Differential Equations UNSW
Differential Equations UNSW
Linear ODEs
Introduction
Singular ODEs
In linear algebra, you have seen the advantages arising from the
compact notation Ax = b for a system of linear equations. A
similarly compact notation is of great value when dealing with a
linear differential equation
Lu = f.
X
m
Lu(x) = aj (x)Dj u(x)
j=0 (1)
= am Dm u + am−1 Dm−1 u + · · · + a0 u,
Example
N(u) = u 00 + u2 u 0 − u is a nonlinear differential operator of
order 2.
Linear initial-value problem
Consider a general mth-order linear differential operator
X
m
Lu = aj (x)Dj u.
j=0
Theorem
Assume that the ODE Lu = f is not singular with respect to [a, b],
and that f is continuous on [a, b]. Then the IVP (2) and (3) has a
unique solution.
Homogeneous problem
Theorem
Assume that the linear, mth-order differential operator L is not
singular on [a, b]. Then the set of all solutions to the homogeneous
equation Lu = 0 on [a, b] is a vector space of dimension m.
Proof.
Let V = { u : Lu = 0 on [a, b] } and define the linear
transformation Θ : V → Rm by
Example
The general solution to u 00 − u 0 − 2u = 0 is
u 00 + p(x)u 0 + q(x)u = 0,
u1 w 0 + (2u10 + pu1 )w = 0.
Reduction of order (continued)
Writing the ODE for w in the standard form
w 0 + (2u10 u−1
1 + p)w = 0,
we seek an integrating
R factor R
0 −1
I(x) = exp (2u1 u1 + p) dx = u21 exp( p dx), so that
d
(Iw) = Iw 0 + I 0 w = I w 0 + (2u10 u−1
1 + p)w = 0.
dx
Then Iw = C for some constant C, and so
Z
C
v= dx.
I(x)
Example
For the ODE u 00 − 6u 0 + 9u = 0, take u1 = e3x and find v.
Differential operators with constant coefficients
X
m
p(z) = aj zj = am zm + am−1 zm−1 + · · · + a1 z + a0 ,
j=0
and so
p(D)eλx = 0 ⇐⇒ p(λ) = 0.
Factorization
k1 + k2 + · · · + kr = m.
Lemma
(D − λ)xj eλx = jxj−1 eλx for j > 0.
Lemma
(D − λ)k xj eλx = 0 for j = 0, 1, . . . , k − 1.
Proof
Lemma
If (z − λ)k is a factor of p(z) then the function u(x) = xj eλx is a
solution of Lu = 0 for 0 6 j 6 k − 1.
Proof.
Write p(z) = (z − λ)k q(z), so that q(z) is a polynomial of
degree m − k. It follows that
and so for 0 6 j 6 k − 1,
Theorem
For the constant-coefficient case, the general solution of the
homogeneous equation Lu = 0 is
X
r kX
q −1
X
r kX
q −1
X
r kX
l −1
Example
From the factorization
is
u = c1 e−3x + c2 + c3 ex + c4 e4x .
Repeated real root
Example
From the factorization
is
u = c1 ex + c2 e−2x + c3 xe−2x + c4 e−3x .
Complex root
Example
From the factorization
is
d2 x
m ẍ = m = f(t) − r(v)v − k(x)x,
dt2
leads to a second-order differential equation
m ẍ + r0 ẋ + k0 x = f(t).
(k0 − mω2 )C + r0 ωE = 0,
−r0 ωC + (k0 − mω2 )E = F.
u1 u2
W(x) = = u1 u20 − u2 u10 ,
u10 u20
and when m = 3,
u1 u2 u3
W(x) = u10 u20 u30 .
u100 u200 u300
Example
The Wronskian of the functions u1 = e2x , u2 = xe2x and
u3 = e−x is
Example
The Wronskian of the functions u1 = ex cos 3x and u2 = ex sin 3x
is
ex cos 3x ex sin 3x
W= = 3e2x .
ex cos 3x− 3e sin 3x e sin 3x + 3ex cos 3x
x x
Linearly dependent functions
Lemma
If u1 , . . . , um are linearly dependent over an interval [a, b] then
W(x; u1 , . . . , um ) = 0 for a 6 x 6 b.
Example
The functions u1 = cosh x, u2 = sinh x and u3 = ex are linearly
dependent because
ex + e−x ex − e−x
cosh x + sinh x = + = ex .
2 2
Their Wronskian is
cosh x sinh x ex
W = sinh x cosh x ex = 0.
cosh x sinh x ex
Proof (for m = 3)
Assume that u1 , u2 , u3 are linearly dependent on the
interval [a, b], that is, there exist constants c1 , c2 , c3 , not all zero,
such that
so
u1 (x) u2 (x) u3 (x) c1 0
u 0 (x) u 0 (x) u 0 (x) c2 = 0 for a 6 x 6 b.
1 2 3
u100 (x) u200 (x) u300 (x) c3 0
Example
The second-order ODE
u 00 + 3u 0 − 4u = 0
ex e−4x
= −5e−3x ,
e −4e−4x
x
and satisfies W 0 + 3W = 0.
Proof (for m = 2)
so
Theorem
Let u1 , u2 , . . . , um be solutions of a non-singular, linear,
homogeneous, mth-order ODE Lu = 0 on the interval [a, b].
Either
W(x) = 0 for a 6 x 6 b and the m solutions are linearly
dependent,
or else
W(x) 6= 0 for a 6 x 6 b and the m solutions are linearly
independent.
Proof
The Wronskian satisfies
am−1
W 0 + pW = 0 for a 6 x 6 b, where p= .
am
Define an integrating factor
Z
I(x) = exp p(x) dx 6= 0,
u1 , u2 , u3 linearly dependent =⇒ W ≡ 0.
then u1 has exactly one zero between any two successive zeros
of u2 in the interval [a, b].
Example
The functions u1 (x) = cos x and u2 (x) = sin x are linearly
independent solutions of u 00 + u = 0 on any interval [a, b].
Example
What about u 00 − u = 0?
Proof
Suppose a 6 α < β 6 b with
Case 1: W(x) > 0 for a 6 x 6 b and u2 (x) > 0 for α < x < β.
Thus,
u2 (α + h)
u20 (α) = lim+ > 0,
h→0 h
u2 (β + h)
u20 (β) = lim− 6 0.
h→0 h
Case 3: W(x) < 0 for a 6 x 6 b and u2 (x) > 0 for α < x < β.
Similar argument shows u1 (α) < 0 and u1 (β) > 0.
Case 4: W(x) < 0 for a 6 x 6 b and u2 (x) < 0 for α < x < β.
Similar argument shows u1 (α) > 0 and u1 (β) < 0.
In all cases, u1 (α) and u1 (β) have opposite signs, and so by the
Intermediate Value Theorem u1 (ξ) = 0 for at least one ξ ∈ (α, β).
Theorem
Assume that L1 6= 0, or equivalently a0 = p(0) 6= 0. For any
integer r > 0, there exists a unique polynomial uP of degree r such
that LuP = xr .
For simplicity, we prove the result only for the case m = 2. Thus,
Lu = a2 u 00 + a1 u 0 + a0 u,
xr xr−1
= a 0 cr + a0 cr−1 + a1 cr
r! (r − 1)!
X
r−2
xj
+ a2 cj+2 + a1 cj+1 + a0 cj ,
j!
j=0
a0 cj + a1 cj+1 + a2 cj+2 = 0, 0 6 j 6 r − 2,
a0 cr−1 + a1 cr = 0,
a0 cr = r!.
so
2C − E + 6F = 0,
2E − 2F + 18G= 0,
2F − 3G= 0,
2G = 8
eµx
uP (x) =
p(µ)
u 00 + 4u 0 − 3u = 3e2x
is
e2x
uP = .
3
Product of polynomial and exponential
Theorem
Let L = p(D) and assume that p(µ) 6= 0. For any integer r > 0,
there exists a unique polynomial v of degree r such that
uP = v(x)eµx satisfies LuP = xr eµx .
Proof.
Again, for simplicity, we prove the result only for m = 2.
Put v = e−µx u so that u = veµx , and observe that
so
3C = 9 and − 7C + 3E = 0.
Thus, C = 3 and E = 7, giving
uP = (3x + 7)e−2x .
Polynomial solutions: the remaining case
Theorem
Let L = p(D) and assume p(0) = p 0 (0) = · · · = p(k−1) (0) = 0
but p(k) (0) 6= 0 where 1 6 k 6 m − 1. For any integer r > 0,
there exists a unique polynomial v of degree r such that
up (x) = xk v(x) satisfies LuP = xr .
For simplicity, we discuss only the case m = 2 and k = 1; thus,
Write
X
r
xj
v(x) = cj
(j + 1)!
j=0
so that
X
r
xj+1 X
r+1
xj
uP (x) = xv(x) = cj = cj−1 .
(j + 1)! j!
j=0 j=1
Since
X
r+1
xj−1 X xj
r
uP0 (x) = cj−1 = cj
(j − 1)! j!
j=1 j=0
and
X
r
xj−1 X
r−1
xj
uP00 (x) = cj = cj+1 ,
(j − 1)! j!
j=1 j=0
we have
xr X
r−1
xj
LuP = a1 cr + a2 cj+1 + a1 cj .
r! j!
j=0
a1 cj + a2 cj+1 = 0, 0 6 j 6 r − 1,
a1 cr = r!.
An example
Let Lu = u 000 + 2u 00 and seek a particular solution to Lu = 12x2 .
The theorem ensures that
so
4C + 6 E = 0,
12E + 24F = 0,
24F = 12
x2
uP = (3 − 2x + x2 ).
2
Exponential times polynomial: remaining case
Lemma
If u(x) = w(x)eµx then
X
m
zj
p(D)u = eµx q(D)w where q(z) = p(j) (µ) .
j!
j=0
Theorem
Let L = p(D) and assume p(µ) = p 0 (µ) = · · · = p(k−1) (µ) = 0
but p(k) (µ) 6= 0, where 1 6 k 6 m − 1. For any integer r > 0,
there exists a unique polynomial v of degree r such that
uP (x) = xk v(x)eµx satisfies LuP = xr eµx .
Proof.
Since q(j) (0) = p(j) (µ) for all j, there is a unique polynomial v of
degree r such that w(x) = xk v(x) satisfies q(D)w = xr and hence
p(D)uP = eµx q(D)w = eµx xr .
Proof of Lemma
X
m X
m X
k
k
µx k µx
Dj w Dk−j eµx
p(D)we = ak D we = ak
j
k=0 k=0 j=0
X
m X
k
k!
= ak Dj w Dk−j eµx
j!(k − j)!
k=0 j=0
Xm
DwX
j m
= ak k(k − 1)(k − 2) · · · (k − j + 1)µk−j eµx
j!
j=0 k=j
X
m
Dj w X
m
= ak k(k − 1)(k − 2) · · · (k − j + 1)µk−j eµx
j!
j=0 k=j
Xm
Dj w (j) X Dj w
m
= p (µ)eµx = eµx p(j) (µ)
j! j!
j=0 j=0
µx
=e q(D)w.
An example
Consider the ODE
uP = Cx2 e2x
and find
LuP = (−8F + 12G)ex cos 2x + (−12F − 8G)ex sin 2x = 13ex sin 2x,
so
xex
uP = − 3 cos 2x + 2 sin 2x .
4
Variation of parameters
What if f is not a polynomial times an exponential, or if L does
not have constant coefficients?
Consider a linear, second-order, inhomogeneous ODE with leading
coefficient 1:
Lu1 = 0, Lu2 = 0, W 6= 0.
is a (particular) solution to Lu = f.
Variation of parameters (continued)
To simplify the expression
u 0 = v1 u10 + v2 u20 .
Lu = v1 Lu1 + v2 Lu2 + v10 u10 + v20 u20 = v10 u10 + v20 u20 ,
v10 u1 + v20 u2 = 0,
v10 u10 + v20 u20 = f.
Thus, we have a pair of equations for the unknown v10 and v20 . In
matrix form
u1 (x) u2 (x) v10 (x)
0
= ,
u10 (x) u20 (x) v20 (x) f(x)
so
v10 (x)
0
1 u2 (x) −u2 (x) 0
= ,
v20 (x) W(x) −u10 (x) u1 (x) f(x)
or in other words,
−u2 (x)f(x) u1 (x)f(x)
v10 (x) = and v20 (x) = .
W(x) W(x)
Example
Find the general solution to
A0 = 1, A1 = 2, A2 = 2, A3 = 3, ...,
giving
u(x) = 1 + 2x + 2x2 + 3x3 + · · · .
Since
Ak+2 xk+2 k+2 2
lim k
= lim x = x2 ,
k→∞ Ak x k→∞ k + 1
P P∞
the ratio test shows that ∞ j=0 A2j x
2j and
j=0 A2j+1 x
2j+1
2 2
converge for x < 1 but diverge for x > 1.
General case
Consider a general second-order, linear, homogeneous ODE
Equivalently,
u 00 + p(x)u 0 + q(x)u = 0,
where
a1 (x) a0 (x)
p(x) = and q(x) = .
a2 (x) a2 (x)
Assume that aj is analytic at 0 for 0 6 j 6 2, and that a2 (0) 6= 0.
Then p and q are analytic at 0, that is, they admit power series
expansions
∞
X ∞
X
p(z) = pk z k
and q(z) = q k zk for |z| < ρ,
k=0 k=0
If
∞
X
u(z) = A k zk
k=0
Lu(z) = (2A2 + p0 A1 + q0 A0 )
+ (6A3 + 2p0 A2 + p1 A1 + q0 A1 + q1 A0 )z + · · · ,
X
n−1
(n + 1)nAn+1 + (n − j)pj An−j + qj An−1−j .
j=0
Convergence theorem
−1 X
n−1
An+1 = (n − j)pj An−j + qj An−1−j , n > 1.
n(n + 1)
j=0
Theorem
If the coefficients p(z) and q(z) are analytic for |z| < ρ, then the
formal power series for the solution u(z), constructed above, is
also analytic for |z| < ρ.
Previous example
Earlier we considered
In this case,
X ∞
−5z
p(z) = = −5 z2k+1
1 − z2
k=0
and
X ∞
−4
q(z) = = −4 z2k
1 − z2
k=0
are analytic for |z| < 1, so the theorem guarantees that u(z), given
by the formal power series, is also analytic for |z| < 1.
Expansion about a point other than 0
d2 u du
+ p(Z + c) + q(Z + c)u = 0.
dZ2 dZ
Now compute the Ak using the series expansions of p(Z + c) and
q(Z + c) in powers of Z.
Example
so
A0 Ak−1 + Ak
A2 = and Ak+2 = for k > 1.
2 (k + 2)(k + 1)
We find that
and suppose that a2 (x0 ) = 0 for some x0 with a < x0 < b, but
a2 (x) 6= 0 if x 6= x0 . Put
Noticing that
Lxr = ar(r − 1) + br + c xr ,
Lemma
If r1 = r2 then the general solution of the homogeneous
Cauchy–Euler equation Lu = 0 is
Example
Solve x2 u 00 − xu 0 + u = 0.
Example
Solve 2(x − 2)2 u 00 − 3(x − 2)u 0 − 3u = 0.
Proof of the lemma
Since r1 = r2 the function F(x, r) = xr satisfies
z2 u 00 + zP(z)u 0 + Q(z)u = 0,
z2 u 00 + P0 zu 0 + Q0 u = 0.
Consider
Lu = 2z2 u 00 + 7zu 0 − (z2 + 3)u = 0.
Here, P(z) = 7/2 and Q(z) = −(z2 + 3)/2 are trivially analytic
at z = 0 (since they are polynomials).
2z2 u 00 + 7zu 0 − 3u = 0.
we find that
and
∞
X ∞
X
Ak zk+r+2 = Ak−2 zk+r
k=0 k=2
it follows that
Ak−2
A1 = 0, Ak = for all k > 2.
(2k + 2r − 1)(k + r + 3)
Ak−2
A1 = 0, Ak = for all k > 2,
k(2k + 7)
so
A0 A1 A2 A0
A2 = , A3 = = 0, A4 = = , ...
22 39 60 1320
and
z2 z4
1/2
u(z) = A0 z 1+ + + ··· .
22 1320
Second solution: r = −3 with
Ak−2
A1 = 0, Ak = for all k > 2,
k(2k − 7)
so
A0 A1 A2 A0
A2 = − , A3 = − = 0, A4 = − = , ...
6 3 4 24
and
z2 z4
−3
u(z) = A0 z 1− + + ··· .
6 24
General solution of Lu = 0:
z2 z4
u(z) = Az1/2 1 + + + ···
22 1320
z2 z4
−3
+ Bz 1− + + ··· .
6 24
General case
Now consider
z2 u 00 + zP(z)u 0 + Q(z)u = 0
for P(z) and Q(z) satisfying (6). Formal manipulations show that
Lu(z) equals
∞
X X
k−1
r
I(r)A0 z + I(k + r)Ak + [(j + r)Pk−j + Qk−j ]Aj zk+r ,
k=1 j=0
−1 X
k−1
Ak (r) = [(j + r)Pk−j + Qk−j ]Aj (r), k > 1,
I(k + r)
j=0
z2 F 00 + zP(z)F 0 + Q(z)F = (r − r1 )2 zr .
(3z)2 (3z)3 z2 z4
2 00 0
z u +3 3z+ + +· · · u +15 1+ + +· · · u = 0.
2! 3! 2! 4!
z2 u 00 + zu 0 + (z2 − ν2 )u = 0.
(z/2)2 (z/2)4
ν
u(z) = A0 z 1 − +
1+ν 2(2 + ν)(1 + ν)
(z/2)6
− + ··· .
3!(3 + ν)(2 + ν)(1 + ν)
Bessel function
With the normalisation
1
A0 =
2ν Γ (1 + ν)
and so
∞
X (−1)k (z/2)2k+ν
Jν (z) = .
k!Γ (k + 1 + ν)
k=0
Bessel function of negative order
Also, since 1/Γ (z) = 0 for z = 0, −1, −2, . . . , we find that Jn and
J−n are linearly dependent; in fact,
(ν − 1)(ν + 2) 3
u1 (z) = z − z
3!
(ν − 3)(ν − 1)(ν + 2)(ν + 4) 5
+ z − ··· .
5!
Pn (1) = 1.
Dynamical Systems
Introduction
Stability
Interpretation: if the foxes fail to catch any rabbits, then the foxes
starve (F → 0) and the rabbits multiply without limit (R → ∞).
Interaction terms
Rewrite equations as
1 dF
= −a + αR,
F dt
1 dR
= b − βF.
R dt
So
relative rate of increase in fox popula-
αR =
tion due to predation on rabbits
and
relative rate of decrease in rabbit pop-
βF = .
ulation due to predation by foxes
Periodic solutions
Recall
dF
= −aF + αFR, F(0) = F0 ,
dt
dR
= bR − βFR, R(0) = R0 .
dt
So defining
F −aF + αFR F
x= , F(x) = , x0 = 0
R bR − βFR R0
we have
dx dF/dt −aF + αFR
= = = F(x),
dt dR/dt bR − βFR
with x(0) = x0 .
Geometric viewpoint
Think of the trajectory x(t) as a parametric curve in the phase
N
space R . Then dx/dt = F(x) means that along any trajectory,
F x(t) always points in the forward tangent direction to the
trajectory, with the speed of x(t) equal to the magnitude of F.
Non-autonomous ODEs
d2 x
dx dx
2
= f x, , t , with x = x0 and = y0 at t = 0.
dt dt dt
d2 x
dy dx
= 2 = f x, , t = f(x, y, t),
dt dt dt
dx
= y, x(0) = x0 ,
dt
dy
= f(x, y, t), y(0) = y0 .
dt
Simple oscillator as first-order system
The second-order ODE
mẍ + r0 ẋ + k0 x = f(t)
ẋ = v,
1
v̇ = f(t) − r0 v − k0 x .
m
That is,
dx
= F(x, t),
dt
where
x v
x= and F(x, t) = .
v m−1 f(t) − r0 v − k0 x
Immune response to a viral infection
dE
= Λ − µE,
dt
for constants Λ > 0 (recruitment rate from bone marrow) and
µ > 0 (death rate).
If E0 is the initial density of effector cells, then
Λ
E(t) = E0 e−µt + (1 − e−µt )
µ
so E(t) → Ê ≡ Λ/µ as t → ∞.
Definition
The number L is a Lipschitz constant for a function f : [a, b] → R
if
|f(x) − f(y)| 6 L|x − y| for all x, y ∈ [a, b].
Example
Consider f(x) = 2x2 − x + 1 for 0 6 x 6 1. Since
Theorem
If f is Lipschitz then f is (uniformly) continuous.
Proof.
Suppose L is a Lipschitz constant for f : [a, b] → R. Given > 0,
if δ = /L then
|f(x) − f(y)| 1
L> =√ √
|x − y| x+ y
Definition
A function f : I → R is Ck if f, f 0 , f 00 , . . . , f(k) all exist and are
continous on the interval I.
Theorem
For any closed and bounded interval I = [a, b], if f is C1 on I then
L = maxx∈I |f 0 (x)| is a Lipschitz constant for f on I.
Proof.
Given a 6 x < y 6 b, the Mean Value Theorem says that there
exists a number c (depending on x and y) such that
dx
= f(x) for t > 0, with x(0) = x0 . (8)
dt
If x = x(t) is a solution then
Zt
x(t) = x0 + f x(s) ds. (9)
0
x1 (t) = x0 ,
Zt
x2 (t) = x0 + f x1 (s) ds,
0
Zt
x3 (t) = x0 + f x2 (s) ds,
0
and in general,
Zt
xk (t) = x0 + f xk−1 (s) ds for k > 1. (10)
0
Increments
Subtracting Zt
xk (t) = x0 + f xk−1 (s) ds
0
from Zt
xk+1 (t) = x0 + f xk (s) ds
0
gives
Zt
xk+1 (t) − xk (t) = f xk (s) − f xk−1 (s) ds,
0
We have Zt
δk (t) 6 L δk−1 (s) ds
0
so for 0 6 t 6 T ,
Zt Zt
δ2 (t) 6 L δ1 (s) ds 6 L ∆1 (T ) ds 6 ∆1 (T )Lt,
0 0
Zt Zt
δ3 (t) 6 L δ2 (s) ds 6 L ∆1 (T )Ls ds = ∆1 (T )L2 × 12 t2 ,
0 0
Zt Zt
δ4 (t) 6 L δ3 (s) ds 6 L ∆1 (T )L2 21 s2 ds = ∆1 (T )L3 × 3!1 3
t ,
0 0
..
.
Convergence of the Picard iterates
By induction on k,
(Lt)k−1
δk (t) 6 ∆1 (T ) for |t| 6 T .
(k − 1)!
dx
and therefore = f x(t) with x(0) = x0 .
dt
Lipschitz vector field
Here,
X
N 1/2
kxk = x2j
j=1
Theorem
Let x0 ∈ RN , fix r > 0 and τ > 0, and put
dx
= F(x, t) for |t| 6 min(r/M, τ), with x(0) = x0 .
dt
Proof.
See Technical Proofs handout.
Example of non-uniqueness
The initial-value problem
dx
= 3x2/3 for t > 0, with x(0) = 0,
dt
has infinitely many solutions, namely, for any a > 0,
0, 0 6 t 6 a,
x(t) = 3
(t − a) , t > a.
|f(x)| 6 M = r2 + 2r + 2 for |x − x0 | 6 r,
Therefore, the two solutions trace out the same trajectory in phase
space.
Practical solution methods
tp = p ∆t for p = 0, 1, 2, . . . , P,
so
0 = t0 < t1 < t2 < · · · < tP = T.
x(tp ) ≈ Xp for 1 6 p 6 P.
Finite difference approximation
Since
dx ∆x x(t + ∆t) − x(t)
= lim = lim ,
dt ∆t→0 ∆t ∆t→0 ∆t
if ∆t is small (and thus P is large), then
X1 = X0 + f(X0 ) ∆t,
X2 = X1 + f(X1 ) ∆t,
X3 = X2 + f(X2 ) ∆t,
..
.
XP = XP−1 + f(XP−1 ) ∆t.
Xp = (1 + a ∆t)p x0 .
at∗ p
(1 + a ∆t)p = 1 + → eat∗ ,
p
Definition
We write
φ(y) = O ψ(y) as y → a,
if there exist constants C > 0 and δ > 0 such that
Interpret
φ(y) = γ(y) + O ψ(y)
as
φ(y) − γ(y) = O ψ(y) .
Taylor expansions
Recall that if f is Cm+1 then then
(x − a)2
f(x) = f(a) + f 0 (a)(x − a) + f 00 (a) + ···
2
(x − a)m
+ f(m) (a) + Rm (x),
m!
where the remainder term is
Zx
(x − y)m
Rm (x) = f(m+1) (y) dy
a m!
Theorem
If f is Cm+1 on a neighbourhood of a, then
Rm (x) = O (x − a)m+1
as x → a.
Proof.
There exist constants C and δ such that |f(m+1) (y)| 6 C for
|x − a| 6 δ. If 0 6 x − a 6 δ, then
Zx
(x − y)m C
|Rm (x)| 6 C dy = (x − a)m+1 ,
a m! (m + 1)!
Example
As x → 0,
sin x = x + O(x3 ),
cos x = 1 − 12 x2 + O(x4 ),
log(1 + x) = x − 12 x2 + O(x3 ),
1
= 1 − x2 + O(x4 ).
1 + x2
How accurate is Euler’s method?
Taylor expansion shows that if x(t) is C2 , then as ∆t → 0,
so in particular when t = t0 ,
X1 = X0 + f(X0 ) ∆t
= x(t0 ) + f x(t0 ) ∆t
= x(t1 ) + O(∆t2 ).
Xp = x(tp ) + O(∆t) as ∆t → 0.
Example
How many steps are needed to ensure the error is less than 10−4 if
T = 5, L = 1 and |ẍ(t)| 6 2? Sufficient to satisfy C ∆t 6 10−4 .
T
T eLT 6 10−4 ⇐⇒ P > T 2 eLT × 104 ≈ 37, 103, 290.
P
A more efficient method
d2 x dx
ẍ = = f 0 (x) = f 0 (x)f(x)
dt2 dt
so by Taylor expansion,
Xp = x(tp ) + O(∆t) as ∆t → 0.
Can show in a similar way that for the Taylor method of order 2,
Xp = x(tp ) + O(∆t2 ) as ∆t → 0.
Example
When T = 1 we have ∆t = 1/P, so
P ∆t ∆t2
10 0.1 0.01
100 0.01 0.0001
1000 0.001 0.000001
10000 0.0001 0.00000001
100000 0.00001 0.0000000001
Example
Consider
dx
= cos2 x for t > 0, with x(0) = 0,
dt
which has the solution x = tan−1 (t). Since
for p = 0, 1, 2, . . . , with X0 = 0.
Comparison using P = 8 steps
Systems of ODEs
Φ1 = F(Xp ),
Φ2 = F(Xp + ∆t Φ1 ),
1
Xp+1 = Xp + 2 ∆t (Φ1 + Φ2 ),
we have
Φ1 = F(Xp ),
Φ2 = F(Xp + 12 ∆t Φ1 ),
Φ3 = F(Xp + 12 ∆t Φ2 ),
Φ4 = F(Xp + ∆t Φ3 ),
1
Xp+1 = Xp + 6 ∆t Φ1 + 2Φ2 + 2Φ3 + Φ4 ,
EP ≈ C ∆tr .
EP/2 CT r (P/2)−r
≈ = 2r
EP CT r P−r
and thus
r ≈ log2 (EP/2 /EP ).
Simple test problem
For the system
dx
= 2x + 3y, x(0) = 2,
dt
dy
= −3x + 2y, y(0) = −1,
dt
we obtain the following maximum errors and convergence rates
(taking T = 1).
Theorem
If A(t) and b(t) are continuous for 0 6 t 6 T , then the linear
initial-value problem
dx
= A(t)x + b(t) for 0 6 t 6 T , with x(0) = x0 ,
dt
has a unique solution x(t) for 0 6 t 6 T .
dx
= λeλt v = eλt (λv) = eλt (Av) = A(eλt v) = Ax
dt
that is, x is a solution of dx/dt = Ax.
If Avj = λj vj for 1 6 j 6 N, then the linear combination
X
N
x(t) = cj eλj t vj (11)
j=1
X
N
x(0) = cj vj = x0 .
j=1
Example
Consider
dx
= −5x + 2y, x(0) = 5,
dt
dy
= −6x + 3y, y(0) = 7.
dt
In this case,
−5 2 1 1
A= , λ1 = −3, v1 = , λ2 = 1, v2 = ,
−6 3 1 3
so the general solution is
−3t 1 t 1
x(t) = c1 e + c2 e ,
1 3
and the initial conditions imply c1 = 4 and c2 = 1, so
x = 4e−3t + et ,
−3t 1 t 1
x(t) = 4e +e and
1 3 y = 4e−3t + 3et .
Exponential of a matrix
Recall that the Taylor series
∞
X
z zk z2 z3
e = =1+z+ + + ···
k! 2! 3!
k=0
t2 A 2 tk A k
tA
x(t) = e x0 = I + tA + + ··· + + · · · x0 .
2! k!
Since
d tA tk−1 Ak
e = 0 + A + tA2 + · · · + + ···
dt (k − 1)!
tk−1 Ak−1
= A I + tA + · · · + + · · · = AetA ,
(k − 1)!
we have
dx
= AetA x0 = Ax and x(0) = Ix0 = x0 .
dt
But how can we calculate etA explicitly?
Diagonalising a matrix
Definition
A square matrix A ∈ CN×N is diagonalisable if there exists a
non-singular matrix Q ∈ CN×N such that Q−1 AQ is diagonal.
Theorem
A square matrix A ∈ CN×N is diagonalisable if and only if there
exists a basis {v1 , v2 , . . . , vN } for CN consisting of eigenvectors
of A. Indeed, if
Avj = λj vj for j = 1, 2, . . . , N,
A = QΛQ−1 .
Thus
and
A3 = A2 A = QΛ2 Q−1 QΛQ−1 = QΛ3 Q−1 .
In general, we see by induction on k that
and in general,
λk
1
λk
2
Λk = for k = 0, 1, 2, 3, . . . .
..
.
λk
N
Example
If
−5 2
A=
−6 3
then
−3 0 1 1 −1 1 3 −1
Λ= , Q= , Q =
0 1 1 3 2 −1 1
so
1 1 1 (−3)k 0
k k −1 3 −1
A = QΛ Q =
2 1 3 0 1 −1 1
k k+1 k+1 × 3k + 1
1 (−1) × 3 − 1 (−1)
= .
2 (−1)k × 3k+1 − 3 (−1)k+1 × 3k + 3
Polynomial of a matrix
For any polynomial
p(z) = c0 + c1 z + c2 z2 + · · · + cm zm
p(A) = c0 I + c1 A + c2 A2 + · · · + cm Am .
= Q c0 I + c1 Λ + c2 Λ2 + · · · + cm Λm Q−1
= Qp(Λ)Q−1 .
Lemma
For any polynomial p and any diagonal matrix Λ,
p(λ1 )
p(λ2 )
p(Λ) = .
. .
.
p(λN )
Theorem
If two polynomials p and q are equal on the spectrum of a
diagonalisable matrix A, that is, if
and observe
We find
2 9 −4
p(A) = A − 4I = = −2A − I = q(A).
12 −7
Exponential of a diagonalisable matrix
Theorem
If A = QΛQ−1 is diagonalisable, then
eλ1
eλ2
eA = QeΛ Q−1 and eΛ =
..
.
eλN
Proof.
eA equals
∞
X ∞
X ∞
X ∞
X
Ak QΛk Q−1 Λk Q−1 Λk
= =Q =Q Q−1 .
k! k! k! k!
k=0 k=0 k=0 k=0
Example
Again put
−5 2
A= .
−6 3
We have
1 1 1 e−3 0
A Λ −1 3 −1
e = Qe Q =
2 1 3 0 e1 −1 1
1 3e − e −e−3 + e
−3
= .
2 3e−3 − 3e −e−3 + 3e
Notice tA = Q(tΛ)Q−1 so
1 1 1 e−3t 0
tA tΛ −1 3 −1
e = Qe Q =
2 1 3 0 et −1 1
1 3e−3t − et −e−3t + et
= .
2 3e−3t − 3et −e−3t + 3et
A (maybe) simpler method
The following trick lets you compute eA without finding Q.
then
eA = QeΛ Q−1 = Qp(Λ)Q−1 = p(A).
X
N Y
N
λ − λr
λj
p(λ) = e .
λj − λ r
j=1 r=1
r6=j
A 3 × 3 example
(λ − λ2 )(λ − λ3 ) (λ − λ1 )(λ − λ3 )
p(λ) = eλ1 t + eλ2 t
(λ1 − λ2 )(λ1 − λ3 ) (λ2 − λ1 )(λ2 − λ3 )
(λ − λ1 )(λ − λ2 )
+ eλ3 t
(λ3 − λ2 )(λ3 − λ2 )
(λ − 7)(λ − 8) (λ − 6)(λ − 8)
= e6t + e7t
(6 − 7)(6 − 8) (7 − 6)(7 − 8)
(λ − 6)(λ − 7)
+ e8t
(8 − 6)(8 − 7)
so that
e6t
etA = p(A) = (A − 7I)(A − 8I) − e7t (A − 6I)(A − 8I)
2
e8t
+ (A − 6I)(A − 7I)
2
−1 0 2 −4 2 6 −2 2 4
= e6t 1 0 −2 − e7t 0 0 0 + e8t −1 1 2 ,
−1 0 2 −2 1 3 −1 1 2
which equals
6t
−e + 4e7t − 2e8t −2e7t + 2e8t 2e6t − 6e7t + 4e8t
e6t − e8t e8t −2e6t + 2e8t .
6t 7t
−e + 2e − e 8t 7t
−e + e 8t 2e − 3e7t + 2e8t
6t
Digression: higher-order, linear, scalar equations
X
m
Lu = aj (x)Dj u.
j=0
Theorem
Assume that the ODE Lu = f is not singular with respect to [a, b],
and that f is continuous on [a, b]. Then the IVP (2) and (3) has a
unique solution.
Equivalent first-order system
Sufficient to look at the case m = 3,
Lu = a3 u 000 + a2 u 00 + a1 u 0 + a0 u.
Can now see why to expect trouble if L is singular, that is, if the
leading coefficient a3 vanishes at any x in the interval of interest.
Equivalent first-order system: matrix version
We can write Lu = f as
dy
= A(x)y + b(x),
dx
where, when m = 3,
0 1 0
A(x) = 0 0 1 ,
−a0 /a3 −a1 /a3 −a2 /a3
u 0
y(x) = u 0 , b(x) = 0 .
u 00 f/a3
Existence and uniqueness
Theorem
Assume that L is not singular on [a, b] and that f is continuous
on [a, b]. Then the IVP (2) and (3) has a unique solution.
Proof.
For m = 3, the IVP (2) and (3) is equivalent to
ν0
dy
= A(x)y + b(x), y(0) = ν1
dx
ν2
where y1 = u, y2 = u 0 , y3 = u 00 and
0 1 0 0
A(x) = 0 0 1 , b(x) = 0 .
−a0 /a3 −a1 /a3 −a2 /a3 f/a3
Stability
Λ − µE + VE = 0,
rV − σVE = 0.
Definition
An equilibrium point a is stable if for every > 0 there exists
δ > 0 such that whenever kx0 − ak < δ the solution of
dx
= F(x) for t > 0, with x(0) = x0
dt
satisfies
kx(t) − ak < for all t > 0.
(In particular, x(t) must exist for all t > 0.)
Definition
Let D be an open subset of RN that contains an equilibrium
point a. We say that a is asymptotically stable in D if a is stable
and, whenever x0 ∈ D, the solution of
dx
= F(x) for t > 0, with x(0) = x0
dt
satisfies
x(t) → a as t → ∞.
In this case D is called a domain of attraction for a.
Linear, constant-coefficient case
Consider
dx
= Ax + b with x(0) = x0 . (13)
dt
Since
Ax + b = 0 ⇐⇒ x = −A−1 b,
the only equilibrium point is a = −A−1 b. Moreover,
because
dx
= AetA (x0 − a) = A(x − a) = Ax − Aa = Ax + b
dt
and
x(0) = a + I(x0 − a) = x0 .
Criteria for stability
Theorem
Let A be a diagonalisable matrix with eigenvalues λ1 , λ2 , . . . , λN .
The equilibrium point a = −A−1 b is of (13)
1. stable if and only Re λj 6 0 for all j.
2. asymptotically stable if and only Re λj < 0 for all j.
In the second case, the domain of attraction is the whole of RN .
Proofs
X
N
2
X
N
ke tΛ 2
wk = e λj t
wj 6 |wj |2 = kwk2 ,
j=1 j=1
and thus
tA tΛ −1 cos t − 2 sin t 5 sin t
e = Qe Q = .
− sin t cos t − 2 sin t
Example: asymptotically stable
Consider
dx 14 −9
= Ax where A= .
dt 30 −19
Eigenvalues λ1 = −1 and λ2 = −4 so Re λ1 < 0 and Re λ2 < 0,
hence asymptotically stable.
Eigenvalue decomposition A = QΛQ−1 where
−1 0 3 1
Λ= , Q = [v1 v2 ] = ,
0 −4 5 2
2 −1
Q−1 = ,
−5 3
and thus
6e−t − 5e−4t −3e−t + 3e−4t
tA tΛ −1
e = Qe Q = .
10e−t − 10e−4t −5e−t + 6e−4t
Example: unstable
Consider
dx −26 36
= Ax where A= .
dt −18 25
Eigenvalues λ1 = 1 and λ2 = −2 so Re λ1 > 0 and Re λ2 < 0,
hence unstable.
Eigenvalue decomposition A = QΛQ−1 where
1 0 4 3
Λ= , Q = [v1 v2 ] = ,
0 −2 3 2
−2 3
Q−1 = ,
3 −4
and thus
−8et + 9e−2t 12et − 12e−2t
tA tΛ −1
e = Qe Q = .
−6et + 6e−2t 9et − 8e−2t
Linearization
Suppose that x0 is close to an equilibrium point a. If
dx
= F(x) for all t, with x(0) = x0 , (14)
dt
then for small t the difference y = x − a is small and satisfies
dy dx
= = F(x) = F(a + y) ≈ F(a) + F 0 (a)y.
dt dt
Recall that,
dE/dt Λ − µE + VE
= F(E, V) = .
dV/dt rV − σVE
Thus,
0 −µ + V E
F (E, V) = ,
−σV r − σE
so at the first equilibrium point,
Λ 0 −µ Ê
E= = Ê, V = 0, F (Λ/µ, 0) = .
µ 0 r − σÊ
(θ, φ) = (nπ, 0)
λ −1
det(λI − A) = = λ2 + rλ + (−1)n k = 0.
(−1)n k λ + r
When n is even, we solve λ2 + rλ + k = 0 to obtain
√ √
1
2
2 √−r ± i 4k − r , 0 6 r < 2 k,
√
λ = λ± = − k, r = 2 k,
1 √ √
−r ± r2 − 4k , r > 2 k.
2
r=0√ Re λ+ = Re λ− = 0 stable
n even 0 < r 6√2 k Re λ+ = Re λ− < 0 asymptot. stable
r>2 k λ− < λ+ < 0 asymptot. stable
n odd r>0 λ− < 0 < λ+ unstable
dx
= F(x)
dt
if G x(t) is constant for every solution x(t).
By the chain rule,
d X ∂G dxj
N
dx
G x(t) = = ∇G(x) · = ∇G(x) · F(x).
dt ∂xj dt dt
j=1
Consider
d x ẋ
ẍ = f(x) or equivalently = , (15)
dt ẋ f(x)
d ẋ2 d dV dx
= ẋ ẍ and V(x) = = −f(x)ẋ,
dt 2 dt dx dt
if x = x(t) is a solution of (15) then
d 2
ẋ /2 + V(x) = ẍ − f(x) ẋ = 0,
dt
so the function G(x, ẋ) = 12 ẋ2 + V(x) is a first integral.
Example: undamped pendulum
The angular deflection of an undamped pendulum satisfies
or r
2 k
θ̈ + ω sin θ = 0, ω= ,
m
which has the form
dV
θ̈ = f(θ) = −
dθ
with f(θ) = −ω2 sin θ and V(θ) = −ω2 cos θ. Thus,
d 1 2
− ω2 cos θ = θ̇θ̈ + ω2 θ̇ sin θ = θ̇ θ̈ + ω2 sin θ = 0
dt 2 θ̇
We necessarily have k 6 N − 1.
http://www.youtube.com/watch?v=FYE4JKAXSfY
Part III
Heat equation
Two-point boundary value problems
Example
No solution exists satisfying
Example
Infinitely many solutions u(x) = C sin x exist satisfying
where
Lu = a2 u 00 + a1 u 0 + a0 u
is a 2nd-order linear differential operator, and the boundary
functionals have the form
Example
B1 (c1 u1 + c2 u2 + uP ) = α1 ,
B2 (c1 u1 + c2 u2 + uP ) = α2 ,
Example
Later we will show that the Legendre polynomials are orthogonal
over the interval (−1, 1):
Z1
Pn (x)Pk (x) dx = 0 if n 6= k.
−1
Key properties
The inner product and norm for functions behave like the dot
product and norm for vectors in Rn .
Proof.
If f = 0 or g = 0 then hf, gi = 0 and kfkkgk = 0 so the result
reduces to 0 6 0.
Otherwise, f 6= 0 and g 6= 0 so kfk =
6 0 and kgk =
6 0, and we may
define α = 1/kfk and β = 1/kgk. Since
Corollary
kf + gk 6 kfk + kgk.
Proof.
kf + gk2 = hf + g, f + gi
= kfk2 + 2hf, gi + kgk2
6 kfk2 + 2kfkkgk + kgk2
= (kfk + kgk)2 .
Self-adjoint differential operators
Lu = a2 u 00 + a1 u 0 + a0 u.
Integrating by parts,
Zb Zb
0
b
u(a1 v) 0 dx,
(a1 u )v dx = u(a1 v) a
−
a a
Zb Zb
00 0 b
0
u(a2 v) 00 dx,
(a2 u )v dx = u (a2 v) − u(a2 v) a +
a a
so
b
hLu, vi = u 0 (a2 v) − u(a2 v) 0 + u(a1 v) a
+ u, (a2 v) 00 − (a1 v) 0 + a0 v .
Adjoint operator and Lagrange identity
Thus, defining the formal adjoint
L∗ v = (a2 v) 00 − (a1 v) 0 + a0 v
= a2 v 00 + (2a20 − a1 )v 0 + (a200 − a10 + a0 )v
and
X
m X
j
P(u, v) = (−1)k−1 (Dj−k u)Dk−1 (aj v).
j=1 k=1
Formal self-adjointness
The operator L is formally self-adjoint if L∗ = L.
Theorem
A second-order, linear differential operator L is formally self-adjoint
iff it can be written in the form
P(u, v) = −p(x)(u 0 v − uv 0 ).
Proof
If L = −(pu 0 ) 0 + qu = −pu 00 − p 0 u 0 + qu then
L∗ v = (−pv) 00 + (p 0 v) 0 + qv
= −(p 00 v + 2p 0 v 0 + pv 00 ) + (p 00 v + p 0 v 0 ) + qv
= −(pv 00 + p 0 v 0 ) + qv = −(pv 0 ) 0 + qv = Lv,
a2 u 00 + a1 u 0 + a0 u
= a2 u 00 + (2a20 − a1 )u 0 + (a200 − a10 + a0 )u
x2 u 00 + xu 0 + (x2 − ν2 )u = f(x).
Example
The Legendre equation
Lemma
Any formally self-adjoint operator L = −(pu 0 ) 0 + qu satisfies the
identity
X
2
hLu, vi − hu, Lvi = Bi u Ri v − Ri u Bi v , (22)
i=1
and at x = b,
0 0 0
pv
−p(u v − uv ) = b21 u + b20 u −
b21
pu
b21 v 0 + b20 v
− −
b21
= (B2 u)(R2 v) − (R2 u)(B2 v).
hf, vi = α1 R1 v + α2 R2 v. (24)
Sufficient condition for existence?
It can be shown that, provided a2 (x) 6= 0 for a 6 x 6 b, the
condition (24) is also sufficient for the existence of u, i.e., u exists
iff (24) holds for all v satisfying (23).
Example
The 2-point BVP
Theorem
Either the homogeneous problem (18) has only the trivial
solution v ≡ 0, in which case
the inhomogeneous problem (17) has a unique solution u
for every choice of f, α1 and α2 ,
or else the homogeneous problem admits non-trivial solutions, in
which case
the inhomogeneous problem (17) has a solution u iff f,
α1 and α2 satisfy (24) for every solution v of the
homogeneous problem (18).
Definition
Let D be a subspace of a vector space V with inner product h·, ·i,
and let L be a linear operator on V with domain D. We say that L
is self-adjoint if
B1 u = 0 at x = a,
B2 u = 0 at x = b.
Example: Bessel operator
and let
B1 u = u 0 , R1 u = u, at x = 1,
0
B2 u = 2u − u, R2 u = −u, at x = 2,
R2
so that with hf, gi = 1 f(x)g(x) dx the Lagrange identity implies
Let
longitudinal momentum = 0,
∂u
transverse momentum = ρ∆x .
∂t
No motion in the longitudinal direction so
T cos θ = T0
∂2 u ∂ ∂
ρ 2
= (T sin θ) = T0 (tan θ).
∂t ∂x ∂x
But tan θ = ∂u/∂x so u satisfies the wave equation
∂2 u ∂2 u
ρ = T 0 .
∂t2 ∂x2
Initial-boundary value problem (IBVP)
p
Put c = T0 /ρ and suppose that the string is initially at rest with
a known deflection u0 (x), then
∂2 u 2
2 ∂ u
− c = 0, 0 < x < `, t > 0,
∂t2 ∂x2
u(0, t) = 0, t > 0,
u(`, t) = 0, t > 0, (27)
XT 00 − c2 X 00 T = 0.
Thus, we want
T 00 X 00
= .
c2 T X
Since T 00 /(c2 T ) is a function of t only, and X 00 /X is a function of
x only, there must be a separation constant λ such that
T 00 X 00
= −λ = .
c2 T X
A homogeneous BVP
Since
B sin ω` = 0.
If B = 0, then X ≡ 0.
If B 6= 0, then sin ω` = 0 so ω` = nπ for some n ∈ {1, 2, 3, . . . }.
In this case, ω = nπ/` so
2
nπ nπx
λ= and X = B sin .
` `
Case λ = 0
X = Ax + B.
A` = 0.
B sinh ω` = 0.
If B = 0, then X ≡ 0.
∂2 un 2
2 ∂ un
− c = 0,
∂t2 ∂x2
un (0, t) = 0,
(28)
un (`, t) = 0,
∂un
(x, 0) = 0,
∂t
but since
nπ
un (x, 0) = sin x (29)
`
the inhomogeneous initial condition will not be satisfied unless
u0 (x) happens to be sin(nπx/`).
Superposition principle
Each equation in (28) is linear and homogeneous, so any
superposition of normal modes,
∞
X
nπ nπc
u(x, t) = An sin x cos t , (30)
` `
n=1
u(x, 0) = u0 (x), 0 6 x 6 `.
Example
If ` = 1 and the initial deflection is given by
2x, 0 < x < 21 ,
u0 (x) =
2(1 − x), 12 < x < 1,
we find that
∞
8 X (−1)k−1
u(x, t) = 2 2
sin (2k − 1)πx cos (2k − 1)πct .
π (2k − 1)
k=1
Partial sums
Heat equation
K = thermal conductivity,
ρ = mass density,
σ = specific heat,
Fourier’s law of heat conduction asserts that the heat flux q in the
x-direction is given by
∂u
q = −K
∂x
Conservation law for thermal energy
In a (fixed) region x1 < x < x2 , the thermal energy per unit of
cross sectional area equals
Z x2
ρσu dx.
x1
The heat flux entering the region at x = x1 equals q(x1 ), and the
heat flux entering the region at x = x2 equals −q(x2 ).
The rate of change of thermal energy must equal the total heat
flux entering the region, so
Z
d x2
ρσu dx = q(x1 ) − q(x2 ),
dt x1
Z x2 Z x2
∂u ∂q
ρσ dx = − dx.
x1 ∂t x1 ∂x
Heat equation
Z x2
∂u ∂q
ρσ + dx = 0.
x1 ∂t ∂x
By Fourier’s law, q = −K∂u/∂x, so
Z x2
∂2 u
1 ∂u
ρσ − K 2 dx = 0.
x2 − x1 x1 ∂t ∂x
The LHS equals the average value of the integrand over the
interval (x1 , x2 ), which tends to the pointwise value at x = x1 if
we let x2 → x1 . Since the choice of x1 is arbitrary, we conclude
that u must satisfy the heat equation,
∂u ∂2 u
ρσ − K 2 = 0,
∂t ∂x
or equivalently,
K
ut − auxx = 0, where a = .
ρσ
Part IV
Sturm–Liouville problems
Fourier–Bessel series
Schrödinger equation
Complete orthogonal systems
If w : (a, b) → R satisfies
T0 (x) = 1, T1 (x) = x,
Tn+1 (x) = 2xTn (x) − Tn−1 (x) for n = 1, 2, 3, . . . .
Tn (cos θ) = cos(nθ).
Lemma
If S is orthogonal then S is linearly independent.
Proof
Suppose that φ1 , φ2 , . . . , φN ∈ S satisfy
X
N
Cj φj (x) = 0 for a 6 x 6 b.
j=1
On the one hand,
X
N
φk , Cj φj = hφk , 0iw = 0 for 1 6 k 6 N,
j=1 w
Lemma
If {φ1 , . . . , φN } is orthogonal then, for any C1 , . . . , CN ∈ R,
X
N 2 X
N
Cj φj = C2j kφj k2w .
j=1 w j=1
Proof.
X
N X
N X
N X
N
LHS = Cj φj , Ck φk = Cj Ck hφj , φk iw
j=1 k=1 w j=1 k=1
X X X
N
= + = C2j hφj , φj iw + 0 = RHS.
j=k j6=k j=1
Generalised Fourier coefficients
Lemma
If f is in the span of an orthogonal set of functions
{φ1 , φ2 , . . . , φN } in L2 (a, b, w), then the coefficients in the
representation
XN
f(x) = Aj φj (x)
j=1
are given by
hf, φj iw
Aj = for 1 6 j 6 N. (33)
kφj k2w
Since
X
N
f(x) = Aj φj (x),
j=1
if 1 6 k 6 N then
X
N X
N
hf, φk iw = Aj φj , φk = Aj hφj , φk iw
j=1 w j=1
X
= Ak hφk , φk iw + = Ak kφk k2w + 0,
j6=k
so
hf, φk iw
Ak = .
kφk k2w
Classical Fourier sine coefficients
The sequence of functions
nπx
φn (x) = sin , n = 1, 2, 3, . . . ,
`
is orthogonal with respect to the weight w(x) = 1 over the
interval (a, b) = (0, `), that is,
Z`
nπx jπx
hφn , φj i = sin sin dx = 0 if n 6= j.
0 ` `
Since Z`
2 nπx `
kφn k = sin2 dx = ,
0 ` 2
the Fourier (sine) coefficients of a function f(x) are
Z`
hf, φj i 2 jπx
Aj = 2
= f(x) sin dx.
kφj k ` 0 `
Least-squares approximation
X
N
f(x) ≈ Cj φj (x) for a 6 x 6 b.
j=1
X
N 2 Z b X
N 2
f− Cj φj = f(x) − Cj φj (x) w(x) dx
j=1 w a j=1
is as small as possible.
Least-squares approximation
Lemma
For all C1 , C2 , . . . , CN , the weighted mean-square error satisfies
X
N 2 X
N
f− Cj φj = kfk2w − A2j kφj k2w
j=1 w j=1
X
N
+ (Cj − Aj )2 kφj k2w ,
j=1
X
N 2 X
N X
N
f− Cj φj = f− Cj φj , f − Ck φk
j=1 w j=1 k=1 w
X N X
N
= hf, fiw − f, Ck φk − Cj φj , f
k=1 w j=1 w
X
N X
N
+ Cj φj , Ck φk
j=1 k=1 w
X
N X
N
= kfk2w − Ck hf, φk iw − Cj hφj , fiw
k=1 j=1
X
N X
N
+ Cj Ck hφj , φk iw .
j=1 k=1
Since
X
N X
N X X X
N
Cj Ck hφj , φk iw = + = C2j kφj k2w + 0
j=1 k=1 j=k j6=k j=1
X
N 2 X
N X
N
f− Cj φj = kfk2w − 2 Cj hf, φj iw + C2j kφj k2w
j=1 w j=1 j=1
X
N X
N
= kfk2w − 2 Cj Aj kφj k2w + C2j kφj k2w
j=1 j=1
X
N
= kfk22 + C2j − 2Cj Aj + A2j − A2j kφj k2w
j=1
X
N X
N
2
= kfk22 − A2j kφj k2w + Cj − Aj kφj k2w .
j=1 j=1
Conclusion: Let f ∈ L2 (a, b, w) and put
hf, φj iw
Aj = .
kφj k2w
X
N
f(x) = Aj φj (x), a 6 x 6 b,
j=1
X
N
f(x) ≈ Aj φj (x), a 6 x 6 b,
j=1
We will see later that these functions are orthgonal with respect to
Z1
hf, gi = f(x)g(x) dx,
−1
that is,
hPn , Pj i = 0 if n 6= j.
It can also be shown that
Z1
2
kPn k2 = Pn (x)2 dx = .
−1 2n + 1
Example of polynomial least-squares approximation
Let
1, 0 < x < 1/2,
f(x) =
0, −1 < x < 0 or 1/2 < x < 1.
We find that
j 0 1 2 3 4
hf, Pj i 1/2 1/8 −3/16 −19/128 15/256
kPj k2 2 2/3 2/5 2/7 2/9
Aj 1/4 3/16 −15/32 −133/256 135/512
so
X
4 X
4
hf, Pj i
f(x) ≈ Aj Pj (x) = Pj (x)
kPj k2
j=0 j=0
1 3 15 133 135
= P0 (x) + P1 (x) − P2 (x) − P3 (x) + P4 (x).
4 16 32 256 512
Orthogonal sequences of functions
hf, φj iw
Aj = for all j > 1.
kφj k2w
Example
Fourier sine series: for f ∈ L2 (0, π),
∞
X Zπ
2
f(x) ∼ Aj sin jx, Aj = f(x) sin jx dx.
π 0
j=1
Example
Fourier–Legendre series: for f ∈ L2 (−1, 1),
∞
X Z1
2j + 1
f(x) ∼ Aj Pj (x), Aj = f(x)Pj (x) dx.
2 −1
j=0
Example
Fourier–Tchebyshev series: recall that
are orthogonal
√ over (−1, 1) with respect to the weigth function
w(x) = 1/ 1 − x2 . The substitution x = cos θ implies that
Z1
2 Tj (x)2 dx π, j = 0,
kTj kw = √ =
−1 1 − x2 π/2, j > 1,
1, cos x, sin x, cos 2x, sin 2x, cos 3x, sin 3x, ...
so if f ∈ L2 (−π, π) then
∞
A0 X
f(x) ∼
+ Aj cos jx + Bj sin jx
2
j=1
where
Zπ Zπ
1 1
Aj = f(x) cos jx dx and Bj = f(x) sin jx dx.
π −π π −π
Bessel inequality
For every N,
X
N 2 X
N
06 f− Aj φj = kfk2w − A2j kφj k2w
j=1 w j=1
so
X
N
A2j kφj k2w 6 kfk2w ,
j=1
Example
The Fourier–Legendre coefficients of f ∈ L2 (−1, 1) satisfy
∞
X Z1
A2j 1
6 f(x)2 dx.
2j + 1 2 −1
j=0
Example
√
For w(x) = 1/ 1 − x2 , the Fourier–Tchebyshev coefficients
of f ∈ L2 (−1, 1, w) satisfy
∞ Z1
A20 X 2 1 f(x)2 dx
+ Aj 6 √ .
2 π −1 1 − x2
j=1
Example
The trigonometric Fourier coefficients of f ∈ L2 (−π, π) satisfy
∞ Zπ
A20 X 2 1
+ Aj + B2j 6 f(x)2 dx.
2 π −π
j=1
Completeness
Example
The set S = { sin jx : j > 1 and j 6= 7 } is not complete in L2 (0, π)
because sin 7x is orthogonal to every function in S.
Equivalent definitions of completeness
Theorem
If S = {φ1 , φ2 , . . .} is orthogonal in L2 (a, b, w), then the following
properties are equivalent:
1. S is complete;
2. for each f ∈ L2 (a, b, w), if Aj denotes the jth Fourier
coefficient of f then
X
N
f− Aj φj →0 as N → ∞;
j=1 w
X
N 2 X
N
f− A j φj = kfk2w − A2j kφj k2w ,
j=1 w j=1
so
LHS → 0 iff RHS → 0.
Now observe that RHS → 0 means the same thing as
∞
X
kfk2w = A2j kφj k2w .
j=1
Proof that 3 =⇒ 1
Since
∞
X
hg, φj iw = hf, φj iw − Ak hφj , φk iw
k=1
= hf, φj iw − Aj kφj k2w = 0
Corollary
If S = {φ1 , φ2 , φ3 , . . .} is a complete orthogonal sequence
in L2 (a, b, w), then for any f ∈ L2 (a, b, w),
∞
X
keN k2 = Aj kφj k2w .
j=N+1
Proof
We saw in the proof of the theorem (2 ⇐⇒ 3) that
X
N
keN k2w = kfk2w − A2j kφj k2w .
j=1
and therefore
∞
X X
N
keN k2w = A2j kφj k2w − A2j kφj k2w
j=1 j=1
∞
X
= A2j kφj k2w .
j=N+1
Parseval identity
Each of our four examples of orthgonal sequences is in fact
complete.
Example
The Fourier sine coefficients of f ∈ L2 (0, π) satisfy
∞
X Zπ
2
A2j = f(x)2 dx.
π 0
j=1
Example
The Fourier–Legendre coefficients of f ∈ L2 (−1, 1) satisfy
∞
X Z1
A2j 1
= f(x)2 dx.
2j + 1 2 −1
j=0
Example
√
For w(x) = 1/ 1 − x2 , the Fourier–Tchebyshev coefficients
of f ∈ L2 (−1, 1, w) satisfy
∞ Z1
A20 X 2 1 f(x)2 dx
+ Aj = √ .
2 π −1 1 − x2
j=1
Example
The trigonometric Fourier coefficients of f ∈ L2 (−π, π) satisfy
∞ Zπ
A20 X 2 1
+ Aj + B2j = f(x)2 dx.
2 π −π
j=1
Estimating the least-squares error
Consider
f(x) = x for 0 < x < π.
This function has the Fourier sine series
∞
X (−1)n+1
f(x) ∼ 2 sin nx
n
n=1
so
X
N
(−1)n+1
f(x) = 2 sin nx + eN (x),
n
n=1
Example
Legendre’s equation
is equivalent to
and
Lv = µrv on (a, b), with B1 v = 0 = B2 v,
and if λ 6= µ, then u and v are orthogonal on the interval (a, b)
with respect to the weight function r(x), i.e.,
Zb
hu, vir = u(x)v(x)r(x) dx = 0.
a
Proof
Theorem
Let L be a Sturm–Liouville differential operator (35). If
u : [a, b] → C is not identically zero and satisfies
then λ is real.
In the same way, we can prove that for a real, symmetric (or
complex, Hermitian) matrix:
I eigenvectors with distinct eigenvalues are orthogonal;
I every eigenvalue is real.
Proof
implying u ≡ 0, a contradiction.
We conclude λ = λ̄, that is, λ is real.
Completeness of the eigenfunctions
Theorem
The regular Sturm–Liouville problem (36) has an infinite sequence
of eigenfunctions φ1 , φ2 , φ3 , . . . with corresponding eigenvalues
λ1 , λ2 , λ3 , . . . and moreover:
1. the eigenfunctions φ1 , φ2 , φ3 , . . . form a complete
orthogonal system on the interval (a, b) with respect to the
weight function r(x);
2. the eigenvalues satisfy λ1 < λ2 < λ3 < · · · with λj → ∞ as
j → ∞.
In the same way, for every real symmetric (or complex, Hermitian)
n × n matrix A there is an orthgonal basis for Rn consisting of
eigenvectors of A.
Real trigonometric Fourier series
The various trigonometric Fourier series arise from the the
Sturm–Liouville ODE
u 00 + λu = 0.
For the interval [−`, `], periodic boundary conditions
u(0) = u(`) = 0
lead to
2
jπx jπ
φj (x) = sin and λj = for j = 1, 2, 3, . . . ,
` `
whereas homogeneous Neumann boundary conditions
u 0 (0) = u 0 (`) = 0
lead to
2
jπx jπ
φj (x) = cos and λj = for j = 0, 1, 2, . . . .
` `
Least-squares error for cosine expansion
Consider once again the function
This time, instead of the sine series use the cosine series
∞
π X 2
f(x) ∼ + (cos nπ − 1) cos nx
2 πn2
n=1
∞
X
π 4 cos(2k − 1)x
∼ − ,
2 π (2k − 1)2
k=1
π X 2
N
f(x) = + (cos nπ − 1) cos nx + eN (x).
2 πn2
n=1
With N = 2K − 1,
X∞ 2
2 2 π
keN k = 2
(cos nπ − 1)
πn 2
n=N+1 | {z } |{z}
A2n kφn k2
∞
X ∞
4 π 1 X 1
= =
k=K
2
π (2k − 1) 24 8π
k=K
(k − 12 )4
Z∞
1 dt 1 1 2 1
6 1
= 1
= .
8π K (t − 2 ) 4 24π (K − 2 ) 3 3π N3
Thus, for this choice of f, the root mean square error using the
cosine expansion is
keN k = O(N−3/2 ),
x2 u 00 + xu 0 + (k2 x2 − ν2 )u = 0,
(xu 0 ) 0 + k2 x − ν2 x−1 u = 0,
i.e.,
ν2
p(x) = x, r(x) = x, λ = k2 , q(x) = .
x
On the interval [0, `], we do not have a regular Sturm–Liouville
problem because
p(0) = 0
and q(x) is unbounded as x → 0+ .
A condition ensuring λ > 0
Lemma
Consider the ODE
u 0 (`) u 0 (`) c1 0 2
−u 0 (`)u(`) = −c0 u(`) = c1 u 0 (`) = u (`) ,
c0 c0 c0
if c0 6= 0, and
u(`) u(`) c0 2
−u 0 (`)u(`) = −c1 u 0 (`)
= c0 u(`) = u(`) ,
c1 c1 c1
Theorem
Assume that c0 c1 > 0. If c0 6= 0 or ν > 0, then the eigenvalues
λ1 , λ2 , . . . and eigenfunctions φ1 , φ2 , . . . of the Bessel equation
(37) with boundary conditions (38) are given by
λ0 = 0 and φ0 (x) = 1.
Proof
x2 u 00 + xu 0 + (k2 x2 − ν2 )u = 0
which has the general solution u = AJν (kx) + BYν (kx). Since
Yν (kx) is unbounded as x → 0, we must have B = 0, so
u = AJν (kx) and u 0 = AkJν0 (kx). Thus, at x = `,
If c1 6= 0, then
Z` 2
2 1 `c0 2 2
φj (x) x dx = 2 + (kj `) − ν Jν (kj `)2 for j > 1,
0 2kj c1
with
Z` Z`
Aj = f(x)Jν (kj x)x dx Jν (kj x)2 x dx
0 0
Z`
2
= f(x)Jν (kj x)x dx.
`2 Jν+1 (kj `)2 0
Plots when ` = 1, ν = 0, c1 = 0
Partial sums
Another singular Sturm–Liouville problem
Consider the Legendre equation with parameter ν,
i.e.,
ν(ν + 1) = µ(µ + 1) if µ = −ν − 1.
Eigenvalues are positive
[(1 − x2 )u 0 ] 0 + λu = 0 (40)
for n = 0, 1, 2, 3, . . . .
Theorem
The Legendre polynomials are orthogonal over the interval [−1, 1],
Z1
hPn , Pm i = Pn (x)Pm (x) dx = 0 if m 6= n,
−1
Proof
∂ψ h2 ∂2 ψ
ih + − V(x, t)ψ = 0. (42)
∂t 2m ∂x2
Here, m is the mass of the particle, V(x, t) is the potential of the
applied force and h = h/(2π) where
is Planck’s constant.
Separable solutions
Suppose that V does not depend on t, and look for separable
solutions
ψ(x, t) = u(x)T (t).
We find that
ihT 0 h2 u 00
=E=− + V(x)
T 2m u
where the separation constant E has the physical dimensions of
energy.
Thus,
iE
T0 = − T
h
and we may take T (t) = e−iEt/h , so that
h2 00
u + E − V(x) u = 0, (43)
2m
which has the form of a Sturm–Liouville ODE (34) with
h2
p(x) = , λ = E, r(x) = 1, q(x) = V(x).
2m
h2 00
u + E − 12 mω2 x2 u = 0.
(45)
2m
To simplify the coefficients, let
and obtain
E 1
v 00 + (2α + 1 − y2 )v = 0, α= − .
hω 2
Series solutions
Next we put
2 /2
v(y) = e−y w(y)
and find that w satisfies Hermite’s equation
w 00 − 2yw 0 + 2αw = 0,
we see that
Z∞
2
Hn (y)Hk (y)e−y dy = 0, n 6= k.
−∞
Normalisation
It can be shown that
Z∞
2 √
Hn (y)2 e−y dy = 2n n! π,
−∞
so
Z∞ Z∞ 1/2
2
−y2 /2
2 h
φn (x) dx = Cn e Hn (y) dy
−∞ −∞ mω
1/2 Z ∞
h 2
= C2n Hn (y)2 e−y dy
mω −∞
1/2
√
h
= C2n 2n n! π
mω
Elliptic eigenproblems
X
d
div F = ∇ · F = ∂j Fj = ∂1 F1 + ∂2 F2 + · · · + ∂d Fd .
j=1
Second-order linear PDEs in Rd
The most general second-order linear partial differential operator
in Rd has the form
X
d X
d X
d
Lu = − ajk (x)∂j ∂k u + bk (x)∂k u + c(x)u. (47)
j=1 k=1 k=1
Example
The Laplacian is defined by ∇2 u = ∇ · (∇u) = div(grad u), that is,
X
d
2
∇ u= ∂2j u = ∂21 u + ∂22 u + · · · + ∂2d u.
j=1
X
d X
d
Lu = − div(A grad u) = − ∂j ajk (x)∂k u,
j=1 k=1
X
d
bk (x) = − ∂j ajk (x) and c(x) = 0.
j=1
Ellipticity
Definition
A second-order linear partial differential operator (47) is
(uniformly) elliptic in a subset Ω ⊆ Rd if there exists a positive
constant c such that
X
d X
d X
d
ajk (x)ξj ξk > c ξ2k .
j=1 k=1 k=1
Example
The operator L = −∇2 is elliptic (with c = 1) on any Ω ⊆ Rd ,
since
Xd Xd Xd
δjk ξj ξk = ξ2k = kξk2 .
j=1 k=1 k=1
Example
The operator L = −(∂21 + 2∂22 − ∂23 ) is not elliptic in R3 since in
this case the quadratic form
1 0 0 ξ1
ξT Aξ = [ξ1 ξ2 ξ3 ] 0 2 0 ξ2 = ξ21 + 2ξ22 − ξ23
0 0 −1 ξ3
is negative if ξ1 = ξ2 = 0 and ξ3 6= 0.
Symmetry and skew-symmetry
Put
1
asy
jk = ajk + akj = symmetric part of ajk ,
2
1
ask
jk = ajk − akj = skew-symmetric part of ajk ,
2
so that
ajk = asy sk
jk + ajk , asy sy
kj = ajk , ask sk
kj = −ajk
Lemma
X
d X
d X
d X
d
ajk (x)ξj ξk = asy
jk (x)ξj ξk .
j=1 k=1 j=1 k=1
Proof
d X
X d X
d X
d
ask −ask
jk (x)ξj ξk = kj (x) ξj ξk
j=1 k=1 j=1 k=1
X
d X
d
=− ask
jk (x)ξk ξj
k=1 j=1
X
d X
d
=− ask
jk (x)ξk ξj
j=1 k=1
X
d X
d
=− ask
jk (x)ξj ξk
j=1 k=1
so
X
d X
d
2 ask
jk (x)ξj ξk = 0.
j=1 k=1
Theorem
Denote the eigenvalues of the real symmetric matrix [asy jk (x)]
by λj (x) for 1 6 j 6 d. The operator (47) is elliptic on Ω if and
only if there exists a positive constant c such that
Example
Show that the L = −(3∂21 + 2∂1 ∂2 + 2∂22 ) is elliptic.
Example
Show that L = −(∂21 − 4∂1 ∂2 + ∂22 ) is not elliptic.
Proof
Fix x, then the real symmetric matrix A = [asy jk ] is diagonalisable,
d
that is, R has an orthonormal basis of eigenvectors v1 , . . . , vd
and so
QT AQ = diag(λ1 , λ2 , . . . , λd )
where the matrix Q = [v1 v2 · · · vd ] is orthogonal:
QT = Q−1 . Put η = QT ξ, then ξ = Qη so
X
d
ξT Aξ = (Qη)T A(Qη) = ηT (QT AQ)η = λj η2j ,
j=1
X
d
ξT Aξ > ckξk2 ⇐⇒ λj η2j > ckηk2 .
j=1
Three important PDEs
−∇2 u = f.
∂u
− ∇2 u = f.
∂t
3. Wave equation (hyperbolic):
∂2 u
− ∇2 u = f.
∂t2
Here, u = u(x, t) for t > 0 and x ∈ Ω where Ω is a bounded,
open subset of Rd with a piecewise smooth boundary.
Green identities and boundary-value problems
Example
If d = 1 then ∇ = d/dx so Lu = −(pu 0 ) 0 + qu.
Example
If d = 2 then
Lu = −(pux )x − (puy )y + qu
∂ ∂u ∂ ∂u
=− p(x, y) − p(x, y) + qu.
∂x ∂x ∂y ∂y
Ellipticity
ξT Aξ = p(x)kξk2
Example
If p(x) = 1 and q(x) = 0 then L = −∇2 .
Divergence theorem (Math2111)
Theorem
If a vector field F : Ω → Rd is continuously differentiable then
Z I
∇ · F dV = F · n dS,
Ω ∂Ω
Theorem
If u is C2 and v is C1 on Ω, then
Z I
∂u
hLu, viΩ = p∇u · ∇v + quv dV − p v dS. (50)
Ω ∂Ω ∂n
Proof.
Since
∇ · [(p∇u)v] = [∇ · (p∇u)]v + p∇u · ∇v
we have
Proof.
Since
(Lu)v = p∇u · ∇v + quv − ∇ · [(p∇u)v]
and
u Lv = p∇v · ∇u + qvu − ∇ · [(p∇v)u],
the Lagange identity generalizes to
(Lu)v − u Lv = ∇ · p(u∇v − v∇u) ,
so that L is elliptic.
Given a source term f, Dirichlet data gD and Neumann data gN ,
we seek a solution u to the BVP
Lu = f in Ω,
u = gD on ΓD , (53)
∂u
p = gN on ΓN .
∂n
Technical requirement: for u to count as a solution it must have
finite energy: Z
|∇u|2 + |u|2 dV < ∞.
Ω
Fredholm alternative
Either the homogeneous problem
∂v
Lv = 0 in Ω, v = 0 on ΓD , p = 0 on ΓN , (54)
∂n
has only the trivial solution v ≡ 0, in which case
the inhomogeneous problem (53) has a unique solution u
for “every” choice of f, gD and gN ;
or else the homogeneous problem admits finitely many, linearly
independent solutions v1 , v2 , . . . , vm , in which case
the inhomogeneous problem (53) has a solution u iff the
data f, gD and gN satisfy
Z Z Z
∂vk
fvk dV = gD p dS − gN vk dS
Ω ΓD ∂n ΓN
for k = 1, 2, . . . , m.
Partial proof
Write the second Green identity (51) as
I
∂v ∂u
hLu, viΩ − hu, LviΩ = up −p v dS.
∂Ω ∂n ∂n
and so
Z Z Z
∂v
fv dV = gD p dS − gN v dS.
Ω ΓD ∂n ΓN
A pure Neumann problem
Consider the BVP
∂u
−∇2 u = f in Ω, = g on ∂Ω. (55)
∂n
If v is a solution of the homogeneous problem,
∂v
−∇2 v = 0 in Ω, = 0 on ∂Ω,
∂n
then by the first Green identity,
Z Z I
2 2 ∂v
(−∇ v) v dV = k∇vk dV − v dS.
Ω | {z } Ω ∂n
∂Ω |{z}
0
0
R
so Ω k∇vk2 dV = 0. If Ω is connected then v must be constant,
so the inhomogeneous problem (55) is solvable iff the data satisfy
Z I
f dV + g dS = 0.
Ω ∂Ω
Elliptic eigenproblems
p(x) > const > 0 and r(x) > const > 0 for x ∈ Ω,
Writing
Lu = −∇ · (p∇u) + qu
as before, the PDE (56) can be written in the form
Lu = λru in Ω.
Boundary conditions
Lu = λru in Ω,
u=0 on ΓD , (57)
∂u
p =0 on ΓN ,
∂n
then u is said to be an eigenfunction with eigenvalue λ, and we
call (u, λ) an eigenpair.
and
∂u ∂u
(0, y) = (1, y) = 0, 0 < y < 1,
∂x ∂x
are given by
for k = 0, 1, 2, . . . and n = 1, 2, 3, . . . .
The eigenfunction φkn (x, y) = cos(kπx) sin(nπy)
Contours of φkn
Example: eigenproblem on the unit disk
Consider
∇2 u + λu = 0, 0 6 r < 1, −π < θ 6 π,
u(1, θ) = 0, −π 6 θ < π.
The eigenvalues λ = λnj are
for n = 0, 1, 2, . . . and j = 1, 2, 3, . . . .
Ω1 = { (x, y) : x2 + y2 < 32 },
Ω2 = (−1, 1)2 ,
Ω = Ω1 \ Ω2 ,
−∇2 u = λu in Ω,
u=0 on ΓD ,
∂u
=0 on ΓN ,
∂n
where ΓD denotes the inner (square) boundary and ΓN the outer
(circular) boundary. Compute approximate solutions using a finite
element method with 254,336 degrees of freedom.
λ1 = 0.505
λ2 = λ3 = 0.680
λ4 = 1.148 and λ5 = 1.251
Eigenfunctions are orthogonal
Theorem
Let L be the partial differential operator (49). If u, v are (possibly
complex-valued) functions satisfying
Lu = λru, Lv = µrv, in Ω,
u = 0, v = 0, on ΓD ,
∂u ∂v
p = 0, p = 0, on ΓN ,
∂n ∂n
and if λ 6= µ, then u and v are orthogonal on Ω with respect to
the weight function r, i.e.,
Z
hu, vir = uvr dV = 0.
Ω
Eigenvalues are real
Theorem
If u is a nontrivial solution of the elliptic eigenproblem (57) then λ
is real. Moreover, both Re u and Im u are also solutions of (57).
(λ − λ̄)hu, ūir = 0.
R
Since hu, ūir = Ω |u|2 r dV > 0, we conclude that λ = λ̄, or in
other words λ is real.
Finally, because u and ū are eigenfunctions with eigenvalue λ, so
are
u + ū u − ū
Re u = and Im u = .
2 2i
Completeness of the eigenfunctions
As before, let
Lu = −∇ · (p∇u) + qu,
but suppose now that ∂Ω is divided into three parts: Γ , ΓD and ΓN .
If u : Ω → C is not identically zero and if
Lu = 0 in Ω,
∂u
p = λru on Γ ,
∂n (58)
u=0 on ΓD ,
∂u
p =0 on ΓN ,
∂n
then we say that (u, λ) is a Steklov eigenpair.
Example: a Steklov eigenproblem for the unit square
for n = 1, 2, . . . .
The eigenfunction φn (x, y) = sin(nπx) cosh(nπy)
Contours of φn
Steklov eigenvalues are real
The first Green identity with v = ū gives
Z I
∂u
hLu, ūiΩ = p|∇u|2 + q|u|2 ) dV − p ū dS.
Ω ∂Ω ∂n
so Z
hu, vir,Γ = uvr dS = 0 if λ 6= µ.
Γ
Completeness?
The second type of eigenproblem (58) has a sequence of
eigenvalues λ = λj and corresponding eigenfunctions u = φj such
that
1. the eigenfunctions φ1 , φ2 , φ3 , . . . are orthogonal with
respect to r on Γ ;
2. the eigenvalues satisfy λ1 6 λ2 6 λ3 6 · · · with λj → ∞ as
j → ∞.
∂φj
Lφj = λj rφj in Ω, φj = 0 on ΓD , p = 0 on ΓN .
∂n
Then u exists iff
hf, φj iΩ = 0 whenever λj = 0,
∂φj ∂φj
Lφj = 0 in Ω, p = λj rφj on ΓD , p = 0 on ΓN .
∂n ∂n
If the homogeneous problem
∂v
Lv = 0 in Ω, v = 0 on ΓD , p = 0 on ΓN ,
∂n
has only the trivial solution v ≡ 0, then
∞
X hgD , φj ir,Γ
u(x) ∼ D
φj (x), x ∈ Ω.
j=1
kφj k2r,ΓD
Boundary-value problem with f = 0 and gD = 0
Consider
∂u
Lu = 0 in Ω, u = 0 on ΓD , p = gN on ΓN .
∂n
Find the Steklov eigenpairs (φj , λj ) satisfying
∂φj
Lφj = 0 in Ω, φj = 0 on ΓD , p = λj rφj on ΓN .
∂n
Then u exists iff
in which case
X X hgN , φj iΓ
u(x) ∼ Cj φj (x) + 2
N
φj (x), x ∈ Ω.
λj =0 λ 6=0
λ j kφ j kr,ΓN
j
Example
u = u1 + u2 + u3 ,
∂w
w = gD on ΓD and p = gN on ΓN ,
∂n
then we can solve (53) by putting f? = f − Lw, solving
Lu? = f? in Ω,
?
u =0 on ΓD ,
∂u?
p =0 on ΓN ,
∂n
and then putting u = u? + w.
plus
∂u ∂u
(0, y) = 0 = (1, y) for 0 < y < 1.
∂x ∂x
Use a previously solved eigenproblem to show that
∞
2 X sin(2m − 1)πy
u(x, y) ∼ y + 1 + 3
3π (2m − 1)3
m=1
∞ ∞
4 XX cos 2jπx sin(2m − 1)πy
− 5 .
π
j=1 m=1
j (2m − 1) 4j2 + (2m − 1)2
2
Wave and diffusion equations
∂2 u
− c2 ∇2 u = 0 in Ω, for t > 0,
∂t2
∂u (60)
u = u0 and = 0 in Ω, when t = 0,
∂t
u = 0 on ∂Ω, for t > 0.
then
∞
X hu(·, t), φj iΩ
u(x, t) ∼ uj (t)φj (x), uj (t) = ,
hφj , φj iΩ
j=1
and
∞ 2
X
∂2 u
2 2 d uj 2
−c ∇ u= + c λj uj φj (x).
∂t2 dt2
j=1
d2 u j hu0 , φj iΩ duj
+ c2 λj uj = 0, uj (0) = u0j = , (0) = 0.
dt2 hφj , φj iΩ dt
Normal modes of vibration
Solving the IVP for uj , we find that
p
uj (t) = u0j cos(ωj t) where ωj = c λj ,
so
∞
X
u(x, t) ∼ u0j cos(ωj t)φj (x).
j=1
q = −K∇u, (61)
On the LHS, Z Z
d ∂u
u dV = dV,
dt Ω Ω ∂t
and on the RHS the divergence theorem gives
I Z Z
q · (−n) dS = − ∇ · q dV = ∇ · (K∇u) dV.
∂Ω Ω Ω
Diffusion
Thus, Z
1 ∂u
− ∇ · (K∇u) − f dV = 0.
vol Ω Ω ∂t
The LHS tends to the pointwise value of the integrand if we shrink
Ω to a point, so u satisfies the diffusion equation,
∂u
− ∇ · (K∇u) = f(x, t). (62)
∂t
If K is independent of x, then
∇ · (K∇u) = K∇2 u.
Heat equation in 3D
If we let
q = −K∇u,
∂u
− ∇2 u = 0 in Ω, for t > 0,
∂t
∂u (63)
+ bu = 0 on ∂Ω, for t > 0,
∂n
u = u0 on Ω, when t = 0,
∇φ = 0 in Ω and φ = 0 on ∂Ω,
∂2 u
1 ∂ 2 ∂u 1 ∂ ∂u 1
2
ρ + 2
sin ϕ + 2 2
ρ ∂ρ ∂ρ ρ sin ϕ ∂ϕ ∂ϕ ρ sin ϕ ∂θ2
but from symmetry, u is independent of the angular variables
ϕ and θ, that is, u = u(ρ, t). Thus, it suffices to consider radially
symmetric eigenfunctions φ = φ(ρ) satisfying
1 d 2 dφ
− 2 ρ = λφ for 0 < ρ < 1.
ρ dρ dρ
(ρ2 φ 0 ) 0 + k2 ρ2 φ = 0.
A tutorial problem shows that the function v(ρ) = ρ1/2 φ(ρ)
satisfies the parameterised Bessel equation of order 1/2,
ρ2 v 00 + ρv 0 + (k2 ρ2 − 41 )v = 0,
Put Z1
hf, gi = f(ρ)g(ρ) ρ2 dρ
0
then
hφn , φj i = 0 if n 6= j,
since u = φn and λ = k2n satisfy (ρ2 u 0 ) 0 + λρ2 u = 0. Note that
dV = 4πρ2 dρ so if we view φn and φj as functions on Ω, then
Z Z1
φn φj dV = 4π φn φj ρ2 dρ = 0 if n 6= j.
Ω 0
gives
∞
X
ut − ∇ u ∼ 2
u̇j + λj uj )φj (ρ),
j=1
so we require that
Thus,
uj (t) = u0j e−λj t .
The series
∞
X
u(ρ, t) ∼ u0j e−λj t φj (ρ), 0 < ρ < 1, t > 0,
j=1
As expected, u(ρ, t) → 0 as t → ∞.