2017 Book FourierSeriesFourierTransformA PDF
2017 Book FourierSeriesFourierTransformA PDF
Valery Serov
Fourier Series,
Fourier Transform and
Their Applications to
Mathematical Physics
Applied Mathematical Sciences
Volume 197
Editors
S.S. Antman, Institute for Physical Science and Technology, University of Maryland, College Park,
MD, USA
[email protected]
Leslie Greengard, Courant Institute of Mathematical Sciences, New York University, New York,
NY, USA
[email protected]
P.J. Holmes, Department of Mechanical and Aerospace Engineering, Princeton University, Princeton,
NJ, USA
[email protected]
Advisors
J. Bell, Lawrence Berkeley National Lab, Center for Computational Sciences and Engineering, Berkeley,
CA, USA
P. Constantin, Department of Mathematics, Princeton University, Princeton, NJ, USA
R. Durrett, Department of Mathematics, Duke University, Durham, NC, USA
R. Kohn, Courant Institute of Mathematical Sciences, New York University, New York, NY, USA
R. Pego, Department of Mathematical Sciences, Carnegie Mellon University, Pittsburgh, PA, USA
L. Ryzhik, Department of Mathematics, Stanford University, Stanford, CA, USA
A. Singer, Department of Mathematics, Princeton University, Princeton, NJ, USA
A. Stevens, Department of Applied Mathematics, University of Münster, Münster, Germany
A. Stuart, California Institute of Technology, Pasadena, CA, USA
S. Wright, Computer Sciences Department, University of Wisconsin, Madison, WI, USA
Founding Editors
Fritz John, Joseph P. LaSalle, Lawrence Sirovich
More information about this series at http://www.springer.com/series/34
Valery Serov
Mathematics Subject Classification (2010): 26A16, 26A45, 35A08, 35F50, 35J05, 35J08, 35J10, 35J15,
35K05, 35K08, 35L05, 35P25, 35R30
The modern theory of analysis and differential equations in general certainly in-
cludes the Fourier transform, Fourier series, integral operators, spectral theory of
differential operators, harmonic analysis and much more. This book combines all
these subjects based on a unified approach that uses modern view on all these
themes. The book consists of four parts: Fourier series and the discrete Fourier
transform, Fourier transform and distributions, Operator theory and integral equa-
tions and Introduction to partial differential equations and it outgrew from the half-
semester courses of the same name given by the author at University of Oulu, Fin-
land during 2005–2015.
Each part forms a self-contained text (although they are linked by a common
approach) and can be read independently. The book is designed to be a modern
introduction to qualitative methods used in harmonic analysis and partial differential
equations (PDEs). It can be noted that a survey of the state of the art for all parts of
this book can be found in a very recent and fundamental work of B. Simon [35].
This book contains about 250 exercises that are an integral part of the text. Each
part contains its own collection of exercises with own numeration. They are not only
an integral part of the book, but also indispensable for the understanding of all parts
whose collection is the content of this book. It can be expected that a careful reader
will complete all these exercises.
This book is intended for graduate level students majoring in pure and applied
mathematics but even an advanced researcher can find here very useful information
which previously could only be detected in scientific articles or monographs.
Each part of the book begins with its own introduction which contains the facts
(mostly) from functional analysis used thereinafter. Some of them are proved while
the others are not.
The first part, Fourier series and the discrete Fourier transform, is devoted to
the classical one-dimensional trigonometric Fourier series with some applications
to PDEs and signal processing. This part provides a self-contained treatment of all
well known results (but not only) at the beginning graduate level. Compared with
some known texts (see [12, 18, 29, 35, 38, 44, 45]) this part uses many function
spaces such as Sobolev, Besov, Nikol’skii and Hölder spaces. All these spaces are
v
vi Preface
introduced by special manner via the Fourier coefficients and they are used in the
proofs of main results. Same definition of Sobolev spaces can be found in [35]. The
advantage of such approach is that we are able to prove quite easily the precise em-
beddings for these spaces that are the same as in classical function theory (see [1, 3,
26, 42]). In the frame of this part some very delicate properties of the trigonometric
Fourier series (Chapter 10) are considered using quite elementary proofs (see also
[46]). The unified approach allows us also to consider naturally the discrete Fourier
transform and establish its deep connections with the continuous Fourier transform.
As a consequence we prove the famous Whittaker-Shannon-Boas theorem about the
reconstruction of band-limited signal via the trigonometric Fourier series (see Chap-
ter 13). Many applications of the trigonometric Fourier series to the one-dimensional
heat, wave and Laplace equation are presented in Chapter 14. It is accompanied by a
large number of very useful exercises and examples with applications in PDEs (see
also [10, 17]).
The second part, Fourier transform and distributions, probably takes a central role
in this book and it is concerned with distribution theory of L. Schwartz and its ap-
plications to the Schrödinger and magnetic Schrödinger operators (see Chapter 32).
The estimates for Laplacian and Hamiltonian that generalize well known Agmon’s
estimates on the continuous spectrum are presented in this part (see Chapter 23).
This part can be considered as one of the most important because of numerous ap-
plications in the scattering theory and inverse problems. Here we have considered
for the first time some classical direct scattering problems for the Schrödinger op-
erator and for the magnetic Schrödinger operator with singular (locally unbounded)
coefficients including the mathematical foundations of the classical approximation
of M. Born. Also, the properties of Riesz transform and Riesz potentials (see Chap-
ter 21) are investigated very carefully in this part. Before this material could only be
found in scientific journals or monographs but not in textbooks. There is a good con-
nection of this part with Operator theory and integral equations. The main technique
applied here is the Fourier transform.
The third part, Operator theory and integral equations, is devoted mostly to the
self-adjoint but unbounded operators in Hilbert spaces and their applications to in-
tegral equations in such spaces. The advantage of this part is that many important
results of J. von Neumann’s theory of symmetric operators are collected together.
J. von Neumann’s spectral theorem allows us, for example, to introduce the heat
kernel without solving the heat equation. Moreover, we show applications of the
spectral theorem of J. von Neumann (for these operators) to the spectral theory of
elliptic differential operators. In particular, the existence of Friedrichs extension for
these operators with discrete spectrum is provided. Special attention is devoted to
the Schrödinger and the magnetic Schrödinger operators. The famous diamagnetic
inequality is proved here. We follow in this consideration B. Simon [35] (slightly
different approach can be found in [28]). We recommend (in addition to this part)
the reader get acquainted with the books [4, 13, 15, 24, 41]. As a consequence of
the spectral theory of elliptic differential operators the integral equations with weak
singularities are considered in quite simple manner not only in Hilbert spaces but
also in some Banach spaces, e.g. in the space of continuous functions on closed
Preface vii
manifolds. The central point of this consideration is the Riesz theory of compact
(not necessarily self-adjoint) operators in Hilbert and Banach spaces. In order to
keep this part short, some proofs will not be given, nor will all theorems be proved
in complete generality. For many details of these integral equations we recommend
[22]. We are able to investigate in quite simple manner one-dimensional Volterra in-
tegral equations with weak singularities in L∞ (a, b) and singular integral equations
in the periodic Hölder spaces Cα [−a, a]. Concerning approximation methods our
considerations use the general theory of bounded or compact operators in Hilbert
spaces and we follow mostly the monograph of Kress [22].
The fourth part, Introduction to partial differential equations, serves as an in-
troduction to modern methods for classical theory of partial differential equations.
Fourier series and Fourier transform play crucial role here too. An important (and
quite independent) segment of this part is the self-contained theory of quasi-linear
partial differential equations of order one. The main attention in this part is devoted
to elliptic boundary value problems in Sobolev and Hölder spaces. In particular, the
unique solvability of direct scattering problem for Helmholtz equation is provided.
We investigate very carefully the mapping and discontinuity properties of double
and single layer potentials with continuous densities. We also refer to similar prop-
erties of double and single layer potentials with densities in Sobolev spaces H 1/2 (S)
and H −1/2 (S), respectively, but will not prove any of these results, referring for their
proofs to monographs [22] and [25]. Here (and elsewhere in the book) S denotes the
boundary of a bounded domain in Rn and if the smoothness of S is not specified
explicitly then it is assumed to be such that Sobolev embedding theorem holds.
Compared with well known texts on partial differential equations some direct and
inverse scattering problems for Helmholtz, Schrödinger and magnetic Schrödinger
operators are considered in this part. As it was mentioned earlier this type of mater-
ial could not be found in textbooks. The presentation in many places of this part has
been strongly influenced by the monographs [6, 7, 11] (see also [8, 16, 24, 36, 40]).
In closing we note that this book is not as comprehensive as the fundamental
work of B. Simon [35]. But the book can be considered as a good introduction to
modern theory of analysis and differential equations and might be useful not only
to students and PhD students but also to all researchers who have applications in
mathematical physics and engineering sciences. This book could not have appeared
without the strong participation, both in content and typesetting, of my colleague
Adj. Prof. Markus Harju. Finally, a special thanks to professor David Colton from
University of Delaware (USA) who encouraged the writing of this book and who
has supported the author very much over the years.
ix
x Contents
Definition 1.1. A function f (x) of one variable x is said to be periodic with period
T > 0 if the domain D( f ) of f contains x + T whenever it contains x and if for every
x ∈ D( f ), one has
f (x + T ) = f (x). (1.1)
f (x − T ) = f (x).
It follows that if T is a period of f , then mT is also a period for every integer m > 0.
The smallest value of T > 0 for which (1.1) holds is called the fundamental period
of f .
mπ x mπ x mπ x
sin , cos , ei L , m = 1, 2, . . .
L L
are periodic with fundamental period T = 2L m . Note also that they are periodic with
common period 2L.
If some function f is defined on the interval [a, a + T ], with T > 0 and f (a) =
f (a + T ), then f can be extended periodically with period T to the whole line as
Therefore, we may assume from now on that every periodic function is defined on
the whole line.
The set of all such functions is denoted by L p (a, b). When p = 1, we say that f is
integrable.
The following “continuity” in the sense of L p space, 1 ≤ p < ∞, holds: for every
f ∈ L p (a, b) and ε > 0, there is a continuous function g on [a, b] such that
b
1/p
| f (x) − g(x)| dx p
<ε
a
(see e.g., Corollary 5.3). If f is p-integrable and g is p -integrable on [a, b], where
1 1
+ = 1, 1 < p < ∞, 1 < p < ∞ ,
p p
b b
1/p b
1/p
| f (x)g(x)|dx ≤ | f (x)| p dx |g(x)| p dx .
a a a
This inequality is called Hölder’s inequality for integrals. Fubini’s theorem states
that
b d d b b d
F(x, y)dy dx = F(x, y)dx dy = F(x, y)dxdy,
a c c a a c
Exercise 1.1. Prove Hölder’s inequality for integrals for every 1 ≤ p < ∞.
where 1 < p < ∞, 1/p + 1/p = 1, and where for p = ∞ (or p = ∞) we consider
max1≤ j≤n |a j | (or max1≤ j≤n |b j |) instead of the corresponding sums.
Lemma 1.3. If f is periodic with period T > 0 and if it is integrable on every finite
interval, then
a+T T
f (x)dx = f (x)dx (1.4)
a 0
for every a ∈ R.
The difference in the square brackets is equal to zero due to periodicity of f . Thus,
(1.4) holds for a > 0.
If a < 0, then we proceed similarly, obtaining
a+T 0 a+T
f (x)dx = f (x)dx + f (x)dx
a a 0
0 T T
= f (x)dx + f (x)dx − f (x)dx
a 0 a+T
T 0 T
= f (x)dx + f (x)dx − f (x)dx .
0 a a+T
Again, the periodicity of f implies that the difference in brackets is zero. Thus the
lemma is proved.
Definition 1.4. Let us assume that the domain of f is symmetric with respect to
{0}, i.e., if x ∈ D( f ), then −x ∈ D( f ). A function f is called even if
f (−x) = f (x), x ∈ D( f ),
6 Part I: Fourier Series and the Discrete Fourier Transform
and odd if
f (−x) = − f (x), x ∈ D( f ).
Proof. Since
a a 0
f (x)dx = f (x)dx + f (x)dx,
−a 0 −a
Definition 1.6. The notation f (c ± 0) is used to denote the right and left limits
f (c ± 0) := lim f (x).
x→c±
The number
n
Vab ( f ) := sup
x ,x ,...,xn
∑ | f (x j ) − f (x j−1 )| (1.5)
0 1 j=1
1 Introduction 7
is called the total variation of f on the interval [a, b]. For every x ∈ [a, b] we can
also define Vax ( f ) by (1.5). The class of functions of bounded variation is denoted
by BV [a, b].
is continuous on the interval [0, 1] but is not of bounded variation on [0, 1].
(2) Show that every piecewise constant function on [a, b] is of bounded variation.
Remark 1.9. This exercise shows that C[a, b] and BV [a, b] are not included in each
other, i.e., they represent two different scales of functions.
If f is real-valued, then Exercise 1.4 implies that Vax ( f ) − f (x) is monotone increas-
ing in x. Indeed, for h > 0 we have that
Vax+h ( f ) − f (x + h) − (Vax ( f ) − f (x)) = Vax+h ( f ) −Vax ( f ) − ( f (x + h) − f (x))
= Vxx+h ( f ) − ( f (x + h) − f (x))
≥ Vxx+h ( f ) − | f (x + h) − f (x)| ≥ 0.
a
g(x)d f (x) = lim
Δ →0
∑ g(ξ j )( f (x j ) − f (x j−1 )),
j=1
with some constant C > 0. This inequality is called the Hölder condition with expo-
nent α .
Definition 1.11. A function f is said to belong to Sobolev space Wp1 (a, b),
1 ≤ p < ∞, if f ∈ L p (a, b) and there is g ∈ L p (a, b) such that
x
f (x) = g(t)dt +C (1.8)
a
Definition 1.12. A function f is said to belong to Sobolev space W∞1 (a, b) if there
is a bounded integrable function g such that
x
f (x) = g(t)dt +C (1.9)
a
Lemma 1.14. Suppose that f ∈ Wp1 (a, b), 1 ≤ p ≤ ∞. Then f is of bounded vari-
ation. Moreover, if p = 1, then f is also continuous, and if 1 < p ≤ ∞, then
f ∈ C1−1/p [a, b].
Proof. Let first p = 1. Then there is an integrable function g such that (1.8) holds
with some constant C. Hence for fixed x ∈ [a, b] with x + h ∈ [a, b] we have
x+h
f (x + h) − f (x) = g(t)dt.
x
It follows that
x+h
| f (x + h) − f (x)| = g(t)dt → 0, h → 0,
x
1 Introduction 9
since g is integrable. This proves the continuity of f . At the same time, for every
{x0 , x1 , . . . , xn } such that a = x0 < x1 < · · · < xn = b, we have
x xj b
n j
n n
∑ | f (x j ) − f (x j−1 )| = ∑ x g(t)dt ≤ ∑ |g(t)|dt = |g(t)|dt.
j=1 j=1 j−1 j=1 x j−1 a
Hence, Definition 1.8 is satisfied with constant c0 = ab |g(t)|dt, and f is of bounded
variation.
If 1 < p < ∞, then using Hölder’s inequality for integrals we obtain for h > 0 that
x+h x+h
1/p x+h
1/p
| f (x + h) − f (x)| ≤ |g(t)|dt ≤ dt |g(t)| dt
p
x x x
b
1/p
≤h 1−1/p
|g(t)| dt
p
,
a
Remark 1.15. Since every f ∈ Wp1 (a, b), 1 ≤ p ≤ ∞, is continuous, it follows that
the constant C in (1.8)–(1.9) is equal to f (a).
Definition 1.16. Two functions u and v are said to be orthogonal on [a, b] if the
product uv is integrable and
b
u(x)v(x)dx = 0,
a
form a mutually orthogonal set on the interval [−L, L] as well as on every interval
[a, a + 2L]. In fact,
L
mπ x nπ x 0, m = n,
cos cos dx = (1.10)
−L L L L, m = n,
L
mπ x nπ x
cos sin dx = 0, (1.11)
−L L L
10 Part I: Fourier Series and the Discrete Fourier Transform
L
mπ x nπ x 0, m = n,
sin sin dx = (1.12)
−L L L L, m = n,
and L L
mπ x mπ x
sin dx = cos dx = 0. (1.13)
−L L −L L
Proof. By Lemma 1.3, it is enough to prove the equalities (1.10), (1.11), (1.12), and
(1.13) only for integrals over [−L, L]. Let us derive, for example, (1.12). Using the
equality
1
sin α sin β = (cos(α − β ) − cos(α + β )) ,
2
we have for m = n that
L L
mπ x nπ x 1 (m − n)π x 1 L (m + n)π x
sin sin dx = cos dx − cos dx
−L L L 2 −L L 2 −L L
(m−n)π x L
(m+n)π x L
1 sin L 1 sin L
= − = 0.
2 (m−n)π 2 (m+n)π
L −L L −L
If m = n, we have
L L L
mπ x nπ x 1 1 2mπ x
sin sin dx = 1dx − cos dx = L.
−L L L 2 −L 2 −L L
The other identities can be proved in a similar manner and are left to the reader. The
lemma is proved.
nπ x
Remark 1.18. This lemma holds also for the functions ei L , n = 0, ±1, ±2, . . ., in
the form
L
nπ x m π x 0, n = m,
ei L e−i L dx =
−L 2L, n = m.
Chapter 2
Formulation of Fourier Series
This series consists of 2L-periodic functions. Thus, if the series (2.1) converges for
all x, then the function to which it converges will also be 2L-periodic. Let us denote
this limiting function by f (x), i.e.,
∞
a0 mπ x mπ x
f (x) := + ∑ am cos + bm sin . (2.2)
2 m=1 L L
for each fixed n = 1, 2, . . .. It follows from the orthogonality relations (1.10), (1.11),
and (1.13) that the only nonzero term on the right-hand side is the one for which
m = n in the first summation. Hence
L
1 nπ x
an = f (x) cos dx, n = 1, 2, . . . . (2.3)
L −L L
A similar expression for bn is obtained by multiplying (2.2) by sin nπL x and integrat-
ing termwise from −L to L. The result is
L
1 nπ x
bn = f (x) sin dx, n = 1, 2, . . . . (2.4)
L −L L
Definition 2.1. Let f be integrable (not necessarily periodic) on the interval [−L, L].
The Fourier series of f is the trigonometric series (2.1), where the coefficients
a0 , am and bm are given by (2.5), (2.3), and (2.4), respectively. In that case, we write
∞
a0 mπ x mπ x
f (x) ∼ + ∑ am cos + bm sin . (2.6)
2 m=1 L L
Remark 2.2. This definition does not imply that the series (2.6) converges to f or
that f is periodic.
Definition 2.1 and Lemma 1.5 imply that if f is even on [−L, L], then the Fourier
series of f has the form
∞
a0 mπ x
f (x) ∼ + ∑ am cos , (2.7)
2 m=1 L
The series (2.7) and (2.8) are called the Fourier cosine series and Fourier sine series,
respectively.
If L = π , then the Fourier series (2.6) ((2.7) and (2.8)) transforms to
∞
a0
f (x) ∼ + ∑ (am cos mx + bm sin mx) , (2.9)
2 m=1
where the coefficients a0 , am , and bm are given by (2.5), (2.3), and (2.4) with L = π .
There are different approaches if the function f is defined on an asymmetric
interval [0, L] with arbitrary L > 0.
2 Formulation of Fourier Series 13
Then g(x) is even and its Fourier (cosine) series (2.7) represents f on [0, L].
(2) Odd extension. Define a function h(x) on the interval [−L, L] as
f (x), 0 ≤ x ≤ L,
h(x) =
− f (−x), −L ≤ x < 0.
Then h(x) is odd, and its Fourier (sine) series (2.8) represents f on [0, L].
(3) Define a function f(t) on the interval [−π , π ] as
tL L
f(t) = f + .
2π 2
and L
2 2mπ x
bm ( f) = (−1)m f (x) sin dx = (−1)m bm ( f ).
L 0 L
Hence,
∞
a0
f(t) ∼ + ∑ (am cos mt + bm sin mt) ,
2 m=1
tL L
where a0 , am , and bm are the same and x = + .
2π 2
These three alternatives allow us to consider (for simplicity) only the case of a sym-
metric interval [−π , π ] such that the Fourier series will be of the form (2.9) i.e.
14 Part I: Fourier Series and the Discrete Fourier Transform
∞
a0
f (x) ∼ + ∑ (am cos mx + bm sin mx) .
2 m=1
Using Euler’s formula, we will rewrite this series in the complex form
∞
f (x) ∼ ∑ cn einx , (2.10)
n=−∞
Exercise 2.5. Find the Fourier series of f (x) = cos(x/2), |x| ≤ π . Using this series,
show that
∞
(−1)k+1
(1) π = 2 + ∑ ;
k=1 k − 1/4
2
∞
π (−1)k+1
(2) =∑ ;
4 k=1 2k − 1
∞
1 1
(3) =∑ 2 .
2 k=1 4k − 1
Exercise 2.6. Show that if N is odd, then sinN x can be written as a finite sum of the
form
N
∑ ak sin kx,
k=1
which means that this finite sum is the Fourier series of sinN x and the coefficients
ak (which are real) are the Fourier coefficients of sinN x.
Exercise 2.7. Show that if N is odd, then cosN x can be written as a finite sum of
the form
N
∑ ak cos kx.
k=1
Chapter 3
Fourier Coefficients and Their Properties
is said to
(1) converge pointwise if for each x ∈ [−π , π ] the limit
lim
N→∞
∑ cn einx
|n|≤N
exists,
(2) converge uniformly in x ∈ [−π , π ] if the limit
lim
N→∞
∑ cn einx
|n|≤N
exists uniformly,
(3) converge absolutely if the limit
lim
N→∞
∑ |cn |
|n|≤N
exists, or equivalently, if
∞
∑ |cn | < ∞.
n=−∞
These three different types of convergence appear frequently in the sequel, and they
are presented above from the weakest to the strongest. In other words, absolute
where the upper bound does not depend on n. Let us assume that a sequence
{cn }∞
n=−∞ is such that
∞
∑ |cn | < ∞.
n=−∞
for some integer k > 0. Then the series (3.2) defines a function that is k times differ-
entiable, with
∞
f (k) (x) = ∑ (in)k cn einx (3.3)
n=−∞
a continuous function. This follows from the fact that the series (3.3) converges
uniformly with respect to x ∈ [−π , π ].
Let us consider a useful example in which Fourier coefficients are applied. If
0 ≤ r < 1, then the geometric series gives
∞
1
= ∑
1 − reix n=0
rn einx , (3.4)
and this series converges absolutely. Using the definition of the Fourier coefficients,
we obtain
1 π e−inx
r =
n
dx, n = 0, 1, 2, . . . ,
2π −π 1 − reix
3 Fourier Coefficients and Their Properties 19
and π
1 einx
0= dx, n = 1, 2, . . . .
2π −π 1 − reix
and
∞
r sin x 1 i ∞
= Im = ∑ rn sin nx = − ∑ r|n| sgn(n)einx . (3.6)
1 − 2r cos x + r 2 1 − reix
n=1 2 n=−∞
and
∞
2r sin x
−i ∑ sgn(n)r|n| einx =
1 − 2r cos x + r2
=: Qr (x). (3.8)
n=−∞
Definition 3.2. The function Pr (x) is called the Poisson kernel, while Qr (x) is
called the conjugate Poisson kernel.
Exercise 3.2. Prove that both Pr (x) and Qr (x) are solutions of the Laplace equation
ux1 x1 + ux2 x2 = 0
in the disk x12 + x22 < 1, where x1 + ix2 = reix with 0 ≤ r < 1 and x ∈ [−π , π ].
20 Part I: Fourier Series and the Discrete Fourier Transform
Δn a = an − an−1 .
Then for any two sequences an and bn and integers M < N, the formula
N N
∑ ak Δk b = aN bN − aM bM − ∑ bk−1 Δk a (3.9)
k=M+1 k=M+1
1 − eix(n+1)
bn = , x = 0,
1 − eix
it follows that
2 1
|bn | ≤ =
|1 − e | | sin 2x |
ix
for x ∈ [−π , π ] \ {0}. Applying (3.9) shows that for M < N we have
N N k k−1 N
∑ ikx
ck e = ∑ ck ∑e − ∑e
ilx ilx
= ∑ ck Δk b
k=M+1 k=M+1 l=0 l=0 k=M+1
N
= cN bN − cM bM − ∑ bk−1 Δk c.
k=M+1
Thus
3 Fourier Coefficients and Their Properties 21
N N
ikx
∑ ck e ≤ cN |bN | + cM |bM | + ∑ |bk−1 ||Δk c|
k=M+1 k=M+1
N
1
≤ cN + cM + ∑ |ck − ck−1 |
| sin 2x | k=M+1
1 2cM
= x (cN + cM + cM − cN ) = →0
| sin 2 | | sin 2x |
En := ∑ |ck |, n = 0, 1, 2, . . . . (3.11)
|k|>n
There is a good connection between the modulus of continuity (1.7) and (3.11).
Indeed, if f (x) denotes the series (2.10) and h > 0, we have
∞
| f (x + h) − f (x)| ≤ ∑ |cn ||einh − 1| = ∑ |cn ||einh − 1| + ∑ |cn ||einh − 1|
n=−∞ |n|h≤1 |n|h>1
≤h ∑ |n||cn | + 2 ∑ |cn | =: I1 + I2 ,
|n|≤[1/h] |n|>[1/h]
where [x] denotes the entire part of x. If we denote [1/h] by Nh , then I2 = 2ENh and
Nh
I1 = h ∑ |n| E|n|−1 − E|n| = −h ∑ l (El − El−1 ) .
|n|≤Nh l=1
22 Part I: Fourier Series and the Discrete Fourier Transform
Nh
1 Nh −1
ωh ( f ) ≤ 2ENh − hNh ENh + h ∑ En−1 ≤ 2ENh + ∑ En . (3.12)
n=1 Nh n=0
Here and throughout, the notation A = O(B) on a set X means that |A| ≤ C|B| on X
with some constant C > 0. Similarly, A = o(B) means that A/B → 0.
Exercise 3.4. Prove the second relation in (3.13).
We summarize (3.13) as follows: if the tail of the trigonometric series (2.10) behaves
as O(n−α ) for some 0 < α < 1, then the function f to which it converges belongs
to the Hölder space Cα [−π , π ].
Chapter 4
Convolution and Parseval’s Equality
Then the function f to which it converges is continuous and periodic. If g(x) is any
continuous function, then the product f g is also continuous and hence integrable on
[−π , π ] and
π ∞ π ∞
1 1
2π −π
f (x)g(x)dx =
2π ∑ cn ( f )
−π
g(x)einx dx = ∑ cn ( f )c−n (g), (4.1)
n=−∞ n=−∞
where integration of the series term by term is justified by the uniform convergence
of the Fourier series. Putting g = f in (4.1) yields
π ∞ ∞
1
2π −π
| f (x)|2 dx = ∑ cn ( f )c−n ( f ) = ∑ |cn ( f )|2 (4.2)
n=−∞ n=−∞
by (2.13).
Definition 4.1. Equality (4.2) is called the Parseval’s equality for the trigonometric
Fourier series.
The formula (4.1) can be generalized as follows.
Exercise 4.1. Let a periodic function f be defined by the absolutely convergent
Fourier series (2.10) and let g be integrable and periodic. Prove that
π ∞
1
2π −π
f (x)g(y − x)dx = ∑ cn ( f )cn (g)einy
n=−∞
cn ( f1 ∗ f2 ) = cn ( f1 )cn ( f2 ),
Proof. We note first that by Fubini’s theorem the convolution (4.3) is well defined
as an L1 function. Indeed,
π π π π −y
| f1 (y)| · | f2 (x − y)|dy dx = | f1 (y)| | f2 (z)|dz dy
−π −π −π −π −y
π π
= | f1 (y)| | f2 (z)|dz dy
−π −π
by Lemma 1.3. The Fourier coefficients of the convolution (4.3) are equal to
1 π
cn ( f1 ∗ f2 ) = ( f1 ∗ f2 )(x)e−inx dx
2π −π
π π
1
= f 1 (y) f 2 (x − y)dy e−inx dx
(2π )2 −π −π
π π
1 −inx
= f 1 (y) f 2 (x − y)e dx dy
(2π )2 −π −π
π π −y
1 −in(y+z)
= f 1 (y) f 2 (z)e dz dy
(2π )2 −π −π −y
π π
1 −iny −inz
= f 1 (y)e f 2 (z)e dz dy = cn ( f1 )cn ( f2 )
(2π )2 −π −π
Exercise 4.2. Prove that if f1 and f2 are integrable and periodic, then their convo-
lution is symmetric and periodic.
and that Pr ∗ f satisfies the Laplace equation, i.e., (Pr ∗ f )x1 x1 + (Pr ∗ f )x2 x2 = 0,
where x12 + x22 = r2 < 1 with x1 + ix2 = reix .
Remark 4.3. We are going to prove in Chapter 10 that for every periodic continuous
function f , the limit limr→1− ( f ∗ Pr )(x) = f (x) exists uniformly in x.
Chapter 5
Fejér Means of Fourier Series. Uniqueness
of the Fourier Series.
Let us denote the partial sum of the Fourier series of f ∈ L1 (−π , π ) (not necessarily
periodic) by
SN ( f ) := ∑ cn ( f )einx
|n|≤N
= ∑ (N + 1 − |k|)ck ( f )eikx ,
|k|≤N
where
1, |k| ≤ N,
1[−N,N] (k) =
0, |k| > N.
Let us assume now that f is periodic. Then Exercise 4.1 and Theorem 4.2 lead to
π ∞
1 |k|
f (y)KN (x − y)dy = ∑ 1[−N,N] (k) 1 − ck ( f )eikx . (5.5)
2π −π k=−∞ N +1
The properties (5.3), (5.4), and (5.6) allow us to prove the following result.
Theorem 5.1. Let f ∈ L p (−π , π ) be periodic with 1 ≤ p < ∞. Then
π
1/p
lim |σN ( f )(x) − f (x)| p dx = 0. (5.7)
N→∞ −π
If, in addition, f has right and left limits f (x0 ± 0) at a point x0 ∈ [−π , π ], then
1
lim σN ( f )(x0 ) = ( f (x0 + 0) + f (x0 − 0)) . (5.8)
N→∞ 2
5 Fejér Means of Fourier Series. Uniqueness of the Fourier Series. 29
Proof. Let us first prove (5.7). Indeed, (5.4) and (5.6) give
π
1/p π
1/p
|σN ( f )(x) − f (x)| p dx = |( f ∗ KN )(x) − f (x)| p dx
−π −π
π 1 π p 1/p
1 π
= KN (y) f (x − y)dy − KN (y) f (x)dy dx
−π 2π −π 2π −π
π π p 1/p
1
= KN (y)( f (x − y) − f (x))dy dx
2π
−π −π
π p 1/p
1
≤ K (y)( f (x − y) − f (x))dy dx
2π
−π |y|<δ
N
π p 1/p
1
+ K (y)( f (x − y) − f (x))dy dx =: I1 + I2 .
2π
−π π ≥|y|>δ
N
π 1/p
1 p
I2 ≤ KN (y) | f (x − y) − f (x)| dx dy
2π π ≥|y|>δ −π
π 1/p
p 1
≤2 | f (x)| dx KN (y)dy, (5.10)
−π 2π π ≥|y|>δ
1 1 1 π2
KN (y) ≤ · 2y ≤ · 2
N + 1 sin 2 N +1 δ
30 Part I: Fourier Series and the Discrete Fourier Transform
and
1 π 2 2(π − δ ) π2 π2
KN (y)dy ≤ · < 2 < √ →0
2π π ≥|y|>δ 2πδ 2 N +1 δ (N + 1) N
1
σN ( f )(x0 ) − ( f (x0 + 0) + f (x0 − 0))
2
1 π 1
= KN (y) f (x0 − y) − ( f (x0 + 0) + f (x0 − 0)) dy
2π −π 2
π
1
= KN (y)g(y)dy, (5.11)
2π −π
where
1
g(y) = f (x0 − y) − ( f (x0 + 0) + f (x0 − 0)) .
2
Since the Fejér kernel KN (y) is even, we can rewrite the right-hand side of (5.11) as
π
1
KN (y)h(y)dy,
2π 0
where h(y) = g(y) + g(−y). It is clear that h(y) is an L1 function. But we have more,
namely,
lim h(y) = 0. (5.12)
y→0+
uniformly in x ∈ [−π , π ].
Proof. Since π
lim |σN ( f )(x) − f (x)| dx = 0
N→∞ −π
uniformly in x ∈ R. In particular, cn ( f ) → 0 as n → ∞.
by Lemma 1.3. Formula (6.2) shows that to prove (6.1) it is enough to show that the
Fourier coefficients cn ( f ) tend to zero as n → ∞. Indeed,
π π +π /n π
2π cn ( f ) = f (y)e−iny dy = f (y)e−iny dy = f (t + π /n)e−int e−iπ dt
−π −π +π /n −π
Write
cn ( f ) = cn (g) + cn ( f − g).
The first term tends to zero as n → ∞, since g is continuous, whereas the second
term is less than ε /(2π ). This implies that
ε
sup lim |cn ( f )| ≤ .
n→∞ 2π
This fact together with (6.2) gives (6.1). The theorem is thus proved.
Corollary 6.2. Let f be as in Theorem 6.1. If a periodic function g is continuous
on [−π , π ], then π
lim f (x + z)g(z)e−inz dz = 0
n→∞ −π
and
π π
lim f (x + z)g(z) sin(nz)dz = lim f (x + z)g(z) cos(nz)dz = 0
n→∞ −π n→∞ −π
uniformly in x ∈ [−π , π ].
Exercise 6.1. Prove this corollary.
Exercise 6.2. Show that if f satisfies the Hölder condition with exponent α ∈ (0, 1],
then cn ( f ) = O(|n|−α ) as n → ∞.
Exercise 6.3. Suppose that f satisfies the Hölder condition with exponent α > 1.
Prove that f ≡ constant.
Exercise 6.4. Let f (x) = |x|α , where −π ≤ x ≤ π and 0 < α < 1. Prove that cn ( f )
|n|−1−α as n → ∞.
Remark 6.3. The notation a b means that there exist 0 < c1 < c2 such that
Let us introduce for all 1 ≤ p < ∞ and periodic functions f ∈ L p (−π , π ) the L p -
modulus of continuity of f by
π
1/p
p
ω p,δ ( f ) := sup | f (x + h) − f (x)| dx .
|h|≤δ −π
6 The Riemann–Lebesgue Lemma 35
Exercise 6.5. Suppose that ω p,δ ( f ) ≤ Cδ α for some C > 0 and α > 1. Prove that
f is constant almost everywhere.
Hint. First show that ω p,2δ ( f ) ≤ 2ω p,δ ( f ); then iterate this to obtain a contradiction.
Suppose that f ∈ L1 (−π , π ) but f is not necessarily periodic. We can consider the
Fourier series corresponding to f , i.e.,
∞
f (x) ∼ ∑ cn einx ,
n=−∞
where the cn are the Fourier coefficients cn ( f ). The series on the right-hand side
is considered formally in the sense that we know nothing about its convergence.
However, the limit
π π
lim
N→∞ −π
∑ cn ( f )einx dx =
−π
f (x)dx (6.4)
|n|≤N
exists. Indeed,
π π π
∑
−π |n|≤N
cn ( f )einx dx = c0 ( f )
−π
dx + ∑ cn ( f )
−π
einx dx = 2π c0 ( f )
0<|n|≤N
π
= f (x)dx.
−π
Remark 6.4 (Important properties of the Fourier series). The existence of the limit
(6.4) shows us that we can always integrate the Fourier series of an L1 function term
by term.
Chapter 7
The Fourier Series of a Square-Integrable
Function. The Riesz–Fischer Theorem.
We can measure the degree of approximation by the (square of) mean square dis-
tance
1 π
| f (x) − g(x)|2 dx = ( f − g, f − g)L2 (−π ,π ) .
2π −π
or
π
1
| f (x)|2 dx
2π −π
π π
1 1
+
2π ∑
−π |n|≤N
bn einx ∑ bk e−ikx dx −
2π
2Re
−π
f (x) ∑ bn e−inx dx
|k|≤N |n|≤N
π
1 π 1
=
2π −π
| f (x)|2 dx +
2π ∑ |bn |2
−π
dx − 2Re ∑ bn cn ( f )
|n|≤N |n|≤N
π
1
=
2π −π
| f (x)|2 dx + ∑ |bn |2 − 2Re ∑ bn cn ( f ) + ∑ |cn ( f )|2 − ∑ |cn ( f )|2 .
|n|≤N |n|≤N |n|≤N |n|≤N
So
π π
1 1
2π −π
| f (x) − g(x)|2 dx =
2π −π
| f (x)|2 dx − ∑ |cn ( f )|2 + ∑ |bn − cn ( f )|2 .
|n|≤N |n|≤N
and in particular,
∞ π
1
∑ |cn ( f )|2 ≤
2π −π
| f (x)|2 dx. (7.2)
n=−∞
Theorem 7.1. For every periodic function f ∈ L2 (−π , π ) with period 2π , its
Fourier series converges in L2 (−π , π ), i.e.,
2
π
1
lim f (x) − ∑ cn ( f )einx dx = 0,
N→∞ 2π −π |n|≤N
holds.
7 The Fourier Series of a Square-Integrable Function 39
It remains to show that F(x) = f (x) almost everywhere. To do this, we compute the
Fourier coefficients cn (F) by writing
π π
2π cn (F) =
−π
F(x)e −inx
dx =
−π
F(x) − ∑ ck ( f )e
ikx
e−inx dx
|k|≤N
π
+ ∑ ck ( f )
−π
ei(k−n)x dx.
|k|≤N
If N > |n|, then the last sum is equal to 2π cn ( f ). Thus, by Hölder’s inequality,
π
2π |cn (F) − cn ( f )| ≤ F(x) − ∑ ck ( f )eikx dx
−π |k|≤N
⎛ 2 ⎞1/2
√ π
≤ 2π ⎝ F(x) − ∑ ck ( f )eikx dx⎠ → 0
−π |k|≤N
Theorem 7.3. Suppose that f ∈ L2 (−π , π ) is periodic with period 2π and that its
Fourier coefficients satisfy
∞
∑ |n|2 |cn ( f )|2 < ∞. (7.4)
n=−∞
40 Part I: Fourier Series and the Discrete Fourier Transform
Then f ∈ W21 (−π , π ) with the Fourier series for f (x) given by
∞
f (x) ∼ ∑ incn ( f )einx .
n=−∞
Proof. Since (7.4) holds, it follows that by the Riesz–Fischer theorem there is a
unique function g(x) ∈ L2 (−π , π ) such that
∞
g(x) ∼ ∑ incn ( f )einx .
n=−∞
On the other hand, cn (g) = incn ( f ). Thus, by the uniqueness of Fourier series, we
obtain that F(x)− f (x) = constant almost everywhere, or f (x) = g(x) almost every-
where. This means that
x
f (x) = g(t)dt + constant,
−π
where we have used the basic inequality 2ab ≤ a2 + b2 for real numbers a and b.
Using Parseval’s equality (7.3), we can obtain for every periodic function f ∈
L2 (−π , π ) and N = 1, 2, . . . that
2
π
1
2π
∑
−π |n|≤N
cn ( f )e
inx
− f (x) dx = ∑ |cn ( f )|2 .
(7.5)
|n|>N
if M|h| ≤ 1. If (7.7) holds, then the second sum is O(M −2α ). To estimate the first
sum we use summation by parts. Writing
42 Part I: Fourier Series and the Discrete Fourier Transform
In := ∑ |ck ( f )|2 ,
|k|≤n
we have
M
∑ n2 |cn ( f )|2 = M 2 IM − 0 · I0 − ∑ In−1 (n2 − (n − 1)2 )
1≤|n|≤M n=1
M
= M 2 IM − ∑ (2n − 1)In−1 − I0 .
n=2
By hypothesis,
I∞ − In = O(n−2α ), n = 1, 2, . . . .
Thus,
M
M 2 IM − ∑ (2n − 1) I∞ + O((n − 1)−2α )
n=2
M M
= M 2 I∞ + O(M −2α ) − I∞ ∑ (2n − 1) − ∑ (2n − 1)O((n − 1)−2α )
n=2 n=2
M M
= O(M 2−2α ) + I∞ M 2 − ∑ (2n − 1) − ∑ (2n − 1)O((n − 1)−2α )
n=2 n=2
=1
= O(M 2−2α ) + O(M 2−2α ) = O(M 2−2α ).
n=1 (2n − 1) = M ;
(1) ∑M 2
(2) ∑M
n=2 O((2n − 1)(n − 1)−2α ) = O(M 2−2α ) for 0 < α < 1.
Combining these two estimates, we may conclude from (7.9) that there exists C > 0
such that
1 π
| f (x + h) − f (x)|2 dx ≤ C h2 M 2−2α + M −2α .
2π −π
if |h| ≤ 1/2. Here {1/|h|} denotes the fractional part of 1/|h|, i.e., {1/|h|} = 1/|h| −
[1/|h|] ∈ [0, 1).
Conversely, if (7.8) holds, then
∞
∑ (1 − cos(nh))|cn ( f )|2 ≤ Ch2α
n=−∞
with some C > 0. Integrating this inequality with respect to h over the interval [0, l],
l > 0, we have
∞ l l
∑ |cn ( f )|2
0
(1 − cos(nh))dh ≤ C
0
h2α dh,
n=−∞
or
∞
sin(nl)
∑ |cn ( f )| 2
l−
n
≤ Cl 2α +1 ,
n=−∞
or
∞
sin(nl)
∑ |cn ( f )| 2
1−
nl
≤ Cl 2α .
n=−∞
It follows that
sin(nl) 1
Cl 2α
≥ ∑ |cn ( f )| 2
1−
nl
≥ ∑ |cn ( f )|2 .
2 |n|l≥2
(7.10)
|n|l≥2
Exercise 7.3. Suppose that a periodic function f ∈ L2 (−π , π ) satisfies the condi-
tion π
| f (x + h) − f (x)|2 dx ≤ Ch2
−π
For an integrable function f periodic on the interval [−π , π ] let us introduce the
mapping
f
→ {cn ( f )}∞
n=−∞ ,
W2α (−π , π )
for some α ≥ 0 if
∞
∑ |n|2α |cn ( f )|2 < ∞, 00 := 1.
n=−∞
Bα2,θ (−π , π )
H2α (−π , π )
for some α ≥ 0 if
sup ∑
j=0,1,2,... 2 j ≤|n|<2 j+1
|n|2α |cn ( f )|2 < ∞.
Remark 8.5. We shall use later the following sufficient condition (see (3.13)): if
there is a constant C > 0 such that for each n = 1, 2, . . . we have
with some integer k ≥ 0 and some 0 < α < 1, then f belongs to the Hölder space
Ck+α [−π , π ].
and
Cα [−π , π ] ⊂ H2α (−π , π ).
or
∞
4 ∑ |n|2k−2 |cn ( f )|2 sin2 (nh/2) ≤ Ch2 .
n=−∞
It follows that
∑ |n|2k−2 |cn ( f )|2 n2 h2 ≤ Ch2
|nh|≤2
or
∑ |n|2k |cn ( f )|2 ≤ C.
|n|≤2/|h|
48 Part I: Fourier Series and the Discrete Fourier Transform
Letting h → 0 yields
∞
∑ |n|2k |cn ( f )|2 ≤ C
n=−∞
i.e., f ∈ W2k (−π , π ). For noninteger α one needs to interpolate between the spaces
C[α ] [−π , π ] and C[α ]+1 [−π , π ].
Now let us consider the second claim. As above, for f ∈ Cα [−π , π ], we have
π
1
| f (k) (x + h) − f (k) (x)|2 dx ≤ C|h|2(α −k) ,
2π −π
or
∞
2 ∑ |n|2k |cn ( f )|2 (1 − cos(nh)) ≤ C|h|2(α −k) .
n=−∞
It suffices to consider h > 0. If we integrate the last inequality with respect to h > 0
from 0 to l, then
∞
sin(nl)
∑ |n|2k
|cn ( f )|2
1 −
nl
≤ Cl 2(α −k) .
n=−∞
It follows that
∑ |n|2k |cn ( f )|2 ≤ Cl 2(α −k)
2≤|n|l≤4
or equivalently,
∑ |n|2k l 2(k−α ) |cn ( f )|2 ≤ C,
2≤|n|l≤4
where the constant C > 0 does not depend on l. Since 2(k − α ) < 0, it follows that
Theorem 8.7. Assume that α > 1/2 and that α − 1/2 is not an integer. Then
for each j = 0, 1, 2, . . .. Let us estimate the tail (8.2). Indeed, by the Cauchy–
Bunyakovsky–Schwarz inequality and (8.3),
∞
∑ |m|k−1 |cm ( f )| ≤ ∑ ∑ |m|k−1 |cm ( f )|
|m|≥n j= j0 2 j ≤|m|<2 j+1
2 j0 ∼n
⎛ ⎞1/2 ⎛ ⎞1/2
∞
⎝ 1 ⎠
≤ ∑ ∑ |m|2k |cm ( f )|2 ⎠ ⎝ ∑ 2
j= j0 2 j ≤|m|<2 j+1 2 j ≤|m|<2 j+1
m
2 j0 ∼n
⎛ ⎞1/2
√ ∞
1 √ ∞ j j0
≤ C ∑ ⎝ ∑ m2 ⎠ ≤ C ∑ 2− 2 ≤ C2− 2 .
j= j0 2 j ≤|m|<2 j+1 j= j0
2 j0 ∼n 2 j0 ∼n
where the constant C > 0 is independent of n. This means that (see (8.2)) f belongs
to Ck−1+1/2 [−π , π ] = Ck−1/2 [−π , π ].
If α > 1/2 is not an integer and α −1/2 is not an integer, then for f ∈ H2α (−π , π )
we have instead of (8.3) the estimate
⎛ ⎞1 ⎛ ⎞1
2 2
∞
∑ |cm ( f )| ≤ ∑ ⎝ ∑ 2α
|m| |cm ( f )| 2⎠ ⎝ ∑ −2α ⎠
|m|
|m|≥n j= j0 2 j ≤|m|<2 j+1 2 j ≤|m|<2 j+1
2 j0 ∼n
⎛ ⎞1/2
∞ ∞
≤C ∑ ⎝ ∑ |m|−2α ⎠ ≤C ∑ 2−(α −1/2) j ≤ Cn−(α −1/2) ,
j= j0 2 j ≤|m|<2 j+1 j= j0
2 j0 ∼n 2 j0 ∼n
∑ |m|k−1 |cm ( f )| ≤
|m|≥n
⎛ ⎞1/2 ⎛ ⎞1/2
∞
∑ ⎝ ∑ |m| 2k+2{α }
|cm ( f )|2⎠ ⎝ ∑ |m| −2−2{α } ⎠
j= j0 2 j ≤|m|<2 j+1 2 j ≤|m|<2 j+1
2 j0 ∼n
∞
≤C ∑ 2− j− j{α } 2 j/2 ≤ Cn−1/2−{α } .
j= j0
2 j0 ∼n
This means again that f ∈ Ck−1+1/2+{α } [−π , π ] = Cα −1/2 [−π , π ]. In the second
case, 1/2 < {α } < 1, we proceed as follows:
∑ |m|k |cm ( f )| ≤
|m|≥n
⎛ ⎞1/2 ⎛ ⎞1/2
∞
∑ ⎝ ∑ |m|2k+2{α } |cm ( f )|2 ⎠ ⎝ ∑ |m|−2{α } ⎠
j= j0 2 j ≤|m|<2 j+1 2 j ≤|m|<2 j+1
2 j0 ∼n
∞
≤C ∑ 2−({α }−1/2) j ≤ Cn−{α }+1/2 .
j= j0
2 j0 ∼n
This means that f ∈ Ck+{α }−1/2 [−π , π ] = Cα −1/2 [−π , π ]. Hence, the theorem is
completely proved.
Corollary 8.9. Assume that α > 1/2 and α − 1/2 is not an integer. Then
1
cn ( f ) = , n = 0, c0 ( f ) = 1
|n| log(1 + |n|)
1/2
defines a function from the Besov space B2,θ (−π , π ) for every 1 < θ < ∞, but not
for θ = 1.
Exercise 8.4. Prove that the Fourier series (8.1) with coefficients
1
cn ( f ) = , n = 0, c0 ( f ) = 1
|n|3/2 log(1 + |n|)
defines a function from the Besov space B12,θ (−π , π ) for every 1 < θ < ∞, but not
for θ = 1.
Exercise 8.5. Consider the Fourier series (8.1) with coefficients
1
cn ( f ) = , n = 0, c0 ( f ) = 1.
|n|2 logβ (1 + |n|)
Prove that
3/2
(1) f ∈ H2 (−π , π ) if β ≥ 0
3/2
(2) f ∈ W2 (−π , π ) if β > 1/2
(3) f ∈ C1 [−π , π ] if β > 1 but f ∈
/ C1 [−π , π ] if β ≤ 1.
So the embeddings W2α (−π , π ) ⊂ Cα −1/2 [−π , π ] and H2α (−π , π ) ⊂ Cα −1/2 [−π , π ]
are not valid for α − 1/2 an integer (see Theorem 8.6).
Exercise 8.6. Let
∞
∑ ak cos(bk x)
k=0
be a trigonometric series, where b = 2, 3, . . . and 0 < a < 1. Prove that the series
defines a function from C1 [−π , π ] if 0 < ab < 1 and a function from the Hölder
space Cγ [−π , π ], γ < 1 if ab = 1.
Exercise 8.7. Assume that a = 1/b2 in Exercise 8.6. Is it true that this function
belongs to C1 [−π , π ]?
Chapter 9
Absolute Convergence. Bernstein and Peetre
Theorems.
for some constant K > 0 and for all h = 0 sufficiently small. Indeed, due to Parseval’s
equality we have
π ∞
1
2π −π
| f (x + h) − f (x)|2 dx = ∑ |cn ( f )|2 |einh − 1|2
n=−∞
≤ ∑j |n|2 |h|2 |cn ( f )|2 + 4 ∑j |cn ( f )|2 , (9.2)
|n|≤2 0 |n|≥2 0
j0
|h|2 ∑j |n|2 |cn ( f )|2 ≤ |h|2 ∑ ∑ |n|2 |cn ( f )|2
|n|≤2 0 j=0 2 j ≤|n|<2 j+1
j0 2−2α
≤ |h|2 ∑ 2 j+1 ∑ |n|2α |cn ( f )|2
j=0 2 j ≤|n|<2 j+1
j +2
j0 22−2α 0 − 22−2α
≤ C|h| 2
∑ 2 2−2α j+1
= C|h| 2
22−2α − 1
j=0
2−2α
2−2α 1
≤ C|h|2 2 j0 ≤ C|h|2
= C|h|2α
|h|
if 0 < α < 1. We used this condition for α because we considered a geometric sum
with common ratio 22−2α = 1. The second sum on the right-hand side of (9.2) is
estimated from above as
∞ ∞
4 ∑ ∑ |n|2α |n|−2α |cn ( f )|2 ≤ 4 ∑ 2−2α j ∑ |n|2α |cn ( f )|2
j= j0 2 j ≤|n|<2 j+1 j= j0 2 j ≤|n|<2 j+1
∞
≤C ∑ 2−2α j ≤ C2−2 j0 α ≤ C|h|2α ,
j= j0
since 1
|h| ≤ 2 j0 +1 and the criterion of Definition 8.3 is satisfied. Thus, (9.1) is proved.
Conversely, if the L2 Hölder condition (9.1) is fulfilled, then Theorem 7.5 implies
for each N = 1, 2, . . . the inequality
N 2α ∑ |cn ( f )|2 ≤ C,
N≤|n|<2N
where the constant C is independent of N. Thus, we obtain for every integer N > 0
that
∑ |n|2α |cn ( f )|2 ≤ C.
N≤|n|<2N
Since N is arbitrary, we may conclude that f ∈ H2α (−π , π ) for 0 < α ≤ 1 in the
sense of Definition 8.3. Therefore, the L2 Hölder condition (9.1) can be considered
as an equivalent definition of the Nikol’skii space H2α (−π , π ) for 0 < α < 1.
Exercise 9.1. Prove that f belongs to the Nikol’skii space H2α (−π , π ) for every
noninteger α > 0 in the sense of Definition 8.3 if and only if the following L2 Hölder
condition holds:
π
1
| f (k) (x + h) − f (k) (x)|2 dx ≤ K|h|2α −2k
2π −π
∞
∑ |cn ( f )| < ∞.
n=−∞
Proof. Since the L2 Hölder condition (9.1) is equivalent to f ∈ H2α (−π , π ) for
0 < α < 1, there is a constant C > 0 such that
Corollary 9.2. Theorem 9.1 holds for Cα [−π , π ], Bα2,θ (−π , π ), and H2α (−π , π ) for
every α > 1/2 and 1 ≤ θ < ∞.
⎛ ⎞1/2
∞
∑⎝ ∑ |n||cn ( f )|2 ⎠ < ∞. (9.3)
j=0 2 j ≤|n|<2 j+1
Hence we have
∞ ∞
∑ |cn ( f )| = |c0 ( f )| + ∑ ∑ |n|1/2 |cn ( f )||n|−1/2
n=−∞ j=0 2 j ≤|n|<2 j+1
⎛ ⎞1/2 ⎛ ⎞1/2
∞
≤ |c0 ( f )| + ∑ ⎝ ∑ |n||cn ( f )|2 ⎠ ⎝ ∑ |n|−1 ⎠
j=0 2 j ≤|n|<2 j+1 2 j ≤|n|<2 j+1
56 Part I: Fourier Series and the Discrete Fourier Transform
⎛ ⎞1/2
∞ 1/2
≤ |c0 ( f )| + ∑ ⎝ ∑ |n||cn ( f )|2 ⎠ 2− j 2 j
j=0 2 j ≤|n|<2 j+1
⎛ ⎞1/2
∞
= |c0 ( f )| + ∑ ⎝ ∑ |n||cn ( f )|2 ⎠ <∞
j=0 2 j ≤|n|<2 j+1
Exercise 9.5. Prove that H2α (−π , π ) ⊂ B2,1 (−π , π ) if α > 1/2.
1/2
Theorem 9.5. Assume that a 2π -periodic function f belongs to the Sobolev space
Wp1 (−π , π ) with some 1 < p < ∞. Then its trigonometric Fourier series converges
absolutely.
Proof. Since Wp11 (−π , π ) ⊂ Wp12 (−π , π ) for 1 ≤ p2 < p1 , we may assume without
loss of generality that f ∈ Wp1 (−π , π ) with 1 < p ≤ 2. Then there is a function
g ∈ L p (−π , π ) with 1 < p ≤ 2 such that
x π
f (x) = g(t)dt + f (−π ), g(t)dt = 0. (9.4)
−π −π
where 1
p + 1
p = 1. The facts (9.4), (9.5) and Hölder’s inequality imply that
∞
1
∑ |cn ( f )| = |c0 ( f )|+ ∑ |cn (g)|
n=−∞ n=0 |n|
1/p
1/p
1 p
≤ |c0 ( f )| + ∑ p ∑ |cn (g)| < ∞,
n=0 |n| n=0
Remark 9.6. For the Sobolev space W11 (−π , π ), this theorem is not valid, i.e., there
is a function f from W11 (−π , π ) with absolutely divergent trigonometric Fourier
series. More precisely, we will prove in the next chapter that the function
∞
sin nx
f (x) := ∑ n log(1 + n) (9.6)
n=1
belongs to the Sobolev space W11 (−π , π ) and is continuous on the interval [−π , π ],
but its trigonometric Fourier series (9.6) diverges absolutely.
Proof. Since f ∈ W11 (−π , π ), it follows that f is of bounded variation. The period-
icity of f implies that
π π N
1 1
2π −π
| f (x+h)− f (x)|2 dx =
2π N −π k=1
∑ | f (x+kh)− f (x+(k −1)h)|2 dx, (9.7)
1/2
Prove that the function f belongs to the Nikol’skii space H2 (−π , π ) but its
trigonometric Fourier series is not absolutely convergent.
(1) Show that f belongs to the Nikol’skii space H21 (−π , π ) but
π
1 4h2 π
| f (x + h) − f (x)|2 dx ≥ log
2π −π π2 |h|
for 0 < |h| < 1, that is, (9.1) does not hold for α = 1.
(2) Show that f does not belong to the Besov space B12,θ (−π , π ) for any 1 ≤ θ < ∞.
Chapter 10
Dirichlet Kernel. Pointwise and Uniform
Convergence.
The material of this chapter forms a central part of the theory of trigonometric
Fourier series. In this chapter we will answer the following question: to what value
does a trigonometric Fourier series converge?
The Dirichlet kernel DN (x), which is defined by the symmetric finite trigonomet-
ric sum
DN (x) := ∑ einx , (10.1)
|n|≤N
plays a key role in this chapter. If x ∈ [−π , π ] \ {0}, then DN (x) from (10.1) can be
recalculated as follows. Using Euler’s formula, we have
N N 2N
DN (x) = ∑ einx = e−iNx ∑ ei(n+N)x = e−iNx ∑ eikx
n=−N n=−N k=0
1 − ei(2N+1)x e−iNx − ei(N+1)x
= e−iNx =
1 − eix 1 − eix
ei(N+1/2)x −e −i(N+1/2)x sin(N + 1/2)x
= = .
eix/2 − e−ix/2 sin x/2
sin(N + 1/2)x
DN (x) = , x = 0. (10.2)
sin x/2
For x = 0 we have
DN (0) = 2N + 1 = lim DN (x),
x→0
π
1 0, |n| > N,
cn (DN ) =
2π −π
e −inx
∑ e dx = 1, |n| ≤ N.
ikx
|k|≤N
Hence, if f is periodic and integrable, then the partial sum (10.3) can be rewritten
as (see Exercise 4.1)
∞ π
1
SN f (x) = ∑ cn (DN )cn ( f )einx = ( f ∗ DN )(x) =
2π −π
DN (x − y) f (y)dy
n=−∞
π π
1 1 sin(N + 1/2)y
= DN (y) f (x + y)dy = f (x + y) dy. (10.4)
2π −π 2π −π sin y/2
Show that
(1) (SN f ) (x) = 21π (DN (x) − 1);
(2) limN→∞ SN f (x) = f (x), x = 0;
(3) limN→∞ SN f (0) = 0.
Exercise 10.3. Prove that as N → ∞,
π
1 4 log N
|DN (x)|dx = + O(1).
2π −π π2
Since the Dirichlet kernel is an even function (see (10.2)), we can rewrite (10.4) as
π
1 sin(N + 1/2)y
SN f (x) = ( f (x + y) + f (x − y)) dy.
2π 0 sin y/2
Using the normalization of the Dirichlet kernel (see Exercise 10.1), we have for
every function S(x) that
10 Dirichlet Kernel. Pointwise and Uniform Convergence. 61
π
1 sin(N + 1/2)y
SN f (x) − S(x) = ( f (x + y) + f (x − y) − 2S(x)) dy. (10.5)
2π 0 sin y/2
Our aim is to define S(x) so that the limit of the right-hand side of (10.5) is equal to
zero. We will simplify the problem by splitting it into two steps. The first simplifi-
cation is connected with the following technical lemma.
for all z ∈ [−π , π ]. In order to prove this inequality, it is enough to show that
x3
0 < x − sin x <
6
for all 0 < x < π /2. The left inequality is well known. To prove the right inequality
we introduce h(x) as
h(x) = x − sin x − x3 /6.
for all 0 < x < π /2. Thus, h(x) is monotonically decreasing on the interval [0, π /2],
which implies that
0 = h(0) > h(x) = x − sin x − x3 /6
for all 0 < x < π /2. This proves (10.6), which in turn yields
1 2 2 |z/2 − sin(z/2)| |z|3 /24 |z|3 /24 π |z| π 2
sin z/2 z = |z|| sin(z/2)| ≤ |z|| sin(z/2)| ≤ |z||z|/π ≤ 24 ≤ 24 ,
−
since | sin z/2| ≥ |z|/π for all z ∈ [−π , π ]. This finishes the proof.
As an immediate corollary of Lemma 10.1, we obtain for every periodic and inte-
grable function f that the function
1 2
y → ( f (x + y) + f (x − y) − 2S(x)) −
sin y/2 y
62 Part I: Fourier Series and the Discrete Fourier Transform
f (x + y) + f (x − y) − 2S(x)
y
SN f (x) − S(x)
δ
1 sin(N + 1/2)y
= ( f (x + y) + f (x − y) − 2S(x)) dy + o(1) (10.8)
π 0 y
∑Nj=0 S j f (x)
σN f (x) = ,
N +1
1
( f (x + 0) + f (x − 0)) .
2
We can obtain some sufficient conditions when the limit in (10.8) exists.
Proof. The result follows immediately from (10.8) and the Riemann–Lebesgue
lemma.
Remark 10.3. If in (10.10) we have uniform convergence, then S(x) must necessar-
ily be periodic (S(−π ) = S(π )) and continuous on the interval [−π , π ].
uniformly in x ∈ [−π , π ].
for 0 < y < δ . This means that the condition (10.9) holds with S(x) = f (x) uniformly
in x ∈ [−π , π ], from which the statement of the corollary follows.
64 Part I: Fourier Series and the Discrete Fourier Transform
1 1
lim ( f (x + y) + f (x − y)) = ( f (x + 0) + f (x − 0)) =: S(x) (10.11)
y→0+ 2 2
F(y) := f (x + y) + f (x − y) − 2S(x),
where S(x) is defined by (10.11). Note that F(0) = 0. Let us also define
y
sin(N + 1/2)t
GN (y) := dt, 0 < y ≤ δ. (10.12)
0 t
where the last integral is well defined as the Stieltjes integral of the continuous
function GN (y) with respect to the function of bounded variation F(y). Since the
10 Dirichlet Kernel. Pointwise and Uniform Convergence. 65
limit (10.13) holds and GN (y) is continuous, we can consider the limit in (10.14) as
N → ∞. Hence, we obtain
1 π π δ
lim (SN f (x) − S(x)) = F(δ ) − dF(y)
N→∞ π 2 2 0
1
= (F(δ ) − F(δ ) + F(0)) = 0.
2
This completes the proof.
Corollary 10.6. If f is periodic and if f and f are piecewise continuous, then the
Fourier series of f converges to 12 ( f (x + 0) + f (x − 0)) at all points. If in addition
f is continuous on (−∞, ∞), then its Fourier series converges to f (x) uniformly on
(−∞, ∞).
Corollary 10.7. If f is periodic and belongs to the Sobolev space W11 (−π , π ), then
its trigonometric Fourier series converges pointwise to f (x) everywhere.
Proof. Since f ∈ W11 (−π , π ), it is of bounded variation and continuous on the in-
terval [−π , π ]. In this case, the value S(x) from (10.11) equals f (x) at every point
x ∈ [−π , π ]. Thus,
lim SN f (x) = f (x)
N→∞
pointwise in x ∈ [−π , π ].
Remark 10.8. The above proof does not allow us to conclude uniform convergence
of the trigonometric Fourier series of functions from the Sobolev space W11 (−π , π ).
However, uniform convergence is in fact the case, as we will prove later in this
chapter.
is of bounded variation but this function does not satisfy condition (10.9) at x = 0.
Hint. δ
1
0 y log y dy = +∞.
It is clear that F(x) belongs to the Sobolev space W11 (−π , π ) with F(−π ) = F(π ) =
0 (periodicity) and
F (x) = f (x) − c0 ( f ).
This implies
cn (F ) = incn (F) = cn ( f ), n = 0, c0 (F ) = 0.
Corollary 10.7 gives us that F(x) has everywhere convergent trigonometric Fourier
series
cn ( f ) inx
F(x) = c0 (F) + ∑ e . (10.16)
n=0 in
a
f (x)dx = ∑ cn ( f )
a
einx dx = lim
N→∞
∑ cn ( f )
a
einx dx
n=−∞ |n|≤N
b
= lim SN f (x)dx.
N→∞ a
10 Dirichlet Kernel. Pointwise and Uniform Convergence. 67
cn ( f ) cn ( f )(−1)n
∑ n and ∑ n
n=0 n=0
converge.
Proof. Let us assume to the contrary that there is a function f ∈ L1 (−π , π ) such
that
∞
sin(nx) 1 ∞ einx 1 ∞ e−inx
f (x) ∼ ∑ = ∑ log(1 + n) 2i ∑ log(1 + n)
−
n=1 log(1 + n) 2i n=1 n=1
1 ∞ einx 1 −∞ sgn(n)einx 1 sgn(n)einx
= ∑ + ∑ = ∑
2i n=1 log(1 + n) 2i n=−1 log(1 + |n|) 2i n=0 log(1 + |n|)
,
i.e., we have
1 sgn(n)
cn ( f ) = , n = 0, c0 ( f ) = 0.
2i log(1 + |n|)
sgn(n) 1 ∞ 1
∑ 2in log(1 + |n|) i ∑ n log(1 + n)
=
n=0 n=1
must be convergent. But this is not true. This contradiction proves this corollary.
Remark 10.12. If we define the function f by the series in Corollary 10.11, then it
turns out that π π
f (x)dx = 0, | f (x)|dx = +∞.
−π −π
1 − r2
Pr (x) = , 0≤r<1
1 − 2r cos x + r2
Corollary 10.4 shows us that this series converges to Pr (x) uniformly in x ∈ [−π , π ].
Theorem 10.13. Suppose that f ∈ C[−π , π ] is periodic. Then
uniformly in x ∈ [−π , π ] or
∞
lim
r→1− n=−∞
∑ r|n| cn ( f )einx = f (x) (10.17)
we have
1 π
(Pr ∗ f )(x) − f (x) = Pr (y)( f (x − y) − f (x))dy
2π −π
1
= Pr (y)( f (x − y) − f (x))dy
2π |y|≤δ
1
+ Pr (y)( f (x − y) − f (x))dy =: I1 + I2 .
2π δ ≤|y|≤π
= sup | f (x − y) − f (x)| → 0
x∈[−π ,π ],|y|≤δ
1 − r2 1 − r2
Pr (y) = =
1 − 2r cos y + r2 4r sin2 y/2 + (1 − r)2
1−r 1−r π2 1 − r
≤ ≤ = .
2r sin y/2 2rδ /π
2 2 2 2 rδ 2
uniformly in x ∈ [−π , π ]. The equality (10.17) follows from this fact and
Exercise 4.1. This completes the proof.
We will prove now the well known Hardy’s theorem and then apply it to the uniform
convergence of the trigonometric sum SN f (x).
exists, then
lim (σn − sn ) = 0,
n→∞
Indeed,
m
(m + 1)σm − (n + 1)σn − ∑ (m + 1 − j)a j
j=n+1
m n m
= ∑ (m + 1 − j)a j − ∑ (n + 1 − j)a j − ∑ (m + 1 − j)a j
j=0 j=0 j=n+1
n n n
= ∑ (m + 1 − j)a j − ∑ (n + 1 − j)a j = ∑ (m − n)a j = (m − n)sn .
j=0 j=0 j=0
Therefore,
m
(m + 1)σm − (n + 1)σn − ∑ (m + 1 − j)a j − (m − n)σn = (m − n)sn − (m − n)σn
j=n+1
or equivalently,
m
m(σm − σn ) + (σm − σn ) − ∑ (m + 1 − j)a j = (m − n)(sn − σn ),
j=n+1
i.e.,
m+1 m+1 m
j
m−n
(σm − σn ) −
m−n ∑ 1−
m+1
a j = sn − σn .
j=n+1
Let m > n → ∞ be such that m/n ∼ 1 + δ (i.e., limm,n→∞ m/n = 1 + δ ) with some
positive δ to be chosen. Since σm is a Cauchy sequence by (10.19), it follows that
m+1
(σm − σn ) → 0, m > n → ∞.
m−n
× δ − δ 2 /2 + o(δ 2 ) − δ (1 − δ + O(δ 2 ))
if δ is chosen small enough. Since δ is arbitrary, we may conclude that also the
second term converges to zero. This proves the theorem.
Corollary 10.15. Suppose that f ∈ C[−π , π ] is periodic and that its Fourier coeffi-
cients satisfy
M
|cn ( f )| ≤ , n = 0
|n|
with positive constant M that does not depend on n. Then the trigonometric Fourier
series of f converges to f uniformly in x ∈ [−π , π ].
Proof. Since f ∈ C[−π , π ] is periodic, Theorem 5.1 gives the convergence of Fejér
means
lim σN f (x) = f (x)
N→∞
Then
n
sn ≡ ∑ ak (x) = Sn f (x)
k=0
and
1 n 1 n
σn f (x) = ∑
n + 1 k=0
Sk f (x) = ∑ sk .
n + 1 k=0
2M
|ak (x)| ≤ , k = 1, 2, . . . .
k
uniformly in x ∈ [−π , π ].
72 Part I: Fourier Series and the Discrete Fourier Transform
uniformly in x ∈ [−π , π ].
Thus, f = g and
cn (g) = incn ( f ),
or equivalently,
cn (g) M
|cn ( f )| = ≤ , n = 0.
in |n|
Due to embedding (see Lemma 1.14), the function f is continuous on the interval
[−π , π ]. Using again Hardy’s theorem, we obtain
Remark 10.17. Corollary 10.16 and Exercise 10.7 show that for every function f
from the spaces H2α and Bα2,θ with 1 ≤ θ < ∞ and α > 1/2, its trigonometric Fourier
series converges to this function uniformly. Here one must take into account that f
might be changed on a set of measure zero.
and
∞
cos(nx)
f2 (x) = ∑ n log(1 + n) . (10.22)
n=1
10 Dirichlet Kernel. Pointwise and Uniform Convergence. 73
These functions are well defined for all x ∈ [−π , π ] \ {0}; see Theorem 3.3. In
addition, f1 (0) = 0, whereas f2 (0) is not defined, since the series (10.22) diverges
at zero. We will show that f2 (x) does not belong to W11 (−π , π ) but f1 (x) does.
If we assume to the contrary that f2 ∈ W11 (−π , π ), then its derivative f2 has the
Fourier series
∞
sin(nx)
f2 (x) ∼ − ∑ .
n=1 log(1 + n)
But due to Corollary 10.11 this is not a Fourier series of an L1 function. This con-
tradiction proves that f2 ∈ / W11 (−π , π ).
Concerning the series (10.21), let us prove first that it converges uniformly in
x ∈ [−π , π ], i.e., f1 (x) is (at least) continuous on the interval [−π , π ]. Indeed, by
summation by parts we obtain for 0 ≤ M < N that
N N n n−1
sin(nx) 1
∑ n log(1 + n) = ∑ n log(1 + n) ∑ sin(kx) − ∑ sin(kx)
n=M+1 n=M+1 k=1 k=1
sin(Nx) sin(Mx)
= −
N log(1 + N) M log(1 + M)
N+1 n
1 1
− ∑ ∑ sin(kx) (n + 1) log(2 + n) − n log(1 + n) . (10.23)
n=M+2 k=1
The first two terms on the right-hand side of (10.23) converge to zero as N > M → ∞
uniformly in x ∈ [−π , π ]. The sum on the right-hand side of (10.23) becomes, using
(10.24),
Let us consider two cases: n|x| < 1 and n|x| > 1. In the first case,
sin(nx/2) sin((n + 1)x/2) n|x|/2 · 1 π n
≤ = .
sin x/2 |x|/π 2
74 Part I: Fourier Series and the Discrete Fourier Transform
Then
1 1
|x| |x|
π n log (1 + 1/(n + 1)) π n 1n
|I1 | ≤ ∑
2 n=M+2 (n + 1) log(1 + n) log(2 + n)
< ∑
2 n=M+2 n log2 n
π ∞ 1
< ∑
2 n=M+2 n log2 n
→0 (10.25)
Then
N+1
n log (1 + 1/(n + 1)) N+1 n1
|I1 | ≤ π ∑
(n + 1) log(1 + n) log(2 + n)
<π ∑ n
n log2 n
1
n= |x| 1
n= |x|
∞
1
∑
<π → 0, M→∞
1
n log2 n
n= |x|
≥M
N+1 ∞
1 1 1 dt
∑
|I2 | ≤ ≤π
| sin x/2| 1 +1
2
n log n |x| 1
|x|
≥M t 2 logt
n= |x|
⎛ ⎞
∞
1 ⎝ 1 ∞ dt ⎠
=π − + 1
t logt 1 ≥M
|x| |x|
2
≥M t 2 log t
|x|
1/|x| 1 ∞ dt
=π +
[1/|x|] log[1/|x|] |x| |x|1 ≥M t 2 log2 t
[1/|x|] + 1 [1/|x|] + 1 ∞ dt
<π +
[1/|x|] log[1/|x|] [1/|x|] 2
M t log t
1 + 1/M 1 + 1/M
≤π + → 0, M → ∞ (10.26)
log M log M
converges uniformly (see Corollary 3.4) for π ≥ |x| ≥ δ > 0. This means that for
this interval, f1 (x) belongs to C1 . Thus, it remains to investigate the behavior of
the series (10.21) as x → 0+. But the estimates (10.25)–(10.26) show us that (if
we choose M 1/x, x → 0+) the function f1 (x) from (10.21) has the asymptotic
behavior
C
f1 (x) ∼ . (10.27)
log x
It is possible to prove (see [46, Chapter V, formula (2.19)]) that the asymptotic
(10.27) can be differentiated, and we obtain
C
f1 (x) ∼ − .
x log2 x
Let x(t) be a 2π -periodic continuous signal. Assume that x(t) can be represented by
an absolutely convergent trigonometric Fourier series
∞
x(t) = ∑ cm eimt , t ∈ [−π , π ], (11.1)
m=−∞
2π k N N
tk = , k = − , . . . , − 1.
N 2 2
Since ei2π kl = 1 for integers k and l, the series (11.2) can be rewritten as
∞ 2π k (m−lN)
∞ 2π k (m−lN)
∑ cm ei N = ∑ ∑ cm ei N
m=−∞ l=−∞ −N/2≤m−lN≤N/2−1
∞ N/2−1
2π k n
= ∑ ∑ cn+lN ei N
l=−∞ n=−N/2
N/2−1
2π k n
∞ N/2−1
2π k n
= ∑ ei N
∑ cn+lN = ∑ ei N Xn , (11.3)
n=−N/2 l=−∞ n=−N/2
∞
Xn = ∑ cn+lN . (11.4)
l=−∞
In the calculation (11.3) we have used the fact that the series (11.1) converges ab-
solutely. Combining (11.2) and (11.3), we obtain
N/2−1
2π kn
xk := x(tk ) = ∑ Xn ei N . (11.5)
n=−N/2
The formula (11.5) can be viewed as an inverse discrete Fourier transform, and it
appeared quite naturally in the discretization of a continuous periodic signal. More-
over, the formula (11.4) becomes the main property of this approach. Since
N/2−1
i(k−m) 2Nπ n 0, k − m = 0, ±N, ±2N, . . .
∑ e =
N, k − m = 0, ±N, ±2N, . . .
(11.6)
n=−N/2
we solve the linear system (11.5) with respect to Xn , n = −N/2, . . . , N/2 − 1, and
obtain
1 N/2−1 2π kn
Xn = ∑ xk e−i N .
N k=−N/2
(11.7)
In fact, the formulas (11.5) and (11.7) give us the inverse and direct discrete Fourier
transforms, respectively.
N/2−1
Definition 11.1. The sequence {Xn }n=−N/2 of complex numbers is called the dis-
N/2−1
crete Fourier transform (DFT) of the sequence {Yk }k=−N/2 if for each n = −N/2, . . . ,
N/2 − 1 we have
1 N/2−1 2π kn
Xn = ∑ Yk e−i N .
N k=−N/2
(11.8)
Xn = F (Yk )n
or simply X = F (Y ).
N/2−1
Definition 11.2. The sequence {Zk }k=−N/2 of complex numbers is said to be the
N/2−1
inverse discrete Fourier transform (IDFT) of the sequence {Xn }n=−N/2 if for each
k = −N/2, . . . , N/2 − 1 we have
11 Formulation of the Discrete Fourier Transform and Its Properties. 79
N/2−1
2π kn
Zk = ∑ Xn ei N . (11.9)
n=−N/2
Zk = F −1 (Xn )k
or simply Z = F −1 (X).
The properties of the DFT and IDFT are collected in the following lemmas.
Lemma 11.3. The following equalities hold:
(1) F −1 (F (Y )) = Y ;
(2) F (F −1 (X)) = X;
(3)
N/2−1
1 N/2−1
∑ F (Xn )k F (Yn )k = ∑ XnYn .
N n=−N/2
k=−N/2
n=−N/2 l=−N/2
N/2−1 N/2−1
1 2π n(k−l) 1
= ∑ Yl
N l=−N/2 ∑ ei N =
N
Yk N = Yk .
n=−N/2
This proves part (1). Part (2) can be proved in the same manner.
1 N/2−1 N/2−1
∑
N n=−N/2
|Xn |2 = ∑ |F (Xn )k |2 .
k=−N/2
Remark 11.5. Due to the periodicity of the complex exponential, we may extend
the values of Xm , m = −N/2, . . . , N/2 − 1 periodically to any integer by
N/2−1
Corollary 11.6. For a sequence X = {Xn }n=−N/2 we define
N/2−1
Xrev = {XN−n }n=−N/2 .
80 Part I: Fourier Series and the Discrete Fourier Transform
Then
F −1 (X) = NF (Xrev ).
N/2−1
(X ∗Y )k = ∑ Xl Yk−l , (11.11)
l=−N/2
N/2−l−1 N/2−1
2π nm 2π nm
∑ Ym e−i N = ∑ Ym e−i N .
m=−N/2−l m=−N/2
m=−N/2 m=N/2−l
N/2−1 N/2−1
2π nm 2π nm
− ∑ Ym e−i N = ∑ Ym e−i N
m=N/2−l m=−N/2
11 Formulation of the Discrete Fourier Transform and Its Properties. 81
m=−N/2 m=N/2
N/2−l−1 N/2−1
2π n (m−N) 2π nm
− ∑ Ym−N e−i N = ∑ Ym e−i N
m=N/2 m=−N/2
N/2−l−1 N/2−1
∑ Ym = ∑ Ym .
m=−N/2−l m=−N/2
(X ∗Y )k = (Y ∗ X)k
Proof. We have
N/2−1 k+1−N/2
(X ∗Y )k = ∑ Xl Yk−l = ∑ Y j Xk− j
l=−N/2 j=k+N/2
N/2−1+(k+1) N/2−1
= ∑ Y j Xk− j = ∑ Y j Xk− j = (Y ∗ X)k
j=−N/2+(k+1) j=−N/2
by Corollary 11.9.
1 N/2−1 N/2−l−1
2π n
= ∑ Xl ∑
N l=−N/2 m=−N/2−l
Ym e−i N (m+l)
1 N/2−1 2π nl
N/2−l−1
2π nm
= ∑
N l=−N/2
Xl e−i N ∑ Ym e−i N . (11.12)
m=−N/2−l
1 N/2−1 2π nl
N/2−1
2π nm
F (X ∗Y )n = ∑
N l=−N/2
Xl e−i N ∑ Ym e−i N = NF (X)n F (Y )n .
m=−N/2
1 −1
F −1 (X ·Y )n = F (X) ∗ F −1 (Y ) n ,
N
F (X ·Y )n = (F (X) ∗ F (Y ))n ,
N/2−1
where X ·Y denotes the sequence {Xk ·Yk }k=−N/2 .
N/2−1 N/2−1
∞
N/2−1
∑ |Xn − cn | = ∑
∑ cn+lN − cn
= ∑
∑ cn+lN
Similarly, we have
11 Formulation of the Discrete Fourier Transform and Its Properties. 83
N/2−1
N/2−1
−1 i 2πNkn
i 2πNkn
F (Xn )k − ∑ cn e
=
∑ (Xn − cn )e
n=−N/2
n=−N/2
N/2−1
≤ ∑ |Xn − cn | ≤ ∑ |cν |. (11.14)
n=−N/2 |ν |≥N/2
In the formulas (11.13) and (11.14) the numbers cn are the Fourier coefficients of
N/2−1 N/2−1
the signal x(t), and {Xn }n=−N/2 is the DFT of {x(tk )}k=−N/2 with tk = 2π k/N.
Theorem 11.13. If x(t) is periodic and belongs to the Sobolev space W2m (−π , π )
for some m = 1, 2, . . ., then
1
Xn = cn + o , N→∞ (11.15)
N m−1/2
and
N/2−1
2π kn 1
F −1 (Xn )k = ∑ cn ei N +o
N m−1/2
, N→∞ (11.16)
n=−N/2
The first sum on the right-hand side tends to zero as N → ∞ due to Parseval’s equality
for a function from the Sobolev space W2m (−π , π ). The second sum can be estimated
precisely. Namely, since for every m = 1, 2, . . . we have
1/2
1/2
∞
∑ |ν | −2m
N/2
t −2m
dt N −m+1/2 ,
|ν |≥N/2
we conclude that (11.15) and (11.16) follow from the last estimate and (11.13) and
(11.14), respectively.
N/2−1
(2) If a sequence Y = {Yk }k=−N/2 is real, then show that
F (Y )n = F (Y )N−n .
Chapter 12
Connection Between the Discrete Fourier
Transform and the Fourier Transform.
lim F f (ξ ) = 0.
ξ →±∞
Hence ∞
1
2| f(ξ )| ≤ √ | f (x + π /ξ ) − f (x)|dx → 0
2π −∞
| f(ξ + h) − f(ξ )|
1 2
≤√ | f (x)||e−ixh − 1|dx + √ | f (x)|dx =: I1 + I2 .
2π |xh|<δ 2π |xh|>δ
as |h| → 0. If we choose δ = |h|1/2 , then both I1 and I2 tend to zero as |h| → 0. This
completes the proof.
Theorem 12.2 (Fourier inversion formula). Suppose that f belongs to W11 (R).
Then
F −1 (F f )(x) = f (x)
at every point x ∈ R.
It remains to show that the latter limit is equal to zero. In order to prove this fact we
split the above integral as follows:
88 Part I: Fourier Series and the Discrete Fourier Transform
∞
sin(t)
( f (t/n) − f (0)) dt
−∞ t
sin(t) sin(t)
= ( f (t/n) − f (0)) dt + ( f (t/n) − f (0)) dt =: I1 + I2 .
|t|<1 t |t|>1 t
Since f ∈ W11 (R), it is continuous with respect to the definition of Sobolev spaces
in this chapter. Therefore,
sin(t)
|I1 | ≤ sup | f (t/n) − f (0)|
|t|<1
t dt → 0, n → ∞.
|t|<1
Since limz→±∞ f (z) = 0 (see Exercise 12.1) and since f is continuous, we obtain (as
n → ∞)
∞ nz
π sin(t)
I2 → − f (0) − lim f (z) dtdz
2 n→∞ 1/n 0 t
−1/n nz
π sin(t)
− f (0) − lim f (z) dtdz
2 n→∞ −∞ 0 t
12 Connection Between the Discrete Fourier Transform and the Fourier Transform. 89
∞ nz
sin(t)
= −π f (0) − lim f (z) dtdz
n→∞ 1/n 0 t
−1/n nz
sin(t)
− lim f (z) dtdz
n→∞ −∞ 0 t
∞
π 0 π
= −π f (0) − f (z) dz + f (z) dz
0 2 −∞ 2
π π
= −π f (0) + f (0) + f (0) = 0.
2 2
Here we used again the fact that limz→±∞ f (z) = 0 and Lebesgue’s dominated con-
vergence theorem. Thus, (12.1) transforms to
∞ √
p. v. f(ξ )dξ = 2π f (0),
−∞
or equivalently,
F −1 (F f )(0) = f (0).
In order to prove the Fourier inversion formula for every x ∈ R, let us note that
∞
1
fx (y)(ξ ) = f
(x + y)(ξ ) = √ f (x + y)e−iyξ dy
2π −∞
∞
1
=√ f (z)e−izξ eixξ dz = eixξ f(ξ ).
2π −∞
as |ξ | → ∞.
90 Part I: Fourier Series and the Discrete Fourier Transform
Proof. Since f ∈ W1m (R), it follows that f , f , . . . , f (m−1) ∈ W11 (R). By Exercise
12.1, we have
lim f (k) (x) = 0
x→±∞
The equality (12.2) allows us to consider (with respect to the accuracy of calcu-
lations) the Fourier transform only on the interval (−R, R) with R > 0 sufficiently
large, i.e., we may neglect the values of F f (ξ ) for |ξ | > R. This simplification
justifies the following approximation of the inverse Fourier transform:
R
1
f ∗ (x) := √ F f (ξ )eixξ dξ .
2π −R
At the same time and without loss of generality we may assume that the function
f (x) is equal to zero outside some finite interval. In that case it can be proved that
F f (ξ ) is a smooth function for which (12.2) holds.
◦
1 (−R, R) if f ∈ W1 (R) and f ≡ 0 for x ∈
Definition 12.5. We say that f ∈ W m m /
(−R, R).
◦
Theorem 12.6. Suppose that f ∈ W m 1 (−R, R) is supported in a fixed interval [a, b] ⊂
(−R, R) with R > 0 sufficiently large for some m = 2, 3, . . .. Then
N/2−1 x(2n+1)
2 1 2n + 1 i
f (x) = ∑ Ff
π N m/(m+2) n=−N/2 N m/(m+2)
e N m/(m+2)
1
+O (12.3)
N (2m−2)/(m+2)
Let us divide the interval [−R, R] into N + 1 subintervals [ξn , ξn+1 ], where n =
−N/2, . . . , N/2 − 1 such that
where
2Rn 2R
ξn = , ξn+1 − ξn = .
N N
Let us also set
ξn + ξn+1 R
ξn∗ = = (2n + 1).
2 N
Then we obtain
R
1
f ∗ (x) = √ F f (ξ )eixξ dξ
2π −R
N/2−1
1 ∗ 2R
=√
2π
∑ F f (ξn∗ )eixξn
N
n=−N/2
1 N/2−1 ξn+1 ∗
+√ ∑
2π n=−N/2 ξn
F f (ξ )eixξ − F f (ξn∗ )eixξn dξ
3
2 R N/2−1 R
= ∑ F f (R(2n + 1)/N)e
π N n=−N/2
ixR(2n+1)/N
+O
N2
. (12.5)
92 Part I: Fourier Series and the Discrete Fourier Transform
1 ∗
√
2π
∑ F f (ξ )eixξ − F f (ξn∗ )eixξn dξ
n=−N/2 ξn
⎧ 3
⎨O R 2 , supp f = [a, b]
= N
uniformly in x ∈ supp f .
Hint. Use the Taylor expansion for the smooth function F f (ξ )eixξ at the point ξn∗ .
If we combine (12.5) with (12.4) and choose R = N 2/(m+2) , then we obtain (12.3).
This completes the proof.
Remark 12.7. The main part of (12.3) represents some kind of inverse discrete
Fourier transform. In order to reconstruct f at a point x ∈ [−R, R] we need to know
only the Fourier transform of this unknown function at the points
2n + 1
, n = −N/2, . . . , N/2 − 1,
N m/(m+2)
where m is the smoothness index of f . What is more, the formula (12.3) shows us
that it is effective if
2m − 2 m
>
m+2 m+2
◦
or m > 2. This means that f must belong to the Sobolev space W m
1 (−R, R) with
some m ≥ 3.
Chapter 13
Some Applications of the Discrete Fourier
Transform.
lim
N→∞
∑ f (x + 2π n)
|n|≤N
Remark 13.2. It is clear that fp (x) is periodic with period 2π . Hence we will con-
sider it only on the interval [−π , π ].
If in addition
∞
∑ |F f (m)| < ∞,
m=−∞
then
∞ ∞
1
∑ f (x + 2π n) = √2π ∑ F f (m)eimx . (13.2)
n=−∞ m=−∞
∞ ∞
1
∑ f (2π n) = √
2π
∑ F f (m). (13.3)
n=−∞ m=−∞
−π
| fp (x)|dx ≤ ∑
−π n=−∞
| f (x + 2π n)|dx = ∑ | f (x + 2π n)|dx
n=−∞ −π
∞ π +2π n ∞
= ∑ | f (t)|dt = | f (t)|dt < ∞.
n=−∞ −π +2π n −∞
This shows that fp is finite almost everywhere and integrable on [−π , π ]. Applying
the same calculation to fp (x)e−imx allows us to integrate term by term to obtain
π ∞ π
1 1
cm ( fp ) =
2π −π
fp (x)e−imx dx = ∑ 2π −π
f (x + 2π n)e−imx dx
n=−∞
∞ π +2π n
1
= ∑ f (t)e−im(t−2π n) dt
n=−∞ 2π −π +2π n
∞ π +2π n
1 1
=√ ∑ √
2π n=−∞ 2π −π +2π n
f (t)e−imt dt
∞
1 1 1
= √ √ f (t)e−imt dt = √ F f (m).
2π 2π −∞ 2π
Thus, fp (x) can be represented by its Fourier series at least almost everywhere (and
we can redefine fp (x) so that this representation holds pointwise), i.e.,
∞
fp (x) = ∑ cm ( fp )eimx .
m=−∞
13 Some Applications of the Discrete Fourier Transform. 95
Example 13.4. If
1 − x2
f (x) = √ e 4t , x ∈ R,
4π t
where t > 0 is a parameter, then it is very well known (see, e.g., Example 16.7 and
Exercise 16.4) that
1
F f (ξ ) = √ e−t ξ .
2
2π
Remark 13.6. If we set F(ξ ) = 0 for |ξ | > 2πλ , then (13.4) is the inverse Fourier
transform of F ∈ L1 (R). In that case f is bounded and continuous.
Theorem 13.7 (Whittaker, Shannon, Boas). Suppose that F ∈ L1 (R) and F(ξ ) =
0 for |ξ | > 2πλ . If λ ≤ 1/2, then for every t ∈ R we have
∞
sin π (t − n)
f (t) = ∑ f (n)
π (t − n)
, (13.5)
n=−∞
96 Part I: Fourier Series and the Discrete Fourier Transform
Proof. Let
∞
Fp (ξ ) = ∑ F(ξ + 2π n)
n=−∞
F (F)(t) = f (−t),
where F (F) denotes the Fourier transform of F(ξ ). Hence, by the Poisson summa-
tion formula, we have (see (13.2))
∞ ∞ ∞
1 1
∑ F(ξ + 2π n) = √
2π
∑ f (−m)eimξ = √
2π
∑ f (m)e−imξ .
n=−∞ m=−∞ m=−∞
Since every trigonometric Fourier series can be integrated term by term, we obtain
π π ∞
−π
Fp (ξ )eit ξ dξ = ∑
−π n=−∞
F(ξ + 2π n)eit ξ dξ
∞ π
1
=√
2π
∑ f (m)
−π
ei(t−m)ξ dξ
m=−∞
∞
1 ei(t−m)π − e−i(t−m)π
=√
2π
∑ f (m)
i(t − m)
m=−∞
∞
2π sin π (t − m)
=√
2π
∑ f (m)
π (t − m)
.
m=−∞
so that (13.5) is proved. If λ > 1/2, then we cannot expect that F(ξ ) = Fp (ξ ) for
|ξ | > π , but using Definition 13.1 we have
13 Some Applications of the Discrete Fourier Transform. 97
π
1 2
it ξ
f (t) = √ F(ξ )e dξ + F(ξ )eit ξ dξ
2π
−π π π <|ξ |<2πλ
∞
sin π (t − n) 2
= ∑ f(n) + F(ξ )eit ξ dξ ,
n=−∞ π (t − n) π π <| ξ |<2 πλ
where π
1
f(n) = √ F(ξ )einξ dξ .
2π −π
Therefore,
∞
sin π (t − n)
f (t) = ∑ f (n)
π (t − n)
n=−∞
∞ sin π (t − n)
1
+ ∑ f(n) − f (n) +√ F(ξ )eit ξ dξ .
n=−∞ π (t − n) 2π π <|ξ |<2πλ
or
∞
1 sin π (t − n) inξ
−√ F(ξ ) ∑ e dξ .
2π π <|ξ |<2πλ n=−∞ π (t − n)
If F is smooth enough (say F ∈ C2 [−π , π ]), then formula (12.5) gives (R = π and
N 1)
√ N/2−1
2π π n(2k+1) 1
f (n) =
N ∑ Fk ei N +O
N2
k=−N/2
√
2π i π n N/2−1 2π kn 1
= e N ∑ Fk ei N + O ,
N k=−N/2
N2
where Fk denotes the value of F(ξ ) at the point π (2k + 1)/N. Therefore, up to the
accuracy of calculations,
√
2π i π n −1
f (n) ≈ e N F (Fk )n ,
N
i.e., for N a sufficiently large even integer,
√ ∞
2π i π n −1 sin π (t − n)
f (t) ≈ ∑ e N F (Fk )n .
n=−∞ N π (t − n)
Chapter 14
Applications to Solving Some Model
Equations
Let us consider a heat conduction problem for a straight bar of uniform cross sec-
tion and homogeneous material. Let x = 0 and x = L denote the ends of the bar
(the x-axis is chosen to lie along the axis of the bar). Suppose that no heat passes
through the sides of the bar. We also assume that the cross-sectional dimensions
are so small that the temperature u can be considered constant on any given cross
section (Figure 14.1).
Then u is a function only of the coordinate x and the time t. The variation of
temperature in the bar is governed by the partial differential equation
where α 2 is a constant known as the thermal diffusivity. This equation is called the
heat conduction equation or heat equation.
In addition, we assume that the initial temperature distribution in the bar is given
by
u(x, 0) = f (x), 0 ≤ x ≤ L, (14.2)
where f is a given function. Finally, we assume that the temperature at each end of
the bar is given by
c Springer International Publishing AG 2017 99
V. Serov, Fourier Series, Fourier Transform and Their Applications
to Mathematical Physics, Applied Mathematical Sciences 197,
DOI 10.1007/978-3-319-65262-7 14
100 Part I: Fourier Series and the Discrete Fourier Transform
where g0 and g1 are given functions. The problem (14.1), (14.2), (14.3) is an ini-
tial value problem in the time variable t. With respect to the space variable x it is a
boundary value problem, and the conditions of (14.3) are called the boundary con-
ditions. Alternatively, this problem can be considered a boundary value problem in
the xt-plane (Figure 14.2):
x
u(x, 0) = f (x)
x=0 x=L
Fig. 14.2 Geometric illustration of the heat equation as a boundary value problem.
or
X (x) 1 T (t)
= 2 ,
X(x) α T (t)
14 Applications to Solving Some Model Equations 101
in which the variables are separated, that is, the left-hand side depends only on x,
and the right-hand side only on t. This is possible only when both sides are equal to
the same constant:
X 1 T
= 2 = −λ .
X α T
Hence, we obtain two ordinary differential equations for X(x) and T (t):
X + λ X = 0,
T + α 2 λ T = 0. (14.6)
It follows that
X(0) = 0,
X(L) = 0.
So, for the function X(x) we obtain the homogeneous boundary value problem
X + λ X = 0, 0 < x < L,
(14.7)
X(0) = X(L) = 0.
The values of λ for which nontrivial solutions of (14.7) exist are called eigenvalues,
and the corresponding nontrivial solutions are called eigenfunctions. The problem
(14.7) is called an eigenvalue problem.
Lemma 14.1. The problem (14.7) has an infinite sequence of positive eigenvalues
n2 π 2
λn = , n = 1, 2, . . . ,
L2
with the corresponding eigenfunctions
nπ x
Xn (x) = c sin ,
L
where c is an arbitrary nonzero constant.
Proof. Suppose first that λ > 0, i.e., λ = μ 2 . The characteristic equation for (14.7)
is r2 + μ 2 = 0 with roots r = ±iμ , so the general solution is
102 Part I: Fourier Series and the Discrete Fourier Transform
Note that μ is nonzero, and there is no loss of generality if we assume that μ > 0.
The first boundary condition in (14.7) implies
X(0) = c1 = 0,
or
sin μ L = 0,
μ L = nπ , n = 1, 2, . . . ,
or
n2 π 2
λn = , n = 1, 2, . . . .
L2
Hence the corresponding eigenfunctions are
nπ x
Xn (x) = c sin .
L
Since
eμ x + e−μ x eμ x − e−μ x
cosh μ x = and sinh μ x = ,
2 2
this is equivalent to
X(x) = c1 eμ x + c2 e−μ x .
The first boundary condition requires again that c1 = 0, while the second gives
c2 sinh μ L = 0.
Since μ = 0 (μ > 0), it follows that sinh μ L = 0, and therefore we must have c2 = 0.
Consequently, X ≡ 0, i.e., there are no nontrivial solutions for λ < 0.
14 Applications to Solving Some Model Equations 103
X(x) = c1 x + c2 .
T (t) = ce−( ) .
nπα 2 t
L
) sin nπ x
N
∑ cn e−(
nπα 2 t
u(x,t) = L
n=1 L
is also a solution of the same problem. In order to take into account infinitely many
functions (14.8), we assume that
∞
) sin nπ x ,
∑ cn e−(
nπα 2 t
u(x,t) = L (14.9)
n=1 L
where the coefficients cn are still undetermined, and the series converges in some
sense. To satisfy the initial condition from (14.4) we must have
∞
nπ x
u(x, 0) = ∑ cn sin L
= f (x), 0 ≤ x ≤ L. (14.10)
n=1
In other words, we need to choose the coefficients cn so that the series (14.10) con-
verges to the initial temperature distribution f (x). We extend f from [0, L] to [−L, L]
as an odd function and then obtain that
L
2 nπ x
cn = f (x) sin dx.
L 0 L
It is not difficult to prove that for t > 0, 0 < x < L, the series (14.9) converges
(with any derivative with respect to x and t) and solves (14.1) with boundary condi-
tions (14.4). Only one question remains: can every function f (x) be represented by
a Fourier sine series (14.10)? Some sufficient conditions for such a representation
are given in Chapter 10.
104 Part I: Fourier Series and the Discrete Fourier Transform
Remark 14.2. We can consider the boundary value problem for a linear differential
equation
y + p(x)y + q(x)y = g(x) (14.11)
where y0 and y1 are given constants. Let us assume that we have found a fundamen-
tal set of solutions y1 (x) and y2 (x) to the corresponding homogeneous equation
where yp (x) is a particular solution to (14.11) and c1 and c2 are arbitrary constants.
To satisfy the boundary conditions (14.12) we have the linear inhomogeneous
algebraic system
c1 y1 (a) + c2 y2 (a) = y0 − yp (a),
(14.13)
c1 y1 (b) + c2 y2 (b) = y1 − yp (b).
If the determinant
y1 (a) y2 (a)
y1 (b) y2 (b)
is nonzero, then the constants c1 and c2 can be determined uniquely, and therefore
the boundary value problem (14.11)–(14.12) has a unique solution. If
y1 (a) y2 (a)
y1 (b) y2 (b) = 0,
where μ > 0 is fixed. This differential equation has a particular solution yp (x) = 1
μ2
.
Hence, the system (14.13) becomes
14 Applications to Solving Some Model Equations 105
c1 sin 0 + c2 cos 0 = y0 − μ12 ,
c1 sin μ + c2 cos μ = y1 − μ12 ,
or ⎧
⎨c2 = y0 − 12 ,
μ
If
0 1
sin μ cos μ = 0,
i.e., sin μ = 0, then c1 is uniquely determined, and the boundary value problem in
question has a unique solution. If sin μ = 0, then the problem has solutions (in fact,
infinitely many) if and only if
1 1
y1 − 2 = y0 − 2 cos μ .
μ μ
If μ = 2π k, then sin μ = 0 and cos μ = 1 and the following equation must hold:
1 1
y1 − = y0 − 2 ,
μ2 μ
2
y1 + y0 = .
μ2
Suppose now that the ends of the bar are held at constant temperatures T1 and T2 .
The corresponding boundary value problem is then
⎧
⎪
⎨α uxx = ut ,
2 0 < x < L,t > 0,
u(0,t) = T1 , u(L,t) = T2 , t > 0, (14.14)
⎪
⎩
u(x, 0) = f (x).
Then w(x,t) satisfies the homogeneous boundary value problem for the heat equa-
tion: ⎧
⎪
⎨α wxx = wt ,
2 0 < x < L,t > 0,
w(0,t) = w(L,t) = 0, (14.17)
⎪
⎩
w(x, 0) = f˜(x),
T2 − T1
v(x) = x + T1 , (14.18)
L
the solution of (14.17) is
∞
) sin nπ x ,
∑ cn e−(
nπα 2 t
w(x,t) = L (14.19)
n=1 L
A different problem occurs if the ends of the bar are insulated so that there is no
passage of heat through them. Thus, in the case of no heat flow, the boundary value
problem is ⎧
⎪
⎨α uxx = ut , 0 < x < L,t > 0,
2
This problem can also be solved by the method of separation of variables. If we let
u(x,t) = X(x)T (t), it follows that
X + λ X = 0, T + α 2 λ T = 0. (14.23)
nπ
The boundary conditions imply c1 = 0 and μ = L , n = 1, 2, . . ., leaving c2 arbi-
n2 π 2
trary. Thus we have an infinite sequence of positive eigenvalues λn = L2
with the
corresponding eigenfunctions
nπ x
Xn (x) = cos , n = 1, 2, . . . .
L
n2 π 2 nπ x
λn = , Xn (x) = cos , n = 0, 1, 2, . . . ,
L2 L
and
nπ x −( nπα )2 t
un (x,t) = cos e L , n = 0, 1, 2, . . . .
L
Each of these functions satisfies the equation and boundary conditions from (14.22).
It remains to satisfy the initial condition. In order to do so, we assume that u(x,t)
has the form
∞
c0 nπ x −( nπα )2 t
u(x,t) = + ∑ cn cos e L , (14.25)
2 n=1 L
14 Applications to Solving Some Model Equations 109
Thus the unknown coefficients in (14.25) must be the Fourier coefficients in the
Fourier cosine series of period 2L for the even extension of f . Hence
L
2 nπ x
cn = f (x) cos dx, n = 0, 1, 2, . . . ,
L 0 L
and the series (14.25) provides the solution to the heat conduction problem (14.22)
for a bar with insulated ends. The physical interpretation of the term
L
c0 1
= f (x)dx
2 L 0
is that it is the mean value of the original temperature distribution and a steady-state
solution in this case.
Example 14.4. ⎧
⎪
⎨uxx = ut , 0 < x < 1,t > 0,
u(0,t) = u(1,t) = 0,
⎪
⎩
u(x, 0) = ∑∞
n=1 n2 sin(nπ x) := f (x).
1
Since
∞ ∞
1
u(x, 0) = ∑ cn sin(nπ x) = ∑ n2 sin(nπ x),
n=1 n=1
Let us consider a bar with mixed boundary conditions at the ends. Assume that
the temperature at x = 0 is zero, while the end x = L is insulated so that no heat
passes through it: ⎧
⎪
⎨α uxx = ut , 0 < x < L,t > 0,
2
and
T + α 2 λ T = 0, t > 0.
As above, one can show that (14.26) has nontrivial solutions only for λ > 0, namely
∞
2
(2m − 1)π x − (2m−1)πα
∑
t
u(x,t) = cm sin e 2L
m=1 2L
14 Applications to Solving Some Model Equations 111
This is a Fourier sine series but in some specific form. We show that the coefficients
cm can be calculated as
L
2 (2m − 1)π x
cm = f (x) sin dx,
L 0 2L
which is why
∞
(2m − 1)π x
f(x) = ∑ c2m−1 sin ,
m=1 2L
112 Part I: Fourier Series and the Discrete Fourier Transform
where
1 2L (2m − 1)π x
c2m−1 = f (x) sin dx
L 0 2L
1 L (2m − 1)π x 1 2L (2m − 1)π x
= f (x) sin dx + f (2L − x) sin dx
L 0 2L L L 2L
L
2 (2m − 1)π x
= f (x) sin dx,
L 0 2L
ux (0,t) = u(L,t) = 0
Another situation in which the separation of variables applies occurs in the study
of a vibrating string. Suppose that an elastic string of length L is tightly stretched
between two supports, so that the x-axis lies along the string. Let u(x,t) denote the
vertical displacement experienced by the string at the point x at time t. It turns out
that if damping effects are neglected, and if the amplitude of the motion is not too
large, then u(x,t) satisfies the partial differential equation
x
x=0 x=L
u(x,t)
The initial conditions are (since (14.27) is of second order with respect to t)
where f and g are given functions. In order for (14.28) and (14.29) to be consistent,
it is also necessary to require that
x
u(x, 0) = f (x)
ut (x, 0) = g(x)
x=0 x=L
Fig. 14.4 Geometric illustration of the wave equation as a boundary value problem.
X + λ X = 0, T + a2 λ T = 0.
This is the same boundary value problem that we have considered previously.
Hence,
n2 π 2 nπ x
λn = 2 , Xn (x) = sin , n = 1, 2, . . . .
L L
nπ at nπ at
T (t) = k1 cos + k2 sin ,
L L
where k1 and k2 are arbitrary constants. Using the linear superposition principle, we
consider the series
∞
nπ x nπ at nπ at
u(x,t) = ∑ sin L
an cos
L
+ bn sin
L
, (14.31)
n=1
Since (14.30) are fulfilled it follows that equations (14.32) are the Fourier sine series
for f and g, respectively. Therefore,
L
2 nπ x
an = f (x) sin
dx,
L
0 L
L (14.33)
2 nπ x
bn = g(x) sin dx.
nπ a 0 L
Finally, we may conclude that the series (14.31) with the coefficients (14.33) solves
(at least formally) the boundary value problem (14.27)–(14.30).
Each displacement pattern
nπ x nπ at nπ at
is called a natural mode of vibration and is periodic in both the space variable x
and time variable t. The spatial period 2L
n in x is called the wavelength, while the
numbers nπLa are called the natural frequencies.
u(x,t) = ∑ sin L
an cos
L
+ bn sin
L
,
n=1
∞
nπ x −( nπα )2 t
u(x,t) = ∑ cn sin L
e L
n=1
for the wave and heat equations, we see that the second series has an exponential
factor that decays rapidly with n for every t > 0. This guarantees convergence of the
series as well as the smoothness of the sum. This is no longer true for the first series,
because it contains only oscillatory terms that do not decay with increasing n. This
means that the solution of the heat equation is a C∞ function in the corresponding
domain, but the solution of the wave equation is not necessarily smooth.
The boundary value problem for the wave equation with free ends of the string
can be formulated as follows:
⎧
⎪
⎨a uxx = utt , 0 < x < L,t > 0,
2
Let us first note that the boundary conditions imply that f (x) and g(x) must satisfy
b0t + a0 ∞
nπ x nπ at nπ at
and ∞
b0 nπ a nπ x
g(x) = + ∑ bn cos ,
2 n=1 L L
where
2 L nπ x
an = f (x) cos dx, n = 0, 1, 2, . . . ,
L 0 L
2 L
b0 = g(x)dx,
L 0
and L
2 nπ x
bn = g(x) cos dx, n = 1, 2, . . . .
nπ a 0 L
Exercise 14.5. (1) Show that there is no uniqueness for the problem
⎧
⎪
⎨uxx = utt , 0 < x < L,t > 0,
u(0,t) = u(L,t), ux (0,t) = ux (L,t), t ≥ 0,
⎪
⎩
u(x, 0) = f (x), ut (x, 0) = g(x), 0 ≤ x ≤ L,
Let us consider the wave equation on the whole line. It corresponds, so to speak,
to an infinite string. In that case we no longer have the boundary conditions, but we
have the initial conditions
a2 uxx = utt , −∞ < x < ∞,t > 0,
(14.34)
u(x, 0) = f (x), ut (x, 0) = g(x).
Proposition 14.6. The solution u(x,t) of the wave equation is of the form
if and only if
∂ξ ∂η u = 0,
or
u = ψ (ξ ) + ϕ (η ),
where ψ = Θ .
To satisfy the initial conditions, we have
It follows that
1 1 1 1
ϕ (x) = f (x) − g(x), ψ (x) = f (x) + g(x).
2 2a 2 2a
Integrating, we obtain
x x
1 1 1 1
ϕ (x) = f (x) − g(s)ds + c1 , ψ (x) = f (x) + g(s)ds + c2 ,
2 2a 0 2 2a 0
where c1 and c2 are arbitrary constants. But ϕ (x) + ψ (x) = f (x) implies c1 + c2 = 0.
Therefore, the solution of the initial value problem is
x+at
1 1
u(x,t) = ( f (x − at) + f (x + at)) + g(s)ds. (14.35)
2 2a x−at
where
1, |x| ≤ 1,
f (x) =
0, |x| > 1,
1
u(x,t) = ( f (x − t) + f (x + t)) .
2
Some solutions are graphed below (Figure 14.5).
u(x, 0)
x
−1 1
u(x, 1/2)
x
−1 1
u(x, 2)
1/2
x
−1 1
We can also apply d’Alembert’s formula to the finite string. Consider again the
boundary value problem with homogeneous boundary conditions with fixed ends of
the string: ⎧ 2
⎪a uxx = utt , 0 < x < L,t > 0,
⎪
⎪
⎨u(0,t) = u(L,t) = 0,t ≥ 0,
⎪
⎪ u(x, 0) = f (x), ut (x, 0) = g(x), 0 ≤ x ≤ L,
⎪
⎩
f (0) = f (L) = g(0) = g(L) = 0.
and h(x) is 2L-periodic, and let k(x) be the function defined for all x ∈ R such that
g(x), 0 ≤ x ≤ L,
k(x) =
−g(−x), −L ≤ x ≤ 0,
and k(x) is 2L-periodic. Let us also assume that f and g are C2 functions on the inter-
val [0, L]. Then the solution to the boundary value problem is given by d’Alembert’s
formula
1 1 x+at
u(x,t) = (h(x − at) + h(x + at)) + k(s)ds.
2 2a x−at
Remark 14.8. It can be checked that this solution is equivalent to the solution given
by the Fourier series.
solves
a2 uxx = utt + F(x,t), −∞ < x < ∞,t > 0,
u(x, 0) = f (x), ut (x, 0) = g(x), −∞ < x < ∞.
Example 14.9. (Linearized system of the equation of gas dynamics) The isen-
tropic (the entropy is assumed to be constant) flow of an inviscid gas in the one-
dimensional case satisfies the nonlinear equations
14 Applications to Solving Some Model Equations 121
⎧ 2
⎨u + u · u + c ρ = 0,
t x x
ρ
⎩
ρt + u · ρx + ρ ux = 0, −∞ < x < ∞,t > 0,
where u(x,t) is the velocity of the gas at x and time t, ρ (x,t) is the density, and
c = c(ρ ) is the known local speed of sound.
As a first step in understanding the general nature of solutions to this system we
assume that u(x,t) and ρ (x,t) are not very different from their values at time t = 0
and that these values and their derivatives are “small.” Neglecting products of terms
of “small” order, we arrive at the linear system (which is the linearization of the
original system)
⎧ 2 ⎧
⎨u + c0 ρ = 0, ⎨c0 ρxx = ρtt ,
t x
ρ0 so c2
⎩ ⎩ 0 uxx = utt ,
ρt + ρ0 ux = 0, ρ0
where ρ0 is the density of the fluid at rest and c0 = c(ρ0 ). This is just a wave equation
for ρ and u. Thus we have that
c0
u(x,t) = √ ( f (x − c0t) + g(x + c0t))
ρ0
√
ρ (x,t) = ρ0 ( f (x − c0t) − g(x + c0t)) ,
where f and g are arbitrary C2 functions, and they are the same here due to the
linearized system. If in addition
then
1 c0
u(x,t) = (ϕ1 (x − c0t) + ϕ1 (x + c0t)) + (ϕ2 (x − c0t) − ϕ2 (x + c0t)) ,
2 2ρ0
1 ρ0
ρ (x,t) = (ϕ2 (x − c0t) + ϕ2 (x + c0t)) + (ϕ1 (x − c0t) − ϕ1 (x + c0t)) .
2 2c0
One of the most important of all partial differential equations in applied mathematics
is the Laplace equation:
The Laplace equation appears quite naturally in many applications. For example, a
steady-state solution of the heat equation in two space dimensions,
α 2 (uxx + uyy ) = ut ,
satisfies the 2D Laplace equation (14.36). When electrostatic fields are considered,
the electric potential function must satisfy either the 2D or the 3D equation (14.36).
A typical boundary value problem for the Laplace equation is (in dimension two)
uxx + uyy = 0, (x, y) ∈ Ω ⊂ R2 ,
(14.37)
u(x, y) = f (x, y), (x, y) ∈ ∂ Ω ,
where g is given and ∂∂ νu is the outward normal derivative, is called the Neumann
problem (Neumann boundary conditions) (Figure 14.6).
y ν |ν| = 1
Ω ∂Ω
Fig. 14.6 Domain Ω and the outward unit normal vector ν on the boundary ∂ Ω .
for fixed a > 0 and b > 0. The solution of this problem can be reduced to the solu-
tions of
⎧
⎪
⎨uxx + uyy = 0, 0 < x < a, 0 < y < b,
u(x, 0) = u(x, b) = 0, 0 < x < a, (14.38)
⎪
⎩
u(0, y) = g2 (y), u(a, y) = f2 (y), 0 ≤ y ≤ b,
and ⎧
⎪
⎨uxx + uyy = 0, 0 < x < a, 0 < y < b,
u(x, 0) = g1 (x), u(x, b) = f1 (x), 0 < x < a,
⎪
⎩
u(0, y) = 0, u(a, y) = 0, 0 ≤ y ≤ b.
y
u(x, b) = 0
b
x
u(x, 0) = 0 a
and
X − λ X = 0, 0 < x < a. (14.40)
u(x, y) = ∑ sin b
an cosh
b
+ bn sinh
b
. (14.41)
n=1
with b
2 nπ y
an = g(y) sin dy.
b 0 b
At x = a we obtain
∞
nπ y nπ a nπ a
f (y) = ∑ sin b
an cosh
b
+ bn sinh
b
.
n=1
This implies
bn − an cosh nπb a
bn = . (14.42)
sinh nπb a
because cosh α sinh β − sinh α cosh β = sinh(β − α ). Using the properties of sinh α
and cosh α for large α , we may conclude that inside of the rectangle, i.e., for
14 Applications to Solving Some Model Equations 125
0 < x < a, 0 < y < b, we may differentiate this series term by term as often as we
wish, and so u is a C∞ function there.
Exercise 14.9. Find a solution of the problem
⎧
⎪
⎨uxx + uyy = 0, 0 < x < 2, 0 < y < 1,
u(x, 0) = u(x, 1) = 0, 0 < x < 2,
⎪
⎩
u(0, y) = 0, u(2, y) = y(1 − y), 0 ≤ y ≤ 1,
u(a, θ ) = f (θ ), (14.43)
1 1
urr + ur + 2 uθ θ = 0. (14.44)
r r
We apply again the method of separation of variables and assume that
1 1
R T + R T + 2 RT = 0,
r r
or
r2 R + rR − λ R = 0,
T + λ T = 0.
f (0) = f (2π ))
T (0) = T (2π ), T (0) = T (2π ). (14.46)
so that c1 = c2 = 0.
T (θ ) = c1 cos(μθ ) + c2 sin(μθ ).
or
c1 sin2 (μπ ) = c2 sin(μπ ) cos(μπ ),
c2 sin2 (μπ ) = −c1 sin(μπ ) cos(μπ ).
If sin(μπ ) = 0, then
c1 = c2 cot(μπ ),
c2 = −c1 cot(μπ ).
Turning now to R, for λ = 0 we have r2 R + rR = 0, i.e., R(r) = k1 + k2 log r. Since
log r → −∞ as r → 0, we must choose k2 = 0 in order for R (and u) to be bounded.
That is why
R0 (r) ≡ constant. (14.48)
14 Applications to Solving Some Model Equations 127
r2 R + rR − n2 R = 0.
Hence
R(r) = k1 rn + k2 r−n .
Rn (r) = k1 rn , n = 1, 2, . . . . (14.49)
and 2π
1
bn = f (θ ) sin(nθ )dθ .
π an 0
This procedure can be used also to study the Neumann problem, i.e., the problem
in the disk with the boundary condition
∂u
(a, θ ) = f (θ ). (14.51)
∂r
Also in this case the solution u(r, θ ) has the form (14.50). The boundary condition
(14.51) implies that
∂u ∞
(r, θ ) = ∑ nan−1 (an cos(nθ ) + bn sin(nθ )) = f (θ ).
∂r r=a n=1
128 Part I: Fourier Series and the Discrete Fourier Transform
Hence 2π
1
an = f (θ ) cos(nθ )dθ
π nan−1 0
and 2π
1
bn = f (θ ) sin(nθ )dθ .
π nan−1 0
since integrating
∞
f (θ ) = ∑ nan−1 (an cos(nθ ) + bn sin(nθ ))
n=1
Exercise 14.11. Prove (14.52) and then show that we have no uniqueness in this
boundary value problem. Hint: Use the fact that if cn = 0, n = 1, 2, . . ., then we may
uniquely determine a0 , an , bn satisfying the boundary condition u(a, θ ) = f (θ ), and
if an = bn = 0, n = 1, 2, . . ., we may uniquely define cn , n = 1, 2, . . . that depend
parametrically on an arbitrary constant a0 .
Remark 14.12. The solution u(r, θ ) of the Laplace equation in the disk {x ∈ R2 :
|x| < a} subject to the boundary condition uθ (a, θ ) = f (θ ) with (possibly non-
smooth) periodic function f (θ ) can be obtained as
∞ ∞
a0 θ + b0
u(r, θ ) = + ∑ rn (an cos(nθ ) + bn sin(nθ )) + ∑ rn−1/2 cn cos(n − 1/2)θ .
2 n=1 n=1
(14.53)
Exercise 14.12. Prove (14.53) and then show that we have no uniqueness in this
boundary value problem (see the previous exercise).
Part II
Fourier Transform and Distributions
Chapter 15
Introduction
In this part we assume that the reader is familiar with the following concepts:
(1) Metric spaces and their completeness.
(2) The Lebesgue integral in a bounded domain Ω ⊂ Rn and in Rn .
(3) The Banach spaces (L p , 1 ≤ p ≤ ∞, Ck ) and Hilbert spaces (L2 ): If 1 ≤ p < ∞,
then we set
1/p
L (Ω ) := { f : Ω → C measurable : f L p (Ω ) :=
p
| f (x)| dx
p
< ∞},
Ω
while
Moreover,
Ck (Ω ) := { f : Ω → C : f Ck (Ω ) := max ∑
x∈Ω |α |≤k
|∂ α f (x)| < ∞},
(4) Hölder’s inequality: Let 1 ≤ p ≤ ∞, u ∈ L p , and v ∈ L p with
1 1
+ = 1.
p p
Then uv ∈ L1 and
1 1
p
p
p
|u(x)v(x)|dx ≤ |u(x)| dx
p
|v(x)| dx ,
Ω Ω Ω
if f ∈ L1 (X ×Y ).
(7) The divergence theorem: Let Ω ⊂ Rn be a bounded domain with C1 boundary
∂ Ω , and let F be a C1 vector-valued function on Ω . Then
div F(x)dx = ν · Fdσ (x),
Ω ∂Ω
Consider
the Euclidean space Rn , n ≥ 1, with x = (x1 , . . . , xn ) ∈ Rn and with |x| =
x12 + · · · + xn2 and scalar product (x, y) = ∑nj=1 x j y j . The open ball of radius δ > 0
centered at x ∈ Rn is denoted by
Uδ (x) := {y ∈ Rn : |x − y| < δ }.
|α | = α1 + · · · + αn , α ! = α1 ! · · · αn !
and
xα = x1α1 · · · xnαn , 00 = 1, 0! = 1.
α ≤β
∂
∂ α = ∂1α1 · · · ∂nαn , ∂j = .
∂xj
Example 16.2.
f (x) = e−a|x| ∈ S for every a > 0.
2
(1)
f (x) = e−a(1+|x| ) ∈ S for every a > 0.
2 a
(2)
(3) f (x) = e−|x| ∈
/ S.
(4) C0∞ (Rn ) ⊂ S(Rn ), where
| f |α ,β = 0 if and only if f =0
fails to hold for, e.g., a constant function f . The space (S, ρ ) is not normable but it
is a metric space if the metric ρ is defined by
| f − g|α ,β
ρ ( f , g) = ∑ 2−|α |−|β | ·
1 + | f − g|α ,β
.
α ,β ≥0
Theorem 16.3 (Completeness). The space (S, ρ ) is a complete metric space, i.e.,
every Cauchy sequence converges.
16 The Fourier Transform in Schwartz Space 135
Proof. Let { fk }∞k=1 , f k ∈ S, be a Cauchy sequence, that is, for every ε > 0 there
exists n0 (ε ) ∈ N such that
ρ ( fk , fm ) < ε , k, m ≥ n0 (ε ).
It follows that
sup ∂ β ( fk − fm ) < ε
x∈K
for every β ≥ 0 and for every compact set K in Rn . This means that { fk }∞
k=1 is a
Cauchy sequence in the Banach space C|β | (K). Hence there exists a function f ∈
C|β | (K) such that
C|β | (K)
lim fk = f.
k→∞
Thus we may conclude that our function f is in C∞ (Rn ). It remains only to prove
that f ∈ S. It is clear that
Taking k → ∞, we obtain
| fk − f |α ,β → 0, k→∞
for all α , β ≥ 0.
Exercise 16.2. Prove that C0∞ (Rn ) = S, that is, for every f ∈ S there exists { fk }∞
k=1 ,
S
fk ∈ C0∞ (Rn ), such that fk → f , k → ∞.
Now we are in position to define the Fourier transform in S(Rn ).
Definition 16.5. The Fourier transform F f (ξ ) or f(ξ ) of the function f (x) ∈ S is
defined by
F f (ξ ) ≡ f(ξ ) := (2π )−n/2 e−i(x,ξ ) f (x)dx, ξ ∈ Rn .
Rn
136 Part II: Fourier Transform and Distributions
for m > n.
Indeed, we have
β
∂ξ f(ξ ) = (2π )−n/2 (−ix)β e−i(x,ξ ) f (x)dx
Rn
and hence
|x||β |
β
∂ξ f (ξ ) ≤ cm (2π )−n/2 dx < ∞
L∞ (Rn ) Rn (1 + |x|)m
β
∂ξ f(ξ ) = (−ix)β f (x). (16.1)
since ∂xα f (x) ∈ S for every α ≥ 0 if f (x) ∈ S. And also we have the formula
ξ α f = (−i)|α | ∂ α f . (16.2)
x
∂
D j = −i∂ j = −i , Dα = Dα1 1 · · · Dαn n .
∂xj
16 The Fourier Transform in Schwartz Space 137
For this new derivative the formulas (16.1) and (16.2) can be rewritten as
Dξα f = (−1)|α | x
α f, ξ α f = D
α f.
Rn
= (2π )−n/2 e− 2 |ξ |
1 1 2 +2i(x,ξ )−|ξ |2 )
e− 2 (|x|
2
dx
Rn
n ∞
= (2π )−n/2 e− 2 |ξ | ∏ e− 2 (t+iξ j ) dt.
1 2 1 2
j=1 −∞
z2
In order to calculate the last integral, we consider the function f (z) = e− 2 of the
complex variable z and the domain DR depicted in Figure 16.1.
DR
t
−R R
But
R ξj
z2 t2
e− 2 (R+iτ ) dτ
1
e− 2 dz = e− 2 dt + i
2
∂ DR −R 0
−R 0
− 12 (t+iξ j )2
e− 2 (−R+iτ ) dτ .
1 2
+ e dt + i
R ξj
138 Part II: Fourier Transform and Distributions
If R → ∞, then
ξj
e− 2 (±R+iτ ) dτ → 0.
1 2
Hence ∞ ∞
t2
e− 2 (t+iξ j ) dt =
1
e− 2 dt,
2
j = 1, . . . , n.
−∞ −∞
Using Fubini’s theorem and polar coordinates, we can evaluate the last integral as
∞
2 2π ∞
t2 1 r2
e− 2 dt e− 2 (t
2 +s2 )
= dtds = dθ e− 2 rdr
−∞ R2 0 0
∞
−m
= 2π e dm = 2π .
0
Thus ∞ √
e− 2 (t+iξ j ) dt =
1 2
2π
−∞
and
|x|2 n √
)(ξ ) = (2π )− 2 e− 2 |ξ | ∏ 2π = e− 2 |ξ | .
n 1 1
F(e−
2 2
2
j=1
P(D) = ∑ aα Dα ,
|α |≤m
= P(ξ )
with constant coefficients. Prove that P(D)u u.
Definition 16.8. We adopt the following notation for translation and dilation of a
function:
(τh f )(x) := f (x − h), (σλ f )(x) := f (λ x), λ = 0.
Exercise 16.5. Let A be a real-valued n × n matrix such that A−1 exists. Define
fA (x) := f (A−1 x). Prove that
fA (ξ ) = ( f)A (ξ )
(F f , g)L2 =
f (ξ )g(ξ )dξ = (2π )−n/2
g(ξ ) e−i(x,ξ )
f (x)dx dξ
Rn Rn Rn
Remark 16.9. Here F ∗ is the adjoint operator (in the sense of L2 ), which maps S
into S since F : S → S. The inverse Fourier transform F −1 is defined as F −1 :=
F ∗.
Theorem 16.10 (Fourier inversion formula). Let f be a function from S(Rn ). Then
F ∗F f = f .
To this end we will prove first the following (somewhat technical) lemma.
Lemma 16.11. Let f0 (x) be a function from L1 (Rn ) with Rn f0 (x)dx = 1 and let
f (x) be a function from L∞ (Rn ) that is continuous at {0}. Then
x
lim ε −n f0 f (x)dx = f (0).
ε →0+ Rn ε
Proof. Since
x
x
ε −n f0 f (x)dx − f (0) = ε −n f0 ( f (x) − f (0))dx,
Rn ε Rn ε
η
| f (x)| <
f 0 L 1
= η + f L∞ Iε .
where we have used Lebesgue’s dominated convergence theorem in the last step.
This proves that
f (0) = (2π )−n/2 F f (ξ )ei(0,ξ ) dξ = (F ∗ F f )(0).
Rn
Proof. The fact that the Fourier transform preserves the norm of f ∈ S follows from
Note that
(F f , g)L2 = ( f , F ∗ g)L2
16 The Fourier Transform in Schwartz Space 141
means that
f(ξ )g(ξ )dξ = f (x)F ∗ g(x)dx = f (x)F (g)(x)dx.
Rn Rn Rn
or
F f , gL2 (Rn ) =
f , F gL2 (Rn ) .
Chapter 17
The Fourier Transform in L p (Rn ), 1 ≤ p ≤ 2
then there exists a unique linear continuous map Tex : X → Y such that Tex |E = T
and
Tex uY ≤ M uX , u ∈ X.
Lp
C0∞ (Rn ) = L p (Rn ),
or
ϕ − χA L p (Rn ) < ε ,
Remark 17.3. Lemma 17.2 does not hold for p = ∞. Indeed, for a function f ≡ c0 =
0 and for every function ϕ ∈ C0∞ (Rn ), we have that
Hence we cannot approximate a function from L∞ (Rn ) by functions from C0∞ (Rn ).
This means that ∞ L
C0∞ (Rn ) = L∞ (Rn ).
Proof. We know that F f L∞ ≤ (2π )−n/2 f L1 for f ∈ S. Now we apply the pre-
L1
liminary proposition to E = S, X = L1 , and Y = L∞ . Since S = L1 (which follows
L1
from C0∞ ⊂ S and C0∞ = L1 ) for every f ∈ L1 (Rn ), there exists { fk } ⊂ S such that
fk − f L1 → 0 as k → ∞. In that case, we can define
L∞
Fex f := lim F fk .
k→∞
L∞
Since S = Ċ (see Exercise 17.2), it follows that Fex f ∈ Ċ and Fex L1 →L∞ ≤
(2π )−n/2 . On the other hand,
Proof (Alternative proof). If f ∈ L1 (Rn ), then we can define the Fourier transform
F f (ξ ) directly by
F f (ξ ) := (2π )−n/2 e−i(x,ξ ) f (x)dx,
Rn
since
e−i(x,ξ )
f (x)dx ≤ | f (x)|dx = f L1 .
Rn Rn
Also we have
(2π )n/2 f(ξ + h) − f(ξ ) = sup e−i(ξ ,x) −i(h,x)
(e − 1) f (x)dx
L∞ (Rn ) Rn
n ξ ∈R
≤2 ε
| f (x)|dx + ε | f (x)|dx =: I1 + I2 .
|x|> |h| |x||h|≤ε
Here we have used the fact that |eiy − 1| ≤ |y| for y ∈ R with |y| ≤ 1. It is easily seen
that I1 → 0 for |h| → 0 and I2 → 0 for ε → 0, since f ∈ L1 (Rn ).
This means that the Fourier transform f(ξ ) is continuous (even uniformly con-
tinuous) on Rn . Moreover, we have
ξπ
2 f(ξ ) = (2π )−n/2 e −i(x,ξ )
f (x) − f x + 2 dx.
Rn |ξ |
for |ξ | → ∞.
L2 L2
Proof. We know that S = L2 , since C0∞ = L2 . Thus for every f ∈ L2 (Rn ) there exists
{ fk }∞
k=1 ⊂ S(R ) such that f k − f L2 (Rn ) → 0 as k → ∞. By Parseval’s equality in
n
S we get
F fk − F fl L2 = fk − fl L2 → 0, k, l → ∞.
L2
Thus {F fk }∞
k=1 is a Cauchy sequence in L (R ), and therefore, F f k → g, where
2 n
Exercise 17.4. Let us assume that f ∈ L1 (Rn ) and F f (ξ ) ∈ L1 (Rn ). Prove that
f (x) = (2π )−n/2 ei(x,ξ ) F f (ξ )dξ = F −1 F f (x).
Rn
( f1 , f2 )L2 = (F f1 , F f2 )L2 .
1 θ 1−θ 1 θ 1−θ
= + , = + , 0 ≤ θ ≤ 1,
p p1 p2 q q1 q2
where 1
q1 + q1 = 1, 1
q2 + q1 = 1, | f0 | ≤ 1, and |g0 | ≤ 1. The two functions f0 and g0
1 2
will be selected later. We assume also that 0 ≤ Re(z) ≤ 1.
Our aim is to prove the inequality
where
1 θ 1−θ 1 θ 1−θ 1 1
= + , = + , + = 1.
p p1 p2 q q1 q2 q q
z z + 1−z
+ 1−z
Since T is a linear continuous map and F p1 p2 , G q1 q2 are holomorphic func-
tions with respect to z (consider az = ez log a , a > 0), we may conclude that Φ (z) is a
holomorphic function also.
(1) Let us assume now that Re(z) = 0, i.e., z = iy. Then we have
iy
iy
+ 1−iy + 1−iy
Φ (iy) = M1−iy M2−1+iy T ( f0 F p1
p 2 ), g0 G q1 q2
L2 .
Since |aix | = 1 for a, x ∈ R, a > 0, it follows from Hölder’s inequality and the
assumptions on T that
iy 1−iy iy
+ 1−iy
p + p
|Φ (iy)| ≤ M2−1 M2
f0 F 1 2
g0 G q1 q2
q
L p2 L 2
1 1 1 1
q
= p2 |g0 |G q2 p
|
0f |F
≤ FL12 GL12 = 1.
q
L p2 L 2
(2) Let us assume now that Re(z) = 1, i.e., z = 1 + iy. Then we have similarly that
1+iy −iy 1+iy −iy
+
p + p2
|Φ (1 + iy)| ≤ M1−1 M1
f0 F 1 p
g0 G q1 q2
q
L 1 L 1
1 1 1 1
q1
= p1 |g0 |G q1 p
|
0f |F ≤ FL11 G L1
= 1.
q
L p1 L 1
If we apply now the Phragmén–Lindelöf theorem for the domain 0 < Re(z) < 1, we
obtain that |Φ (z)| ≤ 1 for every z such that 0 < Re(z) < 1. Then |Φ (θ )| ≤ 1 also for
0 < θ < 1. But this is equivalent to the estimate
148 Part II: Fourier Transform and Distributions
1 1
|T ( f0 F p ), g0 G q
L2 | ≤ M1θ M21−θ , (17.1)
θ θ 1 θ θ
where 1
p = p1 + 1−
p2 , q = q1 + 1−
q2 and
1
q + q1 = 1. In order to finish the proof of
this theorem let us choose (for arbitrary functions f ∈ L p and g ∈ Lq with p and q
as above) the functions F, G, f0 and g0 as follows:
F = | f1 | p , G = |g1 |q , f0 = sgn f1 , g0 = sgn g1 ,
f g
where f1 = f L p , g1 = g q , and
L
⎧
⎪
⎨1, f > 0,
sgn f = 0, f = 0,
⎪
⎩
−1, f < 0.
1 1
In that case, f1 = f0 F p and g1 = g0 G q . Applying the estimate (17.1), we obtain
f g
T , ≤ M1θ M21−θ ,
f L p gLq L2
which is equivalent to
where 1 ≤ p ≤ 2 and 1
p + 1
p = 1. What is more, we have the norm estimate
−n 1p − 12
Fex L p →L p ≤ (2π ) .
Proof. We know from Theorems 17.4 and 17.5 that there exists a unique extension
Fex of the Fourier transform from S to S for spaces:
17 The Fourier Transform in L p (Rn ), 1 ≤ p ≤ 2 149
n
(1) Fex : L1 (Rn ) → L∞ (Rn ) with norm estimate M1 = (2π )− 2 ;
(2) Fex : L2 (Rn ) → L2 (Rn ) with norm estimate M2 = 1.
Applying now Theorem 17.7, we obtain that Fex : L p → Lq , where
1 θ 1−θ 1 θ 1 θ 1−θ 1 θ
= + = + , = + = − .
p 1 2 2 2 q ∞ 2 2 2
It follows that
1 1
+ = 1,
p q
fk − f L p (Rn ) → 0, k → ∞.
F fk − F fl L p (Rn ) ≤ Cn fk − fl L p (Rn ) .
This means that {F fk }∞
k=1 is a Cauchy sequence in L (R ). We can therefore define
p n
Lp
Fex f := lim F fk .
k→∞
e −izξ
In order to calculate this integral we consider the function F(z) := (z−i ε )2
of a com-
plex variable z ∈ C. It is easily seen that z = iε , ε > 0, is a pole of order 2. We
consider the cases ξ > 0 and ξ < 0 separately; see Figure 17.1.
Im z Im z
i i
−R D+
R
R
Re z Re z
−R R
D−
R
and
I2 → 0, R→∞
due to Jordan’s lemma, since ξ Imz < 0. We therefore may conclude that
+∞ −ixξ
e dx
=0
−∞ (x − iε )2
for ξ > 0.
(2) Let ξ < 0. In this case again ξ Imz < 0. So we may apply Jordan’s lemma again
and obtain
R −ixξ
e dx
F(z)dz = + F(z)dz = 2π iRes F(z).
∂ D+
R −R (x − iε )2 z=iε
|z|=R
Imz>0
17 The Fourier Transform in L p (Rn ), 1 ≤ p ≤ 2 151
Hence ∞ −ixξ
e dx
= 2π i((z − iε )2 F(z)) |z=iε = 2πξ eεξ .
−∞ (x − iε )2
1 √
(ξ ) = 2πξ H(−ξ )eεξ ,
(x − iε )2
where
1, t ≥ 0,
H(t) =
0, t < 0,
1 √
(ξ ) = − 2πξ H(ξ )e−εξ .
(x + iε )2
Example 17.11. Let f1 (x) = x−i ε , where ε > 0 is fixed. It is clear that f 1 ∈
/ L1 (R),
1
√
1 −i 2π H(ξ )e−εξ , ξ = 0,
(ξ ) =
x + iε −i π2 , ξ = 0,
and √
1 i 2π H(−ξ )eεξ , ξ = 0,
(ξ ) =
x − iε i π2 , ξ = 0.
Exercise 17.6. Find the Fourier transforms of the following functions on the line.
e−x , x > 0,
(1) f (x) =
0, x ≤ 0,
(2) f (x) = e−|x| and f (x) = 1
1+x2
,
(3) f3 (x) = (x±i1ε )3 , ε > 0.
where | f (x)| ≤ Meax , x > 0, f (x) = 0, x < 0, and p = p1 + ip2 , p1 > a. Prove that
152 Part II: Fourier Transform and Distributions
√
(1) L(p) = 2π f (x)e −p1 x (p ).
2
(2) Apply the Fourier inversion formula to prove the Mellin formula
p1 +i∞
1
f (x) = L(p)e px dp, p1 > a.
2π i p1 −i∞
Chapter 18
Tempered Distributions
(1) there exists a compact set K ⊂ Rn such that supp ϕk ⊂ K for every k and
(2) for every α ≥ 0 we have
D D D
We denote this fact by ϕk → 0. As usual, ϕk → ϕ ∈ D means that ϕk − ϕ → 0.
Now we are in a position to define the Schwartz distribution space.
Definition 18.2. A functional T : D → C is a Schwartz distribution if it is linear
and continuous, that is,
(1) T (α1 ϕ1 + α2 ϕ2 ) = α1 T (ϕ1 ) + α2 T (ϕ2 ) for every ϕ1 , ϕ2 ∈ D and α1 , α2 ∈ C
(2) for every null sequence ϕk in D, one has T (ϕk ) → 0 in C as k → ∞.
The linear space of Schwartz distributions is denoted by D . The action of T on ϕ
is denoted by T (ϕ ) = T, ϕ .
for some locally integrable function f . All other distributions are singular.
gT, ϕ := T, gϕ , ϕ ∈ D.
by integration by parts. This property is used to define the derivative of any distrib-
ution.
Definition 18.10. Let T be a distribution from D . For a multi-index α we define
the derivative ∂ α T by
∂ α T, ϕ := T, (−1)|α | ∂ α ϕ , ϕ ∈ D.
it follows that
∞
H , ϕ = −H, ϕ = − ϕ (x)dx = ϕ (0) = δ , ϕ .
0
Hence H = δ .
Example 18.12. Let us prove that (log |x|) = p. v. 1x in the sense of Schwartz distri-
butions. Indeed,
∞
(log |x|) , ϕ = −log |x|, ϕ = − log(|x|)ϕ (x)dx
−∞
∞ 0
=− log(x)ϕ (x)dx − log(−x)ϕ (x)dx
0 −∞
∞ ∞
=− log(x)(ϕ (x) + ϕ (−x))dx = − log(x)(ϕ (x) − ϕ (−x)) dx
0 0
∞
ϕ (x) − ϕ (−x) 1
= − log(x) [ϕ (x) − ϕ (−x)]∞
0 + dx = p. v. , ϕ
0 x x
|T, ϕ | ≤ c0 ∑ |ϕ |α ,β
|α |,|β |≤n0
for every ϕ ∈ S.
The space of tempered distributions is denoted by S . In addition, for Tk , T ∈ S the
S C
convergence Tk → T means that Tk , ϕ → T, ϕ for all ϕ ∈ S.
E ⊂ S ⊂ D .
It turns out that members of E have compact support, and they are therefore called
distributions with compact support. But more on that later.
|T f , ϕ | = f ϕ dx ≤ f L1 ϕ L∞ .
Rn −δ δ
This means that R f ϕ dx is well defined in this case and
T f , ϕ := f ϕ dx.
Rn
where 1
p + 1
p = 1. This follows from Hölder’s inequality
1
p
(1 + |x|)−δ | f (x)|dx ≤ (1 + |x|)−δ p dx f L p .
R R
(3) Let T ∈ S and ϕ0 (x) ∈ C0∞ (Rn ) with ϕ0 (0) = 1. The product ϕ0 k T is well
defined in S by x
x
ϕ0 T, ϕ := T, ϕ0 ϕ .
k k
Now we are ready to prove a more serious and more useful fact.
Theorem 18.18. Let T ∈ S . Then there exists Tk ∈ S such that
Tk , ϕ = Tk (x)ϕ (x)dx → T, ϕ , k → ∞,
Rn
S
where ϕ ∈ S. In short, S = S .
Proof. Let j(x) be a function from D ≡ C0∞ (Rn ) with Rn j(x)dx = 1 and j(−x) =
j(x). Let jk (x) := kn j(kx). By Lemma 16.11 we have
lim jk , ϕ = lim jk (x)ϕ (x)dx = ϕ (0)
k→∞ k→∞ Rn
S
for every ϕ ∈ S. That is, jk (x) → δ (x).
The convolution of two integrable functions g and ϕ is defined by
(g ∗ ϕ )(x) := g(x − y)ϕ (y)dy.
Rn
If h and g are integrable functions and ϕ ∈ S, then it follows from Fubini’s theorem
that
h ∗ g, ϕ = ϕ (x)dx h(x − y)g(y)dy = g(y)dy h(x − y)ϕ (x)dx
R R Rn Rn
n n
L T, ϕ := T, Lϕ , T ∈ S .
ϕ
LTk , ϕ = Tk , L ϕ → T, L ϕ =: LT,
as k → ∞.
n
1 = (2π ) 2 δ
in S .
(2) δ = (2π )− 2 · 1, since for ϕ ∈ S we have
n
δ, ϕ = δ , F ϕ = F ϕ (0) = (2π )− 2
n n
e−i(0,x) ϕ (x)dx = (2π )− 2 1, ϕ .
Rn
n
Moreover, F −1 δ = (2π )− 2 · 1 in S .
x2 ξ2
(3) e−a 2 = a− 2 e− 2a , Re a ≥ 0, a
= 0. Indeed, for a > 0 we know that
n
√
x2 ( ax)2 n ξ2
F (e−a 2 ) = F (e− 2 ) = a− 2 e− 2a .
1 n
and u, f ∈ S . This equation can be solved in S using the Fourier transform.
Indeed, we get
u = f,
(1 + |ξ |2 )
or
u = (1 + |ξ |2 )−1 f,
or
u = F −1 ((1 + |ξ |2 )−1 F f ).
18 Tempered Distributions 161
P(D) = ∑ aα Dα
|α |≤m
Corollary 18.24. If Δ u = 0 in S and |u| is less than or equal to some constant, then
u is constant.
1 1
:= lim
x ± i0 ε →0+ x ± iε
In a similar fashion,
1 1
:= lim
(x ± i0)2 ε →0+ (x ± iε )2
and √
1 i 2π H(−ξ )eεξ , ξ
= 0,
(ξ ) =
x − iε i π2 , ξ = 0.
Hence √
1 1 −i 2π H(ξ ), ξ
= 0,
= lim =
x + i0 ε →0+ x + iε −i π2 , ξ = 0,
and √
1 1 i 2π H(−ξ ), ξ
= 0,
= lim =
x − i0 ε →0+ x − iε i π2 , ξ = 0.
and thus
1 1 1
+ = 2 p. v. .
x + i0 x − i0 x
In a similar fashion,
1 1 √ √ √
− = i 2π · 1 = i 2π 2π δ = 2π iδ,
x − i0 x + i0
and so
1 1
− = 2π iδ .
x − i0 x + i0
1 1 1 1
= p. v. − iπδ and = p. v. + iπδ .
x + i0 x x − i0 x
(2)
1 1 1 1 1
+ = 2 p. v. 2 and − = −2π iδ ;
(x + i0)2 (x − i0)2 x (x − i0)2 (x + i0)2
18 Tempered Distributions 163
(3)
1 1 1 1
= p. v. 2 + π iδ and = p. v. 2 − π iδ ;
(x + i0)2 x (x − i0)2 x
(4)
π 1
log |x| = − p. v. ;
2 |ξ |
(5)
xβ = (2π )n/2 i|β | ∂ β δ .
(2)
i 1
ξ ) = − π p. v. .
sgn(
2
ξ
we get
∂
α δ = (2π )− 2 (iξ )α .
n
1 k k √ k (k)
δ
(k) = √ iξ , xk = 2π i δ .
2π
and
1 ∂ ∂
∂ := −i
2 ∂x ∂y
in R2 .
164 Part II: Fourier Transform and Distributions
is the fundamental solution (see Chapter 22) of ∂¯ . Taking the Fourier transform of
(2) gives us
1
∂ (ξ ) = π δ(ξ ),
z
which is equivalent to
1
1 1
(iξ1 − ξ2 ) · (ξ ) = π · (2π )−1 · 1 = ,
2 z 2
or
1 1 1
(ξ ) = = −i , ξ
= 0.
z iξ1 − ξ2 ξ1 + iξ2
Let us check that this is indeed the case. We have, by Example 17.11,
∞ −iξ1 x
1 1 e−i(ξ1 x+ξ2 y) 1 ∞ −iξ2 y e
(ξ ) = dxdy = e dy dx
z 2π R2 x + iy 2π −∞ −∞ x + iy
∞ e−iξ1 x ∞ e−iξ1 x
1 ∞ −iξ2 y 1 0 −iξ2 y
= e dy dx + e dy dx
2π 0 −∞ x + iy 2π −∞ −∞ x + iy
∞ √ √
1
= e−iξ2 y 2π (−i 2π H(ξ1 )e−yξ1 )dy
2π 0
1 0 −iξ2 y √ √
+ e 2π (i 2π H(−ξ1 )e−yξ1 )dy
2π −∞
∞ 0
−y(ξ1 +iξ2 ) −y(ξ1 +iξ2 )
= −i H(ξ1 ) e dy − H(−ξ1 ) e dy .
0 −∞
Hence
1 i
(ξ ) = − ,
z ξ1 + iξ2
Let us consider first the direct product of distributions. Let us assume that T1 , . . . , Tn
are one-dimensional tempered distributions, T j ∈ S (R), j = 1, 2, . . . , n. The product
T1 (x1 ) · · · Tn (xn ) can be formally defined by
T1 (x1 ) · · · Tn (xn ), ϕ (x1 , . . . , xn ) = T1 (x1 ) · · · Tn−1 (xn−1 ), ϕ1 (x1 , . . . , xn−1 )
= T1 (x1 ) · · · Tn−2 (xn−2 ), ϕ2 (x1 , . . . , xn−2 )
= · · · = T1 (x1 ), ϕn−1 (x1 ),
where
But the product T1 (x)T2 (x), where the x are the same, in general case does not exist,
that is, it is impossible to define such a product. We remedy this by recalling the
following definition.
If n = 1, then
∞ −∞
(ϕ ∗ ψ )(x) = ϕ (x − y)ψ (y)dy = − ϕ (z)ψ (x − z)dz
−∞ ∞
∞
= ψ (x − z)ϕ (z)dz = (ψ ∗ ϕ )(x).
−∞
(2) It is also clear that the convolution is well defined for ϕ and ψ from S, and
moreover, for every α ≥ 0,
∂xα (ϕ ∗ ψ )(x) = (∂ α ϕ ∗ ψ )(x) = ∂xα ϕ (x − y)ψ (y)dy
Rn
= (−1)|α | ∂yα ϕ (x − y)ψ (y)dy
Rn
= (−1)2|α | ϕ (x − y)∂yα ψ (y)dy = (ϕ ∗ ∂ α ψ )(x),
Rn
where we integrated by parts and used the fact that ∂x j ϕ (x − y) = −∂y j ϕ (x − y).
We would like to prove that for ϕ and ψ from S it follows that ϕ ∗ ψ from S also. In
fact,
(1) ϕ ∗ ψ ∈ C∞ (Rn ) since ∂ α (ϕ ∗ ψ ) = ϕ ∗ ∂ α ψ and ∂ α : S → S.
(2) ϕ ∗ ψ decreases at infinity faster than any inverse power:
1
ϕ (x − y)ψ (y)dy ≤ c1 |ψ (y)|dy + c |ψ (y)|dy
Rn |y|≤ 2 |x − y|
|x| m 2 |x|
|y|> 2
c1
≤ |ψ (y)|dy + c2 |y|−m |y|m |ψ (y)|dy
|x|m |y|≤ 2
|x| |x|
|y|> 2
c1 c2
≤ + = c|x|−m , m ∈ N.
|x|m |x|m
ϕ ∗ ψ
L∞ (Rn ) ≤
ϕ
L p (Rn ) ·
ψ
L p (Rn ) , (19.1)
where 1
p + 1
p = 1, 1 ≤ p ≤ ∞. This means that the convolution is well defined
even for ϕ ∈ L p (Rn ) and ψ ∈ L p (Rn ). In particular,
19 Convolutions in S and S 169
ϕ ∗ ψ
L∞ (Rn ) ≤
ϕ
L1 (Rn ) ·
ψ
L∞ (Rn ) . (19.2)
(3) Interpolating (19.2) and (19.3) leads us to (see the Riesz–Thorin theorem, The-
orem 17.7)
ϕ ∗ ψ
L p ≤
ψ
L1 ·
ϕ
L p . (19.4)
(4) Interpolating (19.1) and (19.4) leads us to (again by the Riesz–Thorin theorem)
ϕ ∗ ψ
Ls ≤
ψ
Lr ·
ϕ
L p ,
where
1 1 1
1+ = + .
s r p
1 1
T : L p (Rn ) → L∞ (Rn ), + =1
p p
and
T : L1 (Rn ) → L p (Rn );
T : Lr (Rn ) → Ls (Rn ),
where
1 θ 1−θ θ
= + = 1−
r p 1 p
and
1 θ 1−θ 1 θ
= + = − .
s ∞ p p p
This gives
1 1 1
− = 1− .
r s p
e−i(x,ξ ) ϕ (x − y)dx
n
= (2π )− 2 ψ (y)dy
Rn Rn
ψ (y)e−i(y,ξ ) dy ϕ (z)e−i(z,ξ ) dz = (2π ) 2 F ϕ · F ψ ,
n n
= (2π )− 2
Rn Rn
i.e., n
∗ ψ = (2π ) 2 ϕ · ψ
ϕ .
Similarly, n
F −1 (ϕ ∗ ψ ) = (2π ) 2 F −1 ϕ · F −1 ψ .
Hence n
ϕ ∗ ψ = (2π ) 2 F (F −1 ϕ · F −1 ψ ),
or n
ϕ · ψ = (2π )− 2 ϕ ∗ ψ
.
L2 (Rn )
lim ϕε ∗ ψ = ψ.
ε →0+
in L2 . But
x
−n
−n −n −n 1 L∞
ε = ε
ϕ ϕ = ε σ 1 ϕ (ξ ) = ε ϕ(εξ ) = ϕ(εξ ) → ϕ(0)
ε ε ε
19 Convolutions in S and S 171
Hence
n L2
ϕ (εξ ) · ψ
ε ∗ ψ = (2π ) 2 ϕ (ξ ) → ψ
(ξ ), ε → 0+.
L2
ϕε ∗ ψ → ψ
as ε → 0+.
Theorem 19.3. For every fixed function ϕ from S(Rn ) the map ϕ ∗ T has, as a
continuous linear map from S to S (with respect to T ), a unique continuous linear
extension as a map from S to S (with respect to T ) as follows:
ϕ ∗ T, ψ := T, Rϕ ∗ ψ ,
(2) ∂ (ϕ ∗ T ) = ∂ α ϕ ∗ T = ϕ ∗ ∂ α T .
α
Proof. Let us assume that ϕ , ψ , and T belong to the Schwartz space S(Rn ). Then
we have checked already the properties (1) and (2) above. But we can easily check
that for such functions the definition is also true. In fact,
ϕ ∗ T, ψ = (ϕ ∗ T )(x)ψ (x)dx = ϕ (x − y)T (y)dyψ (x)dx
R Rn Rn
n
= T (y) ϕ (x − y)ψ (x)dxdy
R Rn
n
= T (y)dy Rϕ (y − x)ψ (x)dx = T, Rϕ ∗ ψ .
Rn Rn
S
For the case T ∈ S the statement of this theorem follows from the fact that S = S
(see Theorem 18.18).
Example 19.5.
(1) It is true that δ ∗ ϕ = ϕ . Indeed,
δ ∗ ϕ , ψ = δ , Rϕ ∗ ψ = (Rϕ ∗ ψ )(0) = ϕ (y)ψ (y)dy = ϕ , ψ .
Rn
∗ ϕ = (2π ) 2 δ · ϕ = 1 · ϕ = ϕ
δ
n
is equivalent to
δ ∗ϕ = ϕ
in S .
(2) Property (2) of Theorem 19.3 and part (1) of this example imply that
∂ α (δ ∗ ϕ ) = δ ∗ ∂ α ϕ = ∂ α ϕ .
(3) Let us consider again the equation (1 − Δ )u = f for u and f ∈ L2 (or even from
u = f is still valid in L2 and u = (1 + |ξ |2 )−1 f or
S ). Then (1 + |ξ |2 )
−1 1
u(x) = F f
1 + |ξ |2
− 2n −1 1
= (2π ) F ∗f= K(x − y) f (y)dy,
1 + |ξ |2 Rn
where
1 ei(x−y,ξ )
K(x − y) := dξ .
(2π )n Rn 1 + |ξ |2
This is the inverse Fourier transform of a locally integrable function. This func-
tion K is the free space Green’s function of the operator 1 − Δ in Rn . We will
calculate this integral precisely in Chapter 22.
x 19.6. Let j(x) be a function from L (R ) with j(x)dx = 1. Set jε (x) =
Lemma 1 n
Rn
−n
ε j ε , ε > 0. Then
jε ∗ f − f
L p → 0, ε → 0+,
for every function f ∈ L p (Rn ), 1 ≤ p < ∞. In the case p = ∞ we can state only the
fact
( jε ∗ f )gdx → f · gdx, ε → 0+
Rn Rn
Exercise 19.1. Prove Lemma 19.6 and find a counterexample showing that the first
part fails for p = ∞.
Remark 19.7. If j ∈ C0∞ (Rn ) or S(Rn ), then jε ∗ f ∈ C0∞ (Rn ) or S(Rn ) also for every
f ∈ L p (Rn ), 1 ≤ p < ∞.
Chapter 20
Sobolev Spaces
Lemma 20.1. For every function f ∈ L2 (Rn ) the following statements are
equivalent:
∂f
∂ x j (x) ∈ L (R ),
(1) 2 n
we have
ξ j f = D j f L2
L2
by Parseval’s equality.
1 t 1 eit ξ j − 1
Δ j f (ξ ) = (eit ξ j − 1) f(ξ ) = · ξ j f(ξ )
t t tξ j
holds. But
eit ξ j − 1
→i
tξ j
pointwise as t → 0. Hence
1 t L2
Δ f → iξ j f, t →0
t j
i.e. (again due to Parseval’s equality),
1 t L2 ∂ f
Δ f→ , t → 0.
t j ∂xj
L2
Let {gk } be a sequence in S such that gk → g and supp gk ⊂ {|ξ j | < 2}.
L2
ξ j h and
Let {hk } be a sequence in S such that hk → supp hk ⊂ {|ξ j | > 2 }.
1
hk L 2
fk (ξ ) = gk + → g(ξ ) + h(ξ ) = f(ξ ).
ξj
But
∂fk L2
= iξ j gk + ihk → iξ j (g + h) = iξ j f.
∂xj
This means that (by the Fourier inversion formula or Parseval’s equality)
∂ fk L2 −1 ∂f
→ F (iξ j f) = .
∂xj ∂xj
Lemma 20.2. Let f be a function from L2 (Rn ) and let s ∈ N. Then the following
statements are equivalent:
(1) Dα f ∈ L2 (Rn ), |α | ≤ s;
(2) ξ α f ∈ L2 (Rn ), |α | ≤ s;
Δhα f α α1 αn
α exists in L (R ), |α | ≤ s. Here Δ h f := (Δ h · · · Δ h ) f and h ∈ R with
(3) lim 2 n n
h→0 h 1 n
h j = 0 for all j = 1, 2, . . . , n.
L2
(4) There exists fk ∈ S such that fk → f and Dα fk has a limit in L2 (Rn ) for |α | ≤ s.
H s (Rn ) := { f ∈ L2 (Rn ) : ∑
Dα f
L2 < ∞}
|α |≤s
But it is easily seen that there are positive constants c1 and c2 such that
c1 (1 + |ξ |2 )s ≤ ∑ |ξ α |2 ≤ c2 (1 + |ξ |2 )s ,
|α |≤s
or
∑ |ξ α |2 (1 + |ξ |2 )s .
|α |≤s
178 Part II: Fourier Transform and Distributions
H s (Rn ) := { f ∈ S : (1 + |ξ |2 ) 2 f ∈ L2 (Rn )}
s
f ∈ H s (Rn )
if and only if
f ∈ Ls2 (Rn ).
Remark 20.9. Since 1 + |ξ |2s (1 + |ξ |2 )s , 0 < s < 1, the right-hand side of (20.1)
is an equivalent norm in H s (Rn ).
Proof. Denote by I the double integral appearing on the right-hand side of (20.1).
Then
I= | f (y + z) − f (y)|2 |z|−n−2s dydz
Rn Rn
|ei(z,ξ ) − 1|2
= |z|−n−2s dz |ei(z,ξ ) − 1|2 | f(ξ )|2 dξ = | f(ξ )|2 dξ dz
Rn Rn Rn Rn |z|n+2s
2vvT
A := I − , v = ξ − |ξ |e1 , ξ ∈ Rn ,
|v|2
Therefore,
| f (x)|2 dx + As | f (x) − f (y)|2 |x − y|−n−2s dxdy
Rn Rn Rn
= | f(ξ )|2 dξ + |ξ |2s | f(ξ )|2 dξ .
Rn Rn
Example 20.11. 1 ∈ / H s (Rn ) for all s. Indeed, assume that 1 ∈ H s0 (Rn ) for some s0
s0
(it is clear that s0 > 0). This means that (1 + |ξ |2 ) 2 1 ∈ L2 (Rn ). It follows from this
fact that 2 (Rn ), and further, 1 (Rn ). But n
1 ∈ Lloc 1 ∈ Lloc 1 = (2π ) 2 δ , and we know that
δ is not a regular distribution.
it follows that H s (Rn ) is also a separable Hilbert space, and the scalar product
can be defined by
( f , g)H s (Rn ) = (1 + |ξ |2 )s f· gdξ .
Rn
|( f , g)L2 (Rn ) |
f
H −s (Rn ) := sup .
0 =g∈H s (Rn )
g
H s (Rn )
20 Sobolev Spaces 181
This means that H −s (Rn ) is the dual space to H s (Rn ) with respect to the Hilbert
space L2 (Rn ). Indeed, since by Parseval’s equality
( f , g)L2 (Rn ) = ( f, g)L2 (Rn ) = ((1 + |ξ |2 )−s/2 f, (1 + |ξ |2 )s/2 g)L2 (Rn ) ,
it follows that
We have used here the fact that the space L−s2 (Rn ) is dual to L2 (Rn ) for every
s
s ∈ R.
(2) For −∞ < s < t < ∞, it follows that S ⊂ H t (Rn ) ⊂ H s (Rn ) ⊂ S .
1
ξ := (1 + |ξ |2 ) 2 ,
n
then δ ∈ H s (Rn ) is equivalent to (2π )− 2
ξ s ∈ L2 (Rn ), which in turn is equivalent
to s < − n2 .
(3) Let ϕ be a function from H s (Rn ), and ψ a function from H −s (Rn ). Then ϕ ∈
∈ L−s
Ls2 (Rn ) and ψ ·ψ
2 (Rn ), so that ϕ ∈ L1 (Rn ) by Hölder’s inequality. We
may therefore define (temporarily, and with slight abuse of notation)
ϕ , ψ L2 (Rn ) := ϕ · ψ
dξ
Rn
and obtain
|
ϕ , ψ L2 (Rn ) | ≤
ϕ
H s (Rn ) ·
ψ
H −s (Rn ) .
∂δ
For example, if ϕ is a function from H 2 +1+ε (Rn ), ε > 0, and ψ =
n
∂xj , then
∂δ ∂δ n
,ϕ = · ϕdξ = i(2π )− 2 ξ j ϕ(ξ )dξ
∂xj L2 (Rn ) Rn ∂xj Rn
P(D) = ∑ aα Dα .
|α |≤m
P(x, D) = ∑ aα (x)Dα
|α |≤m
with variable coefficients such that |aα (x)| ≤ c0 for all x ∈ Rn and |α | ≤ m. Then
Indeed,
α
P(x, D) f
L2 ≤ c0 ∑
Dα f
L2 = c0 ∑ ξ f
L2
|α |≤m |α |≤m
f
m
≤ c0 (1 + |ξ |2 ) 2 = c0
f
H m .
L2
Lemma 20.13. Let ϕ be a function from S, and f a function from H s (Rn ) for s ∈ R.
Then ϕ · f ∈ H s (Rn ) and
|s|
ϕ f
H s ≤ c (1 + |ξ |2 ) 2 ϕ 1 ·
f
H s .
L
Hence
ξ s
ξ ϕ ϕ(ξ − η )
η s f(η )dη .
n
s
· f = (2π )− 2
Rn
η s
ξ s
η |s| |s|
= ≤ 2 2
η − ξ |s| .
η s
ξ |s|
for all s ∈ R.
(6) Let us now consider distributions with compact support in greater detail than
what we saw in Chapter 18.
It can be proved that T ∈ E if and only if there exist c0 > 0, R0 > 0, and n0 ∈ N0
such that
|
T, ϕ | ≤ c0 ∑ sup |Dα ϕ (x)|
|α |≤n0 |x|≤R0
If we now set
T(ξ ) := (2π )−n/2
T, e−i(x,ξ ) ,
and hence T ∈ C∞ (Rn ). On the other hand, |
T, e−i(x,ξ ) | ≤ c0
ξ n0 implies that
|T(ξ )| ≤ c0
ξ n0 and hence T ∈ Ls2 (Rn ) for s < −n0 − n2 . So, by Exercise 20.2, we
may conclude that every T ∈ E belongs to H s (Rn ) for s < −n0 − n2 .
(7) We have the following lemma.
Lemma 20.15. The closure of C0∞ (Rn ) in the norm of H s (Rn ) is H s (Rn ) for all
Hs
s ∈ R. In short, C0∞ (Rn ) = H s (Rn ).
Proof. Let f be an arbitrary function from H s (Rn ) and let fR be a new function such
that
f(ξ ), |ξ | < R,
fR (ξ ) = χR (ξ ) f (ξ ) =
0, |ξ | > R.
Then fR (x) = F −1 (χR f)(x) = (2π )− 2 (F −1 χR ∗ f )(x). It follows from the above
n
n
H s ⊂ C˙k (Rn ), s > k+ .
2
What is more,
|ξ ||α |
|ξ f(ξ )|dξ ≤ c
α
|ξ | | f(ξ )|dξ = c
|α |
ξ s | f(ξ )|dξ
Rn
ξ
s
Rn Rn
1/2
1/2
|ξ |2|α | 2s
≤c dξ
ξ | f (ξ )| dξ2
≤ c
f
H s (Rn )
Rn
ξ 2s Rn
Remark 20.19. This counterexample shows us that the Sobolev embedding theorem
is sharp.
Lemma 20.20. Let us assume that ϕ and f from H s (Rn ) for s > n2 . Then F (ϕ f ) ∈
L1 (Rn ).
Proof. Since f , ϕ ∈ H s (Rn ), it follows that f, ϕ ∈ Ls2 (Rn ) for s > n2 . But this implies
(see Lemma 20.17) that f and ϕ ∈ L1 (Rn ) and
F (ϕ f ) = (2π )− 2 ϕ ∗ f
n
Exercise 20.7. Prove that Wp1 (Rn ) ·Wp1 (Rn ) ⊂ Wp1 (Rn ) if p > n.
Since
∞ ∞
1 1 1
dξ1 =
2 s dξ1
−∞ (1 + |ξ |2 + ξ1 )s
2 (1 + |ξ |2 )s −∞ ξ1
1+ (1+|ξ |2 )1/2
∞
dρ
= (1 + |ξ |2 )−s+1/2 , s > 1/2,
−∞ (1 + ρ 2 )s
we have
∞
(1 + |ξ |2 )−s+1/2 | f(ξ )|2 ≤ Cs u(ξ )|2 (1 + |ξ |2 )s dξ1 ,
|
−∞
where Cs denotes the latter convergent integral with respect to ρ . Integrating with
respect to ξ in the latter inequality leads to
f
2H s−1/2 (Rn−1 ) ≤ Cs
u
2H s (Rn ) .
188 Part II: Fourier Transform and Distributions
This means that the first part of this proposition is proved, since S(Rn ) is dense in
H s (Rn ). Surjectivity follows also. If g ∈ H s−1/2 (Rn−1 ), s > 1/2, we can define
(1 + |ξ |2 )s/2−1/4
u(ξ ) := g(ξ ) .
(1 + |ξ |2 )s/2
Then u := F −1 (
u(ξ )) defines u ∈ H s (Rn ) and u(0, x ) = Cg(x ) with some nonzero
constant C.
We define the space H0k (Ω ) as the completion of C0∞ (Ω ) with respect to the norm of
H k (Ω ).
Theorem 20.23 (Poincaré’s inequality). Suppose f ∈ H0k (Ω ), k ≥ 1. Then there is
a constant M > 0 such that
f
H k−1 (Ω ) ≤ M ∑ ∂ β f 2 . (20.3)
L (Ω )
|β |=k
and f ∈ C0∞ (Ω ) will continue to be identically zero outside of Ω . Then for all x ∈ Qn
we have x1
f (x) = ∂x1 f (ξ1 , x )dξ1 , x = (x1 , x ). (20.4)
−A
Qn
| f (x)|2 dx ≤ 4A2
Qn
|∂x1 f |2 dx ≤ 4A2 ∑ |∂ β f |2 dx.
|β |=1 Qn
Since C0∞ (Ω ) is dense in H01 (Ω ), the case k = 1 is established. Let us assume that
for all f ∈ H0k (Ω ), k ≥ 1, we have
β
f
H k−1 (Ω ) ≤ M ∑ ∂ f
L 2 (Ω )
.
|β |=k
Thus
n
∑ ∂ j f H k−1 (Ω ) ≤ M ∑
∂ γ f
L2 (Ω ) ,
j=1 |γ |=k+1
or
f
H k (Ω ) ≤ M ∑
∂ γ f
L2 (Ω ) .
|γ |=k+1
For all real s > 0 the space H s (Ω ) with fractional s can be obtained as the inter-
polation space between L2 (Ω ) and H k (Ω ) with some integer k ≥ 1 (see, e.g., [39,
p. 286] for details).
Proof. This fact can be proved also by induction on k ≥ 1 using (20.4). Since C0∞ (Ω )
is complete in H0k (Ω ), it follows that for k = 1 and x = (x1 , x ) ∈ ∂ Ω we have from
(20.4) that x1
f (x) = ∂x1 f (ξ1 , x )dξ1 = 0.
−A
Exercise 20.8. Prove Rellich’s theorem: if Ω is bounded and s > t ≥ 0, then the
inclusion map H0s (Ω ) → H0t (Ω ) is compact. Hint: Use the Ascoli–Arzelà theorem
(Theorem 34.7).
We may say that an element f ∈ H k (Rn+ ) is the restriction on Rn+ of some element
from H k (Rn ). More precisely, the following proposition holds.
E f
H k (Rn ) ≤ C
f
H k (Rn+ ) , (20.6)
Proof. Since S(Rn+ ) is dense in H k (Rn+ ) (see, for example, Lemma 20.15), it follows
that for every f ∈ S(Rn+ ) and integer k ≥ 0 we may define E as
f (x), x1 ≥ 0,
E f (x) =
j=1 a j f (− jx1 , x ),
∑k+1 x1 < 0,
k+1
∑ a j (− j)l = 1, l = 0, 1, . . . , k.
j=1
The determinant of this system is the well known Vandermonde polynomial, and it
is not equal to zero. Hence, this system has a unique solution with respect to the
coefficients a j , j = 1, . . . , k − 1. After these coefficients have been determined, the
inequality (20.6) follows immediately.
Remark 20.26. For arbitrary s ≥ 0 the result of Proposition 20.25 can be obtained
by interpolation of Sobolev spaces H s ; see [39, p. 285].
Let Ω ⊂ Rn be a bounded domain with a C∞ boundary ∂ Ω . Since ∂ Ω is a compact
set, it can be covered by finitely many open sets U j , j = 1, . . . , m, such that
m
∂Ω ⊂ Uj
j=1
and
U j ∩ ∂ Ω → {y ∈ Rn : |y| < 1, y1 = 0}.
For these purposes we may use the extension operator E from Proposition 20.25
such that (s ≥ 0)
m
E : H s (Ω ) → H s (U), U= Uj.
j=1
y1
∂Ω 1
Uj
y
Proposition 20.29. For s > 1/2 the map τ extends uniquely to a continuous linear
map
τ : H s (Ω ) → H s−1/2 (∂ Ω )
τ u
H s−1/2 (∂ Ω ) ≤ C
u
H s (Ω ) .
Proof. The proof is based on Proposition 20.22, the definition of H s (∂ Ω ), and the
following diagram:
τ
H s (Ω) H s−1/2 (∂ Ω)
E in local coordinates
τ
H s (U) H s−1/2 (Rn−1 )
See [39, p. 287] for details.
Chapter 21
Homogeneous Distributions
We begin this chapter with the Fourier transform of a radially symmetric function.
Lemma 21.1. Let f (x) be a radially symmetric function in Rn , i.e., f (x) = f1 (|x|).
Let us assume also that f (x) ∈ L1 (Rn ). Then the Fourier transform f(ξ ) is also
radial and ∞
f(ξ ) = |ξ |1− 2
n n
f1 (r)r 2 J n−2 (r|ξ |)dr,
0 2
R∞
n
e−i|ξ |r(ϕ ,θ ) dθ ,
n
−2
= (2π ) f1 (r)rn−1 dr
0 Sn−1
n−1 π
2π 2
e−i|ξ |r(ϕ ,θ ) dθ = e−i|ξ |r cos ψ (sin ψ )n−2 dψ ,
Sn−1 Γ ( n−1
2 ) 0
where Γ is the gamma function. This fact implies that f(ξ ) is a radial function,
since the last integral depends only on |ξ |. A property of Bessel functions [23] is
that
π J n−2 (r|ξ |)
√ n−1
e−i|ξ |r cos ψ (sin ψ )n−2 dψ = 2 2 −1 πΓ
n
2
n−2 . (21.1)
0 2 (r|ξ |) 2
Remark 21.2. If we put the variable u = cos ψ in the integral I appearing in (21.1),
then we obtain
π 1
I= e−i|ξ |r cos ψ (sin ψ )n−2 dψ = e−i|ξ |ru ( 1 − u2 )n−3 du.
0 −1
1
sin(|ξ |r) √ J 12 (r|ξ |)
I= e−i|ξ |ru du = 2 = 2π 1 ,
−1 |ξ |r (r|ξ |) 2
i.e.,
2 sin(|ξ |r)
J 1 (r|ξ |) = .
2 π (|ξ |r) 21
If n = 2, then
1 −i|ξ |ru
e
I= √ du = π J0 (r|ξ |),
−1 1 − u2
i.e.,
1 −i|ξ |ru
1 e
J0 (r|ξ |) = √ du.
π −1 1 − u2
Remark 21.3. For later considerations we state the small- and large-argument
asymptotics of Jν for ν > −1 as
cν |x|ν , |x| → 0+,
Jν (|x|) ≈ √1
cν cos(Aν |x| + Bν ), |x| → +∞
|x|
(see [23]).
Let us assume now that n = 1, 2, 3, 4. Then the last integral can be understood in the
classical sense. It follows from Lemma 21.1 that
∞ r 2n J n−2 (r|x|)dr
1 (|x|) = (2π )
K1 (x) = K − 2n
|x|1− 2n 2
0 1 + r2
∞ ρ n2 J n−2 (ρ )dρ
− 2n
= (2π ) |x|
2−n 2
.
0 ρ 2 + |x|2
∞ ρ n2 J n−2 (ρ )dρ
− 2n
K−1 (|x|) = (2π ) |x| 2−n 2
.
0 ρ 2 − |x|2
But there is a problem with the convergence of this integral near ρ = |x|. Therefore,
this integral must be regularized as
∞ ρ n2 J n−2 (ρ )dρ
lim 2
.
ε →0+ 0 ρ 2 − |x|2 − iε
Recall that
(1) σλ f (x) := f (λ x), λ = 0 and
(2)
σλ T, ϕ := λ −n
T, σ 1 ϕ , λ > 0.
λ
σλ T, ϕ = λ m
T, ϕ ,
196 Part II: Fourier Transform and Distributions
or
T, ϕ = λ −n−m
T, σ 1 ϕ ,
λ
σλ T, ϕ = λ −n
T, σ 1 ϕ = λ −n
T, σ
1 ϕ = λ
−n
T, λ n σλ ϕ
λ λ
=
T, σλ ϕ = λ −n
σ 1 T, ϕ = λ −n λ −m
T, ϕ = λ −n−m
T, ϕ
λ
for all ϕ ∈ S.
Definition 21.7. We set Hm∗ (Rn ) := {T ∈ Hm (Rn ) : T ∈ C∞ (Rn \ {0})}.
Exercise 21.3. Prove that
(1) if T ∈ Hm∗ , then Dα T ∈ Hm−|
∗ α ∗
α | and x T ∈ Hm+|α | ;
∗ ∗
(2) F : Hm → H−m−n .
Exercise 21.4. Let ρ (x) be a function from C∞ (Rn ) with |Dα ρ (x)| ≤ c
xm−|α | for
all α ≥ 0 and m ∈ R. Prove that ρ(ξ ) ∈ C∞ (Rn \ {0}) and (1 − ϕ )ρ ∈ S, where
ϕ ∈ C0∞ (Rn ) and ϕ ≡ 1 in Uδ (0).
∗ (Rn ). Indeed,
Example 21.8. (1) δ ∈ H−n
σλ δ , ϕ = λ −n
δ , σ 1 ϕ = λ −n σ 1 ϕ (0) = λ −n ϕ (0) = λ −n
δ , ϕ .
λ λ
But supp δ = {0}. This means that δ ∈ C∞ (Rn \ {0}). Alternatively, one could
note that
δ = (2π )− 2 · 1 ∈ H0∗ (Rn )
n
(3) Let now m = −n in part (2) and in addition assume that Sn−1 ω (θ )dθ = 0. Note
that T−n (x) ∈
/ Lloc
1 (Rn ). But we can define T
−n as a distribution from S by
p. v. T−n , ϕ := T−n (x)[ϕ (x) − ϕ (0)ψ (|x|)]dx,
Rn
where ϕ ∈ S(Rn ) and ψ ∈ S(R) with ψ (0) = 1. We assume that ψ is fixed. But
it is clear that this definition does not depend on ψ , because Sn−1 ω (θ )dθ = 0.
where T−n = |x|−n ω |x| , Sn−1 ω (θ )dθ = 0.
x
Hence p.
v. T−n ∈ L∞ (Rn ) by duality.
Finally, if f ∈ L2 (Rn ), then
v. T−n · f,
F (p. v. T−n ∗ f ) = (2π ) 2 p.
n
Next we want to consider a more difficult case than the previous one. Define
1
p. v. , ϕ := |x|−n [ϕ (x) − ϕ (0)ψ (|x|)]dx, (21.2)
|x|n Rn
where ϕ ∈ S and ψ ∈ S with ψ (0) = 1. But now we don’t have the condition
Sn−1 ω (θ )dθ = 0 as above. Therefore, (21.2) must depend on the function ψ (|x|).
We will try to choose an appropriate function ψ . Applying the operator σλ , we get
1 1
σλ p. v. n , ϕ =
p. v. n , λ −n σ 1 ϕ
|x| |x| λ
x
= |x|−n λ −n ϕ − ϕ (0)ψ (|x|) dx
Rn λ
= λ −n |y|−n [ϕ (y) − ϕ (0)ψ (λ |y|)]dy
Rn
= λ −n |y|−n [ϕ (y) − ϕ (0)ψ (|y|)]dy
Rn
− λ −n |y|−n ϕ (0)[ψ (λ |y|) − ψ (|y|)]dy
Rn
1
=
λ −n p. v. , ϕ + Rest,
|x|n
where
Rest = −λ −n ϕ (0) |y|−n [ψ (λ |y|) − ψ (|y|)]dy
Rn
∞
ψ (λ r) − ψ (r)
= −λ −n
δ , ϕ dr dθ
r
0 Sn−1
∞
ψ (λ r) − ψ (r)
= −ωn λ −n
δ , ϕ dr,
0 r
n
and ωn = Γ2π( n2) is the area of the unit sphere Sn−1 . Let us denote the last integral by
2
G(λ ), λ > 0. Then
∞ ∞
1 1 1
G (λ ) = ψ (λ r)dr = ψ (t)dt = − ψ (0) = − .
0 λ 0 λ λ
We also have that G(1) = 0. We may therefore conclude that G(λ ) = − log λ , which
implies that
Rest = ωn λ −n log λ
δ , ϕ ,
21 Homogeneous Distributions 199
and so
1 1
σλ p. v. n = λ −n p. v. n + ωn λ −n log λ · δ (x).
|x| |x|
or
1 1 n
λ −n σ 1 F p. v. n = λ −n F p. v. n + (2π )− 2 ωn λ −n log λ ,
λ |x| |x|
or
1 ξ 1 n
F p. v. n = F p. v. n (ξ ) + (2π )− 2 ωn log λ .
|x| λ |x|
Since p. v. |x|1n for such ψ is a radial homogeneous distribution, we must have that
1
F p. v. n
|x|
ξ
is also a radial homogeneous distribution. Therefore, F p. v. |x|1n |ξ |
depends
only on |ξξ | = 1. So this term is a constant that depends on the choice of ψ . We
will choose our function ψ (|x|) so that this constant is zero. Then finally,
1 n
F p. v. n (ξ ) = −(2π )− 2 ωn log |ξ |.
|x|
Now let us consider T−m = |x|−m , 0 < m < n. It is clear that |x|−m ∈ Lloc
1 (Rn ). Thus
∞ r n2 J n−2 (r|ξ |) ∞
|x|−m = |ξ | 1− 2n 2
dr = |ξ |−n+m
n
ρ 2 −m J n−2 (ρ )dρ .
0 rm 0 2
200 Part II: Fourier Transform and Distributions
−m = C
n−1
|x| n,m |ξ | , < m < n.
m−n
2
In fact, this is true even for m such that 0 < Re(m) < n, which follows by analytic
continuation on m. In order to calculate the constant Cn,m , let us apply this distribu-
|x|2
tion to ϕ = e− 2 . Since ϕ = ϕ , we get
|x|2 |ξ |2
|x|−m , e− 2 =
Cn,m |ξ |m−n , e− 2 .
|ξ |2 n−(n−m)−2
m
Cn,m
|ξ |m−n , e− 2 = Cn,m 2 2 ωnΓ .
2
Therefore,
m−2
m n−m−2 n−m
Cn,m 2 2 ωnΓ =2 2 ωnΓ ,
2 2
which gives us
n −m Γ n−m
Cn,m = 2 2 .
2
Γ m2
Finally, we have
n
−m = 2 2 −m
Γ n−m
|x| m2 · |ξ |m−n . (21.3)
Γ 2
i.e.,
1 f (t)dt
H f (x) = lim , x ∈ R.
π ε →0+ |x−t|≥ε x−t
21 Homogeneous Distributions 201
Remark 21.12. We can rewrite R j (x) in the form R j (x) = |x|−n ω j (x), where ω j (x) =
xj
|x| and conclude that
(1) Sn−1 ω j (θ )dθ = 0;
(2) R j (λ x) = λ −n R j (x), λ > 0.
R j ∗ f = p. v. R j ∗ f ,
because in our previous notation, R j (x) = T−n ∈ H−n∗ (Rn ) is a homogeneous distri-
bution. Let us calculate the Fourier transform of the Riesz kernels. By homogeneity,
it suffices to consider |ξ | = 1. We have
e−i(x,ξ ) x j
Rj (ξ ) = p.
n
v. R j (ξ ) = (2π )− 2 dx
Rn |x|n+1
n e−i(x,ξ ) x j
= lim (2π )− 2 dx.
ε →0+ ε <|x|<μ |x|n+1
μ →+∞
We split
e−i(x,ξ ) x j e−i(x,ξ ) x j e−i(x,ξ ) x j
dx = dx + dx =: I1 + I2 .
ε <|x|<μ |x|n+1 ε <|x|<1 |x|
n+1
1<|x|<μ |x|n+1
202 Part II: Fourier Transform and Distributions
But
−i(x,ξ ) ∂ ∂
x je dσ = i e−i(x,ξ ) dσ = i cos(|ξ | · x1 )dσ
|x|=1 ∂ ξ j |x|=1 ∂ ξ j |x|=1
iξ j
=− x1 · sin(|ξ | · x1 )dσ = −iξ j ·C1 , |ξ | = 1,
|ξ | |x|=1
where we have used the fact that a rotation maps ξ to (|ξ |, 0, . . . , 0). Similarly, we
may conclude that
e−i(x,ξ )
dx = cos(|ξ |x1 )|x|1−n dx = C2 , |ξ | = 1.
|x|<1 |x|
n−1
|x|<1
I1 → Cn iξ j , |ξ | = 1
p. v. R j = iξ j ·Cn
21 Homogeneous Distributions 203
for |ξ | = 1. But we know from Exercise 21.3 that p. v. R j ∈ H0∗ (Rn ). We conclude
ξ
that p. v. R j (ξ ) = iCn j . Moreover, we have
|ξ |
R
n
j · f = iCn ξ j f,
j ∗ f = (2π ) 2 R
|ξ |
or
ξj
Rj ∗ f = iCn F −1 f .
|ξ |
i.e.,
R j ∗ : L2 (Rn ) → L2 (Rn ),
where by (21.3),
1
I1 (x) = cn .
|x|n−1
204 Part II: Fourier Transform and Distributions
Therefore, we have
f (y)dy
I −1 f (x) = cn ,
Rn |x − y|n−1
where
1 Γ ((n − 1)/2)
cn = .
2 π (n+1)/2
∂
It is straightforward to verify that ∂ x j I1 = cn R j and hence
∂ −1
I f = cn R j ∗ f .
∂xj
for some s and σ . Since R j ∗ is a bounded map from L2 (Rn ) to L2 (Rn ), we may
conclude that
∂ −1 2 n
I : L (R ) → L2 (Rn ). (21.4)
∂xj
Now let us assume for simplicity that n ≥ 3. Let us try to prove that
if and only if
1
f ∈ L2 (Rn ).
|ξ |
Indeed,
2/r 1
−2 r2 r
|ξ | | f(ξ )|2 dξ ≤
−2
| f(ξ )|r dξ |ξ | dξ 2 < ∞,
|ξ |<1 |ξ |<1 |ξ |<1
21 Homogeneous Distributions 205
r
since 2 > n−2
n
and < n2 . For |ξ | > 1 the function |ξ |−1 f(ξ ) belongs to L2 (Rn ).
r
2
This fact follows from the inequality |ξ |−1 | f(ξ )| < | f(ξ )| and from the positivity
of σ (see Lemma 20.17). This proves (21.5) for σ > 1.
If we combine (21.4) and (21.5), we obtain that
Let us consider now Lσ∞ (Rn ) for σ > 1. If f ∈ Lσ∞ (Rn ), then | f (x)| ≤ C(1 + |x|)−σ
and thus
(1 + |y|)−σ dy
|I −1 f (x)| ≤ C < ∞.
Rn |x − y|n−1
If we recall the fact that R j ∗ : Ls (Rn ) → Ls (Rn ) for all 1 < s < ∞, then we have
L(x, D) = ∑ aα (x)Dα , x ∈ Rn ,
|α |≤m
LE, ϕ = E, L ϕ ,
Here, L ϕ must be in D(Ω ) for ϕ from D(Ω ). This will be the case, for example,
for aα (x) ∈ C∞ (Ω ).
Two fundamental solutions for L with the same parameter y differ by a solution
of the homogeneous equation Lu = 0. Unless boundary conditions are imposed,
the homogeneous equation will have many solutions, and the fundamental solution
will not be uniquely determined. In most problems there are grounds of symmetry
or causality for selecting the particular fundamental solution for the appropriate
physical behavior.
c Springer International Publishing AG 2017 207
V. Serov, Fourier Series, Fourier Transform and Their Applications
to Mathematical Physics, Applied Mathematical Sciences 197,
DOI 10.1007/978-3-319-65262-7 22
208 Part II: Fourier Transform and Distributions
We also observe that if L has constant coefficients, we can find the fundamental
solution in the form E(x|y) = E(x − y|0) := E(x − y). This fact follows from the
properties of the Fourier transform:
Lx E(x − y) = ∑ aα ξ α E(x − y) = ∑ aα ξ α e−i(ξ ,y) E(x)
|α |≤m |α |≤m
= e−i(ξ ,y) δ
(x) = δ
(x − y),
i.e.,
Lx E(x − y) = δ (x − y).
Exercise 22.1. Let L be a differential operator with constant coefficients. Prove that
u = q ∗ E = E ∗ q solves the inhomogeneous equation
Lu = q
in D .
Remark 22.2. In many cases the fundamental solution is a function. We can there-
fore write u as an integral
u(x) = E(x − y)q(y)dy.
Ω
Remark 22.3. In order for the convolution product E ∗q (or q∗E) to be well defined,
we have to assume that, for example, q vanishes outside a finite sphere.
Remark 22.4. If L does not have constant coefficients, we can no longer appeal to
convolution products; instead, one can often show that
u(x) = E(x|y)q(y)dy.
Ω
Definition 22.5. We denote by a0 (x, ξ ) the main (or principal) symbol of L(x, D)
a0 (x, ξ ) = ∑ aα (x)ξ α , ξ ∈ Rn .
|α |=m
Assume that the aα (x) are “smooth.” An operator L(x, D) is said to be elliptic in Ω
if for every x ∈ Ω and ξ ∈ Rn \{0} it follows that
a0 (x, ξ ) = 0.
Exercise 22.2. Let aα (x) be real for |α | = m. Prove that the previous definition is
equivalent to
22 Fundamental Solution of the Helmholtz Operator 209
(1) m is even,
(2) a0 (x, ξ ) ≥ CK |ξ |m (or −a0 (x, ξ ) ≥ CK |ξ |m ), CK > 0, for every compact set K ⊂
Ω and for all ξ ∈ Rn and x ∈ K.
Let us consider the heat equation
⎧
⎨∂u
= Δ u, t > 0, x ∈ Rn ,
∂t
⎩u(x, 0) = f (x), x ∈ Rn
This initial value problem for an ordinary differential equation has the solution
Hence
where
1 |x|2
P(x,t) = (2π )−n e−t|ξ | ei(x,ξ ) dξ = − 4t
2
n e .
Rn (4π t) 2
Definition 22.6. The function P(x,t) is the fundamental solution of the heat equa-
tion and satisfies ⎧
⎪ ∂
⎨ − Δ P(x,t) = 0, t > 0,
∂t
⎪
⎩ lim P(x,t) = S
δ (x).
t→0+
with constant coefficients. Assume that L(ξ ) = ∑|α |≤m aα ξ α > 0 for all ξ ∈ Rn \{0}.
If we consider PL (x,t) as a solution of
⎧
⎪ ∂
⎨ + L(D) PL (x,t) = 0, t > 0,
∂t
⎪ S
⎩ lim PL (x,t) = δ (x),
t→0+
∂
then PL (x,t) is the fundamental solution of ∂t + L(D) and can be calculated by
PL (x,t) = (2π )−n e−tL(ξ ) ei(x,ξ ) dξ .
Rn
Therefore,
∞
(L(D) + λ )F, ϕ = lim e−λ t (L(D) + λ )PL , ϕ dt
ε →0+ ε
∞ ∞
= lim e−λ t L(D)PL , ϕ dt + λ e−λ t PL , ϕ dt
ε →0+ ε 0
∞
∂
= lim e−λ t − PL , ϕ dt + λ F, ϕ
ε →0+ ε ∂t
∞
= lim −e−λ t PL , ϕ |∞
ε − λ e−λ t PL , ϕ dt + λ F, ϕ
ε →0+ ε
−λ ε
= lim e PL (·, ε ), ϕ = δ , ϕ
ε →0+
for all ϕ ∈ S.
∂
Exercise 22.3. Let us define a fundamental solution Γ (x,t) of ∂t + L(D) as a solu-
tion of
( ∂∂t + L)Γ (x,t) = δ (x)δ (t),
Γ (x, 0) = 0.
22 Fundamental Solution of the Helmholtz Operator 211
Prove that ∞
F(x, λ ) := e−λ t Γ (x,t)dt
0
√
where r = λ |x|. From our previous considerations we know that
1
F(x, λ ) = (2π )−n/2 F −1 (x),
|ξ |2 + λ
It is known that
π i iπ ν (1)
Kν (r) = e 2 Hν (ir), r > 0,
2
(1)
where Hν is the Hankel function of first kind of order ν .
Next we want to obtain estimates for F(x, λ ) for x ∈ Rn , λ > 0, and n ≥ 1. Let
∞ −τ − r2 − n 1 ∞
4τ τ 2 dτ in two parts I1 + I2 =
us consider the integral 0 e 0 + 1 .
Since ⎧
⎪ −1
⎨r , n = 1,
I1 ∼ c log 1r , n = 2,
⎪
⎩
1, n ≥ 3,
as r → 0+, we have ⎧
⎪
⎨1, n = 1,
|I1 | ≤ cn log r , n = 2,
1
⎪
⎩ 2−n
r , n ≥ 3.
1 ∞
r2
r2
r−2 e− 4 , n = 1, 2, 3, 4
e− 4y y− 2 dy
n −z n2 −2
I1 ≤ = cn r 2−n
e z dz ≤ cn
0 r2 2−n
r e − δ r 2
, n ≥ 5,
4
n
where 0 < δ < 14 . The last inequality follows from the fact that z 2 −2 ≤ cε eε z for
2 − 2 > 0 and all ε > 0 (z > 1).
n
Since ∞
r2
I2 ≤ e−y− 4y dy,
1
2
we perform the change of variable z := y + 4y
r
. Then z ≥ r and z → +∞. Thus
∞ ∞
r2 z
e−y− 4y dy = c e−z 1 + √ dz
1 r z2 − r 2
∞ ∞
zdz
=c e−z dz + c e−z √
r r z2 − r 2
∞ ∞
= ce−r + c e−z z2 − r2 + e−z z2 − r2 dz
r r
∞
√
(2) If λ |x| > 1, then
√
|F(x, λ )| ≤ cn e−δ λ |x|
, n ≥ 1.
We will rewrite these estimates in a more appropriate form for all λ > 0 and x ∈
Rn as ⎧
⎪
⎨ λ, n = 1,
⎪ √1
√
−δ λ |x|
|F(x, λ )| ≤ cn e 1 + | log √ |, n = 2,
1
⎪
⎪ λ |x|
⎩|x|2−n , n ≥ 3.
Example 22.10. Recall from Chapter 21 that the solution of the equation
(−1 − Δ )u = f can be written in the form
−1 1
u(x) = K−1 ∗ f = F ∗ f,
|ξ |2 − 1
where
∞ ρ n2 J n−2 (ρ )dρ
K−1 (|x|) = cn |x|2−n
lim 2
.
ε →0+ 0 ρ 2 − |x|2 − iε
−Δ En − λ En = δ (x).
√ √
We define λ with nonnegative imaginary part, i.e., λ = α + iβ , where β ≥ 0
and β = 0 if and only if λ ∈ [0, +∞). We require that En is radially symmetric. Then
for x = 0, En must solve the equation
214 Part II: Fourier Transform and Distributions
(rn−1 u ) + λ rn−1 u = 0.
This equation can be reduced to one of Bessel type by making the substitution u =
n
wr1− 2 . A straightforward calculation shows that
n 2 w
(rw ) − 1 − + λ rw = 0,
2 r
or
w n 2 1
w + + λ − 1− w = 0,
r 2 r2
or
√
√
√ v (r λ ) n 2 1 √
v (r λ ) + √ + 1− 1− v(r λ ) = 0, w(r) = v(r λ ).
r λ 2 λ r2
This is the Bessel equation of order n2 − 1. Its two linearly independent solutions
are the Bessel functions J n2 −1 and Y 2n −1 of the first and second kinds, respectively.
Therefore the general solution is of the form
√ √
w(r) = c0 J n2 −1 ( λ r) + c1Y n2 −1 ( λ r).
For us it is convenient to write it in terms of Hankel functions of the first and second
kinds as √ √
(1) (2)
w(r) = c0 H n −1 ( λ r) + c1 H n −1 ( λ r),
2 2
where
(1) (2)
Hν (z) = Jν (z) + iYν (z), Hν (z) = Jν (z) − iYν (z).
√ (2) √
If λ ∈
/ [0, +∞), then λ has positive imaginary part, and the solution H n −1 ( λ r) is
(1) √
2
exponentially large at z = +∞, whereas H n −1 ( λ r) is exponentially small. Hence
2
we take √
1− 2n (1)
En (x, λ ) = c0 r H n −1 ( λ r).
2
or
∂ En
lim rn−1 ωn = 1,
r→0 ∂r
and
(1) 2i
H0 (r) ∼ log r.
π
It can be proved using Exercise 22.4 that
√ n−2
λ
2
i
c0 = .
4 2π
for all λ = 0. The formula (22.1) is valid also for λ ∈ (0, +∞). This fact follows
from the definition:
√ n−2
i λ + iε
2
(1)
En (x, λ ) = lim En (x, λ + iε ) = lim H n−2 ( λ + iε |x|)
ε →0+ 4 ε →0+ 2π |x| 2
√ n−2
i λ
2
(1)
√
= H n−2 ( λ |x|).
4 2π |x| 2
216 Part II: Fourier Transform and Distributions
√ n−2
1 i λ
2
(1)
√
(2π )−n/2 F −1 = H n−2 ( λ |x|),
|ξ | − λ − i0
2 4 2π |x| 2
where (−Δ − k2 − i0)−1 is an integral operator with kernel En (x, k) from the previ-
ous chapter and δ > 12 . In fact, this estimate allows us to consider the Hamiltonian
∞ -potentials only (if we want to preserve (2, 2)-estimates). But we would
with Lloc
p
like to consider the Hamiltonian with Lloc -potentials. We therefore need to prove
(p, q)-estimates.
1 1
We proved in Example 18.26 that the limit lim x−i ε := x−i0 exists in the sense
ε →0+
of tempered distributions and
1 1
= p. v. + iπδ (x),
x − i0 x
i.e.,
1 ϕ (x)
, ϕ = lim dx + iπϕ (0).
x − i0 δ →0+ |x|>δ x
1
(H(ξ ) − i0)−1 := lim
ε →0+ H(ξ ) − iε
c Springer International Publishing AG 2017 217
V. Serov, Fourier Series, Fourier Transform and Their Applications
to Mathematical Physics, Applied Mathematical Sciences 197,
DOI 10.1007/978-3-319-65262-7 23
218 Part II: Fourier Transform and Distributions
1
(H(ξ ) − i0)−1 = p. v. + iπδ (H(ξ ) = 0),
H(ξ )
n−2
|k| (1)
where G+
2
k (|x|) = H n−2 (|k||x|). On the other hand, we can write
i
4 2π |x|
2
n f(ξ )ei(x,ξ ) dξ
(−Δ − k2 − i0)−1 f = F −1 (F [(−Δ − k2 − i0)−1 f ]) = (2π )− 2
Rn |ξ |2 − k2 − i0
n
1 iπ (2π )− 2
f(ξ )ei(x,ξ ) dξ + δ (H) f(ξ )ei(x,ξ ) dξ
n
= (2π )− 2 p. v.
Rn |ξ |2 − k2 2k Rn
f (ξ )ei(x,ξ ) dξ iπ
f(kθ )eik(x,θ ) dθ
n
= (2π )− 2 p. v. +
|ξ |2 − k2
n
Rn 2k(2π ) 2 Sn−1
f (ξ )ei(x,ξ ) dξ iπ
eik(θ ,x−y) dθ .
n
= (2π )− 2 p. v. + f (y)dy
Rn |ξ |2 − k2 2k(2π )n Rn Sn−1
23 Estimates for the Laplacian and Hamiltonian 219
k instead of (−Δ − k2 −
Remark 23.2. In what follows we will use the notation G
−1
i0) .
Proof. First we prove that if the claim holds for k = 1, then it holds for every k > 0.
So let us assume that
G 1 f p n ≤ C f L p (Rn ) .
L (R )
n
Set Tδ f := f (δ x), δ > 0. It is clear that Tδ f L p (Rn ) = δ − p f L p (Rn ) . It is not dif-
ficult to show that Gk = k−2 Tk G1 T 1 . Indeed, since
k
k f = (2π )−n ei(y,ξ ) f (x − y)dξ dy
G ,
Rn Rn |ξ |2 − k2 − i0
we get
ei(y,ξ ) f ( x−y
k )dξ dy
1 T 1 f = (2π )−n
G .
k Rn Rn |ξ | − 1 − i0
2
It follows that
ei(y,ξ ) f (x − ky )dξ dy
1 T 1 f = (2π )−n
Tk G
k Rn Rn |ξ |2 − 1 − i0
ei(z,η ) k−n f (x − z)kn dzdη
= (2π )−n
|η |2
Rn Rn
k2
− 1 − i0
η )
ei(z, f (x − z)dzdη
= (2π )−n k2 .
Rn Rn |η |2 − k2 − i0
k f p = k−2 Tk G
1 T 1 f p = k−2 k − pn 1 T 1 f p
G L L
G L
k k
− n
−2− pn −2− pn 1 p
n 1 − 1 −2
≤ Ck T 1 f L p = Ck f L p = Ck p p f L p .
k k
220 Part II: Fourier Transform and Distributions
Then
c
|Pε (ξ )| ≤ .
ε
Proof. For Pε we have the following representation:
σε ω (ξ − η )
Pε = p. v. + + dη =: I1 + I2 + I3 .
1−ε ≤|η |≤1+ε |η |<1−ε |η |>1+ε |η |2 − 1
−1
The integrals I2 and I3 can be easily bounded by ε ω L1 , because |η | < 1 − ε
1 1
implies that |η |2 −1 = 1−|η |2 < ε and |η | > 1+ ε implies that |η |2 −1 = |η |21−1 < ε1 .
1 1
where
rn−1 (2 − r)n−1
F(r, ξ ) = σε ω (ξ − rθ ) − σε ω (ξ − (2 − r)θ ) dθ .
Sn−1 r+1 3−r
If we observe that F(1, ξ ) = 0, then we get by the mean value theorem (Lagrange
formulas) that
1−δ
F(r, ξ ) 1−δ F(r, ξ ) − F(1, ξ ) ∂F
dr ≤ (ε − δ ) sup (r, ξ )
1−ε r − 1 dr = 1−ε r−1 1−ε <r<1 ∂ r
∂F
≤ ε sup .
∂r
1−ε <r<1
23 Estimates for the Laplacian and Hamiltonian 221
But
∂F rn−1 rn−1
= σε ω (ξ − rθ )dθ − θ · ∇(σε ω (ξ − rθ ))dθ
∂r r+1 Sn−1 r + 1 Sn−1
(2 − r)n−1
− σε ω (ξ − (2 − r)θ )dθ
3−r Sn−1
(2 − r)n−1
− θ · ∇(σε ω (ξ − (2 − r)θ ))dθ =: θ1 + θ2 + θ3 + θ4 .
3−r Sn−1
By the proof of Lemma 23.4 below we get |θ1 | ≤ c1 ε −1 and |θ3 | ≤ c3 ε −1 , where
the constants c1 and c3 depend on ω . The second integral, θ2 , can be estimated as
(see Lemma 23.4)
n
rn−1 ∂
ε −1 ∑ θ j σε ω (ξ − rθ )dθ ≤ c2 ε −2 .
j=1 r + 1 Sn−1 ∂xj
S
Proof. We can reduce the proof to compactly supported ω , since C0∞ = S. Let
us con-
sider a C0∞ partition of unity in Rn such that ∑∞j=0 ψ j (ξ ) = 1 or even ∑∞j=0 ψ j 1
ξ =
1, where ψ0 is supported in |ξ | < 1 and ψ j = ψ (2− j ξ )
for j = 1, 2, 3, . . . with ψ
supported in the annulus 1/2 < |ξ | < 2. We may therefore write
∞
ξ −θ ξ −θ
Sn−1
σε ω (ξ − θ ) f (θ )dθ = ∑ ε −n ψ j
ε
ω
ε
f (θ )dθ .
j=0 S
n−1
ξ −θ
For j = 1, 2, 3, . . . , the function ψ j ε ω ξ −ε θ is supported in the annulus
2 j−1 ≤ | · | ≤ 2 j+1 . Since ω is rapidly decreasing, we have that in this annulus,
ω ξ − θ ≤ CM
,
ε (1 + 2 j )M
Taking M large enough, we see that the sum in j converges to Cε −1 . To end the
proof of Lemma 23.4, notice that the term for j = 0 satisfies this inequality trivially.
Exercise 23.2. Prove that (−Δ )−1 : Lδ2 (R3 ) → L−
2 (R3 ) for δ > 1.
δ
1 f in the form
Let us return to the proof of Theorem 23.1. We can rewrite G
1 f = C p. v. f(ξ )ei(x,ξ ) dξ
G + I1 f ,
Rn |ξ |2 − 1
where
I1 f = C f(θ )ei(θ ,x) dθ .
Sn−1
Let us take a partition of unity ∑∞j=0 ψ j (x) = 1 such that supp ψ0 ⊂ {|x| < 1} and
supp ψ j ⊂ {2 j−1 < |x| < 2 j+1 }, where ψ j = ψ (2− j x) with a fixed function ψ ∈ S.
We set Ψj := ψ j G+ +
1 and K j f := Ψj ∗ f , where G1 is the kernel of the integral operator
1 . Using the estimates of the Hankel function H (1)
G n−2 (|x|) for |x| < 2, we obtain
2
|Ψ0 | ≤ C|x|2−n , n ≥ 3,
and
|Ψ0 | ≤ C(|log |x|| + 1), n = 2.
Exercise 23.3 (Sobolev inequality). Let 0 < α < n, 1 < p < q < ∞, and 1q = 1p − αn .
Prove that
f (y)dy
Rn |x − y|n−α q ≤ C f L p .
L
because
j · f 2 ≤ Ψ
K j f L2 = F (Ψj ∗ f )L2 = CΨ
j L∞ f 2 ≤ C · 2 j f 2 .
L L L
On the other hand, due to the estimate of the fundamental solution at infinity we can
n−1
obtain that |Ψj (x)| ≤ C · 2− j· 2 and
n−1
K j L1 →L∞ ≤ C · 2− j· 2 .
and supp Ψj (x) ⊂ {x : 2 j−1 < |x| < 2 j+1 }. Interpolating these estimates, we obtain
the self-dual estimates
2 1− 1p − n−1 2 −1
K j L p →L p ≤ C(2 j ) 2 p
.
p − p > n+1 . If we want to get the sharper inequality p − p ≥ n+1 , we have to use
1 1 2 1 1 2
2n 2n + 2
≤ p≤ , n ≥ 3,
n+2 n+3
1 < p ≤ 6/5, n = 2,
k p p ≤ C
G L →L
.
2−n 1p − p1
|k|
k C
G p p ≤ ,
Lδ →L−δ 1−(n−1) 1p − 21
|k|
where 2n+2
n+3 < p ≤ 2, n ≥ 2, and δ > 12 − (n + 1) 1
2p − 14 .
224 Part II: Fourier Transform and Distributions
Theorem 23.5. Assume that the potential q(x) belongs to Lσp (Rn ), n ≥ 2, with n2 <
p ≤ ∞ and σ = 0 for n2 < p ≤ n+1
2 and σ > 1 − 2p for 2 < p ≤ +∞. Then for
n+1 n+1
2p 2p
exists in the uniform operator topology from L σp+1 (Rn ) to L−p−1
σ (R ) with the norm
n
2 2
estimate
q f 2p ≤ C|k|−γ f 2p
G
p−1 p+1
L−σ /2 Lσ /2
with kernel
Proof. Let us prove first that the integral operator K
1
K(x, y) := |q| 2 (x)G+
k (|x − y|)q 1 (y), 2
1
where q 1 (y) = |q(y)| 2 sgn q(y) maps from L2 (Rn ) to L2 (Rn ) with the same norm
2
1
estimate as in Theorem 23.5. Indeed, if f ∈ L2 (Rn ) and q ∈ Lσp (Rn ), then |q| 2 ∈
2p
1
L2p
σ (R ), and therefore, f |q| 2 ∈ L
n p+1
σ (Rn ). Applying Theorem 23.1, we obtain
2 2
k (|q| 2 f )
G
1
2p ≤ C|k|−γ f 2p ,
p−1 p+1
L−σ /2 Lσ /2
k (q 1 f ) ∈
where γ is as in Theorem 23.5. Then by Hölder’s inequality we have |q| 2 G
1
2
L2 (Rn ) as asserted.
q . This operator satisfies the resolvent equa-
Let us consider now the operator G
tion
q = G
G k − Gk qG
q ,
which follows easily from (H − k2 )G q = I. We denote by G l and Gr the integral op-
+ 1 +
erators having kernels Gk (|x − y|)q 1 (y) and |q(x)| 2 Gk (|x − y|), respectively. Then
2
one can show that
Gq = Gk − G l (1 + K)
−1 Gr
2p 2p
r : L σp+1 → L2 , and G
: L2 → L2 , G
for large k. Since K l : L2 → L p−1 , Theorem 23.5
2 − σ2
is proved.
The fundamental solution of the Helmholtz operator that was considered in the pre-
vious chapter can be effectively used for the following scattering problem: find
23 Estimates for the Laplacian and Hamiltonian 225
u ∈ Hloc
2 (Rn ), n ≥ 2 that satisfies
− Δ u + qu = k2 u, x ∈ Rn , k>0 (23.2)
u = u0 + usc , u0 = eik(x,θ ) , θ ∈ Sn−1
∂ usc
lim r(n−1)/2 − ikusc = 0, r = |x|.
r→∞ ∂r
The latter condition is called the Sommerfeld radiation condition at infinity. The
problem (23.2) is called the scattering problem.
Theorem 23.6. Assume that q ∈ L1 (Rn ) ∩ Lσp (Rn ), n/2 < p ≤ ∞, σ > max{0, 1 −
(n + 1)/(2p)}, is real-valued. Then there exists a unique solution u of (23.2) such
that usc ∈ L∞ (Rn ), and this solution u necessarily satisfies the Lippmann–Schwinger
equation
u = u0 − G+
k (|x − y|)q(y)u(y)dy. (23.3)
Rn
Proof. Let us show first that there is a constant C > 0 such that
lim |usc (y)|2 dσ (y) ≤ C. (23.4)
R→∞ |y|=R
Indeed, the Sommerfeld radiation condition at infinity and Green’s identity imply
that
2ik |usc (y)|2 dσ (y) = [usc (y)ikusc (y) − usc (y)(−ikusc (y))]dσ (y)
|y|=R |y|=R
∂ ∂
= [usc (y) usc (y) − usc (y) usc (y)]dσ (y)
|y|=R ∂r ∂r
+ o(1/R(n−1)/2 ) (usc (y) + usc (y))dσ (y)
|y|=R
= [usc (y)Δ usc (y) − usc (y)Δ usc (y)]dy
|y|≤R
1/2
+ o(1) |usc (y)| dσ (y)
2
|y|=R
= q(y)[usc (y)u0 (y) − usc (y)u0 (y)]dy
|y|≤R
1/2
+ o(1) |usc (y)|2 dσ (y) , R → ∞.
|y|=R
This inequality clearly implies (23.4). The next observation is that G+ k clearly satis-
fies the Sommerfeld radiation condition at infinity. Fixing now x ∈ Rn and R > 0 suf-
ficiently large that x ∈ BR = {y : |y| < R} and applying Green’s identity to usc (y) and
G+ ∂
k (|x − y|), we obtain (using the fact that on the sphere |y| = r we have ∂νy = ∂ r )
∂ + ∂
[usc (y) Gk (|x − y|) − G+ (|x − y|) usc (y)]dσ (y)
|y|=R ∂r k ∂r
= [usc (y)(Δy + k2 )G+ +
k (|x − y|) − Gk (|x − y|)(Δ + k )usc (y)]dy.
2
|y|≤R
The integral over the sphere |y| = R can be estimated from above by
1/2 2 1/2
∂
|usc (y)| dσ (y) +
∂ r − ik Gk (|x − y|) dσ (y)
2
|y|=R |y|=R
1/2 2 1/2
∂
|G+
+ k (|x − y|)| dσ (y) ∂ r − ik usc (y) dσ (y) .
2
|y|=R |y|=R
v = v0 − Kv, (23.5)
C0
K ≤ ,
L2 (Rn )→L2 (Rn ) kγ
where C0 as in Theorem 23.5 and γ = 2 − n/2 for n/2 < p ≤ (n + 1)/2 and γ =
1/γ
1 − (n − 1)/(2p) for (n + 1)/2 < p ≤ ∞, we obtain for k > C0 that there is a
unique solution v of (23.5), namely
∞
v= ∑ K j v0 .
j=0
2C0
v − v0 L2 (Rn ) ≤ v0 L2 (Rn )
kγ
For the values of k from the interval 0 < k ≤ (C0 )1/γ we proceed as follows.
Exercise 23.4. Show that the integral operator G
k ◦ q for all k > 0 is a compact
operator in L∞ (Rn ), where q satisfies the conditions of Theorem 23.6.
This exercise implies that the integral operators K and G
q ◦ q are also compact
in L2 (Rn ) and L∞ (Rn ), respectively. Next, using Agmon’s estimate and Theorems
q = (−Δ − k2 + q − i0)−1
23.1 and 23.5, we conclude that for all k > 0 the operator G
exists in the appropriate operator topology (see Theorem 23.5), and therefore for the
solution u of (23.2) (or equivalently, (23.3)) the representation
q ◦ q)u0
u = (I − G (23.7)
where u0 = −(G
k ◦ q)u0 ∈ L∞ (Rn ) and T is a compact operator in L∞ (Rn ). By Riesz
theory (see Chapter 34) we shall obtain the unique solvability of (23.8) if we are able
to show that I − T is injective. But injectivity follows immediately from (23.7). The
theorem is therefore completely proved.
Remark 23.7. For k > 0 large enough, the unique solvability in Theorem 23.6 holds
for a complex-valued potential q.
Corollary 23.8. Let v be the outgoing solution of the inhomogeneous Schrödinger
equation
(H − k2 )v = f ,
i.e.,
v = (H − k2 − i0)−1 f ,
k ( f − qG
v(x) = G q ( f ))(x).
where θ = x
|x| and the function A f , called the scattering amplitude, is defined by
A f (k, θ ) := q ( f ))dy.
e−ik(θ ,y) ( f (y) − q(y)G
Rn
Since k is fixed, |y| ≤ R, and |x| → +∞, we may assume that k|x − y| > 1 for x large
enough. Therefore, as |x| → ∞, we have
n−3
eik|x| k 2
q f )dy
v(x) = C n−1 eik(|x−y|−|x|) ( f − qG
|x| 2 |y|≤R
1 q f )dy =: I1 + I2 .
+ o n−1 ( f − qG
|y|≤R |x − y| 2
where θ = |x|
x
∈ Sn−1 . Thus, Corollary 23.8 is proved when q and f have com-
pact support. The proof in the general case is much more difficult and is therefore
omitted.
Remark 23.9. Hint for the general case: The integral over Rn might be divided into
two parts: |y| < |x|ε and |y| > |x|ε , where ε > 0 is chosen appropriately.
Lemma 23.10 (Optical lemma). For the function A f (k, θ ) the following equality
holds:
1
|A f (k, θ )|2 dθ = − 2 n−2 Im( f v)dx,
Sn−1 C k Rn
Letting m → ∞, we obtain
Im f (x)v(x)dx = −C2 kn−2 |A f (k, θ )|2 dθ .
Rn Sn−1
Exercise 23.5. Let n = 2 or n = 3. Assume that q ∈ L p (Rn )∩L1 (Rn ) with 1 < p ≤ ∞
if n = 2 and 3 < p ≤ ∞ if n = 3. Prove that the generalized eigenfunctions u(x,k),
that is, the solutions of the problem (23.2) with (k,k) = k2 , are uniformly bounded
with respect to x ∈ Rn and |k| sufficiently large.
We will obtain very important corollaries from the optical lemma. Let Aq (k) denote
the linear mapping that takes the inhomogeneity f to the corresponding scattering
amplitude
Aq (k) : f (x) → A f (k, θ ).
Lemma 23.11. Let the potential q(x) satisfy the conditions from Theorem 23.5.
2p
/2 (R ) to L (S
Then Aq is a well defined bounded operator from Lσp+1 n 2 n−1 ) with the
C
Aq 2p ≤ γ n−2 ,
p+1
Lσ /2 →L2 |k| 2+ 2
C
Aq f 2L2 (Sn−1 ) ≤ · |k|−γ f 2 2p .
|k|n−2 p+1
Lσ /2 (Rn )
Let us denote by A0 (k) the operator Aq (k) that corresponds to the potential q ≡ 0,
i.e.,
A0 f (θ ) = e−ik(θ ,y) f (y)dy.
Rn
q is a self-adjoint operator.
since G
Let us prove now that
Then for fixed k > 0 and |x| → ∞, the solution u(x, k, θ ) admits the asymptotic
representation
n−3
ik(x,θ ) eik|x| k 2
1
u(x, k, θ ) = e +Cn n−1 A(k, θ , θ ) + o n−1 ,
|x| 2 |x| 2
where θ = x
|x| and the function A(k, θ , θ ) is called the scattering amplitude and has
the form
A(k, θ , θ ) = e−ik(θ ,y) q(y)u(y, k, θ )dy.
Rn
Proof. If (H − k2 )u = 0 and u = eik(θ ,x) + usc (x, k, θ ), then usc (x, k, θ ) satisfies the
equation
23 Estimates for the Laplacian and Hamiltonian 233
We may therefore apply Corollary 23.8 with v := usc and f := −qeik(θ ,x) to obtain
n−3
eik|x| k 2
1
usc (x, k, θ ) = Cn n−1 A f (k, θ ) + o n−1 ,
|x| 2 |x| 2
where
A f (k, θ ) = q (qeik(θ ,·) )(y))dy
e−ik(θ ,y) (−qeik(θ ,y) + qG
Rn
=− q (qeik(θ ,·) )dy.
e−ik(θ ,y) q(y)(eik(θ ,y) − G
Rn
A f (k, θ ) = − e−ik(θ ,y) q(y)u(y, k, θ )dy =: −A(k, θ , θ ).
Rn
Now let Φ0 (k) and Φ (k) be the operators defined for f ∈ L2 (Sn−1 ) as
eik(x,θ ) f (θ )dθ
1
(Φ0 (k) f )(x) := |q(x)| 2 (23.9)
Sn−1
and
1
(Φ (k) f )(x) := |q(x)| 2 u(x, k, θ ) f (θ )dθ . (23.10)
Sn−1
Lemma 23.13. The operators Φ0 (k) and Φ (k) are bounded from L2 (Sn−1 ) to
L2 (Rn ) with the norm estimates
C
Φ0 (k), Φ (k) ≤ γ n−2 , k > 0,
k 2+ 2
and 1
(Φ (k) f )(x) = |q(x)| 2 (A∗q f )(x), (23.12)
234 Part II: Fourier Transform and Distributions
where A∗0 and A∗q are the adjoint operators for A0 and Aq , respectively. Indeed, if
f ∈ L2 (Sn−1 ) and g ∈ L2 (Rn ), then
f (θ )(A0 g)(θ )dθ = f (θ )dθ eik(θ ,y) g(y)dy
Sn−1 Sn−1 Rn
= g(y)dy eik(θ ,y) f (θ )dθ
Rn Sn−1
= eik(θ ,y) f (θ )dθ g(y)dy.
Rn Sn−1
and (23.11) is immediate. Similarly one proves (23.12). Since (see Lemma 23.11)
C
A0 , Aq 2p ≤ γ n−2 ,
p+1
Lσ /2 →L2 (Sn−1 ) k 2+ 2
we have that
C
A∗0 , A∗q 2p ≤ γ n−2 .
p−1
L2 (Sn−1 )→L−σ /2 (Rn ) k 2+ 2
1 1
Φ0 (k) f L2 (Rn ) = |q| 2 (A∗0 f )L2 (Rn ) ≤ qL2 p (Rn ) A∗0 f 2p
σ p−1
L−σ /2 (Rn )
C 1
≤ γ n−2 qL2 p (Rn ) f L2 (Sn−1 ) ,
k 2+ 2 σ
where we have made use of Hölder’s inequality in the first estimate. It is clear that
the same is true for Φ (k).
1
gH 2 (Rn ) + gH 1 (Rn ) + |k| gL2 (Rn ) ≤ C (Δ + k2 )gL2 (Rn ) ,
|k| −δ −δ −δ δ
where δ > 1/2 and H−s δ (Rn ), s = 0, 1, 2, denotes the weighted Sobolev space (see
below for a precise definition). As a consequence of this estimate, for all f ∈ Lδ2 (Rn ),
δ > 1/2, one has the estimates
23 Estimates for the Laplacian and Hamiltonian 235
(−Δ − k2 − i0)−1 f 2 n ≤ β f 2 n
L−δ (R ) Lδ (R )
|k| (23.13)
(−Δ − k2 − i0)−1 f 1 n ≤ β f 2 n .
H (R ) L (R ) δ
−δ
Here (−Δ − k2 − i0)−1 is the integral operator with kernel G+ k (|x − y|), see (22.1),
and the weighted Sobolev spaces Wp,1 σ (Rn ) (or Hσ1 (Rn ) if p = 2) are understood so
that f belongs to Wp,1 σ (Rn ) if and only if f and ∇ f belong to the weighted Lebesgue
space Lσp (Rn ) (see Example 18.17).
Since the integral operator (−Δ − k2 − i0)−1 is of convolution type, using duality
we can conclude that it maps Hδ−1 (Rn ) to L− 2 (Rn ) with the norm estimate
δ
(−Δ − k2 − i0)−1 f 2 ≤ β f H −1 (Rn ) , |k| ≥ 1, (23.14)
L −δ
(Rn ) δ
where Hδ−1 (Rn ) denotes the dual space of the Sobolev space H−1 δ (Rn ) and the con-
stant β is the same as in (23.13).
We will consider now the scattering problem for the magnetic Schrödinger oper-
ator in Rn , n ≥ 2, of the form
(x))2 · +V (x)·,
Hm := −(∇ + iW (23.15)
(x) and V (x) are assumed to be real and are from the spaces
where the coefficients W
∈ Wp,1 σ (Rn ),
W V ∈ Lσp (Rn ), n < p ≤ ∞, σ > n/p ,
1/p + 1/p = 1.
(23.16)
We are looking for the solutions to the equation Hm u = k2 u, k = 0, with Hm from
(23.15) in the form
u(x) = u0 (x) + usc (x), u0 (x) = eik(x,θ ) , θ ∈ Sn−1 ,
∂ usc (x) (23.17)
limr→∞ r(n−1)/2 ∂r − ikusc (x) = 0, r = |x|.
Using the same procedure as for the Schrödinger operator (see Theorem 23.6), we
conclude that the solution (23.17) necessarily satisfies the Lippmann–Schwinger
integral equation
u(x) = u0 (x) + G+
k (|x − y|)(2i∇(W (y)u(y)) − q(y)u(y))dy,
Rn
+ |W
where q = i∇W |2 +V . This equation can be rewritten as the following integral
equation:
usc (x) = u0 (x) + Lk (usc )(x), u0 (x) = Lk (u0 )(x), (23.18)
Lemma 23.14. Suppose that the conditions (23.16) are fulfilled. Then u0 belongs
σ /2 (R ), and Lk from (23.19) maps L−σ /2 (R ) into itself with σ as (23.16).
2
to L− n 2 n
and |W
It is therefore true that under these conditions the functions V, ∇W |2 belong
to Lσ /2 (R ) and W
2 n ∈ Lσ (R ). Using the first Agmon’s estimate (23.13), one can
∞ n
easily obtain
β
u0 L2 (Rn ) ≤ 2|k| W + qL2 (Rn ) .
−σ /2 |k| Lσ2 /2 (Rn ) σ /2
Hence the first inequality in (23.20) is proved. Next, applying now (23.14), we ob-
tain that
Lk f L2 (Rn ) ≤ β 2 ∇(W f ) −1 n + q f H −1 (Rn )
−σ /2 Hσ /2 (R ) σ /2
≤β 2 W f + q f H −1 (Rn )
Lσ2 /2 (Rn ) σ /2
≤ β 2 W f L2 (Rn ) + q f H (Rn ) .
−1
Lσ∞ (Rn ) −σ /2 σ /2
To estimate the second term q f H −1 (Rn ) we proceed using Hölder’s inequality and
σ /2
the Hausdorff–Young inequalities as follows:
23 Estimates for the Laplacian and Hamiltonian 237
) )
q f H −1 (Rn ) = q f = F (q f ≤ C0 F (
q f 2p/(p−2) n
σ /2 H −1 (Rn ) 2 (Rn )
L−1 L (R )
≤ C0 (2π )−n/p q f 2p/(p+2) n ≤ C0 (2π )−n/p qL p (Rn ) f 2 ,
L (R ) L (Rn )
where p > n, q(x) = (1 + |x|2 )σ /2 q(x), f(x) = (1 + |x|2 )−σ /4 f (x), and C0 is equal to
1/p √ n ∞ 1/p
dx ( π) (n−2)/2 −p/2
C0 = = r (1 + r) dr .
Rn (1 + |x|2 ) p/2 Γ (n/2) 0
Combining this constant C0 with the latter inequality, we obtain Cp from this Lemma
and (23.20).
We denote by α and γ the following constants:
α = 2 W +Cp qLσp (Rn ) , γ = 2 W + qL2 (Rn ) . (23.21)
Lσ∞ (Rn ) Lσ2 /2 (Rn ) σ /2
Theorem 23.15. Assume that the conditions (23.16) are satisfied and assume that
β α < 1 with β and α from (23.13) and (23.21), respectively. Then the integral
σ /2 (R ), and uniformly
2
equation (23.18) has a unique solution usc from the space L− n
βγ
usc L2 (Rn ) ≤ . (23.22)
−σ /2 1−βα
σ /2 (R ) and
2
Proof. Lemma 23.14 says that Lk maps in L− n
Since u0 belongs to L− σ /2 (R ) with the norm estimate β γ , the integral equation
2 n
tions
∞
usc = (I − Lk )−1 (u0 ) = ∑ Lkj+1 (u0 ).
j=0
The estimate (23.22) follows now from Lemma 23.14 and from the latter represen-
tation for usc .
Corollary 23.16. If the constant α from (23.21) is small enough, then for fixed k,
|k| ≥ 1, usc (x, k, θ ) belongs to L∞ (Rn ) in x ∈ Rn and uniformly in θ ∈ Sn−1 .
Proof. For α small enough, u0 ∈ L∞ (Rn ) and Lk maps in L∞ (Rn ) with the norm
estimate
Lk L∞ (Rn )→L∞ (Rn ) ≤ c(k)α . (23.23)
238 Part II: Fourier Transform and Distributions
Lemma 23.17. Under the assumptions of Theorem 23.15, for fixed k ≥ 1 and for
f ∈ L∞ (Rn ) the following asymptotic representation holds:
eik|x| k(n−3)/2
+ q) f (y)dy + o 1
Lk f (x) = C e−ik(θ ,y) (2kθ W ,
|x|(n−1)/2 Rn |x|(n−1)/2
(23.24)
as |x| → ∞, where θ = x/|x| and
π
1 e−i 4 (n+1)
C= .
2 (2π )(n−1)/2
Proof. In this proof we assume (for simplicity) that n ≥ 3. Since f ∈ L∞ (Rn ) and W
vanishes at infinity, integration by parts leads to
Lk f (x) = −2i ∇y G+ G+
k (|x − y|)W (y) f (y)dy − k (|x − y|)q(y) f (y)dy.
Rn Rn
In view of this, one must study the behavior as |x| → ∞ of the functions
(n−2)/2
i k (1)
G+
k (|x − y|) = H(n−2)/2 (k|x − y|)
4 2π |x − y|
and (n−2)/2
x−y i k (1)
∇y G+
k (|x − y|) = k Hn/2 (k|x − y|),
|x − y| 4 2π |x − y|
(1)
where Hν denotes the Hankel function of the first kind of order ν . The behavior
of the latter integrals can be studied by dividing them into two cases: |y| ≤ |x|a and
|y| > |x|a , where a > 0 is a parameter that we can adjust to our liking. In the first
case we have for a < 1/2 that
|x − y| = |x| − (θ , y) + O(|x|2a−1 )
and (as a consequence of it) k|x − y| → ∞ for |x| → ∞. Thus, we use the behavior of
(1)
Hν for large argument (see [23])
(1) eiz 1
H(n−2)/2 (z) = Cn √ ,+O
z z3/2
(23.25)
(1) eiz 1
Hn/2 (z) = −iCn √ + O 3/2 ,
z z
23 Estimates for the Laplacian and Hamiltonian 239
2 −i π4 (n−1)
as |z| → ∞, where Cn = πe , n ≥ 2. Hence we obtain in this case that
iCn eik|x−y| 1
G+
k (|x − y|) = k (n−3)/2
+ O
4(2π )(n−2)/2 |x − y|(n−1)/2 |x|(n+1)/2
ik|x−y|
+ Cn e (n−3)/2 1
∇y Gk (|x − y|) = θk k +O .
4(2π )(n−2)/2 |x − y|(n−1)/2 |x|(n+1)/2
x−y x
|x − y|−(n−1)/2 = |x|−(n−1)/2 + O(|x|−(n−1)/2+a−1 ), = + O(|x|a−1 ),
|x − y| |x|
and
ik|x| −ik(θ ,y) 1
eik|x−y|
=e e +O ,
|x|1−2a
and V belong to L1 (Rn ). For the case |y| > |x|/2 we have two subcases:
since W
(1)
k|x − y| < 1 and k|x − y| > 1. For the first subcase we use the behavior of Hν for
small argument, see [23],
(1)
Hν (z) = cν z−ν + o(z−ν ), z → 0+,
G+
k (|x − y|) = c(k)|x − y|
2−n
+ o(|x − y|2−n ),
∇y G+
k (|x − y|) = c(k)|x − y|
1−n
+ o(|x − y|1−n ).
240 Part II: Fourier Transform and Distributions
since σ > n/p , n < p ≤ ∞, and n ≥ 3. We have used here the estimates for the
convolution of the weak singularities (see, for example, Lemma 34.3).
For the second subcase we can use (23.25) and estimate this part of Lk f (x) from
above by
| + |V |)(1 + |y|)σ
(|W
C f L∞ (Rn ) dy
k|x−y|>1,|y|>|x|/2 |x − y|(n−1)/2 (1 + |y|)σ
1/p
dy
≤ C f L∞ (Rn ) W p n + V Lσp (Rn )
Lσ (R ) |y|>|x|/2 |x − y|(n−1)p /2 |y|σ p
1
=o
|x|(n−1)/2
as |x| → ∞, using the estimates for convolution of weak singularities and conditions
(23.16).
Since usc (x, k, θ ) for fixed k ≥ 1 is an L∞ -function in x, Lemma 23.17 yields the
asymptotic representation for u as
π
ik(x,θ ) 1 e−i 4 (n+1) eik|x| k(n−3)/2 1
u(x, k, θ ) = e + A(k, θ , θ ) + o
2 (2π )(n−1)/2 |x|(n−1)/2 |x|(n−1)/2
as |x| → ∞, where the function A(k, θ , θ ) is called the scattering amplitude for the
magnetic Schrödinger operator and it is defined as
A(k, θ , θ ) = (y) + q(y))u(y, k, θ )dy.
e−ik(θ ,y) (2kθ W (23.26)
Rn
The function AB (k, θ , θ ) is called the direct Born approximation. It can be easily
checked that
)(k(θ − θ )) + F (q)(k(θ − θ ))
AB (k, θ , θ ) = 2kθ F (W
)(k(θ − θ )) + F (|W
= k(θ + θ )F (W |2 +V )(k(θ − θ )), (23.28)
where ξ = 0, ξ̂⊥ is any unit vector that is orthogonal to ξ , and k, θ , θ are defined
by
ξ ξ̂⊥ 2 ξ ξ̂⊥ 2
θ= + 4k − ξ 2 , θ = − + 4k − ξ 2
2k 2k 2k 2k
All these results, in particular the direct Born approximation, are valid also for the
Schrödinger operator (W = 0) as well as the approximation for the backscattering
amplitude (see results below).
One may have interest in the particular case θ = −θ . This case leads to the
so-called direct backscattering Born approximation, i.e.,
|2 +V )(2kθ ).
A(k, −θ , θ ) ≈ AbB (k, −θ , θ ) := F (|W (23.30)
242 Part II: Fourier Transform and Distributions
But the approximation for the backscattering amplitude admits more terms than just
the Born backscattering approximation. Namely, the following theorem holds.
Theorem 23.19. Under the conditions of Theorem 23.15 the backscattering ampli-
tude A(k, −θ , θ ) admits the following representation:
|2 +V )(2kθ ) − F (q)(kθ + η )F (q)(kθ − η )
1
A(k, −θ , θ ) = F (|W dη
(2π )n η 2 − k2 − i0
Rn
)(kθ + η )η F (W
)(kθ − η )
4k θ F (W
+ dη + hrest (kθ ),
(2π )n Rn η − k − i0
2 2
(23.31)
+ |W
where q denotes the complex conjugate of q = i∇W |2 + V and where hrest
∞
belongs to L (R ) and
n
β 2 αγ 2
hrest L∞ (Rn ) ≤ 3 . (23.32)
1−βα
Proof. The formulas (23.27) and (23.28) for the case θ = −θ show that we need
to investigate only
R(k, −θ , θ ) = −2kθ (y)usc (y, k, θ )dy +
eik(θ ,y)W eik(θ ,y) q(y)usc (y, k, θ )dy.
Rn Rn
(23.33)
But since usc = ∑∞j=1 Lkj (u0 ), we see that (23.33) can be rewritten as
R(k, −θ , θ ) = −2kθ (y)Lk u0 (y, k, θ )dy
eik(θ ,y)W
Rn
+ eik(θ ,y) q(y)Lk u0 (y, k, θ )dy
Rn
∞
− 2kθ (y) ∑ L j u0 (y, k, θ )dy
eik(θ ,y)W k
Rn j=2
∞
+ eik(θ ,y) q(y) ∑ Lkj u0 (y, k, θ )dy =: R1 + R2 .
Rn j=2
Thus, the second terms in (23.31) are proved. It remains to estimate hrest (or R2 ).
Indeed, the definition of R2 allows us to obtain (using integration by parts) that
∞
R2 (k, −θ , θ ) = − eik(θ ,y) q(y) ∑ Lkj u0 (y, k, θ )dy
Rn j=2
∞
ik(θ ,y)
+ 2i
Rn
e W (y) · ∇ ∑ Lkj u0 (y, k, θ ) dy.
j=2
Since
Lk H 1 (Rn )→L−
2 (Rn ) ≤ β α,
−σ /2 σ /2
|R2 (k, −θ , θ )|
∞ ∞
≤ qL2 (Rn ) ∑ Lkj u0 + 2 W ∇ ∑ L j
u0
σ /2 j=2 Hσ /2 (R )
1 n j=2
k −1
2
L− σ /2
(Rn ) H− σ /2
(Rn )
β αβ γ β αβ γ β 2 αγ 2
≤ qL2 (Rn ) + 2 W 1 ≤ 3 .
σ /2 1−βα Hσ /2 (Rn ) 1 − β α 1−βα
244 Part II: Fourier Transform and Distributions
Remark 23.20. This theorem (as well as Theorem 23.15) is a generalization of the
corresponding results for the Schrödinger operator. But the difference is that com-
pared with the Schrödinger operator, the magnetic Schrödinger operator is not a
“small” perturbation of the Laplacian. This is a reason for the smallness of norms in
Theorem 23.15. For the Schrödinger operator we do not need this requirement.
Remark 23.21 (One-dimensional case). There is one interesting remark that should
be made here. The asymptotic representations (see Theorem 23.15 of formulas
(23.24)) and (23.26) coincide with well known formulas in the one-dimensional
case. Moreover, the definition (23.26) of the scattering amplitude defines the reflec-
tion and transmission coefficients for the one-dimensional “magnetic” Schrödinger
operator.
Part III
Operator Theory and Integral Equations
Chapter 24
Introduction
Despite the fact that this part is devoted to Hilbert spaces, it is assumed that the fol-
lowing concepts are known (they are necessary mainly for examples and exercises):
(1) the Lebesgue integral in a bounded domain Ω ⊂ Rn and in Rn ;
(2) functions of bounded variation BV [a, b] on an interval [a, b] (see Part I for
details);
(3) the Stieltjes integral of continuous functions on [a, b];
(4) a complete normed space Ck (Ω ), k = 0, 1, 2, . . ., on a closed bounded domain
Ω ⊂ Rn defined by
Ck (Ω ) := { f : Ω → C : f Ck (Ω ) := max ∑
Ω |α |≤k
|∂ α f (x)| < ∞},
(7) the generalized (in the L2 sense) derivatives ∂ α f (x), α = (α1 , . . . , αn ) (see Part
II for details);
whence
sup AxH = sup AH→H < ∞,
A∈F,x=1 A∈F
(1) To every pair x, y ∈ H there corresponds a vector x + y, called the sum, with the
following properties:
(a) x + y = y + x;
(b) x + (y + z) = (x + y) + z ≡ x + y + z;
(c) there exists a unique element 0 ∈ H such that x + 0 = x;
(d) for every x ∈ H there exists a unique element y1 ∈ H such that x + y1 = 0.
We set y1 := −x.
(2) For every x ∈ H and every λ , μ ∈ C there corresponds a vector λ · x such that
(a) λ (μ x) = (λ μ )x ≡ λ μ x;
(b) (λ + μ )x = λ x + μ x;
(c) λ (x + y) = λ x + λ y;
(d) 1 · x = x.
Example 25.2. In the complex Euclidean space H = Cn the standard inner product
is
n
(x, y) = ∑ x jy j,
j=1
for every x ∈ H.
for every x ∈ H.
25 Inner Product Spaces and Hilbert Spaces 251
Definition 25.7 (J. von Neumann, 1925). A Hilbert space is an inner product
space that is complete (with respect to its norm topology).
Exercise 25.4. Prove that in an inner product space the norm induced by this inner
product satisfies the parallelogram law
Exercise 25.5. Prove that if in a normed space H the parallelogram law holds, then
there is an inner product on H such that x2 = (x, x) and that this inner product is
defined by the polarization identity
1
(x, y) := x + y2 − x − y2 + i x + iy2 − i x − iy2 .
4
f = max | f (x)|
x∈[a,b]
T
where B∗ = B .
(3) The sequence space l 2 (C) defined by
∞
l 2 (C) := {x j }∞j=1 , x j ∈ C : ∑ |x j |2 < ∞ .
j=1
The estimates
|x j + y j |2 ≤ 2 |x j |2 + |y j |2 , |λ x j |2 = |λ |2 |x j |2
and
1 2
|x j y j | ≤ |x j | + |y j |2
2
imply that l 2 (C) is a linear space. Let us define the inner product by
∞
(x, y) := ∑ x jy j
j=1
2 ∞
(k) (k) (m)
x − x(m) = ∑ |x j − x j |2 < ε 2
j=1
(k) (m)
|x j − x j | < ε , j = 1, 2, . . . ,
(k)
or that {x j }∞
k=1 is a Cauchy sequence in C for every j = 1, 2, . . .. Since C is a
(k)
complete space, it follows that {x j }∞
k=1 converges for every fixed j = 1, 2, . . .,
i.e., there exists x j ∈ C such that
(k)
x j = lim x j .
k→∞
imply that
l l
(k) (m) (k)
lim
m→∞
∑ |x j − x j |2 = ∑ |x j − x j |2 ≤ ε 2
j=1 j=1
l
(k)
sl := ∑ |x j − x j |2 , k ≥ n0 ,
j=1
∞ l
(k) (k)
∑ |x j − x j |2 = lim
l→∞
∑ |x j − x j |2 ≤ ε 2 ,
j=1 j=1
and x ∈ l 2 (C).
(4) The Lebesgue space L2 (Ω ), where Ω ⊂ Rn is an open set. The space L2 (Ω )
consists of all Lebesgue measurable functions f that are square integrable, i.e.,
| f (x)|2 dx < ∞.
Ω
254 Part III: Operator Theory and Integral Equations
Definition 25.8. Let H be an inner product space. For a linear subspace M ⊂ H the
orthogonal complement of M is defined as
x = u + v,
H = M ⊕ M⊥.
d := d(x, M) ≡ inf x − y ≤ x − u
y∈M
for all u ∈ M. The definition of infimum implies that there exists a sequence
{u j }∞j=1 ⊂ M such that
25 Inner Product Spaces and Hilbert Spaces 255
d = lim x − u j .
j→∞
u = lim u j .
j→∞
Since u − α y ∈ M, we have
This inequality implies that (y, v) = 0, which means that v ∈ M ⊥ . In order to prove
uniqueness, assume that x = u1 + v1 = u2 + v2 , where u1 , u2 ∈ M and v1 , v2 ∈ M ⊥ .
It follows that
u1 − u2 = v2 − v1 ∈ M ∩ M ⊥ .
H = M ⊕ M⊥,
T (x) T (x)
(x, v0 ) = (u, v0 ) + v0 2 = v0 2 ,
T (v0 ) T (v0 )
or
T (v0 ) T (v0 )
T (x) = (x, v0 ) = x, v0 ,
v0 2 v0 2
which is of the desired form. The uniqueness of h can be seen as follows. If T (x) =
2
(x, h) = (x, h) = 0 for all x ∈ H. In particular, h −
h), then (x, h − h = (h − h,
h− h) = 0, i.e., h =
h. It remains to prove the statement about the norm T H→C =
T . Firstly,
T = sup |T (x)| = sup |(x, h)| ≤ h .
x≤1 x≤1
On the other hand, T (h/ h) = h implies that T ≥ h. Thus T = h. This
completes the proof.
Corollary 25.15. If M is a linear subspace of a Hilbert space H, then
⊥
M ⊥⊥ := M ⊥ = M.
⊥
Proof. It is not difficult to check that M ⊥ = M . Therefore,
⊥
⊥
M ⊥⊥ = M ,
25 Inner Product Spaces and Hilbert Spaces 257
span B = H,
i.e., for every x ∈ H and every ε > 0 there exist k ∈ N and {c j }kj=1 ⊂ C such
that
k
x − ∑ c j x j < ε , x j ∈ B.
j=1
i.e.,
k
x − ∑ (x, x j )x j → 0, k → ∞.
j=1
Proof. (1)⇒(2) Suppose that there is z ∈ H, z = 0 such that (z, x j ) = 0 for all j =
1, 2, . . .. Then
z
B := , x1 , x2 , . . .
z
k
x(k) = ∑ (x, x j )x j .
j=1
The Pythagorean theorem and Bessel’s inequality (Exercises 25.1 and 25.2) im-
ply that
2 k
(k)
x = ∑ |(x, x j )|2 ≤ x .
2
j=1
It follows that
∞
∑ |(x, x j )|2
j=1
2 k
(k)
x − x(m) = ∑ |(x, x j )|2 → 0
j=m+1
The continuity of the inner product and the orthonormality of {x j }∞j=1 allow us
to conclude that
∞ ∞ ∞
(x, y) = ∑ ∑ (x, x j )(y, xk )(x j , xk ) = ∑ (x, x j )(y, x j ).
j=1 k=1 j=1
A : D(A) ⊂ H → H,
A(λ x + μ y) = λ Ax + μ Ay
for all λ , μ ∈ C and x, y ∈ D(A). The space D(A) is called the domain of A. The
space
N(A) := {x ∈ D(A) : Ax = 0}
is called the range of A. Both N(A) and R(A) are linear subspaces of H. We say that
A is bounded if there exists M > 0 such that
or equivalently,
AH→H = sup Ax .
x=1
is bounded. Indeed,
2 2
K f 2 = f (s)|2 ds =
|K K(s,t) f (t)dt ds
L (Ω ) Ω ΩΩ
2 2
= (K(s, ·), f ) 2 ds ≤ K(s, ·)2L2 f L2 ds
L
Ω Ω
= |K(s,t)| dt | f (t)| dt ds = K2L2 (Ω ×Ω ) f 2L2 (Ω ) ,
2 2
Ω Ω Ω
The norm
K := KL2 (Ω ×Ω )
HS
is called the Hilbert–Schmidt norm of K.
Example 26.2 (Schur test). Assume that p and q are positive measurable functions
on Ω ⊂ Rn and α and β are positive numbers such that
|K(x, y)|p(y)dy ≤ α q(x), a.e. in Ω
Ω
and
|K(x, y)|q(x)dx ≤ β p(y), a.e. in Ω .
Ω
is bounded and
Then K
K ≤ αβ .
L2 →L2
2
|K(x, y)| · | f (y)|dy dx
Ω Ω
2
|K(x, y)|
= |K(x, y)| p(y) | f (y)|dy dx
Ω Ω p(y)
|K(x, y)|
≤ |K(x, y)|p(y)dy | f (y)|2 dy dx
Ω Ω Ω p(y)
| f (y)|2
≤α |K(x, y)|q(x)dx dy ≤ αβ | f (y)|2 dy
Ω Ω p(y) Ω
Exercise 26.1. Assume that α and β are positive constants such that
|K(x, y)|dy ≤ α , a.e. in Ω
Ω
and
|K(x, y)|dx ≤ β , a.e. in Ω
Ω
The following fundamental result can be used in the theory of bounded linear
operators (see [29]).
uk = 4−k , An uk ≥ 2 An uk , An uk ≥ 2(Mk−1 + k), (26.3)
k k k
3
where M0 = 1 and Mk = supn An (u1 + · · · + uk ). Indeed, by (26.2) there exists An1
with An1 ≥ 24. The definition of the norm of a linear operator allows us to choose
u1 such that u1 = 1 and An1 u1 ≥ 23 An1 . Setting u1 = u1 /4 shows that all
conditions (26.3) are satisfied for k = 1 and with M0 = 1.
Assuming that u1 , u2 , . . . , uk−1 , An1 , An2 , . . . , Ank−1 have been defined, choose Ank
such that
An ≥ 3 · 4k (Mk−1 + k),
k
which is possible by hypothesis With this choicek of Ank there exists uk such
2 (26.2).
that uk = 1 and Ank uk ≥ 3 Ank . Setting uk = uk /4 , we have again that uk =
4−k and
An uk ≥ 2 · 4−k An ≥ 2 · 4−k 3 · 4k (Mk−1 + k) = 2(Mk−1 + k).
k k
3 3
x ∈ [−π , π ] there exists a continuous function whose Fourier partial sums Sn f (x) are
unbounded at x.
Proof. Let us consider on the Banach space C[−π , π ] of continuous and periodic
functions on the interval [−π , π ] the linear operators
f → Sn f (x) = ∑ cn ( f )einx ,
n≤N
where cn ( f ) are the trigonometric Fourier coefficients of f . Since we have the sharp
estimate
1 π
Sn f L∞ (−π ,π ) ≤ |DN (x)|dx f L∞ (−π ,π ) ,
2π −π
where DN (x) is the Dirichlet kernel (see Chapter 10), choosing the sequence
fn (x) := σn ( f0 ) defined by Fejér means with f0 (x) = sgn DN (x), we obtain
π
2π SN fn (0) = fn (x)DN (x)dx
−π
π π
8 log N
→ f0 (x)DN (x)dx = |DN (x)|dx = + O(1), N → ∞,
−π −π π
as is stated in Exercise 10.3. Thus, the linear operators f → SN f (x) are bounded (for
each fixed N) with operator norms
4 log N
SN = + O(1).
π2
Therefore, by the uniform boundedness principle, there exists a continuous function
satisfying the present corollary.
Exercise 26.3. Suppose that f ∈ L1 (−π , π ) is periodic. Prove that if for some x
there exists
f (x + y) + f (x − y)
lim ,
y→0 2
d
A := i
dt
First of all, we have that D(A) = L2 (0, 1); see, e.g., Lemma 17.2. Moreover, inte-
gration by parts gives
1 1 1
(A f , g) = i f (t)g(t)dt = i f g|10 − i f (t)g (t)dt = f (t)ig (t)dt = ( f , Ag)
0 0 0
But
1 2
1
d dt = (nπ )2 (nπ )2
Aun 2L2 = i
dt sin(nπ t) | cos(nπ t)|2 dt = = (nπ )2 un 2L2 .
0 0 2
d2 d
A := p0 + ip1 + p2
dt 2 dt
and with real nonzero constant coefficients p0 , p1 , and p2 . The fact D(A) = L2 and
integration by parts gives
26 Symmetric Operators in Hilbert Spaces 267
1 1 1
(A f , g) = p0 f · gdt + ip1 f · gdt + p2 f · gdt
0 0
0
1 1 1
= p0 f g0 −
f · g dt + ip1 f g|0 − 1
f · g dt + p2 ( f , g)L2
0 0
1 1
= −p0 f · g dt − ip1 f · g dt + ( f , p2 g)L2
0 0
1
f g 0 −
1
= −p0 f · g dt + ( f , ip1 g )L2 + ( f , p2 g)L2
0
1
= p0 f · g dt + ( f , ip1 g )L2 + ( f , p2 g)L2 = ( f , Ag)L2
0
for all f , g ∈ D(A). Moreover, for the sequence un (t) = sin(nπ t) we have (for suffi-
ciently large n) that
1
Aun 2L2 = |p0 (sin(nπ t)) + ip1 (sin(nπ t)) + p2 sin(nπ t)|2 dt
0
1
= (p0 (nπ )2 − p2 )2 sin2 (nπ t) + (nπ )2 p21 cos2 (nπ t) dt
0
1
(nπ )4 2 2
≥ p0 sin (nπ t) + (nπ )2 p21 cos2 (nπ t) dt
0 2
1
≥ (nπ )2 p21 sin2 (nπ t) + cos2 (nπ t) dt
0
1
= 2(nπ )2 p21 = 2(nπ )2 p21 un 2L2 .
2
So A is unbounded, since
A2L2 →L2 ≥ 2(nπ )2 p21
for n → ∞.
From now on we assume that D(A) = H, i.e., that A is densely defined in any case.
Remark 26.10. The graph Γ (A) is a linear subspace of a Hilbert space H × H. The
inner product in H × H can be defined as
The reader is asked to verify that it is also possible to use a seemingly weaker, but
equivalent, criterion:
⎧
⎪
⎨xn ∈ D(A), x ∈ D(A),
w
xn → x, ⇒
⎪
⎩ w y = Ax,
Axn → y
w
where xn → x indicates weak convergence in the sense that
(xn , y) → (x, y)
for all y ∈ H.
Remark 26.12. It is important from the point of view of applications (in particular,
for numerical procedures) that the closedness of an operator guarantees the conver-
gence of some process to the “correct” result.
Definition 26.13. Let A and A1 be two linear operators in a Hilbert space H. The
operator A1 is called an extension of A (or A is a restriction of A1 ) if D(A) ⊂ D(A1 )
and Ax = A1 x for all x ∈ D(A). We denote this fact by A ⊂ A1 and A = A1 |D(A) .
whenever A ⊂ A1 = A1 and A ⊂ A 1 .
1 = A
If A is closable, then Γ A = Γ (A).
26 Symmetric Operators in Hilbert Spaces 269
The operator A∗ with domain D(A∗ ) := D∗ and mapping A∗ v = h is called the adjoint
operator of A.
Remark 26.16. The adjoint operator is maximal among all linear operators B (in the
sense that B ⊂ A∗ ) that satisfy
where
0, 0 ≤ x ≤ 1/n,
χn (x) =
1, 1/n < x ≤ 1.
We conclude that
D(A∗ ) = v ∈ L2 : x−α v ∈ L2 .
Let us show that A is not closed. To see this, we take the sequence
xα , 1/n < x ≤ 1,
fn (x) =
0, 0 ≤ x ≤ 1/n.
But xα ∈ / D(A). This contradiction shows that A is not closed. It is not bounded
either, since α > 0.
Theorem 26.18. Let A be a linear and densely defined operator. Then
(1) A∗ = A∗ .
∗ ∗∗ ∗ ∗
if∗ D(A∗ ) = H. In this case A := (A ) = A.
(2) A is closable if and only
(3) If A is closable, then A = A .
Proof. (1) Let us define in H ×H the linear and bounded operator V as the mapping
It has the property V 2 = −I. The equality (Au, v) = (u, A∗ v) for u ∈ D(A) and
v ∈ D(A∗ ) can be rewritten as
This implies that Γ (A∗ ) ⊥V Γ (A) and Γ (A∗ ) ⊥V Γ (A), which in turn means that
⊥
Γ (A∗ ) ⊂ V Γ (A) . Let us check that the criterion for closedness holds, i.e.,
⎧
⎪ ∗
⎨vn ∈ D(A ), v ∈ D(A∗ ),
vn → v, ⇒
⎪
⎩ ∗ y = A∗ v.
A vn → y
Hence (Au, v) = (u, y). Thus v ∈ D(A∗ ) and y = A∗ v. This proves (1).
(2) Assume D(A∗ ) = H. Then A∗∗ exists and due to part (1) we may conclude that
⊥
Γ (A∗ ) ⊂ V Γ (A) .
26 Symmetric Operators in Hilbert Spaces 271
Then
V Γ (A) ⊂ Γ (A∗ )⊥ .
It follows that
Γ (A) ⊂ (−V Γ (A∗ ))⊥ ,
Thus
(−V Γ (A∗ ))⊥ = {(e1 ; e2 )},
so that
(−A∗ u, e1 )H + (u, e2 )H = 0,
or
(A∗ u, e1 )H = (u, e2 )H .
Therefore, e1 ∈ D(A∗∗ ) and A∗∗ e1 = e2 . This shows that (e1 ; e2 ) ∈ Γ (A∗∗ ) and
hence
(−V Γ (A∗ ))⊥ ⊂ Γ (A∗∗ ).
Therefore,
Γ (A) ⊂ Γ (A∗∗ ),
which means that A ⊂ A∗∗ , and since A∗∗ is closed, A is closable and A ⊂ A∗∗ .
Let us show that in this case, in fact, A = A∗∗ . Indeed, if u ∈ D(A∗∗ ), then
or
(u, A∗ v) = (A∗∗ u, v), v ∈ D(A∗ ),
or
(Au, v) = (A∗∗ u, v), v ∈ D(A∗ ).
It follows that
(u0 , v) = (0, A∗ v),
or
(A∗ v, 0) = (v, u0 ).
Γ (A)⊥(−V Γ (A∗ ))
or
Γ (A)⊥(−V Γ (A∗ ))
since A exists. Since also (0; u0 )⊥(−V Γ (A∗ )) then (0; u0 ) ∈ Γ (A) i.e. 0 =
A(0) = u0 = 0. This contradiction proves (2).
(3) Since A is closable, (1) and (2) imply
∗
A∗ = A∗ = (A∗ )∗∗ = (A)∗∗∗ = (A∗∗ )∗ = A .
It means that
A∗ v = (v, u0 ) f0 .
But (v, u0 ) f0 must belong to L2 (R). Since (v, u0 ) f0 is a constant and f0 = 0, it fol-
lows that (v, u0 ) must be equal to 0. Thus
u0 ⊥D(A∗ ),
u0 ⊥D(A∗ ).
Exercise 26.5. Assume that A is closable. Prove that D(A) can be obtained as the
closure of D(A) by the norm
1/2
Au2 + u2 .
Proof. As a closed subspace of the Hilbert space H × H, the graph Γ (A) is a Hilbert
space (see Exercise 26.5). Let us define the projection mappings P1 and P2 as fol-
lows:
P1 : Γ (A) → H, P1 (u, v) = u,
P2 : Γ (A) → H, P2 (u, v) = v.
Since A is linear, both P1 and P2 are linear. Moreover, P1 is injective and surjective
and P1 and P2 are continuous, since
Hence P1 is a bijective continuous (bounded) linear map of Γ (A) onto H and has a
continuous (bounded) inverse, since it is open; see [5]. But at the same time,
Remark 26.22. A symmetric operator is always closable, and its closure is also
symmetric. Indeed, if A ⊂ A∗ , then D(A) ⊂ D(A∗ ). Hence
H = D(A) ⊂ D(A∗ ) ⊂ H
274 Part III: Operator Theory and Integral Equations
d2
A :=
dx2
It is clear that D(A) = L2 (0, 1) and A is not closed. Moreover, integration by parts
gives
(A f , g)L2 = ( f , Ag)L2
for every f ∈ D(A) and g ∈ W22 (0, 1). That is, A is symmetric such that A ⊂ A∗ and
D(A∗ ) = W22 (0, 1). As we know, A∗ = A∗ always. Now we will show that A is the
◦ ◦
same differential operator of order 2 with D(A) = W 22 (0, 1), where W 22 (0, 1) denotes
the closure of D(A) with respect to the norm of the Sobolev space W22 (0, 1). Indeed,
for every f ∈ D(A) we have
A f 2L2 + f 2L2 ≤ f W
2
2
2
and
1
2
f W 2 2
2 = A f L2 + f L2 + | f |2 dx
2 0
1
3 3
= A f 2L2 + f 2L2 − f f dx ≤ A f 2L2 + f 2L2 .
0 2 2
◦
D(A) = W 22 (0, 1).
So we have finally
◦
D(A) D(A) = W 22 (0, 1) = D(A∗∗ ) W22 (0, 1).
and ∗
A∗ A−1 v = v. (26.4)
and therefore ∗
A∗ v ∈ D A−1
and
276 Part III: Operator Theory and Integral Equations
∗
A−1 A∗ v = v. (26.5)
∗
It follows from (26.4) and (26.5) that (A∗ )−1 exists and (A∗ )−1 = A−1 .
The boundedness of A−1 follows from part (1).
Exercise 26.6. Let A and B be injective operators. Prove that if A ⊂ B, then A−1 ⊂
B−1 .
i.e., −1 D(A−1 )
= H. We conclude that H = D(A−1 ) ⊂
A−1 ∗is also symmetric. But −1 ∗
D A ⊂ H and hence D(A ) = D A−1 = H. Thus A−1 is self-adjoint
and bounded (Hellinger–Toeplitz theorem; see Exercise 26.2). Finally,
∗
A−1 = A−1 = (A∗ )−1
if and only if A = A∗ .
This completes the proof.
Theorem 26.25 (Basic criterion of self-adjointness). If A ⊂ A∗ , then the following
statements are equivalent:
(1) A = A∗ .
(2) A = A and N(A∗ ± iI) = {0}.
(3) R(A ± iI) = H.
This implies that u0 = 0, i.e., N(A∗ − iI) = {0}. The proof of N(A∗ + iI) = {0}
is left to the reader.
(2) ⇒ (3) Since A = A and N(A∗ ± iI) = {0}, it follows, for example, that
the equation A∗ u = −iu has only the trivial solution u = 0. This implies that
R(A − iI) = H. For otherwise, there exists u0 = 0 such that u0 ⊥R(A − iI). This
means that for all u ∈ D(A) we have
((A − iI)u, u0 ) = 0
for u ∈ D(A). It follows that if (A − iI)un → v0 , then Aun and un are conver-
gent, i.e., Aun → v0 , un → u0 , and un ∈ D(A). The closedness of A implies that
u0 ∈ D(A) and v0 = Au0 , i.e., (A − iI)un → Au0 − iu0 = v0 . This means that
R(A − iI) is a closed set, i.e., R(A − iI) = R(A − iI) = H. The proof of R(A + iI) =
H is left to the reader.
(3) ⇒ (1) Assume that R(A ± iI) = H. Since A ⊂ A∗ , it suffices to show that
D(A∗ ) ⊂ D(A). For every u ∈ D(A∗ ) we have (A∗ − iI)u ∈ H. Part (3) implies
that there exists v0 ∈ D(A) such that
But in our case R(A + iI) = H. Hence N(A∗ − iI) = {0} and therefore u = v0 . Thus
D(A) = D(A∗ ).
This concludes the proof.
Example 26.26. Assume that an operator A is symmetric and closed in a Hilbert
space H. Consider the operator A∗ A on the domain
This fact leads to R(A∗ A ± iI) = H, since A∗ A ± iI is invertible in this case. Thus,
Theorem 26.25 gives us that A∗ A is self-adjoint. The same is true for the operator
278 Part III: Operator Theory and Integral Equations
N
PN x := ∑ (x, e j )e j , x ∈ H,
j=1
is a projector.
Proof. (1) Since P = P∗ and P = P2 , we have P = P∗ P. Hence P = P∗ P. But
P∗ P = P2 . Indeed,
and
(I − P)∗ = I − P∗ = I − P
and
(I − P)2 = (I − P)(I − P) = I − 2P + P2 = I − P.
PM x = u.
PM x ≤ x .
This means that PM is a bounded linear operator into M. But PM e j = e j and thus
PM2 x = PM x for all x ∈ H. Next, for all x, y ∈ H we have
∞ ∞ ∞
(PM x, y) = ∑ (x, e j )e j , y = ∑ (x, e j ) (e j , y) = ∑ (x, (y, e j )e j )
j=1 j=1 j=1
∞
= x, ∑ (y, e j )e j = (x, PM y),
j=1
i.e., PM∗ = PM . The case of finite N requires no convergence questions and is left
to the reader.
This completes the proof.
Remark 27.4. In the framework of this definition, (Ax, x) and (Bx, x) must be real
for all x ∈ H.
Proposition 27.5. For two projectors P and Q the following statements are equiv-
alent:
(1) P ≤ Q.
(2) Px ≤ Qx for all x ∈ H.
(3) R(P) ⊂ R(Q).
(4) P = PQ = QP.
Proof. (1) ⇒ (2) Follows directly from (Px, x) = (P2 x, x) = (Px, Px) = Px2 .
(3) ⇒ (4) Assume R(P) ⊂ R(Q). Then QPx = Px or QP = P. Conversely, if QP = P,
then clearly R(P) ⊂ R(Q). Finally, P = QP = P∗ = (QP)∗ = P∗ Q∗ = PQ.
282 Part III: Operator Theory and Integral Equations
(2) ⇒ (4) If (4) holds, then Px = PQx and Px = PQx ≤ Qx for all x ∈ H.
Conversely, if Px ≤ Qx, then Px = QPx + Q⊥ Px implies that
2
Px2 = QPx2 + Q⊥ Px ≤ QPx2 .
Hence
⊥ 2
Q Px = 0,
Exercise 27.1. Let {Pj }∞j=1 be a sequence of projectors with Pj ≤ Pj+1 for each
j = 1, 2, . . .. Prove that lim j→∞ Pj := P exists and that P is a projector.
Ax = x , x ∈ H,
is called an isometry.
lim Es x − Et x = 0
s→t+
for all x ∈ H.
27 John von Neumann’s Spectral Theorem 283
lim Eλ x − x = 0.
λ →+∞
Remark 27.10. It follows from the previous definition and Proposition 27.5 that
Eλ Eμ = Emin{λ ,μ } .
Let now
λ0 < λ1 < · · · < λn .
284 Part III: Operator Theory and Integral Equations
Then
n n
∑ |(Eλ j x, y) − (Eλ j−1 x, y)| = ∑ E(λ j−1 , λ j ]x, y
j=1 j=1
n
= ∑ E(λ j−1 , λ j ]x, E(λ j−1 , λ j ]y
j=1
n
≤ ∑ E(λ j−1 , λ j ]x E(λ j−1 , λ j ]y
j=1
1/2 1/2
n 2
n 2
≤ ∑ E(λ j−1 , λ j ]x ∑ E(λ j−1 , λ j ]y
j=1 j=1
n n
= ∑ E(λ j−1 , λ j ]x ∑ E(λ j−1 , λ j ]y
j=1 j=1
= E(λ0 , λn ]x E(λ0 , λn ]y ≤ x y .
By Proposition 27.11 we can define a Stieltjes integral. Indeed, for every continuous
function f (λ ) we may conclude the equality of limits
n n
lim ∑
Δ →0 j=1
f (λ j∗ ) E(λ j−1 , λ j ]x, y = lim
Δ →0
∑ f (λ j∗ )E(λ j−1 , λ j ]x, y ,
j=1
which we denote by
β
f (λ )dEλ x.
α
Thus
β
β
f (λ )d(Eλ x, y) = f (λ )dEλ x, y , x, y ∈ H.
α α
27 John von Neumann’s Spectral Theorem 285
For the spectral representation of self-adjoint operators one needs integrals not only
over finite intervals but also over the whole line, which is naturally defined as the
limit
∞
β
∞
f (λ )d(Eλ x, y) = lim f (λ )d(Eλ x, y) = f (λ )dEλ x, y
−∞ α →−∞ α −∞
β →∞
if it exists. Deriving first some basic properties of the integral just defined, one can
check that
∞
β
f (λ )d(Eλ Eβ x, y) = f (λ )d(Eλ x, y)
−∞ −∞
β
:= lim f (λ )d(Eλ x, y), x, y ∈ H.
α →−∞ α
Theorem 27.12. Let {Eλ }∞ λ =−∞ be a spectral family on a Hilbert space H and let
f be a real-valued continuous function on the line. Define
∞
D := x ∈ H : | f (λ )| d(Eλ x, x) < ∞
2
−∞
∞
(or D := x ∈ H : −∞ f (λ )dEλ x exists ). Let us define on this domain an operator
A as
∞
(Ax, y) = f (λ )d(Eλ x, y), x ∈ D(A) := D, y ∈ H
−∞
∞
(or Ax = −∞ f (λ )dEλ x, x ∈ D(A)). Then A is self-adjoint and satisfies
exists for x ∈ D and y ∈ H. Thus (Ax, y) is well defined. Let v be an element of H and
let ε > 0. Then by normalization, there exist α < −R and β > R with R sufficiently
large such that
v − E(α , β ]v = v − Eβ v + Eα v ≤ (I − Eβ )v + Eα v < ε .
286 Part III: Operator Theory and Integral Equations
In order to prove that A = A∗ , it remains to show that D(A∗ ) ⊂ D(A). Let u ∈ D(A∗ ).
Then
β
(E(α , β ]z, A∗ u) = (AE(α , β ]z, u) = f (λ )d(Eλ z, u)
α
where the integral exists because (z, A∗ u) exists. Hence u ∈ D(A) and A∗ u = Au. For
the second claim we first calculate
27 John von Neumann’s Spectral Theorem 287
∞
E(α , β ]Ax = Eβ − Eα Ax = Eβ − Eα f (λ )dEλ x
−∞
∞
∞
= f (λ )dEλ Eβ x − f (λ )dEλ Eα x
−∞ −∞
β
α
= f (λ )dEλ x − f (λ )dEλ x
−∞ −∞
β
∞
= f (λ )dEλ x = f (λ )dEλ Eβ − Eα x
α −∞
= A Eβ − Eα x = AE(α , β ]x
for all x ∈ D(A). Since the left-hand side is defined on D(A) and the right-hand side
on all of H, the latter is an extension of the former.
Exercise 27.3. Let A be as in Theorem 27.12. Prove that
∞
Au2 = | f (λ )|2 d(Eλ u, u)
−∞
if u ∈ D(A).
Exercise 27.4. Let H = L2 (R) and Au(t) = tu(t), t ∈ R. Define D(A) on which
A = A∗ and evaluate the spectral family {Eλ }∞
λ =−∞ .
Proof. First we assume that this theorem holds when A is bounded, that is, that there
is a unique spectral family {Fμ }∞
μ =−∞ such that
∞
Au = μ dFμ u, u ∈ H,
−∞
since D(A) = H in this case. But Fμ ≡ 0 for μ < m and Fμ ≡ I for μ > M, where
Let us consider now an unbounded operator that is semibounded from below, i.e.,
with some constant m0 . We assume without loss of generality that (Au,u) ≥ (u, u).
This condition implies that A−1 exists, it is defined over all of H, and A−1 ≤ 1.
Indeed, A−1 exists and is bounded because Au = 0 if and only if u = 0. The norm
estimate follows from
2
(v, A−1 v) ≥ A−1 v , v ∈ D(A−1 ).
0 ≤ (A−1 v, v) ≤ v2 , v ∈ H,
where {Fμ } is the spectral family of A−1 . Let us note that F1 = I and F0 = 0, which
follows from the spectral theorem for bounded operators and from the fact that
A−1 v = 0 if and only if v = 0. Next, let us define the operator Bε , ε > 0, as
1
1
Bε u := dFμ u, u ∈ D(A).
ε μ
lim Bε A−1 v = v
ε →0+
and hence
lim A−1 Bε u = u
ε →0+
1
∞
1
A=− dE 1 = λ dEλ .
0 μ μ 1
Exercise 27.5. Prove that this {Eλ } is a spectral family which is left-continuous.
This proves the theorem for self-adjoint operators that are semibounded from below.
For bounded operators we will only sketch the proof.
Step 1. If A = A∗ and A is bounded, then we can define
pN (A) := a0 I + a1 A + · · · + aN AN , N ∈ N,
Step 2. For every continuous real-valued function f on [m, M], where m and M are
as above, we can define f (A) as an approximation by pN (A), i.e., we can prove
that for every ε > 0 there exists pN (A) such that
L( f ) := ( f (A)u, v).
Then
|L( f )| = |( f (A)u, v)| ≤ f (A) u v ,
f K = max | f (x)|.
x∈K
This set is a linear subspace of H. For every u ∈ D f and v ∈ H let us define the linear
functional
∞
∞
L(v) := f (λ )d(Eλ u, v) = f (λ )dEλ u, v .
−∞ −∞
We set
z := f (A)u, u ∈ Df ,
i.e.,
∞
( f (A)u, v) = f (λ )d(Eλ u, v).
−∞
Remark 27.14. Since in general f is not real-valued, f (A) is not a self-adjoint oper-
ator in general.
λ −i
f (λ ) = , λ ∈ R,
λ +i
∞
β
UA u2 = | f (λ )|2 d(Eλ u, u) = lim d(Eλ u, u)
−∞ α →−∞ α
β →∞
2
= lim (Eβ u, u) − (Eα u, u) = lim Eβ u − Eα u2 = u2
α →−∞ α →−∞
β →∞ β →∞
UA = (A − iI)(A + iI)−1
is equivalent to
I −UA = 2i(A + iI)−1 ,
I +UA = 2A(A + iI)−1 ,
or
A = i(I +UA )(I −UA )−1 .
1
f (λ ) = , λ ∈ R, z ∈ C, Im z = 0.
λ −z
Define
∞
1
Rz := (A − zI)−1 = dE .
−∞ λ −z λ
Then
and therefore
A∗ A f (x) = K(y, z)K(y, x)dy f (z)dz.
Ω Ω
27 John von Neumann’s Spectral Theorem 293
Von Neumann’s spectral theorem gives us for this operator and for all s ≥ 0 that
A2
L2 →L2
(A∗ A)s = λ s dEλ ,
0
Exercise 27.7. Let A = A∗ with spectral family Eλ . Let u ∈ D( f (A)). Prove that
f (A)u ∈ D(g(A)) if and only if u ∈ D((g f )(A)) and that
∞
(g f )(A)u = g(λ ) f (λ )dEλ u.
−∞
(g f )(A) = ( f g)(A)
Definition 28.1. Given a linear operator A on a Hilbert space H with domain D(A),
D(A) = H, the set
ρ (A) = z ∈ C : (A − zI)−1 exists as a bounded operator from H to D(A)
σ (A) = C \ ρ (A)
Theorem 28.2.
(1) If A = A then the resolvent set is open and the resolvent operator Rz :=
(A − zI)−1 is an analytic function from ρ (A) to B(H; H), the set of all linear
bounded operators in H. Furthermore, the resolvent identity
Rz − Rξ = (z − ξ )Rz Rξ , z, ξ ∈ ρ (A)
Proof. (1) Assume that z0 ∈ ρ (A). Then Rz0 is a bounded linear operator from H to
−1
D(A) and thus r := Rz0 > 0. Let us define for |z − z0 | < r the operator
Gz0 := (z − z0 )Rz0 .
−1 ∞ j
I − Gz0 = ∑ Gz0 ,
j=0
A − zI = (A − z0 I)(I − Gz0 ),
or
(A − zI)−1 = (I − Gz0 )−1 Rz0 .
Hence Rz exists with D(Rz ) = H and is bounded. It remains to show that R(Rz ) ⊂
D(A). For x ∈ H we know that
y := (A − zI)−1 x ∈ H.
lim (A − zI)sn x = x.
n→∞
Hence y = (A − zI)−1 x ∈ D(A), and therefore ρ (A) is open. The resolvent iden-
tity is proved by a straightforward calculation:
28 Spectra of Self-Adjoint Operators 297
exists, and hence Rz = (Rz )2 exists, which proves this part.
(2) Assume that A = A∗ . If z ∈ ρ (A), then by definition Rz maps from H to D(A).
Hence there exists Mz > 0 such that
Rz v ≤ Mz v , v ∈ H.
This is equivalent to
1
(A − zI)u ≥ u , u ∈ D(A).
Mz
or
(Au, v0 ) = (zu, v0 ),
or
(u, A∗ v0 ) = (u, zv0 ).
It is easy to check that (A − zI)u2 = (A − zI)u2 for all u ∈ D(A). Therefore,
This implies (see part (2) of Theorem 28.2) that z ∈ ρ (A), which means that
σ (A) ⊂ R. Since A = A∗ and is therefore closed, the spectrum σ (A) is closed as
the complement of an open set (see part (1) of Theorem 28.2).
It remains to prove that σ (A) = 0.
/ Assume to the contrary that σ (A) = 0. / Then
the resolvent Rz is an entire analytic function. Let us prove that Rz is uniformly
bounded with respect to z ∈ C. We introduce the functional
We therefore have
1
Rz ≤ ,
| Im z|
Exercise 28.2. [Weyl’s criterion] Let A = A∗ . Prove that λ ∈ σ (A) if and only if
there exists xn ∈ D(A), xn = 1, such that
Definition 28.4. Let us assume that A = A. The point spectrum σp (A) of A is the
set of eigenvalues of A, i.e.,
This means that (A − λ I)−1 does not exist, i.e., there exists a nontrivial u ∈ D(A)
such that Au = λ u. The complement σ (A) \ σp (A) is called the continuous spectrum
σc (A). The discrete spectrum is the set
σd (A) = λ ∈ σp (A) : dim N(A − λ I) < ∞ and λ is isolated in σ (A) .
The set σess (A) := σ (A) \ σd (A) is called the essential spectrum of A.
In the framework of this definition, the complex plane can be divided into regions
according to
C = ρ (A) ∪ σ (A),
and
σ (A) = σd (A) ∪ σess (A),
Define
∞
Ax = ∑ s j (x, e j )e j , x ∈ D.
j=1
∞
1
(A − zI)−1 x = ∑ s j − z (x, e j )e j
j=1
Exercise 28.5. Prove that the spectrum σ (U) of a unitary operator U lies on the
unit circle in C.
Theorem 28.6. Let A = A∗ and let {Eλ }λ ∈R be its spectral family. Then
(1) μ ∈ σ (A) if and only if Eμ +ε − Eμ −ε = 0 for every ε > 0.
(2) μ ∈ σp (A) if and only if Eμ − Eμ −0 = 0. Here Eμ −0 := limε →0+ Eμ −ε in the
sense of the strong operator topology.
Proof. (1) Suppose that μ ∈ σ (A) but there exists ε > 0 such that Eμ +ε −Eμ −ε = 0.
Then by the spectral theorem we obtain for every x ∈ D(A) that
∞
(A − μ I)x2 = (λ − μ )2 d(Eλ x, x) ≥ (λ − μ )2 d(Eλ x, x)
−∞ |λ −μ |≥ε
μ −ε
∞
≥ ε2 d(Eλ x, x) = ε 2 + d(Eλ x, x)
|λ −μ |≥ε −∞ μ +ε
= ε 2 (Eμ −ε x, x) + x2 − (Eμ +ε x, x) = ε 2 x2 .
This inequality means (see part (2) of Theorem 28.2) that μ ∈ / σ (A) but μ ∈
ρ (A). This contradiction proves (1) in one direction. Conversely, if
Pn := Eμ + 1 − Eμ − 1 = 0
n n
for all n ∈ N, then there is a sequence {xn }∞ n=1 such that xn ∈ R(Pn ), i.e.,
xn = Pn xn , i.e., xn ∈ D(A) and xn = 1. For this sequence it is true that
28 Spectra of Self-Adjoint Operators 301
∞
(A − μ I)xn 2 = (λ − μ )2 d(Eλ Pn xn , Pn xn )
−∞
= (λ − μ )2 d(Eλ xn , xn )
|λ −μ |≤1/n
1 ∞ 1 1
≤ d(Eλ xn , xn ) = xn 2 = 2 → 0
n2 −∞ n2 n
0 = En x0 − Eμ +ε x0 .
0 = E−n x0 − Eμ −ε x0 .
x0 = Eμ x0 , 0 = Eμ −0 x0 .
Hence
x0 = (Eμ − Eμ −0 )x0
and therefore
Eμ − Eμ −0 = 0.
P := Eμ − Eμ −0 .
Eλ y = Eλ Py = Eλ Eμ y − Eλ Eμ −0 y = Py = y.
Eλ y = Eλ Eμ y − Eλ Eμ −0 y = Eλ y − Eλ y = 0.
Hence
∞ ∞
2
(A − μ I)y = (λ − μ ) d(Eλ y, y) =
2
(λ − μ )2 dλ (y, y) = 0.
−∞ μ
Definition 28.8. Let H and H1 be two Hilbert spaces. A bounded linear operator
K : H → H1 is called compact or completely continuous if it maps bounded sets in
H into precompact sets in H1 , i.e., for every bounded sequence {xn }∞
n=1 ⊂ H the
sequence {Kxn }∞n=1 ⊂ H1 contains a convergent subsequence.
n
Pn x := ∑ (x, e j )e j , x ∈ H.
j=1
Since R(I − Pn ) ⊃ R(I − Pn+1 ) (see Proposition 27.5), it follows that {dn }∞
n=1 is a
monotonically decreasing sequence of positive numbers. Hence the limit
28 Spectra of Self-Adjoint Operators 303
lim dn := d ≥ 0
n→∞
d
K(I − Pn )yn = Kyn ≥ .
2
Then
|(yn , x)| = |((I − Pn )yn , x)| = |(yn , (I − Pn )x)| ≤ yn (I − Pn )x → 0
w
as n → ∞ for all x ∈ H. This means that yn → 0. The compactness of K implies that
Kyn → 0. Thus d = 0. Therefore,
dn = K − KPn → 0.
Lemma 28.9. Suppose A = A∗ is compact. Then at least one of the two numbers
± A is an eigenvalue of A.
Proof. Since
A = sup |(Ax, x)|,
x=1
In fact, we can assume that limn→∞ (Axn , xn ) exists and equals, say, a. Otherwise, we
would take a subsequence of {xn }. Since A = A∗ , it follows that a is real and A =
|a|. Due to the fact that every bounded set of a Hilbert space is weakly relatively
compact (the unit ball in our case), we can choose a subsequence of {xn }, say {xkn },
w
that converges weakly, i.e., xkn → x. The compactness of A implies that Axkn → y.
Next we observe that
as n → ∞. Hence
⎧
⎪
⎨Axkn − axkn → 0, xkn → x,
Axkn → y, ⇒
⎪
⎩ w Ax = ax.
xkn → x
304 Part III: Operator Theory and Integral Equations
Remark 28.10. It is not difficult to show that the statement of Lemma 28.9 remains
true if A is just bounded and self-adjoint.
(2) If there are infinitely many eigenvalues, then lim j→∞ λ j = 0 and 0 is the only
accumulation point of {λ j }.
(3) The multiplicity of λ j is finite.
(4) If e j is the normalized eigenvector for λ j , then {e j }∞j=1 is an orthonormal system
and
∞ ∞
Ax = ∑ λ j (x, e j )e j = ∑ (Ax, e j )e j , x ∈ H.
j=1 j=1
Proof. Lemma 28.9 gives the existence of an eigenvalue λ1 ∈ R with |λ1 | = A
and a normalized eigenvector e1 . Introduce H1 = e⊥
1 . Then H1 is a closed subspace
of H, and A maps H1 into itself. Indeed,
lim |λ j | = r.
j→∞
1 1
≤ < ∞.
|λ j | r
as n → ∞. Since
n n
APn x = ∑ (x, e j )Ae j = ∑ λ j (x, e j )e j
j=1 j=1
and
A(I − Pn )x = Ax − APn x → 0, n → ∞,
we have
∞
Ax = ∑ λ j (x, e j )e j ,
j=1
and part (4) follows. Finally, Exercise 28.4 gives immediately that
If N(A) = {0}, then H = R(A). This means that for every x ∈ H and ε > 0 there
exists yε ∈ R(A) such that
x − yε < ε /2.
Hence
∞
x − yε = x − ∑ λ j (xε , e j )e j < ε /2.
j=1
Making use of the Pythagorean theorem, Bessel’s inequality, and Exercise 25.8
yields
n n
x − ∑ (x, e j )e j ≤ x − ∑ λ j (xε , e j )e j
j=1
j=1
∞ ∞
= x − ∑ λ j (xε , e j )e j + ∑ λ j (xε , e j )e j
j=1 j=n+1
∞
< ε /2 + ∑ λ j (xε , e j )e j
j=n+1
1/2
∞
≤ ε /2 + ∑ |λ j |2 |(xε , e j )|2
j=n+1
1/2
∞
≤ ε /2 + |λn+1 | ∑ |(xε , e j )| 2
j=n+1
for n sufficiently large. This means that {e j }∞j=1 is a basis of H, and moreover, it is
an orthonormal basis.
Conversely, if {e j }∞j=1 is complete in H, then R(A) = H (Riesz–Schauder) and
therefore N(A) = {0}.
Remark 28.13. The condition N(A) = {0} means that A−1 exists and H must be
separable in this case.
Proposition 28.14 (Riesz). If A is a compact operator on H and μ ∈ C, then the
null space of I − μ A is a finite-dimensional subspace.
28 Spectra of Self-Adjoint Operators 307
I f = (I − μ A) f + μ A f = μ A f ,
We will prove that f ∈ R(I − A), i.e., there exists g ∈ H such that f = (I − A)g.
Since f = 0, we can assume by the decomposition H = N(I − A) ⊕ N(I − A)⊥ that
gn ∈ N(I − A)⊥ and gn = 0 for all n ∈ N.
Suppose that gn is bounded. Then there is a subsequence {gkn } such that
w
gkn → g.
Agkn → h = Ag.
Next,
gkn = (I − A)gkn + Agkn → f + h.
1
(I − A)ukn = (I − A)gkn → 0.
gkn
and u = Au, i.e., u ∈ N(I − A). But gn ∈ N(I − A)⊥ . Hence ukn ∈ N(I − A)⊥ and
further u ∈ N(I − A)⊥ , because N(I − A)⊥ is closed. Since ukn = 1, we have
u = 1. Therefore, u = 0, while
We are now ready to derive the following fundamental result of Riesz theory.
Proof. If (I − μ A)−1 exists, then (I − μ A∗ )−1 exists too and therefore N(I − μ A∗ ) =
0. Then Riesz’s lemma (Theorem 28.15) and Exercise 26.7 imply H = R(I − μ A),
i.e., I − μ A is surjective.
Conversely, if I − μ A is surjective, then N(I − μ A∗ ) = 0, i.e., I − μ A∗ is injective
and so is I − μ A.
It remains to show that (I − μ A)−1 is bounded on H if I − μ A is injective. Assume
that (I − μ A)−1 is not bounded. Then there exists a sequence fn ∈ H with fn = 1
such that
(I − μ A)−1 fn ≥ n.
Define
fn (I − μ A)−1 fn
gn := , ϕn := .
(I − μ A)−1 fn (I − μ A)−1 fn
ϕn − μ Aϕn = gn ,
and we observe that ϕkn → μϕ and ϕ ∈ N(I − μ A). Hence ϕ = 0, and this contradicts
ϕn = 1.
If μ −1 ∈
/ σ (A), then (μ )−1 ∈
/ σ (A) also. Thus
Since A = A∗ , this means that (I − μ A)−1 exists, and the unique solution is f =
(I − μ A)−1 g.
If μ −1 ∈ σ (A), then R(I − μ A) is a proper subspace of H, and the equation (I −
μ A) f = g has a solution if and only if g ∈ R(I − μ A). Since the equation is linear,
every solution is of the form
f = f0 + u, u ∈ N(I − μ A),
(A f )(t) = t f (t).
Show that the equation A f = f has no nontrivial solutions and that (I − A)−1 does
not exist. This means that the Fredholm alternative does not hold for a noncompact
but self-adjoint operator.
Exercise 28.11. Let H = L2 (Rn ) and let
A f (x) = K(x, y) f (y)dy,
Rn
where K(x, y) ∈ L2 (Rn × Rn ) is such that K(x, y) = K(y, x). Prove that A = A∗ and
that A is compact.
310 Part III: Operator Theory and Integral Equations
Theorem 28.18 (Weyl). If A = A∗ , then λ ∈ σess (A) if and only if there exists an
orthonormal system {xn }∞
n=1 such that
(A − λ I)xn → 0
as n → ∞.
Proof. We will provide only a partial proof. See [5] for a full proof. Suppose that
λ ∈ σess (A). If λ is an eigenvalue of infinite multiplicity, then there is an infinite
orthonormal system of eigenvectors {xn }∞ n=1 , because dim(Eλ −Eλ −0 )H = ∞ in this
case. Since (A − λ I)xn ≡ 0, it is clear that
(A − λ I)xn → 0.
Next, suppose that λ is an accumulation point of σ (A). This means that λ ∈ σ (A)
and
λ = lim λn ,
n→∞
Eλn +ε − Eλn −ε = 0
We can therefore find a normalized vector xn ∈ R(Eλn +rn − Eλn −rn ). Since λn = λm
for n = m, we can find {xn }∞
n=1 as an orthonormal system. By the spectral theorem
we have
∞
(A − λ I)xn 2 = (λ − μ )2 d(Eμ xn , xn )
−∞
∞
= (λ − μ )2 d(Eμ (Eλn +rn − Eλn −rn )xn , xn )
−∞
λn +rn
= (λ − μ )2 d(Eμ xn , xn )
λn −rn
∞
≤ max (λ − μ )2 d(Eμ xn , xn )
λn −rn ≤μ ≤λn +rn −∞
= max (λ − μ )2 → 0, n → ∞.
λn −rn ≤μ ≤λn +rn
T := (A − zI)−1 − (B − zI)−1
Proof. We show first that σess (A) ⊂ σess (B). Take any λ ∈ σess (A). Then there is
an orthonormal system {xn }∞
n=1 such that
(A − λ I)xn → 0, n → ∞.
Due to Bessel’s inequality, every orthonormal system in the Hilbert space converges
w
weakly to 0. Hence yn → 0. We also have
|λ − z|
yn ≥ |λ − z| xn − (A − λ I)xn = |λ − z| − (A − λ I)xn > >0
2
for all n ≥ n0 1. Next we take the identity
(B − zI)−1 − (λ − z)−1 yn = −Tyn − (λ − z)−1 (A − λ I)xn .
w
Since T is compact and yn → 0, we deduce that
(B − zI)−1 − (λ − z)−1 yn → 0.
Introduce
zn := (B − zI)−1 yn .
Then
zn − (λ − z)−1 yn → 0,
or
yn + (z − λ )zn → 0.
|λ −z| |λ −z|
This fact and yn > 2 imply that zn ≥ 3 for all n ≥ n0 1. But
is an inner product.
Then there exists a unique self-adjoint operator A defined by the quadratic form Q
as
Q(x, y) = (Ax, y), x ∈ D(A), y ∈ D(Q)
(see Exercise 29.1). Since Q is closed, D(Q) = D(Q) is a closed subspace of H with
respect to the norm ·Q . This means that D(Q) with this inner product defines a
new Hilbert space HQ . It is clear also that
xQ ≥ x
J : H → HQ , Jx = x∗ .
Hence
(y, x) = (y, Jx)Q , x ∈ H, y ∈ HQ .
29 Quadratic Forms. Friedrichs Extension. 315
Next we prove that J is self-adjoint and that it has an inverse operator J −1 . For all
x, y ∈ H we have
for every y ∈ D(Q). Since D(Q) = H, the last equality implies that x = 0, and there-
fore N(J) = {0} and J −1 exists. Moreover,
and R(J) ⊂ HQ . Now we can define a linear operator A on the domain D(A) ≡ R(J)
as
Ax := J −1 x − (λ + 1)x, λ ∈ R.
It follows that
((A1 − A2 )x, y) = 0.
√
for all x ∈ D(A) and√y ∈ D(Q). This fact means that x ∈ D( A + λ I) and y ∈
√ ∗
D( A + λ I ). But A + λ I is self-adjoint, and therefore,
∗
D( A + λ I) = D( A + λ I ) = D(Q) ≡ HQ .
is densely defined, nonnegative, and symmetric. Let us define a new inner product
Then D(Q) becomes an inner product space. This inner product space has a com-
pletion HQ with respect to the norm
xQ := Q(x, x) + x2 .
Moreover, the quadratic form Q(x, y) has an extension Q1 (x, y) to this Hilbert space
HQ defined by
Q1 (x, y) = lim Q(xn , yn )
n→∞
HQ HQ
whenever x = limn→∞ xn , y = limn→∞ yn , xn , yn ∈ D(Q) and these limits exist. The
quadratic form Q1 is densely defined, closed, nonnegative, and symmetric. There-
fore, Theorem 29.2, applied to Q1 , gives a unique nonnegative self-adjoint operator
AF such that
29 Quadratic Forms. Friedrichs Extension. 317
QB ⊃ Q = Q1 .
Let Ω be a domain in Rn , i.e., an open and connected set. We introduce the following
notation:
1 , . . . , xn ) ∈ Ω ;
(1) x = (x
(2) |x| = x12 + · · · + xn2 ;
(3) α = (α1 , . . . , αn ) is a multi-index, i.e., α j ∈ N0 ≡ N ∪ {0}:
(a) |α | = α1 + · · · + αn ,
(b) α ≥ β if α j ≥ β j for all j = 1, 2, . . . , n,
(c) α + β = (α1 + β1 , . . . , αn + βn ),
(d) α − β = (α1 − β1 , . . . , αn − βn ) if α ≥ β ,
(e) xα = x1α1 · · · xnαn with 00 = 1,
(f) α ! = α1 ! · · · αn ! with 0! = 1;
(4) ∂ j = ∂
∂xj and ∂ α = ∂1α1 · · · ∂nαn .
a(x, ξ ) = ∑ aα (x)ξ α , ξ ∈ Rn
|α |=m
is invertible for all x ∈ Ω and ξ ∈ Rn \ {0}, that is, a(x, ξ ) = 0 for all x ∈ Ω and
ξ ∈ Rn \ {0}.
Under Assumption 30.2 either a(x, ξ ) > 0 or a(x, ξ ) < 0 for all x ∈ Ω and ξ ∈
Rn \ {0}. Without loss of generality we assume that a(x, ξ ) > 0. Assumption 30.2
implies also that m is even and that for every compact set K ⊂ Ω there exists CK > 0
such that
a(x, ξ ) ≥ CK |ξ |m , x ∈ Ω , ξ ∈ Rn .
where
β!
Cβα = .
α !(β − α )!
β
∂ α ( f g) = ∑ Cα ∂ α −β f ∂ β g.
β ≤α
where aαβ = aβ α and this value is real for all α and β . We assume also the gener-
alized ellipticity condition
∑
Ω |α |=|β |=m/2
aαβ (x)∂ α f ∂ β f dx ≥ ν ∑
Ω |α |=m/2
|∂ α f |2 dx, f ∈ C0∞ (Ω ),
Remark 30.5. If the coefficients aαβ of A(x, ∂ ) are constants, then this generalized
ellipticity condition reads
∑ aαβ ξ α +β ≥ ν ∑ ξ 2α .
|α |=|β |=m/2 |α |=m/2
30 Elliptic Differential Operators 321
i.e.,
c|ξ |m ≤ ∑ ξ 2α ≤ C|ξ |m ,
|α |=m/2
n
(Au, v)L2 = − ∑ ∂ j2 u vdx
j=1 Ω
n n
=−∑ ∂ j ((∂ j u) v) dx + ∑ (∂ j u) ∂ j v dx
j=1 Ω j=1 Ω
=− (v∇u, nx )dx + (∇u, ∇v)L2 = (∇u, ∇v)L2 ,
∂Ω
Hence
(−Δ u, u)L2 = ∇u2L2 ≤ uL2 Δ uL2 , u ∈ C0∞ (Ω ).
322 Part III: Operator Theory and Integral Equations
Therefore,
2 2 2 2
uW 2 = uL2 + ∇uL2 + Δ uL2
2
≤ u2L2 + uL2 Δ uL2 + Δ u2L2
3 3 3
≤ u2L2 + Δ u2L2 ≡ u2A ,
2 2 2
u2A := u2L2 + −Δ u2L2 .
2
uW 2 ≤ uA ≤ uW 2
3 2 2
for all u ∈ C0∞ (Ω ). A completion of C0∞ (Ω ) with respect to these norms leads us to
the statement
◦
D(A) = W 22 (Ω ).
◦
Thus A = −Δ on D(A) = W 22 (Ω ). Let us determine D(A∗ ) in this case. By the
definition of D(A∗ ) we have
D((−Δ )∗ ) = v ∈ L2 (Ω ) : there exists v∗ ∈ L2 (Ω ) such that
(−Δ u, v) = (u, v∗ ) for all u ∈ C0∞ (Ω )} .
and A = A and A = (A)∗ , that is, the closure of A does not lead us to a self-adjoint
operator.
◦
Remark 30.8. If Ω = Rn , then W 22 (Rn ) ≡ W22 (Rn ) and therefore
A = A∗ = (A)∗ .
and
u2Q ≡ uW
2
1 (Ω ) .
2
If we apply now the procedure from Theorem 29.4, then we obtain the existence of
Q1 = Q with respect to the norm ·Q , which will also be nonnegative and closed
◦
with D(Q1 ) ≡ W 12 (Ω ). The next step is to obtain the Friedrichs extension AF as
AF = J −1 − I
◦
with D(AF ) ≡ R(J) ⊂ W 12 (Ω ). A more careful examination of Theorem 29.2 leads
us to the fact
◦ ◦
D(AF ) = W 12 (Ω ) ∩ D(A∗ ) = W 12 (Ω ) ∩W22 (Ω ).
D(AF ) = {u ∈ HQ : Au ∈ H} ,
which is equivalent to
D(AF ) = {u ∈ HQ : u ∈ D(A∗ )} .
Exercise 30.3. Let H = L2 (Ω ) and A(x, D) = −Δ + q(x), where q(x) = q(x) and
q(x) ∈ L∞ (Ω ). Define A, A∗ , and AF .
A(x, ∂ ) = −(∇ + IW(x))2
+ q(x),
where aαβ (x) = aβ α (x) are real. Assume that there exists C0 > 0 such that
m
|aαβ (x)| ≤ C0 , |α |, |β | < ,
2
for all ξ ∈ Rn .
30 Elliptic Differential Operators 325
(1 + |ξ |2 )−δ ≤ ε .
Hence
(1 + |ξ |2 )m/2−δ ≤ ε (1 + |ξ |2 )m/2 ,
i.e., the claim holds for every positive constant Cε (δ ). For (1 + |ξ |2 )δ < 1
ε we can
obtain
m/2−δ
1 δ
2 m/2−δ
(1 + |ξ | ) < ≡ Cε (δ ).
ε
This means that Aμ := A + μ I is positive for μ > Cε , and therefore Theorem 29.4
gives us the existence of
Aμ F
≡ (AF )μ = AF + μ I
with domain
◦ m/2
D(AF ) = D( Aμ F ) = W 2 (Ω ) ∩ D(A∗ ),
◦ m/2
where W 2 (Ω ) is the domain of the corresponding closed quadratic form (see The-
orem 29.4). If Ω is bounded with smooth boundary ∂ Ω , then it can be proved that
326 Part III: Operator Theory and Integral Equations
D(A∗ ) = W2m (Ω ).
so that
(AF )−1
μ : L (Ω ) → L (Ω ).
2 2
Secondly,
(AF )μ f −m/2 ≥ C0 f m/2 , C0 > 0,
W2 (Ω ) W2 (Ω )
so that
◦ m/2
(AF )−1
μ : L (Ω ) → W 2 (Ω ).
2
is compact.
◦ m/2
Proof. It is enough to show that for every {ϕk }∞
k=1 ⊂ W 2 (Ω ) with ϕk m/2 ≤1
W2
there exists {ϕ jk }∞
k=1 that is a Cauchy sequence in L (Ω ). Since Ω is bounded, we
2
have
|ϕk (ξ )| ≤ ϕk L2 |Ω |1/2 ,
i.e., the Fourier transform ϕk (ξ ) (see Chapter 16) is uniformly bounded. Thus there
exists ϕjk (ξ ) that converges pointwise in Rn . Next, using Parseval’s equality and the
definition of the Sobolev spaces H s (Rn ) (see Chapter 20), we have
30 Elliptic Differential Operators 327
ϕ j − ϕ j 2 2 = |ϕjk (ξ ) − ϕ
jm (ξ )| dξ
2
k m L
R
n
= |ϕjk (ξ ) − ϕ
jm (ξ )| dξ +
2
|ϕjk (ξ ) − ϕ
jm (ξ )| dξ
2
|ξ |<r |ξ |>r
≤ |ϕjk (ξ ) − ϕ
jm (ξ )| dξ
2
|ξ |<r
1
+ (1 + |ξ |2 )m/2 |ϕjk (ξ ) − ϕ
jm (ξ )| dξ
2
(1 + r2 )m/2 Rn
2
2 −m/2
= |ϕjk (ξ ) − ϕ
jm (ξ )| dξ + (1 + r )
2
ϕ jk − ϕ jm W m/2
|ξ |<r 2
=: I1 + I2 .
Moreover, λ j has finite multiplicity and the ψ j are the corresponding eigenfunctions.
We have also the following representation:
−1 ∞
Aμ F
f= ∑ μ j ( f , ψ j )ψ j , f ∈ L2 (Ω ).
j=1
f D(A) := f H + A f H .
G(t) − I
lim f
t→0 t
exists in H.
Remark 30.19. In the sense of the previous definition we write G (0) = A.
Example 30.20. Let H = L2 (Rn ). Let ω (ξ ) be an infinitely differentiable positive
function on Rn \ {0} that is positively homogeneous of order m > 0, i.e., ω (t ξ ) =
|t|m ω (ξ ). Let us define the family {G(t)}t>0 by the formula
(1)
G(t + s) f = F −1 (e−(t+s)ω (ξ ) F f )
= F −1 (e−t ω (ξ ) F F −1 (e−sω (ξ ) F f )) = G(t)(G(s) f );
(2)
G(t) f L2 = F −1 (e−t ω (ξ ) F f ) = e−t ω (ξ ) F f ≤ F f L2 = f L2 ;
L2 L2
(3)
G(t) f − f L2 = F −1 (e−t ω (ξ ) F f − F f ) = (e−t ω (ξ ) − 1)F f →0
L2 L2
The domain of A is
Then
(1)
∞ ∞
G(t + s) = eI(t+s)λ dEλ = eItλ eIsλ dEλ
0 0
∞ ∞
= eItλ dEλ eIsμ dEμ = G(t)G(s);
0 0
(2) ∞
G(t) f 2 = |eItλ |2 d(Eλ f , f ) = f 2 ;
0
(3) ∞
G(t) f − f 2 = |eItλ − 1|2 d(Eλ f , f ) → 0, t → 0,
0
330 Part III: Operator Theory and Integral Equations
and
∞ Itλ ∞
G(t) f − f e −1
= dEλ f → I λ dEλ f ≡ IAf, t → 0,
t 0 t 0
and
G(t) − I
lim
f − IAf = 0.
t→0 t H
where aαβ = aβ α are real, in C∞ (Ω ), and bounded for all α and β . We assume that
∑
Ω |α |=|β |=m/2
aαβ (x)∂ α f ∂ β f dx ≥ ν ∑
Ω |α |=m/2
|∂ α f |2 dx, ν > 0.
As was proved above, there exists at least one self-adjoint extension of A with
D(A) = C0∞ (Ω ), namely, the Friedrichs extension AF with
◦ m/2
D(AF ) = W 2 (Ω ) ∩W2m (Ω ).
with domain
∞
= f ∈ L2 (Ω ) :
D A λ d(Eλ f , f ) < ∞ .
2
0
like that for the Friedrichs extension AF .
In general, we have no formula for D A
But we can say that
◦
Wm
2 (Ω ) ⊂ D A .
Indeed, since aαβ ∈ C∞ (Ω ) and aαβ is bounded, A(x, ∂ ) can be rewritten in the
usual form
A(x, ∂ ) = ∑ aγ (x)∂ γ
|γ |≤m
A f L2 (Ω ) ≤ c ∑ ∂ γ f L2 (Ω ) ≡ c f W m (Ω ) .
2
|γ |≤m
where θ (x, y, λ ) is called the spectral function and has the properties
(1) θ (x, y, λ ) = θ (y, x, λ ),
(2)
θ (x, y, λ ) = θ (x, z, λ )θ (z, y, λ )dz
Ω
and
θ (x, x, λ ) = |θ (x, z, λ )|2 dz ≥ 0,
Ω
(3)
sup θ (x, ·, λ )L2 (Ω ) ≤ c1 λ k ,
x∈Ω1
θ (x, x, λ ) ≤ c1 λ n/m .
31 Spectral Functions 333
Corollary 31.3. Let z ∈ ρ A . Then (A
− zI)−1 is an integral operator whose ker-
nel G(x, y, z) is called the Green’s function corresponding to A and that has the
properties
(1) ∞
dλ θ (x, y, λ )
G(x, y, z) = ,
0 λ −z
∞
− zI)−1 f =
(A (λ − z)−1 dEλ f .
0
− zI)−1 f =
(A (λ − z)−1 dλ θ (x, y, λ ) f (y)dy
0 Ω
∞
= (λ − z)−1 dλ θ (x, y, λ ) f (y)dy = G(x, y, z) f (y)dy,
Ω 0 Ω
In the case of the Friedrichs extension for a bounded domain, the spectral func-
tion θ (x, y, λ ) and the Green’s function have a special form. We know from Corol-
lary 7.14 that the spectrum σ (AF ) is the sequence {λ j }∞j=1 of eigenvalues with only
one accumulation point at +∞, and the corresponding orthonormal system {ψ j }∞j=1
forms an orthonormal basis in L2 (Ω ) such that
∞
AF f = ∑ λ j ( f , ψ j )ψ j in L2 .
j=1
Hence (see Corollary 31.3) the Green’s function has the form
∞
ψ j (x)ψ j (y)
G(x, y, z) = ∑ λj −z
in L2 .
j=1
If we assume now that n < m, then we obtain that the Green’s function G(x, y, z)
is uniformly bounded in (x, y) ∈ Ω × Ω . Let us assume for simplicity that z = iz2
and AF ≥ I. Then applying Hörmander’s estimate (see Remark 31.2) for the spectral
function, we obtain
∞ ∞
|ψ j (x)||ψ j (y)| |ψ j (x)||ψ j (y)|
|G(x, y, z)| ≤ ∑ =∑ ∑
j=1 λ j2 + z22 k=0 2k ≤λ j <2k+1 λ j2 + z22
⎛ ⎞1 ⎛ ⎞1
2 2
∞
1 ⎝
≤ ∑ 2k
(2 + z 2 )1/2 ∑ |ψ j (x)|2 ⎠ ⎝ ∑ |ψ j (y)|2 ⎠
k=0 2 2k ≤λ <2k+1 2k ≤λ <2k+1
j j
∞
(2k+1 )n/m
≤ ∑ (22k + z2 )1/2 .
k=0 2
There are certain physical problems that are connected with the reconstruction of
the quantum-mechanical potential in the Schrödinger operator H = −Δ + q(x). This
operator is defined in Rn . Here and throughout we assume that q is real-valued.
First of all we have to define H as a self-adjoint operator in L2 (Rn ). Our basic
assumption is that the potential q(x) belongs to L p (Rn ) for n2 < p ≤ ∞ and has the
following special behavior at infinity:
with some μ ≥ 0 and R > 0 sufficiently large. The parameter μ will be specified
later, depending on the situation. We would like to construct the self-adjoint ex-
tension of this operator by Friedrichs’s method, because formally our operator is
defined now only for smooth functions, say for functions from C0∞ (Rn ). In order to
construct such an extension let us consider the Hilbert space H1 defined as follows:
H1 = { f ∈ L2 (Rn ) : ∇ f (x) ∈ L2 (Rn ) and |q(x)|| f (x)|2 dx < ∞}.
Rn
Proof. If p = ∞, then
|(q f , f )L2 | ≤ |q(x)|| f (x)|2 dx ≤ qL∞ (Rn ) f 2L2 (Rn )
Rn
≤ ε ∇ f 2L2 (Rn ) + qL∞ (Rn ) f 2L2 (Rn ) .
If n
2 < p < ∞, then we estimate
|(q f , f )L2 | ≤ |q(x)|| f (x)|2 dx + |q(x)|| f (x)|2 dx
|q(x)|<A |q(x)|>A
≤ |q(x)|| f (x)|2 dx + A f 2L2 (Rn ) .
|q(x)|>A
Let us consider the integral appearing in the last estimate. For n ≥ 3 it follows from
Hölder’s inequality that
2 n−2
n n 2n n
|q(x)|| f (x)| dx ≤
2
|q(x)| dx 2 | f (x)| n−2 dx
|q(x)|>A |q(x)|>A |q(x)|>A
2
n
≤ A( 2 −p) n
n 2 2
|q(x)| dx p
c1 f W 1 (Rn )
|q(x)|>A 2
2p 2p
2
≤ c1 A1− n q n
L p (Rn )
f W 1 (Rn ) .
2
2p
The claim follows now from the last inequality, since A1− n can be chosen suffi-
ciently small for n2 < p < ∞.
Remark 32.2. Lemma 32.1 holds for every potential q ∈ L p (Rn ) + L∞ (Rn ) for
p > n2 , n ≥ 2.
Using Lemma 32.1, we obtain
These two inequalities mean that the new Hilbert space H1 is equivalent to the space
W21 (Rn ) up to equivalent norms. Thus we may conclude that for every f ∈ H1 our
operator is well defined by
Remark 32.4. Let us consider this extension procedure from another point of view.
The inequality
for every f ∈ C0∞ (Rn ). This means that there exists (H + μ0 )−1 that is also defined
for g ∈ C0∞ (Rn ) and satisfies the inequality
338 Part III: Operator Theory and Integral Equations
(1) (H + μ0 )−1 gL2 (Rn ) ≤ c1 gL2 (Rn ) or even
(2) (H + μ0 )−1 gW 1 (Rn ) ≤ c1 gW −1 (Rn ) , where W2−1 (Rn ) is the dual space of
2 2
W21 (Rn ).
L2 W −1
Since (H + μ0 )−1 is a bounded operator and C0∞ (Rn ) = L2 (Rn ) and C0∞ (Rn ) = 2
−1
W2 (Rn ), we can extend (H + μ0 )−1 as a bounded operator onto L2 (Rn ) in the first
case and onto W2−1 (Rn ) in the second. The extension for the differential operator is
H + μ0 = ((H + μ0 )−1 )−1 and D(H + μ0 ) = R((H + μ0 )−1 ) in both cases. It is also
clear that H + μ0 and (H + μ0 )−1 are self-adjoint operators.
Lemma 32.5. Let us assume that q ∈ L p (Rn ) for 2 ≤ p ≤ ∞ if n = 2, 3 and q ∈
L p (Rn ) for n2 < p ≤ ∞ if n ≥ 4. Then
|q f |2 dx = |q f |2 dx + |q f |2 dx
Rn |q|<A |q|>A
4 n−4
n n 2n n
≤ A2 f 2L2 (Rn ) + |q| 2 dx | f | n−4 dx
|q|>A |q|>A
32 The Schrödinger Operator 339
4
n 4 n
≤ A2 f 2L2 (Rn ) +CA( 2 −p) n |q| p dx 2
f W 2 (Rn ) < ∞.
|q|>A 2
if n ≥ 5 and
2 2
p
p
p
|q f | dx ≤
2
|q| dx
p
| f | dx <∞
R4 R4 R4
Proof. The embedding W22 (Rn ) ⊂ D(H) was proved in Lemma 32.5. Let us now
assume that f ∈ D(H), i.e., f ∈ W21 (Rn ) and H f ∈ L2 (Rn ). Note that for
g := H f ∈ L2 we have the following representation:
It follows that σ (H0 ) ⊂ [0, +∞), but in fact, σ (H0 ) = [0, +∞) and even σ (H0 ) =
σc (H0 ) = σess (H0 ) = [0, +∞). In order to understand this fact it is enough to ob-
serve that for every λ ∈ [0, +∞) the homogeneous equation (H0 − λ )u = 0 has a
solution of the form u(x,k) = ei(k,x) , where (k,k) = λ and k ∈ Rn . These solutions
u(x,k) are called generalized eigenfunctions, but u(x,k) ∈ / L2 (Rn ). These solutions
are bounded and correspond to the continuous spectrum of H0 . Consequently, u(x,k)
are not eigenfunctions, but generalized eigenfunctions. If we consider the solutions
of the equation (H0 − λ )u = 0 for λ < 0, then these solutions will be exponentially
increasing at the infinity. This implies that λ < 0 does not belong to σ (H0 ).
For the spectral representation of H0 we have two forms:
(1) the Neumann spectral representation
∞
−Δ f = λ dEλ f , f ∈ W22 (Rn );
0
There are some important remarks to be made about the resolvent (−Δ − z)−1
for z ∈
/ [0, +∞). A consequence of the spectral theorem is that
∞
(−Δ − z)−1 = (λ − z)−1 dEλ , z ∈ C \ [0, +∞),
0
and for such z the operator (−Δ − z)−1 is a bounded operator in L2 (Rn ). Moreover,
/ [0, +∞), the operator (−Δ − z)−1 as an operator-valued function
with respect to z ∈
is a holomorphic function. This fact follows immediately from
∞
((−Δ − z)−1 )z = (λ − z)−2 dEλ = (−Δ − z)−2 .
0
The last integral converges as well as the previous one (even better). Now we are in
a position to formulate a theorem about the spectrum of H = −Δ + q.
Proof. Due to our assumptions on the potential q(x), it can be represented as the
sum q(x) = q1 (x) + q2 (x), where q1 ∈ L p (|x| < R) with the same p and q2 → 0
as |x| → ∞. We may assume (without loss of generality) that q2 is supported in
{x ∈ Rn : |x| > R} and that it is a continuous function. Let us consider first the cases
n = 2, 3. If f ∈ L2 (Rn ), then q1 f ∈ L1 (|x| < R) and (−Δ − z)−1 (q1 f ) ∈ W12 (Rn )
(by the Fourier transform). By the embedding theorem for Sobolev spaces (see, e.g.,
[1, 3]) we have that
2− 2n
(−Δ − z)−1 (q1 f ) ∈ W12 (Rn ) ⊂ W2 (Rn ), n = 2, 3,
or
(−Δ − z)−1 ◦ q1 2 ≤ c q1 L2 ,
L (|x|<R)→L2 (Rn )
for f ∈ L2 (Rn ), and therefore, (−Δ −z)−1 (q1 f ) ∈ Ws2 (Rn ). Again by the embedding
theorem for Sobolev spaces we have
2−n( 1s − 21 )
(−Δ − z)−1 (q1 f ) ∈ W2 (Rn )
The reason is that (−Δ − z)−1 ◦ q1 is actually an integral operator with kernel
Kz (x − y) that tends to 0 as |x| → ∞ uniformly with respect to |y| < R (note that
q1 is supported in |y| < R). We therefore can approximate this kernel Kz by the
functions ϕ j ∈ C0∞ (Rn ). But A j is a compact operator for each j = 1, 2, . . ., because
the embedding
W2α (|x| < R) ⊂ L2 (|x| < R)
is a compact embedding. In order to establish this fact let us consider again ϕ j (x) ∈
L∞
C0∞ (Rn ), |ϕ j (x)| ≤ c and ϕ j → q2 as j → ∞. We can state this because C0∞ = Ċ. That
is why we required such behavior of q(x) at infinity (q → 0 as |x| → +∞). If we set
A := q2 (−Δ − z)−1 and A j := ϕ j (−Δ − z)−1 , then we obtain
A − A j 2 2 ≤ sup |ϕ j − q2 | (−Δ − z)−1 2 2
L →L L →L
x
(32.1)
≤ c sup |ϕ j − q2 | → 0, j → +∞.
x
2
But we know that W2,comp ⊂ Lcomp
2 is a compact embedding. This implies (together
with (32.1)) that A is a compact operator. Since
holomorphic function in C \ [0, +∞). Also we can prove that (I + (−Δ − z)−1 q)−1
exists for all z ∈ C \ R and for real z < −c0 , where −Δ + q ≥ −c0 . Indeed, if z ∈ C
with Im z = 0, then (I + (−Δ − z)−1 q)u = 0, or (−Δ − z)u = −qu, or (Δ u, u) +
z(u, u) = (qu, u). This implies for z, Im z = 0, that (u, u) = 0 if and only if u = 0. In
the real case z < −c0 , the equality (I + (−Δ − z)−1 q)u = 0 implies
It follows that
(−c0 − z) u2L2 ≤ 0
and thus u = 0. These remarks show us that in C \ [0, +∞) our operator I + (−Δ −
z)−1 q may be noninvertible only on [−c0 , 0).
Let us consider an open and connected set Q in C \ [0, +∞) such that [−c0 , 0) ⊂
Q; see Figure 32.1.
It is easily seen that there exists z0 ∈ Q such that (I + (−Δ − z0 )−1 q)−1 exists
also. It is not difficult to show that there exists δ > 0 such that (I + (−Δ − z)−1 q)−1
exists in Uδ (z0 ). Indeed, let us choose δ > 0 such that
1
A(z) − A(z0 )L2 →L2 < (32.2)
(I + A(z0 ))−1 L2 →L2
where B := (A(z) − A(z0 ))(I + A(z0 ))−1 . But B < 1 due to (32.2), and then
(I + B)−1 = I − B + B2 + · · · + (−1)n Bn + · · ·
exists in the strong topology from L2 to L2 . We may therefore conclude that I + A(z)
may be noninvertible only for finitely many points in Q. This fact follows from
the holomorphicity of A(z) with respect to z by analogy with the theorem about
the zeros of a holomorphic function in complex analysis. Moreover, since A(z) is a
compact operator, it follows by Fredholm’s alternative that Ker(I + A(z)) has finite
Im z
Q
Re z
−c0
dimension. We conclude that (I + (−Δ − z)−1 q)−1 does not exist at only a finite
numbers of points (at most) on [−c0 , −ε ] for all ε > 0, and these points are of finite
multiplicity. This completes the proof.
But σess ((H0 + μ )−1 ) = [0, μ1 ] = σc ((H0 + μ )−1 ), from which we conclude that
σess (H + μ ) = [μ , +∞].
Outside of this set we have only points of the discrete spectrum with one possible
accumulation point at μ . This statement is a simple corollary of Lemma 32.10.
Moreover, these points of the discrete spectrum are located on [μ − c0 , μ ) and are of
finite multiplicity. Hence the discrete spectrum σd (H) of H belongs to [−c0 , 0) with
only one possible accumulation point at {0}. And (0, +∞) is the continuous part of
σ (H). There is only one problem. Weyl’s theorem states that the operators H and
H0 do not have the same spectrum but the same essential spectrum. Thus on (0, +∞)
there can be eigenvalues of infinite multiplicity (see the definition of σess ). In order
to eliminate such a possibility and to prove that 0 ∈ σc (H) and σd (H) is finite, let
us assume additionally that our potential q(x) has a special behavior at infinity:
where μ > 2. In that case we can prove that on the interval [−c0 , 0) the operator H
has at most finitely many points of the discrete spectrum. And we prove also that
0 ∈ σc (H).
Assume to the contrary that H contains infinitely many points of the discrete
spectrum or that one of them has infinite multiplicity. This means that in D(H) there
exists an infinite-dimensional space of functions {u} that satisfy the equation
It follows that
(|∇u(x)|2 + q+ (x)|u(x)|2 )dx ≤ q− (x)|u(x)|2 dx,
Rn Rn
32 The Schrödinger Operator 345
where q+ and q− are the positive and negative parts of the potential q(x),
respectively. Let us consider an infinite sequence
of functions {u(x)} that are
orthogonal with respect to the inner product Rn q− (x)u(x)v(x)dx. This sequence
is uniformly bounded in the metric Rn (|∇u|2 + |q||u|2 )dx, and hence in the metric
Rn (|∇u| + |u| )dx. But for every eigenfunction u(x) of the operator H with eigen-
2 2
For n ≥ 3 (for n = 2 the proof needs some changes) and μ > 2 we get
I0 ≤ cr2−μ |x|−2 |u(x)|2 dx ≤ cr2−μ |∇u(x)|2 dx, u ∈ W21 (Rn ).
|x|>r Rn
with r and A fixed. On passing to the limit, these inequalities for I0 , I1 , and I2 show
that
|q(x)||u(x)|2 dx → 0, m, k → ∞.
Rn
Let us return to the proof of (1). By Lemma 32.11 we obtain that our sequence
(which is orthogonal with respect to the inner product Rn q− (x)u(x)v(x)dx) is a
Cauchy sequence in the first metric. But this fact contradicts its orthogonality. Thus
(1) is proved.
(2) Let us discuss (briefly) the situation with a positive eigenvalue on the contin-
uous spectrum. If we consider the homogeneous equation
in the space Ċ(Rn ), then by Green’s formula one can show (see [20] or [21]) that the
n−1
solution f (x) of this equation behaves at infinity as o(|x|− 2 ). We thus conclude
[21] that f (x) ≡ 0 outside some ball in Rn . By the unique continuation principle for
the Schrödinger operator it follows that f ≡ 0 in the whole of Rn .
Let us consider now the spectral representation of the Schrödinger operator H =
−Δ + q(x), with q(x) as in Theorem 32.8 with the behavior O(|x|−μ ), μ > 2, at
infinity (compare with the spectral representation that follows from von Neumann’s
spectral theorem, Theorem 27.13, for the self-adjoint operator −Δ + q in L2 (Rn )).
For all f ∈ D(H), we have
M
H f (x) = (2π )−n k2 u(x,k)dk f (y)u(y,k)dy + ∑ λ j f j u j (x),
Rn Rn j=1
where u(x,k) are the solutions of the equation Hu = k2 u, u j (x) are the orthonormal
eigenfunctions corresponding to the negative eigenvalues λ j , taking into account the
multiplicity of λ j and f j = ( f , u j )L2 (Rn ) . The functions u(x,k) are called generalized
eigenfunctions. When q ≡ 0, the generalized eigenfunctions have the form u(x,k) =
ei(x,k) . This follows by means of the Fourier transform. Indeed,
(−Δ − k2 )u = 0
if and only if
(|ξ |2 − k2 )
u = 0,
32 The Schrödinger Operator 347
or
u = ∑ cα δ (α ) (ξ −k),
α
since
|ξ |2 = k2
if and only if
ξ −k = 0.
Hence
u(x,k) = ∑ cα F −1 (δ (α ) (ξ −k))(x)
α
= ∑ cα ei(x,k) F −1 (δ (α ) (ξ ))(x) = ∑ cα ei(x,k) xα .
α α
But u(x,k) must be bounded, and so u(x,k) = c0 ei(x,k) . We choose c0 = 1. If we have
the Schrödinger operator H = −Δ + q with q = 0, then it is natural to look for the
scattering solutions of Hu = k2 u of the form u(x,k) = ei(x,k) + usc (x,k). Due to this
representation, we have
(−Δ − k2 )(ei(x,k) + usc ) = −qu,
or
(−Δ − k2 )usc = −qu.
In order to find usc , let us recall that from Chapter 22 we know the fundamental
solution of the operator −Δ − k2 . Therefore,
u(x, k) = ei(x,k) − G+
k (|x − y|)q(y)u(y)dy,
Rn
where
n−2
i |k| 2
(1)
G+
k (|x|) = H n−2 (|k||x|)
4 2π |x| 2
is the fundamental solution for the operator −Δ − k2 . This equation is called the
Lippmann–Schwinger integral equation.
Chapter 33
The Magnetic Schrödinger Operator
(x))2 u +V (x)u,
Hm u := −(∇ + iW x ∈ Ω ⊂ Rn , n ≥ 2, (33.1)
The operator Hm of the form (33.1) is symmetric in the Hilbert space L2 (Ω ) on the
domain C0∞ (Ω ). We want to construct the Friedrichs self-adjoint extension of this
operator and to describe the domain of this extension.
Lemma 33.1. Assume that the conditions (33.2) are satisfied for the coefficients of
Hm . Then for all f ∈ C0∞ (Ω ) the following double inequality holds:
Therefore, for ε > 0 sufficiently small we obtain the following double inequality:
2 2
(1 − ε ) ∇ f 2L2 (Ω ) − (1/ε − 1) W f 2 − |V |1/2 f 2 ≤ (Hm f , f )L2 (Ω )
L (Ω ) L (Ω )
2 2
≤ (1 + ε ) ∇ f 2L2 (Ω ) + (1/ε + 1) W f 2 + |V |1/2 f 2 .
L (Ω ) L (Ω )
Due to the conditions (33.2) the functions W and |V |1/2 are from equivalent
spaces (with respect to norm estimates). We shall therefore estimate only the norm
1/2 2 2
|V | f 2 , and the norm f
W can be estimated in the same manner. Let
L (Ω ) L 2 (Ω )
us consider first n ≥ 3 and some p satisfying n/2 ≤ p < ∞. Then for R > 0, using
the Hölder’s inequality we obtain
1/2 2
|V | f 2 ≤ |V || f |2 dx + |V || f |2 dx
L (Ω ) |V (x)|>R |V (x)|≤R
2/n (n−2)/n
≤ |V |n/2
dx |f|2n/(n−2)
dx + R f 2L2 (Ω )
|V (x)|>R |V (x)|>R
2p/n 2 2
≤ C1 R1−2p/n V L p (Ω ∩{x:|V (x)|>R}) f W 1 (Ω ) + R f L2 (Ω ) .
2
In obtaining the latter inequality we have used the fact that n ≥ 3, n/2 ≤ p < ∞, and
the well known embedding [1, 3] W21 (Ω ) → L2n/(n−2) (Ω ) with the norm estimate
f L2n/(n−2) (Ω ) ≤ C1 f W 1 (Ω ) .
2
where δ (R) > 0 can be chosen as small as we want if R is sufficiently large. The
same is true (with some evident changes) for p = ∞ and for n = 2. Hence, we have
for arbitrarily small ε > 0 and for arbitrarily small δ (R) > 0 that
Choosing ε > 0 arbitrarily small and R > 0 such that δ (R) = ε 2 , we obtain the
required estimate (33.3).
33 The Magnetic Schrödinger Operator 351
Exercise 33.1. Prove the previous lemma in the cases p = s = ∞ for n ≥ 2 and
p, s < ∞ for n = 2.
This lemma implies that there exists μ0 > 0 such that Hm + μ0 I is positive and
2
((Hm + μ0 I) f , f )L2 (Ω ) f W 1 (Ω ) .
2
This fact implies that there is a Friedrichs self-adjoint extension of the positive op-
erator Hm + μ0 I, denoted by (Hm + μ0 I)F (see, for example [5]), with the domain
◦
D((Hm + μ0 I)F ) = { f ∈ W 12 (Ω ) : (Hm + μ0 I) f ∈ L2 (Ω )}, (33.4)
◦
where W 12 (Ω ) is the closure of C0∞ (Ω ) with respect to the norm of the Sobolev space
W21 (Ω ). Hence, the Friedrichs extension (Hm )F of Hm can be defined as (Hm )F :=
(Hm + μ0 I)F − μ0 IF with the same domain (33.4).
∈ L∞ (Ω ) and ∇· W
Exercise 33.2. Show that if W ,V ∈ L p (Ω ) with some n ≤ p ≤ ∞
for n ≥ 3 and with some 2 < p ≤ ∞ for n = 2, then
◦
D((Hm )F ) = W 12 (Ω ) ∩W22 (Ω ).
(x)∇u + [|W
Hm u = −Δ − 2iW |2 +V − i∇ · W
]u
and then use the same technique and the same embedding theorems for Sobolev
spaces as in the proof of Lemma 33.1.
we may conclude that for all μ0 > 0, Hm + μ0 I is positive, and thus the Friedrichs
self-adjoint extension exists (see, for example, [5]). But in this so-called “general”
case we cannot characterize the domain of (Hm )F constructively. We can say only
that
352 Part III: Operator Theory and Integral Equations
But even in this “general” case we may prove the diamagnetic inequality (see [35]).
For all t > 0 we may consider (using von Neumann’s spectral theorem, see The-
orem 27.13) the self-adjoint operators
∞
e−t(Hm )F f (x) := e−t λ dEλ f (x),
0
∞ (33.5)
(0)
e−t(−Δ )F f (x) := e−t λ dEλ f (x),
0
(0)
where Eλ and Eλ are the spectral families corresponding to the self-adjoint opera-
tors (Hm )F and (−Δ )F , respectively.
Theorem 33.3. Assume that W ∈ L2 (Ω ), V ∈ L1 (Ω ), V ≥ 0, and that these poten-
tials are real-valued. Then for all f ∈ L2 (Ω ), t > 0, and μ > 0 we have that
|DW
f (x)| ≥ |∇| f (x)|| (33.7)
almost everywhere.
Proof. Indeed,
This is equivalent to
f ).
2| f |∇| f | = 2 Re( f DW
|DW
f || f | ≥ | f ||∇| f ||.
almost everywhere.
33 The Magnetic Schrödinger Operator 353
DW f f ∇| f | f
=ϕ − + ∇ϕ .
|f| | f |2 |f|
so
f |D f |2 ∇| f | Re( f DW f)
Re DW
ϕ DW
f =ϕ W − ϕ 2 Re( f DW
f)+ ∇ϕ .
|f| |f| |f| |f|
f ) for f = f 1 + i f 2 , we obtain
Calculating Re( f DW
f ) = Re[( f 1 − i f 2 )(∇ f 1 + i∇ f 2 + i∇ f 2 − W f 2 )] = f 1 ∇ f 1 + f 2 ∇ f 2 = | f |∇| f |.
Re( f DW
Thus,
f |DW
f|
2
∇| f | | f |∇| f |
Re DW
ϕ DW
f = ϕ − ϕ 2 | f |∇| f | + ∇ϕ
|f| |f| |f| |f|
|DW
f | − |∇| f ||
2 2
=ϕ + ∇ϕ ∇| f | ≥ ∇ϕ ∇| f |
|f|
To end the proof of Theorem 33.3 we consider μ > 0. Using these two lemmas
we obtain
f
∇ϕ ∇| f |dx + μ ϕ | f |dx ≤ Re DW ϕ D f dx + μ ϕ | f |dx
Ω Ω Ω |f| W
Ω
f
≤
DW ϕ DW f dx + μ ϕ | f |dx
,
Ω |f| Ω
f f
ϕ (−Δ + μ )| f |dx ≤
−DW
2
ϕ + μ f ϕ dx
,
f
Ω Ω |f| |f|
i.e.,
f
((−Δ + μ )| f |, f )L2 (Ω ) ≤
ϕ (−DW
2
+ μ ) f dx
Ω |f|
≤ ϕ |(−DW
+ μ ) f |dx = (|(−DW
2
+ μ ) f |, ϕ )L2 (Ω ) .
2
Ω
≤ (|u|, ϕ )L2 (Ω ) = (|u|, (−Δ + μ I)−1 ψ )L2 (Ω ) = ((−Δ + μ I)−1 |u|, ψ )L2 (Ω ) ,
where we have used the self-adjointness of all operators and the notation ψ = (−Δ +
μ I)ϕ . Hence, for arbitrary ψ ≥ 0 sufficiently smooth we obtain the inequality
−1 −1
+ μ I) u|, ψ )L2 (Ω ) ≤ ((−Δ + μ ) |u|, ψ )L2 (Ω ) .
(|(−DW
2
Since ψ is an arbitrary function of such type, we may conclude from here that for
every u ∈ L2 (Ω ) we have that
−1 −1
+ μ I) u(x)| ≤ (−Δ + μ I) |u|(x)
|(−DW
2
almost everywhere. Iterating the latter inequality, we obtain for all m ∈ N that
−m
+ μ I)
|(−DW
2
u(x)| ≤ (−Δ + μ I)−m |u|(x).
−1 −1
+ μ +V ) u|, ψ )L2 (Ω ) ≤ (|u|, (−Δ + μ +V ) ψ )L2 (Ω )
(|(−DW
2
since
(−Δ + μ +V )−1 ψ ≤ (−Δ + μ )−1 ψ
Using this definition, we may conclude that if a heat kernel exists, then for every
μ > 0 the inverse operator (A + μ I)−1 (which exists) is an integral operator with
kernel ∞
G(x, y, μ ) = e−μ t P(x, y,t) f (y)dt. (33.10)
0
Indeed, by von Neumann’s theorem for A (see Chapter 27), we have that for every
f ∈ L2 (Ω ), ∞
e−tA f (x) = e−t λ dEλ f ,
0
There is (at least) one quite general situation in which the heat kernel exists and
has “good” estimates. Let us assume that Ω ⊂ Rn is a bounded domain with smooth
boundary (the smoothness is required for the Sobolev embedding theorem). We con-
sider the magnetic Schrödinger operator Hm in Ω with electric potential V ≥ 0 and
with magnetic potential W satisfying all assumptions of Lemma 33.1. In this case
Hm and H0 = −Δ have Friedrichs self-adjoint extensions, which are denoted by the
same symbols Hm and H0 , respectively. We have the following theorem.
Theorem 33.6. Under the conditions of Lemma 33.1 for μ > 0, the resolvent (Hm +
μ I)−1 is an integral operator with kernel G(x, y, μ ), called the Green’s function
356 Part III: Operator Theory and Integral Equations
Proof. Using Lemma 33.1, we conclude that for μ > 0 sufficiently large and for all
◦
f ∈ W 12 (Ω ) we have
2
((Hm + μ I) f , f )L2 (Ω ) ≥ γ f W 1 (Ω ) , γ > 0.
2
◦
Since the embedding W 12 (Ω ) → L2 (Ω ) is compact (see Lemma 30.15), we have
that (Hm + μ I)−1 is compact. Using now the Riesz–Schauder and Hilbert–Schmidt
theorems (see Theorem 28.10), we conclude that the spectrum σ (Hm ) = {λ j }∞j=1
is discrete and of finite multiplicity with only one accumulation point at infinity.
The corresponding normalized eigenfunctions {ϕ j }∞j=1 form an orthonormal basis
in L2 (Ω ) such that the spectral family for Hm is defined as
Hence
∞ ∞
e−tHm f (x) =
0
e−t λ dEλ f (x) = ∑ e−t λ j f j ϕ j (x)
j=1
∞ ∞
= ∑ e−t λ j Ω
f (y)ϕ j (y)dy ϕ j (x) = ∑ e−t λ j ϕ j (x)ϕ j (y) f (y)dy,
Ω
j=1 j=1
It must be mentioned here that all these equalities (and operations) are considered
in the sense of L2 (Ω ). The equality (33.12) implies that (Hm + μ I)−1 is an integral
operator with kernel G(x, y, μ ) defined by
∞ ∞ ∞ ∞
ϕ j (x)ϕ j (y)
G(x, y, μ ) = e−μ t P(x, y,t)dt = ∑ ϕ j (x)ϕ j (y) e−t(λ j +μ ) dt =
. ∑
0 j=1 0 j=1 λ j + μ
(33.13)
This function G(x, y, μ ) is called the Green’s function of the Friedrichs extension
of the magnetic Schrödinger operator Hm in the bounded domain. To obtain the
estimates (33.11) we proceed as follows. It is known that the heat kernel P0 (x, y,t)
33 The Magnetic Schrödinger Operator 357
At the same time, the heat kernel P0 (x, y,t) of H0 in the bounded domain Ω satisfies
the following boundary value problem:
⎧
⎪
⎨∂t P0 (x, y,t)
⎪
= Δx P0 (x, y,t), x, y ∈ Ω ,t > 0,
P0 (x, y,t)
= 0, x ∈ ∂ Ω , y ∈ Ω , t > 0,
⎪
⎪ ∂Ω
⎩P (x, y, 0) = δ (x − y).
0
If we define P0 (x, y,t) = P0 (x, y,t) + R(x, y,t), then R(x, y,t) has to satisfy
⎧
⎪
⎨∂t R(x, y,t) = Δx R(x, y,t), x, y ∈ Ω , t > 0,
R(x, y,t)|∂ Ω = −P0 (x, y,t), x ∈ ∂ Ω , y ∈ Ω , t > 0
⎪
⎩
R(x, y, 0) = 0, x, y ∈ Ω .
But −P0 (x, y,t) < 0 for x, y ∈ Ω ,t > 0 (see (33.14)). Using then the maximum prin-
ciple for the heat equation (see Theorem 45.7) we obtain that R(x, y,t) ≤ 0 for all
x, y ∈ Ω and t > 0, i.e.,
0 ≤ P0 (x, y,t) ≤ P0 (x, y,t).
The next step is as follows: the diamagnetic inequality (33.6) leads in this case to
P(x, y,t) f (y)dy
≤ P0 (x, y,t)| f (y)|dy,
Ω
Ω
which holds almost everywhere in x ∈ Ω and for all f ∈ L2 (Ω ). Using the Hardy–
Littlewood maximal function (see, e.g., [18]), we can obtain from the latter inequal-
ity that
|P(x, y,t)| ≤ (4π t)−n/2 e−|x−y| /(4t) , x, y ∈ Ω , t > 0.
2
where Kν (z) is the Macdonald function of order ν . Using the asymptotic expansion
for Kν (z) for z → 0 and z → ∞ (see, for example, [23]), we can obtain the following
inequalities (see also the straightforward calculations in Example 22.8):
358 Part III: Operator Theory and Integral Equations
√
|x − y|2−n e− μ |x−y| , n ≥ 3,
|G(x, y, μ )| ≤ C √
1 + | log( μ |x − y|)|, n = 2,
where x, y ∈ Ω and the constant C > 0 depends only on the dimension n. Thus,
Theorem 33.6 is completely proved.
One more application of the diamagnetic inequality concerns the estimates of the
normalized eigenfunctions of Hm .
Ω
1/2
(4π t)−n e−|x−y| /(2t) dy
2
≤ ϕ L 2 (Ω )
Ω
1/2
(4π t)−n e−|x−y|
2 /(2t)
≤ dy
Rn
1/2
−n/2 n/4 −|y|2 /2
= (4π t) t e dy
Rn
Remark 34.2. If K(x, y) is continuous for all x, y ∈ Ω and bounded, then this integral
operator is considered also an operator with weak singularity.
and analogously
(A2 ◦ A1 ) f (x) = K2 (x, y)K1 (y, z)dy f (z)dz,
Ω Ω
So, we may conclude that the compositions A1 ◦ A2 and A2 ◦ A1 are again integral
operators with kernels
K(x, y) = K1 (x, z)K2 (z, y)dz,
Ω (34.1)
y) =
K(x, K2 (x, z)K1 (z, y)dz,
Ω
|K1 (x, y)| ≤ M1 |x − y|α1 −n and |K2 (x, y)| ≤ M2 |x − y|α2 −n , (34.2)
y).
The same estimates hold for the kernel K(x,
and
α1 +α2 −n
|K(x, y)| ≤ M1 M2 |x − y| |u − e0 |α1 −n |u|α2 −n du. (34.4)
Rn
and therefore
α1 −n α2 −n n−α1
|u − e0 | |u| du ≤ 2 |u|α2 −n du
|u|≤1/2 |u|≤1/2
1/2
2n−α1 −α2 n−1
= 2n−α1 rα2 −1 dr dθ = |S |,
0 Sn−1 α2
2 |u|
|u − e0 | ≥ |u| − |e0 | ≥ |u| − 1 ≥ |u| − |u| = ,
3 3
and we have analogously
|u − e0 |α1 −n |u|α2 −n du
|u|≥3/2
∞
2n−α1 −α2 3α2 n−1
≤ 3n−α1 |Sn−1 | rα1 +α2 −n−1 dr = |S |.
3/2 n − α1 − α2
Combining (34.4)–(34.5), we obtain (34.3) for the case α1 + α2 < n. It can be men-
tioned here that the estimate (34.3) in this case holds also in the case of an arbitrary
(not necessarily bounded) domain Ω .
If now α1 + α2 = n, then the proof of (34.3) will be a little bit different, and it
holds only for a bounded domain Ω . Indeed, for every z ∈ Ω and
|x − y| |x − y|
|x − z| ≤ or |z − y| ≤ ,
2 2
we have in both cases that
|K(x, y)| ≤ M1 M2 2n−α2 |x − y|α2 −n |x − z|α1 −n dz
Ω
|x−y|/2
≤ M1 M2 2n−α2 |Sn−1 ||x − y|α2 −n rα1 −1 dr
0
M1 M2 n−1 M1 M2 n−1
= |S | or |S |. (34.6)
α1 α2
362 Part III: Operator Theory and Integral Equations
If z ∈ Ω does not belong to these balls with radius |x − y|/2, then we consider two
cases: |z − x| ≥ |z − y| and |z − x| ≤ |z − y|. In both cases we have
d
dz dr
|K(x, y)| ≤ M1 M2 ≤ M1 M2 |Sn−1 |
Ω \Ω |x − z|n |x−y|/2 r
(34.7)
2d
= M1 M2 |Sn−1 | log ,
|x − y|
where d = diam Ω . The estimates (34.6) and (34.7) give us (34.3) in the case α1 +
α2 = n.
If finally α1 + α2 > n, then since Ω is bounded, we can analogously obtain (34.3)
in this case. This finishes the proof.
| logt| ≤ Cε t −ε , ε > 0,
where
β = sup |x − y|α −n dy < ∞.
x∈Ω Ω
AL2 (Ω )→L2 (Ω ) ≤ M β .
1, 0 ≤ t ≤ σ ,
χσ (t) =
0, t > σ .
34 Integral Operators with Weak Singularities 363
The integral operator with kernel K2 (x, y) is a Hilbert–Schmidt operator for all
σ > 0, since
|K2 (x, y)|2 dxdy ≤ M 2 |x − y|2α −2n dxdy
Ω Ω σ ≤|x−y|≤d
which is also an operator with weak singularity. Using Lemma 34.3, we can estimate
the right-hand side of (34.8) from above as
1
|Kσ (x, y)|| f (x)|| f (y)|dxdy ≤ |Kσ (x, y)|| f (x)|2 dxdy
Ω Ω 2 Ω Ω
1
+ |Kσ (x, y)|| f (y)|2 dxdy,
2 Ω Ω
(34.9)
where Kσ (x, y) is the kernel of the operator with weak singularity, i.e.,
⎧
⎪ 2α −n , α < n/2,
⎨|x − y|
|Kσ (x, y)| ≤ M |x − y|−ε , α = n/2,
⎪
⎩
1, α > n/2,
dσ (y) ≤ c0 ρ n−2 dρ dθ ,
where (ρ , θ ) are the polar coordinates in the tangent plane with origin x, and c0 is
independent of x. According to the dimension n − 1 of the surface ∂ Ω , an integral
operator in L2 (∂ Ω ), i.e.,
A f (x) = K(x, y) f (y)dσ (y),
∂Ω
is said to be with weak singularity if its kernel K(x, y) is continuous for all x, y ∈ ∂ Ω ,
x = y, and there are constants M > 0 and α ∈ (0, n − 1] such that
We will need also the famous Ascoli–Arzelà theorem (for a proof, see [22]).
Theorem 34.7. A set U ⊂ C(Ω ) is relatively compact if and only if
(1) there is a constant M > 0 such that for all ϕ ∈ U we have ϕ L∞ (Ω ) ≤ M (uni-
form boundedness);
(2) for every ε > 0 there is δ > 0 such that
Proof. We give the proof for domains in Rn . The proof for surfaces in Rn is the
same. Let x, y ∈ Ω and |x − y| < δ . Then
|A f (x) − A f (y)| ≤ (|K(x, z)| + |K(y, z)|)| f (z)|dz
|x−z|<2δ
+ |K(x, z) − K(y, z)|| f (z)|dz
Ω \{|x−z|<2δ }
≤ M f L∞ (Ω ) (|x − z|α −n + |y − z|α −n )dz
|x−z|<2δ
+ f L∞ (Ω ) |K(x, z) − K(y, z)|dz =: I1 + I2 .
Ω \{|x−z|<2δ }
|y − z| ≥ |x − z| − |x − y| > 2δ − δ = δ .
K(x, z) − K(y, z) → 0, δ → 0,
(I − μ A) f = g, (I − μ A∗ ) f = g ,
have the unique solutions f and f for any given g and g from H or the correspond-
ing homogeneous equations
(I − μ A) f = 0, (I − μ A∗ ) f = 0 (34.10)
34 Integral Operators with Weak Singularities 367
respectively.
Proof. Riesz’s lemma (see Theorem 28.14) and Exercise 26.7 give
R(I − μ A) = N(I − μ A∗ )⊥ ,
R(I − μ A∗ ) = N(I − μ A)⊥ .
These two dimensions are finite due to Riesz (see Proposition 28.13). Since every
compact operator is a norm limit of a sequence of operators of finite rank (see
Chapter 28 for details), for every μ ∈ C, μ = 0, we have
I − μ A = −μ A0 + (I − μ A1 ),
(I − μ A1 )−1 (I − μ A) = I − μ (I − μ A1 )−1 A0 =: I − A2 ,
where A2 is of finite rank too. Analogously, since (I − μ A∗1 )−1 exists, we must have
where A∗2 is adjoint to A2 and is of finite rank too. These representations allow us to
conclude that
g ∈ N(I − μ A) ⇔ g ∈ N(I − A2 ),
g ∈ N(I − μ A∗2 ) ⇔ (I − μ A∗1 )−1 g ∈ N(I − μ A∗ ).
Thus, it suffices to show that the numbers of independent solutions of the equations
g = A2 g, g = A∗2 g
are equal.
368 Part III: Operator Theory and Integral Equations
Since we know that the ranks of A2 and A∗2 are finite, we may represent the
mappings of the operators I − A2 and I − A∗2 as the mappings of matrices I − M2 and
I − M2∗ with adjoint matrices M2 and M2∗ . But the ranks of the adjoint matrices are
equal, and therefore the numbers of independent solutions of the equations g = A2 g
and g = A∗2 g are equal.
The next step is the following: if R(I − μ A) = H, then N(I − μ A∗ ) = {0}, and
consequently N(I − μ A) = {0} and R(I − μ A∗ ) = H (see Exercise 26.7). This means
that both (I − μ A)−1 and (I − μ A∗ )−1 exist, and the unique solutions of (34.10) are
given by
f = (I − μ A)−1 g, f = (I − μ A∗ )−1 g .
If N(I − μ A) and N(I − μ A∗ ) are not zero, then R(I − μ A) and R(I − μ A∗ ) are proper
subspaces of H, and equations (34.10) have solutions if and only if
Hence, the Fredholm alternative for these operators reads as follows: either the equa-
tions
f (x) − μ K(x, y) f (y)dy = g(x),
Ω (34.11)
f (x) − μ K(y, x) f (y)dy = g (x),
Ω
have the same (finite) number of linearly independent solutions. And in this case,
equations (34.11) are solvable if and only if g and g are orthogonal to every solution
f and f of the equations (34.12), respectively.
Definition 34.12. Equations (34.11) and (34.12) are called integral equations of the
second and first kinds, respectively.
where f ∈ L2 (a, b). Solve this equation and formulate the Fredholm alternative for
it.
on the interval [0, 1] with coefficients f , a2 ∈ L2 (0, 1), a1 ∈ W21 (0, 1) and with
smooth a0 (x) ≥ c0 > 0 subject to the boundary conditions
u(0) = u0 , u(1) = u1 .
Dividing this equation by a0 (x), we may consider the boundary value problem in
the form
u + a1 (x)u + a2 (x)u = f , u(0) = u0 , u(1) = u1 .
y(1 − x), 0 ≤ y ≤ x ≤ 1,
G(x, y) =
x(1 − y), 0 ≤ x ≤ y ≤ 1,
where 1
0 (x) = ϕ0 (x) −
ϕ G(x, y) f (y)dy
0
and
K(x, y) = ∂y G(x, y)a1 (y) + G(x, y)a1 (y) − G(x, y)a2 (y).
Exercise 34.3. (1) Prove that K(x, y) is a Hilbert–Schmidt kernel on [0, 1] × [0, 1].
(2) Prove that the boundary value problem and this integral equation of the second
kind are equivalent.
(3) Formulate the solvability condition for the boundary value problem using the
Fredholm alternative for this integral operator.
Chapter 35
Volterra and Singular Integral Equations
| f (x) − f (y)|
f Cα [a,b] = f L∞ (a,b) + sup ,
x,y∈[a,b] |x − y|α
where 0 < α ≤ 1.
The fact that f belongs to the Hölder space Cα [a, b] is equivalent to the fact that
f ∈ L∞ (a, b) and there is a constant c0 > 0 such that for all h (sufficiently small),
| f (x + h) − f (x)| ≤ c0 |h|α ,
and x
ϕ (x) = f (x) + K(x, y)ϕ (y)dy, (35.1)
a
where x ∈ [a, b] and supx,y∈[a,b] |K(x, y)| < ∞, are called Volterra integral equations
of the first and second kinds, respectively.
Theorem 35.2. For each f ∈ L∞ (a, b) the Volterra integral equation of the second
kind has a unique solution ϕ ∈ L∞ (a, b) such that
(M(x − a)) j
|ϕ j (x)| ≤ f L∞ (a,b) , j = 0, 1, . . . . (35.4)
j!
Indeed, this estimate clearly holds for j = 0. Assume that (35.4) has been proved
for some j ≥ 0. Then
x x
(M(y − a)) j
|ϕ j+1 (x)| ≤ |K(x, y)||ϕ j (y)|dy ≤ M f L∞ (a,b) dy
a a j!
x
(y − a) j (x − a) j+1
= M j+1 f L∞ (a,b) dy = M j+1 f L∞ (a,b) .
a j! ( j + 1)!
∞
(M(x − a)) j
|ϕ (x)| ≤ f L∞ (a,b) ∑ = f L∞ (a,b) eM(x−a) .
j=0 j!
Thus, the function ϕ (x) is well defined by the series (35.5), since this series is uni-
formly convergent with respect to x ∈ [a, b].
It remains now to show that this ϕ (x) solves (35.1). Since the series (35.5) con-
verges uniformly, we may integrate it term by term and obtain
35 Volterra and Singular Integral Equations 373
x ∞ x ∞
a
K(x, y)ϕ (y)dy = ∑ K(x, y)ϕ j (y)dy = ∑ ϕ j+1 (x)
j=0 a j=0
∞
= ∑ ϕ j (x) + ϕ0 (x) − f (x) = ϕ (x) − f (x).
j=1
So (35.1) holds with this ϕ . The estimate (35.3) then follows immediately from
(35.2). Finally, the uniqueness of this solution follows from (35.3) too.
In general, integral equations of the first kind are more delicate with respect to
solvability than equations of the second kind. However, in some cases, Volterra inte-
gral equations of the first kind can be treated by reducing them to equations of the
second kind. Indeed, consider for x ∈ [a, b],
x
K(x, y)ϕ (y)dy = f (x), (35.6)
a
and assume that the derivatives ∂∂Kx (x, y) and f (x) exist and are bounded and that
K(x, x) = 0 for all x ∈ [a, b]. Then, differentiating with respect to x reduces (35.6) to
x
∂K
ϕ (x)K(x, x) + (x, y)ϕ (y)dy = f (x),
a ∂x
or
x ∂K
f (x) ∂ x (x, y)
ϕ (x) = − ϕ (y)dy. (35.7)
K(x, x) a K(x, x)
Exercise 35.1. Show that (35.6) and (35.7) are equivalent if f (a) = 0.
∂K
(x, y)
∂y
exists and is bounded and that K(x, x) = 0 for all x ∈ [a, b]. In this case, setting
x
ψ (x) := ϕ (y)dy, ψ = ϕ
a
374 Part III: Operator Theory and Integral Equations
with ϕ0 = f , then it can be proved by induction that for all x ∈ [a, b] we have
j
M(x − a)1−α
|ϕ j (x)| ≤ f L∞ (a,b) , j = 0, 1, . . . .
1−α
Indeed, since this clearly holds for j = 0, assume that it has been proved for some
j ≥ 0. Then
x
|ϕ j+1 (x)| ≤ |K(x, y)||ϕ j (y)|dy
a
x
Mj
≤M |x − y|−α ((y − a)1−α ) j f L∞ (a,b) dy
(1 − α ) j a
x
M j+1
≤ ((x − a)1−α ) j f L∞ (a,b) (x − y)−α dy
(1 − α ) j a
M j+1 (x − a)1−α
≤ ((x − a)1−α ) j f L∞ (a,b)
(1 − α ) j 1−α
j+1
M(x − a)1−α
= f L∞ (a,b) .
1−α
converges uniformly on the interval [a, b], and the function ϕ defined by
∞
ϕ (x) := ∑ ϕ j (x)
j=0
solves therefore the inhomogeneous integral equation (35.1). Moreover, the follow-
ing estimates hold:
f L∞ (a,b)
|ϕ (x)| ≤ , x ∈ [a, b],
M(x − a)1−α
1−
1−α
and
f L∞ (a,b)
ϕ L∞ (a,b) ≤ .
M(b − a)1−α
1−
1−α
Exercise 35.2. Show that the Volterra integral equation of the first kind
x
ϕ (x) = λ e−(x−y) ϕ (y)dy
a
Definition 35.4. Let 0 < α < 1, ϕ ∈ Cα [−a, a], and suppose that ϕ is periodic, i.e.,
ϕ (−a) = ϕ (a). In this space an integral equation of the form
a
ϕ (x + y)dy
ϕ (x) = f (x) + λ p. v. , λ ∈ C, (35.8)
−a y
and the function ϕ is extended periodically (with period 2a) to the whole line.
Thus
a a a
ϕ (x + y)dy ϕ (x + y) − ϕ (x) ϕ (x + y) − ϕ (x)
p. v. = p. v. dy = dy,
−a y −a y −a y
and the latter integral can be understood in the usual sense for periodic ϕ ∈
Cα [−a, a], since
a a a
ϕ (x + y) − ϕ (x) |y|α aα
dy ≤ c0 dy = 2c0 ξ α −1 dξ = 2c0 . (35.10)
−a y −a |y| 0 α
Inequality (35.10) shows us that for every ϕ ∈ Cα [−a, a] the integral in (35.8) is
uniformly bounded and also periodic with period 2a. But even more is true.
Proposition 35.5. For every 2a-periodic ϕ ∈ Cα [−a, a] with 0 < α < 1 the integral
in (35.8) defines a 2a-periodic function of x that belongs to the same Hölder space
Cα [−a, a].
Proof. Let us denote by g(x) the integral in (35.8). For |h| > 0 sufficiently small we
have
a a
ϕ (x + h + y) − ϕ (x + h) ϕ (x + y) − ϕ (x)
g(x + h) − g(x) = dy − dy
−a y −a y
ϕ (x + h + y) − ϕ (x + h) ϕ (x + y) − ϕ (x)
= dy − dy
|y|≤3|h| y |y|≤3|h| y
ϕ (x + h + y) − ϕ (x) ϕ (x + y) − ϕ (x)
+ dy − dy
|y|≥3|h| y |y|≥3|h| y
=: I1 + I2 .
For the estimation of I2 we first rewrite it as (we change variables in the first integral)
35 Volterra and Singular Integral Equations 377
ϕ (z + x) − ϕ (x) ϕ (z + x) − ϕ (x)
I2 = dz − dz
|z−h|≥3|h| z−h |z|≥3|h| z
1 1
= (ϕ (z + x) − ϕ (x)) − dz
|z|≥3|h| z−h z
ϕ (z + x) − ϕ (x)
− dz.
{|z−h|≥3|h|}\{|z|≥3|h|} z−h
Then we have
|ϕ (z + x) − ϕ (x)||h|dz |ϕ (z + x) − ϕ (x)|
|I2 | ≤ + dz
|z|≥3|h| |z| · |z − h| 2|h|≤|z|≤3|h| |z − h|
z∈[−a,a]
|z|α |z|α
≤ c0 |h| dz + c0 dz
a≥|z|≥3|h| |z| · 2|z|/3 2|h|≤|z|≤3|h| |z|/2
a 3|h|
3c0
= 2· |h| ξ α −2 dξ + 4c0 ξ α −1 dξ
2 3|h| 0
a
ξ α −1 (3|h|)α (3|h|)α −1 aα −1 4c0 3α α
= 3c0 |h| + 4c0 = 3c0 |h| − + |h|
α − 1 3|h| α 1−α 1−α α
3α c0 α 4c0 3α α α 1 4
< |h| + |h| = c0 3 + |h|α , (35.12)
1−α α 1−α α
since 0 < α < 1. Estimates (35.11)–(35.12) show that this proposition is completely
proved.
If we denote by a
ϕ (x + y)dy
Aϕ (x) := p. v. (35.13)
−a y
a periodic linear operator on Cα [−a, a], 0 < α < 1, then Proposition 35.5 gives that
A is bounded in this space. But this operator is not compact there. Nevertheless, the
following holds.
Corollary 35.6. There is λ0 > 0 such that for all λ ∈ C, |λ | < λ0 , and periodic
f ∈ Cα [−a, a], 0 < α < 1, the integral equation (35.8) has a unique solution in
Cα [−a, a], 0 < α < 1.
Proof. Since the operator A from (35.13) is a bounded linear operator in the space
Cα [−a, a], it follows that
ACα →Cα ≤ c0
with some constant c0 > 0. If we choose now λ0 = 1/c0 , then for all λ ∈ C, |λ | < λ0 ,
the operator I − λ A will be invertible in the space Cα [−a, a], since
This fact implies that the integral equation (35.8) can be solved uniquely in this
space, and the unique solution ϕ can be obtained as
ϕ = (I − λ A)−1 f .
In this chapter we will study approximate solution methods for equations in a Hilbert
space H of the form
Aϕ = f , (I − A)ϕ = f (36.1)
with a bounded or compact operator A. The fundamental concept for solving equa-
tions (36.1) approximately is to replace them by the equations
An ϕn = fn , (I − An )ϕn = fn , (36.2)
respectively. For practical purposes, the approximating equations (36.2) will be cho-
sen so that they can be reduced to a finite-dimensional linear system.
We will begin with some general results that are the basis of our considerations.
Theorem 36.1. Let A : H → H be a bounded linear operator with bounded inverse
A−1 . Assume that the sequence An : H → H of bounded linear operators is norm
convergent to A, i.e.,
An − A → 0, n → ∞.
Moreover, the solutions of (36.1) and (36.2) satisfy the error estimate
−1
A
ϕn − ϕ ≤ −1
(An − A)ϕ + fn − f .
1 − A (An − A)
A−1 An = I − A−1 (A − An ).
Since −1
A (An − A) < 1
or −1
A−1 −1
n A = I − A (A − An ) ,
or −1 −1
A−1 −1
n = I − A (A − An ) A .
ϕn − ϕ = A−1 −1
n (A − An )ϕ + An ( f n − f ).
Theorem 36.2. Assume that A−1n : H → H exist for all n ≥ n0 and that their norms
are uniformly bounded for such n. Let An − A → 0 as n → ∞. Then the inverse
operator A−1 exists and
−1
−1 An
A ≤
1 − A−1
n (An − A)
for all n ≥ n0 with A−1
n (An − A) < 1.
Exercise 36.1. Prove Theorem 36.2 and obtain the error estimate in this case.
J = {An ϕ : ϕ ∈ U, n = 1, 2, . . .}
lim An ϕ = Aϕ , ϕ ∈ H.
n→∞
as n → ∞.
the inverse operators (I − An )−1 exist and the solutions of (36.1) and (36.2) satisfy
the error estimate
1 + (I − A)−1 An
ϕn − ϕ ≤ (A n − A)ϕ + f n − f .
1 − (I − A)−1 (An − A)An
Proof. By Riesz’s theorem (see Theorem 28.15), the inverse operator (I − A)−1 ex-
ists and is bounded. Due to Exercise 36.3,
(An − A)An → 0, n → ∞.
This fact allows us to conclude (as in Theorem 36.1) that (I − An )−1 exists and
(I − An )−1 ≤ 1 + (I − A)−1 An
.
1 − (I − A)−1 (An − A)An
The error estimate follows from this inequality and the representation
ϕn − ϕ = (I − An )−1 (An − A)ϕ + fn − f .
Corollary 36.5. Let An be as in Theorem 36.4. Assume that the inverse operators
(I −An )−1 exist and are uniformly bounded for all n ≥ n0 . Then the inverse (I −A)−1
exists if
(I − An )−1 (An − A)A < 1.
Theorem 36.6. Let A : H → H be a bounded linear operator with A < 1. Then
the successive approximations
ϕ = (I − A)−1 f .
It remains only to show that the successive approximations converge to ϕ for all
ϕ0 ∈ H. The definition (36.3) implies
ϕ = lim ϕn .
n→∞
in the form
n
ϕn (x) − ∑ γ j a j (x) = f (x),
j=1
where γ j = (ϕn , b j )L2 (Ω ) . This means that the solution ϕn of (36.5) is necessarily
represented as
n
ϕn (x) = f (x) + ∑ γ j a j (x) (36.6)
j=1
such that the coefficients γ j (which are to be determined) satisfy the linear system
n
γ j − ∑ γk (ak , b j )L2 (Ω ) = ( f , b j )L2 (Ω ) = f j , j = 1, 2, . . . , n. (36.7)
k=1
Hence, the solution ϕn of (36.5) (see also (36.6)) can be obtained whenever we
can solve the linear system (36.7) uniquely with respect to γ j .
Let us consider now integral equations of the second kind with compact self-
adjoint operator (36.4), i.e.,
The main idea is to approximate the kernel K(x, y) from (36.8) by the degenerate
kernel Kn (x, y) from (36.5) such that
as n → ∞ and such that in addition, the inverse operators (I − An )−1 exist and are
uniformly bounded in n.
In that case the system (36.7) is uniquely solvable, and we obtain an approximate
solution ϕn such that
384 Part III: Operator Theory and Integral Equations
ϕ − ϕn L2 (Ω ) → 0, n → ∞.
(ϕ − ϕn ) − An (ϕ − ϕn ) = (A − An )ϕ .
Then
n
K(x, y) − ∑ (K(·, y), e j )L2 e j →0
j=1
L2 (Ω ×Ω )
where b j (y) = (K(·, y), e j )L2 . The system (36.7) transforms in this case to
n
γ j − ∑ γk (ek , (e j , K(·, y))L2 )L2 (Ω ) = f j .
k=1
If, for example, e j are the normalized eigenfunctions of the operator A with corre-
sponding eigenvalues λ j , then the latter system can be rewritten as
γ j − λ jγ j = f j, j = 1, 2, . . . , n.
fj
γj = ,
1−λj
36 Approximate Methods 385
(n) (n)
We assume that the points x j and the weights α j are chosen so that
Aϕ − An ϕ 2L2
2
n
1 (n) (n) (n)
=
Ω Ω
K(x, y)ϕ (y) −
|Ω | ∑ α j K(x, x j )ϕ (x j )dy
dx → 0
j=1
(n) (n)
as n → ∞. The main problem here is to choose the weights α j and the points x j
with this approximation property. The original Nyström method was constructed for
continuous kernels K(x, y).
In Hilbert spaces it is more natural to consider projection methods.
Definition 36.8. Let A : H → H be an injective bounded linear operator. Let Pn :
H → Hn be projection operators such that dim Hn = n. For given f ∈ H, the pro-
jection method generated by Hn and Pn approximates the equation Aϕ = f by the
projection equation
Pn Aϕn = Pn f , ϕn ∈ H. (36.10)
This projection method is said to be convergent if there is n0 ∈ N such that for each
f ∈ H the approximating equation (36.10) has a unique solution ϕn ∈ Hn for all
n ≥ n0 and
ϕn → ϕ , n → ∞,
Theorem 36.9. A projection method converges if and only if there exist n0 ∈ N and
M > 0 such that for all n ≥ n0 the operators
Pn A : H → H
are invertible and the operators (Pn A)−1 Pn A : H → H are uniformly bounded, i.e.,
(Pn A)−1 Pn A ≤ M, n ≥ n0 .
386 Part III: Operator Theory and Integral Equations
ϕn − ϕ ≤ (1 + M) inf ψ − ϕ .
ψ ∈Hn
Remark 36.10. Projection methods make sense, and we can expect convergence
only if the subspaces Hn possess the denseness property
inf ψ − ϕ → 0, n → ∞.
ψ ∈Hn
Proof. By Riesz’s theorem (see Theorem 28.15), the operator I − A has a bounded
inverse. Since Pn ϕ → ϕ as n → ∞, we have Pn Aϕ → Aϕ as n → ∞, too. At the same
time, the sequence Pn A is collectively compact, since A is compact and Pn is of finite
rank. Thus, due to Exercise 36.3 we have
Then the operators (I − Pn A)−1 exist and are uniformly bounded. Indeed, writing
Bn := I + (I − A)−1 Pn A,
we obtain
Bn (I − Pn A) = (I − Pn A) + (I − A)−1 Pn A(I − Pn A)
= I − (I − A)−1 (Pn A − A)Pn A =: I − Sn .
Sn → 0, n → ∞.
36 Approximate Methods 387
(I − Pn A)−1 = (I − Sn )−1 Bn .
(I − Pn A)(ϕn − ϕ ) = Pn Aϕ − Aϕ + Pn f − f ,
Corollary 36.12. Under the assumptions of Theorem 36.11 and provided addition-
ally that
Pn A − A → 0, n → ∞,
ϕn − ϕ ≤ M Pn ϕ − ϕ ,
where M is an upper bound for the norm (I − Pn A)−1 .
Proof. The existence of the inverse operators (I − Pn A)−1 and their uniform bound-
edness follow from
and
(I − A)−1
(I − Pn A)−1 ≤ .
1 − (I − A)−1 (Pn A − A)
From
(ϕ − ϕn ) − Pn A(ϕ − ϕn ) = ϕ − Pn ϕ
388 Part III: Operator Theory and Integral Equations
and
ϕn − ϕ ≤ (I − Pn A)−1 Pn ϕ − ϕ
(Aϕn − f , g) = 0, g ∈ Hn . (36.12)
or
Pn (Aϕn − f ) = 0,
since Hn is considered here to be a Hilbert space. This is the basis for the following
Galerkin projection method.
Assume that {e j }∞j=1 is an orthonormal basis in a Hilbert space H. Considering
Hn = span(e1 , . . . , en ),
we have for the solution ϕn of the projection equation (36.10) the representation
n
ϕn (x) = ∑ γ je j. (36.13)
j=1
The task here is to find (if possible uniquely) the coefficients γ j such that ϕn from
(36.13) solves (36.10). Since (36.12) is equivalent to (36.10), we have from (36.13)
that
(Aϕn , g) = ( f , g), g ∈ Hn ,
or
(Aϕn , ek ) = ( f , ek ) = fk , k = 1, 2, . . . , n,
or
n
∑ γ j (Ae j , ek ) = fk ,
j=1
or
Mγ = f , (36.14)
γ = M −1 f .
∞
Pn ϕ − ϕ 2H = ∑ |(ϕ , e j )|2 → 0, n → ∞,
j=n+1
we may apply Theorem 36.11 and conclude the proof of this theorem.
Part IV
Partial Differential Equations
Chapter 37
Introduction
(x, y) = x1 y1 + · · · + xn yn .
Equality here occurs if and only if x = λ y for some λ ∈ R. By BR (x) and SR (x) we
denote the ball and sphere of radius R > 0 with center x:
We say that Ω ⊂ Rn , n ≥ 2, is an open set if for every x ∈ Ω there is R > 0 such that
BR (x) ⊂ Ω . If n = 1, by an open set we mean an open interval (a, b), a < b.
We say that Ω ⊂ Rn , n ≥ 2, is a closed set if Rn \ Ω is open. This is equivalent
to the fact that Ω ⊂ Ω , where Ω denotes the set of limit points of Ω , i.e.,
Ω = {y ∈ Rn : ∃{x(k) }∞ (k)
k=1 ⊂ Ω , |x − y| → 0, k → ∞}.
∂ Ω = Ω ∩ Rn \ Ω .
with 00 = 1.
We will use the shorthand notation
∂ ∂ |α |
∂j = , ∂ α = ∂1α1 · · · ∂nαn ≡ .
∂xj ∂ x1α1 · · · ∂ xnαn
This part assumes that the reader is familiar also with the following concepts:
(1) The Lebesgue integral in a bounded domain Ω ⊂ Rn and in Rn .
(2) The Banach spaces (L p , 1 ≤ p ≤ ∞, Ck ) and Hilbert spaces (L2 ). If 1 ≤ p < ∞,
then we set
1/p
L (Ω ) := { f : Ω → C measurable : f L p (Ω ) :=
p
| f (x)| dx
p
< ∞},
Ω
while
Moreover,
Ck (Ω ) := { f : Ω → C : f Ck (Ω ) := max ∑
x∈Ω |α |≤k
|∂ α f (x)| < ∞},
(3) Hölder’s inequality: Let 1 ≤ p ≤ ∞, u ∈ L p and v ∈ L p with
1 1
+ = 1.
p p
Then uv ∈ L1 and
1 1
p p
|u(x)v(x)|dx ≤ |u(x)| p dx |v(x)| p dx ,
Ω Ω Ω
if f ∈ L1 (X ×Y ).
β
∂ α ( f g) = ∑ Cα ∂ β f ∂ α −β g,
β ≤α
β α! α −β
Cα = = Cα .
β !(α − β )!
396 Part IV: Partial Differential Equations
Hypersurfaces
S ∩V = {x ∈ V : ϕ (x) = 0} .
By the implicit function theorem we can solve the equation ϕ (x) = 0 near x0 to
obtain
xn = ψ (x1 , . . . , xn−1 )
∇ϕ
ν (x) := ± ,
|∇ϕ |
(∇ψ , −1)
ν (x) = ± .
|∇ψ |2 + 1
∂u ∂u
∂ν u := ν · ∇u ≡ ν1 + · · · + νn .
∂ x1 ∂ xn
x−y 1 n
∂ ∂
ν (x) =
r
and ∂ν =
r ∑ (x j − y j ) ∂ x j = ∂ r .
j=1
37 Introduction 397
provided that the integral in question exists. The basic theorem on the existence of
convolutions is the following (Young’s inequality for convolution):
Proposition 37.3 (Young’s inequality). Let f ∈ L1 (Rn ) and g ∈ L p (Rn ),
1 ≤ p ≤ ∞. Then f ∗ g ∈ L p (Rn ) and
f ∗ gL p ≤ f L1 gL p .
Now let 1 ≤ p < ∞. Then it follows from Hölder’s inequality and Fubini’s theorem
that
p
|( f ∗ g)(x)| dx ≤
p
| f (x − y)||g(y)|dy dx
Rn Rn Rn
p/p
≤ | f (x − y)|dy | f (x − y)||g(y)| p dydx
Rn Rn Rn
p/p
≤ f L1 | f (x − y)||g(y)| p dydx
Rn Rn
p/p
≤ f L1 |g(y)| p dy | f (x − y)|dx
Rn Rn
p/p p/p +1
= f L1 gLp p f L1 = f L1 gLp p .
398 Part IV: Partial Differential Equations
1/p +1/p
f ∗ gL p ≤ f L1 gL p = f L1 gL p ,
f ∗ gr ≤ f p gq .
In particular,
f ∗ gL∞ ≤ f L p gL p .
and thus
Rn uε (x)ϕ (x)dx − ϕ (0)
≤ |x|≤√ε |uε (x)||ϕ (x) − ϕ (0)|dx
+ √ |uε (x)||ϕ (x) − ϕ (0)|dx
|x|> ε
≤ sup
√
|ϕ (x) − ϕ (0)| |uε (x)|dx + 2 ϕ L∞ √ |uε (x)|dx
|x|≤ ε Rn |x|> ε
≤ sup
√
|ϕ (x) − ϕ (0)| · uL1 + 2 ϕ L∞ √ |u(y)|dy → 0
|x|≤ ε |y|>1/ ε
as ε → 0.
37 Introduction 399
Fourier transform
Hence
−n/2
2|F f (ξ )| ≤ (2π ) | f (x + πξ /|ξ |2 ) − f (x)|dx
Rn
= (2π )−n/2
f (· + πξ /|ξ |2 ) − f (·)
L1 → 0
as |ξ | → ∞, since f ∈ L1 (Rn ).
∗ g = (2π )n/2 fg.
Exercise 37.3. Prove that if f , g ∈ L1 (Rn ), then f
∂ α f = (−ix)α f, ∂
α f = (iξ )α f.
It is clear that
F −1 f (x) = F f (−x), F −1 f = F ( f ),
If f ∈ S(Rn ), then (F −1 F ) f = f .
Exercise 37.7. Prove the Fourier inversion formula for f ∈ S(Rn ).
37 Introduction 401
Plancherel’s theorem
Exercise 37.8. Prove that if f ∈ L1 (Rn ) has compact support, then f extends to an
entire holomorphic function on Cn .
Exercise 37.9. Prove that if f ∈ C0∞ (Rn ), i.e., f ∈ C∞ (Rn ) with compact support, is
supported in {x ∈ Rn : |x| ≤ R}, then for every multi-index α we have
Distributions
We say that ϕ j → ϕ in C0∞ (Ω ), Ω ⊂ Rn open, if ϕ j are all supported in a common
compact set K ⊂ Ω and
|
u, ϕ | ≤ CK ∑ ∂ α ϕ ∞ .
|α |≤NK
δ , ϕ = ϕ (0), ϕ ∈ C0∞ (Ω ).
In particular,
lim uε (ξ ) = lim u(εξ ) = (2π )−n/2 .
ε →0+ ε →0+
∂ α u, ϕ = u, (−1)|α | ∂ α ϕ ,
f u, ϕ = u, f ϕ , f ∈ C∞ (Ω ),
,
u ∗ ψ , ϕ =
u, ϕ ∗ ψ ψ ∈ C0∞ (Ω ),
∂ α (u ∗ ψ ) = u ∗ ∂ α ψ .
u u.
∗ ψ = (2π )n/2 ψ
37 Introduction 403
pointwise for all x ∈ Ω . The equation (38.1) is said to be linear if it can be written
as
∑ aα (x)∂ α u(x) = f (x) (38.2)
|α |≤k
for some known functions aα and f . In this case we speak of the (linear) differential
operator
L(x, ∂ ) ≡ ∑ aα (x)∂ α
|α |≤k
∂u
where ϕ ∈ C0∞ (Ω ). Let us list some examples. Here and throughout we set ut = ∂t ,
∂ 2u
utt = ∂ t2
, and so forth.
ut = kΔ u.
where Δ ≡ ∇·∇ = ∂12 +· · ·+ ∂n2 is the Laplacian (or the Laplace operator).
(3) The telegrapher’s equation
utt = c2 Δ u − α ut − m2 u.
utt = c2 Δ u − sin u.
Δ 2 u ≡ Δ (Δ u) = 0.
ut + cu · ux + uxxx = 0.
χL (x, ξ ) = ∑ aα (x)ξ α , ξ ∈ Rn .
|α |=k
χL (x, ν (x)) = 0,
S = {x ∈ Rn : x1
= 0, x2 = · · · = xn = 0}
where A = (a1 , . . . , an ). This implies that charx (L) ∪ {0} is the hyperplane orthog-
onal to A, and therefore, S is characteristic at x if and only if A is tangent to S at x
(A · ν = 0). Then
408 Part IV: Partial Differential Equations
n n
∑ a j (x)∂ j u(x) = ∑ a j (x)∂ j g(x), x ∈ S,
j=1 j=1
(A · ∇)u1 − (A · ∇)u2 = (A · ∇)(ϕγ ) = γ (A · ∇)ϕ + ϕ (A · ∇)γ = 0,
du d n
∂u
dt
= (u(x(t))) =
dt ∑ ẋ j ∂ x j = (A · ∇)u = f − bu ≡ f (x(t)) − bu(x(t)),
j=1
or
du
= f − bu. (38.5)
dt
By the existence and uniqueness theorem for ordinary differential equations there
is a unique solution (unique curve) of (38.4) with x(0) = x0 . Along this curve the
solution u(x) of (38.3) must be the solution of (38.5) with u(0) = u(x(0)) = u(x0 ) =
g(x0 ). Moreover, since S is noncharacteristic, x(t) ∈
/ S for t
= 0, at least for small t,
and the curves x(t) fill out a neighborhood of S. Thus we have proved the following
theorem.
Theorem 38.2. Assume that S is a surface of class C1 that is noncharacteristic
for (38.3), and that a j , b, f , and g are real-valued C1 functions. Then for every
sufficiently small neighborhood U of S in Rn there is a unique solution u ∈ C1 of
(38.3) on U that satisfies u = g on S.
38 Local Existence Theory 409
Remark 38.3. The method that was presented above is called the method of char-
acteristics.
χL (x, ν (x)) = x2 · 0 + x1 · 1 = x1
= 0,
so that the lines x1 > 0 and x1 < 0 are noncharacteristic. The system (38.4)–
(38.5) to be solved is
ẋ1 = x2 , ẋ2 = x1 , u̇ = u,
on S. We obtain
x10 t x10 t
x1 = (e + e−t ), x2 = (e − e−t ), u = g(x10 )et .
2 2
These equations imply
So
x1 + x2
et = ± , x12 > x22 ,
x12 − x22
and thus
x1 + x2
u(x1 , x2 ) = ±g ± x12 − x22 ,
x12 − x22
where we have a plus sign for x1 + x2 > 0 and a minus sign for x1 + x2 < 0.
(2) In R2 , solve x1 ∂1 u + x2 ∂2 u = u with u(x1 , 0) = g(x1 ) on the line x2 = 0.
Compared to previous example, in this case the line x2 = 0 is characteristic,
since x1 · 0 + x2 · 1 = 0 on S. The system (38.4)–(38.5) gives in this case that
This means that u(x1 , x2 ) is a function of only one variable x1 , and the original
equation transforms to
x1 ∂1 u = u1 ,
which has only the solution u1 (x1 ) = cx1 , where c is a constant. But then we
have a contradiction, since the equality cx1 = g(x1 ) is impossible for an arbi-
trary C1 function g.
Let us consider more examples in which we apply the method of characteristics.
on S. We obtain
Therefore,
u(x) = u(x1 , x2 , x3 ) = g(x1 e−x3 , x2 e−2x3 )e3x3 .
χL (x, ν (x)) = 1 · 0 + x1 · 0 − 1 · 1 = −1
= 0,
with
(x1 , x2 , x3 )|t=0 = (x10 , x20 , 1), u(0) = x10 + x20 .
We obtain
t2
x1 = t + x10 , x2 = + tx10 + x20 , x3 = −t + 1, u = (x10 + x20 )et .
2
Then,
t = 1 − x3 , x10 = x1 − t = x1 + x3 − 1,
(1 − x3 )2 1 x2
x20 = x2 − − (1 − x3 )(x1 + x3 − 1) = − x1 + x2 − x3 + x1 x3 + 3 ,
2 2 2
and finally,
x32 1 1−x3
u= + x1 x3 + x2 − e .
2 2
is tangent to the graph y = u(x) at every point. This suggests that we look at the
integral curves of A in Rn+1 given by solving the ordinary differential equations
S∗ := {(x, g(x)) : x ∈ S}
in Rn+1 , then the graph of the solution should be the hypersurface generated by
the integral curves of A passing through S∗ . Again, we need to assume that S is
noncharacteristic in the sense that the vector
ẋ j = a j (x, u(x)), j = 1, 2, . . . , n,
with x j (0) = x0j , then writing the solution u via integral curves as y(t) = u(x(t)), we
obtain that
n n
ẏ = ∑ ∂ j u · ẋ j = ∑ a j (x, u)∂ j u = b(x, u) = b(x, y).
j=1 j=1
Thus, as in the linear case, u solves (38.6) with given initial data g on S.
Example 38.7. In R2 , solve u∂1 u+ ∂2 u = 1 with u = s/2 on the segment x1 = x2 = s,
where s > 0, s
= 2, is a parameter.
Since ϕ (s) = (s, s), it follows that (x = x1 = s)
∂ x1
∂s a1 (s, s, s/2) 1 s/2
det ∂ x2 = det = 1 − s/2
= 0,
∂s a2 (s, s, s/2) 1 1
for s > 0, s
= 2. The system (38.4)–(38.5) for this problem is
ẋ1 = u, ẋ2 = 1, u̇ = 1,
with
x10
(x1 , x2 , u)|t=0 = (x10 , x20 , ) = (s, s, s/2).
2
Then
u = t + s/2, x2 = t + s, ẋ1 = t + s/2,
t2
so that x1 = 2 + st2 + s. This implies
x1 − x2 = t 2 /2 + t(s/2 − 1).
s 1 t2 2(x1 − x2 )
= 1+ x1 − x2 − , t= .
2 t 2 x2 − 2
38 Local Existence Theory 413
Hence
2(x1 − x2 ) x1 − x2 t
u= +1+ −
x2 − 2 t 2
2(x1 − x2 ) x2 − 2 x1 − x2
= +1+ −
x2 − 2 2 x2 − 2
x1 − x2 x2 − 2 x1 − x2 x2 2x1 − 4x2 + x22
= +1+ = + = .
x2 − 2 2 x2 − 2 2 2(x2 − 2)
u∂1 u + ∂2 u = 0
1 h(x1 )
det = 1
= 0,
0 1
and ν (x) = (0, 1). Now we have to solve the ordinary differential equations
ẋ1 = u, ẋ2 = 1, u̇ = 0,
with
(x1 , x2 , u)|t=0 = x10 , 0, h(x10 ) .
We obtain
x2 = t, u ≡ h(x10 ), x1 = h(x10 )t + x10 ,
so that
x1 − x2 h(x10 ) − x10 = 0.
By this condition, the last equation defines an implicit function x10 = g(x1 , x2 ).
Therefore, the solution u of the Burgers equation has the form
ax1 + b 1
u(x1 , x2 ) = , x2
= − .
ax2 + 1 a
u, ∂ν u, . . . , ∂νk−1 u (38.7)
on S are called the Cauchy data of u on S. And the Cauchy problem is to solve
(38.2) with the Cauchy data (38.7). We shall consider Rn , n ≥ 2, to be Rn−1 × R
and denote the coordinates by (x,t), where x = (x1 , . . . , xn−1 ). We can make a change
of coordinates from Rn to Rn−1 × R so that x0 ∈ S is mapped to (0, 0) and a neigh-
borhood of x0 in S is mapped into the hyperplane t = 0. In that case ∂ν = ∂∂t on
S = {(x,t) : t = 0}, and equation (38.2) can be written in the new coordinates as
Let us now formulate and give a sketch of the proof of the famous Cauchy–
Kowalevski theorem for the linear case.
Theorem 38.12. If aα , j (x,t), ϕ0 (x), . . . , ϕk−1 (x) are real-analytic near the origin
in Rn , then there is a neighborhood of the origin on which the Cauchy problem
(38.10)–(38.9) has a unique real-analytic solution.
Proof. The uniqueness of the analytic solution follows from the fact that an analytic
function is completely determined by the values of its derivatives at one point (see
the Taylor formula or the Taylor series). Indeed, for all α and j = 0, 1, . . . , k − 1,
Therefore,
−1
∂tk u|t=0 = a0,k f (x, 0) − ∑ aα , j (x, 0)∂xα ϕ j (x) ,
|α |+ j≤k, j<k
and moreover,
−1
∂tk u(x,t) = a0,k f (x,t) − ∑ aα , j (x,t)∂xα ∂t j u .
|α |+ j≤k, j<k
416 Part IV: Partial Differential Equations
Next, let us denote by yα , j = ∂xα ∂t j u and by Y = (yα , j ) this vector. Then equation
(38.10) can be rewritten as
−1
y0,k = a0,k f− ∑ aα , j yα , j ,
|α |+ j≤k, j<k
or
−1
∂t y0,k−1 = a0,k f− ∑ aα , j ∂x j y(α −j), j ,
|α |+ j≤k, j<k
where Y , B, and Φ are analytic vector-valued functions and the A j are analytic
matrix-valued functions. Without loss of generality we can assume that Φ ≡ 0.
( j)
Let Y = (y1 , . . . , yN ), B = (b1 , . . . , bN ), A j = (aml )Nm,l=1 . We seek a solution Y =
(y1 , . . . , yN ) of the form
(m)
ym = ∑ Cα , j xα t j , m = 1, 2, . . . , N.
(m)
The Cauchy data tell us that Cα ,0 = 0 for all α and m, since we assumed Φ ≡ 0. To
(m)
determine Cα , j for j > 0, we substitute ym into (38.11) and get for m = 1, 2, . . . , N,
that
( j)
∂t ym = ∑ aml ∂x j yl + bm (x, y),
or
(m) ( j) (m) (m)
∑ Cα , j jxα t j−1 = ∑ ∑ aml
βr
xβ t r ∑ Cα , j α j xα − j t j + ∑ bα j xα t j .
j,l β ,r
(m)
It can be proved that this equation determines uniquely the coefficients Cα , j and
therefore the solution Y = (y1 , . . . , yN ).
u1 u2 + u2 u1 = 0,
and
u2 = λ u2
are √ √
u1 = A sin( λ x1 ) + B cos( λ x1 )
and √ √
u2 = C sinh( λ x2 ) + D cosh( λ x2 ),
√
respectively. But u2 (0) = 0, u2 (0) = 1 and u1 (x1 ) = ke− k sin(kx1 ). Thus D = 0,
√ √
B = 0, k = λ , A = ke− k , and C = 1k = √1 . So we finally have
λ
√ 1 √
u(x1 , x2 ) = ke− k
sin(kx1 ) sinh(kx2 ) = e− k sin(kx1 ) sinh(kx2 ).
k
As k → +∞, the Cauchy data √
and their derivatives (for x2 = 0) of all orders tend
uniformly to zero, since e− k decays faster than polynomially. But if x2
= 0 (more
precisely, x2 > 0), then
√
lim e− k
sin(kx1 ) sinh(kx2 ) = ∞,
k→+∞
(k)
if we choose, for example, x2 = 1 and x1 = π /(2k) + 2π . Hence u(x1 , x2 ) is not
bounded. But the solution of the original problem that corresponds to the limiting
case k = ∞ is of course u ≡ 0, since u(x1 , 0) = 0 and ∂2 u(x1 , 0) = 0 in the limiting
case. Hence the solution of the Cauchy problem may not depend continuously on
the Cauchy data. This means by Hadamard that the Cauchy problem for elliptic
operators is “ill-posed,” even when this problem is noncharacteristic.
Remark 38.14. This example of Hadamard’s shows that the solution of the Cauchy
problem may not depend continuously on the Cauchy data. By the terminology of
Hadamard, “the Cauchy problem for the Laplacian is not well posed, but it is ill
418 Part IV: Partial Differential Equations
posed.” Due to Hadamard and Tikhonov, a problem is called well posed if the fol-
lowing conditions are satisfied:
(1) existence;
(2) uniqueness;
(3) stability or continuous dependence on data.
Otherwise, it is called ill posed.
Let us consider one more important example due to H. Lewy. Let L be the first-order
differential operator in R3 ((x, y,t) ∈ R3 ) given by
∂ ∂ ∂
L≡ + i − 2i(x + iy) . (38.12)
∂x ∂y ∂t
Proof. Suppose x2 + y2 < R2 , |t| < R, and set z = x + iy = reiθ . Let us denote by
V (t) the function
2π
V (t) := u(x, y,t)dσ (z) = ir u(r, θ ,t)eiθ dθ ,
|z|=r 0
where u(x, y,t) is the C1 solution of the equation Lu = f with L from (38.12). We
continue to denote u in polar coordinates also by u. By the divergence theorem for
F := (u, iu) we get
∂u ∂u
i ∇ · Fdxdy ≡ i +i dxdy = i (u, iu) · ν dσ (z)
|z|<r |z|<r ∂ x ∂y |z|=r
x
y
=i u + iu dσ (z) = i ueiθ dσ (z)
|z|=r r r |z|=r
2π
= ir ueiθ dθ ≡ V (t).
0
∂u ∂u ∂u ∂u
V (t) ≡ i +i dxdy = i +i (ρ , θ ,t)ρ dρ dθ .
|z|<r ∂x ∂y 0 0 ∂x ∂y
38 Local Existence Theory 419
∂V ∂u ∂u ∂u ∂u dσ (z)
= ir +i (r, θ ,t)dθ = +i (x, y,t)2r
∂r 0 ∂ x ∂ y |z|=r ∂ x ∂ y 2z
∂ u f (t) ∂V dσ (z)
= 2r i + dσ (z) = 2r i + f (t)
|z|=r ∂ t 2z ∂ t |z|=r 2z
∂V
= 2r i + iπ f (t) .
∂t
1 ∂V ∂V
=i + π f (t) . (38.13)
2r ∂ r ∂t
Let us introduce now a new function U(s,t) = V (s) + π F(t), where s = r2 and
F = f . The function F exists because f is continuous. It follows from (38.13) that
1 ∂V ∂V ∂U ∂V ∂U ∂U
≡ , = , =i .
2r ∂ r ∂s ∂s ∂s ∂s ∂t
Hence
∂U ∂U
+i = 0. (38.14)
∂t ∂s
Since (38.14) is the Cauchy–Riemann equation, we have that U is a holomorphic
(analytic) function of the variable w = t + is, in the region 0 < s < R2 , |t| < R, and
U is continuous up to s = 0. Next, since U(0,t) = π F(t) (V = 0 when s = 0, i.e.,
r = 0) and f (t) is real-valued, it follows that U(0,t) is also real-valued. Therefore,
by the Schwarz reflection principle (see complex analysis), the formula
U(−s,t) := U(s,t)
We consider what is perhaps the most important of all partial differential operators,
theLaplace operator (Laplacian) on Rn , defined by
n
Δ= ∑ ∂ j2 ≡ ∇ · ∇.
j=1
We will begin with a quite general fact about partial differential operators.
Definition 39.1. (1) A linear transformation T on Rn is called arotation if T =
T −1 .
(2) Let h be a fixed vector in Rn . Thetranslation transformation Th f (x) := f (x + h)
is called a.
Proof. Let
L(x, ∂ ) ≡ ∑ aα (x)∂ α
|α |≤k
This implies that the aα (x) must be constants (because aα (x) ≡ aα (x + h) for all h),
say aα . Next, since L now has constant coefficients, we have (see Exercise 37.5)
ξ ) = P(ξ )
Lu( u(ξ ),
P(ξ ) = ∑ aα (iξ )α .
|α |≤k
u
◦ T (ξ ) = (
u ◦ T ) (ξ ).
Therefore,
x)(ξ ) = Lu(T
(Lu)(T ξ ),
or
P(ξ )u(T x)(ξ ) = P(T ξ )
u(T ξ ).
This forces
P(ξ ) = P(T ξ ).
shows that P(ξ ) does not depend on the angle θ of ξ . Therefore, P(ξ ) is radial, that
is,
P(ξ ) = P1 (|ξ |) = ∑ aα |ξ ||α | .
|α |≤k
P(ξ ) = ∑ a j |ξ |2 j .
j
Δ j u(ξ ) = (−1) j |ξ |2 j u
(ξ ), j = 0, 1, . . . .
u(ξ )) = F −1 ∑ a j |ξ |2 j u(ξ ) = F −1 ∑ aj Δ
Lu = F −1 (P(ξ ) j u(ξ ) =
∑ aj Δ j u.
j j j
39 The Laplace Operator 423
Conversely, let
Lu = ∑ a j Δ j u.
j
It is clear by the chain rule that the Laplacian commutes with translations Th and
rotations T . By induction, the same is true for any power of Δ , and so for L as
well.
Lemma 39.3. If f (x) = ϕ (r), r = |x|, that is, f is radial, then Δ f = ϕ (r) +
n−1
r ϕ (r).
∂r xj
Proof. Since ∂xj = r , it follows that
n n x
∑ ∂ j (∂ j ϕ (r)) = ∑ ∂ j ϕ (r)
j
Δf =
j=1 j=1 r
n x n x2j
∑ ϕ (r)∂ j +∑ ϕ (r)
j
= 2
j=1 r j=1 r
n 2 n x2j
1 xj
= ∑ − 3
r r
ϕ (r) + ∑
r2
ϕ (r)
j=1 j=1
n 1 n
n−1
= ϕ (r) − 3
r r ∑ x2j ϕ (r) + ϕ (r) = ϕ (r) + r
ϕ (r).
j=1
n−1
ϕ (r) + ϕ (r) = 0.
r
n−1
ψ (r) + ψ (r) = 0,
r
or
rn−1 ψ (r) = 0.
It follows that
ϕ (r) = ψ (r) = cr1−n .
(2)
(vΔ u + ∇v · ∇u) dx = v∂ν udσ .
Ω S
Proof. By taking real and imaginary parts, it suffices to consider real-valued func-
tions. If we let u = v in part (2) of Exercise 39.1, we obtain
|∇u|2 dx = u∂ν udσ (x).
Ω S
39 The Laplace Operator 425
2π n/2
where ωn = Γ (n/2) is the area of the unit sphere in Rn .
Proof. Let us apply Green’s identity (1) with u and v = |y|2−n if n = 2, and v = log |y|
if n = 2 in the domain
x−y x−y d d
∂ν = ν · ∇ = =
r r dr dr
for the sphere. Since u is harmonic, due to Exercise 39.2 we can get from (39.1) that
for all ε > 0, ε < r,
ε 1−n udσ (y) = r1−n udσ (y).
|x−y|=ε |x−y|=r
Therefore,
lim ε 1−n u(y)dσ (y) = lim u(x + εθ )dθ
ε →0 |x−y|=ε ε →0 |θ |=1
= ωn u(x) = r1−n u(y)dσ (y).
|x−y|=r
This proves the theorem, because the latter steps hold for n = 2 also.
426 Part IV: Partial Differential Equations
Exercise 39.3. Assumethat u is harmonic in Ω . Let χ (x) ∈ C0∞ (B1 (0)) be such
that χ (x) = χ1 (|x|) and Rn χ (x)dx = 1. Define an approximation to the identity by
χε (·) = ε −n χ (ε −1 ·). Prove that
u(x) = χε (x − y)u(y)dy
Bε (x)
for x ∈ Ωε := {x ∈ Ω : Bε (x) ⊂ Ω }.
Corollary 39.10. If u is harmonic on Ω , then u ∈ C∞ (Ω ).
Proof. The statement follows from Exercise 39.3, since the function χε is com-
pactly supported and we may thus differentiate under the integral sign as often as
we please.
Corollary 39.11. If {uk }∞
k=1 is a sequence of harmonic functions on an open set
Ω ⊂ Rn that converges uniformly on compact subsets of Ω to a limit u, then u is
harmonic on Ω .
Theorem 39.12 (The maximum principle). Suppose Ω ⊂ Rn is open and con-
nected. If u is real-valued and harmonic on Ω with supx∈Ω u(x) = A < ∞, then
either u < A for all x ∈ Ω or u(x) ≡ A in Ω .
Proof. Since u is continuous on Ω , the set {x ∈ Ω : u(x) = A} is closed in Ω . On
the other hand, we may conclude that if u(x) = A at some point x ∈ Ω , then u(y) = A
for all y in a ball about x. Indeed, if y0 ∈ Bσ (x) and u(y0 ) < A, then u(y) < A for all
y from a small neighborhood of y0 . Hence, by Corollary 39.8, for r ≤ σ ,
n
A = u(x) = u(y)dy
rn ωn |x−y|≤r
n n
= u(y)dy + n u(y)dy
rn ωn |x−y|≤r,|y0 −y|>ε r ωn |y−y0 |≤ε
n n
<A n dy + n dy
r ωn |x−y|≤r,|y0 −y|>ε r ωn |y−y0 |≤ε
n
=A n dy = A,
r ωn |x−y|≤r
39 The Laplace Operator 427
that is, A < A. This contradiction proves our statement. This fact also means that the
set {x ∈ Ω : u(x) = A} is also open. Hence it is either Ω (in this case u ≡ A in Ω )
or the empty set (in this case u(x) < A in Ω ).
Corollary 39.13. Suppose Ω ⊂ Rn is connected, open, and bounded. If u is real-
valued and harmonic on Ω and continuous on Ω , then the maximum and minimum
of u on Ω are achieved only on ∂ Ω .
Corollary 39.14 (The uniqueness theorem). Suppose Ω is as in Corollary 39.13.
If u1 and u2 are harmonic on Ω and continuous in Ω (possibly complex-valued) and
u1 = u2 on ∂ Ω , then u1 = u2 on Ω .
Proof. The real and imaginary parts of u1 − u2 and u2 − u1 are harmonic on Ω .
Hence they must achieve their maxima on ∂ Ω . These maxima are therefore zero, so
u1 ≡ u2 .
Theorem 39.15 (Liouville’s theorem). If u is bounded and harmonic on Rn , then
u ≡ constant.
Proof. For all x ∈ Rn and |x| ≤ R, by Corollary 39.8 we have
n
≤ n
|u(x) − u(0)| = u(y)dy − u(y)dy Rn ωn D |u(y)|dy,
Rn ωn BR (x) BR (0)
where
D = (BR (x)\BR (0)) ∪ (BR (0)\BR (x))
is the symmetric difference of the balls BR (x) and BR (0). Therefore, we obtain
n
u
∞ n
u
R+|x|
|u(x) − u(0)| ≤ dy ≤ n ∞ rn−1 dr dθ
R ωn R−|x|≤|y|≤R+|x|
n R ωn R−|x| |θ |=1
(R + |x|)n − (R − |x|)n 1
=
u
∞ = O
u
∞ .
Rn R
Hence the difference |u(x) − u(0)| vanishes as R → ∞, that is, u(x) = u(0).
Definition 39.16. Afundamental solution for a partial differential operator L is a
distribution K ∈ D such that
LK = δ .
Remark 39.17. Note that a fundamental solution is not unique. Any two fundamen-
tal solutions differ by a solution of the homogeneous equation Lu = 0.
Exercise 39.4. Show that the characteristic function of the set
(x1 , x2 ) ∈ R2 : x1 > 0, x2 > 0
Exercise 39.6. Show that the fundamental solution for the Cauchy–Riemann oper-
ator L = 12 (∂1 + i∂2 ) on R2 is equal to
1 1
.
π x1 + ix2
Since the Laplacian commutes with rotations (Theorem 39.2), it should have a radial
fundamental solution that must be a function of |x| that is harmonic on Rn \ {0}.
Δ Kε , ϕ → ϕ (0), ε →0
Δ Kε → δ .
= f (y)ϕ (y)dy = f , ϕ .
Rn
Hence Δ ( f ∗ K) = f .
As ε → 0, the right-hand side of this equation tends to the right-hand side of (39.5)
for each x ∈ Ω , since for x ∈ Ω and y ∈ S there are no singularities in K. On the other
hand, the left-hand side is just (u ∗ Δ Kε ) (x) if we set u ≡ 0 outside Ω . According to
the proof of Theorem 39.18,
(u ∗ Δ Kε ) (x) → u(x), ε → 0,
∂
Kε (x) = ωn−1 x j (|x|2 + ε 2 )−n/2 ,
∂xj
(39.6)
∂2 −nxi x j (|x|2 + ε 2 )−n/2−1 , i = j,
Kε (x) = ωn−1 −n/2−1
∂ xi ∂ x j (|x| + ε − nx j )(|x| + ε )
2 2 2 2 2 , i = j.
∂
K(x) = ωn−1 x j |x|−n ,
∂xj
(39.7)
∂2 −nωn−1 xi x j |x|−n−2 , i = j,
K(x) =
∂ xi ∂ x j ωn−1 (|x|2 − nx2j )|x|−n−2 , i = j,
for x = 0. Formulas (39.7) show that ∂ j K(x) is a locally integrable function, and
since g is bounded with compact support, it follows that g ∗ ∂ j K is continuous. Next,
g ∗ ∂ j Kε → g ∗ ∂ j K uniformly as ε → 0+. This is equivalent to ∂ j Kε → ∂ j K in the
topology of distributions (see the definition). Hence ∂ j (g ∗ K) = g ∗ ∂ j K.
This argument does not work for the second derivatives, because ∂i ∂ j K(x) is not
integrable. But there is a different procedure for these terms.
Let i = j. Then ∂i ∂ j Kε (x) and ∂i ∂ j K(x) are odd functions of xi (and x j ); see
(39.6) and (39.7). Due to this fact, their integrals over an annulus 0 < a < |x| < b
vanish. For Kε we can even take a = 0.
432 Part IV: Partial Differential Equations
If we let ε → 0, we obtain
lim g ∗ ∂i ∂ j Kε (x)
ε →0
= (g(x − y) − g(x))∂i ∂ j K(y)dy + g(x − y)∂i ∂ j K(y)dy.
|y|<b |y|≥b
1
∂ j2 Kε (x) = ε −n ψ (ε −1 x) + K εj (x),
n
where ψ (x) = nωn−1 (|x|2 +1)−n/2−1 and K εj = ωn−1 (|x|2 −nx2j )(|x|2 + ε 2 )−n/2−1 (see
(39.6)). The integral I j of K εj over an annulus a < |y| < b vanishes. Why is that so?
First of all, I j is independent of j by symmetry in the coordinates, that is, I j = Ii for
i = j. So nI j is the integral of ∑nj=1 K εj . But ∑nj=1 K εj = 0. Hence I j = 0 also. We can
therefore apply the same procedure. Since
g ∗ (ε −n ψ (ε −1 x)) → g, ε →0
Since the convergence in (39.8) and (39.9) is uniform, at this point we have shown
that g ∗ K ∈ C2 . But we need to prove more.
for all 0 < a < b < ∞. Then if g is a Cα function with compact support, 0 < α < 1,
then
h(x) = lim (g(x − z) − g(x))N(z)dz
b→∞ |z|<b
belongs to Cα .
and hence
|h1 (x + y) − h1 (x)| ≤ |h1 (x + y)| + |h1 (x)| ≤ 2c |y|α .
=: I1 + I2 .
434 Part IV: Partial Differential Equations
It is clear that
{3|y| < |z + y|} \ {3|y| < |z|} ⊂ {2|y| < |z|} \ {3|y| < |z|}
= {2|y| < |z| ≤ 3|y|} .
Therefore,
|I2 | ≤ |g(x − z) − g(x)||N(z + y)|dz
2|y|<|z|≤3|y|
≤c |z|α |z + y|−n dz ≤ c |z|α −n dz = c |y|α .
2|y|<|z|≤3|y| 2|y|<|z|≤3|y|
Note that the condition α < 1 is needed here. Collecting the estimates for I1 and I2 ,
we can see that the lemma is proved.
In order to end the proof of Theorem 39.22 it remains to note that ∂i ∂ j K(x)
satisfies all the conditions of Lemma 39.23.
Exercise 39.12. Show that a function K1 is a fundamental solution for Δ 2 ≡ Δ (Δ )
on Rn if and only if K1 satisfies the equation
Δ K1 = K,
(2) n = 2:
|x|2 log |x|
;
8π
39 The Laplace Operator 435
(3) n = 2, 4:
|x|4−n
.
2(4 − n)(2 − n)ωn
Exercise 39.14. Show that (4π |x|)−1 e−c|x| is the fundamental solution for −Δ + c2
on R3 for an arbitrary constant c ∈ C.
Chapter 40
The Dirichlet and Neumann Problems
We assume that Ω is bounded with C1 boundary. But we shall not, however, assume
that Ω is connected. The uniqueness theorem (see Corollary 39.14) shows that the
solution of (D) will be unique (if it exists), at least if we require u ∈ C(Ω ). For (N)
uniqueness does not hold: we can add to u(x) any function that is constant on each
connected component of Ω . Moreover, there is an obvious necessary condition for
solvability of (N). If Ω is a connected component of Ω , then
Δ udx = ∂ν udσ (x) = g(x)dσ (x) = f dx,
Ω ∂Ω ∂Ω Ω
that is,
f (x)dx = g(x)dσ (x).
Ω ∂Ω
It is also clear (by linearity) that (D) can be reduced to the following homoge-
neous problems:
Δ v = f , in Ω
(DA )
v = 0, on S
Δ w = 0, in Ω
(DB )
w = g, on S
and u = v + w.
Definition 40.1. The Green’s function for (D) in Ω is the solution G(x, y) of the
boundary value problem
Δx G(x, y) = δ (x − y), x, y ∈ Ω
(40.1)
G(x, y) = 0, x ∈ S, y ∈ Ω .
Analogously, the Green’s function for (N) in Ω is the solution G(x, y) of the bound-
ary value problem
Δx G(x, y) = δ (x − y), x, y ∈ Ω
(40.2)
∂νx G(x, y) = 0, x ∈ S, y ∈ Ω .
where K is the fundamental solution of Δ in Rn and for all y ∈ Ω , the function vy (x)
satisfies
Δ vy (x) = 0, in Ω
(40.4)
vy (x) = −K(x − y), on S
in the case of (40.2). Since (40.4) guarantees that vy is real, it follows that so is G
corresponding to (40.1).
Lemma 40.2. The Green’s function (40.1) exists and is unique.
Proof. The uniqueness of G follows again from Corollary 39.14, since K(x − y) in
(40.4) is continuous for all x ∈ S and y ∈ Ω (x = y). The existence will be proved
later.
Lemma 40.3. For both (40.1) and (40.2) it is true that G(x, y) = G(y, x) for all
x, y ∈ Ω .
Proof. Let G(x, y) and G(x, z) be the Green’s functions for Ω corresponding to
sources located at fixed y and z, y = z, respectively. Let us consider the domain
Ωε = (Ω \ {x : |x − y| < ε }) \ {x : |x − z| < ε } ,
where
I2 = G(x, y)∂νx G(x, z)dσ (x).
|x−z|=ε
and
1 1−n n−1
I2 ≈ ε ε G(εθ + z, y)dθ → G(z, y), ε → 0.
ωn |θ |=1
This means that G(y, z) = G(z, y) for all z = y. This proof holds for n = 2 (and even
for n = 1) with some simple changes.
Proof. For each fixed y, the function vy (x) := G(x, y) − K(x − y) is harmonic in Ω ;
see (40.4). Moreover, on S = ∂ Ω , vy (x) takes on the positive value
40 The Dirichlet and Neumann Problems 441
|x − y|2−n
−K(x − y) ≡ − .
ωn (2 − n)
By the minimum principle, it follows that vy (x) is strictly positive in Ω . This proves
the first inequality.
Exercise 40.2. Show that for n = 2, Lemma 40.4 has the following form:
1 |x − y|
log < G(x, y) < 0, x, y ∈ Ω ,
2π h
Exercise 40.3. Obtain the analogue of Lemma 40.4 for n = 1. Hint: show that the
d2
Green’s function for the operator dx 2 on Ω = (0, 1) is
x(y − 1), x < y
G(x, y) =
y(x − 1), x > y.
Now we can solve both problems (DA ) and (DB ). Indeed, let us set f = 0 in (DA )
outside Ω and define
v(x) := G(x, y) f (y)dy ≡ ( f ∗ K)(x) + (G(x, y) − K(x − y)) f (y)dy.
Ω Ω
Then the Laplacian of the first term is f (see Theorem 39.19), and the second term
is harmonic in x (since vy (x) is harmonic). Also v(x) = 0 on S, because the same is
true for G. Thus, this v(x) solves (DA ).
Consider now (DB ). We assume that g is continuous on S and we wish to find
w that is continuous on Ω . Applying Green’s identity (1) (together with the same
limiting process as in the proof of Lemma 40.3), we obtain
w(x) = (w(y)Δy G(x, y) − G(x, y)Δ w(y)) dy
Ω
= w(y)∂νy G(x, y)dσ (y) = g(y)∂νy G(x, y)dσ (y).
S S
Let us denote the last integral by (P). Since ∂νy G(x, y) is harmonic in x and continu-
ous in y for x ∈ Ω and y ∈ S, then w(x) is harmonic in Ω . In order to prove that this
w(x) solves (DB ), it remains to prove that w(x) is continuous in Ω and w(x) on S is
g(x). We will prove this general fact later.
442 Part IV: Partial Differential Equations
Definition 40.6. The function ∂νy G(x, y) on Ω × S is called the Poisson kernel for
Ω , and (P) is called the Poisson integral.
Now we are in a position to solve the Dirichlet problem in a half-space. Let
Ω = Rn+1
+ = (x , xn+1 ) ∈ R
n+1
: x ∈ Rn , xn+1 > 0 ,
Δn+1 = Δn + ∂t2 , n = 1, 2, . . . .
because for t, s > 0, −t − s < 0, and therefore, δ (−t − s) = 0. Thus G is the Dirichlet
Green’s function for Rn+1+ . From this we immediately have the solution of (DA ) in
R+ as
n+1
∞
u(x,t) = G(x, y;t, s) f (y, s)dsdy.
Rn 0
To solve (DB ) we compute the Poisson kernel for this case. Since the outward normal
∂
derivative on ∂ Rn+1
+ is − ∂ t , the Poisson kernel becomes
∂ ∂
− G(x, y;t, s)|s=0 = − (K(x − y,t − s) − K(x − y, −t − s)) |s=0
∂s ∂s
2t
= n+1 . (40.6)
ωn+1 (|x − y|2 + t 2 ) 2
2t
Pt (x) := n+1 , (40.8)
ωn+1 (|x|2 + t 2 ) 2
u(·,t) − g(·)∞ → 0
as t → 0+.
Proof. It is clear that for all t > 0, Pt (x) ∈ L1 (Rn ) ∩ L∞ (Rn ); see (40.8). Hence
Pt (x) ∈ Lq (Rn ) for all q ∈ [1, ∞] with respect to x and fixed t > 0. Therefore, the
integral in (40.9) is absolutely convergent, and the same is true if Pt is replaced by
its derivative Δx Pt or ∂t2 Pt (due to Young’s inequality for convolution).
Since G(x, y;t, s) is harmonic for (x,t) = (y, s), it follows that Pt (x) is also har-
monic and
Δx u + ∂t2 u = g ∗ (Δx + ∂t2 )Pt = 0.
u(·,t) − g(·)∞ → 0
Remark 40.8. The solution of this problem is not unique: if u(x,t) is a solution,
then so is u(x,t) + ct for all c ∈ C. However, we have the following theorem.
Proof. Assume for the moment that g has compact support, say g = 0 for |x| > R.
Then g ∈ L1 (Rn ) and
g ∗ Pt ∞ ≤ g1 Pt ∞ ≤ ct −n ,
2t
|u(x,t)| ≤ g1 sup |Pt (x − y)| = g1 sup n+1 ≤ cT |x|−n−1 ,
|y|<R |y|<R ωn+1 (|x − y|2 + t 2 ) 2
for |x| > 2R. Hence u(x,t) → 0 as x → ∞ uniformly for t ∈ [0, T ]. This proves that
u(x,t) vanishes at infinity if g(x) has compact support. For general g, choose a se-
quence {gk } of compactly supported functions that converges uniformly (in L∞ (Rn ))
to g and let
uk (x,t) = (gk ∗ Pt )(x).
Then
uk − uL∞ (Rn+1 ) = sup (gk − g)(y)Pt (x − y)dy
t,x R n
≤ sup gk − gL∞ (Rn ) sup |Pt (x − y)|dy
t x Rn
= gk − gL∞ (Rn ) sup |Pt (y)|dy = gk − gL∞ (Rn ) → 0
t>0 Rn
as k → ∞.
40 The Dirichlet and Neumann Problems 445
Hence u(x,t) vanishes at infinity. Now suppose v is another solution and let
w := v − u. Then w vanishes at infinity and also at t = 0 (see Theorem 40.7). Thus
|w| < ε on the boundary of the cylindrical region {(x,t) : |x| < R, 0 < t < R} for R
sufficiently large, see Figure 40.2.
t =R
R t =0
But since w is harmonic, it follows by the maximum principle that |w| < ε in this
region. Letting ε → 0 and R → ∞, we conclude that w ≡ 0.
Let us consider now the Dirichlet problem in a ball. We use here the following
notation:
B = B1 (0) = {x ∈ Rn : |x| < 1} , ∂ B = S.
for x, y ∈ Rn , x = 0, |y| = 1.
and y = x
|x|2
. Hence,
x
G(x, y) − K(x − y) ≡ −K − y|x|
|x|
is harmonic in y. But the symmetry of G and K shows also that G(x, y) − K(x − y)
is harmonic in x. Thus, G(x, y) is the Green’s function for B. This also makes clear
how to define G at x = 0 (and at y = 0):
1
G(0, y) = (|y|2−n − 1),
(2 − n)ωn
since
x
− y|x| → 1
|x|
as x → 0.
For n = 2 the analogous formulae are
1 x 1
G(x, y) = log |x − y| − log − y|x| ,
G(0, y) = log |y|.
2π |x| 2π
Now we can compute the Poisson kernel P(x, y) := ∂νy G(x, y), x ∈ B, y ∈ S. Since
∂νy = y · ∇y on S, it follows that
⎛ ⎞
1 ⎝ (y, x − y)
x
|x| − y|x|, y|x| 1 − |x|2
P(x, y) = − − n ⎠ ≡ , n ≥ 2. (40.11)
ωn |x − y|n x ωn |x − y|n
|x| − y|x|
with x = ry. This proves (40.12). We claim also that for all y0 ∈ S and for a neigh-
borhood Bσ (y0 ) ⊂ S,
lim P(ry0 , y)dσ (y) = 0. (40.13)
r→1− S\Bσ (y0 )
and therefore
|ry0 − y|−n < (r|y0 − y|)−n ≤ (rσ )−n
as r → 1−.
Exercise 40.8. Show that the Poisson kernel for the ball BR (x0 ) is
R2 − |x − x0 |2
P(x, y) = , n ≥ 2.
ωn R|x − y|n
1−r 1+r
u(0) ≤ u(x) ≤ u(0).
(1 + r)n−1 (1 − r)n−1
in B\Bδ (0). These functions are real (as we can assume without loss of generality),
harmonic, and continuous for δ ≤ |x| ≤ 1. Moreover, gε (x) = 0 on ∂ B and gε (x) < 0
on ∂ Bδ (0) for all δ sufficiently small. By the maximum principle, gε (x) is negative
in B\ {0}. Letting ε → 0, we see that u − v ≤ 0 in B\ {0}. By the same arguments
we may conclude that also v − u ≤ 0 in B\ {0}. Hence u = v in B\ {0}, and we can
extend u to the whole ball by setting u(0) = v(0). This proves the theorem.
Chapter 41
Layer Potentials
as x → 0.
Remark 41.3. This definition implies the following behaviour of u at infinity
o(1), n = 2
u(y) =
o(log |y|), n = 2
as y → ∞.
where ∂r u ≡ d
dr u. Since for n > 2 and for large |x| we have
u(x) = O |x|2−n , ∂r u(x) = O |x|1−n ,
it follows that
u∂r udσ (x)
≤ cr r
2−n 1−n
dσ (x) = cr3−2n rn−1 = cr2−n → 0
∂ Br (0) ∂ Br (0)
41 Layer Potentials 453
as r → ∞. Hence
|∇u|2 dx = 0.
Ω
Definition 41.7. The functions u(x) from (41.2) and (41.3) are called the double
and single layer potentials with moment (density) ϕ , respectively.
and
|I(x, y)| ≤ c1 + c2 |log |x − y|| , α = 0,
Remark 41.9. Note that a continuous kernel of order 0 is also a continuous kernel
of order α , 0 < α < n − 1.
with kernel I.
Proof. It is enough to consider 0 < α < n − 1. Let us assume that f ∈ L1 (S). Then
I f ≤ |I(x, y)|| f (y)|dσ (y)dσ (x)
L1 (S) S S
d
I f ≤ c f L∞ (S) rn−2−α dr = c f L∞ (S) .
L∞ (S) 0
is compact in L2 (S).
On the other hand, due to estimates for convolution,
2 1/2
I f − Iε f ≤c | f (y)||x − y| −α
dσ (y) dσ (x)
L2 (S) |x−y|<ε
ε
≤ c f L2 (S) rn−2−α dr → 0, ε → 0.
0
|y − z| ≥ |x − z| − |x − y| > 2δ − δ = δ .
I(x, z) − I(y, z) → 0, x → y,
=ϕ
Write Iu Iu + (1
− ϕ )Iu =: I0 u + I1 u. By the Cauchy–Bunyakovsky–Schwarz
inequality we have
1/2
|I1 u(x) − I1 u(y)| ≤ u2 |I1 (x, z) − I1 (y, z)|2 dσ (z) → 0, y → x,
S
− I1 u ≡ u + I0 u,
g := u + Iu
then g is continuous for u ∈ L2 (S) by the conditions of this lemma. Since the operator
norm of I0 can be made less that 1 on L2 (S) and L∞ (S) (we can do this due to the
choice of ε > 0 sufficiently small), then
−1
u = I + I0 g,
where I is the identity operator. Since g is continuous and the operator norm is less
than 1, we have
∞ j
u = ∑ −I0 g.
j=0
First of all,
(x − y, ν (y))
∂νy K(x − y) = − . (41.4)
ωn |x − y|n
by Taylor’s expansion.
c|x − y|2
|I(x, y)| ≤ = c |x − y|2−n
ωn |x − y|n
by Lemma 41.12.
Lemma 41.14. ⎧
⎪
⎨1, x ∈ Ω ,
I(x, y)dσ (y) = 0, x ∈ Ω , (41.5)
S ⎪
⎩1
2 , x ∈ S.
458 Part IV: Partial Differential Equations
or
δ 1−n
= I(x, y)dσ (y) − dσ (y) = I(x, y)dσ (y) − 1,
S ωn |x−y|=δ S
or
∂ Bδ+
δ S
Bδ+
Ω x Ω
Bδ−
Sδ Ωδ ∂ Bδ−
0= Δy K(x − y)dy = ∂νy K(x − y)dσ (y) − ∂νy K(x − y)dσ (y).
Ωδ Sδ ∂ B−
δ
lim ∂νy K(x − y)dσ (y) = lim ∂νy K(x − y)dσ (y)
δ →0 Sδ δ →0 ∂ B−
δ
δ 1−n
= lim dσ (y)
δ →0 ωn ∂ B−
δ
δ 1−n n−1 ωn 1
= lim δ + o(δ n−1 ) = .
δ →0 ωn 2 2
This means that the limit in (41.6) exists and (41.5) is satisfied.
Lemma 41.15. There exists c > 0 such that
c
|∂νy K(x − y)|dσ (y) ≤ |x − y|2−n dσ (y) ≤ c1 , x ∈ S.
S ωn S
Next, for x ∈
/ S define dist(x, S) = infy∈S |x − y|.
There are two possibilities now: if dist(x, S) ≥ δ /2, then |x − y| ≥ δ /2 for all
y ∈ S, and therefore
Set Bδ = {y ∈ S : |x0 − y| < δ }. We estimate the integrals of |I(x, y)| over S\Bδ and
Bδ separately. If y ∈ S\Bδ , then
|x − y| ≥ |x0 − y| − |x − x0 | > δ − δ /2 = δ /2
and
|I(x, y)| ≤ cδ 1−n ,
so that the integral over S\Bδ satisfies (41.7), where again c does not depend on δ .
460 Part IV: Partial Differential Equations
The latter inequality follows from Lemma 41.12, since x0 , y ∈ S. Moreover, we have
(due to Lemma 41.12)
1
|x − y|2 ≥ |x − x0 |2 + |x0 − y|2
2
and (see (41.4) and (41.8))
|x − x0 | + |x0 − y|2
|I(x, y)| ≤ c
(|x − x0 |2 + |x0 − y|2 )n/2
|x − x0 | c
≤c + .
(|x − x0 |2 + |x0 − y|2 )n/2 |x0 − y|n−2
This implies
δ
δ n−2
|x − x0 | r
|I(x, y)|dσ (y) ≤ c rn−2 dr + c dr
Bδ 0 (|x − x0 | + r )
2 2 n/2 0 rn−2
∞
arn−2
≤ c δ + c dr,
0 (a2 + r2 )n/2
If we combine all estimates, then we may conclude that there is c0 > 0 such that
ϕ (x)
lim u(x + t ν (x)) = + I(x, y)ϕ (y)dσ (y),
t→0− 2 S
ϕ (x)
lim u(x + t ν (x)) = − + I(x, y)ϕ (y)dσ (y)
t→0+ 2 S
u(x + t ν (x)) = ϕ (y)I(xt , y)dσ (y) = (ϕ (y) − ϕ (x))I(xt , y)dσ (y) + ϕ (x)
S S
Uniform convergence follows from the fact that S is compact and ϕ ∈ C(S).
Corollary 41.17. For x ∈ S,
1
≤ cδ (or δ log for n = 2) + ϕ ∞ |K(x − y) − K(x0 − y)|dσ (y) → 0
δ S\Bδ
as x → x0 and δ → 0.
δ, n > 2,
(|K(x − y)| + |K(x0 − y)|) |ϕ (y)|dσ (y) ≤ c ϕ ∞
Bδ δ log δ1 , n = 2.
(x − y, ν (x))
I ∗ (x, y) := ∂νx K(x − y) ≡ .
ωn |x − y|n
Theorem 41.21. Suppose ϕ ∈ C(S) and u is defined on Rn by the single layer po-
tential (41.3) with moment ϕ . Then for x ∈ S,
ϕ (x)
lim ∂ν u(x + t ν (x)) = − + I ∗ (x, y)ϕ (y)dσ (y),
t→0− 2 S
ϕ (x)
lim ∂ν u(x + t ν (x)) = + I ∗ (x, y)ϕ (y)dσ (y).
t→0+ 2 S
Since
(x − y, ν (y))
I(x, y) = −
ωn |x − y|n
(x − y, ν (x)) (x − y, ν (x0 ))
I ∗ (x, y) = I(y, x) = ≡ . (41.10)
ωn |x − y|n ωn |x − y|n
Hence
(x − y, ν (x0 ) − ν (y))
|x − y||ν (x0 ) − ν (y)|
|I(x, y) + I ∗ (x, y)| =
≤
ωn |x − y|n ωn |x − y|n
|x − y||x0 − y| |x0 − y|
≤c ≤ c = c |x0 − y|2−n ,
ωn |x − y|n |x0 − y|n−1
464 Part IV: Partial Differential Equations
because |x0 − y| ≤ |x0 − x| + |x − y| ≤ 2|x − y|. Here we have also used the fact that
|ν (x0 ) − ν (y)| ≤ c|x0 − y|, since ν is C1 (Figure 41.2).
x0
This estimate allows us to obtain that the corresponding integral over Bδ can be
dominated by
δ
c |x0 − y|2−n dσ (y) = c r2−n rn−2 dr = c δ .
|y−x0 |≤δ 0
1
Iϕ (x) + I∗ ϕ (x) = v− (x) + ∂ν− u(x) = ϕ (x) + Iϕ (x) + ∂ν− u(x).
2
It follows that
ϕ (x) ∗
∂ν− u(x) = − + I ϕ (x).
2
By similar arguments we obtain
1
Iϕ (x) + I∗ ϕ (x) = v+ (x) + ∂ν+ u(x) = − ϕ (x) + Iϕ (x) + ∂ν+ u(x)
2
and therefore
ϕ (x) ∗
∂ν+ u(x) = + I ϕ (x).
2
This completes the proof.
Corollary 41.22.
ϕ (x) = ∂ν+ u(x) − ∂ν− u(x),
ϕ dσ = f dσ , f dσ = 0,
S S S
respectively.
Proof. It follows from (41.10) that
1
f (x)dσ (x) = ϕ (x)dσ (x) + ϕ (y)dσ (y) I ∗ (x, y)dσ (x)
S 2 S S S
1 1
= ϕ (x)dσ (x) + ϕ (y)dσ (y) = ϕ (y)dσ (y),
2 S 2 S S
ϕ (x)dσ (x) = 0,
S
ϕ (x)dσ (x) = 0
S
1
u(x) = log |x − y|ϕ (y)dσ (y)
2π S
1 1
= (log |x − y| − log |x|) ϕ (y)dσ (y) + log |x| ϕ (y)dσ (y).
2π S 2π S
But log |x − y| − log |x| → 0 as x → ∞ uniformly for y ∈ S, and therefore, this term
is harmonic at infinity(we have a removable singularity). Hence u is harmonic at
infinity if and only if S ϕ (x)dσ (x) = 0, and in this case u(x) vanishes at infinity.
This proves part (1).
In part (2), u is harmonic at infinity. If u is constant on Ω , then it solves (ED) with
f ≡ constant on S. But a solution of such a problem must be constant and vanish at
infinity. Therefore, this constant is zero. Thus u ≡ 0.
Remark 41.25. For n > 2 the single layer potential u is a harmonic function at in-
finity without any additional conditions for the moment ϕ .
466 Part IV: Partial Differential Equations
For solvability of the corresponding integral equations in the space C(S) with
integral operators I and I∗ (see Theorems 41.16 and 41.21) we need the Fredholm
alternative (see in addition Theorems 34.8 and 34.9).
Theorem 41.26 (First Fredholm theorem). The null spaces of 12 I − I and 12 I − I∗
have the same finite dimension
1 1 ∗
dim N I − I = dim N I − I < ∞,
2 2
i.e., 2I and 2I∗ are identical on the corresponding null spaces. Since they are com-
pact there, this is possible only when the corresponding null spaces are of finite
dimension. The equality of these dimensions can be checked in the same manner as
in the proof of Theorem 34.11. In this proof, part (2) of Lemma 41.11 must be taken
into account.
and
1 1
R I − I∗ = {g ∈ C(S) : (g, ϕ )L2 (S) = 0 for any ϕ ∈ N
I − I }.
2 2
ϕ
Proof. Let f = 2 − Iϕ for some ϕ ∈ C(S). Then for all ψ ∈ N( 12 I − I∗ ) we have
ϕ
( f , ψ )L2 (S) = ( − I ϕ , ψ )L2 (S)
2
ϕ ψ
= ( , ψ )L2 (S) − (ϕ , I∗ ψ )L2 (S) = (ϕ , − I∗ ψ )L2 (S) = 0.
2 2
1 1 1
0 = ( f , ϕ − Iϕ )L2 (S) = ( f , ϕ )L2 (S) − (I∗ f , ϕ )L2 (S) = ( f − I∗ f , ϕ )L2 (S) .
2 2 2
This means that f ∈ N( 12 I − I∗ ). But at the same time, f ⊥N( 12 I − I∗ ). Thus f = 0.
This contradiction proves the opposite embedding. For the operator 12 I − I∗ the proof
is the same, since (I∗ )∗ = I.
Since the ranges of 12 I − I and 12 I − I∗ are closed due to Riesz’s lemma (see Theo-
rem 28.14) and due to part (2) of Lemma 41.11, we obtain the following result.
Theorem 41.28 (Fredholm alternative). Either 12 I − I and 12 I − I∗ are bijective or
∗
2 I − I and 2 I − I have nontrivial null spaces with finite dimension
1 1
1 1
dim N I − I = dim N I − I∗ < ∞,
2 2
and
1 1
R I − I∗ = {g ∈ C(S) : (g, ϕ )L2 (S) = 0 for any ϕ ∈ N I − I }.
2 2
1 1
ϕ − Iϕ = f , ψ − I∗ ψ = g (41.11)
2 2
have unique solutions ϕ and ψ for every given f and g from C(S), or the corre-
sponding homogeneous equations
1 1
ϕ − Iϕ = 0, ψ − I∗ ψ = 0
2 2
have the same number of linearly independent solutions ϕ1 , . . . , ϕm , ψ1 , . . . , ψm , and
in this case equations (41.11) have solutions if and only if f ⊥ϕ j , j = 1, 2, . . . , m,
and g⊥ψ j , j = 1, 2, . . . , m, respectively.
In fact, it is possible to prove a stronger result (which is the analogue of Theo-
rem 28.15) for the complete normed space C(S); see [22] for a proof.
Theorem 41.29 (Riesz). Let A : C(S) → C(S) be a compact linear operator. Then
I − A is injective if and only if it is surjective. If I − A is injective (and therefore also
bijective), then the inverse operator (I − A)−1 : C(S) → C(S) is bounded.
468 Part IV: Partial Differential Equations
and then
1 1 ∗
1 ∗
1
C(S) = R I −I ⊕N I −I = R I −I ⊕N I −I .
2 2 2 2
Theorem 41.30 (Main theorem). Suppose Ω and Ω are simply connected. Then
(1) (ID) has a unique solution for every f ∈ C(S).
(2) (ED) has a unique solution for every f ∈ C(S).
(3) (IN) has a solution for every f ∈ C(S) if and only if S f dσ = 0. The solution
is unique up to a constant.
(4) (EN) has a unique solution for every f ∈ C(S) if and only if S f dσ = 0.
Proof. We have already proved uniqueness (see Theorem 41.4) and the necessity of
the conditions on f (see Exercise 39.2 and Lemma 41.24). So all that remains is to
establish existence.
For (IN) and (EN) the function f must satisfy the condition
f dσ = 0,
S
or
( f , 1)L2 (S) = 0.
Next, since
i.e., dim N( 1 I − I)
we may conclude first that 1 ∈ N( 12 I − I), ≥ 1, and dim N( 1 I −
2 2
∗
I ) ≤ 1 due to the fact that the single layer potential uniquely (up to a constant; see
Theorem 41.4) solves (IN). Thus, due to Theorem 41.28 (Fredholm alternative), we
have
1 1 ∗
dim N I − I = dim N I − I = 1.
2 2
Using again this alternative, we see that the condition ( f , 1)L2 (S) = 0 is necessary and
sufficient for the solvability of the equation − 12 ϕ + I∗ ϕ = f , which solves (IN). For
(EN) we can solve uniquely the equation 12 ϕ + I∗ ϕ = f if and only if f ⊥N( 12 I + I).
But since the solution of (ED) is unique, the null space of I + I consists only of the
1
2
41 Layer Potentials 469
trivial solution (we have used here Lemma 41.24 for n = 2). Therefore, the necessary
and sufficient condition for f is automatically satisfied.
Concerning (ID) we consider the integral equation 12 ϕ + Iϕ = f . Let us set
1
ψ := f − ϕ + Iϕ ,
2
1
ψ + I∗ ψ = 0.
2
If we consider now the single layer potential with moment ψ ,
Hence ∂ν+ v(x) ≡ 0 for all x ∈ Ω (due to the uniqueness result), which means that
v ≡ constant in Rn \ Ω , and since v is required to be harmonic at infinity, this constant
is equal to zero (see Theorem 41.4 and Lemma 41.24 for n = 2). The final step is that
a single layer potential v is continuous everywhere in the whole of Rn and v+ ≡ 0 on
Ω including S. Thus v ≡ 0 in Ω as well. The latter fact can be proved if we consider
the Dirichlet boundary value problem
Δ v = 0, in Ω ,
v = 0, on S.
where ν > 0 is constant. It is easy to see that for the operator (42.1) with constant
coefficients, the condition (42.2) is equivalent to
Re ∑ aαβ ξ α ξ β ≥ ν ∑ |ξ α |2 .
|α |=|β |=m |α |=m
The Dirichlet boundary value problem for the operator L from (42.1) on the domain
Ω can be formulated as follows: given f ∈ L2 (Ω ), find a function u satisfying Lu = f
on Ω in the distributional sense, i.e., u is a distributional solution satisfying the
boundary conditions
u|S = g0 , ∂ν u|S = g1 , . . . , ∂νm−1 uS = gm−1 , (42.4)
1
L := (∂x2 + 2i∂x ∂y − ∂y2 )
4
Hence we have no uniqueness for this problem in B. The reason is that L is elliptic
but not strongly elliptic.
Exercise 42.1. Show that L from Remark 42.1 is elliptic on R2 but not strongly
elliptic in the sense of (42.2).
us := uH s (Ω ) .
Theorem 42.2. Suppose that the operator L from (42.1) satisfies the ellipticity con-
dition (42.2) and the coefficients aαβ (x), |α |, |β | ≤ m belong to the Sobolev space
W∞k (Ω ), k = 0, 1, 2, . . .. Then there is a constant C > 0 such that for all u ∈ H0m+k (Ω )
we have
42 Elliptic Boundary Value Problems 473
Proof. Since C0∞ (Ω ) is dense in the Sobolev space H0s (Ω ), s ≥ 0, let us consider
u ∈ C0∞ (Ω ). Consider first the case k = 0. Then integration by parts leads to
|(Lu, u)L2 (Ω ) |
= ∑ aαβ (x)∂ β u∂ α udx + ∑ aαβ (x)∂ β u∂ α udx
|α |=|β |=m Ω |α |=|β |≤m−1 Ω
≥ Re ∑ aαβ (x)∂ β u∂ α udx −C ∑ |∂ β u∂ α u|dx
|α |=|β |=m Ω |α |=|β |≤m−1 Ω
≥ν ∑ |∂ α u|2 dx −C u2m−1 ,
|α |=m Ω
where the constant C > 0 limits the norms of the coefficients aαβ (x) in L∞ (Ω ).
Using now the Poincaré inequality (Theorem 20.23) and the properties of Sobolev
space, we obtain from the latter inequality that
Thus (42.5) is proved for k = 0. Consider now the case k = 1. Since the coefficients
where L
of L belong to W∞1 (Ω ), in that case we have L(∂ j u) = ∂ j Lu + Lu, is again
an elliptic operator in divergence form with coefficients from L∞ (Ω ). Hence,
∂ j u ≤ C(∂ j Lu + + ∂ j um−1 )
−m
m
Lu
−m
≤ C(Lu1−m + um + um ).
Here we have used the fact that Lu−m ≤ C um by duality. Using the Poincaré in-
equality (Theorem 20.23), we obtain (42.5) for k = 1. The general case for k follows
by induction. Thus, the theorem is proved.
Proof. The result follows by interpolation of Sobolev spaces H0s (Ω ); see [39].
474 Part IV: Partial Differential Equations
Theorem 42.4 (Regularity in Sobolev spaces). Suppose that the coefficients of the
operator L from (42.1)–(42.2) belong to C∞ (Ω ). Let u and f be distributions on Ω
s (Ω ) for s ≥ 0, then u ∈ H s+2m (Ω ).
satisfying Lu = f . If f ∈ Hloc loc
L(ϕ u) = ϕ f + L1 u, (42.7)
or
u) = f+ L1 u,
L(
That is, u ∈ H01 (Ω ). Applying again Theorem 42.2 with this u ∈ H01 (Ω ), we obtain
that
u2 ≤ C(L u1 ) ≤ C( f +
u0 + L1 u0 + u1 + u1 ).
0
Thus, u ∈ H02 (Ω ), and the starting point of induction is checked. Let us assume
now that for every integer s ≥ 1 and 2m ≥ 2 it is true that f ∈ Hloc s (Ω ) implies
may apply the induction hypothesis, that is, the solution u of Lu = f belongs to
s+2m
Hloc (Ω ). But then we have that (see (42.5) with k = m + s + 1)
us+2m+1 ≤ C(L u2m+s ) ≤ C( f + L1 us+1 +
us+1 + u2m+s )
s+1
≤ C( f + us+2m + u2m+s ) < ∞,
s+1
Proof. The result follows from the proof of Theorems 42.2 and 42.4.
Theorem 42.6 (Gårding’s inequality). Suppose L from (42.1)–(42.2) has L∞ (Ω )
coefficients. Then for all u ∈ H0m (Ω ) we have
Since (42.9) and (42.10) imply that all derivatives of w of order strictly less than
m vanish on S, we reformulate the Dirichlet boundary value problem (42.10) as
follows: given f ∈ L2 (Ω ), find w ∈ H0m (Ω ) such that
u ≤ c0 FH→C ,
Proof. For each fixed u ∈ H the mapping v → a(u, v) is a bounded conjugate linear
functional on H. Hence, the Riesz–Fréchet theorem gives that there is a unique
element w ∈ H such that a(u, v) = (v, w) or a(u, v) = (w, v). Thus we can define an
operator A : H → H mapping u to w as
a(u, v) = (Au, v)
if and only if w = Au. It is clear that A is a bounded linear operator. The linearity
follows from
which implies Au ≤ M u. Let us note that a(u, Au) is real and positive here.
Next we show that A is one-to-one and that the range of A is equal to H. Indeed,
implies that β u ≤ Au ≤ M u. The first inequality implies that A is one-to-one
and R(A) = R(A). Now we will show that in fact, R(A) = H. Let w ∈ R(A)⊥ . Then
and therefore w = 0, i.e., R(A) = H = R(A). Next, again due to the Riesz–Fréchet
∈ H such that
theorem for F we have that there is a unique w
v),
F(v) = (w, v ∈ H,
and w if
= FH→C . But since R(A) = H, we may find u ∈ H such that Au = w
and only if
v) = F(v),
a(u, v) = (Au, v) = (w,
1 1 1
u ≤ = F .
Au = w
β β β
Finally, we need to show that this element u is unique. If there are two elements u1
and u2 such that
a(u1 , v) = F(v), a(u2 , v) = F(v),
β u1 − u2 2 ≤ |a(u1 − u2 , u1 − u2 )| = 0.
Remark 42.10. The Lax–Milgram theorem actually says that the operator A that
was constructed there has a bounded inverse such that
−1 1
A ≤ ,
H→H β
Assume now that the sesquilinear form a(u, v) satisfies all conditions of the Lax–
Milgram theorem and that the sesquilinear form b(u, v) is only bounded, i.e., there
is a constant M > 0 such that
So, we may associate with a(u, v) and b(u, v) two operators A and B, respectively,
such that A has bounded inverse and B is just bounded:
Au + Bu = w
Proof. Write
A + B = A(I − (−A−1 )B) =: A(I − K),
Corollary 42.12. Let A be as in Theorem 42.11 and let B be bounded (not neces-
sarily compact) with small norm, i.e., B < ε with ε sufficiently small. Then A + B
is bijective and (A + B)−1 : H → H is bounded.
where A : H0m (Ω ) → H0m (Ω ) is a linear bounded operator with bounded inverse and
B : H0m (Ω ) → H0m (Ω ) is compact.
Proof. The definition (42.3) of D(u, v) allows us to write
where Dm (u, v) is a sesquilinear form that satisfies (see (42.2)) the conditions of the
Lax–Milgram theorem (Theorem 42.9) and Dm−1 (u, v) has the form
Dm−1 (u, v) = ∑ aαβ (x)∂ β u∂ α vdx. (42.13)
|α |=|β |≤m−1 Ω
Applying the Lax–Milgram theorem, we obtain that there is a bounded linear oper-
ator A : H0m (Ω ) → H0m (Ω ) with bounded inverse such that
Concerning the sesquilinear form Dm−1 (u, v), we may say that (see (42.13)) since
|Dm−1 (u, v)| ≤ C ∑ |∂ β u| · |∂ α v|dx ≤ C uH m−1 (Ω ) vH m−1 (Ω )
|α |=|β |≤m−1 Ω
≤ C uH m (Ω ) vH m (Ω ) ,
or
BuH m (Ω ) ≤ C uH m−1 (Ω ) . (42.16)
Now let u j ∈ H0m (Ω ) be such that u j H m (Ω ) is bounded. Since H0m (Ω ) is com-
pactly embedded in H0m−1 (Ω ), we have that there is a subsequence u jk that con-
verges strongly in H0m−1 (Ω ), i.e., u jk is a Cauchy sequence in H0m−1 (Ω ). Hence
Bu jk is a Cauchy sequence in H0m (Ω ); see (42.16). This means that Bu jk converges
strongly in H0m (Ω ), and thus B is compact in H0m (Ω ). The theorem now follows
from (42.14) and (42.15).
Let us now return to the Dirichlet boundary value problem (42.10)–(42.11) for
the elliptic differential operator L from (42.1)–(42.2). Theorem 42.13 allows us to
rewrite (42.11) as
and the task is to find w ∈ H0m (Ω ) such that (42.17) holds for all v ∈ H0m (Ω ). Since
F : H0m (Ω )
v → (Lg − f , v)L2 (Ω )
(A + B)w = f0 . (42.18)
u = g + (A + B)−1 f0 , (42.19)
or
Lu = 0,
u ∈ H0m (Ω ) = 0,
or
Lu = 0, in Ω ,
u = 0, ∂ν u = 0, . . . , ∂ν u = 0, on ∂ Ω .
m−1
Formula (42.19) follows now from (42.18) and (42.9). The theorem is proved.
Gårding’s inequality (see (42.8)) allows us to get essential information about the
kernel of the operator L and its adjoint. Let us define
and
V = {u ∈ H0m (Ω ) : (L∗ u, v)L2 = 0 if and only if D(v, u) = 0 for all v ∈ H0m (Ω )}.
The latter inequality means that the operator L + c2 I is invertible and its inverse
(L + c2 I)−1 is compact as an operator in L2 (Ω ), since the embedding H0m (Ω ) →
L2 (Ω ) is compact. Moreover, the range R((L + c2 I)−1 ) is in H0m (Ω ). Now we can
see that u ∈ W if and only if
or
482 Part IV: Partial Differential Equations
(L + c2 I)u = c2 u,
or
1
I − (L + c2 I)−1 u = 0,
c2
that is, u belongs to the kernel of the operator c12 I − (L + c2 I)−1 . We can say
also that u ∈ V if and only if u belongs to the kernel of the adjoint operator
−1 ∗
c2 I −((L +c2 I) ) . Therefore, the Fredholm alternative (see Theorem 34.11) gives
1
where λ is real. Then Lemma 32.1 and Theorem 42.14 imply that (42.20) has a
solution if and only if λ = 0 is not a point of the spectrum of this Schrödinger
operator with homogeneous Dirichlet boundary conditions.
Example 42.19. Consider the Dirichlet boundary value problem in a bounded do-
main Ω ⊂ Rn , n ≥ 2, for the Helmholtz operator Δ + k2 n(x) with a complex-valued
function n(x) from L p (Ω ), n/2 < p ≤ ∞: given f ∈ H 1/2 (∂ Ω ), find u ∈ H 1 (Ω ) such
that
(Δ + k2 n(x))u = 0, in Ω ,
(42.21)
u = f, on ∂ Ω ,
where k is a real or complex number. The values k2 for which there exists a nonzero
function u ∈ H01 (Ω ) satisfying (in the distributional sense)
42 Elliptic Boundary Value Problems 483
Δ u + k2 n(x)u = 0, in Ω ,
are called the Dirichlet eigenvalues of the Helmholtz operator, and the correspond-
ing nonzero solutions are called the eigenfunctions for it. It is clear that k2 = 0 is
not an eigenvalue of this Helmholtz operator. The application of Lemma 32.1 and
Theorem 42.11 lead to the solvability of (42.21). Namely, (42.21) has a unique so-
lution if and only if k2 is not a point of the spectrum (i.e., is not an eigenvalue) of
this Helmholtz operator. In the case of a real-valued function n(x), we may prove
even more.
of the Helmholtz operator −Δ − λ n(x). The corresponding eigenvalues {λk }∞ k=1 are
all real and accumulate only at infinity (|λk | → ∞). If in addition n(x) ≥ 0 (n(x) ≡ 0),
then λk > 0 for all k = 1, 2, . . ..
Proof. We may rewrite the eigenvalue problem for (42.21) as (see Lemma 32.1 and
Theorem 42.11)
(A − λ B)u = 0, u ∈ H01 (Ω ), λ = 0, (42.22)
This is an eigenvalue problem for the self-adjoint compact operator A−1/2 BA−1/2 .
Using the Riesz–Schauder and Hilbert-Schmidt theorems (see Part III of this book),
we may conclude that there exists a sequence { λ1 }∞ of eigenvalues of A−1/2 BA−1/2
k k=1
such that
1 1
≥ ≥ ··· ≥ 1 ≥ ··· → 0
λ1 λ2 λ
k
In this chapter we will show that the scattering problem for an imperfect conductor
in Rn , n ≥ 2, is well posed. More precisely, we consider a bounded domain Ω ⊂ Rn
(the conductor) containing the origin with connected complement such that ∂ Ω is
in the class C2 . Our aim is to show the existence of a unique solution u ∈ C2 (Rn \
Ω ) ∩C(Rn \ Ω ) of the exterior impedance boundary value problem
Δ u + k2 u = 0, x ∈ Rn \ Ω , (43.1)
ik(x,θ )
u = u0 + usc , u0 = e , θ ∈S , n−1
(n−1)/2 ∂ usc
lim r − ikusc = 0, r = |x|,
r→+∞ ∂r
∂u
+ iλ u = 0, x ∈ ∂ Ω ,
∂ν
where G+
k is defined in Chapter 32.
Proof. Let x ∈ Rn \ Ω be fixed and let ε > 0 be so small that Bε (x) ⊂ Rn \ Ω . Let
BR (0) be a ball of radius R containing both Ω and Bε (x) (Figure 43.1).
x ε
Ω
0
R
∂νy |x − y| = νy · ∇y |x − y| = −1.
The well known properties of Hankel functions (see, e.g., [23]) give that
(1) (1)
ρ −(n−2)/2 H(n−2)/2 (ρ ) = −ρ −(n−2)/2 Hn/2 (ρ ).
ρ
(1)
Using now the asymptotic behavior of Hn/2 (k|x − y|) for small arguments (see, e.g.,
[23])
(1) −i k|x − y| −n/2
Hn/2 (k|x − y|) ∼ Γ (n/2),
π 2
|x − y|1−n
∂νy G+
k (|x − y|) ∼ , n ≥ 2.
ωn
In order to estimate the latter integral in (43.3) we need the following lemma.
and Lemma 43.2, we can easily obtain from (43.3), by letting R → ∞, the equality
(43.2). Thus, the theorem is completely proved.
Proof. Since G+
k (|x − y|) for x
= y is real-analytic in x ∈ Ω , the representation for-
mula (43.4) implies that the same is true for u(x).
Proof. Let BR (0) contain Ω in its interior. Then Green’s identity and the boundary
conditions imply that (∂ν = ∂∂r on the spheres)
∂ v(y) ∂ v(y)
v(y) − v(y) dσ (y)
|y|=R ∂r ∂r
= v(y)∂νy v(y) − v(y)∂νy v(y) dσ (y) = −2i λ (y)|v(y)|2 dσ (y).
∂Ω ∂Ω
∂ v 2
lim + k2 |v(y)|2 dσ (y) = −2k Im v(y)∂νy v(y)dσ (y). (43.6)
R→+∞ |y|=R
∂r ∂Ω
∂ v 2
lim + k |v(y)| dσ (y) = 0
2 2
R→+∞ |y|=R
∂r
and
Im v(y)∂νy v(y)dσ (y) = 0. (43.7)
∂Ω
Thus, we have from (43.5) and (43.7) (compare with Lemma 43.2)
lim |v(y)|2 dσ (y) = 0
R→+∞ |y|=R
and
λ (y)|v(y)|2 dσ (y) = 0.
∂Ω
This implies that v ≡ 0 and ∂ν v ≡ 0 on ∂ Ω . These two facts and the representa-
tion formula (43.2) for the scattering solution provide that v(x) ≡ 0 in Rn \ Ω . The
theorem is proved.
Corollary 43.6. If the solution of the scattering problem (43.1) exists, then it is
unique.
Proof. If two solutions u1 and u2 of (43.1) exist, then their difference v = u1 − u2 =
(1) (2)
usc − usc satisfies the hypothesis of Theorem 43.5. Hence v ≡ 0, i.e., u1 = u2 .
Theorem 43.7 (Rellich’s lemma). Let u ∈ C2 (Rn \ Ω ) be a solution of the
Helmholtz equation satisfying
490 Part IV: Partial Differential Equations
lim |u(y)|2 dσ (y) = 0.
R→+∞ |y|=R
Then u ≡ 0 in Rn \ Ω .
where
∂2 n−1 ∂ 1
Δ= + + ΔS .
∂ r2 r ∂ r r2
n−1
U (r) + U (r) + (k2 − μ 2 /r2 )U = 0.
r
But the general solution of this equation is given by
(1) (2)
U(r) = K1 r−(n−2)/2 Hν (kr) + K2 r−(n−2)/2 Hν (kr),
(1,2)
where K1 and K2 are arbitrary constants, Hν are Hankel functions of order ν with
(2) (1)
Hν = Hν and ν 2 = μ 2 + ( n−2
2 ) . Since
2 the hypothesis of this theorem implies
that U(r) = o(1/r (n−1)/2 ), we deduce that
K1 and K2 are equal to zero and hence
U(r) ≡ 0 for all r ≥ R0 . The same is true for u(rθ ) due to the completeness of
the eigenfunctions of ΔS on Sn−1 ; see [34]. The claim follows now from the real
analyticity of every solution of the Helmholtz equation.
Then v ≡ 0 in Rn \ Ω .
Remark 43.9. All the results of Theorems 43.1–43.8 and their corollaries remain
true if we consider instead of C2 (Rn \ Ω ) ∩ C(Rn \ Ω ) the Sobolev spaces H s .
Namely, we may assume that the problem (43.1) is considered in the space H 2 (Rn \
Ω ) ∩ H 3/2 (Rn \ Ω ).
Here we have considered mostly uniqueness results for these boundary value
problems. The solvability is provided using the results of Chapters 41 and 42 as
follows. As we know, the single layer potential
usc (x) := ϕ (y)G+
k (|x − y|)dσ (y), x ∈ Rn \ ∂ Ω , (43.8)
∂Ω
Let us note that these properties of the single layer potential are also valid for ϕ ∈
H −1/2 (∂ Ω ), where the integrals are interpreted in the sense of duality pairing [22,
25]. Thus, (43.8) will solve the scattering problem (43.1), provided that
ϕ (x) − 2 ϕ (y)∂νx G+
k (|x − y|)dσ (y) − 2iλ (x) ϕ (y)G+
k (|x − y|)dσ (y)
∂Ω ∂Ω
= 2(∂ν u0 (x) + iλ (x)u0 (x)), x ∈ ∂ Ω , (43.9)
where u0 (x) = eik(x,θ ) . Hence, to establish the existence for the problem (43.1), it
suffices to show the existence of a solution to (43.9) in the normed space C(∂ Ω ). To
this end, we first recall that the integral operators in (43.9) are compact on C(∂ Ω )
(see Theorem 34.9). Hence, by Riesz’s theorem (see Theorem 41.29), it suffices to
show that the homogeneous equation (corresponding to (43.9)) has only the trivial
solution.
Let ϕ be a solution of this homogeneous equation. Then usc from (43.8) will
be a solution of (43.9) with u0 set equal to zero, and hence, by Theorem 43.5 and
Corollary 43.6, we have that this usc (x) is equal to zero for x ∈ Rn \ Ω . By the
continuity of (43.8) across ∂ Ω , usc (x) is a solution of the Helmholtz equation in Ω
as well, and usc (x) = 0 on the boundary ∂ Ω . If we assume now that k2 > 0 is not
a Dirichlet eigenvalue for −Δ in Ω , then usc (x) ≡ 0 in Ω , and by the discontinuity
properties of the single layer potential we have that
i.e., the homogeneous equation under consideration has only the trivial
solution ϕ = 0. Hence by Riesz’s theorem (Theorem 41.29), the corresponding
492 Part IV: Partial Differential Equations
with an arbitrary ball BR (0) of radius R > 0 centered at the origin and containing
Ω . We recall that H −s (∂ Ω ), 0 ≤ s < ∞ is the dual space of H s (∂ Ω ). Then, for
f ∈ H −1/2 (∂ Ω ), a weak solution of
Δ u + k2 u = 0, x ∈ Rn \ Ω , (43.10)
u = u0 + usc , u0 = eik(x,θ ) , θ ∈ Sn−1 ,
∂ usc
lim r(n−1)/2 − ikusc = 0, r = |x|,
r→+∞ ∂r
∂ν u(x) + iλ u(x) = 0, x ∈ ∂Ω,
for all v ∈ H 1 (Rn \ Ω ) that are identically equal to zero outside some ball BR (0)
with radius R > 0 sufficiently large. In that case the analogue of Theorem 43.10 for
Sobolev spaces can be proved (see [6, Chapter 8] for details).
Chapter 44
Some Inverse Scattering Problems for the
Schrödinger Operator
The classical inverse scattering problem is to reconstruct the potential q(x) from the
knowledge of the far field data (scattering amplitude, see p. 232) A(k, θ , θ ), when
k, θ , and θ are restricted to some given set.
If q ∈ L1 (Rn ), then q(y)u(y, k, θ ) ∈ L1 (Rn ) uniformly with respect to θ ∈ Sn−1
due to
q(y)u(y, k, θ ) = q(y)(eik(θ ,y) + usc (y, k, θ )) = q(y)eik(θ ,y) + |q| 2 · q 1 usc (y, k, θ )
1
and Hölder’s inequality. We may therefore conclude that the scattering amplitude
A(k, θ , θ ) is well defined and continuous. Also, the following representation holds:
A(k, θ , θ ) = e−ik(θ ,y) q(y)(eik(θ ,y) + usc )dy
Rn
= e−ik(θ −θ ,y) q(y)dy + R(k, θ , θ )
Rn
= (2π )n/2 (F q)(k(θ − θ )) + R(k, θ , θ ),
or
q(x) ≈ (2π )−n/2 F −1 (A(k, θ , θ ))(x),
where the inverse Fourier transform must be understood in some special sense.
Let us introduce the cylinders M0 = R × Sn−1 and M = M0 × Sn−1 , and the mea-
sures μθ and μ on M0 and M, respectively, as
1
dμθ (k, θ ) = |k|n−1 dk|θ − θ |2 dθ ,
4
1
dμ (k, θ , θ ) = dθ dμθ (k, θ ),
|Sn−1 |
2π n/2
where |Sn−1 | = Γ ( 2n ) is the area of the unit sphere Sn−1 , and dθ and dθ denote the
usual Lebesgue measures on Sn−1 . We shall define the inverse Fourier transform on
M0 and M as
1
(FM−10 ϕ1 )(x) = e−ik(θ −θ ,x) ϕ1 (k, θ )dμθ ,
(2π )n/2 M0
1
(FM−1 ϕ2 )(x) = e−ik(θ −θ ,x) ϕ2 (k, θ , θ )dμ .
(2π )n/2 M
|ξ | ξ
k= , θ = θ − 2(θ , ξ)ξ, ξ = . (44.1)
2(θ , ξ) |ξ |
uθ (k, θ ) = k(θ − θ ),
if ϕ ∈ S is even and
ϕ ◦ uθ (k, θ )dμ (k, θ , θ ) = ϕ (x)dx
M Rn
if ϕ ∈ S;
(3) in addition,
FM−10 (ϕ ◦ uθ ) = F −1 ϕ
if ϕ ∈ S is even and
FM−1 (ϕ ◦ uθ ) = F −1 ϕ
44 Some Inverse Scattering Problems for the Schrödinger Operator 495
Definition 44.1. The inverse Born approximations qθB (x) and qB (x) of the potential
q(x) are defined by
1
qθB (x) = (2π )−n/2 (FM−10 A)(x) = e−ik(θ −θ ,x) A(k, θ , θ )dμθ
(2π )n M0
and
1
qB (x) = (2π )−n/2 (FM−1 A)(x) = e−ik(θ −θ ,x) A(k, θ , θ )dμ
(2π )n M
Proof. It is not difficult to check that if q(x) satisfies the conditions of the present
theorem, then q(x) will satisfy the conditions of Theorem 23.5:
n n+1
q ∈ L p (Rn ), < p≤ ,
2 2
or
n+1 n+1
q ∈ Lσp (Rn ), < p ≤ +∞, σ > 1− .
2 2p
1 |ξ |
qθB (x) = dξ e−i(ξ ,x−y) q(y)v y, ,θ dy.
(2π )n R n R n
2(θ , ξ)
|ξ |
q (ξ ) + (2π )−n/2
θ (ξ ) = q ei(ξ ,y) q(y) v y, ,θ − 1 dy,
2(θ , ξ)
B
Rn
i.e.,
v = 1−G k (qv),
k = e−ik(x−y,θ ) G+ . For k sufficiently large we obtain that
where G k
−1
v = (I + G k q) (1),
or
−ik(x−y,θ ) G , and the integral
where G q is an integral operator with kernel Gq = e q
operator Gq with this kernel also satisfies the equation (H − k2 )G q = I. In order to
prove (44.2), we recall that
q = G
G k − Gk qGq ,
and therefore,
G q = Gk − Gk qGq ,
44 Some Inverse Scattering Problems for the Schrödinger Operator 497
or
−1
G q = (I + Gk q) Gk .
because
−1
(I + G k q) Gk (q) = −(v − 1)
is equivalent to
= −v + 1 − 1 + v + G k (q) = Gk (q).
C
v − 1 2p = G q (q) 2p ≤ q 2p ,
p−1
L−σ /2 (Rn )
p−1
L−σ /2 kγ L
p+1
σ /2
where γ , p, and σ are as in that theorem. It remains only to check that the potential
2p
p
q ∈ Lloc (Rn ) with the special behavior at infinity belongs to Lσp+1
/2 (R ). But that is a
n
q1 (ξ ) − q
q1 (ξ ) − q2 (ξ )| = |
| θ +q
B
θ −q
B q1 (ξ ) − q
2 (ξ )| ≤ | θ | + |q
B
θ −q
B 2 (ξ )|
γ γ
|(ξ, θ )| |(ξ, θ )|
≤ C q1 2p
2
+C q2 2p2
=0
L
p+1
(Rn )
|ξ | L
p+1
(Rn )
|ξ |
σ /2 σ /2
Theorem 44.3 (Saito’s formula). Under the same assumptions for q(x) as in
Theorem 44.2,
(2π )n q(y)dy
lim kn−1 e−ik(θ −θ ,x) A(k, θ , θ )dθ dθ = ,
k→+∞ Sn−1 Sn−1 π Rn |x − y|n−1
498 Part IV: Partial Differential Equations
where the limit holds in the classical sense for n < p ≤ ∞ and in the sense of distri-
butions for n2 < p ≤ n.
Proof. Let us consider only the case n < p ≤ ∞. The proof for n
2 < p ≤ n requires
some changes.
By definition of the scattering amplitude,
I := kn−1 A(k, θ , θ )e−ik(θ −θ ,x) dθ dθ
Sn−1 Sn−1
= kn−1 q(y)dy eik(θ −θ ,y−x) dθ dθ
Rn Sn−1 Sn−1
+ kn−1 q(y)dy e−ik(θ ,y) R(y, k, θ )e−ik(θ −θ ,x) dθ dθ =: I1 + I2 ,
Rn Sn−1 Sn−1
We consider two cases: k|x − y| < 1 and k|x − y| > 1. In the first case, using Hölder’s
inequality the integral I1 over {y : k|x − y| < 1} can be estimated by
|q(y)|(k|x − y|)n−2
|I1 | ≤ Ck dy
|x−y|< 1k |x − y|n−2
1 1
p p
≤ Ck n−1
1
|q(y)| dy
p
1
1 · dy
|x−y|< k |x−y|< k
1 1
p p
−n n −1
= Ckn−1 k p 1
|q(y)| p
dy = Ck p
1
|q(y)| dy
p
→0
|x−y|< k |x−y|< k
44 Some Inverse Scattering Problems for the Schrödinger Operator 499
as k → +∞, since n < p ≤ ∞. This means that for every fixed x (or even uniformly
with respect to x) I1 approaches zero as k → ∞. Hence, we have only to estimate the
integral I1 over {y : k|x − y| > 1}. The asymptotic behavior of the Bessel function
Jν (·) for large argument implies that
q(y)
I1 = (2π ) k n
|x−y|> 1k |x − y|n−2
2
2 nπ π 1
× cos k|x − y| − + +O dy
π k|x − y| 4 4 (k|x − y|)3/2
q(y)
= (2π )n k
|x−y|> 1k |x − y|n−2
2 cos2 (k|x − y| − n4π + π4 ) 1
× +O dy
π k|x − y| (k|x − y|)2
(2π )n q(y)dy
=
π |x−y|> 1k |x − y|n−1
(2π )n q(y) nπ π
+ cos 2k|x − y| − + dy
π |x−y|> 1k |x − y|n−1 2 2
1 |q(y)|O(1) (1) (2) (3)
+ dy =: I1 + I1 + I1 .
k |x−y|> 1k |x − y|n
It is clear that
(1) (2π )n q(y)dy
lim I1 =
k→+∞ π Rn |x − y|n−1
and
(2)
lim I = 0.
k→+∞ 1
The latter fact follows from the following arguments. Since q belongs to L p (Rn ) for
p > n and has the special behavior at infinity, we may conclude that the L1 norm of
q(y)
the function |x−y| n−1 is uniformly bounded with respect to x. Hence it follows from
(2)
the Riemann–Lebesgue lemma that I1 approaches zero uniformly with respect to
(3)
x as k → +∞. For I1 we have the estimate
(3) C |q(y)|dy
|I1 | ≤ .
k1−δ Rn |x − y|n−δ
q(y)dy
If we choose δ such that 1 > δ > np , then Rn |x−y|n−δ will be uniformly bounded
(3)
with respect to x. Therefore, I1 → 0 as k → ∞ uniformly with respect to x. If we
collect all estimates, we obtain that
500 Part IV: Partial Differential Equations
(2π )n q(y)dy
lim I1 = .
k→∞ π Rn |x − y|n−1
where
R(y, k, θ ) = − G+
k (|y − z|)q(z)u(z, k, θ )dz = −Gk (qu),
Rn
1
Kq (x, y) = −|q(x)| 2 Gq (k, x, y)q 1 (y).
2
q L2 →L2 ≤ C
K ,
kγ
where γ is as in that theorem. We can therefore estimate I2 using Hölder’s inequality
as
J 2n−2 (k|x − y|)
C
|I2 | ≤ γ k |q(y)| 2 dy.
k Rn |x − y|n−2
C
|I2 | ≤ →0
kγ
as k → +∞.
Remark 44.4. This proof holds also for n = 2. In dimension n = 1 there is an analo-
gous result in which we replace the double integral on the left-hand side by the sum
of four values of the integrand at θ = ±1 and θ = ±1.
Theorem 44.5. Let us assume that n ≥ 2. Under the same assumptions for q1 (x)
and q2 (x) as in Theorem 44.3 let us assume that the corresponding scattering am-
plitudes Aq1 and Aq2 coincide for some sequence k j → ∞ and for all θ , θ ∈ Sn−1 .
Then q1 (x) = q2 (x) in the sense of L p for n < p ≤ ∞ and in the sense of distributions
for n2 < p ≤ n.
Proof. Saito’s formula shows that we have only to prove that the homogeneous
equation
q(y)dy
ψ (x) := =0
R n |x − y|n−1
has only the trivial solution q(y) ≡ 0. Let us assume that n < p ≤ ∞. Introduce the
space S0 (Rn ) of all functions from the Schwartz space that vanish in some neigh-
borhood of the origin. Due to the conditions for the potential q(x) we may conclude
(as before) that ψ ∈ L∞ (Rn ), and ψ defines a tempered distribution. Then for every
function ϕ ∈ S0 (Rn ) it follows that
Since ϕ (ξ ) ∈ S0 (Rn ), we have |ξ |−1 ϕ ∈ S0 (Rn ) also. Hence, for every h ∈ S0 (Rn )
the following equation holds:
q, h = 0.
This means that the support of q(ξ ) is at most at the origin, and therefore q(ξ ) can
be represented as
q(ξ ) = ∑ Cα Dα δ .
|α |≤m
Hence, q(x) is a polynomial. But due to the behavior at infinity we must conclude
that q ≡ 0. This proves Theorem 44.5.
Let us return now to the Born approximation of q(x). Repeated use of the Lippmann–
Schwinger equation leads to the following representation for the scattering ampli-
tude A(k, θ , θ ):
502 Part IV: Partial Differential Equations
m
A(k, θ , θ ) = ∑ j · (|q| 2 eik(x,θ ) )(y)dy
e−ik(θ ,y) q 1 (y)K
1
j=0 R
n 2
+ m+1 (|q| 21 (u(x, k, θ ))(y)dy,
e−ik(θ ,y) q 1 (y)K
Rn 2
is an
where u(x, k, θ ) is the solution of the Lippmann–Schwinger equation and K
integral operator with kernel
1
K(x, y) = |q(x)| 2 G+
k (|x − y|)q 1 (y). 2
The equality for A can be reformulated in the sense of integral operators in L2 (Sn−1 )
as
m
= ∑ Φ0∗ (k) sgn qK
A j Φ0 (k) + Φ0∗ (k) sgn qK
m+1 Φ (k),
j=0
where Φ0 and Φ (k) are defined by (23.9) and (23.10), and Φ0∗ is the L2 - adjoint of
Φ0 .
Using this equality and the definition of Born’s potential qB (x), we obtain
m
qB (x) = ∑ FM−1 j Φ0 (k) + F −1 Φ0∗ (k) sgn qK
Φ0∗ (k) sgn qK m+1 Φ (k) ,
M
j=0
where the inverse Fourier transform is applied to the kernels of the corresponding
integral operators. If we rewrite the latter formula as
m
qB (x) = ∑ q j (x) + q
m+1 (x),
j=0
q j (x) and q
j (x) in the Born series belong to the Hölder class Cα (Rn ) for all α ≤
1 − (3n − 3)/(2p).
For l = 1, 2 let us define el (θ ) = eik(θ ,xl ) ∈ L2 (Sn−1 ) and El (θ ) = el θ ∈ (L2 (Sn−1 ))2 .
Then the latter difference is equal to
C |k|n−1 dk (e1 , Φ0∗ sgn (q)K j Φ0 e1 )L2 (Sn−1 ) − (e2 , Φ0∗ sgn (q)K j Φ0 e2 )L2 (Sn−1 )
|k|≥k0
j Φ0 E1 )L2 (Sn−1 ) + (E2 , Φ0∗ sgn (q)K
− (E1 , Φ0∗ sgn (q)K j Φ0 E2 )L2 (Sn−1 ) .
(44.4)
Since E1 − E2 L2 (Sn−1 ) = e1 − e2 L2 (Sn−1 ) and el 2L2 (Sn−1 ) = |Sn−1 |, we obtain
from (44.3)–(44.4) the estimate
j Φ0
|q j (x1 ) − q j (x2 )| ≤ Cn |k|n−1 e1 − e2 L2 (Sn−1 ) Φ0∗ sgn (q)K dk.
|k|≥k0
where Jν is the Bessel function of order ν . By Lemma 23.13 and Theorem 23.5 we
get
∗ j Φ0 C
Φ0 sgn (q)K ≤ γ ( j+1)+n−2 ,
|k|
where γ = 1 − (n − 1)/(2p).
If we set r = |x1 − x2 |, we have to estimate the integral
∞ 1/2
kn−1 J(n−2)/2 (kr)
|Sn−1 | − (2π )n/2 dk.
k0 kγ ( j+1)+n−2 (kr)(n−2)/2
We split this integral into two parts: over 1/r < k < ∞ and over k0 < k < 1/r. By
the asymptotics of the Bessel functions for large argument, the first part can be
estimated from above by
504 Part IV: Partial Differential Equations
∞
kn−1
|S n−1
| dk ≤ Cn rγ ( j+1)−2 ,
1/r kγ ( j+1)+n−2
where j is chosen so large that γ ( j + 1) > 2. For the second part of the integral we
use the asymptotics of the Bessel functions for small argument [23], namely,
x(n−2)/2
J(n−2)/2 (x) = (1 + O(x2 )), x → 0.
2(n−2)/2Γ (n/2)
Since |Sn−1 | = 2π n/2 /Γ (n/2), we may estimate the second part from above by
1/r 1/r
kn dk dk
Cr = Cr ≤ Crmin(1,γ ( j+1)−2) .
k0 kγ ( j+1)+n−2 k0 kγ ( j+1)−2
To finish the proof we use the fact that q j ∈ L∞ (Rn ), which holds since
∞
kn−1
|q j (x)| ≤ C ( e L2 (Sn−1 ) + E L2 (Sn−1 ) )dk < ∞
k0 kγ ( j+1)+n−2
ξ , η ) = −Cn (ξ , η ) ,
G(
|ξ |2 |η |2
where we have used the precise value of the principal value integral; see [30, proof
of Lemma 2.4].
Lemma 44.8. Assume that the potential q satisfies all conditions of Theorem 44.6.
Then the first nonlinear term q1 admits the representation
2
x−y
q1 (x) = Cn q(y)dy
Rn |x − y|n
−1
where F2n denotes the 2n-dimensional inverse Fourier transform. The claim fol-
lows now from F −1 (ξ /|ξ |2 ) = Cn x/|x|n .
Lemma 44.9. Under the same assumptions on q as in Theorem 44.6 we have that
(1) for 3(n − 1)/2 < p < ∞, q1 belongs to (Wp,2
1
δ −1 (R )) with 1 − (n + 1)/(2p) <
n 2
2δ < n − n/p;
(2) for p = ∞, q1 belongs to the Hölder space C1 (Rn ).
Proof. We introduce the Riesz potential I −1 and Riesz transform R (see [37] and
Chapter 21) as
−1 −1 f(ξ ) −1 ξ f(ξ )
I f (x) = F (x), R f (x) = F (x).
|ξ | |ξ |
Note that ∇I −1 = R is bounded in L p (Rn ) for all 1 < p < ∞; see [37]. From [27] we
know that
I −1 : Lσp +1 (Rn ) → Lσp (Rn )
for −n/p < σ < n−1−n/p. This proves (1). Part (2) can be proved like [31, Lemma
2.2].
Proof. The proof of this theorem follows immediately from Lemmas 44.7, 44.8, and
44.9 and the Sobolev embedding theorem.
The statement of Theorem 44.10 means that all singularities and jumps of the
unknown potential can be recovered by the Born approximation. In particular, if
the potential is the characteristic function of an arbitrary bounded domain, then this
domain can be uniquely determined from the scattering data using a linear method.
Chapter 45
The Heat Operator
L = ∂t − Δx , (x,t) ∈ Rn × R.
The heat operator is a prototype of parabolic operators. These are operators of the
form
∂t + ∑ aα (x,t)∂xα ,
|α |≤2m
where |x|2
Kt (x) = (2π )−n/2 F −1 e−|ξ | t ≡ (4π t)−n/2 e− 4t , t > 0,
2
(45.2)
as t → 0+.
Proof. For fixed t > 0,
|x−y|2 |x − y|2 n
Δx Kt (x − y) = (4π t)−n/2 e− 4t − ,
4t 2 2t
= ( f (x − z) − f (x))Kt (z)dz
R
n
√
= ( f (x − η t) − f (x))K1 (η )dη .
Rn
for small t and for R large enough. So we can see that u(x,t) is continuous (even
uniformly continuous and bounded) for (x,t) ∈ Rn × [0, ∞) and u(x, 0) = f (x).
Proof. We can differentiate under the integral sign defining u as often as we please,
because the exponential function increases at infinity faster than any polynomial.
Thus, the heat equation takes arbitrary initial data (bounded and uniformly continu-
ous) and smooths them out.
then u ≡ 0.
Then
∂t ψ − Δ ψ = 0, t > 0,
∂t ϕ + Δ ϕ = 0, t < t0 .
we obtain
0= F · ν dσ = u(x, b)Kt0 −b (x − x0 )dx − u(x, a)Kt0 −a (x − x0 )dx
∂Ω |x|≤r |x|≤r
b n
xj
+
a
dt
|x|=r
∑ u(x,t)∂ j Kt0 −t (x − x0 ) − Kt0 −t (x − x0 )∂ j u(x,t)
r
dσ (x).
j=1
and
lim Kt0 −a (x − x0 )u(x, a)dx = 0.
a→0+ Rn
Since
Kt0 −b (x − x0 )dx = 1,
Rn
45 The Heat Operator 511
We divide the latter integral into two parts: |z| < δ and |z| > δ . The first part can be
estimated from above by
sup |u(x0 + z,t0 + τ ) − u(x0 ,t0 )| Kτ (z)dz
|z|<δ |z|<δ
as τ → 0 and δ → 0 due to the continuity of u(x,t) at the point (x0 ,t0 ). The second
part can be estimated from above by (see (45.3))
cε
Kτ (z)eε |x0 +z| dz ≤ e−|z| /(4τ )+ε |z| dz
2 2 2
cε
|z|>δ (4πτ ) n/2 |z|>δ
cε −|y|2 /4+ετ |y|2
= τ n/2
√ e dy → 0
(4πτ ) n/2 |y|>δ / τ
The first term in the latter sum can be estimated from above by
sup |u(x0 + z, a)| Kt0 −a (z)dz ≤ sup |u(x0 + z, a)| → 0
|z|≤R |z|≤R |z|≤R
as R → +∞ and a → 0.
512 Part IV: Partial Differential Equations
Theorem 45.5. The kernel Kt (x) is a fundamental solution for the heat operator.
or
∂t Kε − Δx Kε , ϕ → ϕ (0), ϕ ∈ C0∞ (Rn+1 ).
− dt Kt (x)Δx ϕ (x,t)dx
ε Rn
∞
= Kε (x)ϕ (x, ε )dx + dt ∂t Kt (x)ϕ (x,t)dx
ε
R∞ Rn
n
− dt Δx Kt (x)ϕ (x,t)dx
ε Rn
∞
= Kε (x)ϕ (x, ε )dx + dt (∂t − Δ )Kt (x)ϕ (x,t)dx
ε
R Rn
n
∂Ω
t =T
t =0
Ω
Fig. 45.1 Geometry of the boundary value problem for the Heat equation in Ω and 0 ≤ t ≤ T .
The first basic result concerning such problems is the maximum principle.
Theorem 45.7. Let Ω be a bounded domain in Rn and 0 < T < ∞. Suppose u
is a real-valued continuous function on Ω × [0, T ] that satisfies ∂t u − Δ u = 0 in
Ω × (0, T ). Then u assumes its maximum and minimum either on Ω × {0} or on
∂ Ω × [0, T ].
Proof. Given ε > 0, set v(x,t) := u(x,t) + ε |x|2 . Then ∂t v − Δ v = −2nε . Suppose
0 < T < T . If the maximum of v in Ω × [0, T ] occurs at an interior point of Ω ×
(0, T ), then the first derivatives ∇x,t v vanish there and the second derivative ∂ j2 v
for all j = 1, 2, . . . , n is nonpositive (consider v(x,t) a function of one variable x j ,
j = 1, 2, . . . , n). In particular, ∂t v = 0 and Δ v ≤ 0, which contradicts ∂t v − Δ v =
−2nε < 0 and Δ v = 2nε > 0.
Likewise, if the maximum occurs in Ω × {T }, then we have ∂t v(x, T ) ≥ 0 and
Δ v(x, T ) ≤ 0, which contradicts ∂t v − Δ v < 0. Therefore,
Replacing u by −u, we can obtain the same result for the minimum.
u(x,t) = F(x)G(t).
Then
∂t u − Δ u = FG − GΔx F = 0
if and only if
G ΔF
= := −λ 2 ,
G F
or
G + λ 2 G = 0, Δ F + λ 2 F = 0,
for some constant λ . The first equation has the general solution
G(t) = ce−λ t ,
2
∞ ∞
infinitely many solutions Fj (x) j=1 with corresponding λ j2 . The numbers
j=1
−λ j2 are called eigenvalues, and the Fj (x) are called eigenfunctions of the Lapla-
∞
cian. It is also known that λ j > 0, j = 1, 2, . . ., λ j2 → ∞, and Fj (x) j=1 can be
∞
chosen as a complete orthonormal set in L2 (Ω ) (or Fj (x) j=1 forms an orthonor-
mal basis of L2 (Ω )). This fact allows us to represent f (x) in terms of Fourier series:
∞
f (x) = ∑ f j Fj (x), (45.6)
j=1
j=1 j=1
that is, u(x,t) from (45.7) satisfies the heat equation and u(x,t) = 0 on ∂ Ω × (0, ∞).
It remains to prove that u(x,t) satisfies the initial condition and to determine for
which functions f (x) the series (45.6) converges and in what sense. This is the main
question in the Fourier method.
It is clear that the series (45.6) and (45.7) (for t ≥ 0) converge in the sense of
L2 (Ω ). It is also clear that if f ∈ C1 (Ω ) vanishes at the boundary, then u will vanish
on ∂ Ω × (0, ∞), and one easily verifies that u is a distributional solution of the heat
equation (t > 0). Hence it is a classical solution, since u(x,t) ∈ C∞ (Ω × (0, ∞)) (see
Corollary 45.3).
Similar considerations apply to the problem
⎧
⎪
⎨∂t u − Δ u = 0, in Ω × (0, ∞),
u(x, 0) = f (x), in Ω
⎪
⎩
∂ν u(x,t) = 0, on ∂ Ω × (0, ∞).
This problem boils down to finding an orthonormal basis of eigenfunctions for the
Laplacian with the Neumann boundary condition. Let us remark that for this prob-
lem, {0} is always an eigenvalue and 1 is an eigenfunction.
Exercise 45.3. Prove that u(x,t) of the form (45.7) is a distributional solution of
the heat equation in Ω × (0, ∞).
516 Part IV: Partial Differential Equations
π
Exercise 45.4. Show that 0 |u(x,t)|2 dx is a decreasing function of t > 0, where
u(x,t) is the solution of
ut − uxx = 0, 0 < x < π ,t > 0,
u(0,t) = u(π ,t) = 0, t > 0.
Chapter 46
The Wave Operator
The wave equation is satisfied exactly by the components of the classical electro-
magnetic field in vacuum.
The characteristic variety of (46.1) is
charx (L) = (ξ , τ ) ∈ Rn+1 : (ξ , τ ) = 0, τ 2 = |ξ |2 ,
and
{(ξ , τ ) ∈ charx (L) : τ < 0}
Proof. By considering real and imaginary parts we may assume that u is real. Define
Bt = {x : |x − x0 | < t0 − t}. Let us consider the following integral:
1
E(t) = (ut )2 + |∇x u|2 dx,
2 Bt
n
E (t) = ut utt + ∑ ∂ j u(∂ j u)t dx
Bt j=1
1
− (ut )2 + |∇x u|2 dσ (x) =: I1 + I2 .
2 ∂ Bt
n n
I1 =
Bt
∑ ∂ j [(∂ j u)ut ] − ∑ (∂ j2 u)ut + ut utt dx
j=1 j=1
n
=
Bt
ut (utt − Δx u)dx +
∂ Bt
∑ (∂ j u)ν j ut dσ (x)
j=1
1
≤ |ut | |∇x u|dσ (x) ≤ |ut |2 + |∇x u|2 dσ (x) ≡ −I2 .
∂ Bt 2 ∂ Bt
Hence
dE
≤ −I2 + I2 = 0.
dt
But E(t) ≥ 0 and E(0) = 0 due to the Cauchy data. Therefore, E(t) ≡ 0 if 0 ≤ t ≤ t0
and thus ∇x,t u = 0 in Ω . Since u(x, 0) = 0, it follows that u(x,t) = 0 also in Ω .
(x0 ,t0 )
x0 t0
Fig. 46.1 Geometric illustration of the backward light cone at (x0 ,t0 ).
46 The Wave Operator 519
Remark 46.2. This theorem shows that the value of u at (x0 ,t0 ) depends only on the
Cauchy data of u in the ball {(x, 0) : |x − x0 | ≤ t0 }, see Figure 46.1.
We therefore have
rn−1
∂r rn−1 ∂r Mϕ (x, r) = Δ ϕ (x + ry)dσ (y) ≡ rn−1 Δx Mϕ (x, r).
ωn |y|=1
520 Part IV: Partial Differential Equations
and
k
∂r ∂r
(r ϕ ) =
2k
(r2 ϕ ) = 2ϕ + rϕ .
r r
Assume that
k−1
1
∂ k
(r2k ϕ ).
r
∂r2 ∂r r2k−1 ϕ (r) =
r r
Then
k
k−1
1
1 ∂r 2k+1
∂r2 ∂r r 2k+1
ϕ (r) = ∂r2 ∂r r ϕ
r r r
k−1
2 1
= ∂r ∂r (2k + 1)r2k−1 ϕ + r2k ϕ
r
k−1
k−1
1 1
= (2k + 1)∂r2 ∂r r2k−1 ϕ + ∂r2 ∂r r2k ϕ
r r
k
k
∂r ∂r
= (2k + 1) r2k ϕ + (r2k (rϕ ) )
r r
46 The Wave Operator 521
k
∂r
= (2k + 1)r2k ϕ + r2k (rϕ )
r
k
∂r
= (2k + 1)r2k ϕ + r2k ϕ + r2k+1 ϕ
r
k
∂r
= (2k + 2)r2k ϕ + r2k+1 ϕ
r
k+1
∂r
= r2k+2 ϕ .
r
Corollary 46.5 gives that if u(x,t) is a solution of the wave equation (46.1) in
Rn × R, then Mu (x, r,t) satisfies (46.3), i.e.,
n−1
∂r2 + ∂r Mu = ∂t2 Mu ,
r
for n = 2k + 1, k = 1, 2, . . ..
k−1
∂r k−1 2k−1 2 n−1 ∂r
= r ∂r Mu + ∂r Mu = r2k−1 ∂t2 Mu
r r r
k−1
∂r
= ∂t2 r2k−1 Mu = ∂t2 u.
r
Moreover, the initial conditions are satisfied due to (46.4) and (46.5).
u(x, r,t)
u(x,t) = Mu (x, 0,t) = lim ,
r→0 (n − 2)!!r
or
u(x, r,t)
= Mu + O(r), r → 0.
(n − 2)!!r
Hence
u(x, r,t)
Mu (x, 0,t) = lim .
r→0 (n − 2)!!r
But by definition of Mu we have that Mu (x, 0,t) = u(x,t), where u(x,t) is the solution
of (46.2). The initial conditions in (46.2) are satisfied due to (46.5). Next, since
u(x, r,t) satisfies (46.7), we have
46 The Wave Operator 523
u(x, r,t) 1 f(x, r + t) + f(x, r − t) 1 r+t
lim = lim + g(x, s)ds
r→0 (n − 2)!!r 2(n − 2)!! r→0 r r r−t
1
= ∂r f|r=t + ∂r f|r=−t + g(x,t) − g(x, −t) ,
2(n − 2)!!
because f(x,t) and g(x,t) are odd functions of t. We therefore finally obtain
u(x, r,t) 1
lim = ∂r f|r=t + g(x,t) ,
r→0 (n − 2)!!r (n − 2)!!
The last equality follows from the fact that M f (x,t) is even in t, and so its derivative
vanishes at t = 0.
Remark 46.10. If n = 3, then (46.9) becomes
1
u(x,t) = ∂t t f (x + ty)dσ (y) + t g(x + ty)dσ (y)
4π |y|=1 |y|=1
1
≡ f (x + ty)dσ (y) + t ∇ f (x + ty) · yd σ (y)
4π |y|=1 |y|=1
+t g(x + ty)dσ (y) .
|y|=1
524 Part IV: Partial Differential Equations
The solution of (46.2) for even n is readily derived from the solution for odd n
by the “method of descent”. This is just a trivial observation: if u is a solution of the
wave equation in Rn+1 × R that does not depend on xn+1 , then u satisfies the wave
equation in Rn × R. Thus to solve (46.2) in Rn × R with even n, we think of f and g
as functions on Rn+1 that are independent of xn+1 .
n+4 n+2
Theorem 46.11. Suppose that n is even. If f ∈ C 2 (Rn ) and g ∈ C 2 (Rn ), then
the function
n−2
2 ∂t 2 f (x + ty)
u(x,t) = ∂t t n−1
dy
(n − 1)!!ωn+1 t |y|≤1 1 − y2
n−2
(46.10)
∂t 2 g(x + ty)
+ t n−1
dy
t |y|≤1 1 − y2
1
u(x,t) =
(n − 1)!!ωn+1
n−2
∂t 2
× ∂t t n−1
f (x + ty + tyn+1 )dσ (
y) (46.11)
t y21 +···+y2n +y2n+1 =1
n−2
∂t 2
+ t n−1
g(x + ty + tyn+1 )dσ ( y) ,
t y21 +···+y2n +y2n+1 =1
where y = (y, yn+1 ), solves (46.2) in Rn+1 × R (formally). But if we assume now
that f and g do not depend on xn+1 , then u(x,t) does not depend on xn+1 either and
solves (46.2) in Rn × R. It remains only to calculate the integrals in (46.11) under
this assumption. We have
f (x + ty + tyn+1 )dσ (
y) = f (x + ty)dσ (
y)
|y|2 +y2n+1 =1 |y|2 +y2n+1 =1
dy
=2 f (x + ty) ,
|y|≤1 1 − |y|2
because we have the upper and lower hemispheres of the sphere |y|2 + y2n+1 = 1.
Similarly for the second integral in (46.11). This proves the theorem.
Now we consider the Cauchy problem for the inhomogeneous wave equation
∂t2 u − Δx u = w(x,t),
(46.12)
u(x, 0) = f (x), ∂t u(x, 0) = g(x).
and
∂t2 u2 − Δ u2 = w,
(B)
u2 (x, 0) = ∂t u2 (x, 0) = 0.
For the problem (B) we will use a method known as Duhamel’s principle.
Theorem 46.13. Suppose w ∈ C[ 2 ]+1 (Rn × R). For s ∈ R let v(x,t; s) be the solu-
n
tion of
∂t2 v(x,t; s) − Δx v(x,t; s) = 0,
v(x, 0; s) = 0, ∂t v(x, 0; s) = w(x, s).
Then t
u(x,t) := v(x,t − s; s)ds
0
solves (B).
Let us consider again the homogeneous Cauchy problem (46.2). Applying the
Fourier transform with respect to x gives
∂t2 u(ξ ,t) + |ξ |2 u(ξ ,t) = 0,
u(ξ , 0) = f(ξ ), ∂t u(ξ , 0) = g(ξ ).
But this ordinary differential equation with initial conditions can be easily solved to
obtain
sin(|ξ |t) sin(|ξ |t) sin(|ξ |t)
u(ξ ,t) = f(ξ ) cos(|ξ |t) + g(ξ ) ≡ f (ξ )∂t + g(ξ ) .
|ξ | |ξ | |ξ |
ξ ,t) + |ξ |2 F(
∂t2 F( ξ ,t) = (2π )−n/2 δ (t).
This gives two equations for the four unknown coefficients a, b, c, d. But it is rea-
sonable to require F(x,t) ≡ 0 for t < 0. Hence, a = b = c = 0 and d = (2π )−n/2 |ξ1 | .
46 The Wave Operator 527
Therefore, ξ |t)
(2π )−n/2 sin(| , t > 0,
ξ ,t) =
F( |ξ | (46.14)
0, t < 0.
and
Φ (x,t), t > 0,
Φ+ (x,t) =
0, t < 0,
is the fundamental solution of the wave equation, i.e., F(x,t) with t > 0.
There is one more observation. If we compare (46.9) and (46.10) with (46.13),
then we may conclude that these three formulas are the same. Hence, we may cal-
culate the inverse Fourier transform of
sin(|ξ |t)
(2π )−n/2
|ξ |
in odd and even dimensions respectively with (46.9) and (46.10). In fact, the result
is presented in these two formulas.
In solving the wave equation in the region Ω × (0, ∞), where Ω is a bounded
domain in Rn , it is necessary to specify not only the Cauchy data on Ω ×{0} but also
some conditions on ∂ Ω × (0, ∞) to tell the wave what to do when it hits the bound-
ary. If the boundary conditions on ∂ Ω × (0, ∞) are independent of t, the method of
separation of variables can be used.
Let us (for example) consider the following problem:
⎧
⎪
⎨∂t u − Δx u = 0,
2 in Ω × (0, ∞),
u(x, 0) = f (x), ∂t u(x, 0) = g(x), in Ω , (46.15)
⎪
⎩
u(x,t) = 0, on ∂ Ω × (0, ∞).
We can look for a solution u in the form u(x,t) = F(x)G(t) and get
Δ F(x) + λ 2 F(x) = 0, in Ω ,
(46.16)
F(x) = 0, on ∂ Ω ,
and
G (t) + λ 2 G(t) = 0, 0 < t < ∞. (46.17)
528 Part IV: Partial Differential Equations
where f j = ( f , Fj )L2 and g j = (g, Fj )L2 . It follows from (46.15) and (46.18) that
∞ ∞
u(x, 0) = ∑ a j Fj (x), ut (x, 0) = ∑ λ j b j Fj (x). (46.20)
j=1 j=1
∞
The series (46.18), (46.19), and (46.20) converge in L2 (Ω ), because Fj j=1 is an
orthonormal basis in L2 (Ω ). It remains only to investigate the convergence of these
series in stronger norms (which depends on f and g, or more precisely, on their
smoothness).
The Neumann problem with ∂ν u(x,t) = 0, x ∈ ∂ Ω , can be considered in a similar
manner.
References
24. Maurin K, Methods of Hilbert Spaces, Polish Scientific Publishers, Warsaw, 1967
25. McLean W, Strongly Elliptic Systems and Boundary Integral Equations, Cambridge Univer-
sity Press, 2000
26. Nikolskii S M, Approximation of Functions of Several Variables and Imbedding Theorems,
Springer-Verlag, Berlin, Heidelberg, New York, 1975
27. Nirenberg L and Walker H, Nullspaces of elliptic partial differential operators in Rn , J. Math.
Anal. Appl., 42 (1973), 271–301.
28. Ouhabaz E M, Analysis of Heat Equations on Domains, Princeton University Press, 2005
29. Pinsky M A, Introduction to Fourier Analysis and Wavelets, Brooks/Cole, 2002
30. Päivärinta L and Serov V, Recovery of singularities of a multidimensional scattering potential,
SIAM J. Math. Anal., 29 (1998), 697–711
31. Päivärinta L and Somersalo E, Inversion of discontinuities for the Schrödinger equation in
three dimensions, SIAM J. Math. Anal., 22 (1991), 480–499
32. Saito Y, Some properties of the scattering amplitude and inverse scattering problem, Osaka,
J. Math.19 (1982), 527–547
33. Saranen J and Vainikko G, Periodic Integral & Pseudodifferential Equations with Numerical
Approximation, Springer, 2001
34. Shubin M A, Pseudodifferential Operators and Spectral Theory, Springer-Verlag, Berlin, 1987
35. Simon B, A Comprehensive Course in Analysis, American Mathematical Society, 2015
36. Stakgold I, Green’s Functions and Boundary Value Problems, second edition, John Wiley,
1997
37. Stein E M, Singular Integrals and Differentiability Properties of Functions, Princeton Univer-
sity Press, Princeton, NJ, 1970
38. Stein E M and Shakarchi R, Fourier Analysis: An Introduction, Princeton University Press,
2003
39. Taylor M E, Partial Differential Equations: Basic Theory, Springer-Verlag, New York, 1996
40. Taylor M E, Partial Differential Equations II: Qualitative Studies of Linear Equations,
Springer-Verlag, New York, 1996
41. Titchmarsh E C, Eigenfunction Expansions Associated with Second-order Differential Equa-
tions, vol. 1, Oxford University Press, Oxford, 1946
42. Triebel H, Theory of Function Spaces, Birkhäuser, Basel, 1983
43. Watson G N, A Treatise on the Theory of Bessel Functions, Cambridge University Press, Cam-
bridge, UK, 1996
44. Wong M W, Discrete Fourier Analysis, Springer Basel AG, 2011
45. Vretblad A, Fourier Analysis and Its Applications, Springer-Verlag, New York, 2003
46. Zigmund A, Trigonometric Series, Vol. II, Cambridge University Press, 1959
Index
U
S uniform boundedness principle, 248, 263
scalar product, 249 uniform convergence, 17
scattering amplitude, 228, 232, 240 unitary operator, 282
scattering problem, 224
scattering solutions, 235, 347
Schwartz space, 134, 400 V
self-adjoint operator, 273 vector space, 249
semibounded from below, 313 Volterra integral equations of the first and
separable, 257 second kinds, 371
separation of variables, 100, 514
sequence space, 44, 252
sesquilinear form, 476 W
Sine–Gordon equation, 406 wave equation, 113, 406, 517
single layer potential, 454 wave operator, 517
Sobolev embedding theorem, 185 weak convergence, 268
Sobolev inequality, 222 weak solution, 405
Sobolev space, 8, 46, 86, 177, 188, 254 well-posed problem, 418
Sommerfeld radiation condition, 225 Whittaker–Shannon–Boas theorem, 95
spectral family, 282
spectral function, 332
spectrum, 295 Y
spherical mean, 519 Young’s inequality for convolution, 397