Complex Analysis A Functional Analytic Approach
Complex Analysis A Functional Analytic Approach
Complex Analysis
De Gruyter Graduate
Also of Interest
Linear Algebra. A Course for Physicists and Engineers
Arak Mathai, Hans J. Haubold, 2017
ISBN 978-3-11-056235-4, e-ISBN (PDF) 978-3-11-056250-7,
e-ISBN (EPUB) 978-3-11-056259-0
Complex Analysis
|
A Functional Analytic Approach
Mathematics Subject Classification 2010
30-01, 30H20, 32-01, 32A36, 32W05, 35J10
Author
Prof. Dr Friedrich Haslinger
Fakultät für Mathematik
Universität Wien
Oskar-Morgenstern-Platz 1
A-1090 Wien
Austria
friedrich.haslinger@univie.ac.at;
http://www.mat.univie.ac.at/˜has/
ISBN 978-3-11-041723-4
e-ISBN (PDF) 978-3-11-041724-1
e-ISBN (EPUB) 978-3-11-042615-1
www.degruyter.com
Preface
This book provides a thorough introduction to complex analysis in one variable and
features special topics of several variables connected with the Cauchy–Riemann dif-
ferential equation from a functional analysis point of view. Chapters 1 and 2 are basic
and cover the classical material of the complex plane and of holomorphic functions,
their power series and the Cauchy integral formula with the consequences for the be-
havior of zeroes and the maximum principle. The homology and the homotopy ver-
sion of Cauchy’s theorem are related to important topological concepts and will be
crucial for the characterization of simply connected domains in Chapter 4. The calcu-
lus of residues of meromorphic functions is used to evaluate real integrals and Fourier
transforms without computing antiderivatives. There are only 2 sections which do not
belong to a standard treatment: holomorphic parameter integrals, where the knowl-
edge of basic properties of the Lebesgue integral is assumed, and the inhomogeneous
Cauchy formula, where Stokes’ integral theorem from real analysis is used. This is the
first place where we refer to real methods in complex analysis.
Chapter 3 is devoted to analytic continuation and the monodromy theorem. In
Chapter 4 we continue with real methods such as the 𝒞∞ -partition of unity and
the inhomogeneous Cauchy–Riemann equations. This is also true for the general
treatment of Runge’s approximation theorem using the Hahn–Banach theorem. Now
we are able to prove Mittag-Leffler’s theorem on the existence of global meromor-
phic functions with prescribed singularities and, by solutions of the inhomogeneous
Cauchy–Riemann equations, the cohomology version of Mittag-Leffler’s theorem. In-
finite products are introduced in order to obtain Weierstraß’ factorization theorem on
the existence of holomorphic functions with prescribed zeroes. This gives the tools to
show that each domain in the complex plane is a domain of holomorphy, i.e. has a
holomorphic function which cannot be analytically extended beyond any point of the
boundary of the domain. In addition, normal families of holomorphic functions are
studied, a concept which refers to compactness in spaces of holomorphic functions
and provides the main idea in the proof of the Riemann mapping theorem, which
states that each simply connected domain not equal to ℂ is biholomorphic equivalent
to the unit disc. Chapter 4 ends with a characterization of simply connected domains
by means of properties of holomorphic functions. In Chapter 5 we study harmonic
functions – a concept of real analysis, solve the Dirichlet problem constructing a con-
tinuous function on the closure of a domain which is harmonic in the interior and
coincides with a given continuous function on the boundary of the domain. Further-
more, we collect the basics of subharmonic functions.
We would like to emphasize that this presentation of the classical theory for
one complex variable was inspired by the first 30 pages of Lars Hörmander’s book
“An Introduction to Complex Analysis in Several Variables” [41], which, after more than
50 years, is still singular in its elegance and importance.
https://doi.org/10.1515/9783110417241-201
VI | Preface
Chapter 6 describes the main differences between the univariate and multivariate
theories with emphasis on the inhomogeneous Cauchy–Riemann differential equa-
tions. We explain the concept of pseudoconvexity without giving a full proof of the
characterization of domains of holomorphy by pseudoconvexity.
In the following chapters we treat basic functional analytic results on Hilbert
spaces and spectral theory of operators on Hilbert spaces for bounded and unbounded
operators in order to provide the tools for Bergman spaces of holomorphic functions
(Hilbert spaces with the reproducing property) and for a thorough discussion of the
canonical solution operator to 𝜕. In Chapter 8 the solution operator to 𝜕 restricted
to holomorphic L2 -functions in one complex variable is investigated, pointing out
that the Bergman kernel of the associated Hilbert space of holomorphic functions
plays an important role. We investigate operator properties like compactness and
Schatten-class membership (nuclear and Hilbert–Schmidt operators). At this place
we have the prerequisites to study the space of all holomorphic functions on a domain
endowed with the topology of uniform convergence on all compact subsets of the do-
main. Chapter 9 is devoted to the study of these spaces, which turn out to be complete
metric spaces with the Montel property (nuclear Fréchet spaces). In addition, we de-
scribe the dual spaces of these Fréchet spaces again as certain spaces of holomorphic
functions and represent them as sequence spaces (Köthe sequence spaces).
In Chapter 10 we consider the general 𝜕-complex and derive properties of the com-
plex Laplacian on L2 -spaces of bounded pseudoconvex domains. For this purpose we
first concentrate on basic results about distributions, Sobolev spaces, and unbounded
operators on Hilbert spaces. The key result is the Kohn–Morrey formula, which is pre-
sented in different versions. Using this formula the basic properties of the 𝜕-Neumann
operator – the bounded inverse of the complex Laplacian – are proved. It turns out to
be useful to investigate an even more general situation, namely the twisted 𝜕-complex,
where 𝜕 is composed with a positive twist factor. In this way one obtains a rather gen-
eral basic estimate, from which one gets Hörmander’s L2 -estimates for the solution of
the Cauchy–Riemann equation together with results on related weighted spaces of en-
tire functions. The last chapter contains an account of the application of the 𝜕-methods
to Schrödinger operators, Pauli and Dirac operators and to Witten–Laplacians. We
use the 𝜕-methods and some spectral theory to settle the question whether certain
Schrödinger operators with magnetic field on ℝ2 have compact resolvent.
Most of the material in this book is self-contained with the exception of some parts
in Chapters 6 and 11. Each single chapter contains exercises enhancing and reinforc-
ing the material discussed in the text. Notes at the end of each chapter refer to the
literature and more advanced results. The prerequisites for reading this book are: real
analysis, basic measure theory and point set topology.
I was partially supported by the FWF-grant P 28154 of the Austrian Science Fund.
Bibliography | 331
Index | 335
1 Complex numbers and functions
This chapter is devoted to the complex plane, its field structure and its topological
properties. The differential calculus is extended to the complex plane, leading to the
concept of a holomorphic function. Power series are introduced in order to obtain fur-
ther examples of holomorphic functions. After explaining the basic facts about line
integrals and primitive functions, we describe the most common elementary complex
functions, such as exponential, logarithmic and trigonometric functions.
In this way ℂ becomes a commutative field with zero element (0, 0) and unit element
(1, 0); for (a, b) ≠ (0, 0), we have
a −b
(a, b)−1 = ( , ).
a2 + b2 a2 + b2
For (a, b) ∈ ℂ, we also write a+ib, this means that (a, 0) corresponds to the real number
a and (0, 1) to the imaginary unit i.
The multiplication law from above stems from the formal multiplication
z = (x, −y) = x − iy
https://doi.org/10.1515/9783110417241-001
2 | 1 Complex numbers and functions
is called the complex conjugate of z, and we have the following rules for z, w ∈ ℂ:
(z + w)− = z + w, zw = z w, z = z.
We define |z|2 = zz; |z| is the absolute value of z. We have |z| = √x 2 + y2 , for z =
x + iy, and |z| = |z|. In addition,
1 1
ℜz = (z + z), ℑz = (z − z),
2 2i
|zw| = |z||w|, |z + w| ≤ |z| + |w|, ||z| − |w|| ≤ |z − w|.
Polar representation
Let z = x + iy ≠ 0. We set x = r cos θ and y = r sin θ, where r = |z| is the absolute value
of z and θ = arg z the argument of z. We have z = r(cos θ + i sin θ) = reiθ , which will
become clear after introducing the complex exponential function, see Section 1.10.
Since cos θ = cos(θ + 2kπ) and sin θ = sin(θ + 2kπ), for k ∈ ℤ, there are infinitely many
values of θ corresponding to a single z. The principal argument Arg z of z is to be taken
−π < Arg z ≤ π, in this way the polar representation of z becomes uniquely determined.
where we used the addition rules for the cosine and sine function. Hence one has to
multiply the absolute values and to add the angles. Concerning the principal argu-
ment, one has to be careful in this connection as the following example shows: let
z = −1 and w = i. Then Arg(−1) = π and Arg(i) = π/2, but Arg((−1)i) = Arg(−i) = −π/2 ≠
Arg(−1) + Arg(i).
de Moivre’s formula
Let n ∈ ℕ. Then
One-point compactification
By continuous continuation of ϕ to the whole of S2 , we obtain a topological homeo-
morphism between the compact space S2 and the so-called extended plane (one-point
compactification of ℂ) ℂ:
If we forget about the “complex structure”, we identify ℂn ≅ ℝ2n . We will use the no-
tations
We assume some knowledge about basic topological concepts in ℂn , such as that each
Cauchy3 sequence in ℂn is convergent (ℂn is complete).
The following topological concepts are formulated for sets in ℝn , they have anal-
ogous definitions in ℂ ≅ ℝ2 and ℂn ≅ ℝ2n .
Definition 1.2. Let G ⊆ ℝn . G is open, if for each x ∈ G there exists ϵ > 0 such that
Bϵ (x) ⊆ G. A set U is a neighborhood of the set M, if there exists an open set V such
that M ⊂ V ⊂ U. A ⊆ ℂ is called closed, if ℂ ⧵ A is open.
The union of arbitrarily many open sets is open; the intersection of finitely many
open sets is open; the union of finitely many closed sets is closed; the intersection of
arbitrarily many closed sets is closed.
Let M ⊆ ℝn be an arbitrary set in ℝn . The set
M ∘ ∶= ⋃{U ∶ U ⊆ M, U open}
is called the interior of M; M ∘ is an open set, it is the largest open set which is contained
in M. The set
M ∶= ⋂{A ∶ A ⊇ M, A closed}
is called the closure of M; M is a closed set, it is the smallest closed set which con-
tains M.
The set 𝜕M ∶= M ⧵ M ∘ is called the boundary of M.
Let M ⊆ ℝn be an arbitrary set in ℝn . A subset U ⊆ M is called relatively open in M,
if there is an open set O in ℝn such that U = O ∩ M.
A set X ⊆ ℝn is called (pathwise) connected, if any two points can be joined by a
continuous curve in G.
If X ⊆ ℝn and x ∈ X, we denote by Ex the largest connected set in X containing the
point x. The set Ex is called the connected component of x.
Definition 1.3. Let G ⊆ ℂ and f ∶ G ⟶ ℂ a function. We separate f (z) into its real
and imaginary part: f (z) = u(z) + iv(z) = ℜf (z) + iℑf (z), where u, v ∶ ℂ ⟶ ℝ are real-
valued functions. The set {(z, w) ∶ f (z) = w, z ∈ G} ⊆ ℂ2 is called the graph of f . The sets
{z ∶ ℜf (z) = const.}, {z ∶ ℑf (z) = const.}, {z ∶ |f (z)| = const.} are the level lines of f .
Examples. (a) f (z) = z 2 , ℜf (z) = x2 − y2 , ℑf (z) = 2xy. The level lines are circles and
hyperbolas.
(b) f (z) = az, a ∈ ℂ, a ≠ 0. This is a rotation–dilation. We write a = α + iβ, then we
have
az = αx − βy + i(βx + αy).
x αx − βy α −β x cos γ − sin γ x
( )↦( )=( ) ( ) = (α2 + β2 )1/2 ( )( ),
y βx + αy β α y sin γ cos γ y
where
α β
cos γ = , sin γ = .
(α2 + β2 )1/2 (α2 + β2 )1/2
f is continuous at z0 , if and only if for each sequence (zn )∞n=1 with limn→∞ zn = z0
we have limn→∞ f (zn ) = f (limn→∞ zn ) = f (z0 ).
If f and g are continuous, then f + g, f ⋅ g, gf (g ≠ 0) are continuous.
f is continuous, if and only if ℜf and ℑf are continuous.
6 | 1 Complex numbers and functions
|f (z)| ≥ δ ∀z ∈ K.
The following theorem has the same proof as in the real case.
Theorem 1.6. Let f and g be complex differentiable at z0 . Then f + g and f ⋅ g are com-
plex differentiable at z0 . In addition, λf is complex differentiable at z0 , where λ ∈ ℂ, and
the following rules are valid:
If g(z0 ) ≠ 0, then
f ′ f ′ ⋅ g − f ⋅ g′
( ) = .
g g2
We will need some results about holomorphic functions derived from the Cauchy
theorem to show an inverse function theorem for holomorphic functions.
f −1 ∶ V ⟶ U
(b) f (z) = z.
z − z0 h
= = 1;
z − z0 h
z − z 0 −ih
= = −1.
z − z0 ih
We have that Δ1 (z0 ) = 𝜕g (z ) = gx (z0 ), which is the partial derivative with respect
𝜕x 0
to x, and Δ2 (z0 ) = 𝜕g (z
𝜕y 0
) = g y 0 ) is the partial derivative with respect to y. In order to
(z
see this, we first choose z = x + iy0 , then (1.1) implies
g(z) − g(z0 )
Δ1 (z) = ,
x − x0
and putting x → x0 , we obtain the assertion about the partial derivative with respect
to x. Choosing z = x0 + iy, we get the assertion about the partial derivative with respect
to y.
In addition, we have that each real differentiable function at z0 is also continuous
at z0 .
{ xy2 , if z ≠ 0,
u(z) = { |z|
0, if z = 0.
{
Then u fails to be continuous at z = 0 since
1/n2 1
lim u(1/n, 1/n) = lim = ≠ u(0, 0) = 0.
n→∞ n→∞ 2/n2 2
But the partial derivatives exist and ux (0, 0) = uy (0, 0) = 0, since
u(h, 0) − u(0, 0)
ux (0, 0) = lim = 0.
h→0 h
Using the mean value theorem from real analysis, one can show the following
result: if ux and uy are continuous at z0 , then u is real differentiable at z0 .
ux (z0 ) = vy (z0 ),
vx (z0 ) = −uy (z0 ).
1.4 The Cauchy–Riemann equations | 9
This system of two partial differential equations for the functions u and v is called the
Cauchy–Riemann differential equations.
Hence we have shown the following
Theorem 1.10. Let f be complex differentiable at z0 and split the function into real and
imaginary parts, f = u + iv. Then
ux (z0 ) = vy (z0 ),
vx (z0 ) = −uy (z0 ).
In addition,
We have again
𝜕f 𝜕f
Δ1 (z0 ) = (z ) = fx (z0 ) and Δ2 (z0 ) = (z ) = fy (z0 ).
𝜕x 0 𝜕y 0
Lemma 1.12. The following assertions are equivalent:
(1) f ∶ U ⟶ ℂ is real differentiable at z0 .
(2) Real and imaginary part of f = u + iv are real differentiable at z0 .
(3) There exist functions A1 , A2 ∶ U ⟶ ℂ, both continuous at z0 , such that
where
1 1
A1 (z0 ) = (fx (z0 ) − ify (z0 )) and A2 (z0 ) = (fx (z0 ) + ify (z0 )).
2 2
This shows that (1) and (2) are equivalent and that
where
1 1
A1 (z) = (Δ1 (z) − iΔ2 (z)) and A2 (z) = (Δ1 (z) + iΔ2 (z)).
2 2
The expressions for the functions A1 and A2 lead to the following
ux (z0 ) = vy (z0 ),
vx (z0 ) = −uy (z0 )
{ A2 (z)(z−z
−
z−z0
0)
for z ≠ z0 ,
̃
Δ(z) = {
0 for z = z0 .
{
−
Since A2 is continuous at z0 , A2 (z0 ) = 0 and | z−z0 | = 1, we have limz→z0 Δ(z)
̃ = 0.
(z−z )
0
Hence Δ̃ is continuous at z0 .
Now let Δ = A1 + Δ.̃ Then Δ is continuous at z0 and
Examples. (1) f = u+iv, u(x, y) = x2 −y2 , v(x, y) = 2xy. f is real differentiable on ℂ and
since ux = 2x, uy = −2y, vx = 2y, vy = 2x, the Cauchy–Riemann differential equa-
tions are satisfied. Hence f is complex differentiable on ℂ and holomorphic on ℂ.
We have
f (z) = z 2 = x2 − y2 + 2ixy.
(2) f = u + iv, u(x, y) = x3 − 3xy2 , v(x, y) = 3x2 y − y3 . Also in this case, the Cauchy–
Riemann differential equations are satisfied and f is holomorphic on ℂ; we have
f (z) = z 3 .
12 | 1 Complex numbers and functions
ux uy
Jf = | | = ux vy − uy vx .
vx vy
Jf = u2x + vx2 .
By Theorem 1.10, we have f ′ = ux + ivx , which implies that |f ′ |2 = u2x + vx2 and
Jf = |f ′ |2 .
(Tf (z0 )) ∶ ℂ ⟶ ℂ is called the tangential map of f at the point z0 . It is a ℂ-linear map.
𝜕f 𝜕f
(Tf (z0 ))(h) = (z0 )h + (z0 )h.
𝜕z 𝜕z
This mapping (Tf (z0 )) ∶ ℂ ⟶ ℂ is in general only ℝ-linear.
It is easy to show that f is complex differentiable at z0 , if and only if
(Tf (z0 )) ∶ ℂ ⟶ ℂ
is ℂ-linear.
⟨z, w⟩
cos φ = ,
|z||w|
hence the upper assumption means that the angle between 𝒯z and 𝒯w coincides with
the original angle.
Proof. By Definition 1.18, we have (Tf (z))(h) = f ′ (z)h. We have to show that
γ is differentiable at s ∈ (a, b) if the derivatives x ′ (s) and y′ (s) exist. We set γ ′ (s) =
x′ (s) + iy′ (s).
Let z = γ(s) and suppose that γ ′ (s) ≠ 0. The map
t ↦ z + γ ′ (s) t, t ∈ ℝ,
f ∘ γ ∶ [a, b] ⟶ ℂ.
We have
(f ∘ γ)′ (s) = ux (z)x′ (s) + uy (z)y′ (s) + i[vx (z)x′ (s) + vy (z)y′ (s)]
= 1/2[ux (z) + ivx (z) − i(uy (z) + ivy (z))](x′ (s) + iy′ (s))
+ 1/2[ux (z) + ivx (z) + i(uy (z) + ivy (z))](x′ (s) − iy′ (s))
𝜕f 𝜕f
= (z)(x ′ (s) + iy′ (s)) + (z)(x′ (s) − iy′ (s)) (by Definition 1.13)
𝜕z 𝜕z
= (Tf (z))(γ ′ (s)) (by Definition 1.18).
∢(γ1′ (s), γ2′ (s)) = ∢((Tf (z))(γ1′ (s)), (Tf (z))(γ2′ (s))).
N
A series ∑∞
n=1 fn converges uniformly on A, if the sequence of its partial sums (∑n=1 fn )N
converges uniformly on A.
We write |f |A ∶= supz∈A |f (z)|.
ϵ ϵ
|fn − fm |A ≤ |fn − f |A + |f − fm |A < + = ϵ,
2 2
For a fixed z ∈ A, the sequence (fn (z))n is a Cauchy sequence in ℂ. Since ℂ is complete,
the limit of this sequence exists:
|fn (z) − f (z)| ≤ |fn (z) − fm (z)| + |fm (z) − f (z)| < ϵ + ϵ
Theorem 1.24. ∑∞ n=1 fn converges uniformly on A, if and only if for each ϵ > 0 there exists
nϵ ∈ ℕ such that
Proof. Since
n m
fm+1 (z) + ⋯ + fn (z) = ∑ fk (z) − ∑ fk (z),
k=1 k=1
Suppose that ∑∞
n=1 Mn < ∞. Then the series ∑n=1 fn converges uniformly on A.
∞
|fn (z)| = |z n | ≤ r n → 0 ∀z ∈ K.
n 1
(2) ∑∞
n=0 z = 1−z with uniform convergence on all compact subsets of 𝔻. Let K be as
above in example (1), then
∞ ∞ ∞
| ∑ z n | ≤ ∑ |z n | ≤ ∑ r n < ∞,
n=0 n=0 n=0
zn zn Nn
∞ ∞ ∞
|∑ | ≤ ∑| | ≤ ∑ = eN ,
n=0 n! n=0 n! n=0 n!
Theorem 1.27. If s, M > 0 are two constants with the property that
|an |sn ≤ M ∀n ∈ ℕ0 ,
n
then the power series ∑∞ n=0 an (z − z0 ) converges uniformly and absolutely on each com-
pact subset of Ds (z0 ) = {z ∶ |z − z0 | < s}.
Proof. If K is a compact subset of Ds (z0 ) then there exists a positive r < s with K ⊂
Dr (z0 ). Set q = r/s < 1. Then we have
rn
sup|an (z − z0 )n | ≤ sup |an (z − z0 )n | ≤ |an |r n = |an |sn ≤ Mqn ,
z∈K z∈Dr (z0 ) sn
n
since ∑∞
n=0 Mq < ∞, the assertion follows from Theorem 1.25.
Remark. If |an |sn ≤ M ∀n ∈ ℕ0 , then the sequence (|an |r n )n converges to 0 for each
positive r < s.
n
Definition 1.28. Let ∑∞
n=0 an (z − z0 ) be a power series and
n
Theorem 1.29. Let ∑∞ n=0 an (z − z0 ) be a power series with radius of convergence R.
Then:
n
(1) ∑∞n=0 an (z − z0 ) is uniformly convergent on each compact subset of DR (z0 ).
(2) ∑n=0 an (z − z0 )n fails to be convergent in ℂ ⧵ DR (z0 ).
∞
Proof. (1) If R = 0, then the assertion is trivial. If R > 0, then we have for an arbitrary
positive s < R that the sequence (|an |sn )n is bounded. By Theorem 1.27, the power series
n
∑∞n=0 an (z − z0 ) converges uniformly on each compact subset of Ds (z0 ), and as s was
an arbitrary positive number strictly smaller than R, the first assertion of the theorem
follows.
(2) If |z − z0 | > R, then the sequence (|an | |z − z0 |n )n is unbounded and the power
n
series ∑∞ n=0 an (z − z0 ) fails to be convergent.
In the following theorem we show that the radius of convergence can be computed
in terms of the coefficients of the power series.
R ≤ s ∀s > L.
Let L < s < ∞. Then s−1 < L−1 = lim supn→∞ |an |1/n , hence there exists an infinite subset
M ⊆ ℕ such that
This implies |am |sm > 1 ∀m ∈ M and (|an |sn )n fails to be a null-sequence. Therefore we
must have s ≥ R, because s < R would imply that (|an |sn )n is a null-sequence (see the
remark from above).
If L = ∞, then we get L = R from the first step of the proof.
n n 1/n
Examples. (a) ∑∞
n=0 n z , |an | = n and R = 0.
n 1/n
(b) ∑∞
n=0 z n , |an | = 1 and R = 1.
∞ z
(c) ∑n=0 nn , |an |1/n = 1/n and R = ∞.
zn
(d) ∑∞
n=0 n! , for this example we use the following
n
Theorem 1.31. Let ∑∞ n=0 an (z − z0 ) be a power series with radius of convergence R, and
suppose that an ≠ 0 for almost all n. Then
an a
lim inf| | ≤ R ≤ lim sup| n |.
n→∞ an+1 n→∞ an+1
a
If the limit limn→∞ | a n | exists, then
n+1
an
R = lim | |.
n→∞ an+1
zn a
Example. ∑∞
n=0 , n
n! an+1
= (n+1)!
n!
= n + 1 and R = ∞.
a
Proof. Let S = lim infn→∞ | a n |. If S = 0, we have S ≤ R. If S > 0, it suffices to show that
n+1
s ≤ R for each 0 < s < S. Since S is a lim inf, there exists an l ∈ ℕ such that
an
| | > s ∀n ≥ l.
an+1
Hence |an |s−1 > |an+1 | ∀n ≥ l. Let A = |al |sl . Then the last inequality implies that
|al+m |sl+m ≤ A ∀m ∈ ℕ.
Hence the sequence (|an |sn )n is bounded. Now we get from the definition of the radius
of convergence that s ≤ R.
a
Now let T = lim supn→∞ | a n |. If T = ∞, we have R ≤ T. If T < ∞, it remains to
n+1
show that t ≥ R for each t > T. We get again an l ∈ ℕ with
an
| |<t ∀n ≥ l.
an+1
This implies that |an+1 | > |an |t −1 ∀n ≥ l. One can choose l such that B = |al |t l > 0. Then
and by iteration |am+l |t l+m > B ∀m ∈ ℕ. Hence the sequence (|an |t n )n fails to be a null-
sequence. Therefore we have t ≥ R.
1.8 Line integrals | 21
Remark 1.33. (a) Let φ ∶ [a1 , b1 ] ⟶ [a, b] be a bijective 𝒞1 function such that φ′ > 0
everywhere on [a1 , b1 ], let γ ∶ [a, b] ⟶ U and γ1 ∶= γ ∘ φ ∶ [a1 , b1 ] ⟶ U be paths
in U. Since we supposed that φ′ > 0, we have φ(a1 ) = a and φ(b1 ) = b, which
means that the orientation in γ and γ1 coincides.
For each continuous function f ∶ U ⟶ ℂ we have
b1 b1
∫ f (z) dz = ∫ f (γ1 (t))γ1′ (t) dt = ∫ f (γ(φ(t)))γ ′ (φ(t))φ′ (t) dt
γ1 a1 a1
b
= ∫ f (γ(s))γ ′ (s) ds = ∫ f (z) dz,
a γ
(c) Let γ ∶ [0, 1] ⟶ U be a path. We define the inverse path γ1 to γ by γ1 (t) ∶= γ(1−t), t ∈
[0, 1]. Then γ ∗ = γ1∗ and
1 0
∫ f (z) dz = ∫ f (γ(1 − t))(−γ ′ (1 − t)) dt = ∫ f (γ(s))(−γ ′ (s))(−1) ds
γ1 0 1
1
= − ∫ f (γ(s))γ ′ (s) ds = − ∫ f (z) dz.
0 γ
We write γ1 = γ −1 .
(d) Let γ ∶ [a, b] ⟶ U be a path and f ∶ U ⟶ ℂ a continuous function. The length
L(γ) of the path γ is given by
b
L(γ) = ∫ |γ ′ (t)| dt.
a
We have
b b
|∫ f (z) dz| = |∫ f (γ(t))γ ′ (t) dt| ≤ ∫ |f (γ(t))| |γ ′ (t)| dt ≤ L(γ) max |f (z)|.
γ a a
∗ z∈γ
Example 1.34. (1) Let a ∈ ℂ, r > 0 and γ(t) = a + r(cos t + i sin t), t ∈ [0, 2π] the
positively oriented circle with center a and radius r (once passed through). Let
f ∶ U ⟶ ℂ be a continuous function and suppose that Dr (a) ⊂ U. Then
2π
∫ f (z) dz = r ∫ f (a + r(cos t + i sin t))(− sin t + i cos t) dt.
γ 0
2π
∫ z dz = i ∫ (cos 2t + i sin 2t) dt = 0;
γ 0
– f (z) = z. Here
2π 2π
∫ z dz = i ∫ (cos2 t + sin2 t) dt = i ∫ dt = 2πi;
γ 0 0
1 2π − sin t + i cos t
∫ dz = ∫ dt = 2πi.
γ z 0 cos t + i sin t
1.8 Line integrals | 23
dz 2π − sin t + i cos t 2π 1
∫ =r∫ dt = ir ∫ dt = 2πi.
γ z−a 0 a + r(cos t + i sin t) − a 0 r
(2) Let a, b ∈ ℂ, a ≠ b. The path γ(t) = a + (b − a)t, t ∈ [0, 1] describes the straight line
segment [a, b] joining the points a and b, f ∶ U ⟶ ℂ, where γ ∗ ⊂ U. Then
1
∫ f (z) dz = (b − a) ∫ f (a + (b − a)t) dt.
γ 0
1
∫ z dz = 2 ∫ (−1 + 2t) dt = 2(−t + t 2 )|10 = 0.
γ 0
2π
∫ z dz = i ∫ (cos 2t + i sin 2t) dt = 0.
γ1 π
Taking the function f (z) = z, we see that the line integral does not depend on the
path between −1 and 1. But for f (z) = z we have
1
∫ z dz = 2 ∫ (−1 + 2t) dt = 0,
γ 0
and
2π
∫ z dz = i ∫ (cos2 t + sin2 t) dt = iπ.
γ1 π
(3) Let a, b, c ∈ ℂ and Δ = Δ(a, b, c) be the triangle with vertices a, b, c, with the orien-
tation a → b → c → a. Then
= − ∫ f (z) dz.
𝜕Δ
24 | 1 Complex numbers and functions
Theorem 1.36. Suppose that the continuous function f ∶ U ⟶ ℂ has a primitive func-
tion F on U. Let z0 , z1 ∈ U and γ an arbitrary path in U from z0 to z1 . Then
Proof. Let γ ∶ [a, b] ⟶ U, a = t0 < t1 < ⋯ < tn = b and γ a 𝒞1 map on [tk−1 , tk ] for k =
1, … , n. Then
n tk n tk
∫ f (z) dz = ∑ ∫ f (γ(t))γ ′ (t) dt = ∑ ∫ F ′ (γ(t))γ ′ (t) dt
γ k=1 tk−1 k=1 tk−1
n tk n
= ∑ ∫ (F ∘ γ)′ (t) dt = ∑ [F(γ(tk )) − F(γ(tk−1 ))]
k=1 tk−1 k=1
= F(z1 ) − F(z0 ).
∫ f (z) dz = 0,
γ
Remark. If F ∈ ℋ(U) and F ′ is continuous (later we will see that we do not need the
last assumption), then
∫ F ′ (z) dz = 0,
γ
z n+1
Example 1.38. (a) The function f (z) = z n , n ∈ ℤ, n ≠ −1 has F(z) = n+1
as a primitive
on ℂ for n ≥ 0 and on ℂ∗ for n < −1. We have
1.9 Primitive functions | 25
1
∫ z n dz = (z n+1 − z0n+1 ), ∫ z n dz = 0,
[z0 ,z1 ] n+1 1 γ
∫ f (z) dz = 0.
γ
F(z) = ∫ f (ζ ) dζ .
γz
0 = ∫ f (ζ ) dζ = ∫ f (ζ ) dζ + ∫ f (ζ ) dζ − ∫ f (ζ ) dζ .
γ γz0 [z0 ,z] γz
This implies
F(z) − F(z0 ) = ∫ f (ζ ) dζ − ∫ f (ζ ) dζ = ∫ f (ζ ) dζ
γz γz0 [z0 ,z]
1
= ∫ f (z0 + t(z − z0 ))(z − z0 ) dt = (z − z0 )A(z),
0
1
where A(z) = ∫0 f (z0 + t(z − z0 )) dt, A(z0 ) = f (z0 ). Hence
F(z) − F(z0 )
A(z) = .
z − z0
26 | 1 Complex numbers and functions
In order to show F ′ (z0 ) = f (z0 ), it suffices to prove that A is continuous at z0 . For this
we estimate
∫ f (ζ ) dζ − ∫ f (ζ ) dζ = ∫ f (ζ ) dζ = 0,
γz γ̃ z γz γ̃ −1
z
∫ f (ζ ) dζ = ∫ f (ζ ) dζ .
γz γ̃ z
∫ f (z) dz = 0.
𝜕Δ
F(z) = ∫ f (ζ ) dζ ,
[a,z]
for z ∈ G. By assumption, the straight line segment [a, z] ⊂ G. If z0 ∈ G, then the triangle
Δ with vertices a, z, z0 is contained in G and hence
0 = ∫ f (ζ ) dζ = ∫ f (ζ ) dζ + ∫ f (ζ ) dζ − ∫ f (ζ ) dζ .
𝜕Δ [a,z0 ] [z0 ,z] [a,z]
Remark. If G is not convex, the assertions of Theorem 1.41 are true at least in each
convex neighborhood Uz ⊆ G of an arbitrary point z ∈ G.
1.9 Primitive functions | 27
In the following we study the interchange of limit processes where uniform limits
of sequences of functions and line integrals are involved.
Theorem 1.42. Let γ be a path in ℂ and let (fn )n be a sequence of continuous functions
on γ ∗ . Suppose that the sequence (fn )n converges uniformly on γ ∗ to a function f . Then
|∫ fn (z) dz − ∫ f (z) dz| = |∫ (fn (z) − f (z)) dz| ≤ L(γ) max |fn (z) − f (z)|,
γ γ γ
∗ z∈γ
then
∞ ∞
∫ ( ∑ fn (z)) dz = ∑ ∫ fn (z) dz.
γ n=0 n=0 γ
n
Theorem 1.43. Let P(z) = ∑∞ n=0 an (z − z0 ) be a power series with radius of convergence
R > 0. Then P is holomorphic on DR (z0 ) and
∞
P ′ (z) = ∑ nan (z − z0 )n−1
n=1
Proof. First we show the following assertion. Let R′ be the radius of convergence of
n−1
the power series Q(z) = ∑∞ n=1 nan (z − z0 ) . Then R′ ≥ R.
n
Without loss of generality, we can assume that z0 = 0. The power series ∑∞ n=0 nz
n
has radius of convergence 1, i.e. ∀ρ ∈ [0, 1) the sequence (nρ )n is bounded. Now we
have |an z1n | ≤ M, ∀n ∈ ℕ (M > 0), for an arbitrary z1 ∈ DR (0), in addition we have
n z2 n z n M
|nan z2n−1 | = | | |an z1n | ≤ n| 2 | ,
|z2 | z1 z1 |z2 |
z
for 0 < |z2 | < |z1 |. Since | z2 | < 1, the sequence (|nan z2n−1 |)n is bounded. As |z2 | < |z1 | < R
1
were chosen arbitrarily, we get from the definition of the radius of convergence (see
Definition 1.28) that R′ ≥ R.
Now let γ be an arbitrary closed path in DR′ (0). We will show that Q has P as a
primitive on DR′ (0):
∞ ∞
∫ Q(z) dz = ∫ ( ∑ nan z n−1 ) dz = ∑ nan ∫ z n−1 dz = 0,
γ γ n=1 n=1 γ
28 | 1 Complex numbers and functions
here we interchanged integration and summation (see Theorem 1.42) and used Exam-
ple 1.38 (a). By Theorem 1.39, this implies that there exists a primitive of Q on DR′ (0).
A primitive of Q is
∞ ∞
∫ Q(ζ ) dζ = ∫ ( ∑ nan ζ n−1 ) dζ = ∑ nan ∫ ζ n−1 dζ
[0,z] [0,z] n=1 n=1 [0,z]
n
z
∞ ∞
= ∑ nan = ∑ a zn ,
n=1 n n=1 n
where we used again Theorem 1.42 and Example 1.38 (a). Hence also
∞
P(z) = a0 + ∑ an z n
n=1
is a primitive of Q on DR′ (0). We have also seen that P converges on DR′ (0), hence
R′ = R.
Definition 1.44. We define the complex exponential function by the power series
zn
exp z = ez = ∑∞
n=0 n! .
i.e. (ez )′ = ez .
Remark. We know from the last proof that e−z ez = e0 = 1 for all z ∈ ℂ. Hence the ex-
ponential function has no zeroes.
Next we define the complex sine and cosine function again by power series.
Definition 1.46.
since
ℜez = ex cos y, ℑez = ex sin y, |ez | = |ex (cos y + i sin y)| = ex = eℜz ,
ez = ex (cos y + i sin y)
30 | 1 Complex numbers and functions
arg ez = ℑz.
Next we will determine the zeroes of the sine and cosine functions.
First we claim that there are positive real numbers x such that cos x = 0. Suppose
that this is not true. Then, since cos 0 = 1, we have cos x > 0 for each x > 0. Hence
we have for the derivative sin′ x = cos x > 0 and the function sin would be strictly in-
creasing. As sin 0 = 0, we would have sin x > 0 for x > 0. This would imply that for
0 < x1 < x2 ,
x2
(x2 − x1 ) sin x1 < ∫ sin t dt = cos x1 − cos x2 ≤ 2.
x1
Since we have sin x > 0, the last inequality gives a contradiction if x2 is sufficiently
large.
The zero set of the continuous function cos is closed and cos 0 ≠ 0, hence there
exists a smallest positive number x0 with cos x0 = 0.
We define the number π by
π ∶= 2x0 .
Then cos(π/2) = 0 and since cos2 x + sin2 x = 1 this implies sin(π/2) = ±1. As cos x > 0
in (0, π/2), the function sin is increasing in (0, π/2). Hence sin(π/2) = 1. This implies
π
exp( i) = i.
2
By Theorem 1.45, we get
In this way we get that all real zeroes of the function cos are of the form {(2k + 1)π/2 ∶
k ∈ ℤ}, and all real zeroes of the function sin are of the form {kπ ∶ k ∈ ℤ}.
Suppose that cos z = 0. Set z = x + iy and recall that 2 cos z = eiz + e−iz . It follows
that e2iz + 1 = 0, which implies e−2y cos 2x = −1, e−2y sin 2x = 0. Hence cos z = 0 implies
that z must be real and cos has no other zeroes. Also sin has no other zeroes.
ψ is surjective. To see this, let w ∈ ℂ∗ , then each z = x + iy with x = log |w| and
y = arg w is the pre-image of w, since we have ez = ex eiy = elog |w| ei arg w = |w|ei arg w ,
where the last expression corresponds to the polar representation of w.
ψ fails to be injective. Ker ψ = {z ∶ ψ(z) = 1} = {2πik ∶ k ∈ ℤ} = 2πiℤ, since 1 = ez =
ex (cos y + i sin y) implies x = 0 and y = 2πk, k ∈ ℤ.
The exponential function is periodic with period 2πi, i.e. exp(z + 2πi) = exp(z).
(c) Each strip {z ∈ ℂ ∶ a ≤ ℑz < a + 2π}, a ∈ ℝ, is mapped onto ℂ∗ by the exponential
function. This follows from (b).
In the following we introduce some other elementary functions related to the ex-
ponential function.
Definition 1.47.
sin z
tan z = , z ≠ (k + 1/2)π, k∈ℤ
cos z
cos z
cot z = , z ≠ kπ, k ∈ ℤ.
sin z
We have
1 e2iz − 1 e2iz + 1
tan z = , cot z = i ,
i e2iz + 1 e2iz − 1
1 1
(tan z)′ = , (cot z)′ = − 2 .
cos2 z sin z
Definition 1.48.
1 1
cosh z = (ez + e−z ), sinh z = (ez − e−z ).
2 2
We have
1 1
cosh z = cos iz, sinh z = sin iz, cos z = cosh iz, sin z = sinh iz.
i i
Now we investigate the inverse functions of the elementary functions introduced
above.
Definition 1.49. For each z ∈ ℂ∗ there are infinitely many w ∈ ℂ with ew = z. Each of
these values w is called a logarithm of z.
Each logarithm has the form
Theorem 1.51. Let G0 = {z ∈ ℂ ∶ −π < ℑz < π}. Then the exponential function
exp ∶ G0 ⟶ ℂ− is holomorphic and bijective, the inverse function is the principal
branch of the logarithm, it is a map from ℂ− onto G0 , which is also holomorphic and
bijective.
Log(exp z) = x + iℑz = x + iy = z.
If w ∈ ℂ− , then
Hence exp and Log are inverses to each other, and so both of them are bijective.
Since (exp z)′ = exp z ≠ 0 ∀z ∈ ℂ, it follows from the differentiation rule for inverse
functions (see Theorem 1.7) that the principal branch of the logarithm Log is holomor-
phic on ℂ− , and we have
1
(Log w)′ = , w ∈ ℂ− .
w
Remark. One has to be careful when using the functional equation which is valid for
the real logarithm. The following lines show that one must stay in the domain of the
branch of the logarithm when using the functional equation:
but
(z a )′ = az a−1 .
and
√2
√i = i1/2 = exp(1/2 Log i) = exp(iπ/4) = (1 + i).
2
Also here one has to be careful when using arithmetic rules like
(z α )β = z αβ ,
1.11 Exercises
1. Determine the real and imaginary part and the absolute value of
2 1+i 5 1 + i √3 4
, (1 + i√3)6 , ( ) , ( ) .
1 − 3i 1−i 1−i
1 r2 1 r2
ℜz = (z + ), ℑz = (z − ).
2 z 2i z
3 + 2i 4π 4π
2.718 − 3.010i, , 3(cos + i sin ).
5i − 4 3 3
3+i
6. Let z0 = −1−i . Compute z0123 .
√
9. Give a geometric description of the sets of all points determined by the following
relations:
(i) |z − 2 + 3i| < 5,
(ii) ℑz ≥ ℜz,
(iii) |z − i| + |z − 1| = 2,
(iv) ℜz = |z − 2|.
10. Let |w| < 1. Show that
z−w
| | < 1, for |z| < 1,
wz − 1
and
z−w
| | = 1, for |z| = 1.
wz − 1
11. Let f (z) = U(r, θ) + iV(r, θ), where z = r(cos θ + i sin θ). Let f be complex differen-
tiable at z0 ≠ 0. Show that at z0 we have
𝜕U 1 𝜕V 𝜕V 1 𝜕U
= and =− .
𝜕r r 𝜕θ 𝜕r r 𝜕θ
These are the Cauchy–Riemann differential equations in polar coordinates.
12. Let f (z) = u(x, y) + iv(x, y) and suppose that f satisfies the Cauchy–Riemann differ-
ential equations in an open set G ⊆ ℂ. Let G the set in ℂ obtained by the reflection
of G with respect to the real axis, i.e. (x, y) ∈ G, if (x, −y) ∈ G. Define a function g
on G by
g(z) = f (z), z ∈ G.
Show that g satisfies the Cauchy–Riemann differential equations for |z| > R. (Use
polar coordinates!)
14. Let a, b, c ∈ ℝ. For x, y ∈ ℝ and z = x + iy set P(z) = ax2 + 2bxy + cy2 . Find a neces-
sary and sufficient condition for the existence of a holomorphic function f ∈ ℋ(ℂ)
such that ℜ(f ) = P.
15. Suppose that f is holomorphic in ℂ and real-valued. Show that f is constant.
16. Let f (z) = |z|4 + (ℑz)2 . Compute fz and fz .
17. Let f be a real differentiable function. Prove that
𝜕f 𝜕f 𝜕f 𝜕f
= , = .
𝜕z 𝜕z 𝜕z 𝜕z
1.11 Exercises | 35
𝜕f 𝜕f
= .
𝜕z 𝜕z
𝜕(g ∘ f ) 𝜕g 𝜕f 𝜕g 𝜕f
= + ,
𝜕z 𝜕w 𝜕z 𝜕w 𝜕z
𝜕(g ∘ f ) 𝜕g 𝜕f 𝜕g 𝜕f
= + .
𝜕z 𝜕w 𝜕z 𝜕w 𝜕z
d(f ∘ φ) 𝜕f dφ 𝜕f dφ
= + .
dt 𝜕z dt 𝜕z dt
21. Let
1 1
f (z) = (z + ), z ∈ ℂ ⧵ {0}.
2 z
Show that f is angle preserving on ℂ ⧵ {−1, 0, 1}. Determine the images under f of
the circles |z| = r < 1 and of the rays z = ct, 0 < t < 1, |c| = 1 and use these image
curves in order to check the angle preservation of f .
22. Let
1
fn (z) = , a ≠ 0.
1 + az n
Show that the sequence (fn ) converges to 1 uniformly on each compact subset of
the open unit disc D1 (0) and that (fn ) converges to 0 uniformly on each compact
subset of ℂ ⧵ Dr (0), for r > 1.
23. Show that the series
zn
∞
∑ n n+1 )
n=1 (1 − z )(1 − z
converges to z(1 − z)−2 uniformly on each compact subset of D1 (0) and that it con-
verges to (1 − z)−2 uniformly on each compact subset of ℂ ⧵ D1 (0).
24. Where is the series
zn
∞
∑ n
n=1 1 − z
uniformly convergent?
36 | 1 Complex numbers and functions
an n
∞
∑ z , bn ≠ 0, n ∈ ℕ0
b
n=0 n
satisfies R0 ≥ min(R1 , R2 ).
26. Determine the radius of convergence of the following power series
∞ ∞ ∞
∑ (−1)n (z + i)n , ∑ (−1)n 2n z 2n+2 , ∑ n−1/2 z n ,
n=0 n=0 n=1
∞ n n
2 z n!
∞ ∞
(2n)! n
∑ , ∑ 2
z , ∑ n zn ,
n=1 (2n)! n=0 (n!) n=1 n
∞ ∞ ∞
n
∑ 2−n z 2 , ∑ 2log n z n , ∑ (n + an )z n ,
n=0 n=2 n=0
n2 n
∞ ∞ ∞
√(2n)! 2n n
∑ z , ∑ n−1 z 3 , ∑[ ] zn .
n=0 n! n=1 n=1 (n + 1)(n + 2)
1 2π
∞
∫ |f (reiθ )|2 dθ = ∑ |an |2 r 2n , 0 < r < R.
2π 0 n=0
|f (z)| ≤ M, z ∈ DR (0),
1.11 Exercises | 37
|an | ≤ r −n M(r), n ∈ ℕ0 .
where γ1 is the polygon from (1, 0) to (0, 0) and then to (0, 1), and γ2 is the poly-
gon from (0, 0) to (1, 1). γ3 is the square with vertices (0, 0), (1, 0), (1, 1), (0, 1) passed
through once in positive direction, γ4 is the unit circle |z| = 1 passed through once
in positive direction, and γ5 is a quarter of a circle with vertices (0, 0), (2, 0), (0, 2)
passed through once in positive direction.
29. Let Log denote the principal branch of the complex logarithm. Determine for
which z, w ∈ ℂ one has
(z α )β = z αβ .
31. Show that the complex sine function sin maps the strip
bijectively onto the region G obtained from the plane by deleting the two intervals
(−∞, −1] and [1, ∞). Determine the inverse function of sin on G in terms of the
principal branch of the complex logarithm.
32. Determine the real and imaginary part of tan z and cot z.
33. Show that for z = x + iy one has
sin2 x + sinh2 y sinh2 x + sin2 y
| tan z|2 = and | tanh z|2 = .
cos2 x + sinh2 y sinh2 x + cos2 y
34. Compute the sum of the geometric series
n
∑ ekiθ
k=0
1.12 Notes
We assume basic knowledge of real analysis and topology. The interested reader is
referred to other introductory textbooks for more details: [1, 4, 12, 11, 15, 21, 23, 26, 53,
63]. For interesting historical remarks, see [60].
2 Cauchy’s Theorem and Cauchy’s formula
The first section is devoted to winding numbers of closed paths. This is an essentially
topological concept but has important implications for the behavior of holomorphic
functions. Section 2.1 also contains a holomorphic parameter integral which will be
frequently used in the sequel. The theorem of Cauchy–Goursat1 is presented with the
classical proof of A. Pringsheim.2 One of the most important consequences is Cauchy’s
integral formula, which is a first example of a reproducing formula. This gives the tools
required in Sections 2.3, 2.4, 2.5 and 2.6 to prove the most important local properties
of holomorphic functions, like the local Taylor series expansion, behavior of zeroes of
holomorphic functions and the maximum principle. A more general holomorphic pa-
rameter integral is discussed in Section 2.7. The rest of the chapter is devoted to gener-
alizations of Cauchy’s integral formula – the inhomogeneous Cauchy formula, the ho-
mology and the homotopy version of it – and to meromorphic functions, their residues
and applications of the residue theorem to compute real integrals and Fourier3 trans-
forms.
Theorem 2.1 (Jordan4 Curve Theorem). Let γ ∶ [a, b] ⟶ ℂ be a closed Jordan curve
in ℂ, i.e. γ(a) = γ(b), but γ(s) ≠ γ(t) for any s, t ∈ (a, b), s ≠ t. Then the open set
ℂ ⧵ γ ∗ has two connected components, a bounded one and an unbounded one. The
bounded component is called the interior of γ, the unbounded one is called the exte-
rior of γ. The bounded one is simply connected and γ ∗ is the boundary of each of the
components.
https://doi.org/10.1515/9783110417241-002
40 | 2 Cauchy’s Theorem and Cauchy’s formula
1 dζ
Indγ (z) ∶= ∫ , z ∈ Ω.
2πi γ ζ − z
Then Indγ is an integer-valued function on Ω, this means Indγ (Ω) ⊆ ℤ, and Indγ is con-
stant on each connected component of Ω, in addition, Indγ (z) = 0 for each z belonging
to the unbounded connected component of Ω.
Indγ (z) is called the winding number of γ with respect to z.
Remark. The set γ ∗ is compact, hence there exists a disc D such that γ ∗ ⊂ D. The set
ℂ ⧵ D is connected and is therefore contained in a connected component of Ω. Hence
Ω has exactly one unbounded connected component.
Proof. We consider an arbitrary w ∈ Ω. There exists r > 0 such that Dr (w) ⊆ Ω. We will
show that
∞
f (z) = ∑ cn (z − w)n ,
n=0
where the sum is uniformly convergent on each compact subset of Dr (w). Then, by
Theorem 1.43, the function f is holomorphic on Dr (w) and as w ∈ Ω was an arbitrary
point, we conclude that f is holomorphic on Ω.
For this aim we take z ∈ Dr/2 (w) and observe that the denominator in the integral
can be written as
1 1 ϕ(t) − w 1 1
= = z−w
ϕ(t) − z ϕ(t) − w ϕ(t) − w − (z − w) ϕ(t) − w 1 − ϕ(t)−w
1 z−w n (z − w)n
∞ ∞
= ∑( ) =∑ n+1
,
ϕ(t) − w n=0 ϕ(t) − w n=0 (ϕ(t) − w)
2.1 Winding numbers | 41
since |ϕ(t) − w| > r, ∀t ∈ [a, b] and Ω ∩ ϕ∗ = ∅, the last sum converges uniformly for all
t ∈ [a, b].
Hence
b (z − w)n b (z − w)n
∞ ∞
f (z) = ∫ [ ∑ n+1
]ψ(t) dt = ∑ ∫ n+1
ψ(t) dt
a n=0 (ϕ(t) − w) n=0 a (ϕ(t) − w)
∞ b ψ(t)
∞
= ∑ [∫ dt](z − w)n = ∑ cn (z − w)n ,
n=0 a (ϕ(t) − w)n+1 n=0
where
b ψ(t) b−a
|cn | = |∫ dt| ≤ n+1 max |ψ(t)|.
a (ϕ(t) − w)n+1 r t∈[a,b]
1 dζ 1 b γ ′ (s)
Indγ (z) = ∫ = ∫ ds, z ∈ Ω,
2πi γ ζ − z 2πi a γ(s) − z
t γ ′ (s)
Φ(t) = exp[∫ ds],
a γ(s) − z
Φ′ (t) γ ′ (t)
= ,
Φ(t) γ(t) − z
excluding the finitely many points a = s0 < s1 < ⋯ < sn = b, where γ possibly fails to be
differentiable. In addition, we have
hence
except for finitely many points. On the other hand, the function
Φ(t)
t↦
γ(t) − z
is continuous on [a, b], which follows from the fundamental theorem of calculus,
hence it must be constant on [a, b]. As Φ(a) = e0 = 1, it follows that
1 Φ(a) Φ(t)
= = , t ∈ [a, b]
γ(a) − z γ(a) − z γ(t) − z
and
γ(t) − z
Φ(t) = , t ∈ [a, b].
γ(a) − z
Since γ is a closed path, we have γ(a) = γ(b), and hence Φ(b) = 1, and Indγ (z) ∈ ℤ.
By Lemma 2.3, the function
1 b γ ′ (s)
z ↦ Indγ (z) = ∫ ds
2πi a γ(s) − z
1 b γ ′ (s)
Indγ (z) = ∫ ds
2πi a γ(s) − z
we conclude that | Indγ (z)| < 1, if |z| is sufficiently large. Hence Indγ (z) = 0 on the un-
bounded connected component of Ω.
In the following we explain why Indγ (z) is called a winding number. For this pur-
pose let
t γ ′ (s)
λ(t) = ∫ ds.
a γ(s) − z
Then λ(b) = 2πi Indγ (z) and hence ℑλ(b) = 2π Indγ (z). Using the same notation as in
the proof from above, we get
γ(t) − z
exp(λ(t)) = Φ(t) =
γ(a) − z
2.2 The theorem of Cauchy–Goursat and Cauchy’s formula | 43
and arg Φ(b) = ℑλ(b) = 2π Indγ (z). Now we set z = 0. We have arg Φ(a) = 0 and
This means, if t runs through the interval [a, b], the expression Indγ (0) counts how
many times γ turns around z = 0.
For a circle, we get
{1 for |z − a| < r,
Indγ (z) = {
0 for |z − a| > r.
{
Proof. From (1) in Example 1.34 we know that
1 dζ
∫ = 1.
2πi γ ζ − a
∫ f (z) dz = 0.
𝜕Δ
The paths in the interior of the large triangle cancel each other out.
44 | 2 Cauchy’s Theorem and Cauchy’s formula
Figure 2.1
We denote the triangle where the maximum is attained by Δ(1) and repeat the proce-
dure from above for the triangle Δ(1) instead of Δ obtaining a triangle Δ(2) , and so on.
In this way we get a sequence of triangles Δ ⊃ Δ(1) ⊃ Δ(2) ⊃ ⋯ and after n steps the
inequality
there exists a limit point z0 ∈ Δ of the sequence (zn )n , in addition, z0 also belongs to
Δ(m) , m ∈ ℕ, which follows from the fact that the sequence (zn )∞ n=m+1 ⊂ Δ
(m)
. Hence we
have z0 ∈ ⋂n=1 Δ . z0 is uniquely determined, since L(𝜕Δ ) → 0 as n → ∞.
∞ (n) (n)
∫ (f (z0 ) + (z − z0 )f ′ (z0 )) dz = 0 ∀m ∈ ℕ.
𝜕Δ(m)
2.2 The theorem of Cauchy–Goursat and Cauchy’s formula | 45
Now we have
∫ f (z) dz = 0.
𝜕Δ
Figure 2.2
∫ =∫ +∫+⋯,
𝜕Δ 𝜕{a,x,y}
where we only have to handle the first summand, because the other integrals do not
contain the point p and are therefore 0 by the first step of the proof. For the first sum-
mand we can approach x and y arbitrarily close to a, by which
∫ f (z) dz → 0
𝜕{a,x,y}
as f is continuous.
If p lies in the interior of Δ, we can use the decomposition of the triangle in Fig-
ure 2.3 to reduce everything to the second step of the proof.
46 | 2 Cauchy’s Theorem and Cauchy’s formula
Figure 2.3
∫ f (z) dz = 0,
γ
∫ f (z) dz = 0
𝜕Δ
for each triangle Δ in Ω. Now, by Theorem 1.41, there exists a primitive of f on Ω and
finally, by Corollary 1.37.
∫ f (z) dz = 0
γ
Example 2.8. (a) Figure 2.4 shows a star-shaped domain with center z1 .
(b) Consider ℂ− . Here the center is to be chosen on the positive semi-axis.
Theorem 2.9. Let G be a star-shaped domain and let f ∈ ℋ(G). Then f has a primitive
on G and
∫ f (z) dz = 0,
γ
Figure 2.4
Proof. Theorem 1.41 is also valid for star-shaped domains as is easily seen. Hence we
can use the same proof as in Corollary 2.6.
Example 2.10. We consider ℂ− with 1 as center, we write z = reiϕ ∈ ℂ− , and take the
closed path γ, as shown in Figure 2.5.
Figure 2.5
dζ r dt ϕ ireit
∫ =∫ +∫ it
dt = log r + iϕ.
[1,z] ζ 1 t 0 re
1 f (ζ )
f (z) Indγ (z) = ∫ dζ .
2πi γ ζ − z
Proof. Let
{ f (ζ )−f (z) , if ζ ≠ z,
g(ζ ) = { ζ −z
f ′ (z), if ζ = z.
{
Then g is continuous on Ω and g ∈ ℋ(Ω ⧵ {z}). Hence g satisfies the assumptions of
Corollary 2.6. So we have
1
∫ g(ζ ) dζ = 0.
2πi γ
Finally, we get
1 f (ζ ) − f (z) 1 f (ζ ) 1 dζ
∫ dζ = ∫ dζ − f (z) ∫ = 0.
2πi γ ζ − z 2πi γ ζ − z 2πi γ ζ − z
We mention an important special case. Let γ(t) = z + reit , t ∈ [0, 2π], f ∈ ℋ(DR (z)),
R > r. Then Indγ (z) = 1 and
1 f (ζ )
f (z) = ∫ dζ .
2πi γ ζ − z
1 f (ζ ) 1 f (ζ )
|f (z)| = |∫ dζ | ≤ 2πr max | | = max |f (ζ )|.
2π γ ζ − z 2π ζ ∈γ ζ − z
∗ ζ ∈γ ∗
1 f (ζ ) 1 2π f (z + reit ) it 1 2π
f (z) = ∫ dζ = ∫ it
ire dt = ∫ f (z + reit ) dt,
2πi γ ζ − z 2πi 0 re 2π 0
Example 2.12. Using Cauchy’s formula, we compute the following line integral:
ez
I =∫ dz.
γ (z − 1)(z − 2)
ez eit
(a) If G = D3/4 (0), f (z) = (z−1)(z−2)
, f ∈ ℋ(G), γ(t) = 2
, t ∈ [0, 2π], then I = 0, by Corol-
lary 2.6.
2.2 The theorem of Cauchy–Goursat and Cauchy’s formula | 49
ez 3eit
(b) If G = D11/6 (0), f (z) = z−2
, f ∈ ℋ(G), γ(t) = 2
, t ∈ [0, 2π], then by Theorem 2.11
we have
1 f (ζ ) 1 eζ
f (1) = ∫ dζ = ∫ dζ .
2πi γ ζ − 1 2πi γ (ζ − 2)(ζ − 1)
Hence I = −2πie.
(c) If G = D4 (0), f (z) = ez , f ∈ ℋ(G), γ(t) = 3eit , t ∈ [0, 2π], then by Theorem 2.11 we
have
ez ez f (z) f (z)
I =∫ dz − ∫ dz = ∫ dz − ∫ dz = 2πi(e2 − e).
γ z−2 γ z−1 γ z−2 γ z−1
Theorem 2.13 (Taylor5 series expansion). Let Ω ⊆ ℂ be an open set and let a ∈ Ω, R > 0
be such that DR (a) ⊆ Ω. Suppose that f ∈ ℋ(Ω). Then
∞
f (z) = ∑ an (z − a)n ,
n=0
where the sum converges uniformly on all compact subsets of DR (a) and
f (n) (a)
an = , n = 0, 1, 2, … ,
n!
an are the Taylor coefficients of f in the expansion around the point a.
Proof. Let γ(t) = a + reit , r < R, t ∈ [0, 2π]. Then Indγ (z) = 1, ∀z ∈ Dr (a) (see Theo-
rem 2.4). By Theorem 2.11, we have
1 f (ζ ) 1 2π f (γ(t))γ ′ (t)
f (z) = ∫ dζ = ∫ dt ∀z ∈ Dr (a).
2πi γ ζ − z 2πi 0 γ(t) − z
As in the proof of Lemma 2.3 we expand the last integral into a power series
∞
f (z) = ∑ an (z − a)n ,
n=0
where
1 2π f (γ(t))γ ′ (t) 1 f (ζ )
an = ∫ n+1
dt = ∫ dζ
2πi 0 (γ(t) − a) 2πi γ (ζ − a)n+1
and the series converges uniformly on all compact subsets of Dr (a). Since r was an
arbitrary number such that r < R, we obtain the same assertion for DR (a).
By Theorem 1.43, we obtain the formula for the Taylor coefficients.
The proof of the last theorem contains the following important results:
Theorem 2.15. Under the same assumptions as in Theorem 2.13 and for γ(t) = a +
reit , r < R, t ∈ [0, 2π], we have
n! f (ζ )
f (n) (a) = ∫ dζ , n ∈ ℕ.
2πi γ (ζ − a)n+1
∫ f (z) dz = 0
𝜕Δ
Theorem 2.17. Let Ω ⊆ ℂ be a domain and f ∈ ℋ(Ω). We define the zero set of f by
Then Z(f ) = Ω or Z(f ) is a discrete subset of Ω (i.e. Z(f ) has no limit point in Ω).
In the second case, for each a ∈ Z(f ) there exists a uniquely determined integer ma ∈
ℕ such that
Proof. Let A be the set of all limit points of Z(f ) in Ω. Since f is continuous, we have
A ⊆ Z(f ).
Fix a ∈ Z(f ). There exists r > 0, such that Dr (a) ⊆ Ω. By Theorem 2.13, we can ex-
pand f into a Taylor series
∞
f (z) = ∑ an (z − a)n , z ∈ Dr (a).
n=0
Hence g ∈ ℋ(Dr (a)) and, by the definition of g, we obtain g ∈ ℋ(Ω). Since g(a) =
am ≠ 0 and since g is continuous, there exists an open neighborhood U of a such that
g(z) ≠ 0 for all z ∈ U, which implies that f (z) ≠ 0, for all z ∈ U ⧵ {a}. Now we have shown
that a is an isolated point in Z(f ), i.e. there exists a neighborhood U of a, containing
no other point of Z(f ).
In summary, we can state the following: if a ∈ A, then all an = 0, because otherwise
a would be an isolated point (by Case 2); so, by Case 1, we have Dr (a) ⊆ A and A is an
open set. But A is the set of all limit points of Z(f ) and is therefore also closed in Ω,
and the set B = Ω ⧵ A is open. We have Ω = A ∪ B, which is a union of two disjoint
open sets. We assumed that Ω is connected, hence Ω = A or A = ∅. If A = Ω we have
f ≡ 0 on Ω; if A = ∅ we get that Z(f ) is discrete in Ω. In this case, there are at most
finitely many points of Z(f ) in each compact subset of Ω, otherwise a limit point of
Z(f ) would belong to this compact subset and A ≠ ∅. We can write Ω as a countable
union of compact subsets of Ω:
∞
Ω = ⋃ (Dn (0) ∩ Ω1/n )− ,
n=1
where
The following results are easy but important consequences, which are also re-
ferred to as the identity principles.
Proof. Let ϕ = f − g. Then ϕ(z) = 0 for all z ∈ M and Theorem 2.17 implies that ϕ(z) = 0
on the whole of Ω.
{1, z ∈ D1 (0),
f (z) = {
0, z ∈ D1 (3).
{
Then f ∈ ℋ(Ω), f = 0 on the open subset D1 (3) of Ω, but f is not identically zero
on Ω.
Corollary 2.19. Let G be a domain in ℂ, let f ∈ ℋ(G) and suppose that f (n) (a) = 0, n =
0, 1, 2, …, for some a ∈ G. Then f ≡ 0 on G.
Proof. Let r > 0 be such that Dr (a) ⊆ G. Then, by Theorem 2.13, f can be expanded into
a Taylor series
f (n) (a)
∞
f (z) = ∑ (z − a)n
n=0 n!
for z ∈ Dr (a). Since f (n) (a) = 0, n = 0, 1, 2, …, we have f ≡ 0 on Dr (a). Now Theorem 2.18
implies that f ≡ 0 on G.
It follows that f ∈ 𝒞∞ (ℝ) and f (n) (0) = 0, n = 0, 1, 2, …, but f is not identically zero
on ℝ.
Definition 2.20. Let U ⊆ ℂ be open and f ∈ ℋ(U). We say that f has a holomorphic
logarithm on U, if there exists a function g ∈ ℋ(U) such that f = exp(g) on U.
We say that f has a holomorphic mth root on U, if there exists a function q ∈ ℋ(U)
such that qm = f on U.
2.4 Isolated singularities | 53
Theorem 2.21. Let G be a domain in ℂ, and f ∈ ℋ(G) with f ≠ 0 on G. Then the follow-
ing assertions are equivalent:
(1) f has a holomorphic logarithm on G.
(2) f ′ /f has a primitive on G.
ff ′
h′ = f ′ exp(−F) − fF ′ exp(−F) = f ′ exp(−F) − exp(−F) = 0.
f
Proof. By Theorem 2.21, it suffices to show that f ′ /f has a primitive on G. Now we use
Theorem 2.9: if c is a center for G, then
f ′ (ζ )
g(z) = ∫ dζ + b, f (c) = eb
[c,z] f (ζ )
Theorem 2.24. Let f ∈ ℋ(Ω ⧵ {a}) and denote by D′r (a) = {z ∶ 0 < |z − a| < r} the punc-
tured disc. Suppose that there exists a constant M > 0 such that |f (z)| ≤ M for all z ∈
D′r (a). Then f has a removable singularity at a.
54 | 2 Cauchy’s Theorem and Cauchy’s formula
Proof. Let
where we used that f is bounded on D′r (a). By Theorem 2.13, we can expand h into a
Taylor series in Dr (a),
∞
h(z) = ∑ an (z − a)n , z ∈ Dr (a).
n=0
where the series converges in Dr (a). Hence f ∈ ℋ(Dr (a)), and also that f ∈ ℋ(Ω).
Theorem 2.25. Let Ω ⊆ ℂ be an open set, let a ∈ Ω, and f ∈ ℋ(Ω ⧵ {a}). We distinguish
between three possible cases:
(a) f has a removable singularity at a.
(b) There exist c1 , … , cm ∈ ℂ, m ∈ ℕ, cm ≠ 0, such that
m
ck
f (z) − ∑
k=1 (z − a)k
Remark. The assertion from (c) is known as the Casorati7 –Weierstraß Theorem.
Proof. We will show that case (a) or (b) is valid, if (c) fails.
Suppose that (c) fails, then there exist r > 0 and δ > 0 and there exists some w ∈ ℂ
such that
|f (z) − w| > δ,
1
|g(z)| < ∀z ∈ D′r (a),
δ
hence g is bounded on D′r (a), so g has a removable singularity at a, see Theorem 2.24.
Now we have two cases:
Case 1. g(a) ≠ 0. Then there exist 0 < s < r and ρ > 0 with |g(z)| ≥ ρ, for all z ∈ Ds (a).
Hence
1 1
|f (z)| = | + w| ≤ + |w|,
g(z) ρ
for all z ∈ D′s (a), which means that f is bounded on D′s (a). By Theorem 2.24, f has a
removable singularity at a.
Case 2. g has a zero of order m ∈ ℕ at a (see Theorem 2.17). Therefore,
1
g(z) = = (z − a)m g1 (z)
f (z) − w
on D′r (a), we even have that g1 (z) ≠ 0 ∀z ∈ Dr (a). Now we set h = 1/g1 . Then h ∈
ℋ(Dr (a)) and
Example 2.26. (a) Let f (z) = sin z/z, z ∈ ℂ ⧵ {0}. Then f ∈ ℋ(ℂ ⧵ {0}). Since sin 0 = 0
and cos 0 = (sin z)′ |z=0 , we get from Theorem 2.17 that there exists g ∈ ℋ(ℂ) such
that sin z = zg(z). Hence f has a removable singularity at z = 0 and f ∈ ℋ(ℂ). Since
3 5 2 4
sin z = z − z3! + z5! − ⋯, we have f (z) = 1 − z3! + z5! − ⋯.
1
(b) Let f (z) = sin z . Then f ∈ ℋ(D′1 (0)). Since sin z = zg(z), g ∈ ℋ(ℂ), g(0) ≠ 0, there
exists r > 0 such that g(z) ≠ 0 for all z ∈ Dr (0). Setting h(z) = 1/g(z), we get h ∈
ℋ(Dr (0)). Hence
1 h(z) 1
∞
f (z) = = = ( ∑ bn z n ),
sin z z z n=0
where we used the Taylor series expansion of h on Dr (0) and that h(0) = b0 ≠ 0.
Therefore,
b0
f (z) − = b1 + b2 z + ⋯ ,
z
where the right-hand side is a holomorphic function on Dr (0). This means that
f (z) = sin1 z has a pole of order 1 at z = 0.
(c) Let f (z) = exp(1/z). Then f ∈ ℋ(ℂ∗ ). For n ∈ ℕ we get that f (1/n) → ∞ as n → ∞.
But |f (i/n)| = 1, for all n ∈ ℕ. Hence the assertions (a) and (b) from Theorem 2.25
are not valid. Hence f has an essential singularity at z = 0.
(d) In a similar way one shows that f (z) = exp(−1/z 2 ) has an essential singularity at
z = 0. But f |ℝ ∈ 𝒞∞ (ℝ).
n
Theorem 2.27. Let f (z) = ∑∞
n=0 cn (z − a) be a holomorphic function on DR (a). Let 0 <
r < R. Then
1 2π
∞
∑ |cn |2 r 2n = ∫ |f (a + reiθ )|2 dθ.
n=0 2π 0
1 2π 1 2π
∞ ∞
∫ |f (a + reiθ )|2 dθ = ∫ ( ∑ c r n einθ )( ∑ cn r n e−inθ ) dθ
2π 0 2π 0 n=0 n n=0
∞
= ∑ |cn |2 r 2n .
n=0
n
Proof. Let f (z) = ∑∞
n=0 cn z be a bounded entire function, i.e. there exists a constant
M > 0 such that |f (z)| ≤ M, for all z ∈ ℂ. Then, by Theorem 2.27, we have
∞
∑ |cn |2 r 2n ≤ M 2 ,
n=0
Remark. The assertion from above implies that the function z ↦ |f (z)| has no local
maximum in Ω.
Proof. If
1 2π
∞
∑ |cn |2 r 2n = ∫ |f (a + reiθ )|2 dθ ≤ |f (a)|2 = |c0 |2 .
n=0 2π 0
Hence cn = 0 for all n ∈ ℕ, and f ≡ f (a) on Dr (a). Since Ω is connected, we get from
Theorem 2.18 that f ≡ f (a) on the whole of Ω.
Proof. Suppose that there exists z0 ∈ Ω with maxz∈Ω |f (z)| = |f (z0 )|. Then there exists
r > 0 such that Dr (z0 ) ⊂⊂ Ω. By Theorem 2.30, we conclude that
for each θ ∈ [0, 2π] and each r > 1 + 2|a0 | + |a1 | + ⋯ + |an−1 |.
This assertion implies, that all zeros of p are contained in the disk DR (0) where
R = 1 + 2|a0 | + |a1 | + ⋯ + |an−1 |. In order to prove this assertion, we take r > R. Then
r > 1, and we get
In addition, we have
hence
for all sufficiently large r, which contradicts Theorem 2.30. Hence there exists z1 ∈ ℂ
with p(z1 ) = 0, and by Theorem 2.17 we have
p(z) = (z − z1 )m q(z)
for some m ∈ ℕ and a polynomial q with grad(q) < grad(p). (Expand p into a Taylor
series around the point z1 .)
Induction on the degree of the polynomial concludes the proof.
n!
|f (n) (a)| ≤ M (f ),
r n r,a
n! f (ζ )
f (n) (a) = ∫ dζ
2πi 𝜕Dr (a) (ζ − a)n+1
and hence
n! M (f )
|f (n) (a)| ≤ 2πr r,a .
2π r n+1
In the following we address the problem under which conditions the limit of a
sequence of holomorphic functions is again holomorphic. We point out that we use
the whole theory we have developed up to now. This result will also be fundamental
for the study of topological vector spaces of holomorphic functions and gives a first
occasion to use the terminology of functional analysis.
for all n, m > Nϵ,K . Then there exists a holomorphic function f ∈ ℋ(Ω) with limn→∞ fn = f
uniformly on all compact subsets of Ω.
We say that the space ℋ(Ω) is complete in the sense of uniform convergence on all
compact subsets of Ω.
Proof. f is the uniform limit of continuous functions and therefore again continuous.
Let Δ be a solid triangle in Ω. We get from Theorem 2.5 that
∫ fn (z) dz = 0,
𝜕Δ
∫ f (z) dz = 0.
𝜕Δ
K ⊂ ⋃ Dr (z) = U.
z∈K
The closure U is a compact subset of Ω and we can apply Theorem 2.33 to obtain
1
|fn′ (z) − f ′ (z)| ≤ max|f (w) − f (w)| ∀z ∈ K,
r w∈U n
Remark. We already mentioned that the space ℋ(Ω), endowed with the topology of
uniform convergence on compact subsets of Ω, is a complete topological vector space
(Theorem 2.34). The topology on ℋ(Ω) can be described by a metric. Indeed, consider
a so-called compact exhaustion of Ω, which is a sequence (Kj )j of compact subsets of
Ω such that Kj ⊂ K̊ j+1 , j ∈ ℕ and ⋃∞
j=1 Kj = Ω.
2.6 Open mappings | 61
∞
‖f − g‖j
d(f , g) = ∑ 2−j , f , g ∈ ℋ(Ω).
j=1 1 + ‖f − g‖j
The metric d generates the original topology of uniform convergence on all compact
subsets of Ω on ℋ(Ω), see Exercises.
ℋ(Ω) is a Fréchet space, i.e. a complete, metrizable topological vector space.
Remark. Recall that ϕ ∶ U ⟶ ℂ is continuous, if ϕ−1 (O) is open, for each open subset
O ⊆ ℂ. If ϕ is open and invertible, then ϕ−1 is continuous.
Example 2.36. Let ϕ ∶ ℝ ⟶ ℝ, ϕ(x) = x2 . We have ϕ((−1, 1)) = [0, 1). Hence ϕ is not
open.
Theorem 2.37 (Minimum Principle). Let U ⊆ ℂ be open and f ∈ ℋ(U), let c ∈ U and
let V a disc with center c and V ⊂ U. Suppose that
Proof. Suppose that f has no zero in V . Our assumption implies that there exists an
open neighborhood V1 of V such that V1 ⊆ U, and that f has no zero on V1 . Set g(z) =
1/f (z), z ∈ V1 . Then g ∈ ℋ(V1 ), and by Theorem 2.30 we have
−1
|f (c)|−1 = |g(c)| ≤ max|g(z)| = [min|f (z)|] ,
z∈𝜕V z∈𝜕V
Theorem 2.38 (Open Mapping Theorem). Let U ⊆ ℂ be an open set and f ∈ ℋ(U).
Suppose that there is no open subset of U where f is constant. Then f is an open mapping.
Proof. Let O ⊆ U be open and c ∈ O. We have to show that f (O) contains an open disc
with center f (c). Without loss of generality, we can suppose that f (c) = 0, otherwise
we could consider the function z ↦ f (z) − f (c) instead of f . We supposed that f is not
constant in any neighborhood of c. We claim that there exists a disc V with center c
such that V ⊂ O and 0 ∉ f (𝜕V). (If for each disk V with center c and V ⊂ O there exists
z0 ∈ 𝜕V with f (z0 ) = 0, then, by Theorem 2.18, f ≡ 0 in some neighborhood of c, which
contradicts our assumption on f .)
Now we set 2δ = minz∈𝜕V |f (z)| > 0 and D = Dδ (0). We will show that D ⊆ f (O). For
this aim let b ∈ D be an arbitrary point. As |b| < δ we have
Now we can apply the minimum principle Theorem 2.37 to the function z ↦ f (z) − b,
and get a point z ′ ∈ V with f (z ′ ) − b = 0. Hence f (z ′ ) = b, and finally b ∈ f (O).
{ f (z)−f (w) , z ≠ w, z, w ∈ Ω,
g(z, w) = { z−w
f ′ (z), z = w ∈ Ω.
{
Then g ∶ Ω × Ω ⟶ ℂ is continuous on Ω × Ω.
For z, w ∈ Dr (a) we set ζ (t) = (1 − t)z + tw, t ∈ [0, 1], which describes the straight line
from z to w. It is clear that ζ (t) ∈ Dr (a) ∀t ∈ [0, 1]. Now we compute the integral
2.6 Open mappings | 63
1 1 1 1 1 df dζ
∫ f ′ (ζ (t)) dt = ∫ f ′ (ζ (t))(−z + w) dt = ∫ dt
0 −z + w 0 −z + w 0 dζ dt
1 f (z) − f (w)
= f (ζ (t))|10 = .
−z + w z−w
Hence we get
f (z) − f (w) 1
|g(z, w) − g(a, a)| = | − f ′ (a)| = |∫ [f ′ (ζ (t)) − f ′ (a)] dt|
z−w 0
Theorem 2.40. Let Ω ⊆ ℂ be open, ϕ ∈ ℋ(Ω), z0 ∈ Ω and ϕ′ (z0 ) ≠ 0. Then there exists
an open neighborhood V of z0 , V ⊆ Ω such that
(1) ϕ|V is injective,
(2) the function ψ ∶ ϕ(V) ⟶ V defined by ψ(ϕ(z)) = z, z ∈ V , is holomorphic on
W = ϕ(V), ϕ has a holomorphic inverse on V .
1
|ϕ(z1 ) − ϕ(z2 )| ≥ |ϕ′ (z0 )||z1 − z2 | ∀z1 , z2 ∈ V,
2
for this aim one has to choose V in such a way that
ψ(w) − ψ(w1 ) z − z1
= ;
w − w1 ϕ(z) − ϕ(z1 )
1
ψ′ (w1 ) =
ϕ′ (z1 )
Now we are able to show the local form of a holomorphic function as indicated
above.
64 | 2 Cauchy’s Theorem and Cauchy’s formula
ϕ(z) = (z − z0 ) exp(h(z)/m).
Then
and since exp(h(z0 )/m) ≠ 0, we get ϕ′ (z0 ) ≠ 0. By taking a possibly smaller V we can
also get that ϕ′ ≠ 0 on V . The rest of the proof now follows from Theorem 2.40.
Remark. The converse of the last theorem is false. For example, f (z) = ez , f ′ (z) = ez ≠ 0
on ℂ. But the exponential function is not injective on ℂ.
direction was already used for the Taylor series expansion of holomorphic functions,
see Lemma 2.3. Now we use general properties of L1 -functions; see, for instance, [20].
Theorem 2.43. Let Ω ⊂ ℂ be open and let (X, μ) be a measure space with a positive
measure μ. Let
Proof. Let a ∈ Ω and choose r > 0 such that K ∶= D2r (a) ⊂ Ω. Then, by Cauchy’s formula
(see Theorem 2.11), we have for all z ∈ D2r (a)
1 f (ζ , x)
f (z, x) = ∫ dζ .
2πi 𝜕D2r (a) ζ − z
Now let (wk )k be a sequence in Dr (a) with limk→∞ wk = z, where wk ≠ z for all k; define
1 f (ζ , x)
φk (z, x) ∶= ∫ dζ .
2πi 𝜕D2r (a) (ζ − z)(ζ − wk )
Then
4πrgK (⋅) 2
|φk (z, ⋅)| ≤ = gK (⋅) μ-almost everywhere,
2πr 2 r
since |ζ − a| = 2r and wk , z ∈ Dr (a).
66 | 2 Cauchy’s Theorem and Cauchy’s formula
In addition, we have
f (z, ⋅) − f (wk , ⋅)
φk (z, ⋅) = ,
z − wk
since
f (ζ , x) f (ζ , x) (z − wk )f (ζ , x)
− = ,
ζ −z ζ − wk (ζ − z)(ζ − wk )
and we can apply Cauchy’s formula to φk . Hence φk (z, ⋅) is a measurable function. Con-
sidering the limit wk → z, we observe that f (ζ , x)/[(ζ −z)(ζ −wk )] tends to f (ζ , x)/(ζ −z)2
uniformly for ζ ∈ 𝜕D2r (a). Hence we can interchange limit and integration and obtain
1 f (ζ , x) 𝜕f
lim φk (z, x) = ∫ 2
dζ = (z, x),
k→∞ 2πi 𝜕D2r (a) (ζ − z) 𝜕z
where we used Cauchy’s formula for the first derivative of f with respect to z.
Now we apply the dominated convergence theorem (see [20]) and get the desired
assertion for n = 1. Using Cauchy’s formula for the higher derivatives, we obtain the
general result.
dx ∶ z ↦ (dx)z = x, z = x + iy,
𝜕f 𝜕f
(df )z0 ∶= (z ) dx + (z0 ) dy
𝜕x 0 𝜕y
as the complete differential of f at the point z0 . More general, we say that
α = f dx + g dy
hα ∶= hf dx + hg dy.
dz = dx + i dy and dz = dx − i dy.
2.8 Complex differential forms | 67
𝜕 1 𝜕 𝜕 𝜕 1 𝜕 𝜕
= ( −i ) and = ( + i ),
𝜕z 2 𝜕x 𝜕y 𝜕z 2 𝜕x 𝜕y
𝜕f 𝜕f 𝜕f 𝜕f
df = dx + dy = dz + dz.
𝜕x 𝜕y 𝜕z 𝜕z
ξ η ξ η1
dx ∧ dy (( 1 ) , ( 1 )) = | 1 | = ξ1 η2 − ξ2 η1 .
ξ2 η2 ξ2 η2
If f is a function,
ω = f (dx ∧ dy)
dx ∧ dx = dy ∧ dy = dz ∧ dz = dz ∧ dz = 0.
𝜕g 𝜕f
dα ∶= df ∧ dx + dg ∧ dy = ( − ) dx ∧ dy.
𝜕x 𝜕y
If α = F dz + G dz, we have
𝜕G 𝜕F
dα = ( − ) dz ∧ dz.
𝜕z 𝜕z
𝜕f
d(f dz) = − dz ∧ dz.
𝜕z
Example 2.45. Let z be a fixed point and let f be a real differentiable function, let ω
be the following 1-form
1 f (ζ )
ω(ζ ) = dζ for ζ ≠ z.
2πi ζ − z
68 | 2 Cauchy’s Theorem and Cauchy’s formula
Then
1 𝜕f /𝜕ζ 𝜕 1
dω(ζ ) = − [ + f (ζ ) (ζ ↦ )] dζ ∧ dζ
2πi ζ − z 𝜕ζ ζ −z
1 𝜕f /𝜕ζ
=− dζ ∧ dζ ,
2πi ζ − z
1
since ζ ↦ ζ −z
is a holomorphic function.
Let G be open in ℂ and k ∈ ℕ ∪ {∞}. Then 𝒞k (G) denotes the space of k times
continuously differentiable (in the real sense) functions on G. We also write 𝒞(G) in-
stead of 𝒞0 (G). The space 𝒞k (G) is the space of all functions f in 𝒞k (G) such that all
derivatives of f up to order k extend continuously to G.
∫ ω = ∫ dω.
𝜕G G
Theorem 2.47. Let G be a bounded domain in ℂ with piecewise smooth positively ori-
ented boundary 𝜕G. Let f ∈ 𝒞1 (G). Then for z ∈ G we have
1 f (ζ )
f (z) = ∫ dζ .
2πi 𝜕G ζ − z
Proof. Fix z ∈ G and choose r > 0 such that Dr (z) ⊆ G. We remove the disc Dr (z) from G
and define Gr = G ⧵ Dr (z), the boundary of Gr consists of the positively oriented bound-
ary of G and of the negatively oriented circle κr = 𝜕Dr (z). Walking on 𝜕Gr , the domain
Gr lies always on the left-hand side.
2.9 The inhomogeneous Cauchy formula | 69
as r → 0, since f is continuous.
So the line integral
1 f (ζ ) 1 f (ζ )
∫ dζ tends to ∫ dζ − f (z),
2πi 𝜕Gr ζ − z 2πi 𝜕G ζ − z
as r → 0, and we arrive at the desired result.
70 | 2 Cauchy’s Theorem and Cauchy’s formula
Definition 2.48. Let Ω ⊆ ℂ be open and let γ1 , … , γn be paths in Ω. Let Γ ∗ = ⋃nj=1 γj∗
and define
γ̃ j (f ) ∶= ∫ f (z) dz,
γj
for f ∈ 𝒞(Γ ∗ ). Γ is called a chain in Ω. If all paths γ1 , … , γn are closed, we call Γ a cycle
in Ω.
Remark. (a) Chains and cycles can be represented as sums of paths in many ways.
(b) By −Γ we denote the cycle, where each path γj , j = 1, … , n is replaced by its opposite
path, for f ∈ 𝒞(Γ ∗ ) we have
(c) If Γ1 and Γ2 are chains or cycles, we can form the sum Γ = Γ1 + Γ2 and have
1 dz
IndΓ (α) ∶= ∫ .
2πi Γ z − α
Obviously, we have
n
IndΓ (α) = ∑ Indγj (α).
j=1
1 f (w)
f (z) IndΓ (z) = ∫ dw ∀z ∈ Ω ⧵ Γ ∗ ;
2πi Γ w − z
2.10 General versions of Cauchy’s Theorem and Cauchy’s formula | 71
in addition,
∫ f (w) dw = 0.
Γ
then
Proof. Let
{ f (z)−f (w) , z ≠ w, z, w ∈ Ω,
g(z, w) = { z−w
f ′ (z), z = w ∈ Ω.
{
By Lemma 2.39, g ∶ Ω × Ω ⟶ ℂ is continuous. The first assertion of the theorem is
equivalent to the statement h(z) = 0 ∀z ∈ Ω ⧵ Γ ∗ , where
1
h(z) = ∫ g(z, w) dw,
2πi Γ
because
1 f (z) − f (w) 1 dw 1 f (w)
∫ dw = f (z) ∫ − ∫ dw
2πi Γ z − w 2πi Γ z−w 2πi Γ z − w
1 f (w)
= −f (z) IndΓ (z) + ∫ dw.
2πi Γ w − z
1
lim h(zn ) = lim ∫ g(zn , w) dw
n→∞ 2πi n→∞ Γ
1 1
= ∫ lim g(z , w) dw = ∫ g(z, w) dw = h(z),
2πi Γ n→∞ n 2πi Γ
and h is continuous on Ω; the limit and integral can be interchanged because of uni-
form convergence.
Now let Δ be an arbitrary closed triangle in Ω. Then Fubini’s theorem implies
1 1
∫ h(z) dz = ∫ (∫ g(z, w) dw) dz = ∫ (∫ g(z, w) dz) dw.
𝜕Δ 2πi 𝜕Δ Γ 2πi Γ 𝜕Δ
72 | 2 Cauchy’s Theorem and Cauchy’s formula
For w fixed, the function z ↦ g(z, w) has a removable singularity at z = w (see Theo-
rem 2.24), so it is holomorphic on Ω and, by Theorem 2.5, we have
∫ g(z, w) dz = 0 ∀w ∈ Ω,
𝜕Δ
hence
∫ h(z) dz = 0,
𝜕Δ
{h(z), z ∈ Ω,
ϕ(z) = {
h (z), z ∈ Ω1
{ 1
is holomorphic on Ω ∪ Ω1 .
By assumption, we have IndΓ (α) = 0 ∀α ∉ Ω, hence Ωc ⊆ Ω1 and Ω ∪ Ω1 = ℂ. So
ϕ ∈ ℋ(ℂ) is an entire function.
The set Ω1 contains the unbounded connected component of ℂ ⧵ Γ ∗ , since the
index is always zero there (see Theorem 2.2). Therefore
1 f (w)
lim ϕ(z) = lim h1 (z) = lim ∫ dw = 0,
|z|→∞ |z|→∞ 2πi |z|→∞ Γ w − z
and Liouville’s theorem (see Theorem 2.29) implies ϕ ≡ 0, in particular h(z) = 0 ∀z ∈
Ω ⧵ Γ∗ .
Finally, we have to show that ∫Γ f (z) dz = 0. Let a ∈ Ω ⧵ Γ ∗ and F(z) ∶= (z − a)f (z).
From the first assertion of the theorem, we get for F and z = a that
1 F(w) 1
0 = F(a) IndΓ (a) = ∫ dw = ∫ f (w) dw.
2πi Γ w − a 2πi Γ
For the last assertion of the theorem, we pick two cycles Γ0 and Γ1 with
Define Γ = Γ0 − Γ1 , then IndΓ (α) = 0 ∀α ∉ Ω and, by the first part of the theorem,
Therefore, the assumptions of Theorem 2.50 are satisfied, hence Theorem 2.50 im-
plies Corollary 2.11.
Example 2.51. Let Ω = ℂ ⧵ (D1/2 (−2) ∪ D1/2 (0) ∪ D1/2 (2)), and let γ1 (t) = −2 + 43 eit , γ2 (t) =
3 it
4
e , γ3 (t) = 2 + 43 eit , Γ(t) = 6eit , t ∈ [0, 2π]. Define γ = γ1 + γ2 + γ3 . Then
H ∶ [0, 1] × [0, 1] ⟶ Ω
such that
and
Set γt (s) = H(s, t) for a fixed t ∈ [0, 1] and s ∈ [0, 1]. Since H(0, t) = H(1, t), the curves γt
are also closed. We get a one-parameter family γt , t ∈ [0, 1] of closed curves connecting
γ0 and γ1 .
If a closed curve γ0 is Ω-homotopic to a constant curve (consisting of just one
point), we say that γ0 null-homotopic in Ω.
A domain Ω is said to be simply connected, if every closed curve in Ω is null-
homotopic in Ω.
74 | 2 Cauchy’s Theorem and Cauchy’s formula
Example 2.53. Let Ω be a convex domain and let γ be a closed curve in Ω. Fix z0 ∈ Ω.
Define
Then H(s, 0) = γ(s) and H(s, 1) = z0 ∀s ∈ [0, 1]. Since γ(0) = γ(1), we have
For a fixed s, the expression H(s, t) = (1 − t)γ(s) + tz0 , t ∈ [0, 1] describes the straight
line from γ(s) to z0 , which is contained in Ω, as Ω is convex.
Hence H(s, t) ∈ Ω ∀s, t ∈ [0, 1] and γ is null-homotopic in Ω. Therefore, Ω is simply
connected.
Lemma 2.54. Let γ0 and γ1 be closed paths in ℂ. Let α ∈ ℂ be a complex number such
that
1 dz 1 1 γ ′ (s)
0 = Indγ (0) = ∫ = ∫ ds
2πi γ z 2πi 0 γ(s)
1 1 γ1′ (s) 1 1 γ0′ (s)
= ∫ ds − ∫ ds = Indγ1 (α) − Indγ0 (α).
2πi 0 γ1 (s) − α 2πi 0 γ0 (s) − α
2.10 General versions of Cauchy’s Theorem and Cauchy’s formula | 75
Proof. Let H be a homotopy function between Γ0 and Γ1 . The difficulty of the proof
relies on the fact that the one-parameter family Γt (s) = H(s, t) does not consist of nec-
essarily piecewise differentiable curves. We will approximate the curves Γt by suitable
paths, in our case by polygonal closed paths.
Fix α ∉ Ω. Since H is uniformly continuous on the compact set [0, 1] × [0, 1], there
exists ϵ > 0 such that
if |s − s′ | + |t − t ′ | < 1/n.
Now we define the approximating polygonal paths: for k = 0, 1, … , n and s ∈ [0, 1]
let
j k j−1 k
γk (s) = H( , )(ns + 1 − j) + H( , )(j − ns),
n n n n
for j − 1 ≤ ns ≤ j, j = 1, … , n.
It is easily seen that the curves γ0 , γ1 , … , γn are closed. They are also piecewise dif-
ferentiable, since the variable s only appears as a linear term, and not as an argument
of the function H. By (2.4) and from the definition of γk , we get
k
|γk (s) − H(s, )| < ϵ, (2.5)
n
k j k j−1 k k
|γk (s) − H(s, )| = |H( , )(ns + 1 − j) + H( , )(j − ns) − H(s, )|
n n n n n n
j k j−1 k j k k
≤ (j − ns)|H( , ) − H( , )| + |H( , ) − H(s, )|
n n n n n n n
< ϵ/2 + ϵ/2 = ϵ.
76 | 2 Cauchy’s Theorem and Cauchy’s formula
In particular, for k = 0
and for k = n
j k−1 j k
|γk−1 (s) − γk (s)| ≤ (ns + 1 − j)|H( , ) − H( , )|
n n n n
j−1 k−1 j−1 k
+ (j − ns)|H( , ) − H( , )|
n n n n
< ϵ/2 + ϵ/2 = ϵ.
∀s ∈ [0, 1] and k = 1, … , n.
Similarly,
|γ0 (s) − Γ0 (s)| < |α − Γ0 (s)| and |γn (s) − Γ1 (s)| < |α − Γ1 (s)|,
∀s ∈ [0, 1].
Now we can apply Lemma 2.54 to the pairs
and get
IndΓ0 (α) = Indγ0 (α) = Indγ1 (α) = ⋯ = Indγn−1 (α) = Indγn (α) = IndΓ1 (α).
2.10 General versions of Cauchy’s Theorem and Cauchy’s formula | 77
Corollary 2.56. (a) Let Ω be a simply connected domain in ℂ and let γ be a closed
path in Ω. Then
∫ f (z) dz = 0,
γ
n! f (z)
f (n) (a) Indγ (a) = ∫ dz, n ∈ ℕ,
2πi γ (z − a)n+1
Proof. (b) For n = 1, the assertion follows from Theorem 2.50. Finally, differentiation
inside the integral with respect to the variable a gives the desired result.
dz
∫ =0 ∀α ∉ Ω,
γ z−α
this follows from Theorem 2.55, since the function z ↦ 1/(z − α) is holomorphic and
γ is null-homotopic in Ω. This implies that γ is also null-homologous. Every null-
homotopic path is null-homologous. The converse is false (see Exercises).
But if every closed path in Ω is null-homologous, then every closed path is also
null-homotopic and Ω is simply connected (see Chapter 4).
We conclude this chapter with two useful results about the index of a path.
Definition 2.57. Let G be a domain in ℂ and γ ∶ [0, 1] ⟶ ℂ a path. We say that γ runs
in G from boundary to boundary, if
(1) there exists t1 , t2 ∈ [0, 1], t1 < t2 , γ(t1 ), γ(t2 ) ∈ 𝜕G, γ(t1 ) ≠ γ(t2 );
(2) γ(t) ∈ G, for t1 < t < t2 ;
(3) γ(t) ∉ G, for t ∈ [0, 1] but t ∉ [t1 , t2 ];
(4) G ⧵ γ ∗ has exactly two connected components and γ ∗ ∩ G belongs to the boundary
of both of these components.
Theorem 2.58. Let γ be a closed path in ℂ and let D be a disc. Suppose that γ runs in D
from boundary to boundary. Let t1 , t2 ∈ [0, 1] with t1 < t2 and a = γ(t1 ), b = γ(t2 ), a, b ∈ 𝜕D,
78 | 2 Cauchy’s Theorem and Cauchy’s formula
and γ|[t1 ,t2 ] = γ0 . Let D1 , D2 denote the two connected components of D ⧵ γ ∗ . Suppose that
D1 lies to the left of γ. Then
for z1 ∈ D1 and z2 ∈ D2 .
Remark. Let γ in ℂ be an arbitrary path. The index Indγ (z) = 0 for z in the unbounded
component of ℂ ⧵ γ ∗ . Now one can use the theorem above, in order to compute the
indices of γ one after the other.
Figure 2.6
Then
and hence
Similarly,
This gives
Indγ (z1 ) − Indγ (z2 ) = Indγ0 (z1 ) − Indγ0 (z2 ) + Indγ1 (z1 ) − Indγ1 (z2 )
= Indγ0 (z1 ) − Indγ0 (z2 ) + Indκ1 (z1 ) − Indκ1 (z2 )
= Indγ0 (z1 ) + Indκ1 (z1 )
= Indγ0 (z1 ) + Indκ1 (z1 ) + Indκ2 −γ0 (z1 )
= Indκ1 +κ2 (z1 ) = 1.
Lemma 2.59. Let A ⊂ ℂ be a compact subset and U ⊃ A an open set. Then there exists
a cycle Γ in U ⧵ A such that
Proof. Case 1. First we suppose that A is connected. Let δ > 0 be such that 0 < 2δ <
dist(A, 𝜕U). We use a lattice parallel to the axes with mesh width δ and positively
oriented lattice squares. Since A is compact, there exist finitely many lattice squares
Q1 , … , Qn with Qj ∩ A ≠ ∅, j = 1, … , n. Let Γj be the boundary cycle of Qj and define
Γ = Γ 1 + ⋯ + Γn .
If a belongs to 𝜕Qj for some j, then a belongs to the interior of four adjacent squares
of the lattice, and we get the same result IndΓ (a) = 1.
Now we modify Γ in the following way: we take only line segments [p, q], which
are line segments of exactly one square of our collection, these are line segments [p, q]
with [p, q] ∩ A = ∅; line segments with non-empty intersection with A are passed
through in both directions and drop out. The modified boundary cycle is again de-
noted by Γ.
Now we have Γ ∗ ∩ A = ∅ and, by the choice of the mesh width, that Γ ∗ ⊂ U. Hence
Γ ⊂ U ⧵ A. As A is connected, we get from Theorem 2.2 that IndΓ (a) = 1 ∀a ∈ A.
∗
80 | 2 Cauchy’s Theorem and Cauchy’s formula
and we also choose the lattice such that different aj belong to the interior of different
squares of the lattice.
Then IndΓ (aj ) = 1 and hence IndΓ (a) = 1 for every each a ∈ Aj , j = 1, … , N. Every-
thing else can now be reduced to the first case.
Case 3. In the general case, A could have infinitely many connected components.
For every z ∈ A there exists an open square Q(z) with line segments parallel to the axes
such that z ∈ Q(z) ⊂⊂ U; since A is compact, finitely many of these squares cover A. We
denote them by Q1 , … , Qm . Now let
m
A0 = ⋃ Qj ⊂ U.
j=1
Theorem 2.60. Let Dr,R (a) = {z ∈ ℂ ∶ r < |z − a| < R} be an annulus, we define D0,R (a) =
DR (a) ⧵ {a} and Dr,∞ (a) = {z ∶ |z − a| > r}. Let f ∈ ℋ(Dr,R (a)), and define U1 = Dr,∞ (a),
U2 = DR (a). Then there exist functions f1 ∈ ℋ(U1 ) and f2 ∈ ℋ(U2 ) such that
f = f1 + f2 , on U1 ∩ U2 = Dr,R (a).
The function f1 can be chosen with the property lim|z|→∞ f1 (z) = 0. In this way f1 and f2
are uniquely determined.
1 f (ζ )
f2,ρ (z) = ∫ dζ ,
2πi γρ ζ − z
where γρ (t) = a + ρe2πit , t ∈ [0, 1]. The function in the integral is holomorphic for z in
Dρ (a), hence f2,ρ ∈ ℋ(Dρ (a)). By Theorem 2.55, we have f2,ρ (z) = f2,ρ′ (z) on Dρ (a) for
r < ρ < ρ′ < R.
2.11 Laurent series and meromorphic functions | 81
1 f (ζ )
f2 (z) = ∫ dζ ,
2πi γρ ζ − z
1 f (ζ )
f1 (z) = − ∫ dζ .
2πi γσ ζ − z
Similarly, we get f1 ∈ ℋ(U1 ), and from the definition of f1 we derive immediately that
lim|z|→∞ |f1 (z)| = 0.
For z ∈ Dr,R (a), we choose ρ and σ such that
{f − g1 on U1 ,
h={ 1
g −f on U2 ,
{ 2 2
and get a holomorphic function on U1 ∪ U2 = ℂ.
Hence h ∈ ℋ(ℂ) and lim|z|→∞ |h(z)| = 0. So h is a bounded entire function. By
Liouville’s Theorem (see Theorem 2.29), it follows that h ≡ 0, and f1 = g1 as well as
f2 = g2 .
Since lim|z|→∞ |f1 (z)| = 0, we have limw→0 (f1 ∘ F)(w) = 0. Hence f1 ∘ F has a remov-
able singularity at w = 0 (see Theorem 2.24), in addition, f1 ∘ F ∈ ℋ(D1/r (0)) and we
can expand it as a Taylor series around w = 0 to get
∞
(f1 ∘ F)(w) = ∑ bn wn ,
n=1
Theorem 2.61. Let f ∈ ℋ(Dr,R (a)). Then f can be represented in the form
−∞ ∞
f (z) = ∑ an (z − a)n + ∑ an (z − a)n ,
n=−1 n=0
which is the Laurent9 series of f in Dr,R (a), and the series converges uniformly on all
compact subsets of Dr,R (a). The Laurent coefficients an are given by
1 f (z)
an = ∫ dz, n∈ℤ
2πi γρ (z − a)n+1
Proof. It remains to prove the formula for the Laurent coefficients an . We have
−∞ ∞
(z − a)−n−1 f (z) = ∑ ak+n+1 (z − a)k + ∑ ak+n+1 (z − a)k
k=−1 k=0
f (z) dz
∫ dz = an ∫ = 2πian ,
γρ (z − a)n+1 γρ z − a
1
f (z) = .
z(z − i)2
(a) Its Laurent series expansion in D′1 (0) = {z ∶ 0 < |z| < 1} is
1 1 1 1 z n 1 n+2
∞ ∞
2
=− 2
= − ∑ (n + 1)( ) = − + i ∑ n z n .
z(z − i) z (1 − z/i) z n=0 i z n=0 i
1 1 1
−∞
2
= 3 = ∑ i−n−1 (n + 2)z n .
z(z − i) z (1 − i/z)2 n=−3
1 −i 1 1 −i 1 i
2
= 2
+ − = 2
+ − = ⋯,
z(z − i) (z − i) z − i z (z − i) z − i 1 − i(z − i)
where the last term can be written as the sum of a geometric series.
(2) Consider the Laurent series
zn
∞ ∞
∑ z −n + ∑ n+1
n=1 n=0 2
1
with infinitely many negative powers of z. The first summand converges to z−1 for
1
|z| > 1, and the second summand converges to 2−z for |z| < 2. So the whole series
1 1
converges to the function f (z) = z−1 + 2−z on the annulus D1,2 (0). The function f
is holomorphic at z = 0, although the Laurent series in the annulus D1,2 (0) has
infinitely many negative powers of z.
where a−k ≠ 0.
In addition, it follows that f has an essential singularity in a if and only if the
Laurent expansion of f in the punctured disc D′r (a) has infinitely many terms of the
form a−k (z − a)−k , where k > 0 and a−k ≠ 0.
Example 2.64. (1) Let U = D1 (0) and f (z) = 1/z. Then f ∈ ℳ(U).
Every rational function p/q, where p and q are polynomials, belongs to ℳ(ℂ).
sin z
(2) tan z = cos z
belongs to ℳ(ℂ). It is easily seen that tan z has infinitely many poles.
(3) Let f , g ∈ ℋ(U) and suppose that g ≢ 0 on every connected component of U. Then
f /g ∈ ℳ(U) (see Theorem 2.17).
84 | 2 Cauchy’s Theorem and Cauchy’s formula
Theorem 2.65. Let f ∈ ℳ(U). Then, for each a ∈ U, there exists an open neighborhood
V of a, and g, h ∈ ℋ(V) such that f = g/h on V .
lim |f (z)| = ∞,
z→a
i.e. for each compact subset K ⊂ ℂ there exists δ > 0 with f (D′δ (a)) ⊆ ℂ ⧵ K.
Theorem 2.67. Let Ω be a domain in ℂ. Then ℳ(Ω) is a field with respect to pointwise
addition and multiplication of functions.
∫ f (z) dz,
γs
where γs (t) = a + seit , t ∈ [0, 2π] and 0 < s < r, one observes that the integral can be
computed term by term in the Laurent series expansion and that only one term is left,
namely
c−1
∫ dz = 2πic−1 ,
γs z−a
1
∫ f (z) dz = ∑ Res(f ; a) IndΓ (a).
2πi Γ a∈S f
Proof. First we show that the set B = {a ∈ Sf ∶ IndΓ (a) ≠ 0} is finite, which implies that
the sum in the theorem is a finite sum. For this let W = ℂ ⧵ Γ ∗ . The index IndΓ is con-
stant on every connected component V of W . If V is unbounded or if V ∩ (ℂ ⧵ U) ≠ ∅,
then IndΓ (α) = 0 ∀α ∈ V , by our assumption that IndΓ (α) = 0 ∀α ∉ U.
Sf has no limit point in U, the limit points of Sf can only be on the boundary of U,
therefore limit points can belong to the unbounded component of W or to a component
V with V ∩ (ℂ ⧵ U) ≠ ∅. We have dist(Γ ∗ , 𝜕U) > 0, hence B must be finite.
Let B = {a1 , a2 , … , an }, and let Qj be the principal parts of f in the Laurent expan-
sion around aj , j = 1, … , n. Define
g = f − (Q1 + Q2 + ⋯ + Qn ),
∫ g(z) dz = 0
Γ
86 | 2 Cauchy’s Theorem and Cauchy’s formula
Remark. If U is open and convex and Γ is a closed, positively oriented path with-
out double points in U and f is holomorphic in U except for isolated singularities,
then
1
∫ f (z) dz = ∑ Res(f ; ak ),
2πi Γ
For the applications it will be convenient to know about some simple rules how
to compute residues.
Theorem 2.70. Let U ⊆ ℂ be open and let f and g be holomorphic on U except for
isolated singularities. Then
(a)
Res(f + g; a) = Res(f ; a) + Res(g; a)
and for α1 , α2 ∈ ℂ
1 dn−1
Res(f ; z0 ) = lim { n−1 [(z − z0 )n f (z)]}.
(n − 1)! z→z0 dz
c−1
∞
f (z) = + ∑ c (z − z0 )n ,
z − z0 n=0 n
hence
∞
(z − z0 )f (z) = c−1 + ∑ cn (z − z0 )n+1 ,
n=0
c−1
∞ ∞
g(z) = g(z0 ) + ∑ bn (z − z0 )n , f (z) = + ∑ c (z − z0 )n ,
n=1 z − z0 n=0 n
this implies
g(z0 )c−1
∞
f (z)g(z) = + ∑ dn (z − z0 )n .
z − z0 n=0
(e) We have
c−n c
∞
f (z) = + ⋯ + −1 + ∑ ck (z − z0 )k
(z − z0 )n z − z0 k=0
1 f ′ (z)
Nf − Pf = ∫ dz = IndΓ (0),
2πi γ f (z)
where Γ = f ∘ γ.
Proof. Let ϕ = f ′ /f . Then ϕ ∈ ℳ(Ω). Now let a be a zero of f of order m(a). Then, by
Theorem 2.17,
1 f ′ (z)
∫ dz = ∑ Res(ϕ; a) + ∑ Res(ϕ; b) = ∑ m(a) − ∑ p(b) = Nf − Pf .
2πi γ f (z) a∈A b∈B a∈A b∈B
1 dz 1 1 f ′ (γ(s)) ′ 1 f ′ (z)
IndΓ (0) = ∫ = ∫ γ (s) ds = ∫ dz = Nf − Pf .
2πi Γ z 2πi 0 f (γ(s)) 2πi γ f (z)
Our assumption |g1 (z) − g2 (z)| < |g1 (z)| ∀z ∈ γ ∗ implies that g2 has no zeros on γ ∗ . Let
Γ1 = g1 ∘ γ and let Γ2 = g2 ∘ γ. Then
Now we apply Lemma 2.54 and the first part of the proof to obtain
Example 2.72. Let g(z) = z 4 − 4z + 2. How many zeros does g have in D1 (0)?
On |z| = 1 we have |z|4 = 1 < 2 ≤ |−4z + 2|. Set f (z) = −4z + 2, then on |z| = 1 we have
f has exactly one zero in D1 (0), namely z0 = 1/2, hence, by Theorem 2.71, we get that g
also has exactly one zero in D1 (0).
Example 2.73 (Applications of the Residue Theorem). First we compute the real def-
inite integrals
∞ cos x ∞ sin x
∫ dx and ∫ dx.
−∞ x2 − 2x + 2 −∞ x 2 − 2x + 2
where γR is the path consisting of the line segment on the real axis from −R to R and
the semicircle ΓR from R to −R in the upper half-plane, we take R > 3.
The function eiz /(z 2 − 2z + 2) has poles of the first order at the points 1 + i and 1 − i.
The point 1 + i lies in the interior of γR , and we use Theorem 2.70 (b) to compute the
residue of this function at 1 + i:
eiz
∫ dz = 2πi(−ie−1+i )/2 = πe−1+i . (2.7)
γR z2 − 2z + 2
eiz 1
| |≤ 2 ,
z 2 − 2z + 2 R − 2R − 2
eiz πR
|∫ dz| ≤ 2 ,
ΓR z2 − 2z + 2 R − 2R − 2
and we obtain
eiz
∫ dz → 0 as R → ∞.
ΓR z 2 − 2z + 2
90 | 2 Cauchy’s Theorem and Cauchy’s formula
Taking real and imaginary parts on each side and using the dominated convergence
theorem, we finally obtain
∞ cos x π cos 1 ∞ sin x π sin 1
∫ dx = and ∫ dx = .
−∞ x2 − 2x + 2 e −∞ x2 − 2x + 2 e
In the next example we consider two polynomials P and Q with grad Q ≥ grad P +2.
Suppose that Q(x) ≠ 0 ∀x > 0 and that Q(0) = 0 is a simple zero. Let 0 < α < 1 and
R = P/Q. Using the Residue Theorem we will compute the real definite integral
∞
∫ xα R(x) dx,
0
For δ, ϵ > 0 small and ρ > 0 large, choose the following closed path γ in Ω from
Figure 2.7.
Figure 2.7
2.12 The residue theorem | 91
1
∫ eαg(z) R(z) dz = ∑ Res(f ; a), (2.8)
2πi γ a
where f (z) = eαg(z) R(z) and the sum is taken over all poles of f in the interior of γ.
Observe that
and
M
|R(z)| ≤ , M>0
|z|
Hence
M′
|R(z)| ≤ , M′ > 0
|z|2
|∫ f (z) dz| ≤ 2πρ max (|z|α |R(z)|) ≤ 2πρα+1 M ′ /ρ2 = 2πM ′ ρα−1 ,
C1
∗ z∈C1
Hence we obtain from (2.9) and (2.10), using the dominated convergence theorem,
∞
∫ f (z) dz = ∫ + ∫ + ∫ + ∫ → (1 − e2πiα ) ∫ eα log x R(x) dx,
γ C1 L1 L2 C2 0
zα iα 1
Res(z α /(1 + z 2 ); i) = lim[(z − i) ] = = ei(α−1)π/2 ,
z→i (z + i)(z − i) 2i 2
α
z (−i)α 1 i(α−1)3π/2
Res(z α /(1 + z 2 ); −i) = lim [(z + i) ]= = e ,
z→−i (z + i)(z − i) −2i 2
1 1 eisz
ϕA (s) = ∫ dz.
π 2πi ΓA z
2.12 The residue theorem | 93
Figure 2.8
Then
A sin x itx
∫ e dx = ϕA (t + 1) − ϕA (t − 1).
−A x
The function z ↦ eisz /z has a pole of the first order at 0 with residue 1. Hence, by
the Residue Theorem (see Theorem 2.69), we have
1 eisz
∫ dz = 0,
2πi γ z
where the path γ is shown in Figure 2.9.
Figure 2.9
In addition,
1 eisz
∫ dz = 1,
2πi δ z
1 eisz 1 1 −π iAeiθ
∫ dz = ϕA (s) + ∫ exp(isAeiθ ) iθ dθ = 0,
2πi γ z π 2πi 0 Ae
which implies
1 1 0
ϕA (s) = ∫ exp(isAeiθ ) dθ, (2.12)
π 2π −π
94 | 2 Cauchy’s Theorem and Cauchy’s formula
Figure 2.10
and
1 eisz 1 1 π
∫ dz = ϕA (s) + ∫ exp(isAeiθ ) dθ = 1,
2πi δ z π 2π 0
and finally,
1 1 π
ϕA (s) = 1 − ∫ exp(isAeiθ ) dθ. (2.13)
π 2π 0
{π, s > 0,
lim ϕA (s) = {
A→∞ 0, s < 0.
{
1 ∞ sin x
∫ χ(t)e−itx dt = .
2π −∞ x
2.13 Exercises | 95
2.13 Exercises
35. Compute the following line integrals:
∫ (z − a)k dz,
γ1
where a ∈ ℂ, k ∈ ℤ and γ1 is the unit circle |z| = 1 passed through once in the pos-
itive direction.
36. Let γ(t) = eit , 0 ≤ t ≤ 2π. Compute the following line integrals and compare the
results with the assertions of Cauchy’s Theorem:
dz 2πi
∫ = .
γ (z − a)(z − b) a − b
dz
∫ .
γ (z − a)m (z − b)n
eiz sin z
∫ dz, ∫ dz.
γ z2 γ z3
ez − e−z dz
∫ dz, ∫ .
γ zn γ (z − 1/2)n
Log(z)
∫ dz.
γ zn
dz
∫ .
γ (z + 1)(z − 1)3
e1−z
∫ dz.
γ z 3 (1 − z)
45. Compute
z2 + 1
∫ dz,
γ z(z 2 + 4)
where γ(t) = reit , 0 ≤ t ≤ 2π, first for 0 < r < 2 and then for 2 < r < ∞.
46. Prove that
2π
∫ cos(cos θ) cosh(sin θ) dθ = 2π.
0
Hint: use the mean value property for the function f (z) = cos z.
47. (a) Let U ⊆ ℂ be an open and L a straight line, suppose that f ∶ U ⟶ ℂ is contin-
uous and holomorphic on U ⧵ L. Show that f is holomorphic on the whole of
U (Use Morera’s Theorem!).
(b) Let G be a domain in ℂ, which is symmetric with respect to the real axis, i.e.
if z ∈ G then z ∈ G. Let
f ∶ {z ∈ G ∶ ℑz ≥ 0} ⟶ ℂ
{f (z), for ℑz ≥ 0,
f ̃(z) = {
f (z), for ℑz < 0
{
is a holomorphic function on G (Schwarz’s reflection principle).
48. Examine which functions can be holomorphically extended into the point 0:
z 1
z cot z, , z 2 sin .
ez − 1 z
1 z5 ez 1 − ez
, , , ,
z − z3 (1 − z)2 1 + z2 1 + ez
z 1 1 1 −1
exp( ), (ez − 1)−1 exp( ), exp(tan ), sin(cos ) .
1−z 1−z z z
2.13 Exercises | 97
50. Let a ∈ ℂ, R > 0 and f ∈ ℋ(D′R (a)) and suppose that a is an essential singularity
of f . Let g be a non-constant entire function.
(i) Show that the closure of g(ℂ) equals to ℂ.
(ii) Prove that a is an essential singularity of g ∘ f .
51. Let a ∈ ℂ, R > 0 and f ∈ ℋ(D′R (a)) such that ℜf (z) ≥ 0 for each z ∈ D′R (a).
(i) Show that a is not an essential singularity of f .
(ii) Prove that f can in fact be extended to a holomorphic function on DR (a).
52. Expand the following functions as power series around z0 :
2z + 1 1
ez , z0 = πi; , z0 = 0; , z0 = −i
(z 2 + 1)(z + 1)2 (z − i)3
z
54. Let f (z) = z/(e − 1). Expand f as a power series around z0 = 0, and set
ak k
∞
f (z) = ∑ z .
k=0 k!
n+1 n+1
0 = a0 + ( )a1 + ⋯ + ( )an .
1 n
Use the fact that f (z) + 21 z is an even function in order to show that ak = 0 for k
odd and k > 1.
55. Let an , n ∈ ℕ0 be as in Exercise 54. The numbers B2n = (−1)n−1 a2n , n ≥ 1, are called
the Bernoulli numbers. Compute B2 , B4 , … , B10 .
n
56. Let f (z) = ∑∞n=0 an z be a function in ℋ(D1 (0)) such that
|f (z)| ≤ A + B|z|k ,
59. Let a ∈ ℂ, R > 0 and f ∈ ℋ(D′R (a)). Suppose that a is a pole of f . Let g be a tran-
scendental entire function. Show that a is an essential singularity of g ∘ f .
60. Let f ∈ ℋ(DR (0)) be non-constant. Show that the function r ↦ M(r) =
sup|z|=r |f (z)| is strictly increasing for r ∈ (0, R).
61. Let f be an entire function, α a zero of f and z ∈ ℂ. Show that
for all z ∈ ℂ.
62. Let f be a holomorphic on Dr1 ,r2 (0) = {z ∶ r1 < |z| < r2 }, suppose that f is continuous
on Dr1 ,r2 (0), let Mk = sup|z|=rk |f (z)|, k = 1, 2, and M(r) = sup|z|=r |f (z)|. Show that
ℂ∗ ⧵ {z ∶ z = et(1+i) , t ∈ ℝ};
1 1
(3) G = {(x, y) ∶ 0 < x < 1, 0 < y < 1} ⧵ ⋃∞ n=1 {(x, y) ∶ x = n , 0 < y ≤ 2 };
1
(4) {z ∶ |z| < 1} ⧵ { n ∶ n = 2, 3, …} ⧵ {0}?
65. Find a suitable domain G and a path in G, which is null-homologous but not null-
homotopic in G.
66. Determine the Laurent expansion of the function
(z 2 − 1)
f (z) =
(z + 2)(z + 3)
in
2.13 Exercises | 99
1
f (z) = , 0 < |a| < |b|,
(z − a)(z − b)
in
(1) {z ∶ |a| < |z| < |b|};
(2) {z ∶ |z| > b}.
68. Determine the Laurent expansion of the function
1/2
z
f (z) = [ ] , ℑf (3/2) > 0
(z − 1)(z − 2)
z 2 3 ez−1 eiπz
at , − ; at 0; at 2;
(2 − 3z)(4z + 3) 3 4 ez − 1 16 − z 4
sin z π cos2 z π z+1
at ; at 2π; z tan z at ; at 2i.
1 − 2 cos z 3 (2π − z)3 2 (z 2 + 4)2
70. Let G be a domain in ℂ, which is symmetric with respect to the real axis. Let f ∈
ℳ(G) and suppose that f is real-valued on the real axis. Show that
71. Compare the residue of the function f at a simple pole z = a ≠ 0 with the residue
of the function zf (z 2 ) at the point z = a1/2 .
72. Suppose that the function f has an isolated residue at ∞. The residue of f at ∞ is
defined by
z2 + 3 2z − 3
f (z) = z n , n ∈ ℤ; g(z) = ; h(z) = .
5z 4 − 7z 2 + 6z z2
100 | 2 Cauchy’s Theorem and Cauchy’s formula
75. How many zeros does the function f (z) = z 8 − 4z 5 + z 2 − 1 have in D = {z ∶ |z| < 1}?
76. How many zeros does the function g(z) = 2iz 2 + sin z have in the rectangle R =
{(x, y) ∶ |x| ≤ π/2, |y| ≤ 1}?
77. Prove the following theorem (Hurwitz’11 Theorem): Let G be a domain in ℂ and
(fn )n a sequence of holomorphic functions on G without zeros in G, which con-
verges uniformly on all compact subsets of G to f ∈ ℋ(G). Then f ≡ 0 or f has no
zeros in G. Hint: use Rouché’s Theorem.
What are the properties of the sequence fn (z) = ez /n with respect to this theorem?
78. Let G be a domain in ℂ and let f ∈ ℋ(G) be the limit of a sequence fn ∈ ℋ(G), uni-
formly on all compact subsets of G. Show that the zeros of f are limits of sequences
of zeros of the functions fn .
Find an example of a limit point of zeros of the functions fn at the boundary of G
which is not necessarily a zero of f .
79. Let R(x, y) be a rational function of two variables such that
R(cos t, sin t)
Compute the following definite integrals using the formula from above:
π dt π/2 dt
∫ , a > 1; ∫ ;
0 a + cos t 0 1 + sin2 t
2π sin2 t 2π cos2 2t
∫ dt, a ∈ ℝ; ∫ dt, −1 < a < 1.
0 1 − 2a cos t + a2 0 1 − 2a cos t + a2
80. Let R(z) be a rational function having no poles on ℝ and suppose that the degree
of the denominator is larger than the degree of the numerator. Then
+∞
∫ R(x)eix dx = 2πi ∑ Res(R(ζ )eiζ ; z).
−∞ ℑz>0
Hint: choose positive r1 , r2 , s so large that all poles of R in the upper half-space lie
in the rectangle [r2 , r2 + is, −r1 + is, −r1 , r2 ] and use the Residue Theorem for inte-
gration along the boundary of this rectangle. Finally, take the limits as r1 , r2 → ∞.
Compute the following definite integrals using the formula from above:
∞ cos x ∞ cos ax
∫ dx, a > 0; ∫ dx, a, b > 0.
0 a2 + x 2 0 (x 2 + b2 )2
81. Compute
∞ eitx
∫ dx, t ∈ ℝ.
−∞ 1 + x2
Let γ be a closed path in G without double points and such that the points z1 , … , zn
belong to the interior of γ.
Prove that
1 f (ζ ) ωn (ζ ) − ωn (z)
ℒn−1 (z) = ∫ dζ , z ∉ γ⋆
2πi γ ωn (ζ ) ζ −z
(Lagrange12 interpolation).
Put Rn (z) = f (z) − ℒn−1 (z) and show that
1 f (ζ ) ωn (z)
Rn (z) = ∫ dζ .
2πi γ ζ − z ωn (ζ )
(Newton13 interpolation).
2.14 Notes
Further details on the topics of this chapter can be found in [26, 11, 60, 61]. A thorough
treatment of the applications of the Residue Theorem is given in [56]. For more details
on Lagrange and Newton interpolation methods the reader should consult [53].
3 Analytic continuation
We called a function f holomorphic at a point z0 if it is complex differentiable in an
open neighborhood of z0 . We can choose an open disc DR (z0 ) for this neighborhood
and can ask whether for another point z1 ∈ DR (z0 ) we obtain a disc DR1 (z1 ) where f is
also complex differentiable. We will see that the disc DR1 (z1 ) may protrude beyond the
original disc DR (z0 ). This process yields an extension of the domain of holomorphy
of f . In this chapter we shall describe it in more detail.
{g(z), z ∈ Dr (z0 ),
h(z) = {
f (z), z ∈ DR (a)
{
Example 3.2. Let f (z) = 1/(1 − z). Then f ∈ ℋ(D1 (0)). We consider the boundary point
−1 and expand f around the point −1:
1 1 1 1
∞
= = = ∑ 2−n−1 (z + 1)n .
1 − z 2 − (z + 1) 2 1 − (z + 1)/2 n=0
The radius of convergence of this power series is 2. Hence f ∈ ℋ(D2 (−1)) and the point
−1 is regular for f .
The point 1 is obviously singular for f .
Remark. It is easily seen that the set of regular boundary points for f is an open sub-
set of the boundary which can be empty. The set of singular boundary points for f is
closed.
https://doi.org/10.1515/9783110417241-003
104 | 3 Analytic continuation
The radius of convergence of this power series is 1, hence f ∈ ℋ(D1 (0)). Now we con-
sider the 2m th roots of unity {exp(2πik/2m ) ∶ k, m ∈ ℕ}. They form a dense set in 𝕋 =
𝜕D1 (0) since for arbitrary α ∈ [0, 1] and given ϵ > 0 there exist k, m ∈ ℕ such that |α −
k/2m | < ϵ.
In addition, we have for 0 < r < 1,
m−1
m n m
)2n n
f (re2πik/2 ) = ∑ r 2 e(2πik/2 + ∑ r2 ,
n=0 n≥m
hence
m
lim f (re2πik/2 )
r→1
m
is unbounded and e2πik/2 is a singular point for f . Since the set of all 2m th roots of
unity is dense in 𝕋 and the set of all singular points is closed, all points of 𝕋 must be
singular for f .
n
Theorem 3.4. Let f (z) = ∑∞ n=0 an z be a power series with radius of convergence 1. Then
f has at least one singular point on 𝕋.
Proof. Suppose that all points in 𝕋 are regular for f . Since 𝕋 is compact, there exist
finitely many disks D1 , … , DN with centers on 𝕋 such that
(1) 𝕋 ⊆ ⋃Nk=1 Dk ,
(2) ∃ gk ∈ ℋ(Dk ) with gk = f on D1 (0) ∩ Dk , k = 1, … , N.
If Dk ∩ Dj ≠ ∅, set Vk,j = Dk ∩ Dj ∩ D1 (0). Then Vk,j ≠ ∅, since the line connecting the
centers of Dj and Dk passes through Dk ∩ Dj and belongs to D1 (0).
On Vk,j we have gk = f = gj . Now define
N
Ω = D1 (0) ∪ ⋃ Dk
k=1
and
{f (z), z ∈ D1 (0),
h(z) = {
g (z), z ∈ Dk .
{ k
n
Corollary 3.5. Let f (z) = ∑∞ n=0 an (z − z0 ) be a power series with radius of convergence
R > 0. Let ζ ∈ 𝜕DR (z0 ) and let z1 be a point on the line connecting z0 and ζ , suppose that
z1 ≠ z0 , z1 ≠ ζ . Let Δ = R − |z1 − z0 |.
3.2 Analytic continuation along a curve | 105
(a) If
1/n −1
f (n) (z1 )
Δ = [lim sup | | ] ,
n→∞ n!
1/n −1
f (n) (z1 )
Δ < [lim sup | | ] ,
n→∞ n!
1/n −1
f (n) (z1 )
ρ = [lim sup | | ]
n→∞ n!
Definition 3.7. Let γ ∶ [0, 1] ⟶ ℂ be a curve connecting z0 and z1 , i.e. γ(0) = z0 and
γ(1) = z1 , let ϵ, η > 0 and f ∈ ℋ(Dϵ (z0 )) as well as g ∈ ℋ(Dη (z1 )).
106 | 3 Analytic continuation
Figure 3.1
Remark. (a) The function f (z) = 1/(1 − z) has no analytic continuation along a curve
which passes through the point 1.
(b) If f has an analytic continuation along γ, the result g is uniquely determined by f
and γ, it is independent of the subdivision and of the functions fν and the neigh-
borhoods Uν . This follows from the identity principle of holomorphic functions,
Theorem 2.18.
(c) But the result g depends in general on the route of γ and not only on the endpoint
of γ, see Example 3.6.
(d) In stead of the neighborhoods Uν one can also choose discs Kν such that the center
of Kν+1 lies in Kν , for ν = 1, … , n − 1. The functions fν can be taken as power series
around γ(tν ).
Example 3.8. (a) We are looking for a branch of the logarithm which has the value
log 2 + (2k + 1)πi at the point z1 = −2, where k ∈ ℕ.
For this purpose we start with the principle branch of the logarithm in a suitable
neighborhood of z0 = 1. We choose a path starting at z0 which circles around the
origin k times in the positive direction and ends at the point z1 = −2 coming from
3.3 The Monodromy Theorem | 107
the upper half-plane (compare with Example 3.6). In this way we get the desired
branch of the logarithm.
(b) Let G ⊆ ℂ be a domain, f ∈ ℋ(G). If V ⊆ G is a convex subset, then f has a primitive
F ∈ ℋ(V) on V , i.e. F ′ = f on V .
Let γ be an arbitrary curve in G from z0 to z1 . We choose a suitable subdivision
of the parameter interval and discs Dν as associated neighborhoods. Since all Dν
are convex, there exist primitive functions Fν of f on Dν (see Corollary 2.6). By
the identity principle (see Theorem 2.18), we have Fν = Fν+1 on Dν ∩ Dν+1 for ν =
1, … , n − 1. Hence, local primitive functions have an analytic continuation along
each curve in G.
Definition 3.9. Let U ⊆ ℂ be an open set and let γ0 , γ1 denote two curves in U from
z0 ∈ U to z1 ∈ U.
γ0 is homotopic to γ1 in U, if there exists a continuous function
H ∶ [0, 1] × [0, 1] ⟶ U
Figure 3.2
Example 3.11. Let U = ℂ. Consider the curves γ0 and γ1 from z0 to z1 in Figure 3.2.
Each homotopy from γ0 to γ1 passes through the origin. The logarithm has no analytic
continuation to a neighborhood of the origin. The assumptions of the monodromy the-
orem are not satisfied for any branch of the logarithm. Analytic continuation along
γ0 and along γ1 yields a different result for each branch of the logarithm (see Exam-
ple 3.6).
Proof. (a) Suppose that the continuation of f along γs0 yields gs0 . First we show the
following: if s is sufficiently close to s0 , the continuation along γs also yields gs0 . For
the continuation along γs0 we choose the subdivision points tν , corresponding neigh-
borhoods Uν and holomorphic functions fν ∈ ℋ(Uν ), ν = 1, … , n.
Let 𝒰 = ⋃nν=1 Uν . Then 𝒰 is open. Since H is continuous, H −1 (𝒰) is open in [0, 1] ×
[0, 1]. 𝒰 contains γs∗0 . Hence [0, 1] × {s0 } ⊂ H −1 (𝒰). Since H −1 (𝒰) is open, there exists
3.3 The Monodromy Theorem | 109
By (a), the set J is open [0, 1] and 0 ∈ J. Suppose that J ≠ [0, 1]. Then sup J = s0 ∉ J,
otherwise we would get a contradiction to the openness of J. Apply part (a) for this s0 .
Now we get a contradiction to the assertion that s0 is the supremum of J. Hence we
obtain J = [0, 1].
Theorem 3.12. Let G be a simply connected domain in ℂ. Let z0 ∈ G, f ∈ ℋ(Dϵ (z0 )).
Suppose that f has an analytic continuation along each curve in G with starting point
z0 . Then there exists F ∈ ℋ(G) with f = F on Dϵ (z0 ).
Proof. In a simply connected domain, each two curves γ0 and γ1 with the same start
and end points are homotopic (see Definition 2.52). γ = γ0 − γ1 is a closed curve, which
is homotopic to a constant curve. Let H be the corresponding homotopy. Using H one
can easily construct a homotopy H0 from γ0 to γ1 (see Exercises).
Now let z ∈ G and let γz be a curve in G from z0 to z. Let gz denote the result of the
analytic continuation of f along γz . The function gz is holomorphic in a neighborhood
U(z) ⊂ G of z. By Theorem 3.10, gz is independent of γz .
Let F(w) = gz (w) for w ∈ U(z). In order to show that F ∈ ℋ(G), it suffices to show
the following assertion: if z1 ≠ z2 and U(z1 ) ∩ U(z2 ) ≠ ∅, then gz1 = gz2 on U(z1 ) ∩ U(z2 ).
For this purpose let z∗ ∈ U(z1 ) ∩ U(z2 ). Then we obtain gz∗ by analytic continua-
tion along γz1 + [z1 , z∗ ], and also by analytic continuation of gz1 along the straight
line [z1 , z∗ ] which lies in U(z1 ). By the identity principle, we obtain that gz1 = gz∗
on U(z1 ) ∩ U(z∗ ). But we also get that gz∗ = gz2 on U(z∗ ) ∩ U(z2 ). Hence gz1 = gz2 on
U(z1 ) ∩ U(z2 ).
Remark. By the last theorem one can construct a global holomorphic function start-
ing with a locally defined holomorphic function (a germ of a holomorphic function)
and using analytic continuation.
If G is a simply connected domain which does not contain the origin, one can
extend the logarithm to a globally defined holomorphic function F ∈ ℋ(G).
If G is an annulus around the origin, there exists no global holomorphic function
as an extension of the logarithm.
110 | 3 Analytic continuation
3.4 Exercises
84. Let
∞
f (z) = ∑ ak z k
k=0
be a power series with radius of convergence 1, and where all coefficients ak are
non-negative real numbers. Prove that z = 1 is a singular point for f .
Hint: use Corollary 3.5.
85. Let f be the restriction of the principle branch of the function √z to the disc D1 (1).
Determine the analytic continuations of f along the curves γ(t) = exp(2πit) and
σ(t) = exp(4πit) for 0 ≤ t ≤ 1.
86. Let G be a simply connected domain in ℂ. Show that each two curves with com-
mon start and end points are homotopic in G.
87. Let Ω be a domain in ℂ and choose a fixed w ∉ Ω. Let D be a disc in Ω. Show that
there exists a function f ∈ ℋ(D) such that exp[f (z)] = z − w and f ′ (z) = (z − w)−1 in
D. Prove that f ′ has an analytic continuation along each curve in Ω, which starts
at the center of D, and show that it is necessary to assume simple connectivity in
Theorem 3.12.
3.5 Notes
The fact that analytic continuation along a closed path can lead to a different function
than the one started with, leads to the concept of a Riemann surface, see [15]. The more
algebraic aspects of the topic are treated in [46].
4 Construction and approximation of holomorphic
functions
In this chapter we discuss the question in which domains G ⊆ ℂ holomorphic func-
tions can be approximated by polynomials or by functions which are holomorphic in
a larger domain than G (Runge’s Theorem, Section 4.4). We use the Hahn–Banach The-
orem (Section 4.3) to prove a preferably general assertion which will be helpful later
on. In addition, we construct holomorphic functions with a prescribed zero set and
given multiplicities (Weierstraß’ Factorization Theorem, Section 4.6) and we prove the
existence of a meromorphic function with prescribed principle parts (Mittag-Leffler’s
Theorem, Section 4.5).
For all of these problems the solution of the inhomogeneous Cauchy–Riemann
equation plays an important role (Section 4.2). First 𝒞∞ -functions with the desired
properties are constructed, which are finally corrected by suitable solutions of the
inhomogeneous Cauchy–Riemann equation to become holomorphic functions. This
method has its origin in the theory of holomorphic functions of several variables,
which will be explained in more detail in the second part of this book.
This chapter ends with the Riemann Mapping Theorem showing the existence of
biholomorphic mappings between simply connected domains G ≠ ℂ and the unit disc
D1 (0) (Section 4.9) and with a characterization of simply connected domains in terms
of properties of holomorphic function on these domains (Section 4.10).
supp(Φi ) = {x ∈ ℝn ∶ Φi (x) ≠ 0} ⊆ Ui ∀i ∈ I;
(2) {supp(Φi ) ∶ i ∈ I} is locally finite, i.e. the set {i ∈ I ∶ supp(Φi ) ∩ K ≠ ∅} is finite for
each compact subset K ⊂ Ω.
(3) ∑i∈I Φi ≡ 1 on Ω (by (2) this sum is finite ∀x ∈ Ω).
Lemma 4.2. Let Ω ⊆ ℝn be open and let K ⊂ Ω be a compact subset. Then there exists
a function Φ ∈ 𝒞∞
0 (Ω) (Φ has compact support) with Φ(x) > 0 ∀x ∈ K.
https://doi.org/10.1515/9783110417241-004
112 | 4 Construction and approximation of holomorphic functions
Proof. Let
1
{exp(− 1−t ), t < 1,
ψ(t) = {
0, t ≥ 1.
{
Then ψ ∈ 𝒞∞ (ℝ). Now we set
K ⊂ ⋃ Va .
a∈K
Lemma 4.3. Let Ω ⊆ ℝn be open and let 𝒰 = {Ui ∶ i ∈ I} be an open cover of Ω. Then
there exists a 𝒞∞ -partition of unity with respect to 𝒰.
Proof. Ω ⊆ ℝn is paracompact (see [19]), i.e. 𝒰 has a locally finite refinement, this is a
locally finite open cover {Vj ∶ j ∈ J} of Ω with the property that ∀j ∈ J ∃i ∈ I with Vj ⊆ Ui .
Hence there exists a mapping τ ∶ J ⟶ I such that Vj ⊆ Uτ(j) .
In addition, there exist compact subsets Kj ⊂ Vj such that ⋃j∈J Kj = Ω (see [19]).
By Lemma 4.2, there exist ψj ∈ 𝒞∞ 0 (Vj ) such that ψj (x) > 0 ∀x ∈ Kj . Now we set
ψ = ∑ ψj .
j∈J
Since the cover {Vj ∶ j ∈ J} is locally finite, the sum from above is always a finite sum
and ψ ∈ 𝒞∞ . We also have ψ > 0 on Ω, since ψj > 0 on Kj and ⋃j∈J Kj = Ω.
Let χj = ψj /ψ. The family {χj ∶ j ∈ J} is a 𝒞∞ -partition of unity with respect to {Vj ∶
j ∈ J} because we have
ψj 1
∑ χj = ∑ = ∑ ψ = 1.
j∈J j∈J ψ ψ j∈J j
Theorem 4.4. Let Ω ⊆ ℝn be an open set, let U be an open subset of Ω and X a closed
subset of U, i.e. X ⊂ U ⊂ Ω. Then there exists Φ ∈ 𝒞∞ (Ω) with 0 ≤ Φ ≤ 1 such that Φ|X = 1
and Φ|Ω⧵U = 0.
Proof. Let V = Ω ⧵ X. Then V is open and {U, V} is an open cover of Ω. By Lemma 4.3,
there exists a corresponding 𝒞∞ -partition of unity, denoted by {ΦU , ΦV }. Then we have
ΦU + ΦV = 1 on Ω and supp ΦU ⊆ U, supp ΦV ⊆ V .
Now we set Φ = ΦU . Then Φ = 0 outside of U, and since ΦV = 0 outside of V , we
must have ΦU = 1 on X.
Theorem 4.5. Let Ω ⊆ ℝn be open and let X1 and X2 be two disjoint closed subsets of
Ω. Let Φ1 , Φ2 ∈ 𝒞∞ (Ω). Then there exists Φ ∈ 𝒞∞ (Ω) such that
Proof. By Theorem 4.4, there exists α ∈ 𝒞∞ (Ω) with 0 ≤ α ≤ 1 such that α|X1 = 1 and
α|X2 = 0.
Now set Φ = αΦ1 + (1 − α)Φ2 . Then Φ has the desired properties.
𝜕u
= f.
𝜕z
In the second part of this book we will reconsider this problem using functional anal-
ysis tools such as unbounded operators on Hilbert spaces.
Theorem 4.6. Let G be a bounded domain in ℂ with a piecewise smooth boundary and
let f be a function in a neighborhood U of G, which has continuous first real derivatives.
Then
1 f (ζ )
u(z) = ∫ dζ ∧ dζ , z∈G
2πi G ζ − z
𝜕u
= f,
𝜕z
Remark. For the proof we will use the inhomogeneous Cauchy formula 2.47. Suppose
that u is continuously differentiable on G and a solution of 𝜕u
𝜕z
= f , then we get from 2.47
1 u(ζ ) 1 f (ζ )
u(z) = ∫ dζ + ∫ dζ ∧ dζ ,
2πi 𝜕G ζ − z 2πi G ζ − z
where the boundary of G has positive orientation. The first term is a holomorphic func-
tion at z ∈ G. Differentiation with respect to z gives
𝜕u 1 𝜕 u(ζ ) 1 𝜕 f (ζ )
= ∫ ( ) dζ + [∫ dζ ∧ dζ ]
𝜕z 2πi 𝜕G 𝜕z ζ − z 2πi 𝜕z G ζ − z
1 𝜕 f (ζ )
= [∫ dζ ∧ dζ ].
2πi 𝜕z G ζ − z
1 f (ζ )
u(z) = ∫ dζ ∧ dζ .
2πi G ζ − z
1 (1 − ϕ(ζ ))f (ζ )
u2 (z) = ∫ dζ ∧ dζ ,
2πi G⧵Dr (z0 ) ζ −z
1 ϕ(ζ )f (ζ )
u1 (z) = ∫ dζ ∧ dζ .
2πi ℂ ζ − z
4.2 The inhomogeneous Cauchy–Riemann differential equations | 115
𝜕u 𝜕u 1 𝜕 ϕ(ζ )f (ζ )
(z) = 1 (z) = [∫ dζ ∧ dζ ]
𝜕z 𝜕z 2πi 𝜕z G ζ − z
1 𝜕 ϕ(z + w)f (z + w)
= [∫ dw ∧ dw]
2πi 𝜕z ℂ w
1 𝜕 ϕ(z + w)f (z + w)
= ∫ [ ] dw ∧ dw.
2πi ℂ 𝜕z w
For a differentiable function h, we can write
𝜕h 𝜕h
(z + w) = (ζ ),
𝜕z 𝜕ζ
where ζ = z + w. Now set z = x + iy and ζ = ξ + iη, then
𝜕h h(z + w + t) − h(z + w)
(z + w) = lim
𝜕x t∈ℝ,t→0 t
and so
𝜕h h(ζ + t) − h(ζ )
(ζ ) = lim ;
𝜕ξ t∈ℝ,t→0 t
analogous expressions are obtained for the derivative with respect to η.
Now we can continue the computation of 𝜕u𝜕z
(z) and get
𝜕u 1 𝜕 ϕ(z + w)f (z + w)
(z) = ∫ [ ] dw ∧ dw
𝜕z 2πi ℂ 𝜕z w
1 𝜕/𝜕ζ [ϕ(ζ )f (ζ )]
= ∫ dζ ∧ dζ . (4.1)
2πi ℂ ζ −z
We use Theorem 2.47 for ϕf and for a disc D with D2r (z0 ) ⊂⊂ D. Since ϕf = 0 on 𝜕D, we
obtain for z ∈ Dr/2 (z0 )
1 f (ζ )
u(z) = ∫ dζ ∧ dζ ,
2πi ℂ ζ − z
1 𝜕f /𝜕ζ (ζ )
f (z) = ∫ dζ ∧ dζ on ℂ.
2πi ℂ ζ − z
Theorem 4.8. Let G ⊆ ℂ be open and f ∈ ℋ(G), let Kbe a compact subset of G. Let
α ∈ 𝒞∞
0 (G) be such that α = 1 on K (see Theorem 4.4). Then
1 𝜕α f (ζ )
f (z) = ∫ dζ ∧ dζ ∀z ∈ K.
2πi G 𝜕ζ ζ − z
Proof. Let
{f (z)α(z), z ∈ G,
Φ(z) = {
0, z ∉ G.
{
Then Φ ∈ 𝒞∞ 0 (ℂ). By Lemma 2.59, there exists an open subset G1 ⊂ G such that
supp(α) ⊂ G1 and 𝜕G1 is piecewise smooth. We apply Theorem 2.47 and get
1 Φ(ζ ) 1 𝜕Φ/𝜕ζ
Φ(z) = ∫ dζ + ∫ dζ ∧ dζ .
2πi 𝜕G1 ζ − z 2πi G1 ζ − z
Now
𝜕Φ 𝜕f 𝜕α 𝜕α
= α + f = f,
𝜕ζ 𝜕ζ 𝜕ζ 𝜕ζ
1 Φ(ζ ) 1 𝜕α f (ζ )
f (z) = Φ(z) = ∫ dζ + ∫ dζ ∧ dζ
2πi 𝜕G1 ζ − z 2πi G1 𝜕ζ ζ − z
1 𝜕α f (ζ )
= ∫ dζ ∧ dζ ,
2πi G 𝜕ζ ζ − z
Theorem 4.9 (Hahn1 –Banach2 Theorem). Let X be a vector space over ℝ with a semi-
norm ‖ ⋅ ‖. Let Y be a subspace of X. Suppose that l ∶ Y ⟶ ℝ is a linear functional on Y
such that
l(x) ≤ C‖x‖ ∀x ∈ Y,
L(x) ≤ C‖x‖ ∀x ∈ X.
Proof. Let x0 ∉ Y and Z = ⟨Y, x0 ⟩ be the linear span of Y and x0 . If x ∈ Z, then x = y +cx0 ,
y ∈ Y , c ∈ ℝ, and the representation of x is uniquely determined.
We set ϕ(x) = l(y) + ca0 for a0 ∈ ℝ. Then ϕ is an extension of l on Z. Now we show
that a0 can be chosen such that ϕ(x) ≤ C‖x‖ ∀x ∈ Z, i.e.
y y
a0 ≤ C‖ + x0 ‖ − l( );
c c
y y
a0 ≥ −C‖− − x0 ‖ − l( ).
c c
which implies
hence we can find a0 ∈ ℝ with (4.3). Therefore we have found a linear extension ϕ of l
to Z such that
ϕ(x) ≤ C‖x‖ ∀x ∈ Z.
Now let ℰ be the set of all linear extensions λ of l, which are dominated by C‖ ⋅ ‖ on its
domain of definition dom(λ).
We define a partial order on ℰ: for λ1 , λ2 ∈ ℰ let
U ∶ ⋃ dom(λ) ⟶ ℝ
λ∈ℱ
Zorn’s3 Lemma. Let ℰ be a non-empty partially ordered set, in which each totally or-
dered subset has an upper bound. Then ℰ contains at least one maximal element.
We indicate that Zorn’s Lemma is equivalent to the Axiom of Choice (see [63]).
The assumptions of Zorn’s Lemma are satisfied in our case. Hence there exists a
maximal element L in ℰ.
If dom(L) = M ≠ X, then there exists x0 ∈ X ⧵M and an extension of L to ⟨M, x0 ⟩ with
the required properties and we arrive at a contradiction to the maximality of L.
Theorem 4.10. Let X be a vector space over ℂ with a seminorm ‖⋅‖. Let Y be a subspace
of X. Let l ∶ Y ⟶ ℂ be a linear functional on Y with
|l(x)| ≤ C‖x‖ ∀x ∈ Y,
|L(x)| ≤ C‖x‖ ∀x ∈ X.
Proof. Let l1 (x) = ℜl(x). l is now complex linear. We have il(x) = l(ix) = l1 (ix) + iℑl(ix)
and, on the other hand, il(x) = i(l1 (x) + iℑl(x)) = −ℑl(x) + il1 (x), hence ℑl(x) = −l1 (ix).
l1 is a real linear functional on Y and the assumption implies
By Theorem 4.9, there exists a real linear extension L1 of l1 to the whole of X with
|L1 (x)| ≤ C‖x‖ ∀x ∈ X.
Inspired by the computation at the beginning of the proof, we set now
therefore L is an extension of l.
Since we have L(x) = |L(x)|eiθ for some θ ∈ ℝ, we obtain finally
Theorem 4.11. Let (X, ‖ ⋅ ‖) be a normed vector space over ℂ, x0 ∈ X and Y a subspace
of X. Then x0 belongs to Y , if and only if there is no continuous linear functional L on X
such that L(x) = 0 ∀x ∈ Y but L(x0 ) ≠ 0.
If x0 ∉ Y , then there exists δ > 0 such that ‖x − x0 ‖ > δ ∀x ∈ Y . Let Z = ⟨Y, x0 ⟩ and
define l(x + cx0 ) = c for x ∈ Y and c ∈ ℂ. Then we have
hence
Corollary 4.12. Let (X, ‖ ⋅ ‖) be a normed vector space over ℂ and Y a subspace of X.
The subspace Y is dense in X, if and only if each continuous linear functional on X, which
vanishes on Y , also vanishes on the whole of X.
1 g(ζ )
f (z) = ∫ dζ , z ∉ γ∗ ,
2πi γ ζ − z
n−1
1 g(γ(tj ))
R(z) = ∑ (γ(tj+1 ) − γ(tj )).
2πi j=0 γ(tj ) − z
Theorem 4.14. Let Ω ⊆ ℂ be open and K ⊂ Ω a compact subset. Then each function
f ∈ ℋ(Ω) can be uniformly approximated on K by rational functions that have their
poles at points of Ω ⧵ K.
Proof. By Lemma 2.59, there exists a cycle Γ in Ω such that IndΓ (z) = 1, for each z ∈ K
and IndΓ (z) = 0 for all z ∉ Ω. Using this cycle and the homology-version of Cauchy’s
Theorem (see Theorem 2.50), we have
1 f (ζ )
f (z) = ∫ dζ ,
2πi Γ ζ − z
for all z ∈ K. Let ϵ > 0. From Theorem 4.13, we get a rational function R with poles in
Γ ∗ such that |f − R|K < ϵ.
Corollary 4.16. Let K ⊂ ℂ be a compact set. Then the following assertions are equiva-
lent:
(a) Each function being holomorphic in a neighborhood of K can be uniformly approx-
imated on K by polynomials;
(b) ℂ ⧵ K is connected.
Proof of Theorem 4.15. (c) ⟹ (b). Suppose that (b) is not valid. Then there exists a
connected component O of Ω ⧵ K such that O is compact and O ⊂ Ω. We claim that
𝜕O ⊂ K. In order to show this, suppose that there exists a point a ∈ 𝜕O with a ∉ K.
Then, since O ⊂ Ω, we have a ∉ 𝜕Ω. Hence a belongs to the open set Ω ⧵ K, and there
exists r > 0 such that Dr (a) ⊂ Ω ⧵ K. Since a ∈ O, we get Dr (a) ∩ O ≠ ∅ and the set
Dr (a) ∪ O ⊂ Ω ⧵ K is connected, and therefore a larger connected subset of Ω ⧵ K than O.
We arrive at a contradiction, hence 𝜕O ⊂ K.
Now we use the maximum principle (see Corollary 2.31). Let f ∈ ℋ(Ω). Then
and by (4.5)
By Theorem 2.34, there exists F ∈ ℋ(O), being continuous on O, such that fn → F uni-
formly on O. Using (4.6), we obtain
F(z)(z − ζ ) = 1 ∀z ∈ O,
ϕ(ζ1 ) − ϕ(ζ2 ) 1 1
= [L(fζ1 ) − L(fζ2 )] = L(f − f ). (4.7)
ζ1 − ζ2 ζ1 − ζ2 ζ1 − ζ2 ζ1 ζ2
We get
1 1 ζ1 − ζ2
fζ1 (z) − fζ2 (z) = − = .
z − ζ1 z − ζ2 (z − ζ1 )(z − ζ2 )
For Fζ1 ,ζ2 (z) = 1/[(z − ζ1 )(z − ζ2 )], we obtain from (4.7) that
ϕ(ζ1 ) − ϕ(ζ2 ) 1
= L((ζ1 − ζ2 )Fζ1 ,ζ2 ) = L(Fζ1 ,ζ2 ),
ζ1 − ζ2 ζ1 − ζ2
124 | 4 Construction and approximation of holomorphic functions
where we used that L is linear. Now we let ζ1 → ζ2 and get that Fζ1 ,ζ2 → Fζ2 ,ζ2 uniformly
on K, and we have Fζ2 ,ζ2 (z) = 1/(z − ζ2 )2 . Since L is continuous, we obtain
ϕ(ζ1 ) − ϕ(ζ2 )
lim = L(Fζ2 ,ζ2 ),
ζ1 →ζ2 ζ1 − ζ2
hence ϕ ∈ ℋ(ℂ ⧵ K). In a similar way one can show that the higher derivatives of ϕ
can be expressed as
1
ϕ(k) (ζ ) = k!L(z ↦ ).
(z − ζ )k+1
ϕ(k) (ζ ) = 0 ∀k ∈ ℕ,
1
∞
= ∑ a zn
z − ζ n=0 n
1 𝜕ψ g(ζ )
g(z) = ∫ (ζ ) dζ ∧ dζ ∀z ∈ K,
2πi ω 𝜕ζ ζ −z
1 𝜕ψ g(ζ )
= ∫ (ζ ) dζ ∧ dζ ,
2πi ω⧵K1 𝜕ζ ζ −z
4.4 Runge’s approximation theorems | 125
since on K1 we have 𝜕ψ
= 0. No we apply the functional L to this formula. The integral
𝜕ζ
𝜕ψ g(ζ )
∫ (ζ ) dζ ∧ dζ
ω⧵K1 𝜕ζ ζ −z
𝜕ψ g(ζ )
lim ∑ (ζν ) ν m(Rν ),
ν 𝜕ζ ζ ν −z
where Rν are rectangles covering ω ⧵ K1 , the points ζν ∈ Rν and where m(Rν ) is the mea-
sure of Rν . Taking the limit means that the diameters of Rν tend to zero. Furthermore,
the limit occurs uniformly for all z ∈ K. Since L is continuous, we can interchange the
limit with the application of L and obtain
𝜕ψ g(ζ )
L(g) = L(z ↦ ∫ (ζ ) dζ ∧ dζ )
ω⧵K1 𝜕ζ ζ −z
𝜕ψ 1
=∫ (ζ )g(ζ )L(z ↦ ) dζ ∧ dζ = 0,
ω⧵K1 𝜕ζ ζ − z
K̂ Ω ∶= {z ∈ Ω ∶ |f (z)| ≤ |f |K , ∀f ∈ ℋ(Ω)}
Remark 4.18. (a) Taking the holomorphically convex hull of a compact set K, the
holes of K are plugged in a certain sense, which corresponds to the maximum
principle.
(b) We have
dist(K, Ωc ) = dist(K̂ Ω , Ωc ).
1 1 −1
≤ sup| | = ( inf |w − ζ |) ,
|z − ζ | w∈K w − ζ w∈K
dist(K, Ωc ) ≤ dist(K̂ Ω , Ωc ).
(c) A subset M ⊆ ℝ2 is called convex, if for any two points in M the whole straight line
segment between the two points also lies in M. An equivalent formulation is the
following: for any point (x0 , y0 ) ∈ ℝ2 ⧵ M there exists a straight line g = {(x, y) ∈ ℝ2 ∶
l((x, y)) = ax + by + c = 0} such that (x0 , y0 ) ∈ g and M ⊂ {(x, y) ∈ ℝ2 ∶ l((x, y)) < 0},
which is also called the separation property.
The convex hull K̃ of K is the smallest convex set, which contains K. One can show
that
αz
Since the functions z ↦ e are holomorphic on Ω, it follows that
K̂ Ω ⊆ K,̃
∞ ∞ ∞
⋃ Kn = ⋃ (Kλ(n)
∗ ̂ ⊇ ⋃ Kλ(n)
)Ω ∗
= Ω.
n=1 n=1 n=1
(g) Given M, ϵ > 0 and p ∈ Ω ⧵ K̂ Ω , there is f ∈ ℋ(Ω) with |f |K < ϵ and |f (p)| > M. To
prove this, observe that p ∈ Ω ⧵ K̂ Ω implies that there is h ∈ ℋ(Ω) with |h|K < |h(p)|.
After multiplying with a suitable constant, one may assume that |h|K < 1 < |h(p)|,
now take f = hN with N > 0 sufficiently large.
Theorem 4.19. Let Ω ⊆ ℂ be open and let K ⊂ Ω be a compact subset. Then K̂ Ω is the
union of K with all connected components of Ω ⧵ K, which are relatively compact in Ω.
K̂ Ω ⊆ (K1 ) Ω
̂ = K1 .
Remark. Conditions (a), (b) and (c) of Runge’s approximation Theorem 4.15 are sat-
isfied for K ⊂ Ω if and only if K = K̂ Ω . In this case K is called holomorphically convex.
128 | 4 Construction and approximation of holomorphic functions
Theorem 4.20. Let Ω1 and Ω2 be two open subsets in ℂ with Ω1 ⊆ Ω2 . Then the follow-
ing conditions are equivalent:
(a) Each f ∈ ℋ(Ω1 ) can be approximated by functions in ℋ(Ω2 ) uniformly on all com-
pact subsets of Ω1 , i.e. ℋ(Ω2 ) is dense in ℋ(Ω1 ).
(b) If Ω2 ⧵ Ω1 = L ∪ F, where F ⊂ Ω2 is closed, L is compact and L ∩ F = ∅, then L = ∅.
(c1 ) ∀K ⊂ Ω1 compact, one has K̂ Ω2 = K̂ Ω1 .
(c2 ) ∀K ⊂ Ω1 compact, one has K̂ Ω2 ∩ Ω1 = K̂ Ω1 .
(c3 ) ∀K ⊂ Ω1 compact, the set K̂ Ω2 ∩ Ω1 is compact.
K̂ Ω2 ∶= {z ∈ Ω2 ∶ |f (z)| ≤ |f |K ∀f ∈ ℋ(Ω2 )}
and
K̂ Ω1 ⊃ K̂ Ω2 ∩ Ω1 .
For z ∈ K̂ Ω2 ∩ Ω1 , ϵ > 0 arbitrary and f ∈ ℋ(Ω1 ), we use (a) to find g ∈ ℋ(Ω2 ) such that
|f − g|K∪{z} < ϵ. Then we have
which implies K̂ Ω1 ⊃ K̂ Ω2 ∩ Ω1 .
Since ℋ(Ω1 ) ⊇ ℋ(Ω2 ), we have K̂ Ω1 ⊆ K̂ Ω2 ∩ Ω1 , and we get (c2 ).
(c2 ) ⟹ (c3 ). This follows from Remark 4.18(e).
(c3 ) ⟹ (a). Let K ′ = K̂ Ω2 ∩ Ω1 and K ″ = K̂ Ω2 ∩ Ωc1 . Then K ′ ∪ K ″ = K̂ Ω2 and K ′ ∩
K = ∅. We have K ′ ⊂ Ω1 and K ′ ∪ K ″ ⊂ Ω2 . By (c3 ), the set K ′ is compact and, by
″
{h in a neighborhood of K ′ ,
g={
1 in a neighborhood of K ″ ,
{
4.4 Runge’s approximation theorems | 129
{0 in a neighborhood of K ′ ,
g={
1 in a neighborhood of K ″ ,
{
and, again by Theorem 4.15, there exists an approximating function ϕ ∈ ℋ(Ω2 ) such
that
K̂ Ω1 ⊆ K̂ Ω2 ∩ Ω1 ⊆ K̂ Ω2 ,
we get immediately K̂ Ω1 = K̂ Ω2 ∩ Ω1 .
So far we have shown the equivalences of (a), (c1 ), (c2 ) and (c3 ).
Now we still prove that (b) ⟺ (c1 ), then we are done.
(c1 ) ⟹ (b). For a compact set K ⊂ Ω1 , we have by assumption that K̂ Ω1 = K̂ Ω2 . Let
Ω2 ⧵ Ω1 = L ∪ F, where L is compact, F is closed in Ω2 and L ∩ F = ∅; we have to show
that L = ∅.
For this let ω ⊃ L be an open set, which is relatively compact in Ω2 and for which
ω ∩ F = ∅. We have 𝜕ω ⊂ Ω2 and
therefore ω ⊆ (𝜕ω) Ω
̂ and since L ⊂ ω, we can use condition (c1 ) for the compact set
2
𝜕ω to show that
L ⊂ (𝜕ω) Ω
̂ = (𝜕ω) Ω
2
̂ ⊆ Ω1 .
1
(b) ⟹ (c1 ). Let K ⊂ Ω1 be compact. One always has K̂ Ω1 ⊆ K̂ Ω2 . Hence we still have
to show that K̂ Ω1 ⊇ K̂ Ω2 . For this we use Theorem 4.19. Let O be a connected component
of Ω2 ⧵ K, which is relatively compact in Ω2 . We will show that O ⊆ Ω1 . Then we are
done by Theorem 4.19. From the first part of the proof of Theorem 4.15, we know that
𝜕O ⊆ K ⊂ Ω1 . Now let L = O ∩ (Ω2 ⧵ Ω1 ). Then L is compact, and we have
Corollary 4.21. Let Ω ⊆ ℂ be an open set. Then the polynomials are dense in ℋ(Ω) (i.e.
each f ∈ ℋ(Ω) can be approximated by polynomials uniformly on all compact subsets
of Ω), if and only if ℂ ⧵ Ω is connected.
Proof. Let ℂ ⧵ Ω be connected. Taylor series expansion implies that the polynomials
are dense in ℋ(ℂ). Set Ω2 = ℂ, Ω1 = Ω and use Theorem 4.20: we write Ω2 ⧵ Ω1 =
ℂ ⧵ Ω = L ∪ F, where L is compact, F is closed in ℂ and L ∩ F = ∅. We will show that
L = ∅. For this aim let F̃ = F ∪ {∞}; then we have F̃ ∩ L = ∅ and ℂ ⧵ Ω = F̃ ∪ L, where
F̃ ≠ ∅. Since both sets F̃ and L are closed, we must have L = ∅, otherwise we would get
a contradiction to the assumption that ℂ ⧵ Ω is connected. Now we apply Theorem 4.20
and get that ℋ(ℂ) is dense in ℋ(Ω).
To prove the other direction, suppose that ℂ ⧵ Ω is not connected. It suffices to
show that condition (b) from Theorem 4.20 is not satisfied. Now we can write ℂ ⧵ Ω as
the union L ∪ F,̃ where L is compact in ℂ, the set and F̃ is closed in ℂ and ∞ ∈ F,̃ L ≠ ∅
as well as L ∩ F̃ = ∅. Let F = F̃ ⧵ {∞}. Then F is closed in ℂ and ℂ ⧵ Ω = L ∪ F, where
L ∩ F = ∅. Therefore condition (b) from Theorem 4.20 is not satisfied.
Example 4.22. Let Ω = {z ∈ ℂ ∶ −1 < ℑz < 1}. Then ℂ ⧵ Ω is not connected, whereas
ℂ ⧵ Ω is connected.
be principle parts in zj .
4.5 Mittag-Leffler’s Theorem | 131
We want to construct a meromorphic function f ∈ ℳ(Ω) which has the given prin-
ciple parts in zj . Taking the sum of all principle parts ∑∞
j=1 fj raises the question of con-
vergence of this sum. We will construct terms which leave the given principle parts
unchanged, but ensure convergence. For this purpose we will use Runge’s approxi-
mation theorem.
Theorem 4.23 (Mittag-Leffler’s5 Theorem). Let Ω ⊆ ℂ be an open subset and let (zj )j
be a discrete sequence of points in Ω being different from each other. Let fj be functions
meromorphic in a neighborhood of zj , j ∈ ℕ. Then there exists a meromorphic function
f ∈ ℳ(Ω) such that f − fj is holomorphic in a neighborhood of zj , j ∈ ℕ, i.e. f − fj has a
removable singularity at zj . We call f a solution of the Mittag-Leffler problem.
converges. Since uj ∈ ℋ(Ω), the principle parts of f will then exactly be the given fj .
For this purpose we choose a compact exhaustion (Kj )j of Ω such that (Kj ) Ω ̂ = Kj ,
j ∈ ℕ (see Remark 4.18), furthermore we can assume that zk ∉ Kj ∀k ≥ j. The functions fj
are holomorphic in a neighborhood of Kj , since zj ∉ Kj , and we have (Kj ) Ω
̂ = Kj . Hence,
by Theorem 4.15, there exist uj ∈ ℋ(Ω) with
k−1 ∞
f (z) = ∑ (fj (z) − uj (z)) + ∑(fj (z) − uj (z))
j=1 j=k
where the second series converges uniformly on K. Therefore, f has the desired prop-
erties.
Example 4.24. We are looking for a meromorphic function on ℂ having the principle
parts (z − k)−2 for k ∈ ℤ.
Let f (z) = ∑k∈ℤ (z − k)−2 . If |z| ≤ R and |k| > 2R, then |z − k| ≥ |k| − |z| > |k| − R >
|k| − |k|/2 = |k|/2. Hence |z − k|−2 ≤ 4k −2 , and the series defining f converges uniformly
on compact subsets of ℂ which do not contain the poles k ∈ ℤ. The functions f (z) and
π 2 / sin2 (πz) have the same poles and principle parts. Hence f (z) − π 2 / sin2 (πz) must be
an entire function. It is easily seen that f (z + 1) = f (z). The same is true for π 2 / sin2 (πz),
so that f (z) − π 2 / sin2 (πz) is periodic with period 1. It suffices to consider the periodic
strip {z = x + iy ∶ 0 ≤ x ≤ 1, y ∈ ℝ}. In this strip we have
1 1 1
| 2
|= 2 2
≤ ,
(z − k) y + (k − x) (k − 1)2
for |k| ≥ 2 and |y| ≥ 1. The series for f (z) converges uniformly and f (z) is uniformly
bounded for |y| ≥ 1 and 0 ≤ x ≤ 1. Each summand of f (z) tends to 0 as |y| → ∞. Now
the uniform convergence implies that f (x + iy) → 0 as 0 ≤ x ≤ 1 and |y| → ∞. As we
have
we see that the function π 2 / sin2 (πz) is bounded for |y| ≥ 1 and tends to 0 as |y| → ∞.
Hence the entire function f (z)−π 2 / sin2 (πz) is bounded for 0 ≤ x ≤ 1 and |y| ≥ 1 and also
bounded on the entire vertical strip 0 ≤ x ≤ 1. By periodicity, this function is bounded
on ℂ. By Liouville’s Theorem (see Theorem 2.29) it is constant and this constant is
zero, since both f (z) and π 2 / sin2 (πz) tend to 0 as |y| → ∞. Therefore we have
π2 1
=∑ , (4.8)
sin2 (πz) k∈ℤ (z − k)2
and the series converges uniformly on compact subsets of ℂ which do not contain the
poles k ∈ ℤ. The series is called the partial fractions decomposition of π 2 / sin2 (πz).
We are looking for a meromorphic function f ∈ ℳ(ℂ) having simple poles at the
points an with residues cn , i.e. a meromorphic function with principle parts cn /(z − an )
at the points an for n = 0, 1, … .
Let ϵj > 0 such that ∑∞j=1 ϵj < ∞. The series
1 1 z k
∞
=− ∑( )
z − an an k=0 an
converges uniformly on all compact subsets of {z ∶ |z| < |an |}. Now choose ln ∈ ℕ such
that
l
1 1 n z k ϵ
| + ∑( ) | < n ,
z − an an k=0 an |cn |
Define
l
c0 1 zk
∞ n
f (z) = + ∑ cn ( + ∑ k+1 ).
z n=1 z − an k=0 an
Since the expression in the brackets has absolute value ≤ ϵn /|cn | for |z| ≤ 2n , we have
|a |
f ∈ ℳ(ℂ).
Now we choose ak = k, k ∈ ℤ and ck = 1, k ∈ ℤ. In this case it suffices to take lk = 0
because
1 1 z 2R
| + |=| |≤ 2,
z−k k k(z − k) k
if |z| ≤ R and |k| > 2R. We have then |z − k| ≥ |k| − |z| > |k| − R > |k| − |k|/2 = |k|/2 and the
estimate
1 1 2R
| ∑ ( + )| ≤ ∑ 2 < ∞,
|k|>2R z−k k |k|>2R k
1 1 1
f (z) = + ∑ ( + ) (4.9)
z k∈ℤ⧵{0} z − k k
dζ 1 1
∫ 2
= −( + ),
[0,z] (ζ − k) z−k k
134 | 4 Construction and approximation of holomorphic functions
for k ≠ 0, and
π2 1 1
∫ ( − ) dζ = −π cot(πz) + .
[0,z] sin2 (πζ ) ζ 2 z
1 1 1
π cot(πz) = + ∑ ( + ). (4.10)
z k∈ℤ⧵{0} z − k k
Theorem 4.26. Let Ω ⊂ ℂ be open and f ∈ 𝒞∞ (Ω). Then there exists u ∈ 𝒞∞ (Ω) such
that 𝜕u
𝜕z
= f (compare with Corollary 4.7).
∞
∑ ϕj = 1 on Ω.
j=1
Observe that fϕj ∈ 𝒞∞ 0 (Ω) and define fϕj = 0 on ℂ ⧵ Ω. Then fϕj ∈ 𝒞0 (ℂ), j ∈ ℕ
∞
and, by Corollary 4.7, there exist smooth functions uj ∈ 𝒞 (ℂ) such that
∞
𝜕uj
= fϕj , j ∈ ℕ.
𝜕z
converges uniformly on all compact subsets of Ω, which is also true for all derivatives
of u. Hence u ∈ 𝒞∞ (Ω) and we can interchange differentiation and summation to ob-
tain
∞ ∞ ∞
𝜕u 𝜕uj 𝜕vj
= ∑( − ) = ∑ fϕj = f ∑ ϕj = f .
𝜕z j=1 𝜕z 𝜕z j=1 j=1
Remark. Theorem 4.27 implies Theorem 4.23: if (zj )j is a discrete sequence in Ω and
if the functions fj are meromorphic in a neighborhood of zj , then we choose Ωj to be
neighborhoods of zj such that zk ∉ Ωj ∀k ≠ j. We choose Ω0 to cover the rest of Ω. In
addition, we can assume that fj ∈ ℋ(Ωj ⧵ {zj }). Then fk − fj ∈ ℋ(Ωj ∩ Ωk ). Hence, by
Theorem 4.27, there exists a meromorphic function f ∈ ℳ(Ω) such that f − fj ∈ ℋ(Ωj ),
which is the assertion of Theorem 4.23.
Theorem 4.28. Let Ω = ⋃∞ j=1 Ωj be an open cover of Ω. Let gjk ∈ ℋ(Ωj ∩ Ωk ) such that
gjk = −gkj and gjk + gkl + glj = 0 for all triples (j, k, l), where Ωj ∩ Ωk ∩ Ωl ≠ ∅. (We call (gjk )
a cocycle.) Then there exist holomorphic functions gj ∈ ℋ(Ωj ) such that gjk = gk − gj for
all pairs (j, k) where Ωj ∩ Ωk ≠ ∅. (We say that the first cohomology group disappears.)
Proof of Theorem 4.28. Let (ϕj )j be a 𝒞∞ -partition of unity with respect to the cover
(Ωj )j of Ω; one has ϕj ∈ 𝒞∞
0 (Ωj ) and ∑j=1 ϕj = 1 on Ω. For k ∈ ℕ fixed, we set
∞
hk = ∑ ϕj gjk ,
j
136 | 4 Construction and approximation of holomorphic functions
where the sum is taken over all j for which Ωk ∩ Ωj ≠ ∅. Then the cocycle conditions
on Ωk ∩ Ωl imply
by construction of the 𝒞∞ -partition of unity (see Lemma 4.3). The functions hk sat-
isfy glk = hk − hl , but they are still not holomorphic. Using the solutions of the inho-
mogeneous Cauchy–Riemann differential equations, we construct a term u such that
gk = hk + u is holomorphic on Ωk .
Since glk ∈ ℋ(Ωl ∩ Ωk ), we have
𝜕hk 𝜕hl
= on Ωl ∩ Ωk .
𝜕z 𝜕z
limn→∞ un = 0.
Lemma 4.30. Let un ≠ −1 ∀n ∈ ℕ. If one can choose a value of log(1 + un ) for each n ∈ ℕ
such that ∑∞
n=1 log(1 + un ) is convergent, then the product ∏n=1 (1 + un ) is convergent.
∞
4.6 The Weierstraß Factorization Theorem | 137
Proof. We use the functional equation and the continuity of the exponential function.
Since ∑∞
n=1 log(1 + un ) is convergent, we have
N N N
lim ∏(1 + un ) = lim ∏ exp log(1 + un ) = lim exp( ∑ log(1 + un ))
N→∞ N→∞ N→∞
n=1 n=1 n=1
∞
= exp( ∑ log(1 + un )).
n=1
Lemma 4.31. Let Log be the principle branch of the logarithm. Then ∑∞
n=1 Log(1 + un )
converges absolutely if and only if ∑n=1 un converges absolutely.
∞
|u| 3|u|
≤ |Log(1 + u)| ≤ ,
2 2
from this one gets the assertion of the lemma.
For |u| < 1 one has Log(1 + u) = u − u2 /2 + u3 /3 − ⋯, hence
Log(1 + u)
|1 − | = |1 − 1 + u/2 − u2 /3 + ⋯|
u
|u|
|u|
≤ (1 + |u| + |u|2 + ⋯) = 2 since |u| < 1.
2 1 − |u|
Log(1 + u) 1
|1 − |≤ .
u 2
Remark. Let fn ∈ ℋ(U), n ∈ ℕ. The product ∏∞ n=1 (1 + fn ) converges absolutely and
uniformly on each compact subset K of U, if for each compact subset K of U there
exists n0 ∈ ℕ such that ∑∞n=n0 Log(1 + fn ) converges absolutely and uniformly on K.
The limit function of the product is again holomorphic on U, see Lemma 4.30 and
Lemma 4.31.
Lemma 4.32. Let U ⊆ ℂ be open and suppose that the points a, b belong to one and
the same connected component of ℂ ⧵ U. Then there exists f ∈ ℋ(U) such that
z−a
exp(f (z)) = , z ∈ U,
z−b
i.e. there exists a holomorphic logarithm on z ↦ (z − a)/(z − b) on U.
Proof. Let Gk , k ∈ ℕ, denote the connected components of U. The set U is the union
of the domains Gk .
If a, b belong to one and the same connected component of ℂ ⧵ U, then a, b belong
to one and the same connected component of ℂ ⧵ Gk ∀k. We consider the function
138 | 4 Construction and approximation of holomorphic functions
z−a
g(z) = z−b on Gk . In order to show that g has a holomorphic logarithm on Gk , it suffices
to show that g ′ /g has a primitive on Gk , see Theorem 2.21. By Theorem 1.39, it suffices
to show that
g ′ (z)
∫ dz = 0,
γ g(z)
g ′ (z) z − b 1 z−a 1 1
= ( − 2
)= −
g(z) z − a z − b (z − b) z−a z−b
hence
g ′ (z) dz dz
∫ dz = ∫ −∫ = 2πi(Indγ (a) − Indγ (b)).
γ g(z) γ z − a γ −b
z
a, b belong to one and the same connected component of ℂ ⧵ Gk and therefore also to
one and the same connected component of ℂ ⧵ γ ∗ . By Theorem 2.2, the index is locally
constant, hence Indγ (a) − Indγ (b) = 0.
This gives a function fk ∈ ℋ(Gk ) such that
z−a
exp(fk (z)) = , z ∈ Gk .
z−b
Theorem 4.33 (Weierstraß Factorization Theorem). Let Ω ⊆ ℂ be open and let (zj )j be
a discrete sequence of points in Ω, being different from each other. Furthermore, let
nj ∈ ℤ, j ∈ ℕ.
Then there exists a meromorphic function f on Ω, which is holomorphic and ≠ 0 ex-
cept at the points zj , and such that for all j ∈ ℕ the function f (z)(z − zj )−nj is holomorphic
in a neighborhood of zj and does not vanish there.
This means f has a zero of order nj at the points zj , when nj > 0, and f has a pole of
order |nj | at the points zj , when nj < 0.
Remark. Before we start with the proof, we recall some properties of holomorphically
convex sets which will be important for the construction of the desired meromorphic
function. Let K ⊂ Ω be a compact set such that K̂ Ω = K. Then we have the following
assertions:
(1) ∀z ∈ 𝜕Ω there exists a neighborhood U(z) of z and a connected component U of
Ω ⧵ K such that U ∩ U(z) ≠ ∅.
4.6 The Weierstraß Factorization Theorem | 139
(2) For each connected component U of Ω⧵K there exists z0 ∈ 𝜕Ω such that U ∩U(z0 ) ≠
∅, for each neighborhood U(z0 ) of z0 .
(Assume this is not the case, then U must be relatively compact in Ω.)
We will construct holomorphic factors which do not vanish on Ω and generate con-
vergence of the infinite product.
Let (Kj )j be a compact exhaustion of Ω such that (Kj ) Ω
̂ = Kj , j ∈ ℕ.
We will choose successively rational functions fj having the prescribed zeros and
poles in Kj and we will choose functions gj ∈ ℋ(Ω) such that
fj+1
| exp(gj ) − 1| < ϵj (4.11)
fj
for finitely many zν ∈ {zj ∶ j ∈ ℕ} and mν ∈ {nj ∶ j ∈ ℕ}, be a rational function with the
prescribed zeros and poles in Kj+1 (we have only finitely many points zν ∈ Kj+1 ).
Then we can write
h(z)
= c ∏ (z − wν )mν (4.12)
fj (z) ν∈M
Then fj+1 is a rational function and has the prescribed zeros and poles in Kj+1 . By (4.12),
we get
fj+1 (z) h(z) z − wν mν
=c ∏ (z − wν′ )−mν = c ∏ ( ) .
fj (z) fj (z) ν∈M ν∈M z − wν
′
140 | 4 Construction and approximation of holomorphic functions
z − wν mν
c ∏( )
ν∈M z − wν′
in a neighborhood of Kj . We set
z − wν mν
exp(ϕν (z)) = ( ) , ν∈M
z − wν′
and write
z − wν
ϕν (z) = mν log(ν) .
z − wν′
Now we set
z − wν
log(fj+1 (z)/fj (z)) ∶= log c + ∑ mν log(ν) ,
ν∈M z − wν′
N ∞
f = lim fN+1 ∏ exp(gj ) = f1 ∏[(fj+1 /fj ) exp(gj )].
N→∞
j=1 j=1
By Lemma 4.31, the product ∏∞ j=l [(fj+1 /fj ) exp(gj )] converges uniformly on Kl to a func-
tion being holomorphic in the interior of Kl which does not vanish there. The factor
l−1
f1 ∏[(fj+1 /fj ) exp(gj )]
j=1
For this purpose we first solve the Mittag-Leffler problem for the principle parts
nj /(z − aj ) and obtain a meromorphic function h ∈ ℳ(ℂ) with poles in aj and the cor-
responding principle parts nj /(z − aj ), j ∈ ℕ. By Example 4.25, the function h can be
written as
k
1 1 j z k
∞ ∞
h(z) = ∑ nj [ + ∑ ( ) ] = ∑ hj (z),
j=1 z − aj aj k=0 aj j=1
where the indices kj are chosen in such a way that the series converges uniformly on
compact subsets which contain no poles.
Now let
k n
z j
1 z k+1 j
Let R > 0 and choose j0 so large that |aj | > R for j ≥ j0 . Then uj has no zeros in DR (0) for
j ≥ j0 . Let z ∈ DR (0) and set
u′j (ζ )
vj (z) = ∫ dζ = ∫ hj (ζ ) dζ .
[0,z] uj (ζ ) [0,z]
∑ vj = ∑ log uj .
j≥j0 j≥j0
By construction, the series ∑j≥j hj converges uniformly on DR (0) and the same is true
0
for ∑j≥j vj and also for ∑j≥j log uj .
0 0
By Lemma 4.30, we get uniform convergence of ∏j≥j uj on DR (0).
0
Hence
∞
u(z) = z n0 ∏ uj (z)
j=1
Example 4.35. We are looking for an entire function, which has simple zeroes at the
integer points. From (4.9) we get that
z
z ∏ (1 − ) ez/k
k∈ℤ⧵{0} k
142 | 4 Construction and approximation of holomorphic functions
is convergent and has the desired properties. As the function sin(πz) has the same
zeros, there exists an entire function g such that
z
sin(πz) = z ∏ (1 − ) ez/k eg(z) .
k∈ℤ⧵{0} k
Taking the logarithmic derivative and using the partial fraction decomposition of
π cot(πz) (see (4.10)), it follows that g ′ (z) = 0 and so that g is constant. Since
sin(πz)
lim = π,
z→0 z
z2
∞
sin πz = πz ∏(1 − ).
j=1 j2
z z/aj
cos πz = ∏(1 − )e ,
j∈ℤ aj
where aj = j + 1/2, j ∈ ℤ.
Theorem 4.36 (Interpolation). Let Ω ⊆ ℂ be open and let (aj )j be a discrete sequence
of points in Ω being different from each other and let (bj )j be an arbitrary sequence of
complex numbers. Then there exists a function f ∈ ℋ(Ω) such that f (aj ) = bj , j ∈ ℕ.
Proof. By Theorem 4.33, there exists a function g ∈ ℋ(Ω) such that g(aj ) = 0, j ∈ ℕ,
where all zeros have order 1, i.e. g ′ (aj ) ≠ 0. Now we solve the Mittag-Leffler problem
for the principle parts
bj 1
, j ∈ ℕ.
g ′ (aj ) z − aj
4.7 Some applications of the Mittag-Leffler and Weierstraß Theorems | 143
For the solution h to this problem, we have h ∈ ℋ(Ω ⧵ {aj ∶ j ∈ ℕ}). Now we set f = gh,
then f ∈ ℋ(Ω ⧵ {aj ∶ j ∈ ℕ}) and in the points aj we have
g(z) − g(aj ) bj
f (aj ) = lim g(z)h(z) = lim [ (z − aj )h(z)] = g ′ (aj ) = bj ,
z→aj z→aj z − aj g (a
′
j)
Next we show that a meromorphic function can globally be written as the quotient
of two holomorphic functions.
Theorem 4.37. Let Ω ⊆ ℂ be open and h ∈ ℳ(Ω) a meromorphic function. Then there
exist f , g ∈ ℋ(Ω) such that h = f /g on Ω (compare with Theorem 2.65).
Proof. Let (aj )j be the poles of h with the orders nj . By Theorem 4.33, there exists a
function g ∈ ℋ(Ω) with zeroes of order nj in the points aj . Now define f ∶= hg. Then
f ∈ ℋ(Ω) and h = f /g.
The next result is specific for the complex analysis of one variable, it has no ana-
logue in several variables. It states that each domain has holomorphic functions which
cannot be extended holomorphically to a larger domain.
Proof. First we construct a sequence (aj )j which has limit points close to each bound-
ary point of 𝜕G and then we will find a function f ∈ ℋ(G) with f ≢ 0 and f (aj ) = 0,
j ∈ ℕ. Finally, we will show that this function has the desired properties.
For this aim we construct a sequence of open discs (Dn )n having the following
properties: Dn ⊂ G; (Dn )n is locally finite (i.e. each compact subset of G has non-empty
intersection with only finitely many Dn ); ⋃∞ n=1 Dn = G; the radii rn of Dn tend to 0 as
n → ∞.
Let (Kp )p be a compact exhaustion of G. We choose the discs Dn in the following
way: K1 ⊂ D1 ∪ ⋯ ∪ Dn1 and
furthermore, Dn ∩ Kp = ∅ for n ≥ 1 + np+1 . Let the radius of Dn be < 1/p for np < n ≤ np+1 ,
p ∈ ℕ. Existence of such discs follows from the fact that the sets Kp+1 ⧵ K̊ p are compact.
Choose points aj ∈ Dj such that aj ≠ ak for j ≠ k. The sequence (aj )j has no limit
point in G since (Dn )n is locally finite. By Theorem 4.33, there exists a function f ∈
ℋ(G) such that f (aj ) = 0, j ∈ ℕ, and f ≠ 0 elsewhere.
We claim that this function f has the desired properties.
144 | 4 Construction and approximation of holomorphic functions
Suppose there exists a point a ∈ 𝜕G and a path γ ∶ [0, 1] ⟶ G with γ([0, 1)) ⊂ G
and γ(1) = a, such that there exists F ∈ ℋ(Dρ (a)), ρ > 0, which arose from analytic
continuation of f along γ (see Definition 3.7).
Then we can find ϵ > 0 such that γ(t) ∈ Dρ (a) for 1 − ϵ ≤ t ≤ 1. Let U be the con-
nected component of G ∩ Dρ (a), which contains the set {γ(t) ∶ 1 − ϵ ≤ t < 1}. We have
F|U = f |U . Denote D′ = Dρ/2 (a). By construction we have D′ ∩ U ⊆ ⋃∞
n=1 Dn , but D ∩ U ⊄
′
N
⋃n=1 Dn ∀N ∈ ℕ. Hence there exists a sequence (nk )k of natural numbers such that
Dnk ∩ (D′ ∩ U) ≠ ∅, we can even assume that the radii of the discs Dnk are all < ρ/8, for
this purpose we possibly have to take a further subsequence. Then Dnk ⊆ Dρ (a) ∩ G ∀k
and Dnk ∩ U ≠ ∅. Since U and Dnk are connected, we must have Dnk ⊆ U ∀k. Since
F|U = f |U , we also have F(ank ) = 0 ∀k. Finally, we get
Hence F has infinitely many zeroes in a relatively compact subset of Dρ (a), there exists
a limit point of the zeroes of F in Dρ (a) and, by Theorem 2.18, F ≡ 0 on Dρ (a). This
implies f ≡ 0 on U and, again by Theorem 2.18, f ≡ 0 on G, which yields a contradiction
to the construction of f .
Then G is a domain in ℂ2 with the following properties: for each f ∈ ℋ(G) there exists
F ∈ ℋ(U), U = {z = (z1 , z2 ) ∶ ‖z‖ < 1}, with F|G = f |G , see Example 6.12.
Domains in ℂn for which there exist holomorphic functions with the properties in
Theorem 4.38 are called domains of holomorphy. Theorem 4.38 says that each domain
in ℂ is a domain of holomorphy (see [59] or [51]).
Theorem 4.40 (Montel’s6 Theorem). Let Ω ⊆ ℂ be a domain. Then each bounded fam-
ily is normal.
Proof. Let (Kn )n be a compact exhaustion of Ω. We can find δn > 0 such that D2δn (z) ⊆
Kn+1 ∀z ∈ Kn , n ∈ ℕ. Let z ′ , z ″ ∈ Kn be such that |z ′ − z ″ | < δn ; furthermore, let γ(t) =
z ′ + 2δn e2πit , t ∈ [0, 1]. For arbitrary f ∈ ℱ, we have
1 f (ζ ) 1 f (ζ )
f (z ′ ) − f (z ″ ) = ∫ dζ − ∫ dζ
2πi γ ζ − z ′ 2πi γ ζ − z ″
1 (z ′ − z ″ )f (ζ )
= ∫ dζ . (4.13)
2πi γ (ζ − z ′ )(ζ − z ″ )
1 1 1 M(Kn+1 ) ′
|f (z ′ ) − f (z ″ )| ≤ 4πδn |z ′ − z ″ |M(Kn+1 ) = |z − z ″ |, (4.14)
2π 2δn δn δn
ϵδn
δ= ,
M(Kn+1 )
So far we completed the analytical part of the proof. The remainder consists of the
Arzelà–Ascoli Theorem.
Let (fm )m be a sequence in ℱ. We choose a countable subset E = {wj ∶ j ∈ ℕ} ⊂ Ω
such that E ∩ Kn is dense in Kn for each n ∈ ℕ. Since ℱ is bounded, the sequence
(fm (w1 ))m is bounded in ℂ, hence there exists a subsequence (fm,1 ) of (fm ), which is
convergent at w1 . The sequence (fm,1 (w2 )) is again bounded in ℂ and there exists a
subsequence (fm,2 ) of the sequence (fm,1 ), which is convergent at w2 . In this way we get
sequences (fm,j ), convergent at wj , which are subsequences of (fm,j−1 ) for j = 2, 3, … .
Hence the diagonal sequence (fm,m )m converges at all points of E. We claim that it
converges even uniformly on all compact subsets of Ω. For this aim it suffices to show
uniform convergence on Kn , n ∈ ℕ.
Fix Kn and take ϵ > 0. Choose δ > 0 like in the first part of the proof. Since Kn is
compact, we can find points z1 , … , zp ∈ Kn ∩ E such that Kn ⊂ ⋃pj=1 Dδ (zj ). Furthermore,
there exists N > 0 such that
|fr,r (z) − fs,s (z)| ≤ |fr,r (z) − fr,r (zj )| + |fr,r (zj ) − fs,s (zj )| + |fs,s (zj ) − fs,s (z)| < 3ϵ,
∀z ∈ Kn and r, s > N, where we used the equicontinuity in the first and third term.
Hence (fm,m )m is a uniform Cauchy sequence on Kn , which converges uniformly on
Kn by Theorem 2.34.
Lemma 4.41 (Schwarz7 Lemma). Let f ∈ ℋ(D1 (0)) and supz∈D1 (0) |f (z)| ≤ 1, as well as
f (0) = 0. Then we have:
(a) |f (z)| ≤ |z| ∀z ∈ D1 (0);
(b) |f ′ (0)| ≤ 1.
If we have equality in (a) for some z ∈ D1 (0) ⧵ {0}, or if we have equality in (b), then
f (z) = λz for some λ ∈ ℂ with |λ| = 1.
Proof. Define g(z) = f (z)/z, z ∈ D1 (0) ⧵ {0}. Since f (0) = 0, the function g has a remov-
able singularity in 0 (see Theorem 2.24), and
f (z) − f (0) f (z)
f ′ (0) = lim = lim = g(0).
z→0 z z→0 z
For each ϵ > 0 and z ∈ D1−ϵ (0), we have
f (z) 1
|g(z)| ≤ max |g(z)| = max | |≤ ,
z∈D1−ϵ (0) z∈𝜕D1−ϵ (0) z 1−ϵ
where we have used the maximum principle. Since ϵ > 0 was arbitrary, we get |g(z)| ≤
1 ∀z ∈ D1 (0), and |f (z)| ≤ |z| ∀z ∈ D1 (0). Furthermore, we have |f ′ (0)| = |g(0)| ≤ 1.
If there exists z0 ∈ D1 (0) ⧵ {0} such that |f (z0 )| = |z0 |, then |g(z0 )| = 1 and, again by
the maximum principle, we get g(z) = λ, where λ is a constant with |λ| = 1.
If |f ′ (0)| = 1, then |g(0)| = 1, and the maximum principle yields the same result as
before.
Before we prove the Riemann Mapping Theorem, we investigate holomorphic au-
tomorphisms of the unit disc.
Remark. Each holomorphic automorphism of the unit disc has the form λϕα , where
λ ∈ ℂ is a constant with |λ| = 1 and α ∈ D1 (0). This follows from the Schwarz lemma
(see also Exercises).
Proof of Theorem 4.42. ϕα is holomorphic except for a pole at the point 1/α ∉ D1 (0).
One has
z−α
1−αz
+α z − α + (1 − αz)α z − |α|2 z
ϕ−α (ϕα (z)) = z−α = = = z,
1 + α 1−αz 1 − αz + αz − αα 1 − |α|2
hence ϕα is injective on D1 (0) and has the inverse ϕ−α .
Let z0 ∈ 𝕋, z0 = eit , t ∈ ℝ. Then
eit − α eit − α
|ϕα (eit )| = | | = | | = 1,
1 − eit α e−it − α
hence ϕα (𝕋) ⊆ 𝕋 and ϕ−α (𝕋) ⊆ 𝕋, so we have ϕα (𝕋) = 𝕋.
Now it follows from the maximum principle that ϕα (D1 (0)) ⊆ D1 (0) as well as
ϕ−α (D1 (0)) ⊆ D1 (0) and therefore ϕα (D1 (0)) = D1 (0). The assertions about the deriva-
tive of ϕα are easy to check.
148 | 4 Construction and approximation of holomorphic functions
f ′ (z)
∫ dz = 0
γ f (z)
for each closed path γ in Ω. By Theorem 1.39, the function f ′ /f has a primitive on
Ω and by Theorem 2.21, the function f has a holomorphic square root ϕ on Ω, i.e.
ϕ2 (z) = z − w0 ∀z ∈ Ω.
We claim that ϕ is injective. Indeed, if ϕ(z1 ) = ϕ(z2 ), then ϕ2 (z1 ) = ϕ2 (z2 ), hence
z1 − w0 = z2 − w0 , and z1 = z2 .
ϕ is non-constant. Hence, by Theorem 2.38, ϕ is an open mapping. Furthermore,
we have ϕ ≠ 0 on Ω. Let z0 ∈ Ω and ϕ(z0 ) = a ≠ 0. There exists r > 0 with 0 < r < |a|
such that Dr (a) ⊆ ϕ(Ω).
Now we define ψ(z) = r/(ϕ(z) + a), z ∈ Ω. If ϕ(z) = −a for some z ∈ Ω, then ϕ2 (z) =
a = ϕ2 (z0 ), and we have z = z0 , which is a contradiction. So ψ ∈ ℋ(Ω) and ψ is injec-
2
r r
|ψ(z)| = < =1 ∀z ∈ Ω,
|ϕ(z) + a| r
Let ψ ∈ Σ0 and suppose that ψ(Ω) ≠ D1 (0). We claim that there exists a ψ1 ∈ Σ0
such that |ψ′1 (z0 )| > |ψ′ (z0 )|.
For this aim let ψ ∈ Σ0 and α ∈ D1 (0) such that α ∉ ψ(Ω). Take the function ϕα as
in Theorem 4.42. Then we have ϕα ∘ ψ ∈ Σ and since ϕα vanishes only at z = α, the
function ϕα ∘ ψ has no zero in Ω. As in (a) one shows that there exists a g ∈ ℋ(Ω)
such that g 2 = ϕα ∘ ψ. The function g is again injective so that g ∈ Σ. Let β = g(z0 ) and
ψ1 = ϕβ ∘ g. Then we have ψ1 ∈ Σ0 , since ψ1 (z0 ) = ϕβ (g(z0 )) = ϕβ (β) = 0. Consider the
square function s(w) = w2 . Now we have
Furthermore, we have F(D1 (0)) ⊆ D1 (0) and F(0) = F(ψ1 (z0 )) = ψ(z0 ) = 0; since the
function s is involved we know that F fails to be injective. We can apply the Schwarz
Lemma 4.41 to get that |F ′ (0)| < 1. This implies |ψ′ (z0 )| < |ψ′1 (z0 )| and we have proved
our claim.
We remark that ψ′ (z0 ) ≠ 0, since ψ is injective (see Theorem 2.42).
(c) Fix z0 ∈ Ω and let η = sup{|ψ′ (z0 )| ∶ ψ ∈ Σ0 }. We claim that there exists an h ∈
Σ0 such that η = |h′ (z0 )|. By (b), such a function h must be surjective. For this claim
we will use Montel’s Theorem (see Theorem 4.40). The family Σ0 is bounded, since
|ψ(z)| < 1 ∀z ∈ Ω ∀ψ ∈ Σ0 . Hence, by Definition 4.39, Σ0 is normal.
From the definition of η, we get a sequence (ψn )n in Σ0 such that
Hence there exists a subsequence, again denoted by (ψn )n which converges uniformly
on all compact subsets of Ω to a function h ∈ ℋ(Ω) and |h′ (z0 )| = η. We have η > 0
hence h is non-constant. So h is an open mapping, in particular, h(Ω) is open. Since
ψn (Ω) ⊆ D1 (0) ∀n ∈ ℕ, we have h(Ω) ⊆ D1 (0) and even h(Ω) ⊆ D1 (0), as h(Ω) is open.
It remains to show that h is injective. Let z1 , z2 ∈ Ω, z1 ≠ z2 .
Let α = h(z1 ) and αn = ψn (z1 ). By Theorem 2.18, there exists a disc D ⊂ Ω with center
z2 , such that z1 ∉ D and such that h − α has no zero on the boundary 𝜕D. Since ψn − αn
converges uniformly on D to h − α, we have
on 𝜕D, if n is large enough. Since all functions ψn are injective, the function ψn − αn
has no zeroes in D. By Theorem 2.71, we know that the number of zeroes of ψn − αn in
D coincides with the number of zeroes of h − α in D. Hence h − α has no zeroes in D and
h(z1 ) ≠ h(z2 ).
150 | 4 Construction and approximation of holomorphic functions
Corollary 4.44. Let Ω ⫋ ℂ be a simply connected domain, and let z0 ∈ Ω. Then there
exists a biholomorphic mapping g ∶ Ω ⟶ D1 (0) such that g(z0 ) = 0 and such that g ′ (z0 )
is real and g ′ (z0 ) > 0.
The mapping g is uniquely determined by the above properties.
Proof. The existence of g follows immediately from the proof of Theorem 4.43, by mul-
tiplication with a suitable λ ∈ ℂ, |λ| = 1 one can attain that g ′ (z0 ) > 0.
To show uniqueness, we apply the Schwarz Lemma (see Lemma 4.41): let h be
another function with the above properties. Define ϕ = h ∘ g −1 ∶ D1 (0) ⟶ D1 (0). Then
ϕ(0) = h(g −1 (0)) = h(z0 ) = 0. By Lemma 4.41, we get
Let z = g(w). Then |h(g −1 (g(w)))| ≤ |g(w)| and |h(w)| ≤ |g(w)|. Interchanging the roles of
g and h, one gets |g(w)| ≤ |h(w)|. Hence |h(w)| = |g(w)| ∀w ∈ D1 (0), therefore |ϕ(z)| = |z|
on D1 (0) and, again by Lemma 4.41, ϕ(z) = λz for some λ ∈ ℂ with |λ| = 1.
Furthermore, we have
1
ϕ′ (0) = h′ (g −1 (0))(g −1 )′ (0) = h′ (z0 ) > 0.
g ′ (z0 )
f (z) = z + a2 z 2 + a3 z 3 + ⋯ .
z−i
h(z) = eit , t∈ℝ
z+i
are biholomorphic mappings from the upper half-space onto D1 (0) such that h(i) = 0.
The proof of the Riemann mapping theorem is not constructive, which means that
it is impossible to find nice formulas in most cases. There are powerful approxima-
tion methods which have interesting applications, for instance, in fluid mechanics;
see [40]. Later we will discuss another approach using Hilbert space methods and the
Bergman kernel.
∫ f (z) dz = 0;
γ
Proof. (a) ⟹ (b). Let ϕ ∶ Ω ⟶ D1 (0) be a homeomorphism and let γ a closed curve
in Ω. Define
Then H ∶ [0, 1] × [0, 1] ⟶ Ω is continuous and one has that H(s, 0) = ϕ−1 (0), s ∈ [0, 1]
is constant and H(s, 1) = γ(s), s ∈ [0, 1]; furthermore, H(0, t) = H(1, t), t ∈ [0, 1]. Hence γ
is null-homotopic and Ω is simply connected.
(b) ⟹ (c). Theorem 2.55.
(c) ⟹ (d). Suppose that ℂ ⧵ Ω is not connected. Then there are two non-empty
closed, disjoint subsets H, K of ℂ ⧵ Ω such that H ∪ K = ℂ ⧵ Ω. If ∞ ∈ H, then K is
152 | 4 Construction and approximation of holomorphic functions
∫ pn (z) dz = 0 ∀n ∈ ℕ,
γ
which implies
∫ f (z) dz = 0.
γ
4.11 Exercises
88. Let Ω = {z ∶ |z| < 1 and |2z − 1| > 1}, and let f ∈ ℋ(Ω).
(1) Is there a sequence of polynomials Pn such that Pn → f uniformly on all com-
pact subsets of Ω?
(2) Is there such a sequence converging to f uniformly on Ω?
(3) Is the answer for (2) different, if it is assumed that f is holomorphic in an
open set containing the closure of Ω?
89. Show that there exists a sequence (pn )n of polynomials such that
{1, z ∈ ℝ,
lim pn (z) = {
n→∞ 0, z ∉ ℝ.
{
92. Let U = {z ∶ |z| < 1}, V = {z ∶ 0 < |z| < 1}, K = {z ∶ |z| ≤ 1/2} and L = {z ∶ |z| = 1/2}. De-
termine K̂ U , K̂ ℂ , L̂ U , L̂ V and L̂ ℂ . Discuss the consequences for Runge’s Theorem!
93. Prove that a domain G in ℂ is simply connected if and only if for each compact
subset K of G one has K̂ G = K̂ ℂ .
94. Let Ω be a domain in ℂ and let (ak )k∈ℕ be a discrete sequence in Ω. For each
k ∈ ℕ let Rk (z) be the principle part of a function being meromorphic in a neigh-
borhood of ak ; furthermore, let Dk be a sequence of disjoint discs with center ak
and let ϕk ∈ 𝒞∞ 0 (Dk ). Define
∞
u = ∑ ϕ k Rk .
k=1
sin z/n
∞ ∞ ∞
∏(1 − z n ); ∏ cos z/n; ∏ ;
n=1 n=1 n=1 z/n
1
∞ ∞
n
∏(1 − ); ∏(1 + z 2 )?
n=1 nz n=1
1 π2
∞
∑ 2
= .
k=1 k 6
1 1 1
f (z) = , g(z) = , h(z) = ,
sin z cos z 1 − ez
i.e. represent the functions from above as the sum of their principle parts, where
one has to take care for possible convergence generating summands in the sense
of the Mittag-Leffler Theorem.
99. Show that each holomorphic automorphism ϕ of the unit disc D1 (0) has the form
z−α
ϕ(z) = λ ,
1 − αz
1 − |f (z)|2
|f ′ (z)| ≤ , for |z| < 1
1 − |z|2
(Pick’s Lemma).
Hint: fix z0 ∈ D1 (0) and set w0 = f (z0 ). Use Lemma 4.41 for the function h ∘ f ∘ g,
where
z + z0 w − w0
g(z) = and h(w) = .
1 + z0 z 1 − w0 w
4.12 Notes
The proof of the second version of Runge’s Theorem (Theorem 4.20) and the proof of
Weierstraß’ Factorization Theorem (Theorem 4.33) is based on Hörmander’s presen-
tation [41]. We also refer to Narasimhan’s monograph [57], where the so-called real
method in complex analysis to use the inhomogeneous Cauchy–Riemann equation is
explained for functions of one complex variable. This powerful method has its origin
in the theory of several variables and will be discussed in detail in the last chapters
of the book in combination with L2 -methods. The generalization of Mittag-Leffler’s
Theorem (Theorem 4.28) is inspired by the solution of the Cousin I problem for mero-
morphic functions of several variables, where singularities are never isolated; see, for
instance, [59] or [51].
5 Harmonic functions
Harmonic functions in ℝ2 appear as real and imaginary parts of holomorphic func-
tions. Their properties are crucial for the theory of partial differential equations. Har-
monic functions and the Laplace operator describe many important phenomena in
mathematical physics, such as fluid mechanics, thermodynamics, and electricity. The
solution of the Dirichlet problem in Section 5.2 is a striking example how methods of
complex analysis can be applied for a problem in real partial differential equations.
Therefore the relationship between harmonic and holomorphic functions is of
great interest. Jensen’s formula (Theorem 5.16) stands as an example for the other di-
rection, where properties of harmonic functions are used to explain the zero distribu-
tion of holomorphic functions. Finally, we discuss the basic properties of subharmonic
functions.
Example 5.2. (a) Linear polynomials in the two real variables x and y are always
harmonic.
f (x, y) = x2 fails to be harmonic. But u(x, y) = x2 − y2 and v(x, y) = 2xy are harmonic.
If p(z) = z 2 , z = x + iy, then ℜp(x, y) = x2 − y2 and ℑp(x, y) = 2xy.
(b) If f ∈ ℋ(G) and f = u + iv, then the Cauchy–Riemann differential equations are
valid: ux = vy , uy = −vx . Differentiation with respect to x and y yields uxx = vyx ,
uyy = −vxy , hence uxx + uyy = 0. Similarly, we have vxx + vyy = 0. Hence u and v are
harmonic on G.
x
(c) Let G = ℝ2 ⧵ {(0, 0)} and let h(x, y) = 1/2 log(x2 + y2 ). We have hx = x2 +y 2 and
y2 − x2 x2 − y2
hxx = , hyy = .
(x2 + y2 )2 (x 2 + y2 )2
Hence Δh = 0 on G.
Theorem 5.4. Let G ⊆ ℝ2 be a domain. G is simply connected if and only if each har-
monic function u on G has a harmonic conjugate function v on G.
https://doi.org/10.1515/9783110417241-005
156 | 5 Harmonic functions
y y
vx (x, y) = ∫ uxx (x, t) dt − uy (x, 0) = − ∫ uyy (x, t) dt − uy (x, 0)
0 0
= −uy (x, y) + uy (x, 0) − uy (x, 0) = −uy (x, y).
Theorem 5.5 (Mean value property). Let G ⊆ ℝ2 be open and let h ∶ G ⟶ ℝ be a har-
monic function. Let a ∈ G and R > 0 such that DR (a) ⊂ G. Then
1 2π
h(a) = ∫ h(a + Reit ) dt.
2π 0
Proof. We choose R′ > R with DR′ (a) ⊂ G. By Theorem 5.4, there exists a harmonic func-
tion g on DR′ (a), which is harmonic conjugate to h. Hence the function f = h + ig ∈
ℋ(DR′ (a)).
5.2 The Dirichlet problem | 157
Now let γR (t) = a + Reit , t ∈ [0, 2π]. Then, by Cauchy’s integral formula,
1 f (ζ ) 1 2π
f (a) = ∫ dζ = ∫ f (a + Reit ) dt.
2πi γR ζ − a 2π 0
Taking real and imaginary parts, we get the desired result.
Remark. Later we will see that a certain form of the converse is also true.
Example 5.6. We consider the idealized flow of a fluid with velocity vector
Reit + (z − a)
Pa,R (z, t) = ℜ( ), t∈ℝ
Reit − (z − a)
1 2π
Pa,R (Φ)(z) = ∫ P (z, t)Φ(a + Reit ) dt, z ∈ DR (a)
2π 0 a,R
1+z
∞ ∞
= (1 + z) ∑ z n = 1 + 2 ∑ z n ,
1−z n=0 n=1
hence
1+z
∞
Pr (θ) = ℜ( ) = 1 + 2 ∑ r n cos nθ = ∑ r |n| einθ .
1−z n=1 n∈ℤ
Lemma 5.8. The Poisson kernel Pa,R (z, t) is harmonic as a function of z ∈ DR (a), where t
is fixed. Furthermore, we have Pa,R (z, t) > 0 ∀z ∈ DR (a), ∀t ∈ ℝ. For each r with 0 ≤ r < R,
one has
1 2π
∫ P (a + reiθ , t) dθ = 1.
2π 0 a,R
Reit + (z − a)
z↦
Reit − (z − a)
for z ∈ DR (a) and hence harmonic there.
Set ρ = r/R, then
1 − ρ2
Pa,R (z, t) = P(ρei(θ−t) ) = Pρ (θ − t) = .
1 − 2ρ cos(θ − t) + ρ2
1 − r2
0 < Pr (θ) ≤ ∀r, 0 ≤ r < 1.
1 − cos2 δ
In particular, Pr (θ) tends to 0 as r → 1, even uniformly in θ for δ ≤ θ ≤ 2π − δ.
1 − r2
Pr (θ) ≤ 1 − r 2 ≤ .
1 − cos2 δ
If δ ≤ θ ≤ π/2, we have 0 ≤ cos θ ≤ cos δ and
Therefore,
1 − r2
Pr (θ) ≤ .
1 − cos2 δ
If 3π/2 ≤ θ ≤ 2π − δ, we have 0 ≤ cos θ ≤ cos(2π − δ) = cos δ and hence again
1 − r2
Pr (θ) ≤ .
1 − cos2 δ
Theorem 5.10 (Solution of the Dirichlet problem). Let Φ ∶ 𝕋 ⟶ ℝ be a continuous
function. Let z ∈ D1 (0), z = reiθ ,
1 2π
u(z) = ∫ P (θ − t)Φ(eit ) dt,
2π 0 r
and let u(eit ) = Φ(eit ) on 𝕋.
Then u ∶ D1 (0) ⟶ ℝ is a solution of the Dirichlet problem with boundary values Φ.
160 | 5 Harmonic functions
Proof. First we show that u is harmonic on D1 (0). We can write u in the following form:
1 2π eit + z 1 2π eit + z
u(z) = ∫ ℜ( it )Φ(eit ) dt = ℜ(∫ it
Φ(eit ) dt).
2π 0 e −z 2π 0 e −z
The last integral is a holomorphic parameter integral with respect to z ∈ D1 (0). Hence
u is the real part of a holomorphic function and therefore harmonic on D1 (0).
By Lemma 5.8, we have
1 2π 1 2π
u(z) − Φ(eit0 ) = ∫ Pr (θ − t)Φ(eit ) dt − ∫ P (θ − t)Φ(eit0 ) dt
2π 0 2π 0 r
1 2π
= ∫ P (θ − t)(Φ(eit ) − Φ(eit0 )) dt. (5.3)
2π 0 r
|ei(t−θ) − 1| ≥ δ/2,
1 2π
u(z) − Φ(eit0 ) = ∫ P (θ − t)(Φ(eit ) − Φ(eit0 )) dt
2π 0 r
1
= (∫ P (θ − t)(Φ(eit ) − Φ(eit0 )) dt
2π |eit −eit0 |≥δ r
= I + II.
1 2π ϵ
|II| ≤ ϵ ∫ Pr (θ − t) dt = .
2π 0 2π
To estimate the first summand, we apply Lemma 5.9 and obtain δ′ ≤ (t − θ) ≤ 2π − δ′ for
some δ′ > 0, which depends only on δ (we know that |ei(t−θ) − 1| ≥ δ/2, if θ is sufficiently
close to t0 ). Then
5.3 Jensen’s formula | 161
1
|I| ≤ C(δ)(1 − r 2 ) ∫ |Φ(eit ) − Φ(eit0 )| dt
2π |eit −eit0 |≥δ
2π
≤ C ′ (δ)(1 − r 2 ) ∫ |Φ(eit ) − Φ(eit0 )| dt
0
≤ C ′ (δ)(1 − r 2 )M,
where the constant M > 0 depends only on Φ. Taking the limit as z = reiθ → eit0 , we
observe that r → 1, hence I tends to 0. So we have
Corollary 5.11. If Φ ∶ 𝜕DR (a) ⟶ ℝ is continuous, then u(z) = Pa,R (Φ)(z), z ∈ DR (a), and
u(z) = Φ(z), z ∈ 𝜕DR (a), is a solution of the Dirichlet problem on DR (a) with boundary
values Φ.
1 2π
u∗ (w) = ∫ P (w, t)g ∗ (eit ) dt, w ∈ D1 (0),
2π 0 0,1
then u(z) = u∗ (ψ(z)), z ∈ G, is harmonic on G, and for z ∈ 𝜕G we have
1 2π
u(z) = ∫ u(z + rn eit ) dt ∀n ∈ ℕ.
2π 0
Remark. By Theorem 5.5, we know that harmonic functions have the mean value
property. In the following we will prove the converse of this assertion.
Theorem 5.14. Let u ∶ Ω ⟶ ℝ be a continuous function with the mean value property.
Then u is harmonic on Ω.
Proof. Let a ∈ Ω and R > 0 with DR (a) ⊂ Ω. By Corollary 5.11, the Poisson integral
1 2π Reit + (z − a)
h(z) = ∫ ℜ( it )u(a + Reit ) dt
2π 0 Re − (z − a)
defines a function which is continuous on DR (a), harmonic on DR (a) and which coin-
cides with u on 𝜕DR (a).
Now let v = u − h and m = sup{v(z) ∶ z ∈ DR (a)}. Suppose that m > 0, then the set E =
{z ∈ DR (a) ∶ v(z) = m} is a compact subset of DR (a) since v = 0 on 𝜕DR (a). Hence there
exists a z0 ∈ E with |z0 −a| ≥ |z −a| ∀z ∈ E. If r > 0 is sufficiently small, at least half of the
disk Dr (z0 ) does not belong to E. The function v has the mean value property. But the
mean value of v over 𝜕Dr (z0 ) is smaller than v(z0 ) = m, and we have a contradiction.
Hence m = 0 and v ≤ 0 on DR (a).
The same reasoning for −v yields v ≥ 0. Hence u = h on DR (a). Therefore u is har-
monic on Ω.
Lemma 5.15.
1 2π
∫ log|1 − eiθ | dθ = 0
2π 0
Proof. Let Ω = {z ∶ ℜz < 1}. Then 1 − z ≠ 0 ∀z ∈ Ω. Since Ω is simply connected, we
get from Theorem 4.46 that there exists a holomorphic function h ∈ ℋ(Ω) such that
exp h(z) = 1 − z ∀z ∈ Ω. Setting h(0) = 0, the function h becomes uniquely determined.
We have
For δ > 0 small, let Γ be the path defined by Γ(t) = eit , δ ≤ t ≤ 2π − δ and let γ the circular
line from eiδ to e−iδ in Ω, see Figure 5.1.
Figure 5.1
1 h(z) h(z)
0 = h(0) = (∫ dz − ∫ dz).
2πi Γ z γ z
Hence
Theorem 5.16 (Jensen’s3 formula). Let f ∈ ℋ(DR (0)), f (0) ≠ 0, 0 < r < R, and let
α1 , … , αN be the zeros of f in Dr (0), repeated according to multiplicities. Then
N
r 1 2π
|f (0)| ∏ = exp{ ∫ log|f (reiθ )| dθ}.
n=1 |αn | 2π 0
Proof. Arrange the zeros αj such that α1 , … , αm ∈ Dr (0) and |αm+1 | = ⋯ = |αN | = r. Now
define
m N
r 2 − αn z αn
g(z) = f (z) ∏ ∏ ,
n=1 r(α n − z) n=m+1 n − z
α
then g ∈ ℋ(D), where D = Dr+ϵ (0) for some ϵ > 0. We can choose ϵ > 0 such that g has
no zero in D. We know from Section 5.1 that log |g| is harmonic on D, and Theorem 5.5
implies
1 2π
log|g(0)| = ∫ log|g(reiθ )| dθ. (5.4)
2π 0
In addition, we have
m
r
|g(0)| = |f (0)| ∏ . (5.5)
n=1 n |
|α
r 2 − αn z
| | = 1.
r(αn − z)
N
log|g(reiθ )| = log|f (reiθ )| − ∑ log|1 − ei(θ−θn ) |. (5.6)
n=m+1
Integrating (5.6) with respect to θ from 0 to 2π and using Lemma 5.15, we obtain
2π 2π
∫ log|g(reiθ )| dθ = ∫ log|f (reiθ )| dθ.
0 0
Inserting this into (5.4) and using (5.5), we obtain Jensen’s formula.
Corollary 5.17. Let f ∈ ℋ(DR (0)) and 0 < ρ < R, suppose that f (0) = 0 of order λ (where
λ = 0 means that f (0) ≠ 0). Let n(ρ) be the number of zeros of f in Dρ (0), where multi-
plicities are included and let
Then
ρn(ρ) M(ρ)
≤ .
|α1 | ⋯ |αn(ρ) | |f (λ) (0)/λ!|ρλ
n(ρ)
ρ 1 2π
log |f (λ) (0)/λ!| + log(ρλ ∏ )= ∫ log|f (ρeiθ )| dθ ≤ log M(ρ).
k=1 |αk | 2π 0
Remark. The last result describes an important relationship between the distribution
of the zeroes and the growth behavior of a holomorphic function, which is of special
interest for entire functions, see [11].
1 2π
v(a) ≤ ∫ v(a + Reit ) dt, (5.7)
2π 0
Proof. Let DR (z0 ) ⊂ G. We consider (5.7) taking a radius r with 0 ≤ r ≤ R. Then we mul-
tiply both sides by r and integrate with respect to r from 0 to R. So we get
R 1 R 2π
∫ rv(z0 ) dr ≤ ∫ ∫ v(z0 + reit )r dr dt.
0 2π 0 0
On the left-hand side we have R2 v(z0 )/2, whereas on the right-hand side we have polar
coordinates of the Lebesgue measure dλ in ℝ2 . Hence
1
v(z0 ) ≤ ∫ v(z) dλ(z). (5.8)
πR2 DR (z0 )
This implies
1
∫ (v(z) − v(z0 )) dλ(z) ≥ 0,
πR2 DR (z0 )
the integrand satisfies v(z) − v(z0 ) ≤ 0, as v(z0 ) = C, hence the integrand must be iden-
tically zero on DR (z0 ).
When we say that a function satisfies the maximum principle, we refer to the prop-
erty in the last theorem, that is, we suppose that it does not assume a maximum value
in G unless it is constant.
1 2π 1 2π
v(a) ≤ u(a) = ∫ u(a + Reit ) dt = ∫ v(a + Reit ) dt,
2π 0 2π 0
and v is subharmonic on G.
Immediately we get the following corollary which explains the name subhar-
monic.
Remark 5.22. For many important applications it is necessary to consider a more gen-
eral concept. A function v ∶ G ⟶ ℝ ∪ {−∞} is called subharmonic if v is upper semi-
continuous, i.e. {z ∈ G ∶ v(z) < c} is open for every c ∈ ℝ, and if for every compact set
K ⊂ G and every function h ∈ 𝒞(K) which is harmonic on the interior of K and satisfies
v ≤ h on 𝜕K it follows that v ≤ h on K. For this more general concept one also has the
maximum principle and submean value property. In addition, if {vα ∶ α ∈ A} is a family
of subharmonic functions on G such that v = supα vα is finite and upper semicontin-
uous, then v is subharmonic; if {vj ∶ j ∈ ℕ} is a decreasing sequence of subharmonic
functions on G, then v = limj→∞ vj is subharmonic.
Important examples of subharmonic functions are |f |c for c > 0 and log |f |, where
f ∈ ℋ(G) is a holomorphic function on G.
If one supposes that v ∈ 𝒞2 (G), then v is subharmonic on G if and only if Δv ≥ 0
on G. For further details, see [59].
5.5 Exercises
101. Let u be a harmonic function on a domain G ⊆ ℝ2 ≅ ℂ. Show that ux and uy are
harmonic on G and that f = ux − iuy is holomorphic on G.
102. Let u and v be real-valued harmonic functions on a domain G ⊆ ℝ2 . Under what
conditions is uv harmonic on G? Show that u2 is harmonic on G if and only if u
is constant.
103. Let u be harmonic on a domain G and suppose DR (a) ⊂⊂ G. Show that
1
u(a) = ∫ u(x, y) dλ(x, y).
πR2 DR (a)
𝜕 𝜕U 𝜕2 U
r (r ) + 2 = 0.
𝜕r 𝜕r 𝜕θ
105. Suppose that 0 ∉ G and that u depends only on |z|, i.e. u(z) = φ(|z|). Show that u
is harmonic on G if and only if u(z) = a log |z| + b for some constants a, b ∈ ℝ.
106. Let u ∶ DR (a) ⟶ ℝ be a continuous function, harmonic in DR (a). Suppose that
u ≥ 0 on DR (a). Show that for 0 ≤ r < R and all θ ∈ [0, 2π] we have
R−r R+r
u(a) ≤ u(a + reiθ ) ≤ u(a)
R+r R−r
(Harnack’s inequality).
Hint: use the integral representation in the proof of Theorem 5.14.
168 | 5 Harmonic functions
108. Let f be an entire function, let M(ρ) and n(ρ) be as in Corollary 5.17. Suppose that
f (0) = 1 and show that n(ρ) log 2 ≤ log M(2ρ).
109. Let f ∈ ℋ(G). Show that |f |c for c > 0 and log |f | are subharmonic on G.
110. Suppose that v ∈ 𝒞2 (G). Show that v is subharmonic on G, if and only if Δv ≥ 0
on G.
5.6 Notes
A thorough treatment of the relationship between the growth behavior and the distri-
bution of zeroes of entire functions can be found in several textbooks, see [53, 11, 15].
Subharmonic functions and a general treatment of the Dirichlet problem are basic
tools in potential theory, see [11, 15, 26].
6 Several complex variables
To think that the analysis of several complex variables is more or less the one variable
theory with some more indices turns out to be incorrect. Completely new phenomena
appear which will be exploited in the following. Many differences between the univari-
ate and multivariate theories originate from the Cauchy–Riemann differential equa-
tions which constitute an overdetermined system of partial differential equations for
several complex variables. We start with the basic definitions and complex differential
forms. Section 6.1 also presents the main differences between univariate and multi-
variate analysis, such as the Identity Theorem and Hartogs phenomenon. Section 6.2
provides another important example for this difference, namely in the analysis of the
inhomogeneous Cauchy–Riemann differential equations. In addition, the concept of
the tangential Cauchy–Riemann equation is introduced. This gives the tools required
for the famous Lewy example of a partial differential operator without a solution. In
Section 6.3 we discuss pseudoconvex domains and plurisubharmonic functions and
explain the concept of a domain of holomorphy.
𝜕 1 𝜕 𝜕 𝜕 1 𝜕 𝜕
= ( − i ), = ( +i )
𝜕zj 2 𝜕xj 𝜕yj 𝜕z j 2 𝜕xj 𝜕yj
ω= aJ,K dzJ ∧ dz K ,
′
∑
|J|=p,|K|=q
https://doi.org/10.1515/9783110417241-006
170 | 6 Several complex variables
where ∑ denotes the sum taken only over all increasing multiindices J =
′
|J|=p,|K|=q
(j1 , … , jp ), K = (k1 , … , kq ) and
We call ω a (p, q)-form and write ω ∈ 𝒞k(p,q) (Ω), if ω is a (p, q)-form with coefficients
belonging to 𝒞k (Ω).
The derivative dω of ω is defined by
and we set
In the sequel we will use the symbol bΩ for the boundary of a domain Ω in ℂn . The
symbol 𝜕 is now reserved for differential forms.
Lemma 6.3. let ρ1 and ρ2 be two local defining functions of Ω of class 𝒞k in a neighbor-
hood U of p ∈ bΩ. Then there exists a positive 𝒞k−1 function h on U such that ρ1 = hρ2
on U and dρ1 (x) = h(x)dρ2 (x) for x ∈ U ∩ bΩ.
Proof. Since dρ2 ≠ 0 on the boundary near p, we may assume that p = 0, xn = ρ2 (x) and
U ∩ bΩ = {x ∈ U ∶ xn = 0}, after a 𝒞k change of coordinates. Let x ′ = (x1 , … , xn−1 ). Then
we have ρ1 (x′ , 0) = 0 and by the fundamental theorem of calculus
1 𝜕ρ1 ′
ρ1 (x ′ , xn ) = ρ1 (x′ , xn ) − ρ1 (x′ , 0) = xn ∫ (x , txn ) dt.
0 𝜕xn
Hence ρ1 = hρ2 for some 𝒞k−1 function on U. If k − 1 ≥ 1, we get dρ1 (x) = h(x)dρ2 (x) for
x ∈ U ∩ bΩ, as ρ2 (x) = 0 for x ∈ U ∩ bΩ. If k = 1, we get the same conclusion from the
fact that for a function f differentiable at 0 ∈ ℝn such that f (0) = 0 and for a function
h continuous at 0, one has that f ⋅ h is differentiable at 0 and d(hf )0 = h(0)df0 .
Finally, as dρ1 (x) ≠ 0 and dρ2 (x) ≠ 0 for x ∈ U ∩ bΩ, we get h(x) ≠ 0 for x ∈ U ∩ bΩ.
In addition, since h > 0 on U ⧵ bΩ, and h is continuous, we obtain h > 0 on U.
𝜕f
(z) = 0 for 1 ≤ j ≤ n and z ∈ Ω, (6.1)
𝜕z j
equivalently, if f satisfies 𝜕f = 0.
Theorem 6.5. Let P = P(a, r) be a polydisc in ℂn . Suppose that f ∈ 𝒞1 (P) and that f is
holomorphic on P, i.e. for each z ∈ P and 1 ≤ j ≤ n, the function
1 f (ζ )
f (z) = n
∫ ⋯∫ dζ1 ⋯ dζn , (6.2)
(2πi) γ1 γn (ζ1 − z1 ) ⋯ (ζn − zn )
Proof. We use induction over n. For n = 1, one has the classical Cauchy Formula, see
Theorem 2.47. Suppose that the theorem has been proven for n − 1 variables. For z ∈ P
fixed, we apply the inductive hypothesis with respect to (z2 , … , zn ) and obtain
1 f (z1 , ζ2 , … , ζn )
f (z1 , z2 , … , zn ) = ∫ ⋯∫ dζ2 ⋯ dζn .
(2πi)n−1 γ2 γn 2 − z2 ) ⋯ (ζn − zn )
(ζ
1 f (ζ1 , … , ζn )
f (z1 , ζ2 , … , ζn ) = ∫ dζ1
2πi γ1 ζ1 − z1
Like in the case n = 1, we get also here that multivariate holomorphic functions are
𝒞∞ functions, and all complex derivatives of holomorphic functions are again holo-
morphic, by differentiating under the integral sign in (6.2).
In addition, we get the Cauchy estimates: for f ∈ ℋ(P(a, r)) and α = (α1 , … , αn ) ∈
α α
ℕ0 ∶ let |α| = α1 + ⋯ + αn and α! = α1 ! ⋯ αn !; furthermore, set r α = r1 1 ⋯ rn n , then
n
𝜕|α| f α!
|Dα f (a)| = | α α (a)| ≤ α sup{|f (z)| ∶ z ∈ P(a, r)}. (6.3)
𝜕z1 1 ⋯ 𝜕zn n r
Theorem 6.6. Let f ∈ ℋ(P(a, r)). Then the Taylor series of f at a converges to f uni-
formly on all compact subsets of P(a, r), that is,
Dα f (a)
f (z) = ∑ (z − a)α , (6.4)
α∈ℕn0 α!
Proof. Use the same method as in the proof of Theorem 2.13 for each of iterated inte-
grals in (6.2).
From this we get the following: let Ω ⊆ ℂn be a domain and f ∈ ℋ(Ω), suppose that
there is a ∈ Ω such that Dα (a) = 0 for all α ∈ ℕn0 , then f (z) = 0 for z ∈ Ω. In particular,
if there is a nonempty open set U ⊂ Ω such that f (z) = 0 for z ∈ U, then f ≡ 0 on Ω
(Identity Theorem).
But Theorem 2.18 is not valid for n > 1 ∶ let f (z1 , z2 ) = z1 . Then this function is zero
on {(0, z2 ) ∶ z2 ∈ ℂ}, but f is not identically zero.
The following result is also an easy consequence of the corresponding one vari-
able result.
Theorem 6.7. Let Ω be a domain in ℂn and suppose that f ∈ ℋ(Ω) is not constant.
Then f is an open mapping.
Proof. We refer to Theorem 2.38. It is enough to show that for any ball B(a, r) = {z ∈ ℂn ∶
∑nj=1 |zj − aj |2 < r 2 } the image f (B(a, r)) is a neighborhood of f (a). The restriction of f to
B(a, r) is not constant, otherwise f would have to be constant on Ω. Choose p ∈ B(a, r)
with f (p) ≠ f (a) and define g(ζ ) = f (a + ζp) for ζ ∈ D1 (0). Then g is nonconstant and
holomorphic on D1 (0). By Theorem 2.38, g(D1 (0)) contains a neighborhood of g(0). As
g(0) = f (a) and g(D1 (0)) ⊂ f (B(a, r)), the image f (B(a, r)) is a neighborhood of f (a).
Theorem 6.8 (Hartogs2 ). Let n ≥ 2 and suppose that 0 < rj < 1 for j = 1, … , n. Then every
function f holomorphic on the domain
H(r) = {z ∈ ℂn ∶ |zj | < 1 for j < n, rn < |zn | < 1}
∪ {z ∈ ℂn ∶ |zj | < rj for j < n, |zn | < 1},
see Figure 6.1, has a unique holomorphic extension f ̃ to the polydisc P(0, 1).
Proof. The extension is unique because of the Identity theorem. Fix δ with rn < δ < 1.
Then we define
1 f (z ′ , ζ )
f ̃(z ′ , zn ) = ∫ dζ , (6.5)
2πi γδ ζ − zn
where z ′ = (z1 , … , zn−1 ) and γδ (t) = δeit for t ∈ [0, 2π]. In this way we defined a function
holomorphic on the polydisc P(0, (1′ , δ)), where (1′ , δ) = (1, … , 1, δ). For z ′ ∈ P(0, r ′ ) the
function f (z ′ , ⋅) is holomorphic on |zn | < 1, hence (6.5) implies that f ̃(z ′ , zn ) = f (z ′ , zn )
for (z ′ , zn ) ∈ P(0, (r ′ , δ)). The Identity Theorem implies f ̃ = f on H(r) ∩ P(0, (1′ , δ)), so f ̃
is the desired extension of f to the polydisc P(0, 1).
The reason for this phenomenon can be better understood by studying the inho-
mogeneous Cauchy–Riemann differential equations in several complex variables (CR
equations).
𝜕f = g, (6.6)
6.2 The inhomogeneous CR equations | 175
in other words,
𝜕f
= gj , j = 1, … , n. (6.7)
𝜕z j
Theorem 6.9. Let n ≥ 2 and let g = ∑nj=1 gj dz j be a (0, 1)-form with coefficients gj ∈
𝒞k0 (ℂn ), j = 1, … , n, where 1 ≤ k ≤ ∞ and suppose that 𝜕g = 0. Then there exists f ∈
𝒞k0 (ℂn ) such that 𝜕f = g.
We shall see that this result enables us to explain the Hartogs phenomenon in a
rather general setting.
For n = 1 the above theorem is false:
Suppose that ∫ℂ g(ζ ) dλ(ζ ) ≠ 0 and that there is a compactly supported solution f
of the equation 𝜕z
𝜕f
= g. Then there exists R > 0 such that f (ζ ) = 0 for |ζ | ≥ R. Applying
Stokes’ Theorem (see Example 2.45), we obtain for γ(t) = Reit , t ∈ [0, 2π],
0 = ∫ f (ζ ) dζ
γ
𝜕f
=∫ dζ ∧ dζ
DR (0) 𝜕ζ
= 2i ∫ g(ζ ) dλ(ζ )
DR (0)
≠ 0,
1 g (ζ , z2 , … , zn )
f (z1 , … , zn ) = ∫ 1 dζ ∧ dζ . (6.8)
2πi ℂ ζ − z1
By Corollary 4.7 (a), f ∈ 𝒞k (ℂn ) and = g1 . Now let k > 1. By hypothesis we have
𝜕f 𝜕g1
𝜕z 1 𝜕z k
=
. Since g1 has compact support, we can interchange differentiation and integration
𝜕gk
𝜕z 1
176 | 6 Several complex variables
𝜕f 1 𝜕g1 1
= ∫ (ζ , z2 , … , zn ) dζ ∧ dζ
𝜕z k 2πi ℂ 𝜕z k ζ − z1
1 𝜕gk 1
= ∫ (ζ , z2 , … , zn ) dζ ∧ dζ
2πi ℂ 𝜕ζ ζ − z1
= gk (z1 , … , zn ),
Now we are able to describe the Hartogs phenomenon in a more general way.
Theorem 6.10. Let Ω be a bounded open set in ℂn such that ℂn ⧵ Ω is connected. Sup-
pose that n > 1. Let U be an open neighborhood of the boundary bΩ = Ω ⧵ Ω. Then there
exists an open set V with bΩ ⊂ V ⊂ U having the following property: if f ∈ ℋ(U), then
there exists F ∈ ℋ(Ω) such that for the restriction to V ∩ Ω one has
f |V∩Ω = F|V∩Ω .
{αf on U ∩ Ω,
g={
0 on Ω ⧵ U.
{
Since α = 0 in a neighborhood of bU, we have g ∈ 𝒞∞ (Ω). Next, define
{ 𝜕g on Ω,
ϕk = { 𝜕z k
0 on ℂn ⧵ Ω.
{
Since 𝜕z
𝜕g 𝜕f
= 𝜕z on W ∩ Ω, we have ϕk ∈ 𝒞∞ (ℂn ). Furthermore ϕk = 0 on (ℂn ⧵ Ω) ∪ W ,
k k
n
which implies that supp(ϕk ) ⊂ Ω and ϕk ∈ 𝒞∞ 0 (ℂ ).
n
Next we claim that the (0, 1)-form ϕ = ∑j=1 ϕj dz j satisfies 𝜕ϕ = 0. We have to show
that
𝜕ϕj 𝜕ϕk
= ,
𝜕z k 𝜕z j
6.2 The inhomogeneous CR equations | 177
𝜕ϕj 𝜕2 g 𝜕ϕk
= = .
𝜕z k 𝜕z j 𝜕z k 𝜕z j
n
By Theorem 6.9, there exists u ∈ 𝒞∞
0 (ℂ ) such that 𝜕u = ϕ. Now we set F = g − u. We
have
𝜕F 𝜕g
= − ϕk = 0
𝜕z k 𝜕z k
on W , we have 𝜕z𝜕u
= 0 on Ω0 . Hence u ∈ ℋ(Ω0 ). And since supp(u) is compact and
k
Ω is bounded, ℂn ⧵ Ω must intersect ℂn ⧵ supp(u). In particular, the open set Ω1 =
Ω0 ∩ (ℂn ⧵ supp(u)) ≠ ∅, and u|Ω1 = 0. Since u ∈ ℋ(Ω0 ), the Identity Theorem implies
that u = 0 on Ω0 .
Corollary 6.11. Let Ω be a bounded open set in ℂn such that ℂn ⧵ Ω is connected. Sup-
pose that n > 1. Let U be an open neighborhood of the boundary bΩ = Ω ⧵ Ω. Further-
more, suppose that U ∩ Ω is connected. If f ∈ ℋ(U), then there exists G ∈ ℋ(Ω ∪ U)
such that G|U = f .
Example 6.12. Let |z|2 ∶= |z1 |2 + ⋯ + |zn |2 for z ∈ ℂn . Let Ω = {z ∈ ℂn ∶ |z| < 1} and let
U = {z ∈ ℂn ∶ 1/2 < |z| < 3/2}. Then each f ∈ ℋ(U) has a unique holomorphic extension
to Ω ∪ U = {z ∈ ℂn ∶ |z| < 3/2}, see Figure 6.2 in absolute space.
Theorem 6.13. Let Ω be a bounded open set in ℂn , n > 1. Suppose that ℂn ⧵ Ω is con-
nected and bΩ ∈ 𝒞4 , i.e. there exists a real-valued defining function ρ ∈ 𝒞4 (ℂn ) such that
ρ vanishes precisely on bΩ and dρ ≠ 0 on bΩ. If u ∈ 𝒞4 (Ω) and 𝜕u ∧ 𝜕ρ = 0 on bΩ, one
can then find a function U ∈ 𝒞1 (Ω) such that U ∈ ℋ(Ω) and U = u on bΩ.
Figure 6.2
Proof of Theorem 6.13. First we construct U0 ∈ 𝒞2 (Ω) such that U0 = u on bΩ and 𝜕U0 =
ρ2 v where v is a (0, 1)-form with 𝒞1 coefficients on bΩ. First we claim that 𝜕u = h0 𝜕ρ +
ρh1 , where h0 ∈ 𝒞3 (Ω) and h1 ∈ 𝒞2(0,1) (Ω). For this aim we consider the coefficients of
the (0, 1)-form 𝜕u. Using the assumption that 𝜕u ∧ 𝜕ρ = 0 on bΩ, we see that there exists
h0 ∈ 𝒞3 (Ω) such that
𝜕u 𝜕ρ
− h0 =0 on bΩ, j = 1, … , n.
𝜕z j 𝜕z j
From the proof of Lemma 6.3, we get that there exist h1,j ∈ 𝒞2 (Ω), j = 1, … , n such that
𝜕u 𝜕ρ
− h0 = ρh1,j , j = 1, … , n.
𝜕z j 𝜕z j
Now define the (0, 1)-form h1 = ∑nj=1 h1,j dz j . Then 𝜕u = h0 𝜕ρ + ρh1 . Next we get 𝜕(u −
2
h0 ρ) = ρ(h1 − 𝜕h0 ) = ρh2 , where h2 ∈ 𝒞2(0,1) (Ω). Since 0 = 𝜕 (u − h0 ρ) = 𝜕(ρh2 ) = 𝜕ρ ∧ h2 +
ρ𝜕h2 , we have 𝜕ρ ∧ h2 = 0 on bΩ. As in the first part of the proof, we can again write
h2 = h3 𝜕ρ + ρh4 ,
where h3 ∈ 𝒞2 (Ω) and h4 ∈ 𝒞1(0,1) (Ω). Now set U0 = u − h0 ρ − h3 ρ2 /2. An easy computa-
tion shows that
{𝜕U on Ω,
f ={ 0
0 on ℂn ⧵ Ω.
{
Since f = ρ2 v on bΩ we have f ∈ 𝒞1(0,1) (ℂn ) and f has compact support. By Theorem 6.9
we can find a function V ∈ 𝒞10 (ℂn ) with compact support such that 𝜕V = f . The defi-
nition of f implies that V is holomorphic in the connected set ℂn ⧵ Ω and, as V has
compact support, that V = 0 on ℂn ⧵ Ω. The function U = U0 − V is therefore equal to
U0 = u on bΩ, and 𝜕U = 𝜕U0 − 𝜕V = f − f = 0 in Ω.
The tangential Cauchy–Riemann equations for the Siegel upper half-space 𝕌 are
of special interest. Let n = 2. The function ρ(z1 , z2 ) = − 2i1 (z2 − z 2 ) + z1 z 1 is a defining
function for b𝕌. The boundary can be identified with ℍ2 = ℂ × ℝ via the mapping π ∶
(z1 , t + i|z1 |2 ) ↦ (z1 , t), where z2 = t + is. We call ℍ2 the Heisenberg group, see Exercises.
If 𝜕u ∧ 𝜕ρ = 0 on b𝕌, we have for a function u ∈ 𝒞1 (𝕌)
1 𝜕u 𝜕u
− z1 =0
2i 𝜕z 1 𝜕z 2
𝜕 𝜕 𝜕
L= + i − 2i(x + iy)
𝜕x 𝜕y 𝜕t
on ℍ2 . This operator has a special property giving a partial differential operator with-
out solution.
Theorem 6.15 (H. Lewy3 ). Let f be a continuous real-valued function depending only
on t. If there is a 𝒞1 -function u on (x, y, t) ∈ ℍ2 satisfying Lu = f in some neighborhood
of the origin, then f is analytic at t = 0, i.e. can be expanded into a convergent Taylor
series in a neighborhood of t = 0.
So if one takes a continuous function f which is not analytic at 0, the partial dif-
ferential equation Lu = f has no solution.
Proof. Suppose Lu = f in the set where x2 + y2 < R2 and |t| < R, R > 0. Let γ(θ) = reiθ ,
θ ∈ [0, 2π], 0 < r < R. Consider the line integral
2π
V(r, t) = ∫ u(x, y, t) dz = ir ∫ u(r cos θ, r sin θ, t)eiθ dθ.
γ 0
𝜕U 𝜕U
+i = 0,
𝜕t 𝜕s
which is the Cauchy–Riemann equation. Hence U is a holomorphic function of
w = t + is for 0 < s < R2 , and |t| < R, in addition, U is continuous up to the line
s = 0, and V = 0 when s = 0, therefore U(t, 0) = πF(t) is real-valued. We can apply
the Schwarz’ reflection principle (see Exercise 47 (b)): the definition U(t, −s) = U(t, s)
gives a holomorphic continuation of U to a full neighborhood of the origin. In partic-
ular, U(t, 0) = πF(t) is analytic in t, hence so is f = F ′ .
Proof. Let p ∈ bΩ. The convexity implies that we can find an ℝ-linear function
l ∶ ℂn ⟶ ℝ such that the hyperplane {z ∈ ℂn ∶ l(z) = l(p)} separates Ω and p, i.e.
we may assume that l(z) < l(p) for z ∈ Ω. We can write
n n
l(z) = ∑ αj zj + ∑ βj z j ,
j=1 j=1
1
fp (z) ∶=
h(z) − h(p)
Remark. A domain in ℂ is always holomorphically convex (see Remark 4.18 (e)). The
situation is different in higher dimensions. Let Ω = {z ∈ ℂn ∶ 1/2 < |z| < 2} and K = {z ∈
ℂn ∶ |z| = 1}. Then K̂ Ω = K, if n = 1, but if n > 1, Corollary 6.11 implies that every f ∈
182 | 6 Several complex variables
ℋ(Ω) extends to a holomorphic function f ̃ on B(0, 2). It follows from the maximum
principle applied to f ̃ that for 1/2 < |z| ≤ 1, one has
Proof. Since Ω is holomorphically convex, one can use Remark 4.18 (f).
and
j−1
|fj (pj )| > j + 1 + ∑ |fm (pj )|, j = 2, 3, … . (6.10)
m=1
Then (6.9) implies that ∑m>j |fm (pj )| < ∑m>j 2−m ≤ 1, and hence that |f (pj )| > j.
It is now easy to show that a domain Ω is holomorphically convex, if and only if for
every sequence (pj )j in Ω without limit point in Ω there is f ∈ ℋ(Ω) with supj |f (pj )| =
∞. In addition, one can now use Lemma 6.17 to show that every convex domain in ℂn
is holomorphically convex (see Exercises).
6.3 Domains of holomorphy | 183
Figure 6.3
Proof. Let Ω be like in Definition 6.21. If K ⊂ Ω is compact, then rj ∶= |fj |K < 1, for j =
1, … , k. It follows that
Lemma 6.25. Let (Km )m be a compact exhaustion of the domain Ω in ℂn . Then there
are a subsequence (mj ) of ℕ and a sequence (pj )j of points in Ω such that
(a) pj ∈ Kmj+1 ⧵ Kmj , for j = 1, 2, … , and
(b) for every p ∈ bΩ and every connected neighborhood U of p, each component Ω1 of
U ∩ Ω contains infinitely many points from (pj )j .
Proof. Let (ak )k be an enumeration of the points of Ω with rational coordinates. Let
rk = dist(ak , bΩ). Then the balls Bk = B(ak , rk ) are contained in Ω. Let (Qj )j be a se-
quence of such balls Bk which contains each Bk infinitely many times; for example,
the sequence B1 , B1 , B2 , B1 , B2 , B3 , B1 , … . Now take Km1 = K1 and use induction: assume
that l > 1 and p1 , … , pl−1 and Km1 , … , Kml have been chosen so that (a) holds for j =
1, … , l − 1. Since Ql is not contained in any compact subset of Ω, we may choose pl ∈
Ql ⧵ Kml and ml+1 such that pl ∈ Kml+1 . Then (a) holds for all j = 1, 2, … . We claim that
the points (pj )j satisfy (b): given Ω1 as in (b), there is a point q ∈ bΩ1 ∩ (U ∩ bΩ), see
Lemma 6.24. Hence there is aν ∈ Ω1 with rational coordinates sufficiently close to q so
that Bν ⊂ Ω1 . Since Bν occurs infinitely many times in the sequence (Qj )j , and pj ∈ Qj
for j = 1, 2, … , the ball Bν contains infinitely many points of the sequence (pj )j , and we
are done.
for all t ∈ ℂn , t ≠ 0.
r(p + t) = r(p) + 2ℜ(𝜕rp (t) + Qp (r; t)) + i𝜕𝜕r(t, t)(p) + o(|t|2 ), (6.11)
where t = (t1 , … , tn ) ∈ ℂn ,
n
𝜕r
𝜕rp (t) = ∑ (p)tj , (6.12)
j=1 𝜕zj
n
1 𝜕2 r
Qp (r; t) = ∑ (p)tj tk . (6.13)
2 j,k=1 𝜕zj 𝜕zk
The simplest example of a strictly pseudoconvex domain is a ball B(p, R), the function
r(z) = |z − p|2 − R2 is strictly plurisubharmonic, and B(p, R) = {z ∈ ℂn ∶ r(z) < 0}.
In the following we shall show that a strictly pseudoconvex domain is (at least)
locally a domain of holomorphy.
n n
𝜕r 1 𝜕2 r
F (r) (ζ , z) = ∑ (ζ )(ζj − zj ) − ∑ (ζ )(ζj − zj )(ζk − zk ) (6.14)
j=1 𝜕ζj 2 j,k=1 𝜕ζj 𝜕ζk
Proof. From (6.11) with p = ζ ∈ U and t = z − ζ , we obtain the Taylor expansion of r(z)
at ζ :
Theorem 6.31. Let Ω be a strictly pseudoconvex domain. Then every point p ∈ bΩ has
a neighborhood V such that V ∩ Ω is a (weak) domain of holomorphy.
1
fζ (z) ∶=
F (r) (ζ , z)
for all
n
t ∈ Tp1,0 (bΩ) = {t = (t1 , … , tn ) ∈ ℂn ∶ ∑ tj (𝜕r/𝜕zj )(p) = 0},
j=1
where Tp1,0 (bΩ) is the space of type (1, 0) vector fields which are tangent to the bound-
ary at the point p.
The domain Ω is said to be strictly Levi pseudoconvex at p, if the Levi form is
strictly positive for all such t ≠ 0. Ω is called a Levi pseudoconvex domain if Ω is Levi
pseudoconvex at every boundary point of Ω.
A bounded domain Ω in ℂn is pseudoconvex if Ω has a 𝒞2 strictly plurisubhar-
monic exhaustion function φ ∶ Ω ⟶ ℝ, i.e. the sets {z ∈ Ω ∶ φ(z) < c} are relatively
compact in Ω, for every c ∈ ℝ. (Here there is no assumption on the boundary of Ω.)
It turns out that for bounded domains with 𝒞2 boundary the concepts of (strictly)
Levi pseudoconvex and (strictly) pseudoconvex domains coincide. Furthermore, the
following assertion holds:
Let Ω be a domain in ℂn . The following are equivalent:
(1) Ω is pseudoconvex.
(2) The equation 𝜕u = f always has a solution u ∈ 𝒞∞ (p,q) (Ω) for any form f ∈ 𝒞(p,q+1) (Ω)
∞
with 𝜕f = 0, q = 0, 1, … , n − 1.
(3) Ω is a domain of holomorphy.
The proof is beyond the scope of this book. The most difficult part is the solution of
the Levi problem, to prove that a pseudoconvex domain is a domain of holomorphy,
see [41, 59, 51].
6.4 Exercises
111. Show that the Cayley transform Φ(z1 , … , zn ) = (w1 , … , wn ), where wj = zj /(1 + zn )
for 1 ≤ j ≤ n − 1 and wn = i(1 − zn )/(1 + zn ) is a biholomorphic map from 𝔹 ⟶ 𝕌.
112. Let n > 1. Show that the boundary
of the Siegel upper half-space can be identified with ℂn−1 × ℝ. Show that the
multiplication
(z ′ , t) ⋅ (ζ ′ , τ) = (z ′ + ζ ′ , t + τ + 2ℑ⟨z ′ , ζ ′ ⟩)
turns b𝕌 into a group which is non-Abelian. This group is called the Heisenberg4
group.
1 f (z1 , … , zn−1 , ζn )
f (z1 , … , zn ) = ∫ dζn ,
2πi |ζn |=rn ζn − zn
and show that the integral on the right-hand side depends holomorphically on
z1 , … , zn for all z = (z1 , … , zn ) ∈ P(0, r). Therefore holomorphic functions of sev-
eral variables do not have isolated singularities.
114. Let Ω be a bounded domain in ℂn , n > 1, with a defining function ρ ∈ 𝒞2 which
vanishes precisely on bΩ and dρ ≠ 0 on bΩ. Show that the condition 𝜕u ∧ 𝜕ρ = 0
on bΩ can also be stated as
n
𝜕u
∑ tj =0 on bΩ,
j=1 𝜕z j
6.5 Notes
In Section 6.2 we have followed the expositions of L. Hörmander [41] and R.M. Range
[59]. For a thorough treatment of Lewy’s Theorem (see Theorem 6.15) including inter-
esting consequences for Hardy spaces, the reader should consult E. Stein [67]. Pseudo-
convexity is also crucial for the inhomogeneous Cauchy–Riemann equations, as well
as plurisubharmonic functions, see Chapters 10 and 11. The Levi problem to construct
6.5 Notes | 189
(f , g) = ∫ f (z)g(z) dλ(z),
Ω
for f , g ∈ A2 (Ω).
1 f (ζ )
f (z) = ∫ dζ ,
2πi γs ζ − z
where γs (t) = z + seit , t ∈ [0, 2π], 0 < s ≤ r and Dr (z) = {w ∶ |w − z| < r} ⊂ Ω. Using polar
coordinates and integrating the above equality with respect to s between 0 and r, we
get
1
f (z) = ∫ f (w) dλ(w). (7.1)
πr 2 Dr (z)
https://doi.org/10.1515/9783110417241-007
192 | 7 Bergman spaces
1
sup|f (z)| ≤ ‖f ‖.
z∈K π 1/2 r(K)
n
If K ⊂ Ω ⊂ ℂ , we can find a polycylinder
such that for any z ∈ K we have P(z, r(K)) ⊂ Ω. Hence by iterating the above Cauchy
integrals, we get
Theorem 7.2. Let K ⊂ Ω be a compact set. Then there exists a constant C(K), only de-
pending on K, such that
Proof. If (fk )k is a Cauchy sequence in A2 (Ω), by (7.2), it is also a Cauchy sequence with
respect to uniform convergence on compact subsets of Ω. Hence the sequence (fk )k has
a holomorphic limit f with respect to uniform convergence on compact subsets of Ω.
On the other hand, the original L2 -Cauchy sequence has a subsequence, which con-
verges pointwise almost everywhere to the L2 -limit of the original L2 -Cauchy sequence
(see, for instance, [63]), and so the L2 -limit coincides with the holomorphic function f .
Therefore A2 (Ω) is a closed subspace of L2 (Ω) and itself a Hilbert space.
In the sequel we present basic facts about Hilbert2 spaces and their consequences
for the Bergman spaces.
Theorem 7.4. Let E be a non-empty, convex, closed subset of the Hilbert space H, i.e. for
x, y ∈ E one has tx + (1 − t)y ∈ E, for each t ∈ [0, 1]. Then E contains a uniquely determined
element of minimal norm.
implies that
Theorem 7.5. Let M be a closed subspace of the Hilbert space H. Then there exist
uniquely determined mappings
P ∶ H ⟶ M, Q ∶ H ⟶ M⟂
such that
(1) x = Px + Qx ∀x ∈ H
(2) for x ∈ M we have Px = x, hence P 2 = P and Qx = 0; for x ∈ M ⟂ we have Px = 0,
Qx = x, and Q2 = Q.
(3) The distance of x ∈ H to M is given by
inf{‖x − y‖ ∶ y ∈ M} = ‖x − Px‖.
The left-hand side belongs to M, the right-hand side belongs to M ⟂ , hence both sides
are 0, which proves that P and Q are linear.
Property (3) follows by the definition of Q, whereas property (4) by the fact that
(Px, Qx) = 0 ∀x ∈ H.
In addition, we have
(Px, y) = (Px, Py + Qy) = (Px, Py) and (x, Py) = (Px + Qx, Py) = (Px, Py)
Corollary 7.6. If M ≠ H is a closed, proper subspace of the Hilbert space H, then there
exists an element y ≠ 0 with y ⟂ M.
Theorem 7.7. Let L be a continuous linear functional on the Hilbert space H. Then there
exists a uniquely determined element y ∈ H such that Lx = (x, y) ∀x ∈ H.
Lz |Lz|2
Ly = L(αz) = Lz = = (y, y) = |α|2 (z, z).
‖z‖2 ‖z‖2
For x ∈ H, we define
Lx Lx
x′ = x − y and x″ = y.
(y, y) (y, y)
Lx
(x, y) = (x″ , y) = ( y, y) = Lx.
(y, y)
Corollary 7.8. Let H be a Hilbert space and L ∈ H ′ a continuous linear functional. Then
the dual norm
where y ∈ H corresponds to L.
For fixed z ∈ Ω, (7.2) also implies that the point evaluation f ↦ f (z) is a continu-
ous linear functional on A2 (Ω), hence, by the Riesz representation theorem (see The-
orem 7.7), there is a uniquely determined function kz ∈ A2 (Ω) such that
We set K(z, w) = kz (w). Then w ↦ K(z, w) = kz (w) is an element of A2 (Ω), hence the
function w ↦ K(z, w) is anti-holomorphic on Ω and we have
The function of two complex variables (z, w) ↦ K(z, w) is called Bergman kernel of Ω
and the above identity represents the reproducing property of the Bergman kernel.
Now we use the reproducing property for the holomorphic function z ↦ ku (z),
where u ∈ Ω is fixed:
= ∫ kz (w)K(u, w) dλ(w)
Ω
−
= (∫ K(u, w)kz (w) dλ(w))
Ω
= kz (u),
Theorem 7.9. The Bergman kernel is uniquely determined by the properties that it is an
element of A2 (Ω) in z and that it is conjugate symmetric and reproduces A2 (Ω).
Proof. To see this, let K ′ (z, w) be another kernel with these properties. Then we have
Now let ϕ ∈ L2 (Ω). Since A2 (Ω) is a closed subspace of L2 (Ω) there exists a uniquely
determined orthogonal projection P ∶ L2 (Ω) ⟶ A2 (Ω), see Theorem 7.5. For the func-
tion Pϕ ∈ A2 (Ω), we use the reproducing property and obtain
where we still have used that P is a self-adjoint operator and that Pkz = kz . Hence
N−1
vN = xN − ∑ (xN , uj )uj .
j=1
The element vN is nonzero because xN is not in the linear span of x1 , … , xN−1 and hence
of u1 , … , uN−1 . So we can set uN = vN /‖vN ‖. It is now clear that (uk )Nk=1 has the desired
properties.
Next we prove Bessel’s4 inequality.
Theorem 7.10. If {uα ∶ α ∈ A} is an orthonormal set in the Hilbert space H, then for any
u∈H
Proof. It suffices to show that ∑α∈F |(u, uα )|2 ≤ ‖u‖2 , for any finite set F ⊂ A. We use the
property ‖uα + uβ ‖2 = ‖uα ‖2 + ‖uβ ‖2 , for α, β ∈ A, α ≠ β.
2
0 ≤ ‖u − ∑ (u, uα )uα ‖
α∈F
2
2
= ‖u‖ − 2ℜ(u, ∑ (u, uα )uα ) + ‖ ∑ (u, uα )uα ‖
α∈F α∈F
= ‖u‖2 − 2 ∑ |(u, uα )|2 + ∑ |(u, uα )|2
α∈F α∈F
= ‖u‖2 − ∑ |(u, uα )|2 .
α∈F
Theorem 7.11. If {uα ∶ α ∈ A} is an orthonormal set in the Hilbert space H, then the
following conditions are equivalent:
(1) (Completeness) If (u, uα ) = 0 for all α ∈ A, then u = 0.
(2) (Parseval’s5 equation) ‖u‖2 = ∑α∈A |(u, uα )|2 for all u ∈ H.
(3) u = ∑α∈A (u, uα )uα for each u ∈ H, where the sum has only countably many nonzero
terms and converges in norm to u no matter how these terms are ordered.
Proof. (Claim (1) implies Claim (3)) If u ∈ H, let α1 , α2 , … be any enumeration of those
2
α’s for which (u, uα ) ≠ 0. By Bessel’s inequality, the series ∑∞ j=1 |(u, uαj )| converges, so
M 2 M
‖ ∑ (u, uα )uα ‖ = ∑ |(u, uα )|2 → 0 as m, M → ∞.
j=m j=m
Hence the series ∑∞ j=1 (u, uαj )uαj converges in H. If v = u − ∑j=1 (u, uαj )uαj , then (v, uα ) = 0
∞
Proof. For the proof of this statement, we use the duality for the sequence space l2 to
get
∞ 1/2 ∞ ∞
sup(∑|ϕj (z)|2 ) = sup{|∑ aj ϕj (z)| ∶ ∑ |aj |2 = 1, z ∈ K}
z∈K j=1 j=1 j=1
= sup{|f (z)| ∶ ‖f ‖ = 1, z ∈ K}
≤ C(K), (7.7)
where we have used (7.2) in the last inequality. Now Cauchy–Schwarz inequality gives
∞ ∞ 1/2 ∞ 1/2
2 2
∑|ϕj (z)ϕj (w)| ≤ (∑|ϕj (z)| ) (∑|ϕj (w)| )
j=1 j=1 j=1
7.2 Examples | 199
with uniform convergence at z, w ∈ K. In addition it follows that (ϕj (z))j ∈ l2 and the
function
∞
w ↦ ∑ ϕj (z)ϕj (w)
j=1
belongs to A2 (Ω). Let the sum of the series be denoted by K ′ (z, w). Notice that K ′ (z, w)
is conjugate symmetric and that for f ∈ A2 (Ω) we get
∞
∫ K ′ (z, w)f (w) dλ(w) = ∑ ∫ f (w)ϕj (w) dλ(w) ϕj (z) = f (z)
Ω j=1 Ω
with convergence in the Hilbert space A2 (Ω). But (7.7) implies uniform convergence on
compact subsets of Ω, hence
for all f ∈ A2 (Ω), so K ′ (z, w) is a reproducing kernel. By the uniqueness of the Bergman
kernel, we obtain K ′ (z, w) = K(z, w).
7.2 Examples
1 2π
(f , z n ) = ∫ f (z)z n dλ(z) = ∫ ∫ f (reiθ )r n e−inθ r dr dθ
𝔻 0 0
1 2π f (reiθ ) 1 a
=∫ ∫ re iθ
dθr 2n+1
dr = 2πan ∫ r 2n+1 dr = π n ,
0 0 r n+1 ei(n+1)θ 0 n +1
1 f (z)
an = ∫ n+1 dz,
2πi γr z
200 | 7 Bergman spaces
for γr (θ) = reiθ . Hence, by the uniqueness of the Taylor series expansion, we ob-
tain that (f , ϕn ) = 0, for each n = 0, 1, 2, … implies f ≡ 0. This means that (ϕn )∞
n=0
constitutes an orthonormal basis for A2 (𝔻) and we get
∞
‖f ‖2 = ∑ |(f , ϕn )|2 ,
n=0
which is equivalent to
|an |2
∞ ∞
‖f ‖2 = π ∑ , f (z) = ∑ an z n .
n=0 n + 1 n=0
by the reproducing property of the Bergman kernel. This implies that the Bergman
kernel is of the form
∞
K(z, w) = ∑ ϕn (z)ϕn (w), (7.9)
n=0
where the sum converges uniformly in z on all compact subsets of 𝔻. (This is true
for any complete orthonormal system, as is shown above.) A simple computation
now gives
1 1 1
∞ ∞
K(z, w) = ∑ ϕn (z)ϕn (w) = ∑ (n + 1)(zw)n = . (7.10)
n=0 π n=0 π (1 − zw)2
1 1
f (z) = ∫ f (w) dλ(w).
π 𝔻 (1 − zw)2
7.2 Examples | 201
(b) Theorem 7.13. Let Ωj ⊂ ℂnj , j = 1, 2 be two bounded domains with Bergman kernels
KΩ1 and KΩ2 . Then the Bergman kernel KΩ of the product domain Ω = Ω1 × Ω2 is
given by
(c) For the computation of the Bergman kernel K𝔹n of the unit ball in ℂn , we use the
Beta and Gamma functions
1 Γ(k + 1)Γ(m + 1)
∫ xk (1 − x)m dx = B(k + 1, m + 1) = ,
0 Γ(k + m + 2)
√1−a2
x2 m+1 1 2 k+1
1
∫ x2k+1 (1 − ) dx = (1 − a ) ∫ yk (1 − y)m+1 dy
0 1 − a2 2 0
1
= (1 − a2 )k+1 B(k + 1, m + 2)
2
1 Γ(k + 1)Γ(m + 2)
= (1 − a2 )k+1 .
2 Γ(k + m + 3)
α α
Now we can normalize the orthogonal basis {z α = z1 1 ⋯ zn n } in A2 (𝔹n ) to obtain
π
= ∫ |z |2α1 ⋯ |zn−2 |2αn−2 (1 − |z1 |2 − ⋯ − |zn−2 |2 )αn +1
αn + 1 𝔹n−1 1
|zn−1 |2 αn +1
× |zn−1 |2αn−1 (1 − ) dλ
1 − |z1 |2 − ⋯ − |zn−2 |2
π πΓ(αn−1 + 1)Γ(αn + 2)
=
αn + 1 Γ(αn + αn−1 + 3)
×∫ |z1 |2α1 ⋯ |zn−2 |2αn−2 (1 − |z1 |2 − ⋯ − |zn−2 |2 )αn +αn−1 +2 dλ
𝔹n−2
π πΓ(αn−1 + 1)Γ(αn + 2) πΓ(α1 + 1)Γ(αn + ⋯ + α2 + n)
= ⋯
αn + 1 Γ(αn + αn−1 + 3) Γ(αn + ⋯ + α1 + n + 1)
π n α1 ! ⋯ αn !
= .
(αn + ⋯ + α1 + n)!
(αn + ⋯ + α1 + n)! α α
K𝔹n (z, w) = ∑ z w
α π n α1 ! ⋯ αn !
1 (α + ⋯ + α1 + n)! α α
∞
= ∑ ∑ n z w
π n k=0 |α|=k α1 ! ⋯ αn !
1
∞
= ∑ (k + n)(k + n − 1) ⋯ (k + 1)(z1 w1 + ⋯ + zn wn )k
π n k=0
n! 1
= n .
π (1 − (z1 w1 + ⋯ + zn wn ))n+1
2
(d) In the sequel we will also consider the Fock space A2 (ℂn , e−|z| ) consisting of all
entire functions f such that
ℂn
It is clear, that the Fock space is a Hilbert space with the inner product
2
(f , g) = ∫ f (z)g(z)e−|z| dλ(z).
ℂn
2
Similar as in beginning of this chapter, setting n = 1, we obtain for f ∈ A2 (ℂ, e−|z| )
that
1 2 2
|f (z)| ≤ ∫ e|w| /2 |f (w)|e−|w| /2 dλ(w)
πr 2 Dr (z)
1/2 1/2
1 |w|2 2 −|w|2
≤ (∫ e dλ(w)) (∫ |f (w)| e dλ(w))
πr 2 Dr (z) Dr (z)
1/2
≤ C(∫ |f (w)|2 e−|w| dλ(w))
2
ℂ
≤ C‖f ‖,
7.3 Biholomorphic mappings | 203
where C is a constant only depending on z. This implies that the Fock space
2
A2 (ℂn , e−|z| ) has the reproducing property. The monomials {z α } constitute an
orthogonal basis and the norms of the monomials are
2 2
‖z α ‖2 = ∫ |z1 |2α1 e−|z1 | dλ(z1 ) ⋯ ∫ |zn |2αn e−|zn | dλ(zn )
ℂ ℂ
∞ ∞
n 2α1 +1 −r 2 2
= (2π) ∫ r e dr ⋯ ∫ r 2αn +1 e−r dr
0 0
= π n α1 ! ⋯ αn !.
2
Hence the Bergman kernel of A2 (ℂn , e−|z| ) is of the form
z α wα 1 z α wα 1
∞
K(z, w) = ∑ = ∑ ∑ = exp(z1 w1 + ⋯ + zn wn ). (7.13)
α ‖z α ‖2 π n k=0 |α|=k α1 ! ⋯ αn ! π n
for all z, w ∈ Ω1 .
Since TF is an isometry,
∫ TF −1 f (v)[KΩ2 (v, u)]− dλ(v) = ∫ f (w)[TF KΩ2 (⋅, u)(w)]− dλ(w). (7.17)
Ω2 Ω1
204 | 7 Bergman spaces
which means that the right-hand side of (7.14) has the required reproducing property,
belongs to A2 (Ω1 ) in the variable z and is anti-holomorphic in the variable w, and
hence must agree with KΩ1 (z, w).
Finally, we derive a useful formula for the corresponding orthogonal projections
Pj ∶ L2 (Ωj ) ⟶ A2 (Ωj ), j = 1, 2.
Proof. The left-hand side of (7.18) can be written in the form P1 (TF (g)), hence, by (7.5),
we obtain
Theorem 7.16. Let Ω ⊊ ℂ be a simply connected domain, and let KΩ be the Bergman
kernel of Ω. Let F ∶ Ω ⟶ 𝔻 be the Riemann mapping with the uniqueness properties
F(a) = 0, F ′ (a) > 0 for some a ∈ Ω. Then
π
F ′ (z) = √ K (z, a), z ∈ Ω. (7.19)
KΩ (a, a) Ω
7.3 Biholomorphic mappings | 205
where (⋅, ⋅)Ω denotes the inner product in L2 (Ω) and (⋅, ⋅)𝔻 denotes the inner product
of L2 (𝔻). Similarly, for v ∈ L2 (Ω) and G = F −1 we have
1 i
(u1 , u2 ) = (‖u1 + u2 ‖2 − ‖u1 − u2 ‖2 ) − (‖u1 + iu2 ‖2 − ‖u1 − iu2 ‖2 )
4 4
and
π
(f , F ′ )Ω = (G′ (f ∘ G), 1)𝔻 = πG′ (0)f (G(0)) = f (a).
F ′ (a)
It follows that the function k(z, a) = F π(a) F ′ (z) has the reproducing property and be-
′
π
F ′ (z) = K (z, a).
F ′ (a) Ω
7.4 Exercises
120. Let H be a vector space over ℂ with inner product (⋅, ⋅). Show that
for all x, y ∈ H.
121. Let H be a Hilbert space, x0 ∈ H and let M be a closed subspace of H. Show that
G(y, x1 , x2 , … , xn )
(dist(y, M))2 = .
G(x1 , x2 , … , xn )
and show that F is an entire function. Compute F (n) (0) for n ∈ ℕ0 and use the
uniqueness of the Fourier transform.
Hint: consider the function f (w) = 1/(1 − wz)2 , and use the reproducing property
of the Bergman kernel.
129. Let Ω = {(z1 , z2 ) ∶ |z1 | < |z2 | < 1} ⊂ ℂ2 be the Hartogs triangle. Show that every
holomorphic function on a neighborhood of Ω continues to the bidisc (see The-
orem 6.10) and find an orthonormal basis for A2 (Ω).
Hint: use Laurent series expansion for the variable z2 .
7.5 Notes
The basics on Hilbert spaces can be found, for instance, in [70]. There are numerous
good texts on different aspects of Bergman spaces and their reproducing kernels, the
reader may consult the books by S.-C. Chen and M.-C. Shaw [13], St. Krantz [51], and M.
Jarnicki and P. Pflug [44]. The transformation formula (see Theorem 7.14) is of great im-
portance in the study of the boundary behavior of biholomorphic mappings on strictly
pseudoconvex domains (Fefferman’s Theorem, see [59]). The theorem about the Rie-
mann mapping and the Bergman kernel is contained in St. Bell’s book [7]. Further
results regarding the Bergman kernel and the solution to 𝜕 will be discussed in the
next chapters.
8 The canonical solution operator to 𝜕
In this chapter we will use properties of the Bergman kernel to solve the inhomo-
geneous Cauchy–Riemann equation 𝜕u 𝜕z
= g, where 𝜕z
𝜕
= 21 ( 𝜕x
𝜕
+ i 𝜕y
𝜕
), z = x + iy and
g ∈ A2 (𝔻). It turns out that the Bergman kernel plays an important role for this prob-
lem, especially to find the so-called canonical solution in L2 (𝔻) which is perpendicu-
lar to the corresponding space of holomorphic functions. For this purpose we develop
the spectral theory of compact operators on Hilbert spaces together with basics on
the ideals of Schatten class operators (Section 8.1). In Section 8.2 it is shown that the
canonical solution operator to 𝜕 is a Hilbert–Schmidt operator on A2 (𝔻), whereas on
A2 (𝔹), for the ball 𝔹 ⊂ ℂn , n ≥ 2, the corresponding operator is compact but fails to be
Hilbert–Schmidt (Section 8.3 together with the exercises in Section 8.4).
In this way ℒ(H1 , H2 ) becomes a Banach space. Let 𝒦(H1 , H2 ) denote the subspace of
all compact operators from H1 to H2 .
The following characterization of compactness is useful for the special operators
in the text, see, for instance, [17]:
https://doi.org/10.1515/9783110417241-008
210 | 8 The canonical solution operator to 𝜕
– For every ϵ > 0 there are a C = Cϵ > 0 and a compact operator T = Tϵ ∶ H1 → H2 such
that
– For every ϵ > 0 there are a C = Cϵ > 0 and a compact operator T = Tϵ ∶ H1 → H2 such
that
It is easily seen (small constant – large constant trick) that there is C ′ > 0 such that
hence
To prove the theorem it therefore suffices to show that (8.1) is equivalent to com-
pactness. When S is known to be compact, we choose T = S and C = 1, and (8.1) holds
for every positive ϵ.
For the converse let (vn )n be a bounded sequence in H1 . We want to extract a
Cauchy subsequence from (Svn )n . From (8.1) we have
(vn(N) )n is a subsequence of (vn(N−1) )n , and for any pair v and w in (vn(N) )n we have ‖Sv −
Sw‖ ≤ 1/N.
Let (wk )k be the diagonal sequence defined by wk = vk(k) . Then (wk )k is a subse-
quence of (vn(0) )n and the image sequence under S of (wk )k is a Cauchy sequence. Since
H2 is complete, the image sequence converges and S is compact.
Theorem 8.2. 𝒦(H1 , H2 ) is a closed subspace of ℒ(H1 , H2 ) endowed with the operator
norm.
Proof. Let A ∈ ℒ(H1 , H2 ). Suppose, for each ϵ > 0, there is a compact operator Aϵ such
that ‖A − Aϵ ‖ ≤ ϵ. Then for each u ∈ H1 we have
‖Au − Aϵ u‖ ≤ ϵ‖u‖.
Now we get
‖Au‖ = ‖Au − Aϵ u + Aϵ u‖
≤ ‖Au − Aϵ u‖ + ‖Aϵ u‖
≤ ϵ‖u‖ + ‖Aϵ u‖.
Proof. If (uk )k is a bounded sequence in H1 , then (S(uk ))k is also bounded, because
S is a bounded operator. A is compact, so (A(S(uk )))k has a convergent subsequence.
Thus AS is compact.
To show that TA is compact we use Theorem 8.1:
Corollary 8.4. Let H be a Hilbert space. 𝒦(H, H) forms a two-sided, closed ideal in
ℒ(H, H).
Proof. Suppose that A is compact, then, by Theorem 8.3, AA∗ is also compact. Given
ϵ > 0, it follows that there is a constant C such that
212 | 8 The canonical solution operator to 𝜕
The following theorem is the spectral theorem for compact, self-adjoint operators:
as N → ∞.
hence (Ax, y) + (Ay, x) = 0, now replace x by ix, then i(Ax, y) − i(Ay, x) = 0, therefore
(Ax, y) = 0 for all x, y ∈ H, which implies A = 0.
Claim (b). A = A∗ , if and only if (Ax, x) ∈ ℝ for all x ∈ H.
A = A∗ implies (Ax, x)− = (x, Ax) = (Ax, x) ∈ ℝ. If (Ax, x) = (Ax, x)− for all x ∈ H, then
(Ax, x) = (x, A∗ x)− = (A∗ x, x), and so ((A − A∗ )x, x) = 0 for all x ∈ H. By Claim (a) we have
A = A∗ .
Now we show that for a self-adjoint operator A ∈ ℒ(H) the norm of A is given by
For this let NA = sup‖x‖=1 |(Ax, x)|. Then NA ≤ ‖A‖. We also have
There is a θ ∈ ℝ such that e−iθ (Ax, y) = |(Ax, y)|. Now replace y by e−iθ y, then
1
2|(Ax, y)| ≤ NA (t 2 ‖x‖2 + ‖y‖2 ).
t2
We consider the right-hand side of this inequality as a function of t and get after dif-
ferentiation with respect to t
2
2t‖x‖2 − ‖y‖2 ,
t3
so the right-hand side of the inequality will be minimal, if t 2 = ‖y‖/‖x‖. Hence we obtain
and we show that the corresponding eigenvalue λ0 is real and satisfies |λ0 | = ‖A‖.
By (8.4) we have ‖A‖ = sup‖x‖=1 |(Ax, x)|. Hence there is a sequence (xn )n in H with
‖xn ‖ = 1 and limn→∞ |(Axn , xn )| = ‖A‖. The inner products (Axn , xn ) are real, so there
214 | 8 The canonical solution operator to 𝜕
exists a subsequence, which is again denoted by (xn )n , such that limn→∞ (Axn , xn ) = λ0 .
Then we have λ0 = ‖A‖ or λ0 = −‖A‖.
Now we get
and letting n → ∞
lim ‖Axn − λ0 xn ‖ = 0.
n→∞
H = Hλ0 ⊕ ⋯ ⊕ Hλk−1 ⊕ M,
x = P0 x + P1 x + ⋯ + Pk−1 x + y,
where Ay = 0, therefore
Ax = λ0 P0 x + λ1 P1 x + ⋯ + λk−1 Pk−1 x.
(ii) For each ϵ > 0, the set {λi ∶ λi eigenvalue of A with |λi | ≥ ϵ} is a finite set, other-
wise we would get an infinite orthonormal system {xi } with Axi = λi xi and the sequence
(Axi )i , would have to contain a convergent subsequence. But
which is a contradiction.
Let λ0 , … , λk−1 be the eigenvalues with |λi | ≥ ϵ and let M = (Hλ0 ⊕ ⋯ ⊕ Hλk−1 )⟂ . The
restriction AM of A to M is self-adjoint and compact, hence, by (8.4), there is an eigen-
vector x with AM x = λx and |λ| = ‖AM ‖, and hence Ax = λx. Therefore x ∈ (Hλi )⟂ and
|λ| = ‖AM ‖ < ϵ. So we have
Ax = Ax1 + AM x2 , x1 ∈ M ⟂ , x2 ∈ M.
Letting ϵ → 0 we get
∞
Ax = ∑ λj Pj x.
j=0
|λ0 | ≥ |λ1 | ≥ ⋯ ,
because for each ϵ > 0 there are only finitely many λi with |λi | ≥ ϵ.
Finally, the orthogonality of the eigenvectors xk and Bessel’s inequality (7.6) give
n 2 ∞
2
‖Ax − ∑ λk (x, xk )xk ‖ = ∑ |λk (x, xk )|2 ≤ (‖x‖ sup |λk |) ,
k=1 k=n+1 k>n
and since (λk )k≥1 is a zero-sequence, the above series converges to A in the operator
norm.
216 | 8 The canonical solution operator to 𝜕
Proof. In order to show this, one applies the spectral theorem for the self-adjoint, com-
pact operator A∗ A ∶ H1 ⟶ H1 and gets
∞
A∗ Ax = ∑ s2n (x, en )en , (8.6)
n=0
1 1 s2
(fn , fm ) = (Aen , Aem ) = (A∗ Aen , em ) = n (en , em ) = δn,m .
sn sm sn sm sn sm
For y ∈ H1 with y ⟂ en for each n ∈ ℕ0 , we have by (8.6) that
Hence we have
∞ ∞
Ax = A(x − ∑ (x, en )en ) + A( ∑ (x, en )en )
n=0 n=0
∞ ∞
= ∑ (x, en )Aen = ∑ sn (x, en )fn .
n=0 n=0
Remark 8.9. From the last statement we get the missing direction in the proof of The-
orem 8.6.
The numbers sn are called the s-numbers of A. They are uniquely determined by
the operator A, since they are the square roots of the eigenvalues of A∗ A.
Let 0 < p < ∞. The operator A belongs to the Schatten1 -class Sp , if its sequence
(sn )n of s-numbers belongs to lp . The elements of the Schatten class S1 are called nu-
clear operators, the elements of the Schatten class S2 are called Hilbert–Schmidt2 op-
erators.
Proof. (i) implies (ii). Let (ei )i∈I be as in (i) and (gl )l∈L an orthonormal basis of H2 .
Then, by Parseval’s equality (see Theorem 7.11)
(ii) implies (iii). Let (ei )i∈I be an orthonormal basis of H1 and M a finite subset of I.
We set PM x = ∑i∈M (x, ei )ei . Then
where we used Bessel’s inequality (7.6). By assumption, ∑i∈I ‖Aei ‖2 < ∞, hence we
can approximate A in the operator norm by operators with finite range. Therefore A is
compact and, by Theorem 8.8, can be written as
∞
Ax = ∑ sn (x, xn )fn .
n=0
By Remark 7.1, there exists an orthonormal basis (ξj )j∈J , which contains the orthonor-
mal system (xn )n≥0 . Then we have
∞ 2 ∞
∑ ‖Aξj ‖2 = ∑‖ ∑ sn (ξj , xn )fn ‖ = ∑ s2n < ∞,
j∈J j∈J n=0 n=0
so A is a Hilbert–Schmidt operator.
(iii) implies (i). If A is a Hilbert–Schmidt operator, A can be written as above and
we obtain (i).
Proof. We have
1/2
‖Ax‖ = ‖∑(x, ei )Aei ‖ ≤ (∑ ‖Aei ‖2 ) ‖x‖,
i∈I i∈I
hence ν2 (A) ≥ ‖A‖. It is easily seen that A ↦ (∑i∈I ‖Aei ‖2 )1/2 is a norm.
If H1 and H2 are separable Hilbert spaces, A ∈ ℒ(H1 , H2 ), and (en )n≥1 an orthonor-
mal basis of H1 and (fn )n≥1 an orthonormal basis of H2 , then for each x ∈ H1 we have
∞
Ax = ∑ (x, en )Aen
n=1
The coefficients of Ax with respect to (fj )j≥1 can be computed from the coefficients of x
with respect to (en )n≥1 by means of the infinite matrix
aj,n = (Aen , fj ), j, n ∈ ℕ.
and we obtain
In this case
∞
ν2 (A)2 = ∑ |aj,n |2 .
j,n=1
8.1 Compact operators on Hilbert spaces | 219
Proof. If A has the given form, one can apply Cauchy–Schwarz inequality to get
∞
‖Ax‖22 ≤ ( ∑ |aj,n |2 )‖x‖22 ,
j,n=1
Theorem 8.14. Let S ⊆ ℝn and T ⊆ ℝm be open sets and A ∶ L2 (T) ⟶ L2 (S) a bounded
linear operator. A is a Hilbert–Schmidt operator, if and only if there exists K ∈ L2 (S × T)
such that
Proof. Let (gk )k≥1 and (fj )j≥1 be orthonormal bases of L2 (T) and L2 (S), respectively.
Then (hj,k )j,k≥1 , defined by
Af (s) = ∑ aj,k (f , gk )fj (s) = ∫ ∑ aj,k fj (s)gk (t)f (t) dt = ∫ K(s, t)f (t) dt.
j,k≥1 T j,k≥1 T
220 | 8 The canonical solution operator to 𝜕
for all j, k ∈ ℕ. Bessel’s inequality and Corollary 8.12 imply that A is a Hilbert–Schmidt
operator.
where λ0 ≥ λ1 ≥ ⋯. Then
(i)
(Ax, x)
λ0 = max ,
x∈H (x, x)
where the maximum is attained by an eigenvector with eigenvalue λ0 .
(ii)
(Ax, x)
λj = min max , j ≥ 1,
L∈Nj x∈L
⟂
(x, x)
where Nj is the set of all j-dimensional subspaces of H. The minimum is attained by
the subspace L = Lj = ⟨x0 , … , xj−1 ⟩, i.e.
(Ax, x)
λj = max .
⟂ x∈Lj (x, x)
(Ax, x)
max ≥ λj .
x∈L
⟂
(x, x)
The existence of z0 follows from the fact that for a basis {yk ∶ k = 0, … , j − 1} of L the
system of equations
8.1 Compact operators on Hilbert spaces | 221
j
∑ ai (xi , yk ) = 0, k = 0, … , j − 1,
i=0
j
has a non-trivial solution. Set z0 = ∑i=0 ai xi , then z0 ⟂ L.
Proof. Properties (a), (b) and (c) follow from (8.4) and Theorem 8.8.
(d) If Ax = ∑∞ k=0 sk (A)(x, ek )fk , then A x = ∑k=0 sk (A)(x, fk )ek . The canonical rep-
∗ ∞
(A1 x, x) (A2 x, x)
max ≤ max
x∈L
⟂
(x, x) x∈L⟂ (x, x)
(A1 x, x) (A2 x, x)
λj (A1 ) = min max ≤ min max = λj (A2 ),
L∈Nj x∈L⟂
(x, x) L∈Nj x∈L⟂ (x, x)
for j = 0, 1, 2, …. It follows that (A∗ B∗ BAx, x) = ‖BAx‖2 ≤ ‖B‖2 ‖Ax‖2 = (‖B‖2 A∗ Ax, x),
which implies 0 ≤ A∗ B∗ BA ≤ ‖B‖2 A∗ A, hence
(A∗ Ax, x)
s2j (A) = λj (A∗ A) = min max ,
L∈Nj x∈L
⟂
(x, x)
therefore
‖Ax‖
sj (A) = min max .
L∈Nj x∈L⟂
‖x‖
(K ∗ x, y) = 0 ∀x ⟺ (x, Ky) = 0 ∀x ⟺ Ky = 0.
Hence
‖Ax‖
sj (A) ≤ max .
x∈ker K ‖x‖
sj (A) ≤ ‖A − K‖.
For
j−1
Kx = Kj x = ∑ sk (A)(x, fk )ek
k=0
(A − Kj )x = ∑ sk (A)(x, fk )ek .
k≥j
P ∶ L2 (𝔻) ⟶ A2 (𝔻)
8.2 The canonical solution operator to 𝜕 restricted to A2 (𝔻) | 223
where g(w)
̃ = wg(w). We claim that S(g) is a solution of the inhomogeneous Cauchy–
Riemann equation:
𝜕 𝜕z 𝜕g 𝜕P(g)̃
S(g)(z) = g(z) + z + = g(z),
𝜕z 𝜕z 𝜕z 𝜕z
because g and P(g)̃ are holomorphic functions, therefore 𝜕S(g) = g. In addition, we
have S(g) ⟂ A2 (𝔻), because for arbitrary f ∈ A2 (𝔻) we get
hence
Now set
π
cn2 = ∫ |z|2n dλ(z) = ,
𝔻 n+1
n
and ϕn (z) = z /cn , n ∈ ℕ0 , then the Bergman kernel K(z, w) can be expressed in the
form
z k wk
∞
K(z, w) = ∑ 2
.
k=0 ck
Next we compute
z k wk wn z k−1 wk wn c z n−1
∞ ∞
P(ϕ̃ n )(z) = ∫ ∑ 2 w dλ(w) = ∑ 2 ∫ dλ(w) = n 2 ,
𝔻 k=0 ck cn k=1 ck−1 𝔻 cn cn−1
hence we have
cn z n−1
S(ϕn )(z) = zϕn (z) − 2
, n ∈ ℕ.
cn−1
224 | 8 The canonical solution operator to 𝜕
wk z k zz n cn z n−1
∞
S∗ S(ϕn )(w) = ∫ ∑ 2
(z − w)( − 2 ) dλ(z).
𝔻 k=0 ck cn cn−1
wk z k zz n+1 cn z n
∞
∫ ∑ 2
( − 2 ) dλ(z)
𝔻 k=0 ck cn cn−1
z n+1 wk z k+1 c wk z k
∞ ∞
=∫ ∑ 2
dλ(z) − 2n ∫ z n ∑ 2 dλ(z)
𝔻 cn k=0 ck cn−1 𝔻 k=0 ck
wn wn
= 3 ∫ |z|2n+2 dλ(z) − 2 ∫ |z|2n dλ(z)
cn 𝔻 cn−1 cn 𝔻
2
cn+1 c
=( − 2n )wn .
cn3 cn−1
wk z k zz n cn z n−1
∞
w∫ ∑ 2
( − 2 ) dλ(z)
𝔻 k=0 ck cn cn−1
zn wk z k+1 cn z n−1 wk z k
∞ ∞
=w∫ ∑ dλ(z) − w ∫ ∑ dλ(z)
𝔻 cn k=0 ck2 2
𝔻 cn−1 k=0 ck
2
cn wn−1 cn wn−1
= w( 2
− 2 )
cn−1 cn−1
= 0.
c12
S∗ S(ϕ0 )(w) = ϕ (w).
c02 0
Finally, we get
Theorem 8.17. Let S ∶ A2 (𝔻) ⟶ L2 (𝔻) be the canonical solution operator for 𝜕 and
(ϕk )k the normalized monomials. Then
c12 2
cn+1 cn2
∞
S∗ Sϕ = (ϕ, ϕ 0 )ϕ0 + ∑ ( − )(ϕ, ϕn )ϕn (8.9)
c02 n=1 cn
2 2
cn−1
Since
2
cn+1 c2 1
2
− 2n = →0 as n → ∞,
cn cn−1 (n + 2)(n + 1)
2
cn+1 cn2
∞
∑( − 2
)<∞
n=1 cn2 cn−1
|z − w|2
∫ ∫ dλ(z) dλ(w) < ∞.
𝔻 𝔻 |1 − zw|4
|z − w|2 1
∫ ∫ 4
dλ(z) dλ(w) ≤ ∫ ∫ 2
dλ(z) dλ(w).
𝔻 𝔻 |1 − zw| 𝔻 𝔻 |1 − zw|
Introducing polar coordinates z = reiθ and w = seiϕ , we can write the last integral in
the following form:
1 1 1 2π 2π rs dθ dϕ dr ds
∫ ∫ dλ(z) dλ(w) = ∫ ∫ ∫ ∫
𝔻 𝔻 |1 − zw|2 0 0 0 0 1 − 2rs cos(θ − ϕ) + r 2 s2
1 1 2π 2π 1 − r 2 s2 rs
=∫ ∫ ∫ ∫ dθ dϕ dr ds.
0 0 0 0 1 − 2rs cos(θ − ϕ) + r 2 s2 1 − r 2 s2
Integration of the Poisson kernel with respect to θ (see Lemma 5.8) yields
2π 1 − ρ2
∫ dθ = 2π, 0 < ρ < 1.
0 1 − 2ρ cos(θ − ϕ) + ρ2
Therefore we have
1 1 2π 2π 1 − r 2 s2 rs
∫ ∫ ∫ ∫ dθ dϕ dr ds
0 0 0 0 1 − 2rs cos(θ − ϕ) + r s 1 − r 2 s2
2 2
1 1 rs 1 log(1 − s2 )
= (2π)2 ∫ ∫ dr ds = −(2π) 2
∫ ds < ∞.
0 0 1 − r 2 s2 0 2s
226 | 8 The canonical solution operator to 𝜕
and
n
⟨g(w), z − w⟩ = ∑ gj (w)(z j − wj ),
j=1
Hence the canonical solution operator S can be written in the form S(g) = v − P(v),
where P ∶ L2 (Ω) ⟶ A2 (Ω) is the Bergman projection. If ṽ is another solution to 𝜕u = g,
then v − ṽ ∈ A2 (Ω) hence v = ṽ + h, where h ∈ A2 (Ω). Therefore
Now we get
n n
S(g)(z) = ∑ z j gj (z) − ∫ K(z, w)(∑ wj gj (w)) dλ(w)
j=1 Ω j=1
n n
= ∫ [(∑ z j gj (w))K(z, w) − (∑ wj gj (w))K(z, w)] dλ(w)
Ω j=1 j=1
Remark 8.18. It is pointed out that a (0, 1)-form g = ∑nj=1 gj dz j with holomorphic co-
efficients is not invariant under the pullback by a holomorphic map F = (F1 , … , Fn ) ∶
Ω1 ⟶ Ω. Then
n n n
𝜕F l
F ∗ g = ∑ gl dF l = ∑(∑ gl ) dz j ,
l=1 j=1 l=1 𝜕z j
ω = ω̃ dz1 ∧ ⋯ ∧ dzn ∧ dz 1 ∧ ⋯ ∧ dz n ,
where
(−1)n+j−1
uj (z) = ∫ (z j − wj )K(z, w)ω(w)
̃ dλ(w)
n Ω
(−1)n+j−1
uj (z) = (z j ω(z)
̃ − P(wj ω)(z)),
̃
n
from this we obtain
= ω̃ dz1 ∧ ⋯ ∧ dzn ∧ dz 1 ∧ ⋯ ∧ dz n .
Remark 8.20. The pullback by a holomorphic map F has in this case the form
𝜕Fj 2
F ∗ ω = |det | ω̃ dz1 ∧ ⋯ ∧ dzn ∧ dz 1 ∧ ⋯ ∧ dz n .
𝜕zk
8.4 Exercises
130. Let H be a Hilbert space and A ∈ ℒ(H) a compact operator. Show that the kernel
of I − A is finite dimensional.
131. Let (dn )n≥1 be a sequence of complex numbers with limn→∞ dn = 0. Define the
operator D ∶ l2 ⟶ l2 by D(ξ ) = (dn xn )n≥1 for ξ = (xn )n≥1 ∈ l2 . Show that D is a com-
pact operator and determine its s-numbers.
132. Let A ∈ ℒ(H1 , H2 ) be an operator between the Hilbert spaces H1 and H2 . Show
that A is a nuclear operator if and only if there exist sequences (xj )j≥0 in H1 and
(yj )j≥0 in H2 such that
∞ ∞
Ax = ∑ (x, xj )yj , x ∈ H1 and ∑ ‖xj ‖ ‖yj ‖ < ∞.
j=0 j=0
[(n1 + n2 + 2)!]1/2 n1 n2
Un1 ,n2 (z1 , z2 ) = z1 z2 , n1 , n2 ∈ ℕ0
π(n1 !n2 !)1/2
136. Let S1 ∶ A2(0,1) (𝔹2 ) ⟶ L2 (𝔹2 ) be the canonical solution operator to 𝜕 restricted to
A2(0,1) (𝔹2 ). Compute the images of the basis elements under the canonical solu-
tion operator S1 and show that S1 fails to be a Hilbert–Schmidt operator.
137. Let ϕn (z) = ((n + 1)/π)1/2 z n , n ∈ ℕ0 denote the canonical orthonormal basis in
A2 (𝔻). Consider the following (0, 1)-forms αn in L2(0,1) (𝔻 × 𝔻) with holomorphic
coefficients: αn (z1 , z2 ) ∶= ϕn (z1 ) dz 2 . Show that 𝜕αn = 0 and that the norms in
L2(0,1) (𝔻 × 𝔻) are given by ‖αn ‖ = √π, for all n = 0, 1, 2, ….
138. Show that the canonical solution to 𝜕u = αn is given by un (z1 , z2 ) = ϕn (z1 )z 2 ,
which means that un ∈ A2 (𝔻 × 𝔻)⟂ .
139. Prove that the functions un ∈ L2 (𝔻 × 𝔻) have the properties ‖un ‖ = (π/2)1/2 and
un ⟂ um , if n ≠ m. Conclude from these properties that the canonical solution
operator to 𝜕
S1 ∶ A2(0,1) (𝔻 × 𝔻) ⟶ L2 (𝔻 × 𝔻)
fails to be compact.
8.5 Notes
Most of the material in Section 8.2 can be found in [30]. A general treatment of the
canonical solution operator to 𝜕 restricted to (0, 1)-forms with holomorphic coeffi-
cients is given in [32], with a thorough discussion on Schatten class membership
of the solution operator. We mention that the restriction of the canonical solution
operator to 𝜕 is related to Hankel4 and Toeplitz5 operators (see [65, 64, 29]).
https://doi.org/10.1515/9783110417241-009
232 | 9 Nuclear Fréchet spaces of holomorphic functions
each 0-neighborhood U there exist V ∈ 𝒰 and ϵ > 0 such that ϵV ⊂ U. A family (pα )α∈A
of seminorms is called a fundamental system of seminorms, if the sets Uα = {x ∈ X ∶
pα (x) < 1} constitute a fundamental system of 0-neighborhoods of X. We will write
(X, (pα )α∈A ) to refer to that.
Let X and Y be locally convex vector spaces with fundamental systems (pα )α∈A
and (qβ )β∈B of seminorms. A linear mapping T ∶ X ⟶ Y is continuous, if and only if
for each β ∈ B there exist α ∈ A and a constant C > 0 such that qβ (Tx) ≤ Cpα (x), for all
x ∈ X.
A linear functional x′ on X is continuous, if and only if there exist α ∈ A and a
constant C > 0 such that |x′ (x)| ≤ Cpα (x) for all x ∈ X.
We indicate that the consequences of the Hahn–Banach Theorem (Theorem 4.11
and Corollary 4.12) are also true for locally convex vector spaces, one has to replace the
norm in the proof of Theorem 4.11 by one of the seminorms defining the topology of a
locally convex vector space; a subspace Y of a locally convex vector space X is dense
in X, if and only if each continuous linear functional on X, which vanishes on Y , also
vanishes on the whole of X.
The appropriate concept of a bounded subset in X reads as follows: a subset B of
a locally convex vector space is said to be bounded if to every 0-neighborhood U in X
corresponds a number s > 0 such that B ⊂ tU for every t > s. It is easily seen that B is
bounded, if and only if supx∈B pα (x) < ∞ for all α ∈ A, where (pα )α∈A is a fundamental
system of seminorms for the topology of X; compare with Definition 4.39.
Now let X ′ be the space of all continuous linear functionals on a locally con-
vex vector space (X, (pα )α∈A ). We endow the dual space X ′ with the topology of uni-
form convergence on all bounded subsets of X; which can be expressed in the follow-
ing way: (X ′ , (pB )B∈ℬ ), where pB (x′ ) = supx∈B |x′ (x)| and ℬ denotes the family of all
bounded subsets of X. It is called the strong topology on X ′ .
1 |f − g|m
∞
d(f , g) = ∑ m
. (9.1)
m=1 2 1 + |f − g|m
It is easily seen that d(⋅, ⋅) is a metric which generates the original topology of uniform
convergence on all compact subsets of Ω.
9.2 The space ℋ(DR (0)) and its dual space | 233
is clear. Hence, using the Hilbert norms ‖ ⋅ ‖m , the space (ℋ(DR (0)), (‖ ⋅ ‖m )m∈ℕ ) carries
the original topology of uniform convergence on all compact subsets of DR (0).
Now we consider the Bergman spaces A2 (Drm (0)) endowed with the norm ‖ ⋅ ‖m ,
see Section 7.1. If rm < rℓ < R, we have the inclusions
is a Hilbert–Schmidt operator.
Fix m ∈ ℕ and set
n + 1 zn
ϕℓn (z) ∶= √ ,
π rℓn+1
2
for n = 0, 1, 2, … . Then (ϕℓn )∞
n=0 constitutes an orthonormal basis in A (Drℓ (0)), see Sec-
tion 7.2. By Theorem 8.10, we have to show that
∞
∑ ‖ιℓ,m (ϕℓn )‖2m < ∞. (9.4)
n=0
We just showed that for each m ∈ ℕ there exists an ℓ ∈ ℕ such that the natural
embedding
Then the spaces (ℋ(DR (0)), (| ⋅ |m )m∈ℕ ) and (ΛR , (pm )m∈ℕ ) are topologically isomorphic,
where the isomorphism T ∶ ΛR ⟶ ℋ(DR (0)) is given by
∞
n
T((ξn )∞
n=0 )(z) = ∑ ξn z , z ∈ DR (0),
n=0
and
f (n) (0) ∞
T −1 (f ) = ( ) , f ∈ ℋ(DR (0)).
n! n=0
∞ ∞
n n
n=0 )|m = sup | ∑ ξn z | ≤ ∑ |ξn |rm = pm ((ξn )n=0 ).
|T((ξn )∞ ∞
|z|≤rm n=0 n=0
On the other side, we get from Cauchy’s estimates (see Theorem 2.33) that
f (n) (0) |f |
| | ≤ nℓ ,
n! rℓ
f (n) (0) n
∞
pm (T −1 (f )) = ∑ | |rm
n=0 n!
∞
|f |ℓ n
≤∑ r
n m
n=0 rℓ
n
rm
∞
= (∑ n
)|f |ℓ .
n=0 rℓ
In a similar way, we can describe the dual space of ℋ(DR (0)). Recall that L is a
continuous linear functional on (ℋ(DR (0)), (| ⋅ |m )m∈ℕ ), if and only if there exist an
m ∈ ℕ and a constant C > 0 such that |L(f )| ≤ C|f |m for each f ∈ ℋ(DR (0)).
|ηn |
n=0 ∶ ∃m ∈ ℕ with qm ((ηn )n=0 ) ∶= sup
Λ′R = {(ηn )∞ ∞
n
< ∞}.
n rm
Then the dual space ℋ′ (DR (0)) is isomorphic to the sequence space (Λ′R , (qm )m∈ℕ ).
Proof. We indicate that the seminorms (qm )m∈ℕ are decreasing in m, and that
(Λ′R , (qm )m∈ℕ ) is not a metric space, but a dual metric space.
Let L ∈ ℋ′ (DR (0)). Then there exist an m ∈ ℕ and a constant C > 0 such that
|L(f )| ≤ C|f |m for each f ∈ ℋ(DR (0)), in particular applying L to the monomials z ↦ z n
we obtain a sequence ηn ∶= L(z ↦ z n ) such that
n
|ηn | ≤ Crm , n = 0, 1, 2, … .
f (n) (0)
∞
L(f ) = ∑ ηn ,
n=0 n!
f (0) n
for an arbitrary function f (z) = ∑∞ z in ℋ(DR (0)). Again, from Cauchy’s esti-
(n)
n=0 n!
mates (see Theorem 2.33), we obtain
∞
|f |m+1
|L(f )| ≤ ∑ |ηn | n
n=0 rm+1
n
rm
∞
|ηn |
= |f |m+1 ∑ n n
r r
n=0 m m+1
n
rm
∞
≤ qm ((ηn )∞
n=0 )|f |m+1 ∑ n
.
n=0 rm+1
holomorphic in a neighborhood of ∞, i.e. in a set {w ∈ ℂ ∶ |w| > ℓ}, and with the
property F(∞) = limz→0 F(1/z) = 0. This is done in the following way: suppose that
supn rnn < ∞, then ℓ ∶= lim supn→∞ |ηn |1/n ≤ rm < R. Hence the function
|η |
m
ηn
∞
F(w) = ∑ n+1
,
n=0 w
236 | 9 Nuclear Fréchet spaces of holomorphic functions
is holomorphic in {w ∶ |w| > ℓ} and satisfies F(∞) = 0. We know from the last proof that
the expression
f (n) (0)
∞
L(f ) = ∑ ηn , f ∈ ℋ(DR (0))
n=0 n!
represents an arbitrary continuous linear functional on ℋ(DR (0)). Let ℓ < ρ < R and
γρ (t) = ρeit , t ∈ [0, 2π]. Then
1 1 ηn
∞
∫ F(w)f (w) dw = ∫ ∑ n+1 f (w) dw
2πi γρ 2πi γρ n=0 w
1 f (w)
∞
= ∑ ηn ∫ dw
n=0 2πi γρ wn+1
f (n) (0)
∞
= ∑ ηn
n=0 n!
= L(f ).
1 1 F(w)
L(w ↦ )= ∫ dw = F(z), (9.5)
z−w 2πi γρ z − w
and
1 F (k) (z)
L(w ↦ )= , (9.6)
(z − w) k+1 k!
Theorem 9.3. If W ⊂ ℂ ⧵ DR (0) is a set with multiplicity which has a limit point in
ℂ ⧵ DR (0), then the linear span of ℛ(W) is dense in ℋ(DR (0)).
Proof. To show that the linear span of ℛ(W) is dense in ℋ(DR (0)), we take a continu-
ous linear functional L ∈ ℋ′ (DR (0)) which vanishes on the linear span of ℛ(W). Using
Corollary 4.12, we will be done if we can show that L vanishes on ℋ(DR (0)). The as-
sumptions on ℛ(W) imply that the holomorphic function F corresponding to L by (9.5)
vanishes on a set with a limit point or, using (9.6), has the property that F (k) (ζ ) = 0,
for k = 0, 1, 2, … and some ζ ∈ ℂ ⧵ DR (0). In both cases, the Identity Theorems (see
Theorem 2.18 and Corollary 2.19) imply that F ≡ 0, and hence L = 0 on ℋ(DR (0)).
9.3 Exercises
140. Let (Km )m∈ℕ be a compact exhaustion of the domain Ω ⊆ ℂn and let |f |m ∶=
supz∈Km |f (z)|, for f ∈ ℋ(Ω). Show that
1 |f − g|m
∞
d(f , g) = ∑ m 1 + |f − g|
m=1 2 m
145. Let X be a locally convex vector space. Show that every compact subset K ⊂ X is
bounded.
Hint: Choose a 0-neighborhood U and a balanced 0-neighborhood W such that
W ⊂ U and K ⊂ ⋃∞ n=1 nW .
146. Let (sj )j be a strictly decreasing sequence of positive numbers such that
limj→∞ sj = 0 and V be a bounded subset of the locally convex vector space
X. Show that the collection {sj V ∶ j ∈ ℕ} is a local base for X.
238 | 9 Nuclear Fréchet spaces of holomorphic functions
147. Show that each finite dimensional subspace of a locally convex subspace is
closed.
148. X is locally compact if 0 has a neighborhood whose closure is compact. Show
that each locally compact locally convex vector space has finite dimension.
Hint: take a 0-neighborhood V whose closure is compact, since V is also
bounded, the sets 2−n V , n ∈ ℕ form a local base for X. The compactness of V
shows that there exist x1 , … , xm ∈ X such that
1 1
V ⊂ (x1 + V) ∪ ⋯ ∪ (xm + V).
2 2
k
where f ∈ ℋ(DR (0)) has Taylor series expansion f (z) = ∑∞ k=0 ak z , defines the
original topology of uniform convergence on all compact subsets of DR (0).
151. Show that for each 0 < r < R there exist 0 < ρ < R and a constant C, depending
on r, such that
∞
∑ |ak |r k ≤ C sup|f (z)|,
k=0 |z|≤ρ
k
k=0 ak z .
for each f ∈ ℋ(DR (0)) with Taylor series expansion f (z) = ∑∞
9.4 Notes
For a thorough discussion of locally convex vector spaces related to real and complex
analysis, in particular of nuclear Fréchet spaces, the reader should consult [55] or [50].
The Köthe duality, together with its applications, is presented in [50]. Additional de-
tails and applications to different problems in complex analysis are given in [52].
10 The 𝜕-complex
Our main task will be to solve the inhomogeneous Cauchy–Riemann equation 𝜕u = f ,
where the right-hand side f is a given (0, 1)-form with coefficients in L2 and satisfies the
necessary condition 𝜕f = 0. For n > 1, this is an overdetermined system of partial differ-
ential equations, which will be reduced to a system with equally many unknowns and
equations. As we will now consider a more general setting for the 𝜕-equation (in the
sense of distributions), we also need some background from the theory of unbounded
operators on Hilbert spaces and the basics of the theory of distributions (Section 10.1
and Section 10.2). In Section 10.5 we introduce the complex Laplacian (box opera-
tor), show that, under suitable assumptions, this operator has a bounded inverse, the
𝜕-Neumann operator, and discuss important properties of the 𝜕-Neumann operator
(Section 10.7).
is a closed subspace of H1 × H2 .
The inner product in H1 × H2 is
Lemma 10.2. Let T be a densely defined closable operator. Then there is a closed ex-
tension T, called its closure, whose domain is smallest among all closed extensions.
https://doi.org/10.1515/9783110417241-010
240 | 10 The 𝜕-complex
Proof. Let 𝒱 be the set of x ∈ H1 for which there exist xk ∈ dom(T) and y ∈ H2 such
that limk→∞ xk = x and limk→∞ Txk = y. Since T̃ is a closed extension of T it follows
̃ = y. Therefore y is uniquely determined by x. We define Tx = y
that x ∈ dom(T)̃ and Tx
with dom(T) = 𝒱. Then T is an extension of T and every closed extension of T is also
an extension of T. The graph of T is the closure of the graph of T in H1 × H2 . Hence T
is a closed operator.
Lemma 10.4. Let T be a densely defined operator on H and let S be a bounded operator
on H. Then (T + S)∗ = T ∗ + S∗ .
Proof. Let f ∈ dom(T ∗ + S∗ ) = dom(T ∗ ). Then for all g ∈ dom(T + S) = dom(T) we have
Lemma 10.5. Let T ∶ dom(T) ⟶ H2 be a densely defined linear operator and define
V ∶ H1 × H2 ⟶ H2 × H1 by V((x, y)) = (y, −x). Then
Proof.
Lemma 10.6. Let T ∶ dom(T) ⟶ H2 be a densely defined, closed linear operator. Then
H2 × H1 = V(𝒢(T)) ⊕ 𝒢(T ∗ ).
Lemma 10.7. Let T ∶ dom(T) ⟶ H2 be a densely defined, closed linear operator. Then
dom(T ∗ ) is dense in H2 and T ∗∗ = T.
Proof. Let z⊥ dom(T ∗ ). Hence (z, y)2 = 0 for each y ∈ dom(T ∗ ). We have
V −1 ∶ H2 × H1 ⟶ H1 × H2
where V −1 ((y, x)) = (−x, y), and V −1 V = Id. Now, by Lemma 10.6, we have
Hence (z, y)2 = 0 ⟺ ((0, z), (−T ∗ y, y)) = 0 for each y ∈ dom(T ∗ ). This implies (0, z) ∈
𝒢(T) and therefore z = T(0) = 0, which means that dom(T ∗ ) is dense in H2 .
Since T and T ∗ are densely defined and closed, by Lemma 10.5 we have
Proof. Let v ∈ ker T ∗ and y ∈ im T, which means that there exists u ∈ dom(T) such that
Tu = y. Hence
Lemma 10.9. Let T ∶ dom(T) ⟶ H2 be a densely defined, closed linear operator. Then
ker T is a closed linear subspace of H1 .
Proof. We use Lemma 10.8 for T ∗ and get ker T ∗∗ = (im T ∗ )⊥ . Since, by Lemma 10.7,
T ∗∗ = T, we obtain ker T = (im T ∗ )⊥ and that ker T is a closed linear subspace
of H1 .
For our applications to the 𝜕-equation it will be important to know whether the dif-
ferential operators involved have closed range or are even surjective. In the following
theorems we will explain the functional analysis background and show how inequal-
ities correspond to the closed range property.
First we collect some corresponding results about bounded operators and begin
with the Baire category theorem and the open-mapping theorem.
Since U1 ∩ W is open and nonempty, it contains a ball B(r1 , x1 ), and we can assume that
0 < r1 < 1. We choose sequences {xn }n in X and {rn } in ℝ + inductively as follows: having
chosen xj and rj for j < n, we observe that Un ∩ B(rn−1 , xn−1 ) is open and nonempty, so
we can choose xn and rn such that 0 < rn < 2−n and B(rn , xn ) ⊂ Un ∩ B(rn−1 , xn−1 ). Then,
for n, m ≥ N, we see that xn , xm ∈ B(rN , xN ), and since rn → 0, the sequence {xn }n is a
Cauchy sequence. As X is complete, the limit x = limn→∞ xn exists. Since xn ∈ B(rN , xN )
for all n ≥ N, we have
x ∈ B(rN , xN ) ⊂ UN ∩ B(r1 , x1 ) ⊂ UN ∩ W,
Proof. Let Br denote the open ball of radius r and center 0 in X. We have to show
that T(Br ) contains an open ball about 0 in Y . Since X = ⋃∞ n=1 Bn and T is surjective,
we have Y = ⋃∞ n=1 T(B n ). As Y is complete and the map y ↦ ny is a homeomorphism
of Y which maps T(B1 ) to T(Bn ), we can apply Theorem 10.10 and obtain that T(B1 )
cannot be nowhere dense in Y . Hence there exists y0 ∈ Y and r > 0 such that the ball
B(4r, y0 ) = {y ∈ Y ∶ ‖y − y0 ‖ < 4r} is contained in T(B1 ). Pick y1 = Tx1 ∈ T(B1 ) such that
‖y1 − y0 ‖ < 2r. Then B(2r, y1 ) ⊂ T(B1 ), so if ‖y‖ < 2r, we have
Dividing both sides by 2, we obtain that there exists r > 0 such that if ‖y‖ < r then
y ∈ T(B1 ). If we could replace T(B1 ) by T(B1 ), the proof would be complete.
Dilation invariance implies that if ‖y‖ < r2−n then y ∈ T(B2−n ). Suppose that ‖y‖ <
r/2; we can find x1 ∈ B1/2 such that ‖y − Tx1 ‖ < r/4, and proceeding inductively, we can
find xn ∈ B2−n such that
n
‖y − ∑ Txj ‖ < r2−n−1 . (10.1)
j=1
Proof. Let π1 and π2 be the projections of 𝒢(T) onto X and Y . They are both continuous.
X × Y is also complete, 𝒢(T) is a closed subspace of X × Y and therefore also complete.
π1 is a bijection from 𝒢(T) to X. By Corollary 10.12, π1−1 is bounded. But T = π2 ∘ π1−1 ,
therefore T is bounded.
Theorem 10.14. Suppose X and Y are normed spaces and 𝒜 is a subset of ℒ(X, Y).
(i) If supT∈𝒜 ‖Tx‖ < ∞ for all x in a nonmeager subset of X, then
(ii) If X is a Banach space and supT∈𝒜 ‖Tx‖ < ∞ for all x ∈ X, then
Proof. Let
Then the sets En are closed, so under the hypothesis of (i), some En must contain a
nontrivial closed ball B(r, x0 ). Then E2n ⊃ B(r, 0), for if ‖x‖ ≤ r then x − x0 ∈ En and
where y ∈ H2 .
10.1 Unbounded operators on Hilbert spaces | 245
contradicting ‖Ayk ‖ ≥ ϵ.
For the opposite direction we need a number of prerequisites: first we claim the
following assertion: if (xk )k is a bounded sequence in a Hilbert space H and if ((xk , y))k
is a Cauchy sequence in ℂ for each y in a dense subset M of H, then (xk )k is a weak
Cauchy sequence in H. To show this, let x ∈ H be an arbitrary element and choose
y ∈ M such that ‖x − y‖ < ϵ. Then
Since (xk )k is a bounded sequence, we get that ((xk , x))k is a Cauchy sequence in ℂ
for each x ∈ H. Next we claim that each weak Cauchy sequence in H is also weakly
convergent: a weak Cauchy sequence is bounded, therefore we have |(xk , x)| ≤ C‖x‖,
for some constant C > 0 and for each k ∈ ℕ. Hence
defines a continuous linear functional, for which there exists x0 ∈ H such that
Now we can finish the prove of the opposite direction: let (xk )k be a bounded se-
quence in H1 . We know that it contains a weakly convergent subsequence xkℓ → x0 ,
hence the sequence (xkℓ − x0 )ℓ is a weak null sequence. By assumption we obtain that
Axkℓ → Ax0 in H2 . So A is a compact operator.
Now we continue to derive further important properties of unbounded operators
on Hilbert spaces which will later be used for the 𝜕-operator.
is bijective and continuous and, by the open-mapping theorem (see Theorem 10.11),
also open. This implies the desired inequality.
To prove the other direction, let (fn )n be a sequence in H1 with Tfn → y in H2 . We
have to show that there exists h ∈ H1 with Th = y. Decompose fn = gn + hn , where gn ∈
ker T and hn ∈ (ker T)⟂ . By assumption we have
for all sufficiently large n and m. Hence (hn )n is a Cauchy sequence. Let h = limn→∞ hn .
Then we have
and therefore
Lemma 10.19. Let T be as before. T(H1 ) is closed, if and only if T ∗ (H2 ) is closed.
Proof. Since T ∗∗ = T, it suffices to show one direction. We will show that the closed-
ness of T(H1 ) implies, that (ker T)⟂ = im T ∗ ; since (ker T)⟂ is closed, we will be done.
Let x ∈ im T ∗ . Then there exists y ∈ H2 with x = T ∗ y. Now we get for x ′ ∈ ker T that
λ(Tx) = (x, x ′ )
10.1 Unbounded operators on Hilbert spaces | 247
This implies (y, z)2 = (Tx, z)2 = (x, T ∗ z)1 = (x, x′ )1 , for all x ∈ H1 , and hence x ′ = T ∗ z ∈
im T ∗ .
T̃ ∶ 𝒢(T) ⟶ H2 ,
since
and im T̃ = im T.
By Lemma 10.18, im T is closed, if and only if T|̃ (ker T)̃ ⟂ is bounded from below.
We have ker T̃ = ker T ⊕ {0}, and it remains to show that T|̃ (ker T)̃ ⟂ is bounded from
below, if and only if T|dom(T)∩(ker T)⟂ is bounded from below. But this follows from
C2
‖Tf ‖2 ≥ ‖f ‖2 .
1 − C2
248 | 10 The 𝜕-complex
Proof. (i) and (ii) are equivalent by Lemma 10.19 and the fact that QP = Q∗ P ∗ = (PQ)∗ .
Suppose (ii) holds and let (fn )n and (gn )n be sequences in H with Pfn +
(I − Q)gn → h. Then
By assumption, im(QP) is closed, hence there exists f ∈ H with QPf = Qh; it follows
that Qh = Pf − (I − Q)(Pf ) and
and
In addition, we have
Remark 10.24. The last theorem also holds in the other direction: if T ∶ H1 ⟶ H2 is a
densely defined closed operator and G is a closed subspace of H2 with G = im T, then
T ∗ |dom(T ∗ )∩G is bounded from below. Since in this case G = im T, we have that im T is
closed and hence, by Lemma 10.22, im T ∗ is also closed. Therefore, Lemma 10.20 and
the fact that G = (ker T ∗ )⟂ implies that T ∗ |dom(T ∗ )∩G is bounded from below.
Proof. We have ker T = (im T ∗ )⟂ , hence v ∈ (ker T)⟂ = im T ∗ . In addition, G⟂ ⊆ (im T)⟂ =
ker T ∗ , and therefore
im T ∗ |dom(T ∗ )∩G = im T ∗ ,
which means that im T ∗ is closed and that for v ∈ (ker T)⟂ = im T ∗ there exists an f ∈
dom(T ∗ ) ∩ G with T ∗ f = v. The desired norm-inequality follows from the assumption
that T ∗ |dom(T ∗ )∩G is bounded from below.
250 | 10 The 𝜕-complex
Remark 10.27. (a) If T is a symmetric operator, there are two natural closed exten-
sions. We have dom(T) ⊆ dom(T ∗ ) and T ∗ = T on dom(T). Since T ∗ is closed
(Lemma 10.5), T ∗ is a closed extension of T; it is the maximal closed extension.
T is also closable, by Lemma 10.2, therefore T exists; it is the minimal closed
extension.
(b) If T is essentially self-adjoint, then its self-adjoint extension is unique. To prove
this, let S be a self-adjoint extension of T. Then S is closed and, being an extension
of T, it is also an extension of its smallest extension T. Hence
T ⊂ S = S∗ ⊂ (T)∗ = T
and S = T.
H × H = U(𝒢(T −1 )) ⊕ 𝒢((T −1 )∗ )
10.1 Unbounded operators on Hilbert spaces | 251
and
Consequently,
Hence T −1 is symmetric. (i) implies that T −1 is self-adjoint (and bounded), and now it
follows from (ii) that T = (T −1 )−1 is also self-adjoint.
Proof. Let h ∈ H1 be an arbitrary element and consider (h, 0) ∈ H1 × H2 . From the proof
of Lemma 10.7, we get
I = B + T ∗ C, 0 = TB − C,
which gives
‖h‖2 = ‖(h, 0)‖2 = ‖(f , Tf )‖2 + ‖(T ∗ g, −g)‖2 = ‖f ‖2 + ‖Tf ‖2 + ‖T ∗ g‖2 + ‖g‖2 ,
252 | 10 The 𝜕-complex
and hence
and
Proof. We have to show that 𝒢(T) = 𝒢(T|dom(T ∗ T) ). For this purpose we consider ele-
ments (x, Tx) in the graph of T. We suppose that (x, Tx) ⟂ (y, Ty) for each y ∈ dom(T ∗ T).
Then
Definition 10.33. Let (𝒱, ‖ ⋅ ‖𝒱 ) and (H, ‖ ⋅ ‖H ) be Hilbert spaces such that
𝒱 ⊂ H, (10.4)
and suppose that there exists a constant C > 0 such that for all u ∈ 𝒱 we have
In this situation the space H can be imbedded into the dual space 𝒱′ : for h ∈ H
the mapping
a(u, v) = a(v, u)
for all u, v ∈ 𝒱.
The form a is called 𝒱-elliptic, if there exists a constant α > 0 such that
for all u ∈ 𝒱.
Theorem 10.35. Let a be a continuous, 𝒱-elliptic form on 𝒱 × 𝒱. Using (10.6) and the
Riesz representation theorem (Theorem 7.7), we can define a linear operator
A∶𝒱⟶𝒱
such that
Proof. First we show that A is injective: (10.8) and (10.7) imply that for u ∈ 𝒱 we have
hence
Proof. First we show that S is injective. For each u ∈ dom(S), we get from (10.7) and
(10.5) that
which implies that there exists a uniquely determined w ∈ 𝒱 such that (h, v)H = (w, v)𝒱
for all v ∈ 𝒱. Now we apply Theorem 10.35 and get from (10.8) that a(u, v) = (w, v)𝒱 ,
where u = A−1 w. Since a(u, v) = (h, v)H for each v ∈ 𝒱, we conclude that u ∈ dom(S) and
that Su = h, which shows that S is surjective.
Suppose that (u, h)H = 0 for each u ∈ dom(S). As S is surjective, there is v ∈ dom(S)
such that Sv = h and we get that (u, Sv)H = 0 for each u ∈ dom(S). Using the 𝒱-ellipticity
(10.7), we get for u = v that
for all u ∈ dom(S). Using again the surjectivity of S, we derive that v = v0 ∈ dom(S).
This implies that dom(S) = dom(S∗ ) and that S is self-adjoint. By Lemma 10.28 (iv),
S−1 ∈ ℒ(H).
Finally, we show that dom(S) is dense in 𝒱. Let h ∈ 𝒱 be such that (u, h)𝒱 = 0, for
all u ∈ dom(S). By Theorem 10.35, there exists f ∈ 𝒱 such that Af = h. Then
Definition 10.37. Let Ω ⊆ ℝn an open subset and 𝒟(Ω) = 𝒞∞ 0 (Ω) the space of 𝒞 -func-
∞
𝜕|α| ϕj
α α →0
𝜕x1 1 ⋯ 𝜕xn n
uniformly on K for each α = (α1 , … , αn ).
A distribution is a linear functional u on 𝒟(Ω) such that for every compact subset
K ⊂ Ω there exist a k ∈ ℕ0 = ℕ ∪ {0} and a constant C > 0 with
𝜕|α| ϕ(x)
|u(ϕ)| ≤ C ∑ sup| α α |,
|α|≤k x∈K 𝜕x1 1 ⋯ 𝜕xn n
Definition 10.39. Let f ∈ 𝒞∞ (Ω) and u ∈ 𝒟′ (Ω). The multiplication of u with f is de-
fined by (fu)(ϕ) ∶= u(fϕ) for ϕ ∈ 𝒟(Ω). Notice that fϕ ∈ 𝒟(Ω).
For u ∈ 𝒟′ (ℝn ) and f ∈ 𝒟(ℝn ) the convolution of u and f is defined by
Let
𝜕 𝜕|α|
Dk = and Dα = α α ,
𝜕xk 𝜕x1 1 ⋯ 𝜕xn n
where α = (α1 , … , αn ) is a multiindex. The partial derivative of a distribution u ∈ 𝒟′ (Ω)
is defined by
The support supp(u) of u ∈ 𝒟′ (Ω) consists of x ∈ Ω such that there is no open neighbor-
hood U of x with u|U = 0. We denote the space of distributions with compact support
by ℰ′ (Ω).
A fundamental solution of a differential operator P(D) is a distribution E ∈ ℰ′ (ℝn )
with
P(D)E = δ0 .
In what follows we will find a fundamental solution for 𝜕 and we will investigate
regularity properties of the 𝜕-operator. For this purpose we recall Stokes’ Theorem
in the following way: let Ω ⊂ ℂ be a bounded domain with piecewise 𝒞1 -boundary
γ (positively oriented), let f ∈ 𝒞1 (Ω) and ζ ∈ Ω; for 0 < ϵ < dist(ζ , Ωc ), let Ωϵ ∶= {z ∈ Ω ∶
|z − ζ | > ϵ}, where the boundary of Ωϵ consists of γ and the negatively oriented circle
ρϵ = {z ∶ |z − ζ | = ϵ}. Consider the differential form ω = fz−ζ
(z)
dz, then by Stokes’ Theorem
f (z) 2π
∫ dω = ∫ ω + ∫ ω = ∫ dz − ∫ if (ζ + ϵeiθ ) dθ. (10.12)
Ωϵ γ ρϵ γ z−ζ 0
We have
𝜕 f (z) 𝜕f 1
dω = ( ) dz ∧ dz = − dz ∧ dz,
𝜕z z − ζ 𝜕z z − ζ
2π
and by continuity ∫0 if (ζ + ϵeiθ ) dθ → 2πif (ζ ) as ϵ → 0. Hence we obtain
1 f (z) 1 𝜕f 1
f (ζ ) = ∫ dz + ∫ dz ∧ dz; (10.13)
2πi γ z − ζ 2πi Ω 𝜕z z − ζ
1 𝜕ϕ 1
ϕ(ζ ) = − ∫ dλ(z), (10.14)
π Ω 𝜕z z − ζ
Proof. Let E be the fundamental solution to 𝜕 given in (10.15). We have 𝜕E = δ0 . Now let
Ω′ ⊂⊂ Ω be a relatively compact subset and choose g ∈ 𝒟(Ω) such that g = 1 in some
neighborhood of Ω′ . Let S ∶= 𝜕(gu). Then S ∈ ℰ′ (Ω) with supp(S) ⊂ ℂ ⧵ Ω′ , hence
where the derivatives are taken in the sense of distributions and endow the space with
the norm
1/2
‖f ‖k,Ω = [ ∑ ∫ |𝜕α f |2 dλ] , (10.19)
|α|≤k Ω
𝜕|α| f
𝜕α f = α α .
𝜕x1 1⋯ 𝜕xn n
dense in L2 (Ω), it follows that W00 (Ω) = W 0 (Ω) = L2 (Ω). Using the Fourier transform,
it is also possible to introduce Sobolev spaces of non-integer exponent; see [2, 25].
u(x) = |x|−α , x ∈ 𝔹, x ≠ 0.
Hence
|α|
|∇u(x)| = , x ≠ 0.
|x|α+1
see [24].
Let ϕ ∈ 𝒞∞
0 (𝔹) and let 𝔹ϵ be the open ball around 0 with radius ϵ > 0. Take ω =
𝔹 ⧵ 𝔹ϵ and
where ν(x) = (ν1 (x), … , νn (x)) denotes the inward pointing normal on b𝔹ϵ . If α < n − 1,
then |∇u(x)| ∈ L1 (𝔹), and we obtain
≤ Cϵn−1−α → 0,
as ϵ → 0. Thus
for all ϕ ∈ 𝒞∞
0 (𝔹). As
α
|∇u(x)| = ∈ L2 (𝔹),
|x|α+1
n−2
if and only if 2(α + 1) < n, we get that u ∈ W 1 (𝔹), if and only if α < 2
.
α
Dα (ϕu) = ∑ ( )Dβ ϕDα−β u,
β≤α β
α!
where (αβ ) = β!(α−β)!
.
β
Proof. To prove (i), fix ϕ ∈ 𝒞∞
0 (Ω). Then D ϕ ∈ 𝒞0 (Ω), and
∞
Therefore Dα (ϕu) = ϕDα u + uDα ϕ, as required. The induction step is carried out in a
similar way.
Proof. It is clear that the norm of W k (Ω) stems from an inner product. To prove com-
pleteness, let (um )m be a Cauchy sequence in W k (Ω). Then for each multiindex α with
|α| ≤ k, the sequence (Dα um )m is a Cauchy sequence in L2 (Ω). Since L2 (Ω) is complete,
there exist functions uα ∈ L2 (Ω) such that
Dα um → uα in L2 (Ω).
which proves the claim. Since D um → D u in L2 (Ω) for all |α| ≤ k, we see that um → u
α α
in W k (Ω).
Then
2 1/2
‖f ‖2 = [∫ (∫ F(x, y) dλ(y)) dλ(x)]
ℝn ℝn
For |x| sufficiently small, ‖gx − g‖2 < ϵ/3, hence ‖fx − f ‖2 < ϵ.
n
Let χ ∈ 𝒞∞
0 (ℝ ) be a function with support in the unit ball such that χ ≥ 0 and
∫ χ(x) dλ(x) = 1.
ℝn
10.3 Friedrichs’ lemma | 263
We define χϵ (x) = ϵ−n χ(x/ϵ) for ϵ > 0. Let f ∈ L2 (ℝn ) and define for x ∈ ℝn
= ∫ f (x − ϵx ′ )χ(x′ ) dλ(x′ ).
ℝn
In the first integral we can differentiate under the integral sign to show that fϵ ∈
𝒞∞ (ℝn ).
The family of functions (χϵ )ϵ is called an approximation to the identity.
Proof.
But ‖f−ϵx′ − f ‖2 is bounded by 2‖f ‖2 and tends to 0 as ϵ → 0 by Lemma 10.48. Now set
and we can apply the dominated convergence theorem to get the desired result.
n
If u ∈ 𝒞∞
0 (ℝ ), we have
Dj (u ∗ χϵ ) = (Dj u) ∗ χϵ ,
where Dj = 𝜕/𝜕xj . This also true, if u ∈ L2 (ℝn ) and Dj u is defined in the sense of distri-
butions. We will show even more using these methods for approximating a function
in a Sobolev space by smooth functions.
Let Ω ⊆ ℝn be an open subset and let
For a fixed x ∈ Ωϵ the function ϕ(y) ∶= χϵ (x − y) belongs to 𝒞∞ (Ω). The definition of the
αth weak partial derivative implies
Thus
‖aDj (v ∗ χϵ ) − (aDj v) ∗ χϵ ‖2 → 0 as ϵ → 0,
n
Proof. If v ∈ 𝒞∞
0 (ℝ ), we have
To show (10.22), we may assume that a ∈ 𝒞10 (ℝn ), since v has compact support. We
n
have for v ∈ 𝒞∞
0 (ℝ ),
Let M be the Lipschitz constant for a such that |a(x) − a(x − y)| ≤ M|y|, for all x, y ∈ ℝn .
Then
where
𝜕 𝜕
mj = ∫|y χϵ (y)| dy = ∫|y χ(y)| dλ(y).
𝜕yj 𝜕yj
n n 2 n
This shows (10.22) when v ∈ 𝒞∞
0 (ℝ ). Since 𝒞0 (ℝ ) is dense in L (ℝ ), we have proved
∞
be a first order differential operator with variable coefficients where aj ∈ 𝒞1 (ℝn ) and
a0 ∈ 𝒞(ℝn ). If v ∈ L2 (ℝn ) is a function with compact support and Lv = f ∈ L2 (ℝn ) where
n
Lv is defined in the distribution sense, the convolution vϵ = v ∗ χϵ is in 𝒞∞ 0 (ℝ ) and
2 n
vϵ → v, Lvϵ → f in L (ℝ ) as ϵ → 0.
266 | 10 The 𝜕-complex
Lvϵ − Lv ∗ χϵ = Lvϵ − f ∗ χϵ → 0
and observe that ker T = (im T ∗ )⟂ and ker T ∗ = (im T)⟂ . We claim that the operator
SS∗ + T ∗ T ∶ F ⟶ F
hence SS∗ g = T ∗ Tg = 0, but this gives S∗ g ∈ ker S ∩ im S∗ = ker S ∩ (ker S)⟂ = {0}, and
g ∈ ker S∗ = (im S)⟂ ; from T ∗ Tg = 0 we get Tg ∈ ker T ∗ ∩ im T = (im T)⟂ ∩ im T = {0} and
g ∈ ker T = im S, therefore we obtain g ∈ im S ∩ (im S)⟂ = {0}. So SS∗ + T ∗ T is injective
and as F is finite dimensional SS∗ + T ∗ T is bijective.
Let N = (SS∗ + T ∗ T)−1 . We claim that
u = S∗ Nf
We have
f = SS∗ Nf + T ∗ TNf ,
we have
be the space of (0, 1)-forms with coefficients in L2 (Ω). For u, v ∈ L2(0,1) (Ω), we define the
inner product by
n
(u, v) = ∑(uj , vj ).
j=1
In this way L2(0,1) (Ω) becomes a Hilbert space. (0, 1)-forms with compactly sup-
ported 𝒞∞ coefficients are dense in L2(0,1) (Ω).
then
2
0 (Ω) ⟶ L(0,1) (Ω).
𝜕 ∶ 𝒞∞
Definition 10.54. The domain dom(𝜕) of 𝜕 consists of all functions f ∈ L2 (Ω) such
that 𝜕f , in the sense of distributions, belongs to L2(0,1) (Ω), i.e. 𝜕f = g = ∑nj=1 gj dz j , and
for each ϕ ∈ 𝒞∞ 0 (Ω) we have
𝜕ϕ −
∫ f( ) dλ = − ∫ gj ϕ dλ, j = 1, … , n. (10.24)
Ω 𝜕zj Ω
2
It is clear that 𝒞∞
0 (Ω) ⊆ dom(𝜕), hence dom(𝜕) is dense in L (Ω). Since differenti-
ation is a continuous operation in distribution theory we have
Lemma 10.55. 𝜕 ∶ dom ⟶ L2(0,1) (Ω) has closed graph and ker 𝜕 is a closed subspace of
L2 (Ω).
Proof. We use the arguments of the proof of Theorem 10.45: let (fk )k be a sequence in
dom(𝜕) such that fk → f in L2 (Ω) and 𝜕fk → g in L2(0,1) (Ω). We have to show that 𝜕f = g.
From the proof of Theorem 10.45 we know that 𝜕fk → 𝜕f as distributions. As 𝜕fk → g in
L2(0,1) (Ω), it follows that f ∈ dom(𝜕) and 𝜕f = g.
Now we can apply Lemma 10.9 and get that ker 𝜕 is a closed subspace of L2 (Ω).
ker 𝜕 coincides with the Bergman space A2 (Ω) of all holomorphic functions on Ω
belonging to L2 (Ω). This is due to the fact that 𝜕z
𝜕f
= 0 in the sense of distributions
k
10.5 The 𝜕-Neumann operator | 269
(𝜕 u, ψ) = (u, 𝜕ψ).
∗
Since the compactly supported smooth function are dense in L2 (Ω), we must have
n
𝜕r
∑ ∫ uj ψ dσ = 0,
j=1 bΩ 𝜕zj
on bΩ.
where L2(0,q) (Ω) denotes the space of (0, q)-forms on Ω with coefficients in L2 (Ω). The
𝜕-operator on (0, q)-forms is given by
n
𝜕aJ
𝜕( ∑ aJ dz J ) = ∑ ∑ dz ∧ dz J , (10.26)
′ ′
J j=1 J 𝜕z j j
where ∑′ means that the sum is only taken over strictly increasing multiindices J =
(j1 , … , jq ).
270 | 10 The 𝜕-complex
The derivatives are taken in the sense of distributions, and the domain of 𝜕 con-
sists of those (0, q)-forms for which the right-hand side belongs to L2(0,q+1) (Ω). So 𝜕 is a
densely defined closed operator, and therefore has an adjoint operator from L2(0,q+1) (Ω)
into L2(0,q) (Ω) denoted by 𝜕 .
∗
for 1 ≤ q ≤ n − 1.
We remark that a (0, q + 1)-form u = ∑′J uJ dz J belongs to 𝒞∞
(0,q+1) (Ω) ∩ dom(𝜕 ), if
∗
and only if
n
𝜕r
∑ ukK =0 (10.28)
k=1 𝜕zk
n
𝜕αK
(u, 𝜕α) = ( ∑ uJ dz J , ∑ ∑ dz ∧ dz K )
′ ′
|J|=q+1 j=1|K|=q 𝜕z j j
n
𝜕α
= ∑ ∑ ∫ ujK K dλ
′
j=1|K|=q Ω 𝜕zj
n n
𝜕ujK 𝜕r
αK dλ + ∑ ∑ ∫ ujK αK dσ
′ ′
= −∑ ∑ ∫
j=1|K|=q Ω 𝜕zj j=1|K|=q bΩ 𝜕zj
n n
𝜕ujK 𝜕r
) dz K , ∑ αK dz K ) + ∑ ∫ (∑ ujK )α dσ
′ ′ ′
= ( ∑ (− ∑
|K|=q j=1 𝜕zj |K|=q |K|=q bΩ j=1 𝜕zj K
where
n
𝜕ujK
ϑu = ∑ (− ∑ ) dz K (10.29)
′
and
n
𝜕r
∑ ∫ (∑ ujK )α dσ = − ∫ ⟨θ(ϑ, 𝜕r)u, α⟩ dσ, (10.30)
′
hence we have
where θ(ϑ, dr)u denotes the symbol of ϑ in the 𝜕r direction. Note that for u ∈ dom(𝜕 )
∗
we have 𝜕 u = ϑu.
∗
10.5 The 𝜕-Neumann operator | 271
where
n
𝜕r
∫ ⟨𝜕r ∧ u, α⟩ dσ = ∑ ∑ ∫ uK α dσ. (10.33)
′
bΩ |K|=q k=1 bΩ 𝜕z k kK
as an unbounded, densely defined, closed and self-adjoint operator on L2(0,q) (Ω), for
1 ≤ q ≤ n, which means that □ = □∗ and dom(□) = dom(□∗ ).
Proof. dom(□) contains all smooth forms with compact support, hence □ is densely
defined. Showing that □ is closed depends on the fact that both 𝜕 and 𝜕 are closed:
∗
note that
for u ∈ dom(□). We have to prove that for every sequence uk ∈ dom(□) such that uk → u
in L2(0,q) (Ω) and □uk converges, we have u ∈ dom(□) and □uk → □u. It follows from
(10.34) that
which implies that 𝜕uk converges in L2(0,q+1) (Ω) and that 𝜕 uk converges in L2(0,q−1) (Ω).
∗
Since 𝜕 and 𝜕 are closed operators, we get u ∈ dom(𝜕) ∩ dom(𝜕 ) and 𝜕uk → 𝜕u in
∗ ∗
2 2
L(0,q+1) (Ω) and 𝜕 uk → 𝜕 u in L(0,q−1) (Ω).
∗ ∗
To show that 𝜕u ∈ dom(𝜕 ) and 𝜕 u ∈ dom(𝜕), we first notice that 𝜕 𝜕 uk and 𝜕 𝜕uk
∗ ∗ ∗ ∗
Therefore the convergence of □uk = 𝜕 𝜕 uk + 𝜕 𝜕uk implies that both 𝜕 𝜕 uk and 𝜕 𝜕uk
∗ ∗ ∗ ∗
converge. Now use again that 𝜕 and 𝜕 are closed operators to obtain that 𝜕 𝜕 uk →
∗ ∗
𝜕 𝜕 u and 𝜕 𝜕uk → 𝜕 𝜕u. This implies that □uk → □u. Hence □ is closed.
∗ ∗ ∗
R = 𝜕𝜕 + 𝜕 𝜕 + I
∗ ∗
on dom(□). By Lemma 10.29, both (I + 𝜕 𝜕 )−1 and (I + 𝜕 𝜕)−1 are bounded, self-adjoint
∗ ∗
operators. Consider
L = (I + 𝜕 𝜕 )−1 + (I + 𝜕 𝜕)−1 − I.
∗ ∗
272 | 10 The 𝜕-complex
we have that the range of (I + 𝜕 𝜕 )−1 is contained in dom(𝜕 𝜕 ). Similarly, we have that
∗ ∗
L = (I + 𝜕 𝜕)−1 − 𝜕 𝜕 (I + 𝜕 𝜕 )−1 .
∗ ∗ ∗
2
Since 𝜕 = 0, we have that the range of L is contained in dom(𝜕 𝜕) and
∗
∗ ∗ ∗
𝜕 𝜕L = 𝜕 𝜕(I + 𝜕 𝜕)−1 .
𝜕 𝜕 L = 𝜕 𝜕 (I + 𝜕 𝜕 )−1 .
∗ ∗ ∗
definition of dom(□) are called the 𝜕-Neumann boundary conditions. They amount
to a Dirichlet boundary condition on the normal component of u and to the normal
component of 𝜕u, respectively, see [68] for more details.
Ω ∩ U = {z ∈ ℂn ∶ ℑzn = yn < 0} ∩ U.
Let u = ∑nj=1 uj dz j ∈ 𝒞2(0,1) (Ω) and suppose that the support of u lies in U ∩ Ω. Then
u ∈ dom(□), if and only if
un = 0 on bΩ ∩ U, (10.35)
𝜕uj
=0 on bΩ ∩ U, j = 1, … , n − 1. (10.36)
𝜕z n
Equation (10.35) follows from (10.28), which means that u ∈ dom(𝜕 ), and 𝜕u ∈
∗
dom(𝜕 ) is equivalent to
∗
𝜕uj 𝜕un
− =0 on bΩ ∩ U, j = 1, … , n − 1,
𝜕z n 𝜕z j
𝜕uk 𝜕uj
𝜕u = ∑( − ) dz j ∧ dz k ,
j<k 𝜕z j 𝜕z k
and
n
𝜕uj
𝜕 u = −∑
∗
.
j=1 𝜕zj
and
n 2 n n
2 𝜕uj 𝜕uj 𝜕uk
‖𝜕 u‖ = ‖∑
∗
‖ = (∑ ,∑ ).
j=1 𝜕zj j=1 𝜕zj k=1 𝜕zk
n n
𝜕uk 𝜕uj 𝜕 𝜕uk 𝜕r 𝜕uk
− ∑( , ) = ∑ {( , uj ) − ∫ uj dσ}
j,k=1 𝜕z j 𝜕z k j,k=1 𝜕z k 𝜕z j bΩ 𝜕z k 𝜕z j
n
𝜕 𝜕uk 𝜕r 𝜕uk
= ∑ {( ,u ) − ∫ uj dσ}
j,k=1 𝜕z j 𝜕zk j bΩ 𝜕zk 𝜕z j
n
𝜕uk 𝜕uj 𝜕r 𝜕uk 𝜕r 𝜕uk
= ∑ {−( , )+∫ uj dσ − ∫ uj dσ}
j,k=1 𝜕zk 𝜕zj bΩ 𝜕z j 𝜕zk bΩ 𝜕zk 𝜕z j
n n
𝜕r 𝜕uk 𝜕r 𝜕uk
= −‖𝜕 u‖2 + ∑ ∫ u dσ − ∑ ∫ uj dσ.
∗
Also since u ∈ 𝒟0,1 , the vector field X ∶= ∑nk=1 uk 𝜕z𝜕 is tangent to bΩ. We claim that,
k
if g is a 𝒞2 -function on U ∩ bΩ, where U is a neighborhood of bΩ such that g = 0 on
bΩ, then X(g) = 0 on bΩ. To show this, we can assume that 0 ∈ bΩ and that U ∩ bΩ =
{z ∈ U ∶ ℑzn = 0}. Then the defining function r can be written as r(z) = ℑzn and, setting
z ′ = (z1 , … , zn−1 ), we get
where the last integral defines a 𝒞1 -function h with g = rh. Using a 𝒞∞ partition of
unity, we get a global function h. With the identity g = rh, it is now easy to show that
X(g) = 0 on bΩ.
The function g0 = ∑nj=1 𝜕z
𝜕r
uj = 0 on bΩ. Hence we get
j
n n n n
𝜕 𝜕r 𝜕2 r 𝜕r 𝜕uj
X(g0 ) = ∑ uk (∑ uj ) = ∑ uk uj + ∑ uk = 0.
k=1 𝜕zk j=1 𝜕z j j,k=1 𝜕z j 𝜕zk j,k=1 𝜕z j 𝜕zk
n n
𝜕r 𝜕uk 𝜕2 r
∑ uj = − ∑ uj uk ,
j,k=1 𝜕zk 𝜕z j j,k=1 𝜕zj 𝜕z k
n n
𝜕uk 2 𝜕2 r
‖𝜕u‖2 = ∑ ‖ ‖ − ‖𝜕 u‖2 + ∫ ∑ uj uk dσ.
∗
n 2
𝜕uj
‖𝜕u‖2 + ‖𝜕 u‖2 ≥ ∑ ‖ (10.38)
∗
‖.
j,k=1 𝜕z k
n
𝜕2 r
∑ (p)uj (p)uk (p) ≥ 0,
j,k=1 𝜕zj 𝜕z k
for all p ∈ bΩ and for all u ∈ 𝒟0,1 . Hence the result follows immediately from Theo-
rem 10.59.
10.6 Density in the graph norm | 275
Theorem 10.61. If bΩ is 𝒞k+1 , then 𝒞k(0,q) (Ω) ∩ dom(𝜕 ) is dense in dom(𝜕) ∩ dom(𝜕 )
∗ ∗
in the graph norm u ↦ (‖u‖2 + ‖𝜕u‖2 + ‖𝜕 u‖2 )1/2 . The statement also holds with k + 1 and
∗
k replaced by ∞.
Before we begin with the proof of this important approximation result, we gather
some observations about dom(𝜕) and dom(𝜕 ).
∗
Remark 10.62. (a) It is useful to know that dom(𝜕 ) is preserved under multiplica-
∗
(b) Compactly supported forms are not dense in dom(𝜕) ∩ dom(𝜕 ) in the graph norm:
∗
𝜕uj 2 𝜕uj 2
‖ ‖ =‖ ‖.
𝜕zk 𝜕z k
Hence
‖u‖21 ≤ 2(‖𝜕u‖2 + ‖𝜕 u‖2 ),
∗
where ‖u‖21 denotes the standard Sobolev-1 norm of u on Ω. Therefore the closure
of the compactly supported forms in dom(𝜕) ∩ dom(𝜕 ) in the graph norm is con-
∗
tained in the Sobolev space W01 (Ω) for forms that are 𝒞∞ on Ω, which means that
they are zero on the boundary, which is stronger than the condition
n
𝜕r
∑ uj = 0
j=1 𝜕zj
Lemma 10.63. Let Ω be as in Theorem 10.61. Then 𝒞∞ (0,q) (Ω) is dense in dom(𝜕) ∩
dom(𝜕 ) in the graph norm u ↦ (‖u‖2 + ‖𝜕u‖2 + ‖𝜕 u‖2 )1/2 .
∗ ∗
Proof. By this we mean that if u ∈ dom(𝜕) ∩ dom(𝜕 ), one can construct a sequence
∗
2
um ∈ 𝒞∞(0,q) (Ω) such that um → u, 𝜕um → 𝜕u and ϑum → ϑu in L (Ω), recall (10.29).
We use a method closely related to Friedrichs’ Lemma (Lemma 10.50) and use the
notation from there.
Let (χϵ )ϵ be an approximation of the identity and (δν )ν a sequence of small positive
numbers with δν → 0, and define
Then (Ωδν )ν is a sequence of relatively compact open subsets of Ω with union equal
to Ω. The forms uϵ = u ∗ χϵ belong to 𝒞∞
(0,q) (Ωδν ) and uϵ → u, 𝜕uϵ → 𝜕u and ϑuϵ → ϑu in
L2 (Ωδν ), see Lemma 10.48 and Lemma 10.50.
To see that this can be done up to the boundary, we first assume that the domain
Ω is star-shaped and 0 ∈ Ω is a center. Let Ωϵ = {(1 + ϵ)z ∶ z ∈ Ω} and
z
uϵ (z) = u( ),
1+ϵ
where the dilation is performed for each coefficient of u. Then Ω ⊂⊂ Ωϵ and uϵ ∈ L2 (Ωϵ ).
Also, by the dominated convergence theorem, uϵ → u, 𝜕uϵ → 𝜕u and ϑuϵ → ϑu in
L2 (Ω). Now we regularize uϵ defining
where δϵ → 0 as ϵ → 0 and δϵ is chosen sufficiently small. Then u(ϵ) ∈ 𝒞∞ (0,q) (Ω) and
u(ϵ) → u, 𝜕u(ϵ) → 𝜕u and ϑu(ϵ) → ϑu in L2 (Ω). Thus, 𝒞∞
(0,q) (Ω) is dense in the graph norm
when Ω is star-shaped. The general case follows by using a partition of unity since we
assume that our domain has at least 𝒞2 boundary.
Lemma 10.64. Let Ω be as in Theorem 10.61. Then compactly supported smooth forms
are dense in dom(𝜕 ) in the graph norm u ↦ (‖u‖2 + ‖𝜕 u‖2 )1/2 .
∗ ∗
ϑ ũ = ϑu
̃
where ϑũ = ϑu in Ω and ϑũ = 0 outside of Ω. This can be checked from the definition
n
of 𝜕 , since for any v ∈ 𝒞∞
∗
(0,q−1) (ℂ ),
(u,̃ 𝜕v)L2 (ℂn ) = (u, 𝜕v)L2 (Ω) = (ϑu, v)L2 (Ω) = (ϑu,
̃ v) 2 n .
L (ℂ )
10.6 Density in the graph norm | 277
Lemma 10.65. Let Ω be as in Theorem 10.61. Then 𝒞k(0,q) (Ω) ∩ dom(𝜕 ) is dense in
∗
Proof. By Lemma 10.63, it suffices to show that for any u ∈ 𝒞∞ (0,q) (Ω), one can find a
sequence um ∈ 𝒞(0,q) (Ω) ∩ dom(𝜕 ) such that um → u and 𝜕um → 𝜕u in L2 (Ω).
k ∗
Let r be a 𝒞k+1 defining function such that |dr| = 1 on bΩ. We now introduce
some special vector fields and (1, 0)-forms associated with bΩ. Near a point p ∈ bΩ we
choose fields L1 , L2 , … , Ln−1 of type (1, 0) that are orthonormal and span Tp1,0 (bΩ). This
can be done by choosing a basis, and then using the Gram–Schmidt process. To this
collection add Ln , the complex normal, normalized to have length 1. So Ln is a smooth
multiple of
n
𝜕r 𝜕
∑ .
j=1 𝜕z j 𝜕zj
Now denote by w1 , w2 , … , w2 the (1, 0)-forms such that wj (Lk ) = δjk . Ln is defined
globally, in contrast to L1 , … , Ln−1 . The wj ’s then form an orthonormal basis for the
(1, 0)-forms near p. The (1, 0)-form wn is a smooth multiple of ∑nj=1 𝜕z
𝜕r
dzj , and is again
j
globally defined. Taking wedge products of the wj ’s yields (local) orthonormal bases
for the (1, 0)-forms.
We will regularize near a boundary point p ∈ bΩ. Let U be a small neighborhood
of p. By a partition of unity, we may assume that Ω ∩ U is star-shaped and u is sup-
ported in U ∩ Ω. Shrinking U if necessary, we can choose a special boundary chart
(t1 , t2 , … , t2n−1 , r), where (t1 , t2 , … , t2n−1 , 0) are coordinates on bΩ near p. Let w1 , … , wn
be an orthonormal basis for the (0, 1)-forms on U such that 𝜕r = wn .
Let Lj = ∑ns=1 ajs 𝜕z𝜕 , wj = ∑ns=1 bjs dzs , 1 ≤ j ≤ n. Then
s
n n n
𝜕
δjk = wj (Lk ) = ∑ bjs dzs (∑ akℓ ) = ∑ bjs aks .
s=1 ℓ=1 𝜕zℓ s=1
278 | 10 The 𝜕-complex
Consequently, if f is a function,
n n n
𝜕f
𝜕f = ∑ dz s = ∑ ask (Lk f )bsj wj = ∑(Lj f )wj , (10.41)
s=1 𝜕z s j,k,s=1 j=1
where the superscripts denote the entries of the inverses of the corresponding matri-
ces with subscripts. Since multiplication by functions in 𝒞1 (Ω) preserves dom(𝜕 ),
∗
|J|=q |J|=q
n
= ∑ ∑(Lj uJ )wj ∧ wJ + ∑ uJ 𝜕wJ . (10.43)
′ ′
Indeed, the only boundary terms that arise when proving (10.28) come from integrat-
ing
∫ Ln αK unK dλ
Ω
Hence, (u, 𝜕α) = (ϑu, α) for all α ∈ dom(𝜕), which implies u ∈ dom(𝜕 ) and 𝜕 u = ϑu.
∗ ∗
These arguments also give a formula for ϑ and 𝜕 in special boundary frames:
∗
n
ϑu = ϑ( ∑ uJ wJ ) = − ∑ (∑ Lj ujK )wK + 0th order(u). (10.45)
′ ′
u = uτ + uν ,
10.6 Density in the graph norm | 279
where
uτ = ∑ uJ wJ , uν = ∑ uJ wJ .
′ ′
|J|=q,n∉J |J|=q,n∈J
𝜕r ∧ uν = cwn ∧ ∑ uJ wJ = 0. (10.46)
′
|J|=q,n∈J
k
In order to approximate a form u ∈ 𝒞∞ (0,q) (Ω) by forms in 𝒞(0,q) (Ω) ∩ dom(𝜕 ), we
∗
only change the complex normal part uν and leave the complex tangential part uτ
τ k
unchanged: for u ∈ 𝒞∞ (0,q) (Ω), it follows that u ∈ 𝒞(0,q) (Ω) ∩ dom(𝜕 ), and we denote
∗
𝜕ũ ν = 𝜕u
̃ν
in L2 (ℂn ) in the sense of distributions. This follows from (10.32) and (10.46), since uν ∈
𝒞k(0,q) (Ω) and for α ∈ 𝒞∞ n
(0,q+1) (ℂ ) we have
Since 𝜕 is a first order differential operator with variable coefficients, we get from
Friedrichs’ Lemma (Lemma 10.51)
We set
u(−ϵ) = uτ + uν(−ϵ) .
It follows that u(−ϵ) ∈ 𝒞k(0,q) (Ω) ∩ dom(𝜕 ), since each coefficient of uτ , uν(−ϵ) and wj is
∗
in 𝒞k (Ω ∩ U). Therefore we get u(−ϵ) ∈ 𝒞k(0,q) (Ω) ∩ dom(𝜕 ) and u(−ϵ) → u in L2 (Ω).
∗
in L2 (Ω) as ϵ → 0.
ize u in each small star-shaped neighborhood near the boundary. We regularize the
complex tangential and normal parts separately by setting
which means that we first consider u(ϵ) as it was defined in (10.39), take the tangential
components uτ(ϵ) , consider u(−ϵ) as it is defined in (10.40) and then take the normal
components uν(−ϵ) . It follows that for sufficiently small ϵ > 0, uν(−ϵ) has coefficients in
τ
𝒞∞0 (Ω) and u(ϵ) has coefficients in 𝒞 (Ω).
∞
k
Thus we see that u((ϵ)) ∈ 𝒞(0,q) (Ω) ∩ dom(𝜕 ). We get from Lemma 10.63 that u(ϵ) →
∗
u in the graph norm u ↦ (‖u‖2 + ‖𝜕u‖2 + ‖𝜕 u‖2 )1/2 , hence uτ(ϵ) → uτ in the graph norm
∗
u ↦ (‖u‖2 + ‖𝜕u‖2 + ‖𝜕 u‖2 )1/2 . From Lemma 10.64 we obtain that u(−ϵ) → u in the graph
∗
norm u ↦ (‖u‖2 + ‖𝜕 u‖2 )1/2 , hence uν(−ϵ) → uν in the graph norm u ↦ (‖u‖2 + ‖𝜕 u‖2 )1/2 .
∗ ∗
Finally, we use Lemma 10.65, in particular formula (10.47), and see that 𝜕uν(−ϵ) →
𝜕ũ ν in L2 (ℂn ), hence u((ϵ)) → u in the graph norm u ↦ (‖u‖2 + ‖𝜕u‖2 + ‖𝜕 u‖2 )1/2 .
∗
This shows that 𝒞k(0,q) (Ω) ∩ dom(𝜕 ) is dense in dom(𝜕) ∩ dom(𝜕 ) in the graph
∗ ∗
2 2 ∗ 2 1/2
norm u ↦ (‖u‖ + ‖𝜕u‖ + ‖𝜕 u‖ ) .
First we will show that (10.48) implies that 𝜕 and 𝜕 have closed range.
∗
10.7 Properties of the 𝜕-Neumann operator | 281
1
‖u‖ ≤ ‖𝜕u‖.
c
Now we can apply Lemma 10.20 to conclude that im 𝜕 is closed. Finally, Theorem 10.22
gives that im 𝜕 is also closed.
∗
The next result describes the implication of the basic estimates (10.48) for the
□-operator.
1
‖Nu‖ ≤ ‖u‖. (10.50)
c
Proof. Since (□u, u) = ‖𝜕u‖2 + ‖𝜕 u‖2 , it follows that for a convergent sequence (□un )n
∗
we get
which implies that (un )n is convergent and, since □ is a closed operator, we obtain
that □ has closed range. If □u = 0, we get 𝜕u = 0 and 𝜕 u = 0 and by (10.48) that u = 0,
∗
hence □ is injective. By Lemma 10.28 (ii), the range of □ is dense, therefore □ is sur-
jective.
We showed that
The basic estimates (10.48) imply that j is a bounded operator with operator norm
1
‖j‖ ≤ .
√c
By (10.48) it follows in addition that dom(𝜕) ∩ dom(𝜕 ) endowed with the graph-norm
∗
Since (u, v) = (u, jv), we have that (u, v) = (j∗ u, v)Q . Equation (10.51) suggests that as
an operator to dom(𝜕) ∩ dom(𝜕 ),N coincides with j∗ and as an operator to L2(0,q) (Ω),
∗
Ñ = j ∘ j∗ , (10.52)
10.7 Properties of the 𝜕-Neumann operator | 283
We claim that the range of Ñ is contained in dom(□). To show this we use an approach
due to F. Berger (see [8]): since □ is self-adjoint, it suffices to show that Nu ̃ ∈ dom(□∗ )
for all u ∈ L2(0,q) (Ω), which means to show that the functional v ↦ (□v, Nu) ̃ is bounded
on dom(□):
(□Nu,
̃ v) = (Nu,
̃ v)Q = (j∗ u, v) = (u, jv) = (u, v),
Q
hence □Nu
̃ = u. In a similar way we obtain, for u ∈ dom(□),
(N□u,
̃ v) = (□u, Nv)
̃ = (u, Nv)
̃ Q = (u, j∗ v) = (ju, v) = (u, v),
Q
Remark 10.68. Similarly one can use Theorem 10.36 in order to prove the existence
and the main properties of the 𝜕-Neumann operator. For this purpose let H = L2(0,q) (Ω)
and 𝒱 = dom(𝜕) ∩ dom(𝜕 ), endowed with the graph-norm u ↦ (‖𝜕u‖2 + ‖𝜕 u‖2 )1/2 . The
∗ ∗
The basic estimates (10.48) show that 𝒱 is continuously embedded into H, which
gives (10.5), and, by definition, the sesquilinear form a(u, v) is 𝒱-elliptic, see (10.7).
The domain of the operator S in Theorem 10.36 coincides with the domain of the
□-operator because the mapping v ↦ a(u, v) is continuous in the L2 -norm exactly, if
we can write
see Exercise 164. We have S = □ and, by Theorem 10.36, we get immediately that
Theorem 10.70. Let Nq denote the 𝜕-Neumann operator on L2(0,q) (Ω). Then
on dom(𝜕)(𝜕) and
Nq−1 𝜕 = 𝜕 Nq , (10.54)
∗ ∗
on dom(𝜕 ).
∗
Since we already know that both operators 𝜕Nq and 𝜕 Nq are bounded, we can
∗
extend both operators Nq+1 𝜕 and Nq−1 𝜕 to bounded operators on L2(0,q) (Ω).
∗
Theorem 10.71. Let α ∈ L2(0,q) (Ω), with 𝜕α = 0. Then u0 = 𝜕 Nq α is the canonical solu-
∗
α = 𝜕 𝜕 Nq α + 𝜕 𝜕Nq α. (10.56)
∗ ∗
0 = 𝜕α = 𝜕𝜕 𝜕Nq α,
∗
Finally, we set u0 = 𝜕 Nq α and derive from (10.56) and (10.57) that for 𝜕α = 0,
∗
α = 𝜕u0 ,
It follows that
‖𝜕 Nq α‖2 = (𝜕 𝜕 Nq α, Nq α)
∗ ∗
= (𝜕 𝜕 Nq α, Nq α) + (𝜕 𝜕Nq α, Nq α)
∗ ∗
‖𝜕 Nq α‖ ≤ c−1/2 ‖α‖.
∗
operator on
Theorem 10.72.
and
where we applied Lemma 10.3. Hence it follows that for u ∈ L2(0,q) (Ω) we have
Nq u = Nq (𝜕 𝜕 + 𝜕 𝜕)Nq u
∗ ∗
Theorem 10.73. Let Pq ∶ L2(0,q) (Ω) ⟶ ker 𝜕 denote the orthogonal projection, which is
the Bergman projection for q = 0. Then
Pq = I − 𝜕 Nq+1 𝜕, (10.59)
∗
on dom(𝜕).
Proof. First we show that the range of the right-hand side of (10.59), which we denote
by P,̃ coincides with ker 𝜕. Indeed, for u ∈ dom(𝜕), we have
u − 𝜕 Nn+1 𝜕u = u
∗
and
2
P̃ u = Pu
̃ − 𝜕∗ Nq+1 𝜕Pu
̃
= Pu
̃ − 𝜕 Nq+1 𝜕u + 𝜕 Nq+1 𝜕 𝜕 Nq+1 𝜕u
∗ ∗ ∗
= Pu
̃ − 𝜕∗ Nq+1 𝜕u + 𝜕∗ Nq+1 (□ − 𝜕∗ 𝜕)Nq+1 𝜕u
= Pu.
̃
(𝜕 𝜕 Nq u, 𝜕 𝜕Nq u) = (𝜕 𝜕 𝜕 Nq u, 𝜕Nq u) = 0.
∗ ∗ ∗
Hence
10.8 Exercises
152. Let H be a Hilbert space. Show that for each ϵ > 0 there exists a constant Cϵ > 0
such that
153. Show that a densely defined operator T ∶ dom(T) ⟶ H2 between Hilbert spaces
is closable, if and only if the following holds: if (fn )n is a sequence in dom(T)
such that limn→∞ fn = 0, and the sequence (Tfn )n in H2 is convergent, then we
have limn→∞ Tfn = 0.
154. Let T be a densely defined operator. For f , g ∈ dom(T), define a scalar product by
Show that T is closed if and only if (dom(T), (⋅, ⋅)T ) is a Hilbert space.
155. Let X and Y be Fréchet spaces and T ∶ X ⟶ Y a continuous linear operator
which is surjective. Show that T is open.
Hint: use Theorem 10.10 and proceed similarly as in Theorem 10.11.
156. Let H be a Hilbert space and X be a Banach space. Show that an operator A ∈
ℒ(H, X) is compact, if and only if for each orthonormal sequence (en )n in H one
has that limn→∞ Aen = 0.
157. Let A ∈ ℒ(H1 , H2 ) be a bounded operator between Hilbert spaces. Show that A is
compact, if and only if for arbitrary orthonormal sequences (en )n in H1 and (fn )n
in H2 one has that limn→∞ (Aen , fn ) = 0.
158. Let A ∈ ℒ(H) be a compact operator on the Hilbert space H. Show that I − A has
closed range.
Hint: use Theorem 10.11 for the operator I − A.
159. Show that every symmetric operator T on the Hilbert space H is closable and T
is also symmetric.
160. Let Ω ⊂ ℂn be a bounded domain, n ≥ 2. Show that the kernel of 𝜕 in L2 (Ω) co-
incides with the Bergman space A2 (Ω). Show that the kernel of 𝜕 in L2(0,q) (Ω) for
1 ≤ q ≤ n − 1 is larger than the space of (0, q)-forms with coefficients in A2 (Ω).
161. Let Ω ⊂ ℂn be a bounded pseudoconvex domain. Consider the operator T ∶
L2(0,1) (Ω) ⟶ L2 (Ω) defined by
n
⟨f (w), z − w⟩ = ∑ fk (w)(z k − wk ).
k=1
288 | 10 The 𝜕-complex
Conclude from this property that the canonical solution operator S from (8.10)
is compact as an operator from A2(0,1) (Ω) to L2 (Ω), if and only if the operator T is
compact as an operator from L2(0,1) (Ω) to L2 (Ω).
163. Show that 𝜕 N restricted to A2(0,1) (Ω) is compact, if and only if the commutators
∗
now consider (𝜕u, ̄ 𝜕f̄ ) and show that 𝜕f̄ ∈ dom(𝜕̄ ). Similarly, consider (𝜕̄ w, 𝜕̄ ∗ f ),
∗ ∗
where
10.9 Notes
The basics on functional analysis, such as the open-mapping, the closed-graph theo-
rem and the uniform boundedness principle were taken from [24]. A thorough treat-
ment of unbounded operators on Hilbert spaces can be found in [70]. The construction
of self-adjoint operators by means of sesquilinear forms acting on different Hilbert
spaces is mainly based on the assumption of the 𝒱-ellipticity, sometimes also known
as the Lax–Milgram situation, see [10, 34]. Basic facts about pseudoconvex domains
are given, for instance, in [41, 51, 44]. We point out that Theorem 10.67, and the fol-
lowing results on the 𝜕-Neumann operator, also hold for general bounded pseudo-
convex domains without regularity conditions for the boundary of the domain, see
[13, 68]. Theorems 10.71 and 10.73 are due to J.J. Kohn [47, 48]. The approach to the
𝜕-Neumann operator N using the quadratic form Q can be found in [68]. Later on,
the formula N = j ∘ j∗ will be very important to handle the problem of compactness
10.9 Notes | 289
of N. This approach is also inspired by the general study of quadratic forms associ-
ated with non-negative self-adjoint operators and their quadratic roots, see [18]. We
will explain some of these ideas, which are related to the spectral theorem in more
details in Chapter 11. It turns out that the question of compactness of N is crucial for
regularity of N and the Sobolev theory for N. For a thorough treatment of compactness
of N, the reader should consult [68, 30].
11 The twisted 𝜕-complex and Schrödinger operators
We will consider the twisted 𝜕-complex
T S
L2 (Ω) ⟶ L2(0,1) (Ω) ⟶ L2(0,2) (Ω) (11.1)
Suppose (11.2) holds and let α ∈ H2 be such that Sα = 0. Then there exists σ ∈ H1 such
that (i) T(σ) = α and (ii) ‖σ‖21 ≤ ‖P −1 α‖22 .
https://doi.org/10.1515/9783110417241-011
292 | 11 The twisted 𝜕-complex and Schrödinger operators
|(u, α)2 | = |(Pu, P −1 α)2 | ≤ ‖Pu‖2 ‖P −1 α‖2 ≤ (‖T ∗ u‖21 + ‖Su‖23 )1/2 ‖P −1 α‖2
= ‖T ∗ u‖1 ‖P −1 α‖2 ,
if u⊥2 ker S, then (u, α)2 = 0. In addition, we have that T ∗ w = 0 for all w⊥2 ker S, this
follows from the assumption that Tf ∈ ker S, so 0 = (w, Tf )2 ≤ C‖f ‖1 , which means that
w ∈ dom(T ∗ ) and T ∗ w = 0, since (T ∗ w, f )1 = (w, Tf )2 = 0 for all f ∈ dom(T). If T ∗ u =
0, it follows from the above estimate that (u, α)2 = 0, which implies that the linear
functional T ∗ u ↦ (u, α)2 for u ∈ dom(T ∗ ) is well-defined and continuous.
We apply the Hahn–Banach Theorem, where we keep the constant for the esti-
mate of the functional and the Riesz representation theorem to get σ ∈ H1 such that
(T ∗ u, σ)1 = (u, α)2 , which implies that σ ∈ dom(T ∗∗ ). By Lemma 10.7, we have T = T ∗∗ ,
so we get (u, Tσ)2 = (u, α)2 , for all u ∈ dom(T ∗ ). Since dom(T ∗ ) is dense in H2 , we obtain
Tσ = α and, again by the above estimate ‖σ‖1 ≤ ‖P −1 α‖2 .
and for u = ∑nj=1 uj dz j with coefficients uj in 𝒞∞ (Ω), we will write u ∈ Λ0,1 (Ω), we define
𝜕uk 𝜕uj
Su = ∑ √τ( − ) dz j ∧ dz k . (11.4)
j<k 𝜕z j 𝜕z k
We call τ a twist factor. But we also introduce a weight factor φ: for f ∈ 𝒞∞ (Ω) and
u ∈ 𝒟0,1 (see Theorem 10.56),
T ∗ u = √τ 𝜕φ u,
∗
𝜕g 𝜕r −φ
(δℓ f , g)φ = −(f , ) +∫ fg e dσ, (A)
𝜕z ℓ φ bΩ 𝜕zℓ
𝜕f 𝜕r −φ
( , g) = −(f , δk g)φ + ∫ f g e dσ, (B)
𝜕z k φ bΩ 𝜕z k
𝜕 𝜕2 φ
[δℓ , ]f = f, (C)
𝜕z k 𝜕z k 𝜕zℓ
𝜕f
δℓ (fg) = f (δℓ g) + g. (D)
𝜕zℓ
and
n
𝜕g
⟨𝜕g, ξ ⟩(p) ∶= ∑ (p)ξk , (11.6)
k=1 𝜕zk
‖√τ + A 𝜕φ u‖2φ + ‖√τ 𝜕u‖2φ ≥ ‖√τu‖2φ,z + ∫ Θ(u, u)e−φ dλ + ∫ τi𝜕𝜕r(u, u)e−φ dσ,
∗
Ω bΩ
where
|⟨𝜕τ, u⟩|2
Θ(u, u) = τi𝜕𝜕φ(u, u) − i𝜕𝜕τ(u, u) − . (11.8)
A
Proof. Like in the case without twists, we get
𝜕uk 𝜕uj 2
‖√τ 𝜕u‖2φ = ∑‖√τ( − )‖
j<k 𝜕z j 𝜕z k φ
n n
𝜕uk 2 𝜕u 𝜕uj
= ∑ ‖√τ ‖ − ∑ (√τ k , √τ )
j,k=1 𝜕z j φ j,k=1 𝜕z j 𝜕z k φ
294 | 11 The twisted 𝜕-complex and Schrödinger operators
and
n
‖√τ 𝜕φ u‖2φ = ∑ (√τδk uk , √τδj uj )φ ,
∗
j,k=1
hence
n
𝜕uk 𝜕uj
‖√τ 𝜕u‖2φ + ‖√τ 𝜕φ u‖2φ = ‖√τu‖2φ,z + ∑ {(τδk uk , δj uj )φ − (τ
∗
, ) };
j,k=1 𝜕z j 𝜕z k φ
where
n
𝜕r
T1 = ∑ ∫ τ(δk uk )uj e−φ dσ,
j,k=1 bΩ 𝜕z j
and
n
𝜕r 𝜕uk
T2 = − ∑ ∫ τ u e−φ dσ.
j,k=1 bΩ 𝜕zk 𝜕z j j
where
n
𝜕r 𝜕τ
T3 = ∑ ∫ uk u e−φ dσ.
j,k=1 bΩ 𝜕z j 𝜕zk j
Now we get
n
𝜕τ
‖√τ 𝜕u‖2φ + ‖√τ 𝜕φ u‖2φ = (A1) + (A2) + (A3) + T3 − ∑ (δk uk , u)
∗
j,k=1 𝜕zj j φ
n 2
𝜕τ 𝜕τ
− ∑ (uk , δu + u)
j,k=1 𝜕z k j j 𝜕z k 𝜕zj j φ
= (A1) + (A2) + (A3) + T3
n
𝜕τ
− ∫ i𝜕𝜕τ(u, u)e−φ dλ − 2ℜ ∑ (δk uk , u) .
Ω j,k=1 𝜕zj j φ
j,k=1 𝜕zj j φ
Now we move the first term in the last expression to the other side and get the desired
result, since
Ω
0,1
T1 = T3 = 0 for u ∈ 𝒟 , and
n
𝜕2 r
T2 = ∑ ∫ τ u u e−φ dσ = ∫ τi𝜕𝜕r(u, u)e−φ dσ.
j,k=1 bΩ 𝜕z j 𝜕zk j k bΩ
Remark 11.3. In a similar way one can prove an analogous twisted estimate for
(0, q)-forms, 1 ≤ q ≤ n, again using integration by parts repeatedly, see [68].
1
‖𝜕 u‖2 + ‖𝜕u‖2 ≥ ‖u‖2 . (11.9)
∗
4R2
Proof. Since Ω is pseudoconvex, the boundary integral in Theorem 11.2 is ≥ 0. Take
φ = 0 and τ(z) = R2 − |z|2 , then −i𝜕𝜕τ(u, u) = |u|2 and |⟨𝜕τ, u⟩|2 = | ∑nj=1 z j uj |2 ≤ |z|2 |u|2 .
We choose A = 2|z|2 , then |⟨𝜕τ, u⟩|2 /A ≤ 21 |u|2 . Hence from Theorem 11.2 we get
1
‖√R2 + |z|2 𝜕 u‖2 + ‖√R2 − |z|2 𝜕u‖2 ≥ ‖u‖2 , (11.10)
∗
2
for u ∈ 𝒟0,1 . Now, the result follows from Theorem 10.61.
Remark 11.5. In a similar way one can show that for (0, q)-forms u ∈ dom(𝜕) ∩
dom(𝜕 ), 1 ≤ q ≤ n, the corresponding basic estimate reads as
∗
q
‖𝜕 u‖2 + ‖𝜕u‖2 ≥ 2 ‖u‖2 ,
∗
4R
the factor q arises because each strictly increasing multiindex of length q will appear
q times in the corresponding term on the right-hand side of (11.10). One can also dis-
pense with the smoothness requirement on bΩ, see [68].
The next result gives Hörmander’s1 L2 -estimates for the solution of the inhomo-
geneous Cauchy–Riemann equations.
Corollary 11.6. Let Ω be as in Theorem 11.4 and let α ∈ L2(0,1) (Ω) be such that 𝜕α = 0.
Then there exists s ∈ L2 (Ω) such that 𝜕s = α and
Proof. Apply Theorem 11.1 to T = 𝜕∘ √R2 + |z|2 and S = √R2 − |z|2 ∘𝜕, and set P = 1/√2 Id.
Then we have T ∗ = √R2 + |z|2 ∘ 𝜕 , and Theorem 11.4 gives
∗
By Theorem 11.1, we obtain σ ∈ L2 (Ω) such that Tσ = 𝜕(√R2 + |z|2 σ) = α and ‖σ‖2 ≤
2‖α‖2 . Now set s = √R2 + |z|2 σ, then 𝜕s = α and
|s|2
∫ dλ ≤ 2 ∫ |α|2 dλ,
Ω R2 + |z|2 Ω
so we get
1
∫ |s|2 dλ ≤ 2 ∫ |α|2 dλ.
2R2 Ω Ω
where λ denotes the Lebesgue measure, the space L2(0,1) (ℂn , e−φ ) of (0, 1)-forms with
coefficients in L2 (ℂn , e−φ ) and the space L2(0,2) (ℂn , e−φ ) of (0, 2)-forms with coefficients
in L2 (ℂn , e−φ ). Let
(f , g)φ = ∫ f ge−φ dλ
ℂn
‖f ‖2φ = ∫ |f |2 e−φ dλ
ℂn
𝜕 𝜕
L2 (ℂn , e−φ ) ⟶ L2(0,1) (ℂn , e−φ ) ⟶ L2(0,2) (ℂn , e−φ ), (11.12)
⟵
∗
⟵
∗
𝜕φ 𝜕φ
where 𝜕φ is the adjoint operator to 𝜕 with respect to the weighted inner product. For a
∗
n
𝜕 𝜕φ
𝜕φ u = − ∑( (11.13)
∗
− )u .
j=1 𝜕zj 𝜕zj j
and dom(□φ ) is the space of all f ∈ L2(0,1) (ℂn , e−φ ) such that
f ∈ dom(𝜕) ∩ dom(𝜕φ )
∗
bounded inverse of □φ .
In the weighted space L2(0,1) (ℂn , e−φ ) we can give a simple characterization of
dom(𝜕φ ):
∗
Theorem 11.7. Let f = ∑ fj dz j ∈ L2(0,1) (ℂn , e−φ ). Then f ∈ dom(𝜕φ ), if and only if
∗
n
𝜕
eφ ∑ (fj e−φ ) ∈ L2 (ℂn , e−φ ), (11.15)
j=1 𝜕z j
n n
𝜕g 𝜕
(𝜕g, χR f )φ = ∑( , χR fj ) = − ∫ ∑ g (χR f j e−φ ) dλ,
j=1 𝜕z j φ ℂ j=1 𝜕z j
n
Hence by assumption,
11.4 The 𝜕-Neumann operator on weighted (0, q)-forms | 299
n
𝜕
|(𝜕g, f )φ | ≤ ‖g‖φ ‖eφ ∑ (fj e−φ )‖ + M‖g‖φ ‖f ‖φ ≤ C‖g‖φ
j=1 𝜕z j φ
n n
for all g ∈ 𝒞∞0 (ℂ ), and by density of 𝒞0 (ℂ ) this is true for all g ∈ dom(𝜕).
∞
Conversely, let f ∈ dom(𝜕φ ), which means that there exists a uniquely determined
∗
(𝜕g, f )φ = (g, 𝜕φ f )φ .
∗
n
0 (ℂ ). Then g ∈ dom(𝜕) and
Now take g ∈ 𝒞∞
(g, 𝜕φ f )φ = (𝜕g, f )φ
∗
n
𝜕g
= ∑( ,f )
j=1 𝜕z j j φ
n
𝜕
= −(g, ∑ (fj e−φ ))
j=1 𝜕z j L2
n
𝜕
= −(g, eφ ∑ (fj e−φ )) .
j=1 𝜕z j φ
n 2 n −φ
Since 𝒞∞
0 (ℂ ) is dense in L (ℂ , e ), we conclude that
n
𝜕
𝜕φ f = −eφ ∑ (fj e−φ ),
∗
j=1 𝜕z j
n
Lemma 11.8. Forms with coefficients in 𝒞∞
0 (ℂ ) are dense in dom(𝜕) ∩ dom(𝜕φ ) in the
∗
1
graph norm f ↦ (‖f ‖2φ + ‖𝜕f ‖2φ + ‖𝜕φ f ‖2φ ) 2 .
∗
Proof. First we show that compactly supported L2 -forms are dense in the graph norm.
So let {χR }R∈ℕ be a family of smooth radially symmetric cut-offs identically one on
𝔹R and supported in 𝔹R+1 , such that all first order derivatives of the functions in this
family are uniformly bounded in R by a constant M.
Let f ∈ dom(𝜕) ∩ dom(𝜕φ ). Then, clearly, χR f ∈ dom(𝜕) ∩ dom(𝜕φ ) and χR f → f in
∗ ∗
j=1 𝜕z j
hence
n
𝜕
𝜕φ (χR f ) = −eφ ∑ (χR fj e−φ ).
∗
j=1 𝜕z j
300 | 11 The twisted 𝜕-complex and Schrödinger operators
j=1 𝜕zj
Now both terms tend to 0 as R → ∞, and one can see similarly that also 𝜕(χR f ) → 𝜕f
as R → ∞.
So we have density of compactly supported forms in the graph norm, and den-
n
sity of forms with coefficients in 𝒞∞
0 (ℂ ) follows by a standard smoothing process
(Friedrichs’ Lemma (see Lemma 10.51)).
Proof. By Lemma 11.8, it suffices to show (11.16) for each (0, 1)-form u = ∑nk=1 uk dz k
n
with coefficients uk ∈ 𝒞∞
0 (ℂ ), for k = 1, … , n.
For this purpose we set δk = 𝜕z𝜕 − 𝜕z
𝜕φ
and get, since
k k
𝜕uj 𝜕uk
𝜕u = ∑( − ) dz j ∧ dz k ,
j<k 𝜕z k 𝜕z j
that
n
𝜕uj 𝜕uk 2 −φ
‖𝜕u‖2φ + ‖𝜕φ u‖2φ = ∫ ∑| | e dλ + ∫ ∑ δj uj δk uk e−φ dλ
∗
−
ℂn j<k 𝜕z k 𝜕z j ℂn j,k=1
n 2 n
𝜕uj 𝜕uj 𝜕uk −φ
= ∑∫ | | e−φ dλ + ∑ ∫ (δj uj δk uk − )e dλ
j,k=1 ℂ
n 𝜕z k j,k=1 ℂ
n 𝜕z k 𝜕z j
n 2 n
𝜕uj 𝜕
= ∑∫ | | e−φ dλ + ∑ ∫ [δj , ]u u e−φ dλ,
j,k=1 ℂn 𝜕z k j,k=1 ℂn 𝜕z k j k
n
where we used the fact that for f , g ∈ 𝒞∞
0 (ℂ ) we have
𝜕f
( , g) = −(f , δk g)φ
𝜕z k φ
11.4 The 𝜕-Neumann operator on weighted (0, q)-forms | 301
and hence
𝜕 𝜕uj 𝜕uk
([δj , ]uj , uk ) = −( , ) + (δj uj , δk uk )φ .
𝜕z k φ 𝜕z k 𝜕z j φ
Since
𝜕 𝜕2 φ
[δj , ]= ,
𝜕z k 𝜕zj 𝜕z k
we have
n n
𝜕uj 2
𝜕2 φ
‖𝜕u‖2φ + ‖𝜕φ u‖2φ = ∑ ∫ | | e−φ dλ + ∑ ∫ u u e−φ dλ, (11.17)
∗
j,k=1 ℂ
n 𝜕z k j,k=1 ℂ
n 𝜕zj 𝜕z k j k
Remark 11.10. For n = 1, one has 𝜕u = 0, if u is a (0, 1)-form, and (11.17) implies
𝜕u 2 −φ 𝜕2 φ 2 −φ
‖𝜕φ u‖2φ = ∫ | | e dλ + ∫ |u| e dλ, (11.18)
∗
ℂ 𝜕z ℂ 𝜕z𝜕z
for u ∈ dom(𝜕φ ).
∗
Now we generalize formula (11.17) for (0, q)-forms u = ∑′|J|=q uJ dz J with coefficients
n
in 𝒞∞
0 (ℂ ), where 1 ≤ q ≤ n. We notice that
n
𝜕uJ
dz j ∧ dz J
′
𝜕u = ∑ ∑
|J|=q j=1 𝜕z j
and
n
𝜕φ u = − ∑ ∑ δk ukK dz K .
∗ ′
|K|=q−1 k=1
We obtain
n
𝜕uJ 𝜕uM −φ
‖𝜕u‖2φ + ‖𝜕φ u‖2φ = ∑ ∑ ϵjJkM ∫ e dλ
∗ ′
|J|=|M|=q j,k=1 ℂn 𝜕z j 𝜕z k
n
∑ ∫ δj ujK δk ukK e−φ dλ,
′
+ ∑
n
|K|=q−1 j,k=1 ℂ
n n
𝜕uJ 2 𝜕ujK 𝜕ukK −φ
‖ + ∑ ∑ ∫ (δj ujK δk ukK − )e dλ. (11.19)
′ ′
∑ ∑‖
|J|=q j=1 𝜕z j φ |K|=q−1 j,k=1 ℂn 𝜕z k 𝜕z j
302 | 11 The twisted 𝜕-complex and Schrödinger operators
In order to prove this, we first consider the (nonzero) terms where j = k (and hence
M = J). These terms result in the portion of the first sum in (11.19) where j ∉ J. On the
other hand, when j ≠ k, then j ∈ M and k ∈ J, and deletion of j from M and k from
J results in the strictly increasing multiindex K of length q − 1. Consequently, these
terms can be collected into the second sum in (11.19) (in the part with the minus sign,
we have interchanged the summation indices j and k). In this sum, the terms where
j = k compensate for the terms in the first sum where j ∈ J.
Now one can use the same reasoning as in the last proof to get
n n
𝜕uJ 2 𝜕2 φ
‖𝜕u‖2φ + ‖𝜕φ u‖2φ = ∑ ∑‖ u u e−φ dλ. (11.20)
∗ ′ ′
‖ + ∑ ∑∫
|J|=q j=1 𝜕z j φ |K|=q−1 j,k=1 ℂn 𝜕zj 𝜕z k jK kK
Now we suppose that the sum sq of the q smallest eigenvalues of the Levi matrix
𝜕2 φ
Mφ = ( )
𝜕zj 𝜕z k jk
of φ satisfies
for all z ∈ ℂn and some ϵ > 0. If φ(z) = ϵ|z|2 + ψ(z), where ψ is an arbitrary plurisub-
harmonic 𝒞2 -function, then (11.21) is always satisfied.
Theorem 11.11. For a plurisubharmonic weight function φ satisfying (11.21), there exists
a uniquely determined bounded linear operator
such that □φ ∘ Nφ u = u, for any u ∈ L2(0,1) (ℂn , e−φ ). If u ∈ L2(0,1) (ℂn , e−φ ) satisfies 𝜕u = 0,
then 𝜕φ Nφ u is the canonical solution of 𝜕f = u, which means that 𝜕φ Nφ u⊥A2 (ℂn , φ),
∗ ∗
where
given by L(u) = (u, v)φ . Notice that dom(𝜕) ∩ dom(𝜕φ ) is a Hilbert space with the inner
∗
and we claim that Nφ v ∈ dom(□φ ) = dom(□∗φ ), for which we have to show that w ↦
(□φ w, Nφ v)φ is bounded on dom(□φ ):
now we get
hence □φ Nφ v = v, for any v ∈ L2(0,1) (ℂn , φ). If we set u = Nφ v, we get again from Theo-
rem 11.9 that
‖𝜕Nφ v‖2φ + ‖𝜕φ Nφ v‖2φ = Qφ (Nφ v, Nφ v) = (Nφ v, v)φ ≤ ‖Nφ v‖φ ‖v‖φ
∗
hence
where C1 , C2 , C3 , C4 > 0 are constants. Hence we get that Nφ is a continuous linear oper-
ator from L2(0,1) (ℂn , φ) into itself (see also [41] or [13]). The rest is clear from the remarks
made for the 𝜕-Neumann operator without weights in Section 10.7.
Remark 11.12. In this case one can also show that Nφ = jφ ∘ jφ∗ , where
is the embedding and dom(𝜕) ∩ dom(𝜕φ ) is endowed with the graph norm
∗
Theorem 11.13. Let 1 ≤ q ≤ n and suppose that the sum sq of any q (equivalently, the
smallest q) eigenvalues of Mφ satisfies (11.21). Then there exists a uniquely determined
bounded linear operator
Proof. Let μφ,1 ≤ μφ,2 ≤ ⋯ ≤ μφ,n denote the eigenvalues of Mφ and suppose that Mφ is
diagonalized. Then, in a suitable basis,
n n
𝜕2 φ
ujK ukK = ∑ ∑ μφ,j |ujK |2
′ ′
∑ ∑
|K|=q−1 j,k=1 𝜕z 𝜕z
j k |K|=q−1 j=1
≥ sq |u|2 .
The last equality results as follows: for J = (j1 , … , jq ) fixed, |uJ |2 occurs precisely q
times in the second sum, once as |uj1 K1 |2 , once as |uj2 K2 |2 , etc. At each occurrence, it is
multiplied by μφ,jℓ . For the rest of the proof, we use (11.20) and proceed as in the proof
of Theorem 11.11.
and dom(𝜕) ∩ dom(𝜕φ ) is endowed with the graph norm (‖𝜕u‖2φ + ‖𝜕φ u‖2φ )1/2 .
∗ ∗
It also follows in this case that both 𝜕 and 𝜕φ have closed range, see Theorem 10.66.
∗
In Theorem 11.38, we will see that we can replace (11.21) by the weaker condition
For this purpose we will have to use a result about the bottom of the essential spectrum
of an elliptic operator (Persson’s Lemma), together with the Witten Laplacian, which is
a unitarily equivalent form of □φ ; for n = 1, it is a Schrödinger2 operator with magnetic
field.
for u ∈ 𝒟0,1 .
Let P ∶ L2(0,1) (Ω, e−φ ) ⟶ L2(0,1) (Ω, e−φ ) be the multiplication operator by the func-
tion √c. Then it follows from the assumption that
By Theorem 11.1, we get a function f ∈ L2 (Ω, e−φ ) with 𝜕f = g and ‖f ‖φ ≤ √2‖P −1 g‖φ .
For a positive ψ = ∑nj,k=1 ψj,k dzj ∧ dz k ∈ Λ1,1 (Ω) and α = ∑nj=1 αj dz j ∈ Λ0,1 (Ω) we set
n
|α|2ψ ∶= ∑ ψj,k αj αk , (11.25)
j,k=1
Proof. We apply Theorem 11.15 with φ replaced by φ + 2 log(1 + |z|2 ) and use that
n
𝜕2
∑ wj wk log(1 + |z|2 ) = (1 + |z|2 )−2 (|w|2 (1 + |z|2 ) − |(w, z)|2 )
j,k=1 𝜕zj 𝜕z k
≥ (1 + |z|2 )−2 |w|2 ,
so we can take c(z) = 2(1 + |z|2 )−2 to obtain the desired result.
306 | 11 The twisted 𝜕-complex and Schrödinger operators
D = {z ∶ |zj | < r, j = 1, … , n} ⊂ Ω.
for k = 0, 1, … , n.
We shall prove inductively that for every k there is a holomorphic function uk in
Ωk with uk (z0 ) = 1 and
When k = 0, we can take u0 (z) ≡ 1, and un will have the desired properties.
Assume that 0 < k ≤ n and that uk−1 has already been constructed. Choose ψ ∈
𝒞∞
0 (ℂ) so that ψ(zk ) = 0 when |zk | > r/2 and ψ(zk ) = 1 when |zk | < r/3, and set
notice that ψ(zk )uk−1 (z) = 0 in Ωk ⧵ Ωk−1 . To make uk holomorphic we must choose v
as a solution of the equation 𝜕v = zk−1 uk−1 𝜕ψ = f . By the inductive hypothesis, we have
Hence it follows from Theorem 11.17 that v can be found such that
Lemma 11.19. Let ζ ∈ ℂn and K > 0 and define g(z) = log(1 + K|z − ζ |2 ). Then for each
w ∈ ℂn we have
11.5 Weighted spaces of entire functions | 307
K K
2 2
|w|2 ≤ i𝜕𝜕g(w, w)(z) ≤ |w|2 , (11.29)
(1 + K|z − ζ | ) 1 + K|z − ζ |2
𝜕2 g K 2 (z j − ζ j )(zk − ζk ) Kδjk
(z) = − 2 2
+
𝜕zj 𝜕z k (1 + K|z − ζ | ) 1 + K|z − ζ |2
K
= [(1 + K|z − ζ |2 )δjk − K(z j − ζ j )(zk − ζk )].
(1 + K|z − ζ |2 )2
This implies
K
i𝜕𝜕g(w, w)(z) = [(1 + K|z − ζ |2 )|w|2 − K|(w, z − ζ )|2 ],
(1 + K|z − ζ |2 )2
and hence
K
[(1 + K|z − ζ |2 )|w|2 − K|w|2 |z − ζ |2 ]
(1 + K|z − ζ |2 )2
K
≤ i𝜕𝜕g(w, w)(z) ≤ |w|2 .
1 + K|z − ζ |2
We are now able to show that weighted spaces of entire functions are of infinite
dimension if the weight function has an appropriate behavior at infinity.
Theorem 11.20. Let W ∶ ℂn ⟶ ℝ be a 𝒞∞ function and let μ(z) denote the smallest
eigenvalue of the Levi matrix
𝜕2 W(z) n
i𝜕𝜕W(z) = ( ) .
𝜕zj 𝜕zk j,k=1
Suppose that
Then the Hilbert space A2 (ℂn , e−W ) of all entire functions f such that
is of infinite dimension.
Proof. Assumption (11.30) implies that there exists a constant K > 0 such that
for all z, w ∈ ℂn , and that i𝜕𝜕W(z) is strictly positive for large |z|.
308 | 11 The twisted 𝜕-complex and Schrödinger operators
4K
i𝜕𝜕g(w, w)(z) ≥ |w|2 ,
(1 + 4K|z − ζ |2 )2
Since i𝜕𝜕W(w, w)(z) is negative at most on a compact set in ℂn , there exist finitely
many points ζ1 , … , ζM ∈ ℂn such that this compact set is covered by the balls
{z ∶ |z − ζl | < 1/√4K}. Hence
M
φ(z)
̃ ∶= W(z) + ∑ gl (z)
l=1
Let μ̃ 0 ∶= inf{μ(z)
̃ ∶ |z| ≤ R}. Then μ̃ 0 > 0. Set
μ̃ 0
κ=
2(N + M)
and
M
φ(z) ∶= W(z) + ∑ gl (z) − (N + M) log(1 + κ|z|2 ).
l=1
(N + M)κ
i𝜕𝜕φ(w, w)(z) ≥ |w|2 (μ(z)
̃ − ).
1 + κ|z|2
(N + M)κ (N + M)μ̃ 0 μ̃ 0
μ(z)
̃ − 2
≥ μ̃ 0 − (N + M)κ = μ̃ 0 − = > 0,
1 + κ|z| 2(N + M) 2
11.5 Weighted spaces of entire functions | 309
(N + M)κ N + M + 1 N + M 1
μ(z)
̃ − ≥ − = 2,
1 + κ|z|2 |z|2 |z|2 |z|
ℂn
∏M 2
l=1 (1 + 4K|z − ζl | )
|f (z)|2 (1 + |z|2 )N e−φ(z) dλ(z)
̃
=∫ 2 N+M
ℂn (1 + κ|z| )
(1 + |z|2 )N ∏M 2
l=1 (1 + 4K|z − ζl | )
≤ sup { }
z∈ℂn (1 + κ|z|2 )N+M
× ∫ |f (z)|2 (1 + |z|2 )−3n e−φ(z) dλ(z)
ℂn
< ∞.
Hence fp ∈ A2 (ℂn , e−W ) for any polynomial p of degree < Ñ , and since N = Ñ + 3n
was arbitrary, we are done.
∞ ∞ 2 2 4
∞ ∞ 2 2 4
∫ ∫ r22k e−(r1 r2 +r2 ) r1 r2 dr1 dr2 = ∫ (∫ r1 r22 e−r1 r2 dr1 )r22k−1 e−r2 dr2
0 0 0 0
1 ∞ ∞ 4 1 ∞ 4
= ∫ ( ∫ e−s ds)r22k−1 e−r2 dr2 = ∫ r22k−1 e−r2 dr2 < ∞.
0 2 0 2 0
|w|2 zw
i𝜕𝜕φ = ( )
wz |z|2 + 4|w|2
1
μφ (z, w) = (5|w|2 + |z|2 − (9|w|4 + 10|z|2 |w|2 + |z|4 )1/2 )
2
16|w|4
= ,
2(5|w| + |z| + (9|w|4 + 10|z|2 |w|2 + |z|4 )1/2 )
2 2
310 | 11 The twisted 𝜕-complex and Schrödinger operators
hence
RT ∶ ρ(T) ⟶ ℒ(H).
Theorem 11.22. Let T ∈ ℒ(H). Then the spectrum σ(T) of T is a compact subset of ℂ
and |λ| ≤ ‖T‖, for every λ ∈ σ(T).
Let μ ∈ ℂ with |μ| < α. We will show that (λ + μ)I − T has a bounded inverse. Then we
proved that ρ(T) is open. We have
(λ + μ)I − T = λI − T + μI
= (λI − T)[I + μ(λI − T)−1 ]
= (λI − T)(I + μRT (λ)).
Formally,
∞
(I + μRT (λ))−1 = I + ∑ (−1)k (μRT (λ))k ,
k=1
If η ∈ ℂ with |η| > ‖T‖, then I − T/η has a bounded inverse, since
∞
(I − T/η)−1 = I + ∑ (T/η)k .
k=1
This implies that ηI − T has a bounded inverse. Hence, if λ ∈ σ(T), then |λ| ≤ ‖T‖, and
σ(T) is a bounded set.
The resolvent has the following properties:
Theorem 11.24. Let T be a bounded self-adjoint operator. Then σ(T) is real and is con-
tained in [m, M], where
(Tu, u) (Tu, u)
m = inf and M = sup .
u≠0 (u, u) u≠0 (u, u)
Proof. If ℑλ ≠ 0, then
for all u ∈ H, where we used the fact that (Tu, u) = (u, Tu) is real.
This implies that T − λI is injective and has closed range. As
This follows also from Theorem 10.35, once one has observed that
Hence, the spectrum is real. If λ is real and λ > M, we can apply Theorem 10.35 for the
sesquilinear form
We get
In particular,
Now let (un )n be a sequence in H such that ‖un ‖ = 1 for each n ∈ ℕ and (Tun , un ) → M
as n → ∞. By (11.35),
Corollary 11.25. Let T ∈ ℒ(H) be a self-adjoint operator such that σ(T) = {0}. Then
T = 0.
Definition 11.26. The resolvent set of a linear operator T ∶ dom(T) ⟶ H is the set
of all λ ∈ ℂ such that λI − T is an injective mapping of dom(T) onto H whose inverse
belongs to ℒ(H). The spectrum σ(T) of T is the complement of the resolvent set of T.
11.6 Spectral analysis of self-adjoint operators | 313
Lemma 11.27. If the spectrum σ(T) of an operator T does not coincide with the whole
complex plane ℂ, then T must be a closed operator. The spectrum of a linear operator
is always closed. Moreover, if ζ ∉ σ(T) and c ∶= ‖RT (ζ )‖ = ‖(ζI − T)−1 ‖, then the spectrum
σ(T) does not intersect the ball {w ∈ ℂ ∶ |ζ − w| < c−1 }. The resolvent operator RT is a
holomorphic operator-valued function.
Proof. For ζ ∉ σ(T) let S = (ζI − T)−1 which is a bounded operator. Let xn ∈ dom(T) with
limn→∞ xn = x and limn→∞ Txn = y and set un = (ζI − T)xn . Then
therefore
which is a bounded operator. Now Corollary 11.25 gives that T −1 = 0, which contradicts
that T ∘ T −1 = I.
E ∶ M ⟶ ℒ(H)
of M to the algebra ℒ(H) of bounded linear operators on H with the following proper-
ties:
(a) E(∅) = 0 (zero-operator), E(Ω) = I (identity on H).
(b) For each ω ∈ M the image E(ω) is an orthogonal projection on H.
(c) E(ω′ ∩ ω″ ) = E(ω′ )E(ω″ ).
314 | 11 The twisted 𝜕-complex and Schrödinger operators
is a complex measure on M.
We will also need a more detailed study of the spectrum of a self-adjoint operator.
(iii) There exists a complete orthonormal system of eigenvectors {xn }n of T with corre-
sponding eigenvalues μn ≥ 0 which converge to +∞ as n → ∞.
Theorem 11.34. Let T be a positive, self-adjoint operator, which is also injective. Sup-
pose that the bottom of the essential spectrum satisfies inf σe (T) > 0. Then T has a
bounded inverse, i.e. T −1 ∈ ℒ(H).
This is also a consequence of the spectral theorem; see [70, Theorem 8.17].
is a 1-form, and
n 2
𝜕
ΔA = ∑( − iAj ) .
j=1 𝜕xj
The 2-form
𝜕Ak 𝜕Aj
B = dA = ∑( − ) dxj ∧ dxk
j<k 𝜕xj 𝜕xk
is the magnetic field, which is responsible for specific spectral properties of the opera-
tor H(A, V), as will be seen later. It is clear that the operators H(A, V) are second order
elliptic partial differential operators.
Under appropriate assumptions on A and V , the operator H(A, V) acts as an un-
bounded self-adjoint operator on L2 (ℝn ). In many aspects of the spectral theory of
the Schrödinger operator with magnetic field H(A, V), it is convenient to compare this
operator with the ordinary Schrödinger operator
H(0, V) = −Δ + V,
n
− ΔA = ∑ Xj2 , (11.38)
j=1
316 | 11 The twisted 𝜕-complex and Schrödinger operators
n
and for u, v ∈ 𝒞∞
0 (ℝ ) we have (Xj u, v) = (u, Xj v), j = 1, … , n and
n
(−ΔA u, u) = ∑ ‖Xj u‖2 . (11.39)
j=1
V(x) ≥ −C ∀x ∈ ℝn ,
n
where C > 0 is a positive constant. Let dom(H(A, V)) = 𝒞∞
0 (ℝ ). Then H(A, V) is a sym-
2 n
metric, semibounded operator on L (ℝ ).
n
Proof. For u ∈ 𝒞∞
0 (ℝ ), we have
≥ −C‖u‖2 .
Theorem 11.36. Let H(A, V) be as in Theorem 11.35. Then H(A, V) admits a self-adjoint
extension.
Proof. Define
Remark 11.37. The operator H(A, V) of Theorem 11.36 is essentially self-adjoint (see
n
[34]) on 𝒞∞
0 (ℝ ). This implies that its self-adjoint extension is uniquely determined.
−ΔA + B,
where the 2-form dA is given by dA = B dx ∧ dy. So, in this case, the Schrödinger oper-
ator
−ΔA + B = H(A, B)
Dv = g,
where
𝜕 φ/2
D = e−φ/2 e . (11.40)
𝜕z
u is the minimal solution to the 𝜕-equation in L2 (ℂ, e−φ ), if and only if v is the solution
to Dv = g which is minimal in L2 (ℂ).
The formal adjoint of D is D = −eφ/2 𝜕z e . We define
∗ 𝜕 −φ/2
and likewise for D . Then D and D are closed unbounded linear operators from L2 (ℂ)
∗ ∗
and we define D D as D ∘ D on this domain. Any function of the form eφ/2 g, with
∗ ∗
g ∈ 𝒞20 (ℂ) belongs to dom(D D ) and hence dom(D D ) is dense in L2 (ℂ). Since
∗ ∗
𝜕 1 𝜕φ 𝜕 1 𝜕φ
D= and D =−
∗
+ + ,
𝜕z 2 𝜕z 𝜕z 2 𝜕z
we see that
𝜕2 1 𝜕φ 𝜕 1 𝜕φ 𝜕 1 𝜕φ 2 1 𝜕2 φ
𝒮 = DD = − (11.41)
∗
− + + | | + .
𝜕z𝜕z 2 𝜕z 𝜕z 2 𝜕z 𝜕z 4 𝜕z 2 𝜕z𝜕z
It follows that
𝜕2 𝜕2 i 𝜕φ 𝜕 i 𝜕φ 𝜕 1 𝜕φ 2 𝜕φ 2 1
4𝒮 = −( + ) + − + (( ) + ( ) ) + Δφ,
𝜕x2 𝜕y2 2 𝜕x 𝜕y 2 𝜕y 𝜕x 4 𝜕x 𝜕y 2
1
B(x, y) = Δφ(x, y).
2
Both D D and D D are non-negative, self-adjoint operators, see Lemma 10.28 and
∗ ∗
Lemma 10.29. By Theorem 11.36 and Remark 11.37, we now know that the operator 𝒮
with its domain described above is the uniquely determined self-adjoint extension of
the Schrödinger operator with magnetic field 41 (−ΔA + B).
Since 4D D = −ΔA + 21 Δφ, it follows that ((−ΔA + 21 Δφ)f , f ) ≥ 0, for f ∈ 𝒞20 (ℂ). Sim-
∗
1
−2ΔA ≥ −ΔA + Δφ ≥ −ΔA .
2
It follows that
and
∗
□φ = 𝜕 𝜕φ ,
𝜕 𝜕
L2 (ℂn , e−φ ) ⟶ L2(0,1) (ℂn , e−φ ) ⟶ L2(0,2) (ℂn , e−φ ).
It is easy to see that 𝜕u = f for u ∈ L2 (ℂn , e−φ ) and f ∈ L2(0,1) (ℂn , e−φ ), if and only if
D1 v = g, where v = ue−φ/2 and g = fe−φ/2 . It is also clear that the necessary condition
𝜕f = 0 for solvability holds, if and only if D2 g = 0 holds. Here
n
𝜕gj 1 𝜕φ
D2 g = ∑ ( + g ) dz k ∧ dz j
j,k=1 𝜕z k 2 𝜕z k j
and
n
1 𝜕φ 𝜕hkj
D2 h = ∑ ( hkj − ) dz j
∗
= D1 D1 ,
∗
Δ(0,0)
φ
(11.48)
= D1 D1 + D2 D2 .
∗ ∗
Δ(0,1)
φ
and
n n
1 𝜕φ 𝜕gk 1 𝜕φ 𝜕φ
(D1 D1 + D2 D2 )g = ∑ [∑( g
∗ ∗
+
k=1 j=1 2 𝜕zj 𝜕z j 4 𝜕zj 𝜕z j k
1 𝜕φ 𝜕gk 1 𝜕2 φ 𝜕2 gk 𝜕2 φ
− − gk − + g )] dz k . (11.49)
2 𝜕z j 𝜕zj 2 𝜕zj 𝜕z j 𝜕zj 𝜕z j 𝜕zj 𝜕z k j
k=1|J|=q
and
n
Dq h = ∑ ∑ Zk∗ (hJ ) dz k ⌋ dz J ,
∗ ′
k=1|J|=q
320 | 11 The twisted 𝜕-complex and Schrödinger operators
⟨α, dz k ⌋ dz J ⟩ = ⟨dz k ∧ α, dz J ⟩
for q = 1, … , n − 1.
The general D-complex has the form
Dq Dq+1
L2(0,q−1) (ℂn ) ⟶ L2(0,q) (ℂn ) ⟶ L2(0,q+1) (ℂn ).
⟵
∗
⟵
∗
Dq Dq+1
It follows that
We remark that
n n
Dq h = ∑ ∑ Zk∗ (hJ ) dz k ⌋ dz J = ∑ ∑ Zk∗ (hkK ) dz K .
∗ ′ ′
φ v = D1 D1 v = ∑ Zj Zj (v),
∗
Δ(0,0) ∗
j=1
φ g = (D1 D1 + D2 D2 )g
∗ ∗
Δ(0,1)
n
= ∑ {Zj (Zk∗ (gℓ )) dz j ∧ (dz k ⌋ dz ℓ ) + Zk∗ (Zj (gℓ )) dz k ⌋ (dz j ∧ dz ℓ )}
j,k,ℓ=1
n
= ∑ {Zk∗ (Zj (gℓ ))(dz j ∧ (dz k ⌋ dz ℓ ) + dz k ⌋ (dz j ∧ dz ℓ ))
j,k,ℓ=1
= (Δ(0,0)
φ ⊗ I)g + Mφ g,
α ⌋ (a ∧ b) = (α ⌋ a) ∧ b − a ∧ (α ⌋ b),
11.7 Real differential operators | 321
which implies
dz j ∧ (dz k ⌋ dz ℓ ) + dz k ⌋ (dz j ∧ dz ℓ )
= dz j ∧ (dz k ⌋ dz ℓ ) + (dz k ⌋ dz j ) ∧ dz ℓ − dz j ∧ (dz k ⌋ dz ℓ )
= (dz k ⌋ dz j ) ∧ dz ℓ = δkℓ dz ℓ ,
and
n
(Δ(0,0)
φ ⊗ I)g = ∑ Δ(0,0)
φ gk dz k .
k=1
We will now explain how to prove the following theorem, for all details see [31, 30].
where Nφ = □−1φ .
For n = 1, condition (11.51) reads as
φ g = (Δφ
Δ(0,1) ⊗ I)g + Mφ g, (11.54)
(0,0)
and, as Δ(0,0)
φ is a positive operator, obtain
φ g, g) ≥ (Mφ g, g),
(Δ(0,1) (11.55)
where the supremum is taken over all compact subsets K ⊂ ℂn . Using (11.56) together
with the assumption (11.51), we now conclude that the bottom of the essential spec-
trum of Δ(0,1)
φ is strictly positive and Theorem 11.34 implies that Δ(0,1)
φ has a bounded
inverse.
Remark 11.39. Under the same assumptions as before, we have that also Nφ exists
and is bounded, Nφ is self-adjoint and positive. So we get from Theorem 11.31 that Nφ
has a bounded square root Nφ1/2 . Let u ∈ dom(□φ ). Then there exists a w ∈ L2(0,1) (ℂn , e−φ )
such that Nφ w = u. Define v ∶= Nφ1/2 w. It follows that Nφ1/2 v = u, and we have
where
0 1 0 −i
σ1 = ( ), σ2 = ( ).
1 0 i 0
0 𝒜1 − i𝒜2
𝒟=( ).
𝒜1 + i𝒜2 0
We remark that i(𝒜2 𝒜1 − 𝒜1 𝒜2 ) = B and hence it turns out that the square of 𝒟 is
diagonal with the Pauli operators P± on the diagonal:
where
2 2
𝜕 𝜕
P± = (−i − A1 ) + (−i − A2 ) ± B = −ΔA ± B.
𝜕x 𝜕y
By Lemma 10.28 and Lemma 10.29, the Pauli operators P± are non-negative and self-
adjoint.
It follows that
4𝒮 = P+
4Δ(0,0)
φ = P− . (11.59)
Δφ(z) → +∞
= (D D e f,e f ), (11.63)
∗ −φ/2 −φ/2
(11.31).
Δφ ∈ B2 (ℝ2 ). (11.65)
That (11.66) is necessary for compactness follows from a result of Iwatsuka. The
sufficiency of (11.66) is derived form the diamagnetic inequality (see [30]) and a special
form of the Fefferman–Phong inequality (see [31]).
We return to the Dirac and Pauli operators related with the weight function φ:
𝜕 𝜕
𝒟 = (−i − A1 )σ1 + (−i − A2 )σ2 ,
𝜕x 𝜕y
where A1 = − 21 𝜕φ
𝜕y
, A2 = 1 𝜕φ
2 𝜕x
and
0 1 0 −i
σ1 = ( ), σ2 = ( ).
1 0 i 0
P− 0
𝒟2 = ( ),
0 P+
where
2 2
𝜕 𝜕
P± = (−i − A1 ) + (−i − A2 ) ± B = −ΔA ± B,
𝜕x 𝜕y
and B = 21 Δφ.
for each z ∈ ℂ and for each r > 0; in particular μ(ℂ) = ∞, unless μ(ℂ) = 0 (see [67]).
For example, if p(z, z) is a polynomial on ℂ of degree d, then
1
dμ(z) = |p(z, z)|a dλ(z), a>−
d
is of infinite dimension.
μ(ℂ) = ∫ Δφ dλ = ∞. (11.69)
ℂ
Let Kφ (z, w) denote the Bergman kernel of A2 (ℂ, e−φ ). So, if f ∈ A2 (ℂ, e−φ ), then
We claim that
∞
∫ Kφ (z, z)e−φ(z) dλ(z) = ∑ ‖fk ‖2φ , (11.71)
ℂ k=1
for any complete orthonormal basis (fk )k of A2 (ℂ, e−φ ), finite or infinite.
The partial sums
N
FN (z) = ∑ |fk (z)|2
k=1
form a pointwise non-decreasing sequence: FN (z) ≤ FN+1 (z), for each z ∈ ℂ and for
each N ∈ ℕ. In addition, we have
∞
lim FN (z) = ∑ fk (z)fk (z) = Kφ (z, z),
N→∞
k=1
11.10 Exercises | 327
Now we use the following pointwise inequality for the Bergman kernel which is due
to Marzo and Ortega-Cerda [54]: there exists a constant C > 0 such that
for each z ∈ ℂ, where μ(Dρ(z) (z)) = 1 (see [54]), to conclude that A2 (ℂ, e−φ ) is of infinite
dimension.
Proof. Since
𝒟2 has compact resolvent, if and only if 𝒟 has compact resolvent. Suppose that 𝒟
has compact resolvent. Since
P− 0
𝒟2 = ( ),
0 P+
P− = 4D D = 4e−φ/2 𝜕φ 𝜕eφ/2
∗ ∗
and that P− is a non-negative self-adjoint operator. It follows from Theorem 11.45 that
the space of entire functions A2 (ℂ, e−φ ) is of infinite dimension. This means that 0
belongs to the essential spectrum of P− . Hence, by Theorem 11.33, P− fails to have
compact resolvent, and we arrive at a contradiction.
11.10 Exercises
165. Let H be a Hilbert space and A ∈ ℒ(H) a compact self-adjoint operator. Suppose
that μ ∉ σ(A) ∪ {0}. Take an arbitrary y ∈ H and show that the equation
(A − μI)x = y
166. Let A, A′ ∈ 𝒞2 (ℝn , ℝn ) be such that dA = dA′ . Suppose that V ∈ L2loc (ℝn ) and
V ≥ 0. Show that
where g ∈ 𝒞1 (ℝn ) and that σ(H(A, V)) = σ(H(A′ , V)). This is called the gauge in-
variance of the spectrum of H(A, V).
Hint: use the Poincaré lemma to show that A − A′ = dg.
167. Consider the weighted space L2 (ℂ, e−φ ), where φ(z) = |z|2 . Show that
𝜕2 u 𝜕u
□φ u = 𝜕 𝜕φ u = − +z + u,
∗
𝜕z𝜕z 𝜕z
for u ∈ 𝒞∞0 (ℂ).
2
168. Show that the Bergman space A2 (ℂ, e−|z| ) is a subspace of the eigenspace to the
eigenvalue 1 of the operator □φ , where φ(z) = |z|2 , and that 1 ∈ σe (□φ ).
169. Consider the weighted space L2 (ℂ, e−φ ), where φ(z) = |z|α , α > 0 and set β = α/2.
Let uk (z) = z k , k = 0, 1, …. Show that
π k 1
‖uk ‖2φ = Γ( + )
β β β
and
k
‖𝜕φ uk ‖2φ = πβΓ( + 1).
∗
β
170. Let φ(z1 , z2 ) = φ1 (z1 ) + φ2 (z2 ). Consider the space L2 (ℂ2 , e−φ ). Show that
𝜕2 u1 𝜕2 u1 𝜕φ 𝜕u 𝜕φ 𝜕u 𝜕2 φ1
□φ,1 u = (− − + 1 1+ 2 1+ u ) dz 1
𝜕z1 𝜕z 1 𝜕z2 𝜕z 2 𝜕z1 𝜕z 1 𝜕z2 𝜕z 2 𝜕z1 𝜕z 1 1
𝜕2 u2 𝜕2 u2 𝜕φ 𝜕u 𝜕φ 𝜕u 𝜕2 φ2
+ (− − + 1 2+ 2 2+ u ) dz 2 ,
𝜕z1 𝜕z 1 𝜕z2 𝜕z 2 𝜕z1 𝜕z 1 𝜕z2 𝜕z 2 𝜕z2 𝜕z 2 2
2
for u = u1 dz 1 + u2 dz 2 and u1 , u2 ∈ 𝒞∞
0 (ℂ ), and that □φ,2 V equals to
𝜕2 v 𝜕2 v 𝜕φ 𝜕v 𝜕φ2 𝜕v 𝜕2 φ1 𝜕2 φ2
(− − + 1 + + v+ v) dz 1 ∧ dz 2 ,
𝜕z1 𝜕z 1 𝜕z2 𝜕z 2 𝜕z1 𝜕z 1 𝜕z2 𝜕z 2 𝜕z1 𝜕z 1 𝜕z2 𝜕z 2
2
where V = v dz 1 ∧ dz 2 and v ∈ 𝒞∞ 0 (ℂ ).
171. Let φ(z1 , z2 ) = |z1 |2 + |z2 |2 and consider the corresponding 𝜕-Neumann operator
1 1 1
φ(x, y) = (x 2 + y2 ) + (x4 y2 + x 2 y4 ) − (x6 + y6 )
4 24 360
11.11 Notes
Theorem 11.20 is due to I. Shigekawa [66]. A. Iwatsuka [42] and J. Avron, I. Herbst and
B. Simon [6] were the first, who obtained results on the problem of compact resolvent
of Schrödinger operators with magnetic field. Basic facts about Schrödinger operators
with magnetic fields can be found in [16, 34, 35, 37]. The relationship to the 𝜕-equation
appears in [9, 28, 31].
M. Christ [14] initiated the study of weighted spaces of entire functions, where the
weight φ has the property that dμ = Δφ dλ is a non-trivial doubling measure.
Witten Laplacians, as they appear in the context of semiclassical analysis, are
studied in [33, 36, 45]. The Witten Laplacians were introduced by E. Witten on a com-
pact manifold in connection with a Morse function, which is a function with non-
degenerate critical points. The main idea of E. Witten was that the dimension of the
kernels of these Laplacians, which are related by the Hodge theory to the Betti num-
bers, can be estimated from above by the dimension of the eigenspace associated with
the small eigenvalues of these operators (see [72, 73, 33, 38, 39, 45]). A thorough treat-
ment of Dirac and Pauli operators can be found in [16, 37, 69, 62].
Bibliography
[1] M. J. Ablowitz and A. S. Fokas. Complex Analysis, Introduction and Applications. Cambridge
Texts in Applied Mathematics. Cambridge University Press, Cambridge, 1997.
[2] R. A. Adams and J. J. F. Fournier. Sobolev Spaces, volume 140 of Pure and Applied Mathematics.
Academic Press, San Diego, 2006.
[3] S. Agmon. Lectures on Exponential Decay of Solutions of Second-Order Elliptic Equations:
Bounds on Eigenfunctions of N-Body Schrodinger Operations. Princeton University Press,
Princeton, 2014.
[4] L. Ahlfors. Complex Analysis. McGraw–Hill, Princeton, NJ, 1979.
[5] N. Aronszajn. A unique continuation theorem for solutions of elliptic partial differential
equations or inequalities of second order. J. Math. Pures Appl., 36:235–247, 1957.
[6] J. Avron, I. Herbst, and B. Simon. Schrödinger operators with magnetic fields I. Duke Math. J.,
45:847–883, 1978.
[7] St. Bell. The Cauchy Transform, Potential Theory, and Conformal Mapping. Studies in Advanced
Mathematics. CRC Press, Boca Raton, 1992.
[8] F. Berger. The Inhomogeneous Cauchy–Riemann Equation. Bachelorarbeit, Universitaet Wien,
2012.
[9] B. Berndtsson. 𝜕 and Schrödinger operators. Math. Z., 221:401–413, 1996.
[10] H. Brezis. Analyse fonctionnelle, Théorie et applications. Masson, Paris, 1983.
[11] J. Bruna and J. Cufí. Complex Analysis. European Mathematical Society, Zürich, 2013.
[12] R. B. Burckel. An Introduction to Classical Complex Analysis. Birkhäuser Verlag, Basel, 1979.
[13] S.-C. Chen and M.-C. Shaw. Partial Differential Equations in Several Complex Variables,
volume 19 of Studies in Advanced Mathematics. Amer. Math. Soc., Providence, RI, 2001.
[14] M. Christ. On the 𝜕 equation in weighted L2 norms in ℂ1 . J. Geom. Anal., 1:193–230, 1991.
[15] J. B. Conway. Functions of One Complex Variable I. Springer Verlag, New York, 1978.
[16] H. L. Cycon, R. G. Froese, W. Kirsch, and B. Simon. Schrödinger Operators with Applications to
Quantum Mechanics and Global Geometry. Text and Monographs in Physics. Springer Verlag,
Berlin, 1987.
[17] J. P. D’Angelo. Inequalities from Complex Analysis, volume 28 of Carus Mathematical
Monographs. Mathematical Association of America, Washington, DC, 2002.
[18] E. B. Davies. Spectral Theory and Differential Operators, volume 42 of Cambridge Studies in
Advanced Mathematics. Cambridge University Press, Cambridge, 1995.
[19] J. Dieudonné. Foundations of Modern Analysis. Read Books Ltd, Redditch, 2013.
[20] J. Elstrodt. Maß- und Integrationstheorie. Springer Verlag, Berlin, 1996.
[21] W. Fischer and I. Lieb. Funktionentheorie. Vieweg Verlag, Braunschweig, 1992.
[22] W. Fischer and I. Lieb. Ausgewählte Kapitel aus der Funktionentheorie. Vieweg Verlag,
Braunschweig, 1998.
[23] W. Fischer and I. Lieb. A Course in Complex Analysis. Vieweg+Teubner, Wiesbaden, 2012.
[24] G. B. Folland. Real Analysis, Modern Techniques and Their Applications. John Wiley & Sons,
New York, 1984.
[25] G. B. Folland. Introduction to Partial Differential Equations. Princeton University Press,
Princeton, 1995.
[26] T. W. Gamelin. Complex Analysis. Undergraduate Texts in Mathematics. Springer Verlag,
New York, 2001.
[27] E. Goursat. Sur la définition générale des fonctions analytiques, d’après Cauchy. Trans. Amer.
Math. Soc., 1(1):14–16, 1900.
https://doi.org/10.1515/9783110417241-012
332 | Bibliography
[28] F. Haslinger. Schrödinger operators with magnetic fields and the 𝜕-equation. J. Math. Kyoto
Univ., 46:249–257, 2006.
[29] F. Haslinger. The 𝜕-Neumann operator and commutators between multiplication operators and
the Bergman projection. Czechoslov. Math. J., 58:1247–1256, 2008.
[30] F. Haslinger. The 𝜕-Neumann Problem and Schrödinger Operators, volume 59 of De Gruyter
Expositions in Mathematics. Walter De Gruyter, Berlin, 2014.
[31] F. Haslinger and B. Helffer. Compactness of the solution operator to 𝜕 in weighted L2 -spaces.
J. Funct. Anal., 243:679–697, 2007.
[32] F. Haslinger and B. Lamel. Spectral properties of the canonical solution operator to 𝜕. J. Funct.
Anal., 255:13–24, 2008.
[33] B. Helffer. Semiclassical Analysis, Witten Laplacians, and Statistical Mechanics. World
Scientific, Singapore, 2002.
[34] B. Helffer. Spectral Theory and Its Applications, volume 139 of Cambridge Studies in Advanced
Mathematics. Cambridge University Press, Cambridge, 2013.
[35] B. Helffer and A. Mohamed. Caractérisation du spectre essentiel de l’opérateur de Schrödinger
avec un champ magnétique. Ann. Inst. Fourier (Grenoble), 38:95–112, 1988.
[36] B. Helffer and F. Nier. Hypoelliptic Estimates and Spectral Theory for Fokker–Planck Operators
and Witten Laplacians, volume 1862 of Lecture Notes in Mathematics. Springer Verlag, Berlin,
2005.
[37] B. Helffer, J. Nourrigat, and X. P. Wang. Sur le spectre de l’équation de Dirac (dans ℝ2 ou ℝ3 )
avec champ magnétique. Ann. Sci. Éc. Norm. Supér., 22:515–533, 1989.
[38] B. Helffer and J. Sjöstrand. Puits multiples en limite semi-classique IV – Etude du complex de
Witten. Comm. Partial Differential Equations, 10:245–340, 1985.
[39] B. Helffer and J. Sjöstrand. A proof of the Bott inequalities. In: Volume in Honor of M. Sato,
Algebraic Analysis, Vol. 1, pages 171–183. Academic Press, San Diego, 1988.
[40] P. Henrici. Applied and Computational Complex Analysis I, II, III. John Wiley & Sons, New York,
London, 1985.
[41] L. Hörmander. An Introduction to Complex Analysis in Several Variables. North-Holland,
Amsterdam, 1990.
[42] A. Iwatsuka. Magnetic Schrödinger operators with compact resolvent. J. Math. Kyoto Univ.,
26:357–374, 1986.
[43] M. Jarnicki and P. Pflug. Continuous Nowhere Differentiable Functions. Springer Monographs in
Mathematics, 2015.
[44] M. Jarnicki and P. Pflug. Extensions of Holomorphic Functions. volume 34 of De Gruyter
Expositions in Mathematics. Walter De Gruyter, Berlin, 2000.
[45] J. Johnsen. On the spectral properties of Witten Laplacians, their range projections and
Brascamp–Lieb’s inequality. Integral Equations Operator Theory, 36:288–324, 2000.
[46] G. A. Jones and D. Singerman. Complex Functions, an Algebraic and Geometric Viewpoint.
Cambridge University Press, Cambridge, 1987.
[47] J. J. Kohn. Harmonic integrals on strongly pseudoconvex manifolds, i. Ann. of Math.,
78:112–148, 1963.
[48] J. J. Kohn. Harmonic integrals on strongly pseudoconvex manifolds, ii. Ann. of Math.,
79:450–472, 1964.
[49] J. J. Kohn and L. Nirenberg. Non-coercive boundary value problems. Comm. Pure Appl. Math.,
18:443–492, 1965.
[50] G. Köthe. Topological Vector Spaces. Springer, Berlin, Heidelberg, 1983.
[51] S. G. Krantz. Function Theory of Several Complex Variables. 2nd edn., Wadsworth &
Brooks/Cole, Pacific Grove, California, 1992.
Bibliography | 333
[52] B. Luecking and L. A. Rubel. Complex Analysis, a Functional Analysis Approach. Springer
Verlag, New York, 1981.
[53] A. I. Markushevich. Theory of Functions. Chelsea Publishing Company, New York, 1977.
[54] J. Marzo and J. Ortega-Cerda. Pointwise estimates for the Bergman kernel of the weighted Fock
space. J. Geom. Anal., 19:890–910, 2009.
[55] R. Meise and D. Vogt. Einführung in die Funktionalanalysis. Vieweg Studium, 1992.
[56] D. S. Mitrinovic and J. D. Keckic. The Cauchy Method of Residues: Theory and Applications,
volume 9. Springer Science & Business Media, Dordrecht, 1984.
[57] R. Narasimhan. Complex Analysis in One Variable. Birkhäuser Verlag, Basel, 1985.
[58] A. Pringsheim. Über den Goursat’schen Beweis des Cauchy’schen Integralsatzes. Trans. Amer.
Math. Soc., 2(4):413–421, 1901.
[59] R. M. Range. Holomorphic Functions and Integral Representations in Several Complex
Variables. Springer Verlag, New York, 1986.
[60] R. Remmert. Funktionentheorie I. Springer Verlag, Berlin, Heidelberg, 1995.
[61] R. Remmert. Funktionentheorie II. Springer Verlag, Berlin, Heidelberg, 1996.
[62] G. Rozenblum and N. Shirokov. Infiniteness of zero modes for the Pauli operator with singular
magnetic field. J. Funct. Anal., 233:135–172, 2006.
[63] W. Rudin. Real and Complex Analysis. McGraw–Hill, New York, 1987.
[64] N. Salinas. Noncompactness of the 𝜕-Neumann problem and Toeplitz C ∗ -algebras. In: Several
Complex Variables and Complex Geometry, Part 3 (Santa Cruz, CA, 1989), volume 52 of
Proceedings of Symposia in Pure Mathematics, pages 329–334. Amer. Math. Soc., Providence,
RI, 1991.
[65] N. Salinas, A. Sheu, and H. Upmeier. Toeplitz operators on pseudoconvex domains and
foliation C ∗ -algebras. Ann. of Math., 130:531–565, 1989.
[66] I. Shigekawa, Spectral properties of Schrödinger operators with magnetic fields for a spin 1/2
particle. J. Funct. Anal., 101:255–285, 1991.
[67] E. Stein. Harmonic Analysis: Real-Variable Methods, Orthogonality, and Oscillatory Integrals.
Princeton University Press, Princeton, NJ, 1993.
[68] E. J. Straube. The L2 -Sobolev Theory of the 𝜕-Neumann Problem. ESI Lectures in Mathematics
and Physics. EMS, Zürich, 2010.
[69] B. Thaller. The Dirac Equation. Texts and Monographs in Physics. Springer Verlag, Berlin, 1991.
[70] J. Weidmann. Lineare Operatoren in Hilberträumen, Teil I Grundlagen. B. G. Teubner, Leipzig,
2000.
[71] R. O. Wells. Differential Analysis on Complex Manifolds. Prentice–Hall, New York, 1973.
[72] E. Witten. Supersymmetry and Morse inequalities. J. Differential Geom., 17:661–692, 1982.
[73] E. Witten. Holomorphic Morse inequalities. Teubner-Texte Math., 70:318–333, 1984.
Index
A2 (Ω) 191 Cauchy–Riemann differential equation
Tp1,0 (bΩ) 187 – inhomogeneous 12
Δ(0,0)
φ 319 Cauchy–Riemann equations 7
Δ(0,q)
φ 320 Cayley transform 187
𝜕-Neumann operator 281, 298, 303 chain 70
𝒦(H1 , H2 ) 209 closable operator 239
ℒ(H1 , H2 ) 209 closed graph 268
𝒱-elliptic form 253 closed graph theorem 242, 243
𝜕-complex 269, 270, 297, 318 closed operator 239, 270, 313
□ 271, 281, 282 closed range 281
□φ 297 closed set 4
ℂ 3 closure 4
ℋ(U) 6 closure of an operator 239
cocycle 135
absolute value 2 cohomology-group 135
absolutely convex 231 compact exhaustion 60
adjoint operator 223, 239, 297 compact operator 209, 212, 223
analytic continuation 106, 108 compact resolvent 323, 327
analytic polyhedron 183 compact set 5
angle preserving 14 complete differential 66
approximation to the identity 263 complex conjugate 2
argument 2 complex differentiable function 6
Arzela–Ascoli Theorem 146 complex Laplacian 271, 297
automorphism 147
complex measure 314
complex velocity potential 157
Baire category theorem 242
conjugate harmonic function 155
balanced set 237
connected component 4
basic estimate 281
connected set 4
Bergman kernel 196, 200, 226
continuous form 253
Bergman projection 196
continuous function 5
Bergman space 191
continuous linear functional 119
Bernoulli numbers 97
convergent product 136
Bessel’s inequality 197, 215
convex domain 26
Bieberbach’s conjecture 150
convex set 126
biholomorphic mapping 81, 203
core of the operator 252, 322
boundary 4
CR-function 178
bounded family 145
curve 14
bounded subset 232
cycle 70
branching point 62
maximum principle 57, 173 plurisubharmonic function 185, 297, 302, 305,
meager set 242 306, 308, 309, 321
mean value property 48, 156, 162 pointwise convergence 16
Mergelyan’s Theorem 122 Poisson integral 158
meromorphic function 83 Poisson kernel 158
Minimum Principle 61 polar representation 2
Minkowski functional 231 polarization identity 205
Minkowski’s inequality 263 pole of order m 54
Mittag-Leffler’s Theorem 131 polydisc 170
Monodromy Theorem 108 power series 18
Montel’s Theorem 145 primitive function 24
Morera’s Theorem 50 principal argument 2
principal branch 31
principal part 81
neighborhood 4
product of complex numbers 1
Newton interpolation 101
pseudoconvex domain 274, 275, 280, 288, 292,
non-coercive system 273
296, 304–306
norm 117
punctured disc 53
normal family 144
normal operator 310 quadratic form 282
nowhere dense set 242
nuclear Fréchet space 234 radius of convergence 18
nuclear operator 217 real differentiable function 7
null-homologous cycle 73 real part 1
null-homotopic curve 73 reflection principle 96
regular boundary point 103
one-parameter family 73 relatively compact set 5
one-point compactification 3 relatively open set 4
open mapping 61 removable singularity 53
open set 4 reproducing property 196
open-mapping theorem 242, 246 residue 85
operator norm 209 residue theorem 85
order of a zero 50 resolution of the identity 313
orthogonal projection 193, 313 resolvent of an operator 310
orthonormal basis 197, 198 resolvent set 310, 312
restriction of a distribution 257
reverse Hölder class 324
paracompact set 112
Riemann mapping 204
parallelogram rule 193, 213 Riemann Mapping Theorem 148
parameter integral 64 Riemann sphere 3
Parseval’s equation 197 Riesz representation theorem 194
partial fractions decomposition 132 root function 32
partition of unity 111 rotation–dilation 5
path 21 Rouché’s Theorem 87
– inverse 21 Runge’s Theorem 121
pathwise connected set 4
Pauli operator 322, 325 s-number 216
peaking function 122 Schatten-class 216
Persson’s Lemma 304, 322 Schrödinger operator 291, 317
Pick’s Lemma 154 Schrödinger operator with magnetic field 315
338 | Index