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

On plurisubharmonic defining functions for pseudoconvex domains in 2\mathbb{C}^{2}

Anne-Katrin Gallagher Gallagher Tool and Instrument LLC, Redmond, WA 98052, USA [email protected]  and  Tobias Harz Bergische Universitat Wuppertal, Wuppertal, Germany [email protected]
Abstract.

We investigate the question of existence of plurisubharmonic defining functions for smoothly bounded, pseudoconvex domains in 2\mathbb{C}^{2}. In particular, we construct a family of simple counterexamples to the existence of plurisubharmonic smooth local defining functions. Moreover, we give general criteria equivalent to the existence of plurisubharmonic smooth defining functions on or near the boundary of the domain. These equivalent characterizations are then explored for some classes of domains.

Key words and phrases:
Plurisubharmonic defining functions, finite type
2020 Mathematics Subject Classification:
32Txx, 32U05, 32T25

1. Introduction

In this paper, we investigate the question of existence of plurisubharmonic smooth defining functions for pseudoconvex domains with smooth boundary. This basic question has been long resolved for strictly pseudoconvex domains. However, there is a great lack of understanding in the case of weakly pseudoconvex domains, although the existence of plurisubharmonic defining functions is of relevance, e.g., for the classification problem of domains in n\mathbb{C}^{n}.

It is a basic fact that a smoothly bounded domain in n\mathbb{C}^{n} has a plurisubharmonic smooth local defining function near a boundary point if the domain is strictly pseudoconvex or convex near that point; similarly, the domain admits a smooth plurisubharmonic defining function near the boundary if it is strictly pseudoconvex or convex at each boundary point. No other geometric conditions which are sufficient for the existence of plurisubharmonic smooth (local) defining functions are known. Moreover, the existence of local or global plurisubharmonic defining functions may fail on pseudoconvex domains with weakly pseudoconvex boundary points. For instance, the worm domain, constructed by Diederich–Fornæss in [4], does not admit a plurisubharmonic global defining function, although it does admit plurisubharmonic local defining functions near each boundary point. Further, Fornæss [6] constructs a smoothly bounded, pseudoconvex domain in 3\mathbb{C}^{3} for which all 𝒞2\mathcal{C}^{2}-smooth local defining functions fail to be plurisubharmonic on the boundary near some boundary point. Behrens [1] gives an example of a pseudoconvex domain in 2\mathbb{C}^{2} with real-analytic boundary which exhibits the same failure of plurisubharmonicity of 𝒞6\mathcal{C}^{6}-smooth local defining functions. Behrens’ domain is of type 66 at the boundary point in question. No such examples are known for pseudoconvex domains which are of type 44 at the considered boundary point.

In the first part of this paper, we construct a family of counterexamples in the spirit of Behrens’ example in [1]. Namely, we construct domains Ω2k2\Omega_{2k}\subset\mathbb{C}^{2}, k3k\geq 3, such that Ω2k\Omega_{2k} is a domain with real-analytic boundary that is pseudoconvex and of type 2k2k at some boundary point pp but any 𝒞2\mathcal{C}^{2}-smooth defining function of Ω\Omega near pp fails to be plurisubharmonic on the boundary.

The second part of the paper is concerned with introducing new geometric conditions that are sufficient for the existence of plurisubharmonic smooth (local) defining functions. We first give equivalent, yet non-geometric, characterizations for the existence of plurisubharmonic smooth local defining functions, both on and near the boundary, see Proposition 5.2 and Proposition 6.8. These characterizations are then exploited to show the existence of plurisubharmonic defining functions under each of two newly introduced geometric conditions for smoothly bounded pseudoconvex domains.

The first condition pertains to type 44 boundary points pp of pseudconvex domains Ω2\Omega\subset\mathbb{C}^{2} with smooth boundary. We first show that the Levi form λ\lambda of Ω\Omega near pp satisfies

(LL¯λ)(p)|(LLλ)(p)|\displaystyle(L\bar{L}\lambda)(p)\geq|(LL\lambda)(p)| (1.1)

at points pbΩp\in b\Omega where Ω\Omega is weakly pseudoconvex, for all holomorphic tangential vector fields LL near pp, see Theorem 4.10. Note that (1.1) holds trivially at all boundary points at which Ω\Omega is of type larger than 44. Through (1.1), we may then classify type 44 boundary points of Ω\Omega into two groups as follows. We say that Ω\Omega is of strict type 44 at pp if inequality (1.1) is strict, otherwise, call Ω\Omega of weak type 44 at pp. This classification is easily seen to be invariant under biholomorphic coordinate changes, see Lemma 4.14. We then give a description of pseudoconvex domains near strict type 44 boundary points in suitable local holomorphic coordinates in Proposition 4.16. In this form, the notion of strict type 44 boundary points has appeared, albeit implicitly, in the literature. Namely, Kolář showed in [12] that a pseudoconvex domain with real-analytic boundary is convexifiable near any strict type 44 boundary point. He also states in [12] that this result does not hold if the real-analyticity assumption is replaced by mere 𝒞\mathcal{C}^{\infty}-smoothness of the boundary. A proof of this statement is not provided in [12].

We show that if a smoothly bounded, pseudoconvex domain Ω2\Omega\subset\mathbb{C}^{2} is of strict type 44 at some boundary point pp, then there exists a smooth local defining function for Ω\Omega near pp which is plurisubharmonic on bΩb\Omega, see Theorem 5.1. This local defining function is constructed explicitly by solving the following first order partial differential equation on bΩb\Omega near pp. For a given, smooth, complex-valued function FF near pp, find a smooth, real-valued function hh near pp such that

Lh=F+𝒪(λ) on bΩ near p.\displaystyle Lh=F+\mathcal{O}(\sqrt{\lambda})\;\;\text{ on }b\Omega\text{ near }p.

We also prove that if Ω2\Omega\subset\mathbb{C}^{2} admits a smooth defining function which is plurisubharmonic on bΩb\Omega near a type 44 boundary point pp, then Ω\Omega admits a plurisubharmonic smooth defining function, see Theorem 6.19. To construct this defining function, we solve the following system of partial differential equations on bΩb\Omega near pp. For a given, smooth, real-valued function FF near pp, find a smooth, real-valued function hh near pp such that

{Lh=𝒪(λ)LL¯h=F+𝒪(λ) on bΩ near p.\begin{cases}&Lh=\mathcal{O}(\sqrt{\lambda})\\ &L\bar{L}h=F+\mathcal{O}(\sqrt{\lambda})\end{cases}\;\;\text{ on }b\Omega\text{ near }p.

This method of proof fails for higher order boundary points. However, parts of our analysis can be salvaged to give a simplified, short proof of the fact that the Diederich–Fornæss index and Steinness index of Ω2\Omega\subset\mathbb{C}^{2} are 11 if Ω\Omega admits a smooth defining function that is plurisubharmonic on bΩb\Omega, but does not have type 44 boundary points, see Corollary 6.24.

For the second application of our general characterizations of the existence of plurisubharmonic defining functions, see Propositions 5.2 and 6.8, we introduce the notion of sesquiconvexity, a new geometric condition for an open set in 2\mathbb{C}^{2} which may be formulated independently of the choice of a defining function. Sesquiconvexity is sufficient for the existence of (local) defining functions that are plurisubharmonic on the boundary (and inside the domain), see Proposition 7.10 and Corollary 7.15, however, it is not a necessary condition, see Example 7.6.

2. Preliminaries

In this section, we detail our notations and list some known facts for later reference. We note that Section 3 requires almost no prerequisites. As such, readers familiar with basic notions from several complex variables may delay looking through the current section until the study of Section 4 and onward.

The generic term “smooth” always means 𝒞\mathcal{C}^{\infty}-smooth. Domains and functions of finite smoothness classes will only be considered in Section 3.

2.1. Some basic notions from almost complex geometry

Let (M,J)(M,J) be an almost complex manifold. For every pMp\in M, write Tp(M)T_{p}({M}) for the real tangent space to MM at pp. As usual, the complexification Tp(M):=Tp(M)T_{p}({M})^{\scriptscriptstyle\mathbb{C}}\mathrel{\mathop{:}}=T_{p}(M)\otimes\mathbb{C} decomposes as Tp(M)=Tp(M)1,0Tp(M)0,1T_{p}({M})^{\scriptscriptstyle\mathbb{C}}=T_{p}({M})^{1,0}\oplus T_{p}({M})^{0,1} into the +i+i and i-i eigenspaces of JpJ_{p}. Smooth sections over MM in the corresponding bundles T(M)T({M}), T(M)T({M})^{\scriptscriptstyle\mathbb{C}}, T(M)1,0T({M})^{1,0}, T(M)0,1T({M})^{0,1} are denoted by 𝒱(M)\mathcal{V}({M}), 𝒱(M)\mathcal{V}({M})^{\scriptscriptstyle\mathbb{C}}, 𝒱(M)1,0\mathcal{V}({M})^{1,0}, 𝒱(M)0,1\mathcal{V}({M})^{0,1}, and are called, real, complexified, holomorphic and antiholomorphic vector fields on MM, respectively.

Let Tp(M)T^{\ast}_{p}({M}) denote the real cotangent space of MM at pp, and let Jp:Tp(M)Tp(M)J^{\ast}_{p}\colon T^{\ast}_{p}({M})\to T^{\ast}_{p}({M}) be the dual almost complex structure. The complexification Tp(M):=Tp(M)T^{\ast}_{p}({M})_{\scriptscriptstyle\mathbb{C}}\mathrel{\mathop{:}}=T^{\ast}_{p}({M})\otimes\mathbb{C} decomposes as Tp(M)=Tp(M)1,0Tp(M)0,1T^{\ast}_{p}({M})_{\scriptscriptstyle\mathbb{C}}=T^{\ast}_{p}({M})_{1,0}\oplus T^{\ast}_{p}({M})_{0,1} into the +i+i and i-i eigenspaces of JpJ^{\ast}_{p}. Smooth sections over MM in the corresponding bundles T(M)T^{\ast}({M}), T(M)T^{\ast}({M})_{\scriptscriptstyle\mathbb{C}}, T(M)1,0T^{\ast}({M})_{1,0}, T(M)0,1T^{\ast}({M})_{0,1} are denoted by Ω1(M)\Omega^{1}({M}), Ω1(M)\Omega^{1}({M})_{\scriptscriptstyle\mathbb{C}}, Ω1(M)1,0\Omega^{1}({M})_{1,0}, Ω1(M)0,1\Omega^{1}({M})_{0,1}, and are called, real, complexified, holomorphic and antiholomorphic 1-forms on MM, respectively.

For αΩ1(M)\alpha\in\Omega^{1}({M}), write α\alpha_{\scriptscriptstyle\mathbb{C}} to denote the pointwise \mathbb{C}-linear extension of α\alpha to Ω1(M)\Omega^{1}({M})_{\scriptscriptstyle\mathbb{C}}. Then α1,0:=αi(Jα)\alpha_{1,0}\mathrel{\mathop{:}}=\alpha_{\scriptscriptstyle\mathbb{C}}-i(J^{\ast}\alpha)_{\scriptscriptstyle\mathbb{C}} and α0,1:=α+i(Jα)\alpha_{0,1}\mathrel{\mathop{:}}=\alpha_{\scriptscriptstyle\mathbb{C}}+i(J^{\ast}\alpha)_{\scriptscriptstyle\mathbb{C}} define elements in Ω1(M)1,0\Omega^{1}({M})_{1,0} and Ω1(M)0,1\Omega^{1}({M})_{0,1}, respectively. If f:Mf\colon M\to\mathbb{C} is a smooth function, write dfΩ1(M)df\in\Omega^{1}({M}) to denote the differential of ff, and set f:=12(df)1,0\partial f\mathrel{\mathop{:}}=\textstyle\frac{1}{2}(df)_{1,0} and ¯f:=12(df)0,1\bar{\partial}f\mathrel{\mathop{:}}=\textstyle\frac{1}{2}(df)_{0,1}. Clearly, (df)=f+¯f(df)_{\scriptscriptstyle\mathbb{C}}=\partial f+\bar{\partial}f holds.

If MM is an open subset in n\mathbb{C}^{n} with coordinates (z1,,zn)(z^{1},\ldots,z^{n}), zj=xj+iyjz^{j}=x^{j}+iy^{j}, xj,yjx^{j},y^{j}\in\mathbb{R}, and if JJ is the standard almost complex structure, then

f\displaystyle\partial f =j=1nfzjdzj,\displaystyle=\sum_{j=1}^{n}\frac{\partial f}{\partial z^{j}}dz^{j}, ¯f\displaystyle\bar{\partial}f =j=1nfz¯jdz¯j,\displaystyle=\sum_{j=1}^{n}\frac{\partial f}{\partial\bar{z}^{j}}d\bar{z}^{j},

where

zj\displaystyle\frac{\partial}{\partial z^{j}} =12(xjiyj),\displaystyle=\frac{1}{2}\left(\frac{\partial}{\partial x^{j}}-i\frac{\partial}{\partial y^{j}}\right), z¯j\displaystyle\frac{\partial}{\partial\bar{z}^{j}} =12(xj+iyj),\displaystyle=\frac{1}{2}\left(\frac{\partial}{\partial x^{j}}+i\frac{\partial}{\partial y^{j}}\right),
and
dzj\displaystyle dz^{j} =dxj+idyj,\displaystyle=dx^{j}+idy^{j}, dz¯j\displaystyle d\bar{z}^{j} =dxjidyj.\displaystyle=dx^{j}-idy^{j}.

Here, and in what follows, we always suppress the index \mathbb{C} and use the same symbol to denote the vector field V𝒱(M)V\in\mathcal{V}({M}) and its complexification V:=V1𝒱(M)V^{\scriptscriptstyle\mathbb{C}}\mathrel{\mathop{:}}=V\otimes 1\in\mathcal{V}({M})^{\scriptscriptstyle\mathbb{C}}, and the 1-form αΩ1(M)\alpha\in\Omega^{1}({M}) and its complexification αΩ1(M)\alpha_{\scriptscriptstyle\mathbb{C}}\in\Omega^{1}({M})_{\scriptscriptstyle\mathbb{C}}, respectively. Euclidean inner products on 𝒱(M)\mathcal{V}({M}) and Ω1(M)\Omega^{1}({M}) are introduced by declaring that (x1,y1,,xn,yn)(\frac{\partial}{\partial x^{1}},\frac{\partial}{\partial y^{1}},\ldots,\frac{\partial}{\partial x^{n}},\frac{\partial}{\partial y^{n}}) and (dx1,dy1,,dxn,dyn)(dx^{1},dy^{1},\ldots,dx^{n},dy^{n}) are orthonormal bases for 𝒱(M)\mathcal{V}({M}) and Ω1(M)\Omega^{1}({M}), respectively. Similarly, Hermitian inner products on 𝒱(M)\mathcal{V}({M})^{\scriptscriptstyle\mathbb{C}} and Ω1(M)\Omega^{1}({M})_{\scriptscriptstyle\mathbb{C}} are introduced by declaring that (z1,,zn,z¯1,,z¯n)(\frac{\partial}{\partial z^{1}},\ldots,\frac{\partial}{\partial z^{n}},\frac{\partial}{\partial\bar{z}^{1}},\ldots,\frac{\partial}{\partial\bar{z}^{n}}) is an orthogonal basis for 𝒱(M)\mathcal{V}({M})^{\scriptscriptstyle\mathbb{C}} with vectors of constant length 1/21/\sqrt{2}, and (dz1,,dzn,(dz^{1},\ldots,dz^{n}, dz¯1,,dz¯n)d\bar{z}^{1},\ldots,d\bar{z}^{n}) is an orthogonal basis for Ω1(M)\Omega^{1}({M})_{\scriptscriptstyle\mathbb{C}} with vectors of constant length 2\sqrt{2}. Note that the canonical inclusions 𝒱(M)𝒱(M)\mathcal{V}({M})\hookrightarrow\mathcal{V}({M})^{\scriptscriptstyle\mathbb{C}}, VVV\mapsto V^{\scriptscriptstyle\mathbb{C}}, and Ω1(M)Ω1(M)\Omega^{1}({M})\hookrightarrow\Omega^{1}({M})_{\scriptscriptstyle\mathbb{C}}, αα\alpha\mapsto\alpha_{\scriptscriptstyle\mathbb{C}}, are isometries, so the notations |V|\lvert V\rvert and |α|\lvert\alpha\rvert for the corresponding norms are well-defined, even if the index \mathbb{C} is suppressed in the notation.

2.2. Some basic notions from differential geometry

Let MM be a smooth manifold. For every tensor field γ\gamma on MM and every pMp\in M, let γp=γ(p)\gamma_{p}=\gamma(p) denote the value of γ\gamma at pp. If αΩ1(M)\alpha\in\Omega^{1}({M}) and V𝒱(M)V\in\mathcal{V}({M}), then α,V:=α(V)\langle\alpha,V\rangle\mathrel{\mathop{:}}=\alpha(V). Given smooth vector fields V1,,VkV_{1},\ldots,V_{k} on MM and a smooth function f:Mf\colon M\to\mathbb{C}, we write VkV1f:=Vk((V1f))V_{k}\ldots V_{1}f\mathrel{\mathop{:}}=V_{k}(\ldots(V_{1}f)\ldots). Moreover, [V1,V2][V_{1},V_{2}] denotes the Lie bracket of V1V_{1} and V2V_{2}, i.e., [V1,V2]f=V1V2fV2V1f[V_{1},V_{2}]f=V_{1}V_{2}f-V_{2}V_{1}f.

Let MNM\subset\mathbb{R}^{N} be open and let (t1,,tN)(t^{1},\ldots,t^{N}) be the standard Euclidean coordinates on MM. For V,W𝒱(M)V,W\in\mathcal{V}({M}), V=ν=1NVνtνV=\sum_{\nu=1}^{N}V^{\nu}\frac{\partial}{\partial t^{\nu}}, W=ν=1NWνtνW=\sum_{\nu=1}^{N}W^{\nu}\frac{\partial}{\partial t^{\nu}}, we set

VW=ν=1N(μ=1NVμWνtμ)tν.\nabla_{V}W=\sum_{\nu=1}^{N}\bigg{(}\sum_{\mu=1}^{N}V^{\mu}\frac{\partial W^{\nu}}{\partial t^{\mu}}\bigg{)}\frac{\partial}{\partial t^{\nu}}.

Note that VW\nabla_{V}W is precisely the covariant derivative of WW along VV with respect to the Levi-Civita connection in T(M)T({M}) corresponding to the standard Riemannian metric g=ν=1Ndtνdtνg=\sum_{\nu=1}^{N}dt^{\nu}\otimes dt^{\nu} on T(M)T({M}). If for f𝒞(M,)f\in\mathcal{C}^{\infty}(M,\mathbb{R}) we write

Qf(V,W)=ν,μ=1N2ftνtμVνWμ,Q^{\scriptscriptstyle\mathbb{R}}_{f}({V},{W})=\sum_{\nu,\mu=1}^{N}\frac{\partial^{2}f}{\partial t^{\nu}\partial t^{\mu}}V^{\nu}W^{\mu},

then

VWf=Qf(V,W)+(VW)f.\displaystyle VWf=Q^{\scriptscriptstyle\mathbb{R}}_{f}({V},{W})+(\nabla_{V}W)f. (2.1)

Let II\subset\mathbb{R} be an open interval, and let γ:IM\gamma\colon I\to M be a smooth curve. Then γ˙,γ¨:IT(M)\dot{\gamma},\ddot{\gamma}\colon I\to T({M}) denote the vector fields along γ\gamma given by γ˙(τ):=ν=1Nγν(τ)(tj)γ(τ)\dot{\gamma}(\tau)\mathrel{\mathop{:}}=\sum_{\nu=1}^{N}\gamma_{\nu}^{\prime}(\tau)(\frac{\partial}{\partial t_{j}})_{\gamma(\tau)} and γ¨(t):=ν=1Nγν′′(τ)(tj)γ(τ)\ddot{\gamma}(t)\mathrel{\mathop{:}}=\sum_{\nu=1}^{N}\gamma_{\nu}^{\prime\prime}(\tau)(\frac{\partial}{\partial t_{j}})_{\gamma(\tau)}. Observe that γ¨\ddot{\gamma} is the covariant derivative of γ˙\dot{\gamma} with respect to \nabla.

Let MnM\subset\mathbb{C}^{n} be open and let (z1,,zn)(z^{1},\ldots,z^{n}) be the standard Euclidean coordinates on MM. For V,W𝒱(M)1,0V,W\in\mathcal{V}({M})^{1,0}, V=j=1nVjzjV=\sum_{j=1}^{n}V^{j}\frac{\partial}{\partial z^{j}}, W=j=1nWjzjW=\sum_{j=1}^{n}W^{j}\frac{\partial}{\partial z^{j}}, we set

VW=j=1n(k=1nVkWjzk)zj,V¯W¯:=VW¯,V¯W=j=1n(k=1nV¯kWjz¯k)zj,VW¯:=V¯W¯.\begin{split}\nabla_{V}W=\sum_{j=1}^{n}\bigg{(}\sum_{k=1}^{n}V^{k}\frac{\partial W^{j}}{\partial z^{k}}\bigg{)}\frac{\partial}{\partial z^{j}},\quad\nabla_{\bar{V}}\bar{W}\mathrel{\mathop{:}}=\overline{\nabla_{V}W},\\ \nabla_{\bar{V}}W=\sum_{j=1}^{n}\bigg{(}\sum_{k=1}^{n}\bar{V}^{k}\frac{\partial W^{j}}{\partial{\bar{z}}^{k}}\bigg{)}\frac{\partial}{\partial z^{j}},\quad\nabla_{V}\bar{W}\mathrel{\mathop{:}}=\overline{\nabla_{\bar{V}}W}.\end{split}

Note that VW\nabla_{V}W and V¯W\nabla_{\bar{V}}W are precisely the covariant derivatives of WW along VV and V¯\bar{V} with respect to the Chern connection in T(M)1,0T({M})^{1,0} corresponding to the standard Hermitian metric h=j=1ndzjdz¯jh=\sum_{j=1}^{n}dz^{j}\otimes d{\bar{z}}^{j} on T(M)1,0T({M})^{1,0}, respectively. If for f𝒞(M,)f\in\mathcal{C}^{\infty}(M,\mathbb{C}) we write

Qf(V,W)=j,k=1n2fzjzkVjWk,Hf(V,W)=j,k=1n2fzjz¯kVjW¯k,\begin{split}Q^{\scriptscriptstyle\mathbb{C}}_{f}({V},{W})&=\sum_{j,k=1}^{n}\frac{\partial^{2}f}{\partial z^{j}\partial z^{k}}V^{j}W^{k},\\ H^{\scriptscriptstyle\mathbb{C}}_{f}({V},{W})&=\sum_{j,k=1}^{n}\frac{\partial^{2}f}{\partial z^{j}\partial{\bar{z}}^{k}}V^{j}\bar{W}^{k},\end{split}

then

VWf\displaystyle VWf =Qf(V,W)+(VW)f,\displaystyle=Q^{\scriptscriptstyle\mathbb{C}}_{f}({V},{W})+(\nabla_{V}W)f, (2.2)
VW¯f\displaystyle V\bar{W}f =Hf(V,W)+(VW¯)f.\displaystyle=H^{\scriptscriptstyle\mathbb{C}}_{f}({V},{W})+(\nabla_{V}\bar{W})f. (2.3)

Observe that the above formula for Hf(V,W)H^{\scriptscriptstyle\mathbb{C}}_{f}({V},{W}) defines a map

Hf:𝒱(M)1,0×𝒱(M)1,0,Hf(V,W)=¯f(V,W¯).H^{\scriptscriptstyle\mathbb{C}}_{f}\colon\mathcal{V}({M})^{1,0}\times\mathcal{V}({M})^{1,0}\to\mathbb{C},\quad H^{\scriptscriptstyle\mathbb{C}}_{f}(V,W)=\partial\bar{\partial}f(V,\bar{W}).

If ff is real-valued, then HfH^{\scriptscriptstyle\mathbb{C}}_{f} is a sesquilinear form on the 𝒞(M,)\mathcal{C}^{\infty}(M,\mathbb{C})-module 𝒱(M)1,0\mathcal{V}({M})^{1,0}.

The above notations for the Levi-Civita connection and the Chern connection on an open set M2nnM\subset\mathbb{R}^{2n}\simeq\mathbb{C}^{n} are unambiguous in the following sense. If A,B,C,D𝒱(M)A,B,C,D\in\mathcal{V}({M}) such that C+iD𝒱(M)1,0C+iD\in\mathcal{V}({M})^{1,0}, then

A+iB(C+iD)=(AC+iAD)+i(BC+iBD),\nabla_{A+iB}(C+iD)=(\nabla_{A}C+i\nabla_{A}D)+i(\nabla_{B}C+i\nabla_{B}D),

where on the left-hand side \nabla denotes the Chern connection, and on the right-hand side \nabla denotes the Levi-Civita connection.

2.3. Defining functions, pseudoconvexity, and finite type

Let Ωn\Omega\subset\mathbb{C}^{n} be a 𝒞2\mathcal{C}^{2}-smoothly bounded domain and let p0p_{0} in bΩb\Omega. We say that a 𝒞2\mathcal{C}^{2}-smooth function r:Ur\colon U\to\mathbb{R} is a defining function for Ω\Omega, if

  1. (i)

    UU is an open neighborhood of bΩb\Omega,

  2. (ii)

    ΩU={r<0}\Omega\cap U=\{r<0\},

  3. (iii)

    dr0dr\neq 0 on bΩb\Omega.

Moreover, we say that a smooth function r:Ur\colon U\to\mathbb{R} is a local defining function for Ω\Omega (near p0p_{0}), if UU is an open neighborhood of p0p_{0}, and rr satisfies (ii) and (iii).

For every pbΩp\in b\Omega, set Tp(bΩ)1,0:=Tp(n)1,0Tp(bΩ)T_{p}({b\Omega})^{1,0}\mathrel{\mathop{:}}=T_{p}({\mathbb{C}^{n}})^{1,0}\cap T_{p}({b\Omega})^{\scriptscriptstyle\mathbb{C}}. A vector field L𝒱(U)1,0L\in\mathcal{V}({U})^{1,0} is called tangential if LpTp(bΩ)1,0L_{p}\in T_{p}({b\Omega})^{1,0} for every pbΩUp\in b\Omega\cap U. If r:Ur\colon U\to\mathbb{R} is a local defining function for Ω\Omega, then Tp(bΩ)1,0={LpTp(n)1,0:r,L(p)=0}T_{p}({b\Omega})^{1,0}=\{L_{p}\in T_{p}({\mathbb{C}^{n}})^{1,0}:\langle\partial r,L\rangle(p)=0\} for every pbΩUp\in b\Omega\cap U, and LL is tangential if and only if Lr0Lr\equiv 0 on bΩUb\Omega\cap U.

The domain Ω\Omega is called pseudoconvex at pp if Hr(L,L)|bΩU0H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})_{|_{b\Omega\cap U}}\geq 0 near pp for all tangential vector fields L𝒱(U)1,0L\in\mathcal{V}({U})^{1,0} near pp. It is called strictly pseudconvex at pp if in the previous condition the inequality is strict whenever LL is nonvanishing. If Ω\Omega is pseudoconvex at pp but not strictly pseudoconvex at pp, then Ω\Omega is said to be weakly pseudoconvex at pp.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded domain, and let pbΩp\in b\Omega. Let r:Ur\colon U\to\mathbb{R} be a smooth local defining function for Ω\Omega near pp, let LL be a nonvanishing holomorphic tangential vector field near pp, and let λ:=Hr(L,L)|bΩU\lambda\mathrel{\mathop{:}}=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})_{|_{b\Omega\cap U}}. Then the following conditions are equivalent, and independent of the choices of rr and LL.

  1. (1)

    There exists kk\in\mathbb{N} such that

    L1,,Lk{L,L¯}:r,[[[L1,L2],L3],,Lk](p)0,\exists\,L_{1},\ldots,L_{k}\in\{L,\bar{L}\}:\left\langle\partial r,[\ldots[[L_{1},L_{2}],L_{3}],\ldots,L_{k}]\right\rangle(p)\neq 0,

    and kk is the smallest integer with this property.

  2. (2)

    There exists kk\in\mathbb{N} such that

    L1,,Lk2{L,L¯}:(Lk2L1λ)(p)0,\exists\,L_{1},\ldots,L_{k-2}\in\{L,\bar{L}\}:(L_{k-2}\ldots L_{1}\lambda)(p)\neq 0,

    and kk is the smallest integer with this property.

  3. (3)

    There exists kk\in\mathbb{N} and local holomorphic coordinates z,wz,w centered at pp, such that

    r(z,w)=Re(w)+h(z,z¯)+o(|z|k,Im(w)),\displaystyle r(z,w)=\operatorname{Re}(w)+h(z,\bar{z})+o(\lvert z\rvert^{k},\operatorname{Im}(w)), (2.4)

    where hh is a nonvanishing homogeneous polynomial of degree kk without pure terms.

The original definition (1) is given in [11, Definition 2.3]. For the equivalence of (1) and (2), see [11, Proposition 2.8]. The third characterization is implicitly contained in the proof of [11, Lemma 3.16]; see also [2, Theorem 3.3].

If the above properties are satisfied, then bΩb\Omega is said to be of finite type at pp, and the number cp:=cp(bΩ):=kc_{p}\mathrel{\mathop{:}}=c_{p}(b\Omega)\mathrel{\mathop{:}}=k is called the type of bΩb\Omega at pp. If Ω\Omega is pseudoconvex at pp, then cpc_{p} is an even number, see [11, Theorem 3.1], Ω\Omega is strictly pseudoconvex at pp if and only if cp=2c_{p}=2, and Ω\Omega is weakly pseudoconvex at pp if and only if cp4c_{p}\geq 4.

Let Ωn\Omega\subset\mathbb{C}^{n} be a smoothly bounded domain, and let UnU\subset\mathbb{C}^{n} be open. A 𝒞2\mathcal{C}^{2}-smooth function r:Ur\colon U\to\mathbb{R} is called plurisubharmonic if Hr(V,V)0H^{\scriptscriptstyle\mathbb{C}}_{r}({V},{V})\geq 0 for every V𝒱(U)1,0V\in\mathcal{V}({U})^{1,0}, and it is called strictly plurisubharmonic if Hr(V,V)>0H^{\scriptscriptstyle\mathbb{C}}_{r}({V},{V})>0 for every V𝒱(U)1,0V\in\mathcal{V}({U})^{1,0}, V0V\neq 0. Moreover, we say that rr is plurisubharmonic on bΩUb\Omega\cap U if Hr(V,V)|bΩU0H^{\scriptscriptstyle\mathbb{C}}_{r}({V},{V})_{|_{b\Omega\cap U}}\geq 0 for every V𝒱(U)1,0V\in\mathcal{V}({U})^{1,0}, and we say that rr is strictly plurisubharmonic on bΩUb\Omega\cap U if Hr(V,V)|bΩU>0H^{\scriptscriptstyle\mathbb{C}}_{r}({V},{V})_{|_{b\Omega\cap U}}>0 for every V𝒱(U)1,0V\in\mathcal{V}({U})^{1,0}, V0V\neq 0. Note that if pbΩUp\in b\Omega\cap U and rr is plurisubharmonic on bΩUb\Omega\cap U, then in general it does not follow that rr is plurisubharmonic on any open neighborhood UUU^{\prime}\subset U of pp, even if rr is a local defining function for Ω\Omega.

2.4. Canonical vector fields in 2\mathbb{C}^{2}

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded domain, and let r:Ur\colon U\to\mathbb{R} be a smooth local defining function for Ω\Omega. After possibly shrinking UU, we may assume that dr0dr\neq 0 on UU. In this case, define vector fields Lr,Nr𝒱1,0(U)L_{r},N_{r}\in\mathcal{V}^{1,0}(U) by

Lr\displaystyle L_{r} =2|r|(rwzrzw),\displaystyle=\frac{\sqrt{2}}{\lvert\partial r\rvert}\left(r_{w}\frac{\partial}{\partial z}-r_{z}\frac{\partial}{\partial w}\right), (2.5)
Nr\displaystyle N_{r} =2|r|(rz¯z+rw¯w).\displaystyle=\frac{\sqrt{2}}{\lvert\partial r\rvert}\left(r_{\bar{z}}\frac{\partial}{\partial z}+r_{\bar{w}}\frac{\partial}{\partial w}\right). (2.6)

Then

  1. (i)

    (Lr,Nr)(L_{r},N_{r}) is an orthogonal frame for T(U)1,0T({U})^{1,0} such that |Lr|12|Nr|\lvert L_{r}\rvert\equiv\frac{1}{\sqrt{2}}\equiv\lvert N_{r}\rvert,

  2. (ii)

    Lrr0L_{r}r\equiv 0,

  3. (iii)

    Nrr=|r|2N_{r}r=\frac{\lvert\partial r\rvert}{\sqrt{2}}.

Note that, if ρ\rho is some other smooth local defining function for Ω\Omega on UU, then

Lρ\displaystyle L_{\rho} =LrandNρ=Nron bΩU.\displaystyle=L_{r}\quad\text{and}\quad N_{\rho}=N_{r}\quad\text{on }b\Omega\cap U. (2.7)

We will always use the notations L=LrL=L_{r} and N=NrN=N_{r} without explicit reference to the choice of the defining function rr, if we consider these vector fields only on bΩb\Omega.

We will sometimes use the abbreviated notations LL and NN also for the vector fields on the whole open set UU, if it is clear from the context which defining function we are referring to. As an example, given a fixed local defining function for Ω\Omega as above, we introduce X,Y,T,ν𝒱(U)X,Y,T,\nu\in\mathcal{V}({U}) to be the unique real vector fields such that

L=12(X+iY)andN=12(ν+iT).\displaystyle L=\textstyle\frac{1}{2}\left(X+iY\right)\quad\text{and}\quad N=\textstyle\frac{1}{2}\left(\nu+iT\right).

It follows from properties (ii) and (iii) above, that

Xr=Yr=Tr=0andν=gradr|gradr|.\displaystyle Xr=Yr=Tr=0\quad\text{and}\quad\nu=\frac{\operatorname{grad}r}{\lvert\operatorname{grad}r\rvert}. (2.8)

By property (i), and since Y=JXY=-JX and T=JνT=-J\nu, the vector fields X,Y,T,νX,Y,T,\nu are linearly independent at each point qUq\in U. In particular, it follows from (2.7) and (2.8) that the restrictions of X,Y,TX,Y,T to bΩUb\Omega\cap U define a frame for T(bΩU)T({b\Omega\cap U}), which is independent of the choice of rr.

We write r\mathcal{H}^{\scriptscriptstyle\mathbb{C}}_{r} to denote the matrix associated with Hr:𝒱(U)1,0×𝒱(U)1,0H^{\scriptscriptstyle\mathbb{C}}_{r}\colon\mathcal{V}({U})^{1,0}\times\mathcal{V}({U})^{1,0}\to\mathbb{C} relative to the basis (L,N)(L,N), i.e.,

r:=(Hr(L,L)Hr(L,N)Hr(N,L)Hr(N,N)).\displaystyle\mathcal{H}^{\scriptscriptstyle\mathbb{C}}_{r}\mathrel{\mathop{:}}=\begin{pmatrix}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})&H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})\\ H^{\scriptscriptstyle\mathbb{C}}_{r}({N},{L})&H^{\scriptscriptstyle\mathbb{C}}_{r}({N},{N})\end{pmatrix}.

The function rr is plurisubharmonic if and only if r(q)\mathcal{H}^{\scriptscriptstyle\mathbb{C}}_{r}(q) is positive semi-definite at every point qUq\in U.

2.5. Landau symbols

Let MM be a smooth manifold, p0Mp_{0}\in M, and f,g:Mf,g\colon M\to\mathbb{R} be smooth functions. We use the usual notations

f=o(g) for pp0\displaystyle f=o(g)\text{ for }p\to p_{0}\quad :C>0:|f|C|g| in some neighborhood of p0,\displaystyle:\Leftrightarrow\quad\forall\;C>0\;:\lvert f\rvert\leq C\lvert g\rvert\text{ in some neighborhood of }p_{0},
f=𝒪(g) for pp0\displaystyle f=\mathcal{O}(g)\text{ for }p\to p_{0}\quad :C>0:|f|C|g| in some neighborhood of p0.\displaystyle:\Leftrightarrow\quad\exists\;C>0:\lvert f\rvert\leq C\lvert g\rvert\text{ in some neighborhood of }p_{0}.

If it is clear from the context, we usually drop the explicit reference to the point p0p_{0}. Moreover, we write for UMU\subset M open

f=𝒪(g) on U\displaystyle f=\mathcal{O}(g)\text{ on }U\quad :KUC>0:|f|C|g| on K.\displaystyle:\Leftrightarrow\quad\forall\;K\Subset U\,\exists\;C>0:\lvert f\rvert\leq C\lvert g\rvert\text{ on }K.

Roughly speaking, the condition “​f=𝒪(g)f=\mathcal{O}(g) on UU means that ff has at least the same order of vanishing on UU as gg, but it does not contain any information about the growth of ff near the boundary of UU. Similarly, we write

f𝒪(g) on U\displaystyle f\leq\mathcal{O}(g)\text{ on }U\quad :KUC>0:fC|g| on K.\displaystyle:\Leftrightarrow\quad\forall\;K\Subset U\,\exists\;C>0:f\leq C\lvert g\rvert\text{ on }K.
Remark 2.9.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded pseudoconvex domain. If r,ρ:Ur,\rho\colon U\to\mathbb{R} are two local defining functions for Ω\Omega, and if L,LL,L^{\prime} are two nonvanishing tangential holomorphic vector fields on UU, then there exists a nonvanishing smooth function h:Uh\colon U\to\mathbb{R} such that Hr(L,L)|bΩU=hHρ(L,L)|bΩUH^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})_{|_{b\Omega\cap U}}=hH^{\scriptscriptstyle\mathbb{C}}_{\rho}({L^{\prime}},{L^{\prime}})_{|_{b\Omega\cap U}}. In particular, if we write λ:=Hr(L,L)|bΩU\lambda\mathrel{\mathop{:}}=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})_{|_{b\Omega\cap U}}, then the class of smooth functions f:bΩUf\colon b\Omega\cap U\to\mathbb{R} such that, for example,

f=𝒪(λ)on bΩUf=\mathcal{O}(\lambda)\quad\text{on }b\Omega\cap U

is well-defined, i.e., independent of the choice of rr and LL.

2.6. Miscellanea

Let ΩN\Omega\subset\mathbb{R}^{N} be a smoothly bounded domain. Let dbΩ:N[0,)d_{b\Omega}\colon\mathbb{R}^{N}\to[0,\infty),

dbΩ(q):=infpbΩ|qp|,d_{b\Omega}(q)\mathrel{\mathop{:}}=\inf_{p\in b\Omega}\lvert q-p\rvert,

denote the Euclidean distance to the boundary bΩb\Omega, and let δbΩ:N\delta_{b\Omega}\colon\mathbb{R}^{N}\to\mathbb{R},

δbΩ(q):={dbΩ(q), if qNΩdbΩ(q), if qΩ,\delta_{b\Omega}(q)\mathrel{\mathop{:}}=\begin{cases}d_{b\Omega}(q),&\text{ if }q\in\mathbb{R}^{N}\setminus\Omega\\ -d_{b\Omega}(q),&\text{ if }q\in\Omega\end{cases},

be the associated signed distance function. Let UNU\subset\mathbb{R}^{N} be an open neighborhood of bΩb\Omega such that there exists a smooth map π:UbΩ\pi\colon U\to b\Omega with |qπ(q)|=dbΩ(q)\lvert q-\pi(q)\rvert=d_{b\Omega}(q), see, e.g., [5, Lemma 4.11] for existence and [9, Lemma 1 in §15.5] for smoothness of the map π\pi. Finally, let ν\nu denote the outward unit normal vector field along bΩb\Omega, and note that this notation is consistent with the one given in Section 2.4. If f:Uf\colon U\to\mathbb{R} is smooth, then by Taylor’s formula it follows that

f\displaystyle f =fπ+δbΩ((νf)π)+𝒪(dbΩ2) on U.\displaystyle=f\circ\pi+\delta_{b\Omega}((\nu f)\circ\pi)+\mathcal{O}(d_{b\Omega}^{2})\text{ on }U. (2.10)

See also, e.g., (2.1) in [7, 8].

If f:f\colon\mathbb{R}\to\mathbb{R} is a nonnegative 𝒞2\mathcal{C}^{2}-smooth function and f(x0)=0f(x_{0})=0, then it follows readily from L’Hospital’s rule that |f|2Cf\lvert f^{\prime}\rvert^{2}\leq Cf near x0x_{0} for some constant C>0C>0. We will repeatedly need the following generalizations of this simple fact.

Lemma 2.11.

The following assertions hold true.

  1. (1)

    Let UNU\subset\mathbb{R}^{N} be open, and let f:U[0,)f\colon U\to[0,\infty) be a 𝒞2\mathcal{C}^{2}-smooth function. Then for every KUK\Subset U there exists a constant C>0C>0 such that

    |df|2Cf on K.\displaystyle|df|^{2}\leq Cf\text{ on }K. (2.12)
  2. (2)

    Let ΩN\Omega\subset\mathbb{R}^{N} be a smoothly bounded domain, let UNU\subset\mathbb{R}^{N} be open, and let f:bΩU[0,)f\colon b\Omega\cap U\to[0,\infty) be a 𝒞2\mathcal{C}^{2}-smooth function. Then

    Vf=𝒪(f) on bΩU\displaystyle Vf=\mathcal{O}(\sqrt{f})\text{ on }b\Omega\cap U (2.13)

    for every V𝒱(bΩU)V\in\mathcal{V}({b\Omega\cap U}).

Proof.

For a proof of part (1) see, e.g., [8, Lemma 4.3]. In order to prove part (2), let F:U[0,)F\colon U\to[0,\infty) be a 𝒞2\mathcal{C}^{2}-smooth extension of ff, and for given KbΩUK\Subset b\Omega\cap U let C>0C>0 be a constant such that |dF|2CF\lvert dF\rvert^{2}\leq CF on KK, see (2.12). Then |Vf|2=|df,V|2=|dF,V|2|dF|2CF=Cf\lvert Vf\rvert^{2}=\lvert\langle df,V\rangle\rvert^{2}=\lvert\langle dF,V\rangle\rvert^{2}\leq\lvert dF\rvert^{2}\leq CF=Cf on KK for every V𝒱(bΩU)V\in\mathcal{V}({b\Omega\cap U}) such that |V|1\lvert V\rvert\leq 1, which implies (2.13). ∎

3. A counterexample

Consider 2\mathbb{C}^{2} with coordinates (z,w)(z,w), z=x+iyz=x+iy, w=u+ivw=u+iv.

Theorem 3.1.

For fixed kk\in\mathbb{N}, k3k\geq 3, let

r(z,w):=u+1k2|z|2k2(k1)2|z|2k2v+1(k2)2|z|2k4v2+|z|4k2,r(z,w)\mathrel{\mathop{:}}=u+\textstyle\frac{1}{k^{2}}|z|^{2k}-\frac{2}{(k-1)^{2}}|z|^{2k-2}v+\frac{1}{(k-2)^{2}}|z|^{2k-4}v^{2}+|z|^{4k-2},

and set

Ω:={(z,w)2:r(z,w)<0}.\Omega\mathrel{\mathop{:}}=\big{\{}(z,w)\in\mathbb{C}^{2}:r(z,w)<0\big{\}}.

Then the following assertions hold true.

  • (i)

    There exists an open neighborhood U2U\subset\mathbb{C}^{2} of 0bΩ0\in b\Omega, such that ΩU\Omega\cap U is pseudoconvex, and such that the following holds. If k=3k=3, then 0bΩ0\in b\Omega is the only weakly pseudoconvex boundary point of Ω\Omega in bΩUb\Omega\cap U. If k>3k>3, then the set of weakly pseudoconvex boundary points of Ω\Omega in bΩUb\Omega\cap U is (bΩU)({0}×)(b\Omega\cap U)\cap(\{0\}\times\mathbb{C}). Moreover, bΩb\Omega is of finite type c0=2kc_{0}=2k at 0.

  • (ii)

    Let V2V\subset\mathbb{C}^{2} be an open neighborhood of 0 and let ρ:V\rho\colon V\to\mathbb{R} be a 𝒞2\mathcal{C}^{2}-smooth local defining function for Ω\Omega. Then ρ\rho is not plurisubharmonic on bΩVb\Omega\cap V.

Proof.

(i) In a slight deviation from Section 2, set L=rwzrzw.L=r_{w}\textstyle\frac{\partial}{\partial z}-r_{z}\frac{\partial}{\partial w}. Then

Hr(L,L)=rzz¯|rw|22Re[rzw¯rwrz¯]+rww¯|rz|2.\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})=r_{z{\bar{z}}}\lvert r_{w}\rvert^{2}-2\operatorname{Re}[r_{z{\bar{w}}}r_{w}r_{\bar{z}}]+r_{w{\bar{w}}}\lvert r_{z}\rvert^{2}.

Computing the relevant terms, we obtain

rz(z,w)=1kz¯|z|2k22k1z¯|z|2k4v+1k2z¯|z|2k6v2+(2k1)z¯|z|4k4,rw(z,w)=12+i(1(k1)2|z|2k21(k2)2|z|2k4v),rzz¯(z,w)=|z|2k6(|z|2v)2+(2k1)2|z|4k4,rzw¯(z,w)=ik1z¯|z|2k4+ik2z¯|z|2k6v,rww¯(z,w)=12(k2)2|z|2k4.\begin{split}r_{z}(z,w)&=\textstyle\frac{1}{k}{\bar{z}}\lvert z\rvert^{2k-2}-\frac{2}{k-1}{\bar{z}}\lvert z\rvert^{2k-4}v+\frac{1}{k-2}{\bar{z}}\lvert z\rvert^{2k-6}v^{2}+(2k-1){\bar{z}}\lvert z\rvert^{4k-4},\\[2.15277pt] r_{w}(z,w)&=\textstyle\frac{1}{2}+i(\textstyle\frac{1}{(k-1)^{2}}\lvert z\rvert^{2k-2}-\frac{1}{(k-2)^{2}}\lvert z\rvert^{2k-4}v),\\[2.15277pt] r_{z{\bar{z}}}(z,w)&=\textstyle\lvert z\rvert^{2k-6}(\lvert z\rvert^{2}-v)^{2}+(2k-1)^{2}\lvert z\rvert^{4k-4},\\[2.15277pt] r_{z{\bar{w}}}(z,w)&=-\textstyle\frac{i}{k-1}{\bar{z}}\lvert z\rvert^{2k-4}+\frac{i}{k-2}{\bar{z}}\lvert z\rvert^{2k-6}v,\\[2.15277pt] r_{w{\bar{w}}}(z,w)&=\textstyle\frac{1}{2(k-2)^{2}}\lvert z\rvert^{2k-4}.\end{split}

These equations lead straightforwardly to the estimates

(rzz¯|rw|2)(z,w)14[|z|2k6(|z|2v)2+|z|4k4],2Re[rzw¯rwrz¯](z,w)=|z|6k14j=04𝒪(|z|82jvj),(rww¯|rz|2)(z,w)0.\begin{split}(r_{z{\bar{z}}}\lvert r_{w}\rvert^{2})(z,w)&\geq\textstyle\frac{1}{4}\big{[}\lvert z\rvert^{2k-6}(\lvert z\rvert^{2}-v)^{2}+\lvert z\rvert^{4k-4}\big{]},\\ -2\operatorname{Re}[r_{z{\bar{w}}}r_{w}r_{\bar{z}}](z,w)&=\lvert z\rvert^{6k-14}\sum_{j=0}^{4}\mathcal{O}(\lvert z\rvert^{8-2j}v^{j}),\\ (r_{w{\bar{w}}}\lvert r_{z}\rvert^{2})(z,w)&\geq 0.\end{split}

Setting a:=|z|2va\mathrel{\mathop{:}}=\lvert z\rvert^{2}-v, it follows that for |z|\lvert z\rvert and vv sufficiently close to 0

Hr(L,L)(z,w)14(|z|2k6a2+|z|4k4)+|z|6k14j=04𝒪(|z|82jaj)=14|z|2k6[(a2+|z|2k+2)+o(1)(a2+a|z|k+1+|z|2k+2)]18|z|2k6(a2+|z|2k+2).\begin{split}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(z,w)&\geq\textstyle\frac{1}{4}(\lvert z\rvert^{2k-6}a^{2}+\lvert z\rvert^{4k-4})+\lvert z\rvert^{6k-14}\sum_{j=0}^{4}\mathcal{O}(\lvert z\rvert^{8-2j}a^{j})\\ &=\textstyle\frac{1}{4}\lvert z\rvert^{2k-6}\big{[}(a^{2}+\lvert z\rvert^{2k+2})+o(1)(a^{2}+a\lvert z\rvert^{k+1}+\lvert z\rvert^{2k+2})\big{]}\\ &\geq\textstyle\frac{1}{8}\lvert z\rvert^{2k-6}(a^{2}+\lvert z\rvert^{2k+2}).\end{split}

This shows that Ω\Omega is pseudoconvex near 0, and that the set of weakly pseudoconvex points of Ω\Omega has the form as described above. It is clear from (2.4) that c0=2kc_{0}=2k.

(ii) Assume, in order to get a contradiction, that ρ:V\rho\colon V\to\mathbb{R} is a 𝒞2\mathcal{C}^{2}-smooth local defining function for Ω\Omega near 0bΩ0\in b\Omega such that ρ\rho is plurisubharmonic on bΩVb\Omega\cap V. There exists a 𝒞1\mathcal{C}^{1}-smooth function h:Vh\colon V\to\mathbb{R} such that ρ=reh\rho=re^{h}. Thus on bΩVb\Omega\cap V one has (since hδjhh\ast\delta_{j}\to h in the 𝒞1\mathcal{C}^{1}-norm for every Dirac sequence {δj}\{\delta_{j}\})

ρzz¯=(rzz¯+2Re[rzhz¯])eh,ρzw¯=(rzw¯+rzhw¯+rw¯hz)eh.\begin{split}\rho_{z{\bar{z}}}&=\big{(}r_{z{\bar{z}}}+2\operatorname{Re}[r_{z}h_{\bar{z}}]\big{)}e^{h},\\ \rho_{z{\bar{w}}}&=\big{(}r_{z{\bar{w}}}+r_{z}h_{\bar{w}}+r_{\bar{w}}h_{z}\big{)}e^{h}.\end{split} (3.2)

For ε>0\varepsilon>0 sufficiently small, let f:𝔻(0,ε)bΩVf\colon\mathbb{D}(0,\varepsilon)\to b\Omega\cap V be the smooth map

f(ζ):=(ζ,u(|ζ|)+i|ζ|2),u(|ζ|):=(1k22(k1)2+1(k2)2)|ζ|2k|ζ|4k2,\displaystyle f(\zeta)\mathrel{\mathop{:}}=(\zeta,u(|\zeta|)+i\lvert\zeta\rvert^{2}),\,u(|\zeta|)\mathrel{\mathop{:}}=\textstyle-\big{(}\frac{1}{k^{2}}-\frac{2}{(k-1)^{2}}+\frac{1}{(k-2)^{2}}\big{)}\lvert\zeta\rvert^{2k}-\lvert\zeta\rvert^{4k-2}, (3.3)

where 𝔻(0,ε):={ζ:|ζ|<ε}\mathbb{D}(0,\varepsilon)\mathrel{\mathop{:}}=\{\zeta\in\mathbb{C}:\lvert\zeta\rvert<\varepsilon\}. We claim that

(hzf)(ζ)=O(|ζ|2k3).\displaystyle(h_{z}\circ f)(\zeta)=O(\lvert\zeta\rvert^{2k-3}). (3.4)

Indeed, if not, then the number m{0}m\in\mathbb{N}\cup\{0\} such that (hzf)(ζ)=O(|ζ|m)(h_{z}\circ f)(\zeta)=O(\lvert\zeta\rvert^{m}) but (hzf)(ζ)O(|ζ|m+1)(h_{z}\circ f)(\zeta)\neq O(\lvert\zeta\rvert^{m+1}) satisfies m<2k3m<2k-3. Hence, in view of (3.2) and the computations in part (i), we see that (ρzz¯f)(ζ)=O(|ζ|2k1+m)(\rho_{z{\bar{z}}}\circ f)(\zeta)=O(\lvert\zeta\rvert^{2k-1+m}) and (ρzw¯f)(ζ)O(|ζ|m+1)(\rho_{z{\bar{w}}}\circ f)(\zeta)\neq O(\lvert\zeta\rvert^{m+1}). Thus ((ρzz¯ρww¯)f)(ζ)=O(|ζ|2k1+m)((\rho_{z{\bar{z}}}\rho_{w{\bar{w}}})\circ f)(\zeta)=O(\lvert\zeta\rvert^{2k-1+m}) and (|ρzw¯|2f)(ζ)O(|ζ|2m+2)(\lvert\rho_{z{\bar{w}}}\rvert^{2}\circ f)(\zeta)\neq O(\lvert\zeta\rvert^{2m+2}). In view of the inequality 2k1+m>2m+22k-1+m>2m+2, this contradicts the fact that ρ\rho is psh on bΩb\Omega near 0, since this implies that (ρzz¯ρww¯|ρzw¯|2)f0\left(\rho_{z{\bar{z}}}\rho_{w{\bar{w}}}-\lvert\rho_{z{\bar{w}}}\rvert^{2}\right)\circ f\geq 0.

From (3.2), (3.4) and the computations in part (i), we conclude that

(ρzz¯f)(ζ)\displaystyle(\rho_{z{\bar{z}}}\circ f)(\zeta) =O(|ζ|4k4),\displaystyle=O(\lvert\zeta\rvert^{4k-4}),
(ρzw¯f)(ζ)\displaystyle(\rho_{z{\bar{w}}}\circ f)(\zeta) =iμkζ¯|ζ|2k4+12(hzf)(ζ)+O(|ζ|2k2),μk:=1(k1)(k2).\displaystyle=\textstyle i\mu_{k}\bar{\zeta}\lvert\zeta\rvert^{2k-4}+\textstyle\frac{1}{2}(h_{z}\circ f)(\zeta)+O(\lvert\zeta\rvert^{2k-2}),\quad\mu_{k}\mathrel{\mathop{:}}=\frac{1}{(k-1)(k-2)}.

Since (ρzz¯ρww¯|ρzw¯|2)f0\left(\rho_{z{\bar{z}}}\rho_{w{\bar{w}}}-\lvert\rho_{z{\bar{w}}}\rvert^{2}\right)\circ f\geq 0, it follows that

(hzf)(ζ)=2iμkζ¯|ζ|2k4+o(|ζ|2k3).(h_{z}\circ f)(\zeta)=-2i\mu_{k}\bar{\zeta}\lvert\zeta\rvert^{2k-4}+o(\lvert\zeta\rvert^{2k-3}). (3.5)

For given σ>0\sigma>0, define the curve γσ=γ=(γ1,γ2):[0,2π]bΩ\gamma_{\sigma}=\gamma=(\gamma_{1},\gamma_{2})\colon[0,2\pi]\to b\Omega by γ(t):=f(σeit)\gamma(t)\mathrel{\mathop{:}}=f(\sigma e^{it}). Then, for σ>0\sigma>0 sufficiently small, it follows from (3.5) that

γ𝑑h=2Re02πhz(γ(t))γ1˙(t)𝑑t=2Re02π(2iμkσ2k3eit+o(σ2k3))(iσeit)𝑑t=4μk02π(σ2k2+o(σ2k2))𝑑t.\begin{split}\int_{\gamma}dh&=2\operatorname{Re}\int_{0}^{2\pi}h_{z}(\gamma(t))\cdot\dot{\gamma_{1}}(t)\,dt\\ &=2\operatorname{Re}\int_{0}^{2\pi}(-2i\mu_{k}\sigma^{2k-3}e^{-it}+o(\sigma^{2k-3}))\cdot(i\sigma e^{it})\,dt\\ &=4\mu_{k}\int_{0}^{2\pi}(\sigma^{2k-2}+o(\sigma^{2k-2}))\,dt.\end{split}

Thus γ𝑑h0\int_{\gamma}dh\neq 0 whenever σ>0\sigma>0 is sufficiently small. This is a contradiction. ∎

Theorem 3.6.

For fixed kk\in\mathbb{N}, k3k\geq 3, let

r(z,w):=u+1k2|z|2k2(k1)2|z|2k2v+1(k2)2|z|2k4v2+|z|4k2+|w|2,r(z,w)\mathrel{\mathop{:}}=u+\textstyle\frac{1}{k^{2}}|z|^{2k}-\frac{2}{(k-1)^{2}}|z|^{2k-2}v+\frac{1}{(k-2)^{2}}|z|^{2k-4}v^{2}+|z|^{4k-2}+\lvert w\rvert^{2},

and set

Ω:={(z,w)2:r(z,w)<0}.\Omega\mathrel{\mathop{:}}=\{(z,w)\in\mathbb{C}^{2}:r(z,w)<0\}.

Then the following assertions hold true.

  • (i)

    Ω\Omega is a bounded domain with smooth real-analytic boundary.

  • (ii)

    Ω\Omega is pseudoconvex. If k=3k=3, then 0bΩ0\in b\Omega is the only weakly pseudoconvex boundary point of Ω\Omega. If k>3k>3, the set of weakly pseudonconvex boundary points of Ω\Omega is bΩ({0}×)b\Omega\cap(\{0\}\times\mathbb{C}). Moreover, bΩb\Omega is of finite type c0=2kc_{0}=2k at 0.

  • (iii)

    Any 𝒞2\mathcal{C}^{2}-smooth local defining function for Ω\Omega near 0bΩ0\in b\Omega fails to be plurisubharmonic on bΩb\Omega near 0.

Proof.

In order to show (iii), one may proceed exactly as in the proof of part (ii) of Theorem 3.1, after noting that there are two choices for u(|ζ|)u(|\zeta|) in (3.3) to be a solution to r(ζ,u(|ζ|)+i|ζ|2)=0r(\zeta,u(|\zeta|)+i|\zeta|^{2})=0. While the proof of (i) is also straightforward, see below, the difficulty in proving Theorem 3.6 lies in showing that the introduction of the additional term |w|2\lvert w\rvert^{2} in the defining function rr turns Ω\Omega into a globally pseudoconvex domain, which is subject to the precise properties described in (i) and (ii).

In the proofs of (i) and (ii), we repeatedly use the following fact: if A,B,CA,B,C\in\mathbb{R} and A,C0A,C\geq 0, then

ε>0(σ,τ)2:Aσ2+Bστ+Cτ2ε(σ2+τ2)  4ACB2>0.\displaystyle\exists\;\varepsilon>0\;\forall\,(\sigma,\tau)\in\mathbb{R}^{2}:A\sigma^{2}+B\sigma\tau+C\tau^{2}\geq\varepsilon(\sigma^{2}+\tau^{2})\;\;\Leftrightarrow\;\;4AC-B^{2}>0. (3.7)

(i) It follows from (3.7), with σ=|z|2\sigma=\lvert z\rvert^{2} and τ=v\tau=v, that

1k2|z|2k\displaystyle\textstyle\frac{1}{k^{2}}|z|^{2k} 2(k1)2|z|2k2v+1(k2)2|z|2k4v20.\displaystyle-\textstyle\frac{2}{(k-1)^{2}}|z|^{2k-2}v+\frac{1}{(k-2)^{2}}|z|^{2k-4}v^{2}\geq 0.

Thus, for every (z,w)Ω¯(z,w)\in\bar{\Omega}, one has 0r(z,w)u+u20\geq r(z,w)\geq u+u^{2}, and hence u[1,0]u\in[-1,0]. In particular, u+u2[14,0]u+u^{2}\in[-\textstyle\frac{1}{4},0], so that

0\displaystyle 0 r(z,w)u+u2+|z|4k214+|z|4k2,\displaystyle\geq r(z,w)\geq u+u^{2}+|z|^{4k-2}\geq-\textstyle\frac{1}{4}+|z|^{4k-2},
0\displaystyle 0 r(z,w)u+u2+v214+v2.\displaystyle\geq r(z,w)\geq u+u^{2}+v^{2}\geq-\textstyle\frac{1}{4}+v^{2}.

Hence, |z|2412k1|z|^{2}\leq 4^{-\frac{1}{2k-1}} and |v|12|v|\leq\frac{1}{2}. This shows that Ω\Omega is bounded.

To see that bΩb\Omega is smooth, we compute

rz(z,w)\displaystyle r_{z}(z,w) =z¯(1k|z|2k22k1|z|2k4v+1k2|z|2k6v2+(2k1)|z|4k4),\displaystyle={\bar{z}}\left(\textstyle\frac{1}{k}\lvert z\rvert^{2k-2}-\frac{2}{k-1}\lvert z\rvert^{2k-4}v+\frac{1}{k-2}\lvert z\rvert^{2k-6}v^{2}+(2k-1)\lvert z\rvert^{4k-4}\right),
rw(z,w)\displaystyle r_{w}(z,w) =(12+u)+i(1(k1)2|z|2k21(k2)2|z|2k4vv).\displaystyle=\textstyle(\frac{1}{2}+u)+i\left(\textstyle\frac{1}{(k-1)^{2}}\lvert z\rvert^{2k-2}-\frac{1}{(k-2)^{2}}\lvert z\rvert^{2k-4}v-v\right).

It follows from (3.7) that 1k|z|2k22k1|z|2k4v+1k2|z|2k6v20\frac{1}{k}\lvert z\rvert^{2k-2}-\frac{2}{k-1}\lvert z\rvert^{2k-4}v+\frac{1}{k-2}\lvert z\rvert^{2k-6}v^{2}\geq 0, so rz(z,w)0r_{z}(z,w)\neq 0 whenever z0z\neq 0. Moreover, if (0,w)bΩ(0,w)\in b\Omega, then u+u2+v2=0u+u^{2}+v^{2}=0. In this case, either v=0v=0 and u{1,0}u\in\{-1,0\}, and thus ru(0,w)0r_{u}(0,w)\neq 0, or v0v\neq 0, and thus rv(0,w)0r_{v}(0,w)\neq 0.

(ii) Let (z,w)bΩ(z,w)\in b\Omega. As before, write a:=|z|2va\mathrel{\mathop{:}}=\lvert z\rvert^{2}-v. We consider two cases.

Case 1: |z|<110\lvert z\rvert<\frac{1}{10} and |a|<110\lvert a\rvert<\frac{1}{10}. Since then |v|<15\lvert v\rvert<\frac{1}{5}, it follows from r(z,w)=0r(z,w)=0 that

u+u2=1k2|z|2k+2(k1)2|z|2k2v1(k2)2|z|2k4v2|z|4k2v2>316.u+u^{2}=\textstyle-\frac{1}{k^{2}}|z|^{2k}+\frac{2}{(k-1)^{2}}|z|^{2k-2}v-\frac{1}{(k-2)^{2}}|z|^{2k-4}v^{2}-|z|^{4k-2}-v^{2}>-\frac{3}{16}.

Hence u[34,14]u\notin[-\frac{3}{4},-\frac{1}{4}], and thus |ru(z,w)|>14\lvert r_{u}(z,w)\rvert>\frac{1}{4}. In particular,

(rzz¯|rw|2)(z,w)116|z|2k6a2+2516|z|4k4.\displaystyle(r_{z{\bar{z}}}\lvert r_{w}\rvert^{2})(z,w)\geq\textstyle\frac{1}{16}\lvert z\rvert^{2k-6}a^{2}+\frac{25}{16}\lvert z\rvert^{4k-4}. (3.8)

Inserting the equation v=|z|2av=\lvert z\rvert^{2}-a into the formulas for rzr_{z}, rwr_{w} and rzw¯r_{z\bar{w}}, we see that

rz(z,w)\displaystyle r_{z}(z,w) =z¯|z|2k6(2k(k1)(k2)|z|42(k1)(k2)|z|2a+1k2a2+(2k1)|z|2k+2),\displaystyle={\bar{z}}\lvert z\rvert^{2k-6}\left(\textstyle\frac{2}{k(k-1)(k-2)}\lvert z\rvert^{4}-\frac{2}{(k-1)(k-2)}\lvert z\rvert^{2}a+\frac{1}{k-2}a^{2}+(2k-1)\lvert z\rvert^{2k+2}\right),
rw(z,w)\displaystyle r_{w}(z,w) =(12+u)+i(2k3(k1)2(k2)2|z|2k2+1(k2)2|z|2k4a|z|2+a),\displaystyle=(\textstyle\frac{1}{2}+u)+i\left(-\frac{2k-3}{(k-1)^{2}(k-2)^{2}}\lvert z\rvert^{2k-2}+\frac{1}{(k-2)^{2}}\lvert z\rvert^{2k-4}a-\lvert z\rvert^{2}+a\right),
rzw¯(z,w)\displaystyle r_{z\bar{w}}(z,w) =iz¯|z|2k6(1(k1)(k2)|z|21k2a).\displaystyle=i{\bar{z}}\lvert z\rvert^{2k-6}\left(\textstyle\frac{1}{(k-1)(k-2)}\lvert z\rvert^{2}-\frac{1}{k-2}a\right).

Thus, since |z|<1\lvert z\rvert<1 for (z,w)bΩ(z,w)\in b\Omega, we obtain that, for every k3k\geq 3,

|rz(z,w)|\displaystyle\lvert r_{z}(z,w)\rvert |z|2k5((|z|2+|a|)2+(2k1)|z|2k+2),\displaystyle\leq\lvert z\rvert^{2k-5}\left((\lvert z\rvert^{2}+\lvert a\rvert)^{2}+(2k-1)\lvert z\rvert^{2k+2}\right),
|rv(z,w)|\displaystyle\lvert r_{v}(z,w)\rvert 2(|z|2+|a|),\displaystyle\leq 2(\lvert z\rvert^{2}+\lvert a\rvert),
|rzw¯(z,w)|\displaystyle\lvert r_{z\bar{w}}(z,w)\rvert |z|2k5(|z|2+|a|).\displaystyle\leq\lvert z\rvert^{2k-5}(\lvert z\rvert^{2}+\lvert a\rvert).

From |2Re[rzw¯rwrz¯]|=|2Re[rzw¯rvrz¯]|2|rzw¯||rv||rz|\lvert 2\operatorname{Re}[r_{z\bar{w}}r_{w}r_{{\bar{z}}}]\rvert=\lvert 2\operatorname{Re}[r_{z\bar{w}}r_{v}r_{{\bar{z}}}]\rvert\leq 2\lvert r_{z\bar{w}}\rvert\lvert r_{v}\rvert\lvert r_{z}\rvert, it then follows that

|2Re[rzw¯rwrz¯]|(z,w)\displaystyle\left|2\operatorname{Re}[r_{z\bar{w}}r_{w}r_{{\bar{z}}}]\right|(z,w) 4|z|4k10(|z|2+|a|)4+(8k4)|z|6k8(|z|2+|a|)2\displaystyle\leq 4\lvert z\rvert^{4k-10}(\lvert z\rvert^{2}+\lvert a\rvert)^{4}+(8k-4)\lvert z\rvert^{6k-8}(\lvert z\rvert^{2}+\lvert a\rvert)^{2}
=(4|z|2+(8k4)|z|2k4(|z|2+|a|)2)|z|4k4\displaystyle=\left(4\lvert z\rvert^{2}+(8k-4)\lvert z\rvert^{2k-4}(\lvert z\rvert^{2}+\lvert a\rvert)^{2}\right)\cdot\lvert z\rvert^{4k-4}
+16|z|k+1|z|3k5|a|\displaystyle\quad+16\lvert z\rvert^{k+1}\cdot\lvert z\rvert^{3k-5}\lvert a\rvert
+4|z|2k4(6|z|4+4|z|2|a|+|a|2)|z|2k6|a|2.\displaystyle\quad+4\lvert z\rvert^{2k-4}(6\lvert z\rvert^{4}+4\lvert z\rvert^{2}\lvert a\rvert+\lvert a\rvert^{2})\cdot\lvert z\rvert^{2k-6}\lvert a\rvert^{2}.

Since |z|<110\lvert z\rvert<\frac{1}{10} and |a|<110\lvert a\rvert<\frac{1}{10}, this implies that, for every k3k\geq 3,

|2Re[rzw¯rwrz¯]|(z,w)11250|z|2k6|a|2+425|z|3k5|a|+12|z|4k4.\displaystyle\left|2\operatorname{Re}[r_{z\bar{w}}r_{w}r_{{\bar{z}}}]\right|(z,w)\leq\textstyle\frac{11}{250}\lvert z\rvert^{2k-6}\lvert a\rvert^{2}+\frac{4}{25}\lvert z\rvert^{3k-5}\lvert a\rvert+\frac{1}{2}\lvert z\rvert^{4k-4}. (3.9)

From Hr(L,L)rzz¯|rw|2|2Re[rzw¯rwrz¯]|H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})\geq r_{z{\bar{z}}}\lvert r_{w}\rvert^{2}-\lvert 2\operatorname{Re}[r_{z\bar{w}}r_{w}r_{{\bar{z}}}]\rvert, it follows with (3.8) and (3.9) that

Hr(L,L)(z,w)|z|2k6(372000|a|2425|z|k+1|a|+1716|z|2k+2).\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(z,w)\geq\lvert z\rvert^{2k-6}\left(\textstyle\frac{37}{2000}\lvert a\rvert^{2}-\frac{4}{25}\lvert z\rvert^{k+1}\lvert a\rvert+\frac{17}{16}\lvert z\rvert^{2k+2}\right).

Thus, since 43720001716(425)2>04\cdot\frac{37}{2000}\cdot\frac{17}{16}-(\frac{4}{25})^{2}>0, an application of (3.7) shows that

Hr(L,L)(z,w)ε|z|2k6(|a|2+|z|2k+2)\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(z,w)\geq\varepsilon\lvert z\rvert^{2k-6}(\lvert a\rvert^{2}+\lvert z\rvert^{2k+2})

for some constant ε>0\varepsilon>0.

Case 2: |z|>110\lvert z\rvert>\frac{1}{10} or |a|>110\lvert a\rvert>\frac{1}{10}. Set D:=rzz¯rww¯|rzw¯|2D\mathrel{\mathop{:}}=r_{z{\bar{z}}}r_{w{\bar{w}}}-\lvert r_{z{\bar{w}}}\rvert^{2}. Then

D(z,w)|z|2k6|a|2+(2k1)2|z|4k4|z|4k10(k2)2|a|22|z|4k8(k1)(k2)2|a||z|4k6(k1)2(k2)2,\displaystyle D(z,w)\geq\lvert z\rvert^{2k-6}\lvert a\rvert^{2}\!+(2k\!-\!1)^{2}\lvert z\rvert^{4k-4}\!\!-\textstyle\frac{\lvert z\rvert^{4k-10}}{(k-2)^{2}}\lvert a\rvert^{2}\!-\frac{2\lvert z\rvert^{4k-8}}{(k-1)(k-2)^{2}}\lvert a\rvert-\frac{\lvert z\rvert^{4k-6}}{(k-1)^{2}(k-2)^{2}},

so D(z,w)|z|2k6d(z,w)D(z,w)\geq\lvert z\rvert^{2k-6}d(z,w) with

d(z,w):=(1|z|2k4(k2)2)|a|22|z|2k2(k1)(k2)2|a|+((2k1)2|z|2k+2|z|2k(k1)2(k2)2).\displaystyle d(z,w)\mathrel{\mathop{:}}=\textstyle\left(1-\frac{\lvert z\rvert^{2k-4}}{(k-2)^{2}}\right)\lvert a\rvert^{2}-\frac{2\lvert z\rvert^{2k-2}}{(k-1)(k-2)^{2}}\lvert a\rvert+\left((2k-1)^{2}\lvert z\rvert^{2k+2}-\frac{\lvert z\rvert^{2k}}{(k-1)^{2}(k-2)^{2}}\right).

We will show that d(z,w)>0d(z,w)>0. Since, in the currently considered case, we have rzz¯(z,w)c|z|2k6r_{z{\bar{z}}}(z,w)\geq c\lvert z\rvert^{2k-6} with some constant c>0c>0, this implies the claim.

Assume first that |a|>110\lvert a\rvert>\frac{1}{10} and |z|2110\lvert z\rvert^{2}\leq\frac{1}{10}. Then d(z,w)910a21100a14000>0d(z,w)\geq\frac{9}{10}a^{2}-\frac{1}{100}a-\frac{1}{4000}>0. On the other hand, assume now that |z|2>110\lvert z\rvert^{2}>\frac{1}{10}. Then d(z,w)>0d(z,w)>0 provided that

4(1|z|2k4(k2)2)((2k1)2|z|2k+2|z|2k(k1)2(k2)2)4|z|4k4(k1)2(k2)4>0,\displaystyle 4\textstyle\left(1-\frac{\lvert z\rvert^{2k-4}}{(k-2)^{2}}\right)\left((2k-1)^{2}\lvert z\rvert^{2k+2}-\frac{\lvert z\rvert^{2k}}{(k-1)^{2}(k-2)^{2}}\right)-\frac{4\lvert z\rvert^{4k-4}}{(k-1)^{2}(k-2)^{4}}>0,

see (3.7), and this inequality is satisfied if and only if

|z|2k2(k2)2|z|2+1(k1)2(2k1)2<0.\displaystyle\textstyle\lvert z\rvert^{2k-2}-(k-2)^{2}\lvert z\rvert^{2}+\frac{1}{(k-1)^{2}(2k-1)^{2}}<0.

But the left-hand side is negative for |z|2(110,412k1]:=Ik\lvert z\rvert^{2}\in(\frac{1}{10},4^{-\frac{1}{2k-1}}]\mathrel{\mathop{:}}=I_{k}, since the function

fk(t):=tk1(k2)2t+1(k1)2(2k1)2f_{k}(t)\mathrel{\mathop{:}}=t^{k-1}-(k-2)^{2}t+\textstyle\frac{1}{(k-1)^{2}(2k-1)^{2}}

is negative on IkI_{k}: Indeed, for k=3k=3 a straightforward computation shows that f3<0f_{3}<0 on (1265,12+65)I3(\frac{1}{2}-\frac{\sqrt{6}}{5},\frac{1}{2}+\frac{\sqrt{6}}{5})\supset I_{3}, and for k4k\geq 4 note that fk(110)f4(110)<0f_{k}(\frac{1}{10})\leq f_{4}(\frac{1}{10})<0 and fk(1)<0f_{k}(1)<0, so that convexity of fkf_{k} on (0,)(0,\infty) implies fk<0f_{k}<0 on (110,1)Ik(\frac{1}{10},1)\supset I_{k}. ∎

Remark 3.10.

In the case of k>3k>3, the defining function in Theorem 3.1 can be modified in such a way that the origin is the only weakly pseudoconvex boundary point of the modified domain near the origin while maintaining all other properties of (i)-(ii). Similarly, in Theorem 3.6, the defining function can be adapted in such a way that the origin is the only weakly pseudoconvex boundary point of the thereby obtained domain while keeping all other properties of (i)-(iii).

4. Weakly pseudoconvex boundary points of type 4

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain, and p0bΩp_{0}\in b\Omega. Assume that Ω\Omega is weakly pseudoconvex at p0p_{0}. Let r:Ur\colon U\to\mathbb{R} be a smooth local defining function for Ω\Omega on an open neighborhood U2U\subset\mathbb{C}^{2} of p0p_{0}, let LL be a nonvanishing holomorphic tangential vector field on UU, and set

λ:=Hr(L,L)|bΩU.\lambda\mathrel{\mathop{:}}=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})_{|_{b\Omega\cap U}}.

Then λ\lambda attains a local minimum at p0p_{0}. Since LL¯λ=14(XX+YYi[X,Y])λL\bar{L}\lambda=\frac{1}{4}(XX+YY-i[X,Y])\lambda, and since [X,Y][X,Y] is tangential to bΩb\Omega, it follows that (LL¯λ)(p0)(L\bar{L}\lambda)(p_{0}) is real. In this section we will show, in particular, that

(LL¯λ)(p0)|(LLλ)(p0)|.(L\bar{L}\lambda)(p_{0})\geq|(LL\lambda)(p_{0})|. (4.1)
Remark 4.2.

Observe that (4.1) is independent of the choice of rr. Indeed, let ρ:U\rho\colon U\to\mathbb{R} be another smooth local defining function for Ω\Omega. Then there exists a smooth function h>0h>0 on UU such that ρ=rh\rho=rh, and one easily computes that λρ=hλr\lambda_{\rho}=h\lambda_{r}, where λ:=H(L,L)|bΩU\lambda_{\ast}\mathrel{\mathop{:}}=H^{\scriptscriptstyle\mathbb{C}}_{\ast}({L},{L})_{|_{b\Omega\cap U}}. Since λr\lambda_{r} attains a local minimum at p0p_{0}, all tangential derivatives of λr\lambda_{r} at p0p_{0} vanish. It thus follows that all second order tangential derivatives of λr\lambda_{r} and λρ\lambda_{\rho} at p0p_{0} differ by the same constant factor c:=h(p0)>0c\mathrel{\mathop{:}}=h(p_{0})>0. In particular,

(LL¯λρ)(p0)\displaystyle(L\bar{L}\lambda_{\rho})(p_{0}) =c(LL¯λr)(p0),\displaystyle=c(L\bar{L}\lambda_{r})(p_{0}), (4.3)
(LLλρ)(p0)\displaystyle(LL\lambda_{\rho})(p_{0}) =c(LLλr)(p0),\displaystyle=c(LL\lambda_{r})(p_{0}), (4.4)

and thus (4.1) is independent of the local defining function rr.

Note further that (4.1) is also independent of the choice of the holomorphic tangential vector field LL. Namely, for L,LL,L^{\prime} nonvanishing holomorphic tangential vector fields on UU, there exists a nonvanishing complex-valued function hh such that L=hLL^{\prime}=hL on bΩUb\Omega\cap U. Let λ:=Hr(L,L)|bΩU\lambda\mathrel{\mathop{:}}=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})_{|_{b\Omega\cap U}} and λ:=Hr(L,L)|bΩU\lambda^{\prime}\mathrel{\mathop{:}}=H^{\scriptscriptstyle\mathbb{C}}_{r}({L^{\prime}},{L^{\prime}})_{|_{b\Omega\cap U}}. Then, by similar arguments as above, one obtains that

(LL¯λ)(p0)\displaystyle(L^{\prime}\bar{L}^{\prime}\lambda^{\prime})(p_{0}) =|c|4(LL¯λ)(p0),\displaystyle=\lvert c\rvert^{4}(L\bar{L}\lambda)(p_{0}), (4.5)
(LLλ)(p0)\displaystyle(L^{\prime}L^{\prime}\lambda^{\prime})(p_{0}) =c2|c|2(LLλ)(p0),\displaystyle=c^{2}\lvert c\rvert^{2}(LL\lambda)(p_{0}), (4.6)

where c:=h(p0)c\mathrel{\mathop{:}}=h(p_{0}).

The following two Lemmata 4.7 and 4.9 are used to derive (4.1) in Theorem 4.10.

Lemma 4.7.

Let ΩN\Omega\subset\mathbb{R}^{N} be a smoothly bounded domain, p0bΩp_{0}\in b\Omega, and UU an open neighborhood of p0p_{0}. Let VV be a smooth vector field along bΩUb\Omega\cap U. For II\subset\mathbb{R} an open interval containing 0, let γ:IbΩU\gamma:I\longrightarrow b\Omega\cap U be a smooth curve such that γ(0)=p0\gamma(0)=p_{0} and γ˙(τ)=Vγ(τ)\dot{\gamma}(\tau)=V_{\gamma(\tau)} for every τI\tau\in I. Then

(fγ)′′(0)=(VVf)(p0).\displaystyle(f\circ\gamma)^{\prime\prime}(0)=(VVf)(p_{0}). (4.8)

for every smooth function f:bΩUf\colon b\Omega\cap U\to\mathbb{R}.

Proof.

Without loss of generality, assume that ff and VV are defined on UU. Then

(fγ)′′(0)\displaystyle\left(f\circ\gamma\right)^{\prime\prime}(0) =Qf(V,V)(p0)+(df)(p0),γ¨(0).\displaystyle=Q^{\scriptscriptstyle\mathbb{R}}_{f}({V},{V})(p_{0})+\left\langle\left(df\right)(p_{0}),\ddot{\gamma}(0)\right\rangle.

On the other hand, we see from (2.1) that

(VVf)(p0)=Qf(V,V)(p0)+(df)(p0),VV(p0).\displaystyle(VVf)(p_{0})=Q^{\scriptscriptstyle\mathbb{R}}_{f}({V},{V})(p_{0})+\left\langle\left(df\right)(p_{0}),\nabla_{V}V(p_{0})\right\rangle.

Thus, it suffices to show that γ¨(0)=VV(p0)\ddot{\gamma}(0)=\nabla_{V}V(p_{0}). But this is clear, since VV is an extension of γ˙\dot{\gamma}, and since γ¨\ddot{\gamma} is the covariant derivative of γ˙\dot{\gamma} with respect to \nabla. We can also compute this directly as follows. For every ν{1,,N}\nu\in\{1,\ldots,N\} and τI\tau\in I,

γν′′(τ)=(Vνγ)(τ)\displaystyle\gamma_{\nu}^{\prime\prime}(\tau)=(V_{\nu}\circ\gamma)^{\prime}(\tau) =μ=1NVνtμ(γ(τ))γμ(τ)\displaystyle=\sum_{\mu=1}^{N}\frac{\partial V_{\nu}}{\partial t_{\mu}}(\gamma(\tau))\gamma_{\mu}^{\prime}(\tau)
=μ=1N(VμVνtμ)(γ(τ))=V(Vν)(γ(τ)).\displaystyle=\sum_{\mu=1}^{N}\left(V_{\mu}\frac{\partial V_{\nu}}{\partial t_{\mu}}\right)(\gamma(\tau))=V(V_{\nu})(\gamma(\tau)).

Lemma 4.9.

Let ΩN\Omega\subset\mathbb{R}^{N} be a smoothly bounded domain, p0bΩp_{0}\in b\Omega, and UU an open neighborhood of p0p_{0}. Let f:bΩUf\colon b\Omega\cap U\to\mathbb{R} be smooth such that ff attains a local minimum at p0p_{0}. Suppose that V1,,VN1V^{1},\ldots,V^{N-1} are smooth vector fields along bΩUb\Omega\cap U such that (V1,,VN1)p0\smash{\left(V^{1},\ldots,V^{N-1}\right)_{p_{0}}} is a basis for Tp0(bΩ)T_{p_{0}}({b\Omega}). Then the matrix

(V1V1fV1VN1fVN1V1fVN1VN1f)(p0)\begin{pmatrix}V^{1}V^{1}f&\cdots&V^{1}V^{N-1}f\\ \vdots&&\vdots\\ V^{N-1}V^{1}f&\cdots&V^{N-1}V^{N-1}f\end{pmatrix}(p_{0})

is symmetric and positive semi-definite.

Proof.

Consider the function q:Tp0(bΩ)q\colon T_{p_{0}}({b\Omega})\to\mathbb{R} given by

q(Vp0):=(VVf)(p0),q(V_{p_{0}})\mathrel{\mathop{:}}=(VVf)(p_{0}),

where VV is any extension of Vp0V_{p_{0}} to a vector field along bΩUb\Omega\cap U such that VpTp(bΩ)V_{p}\in T_{p}({b\Omega}) for every pbΩUp\in b\Omega\cap U. This is well-defined. To wit, if ¯\bar{\nabla} is any linear connection on bΩb\Omega, then

VVf=Q¯f(V,V)+(¯VV)f,VVf=\bar{Q}_{f}^{\scriptscriptstyle\mathbb{R}}(V,V)+\left(\bar{\nabla}_{V}V\right)f,

where Q¯f=¯2f\bar{Q}_{f}^{\scriptscriptstyle\mathbb{R}}=\bar{\nabla}^{2}f denotes the covariant Hessian. Since p0p_{0} is a local minimum of ff, the derivative (¯VV)f(\bar{\nabla}_{V}V)f vanishes at p0p_{0} independently of the above choice of VV. Furthermore, since Q¯f\bar{Q}_{f}^{\scriptscriptstyle\mathbb{R}} is a tensor, the value Q¯f(V,V)(p0)\bar{Q}_{f}^{\scriptscriptstyle\mathbb{R}}(V,V)(p_{0}) depends only on Vp0V_{p_{0}}.

It follows from (4.8) and an application of the Picard–Lindelöf theorem, that q0q\geq 0. Thus the associated symmetric bilinear form B:Tp0(bΩ)×Tp0(bΩ)B\colon T_{p_{0}}({b\Omega})\times T_{p_{0}}({b\Omega})\to\mathbb{R},

B(Vp0,Wp0):=12(q(Vp0+Wp0)q(Vp0)q(Wp0)),B\!\left(V_{p_{0}},W_{p_{0}}\right)\mathrel{\mathop{:}}=\textstyle\frac{1}{2}\left(q(V_{p_{0}}+W_{p_{0}})-q(V_{p_{0}})-q(W_{p_{0}})\right),

is positive semi-definite. Moreover, note that

B(Vp0j,Vp0k)=(12VjVkf+12VkVjf)(p0)=(VjVkf+12[Vk,Vj]f)(p0).B\!\left(V^{j}_{p_{0}},V^{k}_{p_{0}}\right)=\textstyle\left(\frac{1}{2}V^{j}V^{k}f+\frac{1}{2}V^{k}V^{j}f\right)(p_{0})=\left(V^{j}V^{k}f+\frac{1}{2}[V^{k},V^{j}]f\right)(p_{0}).

Since the vector field [Vk,Vj][V^{k},V^{j}] is tangential to bΩb\Omega, and since ff attains a local minimum at p0p_{0}, it follows that [Vk,Vj]f[V^{k},V^{j}]f vanishes at p0p_{0}. Therefore,

B(Vp0j,Vp0k)=(VjVkf)(p0),\displaystyle B\left(V^{j}_{p_{0}},V^{k}_{p_{0}}\right)=\left(V^{j}V^{k}f\right)(p_{0}),

which proves the claim. ∎

Theorem 4.10.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain. Assume that Ω\Omega is weakly pseudoconvex at p0bΩp_{0}\in b\Omega. Then

(LL¯λ)(p0)|(LLλ)(p0)|.\displaystyle\left(L\bar{L}\lambda\right)(p_{0})\geq|\left(LL\lambda\right)(p_{0})|. (4.11)
Proof.

Let XX, YY and TT be the real vector fields such that

L=12(X+iY)andN=12(ν+iT),\displaystyle L=\textstyle\frac{1}{2}\left(X+iY\right)\quad\text{and}\quad N=\textstyle\frac{1}{2}\left(\nu+iT\right),

and recall that XpX_{p}, YpY_{p}, and TpT_{p} form a basis of Tp(bΩ)T_{p}({b\Omega}) for all pbΩp\in b\Omega, see Section 2.4. Since λ\lambda attains a local minimum at p0p_{0}, it follows from Lemma 4.9 that

(XXλXYλXTλYXλYYλYTλTXλTYλTTλ)(p0)\begin{pmatrix}XX\lambda&XY\lambda&XT\lambda\\ YX\lambda&YY\lambda&YT\lambda\\ TX\lambda&TY\lambda&TT\lambda\end{pmatrix}(p_{0})

is symmetric and positive semi-definite. In particular,

(XXλXYλYXλYYλ)(p0)\displaystyle\begin{pmatrix}XX\lambda&XY\lambda\\ YX\lambda&YY\lambda\end{pmatrix}(p_{0})

is symmetric and positive-semidefinite, i.e.,

(XXλYYλ(XYλ)2)(p0)0.\displaystyle\left(XX\lambda\cdot YY\lambda-(XY\lambda)^{2}\right)(p_{0})\geq 0.

But a straightforward computation shows that

(LL¯λ)2|LLλ|2\displaystyle\left(L\bar{L}\lambda\right)^{2}-|LL\lambda|^{2}
=116((X+iY)(XiY)λ)2116|(X+iY)(X+iY)λ|2\displaystyle=\textstyle\frac{1}{16}\bigl{(}\left(X+iY\right)\left(X-iY\right)\lambda\bigr{)}^{2}-\textstyle\frac{1}{16}\bigl{|}(X+iY)(X+iY)\lambda\bigr{|}^{2}
=116(XXλ+YYλ)2116|XXλYYλ+2iXYλ|2\displaystyle=\textstyle\frac{1}{16}\bigl{(}XX\lambda+YY\lambda\bigr{)}^{2}-\textstyle\frac{1}{16}\bigl{|}XX\lambda-YY\lambda+2iXY\lambda\bigr{|}^{2}
=14(XXλYYλ(XYλ)2).\displaystyle=\textstyle\frac{1}{4}\left(XX\lambda\cdot YY\lambda-(XY\lambda)^{2}\right).

This proves the claim. ∎

Definition 4.12.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain. We say that Ω\Omega is of strict type 4 at p0bΩp_{0}\in b\Omega, if Ω\Omega is weakly pseudoconvex at p0p_{0} and

(LL¯λ)(p0)>|(LLλ)(p0)|.\displaystyle\left(L\bar{L}\lambda\right)(p_{0})>\lvert\left(LL\lambda\right)(p_{0})\rvert. (4.13)

If (4.13) does not hold for p0bΩp_{0}\in b\Omega with cp0=4c_{p_{0}}=4, then we say that Ω\Omega is of weak type 44 at p0p_{0}.

From (4.3) and (4.4), we see that (4.13) is independent of the choice of a local defining function. Therefore, it describes a property of the domain Ω\Omega at p0p_{0}. In the next lemma, it is shown that this property is invariant under biholomorphic transformations.

Lemma 4.14.

Let Ω,Ω2\Omega^{\prime},\Omega\subset\mathbb{C}^{2} be smoothly bounded domains such that Ω\Omega^{\prime} and Ω\Omega are pseudoconvex near p0bΩp_{0}^{\prime}\in b\Omega^{\prime} and p0bΩp_{0}\in b\Omega, respectively. Let Φ:UU\Phi\colon U^{\prime}\to U be a biholomorphic map from an open neighborhood U2U^{\prime}\subset\mathbb{C}^{2} of p0p_{0}^{\prime} to an open neighborhood U2U\subset\mathbb{C}^{2} of p0p_{0} such that Φ(ΩU)=ΩU\Phi(\Omega^{\prime}\cap U^{\prime})=\Omega\cap U and Φ(p0)=p0\Phi(p_{0}^{\prime})=p_{0}. Then p0p_{0}^{\prime} is of strict type 44 for Ω\Omega^{\prime} if and only if p0p_{0} is of strict type 44 for Ω\Omega.

Proof.

Let r:Ur\colon U\to\mathbb{R} be a smooth local defining function for Ω\Omega near p0p_{0}. Let LL^{\prime} be a nonvanishing holomorphic tangential vector field on UU^{\prime}, and let L:=ΦLL\mathrel{\mathop{:}}=\Phi_{\ast}L^{\prime} be the pushforward of LL^{\prime}. Then define

λ=HrΦ(L,L)|bΩUandλ=Hr(L,L)|bΩU,\lambda^{\prime}=H^{\scriptscriptstyle\mathbb{C}}_{r\circ\Phi}({L^{\prime}},{L^{\prime}})_{|_{b\Omega^{\prime}\cap U^{\prime}}}\quad\text{and}\quad\lambda=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})_{|_{b\Omega\cap U}},

and observe that, by the usual transformation law, one has λ=λΦ\lambda^{\prime}=\lambda\circ\Phi. Since, by definition, L(fΦ)=LfL^{\prime}(f\circ\Phi)=Lf for every smooth function f:Uf\colon U\to\mathbb{C}, it thus follows that

(LL¯λ)(p0)=(LL¯λ)(p0)and(LLλ)(p0)=(LLλ)(p0).\displaystyle\left(L^{\prime}\bar{L}^{\prime}\lambda^{\prime}\right)(p_{0}^{\prime})=\left(L\bar{L}\lambda\right)(p_{0})\quad\text{and}\quad\left(L^{\prime}L^{\prime}\lambda^{\prime}\right)(p_{0}^{\prime})=\left(LL\lambda\right)(p_{0}). (4.15)

In view of Remark 4.2, this proves the claim. ∎

In view of Lemma 4.14, it is meaningful to look for local holomorphic coordinates around p0p_{0}, in which the condition (4.13) takes a particularly simple form. The next result shows how this can be achieved.

Proposition 4.16.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain. Assume that 0bΩ0\in b\Omega is a point of weak pseudoconvexity, and let rr be a smooth local defining function for Ω\Omega near 0. If rz(0)=rzz(0)=0r_{z}(0)=r_{zz}(0)=0, then

(LL¯λ)(0)>|(LLλ)(0)|\displaystyle\left(L\bar{L}\lambda\right)(0)>|\left(LL\lambda\right)(0)| (4.17)

if and only if there exists a constant ε>0\varepsilon>0 such that

Hr(L,L)(z,0)ε|z|2+o(|z|2).\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(z,0)\geq\varepsilon|z|^{2}+o(\lvert z\rvert^{2}). (4.18)
Proof.

Let U2U\subset\mathbb{C}^{2} denote the domain of definition of rr, and define Λ:U\Lambda\colon U\to\mathbb{R} by Λ:=Hr(L,L)\Lambda\mathrel{\mathop{:}}=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L}). We will show that both (4.17) and (4.18) are equivalent to the condition that the matrix

A:=(ΛxxΛxyΛyxΛyy)(0)A\mathrel{\mathop{:}}=\begin{pmatrix}\Lambda_{xx}&\Lambda_{xy}\\ \Lambda_{yx}&\Lambda_{yy}\end{pmatrix}(0)

is positive definite, where z=x+iyz=x+iy with x,yx,y\in\mathbb{R}. This proves the claim.

Observe first that Λ|bΩU=λ\Lambda_{|_{b\Omega\cap U}}=\lambda. In particular, Λ(0)=0\Lambda(0)=0. Moreover, since λ\lambda attains a local minimum at 0, it follows that dΛd\Lambda vanishes on T0(bΩ)T_{0}({b\Omega}), and thus Λx(0)=Λy(0)=0\Lambda_{x}(0)=\Lambda_{y}(0)=0. Hence

Λ(z,0)=12(xy)A(xy)+o(|z|2).\Lambda(z,0)=\textstyle\frac{1}{2}\displaystyle\begin{pmatrix}x&y\end{pmatrix}A\begin{pmatrix}x\\ y\end{pmatrix}+o(\lvert z\rvert^{2}).

This shows that (4.18) holds true if and only if A>0A>0.

On the other hand, we claim that rz(0)=rzz(0)=0r_{z}(0)=r_{zz}(0)=0 implies that

(LL¯λ)(0)=Λzz¯(0),(LLλ)(0)=Λzz(0).\begin{split}(L\bar{L}\lambda)(0)=\Lambda_{z\bar{z}}(0),\quad(LL\lambda)(0)=\Lambda_{zz}(0).\end{split} (4.19)

From this, we immediately obtain that

(LL¯λ)(0)\displaystyle(L\bar{L}\lambda)(0) =14(Λxx+Λyy)(0),\displaystyle=\textstyle\frac{1}{4}(\Lambda_{xx}+\Lambda_{yy})(0),
((LL¯Λ)2|LLΛ|2)(0)\displaystyle\left((L\bar{L}\Lambda)^{2}-\lvert LL\Lambda\rvert^{2}\right)(0) =14(ΛxxΛyyΛxy2)(0).\displaystyle=\textstyle\frac{1}{4}\left(\Lambda_{xx}\Lambda_{yy}-\Lambda_{xy}^{2}\right)(0).

Since a real 2×22\times 2 matrix is positive definite if and only if both its trace and determinant are positive, it follows that (4.17)(\ref{E:strictbasictype4estimateagain}) holds true if and only if A>0A>0.

In order to see (4.19) we first note that, in view of Remark 4.2, and after possibly shrinking UU, we can assume without loss of generality that L=LrL=L_{r}. Recall that

LL¯λ\displaystyle L\bar{L}\lambda =HΛ(L,L)+(LL¯)Λ,\displaystyle=H^{\scriptscriptstyle\mathbb{C}}_{\Lambda}({L},{L})+(\nabla_{L}\bar{L})\Lambda,
LLλ\displaystyle LL\lambda =QΛ(L,L)+(LL)Λ.\displaystyle=Q^{\scriptscriptstyle\mathbb{C}}_{\Lambda}({L},{L})+(\nabla_{L}L)\Lambda.

Note that the vectors (LL¯)(0)(\nabla_{L}\bar{L})(0) and (LL)(0)(\nabla_{L}L)(0) are tangential to bΩb\Omega. Indeed, straightforward computations show that, for L=LrL=L_{r} and N=NrN=N_{r},

LL¯\displaystyle\nabla_{L}\bar{L} =1|rz|2+|rw|2(Hr(L,N)L¯Hr(L,L)N¯),\displaystyle=\frac{1}{\lvert r_{z}\rvert^{2}+\lvert r_{w}\rvert^{2}}\Big{(}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})\bar{L}-H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})\bar{N}\Big{)},
LL\displaystyle\nabla_{L}L =1|rz|2+|rw|2(Qr(L,N)LQr(L,L)N).\displaystyle=\frac{1}{\lvert r_{z}\rvert^{2}+\lvert r_{w}\rvert^{2}}\Big{(}Q^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})L-Q^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})N\Big{)}.

Moreover, Hr(L,L)(0)=0H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(0)=0, since Ω\Omega is weakly pseudoconvex at 0, and Qr(L,L)(0)=0Q^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(0)=0, since rz(0)=rzz(0)=0r_{z}(0)=r_{zz}(0)=0, which proves the claim. It follows that both derivatives (L¯L)Λ(\nabla_{\bar{L}}L)\Lambda and (LL)Λ(\nabla_{L}L)\Lambda vanish at 0, since λ\lambda attains a local minimum there. Since the condition rz(0)=0r_{z}(0)=0 implies that L0=(z)0L_{0}=(\partial_{z})_{0}, the claim follows. (For an alternative proof of (4.19) under slightly stronger conditions, see [11, Lemma 3.23].) ∎

Remark 4.20.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain. Then Ω\Omega is of strict type 44 at a weakly pseudoconvex point p0bΩp_{0}\in b\Omega in the sense of Kohn, if

13Re[(LLλ)(p0)]+14(LL¯λ)(p0)>0\displaystyle\textstyle\frac{1}{3}\operatorname{Re}[(LL\lambda)(p_{0})]+\frac{1}{4}(L\bar{L}\lambda)(p_{0})>0 (4.21)

holds for all nonvanishing tangential holomorphic vector fields LL near p0p_{0}, see [11, Definition 2.16] in the case m=3m=3.

If L:=hLL^{\prime}\mathrel{\mathop{:}}=hL for some smooth complex-valued function hh near p0p_{0}, then one has 13Re[(LLλ)(p0)]+14(LL¯λ)(p0)=13Re[h(p0)2(LLλ)(p0)]+14|h(p0)|2(LL¯λ)(p0)\frac{1}{3}\operatorname{Re}[(L^{\prime}L^{\prime}\lambda)(p_{0})]+\frac{1}{4}(L^{\prime}\bar{L}^{\prime}\lambda)(p_{0})=\frac{1}{3}\operatorname{Re}[h(p_{0})^{2}(LL\lambda)(p_{0})]+\frac{1}{4}\lvert h(p_{0})\rvert^{2}(L\bar{L}\lambda)(p_{0}). From this, one easily sees that Ω\Omega is of strict type 4 at p0p_{0} in the sense of Kohn if and only if

(LL¯λ)(p0)>43|(LLλ)(p0)|\displaystyle\left(L\bar{L}\lambda\right)(p_{0})>\textstyle\frac{4}{3}\lvert\left(LL\lambda\right)(p_{0})\rvert (4.22)

for some, and then every, nonvanishing tangential holomorphic vector field LL near p0p_{0}. The condition (4.22) is invariant under biholomorphic transformations by (4.15).

Assume that coordinates are chosen in such a way that p0=0p_{0}=0 and (zjr)(0)=0(\partial_{z}^{j}r)(0)=0 for j{1,,4}j\in\{1,\ldots,4\}. Then (4.22) holds true if and only if there exists a constant ε>0\varepsilon>0 such that

r(z,0)ε|z|4+o(|z|4).\displaystyle r(z,0)\geq\varepsilon|z|^{4}+o(\lvert z\rvert^{4}). (4.23)

Indeed, since (zjz¯kr)(0)=0(\partial_{z}^{j}\partial_{\bar{z}}^{k}r)(0)=0 for j+k2j+k\leq 2, it follows that 0=(Lλ)(0)=λz(0)=Λz(0)=(z2z¯r)(0)0=(L\lambda)(0)=\lambda_{z}(0)=\Lambda_{z}(0)=(\partial_{z}^{2}\partial_{\bar{z}}r)(0), where Λ:=Hr(L,L)\Lambda\mathrel{\mathop{:}}=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L}) and without loss of generality L=LrL=L_{r}. Thus r(z,0)=13Re[(z3z¯r)(0)z3z¯]+14(z2z¯2r)(0)|z|4+o(|z|)4.r(z,0)=\textstyle\frac{1}{3}\operatorname{Re}[(\partial_{z}^{3}\partial_{\bar{z}}r)(0)z^{3}\bar{z}]+\frac{1}{4}(\partial_{z}^{2}\partial_{\bar{z}}^{2}r)(0)\lvert z\rvert^{4}+o(\lvert z\rvert)^{4}. Since (zjz¯kr)(0)=0(\partial_{z}^{j}\partial_{\bar{z}}^{k}r)(0)=0 for j+k3j+k\leq 3, it follows with (4.19) that (LLλ)(0)=(z3z¯r)(0)(LL\lambda)(0)=(\partial_{z}^{3}\partial_{\bar{z}}r)(0) and (L¯Lλ)(0)=(z2z¯2r)(0)(\bar{L}L\lambda)(0)=(\partial_{z}^{2}\partial_{\bar{z}}^{2}r)(0). (The characterization of points of strict type 4 in the sense of Kohn by means of (4.23) is already implicitly contained in Kohn’s original paper, see formulas (3.8) and (3.12) in [11]. See also [2, Theorem 3.3].)

In view of (4.22), the definition of strict type 4 given in Definition 4.12 is more general than the notion of strict type 4 in the sense of Kohn. In particular, if Ω\Omega is of strict type 4 at p0p_{0} in the sense of Kohn, then it is of strict type 4 at p0p_{0} in the sense of Definition 4.12.

Lastly, consider the following example. Let r:2r\colon\mathbb{C}^{2}\to\mathbb{R} be given by

r(z,w):=Rew+Re[az3z¯]+|z|4.r(z,w)\mathrel{\mathop{:}}=\operatorname{Re}w+\operatorname{Re}[az^{3}\bar{z}]+\lvert z\rvert^{4}.

Then Ω:={r<0}\Omega\mathrel{\mathop{:}}=\{r<0\} is pseudoconvex at 0bΩ0\in b\Omega iff |a|43\lvert a\rvert\leq\frac{4}{3}. Moreover, by checking the conditions (4.18) and (4.23), one easily sees that 0 is of strict type 4 in the sense of Definition 4.12 iff |a|<43\lvert a\rvert<\frac{4}{3}, and 0 is of strict type 4 in the sense of Kohn iff |a|<1\lvert a\rvert<1.

5. Plurisubharmonicity on the boundary

The main goal of this section is to prove the following theorem.

Theorem 5.1.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain, and let p0bΩp_{0}\in b\Omega be a point of weak pseudoconvexity for Ω\Omega. If p0p_{0} is of strict type 44, then Ω\Omega admits a smooth local defining function which is plurisubharmonic on bΩb\Omega near p0p_{0}.

We use the following basic lemma to show Theorem 5.1.

Proposition 5.2.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain, p0bΩp_{0}\in b\Omega. Then Ω\Omega admits a smooth local defining function which is plurisubharmonic on bΩb\Omega near p0p_{0} if and only if there exist an open neighborhood UU of p0p_{0} and a smooth local defining function ρ\rho for Ω\Omega on UU such that

|Hρ(L,N)|2=𝒪(Hρ(L,L)) on bΩU.\displaystyle\left|H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})\right|^{2}=\mathcal{O}(H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L}))\text{ on }b\Omega\cap U. (5.3)

In fact, (5.3) holds true for every smooth local defining function ρ:U\rho\colon U\to\mathbb{R} that is plurisubharmonic on bΩUb\Omega\cap U.

Proof.

If ρ:U\rho\colon U\to\mathbb{R} is plurisubharmonic on bΩUb\Omega\cap U, then

ρ=(Hρ(L,L)Hρ(L,N)Hρ(N,L)Hρ(N,N))\displaystyle\mathcal{H}^{\scriptscriptstyle\mathbb{C}}_{\rho}=\begin{pmatrix}H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})&H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})\\ H^{\scriptscriptstyle\mathbb{C}}_{\rho}({N},{L})&H^{\scriptscriptstyle\mathbb{C}}_{\rho}({N},{N})\end{pmatrix}

is positive semi-definite at every point pbΩUp\in b\Omega\cap U. In particular, detρ0\det\mathcal{H}^{\scriptscriptstyle\mathbb{C}}_{\rho}\geq 0 on bΩUb\Omega\cap U, which implies (5.3).

On the other hand, suppose that (5.3) holds for some smooth local defining function ρ:U\rho\colon U\to\mathbb{R} of Ω\Omega near p0p_{0}. Let χ:\chi\colon\mathbb{R}\to\mathbb{R} be a smooth function such that χ(0)=0\chi(0)=0, χ(0)=1\chi^{\prime}(0)=1, and write χ′′(0)=:C\chi^{\prime\prime}(0)=\mathrel{\mathop{:}}C. If ρ^:=χρ\hat{\rho}\mathrel{\mathop{:}}=\chi\circ\rho, then we get

ρ^=(Hρ(L,L)Hρ(L,N)Hρ(N,L)Hρ(N,N)+C|Nρ|2)on bΩU.\displaystyle\mathcal{H}^{\scriptscriptstyle\mathbb{C}}_{\hat{\rho}}=\begin{pmatrix}H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})&H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})\\ H^{\scriptscriptstyle\mathbb{C}}_{\rho}({N},{L})&H^{\scriptscriptstyle\mathbb{C}}_{\rho}({N},{N})+C\lvert N\rho\rvert^{2}\end{pmatrix}\quad\text{on }b\Omega\cap U.

Let VUV\Subset U be another open neighborhood of p0p_{0}. Note that |Nρ|=12|dρ|>c>0\lvert N\rho\rvert=\frac{1}{2}\lvert d\rho\rvert>c>0 on bΩVb\Omega\cap V for some cc\in\mathbb{R}. Thus, if C>0C>0 is sufficiently large, it follows that trρ^>0\operatorname{tr}\mathcal{H}^{\scriptscriptstyle\mathbb{C}}_{\hat{\rho}}>0 and detρ^0\det\mathcal{H}^{\scriptscriptstyle\mathbb{C}}_{\hat{\rho}}\geq 0 on bΩVb\Omega\cap V, where for the second inequality we use assumption (5.3). This means that ρ^0\mathcal{H}^{\scriptscriptstyle\mathbb{C}}_{\hat{\rho}}\geq 0 on bΩVb\Omega\cap V, i.e., ρ^\hat{\rho} is plurisubharmonic on bΩVb\Omega\cap V. ∎

Remark 5.4.

The proof of Proposition 5.2 shows, in particular, the following: if (5.3) holds true, then for every VUV\Subset U there exists a smooth local defining function for Ω\Omega on UU that is plurisubharmonic on bΩVb\Omega\cap V.

Note that (5.3) may be reformulated as

|Hρ(L,N)|=𝒪(λ) on bΩU,\left|H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})\right|=\mathcal{O}(\sqrt{\lambda})\text{ on }b\Omega\cap U,

where λ=Hr(L,L)|bΩU\lambda=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})_{|_{b\Omega\cap U}} for any smooth local defining function r:Ur\colon U\to\mathbb{R} for Ω\Omega and any nonvanishing holomorphic tangential vector field LL on UU, see Remark 2.9.

Proof of Theorem 5.1.

Let r:Ur\colon U\to\mathbb{R} be a smooth local defining function for Ω\Omega near p0p_{0}, and set λ:=Hr(L,L)|bΩU\lambda\mathrel{\mathop{:}}=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})_{|_{b\Omega\cap U}}. Since Ω\Omega is of strict type 44 at p0p_{0}, after possibly shrinking UU, we may assume that B:=|LL¯λ|2|LLλ|2>0B\mathrel{\mathop{:}}=|L\bar{L}\lambda|^{2}-|LL\lambda|^{2}>0 on UU. We claim that, for every smooth function F:UF\colon U\to\mathbb{C}, there exists a smooth function h:Uh\colon U\to\mathbb{R} such that

Lh=F+𝒪(λ)on bΩU.\displaystyle Lh=F+\mathcal{O}(\sqrt{\lambda})\quad\text{on }b\Omega\cap U. (5.5)

Indeed, define smooth functions A1,A2:UA_{1},A_{2}\colon U\to\mathbb{R} by

A1:=2Re((L¯LλLLλ)L¯λ)andA2:=2Re(i(L¯Lλ+LLλ)L¯λ),\displaystyle A_{1}\mathrel{\mathop{:}}=2\operatorname{Re}\left((\bar{L}L\lambda-LL\lambda)\bar{L}\lambda\right)\quad\text{and}\quad A_{2}\mathrel{\mathop{:}}=2\operatorname{Re}\left(i(\bar{L}L\lambda+LL\lambda)\bar{L}\lambda\right),

and set

h:=Re(F)A1B+Im(F)A2B.\displaystyle h\mathrel{\mathop{:}}=\operatorname{Re}(F)\frac{A_{1}}{B}+\operatorname{Im}(F)\frac{A_{2}}{B}. (5.6)

By Lemma 2.11, both LλL\lambda and L¯λ\bar{L}\lambda are of class 𝒪(λ)\mathcal{O}(\sqrt{\lambda}). Thus

Lh=Re(F)LA1B+Im(F)LA2B+𝒪(λ)on bΩU,\displaystyle Lh=\operatorname{Re}(F)\frac{LA_{1}}{B}+\operatorname{Im}(F)\frac{LA_{2}}{B}+\mathcal{O}(\sqrt{\lambda})\quad\text{on }b\Omega\cap U,

and, again on bΩUb\Omega\cap U,

LA1\displaystyle LA_{1} =(L¯LλLLλ)LL¯λ+(LL¯λL¯L¯λ)LLλ+𝒪(λ)=B+𝒪(λ),\displaystyle=(\bar{L}L\lambda-LL\lambda)L\bar{L}\lambda+(L\bar{L}\lambda-\bar{L}\bar{L}\lambda)LL\lambda+\mathcal{O}(\sqrt{\lambda})=B+\mathcal{O}(\sqrt{\lambda}),
LA2\displaystyle LA_{2} =i(L¯Lλ+LLλ)LL¯λi(LL¯λ+L¯L¯λ)LLλ+𝒪(λ)=iB+𝒪(λ).\displaystyle=i(\bar{L}L\lambda+LL\lambda)L\bar{L}\lambda-i(L\bar{L}\lambda+\bar{L}\bar{L}\lambda)LL\lambda+\mathcal{O}(\sqrt{\lambda})=iB+\mathcal{O}(\sqrt{\lambda}).

Hence hh is a solution to (5.5).

Now let h:Uh\colon U\to\mathbb{R} be an arbitrary smooth function, and set ρ:=reh\rho\mathrel{\mathop{:}}=re^{h}. Then

Hρ(L,N)=eh(Hr(L,N)+LhN¯r)on bΩU.\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})=e^{h}\left(H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})+Lh\cdot\bar{N}r\right)\quad\text{on }b\Omega\cap U.

Since N¯r=12|dr|0\bar{N}r=\frac{1}{2}\lvert dr\rvert\neq 0 on bΩUb\Omega\cap U, it follows that (5.3) is satisfied if hh is the solution for (5.5) with F:=2|dr|Hr(L,N)F\mathrel{\mathop{:}}=-\frac{2}{\lvert dr\rvert}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N}). Thus, the claim follows from Proposition 5.2. ∎

Remark 5.7.

The functions A1A_{1}, A2A_{2}, and BB in the proof of Theorem 5.1 depend on the given smooth local defining function r:Ur\colon U\to\mathbb{R}. However, the function hh defined in (5.6) solves (5.5) for any choice of rr.

A global version of Theorem 5.1 easily follows.

Corollary 5.8.

Let Ω2\Omega\Subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain. Assume that all weakly pseudoconvex boundary points of bΩb\Omega are of strict type 44. Then Ω\Omega admits a smooth defining function which is plurisubharmonic on bΩb\Omega.

Proof.

Let r:Vr\colon V\to\mathbb{R} be a smooth defining function for Ω\Omega. Let 𝒲bΩ\mathcal{W}\subset b\Omega denote the set of points at which Ω\Omega is weakly pseudoconvex. Then 𝒲\mathcal{W} is closed in bΩb\Omega. Further, it follows from the hypothesis that B=|LL¯λ|2|LLλ|2B=|L\bar{L}\lambda|^{2}-|LL\lambda|^{2} is strictly positive on some open neighborhood UVU\Subset V of 𝒲\mathcal{W}. As in the proof of Theorem 5.1, we find a smooth function h:Uh\colon U\to\mathbb{R} such that ρ:=reh\rho\mathrel{\mathop{:}}=re^{h} satisfies (5.3) on bΩUb\Omega\cap U.

Let UUU^{\prime}\Subset U be another open neighborhood of 𝒲\mathcal{W}, and let χ\chi be a real-valued, smooth function which is compactly supported in UU and identically 11 on UU^{\prime}. Then ρ~:=reχh\tilde{\rho}:=re^{\chi\cdot h} is a smooth defining function for Ω\Omega such that ρ~=ρ\tilde{\rho}=\rho on UU^{\prime}, hence it satisfies (5.3) on bΩUb\Omega\cap U^{\prime}. Now note that bΩUb\Omega\setminus U^{\prime} is a compact set at whose points Ω\Omega is strictly pseudoconvex. Hence, (5.3) is satisfied on bΩUb\Omega\setminus U^{\prime} by any defining function for Ω\Omega, in particular by ρ~\tilde{\rho}. Thus, ρ~\tilde{\rho} satisfies (5.3) on all of bΩb\Omega. It now follows from Proposition 5.2 that Ω\Omega admits a defining function which is plurisubharmonic on bΩb\Omega. ∎

In the following, we give two examples of smoothly bounded, pseudoconvex domains which admit a plurisubharmonic defining function on the boundary, although they have boundary points of weak type 44.

Example 5.9.

For (z,w)2(z,w)\in\mathbb{C}^{2}, write z=x+iyz=x+iy and w=u+ivw=u+iv. Then define Ω={(z,w)2:r(z,w)<0}\Omega=\{(z,w)\in\mathbb{C}^{2}:r(z,w)<0\} with r(z,w):=u+f(z)r(z,w):=u+f(z) for some smooth, subharmonic function ff. It follows that for all 𝒱(2)1,0\in\mathcal{V}({\mathbb{C}^{2}})^{1,0}, V=V1z+V2wV=V^{1}\frac{\partial}{\partial z}+V^{2}\frac{\partial}{\partial w},

Hr(V,V)=fzz¯|V1|2.H^{\scriptscriptstyle\mathbb{C}}_{r}({V},{V})=f_{z\bar{z}}|V^{1}|^{2}.

Hence, rr is a plurisubharmonic defining function for Ω\Omega, independent of the type of bΩb\Omega at any of its boundary points. In particular, bΩb\Omega may be of weak type 44 at some boundary point p0p_{0}, e.g., if f(z)=x4f(z)=x^{4} and p0=0p_{0}=0.

Example 5.10.

As in the previous example, write z=x+iyz=x+iy and w=u+ivw=u+iv. Set

U={(z,w)2:|x|<π/2},U=\{(z,w)\in\mathbb{C}^{2}:|x|<\pi/2\},

and define Ω={(z,w)U:r(z,w)<0}\Omega=\{(z,w)\in U:r(z,w)<0\} for

r(z,w)=u12(xv)2ln(cos(x)).r(z,w)=u-\textstyle\frac{1}{2}\left(x-v\right)^{2}-\ln(\cos(x)).

We compute that, for every (z,w)U(z,w)\in U,

rz(z,w)\displaystyle r_{z}(z,w) =12(tan(x)(xv)),\displaystyle=\textstyle\frac{1}{2}\left(\tan(x)-\left(x-v\right)\right),
rw(z,w)\displaystyle r_{w}(z,w) =12(1i(xv))\displaystyle=\textstyle\frac{1}{2}(1-i(x-v))
rzz¯(z,w)\displaystyle r_{z\bar{z}}(z,w) =14tan2(x),rzw¯(z,w)=i4,rww¯(z,w)=14,\displaystyle=\textstyle\frac{1}{4}\tan^{2}(x),\;r_{z\bar{w}}(z,w)=\textstyle\frac{i}{4},\;r_{w\bar{w}}(z,w)=-\textstyle\frac{1}{4},

so that

(rzz¯|rw|2)(z,w)\displaystyle(r_{z\bar{z}}|r_{w}|^{2})(z,w) =116tan2(x)(1+(xv)2),\displaystyle=\textstyle{\frac{1}{16}}\tan^{2}(x)(1+(x-v)^{2}),
(rww¯|rz|2)(z,w)\displaystyle(r_{w\bar{w}}|r_{z}|^{2})(z,w) =116(tan(x)(xv))2,\displaystyle=-\textstyle{\frac{1}{16}}\left(\tan(x)-(x-v)\right)^{2},
2Re[rzw¯rwrz¯](z,w)\displaystyle-2\operatorname{Re}[r_{z\bar{w}}r_{w}r_{\bar{z}}](z,w) =18(xv)tan(x)+18(xv)2.\displaystyle=-\textstyle{\frac{1}{8}}(x-v)\tan(x)+\textstyle{\frac{1}{8}}\left(x-v\right)^{2}.

Hence, for L=rwzrzwL=r_{w}\frac{\partial}{\partial z}-r_{z}\frac{\partial}{\partial w}, we obtain

Hr(L,L)(z,w)=116(xv)2sec2(x)0,\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(z,w)=\textstyle{\frac{1}{16}}(x-v)^{2}\sec^{2}(x)\geq 0,

that is, Ω\Omega is pseudoconvex. In fact, Ω\Omega is strictly pseudoconvex except at boundary points satisfying x=vx=v. Moreover, since ln(cos(x))=12x2+112x4+o(x4)-\ln(\cos(x))=\frac{1}{2}x^{2}+\frac{1}{12}x^{4}+o(x^{4}), it follows from (2.4) and Proposition 4.16 that Ω\Omega is of weak type 44 at the origin. In particular, the function hh constructed in the proof of Theorem 5.1, see (5.6), is not defined at the origin. However, a straightforward computation, see Example 7.6 in Section 7, shows that ρ(z,w):=r(z,w)eycosx\rho(z,w):=r(z,w)e^{y}\cos x satisfies (5.3), i.e., for every VUV\Subset U there exists a function ρ^:U\hat{\rho}\colon U\to\mathbb{R} such that ρ^\hat{\rho} is plurisubharmonic on bΩVb\Omega\cap V, see Remark 5.4.

We show in the following that any smooth function hh such that rehre^{h} is plurisubharmonic on bΩb\Omega near the origin must have nonvanishing derivative with respect to yy at the origin, although rr is independent of yy. This is noteworthy because it shows that, in the case of the domain being of weak type 44 at some boundary point, the multiplier function hh, if it exists, can in general not be given as a combination of derivatives of rr as in the strict type 44 case.

Now, suppose hh is a positive, smooth function near the origin such that hy(0)=0h_{y}(0)=0. A straightforward computation, with V=z+iswV=-\frac{\partial}{\partial z}+is\frac{\partial}{\partial w} for s>0s>0, yields

Hr(V,V)(0)\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({V},{V})(0) =12s+𝒪(s2).\displaystyle=-\textstyle\frac{1}{2}s+\mathcal{O}(s^{2}).

Moreover,

(Vr)(0)\displaystyle(Vr)(0) =i2s, and\displaystyle=\textstyle\frac{i}{2}s,\text{ and}
(Vh)(0)\displaystyle(Vh)(0) =12hx(0)+ishw(0).\displaystyle=-\textstyle\frac{1}{2}h_{x}(0)+ish_{w}(0).

Therefore,

2Re(VhV¯r)(0)=𝒪(s2).\displaystyle 2\operatorname{Re}\left(Vh\cdot\bar{V}r\right)(0)=\mathcal{O}(s^{2}).

It then follows that

Hehr(V,V)(0)=eh(0)(Hr(V,V)+2Re(VhV¯r))(0)=12s+𝒪(s2)<0\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{e^{h}r}({V},{V})(0)=e^{h(0)}\Bigl{(}H^{\scriptscriptstyle\mathbb{C}}_{r}({V},{V})+2\operatorname{Re}\left(Vh\cdot\bar{V}r\right)\Bigr{)}(0)=-\textstyle{\frac{1}{2}}s+\mathcal{O}(s^{2})<0

for all s>0s>0 sufficiently close to 0.

6. Plurisubharmonicity near the boundary

In this section, we first consider smoothly bounded, pseudoconvex domains in 2\mathbb{C}^{2} such that all weakly pseudoconvex boundary points are of type 44. In the case of bounded domains, we show that there exists a smooth, plurisubharmonic defining function for the domain whenever there is a smooth defining function which is plurisubharmonic on the boundary of the domain.

In the latter part of this section, we consider smoothly bounded, pseudoconvex domains in 2\mathbb{C}^{2} that are at least of type 6 at their weakly pseudoconvex boundary points. In the case that such a domain admits a smooth defining function which is plurisubharmonic on the boundary, we give a simplified proof that both the Diederich–Fornæss index and the Steinness index are 11.

A lack of understanding of the notion of existence of a plurisubharmonic defining function is rooted in the lack of an equivalent condition which is checkable for any defining function. The following lemma yields a condition which is checkable on a class of defining functions strictly larger than the class of plurisubharmonic defining functions. A version of this lemma in the context of convex domains is given in [10, Proposition 6.17].

Lemma 6.1.

Let Ωn\Omega\subset\mathbb{C}^{n} be a smoothly bounded, pseudoconvex domain, p0bΩp_{0}\in b\Omega. Then Ω\Omega admits a smooth local defining function which is plurisubharmonic near p0p_{0} if and only if there exists a smooth local defining function rr for Ω\Omega near p0p_{0} such that

Hr(ξ,ξ)C(r2|ξ|2+|r,ξ|2)ξn\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({\xi},{\xi})\geq-C\left(r^{2}|\xi|^{2}+\left|\langle\partial r,\xi\rangle\right|^{2}\right)\;\;\hskip 5.69046pt\forall\;\xi\in\mathbb{C}^{n} (6.2)

for some constant C>0C>0.111Here, and occasionally later on, we consider the complex Hessian form HrH^{\scriptscriptstyle\mathbb{C}}_{r} at a point as a sesquilinear form on n\mathbb{C}^{n}.

To put (6.2) in context, we recall that for any smoothly bounded, pseudoconvex domain Ωn\Omega\Subset\mathbb{C}^{n} with smooth defining function rr, there exist an open neighborhood UU of the boundary of Ω\Omega and a constant C>0C>0 such that

Hr(ξ,ξ)C(|r||ξ|2+|r,ξ||ξ|)ξn\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({\xi},{\xi})\geq-C\left(|r|\cdot|\xi|^{2}+\left|\langle\partial r,\xi\rangle\right|\cdot|\xi|\right)\;\;\hskip 5.69046pt\forall\xi\in\mathbb{C}^{n} (6.3)

on UU. That (6.3) holds true on ΩU\Omega\cap U is derived by Range, see (5) in [15], to reprove the result of Diederich–Fornæss [3, Theorem 1] on the existence of bounded, strictly plurisubharmonic exhaustion functions for smoothly bounded, pseudoconvex domains, see [15, Theorem 2]. Arguments similar to the ones in [15] yield (6.3) on UU. We note that, if for every ε>0\varepsilon>0, there exists a smooth defining function r=rεr=r_{\varepsilon} such that (6.3) holds with C=εC=\varepsilon, then both the Diederich–Fornæss index and the Steinness index are 11, see the proof of Corollary 1.6 in [8].

Proof of Lemma 6.1.

Note first that if rr is a smooth local defining function for Ω\Omega which is plurisubharmonic on an open neighborhood UU of p0p_{0}, then (6.2) holds trivially for rr on UU.

Let p0bΩp_{0}\in b\Omega and let UnU\Subset\mathbb{C}^{n} be an open neighborhood of p0p_{0}. Now suppose that r:Ur:U\longrightarrow\mathbb{R} is a smooth local defining function for Ω\Omega on UU such that (6.2) holds. Consider ρ:=r+r2ψ\rho:=r+r^{2}\psi with ψ(z):=K1+K2|z|2\psi(z):=K_{1}+K_{2}|z|^{2} for fixed, positive constants K1K_{1} and K2K_{2} to be determined later. It follows from a straightforward computation that

Hρ(ξ,ξ)=(1+2rψ)Hr(ξ,ξ)+2ψ|r,ξ|2+4K2rRe(r,ξz,ξ)+r2K2|ξ|2\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{\rho}({\xi},{\xi})=(1+2r\psi)H^{\scriptscriptstyle\mathbb{C}}_{r}({\xi},{\xi})+2\psi|\langle\partial r,\xi\rangle|^{2}+4K_{2}r\operatorname{Re}\left(\langle\partial r,\xi\rangle\langle z,\xi\rangle\right)+r^{2}K_{2}|\xi|^{2}

for ξn\xi\in\mathbb{C}^{n}. Next, it follows from

2abεa2+1εb2 for a,b0 and ε>0,\displaystyle 2ab\leq\varepsilon a^{2}+\frac{1}{\varepsilon}b^{2}\;\;\text{ for }a,b\geq 0\text{ and }\varepsilon>0, (6.4)

with a=|r||ξ|a=\lvert r\rvert\lvert\xi\rvert, b=|z||r,ξ|b=\lvert z\rvert\lvert\langle\partial r,\xi\rangle\rvert and ε=14\varepsilon=\frac{1}{4}, that

4K2|rRe(r,ξz,ξ)|r2K22|ξ|2+8K2|z|2|r,ξ|2\displaystyle 4K_{2}\left|r\operatorname{Re}\bigl{(}\langle\partial r,\xi\rangle\langle z,\xi\rangle\bigr{)}\right|\leq r^{2}\frac{K_{2}}{2}|\xi|^{2}+8K_{2}|z|^{2}|\langle\partial r,\xi\rangle|^{2}

holds. Therefore, we obtain

Hρ(ξ,ξ)(1+2rψ)Hr(ξ,ξ)+2|r,ξ|2(K13K2|z|2)+r2K22|ξ|2.\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{\rho}({\xi},{\xi})\geq(1+2r\psi)H^{\scriptscriptstyle\mathbb{C}}_{r}({\xi},{\xi})+2\left|\langle\partial r,\xi\rangle\right|^{2}\left(K_{1}-3K_{2}|z|^{2}\right)+r^{2}\frac{K_{2}}{2}|\xi|^{2}.

Next, for each K:=(K1,K2)K:=(K_{1},K_{2}), there exists an open neighborhood UKUU_{K}\subset U of p0p_{0} such that

1+2r(z)(K1+K2|z|2)3/21+2r(z)(K_{1}+K_{2}|z|^{2})\leq 3/2

holds for all zUKz\in U_{K}. Note that UKU_{K} may be chosen such that bΩU=bΩUKb\Omega\cap U=b\Omega\cap U_{K}. Using (6.2), we then obtain on UKU_{K}

(1+2rψ)Hr(ξ,ξ)\displaystyle(1+2r\psi)H^{\scriptscriptstyle\mathbb{C}}_{r}({\xi},{\xi}) 3C2(r2|ξ|2+|r,ξ|2)\displaystyle\geq-\frac{3C}{2}\left(r^{2}|\xi|^{2}+|\langle\partial r,\xi\rangle|^{2}\right)

for all ξn\xi\in\mathbb{C}^{n}. Therefore,

Hρ(ξ,ξ)r2|ξ|2(K2232C)+2|r,ξ|2(K13K2|z|234C)\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{\rho}({\xi},{\xi})\geq r^{2}|\xi|^{2}\left(\frac{K_{2}}{2}-\frac{3}{2}C\right)+2\left|\langle\partial r,\xi\rangle\right|^{2}\left(K_{1}-3K_{2}|z|^{2}-\frac{3}{4}C\right)

holds on UKU_{K} for all ξn\xi\in\mathbb{C}^{n}. Fix K2K_{2} such that K2>3CK_{2}>3C holds. Let DD be the maximum of |z||z| on UU, then fix K1K_{1} such that K1>3K2D2+34CK_{1}>3K_{2}D^{2}+\frac{3}{4}C holds. It follows easily that there exists a positive contant c>0c>0 such that

Hρ(ξ,ξ)c(ρ2|ξ|2+|ρ,ξ|2)\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{\rho}({\xi},{\xi})\geq c\left(\rho^{2}|\xi|^{2}+|\langle\partial\rho,\xi\rangle|^{2}\right) (6.5)

on UKU_{K} for all ξn\xi\in\mathbb{C}^{n}, i.e., ρ\rho is plurisubharmonic on UKU_{K}. ∎

Remark 6.6.

Note that the function ρ\rho constructed in the proof of Lemma 6.1 is strictly plurisubharmonic on UKbΩU_{K}\setminus b\Omega, see (6.5). Moreover, the complex Hessian of ρ\rho is positive definite at strictly pseudoconvex boundary points of Ω\Omega. To wit, the complex Hessian of ρ\rho is strictly positive in non-zero complex tangential directions at these boundary points by definition, and it is strictly positive in all directions with a non-vanishing normal component to the boundary by (6.5).

We note that a global version of Lemma 6.1 holds if Ω\Omega is bounded and UU is an open neighborhood of bΩb\Omega. Moreover, by a result of Morrow–Rossi [13, Lemma 1.3], see also [14], any smoothly bounded, strictly pseudoconvex, bounded domain in n\mathbb{C}^{n} admits a smooth defining function which is strictly plurisubharmonic in an open neighborhood of the closure of the domain. The same argument as the one used in the proof of Lemma 1.3 in [13] yields the following.

Corollary 6.7.

Let Ωn\Omega\Subset\mathbb{C}^{n} be a smoothly bounded domain. Assume that there exists a smooth defining function rr for Ω\Omega such that

Hr(ξ,ξ)C(r2|ξ|2+|r,ξ|2)ξn\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({\xi},{\xi})\geq-C\left(r^{2}|\xi|^{2}+\left|\left\langle\partial r,\xi\right\rangle\right|^{2}\right)\;\;\forall\xi\in\mathbb{C}^{n}

holds near bΩb\Omega for some constant C>0C>0. Then there exists a smooth defining function ρ\rho for Ω\Omega on an open neighborhood of Ω¯\bar{\Omega} such that

Hρ(ξ,ξ)c(ρ2|ξ|2+|ρ,ξ|2)ξn\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{\rho}({\xi},{\xi})\geq c\left(\rho^{2}|\xi|^{2}+\left|\left\langle\partial\rho,\xi\right\rangle\right|^{2}\right)\;\;\forall\xi\in\mathbb{C}^{n}

holds for some c>0c>0.

Whether a given local defining function actually satisfies condition (6.2) in some open neighborhood of the boundary, can be detected from the behaviour of the complex Hessian of that defining function and its normal derivative on the boundary of the domain as follows.

Proposition 6.8.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain, p0bΩp_{0}\in b\Omega. Then Ω\Omega admits a smooth local defining function which is plurisubharmonic near p0p_{0} if and only if there exist an open neighborhood UU of p0p_{0} and a smooth local defining function rr for Ω\Omega on UU such that

|Hr(L,N)|2\displaystyle\left|H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})\right|^{2} =𝒪(Hr(L,L))onbΩU,and\displaystyle=\mathcal{O}\left(H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})\right)\;\;\text{on}\;\;b\Omega\cap U,\;\;\text{and} (6.9)
|νHr(L,L)|2\displaystyle\left|\nu H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})\right|^{2} =𝒪(Hr(L,L))onbΩU.\displaystyle=\mathcal{O}\left(H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})\right)\;\;\text{on}\;\;b\Omega\cap U. (6.10)
Proof.

Let r:Ur\colon U\to\mathbb{R} be a smooth local defining function for Ω\Omega near p0p_{0}, and let LL be a nonvanishing holomorphic tangential vector field on UU. Note that the normal derivative (νHr(L,L))|bΩU(\nu H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L}))_{|_{b\Omega\cap U}} depends on LL, while Hr(L,L)|bΩUH^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})_{|_{b\Omega\cap U}} depends only on L|bΩUL_{|_{b\Omega\cap U}}. However, in case that (6.9) holds true, the condition (6.10) is in fact independent of the choice of LL. To see this, let LL^{\prime} be another nonvanishing holomorphic tangential vector field on UU. Then there exist a smooth function h:Uh\colon U\to\mathbb{C} and a vector field E𝒱(U)1,0E\in\mathcal{V}({U})^{1,0} such that L=hL+rEL^{\prime}=hL+rE. Thus

Hr(L,L)=|h|2Hr(L,L)+2rRe(hHr(L,E))+r2Hr(E,E),\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({L^{\prime}},{L^{\prime}})=\lvert h\rvert^{2}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})+2r\operatorname{Re}\bigl{(}hH^{\scriptscriptstyle\mathbb{C}}_{r}({L},{E})\bigr{)}+r^{2}H^{\scriptscriptstyle\mathbb{C}}_{r}({E},{E}),

so that

νHr(L,L)=|h|2νHr(L,L)+2|dr|Re(hHr(L,E))+𝒪(λ)on bΩU.\displaystyle\nu H^{\scriptscriptstyle\mathbb{C}}_{r}({L^{\prime}},{L^{\prime}})=\lvert h\rvert^{2}\nu H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})+2\lvert dr\rvert\operatorname{Re}\bigl{(}hH^{\scriptscriptstyle\mathbb{C}}_{r}({L},{E})\bigr{)}+\mathcal{O}(\lambda)\;\;\text{on }b\Omega\cap U.

Since E=aL+bNE=aL+bN on bΩUb\Omega\cap U for smooth functions a,ba,b, it follows from (6.9) that Hr(L,E)=𝒪(λ)H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{E})=\mathcal{O}(\sqrt{\lambda}) on bΩUb\Omega\cap U. In particular, νHr(L,L)=|h|2νHr(L,L)+𝒪(λ)on bΩU\nu H^{\scriptscriptstyle\mathbb{C}}_{r}({L^{\prime}},{L^{\prime}})=\lvert h\rvert^{2}\nu H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})+\mathcal{O}(\sqrt{\lambda})\;\;\text{on }b\Omega\cap U, which shows that (6.10) is well-defined.

Now suppose first that Ω\Omega admits a plurisubharmonic, smooth local defining function r:Ur\colon U\to\mathbb{R} near p0p_{0}. Then (6.9) holds by Proposition 5.2. Moreover, an application of the first part of Lemma 2.11, with f=Hr(L,L)f=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L}), shows that (6.10) is satisfied.

On the other hand, let r:Ur\colon U\to\mathbb{R} be a smooth local defining function for Ω\Omega near p0p_{0} such that (6.9) and (6.10) hold true. After possibly shrinking UU in the direction normal to bΩb\Omega, we may assume that there exists a smooth map π:UbΩ\pi\colon U\to b\Omega such |qπ(q)|=dbΩ(q)\lvert q-\pi(q)\rvert=d_{b\Omega}(q). Fix qUq\in U, and set p:=π(q)p\mathrel{\mathop{:}}=\pi(q). Moreover, fix ξ2\xi\in\mathbb{C}^{2}, and write ξ=aL(q)+bN(q)\xi=aL(q)+bN(q) for some a=a(q,ξ),b=b(q,ξ)a=a(q,\xi),b=b(q,\xi)\in\mathbb{C}. Then

Hr(ξ,ξ)(q)=|a|2Hr(L,L)(q)+2Re\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({\xi},{\xi})(q)=|a|^{2}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(q)+2\operatorname{Re} (ab¯Hr(L,N)(q))+|b|2Hr(N,N)(q).\displaystyle(a\bar{b}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})(q))+|b|^{2}H^{\scriptscriptstyle\mathbb{C}}_{r}({N},{N})(q).

In view of (2.10), with f=Hr(L,L)f=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L}), the Taylor expansion at pp in direction ν\nu gives

Hr(ξ,ξ)(q)\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({\xi},{\xi})(q) =|a|2(Hr(L,L)(p)+δbΩ(q)(νHr(L,L))(p)+𝒪(dbΩ2)(q))\displaystyle=|a|^{2}\left(H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(p)+\delta_{b\Omega}(q)(\nu H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L}))(p)+\mathcal{O}(d_{b\Omega}^{2})(q)\right)
+2Re(ab¯(Hr(L,N)(p)+𝒪(dbΩ)(q)))\displaystyle+2\operatorname{Re}\left(a\bar{b}(H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})(p)+\mathcal{O}(d_{b\Omega})(q))\right)
+|b|2𝒪(1)(q),\displaystyle+\lvert b\rvert^{2}\mathcal{O}(1)(q),

where the 𝒪\mathcal{O}-terms are functions that do not depend on qq and ξ\xi. Fix an open neighborhood VUV\Subset U of p0p_{0}. Using (6.4), it follows from (6.10) and (6.9) that there exist constants C1,C2>0C_{1},C_{2}>0 such that on VV

12Hr(L,L)(p)+δbΩ(q)(νHr(L,L))(p)\displaystyle\textstyle\frac{1}{2}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(p)+\delta_{b\Omega}(q)(\nu H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L}))(p) C1dbΩ2(q),\displaystyle\geq-C_{1}d_{b\Omega}^{2}(q),
12|a|2Hr(L,L)(p)+2Re(ab¯Hr(L,N)(p))\displaystyle\textstyle\frac{1}{2}\lvert a\rvert^{2}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(p)+2\operatorname{Re}\left(a\bar{b}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})(p)\right) C2|b|2.\displaystyle\geq-C_{2}\lvert b\rvert^{2}.

Hence, there exists a constant C3>0C_{3}>0, which does not depend on qq and ξ\xi, such that

Hr(ξ,ξ)(q)C3(|a|2dbΩ2(q)+|b|2)ξnqV.\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({\xi},{\xi})(q)\geq-C_{3}(|a|^{2}d_{b\Omega}^{2}(q)+\lvert b\rvert^{2})\;\;\hskip 5.69046pt\forall\;\xi\in\mathbb{C}^{n}\;\;\forall\;q\in V.

Since |a||ξ||a|\leq|\xi|, |b|=𝒪(|r,ξ|)|b|=\mathcal{O}(|\langle\partial r,\xi\rangle|) on UU, and dbΩ2=𝒪(r2)d_{b\Omega}^{2}=\mathcal{O}(r^{2}) on UU, it follows that r|Vr_{|_{V}} satisfies (6.2). The claim thus follows from Lemma 6.1. ∎

Remark 6.11.

The proofs of Lemma 6.1 and Proposition 6.8 imply the following: if (6.9) and (6.10) hold true, then for every KbΩUK\Subset b\Omega\cap U there exist an open neighborhood VUV\subset U of KK and a smooth local defining function for Ω\Omega on UU that is plurisubharmonic on VV.

Remark 6.12.

Versions of Proposition 5.2 and Proposition 6.8 can also be shown for domains in n\mathbb{C}^{n}, n>2n>2. In this case, LL has to be substituted by a frame {Lj}j=1n1\{L_{j}\}_{j=1}^{n-1} for T(bΩ)1,0T(b\Omega)^{1,0} near p0p_{0}.

Condition (6.10) may always be achieved near boundary points of type 44, independent of whether the smoothly bounded, pseudoconvex domain in consideration actually admits a smooth local defining function which is plurisubharmonic on the boundary.

Lemma 6.13.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain, p0bΩp_{0}\in b\Omega. If bΩb\Omega is of type 44 at p0p_{0}, then for every nonvanishing holomorphic tangential vector field LL near p0p_{0} there exists a smooth local defining function ρ:U\rho\colon U\to\mathbb{R} for Ω\Omega near p0p_{0} such that

|νHρ(L,L)|2\displaystyle\left|\nu H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})\right|^{2} =𝒪(Hρ(L,L))onbΩU.\displaystyle=\mathcal{O}\left(H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})\right)\;\;\text{on}\;\;b\Omega\cap U. (6.14)
Proof.

Let r:Ur\colon U\to\mathbb{R} be a smooth local defining function for Ω\Omega near p0p_{0}, and let LL be a nonvanishing holomorphic tangential vector field near p0p_{0}. After possibly shrinking UU, we may assume that dr0dr\neq 0, and that LL is defined on UU. Set Λ:=Hr(L,L)\Lambda\mathrel{\mathop{:}}=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L}) and λ:=Λ|bΩU\lambda\mathrel{\mathop{:}}=\Lambda_{|_{b\Omega\cap U}}. We claim that, after possibly shrinking UU, for every smooth function F:UF\colon U\to\mathbb{R}, there exists a smooth function h:Uh\colon U\to\mathbb{R} such that

{Lh=𝒪(λ)LL¯h=F+𝒪(λ) on bΩU.\displaystyle\begin{cases}&Lh=\mathcal{O}(\sqrt{\lambda})\\ &L\bar{L}h=F+\mathcal{O}(\sqrt{\lambda})\end{cases}\;\;\text{ on }b\Omega\cap U. (6.15)

Indeed, define smooth functions A,B:UA,B\colon U\to\mathbb{R} by

A:=|LΛ|2andB:=Λ2+|LL¯Λ|2+|LLΛ|2,\displaystyle A\mathrel{\mathop{:}}=|L\Lambda|^{2}\quad\text{and}\quad B\mathrel{\mathop{:}}=\Lambda^{2}+\lvert L\bar{L}\Lambda\rvert^{2}+\lvert LL\Lambda\rvert^{2},

and set

h:=FAB.\displaystyle h\mathrel{\mathop{:}}=F\frac{A}{B}. (6.16)

Since Ω\Omega is of type 4 at p0p_{0}, after possibly shrinking UU, we can assume that B>0B>0 on UU. Thus, hh is well-defined. Moreover, since LΛ=𝒪(λ)L\Lambda=\mathcal{O}(\sqrt{\lambda}) on bΩUb\Omega\cap U by (2.13), it follows that AA and LALA are of class 𝒪(λ)\mathcal{O}(\sqrt{\lambda}) on bΩUb\Omega\cap U, and, in particular, that

Lh=𝒪(λ)on bΩU.\displaystyle Lh=\mathcal{O}(\sqrt{\lambda})\quad\text{on }b\Omega\cap U. (6.17)

Moreover,

LL¯h=FLL¯AB+𝒪(λ)=F|LL¯Λ|2+|LLΛ|2B+𝒪(λ)=F+𝒪(λ) on bΩU.L\bar{L}h=F\frac{L\bar{L}A}{B}+\mathcal{O}(\sqrt{\lambda})=F\frac{\lvert L\bar{L}\Lambda\rvert^{2}+\lvert LL\Lambda\rvert^{2}}{B}+\mathcal{O}(\sqrt{\lambda})=F+\mathcal{O}(\sqrt{\lambda})\;\;\text{ on }b\Omega\cap U.

Now let h:Uh\colon U\to\mathbb{R} be an arbitrary smooth function, and set ρ:=reh\rho\mathrel{\mathop{:}}=re^{h}. Then

Hρ(L,L)=eh(Hr(L,L)+r(Hh(L,L)+|Lh|2)),\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})=e^{h}\left(H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})+r\left(H^{\scriptscriptstyle\mathbb{C}}_{h}({L},{L})+\lvert Lh\rvert^{2}\right)\right),

and thus

νHρ(L,L)=eh(νHr(L,L)+|dr|(Hh(L,L)+|Lh|2))+𝒪(λ)on bΩU.\displaystyle\nu H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})=e^{h}\left(\nu H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})+\lvert dr\rvert\left(H^{\scriptscriptstyle\mathbb{C}}_{h}({L},{L})+\lvert Lh\rvert^{2}\right)\right)+\mathcal{O}(\lambda)\quad\text{on }b\Omega\cap U.

Since Hh(L,L)=LL¯h(LL¯)hH^{\scriptscriptstyle\mathbb{C}}_{h}({L},{L})=L\bar{L}h-(\nabla_{L}\bar{L})h, and since on bΩUb\Omega\cap U it follows from Hr(L,L)=(LL¯)rH^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})=-(\nabla_{L}\bar{L})r on bΩUb\Omega\cap U that the component of LL¯\nabla_{L}\bar{L} normal to bΩb\Omega is of the form 𝒪(λ)\mathcal{O}(\lambda), it follows from (6.17) that

νHρ(L,L)=eh(νHr(L,L)+|dr|LL¯h)+𝒪(λ)on bΩU.\displaystyle\nu H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})=e^{h}\left(\nu H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})+\lvert dr\rvert L\bar{L}h\right)+\mathcal{O}(\sqrt{\lambda})\quad\text{on }b\Omega\cap U.

Thus, if hh is the solution for (6.15) with

F:=1|dr|νHr(L,L),\displaystyle F\mathrel{\mathop{:}}=-\frac{1}{\lvert dr\rvert}\nu H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L}), (6.18)

then ρ\rho satisfies (6.14). ∎

We now can prove the main result of this section.

Theorem 6.19.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain. Suppose p0bΩp_{0}\in b\Omega is such that cp0=4c_{p_{0}}=4. If Ω\Omega admits a smooth local defining function near p0p_{0} which is plurisubharmonic on bΩb\Omega near p0p_{0}, then Ω\Omega admits a smooth local defining function near p0p_{0} which is plurisubharmonic.

Proof.

Let r:Ur\colon U\to\mathbb{R} be a smooth local defining function for Ω\Omega near p0p_{0} such that rr is plurisubharmonic on bΩUb\Omega\cap U. After possibly shrinking UU, we can assume that cp4c_{p}\leq 4 for all pbΩUp\in b\Omega\cap U. Set ρ:=reh\rho\mathrel{\mathop{:}}=re^{h}, where hh is the solution to (6.15) with FF given as in (6.18). Then, as in the proof of Lemma 6.13, we see that ρ\rho satisfies (6.10). On the other hand, since, by Proposition 5.2, rr satisfies (6.9), and since

Hρ(L,N)=eh(Hr(L,N)+LhN¯r)on bΩU,H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})=e^{h}\left(H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})+Lh\cdot\bar{N}r\right)\;\;\text{on }b\Omega\cap U,

it follows with (6.17) that ρ\rho satisfies (6.9). The claim thus follows from Proposition 6.8. ∎

Note that, if rr is a smooth defining function for a smoothly bounded, pseudoconvex domain Ω2\Omega\subset\mathbb{C}^{2} such that Ω\Omega is of type 44 at all its weakly pseudoconvex boundary points, then the function hh defined in (6.16) and (6.18) is defined in an open neighborhood of bΩb\Omega. Moreover, the function hh solves (6.15) on bΩb\Omega. In view of Remark 6.11, this implies the following global result.

Corollary 6.20.

Let Ω2\Omega\Subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain. Suppose that Ω\Omega has a smooth defining function which is plurisubharmonic on bΩb\Omega, and that cp4c_{p}\leq 4 for all pbΩp\in b\Omega. Then Ω\Omega admits a smooth defining function which is plurisubharmonic in an open neighborhood of bΩb\Omega.

An analogon to Theorem 6.19 near higher order boundary points is not apparent, although condition (6.10) always holds at boundary points of type larger than 44, whenever rr is a defining function that is plurisubharmonic on the boundary, as shown by the next Lemma.

Lemma 6.21.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded pseudoconvex domain. Let p0bΩp_{0}\in b\Omega such that cp06c_{p_{0}}\geq 6. If rr is a smooth local defining function of bΩb\Omega near p0p_{0} which is plurisubharmonic on bΩb\Omega near p0p_{0}, and if LL is a nonvanishing holomorphic tangential vector field near p0p_{0}, then

νHr(L,L)(p0)=0.\displaystyle\nu H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(p_{0})=0.
Proof.

Let rr be a smooth local defining function for Ω\Omega on some open neighborhood UU of p0p_{0} such that rr is plurisubharmonic on bΩUb\Omega\cap U. After possibly shrinking UU, we may assume that N:=NrN\mathrel{\mathop{:}}=N_{r} is defined on UU, see (2.6). As usual, we write L=12(X+iY)L=\frac{1}{2}(X+iY) with X,Y𝒱(U)X,Y\in\mathcal{V}({U}).

Consider the function g:bΩUg\colon b\Omega\cap U\to\mathbb{R} given by

g=Hr(L,L)Hr(N,N)|Hr(L,N)|2 on bΩU.g=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})H^{\scriptscriptstyle\mathbb{C}}_{r}({N},{N})-|H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})|^{2}\quad\text{ on }b\Omega\cap U.

Since rr is plurisubharmonic on bΩUb\Omega\cap U, it follows that g0g\geq 0, and since Ω\Omega is weakly pseudoconvex at p0p_{0}, it follows that g(p0)=0g(p_{0})=0. Thus, (XXg)(p0)0(XXg)(p_{0})\geq 0 and (YYg)(p0)0(YYg)(p_{0})\geq 0, see Lemma 4.7. Since LL¯g=14(XXg+YYg)i4[X,Y]gL\bar{L}g=\frac{1}{4}(XXg+YYg)-\frac{i}{4}[X,Y]g, and since the tangential derivative [X,Y]g[X,Y]g vanishes at p0p_{0}, it follows that (LL¯g)(p0)0(L\bar{L}g)(p_{0})\geq 0. Moreover, the fact that cp0>4c_{p_{0}}>4 implies that for λ:=Hr(L,L)|bΩU\lambda\mathrel{\mathop{:}}=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})_{|_{b\Omega\cap U}} the functions λ,Lλ,L¯λ,LL¯λ\lambda,L\lambda,\bar{L}\lambda,L\bar{L}\lambda all vanish at p0p_{0}. Hence, since Hr(L,N)(p0)=0H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})(p_{0})=0,

(LL¯g)(p0)=|LHr(L,N)|2(p0)|L¯Hr(L,N)|2(p0).(L\bar{L}g)(p_{0})=-\left|LH^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})\right|^{2}(p_{0})-\left|\bar{L}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})\right|^{2}(p_{0}).

In particular, it follows that L¯Hr(L,N)(p0)=0\bar{L}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})(p_{0})=0.

However, if we denote coordinates in 2\mathbb{C}^{2} by z=(z1,z2)z=(z_{1},z_{2}), and if we write L=j=12LjzjL=\sum_{j=1}^{2}L^{j}\frac{\partial}{\partial z_{j}}, N=j=12NjzjN=\sum_{j=1}^{2}N^{j}\frac{\partial}{\partial z_{j}}, then

L¯Hr(L,N)\displaystyle\bar{L}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N}) =j,k,=123rz¯zjz¯kL¯LjN¯k+Hr(L¯L,N)+Hr(L,LN),\displaystyle=\sum_{j,k,\ell=1}^{2}\frac{\partial^{3}r}{\partial\bar{z}_{\ell}\partial z_{j}\partial\bar{z}_{k}}\bar{L}^{\ell}L^{j}\bar{N}^{k}+H^{\scriptscriptstyle\mathbb{C}}_{r}({\nabla_{\bar{L}}L},{N})+H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{\nabla_{L}N}), (6.22)
N¯Hr(L,L)\displaystyle\bar{N}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L}) =j,k,=123rz¯zjz¯kN¯LjL¯k+Hr(N¯L,L)+Hr(L,NL).\displaystyle=\sum_{j,k,\ell=1}^{2}\frac{\partial^{3}r}{\partial\bar{z}_{\ell}\partial z_{j}\partial\bar{z}_{k}}\bar{N}^{\ell}L^{j}\bar{L}^{k}+H^{\scriptscriptstyle\mathbb{C}}_{r}({\nabla_{\bar{N}}L},{L})+H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{\nabla_{N}L}). (6.23)

Since Ω\Omega is weakly pseudoconvex at p0p_{0}, one has Hr(L,L)(p0)=Hr(L,N)(p0)=0H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(p_{0})=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})(p_{0})=0, i.e., Hr(L,)(p0):𝒱(U)1,0H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{\,\cdot\,})(p_{0})\colon\mathcal{V}({U})^{1,0}\to\mathbb{R} is identically zero. Moreover, since 0LL¯r=Hr(L,L)+(LL¯)r0\equiv L\bar{L}r=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})+(\nabla_{L}\bar{L})r, it follows that ((LL¯)r)(p0)=0((\nabla_{L}\bar{L})r)(p_{0})=0, i.e., (LL¯)p0=cLp0(\nabla_{L}\bar{L})_{p_{0}}=cL_{p_{0}} for some constant cc\in\mathbb{C}. Thus, the two rightmost terms in both (6.22) and (6.23) vanish at p0p_{0}, which proves that L¯Hr(L,N)(p0)=N¯Hr(L,L)(p0)\bar{L}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})(p_{0})=\bar{N}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(p_{0}). Since all tangential derivatives of Hr(L,L)H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L}) vanish at p0p_{0}, it follows that νHr(L,L)(p0)=2N¯Hr(L,L)(p0)=0\nu H^{\scriptscriptstyle\mathbb{C}}_{r}(L,L)(p_{0})=2\bar{N}H^{\scriptscriptstyle\mathbb{C}}_{r}(L,L)(p_{0})=0, which completes the proof. ∎

This weaker result for boundary points of type greater than 44 leads to a simplified proof of the Diederich–Fornæss index and the Steinness index being 11 for smoothly bounded, pseudoconvex domains which admit a smooth defining function that is plurisubharmonic on the boundary of the domain.

Corollary 6.24.

Let Ω2\Omega\Subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain. Suppose that

  • (i)

    cp4c_{p}\neq 4 for all pbΩp\in b\Omega, and

  • (ii)

    Ω\Omega admits a smooth defining function which is plurisubharmonic on bΩb\Omega.

Then for every η(0,1)\eta\in(0,1) there exist a constant K>0K>0 and an open neighborhood UU of bΩb\Omega such that (rKr2)η-(-r-Kr^{2})^{\eta} is plurisubharmonic on ΩU\Omega\cap U.

Similarly, for every μ>1\mu>1 there exist a constant K>0K>0 and an open neighborhood UU of bΩb\Omega such that (rKr2)μ-(-r-Kr^{2})^{\mu} is plurisubharmonic on ΩcU\Omega^{c}\cap U.

We note that (ii) itself leads to the Diederich–Fornæss index being 1, see [7]. However, the additional condition (i) simplifies the construction in [7, Section 3] considerably.

Proof.

Let rr be a smooth defining function of Ω\Omega which is plurisubharmonic on bΩb\Omega, and assume that L:=LrL\mathrel{\mathop{:}}=L_{r} and N:=NrN\mathrel{\mathop{:}}=N_{r} are defined on some open neighborhood UU of bΩb\Omega. After possibly shrinking UU, we may use (2.10) for f=Hr(L,L)f=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L}) and qUq\in U with cπ(q)6c_{\pi(q)}\geq 6. It then follows from Lemma 6.21 that

Hr(L,L)(q)=𝒪(r2)(q).\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(q)=\mathcal{O}(r^{2})(q).

Similarly, one obtains Hr(L,N)(q)=𝒪(r)(q)H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})(q)=\mathcal{O}(r)(q) and Hr(N,N)(q)=𝒪(1)(q)H^{\scriptscriptstyle\mathbb{C}}_{r}({N},{N})(q)=\mathcal{O}(1)(q). It then follows

Hr(ξ,ξ)(q)=(𝒪(r2)|ξ|2+𝒪(|r,ξ|2))(q)ξnqU with cπ(q)6.\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({\xi},{\xi})(q)=\left(\mathcal{O}(r^{2})|\xi|^{2}+\mathcal{O}(|\langle\partial r,\xi\rangle|^{2})\right)(q)\;\;\hskip 5.69046pt\forall\;\xi\in\mathbb{C}^{n}\;\;\forall\;q\in U\text{ with }c_{\pi(q)}\geq 6.

The arguments following (3.7) in [7] then prove the claim. ∎

7. On a special class of pseudoconvex domains

In this section, we derive a sufficient condition for the existence of local defining functions, which are plurisubharmonic on the boundary, in terms of real coordinates. While this condition, in contrast to the criterion given in Proposition 5.2, is not an equivalent characterization, it has the advantage of being independent of the choice of defining function, and thus is more easily checkable.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded domain, and let r:Ur\colon U\to\mathbb{R} be a smooth local defining function for Ω\Omega near some point p0bΩp_{0}\in b\Omega. After possibly shrinking UU, let LL be a nonvanishing holomorphic tangential vector field on UU, and let N=NrN=N_{r} be defined as in (2.6). Write

L=12(X+iY),N=12(ν+iT)\displaystyle L=\textstyle\frac{1}{2}(X+iY),\;\;N=\frac{1}{2}(\nu+iT)

with X,Y,T,ν𝒱(U)X,Y,T,\nu\in\mathcal{V}({U}). The matrix associated with the real Hessian form Qr:𝒱(U)×𝒱(U)Q^{\scriptscriptstyle\mathbb{R}}_{r}\colon\mathcal{V}({U})\times\mathcal{V}({U})\to\mathbb{R} relative to the basis (X,Y,T,ν)(X,Y,T,\nu) will be denoted by 𝒬r\mathcal{Q}^{\scriptscriptstyle\mathbb{R}}_{r}, i.e.,

𝒬r:=(Qr(X,X)Qr(X,Y)Qr(X,T)Qr(X,ν)Qr(Y,X)Qr(Y,Y)Qr(Y,T)Qr(Y,ν)Qr(T,X)Qr(T,Y)Qr(T,T)Qr(T,ν)Qr(ν,X)Qr(ν,Y)Qr(ν,T)Qr(ν,ν)).\displaystyle\mathcal{Q}^{\scriptscriptstyle\mathbb{R}}_{r}\mathrel{\mathop{:}}=\begin{pmatrix}Q^{\scriptscriptstyle\mathbb{R}}_{r}({X},{X})&Q^{\scriptscriptstyle\mathbb{R}}_{r}({X},{Y})&Q^{\scriptscriptstyle\mathbb{R}}_{r}({X},{T})&Q^{\scriptscriptstyle\mathbb{R}}_{r}({X},{\nu})\\ Q^{\scriptscriptstyle\mathbb{R}}_{r}({Y},{X})&Q^{\scriptscriptstyle\mathbb{R}}_{r}({Y},{Y})&Q^{\scriptscriptstyle\mathbb{R}}_{r}({Y},{T})&Q^{\scriptscriptstyle\mathbb{R}}_{r}({Y},{\nu})\\ Q^{\scriptscriptstyle\mathbb{R}}_{r}({T},{X})&Q^{\scriptscriptstyle\mathbb{R}}_{r}({T},{Y})&Q^{\scriptscriptstyle\mathbb{R}}_{r}({T},{T})&Q^{\scriptscriptstyle\mathbb{R}}_{r}({T},{\nu})\\ Q^{\scriptscriptstyle\mathbb{R}}_{r}({\nu},{X})&Q^{\scriptscriptstyle\mathbb{R}}_{r}({\nu},{Y})&Q^{\scriptscriptstyle\mathbb{R}}_{r}({\nu},{T})&Q^{\scriptscriptstyle\mathbb{R}}_{r}({\nu},{\nu})\end{pmatrix}.

We readily recognize that various convexity-like boundary conditions for Ω\Omega near p0p_{0} may be expressed through conditions on entries of the leading principal 3×33\times 3 submatrix of 𝒬r\mathcal{Q}^{\scriptscriptstyle\mathbb{R}}_{r} for pbΩp\in b\Omega near p0p_{0}.

  • (i)

    Ω\Omega is convex near p0p_{0} if the leading principal 3×33\times 3 submatrix of 𝒬r(p)\mathcal{Q}^{\scriptscriptstyle\mathbb{R}}_{r}(p) is positive semi-definite for all pbΩp\in b\Omega near p0p_{0}.

  • (ii)

    Ω\Omega is \mathbb{C}-convex near p0p_{0} if the leading principal 2×22\times 2 submatrix of 𝒬r(p)\mathcal{Q}^{\scriptscriptstyle\mathbb{R}}_{r}(p) is positive semi-definite for all pbΩp\in b\Omega near p0p_{0}.

  • (iii)

    Ω\Omega is pseudoconvex near p0p_{0} if the trace of the leading principal 2×22\times 2 submatrix of 𝒬r(p)\mathcal{Q}^{\scriptscriptstyle\mathbb{R}}_{r}(p) is non-negative for all pbΩp\in b\Omega near p0p_{0}.

In order to see how to express plurisubharmonicity on the boundary of a smooth local defining function in real coordinates, we need to formulate condition (5.3) in real coordinates. Thus, we compute

4Hr(L,N)=4L(N¯r)4(LN¯)r=(X+iY)(νiT)r(X+iY(νiT))r=Xνr(Xν)r+YTr(YT)r+i(Yνr(Yν)rXTr+(XT)r)=Qr(X,ν)+Qr(Y,T)+i(Qr(Y,ν)Qr(X,T)).\begin{split}4H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})&=4L(\bar{N}r)-4\left(\nabla_{L}\bar{N}\right)r\\ &=(X+iY)(\nu-iT)r-\left(\nabla_{X+iY}(\nu-iT)\right)r\\ &=X\nu r-(\nabla_{X}\nu)r+YTr-(\nabla_{Y}T)r\\ &\quad+i\bigl{(}Y\nu r-(\nabla_{Y}\nu)r-XTr+(\nabla_{X}T)r\bigr{)}\\ &=Q^{\scriptscriptstyle\mathbb{R}}_{r}({X},{\nu})+Q^{\scriptscriptstyle\mathbb{R}}_{r}({Y},{T})+i\bigl{(}Q^{\scriptscriptstyle\mathbb{R}}_{r}({Y},{\nu})-Q^{\scriptscriptstyle\mathbb{R}}_{r}({X},{T})\bigr{)}.\end{split} (7.1)

Proposition 5.2 may now be reformulated in terms of entries of 𝒬r\mathcal{Q}^{\scriptscriptstyle\mathbb{R}}_{r} as follows.

Lemma 7.2.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded, pseudoconvex domain, p0bΩp_{0}\in b\Omega. Then Ω\Omega admits a smooth local defining function which is plurisubharmonic on bΩb\Omega near p0p_{0} if and only if there exists a smooth local defining function ρ:U\rho\colon U\to\mathbb{R} for Ω\Omega on some open neighborhood UU of p0p_{0} such that

(Qρ(X,ν)+Qρ(Y,T))2+(Qρ(Y,ν)Qρ(X,T))2=𝒪(Hρ(L,L))\displaystyle\bigl{(}Q^{\scriptscriptstyle\mathbb{R}}_{\rho}({X},{\nu})+Q^{\scriptscriptstyle\mathbb{R}}_{\rho}({Y},{T})\bigr{)}^{2}+\bigl{(}Q^{\scriptscriptstyle\mathbb{R}}_{\rho}({Y},{\nu})-Q^{\scriptscriptstyle\mathbb{R}}_{\rho}({X},{T})\bigr{)}^{2}=\mathcal{O}\left(H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})\right) (7.3)

on bΩUb\Omega\cap U.

In the proof of Theorem 5.1 it is shown that, given any smooth local defining function r:Ur\colon U\to\mathbb{R} for Ω\Omega, then ρ:=reh\rho\mathrel{\mathop{:}}=re^{h} satisfies (7.3) if h𝒞(U,)h\in\mathcal{C}^{\infty}(U,\mathbb{R}) solves the equation

Lh=2|dr|Hr(L,N)+𝒪(λ)on bΩU,Lh=-\frac{2}{\lvert dr\rvert}H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{N})+\mathcal{O}(\sqrt{\lambda})\quad\text{on }b\Omega\cap U,

with λ:=Hr(L,L)|bΩU\lambda\mathrel{\mathop{:}}=H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})_{|_{b\Omega\cap U}}. This can be reformulated in real coordinates as follows,

Xh\displaystyle Xh =1|dr|(Qr(X,ν)+Qr(Y,T))+𝒪(λ)on bΩU,\displaystyle=\frac{-1}{|dr|}\bigl{(}Q^{\scriptscriptstyle\mathbb{R}}_{r}({X},{\nu})+Q^{\scriptscriptstyle\mathbb{R}}_{r}({Y},{T})\bigr{)}+\mathcal{O}(\sqrt{\lambda})\quad\text{on }b\Omega\cap U, (7.4)
Yh\displaystyle Yh =1|dr|(Qr(X,T)Qr(Y,ν))+𝒪(λ)on bΩU.\displaystyle=\frac{1}{|dr|}\bigl{(}Q^{\scriptscriptstyle\mathbb{R}}_{r}({X},{T})-Q^{\scriptscriptstyle\mathbb{R}}_{r}({Y},{\nu})\bigr{)}+\mathcal{O}(\sqrt{\lambda})\quad\text{on }b\Omega\cap U. (7.5)
Example 7.6.

Let us revisit Example 5.10. There, in 2\mathbb{C}^{2} with coordinates z=x+iyz=x+iy and w=u+ivw=u+iv, we consider Ω={(z,w)U:r(z,w)<0}\Omega=\{(z,w)\in U:r(z,w)<0\} with

U={(z,w)2:|x|<π/2},r(z,w)=u12(xv)2ln(cos(x)).\displaystyle U=\{(z,w)\in\mathbb{C}^{2}:|x|<\pi/2\},\quad r(z,w)=u-\textstyle\frac{1}{2}(x-v)^{2}-\ln(\cos(x)).

We already computed that, with L=rwzrzwL=r_{w}\frac{\partial}{\partial z}-r_{z}\frac{\partial}{\partial w},

Hr(L,L)(z,w)=116(xv)2sec2(x).\displaystyle H^{\scriptscriptstyle\mathbb{C}}_{r}({L},{L})(z,w)=\textstyle\frac{1}{16}(x-v)^{2}\sec^{2}(x).

In particular, note that the function P(z,w):=xvP(z,w)\mathrel{\mathop{:}}=x-v is of class 𝒪(λ)\mathcal{O}(\sqrt{\lambda}) on bΩb\Omega. Considering real vector fields on 2\mathbb{C}^{2} as maps to 4\mathbb{R}^{4}, we can then compute further that

X\displaystyle X =1|dr|(ru,rv,rx,ry)=1|dr|(1,0,tan(x),0)+𝒪(λ),\displaystyle=\frac{1}{\lvert dr\rvert}(r_{u},-r_{v},-r_{x},r_{y})=\frac{1}{\lvert dr\rvert}\left(1,0,-\tan(x),0\right)+\mathcal{O}(\sqrt{\lambda}),
Y\displaystyle Y =1|dr|(rv,ru,ry,rx)=1|dr|(0,1,0,tan(x))+𝒪(λ),\displaystyle=\frac{1}{\lvert dr\rvert}(-r_{v},-r_{u},r_{y},r_{x})=\frac{1}{\lvert dr\rvert}\left(0,-1,0,\tan(x)\right)+\mathcal{O}(\sqrt{\lambda}),
T\displaystyle T =1|dr|(ry,rx,rv,ru)=1|dr|(0,tan(x),0,1)+𝒪(λ),\displaystyle=\frac{1}{\lvert dr\rvert}(r_{y},-r_{x},r_{v},-r_{u})=\frac{1}{\lvert dr\rvert}\left(0,-\tan(x),0,-1\right)+\mathcal{O}(\sqrt{\lambda}),
ν\displaystyle\nu =1|dr|(rx,ry,ru,rv)=1|dr|(tan(x),0,1,0)+𝒪(λ),\displaystyle=\frac{1}{\lvert dr\rvert}(r_{x},r_{y},r_{u},r_{v})=\frac{1}{\lvert dr\rvert}\left(\tan(x),0,1,0\right)+\mathcal{O}(\sqrt{\lambda}),

where, in slight deviation from previous notation, the terms 𝒪(λ)\mathcal{O}(\sqrt{\lambda}) denote vector fields with coefficients that are of class 𝒪(λ)\mathcal{O}(\sqrt{\lambda}) on bΩb\Omega. From this, it follows readily that on bΩb\Omega

Qr(X,ν)\displaystyle Q^{\scriptscriptstyle\mathbb{R}}_{r}({X},{\nu}) =1|dr|2tan3(x)+𝒪(λ),\displaystyle=\frac{1}{\lvert dr\rvert^{2}}\tan^{3}(x)+\mathcal{O}(\sqrt{\lambda}), Qr(Y,T)\displaystyle Q^{\scriptscriptstyle\mathbb{R}}_{r}({Y},{T}) =1|dr|2tan(x)+𝒪(λ),\displaystyle=\frac{1}{\lvert dr\rvert^{2}}\tan(x)+\mathcal{O}(\sqrt{\lambda}),
Qr(X,T)\displaystyle Q^{\scriptscriptstyle\mathbb{R}}_{r}({X},{T}) =1|dr|2+𝒪(λ),\displaystyle=-\frac{1}{\lvert dr\rvert^{2}}+\mathcal{O}(\sqrt{\lambda}), Qr(Y,ν)\displaystyle Q^{\scriptscriptstyle\mathbb{R}}_{r}({Y},{\nu}) =1|dr|2tan2(x)+𝒪(λ).\displaystyle=\frac{1}{\lvert dr\rvert^{2}}\tan^{2}(x)+\mathcal{O}(\sqrt{\lambda}).

Since |dr|2=1+tan2(x)+𝒪(λ)|dr|^{2}=1+\tan^{2}(x)+\mathcal{O}(\lambda) on bΩb\Omega, it follows further that (7.4) and (7.5) are given by

hxtan(x)hu\displaystyle h_{x}-\tan(x)h_{u} =tan(x)+𝒪(λ)on bΩ,\displaystyle=-\tan(x)+\mathcal{O}(\sqrt{\lambda})\quad\text{on }b\Omega,\
hy+tan(x)hv\displaystyle-h_{y}+\tan(x)h_{v} =1+𝒪(λ)on bΩ.\displaystyle=-1+\mathcal{O}(\sqrt{\lambda})\quad\text{on }b\Omega.

It is easy to see that, e.g., h(z,w)=y+ln(cos(x))h(z,w)=y+\ln(\cos(x)) and h(z,w)=y+uh(z,w)=y+u both satisfy these last two equations, so that rehre^{h} satisfies (7.3). Hence, if ρ=reh\rho=re^{h}, then for every VUV\Subset U there exists a smooth defining function for Ω\Omega on UU that is plurisubharmonic on bΩVb\Omega\cap V, see Remark 5.4. In view of Theorem 6.19, this means that Ω\Omega admits plurisubharmonic smooth local defining functions near each boundary point.

Definition 7.7.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded domain. We say that Ω\Omega is sesquiconvex at p0bΩp_{0}\in b\Omega if Ω\Omega is pseudoconvex at p0p_{0} and if there exists a smooth local defining function ρ:U\rho\colon U\to\mathbb{R} for Ω\Omega near p0p_{0} such that

Qρ(Y,T)2+Qρ(X,T)2=𝒪(Hρ(L,L))on bΩU.\displaystyle Q^{\scriptscriptstyle\mathbb{R}}_{\rho}({Y},{T})^{2}+Q^{\scriptscriptstyle\mathbb{R}}_{\rho}({X},{T})^{2}=\mathcal{O}\left(H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})\right)\quad\text{on }b\Omega\cap U. (7.8)
Remark 7.9.

(1) Let ρ:U\rho\colon U\to\mathbb{R} be a smooth local defining function for Ω\Omega near p0p_{0}, and let h:Uh\colon U\to\mathbb{R} be smooth. Then for any two tangential vector fields V,WV,W near p0p_{0} one has Qρeh(V,W)=(VW)(ρeh)=eh(VW)ρ=ehQρ(V,W)Q^{\scriptscriptstyle\mathbb{R}}_{\rho e^{h}}({V},{W})=-(\nabla_{V}W)(\rho e^{h})=-e^{h}(\nabla_{V}W)\rho=e^{h}Q^{\scriptscriptstyle\mathbb{R}}_{\rho}({V},{W}) on bΩUb\Omega\cap U. In particular, this shows that condition (7.8) is independent of the choice of a local defining function ρ\rho.

(2) Since Hρ(L,L)=Qρ(X,X)+Qρ(Y,Y)H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})=Q^{\scriptscriptstyle\mathbb{R}}_{\rho}({X},{X})+Q^{\scriptscriptstyle\mathbb{R}}_{\rho}({Y},{Y}), one easily sees that every domain Ω2\Omega\subset\mathbb{C}^{2} that is convex at p0bΩp_{0}\in b\Omega is sesquiconvex at p0p_{0}. Moreover, it is clear from the definition that if Ω\Omega is strictly pseudoconvex at p0bΩp_{0}\in b\Omega, then Ω\Omega is sesquiconvex at p0p_{0}. On the other hand, the domain Ω\Omega considered in Example 5.10 and Example 7.6 is not sesquiconvex at 0.

In the following, we show that sesquiconvexity at a boundary point p0bΩp_{0}\in b\Omega implies the existence of local defining functions which are plurisubharmonic on a one-sided neighborhood Ω¯U\bar{\Omega}\cap U of p0p_{0}, i.e., ρ(q)0\mathcal{H}^{\scriptscriptstyle\mathbb{C}}_{\rho}(q)\geq 0 for every qΩ¯Uq\in\bar{\Omega}\cap U.

Proposition 7.10.

If Ω2\Omega\subset\mathbb{C}^{2} is sesquiconvex at p0bΩp_{0}\in b\Omega, then Ω\Omega admits a smooth local defining function ρ:U\rho:U\longrightarrow\mathbb{R} near p0p_{0} which is plurisubharmonic on Ω¯U\bar{\Omega}\cap U.

Proof.

Let r:Ur\colon U\to\mathbb{R} be a smooth local defining function for Ω\Omega near p0p_{0}, and assume that dr0dr\neq 0 on UU. We will show that ρ:=r/|dr|\rho\mathrel{\mathop{:}}=r/|dr| satisfies

|Hρ(L,N)|\displaystyle\left|H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})\right| =𝒪(λ)on bΩU, and\displaystyle=\mathcal{O}(\sqrt{\lambda})\;\;\text{on }b\Omega\cap U,\text{ and} (7.11)
νHρ(L,L)\displaystyle\nu H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L}) 𝒪(λ)on bΩU.\displaystyle\leq\mathcal{O}(\sqrt{\lambda})\;\;\text{on }b\Omega\cap U. (7.12)

The claim then follows from a brief analysis of the proofs of Lemma 6.1 and Proposition 6.8.

Let LL be a nonvanishing holomorphic tangential vector field on UU, and let X,Y𝒱(U)X,Y\in\mathcal{V}({U}) such that L=12(X+iY)L=\frac{1}{2}(X+iY). A straightforward computation shows that

0=L(|dρ|2)=Qρ(X,ν)+iQρ(Y,ν)on bΩU.\displaystyle 0=L(|d\rho|^{2})=Q^{\scriptscriptstyle\mathbb{R}}_{\rho}({X},{\nu})+iQ^{\scriptscriptstyle\mathbb{R}}_{\rho}({Y},{\nu})\quad\text{on }b\Omega\cap U. (7.13)

Since Ω\Omega is sesquiconvex at p0p_{0}, it thus follows from (7.8) and the computations in (7.1) that (7.11) is true. The equation in (7.13) may be expressed in complex notation, with N=2ρz¯1z1+2ρz¯2z2N=2\rho_{\bar{z}^{1}}\frac{\partial}{\partial z^{1}}+2\rho_{\bar{z}^{2}}\frac{\partial}{\partial z^{2}}, as

0=L(|ρ|2)=Qρ(L,N)+Hρ(L,N)on bΩU.0=L(|\partial\rho|^{2})=Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})+H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})\;\;\text{on }b\Omega\cap U.

Since (7.11) holds, it then follows that

|Qρ(L,N)|=𝒪(λ)on bΩU.\displaystyle\left|Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})\right|=\mathcal{O}(\sqrt{\lambda})\;\;\text{on }b\Omega\cap U. (7.14)

Moreover, we compute on bΩUb\Omega\cap U

0\displaystyle 0 =L¯L(|ρ|2)=L¯(Qρ(L,N))+L¯(Hρ(L,N))\displaystyle=\bar{L}L\left(|\partial\rho|^{2}\right)=\bar{L}\left(Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})\right)+\bar{L}\left(H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})\right)
=j,k,=123ρz¯zjzkL¯LjNk+Qρ(L¯L,N)+Qρ(L,L¯N)\displaystyle=\sum_{j,k,\ell=1}^{2}\frac{\partial^{3}\rho}{\partial\bar{z}^{\ell}\partial z^{j}\partial z^{k}}\bar{L}^{\ell}L^{j}N^{k}+Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({\nabla_{\bar{L}}L},{N})+Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{\nabla_{\bar{L}}N})
+j,k,=123ρz¯zjz¯kL¯LjN¯k+Hρ(L¯L,N)+Hρ(L,LN).\displaystyle+\sum_{j,k,\ell=1}^{2}\frac{\partial^{3}\rho}{\partial\bar{z}^{\ell}\partial z^{j}\partial\bar{z}^{k}}\bar{L}^{\ell}L^{j}\bar{N}^{k}+H^{\scriptscriptstyle\mathbb{C}}_{\rho}({\nabla_{\bar{L}}L},{N})+H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{\nabla_{L}N}).

In view of (6.23), it follows from (7.11) that on bΩUb\Omega\cap U

2Re(j,k,=123ρz¯zjzkL¯LjNk)=2Re(N)Hρ(L,L)+𝒪(λ).\displaystyle 2\operatorname{Re}\left(\sum_{j,k,\ell=1}^{2}\frac{\partial^{3}\rho}{\partial\bar{z}^{\ell}\partial z^{j}\partial z^{k}}\bar{L}^{\ell}L^{j}N^{k}\right)=2\operatorname{Re}(N)H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})+\mathcal{O}(\sqrt{\lambda}).

Moreover, it follows from (7.11) that

|Hρ(L,LN)|=𝒪(λ)on bΩU.\left|H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{\nabla_{L}N})\right|=\mathcal{O}(\sqrt{\lambda})\quad\text{on }b\Omega\cap U.

Furthermore, since Hρ(L,L)=(L¯L)ρH^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})=-(\nabla_{\bar{L}}{L})\rho on bΩUb\Omega\cap U, it follows that the normal component of L¯L\nabla_{\bar{L}}{L} is 𝒪(Hρ(L,L))\mathcal{O}(H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})) on bΩUb\Omega\cap U. This, together with (7.11) and (7.14), implies that

|Qρ(L¯L,N)|\displaystyle\left|Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({\nabla_{\bar{L}}L},{N})\right| =𝒪(λ)on bΩU, and\displaystyle=\mathcal{O}(\sqrt{\lambda})\;\;\text{on }b\Omega\cap U,\text{ and}
|Hρ(L¯L,N)|\displaystyle\left|H^{\scriptscriptstyle\mathbb{C}}_{\rho}({\nabla_{\bar{L}}L},{N})\right| =𝒪(λ)on bΩU.\displaystyle=\mathcal{O}(\sqrt{\lambda})\;\;\text{on }b\Omega\cap U.

Therefore we obtain

2Re(N)Hρ(L,L)=Qρ(L,L¯N)+𝒪(λ)on bΩU.\displaystyle 2\operatorname{Re}(N)H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})=-Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{\nabla_{\bar{L}}N})+\mathcal{O}(\sqrt{\lambda})\;\;\text{on }b\Omega\cap U.

Since

L¯N=2j,k=12L¯kρz¯jz¯kzj,\nabla_{\bar{L}}N=2\sum_{j,k=1}^{2}\bar{L}^{k}\rho_{\bar{z}^{j}\bar{z}^{k}}\frac{\partial}{\partial z^{j}},

it follows easily that Qρ(L,L¯N)=|Qρ(L,.)|2Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{\nabla_{\bar{L}}N})=|Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{.\,})|^{2}, where Qρ(L,.)𝒱(U)1,0Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{.\,})\in\mathcal{V}({U})^{1,0} is defined via the condition Qρ(L,.),V:=Qρ(L,V)\langle Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{.\,}),V\rangle\mathrel{\mathop{:}}=Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{V}) for all V𝒱(U)1,0V\in\mathcal{V}({U})^{1,0}. Hence

νHρ(L,L)=2Re(N)Hρ(L,L)=|Qρ(L,.)|2+𝒪(λ)on bΩU,\displaystyle\nu H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})=2\operatorname{Re}(N)H^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})=-|Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{.\,})|^{2}+\mathcal{O}(\sqrt{\lambda})\;\;\text{on }b\Omega\cap U,

i.e., (7.12) holds and the claim follows. ∎

A global analog of Proposition 7.10 easily follows for sesquiconvex, bounded domains.

Corollary 7.15.

If Ω2\Omega\Subset\mathbb{C}^{2} is sesquiconvex, then Ω\Omega admits a smooth global defining function ρ:U\rho:U\longrightarrow\mathbb{R} which is plurisubharmonic on Ω¯U\bar{\Omega}\cap U.

Remark 7.16.

Let Ω2\Omega\subset\mathbb{C}^{2} be a smoothly bounded domain, let N=NρN=N_{\rho} for some smooth local defining function ρ:U\rho\colon U\to\mathbb{R} for Ω\Omega, see (2.6), and let LL be a nonvanishing holomorphic tangential vector field on UU. A straightforward computation shows that then

L¯N=Qρ(L,L)¯2|L|2L+Qρ(L,N)¯2|N|2Non bΩU,\displaystyle\nabla_{\bar{L}}N=\frac{\overline{Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})}}{2|L|^{2}}L+\frac{\overline{Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})}}{2|N|^{2}}N\;\;\text{on }b\Omega\cap U,

and hence, in particular,

Qρ(L,L¯N)=|Qρ(L,L)|22|L|2+|Qρ(L,N)|22|N|2on bΩU.\displaystyle Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{\nabla_{\bar{L}}N})=\frac{\lvert Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})\rvert^{2}}{2|L|^{2}}+\frac{\lvert Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{N})\rvert^{2}}{2|N|^{2}}\;\;\text{on }b\Omega\cap U.

Moreover, it is easy to see that if Ω\Omega is \mathbb{C}-convex at p0bΩUp_{0}\in b\Omega\cap U, then |Qρ(L,L)|=𝒪(λ)|Q^{\scriptscriptstyle\mathbb{C}}_{\rho}({L},{L})|=\mathcal{O}(\lambda) on bΩb\Omega near p0p_{0}. In view of (7.14), it thus follows from the arguments in the proof of Proposition 7.10, that Ω\Omega admits a plurisubharmonic smooth local defining function near p0p_{0} whenever Ω\Omega is both sesquiconvex and \mathbb{C}-convex at p0p_{0}. Similarly, if Ω2\Omega\Subset\mathbb{C}^{2} is sesquiconvex and \mathbb{C}-convex, then Ω\Omega admits a plurisubharmonic smooth defining function.

References

  • [1] Behrens, M. Plurisubharmonic defining functions of weakly pseudoconvex domains in 𝐂2{\bf C}^{2}. Math. Ann. 270, 2 (1985), 285–296.
  • [2] Bloom, T. Remarks on type conditions for real hypersurfaces in 𝐂n{\bf C}^{n}. In Several complex variables (Cortona, 1976/1977). Scuola Norm. Sup. Pisa, Pisa, 1978, pp. 14–24.
  • [3] Diederich, K., and Fornæss, J. E. Pseudoconvex domains: bounded strictly plurisubharmonic exhaustion functions. Invent. Math. 39, 2 (1977), 129–141.
  • [4] Diederich, K., and Fornæss, J. E. Pseudoconvex domains: an example with nontrivial Nebenhülle. Math. Ann. 225, 3 (1977), 275–292.
  • [5] Federer, H. Curvature measures. Trans. Amer. Math. Soc. 93 (1959), 418–491.
  • [6] Fornæss, J. E. Plurisubharmonic defining functions. Pacific J. Math. 80, 2 (1979), 381–388.
  • [7] Fornæss, J. E., and Herbig, A.-K. A note on plurisubharmonic defining functions in 𝐂2{\bf C}^{2}. Math. Z. 257 (2007), 285–296.
  • [8] Fornæss, J. E., and Herbig, A.-K. A note on plurisubharmonic defining functions in 𝐂n{\bf C}^{n}. Math. Ann. 342 (2008), 285–296.
  • [9] Gilbarg, D., and Trudinger, N. Elliptic partial differential equations of second order, 2nd Ed., vol. 224 of Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, New York, 1983.
  • [10] Herbig, A.-K., and McNeal, J. D. Convex defining functions for convex domains. J. Geom. Anal. 22, 2 (2012), 433–454.
  • [11] Kohn, J. J. Boundary behavior of ¯\bar{\partial} on weakly pseudo-convex manifolds of dimension two. J. Differential Geometry 6 (1972), 523–542.
  • [12] Kolář, M. On local convexifiability of type four domains in 2\mathbb{C}^{2}. In Differential geometry and applications. Proceedings of the 7th international conference, DGA 98, and satellite conference of ICM in Berlin, Brno, Czech Republic, August 10–14, 1998. Brno: Masaryk University, 1999, pp. 361–370.
  • [13] Morrow, J., and Rossi, H. Some theorems of algebraicity for complex spaces. J. Math. Soc. Japan 27 (1975), 167–183.
  • [14] Morrow, J., and Rossi, H. Correction to: “Some theorems of algebraicity for complex spaces”. J. Math. Soc. Japan 29 (1977), 783.
  • [15] Range, R. M. A remark on bounded strictly plurisubharmonic exhaustion functions. Proc. Amer. Math. Soc. 81, 2 (1981), 220–222.