .
n=1
l l 4
35
3.5 Neumann Boundary Conditions
Separation of variables, starting with X(x) as previously found for the wave equation:
0
0 = X (0)
= 0 + Dβ(1),
this gives
2 kt nπx
u(x, t)n = An e−(nπ/l) cos (3.5.3)
l
For Dirichlet boundary conditions in last section, it could be shown that λ 6= 0; but
for Neumann’s boundary conditions, λ can be zero and λ = 0 adds the term, 12 A0 to
36
what we have; so now (3.5.3) looks like
∞
1 X 2 nπx
u(x, t) = A0 + An e−(nπ/l) kt cos , (3.5.4)
2 n=1
l
nπ 2
λ = β2 = for n = 0, 1, 2, 3, . . . . (3.5.5)
l
For Neumann BCs in (3.5.4), we have the cosine series in it, instead of the sine series
of Dirichlet BCs. With Neumann’s, now the initial conditions (at t = 0) looks like
∞
1 X nπx
u(x, 0) = φ(x) = A0 + An cos . (3.5.6)
2 n=1
l
The cosine series. For Neumann’s boundary conditions, ux (0, t) = ux (l, t) , we get
∞
1 X 2 nπx
φ (x) = u (x, 0) = A0 + An e−(nπ/l) k(0) cos
2 n=1
l
∞
1 X nπx
= A0 + An cos .
2 n=1
l
Example 3.3. Solve the diffusion problem ut = kuxx in 0 < x < l, with the mixed
37
So X 00 (x) + λX (x) = 0, then if λ > 0, we get
The boundary condition u (0, t) = 0 implies X (0) T (t) = 0 for all t =⇒ X (0) = 0,
= C + D (0)
=⇒ C = 0,
cos βl = 0
n + 12 π
1
⇒ βn l = n + π so βn = ,
2 l
1
!2
n + π
thus λn = βn2 = 2
, n = 0, 1, 2, . . . .
l
n + 12 π
And Xn (x) = Dn sin x.
l
For T,
38
Note: there are no negative eigenvalues, for if λ < 0, we can write λ = γ 2 (where WLOG γ > 0) ,
then
0 = X (0) = A · 1 + B · 0
⇒ A = 0,
Figure 3: cosh γl 6= 0
39
3.6 The Robin boundary conditions
X 0 − a0 X = 0 at x = 0
X 0 + al X = 0 at x = l.
where a0 and al are constants. For X (x) = C cos βx + D sin βx with the Robin
0 = X 0 (0) − a0 X (0)
= βD − a0 (C)
a0 (C)
gives D =
β
similarly, at x = l
0 = X 0 (l) + al X (l)
40
β 2 − al a0 sin βl = (a0 + al ) β cos βl
(Struass, p.94).
a0 (C)
If C 6= 0, and D = , we get the corrsponding eigenfunction
β
Ca0
X (x) = C cos βx + sin βx.
β
so will use graphing to analyze numerical values of β. Dividing (3.6.1) by cos βl, we
sin βl
get the trigonometric identity = tan βl and write (3.6.1) as
cos βl
(a0 + al ) β
tan βl = ,
β 2 − al a0
(a0 + al ) β
and find intersections of tan βl and , as functions of β > 0.
β 2 − al a0
41
In figure (7), the eigenvalues are between zeros of tan βl; this, and the inter-
π2 2π
2
n2 < βn
2
= λn < (n + 1) (n = 0, 1, 2, 3, . . .). (3.6.2)
l2 l2
π π
Note that in figure (7), when cos βl = 0, βl = , and sin = 1, (β 2 − al a0 ) sin βl =
2 2
(a0 + al ) β cos βl, gives
β 2 − al a0 sin βl
0 =
β 2 − al a0
=
√
=⇒ β = al a0 ,
this is when “the tangent function and rational function ’intersect at infinity’”
[4, p 94].
For the case of a0 < 0 al > 0, and a0 + al > 0, the maximum occurs at
p (−a0 + al ) β
|−a0 al |; this can be shown by finding the critical point where 2 reaches
β − (−a0 )al
the maximum (see Figure (5)). We use the quotient rule for derivatives:
0
(−a0 + al ) (β 2 + a0 al ) − (−a0 + al ) β (2β)
(−a0 + al ) β
=
β 2 + a0 al (β 2 + a0 al )2
and set the right side to zero, (implies the numerator is zero) and get
(−a0 + al ) β 2 + a0 al = (−a0 + al ) 2β 2
⇒ β 2 + a0 al = 2β 2
⇒ β 2 = a0 al
p
⇒ β= |a0 al | .
42
(−a0 + al ) β
The intersections of tan βl = are shown in figure (5),
β 2 + al a0
(−a0 + al ) β
Figure 5: a0 < 0, al > 0 and a0 + al > 0, tan βl = .
β 2 + al a0
Example 3.4. Find the eigenvalues graphically for the boundary conditions X (0) =
Solution: We have
0 = βD cos βl + aD sin βl
43
β
The intersection of functions − and tanβl, will give the eigenvalues:
a
Case 1: a > 0; so y = − βa < 0 : The discontinuities are at β = n − 21 π/l and the
n − 21 π
nπ
< βn < ,
l l
(n− 12 )π
and also the graph shows that limn→∞ βn − l
= 0.
44
n + 21 π n + 21 π
nπ
shows that < βn < and βn − →0 as n→∞ . So, larger
l l l
2
π2
1
eigenvalues get closer to 2 n + .
l 2
Example 3.5. We will use Newton’s Method of iterations to comopute the first
a0 = al = 1.9, where
f (βn )
βn+1 = β − ,
f 0 (βn )
(tan πβ) (β 2 − 3. 61) − 3. 8β
βn+1 = β −
π (tan2 πβ + 1) (β 2 − 3. 61) + (tan πβ) 2β − 3. 8
45
CHAPTER 4
Eigenvalues
Av = λv (4.0.1)
where v 6= 0.
d2
Analogously, for the differential operator − we have that
dx2
00
X (x) = −λX(x)
d2
=⇒ − X(x) = λX(x),
dx2
d2
where λ is an eigenvalue of − for a nonzero function X(x).
dx2
If λ = 0, then
X 00 = 0
X 0 (x) = D
X (x) = Dx + C
implies C = 0
so X(x) = Dx
46
for X(l) = 0: 0 = X(l) = D(l)
Negative Eigenvalues?
X 00 = − −γ 2 X = γ 2 X.
Suppose λ is negative, then we have two real solutions of the form erx , then by
section (3.2), X (x) = c1 eγx +c2 e−γx are solutions. For Dirichlet boundary conditions,
X(x) = 0, we have
= c1 (1) + c2 (1)
this gives − c1 = c2
= c1 eγl − e−γl
= eγl − e−γl
we get γl = −γl
47
not allow negative eigenvalues.
ux (0, t) = ux (l, t) = 0.
Zero Eigenvalues
constants.
left BC X 0 (0) = A
gives A = 0
then X (x) = B
and X 0 (x) = 0.
righ BC X 0 (l) = 0
Then X (x) = B is not zero; therefore, for the Neumann’s boundary conditions we
Negative Eigenvalues?
method
X 00 = − (−γ 2 ) X = γ 2 X
implies r = ±γ.
48
at X 0 (0): 0 = X 0 (0) = γc1 eγ(0) − γc2 e−γ(0)
0 γ
= (c1 − c2 )
γ γ
gives − c1 = c2
For − X 00 = λX
Zero Eigenvalues?
implies 0 = D − a0 (C + D (0))
= D − a0 C
49
then D = a0 C
so X (x) = C + (a0 C) x
gives = a0 C + al (C + a0 Cl)
= a0 C + al C + al a0 Cl
that is, if C = 0 or 0 = a0 + al + al a0 l
so 0 = a0 + al + al a0 l
Negative Eigenvalues?
= 0 + Bγ − a0 A − 0
a0 A
we get B =
γ
a0 A
so X (x) = A cosh γx + sinh γx
γ
Then 0 = X 0 (l) + al X (l)
50
a0 A
= Aγ sinh γl + a0 Acoshγl + al A cosh γx + sinh γl
γ
γ
0 = γ 2 sinh γl + a0 γ cosh γl + al γ cosh γl + al a0 sinh γx
A
γ 2 + al a0 sinh γl
= − (a0 + al ) γ cosh γl
sinh γl
γ 2 + al a0 = − (a0 + al ) γ
cosh γl
(a0 + al ) γ
tanh γx = − 2
γ + al a0
We graph both sides; if there is an intersection, then we will have a negative eigen-
51
Figure 11: for a0 < 0, a0 + a1 < −a0 al l, one negative eigenvalue where the functions
intersect.
52
CHAPTER 5
Coefficients
tion u(x, 0)
∞
X nπc (0) nπc (0) nπx
φ (x) = u (x, 0) = Cn cos + Dn sin sin
n=1
l l l
∞
X nπx
= (Cn + 0) sin
n=1
l
∞
X nπx
= Cn sin
n=1
l
∞
X nπx mπx
This gives φ (x) = Cn sin , then multiply both sides by sin , and integrate
n=1
l l
from 0 to l term by term
Z l ∞ Z l
mπx X nπx mπx
φ (x) sin dx = Cn sin sin dx
0 l n=1 0 l l
combine cosines
cos (a − b) − cos (a + b) = cos a cos b + sin a sin b − (cos a cos b − sin a sin b)
= 2 sin a sin b
53
1 1
and sin a sin b = cos (a − b) − cos (a + b)
2 2
n−m
Now we find the coefficient; first, we integrate for m 6= n; and we will need θ1 = πx,
l
l n+m l
dθ1 = dx, similarly, θ2 = πx, dθ2 = dx
n−m l n+m
Z l
1 nπx mπx 1 nπx mπx
cos − − cos + dx
0 2 l l 2 l l
Z l
1 n−m 1 n+m
= cos πx − cos πx dx
0 2 l 2 l
Z l Z l
1 n−m 1 n+m
= cos πx dx − cos πx dx
0 2 l 0 2 l
Z l Z l
1 l 1 l
= cos (θ1 ) dθ1 − cos (θ2 ) dθ2
0 2 (n − m) π 0 2 (n + m) π
l l
1 l n−m 1 l n+m
= sin πx − sin πx
2 (n − m) π l 0 2 (n + m) π l 0
1 l 1 l
= (sin ((n − m) π) − (sin 0)) − (sin ((n + m) π) − (sin 0))
2 (n − m) π 2 (n + m) π
= 0
nπx nπ
Integrate when m is fixed and m = n, we will need θ3 = and dθ3 = dx, so
l l
l
that dθ3 = dx is used in substitution at a step of integration; also, we make use
nπ
1 1
of the trigonometric identity sin2 x = − cos2x
2 2
Z l
nπx mπx
sin sin dx
0 l l
Z l
nπx
= sin2 dx
0 l
Z l
1 1 l
Z
2nπx
= − cos dx
0 2 2 0 l
54
Z l Z l
1 l 1
= dx − cos θ3 dθ3
0 2 nπ 0 2
1 1 1 l
= l− 0− (sin 2nπ − sin 0)
2 2 2 2nπ
1 1 l
= l−0− (0 − 0)
2 2 2nπ
1
= l
2
Z l
nπx nπx 0, if m 6= n,
So sin sin dx = 1
0 l l 2
l, if m = n.
Z l
mπx
And φ (x) sin dx
0 l
∞ Z l
X nπx mπx
= Cn sin sin dx
n=1 0 l l
∞ Z l
X nπx mπx
= Cn sin sin dx
n=1 0 l l
Z l
(1) πx mπx
= C1 sin sin dx
0 l l
Z l
2πx mπx
+C2 sin sin dx
0 l l
Z l
mπx mπx
+ · · · + Cm sin sin dx
0 l l
Z l
nπx (m + 1) πx
+Cm+1 sin sin dx + · · ·
0 l l
1
= 0 + 0 + · · · +Cm · + 0 + 0 + · · ·
2
Z l Z l
mπx m=m l 2 mπx
This gives φ (x) sin dx = Cm · , or Cm = , φ (x) sin dx
0 l 2 l 0 l
this is the Fourier coefficient formula for Cn . The other initial condition, ut (x, 0) , for
∞
X nπc nπct nπc nπct nπx
ut (x, t) = − Cn sin + Dn cos sin
n=1
l l l l l
55
∞
X nπc nπc nπx
and ut (x, 0) = − Cn sin 0 + Dn cos 0 sin
n=1
l l l
∞
X nπc nπx
= Dn sin
n=1
l l
∞
X nπc nπx
so ψ (x) = Dn sin
n=1
l l
Z l
nπc 2 mπx
and we get Dn = ψ (x) sin dx by the same process we got Cn .
l l 0 l
A0 ) and follow the same steps as we did in finding coefficients for Dirichlet’s boundary
then cos (a − b) + cos (a + b) = cos a cos b + sin a sin b + (cos a cos b − sin a sin b)
= 2 cos a cos b
56
and for the other integral, as we did before, we fix m and set it equal to n; plus we
Z l
2 1 cos 2x 1 cos 2x 1
use the identity cos x = + , where + dx = l + 0,
2 2 0 2 2 2
Z l Z l
nπx mπx nπx
sin sin dx = cos2 dx
0 l l 0 l
1
= l
2
2 l
Z
mπx
so we get Am ≡ φ (x) cos dx where m = 0, 1, 2, . . . .
l 0 l
∞
1 X nπx
A0 + An cos .
2 n=1
l
Example 5.1. Solve the Diffusion problem with Dirichlet boundary conditions and
initial conditions.
ut = kuxx
u (0, t) = 0 = u (l, t)
u (x, 0) = 1 = φ (x)
We know that
∞
X 2 nπx
u (x, t) = An e−(nπ/l) kt
sin .
n=1
l
∞
X 2 nπx
1 = u (x, 0) = An e−(nπ/l) k(0)
sin
n=1
l
∞
X nπx
= An sin
n=1
l
Z l
2 mπx
and Am = φ (x) sin dx, is the formula to find A.
l 0 l
57
mπx mπ l
Letting θ = , θd = dx, then θd = dx,
l l mπ
2 l
Z
l
Am = sin θ θd
l 0 mπ
l
2
= − cos θ
mπ 0
l
2 mπx
= − cos
mπ l 0
2 2
= − cos mπ + ,
mπ mπ
4
2 m 2 if m is odd
− (−1) + = mπ
mπ mπ 0 if m is even,
so we get
∞ ∞
X nπx X 4 nπx
1= An sin = sin
n=1
l n=1
mπ l
4 πx 2πx 3πx
= sin + sin + sin + ··· .
mπ l l l
The method of separation of variables is very helpful for solving linear PDEs of
2nd order. Hwoever, it has its limitations. For problems with non-constant coefficients
or for those with non-symmetric boundary conditions, the method will not work.
58
REFERENCES
[1] P. R. Garabedian,: Partial Differential Equations, John Wiley & Sons. Inc, New
[3] Carlos, A. Smith and W. Campbell, Scott,: A First Course in Differential Equa-
tions: modeling and simulation, CRC Press, Boca Raton FL, John Wiley & Sons.
[4] W. A. Strauss,: Partial Differential Equations (An introduction), John Wiley &
59