PDE Elementary PDE Text
PDE Elementary PDE Text
PDE Elementary PDE Text
William V. Smith
Introduction.
1
to bring the bibliography up to date more than 1600 items would have to be
added. The process has continued to accelerate.
At the present time, it is impossible to present in a single course a com-
plete survey of what is known as PDEs and the properties of their solutions.
Many advanced monographs exist and in many cases their contents scarcely
overlap.
Plan for this course
A study of classical theories for some of the simplest PDEs. We shall
use as a source, V. Smirnov, A Course in Higher Mathematics, vols. II and
IV, and C. H. Wilcox, “Notes on PDEs.”1 The modern functional analytic
theories of PDEs must wait for further courses.
Contents
2
Chapter 1. Heat Conduction in a Slab.
∂v ∂ 2v ∂ 2v ∂ 2v
= k∆v = k( 2 + 2 + 2 ) Heat or Diffusion Equation
∂t ∂x ∂y ∂z
∂ 2v
2
= c2 ∆v Wave Equation
∂t
∆v = 0 Laplace’s Equation
———
2
See appendix I p. 117 for other examples.
3
Temperature = v(x, t).
Equations:
∂v ∂ 2v
=k 2 for 0 < x < l, t > 0
∂t ∂x
Existence of a Solution?
Uniqueness of Solution?
∂ 2 vs
= 0, 0 < x < l, and vs (0) = vo , vs (l) = v1
∂x2
4
It follows that
l−x x
vs (x) = vo ( ) + v1 ( )
l l
∂v ∂u
= ,
∂t ∂t
∂ 2v ∂ 2u
=
∂x2 ∂x2
Thus,
∂u ∂ 2u
(1) =k 2 for 0 < x < l, t > 0
∂t ∂x
5
XT 0 = kX 00 T
X 00 (x) T 0 (t)
(1) = = const. = −λ
X(x) kT (t)
and
X 00 (x) + λX(x) = 0 for 0 < x < l
πnx
X(x) = Xn (x) = sin( ), n = 1, 2, 3, . . .
l
πn 2
λ = λn = ( )
l
πn 2
T (t) = Tn (t) = e−k( l
) t
and
nπx −k( nπ )2 t
u(x, t) = un (x, t) = sin( )e l
l
6
Superposition. To solve (1), (2), (3) try
∞
X nπx −k( nπ )2 t
(4) u(x, t) = cn sin( )e l
n=1
l
Formal Solution.
∞
X nπx
u(x, 0) = f (x) = cn sin( ) (Fourier Sine Series)
n=1
l
Z l
2 nπx
(5) cn = f (x) sin( )dx, n = 1, 2, 3, . . .
l 0 l
(5) follows from (4) by the orthogonality relations.3 The formal solution
is defined by (4), (5). Convergence theory for Fourier series may now be
applied.
7
Then the Fourier sine series coefficients (5) satisfy
∞
X
(6) |cn | < ∞
n=1
moreover
∞
X nπx
f (x) = cn sin( ) for 0 ≤ x ≤ l
n=1
l
——————————————
Existence Theorem. If f (x) satisfies (a), (b), (c) then the formal so-
lution (4), (5) converges uniformly on Ω and defines a classical solution.
Proof. (6) shows that (4) converges uniformly on Ω and hence u ∈ C(Ω)
bounded and has left and right-hand limits at each point in its domain. Naturally, all
bounded continuous functions are also sectionally continuous, the converse of course, is
not true.
5
R. V. Churchill, Fourier Series and Boundary Value Problems, 2nd ed. McGraw-Hill
1963.
8
and (2) and (3) hold (for a detailed argument that this is so, see pages .
∂u ∂ 2 u
To prove that ,
∂t ∂x2
∈ C(Ω) and (1) holds note that
nπx −k( nπ )2 t nπ 2
|cn sin e l | ≤ |cn |e−k( l ) to
l
∞
X nπx −k( nπ )2 t
n2 cn sin e l
n=1
l
∂u ∂u ∂ 2 u
converges uniformly for all x ∈ R and t ≥ to .6 It follows that , ,
∂t ∂x ∂x2
∈
C(Ω) can all be calculated by termwise differentiation and are continuous for
all x ∈ R, t > 0. Finally, (1) holds because each term in (4) is a solution of
the heat equation. QED
ΩT = Ω ∩ {(x, t) : t < T }
ΩT = closure of ΩT = Ω ∩ {(x, t) : t ≤ T }
∂Ω = boundary of Ω = Ω − Ω
6
The Weierstrass M-Test may be used. See page 142-3.
9
ΓT = ∂Ω ∩ ΩT
Proof. (By contradiction.) Assume the conclusion is false, i.e., maxΩT u (>
maxΓT u) occurs at (xo , to ) ∈ ΩT − ΓT . Define the function
∂ 2v ∂v
(x1 , t1 ) ≤ 0, (x1 , t1 ) ≥ 0
∂x2 ∂t
10
whence
∂ 2u ∂ 2v
(x 1 , t1 ) = (x1 , t1 ) ≤ 0
∂x2 ∂x2
but
∂u ∂v
(x1 , t1 ) = (x1 , t1 ) + > 0
∂t ∂t
max u = max u = 0
ΩT ΓT
i.e.7
u(x, t) ≤ 0 ∀(x, t) ∈ ΩT
7
∀ is a logical symbol which simply means “for all.”
11
The maximum principle implies that if f (x) satisfies (a), (b), (c) and
u(x, t) is the corresponding classical solution then
This implies
12
Chapter 2. Wave Propagation on a Taut String.
∂ 2u 2
2∂ u
= c , c>0
∂t2 ∂x2
Interpretation
∂ 2u ∂ 2u
(1) =
∂τ 2 ∂x2
13
Assume that u ∈ C 2 (Ω) and (1) holds for all (x, τ ) ∈ Ω. Then there exist
functions f (τ ), g(τ ) such that
ξ = x − τ, η =x+τ
and let
u(x, τ ) = v(ξ, η)
∂u ∂v ∂v ∂u ∂v ∂v
= + , =− +
∂x ∂ξ ∂η ∂τ ∂ξ ∂η
∂ 2u ∂ 2v ∂ 2v ∂v ∂ 2u ∂ 2u ∂ 2v ∂ 2v
= + 2 + , = − 2 +
∂x2 ∂ξ 2 ∂ξ∂η ∂η 2 ∂τ 2 ∂ξ 2 ∂ξ∂η ∂η 2
9
That is, the pair (f, g) is equivalent to the pair (f + C, g − C) where C is any constant.
14
Thus
∂ 2u ∂ 2u ∂ 2v
− = 4 = 0 in Ω0 = {(ξ, η) : (x, τ ) ∈ Ω}
∂x2 ∂τ 2 ∂ξ∂η
∂v(ξ, ηo )
= G(ηo ) ∈ C 1 (b1 , b2 ).
∂η
Z
g(η) = G(η)dη ∈ C 2 (b1 , b2 )
15
[See chalkboard illustration]
Let u ∈ C 2 (Ω) describe a motion of the string. By the lemma the values
of u in ∆ ⊂ ∆o are independent of what happens at the ends of the string
and ∃f, g ∈ C 2 (a, b)10 such that
u(x, τ ) = f (x − τ ) + g(x + τ ) in ∆
∂u(x, 0)
u(x, 0) = uo (x) and = u1 (x), a<x<b
∂τ
Indeed,
∂u(x, 0)
= −f 0 (x) + g 0 (x) = u1 (x)
∂τ
Thus
Z x Z x
2f (x) = uo (x) − u1 (ξ)dξ + C, 2g(x) = uo (x) + u1 (ξ)dξ + C 0
a a
10
∃ is a logical symbol meaning “there exists.”
16
Hence
whence
C + C 0 = 0 or C 0 = −C
Thus
1 x−τ
Z
1
u(x, τ ) = uo (x − τ ) − u1 (ξ)dξ + C
2 2 a
1 x+τ
Z
1
+ uo (x + τ ) + u1 (ξ)dξ − C
2 2 a
or
Z x+τ
1
(2) u(x, τ ) = {uo (x − τ ) + uo (x + τ )} + u1 (ξ)dξ, (x, τ ) ∈ ∆.
2 x−τ
17
∂ 2u ∂ 2u
(3) = for − ∞ < x < ∞, τ > 0
∂τ 2 ∂x2
∂u(x, 0)
(4) u(x, 0) = uo (x) and = u1 (x) for −∞<x<∞
∂τ
Uniqueness. The argument leading to (2) shows that any classical so-
lution must be given by (2). Hence classical solutions are unique.
18
Wave Propagation on a String of Finite Length.
∂ 2u ∂ 2u
(5) = for 0 < x < l, τ > 0
∂τ 2 ∂x2
∂u(x, 0)
(7) u(x, 0) = uo (x) and = u1 (x) for 0 ≤ x ≤ l
∂τ
u(x, τ ) = X(x)T (τ )
XT 00 = X 00 T
19
or
T 00 X 00
= = const. = −λ
T X
consequently
(8) T 00 + λT = 0, τ >0
Observe that (9) is the same BV problem which occurred in the reduction
of the heat equation. Hence
nπ 2 nπx
λ = λn = ( ), X(x) = Xn (x) = sin( ), n = 1, 2, 3, 4, . . .
l l
nπτ nπτ
T (τ ) = Tn (τ ) = An cos( ) + Bn sin( )
l l
20
nπτ nπτ nπx
un (x, τ ) = {An cos( ) + Bn sin( )} sin( ), n = 1, 2, 3, 4, . . .
l l l
∞
X nπτ nπτ nπx
(10) u(x, τ ) = {An cos( ) + Bn sin( )} sin( )
n=1
l l l
∞
∂u(x, τ ) X nπ nπτ nπτ nπx
= {−An sin( ) + Bn cos( )} sin( )
∂τ n=1
l l l l
Formal Solution.
∞
X nπx
u(x, 0) = uo (x) = An sin( )
n=1
l
∞
∂u(x, 0) X nπ nπx
= u1 (x) = Bn sin( )
∂τ n=1
l l
Thus
2 l
Z
nπx
(11) An = uo (x) sin dx
l 0 l
2 l
Z
nπ nπx
Bn = u1 (x) sin dx n = 1, 2, 3, 4, . . .
l l 0 l
21
define the
∂u(x, 0)
= u1 (x) ≡ 0
∂τ
∞
0
X nπτ nπx
(10) u(x, τ ) = An cos sin
n=1
l l
Z l
0 2 nπx
(11) An = uo (x) sin dx
l 0 l
22
nπτ nπx nπ(x − τ ) nπ(x + τ )
(12) 2 cos sin = sin + sin
l l l l
∞ ∞
1X nπ(x − τ ) 1 X nπ(x + τ )
(13) u(x, τ ) = An sin + An sin
2 n=1 l 2 n=1 l
The two series in this expression are Fourier sine series for uo (x) evaluated
at x − τ and x + τ respectively. Now if uo (x) has a convergent sine series
then ∀ξ ∈ R
uo (ξ) if 0 ≤ ξ ≤ l;
∞
X nπξ
Uo (ξ) = An sin = Odd periodic extension of uo (ξ)
n=1
l
with period 2l
for all other values of ξ.
Thus, formally,
1
(14) u(x, τ ) = [Uo (x − τ ) + Uo (x + τ )]
2
This will define a classical solution if and only if Uo ∈ C 2 (R). This is true
23
if and only if
(15) uo ∈ C 2 [0, l]
and
∂ 2u ∂ 2u
(1) u ∈ C 1 (Ω) ∩ C 2 (Ω) and = in Ω
∂τ 2 ∂x2
∂u(x, 0)
(2) u(x, 0) = uo (x) and = u1 (x) for 0 ≤ x ≤ l
∂τ
24
It will be shown that Lemma 1 leads to a unique construction of u(x, τ )
in Ω.
ξ = x − τ, η =x+τ
Step 2.
25
Z ξ
1 1
(4) f (ξ) = uo (ξ) − u1 (x)dx ∀ 0≤ξ≤l
2 2 0
Z η
1 1
(5) g(η) = uo (η) − u1 (x)dx ∀ 0≤η≤l
2 2 0
Step 3.
or
Step 4.
or
(7) f (l − τ ) = −g(l + τ ), ∀τ ≥ 0
26
(8) g(τ + 2l) = g(τ ), ∀τ ≥ 0
Proof.
14
In other words, g is 2l- periodic and f nearly so.
27
• (5), (6) extend f to [−l, 0]
Specifically,
1 1
Rξ
u (ξ) − u1 (x)dx, 0≤ξ≤l
2 o
2 0
f (ξ) =
− 1 uo (−ξ) − 1 −ξ u1 (x)dx,
R
−l ≤ ξ ≤ 0
2 2 0
1 0
u (ξ) − 12 u1 (ξ), 0≤ξ≤l
2 o
f 0 (ξ) =
1 u0 (−ξ) + 1 u1 (−ξ),
−l ≤ ξ ≤ 0
2 o 2
28
1 1
Rη
u (η) + u1 (x)dx, 0≤η≤l
2 o
2 0
g(η) =
− 1 uo (2l − η) + 1 2l−ξ u1 (x)dx,
R
l ≤ ξ ≤ 2l
2 2 0
1 0
u (η) + 12 u1 (η), 0≤η≤l
2 o
g 0 (η) =
1 u0 (2l − η) − 1 u1 (2l − η),
l ≤ η ≤ 2l
2 o 2
If f and g have sufficient smoothness, they will verify (16) [see page 16]:
Recall that v(ξ, η) = f (ξ)+g(η). This holds in particular for ξ ≤ l, η ≥ 0.
Along the line segment ξ = 0, 0 ≤ η ≤ l we have by smoothness of v,
29
∂v(0+ , η) ∂v(0− , η)
= ⇔ f 0 (0+ ) = f 0 (0− ) ⇔ u1 (0+ ) = −u1 (0+ ) or u1 (0+ ) = 0
∂ξ ∂ξ
∂ 2 v(0+ , η) ∂ 2 v(0− , η)
= ⇔ f 00 (0+ ) = f 00 (0− ) ⇔ u00o (0+ ) = −u00o (0+ ) or u00o (0+ ) = 0
∂ξ 2 ∂ξ 2
30
Existence Theorem If uo , u1 satisfy (10), (11), (12) then ∃ a solution
of (1), (2), (3a), (3b). The proof is the construction of f and g above.
un (x, τ ) → 0 uniformly on Ω
Proof. This is evident from (4) and (5) and their extensions.
31
Chapter 3. Steady Temperature in a Circular Cylinder.
∂u ∂ 2u ∂ 2u ∂ 2u
= K( 2 + 2 + 2 )
∂t ∂x ∂y ∂z
Assume that
32
Polar Coordinates. Put T = us (x, y) = v(r, θ) where
x = r cos θ, y = r sin θ
Then
p y
r= x2 + y 2 , θ = tan−1
x
Differentiation gives
∂r x ∂r y
= = cos θ, = = sin θ
∂x r ∂y r
∂θ −y − sin θ ∂θ x cos θ
= 2 = , = 2 =
∂x r r ∂y r r
∂us ∂v ∂v ∂r ∂v ∂θ ∂v sin θ ∂v
= = + = cos θ − ≡ F1 (r, θ)
∂x ∂x ∂r ∂x ∂θ ∂x ∂r r ∂θ
∂us ∂v ∂v ∂r ∂v ∂θ ∂v cos θ ∂v
= = + = sin θ + ≡ F2 (r, θ)
∂y ∂y ∂r ∂y ∂θ ∂y ∂r r ∂θ
33
∂ 2 us ∂ 2v ∂F1 ∂r ∂F1 ∂θ 2
2 ∂ v sin2 θ ∂ 2 v sin2 θ ∂v
= = + = cos θ + +
∂x2 ∂x2 ∂r ∂x ∂θ ∂x ∂r2 r2 ∂θ2 r ∂r
∂ 2 us ∂ 2v ∂F2 ∂r ∂F2 ∂θ 2
2 ∂ v cos2 θ ∂ 2 v cos2 θ ∂v
= = + = sin θ + +
∂y 2 ∂y 2 ∂r ∂y ∂θ ∂y ∂r2 r2 ∂θ2 r ∂r
Thus
∂ 2 us ∂ 2 us ∂ 2 v 1 ∂v 1 ∂ 2v
+ = 2+ +
∂x2 ∂y 2 ∂r r ∂r r2 ∂θ2
BV Problem for v(r, θ). If we write f (θ) for g(a cos θ, a sin θ) then
∂ 2 v 1 ∂v 1 ∂ 2v
(1) + + = 0 for 0 < r < a, and all θ
∂r2 r ∂r r2 ∂θ2
34
The condition (3) is an obvious geometric fact. Rotation around the ver-
tical axis of the cylinder by 2π should not effect the temperature. Condition
(4) is certainly true for the steady-state temperature along the vertical axis
of the cylinder. But it is a necessary condition for the solution of the prob-
lem, since (1) has a singularity at r = 0. (4) excludes non-physical solutions.
v(r, θ) = R(r)Θ(θ)
1 1
R00 Θ + R0 Θ + 2 RΘ00
r r
R00 R0 Θ00
r2 +r =− =µ
R R Θ
35
The solutions of this eigenvalue problem are µ = µn = n2 , n = 0, 1, 2, 3, ...,
whence
α2 − n2 = 0, or α = ±n, n = 1, 2, 3, ...
For n = 0, rR00 + R0 = 0 =⇒ R0 = Do
r
, Ro = Co + Do ln r
(4) requires that Do = 0. Thus the separated solutions are
∞
X
(5) v(r, θ) = Ao + rn (An cos nθ + Bn sin nθ)
n=1
Formal Solution
36
∞
X
(6) v(a, θ) = f (θ) = Ao + (an An cos nθ + an Bn sin nθ)
n=1
∞
1 X
f (θ) = ao + (an cos nθ + bn sin nθ)
2 n=1
1
Rπ
an = f (φ) cos(nφ)dφ, n = 0, 1, 2, ...
π −π
(7)
1
Rπ
bn =
f (φ) sin(nφ)dφ, n = 1, 2, ...
π −π
1
Ao = ao , an An = an , an Bn = bn , n = 1, 2, ...
2
∞
1 X r
(8) v(r, θ) = ao + ( )n (an cos nθ + bn sin nθ)
2 n=1
a
The existence of solutions of the Dirichlet problem for the disk can be
discussed by applying convergence theorems for Fourier series to the formal
solution (8). However, we shall take a different approach to show that the
series in (8) can be summed.
37
The Poisson Integral. Still proceeding formally we have, from (7) and
(8),
π ∞
r n1 π
Z Z
1 X
v(r, θ) = f (φ)dφ+ ( ) f (φ){cos nφ cos nθ+sin nφ sin nθ}dφ
2π −π n=1
a π −π
π ∞
r n1 π
Z Z
1 X
= f (φ)dφ + ( ) f (φ) cos n(θ − φ)dφ
2π −π n=1
a π −π
Z π ∞
1 X r
= f (φ){1 + 2 ( )n cos n(θ − φ)}dφ
2π −π n=1
a
Remark. The last step is easy to justify if r < a and f ∈ L1 (−π, π).
Now18
r r reiψ n
( )n cos nψ = Re{( )n einψ } = Re{( ) }
a a a
and
∞ ∞
X reiψ n X n z
( ) = z = z + z 2 + z 3 + ... =
n=1
a n=1
1−z
r iψ
a
e r iψ
a
e (1 − ar e−iψ )
= =
1 − ar eiψ |1 − ar eiψ |2
18
Re stands for “real part.” Thus for any complex number z = c + ib, Re(z) = c.
38
r iψ
a
e − ( ar )2
=
(1 − ar cosψ)2 + ar sin2 ψ
Thus
∞ r
X r n a
cos ψ − ( ar )2
( ) cos nψ =
n=1
a 1 − 1 ar cos ψ + ( ar )2
Hence
∞
X r 1 − 2 ar cos ψ + ( ar )2 + 2 ar cos ψ − 2( ar )2
1+2 ( )n cos nψ =
n=1
a 1 − 2 ar cos ψ + ( ar )2
Simplifying,
∞
X r a2 − r 2
1+2 ( )n cos nψ = 2
n=1
a a + r2 − 2ra cos ψ
π
a2 − r 2
Z
1
(9) v(r, θ) = f (φ) 2 dφ, for 0 ≤ r < a
2π −π a + r2 − 2ra cos(θ − φ)
π
a2 − x 2 − y 2
Z
1
(10) us (x, y) = g(a cos φ, a sin φ)dφ
2π −π a2 + x2 + y 2 − 2a(x cos φ + y sin φ)
39
existence of solutions.
Classical Solutions. us (x, y) is a classical solution of the Dirichlet
problem for the Laplace equation in the disk Ω = {(x, y) : x2 + y 2 < a2 } ⇔
us ∈ C 2 (Ω) ∩ C(Ω) and satisfies
∂ 2 us ∂ 2 us
(11) + = 0 in Ω
∂x2 ∂y 2
a2 −x2 −y 2
1
Rπ
g(a cos φ, a sin φ)dφ, r<a
2π −π a2 +x2 +y 2 −2a(x cos φ+y sin φ)
us (x, y) =
g(x, y),
r=a
40
attack the easier parts first.
∂ 2 us ∂ 2 us
Step 1. Show us ∈ C 2 (Ω) and ∂x2
+ ∂y 2
= 0 in Ω.
To show this, let 0 < δ < a and define
p
Ωδ = {(x, y) : r = x2 + y 2 < a − δ}
Then if ψ = θ − φ
Hence
∞
a2 − r 2 X r
(13) 2 2
=1+2 ( )n cos n(θ − φ), r < a
a + r − 2ra cos(θ − φ) n=1
a
defines a function which has partial derivatives of all orders and satisfies
Laplace’s equation on Ωδ for each δ such that 0 < δ < a. Hence (10) defines
a function us ∈ C 2 (Ω) (indeed C ∞ (Ω)) such that (11) holds.
a2 − r 2
• → 0 when r → a if φ 6= θ.
a2 + r2 − 2ra cos(θ − φ)
41
a2 − r 2 a2 − r2 a+r
• = = → +∞ if φ = θ.
a2 + r2 − 2ra cos(θ − φ) (a − r)2 a−r
Z π
1 a2 − r 2
• dφ ≡ 1, ∀r < a, −π ≤ θ ≤ π (by 13)
2π −π a2 + r2 − 2ra cos(θ − φ)
π
a2 − r 2
Z
1
v(r, θ) − f (θo ) = {f (φ) − f (θo )}dφ19
2π −π a2 + r2 − 2ra cos(θ − φ)
θo +δ
a2 − r 2
Z
1
= {f (φ) − f (θo )}dφ
2π θo −δ a2 + r2 − 2ra cos(θ − φ)
a2 − r 2
Z
1
+ {f (φ) − f (θo )}dφ
2π |θo −φ|≥δ a2 + r2 − 2ra cos(θ − φ)
= I1 (r, θ) + I2 (r, θ)
|f (φ) − f (θo )| < ∀ φ with |φ − θo | ≤ δ1 ()
2
19
Recall a2 + r2 − 2ra cos(θ − φ) > 0.
20
Observe by definition (see (6)) that g(x, y) ∈ C(∂Ω) ⇔ f (θ) ∈ C(R) and f (θ + 2π) =
f (θ), ∀ θ ∈ R.
42
Then
θo +δ
a2 − r 2
Z
1
(15) |I1 (r, θ)| ≤ |f (φ) − f (θo )|dφ
2π θo −δ a2 + r2 − 2ra cos(θ − φ)
≤ , ∀θ ∈ R, r < a.
2
θ−φ
a2 + r2 − 2ra cos(θ − φ) ≥ 2ra(1 − cos(θ − φ)) = 4ra sin2
2
Take
δ1 ()
|θ − θo | < , |θo − φ| ≥ δ1 ()
2
Then
δ1 ()
|θ − φ| ≥ |θo − φ| − |θ − θo | >
2
and hence
δ1 ()
a2 + r2 − 2ra cos(θ − φ) ≥ 2ra(1 − cos(θ − φ) > 4ra sin2 >0
4
43
Thus if
M = max |f (θ)|
0≤θ≤2π
a2 − r 2
Z
1
|I2 (r, θ)| ≤ {|f (φ)| + |f (θo )|}dφ
2π |θo −φ|≥δ1 () a2 + r2 − 2ra cos(θ − φ)
(16)
a2 − r2 a2 − r 2
Z
2M
≤ δ1 ()
dφ ≤ 2M δ1 ()
<
2π |θo −φ|≥δ1 () 4ra sin2 4ra sin2 2
4 4
provided r > a − δ2 () for δ2 () sufficiently small, and |θ − θo | < δ1 ()/2.
δ1 ()
∀ (x, y) = (r cos θ, r sin θ) with a − δ2 () < r < a, |θ − θo | < .
2
44
The Mean Value Theorem for Laplace’s Equation. Let
∂ 2 us ∂ 2 us
(2) u ∈ C 2 (Ω) and + = 0 in Ω.
∂x2 ∂y 2
Then
Z π
1
(3) u(xo , yo ) = u(xo + r cos θ, yo + r sin θ)dθ for 0 ≤ r ≤ R.
2π −π
In other words, at any point in the domain, u is equal to its mean value
along any circle surrounding it in Ω.
Proof. Define
Then
45
∂ 2 v 1 ∂v 1 ∂ 2v
(5) + + = 0 for 0 < r < R, θ ∈ R.
∂r2 r ∂r r2 ∂θ2
Now let
Z π Z π
1 1
(7) A(r) = v(r, θ)dθ = u(xo + r cos θ, yo + r sin θ)dθ
2π −π 2π −π
Then
π
∂2
Z
1 1 1 ∂
00
A (r) + A0 (r) = ( 2
+ )v(r, θ)dθ
r 2π −π ∂r r ∂r
π
∂ 2v
Z
1
=− dθ = 0 by (5), (6)
2πr2 −π ∂θ2
46
constant in Ω then it can have no local maximum or minimum in Ω.
Corollary. If Ω is compact,
u ∈ C(Ω) ∩ C 2 (Ω)
and
∂ 2u ∂ 2u
+ = 0 in Ω
∂x2 ∂y 2
u1 (x, y) ≡ u2 (x, y)
QED.
47
Chapter 4. Basic Concepts in the Theory of Heat Conduc-
tion.
Temperature. The notion of the temperature of a body (at a point) is an
intuitive concept. A more precise definition can be based on thermodynam-
ics. In any case it is measurable by thermometers, thermocouples and many
other devices, The temperature relative to a fixed scale is measured by a real
number. Scales include
TK ≥ 0
Quantity of Heat. Heat is a form of energy. The basic unit of heat energy
is the calory (also spelled calorie), defined by the property that 1 calory =
Quantity of heat needed to heat 1 gram of water from 14.5◦ C to 15.5◦ C.
Specific Heat of a Solid. The lower case c will be used to represent this
48
quantity. For each substance, the amount of heat required to raise the tem-
perature of 1 gram by 1◦ C is called the specific heat of the substance. Actu-
ally, the specific heat varies slightly with the temperature, but we shall treat
it as a constant. Some approximate values for familiar substances are
Substance Specific Heat
Water 1.0000 (at 15◦ C)
Glass .20
Cork .48
Copper .0914
Silver .0556
49
In solid bodies heat transfer in the interior is assumed to take place by
pure conduction. It may be necessary to consider convection and/or radia-
tion at the surface of a solid.
Net Flux of heat through the surface of V during the same time interval.
II. Fourier’s Law of Heat Conduction. This states that the rate of flow of
heat at any point in a solid is a function of the temperature gradient at that
point.
This will also be quantified in several different cases below. To begin, the
case of 1-dimensional heat flow in a plate (slab) is discussed.
50
Heat Flow in a Plate. Imagine a large uniform plate (or wall) with
Area = A, Thickness = l.
Assume that the two faces are kept at fixed temperature To and T1 6= To ,
T (x) being the temperature (independent of time) on a plane parallel to the
faces of the plate at depth x21 and let
(T1 − To )
Q∝ l
A
(T1 − To )At
(1) Q = −K
l
51
plate. The value of K is characteristic of the material of which the plate is
made (it may vary somewhat with the thermal state of the material, but we
shall assume it is constant).
Remarks
a. Heat flows from high temperature regions to low temperature regions.
Hence, with the definition of Q given above, K > 0.
b. The insulating value of a layer of insulation is proportional to its thick-
ness.
c. Our confidence in Fourier’s law is based on both direct experiment and
the accuracy of many predictions based on the law.
Q (T1 − To ) cal.
q= = −K
At l m2 sec.
is called the heat flux through the plate. Some representative values of the
thermal conductivity are
Substance Thermal Conductivity
Water .00144
Glass .0028
Cork .0001
Copper .93
Silver 1.00
52
Steady Temperature Profile in a Plate. Introduce a coordinate x normal
to the surface of the plate, the left side corresponding to x = 0 and xo corre-
sponding to some interior point. Between any two interior points, xo , xo +∆x
the net increase in heat content is determined as specific heat times volume
times net temperature increase over time is approximately (assuming ∆x is
small) = cρA∆x[T (xo , t + ∆t) − T (xo , t)] where ρ is mass density. Mean-
while, the heat energy entering [xo , xo + ∆x] over the time interval [t, t + ∆t]
is (q(xo , t) − q(xo + ∆x, t))A∆t. Applying the principle of conservation of
heat energy tells us that these two quantities should be equal and using
the fact that T (and therefore q) is not a function of time, we have that
q(xo ) = const. = q.
Thus
T (xo ) − To
q = −K
xo
q T1 − To
T (x) = To − x = To + x
K l
53
To = To (t), T1 = T1 (t)
Fourier’s law for steady temperatures suggests that the instantaneous flux
of heat through this slab, at time t, is approximately
∂T (xo , t)
q(xo , t) = −K
∂x
As in the case of steady heat flow, our confidence in Fourier’s law is based
on the accuracy of many predictions based on it.
54
conservation of energy principle to a portion (x, x + ∆x) of the plate and a
time interval (t, t + ∆t):
The net increase of heat content of (x, x + ∆x) during (t, t + ∆t)
mass
z }| {
= c ρ A∆x
| {z } [T (x, t + ∆t) − T (x, t)]
| {z }
volume net temperature increase
Z t+∆t Z x+∆x
∂q
−A dt = cρA∆x[T (x, t + ∆t) − T (x, t)]
t x ∂x
or
t+∆t x+∆x
T (x, t + ∆t) − T (x, t)
Z Z
1 1 ∂q
− dt = cρ
∆t t ∆x x ∂x ∆t
∂q(x, t) ∂T (x, t)
− = cρ
∂x ∂t
55
∂ 2T ∂T
K 2
= cρ
∂x ∂t
or
∂T ∂ 2T
=k 2
∂t ∂x
where
K
k= cρ
= “thermal diffusivity” or the “diffusion coefficient.”
k measures the speed with which heat is conducted through a substance. The
following table gives approximate values of k for some familiar substances
56
we must know T at some initial time (say t = 0). This gives
Initial Condition.
∂T (0, t)
q(0, t) = −K = qo (t) (a given function) t ≥ 0
∂x
or
57
∂T (0, t) qo (t)
= go (t) (= ) for t ≥ 0
∂x −K
Similarly
∂T (l, t)
= g1 (t) for t ≥ 0
∂x
may be given.
∂T (0, t)
−q(0, t) = K = H(T (0, t) − Te )
∂x
where h = H/K. Note that since heat flows from hot to cold, H ≥ 0, h ≥ 0.
58
scribed by a function
T = T (x, t)
If the rod is cooling through its surface into an environment with tem-
perature Te (x) then (cf. p. 59)
Z t+∆t
=− [A(x + ∆x)q(x + ∆x, t) − A(x)q(x, t)]dt
t
59
Z x+∆x Z t+∆t
− H(x)[T (x, t) − Te (x)]P (x)dtdx
x t
and
Net increase in heat content of (x, x+∆x) during (t, t+∆t) (assuming ∆x is small)
t+∆t
A(x + ∆x)q(x + ∆x, t) − A(x)q(x, t)]
Z
1
− [ dt
∆t t ∆x
Z x+∆x Z t+∆t
1 1
− H(x)[T (x, t) − Te (x)]P (x)dtdx
∆x x ∆t t
∂ ∂T (x, t)
− (A(x)q(x, t)) − H(x)P (x)[T (x, t) − Te (x)] = c(x)ρ(x)
∂x ∂t
60
Combining this and Fourier’s law for a non-uniform rod:
∂T (x, t)
q(x, t) = −K(x)
∂x
∂T ∂ ∂T
c(x)ρ(x)A(x) = (A(x)K(x) ) − H(x)P (x)[T (x, t) − Te (x)]
∂t ∂x ∂x
∂ 2u ∂u
Lu = po (x) 2
+ p1 (x) + p2 (x)u
∂x ∂x
with
K(x)
po (x) = >0
c(x)ρ(x)
1 d
p1 (x) = (A(x)K(x))
ρ(x)c(x)A(x) dx
H(x)P (x)
p2 (x) = − <0
ρ(x)c(x)A(x)
61
It is interesting that the most general linear second order operator23 L
can arise in this way; i.e., by suitable choice of A(x), P (x), etc. The only
restrictions are that po (x) > 0, p2 (x) < 0.
The initial and boundary conditions given on pp. 58-59 are appropriate
for the non-uniform rod. In addition we will consider the
∂T (0, t) ∂T (l, t)
T (0, t) = T (l, 0) and K(0) = K(l) , t≥0
∂x ∂x
23
See appendix I.
62
Diffusion of Heat in 3 Space Dimensions. Consider a heat con-
ducting solid body occupying a domain Ω ⊂ R3 . The thermal state of the
body is characterized by a temperature field
(x is a vector quantity).
Z
Q(t) = c(x)ρ(x)T (x, t)dx = Total quantity of heat in V at time t
V
where dx = dx1 dx2 dx3 .24 In the important case where the body is homo-
R
geneous c and ρ are constants and Q(t) = cρ V T (x, t)dx.
~q(x, t) = (q1 (x, t), q2 (x, t), q3 (x, t)) = Heat Flux Field
24
The single integral sign is customary in modern mathematics. In elementary calculus
one often sees multiple integral signs corresponding to the space dimension.
63
To interpret ~q let dS be a (small) surface element in Ω with unit normal
vector ~ν .
Then
Z
~q(x, t) • ~ν (x)dS = Quantity of heat leaving V per unit time at time t.
∂V
Z
dQ
~q(x, t) • ~ν (x)dS = −
∂V dt
64
Z Z
~ =
∇ • Adx ~
~ν • AdS
V ∂V
or
Z Z
∂A1 ∂A2 ∂A3
( + + )dx1 dx2 dx3 = (ν1 A1 + ν2 A2 + ν3 A3 )dS
V ∂x1 ∂x2 ∂x3 ∂V
Z Z
∇ • ~qdx = ~ν • ~qdS
V ∂V
Z Z
∂T (x, t)
∇ • ~qdx = − c(x)ρ(x) dx
V V ∂t
or
Z
∂T (x, t)
∇ • ~qdx + c(x)ρ(x) dx = 0
V ∂t
If T (x, t), ~q(x, t) are C 1 functions (in Ω × R) and c(x), ρ(x) ∈ C 1 (Ω) then
the integrand in the last integral is continuous in Ω (for any fixed t). Since
the identity holds for all volumes V ⊂ Ω it follows that
∂T (x, t)
(1) ∇ • ~q + c(x)ρ(x) = 0, x ∈ Ω, t ∈ R
∂t
65
Indeed, if
∂T (xo , t)
∇ • ~q(xo , t) + c(xo )ρ(xo ) >0
∂t
∂T (x, t) 1
(3) − ∇ • (K(x)∇T (x, t)) = 0
∂t c(x)ρ(x)
66
tion away from the temperature gradient. This can be expressed mathemat-
ically by allowing K to be a matrix quantity. In the isotropic, homogeneous
case (3) becomes
where
Initial Condition. To determine T (x, t) for a given body one must con-
struct a solution of (3) with a given initial temperature distribution
67
(8) ~q(x, t) • ~ν (x) = −K(x)∇T (x, t) • ~ν (x) = φ(x, t)
for x ∈ ∂Ω, t ≥ 0
where
H = H(x) = “outer conductivity” of ∂Ω at x (H ≥ 0)
and
To (x) = exterior temperature at x ∈ ∂Ω. This can be written (if ∂T /∂~ν =
∇T • ~ν , ~ν out of Ω)
∂T
(10) + hT = hTo , x ∈ ∂Ω, t ≥ 0
∂~ν
25
After Victor Gustave Robin, 19th century physicist.
68
Mixed Boundary Conditions are also possible where ∂Ω = S1 ∪S2 ∪. . . Sk
and one of the above BCs holds on each Sj .
∂u ∂ 2u ∂ 2u ∂ 2u
(11) − k( 2 + 2 + 2 ) = 0, x ∈ Ω, t > 0
∂t ∂x1 ∂x2 ∂x3
where f (x) and φ(x, t) are prescribed functions on Ω and ∂Ω × [0, ∞], re-
spectively.
69
BV Problem 2 (Neumann Condition on ∂Ω).26 Find a function u(x, t), x ∈
Ω, t ≥ 0 such that
∂u ∂ 2u ∂ 2u ∂ 2u
(14) − k( 2 + 2 + 2 ) = 0, x ∈ Ω, t > 0
∂t ∂x1 ∂x2 ∂x3
∂u(x, t)
(16) ≡ ∇u(x, t) • ν(x) = φ(x, t), x ∈ ∂Ω, t ≥ 0
∂ν
∂u ∂ 2u ∂ 2u ∂ 2u
(17) − k( 2 + 2 + 2 ) = 0, x ∈ Ω, t > 0
∂t ∂x1 ∂x2 ∂x3
26
Named for German mathematician Carl Gottfried Neumann, cofounder of the math-
ematical research journal Mathematische Annalen.
70
(18) u(x, 0) = f (x), x ∈ Ω
∂u(x, t)
(19) + hu = φ(x, t), x ∈ ∂Ω, t ≥ 0
∂ν
where f (x) and φ(x, t) are prescribed functions and h > 0 is a prescribed
constant.
71
The Maximum Principle. The maximum theorem for the heat equation
in one space dimension formulated and proved on pages 10-11, can be gen-
eralized to higher dimensions. It may be formulated as follows.
∂u ∂ 2u ∂ 2u ∂ 2u
− k( 2 + 2 + 2 ) = 0, x ∈ Ω, t > 0
∂t ∂x1 ∂x2 ∂x3
Let
ΓT = (Ω × {0}) ∪ {(x, t) : x ∈ ∂Ω, 0 ≤ t ≤ T }
Then ∀ T > 0
The proof is essentially the same as the one on pages 6-7. In fact, the
proof works for any number n ≥ 1 of space variables.
72
Uniqueness Theorems for BV Problems 2 and 3 in Bounded Domains. In
the case of BV Problems 2 and 3 the uniqueness of classical solutions does
not follow immediately from the maximum principle as it does for BV Prob-
lem 1. Another method of proving uniqueness will now be given that works
for these two problems. The following notation will be used in the proof.
∂u ∂u ∂u
∇u = ( , , )
∂x1 ∂x2 ∂x3
∂ 2u ∂ 2u ∂ 2u
4u = ∇ • ∇u = + +
∂x21 ∂x22 ∂x23
(20) ∇ • (u∇v) = ∇u • ∇v + u 4 v
To prove that at most one classical solution exists we suppose that u1 (x, t)
and u2 (x, t) are any two classical solutions with the same “data” f (x), φ(x, t)
and consider the difference
73
u(x, t) = u1 (x, t) − u2 (x, t)
Z
J(t) = u(x, t)2 dx, t≥0
Ω
and
Z
0 ∂u(x, t)
J (t) = 2 u(x, t) dx, t ≥ 0
Ω ∂t
Z
0
J (t) = 2 u(x, t) 4 u(x, t)dx
Ω
Z Z
0
J (t) = 2 ∇ • (u∇u)dx − 2 |∇u(x, t)|2 dx
Ω Ω
74
Z Z
0
J (t) = 2 u∇u • ν(x)dS − 2 |∇u|2 dx
∂Ω Ω
Z Z
∂u
=2 u dS − 2 |∇u|2 dx
∂Ω ∂ν Ω
Z Z
0 2
J (t) = −2h u(x, t) dS − 2 |∇u(x, t)|2 dx ≤ 0, ∀t > 0
∂Ω Ω
Z t Z t
0
(21) J(t) = J(0) + J (τ )dτ = J 0 (τ )dτ ≤ 0, ∀ t ≥ 0
0 0
i.e.,
u1 (x, t) = u2 (x, t)
QED.
Remark 1. The same method also works for BV problem 1 if the diver-
gence theorem holds for the domain Ω.
27
by the definition of J
75
Remark 2. The proof works for any number n ≥ 1 of space dimensions.28
28
The divergence theorem for n = 1 dimensions is just the integration by parts formula.
76
Chapter 5.
∂u ∂ 2u
(1) = for − ∞ < x < ∞, t > 0
∂t ∂x2
K
(3) k= =1
cρ
which also can be achieved by the proper choice of time unit. It will also be
assumed that
77
(4) cρ = 1 (choice of unit of heat)
and
(5) A=1 (choice of unit of length)
Let u∆x (x, xo , t) denote the subsequent temperature distribution, i.e., the
78
solution of (1), (2) with uo (x) defined by (6) (it is assumed for the moment
to exist).
The total quantity of heat in the infinite rod for any time t, is
Z ∞ Z ∞
(7) Q∆x = cρAu∆x (x, xo , t)dx = u∆x (x, xo , t)dx
−∞ −∞
∞ ∞
∂ 2 u∆x (x, xo , t)
Z Z
∂u∆x (x, xo , t)
Q0∆x (t) = dx = dx = 0
−∞ ∂t −∞ ∂x2
provided ∂u∆x (x, xo , t)/∂x → 0 when x → ±∞. This is clearly the case.
Thus
79
exists. This will be verified below. The limiting function φ(x, xo , t) should
describe the temperature distribution at time t > 0 due to a unit amount of
heat released at the point xo at time t = 0.
Intuitively, this corresponds to the Cauchy problem (1), (2) with the
initial distribution
but use of distribution theory will be avoided here. The limit function (9)
may be expected to have the following properties
∂φ ∂ 2φ
(10) = ∀ x, xo ∈ R and t > 0
∂t ∂x2
Z ∞
(12) φ(x, xo , t)dx = 1 ∀ xo , t > 0
−∞
80
by phrase, “uniqueness implies existence.”
Step 1. Let φ(x, xo , t) be the function having properties (10)-(13) and
consider the function φ0 (x, xo , t) defined by
∂φ(x, t) ∂ 2 φ(x, t)
(18) = ∀ x ∈ R and t > 0
∂t ∂x2
81
(19) lim φ(x, t) = 0 ∀ x 6= 0
t→0+
Z ∞
(20) φ(x, t)dx = 1 ∀ t>0
−∞
For any fixed α > 0 it satisfies (18), (19) and (21). Moreover,
Z ∞ Z ∞ Z ∞
0 n 2 n−1
φ (x, t)dx = α φ(αx, α t)dx = α φ(y, α2 y)dy = 1
−∞ −∞ −∞
1
Taking α = t− 2 gives
1 1
(24) φ(x, t) = t− 2 φ(t− 2 x)
82
where φ(x) is the function of x ∈ R defined by
∂φ(x, t) 1 3 x 1 1 3 x 1 3
= − t− 2 φ( 1 ) − t− 2 t− 2 xφ0 ( 1 ) = − t 2 (φ(ξ) + ξφ0 (ξ))
∂t 2 t2 2 t2 2
where
x
ξ= 1
t2
Similarly,
∂φ(x, t) 1 x ∂ 2φ 1
= φ0 ( 1 ), 2
= 3 φ00 (ξ)
∂x t t2 ∂x t2
Whence,
∂ 2 φ(x, t) ∂φ(x, t) 1 ξ 1
2
− = 3 {φ00 (ξ) + φ0 (ξ) + φ(ξ)} = 0
∂x ∂t t2 2 2
Also,
Z ∞ Z ∞ Z ∞
1
− 12
φ(x, t)dx = t 2 φ(t x)dx = φ(ξ)dξ = 1
−∞ −∞ −∞
and
83
1 x x 1
lim+ φ(x, t) = lim+ 1 φ( 1 ) = lim ξφ(ξ) = 0
t→0 t→0 x t2 t2 x ξ→±∞
d2 φ ξ dφ 1
(26) + + φ=0 ∀ξ∈R
dξ 2 2 dξ 2
d dφ ξ d2 φ ξ dφ 1
( + φ) ≡ 2 + + φ
dξ dξ 2 dξ 2 dξ 2
dφ ξ
(30) + φ = K = const. ξ ∈ R
dξ 2
84
−ξ2
φ = ce 4
Z ξ
−ξ 2 /4 0 −ξ 2 /4 2 /4
(31) ce +ce eτ dτ
0
Rξ 2
ξ
ξ eτ /4 dτ
Z
−ξ 2 /4 τ 2 /4 0
lim ξe e dτ = lim
ξ→±∞ 0 ξ→±∞ eξ2 /4
Rξ 2 /4 2 /4
0
eτ dτ + ξeξ
= lim 1
ξ→±∞
2
ξeξ2 /4
2 /4
eξ
= 2 + lim ξ 2 ξ 2 /4
=2
ξ→±∞ 1 ξeξ 2 /4 + e
2 4
2 /4
(32) φ(ξ) = ce−ξ
85
Z ∞ Z ∞
2 /4
φ(ξ)dξ = c eξ dξ = c(4π)1/2 = 1
−∞ −∞
2 /4
(33) φ(ξ) = (4π)−1/2 eξ
Note that (29) is also satisfied. Finally, combining (33), (24) and (16)
gives
2 /4t
(34) φ(x, xo , t) = (4πt)−1/2 e−(x−xo )
The corollary follows from the derivation and can also be verified directly.
86
Formal Solution of the Cauchy Problem. Returning to the problem (1),
(2) (p. 78), let us assume that uo ∈ C(R) and approximate it by a step
function:
∞
X
uo (x) ≈ uo (x0i )χi (x)
i=−∞
1 xi ≤ x < xi+1
χi (x) =
0 elsewhere
is (see p. 79)
∞
X
u(x, t) ≈ φ(x, xi , t)uo (x0i )∆x
−∞
87
Z ∞
u(x, t) = φ(x, ξ, t)uo (ξ)dξ
−∞
Z ∞
1 2 /4t
(35) u(x, t) = e−(x−ξ) uo (ξ)dξ
(4πt)1/2 −∞
Z xo +δ
1 2 /4t
u(xo , t) ≈ e−(xo −ξ) uo (ξ)dξ
(4πt)1/2 xo −δ
Z xo +δ
uo (xo ) 2 /4t
≈ e−(xo −ξ) dξ)
(4πt)1/2 xo −δ
≈ uo (xo )
88
Notation. Let Ω ⊂ R2 be a domain (a connected open set). Then
∂u ∂ 2u ∂u ∂ 2u
(36) H(Ω) = {u(x, t) : ∈ C(Ω), 2 ∈ C(Ω), = in Ω)}
∂t ∂x ∂t ∂x2
2
(37) uo (x)e−ax ∈ L1 (R)
Then31 the integral in (35) converges for all (x, t) in the domain Ω1/4a
where
(38) ΩT = R × (0, T )
∞ ∞
∂ 2u
Z Z
∂u 1 2 1
2
= = 2 (x − ξ) φ(x, ξ, t)uo (ξ)dξ − φ(x, ξ, t)uo (ξ)dξ
∂x ∂t 4t −∞ 2t −∞
The theorem will be established if it is shown that the last two integrals
converge uniformly in some neighborhood of each point (xo , to ) ∈ Ω1/4a .
31
(37) is a growth limiting condition on uo . It may grow fast but not too fast!
89
Choose δ > 0 such that
= Cφ(xo + δ, ξ, to + δ)
where r
to + δ
C=
to − δ
2
φ(xo + δ, ξ, to + δ)eaξ ≤ K1 ∀ ξ ∈ [R, ∞)
90
2
(ξ − xo + δ)2 φ(xo + δ, ξ, to + δ)eaξ ≤ K2 ∀ ξ ∈ [R, ∞)
Then
2
|φ(x, ξ, t)uo (ξ)| ∈ CK1 |uo (ξ)e−aξ | ∈ L1 ([R, ∞))
for all (x, t) ∈ N (xo , to ). The remaining integrals can be treated in the same
way. QED.
Theorem 2. Assume that uo satisfies (37) and that the limits uo (xo +)
and uo (xo −) exist. Then the function u ∈ H(Ω1/4a ) defined by (35) satisfies32
(39) lim sup |u(x, t)| ≤ max(|uo (xo +)|, |uo (xo −)|)
x→xo ,t→0+
Proof of Theorem 2. Let M = max(|uo (xo +)|, |uo (xo −)|). Then ∀ >
0 ∃ δ > 0 such that
91
Thus if
Z xo −δ Z xo +δ Z ∞
u(x, t) = φ(x, ξ, t)uo (ξ)dξ+ φ(x, ξ, t)uo (ξ)dξ+ φ(x, ξ, t)uo (ξ)dξ
−∞ xo −δ xo +δ
= I1 + I2 + I3
then
Z ∞
|I2 | ≤ (M + ) φ(x, ξ, t)uo (ξ)dξ = M + , ∀ x ∈ R, t > 0.
−∞
Moreover, if |x−xo | < ρ < δ, ξ ≥ xo +δ then −xo −ρ < x < xo +ρ, ξ−x ≥
ξ − xo − ρ and hence
1
φ(x, ξ, t) ≤ e−(ξ−xo −ρ)/4t = φ(xo + ρ, ξ, t)
(4πt)1/2
Z ∞
|I3 | ≤ φ(xo + ρ, ξ, t)|uo (ξ)|dξ
xo +δ
Now
2
φ(xo + ρ, ξ, t)eaξ
92
2 2
φ(xo + ρ, ξ, t)eaξ ≤ φ(xo + ρ, xo + δ, t)ea(xo +δ) , ξ ≥ xo + δ
or
2 2
φ(xo + ρ, ξ, t) ≤ [φ(ρ, δ, t)ea(xo +δ) ]e−aξ , ξ ≥ xo + δ
and
Z ∞
a(xo +δ)2 2
|I3 | ≤ [φ(ρ, δ, t)e ] e−aξ |uo (ξ)dξ
xo +δ
Z ∞
u(x, t) − uo (xo +) = φ(x, ξ, t)[uo (ξ) − uo (xo +)]dξ = I10 + I20 + I30
−∞
93
|I20 (x, t)| ≤ sup |uo (ξ) − uo (xo +)| +
|x−xo |≤δ
lim sup |u(x, t) − uo (xo +)| ≤ sup |uo (ξ) − uo (xo +)|, ∀ δ > 0
x→xo ,t→0+ |ξ−xo |≤δ
Ω̇T = R × [0, T )
will be used.
Definition. A classical solution of the Cauchy problem (1), (2) (p. 78) in
a domain ΩT is a function
u ∈ H(ΩT ) ∩ C(Ω̇T )
94
Existence Theorem. Assume that
(a) uo ∈ C(R)
2
(b) ∃ a > 0 such that uo (x)e−ax ∈ L1 (R)
Then the integral (35) defines a classical solution of the Cauchy problem
in the domain Ω1/4a .
Proof. Theorems 1 and 2 imply that u ∈ H(Ω1/4a ) and that
for all xo . This property and the continuity of uo imply that u ∈ C(Ω̇1/4a )
and u(x, 0) = uo (x), ∀ x ∈ R. QED.
Corollary. Assume that
(a) uo ∈ C(R)
.
Proof. The boundedness condition (b) implies that condition (b) of the
95
Existence Theorem holds for all a > 0. (35) implies the second part of the
corollary. QED.
————
The heat equation is strongly linked to the notion of absolute zero. In
fact it is possible to show that:
Widder’s Representation Theorem. A real-valued function u(x, t) has the
properties
(1) u ∈ H(ΩT )
Z ∞
(3) u(x, t) = φ(x, ξ, t)dα(ξ) ∀ (x, t) ∈ ΩT
−∞
The class of solutions defined by (3) is more general than that defined by
Poisson’s integral with locally integrable uo (x). However we will not prove
this theorem since it requires some elementary measure theory.
96
Chapter 6. Steady Temperature in a Finite Cylinder.
∂ 2 us ∂ 2 us ∂ 2 us
(1) + + = ∆us = 0, x = (x1 , x2 , x3 ) ∈ Ω
∂x21 ∂x22 ∂x23
97
∂ 2 v 1 ∂v 1 ∂ 2v ∂ 2v
(5) + + + = 0, 0 < r < a, all θ, 0 < x3 < l
∂r2 r ∂r r2 ∂θ2 ∂x23
∂ 2 v 1 ∂v ∂ 2 v
(10) + + = 0, 0 < r < a, all θ, 0 < x3 < l
∂r2 r ∂r ∂x23
98
(14) v(0, x3 ) is finite.
∂ 2 v 1 ∂v ∂ 2 v
(15) + + = 0, 0 < r < a, all θ, 0 < x3 < l
∂r2 r ∂r ∂x23
(18) v(a, x3 ) = 0, 0 ≤ x3 ≤ l
BV Problem 2:
∂ 2 v 1 ∂v ∂ 2 v
(20) + + = 0, 0 < r < a, all θ, 0 < x3 < l
∂r2 r ∂r ∂x23
99
(24) v(0, x3 ) is finite.
BV Problem 3:
∂ 2 v 1 ∂v ∂ 2 v
(25) + + = 0, 0 < r < a, all θ, 0 < x3 < l
∂r2 r ∂r ∂x23
1
(rR0 )0 X + X 00 R = 0, 0 < r < a, 0 < x3 < l
r
and hence
1
(rR0 )0 X 00
r
=− = −λ2
R X
100
(31) R(0) finite
(32) R(a) = 0
(33) X 00 − λ2 X = 0
(34) X(l) = 0
(31), (32) and (34) are required by (19), (18) and (17), respectively. (30) is
a form of Bessel’s equation.33
Fortunately, the solutions to (33) are easily found. Indeed, X(x3 ) =
A cosh λx3 + B sinh λx3 . (34) will be used later.
∆w + λ2 w = 0
1 ∂ ∂v 1 ∂ 2v
(35) (r ) + 2 2 + λ2 v = 0
r ∂r ∂r r ∂θ
33
Named for German astronomer and mathematician Friedrich Bessel. The solutions of
(30)-(32) are known as Bessel functions, or (for obvious reasons) cylinder functions. Bessel
established basic properties of the solutions of (30)-(32) in 1824.
101
(36) v(r, θ + 2π) = v(r, θ)
∂ 2 v 1 ∂v 1 ∂ 2v
+ + + λ2 v = 0
∂r2 r ∂r r2 ∂θ2
π π π π
∂ 2v ∂ 2v
Z Z Z Z
1 ∂v 1
2
dθ + dθ + 2 dθ + λ2 vdθ = 0
−π ∂r r −π ∂r r −π ∂θ2 −π
The third integral may be computed and based on (36) is zero. Thus we have
π π π
∂ 2v
Z Z Z
1 ∂v
2
dθ + dθ + λ2 vdθ = 0
−π ∂r r −π ∂r −π
or
π π π
∂2
Z Z Z
1 ∂ 2
vdθ + vdθ + λ vdθ = 0
∂r2 −π r ∂r −π −π
Z π
(38) R(r) = vdθ
−π
102
may exploit this to obtain information about solutions of (30) which has
no solutions in terms of standard elementary functions. Suppose that w =
eax1 +bx2 . Substituting this into ∆w + λ2 w = 0 yields a2 + b2 + λ2 = 0. We
may choose a and b in any way, so select a = 0, b = iλ. This gives w = eix2 λ .
Introducing polor coordinates again gives
Z π
(40) R(r) = eirλ sin θ dθ
−π
Z π
1
(41) R(r) = eirλ sin θ dθ
2π −π
This specifies that R(0) = 1. Using Euler’s formula and that sin is an
odd function,
Z π
1
(42) R(r) = cos(irλ sin θ)dθ
π 0
Now substitution of the Maclaurin series for cosine into (42) gives
∞
1 X (−1)j (λr)2j π 2j
Z
R(r) = sin θdθ
π j=0 (2j)! 0
103
Reduction formulae can now be used to write
Z π
π(2j)!
sin2j θdθ =
0 22j (j!)2
∞
X (−1)j (λr)2j
(43) R(r) =
j=0
j!2 22j
Z π
2 2
(44) R(r) = cos(λr cos ω)dω
π 0
Z 1
2 cos(λrt)
(45) R(r) = √ dt
π 0 1 − t2
The standard notation for R(r) is Jo (rλ). From (42) we observe that
104
substituting λr for r.) Then
Z 1
2 cos rt
(47) Jo (r) = √ dt
π 0 1 − t2
r
cos(r − τ )
Z
2
(48) Jo (r) = √ p τ
dτ
π o 2rτ 1 − 2r
Z r Z r
2 cos(τ ) 2 sin(τ )
(49) Jo (r) = √ √ p τ
dτ cos r+ √ √ p τ
dτ sin r
π r o 2τ 1 − 2r
π r o 2τ 1 − 2r
As r → ∞
Z r Z ∞
cos(τ ) cos(τ )
√ p τ
dτ → √ dτ
o 2τ 1 − 2r o 2τ
with a similar result for the sine integral. Both limits can be computed
explicitly, i.e.,
Z ∞ Z ∞ √
sin(τ ) cos(τ )
√ dτ = √ dτ = π/2
o 2τ o 2τ
Consequently,
√
2 π
Jo (r) = √ [cos(r − ) + o(r)] as r → ∞
πr 4
105
Thus Jo (λr) behaves like cos(λr − π/4) for large r. Consequently when
r is large, J will have zeros in each of the intervals [ mπ
λ
+ π4 , (m+1)π
λ
+ π4 ].
Arranging the zeros of Jo (r) as
we see that Jo (r) has properties similar to sine and cosine, although it is not
periodic.
Since (30) is a second order equation, it has two linearly independent
solutions. Of course there are many such pairs. Selecting another solution
in addition to J is partly a matter of usefulness. None of these solutions are
important for our heat conduction problem however: they are all unbounded
at r = 0 which violates (31). However, a common second solution, known as
Weber’s function is given by
2 X
Yo (λr) = Jo (λr) ln r + ck r2k
π
106
Recall that X(l) = 0. Hence, An cosh λn l + Bn sinh λn l = 0. This implies
that
An = −Bn tanh λn l
∞
X
(50) v(r, x3 ) = vn (r, x3 ), v(r, 0) = f (r)
n=1
Then the question of convergence for the series (50) is decided in part by
Fourier-Bessel convergence theory.
Theorem. Let f 0 (r) be sectionally continuous on [0, a]. Then for each r,
0 < r < a, the series,
∞
X f (r+) + f (r−)
An Jo (λn r) =
n=1
2
Z a
2
An = 2 0 f (r)Jo (λn r)rdr
a Jo (βn ) 0
107
theorem in much the same was as we have already done. However, we shall
instead consider a similar problem.
Wave Motion in Two Dimensions.
Le Ω be a circular membrane (e.g., drumhead) of radius a, occupying the
region Ω = {(x1 , x2 ) : x21 + x22 ≤ a2 }.
If the membrane or drumhead is under tension, it will vibrate when struck.
For a low-mass drumhead, the transverse displacement from equilibrium u
satisfies the equation
∂ 2u 2
2 ∂ u ∂ 2u
(1) = c ( + ), x21 + x22 < a2 , t > 0
∂t2 ∂x21 ∂x22
∂u(x1 , x2 , 0)
(3) = u1 (x1 , x2 ), x21 + x22 ≤ a2
∂t
Conditions (2) and (3) specify the initial shape and velocity of the drum-
head, while condition (4) says that the drum skin is secured at the edge. As
with the vibrating string problem, we make the simplifying assumption that
u1 ≡ 0 (see p. 22). To further simplify our computations, let us assume that
a change of variable has been done to allow us to substitute 1 for c (see p.
13).
Changing to polar coordinates (u(x1 , x2 , t) = v(r, θ, t)) gives us the prob-
lem
108
∂ 2v 1 ∂ ∂v 1 ∂ 2v
(5) = (r ) + , 0 < r < a, t > 0
∂t2 r ∂r ∂r r2 ∂θ2
∂v(r, θ, 0)
(7) = 0, ∀ 0 ≤ r ≤ a, θ ∈ R
∂t
(8) v(r, θ, t) = 0, r = a, ∀ t ≥ 0
(11) v(a, θ, 0) = 0, ∀ θ ∈ R
Separation of Variables.
We perform this in two stages:
Stage 1. Assume v(r, θ, t) = w(r, θ)T (t). Substitution into (5) gives
1 ∂ ∂w 1 ∂ 2w
T 00 w = (r )T + 2 2 T
r ∂r ∂r r ∂θ
or
1 ∂ 1 ∂2w
T 00 (r ∂w ) + r2 ∂θ2
= r ∂r ∂r
= −λ2
T w
Thus
T 00 + λ2 T = 0, T 0 (0) = 0 (by (7)
109
and
1 ∂ ∂w 1 ∂ 2w
(r ) + 2 2 + λ2 w = 0
r ∂r ∂r r ∂θ
The last equation will be familiar from the previous steady-temperature prob-
lem.
Stage 2. Assume that w(r, θ) = R(r)Θ(θ). Substitution into the last
equation gives
1 ∂ ∂R 1 ∂ 2Θ
(r )Θ + 2 2 R + λ2 RΘ = 0
r ∂r ∂r r ∂θ
It follows that
1 ∂ ∂2Θ
(r ∂R ) 2 2 2 ∂θ2
r ∂r ∂r
r +λ r =− = µ2
R Θ
This implies
1 ∂ ∂R µ2
(12) (r ) + (λ2 − 2 )R = 0
r ∂r ∂r r
(13) Θ00 + µ2 Θ = 0,
110
(12) becomes
1 ∂ ∂R n2
(15) (r ) + (λ2 − 2 )R = 0 n = 1, 2, 3, ...
r ∂r ∂r r
1 ∂ ∂w 1 ∂ 2w
(16) (r ) + 2 2 + λ2 w = 0
r ∂r ∂r r ∂θ
π π π
∂ 2 w −inθ
Z Z Z
1 ∂ ∂ −inθ 1
(r we dθ) + 2 e dθ + λ2 we−inθ dθ = 0
r ∂r ∂r −π r −π ∂θ2 −π
π π π
n2
Z Z Z
1 ∂ ∂ −inθ −inθ
(17) (r we dθ) − 2 we dθ + λ 2
we−inθ dθ = 0
r ∂r ∂r −π r −π −π
Rπ
This implies that R(r) = −π
we−inθ dθ is a solution to (15). Equation
(39), p. 105 gives a possible solution to (16) which leads to
111
Z π
(18) R(r) = eirλ sin θ e−inθ dθ
−π
Z π
1
(19) R(r) = Jn (λr) = cos(rλ sin θ − nθ)dθ
π 0
∞
f (r+) + f (r−) X
= Cnm Jn (λnm r)
2 m=1
Moreover, the function Jn (x) has zeros 0 < βn1 < βn2 < βn3 < .... → ∞ and
λnm = βnm /a. Further the coefficients Cnm are found from
Z a
2
(20) Cnm = 2 2 f (r)Jn (λnm r)rdr
a Jn+1 (λnm a) 0
By the last equation on page 111, we have T (t) = A cos λt. This suggests
the solution of (5) is
112
∞ X
X ∞
(21) v(r, θ, t) = Cnm Jn (λnm r) (An cos nθ + Bn sin nθ) Dn cos λnm t
| {z }| {z }| {z }
n=0 m=1 T
R Θ
∞ X
X ∞
(22) f (r, θ) = Cnm Jn (λnm r)(An cos nθ + Bn sin nθ)
n=0 m=1
Z a Z π
2
Cnm An = 2 2 Jn (λnm r) cos(nθ)f (r, θ)rdθdr
πa Jn+1 (λnm a) 0 −π
113
(5)-(11) given by (21).
The functions
vnm (r, θ, t) = Cnm Jn (λnm r) (An cos nθ + Bn sin nθ) Dn cos λnm t
| {z }| {z }| {z }
R Θ T
are called modes of vibration or harmonics for the drumhead. The vibration
of a drumhead has been the source of many interesting questions. For ex-
ample, if the shape is not circular, can the shape be determined by merely
listening to it? The answer is yes if the drumhead is convex and has a rea-
sonably smooth edge.34 Otherwise, the answer is no in general.
34
Zelditch, “Spectral Determination of Analytic Bi-axisymmetric Plane Domains,” Ge-
ometric and Functional Analysis 10/3 (2000): 628-677.
114
Chapter 7. The Propagation of Electromagnetic
115
Appendix I: Classification of PDEs35
PDEs can be classified in different ways. Some of these are (A) geomet-
ric form, (B) historical significance, (C) associated persons, (D) “algebraic”
form.
As an example of (A) consider an equation of the type
∂ 2u ∂ 2u ∂ 2u ∂u ∂u
α 2
+ β + γ 2
+δ + = F (x, y) (1)
∂x ∂x∂y ∂y ∂x ∂y
This equation is classified as “2nd order” because the highest order derivative
is 2. The values of the coefficients may be used to give a further useful
classification. This is based on obvious correspondence to the equation
Geometrically, this equation describes one of the conic sections: ellipse, hy-
perbola, parabola, line. Which one depends on the relative value of the co-
efficients. The equation (1) is classified according as equation (2) is: elliptic,
hyperbolic, parabolic. For example,
y − x2 = 0 (3)
35
The perusal of any university library will reveal many monographs and research articles
on equations of various types. The survey of types of equations below is by no means
exhaustive.
116
is a parabola. Hence
∂u ∂ 2 u
− =0 (4)
∂y ∂x2
∂ 2u ∂ 2u
= x
∂x2 ∂y 2
∂u ∂ 3 u
u − =x (5)
∂x ∂y 3
The first term, u∂u/∂x, is identified as “nonlinear.”36 The reason for this is
that the form does not distribute over sums and constant multiples.
∂ 3u
4 3 (6)
∂z
is linear since
36
Nonlinear equations often present unique challenges and require customized techiques
to study their solutions and properties.
117
∂ 3 (u1 + u2 ) ∂ 3 u1 ∂ 3 u2
4 = 4 + 4 (7)
∂z 3 ∂z 3 ∂z 3
while
X ∂ |β| u
αn1 n2 n3 ...nk (x) = F (x) (9)
n1 ,n2 ,...,nk
∂xβ
where β = {n1 , n2 , ..., nk }, x = (x1 , x2 , ..., xk ), ∂xβ = ∂xn1 1 ∂xn2 2 ...∂xnk k and
|β| = n1 + n2 + n3 + ... + nk .
No terms involving products or other nonlinear functions of u and/or its
derivatives are linear.
Classification (C) by name of person or thing. The heat and wave equa-
tions are examples. Examples of (C) and (D) classification include a nonlinear
variant of the wave equation
∂ 2u ∂ 2u
− 2 = sin u(x, t) (10)
∂t2 ∂x
118
KdV describes shallow water waves, internal waves in density stratified me-
dia, some kinds of plasma waves, etc.
Sine-Gordon is a nonlinear hyperbolic equation whose name arose as a
parody of the Klein-Gordon equation:
∂ 2u ∂ 2u
− 2 = m2 u(x, t) (12)
∂t2 ∂x
∂u(x, t) 1
i =− ∆u + V (x)u (13)
∂t 2m
119
ample, Maxwell’s equations (in a vacuum) may be written:
0 0 0 0 − ∂x∂ 3 ∂
∂x2
∂
0
0 0 ∂x3
0 − ∂x∂ 1
0 0 0 − ∂x∂ 2 ∂
0
∂~u ∂x1
= ~u (14)
∂t
0 ∂
− ∂x∂ 2 0 0 0
∂x3
− ∂ 0 ∂
0 0 0
∂x3 ∂x1
∂
∂x2
− ∂x∂ 1 0 0 0 0
where
e1 (x1 , x2 , x3 , t)
e2 (x1 , x2 , x3 , t)
e3 (x1 , x2 , x3 , t)
~u(x1 , x2 , x3 , t) =
(15)
h1 (x1 , x2 , x3 , t)
h2 (x1 , x2 , x3 , t)
h3 (x1 , x2 , x3 , t)
0 0 0 0 − ∂x∂ 3 ∂
∂x2
∂
0
0 0 ∂x3
0 − ∂x∂ 1
0 0 0 − ∂x∂ 2 ∂
0
∂~u ∂x1
E = ~u (16)
∂t
0 ∂
− ∂x∂ 2 0 0 0
∂x3
− ∂ 0 ∂
0 0 0
∂x3 ∂x1
∂
∂x2
− ∂x∂ 1 0 0 0 0
120
or
0 0 0 0 − ∂x∂ 3 ∂
∂x2
∂
0
0 0 ∂x3
0 − ∂x∂ 1
0 0 0 − ∂x∂ 2 ∂
0
∂~u −1 ∂x1
=E ~u (17)
∂t 0 ∂
− ∂x∂ 2 0 0 0
∂x3
− ∂ 0 ∂
0 0 0
∂x3 ∂x1
∂
∂x2
− ∂x∂ 1 0 0 0 0
n X
n n
X ∂ 2u X ∂u
aij + bk + cu = 0 (18)
i=1 j=1
∂xi ∂xj k=1 ∂xk
Then (18) is classified as elliptic when the matix (aij ) has all positive eigen-
values, or all negative; hyperbolic when only one eigenvalue is negative and
the rest are positive or vice versa; parabolic when one eigenvalue is zero and
the others have the same sign. Obviously this does not cover all possible
cases - there could be two positive and two negative eigenvalues for example.
The geometric classification may be extended to systems of equations like
(17).
121
Appendix II: Bessel Functions and Sturm-Lioville Problems
1 ∂ ∂
B= (r ) (19)
r ∂r ∂r
so that
1 ∂ ∂u
Bu = (r ) (20)
r ∂r ∂r
1 ∂ ∂u ∂ 2 u 1 ∂u
Bu = (r ) = 2 + (21)
r ∂r ∂r ∂r r ∂r
lim u(r) < ∞, lim+ ru0 (r) = 0, lim+ v(r) < ∞, lim+ rv 0 (r) = 0 (22)
r→0+ r→0 r→0 r→0
Observe that
Z a Z a
1 ∂ ∂u 1 ∂ ∂v
[vBu − uBv]rdr = [v (r ) − u (r )]rdr
c c r ∂r ∂r r ∂r ∂r
37
Phillip Hartman, Ordinary Differential Equations. Cambrige: Cambridge Univ. Press,
2002.
122
Z a Z a
∂ ∂u ∂ ∂v
= v (r )dr − u (r )dr
c ∂r ∂r c ∂r ∂r
By (22)
Z a Z a
lim+ [vBu − uBv]rdr = [vBu − uBv]rdr (23)
c→0 c 0
Thus
Z a
[vBu − uBv]rdr = 0 (25)
0
123
Z a Z a
[u2 Bu1 − u1 Bu2 ]rdr = (λ21 − λ22 ) u1 u2 rdr (28)
0 o
By (23)-(24), (26),
Z a
u1 u2 rdr = 0 (29)
o
Now observe that (26) is Bessel’s equation and Jo (λr) satisfies (26) and (22)
( by (19) p. 114). Hence
Z a
Jo (λn r)Jo (λm r)rdr = 0 (30)
o
Multiply both sides of (31) by Jo (λn r)r and integrate. This gives
Z a Z ∞
aX
f (r)Jo (λn r)rdr = Am Jo (λm r)Jo (λn r)rdr (32)
0 0 m=1
∞ Z
X a Z a
= Am Jo (λm r)Jo (λn r)rdr = An Jo (λn r)Jo (λn r)rdr (33)
m=1 0 0
Ra
f (r)Jo (λn r)rdr
An = 0 R a 2 (34)
0 o
J (λn r)rdr
124
Now recall
Z a
[vBu − uBv]rdr = −[rv(r)u0 (r) − ru(r)v 0 (r)]|ac (35)
c
a
aλm Jo (µa)Jo0 (aλm )
Z
Jo (µr)Jo (λm r)rdr = (36)
0 µ − λm
Since Jo (aλm ) = 0, take the limit of both sides of (36) (l’hôpital’s rule
can be used on the right hand side) as µ → λm
a
a2 0
Z
Jo2 (λm r)rdr = [J (λm a)]2 (37)
0 2 o
—————
The Bessel functions Jn (λr) of order n also satisfy orthogonality relations.
Indeed, there are numerous identities satisfied by these functions. Setting
n2
B1 u = Bu + r2
u we see Bessel’s equation (µ = n) can be written as
B1 u = λ2 u (38)
Now
a a
n2 n2
Z Z
(vB1 u − uB1 v)rdr = (vBu − uBv + v u − u v)rdr (39)
c c r2 r2
Z a
= (vBu − uBv)rdr = −[rv(r)u0 (r) − ru(r)v 0 (r)]|ac (40)
c
125
The same argument used above now shows that
Z a
Jn (λnm2 r)Jn (λnm1 r)rdr = 0 (41)
o
when m1 6= m2 .
The coefficients in (20), p. 114 are computed following the same method
as (32)-(34) above.
126
Appendix III: Fourier Series
Z l
2 nπx
cn = sin( )f (x)dx (42)
l 0 l
Z l Z l
2 nπx 2 nπx 2 nπx 0
cn = sin( )f (x)dx = − cos( )f (x)|l0 + cos( )f (x)dx
l 0 l nπ l nπ 0 l
(43)
Note that if f (0) = f (l) = 0 then the first term on the right side of (43)
vanishes. Next, suppose that f 00 (x) is sectionally continuous. Then we may
repeat the integration by parts:
Z l
2 nπx
cn = sin(
)f (x)dx
l
0 l
Z l
2l nπx 0 l 2l nπx 00
= − 2 2 sin( )f (x)|0 + 2 2 sin( )f (x)dx (44)
nπ l nπ 0 l
127
cn sin( nπx
P
on page 8) and that l
) also converges uniformly and absolutely.
The theorem on page 7 is rather better than this, since it requires nothing
of f 00 (x).
One can conclude from the theorem on page 7 that if f 0 (0) = f 0 (l) and
f 00 (x) is sectionally continuous on [0, l] that it is possible to expand f 0 (x) as
a Fourier series as well, etc.
Continuity is not necessary in the consideration of Fourier series. Indeed,
the most natural requirement turns out to be a condition like38
Z l
|f (x)|2 dx < ∞ (45)
0
128
Moreover, Sturm-Lioville problems always lead to solutions with properties
which are Fourier-like. Indeed, such series are often called generalized Fourier
series. Thus Bessel series are generalized Fourier series.
The question of whether a given function can be expanded as a Fourier
series is a complicated one and has led to very complex mathematics. Func-
tions, even highly discontinuous ones, which satisfy (45) always have conver-
gent Fourier series, at least in a certain general sense:
Z l n
X kπx kπx
lim | ak cos( ) + bk sin( ) − f (x)|2 dx = 0 (47)
n→∞ −l k=0
l l
A very complex result, due to Lennart Carleson shows that functions satis-
fying (45) may be expanded as a Fourier series in the usual sense, except at
certain exceptional points which together form a set of zero length.40 On the
other hand, it has been shown that if one only requires that
Z l
|f (x)|dx < ∞ (48)
0
then the Fourier series for f (x) may not converge at any point at all. Func-
tions satisfying (45) also satisfy (48), but not necessarily the other way
around. Hence such extreme examples must satisfy (48) but not (45).
40
“On convergence and growth of partial sums of Fourier series.” Acta Mathematica 116
(1966): 135-157. Carleson’s proof is so complex that it took considerable time for other
mathematicians to verify it. No simplified proof has been found. Condition (45) was
Rl
generalized to −l |f (x)|p dx < ∞ for any fixed real value p > 1 by Richard Hunt in 1967.
129
Orthogonality Relations.
The trigonometric functions satisfy orthogonality relations (see footnote
3 p. 7) which may be established directly or by means similar to those used
in appendix II above. Such relations include
Z l
nπx kπx
sin( ) sin( )dx = 0 when n 6= k (49)
0 l l
Z l
nπx kπx
sin( ) sin( )dx = 0 when n 6= k (50)
−l l l
Z l
nπx kπx
sin( ) cos( )dx = 0 when n 6= k (51)
−l l l
Z l
nπx kπx
cos( ) cos( )dx = 0 when n 6= k (52)
−l l l
Z l
nπx kπx
cos( ) cos( )dx = 0 when n 6= k (53)
0 l l
Thus
130
or
nπx kπx 1 n+k n−k
cos( ) cos( ) = (cos( πx) + cos( πx))
l l 2 l l
l
n−k
Z
nπx kπx l n+k l
cos( ) cos( )dx = sin( πx)|l0 + sin( πx)|l0
0 l l 2π(n + k) l 2π(n − k) l
(54)
which is equal to zero as long as n 6= ±k.
The convergence theorem of chapter 1 was originally proved by Fourier
himself. A proof essentially constructed by Dirichlet will now be given.
where
Z π Z π
1 1
an = cos nxf (x)dx, bn = sin nxf (x)dx, n = 1, 2, ... (56)
π −π π −π
and
Z π
1
a0 = f (x)dx (57)
2π −π
131
N
X
SN (x) = an cos nx + bn sin nx (58)
n=0
(a partial sum of the Fourier series). We shall sum this expression using
Dirichlet’s technique. We insert the equations from (56)-(57) into (58) to get
N
X
SN (x) = an cos nx + bn sin nx (59)
n=0
π N
1 π 1 π
Z Z Z
1 X
= f (y)dy + cos nyf (y)dy cos nx + sin nyf (y)dy sin nx
2π −π n=1
π −π π −π
(60)
Z π N
1 1 X
= f (y)[ + cos ny cos nx + sin ny sin nx]dy (61)
π −π 2 n=1
Z π N
1 1 X
= f (y)[ + cos n(y − x)]dy (62)
π −π 2 n=1
N
1 X
DN (z) = + cos nz (63)
2 n=1
N
z X z
2 sin(z/2)DN (z) = sin + 2 cos nz sin (64)
2 n=1 2
N
z X
= sin + sin((n + 1/2)z) − sin((n − 1/2)z) (65)
2 n=1
132
The sum telescopes to a single term. Thus
sin((N + 1/2)z)
DN (z) = (66)
2 sin(z/2)
Thus
Z π
1
SN (x) = f (y)DN (y − x)dy (67)
π −π
Step 2. Define
f (x + y) − f (x+)
g(y) = , 0<y≤π (68)
2 sin(y/2)
Then g(y) is sectionally continuous (note that g(0) = f 0 (x+)). We shall show
that
Z π
g(y) sin((N + 1/2)y)dy → 0, as N → ∞ (69)
0
First, if g(x) is any continuous function on [−π, π], let [a, b] be any subinterval
of [−π, π]. Then
Z b
1
sin nxdx = (cos nb − cos na) → 0, as n → ∞ (70)
a n
Now, let
k
X
gk (x) = g(xj )χ[xj ,xj+1 ) (x) (71)
j=1
where {xj } is a regular partition of [−π, π]. Then it easy to see that gk (x) →
133
g(x) uniformly. Moreover (70) shows that
Z π
gk (x) sin(nx)dx → 0, as n → ∞ (72)
−π
Now observe
Z π
g(x) sin(nx)dx (73)
−π
Z π Z π Z π
= gk (x) sin(nx)dx − { gk (x) sin(nx)dx − g(x) sin(nx)dx} (74)
−π −π −π
Z π Z π
= gk (x) sin(nx)dx − (gk (x) − g(x)) sin(nx)dx (75)
−π −π
Therefore
Z π Z π Z π
| g(x) sin(nx)dx| ≤ | gk (x) sin(nx)dx| + |gk (x) − g(x)|| sin(nx)|dx
−π −π −π
(76)
Z π Z π
≤| gk (x) sin(nx)dx| + |gk (x) − g(x)|dx (77)
−π −π
Z π
|gk (x) − g(x)|dx < /2 (78)
−π
Z π
| gk (x) sin(nx)dx| < /2 (79)
−π
134
Then for all such n,
Z π
| g(x) sin(nx)dx| < (80)
−π
Z π
DN (z)dz = 1 (81)
−π
Z π Z 0
1
= DN (z)dz = DN (z)dz (82)
2 0 −π
Step 4.
Z π Z π
1 1
SN (x) = f (y)DN (y − x)dy = f (y)DN (y − x)dy (83)
π −π π −π
Z π−x Z π
1 1
= f (t + x)DN (t)dt = f (t + x)DN (t)dt (84)
π −π−x π −π
Step 5.
Z 0
f (x+) + f (x−) 1 1
SN (x) − = f (x + t)DN (t)dt − f (x−)
2 π −π 2
41
Actually g being sectionally continuous, requires a slight modification of the proof,
breaking up the integral over the intervals of continuity. The conclusion is the same.
Moreover, the same conclusion holds for the cosine. The result of step 2 is a version of
the “Riemann-Lebesgue Lemma” which holds for much less well-behaved functions g.
135
Z π
1 1
+ f (x + t)DN (t)dt − f (x+)
π 0 2
Now
Z 0 Z 0 Z 0
1 1 1 1
f (x+t)DN (t)dt− f (x−) = f (x+t)DN (t)dt− DN (t)dtf (x−)
π −π 2 π −π π −π
0
1 0
Z Z
1
= f (x + t)DN (t)dt − f (x−)DN (t)dt
π −π π −π
1 0
Z
= (f (x + t) − f (x−))DN (t)dt
π −π
QED.
Corollary. If f (x) is an odd function, then only the sine terms appear
in the expansion. Alternatively, if f (x) is even, then only the cosine terms
appear in its Fourier series expansion.
Proof. This a simple consequence of (56), (57) above. In the odd case for
example, an will clearly be zero, since cos nxf (x) will be odd, therefore its
integral is even. Hence integrating over the symmetric interval must give zero.
136
cosine expansion. Moreover in this case, the integrals in (56), (57) can be
written as integrals over [0, π].
The Fourier series itself is defined on all (−∞, ∞) and therefore gives a
periodic extension of f (x) to all of R.
∞
X
un (x)
n=1
137
Moreover, if the functions un are differentiable and the series
∞
X dun
n=1
dx
du
converges uniformly on [a, b] then indeed dx
exists and
∞
X dun du
= .
n=1
dx dx
k
X
| un (x) − u(x)| <
n=1
for all x ∈ Ω.
Now we construct a proof of the theorem on uniform convergence. First
we dispose of the continuity of limits portion. So suppose each un (x) is
138
P
continuous on [a, b] and that un (x) converges uniformly to a function u(x)
on [a, b]. We want to show that u(x) is continuous on [a, b]. To this end,
suppose > 0. Let N be large enough so that
k
X
| un (x) − u(x)| <
n=1
for all x ∈ [a, b] whenever k ≥ N . Now each un (x) being continuous means
that for any xo ∈ [a, b] there exists δn > 0 so that if |x − xo | < δn then
|un (x) − un (xo )| < /k. So choose x so that |x − xo | < min δn Thus
k
X k
X k
X k
X
|u(x)−u(xo )| = |u(x)− un (x)+ un (x)+ un (xo )− un (xo )−u(xo )|
n=1 n=1 n=1 n=1
k
X k
X k
X k
X
≤ |u(x) − un (x)| + |u(xo ) − un (xo )| + | un (x) − un (xo )|
n=1 n=1 n=1 n=1
k
X k
X
< 2 + |un (x) − un (xo )| < 2 + /k = 3
n=1 n=1
Hence if |x − xo | < min δn ≡ δ then |u(x) − u(xo )| < 3. This shows that
u is continuous on [a, b].
For the second portion of the proof concerning differentiability it is some-
what simpler to prove the following lemma:
139
that fn0 (x) converges uniformly to g(x). Then f 0 (x) exists and is equal to
g(x).
Proof of Lemma. Observe that the mean value theorem implies that for any
x, y ∈ [a, b] there a z ∈ [x, y] (note that z depends on both n and m) such
that,
0
(fm (x) − fn (x)) − (fm (y) − fn (y)) = (x − y)(fm (z) − fn0 (z))
If follows that
0
|fm (z)−fn0 (z)| = |fm
0
(z)−g(z)+g(z)−fn0 (z)| ≤ |fm
0
(z)−g(z)|+|g(z)−fn0 (z)| ≤
140
f (x) − f (y)) fn (x) − fn (y)
| − |<
x−y x−y
for |x − y| close to zero. This shows that not only does f 0 (y) exist, but it
must be equal to g(y), which is the conclusion of the lemma.
Now to apply the lemma, we take f (x) = u(x), g(x) = u0n (x), fn (x) =
P
Pn
k=1 uk (x). QED.
141
An often useful theorem attributed to Weierstrass is the “Weierstrass M-
Test.” The name arises from the statement of the theorem:
P
Theorem Weierstrass M-Test. Suppose un (x) is an infinite series of
functions defined on [a, b]. Suppose that there is a sequence of constants
{Mn } with the property |un (x)| ≤ Mn for each n = 1, 2, 3, . . . . Suppose also
P P
that Mn converges (as a series of non-negative constants). Then un (x)
converges uniformly.
P
Proof. By the hypotheses, un (x) is absolutely convergent and therefore
converges. Let
N
X
UN (x) = un (x)
n=1
and let
∞
X
g(x) = un (x).
n=1
∞
X
Mn <
n=K+1
It follows that
∞
X
|un (x)| < .
n=K+1
142
Now we have
∞
X XK ∞
X ∞
X
|g(x)−UK (x)| = | un (x)− un (x)| = | un (x)| ≤ |un (x)| < .
n=1 n=1 n=K+1 n=K+1
Remark 1. The test may be used with functions of more than one vari-
able. The proof is the same.
∞
∂u X ∂un
=
∂xi n=1
∂xi
for all x ∈ Ω.
143
Proof. The proof consists of reducing this general case to the one variable
proof. First, by hypothesis, Ω ⊆ Rk . Hence x ∈ Ω =⇒ x = (x1 , x2 , x3 , ...xk ).
Fix x ∈ Ω and consider a small box, Bδ = {y ∈ Ω : yj = −δ ≤ xj ≤ δ} ⊆ Ω
containing x. In the proof of the one variable case, replace [a, b] by [−δ, δ]
and x by xi .
144
Appendix IV - Limits
The purpose here is to give some brief definitions of useful limit ideas.
The most important are
To define these, we need to know what sup and inf are. Sup stands for
supremum, inf stands for infimum. These in turn are synonyms for least
upper bound and greatest lower bound respectively. It is a fact that any set
of real numbers has both a supremum and infimum. For example,
sup{x|x < 4} = 4
and
1
inf{ |n = 1, 2, 3, 4, . . .} = 0
n
and
sup{x|x > 0} = ∞
145
Observe that when 1 > 2 , {f (x)|x ∈ N (a, 1 )} ⊇ {f (x)|x ∈ N (a, 2 )}.
Thus sup{f (x)|x ∈ N (a, 1 )} ≥ sup{f (x)|x ∈ N (a, 2 )}. It follows that
whereas
inf{f (x)|x ∈ N (a, 1 )} ≤ inf{f (x)|x ∈ N (a, 2 )}
whenever 1 > 2 .
When more than one variable is involved, the definitions remain essen-
tially the same.
Let N ((a, b), ) = {(c, d)||(c, d) − (a, b)| < }. Then we define
lim sup f (x, t) = lim sup{f (x, t)||(x, t) ∈ N ((a, b), )}
(x,t)→(a,b) →0
146
lim sup f (x) = lim sup{f (x)|y ≤ x < a}
x→a− y→a
Observe that
whenever y1 ≤ y2 because
lim inf
−
f (x) = lim inf{f (x)|y ≤ x < a}
x→a y→a
and
147
with
lim inf f (x) = lim inf{f (x)|a < x ≤ y}
x→a+ y→a
Theorem
lim f (x)
x→a
The same statement is true for one-sided limits as well as for functions of
more than one variable.
Example.
Let
1
f (x) = sin( ), x 6= 0
x
It is clear that
lim f (x) does not exist.
x→0
However,
lim sup f (x) = 1, lim inf f (x) = −1
x→0 x→0
lim sup and lim inf give information about a function’s behavior at a
148
point and can be used to control or estimate such behavior as in Theorem 2
on page 91.
When more than one variable is involved, we don’t have a concept of
“one-sided” limit exactly, but sometimes it is useful to restrict the idea of
lim sup, lim inf in a somewhat similar way. For example,
———————————————————————
Promised extended argument from page 9 above:
∞
X π nπx −k( nπ )2 t
−n2 ( )2 cn sin e l
n=1
l l
which is the (term by term) derivative of u(x, t), with respect to t, suggests
that we consider the following estimate:
149
nπ 2
≤ n2 e−k( l
) t
provided n is large enough so that |cn |( πl )2 ≤ 1, say for all n ≥ m for some
m. Now fix to > 0 and δ > 0. For t > to + δ we claim that for large enough
n, that
nπ 2 nπ 2
n2 e−k( l
) t
≤ e−k( l
) to
2 2t
n2 e−cn t ≤ e−cn o
This is equivalent to
2 (t−t )
n2 e−cn o
≤1
n2
→0
ecn2 δ
n2
≤1
ecn2 δ
150
or
2 (t−t )
n2 e−cn o
≤1
2t
Taking Mn = e−cn o
gives us the correct hypothesis for the Weierstrass M-
test provided we can show that
∞
X
Mn < ∞
n=m
n2
→0
ecn2 to
1 2
2
≥ e−cn to = Mn
n
∞
X π nπx −k( nπ )2 t
−n2 ( )2 cn sin e l
n=1
l l
π nπx −k( nπ )2 t
|n2 ( )2 cn sin e l | ≤ Mn
l l
151
have term by term differentiability for all t. Also, since the second partial
derivative with respect to x of u(x, t) delivers a similar series, we have that
both sides of the heat equation are defined for u(x, t) (provided that f satisfies
the conditions of the Fourier convergence theorem).
152