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

Some remarks on the Kobayashi–Fuks metric on strongly pseudoconvex domains

Diganta Borah Indian Institute of Science Education and Research, Pune, India dborah@iiserpune.ac.in Debaprasanna Kar111The second author was supported by the Council of Scientific & Industrial Research (File no. 09/936(0221)/2019-EMR-I) and by the Ph.D. program at the Indian Institute of Science Education and Research, Pune. Indian Institute of Science Education and Research, Pune, India debaprasanna.kar@students.iiserpune.ac.in
Abstract

The Ricci curvature of the Bergman metric on a bounded domain DnD\subset\mathbb{C}^{n} is strictly bounded above by n+1n+1 and consequently log(KDn+1gB,D)\log(K_{D}^{n+1}g_{B,D}), where KDK_{D} is the Bergman kernel for DD on the diagonal and gB,Dg_{B,D} is the Riemannian volume element of the Bergman metric on DD, is the potential for a Kähler metric on DD known as the Kobayashi–Fuks metric. In this note we study the localization of this metric near holomorphic peak points and also show that this metric shares several properties with the Bergman metric on strongly pseudoconvex domains.

keywords:
Kobayashi–Fuks metric , Bergman kernel
MSC:
[2020] 32F45 , 32A36 , 32A25

1 Introduction

For a bounded domain DnD\subset\mathbb{C}^{n} the space

A2(D)={f:D holomorphic and fD2:=D|f|2𝑑V<},A^{2}(D)=\left\{\text{$f:D\to\mathbb{C}$ holomorphic and $\|f\|^{2}_{D}:=\int_{D}|f|^{2}\,dV<\infty$}\right\},

where dVdV is the Lebesgue measure on n\mathbb{C}^{n} is a closed subspace of L2(D)L^{2}(D) and is a reproducing kernel Hilbert space. The associated reproducing kernel denoted by KD(z,w)K_{D}(z,w) is uniquely determined by the following properties: KD(,w)A2(D)K_{D}(\cdot,w)\in A^{2}(D) for each wDw\in D, it is anti-symmetric, i.e., KD(z,w)=KD(w,z)¯K_{D}(z,w)=\overline{K_{D}(w,z)}, and it reproduces A2(D)A^{2}(D):

f(w)=Df(z)KD(z,w)¯𝑑V(z),fA2(D).f(w)=\int_{D}f(z)\overline{K_{D}(z,w)}\,dV(z),\quad f\in A^{2}(D).

It also follows that for any complete orthonormal basis {ϕk}\{\phi_{k}\} of A2(D)A^{2}(D),

KD(z,w)=kϕk(z)ϕk(w)¯,K_{D}(z,w)=\sum_{k}\phi_{k}(z)\overline{\phi_{k}(w)},

where the series converges uniformly on compact subsets of D×DD\times D. The reproducing kernel KD(z,w)K_{D}(z,w) is called the Bergman kernel for DD. Denote by KD(z)=KD(z,z)K_{D}(z)=K_{D}(z,z) its restriction to the diagonal. It is known that logKD\log K_{D} is a strongly plurisubharmonic function and thus is a potential for a Kähler metric which is called the Bergman metric for DD and is given by

dsB,D2=α,β=1ngαβ¯B,D(z)dzαdz¯β,ds^{2}_{B,D}=\sum_{\alpha,\beta=1}^{n}g^{B,D}_{\alpha\overline{\beta}}(z)\,dz_{\alpha}d\overline{z}_{\beta},

where

gαβ¯B,D(z)=2logKDzαz¯β(z).g^{B,D}_{\alpha\overline{\beta}}(z)=\frac{\partial^{2}\log K_{D}}{\partial z_{\alpha}\partial\overline{z}_{\beta}}(z).

Let

GB,D(z)=(gαβ¯B,D(z))n×nandgB,D(z)=detGB,D(z).G_{B,D}(z)=\begin{pmatrix}g^{B,D}_{\alpha\overline{\beta}}(z)\end{pmatrix}_{n\times n}\quad\text{and}\quad g_{B,D}(z)=\det G_{B,D}(z).

The components of the Ricci tensor of dsB,D2ds^{2}_{B,D} are defined by

Ricαβ¯B,D(z)=2loggB,Dzαz¯β(z),\operatorname{Ric}_{\alpha\overline{\beta}}^{B,D}(z)=-\frac{\partial^{2}\log g_{B,D}}{\partial z_{\alpha}\partial\overline{z}_{\beta}}(z), (1.1)

and the Ricci curvature of dsB,D2ds^{2}_{B,D} is given by

RicB,D(z,u)=α,β=1nRicαβ¯B,D(z)uαu¯βα,β=1ngαβ¯B,D(z)uαu¯β.\text{Ric}_{B,D}(z,u)=\frac{\sum_{\alpha,\beta=1}^{n}\operatorname{Ric}_{\alpha\overline{\beta}}^{B,D}(z)u^{\alpha}\overline{u}^{\beta}}{\sum_{\alpha,\beta=1}^{n}g^{B,D}_{\alpha\overline{\beta}}(z)u^{\alpha}\overline{u}^{\beta}}. (1.2)

Kobayashi [1] showed that the Ricci curvature of the Bergman metric on a bounded domain in n\mathbb{C}^{n} is strictly bounded above by n+1n+1 and hence the matrix

GB~,D(z)=(gαβ¯B~,D(z))n×nG_{\tilde{B},D}(z)=\begin{pmatrix}g^{\tilde{B},D}_{\alpha\overline{\beta}}(z)\end{pmatrix}_{n\times n}

where

gαβ¯B~,D(z)=(n+1)gαβ¯B,D(z)Ricαβ¯B,D(z)=2log(KDn+1gB,D)zαz¯β(z),g^{\tilde{B},D}_{\alpha\overline{\beta}}(z)=(n+1)g^{B,D}_{\alpha\overline{\beta}}(z)-\operatorname{Ric}^{B,D}_{\alpha\overline{\beta}}(z)=\frac{\partial^{2}\log(K_{D}^{n+1}g_{B,D})}{\partial z_{\alpha}\partial\overline{z}_{\beta}}(z),

is positive definite (see also Fuks [2]). Therefore,

dsB~,D2=α,β=1ngαβ¯B~,D(z)dzαdz¯βds^{2}_{\tilde{B},D}=\sum_{\alpha,\beta=1}^{n}g^{\tilde{B},D}_{\alpha\overline{\beta}}(z)\,dz_{\alpha}d\overline{z}_{\beta}

is a Kähler metric with Kähler potential log(KDn+1gB,D)\log(K_{D}^{n+1}g_{B,D}). Moreover, if F:DDF:D\to D^{\prime} is a biholomorphism, then

GB~,D(z)=F(z)tGB~,D(F(z))F¯(z),G_{\tilde{B},D}(z)=F^{\prime}(z)^{t}\,G_{\tilde{B},D^{\prime}}\big{(}F(z)\big{)}\overline{F}^{\prime}(z), (1.3)

where F(z)F^{\prime}(z) is the Jacobian matrix of FF at zz. This implies that dsB~,D2ds^{2}_{\tilde{B},D} is an invariant metric. We will call this metric the Kobayashi–Fuks metric on DD.

The boundary asymptotics of the Bergman metric and its Ricci curvature on strongly pseudoconvex domains are known from which it turns out that the Kobayashi–Fuks metric is complete on such domains (to be seen later in Section 5). Dinew [3] showed that on any bounded hyperconvex domain the Kobayashi–Fuks metric is complete, and hence in particular, by a result of Demailly [4], this metric is complete on any bounded pseudoconvex domain with Lipschitz boundary. Dinew [5] also observed that the Kobayashi–Fuks metric is useful in the study of Bergman representative coordinates. Invariant metrics play an important role in understanding the geometry of a domain which makes their study of natural interest in complex analysis and the purpose of this note is to show that the Kobayashi–Fuks metric shares several properties with the Bergman metric. Let us fix some notations before we state our results. We will denote by hh any of BB or B~\tilde{B}. We write

Gh,D(z)=(gαβ¯h,D(z))n×nandgh,D(z)=detGh,D(z).G_{h,D}(z)=\begin{pmatrix}g^{h,D}_{\alpha\overline{\beta}}(z)\end{pmatrix}_{n\times n}\quad\text{and}\quad g_{h,D}(z)=\det G_{h,D}(z).

The length of a vector uu at a point zDz\in D in dsh,D2ds^{2}_{h,D} will be denoted by τh,D(z,u)\tau_{h,D}(z,u), i.e.,

τh,D2(z,u)=α,β=1ngαβ¯h,D(z)uαu¯β.\tau_{h,D}^{2}(z,u)=\sum_{\alpha,\beta=1}^{n}g^{h,D}_{\alpha\overline{\beta}}(z)u^{\alpha}\overline{u}^{\beta}.

The holomorphic sectional curvature of dsh,D2ds^{2}_{h,D} is defined by

Rh,D(z,u)=α,β,γ,δ=1nRα¯βγδ¯h,D(z)u¯αuβuγu¯δ(α,β=1ngαβ¯h,D(z)uαu¯β)2,R_{h,D}(z,u)=\frac{\sum_{\alpha,\beta,\gamma,\delta=1}^{n}R^{h,D}_{\overline{\alpha}\beta\gamma\overline{\delta}}(z)\overline{u}^{\alpha}u^{\beta}u^{\gamma}\overline{u}^{\delta}}{\big{(}\sum_{\alpha,\beta=1}^{n}g^{h,D}_{\alpha\overline{\beta}}(z)u^{\alpha}\overline{u}^{\beta}\big{)}^{2}}, (1.4)

where

Rα¯βγδ¯h,D(z)=2gβα¯h,Dzγz¯δ(z)+μ,νgh,Dνμ¯(z)gβμ¯h,Dzγ(z)gνα¯h,Dz¯δ(z),R^{h,D}_{\overline{\alpha}\beta\gamma\overline{\delta}}(z)=-\frac{\partial^{2}g^{h,D}_{\beta\overline{\alpha}}}{\partial z_{\gamma}\partial\overline{z}_{\delta}}(z)+\sum_{\mu,\nu}g_{h,D}^{\nu\overline{\mu}}(z)\frac{\partial g^{h,D}_{\beta\overline{\mu}}}{\partial z_{\gamma}}(z)\frac{\partial g^{h,D}_{\nu\overline{\alpha}}}{\partial\overline{z}_{\delta}}(z), (1.5)

gh,Dνμ¯(z)g_{h,D}^{\nu\overline{\mu}}(z) being the (ν,μ)(\nu,\mu)th entry of the inverse of the matrix Gh,D(z)G_{h,D}(z). The Ricci curvature of dsh,D2ds^{2}_{h,D} is defined by (1.2) with BB replaced by hh. Finally, note that in dimension one, the metric dsh,D2ds^{2}_{h,D} has the form

dsh,D2=gh,D(z)|dz|2,τh,D(z,u)=gh,D(z)|u|ds^{2}_{h,D}=g_{h,D}(z)|dz|^{2},\quad\tau_{h,D}(z,u)=\sqrt{g_{h,D}(z)}|u|

and both the holomorphic sectional curvature and the Ricci curvature at a point are independent of the tangent vector uu and are simply the Gaussian curvature

Rh,D(z)=1gh,D(z)2loggh,Dzz¯(z).R_{h,D}(z)=-\frac{1}{g_{h,D}(z)}\frac{\partial^{2}\log g_{h,D}}{\partial z\partial\overline{z}}(z).

Our first result is on the localization of the Kobayashi–Fuks metric near holomorphic peak points.

Theorem 1.1

Let DnD\subset\mathbb{C}^{n} be a bounded pseudoconvex domain with a holomorphic peak point p0Dp^{0}\in\partial D. If UU is a sufficiently small neighborhood of p0p^{0}, then

  • (I)

    limzp0τB~,D(z,u)τB~,UD(z,u)=1\lim_{z\to p^{0}}\frac{\tau_{\tilde{B},D}(z,u)}{\tau_{\tilde{B},U\cap D}(z,u)}=1 uniformly in unit vectors unu\in\mathbb{C}^{n}.

  • (II)

    limzp0gB~,D(z)gB~,UD(z)=1\lim_{z\to p^{0}}\frac{g_{\tilde{B},D}(z)}{g_{\tilde{B},U\cap D}(z)}=1.

  • (III)

    If n=1n=1, then limzp02RB~,D(z)2RB~,UD(z)=1\lim_{z\to p^{0}}\frac{2-R_{\tilde{B},D}(z)}{2-R_{\tilde{B},U\cap D}(z)}=1.

A crucial step in the proof of this theorem is to obtain Bergman–Fuks type results for the Kobayashi–Fuks metric and its related invariants, i.e., to express them in terms of certain maximal domain functions. For the holomorphic sectional curvature of the Kobayashi–Fuks metric, we derive such a result only in dimension one, though we believe that in higher dimensions also, an analog of this and hence of (III) above should hold.

Next we investigate the boundary behavior of the Kobayashi–Fuks metric on strongly pseudoconvex domains in n\mathbb{C}^{n}. We will denote by δD(z)\delta_{D}(z) the Euclidean distance from the point zDz\in D to the boundary D\partial D. For zz close to the boundary D\partial D, let π(z)D\pi(z)\in\partial D be the nearest point to zz, i.e., δD(z)=|zπ(z)|\delta_{D}(z)=|z-\pi(z)|, and for a tangent vector unu\in\mathbb{C}^{n} based at zz, let u=uH(z)+uN(z)u=u_{H}(z)+u_{N}(z) be the decomposition along the tangential and normal directions respectively at π(z)\pi(z).

Theorem 1.2

Let DnD\subset\mathbb{C}^{n} be a C2C^{2}-smoothly bounded strongly pseudoconvex domain and p0Dp^{0}\in\partial D. Then there are holomorphic coordinates zz near p0p^{0} in which

  • (I)

    δD(z)τB~,D(z,u)12(n+1)(n+2)|uN(p0)|\delta_{D}(z)\,\tau_{\tilde{B},D}(z,u)\to\frac{1}{2}\sqrt{(n+1)(n+2)}\,|u_{N}(p^{0})| ,

  • (II)

    δD(z)τB~,D(z,uH(z))12(n+1)(n+2)D(p0,uH(p0))\sqrt{\delta_{D}(z)}\,\tau_{\tilde{B},D}\big{(}z,u_{H}(z)\big{)}\to\sqrt{\frac{1}{2}(n+1)(n+2)\mathcal{L}_{\partial D}\big{(}p^{0},u_{H}(p^{0})\big{)}} ,

  • (III)

    δD(z)n+1gB~,D(z)(n+1)n(n+2)n2n+1\delta_{D}(z)^{n+1}g_{\tilde{B},D}(z)\to\frac{(n+1)^{n}(n+2)^{n}}{2^{n+1}},

  • (IV)

    If n=1n=1, then RB~,D(z)13R_{\tilde{B},D}(z)\to-\frac{1}{3},

as zp0z\to p^{0}. Here, D\mathcal{L}_{\partial D} is the Levi form of D\partial D with respect to some defining function for DD.

Note that Theorem 1.2 (I), (II) are analogs of Graham’s result [6] for the Kobayashi and Carathéodory metrics. Also, Theorem 1.2 (IV) combined with Theorem 1.17 of [7] immediately yields

Corollary 1.3

Let D,DD,D^{\prime}\subset\mathbb{C} be C2C^{2}-smoothly bounded domains equipped with the metrics dsB~,D2ds^{2}_{\tilde{B},D} and dsB~,D2ds^{2}_{\tilde{B},D^{\prime}} respectively. Then every isometry f:(D,dsB~,D2)(D,dsB~,D2)f:(D,ds^{2}_{\tilde{B},D})\to(D^{\prime},ds^{2}_{\tilde{B},D^{\prime}}) is either holomorphic or conjugate holomorphic.

Our final result is motivated by a theorem of Herbort [8, Theorem 1.2] on the existence of closed geodesics for the Bergman metric on strongly pseudoconvex domains.

Theorem 1.4

Let DnD\subset\mathbb{C}^{n} be a smoothly bounded strongly pseudoconvex domain which is not simply connected. Then every nontrivial homotopy class in π1(D)\pi_{1}(D) contains a closed geodesic for dsB~,D2ds^{2}_{\tilde{B},D}.

Acknowledgements: The authors thank Kaushal Verma for his support and encouragement.

2 Some examples

Proposition 2.5

For the unit ball 𝔹nn\mathbb{B}^{n}\subset\mathbb{C}^{n},

dsB~,𝔹n2=(n+2)dsB,𝔹n2=(n+1)(n+2)α,β=1n(δαβ¯1|z|2+z¯αzβ(1|z|2)2)dzαdz¯β.ds^{2}_{\tilde{B},\mathbb{B}^{n}}=(n+2)ds^{2}_{B,\mathbb{B}^{n}}=(n+1)(n+2)\sum_{\alpha,\beta=1}^{n}\left(\frac{\delta_{\alpha\overline{\beta}}}{1-|z|^{2}}+\frac{\overline{z}_{\alpha}z_{\beta}}{(1-|z|^{2})^{2}}\right)dz_{\alpha}d\overline{z}_{\beta}.
Proof 1

Recall that for the unit ball 𝔹nn\mathbb{B}^{n}\subset\mathbb{C}^{n},

K𝔹n(z)=n!πn1(1|z|2)n+1,K_{\mathbb{B}^{n}}(z)=\frac{n!}{\pi^{n}}\frac{1}{(1-|z|^{2})^{n+1}},

and so

gαβ¯B,𝔹n(z)=(n+1)2zαz¯βlog11|z|2=(n+1)(δαβ¯1|z|2+z¯αzβ(1|z|2)2).g_{\alpha\overline{\beta}}^{B,\mathbb{B}^{n}}(z)=(n+1)\frac{\partial^{2}}{\partial z_{\alpha}\partial\overline{z}_{\beta}}\log\frac{1}{1-|z|^{2}}=(n+1)\left(\frac{\delta_{\alpha\overline{\beta}}}{1-|z|^{2}}+\frac{\overline{z}_{\alpha}z_{\beta}}{(1-|z|^{2})^{2}}\right).

Denoting the matrix z¯zt\overline{z}z^{t} by AzA_{z} and using the fact that its characteristic polynomial is det(λ𝕀Az)=λn|z|2λn1\det(\lambda\mathbb{I}-A_{z})=\lambda^{n}-|z|^{2}\lambda^{n-1}, we obtain

gB,𝔹n(z)=(n+1)n(1|z|2)n+1,g_{B,\mathbb{B}^{n}}(z)=\frac{(n+1)^{n}}{(1-|z|^{2})^{n+1}},

and hence

Ricαβ¯B,𝔹n(z)=(n+1)2zαz¯βlog11|z|2=gαβ¯B,𝔹n(z).\operatorname{Ric}^{B,\mathbb{B}^{n}}_{\alpha\overline{\beta}}(z)=-(n+1)\frac{\partial^{2}}{\partial z_{\alpha}\partial\overline{z}_{\beta}}\log\frac{1}{1-|z|^{2}}=-g^{B,\mathbb{B}^{n}}_{\alpha\overline{\beta}}(z).

It follows that

gαβ¯B~,𝔹n(z)=(n+2)gαβ¯B,𝔹n(z),g^{\tilde{B},\mathbb{B}^{n}}_{\alpha\overline{\beta}}(z)=(n+2)g^{B,\mathbb{B}^{n}}_{\alpha\overline{\beta}}(z),

which completes the proof of the proposition.\qed

Proposition 2.6

For the unit polydisc Δnn\Delta^{n}\subset\mathbb{C}^{n},

dsB~,Δn2=(n+2)dsB,Δn2=2(n+2)α=1n1(1|zα|2)2dzαdz¯α.ds^{2}_{\tilde{B},\Delta^{n}}=(n+2)ds^{2}_{B,\Delta^{n}}=2(n+2)\sum_{\alpha=1}^{n}\dfrac{1}{(1-|z_{\alpha}|^{2})^{2}}dz_{\alpha}d\overline{z}_{\alpha}.
Proof 2

For the unit polydisc Δnn\Delta^{n}\subset\mathbb{C}^{n}, recall that by the product formula for the Bergman kernel,

KΔn(z)=j=1nKΔ(zj)=1πnj=1n1(1|zj|2)2,K_{\Delta^{n}}(z)=\prod_{j=1}^{n}K_{\Delta}(z_{j})=\frac{1}{\pi^{n}}\prod_{j=1}^{n}\frac{1}{(1-|z_{j}|^{2})^{2}},

and therefore

gαβ¯B,Δn(z)=22zαz¯βj=1nlog11|zj|2=2δαβ(1|zα|2)2.g_{\alpha\overline{\beta}}^{B,\Delta^{n}}(z)=2\,\frac{\partial^{2}}{\partial z_{\alpha}\partial\overline{z}_{\beta}}\sum_{j=1}^{n}\log\frac{1}{1-|z_{j}|^{2}}=\dfrac{2\delta_{\alpha\beta}}{(1-|z_{\alpha}|^{2})^{2}}.

Thus

gB,Δn(z)=2nj=1n1(1|zj|2)2,g_{B,\Delta^{n}}(z)=2^{n}\prod_{j=1}^{n}\dfrac{1}{(1-|z_{j}|^{2})^{2}},

and hence

Ricαβ¯B,Δn(z)=22zαz¯βj=1nlog11|zj|2=gαβ¯B,Δn(z).\operatorname{Ric}^{B,\Delta^{n}}_{\alpha\overline{\beta}}(z)=-2\,\frac{\partial^{2}}{\partial z_{\alpha}\partial\overline{z}_{\beta}}\sum_{j=1}^{n}\log\frac{1}{1-|z_{j}|^{2}}=-g^{B,\Delta^{n}}_{\alpha\overline{\beta}}(z).

It follows that

gαβ¯B~,Δn(z)=(n+2)gαβ¯B,Δn(z)=2(n+2)(1|zα|2)2δαβ,g^{\tilde{B},\Delta^{n}}_{\alpha\overline{\beta}}(z)=(n+2)g^{B,\Delta^{n}}_{\alpha\overline{\beta}}(z)=\dfrac{2(n+2)}{(1-|z_{\alpha}|^{2})^{2}}\delta_{\alpha\beta},

and the proof of the proposition is complete.\qed

In general, if DD is a bounded domain with a transitive group of holomorphic automorphisms, the Bergman metric is Kähler-Einstein, and so dsB~,D2ds^{2}_{\tilde{B},D} is a constant multiple of dsB,D2ds^{2}_{B,D}.

3 Localization

Our goal in this section is to prove Theorem 1.1. The localization of the Bergman kernel, the Bergman metric, and various invariants related to them near holomorphic peak points on pseudoconvex domains are well known, see for example [9, 10] for bounded domains, and [11, 12, 13] for unbounded domains. As in the case of the Bergman metric, the main idea is to express the Kobayashi–Fuks metric and the other invariants in terms of certain maximal domain functions. We will show that τB~,D\tau_{\tilde{B},D} can be expressed in terms of a maximal domain function introduced by Krantz and Yu in [10]. However, for the Gaussian curvature of dsB~,D2ds^{2}_{\tilde{B},D}, we will require some new maximal domain functions. We begin by recalling the definition of the domain function of Krantz and Yu: For a bounded domain DnD\subset\mathbb{C}^{n}, z0Dz_{0}\in D, and a nonzero vector unu\in\mathbb{C}^{n}, let

ID(z0,u)=sup{utf′′(z0)G¯B,D1(z0)f′′¯(z0)u¯:fD=1,f(z0)=f(z0)=0}.I_{D}(z_{0},u)=\sup\left\{u^{t}f^{\prime\prime}(z_{0})\overline{G}_{B,D}^{-1}(z_{0})\overline{f^{\prime\prime}}(z_{0})\overline{u}:\|f\|_{D}=1,f(z_{0})=f^{\prime}(z_{0})=0\right\}. (3.1)

Here f′′(z0)f^{\prime\prime}(z_{0}) is the symmetric matrix

f′′(z0)=(2fzizj(z0))n×n.f^{\prime\prime}(z_{0})=\begin{pmatrix}\frac{\partial^{2}f}{\partial z_{i}\partial z_{j}}(z_{0})\end{pmatrix}_{n\times n}.

It was shown in Proposition 2.1 (ii) of [10] that

RicB,D(z0,u)=(n+1)1τB,D2(z0,u)KD(z0)ID(z0,u).\operatorname{Ric}_{B,D}(z_{0},u)=(n+1)-\frac{1}{\tau^{2}_{B,D}(z_{0},u)K_{D}(z_{0})}I_{D}(z_{0},u). (3.2)

Also, from the definition of the Kobayashi–Fuks metric, note that

τB~,D(z0,u)=τB,D(z0,u)n+1RicB,D(z0,u).\tau_{\tilde{B},D}(z_{0},u)=\tau_{B,D}(z_{0},u)\sqrt{n+1-\operatorname{Ric}_{B,D}(z_{0},u)}. (3.3)

Combining (3.2) and (3.3) we obtain

Proposition 3.7

Let DD be a bounded domain in n\mathbb{C}^{n}, z0Dz_{0}\in D, and unu\in\mathbb{C}^{n}. Then we have

τB~,D2(z0,u)=ID(z0,u)KD(z0).\displaystyle\tau^{2}_{\tilde{B},D}(z_{0},u)=\frac{I_{D}(z_{0},u)}{K_{D}(z_{0})}.

From this proposition, we immediately obtain the localization of τB~,D\tau_{\tilde{B},D}.

Proof of Theorem 1.1 (i) 1

It was shown in [9] that

limzp0KD(z)KUD(z)=1\lim_{z\to p^{0}}\frac{K_{D}(z)}{K_{U\cap D}(z)}=1 (3.4)

and in [10] that

limzp0ID(z,u)IUD(z,u)=1\lim_{z\to p^{0}}\frac{I_{D}(z,u)}{I_{U\cap D}(z,u)}=1 (3.5)

uniformly in unit vectors uu, and hence (I) follows from Proposition 3.7. \square

For the localization of gB~,Dg_{\tilde{B},D}, we will need the following lemma. The notation 𝕀n\mathbb{I}_{n} or simply 𝕀\mathbb{I}, which we have already used in the previous section, stands for the n×nn\times n identity matrix.

Lemma 3.8

Let D1,D2D_{1},D_{2} be two bounded domains in n\mathbb{C}^{n} such that D2D1D_{2}\subset D_{1}. For any z0D2z_{0}\in D_{2}, there exist a nonsingular matrix QQ and positive real numbers d1,,dnd_{1},\ldots,d_{n} such that

QtGB~,D1(z0)Q¯=diag{d1,,dn}andQtGB~,D2(z0)Q¯=𝕀.Q^{t}G_{\tilde{B},D_{1}}(z_{0})\overline{Q}=\operatorname{diag}\{d_{1},\ldots,d_{n}\}\quad\text{and}\quad Q^{t}G_{\tilde{B},D_{2}}(z_{0})\overline{Q}=\mathbb{I}.
Proof 3

Note that as GB~,D2(z0)G_{\tilde{B},D_{2}}(z_{0}) is a positive definite Hermitian matrix one can find an invertible matrix AA such that

AtGB~,D2(z0)A¯=𝕀.A^{t}G_{\tilde{B},D_{2}}(z_{0})\overline{A}=\mathbb{I}.

By the transformation rule (1.3) applied to A:A1D1D1A:A^{-1}D_{1}\to D_{1} and A:A1D2D2A:A^{-1}D_{2}\to D_{2},

GB~,A1D1(A1z0)=AtGB~,D1(z0)A¯andGB~,A1D2(A1z0)=AtGB~,D2(z0)A¯=𝕀.G_{\tilde{B},A^{-1}D_{1}}(A^{-1}z_{0})=A^{t}G_{\tilde{B},D_{1}}(z_{0})\overline{A}\quad\text{and}\quad G_{\tilde{B},A^{-1}D_{2}}(A^{-1}z_{0})=A^{t}G_{\tilde{B},D_{2}}(z_{0})\overline{A}=\mathbb{I}.

From the first identity above, AtGB~,D1(z0)A¯A^{t}G_{\tilde{B},D_{1}}(z_{0})\overline{A} is a positive definite Hermitian matrix and hence there exists a unitary matrix BB such that

Bt(AtGB~,D1(z0)A¯)B¯=diag{d1,,dn}for somed1,,dn>0.B^{t}(A^{t}G_{\tilde{B},D_{1}}(z_{0})\overline{A})\overline{B}=\operatorname{diag}\{d_{1},\ldots,d_{n}\}\quad\text{for some}\quad d_{1},\ldots,d_{n}>0.

Now letting Q=ABQ=AB the lemma follows.\qed

Proof of Theorem 1.1 (II) 2

By Lemma 3.8, there exist an invertible matrix Q(z)Q(z) and positive real numbers d1(z),,dn(z)d_{1}(z),\ldots,d_{n}(z) such that

Qt(z)GB~,D(z)Q¯(z)=diag{d1(z),,dn(z)}andQt(z)GB~,UD(z)Q¯(z)=𝕀.Q^{t}(z)G_{\tilde{B},D}(z)\overline{Q}(z)=\operatorname{diag}\{d_{1}(z),\ldots,d_{n}(z)\}\quad\text{and}\quad Q^{t}(z)G_{\tilde{B},U\cap D}(z)\overline{Q}(z)=\mathbb{I}.

Taking determinant on both sides of these equations yields

gB~,D(z)gB~,UD(z)=j=1ndj(z).\frac{g_{\tilde{B},D}(z)}{g_{\tilde{B},U\cap D}(z)}=\prod\limits_{j=1}^{n}d_{j}(z).

Also, by Proposition 3.7,

τB~,D2(z,u)=KUD(z)KD(z)ID(z,u)IUD(z,u)τB~,UD2(z,u).\tau^{2}_{\tilde{B},D}(z,u)=\frac{K_{U\cap D}(z)}{K_{D}(z)}\frac{I_{D}(z,u)}{I_{U\cap D}(z,u)}\,\tau^{2}_{\tilde{B},U\cap D}(z,u).

Putting u=Q(z)eju=Q(z)e_{j} in the above equation, we get

dj(z)=KUD(z)KD(z)ID(z,Q(z)ej)IUD(z,Q(z)ej)forj=1,,n.d_{j}(z)=\frac{K_{U\cap D}(z)}{K_{D}(z)}\frac{I_{D}\big{(}z,Q(z)e_{j}\big{)}}{I_{U\cap D}\big{(}z,Q(z)e_{j}\big{)}}\quad\text{for}\quad j=1,\ldots,n.

Therefore,

gB~,D(z)gB~,UD(z)=(KUD(z)KD(z))nj=1nID(z,Q(z)ej)IUD(z,Q(z)ej)=(KUD(z)KD(z))nj=1nID(z,vj(z))IUD(z,vj(z))\frac{g_{\tilde{B},D}(z)}{g_{\tilde{B},U\cap D}(z)}=\left(\frac{K_{U\cap D}(z)}{K_{D}(z)}\right)^{n}\,\prod\limits_{j=1}^{n}\frac{I_{D}\big{(}z,Q(z)e_{j}\big{)}}{I_{U\cap D}\big{(}z,Q(z)e_{j}\big{)}}=\left(\frac{K_{U\cap D}(z)}{K_{D}(z)}\right)^{n}\,\prod\limits_{j=1}^{n}\frac{I_{D}\big{(}z,v_{j}(z)\big{)}}{I_{U\cap D}\big{(}z,v_{j}(z)\big{)}}

where vj(z)=Q(z)ej/Q(z)ejv_{j}(z)=Q(z)e_{j}/\|Q(z)e_{j}\|. Now (II) follows immediately from (3.4) and (3.5). \qed

In general, the Kobayashi–Fuks metric and its associated objects do not satisfy monotonicity property. Nevertheless, we show that they can be compared. Recall that the Bergman canonical invariant on DD is the function defined by

JD(z)=detGB,D(z)KD(z)=gB,D(z)KD(z).J_{D}(z)=\frac{\det G_{B,D}(z)}{K_{D}(z)}=\frac{g_{B,D}(z)}{K_{D}(z)}.

From the transformation rule for the Bergman kernel it is evident that JDJ_{D} is a biholomorphic invariant.

Proposition 3.9

Let D1,D2D_{1},D_{2} be two bounded domains in n\mathbb{C}^{n} such that D2D1D_{2}\subset D_{1}. For any z0D2z_{0}\in D_{2} and unu\in\mathbb{C}^{n}, we have

  • (a)

    τB~,D12(z0,u)(KD2(z0)KD1(z0))n+1(JD2(z0)JD1(z0))τB~,D22(z0,u),\tau^{2}_{\tilde{B},D_{1}}(z_{0},u)\leq\left(\frac{K_{D_{2}}(z_{0})}{K_{D_{1}}(z_{0})}\right)^{n+1}\left(\frac{J_{D_{2}}(z_{0})}{J_{D_{1}}(z_{0})}\right)\,\tau^{2}_{\tilde{B},D_{2}}(z_{0},u),

  • (b)

    utG¯B~,D11(z0)u¯(KD1(z0)KD2(z0))n+1(JD1(z0)JD2(z0))utG¯B~,D21(z0)u¯.u^{t}\overline{G}_{\tilde{B},D_{1}}^{-1}(z_{0})\overline{u}\geq\left(\frac{K_{D_{1}}(z_{0})}{K_{D_{2}}(z_{0})}\right)^{n+1}\left(\frac{J_{D_{1}}(z_{0})}{J_{D_{2}}(z_{0})}\right)u^{t}\overline{G}_{\tilde{B},D_{2}}^{-1}(z_{0})\overline{u}.

Proof 4

Fix z0D2z_{0}\in D_{2} and unu\in\mathbb{C}^{n}. For simplicity of notations, we will write KiK_{i} for KDi(z0)K_{D_{i}}(z_{0}), JiJ_{i} for JDi(z0)J_{D_{i}}(z_{0}), GiG_{i} for GB,Di(z0)G_{B,D_{i}}(z_{0}), and G~i\tilde{G}_{i} for GB~,Di(z0)G_{\tilde{B},D_{i}}(z_{0}) for i=1,2i=1,2.

(a) In view of Proposition 3.7, it is enough to prove that

ID1(z0,u)(K2K1)nJ2J1ID2(z0,u).I_{D_{1}}(z_{0},u)\leq\left(\frac{K_{2}}{K_{1}}\right)^{n}\frac{J_{2}}{J_{1}}\,I_{D_{2}}(z_{0},u). (3.6)

From the proof of Proposition 2.2 in [10] (see page 236) there exists a nonsingular matrix PP (depending on z0z_{0}) such that for every vnv\in\mathbb{C}^{n},

vtP1ad¯G1(P)1v¯(K2K1)n1vtP1ad¯G2(P)1v¯,v^{t}{P}^{-1}\,\overline{\operatorname{ad}}G_{1}(P^{*})^{-1}\overline{v}\leq\left(\frac{K_{2}}{K_{1}}\right)^{n-1}v^{t}P^{-1}\,\overline{\operatorname{ad}}G_{2}(P^{*})^{-1}\overline{v}, (3.7)

where by adA\operatorname{ad}A and AA^{*}, we mean the adjugate and the conjugate transpose of the matrix AA respectively. Now consider fA2(D1)f\in A^{2}(D_{1}) such that fD1=1,f(z0)=f(z0)=0\|f\|_{D_{1}}=1,f(z_{0})=f^{\prime}(z_{0})=0. We can write f′′(z0)=(Pt)1Af^{\prime\prime}(z_{0})=(P^{t})^{-1}A for some matrix AA. Putting v=Auv=Au in (3.7) and using the fact that f′′(z0)f^{\prime\prime}(z_{0}) is symmetric, we get

utf′′(z0)ad¯G1f′′¯(z0)u¯(K2K1)n1utf′′(z0)ad¯G2f′′¯(z0)u¯.u^{t}f^{\prime\prime}(z_{0})\,\overline{\operatorname{ad}}G_{1}\overline{f^{\prime\prime}}(z_{0})\overline{u}\leq\left(\frac{K_{2}}{K_{1}}\right)^{n-1}\,u^{t}f^{\prime\prime}(z_{0})\overline{\operatorname{ad}}G_{2}\overline{f^{\prime\prime}}(z_{0})\overline{u}. (3.8)

Define g:D2g:D_{2}\to\mathbb{C} by

g(z)=f(z)fD2.g(z)=\frac{f(z)}{\|f\|_{D_{2}}}.

Then gA2(D2)g\in A^{2}(D_{2}), gD2=1\|g\|_{D_{2}}=1, g(z0)=g(z0)=0g(z_{0})=g^{\prime}(z_{0})=0. Since f′′(z0)=fD2g′′(z0)f^{\prime\prime}(z_{0})=\|f\|_{D_{2}}g^{\prime\prime}(z_{0}), fD21\|f\|_{D_{2}}\leq 1, and adGi=(detGi)Gi1\text{ad}\,G_{i}=(\text{det}\,G_{i})G_{i}^{-1}, we have from (3.8)

utf′′(z0)G¯11f′′¯(z0)u¯(K2K1)n1detG2detG1(utg′′(z0)G¯21g′′¯(z0)u¯)(K2K1)nJ2J1ID2(z0,u).u^{t}f^{\prime\prime}(z_{0})\,\overline{G}_{1}^{-1}\overline{f^{\prime\prime}}(z_{0})\overline{u}\leq\left(\frac{K_{2}}{K_{1}}\right)^{n-1}\frac{\text{det}\,G_{2}}{\text{det}\,G_{1}}\big{(}u^{t}g^{\prime\prime}(z_{0})\,\overline{G}_{2}^{-1}\overline{g^{\prime\prime}}(z_{0})\overline{u}\big{)}\\ \leq\left(\frac{K_{2}}{K_{1}}\right)^{n}\frac{J_{2}}{J_{1}}I_{D_{2}}(z_{0},u). (3.9)

Taking supremum over ff in (3.9) and using Proposition 3.7, we obtain (3.6) and hence (a) is proved.

(b) Let QQ be as in Lemma 3.8. Let eje_{j} denote the j-th standard basis vector in n\mathbb{C}^{n}, i.e., ej=(0,,0,1,0,,0)e_{j}=(0,\ldots,0,1,0,\ldots,0). Taking u=Qeju=Qe_{j} in (a), we get

dj(K2K1)n+1J2J1forj=1,,n.d_{j}\leq\left(\frac{K_{2}}{K_{1}}\right)^{n+1}\frac{J_{2}}{J_{1}}\quad\text{for}\quad j=1,\ldots,n. (3.10)

From Lemma 3.8, it follows that

Q1(G1~¯)1(Q)1=diag{1d1,,1dn}andQ1(G2~¯)1(Q)1=𝕀.Q^{-1}\big{(}\overline{\tilde{G_{1}}}\big{)}^{-1}(Q^{*})^{-1}=\text{diag}\left\{\frac{1}{d_{1}},\ldots,\frac{1}{d_{n}}\right\}\quad\text{and}\quad Q^{-1}\big{(}\overline{\tilde{G_{2}}}\big{)}^{-1}(Q^{*})^{-1}=\mathbb{I}.

Hence for any vnv\in\mathbb{C}^{n}, using the inequality (3.10) we get

vtQ1(G1~¯)1(Q)1v¯=vtdiag{1d1,,1dn}v¯(K1K2)n+1J1J2(vt𝕀v¯)=(K1K2)n+1J1J2(vtQ1(G2~¯)1(Q)1v¯).v^{t}Q^{-1}\big{(}\overline{\tilde{G_{1}}}\big{)}^{-1}(Q^{*})^{-1}\overline{v}=v^{t}\operatorname{diag}\left\{\frac{1}{d_{1}},\ldots,\frac{1}{d_{n}}\right\}\overline{v}\geq\left(\frac{K_{1}}{K_{2}}\right)^{n+1}\frac{J_{1}}{J_{2}}(v^{t}\mathbb{I}\overline{v})\\ =\left(\frac{K_{1}}{K_{2}}\right)^{n+1}\frac{J_{1}}{J_{2}}\big{(}v^{t}Q^{-1}\big{(}\overline{\tilde{G_{2}}}\big{)}^{-1}(Q^{*})^{-1}\overline{v}\big{)}.

Putting u=(Qt)1vu=(Q^{t})^{-1}v in the above inequality, we get (b).\qed

We now introduce two maximal domain functions on planar domains for the purpose of localizing the Gaussian curvature of the Kobayashi–Fuks metric. For a bounded domain DD\subset\mathbb{C}, let

ID(z0)\displaystyle I^{\prime}_{D}(z_{0}) =sup{gB~,D1(z0)|f(z0)|2:fA2(D),fD=1,f(z0)=0},\displaystyle=\sup\{g^{-1}_{\tilde{B},D}(z_{0})|f^{\prime}(z_{0})|^{2}:f\in A^{2}(D),\|f\|_{D}=1,f(z_{0})=0\},
ID′′(z0)\displaystyle I^{\prime\prime}_{D}(z_{0}) =sup{gB~,D3(z0)|f′′′(z0)|2:fA2(D),fD=1,f(z0)=f(z0)=f′′(z0)=0}.\displaystyle=\sup\{g^{-3}_{\tilde{B},D}(z_{0})|f^{\prime\prime\prime}(z_{0})|^{2}:f\in A^{2}(D),\|f\|_{D}=1,f(z_{0})=f^{\prime}(z_{0})=f^{\prime\prime}(z_{0})=0\}.

Note that, as DD is bounded, the functions IDI^{\prime}_{D} and ID′′I^{\prime\prime}_{D} are well-defined and strictly positive. It is also evident that the supremums are achieved. Moreover, under biholomorphisms they transform by the same rule as that of the Bergman kernel which we establish in the following proposition:

Proposition 3.10

Let F:D1D2F:D_{1}\to D_{2} be a biholomorphism between two bounded domains in \mathbb{C}. Then

ID1(z0)=ID2(F(z0))|F(z0)|2andID1′′(z0)=ID2′′(F(z0))|F(z0)|2.I^{\prime}_{D_{1}}(z_{0})=I^{\prime}_{D_{2}}\big{(}F(z_{0})\big{)}\big{|}F^{\prime}(z_{0})\big{|}^{2}\quad\text{and}\quad I^{\prime\prime}_{D_{1}}(z_{0})=I^{\prime\prime}_{D_{2}}\big{(}F(z_{0})\big{)}\big{|}F^{\prime}(z_{0})\big{|}^{2}.
Proof 5

We will prove the transformation rule only for ID′′I^{\prime\prime}_{D}, as the case of IDI^{\prime}_{D} is even simpler and follows from similar arguments. Suppose gA2(D2)g\in A^{2}(D_{2}) is a maximizer for ID2′′(F(z0))I^{\prime\prime}_{D_{2}}\big{(}F(z_{0})\big{)}. Now set

f(z)=g(F(z))F(z).f(z)=g\big{(}F(z)\big{)}F^{\prime}(z).

It is straightforward to check that fD1=gD2=1\|f\|_{D_{1}}=\|g\|_{D_{2}}=1, f(z0)=f(z0)=f′′(z0)=0f(z_{0})=f^{\prime}(z_{0})=f^{\prime\prime}(z_{0})=0, and

f′′′(z0)=g′′′(F(z0))(F(z0))4.f^{\prime\prime\prime}(z_{0})=g^{\prime\prime\prime}\big{(}F(z_{0})\big{)}\big{(}F^{\prime}(z_{0})\big{)}^{4}.

Therefore,

gB~,D13(z0)|f′′′(z0)|2=gB~,D13(z0)|g′′′(F(z0))|2|F(z0)|8.g^{-3}_{\tilde{B},D_{1}}(z_{0})\big{|}f^{\prime\prime\prime}(z_{0})\big{|}^{2}=g^{-3}_{\tilde{B},D_{1}}(z_{0})\big{|}g^{\prime\prime\prime}\big{(}F(z_{0})\big{)}\big{|}^{2}\big{|}F^{\prime}(z_{0})\big{|}^{8}. (3.11)

Note that from the transformation rule for the Kobayashi–Fuks metric, we have

gB~,D11(z0)|F(z0)|2=gB~,D21(F(z0)).g^{-1}_{\tilde{B},D_{1}}(z_{0})|F^{\prime}(z_{0})|^{2}=g^{-1}_{\tilde{B},D_{2}}\big{(}F(z_{0})\big{)}.

Applying this on the right hand side of (3.11), we get

gB~,D13(z0)|f′′′(z0)|2=gB~,D23(F(z0))|g′′′(F(z0))|2|F(z0)|2.g^{-3}_{\tilde{B},D_{1}}(z_{0})|f^{\prime\prime\prime}(z_{0})|^{2}=g^{-3}_{\tilde{B},D_{2}}\big{(}F(z_{0})\big{)}\big{|}g^{\prime\prime\prime}\big{(}F(z_{0})\big{)}\big{|}^{2}\big{|}F^{\prime}(z_{0})\big{|}^{2}.

As ff is a candidate for ID1′′(z0)I^{\prime\prime}_{D_{1}}(z_{0}) and gg is a maximizer for ID2′′(F(z0))I^{\prime\prime}_{D_{2}}(F(z_{0})), we obtain

ID1′′(z0)ID2′′(F(z0))|F(z0)|2.I^{\prime\prime}_{D_{1}}(z_{0})\geq I^{\prime\prime}_{D_{2}}\big{(}F(z_{0})\big{)}\,|F^{\prime}(z_{0})|^{2}.

Similar arguments when applied to the map F1:D2D1F^{-1}:D_{2}\to D_{1} gives the reverse inequality and hence it is an equality.\qed

The main ingredient for the localization of the Gaussian curvature of the Kobayashi–Fuks metric is the following Bergman–Fuks type result:

Proposition 3.11

Let DD\subset\mathbb{C} be a bounded domain and z0Dz_{0}\in D. Then the Gaussian curvature of the Kobayashi–Fuks metric on DD satisfies

RB~,D(z0)=2ID(z0)KD(z0)ID′′(z0)ID(z0).R_{\tilde{B},D}(z_{0})=2-\dfrac{I^{\prime}_{D}(z_{0})}{K_{D}(z_{0})}-\dfrac{I^{\prime\prime}_{D}(z_{0})}{I^{\prime}_{D}(z_{0})}\,. (3.12)

Observe that both the sides of (3.12) are invariant under biholomorphisms and we will establish their equality by computing them in terms of a suitable orthonormal basis of A2(D)A^{2}(D) in some special coordinates. To this end, we fix z0Dz_{0}\in D and consider the closed subspaces of A2(D)A^{2}(D) given by

A1(z0)\displaystyle A_{1}(z_{0}) ={fA2(D):f(z0)=0},\displaystyle=\big{\{}f\in A^{2}(D):f(z_{0})=0\big{\}},
A2(z0)\displaystyle A_{2}(z_{0}) ={fA2(D):f(z0)=f(z0)=0},\displaystyle=\big{\{}f\in A^{2}(D):f(z_{0})=f^{\prime}(z_{0})=0\big{\}},
A3(z0)\displaystyle A_{3}(z_{0}) ={fA2(D):f(z0)=f(z0)=f′′(z0)=0}.\displaystyle=\big{\{}f\in A^{2}(D):f(z_{0})=f^{\prime}(z_{0})=f^{\prime\prime}(z_{0})=0\big{\}}.

Observe that the orthogonal complement of A1(z0)A_{1}(z_{0}) in A2(D)A^{2}(D) has dimension one and let h0h_{0} be a unit vector in this orthogonal complement. It is easy to see that the orthogonal complement of A2(z0)A_{2}(z_{0}) in A1(z0)A_{1}(z_{0}) has dimension at most one. Since DD is bounded, this space contains the function zz0z-z_{0}, and hence its dimension is exactly one. Similarly, the orthogonal complement of A3(z0)A_{3}(z_{0}) in A2(z0)A_{2}(z_{0}) has dimension one. Let {ϕ,ψ,h1,h2,}\{\phi,\psi,h_{1},h_{2},\ldots\} be an orthonormal basis for A1(z0)A_{1}(z_{0}) such that ϕ\phi is a unit vector in A1(z0)A2(z0)A_{1}(z_{0})\setminus A_{2}(z_{0}), ψ\psi is a unit vector in A2(z0)A3(z0)A_{2}(z_{0})\setminus A_{3}(z_{0}), and {h1,,hj,}\{h_{1},\ldots,h_{j},\ldots\} is an orthonormal basis for A3(z0)A_{3}(z_{0}). Note that

KD(z)=|h0(z)|2+|ϕ(z)|2+|ψ(z)|2+j=1|hj(z)|2,zD.K_{D}(z)=\big{|}h_{0}(z)\big{|}^{2}+\big{|}\phi(z)\big{|}^{2}+\big{|}\psi(z)\big{|}^{2}+\sum\limits_{j=1}^{\infty}\big{|}h_{j}(z)\big{|}^{2},\quad z\in D. (3.13)

Hence KD(z0)=|h0(z0)|2K_{D}(z_{0})=\big{|}h_{0}(z_{0})\big{|}^{2}, which in particular implies h0(z0)0h_{0}(z_{0})\neq 0.

Lemma 3.12

In normal coordinates for the Kobayashi–Fuks metric at z0z_{0},

  • (a)

    ID(z0)=|ϕ(z0)|2I^{\prime}_{D}(z_{0})=\left|\phi^{\prime}(z_{0})\right|^{2}, and

  • (b)

    ID′′(z0)=j=1|hj′′′(z0)|2I^{\prime\prime}_{D}(z_{0})=\sum_{j=1}^{\infty}\left|h_{j}^{\prime\prime\prime}(z_{0})\right|^{2}.

Proof 6

(a) Observe that in normal coordinates at z0z_{0}, IDI^{\prime}_{D} is reduced to

ID(z0)=sup{|f(z0)|2:fA1(z0),fD=1}.I^{\prime}_{D}(z_{0})=\sup\left\{|f^{\prime}(z_{0})|^{2}:f\in A_{1}(z_{0}),\|f\|_{D}=1\right\}.

Since ϕ\phi is a candidate for ID(z0)I^{\prime}_{D}(z_{0}),

ID(z0)|ϕ(z0)|2.I^{\prime}_{D}(z_{0})\geq\left|\phi^{\prime}(z_{0})\right|^{2}.

To see the reverse inequality, consider any fA1(z0)f\in A_{1}(z_{0}) with fD=1\|f\|_{D}=1. Since ff can be represented as

f(z)=f,ϕϕ(z)+f,ψψ(z)+j=1f,hjhj(z),f(z)=\langle f,\phi\rangle\,\phi(z)+\langle f,\psi\rangle\,\psi(z)+\sum_{j=1}^{\infty}\langle f,h_{j}\rangle\,h_{j}(z),

using fD=1\|f\|_{D}=1, we have

|f(z0)|2=|f,ϕ|2|ϕ(z0)|2|ϕ(z0)|2.|f^{\prime}(z_{0})|^{2}=\big{|}\langle f,\phi\rangle\big{|}^{2}\big{|}\phi^{\prime}(z_{0})\big{|}^{2}\leq\big{|}\phi^{\prime}(z_{0})\big{|}^{2}.

Now taking supremum in the left hand side of the above inequality, we get

ID(z0)|ϕ(z0)|2.I^{\prime}_{D}(z_{0})\leq\big{|}\phi^{\prime}(z_{0})\big{|}^{2}.

(b) Clearly the right hand side is finite thanks to Cauchy estimates. Observe that in normal coordinates at z0z_{0},

ID′′(z0)=sup{|f′′′(z0)|2:fA3(z0),fD=1}.I^{\prime\prime}_{D}(z_{0})=\sup\left\{|f^{\prime\prime\prime}(z_{0})|^{2}:f\in A_{3}(z_{0}),\|f\|_{D}=1\right\}.

Note that if f(z)=j=1ajhj(z)f(z)=\sum_{j=1}^{\infty}a_{j}h_{j}(z) is an arbitrary member of A3(z0)A_{3}(z_{0}), then

|f′′′(z0)|2=|j=1ajhj′′′(z0)|2=|j=1ajHj|2,\big{|}f^{\prime\prime\prime}(z_{0})\big{|}^{2}=\left|\sum\limits_{j=1}^{\infty}a_{j}h_{j}^{\prime\prime\prime}(z_{0})\right|^{2}=\left|\sum\limits_{j=1}^{\infty}a_{j}H_{j}\right|^{2},

where Hj=hj′′′(z0)H_{j}=h_{j}^{\prime\prime\prime}(z_{0}). Moreover, as A3(z0)A_{3}(z_{0}) is linearly isometric to 2\ell_{2} via the orthonormal basis {hj}\{h_{j}\}, we also have fD=a2\|f\|_{D}=\|a\|_{\ell_{2}}, where a={aj}j1a=\{a_{j}\}_{j\geq 1}. Hence, we arrive at

ID′′(z0)=sup{|j=1ajHj|2:a={aj}j12,a=1}.I^{\prime\prime}_{D}(z_{0})=\sup\left\{\bigg{|}\sum\limits_{j=1}^{\infty}a_{j}H_{j}\bigg{|}^{2}:a=\{a_{j}\}_{j\geq 1}\in\ell_{2},\|a\|=1\right\}. (3.14)

Now let H={Hj}j=1H=\{H_{j}\}_{j=1}^{\infty} and define LH:2L_{H}:\ell_{2}\to\mathbb{C} by

LH(a)=j=1ajHj.L_{H}(a)=\sum_{j=1}^{\infty}a_{j}H_{j}.

Then LHL_{H} is a bounded linear operator on 2\ell_{2} and denoting its operator norm by LH\|L_{H}\|, observe that

LH2=supa=1|LH({aj})|2=supa=1|j=1ajHj|2=ID′′(z0)\|L_{H}\|^{2}=\sup\limits_{\|a\|=1}\Big{|}L_{H}\big{(}\{a_{j}\}\big{)}\Big{|}^{2}=\sup\limits_{\|a\|=1}\bigg{|}\sum\limits_{j=1}^{\infty}a_{j}H_{j}\bigg{|}^{2}=I^{\prime\prime}_{D}(z_{0}) (3.15)

by (3.14). Also, from the canonical isometry of 2\ell_{2}^{\prime} with 2\ell_{2}, we have

LH2=H22=j=1|Hj|2=j=1|hj′′′(z0)|2.\|L_{H}\|^{2}=\|H\|_{\ell_{2}}^{2}=\sum\limits_{j=1}^{\infty}|H_{j}|^{2}=\sum\limits_{j=1}^{\infty}\left|h_{j}^{\prime\prime\prime}(z_{0})\right|^{2}. (3.16)

From (3.15) and (3.16), the lemma follows immediately.\qed

Proof of Proposition 3.11 3

We work in normal coordinates for the Kobayashi–Fuks metric at z0z_{0}. Without loss of generality, we will denote the new coordinates by zz, same as the previous ones. Note that in normal coordinates at z0z_{0}, we have

gB~,D(z0)=1,gB~,Dz(z0)=0,andRB~,D(z0)=2gB~,Dzz¯(z0).g_{\tilde{B},D}(z_{0})=1,\quad\frac{\partial g_{\tilde{B},D}}{\partial z}(z_{0})=0,\quad\text{and}\quad R_{\tilde{B},D}(z_{0})=-\frac{\partial^{2}g_{\tilde{B},D}}{\partial z\partial\overline{z}}(z_{0}). (3.17)

Next, we express the above equations in terms of the basis expansion of KDK_{D}. Recall that the Kähler potential of the Kobayashi–Fuks metric in dimension 11 is logA(z)\log A(z) where

A=KD2gB,D=KD22logKDzz¯=KD2KDzz¯KDzKDz¯.A=K_{D}^{2}g_{B,D}=K_{D}^{2}\frac{\partial^{2}\log K_{D}}{\partial z\partial\overline{z}}=K_{D}\tfrac{\partial^{2}K_{D}}{\partial z\partial\overline{z}}-\tfrac{\partial K_{D}}{\partial z}\tfrac{\partial K_{D}}{\partial\overline{z}}. (3.18)

Thus

gB~,D=2logAzz¯=1A2Azz¯1A2AzAz¯.\displaystyle g_{\tilde{B},D}=\frac{\partial^{2}\log A}{\partial z\partial\overline{z}}=\dfrac{1}{A}\dfrac{\partial^{2}A}{\partial z\partial\overline{z}}-\dfrac{1}{A^{2}}\dfrac{\partial A}{\partial z}\dfrac{\partial A}{\partial\overline{z}}. (3.19)

Using the expansion of the Bergman kernel as in (3.13), we get from (3.18)

A(z0)\displaystyle A(z_{0}) =|h0(z0)|2|ϕ(z0)|2,\displaystyle=\big{|}h_{0}(z_{0})\big{|}^{2}\,\left|\phi^{\prime}(z_{0})\right|^{2}, (3.20)
Az(z0)\displaystyle\frac{\partial A}{\partial z}(z_{0}) =|h0(z0)|2ϕ(z0)¯ϕ′′(z0),\displaystyle=\big{|}h_{0}(z_{0})\big{|}^{2}\,\overline{\phi^{\prime}(z_{0})}\,\phi^{\prime\prime}(z_{0}),
2Azz¯(z0)\displaystyle\dfrac{\partial^{2}A}{\partial z\partial\overline{z}}(z_{0}) =|h0(z0)|2(|ϕ′′(z0)|2+|ψ′′(z0)|2),\displaystyle=\big{|}h_{0}(z_{0})\big{|}^{2}\left(\left|\phi^{\prime\prime}(z_{0})\right|^{2}+\left|\psi^{\prime\prime}(z_{0})\right|^{2}\right),
2Az2(z0)\displaystyle\dfrac{\partial^{2}A}{\partial z^{2}}(z_{0}) =|h0(z0)|2ϕ¯(z0)ϕ′′′(z0)+h0(z0)¯h0(z0)ϕ¯(z0)ϕ′′(z0)\displaystyle=\big{|}h_{0}(z_{0})\big{|}^{2}\overline{\phi^{\prime}}(z_{0})\phi^{\prime\prime\prime}(z_{0})+\overline{h_{0}(z_{0})}\,h_{0}^{\prime}(z_{0})\overline{\phi^{\prime}}(z_{0})\phi^{\prime\prime}(z_{0})
h0(z0)¯h0′′(z0)|ϕ(z0)|2,and\displaystyle\hskip 113.81102pt-\overline{h_{0}(z_{0})}\,h_{0}^{\prime\prime}(z_{0})\left|\phi^{\prime}(z_{0})\right|^{2},\quad\text{and}
3Az2z¯(z0)\displaystyle\dfrac{\partial^{3}A}{\partial z^{2}\partial\overline{z}}(z_{0}) =|h0(z0)|2(ϕ′′¯(z0)ϕ′′′(z0)+ψ′′¯(z0)ψ′′′(z0))h0(z0)¯h0′′(z0)ϕ(z0)ϕ′′¯(z0)\displaystyle=\big{|}h_{0}(z_{0})\big{|}^{2}\bigg{(}\overline{\phi^{\prime\prime}}(z_{0})\phi^{\prime\prime\prime}(z_{0})+\overline{\psi^{\prime\prime}}(z_{0})\psi^{\prime\prime\prime}(z_{0})\bigg{)}-\overline{h_{0}(z_{0})}\,h_{0}^{\prime\prime}(z_{0})\phi^{\prime}(z_{0})\overline{\phi^{\prime\prime}}(z_{0})
+h0(z0)¯h0(z0)(|ϕ′′(z0)|2+|ψ′′(z0)|2).\displaystyle\hskip 85.35826pt+\overline{h_{0}(z_{0})}\,h_{0}^{\prime}(z_{0})\left(\left|\phi^{\prime\prime}(z_{0})\right|^{2}+\left|\psi^{\prime\prime}(z_{0})\right|^{2}\right).

Also, differentiating (3.19) one immediately obtains

gB~,Dz=1A3Az2z¯2A2Az2Azz¯1A2Az¯2Az2+2A3Az¯(Az)2,\dfrac{\partial g_{\tilde{B},D}}{\partial z}=\dfrac{1}{A}\dfrac{\partial^{3}A}{\partial z^{2}\partial\overline{z}}-\dfrac{2}{A^{2}}\dfrac{\partial A}{\partial z}\dfrac{\partial^{2}A}{\partial z\partial\overline{z}}-\dfrac{1}{A^{2}}\dfrac{\partial A}{\partial\overline{z}}\dfrac{\partial^{2}A}{\partial z^{2}}+\dfrac{2}{A^{3}}\dfrac{\partial A}{\partial\overline{z}}\bigg{(}\dfrac{\partial A}{\partial z}\bigg{)}^{2}, (3.21)

and

2gB~,Dzz¯(z)=1A4Az2z¯2+2A2Az¯3Az2z¯+2A2Az3Azz¯2+2A2(2Azz¯)2+1A22Az22Az¯2+6A4(Az)2(Az¯)22A3(Az¯)22Az22A3(Az)22Az¯28A3AzAz¯2Azz¯.-\dfrac{\partial^{2}g_{\tilde{B},D}}{\partial z\partial\overline{z}}(z)=-\dfrac{1}{A}\dfrac{\partial^{4}A}{\partial z^{2}\partial\overline{z}^{2}}+\dfrac{2}{A^{2}}\dfrac{\partial A}{\partial\overline{z}}\dfrac{\partial^{3}A}{\partial z^{2}\partial\overline{z}}+\dfrac{2}{A^{2}}\dfrac{\partial A}{\partial z}\dfrac{\partial^{3}A}{\partial z\partial\overline{z}^{2}}\\ +\dfrac{2}{A^{2}}\left(\dfrac{\partial^{2}A}{\partial z\partial\overline{z}}\right)^{2}+\dfrac{1}{A^{2}}\dfrac{\partial^{2}A}{\partial z^{2}}\dfrac{\partial^{2}A}{\partial\overline{z}^{2}}+\dfrac{6}{A^{4}}\left(\dfrac{\partial A}{\partial z}\right)^{2}\left(\dfrac{\partial A}{\partial\overline{z}}\right)^{2}\\ -\dfrac{2}{A^{3}}\left(\dfrac{\partial A}{\partial\overline{z}}\right)^{2}\dfrac{\partial^{2}A}{\partial z^{2}}-\dfrac{2}{A^{3}}\left(\dfrac{\partial A}{\partial z}\right)^{2}\dfrac{\partial^{2}A}{\partial\overline{z}^{2}}-\dfrac{8}{A^{3}}\dfrac{\partial A}{\partial z}\dfrac{\partial A}{\partial\overline{z}}\dfrac{\partial^{2}A}{\partial z\partial\overline{z}}. (3.22)

Now the relation gB~,D(z0)=1g_{\tilde{B},D}(z_{0})=1 using (3.19) and (3.20) gives

|h0(z0)|4|ϕ(z0)|2|ψ′′(z0)|2=|h0(z0)|4|ϕ(z0)|4.\big{|}h_{0}(z_{0})\big{|}^{4}\left|\phi^{\prime}(z_{0})\right|^{2}\left|\psi^{\prime\prime}(z_{0})\right|^{2}=\big{|}h_{0}(z_{0})\big{|}^{4}\left|\phi^{\prime}(z_{0})\right|^{4}.

Observe that h0(z0)0h_{0}(z_{0})\neq 0 and ϕ(z0)0\phi^{\prime}(z_{0})\neq 0. Hence the above identity implies

|ψ′′(z0)|2=|ϕ(z0)|2.\left|\psi^{\prime\prime}(z_{0})\right|^{2}=\left|\phi^{\prime}(z_{0})\right|^{2}.

As a consequence, applying a unitary transformation to the basis {ψ}\{\psi\}, we may assume that

ψ′′(z0)=ϕ(z0).\psi^{\prime\prime}(z_{0})=\phi^{\prime}(z_{0}). (3.23)

The relation gB~,Dz(z0)=0\tfrac{\partial g_{\tilde{B},D}}{\partial z}(z_{0})=0 using (3.21) and (3.20) gives

|h0(z0)|2|ϕ(z0)|2ψ′′¯(z0)ψ′′′(z0)+h0(z0)¯h0(z0)|ϕ(z0)|2|ψ′′(z0)|22|h0(z0)|2ϕ¯(z0)ϕ′′(z0)|ψ′′(z0)|2=0.\big{|}h_{0}(z_{0})\big{|}^{2}\,\left|\phi^{\prime}(z_{0})\right|^{2}\,\overline{\psi^{\prime\prime}}(z_{0})\,\psi^{\prime\prime\prime}(z_{0})+\overline{h_{0}(z_{0})}\,h_{0}^{\prime}(z_{0})\,\left|\phi^{\prime}(z_{0})\right|^{2}\,\left|\psi^{\prime\prime}(z_{0})\right|^{2}\\ -2\big{|}h_{0}(z_{0})\big{|}^{2}\,\overline{\phi^{\prime}}(z_{0})\,\phi^{\prime\prime}(z_{0})\,\left|\psi^{\prime\prime}(z_{0})\right|^{2}=0.

Using (3.23) in the above sum, we obtain the relation

h0(z0)ψ′′′(z0)+h0(z0)ϕ(z0)2h0(z0)ϕ′′(z0)=0.h_{0}(z_{0})\,\psi^{\prime\prime\prime}(z_{0})+h_{0}^{\prime}(z_{0})\,\phi^{\prime}(z_{0})-2h_{0}(z_{0})\,\phi^{\prime\prime}(z_{0})=0. (3.24)

Finally, we compute the curvature. It follows from (3.22), by a straightforward but lengthy calculation, that

A4\displaystyle A^{4} 2gB~,Dzz¯=|h0|8|ϕ|6(|ψ′′′|2+j=1|hj′′′|2)2|h0|6|ϕ|6Re(h0h0ψ′′¯ψ′′′)\displaystyle\dfrac{-\partial^{2}g_{\tilde{B},D}}{\partial z\partial\overline{z}}=-|h_{0}|^{8}|\phi^{\prime}|^{6}\Big{(}|\psi^{\prime\prime\prime}|^{2}+\sum_{j=1}^{\infty}|h_{j}^{\prime\prime\prime}|^{2}\Big{)}-2|h_{0}|^{6}\,|\phi^{\prime}|^{6}\,\text{Re}\left(h_{0}\overline{h_{0}^{\prime}\psi^{\prime\prime}}\psi^{\prime\prime\prime}\right)
|h0|6|h0|2|ϕ|6|ψ′′|2|h0|6|ϕ|8|ψ′′|2+4|h0|8|ϕ|4Re(ϕϕ′′¯ψ′′¯ψ′′′)\displaystyle-|h_{0}|^{6}\,|h_{0}^{\prime}|^{2}\,|\phi^{\prime}|^{6}\,|\psi^{\prime\prime}|^{2}-|h_{0}|^{6}|\phi^{\prime}|^{8}|\psi^{\prime\prime}|^{2}+4|h_{0}|^{8}|\phi^{\prime}|^{4}\operatorname{Re}\left(\phi^{\prime}\overline{\phi^{\prime\prime}}\overline{\psi^{\prime\prime}}\psi^{\prime\prime\prime}\right)
+4|h0|6|ϕ|4|ψ′′|2Re(h0h0ϕ¯ϕ′′)+2|h0|8|ϕ|4|ψ′′|44|h0|8|ϕ|4|ϕ′′|2|ψ′′|2\displaystyle+4|h_{0}|^{6}|\phi^{\prime}|^{4}|\psi^{\prime\prime}|^{2}\operatorname{Re}\left(h_{0}\overline{h_{0}^{\prime}\phi^{\prime}}\phi^{\prime\prime}\right)+2|h_{0}|^{8}|\phi^{\prime}|^{4}|\psi^{\prime\prime}|^{4}-4|h_{0}|^{8}|\phi^{\prime}|^{4}|\phi^{\prime\prime}|^{2}|\psi^{\prime\prime}|^{2}

at the point z0z_{0}. In the above equation and in the subsequent steps, if not mentioned, all the terms and partial derivatives are evaluated at the point z0z_{0}. Now making use of the relation (3.23), the above equation can be rewritten as

|h0|8|ϕ|82gB~,Dzz¯(z0)=2|h0|8|ϕ|8|h0|6|ϕ|10|h0|8|ϕ|6j=1|hj′′′|2|h0|6|ϕ|6{|h0|2|ψ′′′|2+|h0|2|ϕ|2+4|h0|2|ϕ′′|2+2Re(h0h0ϕ¯ψ′′′)4|h0|2Re(ϕ′′¯ψ′′′)4Re(h0h0ϕ¯ϕ′′)}.|h_{0}|^{8}|\phi^{\prime}|^{8}\dfrac{-\partial^{2}g_{\tilde{B},D}}{\partial z\partial\overline{z}}(z_{0})=2|h_{0}|^{8}|\phi^{\prime}|^{8}-|h_{0}|^{6}|\phi^{\prime}|^{10}-|h_{0}|^{8}|\phi^{\prime}|^{6}\sum_{j=1}^{\infty}|h_{j}^{\prime\prime\prime}|^{2}-|h_{0}|^{6}|\phi^{\prime}|^{6}\Big{\{}|h_{0}|^{2}|\psi^{\prime\prime\prime}|^{2}\\ +|h_{0}^{\prime}|^{2}|\phi^{\prime}|^{2}+4|h_{0}|^{2}|\phi^{\prime\prime}|^{2}+2\operatorname{Re}\big{(}h_{0}\overline{h_{0}^{\prime}\phi^{\prime}}\psi^{\prime\prime\prime}\big{)}-4|h_{0}|^{2}\operatorname{Re}\big{(}\overline{\phi^{\prime\prime}}\psi^{\prime\prime\prime}\big{)}-4\operatorname{Re}\big{(}h_{0}\overline{h_{0}^{\prime}\phi^{\prime}}\phi^{\prime\prime}\big{)}\Big{\}}.

Here one can see that, the terms inside the curly bracket above is exactly equal to

|h0ψ′′′+h0ϕ2h0ϕ′′|2,|h_{0}\psi^{\prime\prime\prime}+h_{0}^{\prime}\phi^{\prime}-2h_{0}\phi^{\prime\prime}|^{2},

which vanishes by vitue of the relation (3.24). Hence we finally arrive at

RB~,D(z0)=2gB~,Dzz¯(z0)=2|ϕ(z0)|2|h0(z0)|21|ϕ(z0)|2j=1|hj′′′(z0)|2.R_{\tilde{B},D}(z_{0})=-\dfrac{\partial^{2}g_{\tilde{B},D}}{\partial z\partial\overline{z}}(z_{0})=2-\dfrac{\big{|}\phi^{\prime}(z_{0})\big{|}^{2}}{\big{|}h_{0}(z_{0})\big{|}^{2}}-\dfrac{1}{\big{|}\phi^{\prime}(z_{0})\big{|}^{2}}\sum_{j=1}^{\infty}\big{|}h_{j}^{\prime\prime\prime}(z_{0})\big{|}^{2}. (3.25)

The right hand side of the above identity is finite thanks to the Cauchy estimates. The proposition now follows from Lemma 3.12. \qed

The following result is an immediate consequence of Proposition 3.11.

Corollary 3.13

The Gaussian curvature of the Kobayashi–Fuks metric on bounded planar domains is strictly bounded above by 22.

Now we localize the domain functions IDI^{\prime}_{D} and ID′′I^{\prime\prime}_{D}.

Proposition 3.14

Let DD\subset\mathbb{C} be a bounded domain, p0Dp^{0}\in\partial D a local peak point, and UU a sufficiently small neighborhood of p0p^{0}. Then

limzp0ID(z)IUD(z)=limzp0ID′′(z)IUD′′(z)=1.\lim\limits_{z\to p^{0}}\frac{I^{\prime}_{D}(z)}{I^{\prime}_{U\cap D}(z)}=\lim\limits_{z\to p^{0}}\frac{I^{\prime\prime}_{D}(z)}{I^{\prime\prime}_{U\cap D}(z)}=1.
Proof 7

We will present the proof only for ID′′I^{\prime\prime}_{D} here. The proof for IDI^{\prime}_{D} follows in an exactly similar manner. Let hh be a local holomorphic peak function for p0p^{0} defined on a neighborhood UU of p0p^{0}. Shrinking UU if necessary, we can assume that hh is nonvanishing on UD¯U\cap\overline{D}. Now choose any neighborhood VV of p0p^{0} such that VUV\subset\subset U. Then there is a constant b(0,1)b\in(0,1) such that |h|b|h|\leq b on (UV)D¯\overline{(U\setminus V)\cap D}. Let us choose a cut-off function χC0(U)\chi\in C_{0}^{\infty}(U) satisfying χ=1\chi=1 on VV and 0χ10\leq\chi\leq 1 on UU. Given any ζV\zeta\in V, a function fA2(UD)f\in A^{2}(U\cap D), and an integer k1k\geq 1, set

ϕ(z)=8log|zζ|andαk=¯(χfhk).\phi(z)=8\log|z-\zeta|\quad\text{and}\quad\alpha_{k}=\bar{\partial}(\chi fh^{k}).

Then ϕ\phi is a subharmonic function on \mathbb{C} and αk\alpha_{k} is a ¯\overline{\partial}-closed, smooth (0,1)(0,1)-form on DD with suppαk(UV)D\operatorname{supp}\alpha_{k}\subset(U\setminus V)\cap D. Now as in [10], applying Theorem 4.2 of [14], we get a solution uu to the equation ¯u=αk\bar{\partial}u=\alpha_{k} on DD such that

|A|+|B|uzAz¯B(ζ)=0 for all multi-indices A,B with |A|+|B|3,\frac{\partial^{|A|+|B|}u}{\partial z^{A}\partial\overline{z}^{B}}(\zeta)=0\text{ for all multi-indices }A,B\text{ with }|A|+|B|\leq 3, (3.26)

and

uDcbkfUD,\|u\|_{D}\leq cb^{k}\|f\|_{U\cap D}, (3.27)

where cc is a constant independent of ζ\zeta and kk.

Now let fA2(UD)f\in A^{2}(U\cap D) be a maximizing function for IUD′′(ζ)I^{\prime\prime}_{U\cap D}(\zeta), i.e., fUD=1\|f\|_{U\cap D}=1, f(ζ)=f(ζ)=f′′(ζ)=0f(\zeta)=f^{\prime}(\zeta)=f^{\prime\prime}(\zeta)=0, and gB~,UD3(ζ)|f′′′(ζ)|2=IUD′′(ζ)g^{-3}_{\tilde{B},U\cap D}(\zeta)|f^{\prime\prime\prime}(\zeta)|^{2}=I^{\prime\prime}_{U\cap D}(\zeta). Choose uu as above and set Fk=χfhkuF_{k}=\chi fh^{k}-u. Then FkA2(D)F_{k}\in A^{2}(D) and it follows from (3.27) that

FkDχfhkD+uDfUD+cbkfUD=(1+cbk).\|F_{k}\|_{D}\leq\|\chi fh^{k}\|_{D}+\|u\|_{D}\leq\|f\|_{U\cap D}+cb^{k}\|f\|_{U\cap D}=(1+cb^{k}). (3.28)

Therefore, setting fk=Fk/FkDf_{k}=F_{k}/\|F_{k}\|_{D}, we see that fkA2(D)f_{k}\in A^{2}(D), fkD=1\|f_{k}\|_{D}=1, and fk(ζ)=fk(ζ)=fk′′(ζ)=0f_{k}(\zeta)=f_{k}^{\prime}(\zeta)=f_{k}^{\prime\prime}(\zeta)=0. Moreover, by the maximality of ID′′(ζ)I^{\prime\prime}_{D}(\zeta), estimate (3.28), and (b) of Proposition 3.9, one obtains

ID′′(ζ)\displaystyle I^{\prime\prime}_{D}(\zeta) gB~,D3(ζ)|fk′′′(ζ)|2\displaystyle\geq g^{-3}_{\tilde{B},D}(\zeta)|f_{k}^{\prime\prime\prime}(\zeta)|^{2}
=FkD2|hk(ζ)|2gB~,D3(ζ)|f′′′(ζ)|2\displaystyle=\|F_{k}\|_{D}^{-2}|h^{k}(\zeta)|^{2}\,g^{-3}_{\tilde{B},D}(\zeta)|f^{\prime\prime\prime}(\zeta)|^{2}
FkD2|h(ζ)|2k(KD(ζ)KUD(ζ))6(JD(ζ)JUD(ζ))3gB~,UD3(ζ)|f′′′(ζ)|2\displaystyle\geq\|F_{k}\|_{D}^{-2}|h(\zeta)|^{2k}\left(\frac{K_{D}(\zeta)}{K_{U\cap D}(\zeta)}\right)^{6}\left(\frac{J_{D}(\zeta)}{J_{U\cap D}(\zeta)}\right)^{3}\,g^{-3}_{\tilde{B},U\cap D}(\zeta)|f^{\prime\prime\prime}(\zeta)|^{2}
|h(ζ)|2k(1+cbk)2(KD(ζ)KUD(ζ))6(JD(ζ)JUD(ζ))3IUD′′(ζ).\displaystyle\geq\frac{|h(\zeta)|^{2k}}{(1+cb^{k})^{2}}\left(\frac{K_{D}(\zeta)}{K_{U\cap D}(\zeta)}\right)^{6}\left(\frac{J_{D}(\zeta)}{J_{U\cap D}(\zeta)}\right)^{3}\,I^{\prime\prime}_{U\cap D}(\zeta).

This implies that

ID′′(ζ)IUD′′(ζ)|h(ζ)|2k(1+cbk)2(KD(ζ)KUD(ζ))6(JD(ζ)JUD(ζ))3.\frac{I^{\prime\prime}_{D}(\zeta)}{I^{\prime\prime}_{U\cap D}(\zeta)}\geq\frac{|h(\zeta)|^{2k}}{(1+cb^{k})^{2}}\left(\frac{K_{D}(\zeta)}{K_{U\cap D}(\zeta)}\right)^{6}\left(\frac{J_{D}(\zeta)}{J_{U\cap D}(\zeta)}\right)^{3}.

Note that h(p0)=1h(p^{0})=1. By (3.4), Proposition 2.1, and Proposition 2.4 of [10],

limζp0KD(ζ)KUD(ζ)=limζp0JD(ζ)JUD(ζ)=1.\lim_{\zeta\to p^{0}}\frac{K_{D}(\zeta)}{K_{U\cap D}(\zeta)}=\lim_{\zeta\to p^{0}}\frac{J_{D}(\zeta)}{J_{U\cap D}(\zeta)}=1.

Hence, letting ζp0\zeta\to p^{0} in the above inequality, we get

lim infζp0ID′′(ζ)IUD′′(ζ)(1+cbk)2.\liminf\limits_{\zeta\to p^{0}}\frac{I^{\prime\prime}_{D}(\zeta)}{I^{\prime\prime}_{U\cap D}(\zeta)}\geq(1+cb^{k})^{-2}.

Now letting kk\to\infty, as 0b<10\leq b<1, and cc is independent of kk, we obtain

lim infζp0ID′′(ζ)IUD′′(ζ)1.\liminf\limits_{\zeta\to p^{0}}\frac{I^{\prime\prime}_{D}(\zeta)}{I^{\prime\prime}_{U\cap D}(\zeta)}\geq 1. (3.29)

On the other hand, consider a candidate function η\eta for ID′′(ζ)I^{\prime\prime}_{D}(\zeta), i.e., ηA2(D)\eta\in A^{2}(D), ηD=1\|\eta\|_{D}=1, and η(ζ)=η(ζ)=η′′(ζ)=0\eta(\zeta)=\eta^{\prime}(\zeta)=\eta^{\prime\prime}(\zeta)=0. Now for zUDz\in U\cap D, let’s define

γ(z)=η(z)ηUD.\gamma(z)=\dfrac{\eta(z)}{\|\eta\|_{U\cap D}}.

Then clearly γA2(UD)\gamma\in A^{2}(U\cap D) with γUD=1\|\gamma\|_{U\cap D}=1, and γ(ζ)=γ(ζ)=γ′′(ζ)=0\gamma(\zeta)=\gamma^{\prime}(\zeta)=\gamma^{\prime\prime}(\zeta)=0. Therefore the maximality of IUD′′(ζ)I^{\prime\prime}_{U\cap D}(\zeta) implies

IUD′′(ζ)gB~,UD3(ζ)|γ′′′(ζ)|2=ηUD2(gB~,UD(ζ)gB~,D(ζ))3gB~,D3(ζ)|η′′′(ζ)|2(gB~,UD(ζ)gB~,D(ζ))3ID′′(ζ).I^{\prime\prime}_{U\cap D}(\zeta)\geq g^{-3}_{\tilde{B},U\cap D}(\zeta)|\gamma^{\prime\prime\prime}(\zeta)|^{2}=\|\eta\|^{-2}_{U\cap D}\left(\dfrac{g_{\tilde{B},U\cap D}(\zeta)}{g_{\tilde{B},D}(\zeta)}\right)^{-3}g^{-3}_{\tilde{B},D}(\zeta)|\eta^{\prime\prime\prime}(\zeta)|^{2}\\ \geq\left(\dfrac{g_{\tilde{B},U\cap D}(\zeta)}{g_{\tilde{B},D}(\zeta)}\right)^{-3}I^{\prime\prime}_{D}(\zeta).

The last inequality above follows from ηUD1\|\eta\|_{U\cap D}\leq 1 and the fact that η\eta  is an arbitrary candidate function for ID′′(ζ)I^{\prime\prime}_{D}(\zeta). Thus we obtain

ID′′(ζ)IUD′′(ζ)(gB~,UD(ζ)gB~,D(ζ))3.\frac{I^{\prime\prime}_{D}(\zeta)}{I^{\prime\prime}_{U\cap D}(\zeta)}\leq\left(\dfrac{g_{\tilde{B},U\cap D}(\zeta)}{g_{\tilde{B},D}(\zeta)}\right)^{3}. (3.30)

By (II) of Theorem 1.1, the right hand side converges to 11 as ζp0\zeta\to p^{0}, and therefore,

lim supζp0ID′′(ζ)IUD′′(ζ)1.\displaystyle\limsup_{\zeta\to p^{0}}\frac{I^{\prime\prime}_{D}(\zeta)}{I^{\prime\prime}_{U\cap D}(\zeta)}\leq 1. (3.31)

From (3.29) and (3.31), we conclude that

limζp0ID′′(ζ)IUD′′(ζ)=1,\lim\limits_{\zeta\to p^{0}}\frac{I^{\prime\prime}_{D}(\zeta)}{I^{\prime\prime}_{U\cap D}(\zeta)}=1,

as required. \qed

Lemma 3.15

Suppose {aj},{bj},{cj}\{a_{j}\},\{b_{j}\},\{c_{j}\} and {dj}\{d_{j}\} are real sequences with bj,dj>0b_{j},d_{j}>0 and

limjajbj=1andlimjcjdj=1.\lim_{j\to\infty}\dfrac{a_{j}}{b_{j}}=1\quad\text{and}\quad\lim_{j\to\infty}\dfrac{c_{j}}{d_{j}}=1.

Then we have

limjaj+cjbj+dj=1.\lim_{j\to\infty}\dfrac{a_{j}+c_{j}}{b_{j}+d_{j}}=1.
Proof 8

To verify this claim let ϵ>0\epsilon>0. Then there exists NN\in\mathbb{N} such that

|ajbj1|<ϵ2and|cjdj1|<ϵ2forjN.\bigg{|}\dfrac{a_{j}}{b_{j}}-1\bigg{|}<\dfrac{\epsilon}{2}\quad\text{and}\quad\bigg{|}\dfrac{c_{j}}{d_{j}}-1\bigg{|}<\dfrac{\epsilon}{2}\quad\text{for}\,\,j\geq N.

This, in particular implies that

|ajbj|bj+dj<ϵ2and|cjdj|bj+dj<ϵ2forjN.\dfrac{|a_{j}-b_{j}|}{b_{j}+d_{j}}<\dfrac{\epsilon}{2}\quad\text{and}\quad\dfrac{|c_{j}-d_{j}|}{b_{j}+d_{j}}<\dfrac{\epsilon}{2}\quad\text{for}\,\,j\geq N.

Hence we have

|aj+cjbj+dj1||ajbj|bj+dj+|cjdj|bj+dj<ϵforjN,\displaystyle\bigg{|}\dfrac{a_{j}+c_{j}}{b_{j}+d_{j}}-1\bigg{|}\leq\dfrac{|a_{j}-b_{j}|}{b_{j}+d_{j}}+\dfrac{|c_{j}-d_{j}|}{b_{j}+d_{j}}<\epsilon\quad\text{for}\,\,j\geq N,

proving the lemma. \qed

We are now in a state to give a proof of Theorem 1.1.

Proof of Theorem 1.1 (III) 4

By Proposition 3.11, let us write

RB~,D(z)=2E~D(z)F~D(z),R_{\tilde{B},D}(z)=2-\tilde{E}_{D}(z)-\tilde{F}_{D}(z),

where

E~D(z)=ID(z)KD(z)andF~D(z)=ID′′(z)ID(z).\tilde{E}_{D}(z)=\dfrac{I^{\prime}_{D}(z)}{K_{D}(z)}\quad\text{and}\quad\tilde{F}_{D}(z)=\dfrac{I^{\prime\prime}_{D}(z)}{I^{\prime}_{D}(z)}.

From Proposition 3.14 and (3.4), we have

limzp0E~D(z)E~UD(z)=1andlimzp0F~D(z)F~UD(z)=1,\lim\limits_{z\to p^{0}}\frac{\tilde{E}_{D}(z)}{\tilde{E}_{U\cap D}(z)}=1\quad\text{and}\quad\lim\limits_{z\to p^{0}}\frac{\tilde{F}_{D}(z)}{\tilde{F}_{U\cap D}(z)}=1,

for a small enough neighborhood UU of p0p^{0}. Now (III) follows immediately from Lemma 3.15. \qed

4 Boundary behavior

The proof of Theorem 1.2 is based on Pinchuk’s scaling method and we begin by recalling the change of coordinates associated to this method. We will use the standard notation z=(z,zn)z=(^{\prime}z,z_{n}), where z=(z1,,zn1){}^{\prime}z=(z_{1},\ldots,z_{n-1}). Throughout this section DD is a C2C^{2}-smoothly bounded strongly pseudoconvex domain in n\mathbb{C}^{n} and ρ\rho is a C2C^{2}-smooth local defining function for DD defined on a neighborhood UU of a point p0Dp^{0}\in\partial D. Without loss of generality, we assume that

z¯ρ(p0)=(0,1) and ρzn(z)0 for all zU.\nabla_{\overline{z}}\rho(p^{0})=({{}^{\prime}0},1)\text{ and }\frac{\partial\rho}{\partial z_{n}}(z)\neq 0\text{ for all }z\in U. (4.1)

Here, zρ=(ρ/z1,,ρ/zn)\nabla_{z}\rho=(\partial\rho/\partial z_{1},\ldots,\partial\rho/\partial z_{n}) and we write z¯ρ=zρ¯\nabla_{\overline{z}}\rho=\overline{\nabla_{z}\rho}. Note that the gradient ρ=2z¯ρ\nabla\rho=2\nabla_{\overline{z}}\rho.

4.1 Change of coordinates

The following lemma from [15] illustrates the change of coordinates near strongly pseudoconvex boundary points.

Lemma 4.16

There exist a family of biholomorphic mappings hζ:nnh^{\zeta}:\mathbb{C}^{n}\to\mathbb{C}^{n} depending continuously on ζDU\zeta\in\partial D\cap U, satisfying the following conditions:

  • (a)

    hp0=𝕀h^{p^{0}}=\mathbb{I}.

  • (b)

    hζ(ζ)=0h^{\zeta}(\zeta)=0.

  • (c)

    The local defining function ρζ=ρ(hζ)1\rho^{\zeta}=\rho\,\circ(h^{\zeta})^{-1} of the domain Dζ=hζ(D)D^{\zeta}=h^{\zeta}(D) near the origin has the form

    ρζ(z)=2Re(zn+Kζ(z))+Hζ(z)+o(|z2|)\rho^{\zeta}(z)=2\text{Re}\big{(}z_{n}+K^{\zeta}(z)\big{)}+H^{\zeta}(z)+o(|z^{2}|)

    in a neighborhood of the origin, where

    Kζ(z)=μ,ν=1naμν(ζ)zμzνandHζ(z)=μ,ν=1naμν¯(ζ)zμz¯νK^{\zeta}(z)=\sum\limits_{\mu,\nu=1}^{n}a_{\mu\nu}(\zeta)z_{\mu}z_{\nu}\quad\text{and}\quad H^{\zeta}(z)=\sum\limits_{\mu,\nu=1}^{n}a_{\mu\overline{\nu}}(\zeta)z_{\mu}\overline{z}_{\nu}

    with Kζ(z,0)0K^{\zeta}(^{\prime}z,0)\equiv 0 and Hζ(z,0)|z|2H^{\zeta}(^{\prime}z,0)\equiv|^{\prime}z|^{2}.

  • (d)

    The biholomorphism hζh^{\zeta} takes the real normal ηζ={z=ζ+2tz¯ρ(ζ):t}\eta_{\zeta}=\{z=\zeta+2t\nabla_{\overline{z}}\rho(\zeta):t\in\mathbb{R}\} to D\partial D at ζ\zeta into the real normal {z=yn=0}\{^{\prime}z=y_{n}=0\} to Dζ\partial D^{\zeta} at the origin.

The definition of the map hζh^{\zeta} and its derivative will play an important role in the computation of the boundary asymptotics and so we quickly recall its construction. We fix ζDU\zeta\in\partial D\cap U. The map hζh^{\zeta} is a polynomial automorphism of n\mathbb{C}^{n} defined as the composition hζ(z)=ϕ3ζϕ2ζϕ1ζ(z)h^{\zeta}(z)=\phi_{3}^{\zeta}\circ\phi_{2}^{\zeta}\circ\phi_{1}^{\zeta}(z), where the maps ϕiζ:nn\phi_{i}^{\zeta}:\mathbb{C}^{n}\to\mathbb{C}^{n} are biholomorphisms defined as follows: The map w=ϕ1ζ(z)w=\phi_{1}^{\zeta}(z) is an affine transformation given by

wj\displaystyle w_{j} =ρz¯n(ζ)(zjζj)ρz¯j(ζ)(znζn)forj=1,,n1,\displaystyle=\frac{\partial\rho}{\partial\overline{z}_{n}}(\zeta)(z_{j}-\zeta_{j})-\frac{\partial\rho}{\partial\overline{z}_{j}}(\zeta)(z_{n}-\zeta_{n})\quad\text{for}\quad j=1,\ldots,n-1, (4.2)
wn\displaystyle w_{n} =ν=1nρzν(ζ)(zνζν).\displaystyle=\sum\limits_{\nu=1}^{n}\frac{\partial\rho}{\partial z_{\nu}}(\zeta)(z_{\nu}-\zeta_{\nu}).

The map ϕ1ζ\phi_{1}^{\zeta} is nonsingular by (4.1) and it takes the point ζ\zeta to the origin. We relabel the new coordinates ww as zz. Then the Taylor series expansion of the local defining function ρ(ϕ1ζ)1\rho\circ(\phi_{1}^{\zeta})^{-1} for the domain ϕ1ζ(D)\phi_{1}^{\zeta}(D) near the origin has the form

2Re(zn+μ,ν=1naμν(ζ)zμzν)+Hζ(z)+o(|z|2),2\text{Re}\left(z_{n}+\sum\limits_{\mu,\nu=1}^{n}a_{\mu\nu}(\zeta)z_{\mu}z_{\nu}\right)+H^{\zeta}(z)+o(|z|^{2}), (4.3)

where Hζ(z)H^{\zeta}(z) is a Hermitian form.

The map w=ϕ2ζ(z)w=\phi_{2}^{\zeta}(z) is given by

w=(z,zn+μ,ν=1n1aμν(ζ)zμzν)w=\left({}^{\prime}z,z_{n}+\sum\limits_{\mu,\nu=1}^{n-1}a_{\mu\nu}(\zeta)z_{\mu}z_{\nu}\right) (4.4)

and is a polynomial automorphism. Relabelling the new coordinates ww as zz, the Taylor series expansion of the local defining function ρ(ϕ1ζ)1(ϕ2ζ)1\rho\circ(\phi^{\zeta}_{1})^{-1}\circ(\phi^{\zeta}_{2})^{-1} for the domain ϕ2ζϕ1ζ(D)\phi_{2}^{\zeta}\circ\phi_{1}^{\zeta}(D) has the form (4.3) with aμν=0a_{\mu\nu}=0 for 1μ,νn11\leq\mu,\nu\leq n-1.

Finally, the map ϕ3ζ\phi^{\zeta}_{3} is chosen so that the Hermitian form Hζ(z)H^{\zeta}(z) satisfies Hζ(z,0)=|z|2H^{\zeta}(^{\prime}z,0)=|^{\prime}z|^{2}. Since DD is strongly pseudoconvex and the complex tangent space to D\partial D at ζ\zeta is given by zn=0z_{n}=0 in the current coordinates, the form Hζ(z,0)H^{\zeta}(^{\prime}z,0) is strictly positive definite. Hence there exists a unitary map Uζ:n1n1U^{\zeta}:\mathbb{C}^{n-1}\rightarrow\mathbb{C}^{n-1} such that Hζ(Uζ(z),0)H^{\zeta}(U^{\zeta}(^{\prime}z),0) is diagonal with diagonal entries λ1ζ,,λn1ζ>0\lambda_{1}^{\zeta},\ldots,\lambda_{n-1}^{\zeta}>0. Now consider the stretching map Lζ=diag{(λ1ζ)1/2,,(λn1ζ)1/2}L^{\zeta}=\text{diag}\{(\lambda_{1}^{\zeta})^{-1/2},\ldots,(\lambda_{n-1}^{\zeta})^{-1/2}\}. Then the linear map Aζ:n1n1A^{\zeta}:\mathbb{C}^{n-1}\to\mathbb{C}^{n-1} given by Aζ:=LζUζA^{\zeta}:=L^{\zeta}\circ U^{\zeta} satisfies Hζ(Aζ(z),0)=|z|2H^{\zeta}(A^{\zeta}(^{\prime}z),0)=|^{\prime}z|^{2}. Note that UζU^{\zeta} and LζL^{\zeta}, and hence AζA^{\zeta} can be chosen to depend continuously on ζ\zeta. Thus, if we define w=ϕ3ζ(z)w=\phi^{\zeta}_{3}(z) by

w=(Aζ(z),zn),w=(A^{\zeta}(^{\prime}z),z_{n}),

and relabel ww as zz, then the local defining function ρζ=ρ(hζ)1\rho^{\zeta}=\rho\circ(h^{\zeta})^{-1} for the domain Dζ=hζ(D)D^{\zeta}=h^{\zeta}(D) near the origin has the Taylor series expansion as in (c). Also, it is evident from the construction of the maps hζh^{\zeta} that they satisfy (a), (b), (d), and that they depend continuously on ζ\zeta.

4.2 Scaling of DD

By strong pseudoconvexity, shrinking UU if necessary, there exist local holomorphic coordinates z1,,znz_{1},\ldots,z_{n} on UU in which p0=0p^{0}=0, and

ρ(z)=2Rezn+|z|2+o(Imzn,|z|2),zU,\rho(z)=2\operatorname{Re}z_{n}+|^{\prime}z|^{2}+o\big{(}\operatorname{Im}z_{n},|^{\prime}z|^{2}\big{)},\quad z\in U, (4.5)

and a constant 0<r<10<r<1 such that

UDΩ:={zn:2Rezn+r|z|2<0}.\displaystyle U\cap D\subset\Omega:=\big{\{}z\in\mathbb{C}^{n}:2\operatorname{Re}z_{n}+r|^{\prime}z|^{2}<0\big{\}}. (4.6)

Henceforth, we will be working in the above coordinates, and with U,ρU,\rho and p0p^{0} as above.

Let us consider a sequence of points pjp^{j} in DD that converges to p0=0p^{0}=0 on D\partial D. For jj sufficiently large, and without loss of generality we assume that for all jj, pjUp^{j}\in U and there exists a unique ζjDU\zeta^{j}\in\partial D\cap U that is closest to pjp^{j}. Define δj:=d(pj,D)=|pjζj|\delta_{j}:=\text{d}(p^{j},\partial D)=|p^{j}-\zeta^{j}|. Note that ζjp0=0\zeta^{j}\to p^{0}=0 and δj0\delta_{j}\to 0 as jj\to\infty. For each ζj\zeta^{j}, denote by hjh_{j} the map hζjh^{\zeta^{j}} given by Lemma 4.16. Denoting ϕiζj\phi_{i}^{\zeta^{j}} by ϕij\phi_{i}^{j} for j=1,2,3j=1,2,3, we have hj=ϕ1jϕ2jϕ3jh_{j}=\phi_{1}^{j}\circ\phi_{2}^{j}\circ\phi_{3}^{j}. Also set ρj=ρζj\rho_{j}=\rho^{\zeta^{j}}. Then by Lemma 4.16, near the origin,

ρj(z)=2Re(zn+Kj(z))+Hj(z)+o(|z|2),\displaystyle\rho_{j}(z)=2\text{Re}\,\big{(}z_{n}+K_{j}(z)\big{)}+H_{j}(z)+o\big{(}|z|^{2}\big{)},

where Kj=KζjK_{j}=K^{\zeta^{j}} and Hj=HζjH_{j}=H^{\zeta^{j}}. Moreover, thanks to the strong pseudoconvexity of D\partial D near p0=0p^{0}=0, shrinking UU if necessary and taking a smaller rr in (4.6), we have

hj(UD)Ω\displaystyle h_{j}(U\cap D)\subset\Omega (4.7)

for all large jj. Note that by Theorem 1.1, it is enough to prove Theorem 1.2 for the domain UDU\cap D, shrinking UU if necessary. Set Dj=hj(UD)D_{j}=h_{j}(U\cap D), qj=hj(pj)q^{j}=h_{j}(p^{j}), and ηj=d(qj,Dj)\eta_{j}=\text{d}(q^{j},\partial D_{j}).

Now consider the anisotropic dilation map Λj:nn\Lambda_{j}:\mathbb{C}^{n}\to\mathbb{C}^{n} defined by

Λj(z)=(z1ηj,,zn1ηj,znηj).\Lambda_{j}(z)=\left(\frac{z_{1}}{\sqrt{\eta_{j}}},\ldots,\frac{z_{n-1}}{\sqrt{\eta_{j}}},\frac{z_{n}}{\eta_{j}}\right). (4.8)

Set D~j=Λj(Dj)=Λjhj(UD)\tilde{D}_{j}=\Lambda_{j}(D_{j})=\Lambda_{j}\circ h_{j}(U\cap D). We will call the maps Sj:=ΛjhjS_{j}:=\Lambda_{j}\circ h_{j} the scaling maps and the domains D~j=Sj(UD)\widetilde{D}_{j}=S_{j}(U\cap D) the scaled domains. Note that since Sj(pj)=Λjhj(pj)=(0,1)S_{j}(p^{j})=\Lambda_{j}\circ h_{j}(p^{j})=(^{\prime}0,-1), each D~j\tilde{D}_{j} contains the point (0,1)(^{\prime}0,-1) and we will denote this point by bb^{*}. A defining function for D~j\tilde{D}_{j} near the origin, is given by

ρ~j(z)=1ηjρj(Λj1(z))=2Re(zn+1ηjKj(Λj1(z)))+1ηjHj(Λj1(z))+o(ηj1/2|z|2).\tilde{\rho}_{j}(z)=\frac{1}{\eta_{j}}\rho_{j}\big{(}\Lambda_{j}^{-1}(z)\big{)}=2\text{Re}\Big{(}z_{n}+\frac{1}{\eta_{j}}K_{j}\big{(}\Lambda_{j}^{-1}(z)\big{)}\Big{)}+\frac{1}{\eta_{j}}H_{j}\big{(}\Lambda_{j}^{-1}(z)\big{)}+o\big{(}\eta_{j}^{1/2}|z|^{2}\big{)}.

Since Kj(z)K_{j}(z) and Hj(z)H_{j}(z) satisfy condition (c) of Lemma 4.16, it follows that

limj1ηjKj(Λj1z)=0andlimj1ηjHj(Λj1z)=|z|2\lim_{j\to\infty}\frac{1}{\eta_{j}}K_{j}(\Lambda_{j}^{-1}z)=0\quad\text{and}\quad\lim_{j\to\infty}\frac{1}{\eta_{j}}H_{j}(\Lambda_{j}^{-1}z)=|^{\prime}z|^{2}

in C2C^{2}-topology on compact subsets of n\mathbb{C}^{n}. Evidently, o(ηj1/2|z|2)0o(\eta_{j}^{1/2}|z|^{2})\to 0 as jj\to\infty in C2C^{2}-topology on any compact set of n\mathbb{C}^{n}. Thus, the defining functions ρ~j\tilde{\rho}_{j} converge in C2C^{2} topology on compact subsets of n\mathbb{C}^{n} to

ρ(z)=2Rezn+|z|2.\rho_{\infty}(z)=2\text{Re}\,z_{n}+|^{\prime}z|^{2}.

Hence our scaled domains D~j\tilde{D}_{j} converge in the local Hausdorff sense to the Siegel upper half-space

D={zn:2Rezn+|z|2<0}.D_{\infty}=\big{\{}z\in\mathbb{C}^{n}:2\text{Re}\,z_{n}+|^{\prime}z|^{2}<0\big{\}}.

4.3 Stability of the Kobayashi–Fuks metric

Proposition 4.17

For zDz\in D_{\infty} and un{0}u\in\mathbb{C}^{n}\setminus\{0\}, we have

gB~,D~j(z)gB~,D(z),τB~,D~j(z,u)τB~,D(z,u)andRicB~,D~j(z,u)RicB~,D(z,u)g_{\tilde{B},\tilde{D}_{j}}(z)\to g_{\tilde{B},D_{\infty}}(z),\quad\tau_{\tilde{B},\tilde{D}_{j}}(z,u)\to\tau_{\tilde{B},D_{\infty}}(z,u)\quad\text{and}\quad\operatorname{Ric}_{\tilde{B},\tilde{D}_{j}}(z,u)\to\operatorname{Ric}_{\tilde{B},D_{\infty}}(z,u)

as jj\to\infty. Moreover, the first convergence is uniform on compact subsets of  DD_{\infty} and the second and third convergences are uniform on compact subsets of  D×nD_{\infty}\times\mathbb{C}^{n}.

Proof 9

Since the Kobayashi–Fuks metric on the domain DD has Kähler potential log(KDn+1gB,D)\log(K^{n+1}_{D}g_{B,D}), i.e.,

gαβ¯B~,D=2log(KDn+1gB,D)zαz¯β,g^{\tilde{B},D}_{\alpha\overline{\beta}}=\frac{\partial^{2}\log(K^{n+1}_{D}g_{B,D})}{\partial z_{\alpha}\partial\overline{z}_{\beta}},

all that is required is to show that

KD~jn+1gB,D~jKDn+1gB,DK^{n+1}_{\tilde{D}_{j}}g_{B,\tilde{D}_{j}}\to K^{n+1}_{D_{\infty}}g_{B,D_{\infty}}

uniformly on compact subsets of DD_{\infty}, together with all derivatives. But this is an immediate consequence of the fact that KD~jKDK_{\tilde{D}_{j}}\to K_{D_{\infty}} together will all derivatives on compact subsets of DD_{\infty}. This fact can be established from a Ramadanov type result [16, Lemma 2.1], and for the details we refer the reader to Lemma 5.3 of [17] with the note that by taking d=0d=0 there, KD~j=KD~j,0K_{\tilde{D}_{j}}=K_{\tilde{D}_{j},0} and KD=KD,0K_{D_{\infty}}=K_{D_{\infty},0}. \qed

4.4 Boundary asymptotics

Recall that Sj=ΛjhjS_{j}=\Lambda_{j}\circ h_{j}, Sj(UD)=D~jS_{j}(U\cap D)=\tilde{D}_{j}, and Sj(pj)=b=(0,1)S_{j}(p^{j})=b^{*}=(^{\prime}0,-1). Denoting the matrix of a linear map by itself, we have

Sj(pj)=Λjhj(pj)=Λjϕ3j(ϕ2j)(ϕ1j(pj))(ϕ1j)(pj).S_{j}^{\prime}(p^{j})=\Lambda_{j}h_{j}^{\prime}(p^{j})=\Lambda_{j}\cdot\phi_{3}^{j}\cdot(\phi_{2}^{j})^{\prime}\big{(}\phi_{1}^{j}(p^{j})\big{)}\cdot(\phi_{1}^{j})^{\prime}(p^{j}). (4.9)

Note that from the definition of ϕ1j\phi_{1}^{j},

(ϕ1j)(pj)=(ρz¯n(ζj)00ρz¯1(ζj)0ρz¯n(ζj)0ρz¯2(ζj)00ρz¯n(ζj)ρz¯n1(ζj)ρz1(ζj)ρz2(ζj)ρzn1(ζj)ρzn(ζj)).(\phi_{1}^{j})^{\prime}(p^{j})=\begin{pmatrix}\dfrac{\partial\rho}{\partial\overline{z}_{n}}(\zeta^{j})&0&\cdots&0&-\dfrac{\partial\rho}{\partial\overline{z}_{1}}(\zeta^{j})\\ 0&\dfrac{\partial\rho}{\partial\overline{z}_{n}}(\zeta^{j})&\cdots&0&-\dfrac{\partial\rho}{\partial\overline{z}_{2}}(\zeta^{j})\\ \vdots&\vdots&\ddots&\vdots&\vdots\\ 0&0&\cdots&\dfrac{\partial\rho}{\partial\overline{z}_{n}}(\zeta^{j})&-\dfrac{\partial\rho}{\partial\overline{z}_{n-1}}(\zeta^{j})\\ \dfrac{\partial\rho}{\partial z_{1}}(\zeta^{j})&\dfrac{\partial\rho}{\partial z_{2}}(\zeta^{j})&\cdots&\dfrac{\partial\rho}{\partial z_{n-1}}(\zeta^{j})&\dfrac{\partial\rho}{\partial z_{n}}(\zeta^{j})\end{pmatrix}. (4.10)

Also, since

pj=ζjδjz¯ρ(ζj)|z¯ρ(ζj)|,p^{j}=\zeta^{j}-\delta_{j}\frac{\nabla_{\overline{z}}\rho(\zeta^{j})}{\big{|}\nabla_{\overline{z}}\rho(\zeta^{j})\big{|}}, (4.11)

we have

ϕ1j(pj)=(0,δj|z¯ρ(ζj)|).\phi_{1}^{j}(p^{j})=\Big{(}{}^{\prime}0,-\delta_{j}\big{|}\nabla_{\overline{z}}\rho(\zeta^{j})\big{|}\Big{)}. (4.12)

Therefore, from the definition of ϕ2j\phi_{2}^{j}, we have

(ϕ2j)(ϕ1j(pj))=𝕀n.\displaystyle(\phi_{2}^{j})^{\prime}\big{(}\phi_{1}^{j}(p^{j})\big{)}=\mathbb{I}_{n}.

Finally, recall that ϕ3j(z)=(Aj(z),zn)\phi^{j}_{3}(z)=(A^{j}(^{\prime}z),z_{n}) where Aj:=Aζj:n1n1A^{j}:=A^{\zeta^{j}}:\mathbb{C}^{n-1}\to\mathbb{C}^{n-1} are linear maps satisfying Aj𝕀n1A^{j}\to\mathbb{I}_{n-1}. Therefore,

ϕ3j=[Ap,qj001],\phi^{j}_{3}=\begin{bmatrix}A^{j}_{p,q}&0\\ 0&1\end{bmatrix},

where Ap,qjA^{j}_{p,q} are the entries of the matrix of AjA^{j}. Thus,

hj(pj)=(A1,1jρz¯n(ζj)A1,n1jρz¯n(ζj)ν=1n1A1,νjρz¯ν(ζj)An1,1jρz¯n(ζj)An1,n1jρz¯n(ζj)ν=1n1An1,νjρz¯ν(ζj)ρz1(ζj)ρzn1(ζj)ρzn(ζj))𝕀nh_{j}^{\prime}(p^{j})=\begin{pmatrix}A^{j}_{1,1}\dfrac{\partial\rho}{\partial\overline{z}_{n}}(\zeta^{j})&\cdots&A^{j}_{1,n-1}\dfrac{\partial\rho}{\partial\overline{z}_{n}}(\zeta^{j})&-\sum\limits_{\nu=1}^{n-1}A^{j}_{1,\nu}\dfrac{\partial\rho}{\partial\overline{z}_{\nu}}(\zeta^{j})\\ \vdots&\cdots&\vdots&\vdots\\ A^{j}_{n-1,1}\dfrac{\partial\rho}{\partial\overline{z}_{n}}(\zeta^{j})&\cdots&A^{j}_{n-1,n-1}\dfrac{\partial\rho}{\partial\overline{z}_{n}}(\zeta^{j})&-\sum\limits_{\nu=1}^{n-1}A^{j}_{n-1,\nu}\dfrac{\partial\rho}{\partial\overline{z}_{\nu}}(\zeta^{j})\\ \dfrac{\partial\rho}{\partial z_{1}}(\zeta^{j})&\cdots&\dfrac{\partial\rho}{\partial z_{n-1}}(\zeta^{j})&\dfrac{\partial\rho}{\partial z_{n}}(\zeta^{j})\end{pmatrix}\to\mathbb{I}_{n} (4.13)

in the operator norm.

We also note that as ϕ2j\phi_{2}^{j} and ϕ3j\phi_{3}^{j} fix points on the Rezn\operatorname{Re}z_{n}-axis, we have from (4.12),

qj=hj(pj)=(0,δj|z¯ρ(ζj)|).q^{j}=h_{j}(p^{j})=\Big{(}{}^{\prime}0,-\delta_{j}\big{|}\nabla_{\overline{z}}\rho(\zeta^{j})\big{|}\Big{)}.

As the normal to Dj\partial D_{j} at 0 is the Rezn\operatorname{Re}z_{n}-axis and ηj=d(qj,Dj)\eta_{j}=\text{d}(q^{j},\partial D_{j}), we have ηj=δj|z¯ρ(ζj)|\eta_{j}=\delta_{j}\,|\nabla_{\overline{z}}\rho(\zeta^{j})| and hence

limjηjδj=1.\lim_{j\to\infty}\frac{\eta_{j}}{\delta_{j}}=1. (4.14)

Now consider the Cayley transform Φ\Phi defined by

Φ(z1,,zn)=(2zzn1,zn+1zn1).\Phi(z_{1},\ldots,z_{n})=\left(\dfrac{\sqrt{2}\,{}^{\prime}z}{z_{n}-1},\dfrac{z_{n}+1}{z_{n}-1}\right). (4.15)

It can be shown that Φ\Phi maps DD_{\infty} biholomorphically onto 𝔹n\mathbb{B}^{n}. We also note that b=(0,1)Db^{*}=(^{\prime}0,-1)\in D_{\infty}, Φ(b)=0\Phi(b^{*})=0, and

Φ(b)=diag{1/2,,1/2,1/2}.\Phi^{\prime}(b^{*})=-\operatorname{diag}\{1/\sqrt{2},\ldots,1/\sqrt{2},1/2\}. (4.16)

We now present the proof of Theorem 1.2.

Proof of Theorem 1.2 5

Note that by the localization result Theorem 1.1, it is enough to compute the asymptotics for the domain UDU\cap D.

(I) By invariance of the Kobayashi–Fuks metric,

τB~,UD(pj,u)=τB~,D~j(b,Sj(pj)u).\displaystyle\tau_{\tilde{B},U\cap D}(p^{j},u)=\tau_{\tilde{B},\tilde{D}_{j}}\big{(}b^{*},S_{j}^{\prime}(p^{j})u\big{)}.

Note that

Sj(pj)u=Λj(hj(pj)u)=(ηj1/2(hj(pj)u),ηj1(hj(pj)u)n),S_{j}^{\prime}(p^{j})u=\Lambda_{j}\big{(}h_{j}^{\prime}(p^{j})u\big{)}=\Big{(}\eta_{j}^{-1/2}\,\,{{}^{\prime}\big{(}h_{j}^{\prime}(p^{j})u\big{)}},\eta_{j}^{-1}\big{(}h_{j}^{\prime}(p^{j})u\big{)}_{n}\Big{)},

and so by (4.13),

ηjSj(pj)u(0,un)\eta_{j}S_{j}^{\prime}(p^{j})u\to(^{\prime}0,u_{n})

uniformly in unit vectors uu. Therefore, by Proposition 4.17,

limjδjτB~,UD(pj,u)=limjδjηjτB~,D~j(b,ηjSj(pj)u)=τB~,D(b,(0,un))\displaystyle\lim_{j\to\infty}\delta_{j}\,\tau_{\tilde{B},U\cap D}(p^{j},u)=\lim_{j\to\infty}\frac{\delta_{j}}{\eta_{j}}\tau_{\tilde{B},\tilde{D}_{j}}\big{(}b^{*},\eta_{j}S_{j}^{\prime}(p^{j})u\big{)}=\tau_{\tilde{B},D_{\infty}}\big{(}b^{*},(^{\prime}0,u_{n})\big{)}

uniformly in unit vectors uu. Now, all that is required is to compute the right hand side using the Cayley transform Φ\Phi from (4.15), its derivative from (4.16), and the transformation rule. Thus,

τB~,D(b,(0,un))=τB~,𝔹n(0,(0,un2))=12(n+1)(n+2)|un|,\tau_{\tilde{B},D_{\infty}}\big{(}b^{*},(^{\prime}0,u_{n})\big{)}=\tau_{\tilde{B},\mathbb{B}^{n}}\bigg{(}0,\left({}^{\prime}0,-\dfrac{u_{n}}{2}\right)\bigg{)}=\dfrac{1}{2}\sqrt{(n+1)(n+2)}|u_{n}|,

where the last equality follows from Proposition 2.5, and this proves (I).

(II) For brevity, we write uj=uH(pj)u^{j}=u_{H}(p^{j}) and u0=uH(p0)u^{0}=u_{H}(p^{0}). By invariance of the Kobayashi–Fuks metric,

τB~,UD(pj,uH(pj))=τB~,D~j(b,Sj(pj)uj).\displaystyle\tau_{\tilde{B},U\cap D}\big{(}p^{j},u_{H}(p^{j})\big{)}=\tau_{\tilde{B},\tilde{D}_{j}}\big{(}b^{*},S_{j}^{\prime}(p^{j})u^{j}\big{)}.

Note that, since ujHqj(D)u^{j}\in H_{q^{j}}(\partial D), we have from (4.13)

hj(pj)uj=(v1j,,vn1j,0),h_{j}^{\prime}(p^{j})u^{j}=(v_{1}^{j},\ldots,v_{n-1}^{j},0),

where

vlj=ν=1n1Al,νj(uνjρz¯n(ζj)unjρz¯ν(ζj)),l=1,,n1.v^{j}_{l}=\sum_{\nu=1}^{n-1}A^{j}_{l,\nu}\left(u^{j}_{\nu}\frac{\partial\rho}{\partial\overline{z}_{n}}(\zeta^{j})-u^{j}_{n}\frac{\partial\rho}{\partial\overline{z}_{\nu}}(\zeta^{j})\right),\quad l=1,\ldots,n-1.

Therefore,

Sj(pj)uj=Λj(hj(pj)uj)=(v1jηj,,vn1jηj,0).\displaystyle S_{j}^{\prime}(p^{j})u^{j}=\Lambda_{j}\big{(}h_{j}^{\prime}(p^{j})u^{j}\big{)}=\left(\frac{v^{j}_{1}}{\sqrt{\eta_{j}}},\ldots,\frac{v^{j}_{n-1}}{\sqrt{\eta_{j}}},0\right).

Observe that vljul0v^{j}_{l}\to u^{0}_{l} and the convergence is uniform on unit vectors uu and so

ηjSj(pj)uj(u10,,un10,0)\sqrt{\eta_{j}}S_{j}^{\prime}(p^{j})u^{j}\to(u^{0}_{1},\ldots,u^{0}_{n-1},0)

and the convergence is uniform in unit vectors uu. Hence, by Proposition 4.17,

δjτB~,UD(pj,uH(pj))=δjηjτB~,D~j(b,ηjSj(pj)uj)τB~,D(b,(u0,0))\sqrt{\delta_{j}}\tau_{\tilde{B},U\cap D}\big{(}p^{j},u_{H}(p^{j})\big{)}=\sqrt{\frac{\delta_{j}}{\eta_{j}}}\tau_{\tilde{B},\tilde{D}_{j}}\big{(}b^{*},\sqrt{\eta_{j}}S_{j}^{\prime}(p^{j})u^{j}\big{)}\to\tau_{\tilde{B},D_{\infty}}\big{(}b^{*},(^{\prime}u^{0},0)\big{)}

uniformly in unit vectors uu. Again, using the Cayley transform Φ\Phi from (4.15) and its derivative from (4.16), the transformation rule gives

τB~,D(b,(u0,0))=τB~,𝔹n(0,(u02,0))=12(n+1)(n+2)|u0|2,\displaystyle\tau_{\tilde{B},D_{\infty}}\big{(}b^{*},(^{\prime}u^{0},0)\big{)}=\tau_{\tilde{B},\mathbb{B}^{n}}\bigg{(}0,\bigg{(}-\dfrac{{}^{\prime}u^{0}}{\sqrt{2}},0\bigg{)}\bigg{)}=\sqrt{\dfrac{1}{2}(n+1)(n+2)|^{\prime}u^{0}|^{2}},

by Proposition 2.5. This proves (II) once we observe from (4.5) that ρ(p0,uH(p0))=|u0|2\mathcal{L}_{\rho}\big{(}p^{0},u_{H}(p^{0})\big{)}=|^{\prime}u^{0}|^{2}.

(III) By the transformation rule for the Kobayashi–Fuks metric, we have

gB~,UD(pj)=gB~,D~j(b)|detSj(pj)|2.\displaystyle g_{\tilde{B},U\cap D}(p^{j})=g_{\tilde{B},\tilde{D}_{j}}(b^{*})\big{|}\det S^{\prime}_{j}(p^{j})\big{|}^{2}. (4.17)

Note that

detSj(pj)=detΛjdethj(pj)=ηj(n+1)/2dethj(pj),\det S^{\prime}_{j}(p^{j})=\det\Lambda_{j}\det h_{j}^{\prime}(p^{j})=\eta_{j}^{-(n+1)/2}\det h_{j}^{\prime}(p^{j}),

and so by (4.13),

ηjn+1|detSj(pj)|21.\eta_{j}^{n+1}|\det S^{\prime}_{j}(p^{j})|^{2}\to 1.

Therefore,

δjn+1gB~,UD(pj)=(δjηj)n+1gB~,D~j(b)ηjn+1|detSj(pj)|2gB~,D(b).\delta_{j}^{n+1}g_{\tilde{B},U\cap D}(p^{j})=\left(\frac{\delta_{j}}{\eta_{j}}\right)^{n+1}g_{\tilde{B},\tilde{D}_{j}}(b^{*})\eta_{j}^{n+1}\big{|}\det S^{\prime}_{j}(p^{j})\big{|}^{2}\to g_{\tilde{B},D_{\infty}}(b^{*}).

As before, using the Cayley transform Φ\Phi from (4.15) and its derivative from (4.16), we obtain from the transformation rule,

gB~,D(b)=gB~,𝔹n(0)|detΦ(b)|2=(n+1)n(n+2)n2n+1,g_{\tilde{B},D_{\infty}}(b^{*})=g_{\tilde{B},\mathbb{B}^{n}}(0)\,|\text{det}\,\Phi^{\prime}(b^{*})|^{2}=\dfrac{(n+1)^{n}(n+2)^{n}}{2^{n+1}},

by Proposition 2.5. This completes the proof of (III).

(IV) By invariance and Proposition 4.17,

RB~,UD(pj)=RB~,D~j(b)RB~,D(b)=RB~,Δ(0).R_{\tilde{B},U\cap D}(p^{j})=R_{\tilde{B},\tilde{D}_{j}}(b^{*})\to R_{\tilde{B},D_{\infty}}(b^{*})=R_{\tilde{B},\Delta}(0).

By Proposition 2.5, we have for zΔz\in\Delta,

gB~,Δ(z)=6(1|z|2)2,g_{\tilde{B},\Delta}(z)=\frac{6}{(1-|z|^{2})^{2}},

and therefore,

RB~,Δ(z)=1gB~,Δ(z)2loggB~,Δzz¯(z)=13,R_{\tilde{B},\Delta}(z)=-\frac{1}{g_{\tilde{B},\Delta}(z)}\frac{\partial^{2}\log g_{\tilde{B},\Delta}}{\partial z\partial\overline{z}}(z)=-\frac{1}{3},

which completes the proof of (IV) and the theorem. \qed

Remark 4.18

The proof of Theorem 1.2 (IV) also follows from the observation that we can choose a sufficiently small disc UU centered at p0p^{0} such that UDU\cap D is simply connected thanks to the smoothness of D\partial D. Therefore, UDU\cap D is biholomorphic to the unit disc Δ\Delta and hence the Kobayashi–Fuks metric of UDU\cap D has the constant Gaussian curvature 1/3-1/3 which follows from Proposition 2.5. Now, applying Theorem 1.1, we immediately obtain RB~,D(pj)1/3R_{\tilde{B},D}(p^{j})\to-1/3.

5 Existence of closed geodesics with prescribed homotopy class

The proof of Theorem 1.4 is based on the following result:

Theorem 5.19 (Herbort, [8, Theorem1.1])

Let GNG\subset\mathbb{R}^{N} be a bounded domain such that π1(G)\pi_{1}(G) is nontrivial and the following conditions are satisfied:

  • (i)

    For each pG¯p\in\overline{G} there is an open neighborhood UnU\subset\mathbb{R}^{n}, such that the set GUG\cap U is simply connected.

  • (ii)

    The domain GG is equipped with a complete Riemannian metric gg which possesses the following property:

    • (P)

      For each S>0S>0 there is a δ>0\delta>0 such that for every point pGp\in G with d(p,G)<δ\text{d}(p,\partial G)<\delta and every XnX\in\mathbb{R}^{n}, g(p,X)SX2g(p,X)\geq S\|X\|^{2}.

Then every nontrivial homotopy class in π1(G)\pi_{1}(G) contains a closed geodesic for gg.

Proof of Theorem 1.4 6

We will show that both the conditions in Theorem 5.19 hold for G=DG=D and g=dsB~,D2g=ds^{2}_{\tilde{B},D}. By the smoothness of D\partial D, it is evident that condition (i) is satisfied. For condition (ii), note that from (3.3) and the fact that RicB,D(z,u)\text{Ric}_{B,D}(z,u) approaches 1-1 near the boundary of a strongly pseudoconvex domain (see [10, Corollary 2]), there exists C=C(D)>0C=C(D)>0 such that

τB~,D(z,u)CτB,D(z,u)\displaystyle\tau_{\tilde{B},D}(z,u)\geq C\tau_{B,D}(z,u) (5.1)

for zz near the boundary of DD and unit vectors uu. As both the Bergman and Kobayashi–Fuks metric are Kähler, this relation also holds for zz on any compact subset of DD and unit vectors uu. Thus (5.1) holds for all zDz\in D and unu\in\mathbb{C}^{n}. This has the following two consequences. Firstly, since the Bergman metric dominates the Carathéodory metric on bounded domains (see Hahn [18]) and the Carathéodory metric is complete on smoothly bounded strongly pseudoconvex domains (see [19], p.539), (5.1) implies that the Kobayashi–Fuks metric on DD is complete. Secondly, as the Bergman metric on DD satisfies property (P) which was observed in the proof of Theorem 1.2 in [8], (5.1) also implies that the Kobayashi–Fuks metric on DD satisfies property (P) as well, and hence condition (ii) holds. This completes the proof of the theorem. \qed

6 Some questions

We conclude this article with the following questions:

  • (i)

    Does the localization of holomorphic sectional curvature and Ricci curvature of the Kobayashi–Fuks metric near holomorphic peak points hold in dimensions n2n\geq 2?

  • (ii)

    Herbort studied the existence of geodesic spirals for the Bergman metric on strongly pseudoconvex domains [8, Theorem 3.2]. Does the analog of this result hold for the Kobayashi–Fuks metric?

  • (iii)

    Is there an analog of the Donnelly-Fefferman’s result on the L2L^{2}-cohomology of the Bergman metric [20, Theorem 1.1] for the Kobayashi–Fuks metric? See also [21, 22] for simpler proofs based on the fact that the Bergman metric is given by a global potential. Note that the Kobayashi–Fuks metric is also given by a global potential.

References

References

  • [1] S. Kobayashi, Geometry of bounded domains, Trans. Amer. Math. Soc. 92 (1959) 267–290. doi:10.2307/1993156.
  • [2] B. A. Fuks, The Ricci curvature of the Bergman metric invariant with respect to biholomorphic mappings, Dokl. Akad. Nauk SSSR 167 (1966) 996–999.
  • [3] Ż. Dinew, On the completeness of a metric related to the Bergman metric, Monatsh. Math. 172 (3-4) (2013) 277–291. doi:10.1007/s00605-013-0501-6.
  • [4] J.-P. Demailly, Mesures de Monge-Ampère et mesures pluriharmoniques, Math. Z. 194 (4) (1987) 519–564. doi:10.1007/BF01161920.
  • [5] Ż. Dinew, On the Bergman representative coordinates, Sci. China Math. 54 (7) (2011) 1357–1374. doi:10.1007/s11425-011-4243-4.
  • [6] I. Graham, Boundary behavior of the Carathéodory and Kobayashi metrics on strongly pseudoconvex domains in CnC^{n} with smooth boundary, Trans. Amer. Math. Soc. 207 (1975) 219–240. doi:10.2307/1997175.
  • [7] R. E. Greene, S. G. Krantz, Deformation of complex structures, estimates for the ¯\bar{\partial} equation, and stability of the Bergman kernel, Adv. in Math. 43 (1) (1982) 1–86. doi:10.1016/0001-8708(82)90028-7.
  • [8] G. Herbort, On the geodesics of the Bergman metric, Math. Ann. 264 (1) (1983) 39–51. doi:10.1007/BF01458049.
  • [9] K.-T. Kim, J. Yu, Boundary behavior of the Bergman curvature in strictly pseudoconvex polyhedral domains, Pacific J. Math. 176 (1) (1996) 141–163.
  • [10] S. G. Krantz, J. Yu, On the Bergman invariant and curvatures of the Bergman metric, Illinois J. Math. 40 (2) (1996) 226–244.
  • [11] N. Nikolov, Localization of invariant metrics, Arch. Math. (Basel) 79 (1) (2002) 67–73. doi:10.1007/s00013-002-8286-1.
  • [12] N. Nikolov, Comparison of invariant functions on strongly pseudoconvex domains, J. Math. Anal. Appl. 421 (1) (2015) 180–185. doi:10.1016/j.jmaa.2014.07.007.
  • [13] N. Nikolov, M. Trybuła, Estimates for the squeezing function near strictly pseudoconvex boundary points with applications, J. Geom. Anal. 30 (2) (2020) 1359–1365. doi:10.1007/s12220-019-00348-3.
  • [14] L. Hörmander, An introduction to complex analysis in several variables, second revised edition Edition, North-Holland Publishing Co., Amsterdam-London; American Elsevier Publishing Co., Inc., New York, 1973, north-Holland Mathematical Library, Vol. 7.
  • [15] S. I. Pinčuk, Holomorphic inequivalence of certain classes of domains in 𝐂n{\bf C}^{n}, Mat. Sb. (N.S.) 111(153) (1) (1980) 67–94, 159.
  • [16] G. P. Balakumar, D. Borah, P. Mahajan, K. Verma, Remarks on the higher dimensional Suita conjecture, Proc. Amer. Math. Soc. 147 (8) (2019) 3401–3411. doi:10.1090/proc/14421.
  • [17] D. Borah, K. Verma, Narasimhan–Simha type metrics on strongly pseudoconvex domains in n\mathbb{C}^{n}, arXiv:2105.07723.
  • [18] K. T. Hahn, On completeness of the Bergman metric and its subordinate metric, Proc. Nat. Acad. Sci. U.S.A. 73 (12) (1976) 4294. doi:10.1073/pnas.73.12.4294.
  • [19] M. Jarnicki, P. Pflug, Invariant distances and metrics in complex analysis, second extended edition Edition, Vol. 9 of De Gruyter Expositions in Mathematics, Walter de Gruyter GmbH & Co. KG, Berlin, 2013. doi:10.1515/9783110253863.
  • [20] H. Donnelly, C. Fefferman, L2L^{2}-cohomology and index theorem for the Bergman metric, Ann. of Math. (2) 118 (3) (1983) 593–618. doi:10.2307/2006983.
  • [21] H. Donnelly, L2L_{2} cohomology of pseudoconvex domains with complete Kähler metric, Michigan Math. J. 41 (3) (1994) 433–442. doi:10.1307/mmj/1029005071.
  • [22] H. Donnelly, L2L_{2} cohomology of the Bergman metric for weakly pseudoconvex domains, Illinois J. Math. 41 (1) (1997) 151–160.