Asymptotic and Perturbation Methods
Asymptotic and Perturbation Methods
Methods
Preface 5
3
4 Contents
3.1.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.2 Forced Motion Near Resonance . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
3.3 Periodically Forced Nonlinear Oscillators . . . . . . . . . . . . . . . . . . . . . . 62
3.3.1 Isochrones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3.3.2 Phase equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
3.3.3 Phase resetting curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.3.4 Averaging theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
3.3.5 Phase-locking and synchronisation . . . . . . . . . . . . . . . . . . . . . . 67
3.3.6 Phase reduction for networks of coupled oscillators . . . . . . . . . . . . 68
3.4 Partial Differential Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
3.4.1 Elastic string with weak damping . . . . . . . . . . . . . . . . . . . . . . 70
3.4.2 Nonlinear wave equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
3.5 Pattern Formation and Amplitude Equations . . . . . . . . . . . . . . . . . . . . 73
3.5.1 Neural field equations on a ring . . . . . . . . . . . . . . . . . . . . . . . 73
3.5.2 Derivation of amplitude equation using the Fredholm alternative . . . . . 75
3.6 Problems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
These notes are largely based on Math 6730: Asymptotic and Perturbation Methods course,
taught by Paul Bressloff in Fall 2017, at the University of Utah. The main textbook is [Hol12],
but additional examples or remarks or results from other sources are added as we see fit, mainly
to facilitate our understanding. These notes are by no means accurate or applicable, and any
mistakes here are of course our own. Please report any typographical errors or mathematical
fallacy to us by email hkim@math.utah.edu or tan@math.utah.edu.
5
6
Chapter 1
Introduction to Asymptotic
Approximation
Our main goal is to construct approximate solutions of differential equations to gain insight
of the problem, since they are nearly impossible to solve analytically in general due to the
nonlinear nature of the problem. Among the most important machinery in approximating
functions in some small neighbourhood is the Taylor’s theorem: Given f ∈ C (N +1) (Bδ (x0 )),
for any x ∈ Bδ (x0 ) we can write f (x) as
N
X f (k) (x0 )
f (x) = (x − x0 )k + RN +1 (x),
k=0
k!
f (N +1) (ξ)
RN +1 (x) = (x − x0 )N +1
(N + 1)!
and ξ is a point between x and x0 . Taylor’s theorem can be used to solve the following problem:
Remark 1.0.1. If the given function is sufficiently differentiable, then Taylor’s theorem offers
a reasonable approximation and we can easily analyse the error as well.
7
8 1.1. Asymptotic Expansion
where ∞
e−t
Z
N +1 N +1
RN (ε) = (−1) (N + 1)!ε dt.
0 (1 + εt)N +2
Since ∞ ∞
e−t
Z Z
dt ≤ e−t dt = 1,
0 (1 + εt)N +2 0
it follows that
|RN (ε)| ≤ |(N + 1)!εN +1 |.
Thus, for fixed N > 0 we have that
f (ε) − PN a εk
k=0 k
lim =0
ε→0 ε N
or
N
X N
X
ak ε k + o ε N = ak εk + O εN +1 .
f (ε) =
k=0 k=0
N
X
The formal series ak εk is said to be an asymptotic expansion of f (ε) such that for fixed N ,
k=0
it provides a good approximation to f (ε) as ε −→ 0. However, this expansion is not convergent
for any fixed ε > 0, since
(−1)N N !εN −→ ∞ as ε −→ 0,
i.e. the correction term actually blows up!
|f (ε)| ≤ M |g(ε)| as ε −→ 0.
N
X
We typically writes f (ε) ∼ ak φk (ε) as ε −→ 0.
k=0
which will tend to be valid over some range of times t. It is often useful to characterise the
time interval over which the asymptotic expansion exists. We say that this estimate is valid
on a time-scale 1/δ̂(ε) if
x(t, ε) − PN a (t)φ (ε)
k=0 k k
lim = 0 for 0 ≤ δ̂(ε)t ≤ C,
ε→0 φN (ε)
d d d
f (x, ε) ∼ φ1 (x, ε) + φ2 (x, ε) as ε −→ 0.
dx dx dx
There are two possible scenarios:
Example 1.1.4. Consider f (x, ε) = e−x/ε sin ex/ε . Observe that for x > 0 we have that
f (x, ε)
lim = 0 for all finite n,
ε→0 εn
Example 1.1.5. Even if {φk (ε)} is an ordered asymptotic sequence, its derivative {φ0k (ε)}
need not be. Consider φ1 (x) = 1 + x, φ2 (x) = ε sin(x/ε) for x ∈ (0, 1). Then φ2 = o(φ1 ) but
and if
d
f (x, ε) ∼ b1 (x)φ1 (ε) + b2 (x)φ2 (ε) as ε −→ 0, (1.1.2)
dx
dak df
then bk = , i.e. the asymptotic expansion for can be obtained from term by term
dx dx
differentiation of (1.1.1). Throughout this course, we will assume that (??) holds whenever
we are given (1.1.1) which almost holds in practice. Integration, on the other hand, is less
problematic. If
x(ε) ∼ x0 + εα x1 + . . . as ε −→ 0, (1.2.2)
y y = 1 − 2x
y = εx2
x
(1/2,0)
Since the right-hand side is zero, the O(ε) in 1 must be balanced by the O(εα ) term in 2 .
This means that we must choose α = 1 and the O(ε) equation is
1
x20 + 2x1 = 0 =⇒ x1 = − .
8
Consequently, a two-term expansion of one of the roots is
1 ε
x(1) (ε) ∼ − + ... as ε −→ 0.
2 8
The chosen ansatz (1.2.2) produce an approximation for the root near x = 1/2 and we missed
the other root because it approaches negative infinity as ε −→ 0. One possible method to
generate the other root is to consider solving
ε(x − x1 )(x − x2 ) = 0,
but a more systematic method which is applicable to ODEs is to avoid the O(1) solution. Take
x ∼ εγ (x0 + εα x1 + . . . ) as ε −→ 0, (1.2.4)
Introduction to Asymptotic Approximation 13
The terms on the LHS must balance to produce zero, and we need to determine the order of
the problem that comes from this balancing. There are 3 possibilities on leading-order:
1
1 + 2γ = 0 =⇒ γ = − ,
2
so that the leading-order term in 1 , 3 are of O(1), whilst 2 = O(ε−1/2 ) which is lower
order that 1 . This is not possible.
1 + 2γ = γ =⇒ γ = −1,
so that the leading-order term in 1 , 2 are of O(ε−1 ) and 3 = O(1). This is consistent
with the assumption!
The solution x0 = 0 gives rise to the root x(1) (ε) by choosing α = 1, so the new root is obtained
by taking x0 = −2. Balancing the equation as before means we must choose α = 1 and the
O(ε) equation is
1
2x0 x1 + 2x1 − 1 = 0 =⇒ x1 = − .
2
Hence, a two-term expansion of the second root of (1.2.2) is
1 ε
x(2) (ε) ∼ −2 − as ε −→ 0.
ε 2
Remark 1.2.1. We may choose x0 = 1/2 in (1.2.2) since one of the root should be close to
x = 1/2 as we “switch on” ε in the term εx2 .
14 1.2. Algebraic and Transcendental Equations
x2 + eεx = 5 (1.2.7)
From Figure 1.2, we see that there are two real solutions nearby x = ±2. We assume an
asymptotic expansion of the form
x(ε) ∼ x0 + εα x1 + . . . as ε −→ 0, (1.2.8)
for some α > 0. Substituting (1.2.8) into (1.2.7) and expanding the exponential term eεx
around x = 0 we obtain
Balancing the O(εα ) term in 1 and the O(ε) term in 2 gives α = 1 and the O(ε) equation
is
1
2x0 x1 + x0 = 0 =⇒ x1 = − .
2
Hence, a two-term asymptotic expansion of each solution is
ε
x(ε) ∼ ±2 − as ε −→ 0.
2
y y = 5 − x2
y = eεx
√ √ x
(- 5,0) ( 5,0)
where we impose the condition µ(ε) 1 when ε 1. Substituting (1.2.11) into (1.2.10) we
obtain
x0 µ(ε)
[x0 + µ(ε)] + 1 + ε sech + = 0. (1.2.12)
ε ε
The O(1) equation remains x0 = −1 and (1.2.12) reduces to
x0 µ(ε)
µ(ε) + ε sech + = 0.
ε ε
Since
x0 µ(ε) 1 2
sech + ∼ sech − = 1/ε −1/ε
∼ 2e−1/ε ,
ε ε ε e +e
we require
µ(ε) = −2εe−1/ε = o(1) as ε −→ 0.
To construct the third term in the expansion, we would extend (1.2.11) into
x ∼ −1 − 2εe−1/ε + ν(ε),
y y = −x − 1
y = ε sech(x/ε)
(0,ε)
x
(-1,0)
d2 x gR2
2
≈ − 2
= −g, x(0) = 0, x0 (0) = v0 , (1.3.2)
dt R
where v0 is some initial velocity. The solution is
gt2
x(t) = − + v0 t. (1.3.3)
2
x(t)
2
v0 v0
,
g 2g
2v0
g
,0
t
Figure 1.4: Graph of x(t) versus t of the first approximation problem (1.3.2).
Unfortunately, this simplification does not determine a correction to the approximate solu-
tion (1.3.3). To this end, we nondimensionalise (1.3.1) with dimensionless variables
t x
τ= , y= ,
tc xc
Introduction to Asymptotic Approximation 17
where tc = v0 /g and xc = v02 /g are the chosen characteristic time and length scales respectively.
This results in the dimensionless initial-value problem
d2 y 1
2
= − 2
, y(0) = 0, y 0 (0) = 1. (1.3.4)
dτ (1 + εy)
xc v2
Observe that the dimensionless parameter ε = = 0 measures how high the projectile gets
R gR
in comparison to the radius of the Earth. Consider an asymptotic expansion
y(τ ) ∼ y0 (τ ) + εα y1 (τ ) + . . . as ε −→ 0. (1.3.5)
where the exponent α > 0 is included since a-priori there is no reason to assume α = 1.
Assuming we can differentiate (1.3.5) term by term, we obtain using generalised Binomial
theorem h i 1
y000 + εα y100 + . . . = − ∼ −1 + 2εy0 + . . . ,
[1 + εy0 + . . . ]2
with
y0 (0) + εα y1 (0) + · · · = 0, y00 (0) + εα y10 (0) + · · · = 1.
The O(1) problem is
τ2
y000 = −1, y0 (0) = 0, y00 (0) = 1 =⇒ y0 (τ ) = − + τ,
2
and we must choose α = 1 to balance the term 2εy0 . Consequently, the O(ε) problem is
τ3 τ4
y100 = 2y0 , y1 (0) = 0, y10 (0) = 0 =⇒ y1 (τ ) = − .
3 12
Hence, a two-term asymptotic expansion of the solution of (1.3.4) is
1 1 τ
y(τ ) ∼ τ 1 − τ + ετ 3 1 − .
2 3 4
Note that the O(1) term is the scaled solution of (1.3.1) in a uniform gravitational field and
the O(ε) term (first-order correction) contains the nonlinear effect of the problem.
where the αi are positive constants and zi is the valence of the ith ionic species. The whole
system must be neutral and this gives the electroneutrality condition
k
X
αi zi = 0. (1.3.7)
i=1
18 1.3. Differential Equations: Regular Perturbation Theory
We impose the Neumann boundary condition in which we assume the charge is uniform on the
boundary:
∇φ · n = ∂n φ = ε on ∂Ω, (1.3.8)
where n is the unit outward normal to ∂Ω.
This nonlinear problem has no known solutions. To deal with this, we invoke the classical
Debye-Hückle theory in electrochemistry which assumes that the potential is small enough
so that the Poisson-Boltzmann equation can be linearised. Because of the boundary condition
(1.3.8), we may assume the zeroth-order solution is 0 and guess an asymptotic expansion of
the form
φ ∼ ε (φ0 (x) + εφ1 (x) + . . . ) as ε −→ 0, (1.3.9)
where a small potential means ε is small. Substituting (1.3.9) into (1.3.6) and expanding the
exponential function around the point 0 yields
k
X
2 2
αi zi e−εzi (φ0 +εφ1 +... )
ε ∇ φ0 + ε∇ φ + . . . = −
i=1
k
X 1
=− αi zi 1 − εzi (φ0 + εφ1 + . . . ) + ε2 zi2 (φ0 + εφ1 + . . . )2 + . . .
i=1
2
k
X
2 1 2 2
=− αi zi 1 − εzi φ0 + ε −zi φ1 + zi φ0 + . . .
i=1
2
k
! k !
X X 1
∼ε αi zi2 φ0 + ε2 αi zi2 φ1 − zi φ20 .
i=1 i=1
2
k
X
Setting κ2 = αi zi2 , the O(ε) equation is
i=1
∇2 φ0 = κ2 φ0 in Ω, (1.3.10a)
∂n φ0 = 1 on ∂Ω. (1.3.10b)
k
1X
Setting λ = αi zi3 , the O(ε2 ) equation is
2 i=1
∇2 − κ2 φ1 = −λφ20
in Ω, (1.3.11a)
∂n φ1 = 0 on ∂Ω. (1.3.11b)
Take Ω to be the region outside the unit sphere, which is radially symmetric. Writing the
Laplacian operator ∇2 in terms of spherical coordinates, the solution must be independent of
the angular variables since the boundary condition is independent of the angular variables.
With φ0 = φ0 (r), the O(ε) equation now has the form
1 d 2 dφ0
r − κ2 φ0 = 0 for 1 < r < ∞, (1.3.12a)
r2 dr dr
φ00 (1) = −1, (1.3.12b)
Introduction to Asymptotic Approximation 19
where the negative sign is due to n = −r̂. The bounded solution of (1.3.12) is
1
φ0 (r) = eκ(1−r) ,
(1 + κ)r
where the exponential term is the screening term. With φ1 = φ1 (r), the O(ε2 ) equation takes
the form
1 d 2 dφ0 λ
2
r − κ2 φ1 = − e2κ(1−r) for 1 < r < ∞, (1.3.13a)
r dr dr (1 + κ)2 r2
φ01 (1) = 0. (1.3.13b)
Spectral problems are widely studied in the context of time-dependence PDEs when time-
harmonic solutions are sought for instance, and we are interested in the behaviour of the
spectrum of L0 as we perturb L0 . Suppose further that for ε = 0, the unperturbed equation
has a unique solution (λ0 , φ0 ) with λ0 non-degenerate. For simplicity, take L0 to be self-adjoint,
that is
hf, L0 gi = hL0 f, gi.
Since L0 , L1 are linear, we introduce the asymptotic expansions for both the eigenfunction
φ and eigenvalue λ with asymptotic sequence {1, ε, ε2 , . . . }
φ ∼ φ0 + εφ1 + ε2 φ2 + . . .
λ ∼ λ0 + ελ1 + ε2 λ2 + . . . .
We obtain
1.4 Problems
1. Consider the transcendental equation
√
1+ x2 + ε = e x . (1.4.1)
Explain why there is only one small root for small ε. Find the three term expansion of
the root
x ∼ x0 + x1 εα + x2 εβ , β > α > 0.
√
Solution: Consider two graph f (x) = x2 + ε and g(x) = ex − 1. If x < 0, then
f (x) > 0 > g(x). It means that there is no solution in negative region. If x > 0, then
f (x) → x as x → ∞ starting its curve from f (0) = ε. One can draw graph of f (x)
and g(x) on x > 0, then it yields there is only one solution.
To obtain first expansion, set ε = 0. Then we get
1 + x = ex =⇒ x = 0.
Since there is only on solution for all ε > 0, then one can set x0 = 0 and expand x
further at this point. To do so, rewrite the equation (1.4.1) as
x2 + ε = (ex − 1)2
Before we balance both sides, consider the leading-order of both sides. Without
doubt, the leading-order is ε2α with coefficient x21 . For LHS, we have three cases
λ = tan(λ).
(a) After sketching the two functions in the equation, establish that there is an infinite
number of solutions, and for sufficiently large λ takes the form
π
λ = πn + − xn ,
2
with xn small.
λ ∼ ε−α λ0 + εβ λ1 ,
and determine ε, α, β, λ0 , λ1 .
22 1.4. Problems
Solution:Set λ = 1/ε and see the asymptotic behavior of xn . Then one can
figure out an asymptotic expansion of λ. For convenience, set xn = x. Then
one can get
1 1
− x = tan − x = cot(x)
ε ε
because tan(1/ε) = 0. By multiplying ε tan(x) on both sides and we get
tan(x) − εx tan(x) = ε.
Since θ > 0, the leading-order of LHS is εθ . To balance both sides with O(ε),
set θ = 1 and get x0 = 1. Therefore, an asymptotic expansion of λ is
1 1
λ= − x ∼ − ε = ε−1 (1 + ε2 (−1)).
ε ε
It follows that α = −1, β = 2, λ0 = 1 and λ1 = −1.
3. In the study of porous media one is interested in determining the permeability k(s) =
F 0 (c(s)), where
Z 1
F −1 (c − εr) dr = s
0
−1
F (c) − F −1 (c − ε) = β,
and β is a given positive constant. The functions F (c) and c both depend on ε, whereas
s and β are independent of ε. Find the first term in the expansion of the permeability
for small ε. Hint: consider an asymptotic expansion of c and use the fact that s is
independent of ε.
From the second condition, expanding F −1 as Taylor series follows provides that
dF −1 dF −1
F −1 (c0 ) + εc1 (c0 ) − F −1 (c0 ) − ε(c1 − 1) (c0 ) + O(ε2 ) = β.
dc dc
It follows that
1 ε
ε + O(ε2 ) = β =⇒ k(s) = F 0 (s) ∼ .
F 0 (s) β
(a) Suppose A is symmetric and has n distinct eigenvalues. Find a two-term expansion
of the eigenvalues of the perturbed matrix A + εD, where D is positive definite.
λ ∼ λ0 + ελ1 + ε2 λ2 + . . .
x ∼ x0 + εx1 + ε2 x2 + . . . .
The O(1) equation is Ax0 = λ0 x0 which means that (λ0 , x0 ) is the eigenpair of
the matrix A. The O(ε) equation is
Ax1 + Dx0 = λ0 x1 + λ1 x0 ,
or
Lx1 = (A − λ0 I)x1 = λ1 x0 − Dx0 .
It follows from the Fredholm Alternative that the solvability condition for λ1 is
Consequently,
xT0 Dx0
0 = xT0 Lx1 = xT0 (λ1 x0 − Dx0 ) =⇒ λ1 = .
xT0 x0
√
0 1
Solution: The perturbed matrix A + εD = has eigenvalues λ = ± ε,
ε 0
which is not of O(ε) and is not differentiable at ε = 0.
5. The eigenvalue problem for the vertical displacement y(x) of an elastic string with variable
density is
y 00 + λ2 ρ(x, ε)y = 0, 0 < x < 1,
where y(0) = y(1) = 0. For small ε, assume ρ ∼ 1 + εµ(x), where µ(x) is positive and
continuous. Consider the asymptotic expansions
Solution: Substituting the given asymptotic expansions together with the ap-
proximation ρ ∼ 1 + εµ(x) gives
d2
Using integration by parts, one can show that the linear operator L = 2 + λ20
dx
with domain
D(L) = f ∈ C 2 [0, 1] : f (0) = f (1) = 0 ,
is self-adjoint with respect to the L2 inner product over [0, 1]. Moreover, for a
fixed λ0 it has a one-dimensional kernel ker(L) = span(sin(λ0 x)). We can now
determine λ1 using Fredholm alternative, this results in
since
1 1
1 − cos(2nπx) A2
Z Z
2 2 2
hy0 , y0 i = A sin (nπx) dx = A dx = .
0 0 2 2
(b) Using the equation for y1 , explain why the asymptotic expansion can break down
when λ0 is large.
Solution: From the previous results, one can find the equation for y1 as
Z 1
00 2 2 2
y1 (x) + λ0 y1 (x) = λ0 −µ(x)y0 (x) + 2 µ(s)y0 (s)ds .
0
Notice that the RHS proportional to λ20 and it follows that the particular solution
of y1 is proportional to λ20 , then it implies that y1 → ∞. This can break down
the expansion mixed with ε.
This is a Fredholm integral equation, where the kernel K(x, d) is known and is assumed
to be smooth and positive. The eigenfunction y(x) is taken to be positive and normalized
so that Z a
y 2 (s) ds = a.
0
Both y(x) and λ depend on the parameter a, which is assumed to be small.
(a) Find the first two terms in the expansion of λ and y(x) for small a.
(c) From part (b) find the two-term expansion for λ and φ(ξ) for small a.
Then balance O(1) terms in both sides of the equations and we get
( R1
K(0, 0) 0 φ0 (r)dr = λ0 φ0 (ξ)
R1 .
0
(φ0 (r))2 dr = 1
Similarly, balance O(a) terms of the equations and one can obtain
R1 R1 R1
K(0, 0) 0 φ1 (r)dr + ξKx (0, 0) 0 φ0 (r)dr + Kd (0, 0) 0 rφ0 (r)dr
= λ0 φ1 (ξ) + λ1 φ0 (ξ)
R 1
φ (r)φ1 (r)dr = 0
0 0
R1
It follows that 0
φ1 (r)dr = 0 and one can have
1
ξKx (0, 0) + Kd (0, 0) = λ0 φ1 (ξ) + λ1 .
2
28 1.4. Problems
(d) Explain why the expansions in parts (a) and (c) are the same for λ but not the
eigenfunction.
7. In quantum mechanics, the perturbation theory for bound states involves the time-
independent Schrodinger equation
where ψ(−∞) = ψ(∞) = 0. In this problem, the eigenvalue E represents energy and V1
is a perturbing potential. Assume that the unperturbed (ε = 0) eigenvalue is nonzero
and nondegenerate.
(a) Assuming
write down the equation for ψ0 (x) and E0 . We will assume in the following that
Z ∞ Z ∞
2
ψ0 (x) dx = 1, |V1 (x)| dx < ∞.
−∞ −∞
(b) Substituting ψ(x) = eφ(x) into the Schrodinger equation and derive the equation for
φ(x).
Plug them into the Schrodinger equation and drop common term eφ(x) . Then it
follows that
φ00 (x) + (φ0 (x))2 − (V0 (x) + εV1 (x)) = −E.
Introduction to Asymptotic Approximation 29
Solution: Assume that φ(x) ∼ φ0 (x) + εφ1 (x) + ε2 φ2 (x). Substituting it into
the new Schrodinger equation and balance O(ε) terms. Then one can obtain
Define an differential operator L = d2 /dx2 +2φ00 ·d/dx. Notice that for sufficiently
smooth f , Z ∞
hψ02 , Lf i = e2φ0 (x) (f 00 (x) + 2φ00 (x)f 0 (x))dx.
−∞
and it follows that hψ02 , Lf i = 0. From the observation, take inner product with
ψ02 to the first order balance equation and get
Therefore, Z ∞
E1 = hψ02 , V1 i = V1 (x)[ψ0 (x)]2 dx. (1.4.11)
−∞
To find φ1 , solve the first order inhomogeneous ODE of φ01 by integrating factora ,
or observe that
Z x Z x
d 2 dφ1
2
ψ0 Lφ1 dy = ψ0 dy = ψ02 (x)φ01 (x)
−∞ dy dy
Z−∞
x
= ψ02 (V1 − E1 )dy
−∞
For most of singular perturbation problem of differential equations, the solution has extreme
changes because a singular problem converges to a differential equation with different order or
behavior as → 0. If we apply the regular asymptotic expansion, it fails to represent such
drastic change and to match all boundary condition. To resolve the problem, we introduce
matched asymptotic expansion which approximates the exact solution by zooming in the ex-
treme changing zones, such as inner or boundary layers, together with the regular expansion
for outer region.
If = 0, then we have a first order ODE. It only needs one boundary condition. It yields to
have drastic dynamics on boundary layer. Remark that boundary layer could be interior, not
only near boundary of domain.
(y000 (x) + y100 (x) + · · · ) + 2(y00 (x) + y10 (x) + · · · ) + 2(y0 (x) + y1 (x) + · · · ) = 0.
It leads to dilemma that the solution has only one arbitrary constant but we have two boundary
conditions. It is over-determined. Moreover, the outer solution cannot describe solution over
the whole domain [0, 1]. The following question is which boundary layer would we use?
31
32 2.1. Introductory example
Figure 2.1: Two choices of boundary layers. It can be chosen by investigating the sign of y 00
near the boundary layer by looking at the concavity of the function.
d2 Y dY
1−2α 2
+ 2−α + 2Y = 0, Y (0) = 0. (2.1.2)
dx̃ dx̃
Try a solution of a form
d2 d
1−2α 2 (Y0 + γ Y1 + · · · ) + 2−α (Y0 + γ Y1 + · · · ) + 2(Y0 + γ Y1 + · · · ) = 0.
| dx̃ {z } | dx̃ {z } | {z }
(3)
(1) (2)
• Balance (1) and (3) with taking (2) is higher order. Then it requires α = 1/2. Then (1),
(3) = O(1), but (3) = O−1/2 . (×)
• Balance (1) and (2) with taking (3) is higher order. Then it requires α = 1. Then (1),
(2) = O(−1 ) and (3) = O(1). (Yay!)
Choosing the last balance, one can obtain an equation from O(−1 ) terms
One can get inner solution Y0 (x̃) = A(1 − e−2x̃ ), where A is unknown constant.
Matched Asymptotic Expansions 33
2.1.3 Matching
It remains to determine the constant A. The inner and outer solutions are both approximations
of the same function. Hence they sould agree in the transition zone between inner and outer
layers. Thus
lim Y0 (x̃) = lim+ y0 (x), (2.1.3)
x̃→∞ x→0
1. change from x to xη in outer expansion youter (xη ). Assume there is η1 () such that youter
is valid for η1 () η() ≤ 1.
2. Change variable x̃ to xη in inner expansion to obtain yinner (xη ). Assume there is η2 ()
such that inner is valid for η() η2 ().
Return to our particular example. Let xη = x/β with 0 < β < 1. Then
1−β
yinner ∼ A(1 − e−2xη / ) ∼ A + O(β−1 ),
and
β
youter ∼ e1−xη ∼ e + O(β ).
These are hard to match so we consider higher-order term. Find the second balance equation
1 1
y100 + y1 = − y000 , y1 (1) = 0 =⇒ y1 (x) = (1 − x)e1−x
2 2
from O() terms of outer expansion and
1
∼ e · 1 − e · xη β + · e · 1 + e · x2η 2β + · · ·
2 x
2
η
yinner ∼ e(1 − eξ ) + B(1 − eξ ) − 1−β e(1 + eξ ) , ξ = −2xη /1−β
β
∼ e − xη · e + · B + · · · ,
which is singular and non-linear. Note that in case when = 0, we get y = ex and it does
not match any boundary conditions. This solution is the first term in the outer solution, i.e.
y0 (x) = ex .
Start to find inner solution at x = 0. Set x̃ = x/α and Y (x̃) = y(x). Then we have
d2 d α
2−2α
2
Y = −e−x̃ = −(1 + x̃α + · · · ) .
Y + x̃ Y − |{z}
| {zdx̃ } | {z dx̃ }
(3)
| {z }
(4)
(1) (2)
In order to balance (1),(3) and (4), we require α = 1. Then taking Y ∼ Y0 + · · · yields the
following balance equation for O(1)
d2 d β
2−2β
2
Y + (1 + β x̃)1−β Y − Y = e−1+ x̃ .
dx̃ dx̃
Achieve balance for β = 1 and we obtain
00 0
Y 0 + Y 0 − Y 0 = −e, − ∞ < x̃ < 0, with Y 0 (0) = 1.
with y(0) = 1 and y(1) = −1. For its outer equation, setting y ∼ y0 + · · · yields
y0 (y00 − 1) = 0 =⇒ y0 = 0 or y0 (x) = x + a
for some constant a. Since the outer equation does not satisfy both boundary condition at
once, we need to find a boundary layer to fit boundary conditions.
Assume that boundary layer is at x = 0. In the boundary layer, y 00 > 0 and y 0 < 0. Since y
can be positive, we cannot match signs of differential equation everywhere. If boundary layer
is at x = 1, then y 00 < 0 and y 0 − 1 < 0. Since y can be negative, it cannot match signs
everywhere in boundary layer. What if it has interior layer at x = x0 ? For x − x0 = 0− , we
have y 00 < 0, y 0 − 1 > 0, but y > 0. For x − x0 = 0+ , we have y 00 > 0, y 0 − 1 < 0, but y < 0.
Thus interior layer can match the signs.
From the argument of interior layer argument, find inner solution by setting x̃ = (x−x0 )/α ,
0 < x0 < 1. Then we have two outer regions 0 ≤ x < x0 and x0 < x ≤ 1. The inner equation
is
1−2α Y 00 = −α Y Y 0 − Y,
and one can balance if α = 1. Setting Y (x̃) ∼ Y0 (x̃) + · · · gives
1
Y000 = Y0 Y00 =⇒ Y00 = Y02 + A.
2
It has three general solution depending on sign of A:
36 2.3. Partial differential equations
h i
1−DeB x̃
1. Y0 = B 1+DeB x̃
if A > 0,
B x̃
2. Y0 = B tan C − 2
if A < 0,
2
3. Y0 = C−x̃
if A = 0.
Three forms (rather than a single general solution) reflects non-linearity.
Next, match the inner solution with outer solution
(
x + 1 , x < x0
y0 (x) = .
x − 2 , x0 < x
Only inner solution 1. can match these outer solutions. Without loss of generality, assume
B > 0 and we get
−B = Y0 (+∞) = y0 (x+
0 ) = x0 − 2 and B = Y0 (−∞) = y0 (x−
0 ) = x0 + 1.
It yields x0 = 1/2 and B = 3/2. What about D? Remember that y(x0 ) = 0. This implies that
3 1−D
Y0 (x0 ) = 0 = · =⇒ D = 1.
2 1+D
Therefore,
3 1 − e3x̃/2
· Y0 (x̃) ∼
.
2 1 + e3x̃/2
Finally, the composite solution can be constructed in the two domains [0, x0 ) and (x0 , 1]
(
1−e3(2x−1)/4
x + 1 + 32 · 1+e 3
3(2x−1)/4 − 2 , 0 ≤ x < x0
y(x) ∼ 3 1−e3(2x−1)/4 3
.
x − 2 + 2 · 1+e3(2x−1)/4 + 2 , x0 < x ≤ 1
Notice that this perturbation problem is singular because type of solution is changed from
parabolic to hyperbolic when > 0 to = 0. Assume that φ(x) is smooth and bounded except
for a jump continuity at x = 0 with φ(0− ) > φ(0+ ) and φ0 ≥ 0. For concreteness, set
(
1, x < 0
u(x, 0) = .
0, 0 < x
This is an example of a Riemann problem – evolves into a traveling front that sharpens as
→ 0. We can handle it in similar way for boundary layer problem. For outer solution,
expanding u(x, t) ∼ u0 (x, t) + · · · gives balance equation for O(1) terms
∂t u0 + u0 · ∂x u0 = 0.
Matched Asymptotic Expansions 37
x = x0 + φ(x0 )t.
Characteristic into set at the shock x = s(t) with determined using the Rankine-Hugoniot
equation
− 2
1 [φ(x+ 2
0 )] − [φ(x0 )] 1 −
ṡ = · − = [φ(x+
0 ) + φ(x0 )]. (2.3.3)
2 φ(x+0 ) − φ(x0 ) 2
We will derive an equation for s(t) using match asymptotics. Introduce a moving inner layer
around s(t)
x − s(t)
x̃ = .
α
The inner PDE for U (x̃, t) = u(x, t)
lim U0 = u−
0 and lim U0 = u+
0
x̃→−∞ x̃→+∞
where u±
0 = limx→s(t)± u0 (x, t). Since U0 (x̃, t) is a constant for x̃ → ±∞, we have ∂x̃ U0 → 0 as
x̃ → ±∞. Then we have
1
0 = [u− 2 0 −
0 ] − s (t)u0 + A(t),
2
1 0
0 = [u+ 2 +
0 ] − s (t)u0 + A(t).
2
1 −
∂x̃ U0 = (U0 − u+
0 )(U0 − u0 ),
2
38 2.4. Strongly localized perturbation theory
with u± ±
0 = u0 (t). Then one can achieve the following equations
Z Z
1 1 1 −
dU0 + − − = dx̃(u+
0 − u0 )
U0 − u0 U0 − u0 2
U0 − u+
0
1 + −
=⇒ log − = (u0 − u0 )x̃ + C(t)
U0 − u0 2
U0 − u+ + −
=⇒ − 0
= b(x̃, t) = B(t)e(u0 −u0 )x̃/2
u0 − U0
where B(t) = eC(t) . Therefore,
−
u+
0 + b(x̃, t)u0
U0 (x̃, t) = .
1 + b(x̃, t)
In order to determine B(t), we have to go to next order [See Holmes for more details]. You
may find, in the end, s
1 + tφ0 (x+
0)
B(t) = .
1 + tφ (x−
0
0)
Construct the inner solution near the hole. Let y = (x − x0 )/ and set V (x; ) = u(x0 + y).
Then we find that V satisfies
A classical result from PDE theory is Vc ∼ C/|y| as |y| → ∞ where C is electrostatic capaci-
tance of Ω0 , determined by shape and size of Ω0 . We now have
C
V0 (x) ∼ φ0 (x0 ) 1 − .
|x − x0 |
It has to match
φ0 (x0 ) + ν()u1
as x → x0 . This yields that ν() = and u( x) → −φ0 (x0 )C/|x − x0 | as x → x0 . To evaluate
perturbed eigenvalue λ1 , return to equation (2.4.3). Since u1 → 4πφ0 (x0 )C · (−1/4π|x − x0 |)
as x → x0 , then we have the modified problem
(
Lu1 ≡ 4u1 + µu1 = −λ1 φ0 + 4πCφ0 (x0 )δ(x − x0 ), in Ω
. (2.4.4)
u1 = 0, on ∂Ω
40 2.4. Strongly localized perturbation theory
and it yields
4πCφ20 (x0 )
λ1 = R 2 .
φ dx
Ω 0
Therefore, λ ∼ µ0 + λ1 .
Remark 2.4.1. 1. Let us assume that u = 0 on ∂Ω is replaced by the no-flux condition on
∂Ω. Then = 0 problem becomes
4φ + µφ = 0,
∂n φ = 0, ∂Ω .
R 2
Ω
φ dx = 1
The principal eigenvalues µ0 = 0 and φ0 (x) = 1/|ω|1/2 . In this case, λ1 ∼ 4πC/|Ω| (to
leading order it is independent of location x0 .)
holes Ωj for j R= 1, · · · n and well-separated, its eigenvalue expansion is
2. For multiple P
λ ∼ µ0 + 4π j cj [φ0 (xj )]2 / Ω φ20 dx.
gives that
Vc ∼ log |y| − log d + O(1/|y|), as |y| → ∞
where d is logarithmic capacitance determined by shape of Ω0 . It is interesting enough to notice
the logarithmic capacitance of simple objects in the table. Then write inner solution in outer
variable
|y|
u(x) ∼ ν()A0 log ∼ ν()A0 [− log(d) + log |x − x0 |] .
d
Matching solution yields that
as x → x0 . In order to match the conditions, set ν() = −1/ log(d). Then unknown constant
A0 = φ0 (x0 ). Thus,
u1 (x) ∼ φ0 (x0 ) log |x − x0 |,
as x → x0 . Hence, by the same procedure in 3D case by using Green’s identity, one can find
eigenvalue expansion
[φ0 (x0 )]2
λ ∼ µ0 + 2π · ν() R 2 .
φ dx
Ω 0
λ ∼ µ0 + A1 ν + A2 ν 2 + A3 ν 3 + · · · .
Its potential problem is that the log decreases very slowly as decreases. Then the remaining
term is quite large and break the asymptotic expansions. By summing the log series, one can
solve the problem.
In the outer region, set w(x; ) = w0 (x; ν()) + σ()w1 (x; ν()) + · · · where ν() = −1/ log(d)
and σ ν k for any k > 0. It gives the outer equation
4w0 = −B,
in Ω\{x0 }
w = 0, on ∂Ω . (2.4.7)
w is singular, as x → x0
In the inner region, set y = (x−x0 )/ and V (y; ) = w(x0 +y; ). Expand V (y; ) = V0 (y; ν())+
µ0 ()V1 (y; ν()) + · · · where µ0 ν k for all k > 0. Then V0 satisfies
(
4y V0 = 0, outside Ω0
. (2.4.8)
V0 = 0, on ∂Ω0
42 2.5. Exercises
Introduce an unknown function γ = γ(ν) with γ(0) = 1 and let V0 (y; ν) = νγVc (y). Then it
follows that
4y Vc = 0,
outside Ω0
Vc = 0, on ∂Ω0 .
Vc ∼ log |y|, as |y| → ∞
Thus, Vc (y) ∼ log |y| − log d + O(1/|y|) for 1 |y|. In original coordinate,
V0 (y; ν) ∼ γ + νγ log |x − x0 |.
2.5 Exercises
1. Find a composite expansion of the solution to the following problems on x ∈ [0, 1] with
a boundary layer at the end x = 0:
0 + 2y00 + y03 = 0.
Now, construct inner expansion near x = 0. Setting x̃ = x/α and Y (x̃) = y(x)
yields ODE for inner solution
1−2α Y 00 + 2−α Y 0 + Y 3 = 0.
Y000 + 2Y00 = 0.
A general solution of Y0 is
y(0) = 0 =⇒ y(0) = 0.
Since f is positive function limx→0 y0 (x) = f (0) > 0, which does not match
boundary condition. It implies that the expansion has boundary layer at x = 0.
Scale near x = 0 by taking a new coordinate x̃ = x/α and Y (x̃) = y(x). In this
coordinate, the smooth function f (x) can be count as a constant f (x) ∼ f (0).
It follows the ODE for inner expansion
q0
y 0 = y − q(y − f )x,
q
and one can rewrite it as
0
y y
+ xq = xf =⇒ z 0 + xqz 0 = xf
q q
by setting z = y/q. In the same fashion in part 1., obtain outer expansion by
balancing O(1)
f (x)
z(x) ∼ z0 (x) = =⇒ y(x) ∼ y0 (x) = f (x).
q(x)
Since f, q are positive, then z has boundary layer at x = 0. With the same
argument in part 1., one can get the ODE for inner expansion Z(x̃)
f (0) 2
Z 0 + xq(0)Z = xf (0) =⇒ Z(x̃) = 1 − e−q(0)x̃ /2 ,
q(0)
−q(0)x̃2 /2
that is Y (x̃) ∼ f (0) 1 − e . Therefore, its composite expansion is
2 /2
y(x) ∼ f (x) − f (0)e−q(x)x .
(c) Show that solution of part 2. still holds if q(x) is continuous but not differentiable
everywhere on [0, 1].
Solution: The basic idea showing the claim is to derive all the expansions from
46 2.5. Exercises
Now, one can take derivative and get the same differential equation for inner
expansion. Therefore, one can achieve the same composite expansion.
3. (Boundary layer at both ends) Find a composite expansion of the following problem on
[0, 1] and sketch the solution:
0 = y0 + x − 1 =⇒ y0 (x) = 1 − x.
It does not satisfy neither boundary conditions. Hence there are two boundary layer
at x = 0 and x = 1. First, consider boundary layer at x = 0 by setting x̃ = x/α for
α > 0 and U (x̃) = y(x). It follows ODE for U
Since α > 0, the smallest order of LHS is O(1 − 2α) and RHS is O(1). To balance
them, require α = 1/2 and setting U ∼ U0 provides
Since β > 0, the smallest order of LHS is O(1 − 2α) and RHS is O(1). To balance
them, require β = 1/2 and setting U ∼ U0 provides
4. (Matched asymptotics can also be used in the time domain) The Michaelis-Menten reac-
tion scheme for an enzyme catalyzed reaction is
ds
= −s + (µ + s)c,
dt
dc
= s − (κ + s)c,
dt
where s(0) = 1, c(0) = 0. Here s(t) is the concentration of substrate, c(t) is the concen-
tration of the catalyzed chemical product, and µ, κ are positive constants with µ < κ.
Find the first term in the expansions in the outer layer, the initial layer around t = 0,
and the composite expansion.
Solution: Find the expansions in the outer layer by setting s ∼ s0 + · · · and c ∼
c0 + · · · and balancing O(1) terms
ds0
= −s0 + (µ + s0 )c0 ,
dt
0 = s0 − (κ + s0 )c0 .
It yields that
s0 (t)
s0 (t) − 1 + κ log s0 (t) = (µ − κ)t, c0 (t) = .
s0 (t) + κ
Notice that s0 is implicitly determined. One can observe that c has a layer near t = 0.
Setting t̃ = t/α , S(t̃) = s(t) and C(t̃) = c(t) gives the system of ODE
dS
−α = −S + (µ + S)C,
dt̃
dC
1−α = S − (κ + S)C
dt̃
48 2.5. Exercises
It requires that α = 1 to balance equation for C not same with outer expansion. By
setting S ∼ S0 and C ∼ C0 + · · · , it follows that
dS0
= 0,
dt̃
dC0
= S0 − (κ + S0 )C0 .
dt̃
First equation with initial condition s(0) = 1 gives that S0 (t̃) = 1. Hence we write
ODE for C0 as
dC0 1
= 1 − (κ + 1)C0 =⇒ C0 (t̃) = 1 − e−(κ+1)t̃ .
dt̃ κ+1
Fortunately, this solution satisfies matching condition
1 s0 (0)
lim C0 (t̃) = = = lim c0 (t).
t̃→∞ κ+1 s0 (0) + κ t→0
5. (Implicit inner solution) A classical model in gas lubrication theory is the Reynolds
equation
d d
H 3 yy 0 =
(Hy), 0 < x < 1,
dx dx
where y(0) = y(1) = 1. Here H(x) is a known, smooth, positive function with H(0) 6=
H(1).
(a) Suppose that there is a boundary layer at x = 1. Construct the first terms of the
outer and inner solutions. Note that the leading order term Y0 of the inner solution
is defined implicitly according to (x − 1)/ = F (Y0 ). Calculate the function F .
Solution: Setting y ∼ y0 and balancing O(1) terms yields outer solution equa-
tion
d C
0= (Hy0 ) =⇒ y0 (x) =
dx H(x)
where C is constant. Since we have boundary layer at x = 1, then applying
boundary condition at x = 0 to outer solution gives y0 (x) = H(0)/H(x). In the
inner layer, setting x̃ = (x − 1)/α and Y (x̃) = y(x) provides the ODE for inner
solution
d d
1−2α H 3 Y Y 0 = −α (HY ).
dx̃ dx̃
Since the inner layer near x = 1, then continuous function H can be approxi-
Matched Asymptotic Expansions 49
d d
1−2α H 3 (1) (Y0 Y00 ) = −α H(1) Y0 .
dx̃ dx̃
for first expansion term Y0 of Y . To balance the equation, it requires α = 1 and
now get
d 1 d
(Y0 Y00 ) = 2 Y0 .
dx̃ H (1) dx̃
The general solution of ODE is given by
Y x̃
Y0 (x̃) − 1 − C log 1 + = 2
C H (1)
H(0)
lim Y0 (x̃) = lim y0 (x) = .
x̃→∞ x→1 H(1)
Solution: By the result from part 1., one can write the composite solution as
H(0) −1 x−1 H(0)
y(x) ∼ + F − .
H(x) H(1)
(c) Show that if the boundary layer was assumed to be at x = 0, then the inner and
outer solutions would not match.
Solution: It follows the same procedure in part 2., but achieve different F
2
H(0)Y0
F (Y0 ) = H (0)(Y0 − 1) + H(0)H(1) log 1 −
= x̃.
H(1)
However, as x̃ → ∞, the RHS tends to negative infinity. It does not match the
conditions.
6. (Boundary layer at both ends) In a one-dimensional bounded domain, the potential φ(x)
of an ionized gas satisfies
d2 φ
− + h(φ/) = α, 0 < x < 1,
dx2
with boundary conditions
φ0 (0) = −γ, φ0 (1) = γ.
50 2.5. Exercises
The function h(s) is smooth and strictly increasing with h(0) = 0. The positive constants
α and β are known (and independent of ), and the constant γ is determined from the
conservation equation.
(b) Find the exact solution for the potential when h(s) = s. Sketch the solution for
γ < 0 and small , and describe the boundary layers that are present.
d2 φ φ
− + = α,
dx2
and its general solution is
x x
φ(x) = A sinh √ + B cosh √ + α,
For x 6= 0, 1, then φ0 (x) decays to zero as → 0. Since φ0 (0) and φ0 (0) are
nonzero, then it implies that φ has boundary layers at x = 0, 1.
(c) Suppose that h(s) = s2k+1 , where k is a positive integer, and assume β < α. Find
the first term in the inner and outer expansions of the solution.
d2 φ
− 2k+1 + φ2k+1 = 2k+1 α, (2.5.1)
dx2
with same boundary conditions. For = 0, φ has a trivial solution. Thus we
expand φ as
φ ∼ p (φ0 + q φ1 + · · · ),
and its derivatives are
d d dx̃ d
→ = −r .
dx dx dx dx̃
It allows the governing equation in the boundary layer at x = 0 to be
and it requires r = p and gives Φ0 (0) = −γ. Then (2.5.2) turns out to be
− Φ000 + Φ2k+1
0 = 0. (2.5.4)
52 2.5. Exercises
1 0 2 Φ2k+2
− (Φ0 ) + 0 = C,
2 2k + 2
for some constant C. As x̃ → ∞, Φ0 matches with the outer solution φ(x) = 0
for 0 < x < 1. It implies that C = 0. Then we have its general solutions
−1/k
k
Φ0 (x̃) = √ (±x̃ − D) ,
k+1
for some constant D. Its derivative becomes
−(k+1)/k
0 1 k k
Φ0 (x̃) = − √ (±x̃ − D) · ±√ . (2.5.5)
k k+1 k+1
Imposing boundary condition at x = 0 gives
−(k+1)/k
1 kD k
− −√ · ±√ = −γ.
k k+1 k+1
Since γ = (β − α)/2 < 0, then we choose the negative sign and determine D
such that
kD √
−√ = (−γ k + 1)−k/(k+1) := λ.
k+1
√
Setting κ = k/ k + 1 gives
Since the boundary layer at x = 1 satisfies the same differential equation (2.5.4),
then one can derive the lowest order boundary layer Ψ0 with rescaling x̂ =
(1 − x)/r
Ψ0 (x̂) = (λ + κx̂)−1/k . (2.5.7)
Therefore, the match asymptotic expansion of the differential equation is
−1/k
h κx i−1/k κ(1 − x)
y(x) ∼ λ − r + λ+ , (2.5.8)
r
(d) Can one construct a composite solution using the first terms?
Solution: Not exactly :)
(a) Find the first term in the expansion of the outer solution. Assume that this function
satisfies the boundary condition at x = 0.
y0 (1 − y0 )y00 − xy0 = 0.
Since we assume that it satisfies y(0) = 2, then y0 (x) 6= 0. Then it follows that
√
(1 − y0 )y00 = x =⇒ y0 (x) = 1 + 1 − x2 .
(b) Explain why there cannot be a boundary layer at x = 1, which links the boundary
condition at x = 1 with the outer solution of part 1. evaluated at x = 1.
(c) Assume that there is an interior layer at some point x0 , which links the outer solution
calculated in (a) for 0 ≤ x < x0 with
√ the outer solution y ∼ 0 for x0 < x < 1. From
the matching show that x0 = 3/2. Note that there will be an undetermined
constant.
Solution: In the interior layer, scale the coordinate as x̃ = (x − x0 )/α and set
Y (x̃) = y(x). Then one can achieve equation for the interior solution
Y 00 + Y (1 − Y )Y 0 = 0.
(d) Given the interior layer at x0 , construct the first term in the expansion of the inner
solution at x = 1.
Solution: In the similar fashion, setting ξ = (x − 1)/β and V (ξ) = y(x). Then
we get the same ODE with the left matching condition
V 00 + V (1 − V )V 0 = 0.
which models a linear oscillator with weak damping. This reduces to the linear oscillator model
when ε = 0.
Substituting (3.1.2) into (3.1.1) and collecting terms in equal powers of ε yields
y000 + y0 = 0
yn00 + yn = −yn−1
0
for n ≥ 1,
55
56 3.1. Introductory Example
y(t)
t
10 20 30 40 50 60 70
−1
Figure 3.1: Comparison between the regular asymptotic approximation (3.1.4) and the exact
solution (3.1.4) for ε = 0.1.
capture the correct behaviour of the exact solution. Indeed, (3.1.1) is a constant-coefficient
linear ODE and it can be solved exactly:
1 p
y(t) = p e−εt/2 sin t 1 − ε2 /4 (3.1.4)
1 − ε2 /4
It is clear that the exact solution decays but the first term in our regular asymptotic approx-
imation (3.1.3) does not. Also, we will pick up the secular terms if we naively expand the
exponential function around t = 0, since
εt ε2 t2
y(t) ≈ 1 − + + . . . sin(t).
2 8
t1 = t, t2 = εα t, α > 0,
where t2 is called the slow time-scale because it does not affect the asymptotic expansion until
εα t ∼ 1. We treat these two time-scales as independent variables and consequently the original
time derivative becomes
d dt1 ∂ dt2 ∂ ∂ ∂
−→ + = + εα . (3.1.5)
dt dt ∂t1 dt ∂t2 ∂t1 ∂t2
Method of Multiple Scales 57
Unlike the original problem, additional constraints are needed for (3.1.6) to have a unique
solution, and it is precisely this degree of freedom that allows us to eliminate the secular terms!
We now introduce an asymptotic expansion
∂t21 + 1 y0 = 0,
where a0 (0) = 1, b0 (0) = 0. Note that y0 (t1 , t2 ) consists of purely harmonic components with
slowly varying amplitude. We now need to determine α in the slow time-scale t2 . Observe that
for α > 1 the O(ε) equation is
∂t21 + 1 y1 = −∂t1 y0 ,
and the inhomogeneous term ∂t1 y0 will generate secular terms, since it belongs to the kernel
of homogeneous linear operator ∂t21 + 1 . More importantly, there is no way to generate non-
trivial solution that will cancel the secular term. This can be prevented by choosing α = 1.
The O(ε) equation is
Substituting y0 gives
with a1 (0) = b00 (0), b1 (0) = 0. We can choose the functions a0 , b0 to remove the secular terms,
which results in
and
2a00 + a0 = 0 =⇒ a0 (t2 ) = α0 e−t2 /2 = e−t2 /2 , since a0 (0) = 0.
Hence, a first term approximation of the solution y(t) of (3.1.1) is
y ∼ e−εt/2 sin(t).
One can prove that this asymptotic expansion is uniformly valid for 0 ≤ t ≤ O (1/ε).
3.1.3 Discussion
1. Many problems have the O(1) equation as
y000 + ω 2 y0 = 0.
If the original problem is nonlinear and the O(1) equation is as above, then it is usually
more convenient to use a complex representation of y0 , i.e.
These complex representations make identify the secular terms much easier.
yn00 + ω 2 yn = f (t).
A secular term arises if f (t) contains a solution of the O(1) problem, e.g. cos(ωt) or
sin(ωt). We can avoid secular terms by requiring the t2 -dependent coefficients of cos(ωt1 )
and sin(ωt1 ) to vanish. For example, there are no secular terms if
3. The time scales should be modified depending on the problem. Some possibilities include:
1 + ω1 ε + ω2 ε2 + . . . t, t2 = εt.
t1 =
| {z }
expansion of the effective frequency
t1 = εα t, t2 = εβ t, α < β.
Although we expect the leading-order term in the expansion to be O(ε), the solution can
become larger near a resonant frequency. Because it is not clear what amplitude the solution
actually reaches, we guess a general asymptotic expansion of the form
We also assume that β < 1 due to the resonance effect. Substituting (3.2.6) into (3.2.5) gives
h i h i
εβ ∂t21 y0 + 2ε 1+β
∂t1 ∂t2 y0 + εγ ∂t21 y1 1+β
+ . . . + ε λ∂t1 y0 + . . .
| {z }
| {z } | {z }
4 1 2
h i h i
+ εβ y0 + εγ y1 + . . . + ε1+3β κy03 + . . . = ε cos (t1 + εwt1 ) .
|{z} | {z } | {z }
1 2 3
∂t21 + 1 y0 = 0,
Thus, we can remove the secular terms sin(t1 + θ) and cos(t1 + θ) by requiring
Figure 3.2: Nullcline for φτ . (a) F (r, β) as a function of r with varying β. (b) Nullcline for φτ
with varying β. Parameter are given by γ = 0.75 and β = −βc , βc /2, βc , 1.5βc , respectively.
It remains to solve (3.2.7) with initial conditions A(0) = 0, θ(0) = −π/2 to find the ampli-
tude function
√ A(t2 ) and phase function θ(t2 ). For the analytic simplicity, changing variables
as r = κA/2 and φ = θ − wt2 gives
(
2r0 = −λr − γ2 sin φ,
γ (3.2.8)
2φ0 = β + 3r2 − 2r cos φ.
√
where γ = κ and β = −2ω. We now analyze the rewritten amplitude equation (3.2.8). The
nullcline for rτ is r = −γ sin θ/2λ. Similarly, nullcline for φτ is given by cos θ = 2r(β +3r2 )/γ ≡
F (r, β), see Fig. 3.2:
• If 0 > β > βc where minr F (r, βc ) = −1 (and it turns out that βc3 = −81γ 2 /16), then
there are two values of r for each cos θ in some interval (−z, 0) for some z ∈ [0, 1]. See
the red line.
• If β < βc , then two values of r exist for all cos θ between −1 and 0. See the purple line.
For 0 > β > βc , then the non-trivial fixed point (FB) stability of (3.2.8) with varying the
nullcline rτ for λ ≥ 0 is the following, see Fig. 3.3:
• For small λ, only one stable fixed point, see the curve A intersecting with the red line.
B,C If λ = λ1C , there is a SN bifurcation, that is, saddle and a stable FP. See the curve B
and B intersecting with the red line.
• At λ = λ2C , there is a second SN bifurcation in which saddle and other stable FP (from
A) annihilate leaning the stable FP (from B). See the curve D and E intersecting with
the red line.
62 3.3. Periodically Forced Nonlinear Oscillators
Figure 3.3: Non-trivial fixed points as a function of λ and its bifurcation diagram. (a) Inter-
sections of nullcline φτ and rτ with varying λ. (b) Bifurcation diagram of fixed radius rF P as
a function of λ.
3.3.1 Isochrones
Roughly speaking, the idea is to define the phase variable in such a way that it rotates uniformly
on the limit cycle as well as its neighbourhood. Suppose that we observe the unperturbed
system stroboscopically at time intervals of length ∆0 . This leads to a Poincaré mapping
u(t) −→ u(t + ∆0 ) ≡ G(u(t)).
The map G has all points on the limit cycle as fixed points. Choose a point U ∗ on the limit
cycle and consider all points in a neighbourhood of U ∗ in RM that are attracted to it under the
action of Φ. They form an (M − 1)-dimensional hypersurface I, called an isochrone, crossing
the limit cycle at U ∗ . A unique isochrone can be drawn through each point on the limit cycle
so we can parameterise the isochrones by the phase φ, i.e. I = I(φ). Finally, we extend the
definition of phase to the vicinity of the limit cycle by taking all points u ∈ I(φ) to have the
same phase, Φ(u) = φ, which then rotates at the natural frequency ω0 (in the unperturbed
case).
Example 3.3.1. Consider the following complex amplitude equation that arises for a limit
cycle oscillator close to a Hopf bifurcation:
dA
= (1 + iη)A − (1 + iα)|A|2 , A ∈ C.
dt
In polar coordinates A = Reiθ , we have
dR
= R(1 − R2 )
dt
dθ
= η − αR2 .
dt
Observe that the origin is unstable and the unit circle is a stable limit cycle. The solution for
arbitrary initial data R(0) = R0 , θ(0) = θ0 is
−1/2
1 − R02 −2t
R(t) = 1 + e
R0
α
θ(t) = θ0 + ω0 t − ln R02 + (1 − R02 )e−2t ,
2
where ω0 = η − α is the natural frequency of the stable limit cycle at R = 1. Strobing the
solution at time t = n∆0 , we see that
lim θ(n∆0 ) = θ0 − α ln R0 .
n→∞
Now consider the perturbed system (3.3.3) but with the “’unperturbed” definition of the phase:
M M
dΦ(u) X ∂Φ X ∂Φ
= fk (u) + εPk (u, t) = ω0 + ε Pk (u, t).
dt k=1
∂u k
k=1
∂u k
Because the sum is O(ε) and the deviations of u from the limit cycle U are small, to a
first approximation, we can neglect these deviations and calculate the sum on the limit cycle.
Consequently,
M
dΦ(u) X ∂Φ(U )
= ω0 + ε Pk (U , t).
dt k=1
∂u k
Finally, since points on the limit cycle are in one-to-one correspondence with the phase θ, we
obtain the closed phase equation
dφ
= ω0 + εQ(φ, t), (3.3.4)
dt
where
M
X ∂Φ(U (φ))
Q(φ, t) = Pk (U (φ), t) (3.3.5)
k=1
∂uk
is a 2π-periodic function of φ and a ∆-periodic function of t. The phase equation (3.3.4) de-
scribes the dynamics of the phase of a periodic oscillator in the presence of a small periodic
external force and Q(φ, t) contains all the information of the dynamical system. This is known
as the phase reduction method.
dx
= x − ηy − x2 + y 2 (x − ηy) + ε cos(ωt)
dt
dy
= y + ηy − x2 + y 2 (y + αx)
dt
where we periodically force the nonlinear oscillator in the x-direction. The isochrone is given
by y α
− ln x2 + y 2 ,
Φ = arctan
x 2
and differentiating with respect to x yields
∂Φ y αx
=− 2 − .
∂x x + y 2 x2 + y 2
Method of Multiple Scales 65
where ω = 2π/∆. Thus Q contains fast oscillating terms (compared to the time scale 1/ε)
together with slowly varying terms, the latter satisfy the resonance condition
kω0 + lω ≈ 0.
Substituting this double Fourier series into the phase equation (3.3.4), we see that the fast
oscillating terms lead to O(ε) phase deviations, while the resonant terms can lead to large
variations of the phase and are mostly important for the dynamics. Thus we have to average
the forcing term Q keeping only the resonant terms. We now identify the resonant terms using
the resonance condition above:
1. The simplest case is ω ≈ ω0 for which the resonant terms satisfy l = −k. This results in
an averaged forcing X
Q(φ, t) ≈ a−k,k eik(φ−ωt) = q(φ − ωt)
k
2. The other case is ω ≈ mω0 /n, where m and n are coprime. The forcing term becomes
X
Q(φ, t) ≈ a−nk,mk eik(mφ−nωt) = qb(mφ − nωt)
k
dψ
= −∆ω + εQ(ψ + ωt, t) = O(ε).
dt
Define Z T
1
q(ψ) = lim Q(ψ + ωt, t) dt,
T −→∞ T 0
Method of Multiple Scales 67
N
dui X
= f (ui ) + ε aij H(uj ), i = 1, . . . , N, (3.3.7)
dt j=1
where the first term represents the local autonomous dynamics and the second term describes
the interaction between oscillators. In a similar fashion to a single periodically forced oscillator,
we can write down the phase equation:
N
!
dφi (ui ) ∂φi X
= ω0 + ε · aij H(uj ) , i = 1, . . . , N. (3.3.8)
dt ∂ui j=1
N
dφi X
= ω0 + ε aij Qi (φi , φj ), i = 1, . . . , N, (3.3.9)
dt j=1
where
∂φi
Qi (φi , φj ) = (U (φi )) · H(U (φj )). (3.3.10)
∂ui
The final step is to use the method of averaging to obtain the phase-difference equation.
Introducing ψi = φi − ω0 t, we obtain
N
dψi X
=ε aij Qi (ψi + ω0 t, ψj + ω0 t) .
dt j=1
N
dψi X
=ε aij h(ψj − ψi ), (3.3.11)
dt j=1
where
N
!
Z ∆0
1 X
h(ψj − ψi ) = R(ψi + ω0 t) · H(U (ψj + ω0 t)) dt
∆0 0 j=1
N
!
Z 2π
1 X
= R(φ + ψi − ψj ) · H(U (φ)) dφ,
2π 0 j=1
Phase-locked solutions
We define a one-to-one phase-locked solutions to be
ψi (t) = t∆w + ψ i , (3.3.12)
where ψ i is constant. Taking time derivative on (3.3.12) and imposing (3.3.11) yields
N
X
∆ω = ε aij h ψ j − ψ i , i = 1, . . . , N. (3.3.13)
j=1
Since we have N equations in N unknowns ∆ω and N − 1 phases ψ j − ψ 1 , then one can find
the phase-locked solutions (We only care about phase difference.)
Stability
In order to determine local stability, we set
ψi (t) = ψ i + t∆ω + ∆ψi (t). (3.3.14)
To linearize it, taking time derivative on (3.3.14) and imposing the phase-locked solutions
(3.3.13) gives
N
d∆ψi X
=ε H
b ij (Φ) ∆ψj , (3.3.15)
dt j=1
where Φ = ψ̄1 , . . . ψ̄N and
X
Hb ij (Φ) = aij h ψ j − ψ i − δij aik h ψ k − ψ i . (3.3.16)
k
Similar to the weakly damped oscillator, we introduce two separate time scales t1 = t, t2 = εt.
In this case, (3.4.1) becomes
h i h i
∂x2 u = ∂t21 + 2ε∂t1 ∂t2 + ε2 ∂t22 u + ε ∂t1 + ε∂t2 u, (3.4.2a)
u = 0 at x = 0 and x = 1, (3.4.2b)
h i
u(x, 0) = g(x), ∂t1 + ε∂t2 u = 0. (3.4.2c)
t1 =t2 =0
As before, the solution of (3.4.2) is not unique and we will use this degree of freedom to
eliminate the secular terms.
We try a regular asymptotic expansion of the form
The initial conditions will be imposed once we determine an (t2 ) and bn (t2 ). The O(ε) equation
is
where
An = (2a0n + an ) λn cos(λn t1 ) − (2b0n + bn ) λn sin(λn t1 ).
Given the zero boundary conditions in (3.4.1), it is appropriate to introduce the Fourier ex-
pansion
∞
X
u1 = Vn (t1 , t2 ) sin(λn x).
n=1
2a0n + an = 0, 2b0n + bn = 0,
It describes the motion of a string on an elastic foundation as well as the waves in a cold
electron plasma.
As usual, let us consider (3.4.8) with ε = 0:
−k 2 u
b = ∂tt u
b+ub, (3.4.10a)
u
b(k, 0) = Fb(k), ∂t u
b(k, 0) = G(k).
b (3.4.10b)
Solving (3.4.10) and applying the inverse Fourier transform we obtain the general solution of
(3.4.9): Z ∞ Z ∞
i(kx−ω(k)t)
u(x, t) = A(k)e dk + B(k)ei(kx+ω(k)t dk, (3.4.11)
−∞ −∞
72 3.4. Partial Differential Equations
where A(k) and B(k) are determined from the initial conditions in (3.4.10). This shows that
the solution
of (3.4.9)
can be written as the superposition of the plane wave solutions u± (x, t) =
exp i(kx + ω(k)t) . We would like to investigate how the nonlinearity affects a right-moving
√
plane wave u(x, t) = cos(kx − ωt), where k > 0 and ω = 1 + k 2 .
A regular asymptotic expansion of the form
will lead to secular terms, and thus we use multiple scales to find an asymptotic approximation
of the solution of (3.4.8). We take three independent variables
where we use the dispersion relation −k 2 = −ω 2 +1. We assume a regular asymptotic expansion
of the form
u(x, t) ∼ u0 (θ, x2 , t2 ) + εu1 (θ, x2 , t2 ) + . . . .
The O(1) equation is
∂θ2 + 1 u0 = 0,
These constitute the amplitude-phase equations and can be solved using characteristic
coordinates. Specifically, let
∂r A = 0
3
∂r φ = − A2
16ωk
and solving this yields
3
A = A(s) and φ = − A2 r + φ0 (s).
16ωk
Hence, a first term approximation of the solution of (3.4.8) is
3 2
u ∼ A(wx2 − kt2 ) cos (kx − ωt) − (ωx2 + kt2 )A + φ0 (ωx2 − kt2 ) . (3.4.14)
16ωk
We can now attempt to answer our main question: how does the nonlinearity affects the
plane wave solution? Consider the plane wave initial conditions
We see that the nonlinearity increases the phase velocity since it increases from c = ω/k to
c=ωb /k.
1
f (a) = , (3.5.1b)
1 + exp(−η(a − k))
where a(θ, t) denotes the activity at time t of a local population of cells at position θ ∈ [0, π),
w(θ−θ0 ) is the strength of synaptic weights between cells at θ0 and θ and the firing rate function
f is a sigmoid function. Assuming w is an even π-periodic function, it can be expanded as a
Fourier series: X
w(θ) = W0 + 2 Wn cos(2πθ), Wn ∈ R. (3.5.2)
n≥1
Since the linear operator L is compact on L2 (S 1 ), it has a discrete spectrum with eigenvalues
λn = −1 + f 0 (ā)Wn , n ∈ Z,
These are obtained by integrating the eigenvalue equation against cos(2nθ) over [0, π]: CHT:
Check this again, unsure about this
π π
f 0 (ā)
Z Z
0 0 0
λn an = −an + w(θ − θ )a(θ ) dθ cos(2nθ) dθ
π 0 0
!
π π
f 0 (ā)
Z Z X
0 0
= −an + a(θ ) Wm cos(2m(θ − θ )) cos(2nθ) dθ dθ0
π 0 0 m∈Z
0 XZ π Z π
f (ā)
= −an + Wm a(θ0 ) cos(2nθ0 ) dθ0 cos(2mθ) cos(2nθ) + sin(2mθ) cos(2nθ) dθ
π m∈Z 0 0
f 0 (ā) X hπ i
= −an + Wm am δ±m,n
π m∈Z 2
f 0 (ā) h i
= −an + Wn an + W−n a−n
2
0
= −an + f (ā)Wn an ,
where Z π
an = a(θ) cos(2nθ) dθ = a−n .
0
Method of Multiple Scales 75
The eigenvalue expression reveals the bifurcation parameter µ = f 0 (ā). For sufficiently
small µ, corresponding to a low activity state, λn < 0 for all n and the fixed point is stable.
As µ increases beyond a critical value µc , the fixed point becomes unstable due to excitation
of the eigenfunctions associated with the largest Fourier component of w(θ). Suppose that
W1 = maxm Wm . Then λn > 0 for all n if and only if
1
1 < µWn ≤ µW1 =⇒ µ > = µc .
W1
Consequently, for µ > µc , the excited modes will be
where z = |z|e−2iθ0 . We expect this mode to grow and stop at a maximum amplitude as µ
approaches µc , mainly because of the saturation of f .
If ∆µ = O(1), then µ − µc = O(ε) and we can carry out a perturbation expansion in powers
of ε. We first Taylor expand the nonlinear function f around a = ā:
The dominant temporal behaviour just beyond bifurcation is the slow growth of the excited
mode eε∆µt and this motivates the introduction of a slow time scale τ = εt. Substituting
(3.5.5), (3.5.6) and (3.5.7) into (3.5.1) yields
h ih √ i
∂t + ε∂τ ā + εa1 + εa2 + ε3/2 a3 + . . .
h √ i
= − ā + εa1 + εa2 + ε3/2 a3 + . . .
1 π
Z
+ w(θ − θ0 )f (ā) dθ0
π 0
1 π h√
Z i
+ w(θ − θ0 ) µc + ε∆µ εa1 + εa2 + ε3/2 a3 + . . . dθ0
π 0
1 π h√
Z i2
+ w(θ − θ0 )g2 εa1 + εa2 + . . . dθ0
π 0
1 π h√
Z i3
+ w(θ − θ )g3 εa1 + εa2 + . . . dθ0
0
π 0
76 3.5. Pattern Formation and Amplitude Equations
Collecting terms with equal powers of ε then leads to a hierarchy of equations of the form:
ā = W0 f (ā)
La
b 1=0
b 2 = V2 := −g2 w ∗ a2
La 1
Using integration by parts, it is easy to see that L b is self-adjoint with respect to this particular
±2iθ
inner product and since Lã
b = 0 for ã = e , we have
hã, La
b n i = hLã,
b an i = 0.
Since La
b n = Vn , it follows from the Fredholm alternative that the set of solvability conditions
are
hã, Vn i = 0 for n ≥ 2.
The O(ε) solvability condition hã, V2 i = 0 is automatically satisfied. The O(ε3/2 ) solvability
condition can be expanded into
To deal with the convolution terms, observe that since w is even, for any function b(θ) we have
he2iθ , w ∗ bi = hw ∗ e2iθ , bi
1 π 1 π
Z Z
0 −2iθ0 0
= w(θ − θ )e dθ b(θ) dθ
π 0 π 0
Method of Multiple Scales 77
!
Z π Z π
1 1 X
0 −2iθ0 0
= 2Wn cos(2n(θ − θ ))e dθ b(θ) dθ
π 0 π 0 n≥1
!
Z π Z π
1 1 X
2in(θ−θ0 ) −2in(θ−θ0 )
−2iθ0
= Wn e +e e dθ0 b(θ) dθ
π 0 π 0 n≥1
Z π Z π
1 1 −2iθ
= W1 e dθ0 b(θ) dθ
π 0 π 0
1 π
Z
= W1 e−2iθ b(θ) dθ
π 0
= W1 he2iθ , bi.
Set W1 = µ−1
c . From the identity above we then have
and
µc π
Z
−La2 = a2 −
b w(θ − θ0 )a2 (θ0 ) dθ0
π 0
g2 π
Z
= w(θ − θ0 )a21 (θ0 ) dθ0
π 0
g2 π n
Z oh i
0 0 0 0
X
= W0 + Wn e2in(θ−θ ) + e−2in(θ−θ ) z 2 e4iθ + 2|z|2 + (z ∗ )2 e−4iθ dθ0
π 0 n≥1
h i
= g2 2|z|2 W0 + z 2 W2 e4iθ + (z ∗ )2 W2 e−4iθ . (3.5.12)
Let
a2 (θ) = A+ e4iθ + A− e−4iθ + A0 + ζa1 (θ). (3.5.13)
The constant ζ remains undetermined at this order of perturbation but does not appear in the
amplitude equation for z(τ ). Substituting (3.5.13) into (3.5.12) yields
g2 z 2 W2 g2 (z ∗ )2 W2 2g2 |z|2 W0
A+ = , A− = , A0 = . (3.5.14)
1 − µc W2 1 − µc W 2 1 − µc W0
78 3.6. Problems
Consequently,
Finally, substituting (3.5.9), (3.5.10), (3.5.11) and (3.5.16) into the O(ε3/2 ) solvability condition
(3.5.8), we obtain the Stuart-Landau equation
dz
= z(∆µ − Λ|z|2 ), (3.5.17)
dτ
where
W2 2W0
Λ = −3g3 − 2g22 + . (3.5.18)
1 − µc W2 1 − µc W0
Note that we also absorbed a factor of µc into τ .
3.6 Problems
1. Find a first-term expansion of the solution of the following problems using two time
scales.
The secular terms are eliminated provided F (τ ) = 0. Writing A(τ ) = R(τ )eiθ(τ ) ,
F (τ ) becomes
2 Rτ e + iRθτ e + 3Reiθ R2 = 0,
iθ iθ
or
2 Rτ + iRθτ + 3R3 = 0.
Consequently, we have
θτ = 0 =⇒ θ(τ ) = θ0
and
2Rτ 1 1
2Rτ + 3R3 = 0 =⇒ = −3 =⇒ = 3τ + C =⇒ R(τ ) = √ .
R3 R2 3τ + C
Therefore, (3.6.2) becomes
which means
i 1 3π
R(0)eiθ0 = − = √ eiθ0 =⇒ C = 4 and θ0 = .
2 C 2
It can be easily seen from the initial conditions of the O(1) problem that A(0) =
0, and so Y0 ≡ 0. Before we proceed any further, note that
Imposing the initial condition Y1 (0, 0) = 0 and (∂T Y1 +∂τ Y0 )(0, 0) = ∂T Y1 (0, 0) =
1, we obtain
B(0) + B ∗ (0) = 0
h i
i B(0) − B ∗ (0) = 1,
√
which gives B(0) = −i/2. Hence, the O( ε) solution is
Substituting (3.6.7) into (3.6.6) and expanding both sin(φ) and cos(φ) around φ = φ0
we obtain:
h i
∂t2 + 2ε∂t ∂τ + ε2 ∂τ2 φ0 + εφ1 + ε2 φ2 + . . .
h i
2
∂t + ε∂τ φ0 + εφ1 + ε2 φ2 + . . .
+ ε 1 + γ cos φ0 − sin φ0 εφ1 + ε φ2 + . . .
+ sin φ0 + cos φ0 εφ1 + ε2 φ2 + . . . = εα,
φ0 + εφ1 + ε2 φ2 + . . . (0, 0) = 0
h i
∂t + ε∂τ φ0 + εφ1 + ε2 φ2 + . . . (0, 0) = 0.
To solve this nonlinear problem ,we approximate sin φ0 ≈ φ0 and the general solution
of the problem is approximately
∂t2 φ1 + φ1 = −2∂t ∂τ φ0 − (1 + γ) ∂t φ0 + α
= −2 (−A0 sin(t) + B 0 cos(t)) − (1 + γ) (−A sin t + B cos(t)) + α
= [2A0 + A(1 + γ)] sin(t) − [2B 0 + B(1 + γ)] cos(t) + α.
The secular terms are eliminated provided the coefficients of cos(t) and sin(t) vanish.
This yields two initial value problems
(
2A0 + A(1 + γ) = 0, A(0) = 0
0
2B + B(1 + γ) = 0, B(0) = 0
which has solutions A(τ ) = B(τ ) ≡ 0. It follows from (3.6.8) that φ0 ≡ 0 and we
need to investigate the O(ε2 ) problem. The general solution of the O(ε) problem is
The secular terms are eliminated provided the coefficients of cos(t) and sin(t) vanish.
This yields two initial value problems
(
2C 0 + C(1 + γ) = 0, C(0) = −α
0
2D + D(1 + γ) = 0, D(0) = 0
Since the first two terms belongs to the kernel of the homogeneous operator, the
corresponding particular solution has the form F (τ ) and G(τ )te−t and only the first
one blows up as t −→ ∞, since
G(τ )te−t −→ 0 as t −→ ∞.
∂τ A = A − A2 . (3.6.11)
A phase-plane analysis shows that the system (3.6.11) has an unstable fixed point
at A = 0 and a stable fixed point at A = 1. Thus, we conclude that A(τ ) −→ 1 as
τ −→ ∞, provided A(0) > 0.
Let y = ẋ.
(a) Show that if E(t) = E(x(t), y(t)) = (x(t)2 + y(t)2 )/2, then
Therefore
Ė(x, y) = xẋ + y ẏ
= xy + y (−x − εf (x, y))
= −εf (x, y)y.
x ∼ x0 + εx1 + . . .
E ∼ E0 + εE1 (t) + . . . .
dE1
= −f (x, y)y = −f (x0 + εx1 + . . . , x˙0 + εx˙1 + . . . ) x˙0 + εx˙1 + . . .
dt
= −f (x0 , x˙0 )x˙0 + O(ε).
Therefore, to O(1),
Z t
E1 (t) = E1 (0) − f (x0 (τ ), x˙0 (τ ))x˙0 (τ ) dτ
0
Z t
= E1 (0) + A0 f (A0 cos(τ ), −A0 sin(τ )) sin(τ ) dτ.
0
(b) Suppose that the periodicity condition on part (a) does not hold. Let En =
E(x(2πn), y(2πn)). Show that to lowest order En satisfies a difference equation
of the form
En+1 = En + εF (En ),
with Z 2π p p p
F (En ) = 2En f 2En cos τ, − 2En sin τ sin τ dτ.
0
√
Hint: Take x ∼ A cos t with A = 2E slowly varying over a single period of length
2π.
Solution: Since
A20
E(t) ∼ E0 + εE1 (t) ∼ ,
2
we have p
A0 (t) ≈ 2E(t) + O(ε).
From part (a), we then have
√
(c) Hence, deduce that a periodic orbit with approximate amplitude A∗ = 2E ∗ exists
if F (E ∗ ) = 0 and this orbit is stable if
dF ∗
ε (E ) < 0.
dE
Hint: Spiralling orbits close to the periodic orbit x = A∗ cos(t) + O(ε) can be ap-
proximated by a solution of the form x = A cos(t) + O(ε).
(d) Using the above result, find the approximate amplitude of the periodic orbit of the
Van der Pol equation
ẍ + x + ε(x2 − 1)ẋ = 0
and verify that it is stable.
with Γ = O(1) and ω 6= 1/3, 1, 3. Use the method of multiple scales to show that the
solution is attracted to
Γ
x(t) = cos(ωt) + O(ε)
1 − ω2
when Γ2 ≥ 2(1 − ω 2 )2 and
1/2
Γ2
Γ
x(t) = 2 1 − cos t + cos(ωt) + O(ε)
2(1 − ω 2 )2 1 − ω2
when Γ2 < 2(1 − ω 2 )2 . Explain why this result breaks down when ω = 1/3, 1, 3.
Solution: Introducing the slow scale τ = εt and substituting the asymptotic expan-
sion
x ∼ x0 (t, τ ) + εx1 (t, τ ) + . . .
into the Van der Pol equation we obtain
h i
2 2
∂t +2ε∂t ∂τ + ε ∂τ x0 + εx1 + . . . + x0 + εx1 + . . .
h 2 ih i
+ ε x0 + εx1 + . . . − 1 ∂t + ε∂τ x0 + εx1 + . . . = Γ cos(ωt).
Γ
−ω 2 B + B = Γ =⇒ B = = δ.
1 − ω2
Thus the general solution of the O(1) equation is
A3 A3 h i
sin(2Ω) cos(Ω) = sin(3Ω) + sin(Ω)
2 4
Aδ 2 h i
Aδ 2 sin(Ω) cos2 (ωt) = sin(Ω) 1 + cos(2ωt)
2
Aδ 2 h i
= sin(Ω) + sin(Ω) cos(2ωt)
2
Aδ 2 h i
= 2 sin(Ω) + sin(Ω + 2ωt) + sin(Ω − 2ωt)
4
Method of Multiple Scales 89
A2 δ h i
A2 δ sin(2Ω) cos(ωt) = sin(2Ω + ωt) + sin(2Ω − ωt)
2
A2 δω h i
A2 δω cos2 (Ω) sin(ωt) = sin(ωt) 1 + cos(2Ω)
2
2
A δω h i
= sin(ωt) + sin(ωt) cos(2Ω)
2
A2 δω h i
= 2 sin(ωt) + sin(ωt + 2Ω) + sin(ωt − 2Ω)
4
2
A δω h i
= 2 sin(ωt) + sin(2Ω + ωt) − sin(2Ω − ωt)
4
δ3ω δ3ω h i
sin(2ωt) cos(ωt) = sin(3ωt) + sin(ωt)
w 4
Aδ 2 ω h i
Aδ 2 ω cos(Ω) sin(2ωt) = sin(2ωt + Ω) + sin(2ωt − Ω)
2
Aδ 2 ω h i
= sin(Ω + 2ωt) − sin(Ω − 2ωt) .
2
Combining everything, the O(ε) equation takes the form
A3 Aδ 2
3
2
h i A
∂t + 1 x1 = 2Aθτ cos(Ω) + 2Aτ − A + + sin(Ω) + sin(3Ω)
4 2 4
A2 δω δ 3 ω
3
δ ω
+ −δω + + sin(ωt) + sin(3ωt)
2 4 4
2
Aδ 2 ω
2
Aδ 2 ω
Aδ Aδ
+ + sin(Ω + 2ωt) + − sin(Ω − 2ωt)
4 2 4 2
2
A δ A2 δω
2
A δ A2 δω
+ + sin(2Ω + ωt) + − sin(2Ω − ωt).
2 4 2 4
Terms of the form sin(ωt) and sin(3ωt) will be resonant if ω = 1, 1/3. Terms of the
form sin(Ω ± 2ωt) will be resonant if
t ± 2ωt = (1 ± 2ω)t = ±t ⇐⇒ ω = 0, 1.
2t ± ωt = (2 ± ω)t = ±t ⇐⇒ ω = 1, 3.
Therefore, if we assume that ω 6= 1/3, 1, 3 then the only resonant terms on the right-
hand side of the O(ε) equation are those involving cos(Ω) and sin(Ω) and we require
their coefficients to vanish, i.e.
A3 Aδ 2
2Aθτ = 0 and 2Aτ = A − −
4 2
2
δ2
A
=A 1− −
4 2
90 3.6. Problems
A2 δ2
= −A − C(δ) , where C(δ) = 1 − .
4 2
(a) When C(δ) < 0, that is, δ 2 > 2, Aτ has only one fixed point at A = 0 and this
is asymptotically stable, i.e. A(τ ) −→ 0 as τ −→ ∞ for any initial conditions
A(0). Therefore, the solution is attracted to
Γ
x(t) = δ cos(ωt) + O(ε) = cos(ωt) + O(ε).
1 − ω2
p
(b) When C(δ) > 0, that is, δ 2 < 2, Aτ has three fixed points A0 = 0, ±2 C(δ) and
a phase-plane analysis shows that the fixed point A0 = 0 becomes unstable and
the other two are stable. Therefore, as τ −→ ∞ we have
( p
2 C(δ) if A(0) > 0,
A(τ ) −→ p
−2 C(δ) if A(0) < 0.
6. Multiple scales with nonlinear wave equations. The Korteweg-de Vries (KdV)
equation is
ut + ux + αuux + βuxxxx = 0, x ∈ R, t > 0,
where α, β are positive real constants and u(x, 0) = εf (x) for 0 < ε 1.
(a) Let θ = kx − ωt and seek traveling wave solutions using an expansion of the form
where ω = k − βk 3 and k > 0 is a constant. Show that this can lead to secular
terms.
Solution:
(b) Use multiple scales (variables θ, εx, εt) to eliminate the secular terms in part (a) and
find a first-term expansion. In the process, show that f (x) must have the form
f (x) = A cos(kx + φ)
for constants A, B, φ in order to generate a traveling wave? Hint: Use the fact that
f (x) is independent of ε.
Method of Multiple Scales 91
Solution:
92 3.6. Problems
Chapter 4
The Wentzel-Kramers-Brillouin
(WKB) Method
The WKB method, named after Wentzel, Kramers and Brillouin, is a method for finding
approximate solutions to linear differential equations with spatially varying coefficients. The
origin of WKB theory dates back to 1920s where it was developed by Wentzel, Kramers and
Brillouin to study time-independent Schrodinger equation. This often arises from the following
problem:
d2 y
− q(εx)y = 0,
dx2
with the slowly varying potential energy. To handle such problem, the WKB method introduces
an ansatz of the expansion term as a product of slowly varying and exponenetially rapidly
varying terms.
and the solution either blows up (q > 0) or oscillates (q < 0) rapidly on a scale of O(ε). The
hypothesis of the WKB method is that this exponential solution can be generalised to obtain
an approximate solution of the full problem (4.1.1).
We start with the following general WKB ansatz:
α
y(x) ∼ eθ(x)/ε [y0 (x) + εα y1 (x) + . . . ] as ε −→ 0 (4.1.2)
for some α > 0. Here, we assume that the solution varies exponentially with respect to the
fast variation. From (4.1.2) we obtain:
α
y 0 ∼ ε−α θx y0 + y00 + θx y1 + . . . eθ/ε
(4.1.3a)
00
−2α 2 −α 0 2
θ/εα
y ∼ ε θx y0 + ε θxx y0 + 2θx y0 + θx y1 + . . . e (4.1.3b)
93
94 4.1. Introductory Example
Figure 4.1: An example of turning points: quantum tunneling. Depending on effective potential
energy, the solutions have different behavior and need to be matched (Taken from Wikipedia
Commons).
α
y 000 ∼ ε−3α θx3 y0 + ε−2α θx 3θx y00 + 3θxx y0 + θx2 y1 + . . . eθ/ε
(4.1.3c)
α
y 0000 ∼ ε−4α θx4 y0 + ε−3α θx2 6θxx y0 + 4θx y00 + θx2 y1 + . . . eθ/ε
(4.1.3d)
Substituting both (4.1.2) and (4.1.3) into (4.1.1) and cancelling the exponential term yield
2
2 θx y0 1 0 2
+ α θxx y0 + 2θx y0 + θx y1 + . . . − q(x) [y0 + εα y1 + . . . ] = 0.
ε (4.1.4)
ε2α ε
To determine y0 (x), we need to solve the O(ε) equation which is the transport equation:
The y1 terms cancel out due to the eikonal equation (4.1.5) and (4.1.7) reduces to
y00 θxx
=−
y0 2θx
1
ln |y0 | = − ln |θx | + C
2 p
ln |y0 | = − ln |θx | + C
The Wentzel-Kramers-Brillouin (WKB) Method 95
C
y0 (x) = √ = Cq(x)−1/4 ,
θx
where C is an arbitrary nonzero constant and the last line follows from (4.1.6). Hence, a
first-term asymptotic approximation of the general solution of (4.1.1) is
1 xp
Z Z x
−1/4 1 p
y(x) ∼ q(x) a0 exp − q(s) ds + b0 exp q(s) ds , (4.1.9)
ε ε
where a0 , b0 are arbitrary constants, possibly complex. It is evident that (4.1.9) is valid if
q(x) 6= 0 on [0, 1]. The x-values where q(x) = 0 are called turning points and this nontrivial
issue will be addressed in Section 4.2.
Example 4.1.1. Choose q(x) = −e2x . Then the WKB approximation (4.1.9) is
x x
y(x) ∼ e−x/2 a0 e−ie /ε + b0 eie /ε = e−x/2 [α0 cos(λex ) + β0 sin(λex )] ,
To measure the error of the WKB approximation (4.1.9), we look at the O(ε2 ) equation
which has the form
The y2 terms vanish due to the eikonal equation (4.1.5) and so (4.1.10) reduces to
Because the first two terms of (4.1.11) are similar to the transport equation (4.1.8), we make
an ansatz y1 (x) = y0 (x)w(x). (4.1.11) reduces to
Suppose q(x) > 0 so that θx is a real-valued function. Rearranging (4.1.12) in terms of w0 and
integrating by parts with respect to x we obtain
2θx y0 w0 = −y000
2Cθx w0 d2
C d Cθxx
√ =− 2 √ =
θx dx θx dx 2θx3/2
0 1 d θxx 1
w = √
4 dx θx3/2 θx
Z x
1 d θxx 1
w(x) = √ ds
4 dx θx3/2 θx
1 x θxx
Z
1 θxx d 1
=d+ − 3/2
√ ds
4 θx2 4 θx dx θx
1 x θxx
Z 2
1 θxx
=d+ + ds,
4 θx2 8 θx3
where d is an arbitrary constant. On the other hand, θx is a complex-valued function if q(x) < 0,
√
i.e. θx = ±i −q. We then have
i −qx iqx
θxx =± √ =∓ √
2 −q 2 −q
θxx iqx iqx
2
=∓ √ =±
θx 2q −q 2(−q)3/2
−qx2 q2
2
θxx = = x
4(−q) 4q
√ 3
θx3 = (±i)3 −q = ∓i (−q)3/2
2
θxx qx2 iqx2
= =∓ .
θx3 ∓4iq(−q)3/2 4(−q)5/2
The Wentzel-Kramers-Brillouin (WKB) Method 97
Consequently,
Z x 2
1 qx 1 qx p
d+ + ds if θx (x) = q(x),
3/2
8q 32 q 5/2
Z x 2
1 qx 1 qx
p
d − − if θx (x) = −
3/2
ds q(x),
8q 32 q 5/2
w(x) = Z x
iqx2
1 iqx 1 p
d+ 3/2
− 5/2
ds if θx (x) = i −q(x),
8 (−q) 32 (−q)
Z x 2
1 iq x 1 iq x
p
d − + ds if θx (x) = −i −q(x).
8 (−q)3/2 32 (−q)5/2
Finally, for small ε the WKB ansatz (4.1.2) is well-ordered provided
|εy1 (x)| |y0 (x)|, or |εw(x)| 1.
In terms of the function q(x) and its first derivatives, for x ∈ [x0 , x1 ] we will have an accurate
approximation if Z x1
1 qx qx
ε |d| + 3/2
4+ dx 1,
32 q
x0
q
where | · | := k · k∞ over the interval [x0 , x1 ]. We stress that this condition holds if the interval
[x0 , x1 ] does not contain a turning point.
Remark 4.1.2. The constants a0 , b0 in (4.1.9) and d in w(x) are determined from boundary
conditions. However, it is very possible that these constants depend on ε. It is therefore
necessary to make sure this dependence does not interfere with the ordering assumed in the
WKB ansatz (4.1.2).
d 1 d
= β
dx ε dx̃
gives the inner equation
ε2−2β Y 00 − εβ x̃qt0 + . . . Y = 0,
(4.2.3)
where qt0 := q 0 (xt ). Balancing leading-order terms in (4.2.3) means we require
2
2 − 2β = β =⇒ β = .
3
Since it is not clear what the asymptotic sequence should be, we take the asymptotic
expansion to be
Y ∼ εγ Y0 (x̃) + . . . . (4.2.4)
The O(ε2/3 ) equation is
Y000 − x̃qt0 Y0 = 0, −∞ < x̃ < ∞. (4.2.5)
Performing a coordinate transformation s = (qt0 )1/3 x̃, (4.2.5) becomes the Airy’s equation:
d2 Y0
− sY0 = 0, −∞ < s < ∞, (4.2.6)
ds2
and this can be solved either using power series expansion or Laplace transform. The general
solution of (4.2.6) is
Y0 (s) = aAi(s) + bBi(s), (4.2.7)
where Ai(·) and Bi(·) are Airy functions of the first and the second kinds respectively. It is
well-known that
∞
1 X 1 k+1 2π k
Ai(x) = 2/3 Γ sin (k + 1) 31/3 x
3 π k=0 k! 3 3
1 3 0 1 4
= Ai(0) 1 + x + . . . + Ai (0) x + x + . . .
6 12
iπ/6 2πi/3 −iπ/6 −2πi/3
Bi(x) = e Ai xe +e Ai xe
1 3 0 1 4
= Bi(0) 1 + x + . . . + Bi (0) x + x + . . . ,
6 12
The Wentzel-Kramers-Brillouin (WKB) Method 99
Figure 4.2: Plot of the two Airy functions (Taken from Wikipedia Commons).
2
where Γ(·) is the gamma function. Setting ξ = |x|3/2 , we also have that
3
1 π 5 π
√π|x|1/4 cos ξ − 4 + 72ξ sin ξ − 4 if x −→ −∞,
Ai(x) ∼ (4.2.8a)
1 −ξ 5
√
e 1− ξ if x −→ +∞,
2 π|x|1/4 72
1 π 5 π
√π|x|1/4 cos ξ + 4 + 72ξ sin ξ + 4 if x −→ −∞,
Bi(x) ∼ (4.2.8b)
1 ξ 5
√
e 1+ ξ if x −→ +∞.
π|x|1/4 72
4.2.2 Matching
From (4.2.7), the general solution of (4.1.1) in the transition layer is
h i h i
0 1/3 0 1/3
Y0 (x̃) = aAi (qt ) x̃ + bBi (qt ) x̃ . (4.2.9)
We now have 6 undetermined constants from (4.2.2) and (4.2.9), but these are all connected
since the inner solution (4.2.9) must match the outer solutions (4.2.2). These will results in
two arbitrary constants in the general solution (4.2.1). Since the inner solution is unbounded,
we introduce an intermediate variable
x − xt 2
xη = η
, 0<η< ,
ε 3
where the interval for η comes from the requirement that the scaling for the intermediate vari-
able must lie between the outer scale, O(1) and the inner scale, O(ε2/3 ).
100 4.2. Turning Points
Y ∼ εγ Y0 εη−2/3 xη + . . .
h i
1/3 1/3
∼ εγ aAi (qt0 ) εη−2/3 xη + bBi (qt0 ) εη−2/3 xη + . . .
∼ εγ [aAi(r) + bBi(r)] + . . .
γ a 2 3/2 b 2 3/2
∼ε √ exp − r + √ 1/4 exp r , (4.2.10)
2 πr1/4 3 πr 3
where r = q 0 (xt )1/3 εη−2/3 xη > 0 and the last line follows from (4.2.8). On the other hand, since
Z x p Z xt +εη xη p
q(s) ds ∼ (s − xt )qt0 ds
xt xt
xt +εη xη
0 2
p 3/2
= qt (s − xt )
3
xt
2
qt0 (εη xη )3/2
p
=
3
2
= εr3/2
3
and
−1/4 −1/4 −1/6
q(x)−1/4 ∼ [q(xt ) + (x − xt )qt0 ] = [εη xη qt0 ] = ε−1/6 (qt0 ) r−1/4 ,
ε−1/6
2 3/2 2 3/2
yR ∼ aR exp − r + bR exp r . (4.2.11)
(qt0 )1/6 r1/4 3 3
Consequently, matching (4.2.10) the right outer solution yR with (4.2.11) the inner solution Y
yields the following:
1 a 1/6 b 1/6
γ = − , aR = √ (qt0 ) , bR = √ (qt0 ) . (4.2.12)
6 2 π π
Y ∼ εγ [aAi(r) + bBi(r)] + . . .
The Wentzel-Kramers-Brillouin (WKB) Method 101
γ a 2 3/2 π b 2 3/2 π
∼ε √ cos |r| − +√ cos |r| + .
π|r|1/4 3 4 π|r|1/4 3 4
Using the identity cos θ = (eiθ + e−iθ )/2, a more useful form of the inner expansion Y as
r −→ −∞ is
εγ h
−iπ/4 iπ/4
iζ iπ/4 −iπ/4
−iζ i
Y ∼ √ ae + be e + ae + be e , (4.2.13)
2 π|r|1/4
2
where ζ = |r|3/2 . On the other hand, since
3
Z xt p Z xt p
q(s) ds ∼ (s − xt )qt0 ds
x η
xt +ε xη
xt
0 2
p 3/2
= qt (s − xt )
3
xt +εη xη
2
qt0 (εη xη )3/2
p
=−
3
2
= − ε|r|3/2 (−1)3/2
3
2
= iε|r|3/2 ,
3
and
−1/4 −1/6
q(x)−1/4 ∼ [εη xη qt0 ] = ε−1/6 (qt0 ) |r|−1/4 (−1)−1/4
−1/6
= ε−1/6 (qt0 ) |r|−1/4 e−iπ/4 ,
ε−1/6 e−iπ/4
aL e−iζ + bL eiζ .
yL ∼ (4.2.14)
(qt0 )1/6 |r|1/4
Consequently, matching (4.2.14) the left outer solution yL with (4.2.13) the inner solution Y
yields the following:
bR i
aL = iaR + , b L = aR + b R (4.2.16)
2 2
or in matrix form
aL i 1/2 aR
= . (4.2.17)
bL 1 i/2 bR
102 4.2. Turning Points
4.2.5 Conclusion
Because we assume q(t) < 0 for x < xt , this introduces complex numbers on yL :
In conclusion, we have (
yL (x, xt ) if x < xt ,
y(x) =
yR (x, xt ) if x > xt ,
where
1 bR −iθ(x)/ε −iπ/4 ibR iθ(x)/ε −iπ/4
yL (x, xt ) = iaR + e e + aR + e e
|q(x)|1/4 2 2
1 −iθ(x)/ε iπ/4 iθ(x)/ε −iπ/4
bR −iθ(x)/ε −iπ/4 iθ(x)/ε iπ/4
= aR e e +e e + e e +e e
|q(x)|1/4 2
1 1 π 1 π
= 2aR cos θ(x) − + bR cos θ(x) +
|q(x)|1/4 ε 4 ε 4
1
aR e−κ(x)/ε + bR eκ(x)/ε
yR (x, xt ) = 1/4
q(x)
Z xt p
θ(x) = |q(s)| ds
Zx x p
κ(x) = |q(s)| ds.
xt
Example 4.2.1. Consider q(x) = x(2 − x), where −1 < x < 1. The simple turning point is at
xt = 0, with q 0 (0) = 2 > 0. One can compute and show that
1 p 1 h p i
θ(x) = (1 − x) x(x − 2) − ln 1 − x + x(x − 2) , x < 0
2 2
1 p 1 π
κ(x) = (x − 1) x(2 − x) − arcsin(x − 1) + , x > 0.
2 2 4
Balancing the first terms on each side of this equation gives γ = 1. The O(ω 2 ) = O(1/ε2 )
equation is the eikonal equation:
θx2 = µ2 (x), (4.3.3)
and its solutions are Z x
θ(x) = ± µ(s) ds. (4.3.4)
0
We choose the positive solution as we are considering the right-moving waves. The O(ω) =
O(1/ε) equation is the transport equation:
− θx2 u1 + iθx ∂x u0 + i (θx ∂x u0 + θxx u0 ) = −µ2 u1 − iαu0 . (4.3.5)
The u1 terms cancel out due to the eikonal equation (4.3.3), so (4.3.5) reduces to
θxx u0 + 2θx ∂x u0 = −αu0 , . (4.3.6)
With θx = µ(x), we can rearrange (4.3.6) and obtain a first order ODE in u0 :
µx + α
∂x u0 + u0 = 0, (4.3.7)
2µ
104 4.3. Wave Propagation and Energy Methods
which can be solved using the method of integrating factor. The integrating factor is given by
Z x Z x
µs (s) + α(s) p 1 α(s)
I(x) = exp ds = µ(x) exp ds ,
0 2µ(s) 2 0 µ(s)
and so (4.3.7) can be written as
1 x α(s)
Z
d a0 a0
(I(x)u0 ) = 0, u0 = =p exp − ds . (4.3.8)
dx I(x) µ(x) 2 0 µ(s)
Finally, imposing the boundary condition at x = 0 we obtain a first-term asymptotic expansion
of the travelling-wave solution of (4.3.1)
s
1 x α(s)
Z Z x
µ(0)
u(x, t) ∼ exp − ds cos ωt − ω µ(s) ds . (4.3.9)
µ(x) 2 0 µ(s) 0
Observe that in (4.3.9) the amplitude and phase of the travelling wave depend on the spatial
position x. Interestingly, (4.3.9) is independent of β(x).
It follows that
1
E(x, t) ∼ A2 µ2 ω 2 + ϕ2x sin2 [ωt − ϕ(x)]
(4.3.12a)
2
S(x, t) ∼ ωϕx A2 sin2 [ωt − ϕ(x)] (4.3.12b)
Φ(x, t) ∼ αω 2 A2 sin2 [ωt − ϕ(x)] . (4.3.12c)
Note that we neglect A0 since A is slowly changing. Suppose we choose xi (t) satisfying
ω
ẋi = = phase velocity.
ϕx (xi )
1 ωA2 2 2
µ ω + ϕ2x sin2 [ωt − ϕ(x)] − ωϕx A2 sin2 [ωt − ϕ(x)]
E ẋ − S ∼
2 ϕx
1 ωA2 2 2
µ ω − ϕ2x sin2 [ωt − ϕ(x)] = 0,
=
2 ϕx
since θ(x) = ϕ(x)/ω satisfies the eikonal equation (4.3.3). Hence, if x2 − x1 = O(1/ω) then it
dE
follows from (4.3.10) that ≈ 0, i.e. the total energy remains constant (to the first term)
dt
between any two phase lines x1 (t), x2 (t) that are O(1/ω) apart.
Recall the energy equation that
∂t E + ∂x S = −Φ.
where the average of ∂t E over one period vanishes using (4.3.12) for E. Substituting (4.3.12)
for S and Φ, we obtain
∂x ϕx A2 = −αωA2
∂x θx A2 = −αA2
which implies that A = u0 since the last equation is precisely the transport equation (4.3.6).
Physically, this means that the transport equation corresponds to the balance of energy over
one period in time.
106 4.4. Higher-Dimensional Waves - Ray Methods
Figure 4.3: Instructive case of the multi-dimensional wave equation. In R2 , the wave propagates
from the circle with radius a.
Thus we have a WKB-like solution for constant µ. Radial lines in this example correspond
to rays and from (4.4.4) we see that along a ray (i.e. ϕ is fixed), the solution has
p a highly
oscillatory component that is multiplied by a slowly varying amplitude V0 = f (ϕ) a/ρ that
decays as ρ increases.
Then
The O(1) equation is the eikonal equation which is now nontrivial to solve:
∇θ · ∇θ = µ2 . (4.4.7)
After cancelling the V1 term using the eikonal equation (4.4.7), the O(1/ω) equation is the
transport equation
2∇θ · ∇V0 + ∇2 θ V0 = 0.
(4.4.8)
Both ±θ are solutions to the eikonal equation and we choose the positive solution +θ since
this corresponds to the outward propagating waves.
108 4.4. Higher-Dimensional Waves - Ray Methods
Figure 4.4: Schematic figure of wave fronts in R3 and the path followed by one of the points
in the wave front (Taken from [Hol12, page 267]).
Θ(x, 0) = ωc.
As t increases, the points where Θ = ωc change, and therefore points forming Sc move and
form a new surface Sc+t = {θ(x) = c + t}. We still have
Θ(x, t) = ωc.
The path each point takes to get from Sc to Sc+t is obtained from the solution of the eikonal
equation and in the WKB method these paths are called rays.
The evolution of the wave front generates a natural coordinate system (s, α, β) where α, β
comes from parameterising the wave front and s from parameterising the rays. Note that
these coordinates are not unique as there are no unique parameterisation for the surfaces and
rays. It turns out that determining these coordinates is crucial in the derivation of the WKB
approximation.
The Wentzel-Kramers-Brillouin (WKB) Method 109
Example 4.4.1. Suppose we know a-priori that θ(x) = x · x. In this case, the surface Sc+t is
described by the equation |x|2 = c + t, which is just the sphere with radius c + t. The rays are
now radial lines and so the points forming Sc move along radial lines to form the surface Sc+t .
To this end, we use a modified version of spherical coordinates:
with
0 ≤ α < π, 0 ≤ β ≤ 2π, 0 ≤ s.
The function ρ(s) is required to be smooth and strictly increasing. Examples are ρ = s,
ρ = es − 1 or ρ = ln(1 + s).
∂s θ = λµ2 (4.4.10)
assuming we can find such a coordinate system (s, α, β). This amounts to solving (4.4.9) which
is generally nonlinear and requires the assistance of numerical method. Nonetheless, we still
have the freedom of choosing the function λ.
2∂s V0 + λ ∇2 θ V0 = 0.
(4.4.12)
This is true provided θ(0, α, β) = 0 in (4.4.11) since otherwise we will get an additional expo-
nential term from the WKB ansatz (4.4.5)
eiωθ(x0 ) = eiωθ(0,α,β) .
We now prove the identity (4.4.13) in R2 but this easily extends to R3 . The transformation
in R2 is x = X(s, α) and its Jacobian is
∂(x, y)
J = = ∂s x∂α y − ∂α x∂s y.
∂(s, α)
X|s=0 = x0 . (4.4.17)
We can also rewrite the ray equation (4.4.9) by taking the dot product of (4.4.9) against
∂s X:
∂X ∂X
· = λ2 ∇θ · ∇θ = λ2 µ2 .
∂s ∂s
If ` be the arc length along a ray, then
Z s Z s
`= k∂s Xk ds = λµ ds.
0 0
Hence, s equals the arc length along a ray if we choose λµ = 1. Another common choice is
λ = 1.
where s is the value for which the solution of (4.4.19) satisfies X(s) = x.
4.5 Problems
1. Use the WKB method to find an approximation of the following problem on x ∈ [0, 1]:
εy 00 + 2y 0 + 2y = 0, y(0) = 0, y(1) = 1.
Balancing leading order terms of the first two terms we obtain α = 1 and the O(1/ε)
equation is the eikonal equation
where c1 , c2 are arbitrary constants. The O(1) equation, after simplifying using the
eikonal equation, is the following:
Suppose θx = 0, then (4.5.2) reduces to 2y00 + 2y0 = 0 and its general solution is
y0 (x) = a0 e−x .
114 4.5. Problems
Suppose θx = −2, then (4.5.2) reduces to −2y00 + 2y0 = 0 and its general solution is
y0 (x) = b0 ex .
y ∼ b0 −e−x + ex−2x/ε
2. Consider seismic waves propagating through the upper mantle of the Earth from a source
on the Earth’s surface. We want to use a WKB approximation in R3 to solve the equation
∇2 v + ω 2 µ2 (r)v = 0,
Figure 4.5: Rays representing waves propagating inside the earth from a source on the surface
of the earth.
The Wentzel-Kramers-Brillouin (WKB) Method 115
(a) Use the ray equation to show that the vector p = r × (µ∂s r) is independent of s.
Hence, show that krkµ sin(χ) is constant along a ray, where χ is the angle between
r and ∂s r.
Solution: With the choice λ = 1/µ, the ray equation (4.4.16) reduces to
∂ ∂
µ X = ∇µ(X), with x = r = X(s, α, β).
∂s ∂s
The first term vanishes because the cross product of any vector with itself is
zero and the second term vanishes since r and ∇µ(r) are parallel. Therefore
∂s p = 0 and so p is independent of s.
where χ is the angle between the vectors r and ∂s r. Multiplying each side by
the positive scalar function µ we obtain
(b) Part (a) implies that each ray lies in a plane containing the origin of the sphere. Let
(ρ, ϕ) be polar coordinates of this plane. It follows that for a polar curve ρ = ρ(ϕ),
the angle χ satisfies
ρ
sin(χ) = q . (4.5.4)
ρ2 + (∂ϕ ρ)2
116 4.5. Problems
Solution: Given a ray, let (ρ, ϕ) be polar coodinates of the plane containing
such ray. Since this plane contains the origin of the sphere, we can identify ρ as
the magnitude of the radial (position) vector r and from (4.5.3) we know that
κ κ
sin(χ) = = . (4.5.5)
krkµ µρ
(c) Use the definition of arc length, show that for a polar curve
q
µds = ρ2 + (∂ϕ ρ)2 dϕ. (4.5.6)
Combining this result with part (b), show that the solution of the eikonal equation
is given by
1 ϕ 2 2
Z
θ= µ ρ dϕ.
κ ϕ0
The Wentzel-Kramers-Brillouin (WKB) Method 117
Solution: First of all, we must distinguish the ray parameter s with the ar-
clength parameter ` of a ray. For a polar curve (ϕ, ρ(ϕ)), we have x = ρ(ϕ) cos ϕ
and y = ρ(ϕ) sin ϕ and so
s
2 2
dx dy
d` = + dϕ
dϕ dϕ
q
= (−ρ sin ϕ + ∂ϕ ρ cos ϕ)2 + (ρ cos ϕ + ∂ϕ ρ sin ϕ)2 dϕ
q
= ρ2 sin2 ϕ + cos2 ϕ + (∂ϕ ρ)2 cos2 ϕ + sin2 ϕ dϕ
q
= ρ2 + (∂ϕ ρ)2 dϕ.
Method of Homogenization
with u(0) = a and u(1) = b. In many physical problems, D is known as the conductivity
tensor and we are interested in D = D(x, x/ε), where it includes a slow variation in x as well
as a fast variation over a length scale that is O(ε). A physical realisation of this is a material
having micro and macrostructures with spatial variation. For example, we might have
1
D(x, y) = , (5.1.2)
1 + αx + βg(x) cos y
with
α = 0.1, β = 0.1, ε = 0.01, g(x) = e4x(x−1) .
Our main goal is to try to replace, if possible, D(x, x/ε) = D(x, y) with some effective
(averaged) D that is independent of ε. A naive guess would be to simply average over the fast
variation, i.e.
1 y
Z
hDi∞ = lim D(x, r) dr. (5.1.3)
y→∞ y 0
It turns out that this is not a good approximation because the solution of (5.1.1) with hDi∞
might be a bad approximation of the solution of (5.1.1).
Because of the two different length scales in (5.1.1), it is natural to invoke the method of
multiple scales, but with an important distinction. Here, we want to eliminate the fast length
scale y = x/ε, as opposed to the standard multiple scales where we keep both the slow and
normal scales. For the existence of solution of (5.1.1), we assume D(x, y) is smooth and satisfies
119
120 5.1. Introductory Example
Figure 5.1: Rapidly varying coefficient D and its average. The red line depicts rapidly varying
D(x, x/) in (5.1.2) and the blue dotted line shows its effective mean D(x) = 1/(1 + αx).
for all x ∈ [0, 1] and y > 0, where Dm (x) and DM (x) are both continuous. With the fast scale
y = x/ε and the slow scale x, the derivative becomes
d 1
−→ ∂y + ∂x
dx ε
and (5.1.1) becomes
(∂y + ε∂x )[D(x, y)(∂y + ε∂x )u] = ε2 f (x). (5.1.6)
We assume a regular perturbation expansion of the form
∂y [D(x, y)∂y u0 ] = 0,
i.e. the integral is unbounded but its growth is confined by linear functions in y as y −→ ∞.
The O(ε) equation is
∂y [D(x, y)∂y u1 ] = −∂x u0 · ∂y D. (5.1.9)
Integrating this with respect to y twice and using the fact that u0 = u0 (x) yields
Observe that 2 and 3 increases linearly with y for large y, and analogous to removing
secular terms in multiple scales, we require that these two terms cancel each other so that u1
is bounded. This means that we must impose
Z y
1 ds
lim b0 (x) − y∂x u0 = 0.
y→∞ y y0 D(x, s)
In general multiple scales problem, it is enough to get information from O(ε) terms to
obtain a first-term approximation. However, for homogenization problems, we need to proceed
to O(ε2 ) equation to determine u0 (x). The O(ε2 ) equation is
The last integral is O(y 2 ) for large y and cannot be cancelled by other terms in (5.1.12).
Therefore, we require b00 (x) = f (x). Finally, rearranging (5.1.11) and differentiating with
respect to x we obtain
∂x [D(x)∂x u0 (x)] = b00 (x) = f (x), (5.1.13)
122 5.2. Multi-dimensional Problem: Periodic Substructure
We called (5.1.13) the homogenized differential equation with the homogenized, or ef-
fective, coefficient D.
Figure 5.2: Exact and averaged solution. The red line depicts the exact solution and the blue
line shows its homogenized solution.
Figure 5.3: Fundamental domain with periodic substructure. On the fundamental domain,
function has a same set of values.
and we assume that u0 , u1 , u2 , . . . are periodic in y with period y p due to the periodicity
assumption on D.
The O(1) equation is
∇y (D∇y u0 ) = 0,
and the general solution of this, which is bounded, is u0 = u0 (x). If D were constant, then
it follows from Liouville’s theorem that bounded solutions of Laplace’s equation over R2 are
constants. One can argue similarly in the case where D is not constant. The O(ε) equation is
Because u1 is periodic in y, it suffices to solve (5.2.5) in a cell Ωp and then simply extend the
solution using periodicity. Observe that (5.2.5) is linear with respect to y and u0 does not
depend on y. Thus the general solution of (5.2.5) follows from superposition principle
To derive the homogenized equation for u0 , we introduce the cell average of a function v(x, y)
over Ωp : Z
1
hvip (x) = v(x, y) dVy .
|Ωp | Ωp
Averaging the first term of (5.2.8) and applying the divergence theorem gives
Z
D E 1
∇y · [D(∇y u2 + ∇x u1 )] = ∇y · [D(∇y u2 + ∇x u1 )] dVy
p |Ωp | Ωp
Z
1
= Dn · (∇y u2 + ∇x u1 ) dSy
|Ωp | ∂Ωp
=0
Method of Homogenization 125
since u1 , u2 are periodic over the cell Ωp . Next, using (5.2.6) we have
Similarly,
D E
hD∂xi u0 ip = hDip ∂xi u0 =⇒ ∇x · (D∇x u0 ) = ∇x · hDip ∇x u0
p
and the functions ai are smooth periodic solutions of the cell problem
Example 5.2.1. Consider the cell Ωp = [0, a] × [0, b] in R2 . To determined the homogenized
coefficients in (5.2.10), it is necessary to solve the cell problem (5.2.11). Consider a “separable”
coefficient function D:
D(x, y) = D0 (x1 , x2 )eα(y1 ) eβ(y2 ) ,
where α(y1 ) and β(y2 ) are periodic with period a and b respectively. The cell equations for
a1 , a2 are
and
Z y1
a1 (y1 ) = −y1 + κ1 e−α(s) ds
0
126 5.3. Problem
Z y2
a2 (y2 ) = −y2 + κ2 e−β(s) ds.
0
Now, since ∂y2 a1 = ∂y1 a2 = 0, it follows from (5.2.10) that D12 = D21 = 0. Moreover,
1 a b
Z Z
α(y1 )+β(y2 ) −α(y1 )
hD∂y1 a1 ip = D0 (x)e − 1 + κ1 e dy1 dy2
ab 0 0
1 a b 1 a b
Z Z Z Z
α(y1 )+β(y2 )
=− D0 (x)e dy1 dy2 + D0 (x)κ1 eβ(y2 ) dy1 dy2
ab 0 0 ab 0 0
Z b
1 β(s)
= −hDip + D0 (x)κ1 e ds
b 0
κ1
= −hDip + D0 (x) ,
κ2
and similarly
Z a Z b
1
hD∂y2 a2 ip = D0 (x)eα(y1 )+β(y2 ) − 1 + κ2 e−β(y2 ) dy1 dy2
ab0 0
1 a b 1 a b
Z Z Z Z
α(y1 )+β(y2 )
=− D0 (x)e dy1 dy2 + D0 (x)κ2 eα(y1 ) dy1 dy2
ab 0 0 ab 0 0
Z a
1
= −hDip + D0 (x)κ2 eα(s) ds
a
0
κ2
= −hDip + D0 (x) .
κ1
Consequently, the homogenized differential equation (5.2.9) for u0 is
5.3 Problem
1. Consider the equation
for some continuous functions Dm , DM in [0, 1]. We introduce y = x/ε and designate
the slow scale simply as x. The derivative transforms into
d ∂ 1 ∂ 1
−→ + = ∂x + ∂y ,
dx ∂x ε ∂y ε
with y0 fixed. We deduce from the lecture that c0 (x) must be zero and consequently
u0 is a function of x only, i.e. u0 (x, y) = u0 (x).
We deduce from the lecture that the following equation must be true to prevent u1
from blowing up:
∂x u0 = hD−1 i∞ b0 (x), (5.3.4)
where hD−1 i∞ = (D)−1 .
Since the last integral is O(y 2 ) for large y and there are no other terms in the
expression of u2 (x, y) that can cancel this growth, it is necessary to impose
1 y
Z h Z τ
1 i
lim − ∂x b0 + g(u0 ) τ + f (x, s) ds dτ = 0,
y→∞ y 2 y D(x, τ ) y0
0
or equivalently Z τ
1
∂x b0 + g(u0 ) = lim f (x, s) ds = hf i∞ . (5.3.5)
τ →∞ τ y0
[Bre14] P. C. Bressloff. “Waves in Neural Media”. In: Lecture Notes on Mathematical Mod-
elling in the Life Sciences (2014). doi: 10.1007/978-1-4614-8866-8.
[BEW08] P. C. Bressloff, B. A. Earnshaw, and M. J. Ward. “Diffusion of Protein Receptors on
a Cylindrical Dendritic Membrane with Partially Absorbing Traps”. In: SIAM Jour-
nal on Applied Mathematics 68.5 (2008), pp. 1223–1246. doi: 10.1137/070698373.
[CSW09] D. Coombs, R. Straube, and M. Ward. “Diffusion on a Sphere with Localized
Traps: Mean First Passage Time, Eigenvalue Asymptotics, and Fekete Points”. In:
SIAM Journal on Applied Mathematics 70.1 (2009), pp. 302–332. doi: 10.1137/
080733280.
[Hol12] M. H. Holmes. Introduction to Perturbation Methods. 2nd. Vol. 20. Texts in Applied
Mathematics. Springer Science & Business Media, 2012. doi: 10.1007/978- 1-
4614-5477-9.
[Kur+15] V. Kurella et al. “Asymptotic Analysis of First Passage Time Problems Inspired
by Ecology”. In: Bulletin of mathematical biology 77.1 (2015), pp. 83–125. doi:
10.1007/s11538-014-0053-5.
[PRK03] A. Pikovsky, M. Rosenblum, and J. Kurths. Synchronization: A Universal Concept
in Nonlinear Sciences. Vol. 12. Cambridge Nonlinear Science. Cambridge University
Press, 2003.
[Pil+10] S Pillay et al. “An asymptotic Analysis of the Mean First Passage Time for Nar-
row Escape Problems: Part I: Two-dimensional Domains”. In: Multiscale Modeling
& Simulation 8.3 (2010), pp. 803–835. doi: 10.1137/090752511.
[SWF07] R. Straube, M. J. Ward, and M. Falcke. “Reaction Rate of Small Diffusing Molecules
on a Cylindrical Membrane”. In: Journal of Statistical Physics 129.2 (2007), pp. 377–
405. doi: 10.1007/s10955-007-9371-4.
129