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

Algebraicity of the Bergman Kernel

Peter Ebenfelt Department of Mathematics, University of California at San Diego, La Jolla, CA 92093, USA pebenfelt@ucsd.edu Ming Xiao Department of Mathematics, University of California at San Diego, La Jolla, CA 92093, USA m3xiao@ucsd.edu  and  Hang Xu Department of Mathematics, University of California at San Diego, La Jolla, CA 92093, USA h9xu@ucsd.edu
Abstract.

Our main result introduces a new way to characterize two-dimensional finite ball quotients by algebraicity of their Bergman kernels. This characterization is particular to dimension two and fails in higher dimensions, as is illustrated by a counterexample in dimension three constructed in this paper. As a corollary of our main theorem, we prove, e.g., that a smoothly bounded strictly pseudoconvex domain GG in 2\mathbb{C}^{2} has rational Bergman kernel if and only if there is a rational biholomorphism from GG to 𝔹2\mathbb{B}^{2}.

2010 Mathematics Subject Classification:
32A36, 32C20, 32S99
The first and second authors were supported in part by the NSF grants DMS-1900955 and DMS-1800549, respectively.

1. Introduction

The Bergman kernel, introduced by S. Bergman in [6, 7] for domains in n\mathbb{C}^{n} and later cast in differential geometric terms by S. Kobayashi [33], plays a fundamental role in several complex variables and complex geometry. Its biholomorphic invariance properties and intimate connection with the CR geometry of the boundary make it an important tool in the study of open complex manifolds. The use of the Bergman kernel, e.g., in the study of biholomorphic mappings and the geometry of bounded strictly pseudoconvex domains in n\mathbb{C}^{n} was pioneered by C. Fefferman [16, 17, 18], who developed a theory of Bergman kernels in such domains and initiated a now famous program to describe the boundary singularity in terms of the local invariant CR geometry; see also [1], [25] for further progress on Fefferman’s program.

A broad and general problem of foundational importance is that of classifying complex manifolds, or more generally analytic spaces, in terms of their Bergman kernels or Bergman metrics. For example, a well-known result of Q. Lu [36] implies that if a relatively compact domain in an nn-dimensional Kähler manifold has a complete Bergman metric with constant holomorphic sectional curvature, then the domain is biholomorphic to the unit ball 𝔹n\mathbb{B}^{n} in n\mathbb{C}^{n}. Another example is the conjecture of S.-Y. Cheng [12], which states that the Bergman metric of a smoothly bounded strongly pseudoconvex domain in n\mathbb{C}^{n} is Kähler–Einstein (i.e., has Ricci curvature equal to a constant multiple of the metric tensor) if and only if it is biholomorphic to the unit ball 𝔹n\mathbb{B}^{n}. This conjecture was confirmed by Fu-Wong [21] and Nemirovski–Shafikov [38] in the two dimensional case, and in the higher dimensional case by X. Huang and the second author [32].

In this paper, we introduce a new characterization of the two-dimensional unit ball 𝔹22\mathbb{B}^{2}\subset\mathbb{C}^{2} and, more generally, two-dimensional finite ball quotients 𝔹2/Γ\mathbb{B}^{2}/\Gamma in terms of algebraicity of the Bergman kernel. It is interesting, and perhaps surprising then, to note that such a characterization fails in the higher dimensional case. Indeed, in Section 6 below we construct a relatively compact domain GG with smooth strongly pseudoconvex boundary in a three-dimensional algebraic variety V4V\subset\mathbb{C}^{4}, with an isolated normal singularity in the interior of GG, such that the boundary G\partial G is not spherical and, furthermore, GG is not biholomorphic to any finite ball quotient; recall that a CR hypersurface MM of dimension 2n12n-1 is said to be spherical if near each point pMp\in M, it is locally CR diffeomorphic to an open piece of the unit sphere S2n1nS^{2n-1}\subset\mathbb{C}^{n}. Nevertheless, in two dimensions it turns out that algebraicity of the Bergman kernel does characterize finite ball quotients:

Theorem 1.1.

Let VV be a 22-dimensional algebraic variety in N\mathbb{C}^{N}, and GG a relatively compact domain in VV. Assume that every point in G¯\overline{G} is a smooth point of VV except for finitely many isolated normal singularities inside GG, and that GG has a smooth strongly pseudoconvex boundary. Then the Bergman kernel form of GG is algebraic if and only if there is an algebraic branched covering map FF from 𝔹2\mathbb{B}^{2} onto GG, which realizes GG as a ball quotient 𝔹2/Γ\mathbb{B}^{2}/\Gamma where Γ\Gamma is a finite unitary group with no fixed points on 𝔹2\partial\mathbb{B}^{2}.

Remark 1.2.

We note that in addition to showing that Theorem 1.1 fails in dimension 3\geq 3, our example in Section 6 also shows that the Ramanadov Conjecture for the Bergman kernel fails for higher dimensional normal Stein spaces. Recall that the Ramadanov Conjecture (c.f., [40], [14, Question 3]) proposes that if the logarithmic term in Fefferman’s asymptotic expansion [17] of the Bergman kernel vanishes to infinite order at the boundary of a normal reduced Stein space with compact, smooth strongly pseudoconvex boundary, then the boundary is spherical. The Ramadanov Conjecture has been established in two dimensions by the work of D. Burns and R. C. Graham (see [22]). The normal reduced Stein space constructed in Section 6 gives a 33-dimensional counterexample with one isolated singularity. The counterexamples in [14] are smooth, but not Stein.

Theorem 1.1 has two immediate consequences in the non-singular case:

Corollary 1.3.

Let VV be a 22-dimensional algebraic variety in N\mathbb{C}^{N}, and let GG be a relatively compact domain in VV with smooth strongly pseudoconvex boundary. Assume that every point in G¯\overline{G} is a smooth point of VV. Then the Bergman kernel form of GG is algebraic if and only if GG is biholomorphic to 𝔹2\mathbb{B}^{2} by an algebraic map.

Corollary 1.4.

Let GG be a bounded domain in 2\mathbb{C}^{2} with smooth strongly pseudoconvex boundary. Then the Bergman kernel of GG is rational (respectively, algebraic) if and only if there is a rational (respectively, algebraic) biholomorphic map from GG to 𝔹2\mathbb{B}^{2}.

We remark that although Theorem 1.1 fails in higher dimension, Corollary 1.3 and 1.4 might still be true. For instance, it is clear from the proof below of Theorem 1.1 (see Remark 5.3) that if the Ramadanov Conjecture is proved to hold for, e.g., strongly pseudoconvex bounded domains in n\mathbb{C}^{n}, which is still a possibility despite Remark 1.2 above, then Corollary 1.4 also holds in n\mathbb{C}^{n}.

We also remark that the rationality of the biholomorphic map G𝔹2G\to\mathbb{B}^{2} in Corollary 1.4, once its existence has been established, follows from the work of S. Bell [5]. For the reader’s convenience, a self-contained proof of the rationality is given in Section 5.

As a final remark in this introduction, we note that, by Lempert’s algebraic approximation theorem [35], if GG is a relatively compact domain in a reduced Stein space XX with only isolated singularities, then there exist an affine algebraic variety VV, a domain ΩV\Omega\subset V, and a biholomorphism FF from a neighborhood of G¯\overline{G} to a neighborhood of Ω¯\overline{\Omega} with F(Ω)=GF(\Omega)=G. We shall say such a domain Ω\Omega is an algebraic realization of GG. Theorem 1.1 implies the following corollary.

Corollary 1.5.

Let GG be a relatively compact domain in a 22-dimensional reduced Stein space XX with smooth strongly pseudoconvex boundary and only isolated normal singularities. If GG has an algebraic realization with an algebraic Bergman kernel, then GG is biholomorphic to a ball quotient 𝔹2/Γ\mathbb{B}^{2}/\Gamma, where Γ\Gamma is a finite unitary group with no fixed point on 𝔹2\partial\mathbb{B}^{2}.

To prove the "only if" implication in Theorem 1.1, we use the asymptotic boundary behavior of the Bergman kernel to establish algebraicity and sphericity of the boundary of GG. Fefferman’s asymptotic expansion [17] and the Riemann mapping type theorems due to X. Huang–S. Ji ([29]) and X. Huang ([28]) play important roles in the proof. To prove the converse ("if") implication in the theorem, we will need to compute the Bergman kernel forms of finite ball quotients. In order to do so, we shall establish a transformation formula for (possibly branched) covering maps of complex analytic spaces. This formula generalizes a classical theorem of Bell ([3], [4]):

Theorem 1.6.

Let M1M_{1} and M2M_{2} be two complex analytic sets. Let V1M1V_{1}\subset M_{1} and V2M2V_{2}\subset M_{2} be proper analytic subvarieties such that M1V1,M2V2M_{1}-V_{1},M_{2}-V_{2} are complex manifolds of the same dimension. Assume that f:M1V1M2V2f:M_{1}-V_{1}\rightarrow M_{2}-V_{2} is a finite (mm-sheeted) holomorphic covering map. Let Γ\Gamma be the deck transformation group for the covering map (with |Γ|=m|\Gamma|=m), and denote by Ki(z,w¯)K_{i}(z,\bar{w}) the Bergman kernels of MiM_{i} for i=1,2i=1,2. Then the Bergman kernel forms transform according to

(1.1) γΓ(γ,id)K1=γΓ(id,γ)K1=(f,f)K2on(M1V1)×(M1V1),\sum_{\gamma\in\Gamma}(\gamma,\mathrm{id})^{*}K_{1}=\sum_{\gamma\in\Gamma}(\mathrm{id},\gamma)^{*}K_{1}=(f,f)^{*}K_{2}\quad\leavevmode\nobreak\ \text{on}\leavevmode\nobreak\ (M_{1}-V_{1})\times(M_{1}-V_{1}),

where id:M1M1\mathrm{id}:M_{1}\rightarrow M_{1} is the identity map.

See Section 2 for the notation used in the formula in Theorem 1.6. We expect that this formula will be useful in other applications as well. In an upcoming paper [13], the authors apply it to study the question of when the Bergman metric of a finite ball quotient is Kähler–Einstein. (This is always the case for finite disk quotients, i.e., one-dimensional ball quotients, by recent work of X. Huang and X. Li [31].)

The paper is organized as follows. Section 2 gives some preliminaries on algebraic functions and Bergman kernels of complex analytic spaces. Section 3 is devoted to establishing the transformation formula in Theorem 1.6. Then in Section 4 we apply it to show that every standard algebraic realization (in particular, Cartan’s canonical realization) of a finite ball quotient must have algebraic Bergman kernel, and thus prove the "if" implication in Theorem 1.1. Section 5 gives the proof of the "only if" implication in Theorem 1.1, as well as those of Corollaries 1.3 and 1.4. In Section 6 and Appendix 7, we construct the counterexample mentioned above to the corresponding statement of Theorem 1.1 in higher dimensions.

Acknowledgment. The second author thanks Xiaojun Huang for many inspiring conversations on quotient singularities.

2. Preliminaries

2.1. Algebraic Functions

In this subsection, we will review some basic facts about algebraic functions. For more details, we refer the readers to [2, Chapter 5.4] and [30].

Definition 2.1 (Algebraic functions and maps).

Let 𝕂\mathbb{K} be the field \mathbb{R} or \mathbb{C}. Let U𝕂nU\subset\mathbb{K}^{n} be a domain. A 𝕂\mathbb{K}-analytic function f:U𝕂f:U\rightarrow\mathbb{K} is said to be 𝕂\mathbb{K}-algebraic (i.e., real/complex-algebraic) on UU if there is a non-trivial polynomial P(x,y)𝕂[x,y]P(x,y)\in\mathbb{K}[x,y], with (x,y)𝕂n×𝕂(x,y)\in\mathbb{K}^{n}\times\mathbb{K}, such that P(x,f(x))=0P(x,f(x))=0 for all xUx\in U. We say that a 𝕂\mathbb{K}-analytic map F:UNF:U\rightarrow\mathbb{C}^{N} is 𝕂\mathbb{K}-algebraic if each of its components is so on UU.

Remark 2.2.

We make two remarks:

  • (i)

    If f(x)f(x) is an 𝕂\mathbb{K}-analytic function in a domain U𝕂nU\subset\mathbb{K}^{n}, then ff is 𝕂\mathbb{K}-algebraic if and only if it is 𝕂\mathbb{K}-algebraic in some neighborhood of any point x0Ux_{0}\in U.

  • (ii)

    If f(x)f(x) is an \mathbb{R}-analytic function in a domain UnU\subset\mathbb{R}^{n}, then there is domain U^n\hat{U}\subset\mathbb{C}^{n} containing UnnU\subset\mathbb{R}^{n}\subset\mathbb{C}^{n} and a \mathbb{C}-analytic (i.e., holomorphic) function g(x+iy)g(x+iy) in U^\hat{U} such that f=g|Uf=g|_{U}; i.e., f(x)=g(x)f(x)=g(x) for xUx\in U. Moreover, ff is \mathbb{R}-algebraic if and only if gg is \mathbb{C}-algebraic.

We say a differential form on Un2nU\subset\mathbb{C}^{n}\cong\mathbb{R}^{2n} is real-algebraic if each of its coefficient functions is so. We can also define real-algebraicity of a differential form on an affine (algebraic) variety.

Definition 2.3.

Let VNV\subset\mathbb{C}^{N} be an affine variety and write RegV\operatorname{Reg}V for the set of its regular points. Let ϕ\phi be a real analytic differential form on RegV\operatorname{Reg}V. We say ϕ\phi is real-algebraic on VV if for every point z0RegVz_{0}\in\operatorname{Reg}V, there exists a real-algebraic differential form ψ\psi in a neighborhood UU of z0z_{0} in N2N\mathbb{C}^{N}\cong\mathbb{R}^{2N} such that

ψ|V=ϕ, on UV.\psi|_{V}=\phi,\quad\mbox{ on }U\cap V.

Let Tz0VTz01,0VT_{z_{0}}V\cong T^{1,0}_{z_{0}}V be the complex tangent space of VV at a smooth point z0Vz_{0}\in V considered as an affine complex subspace in n\mathbb{C}^{n} through z0z_{0}, and let ξ=(ξ1,,ξn)\xi=(\xi_{1},\cdots,\xi_{n}) be affine coordinates for Tz0VT_{z_{0}}V. Since VV can be realized locally as a graph over Tz0VT_{z_{0}}V, the real and imaginary parts of ξ\xi also serve as local real coordinates for VV near z0z_{0}. We call such coordinates the canonical extrinsic coordinates at z0z_{0}. Then the following statements are equivalent.

  • (a)

    ϕ\phi is real-algebraic on RegV\operatorname{Reg}V (in the sense of Definition 2.3).

  • (b)

    For any z0RegVz_{0}\in\operatorname{Reg}V, ϕ\phi is real-algebraic in canonical extrinsic coordinates at z0z_{0}.

If in addition, there is a domain GnG\subset\mathbb{C}^{n} and a \mathbb{C}-algebraic (i.e., holomorphic algebraic) immersion f:GNf:G\rightarrow\mathbb{C}^{N} such that f(G)=RegVf(G)=\operatorname{Reg}V, then (a) and (b) are further equivalent to

  • (c)

    fϕf^{*}\phi is real-algebraic on GG.

Remark 2.4.

We can define complex-algebraicity of (p,0)(p,0)-forms, p>0,p>0, on an complex affine (algebraic) variety in a similar manner as in Definition 2.3.

2.2. The Bergman Kernel

In this section, we will briefly review some properties of the Bergman kernel on a complex manifold. More details can be found in [34].

Let MM be an n-dimensional complex manifold. Write L(n,0)2(M)L^{2}_{(n,0)}(M) for the space of L2L^{2}-integrable (n,0)(n,0) forms on M,M, which is equipped with the following inner product:

(2.1) (φ,ψ)L2(M):=in2Mφψ¯,φ,ψL(n,0)2(M),(\varphi,\psi)_{L^{2}(M)}:=i^{n^{2}}\int_{M}\varphi\wedge\overline{\psi},\quad\varphi,\psi\in L^{2}_{(n,0)}(M),

Define the Bergman space of MM to be

(2.2) A(n,0)2(M):={φL(n,0)2(M):φ is a holomorphic (n,0) form on M}.A^{2}_{(n,0)}(M):=\bigl{\{}\varphi\in L^{2}_{(n,0)}(M):\varphi\mbox{ is a holomorphic $(n,0)$ form on $M$}\}.

Assume A(n,0)2(M){0}A^{2}_{(n,0)}(M)\neq\{0\}. Then A(n,0)2(M)A^{2}_{(n,0)}(M) is a separable Hilbert space. Taking any orthonormal basis {φk}k=1q\{\varphi_{k}\}_{k=1}^{q} of A(n,0)2(M)A^{2}_{(n,0)}(M) (here 1q1\leq q\leq\infty), we define the Bergman kernel (form) of MM to be

KM(x,y¯)=in2k=1qφk(x)φk(y)¯.K_{M}(x,\bar{y})=i^{n^{2}}\sum_{k=1}^{q}\varphi_{k}(x)\wedge\overline{\varphi_{k}(y)}.

Then, KM(x,x¯)K_{M}(x,\bar{x}) is a real-valued, real analytic form of degree (n,n)(n,n) on MM and is independent of the choice of orthonormal basis. When MM is also (the set of regular points on) an affine variety, we say that the Bergman kernel of MM is algebraic if KM(x,x¯)K_{M}(x,\bar{x}) is real-algebraic in the sense of Definition 2.3. The following definitions and facts are standard in literature.

Definition 2.5 (Bergman projection).

Given gL(n,0)2(M)g\in L^{2}_{(n,0)}(M), we define for xMx\in M

Pg(x)=Mg(ζ)KM(x,ζ¯):=in2k=1q(Mg(ζ)φk(ζ)¯)φk(x).Pg(x)=\int_{M}g(\zeta)\wedge K_{M}(x,\bar{\zeta}):=i^{n^{2}}\sum_{k=1}^{q}\Bigl{(}\int_{M}g(\zeta)\wedge\overline{\varphi_{k}(\zeta)}\Bigr{)}\varphi_{k}(x).

P:L(n,0)2A(n,0)2(M)P\colon L^{2}_{(n,0)}\to A^{2}_{(n,0)}(M) is called the Bergman projection, and is the orthogonal projection to the Bergman space A(n,0)2(M)A^{2}_{(n,0)}(M).

The Bergman kernel form remains unchanged if we remove a proper complex analytic subvariety. The following theorem is from [33].

Theorem 2.6 ([33]).

If MM^{\prime} is a domain in an nn-dimensional complex manifold MM and if MMM-M^{\prime} is a complex analytic subvariety of MM of complex dimension n1\leq n-1, then

KM(x,y¯)=KM(x,y¯) for any y M.K_{M}(x,\bar{y})=K_{M^{\prime}}(x,\bar{y})\quad\mbox{ for any y }\in M^{\prime}.

This theorem suggests the following generalization of the Bergman kernel form to complex analytic spaces.

Definition 2.7.

Let MM be a reduced complex analytic space, and let VMV\subset M denote its set of singular points. The Bergman kernel form of MM is defined as

KM(x,y¯)=KMV(x,y¯) for any x,yMV,K_{M}(x,\bar{y})=K_{M-V}(x,\bar{y})\quad\mbox{ for any }x,y\in M-V,

where KMVK_{M-V} denotes the Bergman kernel form of the complex manifold consisting of regular points of MM.

Let N1,N2N_{1},N_{2} be two complex manifolds of dimension nn. Let γ:N1M\gamma:N_{1}\rightarrow M and τ:N2M\tau:N_{2}\rightarrow M be holomorphic maps. The pullback of the Bergman kernel KM(x,y¯)K_{M}(x,\bar{y}) of MM to N1×N2N_{1}\times N_{2} is defined in the standard way. That is, for any zN1,wN2z\in N_{1},w\in N_{2},

((γ,τ)K)(z,w¯)=k=1qγφk(z)τφk(w)¯.\bigl{(}(\gamma,\tau)^{*}K\bigr{)}(z,\bar{w})=\sum_{k=1}^{q}\gamma^{*}\varphi_{k}(z)\wedge\overline{\tau^{*}\varphi_{k}(w)}.

In terms of local coordinates, writing the Bergman kernel form of MM as

KM(x,y¯)=K~(x,y¯)dx1dxndy1¯dyn¯,K_{M}(x,\bar{y})=\widetilde{K}(x,\bar{y})dx_{1}\wedge\cdots dx_{n}\wedge d\overline{y_{1}}\wedge\cdots\wedge d\overline{y_{n}},

we have

((γ,τ)KM)(z,w¯)=K~(γ(z),τ(w)¯)Jγ(z)Jτ(w)¯dz1dzndw1¯dwn¯,\bigl{(}(\gamma,\tau)^{*}K_{M}\bigr{)}(z,\bar{w})=\widetilde{K}(\gamma(z),\overline{\tau(w)})\,J_{\gamma}(z)\,\overline{J_{\tau}(w)}\,dz_{1}\wedge\cdots dz_{n}\wedge d\overline{w_{1}}\wedge\cdots\wedge d\overline{w_{n}},

where JγJ_{\gamma} and JτJ_{\tau} are the Jacobian determinants of the maps γ\gamma and τ\tau, respectively.

3. The transformation law for the Bergman kernel

In this section, we shall prove Theorem 1.6. For this, we shall adapt the ideas in [4] to our situation. More precisely, we shall first prove the following transformation law for the Bergman projections. Then (1.1) will follow readily by comparing the associated distributional kernels for the projection operators.

Proposition 3.1.

Under the assumptions and notation in Theorem 1.6, we denote by nn the complex dimension of M1V1M_{1}-V_{1} and M2V2M_{2}-V_{2}. Let Pi:L(n,0)2(MiVi)A(n,0)2(MiVi)P_{i}:L^{2}_{(n,0)}(M_{i}-V_{i})\rightarrow A^{2}_{(n,0)}(M_{i}-V_{i}) e the Bergman projection for i=1,2i=1,2. Then the Bergman projections transform according to

(3.1) P1(fϕ)=f(P2ϕ) for any ϕL(n,0)2(M2V2).P_{1}(f^{*}\phi)=f^{*}(P_{2}\phi)\quad\mbox{ for any }\phi\in L^{2}_{(n,0)}(M_{2}-V_{2}).

We first check that fϕL(n,0)2(M1V1)f^{*}\phi\in L^{2}_{(n,0)}(M_{1}-V_{1}) if ϕL(n,0)2(M2V2)\phi\in L^{2}_{(n,0)}(M_{2}-V_{2}) in the next lemma. Recall that ff is an mm-sheeted covering map M1V1M2V2M_{1}-V_{1}\rightarrow M_{2}-V_{2}.

Lemma 3.2.
fϕL2(M1V1)=m12ϕL2(M2V2) for any ϕL(n,0)2(M2V2).\|f^{*}\phi\|_{L^{2}(M_{1}-V_{1})}=m^{\frac{1}{2}}\|\phi\|_{L^{2}(M_{2}-V_{2})}\quad\mbox{ for any }\phi\in L^{2}_{(n,0)}(M_{2}-V_{2}).
Proof.

Let {Uj}\{U_{j}\} be a countable, locally finite open cover of M2V2M_{2}-V_{2} such that

  • each UjU_{j} is relatively compact;

  • f1(Uj)=k=1mVj,kf^{-1}(U_{j})=\cup_{k=1}^{m}V_{j,k} for some pairwise disjoint open sets {Vj,k}k=1m\{V_{j,k}\}_{k=1}^{m} on M1V1M_{1}-V_{1};

  • f:Vj,kUjf:V_{j,k}\rightarrow U_{j} is a biholomorphsm for each j=1,2,mj=1,2,\cdots m.

Let {ρj}\{\rho_{j}\} be a partition of unity subordinate to the cover {Uj}\{U_{j}\}. Then

in2M2V2ϕϕ¯=jin2Ujρjϕϕ¯=1mjk=1min2Vj,k(fρj)fϕfϕ¯.\displaystyle i^{n^{2}}\int_{M_{2}-V_{2}}\phi\wedge\overline{\phi}=\sum_{j}i^{n^{2}}\int_{U_{j}}\rho_{j}\phi\wedge\overline{\phi}=\frac{1}{m}\sum_{j}\sum_{k=1}^{m}i^{n^{2}}\int_{V_{j,k}}(f^{*}\rho_{j})\,f^{*}\phi\wedge\overline{f^{*}\phi}.

Note that {fρj}\{f^{*}\rho_{j}\} is a partition of unity subordinate to the countable, locally finite open cover {k=1mVj,k}\{\cup_{k=1}^{m}V_{j,k}\} of M1V1M_{1}-V_{1}. Thus,

1mjk=1min2Vj,k(fρj)fϕfϕ¯=\displaystyle\frac{1}{m}\sum_{j}\sum_{k=1}^{m}i^{n^{2}}\int_{V_{j,k}}(f^{*}\rho_{j})\,f^{*}\phi\wedge\overline{f^{*}\phi}= 1mjin2k=1mVj,k(fρj)fϕfϕ¯\displaystyle\frac{1}{m}\sum_{j}i^{n^{2}}\int_{\cup_{k=1}^{m}V_{j,k}}(f^{*}\rho_{j})\,f^{*}\phi\wedge\overline{f^{*}\phi}
=\displaystyle= 1min2M1V1fϕfϕ¯.\displaystyle\frac{1}{m}i^{n^{2}}\int_{M_{1}-V_{1}}f^{*}\phi\wedge\overline{f^{*}\phi}.

The result therefore follows immediately. ∎

Let F1,F2,,FmF_{1},F_{2},\cdots,F_{m} be the mm local inverses to ff defined locally on M2V2M_{2}-V_{2}. Note that k=1mFk\sum_{k=1}^{m}F_{k}^{*} is a well-defined operator on L(n,0)2(M1V1)L_{(n,0)}^{2}(M_{1}-V_{1}), though each individual FkF_{k} is only locally defined.

Lemma 3.3.

Let vL(n,0)2(M1V1)v\in L^{2}_{(n,0)}(M_{1}-V_{1}) and ϕL(n,0)2(M2V2)\phi\in L^{2}_{(n,0)}(M_{2}-V_{2}). Then k=1mFk(v)L(n,0)2(M2V2)\sum_{k=1}^{m}F_{k}^{*}(v)\in L^{2}_{(n,0)}(M_{2}-V_{2}) and

(3.2) (v,fϕ)L2(M1V1)=(k=1mFk(v),ϕ)L2(M2V2).\bigl{(}v,f^{*}\phi\bigr{)}_{L^{2}(M_{1}-V_{1})}=\bigl{(}\sum_{k=1}^{m}F_{k}^{*}(v),\phi\bigr{)}_{L^{2}(M_{2}-V_{2})}.
Proof.

We first verify k=1mFk(v)L(n,0)2(M2V2)\sum_{k=1}^{m}F_{k}^{*}(v)\in L^{2}_{(n,0)}(M_{2}-V_{2}). For that we note

fk=1mFk(v)=γΓγv.f^{*}\sum_{k=1}^{m}F_{k}^{*}(v)=\sum_{\gamma\in\Gamma}\gamma^{*}v.

By the same argument as in Lemma 3.2, we have

(3.3) γΓγvL2(M1V1)=m12k=1mFk(v)L2(M2V2).\bigl{\|}\sum_{\gamma\in\Gamma}\gamma^{*}v\bigr{\|}_{L^{2}(M_{1}-V_{1})}=m^{\frac{1}{2}}\bigl{\|}\sum_{k=1}^{m}F_{k}^{*}(v)\bigr{\|}_{L^{2}(M_{2}-V_{2})}.

Since each deck transformation γ:M1V1M1V1\gamma:M_{1}-V_{1}\rightarrow M_{1}-V_{1} is biholomorphic, it follows that

γΓγvL2(M1V1)γΓγvL2(M1V1)=mvL2(M1V1).\displaystyle\bigl{\|}\sum_{\gamma\in\Gamma}\gamma^{*}v\bigr{\|}_{L^{2}(M_{1}-V_{1})}\leq\sum_{\gamma\in\Gamma}\bigl{\|}\gamma^{*}v\bigr{\|}_{L^{2}(M_{1}-V_{1})}=m\|v\|_{L^{2}(M_{1}-V_{1})}.

Therefore by (3.3), k=1mFk(v)L(n,0)2(M2V2)\sum_{k=1}^{m}F_{k}^{*}(v)\in L_{(n,0)}^{2}(M_{2}-V_{2}).

Now we are ready to prove (3.2). Let {Uj}\{U_{j}\}, {Vj,k}\{V_{j,k}\} and {ρj}\{\rho_{j}\} be the open covers and partition of unity as in Lemma 3.2. Then

(k=1mFk(v),ϕ)L2(M2V2)=jin2Ujρjk=1mFk(v)ϕ¯.\displaystyle\bigl{(}\sum_{k=1}^{m}F_{k}^{*}(v),\phi\bigr{)}_{L^{2}(M_{2}-V_{2})}=\sum_{j}i^{n^{2}}\int_{U_{j}}\rho_{j}\sum_{k=1}^{m}F_{k}^{*}(v)\wedge\overline{\phi}.

Note that every Fk:UjVj,kF_{k}:U_{j}\rightarrow V_{j,k} is biholomorphic and the inverse of f:Vj,kUjf:V_{j,k}\rightarrow U_{j}. Thus,

jin2Ujρjk=1mFk(v)ϕ¯=jk=1min2Vj,k(fρj)vfϕ¯=(v,fϕ)L2(M1V1).\displaystyle\sum_{j}i^{n^{2}}\int_{U_{j}}\rho_{j}\sum_{k=1}^{m}F_{k}^{*}(v)\wedge\overline{\phi}=\sum_{j}\sum_{k=1}^{m}i^{n^{2}}\int_{V{j,k}}(f^{*}\rho_{j})v\wedge\overline{f^{*}\phi}=(v,f^{*}\phi)_{L^{2}(M_{1}-V_{1})}.

The last equality follows from the fact that {fρj}\{f^{*}\rho_{j}\} is a partition of unity subordinate to the countable, locally finite open cover {k=1mVj,k}\{\cup_{k=1}^{m}V_{j,k}\} of M1V1M_{1}-V_{1}. This proves (3.2). ∎

We are now ready to prove Proposition 3.1.

Proof of Proposition 3.1.

If ϕA(n,0)2(M2V2)\phi\in A^{2}_{(n,0)}(M_{2}-V_{2}), then fϕA(n,0)2(M1V1)f^{*}\phi\in A^{2}_{(n,0)}(M_{1}-V_{1}) by Lemma 3.2, whence (3.1) holds trivially. It thus suffices to prove (3.1) for ϕA(n,0)2(M2V2)\phi\in A^{2}_{(n,0)}(M_{2}-V_{2})^{\perp}. In this case, (3.1) reduces to

P1(fϕ)=0 for any ϕA(n,0)2(M2V2);P_{1}(f^{*}\phi)=0\quad\mbox{ for any }\phi\in A^{2}_{(n,0)}(M_{2}-V_{2})^{\perp};

i.e., ϕA(n,0)2(M2V2)\phi\in A^{2}_{(n,0)}(M_{2}-V_{2})^{\perp} implies that fϕA(n,0)2(M1V1)f^{*}\phi\in A^{2}_{(n,0)}(M_{1}-V_{1})^{\perp}. To prove this, we note that for any vA(n,0)2(M1V1)v\in A^{2}_{(n,0)}(M_{1}-V_{1}), we have by Lemma 3.3

(v,fϕ)L2(M1V1)=(k=1mFk(v),ϕ)L2(M2V2)=0.\displaystyle\bigl{(}v,f^{*}\phi)_{L^{2}(M_{1}-V_{1})}=\bigl{(}\sum_{k=1}^{m}F_{k}^{*}(v),\phi\bigr{)}_{L^{2}(M_{2}-V_{2})}=0.

The last equality follows from the fact ϕA(n,0)2(M2V2)\phi\in A^{2}_{(n,0)}(M_{2}-V_{2})^{\perp}. Thus, fϕA(n,0)2(M1V1)f^{*}\phi\in A^{2}_{(n,0)}(M_{1}-V_{1})^{\perp} and the proof is completed. ∎

We are now in a position to prove Theorem 1.6.

Proof of Theorem 1.6.

Let idMi\mathrm{id}_{M_{i}} be the identity map on MiM_{i} for i=1,2i=1,2. Recall that {Fk}k=1m\{F_{k}\}_{k=1}^{m} are local inverses of ff. Note that k=1m(idM1,Fk)K1\sum_{k=1}^{m}(\mathrm{id}_{M_{1}},F_{k})^{*}K_{1} is a well-defined (n,n)(n,n) form on (M1V1)×(M2V2)(M_{1}-V_{1})\times(M_{2}-V_{2}) though each (idM1,Fk)K1(\mathrm{id}_{M_{1}},F_{k})^{*}K_{1} is only locally defined.

We shall write out the Bergman projection transformation law (3.1) in terms of integrals of the Bergman kernel forms. For any ϕL(n,0)2(M2V2)\phi\in L^{2}_{(n,0)}(M_{2}-V_{2}), by Lemma 3.3 we have for any zM1V1z\in M_{1}-V_{1},

P1(fϕ)(z)=M1V1fϕ(η)K1(z,η)=M2V2ϕ(η)k=1m(idM1,Fk)K1(z,η).\displaystyle P_{1}(f^{*}\phi)(z)=\int_{M_{1}-V_{1}}f^{*}\phi(\eta)\wedge K_{1}(z,\eta)=\int_{M_{2}-V_{2}}\phi(\eta)\wedge\sum_{k=1}^{m}(\mathrm{id}_{M_{1}},F_{k})^{*}K_{1}(z,\eta).

On the other hand,

P2(ϕ)(ξ)=M2V2ϕ(η)K2(ξ,η) for any ξM2V2.\displaystyle P_{2}(\phi)(\xi)=\int_{M_{2}-V_{2}}\phi(\eta)\wedge K_{2}(\xi,\eta)\quad\mbox{ for any }\xi\in M_{2}-V_{2}.

If we pull back the forms on both sides by ff, then

fP2(ϕ)(z)=M2V2ϕ(η)(f,idM2)K2(z,η) for any zM1V1.\displaystyle f^{*}P_{2}(\phi)(z)=\int_{M_{2}-V_{2}}\phi(\eta)\wedge(f,\mathrm{id}_{M_{2}})^{*}K_{2}(z,\eta)\quad\mbox{ for any }z\in M_{1}-V_{1}.

Therefore, the Bergman projection transformation law (3.1) translates to

M2V2ϕ(η)k=1m(idM1,Fk)K1(z,η)=M2V2ϕ(η)(f,idM2)K2(z,η).\displaystyle\int_{M_{2}-V_{2}}\phi(\eta)\wedge\sum_{k=1}^{m}(\mathrm{id}_{M_{1}},F_{k})^{*}K_{1}(z,\eta)=\int_{M_{2}-V_{2}}\phi(\eta)\wedge(f,\mathrm{id}_{M_{2}})^{*}K_{2}(z,\eta).

As this equality holds for any ϕL(n,0)2(M2V2)\phi\in L^{2}_{(n,0)}(M_{2}-V_{2}), it follows that for any zM1V1z\in M_{1}-V_{1} and ηM2V2\eta\in M_{2}-V_{2},

(3.4) k=1m(idM1,Fk)K1(z,η)=(f,idM2)K2(z,η).\sum_{k=1}^{m}(\mathrm{id}_{M_{1}},F_{k})^{*}K_{1}(z,\eta)=(f,\mathrm{id}_{M_{2}})^{*}K_{2}(z,\eta).

If we further pull back the forms on both sides by (idM1,f):(M1V1)×(M1V1)(M1V1)×(M2V2)(\mathrm{id}_{M_{1}},f):(M_{1}-V_{1})\times(M_{1}-V_{1})\rightarrow(M_{1}-V_{1})\times(M_{2}-V_{2}), then we obtain for z,wM1V1z,w\in M_{1}-V_{1},

(3.5) k=1m(idM1,Fkf)K1(z,w)=(f,f)K2(z,w).\displaystyle\sum_{k=1}^{m}(\mathrm{id}_{M_{1}},F_{k}\circ f)^{*}K_{1}(z,w)=(f,f)^{*}K_{2}(z,w).

By using the notation γk\gamma_{k} for the deck transformation FkfF_{k}\circ f, we may write this as

(3.6) k=1m(idM1,γk)K1(z,w)=(f,f)K2(z,w).\displaystyle\sum_{k=1}^{m}(\mathrm{id}_{M_{1}},\gamma_{k})^{*}K_{1}(z,w)=(f,f)^{*}K_{2}(z,w).

Note that

k=1m(idM1,γk)K1(z,w)=k=1m(γkγk1,γkidM1)K1(z,w)=k=1m(γk1,idM1)(γk,γk)K1(z,w).\displaystyle\sum_{k=1}^{m}(\mathrm{id}_{M_{1}},\gamma_{k})^{*}K_{1}(z,w)=\sum_{k=1}^{m}(\gamma_{k}\circ\gamma_{k}^{-1},\gamma_{k}\circ\mathrm{id}_{M_{1}})^{*}K_{1}(z,w)=\sum_{k=1}^{m}(\gamma_{k}^{-1},\mathrm{id}_{M_{1}})^{*}(\gamma_{k},\gamma_{k})^{*}K_{1}(z,w).

Since γk\gamma_{k} is a biholomorphism on M1V1M_{1}-V_{1}, we have

(γk,γk)K1(z,w)=K1(z,w),(\gamma_{k},\gamma_{k})^{*}K_{1}(z,w)=K_{1}(z,w),

and hence

k=1m(idM1,γk)K1(z,w)=k=1m(γk1,idM1)K1(z,w)=k=1m(γk,idM1)K1(z,w).\displaystyle\sum_{k=1}^{m}(\mathrm{id}_{M_{1}},\gamma_{k})^{*}K_{1}(z,w)=\sum_{k=1}^{m}(\gamma_{k}^{-1},\mathrm{id}_{M_{1}})^{*}K_{1}(z,w)=\sum_{k=1}^{m}(\gamma_{k},\mathrm{id}_{M_{1}})^{*}K_{1}(z,w).

Theorem 1.6 now follows by combining the above identity with (3.6). ∎

4. Proof of Theorem 1.1, part I: Bergman kernels of ball quotients

In this section, we will apply the transformation law in Theorem 1.6 to study the Bergman kernel form of a finite ball quotient and prove the "if" implication in Theorem 1.1. For this part, the restriction of the dimension of the algebraic variety to two is not needed, and we shall therefore consider the situation in an arbitrary dimension nn.

Let 𝔹n\mathbb{B}^{n} denote the unit ball in n\mathbb{C}^{n} and Aut(𝔹n)\text{Aut}(\mathbb{B}^{n}) its (biholomorphic) automorphism group. Let Γ\Gamma be a finite subgroup of Aut(𝔹n)\text{Aut}(\mathbb{B}^{n}). As the unitary group U(n)U(n) is a maximal compact subgroup of Aut(𝔹n)\operatorname{Aut}(\mathbb{B}^{n}), by basic Lie group theory, there exists some ψAut(𝔹n)\psi\in\operatorname{Aut}(\mathbb{B}^{n}) such that Γψ1U(n)ψ\Gamma\subset\psi^{-1}\cdot U(n)\cdot\psi. Thus without loss of generality, we can assume ΓU(n)\Gamma\subset U(n), i.e., Γ\Gamma is a finite unitary group. Note that the origin 0n0\in\mathbb{C}^{n} is always a fixed point of every element in Γ\Gamma. We say Γ\Gamma is fixed point free if every γΓ{id}\gamma\in\Gamma-\{\mathrm{id}\} has no other fixed point, or equivalently, if every γΓ{id}\gamma\in\Gamma-\{\mathrm{id}\} has no fixed point on 𝔹n\partial\mathbb{B}^{n}. In this case, the action of Γ\Gamma on 𝔹n\partial\mathbb{B}^{n} is properly discontinuous and 𝔹n/Γ\partial\mathbb{B}^{n}/\Gamma is a smooth manifold.

By a theorem of Cartan [11], the quotient n/Γ\mathbb{C}^{n}/\Gamma can be realized as a normal algebraic subvariety VV in some N\mathbb{C}^{N}. To be more precise, we write 𝒜\mathcal{A} for the algebra of Γ\Gamma invariant holomorphic polynomials, that is,

𝒜:={p[z1,,zn]:pγ=p for all γΓ}.\mathcal{A}:=\big{\{}p\in\mathbb{C}[z_{1},\cdots,z_{n}]:p\circ\gamma=p\,\mbox{ for all }\gamma\in\Gamma\big{\}}.

By Hilbert’s basis theorem, 𝒜\mathcal{A} is finitely generated. Moreover, we can find a minimal set of homogeneous polynomials {p1,,pN}\{p_{1},\cdots,p_{N}\} such that every p𝒜p\in\mathcal{A} can be expressed in the form

p(z)=q(p1(z),,pN(z)) for zn,p(z)=q(p_{1}(z),\cdots,p_{N}(z))\quad\mbox{ for }z\in\mathbb{C}^{n},

where qq is some holomorphic polynomial in N\mathbb{C}^{N}. The map Q:=(p1,,pN):nNQ:=(p_{1},\cdots,p_{N}):\mathbb{C}^{n}\rightarrow\mathbb{C}^{N} is proper and induces a homeomorphism of n/Γ\mathbb{C}^{n}/\Gamma onto V:=Q(n)V:=Q(\mathbb{C}^{n}). By Remmert’s proper mapping theorem (see [24]), VV is an analytic variety. As QQ is a polynomial holomorphic map, VV is furthermore an algebraic variety. The restriction of QQ to the unit ball 𝔹n\mathbb{B}^{n} maps 𝔹n\mathbb{B}^{n} properly onto a relatively compact domain GVG\subset V. In this way, 𝔹n/Γ\mathbb{B}^{n}/\Gamma is realized as GG by QQ. Following [41], we call such QQ the basic map associated to Γ\Gamma. The ball quotient G=𝔹n/ΓG=\mathbb{B}^{n}/\Gamma is nonsingular if and only if the group Γ\Gamma is generated by reflections, i.e., elements of finite order in U(n)U(n) that fix a complex subspace of dimension n1n-1 in n\mathbb{C}^{n} (see [41]); thus, if Γ\Gamma is fixed point free and nontrivial, then G=𝔹n/ΓG=\mathbb{B}^{n}/\Gamma must have singularities. Moreover, GG has smooth boundary if and only if Γ\Gamma is fixed point free (see [19] for more results along this line).

We are now in a position to state the following theorem, which implies the "if" implication in Theorem 1.1.

Theorem 4.1.

Let GG be a domain in an algebraic variety VV in N\mathbb{C}^{N} and ΓU(n)\Gamma\subset U(n) a finite unitary subgroup with |Γ|=m|\Gamma|=m. Suppose there exist proper complex analytic varieties V1𝔹nV_{1}\subset\mathbb{B}^{n}, V2GV_{2}\subset G and F:𝔹nV1GV2F:\mathbb{B}^{n}-V_{1}\rightarrow G-V_{2} such that FF is an m-sheeted covering map with deck transformation group Γ\Gamma. If FF is algebraic, then the Bergman kernel form of GG is algebraic.

Proof.

Note that the Bergman kernel form of GG coincides with that of G~:=GV2\widetilde{G}:=G-V_{2} by Theorem 2.6, and likewise the Bergman kernel form K𝔹nK_{\mathbb{B}^{n}} of 𝔹n\mathbb{B}^{n} coincides with that of B~:=𝔹nV1\widetilde{B}:=\mathbb{B}^{n}-V_{1}. By the transformation law in Theorem 1.6, we have

γΓ(id𝔹n,γ)K𝔹n=(F,F)KG on B~×B~.\sum_{\gamma\in\Gamma}(\mathrm{id}_{\mathbb{B}^{n}},\gamma)^{*}K_{\mathbb{B}^{n}}=(F,F)^{*}K_{G}\quad\mbox{ on }\widetilde{B}\times\widetilde{B}.

Since all γΓ\gamma\in\Gamma and K𝔹nK_{\mathbb{B}^{n}} are rational, so is the right hand side of the equation. This implies that KGK_{G} is algebraic (see the equivalent condition (c) of algebraicity in §2.1). ∎

Theorem 4.1 applies in particular to Cartan’s canonical realization of ball quotient.

Corollary 4.2.

Let ΓU(n)\Gamma\subset U(n) be a finite unitary group. Suppose Q:nNQ:\mathbb{C}^{n}\rightarrow\mathbb{C}^{N} is the basic map associated to Γ\Gamma. Let G=Q(𝔹n)G=Q(\mathbb{B}^{n}), which is a relatively compact domain in the algebraic variety V=Q(n)V=Q(\mathbb{C}^{n}). Then the Bergman kernel form of GG is algebraic.

Proof.

We let

Z={zn:the Jacobian of Q at z is not full rank}.Z=\{z\in\mathbb{C}^{n}:\mbox{the Jacobian of $Q$ at $z$ is not full rank}\}.

Clearly, ZZ is a proper complex analytic variety in n\mathbb{C}^{n}. By Remmert’s proper mapping theorem, Q(Z)VQ(Z)\subset V is a proper complex analytic variety. Moreover, Q:𝔹nZGQ(Z)Q:\mathbb{B}^{n}-Z\rightarrow G-Q(Z) is a covering map with mm sheets, where m=|Γ|m=|\Gamma|, and Γ\Gamma is its deck transformation group (Note that Q1(Q(Z))=ZQ^{-1}(Q(Z))=Z; see [11]). The conclusion now follows from Theorem 4.1. ∎

Remark 4.3.

Note that the "if" implication in Theorem 1.1 in fact holds under a much weaker assumption than that stipulated in the theorem. In Theorem 4.1 we do not assume n=2n=2 nor that the group Γ\Gamma is fixed point free. We remark that the formula for the Bergman kernel of the finite ball quotient is also obtained by Huang-Li [31].

5. Proof of Theorem 1.1, part II

In this section, we prove one of the main results of the paper—the "only if" implication in Theorem 1.1. We also prove Corollary 1.3 and 1.4.

Proof of the "only if" implication in Theorem 1.1..

Let VV and GG be as in Theorem 1.1 and assume that GG has algebraic Bergman kernel. We shall prove that GG is a finite ball quotient. We proceed in several steps.

Step 1. In this step, we prove G\partial G is real analytic, and furthermore, real algebraic. For this step, we do not need to assume that the dimension of VV is two.

Proposition 5.1.

Let GG be a relatively compact domain in an nn-dimensional (n2n\geq 2) algebraic variety VNV\subset\mathbb{C}^{N} with smooth strongly pseudoconvex boundary. If the Bergman kernel KGK_{G} of GG is algebraic, then the boundary G\partial G of GG is Nash algebraic, i.e., G\partial G is locally defined by a real algebraic function.

Proof.

Fix a point pGp\in\partial G. Then there exists a neighborhood UU of pp in VV with canonical extrinsic coordinates z=(z1,,zn)z=(z_{1},\cdots,z_{n}) on UU (see Section 2). Write the Bergman kernel form KGK_{G} of GG as

KG=K(z,z¯)dzdz¯ on UG,K_{G}=K(z,\bar{z})dz\wedge d\overline{z}\quad\mbox{ on }U\cap G,

where dz=dz1dzndz=dz_{1}\wedge\cdots\wedge dz_{n}, dz¯=dz1¯dzn¯d\overline{z}=d\overline{z_{1}}\wedge\cdots\wedge d\overline{z_{n}} and K(z,z¯)K(z,\bar{z}) is a real algebraic function on UGU\cap G.

As KK is real algebraic, there exist real-valued polynomials a1(z,z¯),,aq(z,z¯)a_{1}(z,\bar{z}),\cdots,a_{q}(z,\bar{z}) in n2n\mathbb{C}^{n}\cong\mathbb{R}^{2n} with aq0a_{q}\neq 0 such that

(5.1) aq(z,z¯)K(z,z¯)q++a1(z,z¯)K(z,z¯)+a0(z,z¯)=0,on UG.a_{q}(z,\bar{z})K(z,\bar{z})^{q}+\cdots+a_{1}(z,\bar{z})K(z,\bar{z})+a_{0}(z,\bar{z})=0,\quad\mbox{on }U\cap G.

Note that when zGz\rightarrow\partial G, we have K(z,z¯)K(z,\bar{z})\rightarrow\infty as G\partial G is strictly pseudoconvex. We divide both sides of (5.1) by K(z,z¯)qK(z,\bar{z})^{q} and let zGz\rightarrow\partial G to obtain

aq(z,z¯)=0, on UG.a_{q}(z,\bar{z})=0,\quad\mbox{ on }U\cap\partial G.

Write zk=xk+iykz_{k}=x_{k}+iy_{k} for 1kn1\leq k\leq n, z=(z1,,zn1)z^{\prime}=(z_{1},\cdots,z_{n-1}) and x=(x1,y1,,xn1,yn1,xn)x^{\prime}=(x_{1},y_{1},\cdots,x_{n-1},y_{n-1},x_{n}). By rotation, we can assume that G\partial G near pp is locally defined by

yn=φ(x),y_{n}=\varphi(x^{\prime}),

where φ\varphi is a smooth function. We then have

aq(z,xn+iφ(x),z¯,xniφ(x))=0.a_{q}\bigl{(}z^{\prime},x_{n}+i\varphi(x^{\prime}),\overline{z^{\prime}},x_{n}-i\varphi(x^{\prime})\bigr{)}=0.

By Malgrange’s theorem (see [37] and references therein), φ\varphi is real analytic and thus, since aqa_{q} is a polynomial, also real algebraic. Hence, G\partial G is Nash algebraic. ∎

Step 2. We now return to the case where VV is two-dimensional. We shall prove that G\partial G is spherical, where GG is as in Theorem 1.1. Fix pGp\in\partial G, and a canonical extrinsic coordinates chart (U,z)(U,z) of VV at pp, where z=(z1,z2)z=(z_{1},z_{2}). We again write

KG(z,z¯)=K(z,z¯)dzdz¯ on UG,K_{G}(z,\bar{z})=K(z,\bar{z})dz\wedge d\overline{z}\quad\mbox{ on }U\cap G,

where dz=dz1dz2dz=dz_{1}\wedge dz_{2} and dz¯=dz1¯dz2¯d\overline{z}=d\overline{z_{1}}\wedge d\overline{z_{2}}. Choose a strongly pseudoconvex domain DUGD\Subset U\cap G such that

B(p,δ)D=B(p,δ)Gfor some small δ>0.B(p,\delta)\cap D=B(p,\delta)\cap G\quad\mbox{for some small }\delta>0.

Here B(p,δ)={zU:zp<δ}B(p,\delta)=\{z\in U:\|z-p\|<\delta\} is the ball centered at pp with radius δ\delta with respect to the coordinates (U,z)(U,z). Write KDK_{D} for the Bergman kernel of DD, which is now considered as a function. Then KDKK_{D}-K extends smoothly across B(p,δ)DB(p,\delta)\cap\partial D (see [16, 9], see also [31] for a nice and detailed proof of this fact). Consequently,

KD(z,z¯)=K(z,z¯)+h(z,z¯)on D,K_{D}(z,\bar{z})=K(z,\bar{z})+h(z,\bar{z})\quad\mbox{on }D,

where h(z,z¯)h(z,\bar{z}) is real analytic in DD and extends smoothly across B(p,δ)DB(p,\delta)\cap\partial D. Let rr be a Fefferman defining function of DD and express the Fefferman asymptotic expansion of KDK_{D} as

KD(z,z¯)=ϕ(z,z¯)r(z)3+ψ(z,z¯)logr(z) on D,K_{D}(z,\bar{z})=\frac{\phi(z,\bar{z})}{r(z)^{3}}+\psi(z,\bar{z})\log r(z)\quad\mbox{ on }D,

where ϕ\phi and ψ\psi are smooth functions on DD that extend smoothly across B(p,δ)DB(p,\delta)\cap\partial D; see [17]. Thus,

(5.2) K(z,z¯)=ϕ(z,z¯)h(z,z¯)r(z)3r(z)3+ψ(z,z¯)logr(z) on D.K(z,\bar{z})=\frac{\phi(z,\bar{z})-h(z,\bar{z})r(z)^{3}}{r(z)^{3}}+\psi(z,\bar{z})\log r(z)\quad\mbox{ on }D.

As in Step 1, there exist real-valued polynomials a1(z,z¯),,aq(z,z¯)a_{1}(z,\bar{z}),\cdots,a_{q}(z,\bar{z}) in 24\mathbb{C}^{2}\cong\mathbb{R}^{4} with aq0a_{q}\neq 0 for some q1q\geq 1, such that

aqKq++a1K+a0=0 on D.a_{q}K^{q}+\cdots+a_{1}K+a_{0}=0\quad\mbox{ on }D.

If we substitute (5.2) into the above equation and multiply both sides by r3qr^{3q}, then

(5.3) aqψqr3q(logr)q+j=0q1bj(logr)j=0 on D,a_{q}\psi^{q}r^{3q}(\log r)^{q}+\sum_{j=0}^{q-1}b_{j}(\log r)^{j}=0\quad\mbox{ on }D,

where all bjb_{j} for 0jq10\leq j\leq q-1 are smooth on DD and extend smoothly across B(p,δ)DB(p,\delta)\cap\partial D. We recall the following lemma from [21].

Lemma 5.2 ([21]).

Let f0(t),,fq(t)C(ε,ε)f_{0}(t),\cdots,f_{q}(t)\in C^{\infty}(-\varepsilon,\varepsilon) for ε>0\varepsilon>0. If

f0(t)+f1(t)logt++fq(t)(logt)q=0f_{0}(t)+f_{1}(t)\log t+\cdots+f_{q}(t)(\log t)^{q}=0

for all t(0,ε)t\in(0,\varepsilon), then each fj(t)f_{j}(t) for 0jq0\leq j\leq q, vanishes to infinite order at 0.

It follows from the above lemma and (5.3) that the coefficient ψ\psi of of the logarithmic term vanishes to infinite order at G\partial G near pp. Since GG is two-dimensional, it follows that GG is locally spherical near pp by [23] (see page 129 where the result is credited to Burns) and [8] (see page 23).

Remark 5.3.

Recall from the introduction that the sphericity near pp above follows from the affirmation of the Ramadanov Conjecture in two dimensions. This is also the only place where the fact that GG is two dimensional is essentially used.

Step 3. In this step, we will prove there is an algebraic branched covering map F:𝔹2GF:\mathbb{B}^{2}\rightarrow G with finitely many sheets. Since we have already shown that G\partial G is a Nash algebraic and spherical CR submanifold in N\mathbb{C}^{N}, by a theorem of Huang (see Corollary 3.3 in [28]), it follows that G\partial G is CR equivalent to a CR spherical space form 𝔹2/Γ\partial\mathbb{B}^{2}/\Gamma with ΓU(n)\Gamma\subset U(n) a finite group with no fixed points on 𝔹2\partial\mathbb{B}^{2}. In particular, there is a CR covering map f:𝔹2Gf:\partial\mathbb{B}^{2}\rightarrow\partial G (see the proof of Theorem 3.1 in [28] and also [29]). By Hartogs’s extension theorem, ff extends as a smooth map F:𝔹2¯VF\colon\overline{\mathbb{B}^{2}}\to V, holomorphic in 𝔹2\mathbb{B}^{2} and sending 𝔹2\partial\mathbb{B}^{2} onto G\partial G. The latter implies that FF is moreover algebraic by X. Huang’s algebraicity theorem [26]. It is not difficult to see that FF sends 𝔹2\mathbb{B}^{2} into GG. Since FF maps 𝔹2\partial\mathbb{B}^{2} to G\partial G, we conclude that FF is a proper algebraic mapping 𝔹2G\mathbb{B}^{2}\to G.


Claim 1. F:𝔹2GF\colon\mathbb{B}^{2}\to G is surjective.

Proof of Claim 1.

By the properness of FF, F(𝔹2)F(\mathbb{B}^{2}) is closed in GG. Let us denote by

Z:={z𝔹2:F is not full rank at z}.Z:=\{z\in\mathbb{B}^{2}:F\mbox{ is not full rank at }z\}.

Since FF is a local biholomorphic map at every point of 𝔹2\partial\mathbb{B}^{2}, ZZ is a finite set. We also note that if p𝔹2Zp\in\mathbb{B}^{2}-Z, then F(p)F(p) is a smooth point of VV, and F(p)F(p) is an interior point of F(𝔹2)F(\mathbb{B}^{2}). Assume, in order to reach a contradiction, that F(𝔹2)GF(\mathbb{B}^{2})\neq G. Since F(𝔹2)F(\mathbb{B}^{2}) is closed in GG, its complement GF(𝔹2)G\setminus F(\mathbb{B}^{2}) is then a non-empty open subset of GG. Note that any boundary point of F(𝔹2)F(\mathbb{B}^{2}) in GG can only be in F(Z)F(Z). But F(Z)F(Z) is a finite set, which cannot separate the (non-empty) interior of F(𝔹2)F(\mathbb{B}^{2}) and the (non-empty) open complement GF(𝔹2)G\setminus F(\mathbb{B}^{2}) in the domain GG. This is the desired contradiction and, hence, F(𝔹2)=GF(\mathbb{B}^{2})=G. ∎

Now, we let T:=F1(F(Z))ZT:=F^{-1}(F(Z))\supset Z. Then TT is a compact analytic subvariety of 𝔹2\mathbb{B}^{2} and thus is a finite set. Consider the restriction of FF:

F|𝔹2T:𝔹2TGF(Z),F|_{\mathbb{B}^{2}-T}:\mathbb{B}^{2}-T\rightarrow G-F(Z),

still denoted by FF. Clearly, FF is a proper surjective map. Since FF is also a local biholomorphism, FF is a finite covering map.

Note that 𝔹2T\mathbb{B}^{2}-T is simply connected. It follows that the deck transformation group Γ~={γ~k}k=1m\widetilde{\Gamma}=\{\widetilde{\gamma}_{k}\}_{k=1}^{m} of the covering map F:𝔹2TGf(Z)F:\mathbb{B}^{2}-T\rightarrow G-f(Z) acts transitively on each fiber. Since each γ~k\widetilde{\gamma}_{k} is a biholomorphism from 𝔹2T\mathbb{B}^{2}-T to 𝔹2T\mathbb{B}^{2}-T, it extends to an automorphism of 𝔹2\mathbb{B}^{2}. Consequently,

Γ~={γ~Aut(𝔹2):Fγ~=F on 𝔹2}.\widetilde{\Gamma}=\bigl{\{}\widetilde{\gamma}\in\operatorname{Aut}(\mathbb{B}^{2}):F\circ\widetilde{\gamma}=F\mbox{ on }\mathbb{B}^{2}\bigr{\}}.

Recall that Γ\Gamma is the deck transformation group of the original covering map f:𝔹2Gf:\partial\mathbb{B}^{2}\rightarrow\partial G. From this, it is clear that we can identify Γ\Gamma with Γ~\widetilde{\Gamma}. From now on, we will simply use the notation Γ\Gamma for either group.

Note that ZZ and TT are both closed under the action of Γ\Gamma, and (𝔹2T)/Γ(\mathbb{B}^{2}-T)/\Gamma is biholomorphic to Gf(Z)G-f(Z).

Claim 2: If z,w𝔹2z,w\in\mathbb{B}^{2} satisfy F(z)=F(w),F(z)=F(w), then w=γ(z)w=\gamma(z) for some γΓ.\gamma\in\Gamma. Consequently, T=ZT=Z.

Proof of Claim 2.

We only need to prove the first assertion. If both z,wz,w are in 𝔹2T,\mathbb{B}^{2}-T, then the conclusion is clear as Γ\Gamma acts transitively on each fiber of the covering map F:𝔹2TGf(Z)F:\mathbb{B}^{2}-T\rightarrow G-f(Z). Next we assume one of zz and ww is in TT. Seeking a contradiction, suppose wγ(z)w\neq\gamma(z) for every γΓ.\gamma\in\Gamma. Writing q:=F(z)=F(w),q:=F(z)=F(w), there are then points in distinct orbits of Γ\Gamma that are mapped to q.q. Writing tt for the number of orbits of Γ\Gamma that are mapped to qq, we must then have t2.t\geq 2. Pick p1,,ptp_{1},\cdots,p_{t} from these tt distinct orbits of Γ.\Gamma. Since TT is a finite set, we can choose, for each 1it,1\leq i\leq t, some disjoint neighborhoods UiU_{i} of pip_{i} such that UiT{pi}.U_{i}\cap T\subseteq\{p_{i}\}. Moreover, we can make γ(Ui)Uj=\gamma(U_{i})\cap U_{j}=\emptyset for all γΓ\gamma\in\Gamma if ij.i\neq j. Consequently, F(Ui{pi})F(Uj{pj})=.F(U_{i}-\{p_{i}\})\cap F(U_{j}-\{p_{j}\})=\emptyset. Note there is a small open subset WW containing qq such that Wi=1tF(γΓγ(Ui))=i=1tF(Ui)W\subseteq\cup_{i=1}^{t}F(\cup_{\gamma\in\Gamma}\gamma(U_{i}))=\cup_{i=1}^{t}F(U_{i}). Thus W{q}i=1tF(Ui{pi}).W-\{q\}\subseteq\cup_{i=1}^{t}F(U_{i}-\{p_{i}\}). But the sets F(Uipi)(W{q})F(U_{i}-p_{i})\cap(W-\{q\}) are open and disjoint, and we can choose W{q}W-\{q\} to be connected. This is a contradiction. Thus we must have t=1t=1 and w=γ(z)w=\gamma(z) for some γΓ.\gamma\in\Gamma.

Recall that 0 is assumed to be the only fixed point for elements in Γ\Gamma. We write q0:=F(0)q_{0}:=F(0) and prove that q0q_{0} is the only possible singularity in GG. Also, recall that all singularities of GG are assumed to be isolated and normal.

Claim 3. GG can only have a singularity at q0q_{0}.

Proof of Claim 3.

Suppose q1q_{1} is a (normal) singular point in GG and q1q0q_{1}\neq q_{0}. Since FF is onto, there exists some p1𝔹2p_{1}\in\mathbb{B}^{2} such that f(p1)=q1f(p_{1})=q_{1}.

First, note that we can find a small neighborhood U0U_{0} of p1p_{1}, and a small neighborhood WW of q1q_{1} in GG such that

(i)U0T={p1};(ii)Fis injective onU0;(iii)WF(U0)andWF(Z)={q1}.\text{(i)}\leavevmode\nobreak\ U_{0}\cap T=\{p_{1}\};\quad\text{(ii)}\leavevmode\nobreak\ F\leavevmode\nobreak\ \text{is injective on}\leavevmode\nobreak\ U_{0};\quad\text{(iii)}\leavevmode\nobreak\ W\subseteq F(U_{0})\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ W\cap F(Z)=\{q_{1}\}.

It is easy to see that we can make (i) and (ii) hold. It is guaranteed by Claim 2 (see its proof) that we can find WF(U0)W\subseteq F(U_{0}); the second condition in (iii) is then easy to satisfy, since F(Z)F(Z) is a finite set. Now, we let U:=U0F1(W)U:=U_{0}\cap F^{-1}(W), which is an open subset of 𝔹2\mathbb{B}^{2} containing p1p_{1}. Then F:U{p1}W{q1}F:U-\{p_{1}\}\rightarrow W-\{q_{1}\} is a biholomorphism. We let g:W{q1}U{p1}g:W-\{q_{1}\}\rightarrow U-\{p_{1}\} denote its inverse. By the normality of q1q_{1}, we can assume that gg is the restriction of some holomorphic map g^\widehat{g} defined on some open set W^N\widehat{W}\subset\mathbb{C}^{N}, where W^\widehat{W} contains WW. Since gF|U{p1}g\circ F|_{U-\{p_{1}\}} equals the identity map, g^F\widehat{g}\circ F equals identity on UU by continuity. Similarly F(g^|W)F\circ(\widehat{g}|_{W}) equals the identity on WW. Therefore, q1q_{1} cannot be a singular point. ∎

By Claim 2 and Claim 3, we also see that T=Z={0} or T=Z=\{0\}\mbox{ or }\varnothing. Therefore, FF gives a holomorphic algebraic branched covering map from 𝔹2\mathbb{B}^{2} to GG with a possible branch point at 0. This completes the proof of the "only if" implication in Theorem 1.1. ∎

Remark 5.4.

We reiterate (see Remark 5.3 above) that in the above proof, the condition that dimV=2\dim V=2 is only used in the second step where we apply the affirmative solution of the Ramadanov conjecture in 2\mathbb{C}^{2} ([23], [8]).

We shall now prove Corollaries 1.3 and 1.4.

Proof of Corollary 1.3.

By Theorem 1.1, it follows that GG can be realized as a finite ball quotient 𝔹2/Γ\mathbb{B}^{2}/\Gamma by an algebraic map for some finite unitary group Γ\Gamma with no fixed point on 𝔹2\partial\mathbb{B}^{2}. We must prove that Γ={id}\Gamma=\{\mathrm{id}\}. Suppose not. But, then GG must have a singular point (see [41]), which is a contradiction. ∎

Proof of Corollary 1.4.

The algebraic case follows immediately from Corollary 1.3. Thus, we only need to consider the rational case. First, as a consequence of the algebraic case, there exists an algebraic biholomorphic map f:G𝔹2f:G\rightarrow\mathbb{B}^{2}. It remains to establish that ff is in fact rational. This follows immediately from a result by Bell [5]. For the convenience of the readers, however, we sketch an independent proof here. Denote by KGK_{G} and K𝔹2K_{\mathbb{B}^{2}} the Bergman kernels (now considered as functions) of GG and 𝔹2\mathbb{B}^{2}, respectively. By the transformation law, they are related as

(5.4) KG(z,w)=det(Jf(z))KB(f(z),f(w))det(Jf(w))¯=2!π2det(Jf(z))det(Jf(w))¯1(1f(z)f(w)¯)3.\displaystyle\begin{split}K_{G}(z,w)=&\det\big{(}Jf(z)\bigr{)}\cdot K_{B}\bigl{(}f(z),f(w)\bigr{)}\cdot\overline{\det\bigl{(}Jf(w)\bigr{)}}\\ =&\frac{2!}{\pi^{2}}\det\big{(}Jf(z)\bigr{)}\cdot\overline{\det\bigl{(}Jf(w)\bigr{)}}\cdot\frac{1}{\bigl{(}1-f(z)\cdot\overline{f(w)}\bigr{)}^{3}}.\end{split}

We may assume 0G0\in G by translating GG if necessary and, by composing ff with an automorphism of 𝔹2\mathbb{B}^{2}, we may also assume f(0)=0f(0)=0. Thus, at w=0w=0, we have

KG(z,0)=2!π2det(Jf(z))det(Jf(0))¯.K_{G}(z,0)=\frac{2!}{\pi^{2}}\det\big{(}Jf(z)\bigr{)}\cdot\overline{\det\bigl{(}Jf(0)\bigr{)}}.

It follows that

(5.5) det(Jf(z))=det(Jf(0))KG(z,0)KG(0,0).\det\bigl{(}Jf(z)\bigr{)}=\det\bigl{(}Jf(0)\bigr{)}\frac{K_{G}(z,0)}{K_{G}(0,0)}.

In particular, this implies that KG(z,0)0K_{G}(z,0)\neq 0 for any zDz\in D. We evaluate (5.4) on the diagonal w=zw=z and use (5.5) to obtain

KG(z,z)=2!π2|det(Jf(0))|2|KG(z,0)|2|KG(0,0)|21(1f(z)2)3.K_{G}(z,z)=\frac{2!}{\pi^{2}}\bigl{|}\det\big{(}Jf(0)\bigr{)}\bigr{|}^{2}\frac{|K_{G}(z,0)|^{2}}{|K_{G}(0,0)|^{2}}\frac{1}{\bigl{(}1-\|f(z)\|^{2}\bigr{)}^{3}}.

Taking the logarithm of both sides yields

logKG(z,z)+3log(1f(z)2)=log2!π2+log|det(Jf(0))|2+log|KG(z,0)|2log|KG(0,0)|2.\log K_{G}(z,z)+3\log\bigl{(}1-\|f(z)\|^{2}\bigr{)}=\log\frac{2!}{\pi^{2}}+\log\bigl{|}\det\big{(}Jf(0)\bigr{)}\bigr{|}^{2}+\log|K_{G}(z,0)|^{2}-\log|K_{G}(0,0)|^{2}.

For j=1,2j=1,2, we apply the derivative zj¯\frac{\partial}{\partial\overline{z_{j}}} to both sides and obtain

1KG(z,z)KG(z,z)zj¯31f(z)2i=12fi(z)¯zj¯fi(z)=1KG(z,0)¯KG(z,0)¯zj¯.\displaystyle\frac{1}{K_{G}(z,z)}\frac{\partial K_{G}(z,z)}{\partial\overline{z_{j}}}-\frac{3}{1-\|f(z)\|^{2}}\sum_{i=1}^{2}\frac{\partial\overline{f_{i}(z)}}{\partial\overline{z_{j}}}f_{i}(z)=\frac{1}{\overline{K_{G}(z,0)}}\frac{\partial\overline{K_{G}(z,0)}}{\partial\overline{z_{j}}}.

Complexifying the above equation and evaluating it at w=0w=0, after rearrangement, we obtain

i=12fi¯zj¯(0)fi(z)=13(1KG(z,0)KGzj¯(z,0)1KG(0,0)¯KG¯zj¯(0,0)).\displaystyle\sum_{i=1}^{2}\frac{\partial\overline{f_{i}}}{\partial\overline{z_{j}}}(0)f_{i}(z)=\frac{1}{3}\left(\frac{1}{K_{G}(z,0)}\frac{\partial K_{G}}{\partial\overline{z_{j}}}(z,0)-\frac{1}{\overline{K_{G}(0,0)}}\frac{\partial\overline{K_{G}}}{\partial\overline{z_{j}}}(0,0)\right).

Note this is a linear system for f(z)=(f1(z),f2(z))f(z)=(f_{1}(z),f_{2}(z)) and the coefficient matrix Jf(0)¯\overline{Jf(0)} is non-singular. By solving this linear system for ff, it is immediately clear that the rationality of KGK_{G} implies that of ff. ∎

Remark 5.5.

Corollary 1.3 implies, in particular, that the Burns–Shnider domains in 2\mathbb{C}^{2} (see page 244 in [10]) cannot have algebraic Bergman kernels. In fact, this holds for any Burns–Shnider domain in n\mathbb{C}^{n} for n2n\geq 2, which can be seen as follows. By Proposition 5.1, if the Bergman kernel were algebraic, then the boundary would be Nash algebraic. While this can be seen to not be so by inspection, a contradiction would also be reached by the Huang–Ji Riemann mapping theorem [30] since the boundary of a Burns–Shnider domain is spherical while the domain itself is not biholomorphic to the unit ball.

6. Counterexample in higher dimension

In this section, we construct a 33-dimensional reduced Stein space GG with only one normal singularity and compact, smooth strongly pseudoconvex boundary, realized as a relatively compact domain in a complex algebraic variety VV in 4\mathbb{C}^{4}. We will show that its Bergman kernel is algebraic, while GG is not biholomorphic to any finite ball quotient 𝔹n/Γ\mathbb{B}^{n}/\Gamma, which shows that Theorem 1.1 cannot hold in higher dimensions.

Let GG be defined as

G={w=(w1,w2,w3,w4)4:|w1|2+|w2|2+|w3|2+|w4|2<1,w1w4=w2w3}.G=\bigl{\{}w=(w_{1},w_{2},w_{3},w_{4})\in\mathbb{C}^{4}:|w_{1}|^{2}+|w_{2}|^{2}+|w_{3}|^{2}+|w_{4}|^{2}<1,\quad w_{1}w_{4}=w_{2}w_{3}\bigr{\}}.

Then GG is a relatively compact domain in the complex algebraic variety

(6.1) V={w4:w1w4=w2w3}.V=\bigl{\{}w\in\mathbb{C}^{4}:w_{1}w_{4}=w_{2}w_{3}\bigr{\}}.

Since GG is a closed algebraic subvariety of 𝔹44\mathbb{B}^{4}\subset\mathbb{C}^{4}, GG is a reduced Stein space. Note that 0 is the only singularity of VV. Moreover it is a normal singularity as it is a hypersurface singularity of codimension 3 (>2>2; see [42]). It is also easy to verify that GG has smooth strongly pseudoconvex boundary in VV.

Proposition 6.1.

The boundary M=GM=\partial G of GG is homogeneous and non-spherical.

Proof.

Consider the product complex manifold 1×1\mathbb{CP}^{1}\times\mathbb{CP}^{1}. For j=1,2j=1,2, let πj:1×11\pi_{j}:\mathbb{CP}^{1}\times\mathbb{CP}^{1}\rightarrow\mathbb{CP}^{1} be the projection map to the jj-th component and let (L0,h0)1(L_{0},h_{0})\rightarrow\mathbb{CP}^{1} be the tautological line bundle L0L_{0} with its standard Hermitian metric h0h_{0}. We set the Hermitian line bundle (L,h)(L,h) over 1×1\mathbb{CP}^{1}\times\mathbb{CP}^{1} to be:

(L,h):=π1(L0,h0)π2(L0,h0).(L,h):=\pi_{1}^{*}(L_{0},h_{0})\otimes\pi_{2}^{*}(L_{0},h_{0}).

We begin the proof with the following claim.

Claim 1. Let (L,h)1×1(L,h)\rightarrow\mathbb{CP}^{1}\times\mathbb{CP}^{1} be as above and let S(L)1×1S(L)\rightarrow\mathbb{CP}^{1}\times\mathbb{CP}^{1} be its unit circle bundle. Then MM is CR diffeomorphic to S(L)S(L) by the restriction of biholomorphic map.

Proof of Claim 1.

Note that the circle bundle S(L)1×1S(L)\rightarrow\mathbb{CP}^{1}\times\mathbb{CP}^{1} can be written as

(6.2) S(L)={(λ(ζ1,z1)(ζ2,z2),[ζ1,z1],[ζ2,z2]):[ζ1,z1]1,[ζ2,z2]1|λ|2(|ζ1|2+|z1|2)(|ζ2|2+|z2|2)=1}.S(L)=\left\{\Bigl{(}\lambda(\zeta_{1},z_{1})\otimes(\zeta_{2},z_{2}),[\zeta_{1},z_{1}],[\zeta_{2},z_{2}]\Bigr{)}:\begin{array}[]{l}[\zeta_{1},z_{1}]\in\mathbb{CP}^{1},\quad[\zeta_{2},z_{2}]\in\mathbb{CP}^{1}\\ |\lambda|^{2}(|\zeta_{1}|^{2}+|z_{1}|^{2})(|\zeta_{2}|^{2}+|z_{2}|^{2})=1\end{array}\right\}.

Define F:L4F:L\rightarrow\mathbb{C}^{4} as

(6.3) F(λ(ζ1,z1)(ζ2,z2),[ζ1,z1],[ζ2,z2])=(λζ1ζ2,λz1ζ2,λζ1z2,λz1z2).F\Bigl{(}\lambda(\zeta_{1},z_{1})\otimes(\zeta_{2},z_{2}),[\zeta_{1},z_{1}],[\zeta_{2},z_{2}]\Bigr{)}=\Bigl{(}\lambda\zeta_{1}\zeta_{2},\lambda z_{1}\zeta_{2},\lambda\zeta_{1}z_{2},\lambda z_{1}z_{2}\Bigr{)}.

Then the map FF gives a biholomorphism that sends a neighborhood of S(L)S(L) in LL to a neighborhood of MM in V4V\subset\mathbb{C}^{4}. This proves the claim. ∎

Note that S(L)S(L) is homogeneous (see [14]) and non-spherical by Theorem 12 in [43]. Thus, MM is homogeneous and non-spherical. ∎

Proposition 6.2.

The Bergman kernel form KGK_{G} of GG is algebraic.

Proof.

Set

(6.4) Ω:={(λ,z)=(λ,z1,z2)3:|λ|2(1+|z1|2)(1+|z2|2)<1}.\Omega:=\bigl{\{}(\lambda,z)=(\lambda,z_{1},z_{2})\in\mathbb{C}^{3}:|\lambda|^{2}(1+|z_{1}|^{2})(1+|z_{2}|^{2})<1\bigr{\}}.

Note that Ω\Omega is an unbounded domain with smooth boundary in 3\mathbb{C}^{3}. Moreover, Ω\Omega has a rational Bergman kernel form KΩK_{\Omega} (see Appendix 7 for a proof of this fact). Define the map F:34F:\mathbb{C}^{3}\rightarrow\mathbb{C}^{4} as

F(λ,z1,z2):=(λ,λz1,λz2,λz1z2),F(\lambda,z_{1},z_{2}):=(\lambda,\lambda z_{1},\lambda z_{2},\lambda z_{1}z_{2}),

We note that F(3)F(\mathbb{C}^{3}) is contained in VV as defined by (6.1). And FF is a holomorphic embedding on 3{λ=0}\mathbb{C}^{3}-\{\lambda=0\}. Moreover, F(Ω)GF(\Omega)\subset G and

F:Ω~:=Ω{λ=0}G~:=G{w1=0}F:\widetilde{\Omega}:=\Omega-\{\lambda=0\}\rightarrow\widetilde{G}:=G-\{w_{1}=0\}

is a biholomorphism. By Theorem 2.6, the Bergman kernel form KΩ~K_{\widetilde{\Omega}} of Ω~\widetilde{\Omega} is the restriction (pullback) of KΩK_{\Omega} to Ω~\widetilde{\Omega}. Thus, KΩ~K_{\widetilde{\Omega}} is rational. By the transformation law (1.1), we have

KΩ~=(F,F)KG~.K_{\widetilde{\Omega}}=(F,F)^{*}K_{\widetilde{G}}.

This implies that KG~K_{\widetilde{G}} is algebraic (see the equivalent condition (c) in §2.1), and thus KGK_{G} is also algebraic by Theorem 2.6. ∎

Before we prove GG is not biholomorphic to any finite ball quotient, we pause to study the following bounded domain UU in 3\mathbb{C}^{3}:

U:={(w1,w2,w3)3:|w1|4+|w1|2(|w2|2+|w3|2)+|w2w3|2<|w1|2}.U:=\Bigl{\{}(w_{1},w_{2},w_{3})\in\mathbb{C}^{3}:|w_{1}|^{4}+|w_{1}|^{2}(|w_{2}|^{2}+|w_{3}|^{2})+|w_{2}w_{3}|^{2}<|w_{1}|^{2}\Bigr{\}}.
Proposition 6.3.

The domain UU has algebraic Bergman kernel and its boundary is non-spherical at every smooth boundary point.

Proof.

Let π:43\pi:\mathbb{C}^{4}\rightarrow\mathbb{C}^{3} be the projection map defined by

π(w1,w2,w3,w4):=(w1,w2,w3).\pi(w_{1},w_{2},w_{3},w_{4}):=(w_{1},w_{2},w_{3}).

Let G¯\overline{G} be the closure of GG in 4\mathbb{C}^{4}. Then the image of G¯\overline{G} under the projection π\pi is

U^:=π(G¯)=\displaystyle\widehat{U}:=\pi(\overline{G})= {(w1,w2,w3)3:|w1|4+|w1|2(|w2|2+|w3|2)+|w2w3|2|w1|2,w10}\displaystyle\bigl{\{}(w_{1},w_{2},w_{3})\in\mathbb{C}^{3}:|w_{1}|^{4}+|w_{1}|^{2}(|w_{2}|^{2}+|w_{3}|^{2})+|w_{2}w_{3}|^{2}\leq|w_{1}|^{2},w_{1}\neq 0\bigr{\}}
{(0,w2,w3)3:|w2|2+|w3|21,w2w3=0}\displaystyle\quad\cup\bigl{\{}(0,w_{2},w_{3})\in\mathbb{C}^{3}:|w_{2}|^{2}+|w_{3}|^{2}\leq 1,w_{2}w_{3}=0\bigr{\}}
=\displaystyle= {(w1,w2,w3)3:|w1|4+|w1|2(|w2|2+|w3|2)+|w2w3|2|w1|2,|w2|2+|w3|21}.\displaystyle\bigl{\{}(w_{1},w_{2},w_{3})\in\mathbb{C}^{3}:|w_{1}|^{4}+|w_{1}|^{2}(|w_{2}|^{2}+|w_{3}|^{2})+|w_{2}w_{3}|^{2}\leq|w_{1}|^{2},|w_{2}|^{2}+|w_{3}|^{2}\leq 1\bigr{\}}.

Note that U^o=UandU^=U¯\widehat{U}^{\mathrm{o}}=U\quad\mbox{and}\quad\widehat{U}=\overline{U}, where U^o\widehat{U}^{\mathrm{o}} denotes the interior of U^\widehat{U} . But Uπ(G)U\neq\pi(G). On the other hand if we remove the variety {w1=0}\{w_{1}=0\}, then the projection map

π:G{w1=0}U\pi:G-\{w_{1}=0\}\rightarrow U

is an algebraic biholomorphism. Consequently, by Theorem 2.6 the Bergman kernel form KUK_{U} of UU is algebraic. This proves the first part of the proposition.

To prove the second part of the proposition (i.e., the non-sphericity), we observe that the boundary U\partial U of UU is given by

U=\displaystyle\partial U= {(w1,w2,w3)3:|w1|4+|w1|2(|w2|2+|w3|2)+|w2w3|2=|w1|2,w10}\displaystyle\bigl{\{}(w_{1},w_{2},w_{3})\in\mathbb{C}^{3}:|w_{1}|^{4}+|w_{1}|^{2}(|w_{2}|^{2}+|w_{3}|^{2})+|w_{2}w_{3}|^{2}=|w_{1}|^{2},w_{1}\neq 0\bigr{\}}
{(0,w2,w3)3:|w2|2+|w3|21,w2w3=0}\displaystyle\quad\cup\bigl{\{}(0,w_{2},w_{3})\in\mathbb{C}^{3}:|w_{2}|^{2}+|w_{3}|^{2}\leq 1,w_{2}w_{3}=0\bigr{\}}
=\displaystyle= {(w1,w2,w3)3:|w1|4+|w1|2(|w2|2+|w3|2)+|w2w3|2=|w1|2,|w2|2+|w3|21}.\displaystyle\bigl{\{}(w_{1},w_{2},w_{3})\in\mathbb{C}^{3}:|w_{1}|^{4}+|w_{1}|^{2}(|w_{2}|^{2}+|w_{3}|^{2})+|w_{2}w_{3}|^{2}=|w_{1}|^{2},|w_{2}|^{2}+|w_{3}|^{2}\leq 1\bigr{\}}.

Write

U=(U{w10})(U{w1=0}).\partial U=\bigl{(}\partial U\cap\{w_{1}\neq 0\}\bigr{)}\cup\bigl{(}\partial U\cap\{w_{1}=0\}\bigr{)}.

Since the projection map π\pi is a biholomorphism from G¯{w1=0}\overline{G}-\{w_{1}=0\} to U^{w1=0}\widehat{U}-\{w_{1}=0\}, every point pU{w10}p\in\partial U\cap\{w_{1}\neq 0\} is a smooth point of U\partial U, and, moreover, U\partial U is strictly pseudoconvex and non-spherical at pp. We note that a defining function for U\partial U near pp is given by

ρ=|w1|4+|w1|2(|w2|2+|w3|2)+|w2w3|2|w1|2.\rho=|w_{1}|^{4}+|w_{1}|^{2}(|w_{2}|^{2}+|w_{3}|^{2})+|w_{2}w_{3}|^{2}-|w_{1}|^{2}.

Furthermore, it is easy to verify that every other point qU{w1=0}q\in\partial U\cap\{w_{1}=0\} is not a smooth boundary point of UU. This proves the second part of the assertion. ∎

We are now ready to show that GG is indeed a counterexample to the conclusion of Theorem 1.1 in three dimensions.

Proposition 6.4.

GG is not biholomorphic to any finite ball quotient.

Proof.

Seeking a contradiction, we suppose GG is biholomorphic to a finite ball quotient 𝔹3/Γ\mathbb{B}^{3}/\Gamma, where ΓU(n)\Gamma\subset U(n) is a finite unitary group. We realize 𝔹3/Γ\mathbb{B}^{3}/\Gamma as the image G0NG_{0}\subset\mathbb{C}^{N} of 𝔹3\mathbb{B}^{3} under the basic map QQ associated to Γ\Gamma, where Q=(p1,,pN):3NQ=(p_{1},\cdots,p_{N}):\mathbb{C}^{3}\rightarrow\mathbb{C}^{N} gives a proper map from 𝔹3\mathbb{B}^{3} to G0.G_{0}. Let FF be a biholomorphism from G0𝔹3/ΓG_{0}\cong\mathbb{B}^{3}/\Gamma to GG. Then there is an analytic variety W0G0W_{0}\subset G_{0} such that

F:G0W0G{w1=0} is a biholomorphism.F:G_{0}-W_{0}\rightarrow G-\{w_{1}=0\}\mbox{ is a biholomorphism}.

There also exists an analytic variety WW such that W=Q1(W0)W=Q^{-1}(W_{0}) and thus Q:𝔹3WG0W0Q:\mathbb{B}^{3}-W\rightarrow G_{0}-W_{0} is proper and onto. Set

f:=πFQ:𝔹3WU=π(G{w1=0}),f:=\pi\circ F\circ Q:\mathbb{B}^{3}-W\rightarrow U=\pi(G-\{w_{1}=0\}),

where π\pi is the projection defined in the proof of Proposition 6.3 and is a biholomorphism from G{w1=0}G-\{w_{1}=0\} to UU. Note that ff is proper. Since U3U\subset\mathbb{C}^{3}, we can write ff as (f1,f2,f3)(f_{1},f_{2},f_{3}).

Claim 2. There is a sequence {ζi}𝔹3W\{\zeta_{i}\}\subset\mathbb{B}^{3}-W with ζiζ𝔹3W¯\zeta_{i}\rightarrow\zeta^{*}\in\partial\mathbb{B}^{3}-\overline{W} such that

f(ζi)pU{w10}.f(\zeta_{i})\rightarrow p^{*}\in\partial U\cap\{w_{1}\neq 0\}.
Proof of Claim 2.

Suppose not. Then for any {ζi}𝔹nW\{\zeta_{i}\}\subset\mathbb{B}^{n}-W with ζiζ𝔹nW¯\zeta_{i}\rightarrow\zeta^{*}\in\partial\mathbb{B}^{n}-\overline{W}, every convergent subsequence of f(ζi)f(\zeta_{i}) converges to some point in U{w1=0}\partial U\cap\{w_{1}=0\}. That is to say, if f(ζik)f(\zeta_{i_{k}}) is convergent, then f1(ζik)0f_{1}(\zeta_{i_{k}})\rightarrow 0. Note that UU is bounded. Thus, f1(ζi)0f_{1}(\zeta_{i})\rightarrow 0 for any {ζi}𝔹3W\{\zeta_{i}\}\subset\mathbb{B}^{3}-W with ζiζ𝔹3W¯\zeta_{i}\rightarrow\zeta^{*}\in\partial\mathbb{B}^{3}-\overline{W}. By a standard argument using analytic disks attached to 𝔹3W¯\partial\mathbb{B}^{3}-\overline{W}, we see that

f1=0 on 𝔹3W.f_{1}=0\quad\mbox{ on }\mathbb{B}^{3}-W.

This is a contradiction. ∎

Let ζi,ζ\zeta_{i},\zeta^{*} and pp^{*} be as in Claim 2. Note that ζ\zeta^{*} is a smooth strictly pseudoconvex boundary point of 𝔹3W\mathbb{B}^{3}-W, and pp^{*} is a smooth strictly pseudoconvex boundary point of UU (see the proof of Proposition 6.3). It follows from [20] (see page 239) that ff extends to a Hölder-12\frac{1}{2} continuous map on a neighborhood of ζ\zeta^{*} in 𝔹3¯\overline{\mathbb{B}^{3}}. Since ff is proper 𝔹3WU\mathbb{B}^{3}-W\to U, its extension to the boundary is a (Hölder-12\frac{1}{2}) continuous, nonconstant CR map sending a piece of 𝔹3\partial\mathbb{B}^{3} containing ζ\zeta^{*} to a piece of U\partial U containing pp^{*}. By [39], ff extends holomorphically to a neighborhood of ζ\zeta^{*}, since both boundaries are real analytic (in fact, real-algebraic). Now, since ff is non-constant and sends a strongly pseudoconvex hypersurface to another, it must be a CR diffeomorphism, which would mean that U\partial U is locally spherical near pp^{*}. This contradicts Proposition 6.1. ∎

We conclude this section by a couple of remarks.

Remark 6.5.

Since the Bergman kernel forms of GG and UU are algebraic, it follows from the proof of Theorem 1.1 in Section 5 (see Step 2) that the coefficients of the logarithmic term in Fefferman’s expansions of KGK_{G} and KVK_{V} both vanish to infinite order at every smooth boundary point. The reduced normal Stein space GG gives the counterexample mentioned in Remark 1.2. The domain U3U\subset\mathbb{C}^{3} establishes the following fact, which implies that the Ramadanov conjecture fails for non-smooth domains in higher dimension.

There exists a bounded domain in 3\mathbb{C}^{3} with smooth, real-algebraic boundary away from a 1-dimensional complex curve such that every smooth boundary point is strongly pseudoconvex and non-spherical, while the coefficient of the logarithmic term in Fefferman’s asymptotic expansion of the Bergman kernel vanishes to infinite order at every smooth boundary point.

Remark 6.6.

Using the same idea as in the above example, we can actually construct significantly more general examples of higher dimensional domains in affine algebraic varieties VNV\subset\mathbb{C}^{N} with similar properties. Indeed, let XX be a compact Hermitian symmetric space of rank at least 22. Write

X=X1××Xt,t1,X=X_{1}\times\cdots\times X_{t},\quad t\geq 1,

where X1,,XtX_{1},\cdots,X_{t} are the irreducible factors of XX. Fix a Kähler -Einstein metric ωj\omega_{j} on XjX_{j} and let (L^j,h^j)(\widehat{L}_{j},\widehat{h}_{j}) be the top exterior product ΛnT1,0\Lambda^{n}T^{1,0} of the holomorphic tangent bundle over XjX_{j} with the metric induced from ωj\omega_{j}. Then there is a homogeneous line bundle (Lj,hj)(L_{j},h_{j}) with a Hermitian metric hjh_{j} such that its pjp_{j}-th tensor power gives (L^j,h^j)(\widehat{L}_{j},\widehat{h}_{j}), where pjp_{j} is the genus of Xj.X_{j}. (see [14] for more details).

Let πj\pi_{j} be the projection from XX onto the jj-th factor XjX_{j} for 1jt1\leq j\leq t. Define the line bundle LL over XX with a Hermitian metric hh to be:

(L,h):=π1(L1,h1)πt(Lt,ht).(L,h):=\pi_{1}^{*}(L_{1},h_{1})\otimes\cdots\otimes\pi_{t}^{*}(L_{t},h_{t}).

Let (L,h)(L^{*},h^{*}) be the dual line bundle of (L,h)(L,h). Write D(L)D(L^{*}) and S(L)S(L^{*}) for the associated unit disc and unit circle bundle. The specific example above is the special case t=2t=2 and X1=X2=1X_{1}=X_{2}=\mathbb{CP}^{1}. Proceeding as in that example, one finds that there is a canonical way to map LL^{*} to N\mathbb{C}^{N}, for some NN, induced by the minimal embedding of XX into some complex projective space (see [15]). If we denote this map LNL^{*}\to\mathbb{C}^{N} by FF (in the example above, the map FF is as given by (6.3)), then FF sends the zero section of LL^{*} to the point 0 and is a holomorphic embedding away from the zero section. It follows that the image of D(L)D(L^{*}) under the map FF is a domain GG with a singular point at 0. The boundary of GG is given by the image of S(L)S(L^{*}). It is not spherical since S(L)S(L^{*}) is not by [43]. Moreover, as the Bergman kernel form of D(L)D(L^{*}) is algebraic by [14], the Bergman kernel form of GG is also algebraic by Theorem 2.6.

7. Appendix

In this section, we will prove the claim that the Bergman kernel of the domain Ω\Omega in 3\mathbb{C}^{3} as defined in (6.4) is rational. This fact actually follows from a general theorem in [14] (see Theorem 3.3 in [14] and its proof). We include here a proof in this particular example for the convenience of readers and self-containedness of this paper. In fact, we shall compute the Bergman kernel of Ω\Omega explicitly (Theorem 7.3 below).

Recall that

(7.1) Ω:={(z,λ)=(z1,z2,λ)3:|λ|2(1+|z1|2)(1+|z2|2)<1}.\Omega:=\bigl{\{}(z,\lambda)=(z_{1},z_{2},\lambda)\in\mathbb{C}^{3}:|\lambda|^{2}(1+|z_{1}|^{2})(1+|z_{2}|^{2})<1\bigr{\}}.

We let

h(z):=(1+|z1|2)(1+|z2|2),h(z):=(1+|z_{1}|^{2})(1+|z_{2}|^{2}),

and denote the defining function by

ρ(z,λ):=|λ|2(1+|z1|2)(1+|z2|2)1.\rho(z,\lambda):=|\lambda|^{2}(1+|z_{1}|^{2})(1+|z_{2}|^{2})-1.

We recall that the Bergman space on Ω\Omega is defined as

(7.2) A2(Ω):={f(z,λ) is holomorphic in Ω:iΩ|f(z,λ)|2𝑑zdλdz¯dλ¯<},A^{2}(\Omega):=\bigl{\{}f(z,\lambda)\mbox{ is holomorphic in }\Omega:i\int_{\Omega}|f(z,\lambda)|^{2}dz\wedge d\lambda\wedge d\overline{z}\wedge d\overline{\lambda}<\infty\bigr{\}},

and let

(7.3) Am2(Ω):={f(z) is holomrphic in 2:λmf(z)A2(Ω)}.A^{2}_{m}(\Omega):=\bigl{\{}f(z)\text{ {\rm is holomrphic in $\mathbb{C}^{2}$}}\colon\lambda^{m}f(z)\in A^{2}(\Omega)\bigr{\}}.

Note that the L2L^{2} norm of λmf(z)\lambda^{m}f(z) is given by

λmf(z)2=\displaystyle\|\lambda^{m}f(z)\|^{2}= iΩ|λ|2m|f(z)|2𝑑zdλdz¯dλ¯\displaystyle i\int_{\Omega}|\lambda|^{2m}|f(z)|^{2}dz\wedge d\lambda\wedge d\overline{z}\wedge d\overline{\lambda}
=\displaystyle= z2(|λ|2<h(z)1|λ|2mi𝑑λdλ¯)|f(z)|2𝑑zdz¯\displaystyle\int_{z\in\mathbb{C}^{2}}\Bigl{(}\int_{|\lambda|^{2}<h(z)^{-1}}|\lambda|^{2m}i\,d\lambda\wedge d\overline{\lambda}\Bigr{)}|f(z)|^{2}dz\wedge d\overline{z}
=\displaystyle= z2(|λ|2<h(z)1|λ|2mi𝑑λdλ¯)|f(z)|2𝑑zdz¯.\displaystyle\int_{z\in\mathbb{C}^{2}}\Bigl{(}\int_{|\lambda|^{2}<h(z)^{-1}}|\lambda|^{2m}i\,d\lambda\wedge d\overline{\lambda}\Bigr{)}|f(z)|^{2}dz\wedge d\overline{z}.

We can rewrite the inner integral as follows:

|λ|2<h(z)1|λ|2mi𝑑λdλ¯=02π01h(z)r2m2r𝑑r𝑑θ=2π01h(z)rm𝑑r=2πm+1h(z)(m+1).\displaystyle\int_{|\lambda|^{2}<h(z)^{-1}}|\lambda|^{2m}i\,d\lambda\wedge d\overline{\lambda}=\int_{0}^{2\pi}\int_{0}^{\frac{1}{\sqrt{h(z)}}}r^{2m}2rdrd\theta=2\pi\int_{0}^{\frac{1}{h(z)}}r^{m}dr=\frac{2\pi}{m+1}h(z)^{-(m+1)}.

Thus,

(7.4) λmf(z)2=2πm+12|f(z)|2h(z)(m+1)𝑑zdz¯.\|\lambda^{m}f(z)\|^{2}=\frac{2\pi}{m+1}\int_{\mathbb{C}^{2}}|f(z)|^{2}h(z)^{-(m+1)}dz\wedge d\overline{z}.

If we introduce the weighted Bergman space

A2(2,h(m+1))={f(z) is holomorphic in 2:2|f(z)|2h(z)(m+1)𝑑zdz¯<},A^{2}(\mathbb{C}^{2},h^{-(m+1)})=\bigl{\{}f(z)\mbox{ is holomorphic in }\mathbb{C}^{2}:\int_{\mathbb{C}^{2}}|f(z)|^{2}h(z)^{-(m+1)}dz\wedge d\overline{z}<\infty\bigr{\}},

then

(7.5) Am2(Ω)={λmf(z):f(z)A2(2,h(m+1))}.A^{2}_{m}(\Omega)=\bigl{\{}\lambda^{m}f(z):f(z)\in A^{2}(\mathbb{C}^{2},h^{-(m+1)})\bigr{\}}.

We note that Am12(Ω)A_{m_{1}}^{2}(\Omega) and Am22(Ω)A_{m_{2}}^{2}(\Omega) are orthogonal to each other if m1m2m_{1}\neq m_{2}. We can therefore orthogonally decompose A2(Ω)A^{2}(\Omega) into a direct sum as follows.

Lemma 7.1.
A2(Ω)=m=0Am2(Ω).A^{2}(\Omega)=\bigoplus_{m=0}^{\infty}A_{m}^{2}(\Omega).
Proof.

Let f(z,λ)A2(Ω)f(z,\lambda)\in A^{2}(\Omega). If we fix z2z\in\mathbb{C}^{2}, then λ\lambda is contained in the disc {λ,|λ|2<h(z)1}\{\lambda\in\mathbb{C},|\lambda|^{2}<h(z)^{-1}\}. By taking the Taylor expansion at λ=0\lambda=0, we obtain

f(z,λ)=j=0aj(z)λj, for |λ|2<h(z)1,f(z,\lambda)=\sum_{j=0}^{\infty}a_{j}(z)\lambda^{j},\quad\mbox{ for }|\lambda|^{2}<h(z)^{-1},

where each aj(z)a_{j}(z) is holomorphic on 2\mathbb{C}^{2}. We shall first write f(z,λ)2\|f(z,\lambda)\|^{2} in terms of {aj(z)}j=0\{a_{j}(z)\}_{j=0}^{\infty}. We have

f(z,λ)2=\displaystyle\|f(z,\lambda)\|^{2}= z2(|λ|2<h1(z)|f(z,λ)|2i𝑑λdλ¯)𝑑zdz¯.\displaystyle\int_{z\in\mathbb{C}^{2}}\left(\int_{|\lambda|^{2}<h^{-1}(z)}|f(z,\lambda)|^{2}i\,d\lambda\wedge d\overline{\lambda}\right)dz\wedge d\overline{z}.

The inner integral can be computed as

|λ|2<h1(z)|f(z,λ)|2i𝑑λdλ¯=\displaystyle\int_{|\lambda|^{2}<h^{-1}(z)}|f(z,\lambda)|^{2}i\,d\lambda\wedge d\overline{\lambda}= limε0+s=0t=0as(z)at(z)¯|λ|2<h1(z)ελsλt¯i𝑑λdλ¯\displaystyle\lim_{\varepsilon\rightarrow 0^{+}}\sum_{s=0}^{\infty}\sum_{t=0}^{\infty}a_{s}(z)\overline{a_{t}(z)}\int_{|\lambda|^{2}<h^{-1}(z)-\varepsilon}\lambda^{s}\overline{\lambda^{t}}i\,d\lambda\wedge d\overline{\lambda}
=\displaystyle= limε0+j=02πj+1|aj(z)|2(h(z)1ε)j+1\displaystyle\lim_{\varepsilon\rightarrow 0^{+}}\sum_{j=0}^{\infty}\frac{2\pi}{j+1}|a_{j}(z)|^{2}\bigl{(}h(z)^{-1}-\varepsilon\bigr{)}^{j+1}
=\displaystyle= j=02πj+1|aj(z)|2h(z)(j+1),\displaystyle\sum_{j=0}^{\infty}\frac{2\pi}{j+1}|a_{j}(z)|^{2}h(z)^{-(j+1)},

where the last equality follows from the Monotone Convergence theorem. Therefore,

f(z,λ)2=\displaystyle\|f(z,\lambda)\|^{2}= j=02πj+1z2|aj(z)|2h(z)(j+1)𝑑zdz¯,\displaystyle\sum_{j=0}^{\infty}\frac{2\pi}{j+1}\int_{z\in\mathbb{C}^{2}}|a_{j}(z)|^{2}h(z)^{-(j+1)}dz\wedge d\overline{z},

which immediately implies that each aj(z)a_{j}(z) is contained in A2(2,h(j+1))A^{2}(\mathbb{C}^{2},h^{-(j+1)}).

Suppose f(z,λ)Am2(Ω)f(z,\lambda)\perp A^{2}_{m}(\Omega). Then for any λmg(z)Am2(Ω)\lambda^{m}g(z)\in A^{2}_{m}(\Omega),

0=\displaystyle 0= iΩf(z,λ)λmg(z)¯𝑑zdλdz¯dλ¯\displaystyle i\int_{\Omega}f(z,\lambda)\overline{\lambda^{m}g(z)}dz\wedge d\lambda\wedge d\overline{z}\wedge d\overline{\lambda}
=\displaystyle= z2(|λ|2<h(z)1f(z,λ)λm¯i𝑑λdλ¯)g(z)¯𝑑zdz¯.\displaystyle\int_{z\in\mathbb{C}^{2}}\left(\int_{|\lambda|^{2}<h(z)^{-1}}f(z,\lambda)\overline{\lambda^{m}}\,i\,d\lambda\wedge d\overline{\lambda}\right)\overline{g(z)}dz\wedge d\overline{z}.

The inner integral can be computed as follows

|λ|2<h1(z)f(z,λ)λm¯i𝑑λdλ¯=\displaystyle\int_{|\lambda|^{2}<h^{-1}(z)}f(z,\lambda)\overline{\lambda^{m}}i\,d\lambda\wedge d\overline{\lambda}= limε0+j=0aj(z)|λ|2<h(z)1ελjλm¯i𝑑λdλ¯\displaystyle\lim_{\varepsilon\rightarrow 0^{+}}\sum_{j=0}^{\infty}a_{j}(z)\int_{|\lambda|^{2}<h(z)^{-1}-\varepsilon}\lambda^{j}\overline{\lambda^{m}}i\,d\lambda\wedge d\overline{\lambda}
=\displaystyle= limε0+2πm+1am(z)(h(z)1ε)m+1\displaystyle\lim_{\varepsilon\rightarrow 0^{+}}\frac{2\pi}{m+1}a_{m}(z)\bigl{(}h(z)^{-1}-\varepsilon\bigr{)}^{m+1}
=\displaystyle= 2πm+1am(z)h(z)(m+1).\displaystyle\frac{2\pi}{m+1}a_{m}(z)h(z)^{-(m+1)}.

Therefore,

0=\displaystyle 0= 2πm+1z2am(z)g(z)¯h(z)(m+1)𝑑zdz¯, for any gA2(2,h(m+1)).\displaystyle\frac{2\pi}{m+1}\int_{z\in\mathbb{C}^{2}}a_{m}(z)\overline{g(z)}h(z)^{-(m+1)}dz\wedge d\overline{z},\quad\mbox{ for any }g\in A^{2}(\mathbb{C}^{2},h^{-(m+1)}).

Since ama_{m} belongs to the space A2(2,h(m+1))A^{2}(\mathbb{C}^{2},h^{-(m+1)}), we get am=0a_{m}=0. Therefore, the direct sum of Am2(Ω)A_{m}^{2}(\Omega) for 0m<0\leq m<\infty generates A2(Ω)A^{2}(\Omega). ∎

Since Am2(Ω)A^{2}_{m}(\Omega) can be identified with the weighted Bergman space A2(2,h(m+1))A^{2}(\mathbb{C}^{2},h^{-(m+1)}) as in (7.5), we can find an explicit orthonormal basis and compute its reproducing kernel.

Proposition 7.2.

Let m1.m\geq 1. The reproducing kernel of Am2(Ω)A^{2}_{m}(\Omega) is

(7.6) Km(z,λ,w¯,τ¯)=(m+1)m2(2π)3λmτ¯m(1+z1w1¯)m1(1+z2w2¯)m1,K_{m}^{*}(z,\lambda,\overline{w},\overline{\tau})=\frac{(m+1)m^{2}}{(2\pi)^{3}}\lambda^{m}\overline{\tau}^{m}(1+z_{1}\overline{w_{1}})^{m-1}(1+z_{2}\overline{w_{2}})^{m-1},

where (z,λ),(w,τ)(z,\lambda),(w,\tau) are points in Ω\Omega.

Proof.

Denote

zα=z1α1z2α2, for any multi-index α=(α1,α2)02.z^{\alpha}=z_{1}^{\alpha_{1}}z_{2}^{\alpha_{2}},\quad\mbox{ for any multi-index }\alpha=(\alpha_{1},\alpha_{2})\in\mathbb{Z}_{\geq 0}^{2}.

By (7.5), since Ω\Omega is Reinhardt, it is easy to see that

{λmzα:zαA2(2,h(m+1))}\bigl{\{}\lambda^{m}z^{\alpha}:z^{\alpha}\in A^{2}(\mathbb{C}^{2},h^{-(m+1)})\bigr{\}}

forms an orthogonal basis of Am2(Ω)A^{2}_{m}(\Omega). We shall compute the norm for each λmzα\lambda^{m}z^{\alpha}. Using (7.4), we have

λmzα2=\displaystyle\|\lambda^{m}z^{\alpha}\|^{2}= 2πm+12|zα|2(1+|z1|2)(m+1)(1+|z2|2)(m+1)𝑑zdz¯\displaystyle\frac{2\pi}{m+1}\int_{\mathbb{C}^{2}}|z^{\alpha}|^{2}(1+|z_{1}|^{2})^{-(m+1)}(1+|z_{2}|^{2})^{-(m+1)}dz\wedge d\overline{z}
=\displaystyle= 2πm+1|z1|2α1(1+|z1|2)(m+1)i𝑑z1dz1¯|z2|2α2(1+|z2|2)(m+1)i𝑑z2dz2¯\displaystyle\frac{2\pi}{m+1}\int_{\mathbb{C}}|z_{1}|^{2\alpha_{1}}(1+|z_{1}|^{2})^{-(m+1)}\,i\,dz_{1}\wedge d\overline{z_{1}}\cdot\int_{\mathbb{C}}|z_{2}|^{2\alpha_{2}}(1+|z_{2}|^{2})^{-(m+1)}\,i\,dz_{2}\wedge d\overline{z_{2}}
=\displaystyle= (2π)3m+10r1α1(1+r1)(m+1)𝑑r10r2α2(1+r2)(m+1)𝑑r2.\displaystyle\frac{(2\pi)^{3}}{m+1}\int_{0}^{\infty}r_{1}^{\alpha_{1}}(1+r_{1})^{-(m+1)}dr_{1}\cdot\int_{0}^{\infty}r_{2}^{\alpha_{2}}(1+r_{2})^{-(m+1)}dr_{2}.

By the elementary integral identity

(7.7) 0rp1(1+r)q𝑑r=(qp2)!p!(q1)!, for any nonnegative integers p,q with qp+2,\int_{0}^{\infty}r^{p}\frac{1}{(1+r)^{q}}dr=\frac{(q-p-2)!p!}{(q-1)!},\quad\mbox{ for any nonnegative integers $p,q$ with }q\geq p+2,

we get

λmzα2={(2π)3m+1(mα11)!(mα21)!α!m!2if α1,α2m1,+otherwise.\|\lambda^{m}z^{\alpha}\|^{2}=\begin{dcases}\frac{(2\pi)^{3}}{m+1}\frac{(m-\alpha_{1}-1)!(m-\alpha_{2}-1)!\alpha!}{m!^{2}}&\mbox{if }\alpha_{1},\alpha_{2}\leq m-1,\\ +\infty&\mbox{otherwise}.\end{dcases}

Thus, {λmzαλmzα:0α1,α2m1}\bigl{\{}\frac{\lambda^{m}z^{\alpha}}{\|\lambda^{m}z^{\alpha}\|}:0\leq\alpha_{1},\alpha_{2}\leq m-1\bigr{\}} is an orthonormal basis of Am2(Ω)A^{2}_{m}(\Omega), and the reproducing kernel of Am2(Ω)A^{2}_{m}(\Omega) is given by

Km(z,λ,w¯,τ¯)\displaystyle K_{m}^{*}(z,\lambda,\overline{w},\overline{\tau}) =0α1,α2m1zαλmw¯ατ¯mzαλm2\displaystyle=\sum_{0\leq\alpha_{1},\alpha_{2}\leq m-1}\frac{z^{\alpha}\lambda^{m}\overline{w}^{\alpha}\overline{\tau}^{m}}{\|z^{\alpha}\lambda^{m}\|^{2}}
=(m+1)m2(2π)3λmτ¯mα1=0m1(m1α1)z1α1w1¯α1α2=0m1(m1α2)z2α2w2¯α2\displaystyle=\frac{(m+1)m^{2}}{(2\pi)^{3}}\lambda^{m}\overline{\tau}^{m}\sum_{\alpha_{1}=0}^{m-1}\binom{m-1}{\alpha_{1}}z_{1}^{\alpha_{1}}\overline{w_{1}}^{\alpha_{1}}\sum_{\alpha_{2}=0}^{m-1}\binom{m-1}{\alpha_{2}}z_{2}^{\alpha_{2}}\overline{w_{2}}^{\alpha_{2}}
=(m+1)m2(2π)3λmτ¯m(1+z1w1¯)m1(1+z2w2¯)m1.\displaystyle=\frac{(m+1)m^{2}}{(2\pi)^{3}}\lambda^{m}\overline{\tau}^{m}(1+z_{1}\overline{w_{1}})^{m-1}(1+z_{2}\overline{w_{2}})^{m-1}.

Now we are ready to compute the Bergman kernel form of Ω\Omega.

Theorem 7.3.

The Bergman kernel form of the domain Ω3\Omega\subset\mathbb{C}^{3} in (7.1)\eqref{eq:Omega} is given by

KΩ(z,λ,w¯,τ¯)=iK(z,λ,w¯,τ¯)dzdλdw¯dτ¯,K_{\Omega}(z,\lambda,\overline{w},\overline{\tau})=iK^{*}(z,\lambda,\overline{w},\overline{\tau})dz\wedge d\lambda\wedge d\overline{w}\wedge d\overline{\tau},

where

K(z,λ,w¯,τ¯)=m=1(m+1)m2(2π)3λmτ¯m(1+z1w1¯)m1(1+z2w2¯)m1.K^{*}(z,\lambda,\overline{w},\overline{\tau})=\sum_{m=1}^{\infty}\frac{(m+1)m^{2}}{(2\pi)^{3}}\lambda^{m}\overline{\tau}^{m}(1+z_{1}\overline{w_{1}})^{m-1}(1+z_{2}\overline{w_{2}})^{m-1}.

It can be written in terms of the complexified defining function

ρ(z,λ,w¯,τ¯)=λτ¯(1+z1w1¯)(1+z2w2¯)1\rho(z,\lambda,\overline{w},\overline{\tau})=\lambda\overline{\tau}(1+z_{1}\overline{w_{1}})(1+z_{2}\overline{w_{2}})-1

as

K(z,λ,w¯,τ¯)=1(2π)3(4λτ¯ρ(z,λ,w¯,τ¯)3+6λτ¯ρ(z,λ,w¯,τ¯)4).K^{*}(z,\lambda,\overline{w},\overline{\tau})=\frac{1}{(2\pi)^{3}}\Bigl{(}\frac{4\lambda\overline{\tau}}{\rho(z,\lambda,\overline{w},\overline{\tau})^{3}}+\frac{6\lambda\overline{\tau}}{\rho(z,\lambda,\overline{w},\overline{\tau})^{4}}\Bigr{)}.
Proof.

By Lemma 7.1, we immediately get the reproducing kernel of A2(Ω)A^{2}(\Omega) by adding up the reproducing kernels of Am2(Ω)A^{2}_{m}(\Omega) for all mm. Since A02(Ω)={0}A^{2}_{0}(\Omega)=\{0\}, we obtain

K(z,λ,w¯,τ¯)=\displaystyle K^{*}(z,\lambda,\overline{w},\overline{\tau})= m=1Km(z,λ,w¯,τ¯)\displaystyle\sum_{m=1}^{\infty}K_{m}^{*}(z,\lambda,\overline{w},\overline{\tau})
=\displaystyle= m=0(m+2)(m+1)2(2π)3λm+1τ¯m+1(1+z1w1¯)m(1+z2w2¯)m.\displaystyle\sum_{m=0}^{\infty}\frac{(m+2)(m+1)^{2}}{(2\pi)^{3}}\lambda^{m+1}\overline{\tau}^{m+1}(1+z_{1}\overline{w_{1}})^{m}(1+z_{2}\overline{w_{2}})^{m}.

It remains to write K(z,λ,w¯,τ¯)K^{*}(z,\lambda,\overline{w},\overline{\tau}) in terms of the defining function ρ(z,λ,w¯,τ¯)\rho(z,\lambda,\overline{w},\overline{\tau}). We use the Taylor expansion of 1/(1x)j+11/(1-x)^{j+1} for 0j30\leq j\leq 3 to obtain

1(ρ(z,λ,w¯,τ¯))j+1=1(1(1+z1w1¯)(1+z2w2¯)λτ¯)j+1=m=0(m+jj)(1+z1w1¯)m(1+z2w2¯)mλmτ¯m.\frac{1}{(-\rho(z,\lambda,\overline{w},\overline{\tau}))^{j+1}}=\frac{1}{(1-(1+z_{1}\overline{w_{1}})(1+z_{2}\overline{w_{2}})\lambda\overline{\tau})^{j+1}}=\sum_{m=0}^{\infty}\binom{m+j}{j}(1+z_{1}\overline{w_{1}})^{m}(1+z_{2}\overline{w_{2}})^{m}\lambda^{m}\overline{\tau}^{m}.

Note that (m+2)(m+1)2(m+2)(m+1)^{2} is a polynomial in mm of degree 33. Since {(m+jj)}j=03\{\binom{m+j}{j}\}_{j=0}^{3} is a basis of polynomials in mm with degree 3\leq 3, we can write

(m+2)(m+1)2=j=03aj(m+jj).(m+2)(m+1)^{2}=\sum_{j=0}^{3}a_{j}\binom{m+j}{j}.

One can check that the coefficients are given by a0=a1=0a_{0}=a_{1}=0, a2=4a_{2}=-4 and a3=6a_{3}=6. Therefore,

K(z,λ,w¯,τ¯)=\displaystyle K^{*}(z,\lambda,\overline{w},\overline{\tau})= 1(2π)3m=0j=03aj(m+jj)λm+1τ¯m+1(1+z1w1¯)m(1+z2w2¯)m\displaystyle\frac{1}{(2\pi)^{3}}\sum_{m=0}^{\infty}\sum_{j=0}^{3}a_{j}\binom{m+j}{j}\lambda^{m+1}\overline{\tau}^{m+1}(1+z_{1}\overline{w_{1}})^{m}(1+z_{2}\overline{w_{2}})^{m}
=\displaystyle= 1(2π)3j=03ajλτ¯(ρ(z,λ,w¯,τ¯))j+1,\displaystyle\frac{1}{(2\pi)^{3}}\sum_{j=0}^{3}a_{j}\frac{\lambda\overline{\tau}}{(-\rho(z,\lambda,\overline{w},\overline{\tau}))^{j+1}},

and the result follows. ∎

References

  • [1] Toby N. Bailey, Michael G. Eastwood, and C. Robin Graham. Invariant theory for conformal and CR geometry. Ann. of Math. (2), 139(3):491–552, 1994.
  • [2] M. Salah Baouendi, Peter Ebenfelt, and Linda Preiss Rothschild. Real submanifolds in complex space and their mappings, volume 47 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1999.
  • [3] Steven R. Bell. Proper holomorphic mappings and the Bergman projection. Duke Math. J., 48(1):167–175, 1981.
  • [4] Steven R. Bell. The Bergman kernel function and proper holomorphic mappings. Trans. Amer. Math. Soc., 270(2):685–691, 1982.
  • [5] Steven R. Bell. Proper holomorphic mappings that must be rational. Trans. Amer. Math. Soc., 284(1):425–429, 1984.
  • [6] Stefan Bergmann. Über die Kernfunktion eines Bereiches und ihr Verhalten am Rande. I. J. Reine Angew. Math., 169:1–42, 1933.
  • [7] Stefan Bergmann. Über die Kernfunktion eines Bereiches und ihr Verhalten am Rande. II. J. Reine Angew. Math., 172:89–128, 1935.
  • [8] Louis Boutet de Monvel. Singularity of the Bergman kernel. In Complex geometry (Osaka, 1990), volume 143 of Lecture Notes in Pure and Appl. Math., pages 13–29. Dekker, New York, 1993.
  • [9] Louis Boutet de Monvel and Johannes Sjöstrand. Sur la singularité des noyaux de Bergman et de Szegö. In Journées: Équations aux Dérivées Partielles de Rennes (1975), pages 123–164. Astérisque, No. 34–35. 1976.
  • [10] D. Burns, Jr. and S. Shnider. Spherical hypersurfaces in complex manifolds. Invent. Math., 33(3):223–246, 1976.
  • [11] Henri Cartan. Quotient d’un espace analytique par un groupe d’automorphismes. In Algebraic geometry and topology, pages 90–102. Princeton University Press, Princeton, N. J., 1957. A symposium in honor of S. Lefschetz,.
  • [12] Shiu-Yuen Cheng. Open Problems, Conference on Nonlinear Problems in Geometry held in Katata, Sep. 3-8, 1979. Tohoku University, Department of Math., Sendai, 1979.
  • [13] Peter Ebenfelt, Ming Xiao, and Hang Xu. in preparation, 2020.
  • [14] Miroslav Engliš and Genkai Zhang. Ramadanov conjecture and line bundles over compact Hermitian symmetric spaces. Math. Z., 264(4):901–912, 2010.
  • [15] Hanlong Fang, Xiaojun Huang, and Ming Xiao. Volume-preserving mappings between Hermitian symmetric spaces of compact type. Adv. Math., 360:106885, 74, 2020.
  • [16] Charles Fefferman. The Bergman kernel and biholomorphic mappings of pseudoconvex domains. Invent. Math., 26:1–65, 1974.
  • [17] Charles Fefferman. Monge-Ampère equations, the Bergman kernel, and geometry of pseudoconvex domains. Ann. of Math. (2), 103(2):395–416, 1976.
  • [18] Charles Fefferman. Parabolic invariant theory in complex analysis. Adv. in Math., 31(2):131–262, 1979.
  • [19] Franc Forstnerič. Proper holomorphic maps from balls. Duke Math. J., 53(2):427–441, 1986.
  • [20] Franc Forstnerič and Jean-Pierre Rosay. Localization of the Kobayashi metric and the boundary continuity of proper holomorphic mappings. Math. Ann., 279(2):239–252, 1987.
  • [21] Siqi Fu and Bun Wong. On strictly pseudoconvex domains with Kähler-Einstein Bergman metrics. Math. Res. Lett., 4(5):697–703, 1997.
  • [22] C. Robin Graham. Scalar boundary invariants and the Bergman kernel. In Complex analysis, II (College Park, Md., 1985–86), volume 1276 of Lecture Notes in Math., pages 108–135. Springer, Berlin, 1987.
  • [23] C. Robin Graham. Scalar boundary invariants and the Bergman kernel. In Complex analysis, II (College Park, Md., 1985–86), volume 1276 of Lecture Notes in Math., pages 108–135. Springer, Berlin, 1987.
  • [24] Phillip Griffiths and Joseph Harris. Principles of algebraic geometry. Wiley Classics Library. John Wiley & Sons, Inc., New York, 1994. Reprint of the 1978 original.
  • [25] Kengo Hirachi. Construction of boundary invariants and the logarithmic singularity of the Bergman kernel. Ann. of Math. (2), 151(1):151–191, 2000.
  • [26] Xiaojun Huang. On the mapping problem for algebraic real hypersurfaces in the complex spaces of different dimensions. Ann. Inst. Fourier (Grenoble), 44(2):433–463, 1994.
  • [27] Xiaojun Huang. Schwarz reflection principle in complex spaces of dimension two. Comm. Partial Differential Equations, 21(11-12):1781–1828, 1996.
  • [28] Xiaojun Huang. Isolated complex singularities and their CR links. Sci. China Ser. A, 49(11):1441–1450, 2006.
  • [29] Xiaojun Huang and Shanyu Ji. Global holomorphic extension of a local map and a Riemann mapping theorem for algebraic domains. Math. Res. Lett., 5(1-2):247–260, 1998.
  • [30] Xiaojun Huang and Shanyu Ji. Cartan-Chern-Moser theory on algebraic hypersurfaces and an application to the study of automorphism groups of algebraic domains. Ann. Inst. Fourier (Grenoble), 52(6):1793–1831, 2002.
  • [31] Xiaojun Huang and Xiaoshan Li. Bergman-Einstein metric on a Stein space with a strongly pseudoconvex boundary. 2020.
  • [32] Xiaojun Huang and Ming Xiao. Bergman-Einstein metrics, hyperbolic metrics and Stein spaces with spherical boundaries, 2016.
  • [33] Shoshichi Kobayashi. Geometry of bounded domains. Transactions of the American Mathematical Society, 92(2):267–290, 1959.
  • [34] Shoshichi Kobayashi and Katsumi Nomizu. Foundations of differential geometry. Vol. II. Wiley Classics Library. John Wiley & Sons, Inc., New York, 1996. Reprint of the 1969 original, A Wiley-Interscience Publication.
  • [35] László Lempert. Algebraic approximations in analytic geometry. Invent. Math., 121(2):335–353, 1995.
  • [36] Qi-Keng Lu. On Kaehler manifolds with constant curvature. Chinese Math.–Acta, 8:283–298, 1966.
  • [37] Tejinder S. Neelon. On solutions of real analytic equations. Proc. Amer. Math. Soc., 125(9):2531–2535, 1997.
  • [38] Stefan Yu Nemirovski and Ruslan G. Shafikov. Conjectures of Cheng and Ramadanov. Uspekhi Mat. Nauk, 61(4(370)):193–194, 2006.
  • [39] Sergey Ivanovich Pinchuk and Sh I. Tsyganov. The smoothness of CR-mappings between strictly pseudoconvex hypersurfaces. Mathematics of the USSR-Izvestiya, 35(2):457, 1990.
  • [40] I. P. Ramadanov. A characterization of the balls in 𝐂n{\bf C}^{n} by means of the Bergman kernel. C. R. Acad. Bulgare Sci., 34(7):927–929, 1981.
  • [41] Walter Rudin. Proper holomorphic maps and finite reflection groups. Indiana Univ. Math. J., 31(5):701–720, 1982.
  • [42] Igor R. Shafarevich. Basic algebraic geometry. 1. Springer, Heidelberg, third edition, 2013. Varieties in projective space.
  • [43] Xiaodong Wang. Some recent results in CR geometry. Proc. of the 3rd Conference of Tsinghua Sanya Math Forum.