Power series as Fourier Series
Abstract.
An abstract theory of Fourier series in locally convex topological vector spaces is developed. An analog of Fejér’s theorem is proved for these series. The theory is applied to distributional solutions of Cauchy-Riemann equations to recover basic results of complex analysis. Some classical results of function theory are also shown to be consequences of the series expansion.
2010 Mathematics Subject Classification:
32A05,22D121. Introduction
1.1. Motivation
Two types of series expansion ubiquitous in mathematics are the power series of an analytic function, and the Fourier series of an integrable function on the circle. In the complex domain, well-known theorems of elementary complex analysis guarantee the existence of locally uniformly convergent power series expansions of holomorphic functions in domains with ample symmetry such as disks and annuli, where “holomorphic” can be taken in the sense of Goursat, i.e., the function is complex-differentiable at each point. On the other hand, the convergence of a Fourier series is a subtle matter, and its study has led to many developments in analysis (see [Zyg02]). The close connection between these two types of outwardly different series expansions is a recurring theme in many areas of classical analysis, e.g., in the theory of Hardy spaces.
It is however not difficult to see that at a certain level of abstraction, Fourier series and power series are in fact two examples of the same phenomenon, the representation theory of the circle group. The aim of this article is to take this idea seriously, and use it to recapture some basic results of complex analysis. The take-home message is that many properties of holomorphic functions, such as the almost-supernatural regularity phenomena are profitably thought of as expressions of symmetry, more precisely the invariance of certain locally convex spaces under the Reinhardt action of the torus group on . It is hoped that the point of view taken here has pedagogical as well as conceptual value, and will be of interest to students of complex analysis.
1.2. Abstract Fourier series
We begin in Section 2 with an account of Fourier series associated to a continuous representation of the -dimensional torus on locally convex topological vector spaces, using some ideas of [Joh76]. This provides the unifying language of sufficient generality to encompass both classical Fourier expansions and the power series representations of complex analysis. Even at this very general and “soft” level, one can establish a version of Fejér’s theorem on the summability in the Cesàro sense (Theorem 2.1), which can be thought of as a “completeness” statement for the holomorphic monomials.
1.3. Analyticity of holomorphic distributions
Using the abstract framework of Section 2, we recapture in Section 3, in a novel way, some of the basic classical facts about holomorphic functions. We show that on a Reinhardt domain, a distribution which satisfies the Cauchy-Riemann equations (which we call a holomorphic distribution) has a complex power series representation that converges uniformly along with derivatives of all orders on compact subsets of a “relatively complete” log-convex Reinhardt domain (see Theorem 3.1 below), and thus a holomorphic distribution is a function, in fact an analytic function. Traditionally, to prove such an assertion, one would start by showing that a holomorphic distribution is actually a (smooth) function. This can be done either by ad hoc arguments for the Laplacian going back to Weyl (see [Wey40]), or in a more general way, by noticing that fundamental solution of the Cauchy-Riemann operator is smooth, in fact real-analytic, away from the singularity at the origin. It then follows by standard arguments about convolutions (see [Hör03, Theorem 4.4.3]), that any distributional solution of is real-analytic. Therefore, a holomorphic distribution is a holomorphic function in the sense of Goursat, and classical results of elementary complex analysis give the power series expansion via the Cauchy integral formula. The extension of the domain of convergence to the envelope of holomorphy can be obtained by convexity arguments (see [Ran86]).
While this classical argument has the admirable advantage of placing holomorphic functions in the context of solutions of hypoelliptic equations, it also has the shortcoming that the crucial property of complex-analyticity (and the associated Hartogs phenomenon in several variables) of holomorphic distributions is proved not from an analysis of the action of a differential operator on distributions, but by falling back on the Cauchy integral formula. After using the real-analytic hypoellipticity of to conclude that holomorphic distributions are real-analytic, we discard all of this information, except that holomorphic functions are and satisfy the Cauchy-Riemann equations in the classical sense. In our approach here, however, complex analyticity of holomorphic distributions is proved directly in a conceptually straightforward way, by expanding a holomorphic distribution on a Reinhardt domain in a Fourier series, and then showing that the resulting series (the Laurent series of the holomorphic function) converges in the -topology, not only on the original set where the distribution was defined, but possibly on a larger domain, thus underlining the fact that Hartogs phenomenon can be thought of as a regularity property of the solutions of the same nature as smoothness. Our proof also clearly locates the origin of the remarkable regularity of holomorphic distributions in
-
(a)
the invariance properties of the space of holomorphic distributions under Reinhardt rotations and translations,
-
(b)
symmetry and convexity properties of the Laurent monomial functions
and -
(c)
the fact that radially symmetric holomorphic distributions are constants.
It is also interesting that the fact that is a fundamental solution of the Cauchy-Riemann operator, which is key to many results of complex analysis including the Cauchy integral formula, does not play any role in our approach.
The method of proving analyticity via Fourier expansion can be used in other contexts. For example, replacing the representation theory of by that of the special orthogonal group , the method can be used to show that a harmonic distribution in a ball of is in fact a real analytic function and admits an expansion in solid harmonics (see, e.g., [CH53, pp. 316-317]), which converges uniformly along with all derivatives on compact subsets of the ball. Similarly, one can obtain the Taylor/Laurent expansion of a monogenic function of a Clifford-algebra variable in “spherical monogenics”, the analogs for functions of a Clifford-algebra variable of the monomial functions (see [BDS82]).
1.4. Analyticity of continuous holomorphic functions
While the theory of generalized functions forms the natural context for solution of linear partial differential equations such as the Cauchy-Riemann equations, for aesthetic and pedagogical reasons it is natural to ask whether it is possible to develop the Fourier approach to power series, as outlined in Section 3, using only classical notions of functions as continuous mappings, and derivatives as limits of difference quotients, and without anachronistically invoking distributions. We accomplish this in the last Section 5 of this paper. We start from the assumption that a continuous function on the complex complex plane satisfies the hypothesis of the Morera theorem, and show directly that it is complex-analytic without any recourse to the Cauchy integral representation formula (which, being a case of Stoke’s theorem, requires the differentiability of the function, at least at each point). Not unexpectedly, one of the steps of the proof uses the classical triangle-division used in the standard proof of the Cauchy theorem for triangles.
1.5. Acknowledgements
The first author would like to thank Luke Edholm and Jeff McNeal for many interesting conversations about the topic of power series. He would also like to thank the students of MTH 636 and MTH 637 at Central Michigan University over the years for their many questions, which led him to think about the true significance of power series expansions.
Sections 2 and 4 of this paper are based on part of the Ph.D. thesis of the second author under the supervision of the first author. The second author would also like to thank his other committee members, Dmitry Zakharov and Sonmez Sahutoglu for their support. Other results from the thesis have appeared in the paper [Daw21].
2. Fourier Series in Locally Convex spaces
2.1. Nets, series and integrals in LCTVS
We begin by recalling some notions and facts about functional analysis in topological vector spaces. See the textbooks [Trè67, Rud91, Bou04] for more details on these matters.
Let be a locally convex Hausdorff topological vector space (we use the standard abbreviation LCTVS). Recall that the topology of can also be described by prescribing the family of continuous seminorms on : a net in converges to , if and only if for each continuous seminorm on , we have as a net of real numbers. In practice, we describe the topology of an LCTVS by specifying a generating family of seminorms (analogous to describing a topology by a subbasis): a collection of continuous seminorms on is said to generate the topology of if for every continuous seminorm on , there exists a finite subset and a such that
(2.1) |
and further, for every nonzero there exists at least one such that (this separating property ensures that the topology of is Hausdorff). Then clearly a net converges in if and only if for each .
Let be a directed set with order . Recall that a net in is said to be Cauchy if for every and every continuous seminorm on , there exists such that whenever and , we have . The space is said to be complete if every Cauchy net in converges. Observe that in the above definition we can use a generating family of seminorms rather than all continuous seminorms.
If is a sequence in a vector space, we can define the sequence of the corresponding Cesàro means by
The following is the analog for sequences in LCTVS of an elementary fact well-known for sequences of numbers:
Proposition 2.1.
Let be a convergent sequence in an LCTVS . Then the sequence of Cesàro means is also convergent, and has the same limit.
Proof.
Let . If is a continuous seminorm on and , there is such that for . Set . Then for , we have
Therefore, if we choose so large that , then for we have , so in . ∎
For a formal series in an LCTVS , convergence is defined in the usual way, i.e. the sequence of partial sums converges in . A formal sum over a countable index set , where are vectors in an LCTVS , is said to be absolutely convergent if there exists a bijection such that for every continuous seminorm on , the series of non-negative real numbers is convergent (see [KK97]). To check that a series is absolutely convergent, we only need to check the convergence of the above series for seminorms in a fixed generating family. If is a locally compact Hausdorff space, absolute convergence in the Fréchet space of continuous complex valued functions on is what is classically called normal convergence (see [Rem91, pp. 104 ff.]). Absolute convergence is typical for many spaces of holomorphic functions, e.g. in the space of holomorphic functions on a Reinhardt domain smooth up to the boundary (see [Daw21]).
The following result, whose proof mimics the corresponding result for numbers, shows that absolutely convergent series in LCTVS behave very much like absolutely convergent series of numbers:
Proposition 2.2.
Let be a complete LCTVS, and let be an absolutely convergent series of elements of . Then the series is unconditionally convergent: there is an such that for every bijection , the series converges in to .
The element is naturally called the sum of the series, and we write .
Proof.
By definition, there exists a bijection , such that for each continuous seminorm on , the series converges. Let and . Since converges, for there exists such that whenever with , . Therefore for ,
(2.2) |
Therefore is Cauchy sequence in the complete LCTVS , and therefore converges to an . In order to complete the proof, it suffices to show that for every bijection , the series converges to the same sum . Let . Choose such that the set of integers is contained in the set . Then, if , the elements get cancelled in the difference and we have by (2.2). This proves that the sequences and converge to the same limit. So, as . ∎
By a famous theorem of Dvoretzky and Rogers, the converse of the above result fails when is an infinite dimensional Banach space ([KK97, Chapter 4]).
We will also need the notion of the weak (or Pettis) integral of a LCTVS valued function which we will now recall (see [Bou04, p. INT III.32-39] for details). Let be a compact Hausdorff space and let be a Borel measure on , and let be a continuous map from to an LCTVS . An element is called a Pettis integral of on with respect to if for all ,
(2.3) |
where denotes the dual space of (the space of continuous linear functionals on ), and the right hand side of (2.3) is an integral of a continuous function. If is complete, one can show that there exists a unique such that (2.3) holds and we denote the Pettis integral of on with respect to by . In fact, the integral exists uniquely, as soon as the space is quasi-complete, i.e. if each bounded Cauchy net in converges, where a net is bounded if for each continuous seminorm , the net of real numbers is bounded. While there are situations in which this more refined existence theorem for Pettis integrals is useful (e.g. when the space is a dual space with weak-* topology), in this paper, we only consider integrals in complete LCTVS. Since each Hausdorff TVS has a unique Hausdorff completion, we can define Pettis integrals in any LCTVS, provided we allow the integral to have a value in the completion.
If is a continuous linear map of complete LCTVSs, and is continuous then we have
(2.4) |
since for any linear functional ,
using the fact that and the definition (2.3) of the Pettis integral.
2.2. Representation of the torus on an LCTVS .
Let
be the -dimensional unit torus. With the subspace topology inherited from and binary operation defined as , is a compact abelian topological group. For a LCTVS , and a continuous function , we denote the Pettis integral of with respect to the Haar measure of (normalized to be a probability measure) by
Let be an LCTVS, and let be a continuous representation of on . Recall that this means that for each , the map is an automorphism (i.e. linear self-homeomorphism) of as a topological vector space, the map is a group homomorphism from to , and the associated map
is continuous.
Given a representation of the group on an LCTVS , a continuous seminorm on is said to be invariant (with respect to ) if for all and .
Proposition 2.3.
A representation of on an LCTVS is continuous if and only if the following two conditions are both satisfied:
-
(a)
the topology of is generated by a family of invariant seminorms, and
-
(b)
for each the function from to given by is continuous at the identity element of .
Proof.
Assume that is continuous, i.e., is continuous, so for , the function is continuous on , in particular at the identity. Therefore (b) follows.
For a continuous seminorm on , define , which is finite since is compact, and which is easily seen to be a seminorm. To show is invariant, we note that
It remains to show that is continuous. For , we have
so that for all , and it follows that the seminorm is continuous on if and only if it is continuous at . We show that for each , there exists a neighbourhood of in such that for all , on . For each , since and are continuous, there exists a neighborhood of in and a neighbourhood of in such that for all and . The collection forms an open cover of . Since is compact, let be a finite subcover of corresponding to the open cover. Then for all and , we have .
Now assume the two conditions (a) and (b). Let be a directed set and let and be nets in and respectively with in . We need to show in , i.e., for each invariant continuous seminorm of . But we have, by the invariance of :
The term goes to zero since , and the term also goes to zero since and is continuous at . The result follows. ∎
2.3. Abstract Fejér Theorem
Let be a complete LCTVS and let be a continuous representation of on . For each and , define
(2.5) |
the Pettis integral of the continuous function on with respect to the Haar probability measure of . We will say that is the -th Fourier component of with respect to the representation . We use the standard convention with respect to multi-index powers, i.e., . We will say that the subspace of defined as
(2.6) |
is the -th Fourier mode of the space , and we will call the map the -th Fourier projection, both with respect to the representation . We note the following facts:
Proposition 2.4.
As above, let be a complete LCTVS and be a continuous representation of on .
-
(1)
For each , the -th Fourier mode is a closed -invariant subspace of , and the Fourier projection is a continuous linear projection from onto .
-
(2)
Let be another complete LCTVS, and let be a continuous representation of on , and let be a continuous linear map intertwining and , i.e., for each , Then for each , we have
(2.7)
Proof.
-
(1)
Linearity of follows from the linearity of . Recall from Proposition 2.3 that there exists a family of continuous invariant seminorms that generates the locally convex topology of . To see the continuity of , observe that for each , we have
(2.8) where the inequality is due to Proposition 6 in [Bou04, p. INT III.37]. Now, if , then
so that . To prove that , notice that for each , each and ,
using (2.4) - (2)
∎
Remark.
The inequality (2.8)
(2.9) |
which holds for each for an invariant seminorm can be thought of as an abstract form of the familiar Cauchy inequalities of complex analysis.
If is , the Banach space of integrable functions on , and if is the continuous representation of on given by
an easy computation shows that for , and , we have
the -th term of the Fourier series of .
It is therefore natural to define, for , the Fourier series of with respect to to be the formal series
(2.10) |
For an integer , define the -th square partial sum of the Fourier series in (2.10) by
(2.11) |
where . We are ready to state an abstract version of Fejér’s theorem.
Theorem 2.1.
Let be a continuous representation of on an LCTVS and let . Then the Cesàro means of the square partial sums of the Fourier series of (with respect to ) converge to in the topology of .
Proof.
Write the Cesàro means of the square partial sums of the Fourier series of as,
where
is the classical Fejér kernel. Introducing polar coordinates, on , and summing, we obtain the classical representation
It is well-known that the Fejér kernel has the properties that
-
(a)
for all ,
-
(b)
and
-
(c)
For each , uniformly on , where is the -dimensional ball centered at and radius .
Let be a continuous -invariant seminorm on . Then for , we have
using property (b) of | ||||
using (2.4) | ||||
(2.12) | ||||
using [Bou04, Prop. 6, p. INT III.37] and the positivity of |
Since is continuous and is a continuous seminorm, there exists such that whenever Then on the set we have
By property (c) of , there exists such that whenever , for all we have,
(2.13) |
Now from (2.12), we have
Since by Proposition 2.3 the topology of is generated by -invariant seminorms, the result follows. ∎
Corollary 2.5.
Let be a complete LCTVS and suppose that we are given a continuous representation of the group on . Then if the Fourier series of an element with respect to this representation is absolutely convergent in , the sum of the Fourier series equals .
Proof.
Since the series is absolutely convergent, by Proposition 2.2, there exists the sum of the series, i.e., an such that for every bijection we have . Let ; then the sequence of partial sums converges to in . By Proposition 2.1, the sequence of Cesàro means of the partial sums converges to as well. However, by Theorem 2.1, the Cesàro means converge to . Therefore . ∎
3. Recapturing Complex analysis
We now use the machinery developed in the previous section to give a conceptually simple account of the remarkable regularity properties of holomorphic distributions. So we will pretend that we have forgotten everything about complex analysis, but do remember the rudiments of the theory of distributions, accounts of which can be found in the classic treatises [Hör03, Trè67, Sch66]. First we clarify notations and recall a few facts.
3.1. The basic spaces
For an open , the space of test functions is the -space of smooth compactly supported complex valued functions, topologized as the inductive limit of the Fréchet spaces consisting, for a given compact , of those elements of which have support in . Recall that a subset is bounded, if and only if there is a compact such that , and for each nonnegative multi-index , we have
(3.1) |
where, here and later, we will use standard multi-index conventions such as
(3.2) |
The space of distributions on is the dual of , consisting of continuous linear forms on . We denote the value of a distribution at a test function by The space is endowed with the usual strong dual topology. Recall that this topology is generated by the family of seminorms
(3.3) |
In this topology, the space is complete.
Given a locally integrable function , we can associate to a distribution defined by
where denotes the Lebesgue measure of . Then the locally integrable distribution is said to be generated by , and as usual we grant ourselves the the right to abuse language by identifying the distribution with the function .
We will use the abbreviations
(3.4) |
for the basic constant coefficient differential operators of complex analysis, acting on functions or distributions on . If is an open set, define
(3.5) |
the space of holomorphic distributions on . The subspace is closed in the space by the continuity of the operators , and is therefore a complete LCTVS in the subspace topology.
3.2. The main theorem
We begin with some definitions and notational conventions. For we denote by the Reinhardt rotation of by the element , the linear automorphism of the vector space given by
(3.6) |
A domain is defined to be Reinhardt if and only if for each . Let denote the union of the coordinate hyperplanes of :
(3.7) |
Recall that a Reinhardt domain is said to be log-convex, if whenever , the point belongs to if there is a such that for
(3.8) |
For , let be the monomial function given by
(3.9) |
For a Reinhardt subset and , let
(3.10) |
where is the closed disk. This can be thought of as the result of “completing” in the -th coordinate. Following [JP08], we say that the Reinhardt domain is relatively complete if for each , whenever we have , we also have
We prove the following well-known structure theorem for holomorphic distributions on Reinhardt domains, as an application of the ideas of Section 2:
Theorem 3.1.
Let be a Reinhardt domain in and let
(3.11) |
Let be the smallest relatively complete log-convex Reinhardt domain in that contains . The for each there is a continuous linear functional such that for each , the series
(3.12) |
converges absolutely in to a function , and generates the distribution .
Remarks:
-
(1)
For , all Reinhardt domains in the plane, i.e., disks and annuli, are automatically relatively complete and log-convex. For , it is easy to give examples of Reinhardt domains which are not log-convex, or not relatively complete, or perhaps both. For such a domain , it follows that each holomorphic distribution extends to a holomorphic function . This is the simplest example of Hartogs phenomenon, the compulsory extension of all holomorphic functions from a smaller domain to a larger one, characteristic of domains in several complex variables.
-
(2)
The functionals are called the coefficient functionals, and the series (3.12) is of course the Laurent series of the function (the Taylor series if ).
-
(3)
It is known (by a direct construction of a plurisubharmonic exhaustion) that relatively complete log-convex Reinhardt domains are pseudoconvex. This means that such a domain admits a holomorphic distribution whose Laurent expansion converges absolutely precisely on .
As an immediate consequence of Theorem 3.1 we have the following:
Corollary 3.1.
Let be open. Then each distribution is complex-analytic, i.e., for each there is a neighborhood of , where the function generating is represented by a Taylor series centered at .
3.3. Holomorphic functions and maps
A holomorphic distribution will be called a holomorphic function if it is generated by a function . We denote the space of holomorphic functions on temporarily by . Once Theorem 3.1 is proved, it will follow that .
Let be domains in . By a holomorphic map , we mean a mapping each of whose components is a holomorphic function on . A holomorphic map is a biholomorphism, if it is a bijection, and its set-theoretic inverse is also a holomorphic map. (It is of course known that the assumption of the holomorphicity of the inverse map is redundant, but this is a consequence of complex-analyticity, which is exactly what we are proving here). If is a biholomorphism, then for a distribution , we can define in the usual way the pullback distribution : if is generated by a test function , then is the distribution generated by the function , and for general , we extend this definition by continuity, using the density of test functions in , see [Hör03, Theorem 6.1.2] for details. Extending the chain rules for the complex derivative operators from test functions to distributions, we have the following relations analogous to [Hör03, formula (6.1.2)] for the Wirtinger derivatives (3.4):
(3.13) |
and
(3.14) |
where as above, is a biholomorphism of domains in , written in components as , and .
Therefore, we have the following immediate consequence of (3.14):
Proposition 3.2.
If is a biholomorphism and , then we have . If , then .
Therefore the spaces of holomorphic distributions and functions are invariant under pullbacks under biholomorphic maps. In fact, in the proof of Theorem 3.1, only two simple special cases of Proposition 3.2 noted below are needed:
-
(1)
Translation by a vector is the map
(3.15) which is obviously a biholomorphic automorphism of . For a domain , we therefore have a pullback isomorphism of spaces of holomorphic distributions . This can be thought of as an expression of the fact that the operator is translation invariant.
-
(2)
The Reinhardt rotations of (3.6) are clearly biholomorphic automorphisms of Reinhardt domains, and the pullback operation induces a representation of the group on the space of holomorphic distributions (see (3.32) below).
A domain is said to be Reinhardt centered at for an if there is a Reinhardt domain such that . Every open set in has local Reinhardt symmetry, in the sense that each point has neighborhood which is a Reinhardt domain centered at that point.
3.4. Mean value property of the monomials
It is easily verified by direct computation that for each , we have , where is the monomial of (3.9). We now note a remarkable symmetry property (the Mean-Value Property) of the functions :
Lemma 3.3.
Let , let , and let be a test function which has radial symmetry in each variable around the point , i.e., there are functions such that , and whose integral is 1:
(3.16) |
Then we have
(3.17) |
Proof.
First consider the case . If , the formula
(3.18) |
holds for , by expanding the integrand using the binomial formula, and integrating the finite sum term by term. The formula (3.18) also holds for , provided , and . This follows on noticing that we have an infinite series expansion
where by the -test, the convergence is uniform in , and then integrating the series on the right term by term.
3.5. The representation
If is a Reinhardt domain, then for each , the map of (3.6) is a biholomorphic automorphism of . Define a representation of on the space of test functions by
(3.20) |
Recall that a net converges in the space , if each is supported in a fixed compact , and the net converges in the Fréchet space i.e. all partial derivatives converge uniformly on . Using this it is easily verified that is a continuous representation of in the space .
Notice that is a Reinhardt domain. For a positive integer we define the norm with respect to the polar coordinates for a function in :
(3.21) |
where the tuples are the polar coordinates on specified by , and are partial derivatives operators in the polar coordinates defined as in (3.2). From the formulas
(3.22) |
we see that for , there is a constant such that for each compact set such that
we have
(3.23) |
We will need the following elementary estimate:
Proposition 3.4.
Let be the representation of on given by (3.20). For integers , and a compact there is a constant such that for each and each , such that for , we have
Proof.
For , define the -th Fourier coefficient of by
(3.24) |
where , and is the Lebesgue measure. We will also write for whenever convenient. We note the following properties:
-
(1)
We clearly have
(3.25) -
(2)
If , by a standard integration by parts argument, for each ,
(3.26) where, as usual, we set for and , .
-
(3)
Differentiating under the integral sign we have, for any ,
and combining this with (3.26), we see that if , we have for each integer that
(3.27)
Now for the evaluation , is continuous, so using (2.4) we see that for each and each , we have
(3.28) |
where in the last step we make a change of variables in the integral from to . Therefore, if with for , then
Now in the above formula, if we have , and let
where is the multi-index each of whose entries is 1, we obtain, after taking absolute values of both sides
(3.29) |
In the first factor on the right hand side, we have by hypothesis for each that and . Therefore, we have . The first factor can be estimated as
(3.30) |
Using (3.25), the second factor can be estimated as
(3.31) |
where in the last step we use the norm introduced in (3.21), and used the fact that
since each . Combining (3.29), (3.30) and (3.31) we see that
from which, taking a supremum on the left hand side over , and remembering that , we conclude that
Recall that is a compact set in such that , i.e. , the support of is contained in . Suppose that are such that , where is the product of annuli . Formulas (3.24) and (3.28) show that the compact support of is also contained in the set . Therefore, passing to the equivalent -norms using (3.23), we have
which completes the proof of the result. ∎
3.6. The dual representation
Let be a Reinhardt domain. Since for each , the map maps biholomorphically (and therefore diffeomorphically) to itself. Consequently, we can define a representation of on the space of distributions using the pullback operation
(3.32) |
The representation is closely related to the representation introduced in (3.20). Clearly, is an invariant (dense) subspace of , on which restricts to . So is simply the extension of the representation of (3.20) by continuity to the space of distributions.
The representation on is also “dual” to the representation on :
(3.33) |
so that is the transpose of the map . The second equality in the chain may be proved by change of variables when is a test function, and then using density.
Proposition 3.5.
The representation of on defined in (3.32) is continuous.
Proof.
Let be a sequence in converging to the identity element. Since pointwise convergence of a sequence in on each test function implies convergence in the strong dual topology (see [Trè67]), to show that in we need to show that for we have . But by (3.33), and by the continuity of the representation , to follows that in .
To complete the proof, by Proposition 2.3, we need to show that the topology of is generated by a collection of -invariant seminorms. Let be a bounded subset of . Let
Then it is clear from (3.1) that is also a bounded subset of . With notation as it (3.3), it is clear that for each , we have . Therefore the topology of is generated by the family of continuous invariant seminorms . ∎
3.7. Fourier series of distributions on Reinhardt domains
Let be a Reinhardt domain and let be a distribution. Then by the results of Section 2 we can expand in a formal Fourier series with respect to the representation of (3.32). For simplicity of notation, whenever there is no possibility of confusion, we denote the Fourier components of by
(3.34) |
so that the Fourier series of is written as . We notice the following properties of the Fourier components:
Proposition 3.6.
Let and be as above and let .
-
(a)
If the distribution lies in one of the linear subspaces or of , the Fourier component also lies in the same subspace.
-
(b)
If is another Reinhardt domain such that , we have
(3.35)
Proof.
All these are consequences of Part 2 of Proposition 2.4. To see Part (a), let be the space and let be one of the spaces or . Each of these has a natural structure of a complete LCTVS (the space as a closed subspace of , in its Fréchet topology and in its -topology), and for each one of these topologies, the representation restricts to a continuous representation (in the stronger topology of the subspace), which we will call (this coincides with the introduced in (3.20)). If is the inclusion map of any one of these subspaces into , then clearly it is continuous (where the subspace has the natural topology). Therefore, by (2.7), we have . By definition, the right hand side is precisely , i.e. the Fourier component of as an element of . Notice that , since it is the Fourier component of as an element of . Since is the inclusion of in , it now follows that as well.
Since the result is true for and it follows for .
For part (b), in part 2 of Proposition 2.4, let with representation defined in (3.32), which we call for clarity, and let with the representation and let be the restriction map of distributions, which is clearly continuous and intertwines the two representations. Then (2.7) takes the form . Applying this to the distribution , we see that
∎
We establish the following lemma:
Lemma 3.7.
Let be a Reinhardt domain, and let
Suppose that for each the Laurent series converges absolutely in the Fréchet space . Then the Laurent series of each element of converges absolutely in the space .
Proof.
For a function and a compact set , let be the seminorm:
(3.36) |
Notice that the family is a generating family of seminorms for , and the hypothesis on can be expressed as
(3.37) |
for each compact .
The Fréchet topology of is generated by the seminorms , where ranges over compact subsets of , is a constant coefficient differential operator on , and
We will continue to write for the seminorm (3.36) corresponding to . From the distributional chain rule (3.13), we have for each that
Let , and denote . Then, from the above we have
where . It follows that all Fourier coefficients of the distribution are in , so . Therefore,
It follows therefore that
using assumption (3.37) on elements of . A partial derivative on can be rewritten as a polynomial in the commuting differential operators , and by hypothesis the derivatives are zero. Therefore, it follows that . Consequently the series is absolutely convergent in .
∎
3.8. Proof of Theorem 3.1: case of annular domains
We now prove a weaker version of Theorem 3.1. A Reinhardt domain will be called annular if it is disjoint from the coordinate hyperplanes:
(3.38) |
where is as in (3.7). We have the following:
Proposition 3.8.
Let be an annular Reinhardt domain in . Then for each there is a continuous linear functional such that for each , the series of terms in given by converges absolutely in to a function , and this -smooth function generates the distribution .
Proof.
Let and let be its -th Fourier component as in (3.34), and let . The distributional Leibniz rule
(3.39) |
shows that , since by Part (1) of Proposition 3.6. By Part (1) of Proposition 2.4, the Fourier component lies in the -th Fourier mode of , i.e., , so
Therefore, using polar coordinates , with (with in the -th position) we have
Since , and in polar coordinates we have
(3.40) |
we see that
also. Now from the representations (3.22) we have that and in for . Therefore, by a classical result in the theory of distributions (see [Sch66, Theorème VI, pp. 69ff.]) we have that is locally constant on , or more precisely, is generated by a locally constant function. Since is connected, it follows that there is a constant which generates the distribution , so that This can also be written as . Since multiplication by the fixed -function is a continuous map on (and therefore on the subspace ), and is continuous by part 1 of Proposition 2.4, it follows that is continuous. But we know that takes values in the subspace of distributions generated by constants. By a well-known theorem, on a finite-dimensional topological vector space, there is only one topology. Therefore the topology induced from on the subspace of constants coincides with the natural topology of . Therefore is continuous.
By the above, the Fourier components of the Fourier series of actually lie in , and thus any partial sum (i.e. sum of finitely many terms) of this series also lies in . We will now show that the series (3.34) is absolutely convergent in the topology of , in the sense of Proposition 2.2. Let be compact. Let , and let be a test function such that
i.e. is radial around the origin in each complex coordinate. For , define so that is radially symmetric around in each complex direction. Therefore, for , we have, by Lemma 3.3 for each that
By the continuity of the mapping , there is a constant and an integer such that for all with support in , we have that
(3.41) |
Also notice that for any we have
using the invariance of Haar measure of a compact group under inversion | ||||
(3.42) |
In the following estimates, let denote a constant that depends only on the compact and the distribution , and may have different values at different occurrences. By combining (3.41) with (3.42), we see that for and each with , using Lemma 3.3
recalling that the local order depends only on the distribution and the compact set . Now, for each we have by translation invariance of the norm. Therefore, for each we obtain the estimate for the seminorm (with the same convention as above on the constant )
(3.43) |
Clearly therefore . By Lemma 3.7, the series converges absolutely in . Let be its sum. Since the inclusion is continuous, we see easily that the Fourier series of converges absolutely in . Now it follows from Corollary 2.5 that the sum of the series is . Thus is the distribution generated by , and this completes the proof of the proposition. ∎
3.9. Smoothness of holomorphic distributions
Corollary 3.9.
Let be an open set. Then each holomorphic distribution on is generated by a -smooth function, i.e. .
Proof.
Let . It suffices to show that is a smooth function in a neighborhood of each point . Without loss of generality, thanks to the invariance of the space of holomorphic functions under translations (Proposition 3.2 and comments after it) , we can assume that . Let be such that the polydisc given by
is contained in . We will show that is a smooth function on , i.e., .
By the previous section, holomorphic distributions on an annular Reinhardt domain are smooth. By translation invariance, the same is true if the annular domain is centered at a point different from the origin, i.e. for a domain of the form where is an annular Reinhardt domain centered at the origin, with as in (3.15) is the translation by .
Indeed, let , and set , so that is the annular Reinhardt domain centered at given by
for which it is easy to verify that . Now since is a holomorphic function, the result follows. ∎
3.10. Extension to the log-convex hull
In this section, we prove the following special case of Theorem 3.1 for annular Reinhardt domains:
Proposition 3.10.
Let be an annular Reinhardt domain in . Then the Laurent series of a distribution converges absolutely in where is the smallest log-convex annular Reinhardt domain containing .
Introduce the following notation: for a compact subset , we let
(3.44) |
Lemma 3.11.
Let be an annular Reinhardt domain in and let be the log-convex annular Reinhardt domain containing . Then for each point there is a compact neighborhood of in and a compact subset such that .
Proof.
Let be the map
(3.45) |
From (3.8), it follows that an annular Reinhardt domain is log-convex if and only if the subset of (called the logarithmic shadow of ) is convex. The set is characterized by the fact that it is the convex hull of . Therefore, if , there exist points and positive numbers , with such that for , we have , i.e.,
In other words lies in the convex hull of the -element set . (By a theorem of Carathéodory, , but we do not need this fact.) We will show that we can “fatten” one the points in this set so that the convex hull of the fattened set contains a neighborhood of the point .
Without loss of generality, we can assume that . Consider the affine self-map of given by
which is clearly an affine automorphism of . Therefore if is a compact neighborhood of in , then is a compact neighborhood of , which after shrinking can be taken to be contained in Let , and set
We claim that contains the compact neighborhood of the point . Let . Then there is a point such that , i.e. for each we have
For simplicity of writing, introduce new notation as follows: we let for and . Then for a multi-index and the point we have
(3.46) | ||||
(3.47) |
this shows that , and completes the proof. ∎
Proof of Proposition 3.10.
By Lemma 3.7 it is sufficient to show that the Laurent series (whose existence was proved in Proposition 3.8, and whose terms lie in ) converges absolutely in . It is sufficient to show that each point of has a compact neighborhood such that we have
Now by Lemma 3.11, there is a compact subset such that . We observe, using the definition (3.44) that
so that
using the estimate (3.43) which holds since . ∎
3.11. Laurent series on non-annular Reinhardt domains
We now consider the case of a general (i.e. possibly non-annular) Reinhardt domain . We begin by showing that the only monomials that occur in the Laurent series are the ones smooth on
Proposition 3.12.
Let be a (possibly non-annular) Reinhardt domain in . Then for each (where is as in (3.11)), there is a continuous linear functional such that the Fourier series of a is of the form
(3.48) |
Proof.
Since we have . Notice that is an annular Reinhardt domain and therefore the proof of Proposition 3.8 shows that the Fourier components are given for by
where is the -th coefficient functional associated to the domain (see Proposition 3.8). Thanks to (3.35), we know that , so we have
By Proposition 3.6, the Fourier components of a holomorphic distribution are holomorphic distributions, so we have since . Therefore, since by Corollary 3.9, each holomorphic distribution is a smooth function, we know that . Therefore, the function admits a extension through . If for some , this means that itself admits a extension to , i.e., , where is as in (3.11). Therefore if , the corresponding term in the Laurent series of vanishes, and the series takes the form
Now for each define the map by Since both the restriction map and the coefficient functional are continuous, it follows that is continuous. The extension of from to is still the monomial , so the Fourier series of in is of the form (3.48). ∎
3.12. Extension to the relative completion
Given a Reinhardt domain , its relative completion is the smallest relatively complete domain containing (see above before the statement of Theorem 3.1 for the definition of relative completeness of a domain.) Notice that the relative completion of coincides with the unions of the sets of (3.10), where the union is taken over those for which is nonempty. The following general proposition, which encompasses classical examples of the Hartogs phenomenon, e.g. in the “Hartogs figure”, will be needed to complete the proof of Theorem 3.1.
Proposition 3.13.
Let be a Reinhardt domain. Then each holomorphic function on extends holomorphically to the relative completion of .
Proof.
We may assume that since each Reinhardt domain in the plane is automatically relatively complete. If for each , the intersection , then the domain is annular, and its relative completion is itself so there is nothing to prove. Suppose therefore that there is such that . We need to prove that each function in extends holomorphically to .
Without loss of generality we can assume that . Write the coordinates of a point as where . Let . By Proposition 3.12 admits a Laurent series representation
with as in (3.11), and the series converges absolutely in . To prove the proposition it suffices to show that the series in fact converges absolutely in , which by Lemma 3.7 is equivalent to the following: for each point there is a compact neighborhood of in such that
(3.49) |
We claim the following: for each there is a compact neighborhood of and a compact subset such that for each . Assuming the claim for a moment, we see that we have
so that (3.49) follows since by (3.43) we do have for a compact subset of . Since each point has such a neighborhood this completes the proof, modulo the claim above.
To establish the claim, we may assume that , since otherwise there is nothing to prove. Therefore , and consequently, there is a and a such that , where . Let be a compact neighborhood of in of the form , where and is a closed disk. Let be the disk , so that , and
is a point of maximum modulus (i.e. maximum distance from the origin) in both sets and . Note that
We set . We will show that these sets satisfy the conditions of the claim.
Now let . Since , it follows that . Let , and set
so that we have
On the other hand
Consequently, in fact we have , and the claim is proved, thus completing the proof. ∎
3.13. End of proof of Theorem 3.1
It only remains to put together the various pieces to note that all parts of Theorem 3.1 have been established. If is annular, i.e. has empty intersection with the set of (3.7), then Proposition 3.8 takes care of the complete proof. When is allowed to be non-annular, we see from Proposition 3.12 that the Laurent series representation has only monomials which are smooth functions on . Now it is not difficult to see that the smallest log-convex relatively complete Reinhardt domain containing can be constructed from in two steps. First, we construct the log-convex hull of the set . Notice that and are both annular. The second step consists of constructing the relative completion of the domain , thus obtaining the domain . Now by Proposition 3.10, the Laurent series of a holomorphic distribution on converges absolutely in . Applying Proposition 3.13 (with ), we see that the Laurent series actually converges absolutely in the space . The sum of this series is the required holomorphic extension of a given holomorphic distribution on . The result has been completely established.
4. Missing monomials
In Theorem 3.1, we considered the natural representation of the torus on the space of holomorphic functions on a Reinhardt domain . In applications, one often deals with a subspace of functions such that
-
(1)
the subspace is invariant under the natural representation , i.e., if then for each , where is as in (3.6),
-
(2)
there is a locally convex topology on in which it is complete, and such that the inclusion map
is continuous, and
-
(3)
when is given this topology, the representation restricts to a continuous representation on .
The locus classicus here is the theory of Hardy spaces on the disc. We can make the following elementary observation:
Proposition 4.1.
Let and be as above, and set
Then the Laurent series of a function is of the form
Proof.
It suffices to show that if for , if the monomial does not belong to , we have that , where is the Laurent coefficient of as in (3.12). By part (1) of Proposition 2.4, we see that . By part (2) of the same proposition, taking to be the space , for we have that , and from the description of the Fourier components of a holomorphic function in the proof of Theorem 3.1, we see that . Therefore , which contradicts the fact that unless . ∎
This simple observation can be called the “principle of missing monomials”, since it says that certain monomials cannot occur in the Laurent series of the function . It can be thought to be the reason behind several phenomena associated to holomorphic functions. We consider two examples:
-
(1)
Bergman spaces in Reinhardt domains: Let be a Reinhardt domain in and let be a radial weight on , i.e., for we have
The -Bergman space is defined to be the subspace of the weighted -space consisting of holomorphic functions, where the norm on the weighted -space is given by
where is the Lebesgue measure. It is well-known that is a closed subspace of the Banach space and therefore a Banach space ([DS04]). It is also easy to see (using standard facts about -spaces) that the natural representation of on is continuous for , so it follows that the representation on is also continuous. It now follows from Proposition 4.1 that the Laurent series expansion of a function consists only of terms with monomials such that . The case of this fact was deduced by a different argument in [CEM18].
-
(2)
Extension of holomorphic functions smooth up to the boundary: Let be a Reinhardt domain such that the origin (which is the center of symmetry) is on the boundary of . A classic example of this is the Hartogs triangle in . In [Cha19], the following extension theorem was proved: there is a complete Reinhardt domain such that and each function in the space of holomorphic functions on smooth up to the boundary extends holomorphically to the domain . This was noted for the Hartogs triangle in [Sib75], where is the unit bidisk.
To deduce this from the principle of missing monomials, it suffices to consider the case when is bounded. We notice that is a Fréchet space in its usual topology of uniform convergence on with all partial derivatives. A generating family of seminorms for this topology is given by the norms where . The natural representation on restricts to a continuous representation on . Therefore, the principle of missing monomials applies and the only monomials that occur in the Laurent expansion of a function are such that . For such a multi-index , write the multiindex , where and . Then , and we can apply the differential operator to obtain
Since this means that , which is possible only if . Thus , and the Laurent series of each function is a Taylor series which converges in some complete (log-convex) Reinhardt domain , and this must strictly contain , since is not complete.
5. Classical characterizations of holomorphic functions
In this section we show how one can avoid the machinery of generalized functions and weak derivatives altogether, and still use Fourier methods to prove the basic facts of function theory. We confine ourselves to one complex variable and the simple geometry of the disk.
5.1. Goursat’s characterization
The starting point of a traditional account of holomorphic functions of a single variable is typically Goursat’s definition ( [Gou00]): a function on an open subset is holomorphic, if it is complex differentiable, i.e., for each point , the limit
(5.1) |
exists. The result that a holomorphic function in this sense is infinitely many times complex differentiable and even admits a convergent power series representation near each point is rightly celebrated as one of the most elegant and surprising in all of mathematics. Unfortunately, we cannot use it as a definition, if we want to apply the theory of abstract Fourier expansions as developed in Section 2. Denoting by the collection of holomorphic functions in the sense of Goursat in an open set , we notice that the space does not have a nice a priori linear locally convex topology in which it is complete and such that when is a disc or an annulus, the natural action of the group on the space is a continuous representation. Though Goursat’s definition carries the weight of a century of academic tradition, we will start from an alternative definition which lends itself better to the application of the methods of Section 2. We also note that the characterization of holomorphic functions by complex-differentiability cannot be used for natural generalizations of complex analysis, e.g. quaternionic analysis, analysis on Clifford algebras etc. (see [GM91, pp. 87–93]).
5.2. Morera’s definition
Let be an open subset of the complex plane , and let be a continuous function. In honor of [Mor02], let us say that the function is holomorphic in the sense of Morera (Morera-holomorphic for short) if for each triangle contained (with its interior) in , we have the vanishing of the complex line integral of around the boundary of :
(5.2) |
where denotes the boundary of , oriented counterclockwise. For an open set , let us denote by the collection of Morera-holomorphic functions on . It is known by Morera’s theorem that a Morera-holomorphic function is Goursat-holomorphic, and one can develop function theory starting from Morera’s definition (see [Hef55, MW67]). Notice that the a priori regularity of Morera-holomorphic functions (assumed to be only continuous) is even less than that assumed for Goursat-holomorphic functions (assumed also to admit the limit (5.1) at each .)
It is immediate from the definition that is a closed linear subspace of the Fréchet space of continuous functions. “Closed” means that the limit of a sequence of Morera-holomorphic functions converging uniformly on compact subsets of is itself Morera-holomorphic, a fact that was already noted in [Mor86]. The proof of this crucial fact starting from the Goursat definition must pass through a lengthy development of integral representations, so this is definitely a pedagogical advantage of Morera’s definition over Goursat’s.
The notion of Morera-holomorphicity is local: i.e., if and only if there is an open cover of such that for each . One half of this claim is trivial, and for the other half, for a triangle in , we can perform repeated barycentric subdivisions till the triangles so formed are each contained in some element of the open cover . We therefore conclude that is a sheaf of Fréchet spaces on .
The following local description of Morera-holomorphic functions is well-known, and a proof can be found in e.g. [Rem91, pp. 186-189].
Proposition 5.1.
Let be convex. Then the following statements about a continuous function are equivalent:
-
(A)
.
-
(B)
has a holomorphic primitive, i.e., there is an which is complex-differentiable on and .
-
(C)
for each piecewise closed path in we have
(5.3)
Recall that assuming (A), the primitive in (B) is constructed by fixing , and setting where denotes the line segment from to . Proposition 5.1 allows us to give examples of Morera-holomorphic functions. Recall from (3.9) that for an integer , we use the notation for the holomorphic monomials.
Proposition 5.2.
If then and if then .
Proof.
First note that is continuous, on all of if and on if . If for , we can verify from the definition (5.1) that is complex-differentiable and so that by part (B) of Proposition 5.1 the result follows for . For , we can construct for each , a local primitive of near by setting
where denotes a branch of the argument defined near the point . A direct computation shows that near , so that again we see that ∎
5.3. Products of Morera-holomorphic functions
In the proof of Theorem 3.1, an important role is played by the fact that if is a holomorphic distribution (in the sense of (3.5))and is a holomorphic function (i.e. a holomorphic distribution which is , Section 3.3 above), then the product distribution is also a holomorphic distribution. This is an immediate consequence of the distributional Lebniz formula (3.39). A similar result, proved in [MW67], will be needed in order to develop the properties of holomorphic functions starting from Morera’s definition.
Proposition 5.3.
Let , and assume that is locally Lipschitz at each point, i.e. for each and each compact such that , there is an such that for we have
(5.4) |
Then the product also belongs to .
The proof is based on a version of the classical Goursat lemma ([Gou00, Pri01]). This is of course the main ingredient in the standard textbook proof of the Cauchy theorem for triangles for Goursat-holomorphic functions. Recall that two triangles are similar if they have the same angles.
Lemma 5.4.
Let be an open subset of and let be a complex valued function defined on the set of triangles contained in such that the following two conditions are satisfied:
-
(1)
is additive in the following sense: if a triangle is represented as a union of smaller triangles with pairwise disjoint interiors then
(5.5) -
(2)
For each and each triangle , we have
(5.6) where denotes the area of the triangle , and the limit is taken along the family of triangles similar to and containing the point , as these triangles shrink to the point .
Then .
Proof.
For completeness, we recall the classic argument. Let be a triangle contained in . We construct a sequence of triangles with using the following recursive procedure. Assuming that has been constructed, we divide into four similar triangles with half the diameter of by three line segments each parallel to a side of and passing through the midpoints of the other two sides. Denote the four triangles so obtained by . Then, by (5.5), we have
Choose to be one of such that the value of is the largest. Then, by the triangle inequality we have and by induction it follows that
(5.7) |
where in the last step we have used the fact that has one-fourth the area of , so . Since the diameters of the go to zero, by compactness, there is a unique point in the intersection . Since the family is a subfamily of all the triangles containing , and each is similar to , therefore by letting in (5.7) and using (5.6) the result follows. ∎
Proof of Proposition 5.3.
For a triangle , define
To prove the result, we need to show that . Since condition (5.5) of Lemma 5.4 is obvious, we need to show (5.6) to complete the proof. Let , let be a compact neighborhood of in , and denote by the Lipschitz constant corresponding to this and this in (5.4). Now, let be a triangle and let be a triangle similar to such that . Then observe that, by the hypothesis of Morera-holomorphicity of and we have
Therefore, denoting the perimeter of the triangle by ,
where in the last step we use the fact that and are similar, so the quantity in parentheses (which is clearly invariant under dilations) is the same. So we have
for a constant independent of the triangle as long as is similar to and . Letting shrink to , we have (5.6) and the proof is complete. ∎
5.4. Fourier expansion of a Morera-holomorphic function
We will now prove the following analog of Theorem 3.1. In particular, it shows that holomorphic functions in the sense of Morera are identical to the holomorphic distributions considered in Section 3.
Theorem 5.1.
Let where is the open unit disc. Then there is a sequence of complex numbers, such that
(5.8) |
where is as in (3.9), and the series on the right converges absolutely in to the function .
Let be the natural representation of on given by
(5.9) |
Proposition 5.5.
The space is invariant under and the resulting representation of on is continuous.
Proof.
Let , let be a triangle in . Notice that
is itself a triangle, and we have
It follows that is invariant under .
It suffices to show that the representation is continuous on . For , let be the seminorm on given by
(5.10) |
It is clear that for each and . Also in if and only if for each , we have , so the family is a -invariant family of seminorms that generates the topology of . Further, given , by uniform continuity, for each , we have
so that in the space . Therefore both conditions of Proposition 2.3 are satisfied, and the representation is continuous. ∎
In view of the above, the machinery of Section 2 applies. We now compute the Fourier components (2.5).
Proposition 5.6.
For , and with the natural representation (5.9), the Fourier components of are of the form:
where and .
The proof will use the following lemma:
Lemma 5.7.
A radial function in is constant.
Proof.
Let be radial, and define the complex valued continuous function on the interval by restriction, , so that we have by the radiality of . To prove the theorem, it suffices to show that is a constant .
Fix and . For in the interval consider the curvilinear quadrilateral defined by
(5.11) |
and notice that lies in the upper half disc, which is convex. The region is bounded by the two circular arcs
along with the two radial line segments
Orient the boundary counterclockwise. We can write
where the vanishing of the integral follows from part (c) of Proposition 5.1 above. Parametrizing by where , and using we get
Similarly,
Now, parametrizing by , we have , so
Similarly,
Therefore, adding the four integrals we have
so that we have the relation
(5.12) |
which can be written as
Since is continuous, the right hand side is a function of which is continuously differentiable on . Thus . Differentiating both sides of (5.12) with respect to we have so that we have for . But can be chosen arbitrarily, so it follows that on and the result follows. ∎
Proof of Proposition 5.6.
Let and for simplicity of notation, let . Then by Proposition 2.4, we see that lies in the Fourier mode so that and for we have . Since clearly , it follows by Proposition 5.3 that the function on defined by
lies in . Further for we have
so is radial, and hence by Lemma 5.7, is a constant, which we call . Therefore, on , we have , so the product extends for each to a continuous function on . If , this is possible only if , and this completes the proof. ∎
5.5. Conclusion of the proof of Theorem 5.1
We will first show that the series converges absolutely in the space . It suffices to show that for each , we have
where is as in (5.10). Fix an with and let be such that . Applying the Cauchy estimate (2.9) to the seminorm we see that for each :
Thanks to Proposition 5.6 we have for each ,
Therefore we have
This proves that the series converges absolutely in . Let be its sum. Since the partial sums are all in by Proposition 5.2, we see that . However, it follows from Corollary 2.5 that the sum of the series is , therefore , and the proof of the theorem is complete.
5.6. Pompeiu’s characterization of holomorphic functions
Goursat’s definition is non-quantitative, since it is framed in terms of the existence of the limit (5.1), and does not provide a way to measure the degree of non-holomorphicity of a function. For example, if is small and nonzero, all it says about the functions and is that both are non-holomorphic. Morera’s definition of a holomorphic function does not have this shortcoming, as the integral gives a measure of the amount of non-holomorphicity of on the triangle .
It is possible to normalize and localize this measure of non-holomorphicity, as was realized by Pompeiu (see [Pom12, MŞ98]). The quantity
(5.13) |
is a numerical measure of the “average density of non-holomorphicity” of a continuous function on a triangle (or other region with piecewise smooth boundary) , where denotes the area of . To localize this, we can consider, for a point , and a continuous function , the following limit (called the “areolar derivative” by Pompieu) as a measure of the degree of non-holomorphicity of at the point :
(5.14) |
where the limit is taken over the family of triangles containing the point and contained in , as these triangles shrink to the point . Let us say that a function is Pompeiu-holomorphic if the limit (5.14) exists at each , and is equal to zero. The following simple and well-known observations clarify the meaning of this notion:
Proposition 5.8.
-
(a)
A continuous function is Pompieu-holomorphic if and only if it is Morera-holomorphic.
-
(b)
If the function is in , then the limit (5.14) exists, and is equal to .
Proof.
For part (a), if is Morera-holomorphic, then the quantity (5.13) vanishes for each triangle , so the limit (5.14) vanishes. Conversely, suppose that the limit (5.14) vanishes at each point , and let be a triangle in . In Lemma 5.4, if we take , then the additivity condition (5.5) is clear, and the limit condition (5.6) holds by hypothesis. The result follows.
For part (b), using Stoke’s theorem
Since the last integral has a continuous integrand, we may take the limit as the triangle shrinks to the point to obtain:
∎
References
- [BDS82] F. Brackx, Richard Delanghe, and F. Sommen. Clifford analysis, volume 76 of Research Notes in Mathematics. Pitman (Advanced Publishing Program), Boston, MA, 1982.
- [Bou04] Nicolas Bourbaki. Integration. I. Chapters 1–6. Elements of Mathematics (Berlin). Springer-Verlag, Berlin, 2004. Translated from the 1959, 1965 and 1967 French originals by Sterling K. Berberian.
- [CEM18] D. Chakrabarti, L. D. Edholm, and J. D. McNeal. Duality and approximation of Bergman spaces. ArXiv e-prints, April 2018.
- [CH53] R. Courant and D. Hilbert. Methods of mathematical physics. Vol. I. Interscience Publishers, Inc., New York, N.Y., 1953.
- [Cha19] Debraj Chakrabarti. On an observation of Sibony. Proc. Amer. Math. Soc., 147(8):3451–3454, 2019.
- [Daw21] Anirban Dawn. Laurent series of holomorphic functions smooth up to the boundary. Complex Anal. Synerg., 7(2):Paper No. 13, 8, 2021.
- [DS04] Peter Duren and Alexander Schuster. Bergman spaces, volume 100 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2004.
- [GM91] John E. Gilbert and Margaret A. M. Murray. Clifford algebras and Dirac operators in harmonic analysis, volume 26 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1991.
- [Gou00] E. Goursat. Sur la définition générale des fonctions analytiques, d’après Cauchy. Trans. Amer. Math. Soc., 1(1):14–16, 1900.
- [Hef55] Lothar Heffter. Begründung der Funktionentheorie auf alten und neuen Wegen. Springer-Verlag, Berlin-Göttingen-Heidelberg, 1955.
- [Hör03] Lars Hörmander. The analysis of linear partial differential operators. I. Classics in Mathematics. Springer-Verlag, Berlin, 2003. Distribution theory and Fourier analysis, Reprint of the second (1990) edition [Springer, Berlin; MR1065993 (91m:35001a)].
- [Joh76] Russell A. Johnson. Representations of compact groups on topological vector spaces: some remarks. Proc. Amer. Math. Soc., 61(1):131–136 (1977), 1976.
- [JP08] Marek Jarnicki and Peter Pflug. First steps in several complex variables: Reinhardt domains. EMS Textbooks in Mathematics. European Mathematical Society (EMS), Zürich, 2008.
- [KK97] Mikhail I. Kadets and Vladimir M. Kadets. Series in Banach spaces, volume 94 of Operator Theory: Advances and Applications. Birkhäuser Verlag, Basel, 1997. Conditional and unconditional convergence, Translated from the Russian by Andrei Iacob.
- [Mor86] Giacinto Morera. Un teorema fondamentale nella teorica delle funzioni di una variable complessa. Ist. Lombardo, Rend., II. Ser., 19:304–307, 1886.
- [Mor02] Giacinto Morera. Sulla definizione di funzione di una variabile complessa. Atti della Accademia delle Scienze di Torino, 37:99–102, 1902.
- [MŞ98] Marius Mitrea and Florin Şabac. Pompeiu’s integral representation formula. History and mathematics. Rev. Roumaine Math. Pures Appl., 43(1-2):211–226, 1998. Collection of papers in memory of Martin Jurchescu.
- [MW67] A. J. Macintyre and W. John Wilbur. A proof of the power series expansion without differentiation theory. Proc. Amer. Math. Soc., 18:419–424, 1967.
- [Pom12] Dimitrie Pompéiu. Sur une classe de fonctions d’une variable complexe. Rend. Circ. Mat. Palermo, 33:108–113, 1912.
- [Pri01] Alfred Pringsheim. Ueber den Goursat’schen Beweis des Cauchy’schen Integralsatzes. Trans. Amer. Math. Soc., 2(4):413–421, 1901.
- [Ran86] R. Michael Range. Holomorphic functions and integral representations in several complex variables, volume 108 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1986.
- [Rem91] Reinhold Remmert. Theory of complex functions, volume 122 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1991. Translated from the second German edition by Robert B. Burckel, Readings in Mathematics.
- [Rud91] Walter Rudin. Functional analysis. International Series in Pure and Applied Mathematics. McGraw-Hill, Inc., New York, second edition, 1991.
- [Sch66] Laurent Schwartz. Théorie des distributions. Publications de l’Institut de Mathématique de l’Université de Strasbourg, No. IX-X. Nouvelle édition, entiérement corrigée, refondue et augmentée. Hermann, Paris, 1966.
- [Sib75] Nessim Sibony. Prolongement des fonctions holomorphes bornées et métrique de Carathéodory. Invent. Math., 29(3):205–230, 1975.
- [Trè67] François Trèves. Topological vector spaces, distributions and kernels. Academic Press, New York-London, 1967.
- [Wey40] Hermann Weyl. The method of orthogonal projection in potential theory. Duke Math. J., 7:411–444, 1940.
- [Zyg02] A. Zygmund. Trigonometric series. Vol. I, II. Cambridge Mathematical Library. Cambridge University Press, Cambridge, third edition, 2002. With a foreword by Robert A. Fefferman.