Dispersion3 Merged
Dispersion3 Merged
Bernard Deconinck
Department of Applied Mathematics
University of Washington
Campus Box 352420
Seattle, WA, 98195, USA
Prolegomenon
These are the lecture notes for Amath 353: Partial Differential Equations and Waves. This
is the first year these notes are typed up, thus it is guaranteed that these notes are full of
mistakes of all kinds, both innocent and unforgivable. Please point out these mistakes to
me so they may be corrected for the benefit of your successors. If you think that a different
phrasing of something would result in better understanding, please let me know.
These lecture notes are not meant to supplant the textbook used with this course. The
main textbook is Roger Knobel’s “An introduction to the mathematical theory of waves”,
American Mathematical Society 1999, Student Mathematical Library Vol 3.
These notes are not copywrited by the author and any distribution of them is highly
encouraged, especially without express written consent of the author.
ii
Contents
1 Introduction 1
1.1 An introduction to waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 A mathematical representation of waves . . . . . . . . . . . . . . . . . . . . 3
1.3 Partial differential equations . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3 Fourier series 49
3.1 Superposition of standing waves . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.2 Fourier series . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.3 Fourier series solutions of the wave equation . . . . . . . . . . . . . . . . . . 56
3.4 The heat equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
3.5 Laplace’s equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
3.6 Laplace’s equation on the disc . . . . . . . . . . . . . . . . . . . . . . . . . . 64
iii
iv CONTENTS
Introduction
ACMS and Mathematics: The study of wave phenomena has produced some of
the biggest mathematical breakthroughs of the last decades. Waves have helped our
understanding of geometry, algebraic geometry, analysis, and just about any other area
of mathematics. And, obviously, as an ACMS major, all applications listed here are
relevant to you.
Biology: Dispersal of seeds by wind waves is one of the most important means of
plant regeneration.
1
2 CHAPTER 1. INTRODUCTION
end, they are transmitted, and finally they are received on the other end. If we’re
dealing with digital communication, the waves we’re talking about are sequences of
zeros and ones.
Mechanical Engineering: The understanding of tidal and other water waves is im-
portant for harbor and ship design, see https://www.dropbox.com/s/mlittwqaxk6j3we/
hexagons-bw.jpg?dl=0.
Psychology: A delta wave is a brain wave that occurs during what is known as
deep sleep. It is known that delta wave activity is vastly reduced in people with
schizophrenia.
What do all the phenomena mentioned above have in common? It turns out that it is
not so easy to give a precise definition of a wave that captures everything we want. We can
agree on the following.
1. A wave is the result of a disturbance propagating through a medium, with finite ve-
locity, and
2. Associated with waves are signals: the result of any kind of measurement of a wave.
The outcome could be called Amplitude, Frequency, etc.
1.2. A MATHEMATICAL REPRESENTATION OF WAVES 3
u(x, tn )
u(x, 0) u(x, 3)
u(x, 2)
u(x, 1)
Figure 1.1: A wave of decreasing height (i.e., amplitude) traveling to the right.
Example. Ripples in a pond. Waves travel horizontally across the surface of the
pond. A signal might be the vertical displacement of the crests.
Example. Waves in traffic disturbances can be caused by an accident, a police car,
a traffic light, a car merging, etc. We will discuss this example near the end of the course.
And I apologize: this knowledge will not let you race through Seattle traffic. But it will tell
you why you are stopped. That should make you happier, in a zen-like way, no?
Example. The wave in a sports stadium.
u(x, tn )
u(x, 0)
u(x, 2.6)
gives a profile f (x) that is moving to the right (because v is positive) with speed v. If v < 0,
the profile would be moving to the left, with speed |v|. Indeed, at t = 0, we have
u(x, 0) = f (x),
so this is our initial profile. At a later time t > 0, we get the same profile, but it has moved.
Suppose we wanted to track the value f (0). At any time t, this value can be found at the
position given by
x − vt = 0 ⇒ x = vt,
which is positive, and increasing linearly with time t. Specifically, the speed at which the
value f (0) moves to the right is given by
dx
= v,
dt
which verifies our claim. The same reasoning works for tracking any value of f (x)1 .
Example. As an example of our example (Really? Yup, really.), we consider
u(x, t) = sech2 (x − 5t).
I know you all love hyperbolic functions, so I won’t waste much time recalling what they
are. We have that
1
sech(x) = ,
cosh(x)
and
ex + e−x
cosh(x) = .
2
1
When I make such gratuitous statements, you should check them. I do not make them to sound smart.
If I did, I’d make them in Latin.
1.2. A MATHEMATICAL REPRESENTATION OF WAVES 5
Figure 1.3: The graphs for cosh(x) (left), sech(x) (middle), and sech2 (x) (right).
Figure 1.3 shows what the plot of this function looks like at time t = 0. Once we know that,
we know that u(x, t) has the same plot at all time, but translated to the right by an amound
5t.
Example. The wave u(x, t) = sin(x + t) represents a sine function at t = 0, which moves
to the left with speed 1.
Example. Consider
u(x, t) = H(x − 7t),
where H(x) is the Heaviside step function:
1 if x ≥ 0,
H(x) =
0 if x < 0.
Its graph is drawn in Fig. 1.4. Thus u(x, t) is a step profile, moving to the right with
velocity 7.
Slice plots. We create a bunch of slices and we plot them together, either in 2-D or
in 3-D. Great for paper communication.
6 CHAPTER 1. INTRODUCTION
H(x)
0
x
Figure 1.4: The graphs for cosh(x) (left), sech(x) (middle), and sech2 (x) (right).
Surface plots. We plot u(x, t) as a surface depending on the two variables x and
t. Also good for paper communication, provided that the three dimensions come out
well. This depends heavily on u(x, t).
(x, t)-plots, contour plots. We plot u(x, t) as a function in the (x, t) plane, but
looked at from above. We use different colors or grey scale to indicate the height
u(x, t). Great on paper.
ut + cux = 0,
where c is a parameter. This equation is linear, since it contains no products of u with itself
or any of its derivatives, and first order in both x and t, since no derivatives of higher than
1.3. PARTIAL DIFFERENTIAL EQUATIONS 7
first order appear. The equation is also homogeneous, because u = 0 is a solution, albeit2
a not very interesting one.
Example. The diffusion equation is given by
ut = σuxx ,
where σ > 0 is a parameter. This equation is also linear, but is is of second order in x, first
order in t. The equation is homogeneous. Since it arises in the study of heat transport, it is
also known as the heat equation.
Example. The three-dimensional diffusion or heat equation is
As before, the equation is linear and homogeneous. It is first order in t, and second order in
the spatial variables x, y and z.
Example. The Burgers equation is given by
ut + uux = 0.
It is first order in x and t. It is our first example of a nonlinear PDE, because of the second
term. It is still homogeneous, but for nonlinear equations that won’t buy us much.
Example. The equation
ut + uux + uxxx = g(t)
is nonlinear (evil second term), first order in t, third order in x, and nonhomogeneous if
g(t) 6≡ 0.
Example. The sine-Gordon equation3 is given by
uxt = sin u.
uxt = cos u,
uxx > 0
uxx > 0
uxx < 0
Figure 1.5: An initial profile for the heat equation and how it will evolve.
is the rate of change of u as x changes, at a fixed time. This corresponds to taking a snapshot
of u, and looking at the slope of the curve in the picture. It follows that both ux and ut are
interpreted as velocities. Similarly, both uxx and utt can be seen as accelerations.
Example. Let u represent the temperature in a metal rod at position x, at time t. Then
u satisfies the so-called heat equation
ut = Duxx ,
where D > 0 is the heat conductivity of the rod, assumed to be constant here. Suppose we
start with an initial temperature profile, as in Fig. 1.5. In those regions where uxx > 0 (in
other words, the function is concave up), the heat equation states that ut will be positive,
so u will increase in time. For those regions where the profile is concave down, the heat
equation gives that ut will be negative, i.e., u will decrease in time, as indicated in the
figure. It follows that the heat equation likes to smear out profiles to a constant value.
Example. Next, we consider u to be a solution of the transport equation
ut = ux .
We choose an initial profile u(x, 0) as in Fig. 1.6. Where u(x, 0) is increasing as a function
of x, ux > 0 and therefore ut = ux > 0, thus u is increasing in time. Similarly, if u(x, 0)
decreases as a function of x, ux < 0 and it follows that u decreases in time, as indicated by
the arrows in Fig. 1.6. It follows that the overall shape of the profile will move to the left.
1.3. PARTIAL DIFFERENTIAL EQUATIONS 9
ux < 0
ux > 0
Figure 1.6: An initial profile for the transport equation and how it will evolve.
10 CHAPTER 1. INTRODUCTION
Chapter 2
u(x, t) = f (x − vt)
represent profiles f (x) that move to the right with velocity v is v > 0, and to the left if
v < 0.
In general, solutions of PDEs are functions of both x and t independently. Sometimes
they have solutions where x and t always show up in the special combination x − vt. We call
such solutions traveling waves. We are especially interested in the case when f (x) is not
constant, or in the case where f (x) is bounded for all values of x. Unbounded signals (i.e.,
u → ±∞) are usually1 not relevant for applications. Then f (x − vt) represents a disturbance
moving through a medium with velocity v.
Example. Consider
u(x, t) = sin(3x − t).
This is a traveling wave moving to the right with velocity v = 1/3. Indeed,
u(x, t) = sin(3x − t)
= sin 3(x − t/3)
= f (x − vt),
utt = a2 uxx ,
1
In the “always” sense.
11
12 CHAPTER 2. TRAVELING AND STANDING WAVES
u = f (x − vt).
Our task is to find the function f and the constant v. Let z = x − vt. We get
u(x, t) = f (z)
⇒ ux = f 0 ,
⇒ uxx = f 00 ,
⇒ ut = −vf 0 ,
⇒ utt = (−v)2 f 00 = v 2 f 00 ,
where we have used the chain rule and the fact that
∂z ∂z
= 1, = −v.
∂x ∂t
Thus, traveling wave solutions of the PDE
utt = a2 uxx
v 2 f 00 = a2 f 00
⇒ (v 2 − a2 )f 00 = 0.
f 00 = 0 ⇒ f (z) = Az + B,
independent of what v is. Here A and B are constants. The second possibility results in
v 2 = a2 ⇒ v = ±a,
independent of what f (z) is. In summary, we have obtained the following solutions:
A(x − vt) + B, for any value of A, B, v. These solutions are not very interesting:
in order for them to be bounded, we need A to be zero. But that leaves us with a
constant solution, which is unexciting.
Since the equation is linear, we can superimpose the solutions to get a more general
solution:
u(x, t) = A(x − vt) + B + f1 (x − at) + f2 (x + at).
Some remarks are in order.
Note that the superposition of different traveling waves is not necessarily a traveling
wave! Our superposition consists of three different parts. Two of these parts (f1 (x−at)
and f2 (x + at)) even move in opposite directions.
First we substitute u = f (z), z = x − vt into the equation. We want to find an ODE for
f (z), which we want to use to determine f (z) and perhaps also v. In this case, we get
v 2 f 00 = f 00 − sin f
⇒ (1 − c2 )f 00 = sin f.
This is a messy second-order ODE. We can reduce it to a first order ODE by multiplying by
f 0 , which results in an equation we may integrate once:
(1 − c2 )f 0 f 00 = f 0 sin f
02
d 2 f d
⇒ (1 − c ) = [− cos f ]
dz 2 dz
⇒ (1 − c2 )f 02 = A − 2 cos f,
where we have used that f 0 f 00 is the derivative of f 02 /2 and that f 0 sin f is the derivative of
− cos f . Here A is an arbitrary constant. Our new equation is a first-order ODE for f (z).
This is progress! It’s still a messy ODE, but we can actually solve this ODE using separation
of variables. Here I just give one class of solutions. You should check that these are, in fact,
solutions2 . √
z/ 1−c2
f (z) = 4 arctan e ,
2
You should never, ever, trust me.
14 CHAPTER 2. TRAVELING AND STANDING WAVES
u(x, t)
Figure 2.1: The profile of a traveling wave solution of the sine-Gordon equation. For this
specific profile, c = 1/2 and the whole graph moves to the right with velocity 1/2.
“I was observing the motion of a boat which was rapidly drawn along a narrow channel
by a pair of horses, when the boat suddenly stopped - not so the mass of water in the
channel which it had put in motion; it accumulated round the prow of the vessel in a
state of violent agitation, then suddenly leaving it behind, rolled forward with great
velocity, assuming the form of a large solitary elevation, a rounded, smooth and well-
defined heap of water, which continued its course along the channel apparently without
change of form or diminution of speed. I followed it on horseback, and overtook it still
rolling on at a rate of some eight or nine miles an hour, preserving its original figure
some thirty feet long and a foot to a foot and a half in height. Its height gradually
3
I am not making this up!
2.3. THE KORTEWEG-DE VRIES EQUATION 15
diminished, and after a chase of one or two miles I lost it in the windings of the channel.
Such, in the month of August 1834, was my first chance interview with that singular
and beautiful phenomenon which I have called the Wave of Translation”
It took 10 years for J. S. Russell to publish his results. It took another 50 for them
to become appreciated (no Twitter yet). Two Dutchmen Korteweg (PhD advisor) and de
Vries (his student) derived the equation that now bears their name, although we usually
abbreviate it the KdV equation:
ut + uux + uxxx = 0.
They derived the equation to describe long waves in shallow water, like tsunamis. But it
describes much more than this: in general it describes the propagation of long waves in a
dispersive medium4 . This covers water waves, but also waves in plasmas, like the northern
lights, or like the light flickering we sometimes see in (large) LED lights.
Let’s look for traveling wave solutions of the KdV equation. Thus we let
u(x, t) = f (z), z = x − vt,
where v is the velocity of the wave. For simplicity we will look for waves traveling to the
right, thus v > 0. Further, we will limit our investigations to waves like the ones that Russell
saw, namely waves that decay to zero as x → ∞ or x → −∞. As before, we have
where B is a second constant of integration. evaluating the left-hand side once again at ±∞,
we find that B = 0. Thus, we find the first-order ordinary differential equation
v 1 1
− f 2 + f 3 + f 02 = 0.
2 6 2
It follows that
3f 02 = (3v − f )f 2 .
Since v > 0 (by choice), we find that f ≤ 3v. That’s good: we are looking for a bounded
solution, and it looks like we’ll find one. It remains to solve the first-order equation using
separation of variables. Ready to have a good time with integration? We have
√ 0
3f
√ =1
f 3v − f
√ Z df
⇒ 3 √ = z + α,
f 3v − f
where α is a constant of integration. We use substitution to simplify the integral. The
hardest part is the square root in the denominator. Let
p
g = 3v − f ⇒ f = 3v − g 2 .
v
3v
Figure 2.2: The profile of a traveling wave solution of the KdV equation. The profile moves
to the right with velocity v and has height 3v.
is a traveling wave solution of the KdV equation, for any v > 0. It is illustrated in Fig. 2.2. Its
amplitude (i.e., height) is 3v, while its velocity is v. Thus higher waves of the KdV equation
travel
√ faster, as we observe at the beach. Further, the width of the profile is proportional to
1/ v, thus taller, faster waves are more narrow. Lastly, the maximum of the wave profile
occurs when x − vt + α = 0, or at x = vt − α.
lim u = k1 , lim u = k2 ,
x→−∞ x→∞
and
k1 6= k2 .
In other words, our profile approaches different limit values at −∞ and +∞. The traveling
wave solution of the sine-Gordon equation is a good example. Note, however, that the
transition between the two limiting values does not have to be monotone.
If on the other hand,
k1 = k2 ,
then we call the solution a pulse. The profile drawn in Fig. 1.6 is an example. Thus, for
pulses, the beginning and end state are the same: the medium returns back to its original
state after a pulse passes through. On the other hand, the passing of a front forever alters
the state of the medium.
Examples of fronts.
18 CHAPTER 2. TRAVELING AND STANDING WAVES
Weather fronts: different meteorological signals are altered by the passing of a weather
front, such as the air pressure.
Sonic booms: the density in the air is altered by the passing of the sonic boom. Of
course, dissipation effects eventually return it to its original value, but for quite a while,
the air density is changed. That is different from a subsonic plane passing through,
which creates a localized disturbance in the density.
Bob Dylan, Beethoven: music was forever altered by both of them.
Examples of pulses.
Optical flashes
Bits
A is the amplitude. This is the largest value the wave train attains. Similarly, −A is
its smallest value.
The wave number k denotes how many oscillations occur during an interval of length
2π. Indeed, the period of the wave train is 2π/k. Since one oscillation occurs per
period, k oscillations occur over an interval of length 2π.
The frequency ω denotes how many oscillations happen during a time interval of
length 2π. The same argument as above shows this, but we consider the temporal
dependence instead of the spatial dependence. It follows that signals with high wave
number or frequency are very oscillatory. That means that if we want to plot them
accurately, we’ll need to use many points at which to sample the function.
2.5. WAVE TRAINS AND DISPERSION 19
The phase shift φ simply shift the origin of space or time. That sounds very philo-
sophical5 , but it really isn’t. We can rewrite the wave train signal as (using the cosine
form, for instance)
u(x, t) = A cos(k(x − x0 ) − ωt) or u(x, t) = A cos(kx − ω(t − t0 )),
where −kx0 = φ and ωt0 = φ. Thus φ simply changes when we start measuring where
or when we start.
The velocity of the traveling wave is v = ω/k, since we can rewrite the signal as
u(x, t) = A cos(k(x − vt) + φ),
where v = ω/k.
Often, the frequency ω and the wave number k are related. This relation is called the dis-
persion relation. Since sin and cos can be written as linear combinations of exponentials,
it suffices to consider expressions of the form
u = Aeikx−iωt .
The idea is to substitute this into the PDE and find the relationship between ω and k, if
there is one. This is easy: note that whenever we take an x derivative, we get
ux = ikAeikx−iωt = iku,
and so on for higher-order derivatives. Thus taking an x derivative simply multiplies our
solution by a factor ik. Similarly, taking a time derivative,
ut = −iωAeikx−iωt = −iωu,
and taking a time derivative results in multiplying u by −iω. Taking more time derivatives
results in more powers of −iω.
Example. Consider the wave equation
utt = a2 uxx , a > 0.
We already know that this equation has traveling wave solutions that travel to the left with
velocity −a and to the right with velocity a. Plugging in
u = Aeikx−iωt ,
we get
(−iω)2 u = a2 (ik)2 u
⇒ −ω 2 = −k 2
⇒ ω 2 = k 2 a2 ,
5
That’s OK: I have a doctorate in philosophy. Really.
20 CHAPTER 2. TRAVELING AND STANDING WAVES
Thus there are two possible branches of the dispersion relation. They give rise to the solutions
u1 = A1 eik(x−at) ,
and
u2 = A2 eik(x+at) .
(−iω)2 u = a(ik)2 u − bu
⇒ −ω 2 = −ak 2 − b
⇒ ω 2 = ak 2 + b
√
⇒ ω1,2 = ± ak 2 + b,
u1 = A1 eikx−iω1 t ,
and
u2 = A2 eikx−iω2 t .
iϕt = −ϕxx .
to get
i(−iω)ϕ = −(ik)2 ϕ
⇒ ω = k2.
6
because there is no potential
2.6. DISPERSION RELATIONS FOR SYSTEMS OF PDES 21
ut = αux + vxxx
vt = βvx − uxxx .
We wish to look for wave train solutions, whose x and t dependence is of the form
eikx−iωt .
For the scalar case, the amplitude A never came into play, since for linear equations, we
could simply divide it out. That is no longer true here, since there is no reason why u and
v should have the same amplitude.
This situation is similar to what we do with systems of ODEs: if we wish to solve
ay 00 + by 0 + cy = 0,
we guess
y = eλt .
On the other hand, if we want to solve the system
y 0 = Ay,
y = eλt v,
where v is a vector.
We do the same for our system of PDEs. We guess
u U
= eikx−iωt .
v V
we need that
−iω − αik ik 3
3
ik −iω − βik
is a singular matrix. Thus
−iω − αik ik 3
det 3 = 0.
ik −iω − βik
This the dispersion relation, determining the frequency ω as a function of the wave number
k. The rest is algebra. We get
u = eikx−iωt .
Since the equation is linear, we may superimpose such solutions. But what can we say about
these solutions? What are their properties?
u = eik(x−cp (k)t) ,
where
ω(k)
cp (k) =
k
is called the phase velocity. It is the velocity with which a single wave train travels.
dω
cg (k) = ,
dk
2.7. INFORMATION FROM THE DISPERSION RELATION 23
cg
cp
Figure 2.3: A wave packet consisting of two traveling wave trains. The phase speed is the
speed of waves inside the packets, whereas the envelope of the packets moves with the group
speed.
which we call the group velocity. The full importance of the group velocity is hard to
explain (see Amath569), but let’s give it a shot.
Suppose that in a given signal, the most important wave number is k0 , with corre-
sponding frequency ω0 = ω(k0 ). We rewrite exp(ikx − iω(k)t) as
eikx−ω(k)t = eik0 −iω0 t ei(k−k0 )x−i(ω(k)−ω0 )t
0 2 )−ω )t
= eik0 −iω0 t ei(k−k0 )x−i(ω(k0 )−(k−k0 )ω (k0 )+O((δk) 0
0 2 ))t
= eik0 −iω0 t ei(k−k0 )x−i((k−k0 )ω (k0 )+O((∆k)
0
= eik0 −iω0 t ei∆k(x−ω (k0 )t)+... .
The second exponential factor acts as a slowly varying amplitude to the first one.
Indeed, the wave number ∆k and frequency ω 0 (k0 )∆k are both small, assuming that
k is close to k0 . The second, slowly-varying factor moves with the group velocity! A
wave packet is used to illustrate this in Fig. 2.3.
Thus individual waves move with the phase velocity, while wave packets move with the
group velocity. Note that it is perfectly possible for the group and phase velocity to
have opposite signs.
Although we will not show this here, the group velocity is also the velocity with which
the energy associated with a wave moves. Often we care about this much more than
we care about individual waves.
Dispersive vs. non-dispersive waves. An equation or a system is called dispersive
if (a) the phase speed cp is real for real k, and (b) the phase speed is not constant. This
implies that wave trains with different wave numbers will move with different speeds.
24 CHAPTER 2. TRAVELING AND STANDING WAVES
Figure 2.4: A decaying wave train (ωI < 0, left) and a growing wave train (ωI > 0, right).
But there is more! In general, the dispersion relation may be complex, even for real k.
If that happens, then for those values of k the equation is not dispersive.
Example. Consider the PDE
ut = uxx + cux .
In general, we can split ω(k) into its real and imaginary parts: ω = ωR + iωI , where ωR
and ωI are the real and imaginary parts of ω, respectively. It follows that our wave train
solutions are of the form
Thus, if ωI > 0, the signal will grow in time. If ωI < 0, the signal will decay in time. Both
situations are illustrated in Fig. 2.4.
If ωI < 0 and amplitudes decay, we say the PDE is dissipative. if ωI > 0 and amplitudes
grow, we call the system unstable.
2.8. PATTERN FORMATION 25
Let’s investigate the dynamics of solutions close to this solution. Such solutions are
small, so that we may ignore the term uux . Indeed, if
u ∼ Aeikx−iωt ,
ut + uxx + auxxxx = 0.
k 2 − ak 4
so that
2 −ak 4 )t
eikx−iωt = eikx+(k .
It follows that the solution grows in t if k 2 − ak 4 > 0 and it decays if k 2 − ak 4 < 0. We
plot k 2 − ak 4 in Fig. 2.5.
√ √
If k ∈ (−1/ a, 1/ a), the growth rate is positive. Otherwise it is negative. Since the
equation is linear, the general solution is a superposition of a bunch of solutions of the
form
2 4
eikx+(k −ak )t .
We have √ just concluded that any part of this superposition that has k > 1/sqrta or
k < −1/ a will decay. Thus after some time, these parts will not come into play
anymore. Since for all of these parts, |k| is large, the period is small. In other words,
these are highly oscillatory signals. The equation appears to get rid of them quickly.
All the contributions from the other wave numbers k grow (except from k = 0). Which
one grows the most? Let
Thus the growth rate is maximal for k ∗ = ±1/sqrt2a. All other solutions grow slower
than this one. Thus, in comparison, they decay. Indeed, if we have
y = c1 eax + c2 ebx ,
2.9. A DERIVATION OF THE WAVE EQUATION 27
u(x, t)
Figure 2.6: The set-up for the derivation of the wave equation modeling a plucked string.
y = eax c1 + c2 e(b−a)x ,
and the exponential in the parentheses decays! Thus, so sufficiently large x, we see
y ∼ c1 eax .
The same conclusion holds for our case: for large enough t, we see only
∗ x+(k ∗2 )−ak ∗4 t ∗ x+(k ∗2 )−ak ∗4 t
eik and eik .
√
Thus, the equation naturally creates periodic patterns with period 2π/k ∗ = 2π 2a,
independent of the initial conditions!
After some t, these solutions become too big and we can no longer ignore the nonlinear
terms. But, the stage is set and the linear problem has already selected the period of
the solution!
The vibration of the string stays in the plane: there is no dependence on the transverse
variable y.
28 CHAPTER 2. TRAVELING AND STANDING WAVES
T x
x x + ∆x
Tension is uniform: a string extends a force only in the direction parallel to the string.
In other words, the force a piece of string exerts on neighboring pieces , keeping the
string together, is tangential to the string.
We assume that tension is constant anywhere along the string.
There are no other forces.
All vibrations are small. This is a physical way of saying that, mathematically, we will
be ignoring nonlinear effects.
Next, we apply Newton’s law of motion to an itty bitty7 piece of string, lying between x
and x + ∆x, where ∆x is very small. We have
Mass of S × Acceleration of S = Net force on S.
This is Newton’s law, perpendicular to the x axis. We could also write it parallel to the x
axis, but that would result in a perfect force balance. This would offer no information about
u(x, t), which is the vertical displacement.
First, we get an expression for the mass of S.
Mass of S = ρ × arclength
Z x+∆x p
=ρ 1 + u2x (s, t)ds
x
Z x+∆x
1 2
≈ρ 1 + ux (s, t) + . . . ds
x 2
≈ ρ∆x,
7
Technical term.
2.9. A DERIVATION OF THE WAVE EQUATION 29
where we have used that the vibrations are small, thus all nonlinear terms are ignored.
Next, the acceleration of S is simply utt (x, t), by definition. Last, we turn to the net
force. The net force is pulling on the left and right ends of S by the string parts to the
immediate left and right of S. On the left end, the tension pulls with magnitude T in the
direction of the tangent vector. This normalized tangent vector is given by
−(1, ux (x, t))
p .
1 + u2x (x, t)
Thus the left force is
−T (1, ux (x, t)) 1 2
p ≈ −T (1, ux ) 1 − ux + . . .
1 + u2x (x, t) 2
≈ −T (1, ux ).
It follows that the left force in the vertical direction is −T ux (x, t). We repeat these consid-
erations on the right end. The tangent vector is
(1, u (x + ∆x, t))
p x ,
1 + u2x (x + ∆x, t)
so that the force is
T (1, ux (x + ∆x, t)) 1 2
p ≈ −T (1, ux (x + ∆x)) 1 − ux (x + ∆x, t) + . . .
1 + u2x (x + ∆x, t) 2
≈ −T (1, ux (x + ∆x, t)).
Thus, after ignoring the nonlinear terms, the right force in the vertical direction is given by
T ux (x + ∆x, t).
Combining all of this, Newton’s law becomes
ρ∆xutt (x, t) = T (ux (x + ∆x, t) − ux (x, t))
ux (x + ∆x, t) − ux (x, t)
⇒ ρutt (x, t) = T .
∆x
Taking the limit ∆x → 0, we obtain
ρutt = T uxx ,
or
utt = c2 uxx ,
where
T
c2 = ,
ρ
and we have derived the wave equation! We have already seen that f (x − ct) and f (x + ct)
are traveling wave solutions of the wave equation. It follows that their velocity is given by
s
T
±c = ± .
ρ
30 CHAPTER 2. TRAVELING AND STANDING WAVES
utt = c2 uxx .
We already know that this equation has traveling wave solutions f (x − ct) and g(x + ct),
for arbitrary profiles f and g. We show that any solution of the wave equation is a linear
combination of such traveling waves.
Inspired by the form of the traveling waves, we use a coordinate transformation
ξ = x − ct,
η = x + ct.
∂u ∂u ∂ξ ∂u ∂η
= +
∂x ∂ξ ∂x ∂η ∂x
= uξ · 1 + uη · 1
= uξ + uη .
∂ 2u ∂
2
= (uξ + uη )
∂x ∂x
∂ ∂ξ ∂ ∂η
= (uξ + uη ) + (uξ + uη )
∂ξ ∂x ∂η ∂x
∂ ∂
= (uξ + uη ) · 1 + (uξ + uη ) · 1
∂ξ ∂η
∂ ∂
= (uξ + uη ) + (uξ + uη )
∂ξ ∂η
= (uξξ + uηξ ) + (uξη + uηη )
= uξξ + 2uξη + uηη ,
where we have assumed that u is smooth, so that uξη = uηξ . Next, we do the same for the t
derivatives.
∂u ∂u ∂ξ ∂u ∂η
= +
∂t ∂ξ ∂t ∂η ∂t
= uξ · (−c) + uη · c
= −cuξ + cuη .
2.10. D’ ALEMBERT’S SOLUTION OF THE WAVE EQUATION 31
where g is any function of η. The first term is the anti-derivative of any function of ξ, so it’s
another arbitrary function of ξ. Let’s denote it by f . Thus we have that any solution of the
wave equation can be written in the form
u = f (ξ) + g(η),
or, returning to the original variables:
u = f (x − ct) + g(x + ct),
which proves our claim.
Example. You can easily show that u = cos t sin x solves the wave equation with c = 1.
Strange. . . . It doesn’t appear to be of the form given above. But it is, as we show now. Fun
with trig identities! From the trig addition formulas, we have
cos t sin x + sin t cos x = sin(x + t)
cos t sin x − sin t cos x = sin(x − t),
32 CHAPTER 2. TRAVELING AND STANDING WAVES
utt = c2 uxx
u(x, 0) = f (x)
ut (x, 0) = g(x),
where f and g are given functions. Here f (x) represents the initial position of the string and
g(x) represents its initial velocity.
We know that u(x, t) can be written as
Our task, should we choose to accept it8 , is to find F and G in terms of f and g. At t = 0,
we have u(x, t) = f (x), thus
f (x) = F (x) + G(x).
Taking a time derivative of our solution formula, we have
1
−F 0 (x) + G0 (x) = g(x)
cZ
1 x
⇒ −F (x) + G(x) = g(s)ds + α,
c 0
8
We will. We’re awesome. We’re fearless.
2.11. CHARACTERISTICS FOR THE WAVE EQUATION 33
1 x−ct 1 x+ct
Z Z
1 α 1 α
u(x, t) = f (x − ct) − g(s)ds − + f (x + ct) + g(s)ds +
2 2c 0 2 2 2c 2
Z 0 Z x+ct 0
1 1 1 1
= f (x − ct) + g(s)ds + f (x + ct) + g(s)ds
2 2c x−ct 2 2c 0
1 x+ct
Z
1
= (f (x − ct) + f (x + ct)) + g(s)ds.
2 2c x−ct
This is the d’ Alembert solution for the initial-value problem for the wave equation.
Don’t take this for granted. The wave equation is one of very few PDEs for which we
can write down the solution of the initial-value problem so explicitly!
Example. Consider the initial-value problem
utt = c2 uxx
2
u(x, 0) = e−x
ut (x, 0) = 0,
x0
ξ = x − ct,
τ = t,
u = f (x − ct),
where f (x) = u(x, 0), the initial profile. Thus, the transport equation simply moves the
initial profile to the right with speed c, where we have assumed that c > 0.
This is illustrated in Fig. 2.8. The straight lines x − ct = x0 , originating at (x0 , 0) are
called the characteristics. It follows that the value of u(x, t) along these characteristics is
given by
u(x, t) = f (x − ct) = f (x0 ).
Thus, u is constant along the characteristics: if we know u anywhere along a characteristic,
when we know it anywhere on that characteristic.
We aim to get a similar understanding for the dynamics of the wave equation
utt = c2 uxx .
t0
x0 − ct0 x0 x0 + ct0
Figure 2.9: The domain of dependence (grey) for the solution of the wave equation.
It is clear from this formula that the value of u at (x = x0 , t = t0 ) is determined only by the
initial values for x ∈ [x0 − ct0 , x0 + ct0 ]. This is illustrated in Fig. 2.9.
The cone given by the grey region in Fig. 2.9 is called the domain of dependence. It
indicates that the solution at place x0 and time t0 depends on all values in that cone, but
on no values outside of it.
If we consider the special situation where
ut (0, t) = 0,
in other words, the string starts with an initial profile, but is released without any initial
velocity, then d’Alembert giveth
1
u(x, t) = (u(x − ct, 0) + u(x + ct, 0)),
2
which shows that u depends only on the initial condition on the boundary of the domain
of dependence. These two boundary lines are called the characteristics for the wave
equation. Thus, the characteristics determine the domains of dependence.
By turning all of the above around, we can answer the question of which (x, t) points
have a given point (x0 , t0 ) in their domain of dependence. This is known as the domain of
influence. It is illustrated in Fig. 2.10. The domain of influence is also bordered by the
characteristics, but now on the lower side.
The domain of influence shows that information in the wave equation travels at a finite
speed c: it takes a definite time for the effect from x0 to be felt at any other x.
Example. Sometimes the characteristics may be used to examine the solution of an
36 CHAPTER 2. TRAVELING AND STANDING WAVES
(x0 , t0 )
Figure 2.10: The domain of influence (grey) of the point (x0 , t0 ) for the wave equation.
utt = 4uxx
1 if x ∈ [0, 1]
u(x, 0) =
0 if x ∈
6 [0, 1]
ut (x, 0) = 0.
c2 = 4 ⇒ c = 2.
1/2
1/2 1/2
0 (1/2, 1/4) 0 x
1
0 0 1 1 0
Figure 2.11: The different solution regions for the example problem. The values in blue are
values of u(x, t) in that region. The characteristics drawn have slope 1/2.
condition on the function we are looking for, given at the end of our physical domain.
Thus, we want to solve the boundary-value problem
utt = c2 uxx , x ∈ (0, ∞)
u(0, t) = 0, for all t > 0
,
u(x, 0) = f (x), x ∈ (0, ∞)
ut (x, 0) = g(x), x ∈ (0, ∞)
where f (x) and g(x) are given functions. We will solve this using d’Alembert’s solution and
the insight we gained from using characteristics. From d’Alembert,
Z x+ct
1 1
u(x, y) = (f (x − ct) + f (x + ct)) + g(s)ds.
2 2c x−ct
This formula can only be valid if x − ct > 0. Otherwise, we’d be asking to evaluate f (x) at
arguments that are negative, but we have been given f (x) only for positive arguments. A
conundrum! Thus the solution ceases to be valid when
t=0
x
0 0
0 1
1
t = 1/4 1/2 x
0 0
1/2
1/2 1/2 x
t > 1/4 0 0
0
1/2
Figure 2.12: The different stages of the solution. As before, values of u are in blue.
since all solutions of the wave equation are of this form. Using the boundary condition, we
have
0 = F1 (−ct) + G(ct)
⇒ F1 (z) = −G(−z),
so that
u(x, t) = −G(−x + ct) + G(x + ct).
as before. Note that to determine G(x + ct), it is fine to use the initial conditions, since the
information in G(x + ct) comes only from positive values of x. Alternatively, we can impose
the continuity of our solution at x = ct, finding the same result for G(x). Either way, it
2.12. THE WAVE EQUATION ON THE SEMI-INFINITE DOMAIN 39
Figure 2.13: The wave equation string, on the half line x > 0 with fixed end u(0, t) = 0, for
all t.
follows that
1 x+ct 1 −x+ct
Z Z
1 1
u(x, t) = f (x + ct) + g(s)ds − f (−x + ct) − g(s)ds
2 2c 0 2 2c 0
1 ct+x
Z
1
= (f (x + ct) − f (ct − x)) + g(s)ds,
2 2c ct−x
which is valid for t > x/c, while we still have that
Z x+ct
1 1
u(x, t) = (f (x − ct) + f (x + ct)) + g(s)ds,
2 2c x−ct
for t ≤ x/c.
Let’s examine the characteristics to make sense out of this, see Fig. 2.15. At x = 0,
the region of dependence would usually come partially from x < 0, but that is not allowed
now. Rather, all the information comes from x > 0, and from the boundary itself. We can
pretend there is an x < 0 region, as long as we always satisfy u(0, t) = 0, for all t. This is
easily done: let us extend u for x < 0 to be the opposite of u for the corresponding positive
value −x. In other words, the overall u(x, t) is odd, as a function of x: u(−x, t) = −u(x, t):
u(x, t), x≥0
û(x, t) = ,
−u(−x, t) x≤0
then clearly û(0, t) = 0. Also, the function behaves as if it switches sign every time it hits the
boundary, as the derivative will be even9 . Indeed, when the value from point (a) in Fig. 2.15
hits the boundary, it meets the value from (b), which has the opposite value. This opposite
value continues on for x > 0. Thus, in effect, the value flips at the boundary.
9
Recall that the derivative of an even function is odd and the derivative of an odd function is even. Whoa!
40 CHAPTER 2. TRAVELING AND STANDING WAVES
t = x/c
Figure 2.14: Our original d’Alembert solution is not valid in the shaded region, which
corresponds to t > x/c.
Free end
We consider the boundary-value problem
utt = c2 uxx , x ∈ (0, ∞)
ux (0, t) = 0, for all t > 0
,
u(x, 0) = f (x), x ∈ (0, ∞)
ut (x, 0) = g(x), x ∈ (0, ∞)
Thus, according to the boundary condition, the string is horizontal at the boundary, for all
time. We proceed as before. By d’Alembert,
Z x+ct
1 1
u(x, t) = (f (x − ct) + f (x + ct)) + g(s)ds.
2 2c x−ct
By the same reasoning as for the fixed-end case, this is only valid when x − ct > 0, and
another solution form has to be found in the shaded region of Fig. 2.14. Again this is
expected: in the shaded region, part of the solution would come from x < 0, but we have no
information there.
As always, we have that in the shaded region10
u=0
(a) (b) x
x<0 0 x>0
Figure 2.15: The characteristics for the wave equation on the half line, including the
“phantom” characteristics for x < 0.
where G(z) is determined as before (once again: the information in G(z) comes only from
positive x values):
1 z
Z
1
G(z) = f (z) + g(s)ds.
2 2c 0
Thus
1 x+ct
Z
1
u(x, t) = F2 (x − ct) + f (x + ct) + g(s)ds.
2 2c 0
In order to impose the boundary condition, we calculate
1 1
ux (x, t) = F20 (x − ct) + f 0 (x + ct) + g(x + ct).
2 2c
Evaluating this at x = 0, we get
1 1
0 = F20 (−ct) + f 0 (ct) + g(ct)
2 2c
0 1 0 1
⇒ F2 (−ct) = − f (ct) − g(ct).
2 2c
Equating −ct = z, we get
1 1
F20 (z) = − f 0 (−z) − g(−z)
2 2c
1 −z
Z
1
⇒ F2 (z) = f (−z) + g(s)ds,
2 2c 0
by the fundamental theorem of calculus. Thus
Z 0
1 1
F2 (x − ct) = f (ct − x) − g(s)ds,
2 2c ct−x
42 CHAPTER 2. TRAVELING AND STANDING WAVES
If we let g(x) ≡ 0, we can understand this result, using the characteristics, see Fig. 2.15.
In order to satisfy the boundary condition, we extend u(x, 0) to negative values to be an
even function. Then its derivative will be odd, and its value at 0 will be 0. We see that at
(x, t), there are two contributions:
The first one is f (x + ct)/2, the second is given by f (ct − x)/2, by our result above.
represents a standing wave. It looks like 2 sin x at all time t, but multiplied by a time-
dependent amplitude cos t. It is illustrated in Fig. 2.16. It is clear11 that standing waves can
often be written as linear combinations of traveling waves. Here we have
utt = c2 uxx ,
we let
u(x, t) = T (t)X(x).
Thus X(x) represents the spatial profile, while T (t) gives the temporal part of the profile,
its time-dependent amplitude. We get
t=0
t = π/4
t = π/2 x
t=π
Figure 2.16: The standing wave profile u = 2 cos t sin x, for different values of t.
Note that the left-hand side depends on t only, while the right-hand side is a function of
only x. It follows that both sides have to be constant. Indeed, if we take an x derivative of
the equation, we get 00 0
2 X
0=c ,
X
since the left-hand side does not depend on x. Thus c2 X 00 /X is constant. The same argument,
but by taking a derivative with respect to t, gives that T 00 /T is constant too. Thus we have
T 00 X 00
= c2 = λ,
T X
where λ is a constant. We get two ordinary differential equations:
00
T = λT,
X 00 = cλ2 X.
Here the first equation is an ODE in t, while the second one depends on x.
As you know, we get different kinds of solutions, depending on the sign of λ. Let us
investigate all possibilities.
λ = 0. we get T 00 = 0 and X 00 = 0. It follows that T = A + Bt and X = C + Dx. Thus
u = (C + Dx)(A + Bt).
In particular, we want D = 0 and B = 0, if x ∈ R, since we want bounded solutions.
On the other hand, if x is restricted to a smaller domain, the solution with D 6= 0 6= B
may be perfectly acceptable.
44 CHAPTER 2. TRAVELING AND STANDING WAVES
λ < 0. For the last case, we write λ = −r2 , with r > 0. We have to solve
00
T = −r2 T,
2
X 00 = − rc2 X,
T = A cos(rt + B sin(rt),
and
X = C cos(rx/c) + D sin(rx/c),
resulting in
All of these result in standing wave solutions of the wave equation. Next, we impose
some boundary conditions.
utt = c2 uxx ,
u(0, t) = 0,
u(L, t) = 0.
12
Or, you could use hyperbolic functions. Always fun!
2.13. STANDING WAVE SOLUTIONS OF THE WAVE EQUATION 45
We have that
u(x, t) = X(x)T (t),
for standing waves. We want u(0, t) = 0, which implies that
X(0)T (t) = 0.
Since we don’t want that T (t) = 013 , we need that X(0) = 0. Similarly, we need that
X(L) = 0. We already know the allowed forms for X(x):
X(x) = C + Dx,
or X(x) = Cerx/c + De−rx/c ,
or X(x) = C cos(rx/c) + D sin(rx/c).
We now impose the conditions
X(0) = 0, X(L = 0),
on these possibilities.
With X = C + Dx, we get
X(0) = 0 = C,
X(L) = 0 = C + DL,
from which it follows that both C and D are zero, so that X = 0. Not interesting.
Next, we consider X = Cerx/c + De−rx/c . Imposing X(0) = 0, we get
C + D = 0 ⇒ D = −C.
This allows us to rewrite X(x) as
X(x) = C(erx/c − e−rx/c ) = 2C sinh(rx/c).
Next, we impose X(L) = 0. We get
0 = 2C sinh(rL/c).
Clearly14 , we don’t want C = 0, since this would result in u(x, t) = 0. That is not
exciting. If we want excitement, we have to impose
sinh(rL/c) = 0.
Unfortunately, the sinh function is zero only when it’s argument is zero, which would
imply r = 0, since L 6= 0 (otherwise we’d have no string. But r = 0 is not allowed for
this case, since r = 0 implies λ = 0, which was the previous case. Bummer. All that
work and no solutions. What is this? A course on how not to find solutions to PDEs?
Patience, my young apprentices. . .
13
Resulting in only the extremely boring u = 0 solution. . .
14
The most hated word in a math text, with the possible exception of “obviously”.
46 CHAPTER 2. TRAVELING AND STANDING WAVES
Inevitably, we end up with all our money on the last case15 . Imposing X(0) = 0, with
we get
C = 0.
Not a good start! We’ve just thrown out half of our last remaining non-zero solutions.
This leaves us with
X(x) = D sin(rx/c).
Next we impose X(L) = 0. We obtain
D sin(rL/c) = 0.
rL
= nπ,
c
where n is any integer. Since r is not allowed to be zero, we have to exclude n = 0.
Further, since we may assume that r > 0, we need only consider n ∈ Z+ 0 , the set of
strictly positive integers: n ∈ {1, 2, 3, 4, . . .}. Thus we have found
nπx
Xn (x) = sin , n = 1, 2, . . .
L
We can ignore the constant multiplication factor D, since it can be absorbed into the
multiplying function T (t), which we still have to determine. We have endowed X(x)
with an index n, to distinguish the different solutions we have found.
In summary, we find that the only standing wave solutions of the wave equation that
satisfies the boundary conditions for a fixed finite-length string are given by
nπct nπct nπx
un (x, t) = (A cos + B sin ) sin ,
L L L
for all n = 1, 2, . . ..
Modes of vibration
We call un the n-th mode of vibration of the string. The wave number of the n-th mode is
nπ/L. Over the domain x ∈ (0, L), the solution has exactly n − 1 zeros16 Note that all of
the Xn are sin functions, with increasing wave number as n increased. In other words, for
15
This statement is in no way an endorsement of any gambling activity. The University of Washington
and its employees do not encourage gambling in any way or form. Except for solving PDEs.
16
You should convince yourself of this statement. If you have extra time, convince your neighbor too.
2.13. STANDING WAVE SOLUTIONS OF THE WAVE EQUATION 47
larger n, Xn has a smaller period. For increasing n we are simply cramming more periods
in the interval [0, L].
We briefly revisit the x-problem we solved:
c2 X 00 = λX,
X(0) = 0,
X(L) = 0.
This is called a Sturm-Liouville problem for X(x). The constant λ is called the eigen-
value, X(x) the eigenfunction corresponding to λ.
The method we have used to solve for the modes of the wave equation is called Sepa-
ration of Variables, because we looked for that solutions that are multiples of functions
that depend on x and t separately.
In the next chapter, we look at how we can construct very general solutions from linear
superpositions of these standing wave solutions.
48 CHAPTER 2. TRAVELING AND STANDING WAVES
Chapter 3
49
50 CHAPTER 3. FOURIER SERIES
As you repeat the calculation above for this problem, you encounter a problem with the
second boundary condition. Bummer!
The above implies we can add our standing wave solutionsto get new, more general
solutions of the wave equation. Let
N
X
u(x, t) = un (x, t)
n=1
N
X nπct nπct nπx
= An cos + Bn sin sin .
n=1
L L L
Let
N
X
u(x, t) = [An cos (nπt) + Bn sin (nπt)] sin (nπx) .
n=1
At this point, we have satisfied the PDE and the boundary conditions. From the initial
conditions, we get
N
X
u(x, 0) = 0 = An sin(nπx),
n=1
3.1. SUPERPOSITION OF STANDING WAVES 51
and
N
X
ut (x, 0) = [(−An )nπ sin(nπt) + Bn nπ cos(nπt) sin(nπx)]|t=0
n=1
XN
= Bn nπ sin(nπx)
n=1
= 2 sin πx − 3 sin 4πx.
The first equation is easily satisfied by choosing
An = 0,
for all n. Next, we can choose N = 4, so that the second equation becomes
B1 π sin πx + B2 2π sin 2πx + B3 3π sin 3πx + B4 4π sin 4πx = 2 sin πx − 3 sin 4πx,
which is solved by choosing
2 −3
B1 =
, B2 = 0, B3 = 0, B4 = .
π 4π
This solves the given initial-value problem.
This whole thing feels like a cheat, right? Those were very special initial conditions! Can
we do something in general? For starters, can we show that the above is the unique solution
to the problem? The answers to these questions will be “Yes!” and “Yes!”1 .
Let’s kick this up a notch© . How about if we include all standing wave solutions we
know? Can we let N → ∞? We would have
∞
X nπct nπct nπx
u(x, t) = An cos + Bn sin sin .
n=1
L L L
Does this make sense? Perhaps. If An , Bn → 0 sufficiently fast, we might get a convergent
series. Such a convergent series is called a Fourier series. The big trick is to see whether
we can find An and Bn such that we can satisfy general initial conditions, not just the really
special ones we used above. Perhaps this is possible.
Example. Suppose the initial condition is given as
1 1
u(x, 0) = sin πx − sin 3πx + sin 5x − . . .
9 25
∞
X (−1)n+1
= sin(2n − 1)πx,
n=1
(2n − 1)2
and ut (x, 0) = 0. Clearly, we would need
1 1
A1 = 1, A2 = 0, A3 = − , A4 = 0, A5 = , . . .
9 25
Can we do this for more general initial conditions?
1
Yes, with exclamation points.
52 CHAPTER 3. FOURIER SERIES
since the integral of the second term is zero. Combining these two results, we get the desired
identity: Z L nπx mπx
sin sin dx = Lδnm .
−L L L
The others are proven in a similar way.
Back to our main business: we wish to find an and bn such that
∞
X nπx nπx
f (x) = A + an cos + bn sin .
n=1
L L
Since the average of all the sines and cosines is zero, we get
Z L
1
A= f (x)dx.
2L −L
1 L
Z mπx
bm = f (x) sin dx, m = 1, 2, . . . .
L −L L
Similarly, we get
Z L
1 mπx
am = f (x) cos dx, m = 1, 2, . . . .
L −L L
with
1 L
Z nπx
an = f (x) cos dx, n = 0, 1, 2, . . . ,
L −L L
1 L
Z nπx
bn = f (x) sin dx, n = 1, 2, . . . .
L −L L
This series for f (x) using trig functions is called its Fourier series. It arises not only in
wave and PDE problems, but also in image processing, data analysis, etc.
Example. Consider
f (x) = x2 , x ∈ [−1, 1],
as plotted in Fig. 3.1. For this example, L = 1, and we get
1 L
Z nπx
an = f (x) cos dx
L −L L
Z 1
= x2 cos(nπx)dx
−1
4nπ(−1)n
=
n3 π 3
4(−1)n
= 2 2 ,
nπ
where we have used integration by parts a few times4 . Next,
1 L
Z nπx Z 1
bn = f (x) sin dx = x2 sin(nπx)dx = 0,
L −L L −1
4
Per the Constitution, you should check this.
3.2. FOURIER SERIES 55
f (x)
Figure 3.1: The function f (x) = x2 over its domain of definition x ∈ [−L, L].
since the integrand is an odd function. It follows that for x ∈ [−1, 1],
∞
a0 X
x2 = + an cos nπx
2 n=1
∞
1 4 X (−1)n
= + 2 cos nπx.
3 π n=1 n2
f (x)
is the most general solution of the wave equation that satisfies the boundary conditions,
found using separation of variables. It remains to impose the initial conditions. First, by
letting t = 0, we get
∞
X nπx
f (x) = A + an sin .
n=1
L
Next, we take a derivative of u(x, t) with respect to t, and we let t = 0. This results in
∞
X nπc nπx
g(x) = bn sin .
n=1
L L
Because of the boundary conditions (fixed end), we use an odd extension of f (x). Using the
Fourier sine series, we get
A = 0,
and Z L
2 nπx
an = f (x) sin dx,
L 0 L
for n = 1, 2, . . .. Similarly, we use an odd extension for g(x), since the above indicates we
wish to use a Fourier sine series. We get
2 L
Z
nπc nπx
bn = g(x) sin dx,
L L 0 L
Z L
2 nπx
⇒ bn = g(x) sin dx.
nπc 0 L
also for n = 1, 2, . . ..
This completely determines the solution of the wave equation problem with the given
initial data.
Let’s consider a different problem, namely that with two free ends.
We begin by using separation of variables to find the standing wave solutions. Let
Then
T 00 X 00
= = λ,
c2 T X
so that λ has to be constant. It follows that
X 00 = λX.
Note that the boundary conditions imply that X 0 (0) = 0 = X 0 (L). We consider three
different cases.
λ = 0. We have
X 00 = 0 ⇒ X = D,
where we have already used the boundary conditions. This case results in a constant
(as a function of x) standing wave, given by
u0 (x, t) = A0 + B0 t,
where we have equated D = 1, since we can absorb the constant in the values of A0
and B0 .
0 = c1 r sinh(rL),
which requires c1 = 0, so that the exponential case does not result in any solutions.
λ < 0. Now we let λ = −r2 , again with r > 0. We know we get solutions of the form
−c1 r sin(rL) = 0.
Using superposition, we get that the most general solution satisfying the boundary
conditions is given by
∞ nπx
X nπct nπct
u(x, t) = A0 + B0 t + cos An cos + Bn sin .
n=1
L L L
and Z L
1
A0 = f (x)dx,
L 0
using the formulae for the Fourier cosine series. Lastly, to impose the second initial
condition, we take a derivative with respect to t, and we let t = 0. We get
∞
nπcX nπx
g(x) = B0 + Bn cos .
n=1
L L
60 CHAPTER 3. FOURIER SERIES
1 L
Z
B0 = g(x)dx,
L 0
and
2 L
Z
nπc nπx
Bn = g(x) cos dx
L L 0 L
Z L
2 nπx
⇒ Bn = g(x) cos dx.
nπc 0 L
Note that in a practical problem, you might have to impose that the average of the
initial velocity g(x) is zero. If this is not the case, your solution to the string-with-
free-ends problem will have a component that is linearly growing!
Here σ > 0 is the heat conductivity coefficient. The heat equation describes heat flow in a
medium with heat conductivity σ. We will consider only the one-dimensional heat equation.
As for the wave equation, we begin by looking for solutions of the form
We get
XT 0 = σX 00 T
T0 X 00
⇒ = = λ.
σT X
Here λ is a separation constant, using the same argument we used for the wave equation:
since the left-hand side depends only on t, and the right-hand side depends only on x, they
must both be constant. It follows that X(x) satisfies the following problem:
X 00 − λX = 0,
X(0) = 0,
X(L) = 0.
3.5. LAPLACE’S EQUATION 61
This is the same exact problem for X as we had for the wave equation with fixed ends
(Dirichlet boundary conditions). As a consequence, we know we get solutions only for
λ = −r2 < 0, and nπx
Xn (x) = sin , n = 1, 2, . . . ,
L
and
n2 π 2
λn = − 2 .
L
It follows that Tn satisfies the ordinary differential equation
Tn0 n2 π 2 2 π 2 t/L2
= 2 ⇒ Tn = e−σn .
σTn L
We could have included a multiplicative constant, but this is not necessary, as the next step
is to take a linear superposition of the solutions un (x, t) = Xn (x)Tn (t) we have just found:
2 2 2
nπx
un (x, t) = e−σn π t/L sin .
L
The superposition results in the general solution
∞ nπx
−σn2 π 2 t/L2
X
u= cn e sin .
n=1
L
since all the exponentials become 1, when evaluated at t = 0. Using an odd extension of
f (x), we get that
2 L
Z nπx
cn = f (x) sin dx, n = 1, 2, . . . .
L 0 L
Thus, we have solved the initial-value problem for the heat equation with prescribed zero
temperature at the ends5 .
This is Laplace’s equation, posed on a rectangle. You could imagine getting to Laplace’s
equation by wanting to find time-independent (or stationary) solutions of the multi-dimensional
5
That was quick! We’re getting to be good at this.
62 CHAPTER 3. FOURIER SERIES
M
u(x, M ) = 0
ux (0, y) = 0 ux (L, y) = 0
0 u(x, 0) = f (x) L
wave equation utt = c2 (uxx +uyy ) or of the multi-dimensional heat equation ut = σ(uxx +uyy ).
We impose boundary conditions as follows, see Fig. 3.3:
u(x, 0) = f (x),
ux (0, y) = 0,
ux (L, y) = 0,
u(x, M ) = 0.
and we get
X 00 Y 00
X 00 Y + XY 00 = 0 ⇒ =− = λ.
X Y
As before, λ is a constant, since X 00 /X and −Y 00 /Y are dependent on x and y separately.
We have that
X 00 − λX = 0,
X 0 (0) = 0,
X 0 (L) = 0.
λ > 0. You should check that this does not result in any solutions for X(x).
3.5. LAPLACE’S EQUATION 63
X = c1 cos(rx) + c2 sin(rx).
0 = c2 r ⇒ c2 = 0.
n2 π 2
Yn00 − Yn = 0
L2 nπ nπ
⇒ Yn = An cosh (y − M ) + Bn sinh (y − M ) ,
L L
where we have opted to write the solution using hyperbolic functions of a shifted
argument. Imposing the boundary condition yn (M ) = 0, we find that
An = 0.
It follows that nπ
Yn (y) = Bn sinh (y − M ) .
L
The linear superposition of all solutions is7
∞
X nπ nπx
u(x, y) = B0 (y − M ) + Bn sinh (y − M ) cos .
n=1
L L
6
For now. Don’t worry. It’ll come back.
7
It’s back!
64 CHAPTER 3. FOURIER SERIES
the rest is some simple Fourier series stuff, using an even extension8 of f (x)!
For n = 0,
Z L
1
−M B0 = f (x)dx
L 0
Z L
1
⇒ B0 = − f (x)dx.
ML 0
For n = 1, 2, . . .,
Z L
nπM 2 nπx
−Bn sinh = f (x) cos dx
L L 0 L
Z L
2 nπx
⇒ Bn = − nπM
f (x) cos dx.
L sinh L 0 L
This completely determines the solution of the Laplace problem on the rectangle.
Then9
∂u ∂v ∂r ∂v ∂θ
ux = = + .
∂x ∂r ∂x ∂θ ∂x
From
r 2 = x2 + y 2 ,
8
Because the right-hand side is a cosine series.
9
You can see the chain rules coming a mile away!
3.6. LAPLACE’S EQUATION ON THE DISC 65
x2 + y 2 = R 2
it follows that
∂r ∂r x
2r = 2x ⇒ = .
∂x ∂x r
similarly, from
y
tan θ = ,
x
we get
∂θ y r2 ∂θ y ∂θ y
sec2 θ =− 2 ⇒ = − ⇒ = − 2.
∂x x x2 ∂x x2 ∂x r
Thus
x y
ux = vr − vθ 2
r r
sin θ
= vr cos θ − vθ .
r
Next, we apply this same process to get uxx = (ux )x : in other words, we repeat the above,
but with v replaced by the expression we found for ux . This gives
sin θ sin θ sin θ
uxx = vr cos θ − vθ cos θ − vr cos θ − vθ
r r r θ r
sin θ sin θ sin θ cos θ sin θ
= vrr cos θ − vrθ + vθ 2 cos θ− vrθ cos θ − vr sin θ − vθθ − vθ
r r r r θ r
2 2
sin θ cos θ sin θ cos θ sin θ sin θ
= vrr cos2 θ − 2vrθ + 2vθ + v r + vθθ .
r r2 r r2
66 CHAPTER 3. FOURIER SERIES
Similarly, we find
We get
00 S 0T T 00 S
S T+ + 2 =0
r r
00 0
S rS T 00
⇒ r2 + + =0
S S T
S 00 rS 0 T 00
⇒ r2 + =− = λ,
S S T
where λ is a separation constant. Indeed, in this last line, the left-hand side is a function of
r only, while −T 00 /T depends only on θ. Thus both are constant.
Since the equation for T is the simplest, we solve it first.
If λ = 0, then
T 00 = 0 ⇒ T = a + bθ.
We want v(r, θ) to be single-valued as a function of both r and θ. This implies that T
should be a periodic function of θ, with period 2π. Thus, b = 0. This leaves us with
T0 = 1,
α = n,
With λ0 = 0, then
r2 S000 + rS00 =0
⇒ rS000 + S00 =0
⇒ (rS00 )0 =0
⇒ rS00 = c1
⇒ S0 = c1 ln r + c2 .
r2 Sn00 + rSn0 − n2 Sn = 0.
p(p − 1) + p − n2 = 0
⇒ p2 = n2
⇒ p = ±n.
11
Don’t make me say it. OK, fine: you should check this.
12
Nothing goes to infinity on our watch!
68 CHAPTER 3. FOURIER SERIES
This gives
Sn = cn rn + dn r−n .
Sn = r n ,
up to a multiplicative constant.
v(R, θ) = f (θ).
This results in ∞
a0 X n
f (θ) = + R (an cos nθ + bn sin nθ).
2 n=1
S x
x=a x=b
Figure 4.1: A one-dimensional domain, setting up for the derivation of a conservation law.
69
70 CHAPTER 4. THE METHOD OF CHARACTERISTICS
2. Q is created or destroyed in S.
It follows that the rate of change of Q, dQ/dt, is given by the rate at which Q enters or
leaves at x = a, plus the rate at which Q enters or leaves at x = b, plus the rate at which Q
is created or destroyed in S. In equations,
Z b Z b
d
u(x, t)dx = φ(a, t) − φ(b, t) + f (x, t)dx,
dt a a
where φ(x, t) is the rate at which Q moves past x at time t. If φ(x, t) > 0, then the flow is
in the positive x direction, otherwise it is in the negative x direction. Thus, the net rate at
which Q enters through the ends of S is
The − sign for the second term is a consequence of our flow convention: if the flow of Q
through x = b is to the right, it is leaving S, thus it results in a decrease.
Lastly, if Q is created or destroyed in S, this happens with a source or sink function
f (x, t), resulting in an amount of Q that is added equal to
Z b
f (x, t)dx.
a
ut + φx = f.
Constitutive relations
Even if we consider f (x, t) as given, we still have one partial differential equation for two
quantities u and φ. A constitutitve relation relates u and φ. In many cases such a relation
gives φ as a function of u. Then φ = φ(u), and we get
ut + φ0 (u)ux = f.
ut + uux = 0,
is a conservation law with f = 0, φ = u2 /2. However, there are more possibilities. The same
equation can be written as
2 1 2 1 3
uut + u ux = 0 ⇒ u + u = 0.
2 t 3 x
If we let
1
v = u2 ,
2
then
u = (2v)1/2 ,
and
u3 1
φ= = (2v)3/2 ,
3 3
and the Burgers equation can be rewritten as
1 3/2
vt + (2v) = 0.
3 x
Which form of the equation we choose will matter in the following lectures. In practice, it
is of course dictated by the application we are working on.
ut + φx = f
is our conservation law. Assume that we have no magical pollutant eating piranhas in the
pipe, and no pollutant is destroyed or created, then f = 0. Next, we need to relate the flux
function φ to the concentration u.
72 CHAPTER 4. THE METHOD OF CHARACTERISTICS
Suppose we have an initial concentration u0 (x) shown below in Fig. 4.3. We expect that
pollutant will flow from areas where there is a lot to areas where there is less. If ux is
positive, as in (a), then φ should be negative so that pollutant flows to the left. If, at (b),
ux is negative, then φ should be positive and u will flow to the right. The simplest way to
make this happen is Fick’s Law:
ut + φx = 0
⇒ ut + (−Dux )x = 0
⇒ ut = Duxx ,
Traffic flow
This goes back to studies by Whitham and Lighthill in the 50s. We approximate the number
of cars per unit length by a continuous function u(x, t). Assuming there are no exits or
entrances, we get
f = 0,
4.3. THE METHOD OF CHARACTERISTICS 73
ut + φx = 0.
φ = cars/time unit
= Rate at which cars are passing x at t
= u × v,
where u is the car density and v is the velocity. Thus, we need to relate the velocity v to u,
and we will be done. Clearly1 , if u is high then v will be low, and if u is near zero, v should
be maximal, the speed limit. The simplest way to do this is
v = v1 − au,
u2
u
φ = uv = v1 u 1 − = v1 u − .
u1 u1
This last form is less instructional, but it is easier to take a derivative. The conservation law
becomes
ut + φx = 0
2uux
⇒ ut + v1 ux − =0
u1
2u
⇒ ut + v1 1 − ux = 0.
u1
ut + φx = f,
u(x, 0) = u0 (x),
1
That word again!
74 CHAPTER 4. THE METHOD OF CHARACTERISTICS
x(t)
x0
ut + cux = 0,
u(x, 0) = u0 (x).
The method of characteristics looks for special curves in the (x, t) plane along which our
PDE becomes an ODE. In other words, we wish to find x(t) in the (x, t) plane such that we
only have to solve ODEs along this curve, see Fig. 4.4.
Along these curves u(x, t) becomes a function of t only. Then
d dx
u(x(t), t) = ux + ut .
dt dt
We compare this expression with the PDE we wish to solve
ut + cux = 0.
We see that if we pick the curves (called the characteristic curves or characteristics)
such that
dx
= c,
dt
then our PDE simply becomes
du
(x(t), t) = 0.
dt
Thus u is constant along characteristic curves in this case. The value of the constant is of
course determined by the initial conditions.
4.3. THE METHOD OF CHARACTERISTICS 75
ut + 4ux = 0,
u(x, 0) = arctan(x).
We have to solve
du
= 0,
dt
along curves for which
dx
= 4 ⇒ x = 4t + x0 .
dt
It follows that
u = arctan(x0 ) = arctan(x − 4t).
slope 1/c
x−x0
t= c
x
x0
Proceeding as before, we look for curves x(t) so that we get ODEs. Along these curves,
d dx
u(x(t), t) = ut + ux .
dt dt
Thus we solve
dx
= c,
dt
x(0) = x0 .
The characteristics are the solutions of this system. Along the characteristics, we have to
solve
du
= f (x(t), t),
dt
which is easily solved by simply integrating both sides.
Example. Consider the nonhomogeneous problem
ut + 4ux = 1,
u(x, 0) = arctan(x).
We have to solve
dx
= 4 ⇒ x = 4t + x0 .
dt
Along these straight-line characteristics, we solve
du
= 1 ⇒ u = t + A,
dt
4.3. THE METHOD OF CHARACTERISTICS 77
u(x0 , 0) = A = arctan(x0 ).
It follows that
u = t + arctan(x − 4t).
u(x, 0) = u0 (x).
ut + xux = 0,
1
u(x, 0) = .
1 + x4
We solve for the characteristics first. We have
dx
=x
dt
d
⇒ ln |x| = 1
dt
x
⇒ ln =t
x0
⇒ x = x0 et .
Next, since
du
= 0,
dt
78 CHAPTER 4. THE METHOD OF CHARACTERISTICS
It follows that
1
u= ,
1 + x4 e−4t
where we have used that x0 = x exp(−t). The characteristics are not straight lines in this
case. They are shown in Fig. 4.6.
ut + φ0 (u)ux = f.
Define
c(u) := φ0 (u).
4.3. THE METHOD OF CHARACTERISTICS 79
du
= 0 ⇒ u = u0 (x0 ).
dt
Thus this equation tells us that u is constant along characteristics, even though at this point
we do not yet know what these characteristics are. We can substitute this result into our
ODE for the characteristics:
dx
= c(u0 (x0 )).
dt
Since the right-hand side is constant, we get
It follows that the characteristics are all straight lines, but with varying slopes depending
on x0 , as illustrated in Fig. 4.7
The slope depends not only on where (x0 ) we start, but also on what the initial condition
u0 is there. In order to get the full solution to the problem, we need to solve the characteristic
equation x = x0 + tc(u0 (x0 )) for x0 as a function of x and t. More often than not, this is
not possible. In most cases, we have to be satisfied with an implicit representation of the
solution.
Example. Let φ = u2 /2. Then c(u) = u, and we consider the initial-value problem
ut + uux = 0,
0, x ≤ 0,
u(x, 0) = −1/x
e , x > 0.
x = x0 , x0 ≤ 0,
−1/x0
x = x0 + te , x0 > 0.
80 CHAPTER 4. THE METHOD OF CHARACTERISTICS
x
x0
slope= c(u01(x0 ))
Figure 4.7: The characteristics for different values of x0 with different slopes 1/c(u0 (x0 )).
There is no way for us2 to solve this3 for x0 as a function of x and t. However, we can obtain
a perfectly fine implicit solution:
0, x ≤ 0,
u(x, t) =
e−1/x0 , x > 0,
x = x0 + te−1/x0 .
are all straight lines, with slope 1/c(u0 (x0 )), where x0 is the starting point of the character-
istic at t = 0.
Even though the characteristics are just straight lines, things4 can get very interesting.
Let’s see what can happen.
2
Or anyone!
3
Other than numerically.
4
Technical term.
4.4. BREAKING AND GRADIENT CATASTROPHES 81
Figure 4.8: The characteristics for the problem with the piecewise defined initial condition.
u0 (x)
2. The characteristics are spreading out. Starting from an arctan-like profile, we get
the situation depicted in Fig. 4.11. Now the values of the solution get spread out, as
the values follow the characteristics. Thus, spreading characteristics lead to smoother
solutions.
82 CHAPTER 4. THE METHOD OF CHARACTERISTICS
u0 (x − ct)
u0 (x)
Figure 4.10: Moving the initial condition along the parallel characteristics.
3. The characteristics are crossing, as in Fig. 4.12. If the characteristics cross, then
the values of u0 (x) between two crossing characteristics starting at a and b get squeezed
together as t increases. Thus the solution becomes locally steeper, since rise/run→ 0,
since run→ 0 as t approaches the time at which the characteristics cross. At that time,
the solution becomes infinitely steep (i.e., it has a vertical tangent), which implies
that the differential form of the conservation law is no longer valid. Indeed,
to derive the differential form, we assumed that all derivatives of u existed and were
continuous. If one or all of these derivatives → ∞, we have to revisit the integral form
of the conservation law. Up to the time where we first get the vertical tangent, the
differential form works well. Past that time, we have a problem. As is seen in Fig. 4.12,
the crossing characteristics give rise to a wedge-like region where any point has multiple
characteristics passing through it. Outside of this wedge, we can immediately see what
the value of u is by tracing back the unique characteristic going through the point. The
value of the initial condition on that characteristic is the value of the solution. For
points in the wedge, this does not work, as it is unclear what characteristic to follow
back.
The formation of the solution with a vertical tangent is called a gradient catastro-
phe5 .
Let’s do a few examples.
Example. Consider the problem
ut + uux = 0,
5
Wow. Sounds serious. It is.
4.4. BREAKING AND GRADIENT CATASTROPHES 83
Figure 4.11: Moving the initial condition along the spreading characteristics.
u(x, 0) = arctan(x).
u(x, t) = arctan(x0 ),
x = x0 + tc(u0 (x0 ))
= x0 + tu0 (x0 )
= x0 + t arctan(x0 ).
We cannot solve this last equation for x0 as a function of x and t, and an implicit solution
is the best we can do. However, we can plot the characteristics. We have
x − x0
t= .
arctan(x0 )
These characteristics, for varying x0 are plotted in Fig. 4.11. We see that the characteristics
are spreading out, and the initial profile of a front-like arctan becomes less steep as time
progresses, since the characteristics in the region of the front are fanning out.
Example. Consider the problem
ut + uux = 0,
u(x, 0) = − arctan(x).
84 CHAPTER 4. THE METHOD OF CHARACTERISTICS
u(x, t = 0.8)
u0 (x)
x
Figure 4.12: Moving the initial condition along the crossing characteristics.
This is almost the same problem as above, but the sign of the initial condition is flipped.
The implicit solution to this problem is given by
u(x, t) = − arctan(x0 ),
x = x0 + tc(u0 (x0 ))
= x0 + tu0 (x0 )
= x0 − t arctan(x0 ).
As before, we cannot solve this last equation for x0 as a function of x and t, and an implicit
solution is the best we can do. However, we can plot the characteristics. We have
−x + x0
t= .
arctan(x0 )
These characteristics, for varying x0 are plotted in Fig. 4.13 (left). We see that the charac-
teristics are crossing, with an apparent gradient catastrophe occurring at t = 16 . A wedge
region where the characteristics cross is formed. A few time slices of solution profiles are
shown in the right panel of Fig. 4.13. We note that the front profile becomes steeper as we
approach the gradient catastrophe time, referred to as the breaking time.
As discussed, if the characteristics cross, we will get ux , ut → ∞, as t → tb , the so-called
breaking time. Now what? The PDE is no longer and we need to rethink what we are doing.
6
We’ll check whether this is correct soon.
4.4. BREAKING AND GRADIENT CATASTROPHES 85
t u(x, 1)
u(x, 0.8)
u(x, 0.4) x
u(x, 0)
Figure 4.13: Crossing characteristics and the solution profiles that go with them.
We know that the slope of the characteristics is the inverse of the velocity:
1
slope = .
c(u0 (x0 ))
Let’s look at the effect this has on different profiles. Let’s assume, as in our examples that
c(u) is an increasing function of u. In other words, higher values of u will move with higher
velocities. Suppose we start with an initial profile u0 (x) that resembles an increasing front,
as in Fig. 4.14. Since higher values of u travel faster, the top part of the profile will move
more ahead of the bottom part, and the overall effect is that of the spreading out of the
solution, i.e., the solution becomes less steep. As shown in the figure, this corresponds to
two characteristics where the right one has a lesser slope 1/c2 than the left one 1/c1 , leading
to the characteristics fanning out. Thus the values of u in between u1 and u2 are being
spread out over a larger x interval.
On the other hand, if we start with a downward front, we obtain the situation depicted
in Fig. 4.15. Now the higher velocities of the higher values of u lead to the steepening of the
profile. The characteristics cross, and the values of u between u1 and u2 are condensed in an
x interval that shrinks to a point at the crossing of the characteristics, leading to a profile
that is infinitely steep: the solution becomes steeper as the interval length decreases, until
it becomes vertical.
It might be tempting to guess that the solution behaves as illustrated in Fig. 4.16. This
is incorrect! The evolution from the first panel to the second one is correct. But at the
second panel, the solution has a vertical tangent, and we cannot rely on the characteristics
or anything else coming from the PDE anymore. Since the PDE is no longer valid, we cannot
86 CHAPTER 4. THE METHOD OF CHARACTERISTICS
u0 (x)
u2
c2 > c1
u1
c1
x
t
1
c1
1 1
c2
< c1
use it to move from the second to the last panel. Of course, the solution u(x, t) is supposed
to be a single-valued function of x and t. In the third panel, there exists an entire interval
of x values for which the solution is tripple valued. Woe!
Before we figure out what happens after a shock (a profile with vertical tangent) forms,
we should first determine when a shock forms. Thus, we want to determine the so-called
breaking time tb ≥ 0.
Example. We revisit the example we plotted the characteristics and the solution profiles
for in Fig. 4.12. The problem is given by
ut + uux = 0,
2
u(x, 0) = e−x .
2 2
x = x0 + te−x0 ⇒ t = (x − x0 )ex0 .
The plot seems to indicate that tb ≈ 1.2. How can we find this value?
4.4. BREAKING AND GRADIENT CATASTROPHES 87
u0 (x)
u1 c2 > c1
c1
u2
1
c2
1
c1
u = u0 (x0 ),
∂x0
⇒ ut = u00 (x0 ) , and
∂t
∂x0
⇒ ux = u00 (x0 ) ,
∂x
where we have used the chain rule, since x0 depends implicitly on x and t. Assuming that
u0 (x0 ) is a nice profile (i.e., no vertical tangents), we see that we need to determine when
∂x0 /∂x and/or ∂x0 /∂t are infinite. We have
x x
∂x0 1
⇒ = .
∂x 1 + tc (u0 (x0 ))u00 (x0 )
0
and
2
u0 (x) = e−x .
The expression for the breaking time becomes
−1
tb = min 2
x0 −2x0 e−x0
1
= min 2.
x0 2x0 e−x0
We have
2 2
F 0 = 2e−x − 4x2 e−x
2
= 2e−x (1 − 2x2 ),
Once the derivative becomes infinite, we should expect the solution to develop a discontinuity,
known as a shock. In other words, u(x, t) will have different values u− (before) and u+ (after)
the shock, where the derivative is vertical. At either side of the shock, the solution satisfies
the PDE, but the PDE cannot capture the shock itself.
Suppose we have a wedge-like region of the (x, t) plane where the characteristics cross,
as in Fig. 4.17. The idea is to insert a path in the (x, t) plane along which the shock will
propagate. Up until the shock path, our previous solution is valid, and we continue to follow
the characteristics, as before, until they hit the shock. Such a shock path x = xs (t) is
inserted in red in Fig. 4.17. Once we have inserted the shock path, we can merrily continue
the characteristics until the hit the shock, even into the wedge region!
So, how do we find how the shock moves? The integral form is given by
d b
Z
u(x, t)dx = φ(a, t) − φ(b, t).
dt a
Suppose a shock exists at x = xs (t), in between x = a and x = b. Then
Z xs Z b
d
u(x, t)dx + u(x, t)dx = φ(a, t) − φ(b, t).
dt a xs
Now we use the Leibniz rule7 , which is really just a big ol’ chainrule. Since xs depends on t,
we get
Z xs Z b
− dxs dxs
ut (x, t)dx + u(xs , t) + ut (x, t)dx − u(x+
s , t) = φ(a, t) − φ(b, t),
a dt xs dt
R b(t) R b(t)
7
Recall, d
dt a(t)
f (x, t)dx = a(t)
ft (x, t)dx + f (b(t), t)b0 (t) − f (a(t), t)a0 (t)
4.5. SHOCK WAVES 91
t
x = xs (t)
tb
x0b xb
Figure 4.17: The wedge region where characteristics cross, with a shock curve inserted.
where
u− := u(x−
s , t) = lim− u(x, t),
x→xs
+
u := u(x+
s , t) = lim+ u(x, t),
x→xs
are the limits as x approaches xs from the left and right, respectively.
Next, since a and b are completely arbitrary in this process, we now let a → x− s and
+
b → xs . This eliminates the integrals above, since they are integrals of bounded functions
over an interval that shrinks to zero. We are left with
u− x0s − u+ x0s = φ− − φ+
dxs φ− − φ+ ∆φ
⇒ = − +
:= .
dt u −u ∆u
This is known as the Rankine-Hugoniot condition. It dictates the speed at which the
shock moves. Note that for the PDE
ut + uux = 0,
and the shock speed is simply the average value of the solution to the left and right of it.
Let’s see how the Rankine-Hugoniot condition works.
Example. Consider the problem
ut + uux = 0,
1, x ≤ 0,
u(x, 0) = u0 (x) =
0, x > 0.
The implicit solution is given by
u(x, t) = u0 (x0 ),
x = x0 + tu0 (x0 ).
u=1 u=0
Example. We consider a second example, from traffic flow. Recall that we the traffic
flow model is governed by
ut + φx = 0,
with
u
φ = v1 u 1 − .
u1
Let’s assume that the speed limit v1 = 45 miles/hour. Similarly, we work with a maximal
density of cars of u1 = 300 cars/mile. Thus
u2
φ = 45 u − .
300
u0 (x)
300 (stopped)
x
0
Figure 4.20: The initial condition for the traffic jam problem.
Figure 4.21: The characteristics for the traffic jam problem, without the insertion of a shock
path.
Once again, there is a (very large) triangular region where characteristics cross. We use
the Rankine-Hugoniot condition, because it’s what the cool kids do. We have
dxs ∆φ
= .
dt ∆u
On the left-side of the shock, we will have u− = 100, thus φ− = 3000. On the right side of
the shock, u+ = 300, so that φ+ = 0. Thus
3000
x0s = = −15 ⇒ xs = −15t,
−200
where we have used that the shock starts at xs = 0 at t = 0. The characteristics with this
shock line inserted are shown in Fig. 4.22.
Finally, our solution is given by
300, x > −15t,
u(x, t) =
100, x < −15t.
u = 300
u = 100
Figure 4.22: The characteristics for the traffic jam problem, with the insertion of a shock
path.
In real applications, a shock never forms. Instead, we get a solution with a very steep,
but not vertical profile. The presence of extra physical effects precludes the formation of
actual shocks.
As an example, we consider once again traffic flow. How shall we update our model to
take into account extra effects which presumably will prevent shock formation?
We have
ut + φx = 0.
So far, we have used φ = uv, with
u
v = v1 1 −
u1
as our velocity profile. We have assumed that drivers adjuct their speed based on the car
density they observe where they are. Let’s give drivers a bit more credit8 . Let’s assume that
drivers can adjust their speed based on what they observe ahead of them.
For instance, suppose that drivers observe that the traffic is becoming more dense. In
other words, ux > 0. It would seem natural that they would decrease their speed. Further,
if u is very small, any chance appears huge and might have a big effect. Similarly, if u
is large, and change is not very important, this it is the relative change that matters: we
wish to modify our equation for v as a function of u with a term −rux /u. Here ux /u is the
relative density change, and r is a positive proportionality constant. The − sign ensures that
the velocity increases as the density increases. Notice that the opposite happens if we see
8
You can let me know in 30 years whether we should or not.
4.6. SHOCK WAVES AND THE VISCOSITY METHOD 97
the density is lighter ahead: the drivers will start to speed up. Thus, our new velocity law
becomes
u ux
v = v1 1 − −r .
u1 u
It follows that
u2
φ = uv = v1 u−− rux
u1
2u
⇒ φx v1 1 − ux − ruxx .
u1
lim u(x, t) = u1 ,
x→∞
lim u(x, t) = u0 < u1 .
x→−∞
In other words, we have gridlock on the right, and some lower-density moving traffic moving
into the gridlock. In addition, we have
lim ux (x, t) = 0,
x→∞
lim u(x, t) = 0.
x→−∞
u(x, t) = f (x − ct),
we get
0 2f
−cf + v1 1 − f 0 = rf 00
u1
v1
⇒ −cf + v1 f − f 2 = rf 0 + k.
u1
At this point, we can find c and k, using the boundary conditions. Evaluating the above at
+∞, we get
−cu1 + v1 u1 − v1 u1 − r · 0 + k ⇒ k = −cu1 .
98 CHAPTER 4. THE METHOD OF CHARACTERISTICS
u(x, t)
u = u1 = 300
c = −v1 u0 /u1
x
u = u0 = 100
Figure 4.23: The traveling-wave solution for the traffic flow problem with smart drivers.
Here u1 = 300, u0 = 100, v1 = 45. Different values of r are used, ranging from 20 to 5.
v1 u20 u0
−cu0 + v1 u0 − = r · 0 + k ⇒ c = −v1 .
u1 u1
This solution is drawn in Fig. 4.23, for different values of r. Note that as r → 0,
u1 , x − ct > 0,
u(x, t) =
u0 , x − ct < 0.
Thus, the solution becomes steeper and steeper as r → 0, ultimately resulting in the shock
solution as r → 0.
Unfortunately, we cannot devote more time to the viscosity method at this point. This
example gives a flavor, but it also indicates that the way to introduce extra effects which
might arrest shock formation is very problem dependent.
u(x, 0) = u0 (x)
x
0
0
Figure 4.24: The initial condition u0 (x) for the rarefaction example.
ut + uux = 0,
0, x < 0,
u(x, 0) = u0 (x) =
1, x > 0.
u(x, 0) = u0 (x)
− x
0
x = x0 + u0 (x0 )t,
x0 < 0 : x = x0 ,
x0 > 0 : x = x0 + t.
These characteristics are shown in Fig. 4.25. We face the immediate question what the value
of the solution is in the red region without characteristics.
To answer this question, we revisit the same problem, but now with an initial condition
that is smooth as opposed to discontinuous. Such an initial condition is drawn in Fig. 4.26.
It is zero outside of |x| > , but it changes smoothly from zero to one in |x| < . The new
initial condition is constructed so that it limits to the discontinuous one, u0 (x), as → 0.
This gives rise to a new characteristics picture, shown in Fig. 4.27. The characteristics
are given by
x − x0
t= ,
u0 (x0 )
and it is clear that for x0 between − and , the characteristics will start at x0 and have a
slope that varies smoothly from 0 for x0 = − to 1 for x0 = .
In the limit as → 0, we get a fan of characteristics, all starting at (0, 0), but with slope
ranging from ∞ (on the left) to 1 on the right. This seems like a reasonable way to fill in the
empty region: we assume an initial profile with high steepness that is smooth, and we let
the steepness → ∞. The limiting characteristic plane is shown in Fig. 4.28. We call the red
region in Fig. 4.25 the rarefaction region and the characteristics that fill it the rarefaction
fan.
4.7. RAREFACTION WAVES 101
− x
Now that we have determined the rarefaction characteristics, what is the solution u(x, t)
there? We know that u(x, t) should be constant along characteristics. Since these are of the
form
x = ωt,
it follows that
u(x, t) = u(ωt, t).
Since u(x, t) is supposed to be constant along these characteristics, it follows that this last
quantity has to be independent of t. Thus
u(x, t) = g(ω) = g(x/t).
Now we have to find the function g(ω). Using the PDE ut + uux = 0, we get
∂ω x
ut = g 0 (ω) = − 2 g 0 (ω),
∂t t
∂ω 1
ux = g 0 (ω) = g 0 (ω).
∂x t
Thus
x 1
− 2 g 0 (ω) + g(ω) g 0 (ω) = 0
t t
x 0
⇒ − g (ω) + g(ω)g 0 (ω) = 0
t
⇒ −ωg 0 (ω) + g(ω)g 0 (ω) = 0
⇒ g 0 (ω)(g(ω) − ω) = 0.
102 CHAPTER 4. THE METHOD OF CHARACTERISTICS
1. g 0 (ω) = 0 ⇒ g(ω) =constant. This would mean that u(x, t) is still discontinuous,
leading to a shock. But the characteristics do not cross for t > 0. Thus this doesn’t
work. We may ignore this possibility.
2. g(x) = ω, so that
x
u(x, t) = ,
t
0, x < 0,
u(x, t) = x/t, 0 < x < t,
1, x > t.
A few time slices of the solution are shown in Fig. 4.29. The time slices make sense: due
to the fanning out of the characteristics in the rarefaction region, the values of u get spread
out over a larger region, as t increases.
4.8. RAREFACTION AND SHOCK WAVES COMBINED 103
u(x, 0)
x
0
u(x, 1)
x
0
u(x, 3)
x
0
ut + uux = 0,
1, x ∈ (0, 1),
u(x, 0) = u0 (x) =
0, x 6∈ (0, 1).
The initial condition u0 (x) for this problem is shown in Fig. 4.30. The characteristics are
given by
x = x0 + tu0 (x0 ),
or,
x0 ∈ (0, 1) : x = x0 ,
x0 6∈ (0, 1) : x = x0 + t.
104 CHAPTER 4. THE METHOD OF CHARACTERISTICS
u0 (x)
0 0 x
0
2. In region E, we use the Rankine-Hugoniot condition to insert a shock path. The shock
will receive the value u− = 1 from the left (the characteristics from region B) and the
value u+ = 0 from the right (the characteristics from region C). Thus
1 1 1
x0s = (u− + u+ ) = ⇒ xs = t + 1,
2 2 2
where the integration constant has been chosen to ensure that the shock line starts at
x0 = 1.
The characteristic plane with shock and fan inserted is shown in Fig. 4.32. We have
resolved all problems at t = 0 and we can move the initial condition forward in time. The
4.8. RAREFACTION AND SHOCK WAVES COMBINED 105
D B C
E
x
solution is perfectly well defined in regions A, B, c, D, and we have a beautiful shock moving
along shock path E. But, we’re not quite done yet! At (x, t) = (2, 2), the shock line xs
(in red) hits the first characteristic (in bold black) of the rarefaction fan in region B. This
implies that the left and right values we used in the Rankine-Hugoniot condition will have
to be altered. Past t = 2, the value of u+ (coming in from the right) remains at 0, following
the characteristics in region C. However, the value u− will no longer be 1, as now the values
come from the rarefaction fan. Thus
xs
u− = ,
t
where we have imposed that x = xs on the shock line. The problem determining our new
shock path is
dxs 1 xs
= (u− + u+ ) = ,
dt 2 2t
xs (2) = 2,
where the last equation simply state that the new shock path is a continuation of the old
one. We get
dxs 1 dt
=
xs 2 t
1
⇒ ln xs = ln t + c
2√
⇒ ln xs = ln Ct,
106 CHAPTER 4. THE METHOD OF CHARACTERISTICS
A
E
D
B C
x
Figure 4.32: The new and improved characteristics, now with shock and fan.
where c and C are constants, related by c = ln(C)/2. Applying the initial condition, we get
√
ln 2 = ln 2C ⇒ C = 2,
3. If x > t/2 + 1, t ∈ (0, 2), then u = 0. This corresponds to the part of region C below
the red straight-line shock path in Fig. 4.33.
4.8. RAREFACTION AND SHOCK WAVES COMBINED 107
A E
D
B C x
Figure 4.33: The newer and more improved characteristics, now with complete shock and
fan.
4. If x ∈ (0, t), t ∈ (0, 2), then u = x/t. This corresponds to the lower part of region D
in Fig. 4.33, to the left of the black line.
√
5. Similarly, if x ∈ (0, 2t), t > 2, then u = x/t, corresponding to the upper part of
region D in Fig. 4.33, to the left of the blue path.
√
6. Lastly, if x > 2t and t > 2, then u = 2, corresponding to the part of region C, below
the blue path.
ut + uux = 0
Z ∞ Z ∞
d
⇒ udx + uux dx = 0
dt −∞ −∞
Z ∞ ∞
d 1
⇒ udx + u2 =0
dt −∞ 2 −∞
d ∞
Z
⇒ udx = 0
dt −∞
Z ∞
⇒ u(x, t)dx = 1.
−∞
The last line comes from the fact that the area underneath the solution is 1 initially, while
the line before it says that this area is conserved. We have also used that limx→±∞ u = 0. We
easily verify this: at t = 1, the area is the area of a trapezoid. It is equal to 1×(1/2+3/2)/2 =
1. Next, at t = 2, we need the √ p a triangle, equal to 2 × 1/2 = 1. Lastly, at t = 3, we
area of
still have a triangle, with area 6 × 2/3/2 = 1. It also follows from this argument that as
108 CHAPTER 4. THE METHOD OF CHARACTERISTICS
u=1
x0s = 1/2
t=0
u=0 u=0
x=0 x=1
u=1
we march on, the top of the triangle is lowered, while its base grows so that its area remains
unchanged.
A warning
Let’s reconsider the example on page 92. We have
ut + uux = 0,
1, x < 0,
u(x, 0) =
0, x > 0.
which we used to determine a perfectly fine shock. This condition is based on the conservation
of u.
However, consider the same problem, but written as
uut + u2 ux = 0,
1, x < 0,
u(x, 0) =
0, x > 0.
Now, the PDE is written
ρt + Φx = 0,
with ρ = u2 /2 and Φ = u3 /3. The Rankine-Hugoniot condition becomes
1 − 3 3
∆Φ u − 1 u+ 2
x0s = = 31 − 2 31 + 2 = ,
∆ρ 2
u − 2u 3
and we find a different, but still equally fine shock. Thus, the same PDE can give rise to
two (or many more) conclusions about the shock velocity. How do we choose? Which one is
correct?
It turns out the answer is we don’t choose. We are lowly mathematicians, who solve
problems given to us by our experimentalist friends10 . The application we’re dealing with
dictates which quantity is conserved, which implies how we should write the PDE. One
we know this, we can propagate the shock. But simply knowing the PDE is not enough
information to work with the shock, as the PDE breaks down once we have a shock.
way to characterize weak solutions of the PDE, through the so-called weak formulation of
the PDE.
The denominator in the exponent is always positive and approaches zero. Thus −1/(1 −
x2 ) → −∞ and
lim T (x) = 0.
x<1
The same argument holds for all derivatives: at all orders, we get the same exponential12
and some rational function of x. Thus
lim T (n) (x) = 0, for n ≥ 0.
x<1
11
No, this is not 111 . Rather, it’s the obligatory reminder that you should check this.
12
Exponentials do that, you know.
4.9. WEAK SOLUTION OF PDES 111
Remark. As a side note, this implies that the Taylor series of T (x) around x = 1 is
given by
∞
X T (n) (1)
T (x) = (x − 1)2 ≡ 0.
n=0
n!
Quite a weird result!
We call T (x) a test function or, sometimes, a Schwarz function. It is a function that is
infinitely differentiable but has compact support. This means that it is nonzero only in
a finite region. By shifting x and rescaling, we can control where the support of the test
function is located and how large it is.
Similarly, consider
−(x2 +t2 )/(1−(x2 +t2 ))
e , x2 + t2 < 1,
T (x, t) =
0, x2 + y 2 ≥ 0.
This is a test function in the (x, t) plane. Basically, it looks like a hat with compact support.
Suppose we wish to solve
ut + φx = 0,
u(x, 0) = u0 (x),
for x ∈ R, t¿0. Let T (x, t) be a test function. Then T (x, t)u(x, t) is zero at every point
outside the circle x2 + y 2 = 1, so we have isolated a small part of u(x, t). Then
Here the last line is just a less confusing version of the previous line.
Now we integrate by parts13 .
Z ∞ Z ∞
t=∞
dx u(x, t)T (x, t)|t=0 − dtu(x, t)Tt (x, t) +
−∞ 0
Z ∞ Z ∞
x=∞
dt φ(x, t)T (x, t)|x=−∞ − dxφ(x, t)Tx (x, t) = 0
0 −∞
Z ∞ Z ∞
⇒ dx 0 − u0 (x)T (x, 0) − u(x, t)Tt (x, t) +
−∞ 0
13
Secret: integration by parts is the main weapon of a good applied mathematician.
112 CHAPTER 4. THE METHOD OF CHARACTERISTICS
Z ∞ Z ∞
dt 0 − dxφ(x, t)Tx (x, t) = 0
0 −∞
Z ∞ Z ∞ Z ∞
⇒ − u0 (x)T (x, 0) − dt dx (u(x, t)Tt (x, t) + φ(x, t)Tx (x, t)) = 0.
−∞ 0 −∞
This is known as the weak form of the PDE. It involves derivatives only of the test function,
and just the value of u and φ, which are defined everywhere, even for weak solutions. We
should remark that the weak form of the equation gives us back the PDE, provided that
the solutions are classical solutions. In that case, we can start from the weak form, use
integration by parts and get the PDE.
There are many reasons for wanting to work with the weak form of an equation. Here
are two.
1. Larger function spaces. The modern theory of PDEs likes to talk about solutions
is specific function spaces. Examples are C(R), the space of continuous functions,
defined for all x ∈ R, C 1 (R), the space of differentiable functions on R, etc. Clearly,
C 1 (R) ⊂ C(R), and so on. In general, we like to solve PDEs in the largest function
space possible: a larger space implies fewer conditions on the solution, so the solution
is more general. Since the weak form of the PDE has fewer conditions on the solution
u(x, t), by only needing its value, and not its derivatives, it allows us to work in larger
function spaces.
Chapter 5.
Dissipation, Dispersion, and Group Velocity
! !
i! !
was described in Chapter 2, and examples were given in x3.1. For a PDE of
rst order in t, the result is
Z 1
u(xt) = 2 1 i(x+! ( )t)
;1 e u^0( ) d: (5:1:7)
Since most partial di erential equations of practical importance have variable
coecients, nonlinearity, or boundary conditions, it is rare that this integral
representation is exactly applicable, but it may still provide insight into local
behavior.
Discrete approximations to di erential equations also admit plane wave
solutions (5.1.1), at least if the grid is uniform, and so they too have dispersion
relations. To begin with, let us discretize in x only so as to obtain a semidis-
crete formula. Here are the dispersion relations for the standard centered
semidiscretizations of (5.1.3){(5.1.6):
ut = 0u : ! = h1 sin h (5:1:8)
utt = u : !2 = h42 sin2 h 2 (5:1:9)
ut = u : i! = ; h42 sin2 h 2 (5:1:10)
ut = i u : ! = ; h42 sin2 h 2
: (5:1:11)
These formulas are obtained by substituting (5.1.1) into the nite di erence
formulas with x = xj . In keeping with the results of x2.2, each dispersion
relation is 2=h-periodic in , and it is natural to take 2 ;=h=h] as
a fundamental domain. The dispersion relations are plotted in Figure 5.1.2,
superimposed upon dotted curves from Figure 5.1.1 for comparison.
Stop for a moment to compare the continuous and semidiscrete curves in
Figure 5.1.2. In each case the semidiscrete dispersion relation is an accurate
approximation when is small, which corresponds to many grid points per
wavelength. (The number of points per spatial wavelength for the wave (5.1.1)
is 2=h.) In general, the dispersion relation for a partial di erential equation
is a polynomial relation between and !, while a discrete model amounts to
a trigonometric approximation. Although other design principles are possible,
the standard discrete approximations are chosen so that the trigonometric
function matches the polynomial to as high a degree as possible at the origin
= ! = 0. To illustrate this idea, Figure 5.1.3 plots dispersion relations for the
standard semidiscrete nite di erence approximations to ut = ux and ut = iuxx
of orders 2, 4, and 6. The formulas were given in x3.3.
5.1. DISPERSION RELATIONS TREFETHEN 1994 196
! !
i! !
(c)ut = u (d) ut = i u
6
4
2
ut = ux
(a) (b) ut = iuxx
Now let us turn to fully discrete nite di erence formulas: discrete in time
as well as space. The possibilities become numerous. For example, substituting
the plane wave
vjn = ei(xj +!tn) = ei(jh+!nk)
into the leap frog approximation of ut = ux yields the dispersion relation
ei!k ; e;i!k = (eix ; e;ix )
where = k=h, that is,
sin !k = sin h: (5:1:12)
Similarly, the Crank-Nicolson approximation of ut = uxx has the dispersion
relation
ei!k ; 1 = (e 2 +1) eih ; 2+ eih
i!k h i
which reduces to
i tan !k
2 = ;2 sin2 :
h
2 (5:1:13)
5.1. DISPERSION RELATIONS TREFETHEN 1994 198
Wiley, 1976).
5.2. DISSIPATION TREFETHEN 1994 199
Figure 5.1.4. Dispersion relation for the leap frog model of utt =
uxx + uyy in the limit ! 0. The region shown is the domain ;=h
=h]2 of the ( ) plane. The concentric curves are lines of constant
! for !h = 41 21 ::: 114 .
5.2. Dissipation
Even though a partial dierential equation may conserve energy in the L2 norm, its
nite dierence models will often lose energy as t increases, especially in the wave numbers
comparable to the grid size. This property is numerical dissipation, and it is often advan-
tageous, since it tends to combat instability and unwanted oscillations. In fact, articial
dissipation* is often added to otherwise nondissipative formulas to achieve those ends. An
example of this kind appeared in Exercise 3.2.1.
To make the matter quantitative, suppose we have a linear partial dierential equation
or nite dierence approximation that admits waves (5.1.1) with ! given by a dispersion
relation (5.1.2). Since is assumed to be real, it follows that the wave has absolute value
jei(x+!t) j = e;t Im ! (5:2:1)
*In computational uid dynamics one encounters the more dramatic term articial viscosity.
5.2. DISSIPATION TREFETHEN 1994 201
As an extreme case the heat equation ut = uxx, with dispersion relation ! = i 2 , dissipates
nonzero wave numbers strongly|and indeed it does nothing else at time t, only wave
p
numbers = O( t ) remain with close to their initial amplitudes. But it is principally the
dissipation introduced by nite dierence formulas that we are concerned with here.
The following denitions are standard:
A nite dierence formula is nondissipative if Im ! = 0 for all . It is dissipative if
Im ! > 0 for all = 0. It is dissipative of order 2r if ! satises
6
Im !k 2r ei !k j 1 ; 2 (h)2r
1 (h) i.e., j (5:2:3)
for some constants j > 0. In each of these statements, varies over the interval
=h=h], and ! represents all possible values ! corresponding to a given . For
;
For example, the leap frog and Crank-Nicolson models of ut = ux are nondissipative,
while the upwind and Lax-Wendro formulas are dissipative.
Dissipative and nondissipative are mutually exclusive, but not exhaustive: a nite
dierence formula can be neither dissipative nor nondissipative. See Exercise 5.2.1.
According to the denition, no consistent nite dierence approximation of ut = ux + u
could be dissipative, but customary usage would probably use the term dissipative sometimes
for such a problem anyway. One could modify the denition to account for this by including
a term O(k) in (5.2.3).
A more serious problem with these standard denitions arises in the case of multistep
formulas, for which each corresponds to several values of !. In such cases the denition
of dissipative, for example, should be interpreted as requiring Im ! > 0 for every value of !
that corresponds to some = 0. The diculty arises because for multistep formulas, that
6
condition ensures only that the formula dissipates oscillations in space, not in time. For
example, the leap frog model becomes dissipative if a small term such as k vn is added
to it, according to our denitions, yet the resulting formula still admits the wave (5.1.1)
with = 0, ! = =h, which is sawtoothed in time. To exclude possibilities of that kind it is
sometimes desirable to use a stronger denition:
A nite dierence formula is totally dissipative if it is dissipative and in addition,
Im ! = 0 implies ! = 0.
EXERCISES
. 5.2.1. Determine whether each of the following models of ut = ux is nondissipative, dissi-
pative, or neither. If it is dissipative, determine the order of dissipativity.
(a) Lax-Wendro, (b) Backward Euler, (c) Fourth-order leap frog, (d) Box, (e) Upwind.
5.3. DISPERSION AND GROUP VELOCITY TREFETHEN 1994 202
parameters (1 !1 ) and (2 !2 ). It is then readily seen that the superposition consists of a
smooth envelope times a carrier wave at frequency 21 (!1 + !2) and wave number 21 (1 + 2 ),
and the envelope moves at velocity (!2 !1)=(2 1 ), which approaches (5.3.2) in the
; ; ;
and ! real, and change slightly to a complex value + i \in order to visualize which
way the envelope is moving." If the eect on ! is to make it change to ! + i !, it is readily
calculated that the resulting evanescent wave has an envelope that moves laterally at the
velocity != , and this again approaches (5.3.2) in the limit 0, ! 0. A third,
; ! !
more \PDE-style" explanation is based upon advection of local wave number according to
a simple hyperbolic equation with coecient cg see Lighthill.
Group velocity in multiple dimensions. If there are several space dimensions, the group
velocity becomes the gradient of ! with respect to the vector , i.e., cg = !.
;r
Phase and group velocities on a grid. There is another sense in which group velocity
has more physical meaning than phase velocity on a nite dierence grid: the former is
well-dened, but the latter is not. On a periodic grid, any Fourier mode can be represented
in terms of innitely many possible choices of and ! that are indistinguishable physically,
and according to (5.3.1), each choice gives a dierent phase velocity. What's going on here
is that naturally one can't tell how fast a pure complex exponential wave is moving if one
sees it at only intermittent points in space or time, for one wave crest is indistinguishable
from another. By contrast, the group velocity is well-dened, since it depends only on the
slope, which is a local property formula (5.3.2) has the same periodicity as the dispersion
relation itself.
Computation of a group velocity on a grid. To compute the group velocity for a
nite dierence formula, dierentiate the dispersion relation implicitly and then solve for
cg = d!=d . For example, the wave equation ut = ux has c = cg = 1 for all . For the leap
; ;
frog approximation the dispersion relation is sin !k = sin ! , which implies k cos !k d! =
h cos ! d , and since k = h, cg (! ) = cos h= cos !k .
;
Parasitic waves. Many nite dierence formulas admit parasitic waves as solutions, i.e.,
waves that are sawtoothed with respect to space or time. These correspond to = =h,
! = =k , or both. It is common for such waves to have group velocities opposite in sign
to what is correct physically. In the example of the leap frog formula, all four parasitic
modes 1, ( 1)j , ( 1)n , and ( 1)j+n are possible, with group velocities 1, 1, 1, and 1,
; ; ; ; ;
respectively.
Spurious wiggles near interfaces and boundaries. It is common to observe spurious
wiggles in a nite dierence calculation, and they appear most often near boundaries, in-
terfaces, or discontinuities in the solution itself. The explanation of where they appear is
usually a matter of group velocity. Typically a smooth wave has passed through the discon-
tinuity and generated a small reected wave of parasitic form, which propagates backwards
into the domain because its group velocity has the wrong sign. More on this in the next
chapter.
Waves in crystals. Dispersion relations for vibrations in crystals are also periodic with
respect to . As a result, sound waves in crystals exhibit dispersive eects much like those
associated with nite dierence formulas, including the existence of positive and negative
group velocities.
5.3. DISPERSION AND GROUP VELOCITY TREFETHEN 1994 204
Figure 5.3.1. Dispersion under the leap frog model of ut = ux with = 0:5. The
lower mesh is twice as ne as the upper.
5.3. DISPERSION AND GROUP VELOCITY TREFETHEN 1994 205
EXERCISES
. 5.3.1. The Box formula.
(a) Write out the BOXx formula of Table 3.2.1 in terms of vjn , vjn+1 , etc.
(b) Determine the dispersion relation (expressed in as simple a form as possible).
(c) Sketch the dispersion relation.
(d) Determine the group velocity as a function of and !.
. 5.3.2. Schrodinger equation.
(a) Calculate and plot the dispersion relation for the Crank-Nicolson model of ut = iuxx of
Exercise 3.2.1(f).
(b) Calculate the group velocity. How does it compare to the group velocity for the equation
ut = iuxx itself?
. 5.3.3. A paradox. Find the resolution of the following apparent paradox, and be precise in
stating where the mistake is. Draw a sketch of an appropriate dispersion relation to explain
your answer.
One the one hand, if we solve ut = ux by the leap frog formula with = 1, the results will
be exact, and in particular, no dispersion will take place.
On the other hand, as discussed above, the dispersion relation on any discrete grid must be
periodic, hence nonlinear|and so dispersion must take place after all.
Chapter 2
The main reference for this chapter is §1.1 of the book [Drazin and Johnson, 1989].
2.1 Dispersion
Usually, localised waves spread out (“disperse”) as they travel. This prevents them from
being solitons. Let’s understand this phenomenon first.
EXAMPLES:
1
ut ` ux “ 0 (2.1)
v
ÝÑ Solution
upx, tq “ f px ´ v tq for any function f ,
i.e. a wave moving with velocity v (right-moving if v ą 0, left-moving if v ă 0). The
wave keeps a fixed profile f pξ q and moves rigidly at velocity v (indeed ξ “ x ´ v t):
12
CHAPTER 2. WAVES, DISPERSION AND DISSIPATION 13
1
utt ´ uxx “ 0 pv ą 0 wlogq (2.2)
v2
ÝÑ Solution
All waves move at the same speed, so there is no dispersion, but there is no interaction
either, so this is also not very interesting for our purposes.
1
utt ´ uxx ` m2 u “ 0 , (2.3)
v2
where we take v ą 0 wlog.
This is a more interesting equation. Let us try a complex “plane wave” solution2
Substituting the plane wave (2.4) in the Klein-Gordon equation (2.3), we find:
ω 2 ipkx´ωtq
´ e ` k 2 eipkx´ωtq ` m2 eipkx´ωtq “ 0
v2
ω2
ùñ ´ 2 ` k 2 ` m2 “ 0 .
v
1
This is the first relativistic wave equation (with v the speed of light). It was introduced independently by
Oskar Klein [Klein, 1926] and Walter Gordon [Gordon, 1926], who hoped that their equation would describe
electrons. It doesn’t, but it describes massive elementary particles without spin, like the pion or the Higgs boson.
2
This is called a “plane wave” because its three-dimensional analogue up⃗x, tq “ exprip⃗k ¨ ⃗x ´ ω tqs has constant
u along a plane ⃗k ¨ ⃗x “ const at fixed t. Unless specified, in this course we are interested in real fields u. It is
nevertheless convenient to use complex plane waves (2.4) and eventually take the real or imaginary part to find
a real solution, rather than working with the real plane waves cospk x ´ ω tq and sinpk x ´ ω tq from the outset.
CHAPTER 2. WAVES, DISPERSION AND DISSIPATION 14
So the plane wave (2.4) is a solution of the Klein-Gordon equation (2.3) provided that ω
satisfies ?
ω “ ω pk q “ ˘ v k 2 ` m2 . (2.5)
We will usually ignore the sign ambiguity and only consider the ` sign in (2.5) and
similar equations.3
VOCABULARY:
k wavenumber λ“ 2π
k
wavelength (periodicity in x)
ω angular frequency τ“ 2π
ω
period (periodicity in t)
A formula like (2.5) relating ω to k : dispersion relation.
The maxima of a real plane wave, like for instance Re eipkx´ωpkqtq or Im eipkx´ωpkqtq , are
called “wave crests”. By a slight abuse of terminology, we will refer to the wave crests
of the real or imaginary part of a complex plane wave like (2.4) simply as the wave crests
of the complex plane wave.
By rewriting the complex plane wave solution (2.4) of the Klein-Gordon equation as
eikpx´cpkqtq , we see that its wave crests move at the velocity
c
ω pk q m2
cpk q “ “ v 1 ` 2 signpk q .
k k
it will disperse.
- PHASE VELOCITY
ω pk q
cpk q “ , (2.7)
k
which is the velocity of wave crests.
3
We do not lose generality here, since we can obtain the plane wave solution with opposite ω by taking the
complex conjugate plane wave solution and sending k Ñ ´k .
CHAPTER 2. WAVES, DISPERSION AND DISSIPATION 15
- GROUP VELOCITY
dω pk q
cg pk q “ , (2.8)
dk
which is the velocity of the lump of field while it disperses.
We will understand better the relevance of the group velocity in the next section.
REMARK:
The energy (and information) carried by a wave travels at the group velocity, not at the
phase velocity. For a relativistic wave equation with speed of light v , no signals can be
transmitted faster than the speed of light. So it should be the case that |cg pk q| ď v for all
wavenumbers k , but there is no analogous bound on the phase velocity. For example, for the
Klein-Gordon equation (2.3), we can calculate
- |Group velocity|: ˇ ˇ
ˇ dω pk q ˇ
|cg pk q| “ ˇ
ˇ ˇ“ b v ďv
dk ˇ 1` m2
k2
- |Phase velocity|: c
ˇ ˇ 2
ˇ ω pk q ˇ
|cpk q| “ ˇˇ ˇ“v 1` m ěv,
k ˇ k2
which is faster than the speed of light v for all k , but this is not a problem.
average wavenumber k̄
spread of wavenumber „ 1{a ,
?
provided that ω pk q “ v k 2 ` m2 .4
Since most of the integral (2.9) comes from the region k « k̄ , we can obtain a good approxi-
mation to (2.9) by Taylor expanding ω pk q about k “ k̄ . Expanding to first order in pk ´ k̄ q we
obtain
ω pk q “ ω pk̄ q ` ω 1 pk̄ q ¨ pk ´ k̄ q ` Oppk ´ k̄ q2 q
“ ω pk̄ q ` cg pk̄ q ¨ pk ´ k̄ q ` Oppk ´ k̄ q2 q
« ω pk̄ q ` cg pk̄ q ¨ pk ´ k̄ q ,
where in the second line we used (2.8) and in the third line we introduced a short-hand « to
4
z px, tq is a complex solution of the Klein-Gordon equation. Since the Klein-Gordon equation is a linear
equation with real coefficients, the complex conjugate z px, tq˚ is also a solution of the Klein-Gordon equation,
as are Re z px, tq and Im z px, tq.
CHAPTER 2. WAVES, DISPERSION AND DISSIPATION 17
where in the second line we factored out a plane wave with k “ k̄ , in the third line we
changed integration variable replacing k by k ` k̄ , in the fourth line we completed the square
2
Ak 2 ` B k “ Apk ` 2BA q2 ´ B
4A
, and in the last line we used the Gaussian integral formula
ż `8`ic c
2 π
e´Ak “ ,
´8`ic A
which holds for all A ą 0 and c P R. The final result is the product of a:
1. “CARRIER WAVE”:
a plane wave moving at the phase velocity
ω pk̄ q
cpk̄ q “
k̄
2. “ENVELOPE”:
a localised profile (or “wave packet”) mov-
ing at the group velocity
cg pk̄ q “ ω 1 pk̄ q .
Click here to see an animation of a Gaussian wavepacket with a (Gaussian) envelope and a
carrier wave moving at different velocities. In the animation the phase velocity is much larger
than the group velocity.
To this order of approximation, the spatial width of the lump has the parametric dependence
CHAPTER 2. WAVES, DISPERSION AND DISSIPATION 18
WIDTH „ a,
meaning that the width doubles if a is doubled, and is constant in time. (Indeed, a simultaneous
rescaling of x ´ cg pk̄ qt and a by the same constant λ leaves the envelope invariant.)
and that the amplitude of the wave packet also decreases as time increases.
This leads to the phenomenon of DISPERSION, whereby the profile of the wave packet
changes as it propagates. In particular, starting from a localised wave packet, dispersion makes
the wave packet spread out: the width of the initial wave packet grows and the amplitude de-
creases as time increases. See this animation of the time evolution of the Gaussian wave-packet
up to second order in pk ´ k̄ q.
2.3 Dissipation
So far we have considered wave equations which lead to a real dispersion relation, so ω pk q P R.
If instead ω pk q P C, then a new phenomenon occurs: DISSIPATION, where the amplitude
of the wave decays (or grows) exponentially in time. For a plane wave
upx, tq “ eipkx´ωpkqtq “ eipkx´ Re ωpkq¨tqq e Im ωpkq¨t (2.10)
and we have two cases:
• Im ω pk q ă 0: “PHYSICAL DISSIPATION”
The amplitude decays exponentially with time.
• Im ω pk q ą 0: “UNPHYSICAL DISSIPATION”
The amplitude grows exponentially with time (physically unacceptable).
EXAMPLES:
1.
1
ut ` ux ` αu “ 0 pα ą 0, v ą 0q (2.11)
v
Sub in a plane wave u “ eipkx´ωtq :
ω
´i ` ik ` α “ 0 ùñ ω pk q “ v pk ´ iαq ,
v
CHAPTER 2. WAVES, DISPERSION AND DISSIPATION 19
2. HEAT EQUATION:
ut ´ αuxx “ 0 pα ą 0q (2.12)
˚ EXERCISE: Sub in a plane wave and derive the dispersion relation ωpkq “ ´iαk . 2
2.4 Summary
• Linear wave equation ÝÑ (Complex) plane wave solutions u “ eipkx´ωtq .
Sub in to get ω “ ω pk q dispersion relation.
• Dispersion (real ω , width increases and amplitude decreases) and dissipation (complex
ω , amplitude decreases exponentially) smooth out and destroy localised lumps of energy
in linear wave (or field) equations.
• Non-linearity can have an opposite effect (steepening and breaking, see chapter 1).
• For solitons the competing effects counterbalance one another precisely, leading to
stable lumps of energy, unlike for ordinary waves.