1 December 2004
2000 Mathematics Subject Classification. Primary 32D15, 32D10.
Generalization of Drużkowski's and Gonchar's ”Edge-of-the-Wedge” Theorems
Peter Pflug
Viê ̣ t-Anh Nguyên
Peter Pflug, Carl von Ossietzky Universitat Oldenburg, Fachbereich Mathematik, Postfach 2503, D–26111, Oldenburg, Germany E-mail address : pflug@mathematik.uni-oldenburg.de Vie t-Anh Nguyen, Carl von Ossietzky Universitat Oldenburg, Fachbereich Mathematik, Postfach 2503, D–26111, Oldenburg, Germany E-mail address : nguyen@mathematik.uni-oldenburg.de
-
Abstract.
Let
be two open sets, let
(resp.
) be a subset of
(resp.
), and let
be the
-fold cross
Suppose in addition that
(resp.
) is locally rectifiable on
(resp.
) and that
and
are of positive one-dimensional Hausdorff measure. We determine the ”envelope of holomorphy”
of
in the sense that any function locally bounded on
measurable on
and separately holomorphic on
”extends” to a function holomorphic on the interior of
Generalizations of this result for an
-fold cross are also given.
1 Introduction
The ”Edge-of-the-Wedge” type theorems deal with the continuation of holomorphic functions of several complex variables. The first theorem was discovered and proved by N. N. Bogolyubov in 1956 in connection with quantum field theory and dispersion relations. Since then, there is a long list of papers dealing with this theorem and its generalizations under various assumptions (see [16] , [20] and the references therein).
Here we consider a one-sided version of the ”Edge-of-the-Wedge” type theorem in the spirit of the pioneer work of Malgrange–Zerner [21] . Epstein's survey article [3] gives a historical discussion and motivation for this version of an ”Edge-of-the-Wedge” theorem, as well as its natural connections with theorems on separate analyticity.
The first results in this direction are obtained by Komatsu [10] and Drużkowski [2] , but only for some special cases. Recently, Gonchar [5, 6] has proved a remarkable more general result for the one-dimensional case. In a recent work [14] , the authors are able to generalize Gonchar's result to the higher dimensional case.
However, in all these cases the hypotheses on the function to extend and its domain of definition are, in some sense, rather restrictive and strong. Therefore, the main goal of our work is to establish one-sided ”Edge-of-the-Wedge” type theorems in some more general one-dimensional cases with more optimal hypotheses. Perhaps, this will be the first step towards understanding the higher dimensional case and the manifold settings in its full generality.
This paper is organized as follows.
In Section 2 we gather some necessary notions and auxiliary results. This preparation will enable us, at the end of this section, to formulate the above mentioned results of Drużkowski and Gonchar and to discuss in more details the motivation for our work.
The statements of our main results as well as an outline of their proofs are given in Section 3.
The tools which are needed for the proof of the main results are developed in Sections 4, 5, and 8.
The proofs of the main results are given in Sections 6, 7, 9, and 10.
Section 11 gives examples showing the optimality of our results. Finally, we conclude the article with some remarks and open questions.
Our approach is based on our previous work [14] , the Gonchar–Carleman operator developed in [5, 6] , conformal mapping theory, and a thorough geometric study of harmonic measures.
Acknowledgment. The paper was written while the second author was visiting the Carl von Ossietzky Universität Oldenburg being supported by The Alexander von Humboldt Foundation. He wishes to express his gratitude to these organizations.
2 Preliminaries
In order to recall the classical one-sided versions of the ”Edge-of-the-Wedge” theorem and to discuss in more detail our motivation, we need to introduce some notation and terminology. In fact, we keep the main notation from the previous work [14] .
2.1 Planar domains with partly rectifiable boundary
We collect here some classical facts from the books by Goluzin, Koosis and Pommerenke ([4] , [11] , [13] ).
A Jordan curve is the image
of a continuous one-to-one map
where
The set
is said to be the interior of the Jordan curve. A closed Jordan curve is the image
of a continuous map
which is one-to-one in
and which satisfies
The map
is called a parametrization of
Moreover,
is said to be rectifiable if
where the supremum is taken over all possible positive integers
and sequences of values
such that
It is easy to see that this supremum is independent of the choice of a parametrization. It is called the length of
A (rectifiable) Jordan domain is a bounded domain in
whose boundary is a (rectifiable) closed Jordan curve.
Consider an open set
Then
is said to be locally rectifiable at a point
if there is a neighborhood
of
in
such that
is the interior of a rectifiable Jordan curve. Moreover,
is said to be locally rectifiable on a subset
of
if
is locally rectifiable at all points of
Now let
be an open set which is locally rectifiable on a set
We consider the induced topology on
with respect to the Euclidean topology of
Therefore, there is an open set
in
such that
and
is locally rectifiable on
Denote by
the linear measure (i.e. the one-dimensional Hausdorff measure) in
Notice that when
is a Jordan curve, then the length of
coincides with
We fix the following terminology: A set
is said to be linearly measurable if it is measurable with respect to the linear measure.
Let
be a point with the following property: There is a closed neighborhood
of
such that
is a Jordan curve whose interior contains
Then
is said to be of type I if there is a neighborhood
of
such that
is a Jordan domain.
Otherwise,
is said to be of type II. We see easily that if
is of type II, then there are an open neighborhood
of
and two Jordan domains
such that
Let
be a point of type I or II.
is said to be a point where
admits a tangent if there are a continuous one-to-one map
and a point
such that
and the following limit exists
The tangent line of
at
is, by definition, the real line
In Proposition 5.6 below we shall see that all the above notions are independent of the choice of a parametrization
Let
denote the set of all points
where
admits a tangent at
Given a linearly measurable set
such that
is locally rectifiable on
then it is well-known (see [11,p.68–69] ) that
We define the concept of angular approach regions at every point of
as follows. Let
be the tangent line of
at
and let
be any element of
Then, for any
the Stolz region or angular approach region
is given by
where
is as usual the argument function. Geometrically,
is the intersection of
with two cones of aperture
and vertex
Let
and let
be an open neighborhood of
We say that a function
defined on
admits the angular limit
at
if
for all
We conclude this subsection with a simple example which may clarify the above definitions. Let
be the open square whose four vertex are
and
Define the domain
Then
is locally rectifiable on
Every point of
is of type I and every point of
is of type II. Moreover,
2.2 Harmonic measure for an open set of
Let
be a proper open subset of
such that the boundary
(with respect to
) is non-polar (and
a subset of
) Let
be a subset of
Consider the characteristic function
Then the harmonic measure of the set
(denoted by
) is the Perron solution of the generalized Dirichlet problem with boundary data
In other words, one has
where
denotes the family of all subharmonic functions
on
such that
for each
It is well-known (see, for example, the book of Ransford [15] ) that
is harmonic on
Let
be an open set which is locally rectifiable on a subset
of
We say that a point
is a locally regular point relative to
if
for any
and any open neighborhood
of
Obviously,
If, moreover,
then
is said to be a locally regular point of
The set of all locally regular points relative to
is denoted by
Observe that in general
However, if
is open, then
As an immediate consequence of the Subordination Principle for the harmonic measure (see Corollary 4.3.9 in [15] ), one gets
|
(2.1)
|
We extend the function
to
by simply setting
Geometric properties of the harmonic measure will be discussed in Section 5 below.
2.3 Cross and separate holomorphicity
Let
and let
be a planar domain which is locally rectifiable on a linearly measurable subset
of
We define an
-fold cross
its regular part
its interior
its edge
and its regular edge
as
| |
| |
| |
| |
| |
| |
Moreover, put
It is clear that
is harmonic.
For an
-fold cross
define its wedge
Then the set of all interior points of the wedge
is given by
In particular, if
is an open set of
and
one has
and
We say that a function
is separately holomorphic on
and write
if for any
and
the function
is holomorphic on
We say that a function
(resp.
) is separately continuous on
(resp. on
) and write
(resp.
), if for any
and
the function
(resp.
) is continuous. In the sequel, for a subset
of
we write
Moreover, one often identifies
with
where
and
In this paragraph, suppose that
We say that a function
is superholomorphic on
and write
if for every proper subset
of
(including
) and any
the restricted function
given by
where
is holomorphic.
For any
and
let
denote the ball with center
with radius
Using identity ( 2.1 ) we make the following observation: For any
with
and
and any
there is a neighborhood
of
and an
such that
We say that a function
admits an angular limit
at
if
Throughout the paper, for a subset
of an Euclidean space,
denotes the space of all continuous functions
equipped with the sup-norm
Moreover, a function
is said to be locally bounded on
if, for any point
there are an open neighborhood
of
(with respect to the induced topology on
) and a positive number
such that
Finally, for an open set
(resp.
) denotes the set of all subharmonic (resp. plurisubharmonic, holomorphic) functions on
2.4 Motivations for our work
We are now able to formulate what, in the sequel, we quote as the classical one-sided version of the ”Edge-of-the-Wedge” theorem.
Theorem 1.
(Gonchar [
5,
6]
) Let
be a rectifiable Jordan domain and
an open set of the boundary
Then, for any function
there is a unique function
such that
on
Moreover, if
then
where
,
, and
denote the
-fold cross, its interior and its wedge, respectively, associated to the
,
.
Theorem 1 admits various generalizations. The following theorem is announced by Gonchar in [5] .
Theorem 2.
Let
be a Jordan domain and let
be an open set of the boundary
such that
is locally rectifiable on
Let
be a function defined on the
-fold cross
with the following properties:
-
(i)
-
(ii)
is locally bounded on
-
(iii)
for any
there is a function
defined on
such that for any
the holomorphic function
has the angular limit
at
for a.e.
and
a.e. on
1) Then there is a unique function
such that
2) If, moreover,
then
3) If, moreover,
is continuous at a point
then
On the other hand, the following result due to Drużkowski [2] gives a different flavor.
Theorem 3.
Let
be a rectifiable Jordan domain and let
be an open connected set of the boundary
Let
be a function defined on
with the following properties:
-
(i)
-
(ii)
is locally bounded on
-
(iii)
is continuous on
Then all conclusions of Theorem 1 still hold.
Observe that all these theorems require the following very strong hypothesis:
are rectifiable Jordan domains and the edge
is an open set of
Moreover, the assumptions on the boundedness and continuity of
are rather restrictive.
The question naturally arises whether Theorems 1–3 are still true if
are open sets in
and the edge
is not necessarily an open set of
In addition, if one drops the hypothesis on the local boundedness and the continuity of
can one obtain a holomorphic extension of
and what are its properties? These matters seem to be of interest especially when one seeks to generalize Theorems 1–3 to higher dimensions.
The present paper is motivated by these questions. Our first purpose is to generalize Gonchar's theorems to a very general situation, where
are, in some sense, almost general open subsets of
and where the boundary sets
are almost general subsets of
Our second goal is to establish, in this general context, an extension theorem analogous to Drużkowski's theorem with a minimum of hypothesis on
3 Statement of the main results and outline of the proofs
We are now ready to state the first main result.
Theorem A.
Let
be an open set and
a linearly measurable subset of
such that
is locally rectifiable on
Let
be such that:
-
(i)
is locally bounded on
and
-
(ii)
is measurable with respect to the
-dimensional Hausdorff measure on
-
(iii)
for any
there is a function
such that, for any
the holomorphic function
(see (i)) has the angular limit
at
for a.e.
and
a.e. on
Then there exists a unique function
with the following property:
1) There are subsets
such that
-
1a)
1
-
1b)
can be extended to a function (still denoted by)
which admits the angular limit
at every point
-
1c)
on
In addition,
enjoys the following properties:
2) If
then
3) For any
and
if
exists, then
admits the angular limit
at
4) For any
if
exists, then
admits the angular limit
at
5) If
can be extended to a continuous function defined on
then
can be extended to a unique continuous function (still denoted by)
defined on
and
admits the angular limit
at every
and
on
Theorem A has an immediate consequence.
Corollary A'.
We keep the hypotheses and the notation of Theorem A. Suppose in addition that
Then there exists a unique function
with the following property:
-
1a')
can be extended to a function (still denoted by)
which admits the angular limit
at every point
-
1b')
on
It is worthy to note that Theorem A and Corollary A' generalize, in some sense, Theorems 1–3.
Now we drop the hypothesis on local boundedness and continuity of
Then the examples of Drużkowski in [2] (see Section 11 below) show that, without these conditions, the extended function
(if it does exist) is, in general, not continuous on
However, our second main result gives a partially positive answer to this question.
Theorem B.
Let
be an open set and
a linearly measurable subset of
such that
is locally rectifiable on
and
Let
satisfy the following properties:
-
(i)
and
-
(ii)
for any
and for any
the function
is locally bounded on
and the (holomorphic) restriction function
has the angular limit
at
for every
Then there are subsets
and a unique function
with the following properties:
1)
2)
can be extended to a function (still denoted by)
which admits the angular limit
at every point
3)
on
Observe that if
then conditions (i)–(ii) above are fulfilled.
Below we give some ideas how to prove Theorems A and B. Our method consists of two steps. In the first step we suppose that each
is a Jordan domain,
In the second one we treat the general case. The key technique here is to use level sets of the harmonic measure. More precisely,we exhaust each
by the level sets of the harmonic measure
i.e. by
In order to carry out the first step, we improve Gonchar's method [5, 6] and make intensive use of Carleman's formula and of geometric properties of the level sets of harmonic measures.
In the second step we apply some mixed cross type theorems (see [14] ) in order to prove Theorems A and B with
replaced by
Then we construct the solution for the original domains
by means of a gluing procedure.
Although our results have been stated for the general case
the proofs will be presented only in the case
At the end of Section 9 we will give the ideas how to get the general results.
4 Gonchar–Carleman operator
In this section we reformulate a result due to Gonchar [5, 6] to our context. This result will play an important role for the proof of Theorems A and B. Let
be a rectifiable Jordan domain and let
be a linearly measurable subset of
such that
Let
be a function defined on
with the following properties:
-
(i)
is measurable;
-
(ii)
-
(iii)
there exist a constant
and
functions
such that for any
and
has the angular limit
at
for a.e.
and
a.e. on
For
let
be the conjugate function of
such that
for a certain fixed point
. Therefore, we can define the holomorphic functions
and
Each function
is bounded on
Therefore, in virtue of [4,p.
439] , we may define
for a.e.
to be the angular boundary limit of
at
In virtue of (i), for each positive integer
we define the Gonchar–Carleman operator as follows
|
(4.1)
|
where
Moreover, we can extend
to
in the following way. Let
and
the set of
. Set
In the sequel one often identifies
with
where
and
Then we define
(4.2)
K
M
1
,
…
,
M
N
(
z
)
=
K
M
1
,
…
,
M
N
[
f
]
(
z
)
:
=
{
f
(
z
)
,
J
′
′
=
∅
,
1
(
2
π
i
)
N
∫
A
′
′
e
−
∑
j
∈
J
′
′
M
j
(
g
j
(
a
j
)
−
g
(
z
j
)
)
f
(
z
′
,
a
′
′
)
d
a
′
′
a
′
′
−
z
′
′
,
J
′
′
≠
∅
,
where
,
, and
Moreover, we will write simply
in place of
for any
This notation is in accordance with ( 4.1 ).
The following Carleman Theorem due to Goluzin and Krylov (see, for example, [1,p.2] ) will be very useful.
Theorem 4.1.
Let
be a rectifiable Jordan domain and
a linearly measurable subset of
such that
Then for any function
any relatively compact subset
, and any
there exists a positive integer
(depending only on the sets
and the number
) such that
where
is given by (4.1) for
and
is the angular boundary limit of
on
(Notice that by [
4,p.439]
,
is linearly measurable).
-
Proof.
For every
applying the Cauchy formula to the function
we obtain
Since
for a.e.
the theorem follows. □
Theorem 4.2.
Let
be a rectifiable Jordan domain and
a linearly measurable subset of
such that
Let
be a function defined on
which satisfies (i)–(iii) above. Then 1) the following limit
exists for all
and its limit is uniform on compact subsets of
2) there is a finite constant
such that
where
and
is the constant given in (iii) above.
-
Proof.
We briefly recall the argument of Gonchar in [6] . Write the difference
as follows:
|
(4.3)
|
where
Reasoning as in formulas (6)–(9) in [6] and taking (iii) into account, we see that there is a constant
such that
|
(4.4)
|
for
Since
the desired conclusion of Part 1) follows immediately from this estimate. Finally, using ( 4.4 ), Part 2) follows. □
The following version of Privalov's Uniqueness Theorem will be also needed.
Theorem 4.3.
Let
be a domain which is locally rectifiable on a linearly measurable subset
of
with
Let
be a holomorphic function on
such that the angular limits of
on the set
are equal to
Then
-
Proof.
Using the hypothesis, one may find a rectifiable Jordan domain
such that
and
admits the angular limit
at almost every point of
By Privalov's Uniqueness Theorem (see [11] ),
Hence,
□
Finally, the following result will play a key role in the proof of Theorems A and B.
Lemma 4.4.
Let
be a rectifiable Jordan domain,
a sequence of holomorphic functions on
, and
a linearly measurable subset of
with
Assume that
admits the angular limit (denoted by)
at every point
, 2
that
and that
Then
converges uniformly on every compact subset of
to
-
Proof.
Let
where
is the conjugate harmonic function of
Since
is bounded, it follows from [4,p.439] that for a.e.
admits an angular limit (denoted by
) at
Fix an arbitrary compact subset
of
Recall that
Consequently, applying Theorem 4.1 we see that, for any
there is an
such that, for any
|
(4.5)
|
On the other hand, using that
and applying Lebesgue's Dominated Convergence Theorem, we see that
for all
and
This, combined with ( 4.5 ), implies that
Hence, by the Montel Theorem the sequence
converges uniformly on compact subsets of
to
This completes the proof. □
5 Properties of the harmonic measure and some elements of Conformal mapping theory
In this section we develop the tools needed for the proofs of Theorems A and B. In the sequel,
is an open set and
is a linearly measurable nonempty subset of
such that
is locally rectifiable on
Observe that under the above assumption
is non-polar. Let
be the set of all points
where
admits a tangent. Recall from Subsection 2.1 that
Let
be the generalized Poisson integral of
If, in addition,
is a Borel set, then, by Theorem 4.3.3 of [15] , the harmonic measure of
is given by
|
(5.1)
|
Next, let
denote the unit disc of
Then, for an
a point
is said to be a Lebesgue point of
if
where
is the linear measure defined on
. If
is a linearly measurable subset of
and
then every point of
that is a Lebesgue point of
is called a density point of
Proposition 5.1.
1) Let
Then a.e. points of
are Lebesgue points of
and
for every Lebesgue point
of
2) Let
be a subset of positive linear measure of
and
the set of all density points of
Then
and
for every
Moreover, if
is a subset of
with
then
for all interior points
of
-
Proof.
Part 1) is classical (see, for example, Theorem 5.4.8 in [17] ). The first assertion of Part 2) is also classical. Applying Part 1) to the function
the second assertion of Part 2) follows. The last assertion is almost trivial. □
We recall the following well-known result due to Carathéodory and F. and M. Riesz (see [4,p.44,p.420] ).
Theorem 5.2.
Let
be a Jordan domain and let
be a conformal mapping from
onto the unit disc
1) Then
extends to a (unique) homeomorphic map (still denoted by)
from
onto
If, in addition,
is a rectifiable Jordan domain, then, for any linearly measurable subset
of
if and only if
2) Suppose that
is a one-to-one real analytic map and
is an open set such that
and
is a Jordan domain. Then
extends to a conformal mapping from
onto
(i.e.
exists and is nonzero for
).
Now we are ready to formulate the following
Definition 5.3.
Let
be an open set which is locally rectifiable on a linearly measurable subset
of
A point
is said to be a density point of
3
if one of the following cases happens:
Case 1:
is of type I. There are an open neighborhood
of
such that
is a Jordan domain and a conformal mapping
from
onto the unit disc
which extends homeomorphically from
onto
such that
is a density point of the set
In other words,
Case 2:
is of type II. There are an open neighborhood
of
such that
with Jordan domains
and conformal mappings
from
onto
which extends homeomorphically from
onto
such that
is a density point of the set
In other words,
Proposition 5.4.
Let
be an open set which is locally rectifiable on a linearly measurable subset
of
1) Then Definition 5.3 is independent of the choice of
and
in Case 1 (resp.
in Case 2).
2) Let
denote the set of all density points of
Then
-
Proof.
To prove Part 1), let
be a density point of
We consider two cases.
Case 1:
is of type I. In virtue of Definition 5.3 and the assumption that
is of type I, for each
let
be an open neighborhood of
such that
is a Jordan domain and let
be a conformal mapping from
onto
which extends homeomorphically from
onto
Suppose that
is a density point of
We would like to show that
is a density point of
Let
be an open neighborhood of
such that
is a rectifiable Jordan domain and
is connected. We deduce that
is an open arc of
(which is obviously real analytic) and
is a Jordan subdomain of
for
Let
be a conformal mapping from the Jordan domain
onto
which extends homeomorphically from
onto
By Part 2) of Theorem 5.2 and the fact that
is an open real analytic arc, we see that
extends conformally to the arc
and
is an arc of
Next, consider the conformal mapping
given by
Since by Part 1) of Theorem 5.2
is a Jordan domain,
extends homeomorphically from
onto
and satisfies the equation
Since
is an open real analytic arc, applying again Part 2) of Theorem 5.2 , we deduce that
extends conformally to the arc
and
is an arc of
We summarize what has been done so far: We have shown that
is conformal from the arc
onto the arc
and
Hence,
Choosing
and noting that
we deduce that
| |
| |
| |
| |
where the last identity follows from the assumption that
is a density point of
Consequently,
is a density point of
Hence the proof of Part 1) is complete in this first case.
Case 2:
is of type II. Then there is an open neighborhood
of
such that
where
are some Jordan domains. We apply the result of Case 1 to each Jordan domain
Hence the proof of Part 1) is finished.
It remains to prove Part 2). We may find a sequence
of open sets of
such that
is either a rectifiable Jordan domain or the disjoint union of two rectifiable Jordan domains and
Using Part 1) and Proposition 5.1 , we see that almost every point in
is a density point of this set, i.e. belongs to
On the other hand, clearly
Consequently,
Hence,
is linearly measurable and
This completes the proof of Part 2). □
The following classical result will be needed to prove Proposition 5.6 below (see [11,p.60] or [13,p.51] ).
Theorem 5.5.
(A theorem of Lindelöf ) Let
be a Jordan domain and let
be a conformal mapping from
onto
By Part 1) of Theorem 5.2 we still denote by
its homeomorphic extension from
onto
Let
and let
Then
has a corner 4
of opening
at
if and only if
The next proposition justifies the coherence of the definitions of the tangent line and the angular approach regions given in Subsection 2.1.
Proposition 5.6.
Let
be an open set and let
1) Suppose that
is either of type I or of type II. For
let
be a parametrization of one Jordan curve contained in
such that
,
. Suppose that the following limit exists
Then the following limit also exists
and
In other words, the tangent line at
is independent of the choice of parametrizations.
2) Suppose that
is of type I. Suppose that
is a neighborhoods of
such that
is a Jordan domain. Let
be a conformal mapping from
onto
such that
where
is defined as in Part 1) of Theorem 5.2 . Then for any
and any neighborhood
of
there are neighborhoods
of
and
of
such that
3) Suppose that
is of type II. Suppose that
is a neighborhood of
such that
with some disjoint Jordan domains
satisfying
For
let
be a conformal mapping from
onto
such that
Then for any
and any neighborhood
of
there are neighborhoods
of
and
of
such that
-
Proof.
First one proves Part 1). Without loss of generality suppose that
Then the function
is a bijective continuous map and
Therefore, it is either monotone increasing or monotone decreasing. Observe that
| |
| |
where
Hence,
and the proof of Part 1) is complete.
To prove Part 2), let
be an open neighborhood of
such that
and
is a Jordan domain. Then
is a Jordan subdomain of
Let
be a conformal mapping from
onto the Jordan domain
such that
Then
is a conformal mapping from
onto the Jordan domain
and
By Part 1) of Theorem 5.2 , we have that
|
(5.2)
|
On the other hand, observe that
as
Moreover,
because
contains a neighborhood of
in
Consequently, applying Theorem 5.5 to the two conformal mappings
and
we obtain the existence of the following two limits
|
(5.3)
|
where
denotes the argument of a complex number
Choose an
such that
Then using ( 5.2 ) and ( 5.3 ), it follows that there is an open neighborhood
of
such that
|
(5.4)
|
Let
be a neighborhood of
such that
and
a neighborhood of
such that
Then we deduce immediately from ( 5.4 ) that
The remaining inclusion of Part 2) can be proved in exactly the same manner.
For Part 3) one argues as in the proof of Part 2). Hence the proof of the proposition is complete. □
The following two lemmas will be very useful.
Lemma 5.7.
Let
be a bounded domain with
smooth boundary and let
be a linearly measurable subset of
Then
-
Proof.
Since
is bounded with
boundary, using ( 5.1 ) it is a classical fact (see, for example, [19] ) that
|
(5.5)
|
for any Borel subset
of
Here
is the Poisson kernel of
and
is the Lebesgue boundary measure of
Since
is linearly measurable, there are two Borel sets
such that
and
It follows from definition that
|
(5.6)
|
On the other hand, ( 5.5 ) and the assumption on
and
imply that
This, combined with ( 5.6 ), gives the desired conclusion. □
Lemma 5.8.
Let
be the unit disc,
a linearly measurable subset of
and
a subharmonic function defined on
with
. Let
be such that
Then
on
-
Proof.
Fix a point
We wish to show that
|
(5.7)
|
As in the proof of Lemma 5.7 , we may assume (without loss of generality) that
is a Borel set. Moreover, we may suppose that
is continuous at
Otherwise, one considers a Poisson Modification
of
given by
where
is chosen so that
Then by the hypothesis on
and by Lemma 4.1.3 in [15] , we deduce that
is subharmonic on
on
and
is harmonic.
For
put
We obtain
| |
| |
where the second estimate holds because of Fatou's Lemma, the third one follows from the hypothesis on
and the last equality is a consequence of ( 5.1 ). This proves ( 5.7 ). Hence, the proof is complete. □
The following estimate will be crucial for the future development.
Proposition 5.9.
Let
be an open set which is locally rectifiable on linearly measurable subsets
of
with
Then 1) all density points of
which are elements of
are locally regular points of
and
2) for every
and every open neighborhood
of
it holds
and
3) for any interior point
of
we have
-
Proof.
To prove Part 1), let
be a density point of
There are two cases to consider according to the type of
We only treat the case when
is of type I. The second case where
is of type
is almost analogous, and therefore, left to the interested reader.
Since
is of type I and
is rectifiable at
there is an open neighborhood
of
such that
is a rectifiable Jordan domain. Fix a conformal mapping
from
onto
which extends to a homeomorphism from
onto
By Definition 5.3 and Proposition 5.4 ,
is a density point of the linearly measurable subset
of
Consequently, applying identity ( 5.1 ) and Part 1) of Proposition 5.1 yields that
|
(5.8)
|
On the other hand, one has
|
(5.9)
|
Applying Theorem 5.5 to
at
and combining ( 5.8 ) and ( 5.9 ), it follows that
Hence,
This proves the first assertion of Part 1). Let
denote the set of all density points of
By Part 2) of Proposition 5.4 ,
On the other hand, we have already shown that
Consequently,
which completes the proof of the last assertion of Part 1).
Part 2) follows from the definition of regular points and the Subordination Principle (see Corollary 4.3.9 in [15] ).
In virtue of Lemma 5.8 and the second assertion of Part 2) of Proposition 5.1 , the proof of Part 3) follows along the same lines as that of Part 2). □
In the sequel we formulate some important stability property of the angular harmonic measure. Recall that
is a proper open subset of
such that the boundary
(with respect to
) is non-polar. Let
be a linearly measurable subset of
Let
be a bounded function. The associated Perron function
is defined by
|
(5.10)
|
where
denotes the family of all subharmonic functions
on
such that
| |
| |
In the sequel,
will stand for
Proposition 5.10.
(Maximum Principle) Let
be such that
is bounded from the above and
| |
| |
Then
on
-
Proof.
Suppose that
for some
Let
be an arbitrary point of
Fix a rectifiable Jordan domain
such that
and
is a neighborhood of
in
Applying Lemma 5.8 and Part 3) of Proposition 5.9 to
yields that
Since
is an arbitrary point of
, the desired conclusion follows from the classical Maximum Principle (see Theorem 2.3.2 in [15] ). □
Using the above proposition, the corresponding results in Sections 4.1 and 4.2 of [15] with respect to
(instead of
) are still valid making the obviously necessary changes. In particular, we have the following (see Corollary 4.2.6 in [15] ):
Proposition 5.11.
Let
be a proper open subset of
such that the boundary
(with respect to
) is non-polar. Let
be a linearly measurable subset of
and
a bounded function which is continuous n.e. on
Then there exists a unique bounded harmonic function
on
such that
for n.e.
Moreover,
In virtue of this result, Theorem 4.3.3 in [15] is still valid in the context of
More precisely,
Proposition 5.12.
Let
be a proper open subset of
such that
(with respect to
) is non-polar. Let
be a linearly measurable subset of
and
a bounded Borel function. Then
We also need the following
Proposition 5.13.
Let
be a proper open subset of
such that
(with respect to
) is non-polar. Let
be a Borel subset of
such that
and
Then
on
-
Proof.
Suppose without loss of generality that
is locally rectifiable on the interval
and that
is a Borel subset of
with
Since
it follows from the Subordination Principle that
Therefore, it suffices to show that
on
To this end consider the conformal mapping
which maps
onto
It is not difficult to show that
This concludes the proof. □
Now we arrive at one of the main results of the section
Theorem 5.14.
Let
be a proper open subset of
such that
(with respect to
) is non-polar. Suppose that
is locally rectifiable on a linearly measurable subset
of
Let
be a linearly measurable subset of
with
and
Then
for
In particular, for every subset
of
with
we have
for
-
Proof.
Replacing
by two Borel sets
such that
and
and
one gets that
Then we conclude by the Subordination Principle and Proposition 5.13 that
In virtue of Proposition 5.12 , we have that
On the other hand, it follows from the definition that
Combining the above three estimates, the proof of the first assertion of the theorem follows.
The second assertion is a direct consequence of the first one. □
Proposition 5.15.
Let
be an open set which is locally rectifiable on a linearly measurable subset
of
Let
be a sequence of open subsets
of
and
a sequence of measurable subsets of
such that
-
(i)
and
-
(ii)
and
and
is locally rectifiable on
and
-
(iii)
for any point
there is an open neighborhood
of
in
such that
=
for some
Then
Remark 5.16.
We may always choose sequences
and
such that (i)–(iii) are satisfied and the open sets
are bounded. For example, take
and
-
Proof.
Using the Subordination Principle it is easy to see that the sequence
is decreasing and the following limit
exists and defines a subharmonic function in
By the Subordination Principle again, we have
Therefore, it remains to establish the converse inequality. In virtue of (iii) and of Part 2) of Proposition 5.9 , we conclude that
|
(5.11)
|
where
On the other hand, in virtue of Part 1) of Proposition 5.9 , we have that
Consequently, applying Theorem 5.14 , we deduce from ( 5.11 ) that
This completes the proof. □
Next, we introduce a notion which will be relevant for our further study.
Definition 5.17.
Let
be two open sets such that
A point
is said to be an end-point of
in
if, for every
there is an open neighborhood
of
such that
The set of all end-points of
in
is denoted by
The remaining part of this section is devoted to the study of level sets of the harmonic measure. We begin with the following important properties of these sets.
Theorem 5.18.
Let
be an open set and
a linearly measurable subset of
such that
is locally rectifiable on
and
Then, for any
the ”
-level set”
enjoys the following properties:
-
(i)
Let
be arbitrary distinct connected components of
then
-
(ii)
For any point
there is exactly one connected component
of
such that
-
(iii)
is linearly measurable and
for every connected component
of
-
Proof.
To prove (i), suppose, in order to reach a contradiction, that
Fix a point
Then, for any
there is an open neighborhood
of
such that
This implies that
Hence,
which contradicts the hypothesis that
The proof of (i) is complete.
Next, we turn to the proof of (ii). Fix a
In virtue of assertion (i), it suffices to show the existence of a connected component
of
such that
Applying Part 2) of Proposition 5.9 , we see that, for every
there is an open neighborhood
of
such that
|
(5.12)
|
Fix an arbitrary
and let
be the connected component of
containing
Since
we deduce from ( 5.12 ) that
also contains
for every
Hence
The proof of (ii) is finished.
Finally, we prove (iii). First, we may find a sequence
of open sets of
such that
is either a rectifiable Jordan domain or the disjoint union of two rectifiable Jordan domains and
Since
is linearly measurable, we see that in order to prove the measurability of
it is sufficient to check that
is linearly measurable for every
To prove the latter assertion, fix an
and let
Let
be a conformal mapping from
onto
which extends to a homeomorphic mapping (still denoted by)
from
onto
Applying Part 2) of Proposition 5.6 and using Definition 5.17 , we see that, for any
if and only if
We shall prove, in the sequel, that
is a Borel subset of
Taking this for granted, then
is also a Borel set. Consequently,
is linearly measurable.
To check that
is a Borel set, put
|
(5.13)
|
For any
let
(5.14)
T
n
m
p
:
=
{
η
∈
∂
E
:
A
n
,
m
(
η
)
⊂
Φ
(
G
∩
U
)
and
ω
(
Φ
−
1
(
w
)
,
A
,
D
)
≤
1
−
ε
−
1
p
,
∀
w
∈
A
n
,
m
(
η
)
}
.
We observe the following:
Geometric fact. Let
and
such that
Then
The proof of this fact follows immediately from the geometric shape of the cone
given in ( 5.13 ).
Let
such that
Using the above geometric fact, we see that
This, combined with ( 5.14 ) and the continuityof
implies that
Hence, the set
is closed. Clearly, we have
It follows immediately from this identity that
is a Borel set.
Consequently, as was already discussed before,
is linearly measurable.
To finish assertion (iii), it remains to prove that
Suppose, in order to reach a contradiction, that
Consider the following function
Then clearly
and
In virtue of assertions (i) and (ii) and Part 2) of Proposition 5.9 , we have that
Consequently, using the notation in ( 5.10 ), we conclude that
where
Since, by our above assumption,
it follows from Theorem 5.14 that
But on the other hand, one has
for
This leads to the desired contradiction. Hence, the proof of (iii) is finished. □
Theorem 5.19.
Let
be an open set and
a linearly measurable subset of
such that
is locally rectifiable on
and
For any
let
1) For any subset
of
such that
let
Then
2) Define the ”harmonic measure of the
-level set”
as
Then
-
Proof.
Clearly, by definition,
To prove the converse inclusion, fix an arbitrary
Consider the following function
Then
and
Moreover, in virtue of Part 2) of Proposition 5.9 , we have that
Consequently, for every
(5.15)
sup
0
<
α
<
π
2
limsup
z
→
ζ
,
z
∈
A
α
(
ζ
)
u
^
(
z
)
≤
max
{
sup
0
<
α
<
π
2
limsup
z
→
ζ
,
z
∈
A
α
(
ζ
)
u
(
z
)
,
sup
0
<
α
<
π
2
limsup
z
→
ζ
,
z
∈
A
α
(
ζ
)
ω
(
z
,
A
,
D
)
}
.
Observe that the first term in the latter line of ( 5.15 ) is equal to
because
In addition, in virtue of Part 2) of Proposition 5.9 , the second term in the latter line of ( 5.15 ) is also equal to
Hence,
Consequently, by Theorem 5.14 ,
In particular, one has
|
(5.16)
|
On the other hand, as an immediate consequence of Part 2) of Proposition 5.9 , we get that
This, combined with ( 5.16 ), implies the desired conclusions of Part 1) and Part 2). □
An immediate consequence of Theorem 5.19 is the following Two-Constant Theorem for level sets.
Corollary 5.20.
Let
be an open set and let
and
be two linearly measurable subsets of
such that
is locally rectifiable on
and
Let
and put
If
satisfies
on
and
then
6 Boundary behaviour of the Gonchar–Carleman operator
Before investigating the boundary behavior of the Gonchar–Carleman operator, we first introduce the following notion and study its properties.
6.1 Angular Jordan domains
Let
be the unit disc. We begin with the
Definition 6.1.
For every closed subset
of
and any real number
such that
and
the open set
is called the angular Jordan domain with base
and height
We give below a list of properties of such angular Jordan domains.
Proposition 6.2.
Let
be an angular Jordan domain.
1) Then there exist exactly two points
such that
and
where
is the (small) closed arc of
which is oriented in the positive sense and which starts from
and ends at
2) Write the open set
as the union of disjoint open arcs
where
is the (small) open arc of
which goes from
to
and which is oriented in the positive sense, and the index set
is finite or countable.
For
we construct the isosceles triangle with the three vertices
and
such that the base of the isosceles triangle is the segment connecting
to
and
satisfies
Let
(resp.
) denote the segment connecting
to
(resp.
the segment connecting
to
). Put
Then
is a rectifiable Jordan curve starting from
and ending at
3) Let
(resp.
) be the unique point in the circle
such that
and that
(resp.
) is minimal. Let
(resp.
) denote the segment connecting
to
(resp. the segment connecting
to
).
Let
be the (small) closed arc of the circle
which starts from
and ends at
and which is oriented in the negative sense.
Then
is a rectifiable Jordan domain and its boundary
consists of the rectifiable Jordan curve
two segments
and the closed arc
4) For every
define the dilatation
as follows
Put
Then
is a rectifiable Jordan domain and its boundary
consists of the rectifiable Jordan curve
a sub-segment
of
a sub-segment
of
and a closed arc
of
5) Consider the projection
given by
For every
notice that
Then the two maps
| |
| |
are one-to-one. In addition, for any linearly measurable subset
of
6)
as
7) For any closed Jordan curve
contained in
there is an
such that
8)
Theorem 6.3.
Let
be an open set which is locally rectifiable on a linearly measurable subset
of
with
Then, for any
and any connected component
of
there are an open set
a conformal mapping
, and an angular Jordan domain
such that
-
(i)
is either a rectifiable Jordan domain or the disjoint union of two rectifiable Jordan domains;
-
(ii)
maps
conformally onto one connected component of
(notice that, in virtue of (i),
has at most two connected components);
-
(iii)
and
-
Proof.
We have already shown in the proof of (iii) of Theorem 5.18 that there is a sequence
of open sets of
such that
is either a rectifiable Jordan domain or the disjoint union of two rectifiable Jordan domains, and
and
Consequently, there is an index
such that
|
(6.1)
|
where
Suppose without loss of generality that
is a rectifiable Jordan domain. The remaining case where
is the disjoint union of two rectifiable Jordan domains may be proved in the same way. Let
be a conformal mapping from
onto
By Theorem 5.2 ,
extends to a homeomorphic map (still denoted by)
from
onto
Hence, (i) and (ii) are satisfied.
In virtue of Theorem 5.2 , it follows from ( 6.1 ) that
|
(6.2)
|
For any
let
|
(6.3)
|
where
is given by formula ( 5.13 ).
Using the Geometric fact just after ( 5.14 ), we see that
is closed. On the other hand, applying Proposition 5.6 leads to
Therefore, in virtue of ( 6.2 ), there is an index
such that
Put
By the latter estimate one may find a closed set
contained in
such that
and
Since
a geometric argument shows that
This together with ( 6.3 ) implies that
Hence, (iii) is verified.
This completes the proof. □
In the sequel, the following uniqueness theorem will play a vital role.
Theorem 6.4.
Let
be an open set which is locally rectifiable on a linearly measurable subset
of
with
Let
with
Let
and
a connected component of
If
admits the angular limit
at every point of
then
-
Proof.
Applying Theorem 6.3 we obtain an open set
in
a conformal mapping
from
onto
which extends homeomorphically to
and an angular Jordan domain
satisfying assertions (i)–(iii) listed in that theorem.
Consider the function
By the hypothesis and by Proposition 5.6 ,
admits the angular limit
at a.e point in
Since
Theorem 4.3 gives that
on
Hence,
on the subdomain
of
This proves
□
6.2 Main result of the section
The boundary behavior of Gonchar–Carleman operator is described below.
Theorem 6.5.
Let
be the unit disc
and
a linearly measurable subset of
such that
Let
be a function defined on
with the following properties:
-
(i)
is measurable;
-
(ii)
-
(iii)
there exist a constant
and
functions
such that for any
and
has the angular limit
at
for a.e.
and
a.e. on
Let
such that
and let
be any connected component of
Then there is an angular Jordan domain
such that
and the Gonchar–Carleman operator
(see formulas ( 4.1 )–( 4.2 )) satisfies
for a.e.
6.3 Preparatory results
For the proof of Theorem 6.5 we need the following results.
In the sequel, for every function
let
denote the Cauchy integral
For a function
the radial maximal function
is defined by
Now we are able to state the following classical result (see Theorem 6.3.1 in Rudin's book [16] )
Theorem 6.6.
(Korányi-Vági type theorem) There is a constant
such that (changes)
for every
We recall the definition of the Smirnov class
on rectifiable Jordan domains.
Definition 6.7.
Let
and
a rectifiable Jordan domain. A function
is said to belong to the Smirnov class
if there exists a sequence of rectifiable closed Jordan curves
in
tending to the boundary in the sense that
eventually surrounds each compact subdomain of
such that
Next, we rephrase some facts concerning the Smirnov class
on rectifiable Jordan domains in the context of angular Jordan domains
Theorem 6.8.
1) Let
be a rectifiable Jordan domain. Then, for every
admits the angular limit
a.e. on
2) Let
be an angular Jordan domain and let
For any
let
be the rectifiable closed Jordan curve defined in Part 4) of Proposition 6.2 . Then
if
In addition, for an
,
it holds that
3) Every
has a Cauchy representation
Conversely, if
and
then
and
coincides with
a.e. on
-
Proof.
For the proof of Parts 1) and 3), see [4,p.438–441] . Taking into account Parts 6) and 7) of Proposition 6.2 , Part 2) also follows from the results in [4,p.438–441] .
Hence, the proof is complete. □
6.4 Proof of Theorem 6.5 .
We only give the proof of this theorem in the case
The proof for the general case is quite similar, therefore, it is left to the interested reader.
We fix an arbitrary
,
such that
and an arbitrary connected component
of
Applying Theorem 6.3 , we may find an angular Jordan domain
such that
Without loss of generality we may assume that
since the case where
is trivial because of (iii). In the course of the proof, the letter
will denote a positive constant that is not necessarily the same at each step.
Applying Theorem 4.1 , we have
| |
| |
Consequently,
is measurable. In addition, by (iii) this function is bounded.
Therefore, for every
we are able to define the function
|
(6.4)
|
Since, in virtue of (ii)–(iii),
and
for
it follows from Theorem 4.1 that
|
(6.5)
|
and the above convergence is uniform with respect to
On the other hand, by ( 6.4 ) we see that
is measurable and bounded.
In addition, for any
taking (ii) into account, we have that
where the first equality follows from an application of Fubini's Theorem and the second one from an application of Part 3) of Theorem 6.8 to
Consequently, in virtue of Part 3) of Theorem 6.8 , we can extend
to
by setting
|
(6.6)
|
Then the following identity holds
|
(6.7)
|
for a.e.
Now we come back to the angular Jordan domain
. We keep the notation introduced in Proposition 6.2 . Put
and
(see formulas ( 4.1 )–( 4.2 )). For any
and any
applying the Cauchy integral formula, we obtain
(6.8)
K
∞
,
M
(
z
1
,
z
2
0
)
−
K
M
,
M
(
z
1
,
z
2
0
)
=
1
(
2
π
i
)
2
∫
∂
D
1
\
A
1
∫
A
2
e
M
(
g
1
(
z
1
)
−
g
1
(
t
1
)
)
+
M
(
g
2
(
z
2
0
)
−
g
2
(
t
2
)
)
f
(
t
1
,
t
2
)
d
t
1
d
t
2
(
t
1
−
z
1
)
(
t
2
−
z
2
0
)
=
e
M
(
g
1
(
z
1
)
−
(
1
−
δ
1
)
)
∫
∂
D
1
p
N
(
t
1
)
d
t
1
t
1
−
z
1
.
Using the choice of
and the hypothesis on
and
it can be checked that
|
(6.9)
|
Therefore, recalling the projection
is defined in Part 5) of Proposition 6.2 , we estimate
(6.10)
∫
Γ
ε
|
K
∞
,
M
(
z
1
,
z
2
0
)
−
K
M
,
M
(
z
1
,
z
2
0
)
|
2
|
d
z
1
|
≤
C
∫
Γ
ε
|
M
rad
C
[
p
N
]
(
τ
(
z
1
)
)
|
2
|
d
z
1
|
≤
10
C
∫
τ
(
F
0
ε
∪
F
1
ε
∪
F
2
ε
)
|
M
rad
C
[
p
N
]
(
a
1
)
|
2
|
d
a
1
|
+
10
C
∫
τ
(
F
3
ε
)
|
M
rad
C
[
p
N
]
(
a
1
)
|
2
|
d
a
1
|
≤
20
C
∫
∂
E
|
M
rad
C
[
p
N
]
(
a
1
)
|
2
|
d
a
1
|
≤
C
∫
∂
E
|
p
N
(
a
1
)
|
2
|
d
a
1
|
≤
C
e
−
M
(
δ
1
−
δ
2
)
,
where the first estimate follows from ( 6.8 )–( 6.9 ) and the definition of the radial maximal function, the second and the third one from Part 5) of Proposition 6.2 , the fourth estimate holds by an application of Theorem 6.6 , and the last one follows from ( 6.9 ).
On the other hand, for any
(6.11)
∫
Γ
ε
|
K
M
+
1
,
M
+
1
(
z
1
,
z
2
0
)
−
K
M
,
M
(
z
1
,
z
2
0
)
|
2
|
d
z
1
|
≤
2
∫
Γ
ε
|
K
M
1
(
z
1
,
z
2
0
)
|
2
|
d
z
1
|
+
2
∫
Γ
ε
|
K
M
2
(
z
1
,
z
2
0
)
|
2
|
d
z
1
|
≤
C
e
−
M
δ
0
,
where the latter estimate follows from the same argument as in the proof of ( 6.8 )–( 6.10 ). We recall from Part 1) of Theorem 4.2 that
and that the convergence is uniform with respect to
This, combined with ( 6.10 )–( 6.11 ), implies that
|
(6.12)
|
Since we have already shown that
in virtue of Part 2) of Theorem 6.8 , we deduce from ( 6.12 ) that
For every
let
denote the angular limit of
at
(if the limit exists). It follows from ( 6.12 ) and Part 2) of Theorem 6.8 that
This, combined with ( 6.5 ) and Part 8) of Proposition 6.2 , implies finally that
Hence, Theorem 6.5 has been proved.
7 Proof of Theorem A in a special case
In this section, we prove Theorem A under the following hypotheses:
are rectifiable Jordan domains and
We first prove that Theorem A under the above hypotheses may be reduced to the case when
Indeed, let
be a conformal mapping from
onto
which extends to a homeomorphic map (still denoted by)
from
onto
Consider the function
given by
Then we deduce easily from the hypothesis on
that
satisfies the hypotheses (i)–(iii) on the cross
Let
denote the extension function of
provided by Theorem A in the case where
Using Proposition 5.6 and Part 1) of Theorem 5.2 , we see that the function
defined by
satisfies the conclusions of Parts 1)–4). This proves the above reduction.
We give here only the proof of Theorem A for the case
Since the general case is quite similar and does not require any new ideas, it is therefore left to the interested reader. In summary, from now on we assume that
and
Using hypotheses (i)–(iii) and
we are able to apply Theorem 6.5 and obtain a function
(see the notation in Subsection 2.3). We define the desired extension function
as follows
|
(7.1)
|
The remaining part of the proof is divided into several steps.
Step 1: Proof of the estimate
Proof of Step 1. Let
be an arbitrary point of
Then we may find an
such that
Let
be the connected component of
that contains
. By Theorem 6.3 we may find an angular Jordan domain
contained in
such that
In addition, for every
applying Theorem 6.5 to the function
we obtain the function
with the following property
for a.e.
Consequently, arguing as in the proof of Theorem 6.4 , we get
Since
is arbitrarily chosen, it follows from the latter identity that
|
(7.2)
|
Now we are able to conclude the proof in the same way as in [6,p.23] . More precisely, taking into account ( 7.1 )–( 7.2 ) and Part 2) of Theorem 4.2 , one gets that
Extracting the
th roots of both sides and letting
tend to
the desired estimate of Step 1 follows.
Step 2: We shall prove that
is the unique function
which verifies Property 1).
Proof of Step 2. First we show that the function
given by ( 7.1 ) satisfies Property 1). Without loss of generality, it suffices to prove that there is a subset
of
such that
and
admits the angular limit
at every point of
For any
put
By hypothesis (iii), we have
Consequently, applying Fubini's Theorem, we obtain that
Hence,
|
(7.3)
|
The same reasoning also gives that
|
(7.4)
|
Set
(7.5)
A
~
2
:
=
{
a
2
∈
A
2
∩
A
2
*
:
mes
(
{
a
1
∈
A
1
:
a
2
∈
A
a
1
′
}
)
=
mes
(
A
1
)
and
mes
(
{
a
1
∈
A
1
:
f
(
a
1
,
a
2
)
=
f
1
(
a
1
,
a
2
)
}
)
=
mes
(
A
1
)
}
.
We deduce from ( 7.3 )–( 7.5 ) that
|
(7.6)
|
Fix an arbitrary point
and let
be an arbitrary sequence of
such that
and
for some fixed number
Fix an arbitrary point
of
and let
be an arbitrary sequence of
such that
Clearly, we may find
such that
|
(7.7)
|
Fix an
such that
Since
is locally regular relative to
and
and
there is a sufficiently large number
with
|
(7.8)
|
Let
be that connected component of the following open set
which contains
(see ( 7.7 )). Applying Theorem 6.3 , we may find an angular Jordan domain
contained in
such that
Let
be a rectifiable Jordan domain with
, and
some neighborhood of the base
of
with
In virtue of ( 7.8 ) and of the fact that
we obtain that
|
(7.9)
|
Consequently, Theorem 6.5 yields that for any
|
(7.10)
|
for a.e.
Next, for any
let
| |
It follows from ( 7.5 ), ( 7.10 ) and the fact that
that
Hence
|
(7.11)
|
In virtue of ( 7.9 ), consider the following holomorphic functions on
|
(7.12)
|
Since we have already shown in Step I that
or
applying Part 1) of Theorem 6.8 , we may find a subset
of
with
such that
(resp.
) admits the angular limit
(resp.
at
Observe that by ( 7.5 ) and the fact that
we have that
Using this and ( 7.12 ), we are able to apply Lemma 4.4 to the sequence
Consequently, one gets
This shows that
admits the angular limit
at every point of
Hence,
satisfies Property 1).
In order to complete Step 2 we need to show the uniqueness of
. To do this, let
be a function with the following property: There is a subset
of
such that
and
admits the angular limit
at every point of
Fix an arbitrary point
Let
be the connected component containing
of the following open set
We deduce from the property of
and
that both holomorphic functions
and
admit the angular limit
at every point of
Consequently, applying Theorem 6.4 yields that
on
Hence,
Since
is arbitrary, the uniqueness of
is established. This completes Step 2.
Step 3: Proof of Part 2).
Proof of Step 3. Fix
For every
we have
Therefore, the Two-Constant Theorem (see Theorem 2.2 in [14] ) implies that
|
(7.13)
|
Let
and consider the
-level set
Clearly,
Recall from Step 2 that
and
|
(7.14)
|
Consider the following function
defined by
|
(7.15)
|
Clearly,
On the other hand, in virtue of ( 7.15 ) and the result of Step 1, we have
|
(7.16)
|
In addition, applying Corollary 5.20 and taking ( 7.14 )–( 7.15 ) into account yields
where, by Theorem 5.19 ,
This, combined with ( 7.13 )–( 7.16 ), implies that
| |
Hence Part 2) for the point
is proved.
Step 4: Proof of Part 3).
Proof of Step 4. Let
be such that the following limit exists
We like to show that
admits the angular limit
at
In virtue of assumption
we may suppose (without loss of generality) that
For any
we may find an open neighborhood
of
in
and a positive number
such that
and
|
(7.17)
|
Put
|
(7.18)
|
Since
it follows from Part 2) of Proposition 5.9 that
Next, consider the level set
In virtue of ( 7.18 ), we can define
|
(7.19)
|
Clearly,
|
(7.20)
|
By ( 7.19 ) and using the result of Step 2, we know that for every
the holomorphic function
admits the angular limit
at
for
where
is given in Step 2. Consequently, applying Corollary 5.20 and taking ( 7.17 ) and ( 7.20 ) into account, we see that
Let
In virtue of Theorem 5.19 and the hypothesis that
we deduce that
Consequently, there is an
such that
This completes the above assertion.
Similarly, we can prove that
admits the angular limit
at any point
if the latter limit exists. Hence the proof of Step 4 (i.e. Part 3)) is finished.
Step 5: Proof of Part 4).
Proof of Step 5. Let
be such that the following limit exists
We like to show that
admits the angular limit
at
Suppose without loss of generality that
and fix an arbitrary
Since
we may find an open neighborhood
of
in
(resp.
an open neighborhood
of
in
such that
|
(7.21)
|
By Part 2) of Proposition 5.9 , one gets
and
Consider the function
|
(7.22)
|
Clearly,
|
(7.23)
|
Applying the results of Steps 1–3 to
we obtain the function
|
(7.24)
|
so that
admits the angular limit
on
where
are given by Step 2. Clearly,
Consequently, arguing as in Step 1 and taking into account the above mentioned angular limit of
we conclude that
Consequently, applying Step 3 and taking into account ( 7.21 )–( 7.24 ) and the inequality
we see that
| |
Therefore, for all
satisfying
|
(7.25)
|
we deduce from the latter estimate that
|
(7.26)
|
Since
(resp.
) is locally regular relative to
(resp.
), there is an
such that ( 7.25 ) is fulfilled for
This, combined with ( 7.26 ), completes the proof. Hence Step 5 (i.e. Part 4)) is finished.
Step 6: Proof of Part 5).
Proof of Step 6. In virtue of Step 5, we only need to show that
admits the angular limit
on
To do this let
and choose an arbitrary
Fix a compact subset
of
such that
and a sufficiently large
such that
|
(7.27)
|
Using the hypothesis that
can be extended to a continuous function on
we may find an open neighborhood
of
in
such that
|
(7.28)
|
On the other hand,
|
(7.29)
|
For
applying the Two-Constant Theorem to the function
and taking ( 7.27 )–( 7.29 ) into account, we deduce that
|
(7.30)
|
Since
is a bounded holomorphic function for
there is an open neighborhood
of
such that
This, combined with ( 7.30 ), implies that
| |
| |
Therefore,
is continuous at
Consequently, we conclude, by Step 5, that
admits the angular limit
at
Similarly, we may also show that
admits the angular limit
at every point
This completes the proof of the last step.
8 Preparatory results
We first develop some auxiliary results. This preparation will enable us to generalize the results of section 7 to the general case considered in Theorem A.
Definition 8.1.
Let
be an open set and let
Define
| |
| |
The function
is called the plurisubharmonic measure of
relative to
Proposition 8.2.
Let
be an open set which is locally rectifiable on a linearly measurable subset
of
with
Let
be a finite or countable subset of
with the following properties:
-
(i)
For any
there is an open neighborhood
of
such that
is either a rectifiable Jordan domain or the disjoint union of two rectifiable Jordan domains;
-
(ii)
For any
define
| |
Then: 1)
and
2)
-
Proof.
To prove Part 1), let
and fix an
such that
Then
Consequently, for every
there is an open neighborhood
of
such that
This proves
To prove the second assertion of Part 1), one invokes the Subordination Principle and obtains for
|
(8.1)
|
Hence,
This implies that
In addition, since
it follows that
Hence, Part 1) is proved.
We turn to Part 2). Since
is an open set and, by Part 1),
it follows from Definitions 5.17 , 8.1 that
Hence, in virtue of Proposition 5.9 and Theorem 5.14 , it follows that
which proves the second estimate of Part 2).
To complete Part 2), let
Choose
such that
We deduce from ( 8.1 ) that
Hence,
On the other hand,
Consequently, the first estimate of Part 2) follows. The proof of the lemma is finished. □
The next result gives a nice geometric property of the interior of a wedge.
Proposition 8.3.
Let
a domain and
such that
is locally rectifiable on
Put
Then
is a domain.
-
Proof.
We only give the proof for the case
The general case when
can be proved in exactly the same way. Fix arbitrary points
and
of
Let
|
(8.2)
|
For
,
let
(resp.
) be the connected component containing
(resp.
) of the following open set
|
(8.3)
|
where
is given in ( 8.2 ).
By Theorem 5.18 (iii), for
we may find a point
(resp.
).
For
fix a connected open neighborhood
of
and an open neighborhood
of
such that
|
(8.4)
|
Combining ( 8.2 )–( 8.4 ), we see that
and
are in the same connected component of
The same argument also shows that
and
are in the same connected component of
Hence,
and
are in the same connected component of
We apply the same argument as above to
Consequently, one may find, for
a connected open neighborhood
of
and an open neighborhood
of
such that
|
(8.5)
|
and
and
are in the same connected component of
Recall that
and
are domains. Then in virtue of ( 8.4 ) and ( 8.5 ), one may find, for
a Jordan curve
such that
|
(8.6)
|
Let
|
(8.7)
|
Next, we shrink
such that
This, combined with ( 8.4 )–( 8.7 ), implies that
and
are in the same connected component of
Similarly, by shrinking
if necessary, we see that
and
are in the same connected component of
Hence,
and
are in the same connected component of
In summary, combining the above fact, ( 8.5 ) and the similar conclusion for
we have shown that
and
are in the same connected component of
Consequently,
and
are in the same connected component of
which completes the proof. □
The main ingredient in the proof of Theorem A is the following mixed cross theorem.
Theorem 8.4.
Let
be a domain of holomorphy,
an open set,
and
a linearly measurable subset of
Assume that
with
locally pluriregular5
compact subsets of
In addition,
is locally rectifiable on
with
and
For
put
Let
,
and (using the notation of Theorem 5.19 )
Let
be such that
-
(i)
-
(ii)
is measurable and bounded on
-
(iii)
for any
Then there is a unique function
such that
on
and
for every
and
Moreover,
-
Proof.
First one proves the existence and uniqueness of
To do this we argue as in the proof of Theorem 3.5.1 in [7] . For the sake of completeness, we give here a sketchy proof. Fix an
which satisfies (i)–(iii) above.
Step I: Reduction to the case where
is strongly pseudoconvex and
is a locally pluriregular compact subset of
One proceeds as in the first and second step in that proof. More precisely, since
is a domain of holomorphy, we may find an exhaustion sequence
of relatively compact, strongly pseudoconvex subdomains
of
with
By reduction assumption, for each
there exists an
such that
admit the angular limit
on
We would like to show that
on
Indeed, fix an arbitrary
and an arbitrary point
Let
such that
Let
be the connected component containing
of the following open set
Observe that both functions
and
are holomorphic and
Since
is non-pluripolar, we deduce that
Hence,
which proves the above assertion.
On the other hand, by Proposition 5.15 one gets
as
Therefore, we may glue
together to obtain a function
such that
admits the angular limit
on
and
on
The uniqueness of such an extension
can be proved using the argument given in the previous paragraph.
This completes Step I. Step II: The case where
is strongly pseudoconvex and
is a locally pluriregular compact subset of
Suppose without loss of generality that
The key observation is that we are still able to apply the classical method of doubly orthogonal bases of Bergman type.
Next one observes that Lemma 3.5.10 in [7] is still valid in the present context.
Look at Step 3 in that proof. In the sequel, we will use the notation from [7] . Let
the closure of
in
and let
be the basis from Lemma 3.5.10 in [7] ,
with the following property:
|
(8.8)
|
For any
we have
and
Hence
|
(8.9)
|
where
|
(8.10)
|
Taking the hypotheses (i)–(iii) into account and applying Lebesgue's Dominated Convergence Theorem, we see that the formula
|
(8.11)
|
defines a bounded function which is holomorphic in
Moreover, by (iii) and ( 8.10 )–( 8.11 ) it follows that
|
(8.12)
|
Observe, as in [7] and using ( 8.10 )–( 8.12 ), that we obtain the following estimates
| |
| |
where
is the volume of
with respect to the Lebesgue measure of
This shows that for any
there is a sufficiently large
such that for all
|
(8.13)
|
Take a compact set
and let
Choose an
so small that
Consider the open set
By ( 8.13 ) there is a constant
such that
|
(8.14)
|
Now we wish to show that
|
(8.15)
|
converges locally uniformly in
Indeed, by ( 8.8 ), ( 8.14 ), and Lemma 3.5.10 in [7] ,
|
(8.16)
|
which gives the normal convergence on
Since the compact set
and
are arbitrary, the series in ( 8.15 ) converges uniformly on compact subsets of
Let
denote this limit function in ( 8.15 ).
Fix
and
We choose a compact
of
as above. Moreover, we may suppose without loss of generality that
contains a neighborhood of
Let
In virtue of ( 8.16 ), there is an
such that
|
(8.17)
|
On the other hand, in virtue of ( 8.9 )–( 8.12 ), we may find, for any
an open neighborhood
of
such that
This, combined with ( 8.15 ) and ( 8.17 ), implies that
Since
and
can be arbitrarily chosen, we conclude that
To complete Step II, it remains to show that
on
To do this, fix an arbitrary
Let
be the connected component of
containing
.
Recall that
Then observe that both functions
and
admit the same angular limit
on
Consequently, applying Theorem 6.4 yields that
Hence,
which proves the above assertion.
This completes the proof of Step II. It remains to prove the estimate
In order to reach a contradiction assume that there is a point
such that
Put
and consider the function
|
(8.18)
|
Using the above assumption, it can be checked that
satisfies hypotheses (i)–(iii) of Theorem 8.4 . Hence applying the first assertion of the theorem, there is exactly one function
with
on
Therefore, by ( 8.18 ) we have on
Thus
on
In particular,
a contradiction. Hence the inequality
is proved. □
Finally, we conclude this section with two uniqueness results.
Proposition 8.5.
Let
be an open set and
a linearly measurable subset of
such that
is locally rectifiable on
Let
be a domain,
and let
be a linearly measurable subset of
such that
is locally rectifiable on
and
Put
| |
| |
Let
and
be such that both
and
admit the same angular limit at
for a.e.
Then
for every
-
Proof.
Fix an arbitrary
such that
Let
and put
Applying Theorem 6.4 to
and
it follows that
Hence, the proof is finished. □
Now we are able to prove the uniqueness stated in Theorem A.
Corollary 8.6.
We keep the hypotheses and the notation of Theorem A. Then there is at most one function
which satisfies Property 1) of Theorem A.
-
Proof.
It follows immediately from Proposition 8.5 . □
9 Proof of Theorem A
We mainly consider the case
In this case we proceed by four steps. Recall that by Corollary 8.6 , the function
satisfying Part 1) is uniquely determined (if exists).
In the first two steps we mainly prove the following assertion:
There are a function
and a subset
of
such that
and
admits the angular limit
at every point of
In fact, in the first two steps we always assume that
Using this and the previous observation and taking
for granted, we conclude the proof of Steps 1 and 2 below in exactly the same way as we did in Section 7 starting from Step 2 of that section.
Step 1: Proof of Theorem A for the case where
and
is a rectifiable Jordan domain.
Proof of Step 1. In virtue of Proposition 8.2 , let
be a finite or countable subset of
with the following properties:
-
For any
there is an open neighborhood
of
such that
is either a rectifiable Jordan domain or the disjoint union of two rectifiable Jordan domains (according to the type of
);
-
For any
define
| |
Moreover, for every
let
|
(9.1)
|
Using the hypotheses on
we conclude that
satisfies (i)–(iii) of Theorem A. Moreover,
Since
is a rectifiable Jordan domain and
is either a rectifiable Jordan domain or the disjoint union of two rectifiable Jordan domains, we are able to apply the result of Section 7 to
Consequently, we obtain, for
a unique function
a subset
of
a subset
of
such that
|
(9.2)
|
Put
|
(9.3)
|
In virtue of Proposition 8.5 , we are able to collect the family
in order to obtain a function
Next, consider the function
given by
|
(9.4)
|
In virtue of ( 9.1 )–( 9.4 ), we deduce that
|
(9.5)
|
and
|
(9.6)
|
In virtue of ( 9.4 )–( 9.6 ),
satisfies the hypotheses (i)–(iii) of Theorem 8.4 . Applying this theorem to
we obtain, for every
a function
In virtue of ( 9.6 ), we see that
|
(9.7)
|
We are now in a position to define the desired extension function
Indeed, one glues
together to obtain
in the following way
|
(9.8)
|
Now one has to check that the limit ( 9.8 ) exists and possesses all the required properties. This will be an immediate consequence of the following
Lemma 9.1.
For any point
put
|
(9.9)
|
Then
for all
Proof of Lemma 9.1 . Fix an arbitrary point
and let
Let
Then, by Proposition 8.2 , we see that
and
| |
| |
where the latter estimate follows from formula ( 9.9 ). Consequently,
|
(9.10)
|
On the other hand, using Part 1) of Proposition 8.2 , it is clear that
|
(9.11)
|
Moreover, in virtue of ( 9.4 ) and ( 9.7 ), we have
|
(9.12)
|
Next, let
be the connected component containing
of the following open set
Observe that, in virtue of ( 9.10 )–( 9.11 ), both functions
and
are holomorphic and
is a nonempty open set. Therefore, we deduce from ( 9.12 ) that
on
Hence,
which completes the proof of the lemma.
We complete the proof as follows. An immediate consequence of Lemma 9.1 is that
Next, we apply Lemma 9.1 and make use of ( 9.4 )–( 9.9 ) and of the fact that
as
Consequently, we conclude that
satisfies
Taking into account the remark made at the beginning of this section, the proof of the theorem in Step 1 is finished.
Step 2: Proof of Theorem A for the case
and
Proof of Step 2. We proceed using Step 1 in exactly the same way as we proved Step 1 using the result of Section 7. Hence, Step 2 is finished.
Step 3: Proof of Theorem A for the case when
and
are compacts.
Proof of Step 3. Recall from the hypothesis (i) of Theorem A that
is locally bounded. Since
are compact, a compactness argument shows that there are a positive number
and an open neighborhood
(resp.
) of
(resp.
) such that
|
(9.13)
|
For
fix a sequence
of subdomains of
such that
|
(9.14)
|
Consider the sequence
of open sets of
the sequence of crosses
and the sequence of the interior of their wedges
defined as
|
(9.15)
|
Since
is locally bounded,
are compact and, by ( 9.14 ),
a compactness argument shows that for every
there is a number
such that
and
This, combined with ( 9.13 ) and ( 9.15 ), implies that
|
(9.16)
|
On the other hand, in virtue of ( 9.14 ) and ( 9.15 ), we see that the sequences
and
satisfy the hypotheses of Proposition 5.15 .
Consequently, applying this proposition and taking ( 9.15 ) into account, we obtain
|
(9.17)
|
In virtue of ( 9.16 ) and the hypothesis on
we see that
satisfies the hypotheses of Step 2. Consequently, applying the result of Step 2 to
we obtain a unique function
and a subset
(resp.
) of
(resp.
) such that
|
(9.18)
|
Put
|
(9.19)
|
Using ( 9.18 ), we deduce from ( 9.19 ) that
|
(9.20)
|
Using ( 9.17 ), ( 9.20 ) and applying Corollary 8.6 , we obtain
|
(9.21)
|
Therefore, we may glue
together in order to obtain the desired extension function
as
|
(9.22)
|
In virtue of this formula and ( 9.17 ), ( 9.20 )–( 9.21 ), we conclude that
satisfies
Hence, the proof of Part 1) of Theorem A is complete.
Using
and arguing as in Step 3 of Section 7, Part 2) of Theorem A follows. Now we turn to Part 3) of Theorem A. Let
be such that the following limit exists
We would like to show that
admits the angular limit
at
Let
be a sufficiently large integer such that
Then in virtue of ( 9.17 ) and ( 9.20 )–( 9.21 ),
|
(9.23)
|
In addition, applying Part 2) of Theorem A established in Step 2 and taking ( 9.16 ) into account, we see that
|
(9.24)
|
Consequently, the proof of Step 4 in Section 7 applied to
still works in this context making the obviously necessary changes. Hence,
and then
by ( 9.23 ) admits the angular limit
at
This completes the proof of Part 3) of Theorem A. Using
and ( 9.23 )–( 9.24 ) (for a sufficiently large
), the proof of Parts 4) and 5) of Theorem A given in Steps 5 and 6 of Section 7 still works in this context making the obviously necessary changes. This completes Step 3.
Step 4: Proof of Theorem A for the case
Proof of Step 4. For each
since
is linearly measurable, one may find a sequence
of subsets of
such that
-
(a)
is a compact subset of
and
-
(b)
-
(c)
For
let
|
(9.25)
|
On the other hand, in virtue of (a)–(c), we see that the sequences
and
satisfy the hypotheses of Proposition 5.15 . Consequently, applying this proposition and taking ( 9.25 ) into account, we obtain
|
(9.26)
|
In virtue of (a) and the hypothesis on
we see that
satisfies the hypotheses of Step 3. Consequently, applying the result of Step 3 to
we obtain a unique function
and a subset
(resp.
) of
(resp.
) such that
|
(9.27)
|
Put
|
(9.28)
|
Using ( 9.27 ), we deduce from ( 9.28 ) that
|
(9.29)
|
Applying ( 9.26 ), ( 9.29 ), and Corollary 8.6 , it follows that
|
(9.30)
|
Therefore, we may glue
together in order to obtain the desired extension function
as
|
(9.31)
|
In virtue of this formula and ( 9.26 ), ( 9.29 )–( 9.30 ), we conclude that
satisfies
Hence, the proof of Part 1) of Theorem A is complete.
Using
and arguing as in Step 3 of Section 7, Part 2) of Theorem A follows.
The remaining parts of Theorem A can also be proved using Step 5 and 6 of Section 7 and making the obviously necessary changes.
This completes the last step of Theorem A in the case
Finally, we present a sketch of the argument for the general case
In order to prove Theorem A in its full generality, we use induction twice. More precisely, we proceed by induction (I) on
Suppose the theorem is true for
We have to discuss the case of an
-fold cross
where
and
are linearly measurable subsets of
such that
is locally rectifiable on
We proceed again by induction (II) on the positive integer
such that there are at least
rectifiable Jordan domains among the open sets
For
we are reduced to Section 7.
In fact, our proof follows essentially the scheme of the works in [14] , [12] , and that of Sections 7 and the previous proof for
10 Proof of Theorem B
We will only give the proof of Theorem B for the case when
and
are rectifiable Jordan domains. Since the general case can be proved using the scheme of Section 7 and 9, it is left to the interested reader. The proof is divided into two steps.
Step 1: Proof of Theorem B for the case when the slice functions
and
are bounded for every
and
Proof of Step 1. For any
let
|
(10.1)
|
Using the assumption of Step 1 and ( 10.1 ), we obtain
|
(10.2)
|
Now we would like to show that for
|
(10.3)
|
To do this, suppose without loss of generality that
and fix an arbitrary
and a point
Let
be a sequence in
such that
Consequently, using the hypothesis (i), we see that
|
(10.4)
|
On the other hand, it follows from the assumption
and the hypothesis of Step 1 that
Combining this and ( 10.4 ), we are able to apply Lemma 4.4 to the sequence
where
Consequently, the sequence
converges uniformly on compact subsets of
to
This completes the proof of ( 10.3 ).
On the other hand, by hypothesis (ii), the holomorphic function
admits the angular limit
at
Hence, it follows that
is measurable.
Moreover, by ( 10.1 ),
for every
In addition, in virtue of ( 10.2 ), there exists a sufficiently large integer
such that
for
and
Consequently, we are in a position to apply Theorem A to the function
restricted to the cross
for
Therefore, we obtain a function
and a subset
(resp.
) of
(resp.
) for
such that
|
(10.5)
|
Put
|
(10.6)
|
Using ( 10.5 ), we deduce from ( 10.6 ) that
|
(10.7)
|
Applying ( 10.2 ), ( 10.7 ), and Corollary 8.6 , we obtain
|
(10.8)
|
Therefore, we may glue the
together to obtain the desired extension function
as
|
(10.9)
|
Next, for every
and
in virtue of ( 10.2 )–( 10.3 ) and ( 10.5 ), one may find a sequence
of compact subsets of
such that
|
(10.10)
|
Moreover, for any
and for any
put
|
(10.11)
|
Since, by hypothesis (i),
we deduce from ( 10.10 ) and ( 10.11 ) that
is a closed subset of
and
|
(10.12)
|
Consequently, there is an
such that
for any
Now we are in a position to apply Theorem A to the function
restricted on the cross
Using ( 10.7 )–( 10.9 ) and Corollary 8.6 , we obtain exactly the function
restricted to
Let
|
(10.13)
|
where
denotes as usual the set of locally regular points relative to
Taking ( 10.11 )–( 10.13 ) into account and arguing as in Step 5 of Section 7, we may show that
|
(10.14)
|
for every
Now it suffices to put
Combining this and ( 10.14 ), ( 10.12 ), ( 10.2 ), we may check that all the conclusions 1)–3) of Theorem B are satisfied. Hence the proof is complete in this first step.
Step 2: The general case.
Proof of Step 2. Arguing as in Section 7, it suffices to treat the case when
the unit disc
We begin with the following
Definition 10.1.
For every closed subset
of
and any
define the following open set
The reader should compare this definition with Definition 6.1 . Below we give a list of properties of such open sets.
Proposition 10.2.
Let
be a closed subset of
1) Let
be as in Definition 10.1 , then
is a rectifiable Jordan domain and
2)
as
3) Consider a locally bounded function
Then
for every
4) There holds the following equality
Proof of Proposition 10.2 . Part 1) may be done as in the proof of Proposition 6.2 .
Part 2) is an immediate consequence of Definition 10.1 .
Part 3) follows immediately from the compactness of
The proof of Proposition 5.15 still works in the context of Part 4) making the obviously necessary changes. This completes Part 4).
Now we are in a position to complete Step 2. Indeed, first suppose that both
and
are closed. Then for
consider the sequence of rectifiable Jordan domain
given by
For
let
In virtue of Proposition 10.2 , we are able to apply the result of Step 1 to
Consequently, we obtain a function
Therefore, we may glue
together in order to obtain the desired extension function
as
Because of Proposition 10.2 , we can show that
possesses all the assertions of Theorem B. The case when
and
are only measurable is similar. It suffices to find a sequence
of subsets of
such that
is compact and
Then we may apply the previous discussion to
in order to obtain a function
Finally, the desired extension function
is defined by
This completes the proof in this last step.
11 Examples and Concluding remarks
The following examples of Drużkowski [2] show the optimality of Theorem A and B. Consider
and
Example 1. Define a function
as follows
where
is the principal branch of logarithm.
Put
As in [2] observe that
is measurable,
but
is not continuous at
Since
using the uniqueness established in Theorem A, we conclude that the solution
provided by Theorem A and B satisfies
In addition, we see that, for
the angular limit of
at
does not exist. Thus the condition in assertion 3) of Theorem A is necessary. Moreover, the sets
given by Theorem B do depend on
Example 2. Define a function
as follows
where
Define
Then
As in [2] observe that
is continuous,
but
is not locally bounded on
In addition, for
consider the functions
given by
| |
| |
We may prove that there is an
and a neighborhood
of
in
such that
In addition, it can be checked that
This shows that the assumption of the local boundedness on
is necessary in Theorem A. Finally, we conclude the article by some remarks and open questions.
1. We may generalize Theorems A and B in the following directions: Let
be an open set.
-
1a.
We extend
to the set of all points
where
admits a corner at
(see [13] ). Note that there are at most countably many points where
admits a corner but not a tangent (see Exercises 3.4 in [13] ).
-
1b.
We extend the notion type of a point in
as follows. A point
is said to be of type
(
) if there are an open neighborhood
of
and
disjoint rectifiable Jordan domains
such that
-
1c.
Instead of the notion local rectifiability, we use the weaker notion local Jordan curve. More precisely,
is said to be locally Jordan curve like at a point
if there is a neighborhood
of
in
such that
is the union of
disjoint Jordan curves and the interior of at least one of them contains
Since the corresponding statements of Theorems A and B under these generalizations are somewhat complicated and their proofs do not require any new method, we leave the interested reader to develop these ideas.
2. It may be proved that
provided by Theorem A is the maximal domain of holomorphic extension of the function
We postpone the proof of this result to an ongoing work.
3. It seems to be of interest to consider Theorem A and B under the following general settings: Let
be a complex manifold of dimension
and
an open set,
Let
be a subset of positive
-dimensional Hausdorff measure of
where
is a real
-dimensional generating submanifold contained in
etc. We postpone this issue to an ongoing work.
4. Does Theorem A still hold if we omit the assumption (ii) ”
is measurable”?5. Does Theorem B still hold if we omit the assumption that
?
References
-
L. Aizenberg, Carleman's formulas in complex analysis. Theory and applications, Transl. from the Russian, updated, enlarged and revised. (English) Mathematics and its Applications (Dordrecht). 244. Dordrecht: Kluwer Academic Publishers. xx, 299 p. (1993).
-
L. M. Drużkowski, A generalization of the Malgrange–Zerner theorem, Ann. Polon. Math., 38, (1980), 181–186.
-
H. Epstein, Generalization of the ”Edge of the Wedge” Theorem, J. Math. Phys., 1, (1960), 524–531.
-
G. M. Goluzin, Geometric theory of functions of a complex variable, (English), Providence, R. I.:American Mathematical Society (AMS). VI, (1969), 676 p.
-
A. A. Gonchar, On analytic continuation from the ”edge of the wedge” theorem, Ann. Acad. Sci. Fenn. Ser. A.I: Mathematica, 10, (1985), 221–225.
-
A. A. Gonchar, On Bogolyubov's ”edge of the wedge” theorem, Proc. Steklov Inst. Math., 228, (2000), 18–24.
-
M. Jarnicki, P. Pflug, Extension of Holomorphic Functions, de Gruyter Expositions in Mathematics 34, Walter de Gruyter, 2000.
-
M. Jarnicki, P. Pflug, An extension theorem for separately holomorphic functions with analytic singularities, Ann. Pol. Math., 80, (2003), 143–161.
-
M. Klimek, Pluripotential theory, London Mathematical society monographs, Oxford Univ. Press., 6, (1991).
-
H. Komatsu, A local version of Bochner's tube theorem, J. Fac. Sci., Univ. Tokyo, Sect. I A 19, (1972), 201–214.
-
P. Koosis, Introduction to
spaces. With an appendix on Wolff 's proof of the corona theorem, London Mathematical Society Lecture Note Series. 40. Cambridge etc.: Cambridge University Press. XV, 376 p. (1980). (1991), 183–194.
-
V.-A. Nguyên, A general version of the Hartogs extension theorem for separately holomorphic mappings between complex analytic spaces, preprint of the University of Oldenburg (2004), 33 p.
-
Ch. Pommerenke, Boundary behaviour of conformal maps, Grundlehren der Mathematischen Wissenschaften, 299, Berlin: Springer-Verlag, 300 p. (1992).
-
P. Pflug and V.-A. Nguyên, A boundary cross theorem for separately holomorphic functions, Ann. Polon. Math., 84, (2004), no. 3, 237–271.
-
T. Ransford, Potential theory in the complex plane, London Mathematical Society Student Texts, 28, Cambridge: Univ. Press., (1995).
-
W. Rudin, Lectures on the edge-of-the-wedge theorem, (English) Conference Board of the Mathematical Sciences. Regional Conference Series in Mathematics. No.6. Providence, R.I.:American Mathematical Society (1971).
-
W. Rudin, Function theory in the unit ball of
, (English) Grundlehren der Mathematischen Wissenschaften 241, Springer-Verlag, New York-Berlin, 1980.
-
W. Rudin, Function theory in polydiscs (English) Mathematics Lecture Note Series. New York-Amsterdam: W.A. Benjamin, Inc., 188 p. (1969).
-
E. M. Stein, Boundary behavior of holomorphic functions of several complex variables, (English) Mathematical Notes. Princeton, N. J.: Princeton University Press. IX, 72 p. (1972).
-
V. S. Vladimirov, Methods of the Theory of Functions of Several Complex Variables, (Russian), Moscow: 'Nauka' Publisher, 411 p. (1964).
-
M. Zerner, Quelques résultats sur le prolongement analytique des fonctions de variables complexes, Séminaire de Physique Mathématique.
Peter Pflug, Carl von Ossietzky Universitat Oldenburg, Fachbereich Mathematik, Postfach 2503, D–26111, Oldenburg, Germany E-mail address : pflug@mathematik.uni-oldenburg.de Vie t-Anh Nguyen, Carl von Ossietzky Universitat Oldenburg, Fachbereich Mathematik, Postfach 2503, D–26111, Oldenburg, Germany E-mail address : nguyen@mathematik.uni-oldenburg.de