A Course On Partial Differential Equations - Walter Craig
A Course On Partial Differential Equations - Walter Craig
I N M AT H E M AT I C S 197
A Course on
Partial Differential
Equations
Walter Craig
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
10.1090/gsm/197
GRADUATE STUDIES
I N M AT H E M AT I C S 197
A Course on
Partial Differential
Equations
Walter Craig
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
EDITORIAL COMMITTEE
Daniel S. Freed (Chair)
Bjorn Poonen
Gigliola Staffilani
Jeff A. Viaclovsky
Copying and reprinting. Individual readers of this publication, and nonprofit libraries acting
for them, are permitted to make fair use of the material, such as to copy select pages for use
in teaching or research. Permission is granted to quote brief passages from this publication in
reviews, provided the customary acknowledgment of the source is given.
Republication, systematic copying, or multiple reproduction of any material in this publication
is permitted only under license from the American Mathematical Society. Requests for permission
to reuse portions of AMS publication content are handled by the Copyright Clearance Center. For
more information, please visit www.ams.org/publications/pubpermissions.
Send requests for translation rights and licensed reprints to [email protected].
c 2018 by the American Mathematical Society. All rights reserved.
Printed in the United States of America.
∞ The paper used in this book is acid-free and falls within the guidelines
established to ensure permanence and durability.
Visit the AMS home page at https://www.ams.org/
10 9 8 7 6 5 4 3 2 1 23 22 21 20 19 18
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Contents
Preface vii
Chapter 1. Introduction 1
§1.1. Overview of the subject 1
§1.2. Examples 2
Exercises: Chapter 1 8
iii
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
iv Contents
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Contents v
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Preface
vii
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
viii Preface
many editorial choices that one must make. The area of PDEs is not one
unified topic; rather it is a union of numerous directions of research with
different sources of problems, motivations, and goals. My own point of view
in this area has been strongly influenced by my teachers and colleagues,
including Louis Nirenberg, Jürgen Moser, Henry McKean, Joe Keller, and
Constantine Dafermos, as well as my many friends and collaborators, in-
cluding Catherine Sulem, Claude Bardos, Jean-Claude Saut, Eugene Wayne,
Sergei Kuksin, and Håkan Eliasson. Members of my McMaster PDE class
in 2015–2016 helped with the preparation of the text and the figures for
this publication; for their input I would like to thank in particular Alexandr
Chernyavsky, Nikolay Hristov, Kyle MacDonald, and Adam Sliwiak. I also
want to extend a special thanks to Sarnia (Nia) Sulaiman for her rendering
of numerous figures, to Ramsha Khan for her work on the images, to Emma
Holmes for her proofreading, and especially to Deirdre Haskell who read
the text thoroughly and critically and suggested many improvements to the
presentation. However, the material itself and its presentation in this text
is based on my own judgement and experiences, both scientific and peda-
gogical, and responsibility for omissions and incompleteness is entirely my
own.
The present form of the course material follows from having given courses
on PDEs many times over the years; certainly I have taught similar courses
during my time on the faculty at McMaster University. But the kernel of
the course material also comes from previous experiences with the gradu-
ate Mathematical Physics course at Stanford University, the two-semester
sequence of courses on PDEs at Brown University, and a yearlong graduate
course in PDEs at the Fields Institute in Toronto, in conjunction with the
yearlong thematic program on PDEs.
Suggestions on how to use this book. The first chapters, plus a selection
of the more in-depth material of the book, in particular Chapters 5–7 and
sections 8.1 and 8.2, make quite a complete and interesting first semester
graduate course in PDEs. If one includes the full content of Chapters 4,
6, and 8, it provides a more challenging semester-long first year graduate
level course. Each chapter concludes with a set of exercises; these are on
topics that augment the text, and to a large extent they are selected for
their independent interest. In addition to the exercises there are a number
of more challenging in-depth and open-ended projects in most chapters.
Alternatively, I have taught a pleasant semester-long undergraduate course
on PDEs at McMaster University based on the material in Chapters 2–4
and sections 7.1 and 7.2, including some of the material from Chapter 5
for analytic rigor. The course started with an introductory lecture that is
appropriate for this level of class.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Preface ix
I am hopeful that the point of view adopted in this text, which brings
phenomenology and applications along with the theory and intends to in-
troduce the class to a spectrum of techniques, will be useful to students
entering the field.
This work was partially supported by the Canada Research Chairs Pro-
gram, the Fields Institute, and by NSERC with grant number 238452-16.
Walter Craig,
Hamilton, Ontario
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
10.1090/gsm/197/01
Chapter 1
Introduction
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
2 1. Introduction
introduction that will motivate the material we are going to address in this
course, so we can at least start with a sense of its content.
The first questions are possibly “Where do partial differential equations
(PDEs) arise, and why are they useful?”. In fact, the language of the sciences
is mathematics (the joke has it that the language of the sciences is English
with an accent). Many if not most statements in the physical sciences are
in the form of mathematical equations, and the vast majority of these are
differential equations, quantifying the change of one quantity in terms of
others. Indeed, equations describing physical, chemical, and sometimes bi-
ological phenomena are for the most part PDEs, and the same statement
holds for the engineering sciences.
Disciplines of mathematics such as geometry and dynamical systems also
often give rise to PDEs. Conditions such as (i) a surface is of minimal area,
(ii) a submanifold is invariant under a flow, (iii) a mapping is conformal,
or (iv) the curvature tensor satisfies a particular property, are PDEs. For
example, the statement that “the tensor (gμν ) is an Einstein metric for a
manifold M ” is a system of PDEs.
The course material will discuss the most commonly occurring PDEs
and the implications they have for a function to be a solution. We are
particularly interested in knowing whether solutions exist, whether they are
unique, and their properties of smoothness, positivity, etc.
1.2. Examples
Here is a set of examples of some common PDEs with a brief overview of
the settings where they appear, either as a characterization of mathematical
properties of a geometrical object or as a description of a physical phenom-
enon.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
1.2. Examples 3
xn uN
x1
u1
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
4 1. Introduction
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
1.2. Examples 5
The three equations above, namely the wave equation, the heat equa-
tion, and Laplace’s equation, are the standard second-order linear constant
coefficient PDEs. Most of the elementary theory of PDEs is developed to
describe properties of their solutions. However, this list is far from being
exhaustive; many other forms of equations, both linear and nonlinear, occur
frequently and are of great interest and importance.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
6 1. Introduction
Δf = tr(∂xj ∂x f ) .
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
1.2. Examples 7
In the case that the mapping u is given as a graph of a function f (x) over
D ⊆ Rn (where N = n + 1), namely that u(x) = (x1 , . . . , xn , f (x)), then
A(f ) = 1 + (∂x f )2 dx .
D
A minimal surface is the result of a mapping for which the area functional
A(u) is minimal (or at least a critical point), generally with given additional
constraints such as boundary conditions. That is to say, the variations of A
about u all vanish; namely,
d
0 = δA(u) · v := A(u + τ v)
dτ τ =0
for all variations v(x) such that u + τ v continues to satisfy the constraints.
For the case where n = 2 and for a graph, if f is such a critical point, then
it must satisfy the PDE
x3
x1
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
8 1. Introduction
Exercises: Chapter 1
Exercise 1.1. Prove that statement (1.1) characterizing conformal map-
pings implies that either (1.2) or (1.3) holds.
Exercise 1.2. Determine the class of conformal mappings in the case that
n = N ≥ 3. The conditions (1.1) are much more rigid than in two dimen-
sions.
Exercise 1.3. Prove that the Laplacian (1.6) is invariant under Euclidean
symmetries (1.7).
Exercise 1.4. In dimension n = 2, some elementary harmonic functions
are polynomials. They include the polynomials of degrees one and two:
u(x) = xj , j = 1, 2 ,
u(x) = x21 − x22 , u(x) = x1 x2 .
How many independent harmonic polynomials are there of general degree
?
When the dimension n ≥ 3, how many independent harmonic polyno-
mials are there of degree ?
Exercise 1.5. Prove the vector calculus identity
∇ × (∇ × V ) = −ΔV + ∇(∇ · V )
for a vector field V (x), where in our notation the gradient of a function F is
∇F = (∂x1 F, . . . , ∂xn F ) .
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
10.1090/gsm/197/02
Chapter 2
Wave equations
We will start the topic of PDEs and their solutions with a discussion of a
class of wave equations, initially with several transport equations, which will
lead us to the study of the standard second-order wave equation (1.9) in one
space dimension. These transport equations are often called first-order wave
equations for reasons that will become clear as we describe their solutions.
We will use this study of wave equations as motivation to introduce the
Fourier transform as well as to give examples of the concept of a field of
characteristic curves.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
10 2. Wave equations
(2.1) ∂t u + c(t, x) · ∂x u = 0 .
This typically gives rise to an initial value problem where one asks that
u(0, x) = f (x) .
The function f (x) is called the initial data. The problem is to describe how
the solution u(t, x) evolves given its initial “shape” f (x) at t = 0. Often
these equations are known as first-order wave equations for the unknown
function u(t, x). The coefficient c(t, x) is a vector known as the “speed of
propagation”; we will see the reason for this when we describe the solution
process.
A first example is given by the case in which the coefficient c is constant.
We will see in the solution why c is known as the speed of propagation. We
will first solve this equation using the Fourier transform. For convenience
we restrict ourselves to one space dimension. The process is elementary, but
nevertheless it serves to introduce a surprisingly powerful technique for the
study of PDEs.
At a later point we will discuss the hypotheses under which this theorem
holds in much more detail; for now a perfectly reasonable condition would
be that f ∈ L2 (R), that is,
+∞ 1
2
|f (x)| dx := f L2 (R) < +∞ .
2
−∞
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
2.1. Transport equations: The Fourier transform 11
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
12 2. Wave equations
(x0+ct, t)
x0 x
not in general, nor for a time-independent speed c(x). However, when c(t) is
independent of x, the answer is “yes” and a similar solution representation
is possible. This is given as an exercise at the end of the section.
u(0, x) = f (x) .
This is the statement that the directional derivative of u(t, x) in the (1, c)
direction in the (t, x) plane vanishes:
(1, c) · (∂t , ∂x )u = 0 .
That is to say, u(t, x) is constant along lines with tangent direction (1, c).
Therefore for all lines defined by x, β such that (x − β) = ct, we have
u(t, x) = u(0, β) = f (β) = f (x − ct). Verifying via the chain rule that
this expression satisfies the transport equation (2.4) for given initial data
f ∈ C 1 (R1 ),
and the sum of the two terms cancels. The lines in (t, x) space defined by
(x − β) = ct, exhibited in Figure 2, are the characteristic curves for this
equation.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
2.2. Transport equations: The method of characteristics 13
c
(c, 1)
(x, t)
x
x = x0 (x-ct)
This observation extends to give a solution method for the general trans-
port equation, which uses ODEs and is known as the method of character-
istics. Consider the equation
(2.5) ∂t u + c(t, x)∂x u = 0 ,
u(0, x) = f (x) ,
where the speed of propagation c(t, x) is allowed to vary. We will assume
that c(t, x) ∈ C 1 . Interpret equation (2.5) in terms of directional derivatives
(1, c(t, x)) · (∂t , ∂x )u = 0 ,
which in turn is the statement that u(t, x) is constant along integral curves
of the vector field X = (1, c(t, x)) in the (t, x) plane:
(2.6) X u(t, x) = (∂t + c(t, x)∂x )u = 0 .
The characteristic equations for the transport equation (2.5) are
dt dx
(2.7) =1, = c(t, x) ,
ds ds
with initial conditions
(2.8) x(0) = β
whose solutions are the integral curves for X; these are the characteristic
curves for equation (2.5). This family of curves (t(s), x(s)) = (s, x(s, β)) is
parametrized by the points β ∈ R1 at which the curves intersect the initial
surface {t = 0}. A family of characteristic curves is illustrated in Figure 3.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
14 2. Wave equations
(c (x, t), 1)
(x (t, ), t)
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
2.3. Conservation laws 15
Exercise 2.3 asks to solve these two equations, giving a solution representa-
tion for the initial value problem u(0, x) = f (x) using explicit expressions
for the respective functions β(t, x).
Proof. Recall that the characteristic curves (t, x1 (t)) and (t, x2 (t)) satisfy
the characteristic equations (2.7) with initial conditions x1 (0) = β1 and
x2 (0) = β2 . The two curves do not intersect (as long as β1 = β2 ) because
solutions of ODEs such as the characteristic vector field are unique. Differ-
entiate the quantity in question with respect to time:
(2.12)
x2 (t) x2 (t)
dx2 dx1
∂t v(t, y) dy = ∂t v(t, y) dy + v(t, x2 (t)) − v(t, x1 (t)) .
x1 (t) x1 (t) dt dt
Now use the fact that x1 (t) and x2 (t) satisfy the characteristic equations:
dx2 dx1
v(t, x2 (t)) − v(t, x1 (t)) = v(t, x2 (t))c(t, x2 (t)) − v(t, x1 (t))c(t, x1 (t))
dt dt
x2 (t)
= ∂x c(t, y)v(t, y) dy .
x1 (t)
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
16 2. Wave equations
Returning to the original variables (t, x), the solution takes the form
(2.15) u(t, x) = V (x + t) + W (x − t) ,
which has the interpretation of a decomposition of the solution u(t, x) into
a left-moving component V and a right-moving component W . That is, any
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
2.4. The d’Alembert formula 17
Differentiating the first line with respect to x and solving for V and W , we
find
1 1 1 1
V (x) = f (x) + g(x) , W (x) = f (x) − g(x) ;
2 2 2 2
thus
1 1 x
V (x) = f (x) + g(x ) dx + c1 ,
2 2 0
1 1 x
W (x) = f (x) − g(x ) dx + c2 ,
2 2 0
where necessarily c1 + c2 = 0. This is to say that
u(t, x) = V (x + t) + W (x − t)
x+t
1 1
(2.16) = f (x + t) + f (x − t) + g(x ) dx .
2 2 x−t
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
18 2. Wave equations
(x, t)
x
x -t x +t
supp(h) := {x ∈ R1 : h(x) = 0} ,
and taking both (f, g) to have compact support, then if supp(f ), supp(g) ⊆
[−R, +R], then the support of the solution u(t, x) satisfies supp(u(t, x)) ⊆
[−R − |t|, R + |t|], as can be seen in Figure 5. At time t the solution vanishes
outside of the region {−R−|t| ≤ x ≤ R+|t|}, which is to say that information
given by the initial data in the interval [−R, +R] does not travel faster
than speed c = 1 and therefore at time t > 0 has not propagated farther
than [−(R + t), (R + t)]. This property of finite propagation speed is called
Huygens’ principle and is a common feature of solutions of wave equations.
Furthermore, it is evident from the d’Alembert formula and Figure 5 that
the solution u(t, x) is constant inside the union of light cones emanating from
the support of the initial data. Indeed, for t > R and then for |x| < t − R,
1 x+t 1 +R
u(t, x) = g(x ) dx = g(x ) dx
2 x−t 2 −R
which is a constant independent of both (t, x). This property is called the
strong Huygens’ principle; it holds for solutions of the wave equation (1.9) in
odd space dimensions, but not in even space dimensions nor for most other
wave equations.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
2.5. Duhamel’s principle 19
(x, t)
x
x-t -R R x+t
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
20 2. Wave equations
To check that this indeed solves the wave equation with the proper initial
data, one can either take the time derivative ∂t W(t)f (x) directly on the
integral expression for W(t), or check that for continuous f (x) and g(x),
the desired initial values are taken on:
To verify that this is indeed the solution we seek, check the following:
u2 (0, x) = 0 ,
t
∂t u2 (t, x) = ∂t W(t − s)h(s, ·)(x) ds = W(0)h(t, ·)(x) = 0 ,
0 t=0
t
2
∂t u2 (t, x) = ∂t2 W(t − s)h(s, ·)(x) ds + ∂t W(0)h(t, ·)(x)
0
= Δu2 (t, x) + h(t, x) .
In the case of the one-dimensional wave equation, using the explicit form
of the operator W(t), the solution formula for the inhomogeneous problem
(2.17) is similar to the d’Alembert formula itself. Indeed,
The term with the double integral accounts for the component u2 (t, x) of
the solution; it is the integral of h(t, x) over the interior of the backwards
light cone centered at the point (t, x).
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
2.6. The method of images 21
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
22 2. Wave equations
other boundary conditions, which we will not pursue here. The method
as applied to the problem of Dirichlet boundary conditions is based on the
simple fact that a function that is odd, h(−x) = −h(x), and continuous at
zero must necessarily satisfy h(0) = 0. Since the spatial domain is R1+ , the
initial data f (x), g(x) are only defined for x > 0. Assume for simplicity that
both f, g ∈ C(R1+ ). We require that both f (0), g(0) = 0 for compatibility
between the initial and boundary values of the solution. The method of
images consists of defining extensions f (x) = −f (−x) and g(x) = −g(−x)
for x < 0, the odd reflections of f (x) and g(x) through the origin; these are
defined and continuous on all of x ∈ R1 . The d’Alembert formula gives a
solution to the Cauchy problem, namely
1 1 x+t
u(t, x) = f (x + t) + f (x − t) + g(x ) dx .
2 2 x−t
This expression is clearly an odd function of x, as the extended values of f
and g are odd. For any x > t, this expression only involves the arguments
of f, g for positive x; however, for x < t, the expression involves negative
arguments even though x > 0. To remedy this we use the fact that the
extended f, g are odd to rewrite the expression as
1 1 x+t
(2.23) u(t, x) = f (t + x) − f (t − x) + g(x ) dx .
2 2 t−x
The effect is that the solution depends only upon values of the initial data
(f (x), g(x)) for x > 0. Examining this formula, it is as though the solution
is reflected off of the boundary with a negative mirror image of the incident
signal, as in Figure 6.
In the case of Neumann boundary conditions at x = 0, ask that the
boundary data f, g ∈ C 1 (R1+ ) and that it satisfy the compatibility conditions
that
∂x f (0) = 0 = ∂x g(0) .
Then reflect f, g to be even functions of x ∈ R1 , which is possible because
an even function h(x) whose first derivative is continuous at x = 0 must
satisfy ∂x h(0) = 0. Using the even property in the d’Alembert formula, for
x < t,
1 1 x+t
u(t, x) = f (x + t) + f (x − t) + g(x ) dx
2 2 x−t
t−x
1 1 t+x
= f (t + x) + f (t − x) + g(x ) dx + g(x ) dx .
2 2 t−x 0
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
2.6. The method of images 23
The wave equation is linear so that a sum of two solutions is also a solution
of the PDEs. Therefore we may take the solution to be a sum of two compo-
nents: u1 = u1 (t, x) which solves (2.13) with zero boundary data (2.21), and
a second solution u2 (t, x) which has homogeneous initial data f = g = 0 but
satisfies the requested boundary conditions (2.24). The first component u1
can be expressed as above in (2.23). The second component can be taken to
vanish for t < x as the backwards light cone does not intersect the boundary
for positive times t. However, for t > x define
u2 (t, x) = h(t − x) ,
which satisfies both the wave equation (2.13) with f (x) = 0 = g(x) and the
boundary conditions (2.21). The full solution is then u(t, x) = u1 (t, x) +
u2 (t, x). The initial and boundary data for this solution have their own
compatibility conditions to satisfy. Namely, they must have h(0) = f (0)
to avoid discontinuities in u(t, x) being introduced at u(0, 0), then also
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
24 2. Wave equations
∂t h(0) = g(0) = ∂t u(0, 0), and finally ∂t2 h(0) = ∂t2 u(0, 0) = ∂x2 u(0, 0) =
∂x2 f (0), all to avoid incompatibilities in the first and second derivatives of
the solution at (t, x) = (0, 0).
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
2.7. Separation of variables 25
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
26 2. Wave equations
Given the desired initial data f (x), g(x), the remaining task is to cal-
culate the coefficients ak of their eigenfunction expansion (2.29) for f and,
respectively, bk in the expansion (2.30) for g. It is here that the utility of
an orthonormal set of basis elements becomes clear. Using our example of
the sine series wk (x) = √1π sin( k2 x), one calculates that
2π
1 2π k
wk (x)w (x) dx = sin( x) sin( x) dx = 0 for k = ,
0 π 0 2 2
2π 2π
1 k
wk2 (x) dx = sin2 ( x) dx = 1 .
0 π 0 2
Denote the inner product between two functions w(x) and z(x) (in the
Hilbert space L2 ([0, 2π])) by
2π
w, z := w(x)z(x) dx .
0
Then the expressions above for the various integrals between the eigenfunc-
tions wk (x) are manifestations of the fact that they form an orthonormal set
in L2 ([0, 2π]), namely wk , w = 0 for k = and wk , wk := wk 2L2 = 1.
Theorem 2.5 also asserts that this set forms a complete set, or basis, for the
Hilbert space, an important result from the theory of Fourier series (again
a proof that we defer for the sake of continuity of the presentation).
To calculate the coefficients ak , bk of the sine series for f and g, take the
inner product of (2.29) and (2.30), respectively, with the basis element wk ,
and using the orthonormality relations,
+∞
f, wk = a w , wk = ak ,
=1
+∞
g, wk = b w , wk = bk .
=1
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
2.7. Separation of variables 27
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
28 2. Wave equations
Hydrogen atom. In this case x ∈ R3 , and one solves for the orbitals of
electrons in the presence of the attractive Coulomb potential of the atomic
nucleus. The Schrödinger equation describing this situation is
1 Z
(2.31) i∂t ψ = − Δ2 ψ − ψ.
2 |x|
Again, to solve this problem we separate variables in time and space to
reduce the question to a spectral problem:
1 Z
(2.32) − Δ2 ϕ − ϕ = λϕ .
2 |x|
n = 1, l = 0 n = 2, l = 0 n = 3, l = 0
n = 2, l = 1 n = 3, l = 1
n = 3, l = 2
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
2.7. Separation of variables 29
one finds that Φm (φ) are given in terms of complex exponentials, while
Θm (θ) are associated Legendre functions of cos(θ), and the radial depen-
dence Rn (r) is given in terms of associated Laguerre functions. The op-
erator (2.32) has continuous spectrum for λ ≥ 0 and an infinite number of
Z2
discrete negative eigenvalues λn = − 2n 2 , n = 1, 2, . . . , corresponding to the
energies of bound electron orbitals.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
30 2. Wave equations
0 2
One may check the fact that the hammers of a piano are designed to
strike the string at one-seventh the distance from the end, which tends to
suppress the presence of the anharmonic seventh overtone in the sound of
the note.
Woodwinds. The vibrations in an air column of a wind instrument obey the
Euler equations of fluid dynamics, which, in the case of the small air veloc-
ities and small pressure variations that hold in the body of the instrument,
can be assumed to satisfy the linearized equation for variations around the
state of constant ambient density ρ0 and zero velocity. This is the system
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
2.7. Separation of variables 31
of equations
(2.33) ∂t ρ + ρ0 ∂x v = 0 ,
ρ0 ∂t v + p (ρ0 )∂x ρ = 0
for small density variations ρ(t, x) and small air velocities v(t, x) in the air
column of the instrument. The pressure p(ρ) is determined as a function of
ρ by a constitutive law, given by the theory of thermodynamics. Taking an
additional time derivative of the first (or second) equation of (2.33) and then
using the second (or the first, respectively), we are led to the two essentially
equivalent wave equations
(2.34) ∂t2 v − p (ρ0 )∂x2 v = 0 ,
∂t2 ρ − p (ρ0 )∂x2 ρ = 0 ,
which differ through treatment at the boundaries x = 0, 2π. The wave
speed is c = p (ρ0 ), and we will take the liberty of choosing units so that,
conveniently, c = 1. Boundary conditions for ρ(t, x), respectively for v(t, x),
vary according to the design of the instrument.
The flute. This instrument has large holes at the mouthpiece and at the
end of the active part of the flute body, which effectively imposes that the
pressure of the air column in the flute coincides with atmospheric pressure
at these two boundary points; this gives boundary conditions
ρ(t, 0) = 0 = ρ(t, 2π) , equivalently ∂x v(t, 0) = 0 = ∂x v(t, 2π) .
Solving for the density ρ(t, x) using the method of separation of variables
gives us the eigenmodes wk (x) = π −1/2 sin( 12 kx), resulting in the expression
sin(ωk t) 1 1
ρ(t, x) = ak cos(ωk t) + bk √ sin( kx) .
ωk π 2
k≥1
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
32 2. Wave equations
0 2
eigenmodes are wk (x) = π −1/2 sin((2k − 1)x/4) and the frequencies of the
overtones are ωk = 12 (k − 12 ). The first five eigenmodes are illustrated in
Figure 10. Evidently ω1 = 1/4, ω2 = 3/4 = 3ω1 , and ω3 = 5/4 = 5ω1 , etc.
We remark that all of the even multiples of the fundamental frequency are
absent and that the lowest note of a clarinet is about one octave lower than
that of a flute, while the instruments are about the same size. A clarinetist
overblowing to the second overtone to achieve the second register of the in-
strument increases the tone by one-twelfth, or one octave and a major fifth.
This absence of even overtones is an explanation for the composer’s rule
of thumb to write the clarinet parts above the string parts. Otherwise the
strings may be obliged to accord their notes with overtones that are absent
in the notes of the clarinet.
The oboe and the saxophone. Both the oboe and the saxophone are reed
instruments; the saxophone uses a single reed similar to the clarinet, while
the oboe has a famously difficult double reed arrangement. Both impose
boundary conditions at the mouthpiece and at the bell, similar to the clar-
inet. However, both overblow an octave rather than a twelfth. The answer
to this conundrum is that they are both conical bore instruments, unlike
the cylindrical bore clarinet, and must be analyzed in terms of a three-
dimensional rather than one-dimensional air column. Making the reason-
able assumption that the density variation in the air column (expressed in
polar coordinates) is ρ(t, r), namely independent of angles (θ, φ), the wave
equation (2.34) becomes
2
(2.35) ∂t2 ρ − p (ρ0 )(∂r2 + ∂r )ρ = 0
r
with boundary conditions ∂r ρ(t, 0) = 0 and ρ(t, 2π) = 0. Making the sub-
stitution R = rρ(t, r) (d’après Beals [3]) leads to the equation
∂t2 R − p (ρ0 )∂r2 R = 0 ,
satisfying boundary conditions R(t, 0) = 0 = R(t, 2π). This is now addressed
as before by separation of variables techniques.
The fundamental is R1 (r) =
π −1/2 1 1
sin( 2 r) with frequency ω1 = 2 p (ρ0 ) while, just as in the case of the
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
2.7. Separation of variables 33
0 2
stringed instruments or the flute, the higher eigenfrequencies are 12 p (ρ0 ) k.
Again we find that ω2 = 2ω1 , ω3 = 3ω1 , etc. Eigenmodes in terms of the
original variables are given by ρn (r) = Rnr(r) . Alternatively, the problem
can be phrased in terms of the velocity v(t, r); see Project 2.14 at the end
of this chapter. A similar analysis can be performed in terms of the velocity
variables v(r), giving eigenmodes wn (r) and the same sequence of eigenvalues
ωn ; this is the topic of Project 2.14.
The drum. The eigenmodes of a drum correspond to those of a two-
dimensional disk BA (0) = {x = (x1 , x2 ) : |x|2 < A2 }. By separation of vari-
ables one seeks solutions (wk (x), λk ) of the eigenvalue problem for Laplace’s
equation, which in polar coordinates is
1 1
(∂r2 + ∂r )wk + 2 ∂θ2 wk = λk wk , w(A, θ) = 0 .
r r
The angular dependence can be described in terms of complex exponentials
eiθ . However, the radial dependence involves Bessel functions whose√zeros
do not generally give rise to an arithmetic series of frequencies ωk = −λk ,
as is the case for the one-dimensional instruments described above. And as
we know, drums are quite noisy instruments.
Synthesizers have allowed musicians to create new instruments with the
sound of a note determined by a chosen overtone series. D. DeTurck has
used this idea to synthesize the sound of an imaginary piano-like instrument,
where the role of the strings is replaced by higher-dimensional geometrical
objects such as spheres, tori, or hyperbolic Riemann surfaces, and the sub-
sequent overtone series of each note is determined by the square roots of the
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
34 2. Wave equations
Exercises: Chapter 2
Exercise 2.1. For the wave speed c(t) which is a general (locally integrable)
function of time, give a Fourier integral representation of the solution of the
transport equation (2.1).
Exercise 2.2. Find an explicit expression for the solution of the equation
∂t u + c∂x u + ru = 0 , u(0, x) = f (x) ,
using either (or both) of the above solution methods. When r > 0, what
happens to the solution u(t, x) asymptotically for t → +∞?
Exercise 2.3. Give a solution formula of the following equations using the
method of characteristics:
∂t u + x∂x u = 0 , u(0, x) = f (x) ;
∂ t u + x2 ∂ x u = 0 , u(0, x) = f (x) .
Then differentiate to check your answer. These two examples appear in
the text as (2.9) and (2.10). Explain the differences between the two cases
with regards to the behavior of solutions. Are the solutions u(t, x) unique
as functions defined over all of (t, x) ∈ R2 in both cases? What are the
asymptotic states
v(x) = lim u(t, x) ,
t→±∞
if the limit exists, that is?
Exercise 2.4. Express the solution to the following transport equation on
x ∈ R1 :
∂t u − xt∂x u = 0
for general initial data u(0, x) = f (x).
Suppose that the initial data f (x) has compact support, that is, it is
zero outside of an interval [−R, R]. For x = 0, what is the limit
lim u(t, x) ?
t→+∞
Exercise 2.5. Show that if the function u(t, x) solves the transport equation
(2.1), then the function v(t, x) = ∂x u(t, x) solves the conservation law (2.11).
Exercise 2.6. Give an explicit expression for the solution of the dispersive
equation
∂t u ± x∂x3 u = 0 , u(0, x) = f (x) ,
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Exercises: Chapter 2 35
using the Fourier transform and then the method of characteristics [4]. Com-
ment on properties of the solution when the indicated choice of sign is +
and when it is −.
∂t2 u − ∂x2 u = 0 ,
u(0, x) = f (x) , ∂t u(0, x) = g(x)
u(t, x) = V (x + t) + W (x − t) .
(ii) Suppose that ∂x f (x) = −g(x). Show that the left-moving component of
the solution vanishes.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
36 2. Wave equations
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 2 37
where ϑ is a parameter on the circle 0 ≤ ϑ < 2π. Use the method of sepa-
ration of variables to give an expression for the solution to the Schrödinger
equation in this setting.
In your analysis show that the eigenvalue problem
−∂x2 ϕk (x) = λk ϕk (x) , ϕk (x + 2π) = eiϑ ϕk (x)
gives rise to a symmetric operator A on L2 (T1 ) and that all of the eigenvalues
λk (ϑ) are real. Indexing the eigenvalues in increasing order, show that they
are 2π-periodic in ϑ. For which values of ϑ are the eigenvalues simple, and
for which values are they double? Comment on the properties of continuity
of these eigenvalues and of the associated eigenfunctions.
Projects: Chapter 2
Project 2.14 (Eigenmodes of the oboe). Formulated in terms of the density
ρ(t, r), we will assume that the air column in an oboe (or a saxophone) de-
pends only upon the radial coordinate r, where (r, θ, φ) are polar coordinates
in R3 .
(i) Making the substitution R(r) = rρ(r), show that the wave equation
(2.35) can indeed be rephrased for the radial function R(r) as
∂t2 R − p (ρ0 )∂r2 R = 0
with boundary conditions
R(t, 0) = 0 = R(t, 2π) .
See Beals [3] for a reference.
(ii) Using the method of separation of variables, the eigenmodes for the air
column are given by an eigenvalue problem for ρ(r), which in turn becomes
2
an eigenvalue problem for R(r), with evidently the eigenvalues λn = n4 and
eigenmodes
1 n
Rn (r) = sin( r) .
cn 2
Therefore rρn (r) = Rn (r) solves the eigenvalue problem (2.35). The L2
inner product for radial functions defined on B2π (0) is given by
2π
f, g = f (r)g(r) r2 dr .
0
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
38 2. Wave equations
(iii) The linearized Euler’s equations in three dimensions can be written for
(ρ, v), where ρ is the density as above and v is a velocity vector field:
∂t ρ + ρ0 ∇ · v = 0 ,
ρ0 ∂t v + p (ρ0 )∇ρ = 0 .
Show that the vector field v satisfies the wave equation
∂t2 v − p (ρ0 )∇(∇ · v) = 0
with boundary conditions v(t, 0) = 0 and ∇ · v(t, 2π) = 0.
(iv) The eigenvalue problem for the normal modes of the oboe can also be
phrased in terms of the velocity variables v, which is a vector field in the
three-dimensional problem. In polar coordinates (r, θ, φ) it can be written
v = (vr , vθ , vφ ) .
The gradient of a function f is given in polar coordinates by
1 1
∇f = ∂r f r̂ + ∂θ f θ̂ + ∂φ f φ̂
r r sin(θ)
and the divergence of a vector field u is given by
1 1 1
∇ · u = 2 ∂r (r2 ur ) + ∂θ (uθ sin(θ)) + ∂φ uφ .
r r sin(θ) r sin(θ)
Assuming that the motion of the air column in the instrument is radial, that
is,
v(r) = (vr (t, r), 0, 0) ,
show that the component vr (t, r) satisfies the scalar equation
1
∂t2 vr − p (ρ0 )∂r ∂r (r2 vr ) = 0
r2
with boundary conditions
1
vr (t, 0) = 0 , ∂r (r2 vr ) = 0 .
r2
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 2 39
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
40 2. Wave equations
(c) For most values of λ, the coefficients of this series grow rapidly and
the resulting formal power series does not converge. However, the series
terminates when
1
λ = λk = k + .
2
The resulting polynomials H(x) = Hk (x) are the Hermite polynomials, and
the eigenfunctions and eigenvalues of the quantum harmonic oscillator are
given by
1
ϕk (x) = ck Hk (x)e−x /2 , λk = k + .
2
2
The normalizing constants ck for the Hermite functions hk (x) = Hk (x)e−x /2
2
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 2 41
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
42 2. Wave equations
of the eigenstate is L = ( + 1). Once given k = ( + 1), the third equa-
tion is a spectral problem for the function R(r), which has eigenfunctions
that are ordered by n ∈ N and whose eigenvalues
Z2
λn = −
2n2
are determined by the solution process of the third equation. These are the
energies of the eigenstates. The full orbital is given by the eigenfunction
ϕ(x) = Rn (r)Φm (φ)Θm
(θ) .
The three numbers (n, , m) are the quantum numbers of the eigenstates;
they satisfy
n = 1, 2, 3 . . . , = 0, 1, 2, . . . , n − 1 , m = − , −( − 1) . . . ( − 0, ) .
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
10.1090/gsm/197/03
Chapter 3
43
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
44 3. The heat equation
where fˆ(ξ) is the Fourier transform of the initial data. With this we can
express the solution in integral operator form:
+∞
1
eiξx e− 2 ξ t fˆ(ξ) dξ
1 2
u(t, x) = √
2π −∞
1
eiξ(x−y) e− 2 ξ t f (y) dydξ
1 2
(3.2) =
2π
1
eiξ(x−y) e− 2 ξ t dξ f (y) dy := H(t, x − y)f (y) dy .
1 2
=
2π
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
3.1. The heat kernel 45
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
46 3. The heat equation
The heat kernel H(t, x − y) for t > 0 has its maximum at x √ = y, where
1
the maximum value is √2πt , and width, or standard deviation, is t (defined
as the distance between its center and inflection point). A graph of the heat
kernel taken at several snapshots in time is given in Figure 1.
It is reasonable to expect that we would obtain the same solution u(t, x)
at time t > 0 starting with data f (x) at time t = 0 if we alternatively solved
the heat equations up to time 0 < s < t, and then, starting again at s
with data u(s, x), continued the solution for time t − s. That is, in operator
notation
(3.5) H(t)f = H(t − s)H(s)f , 0<s<t.
This is the semigroup property, and it is related to the question of uniqueness.
In these terms the evolution operator for the heat equation can be writ-
ten as
+∞
u(t, x) = H(t)f (x) = H(t, ·) ∗ f (x) = H(t, x − y)f (y) dy .
−∞
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
3.2. Convolution operators 47
Proof. These assertions will follow from the proof that g(x) is integrable.
To this end, estimate
|g(x)| dx = h(x − y)f (y) dy dx
R1 R1 R1
≤ |h(x − y)||f (y)| dydx = h L1 (R1 ) f L1 (R1 ) ,
R1 R1
the last step following after an exchange of order of integration.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
48 3. The heat equation
The conclusion is powerful because of the fact that the existence of one
zero of the solution implies that both the whole solution and the initial data
must vanish identically.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
3.4. Initial and initial-boundary value problems 49
sets of zero Lebesgue measure), then Bδ is nonempty for some δ and there
is a bounded set A ⊆ Bδ with positive measure. Then
+∞
u(t, x) = H(t, x − y)f (y) dy
−∞
≥ H(t, x − y)f (y) dy
A
≥ H(t, x − y)δ dy .
A
This last expression is surely positive by the positive character of the heat
kernel H(t, x). There is more precise information available on a lower bound
for u(t, x) if we consider the Gaussian nature of the heat kernel:
1 (x−y)2
u(t, x) ≥ δ meas(A) inf √ e− 2t .
y∈A 2πt
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
50 3. The heat equation
Theorem 3.7. Suppose that f ∈ L1 (R1 ) and additionally that f (x) is con-
tinuous at a point x = x0 . Then
+∞
lim H(t, x0 − y)f (y) dy = f (x0 ) .
t→0+ −∞
for any continuous function f (x). Its mathematical sense is defined in terms
of a limit, as in the limit of the statement of the theorem.
for all t > 0 because of the normalization of the heat kernel. Furthermore,
for y = x0 the heat kernel satisfies
1 − (x0 −y)2
lim √ e 2t =0.
t→0+ 2πt
This limit is uniform for y bounded away from x0 , and in fact for δ > 0
fixed, for any ε > 0 there exists a time t(ε) > 0 such that for any smaller
0 < t < t(ε), both
1 − (x0 −y)2 ε
(3.7) √ e 2t f L1 < for any |y − x0 | ≥ δ ,
2πt 3
1 − (x0 −y)2 ε
(3.8) √ e 2t dy |f (x0 )| < .
|y−x0 |≥δ 2πt 3
Now split up the integral in expression (3.4) for the solution of the heat
equation in the following way. Given ε > 0, choose δ > 0 so that (3.6) holds
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
3.4. Initial and initial-boundary value problems 51
For 0 < t < t(ε), the first two integrals are bounded in terms of ε using, re-
spectively, (3.8) for the terms containing f (x0 ) and (3.7) for those containing
f (y):
x0 −δ +∞
1 − (x0 −y)2
+ √
e 2t [f (x0 ) − f (y)] dy
−∞ x0 +δ 2πt
x0 −δ +∞ 1
(x0 −y)2
≤ |f (x0 )| + √ e− 2t dy
−∞ x0 +δ 2πt
+∞
1 − (x0 −y)2
+ |f (y)| dy sup √ e 2t
−∞ |x0 −y|≥δ 2πt
2ε
≤ .
3
Using (3.6), the remaining integral over the interval [x0 − δ, x0 + δ] is also
shown to be bounded in terms of ε:
x0 +δ ε x0 +δ 1
1 − (x0 −y)2 (x0 −y)2
√ e 2t [f (x0 ) − f (y)] dy ≤ √ e− 2t dy
x0 −δ 2πt 3 x0 −δ 2πt
+∞ 2
ε 1 − (x0 −y) ε
≤ √ e 2t dy = .
3 −∞ 2πt 3
Since ε can be chosen arbitrarily small, this proves the statement of the
theorem.
Method of images. As with the wave equation, there are methods for
solving the heat equation in a region with a boundary, where boundary
conditions are imposed. Not surprisingly, this involves the method of images.
The two basic boundary conditions are homogeneous Dirichlet conditions,
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
52 3. The heat equation
1
∂t u = ∂x2 u , x > 0 , t > 0 ,
2
u(0, x) = f (x) , x > 0 ,
u(t, 0) = 0 ,
∂x u(t, 0) = 0 .
1 (x−y)2 (x+y)2
HD (t, x, y) = [H(t, x − y) − H(t, x + y)] = √ [e 2t − e 2t ] ,
2πt
where both x > 0 and y > 0. We may denote the solution operator for heat
flow on R+1 with Dirichlet boundary conditions by
+∞
HD (t)f (x) = HD (t, x, y)f (y) dy .
0
Similarly, initial data for the case of Neumann boundary conditions use
the method of images with an even reflection of the initial data, defining
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
3.4. Initial and initial-boundary value problems 53
1 (x−y)2 (x+y)2
HN (t, x, y) = [H(t, x − y) + H(t, x + y)] = √ [e 2t + e 2t ] ,
2πt
where again both x > 0 and y > 0. One may similarly denote the solution
1 with Neumann boundary conditions by
operator for heat flow on R+
+∞
HN (t)f (x) = HN (t, x, y)f (y) dy .
0
which we consider first for x ∈ R1 . The general principle given in section 2.5
indicates the solution process. As before, we decompose the solution as
u(t, x) = u1 (t, x) + u2 (t, x) where u1 is a solution of the initial value problem
with initial data f (x), for which we take the inhomogeneous term to be zero.
The component u2 (t, x) then accounts for the presence of a nonzero h(t, x),
which, following Duhamel’s principle, is given by the expression
t
u2 (t, x) = H(t − s)h(s, ·)(x) ds
0
t +∞ (x−y)2
1
= e 2(t−s) h(s, y) dyds .
0 −∞ 2π(t − s)
One can check that u2 (0, x) = 0, so that the component u1 (t, x) accounts
for the initial data f (x), and that
1
∂t u2 = ∂x2 u2 + h(t, x) .
2
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
54 3. The heat equation
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
3.4. Initial and initial-boundary value problems 55
and therefore
t +∞
1
(3.11) H(t − s, x − y) ∂s u − ∂y2 u dyds
0 −∞ 2
+∞
= u(t, x) − H(t, x − y)u(0, y) dy .
−∞
This is known as Green’s identity for the heat equation. From it we recover
the statement that if u(t, x) is a solution of the heat equation (3.1) which
decays at |x| → +∞ sufficiently well, then
+∞
u(t, x) = H(t, x − y)f (y) dy .
−∞
The heat kernel HD (t − s, x, y) satisfies (3.10). We also know that the limit
as s → t− of the heat kernel HD (t, s, x, y) = δx (y). Using this as well as the
fact that HD (t − s, x, 0) = 0, and furthermore assuming that the solution
u(s, y) does not grow faster than the decay of the heat kernel as y → +∞,
the above formula gives us a second version of Green’s identities (3.11),
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
56 3. The heat equation
namely
t +∞
1
HD (t, x, y) ∂s u − ∂y2 u dyds
0 0 2
+∞
(3.12) = u(t, x) − HD (t, x, y)u(0, y)dy
0
t
1
+ ∂y HD (t − s, x, 0)u(s, 0) ds .
0 2
Initial data u(0, x) = f (x), boundary data u(t, 0) = g(t), and inhomogeneous
data h(t, x) gives a full solution formula:
+∞
u(t, x) = HD (t, x, y)f (y)dy
0
t
1
+ ∂y HD (t − s, x, 0)g(s) ds
0 2
t +∞
+ HD (t, x, y)h(s, y) dyds .
0 0
A similar line of reasoning gives a formula for solutions of the heat equa-
tion on R1+ with prescribed inhomogeneous Neumann data on the bound-
ary x = 0. Namely, given initial data u(0, x) = f (x), boundary data
∂x u(t, 0) = g(t), and inhomogeneous data h(t, x), one finds the following:
+∞
u(t, x) = HN (t, x, y)f (y)dy
0
t
1
(3.13) − HN (t − s, x, 0)g(s) ds
0 2
t +∞
+ HN (t, x, y)h(s, y) dyds .
0 0
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
3.5. Conservation laws and the evolution of moments 57
where F (a) is interpreted to be the flux of heat across the boundary point
a into the interval, and F (b) is interpreted to be the flux of heat out of the
interval across the boundary point b. Fourier’s law for the heat flux gives
the form as
1
F (a) := − ∂x u(t, a) ,
2
and similarly for b so that
b
∂t Pt [a, b) = ∂t u(t, x) dx = F (a) − F (b)
a
b
1 1 1 2
= ∂x u(t, b) − ∂x u(t, a) = ∂ u(t, x) dx .
2 2 a 2 x
Since the interval [a, b) is arbitrary, the heat density u(t, x) must satisfy the
heat equation ∂t u = 12 ∂x2 u.
This interpretation must satisfy two conditions in order to be consis-
tent: (i) the total amount of heat must be conserved, and (ii) if the initial
heat distribution f (x) ≥ 0, then for all t ≥ 0 we should have u(t, x) ≥ 0.
The second condition is already satisfied for us by the maximum principle.
Regarding the first, there is the following result.
Theorem 3.8. If u(t, x) ∈ L1 (R1 ) is a solution of (3.1), then for all t ≥ 0
+∞ +∞
u(t, x) dx = f (x) dx .
−∞ −∞
Proof.
+∞ +∞
∂t u(t, x) dx = ∂t u(t, x) dx
−∞ −∞
+∞
1 2 1
= ∂x u(t, x) dx = lim ∂x u(t, R) − ∂x u(t, −R) = 0 .
−∞ 2 2 R→+∞
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
58 3. The heat equation
In fact the proof is incomplete, for we should also have shown that
when f (x) ∈ L1 (R1 ), then for t > 0 we have ∂x2 u(t, x) ∈ L1 (R1 ), and then
justified the exchanges of differentiation and integration and the integrations
by parts. These facts will be addressed in the later chapter on the properties
of the Fourier transform.
The interpretation of a solution u(t, x) of (3.1) as the density of a prob-
ability measure motivates a discussion of the time evolution of its moments.
For a given distribution f (x) dx, define the kth moment to be the quantity
+∞
(3.14) mk (f ) = xk f (x) dx .
−∞
For an arbitrary f (x) ∈ L1 (R1 ) this may be infinite or not well defined, but
for sufficiently well-localized functions it makes perfectly good sense. To
give examples, take the case of the heat kernel itself, then
+∞
1 − x2
m0 (H(t, ·)) = √ e 2t dx = 1 ,
−∞ 2πt
+∞
1 x2
m1 (H(t, ·)) = √ xe− 2t dx = 0 ,
−∞ 2πt
+∞
1 x2
m2 (H(t, ·)) = √ x2 e− 2t dx = t .
−∞ 2πt
Now consider the initial value problem for the heat equation; our solution
procedure yields u(t, x) = H(t, ·) ∗ f (x), and the resulting heat distribution
u(t, x) dx has moments which are given by
+∞
1 − (x−y)2
mk (u(t, ·)) = xk u(t, x) dx = xk √ e 2t f (y) dydx .
−∞ 2πt
Proposition 3.9. The first two moments of the solution u(t, x) are con-
served:
m0 (u(t, ·)) = m0 (f ) , m1 (u(t, ·)) = m1 (f ) .
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
3.5. Conservation laws and the evolution of moments 59
which is the first statement of the proposition. To show the second state-
ment,
√
g(x) dx = e−iξx g(x) dxξ=0 = 2πĝ(0) .
Hypotheses on the function g(x) are necessary in order to justify the inte-
grations by parts and the limits.
In particular for k = 0,
√
m0 (u(t, ·)) = 2π Ĥ(t, 0)fˆ(0) = 2π fˆ(0) = m0 (f ) ,
which gives the first statement of Proposition 3.9. For k = 1,
ˆ
m1 (u(t, ·)) = 2πi∂ξ Ĥ(t, ξ)f(ξ)
ξ=0
= 2πi ∂ξ Ĥ(t, ξ)f(ξ) + Ĥ(t, ξ)∂ξ fˆ(ξ) ξ=0
ˆ
= 2πiĤ(t, 0)∂ξ fˆ(ξ)ξ=0 = m1 (f ) .
This proves the second statement of Proposition 3.9, for initial data with
finite zeroth and first moments, at least. The general case is now clear:
ˆ
mk (u(t, ·)) = 2π(i∂ξ )k Ĥ(t, ξ)f(ξ) |ξ=0
k
k
= 2πik ∂ξj Ĥ(t, ξ)∂ξk−j fˆ(ξ) |ξ=0
j
j=0
k
k
= mj (H(t, ·))mk−j (f ) .
j
j=0
j even
We used the Leibnitz product rule for differentiation in the second line, and
in the third line we used the fact that odd moments of the heat kernel are
zero, as the kernel itself is an even function of x.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
60 3. The heat equation
Proof. By nonincreasing we mean that the norm in question could stay the
same or it could decrease in time; in any case it cannot increase. The easy
case is for f ∈ L1 to be of one sign; say for convenience that f (x) ≥ 0. Then
the maximum principle implies that u(t, x) ≥ 0 as well, so that
∂t |u(t, x)| dx = ∂t u(t, x) dx
1 2
= ∂t u(t, x) dx = ∂ u(t, x) dx = 0 ,
2 x
and the L1 -norm is unchanged by the flow. However, when the initial data
f (x) changes sign, the L1 -norm will in fact be decreasing in time. To show
this, decompose the initial data f (x) = f+ (x) − f− (x) into, respectively, its
positive and negative parts. Both f± (x) ≥ 0, and furthermore they have
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
3.6. The heat equation in Rn 61
Figure 2. Decomposition of initial data into positive (f+ (x)) and neg-
ative (−f− (x)) components, and their subsequent evolution under heat
flow, respectively u+ (t, x) and u− (t, x).
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
62 3. The heat equation
3.7. Entropy
The interpretation of nonnegative and normalized solutions of the heat equa-
tion as probability distributions has been discussed in section 3.5. Associ-
ated with such probability measures are a variety of concepts that concern
the entropy of a measure. Given a density f ∈ L1 (Rn ) with f (x) ≥ 0 and
f L1 = 1, the entropy is defined as
(3.19) S(f ) := f (x) log(f (x)) dx
Rn
with the entropy density given by
(3.20) s(x) = f (x) log(f (x)) .
The concept of entropy dates to the work of Boltzmann, who recognized
the quantity as a measure of the disorder of a distribution of particles in
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
3.7. Entropy 63
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
64 3. The heat equation
The first term of the RHS we know from (3.21). The second term evolves
as follows:
1
d ∂t e 1 Δe
u log(e) dx = ∂t u log(e) + u dx = Δu log(e) + u 2 dx
dt Rn R n e R n 2 e
1 ∇e 1 u ∇e 1 |∇e|2
=− ∇u · + ∇ · ∇e dx = − ∇u · − u 2 dx .
Rn 2 e 2 e Rn e 2 e
Since e(t, x) = H(t, x), the heat kernel, the latter quantity is explicit:
∇e x
=− .
e t
We can therefore compute that
∇e x n
∇u · dx = − ∇u · dx = u dx ,
Rn e Rn t t Rn
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
3.8. Gradient flow 65
and furthermore
1 |∇e|2 1 |x|2 1
− u 2 dx = − u 2 dx = − 2 |x|2 u(t, x) dx .
R n 2 e R n 2 t 2t R n
The latter is the variance of the probability distribution dP1 = u(t, x)dx,
which from section 3.5 we know satisfies
1 1
∂t |x|2 u dx = |x|2 Δu dx = Δ(|x|2 ) u dx = n u(t, x) dx .
Rn 2 Rn 2 Rn Rn
Since by Proposition 3.9 the moment m0 (u(t, ·)) = m0 (f ) is a constant, a
conserved quantity, we then have that
∇e 1 |∇e|2
− ∇u · − u 2 dx
Rn e 2 e
n
n 1
=− u dx + u dx + 2 |x|2 f (x) dx .
t Rn 2t Rn 2t Rn
Putting these together, we have a quite explicit form for the evolution of
the relative entropy S(u|e), namely
d 1 |∇u|2 n 1
(3.22) S(u|e) = − dx − m0 (f ) − 2 m2 (f ) .
dt 2 Rn u 2t t
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
66 3. The heat equation
Exercises: Chapter 3
Exercise 3.1. This problem concerns the semigroup property of the solution
process for the heat equation. Show that the heat kernel satisfies the identity
+∞
H(t, x) = H(t − s, x − y)H(y, s) dy ∀0 < s < t .
−∞
Now show that the heat operator satisfies the property that for all 0 < s < t,
H(t) = H(t − s)H(s) .
Conclude that formula (3.5) holds for the one-parameter family of solution
operators {H(t)}t∈R1+ of the heat equation.
Exercise 3.2. Justify on a rigorous level of analysis the exchange of integra-
tions in the proof of Proposition 3.4(iii), therefore completing the rigorous
proof of the proposition’s three parts.
Exercise 3.3. Solve the following initial value problems for the heat equa-
tion in explicit terms.
(1) f (x) = x
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Exercises: Chapter 3 67
(2) f (x) = x2
(3) f (x) = 0 for x < 0 and f (x) = 1 for x ≥ 0
(4) f (x) = eαx
(5) f (x) = sin(kx)
What is the asymptotic behavior of u(t, x) as t → +∞? What is their
asymptotic behavior as t → +∞? On what properties of the initial data
does this behavior depend? Does it affect the asymptotic behavior if f (x) ∈
L1 (R1 )?
Exercise 3.4 (Moments of the heat kernel). (i) Show that the second mo-
ment of the heat kernel
1 − x2
H(t, x) = √ e 2t = Pt (x)
2πt
is its variance and that
m2 (Pt ) = t .
Hint: Calculate the following integral:
+∞
1 x2
G(α) = √ e−α 2 dx ,
2π −∞
and then use the fact that
+∞
1 x2
√ e− 2 x2 dx = −2∂α G(1) .
2π −∞
(ii) Also calculate all the moments mj (Pt ) for j any odd integer.
(iii) Using the same technique as in (i), give an expression for all of the
moments mj (Pt ) for j even.
Exercise 3.5. Suppose that the initial data f (x) ∈ L1 (R1 ) for the heat
equation is piecewise continuous, meaning that it contains a possibly finite
number of jump discontinuities
lim f (x) = aj , lim f (x) = bj
x→x−
j x→x+
j
at points xj , j = 1, N . Show that the limit of the solution (3.4) of the heat
equation satisfies
1
lim u(t, xj ) = (aj + bj ) .
t→0 + 2
Exercise 3.6 (Method of images). Give an expression for the solution of
the heat equation over the bounded interval 0 ≤ x < 2π for the following
cases of boundary conditions.
(1) Dirichlet: u(t, 0) = 0 = u(t, 2π). Denote this solution by uD (t, x).
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
68 3. The heat equation
(ii) Clearly u(0, 2π) = 0. Show that for all t > 0 the solution nevertheless
satisfies u(t, 2π) = 0.
Exercise 3.8. Consider the three solutions with initial data 0 ≤ f (x) for
the three different boundary conditions in Exercise 3.6. Show that for all
0 ≤ x ≤ 2π and all t > 0, the three solutions satisfy
uD (t, x) < up (t, x) and uD (t, x) < uN (t, x) .
Is it also true that
up (t, x) < uN (t, x) ?
Exercise 3.9. Suppose that the initial data f (x) ∈ L1 (Rn ), and consider
the solution u(t, x) given by convolution with the heat kernel. Show that
lim u(t, x) = 0
t→+∞
uniformly in x ∈ Rn . Furthermore, give a decay rate for the solution by
showing that it is bounded by
1
|u(t, x)| ≤ √ n f L1 (Rn ) .
2πt
Exercise 3.10 (Entropy). (i) Define two distributions by their densities:
f1 (x) = χ−1/2,+1/2(x) , f2 (x) = 2χ−1/4,+1/4 .
They both have total mass m0 (fj ) = 1 for j = 1, 2. Compare their entropies.
Which is larger (less disordered)?
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 3 69
Projects: Chapter 3
Project 3.12 (Tikhonov-nonuniqueness of solutions of the heat equation).
It turns out that solutions of the heat equations are not unique, at least
if one admits solutions that grow very rapidly and oscillate as |x| → +∞.
This is a surprising result that appeared in Tikhonov [20]. To show this we
will solve the heat equation with data on the positive t-axis in the form
1
∂t u = ∂x2 u , x ∈ R1 ,
2
u(t, 0) = g(t) , ∂x u(t, 0) = 0 , t ∈ R+ .
(i) Derive a formal power series expansion in x for a solution u(t, x) in the
form
u(t, x) = c j xj ,
j≥0
and use the heat equation to show that for all j ∈ N,
2j j
c2j = ∂ g(t) , c2j+1 = 0 .
(2j)! t
(ii) Give functions g(t) such that this formal series for u(t, x) converges for
all x ∈ R1 . For example, functions in the Gevrey class will do, such as
g(t) = e− t2 ,
1
t>0.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
70 3. The heat equation
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
10.1090/gsm/197/04
Chapter 4
Laplace’s equation
71
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
72 4. Laplace’s equation
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
4.2. Green’s identities 73
The notation is that dSx is the differential of surface area in the integral
over the boundary. To ensure that the manipulations in this formula are
valid, we take Ω ⊆ Rn to be bounded and ask that u, v ∈ C 2 (Ω) ∩ C 1 (Ω);
that is, all derivatives of u and v up to second order are continuous in the
interior of Ω and at least their first derivatives have continuous limits on the
boundary ∂Ω.
Integrating again by parts (or alternatively using (4.5) in a symmetric
way with the roles of u and v reversed), we obtain Green’s second identity
(4.6) vΔu dx − Δv u dx = v∂N u − ∂N vu dSx
Ω Ω ∂Ω
For a second choice, let v(x) = u(x) itself. Then Green’s first identity
(4.5) is an identity involving the L2 -norm of ∇u, which is a form of an
“energy” identity
(4.7) |∇u(x)|2 dx + uΔu dx = u∂N u dSx .
Ω Ω ∂Ω
One consequence of this is a uniqueness theorem.
Theorem 4.2. Suppose that u ∈ C 2 (Ω) ∩ C 1 (Ω) satisfies the Dirichlet prob-
lem (4.1) with f (x) = 0, the Poisson problem (4.2) with h(x) = 0, or the
Neumann problem (4.3) with g(x) = 0. Then
|∇u(x)| dx =
2
u∂N u dSx = 0 .
Ω ∂Ω
Therefore u(x) = 0 in the cases of the Dirichlet problem (4.3) and the Pois-
son problem (4.2). In the case of the Neumann problem, the conclusion is
that u(x) is constant.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
74 4. Laplace’s equation
in each of the three cases, which in turn implies that ∇u = 0 almost ev-
erywhere in Ω. Therefore u(x) must be constant as we have assumed that
u ∈ C 2 (Ω). In cases (4.1) and (4.2), this constant must vanish in order to
satisfy the boundary conditions. In the case of the Neumann problem (4.3),
we only conclude that u(x) is constant. In this situation we conclude that
the constant functions u(x) ≡ C span the null space of the Laplace operator
with Neumann boundary conditions.
C 1
Γ(r) = , n≥3,
2−nr n−2
The function Γ(r) is harmonic for 0 < r < +∞ but singular for r = 0. It
is called a fundamental solution, which is part of the content of Theorem 4.3.
Its proof is the outcome of the following computation that uses Green’s
identities. Suppose that u ∈ C 2 (Ω) ∩ C 1 (∂Ω) and consider y ∈ Ω a point
inside the domain under consideration. Take ρ > 0 sufficiently small so that
the ball Bρ (y) ⊆ Ω, as in Figure 1. Using Green’s second identity (4.6) over
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
4.3. The fundamental solution 75
x2
y
B (y)
x1
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
76 4. Laplace’s equation
and as ρ tends to zero, by the continuity of u(x) at x = y, this has the limit
(4.11)
C(2 − n) C(2 − n)
lim u dSx = lim ωn ρn−1 u(y) = C(2 − n)ωn u(y) ,
ρ→0 ρn−1 Sρ (y) ρ→0 ρ n−1
where the quantity ωn is the surface area of the unit sphere Sn−1 ⊆ Rn .
Theorem 4.3. For u ∈ C 2 (Ω) ∩ C 1 (∂Ω) and for y ∈ Ω we have the identity
(4.12) Γ(|x −y|)Δu(x) dx = C(2−n)ωn u(y) + Γ∂N u−∂N Γ u dSx .
Ω ∂Ω
Setting C = ((2−n)ωn )−1 , we recover precisely the value of u(y) with this
identity. In fact this procedure need not be restricted to bounded
domains
Ω. If we specify that u ∈ C (R ) is such that limR→+∞ SR (0) (Γ∂N u −
2 n
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
4.4. Maximum principle 77
In words, this result states that a harmonic function u(y) is equal to its
average values over spheres Sρ (y) about y. This fact is consistent with dis-
cretizations of the Laplace operator in numerical simulations. For example,
the simplest version of a finite difference approximation of the Laplacian at
y ∈ Zn is
1
Δh u(y) := 2 u(x) − 2nu(y) .
h n x∈Z :|x−y|=1
A discrete harmonic function Δh u = 0 explicitly satisfies the property that
u(y) is equal to the average of its values at the nearest neighbor points
|x − y| = 1.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
78 4. Laplace’s equation
Proof. First of all, start with the stronger assumption of strict inequality,
Δu(x) > 0 in Ω. Suppose that at some point x0 ∈ Ω the function u(x)
achieves its maximum, u(x0 ) = maxx∈Ω (u(x)). Then ∇u(x0 ) = 0 and the
Hessian matrix of u satisfies
n
H(u) = ∂xj ∂x u(x0 ) j=1
≤0.
where the matrices (aj (x))nj,=1 are positive definite. The disadvantage is
that it relies upon the hypothesis that u(x) ∈ C 2 (Ω), which in the present
situation is stronger than necessary. The result on Gauss’ law of arithmetic
mean (4.13) in the previous section allows us to discuss maximum principles
for functions which are not C 2 (Ω) nor even C 1 (Ω), but which merely are
continuous.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
4.4. Maximum principle 79
The latter result is known as the strong maximum principle for subhar-
monic function. If u(x) is superharmonic, then −u(x) is subharmonic, and
therefore u satisfies a minimum principle. A corollary of Theorem 4.8 is that
harmonic functions, which are both sub- and superharmonic, satisfy upper
and lower estimates in the supremum norm. Namely, if u(x) is harmonic in
Ω and u(x) = f (x) for x ∈ ∂Ω, then the weak maximum principle implies
that
min (f (x)) ≤ u(x) ≤ max (f (x))
x∈∂Ω x∈∂Ω
for all x ∈ Ω.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
80 4. Laplace’s equation
Proof of Theorem 4.8. The proof of this second version of the maximum
principle is more specific to Laplace’s equation, using the property of subhar-
monicity. Consider Ω a domain which is connected and u(x) a subharmonic
function, and set M := supx∈Ω (u(x)). Decompose the domain into two
disjoint subsets:
Ω = {x ∈ Ω : u(x) = M } ∪ {x ∈ Ω : u(x) < M } := Ω1 ∪ Ω2 .
Since u(x) ∈ C(Ω), then Ω2 is open as a subset of Ω because an inequality
is an open condition. The claim is that the set Ω1 is also open in Ω, which
we will prove using the property of subharmonicity. Therefore, because of
the fact of being connected, either Ω = Ω2 and Ω1 = ∅, whereupon strict
inequality holds throughout Ω, or else Ω = Ω1 and Ω2 = ∅, in which case
u(x) = M a constant.
To prove the claim above that Ω1 is open, consider x0 ∈ Ω1 and refer
to the property of subharmonicity on all sufficiently small spheres Sρ (x0 )
about x0 which lie in Ω. This property implies that
1
0≤ u(x) − u(x0 ) dSx .
ωn ρn−1 |x−x0 |=ρ
The integrand is nonpositive because u(x) ≤ M = u(x0 ) in Ω. But the
integrand cannot be negative anywhere near x0 either, for that would violate
the above inequality. Therefore we must have Sρ (x0 ) ⊆ Ω1 for all sufficiently
small ρ, which is the statement that Ω1 is open.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
4.5. Green’s functions and Dirichlet–Neumann operators 81
The Green’s function for Rn+ . While it is usually not possible to have
explicit formulae for the Green’s function G(x, y) for a general domain, it
is straightforward for the domain consisting of the half space Rn+ . Let y =
(y , yn ) ∈ Rn+ so that yn > 0, and denote the point that is its reflection
through the boundary {xn = 0} by y ∗ := (y , −yn ) ∈ Rn− . The function
w(x, y) = −Γ(|x − y ∗ |) satisfies the property that it is harmonic in Rn+
(since its singularity at x = y ∗ is in Rn− ), and Γ(|x − y ∗ |) = Γ(|x − y|) when
x = (x , 0) ∈ ∂Rn+ . This gives rise to the following expression for the Green’s
function for the domain Rn+ :
(4.20) G(x, y) = Γ(|x − y|) − Γ(|x − y ∗ |)
1 1
=
(2 − n)ωn (|x − y | + (xn − yn )2 )(n−2)/2
2
1
−
(|x − y | + (xn + yn )2 )(n−2)/2
2
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
82 4. Laplace’s equation
x3
x’
x2
x1
when n ≥ 3. When n = 2,
1 (x − y )2 + (x − y )2
1 1 2 2
(4.21) G(x, y) = log .
4π (x1 − y1 )2 + (x2 + y2 )2
This procedure is similar to the method of images, as one can see. Actually,
in two dimensions there is a connection with the theory of complex variables;
because of the Riemann mapping theorem there is an expression for the
Green’s function for arbitrary simply connected domains in terms of formula
(4.21) and the Riemann mapping of the domain to the upper half plane R2+ .
The Green’s function for BR (0). A second case of a domain Ω for which
the Green’s function has an explicit expression is the ball BR (0) ⊆ Rn .
Given x, y ∈ Ω and another point y ∗ ∈ Ω, the function constructed from the
fundamental solution Γ(|x − y|) given by
G(x, y) = Γ(|x − y|) − c(y)Γ(|x − y ∗ |)
is harmonic as a function of x ∈ Ω,
Δx G(x, y) = 0 , x=y ,
with the singularity of the fundamental solution at x = y. In the special
case of Ω = BR (0), one may choose the exterior point y ∗ to be the inversion
of y in the sphere Sn−1
R = ∂BR (0), given by
R2
y∗ = y.
|y|2
For x ∈ Sn−1
R , this choice satisfies the special geometrical property that
|x − y ∗ | R
(4.22) = ,
|x − y| |y|
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
4.5. Green’s functions and Dirichlet–Neumann operators 83
The Poisson kernel. The Green’s function G(x, y) is the integral kernel
for the solution of the Poisson problem (4.2), but it also is relevant for the
Dirichlet problem (4.1). Let u ∈ C 2 (Ω) ∩ C 1 (Ω) be harmonic. Then by
Green’s identities
0= G(x, y)Δu(x) dx
Ω
= u(y) + G(x, y)∂N u(x) − ∂Nx G(x, y)u(x) dSx .
∂Ω
Using the fact that G(x, y) = 0 for x ∈ ∂Ω, we deduce that
(4.24) u(y) = u(x)∂Nx G(x, y) dSx
∂Ω
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
84 4. Laplace’s equation
In the case that Ω = Rn+ , the explicit form of the Green’s function (4.20)
(and (4.21) in the case in which n = 2) gives an equally explicit expression
for the Poisson kernel. This is
2 yn
(4.25) −∂xn G(x, y)xn =0 = D(x − y , yn ) =
.
ωn (|x − y |2 + yn2 )n/2
In the case of the ball Ω = BR (0), the Poisson kernel in (4.24) is given
by the expression
1 R2 − |y|2
(4.26) D(x, y) = ,
Rωn |x − y|n
as one finds from (4.22) and (4.23) after a short calculation.
1
+
(|x − y | + (xn + yn )2 )(n−2)/2
2
when n ≥ 3. When n = 2,
(4.29)
1
N (x, y) = log (x1 − y1 )2 + (x2 − y2 )2 (x1 − y1 )2 + (x2 + y2 )2 .
4π
Setting xn = 0 gives rise to the single layer potential
2 1
(4.30) S(x − y , yn ) =
(2 − n)ωn (|x − y | + yN
2 2 )(n−2)/2
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
4.5. Green’s functions and Dirichlet–Neumann operators 85
with which one solves the Neumann boundary value problem with data g(x),
namely
u(y) = S(x − y , xn )g(x ) dx := Sg(y) .
x ∈Rn−1
The physical interpretation of this expression is that it represents the elec-
trostatic potential induced by a charge distribution of density g(x ) on the
boundary ∂Rn+ = Rn−1 of the domain Rn+ . Similarly, the double layer po-
tential is given by
1 yn 1
u(y) =
f (x ) dx = Df (y) ,
x ∈Rn−1 ωn (|x − y | + yn )
2 2 n/2 2
and its interpretation is the resulting electrostatic potential from a distri-
bution of dipole charges with density f (x ) on the boundary of Rn+ .2 The
potential associated to a distribution ρ(x) of charge in Rn is given by a
convolution with the fundamental solution itself
u(y) = Γ(x − y)ρ(x) dx .
Rn
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
86 4. Laplace’s equation
n−1
−ξ j + |ξ |2 eiξ ·x e−|ξ |xn = 0 .
2
=
j=1
The other possible solution is eiξ ·x e+|ξ |xn , but this is ruled out by its growth
as xn → +∞. The Fourier transform allows us to decompose a general
3 Analogous
to the case above, technically the operator G is symmetric on C 1 (∂Ω) and self-
adjoint on an appropriate subspace H 1/2 (∂Ω) ⊆ L2 (∂Ω).
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
4.6. Poisson kernel on Rn+ 87
The function D(x , xn ) is an integral kernel for the solution operator, often
called the Poisson kernel for Rn+ or the double layer potential, and the solu-
tion u(x) is evidently given by convolution of the Dirichlet boundary data
with D(x , xn ). Evaluating the above Fourier integral expression (4.35), we
get an explicit expression:
1
D(x , xn ) = n−1
eiξ ·x e−|ξ |xn dξ
(2π)
2 xn
(4.36) =
.
ωn (|x | + x2n )n/2
2
The second line of (4.36) has already been derived as (4.25) in section 4.5
using the method of images. The constant ωn is the surface area of the unit
sphere in Rn .
Theorem 4.10. The solution of the Dirichlet problem on Rn+ with data
f (x ) ∈ L2 (Rn−1 ) is given by
(4.37) u(x) = D(x − y, xn )f (y ) dy
R n−1
1
= eiξ ·(x −y ) e−|ξ |xn dξ f (y ) dy .
(2π)n−1 Rn−1 Rn−1
For xn > 0, this function is C ∞ (differentiable an arbitrary number of
times).
Proof. For xn > 0, both of the above integrals in (4.37) converge absolutely,
as indeed we have
iξ ·x −|ξ |x
e e n ˆ
f (ξ ) dξ ≤ fˆ L2 (Rn−1 )
e−2|ξ |xn dξ
1/2
Rn−1 Rn−1
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
88 4. Laplace’s equation
The issue is whether the harmonic function u(x , xn ) that we have pro-
duced converges to f (x ) as xn → 0, and in what sense. In this paragraph
we will show that for every xn > 0, u(x , xn ) is an L2 (Rn−1 ) function of the
horizontal variables x ∈ Rn−1 , and that u(x , xn ) → f (x ) in the L2 (Rn−1 )
sense as xn → 0. By Plancherel,
u(x , xn ) − f (x ) L2 (Rn−1 ) = e−|ξ |xn fˆ(ξ ) − fˆ(ξ ) L2 (Rn−1 ) .
Because fˆ L2 < +∞, for any δ > 0 there is a (possibly large) R > 0 such
that
|fˆ(ξ )|2 dξ < δ .
|ξ |>R
We now estimate
u(x , xn ) − f (x ) 2L2 (Rn−1 ) = e−|ξ |xn − 1 fˆ(ξ ) 2L2 (Rn−1 )
−|ξ |x 2 −|ξ |x 2
= e n
− 1 fˆ(ξ ) dξ + e n
− 1 fˆ(ξ ) dξ
|ξ |≤R |ξ |>R
≤ e−Rxn − 1 |fˆ(ξ )|2 dξ + 2δ .
|ξ |≤R
The first term of the RHS vanishes in the limit as δ → 0. Since δ is arbitrary,
we conclude that u(x , xn ) converges to f (x ) as xn → 0 in the L2 (Rn−1 )
sense, which is that the L2 -norm of their difference tends to zero.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
4.6. Poisson kernel on Rn+ 89
Recall that 1
i ∂xj u(ξ) = ξj û(ξ), which motivates the notation for differential
operators that Dj = 1i ∂xj and the definition of a general Fourier multiplier
operator
1
m(D )f (x ) = √ n−1 eiξ ·x m(ξ )fˆ(ξ ) dξ
2π Rn−1
1
= n−1
eiξ ·(x −y ) m(ξ )f (y ) dξ dy .
(2π) Rn−1 ×Rn−1
In these terms the harmonic extension u(x) of the boundary conditions f (x )
can be written as
u(x) = e−xn |D | f (x ) := (Df )(x , xn ) .
Therefore
|∇u(x)|2 dx
Rn
+
+∞
= |iξ fˆ(ξ )|2 e−2|ξ |xn + |ξ |2 |fˆ(ξ )|2 e−2|ξ |xn dξ dxn ,
0 Rn−1
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
90 4. Laplace’s equation
Therefore whenever a solution of the Dirichlet problem (4.33) has the prop-
erty that f (x ) ≥ 0,
u(x) = D(x − y , xn )f (y ) dy > 0 ;
Rn−1
this follows from an argument that is very similar to the one we used for
Theorem 3.5 on the heat equation.
Theorem 4.11. Suppose that f (x ) ≥ 0; then for x ∈ Rn+ we have u(x) ≥ 0,
and in fact if at any point x ∈ Rn+ (meaning, with xn > 0) it happens that
u(x) = 0, then we conclude that u(x) ≡ 0 and f (x ) = 0.
This is our third encounter with the recurring theme in elliptic and
parabolic PDEs of comparison and maximum principles. In the case of
Laplace’s equation and other elliptic equations, analog results closely hold
on essentially arbitrary domains as well, although the Poisson kernel is not
generally so explicit and the proof is different. In our present setting, the
form of the Poisson kernel gives us a lower bound on the decay rates of
solutions u(x , xn ) for large xn .
Corollary 4.13. Suppose that f (x ) ≥ 0 and that A ⊆ {x : f (x ) ≥ δ} is
a bounded set of positive measure meas(A) > 0. Then
2 xn
u(x , xn ) ≥ δ meas(A) ,
ωn supy ∈A (|x − y |2 + x2n )n/2
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
4.8. Oscillation and attenuation estimates 91
≥ D(x − y , xn )f (y ) dy
A
≥ inf
D(x − y , xn ) f (y ) dy ,
y ∈A A
and of course
f (y ) dy ≥ δ meas(A)
A
while xn
inf D(x − y , xn ) ≥ .
y ∈A supy ∈A (|x − y |2 + x2n )n/2
Theorem 4.14. Suppose that f (x ) ∈ L2 (Rn−1 ) is the Dirichlet data for
(4.33) on Rn+ , and suppose in addition that
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
92 4. Laplace’s equation
Proof. Since dist supp(fˆ(ξ )), 0 > ρ, there is a δ > 0 such that inf ξ ∈supp(fˆ) |ξ|
> ρ(1 + δ). Using the Fourier representation for u(x),
1
u(x) = √ n−1 eiξ ·x e−|ξ |xn fˆ(ξ ) dξ ,
2π supp(fˆ)
we conclude that
1 −|ξ |x
|u(x , xn )| ≤ √ n−1 e n ˆ
f (ξ ) dξ
2π supp(fˆ)
1
e−|ξ |xn ( 1+δ ) e−|ξ |xn ( 1+δ ) |fˆ(ξ )| dξ
1 δ
≤ √ n−1
2π supp(fˆ)
|ξ |x
− inf ξ ∈supp(fˆ) ( 1+δn ) 1
e−|ξ |xn ( 1+δ ) |fˆ(ξ )| dξ
δ
≤e √ n−1
2π supp(fˆ)
1/2
1/2
e−|ξ |xn ( 1+δ ) dξ
2δ
≤ e−ρ|xn | |fˆ|2 dξ .
We have used the Cauchy–Schwarz inequality on the last line. We thus have
the estimate on the decay of u(x , xn ) for large xn , namely
C(δ)
|u(x)| ≤ e−ρ|xn | f L2 .
|xn |(n−1)/2
and assuming the hypotheses on the support of the Dirichlet data f (x) as
in Theorem 4.14, one follows a similar argument as in the proof above to
show that
||ξ |αn |fˆ(ξ )|e−|ξ |xn ( 1+δ ) dξ
δ
−ρ αn−1
|∂x u(x , xn )| ≤ e
α
|ξ1α1 . . . ξn−1
C(δ, α)
≤ n−1 e−ρ|xn | f L2 .
|xn | 2
+|α|
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
4.9. Hadamard variational formula 93
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
94 4. Laplace’s equation
Given one of the domains Ω(η), the boundary values of ϕk (x) on ∂Ω(η) are
eik·x e−|k|η(x ) ; therefore for xn > η(x ), we know that
D(η) eik·x e−|k|η(x ) = eik·x e−|k|xn .
where u(x) is the bounded harmonic extension of the Dirichlet data f (x )
to the domain Ω(η). Using again the family of harmonic functions ϕk (x) =
eik·x e−|k|xn , we compute its boundary values f (x ) and its normal derivative
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Exercises: Chapter 4 95
Exercises: Chapter 4
Exercise 4.1. Derive expression (4.36) for the case n = 2 from the Fourier
integral, and show that ω2 = 2π. Hint: Complex variables techniques would
be useful.
Exercise 4.2. In the case n = 2, the fundamental solution has the prop-
erty that ΔΓ(|x − y|) = δy (x). Show this by proving that the limit of the
expression in (4.10) vanishes and that the limit in (4.11) holds.
Exercise 4.3 (Separation of variables). Consider the boundary value prob-
lem for the domain B1 (0) ⊆ R2 for Dirichlet boundary data u(x) = f (x)
when |x| = 1:
Δu(x) = 0 , u(x) = f (x) for x ∈ S1 (0) .
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
96 4. Laplace’s equation
(ii) Show that the resulting solutions v(r) are therefore of the form
Δu = 0 .
(1) Dirichlet boundary data on the side {(x1 , 0) : 0 < x1 < 2π}
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Exercises: Chapter 4 97
(2) Dirichlet boundary data on opposite sides {(x1 , 0) : 0 < x1 < 2π}
and {(x1 , 2π) : 0 < x1 < 2π}
f (x) = u(x1 , 0) = x1 (2π − x1 ) ,
f (x) = u(x1 , 2π) = −2x1 (2π − x1 ) ,
while on the remaining two sides set f (x) = 0.
(3) Dirichlet boundary data on adjacent sides {(x1 , 0) : 0 < x1 < 2π}
and {(0, x2 ) : 0 < x2 < 2π}
f (x) = u(x1 , 0) = 2π − x1 , f (x) = u(0, x2 ) = 2π − x2 ,
and f (x) = 0 on the remaining two sides.
Exercise 4.6 (Dirichlet Laplace operator on the square domain D). Eigen-
functions of the Laplace operator on a domain Ω with Dirichlet boundary
conditions satisfy the equation
Δu = λu , u(x) = 0 for x ∈ ∂D ,
and λ is the associated eigenvalue. In the case that Ω = D is the square,
eigenfunction-eigenvalue pairs can be expressed by separation of variables
methods.
(1) Use the Ansatz u(x) = v(x1 )w(x2 ) to solve for eigenfunctions and
eigenvalues of the Dirichlet Laplacian on the square D and their
associated eigenvalues.
(2) Show that all of the eigenvalues satisfy λ < 0 and that the eigen-
function u(x) associated with the eigenvalue λ that is smallest in
absolute value satisfies
u(x) = 0 for x ∈ D .
(3) Show that this set of eigenfunctions is complete, meaning that
linear combinations of them are dense in L2 (D). Enumerate the
eigenfunction-eigenvalue pairs (uk (x), λk ) as you prefer, and nor-
malize the eigenfunctions so that
|uk (x)| dx = 1 ,
2
uk (x)u (x) dx = 0
D D
for k = .
(4) Use an eigenfunction expansion to give an expression for the Green’s
function for D in the form
1
G(x, y) = uk (x)uk (y) .
λk
k
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
98 4. Laplace’s equation
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 4 99
Projects: Chapter 4
Project 4.15 (Hadamard variational formula). This project is an extension
of the discussion in section 4.9 on perturbations of harmonic functions with
respect to their domain of definition. In particular, it studies perturbations
of the Green’s function of the domain. The goal is to prove the formula
that under perturbation of a domain Ω by small displacements η(x) in the
normal direction, the Green’s function for the domain, G(x, y) for x, y ∈ Ω,
satisfies the variational formula
(4.44) δG(x, y) = − Nz · ∇z G(x, z) η(z) Nz · ∇z G(z, y) dSz .
∂Ω
This formula was derived by J. Hadamard in 1915; it is from one point of
view a version of the resolvant identity.
Consider a compact domain Ω ⊆ Rn with real analytic boundary, and
a nearby domain Ω1 ⊆ Rn which is a real analytic perturbation of Ω, in
the sense that the boundary ∂Ω1 can be parametrized by x1 ∈ ∂Ω1 , where
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
100 4. Laplace’s equation
∇z G(x, z) · ∇z G1 (z, y) dz
Ω1
= Nz · G(x, z) G1 (z, y) dSz − G1 (x, y) = −G1 (x, y) .
∂Ω1
Conclude that
G1 (x, y)−G(x, y) = − ∇G(x, z) · ∇G1 (z, y) dz+ ∇G(x, z) · ∇G1(z, y) dz .
Ω1 Ω
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 4 101
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
10.1090/gsm/197/05
Chapter 5
Properties of the
Fourier transform
x2
Normalised Hermite function hn (x) = √
1
1 Hn (x)e− 2 , the graphs
(2n πn!) 2
of h1 , h2 , and h13 .
103
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
104 5. Properties of the Fourier transform
Proposition 5.2 (Elementary properties of L2 (Rn )). (1) The space of func-
tions L2 (Rn ) is a linear space; namely, if f, g ∈ L2 , then this implies that
af + bg ∈ L2 for all a, b ∈ C.
(2) The topology on L2 (Rn ) is given by a norm
1/2
f L2 = |f (x)|2 dx , f L2 = 0 ⇔ f = 0 .
Rn
Norms are homogeneous of degree 1:
αf L2 = |α| f L2 .
(3) Norms satisfy the triangle (or Minkowski) inequality:
f +g L2 ≤ f L2 + g L2 .
(4) The linear space L2 (Rn ) is complete with respect to the norm f L2 .
That is, if {fn }∞
n=1 ⊆ L (R ) is a Cauchy sequence of functions, then there
2 n
exists a limit function f (x) ∈ L2 (Rn ) such that fn (x) →L2 f (x) (meaning
fn (x) − f (x) L2 → 0).
(5) Furthermore, the norm on L2 (Rn ) is given by an inner product:
(5.1) f 2L2 = f, f where f, g = f (x)g(x) dx .
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
5.1. Hilbert spaces 105
Our favorite Hilbert space L2 (Rn ) has its inner product defined in (5.1).
One reason that L2 (Rn ) is a natural setting for the Fourier transform is
that it is preserved under the transform. In Rn , we define the transform as
follows. For ξ ∈ Rn ,
1
ˆ
f (ξ) = √ n e−iξ·x f (x)dx = (F f )(ξ) .
2π Rn
This equality between the L2 -norms of a function and its Fourier transform
is known as the Plancherel identity; it is a general fact about the Fourier
transform that holds in many settings. The proof of Theorem 5.4 is deferred
until the end of our discussion of Schwartz class.
Other examples of Hilbert spaces and Banach spaces as tools of analysis
include the following.
(1) Sobolev spaces H s (Rn ) (Hilbert spaces based on L2 -norms): For s ∈ N,
1
2
H (R ) = {f (x) :
s n
|∂xα f (x)|2 dx < +∞} .
0≤|α|≤s
The norm of a function f in this Sobolev space measures the size of deriva-
tives of f as well as the size of f itself. We are using our previously intro-
duced notation that a convenient device for multivariable objects involves
multi-indices α, β: α = (α1 , α2 , . . . , αn ), β = (β1 , β2 , . . . , βn ), with each
αi , βj ∈ N. The order of a multi-index is given by |α| = |α1 |+· · ·+|αn | so that
xα = xα1 1 xα2 2 . . . xαnn is a monomial of degree |α| and ∂xβ = ∂xβ11 ∂xβ22 . . . ∂xβnn =
β
Π∂xjj is a differential operator of order |β|. The family H s (Rn ) is a scale of
spaces which are nested in terms of decreasing index s: H s ⊆ H s−1 ⊆ · · · ⊆
L2 (Rn ).
(2) Lp spaces (Banach spaces): For 1 ≤ p < +∞,
1
p
L (R ) = {f (x) :
p n
|f (x)|p dx < +∞} .
Rn
For p = ∞, the space of bounded measurable functions on Rn is denoted by
L∞ (Rn ) and has a norm given by
f L∞ := sup |f (x)| .
x∈Rn
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
106 5. Properties of the Fourier transform
(3) Sobolev spaces W s,p (Rn ) (Banach spaces modeled on Lp -norms, which
again measure derivatives of f ):
1
p
W s,p (Rn ) = {f : |∂xα f (x)|p dx < +∞} .
0≤|α|≤s
(4) Schauder spaces C s,γ (Ω) (Banach spaces modeled on C 0 -norms): For
Ω ⊆ Rn a domain, that is, an open set with smooth boundary,
The many relationships of inclusion among these spaces, and their as-
sociated inequalities of norms, are important information for the analysis
of PDEs. A principal one that is very often used in PDEs is the Sobolev
Embedding Theorem. A version of this result is as follows.
(1 + |ξ|2 )s |fˆ(ξ)|2 dξ
2 2
(5.2) |f (x)| ≤ √ n dξ .
2π (1 + |ξ|2 )s
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
5.2. Schwartz class 107
where we have used the Plancherel identity for f (x) and its derivatives in
the last line. This holds of course for any integer s. It remains to bound the
RHS of (5.2),
1 1 +∞
1 1
2 n−1 2
dξ = r dr dSϕ ,
(1 + |ξ|2 )s S n−1 0 (1 + r 2 )s
and if s > n2 , then this quantity is finite, which finishes the proof.
with α! = α1 !α2 ! . . . αn !.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
108 5. Properties of the Fourier transform
That is to say, a sequence {fn }∞n=1 converges to f if and only if for all
multi-indices α, β, the seminorms
fn − f α,β = sup |xα ∂xβ (fn (x) − f (x))| → 0
x
and it is complete with respect to this metric. Clearly {fn (x)}∞ n=1 converges
in S to f (x) if and only if ρ(fn , f ) → 0 if and only if all fn − f α,β → 0 as
n → ∞. The notation for limits of sequences of functions in this Schwartz
class sense is to write S − lim fn = f . Metric spaces such as this one whose
topology is given by a countable family of seminorms are called Fréchet
spaces.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
5.2. Schwartz class 109
Lemma 5.7 (Technical lemma). This result quantifies the properties of scal-
ing limits in two cases, and these limits give rise to well-defined functionals
on S.
(i) (Approximations of the identity). Let g ∈ S with g(0) = 1. Then for all
f ∈ S,
S − lim (g(εx)f (x)) = f (x) .
ε→0
and by the Lebesgue dominated convergence theorem, the RHS of the last
term with r fixed tends to zero with ε → 0.
Proof. This uses Lemma 5.7: Take any g ∈ C0∞ with g(0) = 1. Then for
f ∈ S arbitrary,
g(εx)f (x) ∈ C0∞
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
110 5. Properties of the Fourier transform
and
S − lim (g(εx)f (x)) = f (x) .
ε→0
(4) e
ihx f (ξ) = fˆ(ξ − h), a momentum boost.
(5) σλ f (ξ) = 1 ˆ 1 ˆ
|λ|n σ(1/λ) (f )(ξ) = |λ|n f (ξ/λ).
The principal reason why Schwartz class is well suited for Fourier anal-
ysis is that it is invariant under the Fourier transform.
Theorem 5.10. The Fourier transform maps S into itself; that is, whenever
f ∈ S, then fˆ = F (f ) ∈ S.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
5.2. Schwartz class 111
Proof. For f (x) ∈ S, at the very least fˆ(ξ) makes sense as a convergent
integral. In order to check that fˆ(ξ) ∈ S, we have to check that all of
the seminorms are finite, which we do using the properties described in
Proposition 5.9:
fˆ α,β = sup |ξ α ∂ β fˆ(ξ)| ξ
ξ
1
1 1
= sup √ n e−iξ·x (( ∂x )α ( x)β f (x)) dx.
ξ 2π i i
This integrand is still bounded by Cβ+N,α (1 + |x|2 )− 2 , and hence the supre-
N
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
112 5. Properties of the Fourier transform
Now apply (5.7) to the result. Since G(ξ)ξ=0 = √2π 1
n,
1
f (0) = lim fˆ(ε)G(ε/ξ) dξ = √ n fˆ(ξ) dξ .
ε→0 2π
1
This recovers f (0) = √2π n fˆ(ξ) dξ. To generalize this procedure to cover all
x ∈ Rn is in fact simple, again using the list of properties in Proposition 5.9:
1 1
f (x) = (τ−x f )(0) = √ n F (τ−x f )(ξ) dξ = √ n eiξ·x fˆ(ξ) dξ .
2π 2π
We can finish this circle of ideas with a proof of the L2 (Rn ) Fourier
inversion theorem. We have already stated that Schwartz class S ⊆ L2 (Rn )
is a dense subspace. Lemma 5.12 states that for f, g ∈ S,
f, F −1 g = F f, g ,
where we also note that
(F −1 g)(x) = ĝ(−x) .
Therefore for every f ∈ S,
f 2
L2 = f, F −1 (F f )
= F (f ), F (f )
= fˆ 2
L2 ,
so the Fourier transform acts on the the subspace S ⊆ L2 (Rn ) isometrically
(in the sense of the L2 -norm). Since S is dense, given an arbitrary f ∈
L2 (Rn ), there is a sequence fn → f in the L2 (Rn ) sense of convergence,
with fn ∈ S. Then we define
F (1) (f ) := lim F (fn ) .
n→∞
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
5.3. Fourier transform of L1 -integrable functions 113
This completes the proof of Theorem 5.4. We will redefine the “extension
notation” F (1) := F at this time, dropping the extra superscript. The
statement is then that f L2 = F f L2 for all f ∈ L2 , which is the desired
result.
To finish our remarks we should note that complex isometries U have
the property that U∗ = U−1 ; namely, they are unitary operators. Let’s
verify this with the Fourier transform.
Definition 5.14. The adjoint T∗ of an operator T on a Hilbert space H
satisfies T∗ f, g = f, Tg for all f, g ∈ H.
= F −1 (g)(x)(f (x)) dx ,
Rn
which states that F∗ = F −1 , as desired.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
114 5. Properties of the Fourier transform
lim fˆ(ξ) = 0 .
|ξ|→+∞
Proof. The statement (i) of boundedness follows directly from the expres-
sion of the Fourier transform; indeed
1 1
|f (ξ)| = √
ˆ e −iξy
f (y) dy ≤ √ |f (y)| dy .
2π 2π
To prove the property of continuity (ii), consider the difference
1
fˆ(ξ1 ) − fˆ(ξ2 ) = √ [e−iξ1 y − e−iξ2 y ]f (y) dy .
2π
Given f ∈ L1 , for any ε > 0 we will show that if |ξ1 − ξ2 | is sufficiently small,
then this difference between f (ξ1 ) and f (ξ2 ) is less than ε in absolute value.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
5.3. Fourier transform of L1 -integrable functions 115
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
116 5. Properties of the Fourier transform
Proof. First of all we can describe the integrals over [−R, +R] in terms of
a convolution
√ +R iξx +R +∞
2π ˆ
e f (ξ) dξ = eiξ(x−y) f (y) dydξ
−R −R −∞
+∞ +R
iξ(x−y)
= e dξ f (y) dy
−∞ −R
+∞ iR(x−y)
e − e−iR(x−y)
= f (y) dy .
−∞ i(x − y)
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
5.3. Fourier transform of L1 -integrable functions 117
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
118 5. Properties of the Fourier transform
The remaining term of the RHS of (5.4) is also bounded by ε/2 for T suffi-
ciently large, namely
−δ +∞ +∞ 1 sin2 (T y)
+ [f (x − y) − f (x)] dx dy
2
−∞ +δ −∞ T y
−δ +∞ 1 sin2 (T y)
≤ 2 f L1 + dy
−∞ +δ T y2
−T δ +∞ sin2 (y )
≤ 2 f L1 + )2
dy .
−∞ +T δ (y
Since the function sin2 (y )/(y )2 is integrable, for δ > 0 fixed, one can take
T large so that this latter term is also bounded above by ε/2.
Proof. The heat kernel has Fourier transform H(t, ξ) = e− 21 |ξ|2 t of course,
and for t > 0, it is a member of Schwartz class. From Theorem 5.15(ii), we
know that fˆ(ξ) is bounded; thus the product satisfies e− 2 |ξ| t fˆ(ξ) ∈ L1 (R1 ).
1 2
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Exercises: Chapter 5 119
As long as 0 < t is taken sufficiently small, this quantity can also be shown
to be less than ε/2, as in the estimate (3.8) of the heat kernel.
The results of this subsection are related in spirit to Theorem 3.7 which
concerns the pointwise behavior of convolutions with the heat kernel as
t → 0+. It is the statement that at points x0 at which f (x) is continuous,
limt→0+ u(t, x0 ) = f (x0 ). A similar result holds for the Cesaro integrals,
with a similar proof. Namely, if f (x) is continuous at x0 , then
lim FT (x0 − y)f (y) dy = f (x0 ) .
T →+∞
Exercises: Chapter 5
Exercise 5.1. Show that the Fourier transform preserves angles between
vectors. For f, g ∈ L2 ,
Ref, g = f L2 g L2 cos θ
and
Refˆ, ĝ = fˆ L2 ĝ L2 cos ϕ .
Show that θ = ϕ.
Exercise 5.2. The Lp (Ω)-norm for functions on a domain Ω ⊆ Rn is defined
as
1/p
f Lp = |f (x)|p dx
Ω
for 1 ≤ p < +∞. The Banach space L∞ (Ω) is defined for p = +∞ by
the norm
f L∞ = sup |f (x)| .
x∈Ω
Because of Lebesgue measure, this supremum is taken modulo sets of zero
measure (the essential supremum). The Hölder inequality states that
f (x)g(x) dx ≤ f Lp g p ,
L
Ω
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
120 5. Properties of the Fourier transform
Exercise 5.4. Prove part (i) of the technical Lemma 5.7 on Schwartz class
functions. Namely, show that for g ∈ S(Rn ) such that g(0) = 1, then for all
f ∈ S(Rn ),
S − lim g(εx)f (x) = f (x) .
ε→0
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
10.1090/gsm/197/06
Chapter 6
Wave equations on Rn
121
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
122 6. Wave equations on Rn
The given initial data u(0, x) = f (x) is often referred to as the initial position
or displacement of the field u(t, x), while the data for ∂t u(0, x) = g(x) is
called the initial velocity, or momentum, as it specifies the time derivative
of the field u(t, x) at the initial moment.
gives a weak solution to the wave equation in the sense that (6.2) is satisfied.
Proof. Using expression (6.3) for the Fourier transform of the solution, we
give a similar formula for the vector (∂x u(t, x), ∂t u(t, x))T :
iξ
∂x u(t, x) 1 cos(|ξ|t) |ξ| sin(|ξ|t) iξ fˆ(ξ)
=√ n e iξ·x
iξ dξ .
∂t u(t, x) 2π |ξ| sin(|ξ|t) cos(|ξ|t) ĝ(ξ)
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
6.2. Lorentz transformations 123
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
124 6. Wave equations on Rn
while all other ∂xj = ∂x j are invariant. The Lorentz group of transforma-
tions on the space-time R1t × Rnx is generated by all spatial rotations R and
by the hyperbolic rotations H(ψ). It is a Lie group known as SO(1, n). The
set of space-time translations (t , x ) = (t + b, x + c) leave the wave equation
invariant, and so do the elements of the Lorentz group.
Proposition 6.3. The Lorentz transformations leave the d’Alembertian op-
erator invariant.
Proof. Since the group generators involving translations obviously leave the
d’Alembertian invariant, and all spatial rotations R just involve the Lapla-
cian, it suffices to check invariance under the hyperbolic rotations H(ψ).
Write the d’Alembertian operator using matrix notation:
T
∂t 1 0 ∂t
u = (∂t − Δ)u =
2
u.
∂x 0 −In×n ∂x
Use the transformation rule (6.6) for vector fields to check the invariance of
the d’Alembertian:
T T
∂ t H2×2 0 1 0 H2×2 0 ∂ t
.
∂x 0 I(n−1)×(n−1) 0 −In×n 0 I(n−1)×(n−1) ∂x
The only nontrivial part of this matrix calculation is the upper left hand
2 × 2 block,
T 1 0 cosh(ψ) sinh(ψ) 1 0 cosh(ψ) sinh(ψ)
H2×2 H2×2 =
0 −1 sinh(ψ) cosh(ψ) 0 −1 sinh(ψ) cosh(ψ)
cosh (ψ) − sinh (ψ)
2 2
0 1 0
= = ,
0 sinh2 (ψ) − cosh2 (ψ) 0 −1
which is precisely to say that the d’Alembertian is invariant under H(ψ).
This is the stated result.
The effects of the Lorentz transformation H(ψ) on the time axis {x = 0},
exhibited in Figure 1, are given by
x 1 = sinh(ψ)t , t = cosh(ψ)t ;
therefore their ratio
x 1
= tanh(ψ) := v
t
gives the velocity of the new frame of reference with respect to the old. We
note that |v| = | tanh(ψ)| < 1. In terms of v, the hyperbolic rotation can be
written in a more familiar form:
⎛ ⎞
√ 1 2 √ −v 2
H = ⎝ −v
1−|v| 1−|v| ⎠
.
√ 2
√ 1
2 1−|v| 1−|v|
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
6.2. Lorentz transformations 125
Figure 1. This is the image of the x1 - and t-axes and the light cone
under a Lorentz transformation consisting of a hyperbolic rotation of
the (t, x1 ) coordinates.
The set that remains invariant under the transformations of the Lorentz
group is the light cone LC = {t2 − |x|2 = 0}. This fact is checked as follows:
take (t, x) ∈ LC, then
T
t 1 0 t
=0.
x 0 −In×n x
As we have checked above,
T 1 0 1 0
L L= ;
0 −In×n 0 −In×n
therefore under a Lorentz transformation L : (t, x) → (t , x ), we have
T T
t 1 0 t t T 1 0 t
= L L =0.
x 0 −In×n x x 0 −In×n x
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
126 6. Wave equations on Rn
on the space-time R1t × Rnx . What we have just shown is that this metric is
invariant under the Lorentz group.
The invariance of the wave equation under the Lorentz group of trans-
formations was part of a paradox that physics faced towards the end of the
nineteenth century. As we have seen in (1.1), Maxwell’s equations are in-
timately tied to the wave equation. The equations of classical mechanics
are invariant under the Galilean transformation group, that is, the trans-
formation group consisting of rotations and translations of space along with
constant velocity translations. And the theory of electricity and magnetism
is invariant under the Lorentz group. The fact is that these are incom-
patible. Compatibility was restored in 1905 when Einstein introduced the
special theory of relativity. However, this was a revolutionary change in our
perception of the universe, for which many intuitive ideas about space-time
had to be modified. One of these we have seen above; the action of a hyper-
bolic rotation transforms the plane {t = 0} of spatial coordinates as well as
the time axis. Since we think of the spatial coordinate plane as being the
“present” state of the universe, it is a new idea that there is no universally
valid instant that can globally be considered to be the present, and that the
sense of the present is relative to the frame of reference of the observer. The
concept of simultaneity has to be sacrificed. In terms of the wave equation,
any one of the hypersurfaces {t = 0} can be used as a hypersurface on which
to pose initial data for the wave equation.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
6.3. Method of spherical means 127
Definition 6.4. Given a function h(x) ∈ C(Rn ), its spherical mean centered
about the point x ∈ Rn is
1
(6.7) M (h)(x, r) := h(y) dSy ,
ωn rn−1 |x−y|=r
Recall that ωn rn−1 is the surface area of the sphere Sn−1 of radius r
in n dimensions. When n = 1, the spherical mean of a function f (x) is
M (f )(x, r) = 12 (f (x + r) + f (x − r)), reminding one of the first term of the
d’Alembert formula. Changing variables in the integral (6.7), y → x + rξ,
ξ ∈ S1 (0), there is another useful expression for the spherical mean:
1
(6.8) M (h)(x, r) = h(x + rξ) dSξ .
ωn |ξ|=1
Expression (6.8) is defined for r ≥ 0, but it is clear from this second expres-
sion that M (h)(x, r) is an even function of r:
n−1
Δx M (h)(x, r) = ∂r2 + ∂r M (h)(x, r) ,
r
We notice that nj=1 ∂xj h(x + rξ)ξj = ∇h · N , the outwards normal de-
rivative of h on the unit sphere. Continue this line of calculation using
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
128 6. Wave equations on Rn
Green’s theorem:
n
1 1
∂xj h(x + rξ)ξj dSξ = rΔx h(x + rξ) dξ
ωn |ξ|=1 ωn |ξ|<1
j=1
1
= Δx h(y) dy
ωn rn−1 |x−y|<r
1 r
= Δx h(y) dSy dρ
ωn rn−1 0 |x−y|=ρ
1 r
= n−1 Δx ρn−1 M (h)(x, ρ) dρ .
r 0
Therefore
1
(6.9) Δx M (h)(x, r) = ∂r rn−1 ∂r M (h)(x, r)
rn−1
n−1
= ∂r2 + ∂r M (h)(x, r) ,
r
which is the result of the lemma.
Proposition 6.6. (i) For h(x) ∈ C(Rn ), the value of h(x) at any x ∈ Rn
can be recovered from its spherical mean:
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
6.3. Method of spherical means 129
Taking time derivatives and using the fact that u(t, x) is a solution of (6.1),
one obtains
2 1
∂t M (u)(t, x, r) = ∂ 2 u(t, x + rξ) dSξ
ωn |ξ|=1 t
1
= Δx u(t, x + rξ) dSξ = Δx M (u)(t, x, r) .
ωn |ξ|=1
Now use the Darboux equation of Lemma 6.5,
n−1
(6.10) ∂t2 M (u)(t, x, r) = ∂r2 +
∂r M (u)(t, x, r) ,
r
a PDE in the two variables (t, r) ∈ R2 . This is known as the Euler–Poisson–
Darboux equation.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
130 6. Wave equations on Rn
r−t
We have used the fact that t−r ρM (g)(x, ρ) dρ = 0, which holds because
rM (g)(x, r) is odd. Taking the limit as r → 0 and using Proposition 6.6, we
recover a representation for the solution u(t, x).
Corollary 6.8. Given initial data g(x) ∈ C 2 (R3 ) and f (x) ∈ C 3 (R3 ), the
spherical means solution given in (6.12) is a solution u(t, x) ∈ C 2 (R1t × R3x ).
Spherical means in Rn for odd n. There are similar expressions for the
solution of the wave equation for x ∈ Rn , for n ≥ 3 and odd. Returning
to the Euler–Poisson–Darboux equation (6.10) for the n-dimensional wave
equation
n−1
∂t2 M (u) = ∂r2 + ∂r M (u) ,
r
we seek an algebraic reduction to the wave equation in two dimensions in
the variables (t, r). This is able to be carried out in the odd-dimensional
case.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
6.3. Method of spherical means 131
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
132 6. Wave equations on Rn
Proof. (1) of Huygens’ principle. The result follows from the form of the
Kirchhoff formula. Express the solution u(t, x) in terms of spherical means
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
6.4. Huygens’ principle 133
over spheres of radius |t| in the initial hyperplane {t = 0} that are of the form
{y : |x − y| = |t|} (the intersection of the backwards light cone emanating
from (t, x) and the Cauchy hypersurface). If |x| > R+|t|, then the associated
sphere does not intersect BR (0), which contains the support of the initial
data, and hence the solution at that space-time point vanishes.
(2) of the strong Huygens’ principle. When |t| > R and |x| < |t| − R,
again the backwards light cone emanating from the point (t, x) does not
intersect the support of the initial data, in this case because the sphere
{y : |x − y| = |t|} is too big and has passed outside of BR (0).
One notes that this proof is geometric, only depending upon the fact
that the solution is represented as a spherical mean. Therefore the result
holds for solutions of the wave equation in arbitrary odd dimensions, as
exhibited in (6.13).
xn
x1
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
134 6. Wave equations on Rn
Corollary 6.12. If the initial data f (x), g(x) vanish on BR (0), then for 0 <
|t| < R, the solution u(t, x) must vanish on the cone {(t, x) : |x| < R − |t|}.
In particular if f, g are identically zero, then u(t, x) will also be zero.
HR - |t|, v
xn
x1
The results of Theorem 6.17 and Corollary 6.12 imply the following
pictures for the domains of influence and dependence of solutions of the
wave equation, given respectively in Figures 4 and 5.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
6.5. Paley–Wiener theory 135
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
136 6. Wave equations on Rn
The content of Paley–Wiener theory is to say that the converse also holds.
Theorem 6.16 (Paley–Wiener). Suppose that g(ζ) is an entire function of
exponential type R. Then there is a function f (x) ∈ C0∞ (Rn ) with supp(f ) ⊆
BR (0) such that
1
g(ζ) = f (ζ) = √ n e−iζ·x f (x) dx .
ˆ
2π
Proof. It is natural to restrict g to ξ ∈ Rn , and define
1
f (x) = √ n eiξ·x g(ξ)dξ.
2π R n
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
6.5. Paley–Wiener theory 137
and the integral remains absolutely convergent. Thus f ∈ C ∞ (Rn ) and the
question that remains has to do with its support.
Using the fact that g(ζ) is holomorphic, consider deformations of the
region of integration off of the real axis in the complex space Cn ,
1
√ n ei((ξ1 +iη1 )x1 +ξ ·x ) g(ξ1 + iη1 , ξ ) dξ1 dξ
2π
for η1 ∈ R. This is independent of η1 , as can be shown by Cauchy’s theorem,
taking the limit of an integral in ζ1 = ξ1 + iη1 over the contour given by
Figure 6 and letting T → +∞. The decay condition (6.14) implies that there
ξ1
-T T
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
138 6. Wave equations on Rn
Then by Proposition 6.15, both fˆ(ξ) and ĝ(ξ) extend to entire functions
on Cn of exponential type R. Let us examine the Fourier transform of the
solution
ˆ + sin(|ξ|t) ĝ(ξ) .
û(ξ, t) = cos(|ξ|t)f(ξ)
|ξ|
√
sin( ζ 2 t)
The individual functions cos( ζ 2 t)fˆ(ζ) and √ 2 ĝ(ζ) are entire func-
ζ
tions of ζ ∈ Cn , where we are using the notation that ζ 2 = ζ12 + ζ22 + · · · + ζn2
for ζj ∈ C. Furthermore,
sin(ζ 2 t)
|im(ζ)||t|
cos( ζ 2 t) ≤ e , ≤ e|im(ζ)||t| ;
ζ 2
ˆ
therefore the two products cos( ζ t)f(ζ) and sin( ζ 2 t)/ ζ 2 ĝ(ζ) are of
2
exponential type (R + |t|). We can conclude that the solution u(x, t) of the
wave equation is the Fourier transform of an entire function of exponential
type (R + |t|), and thus by the Paley–Wiener theorem, the solution u(x, t)
has its support in the ball BR+|t| (0) of radius R + |t|. This is a second proof
of Huygens’ principle for the wave equation.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
6.6. Lagrangians and Hamiltonian PDEs 139
From the assumption that v(0, x) = v(T, x) = 0, the last term of the RHS
vanishes. The conclusion is that because v(t, x) is otherwise arbitrary, we
must have that
∂t2 u − Δu = 0 .
These are the Euler–Lagrange equations for the action (6.17). Notice that,
ironically, the formal principle of stationary action allows us to give an initial
position u(0, x) = f (x) but does not allow for setting the initial momentum
∂t u(0, x) = g(x). Furthermore this formal principle asks us to specify the
final position u(T, x), so it is not in fact compatible with an initial value
problem. Nonetheless, the principle of least action, or more generally of
stationary action, remains a guiding principle for many equations in physics,
while on a rigorous mathematical level the principle remains a formal one
in this and other cases.
A Lagrangian L and subsequently an action integral S can be defined
for more general systems; indeed, this is how wave equations are derived in
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
140 6. Wave equations on Rn
where for clarity we have simplified the Lagrangian (6.18) above, and we are
using the notation that ∂t u = u̇. The action integral is as before
T
S= L dt,
0
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
6.6. Lagrangians and Hamiltonian PDEs 141
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
142 6. Wave equations on Rn
Proof. The formal equivalence of (6.20) and (6.22) is a general fact. Firstly,
H = u̇ p dx − L
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
6.6. Lagrangians and Hamiltonian PDEs 143
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
144 6. Wave equations on Rn
dinger equation, where the + sign is the defocusing case and the − sign is
the focusing case.
Exercises: Chapter 6
Exercise 6.1. This problem concerns the Klein–Gordon equation
∂t2 u − Δu + m2 u = 0 ,
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Exercises: Chapter 6 145
Exercise 6.4 (Global existence with small initial data for certain nonlin-
ear wave equations). This question is to show that certain nonlinear wave
equations possess smooth solutions for all t ∈ R; see [12]. This contrasts
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
146 6. Wave equations on Rn
with other cases where solutions form singularities in finite time. Consider
the equation
(6.26) ∂t2 v − Δv + (∂t v)2 − |∇v|2 = 0 ,
v(0, x) = f (x) , ∂t v(0, x) = g(x) ,
with (f (x), g(x)) ∈ C 1 (R3 ) × C(R3 ) supported in a compact set.
(i) Setting u = ev − 1, show that u(t, x) satisfies the wave equation
∂t2 u − Δu = 0 ,
u(0, x) = ef (x) − 1 := F (x) , ∂t u(0, x) = g(x)ef (x) := G(x) ,
(F (x), G(x)) ∈ C 1 (R3 ) × C(R3 ) .
Explain why (F, G) has compact support.
(ii) Restricting our attention to the case x ∈ R3 , show that for sufficiently
small F C 1 (R3 ) , G C(R3 ) , the solution u(t, x) is bounded by
|u(t, x)| < 1 ,
using the result of Exercise 6.3.
In this case the transformation v → u is invertible for all (t, x) ∈ R1t ×R3x ,
giving rise to a global solution v(t, x) of equation (6.26).
Exercise 6.5 (Method of descent for the wave equation for x ∈ R2 ).
(i) Show that if x ∈ R3 but the Cauchy data for the wave equation only
depends upon (x1 , x2 ), namely
(6.27) f = f (x1 , x2 ) , g = g(x1 , x2 ) ,
then the solution of the wave equation in R1t × R3x is also independent of x3 ,
u = u(t, x1 , x2 ) ,
and therefore u(t, x1 , x2 ) satisfies the wave equation in two space dimensions:
∂t2 u − (∂x21 + ∂x22 )u = 0 .
(ii) Use the Kirchhoff formula to express the solution to the wave equation
in R3 for data satisfying (6.27).
(iii) In the expression in (ii), reparametrize the spherical integrals by their
projection onto the (x1 , x2 ) plane, e.g.,
f (y) dSy
S2 :|x−y|=t
= f (y1 , y2 ) t2 − ((x1 − y1 )2 + (x2 − y2 )2 ) dy1 dy2 ,
|(x1 −y1 ,x2 −y2 )|<t
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 6 147
which gives a general formula in R2 for the solution of the wave equation
u = 0 .
(iv) Describe the nature of this solution in the case that the support of f
and g as functions on R2 is compact, say, supported in the ball BR (0) =
{|(x1 , x2 )| < R}. In particular comment on Huygens’ principle. Does the
solution violate the strong form of Huygens’ principle, and why? Describe
what an observer sees as time progresses when they are situated farther than
R from the origin.
Exercise 6.6 (Equipartition of energy). The energy function for the wave
equation is
1
E(u(t, ·)) = |∂t u(t, x)|2 + |∇u(t, x)|2 dx .
2 Rn
By reason of physical analogy, the first term
1
EK (u(t, ·)) = |∂t u(t, x)|2 dx
2 Rn
is called the kinetic energy, and the second term
1
EP (u(t, ·)) = |∇u(t, x)|2 dx
2 Rn
is called the potential energy. The total energy is a conserved quantity for
solutions of the wave equation.
Show that, on average, the kinetic energy and the potential energy for
a solution of the wave equation are equal; this is in the sense that
1 T 1 T
lim EK (u(t, ·)) dt = lim EP (u(t, ·)) dt .
T →∞ T 0 T →∞ T 0
Hint: You may assume that the initial data (f, g) are in Schwartz class. And
you may consider using the solution by Fourier synthesis for this problem
and apply the Plancherel identity, that is, the statement that
f 2 2 = fˆ 2 2 . L L
Projects: Chapter 6
Project 6.7 (Method of descent in Rn for even dimensions n = 2m). The
solution of the wave equation in even spatial dimensions is different in char-
acter than in odd dimensions. This is described in the case of R2 in Ex-
ercise 6.5 above. A similar procedure of descent gives an explicit spherical
means solution formula for solutions to the wave equation in higher even-
dimensional Euclidean space Rn , n = 2m. Derive a similar version of the
spherical means formulae to express the solution. You will note that, simi-
lar to the case of dimension n = 2, the solution does not satisfy the strong
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
148 6. Wave equations on Rn
Huygens’ principle. Assume that the initial data is smooth and compactly
supported, derive a decay rate for the solution (in the L∞ -norm), and show
that this rate is achieved on the light cones supported over the initial data.
Show that inside the light cones the solution decays faster, and derive the
rate of decay.
Project 6.8 (The Cauchy problem for the wave equation in the Fried-
man–Robertson–Walker space-time) [1],[13]). The metric for a Friedman–
Robertson–Walker (FRW) space-time is given in terms of the line element
in the form
ds2 = −dt2 + S 2 (t)dσ 2 , S(0) = 0 ,
defined on the half space-time R1+ × R3x , where dσ 2 = dx21 + dx22 + dx23 is the
Euclidean metric of each space-like hypersurface {(t, x) : t = Const.} R3 .
Changing time variable
dt
= S(τ ) = τ 2 ,
dτ
the metric becomes
ds2 = S 2 (τ )(−dτ 2 + dσ 2 ) .
This metric describes an emerging space-time from a Big Bang at t = τ = 0.
Consider the wave equation on R1+ × R3x in this metric,
1 2 2Ṡ 1
∂ u + 3 ∂τ u − 2 Δσ u = 0 .
u :=
S2 τ S S
Initial data is given on the Cauchy hypersurface {(τ0 , x)} R3 ,
u(τ0 , x) = g(x) , ∂τ u(τ0 , x) = h(x) .
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 6 149
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
10.1090/gsm/197/07
Chapter 7
Dispersion
151
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
152 7. Dispersion
Reconstituting the wave function with the inverse Fourier transform gives
the following expression of the solution:
1
ψ(t, x) = √ n eiξ·x e− 2 |ξ| t ψ̂0 (ξ) dξ
i 2
2π
1
eiξ·(x−y) e− 2 |ξ| t ψ0 (y) dydξ .
i 2
(7.2) = n
(2π)
Proposition 7.1. Evaluation of the complex Gaussian integral in (7.2) gives
the expression
1 1 |x−y|2
iξ·(x−y) − 2i |ξ|2 t
e e dξ = √ n e i 2t
.
(2π)n 2πit
Thus the Fourier representation of the solution leads to the expression for
the solution involving the Schrödinger kernel
1 |x−y|2
(7.3) ψ(t, x) = √ ne
i 2t
ψ0 (y) dy
Rn 2πit
:= S(t, x − y)ψ0 (y) dy = S(t)ψ0 (x) .
Rn
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
7.1. Schrödinger’s equation 153
One notices formal similarities of the Schrödinger kernel with the heat
kernel and its analogous representation of solutions of the heat equation.
On the other hand, the presence of the complex unit i in the exponent
radically changes the character of the Schrödinger kernel and the behavior
of solutions. Indeed, the heat kernel satisfies
1 |x|2
−
H(t, x) = √ ne 2t ∈S ∀t > 0 ,
2πt
and in particular for t > 0, the heat kernel H(t, x) decays very rapidly as
|x| → +∞ (faster than exponential decay). On the other hand,
1 i|x|2
S(t, x) = √ ne 2t ∈ C∞ ∀t ∈ R ,
2πit
but it is not Schwartz class as it does not decay as |x| → +∞, rather it
oscillates rapidly.
Theorem 7.2. (i) For initial data ψ0 (x) ∈ S, the solution expression (7.3)
gives rise to an absolutely convergent integral, and for all t ∈ R1 , ψ(t, x) ∈ S.
(ii) For initial data ψ0 (x) ∈ L2 (Rn ), the expression (7.3) for the solution
converges in the L2 sense, and for all t ∈ R1 , ψ(t, x) ∈ L2 (Rn ).
This last fact, that the L2 -norm is preserved by the Schrödinger flow,
is central to the interpretation of the wave function in quantum mechanics.
Consider initial data ψ0 (x) ∈ L2 (Rn ), normalized to satisfy ψ0 L2 = 1; this
can be viewed as a rescaling of ψ0 to lie on the unit sphere S∞1 ⊆ L . Then
2
the quantity
|ψ(t, x)|2 dx := dPt (x)
is a time-dependent distribution on x ∈ Rn with initial distribution
|ψ0 (x)|2 dx = dP0 (x) .
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
154 7. Dispersion
1 − i |ξ|2 t 2
= ξe 2 ψ̂0 (ξ) dξ = E(ψ0 ) .
Rn 2
The probability distribution dPt (x) is called the position distribution. One
can also study the momentum distribution
dP̂t (ξ) := |ψ̂(t, ξ)|2 dξ ,
which has initial distribution dP̂0 (ξ) = |ψ̂0 (ξ)|2 dξ. Again from the Plancherel
identity,
dP̂t (ξ) = |ψ̂(t, ξ)|2 dξ = |ψ(t, x)|2 dx = dP0 (x) ;
Rn Rn Rn Rn
thus the total momentum probability, or the zeroth moment of dP̂t (ξ), is
preserved under Schrödinger evolution. Furthermore, all moments of the
momentum distribution are preserved.
Theorem 7.3. Define a Fourier multiplier operator m(D) by the formula
1
m(D)ψ(x) = √ n eiξ·x m(ξ)ψ̂(ξ) dξ .
2π Rn
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
7.1. Schrödinger’s equation 155
Moments of the momentum distribution dP̂t (ξ) are of this form. There-
fore all moments of the momentum distribution are preserved under free
Schrödinger flow.
Proof. The result follows from a calculation of the integrals in the statement
of the theorem,
|m(D)ψ(t, x)|2 dx = |m(ξ)ψ̂(t, ξ)|2 dξ
= |m(ξ)ψ̂0 (ξ)| dξ = |m(D)ψ0 (x)|2 dx .
2
In particular, it follows that for k = 2s, these are Sobolev estimates of the
solution
m̂2s (ψ(t, ·)) = |ξ ψ(t, ξ)| dξ = |∂xs ψ(t, x)|2 dx
s 2
= |∂xs ψ0 (x)|2 dx = m̂2s (ψ0 ) .
It turns out that for ψ0 ∈ L2 (Rn ), the quantities mk (ψ(t, ·)) are not neces-
sarily bounded; see Projects 7.7 and 7.8 in this chapter. This is in contrast
to the recursive control of the moments of the analogous probability measure
given by the flow of the heat equation.
A second proof of the conservation law (7.4) for ψ(t, ·) 2L2 goes as fol-
lows. Take a domain A ⊆ Rn , that is, an open subset with sufficiently
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
156 7. Dispersion
d
ψ(t, ·) 2L2 = 0 ;
dt
hence ψ(t, ·) 2L2 = ψ0 2L2 for all t ∈ R1t . The advantage to this proof is that
it is very general, and in particular it applies to the Schrödinger equation
with potential
1
(7.5) i∂t ψ = − Δψ + V (x)ψ ,
2
where V (x) represents the interaction of the quantum particle in question
with a background potential energy. Two examples that we have already dis-
cussed in Chapter 2 are the quantum harmonic oscillator and the hydrogen
atom. The potential for the first is given by
1
V (x) = |x|2 ,
2
for which solutions of the associated Schrödinger equation (7.5) correspond
to the dynamics of a quantum particle trapped in a confining potential. For
the second example, let x ∈ R3 . The potential energy of the nucleus of the
hydrogen atom is given by
Z
V (x) = − ,
|x|
where the associated wave function ψ(t, x) describes the dynamics of the
electron of the hydrogen atom, which is a linear superposition of electron
orbitals and extended, or radiating, states.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
7.2. Heisenberg uncertainty principle 157
We have furthermore shown that the moments m0 (ψ(t, ·)) and m̂k (ψ(t, ·))
are constants of Schrödinger evolution. Introducing the Dirac “bra” :=
and “ket” := notation for inner products, the first moments of position
and momentum of an arbitrary wave function ψ(x) can be written as
m1 (ψ) = x|ψ(x)|2 dx = ψ|x|ψ ,
R
n
1 1
m̂1 (ψ) = ξ|ψ̂(ξ)| dξ = ψ(x) ∂x ψ(x) dx = ψ| ∂x |ψ .
2
i i
Since we are working in the setting x ∈ Rn , the operators 1i ∂x and x are
vectorial, and therefore so are the moments m1 (ψ) and m̂1 (ψ).
Proposition 7.4. Assume that the wave function ψ(x) ≡ 0 and that
m1 (ψ), m̂1 (ψ) < +∞. Under translation and change of phase, ψ(x) may
be adjusted so that m1 (ψ) = 0 = m̂1 (ψ).
= m1 (τx0 ψ) − x0 m0 (τx0 ψ) .
The variance Δx (ψ) of a wave function quantifies its deviation from its
mean value, which we can write as follows. Assume that m1 (ψ) = 0 = m̂1 (ψ)
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
158 7. Dispersion
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
7.3. Phase and group velocities 159
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
160 7. Dispersion
The wave equation gives a good example of this method; in matrix form,
1
0 1 P (ξ)t cos(|ξ|t) |ξ| sin(|ξ|t)
P (ξ) = and e = .
−|ξ|2 0 −|ξ| sin(|ξ|t) cos(|ξ|t)
Since the above sequence of calculation can be reproduced for any constant
coefficient operator P (D), one should guess that the Fourier kernel eP (ξ)t of
the solution operator cannot give rise to a well-defined distribution kernel
in all cases. There is, however, a simple criterion in the scalar case for
guaranteeing that we can proceed in this fashion.
Definition 7.6. Suppose that the polynomial iP (ξ) is real valued for all
ξ ∈ Rn . Then P (ξ) is said to be of real type, and
ω(ξ) = iP (ξ)
is known as the dispersion relation.
Proof. Well-posedness will mean in this context that if the initial data
u0 (x) ∈ L2 (Rn ), then there exists a solution u(x, t) of (7.15) with the prop-
erties that for each t ∈ R, u(x, t) ∈ L2 (Rn ), and that limt→0 u(x, t) −
u0 (x) L2 (Rn ) = 0, which is to say that the solution is continuous in L2 . To
show this, we will use the Fourier integral expression for a solution
1
(7.16) u(x, t) = √ n eiξ·x e−iω(ξ)t û0 (ξ) dξ .
2π
Using again the Plancherel identity, u(x, t) L2 (Rn
x)
= e−iω(ξ)t û0 (ξ) L2 (Rnξ ) =
û0 (ξ) L2 (Rnξ ) = u0 L2 (Rnx ) . The continuity in L2 follows arguments that we
have also seen before.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
7.3. Phase and group velocities 161
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
162 7. Dispersion
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
7.4. Stationary phase 163
where sgn ∂ξ2 ω(k) is the signature of the Hessian matrix (the signature of
a symmetric matrix Λ, denoted by sgn Λ, is the number of positive eigen-
values of Λ minus the number of negative ones). Expression (7.19) follows
from contour deformation of ξ into the complex plane; it gives a valid as-
ymptotic formula under a number of further assumptions, including the
hypothesis that det ∂ξ2 ω(k) = 0. This asymptotic calculation is discussed
in Exercise 7.6. If the hypothesis on the nondegeneracy of the Hessian is
not satisfied, the contribution to the asymptotic behavior of the solution
can be more prominent as the resulting asymptotic expression decays less
rapidly; this plays a role in the Kelvin ship wake problem which is the topic
of Project 7.10. In any case, the conclusion is that the principal contribution
to the solution u(t, x) is for the space-time points such that
x
= ∂ξ ω(k) ,
t
which is to say that the wave packet propagates at the group velocity cg (k).
The second part of this section on the method of stationary phase is
dedicated to a more robust analysis of the Fourier integral expression for
the solution
1 x·ξ
(7.20) u(t, x) = √ n eit( t −ω(ξ)) uˆ0 (ξ) dξ .
2π
Consider the data uˆ0 (ξ) supported in Br (ξ0 ), and take any point (x, t) ∈
Rn+1 such that for ξ ∈ supp(uˆ0 (·)), | xt − ∂ξ ω(ξ)| > R. Make use of the
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
164 7. Dispersion
identity
−i( xt − ∂ξ ω(ξ))
x·ξ
−ω(ξ)) x·ξ
eit( t = · ∂ξ (eit( t −ω(ξ)) )
t| t − ∂ξ ω(ξ)|
x 2
C1
≤ ,
R|t|
for R > 1 at least. Thus the solution is shown to decay in time along the ray
x
t = ∂ξ ω(ξ). Repeating the integration by parts that we performed above,
1 i( x − ∂ξ ω(ξ)) N
x·ξ
|u(x, t)| = √ n eit( t −ω(ξ)) (i∂ξ · xt ) u
ˆ0 (ξ) dξ
2π t| t − ∂ξ ω(ξ)|2
CN
≤ N N
R |t|
for large |t|. The picture of this is as follows.
Theorem 7.10. For initial data u0 (x) ∈ S such that supp(uˆ0 (ξ)) ⊆ Br (ξ0 ),
the solution u(x, t) is also in S for every t. Furthermore, on the set B =
{(x, t) : inf ξ∈Br (ξ0 ) | xt − ∂ξ ω(ξ)| > R}, we have for all N
CN
|u(x, t)| ≤ .
R |t|N
N
This is the statement that away from the rays of stationary phase { xt −
∂ξ ω(ξ) = 0; ξ ∈ supp(uˆ0 (·))}, the solution decays rapidly in space and time.
It may seem as though the above result only applies to quite special
initial functions u0 (x), but in fact it is rather generally applicable. For
example, the regions describing stationary phase asymptotics for two wave
packet solutions are illustrated in Figure 1. Indeed, we may furthermore
decompose any initial data u0 (x) into a sum of wave packets through a
partition of unity.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
7.4. Stationary phase 165
Proof. This can be constructed from a test function χ̃0 (y) ∈ C∞ 0 which
satisfies supp(χ̃0 ) ⊆ 32 Q, χ̃0 ≥ 0, and χ̃0 (y) = 0 for y ∈ Q. Set χ̃j (y) =
χ̃0 (y − j) for all j ∈ Zn ; it is supported in the cube j + 32 Q. We finally
have to adjust the magnitude. But this is easy as we notice that for every
y ∈ m + Q,
χ̃m (y) = χ̃k (y) > 0 .
m |k−m|<2
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
166 7. Dispersion
Theorem 7.12. For initial data u0 ∈ L2 (Rn ), the solution u(x, t) of (7.8)
is in L2 (Rn ). Furthermore the initial data can bedecomposed into a su-
perposition of compactly supported L2 functions j∈Zn (u0 )j where each
(u0 )j (x) ∈ L (j + 2 Q). The resulting solution is a superposition of wave
2 3
Proof. The initial data u0 (x) can be decomposed into a sum of compactly
supported L2 functions using a partition of unity from Lemma 7.11:
u0 (x) = (u0 )j (x) = (χj u0 )(x) .
j∈Zn j∈Zn
wave packet, asymptotically being very well localized near rays determined
by the relevant group velocity { xt = ∂ξ ω(j)}.
Exercises: Chapter 7
Exercise 7.1. The Schrödinger equation with a potential is given as
1
i∂t ψ = − Δψ + V (x)ψ .
2
(i) Show that for a potential V (x) that is bounded, the L2 (Rn )-norm of the
solution is conserved by the equation, namely
d
|ψ(t, x)|2 dx = 0 .
dt Rn
(ii) What is the quantum mechanical flux F in this case?
Exercise 7.2 (Conservation laws). There is an analogy between conserva-
tion laws and the Schrödinger equation, with or without a potential term.
The density function for the position distribution probability is given by
u(t, x) = |ψ(t, x)|2 .
Derive an evolution equation for the function u(t, x) of the form of a con-
servation law, namely
∂t u + ∇ · F = 0 .
What is the function F in terms of the wave function ψ(t, x)? In this equa-
tion, the vector function F plays the role of the quantum mechanical flux.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Exercises: Chapter 7 167
Exercise 7.3. Consider the initial value problem for Schrödinger’s equation
for x ∈ [0, 2π], namely
1
i∂t ψ = − ∂x2 ψ ,
2
ψ(0, x) = ψ0 (x) .
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
168 7. Dispersion
Projects: Chapter 7
Project 7.7 (Gaussian wave packets for the Schrödinger equation). Give
an explicit solution to the Schrödinger equation
1
i∂t ψ = − Δψ
2
for all t ∈ R , with initial data given by
1
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 7 169
(iii) However, for j = 2p even, show that there is a constant C2p such that
m2p (ψ(t, ·)) ≤ C2p m2p (ψ0 ) + m
2p (ψ0 ) + m0 (ψ0 ) .
Compare this situation with what is true for t > 0 in the case of higher
moments of the heat equation.
Project 7.9 (Electron orbitals for the hydrogen atom). In Project 2.16 of
Chapter 2 we studied the orbitals of electrons in the quantum mechanical
potential of the hydrogen atom nucleus. In that project we derived the
equations for the components of the orbitals using the method of separation
of variables. In this project we will complete the picture by solving these
equations for the special functions that arise from the method. Recall that
the dynamics of the wave function for an electron in the potential of a point-
like nucleus are given by the Schrödinger equation
1 Z
i∂t ψ = − ∂x2 ψ − ψ.
2 |x|
The coefficient Z is related to the atomic number of the nucleus and the
charge of the electron. The Ansatz for a wave function from the method of
separation of variables is that
ψ(t, x) = e−iλt ϕ(x) = e−iλt R(r)Φ(φ)Θ(θ) ,
where (r, φ, θ) are spherical polar coordinates. The three functions Φ(φ),
Θ(θ), and R(r) satisfy the following three equations. Firstly, the dependence
of the wave function upon the longitudinal angle φ ∈ T1 is given by
d2 Φ
= −m2 Φ , m∈Z,
dφ2
for which the solution is evidently Φ(φ) = eimφ for some m ∈ Z. Secondly,
dependence on the azimuthal angle is given by solutions of the associated
Legendre equation
1 d d m2
(7.22) − sin(θ) Θ + Θ = kΘ .
sin(θ) dθ dθ sin2 (θ)
Finally, the radial dependence of the wave function is given through the
associated Laguerre equation,
1 d 2d Z k
(7.23) r R + λ+ R = 2R .
2
2r dr dr |x| 2r
The goal of this project is to describe the solutions Θ(θ) and R(r) of the
latter two equations.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
170 7. Dispersion
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 7 171
(b) For large ρ, bounded solutions behave as R(ρ) ∼ e−ρ/2 which motivates
the Ansatz
R(ρ) = e−ρ/2 ρ G(ρ) .
The Lagrangian for this problem is
+∞ +∞
1 dR 2 2 1 1 ( + 1) γ
L= ρ dρ + − − + R2 dρ .
0 2 dρ 0 2 4 ρ2 ρ
Substitute the Ansatz into the Lagrangian and derive the Euler–Lagrange
equations δL = 0:
d2 G 2 dG γ − 2
+ − 1 + − 2 G=0.
dρ2 ρ dρ ρ ρ
(c) As with the case of the Legendre polynomials, substitute a formal power
series for the solution
G(ρ) = gj ρj .
j≥0
In order to satisfy the Euler–Lagrange equations, the coefficients must satisfy
the recursion relation
[(j + 2)(j + 1) + 2 (j + 1)]gj+2 = (j + 1 + − γ)gj+1 .
The series terminates when n = γ ≥ ( +1) is an integer, giving an associated
Laguerre polynomial Gn (ρ) as a solution.
Following our rescaling of the problem, γ = n; therefore
Z 2β 2 Z2
, λ = λn = −
β= =− 2 .
n 4 2n
2 3
These λn are the L (R ) eigenvalues of the hydrogen atom, while the radial
and angular dependence of the eigenfunctions are determined by the above
products of associated Laguerre functions Rn (r), associated Legendre func-
tions Θm
(θ), and finally the complex exponentials Φm (φ). The possible
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
172 7. Dispersion
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 7 173
The first equation expresses the fact that η(t, x) moves vertically accord-
ing to the vertical component of the fluid velocity. The second equation
expresses Newton’s law of forces; here g is the acceleration of the restoring
force of gravity, and in the absence of other forces the value of the pressure
on the free surface is taken to be zero. The function F describes any other
additional forces that impinge upon the upper free surface.
(i) Equations (7.24) describe the evolution of the free surface in terms of the
variables η(t, x) and ϕ(t, x, y), for which the only reference to the motion of
the body of the fluid is through the term ∂y ϕ(t, x, 0), where ϕ(t, x, y) is a
harmonic function in the fluid domain. Conveniently, this is the quantity
expressed by the Dirichlet–Neumann operator for the fluid domain, which
in the case of Rn+ is given in formula (4.38). Denoting the surface bound-
ary values of ϕ(x, y) by ξ(x) = ϕ(x, 0), show that the Dirichlet–Neumann
operator for R3− is given by
∂y ϕ(x, y) = |Dx |ξ(x) .
y=0
Therefore the equations of motion (7.24) for the linear water wave problem
are given by
η 0 |Dx | η 0
(7.25) ∂t = + .
ξ −g 0 ξ F
Eigenvalues of the matrix of the RHS are ±iω(k), giving the dispersion
relation, which in
our case of the water wave equations in deep water is
given by ω(k) = g|k|. This solution can be conveniently expressed in
complex coordinates, writing
"
1 4 g |k| ˆ
ẑ(t, k) = √ η̂ + i 4 ξ .
2 |k| g
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
174 7. Dispersion
(iv) The inverse Fourier transform of this expression gives the solution in
complex coordinates in the spatial variables. Transforming to coordinates
in which the ship is stationary, X = x + U t, show that the solution is in the
form
1
z(t, X) = eik·(X−U t) eiω(k)t ẑ(0, k) dk
2π
" ik·X
ip 4 |k| e − ei(k·(X−U t)+ω(k)t)
+ √ dk .
4 2π 2 g i(k · U − ω(k))
(v) Instead of solving the initial value problem from t = 0, suppose that
motion started at some asymptotically large past time −T , from which we
assume that all time-dependent transient waves are now negligible. In this
situation the stationary wave pattern of the ship wake is given by
"
ip 4 |k| eik·X
z(X) = √ dk .
4 2π 2 g i(k · U − ω(k))
The integrand has a denominator that vanishes along two curves. Assume
that U = (U1 , 0), parametrize these curves in polar coordinates in the Fourier
transform variables as (k1 (α), k2 (α)) = (κ(α) cos(α), κ(α) sin(α)) where
g 1
κ(α) = 2 ,
U1 cos(α)
and use the calculus of residues to express this two-dimensional integral as
a line integral.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 7 175
(vi) The line integrals from part (v) take the form
+π/2
z(X) = eiΘ(α,X) A(α) dα ,
−π/2
where the phase function is given by
Θ(α, X) = κ(α) cos(α)X1 + κ(α) sin(α)X2 .
In addition, the spatial variable can be represented in polar coordinates,
X = (X1 , X2 ) = (r cos(β), r sin(β)). For fixed X, show that the point of
stationary phase occurs when
tan(α) + (1 + 2 tan2 (α)) tan(β) = 0 .
Observe that this equation has two real roots when
1
| tan(β)| ≤ √ ,
2 2
which are distinct when the inequality is strict. These X = (r cos(β), r sin(β))
are in the region of the wake behind the ship. In the interior of the wake
region, two distinct wavenumbers are dominant at each point X.
(vii) As mentioned in section 7.4, points at which the derivative of the
phase degenerates have the potential to be more prominent features in the
solution. This is indeed the case with the wake of a ship. Show that the
second derivative of the phase function Θ(α, X) with respect to α vanishes
for
(1 + tan2 (α))(1 + 4 tan(α) tan(β)) = 0 .
Therefore both ∂α Θ(α, X) = 0 and ∂α2 Θ(α, X) = 0 when
1
| tan(β)| = √ ∼ 19.47◦ ,
2 2
that is to say, at the outer boundaries of the wake region described above.
Conclusion: The fact that people find remarkable is that this angle of the
most prominent features of a ship wake is independent of the speed |U | of the
ship. In addition, the wave number k(α) √associated with the waves produced
at this angle is such that α = arctan(1/ 2) ∼ 35.26◦ is independent of the
ship speed. What does change is the magnitude |k(α)|, as well as other
details of the wake. A continuation of this project is to discuss a similar
analysis of the wake of a ship which is in motion in a body of water over a
bottom of constant finite depth.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
10.1090/gsm/197/08
Chapter 8
177
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
178 8. Conservation laws and shocks
equation
(8.1) ∂t u + u ∂x u = 0 , x ∈ R1 ,
u(0, x) = h(x) .
∂t u + c ∂x u = 0 ,
Namely, the total quantity contained in the interval [a, b) changes in time t
only through flux out of the interval through the point x = b and flux into
the interval through the point x = a.
If f (u(t, x)) is differentiable at all x ∈ [a, b), (8.2) can be rewritten as
b b
0 = ∂t u(t, x) dx + ∂x f (u) dx
a a
b
= ∂t u(t, x) + ∂x f (u) dx .
a
Of course if u(t, ·) ∈ C 1 (R1 ), then this holds for any interval [a, b); therefore
u(t, x) satisfies the PDE
∂t u + ∂x f (u) = 0 ,
which is to say
(8.3) ∂t u + f (u)∂x u = 0 .
1 2
When the flux function is f (u) = 2u , this equation is precisely Burger’s
equation (8.1).
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
8.1. First-order quasilinear equations 179
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
180 8. Conservation laws and shocks
lie on the graph z = u(t, x). A solution to (8.5) is given as the union
of a parameter family of such characteristic curves. This family is locally
specified by a condition that the characteristic curves pass through another
given curve Γ = (T (β), X(β), Z(β)). In the above case of an initial value
problem (8.1), this is the curve Γ = (0, β, h(β)) for β ∈ R1 . To ensure that
the solution surface is well defined, one requires that the tangent vectors
to the curve Γ and the tangent vectors to the characteristic curves (8.6)
at points on Γ are transversal, and that they locally span a tangent space
which itself is a graph over the coordinate plane (t, x). This is the condition
on the curve Γ that it be noncharacteristic. That is,
a b
det =0.
∂β T ∂β X
In the case (8.1) and for the initial value problem, this is to say that
1 z
det =0,
0 1
a condition which, in the case of Burger’s equation, holds independently of
the initial data z = h(β).
Applying this construction to Burger’s equation (8.4), we have derived
the solution in the form
(8.7) u(t, x) = z(s, β) ,
where t=s, x(s, β) = β + sz , z(s, β) = h(β) .
Describing this solution process in words, the values of the solution u(t, x)
are given by h(β) where (β, h(β)) is the point where the characteristic curve
(t(s, β), x(s, β), z(s, β)) intersects the {t = 0} coordinate plane. Let us as-
sume that the initial data satisfies h ∈ C 1 (R1 ). To present the solution as a
graph, we need to solve the above relations (8.7) for u(t, x) = z(s, β). That
is, one considers the Jacobian of the mapping (s, β) → (t, x), which is given
by
t s xs 1 z(0, β)
(8.8) det = det .
t β xβ 0 1 + s∂β z(s, β)
For small times t = s, the Jacobian will be invertible, but the fact is that
for larger times t, the Jacobian determinant has the possibility to vanish.
Indeed, this is the case whenever 1 + t∂β h(β) = 0, in other words when
−1
t= .
∂β h(β)
Therefore, for h ∈ C 1 (R1 ), ∂β h is bounded and there is an interval of time
for which the above construction gives a C 1 solution to (8.1). However, at
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
8.1. First-order quasilinear equations 181
the first time t for which the Jacobian vanishes, the graph z = u(t, x) will
fold, developing a vertical tangent, and a discontinuity, or shock, will form.
Theorem 8.1. If the initial data h(x) for Burger’s equation is decreasing at
some point x = β, then the solution will become singular at a time T ∗ > 0,
where
−1
T∗ = .
inf β ∂x h(β)
Proof. We will give two proofs of this result. The first follows from the
Jacobian calculation that appears above. Let β0 be a point which minimizes
−1
∂x h(x). As t → T ∗ = − ∂x h(β0 ) , the Jacobian determinant (8.8) tends
to zero as the tangent plane of the graph z = u(t, x) becomes vertical. In the
case where there is no point which achieves the minimum, nonetheless the
Jacobian will tend to zero on a minimizing sequence (tn , xn ) where |xn | →
+∞ and tn → T ∗ as n → +∞, implying that the solution u(t, x) has an
unbounded C 1 -norm in the limit.
A second proof of this theorem on the inevitable formation of shocks
comes from an implicit relation satisfied by the solution. Namely, since
u = z(s, β) = z(0, β) = h(β), t = s, and β = x − sz(s, β),
u(t, x) = h(β) = h(x − tu) .
Taking a derivative of this relation with respect to x,
∂x u(t, x) = h (x − ut)(1 − t∂x u) ,
and solving for ∂x u, we get
h (β)
∂x u(t, x) = .
1 + th (β)
If h (β0 ) < 0 for some β0 , this expression becomes singular for t = −(h (β0 ))−1 .
If β0 is the infimum of h (β), then the first time t > 0 at which this expression
becomes singular is precisely T ∗ .
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
182 8. Conservation laws and shocks
One could ask about the variety of scalar conservation laws for different
flux functions f (u). Consider the general case as in (8.3),
∂t u + f (u)∂x u = 0 .
Suppose that u(t, x) ∈ C 1 in some neighborhood Ω ∈ R2tx , and assume that
f (u) is strictly convex, which is to say that f (u) is a strictly increasing func-
tion in u. Then g(v) = (f )(−1) (v) is well defined, and the transformation
u = g(v) is such that in Ω, the function v(t, x) satisfies
∂t u = g (v)∂t v , f (u)∂x u = f (g(v))g (v)∂x v .
Since f (g(v)) = v, v(t, x) satisfies
∂t v + v∂x v = 0 ,
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
8.2. The Riemann problem 183
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
184 8. Conservation laws and shocks
x
h = -1 h = +1
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
8.2. The Riemann problem 185
Figure 4. A shock wave for Burger’s equation (8.11) that does not
satisfy the entropy condition.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
186 8. Conservation laws and shocks
∂t u + ∂x f (u) = 0 ,
(8.13) ∂t u + a(u)∂x u = 0 ,
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
8.3. Lax–Olenik solutions 187
f3
f
2
f’ (v) (u - v)
f1
u-v
u
v u
Convexity and convex duality. The properties of f (u) and a(u) are used
in several ways. Firstly, the following inequality holds for arbitrary u and v:
(8.14) f (v) + a(v)(u − v) ≤ f (u) .
This inequality is a consequence of convexity, as illustrated in Figure 5.
Because of strict convexity, equality holds only for u = v. Being convex,
f (v) has a convex dual defined by
g(z) := a(v)v − f (v) , where z = a(v) .
The dual has the alternative definition
g(z) = min zv − f (v) ,
v
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
188 8. Conservation laws and shocks
f(v)
u
v
g (z)
and
1 1 1 1
f (v) = v p , g(z) = z p , where + =1,
p p p p
all of which are straightforward to check.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
8.3. Lax–Olenik solutions 189
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
190 8. Conservation laws and shocks
The elegant aspect of the theory is that the converse holds. Namely, the
variational formula given by (8.18) and (8.19) defines a weak solution to
(8.13) that in addition satisfies the entropy condition at every discontinuity.
Theorem 8.6. Formulae (8.18) and (8.19) define a possibly discontinuous
function u(t, x) for arbitrary initial data h(x) ∈ L1 (R1 ), which is a weak so-
lution with at most countably many jump discontinuities, and which satisfies
the entropy condition at such discontinuities.
The function u(t, x) defined by (8.18) and (8.19) is called the Lax–Olenik
solution.
Proof. First of all, because h(x) ∈ L1 , H(x) is bounded, and the convex
dual g(z) = a(v)v − f (v) for v = b(z) taken over the range of H(y) has
its minimum at some z = c where ∂z g(z) = 0. Therefore the function
G(x, y; t) defined in (8.19) is minimized and takes on its minimum at some
finite y. This minimizer is possibly nonunique. If y is the unique minimizer
of G(x, y; t), then ∂y G(x, y; t) = 0, at least for a.e. y at which H exists,
which implies that
x−y x−y
0 = H (y) − g = h(y) − b .
t t
In fact this y = y(t, x) is the minimizer for all points on the ray between
(0, y) and (t, x); indeed, for 0 < t1 < t and for x1 such that
x1 − y x−y
= ,
t1 t
∂y G(x1 , y; t1 ) = 0 so that h(y) = b( x1t−y
1
) as well. Thus the PDE (8.13) is
satisfied along this ray, and u(t, x) = h(y) = u(t1 , x1 ).
However, it could be that for given (t, x), G(x, y; t) has more than one
minimizer y(t, x); this situation is illustrated in Figure 8. In this case, the
following property of y(t, x) is useful.
Lemma 8.7. Fix t > 0 and denote y = y(t, x) the minimizers of G(x, y; t).
Then y(t, x) is a nondecreasing function of x.
Proof of Lemma 8.7. The result will follow from a fact that we shall prove
geometrically. Take x2 > x1 , then G(x2 , y1 ) < G(x2 , y) for all y < y1 . This
inequality depends upon the convexity of the function g:
x − y x − y x − y x − y
2 1 1 1 1 2 1
(8.20) g +g <g +g .
t t t t
The intervals [(x2 − y1 )/t, (x1 − y)/t] and [(x1 − y1 )/t, (x2 − y)/t] share
the same midpoint, and by convexity, the sketch in Figure 7 proves the
inequality.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
8.3. Lax–Olenik solutions 191
u
x1-y1 x1-y x2-y1 x2-y
The first conclusion we draw from this monotonicity is that for fixed
t > 0, the mapping x → y(t, x) is increasing in x with only countably many
jump discontinuities. Suppose that G(x, y; t) has more than one minimizer,
and let y − be the least minimizer of G(x, y; t) and y + > y − the greatest.
Because of the property of monotonicity, for fixed t > 0 if x1 < x, then any
y1 that is a minimizer of G(x1 , y; t) must satisfy y1 < y − , while if x1 > x,
then y1 > y + . Therefore the interval [y − , y + ] is a jump discontinuity of
y(t, x), and it follows that there can only be countably many points x for
which the minimizer y = y(t, x) is not unique.
Now consider t1 > t; for all x1 ∈ R1 , a minimizer y(t1 , x1 ) either sat-
isfies y(t1 , x1 ) < y − (t, x) or else y + (t, x) < y(t1 , x1 ). This is because for
any y − (t, x) < y < y + (t, x), the ray from (t1 , x1 ) to (0, y) will intersect one
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
192 8. Conservation laws and shocks
(t,x1) (t,x2)
x
y-(t,x1) y+(t,x1) y(t,x2)
of the rays (t, x) to (0, y ± (t, x)) and therefore cannot have the minimizing
slope of the latter. Hence when t1 > t, for each interval [y − (t, x), y + (t, x)]
of jump discontinuity of y(t, ·), there is a point x1 with multiple minimizers
of G(x1 , y; t1 ) and a corresponding interval [y − (t1 , x1 ), y + (t1 , x1 )] represent-
ing a jump in y(t1 , x). In particular, if {xk } and {xj1 } are the points of
discontinuity of y(t, x) and y(t1 , x) at times t and t1 > t, respectively, then
# #
[y − (t, xk ), y + (t, xk )] ⊆ [y − (t1 , xj1 ), y + (t1 , xj1 )] .
k j
The union of these points with nonunique minimizers form a set of shock
curves, which by the above property may merge and come into existence,
but may not split as t increases. The presence of such space-time points
for which G(x, y; t) has nonunique minimizers implies that a solution given
through the Lax–Olenik variational principle is not a classical solution.
We now show that it is a weak solution in the sense that it satisfies the
jump condition (8.9) at points on shock curves.
Consider t1 > t and points (t, x) and (t1 , x1 ) on a shock curve, where
each has nonunique minimizers y ± and y1± , respectively, and such that
[y − (t, x), y + (t, x)] ⊆ [y1− (t1 , x1 ), y1+ (t1 , x1 )] .
Since they are nonunique minimizers, from the Lax–Olenik variational prin-
ciple we know that
(8.21) G(x, y − ; t) = G(x, y + ; t) , G(x1 , y1− ; t1 ) = G(x1 , y1+ ; t1 ) ,
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
8.3. Lax–Olenik solutions 193
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
194 8. Conservation laws and shocks
where one uses in the second line that b(z) has Lipschitz constant Cb :
b(z2 ) − b(z1 ) ≤ Cb (z2 − z1 ) for z1 ≤ z2 .
Therefore approaching a point x of discontinuity of u(t, x) = b((x − y)/t)
from the left by x1 and from the right by x2 , our solution satisfies
u− (t, x) ≥ u+ (t, x) ,
which is the entropy condition for a jump discontinuity.
where yn (t) are the points of sign changes of the difference u1 (t, x)−u2 (t, x),
with sgn(u1 (t, x) − u2 (t, x)) = (−1)n over the interval (yn (t), yn+1 (t)). We
will analyze the behavior of the individual integrals of this sum.
Case (1). Suppose that yn (t) is a point of continuity for both u1 (t, x) and
u2 (t, x). In this case, both these solutions are constant along characteristics,
which in turn are linear. Therefore yn (t) is linear in t, and u1 (t, yn (t)) −
u2 (t, yn (t)) = 0.
Case (2). When yn (t) is a point of discontinuity of either u1 (t, x) or u2 (t, x),
or both, yn (t) is a shock curve.
Using the nature of conservation laws and their weak solutions, the time
derivative of the difference satisfies
d d yn+1 (t)
u1 (t, x) − u2 (t, x) L1 = (−1)n (u1 (t, x) − u2 (t, x)) dx
dt n
dt yn (t)
yn+1 (t)
+ (−1)n ∂t (u1 (t, x) − u2 (t, x)) dx
n yn (t)
dyn+1
+ (−1)n (u1 (t, yn+1 (t)) − u2 (t, yn+1 (t)))
n
dt
dyn (t)
− u1 (t, yn (t)) − u2 (t, yn (t)) .
dt
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Exercises: Chapter 8 195
u − u+ u−
2 − 1 − u2
≤ f (u2 ) − −
1
+ f (u1 ) + − + f (u1 ) ≤ 0 .
+
u1 − u1 u1 − u1
Hence this boundary term gives a negative contribution:
dy
(−1)n (u1 − u2 )(y) ≤0.
dt y=yn+1 (t)
Inspecting the other cases, they also only give negative contributions. Hence
the L1 -norm of the difference is decreasing. In the case that the initial data
satisfies h1 = h2 , this implies the uniqueness of entropy condition-satisfying
solutions and thus uniqueness for the class of Lax–Olenik solutions.
Exercises: Chapter 8
Exercise 8.1. Suppose that f (x, u) = b(x)u is the flux function for a linear
scalar conservation law, and consider a solution u(t, x) ∈ C 1 (R2 ):
∂t u + ∂x (b(x)u) = 0 .
Suppose that x = X1 (t) and x = X2 (t) are two characteristic curves such
that X1 (t) < X2 (t). Show that for all t ∈ R1 ,
X2 (t) X2 (0)
u(t, x) dx = u(0, x) dx .
X1 (t) X1 (0)
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
196 8. Conservation laws and shocks
Exercise 8.2. Solve the linear scalar conservation law with flux function
f (x, u) = x3 u:
u(0, x) = h(x) ;
h(x) = h− , x<0, h(x) = h+ , x≥0.
(ii) Work out the more complicated but still explicit (entropy condition-
satisfying) solution to the family of problems with piecewise constant initial
data u(0, x) = h(x), where
Note: These solutions are global in (t, x) ∈ R1+ × R1 , and they involve shock
interactions and interactions between shocks and rarefaction waves. There
are eight cases depending on the relative inequalities between h− , h0 , h+ .
Projects: Chapter 8
Project 8.5 (The Euler equations and shear flows in fluid dynamics). The
motion of a fluid is often represented by its velocity field
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 8 197
(i) Shear flows are special solutions of the Euler equations in the domain
Ω = Rd+ . Considering the case d = 2, a shear flow is a velocity vector field
given by
u(t, x1 , x2 ) = (u1 (x2 ), 0) ,
where the function u1 (x2 ) is the shear profile. Remark that this flow is
time independent, so that if initially u(t = 0, x) is Hölder continuous with
exponent 0 ≤ α ≤ 1,
|u(t, x) − u(t, x )| ≤ C|x − x |α ∀x, x ∈ R2+ ,
then it remains so for all time t ∈ R1 .
(a) Show that this vector field is incompressible and that it satisfies the
Euler equations (8.26).
(b) Calculate the vorticity of the velocity vector field.
(c) Show that the velocity vector field satisfies the boundary conditions on
{x2 = 0}.
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
198 8. Conservation laws and shocks
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Projects: Chapter 8 199
start with a solution that forms a shock, and consider the behavior of nearby
initial data and the resulting solutions. If instead the function class one con-
siders is that of continuous functions with piecewise continuous derivatives,
is the solution map continuous in this case?
(ii) Show that the solution map ϕt (u0 ) for Burger’s equation is continuous,
locally in time, when C is the Sobolev space H s (R1 ) for s > 1. That is, for
u0 ∈ K ⊆ H s (R1 ) a bounded open set, there exists a time T = T (K) > 0
such that the initial value problem satisfies the well posedness criterion of
Hadamard up to times |t| < T .
(iii) However, the solution map on the class H s (R1 ) is not smooth, nor even
Hölder continuous; this is a theorem of Kato [10]. His counterexamples to
any possible Hölder modulus of continuity are constructed as follows.
(a) Define a family of initial data for Burger’s equation to be
where ϕ(x) is a C0∞ cutoff function such that ϕ(x) ≡ 1 for −1 < x < +1.
The notation is that
x+ = 0 (x < 0) , x+ = x (x > 0) ,
and the exponent α > 0 will be taken large, according to the Sobolev regular-
ity exponent s. Show that for α > s−1/2, both uλ0 (x) and u00 (x) are H s (R1 ).
(b) The characteristics for this problem for λ = 0 and λ > 0 are, respectively,
x = y 0 + u00 t = y 0 + (y+
0 α+1
) t, x = y 0 + uλ0 t = y λ + (λ + (y+
λ α+1
) )t ,
where we are restricting our attention to the region where |x|, |y| < 1. Show
that the respective solutions u0 (t, x) and uλ (t, x) satisfy
Therefore one concludes that ∂xs uλ (t, x) ∈ L2 (R1 ) and the same for ∂xs u0 (t, x).
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
200 8. Conservation laws and shocks
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Bibliography
[1] B. Abbasi and W. Craig, On the initial value problem for the wave equation in Friedmann-
Robertson-Walker space-times, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 470 (2014),
no. 2169, 20140361, 13, DOI 10.1098/rspa.2014.0361. MR3238168
[2] C. Bardos and E. S. Titi, Loss of smoothness and energy conserving rough weak solutions
for the 3d Euler equations, Discrete Contin. Dyn. Syst. Ser. S 3 (2010), no. 2, 185–197, DOI
10.3934/dcdss.2010.3.185. MR2610558
[3] R. Beals, Analysis: An introduction, Cambridge University Press, Cambridge, 2004.
MR2098699
[4] W. Craig and J. Goodman, Linear dispersive equations of Airy type, J. Differential Equations
87 (1990), no. 1, 38–61, DOI 10.1016/0022-0396(90)90014-G. MR1070026
[5] H. Dym and H. P. McKean, Fourier series and integrals, Probability and Mathematical
Statistics, vol. 14, Academic Press, New York-London, 1972. MR0442564
[6] L. C. Evans, Partial differential equations, 2nd ed., Graduate Studies in Mathematics, vol. 19,
American Mathematical Society, Providence, RI, 2010. MR2597943
[7] G. B. Folland, Introduction to partial differential equations, 2nd ed., Princeton University
Press, Princeton, NJ, 1995. MR1357411
[8] P. R. Garabedian, Partial differential equations, John Wiley & Sons, Inc., New York-London-
Sydney, 1964. MR0162045
[9] F. John, Partial differential equations, 4th ed., Applied Mathematical Sciences, vol. 1,
Springer-Verlag, New York, 1982. MR831655
[10] T. Kato, The Cauchy problem for quasi-linear symmetric hyperbolic systems, Arch. Rational
Mech. Anal. 58 (1975), no. 3, 181–205, DOI 10.1007/BF00280740. MR0390516
[11] W. Thomson, On ship waves, Institution of Mechanical Engineers, Proceedings 38 (1887),
409–434.
[12] S. Klainerman, Global existence for nonlinear wave equations, Comm. Pure Appl. Math. 33
(1980), no. 1, 43–101, DOI 10.1002/cpa.3160330104. MR544044
[13] S. Klainerman and P. Sarnak, Explicit solutions of cmu = 0 on the Friedmann-Robertson-
Walker space-times, Ann. Inst. H. Poincaré Sect. A (N.S.) 35 (1981), no. 4, 253–257 (1982).
MR650132
[14] P. Lax, Hyperbolic Systems of Conservation Laws and the Mathematical Theory of Shock
Waves, CBMS-NSF Regional Conference Series in Applied Mathematics, Society for Indus-
trial and Applied Mathematics (1973). ISBN: 978-0-89871-177-6.
201
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
202 Bibliography
[15] P.-L. Lions, Mathematical topics in fluid mechanics. Vol. 2. Compressible Models, Oxford
Lecture Series in Mathematics and its Applications, vol. 10, Oxford Science Publications, The
Clarendon Press, Oxford University Press, New York, 1998. MR1637634
[16] R. McOwen, Partial Differential Equations: Methods and Applications, Second Edition, Pear-
son (2002). ISBN-10: 0130093351.
[17] L. Nirenberg, Lectures on linear partial differential equations, Conference Board of the Mathe-
matical Sciences Regional Conference Series in Mathematics, vol. 17, American Mathematical
Society, Providence, R.I., 1973. MR0450755
[18] J. Rauch, Partial differential equations, Graduate Texts in Mathematics, vol. 128, Springer-
Verlag, New York, 1991. MR1223093
[19] M. E. Taylor, Pseudodifferential operators, Princeton Mathematical Series, vol. 34, Princeton
University Press, Princeton, N.J., 1981. MR618463
[20] A. N. Tikhonov and V. Y. Arsenin, Solutions of ill-posed problems, Scripta Series in Math-
ematics, V. H. Winston & Sons, Washington, D.C.: John Wiley & Sons, New York-Toronto,
Ont.-London, 1977. Translated from the Russian. MR0455365
[21] A. Vasy, Partial differential equations: An accessible route through theory and applications,
Graduate Studies in Mathematics, vol. 169, American Mathematical Society, Providence, RI,
2015. MR3410751
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Index
Cauchy–Riemann equations, 3
Féjer kernel, 117
Cesaro integrability, 116
first-order quasilinear equation, 178
complete orthonormal basis, 26
arrow of time, 185
conformal mapping, 2, 8, 99
Burger’s equation, 178
conservation law, 15, 34
characteristic curves, 179
constituitive law, 31
characteristic equations, 179
convex dual, 187
convex duality, 187 conservation laws, 178
convolution, 46 entropy condition, 185
convolution algebra, 114 Hamilton–Jacobi equation, 188
jump condition, 182
d’Alembert’s formula, 17 Lax–Olenik solutions, 186
Darboux equation, 127 Lax–Olenik variational principle, 189
diffusion equation, 56 Lax–Olenik weak solutions, 190
Dirac δ-function, 49, 109, 110 noncharacteristic Cauchy surface, 180
Dirichlet–Neumann operator, 86, 88, 94 Riemann problem, 183
dispersion, 151 shock curves, 192
dispersion relation, 160 shocks, 181
dispersive equation, 34 uniqueness, 194
distributions, 110 weak solutions, 181
double layer potential, 85, 87 flow of a dynamical system, 200
Duhamel’s principle, 19 fluid dynamics, 196
heat equation, 53 Fourier multiplier, 154
wave equation, 19 Fourier transform, 10, 105
Fourier inversion formula, 10
Einstein’s equations, 6 isometry, 105
203
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
204 Index
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Index 205
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.
Does entropy really increase no matter what we do? Can light
pass through a Big Bang? What is certain about the Heisenberg
uncertainty principle? Many laws of physics are formulated in
terms of differential equations, and the questions above are about
GSM/197
www.ams.org
Licensed to Univ Nac Autonoma de Mexico. Prepared on Sun Jul 16 20:20:10 EDT 2023for download from IP 132.248.9.34.