A SEMILINEAR PARTIAL DIFFERENTIAL EQUATION INDUCED BY HERMITIAN YANG-MILLS METRICS
School of Mathematical Sciences, Tongji University, Shanghai 200082, China
Received date: 2022-04-04; Accepted date: 2022-06-06
Foundation item: Supported by the National Natural Science Foundation of China (11701426).
Biography:
Li Yuxuan (1997-), male, born at Urumqi, Xinjiang, postgraduate, major in differentialgeometry. E-mail:
1653454@tongji.edu.cn.
1 Introduction
Let $ X $ be a Kähler manifold with a family of Kähler metrics $ \omega_\varepsilon $, and let $ V $ be a slope stable holomorphic vector bundle over $ X $. According to the Donaldson-Uhlenbeck-Yau theorem [1], $ V $ admits unique fully irreducible Hermitian-Yang-Mills metrics $ H_\varepsilon $ associated to each $ \omega_\varepsilon $. Similar to study the limiting behaviour Ricci flat metrics, Professor Jixiang Fu [2] studied the limiting behaviour of Hermitian Yang-Mills metrics $ H_\varepsilon $ when $ \omega_\varepsilon $ goes to a large Kähler metric limit. A critical step in [2] is to explicitly construct a family of Hermitian-Yang-Mills metrics by solving the following semilinear partial differential equation on unit ball $ B_1(0) $ of $ \mathbb{R}^2 $
$ \begin{equation} \left\{ \begin{aligned} \Delta u &=\varepsilon^{-2}\left(e^{u}-\left(x^{2}+y^{2}\right) e^{-u}\right) &{\text { in }} B_1(0),\\ u&=0 &{\text { on }} \partial B_1(0). \end{aligned} \right. \end{equation} $ |
(1.1) |
Here $ \varepsilon $ is a constant and $ (x,y) $ is the coordinate of $ \mathbb{R}^2 $. By the symmetry of the domain $ B_1(0) $ and using reference [3], Jixiang Fu proved the equation (1.1) has a unique radially symmetric solution. However, this method can not be applied to non-symmetric domain $ \Omega $ in $ \mathbb{R}^2 $.
In this paper, we first study the following equation defined a bounded connected domain $ \Omega\subset \mathbb{R}^2 $ with Dirichlet boundary value
$ \begin{equation} \left\{ \begin{aligned} \Delta u &=\varepsilon^{-2}\left(e^{u}-\left(x^{2}+y^{2}\right) e^{-u}\right) &{\text { in }} \Omega,\\ u&=g &{\text { on }} \partial \Omega. \end{aligned} \right. \end{equation} $ |
(1.2) |
For existence and uniqueness of the solution of (1.2), we have the following theorem.
Theorem 1.1 If $ \partial\Omega $ is $ C^{2,\alpha} $ and $ g\in C^{2,\alpha}(\partial\Omega) $, there is a unique solution $ u\in C^{2,\alpha}(\Omega) $ to equation (1.2). Especially if $ \partial\Omega $ and $ g $ are smooth, the solution $ u $ is smooth.
This theorem will give a Hermitian Yang-Mills metrics on a certain Kähler manifold given by [2]. On the other hand, the equation (1.1) can be defined on whole space $ \mathbb{R}^2 $. It is natural to explore whether the global solution of (1.1) is radially symmetric. The symmetry of global solutions of some semilinear equations has been investigated in [4] and [3] under the assumption $ u(x,y) $ decays to zero at a certain rate as $ r^2=x^2+y^2\to +\infty $. But they do not fit the equation(1.1) since one can see the global solution $ u $ is not bounded. Similar to [3, 4], by using moving plane method and maximum principle, we get the following theorem.
Theorem 1.2 For any given constant $ c $, if the global $ C^{2} $ solution $ u $ of
$ \begin{equation} \Delta u =\varepsilon^{-2}\left(e^{u}-\left(x^{2}+y^{2}\right) e^{-u}\right) \qquad {\text { in }} \mathbb{R}^2 \end{equation} $ |
(1.3) |
satisfies
$ u(s)-u(t) \rightarrow 0 \quad{\text { as}}\quad |s|,|t| \rightarrow \infty\quad {\text {and}}\quad |s|-|t|=c, $ |
then $ u $ is radially symmetric and $ \frac{\partial u}{\partial r} \geqslant0 $. Here $ s,t \in \mathbb{R}^2 $.
One may observe $ \frac{1}{2}\log(x^2+y^2) $ is a singular solution to the equation (1.3) and also satisfies $ \log (|s|)-\log(|t|)\to 0 $ as $ |t|-|s|=c $ and $ |s|, |t|\to \infty $, so the assumption of Theorem1.2 is natural and reasonable.
The next part of this paper will give the detailed proof of Theorem1.1 and Theorem1.2.
2 Existence of Solution of the Dirichlet Boundary Value Problem
In this section we will prove Theorem1.1. One can use Chapter 14 in [5] to show the existence of the equation 1.2 by using the variational method. Here we take Leray-Schauder existence theorem to prove it.
Let $ \Omega $ be a $ C^{2,\alpha} $ bounded domain in $ \mathbb{R}^2 $ and $ g\in C^{2,\alpha}(\partial\Omega) $ with $ \alpha\in (0,1 ) $. We first have a $ C^{0}(\Omega) $ estimate.
Lemma 2.1 Let $ \Phi $ be the $ C^{2,\alpha} $solution of Dirichlet boundary value problem
$ \begin{equation} \left\{ \begin{aligned} \Delta \Phi &=\varepsilon^{-2}(1-\left(x^{2}+y^{2}\right)) &{\text { in }} \Omega,\\ \Phi&=g &{\text { on }} \partial \Omega. \end{aligned} \right. \end{equation} $ |
(2.1) |
Then a solution u to (1.2) satisfy
$ \begin{equation} \sup\limits _{\bar\Omega} |u |\leqslant \sup\limits_{\bar\Omega }2|\Phi|. \end{equation} $ |
(2.2) |
Proof The existence of $ \Phi $ is from Green formula (one can see [6]). We consider $ u-\Phi $ for $ u>0 $. We note that on $ \mathcal{O}=\{x \in \Omega: $ $ u(x)>0\} $
$ \Delta(u-\Phi) =\varepsilon^{-2}(e^{u}-1+\left(x^{2}+y^{2}\right)\left(1-e^{-u}\right))>0. $ |
By maximum principle, we have
$ \sup\limits _{\mathcal{O}}(u-\Phi)=\sup\limits _{\partial \mathcal{O}}(u-\Phi) \leqslant \sup\limits _{\Omega}\{-\Phi, 0\} \leqslant\sup\limits_\Omega |\Phi|. $ |
It follows
$ \begin{equation} \sup\limits _{\Omega} u \leqslant \sup\limits_\Omega 2|\Phi|. \end{equation} $ |
(2.3) |
Similarly, if $ u<0 $, $ \Delta(u-\Phi) =\varepsilon^{-2}(e^{u}-1+\left(x^{2}+y^{2}\right)\left(1-e^{-u}\right))<0. $ Hence we obtain on $ \mathcal{O}^{-}=\{x \in \Omega: u(x)<0\} $
$ \sup\limits _{\mathcal{O}^{-}}(\Phi-u)=\sup\limits _{\partial \mathcal{O}^{-}}(\Phi-u) \leqslant\sup\limits _{\Omega} \{\Phi ,0\}\leqslant \sup\limits_\Omega |\Phi| $ |
which implies
$ \begin{equation} \sup\limits _{\Omega} {-u} \leqslant \sup\limits_\Omega 2|\Phi|. \end{equation} $ |
(2.4) |
Therefore from (2.3) and (2.4) one can get the estimate (2.2).
Second, we give the gradient estimate of $ u $.
Lemma 2.2 Suppose $ u\in C^{2}(\Omega) $ satisfies the equation (1.2) in $ \Omega $, then there is positive constant $ C $ depending only on $ \Omega $ and $ g $ such that
$ \begin{equation} \sup\limits_{\bar \Omega}|\nabla u|\leqslant C. \end{equation} $ |
(2.5) |
Proof From the equation (1.2) and by the standard regularity, one can see $ u $ is $ C^4 $ since $ u $ and $ g $ are $ C^2 $. Then we have
$ \begin{align} \Delta|\nabla u|^2&=<\nabla \Delta u, \nabla u>+|\nabla^2u|^2\\ &=\varepsilon^{-2}(e^u+r^2e^{-u})|\nabla u|^2-\varepsilon^{-2}e^{-u}<\nabla r^2,\nabla u>+|\nabla^2 u|^2. \end{align} $ |
(2.6) |
If $ |\nabla u|^2 $ attains its maximum on the boundary $ \partial \Omega $, we have $ \sup |\nabla u|=\sup |\nabla g| $ which leads to (2.5). Now we assume $ |\nabla u|^2 $ attains its maximum at $ z_0\in \Omega $. Then from (2.6), at the point $ z_0 $ we have
$ \varepsilon^{-2}(e^u+r^2e^{-u})|\nabla u|^2-\varepsilon^{-2}e^{-u}<\nabla r^2,\nabla u>\leqslant 0 $ |
or
$ (e^u+r^2e^{-u})|\nabla u|^2\leqslant e^{-u} |\nabla r^2| |\nabla u|. $ |
Since $ |u| $ is bounded from Lemma2.1, there is a constant $ C $ dependent on $ \Omega $ and $ g $ such that $ |\nabla u|(z_0)\leqslant C. $ Then we finish the proof.
Now we give the proof of Theorem1.1.
Proof Let $ \sigma\in [0, 1] $, we claim if $ u_\sigma\in C^{2,\alpha}(\Omega) $ is the solution of boundary value problem
$ \begin{equation} \left\{ \begin{aligned} \Delta u &=\sigma\varepsilon^{-2}\left(e^{u}-\left(x^{2}+y^{2}\right) e^{-u}\right) &{\text { in }} \Omega,\\ u&=\sigma g &{\text { on }} \partial \Omega, \end{aligned} \right. \end{equation} $ |
(2.7) |
then there is a constant $ M $ independent of $ u_\sigma $ and $ \sigma $ such that
$ \begin{equation} ||u_\sigma||_{C^{1,\alpha}(\bar{\Omega})}\leqslant M. \end{equation} $ |
(2.8) |
Then one can use the Leray-Schauder existence theorem (see Theorem 6.23 in [6]) to show the Dirichlet problem (1.2) is solvable in $ C^{2,\alpha}(\bar{\Omega}) $.
In fact, one can see $ \sigma\Phi $ solves
$ \begin{equation} \left\{ \begin{aligned} \Delta \Phi&=\sigma\varepsilon^{-2} (1-(x^{2}+y^{2}))&{\text { in }} \Omega,\\ \Phi&=\sigma g &{\text { on }} \partial \Omega. \end{aligned} \right. \end{equation} $ |
(2.9) |
Then from Lemma2.1, we have
$ \begin{equation} ||u_\sigma||_{C^{0}(\bar{\Omega})}\leqslant \sup\limits_\Omega 2|\sigma\Phi |\leqslant \sup\limits_\Omega 2|\Phi|. \end{equation} $ |
(2.10) |
Therefore, from (2.7), there is a constant C independent on $ \sigma $ and $ u_\sigma $, such that $ |\Delta u|\leqslant C. $ This means $ |\nabla^2 u| $ is also bounded. From Lemma2.2 and using interpolation inequality in Hölder space, there is a constant $ M $ independent on $ u $ and $ \sigma $ such that (2.8) is satisfied.
In the end, by standard bootstrap argument of the regularity we have $ u $ is smooth if $ \Omega $ and $ g $ are smooth. For the uniqueness, one can use the method in Lemma 2.1 to prove that if there are two solutions $ u_1 $ and $ u_2 $, then $ u_1=u_2 $.
3 Radial Symmetry of the Global $ C^{2} $ Solution of the Equation in $ \mathbb{R}^{2} $
In this section we will prove Theorem1.2. In [3] and [4], the radially symmetry of the $ C^{2} $ positive solutions of the following second order elliptic equation is studied$ \Delta u+f(u)=0 {\text { in }} \mathbb{R}^{n} $ under the assumption on $ f $ and $ u $. For example, they assumed $ u(x)\to 0 $ as $ x\to \infty $. Obviously, our equation (1.2) is different from this type since $ e^{u}-r^2e^{-u} $ has the term $ r^2 $. Also we cannot assume $ |u|\to 0 $ as $ r\to +\infty $. In fact, it will lead $ \Delta u\to -\infty $ and then $ u $ is unbounded. It contradicts the hypothesis. In this paper, we assume for any finite constant $ c, $
$ \begin{equation} u(s)-u(t) \rightarrow 0\quad {\text {and}}\quad |s|-|t|=c, \quad{\text { as}}\quad |s|,|t| \rightarrow \infty, \end{equation} $ |
(3.1) |
where $ s,t\in \mathbb{R} ^2 $.
Proof of Theorem 1.2 Since the partial differential equation (1.3) is rotationally symmetric, we only have to prove the symmetry about a line across origin. Here we choose the line $ y $ axis. Define $ \Sigma(\lambda)=\left\{\left(x, y\right) \in \mathbb{R}^{2} \mid x<\lambda\right\} $ and let $ v=u(2\lambda-x,y),\ x^{\lambda}=2\lambda-x. $
In $ \Sigma(\lambda) $ we define $ w(x,\lambda)=v-u. $ When $ \lambda = 0 $ and $ (x,y)\in \Sigma(\lambda) $, we have $ x+x^\lambda = 0 $ and $ x<x^\lambda $. Then $ x^2= (x^\lambda)^2 $ and
$ \begin{equation} \Delta v -\varepsilon^{-2}\left(e^{v}-\left(({x^{\lambda}})^{2}+y^{2}\right) e^{-v}\right)=\Delta v -\varepsilon^{-2}\left(e^{v}-\left({x}^{2}+y^{2}\right) e^{-v}\right)= 0. \end{equation} $ |
(3.2) |
By the mean value theorem, we have $ \Delta w+\bar{c} w = 0 $ where $ \bar{c}=-\int_{0}^{1} \varepsilon ^{-2} (e^{ u+t w}+r^2e^{-u-wt} )d t<0. $ Then from the assumption (3.1) and $ w(0,0)=0 $ on the $ y $ axis, we have by maximum principle and minimum principle $ w=0 $ in $ \Sigma (0) $. That's to say the global solution of (1.3) in $ \mathbb{R}^{2} $ is symmetric about $ y $ axis.
In the end, assuming $ \lambda > 0 $ and $ x\in \Sigma(\lambda) $, then we have $ x+x^\lambda > 0 $ and $ x<x^\lambda $. It follows $ x^2< (x^\lambda)^2 $ which implies
$ \begin{equation} \Delta v -\varepsilon^{-2}\left(e^{v}-\left({x}^{2}+y^{2}\right) e^{-v}\right)< 0. \end{equation} $ |
(3.3) |
Then by the mean value theorem, in $ \Sigma(\lambda), $ $ \Delta w+\bar{c} w < 0 $ with $ \bar{c}<0 $. Using the infinite boundary condition (3.1) and $ w(\lambda,\lambda)=0 $, we have by maximum principle, in $ \Sigma(\lambda), $ $ w\geqslant 0. $
Then if $ x>0 $ and let $ x_\lambda \to x $, we have $ \frac{\partial u}{\partial x}\geqslant 0 $. Since $ u $ is radially symmetric and from
$ \frac{\partial u}{\partial x}=\frac{\partial u}{ \partial r} \frac{\partial r}{\partial x}=\frac{\partial u}{\partial r} \frac{x}{r}, $ |
it follows $ \frac{\partial u}{ \partial r} \geqslant0 $ and we finish the proof.