Hostname: page-component-745bb68f8f-5r2nc Total loading time: 0 Render date: 2025-02-06T10:24:45.929Z Has data issue: false hasContentIssue false

$L^p$-regularity of the Bergman projection on quotient domains

Published online by Cambridge University Press:  08 February 2021

Chase Bender*
Affiliation:
Department of Mathematics, Central Michigan University, Mt Pleasant, MI48859, USA e-mail: chakr2d@cmich.edumaink1m@cmich.edu URL: http://people.cst.cmich.edu/chakr2d
Debraj Chakrabarti
Affiliation:
Department of Mathematics, Central Michigan University, Mt Pleasant, MI48859, USA e-mail: chakr2d@cmich.edumaink1m@cmich.edu URL: http://people.cst.cmich.edu/chakr2d
Luke Edholm
Affiliation:
Department of Mathematics, University of Michigan, Ann Arbor, MI48109, USA e-mail: edholm@umich.edu
Meera Mainkar
Affiliation:
Department of Mathematics, Central Michigan University, Mt Pleasant, MI48859, USA e-mail: chakr2d@cmich.edumaink1m@cmich.edu URL: http://people.cst.cmich.edu/chakr2d
Rights & Permissions [Opens in a new window]

Abstract

We obtain sharp ranges of $L^p$ -boundedness for domains in a wide class of Reinhardt domains representable as sublevel sets of monomials, by expressing them as quotients of simpler domains. We prove a general transformation law relating $L^p$ -boundedness on a domain and its quotient by a finite group. The range of p for which the Bergman projection is $L^p$ -bounded on our class of Reinhardt domains is found to shrink as the complexity of the domain increases .

MSC classification

Type
Article
Copyright
© Canadian Mathematical Society 2021

1 Introduction

1.1 Main result

Let $n\geq 2$ , and for each $1\leq j \leq n$ , let $b^j=(b^j_1,\dots , b^j_n)\in \mathbb {Q}^n$ be an n-tuple of rational numbers. Let $\mathscr {U}\subset {\mathbb {C}}^n$ be a bounded domain (open connected subset) of the form

(1.1) $$ \begin{align} \mathscr{U}=\left\{z\in{\mathbb{C}}^n: \text{ for } 1\leq j \leq n,\quad \prod_{k=1}^n\left\vert z_k\right\vert{{\hspace{-0.0001pt}}}^{b^j_k} <1 \right\}, \end{align} $$

where it is understood that a point $z\in {\mathbb {C}}^n$ does not belong to $\mathscr {U}$ if for some $1\leq j \leq n$ , the quantity $ \prod _{k=1}^n\left \vert z_k\right \vert {{\hspace{-0.0001pt}}}^{b^j_k}$ is not defined due to division by zero. We call a domain such as $\mathscr {U}$ a monomial polyhedron.

We refer the reader to Section 1.2 for a discussion of the significance of monomial polyhedra in complex analysis. Our main result is the following:

Theorem 1.2 Suppose that the monomial polyhedron $\mathscr {U}$ of ( 1.1 ) is bounded. Then there is a positive integer $\kappa (\mathscr {U})$ (the complexity of $\mathscr {U}$ , whose computation is described below) such that the Bergman projection on $\mathscr {U}$ is bounded on $L^p(\mathscr {U})$ if and only if

(1.3) $$ \begin{align} \frac{2\kappa(\mathscr{U})}{\kappa(\mathscr{U})+1}<p< \frac{2\kappa(\mathscr{U})}{\kappa(\mathscr{U})-1}. \end{align} $$

To compute $\kappa (\mathscr {U})$ , first define for a vector $x\in \mathbb {Q}^n\setminus \{0\}$ , the positive integer $\mathsf {h}(x)$ (the projective height of x) as follows. If we think of x as the homogeneous coordinates of a point $[x]$ in the rational projective space $\mathbb {P}^{n-1}(\mathbb {Q})$ , there is clearly an integer vector $y\in \mathbb {Z}^n$ such that $[y]=[x]$ (i.e., there is a $\lambda \in \mathbb {Q}\setminus \{0\}$ such that $y=\lambda x$ ), and we have additionally that $\gcd (y_1,\dots , y_n)=1$ . We then set

(1.4) $$ \begin{align} \mathsf{h}(x)=\sum_{j=1}^n \left\vert y_j\right\vert. \end{align} $$

We can think of $\mathsf {h}$ as a height function on $\mathbb {P}^{n-1}(\mathbb {Q})$ in the sense of Diophantine geometry, uniformly comparable to the standard multiplicative height function (see [Reference Hindry and SilvermanHS00, pp. 174 ff.]).

Let B be the $n\times n$ matrix whose entry in the jth row and kth column is $b^j_k$ , i.e., the jth row of B is the multi-index $b^j$ in (1.1). It will follow from our work below (Proposition 3.2) that the matrix $B\in M_n(\mathbb {Q})$ is invertible. We define

(1.5) $$ \begin{align} \kappa(\mathscr{U})= \max_{1\leq k \leq n} \mathsf{h}(B^{-1}e_k), \end{align} $$

where $e_k$ denotes the $n\times 1$ column vector all whose entries are zero, except the kth, which is 1. Notice that $B^{-1}e_k$ is simply the kth column of the matrix $B^{-1}$ , that is, the arithmetic complexity of the monomial polyhedron $\mathscr {U}$ is the maximum projective height of the columns of $B^{-1}$ , where B is the rational matrix whose rows are the multi-indices occurring in the n inequalities that define $\mathscr {U}$ in (1.1). It will be shown below in Proposition 3.5 that the integer $\kappa (\mathscr {U})$ is determined only by the domain $\mathscr {U}$ and not the particular representation on the right hand side of (1.1).

1.2 Singular Reinhardt domains in complex analysis

Except in the degenerate case when it reduces to a polydisc (e.g., when $b^j=e_j$ , the jth natural basis vector of $\mathbb {Q}^n$ ), the domain $\mathscr {U}$ is a Reinhardt pseudoconvex domain (with center of symmetry at the origin) such that the origin is a boundary point. These singular Reinhardt domains (their boundaries are not Lipschitz at 0) display pathological holomorphic extension phenomena: the best-known example is that of the Hartogs triangle $\{\left \vert z_1\right \vert <\left \vert z_2\right \vert <1\}\subset {\mathbb {C}}^2$ , corresponding to a $\mathscr {U}$ with $b^1=(1,-1), b^2=(0,1)$ (see [Reference BehnkeBeh33, Reference SibonySib75]). For example, on a singular Reinhardt domain, each holomorphic function smooth up to the closure extends holomorphically to a fixed neighborhood of the closure (see [Reference ChakrabartiCha19]), something which is impossible for smoothly bounded pseudoconvex domains [Reference Hakim and SibonyHS80, Reference CatlinCat80]. Therefore, a profound understanding of function theory on these domains is an important step in extending classical results on the regularity of the $\overline \partial $ -problem (and associated operators such as the Bergman projection) to new and more general settings (see [Reference Jarnicki and PflugJP08]). Monomial polyhedra are an interesting class of such singular Reinhardt pseudoconvex domains with tractable geometry and some very interesting properties. They can be compared to analytic polyhedra in the classical theory of pseudoconvex domains: model exhausting domains where explicit computations are possible (see [Reference VladimirovVla66, Section 24]).

The striking phenomenon observed in Theorem 1.2 was first noticed (see [Reference Edholm and McNealEM17]) in the setting of the so-called generalized Hartogs triangles, defined for coprime positive integers $k_1,k_2$ as

(1.6) $$ \begin{align} H_{k_1/k_2}=\{(z_1,z_2)\in {\mathbb{C}}^2: \left\vert z_1\right\vert{{\hspace{-0.0001pt}}}^{k_1/k_2}<\left\vert z_2\right\vert<1\}, \end{align} $$

which corresponds to $b^1=\left (\frac {k_1}{k_2},-1\right )$ and $b^2=(0,1)$ . It is striking that the range (1.3) should depend, not on the shape of the domain as a subset of ${\mathbb {C}}^2$ (which is determined in the case of $H_{k_1/k_2}$ by the “fatness exponent” $\frac {k_1}{k_2}$ ), but on the complexity (which, for, $H_{k_1/k_2}$ is $k_1+k_2$ ) , the range becoming narrower as the complexity rises. As a limiting case, if $\gamma>0$ is irrational, on the domain $\{\left \vert z_1\right \vert {{\hspace{-0.0001pt}}}^\gamma <\left \vert z_2\right \vert <1\}\subset {\mathbb {C}}^2$ (a domain of infinite complexity!), the Bergman projection is bounded in the $L^p$ -norm, only if $p=2$ . The proofs of the precise range of $L^p$ -boundedness of the Bergman projection on $H_{k_1/k_2}$ and its generalizations in earlier work [Reference Edholm and McNealEM17, Reference ChenChe17, Reference HuoHuo18, Reference ZhangZha20, Reference ZhangZha19] consist of an explicit computation of the Bergman kernel, followed by an application of Schur’s theorem on $L^p$ boundedness of operators defined by integral kernels to determine the range of $L^p$ -boundedness. Other authors have used techniques of classical harmonic analysis, such as weak-type endpoint estimates along with interpolation in $L^p$ -spaces, Muckenhoupt $A_p$ weights etc. to study related questions. See [Reference Chakrabarti and ZeytuncuCZ16, Reference EdholmEdh16, Reference Edholm and McNealEM16, Reference Chakrabarti, Edholm and McNealCEM19, Reference Huo and WickHW19, Reference Chen, Krantz and YuanCKY20, Reference Edholm and McNealEM20, Reference Chen, Jin and YuanCJY20] for other results in this circle of ideas.

Theorem 1.2 not only encompasses the known examples of domains on which the relation between the regularity of the Bergman projection and arithmetic complexity has been observed, but also substantially extends this class of domains. Its proof is based on an understanding of the geometry of monomial polyhedra as quotient domains. It is hoped that this will eventually lead to a deeper understanding of this mysterious notion of complexity, and its extension to other contexts.

1.3 Ingredients in the proof of Theorem 1.2

The range of $L^p$ -boundedness of the Bergman projection on a domain is a function theoretic property determined by its Hermitian geometry, but the full extent of this relationship is yet to be understood. This article brings to bear a new perspective on this problem in the case of monomial polyhedra: one in which the domain is realized as a quotient of a simpler domain under the action of a group of biholomorphic automorphisms $\Gamma $ (Theorem 3.12 below).

Our approach to the geometry of $\mathscr {U}$ in Section 3 is inspired by the observation that in the “log-absolute coordinates” $\xi _k=\log \left \vert z_k\right \vert $ , it is represented as

(1.7) $$ \begin{align} \sum_{k=1}^n b^j_k\xi_k<0,\quad \text{ for each } 1\leq j \leq n, \end{align} $$

which is an open polyhedral cone (intersection of open half-spaces) in the sense of convex geometry. By a classical result (see [Reference ZieglerZie95, Theorem 1.3, p. 30], also [Reference GrünbaumGrü03, Section 3.1]), such a polyhedral cone can also be represented as the cone generated by its extreme points, i.e., it is the image of an orthant (in a possibly higher dimensional Euclidean space) under a linear map. In Section 3, we prove an analogous statement for $\mathscr {U}$ : there is a domain $\mathbb {D}^n_{\mathsf {L}(B)}\subset {\mathbb {C}}^n$ , which is the product of a certain number of unit discs with a certain number of punctured unit discs, and a proper holomorphic map $\Phi _A:\mathbb {D}^n_{\mathsf {L}(B)}\to \mathscr {U}$ which is of the “quotient type” (see Definition 3.10), i.e., off some small analytic sets in the source and target, it is essentially a quotient map by a group $\Gamma $ of automorphisms of the source. It turns out that the map $\Phi _A$ is of “monomial type” in the sense of [Reference Nagel and PramanikNP20], i.e., an n-dimensional analog of the branched covering map from the disc to itself given by $z\mapsto z^a$ for an integer $a>0$ . This fact has several pleasant consequences and facilitates computations.

One of the consequences of the existence of the map $\Phi _A$ is that it allows us to compute the Bergman kernel of $\mathscr {U}$ explicitly. This has been a crucial step in the study of the Bergman projection on such domains in all prior investigations. In this paper, however, we avoid computing Bergman kernels and directly study the transformation properties of $L^p$ -Bergman spaces. However, we do show in Proposition 3.22, using Theorem 3.12, that the Bergman kernel of $\mathscr {U}$ is a rational function.

In Section 4, we study how $L^p$ -Bergman spaces and the Bergman projection acting on $L^p$ -Bergman spaces transform under proper holomorphic maps of quotient type. This point of view leads to a transformation law (Theorem 4.15 below) relating the $L^p$ -Bergman spaces and Bergman projections of the source and target—one that is closely connected to the well-known Bell’s transformation law relating the Bergman kernels. One new ingredient here is the use of subspaces invariant under the action of the deck transformation group of the proper holomorphic map, which allows us to state a sharp result which can be used in the Proof of Theorem 1.2. We believe that the considerations of Section 4 have an independent interest beyond their application here.

The transformation law Theorem 4.15 can be used to pull back the problem from $\mathscr {U}$ to the well-understood domain $\mathbb {D}^n_{\mathsf {L}(B)}$ , which is the polydisc, except for a missing analytic hypersurface. The pulled-back problem can be solved using classical estimates on the polydisc. This is achieved in Sections 5 and 6, completing the proof of Theorem 1.2.

The germ of the idea of relating the $L^p$ -regularity of the Bergman projection with the properties of a “uniformizing” map from a simpler domain may already be found in [Reference Chakrabarti and ZeytuncuCZ16, Reference Chen, Krantz and YuanCKY20]. The sharper version of this technique presented in this paper may be thought of as a step toward a unified understanding of the way in which boundary singularities affect the mapping properties of the Bergman projection.

1.4 Examples in ${\mathbb {C}}^2$

It is not difficult to see that a monomial polyhedron in ${\mathbb {C}}^n$ is bounded by n Levi-flat “faces” (in the log-absolute representation (1.7), these faces are linear hyperplanes; see the proof of Proposition 3.5). The generalized Hartogs triangles of (1.6) are special in that one of the faces is a “coordinate face”, i.e., represented by a coordinate hyperplane in log-absolute coordinates. In (1.6), this is, $\{\left \vert z_2\right \vert =1\}$ which corresponds to $\{\xi _2=0\}$ in the log-absolute representation. In two dimensions, the generic monomial polyhedron has two noncoordinate faces, and can be thought of as an intersections of two generalized Hartogs triangles. The Reinhardt shadows in ${\mathbb {R}}^2$ (the image of $z\mapsto (\left \vert z_1\right \vert , \left \vert z_2\right \vert )$ ) of three monomial polyhedra in ${\mathbb {C}}^2$ can be seen in Figure 1. The domains corresponding to $(a), (b),$ and $(c)$ are respectively given by:

$$ \begin{align*} \left\{ \left\vert z_1\right\vert{{\hspace{-0.0001pt}}}^4 < \left\vert z_2\right\vert < \left\vert z_1\right\vert{{\hspace{-0.0001pt}}}^{1/3} \right\}, \quad \left\{ \left\vert z_1\right\vert{{\hspace{-0.0001pt}}}^{1/2} < \left\vert z_2\right\vert < \left\vert z_1\right\vert{{\hspace{-0.0001pt}}}^{1/4} \right\}, \quad \text{ and }\left\{ \left\vert z_1\right\vert{{\hspace{-0.0001pt}}}^{1/2} < \left\vert z_2\right\vert < 1 \right\}. \end{align*} $$

Figure 1: Reinhardt shadows of some monomial polyhedra.

1.5 $L^p$ -theory of the Bergman projection

We collect here some general information about the Bergman projection, and set up notation for later use.

Let $\Omega $ be a domain (an open connected set) in ${\mathbb {C}}^n$ . The Bergman space $A^2(\Omega )$ is the Hilbert space of holomorphic functions on $\Omega $ which are square integrable with respect to the Lebesgue measure; see [Reference KrantzKra13] for a modern treatment. The space $A^2(\Omega )$ is a closed subspace of $L^2(\Omega )$ , the usual Hilbert space of measurable functions square integrable with respect to the Lebesgue measure. The Bergman projection is the orthogonal projection

$$ \begin{align*} \boldsymbol{B}_{\Omega}:L^2(\Omega)\to A^2(\Omega). \end{align*} $$

The construction of Bergman spaces has a contravariant functorial character. If $\phi :\Omega _1\to \Omega _2$ is a suitable holomorphic map of domains, we can associate a continuous linear mapping of Hilbert spaces $\phi ^\sharp : L^2(\Omega _2)\to L^2(\Omega _1)$ defined for each $f\in L^2(\Omega _2)$ by

(1.8) $$ \begin{align} \phi^\sharp(f)= f\circ \phi\cdot \det \phi', \end{align} $$

where $\phi '(z):{\mathbb {C}}^n\to {\mathbb {C}}^n$ is the complex derivative of the map $\phi $ at $z\in \Omega _1$ . It is clear that $\phi ^\sharp $ restricts to a map $A^2(\Omega _2) \to A^2(\Omega _1)$ . We will refer to $\phi ^\sharp $ as the pullback induced by $\phi $ . It is not difficult to see that if $\phi $ is a biholomorphism, then the pullback $\phi ^\sharp $ is an isometric isomorphism of Hilbert spaces $L^2(\Omega _2)\cong L^2(\Omega _1)$ , and restricts to an isometric isomorphism $A^2(\Omega _2) \cong A^2(\Omega _1)$ . This biholomorphic invariance of Bergman spaces can be understood intrinsically by interpreting the Bergman space as a space of top-degree holomorphic forms (see [Reference KobayashiKob59] or [Reference KrantzKra13, p. 178 ff.]), and the map $\phi ^\sharp $ as the pullback map of forms induced by the holomorphic map $\phi $ . This invariance can be extended to proper holomorphic mappings via Bell’s transformation formula, and lies at the heart of classical applications of Bergman theory to the boundary regularity of holomorphic maps; see [Reference BellBel81, Reference BellBel82, Reference Diederich and FornaessDF82, Reference Bell and CatlinBC82].

For $0<p<\infty $ , define $L^p$ -Bergman spaces $A^p(\Omega )$ of pth power integrable holomorphic functions on $\Omega $ . For $p\ge 1$ , these are Banach spaces when equipped with the $L^p$ -norm. An extensive theory of these spaces on the unit disc has been developed, in analogy with the theory of Hardy spaces (cf. [Reference Duren and SchusterDS04, Reference Hedenmalm, Korenblum and ZhuHKZ00]). Unlike the $L^2$ -Bergman space, the general $L^p$ -Bergman space is not invariantly determined by the complex structure alone, but also depends on the Hermitian structure of the domain as a subset of ${\mathbb {C}}^n$ . An important question about these spaces is the boundedness of the Bergman projection in the $L^p$ -norm. After initial results were obtained for discs and balls ([Reference Zaharjuta and JudovicZJ64, Reference Forelli and RudinFR75]) the problem was studied on various classes of smoothly bounded pseudoconvex domains using estimates on the kernel (e.g., [Reference Phong and SteinPS77, Reference McNeal and SteinMS94]). On these domains the Bergman projection is bounded in $L^p$ for $1<p<\infty $ . Many examples have been given which show that there are domains on which the Bergman projection fails to be bounded in $L^p$ for certain p. See [Reference BarrettBar84, Reference Barrett and ŞahutoğluBŞ12, Reference Krantz and PelosoKP08, Reference HedenmalmHed02], in addition to the singular Reinhardt domains already mentioned in [Reference Chakrabarti and ZeytuncuCZ16, Reference Edholm and McNealEM16, Reference Edholm and McNealEM17, Reference ChenChe17, Reference Chen, Krantz and YuanCKY20, Reference HuoHuo18, Reference Huo and WickHW19]. This paper proves sharp $L^p$ -regularity results on a large class of singular Reinhardt domains.

1.6 Two examples of Theorem 1.2

We illustrate the application of Theorem 1.2 to two families of domains generalizing the Hartogs triangle to higher dimensions, recapturing the results of [Reference ZhangZha20, Reference ZhangZha19].

Let $k = (k_1, \ldots , k_n)$ be an n-tuple of positive integers. The domain

(1.9) $$ \begin{align} \mathscr{H}_{k}=\left\{z\in \mathbb{D}^n: \left\vert z_1\right\vert{{\hspace{-0.0001pt}}}^{k_1} < \prod_{j=2}^n \left\vert z_j\right\vert{{\hspace{-0.0001pt}}}^{k_j}\right\}, \end{align} $$

was introduced in [Reference Chakrabarti, Konkel, Mainkar and MillerCKMM20] where it was called an elementary Reinhardt domain of signature 1, and its Bergman kernel was computed explicitly. We see that $\mathscr {H}_{k}$ is a monomial polyhedron as in (1.1), since

(1.10) $$ \begin{align} \mathscr{H}_{k}=\left\{z\in \mathbb{C}^n: \left\vert z_1\right\vert{{\hspace{-0.0001pt}}}^{k_1} \left\vert z_2\right\vert{{\hspace{-0.0001pt}}}^{-k_2}\cdots \left\vert z_n\right\vert{{\hspace{-0.0001pt}}}^{-k_n}< 1,\text{ and } \left\vert z_j\right\vert < 1 \text{ for } 2 \leq j \leq n \right\}. \end{align} $$

The matrix B whose rows are multi-indices occurring in the inequalities in (1.10) is then given by

where ${I}_{n-1}$ is the identity matrix of size $n-1$ . The projective height of the first column of $B^{-1}$ is 1, and for $2\leq j\leq n$ that of the jth column is $\dfrac {k_1+k_j}{\gcd (k_1,k_j)}$ , using (1.4). Therefore, we get the complexity of $\mathscr {H}_{k}$ as (see (1.5)):

$$ \begin{align*}\displaystyle{\kappa({\mathscr{H}_{k}}) = \max \left\{ 1, \frac{k_1+k_2}{\mathrm{gcd}(k_1, k_2)}, \ldots, \frac{k_1+k_n}{\mathrm{gcd}(k_1, k_n)} \right\}}= \max_{2\leq j \leq n} \left\{\frac{k_1+k_j}{\mathrm{gcd}(k_1, k_j)}\right\}. \end{align*} $$

Noting that the function $x\mapsto \frac {x}{x-1}$ is decreasing for $x>1$ , and the function $x\mapsto \frac {x}{x+1}$ is increasing, by Theorem 1.2, the Bergman projection on $\mathscr {H}_{k}$ is bounded on $L^p(\mathscr {H}_{k})$ if and only if

$$ \begin{align*} \max_{2\leq j \leq n}\, \frac{2(k_1+k_j)}{k_1+k_j+\mathrm{gcd}(k_1, k_j)} < p < \min_{2\leq j \leq n}\, \frac{2(k_1 +k_j)}{k_1+k_j- \mathrm{gcd}(k_1, k_j)}, \end{align*} $$

in consonance with the result of [Reference ZhangZha19].

The second family of domains, which we denote by $\mathscr {S}_{k}$ , gives a different type of generalization of the Hartogs triangle. For an n-tuple of positive integers $(k_1,\dots , k_n)$ we define:

(1.11) $$ \begin{align} \mathscr{S}_{k} = \left\{ z\in \mathbb{C}^n: \left\vert z_1\right\vert{{\hspace{-0.0001pt}}}^{k_1}< \left\vert z_2\right\vert{{\hspace{-0.0001pt}}}^{k_2}<\dots <\left\vert z_n\right\vert{{\hspace{-0.0001pt}}}^{k_n}<1\right\}. \end{align} $$

In [Reference ParkPar18], the Bergman kernel of $\mathscr {S}_{k}$ was explicitly computed for $n=3$ , and in [Reference ChenChe17] the special case $k=(1,1,\dots , 1)$ was considered and it was shown that the Bergman projection is $L^p$ -bounded on $\mathscr {S}_{(1,1,\dots ,1)}$ if and only if $\frac {2n}{n+1}<p< \frac {2n}{n-1}$ . In [Reference ZhangZha20] the Bergman kernel of $\mathscr {S}_{k}$ was computed in general, and the range of $L^p$ -boundedness of the Bergman projection was determined. Since $\mathscr {S}_{k}$ is a monomial polyhedron given by

(1.12) $$ \begin{align}\mathscr{S}_{k}=\left\{z\in \mathbb{C}^n: \text{ for } 1 \leq j \leq n-1, \left\vert z_j\right\vert{{\hspace{-0.0001pt}}}^{k_j} \left\vert z_{j+1}\right\vert{{\hspace{-0.0001pt}}}^{-k_{j+1}} <1, \text{ and } \left\vert z_n\right\vert{{\hspace{-0.0001pt}}}^{k_n} < 1 \right\}, \end{align} $$

the matrix B in the definition of complexity and its inverse $B^{-1}$ are given as below:

$$ \begin{align*}B = \begin{pmatrix} k_1 & -k_2 & 0 & 0 & \cdots& 0 \\ 0& \phantom{-} k_2 & -k_3 & 0& \cdots & 0\\ \vdots & \phantom{k_2} & \ddots & \ddots & \phantom{0} &\vdots \\ \vdots & \phantom{k_2} & \phantom{k_2} & \ddots & \ddots& \vdots \\ 0 & 0& \cdots & 0 & k_{n-1} & - k_n \\ 0 & 0 & 0 & \cdots& 0 & \phantom{-} k_n \end{pmatrix}, \quad \ B^{-1} = \frac{1}{K} \begin{pmatrix} \ell_1 & \ell_1 & \ell_1 & \cdots& \cdots & \ell_1 \\ 0& \ell_2 & \ell_2 & \cdots & \cdots & \ell_2\\ 0 & 0 & \ell_3 & \cdots & \cdots& \ell_3 \\ \vdots & \vdots & \phantom{k_2} & \ddots & \phantom{k_2} & \vdots \\ 0 & 0 & \cdots& \cdots& 0 & \ell_n \end{pmatrix},\end{align*} $$

where $K = \prod _{j=1}^n k_j$ and $\ell _j =\frac {K}{k_j} $ for $1 \leq j \leq n.$ Note that the height of the mth column of $B^{-1}$ is

$$ \begin{align*} h_m= \frac{\sum_{j=1}^m \ell_j}{\mathrm{gcd}(\ell_1, \ldots, \ell_m)}. \end{align*} $$

As m increases, the numerator $\sum _{j=1}^m \ell _j$ of $h_m$ increases and the denominator $\gcd (\ell _1, \ldots , \ell _m)$ decreases, so $h_m$ increases with m. Therefore, the complexity of $\mathscr {S}_{k}$ is

$$ \begin{align*}\kappa(\mathscr{S}_{k}) = \max_{1\leq m\leq n} h_m =h_n= \frac{\sum_{j=1}^n \ell_j}{\mathrm{gcd}(\ell_1, \ldots, \ell_n)} .\end{align*} $$

Therefore, Theorem 1.2 shows that the Bergman projection on $\mathscr {S}_{k}$ is bounded on $L^p(\mathscr {S}_{k})$ if and only if

$$ \begin{align*} \frac{2 \sum_{j=1}^n \ell_j}{ \sum_{j=1}^n \ell_j + \gcd (\ell_1, \ldots, \ell_n)} < p < \frac{2 \sum_{j=1}^n \ell_j}{ \sum_{j=1}^n \ell_j - \gcd (\ell_1, \ldots, \ell_n)},\end{align*} $$

recapturing the main result of [Reference ZhangZha20].

2 Notation

2.1 Elementwise operations on matrices

Let A be an $m\times n$ matrix. We denote the entry of A at the jth row and kth column by $a^j_k$ , where $1\leq j \leq m, 1\leq k \leq n$ . We use two kinds of products of matrices: one is the standard matrix multiplication, denoted by simple juxtaposition $AB$ , or sometimes a dot for clarity: $A\cdot B$ .

We also need a second type of multiplication, the elementwise, or Hadamard–Schur product, in which the product of two $m\times n$ matrices $A=(a^j_k)$ and $B=(b^j_k)$ is the $m\times n$ matrix C, for which we have $c^j_k=a^j_kb^j_k$ , i.e., the element at a certain position of the product is the product of the corresponding entries of the factors. We denote this by

(2.1) $$ \begin{align} C=A\odot B. \end{align} $$

It will be important to distinguish between column and row vectors. We denote the R-module of $n\times 1$ column vectors by $R^n$ , where R can be one of $\mathbb {Z},\mathbb {Q},{\mathbb {R}}, {\mathbb {C}}$ . The R-module of $1\times n$ row vectors is denoted by $(R^n)^\dagger $ . We write the entries of the row vector a (resp. the column vector b) as $(a_1,\dots , a_n)$ (resp. as $(b_1,\dots ,b_n)^T$ ). We let denote a $1\times n$ row vector, all whose entries are 1. The positive integer n will be clear from the context:

(2.2)

Similarly, is the $n\times 1$ column vector, all whose entries are 1. Notice that these are identity elements for the elementwise multiplication of row and column vectors.

For real matrices of the same size $A,B$ , the notations

(2.3) $$ \begin{align} A \succ B, A \succeq B, A\prec B, A \preceq B \end{align} $$

stand for the natural elementwise order, e.g., $A \succ B$ denotes that $a^j_k> b^j_k.$ Note that this elementwise ordering of matrices is distinct from other notions of matrix ordering, such as positive (semi-)definiteness.

2.2 Vector and matrix powers

If $\alpha $ is a row vector of size n, and z is column vector of the same size, we will denote

$$ \begin{align*}z^\alpha =z_1^{\alpha_1}\dots z_n^{\alpha_n}=\prod_{j=1}^n z_j^{\alpha_j},\end{align*} $$

whenever the powers $z_j^{\alpha _j}$ make sense, and where we use the convention $0^0=1.$ For example, we could have $\alpha \in (\mathbb {Z}^n)^\dagger $ and $z\in {\mathbb {C}}^n$ such that for each j such that $\alpha _j<0$ , we have $z_j\not =0$ . In this context $\alpha $ is typically called a multi-index and $z^\alpha $ a (Laurent) monomial. We also set

(2.4) $$ \begin{align} \varphi_\alpha(z) = z^\alpha. \end{align} $$

Informally, we think of a monomial function as a “vector power” of a vector variable.

We will also use “matrix powers.” Recall from the previous subsection that for an $n\times n$ matrix A, we denote the element at the jth row and kth column of A by $a_k^j$ . Let $a^j$ denote the jth row of A, so that each row can be thought of as a multi-index. We then define for a column vector z of size n,

(2.5) $$ \begin{align} z^A=\begin{pmatrix} z^{a^1}\\\vdots\\ z^{a^n} \end{pmatrix}= \begin{pmatrix} z_1^{a^1_1}z_2^{a^1_2}\cdots z_n^{a^1_n}\\\vdots\\ z_1^{a^n_1}z_2^{a^n_2}\cdots z_n^{a^n_n} \end{pmatrix} \end{align} $$

provided each of the monomials is defined. We will also set

(2.6) $$ \begin{align} \Phi_A(z)=z^A, \end{align} $$

which is called a monomial map.

For row vectors $\alpha , \beta $ , square matrices $A,B$ and column vectors $z,w$ such that the vector and matrix powers are well-defined, the following simple algebraic properties of monomials and monomial mappings can be easily verified. Not unexpectedly, these mirror familiar rules of elementary algebra for exponents.

(2.7a) $$ \begin{align} (z\odot w)^\alpha =z^\alpha\cdot w^\alpha \\[-25pt]\nonumber \end{align} $$
(2.7b) $$ \begin{align} (z\odot w)^A=z^A\odot w^A \\[-25pt]\nonumber \end{align} $$
(2.7c) $$ \begin{align} z^{\alpha+\beta}=z^\alpha\cdot z^\beta \\[-25pt]\nonumber \end{align} $$
(2.7d) $$ \begin{align} z^{A+B}=z^A\odot z^B \\[-25pt]\nonumber \end{align} $$
(2.7e) $$ \begin{align} (z^A)^\alpha&=z^{\alpha A}& \text{ i.e.,}\quad \varphi_\alpha\circ \Phi_A= \varphi_{\alpha A} \\[-25pt]\nonumber \end{align} $$
(2.7f) $$ \begin{align} (z^A)^B&=z^{BA}& \text{ i.e.,}\quad \Phi_B\circ \Phi_A= \Phi_{B A} \end{align} $$

2.3 Two other maps

  1. (1) The (elementwise) exponential map

    (2.8) $$ \begin{align} \exp: {\mathbb{C}}^n \to ({\mathbb{C}}^*)^n, \quad \exp(z)=(e^{z_1},\dots, e^{z_n})^T \end{align} $$
    can be thought of as the exponential map associated to the abelian Lie group $({\mathbb {C}}^*)^n$ . Notice that for $\alpha \in (\mathbb {Z}^n)^\dagger $ and $A\in M_n(\mathbb {Z})$ we have:
    (2.9) $$ \begin{align} \exp(z)^\alpha=e^{\alpha z}\text{ and } \exp(z)^A=\exp(Az). \end{align} $$
  2. (2) The multi-radius map

    (2.10) $$ \begin{align} \rho : {\mathbb{C}}^n\to {\mathbb{R}}^n \quad z\mapsto(\left\vert z_1\right\vert, \dots , \left\vert z_n\right\vert )^T \end{align} $$
    restricts to a surjective group homomorphism $({\mathbb {C}}^*)^n \to ({\mathbb {R}}^+)^n$ , and satisfies, for $\alpha \in (\mathbb {Z}^n)^\dagger $ and $A\in M_n(\mathbb {Z})$
    (2.11) $$ \begin{align} \rho(z)^\alpha=\left\vert z^\alpha\right\vert \text{ and } \rho(z)^A=\rho(z^A). \end{align} $$

3 Geometry of monomial polyhedra

The main goal of this section is Theorem 3.12, which gives a “uniformization” of a monomial polyhedron by a proper holomorphic monomial map. This construction is crucial for everything that follows. We also show that the Bergman kernel of a monomial polyhedron is rational, although this fact is not used in the sequel.

3.1 The matrix B

The definition of the domain $\mathscr {U}$ as given in (1.1) can be succinctly rewritten, using the notation introduced in the previous Section 2 as

(3.1)

where $B\in M_n(\mathbb {Q})$ is the matrix whose jth row is $b^j$ . In the next proposition, we show that the matrix B in the definition of $\mathscr {U}$ can always be taken to be an integer matrix which is monotone in the sense of Collatz (see [Reference CollatzCol66, p. 376 ff.]):

Proposition 3.2 In the representation ( 3.1 ) of the bounded monomial polyhedron $\mathscr {U}$ of Theorem 1.2, we may assume without loss of generality (after switching two rows, if necessary) that

(3.3) $$ \begin{align} B\in M_n(\mathbb{Z}), \quad \det B>0, \quad \text{ and } \quad B^{-1}\succeq 0. \end{align} $$

Proof Let $B\in M_n(\mathbb {Q})$ be the matrix whose rows are $b^1,\dots ,b^n$ , where for $1\leq j \leq n$ , the vector $b^j\in (\mathbb {Q}^n)^\dagger $ , $b^j=(b^j_1,\dots , b^j_n)$ is as in (1.1), which can be written in the form (3.1) using the notation introduced above. Notice that if any one of the vectors $b^j$ is 0, then $\mathscr {U}$ is empty, since the inequality $\rho (z)^{b^j}<1$ becomes $1<1$ .

If $\delta _j>0$ is a common denominator for the rational numbers $b^j_1,\dots , b^j_n$ , then notice that for a $z\in {\mathbb {C}}^n$ , the quantity $\rho (z)^{b^j}$ is defined and less than 1 if and only if $\rho (z)^{\delta _j b^j}$ is defined and less than 1. Therefore, we can assume without loss of generality that the matrix B has integer entries. Note also that interchanging the rows of B simply corresponds renumbering the equations in (1.1), so we can further assume without loss of generality that $\det B\geq 0$ .

Now, assuming that $B\in M_n(\mathbb {Z})$ and $\det B\geq 0$ , suppose for a contradiction that $\det B=0$ . We will show that $\mathscr {U}$ is unbounded, which will contradict the hypothesis of Theorem 1.2. Since $\det B=0$ , there is a nonzero vector $x\in {\mathbb {R}}^n$ such that $Bx=0$ . Let $r\in \mathscr {U}$ be such that $r \succ 0$ (such an r exists since $\mathscr {U}$ is nonempty, open, and Reinhardt). Consider the curve $f : {\mathbb {R}} \to {\mathbb {C}}^n$ parametrized by

$$ \begin{align*} f(t)=r\odot \exp(tx),\end{align*} $$

where $\exp $ is as in (2.8). Observe that

$$ \begin{align*}f(t)^B=r^B\odot \exp(tx)^B=r^B\odot \exp(tBx)= r^B, \end{align*} $$

since $Bx=0$ . Therefore,

$$ \begin{align*}\rho(f(t))^B= \rho(f(t)^B)=r^B. \end{align*} $$

But since $r \in \mathscr {U}$ , it follows that , so in fact f is a curve in $\mathscr {U}$ . As $x \in {\mathbb {R}}^n $ is nonzero, some component $x_k$ of x is nonzero, so that the kth component of the curve f, given by $f_k(t)= r_k e^{tx_k}$ is unbounded as $t\to \pm \infty $ . Thus, $\mathscr {U}$ contains the unbounded image of the curve f, showing that $\mathscr {U}$ is unbounded if $\det B=0$ . Since by assumption $\mathscr {U}$ is bounded, we have $\det B\neq 0$ . So $\det B>0$ .

To show that $B^{-1}\succeq 0$ , we let A be the adjugate of B, i.e., $A = \operatorname {\mathrm {adj}} B$ , and we show that

$$ \begin{align*} \Phi_A((\mathbb{D}^*)^n) \subset \mathscr{U},\end{align*} $$

where $\Phi _A$ is the monomial map as in (2.6), and $(\mathbb {D}^*)^n$ is the product of n punctured unit disks. Since $\mathscr {U}$ is bounded this implies that $\Phi _A((\mathbb {D}^*)^n)$ is bounded. Let $w \in \Phi _A((\mathbb {D}^*)^n)$ and let $z \in (\mathbb {D}^*)^n$ be such that $z^A=w$ . Notice that $z \in (\mathbb {D}^*)^n$ is equivalent to , and we also have $z^A=w \in ({\mathbb {C}}^*)^n$ . By applying (2.11) and Cramer’s rule, we find

$$ \begin{align*}\rho(w)^{B} = \left(\rho(z)^A\right)^{B}= \rho(z)^{BA}=\rho(z)^{(\det B) I}=\left(\left\vert z_1\right\vert{{\hspace{-0.0001pt}}}^{\det B},\dots, \left\vert z_n\right\vert{{\hspace{-0.0001pt}}}^{\det B}\right)^T.\end{align*} $$

As $\det B> 0$ we find that , i.e., $w \in \mathscr {U}$ , yielding the desired inclusion $\Phi _A((\mathbb {D}^*)^n) \subset \mathscr {U}$ .

If $A \not \succeq 0$ then there exist some $1 \leq j,k \leq n$ such that $a^j_k < 0$ . Let $z\in (\mathbb {D}^*)^n$ be such that all its components, except the kth, are $1/2$ . Then the jth component of $\Phi _A(z)$ is given by

$$ \begin{align*} z^{a^j} = {z_k^{a^j_k}}\cdot \frac{1}{2^m},\quad \text{where }\quad m= \sum_{\substack{1 \leq \ell \leq n \\ \ell \neq k}} a^j_\ell. \end{align*} $$

As $a^j_k < 0$ we see that as $z_k\to 0$ , the monomial $z^{a^j}$ is unbounded, showing that $\Phi _A((\mathbb {D}^*)^n)$ is unbounded. Therefore, the boundedness of $\mathscr {U}$ guarantees that $A = \operatorname {\mathrm {adj}} B \succeq 0$ .

To complete the proof note that $A \succeq 0$ implies that $B^{-1}=(\det B)^{-1} \operatorname {\mathrm {adj}} B\succeq 0$ . ▪

Due to Proposition 3.2, we can easily establish a more concrete bound on monomial polyhedra:

Corollary 3.4 A bounded monomial polyhedron is contained in the unit polydisc $\mathbb {D}^n$ and its boundary contains the unit torus $\mathbb {T}^n=\{\left \vert z_j\right \vert =1, 1\leq j\leq n\}$ .

Proof For a bounded monomial polyhedron $\mathscr {U}$ , let $z\in \mathscr {U}$ so that $\rho (z)^B\prec 1$ , where B is as in Proposition 3.2. Since $B^{-1}\succeq 0$ , we have for any $r\in {\mathbb {R}}^n$ with, $0\preceq r \prec 1$ that $r^{B^{-1}}\prec 1$ . Therefore, we have $(\rho (z)^{B})^{B^{-1}}=\rho (z)\prec 1$ , i.e., $z\in \mathbb {D}^n$ . The second assertion follows on noting that in the log absolute representation (1.7) of $\mathscr {U}$ , the origin (which corresponds to the unit torus of ${\mathbb {C}}^n$ ) is a boundary point.  ▪

In the Definition (3.1), it is clear that the matrix B is not unique: permuting the order of inequalities, does not change to domain $\mathscr {U}$ , and multiplying for each j the row $b^j$ by the same rational $\delta _j>0$ also gives exactly the same $\mathscr {U}$ . Therefore, for any permutation matrix P and any positive diagonal matrix D with rational entries, the domains corresponding to B and $DPB$ are the same. The following proposition shows that the complexity $\kappa (\mathscr {U})$ is independent of the matrix B and depends only on $\mathscr {U}$ .

Proposition 3.5 The complexity $\kappa (\mathscr {U})$ is independent of the choice of the representing matrix B in ( 3.1 ).

Proof Notice that switching two rows of B does not change the complexity of $\mathscr {U}$ , we can assume that the conditions (3.3) hold for the representing matrix. In the log-absolute coordinates $\xi _k=\log \left \vert z_k\right \vert $ the domain $\mathscr {U}$ is represented by the equations (1.7), which define the open polyhedral cone $\mathscr {C}_{\mathscr {U}}=\{B\xi \prec 0\}\subset {\mathbb {R}}^n$ . We claim that for each $1\leq j\leq n$ , the hyperplane $H_j=\{ b^j \xi = 0 \}$ determined by the jth row of the matrix B is a face of $\mathscr {C}_{\mathscr {U}}$ , i.e., the intersection $\overline {\mathscr {C}_{\mathscr {U}}}\cap H_j\not =\emptyset $ . Indeed, since by Proposition 3.2, $B^{-1}\succ 0$ and B is a linear automorphism of $\mathbb {Q}^n$ , it is easy to verify that the image $B^{-1}(\mathscr {C}_{\mathscr {U}})$ is precisely the negative orthant $\{\eta \prec 0\}\subset {\mathbb {R}}^n$ . But the negative orthant clearly has faces $\{\eta _j=0\}, j=1,\dots , n$ , which corresponds to the faces $H_j$ of $\mathscr {C}_{\mathscr {U}}$ .

Now, if $C\in M_n(\mathbb {Q})$ is another matrix (satisfying (3.3)) such that

then the faces of $\mathscr {C}_{\mathscr {U}}$ are also given by $\{ c^j \xi = 0 \}, 1\leq j\leq n$ , so that there exists a permutation $\sigma \in S_n$ such that

$$ \begin{align*} \{ c^j \xi = 0 \} = H_{\sigma(j)}= \{b^{\sigma(j)} \xi = 0 \} \end{align*} $$

for each $1 \leq j \leq n$ . Then the transposes $(c^j)^T$ and $(b^{\sigma (j)})^T$ are both normal to $H_{\sigma (j)}$ , so that there is a $\delta _j \in \mathbb {Q}\setminus \{0\}$ such that $c^j = \delta _j b^{\sigma (j)}$ . Therefore, $C=DPB$ , where D is the diagonal matrix $\mathrm {diag}(\delta _1,\dots ,\delta _n)$ and P is the permutation matrix whose jth column is $e_{\sigma (j)}$ . Note that $P^{-1}$ is also a permutation matrix, and hence $P^{-1} \succ 0$ . Therefore, since $C^{-1}=B^{-1}P^{-1}D^{-1}$ and we have $C^{-1}\succ 0, B^{-1}\succ 0$ by Proposition 3.2, we have $D\succ 0$ since $B^{-1}P^{-1}\succ 0$ . Therefore,

$$ \begin{align*} \max_{1 \leq j \leq n} \mathsf{h}(C^{-1}e_j) &= \max_{1 \leq j \leq n} \mathsf{h}(B^{-1}P^{-1}D^{-1}e_j) = \max_{1 \leq j \leq n} \mathsf{h}(\frac{1}{\delta_j}B^{-1}e_{\sigma(j)}) \\ &=\max_{1 \leq j \leq n} \mathsf{h}(B^{-1}e_{\sigma(j)}) =\max_{1 \leq j \leq n} \mathsf{h}(B^{-1}e_j). \end{align*} $$

This shows that $\kappa (\mathscr {U})$ as in 1.5 is well-defined, being independent of the matrix B. ▪

3.2 The Jacobian determinant of $\Phi _A$

For a subset $E \subseteq \{1 ,\dots , n \}$ , define

(3.6) $$ \begin{align} {\mathbb{C}}^n_E := \{z \in {\mathbb{C}}^n : z_k \neq 0 \text{ for all } k \in E \} = \prod_{k=1}^n \Omega_k, \end{align} $$

where $\Omega _k = {\mathbb {C}}$ if $k \not \in E$ and $\Omega _k = {\mathbb {C}}^*$ if $k \in E$ . Then for $P \in M_n(\mathbb {Q})$ define

(3.7) $$ \begin{align} \mathsf{K}(P) = \{ 1 \leq k \leq n : p^j_k < 0 \text{ for some } 1 \leq j \leq n \}, \end{align} $$

that is, $\mathsf {K}(P)$ is the set of indices of those columns of P which have at least one negative entry. Notice that if $P\in M_n(\mathbb {Z})$ the matrix power $z^P$ is defined for $z\in {\mathbb {C}}^n$ if and only if $z\in {\mathbb {C}}^n_{\mathsf {K}(P)}$ , i.e., in computing $z^P$ we do not raise zero to a negative power. Similarly, for $P\in M_n(\mathbb {Q})$ and a vector $r\in {\mathbb {R}}^n$ with $r_j\geq 0$ , the rational power $r^P$ is defined provided $r \in {\mathbb {R}}^n_{\mathsf {K}(P)} $ for exactly the same reason.

The following analog of the calculus formula $\frac {d}{dx}x^n=nx^{n-1}$ may be found in [Reference Nagel and PramanikNP09, Lemma 4.2], and will be needed frequently in the sequel.

Lemma 3.8 For any $A \in M_n (\mathbb {Z})$ and any $z \in {\mathbb {C}}^n_{\mathsf {K}(A)}$ we have

(3.9)

where $\Phi _A(z)=z^A$ and $\Phi _A'(z):{\mathbb {C}}^n\to {\mathbb {C}}^n$ is the complex derivative of $\Phi _A$ at the point $z\in {\mathbb {C}}^n_{\mathsf {K}(A)}$ .

Proof First assume that z is such that $z_k{\neq }0$ for each k. For $1\leq j,k \leq n$ , the entry in the jth row and kth column of the complex derivative matrix $\Phi _A'(z)$ is $\dfrac {\partial z^{a^j}}{\partial z_k}$ , where recall that $a^j$ denotes the multi-index in $(\mathbb {Z}^n)^\dagger $ whose entries constitute the jth row of A. If $a_k^j \neq 0$ then

$$ \begin{align*}\frac{\partial z^{a^j}}{\partial z_k}=a_k^j z_k^{a_k^j-1} \prod_{\substack{\ell=1 \\ \ell\neq k}}^n z_\ell^{a_\ell^j} = a_k^j\cdot \frac{z^{a^j}}{z_k}.\end{align*} $$

If $a_k^j=0$ , then $\dfrac {\partial z^{a^j}}{\partial z_k}=0$ , so in fact the above formula holds for all $j,k$ . Now, by the representation of a determinant as an explicit polynomial in the matrix entries, we find

where we have used the fact that

Also,

Therefore,

for z such that $z_k{\neq }0$ for each k. The result follows by analytic continuation.  ▪

3.3 A branched covering of the domain $\mathscr {U}$

In this section, we will construct a quotient map under a group action from a domain with simple geometry (a product of some copies of discs with some copies of punctured discs) to the domain $\mathscr {U}$ , and this construction will be fundamental in the proof of Theorem 1.2. We first specify what we mean by a quotient map.

Definition 3.10 Let $\Omega _1, \Omega _2\subset {\mathbb {C}}^n$ be domains, let $\Phi : \Omega _1\to \Omega _2$ be a proper holomorphic mapping. Let $\Gamma \subset \mathrm {Aut}(\Omega _1)$ be a group of biholomorphic automorphisms of $\Omega _1$ . We will say that $\Phi $ is of quotient type with group $\Gamma $ if there exist closed lower-dimensional complex-analytic subvarieties $Z_j\subset \Omega _j, j=1,2$ such that $\Phi $ restricts to a covering map

$$ \begin{align*} \Phi: \Omega_1\setminus Z_1 \to \Omega_2\setminus Z_2\end{align*} $$

and for each $z\in \Omega _2\setminus Z_2$ , the action of $\Gamma $ on $\Omega _1$ restricts to a transitive action on the fiber $\Phi ^{-1}(z)$ . Other names in the literature for such proper maps include regular and Galois proper maps. The group $\Gamma $ will be referred to as the group of deck transformations of $\Phi $ (sometimes called the Galois group).

Notice that the restricted map $\Phi :\Omega _1\setminus Z_1 \to \Omega _2\setminus Z_2$ becomes a so called regular covering map (see [Reference MasseyMas91, p. 135 ff.]), i.e., the covering map gives rise to a biholomorphism between $\Omega _2\setminus Z_2$ and the quotient $(\Omega _1\setminus Z_1)/\Gamma $ , where it can be shown that $\Gamma $ acts properly and discontinuously on $ \Omega _1\setminus Z_1$ . Further, it follows that $\Gamma $ is in fact the full group of deck transformations of the covering map $\Phi :\Omega _1\setminus Z_1 \to \Omega _2\setminus Z_2$ , and that this covering map has exactly $\left \vert \Gamma \right \vert $ sheets. Notice that by analytic continuation, the relation $\Phi \circ \sigma =\Phi $ holds for each $\sigma $ in $\Gamma $ on all of $\Omega _1$ .

Now let B be a matrix, let the set $\mathsf {K}(B)\subset \{1,\dots , n\}$ be as in (3.7), the set of indices of those columns of B which have at least one negative entry. Define a subset $\mathsf {L}(B) \subset \{1,\dots , n\}$ by setting

(3.11) $$ \begin{align} \mathsf{L}(B)=\{1 \leq \ell \leq n: a^k_\ell \neq 0 \text{ for some } k\in \mathsf{K}(B) \}. \end{align} $$

We define the domain $\mathbb {D}^n_{\mathsf {L}(B)} =\mathbb {D}^n \cap {\mathbb {C}}^n_{\mathsf {L}(B)} $ , i.e.,

$$ \begin{align*} \mathbb{D}^n_{\mathsf{L}(B)}=\{ z\in \mathbb{D}^n: z_\ell{\neq}0 \text{ for all } \ell \in \mathsf{L}(B) \}. \end{align*} $$

Notice that $\mathbb {D}^n_{\mathsf {L}(B)}$ is the product of some copies of the unit disc with some copies of the punctured unit disc.

Theorem 3.12 Let $\mathscr {U}$ be the domain of Theorem 1.2, and assume (without loss of generality) that $\mathscr {U}$ is represented as in ( 3.1 ), where the matrix $B=(b^j_k)$ satisfies the conditions ( 3.3 ), and let $A = \operatorname {\mathrm {adj}} B.$ Then the mononial map $\Phi _A$ of ( 2.6 ) maps $\mathbb {D}^n_{\mathsf {L}(B)}$ onto $\mathscr {U}$ , and the map so defined

(3.13) $$ \begin{align} \Phi_{A}: \mathbb{D}^n_{\mathsf{L}(B)} \to \mathscr{U} \end{align} $$

is a proper holomorphic map of quotient type with group $\Gamma $ consisting of the automorphisms $\sigma _\nu : \mathbb {D}^n_{\mathsf {L}(B)} \to \mathbb {D}^n_{\mathsf {L}(B)}$ given by

(3.14) $$ \begin{align} \sigma_\nu(z)= \exp\left(2\pi i A^{-1}\nu\right)\odot z \end{align} $$

for $\nu \in \mathbb {Z}^n$ . Further, the group $\Gamma $ has exactly $\det A$ elements.

Some related results may be found in [Reference Nagel and PramanikNP20, Reference ZwonekZwo99].

Proof By the conditions (3.3), we have $A =(\det B)B^{-1} \succeq 0$ , so $\Phi _A$ is defined on all of $\mathbb {D}^n_{\mathsf {L}(B)}$ . We first claim that the image $\Phi _A(\mathbb {D}^n_{\mathsf {L}(B)})$ is contained in ${\mathbb {C}}^n_{\mathsf {K}(B)}$ . Indeed, if $z\in \Phi _A(\mathbb {D}^n_{\mathsf {L}(B)})$ , then $z_\ell {\neq }0$ if $a^k_\ell {\neq }0$ for some $k\in \mathsf {K}(B)$ . Therefore the kth element of the vector $z^A$ , i.e.,

(3.15) $$ \begin{align} z^{a^k}=z_1^{a^k_1}\dots z_n^{a^k_n} \end{align} $$

is nonzero, so $z^A\in {\mathbb {C}}^n_{\mathsf {K}(B)}.$ Now notice that

since $\rho (z)^B$ (and $z^B$ ) are defined if and only if $z\in {\mathbb {C}}^n_{\mathsf {K}(B)}.$ Therefore, as in the proof of Proposition 3.2, we have

for each point $z\in \mathbb {D}^n_{\mathsf {L}(B)} $ since $\det B>0$ . This shows that $\Phi _A : \mathbb {D}^n_{\mathsf {L}(B)} \to \mathscr {U}$ is a well defined holomorphic map.

We now show that, if we think of $\Phi _A$ as a map from ${\mathbb {C}}^n$ to itself (recall that $A\succeq 0$ ), then we have

(3.16) $$ \begin{align} \Phi_A^{-1}(\mathscr{U})\subset \mathbb{D}^n_{\mathsf{L}(B)}. \end{align} $$

If $z\in {\mathbb {C}}^n$ is such that $\Phi _A(z)=z^A\in \mathscr {U}$ , then $\rho (z^A)^B$ is well defined and

. The fact that $\rho (z^A)^B$ is well defined is equivalent to $z^A\in {\mathbb {C}}^n_{\mathsf {K}(B)}$ , i.e., for each $k\in \mathsf {K}(B)$ we have the entry $z^{a^k}{\neq }0.$ Now from (3.15), we see that whenever $a^k_\ell {\neq }0$ (so $a^k_\ell \geq 1$ since $A\succeq 0$ ), we must have $z_\ell {\neq }0$ , i.e., $z\in {\mathbb {C}}^n_{\mathsf {L}(B)}$ . The other condition

on the point z shows that

Since $\det B>0$ , this shows that

, i.e., $z\in \mathbb {D}^n$ . Together these conditions show that $z\in \mathbb {D}^n_{\mathsf {L}(B)}$ , proving the desired inclusion of (3.16).

We can now show that $\Phi _A:\mathbb {D}^n_{\mathsf {L}(B)}\to \mathscr {U}$ is a proper holomorphic map. Let $K \subset \mathscr {U}$ be compact. Since the topology induced on K from $\mathscr {U}$ is the same as that induced from ${\mathbb {C}}^n$ , it follows that K is closed in ${\mathbb {C}}^n$ (and bounded). As $\Phi _A$ is continuous, $\Phi _A^{-1} (K)$ is closed in ${\mathbb {C}}^n$ and in fact $\Phi _A^{-1} (K) \subseteq \mathbb {D}^n_{\mathsf {L}(B)}$ by (3.16), so that it is bounded as well. Thus, $\Phi _A^{-1} (K)$ is compact for every compact $K \subset \mathscr {U}$ , which is to say $\Phi _A$ is proper.

Now we want to verify that the proper holomorphic map $\Phi _A:\mathbb {D}^n_{\mathsf {L}(B)} \to \mathscr {U}$ is of quotient type in the sense of Definition 3.10 with group $\Gamma = \{\sigma _\nu : \nu \in \mathbb {Z}^n \},$ with $\sigma _\nu $ as in (3.14) above. Let

be the union of the coordinate hyperplanes, which is an analytic variety in ${\mathbb {C}}^n$ of codimension one. Notice that the formula (3.9) shows that the set of regular points of the map $\Phi _A$ (i.e., the set of points $\{z\in \mathbb {D}^n_{\mathsf {L}(B)}: \det \Phi _A'(z){\neq }0 \}$ ) contains $\mathbb {D}^n_{\mathsf {L}(B)}\setminus Z_1$ , and from the fact that $\Phi _A(z)=z^A$ , we see that the regular values of $\Phi _A$ contains the open set $\mathscr {U}\setminus Z_2$ . Notice that both $\mathbb {D}^n_{\mathsf {L}(B)}\setminus Z_1$ and $\mathscr {U}\setminus Z_2$ are open subsets of the “algebraic torus” $({\mathbb {C}}^*)^n$ , and it is clear that $\Phi _A$ maps $({\mathbb {C}}^*)^n$ into itself (and is even a group homomorphism, if $({\mathbb {C}}^*)^n$ is given the group structure from the elementwise multiplication operation $\odot $ ). Since a proper holomorphic map is a covering map restricted to its regular points, it now follows that $\Phi _A: \mathbb {D}^n_{\mathsf {L}(B)}\setminus Z_1\to \mathscr {U}\setminus Z_2 $ is a holomorphic covering map. We need to show that it is regular.

Let $w\in \mathscr {U}\setminus Z_2$ have polar representation $w=\rho (w)\odot \exp (i\theta )$ , where $\theta \in {\mathbb {R}}^n$ . Let $z\in \mathbb {D}^n_{\mathsf {L}(B)}$ be such that $\Phi _A(z)=z^A=w$ . Comparing the radial and angular parts, one such preimage is

$$ \begin{align*}z=\rho(w)^{A^{-1}}\odot \exp\left( i A^{-1}\theta\right). \end{align*} $$

Now notice that the angular vector $\theta $ is known only up to an additive ambiguity of $2\pi \mathbb {Z}^n$ , i.e., two values of $\theta $ corresponding to the same w differ by $2\pi \nu $ for some integer vector $\nu \in \mathbb {Z}^n$ . Therefore two different preimages $z,z'$ of the point w (corresponding to the choices $\theta $ and $\theta +2\pi \nu $ of the angular vector in the polar representation of w) are related by

$$ \begin{align*} z' &= \rho(w)^{A^{-1}}\odot \exp\left( i A^{-1}(\theta+2\pi\nu)\right) = \rho(w)^{A^{-1}}\odot \exp\left( i A^{-1}\theta\right)\odot \exp\left(2\pi i A^{-1}\nu\right)\\ &= \exp\left(2\pi i A^{-1}\nu\right)\odot z =\sigma_\nu(z), \end{align*} $$

which shows that the group $\Gamma $ acts transitively on the fiber $\Phi ^{-1}(w)$ .

Now we want to show that the group $\Gamma $ has exactly $\det A$ elements. Consider the map $\psi : \mathbb {Z}^n\to \Gamma $ given by $\psi (\nu )= \sigma _\nu . $ For any $\nu , \mu \in \mathbb {Z}^n$ and any $z \in \mathbb {D}^n_{\mathsf {L}(B)}$ we find that

$$ \begin{align*} \psi(\nu+\mu)(z)=\sigma_{\nu + \mu}(z) &=\exp \left(2 \pi i A^{-1}(\nu + \mu) \right) \odot z \\ &= \exp \left(2 \pi i A^{-1}\nu \right)\odot \left( \exp\left(2 \pi i A^{-1} \mu \right) \odot z \right) \\ &= \sigma_\nu \circ \sigma_\mu (z)\\ &=\psi(\nu) \circ \psi(\mu) (z). \end{align*} $$

Hence, $\psi : \mathbb {Z}^n \to \Gamma $ is a surjective group homomorphism, and so we have an isomorphism

$$ \begin{align*} \overline{\psi}: \mathbb{Z}^n/\ker\psi\to \Gamma. \end{align*} $$

given by $\psi ([\nu ])=\sigma _\nu $ where $[\nu ] \in \mathbb Z^n/ \ker \psi $ is the class of $\nu \in \mathbb Z^n$ . Notice that $\nu \in \ker \psi $ if and only if . This means that $A^{-1}\nu \in \mathbb {Z}^n$ , so it follows that $\ker \psi = A(\mathbb Z^n)$ . Therefore, the group $\Gamma $ is isomorphic to the quotient $\mathbb {Z}^n/A(\mathbb {Z}^n)$ by the isomorphism $ \overline {\psi }: \mathbb {Z}^n/A(\mathbb {Z}^n)\to \Gamma $

To complete the proof, it is sufficient to recall that for any matrix $A\in M_n(\mathbb {Z})$ with $\det A{\neq }0$ we have

(3.17) $$ \begin{align} \left\vert\mathbb{Z}^n/A(\mathbb{Z}^n)\right\vert=\left\vert\det A\right\vert. \end{align} $$

For completeness, we give two proofs of (3.17), neither of which unfortunately avoids high technology. The first (see [Reference Nagel and PramanikNP20] and [Reference ZwonekZwo99, Theorem 2.1]) is based on the Smith Canonical Form of an integer matrix (see [Reference MunkresMun84, p. 53 ff.]): for any $A \in M_n (\mathbb Z)$ there exist $P,Q \in GL_n(\mathbb Z)$ and a diagonal matrix $D \in M_n(\mathbb Z)$ , say $D = \mathrm {diag}(\delta _1 ,\dots ,\delta _n)$ , such that $A = PDQ$ . Consider the automorphism of $\mathbb {Z}^n$ given by $x\mapsto P^{-1}x$ . This maps $A(\mathbb {Z}^n)$ to $D(\mathbb {Z}^n)$ isomorphically, and therefore leads to an isomorphism $\mathbb {Z}^n/A(\mathbb {Z}^n)$ with $\mathbb {Z}^n/D(\mathbb {Z}^n)$ , but $\mathbb {Z}^n/D(\mathbb {Z}^n)$ is isomorphic to the product abelian group $\prod _{j=1}^n\mathbb {Z}/\delta _j\mathbb {Z}$ , which has exactly $\prod _{j=1}^n\left \vert \delta _j\right \vert $ elements. On the other hand note that

$$ \begin{align*}&\left\vert\det A\right\vert=\left\vert \det(PDQ)\right\vert= \left\vert\det P\det D\det Q\right\vert= \left\vert\pm 1\cdot \prod_{j=1}^n \delta_j \cdot \pm 1\right\vert,\\[-6pt]\end{align*} $$

which completes the first proof.

The second proof of (3.17) is geometric, and based on noticing that the inclusion of abelian groups $A(\mathbb {Z}^n)\hookrightarrow \mathbb {Z}^n$ gives rise to a covering map of tori

(3.18) $$ \begin{align} &p:{\mathbb{R}}^n/A(\mathbb{Z}^n)\to {\mathbb{R}}^n/\mathbb{Z}^n\\[-6pt]\nonumber\end{align} $$

given by $x+A(\mathbb {Z}^n)\mapsto x+\mathbb {Z}^n$ . If we endow ${\mathbb {R}}^n$ with the Euclidean metric (thought of as a Riemannian metric), we obtain quotient (flat) Riemannian structures on both ${\mathbb {R}}^n/A(\mathbb {Z}^n)$ and ${\mathbb {R}}^n/\mathbb {Z}^n$ , so that the covering map (3.18) is actually a local isometry. Therefore, the number of sheets of this covering map coincides with the ratio $\mathrm {vol}({\mathbb {R}}^n/A(\mathbb {Z}^n))/\mathrm {vol}( {\mathbb {R}}^n/\mathbb {Z}^n)$ . But we know that $\mathrm {vol}( {\mathbb {R}}^n/\mathbb {Z}^n)$ is 1, being also the volume of the unit parallelepiped, and similarly $\mathrm {vol}({\mathbb {R}}^n/A(\mathbb {Z}^n))$ is the volume of the image of the unit parallelepiped under the action of A, which is therefore $\left \vert \det A\right \vert $ . On the other hand, by covering space theory, we see that the number of sheets is equal to the order of the group of deck transformations of the covering p, which is in turn isomorphic to $\pi _1({\mathbb {R}}^n/\mathbb {Z}^n)/p_*\pi _1({\mathbb {R}}^n/A(\mathbb {Z}^n))=\mathbb {Z}^n/A(\mathbb {Z}^n)$ . ▪

3.4 Rationality of the Bergman kernel of $\mathscr {U}$

For a domain $\Omega \subset {\mathbb {C}}^n$ , denote by $\mathrm {Rat}\left (\Omega \right )$ the rational functions of ${\mathbb {C}}^n$ restricted to $\Omega $ . An element f of $\Omega $ is defined on $\Omega \setminus Z_f$ , where $Z_f$ is an affine algebraic variety in ${\mathbb {C}}^n$ . Algebraically, $\mathrm {Rat}\left (\Omega \right )$ is a field, and is isomorphic to the field of rational functions ${\mathbb {C}}(z_1,\dots ,z_n)$ in n indeterminates. The map $\Phi _A:\mathbb {D}^n_{\mathsf {L}(B)}\to \mathscr {U}$ of (3.13) induces a mapping of fields

$$ \begin{align*} \Phi_A^*: \mathrm{Rat}\left(\mathscr{U}\right)\to \mathrm{Rat}\left(\mathbb{D}^n_{\mathsf{L}(B)}\right)\end{align*} $$

given by $\Phi _A^*(f)= f\circ \Phi _A$ . Denote by k the image

(3.19) $$ \begin{align} k=\Phi_A^*(\mathrm{Rat}\left(\mathscr{U}\right))\subset\mathrm{Rat}\left(\mathbb{D}^n_{\mathsf{L}(B)}\right).\\[-10pt]\nonumber \end{align} $$

Then k is clearly a subfield of $\mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})$ . The following lemma records a special case of a very general phenomenon:

Lemma 3.20 The field extension $\mathrm {Rat} (\mathbb {D}^n_{\mathsf {L}(B)})/k$ is Galois, and the map

(3.21) $$ \begin{align}& \Theta:\Gamma\to \mathrm{Gal}\left(\mathrm{Rat}\left(\mathbb{D}^n_{\mathsf{L}(B)}\right)/k\right), \quad \Theta(\sigma)(f)=f\circ \sigma^{-1},\\[-8pt]\nonumber \end{align} $$

where $f\in \mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)}), \sigma \in \Gamma $ , is an isomorphism of the group of deck transformations of the proper map $\Phi _A$ with the Galois group of the extension.

Proof Denote by $w_j$ the jth component of the map $\Phi _A$ , where $1\leq j \leq n$ . Notice that $w_j$ can be identified with the function $\Phi _A^*(\pi _j)=\pi _j\circ \Phi _A$ where $\pi _j$ is the jth coordinate function on $\mathscr {U}$ . Since $\mathrm {Rat}(\mathscr {U})={\mathbb {C}}(\pi _1,\dots , \pi _n)$ , the field k is generated over ${\mathbb {C}}$ by the functions $w_1,\dots , w_n$ , i.e., $k={\mathbb {C}}(w_1,\dots , w_n)$ . Denoting the coordinate functions on $\mathbb {D}^n_{\mathsf {L}(B)}$ by $z_1,\dots , z_n$ , we see that $\mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})= k(z_1,\dots , z_n)$ . To show that $\mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})/k$ is Galois, we need to show that each $z_j$ is algebraic over k, and all the conjugates of $z_j$ over k are already present in $\mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)}).$

Notice that by definition for each $1\leq j \leq n$ , we have $w_j=z_1^{a^j_1}\dots z_n^{a^j_n}=z^{a^j}$ , which follows from the fact that $w=\Phi _A(z)=z^A$ . Therefore we have

$$ \begin{align*}z^{\det A\cdot I} = z^{ \operatorname{\mathrm{adj}} A\cdot A} =(z^A)^{ \operatorname{\mathrm{adj}} A} =w^{ \operatorname{\mathrm{adj}} A}. \end{align*} $$

Denoting the jth row of the integer matrix $ \operatorname {\mathrm {adj}} A$ by $c^j$ we see that $z_j^{\det A} = w^{c^j}\in k,$ so that $z_j$ is a root of the polynomial $t^{\det A}-w^{c^j}\in k[t]$ . Further, all the roots of this polynomial (which are of the form $\omega \cdot z_j $ for a $(\det A)$ th root of unity, $\omega \in {\mathbb {C}}$ ) are clearly in $\mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})$ . It follows that $\mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})$ is Galois over the subfield k.

For each $\sigma \in \Gamma $ , it is clear that $\Theta (\sigma )$ is an automorphism of the field $\mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})$ . We need to show that it fixes k. But for $h=\Phi _A^*g\in k$ , where $g\in \mathrm {Rat}\left (\mathscr {U}\right )$ , we have

$$ \begin{align*} &\Theta(\sigma)h=h\circ\sigma^{-1}=g\circ \Phi_A\circ \sigma^{-1}=g\circ \Phi_A=h,\\[-4pt]\end{align*} $$

since $\sigma $ is a deck transformation of $\Phi _A$ . It follows that $\Theta (\sigma )\in \mathrm {Gal}(\mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})/k)$ . It is easily verified by direct computation that $\Theta $ is an injective group homomorphism. To show that $\Theta $ is surjective, for $\gamma \in \mathrm {Gal}(\mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})/k)$ , we find a deck transformation $\theta \in \Gamma $ such that $\gamma (f)=f\circ \theta $ , so that $\gamma =\Theta ( \theta ^{-1})$ . Let $\theta _j\in \mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})$ be the function $\theta _j=\gamma (z_j)$ , where $z_j\in \mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})$ is the jth coordinate function. As we saw above, we have $\theta _j=\omega _j\cdot z_j$ , where $\omega _j$ is a $(\det A)$ th root of unity. Then $\theta =(\theta _1,\dots ,\theta _n)$ defines a diagonal unitary mapping of ${\mathbb {C}}^n$ and therefore maps $\mathbb {D}^n_{\mathsf {L}(B)}$ (a product of discs and punctured discs) into itself. Also, the jth component of $\Phi _A\circ \theta $ is given by

$$ \begin{align*}&\theta^{a^j}=\theta_1^{a^j_1}\dots \theta_n^{a^j_n}= \gamma(z_1)^{a^j_1}\dots \gamma(z_n)^{a^j_n}=\gamma(z^{a^j})=\gamma(w_j)=w_j,\\[-4pt] \end{align*} $$

which is precisely the jth component of $\Phi _A$ . Therefore, $\Phi _A\circ \theta =\Phi _A$ , and $\theta \in \Gamma $ . Finally, for a rational function $f=f(z_1,\dots ,z_n)\in \mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})$ , we have

$$ \begin{align*}\gamma(f)=f(\gamma(z_1),\dots, \gamma(z_n))=f(\theta_1,\dots,\theta_n)=f\circ\theta. \end{align*} $$

 ▪

Proposition 3.22 The Bergman kernel of $\mathscr {U}$ is a rational function.

Proof Choosing a local continuous branch $\theta _j$ of $\arg w_j$ for each coordinate of $w\in \mathscr {U}$ , we can write $w=\rho (w)\odot \exp (i\theta )$ for $\theta \in {\mathbb {R}}^n$ . Then $\psi (w)= w^{A^{-1}}=\rho (w)^{A^{-1}}\odot \exp (iA^{-1}\theta )$ is a local inverse to the proper holomorphic map $\Phi _A:\mathbb {D}^n_{\mathsf {L}(B)}\to \mathscr {U}$ . Since $\Phi _A$ is of the quotient type, it follows that the local branches of $\Phi ^{-1}_A$ are $ \{\sigma \circ \psi : \sigma \in \Gamma \}.$ By Theorem 3.12, the map $\sigma :\mathbb {D}^n_{\mathsf {L}(B)}\to \mathbb {D}^n_{\mathsf {L}(B)}$ is the restriction of a unitary linear map on ${\mathbb {C}}^n$ , so, noting that the argument behind the proof of (3.9) also applies for (locally defined) rational matrix powers, we have

Therefore, by the Bell transformation formula, for $z\in \mathbb {D}^n_{\mathsf {L}(B)}$ and $w\in \mathscr {U}$ , the Bergman kernels $K_{\mathscr {U}}$ and $K_{\mathbb {D}^n_{\mathsf {L}(B)}}=K_{\mathbb {D}^n}$ are related by

For a fixed $z\in \mathbb {D}^n_{\mathsf {L}(B)}$ , consider the function L on $\mathbb {D}^n_{\mathsf {L}(B)}$ given by

where $\overline {z}=(\overline {z_1},\dots , \overline {z_n})\in \mathbb {D}^n_{\mathsf {L}(B)} $ . Recalling that

(3.23) $$ \begin{align} K_{\mathbb{D}^n}(z,w)= \frac{1}{\pi^n}\prod_{j=1}^n \frac{1}{(1-z_j\overline{w_j})^2}, \end{align} $$

we see that $L\in \mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})$ , and

We now claim that $L\in k$ , where $k\subset \mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})$ is as in (3.19). Since the extension $\mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})/k$ is Galois, it suffices to show that for each field-automorphism $\tau \in \mathrm {Gal}(\mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})/k)$ , we have $\tau (L)=L$ . But by Lemma 3.20, for each $\tau \in \mathrm {Gal}(\mathrm {Rat}(\mathbb {D}^n_{\mathsf {L}(B)})/k)$ there is a $\theta \in \Gamma $ such that $\tau (L)=L\circ \theta $ . Notice that thanks to (3.14), we can think of the map $\theta :\mathbb {D}^n_{\mathsf {L}(B)}\to \mathbb {D}^n_{\mathsf {L}(B)}$ as a linear map of ${\mathbb {C}}^n$ represented by a diagonal matrix. Therefore, we have

Consequently

establishing the claim.

Now since $L\in k$ , there is a function $R(z,\cdot )\in \mathrm {Rat}(\mathscr {U})$ such that $L=\Phi _A^* R(z,\cdot )$ , i,e,

$$ \begin{align*} L(\zeta)= R(z, \Phi_A(\zeta))= R(z, \zeta^A).\end{align*} $$

Therefore, we have

which shows that for each fixed $z\in \mathbb {D}^n_{\mathsf {L}(B)}$ , the function $K_{\mathscr {U}}(z,\cdot )$ is rational. By the Reinhardt symmetry of $K_{\mathscr {U}}$ , there is a function $\widetilde {K}$ such that

$$ \begin{align*} K_{\mathscr{U}}\left(z,w\right)=\widetilde{K}(z_1\overline{w_1},\dots, z_n\overline{w_n})= \widetilde{K}(z\odot \overline{w}).\end{align*} $$

Since for a fixed z, the function $w\mapsto \widetilde {K}(z\odot \overline {w})$ is rational, it follows that the function $t\mapsto \widetilde {K}(t)=\widetilde {K}(t_1,\dots , t_n)$ is also rational. Therefore, the function $K_{\mathscr {U}}$ is rational on $\mathscr {U}\times \mathscr {U}$ . ▪

4 Transformation of $L^p$ -Bergman spaces under quotient maps

4.1 Definitions

In order to state our results, we introduce some terminology:

Definition 4.1 We say that a linear map T between Banach spaces $(E_1, \left \Vert \cdot \right \Vert {{\hspace{-0.0001pt}}}_1)$ and $(E_2, \left \Vert \cdot \right \Vert {{\hspace{-0.0001pt}}}_2)$ is a homothetic isomorphism if it is a continuous bijection (and therefore has a continuous inverse) and there is a constant $C>0$ such that for each $x\in E_1$ we have

$$ \begin{align*} \left\Vert Tx\right\Vert{{\hspace{-0.0001pt}}}_2 = C \left\Vert x\right\Vert{{\hspace{-0.0001pt}}}_1.\end{align*} $$

Remark If T is a homothetic isomorphism between Banach spaces, then $\tfrac {1}{\left \Vert T\right \Vert }\cdot T$ is an isometric isomorphism of Banach spaces, so a homothetic isomorphism is simply the product of an isometric isomorphism and a scalar operator. In particular, a homothetic isomorphism between Hilbert spaces preserves angles, and in particular orthogonality of vectors.

The following definition and facts are standard:

Definition 4.2 Let $\Omega \subset {\mathbb {C}}^n$ be a domain, let $\lambda>0$ be a continuous function (the weight), and let $0<p<\infty $ . Then we define

$$ \begin{align*} L^p(\Omega, \lambda)= \left\{f:\Omega \to {\mathbb{C}} \text{ measurable }: \int_\Omega \left\vert f\right\vert{{\hspace{-0.0001pt}}}^p \lambda\, dV < \infty\right\}\end{align*} $$

and

$$ \begin{align*} A^p(\Omega, \lambda)= \left\{f:\Omega \to {\mathbb{C}} \text{ holomorphic }: \int_\Omega \left\vert f\right\vert{{\hspace{-0.0001pt}}}^p\lambda\, dV < \infty\right\},\end{align*} $$

where the latter is called a weighted Bergman space. If $p\geq 1$ , then each of $L^p(\Omega ,\lambda )$ and $A^p(\Omega , \lambda )$ is a Banach space with the natural weighted norm, and $A^p(\Omega ,\lambda )$ is a closed subspace of $L^p(\Omega ,\lambda )$ .

We will make extensive use of the following notion:

Definition 4.3 Given a group G of biholomorphic automorphisms of a domain $\Omega \subset {\mathbb {C}}^n$ , and a space $\mathfrak {F}$ of functions on $\Omega $ , we denote by $[ \mathfrak {F}]^G$ the subspace of $\mathfrak {F}$ consisting of functions which are G-invariant in the following sense

(4.4) $$ \begin{align} \left[\mathfrak{F}\right]^G = \{f\in \mathfrak{F} : f= \sigma^\sharp(f) \text{ for all } \sigma \in G\}, \end{align} $$

where $\sigma ^\sharp $ is the pullback induced by $\sigma $ as in (1.8).

Remark Interpreting $\mathfrak {F}$ as a space of holomorphic forms on $\Omega $ by associating $f\in \mathfrak {F}$ with the form $fdz_1\wedge \dots \wedge dz_n$ , this simply says that the forms in $[\mathfrak {F}]^G$ are invariant under pullback by elements of G.

4.2 Transformation of Bergman spaces

With the above definitions, we are ready to state and prove the following elementary facts. For completeness, we give details of the proofs.

Proposition 4.5 Let $\Omega _1,\Omega _2$ be domains in ${\mathbb {C}}^n$ , and let $\Phi :\Omega _1\to \Omega _2$ be a proper holomorphic map of quotient type with group $\Gamma \subset \mathrm {Aut}(\Omega _1)$ . Then for $1<p<\infty $ , the pullback map $\Phi ^\sharp $ gives rise to a homothetic isomorphism

(4.6) $$ \begin{align} \Phi^\sharp: L^p(\Omega_2)\to \left[L^p\left(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p}\right)\right]^{\Gamma}. \end{align} $$

This restricts to a homothetic isomorphism

(4.7) $$ \begin{align} \Phi^\sharp: A^p(\Omega_2)\to \left[A^p\left(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p}\right)\right]^{\Gamma}. \end{align} $$

Proof Let f be a function on $\Omega _2$ , and let $g=\Phi ^\sharp f$ be its pullback to $\Omega _1$ , then we have for each $\sigma \in \Gamma $ that

$$ \begin{align*} \sigma^\sharp(g)=\sigma^\sharp (\Phi^\sharp f) = (\Phi\circ \sigma)^\sharp f = \Phi^\sharp f =g, \end{align*} $$

where we have used the contravariance of the pullback $(\Phi \circ \sigma )^\sharp =\sigma ^\sharp \circ \Phi ^\sharp $ and the fact that $\Phi \circ \sigma =\Phi $ which follows since the action of $\Gamma $ on $\Omega _2$ restricts to actions on each of the fibers. This shows that the range of $\Phi ^\sharp $ consists of $\Gamma $ -invariant functions. Special cases of this invariance were already noticed [Reference Misra, Roy and ZhangMSRZ13, Reference Chen, Krantz and YuanCKY20].

To complete the proof of (4.6) we must show that

  1. (i) for each $f\in L^p(\Omega _2)$ ,

    (4.8) $$ \begin{align} \left\Vert\Phi^\sharp f\right\Vert{{\hspace{-0.0001pt}}}^p_{L^p\left(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p}\right)}= \left\vert\Gamma\right\vert\cdot \left\Vert f\right\Vert{{\hspace{-0.0001pt}}}^p_{L^p(\Omega_2)}, \end{align} $$
  2. (ii) The image $\Phi ^\sharp (L^p(\Omega _2))$ is precisely $\left [L^p\left (\Omega _1, \left \vert \det \Phi '\right \vert {{\hspace{-0.0001pt}}}^{2-p}\right )\right ]^\Gamma $ .

Let $Z_1, Z_2$ be as in the Definition 3.10 of proper holomorphic maps of quotient type, i.e., $\Phi $ is a regular covering map from $\Omega _1\setminus Z_1$ to $\Omega _2\setminus Z_2$ . Let U be an open set in $\Omega _2\setminus Z_2$ which is evenly covered by $\Phi $ , and let V be an open set of $\Omega _1\setminus Z_1$ which is mapped biholomorphically by $\Phi $ onto U. Then the inverse image $\Phi ^{-1}(U)$ is the finite disjoint union $\bigcup _{\sigma \in \Gamma } \sigma V$ . Therefore, if $f\in L^p(\Omega _2)$ is supported in U, then $\Phi ^\sharp f$ is supported in $\bigcup _{\sigma \in \Gamma } \sigma V$ , and we have

$$ \begin{align*} &\left\Vert\Phi^\sharp f\right\Vert{{\hspace{-0.0001pt}}}^p_{L^p\left(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p}\right)}\\[10pt]&\quad =\int_{\Omega_1} \left\vert f\circ \Phi \cdot \det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{p} \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p} dV \\[10pt]&\quad =\sum_{\sigma \in \Gamma}\int_{\sigma V} \left\vert f\circ \Phi\right\vert{{\hspace{-0.0001pt}}}^p \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^2 dV= \sum_{\sigma\in \Gamma }\int_{U} \left\vert f\right\vert{{\hspace{-0.0001pt}}}^p dV = \left\vert\Gamma\right\vert\cdot\left\Vert f\right\Vert{{\hspace{-0.0001pt}}}_{L^p(\Omega_2)}^p, \end{align*} $$

where we have used the change of variables formula applied to the biholomorphic map $\Phi |_{\sigma V}$ along with the fact that the real Jacobian determinant of the map $\Phi $ is equal to $\left \vert \det \Phi '\right \vert {{\hspace{-0.0001pt}}}^2$ .

For a general $f\in L^p(\Omega _2)$ , modify the proof as follows. There is clearly a collection of pairwise disjoint open sets $\{U_j\}_{j\in J}$ in $\Omega _2\setminus Z_2$ such that each $U_j$ is evenly covered by $\Phi $ and $\Omega _2 \setminus \bigcup _{j\in J} U_j$ has measure zero. Set $f_j = f\cdot \chi _j$ , where $\chi _j$ is the indicator function of $U_j$ , so that each $f_j\in L^p(\Omega _2)$ and $f=\sum _{j}f_j$ . Also, the functions $\Phi ^\sharp f_j$ have pairwise disjoint supports in $\Omega _1$ . Therefore we have

$$ \begin{align*} \left\Vert\Phi^\sharp f\right\Vert{{\hspace{-0.0001pt}}}^p_{L^p\left(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p}\right)}&= \sum_{j\in J} \left\Vert\Phi^\sharp f_j\right\Vert{{\hspace{-0.0001pt}}}^p_{L^p\left(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p}\right)} \\[10pt]&= \left\vert\Gamma\right\vert \sum_{j\in J} \left\Vert f_j\right\Vert{{\hspace{-0.0001pt}}}_{L^p(\Omega_2)}^p = \left\vert\Gamma\right\vert \left\Vert f\right\Vert{{\hspace{-0.0001pt}}}_{L^p(\Omega_2)}^p. \end{align*} $$

To complete the proof, we need to show that $\Phi ^\sharp $ is surjective in both (4.6) and (4.7). Let $g\in \left [L^p\left (\Omega _1, \left \vert \det \Phi '\right \vert {{\hspace{-0.0001pt}}}^{2-p}\right )\right ]^{\Gamma } $ . Let $\{U_j\}_{j\in J}$ be as in the previous paragraph, and set $g_j= g\cdot \chi _{\Phi ^{-1}(U_j)}$ , where $\chi _{\Phi ^{-1}(U_j)}$ is the indicator function of $\Phi ^{-1}(U_j)$ . Notice that $g_j\in \left [L^p\left (\Omega _1, \left \vert \det \Phi '\right \vert {{\hspace{-0.0001pt}}}^{2-p}\right )\right ]^{\Gamma }$ . Let $V_j\subset \Phi ^{-1}(U_j)$ be such that $\Phi $ maps $V_j$ biholomorphically to $U_j$ , and let $\Psi : U_j\to V_j$ be the local inverse of $\Phi $ onto $V_j$ . Define

(4.9) $$ \begin{align} f_j = \Psi^\sharp (g_j). \end{align} $$

We claim that $f_j$ is defined independently of the choice of $V_j$ . Indeed, any other choice is of the form $\sigma V_j$ for some $\sigma \in \Gamma $ and the corresponding local inverse is $\sigma \circ \Psi $ . But we have

$$ \begin{align*} (\sigma\circ \Psi)^\sharp g_j = \Psi^\sharp \circ \sigma^\sharp g_j = \Psi^\sharp g_j =f_j, \end{align*} $$

where we have used the fact that $ \sigma ^\sharp g_j=g_j$ since since $g_j\in [L^p(\Omega _1, \left \vert \det \Phi '\right \vert {{\hspace{-0.0001pt}}}^{2-p})]^{\Gamma }$ .

Now we define $f=\sum _{j}f_j$ . Notice that the $f_j$ have pairwise disjoint support, and it is easily checked that $\Phi ^\sharp f =g$ . This establishes that (4.6) is a homothetic isomorphism.

It is clear that if f is holomorphic on $\Omega _2$ , then $\Phi ^\sharp f$ is holomorphic on $\Omega _1$ , therefore, $\Phi ^\sharp $ maps $A^p(\Omega _2)$ into $[A^p(\Omega _1, \left \vert \det \Phi '\right \vert {{\hspace{-0.0001pt}}}^{2-p})]^{\Gamma }$ . Now, in the argument in the previous paragraph showing that the image $\Phi ^\sharp (L^p(\Omega _2))$ is $[L^p(\Omega _1, \left \vert \det \Phi '\right \vert {{\hspace{-0.0001pt}}}^{2-p})]^{\Gamma }$ , local definition of the inverse map (4.9) shows that if $g\in [A^p(\Omega _1, \left \vert \det \Phi '\right \vert {{\hspace{-0.0001pt}}}^{2-p})]^{\Gamma }$ , then the f constructed by this procedure is holomorphic, and therefore lies in $A^p(\Omega _2)$ . This completes the proof of the proposition.  ▪

4.3 Bell transformation law for quotient maps

The following is a refinement (for the class of proper holomorphic maps of quotient type) of a classic result of Bell (see [Reference BellBel81, Theorem 1], [Reference BellBel82, Equation 2.2]).

Proposition 4.10 Let $\Omega _1,\Omega _2$ be domains in ${\mathbb {C}}^n$ and let $\Phi :\Omega _1\to \Omega _2$ be a proper holomorphic map of quotient type with group $\Gamma \subset \mathrm {Aut}(\Omega _1)$ . Then the following diagram commutes:

(4.11)

In order to prove the proposition, we need the following simple lemma, which shows that the Bergman projection interacts well with the action of automorphisms:

Lemma 4.12 Let $\Omega \subset {\mathbb {C}}^n$ be a domain, and let $\boldsymbol {B}_\Omega $ be its Bergman projection operator.

  1. (1) If $\sigma \in \mathrm {Aut}(\Omega )$ is a biholomorphic automorphism, then

    (4.13) $$ \begin{align} \boldsymbol{B}_{\Omega}\circ \sigma^\sharp = \sigma^\sharp \circ \boldsymbol{B}_{\Omega}. \end{align} $$
  2. (2) If $G\subset \operatorname {\mathrm {Aut}}(\Omega )$ is a group of biholomorphic automorphisms, then $\boldsymbol {B}_{\Omega }$ restricts to the orthogonal projection operator from $[L^2(\Omega )]^G$ onto $[A^2(\Omega )]^G$ .

Proof For (1), note that $\sigma ^\sharp $ is a unitary operator on $L^2(\Omega )$ and $\boldsymbol {B}_{\Omega }$ is an orthogonal projection on $L^2(\Omega )$ , therefore the unitarily similar operator $Q=\sigma ^\sharp \circ \boldsymbol {B}_{\Omega }\circ (\sigma ^\sharp )^{-1}$ is also an orthogonal projection. Since $ \sigma ^\sharp $ (and therefore its inverse) leaves $A^2(\Omega )$ invariant, it follows that the range of Q is $A^2(\Omega )$ . Therefore, $Q=\boldsymbol {B}_{\Omega }$ .

For (2), let $f\in [L^2(\Omega )]^G$ . Then using (4.13), we have for $\sigma \in G$

$$ \begin{align*} \sigma^\sharp (\boldsymbol{B}_{\Omega}f) = \boldsymbol{B}_{\Omega}(\sigma^\sharp (f)) = \boldsymbol{B}_{\Omega}f,\end{align*} $$

which shows that $\boldsymbol {B}_{\Omega }f\in [A^2(\Omega )]^G$ , so that $\boldsymbol {B}_{\Omega }$ maps the G-invariant functions $[L^2(\Omega )]^G$ into the G-invariant holomorphic functions $[A^2(\Omega )]^G$ . Since $\boldsymbol {B}_\Omega $ restricts to the identity on $[A^2(\Omega )]^G$ , it follows that the range of $\boldsymbol {B}_{\Omega }$ is $[A^2(\Omega )]^G$ . Observe that

$$ \begin{align*} \ker\left(\boldsymbol{B}_{\Omega}|_{[L^2(\Omega)]^G}\right)\subseteq \ker \boldsymbol{B}_{\Omega}= A^2(\Omega)^\perp \subseteq \left([A^2(\Omega)]^G\right)^\perp, \end{align*} $$

which shows that kernel of the restriction of $\boldsymbol {B}_{\Omega }$ to $[L^2(\Omega )]^G$ is orthogonal to its range, and therefore an orthogonal projection.  ▪

Proof of Proposition 4.10 By Proposition 4.5, the $\Phi ^\sharp $ represented by the top (resp. bottom) horizontal arrow is a homothetic isomorphism from the Hilbert space $ {L}^2(\Omega _2)$ (resp. $ {A}^2(\Omega _2)$ ) onto the Hilbert space $[{L}^2(\Omega _1)]^\Gamma $ (resp. $[{A}^2(\Omega _1)]^\Gamma $ ). Therefore, $\Phi ^\sharp $ preserves angles and in particular orthogonality. Now consider the map $ P: [{L}^2(\Omega _1)]^\Gamma \to [{A}^2(\Omega _1)]^\Gamma $ defined by

(4.14) $$ \begin{align} P= \Phi^\sharp \circ \boldsymbol{B}_{\Omega_2}\circ (\Phi^\sharp)^{-1}, \end{align} $$

which, being a composition of continuous linear maps, is a continuous linear mapping of Hilbert spaces. Notice that

$$ \begin{align*} P^2 = \Phi^\sharp \circ \boldsymbol{B}_{\Omega_2}\circ (\Phi^\sharp)^{-1}\circ \Phi^\sharp \circ \boldsymbol{B}_{\Omega_2}\circ (\Phi^\sharp)^{-1}= \Phi^\sharp \circ \boldsymbol{B}_{\Omega_2}\circ (\Phi^\sharp)^{-1}=P, \end{align*} $$

so P is a projection in $[{L}^2(\Omega _1)]^\Gamma $ , with range contained in $[{A}^2(\Omega _1)]^\Gamma $ . Since $(\Phi ^\sharp )^{-1}$ and $\Phi ^\sharp |_{A^2(\Omega )}$ are isomorphisms, and $\boldsymbol {B}_{\Omega _2}$ is surjective, it follows that P is a projection onto $[{A}^2(\Omega _1)]^\Gamma $ . We claim that P is in fact the orthogonal projection on to $[{A}^2(\Omega _1)]^\Gamma $ , i.e., the kernel of P is $\left ([{A}^2(\Omega _1)]^\Gamma \right )^\perp $ , the orthogonal complement of $[{A}^2(\Omega _1)]^\Gamma $ in $[{L}^2(\Omega _1)]^\Gamma $ . Since in formula (4.14), the maps $(\Phi ^\sharp )^{-1}$ and $\Phi ^\sharp $ are isomorphisms, it follows that $f\in \ker P$ if and only if $(\Phi ^\sharp )^{-1}f\in \ker \boldsymbol {B}_{\Omega _2}$ . But $\ker \boldsymbol {B}_{\Omega _2}= A^2(\Omega _2)^\perp $ , since the Bergman projection is orthogonal. It follows that $\ker P = \Phi ^\sharp (A^2(\Omega _2)^\perp )$ . Notice that $\Phi ^\sharp $ , being a homothetic isomorphism of Hilbert spaces, preserves orthogonality, and maps $A^2(\Omega _2)$ to $[{A}^2(\Omega _1)]^\Gamma $ isomorphically, therefore, $\Phi ^\sharp ( (A^2(\Omega _2))^\perp )= \left ([{A}^2(\Omega _1)]^\Gamma \right )^\perp $ , which establishes the claim.

Therefore, we have shown that $P=\Phi ^\sharp \circ \boldsymbol {B}_{\Omega _2}\circ (\Phi ^\sharp )^{-1}$ is the orthogonal projection from $ [{L}^2(\Omega _1)]^\Gamma $ to the subspace $[{A}^2(\Omega _1)]^\Gamma $ . To complete the proof, we only need to show that the restriction of the Bergman projection $\boldsymbol {B}_{\Omega _1}$ to the $\Gamma $ -invariant subspace $[{L}^2(\Omega _1)]^\Gamma $ is also the orthogonal projection from $[{L}^2(\Omega _1)]^\Gamma $ onto $[{A}^2(\Omega _1)]^\Gamma $ . But this follows from Lemma 4.12 above.

Thus $P=\boldsymbol {B}_{\Omega _1}|_{[L^2(\Omega _1)]^\Gamma }$ , and the commutativity of (4.11) follows.  ▪

4.4 Transformation of the Bergman projection in $L^p$ -spaces

The following result will be our main tool on studying $L^p$ -regularity of the Bergman projection in the domain $\mathscr {U}$ :

Theorem 4.15 Let $\Omega _1,\Omega _2$ be bounded domains in ${\mathbb {C}}^n$ , let $\Phi :\Omega _1\to \Omega _2$ be a proper holomorphic map of quotient type with group $\Gamma \subset \mathrm {Aut}(\Omega _1)$ . Let $p\geq 1$ . The following two assertions are equivalent:

  1. (1) The Bergman projection $\boldsymbol {B}_{\Omega _2}$ gives rise to a bounded operator mapping

    $$ \begin{align*}L^p(\Omega_2) \to A^p(\Omega_2).\end{align*} $$
  2. (2) The Bergman projection $\boldsymbol {B}_{\Omega _1}$ gives rise to a bounded operator mapping

    $$ \begin{align*}\left[L^p\left(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p}\right)\right]^{\Gamma} \to \left[A^p\left(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p}\right)\right]^{\Gamma}.\end{align*} $$

If one of the conditions ( 1 ) or ( 2 ) holds (and therefore both hold), then the following diagram commutes, where $\boldsymbol {B}_{\Omega _j}, j=1,2$ denote the extension by continuity of the Bergman projections:

(4.16)

Remark Statement (1) in Theorem 4.15 means the following: the restriction of the Bergman projection to a dense subspace of $L^p(\Omega _2)$ given by

$$ \begin{align*} \boldsymbol{B}_{\Omega_2}:L^2(\Omega_2)\cap L^p(\Omega_2) \to A^2(\Omega_2) \end{align*} $$

is bounded in the $L^p$ -norm, i.e., there is a $C>0$ such that for all $f\in L^2(\Omega _2)\cap L^p(\Omega _2)$ ,

$$ \begin{align*} \left\Vert\boldsymbol{B}_{\Omega_2} f\right\Vert{{\hspace{-0.0001pt}}}_{L^p(\Omega_2)} \leq C \left\Vert f\right\Vert{{\hspace{-0.0001pt}}}_{L^p(\Omega_2)}. \end{align*} $$

By continuity $\boldsymbol {B}_{\Omega _2}$ extends to a bounded linear operator from $L^p(\Omega _2)$ to $A^p(\Omega _2)$ .

Similarly, Statement (2) means the following: the restriction of the Bergman projection to the dense subspace of $\left [L^p\left (\Omega _1, \left \vert \det \Phi '\right \vert {{\hspace{-0.0001pt}}}^{2-p}\right )\right ]^{\Gamma }$ given by

$$ \begin{align*} \boldsymbol{B}_{\Omega_1}: [L^2(\Omega_1)]^\Gamma \cap \left[L^p\left(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p}\right)\right]^{\Gamma} \to A^2(\Omega_1) \end{align*} $$

is bounded in the $L^p\left (\Omega _1, \left \vert \det \Phi '\right \vert {{\hspace{-0.0001pt}}}^{2-p}\right )$ -norm, i.e., there is a $C>0$ such that for all $f\in [L^2(\Omega _1)]^\Gamma \cap \left [L^p\left (\Omega _1, \left \vert \det \Phi '\right \vert {{\hspace{-0.0001pt}}}^{2-p}\right )\right ]^{\Gamma }$ ,

$$ \begin{align*} \left\Vert\boldsymbol{B}_{\Omega_1}f\right\Vert{{\hspace{-0.0001pt}}}_{L^p\left(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p}\right)}\leq C \left\Vert f\right\Vert{{\hspace{-0.0001pt}}}_{L^p\left(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p}\right)}. \end{align*} $$

We now see by Lemma 4.12 that

$$ \begin{align*} \boldsymbol{B}_{\Omega_1} \left(\left[L^p\left(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p}\right)\right]^{\Gamma}\right) \subseteq \left[A^p\left(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p}\right)\right]^{\Gamma}, \end{align*} $$

where we have used continuity to extend the operator.

Proof Proposition 4.5 says that $\Phi ^\sharp $ is a homothetic isomorphism, mapping

$$ \begin{align*}L^p(\Omega_2) \to \left[L^p(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p})\right]^\Gamma,\end{align*} $$

and that it restricts to a homothetic isomorphism on the holomorphic subspaces. Similarly, $(\Phi ^\sharp )^{-1}$ has the same properties with the domains and ranges switched.

First assume Statement (2). From the diagram (4.11), we write

(4.17) $$ \begin{align} \boldsymbol{B}_{\Omega_2}= (\Phi^\sharp)^{-1}\circ \boldsymbol{B}_{\Omega_1}\circ\Phi^\sharp. \end{align} $$

By hypothesis, $\boldsymbol {B}_{\Omega _1}$ is a bounded linear operator mapping

$$ \begin{align*}\left[L^p(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p})\right]^\Gamma \to \left[A^p(\Omega_1, \left\vert\det \Phi'\right\vert{{\hspace{-0.0001pt}}}^{2-p})\right]^\Gamma.\end{align*} $$

Consequently, this composition maps $L^p(\Omega _2)$ boundedly into $A^p(\Omega _2)$ , giving statement (1). A similar argument shows that (1) implies (2).

For the commutativity of the diagram, rewrite (4.17) and see that on the subspace $L^p(\Omega _2)\cap L^2(\Omega _2)$ of $L^p(\Omega _2)$ we have the relation

(4.18) $$ \begin{align} \Phi^\sharp\circ\boldsymbol{B}_{\Omega_2}= \boldsymbol{B}_{\Omega_1}\circ\Phi^\sharp. \end{align} $$

Using the hypothesis (for the $\boldsymbol {B}_{\Omega _j}$ ) and Proposition 4.5 (for $\Phi ^\sharp $ ), we see that each of the four maps in the diagram (4.16) extends to the respective domain in that diagram and is continuous. By continuity, (4.18) continues to hold for the extended maps. This shows that the diagram (4.16) is commutative.  ▪

Remark Diagram (4.11) is a special case of diagram (4.16) for $p=2$ .

5 The unboundedness of the Bergman projection on $\mathscr {U}$

Using Proposition 5.1 and the results of Section 4, we will prove in this section the following partial form of Theorem 1.2:

Proposition 5.1 The Bergman projection is not bounded in $L^p(\mathscr {U})$ , provided

(5.2) $$ \begin{align} p \geq \frac{2\kappa(\mathscr{U})}{\kappa(\mathscr{U})-1}, \end{align} $$

where $\kappa (\mathscr {U})\in \mathbb {N}$ is the complexity of the domain $\mathscr {U}$ , as defined in Section 1.1.

5.1 Reinhardt domains

Recall some elementary facts about holomorphic function theory on Reinhardt domains (which are always assumed to be centered at the origin). Let $\Omega \subset \mathbb {C}^n$ be a Reinhardt domain. Every holomorphic function $f\in \mathcal {O}(\Omega )$ admits a unique Laurent expansion

(5.3) $$ \begin{align} f = \sum_{\alpha\in (\mathbb{Z}^n)^\dagger} a_\alpha(f) \varphi_\alpha, \end{align} $$

where for $\alpha \in (\mathbb {Z}^n)^\dagger $ , $\varphi _\alpha (z)$ is the Laurent monomial $z^\alpha $ as in (2.4), and where $a_\alpha (f) \in \mathbb C$ is the $\alpha $ th Laurent coefficient. The Laurent series of f converges absolutely and uniformly to f on every compact subset of $\Omega $ .

When f lies in the Bergman space $A^2(\Omega )$ , we can say more about the series (5.3): it is actually an orthogonal series converging in the Hilbert space $A^2(\Omega )$ , and the family of monomials

(5.4) $$ \begin{align} {\left\{\dfrac{\varphi_\alpha}{\left\Vert\varphi_\alpha\right\Vert{{\hspace{-0.0001pt}}}_{L^2}}: \varphi_\alpha\in L^2(\Omega)\right\}} \end{align} $$

forms an orthonormal basis of $A^2(\Omega )$ . In particular, if $f\in A^2(\Omega )$ then the Laurent series (5.3) can have $a_\alpha (f){\neq }0$ only when $\varphi _\alpha \in L^2(\Omega )$ . It is possible to generalize some of these results to the spaces $A^p(\Omega )$ ; see [Reference Chakrabarti, Edholm and McNealCEM19].

5.2 A criterion for unboundedness of the Bergman projection

We now give an easily checkable condition which shows $L^p$ -Bergman unboundedness on any Reinhardt domain.

Lemma 5.5 Let $\Omega $ be a bounded Reinhardt domain in ${\mathbb {C}}^n$ , and let $p\geq 2$ . Suppose that there is a multi-index $\beta \in (\mathbb {Z}^n)^\dagger $ such that

(5.6) $$ \begin{align} \varphi_\beta\in L^2(\Omega)\setminus L^p(\Omega). \end{align} $$

Then the Bergman projection $\boldsymbol {B}_\Omega $ fails to map $L^p(\Omega ) \to L^p(\Omega )$ .

Proof Define subsets $\mathcal {J}_\beta , \mathcal {K}_\beta \subset \{1,2,\cdots ,n \}$ with

$$ \begin{align*} \mathcal{J}_\beta = \{\,j : \beta_j \ge 0 \}, \quad \mathcal{K}_\beta = \{\,k : \beta_k < 0 \}, \end{align*} $$

and let

$$ \begin{align*} f(w)= \prod_{j\in \mathcal{J}_\beta} w_j^{\beta_j} \times \prod_{k\in \mathcal{K}_\beta } \left(\overline{w_k}\right)^{-\beta_k}. \end{align*} $$

Then f is a bounded function on $\Omega $ , and therefore $f \in L^p(\Omega )$ . We now show that $\boldsymbol {B}_\Omega f = C \varphi _\beta $ for some $C{\neq }0$ . Since $\varphi _\beta \not \in L^p(\Omega )$ , this will show that $\boldsymbol {B}_\Omega $ fails to map the element $f\in L^p(\Omega )$ to a function in $L^p(\Omega )$ . This will imply that $ \boldsymbol {B}_\Omega $ is not bounded in the $L^p$ -norm, since if it were so, it would extend to a map from the dense subspace $L^p(\Omega )\cap L^2(\Omega )$ to the whole of $L^p(\Omega )$ .

Let $\gamma =(\left \vert \beta _1\right \vert , \dots , \left \vert \beta _n\right \vert )\in \mathbb {N}^n$ be the multi-index obtained by replacing each entry of $\beta $ by its absolute value. Write the polar form of w as $w=\rho (w)\odot \exp (i\theta )$ for a $\theta \in {\mathbb {R}}^n$ . Then $ f(w) = \rho (w)^\gamma e^{i \beta \theta }, $ and for $\alpha \in (\mathbb {Z}^n)^\dagger $ , we have

$$ \begin{align*} \varphi_\alpha(w)=w^\alpha= \left(\rho(w)\odot \exp(i\theta) \right)^\alpha= \rho(w)^\alpha e^{i \alpha \theta}. \end{align*} $$

Further, denote by $\left \vert \Omega \right \vert =\{\rho (z):z\in \Omega \}\subset {\mathbb {R}}^n$ the Reinhardt shadow of $\Omega $ , and let $\mathbb {T}^n$ be the unit torus of n dimensions. Then, for each $\alpha \in \mathbb {Z}^n $ , we have

Since (5.4) is an orthonormal basis of $A^2(\Omega )$ it follows that all the Fourier coefficients of $\boldsymbol {B}_\Omega f$ with respect this basis vanish, except the $\beta $ th coefficient, which is nonzero. Therefore, $\boldsymbol {B}_\Omega f = C \varphi _\beta \notin L^p(\Omega ), $ for some constant $C{\neq }0$ . ▪

5.3 Preliminaries

In this section (and the following Section 6), we will use the following default notation and conventions:

  1. (1) $B\in M_n(\mathbb {Z})$ is a matrix such that the domain $\mathscr {U}$ is represented as in (3.1),

  2. (2) B satisfies the properties (3.3), which is not a loss of generality by Proposition 3.2.

  3. (3) We have $A= \operatorname {\mathrm {adj}} B$ .

Observe that then by Theorem 3.12, the monomial map $\Phi _A$ is a proper holomorphic map of quotient type.

We note the following computation:

Proposition 5.7 The upper bound in ( 1.3 ) is given by

(5.8)

where, as usual, $a_j$ is the jth column of the matrix A, where notation is as above.

Proof Recall the definition of the projective height function as in (1.4). Then we have, by projective invariance, for each $1\leq j \leq n$ :

$$ \begin{align*}\mathsf{h}(B^{-1}e_j)=\mathsf{h}(\det B\cdot B^{-1}e_j)=\mathsf{h}_j( \operatorname{\mathrm{adj}} B\cdot e_j)= \mathsf{h}(A e_j)=\mathsf{h}(a_j). \end{align*} $$

Since $A\succ 0$ , it follows that $a_j$ is a vector of non-negative integers, and the vector $\frac {1}{\gcd (a_j)}a_j$ is such that its entries are coprime non-negative integers. Therefore,

So

Since the function $x\mapsto \dfrac {2x}{x-1}$ is strictly decreasing for $x>1$ , we have

 ▪

5.4 p-allowable multi-indices

Let $\beta \in (\mathbb {Z}^n)^\dagger $ be a multi-index. We say that $\beta $ is p-allowable on a Reinhardt domain $\Omega $ if we have that the monomial $\varphi _\beta \in L^p(\Omega )$ , i.e.,

$$ \begin{align*} \int_\Omega \left\vert\varphi_\beta\right\vert{{\hspace{-0.0001pt}}}^p dV<\infty.\end{align*} $$

We denote the collection of p-allowable multi-indices on $\Omega $ by $\mathscr {S}_p(\Omega )$ . We first compute the p-allowable multi-indices on the domain $\mathscr {U}$ . Recall that the conventions introduced in Section 5.3 are in force.

Proposition 5.9 Let $\beta \in (\mathbb {Z}^n)^\dagger $ and $p>0$ . Then $\beta \in \mathscr {S}_p(\mathscr {U})$ if and only if

(5.10)

Proof By Theorem 3.12, $\Phi _A:\mathbb {D}^n_{\mathsf {L}(B)}\to \mathscr {U}$ is a proper holomorphic map of quotient type with $\det A$ sheets. Therefore, by Proposition 4.5 (in particular (4.8)) we see that

where $a_j \in \mathbb Z^n$ is the jth column of A. It is clear that $\int _{\mathscr {U}} \left \vert \varphi _\beta \right \vert {{\hspace{-0.0001pt}}}^p dV<\infty $ if and only if

for each $1\leq j \leq n$ . This completes the proof.  ▪

Remark The above proposition can be seen as a special case of [Reference ZwonekZwo00, Lemma 2.2.1].

We now consider the important case $p=2$ , so that $\mathscr {S}_2(\Omega )$ corresponds to the monomials in the Bergman space. For a matrix $A\in M_n(\mathbb {Z})$ , none of whose columns are zero, denote by $\gcd (a_j)$ the greatest common divisor of the entries in the jth column of A. We then let

(5.11) $$ \begin{align} g(A)= \left(\gcd(a_1),\dots, \gcd(a_n)\right)\in (\mathbb{Z}^n)^\dagger \end{align} $$

be the integer row vector whose jth entry is the greatest common divisor of the jth column of A.

Proposition 5.12

  1. (1) Let $\beta \in (\mathbb {Z}^n)^\dagger $ . Then $\beta \in \mathscr {S}_2(\mathscr {U})$ if and only if

    (5.13)
  2. (2) For $1\leq j \leq n$ let $\Pi _j$ be the integer hypersurface determined by equality in the jth entry of (5.13), that is

    (5.14)
    Then we have
    $$ \begin{align*} \Pi_j\cap \mathscr{S}_2(\mathscr{U}){\neq}\emptyset.\end{align*} $$

The following Lemma will be needed in the proof of part (2) of the proposition.

Lemma 5.15 Let P be an $m\times n$ integer matrix of rank m for positive integers $n\geq m$ , and let $q\in (\mathbb {Z}^n)^\dagger $ . Then there is an $x\in (\mathbb {Z}^m)^\dagger $ such that $xP\succeq q$ .

Proof Let $\mathcal {O}=\{x\in ({\mathbb {R}}^m)^\dagger :xP\succeq q \} .$ Then $\mathcal {O}$ is an unbounded convex set, so at least one of the coordinates $x_1,\dots , x_m$ is unbounded on $\mathcal {O}$ . Rename the coordinates so that $x_1$ is unbounded. It follows that the projection

$$ \begin{align*} \{x_1\in{\mathbb{R}}: (x_1,\dots, x_m)\in \mathcal{O}\} \end{align*} $$

of $\mathcal {O}$ on the coordinate axis $x_1$ is an unbounded convex set, and therefore a ray or the whole of ${\mathbb {R}}$ . For an integer N let $C_N=\{x_1=N\}\cap \mathcal {O}$ , which is naturally thought of as a subset of $({\mathbb {R}}^{m-1})^\dagger $ . Therefore, either there is an $N_1$ such that $C_N$ is nonempty if $N\geq N_1$ or there is an $N_2$ such that $C_N$ is nonempty if $N\leq N_2$ . Assuming the former, we see that the sets $C_N$ are convex subsets of $({\mathbb {R}}^{m-1})^\dagger $ and similar to each other, i.e., they are dilations of the same set. As the size of each $C_N$ becomes infinite as $N\to \infty $ , for large N, the set $C_N$ contains cubes of arbitrarily large size, where a cube is a product of intervals of the same size in each coordinate. As soon as $C_N$ contains a cube of side $(1+\epsilon )$ for some $\epsilon>0$ , we see that there is a point $M\in (\mathbb {Z}^{m-1})^\dagger $ that belongs to $C_N$ . It follows that the point $(N,M)\in \mathbb {Z}\times (\mathbb {Z}^{m-1})^\dagger $ belongs to $\mathcal {O}$ . ▪

Proof of Proposition 5.12 (1) If $p=2$ , the condition (5.10) becomes

, which is equivalent to

(5.16)

Now the jth entry of the row vector on the left of the above equation is given by

which is a positive integer divisible by $\gcd (a_j)=\gcd (a^1_j,\dots , a^n_j).$ It follows that (5.16) holds if and only if

which is precisely the content of (5.13).

(2) Fix $1\leq j \leq n$ . By the Euclidean algorithm, $\Pi _j {\neq }\emptyset $ . Choose $y \in \Pi _j$ . Define a $\mathbb {Z}$ -module homomorphism $\phi : (\mathbb {Z}^n)^\dagger \to \mathbb {Z}$ by setting $\phi (x)=xa_j$ , i.e.,

$$ \begin{align*} \phi(x_1, \ldots, x_n) = \sum_{k=1}^n x_k a^k_j.\end{align*} $$

We then see that $\Pi _j = y + \ker \phi $ . Since $\mathbb {Z}$ is a principal ideal domain, $\ker \phi $ is a free $\mathbb {Z}$ -submodule of $(\mathbb {Z}^n)^\dagger $ of rank $\leq n$ (see [Reference Dummit and FooteDF04, Theorem 4, Chapter 12 (p. 460)]). Moreover as $\phi $ is surjective, the quotient $\mathbb {Z}$ -module $(\mathbb {Z}^n)^\dagger /\ker \phi $ is isomorphic to $\mathbb {Z}$ . It can be seen, by tensoring with $\mathbb Q$ for example, that the rank of $\ker \phi $ is $n-1$ . Let D be an $(n-1)\times n$ integer matrix whose rows are a $\mathbb {Z}$ -basis of $\ker \phi $ . Then the map $f:(\mathbb {Z}^{n-1})^\dagger \to (\mathbb {Z}^n)^\dagger $ given by $ f(t)=y+tD$ is a parametrization of $\Pi _j$ , i.e., it is one-to-one and its range is precisely $\Pi _j$ . To complete the proof of the result, it is sufficient to show that $f^{-1}(\mathscr {S}_2(\mathscr {U}))$ is a nonempty subset of $(\mathbb {Z}^{n-1})^\dagger $ . Notice now that an integer vector $t\in f^{-1}(\mathscr {S}_2(\mathscr {U}))$ , i.e., $f(t)\in \mathscr {S}_2(\mathscr {U})\cap \Pi _j$ if and only if

By Lemma 5.15, there is an integer vector $t\in (\mathbb {Z}^{n-1})^\dagger $ that satisfies the above system of inequalities. This concludes the proof of part (2)  ▪

5.5 Proof of Proposition 5.1

Recall that p satisfies (5.2). Now by (5.8),

so that there is a J with $1\leq J \leq n$ such that

(5.17)

By part 2 of Proposition 5.12, there is an $\beta \in (\mathbb {Z}^n)^\dagger $ which lies in $\mathscr {S}_2(\mathscr {U})\cap \Pi _J$ . Such a $\beta $ satisfies:

(5.18)

and also

(5.19)

with $g(A)$ as in (5.11). By construction, $\beta \in \mathscr {S}_2(\mathscr {U})$ . We now claim that $\beta \not \in \mathscr {S}_p(\mathscr {U})$ . By Lemma 5.5 this shows that the Bergman projection is not bounded in $L^p(\mathscr {U})$ . To establish the claim we note that the Jth entry of the row vector

is

(5.20)

where in the second line we have used (5.18). Thanks to the inequality (5.17) it follows that the quantity in (5.20) is not positive. It follows by Proposition 5.9 that $\beta $ is not in $L^p(\mathscr {U})$ , which establishes the claim and completes the proof.

6 Boundedness of the Bergman projection

In this section, we obtain the following part of Theorem 1.2:

Proposition 6.1 Let

(6.2) $$ \begin{align} 2\leq p < \frac{2\kappa(\mathscr{U})}{\kappa(\mathscr{U})-1}, \end{align} $$

then the Bergman projection is bounded on $L^p(\mathscr {U})$ .

We begin by recalling the following fact, which will be the main “hard analysis” ingredient of the proof:

Proposition 6.3 The Bergman projection on the polydisc $\mathbb {D}^n$ gives rise to a bounded operator $\boldsymbol {B}_{\mathbb {D}^n}:L^p(\mathbb {D}^n) \to A^p(\mathbb {D}^n)$ for all $1<p<\infty $ .

Proof For the polydisc $\mathbb {D}^n$ , the Bergman projection has the well-known integral representation

$$ \begin{align*} \boldsymbol{B}_{\mathbb{D}^n} f (z)= \int_{\mathbb{D}^n} K(z,w) f(w) dV(w),\quad f\in L^2(\mathbb{D}^n),\end{align*} $$

where K is the Bergman kernel of the polydisc, which is easily shown to be given by the well-known formula (3.23).

The case $n=1$ of Propsition 6.3 is by now a staple result in Bergman theory, going back to [Reference Zaharjuta and JudovicZJ64], where it was proved using the $L^p$ -boundedness of a Calderón–Zygmund singular integral operator. Another approach, based on Schur’s test for $L^p$ -boundedness of an integral operator, was used in [Reference Forelli and RudinFR75]. An alternative proof of the main estimate needed in this method can be found in [Reference AxlerAxl88] and in the monograph [Reference Duren and SchusterDS04].

Since $\mathbb {D}^n$ is a product domain, the theorem in higher dimensions follows from a textbook application of Fubini’s theorem to the case $n=1$ . ▪

6.1 Two lemmas

The following two simple lemmas will be used to deduce monomially weighted estimates starting from Proposition 6.3:

Lemma 6.4 Let $1\leq p<\infty $ , let $n\geq 1$ , and let $\gamma \in ({\mathbb {R}}^n)^\dagger $ be such that . Then there is a $C>0$ such that for any $f\in A^p(\mathbb {D}^n)$ we have

(6.5) $$ \begin{align} \int_{\mathbb{D}^n} \left\vert f\right\vert{{\hspace{-0.0001pt}}}^p \rho^\gamma \,dV \leq C \int_{\mathbb{D}^n} \left\vert f\right\vert{{\hspace{-0.0001pt}}}^p dV, \end{align} $$

where as usual, $\rho (z)^\gamma = \prod _{j=1}^n \left \vert z_j\right \vert {{\hspace{-0.0001pt}}}^{\gamma _j}$ .

Proof Throughout this proof, C will denote a constant that depends only on p and $\gamma $ . The actual value of C may change from line to line.

Proceed by induction on the dimension n. First consider the base case $n=1$ . We have, by the Bergman inequality (cf. [Reference Duren and SchusterDS04, Theorem 1]) that there is a $C>0$ such that

$$ \begin{align*} \mathop{\mathrm{sup}}\limits_{\left\vert z\right\vert< \frac{1}{2}} \left\vert f(z)\right\vert\leq C \left\Vert f\right\Vert{{\hspace{-0.0001pt}}}_{L^p(\mathbb{D})} \end{align*} $$

for all $f\in A^p(\mathbb {D})$ . Therefore, for $f\in A^p(\mathbb {D})$ we have the estimate

(6.6) $$ \begin{align} \int_{\left\vert z\right\vert<\frac{1}{2}}\left\vert f(z)\right\vert{{\hspace{-0.0001pt}}}^p \left\vert z\right\vert{{\hspace{-0.0001pt}}}^\gamma dV(z)\leq \mathop{\mathrm{sup}}\limits_{\left\vert z\right\vert< \frac{1}{2}} \left\vert f(z)\right\vert{{\hspace{-0.0001pt}}}^p \cdot \int_{\left\vert z\right\vert<\frac{1}{2}} \left\vert z\right\vert{{\hspace{-0.0001pt}}}^\gamma dV(z) < C \cdot\left\Vert f\right\Vert{{\hspace{-0.0001pt}}}_{L^p(\mathbb{D})}^p, \end{align} $$

where we have used the fact that since $\gamma>-2$ we have

$$ \begin{align*}\int_{\left\vert z\right\vert<\frac{1}{2}} \left\vert z\right\vert{{\hspace{-0.0001pt}}}^\gamma dV(z)=2\pi\int_0^{\frac{1}{2}} r^{\gamma+1}dr <\infty. \end{align*} $$

On the other hand,

(6.7) $$ \begin{align} \int_{\frac{1}{2}\leq \left\vert z\right\vert<1} \left\vert f(z)\right\vert{{\hspace{-0.0001pt}}}^p \left\vert z\right\vert{{\hspace{-0.0001pt}}}^\gamma dV(z)\leq C \left\Vert f\right\Vert{{\hspace{-0.0001pt}}}_{L^p(\mathbb{D})}^p, \end{align} $$

where we have used the fact that $\displaystyle { \mathop {\mathrm {sup}}\limits _{\frac {1}{2}\leq \left \vert z\right \vert <1} \left \vert z\right \vert {{\hspace{-0.0001pt}}}^\gamma <\infty . }$

Adding (6.6) and (6.7), the estimate (6.5) follows in the case $n=1$ .

For the general case, assume the result established in $n-1$ dimensions. Write the coordinates of ${\mathbb {C}}^n$ as $z=(z',z_{n})\in {\mathbb {C}}^{n-1}\times {\mathbb {C}}$ , and $\gamma =(\gamma ',\gamma _n)\in {\mathbb {R}}^{n-1}\times {\mathbb {R}}$ . Then, using Fubini’s theorem

$$ \begin{align*} \int_{\mathbb{D}^n}\left\vert f\right\vert{{\hspace{-0.0001pt}}}^p \rho^{\gamma} dV& = \int_{\mathbb{D}^{n-1}}\rho(z')^{\gamma'}\left(\int_{\mathbb{D}} \left\vert f(z',z_n)\right\vert{{\hspace{-0.0001pt}}}^p\left\vert z_n\right\vert{{\hspace{-0.0001pt}}}^{\gamma_n}dV(z_n)\right)dV(z')\\ &\leq C \int_{\mathbb{D}^{n-1}}\rho(z')^{\gamma'}\left(\int_{\mathbb{D}} \left\vert f(z',z_n)\right\vert{{\hspace{-0.0001pt}}}^p dV(z_n)\right)dV(z')\\ &\leq C \int_{\mathbb{D}}\left(\int_{\mathbb{D}^{n-1}}\left\vert f(z',z_n)\right\vert{{\hspace{-0.0001pt}}}^p \rho(z')^{\gamma'} dV(z')\right)dV(z_n)\\ &\leq C \int_{\mathbb{D}}\left(\int_{\mathbb{D}^{n-1}}\left\vert f(z',z_n)\right\vert{{\hspace{-0.0001pt}}}^p dV(z')\right)dV(z_n)\\ &=C \int_{\mathbb{D}^n}\left\vert f(z',z_n)\right\vert{{\hspace{-0.0001pt}}}^p dV(z',z_n), \end{align*} $$

which proves the result.  ▪

Lemma 6.8 Let $1\leq p<\infty $ , let $n\geq 1$ and let $\lambda \in \mathbb {N}^n$ be a multi-index of non-negative integers. Then there is a $C>0$ such that for all $f\in A^p(\mathbb {D}^n)$ we have the estimate:

(6.9) $$ \begin{align} \int_{\mathbb{D}^n}\left\vert f\right\vert{{\hspace{-0.0001pt}}}^p dV \leq C \int_{\mathbb{D}^n}\left\vert\varphi_\lambda f\right\vert{{\hspace{-0.0001pt}}}^p dV, \end{align} $$

where $\varphi _\lambda (z)=z^\lambda $ is as in ( 2.4 ).

Proof We need only to prove the case in which $\lambda =(1,0,\dots ,0)$ , so that $\varphi _\lambda (z)=z_1$ . Once this special case is established, the general result follows by repeatedly applying it and permuting the coordinates.

In what follows, C will denote some positive constant that depends only on p and $\lambda $ , where the actual value of C may change from line to line. First, consider the one dimensional case, so that we have to show that for a holomorphic function f on the disc we have

$$ \begin{align*} \int_{\mathbb{D}}\left\vert f(z)\right\vert{{\hspace{-0.0001pt}}}^p dV(z) \leq C \int_{\mathbb{D}}\left\vert z f(z)\right\vert{{\hspace{-0.0001pt}}}^p dV(z), \end{align*} $$

where the left hand side is assumed to be finite (and therefore the right hand side is finite.) First note that we obviously have

(6.10) $$ \begin{align} \int_{\frac{1}{2}\leq \left\vert z\right\vert<1}\left\vert f(z)\right\vert{{\hspace{-0.0001pt}}}^p dV(z) \leq 2^p \int_{\frac{1}{2}\leq \left\vert z\right\vert<1}\left\vert z f(z)\right\vert{{\hspace{-0.0001pt}}}^p dV(z). \end{align} $$

On the other hand, if $\left \vert z\right \vert =\frac {1}{2}$ , we have

$$ \begin{align*} \left\vert f(z)\right\vert{{\hspace{-0.0001pt}}}^p &= 2^p \left\vert z f(z)\right\vert{{\hspace{-0.0001pt}}}^p\\ &\leq 2^p \mathop{\mathrm{sup}}\limits_{\left\vert w\right\vert\leq \frac{1}{2}} \left\vert w f(w)\right\vert{{\hspace{-0.0001pt}}}^p& \text{(maximum principle)}\\ &\leq C \int_{\mathbb{D}}\left\vert w f(w)\right\vert{{\hspace{-0.0001pt}}}^p dV(w) &\text{(Bergman's inequality)} \end{align*} $$

The maximum principle now implies

$$ \begin{align*} \mathop{\mathrm{sup}}\limits_{\left\vert z\right\vert\leq \frac{1}{2}} \left\vert f(z)\right\vert{{\hspace{-0.0001pt}}}^p \leq C \int_{\mathbb{D}}\left\vert z f(z)\right\vert{{\hspace{-0.0001pt}}}^p dV(z), \end{align*} $$

so that we have

(6.11) $$ \begin{align} \int_{\left\vert z\right\vert\leq \frac{1}{2}} \left\vert f(z)\right\vert{{\hspace{-0.0001pt}}}^p dV(z) \leq C \int_{\mathbb{D}}\left\vert z f(z)\right\vert{{\hspace{-0.0001pt}}}^p dV(z). \end{align} $$

Combining (6.10) and (6.11) the result follows for $n=1$ .

For the higher-dimensional case, denote the coordinates of ${\mathbb {C}}^n$ as $(z_1, z')$ where $z'=(z_2,\dots , z_n)$ . Then for $f\in A^p(\mathbb {D}^n)$ we have

$$ \begin{align*} \int_{\mathbb{D}^n} \left\vert f(z_1, z')\right\vert{{\hspace{-0.0001pt}}}^pdV(z_1, z')& = \int_{\mathbb{D}^{n-1}} \left(\int_{\mathbb{D}} \left\vert f(z_1,z')\right\vert{{\hspace{-0.0001pt}}}^p dV(z_1)\right)dV(z')\\ &\leq C \int_{\mathbb{D}^{n-1}} \left(\int_{\mathbb{D}} \left\vert z_1f(z_1,z')\right\vert{{\hspace{-0.0001pt}}}^p dV(z_1)\right)dV(z')\\ &= C\int_{\mathbb{D}^{n}} \left\vert z_1f(z_1,z')\right\vert{{\hspace{-0.0001pt}}}^p dV(z_1,z'). \end{align*} $$

 ▪

6.2 Determination of $\Gamma $ -invariant subspaces

In this section and Section 6.3, we use the notation established in Section 5.3 above, so that B and A have the same meaning as there. The group $\Gamma $ is as in Theorem 3.12: the deck transformation group associated to the map $\Phi _A$ . We will now determine the $\Gamma $ -invariant subspaces of holomorphic functions, in the sense of (4.4):

Proposition 6.12 Let $\alpha \in (\mathbb {Z}^n)^\dagger $ . Then the monomial $\varphi _\alpha (z)=z^\alpha $ belongs to the space $\left [\mathcal {O}\left ((\mathbb {D}^*)^n\right )\right ]^\Gamma $ of $\Gamma $ -invariant holomorphic functions on $(\mathbb {D}^*)^n$ if and only if there is a $\beta \in (\mathbb {Z}^n)^\dagger $ such that

Proof Recall that, by definition, the monomial $\varphi _\alpha $ is invariant under the group action of $\Gamma $ if and only if $\sigma _\nu ^\sharp (\varphi _\alpha )= (\varphi _\alpha \circ \sigma _\nu ) \det \sigma _\nu ' = \varphi _\alpha $ for all $\sigma _\nu \in \Gamma $ , where for $\nu \in \mathbb {Z}^n$ , the automorphism $\sigma _\nu \in \Gamma $ is as in (3.14). Denote the rows of $A^{-1}$ by $c^1, \ldots , c^n$ . By the linearity of $\sigma _\nu $

for $\nu \in \mathbb {Z}^n$ . Also, for $z \in (\mathbb {D}^*)^n$ and $ \nu \in \mathbb {Z}^n$ ,

$$ \begin{align*}\varphi_\alpha \circ \sigma_\nu (z) = \varphi_\alpha( \exp( 2 \pi i A^{-1} \nu) \odot z ) = (\exp( 2 \pi i A^{-1} \nu) \odot z )^\alpha = e^{2 \pi i \alpha A^{-1} \nu} z^\alpha. \end{align*} $$

Therefore,

(6.13)

Hence, $\varphi _\alpha \in [\mathcal {O} ({\mathbb {D}^n_{\mathsf {L}(B)}})]^\Gamma $ if and only if $\sigma _\nu ^\sharp (\varphi _\alpha )=\varphi _\alpha $ for all $\nu \in \mathbb {Z}^n$ , i.e.,

for all $z \in (\mathbb {D}^*)^n$ and $ \nu \in \mathbb {Z}^n, $ i.e., if and only if

Now if there is $\beta \in (\mathbb {Z}^n)^\dagger $ such that , then clearly, for all $\nu \in \mathbb {Z}^n$ . Conversely, assume that for all $\nu \in \mathbb {Z}^n$ and let $e_1,\dots , e_n$ be the standard basis of $\mathbb {Z}^n$ . Then the jth column of (where I is the $n\times n$ identity matrix) is which is therefore in $\mathbb {Z}$ . Therefore, we have . It follows that as desired.  ▪

Corollary 6.14 Let $f\in \left [\mathcal {O}(\mathbb {D}^n)\right ]^\Gamma $ . Then there is an $h\in \mathcal {O}(\mathbb {D}^n)$ such that

with $g(A)$ as in (5.11).

Proof Let $\displaystyle { f(z)=\sum _{\alpha \succeq 0}a_\alpha z^\alpha }$ , be the Taylor expansion of f. If f is $\Gamma $ invariant, then we claim that for each $\alpha $ such that $a_\alpha {\neq }0$ , we have that $z^\alpha $ is $\Gamma $ -invariant. Indeed using (6.13) we have

comparing this with the Taylor expansion of f and equating coefficients the claim follows. Therefore, the Taylor expansion of f is of the form

(6.15)

Notice that

. The integer $\beta a_j=\sum _{k=1}^n\beta _k a_j^k$ which occurs in the exponent is divisible by $\gcd (a_j)$ . Further since $\sum _{k=1}^n\beta _k a_j^k\geq 1$ it follows that $\beta a_j\geq \gcd (a_j).$ It follows that the monomial

is of the form

, where $\gamma \succeq 0$ . The corollary follows from (6.15).  ▪

6.3 Proof of Proposition 6.1

By Theorem 3.12, $\Phi _A: \mathbb {D}^n_{\mathsf {L}(B)}\to \mathscr {U} $ is a proper holomorphic map of quotient type with group $\Gamma $ . Therefore, thanks to Theorem 4.15, the Bergman projection

$$ \begin{align*} \boldsymbol{B}_{\mathscr{U}}: L^p(\mathscr{U})\to L^p(\mathscr{U})\end{align*} $$

is bounded if and only if

(6.16) $$ \begin{align} \boldsymbol{B}_{\mathbb{D}^n_{\mathsf{L}(B)}} : \left[ L^p(\mathbb{D}^n_{\mathsf{L}(B)} , \left\vert\det \Phi_A'\right\vert{{\hspace{-0.0001pt}}}^{2-p}) \right]^{\Gamma} \to \left[ A^p(\mathbb{D}^n_{\mathsf{L}(B)} , \left\vert\det \Phi_A'\right\vert{{\hspace{-0.0001pt}}}^{2-p}) \right]^{\Gamma} \end{align} $$

is bounded. We therefore show that the map (6.16) is bounded if $2\leq p < \frac {2\kappa (\mathscr {U})}{\kappa (\mathscr {U})-1}$ .

To do this we express the map of (6.16) as a composition of three maps, and show that each of these three maps is continuous:

  1. (1) We first show that we have an inclusion of Banach spaces

    $$ \begin{align*} \left[L^p(\mathbb{D}^n_{\mathsf{L}(B)} , \left\vert\det \Phi_A'\right\vert{{\hspace{-0.0001pt}}}^{2-p}) \right]^{\Gamma} \subseteq \left[L^p(\mathbb{D}^n_{\mathsf{L}(B)})\right]^\Gamma,\end{align*} $$
    and that the inclusion map
    (6.17) $$ \begin{align} \imath: \left[ L^p(\mathbb{D}^n , \left\vert\det \Phi_A'\right\vert{{\hspace{-0.0001pt}}}^{2-p}) \right]^{\Gamma} \hookrightarrow \left[L^p(\mathbb{D}^n)\right]^\Gamma \end{align} $$
    is bounded.

    It is clearly sufficient to show that there is a continuous inclusion of the full space $ L^p(\mathbb {D}^n_{\mathsf {L}(B)} , \left \vert \det \Phi _A'\right \vert {{\hspace{-0.0001pt}}}^{2-p}) $ into $L^p(\mathbb {D}^n_{\mathsf {L}(B)})$ . For $z\in \mathbb {D}^n$ we have for $p\geq 2$

    where we use the fact that the exponent of $\rho (z)$ is nonpositive, since A has non-negative integer entries and $p\geq 2$ . Thus for any $p \geq 2$ and any $f \in L^p(\mathbb {D}^n_{\mathsf {L}(B)} , \left \vert \det \Phi _A'\right \vert {{\hspace{-0.0001pt}}}^{2-p}) $ we have
    (6.18) $$ \begin{align} \int_{\mathbb{D}^n_{\mathsf{L}(B)}} \left\vert f\right\vert{{\hspace{-0.0001pt}}}^p dV \leq \frac{1}{\left\vert \det A\right\vert{{\hspace{-0.0001pt}}}^{2-p}} \int_{\mathbb{D}^n_{\mathsf{L}(B)}} \left\vert f\right\vert{{\hspace{-0.0001pt}}}^p \left\vert \det \Phi_A'\right\vert{{\hspace{-0.0001pt}}}^{2-p} dV, \end{align} $$
    which proves the inclusion and its continuity.
  2. (2) Let $P_\Gamma = \left .\boldsymbol {B}_{\mathbb {D}^n_{\mathsf {L}(B)}}\right |{{\hspace{-0.0001pt}}}_{\left [ L^p(\mathbb {D}^n_{\mathsf {L}(B)}) \right ]^{\Gamma } }$ be the restriction of the Bergman projection operator to the $\Gamma $ -invariant functions. We claim that the operator of Banach spaces

    (6.19) $$ \begin{align} P_\Gamma: \left[ L^p(\mathbb{D}^n_{\mathsf{L}(B)}) \right]^{\Gamma} \to \left[ A^p(\mathbb{D}^n_{\mathsf{L}(B)}) \right]^{\Gamma} \end{align} $$
    is bounded. Indeed, we have that $L^p(\mathbb {D}^n_{\mathsf {L}(B)})=L^p(\mathbb {D}^n)$ (since $\mathbb {D}^n_{\mathsf {L}(B)}$ is obtained from $\mathbb {D}^n$ by removing an analytic set Z, which is of measure zero) and $ A^p(\mathbb {D}^n_{\mathsf {L}(B)})= A^p(\mathbb {D}^n)$ , in the sense that the analytic set Z is a removable singularity of functions integrable in the pth power, $p\geq 2$ (see [Reference BellBel82, p. 687]). By Proposition 6.3, the Bergman projection maps $L^p(\mathbb {D}^n)$ to $A^p(\mathbb {D}^n)$ boundedly. Finally, by part (2) of Lemma 4.12, the Bergman projection maps $\Gamma $ -invariant functions to $\Gamma $ -invariant functions. The boundedness of $P_\Gamma $ follows.
  3. (3) Finally, we show that there is an inclusion $ [A^p(\mathbb {D}^n_{\mathsf {L}(B)})]^\Gamma \subset [ A^p (\mathbb {D}^n_{\mathsf {L}(B)}, \left \vert \det \Phi _A'\right \vert {{\hspace{-0.0001pt}}}^{2-p})]^\Gamma $ , and the inclusion map so determined is bounded. As noted above $A^p(\mathbb {D}^n_{\mathsf {L}(B)})=A^p(\mathbb {D}^n)$ , so it will suffice to show that there is a continuous inclusion

    (6.20) $$ \begin{align} \jmath :\left[A^p(\mathbb{D}^n)\right]^\Gamma \hookrightarrow{} \left[ A^p (\mathbb{D}^n, \left\vert \det \Phi_A'\right\vert{{\hspace{-0.0001pt}}}^{2-p}) \right]^\Gamma. \end{align} $$
    Let $f \in \left [A^p(\mathbb {D}^n)\right ]^\Gamma $ , so by Corollary 6.14, there exists $h \in \mathcal {O}(\mathbb {D}^n)$ such that
    (6.21)
    In fact, $h\in A^p(\mathbb {D}^n)$ , since
    where the first of the two integrals is clearly finite and the second integral is

    Now we have that

    (6.22)
    Combining Hypotheses (6.2) and (5.8),
    so for each $1\leq j \leq n$ ,
    Therefore, the jth component of the exponent of $\rho $ in (6.22) is
    where we have used the obvious fact that . Therefore, we have
    and so we have, for constants $C_1, C_2$ independent of the function f:
    It follows that the inclusion $\jmath $ of (6.20) is continuous.

Therefore, the map $\boldsymbol {B}_{\mathbb {D}^n_{\mathsf {L}(B)}}$ of Banach spaces in (6.16) can be represented as a composition

$$ \begin{align*} \boldsymbol{B}_{\mathbb{D}^n_{\mathsf{L}(B)}}= \jmath\circ P_\Gamma \circ \imath, \end{align*} $$

where $\imath $ , $P_\Gamma $ and $\jmath $ are as in (6.17), (6.19) and (6.20), respectively and each of which has already been shown to be continuous. The proof of Proposition 6.1 is complete.

7 Conclusion

7.1 End of proof of Theorem 1.2

Recall that, as a consequence of the self-adjointness of the Bergman projection on $L^2$ , the set of p for which the Bergman projection on a domain is $L^p$ -bounded is Hölder-symmetric (see [Reference Edholm and McNealEM16, Reference Chakrabarti and ZeytuncuCZ16]), i.e., p belongs to this set if and only if the conjugate index $p'$ also belongs to it, where $\frac {1}{p'}+\frac {1}{p}=1$ . Notice now that the index conjugate to $\frac {2\kappa (\mathscr {U})}{\kappa (\mathscr {U})-1}$ is $\frac {2\kappa (\mathscr {U})}{\kappa (\mathscr {U})+1}$ . Now by combining Propositions 6.1 and 5.1, we see that for $p\geq 2$ , the Bergman projection on $\mathscr {U}$ is bounded if and only if $p\in \left [2, \frac {2\kappa (\mathscr {U})}{\kappa (\mathscr {U})-1} \right ).$ By Hölder symmetry, for $p\leq 2$ , the Bergman projection is bounded if and only if $p\in \left (\frac {2\kappa (\mathscr {U})}{\kappa (\mathscr {U})+1},2\right ] $ . The bounds claimed in (1.3) are proved, and therefore Theorem 1.2 holds.

7.2 Comments and questions

The methods used to prove Theorem 1.2 are more general than the result itself, and apply to the Bergman projection on various quotient domains of simple domains with known Bergman kernels, provided the quotient type proper holomorphic map is a monomial map. For example, we may deduce using a modification of our arguments, the range of p for which the Bergman projection is bounded in $L^p$ on the domain

$$ \begin{align*} \{(\left\vert z_1\right\vert{{\hspace{-0.0001pt}}}^2+\dots+\left\vert z_{n-1}\right\vert{{\hspace{-0.0001pt}}}^2)^{\frac{k_1}{2}}<\left\vert z_n\right\vert{{\hspace{-0.0001pt}}}^{k_2}<1\}\subset{\mathbb{C}}^n, \end{align*} $$

where $k_1,k_2$ are positive integers.

Thanks to Theorem 1.2, the Bergman projection is no longer bounded in $L^p(\mathscr {U})$ if $p\geq \frac {2\kappa (\mathscr {U})}{\kappa (\mathscr {U})-1}$ . It is natural to ask if there is an alternate projection from $L^p(\mathscr {U})$ to $A^p(\mathscr {U})$ for such p. In the special case $\mathscr {U}=H_{m/n}$ , the generalized Hartogs triangle of (1.6), it is possible to construct for each $p\geq 2$ a sub-Bergman projection which gives rise to a bounded projection on $L^p$ . It would be interesting to see whether a similar statement holds for the monomial polyhedra considered in this paper.

Finally, we would like to understand the precise geometric significance of the arithmetic complexity of $\mathscr {U}$ without reference to the representation in terms of the matrix B. Such a description will pave the way of generalizing the results of this paper to wider classes of domains.

Acknowledgments

We would like to thank Yuan Yuan for very interesting discussions with DC and LE about this problem and the results of [Reference Chen, Krantz and YuanCKY20] during the 2019 Midwest Several Complex Variables Conference at Dearborn, MI, and Steven Krantz for comments on the same paper made to LE during a visit to Washington University at St. Louis, MO, in the same year. DC would like to thank the mathematics department of the University of Michigan, Ann Arbor for its hospitality in Fall 2019. We also thank Shuo Zhang for pointing out his work in [Reference ZhangZha20, Reference ZhangZha19], and Włodzimierz Zwonek for pointing out his work in [Reference ZwonekZwo99, Reference ZwonekZwo00].

Footnotes

Chase Bender was supported by a Student Research and Creative Endeavors grant from Central Michigan University.

Debraj Chakrabarti was partially supported by National Science Foundation grant DMS-1600371.

References

Axler, S., Bergman spaces and their operators . In: Surveys of some recent results in operator theory, Vol. I, Pitman Res. Notes Math. Ser., 171, Longman Sci. Tech., Harlow, 1988, pp. 150.Google Scholar
Barrett, D. and Şahutoğlu, S., Irregularity of the Bergman projection on worm domains in ${\mathbb{C}}^n$ . Michigan Math. J. 61(2012), no. 1, 187198.CrossRefGoogle Scholar
Barrett, D. E., Irregularity of the Bergman projection on a smooth bounded domain in C2 . Ann. Math. (2) 119(1984), no. 2, 431436.CrossRefGoogle Scholar
Behnke, H., Zur Theorie der Singularitäten der Funktionen mehrerer komplexen Veränderlichen . Math. Ann. 108(1933), no. 1, 91104.CrossRefGoogle Scholar
Bell, S. and Catlin, D., Boundary regularity of proper holomorphic mappings . Duke Math. J. 49(1982), no. 2, 385396.CrossRefGoogle Scholar
Bell, S. R., Proper holomorphic mappings and the Bergman projection . Duke Math. J. 48(1981), no. 1, 167175.CrossRefGoogle Scholar
Bell, S. R., The Bergman kernel function and proper holomorphic mappings . Trans. Amer. Math. Soc. 270(1982), no. 2, 685691.Google Scholar
Catlin, D., Boundary behavior of holomorphic functions on pseudoconvex domains . J. Differential Geom. 15(1980/1981), no. 4, 605625.CrossRefGoogle Scholar
Chakrabarti, D., On an observation of Sibony . Proc. Amer. Math. Soc. 147(2019), no. 8, 34513454.CrossRefGoogle Scholar
Chakrabarti, D., Edholm, L. D., and McNeal, J. D., Duality and approximation of Bergman spaces . Adv. Math. 341(2019), 616656.CrossRefGoogle Scholar
Chakrabarti, D., Konkel, A., Mainkar, M., and Miller, E., Bergman kernels of elementary Reinhardt domains . Pacific J. Math. 306(2020), no. 1, 6793.CrossRefGoogle Scholar
Chakrabarti, D. and Zeytuncu, Y. E., Lp mapping properties of the Bergman projection on the Hartogs triangle . Proc. Amer. Math. Soc. 144(2016), no. 4, 16431653.CrossRefGoogle Scholar
Chen, L., The ${L}^p$ boundedness of the Bergman projection for a class of bounded Hartogs domains . J. Math. Anal. Appl. 448(2017), no. 1, 598610.CrossRefGoogle Scholar
Chen, L., Jin, M., and Yuan, Y., Bergman projection on the symmetrized bidisk. Preprint, 2020. arXiv:2004.02785.Google Scholar
Chen, L., Krantz, S. G., and Yuan, Y., Lp regularity of the Bergman projection on domains covered by the polydisc . J. Funct. Anal. 279(2020), no. 2, 108522.CrossRefGoogle Scholar
Collatz, L., Functional analysis and numerical mathematics. Academic Press, New York, London, 1966. Translated from the German by Hansjörg Oser.Google Scholar
Diederich, K. and Fornaess, J. E., Boundary regularity of proper holomorphic mappings . Invent. Math. 67(1982), no. 3, 363384.CrossRefGoogle Scholar
Dummit, D. S. and Foote, R. M., Abstract algebra. 3rd ed., John Wiley & Sons, Inc., Hoboken, NJ, 2004.Google Scholar
Duren, P. and Schuster, A., Bergman spaces . Mathematical Surveys and Monographs, 100, American Mathematical Society, Providence, RI, 2004.CrossRefGoogle Scholar
Edholm, L. D., Bergman theory of certain generalized Hartogs triangles . Pacific J. Math. 284(2016), no. 2, 327342.CrossRefGoogle Scholar
Edholm, L. D. and McNeal, J. D., The Bergman projection on fat Hartogs triangles: ${L}^p$ boundedness . Proc. Amer. Math. Soc. 144(2016), no. 5, 21852196.CrossRefGoogle Scholar
Edholm, L. D. and McNeal, J. D., Bergman subspaces and subkernels: degenerate ${L}^p$ mapping and zeroes . J. Geom. Anal. 27(2017), no. 4, 26582683.CrossRefGoogle Scholar
Edholm, L. D. and McNeal, J. D., Sobolev mapping of some holomorphic projections . J. Geom. Anal. 30(2020), no. 2, 12931311.CrossRefGoogle Scholar
Forelli, F. and Rudin, W., Projections on spaces of holomorphic functions in balls . Indiana Univ. Math. J. , 24(1974/1975), 593602.CrossRefGoogle Scholar
Grünbaum, B., Convex polytopes. 2nd ed., Graduate Texts in Mathematics, 221, Springer-Verlag, New York, 2003. Prepared and with a preface by Volker Kaibel, Victor Klee and Günter M. Ziegler.CrossRefGoogle Scholar
Hakim, M. and Sibony, N., Spectre de $A\left(\bar{\varOmega}\right)$  pour des domaines bornés faiblement pseudoconvexes réguliers . J. Funct. Anal. 37(1980), no. 2, 127135.CrossRefGoogle Scholar
Hedenmalm, H., The dual of a Bergman space on simply connected domains . J. Anal. Math. 88(2002), 311335.CrossRefGoogle Scholar
Hedenmalm, H., Korenblum, B., and Zhu, K., Theory of Bergman spaces . Graduate Texts in Mathematics, 199, Springer-Verlag, New York, 2000.CrossRefGoogle Scholar
Hindry, M. and Silverman, J. H., Diophantine geometry . Graduate Texts in Mathematics, 201, Springer-Verlag, New York, 2000. An introduction.CrossRefGoogle Scholar
Huo, Z., Lp estimates for the Bergman projection on some Reinhardt domains . Proc. Amer. Math. Soc. 146(2018), no. 6, 25412553.CrossRefGoogle Scholar
Huo, Z. and Wick, B. D., Weak-type estimates for the Bergman projection on the polydisc and the Hartogs triangle. Preprint, 2019. arXiv:1909.05902.CrossRefGoogle Scholar
Jarnicki, M. and Pflug, P., First steps in several complex variables: Reinhardt domains . EMS Textbooks in Mathematics, European Mathematical Society (EMS), Zürich, 2008.CrossRefGoogle Scholar
Kobayashi, S., Geometry of bounded domains . Trans. Amer. Math. Soc. 92(1959), 267290.CrossRefGoogle Scholar
Krantz, S. G., Geometric analysis of the Bergman kernel and metric . Graduate Texts in Mathematics, 268, Springer, New York, 2013.CrossRefGoogle Scholar
Krantz, S. G. and Peloso, M. M., Analysis and geometry on worm domains . J. Geom. Anal. 18(2008), no. 2, 478510.CrossRefGoogle Scholar
Massey, W. S., A basic course in algebraic topology. Graduate Texts in Mathematics, 127, Springer-Verlag, New York, 1991.CrossRefGoogle Scholar
McNeal, J. D. and Stein, E. M., Mapping properties of the Bergman projection on convex domains of finite type . Duke Math. J. 73(1994), no. 1, 177199.CrossRefGoogle Scholar
Misra, G., Roy, S. S., and Zhang, G., Reproducing kernel for a class of weighted Bergman spaces on the symmetrized polydisc . Proc. Amer. Math. Soc. 141(2013), no. 7, 23612370.CrossRefGoogle Scholar
Munkres, J. R., Elements of algebraic topology. Addison-Wesley Publishing Company, Menlo Park, CA, 1984.Google Scholar
Nagel, A. and Pramanik, M., Maximal averages over linear and monomial polyhedra . Duke Math. J. 149(2009), no. 2, 209277.CrossRefGoogle Scholar
Nagel, A. and Pramanik, M., Bergman spaces under maps of monomial type . J. Geometr. Anal. 31(2021), no. 5, 45314560.CrossRefGoogle Scholar
Park, J.-D., The explicit forms and zeros of the Bergman kernel for 3-dimensional Hartogs triangles . J. Math. Anal. Appl. 460(2018), no. 2, 954975.CrossRefGoogle Scholar
Phong, D. H. and Stein, E. M., Estimates for the Bergman and Szegö projections on strongly pseudo-convex domains . Duke Math. J. 44(1977), no. 3, 695704.CrossRefGoogle Scholar
Sibony, N., Prolongement des fonctions holomorphes bornées et métrique de Carathéodory . Invent. Math. 29(1975), no. 3, 205230.CrossRefGoogle Scholar
Vladimirov, V. S., Methods of the theory of functions of many complex variables. The M.I.T. Press, Cambridge, MA, London, 1966. Translated from the Russian by Scripta Technica, Inc. Translation edited by Leon Ehrenpreis.Google Scholar
Zaharjuta, V. P. and Judovic, V. I., The general form of a linear functional on  ${H}_p^{\prime }$ . Uspekhi Mat. Nauk. 19(1964), no. 2, 139142.Google Scholar
Zhang, S., Mapping properties of the Bergman projections on elementary Reinhardt domains. Submitted, 2019.Google Scholar
Zhang, S., Lp boundedness for the Bergman projections over $n$ -dimensional generalized Hartogs triangles . Complex Var. Elliptic Equ. (2020). https://doi.org/10.1080/17476933.2020.1769085.Google Scholar
Ziegler, G. M., Lectures on polytopes. Graduate Texts in Mathematics, 152, Springer-Verlag, New York, 1995.CrossRefGoogle Scholar
Zwonek, W., On hyperbolicity of pseudoconvex Reinhardt domains . Arch. Math. (Basel) 72(1999), no. 4, 304314.CrossRefGoogle Scholar
Zwonek, W., Completeness, Reinhardt domains and the method of complex geodesics in the theory of invariant functions . Dissertationes Math. (Rozprawy Mat.) 388(2000), 103.Google Scholar
Figure 0

Figure 1: Reinhardt shadows of some monomial polyhedra.