It is a well-known consequence of the classical uniformization theorem that there is a metric with constant Gaussian curvature in each conformal class of any compact Riemann surface. It is natural to ask how to generalize this classical uniformization theory to compact surfaces with conical singularities and with nonempty boundary, or to find a "best metric" on such surfaces.
Instead of using metrics of constant curvature, Chen[1-2] started to use the extremal Hermitian metrics to generalize the classical uniformization theory to Riemann surfaces with finite conical singularities. On any football there is at least an extremal Hermitian metric and it was claimed that there is at least an extremal Hermitian metric on any surface with boundary (see Ref. [2]).
Wang and Zhu[3] discussed extremal Hermitian metrics with finite energy and area on Riemann surfaces with conical singularities and provided a classification of Kg for M with all singularities satisfying small enough angles. The primary challenge in obtaining this classification arises from the fact that the Gaussian curvature Kg may become unbounded near singular points. This challenge was overcome by studying the asymptotic behavior of the metrics near the singular points, along with conducting numerous meticulous analyzes of the metrics.
During above analytic process, Proposition 2.1 in Ref.[3], an accurate estimate of the conical structure of metrics, is necessary to obtain the asymptotic behavior of the Gaussian curvature and the estimate of the second covariant derivative of curvature. While showing that the Gaussian curvature can be extended continuously to singular points with small enough angles (specifically 2παi, provided that αi≤1 for any i), Wang and Zhu[3] combined the asymptotic behavior of the first derivatives of curvature and the holomorphicity of the second covariant derivative of curvature, and the proposition played an important role as a prerequisite. Also, without studying the conical structure of metrics, the approach to establish the estimate of the second covariant derivative of curvature will be invalid.
However, the proof of Proposition 2.1 given by Wang and Zhu remains several gaps that are eventually correct but lack essential explanation, which may create some obstacles for readers to understand. The main purpose of this paper is to make amends for the neglect and to provide a specific computation to verify such an estimate.
We now state Theorem A, i.e. Proposition 2.1 in Ref.[3], and Theorem B, a weaker estimate given by Chen[2] that will be used to support the proof of Theorem A.
Let M be a compact Riemann surface with nonempty boundary ∂M. For any Hermitian metric g0 on M, consider the set
$ E(g) = \int\limits_M K_g^2 \mathrm{d}g, $ |
where Kg is the Gaussian curvature of g and dg stands for the volume element. A critical point of this function is called an extremal Hermitian metric.
The Euler-Lagrange equation of this function is
$ \Delta_g K_g+K_g^2=C, $ | (1) |
for some constant C. Equivalently, in a local complex coordinate chart,
$ \frac{\partial}{\partial \bar{z}} K_{g, z z}=0, $ | (2) |
where Kg, zz is the second-order covariant derivative of Kg. The equations (1) and (2) are equivalent (see Ref.[1]). In particular, a metric with curvature satisfying Kg, zz=0 is called HCMU (this means that the Hessian of the Curvature of the Metric is Umbilical). There are examples of HCMU metrics that are not metrics of constant curvature in Ref.[1] and Wu[5] studied the character 1-form of an HCMU metric.
Let D be a disk centered at the origin. Suppose that
$ \left\{\begin{array}{l} \Delta K=-K^2 \mathrm{e}^{2 \psi}+C \mathrm{e}^{2 \psi} , \\ \Delta \psi=-K \mathrm{e}^{2 \psi} . \end{array}\right. $ | (3) |
for some constant C.
Wang and Zhu considered the conical structure of extremal Hermitian metrics.
Theorem A (Proposition 2.1[3]). Let
$ \begin{aligned} \lim _{x \rightarrow 0}|x| K(x) \mathrm{e}^{\psi(x)}=0 . \end{aligned} $ |
The above estimate is an improved result of the following estimate given by Chen.
Theorem B (Theorem 2[2]). Let
$ \begin{aligned} \lim _{x \rightarrow 0}|x|^{2} K(x) \mathrm{e}^{2 \psi(x)}=0. \end{aligned} $ |
To begin with, a technical method frequently appears in this paper and it is generalized as the following proposition, which shows an equivalent statement involving certain uniformity to continuity.
Proposition 1.1 For a given point
1)
2)
$ \mathit{Ω} \cap\left\{x_{0}+t y \mid t \in \mathbb{R}\right\} \neq \varnothing . $ |
Then f is continuous at x0 if and only if for an arbitrary sequence of real numbers ri→0,
$ \begin{aligned} \lim _{i \rightarrow \infty} f\left(x_{0}+r_{i} x\right)=f\left(x_{0}\right), \end{aligned} $ |
uniformly for
Proof Without loss of generality, we can set x0=0.
We first prove the necessity. Denote
$ |f(x)-f(0)|<\varepsilon . $ |
Taking an arbitrary sequence
$ \left|r_{i}\right|<\frac{\delta}{d} . $ |
Therefore, for all
$ \left|f\left(r_{i} x\right)-f(0)\right|<\varepsilon, $ |
which completes one direction of the proposition.
For sufficiency, since
$ \lambda_{i} \frac{x_{i}}{\left|x_{i}\right|} \in \mathit{Ω} . $ |
The existence is guaranteed by the property 2) of
$ \left|r_{i}\right|<\frac{\left|x_{i}\right|}{\varepsilon} \rightarrow 0 \quad \text { as } i \rightarrow \infty . $ |
Hence, by the uniformity, as
$ \begin{aligned} \left|f\left(x_{i}\right)-f(0)\right| & =\left|f\left(r_{i} \cdot \lambda_{i} \frac{x_{i}}{\left|x_{i}\right|}\right)-f(0)\right| \\ & \leqslant \sup _{x \in \mathit{Ω}}\left|f\left(r_{i} x\right)-f(0)\right| \rightarrow 0 . \end{aligned} $ |
For the arbitrariness of
Remark 1.1 From the proof, the necessity holds for more general Ω without property 2).
Most uniformity in this paper is from Proposition 1.1, so readers should keep this in mind when encountering uniformity in the sequel.
Next, we will introduce the Newtonian potential (also called Poisson potential) and show some basic properties of it in the context of
Theorem 1.1 (Corollary 2.25[6]). Let
$ \|u * v\|_{r^{\prime}} \leqslant\|u\|_{p^{\prime}}\|v\|_{q^{\prime}}. $ |
Denote
$ \mathit{\Gamma}(x)=\frac{1}{2 \pi} \ln |x| \quad \text { on } \mathbb{R}^{2}-\{0\}, $ |
known as the fundamental solution of the Laplacian equation and in the sequel. For an integrable function f, the integral
$ \mathit{\Gamma} * f(x)=\int_{\mathbb{R}^{2}} \mathit{\Gamma}(x-y) f(y) \mathrm{d} y $ |
is called the Newtonian potential of f. Then we give the following proposition which claims that the Laplacian (in the weak sense) of the Newtonian potential is equal to the function f itself almost everywhere.
Proposition 1.2 Suppose that Ω is a bounded domain of
$ \int_{\mathit{Ω}}(\mathit{\Gamma} * f) \cdot \Delta \phi \mathrm{d} x=\int_{\mathit{Ω}} f \phi \mathrm{~d} x. $ |
Proof First, we prove that
$ \begin{aligned} \int_{\mathit{Ω}}|\mathit{\Gamma} * f| \mathrm{d} x & \leqslant \int_{\mathit{Ω}}|f(y)| \int_{\mathit{Ω}}|\mathit{\Gamma}(x-y)| \mathrm{d} x \mathrm{~d} y \\ & \leqslant G \int_{\mathit{Ω}}|f(y)| \mathrm{d} y=G\|f\|_{L^{1}(\mathit{Ω})} <\infty. \end{aligned} $ |
Thus, Γ*f is well-defined as an integrable function.
To calculate the Laplacian of f, we use the approximation of f (see Corollary 2.30[6]). Specifically, a sequence
$ f_{i}= \begin{cases}f_{i} & \text { on } \mathit{Ω}, \\ 0 & \text { on } \mathbb{R}^{2}-\mathit{Ω},\end{cases} $ |
and
$ f= \begin{cases}f & \text { on } \mathit{Ω}, \\ 0 & \text { on } \mathbb{R}^{2}-\mathit{Ω}.\end{cases} $ |
We shall still denote the above extensions by
$ \begin{aligned} \left\|\mathit{\Gamma} * f_{i}-\mathit{\Gamma} * f\right\|_{L^{1}(\mathit{Ω})} & \leqslant\left\|\chi_{r} \mathit{\Gamma}\right\|_{L^{1}\left(\mathbb{R}^{2}\right)} \cdot\left\|f_{i}-f\right\|_{L^{1}\left(\mathbb{R}^{2}\right)} \\ & =\|\mathit{\Gamma}\|_{L^{1}\left(B_{r}\right)} \cdot\left\|f_{i}-f\right\|_{L^{1}(\mathit{Ω})} \rightarrow 0. \end{aligned} $ |
Thus,
$ \begin{aligned} \lim _{i \rightarrow \infty} \int_{\mathit{Ω}} f_{i} g \mathrm{~d} x=\int_{\mathit{Ω}} f g \mathrm{~d} x \end{aligned} $ |
and
$ \begin{aligned} \lim _{i \rightarrow \infty} \int_{\mathit{Ω}}\left(\mathit{\Gamma} * f_{i}\right) \cdot g \mathrm{~d} x=\int_{\mathit{Ω}}(\mathit{\Gamma} * f) \cdot g \mathrm{~d} x. \end{aligned} $ |
On the other hand, since
$ \begin{aligned} \Delta\left(\mathit{\Gamma} * f_{i}\right)=f_{i} . \end{aligned} $ |
Therefore, for any test function
$ \int_{\mathit{Ω}}\left(\mathit{\Gamma} * f_{i}\right) \cdot \Delta \phi \mathrm{d} x=\int_{\mathit{Ω}} f_{i} \phi \mathrm{~d} x. $ |
Thus, for any test function
$ \begin{aligned} \int_{\mathit{Ω}}(\mathit{\Gamma} * f) \cdot \Delta \phi \mathrm{d} x & =\lim _{i \rightarrow \infty} \int_{\mathit{Ω}}\left(\mathit{\Gamma} * f_{i}\right) \cdot \Delta \phi \mathrm{d} x \\ & =\lim _{i \rightarrow \infty} \int_{\mathit{Ω}} f_{i} \cdot \phi \mathrm{~d} x \\ & =\int_{\mathit{Ω}} f \phi \mathrm{~d} x. \end{aligned} $ |
Corollary 1.1 Let D be a disk at the origin
and
$ \Delta(\mathit{\Gamma} * f)=f \quad \text { on } \quad D^{*}. $ |
Proof Take r>diam(D) for Br. Denoting the zero extensions of f on
$ \|\mathit{\Gamma} * f\|_{L^{2}(D)} \leqslant\|\mathit{\Gamma}\|_{L^{2}\left(B_{r}\right)}\|f\|_{L^{1}(D)}<\infty . $ |
Thus,
$ \int_{D}(\mathit{\Gamma} * f) \cdot \Delta \phi \mathrm{d} x=\int_{D} f \phi \mathrm{~d} x. $ |
Then by Weyl's Lemma (see Proposition 28[8]), we have:
$ \Delta(\mathit{\Gamma} * f)=f \quad \text { on } \quad D^{*} . $ |
One of the gaps in Chen's previous proof for Theorem B is that they lacked enough explanation on how to restrain the Lp-norm of a function by the L1-norm of its Laplacian. The next proposition given by us will provide concrete evidence. To prove the proposition, we list some related theorems.
The following result proved by Gilbarg and Trudinger embraces a special case of the Calderon-Zygmund inequality.
Theorem 1.2 (Theorem 9.9[9]). Let Ω be a bounded domain in
$ \begin{aligned} \left\|D^2 w\right\|_p \leqslant C\|f\|_p, \end{aligned} $ |
where C depends only on n and p.
Notice that p above is required to be larger than 1, our proposition 1.2 indicates that the above inequality may not hold but the equality of f and the Laplacian of its Newtonian potential remains true as for p=1.
Theorem 1.3 (Theorem 9.11[9]). Let Ω be an open set in
$ \begin{aligned} & a^{i j} \in C^{0}(\mathit{Ω}), \quad b^{i}, c \in L^{\infty}(\mathit{Ω}), \quad f \in L^{p}(\mathit{Ω}) ; \\ & a^{i j} \xi_{i} \xi_{j} \geqslant \lambda|\xi|^{2} \quad \forall \xi \in \mathbb{R}^{n} ; \\ & \left|a^{i j}\right|, \left|b^{i}\right|, |c| \leqslant \Lambda, \end{aligned} $ |
where
$ \|u\|_{2, p ; \mathit{Ω}^{\prime}} \leqslant C\left(\|u\|_{p ; \mathit{Ω}}+\|f\|_{p ; \mathit{Ω}}\right), $ |
where
This gives interior estimates for strong solutions (not classical solutions) of second-order elliptic equations. For more details, see Chapter 9 in Ref.[9].
The proof of our proposition also requires the Sobolev embedding theorem. To state the embedding theorem, we suppose that Ω is a domain in
Definition 1.1 A domain Ω satisfies the cone condition if there exists a finite cone C such that each x∈Ω is the vertex of a finite cone Cx contained in Ω and congruent to C.
For a domain
$ \begin{aligned} \|\phi\|_{C_{B}^{m}(\mathit{Ω})}=\max _{0 \leqslant \alpha \leqslant m} \sup _{x \in \mathit{Ω}}\left|D^{\alpha} \phi(x)\right| . \end{aligned} $ |
The following theorem is known as the Sobolev embedding theorem.
Theorem 1.4 (Theorem 4.12[6]). Let Ω be a domain in
$ W^{j+m, p}(\mathit{Ω}) \hookrightarrow C_{B}^{j}(\mathit{Ω}) . $ |
Moreover,
$ W^{m, p}(\mathit{Ω}) \hookrightarrow L^{q}(\mathit{Ω}) \quad \text { for } p \leqslant q \leqslant \infty . $ |
For Sobolev space
Now we are ready to state our another analytic result required to prove Theorem A and present the proof.
Proposition 1.3 Let Ω be a bounded open subset of
$ \left\|F_{i}\right\|_{L^{2}(\mathit{Ω})} \rightarrow 0 \quad \text { and } \quad\left\|\Delta F_{i}\right\|_{L^{1}(\mathit{Ω})} \rightarrow 0 , $ |
as
$ \left\|F_{i}\right\|_{L^{p}\left(\mathit{Ω}^{\prime}\right)} \rightarrow 0 \quad \text { as } i \rightarrow \infty . $ |
Proof For each
$ F_{i}=u_{i}+v_{i} \quad \text { on } \quad \mathit{Ω}, $ |
where
$ \widetilde{u_{i}}(x)=\left(\chi_{r} \mathit{\Gamma}\right) * \widetilde{f_{i}}(x), \quad x \in \mathbb{R}^{2}. $ |
Applying Theorem 1.1, the Young's convolution inequality, with
$ \left\|{\widetilde{u_{i}}}\right\|_{L^{p}(\mathit{Ω})} \leqslant\|\mathit{\Gamma}\|_{L^{p}\left(B_{r}\right)}\left\|f_{i}\right\|_{L^{1}(\mathit{Ω})} \leqslant c_{1}\left\|f_{i}\right\|_{L^{1}(\mathit{Ω})}. $ |
Taking
$ \left\|\widetilde{u_{i}}\right\|_{L^{p}(\mathit{Ω})} \leqslant c_{1}\left\|f_{i}\right\|_{L^{1}(\mathit{Ω})}=c_{1}\left\|\Delta F_{i}\right\|_{L^{1}(\mathit{Ω})} \rightarrow 0. $ |
On the other hand, denote
$ \Delta u_{i}=f_{i} . $ |
Due to the definition of
$ \widetilde{u_{i}}=u_{i} \quad \text { on } \quad \mathit{Ω} . $ |
Thus,
$ \begin{aligned} \left\|v_{i}\right\|_{L^{2}(\mathit{Ω})} & =\left\|F_{i}-u_{i}\right\|_{L^{2}(\mathit{Ω})} \\ & \leqslant\left\|F_{i}\right\|_{L^{2}(\mathit{Ω})}+\left\|u_{i}\right\|_{L^{2}(\mathit{Ω})} \rightarrow 0 . \end{aligned} $ |
By Theorem 1.3, where p=2, the operator
$ \left\|v_{i}\right\|_{H^{2}\left(\mathit{Ω}^{\prime}\right)} \leqslant c_{2}\left\|v_{i}\right\|_{L^{2}(\mathit{Ω})} . $ |
Therefore, as
$ \left\|v_{i}\right\|_{H^{2}\left(\mathit{Ω}^{\prime}\right)} \rightarrow 0 . $ |
Furthermore, by Theorem 1.4, the Sobolev embedding theorem, we have
$ \left\|F_{i}\right\|_{L^{p}\left(\mathit{Ω}^{\prime}\right)} \leqslant\left\|u_{i}\right\|_{L^{p}\left(\mathit{Ω}^{\prime}\right)}+\left\|v_{i}\right\|_{L^{p}\left(\mathit{Ω}^{\prime}\right)} \rightarrow 0 . $ |
Now we will show the complete proofs of Theorem A and Theorem B and explain how we apply our results to the proofs of Theorem A and B.
First, we introduce some notations. For any given
$ \begin{gathered} \psi_i(y)=\psi\left(\left|x_i\right| y\right)+\ln \left|x_i\right| \quad \text { and } \\ K_i(y)=K\left(\left|x_i\right| y\right) . \end{gathered} $ | (4) |
By equation (3),
$ \Delta \psi_{i}=-K_{i} \mathrm{e}^{2 \psi_{i}} . $ |
Without other declaration,
It is necessary to figure out the asymptotic behavior of metrics ψi, given by those coordinate transformations.
Lemma 2.1 (Theorem 2[4]). Let
$ \lim _{i \rightarrow \infty} \psi_{i}(y)=-\infty $ |
uniformly for
The following lemma is an improved result of Chen's estimate (see Ref.[2]).
Lemma 2.2 (Lemma 1.1[3]). Let
$ \begin{aligned} \lim _{r \rightarrow 0} \sup _{x, x^{\prime} \in r T_{1}} \frac{\psi(x)+\ln |x|}{\psi\left(x^{\prime}\right)+\ln \left|x^{\prime}\right|}=1 . \end{aligned} $ |
Although the proof of Theorem B was shown by Chen[2], it omitted several important steps. Here we reprove this estimate and will fix all the gaps.
Proof of Theorem B For any sequence of points
$ \begin{aligned} \lim _{i \rightarrow \infty}\left|K\left(x_{i}\right)\right| \mathrm{e}^{2\left(\psi\left(x_{i}\right)+\ln \left|x_{i}\right|\right)}=0 . \end{aligned} $ |
For each
$ \Delta K_{i}(y)+K_{i}(y)^{2} \mathrm{e}^{2 \psi_{i}(y)}=C \mathrm{e}^{2 \psi_{i}(y)} . $$ $ |
Set
$ a_{i}=\inf _{y \in T_{1}} \psi_{i}(y) \quad \text { and } \quad \tilde{K}_{i}(y)=K_{i}(y) \mathrm{e}^{a_{i}} . $ |
Then
$ \left\|\widetilde{K}_{i}^{2}\right\|_{L^{2}\left(T_{1}\right)}=\int_{T_{1}} K_{i}^{2} \mathrm{e}^{2 a_{i}} \mathrm{~d} y \leqslant \int_{T_{1}} K_{i}^{2} \mathrm{e}^{2 \psi_{i}} \mathrm{~d} y . $ |
Since the original metric
$ \begin{aligned} \lim _{i \rightarrow \infty} \int_{T_{1}} K_{i}^{2} \mathrm{e}^{2 \psi_{i}} \mathrm{~d} y=\lim _{i \rightarrow \infty} \int_{\left|x_{i}\right| T_{1}} K^{2} \mathrm{e}^{2 \psi} \mathrm{~d} x=0 . \end{aligned} $ |
Thus as
$ \left\|\tilde{K}_i\right\|_{L^2\left(T_1\right)} \rightarrow 0 . $ | (5) |
On the other hand,
$ \Delta \tilde{K}_{i}(y)+\tilde{K}_{i}(y) K_{i}(y) \mathrm{e}^{2 \psi_{i}(y)}=C \mathrm{e}^{a_{i}} \mathrm{e}^{2 \psi_{i}}, \quad i \in \mathbb{N}, y \in T_{1}. $ |
Denote
$ a_{i}<b. $ |
Then we have: for sufficiently large i,
$ \begin{aligned} \int_{T_{1}}\left|f_{i}\right| \mathrm{d} y & \leqslant \mathrm{e}^{b}\left(\int_{T_{1}} K_{i}^{2} \mathrm{e}^{2 \psi_{i}} \mathrm{~d} y+|C| \int_{T_{1}} \mathrm{e}^{2 \psi_{i}} \mathrm{~d} y\right) \\ & =\mathrm{e}^{b}\left(\int_{\left|x_{i}\right| T_{1}} K^{2} \mathrm{e}^{2 \psi} \mathrm{~d} x+|C| \int_{\left|x_{i}\right| T_{1}} \mathrm{e}^{2 \psi} \mathrm{~d} x\right). \end{aligned} $ |
Therefore, with the original metric
$ \left\|\Delta \tilde{K}_{i}\right\|_{L^{1}\left(T_{1}\right)}=\int_{T_{1}}\left|f_{i}\right| \mathrm{d} y \rightarrow 0 $ | (6) |
By the Hölder inequality, for any 1 < s < 2,
$ \left\|f_{i}\right\|_{L^{s}\left(T_{1 / 2}\right)} \leqslant\left\|\tilde{K}_{i}\right\|_{L^{\frac{2 s}{2-s}}\left(T_{1 / 2}\right)}\left\|K_{i} \mathrm{e}^{2 \psi_{i}}\right\|_{L^{2}\left(T_{1 / 2}\right)}+ \\ \left\|C \mathrm{e}^{a_{i}} \mathrm{e}^{2 \psi_{i}}\right\|_{L^{s}\left(T_{1 / 2}\right)} $ | (7) |
Denote
$ \left\|\tilde{K}_{i}\right\|_{L^{p}\left(T_{1 / 2}\right)} \rightarrow 0 . $ | (8) |
For sufficiently large i, with the original metric having finite energy and area, we have the following limits
$ \begin{aligned} & \left\|K_{i} \mathrm{e}^{2 \psi_{i}}\right\|_{L^{2}\left(T_{1 / 2}\right)}^{2} \leqslant \mathrm{e}^{2 b} \int_{\left|x_{i}\right| T_{1 / 2}} K^{2} \mathrm{e}^{2 \psi} \mathrm{~d} x \rightarrow 0, \\ & \left\|C \mathrm{e}^{a_{i}} \mathrm{e}^{2 \psi_{i}}\right\|_{L^{s}\left(T_{1 / 2}\right)}^{S} \leqslant C \mathrm{e}^{4 b} \int_{T_{1 / 2}} \mathrm{e}^{2 \psi_{i}} \mathrm{~d} y \rightarrow 0. \end{aligned} $ |
Therefore, we have estimated all the terms in (7) and obtain that as
$ \left\|\tilde{K}_{i}\right\|_{L^{s}\left(T_{1 / 2}\right)}=\left\|f_{i}\right\|_{L^{s}\left(T_{1 / 2}\right)} \rightarrow 0 . $ | (9) |
For 1 < s < 2, by Theorem 1.3 with
$ \left\|\tilde{K}_{i}\right\|_{W^{2, s}\left(T_{1 / 4}\right)} \leqslant c_{3}\left(\left\|\tilde{K}_{i}\right\|_{L^{s}\left(T_{1 / 4}\right)}+\left\|f_{i}\right\|_{L^{s}\left(T_{1 / 4}\right)}\right) \rightarrow 0 . $ |
In particular, by Theorem 1.4, the embedding theorems,
$ \begin{aligned} \sup _{y \in T_{1 / 4}}\left|\tilde{K}_i(y)\right| \rightarrow 0. \end{aligned} $ | (10) |
From Lemma 2.2, we have
$ \begin{aligned} & \lim _{i \rightarrow+\infty} \sup _{x, x^{\prime} \in\left|x_{i}\right| T_{1}}\left|\frac{\psi(x)+\ln |x|}{\psi\left(x^{\prime}\right)+\ln \left|x^{\prime}\right|}\right| \\ = & \lim _{i \rightarrow+\infty} \sup _{x, x^{\prime} \in T_{1}}\left|\frac{\psi\left(\left|x_{i}\right| x\right)+\ln \left|x_{i}\right|}{\psi\left(\left|x_{i}\right| x^{\prime}\right)+\ln \left|x_{i}\right|}\right| \\ = & \lim _{i \rightarrow+\infty} \sup _{x, x^{\prime} \in T_{1}}\left|\frac{\psi_{i}(x)}{\psi_{i}\left(x^{\prime}\right)}\right|=1 . \end{aligned} $ |
Thus,
$ a_{i} \geqslant 2 \psi_{i}(y), \quad \forall y \in T_{1} . $ |
Let
$ \left|x_{i}\right|^{2}\left|K\left(x_{i}\right)\right| \mathrm{e}^{2 \psi\left(x_{i}\right)}=\left|K_{i}\left(y_{i}\right)\right| \mathrm{e}^{2 \psi_{i}\left(y_{i}\right)} \leqslant\left|K_{i}\left(y_{i}\right)\right| \mathrm{e}^{a_{i}}. $ |
Since
$ \left|K_{i}\left(y_{i}\right)\right| \mathrm{e}^{a_{i}} \longrightarrow 0 \quad \text { as } i \rightarrow \infty . $ | (11) |
Chen[4] studied the behavior of the Newtonian potential satisfying certain asymptotic properties near the origin.
Lemma 2.3 Let
$ v(x)=\int_{D} \mathit{\Gamma}(x-y) f(y) \mathrm{d} y . $ |
Then the following two statements hold:
1)
2)
Define the oscillation of a function
$ \begin{aligned} \operatorname{Osc}_{x \in U}\{f(x)\}=\sup _{x \in U} f(x)-\inf _{x \in U} f(x) . \end{aligned} $ |
The following lemma is equivalent to Theorem 3[3] and shows that the metric is asymptotically rotationally symmetric.
Lemma 2.4 Let
$ \begin{aligned} \lim _{i \rightarrow \infty} \operatorname{Osc}_{\theta}\left\{\psi_{i}(r, \theta)\right\}=0, \end{aligned} $ |
uniformly for
Proof First from (3) and the Hölder inequality, we notice that
$ \begin{aligned} \int_{D-\{0\}}|\Delta \psi| \mathrm{d} x & =\int_{D-\{0\}}|K| \mathrm{e}^{2 \psi} \mathrm{~d} x \\ & \leqslant\left(\int_{D-\{0\}} K^{2} \mathrm{e}^{2 \psi} \mathrm{~d} x\right)^{1 / 2}\left(\int_{D-\{0\}} \mathrm{e}^{2 \psi} \mathrm{~d} x\right)^{1 / 2}, \end{aligned} $ |
i.e.
$ \begin{aligned} v(x)=\int_{D} \mathit{\Gamma}(x-y) \Delta \psi(y) \mathrm{d} y. \end{aligned} $ |
From Theorem B, we have
$|x|^{2} \Delta \psi(x)=-|x|^{2} K(x) \mathrm{e}^{2 \psi(x)} \rightarrow 0 \quad \text { as } |x| \rightarrow 0. $ |
Applying Lemma 2.3, we obtain
$ \begin{aligned} \lim _{|x| \rightarrow 0} \frac{v(x)}{\ln |x|}=0 \quad \text { and } \quad \lim _{r \rightarrow 0} \operatorname{Osc}_{\theta}\{v(r, \theta)\}=0 . \end{aligned} $ |
Let
$ \Delta u=0 \quad \text { on } D-\{0\}. $ |
Then from Proposition 1.1 in Ref.[3], we have
$ \begin{aligned} \lim _{|x| \rightarrow 0} \frac{u(x)}{\ln |x|}=\lim _{|x| \rightarrow 0} \frac{\psi(x)}{\ln |x|}=\alpha-1 . \end{aligned} $ |
Take
For
$ w(x)=\bar{u}(x)-\hat{u}(x)+\varepsilon \ln |x| $ |
which is harmonic on
$ \begin{aligned} \lim _{x \rightarrow 0} w(x)=-\infty. \end{aligned} $ |
By the maximum principle: for arbitrary small δ>0,
$ \begin{aligned} \sup _{B_{r}-\bar{B}_{\delta}} w=\sup _{\partial B_{r}} w=\varepsilon \sup _{\partial B_{r}} \ln |x| . \end{aligned} $ |
Taking
$ u(x)=(\alpha-1) \ln |x|+\hat{u}(x) . $ |
In particular, we have
$ \begin{aligned} \lim _{r \rightarrow 0} \operatorname{Osc}_{\theta}\{u(r, \theta)\}=\lim _{r \rightarrow 0} \operatorname{Osc}_{\theta}\{\hat{u}(r, \theta)\}=0 . \end{aligned} $ |
Then we get as r→0,
$ \operatorname{Osc}_{\theta}\{\psi(r, \theta)\} \leqslant \operatorname{Osc}_{\theta}\{u(r, \theta)\}+\operatorname{Osc}_{\theta}\{v(r, \theta)\} \rightarrow 0 . $ | (12) |
Define a function
$ g(t)=\left\{\begin{array}{l} 0, t \leqslant 0, \\ \operatorname{Osc}_{\theta}\{\psi(t, \theta)\}, t>0 . \end{array}\right. $ |
By (12), it is clear that g is continuous at 0. Then by Proposition 1.1 with
It was showed that x=0 is either a weak cusp or a weak conical singular point with angle 2πα>0 in Ref. [2], which is expressed as the following.
Lemma 2.5 If
$ \begin{aligned} \lim _{r \rightarrow 0} \frac{1}{2 \pi} \int_{0}^{2 \pi} \frac{\partial \psi}{\partial r}(r, \theta) \cdot r \mathrm{~d} \theta=\alpha-1 . \end{aligned} $ |
In particular, the singular point is called a cusp if
With plenty of preparation, we now give the proof of Theorem A.
Proof of Theorem A As in Theorem B, it suffices to prove that for any sequence
$ \begin{aligned} \lim _{\left|x_{i}\right| \rightarrow 0}\left|x_{i}\right|\left|K\left(x_{i}\right)\right| \mathrm{e}^{\psi\left(x_{i}\right)}=0. \end{aligned} $ |
From the proof of Theorem B, we have
$ \begin{aligned} \lim _{i \rightarrow \infty}\left|K\left(x_{i}\right)\right| \mathrm{e}^{a_{i}}=0, \end{aligned} $ | (13) |
where
$ \widetilde{\psi}_{i}(y)=\psi\left(\left|x_{i}\right| y\right)+(1-\alpha) \ln \left(\left|x_{i}\right||y|\right) \quad \text { on } T_{2}. $ |
By Lemma 2.5, the function g is continuous at 0, which is given by
$ g(t)=\left\{\begin{array}{l} 0, t \leqslant 0, \\ \frac{1}{2 \pi} \int_{0}^{2 \pi}\left(t \frac{\partial \psi(t, \theta)}{\partial t}+1-\alpha\right) \mathrm{d} \theta, t>0. \end{array}\right. $ |
Let
$ \frac{1}{2 \pi} \int_{0}^{2 \pi}\left(r^{\prime} \frac{\partial \psi\left(r^{\prime}, \theta\right)}{\partial r^{\prime}}+1-\alpha\right) \mathrm{d} \theta \rightarrow 0, $ |
uniformly for
$ \int_{0}^{2 \pi}\left(r^{\prime} \frac{\partial \psi\left(r^{\prime}, \theta\right)}{\partial r^{\prime}}+1-\alpha\right) \mathrm{d} \theta=\int_{0}^{2 \pi} r \frac{\partial \widetilde{\psi}_{i}(r, \theta)}{\partial r} \mathrm{~d} \theta, $ |
it follows that
$ \begin{aligned} \lim _{i \rightarrow \infty} \frac{1}{2 \pi} \int_{0}^{2 \pi} r \frac{\partial \widetilde{\psi}_{i}(r, \theta)}{\partial r} \mathrm{~d} \theta=0, \end{aligned} $ |
uniformly for
$ \begin{aligned} \lim _{i \rightarrow \infty} \frac{\partial}{\partial r} \int_0^{2 \pi} \widetilde{\psi}_i(r, \theta) \mathrm{d} \theta=0, \end{aligned} $ | (14) |
uniformly for
$ \begin{aligned} & \operatorname{Osc}_r\left\{\int_0^{2 \pi} \widetilde{\psi}_i(r, \theta) \mathrm{d} \theta\right\} \\ = & \max _r\left\{\int_0^{2 \pi} \widetilde{\psi}_i(r, \theta) \mathrm{d} \theta\right\}-\min _r\left\{\int_0^{2 \pi} \widetilde{\psi}_i(r, \theta) \mathrm{d} \theta\right\} \\ \leqslant & \sup _r\left\{\frac{\partial}{\partial r} \int_0^{2 \pi} \widetilde{\psi}_i(r, \theta) \mathrm{d} \theta\right\} \cdot 2 \mathrm{e} . \end{aligned} $ |
By (14), as
$ \operatorname{Osc}_{r}\left\{\int_{0}^{2 \pi} \widetilde{\psi}_{i}(r, \theta) \mathrm{d} \theta\right\} \rightarrow 0 . $ | (15) |
For any
$ \begin{gathered} \widetilde{\psi}_i\left(r, \theta_r^i\right)=\frac{1}{2 \pi} \int_0^{2 \pi} \widetilde{\psi}_i(r, \theta) \mathrm{d} \theta, \forall \mathrm{e}^{-1} \leqslant r \leqslant \mathrm{e}^1, \\ \widetilde{\psi}_i\left(\hat{r}_i, \theta_0\right)=\sup _r \widetilde{\psi}_i\left(r, \theta_0\right) \quad \text { and } \\ \widetilde{\psi}_i\left(\check{r}_i, \theta_0\right)=\underset{r}{\inf } \widetilde{\psi}_i\left(r, \theta_0\right) . \end{gathered} $ |
From Lemma 2.4, we know that
$ \begin{aligned} \lim _{i \rightarrow \infty} \operatorname{Osc}_{\theta}\left\{\widetilde{\psi}_{i}(r, \theta)\right\}=0 . \end{aligned} $ | (16) |
uniformly for
$ \begin{aligned} & \operatorname{Osc}_r\left\{\widetilde{\psi}_i\left(r, \theta_0\right)\right\} \\ &=\left(\widetilde{\psi}_i\left(\hat{r}, \theta_0\right)-\widetilde{\psi}_i\left(\hat{r}, \theta_{\hat{r}}^i\right)\right)+\left(\widetilde{\psi}_i\left(\check{r}, \theta_r^i\right)-\widetilde{\psi}_i\left(\check{r}, \theta_0\right)\right)+ \\ &\left(\widetilde{\psi}_i\left(\hat{r}, \theta_{\hat{r}}^i\right)-\widetilde{\psi}_i\left(\check{r}, \theta_r^i\right)\right) \\ & \leqslant \operatorname{Osc}_\theta\left\{\widetilde{\psi}_i(\hat{r}, \theta)\right\}+\operatorname{Osc}_\theta\left\{\widetilde{\psi}_i\left(\check{r}, \theta_0\right)\right\}+ \\ & \frac{1}{2 \pi} \operatorname{Osc}_r\left\{\int_0^{2 \pi} \widetilde{\psi}_i(r, \theta) \mathrm{d} \theta\right\}, \end{aligned} $ |
Combined (15), we have: as
$ \operatorname{Osc}_{r}\left\{\widetilde{\psi}_{i}\left(r, \theta_{0}\right)\right\} \rightarrow 0 . $ | (17) |
Furthermore, for each i, denote
$ \widetilde{\psi}_{i}\left(r_{1}, \theta_{1}\right)=\max _{(r, \theta) \in \bar{T}_{1}}\left\{\widetilde{\psi}_{i}(r, \theta)\right\} $ |
and
$ \widetilde{\psi}_{i}\left(r_{2}, \theta_{2}\right)=\min _{(r, \theta) \in \bar{T}_{1}}\left\{\widetilde{\psi}_{i}(r, \theta)\right\}, $ |
where
$ \begin{aligned} & \operatorname{Osc}_{r, \theta \in \bar{T}_1}\left\{\widetilde{\psi}_i(r, \theta)\right\} \\ = & \widetilde{\psi}_i\left(r_1, \theta_1\right)-\widetilde{\psi}_i\left(r_2, \theta_2\right) \\ = & \widetilde{\psi}_i\left(r_1, \theta_1\right)-\widetilde{\psi}_i\left(r_1, \theta_0\right)+\widetilde{\psi}_i\left(r_2, \theta_0\right)-\widetilde{\psi}_i\left(r_2, \theta_2\right)+ \\ & \widetilde{\psi}_i\left(r_1, \theta_0\right)-\widetilde{\psi}_i\left(r_2, \theta_0\right) \\ \leqslant & \operatorname{Osc}_\theta\left\{\widetilde{\psi}_i\left(r_1, \theta\right)\right\}+\operatorname{Osc}_\theta\left\{\widetilde{\psi}_i\left(r_2, \theta\right)\right\}+\operatorname{Osc}_r\left\{\widetilde{\psi}_i\left(r, \theta_0\right)\right\}, \end{aligned} $ |
by (16) and (17), it follows that
$ \begin{aligned} \lim _{i \rightarrow \infty} \operatorname{Osc}_{(r, \theta) \in \bar{T}_{1}}\left\{\widetilde{\psi}_{i}(r, \theta)\right\}=0 . \end{aligned} $ |
Consequently,
$ \begin{aligned} \operatorname{Osc}_{y \in T_{1}} \psi_{i}(y) & =\operatorname{Osc}_{y \in T_{1}}\left\{\widetilde{\psi}_{i}(y)+\alpha \ln \left(\left|x_{i}\right||y|\right)-\ln |y|\right\} \\ & \leqslant 3(\alpha+1), \end{aligned} $ |
when i is sufficiently large. In view of (5) and the above inequalities, we get
$ \begin{aligned} \lim _{\left|x_{i}\right| \rightarrow 0}\left|x_{i}\right|\left|K\left(x_{i}\right)\right| \mathrm{e}^{\psi\left(x_{i}\right)} & \leqslant \lim _{\left|x_{i}\right| 0}\left|K\left(x_{i}\right)\right|\left|\sup _{y \in T_{1}} \mathrm{e}^{\psi_{i}(y)}\right| \\ & \leqslant \lim _{\mid x_{i} \nmid 0} \mathrm{e}^{3(\alpha+1)}\left|K\left(x_{i}\right)\right| \mathrm{e}^{\alpha_{i}}=0 . \end{aligned} $ |
In this section, we will give an example to illustrate Theorem A. The HCMU metric is a special case of the extremal Hermitian metric. Chen et al.[10-11] have studied the HCMU metrics. They showed one can use the following method to construct HCMU metrics on a compact Riemann surface M. First one can consider the following equation on M
$ \left\{\begin{array}{l} \frac{\mathrm{d} K}{-\frac{1}{3}\left(K-K_{1}\right)\left(K-K_{2}\right)\left(K+K_{1}+K_{2}\right)}=\omega+\bar{\omega}, \\ K\left(p_{0}\right)=K_{0}, \end{array}\right. $ | (18) |
where ω is a meromorphic 1-form on M satisfying:
1) ω only has simple poles,
2) the residues of ω at poles are all real numbers,
3)
$ g=-\frac{4}{3}\left(K-K_{1}\right)\left(K-K_{2}\right)\left(K+K_{1}+K_{2}\right) \omega \bar{\omega} . $ | (19) |
One can prove g is an HCMU metric with finite area and finite energy and K is the Gauss curvature of g. Furthermore, the set of the singularities of g is a subset of the set of zeros and poles of ω. If
1) if p is a zero of ω, then p is a conical singularity with the conical angle
2) denote
3) denote
If
1) if
2) denote
3) if p is a pole of ω with the positive residue of ω, p is a cusp.
Next, we calculate the following limit
$ \begin{aligned} \lim _{z \rightarrow 0}|z| K e^{\psi}, \end{aligned} $ |
at any singularity of g.
If p is a conical singularity of g with the conical angle 2πα, then in a neighborhood of p
$ g=h|z|^{2 \alpha-2}|\mathrm{~d} z|^{2}=\mathrm{e}^{2 \psi}|\mathrm{~d} z|^{2}, $ |
where z is the local coordinate in a neighborhood of p with
$ |z| K \mathrm{e}^{\psi}=|z| K \sqrt{h}|z|^{\alpha-1}=K \sqrt{h}|z|^{\alpha} . $ |
Since both K and
$ \begin{aligned} \lim _{z \rightarrow 0}|z| K \mathrm{e}^{\psi}=\lim _{z \rightarrow 0} K \sqrt{h}|z|^{\alpha}=0 . \end{aligned} $ |
Therefore Theorem A holds in conical singularity case of HCMU metric.
If p is a cusp, then
$ g=-\frac{4}{3}\left(K-K_{1}\right)\left(K-K_{2}\right)^{2} \omega \bar{\omega} . $ |
Suppose
$ \begin{gathered} g=-\frac{4}{3}\left(K-K_{1}\right)\left(K-K_{2}\right)^{2} \frac{|\hat{f}(z)|^{2}}{|z|^{2}}|\mathrm{~d} z|^{2}= \\ \mathrm{e}^{2 \psi}|\mathrm{~d} z|^{2} . \end{gathered} $ |
Therefore
$ \mathrm{e}^{\psi}=\sqrt{-\frac{4}{3}\left(K-K_{1}\right)}\left(K-K_{2}\right) \frac{|\hat{f}(z)|}{|z|} . $ |
Then
$ |z| K \mathrm{e}^{\psi}=K \sqrt{-\frac{4}{3}\left(K-K_{1}\right)}\left(K-K_{2}\right)|\hat{f}(z)| . $ |
Since K is continuous at p with
$ \begin{aligned} =\lim _{z \rightarrow 0} K \sqrt{-\frac{4}{3}\left(K-K_{1}\right)}\left(K-K_{2}\right)|\hat{f}(z)|=0 . \end{aligned} $ |
Therefore Theorem A holds in cusp case of HCMU metric.
[1] |
Chen X X. Obstruction to the existence of metric whose curvature has umbilical hessian in a K-surface[J]. Communications in Analysis and Geometry, 2000, 8(2): 267-299. DOI:10.4310/CAG.2000.v8.n2.a2 |
[2] |
Chen X X. Extremal Hermitian metrics on Riemann surfaces[J]. Calculus of Variations and Partial Differential Equations, 1999, 8(3): 191-232. DOI:10.1007/s005260050123 |
[3] |
Wang G F, Zhu X H. Extremal Hermitian metrics on Riemann surfaces with singularities[J]. Duke Mathematical Journal, 2000, 104(2). DOI:10.1215/S0012-7094-00-10421-8 |
[4] |
Chen X X. Weak limits of Riemannian metrics in surfaces with integral curvature bound[J]. Calculus of Variations and Partial Differential Equations, 1998, 6(3): 189-226. DOI:10.1007/s005260050089 |
[5] |
Wu Y Y. On the character 1-form of an HCMU metric[J]. Journal of the Graduate School of the Chinese Academy of Sciences, 2009, 26(2): 185-193. DOI:10.7523/j.issn.2095-6134.2009.2.006 |
[6] |
Adams R A, Fournier J J F. Sobolev spaces[M]. 2nd ed. Amsterdam: Academic Press, 2003.
|
[7] |
Evans L C. Partial differential equations[M]. Providence, R.I.: American Mathematical Society, 1998.
|
[8] |
Donaldson S K. Riemann surfaces[M]. Oxford: Oxford University Press, 2011.
|
[9] |
Gilbarg D, Trudinger N S. Elliptic partial differential equations of second order[M]. Berlin, Heidelberg: Springer Berlin Heidelberg, 2001. DOI:10.1007/978-3-642-61798-0
|
[10] |
Chen Q, Wu Y Y. Character 1-form and the existence of an HCMU metric[J]. Mathematische Annalen, 2011, 351(2): 327-345. DOI:10.1007/s00208-010-0598-z |
[11] |
Chen Q, Wu Y Y, Xu B. On one-dimensional and singular Calabi's extremal metrics whose Gauss curvatures have nonzero umbilical Hessians[J]. Israel Journal of Mathematics, 2015, 208(1): 385-412. DOI:10.1007/s11856-015-1204-6 |