This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Power series as Fourier Series

Debraj Chakrabarti Department of Mathematics, Central Michigan University, Mount Pleasant, MI 48859, USA [email protected]  and  Anirban Dawn Department of Mathematics, The University of Tampa, Tampa, FL 33606, USA [email protected]
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:
32A05,22D12
Debraj Chakrabarti was partially supported by Simons Foundation Collaboration Grant number 706445.

1. 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 n{\mathbb{C}}^{n}. 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 nn-dimensional torus 𝕋n\mathbb{T}^{n} 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 1πz\frac{1}{\pi z} of the Cauchy-Riemann operator z¯\frac{\partial}{\partial\overline{z}} 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 z¯u=0\frac{\partial}{\partial\overline{z}}u=0 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 z¯\frac{\partial}{\partial\overline{z}} to conclude that holomorphic distributions are real-analytic, we discard all of this information, except that holomorphic functions are 𝒞1\mathcal{C}^{1} 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 𝒞\mathcal{C}^{\infty}-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 zz1α1znαn,z\mapsto z_{1}^{\alpha_{1}}\dots z_{n}^{\alpha_{n}},
    αj\alpha_{j}\in\mathbb{Z} and

  • (c)

    the fact that radially symmetric holomorphic distributions are constants.

It is also interesting that the fact that 1πz\frac{1}{\pi z} 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 𝕋n\mathbb{T}^{n} by that of the special orthogonal group SO(n)SO(n), the method can be used to show that a harmonic distribution in a ball of n{\mathbb{R}}^{n} 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 zzn,nz\mapsto z^{n},n\in\mathbb{Z} (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 XX be a locally convex Hausdorff topological vector space (we use the standard abbreviation LCTVS). Recall that the topology of XX can also be described by prescribing the family of continuous seminorms on XX: a net {xj}\{x_{j}\} in XX converges to xx, if and only if for each continuous seminorm pp on XX, we have p(xxj)0p(x-x_{j})\to 0 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 {pk:kK}\{p_{k}:k\in K\} on XX is said to generate the topology of XX if for every continuous seminorm qq on XX, there exists a finite subset {k1,,kn}K\{k_{1},\dots,k_{n}\}\subset K and a C>0C>0 such that

q(x)Cmax{pk1(x),,pkn(x)}for all xX,q(x)\leq C\cdot\max\{p_{k_{1}}(x),\dots,p_{k_{n}}(x)\}\quad\textrm{for all $x\in X$}, (2.1)

and further, for every nonzero xXx\in X there exists at least one kKk\in K such that pk(x)0p_{k}(x)\neq 0 (this separating property ensures that the topology of XX is Hausdorff). Then clearly a net {xj}\{x_{j}\} converges in XX if and only if pk(xjx)0p_{k}(x_{j}-x)\rightarrow 0 for each kKk\in K.

Let Γ\Gamma be a directed set with order \geq. Recall that a net {xα}αΓ\{x_{\alpha}\}_{\alpha\in\Gamma} in XX is said to be Cauchy if for every ϵ>0\epsilon>0 and every continuous seminorm pp on XX, there exists γΓ\gamma\in\Gamma such that whenever α,βΓ\alpha,\beta\in\Gamma and α,βγ\alpha,\beta\geq\gamma, we have p(xαxβ)<ϵp(x_{\alpha}-x_{\beta})<\epsilon. The space XX is said to be complete if every Cauchy net in XX converges. Observe that in the above definition we can use a generating family of seminorms rather than all continuous seminorms.

If {Sk}k\{S_{k}\}_{k\in\mathbb{N}} is a sequence in a vector space, we can define the sequence of the corresponding Cesàro means {Cn}n\{C_{n}\}_{n\in\mathbb{N}} by

Cn=1n+1k=0nSk.\displaystyle{C_{n}=\frac{1}{n+1}\sum\limits_{k=0}^{n}S_{k}}.

The following is the analog for sequences in LCTVS of an elementary fact well-known for sequences of numbers:

Proposition 2.1.

Let {Sk}k\{S_{k}\}_{k\in\mathbb{N}} be a convergent sequence in an LCTVS XX. Then the sequence of Cesàro means {Cn}n\{C_{n}\}_{n\in\mathbb{N}} is also convergent, and has the same limit.

Proof.

Let S=limkSkS=\lim\limits_{k\to\infty}S_{k}. If pp is a continuous seminorm on XX and ϵ>0\epsilon>0, there is N1N_{1} such that p(SkS)<ϵ/2p(S_{k}-S)<\epsilon/2 for kN1k\geq N_{1}. Set s=k=0N1p(SkS)s=\sum_{k=0}^{N_{1}}p(S_{k}-S). Then for n>N1n>N_{1}, we have

p(CnS)=p(1n+1k=0nSkS)1n+1k=0np(SkS)<sn+1+nN1n+1ϵ2.\displaystyle p(C_{n}-S)=p\left(\frac{1}{n+1}\sum\limits_{k=0}^{n}S_{k}-S\right)\leq\frac{1}{n+1}\sum\limits_{k=0}^{n}p(S_{k}-S)<\frac{s}{n+1}+\frac{n-N_{1}}{n+1}\cdot\frac{\epsilon}{2}.

Therefore, if we choose N>N1N>N_{1} so large that sn+1<ϵ2\frac{s}{n+1}<\frac{\epsilon}{2}, then for nNn\geq N we have p(CnS)<ϵp(C_{n}-S)<\epsilon, so CnSC_{n}\to S in XX. ∎

For a formal series j=0xj\displaystyle{\sum_{j=0}^{\infty}x_{j}} in an LCTVS XX, convergence is defined in the usual way, i.e. the sequence of partial sums converges in XX. A formal sum α𝔄xα{\sum_{\alpha\in\mathfrak{A}}x_{\alpha}} over a countable index set 𝔄\mathfrak{A}, where xαx_{\alpha} are vectors in an LCTVS XX, is said to be absolutely convergent if there exists a bijection τ:𝔄\tau:\mathbb{N}\rightarrow\mathfrak{A} such that for every continuous seminorm pp on XX, the series of non-negative real numbers j=0p(xτ(j))\displaystyle{\sum_{j=0}^{\infty}p(x_{\tau(j)})} 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 XX is a locally compact Hausdorff space, absolute convergence in the Fréchet space 𝒞(X)\mathcal{C}(X) of continuous complex valued functions on XX 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 XX be a complete LCTVS, and let α𝔄xα{\sum_{\alpha\in\mathfrak{A}}x_{\alpha}} be an absolutely convergent series of elements of XX. Then the series is unconditionally convergent: there is an sXs\in X such that for every bijection θ:𝔄\theta:\mathbb{N}\rightarrow\mathfrak{A}, the series j=0xθ(j)\sum_{j=0}^{\infty}x_{\theta(j)} converges in XX to ss.

The element sXs\in X is naturally called the sum of the series, and we write s=α𝔄xαs=\sum_{\alpha\in\mathfrak{A}}x_{\alpha}.

Proof.

By definition, there exists a bijection σ:𝔄\sigma:\mathbb{N}\rightarrow\mathfrak{A}, such that for each continuous seminorm pp on XX, the series j=0p(xσ(j))\sum_{j=0}^{\infty}p(x_{\sigma(j)}) converges. Let yj=xσ(j)y_{j}=x_{\sigma(j)} and sk=j=0kyjs_{k}=\sum_{j=0}^{k}y_{j}. Since j=0p(yj)\sum_{j=0}^{\infty}p(y_{j}) converges, for ϵ>0\epsilon>0 there exists N0N_{0}\in\mathbb{N} such that whenever m,m,\ell\in\mathbb{N} with mN0m\geq\ell\geq N_{0}, j=+1mp(yj)<ϵ\sum_{j=\ell+1}^{m}p(y_{j})<\epsilon. Therefore for mN0m\geq\ell\geq N_{0},

p(sms)=p(j=+1myj)j=+1mp(yj)<ϵ.p(s_{m}-s_{\ell})=p\Big{(}\sum_{j=\ell+1}^{m}y_{j}\Big{)}\leq\sum_{j=\ell+1}^{m}p(y_{j})<\epsilon. (2.2)

Therefore {sk}\{s_{k}\} is Cauchy sequence in the complete LCTVS XX, and therefore converges to an sXs\in X. In order to complete the proof, it suffices to show that for every bijection τ:\tau:\mathbb{N}\rightarrow\mathbb{N}, the series j=0yτ(j)\sum_{j=0}^{\infty}y_{\tau(j)} converges to the same sum ss. Let skτ=j=0kyτ(j)s_{k}^{\tau}=\sum_{j=0}^{k}y_{\tau(j)}. Choose uu\in\mathbb{N} such that the set of integers {0,1,2,,N0}\{0,1,2,\cdots,N_{0}\} is contained in the set {τ(0),τ(1),,τ(u)}\{\tau(0),\tau(1),\cdots,\tau(u)\}. Then, if k>uk>u, the elements y1,,yN0y_{1},\cdots,y_{N_{0}} get cancelled in the difference skskτs_{k}-s_{k}^{\tau} and we have p(skskτ)<ϵp(s_{k}-s_{k}^{\tau})<\epsilon by (2.2). This proves that the sequences {sk}\{s_{k}\} and {skτ}\{s_{k}^{\tau}\} converge to the same limit. So, skτss_{k}^{\tau}\rightarrow s as kk\rightarrow\infty. ∎

By a famous theorem of Dvoretzky and Rogers, the converse of the above result fails when XX 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 KK be a compact Hausdorff space and let μ\mu be a Borel measure on KK, and let ff be a continuous map from KK to an LCTVS XX. An element xXx\in X is called a Pettis integral of ff on KK with respect to μ\mu if for all ϕX\phi\in X^{\prime},

ϕ(x)=K(ϕf)𝑑μ,\phi(x)=\int_{K}(\phi\circ f)\,d\mu, (2.3)

where XX^{\prime} denotes the dual space of XX (the space of continuous linear functionals on XX), and the right hand side of (2.3) is an integral of a continuous function. If XX is complete, one can show that there exists a unique xXx\in X such that (2.3) holds and we denote the Pettis integral of ff on KK with respect to μ\mu by Kf𝑑μ=x\displaystyle{\int_{K}f\hskip 2.84526ptd\mu=x}. In fact, the integral exists uniquely, as soon as the space XX is quasi-complete, i.e. if each bounded Cauchy net in XX converges, where a net {xα}αΓ\{x_{\alpha}\}_{\alpha\in\Gamma} is bounded if for each continuous seminorm pp, the net of real numbers {p(xα)}αΓ\{p(x_{\alpha})\}_{\alpha\in\Gamma} is bounded. While there are situations in which this more refined existence theorem for Pettis integrals is useful (e.g. when the space XX 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 T:XYT:X\to Y is a continuous linear map of complete LCTVSs, and f:KXf:K\to X is continuous then we have

T(Kf𝑑μ)=KTf𝑑μ,T\left(\int_{K}f\,d\mu\right)=\int_{K}T\circ f\,d\mu, (2.4)

since for any linear functional ψY\psi\in Y^{\prime},

ψ(T(Kf𝑑μ))=(ψT)(Kf𝑑μ)=K(ψT)(f)𝑑μ\psi\left(T\left(\int_{K}f\,d\mu\right)\right)=(\psi\circ T)\left(\int_{K}f\,d\mu\right)=\int_{K}(\psi\circ T)(f)\,d\mu

using the fact that ψTX\psi\circ T\in X^{\prime} and the definition (2.3) of the Pettis integral.

2.2. Representation of the torus on an LCTVS XX.

Let

𝕋n={λ=(λ1,,λn)n:|λj|=1for every 1jn}\mathbb{T}^{n}=\{\lambda=(\lambda_{1},...,\lambda_{n})\in{\mathbb{C}}^{n}:\left|\lambda_{j}\right|=1\hskip 2.84526pt\text{for every $1\leq j\leq n$}\}

be the nn-dimensional unit torus. With the subspace topology inherited from n{\mathbb{C}}^{n} and binary operation defined as (λ,ξ)λξ=(λ1ξ1,,λnξn)(\lambda,\xi)\mapsto\lambda{\cdot}\xi=(\lambda_{1}\xi_{1},\cdots,\lambda_{n}\xi_{n}), is a compact abelian topological group. For a LCTVS XX, and a continuous function f:𝕋nXf:\mathbb{T}^{n}\to X, we denote the Pettis integral of ff with respect to the Haar measure of 𝕋n\mathbb{T}^{n} (normalized to be a probability measure) by

Tnf(λ)𝑑λ.\int_{T^{n}}f(\lambda)d\lambda.

Let XX be an LCTVS, and let λσλ\lambda\mapsto\sigma_{\lambda} be a continuous representation of 𝕋n\mathbb{T}^{n} on XX. Recall that this means that for each λ𝕋n\lambda\in\mathbb{T}^{n}, the map σλ\sigma_{\lambda} is an automorphism (i.e. linear self-homeomorphism) of XX as a topological vector space, the map λσλ\lambda\mapsto\sigma_{\lambda} is a group homomorphism from 𝕋n\mathbb{T}^{n} to Aut(X)\mathrm{Aut}(X), and the associated map

σ:𝕋n×XX,σ(λ,x)=σλ(x)\sigma:\mathbb{T}^{n}\times X\to X,\quad\sigma(\lambda,x)=\sigma_{\lambda}(x)

is continuous.

Given a representation σ\sigma of the group 𝕋n\mathbb{T}^{n} on an LCTVS XX, a continuous seminorm pp on XX is said to be invariant (with respect to σ\sigma) if p(σλ(x))=p(x)p(\sigma_{\lambda}(x))=p(x) for all xXx\in X and λ𝕋n\lambda\in\mathbb{T}^{n}.

Proposition 2.3.

A representation σ\sigma of 𝕋n\mathbb{T}^{n} on an LCTVS XX is continuous if and only if the following two conditions are both satisfied:

  1. (a)

    the topology of XX is generated by a family of invariant seminorms, and

  2. (b)

    for each xXx\in X the function from 𝕋n\mathbb{T}^{n} to XX given by λσλ(x)\lambda\mapsto\sigma_{\lambda}(x) is continuous at the identity element of 𝕋n\mathbb{T}^{n}.

Proof.

Assume that σ\sigma is continuous, i.e., σ:𝕋n×XX\sigma:\mathbb{T}^{n}\times X\to X is continuous, so for xXx\in X, the function λσλ(x)\lambda\mapsto\sigma_{\lambda}(x) is continuous on 𝕋n\mathbb{T}^{n}, in particular at the identity. Therefore (b) follows.

For a continuous seminorm qq on XX, define p(x)=supλ𝕋nq(σλ(x))p(x)=\sup_{\lambda\in\mathbb{T}^{n}}q(\sigma_{\lambda}(x)), which is finite since 𝕋n\mathbb{T}^{n} is compact, and which is easily seen to be a seminorm. To show pp is invariant, we note that

p(σλ(x))=supμ𝕋nq(σμ(σλ(x)))=supμ𝕋nq(σμλ(x))=supξ𝕋nq(σξ(x))=p(x).p(\sigma_{\lambda}(x))=\sup\limits_{\mu\in\mathbb{T}^{n}}q(\sigma_{\mu}(\sigma_{\lambda}(x)))=\sup\limits_{\mu\in\mathbb{T}^{n}}q(\sigma_{\mu{\cdot}\lambda}(x))=\sup\limits_{\xi\in\mathbb{T}^{n}}q(\sigma_{\xi}(x))=p(x).

It remains to show that pp is continuous. For x,yXx,y\in X, we have

p(x)=supλ𝕋nq(σλ(y)+σλ(xy))supλ𝕋n(q(σλ(y))+q(σλ(xy)))p(y)+p(xy),p(x)=\sup_{\lambda\in\mathbb{T}^{n}}q\left(\sigma_{\lambda}(y)+\sigma_{\lambda}(x-y)\right)\leq\sup_{\lambda\in\mathbb{T}^{n}}\left(q(\sigma_{\lambda}(y))+q(\sigma_{\lambda}(x-y))\right)\leq p(y)+p(x-y),

so that |p(x)p(y)|p(xy)\left|p(x)-p(y)\right|\leq p(x-y) for all x,yXx,y\in X, and it follows that the seminorm pp is continuous on XX if and only if it is continuous at 0. We show that for each ϵ>0\epsilon>0, there exists a neighbourhood VV of 0 in XX such that for all λ𝕋n\lambda\in\mathbb{T}^{n}, qσλ<ϵq\circ\sigma_{\lambda}<\epsilon on VV. For each ξ𝕋n\xi\in\mathbb{T}^{n}, since qq and σ\sigma are continuous, there exists a neighborhood UξU_{\xi} of ξ\xi in 𝕋n\mathbb{T}^{n} and a neighbourhood VξV_{\xi} of 0 in XX such that q(σλ(x))<ϵq(\sigma_{\lambda}(x))<\epsilon for all xVξx\in V_{\xi} and λUξ\lambda\in U_{\xi}. The collection {Uξ}ξ𝕋n\{U_{\xi}\}_{\xi\in\mathbb{T}^{n}} forms an open cover of 𝕋n\mathbb{T}^{n}. Since 𝕋n\mathbb{T}^{n} is compact, let {Uξ1,,Uξk}\{U_{\xi_{1}},...,U_{\xi_{k}}\} be a finite subcover of 𝕋n\mathbb{T}^{n} corresponding to the open cover. Then for all xj=1kVξjx\in\bigcap_{j=1}^{k}V_{\xi_{j}} and λ𝕋n\lambda\in\mathbb{T}^{n}, we have q(σλ(x))<ϵq(\sigma_{\lambda}(x))<\epsilon.

Now assume the two conditions (a) and (b). Let (Γ,)(\Gamma,\geq) be a directed set and let (λα)αΓ(\lambda_{\alpha})_{\alpha\in\Gamma} and (xα)αΓ(x_{\alpha})_{\alpha\in\Gamma} be nets in 𝕋n\mathbb{T}^{n} and XX respectively with (λα,xα)(λ,x)(\lambda_{\alpha},x_{\alpha})\rightarrow(\lambda,x) in 𝕋n×X\mathbb{T}^{n}\times X. We need to show σλα(xα)σλ(x)\sigma_{\lambda_{\alpha}}(x_{\alpha})\rightarrow\sigma_{\lambda}(x) in XX, i.e., p(σλα(xα)σλ(x))0p(\sigma_{\lambda_{\alpha}}(x_{\alpha})-\sigma_{\lambda}(x))\to 0 for each invariant continuous seminorm pp of XX. But we have, by the invariance of pp:

p(σλα(xα)σλ(x))=p(xασλα1λ(x))p(xαx)+p(xσλα1λ(x)).p\left(\sigma_{\lambda_{\alpha}}(x_{\alpha}\right)-\sigma_{\lambda}(x))=p\left(x_{\alpha}-\sigma_{\lambda_{\alpha}^{-1}\lambda}(x)\right)\leq p\left(x_{\alpha}-x\right)+p\left(x-\sigma_{\lambda_{\alpha}^{-1}\lambda}(x)\right).

The term p(xαx)p\left(x_{\alpha}-x\right) goes to zero since xαxx_{\alpha}\to x, and the term p(xσλα1λ(x))p\left(x-\sigma_{\lambda_{\alpha}^{-1}\lambda}(x)\right) also goes to zero since λαλ\lambda_{\alpha}\to\lambda and μσμ(x)\mu\mapsto\sigma_{\mu}(x) is continuous at μ=𝟏\mu=\mathbf{1}. The result follows. ∎

2.3. Abstract Fejér Theorem

Let XX be a complete LCTVS and let σ\sigma be a continuous representation of 𝕋n\mathbb{T}^{n} on XX. For each α=(α1,,αn)n\alpha=(\alpha_{1},...,\alpha_{n})\in\mathbb{Z}^{n} and xXx\in X, define

𝝅ασ(x)=𝕋nλασλ(x)𝑑λ,\bm{\pi}_{\alpha}^{\sigma}(x)=\int_{\mathbb{T}^{n}}\lambda^{-\alpha}\sigma_{\lambda}(x)\,d\lambda, (2.5)

the Pettis integral of the continuous function λλασλ(x)\lambda\mapsto\lambda^{-\alpha}\sigma_{\lambda}(x) on 𝕋n\mathbb{T}^{n} with respect to the Haar probability measure of 𝕋n\mathbb{T}^{n}. We will say that 𝝅ασ(x)\bm{\pi}^{\sigma}_{\alpha}(x) is the α\alpha-th Fourier component of xx with respect to the representation σ\sigma. We use the standard convention with respect to multi-index powers, i.e., λα=λ1α1λnαn\lambda^{\alpha}=\lambda_{1}^{\alpha_{1}}\dots\lambda_{n}^{\alpha_{n}}. We will say that the subspace of XX defined as

[X]ασ={xX:σλ(x)=λαx for all λ𝕋n}[X]^{\sigma}_{\alpha}=\{x\in X:\sigma_{\lambda}(x)=\lambda^{\alpha}\cdot x\text{ for all }\lambda\in\mathbb{T}^{n}\} (2.6)

is the α\alpha-th Fourier mode of the space XX, and we will call the map 𝝅ασ\bm{\pi}^{\sigma}_{\alpha} the α\alpha-th Fourier projection, both with respect to the representation σ\sigma. We note the following facts:

Proposition 2.4.

As above, let XX be a complete LCTVS and σ\sigma be a continuous representation of 𝕋n\mathbb{T}^{n} on XX.

  1. (1)

    For each αn\alpha\in\mathbb{Z}^{n}, the α\alpha-th Fourier mode [X]ασ[X]^{\sigma}_{\alpha} is a closed σ\sigma-invariant subspace of XX, and the Fourier projection 𝝅ασ\bm{\pi}^{\sigma}_{\alpha} is a continuous linear projection from XX onto [X]ασ[X]^{\sigma}_{\alpha}.

  2. (2)

    Let YY be another complete LCTVS, and let τ\tau be a continuous representation of 𝕋n\mathbb{T}^{n} on YY, and let j:YXj:Y\to X be a continuous linear map intertwining σ\sigma and τ\tau, i.e., for each λ𝕋n\lambda\in\mathbb{T}^{n}, jτλ=σλj.j\circ\tau_{\lambda}=\sigma_{\lambda}\circ j. Then for each αn\alpha\in\mathbb{Z}^{n}, we have

    j𝝅ατ=𝝅ασj.j\circ\bm{\pi}^{\tau}_{\alpha}=\bm{\pi}^{\sigma}_{\alpha}\circ j. (2.7)
Proof.
  1. (1)

    Linearity of 𝝅ασ\bm{\pi}_{\alpha}^{\sigma} follows from the linearity of σλ\sigma_{\lambda}. Recall from Proposition 2.3 that there exists a family 𝒫\mathscr{P} of continuous invariant seminorms that generates the locally convex topology of XX. To see the continuity of 𝝅ασ\bm{\pi}_{\alpha}^{\sigma}, observe that for each p𝒫p\in\mathscr{P}, we have

    p(𝝅ασ(x))=p(𝕋nλασλ(x)𝑑λ)𝕋np(λασλ(x))𝑑λ=𝕋np(x)𝑑λ=p(x),p(\bm{\pi}_{\alpha}^{\sigma}(x))=p\left(\int_{\mathbb{T}^{n}}\lambda^{-\alpha}\sigma_{\lambda}(x)\,d\lambda\right)\leq\int_{\mathbb{T}^{n}}p\left(\lambda^{-\alpha}\sigma_{\lambda}(x)\right)\,d\lambda=\int_{\mathbb{T}^{n}}p(x)\,d\lambda=p(x), (2.8)

    where the inequality is due to Proposition 6 in [Bou04, p. INT III.37]. Now, if x[X]ασx\in[X]^{\sigma}_{\alpha}, then

    𝝅ασ(x)=𝕋nλασλ(x)𝑑λ=𝕋nλαλαx𝑑λ=x(𝕋n𝑑λ)=x,\bm{\pi}_{\alpha}^{\sigma}(x)=\int_{\mathbb{T}^{n}}\lambda^{-\alpha}\sigma_{\lambda}(x)\hskip 2.84526ptd\lambda=\int_{\mathbb{T}^{n}}\lambda^{-\alpha}\cdot\lambda^{\alpha}x\hskip 2.84526ptd\lambda=x\cdot\bigg{(}\int_{\mathbb{T}^{n}}\hskip 2.84526ptd\lambda\bigg{)}=x,

    so that [X]ασ𝝅ασ(X)[X]^{\sigma}_{\alpha}\subset\bm{\pi}_{\alpha}^{\sigma}(X). To prove that 𝝅ασ(X)[X]ασ\bm{\pi}_{\alpha}^{\sigma}(X)\subset[X]^{\sigma}_{\alpha}, notice that for each αn\alpha\in\mathbb{Z}^{n}, each λ𝕋n\lambda\in\mathbb{T}^{n} and xXx\in X,

    σλ(𝝅ασ(x))\displaystyle\sigma_{\lambda}(\bm{\pi}^{\sigma}_{\alpha}(x)) =σλ(𝕋nμασμ(x)𝑑μ)=𝕋nμασλμ(x)𝑑μ\displaystyle=\sigma_{\lambda}\left(\int_{\mathbb{T}^{n}}\mu^{-\alpha}\sigma_{\mu}(x)d\mu\right)=\int_{\mathbb{T}^{n}}\mu^{-\alpha}\sigma_{\lambda\cdot\mu}(x)d\mu using (2.4)
    =λα𝕋n(λμ)ασλμ(x)𝑑μ=λα𝝅ασ(x)\displaystyle=\lambda^{\alpha}\cdot\int_{\mathbb{T}^{n}}(\lambda\mu)^{-\alpha}\sigma_{\lambda\mu}(x)d\mu=\lambda^{\alpha}\cdot\bm{\pi}^{\sigma}_{\alpha}(x) by Haar invariance.\displaystyle\text{by Haar invariance}.
  2. (2)

    For xXx\in X, we have

    j𝝅ατ(x)=j(𝕋nλατλ(x)𝑑λ)=𝕋nλαjτλ(x)𝑑λ=𝕋nλασλj(x)𝑑λ=𝝅ασj(x),j\circ\bm{\pi}^{\tau}_{\alpha}(x)=j\left(\int_{\mathbb{T}^{n}}\lambda^{-\alpha}\tau_{\lambda}(x)d\lambda\right)=\int_{\mathbb{T}^{n}}\lambda^{-\alpha}j\circ\tau_{\lambda}(x)d\lambda=\int_{\mathbb{T}^{n}}\lambda^{-\alpha}\sigma_{\lambda}\circ j(x)d\lambda=\bm{\pi}^{\sigma}_{\alpha}\circ j(x),

    where we have used (2.4) to go from the second to the third step.

Remark.

The inequality (2.8)

p(𝝅ασ(x))p(x),p\left(\bm{\pi}_{\alpha}^{\sigma}(x)\right)\leq p(x), (2.9)

which holds for each α\alpha for an invariant seminorm pp can be thought of as an abstract form of the familiar Cauchy inequalities of complex analysis.

If XX is L1(𝕋)L^{1}(\mathbb{T}), the Banach space of integrable functions on 𝕋\mathbb{T}, and if σ\sigma is the continuous representation of 𝕋\mathbb{T} on L1(𝕋)L^{1}(\mathbb{T}) given by

σλ(f)(μ)=f(λμ),λ,μ𝕋,fL1(𝕋),\sigma_{\lambda}(f)(\mu)=f(\lambda\cdot\mu),\quad\lambda,\mu\in\mathbb{T},f\in L^{1}(\mathbb{T}),

an easy computation shows that for ϕ\phi\in{\mathbb{R}}, and fL1(𝕋)f\in L^{1}(\mathbb{T}), we have

𝝅ασ(f)(eiϕ)=eiαϕf^(α),with f^(α)=12π02πeiαθf(eiθ)𝑑θ,\bm{\pi}_{\alpha}^{\sigma}(f)(e^{i\phi})=e^{i\alpha\phi}\cdot\hat{f}(\alpha),\quad\text{with }\hat{f}(\alpha)=\frac{1}{2\pi}\cdot\int_{0}^{2\pi}e^{-i\alpha\theta}f(e^{i\theta})d\theta,

the α\alpha-th term of the Fourier series of ff.

It is therefore natural to define, for xXx\in X, the Fourier series of xx with respect to σ\sigma to be the formal series

xαn𝝅ασ(x).x\sim\sum\limits_{\alpha\in\mathbb{Z}^{n}}\bm{\pi}_{\alpha}^{\sigma}(x). (2.10)

For an integer k0k\geq 0, define the kk-th square partial sum of the Fourier series in (2.10) by

Skσ(x)=|α|k𝝅ασ(x),S_{k}^{\sigma}(x)=\sum\limits_{\left|\alpha\right|_{\infty}\leq k}\bm{\pi}_{\alpha}^{\sigma}(x), (2.11)

where |α|max{|αj|,1jn}\left|\alpha\right|_{\infty}\coloneqq\max\big{\{}\left|\alpha_{j}\right|,1\leq j\leq n\big{\}}. We are ready to state an abstract version of Fejér’s theorem.

Theorem 2.1.

Let σ\sigma be a continuous representation of 𝕋n\mathbb{T}^{n} on an LCTVS XX and let xXx\in X. Then the Cesàro means of the square partial sums of the Fourier series of xx (with respect to σ\sigma) converge to xx in the topology of XX.

Proof.

Write the Cesàro means of the square partial sums of the Fourier series of xx as,

CNσ(x)\displaystyle C_{N}^{\sigma}(x) =1N+1k=0NSkσ(x)=1N+1k=0N|α|k𝝅ασ(x)=1N+1k=0N|α|k(𝕋nλασλ(x)𝑑λ)\displaystyle=\frac{1}{N+1}\sum\limits_{k=0}^{N}S_{k}^{\sigma}(x)=\frac{1}{N+1}\sum\limits_{k=0}^{N}\sum\limits_{\left|\alpha\right|_{\infty}\leq k}\bm{\pi}_{\alpha}^{\sigma}(x)=\frac{1}{N+1}\sum\limits_{k=0}^{N}\sum\limits_{\left|\alpha\right|_{\infty}\leq k}\left(\int\limits_{\mathbb{T}^{n}}\lambda^{-\alpha}\sigma_{\lambda}(x)\,d\lambda\right)
=1N+1𝕋nk=0N(|α|kλα)σλ(x)dλ=𝕋nFN(λ)σλ(x)𝑑λ\displaystyle=\frac{1}{N+1}\int\limits_{\mathbb{T}^{n}}\sum\limits_{k=0}^{N}\left(\sum\limits_{\left|\alpha\right|_{\infty}\leq k}\lambda^{-\alpha}\right)\sigma_{\lambda}(x)\,d\lambda=\int\limits_{\mathbb{T}^{n}}F_{N}(\lambda)\hskip 2.84526pt\sigma_{\lambda}(x)\,d\lambda

where

FN(λ)=1N+1k0N(|α|kλα)F_{N}(\lambda)=\displaystyle{\frac{1}{N+1}\sum\limits_{k-0}^{N}\left(\sum\limits_{\left|\alpha\right|_{\infty}\leq k}\lambda^{-\alpha}\right)}

is the classical Fejér kernel. Introducing polar coordinates, λj=eiθj\lambda_{j}=e^{i\theta_{j}} on 𝕋n\mathbb{T}^{n}, and summing, we obtain the classical representation

FN(eiθ1,,eiθn)=1N+1j=1nsin2((N+12)θj)sin2(θj2)F_{N}(e^{i\theta_{1}},\cdots,e^{i\theta_{n}})=\frac{1}{N+1}\prod_{j=1}^{n}\frac{\sin^{2}\left(\left(\frac{N+1}{2}\right)\theta_{j}\right)}{\sin^{2}\left(\frac{\theta_{j}}{2}\right)}

It is well-known that the Fejér kernel has the properties that

  • (a)

    FN0F_{N}\geq 0 for all NN,

  • (b)

    𝕋nFN(λ)𝑑λ=1\int\limits_{\mathbb{T}^{n}}F_{N}(\lambda)\,d\lambda=1 and

  • (c)

    For each δ>0\delta>0, FN0F_{N}\rightarrow 0 uniformly on 𝕋nB(𝟏,δ)\mathbb{T}^{n}\setminus B(\mathbf{1},\delta), where B(𝟏,δ)B(\mathbf{1},\delta) is the nn-dimensional ball centered at 𝟏=(1,1,,1)\mathbf{1}=(1,1,...,1) and radius δ\delta.

Let pp be a continuous σ\sigma-invariant seminorm on XX. Then for xXx\in X, we have

p(CN(x)x)\displaystyle p(C_{N}(x)-x) =p(𝕋nFN(λ)σλ(x)𝑑λx𝕋nFN(λ)𝑑λ)\displaystyle=p\left(\int\limits_{\mathbb{T}^{n}}F_{N}(\lambda)\cdot\sigma_{\lambda}(x)\,d\lambda-x\cdot\int\limits_{\mathbb{T}^{n}}F_{N}(\lambda)\,d\lambda\right)
using property (b) of FNF_{N}
=p(𝕋nFN(λ)σλ(x)𝑑λ𝕋nxFN(λ)𝑑λ)\displaystyle=p\left(\int\limits_{\mathbb{T}^{n}}F_{N}(\lambda)\cdot\sigma_{\lambda}(x)\,d\lambda-\int\limits_{\mathbb{T}^{n}}x\cdot F_{N}(\lambda)\,d\lambda\right)
using (2.4)
=p(𝕋n(σλ(x)x)FN(λ)𝑑λ)\displaystyle=p\left(\int\limits_{\mathbb{T}^{n}}\big{(}\sigma_{\lambda}(x)-x\big{)}\cdot F_{N}(\lambda)\,d\lambda\right)
𝕋np(σλ(x)x)FN(λ)𝑑λ\displaystyle\leq\int\limits_{\mathbb{T}^{n}}p(\sigma_{\lambda}(x)-x)\cdot F_{N}(\lambda)\,d\lambda (2.12)
using [Bou04, Prop. 6, p. INT III.37] and the positivity of FN(λ)F_{N}(\lambda)

Since λσλ(x)\lambda\mapsto\sigma_{\lambda}(x) is continuous and pp is a continuous seminorm, there exists δ>0\delta>0 such that p(σλ(x)σ𝟏(x))=p(σλ(x)x)<ϵ2p(\sigma_{\lambda}(x)-\sigma_{\mathbf{1}}(x))=p(\sigma_{\lambda}(x)-x)<\frac{\epsilon}{2} whenever λ𝕋nB(𝟏,δ).\lambda\in\mathbb{T}^{n}\cap B(\mathbf{1},\delta). Then on the set 𝕋nB(𝟏,δ)\mathbb{T}^{n}\setminus B(\mathbf{1},\delta) we have

p(σλ(x)x)p(σλ(x))+p(x)=2p(x).p(\sigma_{\lambda}(x)-x)\leq p(\sigma_{\lambda}(x))+p(x)=2\cdot p(x).

By property (c) of FNF_{N}, there exists N1N_{1}\in\mathbb{N} such that whenever λ𝕋nB(𝟏,δ)\lambda\in\mathbb{T}^{n}\setminus B(\mathbf{1},\delta), for all NN1N\geq N_{1} we have,

FN(λ)<ϵ4p(x)F_{N}(\lambda)<\frac{\epsilon}{4\cdot p(x)} (2.13)

Now from (2.12), we have

p(CN(x)x)\displaystyle p(C_{N}(x)-x) 𝕋nB(𝟏,δ)FN(λ)p(σλ(x)x)𝑑λ+𝕋nB(𝟏,δ)FN(λ)p(σλ(x)x)𝑑λ\displaystyle\leq\int\limits_{\mathbb{T}^{n}\cap B(\mathbf{1},\delta)}F_{N}(\lambda)\cdot p(\sigma_{\lambda}(x)-x)\,d\lambda+\int\limits_{\mathbb{T}^{n}\setminus B(\mathbf{1},\delta)}F_{N}(\lambda)\cdot p(\sigma_{\lambda}(x)-x)\,d\lambda
<ϵ2𝕋nFn(λ)𝑑λ+ϵ4p(x)2p(x)\displaystyle<\frac{\epsilon}{2}\int\limits_{\mathbb{T}^{n}}F_{n}(\lambda)\,d\lambda+\frac{\epsilon}{4\cdot p(x)}\cdot 2\cdot p(x)
=ϵ2+ϵ2=ϵ.\displaystyle=\frac{\epsilon}{2}+\frac{\epsilon}{2}=\epsilon.

Since by Proposition 2.3 the topology of XX is generated by σ\sigma-invariant seminorms, the result follows. ∎

Corollary 2.5.

Let XX be a complete LCTVS and suppose that we are given a continuous representation of the group 𝕋n\mathbb{T}^{n} on XX. Then if the Fourier series of an element xXx\in X with respect to this representation is absolutely convergent in XX, the sum of the Fourier series equals xx.

Proof.

Since the series αnπασ(x)\sum_{\alpha\in\mathbb{Z}^{n}}\pi_{\alpha}^{\sigma}(x) is absolutely convergent, by Proposition 2.2, there exists the sum of the series, i.e., an x~X\widetilde{x}\in X such that for every bijection θ:n\theta:\mathbb{N}\rightarrow\mathbb{Z}^{n} we have j=0πθ(j)σ(x)=x~\sum_{j=0}^{\infty}\pi_{\theta(j)}^{\sigma}(x)=\widetilde{x}. Let SN=j=0Nπθ(j)σ(x)S_{N}=\sum_{j=0}^{N}\pi_{\theta(j)}^{\sigma}(x); then the sequence of partial sums {SN}\{S_{N}\} converges to x~\widetilde{x} in XX. By Proposition 2.1, the sequence of Cesàro means {CN}\{C_{N}\} of the partial sums converges to x~\widetilde{x} as well. However, by Theorem 2.1, the Cesàro means converge to xx. Therefore x~=x\widetilde{x}=x. ∎

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 Ωn\Omega\subset{\mathbb{R}}^{n}, the space 𝒟(Ω)\mathscr{D}(\Omega) of test functions is the LFLF-space of smooth compactly supported complex valued functions, topologized as the inductive limit of the Fréchet spaces 𝒟K\mathscr{D}_{K} consisting, for a given compact KΩK\subset\Omega, of those elements of 𝒟(Ω)\mathscr{D}(\Omega) which have support in KK. Recall that a subset B𝒟(Ω)B\subset\mathscr{D}(\Omega) is bounded, if and only if there is a compact KΩK\subset\Omega such that B𝒟KB\subset\mathscr{D}_{K}, and for each nonnegative multi-index αn\alpha\in\mathbb{N}^{n}, we have

supϕBxK|αxαϕ(x)|<,\sup_{\begin{subarray}{c}\phi\in B\\ x\in K\end{subarray}}\left|\frac{\partial^{\alpha}}{\partial x^{\alpha}}\phi(x)\right|<\infty, (3.1)

where, here and later, we will use standard multi-index conventions such as

αxα=|α|x1α1xnαn,|α|=jαj,forαn.\frac{\partial^{\alpha}}{\partial x^{\alpha}}=\frac{\partial^{\left|\alpha\right|}}{\partial x_{1}^{\alpha_{1}}\dots\partial x_{n}^{\alpha_{n}}},\quad\left|\alpha\right|=\sum_{j}\alpha_{j},\quad\text{for}\quad\alpha\in\mathbb{N}^{n}. (3.2)

The space of distributions 𝒟(Ω)\mathscr{D}^{\prime}(\Omega) on Ω\Omega is the dual of 𝒟(Ω)\mathscr{D}(\Omega), consisting of continuous linear forms on 𝒟(Ω)\mathscr{D}(\Omega). We denote the value of a distribution T𝒟(Ω)T\in\mathscr{D}^{\prime}(\Omega) at a test function ϕ𝒟(Ω)\phi\in\mathscr{D}(\Omega) by T,ϕ.\left\langle T,\phi\right\rangle. The space 𝒟(Ω)\mathscr{D}^{\prime}(\Omega) is endowed with the usual strong dual topology. Recall that this topology is generated by the family of seminorms

{pB:B𝒟(Ω) is bounded},where, pB(T)=supϕB|T,ϕ|.\left\{p_{B}:B\subset\mathscr{D}(\Omega)\text{ is bounded}\right\},\quad\text{where, }p_{B}(T)=\sup_{\phi\in B}\left|\left\langle T,\phi\right\rangle\right|. (3.3)

In this topology, the space 𝒟(Ω)\mathscr{D}^{\prime}(\Omega) is complete.

Given a locally integrable function fLloc1(Ω)f\in L^{1}_{\mathrm{loc}}(\Omega), we can associate to ff a distribution TfT_{f} defined by

Tf,ϕ=Ωfϕ𝑑V, for ϕ𝒟(Ω),\left\langle T_{f},\phi\right\rangle=\int_{\Omega}f\phi\,dV,\textrm{ for }\phi\in\mathscr{D}(\Omega),

where dVdV denotes the Lebesgue measure of n{\mathbb{R}}^{n}. Then the locally integrable distribution TfT_{f} is said to be generated by ff, and as usual we grant ourselves the the right to abuse language by identifying the distribution Tf𝒟(Ω)T_{f}\in\mathscr{D}^{\prime}(\Omega) with the function fLloc1(Ω)f\in L^{1}_{\mathrm{loc}}(\Omega).

We will use the abbreviations

𝑫j=zj,𝑫¯j=zj¯,\bm{D}_{j}=\frac{\partial}{\partial z_{j}},\quad\bm{\overline{D}}_{j}=\frac{\partial}{\partial\overline{z_{j}}}, (3.4)

for the basic constant coefficient differential operators of complex analysis, acting on functions or distributions on n{\mathbb{C}}^{n}. If Ωn\Omega\subset{\mathbb{C}}^{n} is an open set, define

𝒪(Ω)={f𝒟(Ω):𝑫¯jf=0,1jn},\mathscr{O}(\Omega)=\left\{f\in\mathscr{D}^{\prime}(\Omega):\bm{\overline{D}}_{j}f=0,1\leq j\leq n\right\}, (3.5)

the space of holomorphic distributions on Ω\Omega. The subspace 𝒪(Ω)\mathscr{O}(\Omega) is closed in the space 𝒟(Ω)\mathscr{D}^{\prime}(\Omega) by the continuity of the operators 𝑫¯j\bm{\overline{D}}_{j}, and is therefore a complete LCTVS in the subspace topology.

3.2. The main theorem

We begin with some definitions and notational conventions. For λ𝕋n\lambda\in\mathbb{T}^{n} we denote by Rλ\mathrm{R}_{\lambda} the Reinhardt rotation of n{\mathbb{C}}^{n} by the element λ\lambda, the linear automorphism of the vector space n{\mathbb{C}}^{n} given by

Rλ(z)=(λ1z1,,λnzn).\mathrm{R}_{\lambda}(z)=(\lambda_{1}z_{1},\dots,\lambda_{n}z_{n}). (3.6)

A domain Ωn\Omega\subset{\mathbb{C}}^{n} is defined to be Reinhardt if and only if Rλ(Ω)=Ω\mathrm{R}_{\lambda}(\Omega)=\Omega for each λ𝕋n\lambda\in\mathbb{T}^{n}. Let ZZ denote the union of the coordinate hyperplanes of n{\mathbb{C}}^{n}:

Z=j=1n{zn:zj=0}.Z=\bigcup_{j=1}^{n}\{z\in{\mathbb{C}}^{n}:z_{j}=0\}. (3.7)

Recall that a Reinhardt domain Ωn\Omega\subset{\mathbb{C}}^{n} is said to be log-convex, if whenever z,wΩZz,w\in\Omega\setminus Z, the point ζn\zeta\in{\mathbb{C}}^{n} belongs to Ω\Omega if there is a t[0,1]t\in[0,1] such that for 1jn1\leq j\leq n

|ζj|=|zj|t|wj|1t.\left|\zeta_{j}\right|=\left|z_{j}\right|^{t}\left|w_{j}\right|^{1-t}. (3.8)

For αn\alpha\in\mathbb{Z}^{n}, let eαe_{\alpha} be the monomial function given by

eα(z)=zα=z1α1znαn.e_{\alpha}(z)=z^{\alpha}=z_{1}^{\alpha_{1}}\dots z_{n}^{\alpha_{n}}. (3.9)

For a Reinhardt subset Ωn\Omega\subset{\mathbb{C}}^{n} and 1jn1\leq j\leq n, let

Ω(j)={(z1,,ζzj,,zn):zΩ,ζ𝔻¯},{\Omega}^{(j)}=\{(z_{1},\dots,\zeta z_{j},\dots,z_{n}):z\in\Omega,\zeta\in\overline{\mathbb{D}}\}, (3.10)

where 𝔻¯={|ζ|1}\overline{\mathbb{D}}=\{\left|\zeta\right|\leq 1\}\subset{\mathbb{C}} is the closed disk. This can be thought of as the result of “completing” Ω\Omega in the jj-th coordinate. Following [JP08], we say that the Reinhardt domain Ωn\Omega\subset{\mathbb{C}}^{n} is relatively complete if for each 1jn1\leq j\leq n, whenever we have Ω{zj=0}\Omega\cap\{z_{j}=0\}\not=\emptyset, we also have

Ω(j)Ω.\Omega^{(j)}\subset\Omega.

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 Ω\Omega be a Reinhardt domain in n{\mathbb{C}}^{n} and let

𝒮(Ω)={αn:eα𝒞(Ω)}.\mathcal{S}(\Omega)=\{\alpha\in\mathbb{Z}^{n}:e_{\alpha}\in\mathcal{C}^{\infty}(\Omega)\}. (3.11)

Let Ω^\widehat{\Omega} be the smallest relatively complete log-convex Reinhardt domain in n{\mathbb{C}}^{n} that contains Ω\Omega. The for each α𝒮(Ω)\alpha\in\mathcal{S}(\Omega) there is a continuous linear functional aα:𝒪(Ω)a_{\alpha}:\mathscr{O}(\Omega)\to{\mathbb{C}} such that for each T𝒪(Ω)T\in\mathscr{O}(\Omega), the series

α𝒮(Ω)aα(T)eα\sum_{\alpha\in\mathcal{S}(\Omega)}a_{\alpha}(T)e_{\alpha} (3.12)

converges absolutely in 𝒞(Ω^)\mathcal{C}^{\infty}(\widehat{\Omega}) to a function f𝒞(Ω^)f\in\mathcal{C}^{\infty}(\widehat{\Omega}), and f|Ωf|_{\Omega} generates the distribution TT.

Remarks:

  1. (1)

    For n=1n=1, all Reinhardt domains in the plane, i.e., disks and annuli, are automatically relatively complete and log-convex. For n2n\geq 2, 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 Ω\Omega, it follows that each holomorphic distribution f𝒪(Ω)f\in\mathscr{O}(\Omega) extends to a holomorphic function F𝒪(Ω^)F\in\mathscr{O}(\widehat{\Omega}). 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. (2)

    The functionals aαa_{\alpha} are called the coefficient functionals, and the series (3.12) is of course the Laurent series of the function ff (the Taylor series if 𝒮(Ω)=n\mathcal{S}(\Omega)=\mathbb{N}^{n}).

  3. (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 Ω\Omega admits a holomorphic distribution whose Laurent expansion converges absolutely precisely on Ω\Omega.

As an immediate consequence of Theorem 3.1 we have the following:

Corollary 3.1.

Let Ωn\Omega\subset{\mathbb{C}}^{n} be open. Then each distribution T𝒪(Ω)T\in\mathscr{O}(\Omega) is complex-analytic, i.e., for each pΩp\in\Omega there is a neighborhood UU of pp, where the function ff generating TT is represented by a Taylor series centered at pp.

3.3. Holomorphic functions and maps

A holomorphic distribution T𝒪(Ω)T\in\mathscr{O}(\Omega) will be called a holomorphic function if it is generated by a 𝒞\mathcal{C}^{\infty} function ff. We denote the space of holomorphic functions on Ω\Omega temporarily by (𝒪𝒞)(Ω)(\mathscr{O}\cap\mathcal{C}^{\infty})(\Omega). Once Theorem 3.1 is proved, it will follow that (𝒪𝒞)(Ω)=𝒪(Ω)(\mathscr{O}\cap\mathcal{C}^{\infty})(\Omega)=\mathscr{O}(\Omega).

Let Ω1,Ω2\Omega_{1},\Omega_{2} be domains in n{\mathbb{C}}^{n}. By a holomorphic map Φ:Ω1Ω2\Phi:\Omega_{1}\to\Omega_{2}, we mean a mapping each of whose components is a holomorphic function on Ω1\Omega_{1}. 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 Φ:Ω1Ω2\Phi:\Omega_{1}\to\Omega_{2} is a biholomorphism, then for a distribution T𝒟(Ω2)T\in\mathscr{D}^{\prime}(\Omega_{2}), we can define in the usual way the pullback distribution ΦT𝒟(Ω1)\Phi^{*}T\in\mathscr{D}^{\prime}(\Omega_{1}): if TT is generated by a test function f𝒟(Ω2)f\in\mathscr{D}(\Omega_{2}), then ΦT\Phi^{*}T is the distribution generated by the function fΦf\circ\Phi, and for general TT, we extend this definition by continuity, using the density of test functions in 𝒟(Ω)\mathscr{D}^{\prime}(\Omega), 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):

𝑫j(ΦT)=k=1n𝑫jΦkΦ(𝑫kT),\bm{D}_{j}(\Phi^{*}T)=\sum_{k=1}^{n}\bm{D}_{j}{\Phi_{k}}\cdot\Phi^{*}\left(\bm{D}_{k}T\right), (3.13)

and

𝑫¯j(ΦT)=k=1n𝑫¯jΦk¯Φ(𝑫¯kT),\bm{\overline{D}}_{j}(\Phi^{*}T)=\sum_{k=1}^{n}\bm{\overline{D}}_{j}\overline{\Phi_{k}}\cdot\Phi^{*}\left(\bm{\overline{D}}_{k}T\right), (3.14)

where as above, Φ:Ω1Ω2\Phi:\Omega_{1}\to\Omega_{2} is a biholomorphism of domains in n{\mathbb{C}}^{n}, written in components as Φ=(Φ1,,Φn)\Phi=(\Phi_{1},\dots,\Phi_{n}), and T𝒟(Ω2)T\in\mathscr{D}^{\prime}(\Omega_{2}).

Therefore, we have the following immediate consequence of (3.14):

Proposition 3.2.

If Φ:Ω1Ω2\Phi:\Omega_{1}\to\Omega_{2} is a biholomorphism and T𝒪(Ω2)T\in\mathscr{O}(\Omega_{2}), then we have ΦT𝒪(Ω1)\Phi^{*}T\in\mathscr{O}(\Omega_{1}). If f(𝒪𝒞)(Ω2)f\in(\mathscr{O}\cap\mathcal{C}^{\infty})(\Omega_{2}), then Φf(𝒪𝒞)(Ω1)\Phi^{*}f\in(\mathscr{O}\cap\mathcal{C}^{\infty})(\Omega_{1}).

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. (1)

    Translation by a vector ana\in{\mathbb{C}}^{n} is the map Ma:nn\mathrm{M}_{a}:{\mathbb{C}}^{n}\to{\mathbb{C}}^{n}

    Ma(z)=z+a,\mathrm{M}_{a}(z)=z+a, (3.15)

    which is obviously a biholomorphic automorphism of n{\mathbb{C}}^{n}. For a domain Ωn\Omega\subset{\mathbb{C}}^{n}, we therefore have a pullback isomorphism of spaces of holomorphic distributions Ma:𝒪(Ma(Ω))𝒪(Ω)\mathrm{M}_{a}^{*}:\mathscr{O}(\mathrm{M}_{a}(\Omega))\to\mathscr{O}(\Omega). This can be thought of as an expression of the fact that the operator ¯=(𝑫¯1,,𝑫¯n)\overline{\partial}=(\bm{\overline{D}}_{1},\dots,\bm{\overline{D}}_{n}) is translation invariant.

  2. (2)

    The Reinhardt rotations Rλ\mathrm{R}_{\lambda} of (3.6) are clearly biholomorphic automorphisms of Reinhardt domains, and the pullback operation induces a representation of the group 𝕋n\mathbb{T}^{n} on the space of holomorphic distributions (see (3.32) below).

    A domain Ω\Omega is said to be Reinhardt centered at aa for an ana\in{\mathbb{C}}^{n} if there is a Reinhardt domain Ω0\Omega_{0} such that Ω=Ma(Ω0)\Omega=\mathrm{M}_{a}(\Omega_{0}). Every open set in n{\mathbb{C}}^{n} 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 α\alpha, we have eα(𝒪𝒞)(nZ)e_{\alpha}\in(\mathscr{O}\cap\mathcal{C}^{\infty})({\mathbb{C}}^{n}\setminus Z), where eαe_{\alpha} is the monomial of (3.9). We now note a remarkable symmetry property (the Mean-Value Property) of the functions eαe_{\alpha}:

Lemma 3.3.

Let αn\alpha\in\mathbb{Z}^{n}, let znZz\in{\mathbb{C}}^{n}\setminus Z, and let ψ𝒟(nZ)\psi\in\mathscr{D}({\mathbb{C}}^{n}\setminus Z) be a test function which has radial symmetry in each variable around the point zz, i.e., there are functions ρ1,,ρn𝒟()\rho_{1},\dots,\rho_{n}\in\mathscr{D}({\mathbb{R}}) such that ψ(ζ)=j=1nρj(|ζjzj|)\psi(\zeta)=\prod_{j=1}^{n}\rho_{j}(\left|\zeta_{j}-z_{j}\right|), and whose integral is 1:

nZψ(ζ)𝑑V(ζ)=1.\int_{{\mathbb{C}}^{n}\setminus Z}\psi(\zeta)dV(\zeta)=1. (3.16)

Then we have

eα,ψ=eα(z)\left\langle e_{\alpha},\psi\right\rangle=e_{\alpha}(z) (3.17)
Proof.

First consider the case n=1n=1. If α0\alpha\geq 0, the formula

02π(z+reiθ)α𝑑θ=2πzα\int_{0}^{2\pi}(z+re^{i\theta})^{\alpha}d\theta=2\pi z^{\alpha} (3.18)

holds for r>0r>0, by expanding the integrand using the binomial formula, and integrating the finite sum term by term. The formula (3.18) also holds for α<0\alpha<0, provided z0z\not=0, and 0<r<|z|0<r<\left|z\right|. This follows on noticing that we have an infinite series expansion

(z+reiθ)α=zα(1+rzeiθ)α=zαk=0(αk)(rz)keikθ,(z+re^{i\theta})^{\alpha}=z^{\alpha}\left(1+\frac{r}{z}e^{i\theta}\right)^{\alpha}=z^{\alpha}\sum_{k=0}^{\infty}\binom{\alpha}{k}\left(\frac{r}{z}\right)^{k}e^{ik\theta},

where by the MM-test, the convergence is uniform in θ\theta, and then integrating the series on the right term by term.

Now ψ(ζ)=ρ(|ζz|)\psi(\zeta)=\rho(\left|\zeta-z\right|) and ψ\psi is supported in some disc B(z,R)={0}B(z,R)\subset{\mathbb{C}}^{*}={\mathbb{C}}\setminus\{0\}, which means R<|z|R<\left|z\right|. Then the normalization (3.16) is equivalent to

2π0Rρ(r)r𝑑r=1.2\pi\int_{0}^{R}\rho(r)rdr=1. (3.19)

Now

eα,ψ\displaystyle\left\langle e_{\alpha},\psi\right\rangle =eα(ζ)ρ(|ζz|)𝑑V(ζ)=0R02π(z+reiθ)αρ(r)r𝑑θ𝑑r\displaystyle=\int_{{\mathbb{C}}^{*}}e_{\alpha}(\zeta)\rho(\left|\zeta-z\right|)dV(\zeta)=\int\limits_{0}^{R}\int\limits_{0}^{2\pi}(z+re^{i\theta})^{\alpha}\rho(r)rd\theta dr
=0Rρ(r)r(02π(z+reiθ)α𝑑θ)𝑑r=2πzα0Rρ(r)r𝑑r\displaystyle=\int_{0}^{R}\rho(r)r\left(\int\limits_{0}^{2\pi}(z+re^{i\theta})^{\alpha}d\theta\right)dr=2\pi z^{\alpha}\cdot\int_{0}^{R}\rho(r)rdr using (3.18)
=zα=eα(z),\displaystyle=z^{\alpha}=e_{\alpha}(z), using (3.19).\displaystyle\text{using \eqref{eq-mvp2}}.

which establishes the result for n=1n=1. In the general case, notice that for αn\alpha\in\mathbb{Z}^{n} and ζnZ\zeta\in{\mathbb{C}}^{n}\setminus Z we have eα(ζ)=j=1neαj(ζj).e_{\alpha}(\zeta)=\prod_{j=1}^{n}e_{\alpha_{j}}(\zeta_{j}). Therefore, since nZ=××{\mathbb{C}}^{n}\setminus Z={\mathbb{C}}^{*}\times\dots\times{\mathbb{C}}^{*},

eα,ψ\displaystyle\left\langle e_{\alpha},\psi\right\rangle =nZj=1neαj(ζj)ρj(|ζjzj|)dV(ζ)=j=1neαj(ζj)ρj(|ζjzj|)𝑑V(ζj)\displaystyle=\int_{{\mathbb{C}}^{n}\setminus Z}\prod_{j=1}^{n}e_{\alpha_{j}}(\zeta_{j})\rho_{j}(\left|\zeta_{j}-z_{j}\right|)\,dV(\zeta)=\prod_{j=1}^{n}\int_{{\mathbb{C}}^{*}}e_{\alpha_{j}}(\zeta_{j})\rho_{j}(\left|\zeta_{j}-z_{j}\right|)\,dV(\zeta_{j})
=j=1neαj(zj)=eα(z).\displaystyle=\prod_{j=1}^{n}e_{\alpha_{j}}(z_{j})=e_{\alpha}(z).

3.5. The representation τ\tau

If Ωn\Omega\subset{\mathbb{C}}^{n} is a Reinhardt domain, then for each λ\lambda, the map Rλ\mathrm{R}_{\lambda} of (3.6) is a biholomorphic automorphism of Ω\Omega. Define a representation τ\tau of 𝕋n\mathbb{T}^{n} on the space 𝒟(Ω)\mathscr{D}(\Omega) of test functions by

τλψ=ψRλ,ψ𝒟(Ω).\tau_{\lambda}\psi=\psi\circ\mathrm{R}_{\lambda},\quad\psi\in\mathscr{D}(\Omega). (3.20)

Recall that a net {ϕj}\{\phi_{j}\} converges in the space 𝒟(Ω)\mathscr{D}(\Omega), if each ϕj\phi_{j} is supported in a fixed compact KΩK\subset\Omega, and the net {ϕj}\{\phi_{j}\} converges in the Fréchet space 𝒟K,\mathscr{D}_{K}, i.e. all partial derivatives converge uniformly on KK. Using this it is easily verified that τ\tau is a continuous representation of 𝕋n\mathbb{T}^{n} in the space 𝒟(Ω)\mathscr{D}(\Omega).

Notice that nZ{\mathbb{C}}^{n}\setminus Z is a Reinhardt domain. For a positive integer kk we define the norm ψk\left\|\psi\right\|_{k} with respect to the polar coordinates for a function ψ\psi in 𝒟(nZ)\mathscr{D}({\mathbb{C}}^{n}\setminus Z) :

ψk=maxβ,γn|β|+|γ|ksupnZ|βrβγθγψ|,\left\|\psi\right\|_{k}=\max_{\begin{subarray}{c}\beta,\gamma\in\mathbb{N}^{n}\\ \left|\beta\right|+\left|\gamma\right|\leq k\end{subarray}}\,\sup_{{\mathbb{C}}^{n}\setminus Z}\left|\frac{\partial^{\beta}}{\partial r^{\beta}}\frac{\partial^{\gamma}}{\partial\theta^{\gamma}}\psi\right|, (3.21)

where the tuples r(+)n,θnr\in({\mathbb{R}}^{+})^{n},\theta\in{\mathbb{R}}^{n} are the polar coordinates on nZ{\mathbb{C}}^{n}\setminus Z specified by zj=reiθjz_{j}=re^{i\theta_{j}}, and βrβ,γθγ\frac{\partial^{\beta}}{\partial r^{\beta}},\frac{\partial^{\gamma}}{\partial\theta^{\gamma}} are partial derivatives operators in the polar coordinates defined as in (3.2). From the formulas

xj=cosθjrjsinθjrjθj and yj=sinθjrj+cosθjrjθj,{\frac{\partial}{\partial x_{j}}=\cos\theta_{j}\frac{\partial}{\partial r_{j}}-\frac{\sin\theta_{j}}{r_{j}}\frac{\partial}{\partial\theta_{j}}}\text{ and }{\frac{\partial}{\partial y_{j}}={\sin\theta_{j}}\frac{\partial}{\partial r_{j}}+\frac{\cos\theta_{j}}{r_{j}}\frac{\partial}{\partial\theta_{j}}}, (3.22)

we see that for 0<ϱ1<ϱ2<0<\varrho_{1}<\varrho_{2}<\infty, there is a constant Bk(ϱ1,ϱ2)B_{k}(\varrho_{1},\varrho_{2}) such that for each compact set KK such that

K{zn:ϱ1<|zj|<ϱ2},K\subset\{z\in{\mathbb{C}}^{n}:\varrho_{1}<\left|z_{j}\right|<\varrho_{2}\},

we have

1Bk(ϱ1,ϱ2)ψkψ𝒞kBk(ϱ1,ϱ2)ψk.\frac{1}{B_{k}(\varrho_{1},\varrho_{2})}\cdot\left\|\psi\right\|_{k}\leq\left\|\psi\right\|_{\mathcal{C}^{k}}\leq B_{k}(\varrho_{1},\varrho_{2})\cdot\left\|\psi\right\|_{k}. (3.23)

We will need the following elementary estimate:

Proposition 3.4.

Let τ\tau be the representation of 𝕋n\mathbb{T}^{n} on nZ{\mathbb{C}}^{n}\setminus Z given by (3.20). For integers m,k0m,k\geq 0, and a compact KnZK\subset{\mathbb{C}}^{n}\setminus Z there is a constant C>0C>0 such that for each ψ𝒟K\psi\in\mathscr{D}_{K} and each αn\alpha\in\mathbb{Z}^{n}, such that |αj|2k\left|\alpha_{j}\right|\geq 2k for 1jn1\leq j\leq n, we have

𝝅ατψ𝒞kCj=1n|αj|mψ𝒞2nk+nm.\left\|\bm{\pi}^{\tau}_{\alpha}\psi\right\|_{\mathcal{C}^{k}}\leq\frac{C}{\prod\limits_{j=1}^{n}\left|\alpha_{j}\right|^{m}}\cdot\left\|\psi\right\|_{\mathcal{C}^{2nk+nm}}.
Proof.

For r(+)n,αnr\in({\mathbb{R}}^{+})^{n},\alpha\in\mathbb{Z}^{n}, define the α\alpha-th Fourier coefficient of ψ\psi by

ψ^(r,α)=1(2π)n[0,2π]nψ(reiθ)eiα,θ𝑑θ,\widehat{\psi}(r,\alpha)=\frac{1}{(2\pi)^{n}}\int\limits_{[0,2\pi]^{n}}\psi(re^{i\theta})e^{-i\left\langle\alpha,\theta\right\rangle}d\theta, (3.24)

where reiθ=(r1eiθ1,,rneiθn)nZre^{i\theta}=(r_{1}e^{i\theta_{1}},\dots,r_{n}e^{i\theta_{n}})\in{\mathbb{C}}^{n}\setminus Z, α,θ=j=1nαjθj\left\langle\alpha,\theta\right\rangle=\sum_{j=1}^{n}\alpha_{j}\theta_{j} and dθ=dθ1dθnd\theta=d\theta_{1}\dots d\theta_{n} is the Lebesgue measure. We will also write (ψ)(\psi)^{\wedge} for ψ^\widehat{\psi} whenever convenient. We note the following properties:

  1. (1)

    We clearly have

    supr(+)n|ψ^(r,α)|supnZ|ψ|.\sup_{r\in({\mathbb{R}}^{+})^{n}}\left|\widehat{\psi}(r,\alpha)\right|\leq\sup_{{\mathbb{C}}^{n}\setminus Z}\left|\psi\right|. (3.25)
  2. (2)

    If α0\alpha\not=0, by a standard integration by parts argument, for each n\ell\in\mathbb{N}^{n},

    ψ^(r,α)=1(iα)(ψθ)(r,α),\widehat{\psi}(r,\alpha)=\frac{1}{(i\alpha)^{\ell}}\left(\frac{\partial^{\ell}\psi}{\partial\theta^{\ell}}\right)^{\wedge}(r,\alpha), (3.26)

    where, as usual, we set for ζn\zeta\in{\mathbb{C}}^{n} and βn\beta\in\mathbb{N}^{n}, ζβ=ζ1β1ζnβn\zeta^{\beta}=\zeta_{1}^{\beta_{1}}\dots\zeta_{n}^{\beta_{n}}.

  3. (3)

    Differentiating under the integral sign we have, for any βn\beta\in\mathbb{N}^{n},

    (r)βψ^(r,α)=1(2π)n[0.2π]nβψrβ(reiθ)eiβ,θeiα,θ𝑑θ=(βψrβ)(r,αβ),\left(\frac{\partial}{\partial r}\right)^{\beta}\widehat{\psi}(r,\alpha)=\frac{1}{(2\pi)^{n}}\int\limits_{[0.2\pi]^{n}}\frac{\partial^{\beta}\psi}{\partial r^{\beta}}(re^{i\theta})e^{i\left\langle\beta,\theta\right\rangle}e^{-i\left\langle\alpha,\theta\right\rangle}d\theta={\left(\frac{\partial^{\beta}\psi}{\partial r^{\beta}}\right)}^{\wedge}(r,\alpha-\beta),

    and combining this with (3.26), we see that if βα\beta\not=\alpha, we have for each integer 0\ell\geq 0 that

    (r)βψ^(r,α)=1(i(αβ))(θβrβψ)(r,αβ).\left(\frac{\partial}{\partial r}\right)^{\beta}\widehat{\psi}(r,\alpha)=\frac{1}{\left(i(\alpha-\beta)\right)^{\ell}}\left(\frac{\partial^{\ell}}{\partial\theta^{\ell}}\frac{\partial^{\beta}}{\partial r^{\beta}}\psi\right)^{\wedge}(r,\alpha-\beta). (3.27)

Now for reiθnZre^{i\theta}\in{\mathbb{C}}^{n}\setminus Z the evaluation 𝒟(nZ)\mathscr{D}({\mathbb{C}}^{n}\setminus Z)\to{\mathbb{C}}, ψψ(reiθ)\psi\mapsto\psi(re^{i\theta}) is continuous, so using (2.4) we see that for each αn\alpha\in\mathbb{Z}^{n} and each ψ𝒟(nZ)\psi\in\mathscr{D}({\mathbb{C}}^{n}\setminus Z), we have

𝝅ατψ(reiθ)\displaystyle\bm{\pi}^{\tau}_{\alpha}\psi(re^{i\theta}) =𝕋λα(τλψ)(reiθ)𝑑λ=1(2π)n[0,2π]neiα,ϕψ(eiϕreiθ)𝑑ϕ=eiα,θψ^(r,α),\displaystyle=\int_{\mathbb{T}}\lambda^{-\alpha}(\tau_{\lambda}\psi)(re^{i\theta})d\lambda=\frac{1}{(2\pi)^{n}}\int_{[0,2\pi]^{n}}e^{-i\left\langle\alpha,\phi\right\rangle}\psi(e^{i\phi}\cdot re^{i\theta})d\phi=e^{i\left\langle\alpha,\theta\right\rangle}\cdot\widehat{\psi}(r,\alpha), (3.28)

where in the last step we make a change of variables in the integral from ϕ\phi to ϕ+θ\phi+\theta. Therefore, if β,γn\beta,\gamma\in\mathbb{N}^{n} with βjαj\beta_{j}\not=\alpha_{j} for 1jn1\leq j\leq n, then

γθγβrβ(𝝅ατψ)(reiθ)=eiα,θ(iα)γ(i(αβ))(θβrβψ)(r,αβ).\frac{\partial^{\gamma}}{\partial\theta^{\gamma}}\frac{\partial^{\beta}}{\partial r^{\beta}}(\bm{\pi}^{\tau}_{\alpha}\psi)(re^{i\theta})=e^{i\left\langle\alpha,\theta\right\rangle}\cdot\frac{(i\alpha)^{\gamma}}{\left(i(\alpha-\beta)\right)^{\ell}}{\left(\frac{\partial^{\ell}}{\partial\theta^{\ell}}\frac{\partial^{\beta}}{\partial r^{\beta}}\psi\right)}^{\wedge}(r,\alpha-\beta).

Now in the above formula, if we have |β|+|γ|k\left|\beta\right|+\left|\gamma\right|\leq k, and let

=(k+m,,k+m)=(k+m)𝟏,\ell=(k+m,\dots,k+m)=(k+m)\bm{1},

where 𝟏n\bm{1}\in\mathbb{N}^{n} is the multi-index each of whose nn entries is 1, we obtain, after taking absolute values of both sides

|γθγβrβ(𝝅ατψ)(reiθ)|=j=1n|αj|γjj=1n|αjβj|m+k|((m+k)𝟏θ(m+k)𝟏βrβψ)(r,αβ)|.\left|\frac{\partial^{\gamma}}{\partial\theta^{\gamma}}\frac{\partial^{\beta}}{\partial r^{\beta}}(\bm{\pi}^{\tau}_{\alpha}\psi)(re^{i\theta})\right|=\frac{\prod\limits_{j=1}^{n}\left|\alpha_{j}\right|^{\gamma_{j}}}{\prod\limits_{j=1}^{n}\left|\alpha_{j}-\beta_{j}\right|^{m+k}}\cdot\left|\left(\frac{\partial^{(m+k)\bm{1}}}{\partial\theta^{(m+k)\bm{1}}}\frac{\partial^{\beta}}{\partial r^{\beta}}\psi\right)^{\wedge}(r,\alpha-\beta)\right|. (3.29)

In the first factor on the right hand side, we have by hypothesis for each jj that |αj|2k\left|\alpha_{j}\right|\geq 2k and 0βj,γjk0\leq\beta_{j},\gamma_{j}\leq k. Therefore, we have |αjβj|12|αj|\left|\alpha_{j}-\beta_{j}\right|\geq\frac{1}{2}\left|\alpha_{j}\right|. The first factor can be estimated as

j=1n|αj|γjj=1n|αjβj|m+kj=1n|αj|kj=1n(12|αj|m+k)=2m+kj=1n|αj|m.\frac{\prod\limits_{j=1}^{n}\left|\alpha_{j}\right|^{\gamma_{j}}}{\prod\limits_{j=1}^{n}\left|\alpha_{j}-\beta_{j}\right|^{m+k}}\leq\frac{\prod\limits_{j=1}^{n}\left|\alpha_{j}\right|^{k}}{\prod\limits_{j=1}^{n}\left(\frac{1}{2}\left|\alpha_{j}\right|^{m+k}\right)}=\frac{2^{m+k}}{\prod\limits_{j=1}^{n}\left|\alpha_{j}\right|^{m}}. (3.30)

Using (3.25), the second factor can be estimated as

|((m+k)𝟏θ(m+k)𝟏βrβψ)(r,αβ)|supnZ|(m+k)𝟏θ(m+k)𝟏βrβψ|ψ2nk+nm,\left|\left(\frac{\partial^{(m+k)\bm{1}}}{\partial\theta^{(m+k)\bm{1}}}\frac{\partial^{\beta}}{\partial r^{\beta}}\psi\right)^{\wedge}(r,\alpha-\beta)\right|\leq\sup_{{\mathbb{C}}^{n}\setminus Z}\left|\frac{\partial^{(m+k)\bm{1}}}{\partial\theta^{(m+k)\bm{1}}}\frac{\partial^{\beta}}{\partial r^{\beta}}\psi\right|\leq\left\|\psi\right\|_{2nk+nm}, (3.31)

where in the last step we use the norm introduced in (3.21), and used the fact that

|(m+k)𝟏+β|=(m+k)n+|β|(m+k)n+nk,\left|(m+k)\bm{1}+\beta\right|=(m+k)n+\left|\beta\right|\leq(m+k)n+nk,

since each βjk\beta_{j}\leq k. Combining (3.29), (3.30) and (3.31) we see that

|γθγβrβ(𝝅ατψ)(reiθ)|2m+kj=1n|αj|mψ2nk+nm,\left|\frac{\partial^{\gamma}}{\partial\theta^{\gamma}}\frac{\partial^{\beta}}{\partial r^{\beta}}(\bm{\pi}^{\tau}_{\alpha}\psi)(re^{i\theta})\right|\leq\frac{2^{m+k}}{\prod\limits_{j=1}^{n}\left|\alpha_{j}\right|^{m}}\cdot\left\|\psi\right\|_{2nk+nm},

from which, taking a supremum on the left hand side over reiθZre^{i\theta}\in{\mathbb{C}}\setminus Z, and remembering that |β|+|γ|k\left|\beta\right|+\left|\gamma\right|\leq k, we conclude that

𝝅ατψk2m+kj=1n|αj|mψ2nk+nm.\left\|\bm{\pi}^{\tau}_{\alpha}\psi\right\|_{k}\leq\frac{2^{m+k}}{\prod\limits_{j=1}^{n}\left|\alpha_{j}\right|^{m}}\cdot\left\|\psi\right\|_{2nk+nm}.

Recall that KK is a compact set in nZ{\mathbb{C}}^{n}\setminus Z such that ψ𝒟K\psi\in\mathscr{D}_{K}, i.e. , the support of ψ\psi is contained in KK. Suppose that ϱ1,ϱ2\varrho_{1},\varrho_{2} are such that KAK\subset A, where AA is the product of annuli A={zn:ϱ1<|zj|<ϱ2,1jn}A=\{z\in{\mathbb{C}}^{n}:\varrho_{1}<\left|z_{j}\right|<\varrho_{2},1\leq j\leq n\}. Formulas (3.24) and (3.28) show that the compact support of 𝝅ατψ\bm{\pi}^{\tau}_{\alpha}\psi is also contained in the set AA. Therefore, passing to the equivalent 𝒞k\mathcal{C}^{k}-norms using (3.23), we have

𝝅ατψ𝒞kBk(ϱ1,ϱ2)B2nk+nm(ϱ1,ϱ2)2m+kj=1n|αj|mψ𝒞2nk+nm,\left\|\bm{\pi}^{\tau}_{\alpha}\psi\right\|_{\mathcal{C}^{k}}\leq B_{k}(\varrho_{1},\varrho_{2})\cdot B_{2nk+nm}(\varrho_{1},\varrho_{2})\cdot\frac{2^{m+k}}{\prod\limits_{j=1}^{n}\left|\alpha_{j}\right|^{m}}\cdot\left\|\psi\right\|_{\mathcal{C}^{2nk+nm}},

which completes the proof of the result. ∎

3.6. The dual representation

Let Ω\Omega be a Reinhardt domain. Since for each λ𝕋n\lambda\in\mathbb{T}^{n}, the map Rλ\mathrm{R}_{\lambda} maps Ω\Omega biholomorphically (and therefore diffeomorphically) to itself. Consequently, we can define a representation of 𝕋n\mathbb{T}^{n} on the space of distributions 𝒟(Ω)\mathscr{D}^{\prime}(\Omega) using the pullback operation

σλ(T)=(Rλ)(T),T𝒟(Ω).\sigma_{\lambda}(T)=(\mathrm{R}_{\lambda})^{*}(T),\quad T\in\mathscr{D}^{\prime}(\Omega). (3.32)

The representation σλ\sigma_{\lambda} is closely related to the representation τλ\tau_{\lambda} introduced in (3.20). Clearly, 𝒟(Ω)\mathscr{D}(\Omega) is an invariant (dense) subspace of σ\sigma, on which σ\sigma restricts to τ\tau. So σ\sigma is simply the extension of the representation τ\tau of (3.20) by continuity to the space of distributions.

The representation σ\sigma on 𝒟(Ω)\mathscr{D}^{\prime}(\Omega) is also “dual” to the representation τ\tau on 𝒟(Ω)\mathscr{D}(\Omega):

σλ(T),ϕ=(Rλ)(T),ϕ=T,1det(Rλ)ϕRλ1=T,τλ1ϕ,\left\langle\sigma_{\lambda}(T),\phi\right\rangle=\left\langle(\mathrm{R}_{\lambda})^{*}(T),\phi\right\rangle=\left\langle T,\frac{1}{\det_{\mathbb{R}}(\mathrm{R}_{\lambda})}\phi\circ\mathrm{R}_{\lambda}^{-1}\right\rangle=\left\langle T,\tau_{\lambda^{-1}}\phi\right\rangle, (3.33)

so that σλ\sigma_{\lambda} is the transpose of the map τλ1\tau_{\lambda^{-1}}. The second equality in the chain may be proved by change of variables when TT is a test function, and then using density.

Proposition 3.5.

The representation σ\sigma of 𝕋n\mathbb{T}^{n} on 𝒟(Ω)\mathscr{D}^{\prime}(\Omega) defined in (3.32) is continuous.

Proof.

Let {λj}\{\lambda_{j}\} be a sequence in 𝕋n\mathbb{T}^{n} converging to the identity element. Since pointwise convergence of a sequence in 𝒟(Ω)\mathscr{D}^{\prime}(\Omega) on each test function implies convergence in the strong dual topology (see [Trè67]), to show that σλj(T)T\sigma_{\lambda_{j}}(T)\to T in 𝒟(Ω)\mathscr{D}^{\prime}(\Omega) we need to show that for ϕ𝒟(Ω)\phi\in\mathscr{D}(\Omega) we have σλj(T),ϕT,ϕ\left\langle\sigma_{\lambda_{j}}(T),\phi\right\rangle\to\left\langle T,\phi\right\rangle. But by (3.33), σλj(T),ϕ=T,τλj1ϕ\left\langle\sigma_{\lambda_{j}}(T),\phi\right\rangle=\left\langle T,\tau_{\lambda_{j}^{-1}}\phi\right\rangle and by the continuity of the representation τ\tau, to follows that σλj(T)T\sigma_{\lambda_{j}}(T)\to T in 𝒟(Ω)\mathscr{D}^{\prime}(\Omega).

To complete the proof, by Proposition 2.3, we need to show that the topology of Ω\Omega is generated by a collection of σ\sigma-invariant seminorms. Let BB be a bounded subset of 𝒟(Ω)\mathscr{D}(\Omega). Let

B~={σλ(ϕ):λ𝕋n,ϕB}.\widetilde{B}=\{\sigma_{\lambda}(\phi):\lambda\in\mathbb{T}^{n},\phi\in B\}.

Then it is clear from (3.1) that B~\widetilde{B} is also a bounded subset of 𝒟(Ω)\mathscr{D}(\Omega). With notation as it (3.3), it is clear that for each T𝒟(Ω)T\in\mathscr{D}^{\prime}(\Omega), we have pB(T)pB~(T)p_{B}(T)\leq p_{\widetilde{B}}(T). Therefore the topology of 𝒟(Ω)\mathscr{D}^{\prime}(\Omega) is generated by the family of continuous invariant seminorms {pB~}\{p_{\widetilde{B}}\}. ∎

3.7. Fourier series of distributions on Reinhardt domains

Let Ω\Omega be a Reinhardt domain and let T𝒟(Ω)T\in\mathscr{D}^{\prime}(\Omega) be a distribution. Then by the results of Section 2 we can expand TT in a formal Fourier series with respect to the representation σ\sigma of (3.32). For simplicity of notation, whenever there is no possibility of confusion, we denote the Fourier components of TT by

Tα=𝝅ασ(T),αn,T_{\alpha}=\bm{\pi}^{\sigma}_{\alpha}(T),\quad\alpha\in\mathbb{Z}^{n}, (3.34)

so that the Fourier series of TT is written as TαnTαT\sim\sum\limits_{\alpha\in\mathbb{Z}^{n}}T_{\alpha}. We notice the following properties of the Fourier components:

Proposition 3.6.

Let Ω\Omega and TT be as above and let αn\alpha\in\mathbb{Z}^{n}.

  1. (a)

    If the distribution TT lies in one of the linear subspaces 𝒪(Ω),𝒞(Ω),(𝒪𝒞)(Ω)\mathscr{O}(\Omega),\mathcal{C}^{\infty}(\Omega),(\mathscr{O}\cap\mathcal{C}^{\infty})(\Omega) or 𝒟(Ω)\mathscr{D}(\Omega) of 𝒟(Ω)\mathscr{D}^{\prime}(\Omega), the Fourier component TαT_{\alpha} also lies in the same subspace.

  2. (b)

    If UU is another Reinhardt domain such that UΩU\subset\Omega, we have

    Tα|U=(T|U)α.T_{\alpha}|_{U}=\left(T|_{U}\right)_{\alpha}. (3.35)
Proof.

All these are consequences of Part 2 of Proposition 2.4. To see Part (a), let XX be the space 𝒟(Ω)\mathscr{D}^{\prime}(\Omega) and let YY be one of the spaces 𝒪(Ω),𝒞(Ω)\mathscr{O}(\Omega),\mathcal{C}^{\infty}(\Omega) or 𝒟(Ω)\mathscr{D}(\Omega). Each of these has a natural structure of a complete LCTVS (the space 𝒪(Ω)\mathscr{O}(\Omega) as a closed subspace of 𝒟(Ω)\mathscr{D}^{\prime}(\Omega), 𝒞(Ω)\mathcal{C}^{\infty}(\Omega) in its Fréchet topology and 𝒟(Ω)\mathscr{D}(\Omega) in its LFLF-topology), and for each one of these topologies, the representation σ\sigma restricts to a continuous representation (in the stronger topology of the subspace), which we will call τ\tau (this coincides with the τ\tau introduced in (3.20)). If jj is the inclusion map of any one of these subspaces into 𝒟(Ω)\mathscr{D}^{\prime}(\Omega), then clearly it is continuous (where the subspace has the natural topology). Therefore, by (2.7), we have j(𝝅ατT)=𝝅ασ(j(T))j\left(\bm{\pi}^{\tau}_{\alpha}T\right)=\bm{\pi}^{\sigma}_{\alpha}(j(T)). By definition, the right hand side is precisely TαT_{\alpha}, i.e. the Fourier component of TT as an element of 𝒟(Ω)\mathscr{D}^{\prime}(\Omega). Notice that 𝝅ατTY\bm{\pi}^{\tau}_{\alpha}T\in Y, since it is the Fourier component of TT as an element of YY. Since jj is the inclusion of YY in XX, it now follows that TαYT_{\alpha}\in Y as well.

Since the result is true for 𝒪(Ω)\mathscr{O}(\Omega) and 𝒞(Ω)\mathcal{C}^{\infty}(\Omega) it follows for (𝒪𝒞)(Ω)(\mathscr{O}\cap\mathcal{C}^{\infty})(\Omega).

For part (b), in part 2 of Proposition 2.4, let X=𝒟(U)X=\mathscr{D}^{\prime}(U) with representation defined in (3.32), which we call σ\sigma^{\prime} for clarity, and let Y=𝒟(Ω)Y=\mathscr{D}^{\prime}(\Omega) with the representation σ\sigma and let j:YXj:Y\to X be the restriction map of distributions, which is clearly continuous and intertwines the two representations. Then (2.7) takes the form j𝝅ασ=𝝅ασjj\circ\bm{\pi}^{\sigma}_{\alpha}=\bm{\pi}^{\sigma^{\prime}}_{\alpha}\circ j. Applying this to the distribution TT, we see that

Tα|U=j(𝝅ασT)=𝝅ασ(j(T))=(T|U)α.T_{\alpha}|_{U}=j\left(\bm{\pi}^{\sigma}_{\alpha}T\right)=\bm{\pi}^{\sigma^{\prime}}_{\alpha}\left(j(T)\right)=\left(T|_{U}\right)_{\alpha}.

We establish the following lemma:

Lemma 3.7.

Let Ωn\Omega\subset{\mathbb{C}}^{n} be a Reinhardt domain, and let

𝒜={T𝒪(Ω):each Fourier component Tα belongs to 𝒞(Ω)}.\mathscr{A}=\{T\in\mathscr{O}(\Omega):\text{each Fourier component }T_{\alpha}\text{ belongs to }\mathcal{C}^{\infty}(\Omega)\}.

Suppose that for each T𝒜T\in\mathscr{A} the Laurent series αnTα\sum\limits_{\alpha\in\mathbb{Z}^{n}}T_{\alpha} converges absolutely in the Fréchet space 𝒞(Ω)\mathcal{C}(\Omega) . Then the Laurent series of each element of 𝒜\mathscr{A} converges absolutely in the space 𝒞(Ω)\mathcal{C}^{\infty}(\Omega).

Proof.

For a function f𝒞(Ω)f\in\mathcal{C}(\Omega) and a compact set KK, let pKp_{K} be the seminorm:

pK(f)=supK|f|.p_{K}(f)=\sup_{K}\left|f\right|. (3.36)

Notice that the family {pK:KΩcompact}\{p_{K}:K\subset\Omega\quad\text{compact}\} is a generating family of seminorms for 𝒞(Ω)\mathcal{C}(\Omega), and the hypothesis on 𝒜\mathscr{A} can be expressed as

αnpK(Tα)<,\sum\limits_{\alpha\in\mathbb{Z}^{n}}p_{K}(T_{\alpha})<\infty, (3.37)

for each compact KΩK\subset\Omega.

The Fréchet topology of 𝒞(Ω)\mathcal{C}^{\infty}(\Omega) is generated by the seminorms pK,Lp_{K,L}, where KK ranges over compact subsets of Ω\Omega, LL is a constant coefficient differential operator on n{\mathbb{C}}^{n}, and

pK,L(f)=supK|Lf|.p_{K,L}(f)=\sup_{K}\left|Lf\right|.

We will continue to write pKp_{K} for the seminorm (3.36) corresponding to L=1L=1. From the distributional chain rule (3.13), we have for each jj that

𝑫jσλ(T)=𝑫j(RλT)=(𝑫jRλ)Rλ(𝑫jT)=λjσλ(𝑫jT).\bm{D}_{j}\sigma_{\lambda}(T)=\bm{D}_{j}(\mathrm{R}_{\lambda}^{*}T)=(\bm{D}_{j}\mathrm{R}_{\lambda})\cdot\mathrm{R}_{\lambda}^{*}(\bm{D}_{j}T)=\lambda_{j}\cdot\sigma_{\lambda}(\bm{D}_{j}T).

Let βn\beta\in\mathbb{N}^{n}, and denote 𝑫β=𝑫1β1𝑫nβn\bm{D}^{\beta}=\bm{D}_{1}^{\beta_{1}}\dots\bm{D}_{n}^{\beta_{n}}. Then, from the above we have

𝑫βTα=𝑫β𝕋nλασλ(T)𝑑λ=𝕋nλα𝑫βσλ(T)𝑑λ=𝕋nλ(αβ)σλ(𝑫βT)𝑑λ=(𝑫βT)αβ,\displaystyle\bm{D}^{\beta}T_{\alpha}=\bm{D}^{\beta}\int_{\mathbb{T}^{n}}\lambda^{-\alpha}\sigma_{\lambda}(T)d\lambda=\int_{\mathbb{T}^{n}}\lambda^{-\alpha}\bm{D}^{\beta}\sigma_{\lambda}(T)d\lambda=\int_{\mathbb{T}^{n}}\lambda^{-(\alpha-\beta)}\sigma_{\lambda}(\bm{D}^{\beta}T)d\lambda=(\bm{D}^{\beta}T)_{\alpha-\beta},

where (𝑫βT)αβ=𝝅αβσ(𝑫βT)(\bm{D}^{\beta}T)_{\alpha-\beta}=\bm{\pi}^{\sigma}_{{\alpha-\beta}}(\bm{D}^{\beta}T). It follows that all Fourier coefficients of the distribution 𝑫βT\bm{D}^{\beta}T are in 𝒞(Ω)\mathcal{C}^{\infty}(\Omega), so 𝑫βT𝒜\bm{D}^{\beta}T\in\mathscr{A}. Therefore,

pK,𝑫β(Tα)=pK(𝑫βTα)=pK((𝑫βT)αβ).p_{K,\bm{D}^{\beta}}(T_{\alpha})=p_{K}\left(\bm{D}^{\beta}T_{\alpha}\right)=p_{K}\left((\bm{D}^{\beta}T)_{\alpha-\beta}\right).

It follows therefore that

αnpK,𝑫β(Tα)=γnpK((𝑫βT)γ)<,\sum_{\alpha\in\mathbb{Z}^{n}}p_{K,\bm{D}^{\beta}}(T_{\alpha})=\sum_{\gamma\in\mathbb{Z}^{n}}p_{K}\left((\bm{D}^{\beta}T)_{\gamma}\right)<\infty,

using assumption (3.37) on elements of 𝒜\mathscr{A}. A partial derivative L=γxγδyδL=\frac{\partial^{\gamma}}{\partial x^{\gamma}}\frac{\partial^{\delta}}{\partial y^{\delta}} on n{\mathbb{C}}^{n} can be rewritten as a polynomial in the commuting differential operators 𝑫1,,𝑫n,𝑫¯1,,𝑫¯n\bm{D}_{1},\dots,\bm{D}_{n},\bm{\overline{D}}_{1},\dots,\bm{\overline{D}}_{n}, and by hypothesis the derivatives 𝑫¯j\bm{\overline{D}}_{j} are zero. Therefore, it follows that αpK,L(Tα)<\sum_{\alpha}p_{K,L}(T_{\alpha})<\infty. Consequently the series αTα\sum_{\alpha}T_{\alpha} is absolutely convergent in 𝒞(Ω)\mathcal{C}^{\infty}(\Omega).

3.8. Proof of Theorem 3.1: case of annular domains

We now prove a weaker version of Theorem 3.1. A Reinhardt domain Ωn\Omega\subset{\mathbb{C}}^{n} will be called annular if it is disjoint from the coordinate hyperplanes:

ΩZ=,\Omega\cap Z=\emptyset, (3.38)

where ZZ is as in (3.7). We have the following:

Proposition 3.8.

Let ΩnZ\Omega\subset{\mathbb{C}}^{n}\setminus Z be an annular Reinhardt domain in n{\mathbb{C}}^{n}. Then for each αn\alpha\in\mathbb{Z}^{n} there is a continuous linear functional aα:𝒪(Ω)a_{\alpha}:\mathscr{O}(\Omega)\to{\mathbb{C}} such that for each T𝒪(Ω)T\in\mathscr{O}(\Omega), the series of terms in 𝒞(Ω)\mathcal{C}^{\infty}(\Omega) given by αnaα(T)eα\sum_{\alpha\in\mathbb{Z}^{n}}a_{\alpha}(T)e_{\alpha} converges absolutely in 𝒞(Ω)\mathcal{C}^{\infty}({\Omega}) to a function ff, and this 𝒞\mathcal{C}^{\infty}-smooth function generates the distribution TT.

Proof.

Let T𝒪(Ω)T\in\mathscr{O}(\Omega) and let TαT_{\alpha} be its α\alpha-th Fourier component as in (3.34), and let S=eαTαS=e_{-\alpha}\cdot T_{\alpha}. The distributional Leibniz rule

𝑫¯j(fU)=(𝑫¯jf)U+f(𝑫¯jU),f𝒞(Ω) and U𝒟(Ω),\bm{\overline{D}}_{j}(fU)=\left(\bm{\overline{D}}_{j}f\right)\cdot U+f\cdot\left(\bm{\overline{D}}_{j}U\right),\quad f\in\mathcal{C}^{\infty}(\Omega)\text{ and }U\in\mathscr{D}^{\prime}(\Omega), (3.39)

shows that S𝒪(Ω)S\in\mathscr{O}(\Omega), since Tα𝒪(Ω)T_{\alpha}\in\mathscr{O}(\Omega) by Part (1) of Proposition 3.6. By Part (1) of Proposition 2.4, the Fourier component TαT_{\alpha} lies in the α\alpha-th Fourier mode of 𝒪(Ω)\mathscr{O}(\Omega), i.e., σλ(Tα)=λαTα\sigma_{\lambda}(T_{\alpha})=\lambda^{\alpha}\cdot T_{\alpha}, so

σλ(S)=λαeαλαTα=S.\sigma_{\lambda}(S)=\lambda^{-\alpha}e_{-\alpha}\cdot\lambda^{\alpha}T_{\alpha}=S.

Therefore, using polar coordinates zj=rjeiθjz_{j}=r_{j}e^{i\theta_{j}}, with ϵj(h)=(1,,1,eih,,1)𝕋n\epsilon_{j}(h)=(1,\dots,1,e^{ih},\dots,1)\in\mathbb{T}^{n} (with eihe^{ih} in the jj-th position) we have

θjS=limh01h(σϵj(h)(S)S)=0.\frac{\partial}{\partial\theta_{j}}S=\lim_{h\to 0}\frac{1}{h}\left(\sigma_{\epsilon_{j}(h)}(S)-S\right)=0.

Since 𝑫¯jS=0\bm{\overline{D}}_{j}S=0, and in polar coordinates we have

𝑫¯j=eiθj2(rj+irjθj),\bm{\overline{D}}_{j}=\frac{e^{i\theta_{j}}}{2}\left(\frac{\partial}{\partial r_{j}}+\frac{i}{r_{j}}\frac{\partial}{\partial\theta_{j}}\right), (3.40)

we see that

rjS=0\dfrac{\partial}{\partial r_{j}}S=0

also. Now from the representations (3.22) we have that xjS=0\displaystyle{\frac{\partial}{\partial x_{j}}S=0} and yjS=0\displaystyle{\frac{\partial}{\partial y_{j}}S=0} in 𝒟(Ω)\mathscr{D}^{\prime}(\Omega) for 1jn1\leq j\leq n. Therefore, by a classical result in the theory of distributions (see [Sch66, Theorème VI, pp. 69ff.]) we have that SS is locally constant on Ω\Omega, or more precisely, SS is generated by a locally constant function. Since Ω\Omega is connected, it follows that there is a constant aα(T)a_{\alpha}(T)\in{\mathbb{C}} which generates the distribution SS, so that Tα=aα(T)eα.T_{\alpha}=a_{\alpha}(T)e_{\alpha}. This can also be written as aα(T)=eαTα=eα𝝅ασ(T)a_{\alpha}(T)=e_{-\alpha}\cdot T_{\alpha}=e_{-\alpha}\cdot\bm{\pi}^{\sigma}_{\alpha}(T). Since multiplication by the fixed 𝒞\mathcal{C}^{\infty}-function eαe_{-\alpha} is a continuous map on 𝒟(Ω)\mathscr{D}^{\prime}(\Omega) (and therefore on the subspace 𝒪(Ω)\mathscr{O}(\Omega)), and 𝝅ασ:𝒪(Ω)𝒪(Ω)\bm{\pi}^{\sigma}_{\alpha}:\mathscr{O}(\Omega)\to\mathscr{O}(\Omega) is continuous by part 1 of Proposition 2.4, it follows that aα:𝒪(Ω)𝒪(Ω)a_{\alpha}:\mathscr{O}(\Omega)\to\mathscr{O}(\Omega) is continuous. But we know that aαa_{\alpha} 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 𝒪(Ω)\mathscr{O}(\Omega) on the subspace of constants coincides with the natural topology of {\mathbb{C}}. Therefore aα:𝒪(Ω)a_{\alpha}:\mathscr{O}(\Omega)\to{\mathbb{C}} is continuous.

By the above, the Fourier components TαT_{\alpha} of the Fourier series of TT actually lie in (𝒪𝒞)(Ω)(\mathscr{O}\cap\mathcal{C}^{\infty})(\Omega), and thus any partial sum (i.e. sum of finitely many terms) of this series also lies in (𝒪𝒞)(Ω)(\mathscr{O}\cap\mathcal{C}^{\infty})(\Omega). We will now show that the series (3.34) is absolutely convergent in the topology of 𝒞(Ω)\mathcal{C}^{\infty}(\Omega), in the sense of Proposition 2.2. Let KΩK\subset\Omega be compact. Let δ<dist(K,nΩ)\delta<\mathrm{dist}(K,{\mathbb{C}}^{n}\setminus\Omega), and let Ψ𝒟(n)\Psi\in\mathscr{D}({\mathbb{C}}^{n}) be a test function such that

support(Ψ)B(0,δ),Ψ(ζ)𝑑V(ζ)=1, and Ψ(ζ)=Ψ(|ζ1|,,|ζn|) for ζn,\mathrm{support}(\Psi)\subset B(0,\delta),\quad\int_{\mathbb{C}}\Psi(\zeta)dV(\zeta)=1,\quad\text{ and }\Psi(\zeta)=\Psi(\left|\zeta_{1}\right|,\dots,\left|\zeta_{n}\right|)\text{ for $\zeta\in{\mathbb{C}}^{n}$},

i.e. Ψ\Psi is radial around the origin in each complex coordinate. For zKz\in K, define ψz(ζ)=Ψ(ζz),\psi_{z}(\zeta)=\Psi(\zeta-z), so that ψz\psi_{z} is radially symmetric around zz in each complex direction. Therefore, for zKz\in K, we have, by Lemma 3.3 for each α\alpha that

Tα,ψz=aα(T)eα,ψz=aα(T)zα.\left\langle T_{\alpha},\psi_{z}\right\rangle=\left\langle a_{\alpha}(T)e_{\alpha},\psi_{z}\right\rangle=a_{\alpha}(T)z^{\alpha}.

By the continuity of the mapping T:𝒟(Ω)T:\mathscr{D}({\Omega})\to{\mathbb{C}}, there is a constant C>0C>0 and an integer k0k\geq 0 such that for all ψ𝒟(Ω)\psi\in\mathscr{D}({\Omega}) with support in KK, we have that

|T,ψ|Cψ𝒞k(Ω).\left|\left\langle T,\psi\right\rangle\right|\leq C\cdot\left\|\psi\right\|_{\mathcal{C}^{k}({\Omega})}. (3.41)

Also notice that for any ψ𝒟(Ω)\psi\in\mathscr{D}({\Omega}) we have

𝝅ασ(T),ψ\displaystyle\left\langle\bm{\pi}^{\sigma}_{\alpha}(T),\psi\right\rangle =𝕋nλασλ(T)𝑑λ,ψ=𝕋nλασλ(T),ψ𝑑λusing (2.4)\displaystyle=\left\langle\int_{\mathbb{T}^{n}}\lambda^{-\alpha}\sigma_{\lambda}(T)d\lambda,\psi\right\rangle=\int_{\mathbb{T}^{n}}\lambda^{-\alpha}\left\langle\sigma_{\lambda}(T),\psi\right\rangle d\lambda\qquad\text{using \eqref{eq-t}}
=𝕋nλαT,τλ1ψ𝑑λ=T,𝕋nλατλ1ψ𝑑λusing (3.33) and (2.4)\displaystyle=\int_{\mathbb{T}^{n}}\lambda^{-\alpha}\left\langle T,\tau_{\lambda^{-1}}\psi\right\rangle d\lambda=\left\langle T,\int_{\mathbb{T}^{n}}\lambda^{-\alpha}\tau_{\lambda^{-1}}\psi d\lambda\right\rangle\qquad\text{using \eqref{eq-dual} and \eqref{eq-t}}
=T,𝕋nλατλψ𝑑λ\displaystyle=\left\langle T,\int_{\mathbb{T}^{n}}\lambda^{\alpha}\tau_{\lambda}\psi d\lambda\right\rangle
using the invariance of Haar measure of a compact group under inversion
=T,𝝅ατψ.\displaystyle=\left\langle T,\bm{\pi}^{\tau}_{-\alpha}\psi\right\rangle. (3.42)

In the following estimates, let CC denote a constant that depends only on the compact KK and the distribution TT, and may have different values at different occurrences. By combining (3.41) with (3.42), we see that for zK,ϕ𝒟(Ω)z\in K,\phi\in\mathscr{D}({\Omega}) and each α\alpha with |α|2k\left|\alpha\right|\geq 2k , using Lemma 3.3

|aα(T)zα|\displaystyle\left|a_{\alpha}(T)z^{\alpha}\right| =|Tα,ψz|=|𝝅ασT,ψz|=|T,𝝅ατψz|\displaystyle=\left|\left\langle T_{\alpha},\psi_{z}\right\rangle\right|=\left|\left\langle\bm{\pi}^{\sigma}_{\alpha}T,\psi_{z}\right\rangle\right|=\left|\left\langle T,\bm{\pi}^{\tau}_{-\alpha}\psi_{z}\right\rangle\right|
C𝝅ατψz𝒞k(Ω)using (3.41)\displaystyle\leq C\cdot\left\|\bm{\pi}^{\tau}_{-\alpha}\psi_{z}\right\|_{\mathcal{C}^{k}({\Omega})}\quad\text{using \eqref{eq-distn1}}
C|α|2ψz𝒞2nk+2nby Proposition 3.4 with m=2,\displaystyle\leq\frac{C}{\left|\alpha\right|^{2}}\cdot\left\|\psi_{z}\right\|_{\mathcal{C}^{2nk+2n}}\quad\text{by Proposition~{}\ref{prop-est} with $m=2$,}
=C|α|2ψz𝒞2nk+2n,\displaystyle=\frac{C}{\left|\alpha\right|^{2}}\cdot\left\|\psi_{z}\right\|_{\mathcal{C}^{2nk+2n}},

recalling that the local order kk depends only on the distribution TT and the compact set KK. Now, for each zz we have ψz𝒞2nk+2n=Ψ𝒞2nk+2n\left\|\psi_{z}\right\|_{\mathcal{C}^{2nk+2n}}=\left\|\Psi\right\|_{\mathcal{C}^{2nk+2n}} by translation invariance of the norm. Therefore, for each α2k\alpha\geq 2k we obtain the estimate for the seminorm (with the same convention as above on the constant CC)

pK(Tα)=pK(𝝅ασT)=supzK|aα(T)zα|C|α|2Ψ𝒞2nk+2n=C|α|2.p_{K}(T_{\alpha})=p_{K}(\bm{\pi}^{\sigma}_{\alpha}T)=\sup_{z\in K}\left|a_{\alpha}(T)z^{\alpha}\right|\leq\frac{C}{\left|\alpha\right|^{2}}\cdot\left\|\Psi\right\|_{\mathcal{C}^{2nk+2n}}=\frac{C}{\left|\alpha\right|^{2}}. (3.43)

Clearly therefore αnpK(Tα)<\sum\limits_{\alpha\in\mathbb{Z}^{n}}p_{K}(T_{\alpha})<\infty. By Lemma 3.7, the series αTα\sum_{\alpha}T_{\alpha} converges absolutely in 𝒞(Ω)\mathcal{C}^{\infty}(\Omega). Let ff be its sum. Since the inclusion 𝒞(Ω)𝒟(Ω)\mathcal{C}^{\infty}(\Omega)\subset\mathscr{D}^{\prime}(\Omega) is continuous, we see easily that the Fourier series αTα\sum_{\alpha}T_{\alpha} of T𝒟(Ω)T\in\mathscr{D}^{\prime}(\Omega) converges absolutely in 𝒟(Ω)\mathscr{D}^{\prime}(\Omega). Now it follows from Corollary 2.5 that the sum of the series is TT. Thus TT is the distribution generated by ff, and this completes the proof of the proposition. ∎

3.9. Smoothness of holomorphic distributions

Corollary 3.9.

Let Ωn\Omega\subset{\mathbb{C}}^{n} be an open set. Then each holomorphic distribution on Ω\Omega is generated by a 𝒞\mathcal{C}^{\infty}-smooth function, i.e. 𝒪(Ω)=(𝒪𝒞)(Ω)\mathscr{O}(\Omega)=(\mathscr{O}\cap\mathcal{C}^{\infty})(\Omega).

Proof.

Let T𝒪(Ω)T\in\mathscr{O}(\Omega). It suffices to show that TT is a smooth function in a neighborhood of each point pΩp\in\Omega. 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 p=0p=0. Let r>0r>0 be such that the polydisc P(r)P(r) given by

P(r)={|zj|<r,1jn}nP(r)=\{\left|z_{j}\right|<r,1\leq j\leq n\}\subset{\mathbb{C}}^{n}

is contained in Ω\Omega. We will show that TT is a smooth function on P=P(r6)P^{\prime}=P(\frac{r}{6}), i.e., T(𝒪𝒞)(P)T\in(\mathscr{O}\cap\mathcal{C}^{\infty})(P^{\prime}).

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 ana\in{\mathbb{C}}^{n} different from the origin, i.e. for a domain of the form Ma(A)\mathrm{M}_{a}(A) where AA is an annular Reinhardt domain centered at the origin, with Ma\mathrm{M}_{a} as in (3.15) is the translation by aa.

Indeed, let a=(r3,,r3)a=\left(\frac{r}{3},\dots,\frac{r}{3}\right), and set Q=Ma(P(r2)Z)Q=\mathrm{M}_{a}\left(P(\frac{r}{2})\setminus Z\right), so that QQ is the annular Reinhardt domain centered at aa given by

Q={zn:0<|zjr3|<r2,1jn},Q=\left\{z\in{\mathbb{C}}^{n}:0<\left|z_{j}-\frac{r}{3}\right|<\frac{r}{2},1\leq j\leq n\right\},

for which it is easy to verify that PQP(r)P^{\prime}\subset Q\subset P(r). Now since T|Q(𝒪𝒞)(Q)T|_{Q}\in(\mathscr{O}\cap\mathcal{C}^{\infty})(Q) 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 Ω\Omega be an annular Reinhardt domain in n{\mathbb{C}}^{n}. Then the Laurent series of a distribution T𝒪(Ω)T\in\mathscr{O}(\Omega) converges absolutely in 𝒞(Ω^)\mathcal{C}^{\infty}(\widehat{\Omega}) where Ω^\widehat{\Omega} is the smallest log-convex annular Reinhardt domain containing Ω\Omega.

Introduce the following notation: for a compact subset KnZ=()nK\subset{\mathbb{C}}^{n}\setminus Z=({\mathbb{C}}^{*})^{n}, we let

K^={znZ:|eα(z)|supK|eα|,for all αn}.\widehat{K}=\left\{z\in{\mathbb{C}}^{n}\setminus Z:\left|e_{\alpha}(z)\right|\leq\sup_{K}\left|e_{\alpha}\right|,\text{for all $\alpha\in\mathbb{Z}^{n}$}\right\}. (3.44)
Lemma 3.11.

Let Ω\Omega be an annular Reinhardt domain in n{\mathbb{C}}^{n} and let Ω^\widehat{\Omega} be the log-convex annular Reinhardt domain containing Ω\Omega. Then for each point pΩ^p\in\widehat{\Omega} there is a compact neighborhood MM of pp in Ω^\widehat{\Omega} and a compact subset KΩK\subset\Omega such that MK^M\subset\widehat{K}.

Proof.

Let Λ:nZn\Lambda:{\mathbb{C}}^{n}\setminus Z\to{\mathbb{R}}^{n} be the map

Λ(z1,,zn)=(log|z1|,,log|zn|).\Lambda(z_{1},\dots,z_{n})=(\log\left|z_{1}\right|,\dots,\log\left|z_{n}\right|). (3.45)

From (3.8), it follows that an annular Reinhardt domain VnV\subset{\mathbb{C}}^{n} is log-convex if and only if the subset Λ(V)\Lambda(V) of n{\mathbb{R}}^{n} (called the logarithmic shadow of VV) is convex. The set Ω^\widehat{\Omega} is characterized by the fact that it is the convex hull of Λ(Ω)\Lambda(\Omega). Therefore, if pΩ^p\in\widehat{\Omega}, there exist mm points q1,,qmVZq^{1},\dots,q^{m}\in V\setminus Z and positive numbers t1,,tmt_{1},\dots,t_{m}, with k=1mtk=1\sum\limits_{k=1}^{m}t_{k}=1 such that for 1jn1\leq j\leq n, we have Λ(p)=k=1mtkΛ(qk)\Lambda(p)=\sum_{k=1}^{m}t_{k}\Lambda(q^{k}), i.e.,

log|pj|=k1mtklog|qjk|.\log\left|p_{j}\right|=\sum\limits_{k-1}^{m}t_{k}\log\left|q_{j}^{k}\right|.

In other words Λ(p)\Lambda(p) lies in the convex hull of the mm-element set {Λ(q1),,Λ(qm)}\{\Lambda(q^{1}),\dots,\Lambda(q^{m})\}. (By a theorem of Carathéodory, mn+1m\leq n+1, 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 Λ(p)\Lambda(p).

Without loss of generality, we can assume that tm0t_{m}\not=0. Consider the affine self-map FF of n{\mathbb{R}}^{n} given by

F(x)=tmx+k=1m1tkΛ(qjk),F(x)=t_{m}x+\sum_{k=1}^{m-1}t_{k}\Lambda(q_{j}^{k}),

which is clearly an affine automorphism of n{\mathbb{R}}^{n}. Therefore if LL is a compact neighborhood of Λ(p)\Lambda(p) in n{\mathbb{R}}^{n}, then F1(L)F^{-1}(L) is a compact neighborhood of Λ(qm)\Lambda(q^{m}), which after shrinking can be taken to be contained in Λ(Ω).\Lambda(\Omega). Let K0=Λ1(F1(L))K_{0}=\Lambda^{-1}(F^{-1}(L)), and set

K={q1,,qm1}K0.K=\{q^{1},\dots,q^{m-1}\}\cup K_{0}.

We claim that K^\widehat{K} contains the compact neighborhood Λ1(L)\Lambda^{-1}(L) of the point pp. Let ζΛ1(L)\zeta\in\Lambda^{-1}(L). Then there is a point zK0z\in K_{0} such that Λ(ζ)=k=1m1tkΛ(qk)+tmΛ(z)\Lambda(\zeta)=\sum_{k=1}^{m-1}t_{k}\Lambda(q^{k})+t_{m}\Lambda(z), i.e. for each 1jn1\leq j\leq n we have

log|ζj|=k=1m1tklog|qjk|+tmlog|zj|.\log\left|\zeta_{j}\right|=\sum_{k=1}^{m-1}t_{k}\log\left|q^{k}_{j}\right|+t_{m}\log\left|z_{j}\right|.

For simplicity of writing, introduce new notation as follows: we let zk=qkz^{k}=q^{k} for 1km11\leq k\leq m-1 and zm=zz^{m}=z. Then for a multi-index αn\alpha\in\mathbb{Z}^{n} and the point ζΛ1(L)\zeta\in\Lambda^{-1}(L) we have

|eα(ζ)|\displaystyle\left|e_{\alpha}(\zeta)\right| =j=1n|ζj|αj=exp(j=1nαjlog|ζj|)\displaystyle=\prod_{j=1}^{n}\left|\zeta_{j}\right|^{\alpha_{j}}=\exp\left(\sum_{j=1}^{n}\alpha_{j}\log\left|\zeta_{j}\right|\right)
=exp(j=1nαjk=1mtklog|zjk|)=exp(k=1mtkj=1nαjlog|zjk|)\displaystyle=\exp\left(\sum_{j=1}^{n}\alpha_{j}\sum_{k=1}^{m}t_{k}\log\left|z^{k}_{j}\right|\right)=\exp\left(\sum_{k=1}^{m}t_{k}\sum_{j=1}^{n}\alpha_{j}\log\left|z^{k}_{j}\right|\right)
exp(max1kmj=1nαjlog|zjk|)\displaystyle\leq\exp\left(\max_{1\leq k\leq m}\sum_{j=1}^{n}\alpha_{j}\log\left|z^{k}_{j}\right|\right) (3.46)
=max1kmexp(j=1nαjlog|zjk|)\displaystyle=\max_{1\leq k\leq m}\exp\left(\sum_{j=1}^{n}\alpha_{j}\log\left|z^{k}_{j}\right|\right)
=max1km|eα(zk)|=max{max1km1|eα(qk)|,|eα(z)|}\displaystyle=\max_{1\leq k\leq m}\left|e_{\alpha}(z^{k})\right|=\max\left\{\max_{1\leq k\leq m-1}\left|e_{\alpha}(q^{k})\right|,\left|e_{\alpha}(z)\right|\right\}
maxK|eα|,\displaystyle\leq\max_{K}\left|e_{\alpha}\right|, (3.47)

this shows that ζK^\zeta\in\widehat{K}, and completes the proof. ∎

Proof of Proposition 3.10.

By Lemma 3.7 it is sufficient to show that the Laurent series αaα(T)eα\sum_{\alpha}a_{\alpha}(T)e_{\alpha} (whose existence was proved in Proposition 3.8, and whose terms lie in (𝒪𝒞)(Ω)(\mathscr{O}\cap\mathcal{C}^{\infty})(\Omega)) converges absolutely in 𝒞(Ω^)\mathcal{C}(\widehat{\Omega}). It is sufficient to show that each point pp of Ω^\widehat{\Omega} has a compact neighborhood MM such that we have

αpM(aα(T)eα)<.\sum_{\alpha}p_{M}(a_{\alpha}(T)e_{\alpha})<\infty.

Now by Lemma 3.11, there is a compact subset KΩK\subset\Omega such that MK^M\subset\widehat{K}. We observe, using the definition (3.44) that

pM(aα(T)eα)\displaystyle p_{M}(a_{\alpha}(T)e_{\alpha}) pK^(aα(T)eα)=supK^|aα(T)eα|=|aα(T)|supK^|eα|\displaystyle\leq p_{\widehat{K}}(a_{\alpha}(T)e_{\alpha})=\sup_{\widehat{K}}\left|a_{\alpha}(T)e_{\alpha}\right|=\left|a_{\alpha}(T)\right|\cdot\sup_{\widehat{K}}\left|e_{\alpha}\right|
=|aα(T)|supK|eα|=supK|aα(T)eα|=pK(aα(T)eα),\displaystyle=\left|a_{\alpha}(T)\right|\cdot\sup_{{K}}\left|e_{\alpha}\right|=\sup_{{K}}\left|a_{\alpha}(T)e_{\alpha}\right|=p_{K}(a_{\alpha}(T)e_{\alpha}),

so that

αpM(aα(T)eα)αpK(aα(T)eα)<,\sum_{\alpha}p_{M}(a_{\alpha}(T)e_{\alpha})\leq\sum_{\alpha}p_{K}(a_{\alpha}(T)e_{\alpha})<\infty,

using the estimate (3.43) which holds since KΩK\subset\Omega. ∎

3.11. Laurent series on non-annular Reinhardt domains

We now consider the case of a general (i.e. possibly non-annular) Reinhardt domain Ω\Omega. We begin by showing that the only monomials that occur in the Laurent series are the ones smooth on Ω:\Omega:

Proposition 3.12.

Let Ω\Omega be a (possibly non-annular) Reinhardt domain in n{\mathbb{C}}^{n}. Then for each α𝒮(Ω)\alpha\in\mathcal{S}(\Omega) (where 𝒮(Ω)\mathcal{S}(\Omega) is as in (3.11)), there is a continuous linear functional aα:𝒪(Ω)a_{\alpha}:\mathscr{O}(\Omega)\rightarrow{\mathbb{C}} such that the Fourier series of a T𝒪(Ω)T\in\mathscr{O}(\Omega) is of the form

Tα𝒮(Ω)aα(T)eα.T\sim\sum_{\alpha\in\mathcal{S}(\Omega)}a_{\alpha}(T)e_{\alpha}. (3.48)
Proof.

Since T𝒪(Ω)T\in\mathscr{O}(\Omega) we have T|ΩZ𝒪(ΩZ)T|_{\Omega\setminus Z}\in\mathscr{O}(\Omega\setminus Z). Notice that ΩZ\Omega\setminus Z is an annular Reinhardt domain and therefore the proof of Proposition 3.8 shows that the Fourier components are given for αn\alpha\in\mathbb{Z}^{n} by

(T|ΩZ)α=a^α(T|ΩZ)eα,\left(T|_{\Omega\setminus Z}\right)_{\alpha}=\widehat{a}_{\alpha}(T|_{\Omega\setminus Z})e_{\alpha},

where a^α:𝒪(ΩZ)\widehat{a}_{\alpha}:\mathscr{O}(\Omega\setminus Z)\to{\mathbb{C}} is the α\alpha-th coefficient functional associated to the domain ΩZ\Omega\setminus Z (see Proposition 3.8). Thanks to (3.35), we know that (T|ΩZ)α=(Tα)|ΩZ\left(T|_{\Omega\setminus Z}\right)_{\alpha}=(T_{\alpha})|_{\Omega\setminus Z}, so we have

(Tα)|ΩZ=a^α(T|ΩZ)eα.(T_{\alpha})|_{\Omega\setminus Z}=\widehat{a}_{\alpha}(T|_{\Omega\setminus Z})e_{\alpha}.

By Proposition 3.6, the Fourier components of a holomorphic distribution are holomorphic distributions, so we have Tα𝒪(Ω)T_{\alpha}\in\mathscr{O}(\Omega) since T𝒪(Ω)T\in\mathscr{O}(\Omega). Therefore, since by Corollary 3.9, each holomorphic distribution is a smooth function, we know that Tα𝒞(Ω)T_{\alpha}\in\mathcal{C}^{\infty}(\Omega). Therefore, the function Tα|ΩZ=a^α(T|ΩZ)eαT_{\alpha}|_{\Omega\setminus Z}=\widehat{a}_{\alpha}(T|_{\Omega\setminus Z})e_{\alpha} admits a 𝒞\mathcal{C}^{\infty} extension through ZZ. If a^α(T|ΩZ)0\widehat{a}_{\alpha}(T|_{\Omega\setminus Z})\not=0 for some T𝒪(Ω)T\in\mathscr{O}(\Omega), this means that eαe_{\alpha} itself admits a 𝒞\mathcal{C}^{\infty} extension to Ω\Omega, i.e., α𝒮(Ω)\alpha\in\mathcal{S}(\Omega), where 𝒮(Ω)\mathcal{S}(\Omega) is as in (3.11). Therefore if α𝒮(Ω)\alpha\not\in\mathcal{S}(\Omega), the corresponding term in the Laurent series of T|ΩZT|_{\Omega\setminus Z} vanishes, and the series takes the form

T|ΩZ=α𝒮(Ω)a^α(T|ΩZ)eα.T|_{\Omega\setminus Z}=\sum_{\alpha\in\mathcal{S}(\Omega)}\widehat{a}_{\alpha}(T|_{\Omega\setminus Z})e_{\alpha}.

Now for each α𝒮(Ω)\alpha\in\mathcal{S}(\Omega) define the map aα:𝒪(Ω)a_{\alpha}:\mathscr{O}(\Omega)\to{\mathbb{C}} by aα(T)=a^α(T|ΩZ).a_{\alpha}(T)=\widehat{a}_{\alpha}(T|_{\Omega\setminus Z}). Since both the restriction map and the coefficient functional a^α\widehat{a}_{\alpha} are continuous, it follows that aαa_{\alpha} is continuous. The extension of eαe_{\alpha} from ΩZ\Omega\setminus Z to Ω\Omega is still the monomial eαe_{\alpha}, so the Fourier series of TT in 𝒪(Ω)\mathscr{O}(\Omega) is of the form (3.48). ∎

Notice that each term of (3.48) is in 𝒞(Ω)\mathcal{C}^{\infty}(\Omega) and by Proposition 3.8, the series converges absolutely in 𝒞(ΩZ)\mathcal{C}^{\infty}(\Omega\setminus Z) when its terms are restricted to ΩZ\Omega\setminus Z, i.e., it converges uniformly along with all derivatives on those compact sets in Ω\Omega which do not intersect ZZ.

3.12. Extension to the relative completion

Given a Reinhardt domain DnD\subset{\mathbb{C}}^{n}, its relative completion is the smallest relatively complete domain containing DD (see above before the statement of Theorem 3.1 for the definition of relative completeness of a domain.) Notice that the relative completion of DD coincides with the unions of the sets D(j)D^{(j)} of (3.10), where the union is taken over those jj for which D{zj=0}D\cap\{z_{j}=0\} 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 DD be a Reinhardt domain. Then each holomorphic function on DD extends holomorphically to the relative completion of DD.

Proof.

We may assume that n2n\geq 2 since each Reinhardt domain in the plane is automatically relatively complete. If for each jj, the intersection D{zj=0}=D\cap\{z_{j}=0\}=\emptyset, then the domain DD is annular, and its relative completion is itself so there is nothing to prove. Suppose therefore that there is 1jn1\leq j\leq n such that D{zj=0}D\cap\{z_{j}=0\}\not=\emptyset. We need to prove that each function in 𝒪(D)\mathscr{O}(D) extends holomorphically to D(j)D^{(j)}.

Without loss of generality we can assume that j=1j=1. Write the coordinates of a point znz\in{\mathbb{C}}^{n} as z=(z1,z~)z=(z_{1},\widetilde{z}) where z~n1\widetilde{z}\in{\mathbb{C}}^{n-1}. Let f𝒪(D)f\in\mathscr{O}(D). By Proposition 3.12 ff admits a Laurent series representation

f=α𝒮(D)aα(f)eαf=\sum_{\alpha\in\mathcal{S}(D)}a_{\alpha}(f)e_{\alpha}

with S(D)S(D) as in (3.11), and the series converges absolutely in 𝒞(DZ)\mathcal{C}^{\infty}(D\setminus Z). To prove the proposition it suffices to show that the series in fact converges absolutely in 𝒞(D(1))\mathcal{C}^{\infty}(D^{(1)}), which by Lemma 3.7 is equivalent to the following: for each point pD(1)p\in D^{(1)} there is a compact neighborhood MM of pp in D(1)D^{(1)} such that

α𝒮(D)pM(aα(f)eα)<.\sum_{\alpha\in\mathcal{S}(D)}p_{M}(a_{\alpha}(f)e_{\alpha})<\infty. (3.49)

We claim the following: for each pD(1)p\in D^{(1)} there is a compact neighborhood MM of pp and a compact subset KDK\subset D such that pM(eα)pK(eα)p_{M}(e_{\alpha})\leq p_{K}(e_{\alpha}) for each α𝒮(D)\alpha\in\mathcal{S}(D). Assuming the claim for a moment, we see that we have

pM(aα(f)eα)=|aα(f)|pM(eα)|aα(f)|pK(eα)=pK(aα(f)eα),p_{M}(a_{\alpha}(f)e_{\alpha})=\left|a_{\alpha}(f)\right|p_{M}(e_{\alpha})\leq\left|a_{\alpha}(f)\right|p_{K}(e_{\alpha})=p_{K}(a_{\alpha}(f)e_{\alpha}),

so that (3.49) follows since by (3.43) we do have α𝒮(D)pK(aα(f)eα)<\sum_{\alpha\in\mathcal{S}(D)}p_{K}(a_{\alpha}(f)e_{\alpha})<\infty for a compact subset KK of DD. Since each point pD(1)p\in D^{(1)} has such a neighborhood this completes the proof, modulo the claim above.

To establish the claim, we may assume that pDp\not\in D, since otherwise there is nothing to prove. Therefore pD(1)Dp\in D^{(1)}\setminus D, and consequently, there is a zDz\in D and a λ𝔻\lambda\in\mathbb{D} such that p=(λz1,z~)p=(\lambda z_{1},\widetilde{z}), where z~=(z2,,zn)\widetilde{z}=(z_{2},\dots,z_{n}). Let KK be a compact neighborhood of zz in DD of the form K=K1×K~K=K_{1}\times\widetilde{K}, where K~n1\widetilde{K}\subset{\mathbb{C}}^{n-1} and K1={ζ:|ζz1|ϵ}K_{1}=\{\zeta\in{\mathbb{C}}:\left|\zeta-z_{1}\right|\leq\epsilon\} is a closed disk. Let L1L_{1} be the disk {ζ:|ζ||z1|+ϵ}\{\zeta\in{\mathbb{C}}:\left|\zeta\right|\leq\left|z_{1}\right|+\epsilon\}, so that K1L1K_{1}\subset L_{1}, and

γ=z1+z1|z1|ϵ\gamma={z_{1}}+\frac{z_{1}}{\left|z_{1}\right|}\epsilon

is a point of maximum modulus (i.e. maximum distance from the origin) in both sets K1K_{1} and L1L_{1}. Note that

|γ|=|z1|z1||z1|+z1|z1|ϵ|=|z1|+ϵ.\left|\gamma\right|=\left|\frac{z_{1}}{\left|z_{1}\right|}\cdot\left|z_{1}\right|+\frac{z_{1}}{\left|z_{1}\right|}\epsilon\right|=\left|z_{1}\right|+\epsilon.

We set M=L1×K~M=L_{1}\times\widetilde{K}. We will show that these sets K,MK,M satisfy the conditions of the claim.

Now let α𝒮(D)\alpha\in\mathcal{S}(D). Since D{z1=0}D\cap\{z_{1}=0\}\not=\emptyset, it follows that α10\alpha_{1}\geq 0. Let α~=(α2,,αn)n1\widetilde{\alpha}=(\alpha_{2},\dots,\alpha_{n})\in\mathbb{Z}^{n-1}, and set

B=supw~K~|w~α~|where w~=(w2,,wn)n1 and |w~α~|=|w2α2||wnαn|,B=\sup_{\widetilde{w}\in\widetilde{K}}\left|\widetilde{w}^{\widetilde{\alpha}}\right|\quad\text{where $\widetilde{w}=(w_{2},\dots,w_{n})\in{\mathbb{C}}^{n-1}$ and $\left|\widetilde{w}^{\widetilde{\alpha}}\right|=\left|w_{2}^{\alpha_{2}}\right|\cdots\left|w_{n}^{\alpha_{n}}\right|$},

so that we have

pM(eα)=supwM|eα(w)|=supw1L1|w1|α1B=(|z1|+ϵ)α1B.p_{M}(e_{\alpha})=\sup_{w\in M}\left|e_{\alpha}(w)\right|=\sup_{w_{1}\in L_{1}}\left|w_{1}\right|^{\alpha_{1}}\cdot B=(\left|z_{1}\right|+\epsilon)^{\alpha_{1}}\cdot B.

On the other hand

pK(eα)=supwK|eα(w)|=supw1K1|w1|α1B=(|z1|+ϵ)α1B.p_{K}(e_{\alpha})=\sup_{w\in K}\left|e_{\alpha}(w)\right|=\sup_{w_{1}\in K_{1}}\left|w_{1}\right|^{\alpha_{1}}\cdot B=(\left|z_{1}\right|+\epsilon)^{\alpha_{1}}\cdot B.

Consequently, in fact we have pM(eα)=pK(eα)p_{M}(e_{\alpha})=p_{K}(e_{\alpha}), 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 Ω\Omega is annular, i.e. Ω\Omega has empty intersection with the set ZZ of (3.7), then Proposition 3.8 takes care of the complete proof. When Ω\Omega 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 Ω\Omega. Now it is not difficult to see that the smallest log-convex relatively complete Reinhardt domain Ω^\widehat{\Omega} containing Ω\Omega can be constructed from Ω\Omega in two steps. First, we construct the log-convex hull Ω1\Omega_{1} of the set ΩZ\Omega\setminus Z. Notice that ΩZ\Omega\setminus Z and Ω1\Omega_{1} are both annular. The second step consists of constructing the relative completion of the domain Ω1\Omega_{1}, thus obtaining the domain Ω^\widehat{\Omega}. Now by Proposition 3.10, the Laurent series of a holomorphic distribution on Ω\Omega converges absolutely in 𝒞(Ω1)\mathcal{C}^{\infty}(\Omega_{1}). Applying Proposition 3.13 (with D=Ω1D=\Omega_{1}), we see that the Laurent series actually converges absolutely in the space 𝒞(Ω^)\mathcal{C}^{\infty}(\widehat{\Omega}). The sum of this series is the required holomorphic extension of a given holomorphic distribution on Ω\Omega. The result has been completely established.

4. Missing monomials

In Theorem 3.1, we considered the natural representation of the torus 𝕋n\mathbb{T}^{n} on the space 𝒪(Ω)\mathscr{O}(\Omega) of holomorphic functions on a Reinhardt domain Ω\Omega. In applications, one often deals with a subspace of functions Y𝒪(Ω)Y\subset\mathscr{O}(\Omega) such that

  1. (1)

    the subspace YY is invariant under the natural representation σ\sigma, i.e., if fYf\in Y then fRλYf\circ\mathrm{R}_{\lambda}\in Y for each λ𝕋n\lambda\in\mathbb{T}^{n}, where Rλ\mathrm{R}_{\lambda} is as in (3.6),

  2. (2)

    there is a locally convex topology on YY in which it is complete, and such that the inclusion map

    j:Y𝒪(Ω)j:Y\hookrightarrow\mathscr{O}(\Omega)

    is continuous, and

  3. (3)

    when YY is given this topology, the representation σ\sigma restricts to a continuous representation on YY.

The locus classicus here is the theory of Hardy spaces on the disc. We can make the following elementary observation:

Proposition 4.1.

Let YY and σ\sigma be as above, and set

𝒮(Y)={αn:eαY}.\mathcal{S}(Y)=\{\alpha\in\mathbb{Z}^{n}:e_{\alpha}\in Y\}.

Then the Laurent series of a function fYf\in Y is of the form

f=α𝒮(Y)aα(f)eα.f=\sum_{\alpha\in\mathcal{S}(Y)}a_{\alpha}(f)e_{\alpha}.
Proof.

It suffices to show that if for αn\alpha\in\mathbb{Z}^{n}, if the monomial eαe_{\alpha} does not belong to YY, we have that aα(f)=0a_{\alpha}(f)=0, where aα(f)a_{\alpha}(f) is the Laurent coefficient of ff as in (3.12). By part (1) of Proposition 2.4, we see that 𝝅ασ(f)Y\bm{\pi}^{\sigma}_{\alpha}(f)\in Y. By part (2) of the same proposition, taking XX to be the space 𝒪(Ω)\mathscr{O}(\Omega), for fYf\in Y we have that 𝝅ασ(j(f))=j(𝝅ασ(f))\bm{\pi}^{\sigma}_{\alpha}(j(f))=j(\bm{\pi}^{\sigma}_{\alpha}(f)), and from the description of the Fourier components of a holomorphic function in the proof of Theorem 3.1, we see that 𝝅ασ(j(f))=aα(j(f))eα=aα(f)eα\bm{\pi}^{\sigma}_{\alpha}(j(f))=a_{\alpha}(j(f))e_{\alpha}=a_{\alpha}(f)e_{\alpha}. Therefore j(𝝅ασ(f))=aα(f)eαj(\bm{\pi}^{\sigma}_{\alpha}(f))=a_{\alpha}(f)e_{\alpha}, which contradicts the fact that 𝝅ασ(f)Y\bm{\pi}^{\sigma}_{\alpha}(f)\in Y unless aα(f)=0a_{\alpha}(f)=0. ∎

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 ff. It can be thought to be the reason behind several phenomena associated to holomorphic functions. We consider two examples:

  1. (1)

    Bergman spaces in Reinhardt domains: Let Ω\Omega be a Reinhardt domain in n{\mathbb{C}}^{n} and let λ>0\lambda>0 be a radial weight on Ω\Omega, i.e., for zΩz\in\Omega we have

    λ(z1,,zn)=λ(|z1|,,|zn|).\lambda(z_{1},\dots,z_{n})=\lambda(\left|z_{1}\right|,\dots,\left|z_{n}\right|).

    The LpL^{p}-Bergman space Ap(Ω,λ)A^{p}(\Omega,\lambda) is defined to be the subspace of the weighted LpL^{p}-space Lp(Ω,λ)L^{p}(\Omega,\lambda) consisting of holomorphic functions, where the norm on the weighted LpL^{p}-space is given by

    fLp(Ω,λ)p=Ω|f|pλ𝑑V,\left\|f\right\|_{L^{p}(\Omega,\lambda)}^{p}=\int_{\Omega}\left|f\right|^{p}\lambda dV,

    where dVdV is the Lebesgue measure. It is well-known that Ap(Ω,λ)A^{p}(\Omega,\lambda) is a closed subspace of the Banach space Lp(Ω,λ)L^{p}(\Omega,\lambda) and therefore a Banach space ([DS04]). It is also easy to see (using standard facts about LpL^{p}-spaces) that the natural representation σ\sigma of 𝕋n\mathbb{T}^{n} on Lp(Ω,λ)L^{p}(\Omega,\lambda) is continuous for 1p<1\leq p<\infty, so it follows that the representation on Ap(Ω,λ)A^{p}(\Omega,\lambda) is also continuous. It now follows from Proposition 4.1 that the Laurent series expansion of a function fAp(Ω,λ)f\in A^{p}(\Omega,\lambda) consists only of terms with monomials eαe_{\alpha} such that eαLp(Ω,λ)e_{\alpha}\in L^{p}(\Omega,\lambda). The case λ1\lambda\equiv 1 of this fact was deduced by a different argument in [CEM18].

  2. (2)

    Extension of holomorphic functions smooth up to the boundary: Let Ωn\Omega\subset{\mathbb{C}}^{n} be a Reinhardt domain such that the origin (which is the center of symmetry) is on the boundary of Ω\Omega. A classic example of this is the Hartogs triangle {|z1|<|z2|<1}\{\left|z_{1}\right|<\left|z_{2}\right|<1\} in 2{\mathbb{C}}^{2}. In [Cha19], the following extension theorem was proved: there is a complete Reinhardt domain VnV\subset{\mathbb{C}}^{n} such that ΩV\Omega\subset V and each function in the space 𝒪(Ω)𝒞(Ω¯)\mathcal{O}(\Omega)\cap\mathcal{C}^{\infty}(\overline{\Omega}) of holomorphic functions on Ω\Omega smooth up to the boundary extends holomorphically to the domain VV. This was noted for the Hartogs triangle in [Sib75], where VV is the unit bidisk.

    To deduce this from the principle of missing monomials, it suffices to consider the case when Ω\Omega is bounded. We notice that 𝒪(Ω)𝒞(Ω¯)\mathcal{O}(\Omega)\cap\mathcal{C}^{\infty}(\overline{\Omega}) is a Fréchet space in its usual topology of uniform convergence on Ω¯\overline{\Omega} with all partial derivatives. A generating family of seminorms for this topology is given by the norms {pk}\{p_{k}\} where pk(f)=fCk(Ω¯)p_{k}(f)=\left\|f\right\|_{C^{k}(\overline{\Omega})}. The natural representation σ\sigma on 𝒪(Ω)\mathscr{O}(\Omega) restricts to a continuous representation on 𝒪(Ω)𝒞(Ω¯)\mathcal{O}(\Omega)\cap\mathcal{C}^{\infty}(\overline{\Omega}). Therefore, the principle of missing monomials applies and the only monomials eαe_{\alpha} that occur in the Laurent expansion of a function f𝒪(Ω)𝒞(Ω¯)f\in\mathcal{O}(\Omega)\cap\mathcal{C}^{\infty}(\overline{\Omega}) are such that eα𝒪(Ω)𝒞(Ω¯)e_{\alpha}\in\mathcal{O}(\Omega)\cap\mathcal{C}^{\infty}(\overline{\Omega}). For such a multi-index α\alpha, write the multiindex α=βγ\alpha=\beta-\gamma, where βj=max(αj,0)\beta_{j}=\max(\alpha_{j},0) and γj=(αj,0)\gamma_{j}=(-\alpha_{j},0). Then β,γn\beta,\gamma\in\mathbb{N}^{n} , and we can apply the differential operator (z)β\left(\frac{\partial}{\partial z}\right)^{\beta} to obtain

    (z)βeα(z)=(z)βeβ(z)eα(z)=β!zγ.\left(\frac{\partial}{\partial z}\right)^{\beta}e_{\alpha}(z)=\left(\frac{\partial}{\partial z}\right)^{\beta}\frac{e_{\beta}(z)}{e_{\alpha}(z)}=\frac{\beta!}{z^{\gamma}}.

    Since eα𝒞(Ω¯)e_{\alpha}\in\mathcal{C}^{\infty}(\overline{\Omega}) this means that eγ𝒞(Ω¯)e_{\gamma}\in\mathcal{C}^{\infty}(\overline{\Omega}), which is possible only if γ=0\gamma=0. Thus αn\alpha\in\mathbb{N}^{n}, and the Laurent series of each function f𝒪(Ω)𝒞(Ω¯)f\in\mathcal{O}(\Omega)\cap\mathcal{C}^{\infty}(\overline{\Omega}) is a Taylor series which converges in some complete (log-convex) Reinhardt domain VV, and this VV must strictly contain Ω\Omega, since Ω\Omega 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 f:Ωf:\Omega\to{\mathbb{C}} on an open subset Ω\Omega\subset{\mathbb{C}} is holomorphic, if it is complex differentiable, i.e., for each point wΩw\in\Omega, the limit

limzwf(z)f(w)zw{\lim_{z\to w}\frac{f(z)-f(w)}{z-w}} (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 G(Ω)G(\Omega) the collection of holomorphic functions in the sense of Goursat in an open set Ω\Omega\subset{\mathbb{C}}, we notice that the space G(Ω)G(\Omega) does not have a nice a priori linear locally convex topology in which it is complete and such that when Ω\Omega is a disc or an annulus, the natural action of the group 𝕋\mathbb{T} on the space G(Ω)G(\Omega) 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 Ω\Omega be an open subset of the complex plane {\mathbb{C}}, and let f:Ωf:\Omega\to{\mathbb{C}} be a continuous function. In honor of [Mor02], let us say that the function ff is holomorphic in the sense of Morera (Morera-holomorphic for short) if for each triangle TT contained (with its interior) in Ω\Omega, we have the vanishing of the complex line integral of ff around the boundary of TT:

Tf(z)𝑑z=0,\int_{\partial T}f(z)dz=0, (5.2)

where T\partial T denotes the boundary of TT, oriented counterclockwise. For an open set Ω\Omega\subset{\mathbb{C}}, let us denote by 𝒪M(Ω)\mathscr{O}_{M}(\Omega) the collection of Morera-holomorphic functions on Ω\Omega. 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 ww.)

It is immediate from the definition that 𝒪M(Ω)\mathscr{O}_{M}(\Omega) is a closed linear subspace of the Fréchet space 𝒞(Ω)\mathcal{C}(\Omega) of continuous functions. “Closed” means that the limit of a sequence of Morera-holomorphic functions converging uniformly on compact subsets of Ω\Omega 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., f𝒪M(Ω)f\in\mathscr{O}_{M}(\Omega) if and only if there is an open cover {Ωj}jJ\{\Omega_{j}\}_{j\in J} of Ω\Omega such that f|Ωj𝒪M(Ωj)f|_{\Omega_{j}}\in\mathscr{O}_{M}(\Omega_{j}) for each jj. One half of this claim is trivial, and for the other half, for a triangle TT in Ω\Omega, we can perform repeated barycentric subdivisions till the triangles so formed are each contained in some element of the open cover {Ωj}jJ\{\Omega_{j}\}_{j\in J}. We therefore conclude that 𝒪M\mathscr{O}_{M} is a sheaf of Fréchet spaces on {\mathbb{C}}.

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 Ω\Omega\subset{\mathbb{C}} be convex. Then the following statements about a continuous function f𝒞(Ω)f\in\mathcal{C}(\Omega) are equivalent:

  1. (A)

    f𝒪M(Ω)f\in\mathscr{O}_{M}(\Omega).

  2. (B)

    ff has a holomorphic primitive, i.e., there is an FF which is complex-differentiable on Ω\Omega and F=fF^{\prime}=f.

  3. (C)

    for each piecewise 𝒞1\mathcal{C}^{1} closed path γ\gamma in Ω\Omega we have

    γf(z)𝑑z=0.\int_{\gamma}f(z)dz=0. (5.3)

Recall that assuming (A), the primitive FF in (B) is constructed by fixing z0Ωz_{0}\in\Omega, and setting F(z)=[z0,z]f(ζ)𝑑ζ\displaystyle{F(z)=\int_{[z_{0},z]}f(\zeta)d\zeta} where [z0,z][z_{0},z] denotes the line segment from z0z_{0} to zz. Proposition 5.1 allows us to give examples of Morera-holomorphic functions. Recall from (3.9) that for an integer nn, we use the notation en(z)=zne_{n}(z)=z^{n} for the holomorphic monomials.

Proposition 5.2.

If n0n\geq 0 then en𝒪M()e_{n}\in\mathscr{O}_{M}({\mathbb{C}}) and if n<0n<0 then en𝒪M({0})e_{n}\in\mathscr{O}_{M}({\mathbb{C}}\setminus\{0\}).

Proof.

First note that ene_{n} is continuous, on all of {\mathbb{C}} if n0n\geq 0 and on {0}{\mathbb{C}}\setminus\{0\} if n<0n<0. If gn(z)=znn+1g_{n}(z)=\frac{z^{n}}{n+1} for n1n\not=-1, we can verify from the definition (5.1) that gng_{n} is complex-differentiable and gn=eng_{n}^{\prime}=e_{n} so that by part (B) of Proposition 5.1 the result follows for n1n\not=-1. For n=1n=-1, we can construct for each p{0}p\in{\mathbb{C}}\setminus\{0\}, a local primitive of e1e_{-1} near pp by setting

g1(z)=ln|z|+iargz,g_{-1}(z)=\ln\left|z\right|+i\arg z,

where arg\arg denotes a branch of the argument defined near the point pp. A direct computation shows that g1=e1g_{-1}^{\prime}=e_{-1} near pp, so that again we see that e1𝒪M({0}).e_{-1}\in\mathscr{O}_{M}({\mathbb{C}}\setminus\{0\}).

5.3. Products of Morera-holomorphic functions

In the proof of Theorem 3.1, an important role is played by the fact that if UU is a holomorphic distribution (in the sense of (3.5))and ff is a holomorphic function (i.e. a holomorphic distribution which is 𝒞\mathcal{C}^{\infty}, Section 3.3 above), then the product distribution fUfU 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 f,g𝒪M(Ω)f,g\in\mathscr{O}_{M}(\Omega), and assume that gg is locally Lipschitz at each point, i.e. for each wΩw\in\Omega and each compact KΩK\subset\Omega such that wKw\in K, there is an M>0M>0 such that for zKz\in K we have

|g(z)g(w)|M|zw|.\left|g(z)-g(w)\right|\leq M\left|z-w\right|. (5.4)

Then the product fgfg also belongs to 𝒪M(Ω)\mathscr{O}_{M}(\Omega).

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 Ω\Omega be an open subset of {\mathbb{C}} and let λ\lambda be a complex valued function defined on the set of triangles contained in Ω\Omega such that the following two conditions are satisfied:

  1. (1)

    λ\lambda is additive in the following sense: if a triangle Δ\Delta is represented as a union of smaller triangles Δ=k=1nΔk\Delta=\bigcup_{k=1}^{n}\Delta_{k} with pairwise disjoint interiors then

    λ(Δ)=k=1nλ(Δk).\lambda(\Delta)=\sum_{k=1}^{n}\lambda(\Delta_{k}). (5.5)
  2. (2)

    For each wΩw\in\Omega and each triangle Δ0\Delta_{0}, we have

    limΔwΔΔ0λ(Δ)|Δ|=0,\lim_{\begin{subarray}{c}\Delta\downarrow w\\ \Delta\sim\Delta_{0}\end{subarray}}\frac{\lambda(\Delta)}{\left|\Delta\right|}=0, (5.6)

    where |Δ|\left|\Delta\right| denotes the area of the triangle Δ\Delta, and the limit is taken along the family of triangles similar to Δ0\Delta_{0} and containing the point ww, as these triangles shrink to the point ww.

Then λ0\lambda\equiv 0.

Proof.

For completeness, we recall the classic argument. Let TT be a triangle contained in Ω\Omega. We construct a sequence of triangles {Tk}k=0\{T_{k}\}_{k=0}^{\infty} with T0=TT_{0}=T using the following recursive procedure. Assuming that TkT_{k} has been constructed, we divide TkT_{k} into four similar triangles with half the diameter of TkT_{k} by three line segments each parallel to a side of TkT_{k} and passing through the midpoints of the other two sides. Denote the four triangles so obtained by Δj,1j4\Delta_{j},1\leq j\leq 4. Then, by (5.5), we have

λ(Tk)=j=14λ(Δj).\lambda(T_{k})=\sum_{j=1}^{4}\lambda(\Delta_{j}).

Choose Tk+1T_{k+1} to be one of Δj,1j4\Delta_{j},1\leq j\leq 4 such that the value of |λ(Tk+1)|\left|\lambda(T_{k+1})\right| is the largest. Then, by the triangle inequality we have |λ(Tk)|4|λ(Tk+1)|,\left|\lambda(T_{k})\right|\leq 4\left|\lambda(T_{k+1})\right|, and by induction it follows that

|λ(T)|4k|λ(Tk)|=4k|Tk||λ(Tk)||Tk|=|T||λ(Tk)||Tk|,\left|\lambda(T)\right|\leq 4^{k}\left|\lambda(T_{k})\right|=4^{k}\left|T_{k}\right|\cdot\frac{\left|\lambda(T_{k})\right|}{\left|T_{k}\right|}=\left|T\right|\cdot\frac{\left|\lambda(T_{k})\right|}{\left|T_{k}\right|}, (5.7)

where in the last step we have used the fact that Tk+1T_{k+1} has one-fourth the area of TkT_{k}, so |Tk|=4k|T0|=4k|T|\left|T_{k}\right|=4^{-k}\left|T_{0}\right|=4^{-k}\left|T\right|. Since the diameters of the TkT_{k} go to zero, by compactness, there is a unique point ww in the intersection k=0Tk\bigcap_{k=0}^{\infty}T_{k}. Since the family {Tk}\{T_{k}\} is a subfamily of all the triangles containing ww, and each TkT_{k} is similar to T0T_{0}, therefore by letting kk\to\infty in (5.7) and using (5.6) the result follows. ∎

Proof of Proposition 5.3.

For a triangle ΔΩ\Delta\subset\Omega, define

λ(Δ)=Δf(z)g(z)𝑑z.\lambda(\Delta)=\int_{\partial\Delta}f(z)g(z)dz.

To prove the result, we need to show that λ0\lambda\equiv 0. Since condition (5.5) of Lemma 5.4 is obvious, we need to show (5.6) to complete the proof. Let wΩw\in\Omega, let KK be a compact neighborhood of ww in Ω\Omega, and denote by MM the Lipschitz constant corresponding to this ww and this KK in (5.4). Now, let Δ0\Delta_{0} be a triangle and let Δ\Delta be a triangle similar to Δ0\Delta_{0} such that wΔKw\in\Delta\subset K. Then observe that, by the hypothesis of Morera-holomorphicity of ff and gg we have

λ(Δ)=Δ(f(z)f(w))(g(z)g(w))𝑑z.\lambda(\Delta)=\int_{\partial\Delta}\left(f(z)-f(w)\right)\left(g(z)-g(w)\right)dz.

Therefore, denoting the perimeter of the triangle Δ\Delta by |Δ|\left|\partial\Delta\right|,

|λ(Δ)|\displaystyle\left|\lambda(\Delta)\right| supzΔ|(f(z)f(w))(g(z)g(w))||Δ|\displaystyle\leq\sup_{z\in\partial\Delta}\left|\left(f(z)-f(w)\right)\left(g(z)-g(w)\right)\right|\left|\partial\Delta\right|
M|Δ|supzΔ(|f(z)f(w)||zw|)\displaystyle\leq M\cdot\left|\partial\Delta\right|\cdot\sup_{z\in\partial\Delta}\left(\left|f(z)-f(w)\right|\left|z-w\right|\right)
M|Δ|diam(Δ)supzΔ|f(z)f(w)|\displaystyle\leq M\cdot\left|\partial\Delta\right|\cdot\mathrm{diam}(\Delta)\cdot\sup_{z\in\partial\Delta}\left|f(z)-f(w)\right|
=M|Δ|(|Δ|diam(Δ)|Δ|)supzΔ|f(z)f(w)|\displaystyle=M\cdot\left|\Delta\right|\cdot\left(\frac{\left|\partial\Delta\right|\cdot\mathrm{diam}(\Delta)}{\left|\Delta\right|}\right)\cdot\sup_{z\in\partial\Delta}\left|f(z)-f(w)\right|
=M|Δ|(|Δ0|diam(Δ0)|Δ0|)supzΔ|f(z)f(w)|,\displaystyle=M\cdot\left|\Delta\right|\cdot\left(\frac{\left|\partial\Delta_{0}\right|\cdot\mathrm{diam}(\Delta_{0})}{\left|\Delta_{0}\right|}\right)\cdot\sup_{z\in\partial\Delta}\left|f(z)-f(w)\right|,

where in the last step we use the fact that Δ\Delta and Δ0\Delta_{0} are similar, so the quantity in parentheses (which is clearly invariant under dilations) is the same. So we have

|λ(Δ)||Δ|CsupzΔ|f(z)f(w)|,\frac{\left|\lambda(\Delta)\right|}{\left|\Delta\right|}\leq C\cdot\sup_{z\in\partial\Delta}\left|f(z)-f(w)\right|,

for a constant CC independent of the triangle Δ\Delta as long as Δ\Delta is similar to Δ0\Delta_{0} and wΔKw\in\Delta\subset K. Letting Δ\Delta shrink to ww, 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 f𝒪M(𝔻)f\in\mathscr{O}_{M}(\mathbb{D}) where 𝔻={|z|<1}\mathbb{D}=\{\left|z\right|<1\} is the open unit disc. Then there is a sequence {an}n=0\{a_{n}\}_{n=0}^{\infty} of complex numbers, such that

f=n=0anen,f=\sum_{n=0}^{\infty}a_{n}e_{n}, (5.8)

where ene_{n} is as in (3.9), and the series on the right converges absolutely in 𝒞(𝔻)\mathcal{C}(\mathbb{D}) to the function ff.

Let σ\sigma be the natural representation of 𝕋\mathbb{T} on 𝒞(𝔻)\mathcal{C}(\mathbb{D}) given by

σλ(f)(z)=f(λz),λ𝕋,z𝔻.\sigma_{\lambda}(f)(z)=f(\lambda z),\quad\lambda\in\mathbb{T},z\in\mathbb{D}. (5.9)
Proposition 5.5.

The space 𝒪M(𝔻)\mathscr{O}_{M}(\mathbb{D}) is invariant under σ\sigma and the resulting representation of 𝕋\mathbb{T} on 𝒪M(𝔻)\mathscr{O}_{M}(\mathbb{D}) is continuous.

Proof.

Let λ𝕋\lambda\in\mathbb{T}, let TT be a triangle in 𝔻\mathbb{D}. Notice that

λ1T={λ1z:zT}\lambda^{-1}T=\{\lambda^{-1}z:z\in T\}

is itself a triangle, and we have

T(σλf)(z)𝑑z=Tf(λz)𝑑z=λ1(λ1T)f(w)𝑑w=0.\int_{\partial T}(\sigma_{\lambda}f)(z)dz=\int_{\partial T}f(\lambda z)dz=\lambda^{-1}\int_{\partial(\lambda^{-1}T)}f(w)dw=0.

It follows that 𝒪M(𝔻)\mathscr{O}_{M}(\mathbb{D}) is invariant under σ\sigma.

It suffices to show that the representation σ\sigma is continuous on 𝒞(𝔻)\mathcal{C}(\mathbb{D}). For 0<r<10<r<1, let prp_{r} be the seminorm on 𝒞(𝔻)\mathcal{C}(\mathbb{D}) given by

pr(f)=sup|z|r|f(z)|.p_{r}(f)=\sup_{\left|z\right|\leq r}\left|f(z)\right|. (5.10)

It is clear that pr(σλ(f))=pr(f)p_{r}(\sigma_{\lambda}(f))=p_{r}(f) for each 0<r<1,λ𝕋0<r<1,\lambda\in\mathbb{T} and f𝒞(𝔻)f\in\mathcal{C}(\mathbb{D}). Also fjff_{j}\to f in 𝒞(𝔻)\mathcal{C}(\mathbb{D}) if and only if for each rr, we have pr(ffj)0p_{r}(f-f_{j})\to 0, so the family {pr}\{p_{r}\} is a σ\sigma-invariant family of seminorms that generates the topology of 𝒞(𝔻)\mathcal{C}(\mathbb{D}). Further, given f𝒞(𝔻)f\in\mathcal{C}(\mathbb{D}), by uniform continuity, for each 0<r<10<r<1, we have

pr(σλ(f)f)=sup|z|r|f(λz)f(z)|0 as λ1,p_{r}(\sigma_{\lambda}(f)-f)=\sup_{\left|z\right|\leq r}\left|f(\lambda z)-f(z)\right|\to 0\quad\text{ as }\lambda\to 1,

so that limλ1σλ(f)=f\lim_{\lambda\to 1}\sigma_{\lambda}(f)=f in the space 𝒞(𝔻)\mathcal{C}(\mathbb{D}). Therefore both conditions of Proposition 2.3 are satisfied, and the representation σ\sigma 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 f𝒪M(𝔻)f\in\mathscr{O}_{M}(\mathbb{D}), and with σ\sigma the natural representation (5.9), the Fourier components of ff are of the form:

𝝅nσ(f)={anen if n00 if n<0,\bm{\pi}^{\sigma}_{n}(f)=\begin{cases}a_{n}e_{n}&\text{ if }n\geq 0\\ 0&\text{ if }n<0,\end{cases}

where ana_{n}\in{\mathbb{C}} and en(z)=zne_{n}(z)=z^{n}.

The proof will use the following lemma:

Lemma 5.7.

A radial function in 𝒪M(𝔻{0})\mathscr{O}_{M}(\mathbb{D}\setminus\{0\}) is constant.

Proof.

Let f𝒪M(𝔻{0})f\in\mathscr{O}_{M}(\mathbb{D}\setminus\{0\}) be radial, and define the complex valued continuous function uu on the interval (0,1)(0,1) by restriction, u(r)=f(r)u(r)=f(r), so that we have f(z)=u(|z|)f(z)=u(\left|z\right|) by the radiality of ff. To prove the theorem, it suffices to show that uu is a constant .

Fix 0α<β<π0\leq\alpha<\beta<\pi and 0<ρ<10<\rho<1. For RR in the interval (ρ,1)(\rho,1) consider the curvilinear quadrilateral defined by

S(R)={reiθ:ρrR,αθβ},S(R)=\{re^{i\theta}:\rho\leq r\leq R,\alpha\leq\theta\leq\beta\}, (5.11)

and notice that S(R)S(R) lies in the upper half disc, which is convex. The region S(R)S(R) is bounded by the two circular arcs

AB={ρeiθ:αθβ},CD={Reiθ:αθβ},AB=\{\rho e^{i\theta}:\alpha\leq\theta\leq\beta\},\quad CD=\{Re^{i\theta}:\alpha\leq\theta\leq\beta\},

along with the two radial line segments

AD={reiα:ρrR},BC={reiβ:ρrR}.AD=\{re^{i\alpha}:\rho\leq r\leq R\},\quad BC=\{re^{i\beta}:\rho\leq r\leq R\}.
AADDCCBBS(R)S(R)

Orient the boundary S(R)\partial S(R) counterclockwise. We can write

S(R)f(z)𝑑z=AD+DC+CB+BA=0,\int_{\partial S(R)}f(z)dz=\int_{AD}+\int_{DC}+\int_{CB}+\int_{BA}=0,

where the vanishing of the integral follows from part (c) of Proposition 5.1 above. Parametrizing ADAD by z=reiαz=re^{i\alpha} where ρrR\rho\leq r\leq R , and using dz=eiαdrdz=e^{i\alpha}dr we get

ADf(z)𝑑z=ρRf(reiα)eiα𝑑r=eiαρRu(r)𝑑r.\int_{AD}f(z)dz=\int_{\rho}^{R}f(re^{i\alpha})e^{i\alpha}dr=e^{i\alpha}\int_{\rho}^{R}u(r)dr.

Similarly,

CBf(z)𝑑z=eiβρRu(r)𝑑r.\int_{CB}f(z)dz=-e^{i\beta}\int_{\rho}^{R}u(r)dr.

Now, parametrizing DCDC by z=Reiθ,αθβz=Re^{i\theta},\alpha\leq\theta\leq\beta, we have dz=Rieiθdθdz=Rie^{i\theta}d\theta, so

DCf(z)𝑑z=θ=αβf(Reiθ)Rieiθ𝑑θ=Ru(R)αβieiθ𝑑θ=Ru(R)(eiβeiα).\int_{DC}f(z)dz=\int_{\theta=\alpha}^{\beta}f(Re^{i\theta})\cdot Rie^{i\theta}d\theta=Ru({R})\int_{\alpha}^{\beta}ie^{i\theta}d\theta=Ru(R)(e^{i\beta}-e^{i\alpha}).

Similarly,

BAf(z)𝑑z=ρu(ρ)(eiβeiα).\int_{BA}f(z)dz=\rho u(\rho)(e^{i\beta}-e^{i\alpha}).

Therefore, adding the four integrals we have

Sf(z)𝑑z=(ρu(ρ)Ru(R)+ρRu(r)𝑑r)(eiαeiβ)=0,\int_{\partial S}f(z)dz=\left({\rho}u({\rho})-Ru(R)+\int_{\rho}^{R}u(r)dr\right)(e^{i\alpha}-e^{i\beta})=0,

so that we have the relation

ρRu(r)𝑑r=Ru(R)ρu(ρ),\int_{\rho}^{R}u(r)dr=Ru(R)-{\rho}u({\rho}), (5.12)

which can be written as

u(R)=1R(ρRu(r)𝑑r+ρu(ρ)).u(R)=\frac{1}{R}\left(\int_{\rho}^{R}u(r)dr+{\rho}u({\rho})\right).

Since uu is continuous, the right hand side is a function of RR which is continuously differentiable on (ρ,1)(\rho,1). Thus u𝒞1(ρ,1)u\in\mathcal{C}^{1}(\rho,1). Differentiating both sides of (5.12) with respect to RR we have u(R)=Ru(R)+u(R)u(R)=Ru^{\prime}(R)+u(R) so that we have u(R)=0u^{\prime}(R)=0 for R(ρ,1)R\in(\rho,1). But 0<ρ<10<\rho<1 can be chosen arbitrarily, so it follows that u0u^{\prime}\equiv 0 on (0,1)(0,1) and the result follows. ∎

Proof of Proposition 5.6.

Let nn\in\mathbb{Z} and for simplicity of notation, let fn=𝝅nσ(f)f_{n}=\bm{\pi}^{\sigma}_{n}(f). Then by Proposition 2.4, we see that fnf_{n} lies in the Fourier mode [𝒪M(𝔻)]nσ[\mathscr{O}_{M}(\mathbb{D})]_{n}^{\sigma} so that fn𝒪M(𝔻)f_{n}\in\mathscr{O}_{M}(\mathbb{D}) and for λT,z𝔻\lambda\in T,z\in\mathbb{D} we have fn(λz)=λnfn(z)f_{n}(\lambda z)=\lambda^{n}f_{n}(z). Since clearly fn|𝔻{0}𝒪M(𝔻{0})f_{n}|_{\mathbb{D}\setminus\{0\}}\in\mathscr{O}_{M}(\mathbb{D}\setminus\{0\}), it follows by Proposition 5.3 that the function hnh_{n} on 𝔻{0}\mathbb{D}\setminus\{0\} defined by

hn=enfnh_{n}=e_{-n}\cdot f_{n}

lies in 𝒪M(𝔻{0})\mathscr{O}_{M}(\mathbb{D}\setminus\{0\}). Further for λT,z𝔻\lambda\in T,z\in\mathbb{D} we have

hn(λz)=en(λz)fn(λz)=λnen(z)λnfn(z)=hn(z),h_{n}(\lambda z)=e_{n}(\lambda z)f_{n}(\lambda z)=\lambda^{-n}e_{-n}(z)\cdot\lambda^{n}f_{n}(z)=h_{n}(z),

so hnh_{n} is radial, and hence by Lemma 5.7, hnh_{n} is a constant, which we call ana_{n}. Therefore, on 𝔻{0}\mathbb{D}\setminus\{0\}, we have fn=anenf_{n}=a_{n}e_{n}, so the product anena_{n}e_{n} extends for each nn to a continuous function on 𝔻\mathbb{D}. If n<0n<0, this is possible only if an=0a_{n}=0, and this completes the proof. ∎

5.5. Conclusion of the proof of Theorem 5.1

We will first show that the series n=0anen\sum_{n=0}^{\infty}a_{n}e_{n} converges absolutely in the space 𝒞(𝔻)\mathcal{C}(\mathbb{D}). It suffices to show that for each 0<r<10<r<1, we have

n=0pr(anen)<,\sum_{n=0}^{\infty}p_{r}(a_{n}e_{n})<\infty,

where prp_{r} is as in (5.10). Fix an rr with 0<r<10<r<1 and let ρ\rho be such that r<ρ<1r<\rho<1. Applying the Cauchy estimate (2.9) to the seminorm pρp_{\rho} we see that for each nn\in\mathbb{Z}:

pρ(𝝅nσf)pρ(f)p_{\rho}(\bm{\pi}^{\sigma}_{n}f)\leq p_{\rho}(f)

Thanks to Proposition 5.6 we have for each n0n\geq 0,

pr(𝝅nσf)=sup|z|r|anzn|=|an|rn(rρ)n|an|ρn(rρ)npρ(f).p_{r}(\bm{\pi}^{\sigma}_{n}f)=\sup_{\left|z\right|\leq r}\left|a_{n}z^{n}\right|=\left|a_{n}\right|r^{n}\leq\left(\frac{r}{\rho}\right)^{n}\cdot\left|a_{n}\right|\rho^{n}\leq\left(\frac{r}{\rho}\right)^{n}\cdot p_{\rho}(f).

Therefore we have

n=pr(𝝅nσf)=n=0pr(𝝅nσf)(n=0(rρ)n)pρ(f)=11rρpρ(f)<.\sum_{n=-\infty}^{\infty}{p_{r}(\bm{\pi}^{\sigma}_{n}f)}=\sum_{n=0}^{\infty}{p_{r}(\bm{\pi}^{\sigma}_{n}f)}\leq\left(\sum_{n=0}^{\infty}\left(\frac{r}{\rho}\right)^{n}\right)\cdot p_{\rho}(f)=\frac{1}{1-\frac{r}{\rho}}\cdot p_{\rho}(f)<\infty.

This proves that the series n=0anen\sum_{n=0}^{\infty}a_{n}e_{n} converges absolutely in 𝒞(𝔻)\mathcal{C}(\mathbb{D}). Let gg be its sum. Since the partial sums are all in 𝒪M(𝔻)\mathscr{O}_{M}(\mathbb{D}) by Proposition 5.2, we see that g𝒪M(𝔻)g\in\mathscr{O}_{M}(\mathbb{D}). However, it follows from Corollary 2.5 that the sum of the series is ff, therefore f=gf=g, 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 ϵ\epsilon is small and nonzero, all it says about the functions f(z)=z+ϵz¯f(z)=z+\epsilon\overline{z} and g(z)=z¯+ϵzg(z)=\overline{z}+\epsilon{z} is that both are non-holomorphic. Morera’s definition of a holomorphic function does not have this shortcoming, as the integral Tf(z)𝑑z\int_{\partial T}f(z)dz gives a measure of the amount of non-holomorphicity of ff on the triangle TT.

It is possible to normalize and localize this measure of non-holomorphicity, as was realized by Pompeiu (see [Pom12, MŞ98]). The quantity

1|T|Tf(z)𝑑z\frac{1}{\left|T\right|}\int_{\partial T}f(z)dz (5.13)

is a numerical measure of the “average density of non-holomorphicity” of a continuous function ff on a triangle (or other region with piecewise smooth boundary) TT, where |T|\left|T\right| denotes the area of TT. To localize this, we can consider, for a point wΩw\in\Omega, and a continuous function f:Ωf:\Omega\to{\mathbb{C}}, the following limit (called the “areolar derivative” by Pompieu) as a measure of the degree of non-holomorphicity of ff at the point ww:

limTw1|T|Tf(z)𝑑z\lim_{T\downarrow w}\frac{1}{\left|T\right|}\int_{\partial T}f(z)dz (5.14)

where the limit is taken over the family of triangles containing the point ww and contained in Ω\Omega, as these triangles shrink to the point ww. Let us say that a function is Pompeiu-holomorphic if the limit (5.14) exists at each wΩw\in\Omega, and is equal to zero. The following simple and well-known observations clarify the meaning of this notion:

Proposition 5.8.
  1. (a)

    A continuous function f:Ωf:\Omega\to{\mathbb{C}} is Pompieu-holomorphic if and only if it is Morera-holomorphic.

  2. (b)

    If the function ff is in 𝒞1(Ω)\mathcal{C}^{1}(\Omega), then the limit (5.14) exists, and is equal to 2ifz¯(w)\displaystyle{2i\frac{\partial f}{\partial\overline{z}}(w)}.

Proof.

For part (a), if ff is Morera-holomorphic, then the quantity (5.13) vanishes for each triangle TT, so the limit (5.14) vanishes. Conversely, suppose that the limit (5.14) vanishes at each point wΩw\in\Omega, and let TT be a triangle in Ω\Omega. In Lemma 5.4, if we take λ(Δ)=Δf(z)𝑑z\lambda(\Delta)=\int_{\partial\Delta}f(z)dz, 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

1|T|Tf(z)𝑑z\displaystyle\frac{1}{\left|T\right|}\int_{\partial T}f(z)dz =1|T|Tf(z)𝑑x+if(z)dy=1|T|T(ifx(z)fy(z))𝑑xdy\displaystyle=\frac{1}{\left|T\right|}\int_{\partial T}f(z)dx+if(z)dy=\frac{1}{\left|T\right|}\int_{T}\left(i\frac{\partial f}{\partial x}(z)-\frac{\partial f}{\partial y}(z)\right)dx\wedge dy
=2iTfz¯𝑑xdy.\displaystyle=2i\int_{T}\frac{\partial f}{\partial\overline{z}}dx\wedge dy.

Since the last integral has a continuous integrand, we may take the limit as the triangle TT shrinks to the point ww to obtain:

limTw1|T|Tf(z)𝑑z=2ifz¯(w).\lim_{T\downarrow w}\frac{1}{\left|T\right|}\int_{\partial T}f(z)dz=2i\frac{\partial f}{\partial\overline{z}}(w).

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.