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

Geodesic Equations on asymptotically locally Euclidean Kähler manifolds

Qi Yao
Abstract.

In this paper, We solve the geodesic equation in the space of Kähler metrics under the setting of asymptotically locally Euclidean (ALE) Kähler manifolds and establish global 𝒞1,1\mathcal{C}^{1,1} regularity of the solution. The solution of the geodesic equation is then related to the uniqueness of scalar-flat ALE metrics. To this end, we study the asymptotic behavior of ε\varepsilon-geodesics at spatial infinity. We will prove the convexity of Mabuchi KK energy along ε\varepsilon-geodesics under the assumption that the Ricci curvature of a reference ALE Kähler metric is non-positive. However, by testing the Ricci curvature of ALE Kähler metrics, we find that on the line bundle 𝒪(k)\mathcal{O}(-k) over n1\mathbb{C}\mathbb{P}^{n-1} with n2n\geq 2 and knk\neq n, all ALE Kähler metrics cannot have non-positive (or non-negative) Ricci curvature.

1. Introduction

In the paper, we study the geodesic equation in the setting of ALE Kähler cases, assuming relatively weak fall-off conditions. Let (X,J,g)(X,J,g) be a complete non-compact Kähler manifold of complex dimension nn (n2n\geq 2), we say (X,J,g)(X,J,g) is ALE if there is a compact subset KXK\subseteq X such that ψ:XK(nBR)/Γ\psi:X\setminus K\rightarrow(\mathbb{C}^{n}\setminus B_{R})/\Gamma is a diffeomorphism, where BRB_{R} is a closed ball in n\mathbb{C}^{n} with radius RR and Γ\Gamma is a finite subgroup of U(n)U(n) (any ALE Kähler manifold has only one end according to [19, Proposition 1.5, 3.2]) and the metric gg satisfies the following condition on the end XKX\setminus K:

  1. \bullet

    The metric gg is asymptotic to the Euclidean metric δij\delta_{ij} at the end with decay rate τ-\tau for some τ>n1\tau>n-1, i.e., for i=0,1,,ki=0,1,\ldots,k,

    gij=δij+O(rτ),|i((ψ1)g)|g0=O(rτi).\displaystyle g_{ij}=\delta_{ij}+O(r^{-\tau}),\qquad|\nabla^{i}((\psi^{-1})^{*}g)|_{g_{0}}=O(r^{-\tau-i}). (1.1)

The fall-off condition τ>n1\tau>n-1 is the weakest decay rate to make the ADM mass coordinate-invariant in general, referring to Bartnik [4] and Chrus̀ciel [11].

One of the difficulties in building up a general theory of scalar-flat Kähler metrics in the ALE setting is that the decay rate of such metrics to their asymptotic models is not good enough compared to the Ricci-flat case. For instance, consider the family of scalar-flat Kähler metric constructed on 𝒪1(k)\mathcal{O}_{\mathbb{C}\mathbb{P}^{1}}(-k) by LeBrun [23],

g=ds21+A/s2+B/s4+s2[σ12+σ22+(1+As2+Bs4)σ32],\displaystyle g=\frac{ds^{2}}{1+A/s^{2}+B/s^{4}}+s^{2}\Big{[}\sigma_{1}^{2}+\sigma_{2}^{2}+\Big{(}1+\frac{A}{s^{2}}+\frac{B}{s^{4}}\Big{)}\sigma_{3}^{2}\Big{]},

where AA, BB are constants, σ1\sigma_{1}, σ2\sigma_{2}, σ3\sigma_{3} are three invariant vector fields on 33-sphere and ss is a radial function on 𝒪(k)\mathcal{O}(-k). It can be checked that ggeuc=O(r2)g-g_{euc}=O(r^{-2}), where rr denotes the geodesic distance from a fixed basepoint, indicating that the Kähler potential function should be of log\log growth. In Arezzo-Pacard [2, Lemma 7.2], an expansion theorem is proved for scalar-flat Kähler metrics in the complement of BΓ={zn/Γ:|z|1}B_{\Gamma}=\{z\in\mathbb{C}^{n}/\Gamma:|z|\leq 1\} in n/Γ\mathbb{C}^{n}/\Gamma, where Γ\Gamma is a finite subgroup of U(n)U(n), assuming that the ddcdd^{c}-lemma holds in this situation. In [30], the author proved a ddcdd^{c} lemma and an expansion theorem under the setting of asymptotically conical (AC) Kähler manifolds. Here, we only need a theorem of weaker version under the setting of ALE Kähler manifolds.

Theorem 1.1.

(Yao 2022) Let (X,J)(X,J) be an ALE Kähler manifold asymptotic to n/Γ\mathbb{C}^{n}/\Gamma. Let ω1\omega_{1}, ω2\omega_{2} be Kähler forms in the same Kähler class of (X,J)(X,J) with the corresponding metrics satisfying (1.1) and such that the scalar curvatures of ω1\omega_{1} and ω2\omega_{2} are equal, R1R2R_{1}\equiv R_{2}. Then

ω2=ω1+ddcφ, with the potential φ𝒞22τ~\displaystyle\omega_{2}=\omega_{1}+dd^{c}\varphi,\quad\text{ with the potential }\varphi\in\mathcal{C}^{\infty}_{2-2\tilde{\tau}} (1.2)

for some τ~>n1\tilde{\tau}>n-1 depending on (n,τ)(n,\tau).

Let ω\omega be the corresponding Kähler form of gg. According to Theorem 1.1, given two Kähler forms ω1,ω2[ω]\omega_{1},\omega_{2}\in[\omega], if the corresponding ALE Kähler metrics g1,g2g_{1},g_{2} satisfy the decay condition (1.1) and that the scalar curvatures of g1g_{1} and g2g_{2} are identically equal, R(g1)R(g2)R(g_{1})\equiv R(g_{2}), then ω1ω2=ddcf\omega_{1}-\omega_{2}=dd^{c}f and ff decays at infinity with higher rate γ-\gamma, with γ=2τ~2\gamma=2\tilde{\tau}-2, for some τ~>n1\tilde{\tau}>n-1. Hence, for the prescribed scalar curvature problem, we consider the following restricted weighted Kähler potential space,

γ(ω)={φ𝒞^γ:ωφ=ω+ddcφ>0}(γ>2n40),\displaystyle\mathcal{H}_{-\gamma}(\omega)=\{\varphi\in\hat{\mathcal{C}}^{\infty}_{-\gamma}:\omega_{\varphi}=\omega+dd^{c}\varphi>0\}\quad(\gamma>2n-4\geq 0),

where the class of functions, 𝒞^s\hat{\mathcal{C}}^{\infty}_{s}, is defined as follows

𝒞s\displaystyle\mathcal{C}^{\infty}_{s} ={f𝒞(X):|g0jf|g0=O(rsj)for allj0},\displaystyle=\{f\in\mathcal{C}^{\infty}(X):|\nabla_{g_{0}}^{j}f|_{g_{0}}=O(r^{s-j})\;\,\text{for all}\;\,j\geq 0\},
𝒞^s\displaystyle\hat{\mathcal{C}}^{\infty}_{s} ={f^𝒞(X):f^=f+c, for f𝒞s and c is a constant}.\displaystyle=\{\hat{f}\in\mathcal{C}^{\infty}(X):\hat{f}=f+c,\text{ for }f\in\mathcal{C}^{\infty}_{s}\text{ and }c\text{ is a constant}\}.

Define ω0=ω+ddcψ0\omega_{0}=\omega+dd^{c}\psi_{0}, ω1=ω+ddcψ1\omega_{1}=\omega+dd^{c}\psi_{1}, for any two boundary data ψ0,1γ(ω)\psi_{0,1}\in\mathcal{H}_{-\gamma}(\omega). Also introduce the linear reference path ψ(t)=(1t)ψ0+tψ1\psi(t)=(1-t)\psi_{0}+t\psi_{1} in γ(ω)\mathcal{H}_{-\gamma}(\omega). Another path φ(t)\varphi(t) in γ(ω)\mathcal{H}_{-\gamma}(\omega) with the same endpoints ψ0,ψ1\psi_{0},\psi_{1} is called a geodesic in γ(ω)\mathcal{H}_{-\gamma}(\omega) if

φ¨(t)12|ωφ(t)φ˙(t)|ωφ(t)2=0.\displaystyle\ddot{\varphi}(t)-\frac{1}{2}|\nabla_{\omega_{\varphi(t)}}\dot{\varphi}(t)|_{\omega_{\varphi(t)}}^{2}=0. (1.3)

As observed by Donaldson [15] and Semmes [26], the geodesic equation is equivalent to a homogeneous complex Monge-Ampère equation in the product space X×ΣX\times\Sigma, where Σ[0,1]×S1\Sigma\cong[0,1]\times S^{1} can be embedded as an annulus in \mathbb{C}. Notice that any path φ(t)\varphi(t) of functions on XX can be viewed as a function Φ\Phi on X×ΣX\times\Sigma via Φ(,t,eis)=φ(t)\Phi(\cdot,t,e^{is})=\varphi(t). Let ΩΦ=pω+ddcΦ\Omega_{\Phi}=p^{*}\omega+dd^{c}\Phi, where pp is the projection from X×ΣX\times\Sigma to XX and ddcΦdd^{c}\Phi is computed on X×ΣX\times\Sigma. Then the equation (1.3) can be rewritten as follows:

ΩΦn+1=0,\displaystyle\Omega_{\Phi}^{n+1}=0, (1.4)
ΩΦ0,\displaystyle\Omega_{\Phi}\geq 0, (1.5)
Φ|t=0,1=ψ0,1.\displaystyle\Phi|_{t=0,1}=\psi_{0,1}. (1.6)

In [15], Donaldson proposed a program to attack the existence and uniqueness problems regarding canonical metrics by studying the geometric structure of the potential space \mathcal{H}, where the geodesic equation plays a central role. In the cases of compact Kähler manifolds, Chen [9] showed that for any ψ0\psi_{0}, ψ1\psi_{1}\in\mathcal{H}, the geodesic equation has a unique solution up to ddcdd^{c}-regularity. Blocki [6] and He [18] built up direct calculations to prove the gradient estimate and Laplacian estimate. The full 𝒞1,1\mathcal{C}^{1,1} estimate was proved by Chu-Tosatti-Weinkove in [12]. In the other direction, Lempert-Vivas [24] and Darvas-Lempert [14] constructed counter-examples to assert that ddcΨdd^{c}\Psi is not continuous in general, hence the 𝒞1,1\mathcal{C}^{1,1} regularity is optimal in general. In [3], Auvray generalized the ddcdd^{c}-regularity to singular cases (precisely, there exist cusp singularities along simple normal crossings). The main theorem of sections 2-5 is to generalize the full 𝒞1,1\mathcal{C}^{1,1} estimates to ALE Kähler manifolds.

Theorem A.

Let XX be an ALE Kähler manifold and ψ0,ψ1γ(ω)\psi_{0},\psi_{1}\in\mathcal{H}_{-\gamma}(\omega) (γ>0)(\gamma>0). Then ψ0\psi_{0} and ψ1\psi_{1} can be connected by a 𝒞1,1\mathcal{C}^{1,1} geodesic Φ\Phi solving (1.4), (1.5), (1.6). Moreover, there is a uniform constant CC depending only on ψ0𝒞1,1(X,ω)\|\psi_{0}\|_{\mathcal{C}^{1,1}(X,\omega)}, ψ1𝒞1,1(X,ω)\|\psi_{1}\|_{\mathcal{C}^{1,1}(X,\omega)} and on the geometry of (X,ω)(X,\omega) such that

supX×Σ(|Φ|+|ΘΨΦ|ΘΨ+|ΘΨ2Φ|ΘΨ)C.\displaystyle\sup_{X\times\Sigma}\big{(}|\Phi|+|\nabla_{\Theta_{\Psi}}\Phi|_{\Theta_{\Psi}}+|{\nabla}^{2}_{\Theta_{\Psi}}\Phi|_{\Theta_{\Psi}}\big{)}\leq C. (1.7)

Here, ΘΨ\Theta_{\Psi} is a Kähler form on X×ΣX\times\Sigma given by ΘΨ=Θ+ddcΨ\Theta_{\Psi}=\Theta+dd^{c}\Psi with Ψ(,t,eis)=ψ(t)=(1t)ψ0+tψ1\Psi(\cdot,t,e^{is})=\psi(t)=(1-t)\psi_{0}+t\psi_{1} the linear path introduced above, and with Θ=pω+Addct(t1)\Theta=p^{*}\omega+Add^{c}t(t-1), where A>0A>0 is fixed depending only on ψ0𝒞1,1(X,ω)\|\psi_{0}\|_{\mathcal{C}^{1,1}(X,\omega)}, ψ1𝒞1,1(X,ω)\|\psi_{1}\|_{\mathcal{C}^{1,1}(X,\omega)} such that ΘΨ>0\Theta_{\Psi}>0.

Then, we relate the solution of the geodesic equation to the uniqueness of scalar-flat ALE Kähler metrics in each Kähler class. The main idea is to follow the framework of Chen [9] in the compact case, under the assumption that the Ricci curvature of the reference metric is non-positive. This was extended to the non-compact case with Poincaré cusp ends by Auvray [3]. In the ALE case, it is first necessary to prove sufficient decay at infinity of solutions to the ε\varepsilon-geodesic equation.

In Section 6, we discuss the asymptotic behavior of ε\varepsilon-geodesics. Given any two functions

ψ0,ψ1γ(ω)={φ𝒞^γ:ωφ=ω+ddcφ>0}(γ>0),\displaystyle\psi_{0},\psi_{1}\in\mathcal{H}_{-\gamma}(\omega)=\{\varphi\in\hat{\mathcal{C}}^{\infty}_{-\gamma}:\omega_{\varphi}=\omega+dd^{c}\varphi>0\}\quad(\gamma>0),

we set ψ(t)=(1t)ψ0+tψ1\psi(t)=(1-t)\psi_{0}+t\psi_{1} and let Ψ\Psi denote the corresponding function on X×ΣX\times\Sigma. We fix AA large depending on ψ0𝒞1,1(X,ω)\|\psi_{0}\|_{\mathcal{C}^{1,1}(X,\omega)}, ψ1𝒞1,1(X,ω)\|\psi_{1}\|_{\mathcal{C}^{1,1}(X,\omega)} such that ΘΨ:=Θ+ddcΨ\Theta_{\Psi}:=\Theta+dd^{c}\Psi is positive on X×ΣX\times\Sigma, where Θ:=pω+Addct(t1)\Theta:=p^{*}\omega+Add^{c}t(t-1) with p:X×ΣXp:X\times\Sigma\to X the projection. Then, we introduce the following ε\varepsilon-geodesic equations

(Eε){(Θ+ddcΦε)n+1=υ(ε)ΘΨn+1, in X×Σ,Θ+ddcΦε>0, in X×Σ,Φε|t=0,1=ϕ0,1, on X×Σ,\displaystyle(E_{\varepsilon})\qquad\begin{cases}(\Theta+dd^{c}\Phi_{\varepsilon})^{n+1}=\upsilon(\varepsilon)\Theta_{\Psi}^{n+1},\quad&\text{ in }X\times\Sigma,\\ \Theta+dd^{c}\Phi_{\varepsilon}>0,&\text{ in }X\times\Sigma,\\ \Phi_{\varepsilon}|_{t=0,1}=\phi_{0,1},&\text{ on }X\times\partial\Sigma,\end{cases}

where υ(ε)\upsilon(\varepsilon) is a smooth nonnegative function defined in X×Σ×[0,1]X\times\Sigma\times[0,1] satisfies the following conditions

υ(0)0,υ(1)1;υ(ε)>0, for ε(0,1];C1ευ(ε)min(Cε,1),for ε[0,1];|kυ(y,ε)|Cεr(y)ςk, for (y,ε)X×Σ×[0,1],k1\displaystyle\begin{split}&\upsilon(0)\equiv 0,\quad\upsilon(1)\equiv 1;\\ &\upsilon(\varepsilon)>0,\qquad\qquad\qquad\qquad\ \text{ for }\varepsilon\in(0,1];\\ &C^{-1}\varepsilon\leq\upsilon(\varepsilon)\leq\min(C\varepsilon,1),\quad\text{for }\varepsilon\in[0,1];\\ &|\nabla^{k}\upsilon(y,\varepsilon)|\leq C\varepsilon r(y)^{-\varsigma-k},\quad\text{ for }(y,\varepsilon)\in X\times\Sigma\times[0,1],\quad k\geq 1\end{split} (1.8)

where ς\varsigma is an real number with ςγ\varsigma\geq\gamma. In particular, we are interested in the following case. By taking

υ(ε)=ε((1χ(ε))f+χ(ε)),\displaystyle\upsilon(\varepsilon)=\varepsilon((1-\chi(\varepsilon))f+\chi(\varepsilon)), (1.9)

where χ\chi is a smooth increasing function in [0,1][0,1] equal to 0 (resp. 11) in a neighborhood of 0 (resp. 11) and ff is defined as follows

f=A1Θn+1Θψn+1𝒞(X×Σ),\displaystyle f=A^{-1}\frac{\Theta^{n+1}}{\Theta_{\psi}^{n+1}}\in\mathcal{C}^{\infty}(X\times\Sigma), (1.10)

and in this case, |kf|Crγ2k|\nabla^{k}f|\leq Cr^{-\gamma-2-k}, ς=γ+2\varsigma=\gamma+2. By taking ε\varepsilon to be small enough, (Eε)(E_{\varepsilon}) can be written as

(φ¨12|ωφφ˙|ωφ2)ωφn=εωn.\displaystyle\Big{(}\ddot{\varphi}-\frac{1}{2}|\nabla_{\omega_{\varphi}}\dot{\varphi}|_{\omega_{\varphi}}^{2}\Big{)}\omega_{\varphi}^{n}=\varepsilon\omega^{n}.

Due to the positivity of the right hand side of (Eε)(E_{\varepsilon}), it is well known that for every ε(0,1]\varepsilon\in(0,1] there exists a solution Φεk,α𝒞k,α\Phi_{\varepsilon}\in\bigcap_{k,\alpha}\mathcal{C}^{k,\alpha}. We now prove:

Theorem B.

Let Φε\Phi_{\varepsilon} be the ε\varepsilon-geodesic constructed above. Then, there exists a constant C(k,ε1)C(k,\varepsilon^{-1}) depending on k1k\geq 1 and on an upper bound for ε1\varepsilon^{-1} such that

(|X,ωkΦε|ω+|X,ωkΦ˙ε|ω+|X,ωkΦ¨ε|ω)C(k,ε1)rγkfor allk1,\displaystyle\left(|\nabla^{k}_{X,\omega}\Phi_{\varepsilon}|_{\omega}+|\nabla^{k}_{X,\omega}\dot{\Phi}_{\varepsilon}|_{\omega}+|\nabla^{k}_{X,\omega}\ddot{\Phi}_{\varepsilon}|_{\omega}\right)\leq C(k,\varepsilon^{-1})r^{-\gamma-k}\quad\text{for all}\ k\geq 1,

where X,ω\nabla_{X,\omega} denotes the Levi-Civita connection of the ALE Kähler metric ω\omega on XX, acting as a differential operator in the XX directions on X×ΣX\times\Sigma. And

|Φεc(t)|C(ε1)rγ,\displaystyle|\Phi_{\varepsilon}-c(t)|\leq C(\varepsilon^{-1})r^{-\gamma},

where c(t)c(t) is a function only depending on tt. Hence, for any two potentials ψ0\psi_{0}, ψ1\psi_{1} in γ(ω)\mathcal{H}_{-\gamma}(\omega), there exist ε\varepsilon-geodesics in γ(ω)\mathcal{H}_{-\gamma}(\omega) connecting ψ0\psi_{0} and ψ1\psi_{1}.

In section 6, we prove a stronger statement. Let φε=ΦεΨ\varphi_{\varepsilon}=\Phi_{\varepsilon}-\Psi, then φεmax{2γ2,ς}\varphi_{\varepsilon}\in\mathcal{H}_{\max\{-2\gamma-2,-\varsigma\}} due to the fact that Ψ\Psi was chosen to be linear in tt (see section 6 for details).

Hence, while we still cannot define the Mabuchi KK-energy along geodesics, the Mabuchi KK-energy is now actually well-defined along ε\varepsilon-geodesics assuming γ=2τ~2>2n4\gamma=2\tilde{\tau}-2>2n-4.

In Section 7, the second derivative of the Mabuchi KK-energy will be calculated. Throughout section 7, we assume γ=2τ~2>2n4\gamma=2\tilde{\tau}-2>2n-4 (Here it turns out that if ψ0\psi_{0}, ψ1\psi_{1} are only in 42n(ω)\mathcal{H}_{4-2n}(\omega), there would be boundary terms at infinity breaking the positivity of the second derivative. This is a new phenomenon compared to Chen [9] and Auvray [3]). However, under the assumption that the Ricci curvature of some reference ALE Kähler metric, ω\omega, is non-positive, we can then prove the convexity of Mabuchi KK-energy:

Theorem C.

Assume that ω\omega is an ALE Kähler metric on XX such that the Ricci curvature of ω\omega is non-positive, Ric(ω)0\textup{Ric}(\omega)\leq 0. Then, along each ε\varepsilon-geodesic in γ(ω)\mathcal{H}_{-\gamma}(\omega) with γ>2n4\gamma>2n-4, φ(t)\varphi(t), the Mabuchi KK-energy is convex.

A quick corollary of Theorem C is that assuming Ric(ω)0\textup{Ric}(\omega)\leq 0, the scalar-flat Kähler metric, if it exists, is unique in γ(ω)\mathcal{H}_{-\gamma}(\omega). However, if there exists a scalar-flat Kähler metric ω0\omega_{0} in γ(ω)\mathcal{H}_{-\gamma}(\omega), the condition, Ric(ω)0\textup{Ric}(\omega)\leq 0 implies Ric(ω)=0\textup{Ric}(\omega)=0. Hence, the uniqueness of scalar-flat ALE metric can be reduced to the uniqueness result of Ricci-flat ALE Kähle metric, which can be found in reference [20, 28, 13]. The point is that ω0=ω+O(rγ2)\omega_{0}=\omega+O(r^{-\gamma-2}) implies by definition that the ADM masses of ω\omega and ω0\omega_{0} are equal, 𝔪(ω)=𝔪(ω0)\mathfrak{m}(\omega)=\mathfrak{m}(\omega_{0}). According to mass formula by Hein-LeBrun [19], it follows that R(ω)=R(ω0)=0\int R(\omega)=\int R(\omega_{0})=0. The assumption that Ric(ω)0\textup{Ric}(\omega)\leq 0 implies that Ric(ω)=0\textup{Ric}(\omega)=0 (see Remark 7.4 for details). In fact, in Section 8, we will prove that many ALE Kähler manifolds do not admit any ALE Kähler metrics with Ric0\textup{Ric}\leq 0 (or Ric0\textup{Ric}\geq 0) at all:

Theorem D.

Let 𝒪(k)\mathcal{O}(-k) be the standard negative line bundle over n1\mathbb{C}\mathbb{P}^{n-1} with n2n\geq 2, knk\neq n, and let ω\omega be an ALE Kähler metric on 𝒪(k)\mathcal{O}(-k) with decay rate τ-\tau, τ>0\tau>0. Then, the Ricci form of ω\omega, is of mixed type, i.e., neither Ric(ω)0\textup{Ric}(\omega)\geq 0 nor Ric(ω)0\textup{Ric}(\omega)\leq 0 is true.

In Riemannian geometry, AE metrics of negative Ricci curvature are well-known to exist in n\mathbb{R}^{n} by explicit construction in Lohkamp [25]. Theorem D gives a negative answer to this question in the setting of ALE Kähler metrics.

An interesting question in this context is to ask whether some version of the Nonexistence Theorem D holds in general ALE Kähler manifolds or even AC Kähler manifolds.

Question.

Is it true in any ALE Kähler manifold that the Ricci curvature form of an ALE Kähler metric can only be identically zero or of mixed type?

This paper is a part of the Ph.D. thesis of the author. The author would like to express his gratitude to Professor Hans-Joachim Hein and Professor Bianca Santoro for suggesting the problem, and for constant support, many helpful comments, as well as much enlightening conversation. The author is also thankful to professor Gustav Holzegel for providing financial support via his Alexander von Humboldt Professorship during the last semester at University of Münster. The whole project is Funded by the DFG under Germany’s Excellence Strategy EXC 2044-390685587, Mathematics Münster: Dynamics-Geometry-Structure, and by the CRC 1442, Geometry: Deformations and Rigidity, of the DFG.

2. ε\varepsilon-geodesic equations and openness

Recall that ε\varepsilon geodesic equations can be written as follows,

(Eε){(Θ+ddcΦ~)n+1=υ(ε)(Θ+ddcΨ)n+1, in X×Σ,λΘ<Θ+ddcΦ~<ΛΘ, in X×Σ,Φ~|t=0,1=ψ0,1, on X×Σ,\displaystyle(E_{\varepsilon})\qquad\begin{cases}(\Theta+dd^{c}\widetilde{\Phi})^{n+1}=\upsilon(\varepsilon)(\Theta+dd^{c}\Psi)^{n+1},\quad&\text{ in }X\times\Sigma,\\ \lambda\Theta<\Theta+dd^{c}\widetilde{\Phi}<\Lambda\Theta,&\text{ in }X\times\Sigma,\\ \widetilde{\Phi}|_{t=0,1}=\psi_{0,1},&\text{ on }X\times\partial\Sigma,\end{cases}

where ε(0,1]\varepsilon\in(0,1] and 0<λ<Λ0<\lambda<\Lambda are constants depending on ε\varepsilon. The family of equations (Eε)(E_{\varepsilon}) is called the ε\varepsilon-geodesic equations. The idea to solve the equation (Eε)(E_{\varepsilon}) is the following. Firstly, we apply the continuity method to show that there exists a solution of (Eε)(E_{\varepsilon}) in 𝒞k,α\mathcal{C}^{k,\alpha}. In particular, consider the family of equations (Es)(E_{s}), s[ε,1]s\in[\varepsilon,1]. There is a trivial solution at (E1)(E_{1}). Then, we shall prove the openness and closedness of (Es)(E_{s}) in certain regularity. In the current section, we deal with the openness of (Es)(E_{s}).

Assuming that there exists a solution of (Es0)(E_{s_{0}}) in 𝒞k,α\mathcal{C}^{k,\alpha} for some s0[ε,1]s_{0}\in[\varepsilon,1], we will show in this subsection that (Es)(E_{s}) can be solved for all ss in a small open neighborhood of s0s_{0}. For simplicity, we write ΘΨ=Θ+ddcΨ\Theta_{\Psi}=\Theta+dd^{c}\Psi as in Theorem A and φ~=Φ~Ψ\widetilde{\varphi}=\widetilde{\Phi}-\Psi. Then, the equation (Eε)(E_{\varepsilon}) can be written as, (ΘΨ+ddcφ~)n+1=εΘΨn+1(\Theta_{\Psi}+dd^{c}\widetilde{\varphi})^{n+1}=\varepsilon\Theta_{\Psi}^{n+1} in X×ΣX\times\Sigma, with the boundary condition φ~=0\widetilde{\varphi}=0 on X×ΣX\times\partial\Sigma. Then, the Monge-Ampère operator is defined to be

(χ)=(ΘΨ+ddcχ)n+1ΘΨn+1.\displaystyle\mathcal{M}(\chi)=\frac{(\Theta_{\Psi}+dd^{c}\chi)^{n+1}}{\Theta_{\Psi}^{n+1}}.

Let φ~\widetilde{\varphi} be a solution of (Es0)(E_{s_{0}}) for some s0[ε,1]s_{0}\in[\varepsilon,1]. By assumption, φ~\widetilde{\varphi} is ΘΨ\Theta_{\Psi}-plurisubharmonic satisfying cΘΘΨ+ddcφ~CΘc\Theta\leq\Theta_{\Psi}+dd^{c}\widetilde{\varphi}\leq C\Theta. Then, the linearization of Monge-Ampère operator at φ~\widetilde{\varphi} is uniformly elliptic, and is given by

φ~(χ)=(Δφ~χ)(ΘΨ+ddcφ~)n+1ΘΨn+1=s0Δφ~χ,\displaystyle\mathcal{L}_{\widetilde{\varphi}}(\chi)=\big{(}\Delta_{{\widetilde{\varphi}}}\chi\big{)}\cdot\frac{(\Theta_{\Psi}+dd^{c}\widetilde{\varphi})^{n+1}}{\Theta_{\Psi}^{n+1}}=s_{0}\Delta_{\widetilde{\varphi}}\chi,

where Δφ~\Delta_{\widetilde{\varphi}} represents the Laplacian with respect to ΘΨ+ddcφ~\Theta_{\Psi}+dd^{c}\widetilde{\varphi}. Let (𝒞k,α)0(\mathcal{C}^{k,\alpha})_{0} be the functions in 𝒞k,α\mathcal{C}^{k,\alpha} vanishing on the boundary X×ΣX\times\partial\Sigma. Then, we have the following property of φ~\mathcal{L}_{\widetilde{\varphi}}, from which the desired openness is clear by the implicit function theorem.

Proposition 2.1.

Let φ~\widetilde{\varphi} be the solution of (Es0)(E_{s_{0}}), then the linearized operator φ~:(𝒞k,α)0𝒞k2,α\mathcal{L}_{\widetilde{\varphi}}:({\mathcal{C}}^{k,\alpha})_{0}\rightarrow{\mathcal{C}}^{k-2,\alpha} is an isomorphism for all integers k2k\geq 2 and α(0,1)\alpha\in(0,1).

Proof.

Let us first prove the surjectivity. Fixing f𝒞k2,αf\in{\mathcal{C}}^{k-2,\alpha}, the exhaustion argument will be applied to solve the equation φ~u=f\mathcal{L}_{\widetilde{\varphi}}u=f. Take an exhaustive sequence of pre-compact sets, ΩkX×Σ\Omega_{k}\subseteq X\times\Sigma, with smooth boundary. In particular, by taking a sequence of subsets, Brk×ΣB_{r_{k}}\times\Sigma where Brk={xX:r(x)rk}B_{r_{k}}=\{x\in X:r(x)\leq r_{k}\}, and smoothing the corners, we can obtain the exhaustive sequence {Ωk}\{\Omega_{k}\}. Then, we can solve the following Dirichlet problems,

(Lk){φ~uk=f in Ωk,uk=0 on Ωk,\displaystyle(L_{k})\quad\begin{cases}\mathcal{L}_{\widetilde{\varphi}}u_{k}=f\quad&\text{ in }\Omega_{k},\\ u_{k}=0&\text{ on }\partial\Omega_{k},\end{cases}

where f𝒞k2,αf\in{\mathcal{C}}^{k-2,\alpha}. The existence of the solution of (Lk)(L_{k}) is a classic result of the Dirichlet problem on compact Riemannian manifolds with boundary. The key to complete the proof is to give the uniform estimates of uku_{k}. The main idea to show the 𝒞0\mathcal{C}^{0} uniform estimates is to construct barrier functions. Consider the function At(1t)At(1-t). The fact that λΘΘΨ+ddcφ~ΛΘ\lambda\Theta\leq\Theta_{\Psi}+dd^{c}\widetilde{\varphi}\leq\Lambda\Theta implies Δφ~At(1t)λA\Delta_{\widetilde{\varphi}}At(1-t)\leq-\lambda A. If we suppose that fLC0\|f\|_{L^{\infty}}\leq C_{0} and take A=C0/λA=C_{0}/\lambda, then we have Δφ~At(1t)f=Δφ~uk\Delta_{\widetilde{\varphi}}At(1-t)\leq f=\Delta_{\widetilde{\varphi}}u_{k}. Combining with the fact that At(1t)0At(1-t)\geq 0 on the boundary Ωk\partial\Omega_{k}, the maximum principle implies that,

ukLC0λt(1t)C04λ.\displaystyle\|u_{k}\|_{L^{\infty}}\leq\frac{C_{0}}{\lambda}t(1-t)\leq\frac{C_{0}}{4\lambda}. (2.1)

The uniform 𝒞k,α\mathcal{C}^{k,\alpha} estimates follows directly from the standard Schauder estimates. Precisely, for interior points pΩkp\in\Omega_{k} away from the boundary, we pick a pair of balls centered at pp, B14(p)B12(p)ΩkB_{\frac{1}{4}}(p)\subseteq B_{\frac{1}{2}}(p)\subset\Omega_{k}. Then, the interior Schauder estimates implies that ukk,α;B116(p)C(ukL(B18(p))+fk2,α;B18(p))\|u_{k}\|_{k,\alpha;B_{\frac{1}{16}}(p)}\leq C(\|u_{k}\|_{L^{\infty}(B_{\frac{1}{8}}(p))}+\|f\|_{k-2,\alpha;B_{\frac{1}{8}}(p)}). If pΩkp\in\Omega_{k} is close to the boundary, we can apply the boundary Schauder estimate. After straightening the boundary in case the boundary portion on Ωk\partial\Omega_{k} is not flat, we can pick half balls, pB14+(q)B12+(q)p\in B^{+}_{\frac{1}{4}}(q)\subseteq B^{+}_{\frac{1}{2}}(q) for some qΩkq\in\partial\Omega_{k}. Together with the interior estimates, we have

ukk,α;ΩkC(ukL(X×Σ)+fk2,α;X×Σ),\displaystyle\|u_{k}\|_{k,\alpha;\Omega_{k}}\leq C(\|u_{k}\|_{L^{\infty}(X\times\Sigma)}+\|f\|_{k-2,\alpha;X\times\Sigma}), (2.2)

where CC depends only on n,k,α,λ,Λn,k,\alpha,\lambda,\Lambda. After passing to a subsequence, we conclude that the limit function, uu, satisfies φ~u=f\mathcal{L}_{\widetilde{\varphi}}u=f in X×ΣX\times\Sigma and u0u\equiv 0 on X×ΣX\times\partial\Sigma. The uniqueness directly follows from the following maximum principle, Lemma 2.3. ∎

The following lemma comes from Yau’s generalized maximum principle, referring to [10, 29]. To describe the model metric on X×ΣX\times\Sigma, we introduce the asymptotic coordinates of X×ΣX\times\Sigma. Let {z1,,zn}\{z_{1},\ldots,z_{n}\} be asymptotic coordinates of the end of XX and let w=t+isw=t+is be the complex coordinate of Σ\Sigma. Real asymptotic coordinates are given by {x1,,x2n,x2n+1=t,x2n+2=s}\{x_{1},\ldots,x_{2n},x_{2n+1}=t,x_{2n+2}=s\}, where the complex coordinates are written as zi=x2i1+ix2iz_{i}=x_{2i-1}+ix_{2i}. The asymptotic coordinate system will be applied to describe the asymptotic behavior of prescribed Kähler metrics on X×ΣX\times\Sigma.

Lemma 2.2.

Let (X×Σ,ΘΦ~)(X\times\Sigma,\Theta_{\widetilde{\Phi}}) be the noncompact Kähler manifold as above with the Kähler metric g~\widetilde{g} associated with Φ~\widetilde{\Phi} satisfying, for some uniform constant 0<λ<Λ0<\lambda<\Lambda,

λδijg~ijΛδij\displaystyle\lambda\delta_{ij}\leq\widetilde{g}_{ij}\leq\Lambda\delta_{ij}

in the asymptotic coordinates of X×ΣX\times\Sigma. Let uu be a 𝒞loc2\mathcal{C}_{loc}^{2} function bounded from above on X×ΣX\times\Sigma. Suppose that supX×Σu>supX×Σu\sup_{X\times\Sigma}u>\sup_{X\times\partial\Sigma}u, then there exists a sequence {xk}\{x_{k}\} in X×ΣX\times\Sigma^{\circ} such that

limku(xk)=supX×Σu,limk|du(xk)|g~=0,lim supkΔg~u(xk)0.\displaystyle\lim_{k\rightarrow\infty}u(x_{k})=\sup_{X\times\Sigma}u,\quad\lim_{k\rightarrow\infty}|du(x_{k})|_{\widetilde{g}}=0,\quad\limsup_{k\rightarrow\infty}\Delta_{\widetilde{g}}u(x_{k})\leq 0. (2.3)
Proof.

Let rr be the radial function inherited from the asymptotic chart of XX, for instance, r=(i=1n|zi|2)1/2r=(\sum_{i=1}^{n}|z_{i}|^{2})^{1/2}. The radial function can be extended to a non-negative smooth function in the whole space X×ΣX\times\Sigma satisfying the estimate

|g~r|g~C,|Δg~r|C,\displaystyle|\nabla_{\widetilde{g}}r|_{\widetilde{g}}\leq C,\qquad|\Delta_{\widetilde{g}}r|\leq C, (2.4)

for some uniform constant CC. Consider the function u𝐞=u𝐞ru_{\mathbf{e}}=u-\mathbf{e}r. Since u𝐞u_{\mathbf{e}} tends to negative infinity as rr goes to infinity, u𝐞u_{\mathbf{e}} achieves its maximum at some point x𝐞x_{\mathbf{e}}. And x𝐞x_{\mathbf{e}} must be an interior point in X×ΣX\times\Sigma based on the assumption that supX×Σu>supX×Σu\sup_{X\times\Sigma}u>\sup_{X\times\partial\Sigma}u. At x𝐞x_{\mathbf{e}}, the function u𝐞u_{\mathbf{e}} satisfies

0=du𝐞(x𝐞)=du(x𝐞)𝐞dr(x𝐞),\displaystyle 0=du_{\mathbf{e}}(x_{\mathbf{e}})=du(x_{\mathbf{e}})-\mathbf{e}dr(x_{\mathbf{e}}),
0Δg~u𝐞(x𝐞)=Δg~u(x𝐞)𝐞Δg~r(x𝐞)\displaystyle 0\geq\Delta_{\widetilde{g}}u_{\mathbf{e}}(x_{\mathbf{e}})=\Delta_{\widetilde{g}}u(x_{\mathbf{e}})-\mathbf{e}\Delta_{\widetilde{g}}r(x_{\mathbf{e}})

and

u𝐞(x𝐞)u(x)𝐞r(x), for all xX×Σ.\displaystyle u_{\mathbf{e}}(x_{\mathbf{e}})\geq u(x)-\mathbf{e}r(x),\quad\text{ for all }x\in X\times\Sigma.

Choosing {xk}\{x_{k}\} to be points achieving the maximum of u1/ku_{1/k}, then combining with (2.4) and letting kk go to infinity, we complete the proof of (2.3). ∎

The following lemma is a strengthened version of the above maximum principle, based on solving the Dirichlet problem in X×ΣX\times\Sigma.

Lemma 2.3.

Let (X×Σ,g~)(X\times\Sigma,\widetilde{g}) be the same as in Lemma 2.2. Suppose that uu is a function in 𝒞loc2(X×Σ)\mathcal{C}^{2}_{loc}(X\times\Sigma) and bounded from above. Suppose that uu satisfies Δg~u0\Delta_{\widetilde{g}}u\geq 0 in X×ΣX\times\Sigma and u0u\leq 0 on X×ΣX\times\partial\Sigma. Then u0u\leq 0 in X×ΣX\times\Sigma.

Proof.

Assuming uu satisfies supX×Σuδ>0\sup_{X\times\Sigma}u\geq\delta>0. According to the surjectivity part of the proof of Proposition 2.1, there exists a function vv satisfying

{Δg~v=1, in X×Σ,v=0, on X×Σ,\displaystyle\begin{cases}\Delta_{\widetilde{g}}v=-1,\quad&\text{ in }X\times\Sigma,\\ v=0,&\text{ on }X\times\partial\Sigma,\end{cases}

and vLC(n,λ,Λ)\|v\|_{L^{\infty}}\leq C(n,\lambda,\Lambda). Consider the function u𝐞=u𝐞vu_{\mathbf{e}}=u-\mathbf{e}v for 𝐞=δ2C\displaystyle\mathbf{e}=\frac{\delta}{2C}. Then supX×Σu𝐞δ2>0\displaystyle\sup_{X\times\Sigma}u_{\mathbf{e}}\geq\frac{\delta}{2}>0 and Δg~u𝐞𝐞\displaystyle\Delta_{\widetilde{g}}u_{\mathbf{e}}\geq\mathbf{e}. According to Lemma 2.2, there exists a sequence {xk}\{x_{k}\} in X×ΣX\times\Sigma^{\circ} such that limku𝐞(xk)=supX×Σu𝐞\lim_{k\rightarrow\infty}u_{\mathbf{e}}(x_{k})=\sup_{X\times\Sigma}u_{\mathbf{e}}, limk|du𝐞(xk)|g~=0\lim_{k\rightarrow\infty}|du_{\mathbf{e}}(x_{k})|_{\widetilde{g}}=0, lim supkΔg~u𝐞(xk)0\limsup_{k\rightarrow\infty}\Delta_{\widetilde{g}}u_{\mathbf{e}}(x_{k})\leq 0. However, Δg~u𝐞>0\Delta_{\widetilde{g}}u_{\mathbf{e}}>0, which leads to the contradiction. ∎

3. A priori estimate up to 𝒞0\mathcal{C}^{0}

From section 3 to 5, we complete the proof of Theorem A. The key ingredient is to prove uniform a priori estimates up to order 𝒞1,1\mathcal{C}^{1,1} for the solution φ~=Φ~Ψ\widetilde{\varphi}=\widetilde{\Phi}-\Psi of the ε\varepsilon-geodesic equation (EεE_{\varepsilon}). These estimates will be uniform with respect to ε(0,1]\varepsilon\in(0,1] and with respect to the distance from a fixed point in XX. (In section 6, we will also see that for a fixed ε>0\varepsilon>0 it can be proved that φ~\widetilde{\varphi} is decaying at spatial infinity. However, we are currently unable to make these decay estimates uniform with respect to ε\varepsilon.)

These uniform 𝒞1,1\mathcal{C}^{1,1} estimates are then used in two ways:

  • \bullet

    First, they allow us to solve (Eε)(E_{\varepsilon}) for any fixed ε(0,1]\varepsilon\in(0,1] via the continuity method in (𝒞k,α)0(\mathcal{C}^{k,\alpha})_{0} for any k2k\geq 2. Recall this is done by considering the family of equations (Es)(E_{s}) with s[ε,1]s\in[\varepsilon,1], where openness in (𝒞k,α)0(\mathcal{C}^{k,\alpha})_{0} follows from Proposition 2.1. The uniform 𝒞1,1\mathcal{C}^{1,1} estimates that we will prove, together with the general regularity theory of the Monge-Ampère equation, then imply closedness. Here, it is not yet important that the 𝒞1,1\mathcal{C}^{1,1} estimates are uniform in ε\varepsilon, and the higher 𝒞k,α\mathcal{C}^{k,\alpha} estimates will depend on ε\varepsilon because the ellipticity of the equation does. Also, note that these higher-order estimates follow from standard local regularity in the interior and from [8, Section 2.1–2.2] near the boundary because we already have a true 𝒞1,1\mathcal{C}^{1,1} bound.

  • \bullet

    Once (Eε)(E_{\varepsilon}) is actually solved, we can then let ε\varepsilon go to zero and use the uniformity of the 𝒞1,1\mathcal{C}^{1,1} estimates of the ε\varepsilon-geodesic solution φ~\widetilde{\varphi} to extract a subsequential limit φ𝒞1,1\varphi\in\mathcal{C}^{1,1} such that Φ=Ψ+φ\Phi=\Psi+\varphi solves the geodesic equation (1.4), (1.5), (1.6).

We omit these standard arguments and instead focus on the proof of the uniform 𝒞1,1\mathcal{C}^{1,1} a priori estimates of the ε\varepsilon-geodesic solution φ~\widetilde{\varphi}. For this we follow the outline of [7] in the compact case. However, we provide all the necessary details that are required to generalize this theory to the ALE case. In addition, we also make use of the recent advance [12] to obtain a 𝒞1,1\mathcal{C}^{1,1} estimate which is uniform in ε\varepsilon.

In this section, we only deal with the uniform 𝒞0\mathcal{C}^{0} estimate. We begin with a standard comparison principle [5, Proposition 3.1].

Lemma 3.1.

Let DD be a bounded connected domain in n\mathbb{C}^{n} with smooth boundary and u,v𝒞2(D)u,v\in\mathcal{C}^{2}(D), plurisubharmonic functions in DD. If u=vu=v on D\partial D and uvu\geq v, then we have

Ω(ddcu)nΩ(ddcv)n.\displaystyle\int_{\Omega}(dd^{c}u)^{n}\leq\int_{\Omega}(dd^{c}v)^{n}.

Then we can prove the following maximum principle for Monge-Ampère operators.

Theorem 3.2.

Let Θ\Theta be a fixed reference Kähler form and Ω\Omega, the pull-back of a semipositive (1,1)(1,1)-form in XX. Assume that u,v𝒞2(X×Σ)u,v\in\mathcal{C}^{2}(X\times\Sigma) are bounded functions with Ω+ddcv\Omega+dd^{c}v, Ω+ddcu0\Omega+dd^{c}u\geq 0. If for some positive constants λ\lambda, Λ\Lambda, we have the following properties:

(Ω+ddcv)n+1(Ω+ddcu)n+1\displaystyle(\Omega+dd^{c}v)^{n+1}\leq(\Omega+dd^{c}u)^{n+1}  in X×Σ,\displaystyle\quad\text{ in }X\times\Sigma, (3.1)
λΘΩ+ddcuΛΘ\displaystyle\lambda\Theta\leq\Omega+dd^{c}u\leq\Lambda\Theta  in X×Σ,\displaystyle\quad\text{ in }X\times\Sigma, (3.2)
uv\displaystyle u\leq v  on X×Σ,\displaystyle\quad\text{ on }X\times\partial\Sigma, (3.3)

then uvu\leq v in X×ΣX\times\Sigma.

Proof.

Assume u(z0)>v(z0)u(z_{0})>v(z_{0}) at some point z0X×Σz_{0}\in X\times\Sigma. Let 2h=u(z0)v(z0)2h=u(z_{0})-v(z_{0}). Then, we can modify u,vu,v to be u~,v~\tilde{u},\tilde{v} as follows:

v~=v+h,u~=u+h2|τ|2.\displaystyle\begin{split}&\tilde{v}=v+h,\\ &\tilde{u}=u+\frac{h}{2}|\tau|^{2}.\end{split} (3.4)

It can be checked that u~\tilde{u}, v~\tilde{v} are bounded functions satisfying that u~v~\tilde{u}\leq\tilde{v} on X×ΣX\times\partial\Sigma and u~(z0)v~(z0)+h\tilde{u}(z_{0})\geq\tilde{v}(z_{0})+h. By Wu-Yau’s generalized maximum principle, there exists a sequence {pk}\{p_{k}\} in X×ΣX\times\Sigma such that

limk(u~v~)(pk)=supX×Σ(u~v~)h,lim supkddc(u~v~)(pk)0.\lim_{k\rightarrow\infty}(\tilde{u}-\tilde{v})(p_{k})=\sup_{X\times\Sigma}(\tilde{u}-\tilde{v})\geq h,\quad\limsup_{k\rightarrow\infty}dd^{c}(\tilde{u}-\tilde{v})(p_{k})\leq 0.

For a sufficiently small constant δ>0\delta>0, there exist a point pX×Σp\in X\times\Sigma, ddcu~(p)ddcv~(p)δΘdd^{c}\tilde{u}(p)-dd^{c}\tilde{v}(p)\leq\delta\Theta and η0=u~(p)v~(p)supX×Σ(u~v~)δ\eta_{0}=\tilde{u}(p)-\tilde{v}(p)\geq\sup_{X\times\Sigma}(\tilde{u}-\tilde{v})-\delta. Fix a local holomorphic chart around pp, {U,zi:i=1,,n+1}\{U,z^{i}:i=1,\ldots,n+1\} with zn+1=τz^{n+1}=\tau. Without loss of generality, we assume UU contains the unit disk in n+1\mathbb{C}^{n+1} and for any local vector field VT1,0UV\in T^{1,0}U,

C1|V|2Θ(V,V¯)C|V|2,\displaystyle C^{-1}|V|^{2}\leq\Theta(V,\overline{V})\leq C|V|^{2},

where the constant CC only depends on the geometry of XX and the reference metric Θ\Theta. Let 𝐞=2Cδ\mathbf{e}=2C\delta and η=η0Cδ2\eta=\eta_{0}-\frac{C\delta}{2}. To derive the contradiction, we construct the following local functions in UU,

u=u~𝐞|z|2,v=v~+η.\displaystyle\begin{split}&\accentset{\approx}{u}=\tilde{u}-\mathbf{e}|z|^{2},\\ &\accentset{\approx}{v}=\tilde{v}+\eta.\end{split} (3.5)

If we denote the unit ball contained in the coordinate chart of UU by B1(p)B_{1}(p), we have u(p)v(p)=Cδ2>0\accentset{\approx}{u}(p)-\accentset{\approx}{v}(p)=\frac{C\delta}{2}>0 and uv\accentset{\approx}{u}\leq\accentset{\approx}{v} on B1(p)\partial B_{1}(p). Consider the following subset of B1(p)B_{1}(p),

D={zB1(p):u(z)>v(z)}.\displaystyle D=\{z\in B_{1}(p):\accentset{\approx}{u}(z)>\accentset{\approx}{v}(z)\}.

Let ρ\rho be the local potential of Ω\Omega in UU, Ω=ddcρ\Omega=dd^{c}\rho. According to Lemma 3.1,

D[ddc(ρ+v)]n+1D[ddc(ρ+u)]n+1.\displaystyle\int_{D}\big{[}dd^{c}(\rho+\accentset{\approx}{v})\big{]}^{n+1}\geq\int_{D}\big{[}dd^{c}(\rho+\accentset{\approx}{u})\big{]}^{n+1}. (3.6)

Taking 𝐞λ4C\mathbf{e}\leq\frac{\lambda}{4C},

ddc(ρ+u𝐞|z|2)12ddc(ρ+u).\displaystyle dd^{c}(\rho+u-\mathbf{e}|z|^{2})\geq\frac{1}{2}dd^{c}(\rho+u).

Together with the construction of u\accentset{\approx}{u} and v\accentset{\approx}{v} in (3.4), (3.5),

D[ddc(ρ+v)]n+1\displaystyle\int_{D}\big{[}dd^{c}(\rho+v)\big{]}^{n+1} D[ddc(ρ+u+h2|τ|2𝐞|z|2)]n+1\displaystyle\geq\int_{D}\Big{[}dd^{c}\big{(}\rho+u+\frac{h}{2}|\tau|^{2}-\mathbf{e}|z|^{2}\big{)}\Big{]}^{n+1}
hλn2n+1DΘn+1+D[ddc(ρ+u)]n+12𝐞ΛnDΘn+1.\displaystyle\geq\frac{h\lambda^{n}}{2^{n+1}}\int_{D}\Theta^{n+1}+\int_{D}\big{[}dd^{c}(\rho+u)\big{]}^{n+1}-2\mathbf{e}\Lambda^{n}\int_{D}\Theta^{n+1}. (3.7)

By picking 𝐞\mathbf{e} smaller, 𝐞hλn2n+4Λn\mathbf{e}\leq\frac{h\lambda^{n}}{2^{n+4}\Lambda^{n}}, and combining with (3.1), we have,

D[ddc(ρ+v)]n+1\displaystyle\int_{D}\big{[}dd^{c}(\rho+v)\big{]}^{n+1} D[ddc(ρ+u)]n+1+hλn2n+4DΘn+1\displaystyle\geq\int_{D}\big{[}dd^{c}(\rho+u)\big{]}^{n+1}+\frac{h\lambda^{n}}{2^{n+4}}\int_{D}\Theta^{n+1}
D[ddc(ρ+v)]n+1+hλn2n+4DΘn+1.\displaystyle\geq\int_{D}\big{[}dd^{c}(\rho+v)\big{]}^{n+1}+\frac{h\lambda^{n}}{2^{n+4}}\int_{D}\Theta^{n+1}. (3.8)

Since the second term of (3.8) is strictly positive, which leads to a contradiction, we complete the proof. ∎

Let φ~=Φ~Ψ\widetilde{\varphi}=\widetilde{\Phi}-\Psi be the solution of (Eε)(E_{\varepsilon}) after subtracting Ψ\Psi. According to Theorem 3.2, we have a uniform lower bound φ~0\widetilde{\varphi}\geq 0; hence, Φ~Ψ\widetilde{\Phi}\geq\Psi. The upper bound is easy to construct. Consider the function defined in X×ΣX\times\Sigma, H=2t(1t)H=2t(1-t). By restricting to each section Σx0={x0}×Σix0X×Σ\Sigma_{x_{0}}=\{x_{0}\}\times\Sigma\stackrel{{\scriptstyle i_{x_{0}}}}{{\hookrightarrow}}X\times\Sigma, we have

ix0(ΘΨ+ddcH)0<ix0(ΘΨ+ddcφ).\displaystyle i_{x_{0}}^{*}(\Theta_{\Psi}+dd^{c}H)\leq 0<i_{x_{0}}^{*}(\Theta_{\Psi}+dd^{c}\varphi).

Hence, ΔΣHddcφ~\Delta_{\Sigma}H\leq dd^{c}\widetilde{\varphi} in Σx0\Sigma_{x_{0}} and H=φ=0H=\varphi=0 on its boundary Σx0\partial\Sigma_{x_{0}}. The maximum principle on compact manifolds with boundary implies that φ~H\widetilde{\varphi}\leq H on each section. Hence, we get the desired uniform 𝒞0\mathcal{C}^{0} estimate,

ΨΦ~Ψ+H.\displaystyle\Psi\leq\widetilde{\Phi}\leq\Psi+H.

4. A priori estimate up to 𝒞1\mathcal{C}^{1}

For the 𝒞1\mathcal{C}^{1} bound, Blocki gives an explicit estimate in the compact setting in [6]. We generalize this estimate to the non-compact case. The 𝒞1\mathcal{C}^{1} boundary estimate follows directly from the fact that ΨΦ~Ψ+H\Psi\leq\widetilde{\Phi}\leq\Psi+H in X×ΣX\times\Sigma and Ψ\Psi, Φ~\widetilde{\Phi}, Ψ+H\Psi+H agree along X×ΣX\times\partial\Sigma. Let \nabla be the Levi-Civita connection of ΘΨ\Theta_{\Psi} on X×ΣX\times\Sigma. Then we have

|Φ~|ΘΨmax{|Ψ|ΘΨ,|(Ψ+H)|ΘΨ}, on X×Σ.\displaystyle|\nabla\widetilde{\Phi}|_{\Theta_{\Psi}}\leq\max\{|\nabla\Psi|_{\Theta_{\Psi}},|\nabla(\Psi+H)|_{\Theta_{\Psi}}\},\quad\text{ on }X\times\partial\Sigma.

Hence, supX×Σ|Φ~|ΘΨC\sup_{X\times\partial\Sigma}|\nabla\widetilde{\Phi}|_{\Theta_{\Psi}}\leq C, where CC is a uniform constant.

Proposition 4.1.

Let φ~=Φ~Ψ𝒞loc3(X×Σ)\widetilde{\varphi}=\widetilde{\Phi}-\Psi\in\mathcal{C}^{3}_{loc}(X\times\Sigma) be a solution of (Eε)(E_{\varepsilon}) and let \nabla be the Levi-Civita connection of the Kähler metric ΘΨ\Theta_{\Psi} on X×ΣX\times\Sigma. Assume that φ~\widetilde{\varphi} lies in the space 𝒞1(X×Σ,ΘΨ)\mathcal{C}^{1}(X\times\Sigma,\Theta_{\Psi}). Then,

supX×Σ|φ~|ΘΨC,\displaystyle\sup_{X\times\Sigma}|\nabla\widetilde{\varphi}|_{\Theta_{\Psi}}\leq C,

where CC is a positive constant depending only on upper bounds for |φ~||\widetilde{\varphi}|, on lower bounds for the bisectional curvature of ΘΨ\Theta_{\Psi}, and on nn, but not on ε\varepsilon.

Proof.

Suppose that infX×Σφ~=A\inf_{X\times\Sigma}\widetilde{\varphi}=A and supX×Σφ~=B\sup_{X\times\Sigma}\widetilde{\varphi}=B. Consider the following function,

α=logβγφ~,\displaystyle\alpha=\log\beta-\gamma\circ\widetilde{\varphi},

where β=|φ~|ΘΨ2\beta=|\nabla\widetilde{\varphi}|^{2}_{\Theta_{\Psi}} and γ:[A,B]\gamma:[A,B]\rightarrow\mathbb{R} is a smooth function to be determined later. According to the assumption that φ~\widetilde{\varphi} lies in the space 𝒞1\mathcal{C}^{1}, Yau’s maximum principle can be applied here. In particular, there exists a sequence in {xk}\{x_{k}\} in X×ΣX\times\Sigma^{\circ} such that,

limkα(xk)=supX×Σα,limk|α(xk)|ΘΨ=0,lim supkΔα(xk)0,\displaystyle\lim_{k\rightarrow\infty}\alpha(x_{k})=\sup_{X\times\Sigma}\alpha,\quad\lim_{k\rightarrow\infty}|\nabla\alpha(x_{k})|_{\Theta_{\Psi}}=0,\quad\limsup_{k\rightarrow\infty}\Delta\alpha(x_{k})\leq 0,

where Δ=ΔΘΨ\Delta=\Delta_{\Theta_{\Psi}}. Then, for a sufficiently small 𝐞>0\mathbf{e}>0 to be determined later and all k1k\gg 1, we have

α(xk)supX×Σα𝐞,|α(xk)|ΘΨ𝐞,Δα(xk)𝐞.\displaystyle\alpha(x_{k})\geq\sup_{X\times\Sigma}\alpha-\mathbf{e},\quad|\nabla\alpha(x_{k})|_{\Theta_{\Psi}}\leq\mathbf{e},\qquad\Delta\alpha(x_{k})\leq\mathbf{e}. (4.1)

Fixing O=xkO=x_{k} satisfying (4.1), we can pick the normal coordinates around OO. Let gg and g~\widetilde{g} denote the metric tensors corresponding to ΘΨ\Theta_{\Psi} and ΘΦ~=ΘΨ+ddcφ~\Theta_{\widetilde{\Phi}}=\Theta_{\Psi}+dd^{c}\widetilde{\varphi}. Then there exist local holomorphic coordinates near OO such that,

gij¯(O)=δij,gij¯,k(O)=0 and g~ij¯(O) is diagonal.\displaystyle g_{i\overline{j}}(O)=\delta_{ij},\quad g_{i\overline{j},k}(O)=0\quad\text{ and }\quad\widetilde{g}_{i\overline{j}}(O)\text{ is diagonal}.

By taking derivative of α\alpha,

αp=βpβ(γφ~)φ~p.\displaystyle\alpha_{p}=\frac{\beta_{p}}{\beta}-(\gamma^{\prime}\circ\widetilde{\varphi})\cdot\widetilde{\varphi}_{p}.

Combining with condition (4.1), |αp(O)|𝐞|\alpha_{p}(O)|\leq\mathbf{e}. Then, at the point OO, we have

αpp¯βpp¯β[(γ)2+γ′′]|φ~p|2γφ~pp¯𝐞|γ||φ~p|𝐞.\displaystyle\alpha_{p\overline{p}}\geq\frac{\beta_{p\overline{p}}}{\beta}-[(\gamma^{\prime})^{2}+\gamma^{\prime\prime}]|\widetilde{\varphi}_{p}|^{2}-\gamma^{\prime}\widetilde{\varphi}_{p\overline{p}}-\mathbf{e}|\gamma^{\prime}||\widetilde{\varphi}_{p}|-\mathbf{e}. (4.2)

If we write the local potential of g~ij¯\widetilde{g}_{i\overline{j}} as uu near OO, then the ε\varepsilon-geodesic equation is locally given by det(uij¯)=υ(ε)det(gij¯)\det(u_{i\overline{j}})=\upsilon(\varepsilon)\det(g_{i\overline{j}}). The direct derivative of the equation at OO gives,

pupp¯jupp¯=(logυ(ε))j.\displaystyle\sum_{p}\frac{u_{p\overline{p}j}}{u_{p\overline{p}}}=\big{(}\log\upsilon(\varepsilon)\big{)}_{j}. (4.3)

Also, notice that,

βpp¯Dβ+2Rejupp¯jφ~j¯+j|φ~jp|2+φ~pp¯2,\displaystyle\beta_{p\overline{p}}\geq-D\beta+2\operatorname{Re}\sum_{j}u_{p\overline{p}j}\widetilde{\varphi}_{\overline{j}}+\sum_{j}|\widetilde{\varphi}_{jp}|^{2}+\widetilde{\varphi}_{p\overline{p}}^{2},

where D-D is the negative lower bound of bisectional curvature of ΘΨ\Theta_{\Psi}. Recall that we have the assumption C1gij¯uij¯Cgij¯C^{-1}g_{i\overline{j}}\leq u_{i\overline{j}}\leq Cg_{i\overline{j}} and |φ~p|<C|\widetilde{\varphi}_{p}|<C, where CC is the constant from our assumption at the beginning of this section and we will get rid of this constant in the end. Together with (4.2) and (4.3), we have,

C𝐞pαpp¯upp¯(γD)p1upp¯+1βjp|φ~jp|upp¯2Re1βj(logυ(ε))jφ~j¯[(γ)2+γ′′]p|φ~p|2upp¯nγC(|γ|+1)𝐞.\displaystyle\begin{split}C\mathbf{e}\geq\sum_{p}\frac{\alpha_{p\overline{p}}}{u_{p\overline{p}}}\geq(\gamma^{\prime}&-D)\sum_{p}\frac{1}{u_{p\overline{p}}}+\frac{1}{\beta}\sum_{jp}\frac{|\widetilde{\varphi}_{jp}|}{u_{p\overline{p}}}\\ &-2\operatorname{Re}\frac{1}{\beta}\sum_{j}\big{(}\log\upsilon(\varepsilon)\big{)}_{j}\widetilde{\varphi}_{\overline{j}}\\ &-[(\gamma^{\prime})^{2}+\gamma^{\prime\prime}]\sum_{p}\frac{|\widetilde{\varphi}_{p}|^{2}}{u_{p\overline{p}}}-n\gamma^{\prime}-C(|\gamma^{\prime}|+1)\mathbf{e}.\end{split} (4.4)

According to Blocki’s key observation in [6], after modified in our case, at the point OO, we have

1βj,p|φ~jp|2upp¯(γ)2p|φ~p|2upp¯2γ2+C𝐞βC(1+|γ|)𝐞,\displaystyle\frac{1}{\beta}\sum_{j,p}\frac{|\widetilde{\varphi}_{jp}|^{2}}{u_{p\overline{p}}}\geq(\gamma^{\prime})^{2}\sum_{p}\frac{|\widetilde{\varphi}_{p}|^{2}}{u_{p\overline{p}}}-2\gamma^{\prime}-\frac{2+C\mathbf{e}}{\beta}-C(1+|\gamma^{\prime}|)\mathbf{e},

and assuming that β1\beta\geq 1, we have

2βRej(logυ)jφj¯2|logυ(ε)|β2(n+1)|(υ(ε)1n+1)|υ(ε)1n+1Vp1upp¯\displaystyle\frac{2}{\beta}\operatorname{Re}\sum_{j}\big{(}\log\upsilon\big{)}_{j}\varphi_{\overline{j}}\geq-2\frac{|\nabla\log\upsilon(\varepsilon)|}{\sqrt{\beta}}\geq-2(n+1)\frac{\big{|}\nabla\big{(}\upsilon(\varepsilon)^{\frac{1}{n+1}}\big{)}\big{|}}{\upsilon(\varepsilon)^{\frac{1}{n+1}}}\geq-V\sum_{p}\frac{1}{u_{p\overline{p}}}

where VV is a uniform constant satisfying

V2(n+1)|(υ(ε)1n+1)|.\displaystyle V\geq 2(n+1)\big{|}\nabla\big{(}\upsilon(\varepsilon)^{\frac{1}{n+1}}\big{)}\big{|}.

Combining with (4.4),

C(1+|γ|)𝐞(γDV)p1upp¯γ′′p|φ~p|2upp¯(n+2)γ2.\displaystyle C(1+|\gamma^{\prime}|)\mathbf{e}\geq(\gamma^{\prime}-D-V)\sum_{p}\frac{1}{u_{p\overline{p}}}-\gamma^{\prime\prime}\sum_{p}\frac{|\widetilde{\varphi}_{p}|^{2}}{u_{p\overline{p}}}-(n+2)\gamma^{\prime}-2. (4.5)

Now, we choose the function γ\gamma and the small number 𝐞>0\mathbf{e}>0 in (4.5) as follows. Let γ=(D+V+3)(tA)(BA)1(tA)2\gamma=(D+V+3)(t-A)-(B-A)^{-1}(t-A)^{2} and 𝐞C1(D+V+3)1\mathbf{e}\leq C^{-1}(D+V+3)^{-1}, then we have

p1upp¯+2BAp|φ~p|2upp¯3+(n+2)(D+V+3).\displaystyle\sum_{p}\frac{1}{u_{p\overline{p}}}+\frac{2}{B-A}\sum_{p}\frac{|\widetilde{\varphi}_{p}|^{2}}{u_{p\overline{p}}}\leq 3+(n+2)(D+V+3).

Then, it is straightforward to conclude that β(O)max{[(n+3)(D+V+3)]n+1n(BA),1}\beta(O)\leq\max\{[(n+3)(D+V+3)]^{n+1}n(B-A),1\}. Noting that βexp{𝐞+logβ(O)γφ~(O)+γφ~}\beta\leq\exp\{\mathbf{e}+\log\beta(O)-\gamma\circ\widetilde{\varphi}(O)+\gamma\circ\widetilde{\varphi}\}, hence, β\beta is controlled by some uniform constant only depending on φ~L\|\widetilde{\varphi}\|_{L^{\infty}}, DD, VV and nn. ∎

5. A priori estimate up to 𝒞1,1\mathcal{C}^{1,1}

First, we deal with the uniform 𝒞2\mathcal{C}^{2} boundary estimate on X×ΣX\times\partial\Sigma. The technique is to construct local barrier functions near the boundary, which is completely parallel to [8, 9, 17]. The statement is the following:

Lemma 5.1.

Let the data (X×Σ,ΘΨ,φ~)(X\times\Sigma,\Theta_{\Psi},\widetilde{\varphi}) be the same as in Proposition 4.1. Let \nabla denote the Levi-Civita connection of ΘΨ\Theta_{\Psi} on X×ΣX\times\Sigma. Then

supX×Σ|2φ~|ΘΨC,\displaystyle\sup_{X\times\partial\Sigma}|\nabla^{2}\widetilde{\varphi}|_{\Theta_{\Psi}}\leq C,

where the constant CC only depends on supX×Σ|φ~|ΘΨ\sup_{X\times\Sigma}|\nabla\widetilde{\varphi}|_{\Theta_{\Psi}} and on (X×Σ,ΘΨ)(X\times\Sigma,\Theta_{\Psi}).

Proof.

Fixing a point pX×Σp\in X\times\partial\Sigma, we pick the local holomorphic coordinates around the point pp such that the coordinates system is normal in XX and Σ\Sigma direction, we still pick the standard coordinate function of the annulus, denoted by {x1,,x2n,x2n+1=t,x2n=s}\{x_{1},\ldots,x_{2n},x_{2n+1}=t,x_{2n}=s\} and the corresponding holomorphic coordinates, zi=x2i1+ix2iz_{i}=x_{2i-1}+ix_{2i}. Throughout the proof, we assume the metric tensor gg associated with ΘΨ\Theta_{\Psi} satisfies mδijgij¯Mδijm\delta_{ij}\leq g_{i\overline{j}}\leq M\delta_{ij}. In general, we need to prove the boundary C2C^{2} estimate at pp in tangential-tangential, tangential-normal and normal-normal directions respectively. However, the tangential-tangential is trivial in our case and the normal-normal estimate follows directly from the tangential-normal estimate. Here, we briefly summarize the proof of tangential-normal estimate by explicitly constructing the barrier functions.

Consider a small neighborhood near pp, Bδ(p)=(X×Σ)Bδ(p)\displaystyle B^{\prime}_{\delta}(p)=(X\times\Sigma)\cap B_{\delta}(p), where the small constant δ\delta will be determined later. Firstly, we construct the following auxiliary function in Bδ(p)B^{\prime}_{\delta}(p),

v=φ~+Nt(1t),\displaystyle v=\widetilde{\varphi}+Nt(1-t), (5.1)

where NN is a large constant to be determined. Then, it can be easily checked that

Δ~vn+1mig~ii¯Ng~n+1,n+1¯,\displaystyle\widetilde{\Delta}v\leq n+1-m\sum_{i}\widetilde{g}^{i\overline{i}}-N\widetilde{g}^{n+1,\overline{n+1}},

where g~\widetilde{g} again denotes the metric tensor associated with ΘΦ~=ΘΨ+ddcφ~\Theta_{\widetilde{\Phi}}=\Theta_{\Psi}+dd^{c}\widetilde{\varphi} and Δ~\widetilde{\Delta} denotes the corresponding Laplacian. Notice that

m2ig~ii¯Ng~n+1,n+1¯mN1n+12(detg~)1n+1=mN1n+12υ(ε)1n+1(detg)1n+1.\displaystyle-\frac{m}{2}\sum_{i}\widetilde{g}^{i\overline{i}}-N\widetilde{g}^{n+1,\overline{n+1}}\leq-\frac{mN^{\frac{1}{n+1}}}{2(\det\widetilde{g})^{\frac{1}{n+1}}}=-\frac{mN^{\frac{1}{n+1}}}{2\upsilon(\varepsilon)^{\frac{1}{n+1}}}(\det g)^{-\frac{1}{n+1}}.

By taking N=[(n+1)(2/m)]n+1maxBδ(p)(detg)N=[(n+1)(2/m)]^{n+1}\max_{B_{\delta}^{\prime}(p)}(\det g), we have Δ~vm2ig~ii¯\displaystyle\widetilde{\Delta}v\leq-\frac{m}{2}\sum_{i}\widetilde{g}^{i\overline{i}}. Noting that φ~=Φ~Ψ0\widetilde{\varphi}=\widetilde{\Phi}-\Psi\geq 0, we have v0v\geq 0 on Bδ(p)\partial B^{\prime}_{\delta}(p). Then, the barrier functions can be constructed as follows:

w=Av+B|z|2±xkφ~, for 1k2n or k=2n+2.\displaystyle w=Av+B|z|^{2}\pm\frac{\partial}{\partial x_{k}}\widetilde{\varphi},\quad\text{ for }1\leq k\leq 2n\text{ or }k=2n+2.

By differentiating the Monge-Ampère equation (Eε)(E_{\varepsilon}) in the local coordinates,

±Δ~(xkφ~)=±(g~ij¯(g~)ij¯,kg~ij¯gij¯,k)C(1+g~ii¯),\displaystyle\pm\widetilde{\Delta}\Big{(}\frac{\partial}{\partial x_{k}}\widetilde{\varphi}\Big{)}=\pm\big{(}\widetilde{g}^{i\overline{j}}(\widetilde{g})_{i\overline{j},k}-\widetilde{g}^{i\overline{j}}g_{i\overline{j},k}\big{)}\leq C(1+\sum\widetilde{g}^{i\overline{i}}),

where AA and BB are large positive constants to be determined. According to the 𝒞1\mathcal{C}^{1} estimate of φ~\widetilde{\varphi}, we assume that |kφ~|C|\partial_{k}\widetilde{\varphi}|\leq C. By picking a very large constant BB such that, on Bδ(p)\partial B^{\prime}_{\delta}(p), B|z|2±kφ~0B|z|^{2}\pm\partial_{k}\widetilde{\varphi}\geq 0, we have w0w\geq 0 on Bδ(p)\partial B^{\prime}_{\delta}(p). Then, we choose a large constant AA such that Δ~w0\widetilde{\Delta}w\leq 0 in Bδ(p)B^{\prime}_{\delta}(p). Then, by maximum principle, w0w\geq 0 in Bδ(p)B^{\prime}_{\delta}(p). Together with the fact that w(p)=0w(p)=0, we have tw0\partial_{t}w\geq 0 at pp, which implies the tangential-normal estimate on the boundary. ∎

Lemma 5.1 together with Yau’s standard calculation on Laplacian estimate implies the following interior Laplacian estimate, referring to [31].

Lemma 5.2.

Let φ~\widetilde{\varphi} be the solution of (Eε)(E_{\varepsilon}) and Δ\Delta, Δ~\widetilde{\Delta}, the Laplacian operators of g=ΘΨg=\Theta_{\Psi} and g~=ΘΦ~=ΘΨ+ddcφ~\widetilde{g}=\Theta_{\widetilde{\Phi}}=\Theta_{\Psi}+dd^{c}\widetilde{\varphi} respectively. Then, for any constant CC,

Δ~(eCφ~(n+1+Δφ~))\displaystyle\widetilde{\Delta}\big{(}e^{-C\widetilde{\varphi}}(n+1+\Delta\widetilde{\varphi})\big{)}\geq eCφ~(Δlogυ(ε)(n+1)2infil(Rii¯ll¯))\displaystyle\quad e^{-C\widetilde{\varphi}}\big{(}\Delta\log\upsilon(\varepsilon)-(n+1)^{2}\inf_{i\neq l}(R_{i\overline{i}l\overline{l}})\big{)}
CeCφ~(n+1)(n+1+Δφ~)\displaystyle-Ce^{-C\widetilde{\varphi}}(n+1)(n+1+\Delta\widetilde{\varphi})
+(C+infil(Rii¯ll¯))eCφ~(n+1+Δφ~)1+1nυ(ε)1,\displaystyle+(C+\inf_{i\neq l}(R_{i\overline{i}l\overline{l}}))e^{-C\widetilde{\varphi}}(n+1+\Delta\widetilde{\varphi})^{1+\frac{1}{n}}\upsilon(\varepsilon)^{-1},

where RR denotes the curvature tensor of gg. From this, we can deduce the estimate

supX×Σ|Δφ~|C(1+supX×Σ|Δφ~|),\displaystyle\sup_{X\times\Sigma}|\Delta\widetilde{\varphi}|\leq C(1+\sup_{X\times\partial\Sigma}|\Delta\widetilde{\varphi}|),

where CC only depends on supX×Σφ~\sup_{X\times\Sigma}\widetilde{\varphi} and on a negative lower bound of infil(Rii¯ll¯)\inf_{i\neq l}(R_{i\overline{i}l\overline{l}}).

Lemma 5.2, together with Lemma 5.1, implies that there exists a uniform constant CC only depending on supX×ΣΔφ~\sup_{X\times\Sigma}\Delta\widetilde{\varphi} such that εC1gij¯g~ij¯Cgij¯\varepsilon C^{-1}g_{i\overline{j}}\leq\widetilde{g}_{i\overline{j}}\leq Cg_{i\overline{j}}. This is already enough to apply the standard local regularity theory of the Monge-Ampère equation to prove 𝒞k,α\mathcal{C}^{k,\alpha} estimates for any k2k\geq 2 that depend on a positive lower bound for ε\varepsilon. In this way the equation (Eε)(E_{\varepsilon}) can be solved using the continuity path (Es)(E_{s}), s[ε,1]s\in[\varepsilon,1]. However, in order to construct an honest geodesic by letting ε0\varepsilon\to 0, we require a full 𝒞1,1\mathcal{C}^{1,1} estimate which is uniform in ε\varepsilon. In [12], 𝒞1,1\mathcal{C}^{1,1} regularity is proved in the compact case. The method can also be applied in the ALE Kähler setting.

Proposition 5.3.

Let the data (X×Σ,ΘΨ,φ~)(X\times\Sigma,\Theta_{\Psi},\widetilde{\varphi}) be the same as in Proposition 4.1. If φ~\widetilde{\varphi} lies in the space 𝒞2(X×Σ,ΘΨ)\mathcal{C}^{2}(X\times\Sigma,\Theta_{\Psi}), then there exists a constant CC such that

|2φ~|ΘΨC,\displaystyle|\nabla^{2}\widetilde{\varphi}|_{\Theta_{\Psi}}\leq C,

where \nabla again denotes the Levi-Civita connection of the metric ΘΨ\Theta_{\Psi} and CC depends only on (X×Σ,ΘΨ)(X\times\Sigma,\Theta_{\Psi}) and on supX×Σ|φ~|\sup_{X\times\Sigma}|\widetilde{\varphi}|, supX×Σ|φ~|ΘΨ\sup_{X\times\Sigma}|\nabla\widetilde{\varphi}|_{\Theta_{\Psi}}, supX×Σ|Δφ~|\sup_{X\times\Sigma}|\Delta\widetilde{\varphi}|, supX×Σ|2φ~|ΘΨ\sup_{X\times\partial\Sigma}|\nabla^{2}\widetilde{\varphi}|_{\Theta_{\Psi}}.

Proof.

We again write gg for the metric tensor associated with ΘΨ\Theta_{\Psi}. Let λ1(2φ~)\lambda_{1}(\nabla^{2}\widetilde{\varphi}) be the largest eigenvalue of the real Hessian 2φ~\nabla^{2}\widetilde{\varphi}. By observing that there exists a uniform constant CC such that λ1(2φ~)|2φ~|gCλ1(2φ~)+C\lambda_{1}(\nabla^{2}\widetilde{\varphi})\leq|\nabla^{2}\widetilde{\varphi}|_{g}\leq C\lambda_{1}(\nabla^{2}\widetilde{\varphi})+C, it suffices to prove that λ1(2φ~)\lambda_{1}(\nabla^{2}\widetilde{\varphi}) has a uniform upper bound. Consider the following quantity,

Q=logλ1(2φ~)+h(|φ~|g2)Aφ~,\displaystyle Q=\log\lambda_{1}(\nabla^{2}\widetilde{\varphi})+h(|\nabla\widetilde{\varphi}|_{g}^{2})-A\widetilde{\varphi},

where hh is defined to be h(s)=12log(1+supX×Σ|φ~|g2s)\displaystyle h(s)=-\frac{1}{2}\log\big{(}1+\sup_{X\times\Sigma}|\nabla\widetilde{\varphi}|_{g}^{2}-s\big{)} and AA is a uniform large positive constant to be determined later. We can further modify this quantity to Q𝐞=Q𝐞rQ_{\mathbf{e}}=Q-\mathbf{e}r, where 𝐞\mathbf{e} is a small positive constant to be determined later. According to the assumption that |2φ~||\nabla^{2}\widetilde{\varphi}| is bounded and hence so is QQ, the modified quantity Q𝐞Q_{\mathbf{e}} attains its maximum at some point x𝐞X×Σx_{\mathbf{e}}\in X\times\Sigma. The same argument as in Lemma 2.2 implies that lim𝐞0Q(x𝐞)=supX×ΣQ\lim_{\mathbf{e}\rightarrow 0}Q(x_{\mathbf{e}})=\sup_{X\times\Sigma}Q. In the following, we assume 𝐞\mathbf{e} is small enough such that |Q(x𝐞)supX×ΣQ|<1|Q(x_{\mathbf{e}})-\sup_{X\times\Sigma}Q|<1 and always write p=x𝐞p=x_{\mathbf{e}}. Since Q𝐞Q_{\mathbf{e}} might not be smooth at pp if the eigenspace of λ1(2φ~)(p)\lambda_{1}(\nabla^{2}\widetilde{\varphi})(p) has dimension greater than one, a perturbation argument used in [12] can be applied to the quantity Q𝐞Q_{\mathbf{e}} here.

Fix normal coordinates (z1,,zn+1)(z_{1},\ldots,z_{n+1}) with respect to gg at pp such that (φ~ij¯)(\widetilde{\varphi}_{i\overline{j}}) is diagonal at pp. Define the corresponding real coordinates (x1,,x2n)(x_{1},\ldots,x_{2n}) by zi=x2i1+ix2iz_{i}=x_{2i-1}+ix_{2i}. Let λ1λ2λ2n\lambda_{1}\geq\lambda_{2}\geq\ldots\geq\lambda_{2n} be the eigenvalues of 2φ~\nabla^{2}\widetilde{\varphi} at pp and V1,,V2nV_{1},\ldots,V_{2n}, the corresponding unit eigenvectors at pp. The eigenvectors can be extended to vector fields with constant coefficients in a small neighborhood of pp, also denoted by V1,,V2nV_{1},\ldots,V_{2n}, and can be represented by Vα=VαβxβV_{\alpha}=V^{\beta}_{\alpha}{\partial_{x_{\beta}}} in the local coordinates. The perturbation argument is to perturb 2φ~\nabla^{2}\widetilde{\varphi} locally around pp and to ensure that λ1>λ2\lambda_{1}>\lambda_{2} near pp. Precisely, consider the following locally defined tensor field,

P=α,β(δαβV1αV1β)dxαdxβ.\displaystyle P=\sum_{\alpha,\beta}\big{(}\delta_{\alpha\beta}-V^{\alpha}_{1}V^{\beta}_{1}\big{)}dx_{\alpha}\otimes dx_{\beta}.

Let λi=λi(2φ~P)\lambda_{i}^{\prime}=\lambda_{i}(\nabla^{2}\widetilde{\varphi}-P). Then, one can easily check that λ1(p)=λ1(p)\lambda^{\prime}_{1}(p)=\lambda_{1}(p) and λi(p)=λi(p)1\lambda^{\prime}_{i}(p)=\lambda_{i}(p)-1 for i2i\geq 2. Hence, there exists a neighborhood of pp such that λ1>λ2λ2n\lambda^{\prime}_{1}>\lambda^{\prime}_{2}\geq\ldots\geq\lambda^{\prime}_{2n} and λ1λ1\lambda^{\prime}_{1}\leq\lambda_{1}. Consider the following perturbed quantities,

Q^=logλ1+h(|φ~|g2)Aφ~,Q^𝐞=Q^𝐞r.\displaystyle\hat{Q}=\log\lambda^{\prime}_{1}+h(|\nabla\widetilde{\varphi}|^{2}_{g})-A\widetilde{\varphi},\quad\hat{Q}_{\mathbf{e}}=\hat{Q}-\mathbf{e}r.

Therefore, Q^𝐞\hat{Q}_{\mathbf{e}} is a smooth quantity with a local maximum at pp. Then, we have,

|dQ^(p)|gC𝐞,ΔQ^(p)C𝐞.\displaystyle|d\hat{Q}(p)|_{g}\leq C\mathbf{e},\quad\Delta\hat{Q}(p)\leq C\mathbf{e}.

The following inequality follows directly from the calculation in [12, Lemma 2.1]. The only information we need in the calculation is the second derivative of the Monge-Ampère equation at pp. We will not repeat the details here. By assuming λ11\lambda_{1}^{\prime}\geq 1 at pp, and again writing g~\widetilde{g} for the metric tensor associated with ΘΦ~=ΘΨ+ddcφ~\Theta_{\widetilde{\Phi}}=\Theta_{\Psi}+dd^{c}\widetilde{\varphi}, we have

ΔQ^ 2α>1g~ii¯|i(φ~VαV1)|2λ1(λ1λα)+g~ii¯g~jj¯|V1(g~ij¯)|2λ1g~ii¯|i(φ~V1V1)|2λ12+hkg~ii¯(|φ~ik|2+|φ~ik¯|2)+h′′g~ii¯|i|φ~|g2|+(AB)ig~ii¯An,\displaystyle\begin{split}\Delta\hat{Q}\geq&\ 2\sum_{\alpha>1}\frac{\widetilde{g}^{i\overline{i}}|\partial_{i}(\widetilde{\varphi}_{V_{\alpha}V_{1}})|^{2}}{\lambda_{1}(\lambda_{1}-\lambda_{\alpha})}+\frac{\widetilde{g}^{i\overline{i}}\widetilde{g}^{j\overline{j}}\big{|}V_{1}\big{(}\widetilde{g}_{i\overline{j}}\big{)}\big{|}^{2}}{\lambda_{1}}-\frac{\widetilde{g}^{i\overline{i}}|\partial_{i}(\widetilde{\varphi}_{V_{1}V_{1}})|^{2}}{\lambda_{1}^{2}}\\ &+h^{\prime}\sum_{k}\widetilde{g}^{i\overline{i}}\big{(}|\widetilde{\varphi}_{ik}|^{2}+|\widetilde{\varphi}_{i\overline{k}}|^{2}\big{)}+h^{\prime\prime}\widetilde{g}^{i\overline{i}}\big{|}\partial_{i}|\nabla\widetilde{\varphi}|^{2}_{g}\big{|}\\ &+(A-B)\sum_{i}\widetilde{g}^{i\overline{i}}-An,\end{split} (5.2)

where the constant BB only depends on (X×Σ,g)(X\times\Sigma,g) and supX×Σ|φ~|g\sup_{X\times\Sigma}|\nabla\widetilde{\varphi}|_{g}. To cancel the annoying terms, we deal with the third term in (5.2), λ12g~ii¯|i(φ~V1V1)|2\displaystyle{{\lambda_{1}^{-2}}\widetilde{g}^{i\overline{i}}|\partial_{i}(\widetilde{\varphi}_{V_{1}V_{1}})|^{2}}. To estimate the term, we split it into the following two parts,

I1=(12δ)g~ii¯|i(φ~V1V1)|2λ12,\displaystyle I_{1}=(1-2\delta)\frac{\widetilde{g}^{i\overline{i}}|\partial_{i}(\widetilde{\varphi}_{V_{1}V_{1}})|^{2}}{\lambda_{1}^{2}},
I2=2δg~ii¯|i(φ~V1V1)|2λ12,\displaystyle I_{2}=2\delta\frac{\widetilde{g}^{i\overline{i}}|\partial_{i}(\widetilde{\varphi}_{V_{1}V_{1}})|^{2}}{\lambda_{1}^{2}},

where 0<δ<1/40<\delta<1/4 is to be determined later. For I1I_{1}, referring to [12, Lemma 2.2], by assuming that λ1D/δ\lambda_{1}^{\prime}\geq D/\delta, where DD only depends on (X×Σ,g)(X\times\Sigma,g) and supX×ΣΔφ~\sup_{X\times\Sigma}\Delta\widetilde{\varphi}, we have

I1i,jg~ii¯g~jj¯|V1(g~ij¯)|λ1+2α>1ig~ii¯|i(φ~VαV1)|2λ1(λ1λα)+ig~ii¯.\displaystyle I_{1}\leq\sum_{i,j}\frac{\widetilde{g}^{i\overline{i}}\widetilde{g}^{j\overline{j}}\big{|}V_{1}\big{(}\widetilde{g}_{i\overline{j}}\big{)}\big{|}}{\lambda_{1}}+2\sum_{\alpha>1}\sum_{i}\frac{\widetilde{g}^{i\overline{i}}|\partial_{i}(\widetilde{\varphi}_{V_{\alpha}V_{1}})|^{2}}{\lambda_{1}(\lambda_{1}-\lambda_{\alpha})}+\sum_{i}\widetilde{g}^{i\overline{i}}. (5.3)

To estimate I2I_{2}, recall the fact that dQ^𝐞=0d\hat{Q}_{\mathbf{e}}=0 and apply the derivative of eigenvalues referring to [12, Lemma 5.2]. Then, we have

I2=2δig~ii¯|Aφ~i+hi|φ~|g2𝐞ri|28δA2ig~ii¯|φ~i|2+2(h)2ig~ii¯|i|φ~|g2|2+C𝐞ig~ii¯.\displaystyle\begin{split}I_{2}&=2\delta\sum_{i}\widetilde{g}^{i\overline{i}}\big{|}A\widetilde{\varphi}_{i}+h^{\prime}\partial_{i}|\nabla\widetilde{\varphi}|^{2}_{g}-\mathbf{e}r_{i}\big{|}^{2}\\ &\leq 8\delta A^{2}\sum_{i}\widetilde{g}^{i\overline{i}}|\widetilde{\varphi}_{i}|^{2}+2(h^{\prime})^{2}\sum_{i}\widetilde{g}^{i\overline{i}}|\partial_{i}|\nabla\widetilde{\varphi}|^{2}_{g}|^{2}+C\mathbf{e}\sum_{i}\widetilde{g}^{i\overline{i}}.\end{split} (5.4)

Combining (5.2), (5.3), (5.4) and ΔQ^C𝐞\Delta\hat{Q}\leq C\mathbf{e}, then, by assuming λ1D/δ\lambda_{1}^{\prime}\geq D/\delta, we have

C𝐞\displaystyle C\mathbf{e}\geq hkg~ii¯(|φ~ik|2+|φ~ik¯|2)+(h′′2(h)2)g~ii¯|i|φ~|2|\displaystyle\ h^{\prime}\sum_{k}\widetilde{g}^{i\overline{i}}\big{(}|\widetilde{\varphi}_{ik}|^{2}+|\widetilde{\varphi}_{i\overline{k}}|^{2}\big{)}+\big{(}h^{\prime\prime}-2(h^{\prime})^{2}\big{)}\widetilde{g}^{i\overline{i}}\big{|}\partial_{i}|\nabla\widetilde{\varphi}|^{2}\big{|}
8δA2g~ii¯|φ~i|2+(ABC𝐞)ig~ii¯An.\displaystyle-8\delta A^{2}\widetilde{g}^{i\overline{i}}|\widetilde{\varphi}_{i}|^{2}+(A-B-C\mathbf{e})\sum_{i}\widetilde{g}^{i\overline{i}}-An.

Notice that h′′=2(h)2h^{\prime\prime}=2(h^{\prime})^{2}. Picking 𝐞1/C\mathbf{e}\leq 1/C, A=B+2A=B+2 and δ=(8A2(supX×Σ|φ~|2+1))1\displaystyle\delta=\big{(}8A^{2}(\sup_{X\times\Sigma}|\nabla\widetilde{\varphi}|^{2}+1)\big{)}^{-1}, then we have

hkg~ii¯(|φ~ik|2+|φ~ik¯|2)+ig~ii¯An+1.\displaystyle h^{\prime}\sum_{k}\widetilde{g}^{i\overline{i}}\big{(}|\widetilde{\varphi}_{ik}|^{2}+|\widetilde{\varphi}_{i\overline{k}}|^{2}\big{)}+\sum_{i}\widetilde{g}^{i\overline{i}}\leq An+1.

Recall g~ij¯Cgij¯\widetilde{g}_{i\overline{j}}\leq Cg_{i\overline{j}}, where CC only depends on supX×ΣΔφ~\sup_{X\times\Sigma}\Delta\widetilde{\varphi}. Hence, at pp, g~ii¯C1\widetilde{g}^{i\overline{i}}\geq C^{-1}. Then,

λ1(p)max{Dδ,{(An+1)Cn}(1+supX×Σ|φ~|g2)}.\displaystyle\lambda_{1}(p)\leq\max\Big{\{}\frac{D}{\delta},\big{\{}(An+1)C-n\big{\}}(1+\sup_{X\times\Sigma}|\nabla\widetilde{\varphi}|_{g}^{2})\Big{\}}.

Together with the fact that supX×ΣQQ(p)+1\sup_{X\times\Sigma}Q\leq Q(p)+1, we prove that supX×Σλ1\sup_{X\times\Sigma}\lambda_{1} is bounded by some uniform constant. ∎

6. The asymptotic behavior of ε\varepsilon-geodesics

In this section, we prove Theorem B on the asymptotic behavior of ε\varepsilon-geodesics for a fixed ε>0\varepsilon>0. We use the notation introduced before Theorem B and we assume ψ0,ψ1γ(ω)\psi_{0},\psi_{1}\in\mathcal{H}_{-\gamma}(\omega) (γ>0)(\gamma>0). We are really interested in the case when γ=22τ~-\gamma=2-2\tilde{\tau} due to theorem 1.1. In ε\varepsilon-geodesic equation (Eε)(E_{\varepsilon}), the derivatives of function υ(ε)\upsilon(\varepsilon) decays at infinity with order ς-\varsigma, |kυ(ε)|O(rςk)|\nabla^{k}\upsilon(\varepsilon)|\leq O(r^{-\varsigma-k}) with ςγ\varsigma\geq\gamma for k1k\geq 1. Without loss of generality, we assume ςγ>τ\varsigma\geq\gamma>\tau, otherwise theorem B can be proved more easily without iteration (step 3).

We also write φε=ΦεΨ\varphi_{\varepsilon}=\Phi_{\varepsilon}-\Psi, so that the solution is given by Θ+ddcΦε=ΘΨ+ddcφε\Theta+dd^{c}\Phi_{\varepsilon}=\Theta_{\Psi}+dd^{c}\varphi_{\varepsilon} with φε=0\varphi_{\varepsilon}=0 on X×ΣX\times\partial\Sigma. In Aleyasin [1], a rough idea is given to prove the asymptotic behavior of ε\varepsilon-geodesics by constructing barrier functions in the (strictly easier) special case where the asymptotic coordinates are JJ-holomorphic and the decay rate of the ALE Kähler metric to the Euclidean metric is high enough. However, even in this special case, the details are more involved than what is suggested in [1]. Here we give a complete proof in the general setting.

Step 1: Differentiating the Monge-Ampère equation.

The Monge-Ampère equation can be written explicitly in the asymptotic coordinates of X×ΣX\times\Sigma. As the complex structure JJ of XX does not coincide with the Euclidean complex structure J0J_{0} of the asymptotic coordinates in general, we will use real coordinates for clarity. By passing to the universal covering of the end, we are able to work with the global coordinates. Precisely, let {z1,,zn}\{z_{1},\ldots,z_{n}\} be the asymptotic complex coordinates of n\BR\mathbb{C}^{n}\backslash B_{R} and w=t+isw=t+is the complex coordinate of Σ\Sigma. The corresponding real coordinates are {x1,,x2n,\{x_{1},\ldots,x_{2n}, x2n+1=t,x2n+2=s}x_{2n+1}=t,x_{2n+2}={s}\}, where zk=x2k1+ix2kz_{k}=x_{2k-1}+ix_{2k} for k=1,,nk=1,\ldots,n. From now on:

  • \bullet

    Latin indices i,j,i,j,\ldots will denote the real coordinates from 11 to 2n+22n+2.

  • \bullet

    Greek indices α,β,\alpha,\beta,\ldots will denote the real coordinates from 11 to 2n2n.

  • \bullet

    The bold Greek indices 𝝁,𝝂\boldsymbol{\mu},\boldsymbol{\nu} will denote the real coordinates from 2n+12n+1 to 2n+22n+2.

In these coordinates, we write the Riemannian metric tensors corresponding to ΘΨ\Theta_{\Psi} and ΘΨ+ddcφε\Theta_{\Psi}+dd^{c}\varphi_{\varepsilon} as gijg_{ij} and (gφε)ij(g_{\varphi_{\varepsilon}})_{ij}, respectively.

Throughout this section, we work in the asymptotic chart of XX. This allows us to use the Euclidean metric on (2nBR)×Σ(\mathbb{R}^{2n}\setminus B_{R})\times\Sigma as a reference metric to measure derivatives. This is helpful because it enables us to write down equations with a good structure. Let ||0|\cdot|_{0} denote the Euclidean length, 0\nabla_{0} the Euclidean Levi-Civita connection and 0,X\nabla_{0,X} (0,Σ\nabla_{0,\Sigma}) the component of 0\nabla_{0} acting only in the space (time) directions on (2nBR)×Σ(\mathbb{R}^{2n}\setminus B_{R})\times\Sigma.

Then, the equation (Eε)(E_{\varepsilon}) can be written as

det((gφε)ij)=υ(ε)det(gij).\displaystyle\sqrt{\det{\big{(}(g_{{\varphi_{\varepsilon}}})_{ij}\big{)}}}=\upsilon(\varepsilon)\sqrt{\det(g_{ij})}. (6.1)

Recall that υ\upsilon satisfies conditions in (1.8). By differentiating the log of both sides by Dα=/xαD_{\alpha}=\partial/\partial_{x_{\alpha}}, we have

gφεijDα(gφε)ij=gijDαgij+Dαlogυ(ε).\displaystyle g_{\varphi_{\varepsilon}}^{ij}D_{\alpha}(g_{\varphi_{\varepsilon}})_{ij}=g^{ij}D_{\alpha}g_{ij}+D_{\alpha}\log\upsilon(\varepsilon). (6.2)

The first goal is to rewrite the equation (6.2) to be an elliptic equation in terms of DαφεD_{\alpha}{\varphi_{\varepsilon}}. Let e1,,e2n+2e_{1},\ldots,e_{2n+2} represent the real coordinate vector fields of x1,,x2n+2x_{1},\ldots,x_{2n+2}. Notice that (gφε)ij=gij+ddcφε(ei,Jej)(g_{\varphi_{\varepsilon}})_{ij}=g_{ij}+dd^{c}{\varphi_{\varepsilon}}(e_{i},Je_{j}). We compute DαD_{\alpha} of the second term:

Dα[ddcφε(ei,Jej)]=dJd(Dαφε)(ei,Jej)d(DαJ)dφε(ei,Jej)dJdφε(ei,(DαJ)ej).\displaystyle\begin{split}D_{\alpha}[dd^{c}{\varphi_{\varepsilon}}(e_{i},Je_{j})]=&-d\circ J\circ d(D_{\alpha}{\varphi_{\varepsilon}})(e_{i},Je_{j})-d\circ(D_{\alpha}J)\circ d{\varphi_{\varepsilon}}(e_{i},Je_{j})\\ &-d\circ J\circ d{\varphi_{\varepsilon}}(e_{i},(D_{\alpha}J)e_{j}).\end{split} (6.3)

Observe that DαJD_{\alpha}J is completely horizontal because JJ preserves the product structure of the tangent bundle T((2nBR)×Σ)T((\mathbb{R}^{2n}\setminus B_{R})\times\Sigma) and J|TΣJ|_{T\Sigma} is constant. Thus,

DαJ=(DαJ)ξβ(eξeβ),(DαJ)ξ𝝁=0,(DαJ)𝝂β=0,(DαJ)𝝂𝝁=0,D_{\alpha}J=(D_{\alpha}J)^{\beta}_{\xi}(e_{\xi}^{*}\otimes e_{\beta}),\quad(D_{\alpha}J)_{\xi}^{\boldsymbol{\mu}}=0,\quad(D_{\alpha}J)^{\beta}_{\boldsymbol{\nu}}=0,\quad(D_{\alpha}J)^{\boldsymbol{\mu}}_{\boldsymbol{\nu}}=0, (6.4)

where the coefficients (DαJ)ξβ(D_{\alpha}J)^{\beta}_{\xi} depend only on x1,,x2nx_{1},\ldots,x_{2n} and not on x2n+1,x2n+2x_{2n+1},x_{2n+2}. In the same way, we can also see that

|0,Xm(DαJ)|0=O(rτ1m)(allm0),0,Σm(DαJ)=0(allm1).|\nabla_{0,X}^{m}(D_{\alpha}J)|_{0}=O(r^{-\tau-1-m})\;\,(\text{all}\;m\geq 0),\;\,\nabla_{0,\Sigma}^{m}(D_{\alpha}J)=0\;\,(\text{all}\;m\geq 1). (6.5)

Moreover, it is obvious that

Δgφε(Dαφε)=trgφε(ddc(Dαφε)(,J))=gφεijddc(Dαφε)(ei,Jej).\Delta_{g_{\varphi_{\varepsilon}}}(D_{\alpha}{\varphi_{\varepsilon}})=\operatorname{tr}_{g_{{\varphi_{\varepsilon}}}}(dd^{c}(D_{\alpha}{\varphi_{\varepsilon}})(\cdot,J\cdot))=g^{ij}_{{\varphi_{\varepsilon}}}dd^{c}(D_{\alpha}{\varphi_{\varepsilon}})(e_{i},Je_{j}). (6.6)

Then, (6.3)–(6.4) imply that

gφεijDα[ddcφε(ei,Jej)]=Δgφε(Dαφε)+𝐎(rτ1)gφε100,Xφε+𝐎(rτ2)gφε10,Xφε,\displaystyle\begin{split}g_{\varphi_{\varepsilon}}^{ij}D_{\alpha}[dd^{c}{\varphi_{\varepsilon}}(e_{i},Je_{j})]&=\Delta_{g_{\varphi_{\varepsilon}}}(D_{\alpha}{\varphi_{\varepsilon}})+\mathbf{O}(r^{-\tau-1})\circledast g_{\varphi_{\varepsilon}}^{-1}\circledast\nabla_{0}\nabla_{0,X}{\varphi_{\varepsilon}}\\ &+\mathbf{O}(r^{-\tau-2})\circledast g_{\varphi_{\varepsilon}}^{-1}\circledast\nabla_{0,X}\varphi_{\varepsilon},\end{split} (6.7)

where \circledast denotes a contraction and 𝐎\mathbf{O} denotes the following behavior of a tensor TT:

T=𝐎(rρ):|0,XmT|0=O(rρm)(allm0),0,ΣmT=0(allm1).T=\mathbf{O}(r^{-\rho}):\Longleftrightarrow|\nabla_{0,X}^{m}T|_{0}=O(r^{-\rho-m})\;\,(\text{all}\;m\geq 0),\;\,\nabla_{0,\Sigma}^{m}T=0\;\,(\text{all}\;m\geq 1).

Then, abbreviating the estimates

|0,Xm(Dαgij)|0=O(rτ1m)(allm0),|0,Xm0,Σ(Dαgij)|0=O(r2τ1m)(allm0),0,Σm(Dαgij)=0(allm2),\begin{split}|\nabla_{0,X}^{m}(D_{\alpha}g_{ij})|_{0}=&\;O(r^{-\tau-1-m})\;\,(\text{all}\;m\geq 0),\\ |\nabla_{0,X}^{m}\nabla_{0,\Sigma}(D_{\alpha}g_{ij})|_{0}=O(r^{-2\tau-1-m})\;\,&(\text{all}\;m\geq 0),\;\,\nabla_{0,\Sigma}^{m}(D_{\alpha}g_{ij})=0\;\,(\text{all}\;m\geq 2),\end{split}

by Dαgij=𝐎^(rτ1)D_{\alpha}g_{ij}=\widehat{\mathbf{O}}(r^{-\tau-1}) and

|0,Xm0,ΣkDαlogυ(ε)|0=O(rς1m)(all m,k0),\displaystyle|\nabla^{m}_{0,X}\nabla^{k}_{0,\Sigma}D_{\alpha}\log\upsilon(\varepsilon)|_{0}=O(r^{-\varsigma-1-m})\ (\text{all }m,k\geq 0),

by Dαlogυ(ε)=𝐎^^(rς1)D_{\alpha}\log\upsilon(\varepsilon)=\widehat{\widehat{\mathbf{O}}}(r^{-\varsigma-1}) the equation (6.2) can be rewritten as

Δgφε(Dαφε)=𝐎(rτ1)gφε100,Xφε+𝐎(rτ2)gφε10,Xφε+Dαgij(gφεijgij),Dαgij=𝐎^(rτ1),+𝐎^^(rς1)\displaystyle\begin{split}\Delta_{g_{\varphi_{\varepsilon}}}(D_{\alpha}{\varphi_{\varepsilon}})&=\mathbf{O}(r^{-\tau-1})\circledast g_{\varphi_{\varepsilon}}^{-1}\circledast\nabla_{0}\nabla_{0,X}{\varphi_{\varepsilon}}\\ &+\mathbf{O}(r^{-\tau-2})\circledast g_{\varphi_{\varepsilon}}^{-1}\circledast\nabla_{0,X}\varphi_{\varepsilon}\\ &+D_{\alpha}g_{ij}\cdot(g_{\varphi_{\varepsilon}}^{ij}-g^{ij}),\quad D_{\alpha}g_{ij}=\widehat{\mathbf{O}}(r^{-\tau-1}),\\ &+\widehat{\widehat{\mathbf{O}}}(r^{-\varsigma-1})\end{split} (6.8)

We will later use this formula in full but for now it is enough to take absolute values. Using the fact that Λ1εδij(gφε)ijΛδij\Lambda^{-1}\varepsilon\delta_{ij}\leq(g_{{\varphi_{\varepsilon}}})_{ij}\leq\Lambda\delta_{ij}, and according to the uniform estimates of |0φε|0|\nabla_{0}{\varphi_{\varepsilon}}|_{0} and |02φε|0|\nabla_{0}^{2}{\varphi_{\varepsilon}}|_{0} from Theorem A, the formula (6.8) implies that

|Δgφε(Dαφε)|Cε1rτ1\displaystyle|\Delta_{g_{\varphi_{\varepsilon}}}(D_{\alpha}{\varphi_{\varepsilon}})|\leq C\varepsilon^{-1}r^{-\tau-1} (6.9)

for some constant C=C(φε𝒞2(X×Σ,ΘΨ),Λ,g,J)C=C(\|{\varphi_{\varepsilon}}\|_{\mathcal{C}^{2}(X\times\Sigma,\Theta_{\Psi})},\Lambda,g,J) bounded above independently of ε\varepsilon.

Step 2: Barrier estimate of the first derivatives.

The next target is to construct the upper barrier and lower barrier functions to control |Dαφε||D_{\alpha}{\varphi_{\varepsilon}}|. Consider a smooth cutoff function χ:0\chi:\mathbb{R}_{\geq 0}\rightarrow\mathbb{R} satisfying χ(x)=0\chi(x)=0 for x1x\leq 1, χ(x)=1\chi(x)=1 for x2x\geq 2 and |χ(x)|4|\chi^{\prime}(x)|\leq 4, |χ′′(x)|4|\chi^{\prime\prime}(x)|\leq 4 for 1x21\leq x\leq 2. The function DαφεD_{\alpha}{\varphi_{\varepsilon}} can be extended smoothly to X×ΣX\times\Sigma by defining

h=χR0Dαφε,\displaystyle h=\chi_{R_{0}}\cdot D_{\alpha}{\varphi_{\varepsilon}}, (6.10)

where R0R_{0} is a large positive constant to be determined later such that {r(p)R0/2}\{r(p)\geq R_{0}/2\} is contained in the asymptotic chart of XX and χR0(p)=χ(r(p)/R0)\chi_{R_{0}}(p)=\chi(r(p)/R_{0}). From (6.9),

|Δgφεh|{Cε1rτ1,forr2R0,4Λε1(R02|0,Xφε|0+R01|0,X2φε|0)+Cε1R0τ1,forR0r2R0,0,forrR0.\displaystyle\begin{split}|\Delta_{g_{\varphi_{\varepsilon}}}h|\leq\begin{cases}C\varepsilon^{-1}r^{-\tau-1},&\text{for}\;\,r\geq 2R_{0},\\ 4\Lambda\varepsilon^{-1}\big{(}R_{0}^{-2}|\nabla_{0,X}{\varphi_{\varepsilon}}|_{0}+R_{0}^{-1}|\nabla_{0,X}^{2}{\varphi_{\varepsilon}}|_{0}\big{)}+C\varepsilon^{-1}R_{0}^{-\tau-1},&\text{for}\;\,R_{0}\leq r\leq 2R_{0},\\ 0,&\text{for}\;\,r\leq R_{0}.\end{cases}\end{split} (6.11)

Then, we can pick a barrier function as follows:

u1=E{(1χR02)(R02)τ1t(t1)+χR02t(t1)rτ1}0,\displaystyle u_{1}=E\Big{\{}\big{(}1-\chi_{\frac{R_{0}}{2}}\big{)}\Big{(}\frac{R_{0}}{2}\Big{)}^{-\tau-1}t(t-1)+\chi_{\frac{R_{0}}{2}}t(t-1)r^{-\tau-1}\Big{\}}\leq 0, (6.12)

where the constant EE is to be determined later. The barrier function u1u_{1} is defined in X×ΣX\times\Sigma with u1=0u_{1}=0 on X×ΣX\times\partial\Sigma. We also have

Δgφεu1\displaystyle\Delta_{g_{\varphi_{\varepsilon}}}u_{1} =12trgφε(ddcu1(,J))=12{1α,β2ngφεαβ(u1,αβ+u1,Jα,Jβ)\displaystyle=\frac{1}{2}\operatorname{tr}_{g_{\varphi_{\varepsilon}}}(dd^{c}u_{1}(\cdot,J\cdot))=\frac{1}{2}\left\{\sum_{1\leq\alpha,\beta\leq 2n}g^{\alpha\beta}_{{\varphi_{\varepsilon}}}(u_{1,\alpha\beta}+u_{1,J\alpha,J\beta})\right.
+1α2n,2n+1𝝁2n+2gφε𝝁α(u1,α𝝁+u1,Jα,J𝝁)+2n+1𝝁,𝝂2n+2gφε𝝁𝝂(u1,𝝁𝝂+u1,J𝝂,J𝝂)}.\displaystyle\left.+\sum_{\begin{subarray}{c}1\leq\alpha\leq 2n,\\ 2n+1\leq\boldsymbol{\mu}\leq 2n+2\end{subarray}}g^{\boldsymbol{\mu}\alpha}_{{\varphi_{\varepsilon}}}(u_{1,\alpha\boldsymbol{\mu}}+u_{1,J\alpha,J{\boldsymbol{\mu}}})+\sum_{2n+1\leq\boldsymbol{\mu},\boldsymbol{\nu}\leq 2n+2}g_{\varphi_{\varepsilon}}^{\boldsymbol{\mu}\boldsymbol{\nu}}(u_{1,\boldsymbol{\mu}\boldsymbol{\nu}}+u_{1,J\boldsymbol{\nu},J\boldsymbol{\nu}})\right\}.

Using the estimate Λ1εδij(gφε)ijΛδij\Lambda^{-1}\varepsilon\delta_{ij}\leq(g_{{\varphi_{\varepsilon}}})_{ij}\leq\Lambda\delta_{ij}, we obtain that

Δgφεu1{E(Λ1rτ1Λε1rτ2Λε1rτ3)forrR0/2,EΛ1R0τ1forrR0/2.\displaystyle\Delta_{g_{\varphi_{\varepsilon}}}u_{1}\geq\begin{cases}E(\Lambda^{-1}r^{-\tau-1}-\Lambda\varepsilon^{-1}r^{-\tau-2}-\Lambda\varepsilon^{-1}r^{-\tau-3})&\text{for}\;\,r\geq R_{0}/2,\\ E\Lambda^{-1}R_{0}^{-\tau-1}&\text{for}\;\,r\leq R_{0}/2.\end{cases} (6.13)

By taking

R04Λ2ε1,E=8R0τCwithC=C(φε𝒞2(X×Σ,ΘΨ),Λ,g,J),\displaystyle R_{0}\geq 4\Lambda^{2}\varepsilon^{-1},\quad E=8R_{0}^{\tau}C\;\,\text{with}\;\,C=C(\|{\varphi_{\varepsilon}}\|_{\mathcal{C}^{2}(X\times\Sigma,\Theta_{\Psi})},\Lambda,g,J), (6.14)

and comparing with the inequality (6.11), we have Δgφεu1Δgφεh\Delta_{g_{\varphi_{\varepsilon}}}u_{1}\geq\Delta_{g_{\varphi_{\varepsilon}}}h. Together with the fact that u1=h=0u_{1}=h=0 on X×ΣX\times\partial\Sigma, Lemma 2.2 implies that hu1h\geq u_{1} in X×ΣX\times\Sigma. The same method shows the upper bound hu1h\leq-u_{1}, which, together with the lower bound, implies that for each spatial index 1α2n1\leq\alpha\leq 2n,

|Dαφε|C(φεC2(X×Σ,ΘΨ),Λ,g,J,ε1)t(1t)rτ1on{r2R0}×Σ.\displaystyle|D_{\alpha}{\varphi_{\varepsilon}}|\leq C\big{(}\|{\varphi_{\varepsilon}}\|_{C^{2}(X\times\Sigma,\Theta_{\Psi})},\Lambda,g,J,\varepsilon^{-1}\big{)}t(1-t)r^{-\tau-1}\;\,\text{on}\;\,\{r\geq 2R_{0}\}\times\Sigma. (6.15)

Step 3: Barrier estimate of the second derivatives.

Now, it comes to deal with the asymptotic behavior of the second derivative.

For a preliminary estimate, we go back to the full formula (6.8) for Δgφε(Dαφε)\Delta_{g_{\varphi_{\varepsilon}}}(D_{\alpha}\varphi_{\varepsilon}). For every 𝐚(0,1)\mathbf{a}\in(0,1), the Euclidean 𝒞0,𝐚\mathcal{C}^{0,\mathbf{a}} norm of the right-hand side on a restricted unit ball B^1(p)=B1(p)((2nBR)×Σ)\hat{B}_{1}(p)=B_{1}(p)\cap((\mathbb{R}^{2n}\setminus B_{R})\times\Sigma) with r(p)=r2Rr(p)=r\geq 2R is still bounded by C(φεC2(X×Σ,ΘΨ),Λ,g,J,ε1)rτ1C(\|{\varphi_{\varepsilon}}\|_{C^{2}(X\times\Sigma,\Theta_{\Psi})},\Lambda,g,J,\varepsilon^{-1})r^{-\tau-1} thanks to the Evans-Krylov estimates applied to φε\varphi_{\varepsilon} in the interior and the estimates of [8, Sections 2.1–2.2] at the boundary. (The precise dependence of this constant on the ellipticity, and hence on ε1\varepsilon^{-1}, is not clear but also not needed.) Likewise, the 𝒞0,𝐚\mathcal{C}^{0,\mathbf{a}} norm of the coefficient tensor of the PDE, gφε1g_{\varphi_{\varepsilon}}^{-1}, is bounded by C(φεC2(X×Σ,ΘΨ),Λ,g,J,ε1)C(\|{\varphi_{\varepsilon}}\|_{C^{2}(X\times\Sigma,\Theta_{\Psi})},\Lambda,g,J,\varepsilon^{-1}). Applying the classic interior and boundary Schauder estimates to (6.8), we thus obtain from (6.15) that

Dαφε𝒞2,𝐚(B^1(p))C(φεC2(X×Σ,ΘΨ),Λ,g,J,ε1)rτ1.\displaystyle\|D_{\alpha}{\varphi_{\varepsilon}}\|_{\mathcal{C}^{2,\mathbf{a}}(\hat{B}_{1}(p))}\leq C\big{(}\|\varphi_{\varepsilon}\|_{C^{2}(X\times\Sigma,\Theta_{\Psi})},\Lambda,g,J,\varepsilon^{-1}\big{)}r^{-\tau-1}. (6.16)

These estimates will now be used to start a bootstrap to obtain some decay for DβDαφεD_{\beta}D_{\alpha}\varphi_{\varepsilon} using the same barrier method as in Step 2. Differentiate the equation (6.8) again by Dβ=/xβD_{\beta}=\partial/\partial x_{\beta} for 1β2n1\leq\beta\leq 2n. This yields

Δgφε(DβDαφε)=gφε1Dβgφεgφε1020,Xφε+𝐎(rτ2)gφε100,Xφε+𝐎(rτ1)gφε1Dβgφεgφε100,Xφε+𝐎(rτ1)gφε100,X2φε+𝐎(rτ3)gφε10,Xφε+𝐎(rτ2)gφε1Dβgφεgφε10,Xφε+𝐎(rτ2)gφε10,X2φε+DβDαgij(gφεijgij),DβDαgij=𝐎^(rτ2),+𝐎^(rτ1)(gφε1Dβgφεgφε1g1Dβgg1)+𝐎^^(rς2)\displaystyle\begin{split}&\Delta_{g_{\varphi_{\varepsilon}}}(D_{\beta}D_{\alpha}{\varphi_{\varepsilon}})=g_{\varphi_{\varepsilon}}^{-1}\circledast D_{\beta}g_{\varphi_{\varepsilon}}\circledast g_{\varphi_{\varepsilon}}^{-1}\circledast\nabla_{0}^{2}\nabla_{0,X}\varphi_{\varepsilon}\\ &+\mathbf{O}(r^{-\tau-2})\circledast g_{\varphi_{\varepsilon}}^{-1}\circledast\nabla_{0}\nabla_{0,X}\varphi_{\varepsilon}\\ &+\mathbf{O}(r^{-\tau-1})\circledast g_{\varphi_{\varepsilon}}^{-1}\circledast D_{\beta}g_{\varphi_{\varepsilon}}\circledast g_{\varphi_{\varepsilon}}^{-1}\circledast\nabla_{0}\nabla_{0,X}\varphi_{\varepsilon}\\ &+\mathbf{O}(r^{-\tau-1})\circledast g_{\varphi_{\varepsilon}}^{-1}\circledast\nabla_{0}\nabla_{0,X}^{2}\varphi_{\varepsilon}\\ &+\mathbf{O}(r^{-\tau-3})\circledast g_{\varphi_{\varepsilon}}^{-1}\circledast\nabla_{0,X}\varphi_{\varepsilon}\\ &+\mathbf{O}(r^{-\tau-2})\circledast g_{\varphi_{\varepsilon}}^{-1}\circledast D_{\beta}g_{\varphi_{\varepsilon}}\circledast g_{\varphi_{\varepsilon}}^{-1}\circledast\nabla_{0,X}\varphi_{\varepsilon}\\ &+\mathbf{O}(r^{-\tau-2})\circledast g_{\varphi_{\varepsilon}}^{-1}\circledast\nabla_{0,X}^{2}\varphi_{\varepsilon}\\ &+D_{\beta}D_{\alpha}g_{ij}\cdot(g_{\varphi_{\varepsilon}}^{ij}-g^{ij}),\quad D_{\beta}D_{\alpha}g_{ij}=\widehat{\mathbf{O}}(r^{-\tau-2}),\\ &+\widehat{\mathbf{O}}(r^{-\tau-1})\circledast(g_{\varphi_{\varepsilon}}^{-1}\circledast D_{\beta}g_{\varphi_{\varepsilon}}\circledast g_{\varphi_{\varepsilon}}^{-1}-g^{-1}\circledast D_{\beta}g\circledast g^{-1})\\ &+\widehat{\widehat{\mathbf{O}}}(r^{-\varsigma-2})\end{split} (6.17)

As before, we have that Λ1g1gφε1ε1Λg1\Lambda^{-1}g^{-1}\leq g_{\varphi_{\varepsilon}}^{-1}\leq\varepsilon^{-1}\Lambda g^{-1}, and we also have

|Dβgφε|0|Dβg|0+|0,X02φε|0=O(rτ1)|D_{\beta}g_{\varphi_{\varepsilon}}|_{0}\leq|D_{\beta}g|_{0}+|\nabla_{0,X}\nabla_{0}^{2}\varphi_{\varepsilon}|_{0}=O(r^{-\tau-1}) (6.18)

thanks to the preliminary estimate (6.16). Similarly, all derivatives of φε\varphi_{\varepsilon} on the right-hand side of (6.17) are at worst of order 33, with at least one purely spatial derivative, and hence can be bounded by O(rτ1)O(r^{-\tau-1}) thanks to (6.16). In this way, we obtain that

Δgφε(DβDαφε)=𝐎^(rτ2)(gφε1g1)+O(r2τ2).\displaystyle\Delta_{g_{\varphi_{\varepsilon}}}(D_{\beta}D_{\alpha}{\varphi_{\varepsilon}})=\widehat{\mathbf{O}}(r^{-\tau-2})\circledast(g_{\varphi_{\varepsilon}}^{-1}-g^{-1})+O(r^{-2\tau-2}). (6.19)

The majority of terms on the right-hand side actually decay faster than O(r2τ2)O(r^{-2\tau-2}), and the only term that might decay more slowly is 𝐎^(rτ2)(gφε1g1)\widehat{\mathbf{O}}(r^{-\tau-2})\circledast(g_{\varphi_{\varepsilon}}^{-1}-g^{-1}). So far, we can only bound this by O(rτ2)O(r^{-\tau-2}). However, by applying the same method as in the weighted estimate of the first derivative in Step 2, we can then construct the following barrier function for DβDαφεD_{\beta}D_{\alpha}{\varphi_{\varepsilon}}:

u2=E{(1χR02)(R02)τ2t(t1)+χR02t(t1)rτ2},\displaystyle u_{2}=E^{\prime}\Big{\{}\big{(}1-\chi_{\frac{R_{0}}{2}}\big{)}\Big{(}\frac{R_{0}}{2}\Big{)}^{-\tau-2}t(t-1)+\chi_{\frac{R_{0}}{2}}t(t-1)r^{-\tau-2}\Big{\}}, (6.20)

where R0R_{0} is the same constant as in (6.12) and EE^{\prime} is another uniform constant depending on R0R_{0}, φεC2(X×Σ,ΘΨ)\|{\varphi_{\varepsilon}}\|_{C^{2}(X\times\Sigma,\Theta_{\Psi})}, Λ\Lambda, gg, JJ and on the constant of (6.16). Hence, we get the weighted estimate for DβDαφεD_{\beta}D_{\alpha}{\varphi_{\varepsilon}}:

|DβDαφε|C(φεC2(X×Σ,ΘΨ),Λ,g,J,ε1)rτ2.\displaystyle|D_{\beta}D_{\alpha}{\varphi_{\varepsilon}}|\leq C\big{(}\|\varphi_{\varepsilon}\|_{C^{2}(X\times\Sigma,\Theta_{\Psi})},\Lambda,g,J,\varepsilon^{-1}\big{)}r^{-\tau-2}. (6.21)

According to the full formula (6.17) for Δgφε(DβDαφε)\Delta_{g_{\varphi_{\varepsilon}}}(D_{\beta}D_{\alpha}\varphi_{\varepsilon}) and (6.16), in the restricted unit ball B^1(p)\hat{B}_{1}(p), the 𝒞0,𝐚\mathcal{C}^{0,\mathbf{a}} norm of all terms on the right hand side of (6.17) are bounded by C(φεC2,Λ,g,J,ε1)rτ2C(\|{\varphi_{\varepsilon}}\|_{C^{2}},\Lambda,g,J,\varepsilon^{-1})r^{-\tau-2}. Applying the classic interior and boundary Schauder estimates to (6.17), we thus obtain from (6.15) that

DαDβφε𝒞2,𝐚(B^1(p))C(φεC2(X×Σ,ΘΨ),Λ,g,J,ε1)rτ2.\displaystyle\|D_{\alpha}D_{\beta}{\varphi_{\varepsilon}}\|_{\mathcal{C}^{2,\mathbf{a}}(\hat{B}_{1}(p))}\leq C\big{(}\|\varphi_{\varepsilon}\|_{C^{2}(X\times\Sigma,\Theta_{\Psi})},\Lambda,g,J,\varepsilon^{-1}\big{)}r^{-\tau-2}. (6.22)

Step 4: Iterative improvement of the barrier estimates.

In this step, we improve the decay order of the estimates we obtain in (6.16) and (6.22) by an iteration argument. Recall that from Steps 2–3 we have the following weighted estimates to start the iteration process (see (6.16) and (6.22)):

0,Xφε𝒞0,𝐚(B^1(p))+00,Xφε𝒞0,𝐚(B^1(p))+020,Xφε𝒞0,𝐚(B^1(p))=O(rτ1),0,X2φε𝒞0,𝐚(B^1(p))+00,X2φε𝒞0,𝐚(B^1(p))=O(rτ2).\displaystyle\begin{split}\|\nabla_{0,X}\varphi_{\varepsilon}\|_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))}+\|\nabla_{0}\nabla_{0,X}\varphi_{\varepsilon}\|_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))}&\\ +\;\|\nabla_{0}^{2}\nabla_{0,X}\varphi_{\varepsilon}\|_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))}&=O(r^{-\tau-1}),\\ \|\nabla^{2}_{0,X}\varphi_{\varepsilon}\|_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))}+\|\nabla_{0}\nabla_{0,X}^{2}\varphi_{\varepsilon}\|_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))}&=O(r^{-\tau-2}).\end{split} (6.23)

To complete the iteration argument, we need to improve the decay of the term gφε1g1g_{\varphi_{\varepsilon}}^{-1}-g^{-1}. More precisely, this term occurs in a combination (gφεijgij)Dαgij(g_{\varphi_{\varepsilon}}^{ij}-g^{ij})D_{\alpha}g_{ij} in the first derivative estimate (Step 2), and in combinations (gφεijgij)DβDαgij(g_{\varphi_{\varepsilon}}^{ij}-g^{ij})D_{\beta}D_{\alpha}g_{ij} and [gφεikDβ(gφε)klgφεjlgikDβgklgjl]Dαgij[g_{\varphi_{\varepsilon}}^{ik}D_{\beta}(g_{\varphi_{\varepsilon}})_{kl}g_{\varphi_{\varepsilon}}^{jl}-g^{ik}D_{\beta}g_{kl}g^{jl}]D_{\alpha}g_{ij} (to get optimal decay rate of DαDβφεD_{\alpha}D_{\beta}\varphi_{\varepsilon}, we need to analyze this term) in the second derivative estimate (Step 3). We will now analyze these combinations more carefully. All constants in this step may depend on φεC2(X×Σ,ΘΨ),Λ,g,J,ε1\|\varphi_{\varepsilon}\|_{C^{2}(X\times\Sigma,\Theta_{\Psi})},\Lambda,g,J,\varepsilon^{-1}. Let φ\varphi be a continuous function defined in (X\BR)×Σ(X\backslash B_{R})\times\Sigma with at most polynomial growth rate at infinity, for simplicity, we introduce the notation (φ)(\varphi)^{\sharp} to denote the decay rate of φ\varphi and (DXφ)(D_{X}\varphi)^{\sharp}, (DX2φ)(D_{X}^{2}\varphi)^{\sharp} to denote the decay rate of 0,Xφ𝒞0,𝐚(B^1(p))+00,Xφ𝒞0,𝐚(B^1(p))+020,Xφ𝒞0,𝐚(B^1(p))\|\nabla_{0,X}\varphi\|_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))}+\|\nabla_{0}\nabla_{0,X}\varphi\|_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))}+\|\nabla_{0}^{2}\nabla_{0,X}\varphi\|_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))}, 0,X2φ𝒞0,𝐚(B^1(p))+00,X2φ𝒞0,𝐚(B^1(p))\|\nabla^{2}_{0,X}\varphi\|_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))}+\|\nabla_{0}\nabla_{0,X}^{2}\varphi\|_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))} respectively.

The metric tensor (gφε)ij(g_{\varphi_{\varepsilon}})_{ij} and its inverse can be written as (2n+2)×(2n+2)(2n+2)\times(2n+2)-matrices

P~=(Pηtη𝔭),(P~)1=(Qξtξ𝔮),\displaystyle\widetilde{P}=\begin{pmatrix}P&\eta^{t}\\ \eta&\mathfrak{p}\end{pmatrix},\quad(\widetilde{P})^{-1}=\begin{pmatrix}Q&\xi^{t}\\ \xi&\mathfrak{q}\end{pmatrix},

where PP, QQ are 2n×2n2n\times 2n-matrices, 𝔭\mathfrak{p}, 𝔮\mathfrak{q} are 2×22\times 2-matrices and η\eta, ξ\xi are 2×2n2\times 2n-matrices. By direct calculation, we have

Q=P1P1ηtξ,ξ=𝔭1ηQ,𝔮=(I2ξηt)𝔭1.\displaystyle Q=P^{-1}-P^{-1}\eta^{t}\xi,\quad\xi=-\mathfrak{p}^{-1}\eta Q,\quad\mathfrak{q}=(I_{2}-\xi\eta^{t})\mathfrak{p}^{-1}. (6.24)

The fact that Λ1εI2n+2P~ΛI2n+2\Lambda^{-1}\varepsilon I_{2n+2}\leq\widetilde{P}\leq\Lambda I_{2n+2} implies that |ξ|C|η||\xi|\leq C|\eta|. The weighted estimate (6.23), together with the fall-off condition of the metric gg, implies that |η|=O(r(DXφε))|\eta|=O(r^{(D_{X}\varphi_{\varepsilon})^{\sharp}}). Then, from (6.24), we have

Q=P1+O(|η|2),ξ=O(|η|),𝔮=𝔭1+O(|η|2).\displaystyle Q=P^{-1}+O(|\eta|^{2}),\quad\xi=O(|\eta|),\quad\mathfrak{q}=\mathfrak{p}^{-1}+O(|\eta|^{2}). (6.25)

Similarly, let P~\widetilde{P}^{\prime} denote the matrix of gg in asymptotic coordinates. If we write

P~=(P(η)tη𝔭),(P~)1=(Q(ξ)tξ𝔮),\displaystyle\widetilde{P}^{\prime}=\begin{pmatrix}P^{\prime}&(\eta^{\prime})^{t}\\ \eta^{\prime}&\mathfrak{p}^{\prime}\end{pmatrix},\quad(\widetilde{P}^{\prime})^{-1}=\begin{pmatrix}Q^{\prime}&(\xi^{\prime})^{t}\\ \xi^{\prime}&\mathfrak{q}^{\prime}\end{pmatrix},

then we have

Q=(P)1+O(|η|2),ξ=O(|η|),𝔮=(𝔭)1+O(|η|2),\displaystyle Q^{\prime}=(P^{\prime})^{-1}+O(|\eta^{\prime}|^{2}),\quad\xi^{\prime}=O(|\eta^{\prime}|),\quad\mathfrak{q}^{\prime}=(\mathfrak{p}^{\prime})^{-1}+O(|\eta^{\prime}|^{2}), (6.26)

where |η|=O(rγ1)|\eta^{\prime}|=O(r^{-\gamma-1}). According to the estimate (6.23), |PP|=O(r(DX2φε))|P-P^{\prime}|=O(r^{(D_{X}^{2}\varphi_{\varepsilon})^{\sharp}}) and hence |P1(P)1|=O(r(DX2φε))|P^{-1}-(P^{\prime})^{-1}|=O(r^{(D_{X}^{2}\varphi_{\varepsilon})^{\sharp}}) as well because P,PP,P^{\prime} are uniformly bounded. Moreover, 𝔭,𝔮,𝔭,𝔮\mathfrak{p},\mathfrak{q},\mathfrak{p}^{\prime},\mathfrak{q}^{\prime} are all uniformly equivalent to I2I_{2} but there is no reason for 𝔭𝔭\mathfrak{p}-\mathfrak{p}^{\prime} to decay. Then (6.25) and (6.26) imply that

|QQ|=O(rmax{(DX2φε),2(DXφε),2γ2}),|ξξ|=O(rmax{(DXφε),γ1}),|𝔭𝔭|=O(1).\displaystyle\begin{split}|Q-Q^{\prime}|&=O(r^{\max\{(D_{X}^{2}\varphi_{\varepsilon})^{\sharp},2(D_{X}\varphi_{\varepsilon})^{\sharp},-2\gamma-2\}}),\\ |\xi-\xi^{\prime}|&=O(r^{\max\{(D_{X}\varphi_{\varepsilon})^{\sharp},-\gamma-1\}}),\\ |\mathfrak{p}-\mathfrak{p}^{\prime}|&=O(1).\end{split} (6.27)

Then, by calculating blockwise and using that Dαg𝝁𝝂=0D_{\alpha}g_{\boldsymbol{\mu}\boldsymbol{\nu}}=0, we have

(gφεijgij)Dαgij=|QQ|O(rτ1)+|ξξ|O(rγ2),(gφεijgij)DβDαgij=|QQ|O(rτ2)+|ξξ|O(rγ3).\displaystyle\begin{split}(g_{\varphi_{\varepsilon}}^{ij}-g^{ij})D_{\alpha}g_{ij}&=|Q-Q^{\prime}|O(r^{-\tau-1})+|\xi-\xi^{\prime}|O(r^{-\gamma-2}),\\ (g_{\varphi_{\varepsilon}}^{ij}-g^{ij})D_{\beta}D_{\alpha}g_{ij}&=|Q-Q^{\prime}|O(r^{-\tau-2})+|\xi-\xi^{\prime}|O(r^{-\gamma-3}).\end{split} (6.28)

By inserting (6.27), (6.28) into (6.8), we have

|Dαφε|=O(rmax{(DX2φε)τ1,(DXφε)τ1,2γ3,ς1}).\displaystyle\begin{split}|D_{\alpha}\varphi_{\varepsilon}|&=O(r^{\max\{(D_{X}^{2}\varphi_{\varepsilon})^{\sharp}-\tau-1,(D_{X}\varphi_{\varepsilon})^{\sharp}-\tau-1,-2\gamma-3,-\varsigma-1\}}).\end{split} (6.29)

For the last but one term of (6.17),

Dαgij(gφεikgφεjlDβ(gφε)klgikgjlDβgkl)=Dαgijgφεikgφεjl(Dβ(gφε)klDβgkl)+DαgijDβgklgφεjl(gφεikgik)+DαgijDβgklgik(gφεjlgjl).\displaystyle\begin{split}D_{\alpha}g_{ij}\big{(}g_{\varphi_{\varepsilon}}^{ik}g^{jl}_{\varphi_{\varepsilon}}D_{\beta}(g_{\varphi_{\varepsilon}})_{kl}-g^{ik}g^{jl}D_{\beta}g_{kl}\big{)}&=D_{\alpha}g_{ij}g^{ik}_{\varphi_{\varepsilon}}g^{jl}_{\varphi_{\varepsilon}}(D_{\beta}(g_{\varphi_{\varepsilon}})_{kl}-D_{\beta}g_{kl})\\ &+D_{\alpha}g_{ij}D_{\beta}g_{kl}g_{\varphi_{\varepsilon}}^{jl}(g^{ik}_{\varphi_{\varepsilon}}-g^{ik})\\ &+D_{\alpha}g_{ij}D_{\beta}g_{kl}g^{ik}(g^{jl}_{\varphi_{\varepsilon}}-g^{jl}).\end{split} (6.30)

For the first term of right-hand side of (6.30), by using |Dβ(gφε)klDβgkl||020,Xφε||D_{\beta}(g_{\varphi_{\varepsilon}})_{kl}-D_{\beta}g_{kl}|\leq|\nabla_{0}^{2}\nabla_{0,X}\varphi_{\varepsilon}|, we obtain that the decay rate of the first term is given by (DXφε)τ1(D_{X}\varphi_{\varepsilon})^{\sharp}-\tau-1. For the second and third terms, we need to analyze (gφεikgik)(g^{ik}_{\varphi_{\varepsilon}}-g^{ik}). Similar to (6.28), we have

DαgijDβgklgik(gφεjlgjl)=|QQ|O(r2τ2)+|ξξ|O(rτγ3)+O(r2γ4).\displaystyle D_{\alpha}g_{ij}D_{\beta}g_{kl}g^{ik}(g^{jl}_{\varphi_{\varepsilon}}-g^{jl})=|Q-Q^{\prime}|O(r^{-2\tau-2})+|\xi-\xi^{\prime}|O(r^{-\tau-\gamma-3})+O(r^{-2\gamma-4}). (6.31)

By inserting (6.27) into (6.31), we have

Dαgij(gφεikgφεjlDβ(gφε)klgikgjlDβgkl)=O(rmax{(DXφε)τ1,(DX2φε)2τ2,2γ4}).\displaystyle\begin{split}D_{\alpha}g_{ij}\big{(}g_{\varphi_{\varepsilon}}^{ik}g^{jl}_{\varphi_{\varepsilon}}&D_{\beta}(g_{\varphi_{\varepsilon}})_{kl}-g^{ik}g^{jl}D_{\beta}g_{kl}\big{)}=O(r^{\max\{(D_{X}\varphi_{\varepsilon})^{\sharp}-\tau-1,(D_{X}^{2}\varphi_{\varepsilon})^{\sharp}-2\tau-2,-2\gamma-4\}}).\end{split} (6.32)

Then, inserting (6.27), (6.28) into (6.17), we have

|DαDβφε|\displaystyle|D_{\alpha}D_{\beta}\varphi_{\varepsilon}| =O(rmax{(DX2φε)τ1,(DXφε)τ1,2γ4,ς2}).\displaystyle=O(r^{\max\{(D_{X}^{2}\varphi_{\varepsilon})^{\sharp}-\tau-1,(D_{X}\varphi_{\varepsilon})^{\sharp}-\tau-1,-2\gamma-4,-\varsigma-2\}}).

We can go one step further by applying Schauder estimates to (6.8) and (6.17) and to obtain 𝒞2,𝐚\mathcal{C}^{2,\mathbf{a}} estimates for DαφεD_{\alpha}\varphi_{\varepsilon} and DαDβφεD_{\alpha}D_{\beta}\varphi_{\varepsilon} in B^1(p)\hat{B}_{1}(p). Indeed, those terms on the right-hand side of the PDEs (6.8), (6.17) that were known to decay pointwise with rate max{(DX2φε),(DXφε)}τ1\max\{(D_{X}^{2}\varphi_{\varepsilon})^{\sharp},(D_{X}\varphi_{\varepsilon})^{\sharp}\}-\tau-1 already after Step 3 are actually also decaying at rate max{(DX2φε),(DXφε)}τ1\max\{(D_{X}^{2}\varphi_{\varepsilon})^{\sharp},(D_{X}\varphi_{\varepsilon})^{\sharp}\}-\tau-1 in 𝒞0,𝐚(B^1(p))\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p)) norm. This is clear from (6.23). So we just need to find the decay rates of the most difficult terms, (gφεijgij)Dαgij(g_{\varphi_{\varepsilon}}^{ij}-g^{ij})D_{\alpha}g_{ij} in (6.8) and (gφεijgij)DβDαgij(g_{\varphi_{\varepsilon}}^{ij}-g^{ij})D_{\beta}D_{\alpha}g_{ij}, Dαgij(gφεikgφεjlDβ(gφε)klgikgjlDβgkl)D_{\alpha}g_{ij}(g_{\varphi_{\varepsilon}}^{ik}g^{jl}_{\varphi_{\varepsilon}}D_{\beta}(g_{\varphi_{\varepsilon}})_{kl}-g^{ik}g^{jl}D_{\beta}g_{kl}) in (6.17) in 𝒞0,𝐚(B^1(p))\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p)) norm as well. For this, we need to go back and also estimate the 𝒞0,𝐚\mathcal{C}^{0,\mathbf{a}}-norm of QQQ-Q^{\prime} and ξξ\xi-\xi^{\prime} in B^1(p)\hat{B}_{1}(p), as follows. By using (6.23), we have that

[P1(P)1]𝒞0,𝐚(B^1(p))=O(r(DX2φε)),[ξ]𝒞0,𝐚(B^1(p))=O(r(DXφε)).\displaystyle[P^{-1}-(P^{\prime})^{-1}]_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))}=O(r^{(D_{X}^{2}\varphi_{\varepsilon})^{\sharp}}),\quad[\xi]_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))}=O(r^{(D_{X}\varphi_{\varepsilon})^{\sharp}}).

Then, based on (6.25), we have that

[QQ]𝒞0,𝐚(B^1(p))=O(r(DX2φε),2(DXφε),2γ2),[ξξ]𝒞0,𝐚(B^1(p))=O(rmax{(DXφε),γ1}).\displaystyle\begin{split}[Q-Q^{\prime}]_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))}&=O(r^{(D_{X}^{2}\varphi_{\varepsilon})^{\sharp},2(D_{X}\varphi_{\varepsilon})^{\sharp},-2\gamma-2}),\\ [\xi-\xi^{\prime}]_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))}&=O(r^{\max\{(D_{X}\varphi_{\varepsilon})^{\sharp},-\gamma-1\}}).\end{split} (6.33)

Then we can proceed as in (6.28) and (6.32), obtaining that the decay rates of [ΔφεDαφε]𝒞0,𝐚(B^1(p))[\Delta_{\varphi_{\varepsilon}}D_{\alpha}\varphi_{\varepsilon}]_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}_{1}(p))} and [ΔφεDαDβφε]𝒞0,𝐚(B^(p))[\Delta_{\varphi_{\varepsilon}}D_{\alpha}D_{\beta}\varphi_{\varepsilon}]_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}(p))} are max{(DX2φε)τ1,(DXφε)τ1,2γ3,ς2}\max\{(D_{X}^{2}\varphi_{\varepsilon})^{\sharp}-\tau-1,(D_{X}\varphi_{\varepsilon})^{\sharp}-\tau-1,-2\gamma-3,-\varsigma-2\} and max{(DX2φε)τ1,(DXφε)τ1,2γ4,ς2}\max\{(D_{X}^{2}\varphi_{\varepsilon})^{\sharp}-\tau-1,(D_{X}\varphi_{\varepsilon})^{\sharp}-\tau-1,-2\gamma-4,-\varsigma-2\} respectively. According to the classic interior and boundary Schauder estimates, we improve (6.29) to 𝒞2,𝐚(B^1(p))\mathcal{C}^{2,\mathbf{a}}(\hat{B}_{1}(p)) norm,

Dαφε𝒞2,𝐚(B^1(p))=O(rmax{(DX2φε)τ1,(DXφε)τ1,2γ3,ς1}),DαDβφε𝒞2,𝐚(B^1(p))=O(rmax{(DX2φε)τ1,(DXφε)τ1,2γ4,ς2}).\displaystyle\begin{split}\|D_{\alpha}\varphi_{\varepsilon}\|_{\mathcal{C}^{2,\mathbf{a}}(\hat{B}_{1}(p))}&=O(r^{\max\{(D_{X}^{2}\varphi_{\varepsilon})^{\sharp}-\tau-1,(D_{X}\varphi_{\varepsilon})^{\sharp}-\tau-1,-2\gamma-3,-\varsigma-1\}}),\\ \|D_{\alpha}D_{\beta}\varphi_{\varepsilon}\|_{\mathcal{C}^{2,\mathbf{a}}(\hat{B}_{1}(p))}&=O(r^{\max\{(D_{X}^{2}\varphi_{\varepsilon})^{\sharp}-\tau-1,(D_{X}\varphi_{\varepsilon})^{\sharp}-\tau-1,-2\gamma-4,-\varsigma-2\}}).\end{split} (6.34)

Inserting (6.23) into (6.34), and using (6.34) again to improve (6.23), we finally obtain the following estimates:

Dαφε𝒞2,𝐚(B^1(p))=O(rmax{2γ3,ς1}),DαDβφε𝒞2,𝐚(B^1(p))=O(rmax{2γ4,ς2}).\displaystyle\|D_{\alpha}\varphi_{\varepsilon}\|_{\mathcal{C}^{2,\mathbf{a}}(\hat{B}_{1}(p))}=O(r^{\max\{-2\gamma-3,-\varsigma-1\}}),\quad\|D_{\alpha}D_{\beta}\varphi_{\varepsilon}\|_{\mathcal{C}^{2,\mathbf{a}}(\hat{B}_{1}(p))}=O(r^{\max\{-2\gamma-4,-\varsigma-2\}}). (6.35)

Note that according to (6.35), because Ψ\Psi was chosen to be linear in tt, the decay rate of φε\varphi_{\varepsilon} is faster than the decay rate of the boundary data ψ0,ψ1\psi_{0},\psi_{1}.

Step 5: Proof of Theorem B

In Step 4, we have obtained the optimal decay rates in the cases of k=1,2k=1,2 (even though it is not required in the proof of Theorem B). In this step, we give optimal estimates for k3k\geq 3 and complete the proof of Theorem B.

For the higher order derivatives, by differentiating the Monge-Ampère equation (6.2) mm times, similar to (6.8) and (6.17) and writing DK=Dκ1DκmD_{K}=D_{\kappa_{1}}\cdots D_{\kappa_{m}} (1κi2n1\leq\kappa_{i}\leq 2n, for i=1,,mi=1,\ldots,m), instead of giving a full formula as (6.8) and (6.17), we write a simplified formula of ΔφεDKφε\Delta_{\varphi_{\varepsilon}}D_{K}\varphi_{\varepsilon}:

|ΔφεDKφε|i=1mO(rτ2m+i)|0,Xiφε|+i=1mO(rτ1m+i)|00,Xiφε|+i=1m1O(rτm+i)|020,Xiφε|+i=1m|0,Xigjl||0,Xmi(gφεjlgjl)|+O(rςm).\displaystyle\begin{split}|\Delta_{\varphi_{\varepsilon}}D_{K}\varphi_{\varepsilon}|&\leq\sum_{i=1}^{m}O(r^{-\tau-2-m+i})|\nabla_{0,X}^{i}\varphi_{\varepsilon}|+\sum_{i=1}^{m}O(r^{-\tau-1-m+i})|\nabla_{0}\nabla_{0,X}^{i}\varphi_{\varepsilon}|\\ &+\sum_{i=1}^{m-1}O(r^{-\tau-m+i})|\nabla_{0}^{2}\nabla_{0,X}^{i}\varphi_{\varepsilon}|+\sum_{i=1}^{m}|\nabla_{0,X}^{i}g_{jl}||\nabla_{0,X}^{m-i}(g^{jl}_{\varphi_{\varepsilon}}-g^{jl})|\\ &+O(r^{-\varsigma-m}).\end{split} (6.36)

Applying induction on mm, according to iteration process (step 4), we can assume for km1k\leq m-1

0,XkφεB^1(p)=O(rmax{2γ2,ς}k).\displaystyle||\nabla^{k}_{0,X}\varphi_{\varepsilon}||_{\hat{B}_{1}(p)}=O(r^{\max\{-2\gamma-2,-\varsigma\}-k}). (6.37)

To find the optimal decay rates, the most difficult term is i=1m|0,Xigjl||0,Xmi(gφεjlgjl)|\sum_{i=1}^{m}|\nabla_{0,X}^{i}g_{jl}||\nabla_{0,X}^{m-i}(g^{jl}_{\varphi_{\varepsilon}}-g^{jl})|. Notice that by (6.31) and (6.34), we have

|DK1gjlDK2gik(gφεijgij)|=O(r2γ2k1k2),\displaystyle\big{|}D_{K_{1}}{g_{jl}}D_{K_{2}}g_{ik}(g^{ij}_{\varphi_{\varepsilon}}-g^{ij})\big{|}=O(r^{-2\gamma-2-k_{1}-k_{2}}),

where K1K_{1}, K2K_{2} are k1k_{1}-, k2k_{2}-multi-indices respectively. Then, we apply induction on kk to find the decay rate of |DK1gjlDK2gikDK(gφεijgij)||D_{K_{1}}{g_{jl}}D_{K_{2}}g_{ik}D_{K}(g^{ij}_{\varphi_{\varepsilon}}-g^{ij})|, where KK is a kk-multi-index. Applying one derivative to (gφεijgij)(g^{ij}_{\varphi_{\varepsilon}}-g^{ij}), by (6.30), we can prove that

|DK1gjlDK2gikDK(gφεijgij)|=O(r2γ2k1k2k).\displaystyle|D_{K_{1}}{g_{jl}}D_{K_{2}}g_{ik}D_{K}(g^{ij}_{\varphi_{\varepsilon}}-g^{ij})|=O(r^{-2\gamma-2-k_{1}-k_{2}-k}). (6.38)

Then, by (6.30) and (6.38), we have

|0,Xigjl||0,Xmi(gjlφεgjl)|\displaystyle|\nabla_{0,X}^{i}g_{jl}||\nabla_{0,X}^{m-i}(g^{jl}_{\varphi_{\varepsilon}}-g^{jl})| |i0,Xgjl|{|0,Xmi1[gjkφεgslφε(0,X(gφε)ks0,Xgks)]|\displaystyle\leq|\nabla^{i}_{0,X}g_{jl}|\Big{\{}\big{|}\nabla_{0,X}^{m-i-1}\big{[}g^{jk}_{\varphi_{\varepsilon}}g^{sl}_{\varphi_{\varepsilon}}(\nabla_{0,X}(g_{\varphi_{\varepsilon}})_{ks}-\nabla_{0,X}g_{ks})\big{]}\big{|}
+2|mi1[gφεjk(glsφεgik)0,Xgks]|}\displaystyle+2\big{|}\nabla^{m-i-1}\big{[}g_{\varphi_{\varepsilon}}^{jk}(g^{ls}_{\varphi_{\varepsilon}}-g^{ik})\nabla_{0,X}g_{ks}\big{]}\big{|}\Big{\}}
=O(r2γ2m).\displaystyle=O(r^{-2\gamma-2-m}).

Combining with (6.37), we have that the right-hand side of (6.36) is O(r2τ+2k)O(r^{-2\tau+2-k}). Using the construction of barrier functions in Step 2–3, we obtain that |DKφε|Cr2τ+2m|D_{K}\varphi_{\varepsilon}|\leq Cr^{-2\tau+2-m}. To apply Schauder estimates to the mm-th derivative of Monge-Ampère equation, we also need to know the decay rate of [ΔφεDKφε]𝒞0,𝐚(B^(p))[\Delta_{\varphi_{\varepsilon}}D_{K}\varphi_{\varepsilon}]_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}(p))}:

[ΔφεDKφε]𝒞0,𝐚(B^(p))\displaystyle[\Delta_{\varphi_{\varepsilon}}D_{K}\varphi_{\varepsilon}]_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}(p))} i=1mO(rτ2m+i)0,Xiφε𝒞0,𝐚(B^(p))\displaystyle\leq\sum_{i=1}^{m}O(r^{-\tau-2-m+i})\|\nabla_{0,X}^{i}\varphi_{\varepsilon}\|_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}(p))}
+i=1mO(rτ1m+i)00,Xiφε𝒞0,𝐚(B^(p))\displaystyle+\sum_{i=1}^{m}O(r^{-\tau-1-m+i})\|\nabla_{0}\nabla_{0,X}^{i}\varphi_{\varepsilon}\|_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}(p))}
+i=1m1O(rτm+i)020,Xiφε𝒞0,𝐚(B^(p))\displaystyle+\sum_{i=1}^{m-1}O(r^{-\tau-m+i})\|\nabla_{0}^{2}\nabla_{0,X}^{i}\varphi_{\varepsilon}\|_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}(p))}
+i=1m|0,Xigjl||0,Xmi(gjlφεgjl)|𝒞0,𝐚(B^(p))+O(rςm)\displaystyle+\sum_{i=1}^{m}\big{\|}|\nabla_{0,X}^{i}g_{jl}||\nabla_{0,X}^{m-i}(g^{jl}_{\varphi_{\varepsilon}}-g^{jl})|\big{\|}_{\mathcal{C}^{0,\mathbf{a}}(\hat{B}(p))}+O(r^{-\varsigma-m})
=O(rmax{2γ2,ς}m)\displaystyle=O(r^{{\max\{-2\gamma-2,-\varsigma\}-m}})

Hence, we have DKφε𝒞2,𝐚(B^1(p))Crmax{2γ2,ς}m\|D_{K}\varphi_{\varepsilon}\|_{\mathcal{C}^{2,\mathbf{a}}(\hat{B}_{1}(p))}\leq Cr^{{{\max\{-2\gamma-2,-\varsigma\}-m}}}, for m1m\geq 1. To prove that φε\varphi_{\varepsilon} is in γ\mathcal{H}_{-\gamma}, by integrating (φε)r=O(rmax{2γ2,ς}1)({\varphi_{\varepsilon}})_{r}=O(r^{\max\{-2\gamma-2,-\varsigma\}-1}) in the radial direction from infinity to r=Rr=R, we obtain a function φ^ε\hat{\varphi}_{\varepsilon} defined in XBRX\setminus B_{R} with decay rate max{2γ2,ς}\max\{-2\gamma-2,-\varsigma\}. Then,

φεφ^ε=c(θ,t),\displaystyle\varphi_{\varepsilon}-\hat{\varphi}_{\varepsilon}=c(\theta,t), (6.39)

where c(θ,t)c(\theta,t) is a function in XBRX\setminus B_{R} independent of radius rr and θ\theta be viewed as a variable on the link. It suffices to prove that c(θ,t)c(\theta,t) is independent of θ\theta. By taking derivative of (6.39), we have |0,Xc(θ,t)|=O(rmax{2γ2,ς}1)|\nabla_{0,X}c(\theta,t)|=O(r^{\max\{-2\gamma-2,-\varsigma\}-1}). In the case that c(θ,t)c(\theta,t) is not constant with respect to θ\theta, |0,Xc(θ,t)|r1|\nabla_{0,X}c(\theta,t)|\sim r^{-1}, which contradicts to the fact that max{2γ2,ς}<1\max\{-2\gamma-2,-\varsigma\}<-1. Hence we proved that φε=c(t)+O(rmax{2γ2,ς})\varphi_{\varepsilon}=c(t)+O(r^{-\max\{-2\gamma-2,-\varsigma\}}). We conclude that, for Φε=φε+Ψ\Phi_{\varepsilon}=\varphi_{\varepsilon}+\Psi,

sup(2n\BR)×Σ(|k0,XΦε|+|k0,XΦ˙ε|+|k0,XΦ¨ε|)C(k,ε1)rγkfor allk1.\displaystyle\sup_{(\mathbb{R}^{2n}\backslash B_{R})\times\Sigma}\left(|\nabla^{k}_{0,X}\Phi_{\varepsilon}|+|\nabla^{k}_{0,X}\dot{\Phi}_{\varepsilon}|+|\nabla^{k}_{0,X}\ddot{\Phi}_{\varepsilon}|\right)\leq C(k,\varepsilon^{-1})r^{-\gamma-k}\quad\text{for all}\ k\geq 1.

In conclusion, we have proved Theorem B.

7. Convexity of the Mabuchi KK-energy

According to Theorem 1.1 (assuming τ=τ~\tau=\tilde{\tau}), we can restrict ourselves to the space

2τ+2={φ𝒞^2τ+2:ωφ=ω+ddcφ>0},τ>n1,\displaystyle\mathcal{H}_{-2\tau+2}=\{\varphi\in\hat{\mathcal{C}}^{\infty}_{-2\tau+2}:\omega_{\varphi}=\omega+dd^{c}\varphi>0\},\quad\tau>n-1,

and the function υ(ε)\upsilon(\varepsilon) is constructed by (1.9) and (1.10). In the previous section, we proved that for any two given boundary data ψ0,ψ12τ+2\psi_{0},\psi_{1}\in\mathcal{H}_{-2\tau+2}, there exists a solution of the ε\varepsilon-geodesic equation (Eε)(E_{\varepsilon}) in the same space 2τ+2\mathcal{H}_{-2\tau+2}.

The derivative of the Mabuchi KK-energy can be defined as follows: for ψTφ2τ+2\psi\in T_{\varphi}\mathcal{H}_{-2\tau+2},

δψ𝒦(φ)=XψR(ωφ)ωφnn!.\displaystyle\delta_{\psi}\mathcal{K}(\varphi)=-\int_{X}\psi R(\omega_{\varphi})\frac{\omega_{\varphi}^{n}}{n!}.

The integral converges because 22τ<2n-2-2\tau<-2n, equivalently, τ>n1\tau>n-1. In the following proposition, the second derivative of Mabuchi KK-energy will be calculated in M=XR={xX:r(x)R}M=X_{R}=\{x\in X:r(x)\leq R\} containing boundary terms, and it will be clear that these boundary terms go to zero as RR\to\infty. Precisely, we consider Mabuchi KK-energy restricted in MM,

δψ𝒦M(φ)=MψR(ωφ)ωφn.\displaystyle\delta_{\psi}\mathcal{K}_{M}(\varphi)=-\int_{M}\psi R(\omega_{\varphi})\omega_{\varphi}^{n}. (7.1)

The calculation of the second variation of 𝒦M\mathcal{K}_{M} is due to my advisor Bianca Santoro in one of her unpublished notes, several years before I started this project. The limiting case RR\to\infty was previously stated by Aleyasin [1] without details concerning the vanishing of boundary terms.

To simplify the notation, in the following proposition, we write Rφ=R(ωφ)R_{\varphi}=R(\omega_{\varphi}), Ricφ=Ric(ωφ)\textup{Ric}_{\varphi}=\textup{Ric}(\omega_{\varphi}), Δ=gik¯φik¯\Delta=g^{i\overline{k}}_{\varphi}\partial_{i}\partial_{\overline{k}}, ||=||ωφ|\cdot|=|\cdot|_{\omega_{\varphi}}, =ωφ\nabla=\nabla_{\omega_{\varphi}} and 𝒟f=ikfdzidzk\mathcal{D}f=\nabla_{i}\nabla_{k}fdz^{i}dz^{k}, where ikf=f,ik\nabla_{i}\nabla_{k}f=f_{,ik} is a covariant derivative of ff with respect to ωφ\omega_{\varphi}. Recall that 𝒟\mathcal{D} is called the Lichnerowicz operator, and 𝒟f=0\mathcal{D}f=0 if and only if grad1,0f{\rm grad}^{1,0}f is a holomorphic type (1,0)(1,0) vector field.

Proposition 7.1 (Santoro).

Along a path of potentials φ(t)2τ+2\varphi(t)\in\mathcal{H}_{-2\tau+2},

d2𝒦Mdt2=M[φ¨12|φ˙|2]Rφωφn+M|𝒟φ˙|2ωφnn(n1)2Mφ˙dcφ˙Ricφωφn2+niMφ˙gφkl¯(Ricφ)il¯φ˙kdziωφn1niMφ˙gφij¯φ˙,ikj¯dzkωφn1niMgφkl¯φ˙l¯φ˙,kidziωφn1.\displaystyle\begin{split}\frac{d^{2}\mathcal{K}_{M}}{dt^{2}}&=-\int_{M}[\ddot{\varphi}-\frac{1}{2}|\nabla\dot{\varphi}|^{2}]R_{\varphi}\omega_{\varphi}^{n}+\int_{M}|\mathcal{D}\dot{\varphi}|^{2}\omega_{\varphi}^{n}\\ &-\frac{n(n-1)}{2}\int_{\partial M}\dot{\varphi}d^{c}\dot{\varphi}\wedge\textup{Ric}_{\varphi}\wedge\omega_{\varphi}^{n-2}+ni\int_{\partial M}\dot{\varphi}g_{\varphi}^{k\overline{l}}(\textup{Ric}_{\varphi})_{i\overline{l}}{\dot{\varphi}}_{k}dz^{i}\wedge\omega_{\varphi}^{n-1}\\ &-ni\int_{\partial M}\dot{\varphi}g_{\varphi}^{i\overline{j}}{\dot{\varphi}}_{,ik\overline{j}}dz^{k}\wedge\omega_{\varphi}^{n-1}-ni\int_{\partial M}g_{\varphi}^{k\overline{l}}{\dot{\varphi}}_{\overline{l}}{\dot{\varphi}}_{,ki}dz^{i}\wedge\omega_{\varphi}^{n-1}.\end{split} (7.2)

Furthermore, by taking RR\to\infty in (7.2), we have

d2𝒦dt2=X[φ¨12|φ˙|2]Rφωφnn!+X|𝒟φ˙|2ωφnn!.\displaystyle\frac{d^{2}\mathcal{K}}{dt^{2}}=\int_{X}\big{[}\ddot{\varphi}-\frac{1}{2}|\nabla\dot{\varphi}|^{2}\big{]}R_{\varphi}\frac{\omega_{\varphi}^{n}}{n!}+\int_{X}|\mathcal{D}\dot{\varphi}|^{2}\frac{\omega_{\varphi}^{n}}{n!}. (7.3)
Proof.

By taking the second derivative of Mabuchi KK-energy in MM, we have

d2𝒦Mdt2=\displaystyle\frac{d^{2}\mathcal{K}_{M}}{dt^{2}}= ddt[MRφφ˙ωφn]\displaystyle\frac{d}{dt}\left[-\int_{M}R_{\varphi}\dot{\varphi}\,\omega_{\varphi}^{n}\right]
=\displaystyle= Mφ¨RφωφnnMφ˙ddt(Ricφ)ωφn1\displaystyle-\int_{M}\ddot{\varphi}\,R_{\varphi}\,\omega_{\varphi}^{n}-n\int_{M}\dot{\varphi}\frac{d}{dt}(\textup{Ric}_{\varphi})\wedge\omega_{\varphi}^{n-1} (7.4)
n(n1)Mφ˙Ricφωφn2(i¯φ˙).\displaystyle-n(n-1)\int_{M}\dot{\varphi}\textup{Ric}_{\varphi}\wedge\omega_{\varphi}^{n-2}\wedge(i\partial\overline{\partial}\dot{\varphi}).

The second term of (7.4) needs one integration by parts, and we get

nMφ˙ddt(Ricφ)ωφn1\displaystyle-n\int_{M}\dot{\varphi}\frac{d}{dt}(\textup{Ric}_{\varphi})\wedge\omega_{\varphi}^{n-1} =nMφ˙[i¯(ddt(logωφn))ωφn1]\displaystyle=-n\int_{M}\dot{\varphi}\left[-i\partial\overline{\partial}\left(\frac{d}{dt}(\log\omega_{\varphi}^{n})\right)\wedge\omega_{\varphi}^{n-1}\right]
=nMφ˙[i¯(nωφn1i¯φ˙ωφn)ωφn1]\displaystyle=n\int_{M}\dot{\varphi}\left[i\partial\overline{\partial}\left(\frac{n\omega_{\varphi}^{n-1}\wedge i\partial\overline{\partial}\dot{\varphi}}{\omega_{\varphi}^{n}}\right)\wedge\omega_{\varphi}^{n-1}\right]
=Mφ˙(Δ2φ˙)ωφn.\displaystyle=\int_{M}\dot{\varphi}(\Delta^{2}\dot{\varphi})\omega_{\varphi}^{n}.

Now, to the term Mφ˙Ricφωφn2i¯φ˙\int_{M}\dot{\varphi}\textup{Ric}_{\varphi}\wedge\omega_{\varphi}^{n-2}\wedge i\partial\overline{\partial}\dot{\varphi}. For simplicity, φ˙=u\dot{\varphi}=u,

Mφ˙Ricφωφn2i¯φ˙\displaystyle\int_{M}\dot{\varphi}\textup{Ric}_{\varphi}\wedge\omega_{\varphi}^{n-2}\wedge i\partial\overline{\partial}\dot{\varphi}
=iMu¯uRicφωφn2+12MudcuRicφωφn2\displaystyle=-i\int_{M}\partial u\wedge\overline{\partial}u\wedge\textup{Ric}_{\varphi}\wedge\omega_{\varphi}^{n-2}+\frac{1}{2}\int_{\partial M}ud^{c}u\wedge\textup{Ric}_{\varphi}\wedge\omega_{\varphi}^{n-2}
=iMu¯uRic̊φωn2φinMu¯uRφωn1φ\displaystyle=-i\int_{M}\partial u\wedge\overline{\partial}u\wedge\mathring{\textup{Ric}}_{\varphi}\wedge\omega^{n-2}_{\varphi}-\frac{i}{n}\int_{M}\partial u\wedge\overline{\partial}u\wedge R_{\varphi}\omega^{n-1}_{\varphi}
+12MudcuRicφωφn2\displaystyle+\frac{1}{2}\int_{\partial M}ud^{c}u\wedge\textup{Ric}_{\varphi}\wedge\omega_{\varphi}^{n-2}
=iMu¯uRic̊φωn2φ12n2M|u|2Rφωnφ\displaystyle=-i\int_{M}\partial u\wedge\overline{\partial}u\wedge\mathring{\textup{Ric}}_{\varphi}\wedge\omega^{n-2}_{\varphi}-\frac{1}{2n^{2}}\int_{M}|\nabla u|^{2}R_{\varphi}\omega^{n}_{\varphi}
+12MudcuRicφωφn2,\displaystyle+\frac{1}{2}\int_{\partial M}ud^{c}u\wedge\textup{Ric}_{\varphi}\wedge\omega_{\varphi}^{n-2},

where Ric̊\mathring{\textup{Ric}} is the traceless part of Ricci. If ψ\psi is any primitive (1,1)(1,1)-form, then

ψ=1(n2)!ψωn2,and hencen(n1)Ric̊ωφn2=n!(Ric̊).\displaystyle*\psi=\frac{-1}{(n-2)!}\psi\wedge\omega^{n-2},\;\,\text{and hence}\;\,n(n-1)\mathring{\textup{Ric}}\wedge\omega_{\varphi}^{n-2}=-n!({*\mathring{\textup{Ric}}}).

Hence,

n(n1)Miu¯uRic̊φωφn2\displaystyle n(n-1)\int_{M}i\partial u\wedge\overline{\partial}u\wedge\mathring{\textup{Ric}}_{\varphi}\wedge\omega_{\varphi}^{n-2}
=Mn!(Ric̊φ)(iu¯u)\displaystyle=-\int_{M}n!(*\mathring{\textup{Ric}}_{\varphi})\wedge(i\partial u\wedge\overline{\partial}u)
=MRic̊φ,iu¯uωφn\displaystyle=-\int_{M}\langle\mathring{\textup{Ric}}_{\varphi},i\partial u\wedge\overline{\partial}u\rangle\,\omega_{\varphi}^{n}
=MRicφ,iu¯uωφn+M1nRφωφ,iu¯uωφn.\displaystyle=-\int_{M}\langle\textup{Ric}_{\varphi},i\partial u\wedge\overline{\partial}u\rangle\,\omega_{\varphi}^{n}+\int_{M}\langle\tfrac{1}{n}R_{\varphi}\,\omega_{\varphi},i\partial u\wedge\overline{\partial}u\rangle\,\omega_{\varphi}^{n}.

Note that

M1nRφωφ,iu¯uωφn\displaystyle\int_{M}\langle\tfrac{1}{n}R_{\varphi}\,\omega_{\varphi},i\partial u\wedge\overline{\partial}u\rangle\,\omega_{\varphi}^{n} =M(n1)!Rφωφ,iu¯uωφnn!\displaystyle=\int_{M}\langle(n-1)!R_{\varphi}\,\omega_{\varphi},i\partial u\wedge\overline{\partial}u\rangle\,\frac{\omega_{\varphi}^{n}}{n!}
=M(iu¯u)[(n1)!Rφωφ]\displaystyle=\int_{M}(i\partial u\wedge\overline{\partial}u)\wedge*[(n-1)!R_{\varphi}\omega_{\varphi}]
=MRφ(iu¯u)ωφn1\displaystyle=\int_{M}R_{\varphi}(i\partial u\wedge\overline{\partial}u)\wedge\omega_{\varphi}^{n-1}
=12nM|u|2Rφωφn.\displaystyle=\frac{1}{2n}\int_{M}|\nabla u|^{2}R_{\varphi}\,\omega_{\varphi}^{n}.

Thus, we get that

d2𝒦Mdt2=M[φ¨12|φ˙|2]RφωφnMRicφ,iu¯uωφn+Mu(Δ2u)ωφnn(n1)2MudcuRicφωφn2.\displaystyle\begin{split}\frac{d^{2}\mathcal{K}_{M}}{dt^{2}}=&-\int_{M}[\ddot{\varphi}-\frac{1}{2}|\nabla\dot{\varphi}|^{2}]R_{\varphi}\,\omega_{\varphi}^{n}-\int_{M}\langle\textup{Ric}_{\varphi},i\partial u\wedge\overline{\partial}u\rangle\,\omega_{\varphi}^{n}\\ &+\int_{M}u(\Delta^{2}u)\,\omega_{\varphi}^{n}-\frac{n(n-1)}{2}\int_{\partial M}ud^{c}u\wedge\textup{Ric}_{\varphi}\wedge\omega_{\varphi}^{n-2}.\end{split} (7.5)
Lemma 7.2.

Let ff be a smooth function defined on MM. Then we have that

Δ2f=𝒟𝒟fgφik¯gφjl¯(Ricφ)il¯fjk¯gφik¯gφjl¯(k¯(Ricφ)il¯)fj.\displaystyle\Delta^{2}f=\mathcal{D}^{*}\mathcal{D}f-g_{\varphi}^{i\overline{k}}g_{\varphi}^{j\overline{l}}(\textup{Ric}_{\varphi})_{i\overline{l}}f_{j\overline{k}}-g_{\varphi}^{i\overline{k}}g_{\varphi}^{j\overline{l}}(\nabla_{\overline{k}}(\textup{Ric}_{\varphi})_{i\overline{l}})f_{j}.

Hence,

Mu(Δ2u)ωφnMRicφ,iu¯uωφn=M|𝒟u|2ωφn+niMugφkl¯(Ricφ)il¯ukdziωφn1niMugφij¯j¯kiudzkωφn1niMgφkl¯ul¯u,kidziωφn1.\displaystyle\begin{split}\int_{M}u(\Delta^{2}u)\omega_{\varphi}^{n}&-\int_{M}\langle\textup{Ric}_{\varphi},i\partial u\wedge\overline{\partial}u\rangle\omega_{\varphi}^{n}\\ &=\int_{M}|\mathcal{D}u|^{2}\omega_{\varphi}^{n}+ni\int_{\partial M}ug_{\varphi}^{k\overline{l}}(\textup{Ric}_{\varphi})_{i\overline{l}}u_{k}dz^{i}\wedge\omega_{\varphi}^{n-1}\\ &-ni\int_{\partial M}ug_{\varphi}^{i\overline{j}}\nabla_{\overline{j}}\nabla_{k}\nabla_{i}udz^{k}\wedge\omega_{\varphi}^{n-1}-ni\int_{\partial M}g_{\varphi}^{k\overline{l}}u_{\overline{l}}u_{,ki}dz^{i}\wedge\omega_{\varphi}^{n-1}.\end{split} (7.6)
Proof.

Notice that

jk¯if=k¯jif\tensorRimjk¯f\displaystyle\nabla_{j}\nabla_{\overline{k}}\nabla_{i}f=\nabla_{\overline{k}}\nabla_{j}\nabla_{i}f-\tensor{R}{{}_{i}^{m}{}_{j}{}_{\overline{k}}}f

Then, we have

Δ2f\displaystyle\Delta^{2}f =gij¯φgkl¯φl¯kj¯if\displaystyle=g^{i\overline{j}}_{\varphi}g^{k\overline{l}}_{\varphi}\nabla_{\overline{l}}\nabla_{k}\nabla_{\overline{j}}\nabla_{i}f
=gij¯φgkl¯φl¯(j¯kif\tensorRkmij¯fm)\displaystyle=g^{i\overline{j}}_{\varphi}g^{k\overline{l}}_{\varphi}\nabla_{\overline{l}}\big{(}\nabla_{\overline{j}}\nabla_{k}\nabla_{i}f-\tensor{R}{{}_{k}^{m}{}_{i}{}_{\overline{j}}}f_{m}\big{)}
=𝒟𝒟fgφkl¯gφmj¯Rickj¯fml¯gφkl¯gφmj¯(l¯Rickj¯)fm.\displaystyle=\mathcal{D}^{*}\mathcal{D}f-g_{\varphi}^{k\overline{l}}g_{\varphi}^{m\overline{j}}\textup{Ric}_{k\overline{j}}f_{m\overline{l}}-g_{\varphi}^{k\overline{l}}g_{\varphi}^{m\overline{j}}(\nabla_{\overline{l}}\textup{Ric}_{k\overline{j}})f_{m}.

Here 𝒟𝒟=gij¯gkl¯l¯j¯ki\mathcal{D}^{*}\mathcal{D}=g^{i\overline{j}}g^{k\overline{l}}\nabla_{\overline{l}}\nabla_{\overline{j}}\nabla_{k}\nabla_{i}. Then, we have

Mu𝒟𝒟uωφn=Mgφkl¯(ugφij¯j¯kiu)l¯ωφnM(gφkl¯ul¯gφij¯j¯kiu)ωφn.\displaystyle\int_{M}u\mathcal{D}^{*}\mathcal{D}u\omega_{\varphi}^{n}=\int_{M}g_{\varphi}^{k\overline{l}}\big{(}ug_{\varphi}^{i\overline{j}}\nabla_{\overline{j}}\nabla_{k}\nabla_{i}u\big{)}_{\overline{l}}\omega_{\varphi}^{n}-\int_{M}\big{(}g_{\varphi}^{k\overline{l}}u_{\overline{l}}g_{\varphi}^{i\overline{j}}\nabla_{\overline{j}}\nabla_{k}\nabla_{i}u\big{)}\omega_{\varphi}^{n}.

The Stokes’ theorem can be applied to the first term in the above formula by observing that if we write 𝔥=ihkdzk=i(ugφij¯j¯kiu)dzk\mathfrak{h}=ih_{k}dz^{k}=i(ug_{\varphi}^{i\overline{j}}\nabla_{\overline{j}}\nabla_{k}\nabla_{i}u)dz^{k}, then gφkl¯(hk)l¯ωφn=n¯𝔥ωφn1g_{\varphi}^{k\overline{l}}(h_{k})_{\overline{l}}\omega_{\varphi}^{n}=n\overline{\partial}\mathfrak{h}\wedge\omega_{\varphi}^{n-1}. Hence,

Mgφkl¯(ugφij¯j¯kiu)l¯ωφn=Mn𝔥ωφn1.\displaystyle\int_{M}g_{\varphi}^{k\overline{l}}\big{(}ug_{\varphi}^{i\overline{j}}\nabla_{\overline{j}}\nabla_{k}\nabla_{i}u\big{)}_{\overline{l}}\omega_{\varphi}^{n}=\int_{\partial M}n\mathfrak{h}\wedge\omega_{\varphi}^{n-1}.

Similarly,

M(gφkl¯ul¯gφij¯j¯kiu)ωφn\displaystyle-\int_{M}\big{(}g_{\varphi}^{k\overline{l}}u_{\overline{l}}g_{\varphi}^{i\overline{j}}\nabla_{\overline{j}}\nabla_{k}\nabla_{i}u\big{)}\omega_{\varphi}^{n} =Mgφij¯(gφkl¯ul¯kiu)j¯ωφn+M|𝒟u|2ωφn\displaystyle=-\int_{M}g_{\varphi}^{i\overline{j}}\big{(}g_{\varphi}^{k\overline{l}}u_{\overline{l}}\nabla_{k}\nabla_{i}u\big{)}_{\overline{j}}\omega_{\varphi}^{n}+\int_{M}|\mathcal{D}u|^{2}\omega_{\varphi}^{n}
=M|𝒟u|2ωφnniMgφkl¯ul¯u,kidziωφn1.\displaystyle=\int_{M}|\mathcal{D}u|^{2}\omega_{\varphi}^{n}-ni\int_{\partial M}g_{\varphi}^{k\overline{l}}u_{\overline{l}}u_{,ki}dz^{i}\wedge\omega_{\varphi}^{n-1}.

We have

MuΔ2u\displaystyle\int_{M}u\Delta^{2}u =M|𝒟u|2Mugφkl¯gφmj¯(Ricφ)kj¯uml¯Mugφkl¯gφmj¯(l¯(Ricφ)kj¯)um\displaystyle=\int_{M}|\mathcal{D}u|^{2}-\int_{M}ug_{\varphi}^{k\overline{l}}g_{\varphi}^{m\overline{j}}(\textup{Ric}_{\varphi})_{k\overline{j}}u_{m\overline{l}}-\int_{M}ug_{\varphi}^{k\overline{l}}g_{\varphi}^{m\overline{j}}\big{(}\nabla_{\overline{l}}(\textup{Ric}_{\varphi})_{k\overline{j}}\big{)}u_{m}
niMugφij¯j¯kiudzkωφn1niMgφkl¯ul¯u,kidziωφn1.\displaystyle-ni\int_{\partial M}ug_{\varphi}^{i\overline{j}}\nabla_{\overline{j}}\nabla_{k}\nabla_{i}udz^{k}\wedge\omega_{\varphi}^{n-1}-ni\int_{\partial M}g_{\varphi}^{k\overline{l}}u_{\overline{l}}u_{,ki}dz^{i}\wedge\omega_{\varphi}^{n-1}.

Notice that

MRicφ,iu¯uωφn=Mgφij¯gφkl¯(Ricφ)il¯ukuj¯ωφn,\displaystyle-\int_{M}\langle\textup{Ric}_{\varphi},i\partial u\wedge\overline{\partial}u\rangle\omega_{\varphi}^{n}=-\int_{M}g_{\varphi}^{i\overline{j}}g_{\varphi}^{k\overline{l}}(\textup{Ric}_{\varphi})_{i\overline{l}}u_{k}u_{\overline{j}}\omega_{\varphi}^{n},

and integrating by parts,

Mgφij¯gkl¯φ(Ricφ)il¯ukuj¯ωφn\displaystyle-\int_{M}g_{\varphi}^{i\overline{j}}g^{k\overline{l}}_{\varphi}(\textup{Ric}_{\varphi})_{i\overline{l}}u_{k}u_{\overline{j}}\omega_{\varphi}^{n} =Mgij¯φ(gkl¯φ(Ricφ)il¯uku)j¯ωφn+Mugij¯φgkl¯φ(Ricφ)il¯ukj¯ωφn\displaystyle=\int_{M}g^{i\overline{j}}_{\varphi}\big{(}g^{k\overline{l}}_{\varphi}(\textup{Ric}_{\varphi})_{i\overline{l}}u_{k}u\big{)}_{\overline{j}}\omega_{\varphi}^{n}+\int_{M}ug^{i\overline{j}}_{\varphi}g^{k\overline{l}}_{\varphi}(\textup{Ric}_{\varphi})_{i\overline{l}}u_{k\overline{j}}\omega_{\varphi}^{n}
+Mugφij¯gφkl¯(j¯(Ricφ)il¯)ukωφn\displaystyle+\int_{M}ug_{\varphi}^{i\overline{j}}g_{\varphi}^{k\overline{l}}\big{(}\nabla_{\overline{j}}(\textup{Ric}_{\varphi})_{i\overline{l}}\big{)}u_{k}\omega_{\varphi}^{n}
=niMugφkl¯(Ricφ)il¯ukdziωφn1+Mugij¯φgkl¯φ(Ricφ)il¯ukj¯ωφn\displaystyle=ni\int_{\partial M}ug_{\varphi}^{k\overline{l}}(\textup{Ric}_{\varphi})_{i\overline{l}}u_{k}dz^{i}\wedge\omega_{\varphi}^{n-1}+\int_{M}ug^{i\overline{j}}_{\varphi}g^{k\overline{l}}_{\varphi}(\textup{Ric}_{\varphi})_{i\overline{l}}u_{k\overline{j}}\omega_{\varphi}^{n}
+Mugφij¯gφkl¯(j¯(Ricφ)il¯)ukωφn.\displaystyle+\int_{M}ug_{\varphi}^{i\overline{j}}g_{\varphi}^{k\overline{l}}\big{(}\nabla_{\overline{j}}(\textup{Ric}_{\varphi})_{i\overline{l}}\big{)}u_{k}\omega_{\varphi}^{n}.

Hence, we proved that

Mu(Δ2u)ωφn\displaystyle\int_{M}u(\Delta^{2}u)\omega_{\varphi}^{n} MRicφ,iu¯uωφn=M|𝒟u|2ωφn+niMugφkl¯(Ricφ)il¯ukdziωφn1\displaystyle-\int_{M}\langle\textup{Ric}_{\varphi},i\partial u\wedge\overline{\partial}u\rangle\omega_{\varphi}^{n}=\int_{M}|\mathcal{D}u|^{2}\omega_{\varphi}^{n}+ni\int_{\partial M}ug_{\varphi}^{k\overline{l}}(\textup{Ric}_{\varphi})_{i\overline{l}}u_{k}dz^{i}\wedge\omega_{\varphi}^{n-1}
niMugφij¯j¯kiudzkωφn1niMgφkl¯ul¯u,kidziωφn1,\displaystyle-ni\int_{\partial M}ug_{\varphi}^{i\overline{j}}\nabla_{\overline{j}}\nabla_{k}\nabla_{i}udz^{k}\wedge\omega_{\varphi}^{n-1}-ni\int_{\partial M}g_{\varphi}^{k\overline{l}}u_{\overline{l}}u_{,ki}dz^{i}\wedge\omega_{\varphi}^{n-1},

which completes the proof of the lemma. ∎

The integration formula (7.6) in this lemma, together with (7.5), completes the proof of (7.2). It suffices to show that all boundary terms in this formula vanish as RR\to\infty. According to Theorem B, we can check that the decay rates of the integrands integrated on M\partial M are at most 2τ1<2n+1-2\tau-1<-2n+1. This completes the proof. ∎

Theorem 7.3.

Assume that ω\omega is an ALE Kähler metric on XX such that the Ricci curvature of ω\omega is non-positive, Ric(ω)0\textup{Ric}(\omega)\leq 0. Then, along each ε\varepsilon-geodesic in 2τ+2(ω)\mathcal{H}_{-2\tau+2}(\omega), φ(t)\varphi(t), the Mabuchi KK-energy is convex.

Proof.

The proof is parallel to Chen [9]. Here, we just do the calculation in the ALE setting. Define f=φ¨12|φ˙|ωφ2\displaystyle f=\ddot{\varphi}-\frac{1}{2}|\nabla\dot{\varphi}|_{\omega_{\varphi}}^{2}. Then the ε\varepsilon-geodesic equation can be written as

εωnωnφ=f.\displaystyle\varepsilon\frac{\omega^{n}}{\omega^{n}_{\varphi}}=f.

According to (7.3), together with the observation, Ric(ωφ)=Ric(ω)+ddclogf\textup{Ric}(\omega_{\varphi})=\textup{Ric}(\omega)+dd^{c}\log f, we have

d2𝒦dt2\displaystyle\frac{d^{2}\mathcal{K}}{dt^{2}} =X|𝒟φ˙(t)|2ωφωφnXfR(ωφ)ωφn\displaystyle=\int_{X}\big{|}\mathcal{D}\dot{\varphi}(t)\big{|}^{2}_{\omega_{\varphi}}\omega_{\varphi}^{n}-\int_{X}fR(\omega_{\varphi})\,\omega_{\varphi}^{n}
=X|𝒟φ˙(t)|2ωφωφnXftrωφRic(ω)ωφnXfΔωφlogfωnφ\displaystyle=\int_{X}\big{|}\mathcal{D}\dot{\varphi}(t)\big{|}^{2}_{\omega_{\varphi}}\omega_{\varphi}^{n}-\int_{X}f\operatorname{tr}_{\omega_{\varphi}}\textup{Ric}(\omega)\,\omega_{\varphi}^{n}-\int_{X}f\Delta_{\omega_{\varphi}}\log f\,\omega^{n}_{\varphi}
=X|𝒟φ˙(t)|2ωφωφnXftrωφRic(ω)ωφn+X|f|2ωφfωφn0.\displaystyle=\int_{X}\big{|}\mathcal{D}\dot{\varphi}(t)\big{|}^{2}_{\omega_{\varphi}}\omega_{\varphi}^{n}-\int_{X}f\operatorname{tr}_{\omega_{\varphi}}\textup{Ric}(\omega)\,\omega_{\varphi}^{n}+\int_{X}\frac{|\nabla f|^{2}_{\omega_{\varphi}}}{f}\,\omega_{\varphi}^{n}\geq 0.

We have the last equality because f|logf|ωφ=O(r2τ1)f|\nabla\log f|_{\omega_{\varphi}}=O(r^{-2\tau-1}) and 2τ1<(2n1)-2\tau-1<-(2n-1), so that the relevant boundary integral vanishes. Hence, we have proved the convexity of the Mabuchi KK-energy. ∎

Remark 7.4.

A quick corollary of Theorem 7.3 is that assuming Ric(ω)0\textup{Ric}(\omega)\leq 0, the scalar-flat Kähler metric if it exists, is unique in 2τ+2(ω)\mathcal{H}_{-2\tau+2}(\omega). The proof is also parallel to Chen [9]. However, if there exists a scalar-flat Kähler metric in 2τ+2(ω)\mathcal{H}_{-2\tau+2}(\omega), the condition, Ric(ω)0\textup{Ric}(\omega)\leq 0, implies Ric(ω)=0\textup{Ric}(\omega)=0. Hence, the uniqueness of scalar-flat ALE Kähler metric can be reduced to the uniqueness result of Ricci-flat ALE Kähler metric (which can be found in many reference [20, 28, 13]). A short proof is given as follows. Let ω0\omega_{0} be a scalar-flat Kähler metric in 2τ+2(ω)\mathcal{H}_{-2\tau+2}(\omega). The fact, ω0=ω+O(r2τ)\omega_{0}=\omega+O(r^{-2\tau}), implies that the ADM mass of ω\omega and ω0\omega_{0} are equal, 𝔪(ω)=𝔪(ω1)\mathfrak{m}(\omega)=\mathfrak{m}(\omega_{1}). According to the mass formula by Hein-LeBrun [19],

𝔪(ω)=A(n,c1(X),[ω])+B(n)XR(ω)ωnn!,\displaystyle\mathfrak{m}(\omega)=A(n,c_{1}(X),[\omega])+B(n)\int_{X}R(\omega)\frac{\omega^{n}}{n!},

where A(n,c1(X),[ω])A(n,c_{1}(X),[\omega]) is a constant only determined by the dimension nn, the first Chern class of XX and the cohomology class of ω\omega and B(n)B(n) only depends on dimension nn. The fact, 𝔪(ω)=𝔪(ω1)\mathfrak{m}(\omega)=\mathfrak{m}(\omega_{1}), together with the mass formula, implies that

XR(ω)=XR(ω1)=0.\displaystyle\int_{X}R(\omega)=\int_{X}R(\omega_{1})=0.

The assumption that Ric(ω)0\textup{Ric}(\omega)\leq 0 implies that Ric(ω)=0\textup{Ric}(\omega)=0. Then, by a simple argument, we can prove that all scalar-flat ALE Kähler metrics in [ω][\omega] is actually Ricci-flat. The expansion of scalar-flat Kähler metrics (Theorem 1.1) implies that the Ricci form, Ric(ω1)\textup{Ric}(\omega_{1}), decays to zero at infinite with decay rate faster than 2n-2n. The ddbar lemma implies that there exist f𝒞22nf\in\mathcal{C}^{\infty}_{2-2n} such that

Ric(ω1)=ddcf.\displaystyle\textup{Ric}(\omega_{1})=dd^{c}f.

Taking trace with respect to gg, we have that Δf=0\Delta f=0. By solving the Laplacian equation (for instance, see [30, Propsition 2.3]), there is a unique solution in the space 𝒞δ\mathcal{C}^{\infty}_{-\delta} (for δ(,0)\D-\delta\in(-\infty,0)\backslash D). Hence, f0f\equiv 0, which implies that ω\omega is Ricci-flat.

8. Nonexistence of non-positive (or non-negative) Ricci curvature

Consider the standard family of negative line bundles, 𝒪(k)\mathcal{O}(-k), over n1\mathbb{C}\mathbb{P}^{n-1} together with their natural projections π:𝒪(k)n1\pi:\mathcal{O}(-k)\rightarrow\mathbb{C}\mathbb{P}^{n-1}. The total spaces of 𝒪(k)\mathcal{O}(-k) are fundamental examples of ALE Kähler manifolds by viewing 𝒪(k)\mathcal{O}(-k) as a resolution space of n/k\mathbb{C}^{n}/\mathbb{Z}_{k}. Let ω\omega be any ALE Kähler metric on 𝒪(k)\mathcal{O}(-k) asymptotic to the Euclidean metric with decay rate τ-\tau (τ>0\tau>0). In the following, we shall prove the nonexistence of a sign of the Ricci curvature of ω\omega in the case knk\neq n. When k=nk=n, there always exists a Ricci-flat ALE Kähler metric in each compactly supported ALE Kähler class, see [21, 22, 27].

Theorem 8.1.

Let 𝒪(k)\mathcal{O}(-k) be the standard negative line bundle over n1\mathbb{C}\mathbb{P}^{n-1} with n2n\geq 2 and knk\neq n. Let ω\omega be an ALE Kähler metric on 𝒪(k)\mathcal{O}(-k) with decay rate τ-\tau (τ>0)(\tau>0). Then, the Ricci form of ω\omega, ρ\rho, is of mixed type, i.e., neither ρ0\rho\geq 0 nor ρ0\rho\leq 0 is true.

Proof.

Notice that for each integer k1k\geq 1, there is a compactification of 𝒪(k)\mathcal{O}(-k) by adding a divisor at infinity, Dn1D_{\infty}\cong\mathbb{C}\mathbb{P}^{n-1}. We denote the compactified manifold as MkM_{k} and the natural embedding j:𝒪(k)Mkj:\mathcal{O}(-k)\rightarrow M_{k} is holomorphic. MkM_{k} is a 1\mathbb{C}\mathbb{P}^{1}-bundle over n1\mathbb{C}\mathbb{P}^{n-1}. Denote D0D_{0} as the divisor corresponding to the base manifold, n1𝒪(k)Mk\mathbb{C}\mathbb{P}^{n-1}\subset\mathcal{O}(-k)\hookrightarrow M_{k}. Then, the normal line bundles of D0D_{0} and DD_{\infty} are given by

ND0/Mk=𝒪(k),ND/Mk=𝒪(k).\displaystyle N_{D_{0}/M_{k}}=\mathcal{O}(-k),\quad N_{D_{\infty}/M_{k}}=\mathcal{O}(k). (8.1)

The following facts on the geometry of MkM_{k} can be checked by viewing MkM_{k} as a smooth toric variety. MkM_{k} can be described by 2n2n coordinate charts with coordinates {Ui;ui1,,\{U_{i};\ u_{i}^{1},\ldots, uin1,ui}u_{i}^{n-1},u_{i}\}, {Vi;vi1,,vin1,vi}\{V_{i};\ v_{i}^{1},\ldots,v_{i}^{n-1},v_{i}\} (0in10\leq i\leq n-1), where the coordinates are related by

(u1i,,un1i,ui)\displaystyle(u^{1}_{i},\ldots,u^{n-1}_{i},u_{i}) =(1ui0,u01ui0,,ui0ui0^,,un10ui0,(ui0)kun1),1in1,\displaystyle=\left(\frac{1}{u^{i}_{0}},\frac{u_{0}^{1}}{u^{i}_{0}},\ldots,\widehat{\frac{u^{i}_{0}}{u^{i}_{0}}},\ldots,\frac{u^{n-1}_{0}}{u^{i}_{0}},(u^{i}_{0})^{k}u^{n}_{1}\right),\quad 1\leq i\leq n-1,
(v1i,,vn1i,vi)\displaystyle(v^{1}_{i},\ldots,v^{n-1}_{i},v_{i}) =(u1i,,un1i,1ui),0in1.\displaystyle=\left(u^{1}_{i},\ldots,u^{n-1}_{i},\frac{1}{u_{i}}\right),\quad 0\leq i\leq n-1.

The divisor classes of MkM_{k} are generated by the class of D0n1D_{0}\cong\mathbb{C}\mathbb{P}^{n-1}, the zero section of 𝒪(k)Mk\mathcal{O}(-k)\subset M_{k}, and the class of DfD_{f}, the total space of the restriction of the 1\mathbb{C}\mathbb{P}^{1}-bundle MkD0M_{k}\to D_{0} to a linear subspace of D0D_{0}. Restricting D0D_{0} and DfD_{f} to U0U_{0}, we can write

D0=(u0=0)¯,Df=(u01=0)¯.\displaystyle D_{0}=\overline{(u_{0}=0)},\qquad D_{f}=\overline{(u_{0}^{1}=0)}.

The divisor at infinity, DD_{\infty}, can be represented by (u0=)=(v0=0)(u_{0}=\infty)=(v_{0}=0) and DD_{\infty} can be represented in terms of D0D_{0} and DfD_{f} as follows,

D=D0+kDf\displaystyle D_{\infty}=D_{0}+kD_{f} (8.2)

By viewing D0D_{0}, DfD_{f} and DD_{\infty} as smooth complex hypersurfaces of MkM_{k}, the Poincaré duals of D0D_{0}, DfD_{f} and DD_{\infty} have natural explicit representatives denoted by ρ0\rho_{0}, ρf\rho_{f}, ρ\rho_{\infty} respectively. For instance, in U0U_{0},

ρ0|U0\displaystyle\rho_{0}|_{U_{0}} =1nπi¯log(1+j|u0j|2)k|u0|2+1(1+j|u0j|2)k|u0|2,\displaystyle=\frac{1}{n\pi}i\partial\overline{\partial}\log\frac{(1+\sum_{j}|u_{0}^{j}|^{2})^{k}|u_{0}|^{2}+1}{(1+\sum_{j}|u_{0}^{j}|^{2})^{k}|u_{0}|^{2}}, (8.3)
ρf|U0\displaystyle\rho_{f}|_{U_{0}} =1nπi¯log(1+j|u0j|2),\displaystyle=\frac{1}{n\pi}i\partial\overline{\partial}\log\big{(}1+\sum_{j}|u_{0}^{j}|^{2}\big{)}, (8.4)
ρ|U0\displaystyle\rho_{\infty}|_{U_{0}} =1nπi¯log[(1+j|u0j|2)k|u0|2+1].\displaystyle=\frac{1}{n\pi}i\partial\overline{\partial}\log\big{[}\big{(}1+\sum_{j}|u_{0}^{j}|^{2}\big{)}^{k}|u_{0}|^{2}+1\big{]}. (8.5)

Step 1: Extension of the ALE Ricci form to MkM_{k}.

Recall that the diffeomorphism Φ:(n)/k𝒪(k)D0\Phi:(\mathbb{C}^{n})^{*}/\mathbb{Z}_{k}\rightarrow\mathcal{O}(-k)\setminus D_{0} gives a holomorphic asymptotic chart of 𝒪(k)\mathcal{O}(-k). The diffeomorphism Φ\Phi can be explicitly written as

Φ:(n)𝒪(k)D0,Φ(z1,,zn)|U0=(z2z1,,znz1,z1k).\displaystyle{\Phi}:(\mathbb{C}^{n})^{*}\rightarrow\mathcal{O}(-k)\setminus D_{0},\quad{\Phi}(z_{1},\ldots,z_{n})\big{|}_{U_{0}}=\Big{(}\frac{z_{2}}{z_{1}},\ldots,\frac{z_{n}}{z_{1}},{z_{1}^{k}}\Big{)}.

In the coordinate chart {U0;u01,,u0n1,u0}\{U_{0};\ u_{0}^{1},\ldots,u_{0}^{n-1},u_{0}\}, we have r2k=(1+j|u0j|2)k|u0|2r^{2k}=(1+\sum_{j}|u_{0}^{j}|^{2})^{k}|u_{0}|^{2}. By the asymptotic condition of ω\omega, in an asymptotic chart of 𝒪(k)\mathcal{O}(-k), log(ωn/ω0n)\log(\omega^{n}/\omega_{0}^{n}) can be viewed as a function of decay order O(rτ)O(r^{-\tau}), where ω0\omega_{0} is the standard Euclidean metric on the asymptotic chart. Thus, the Ricci form satisfies

ρ=i¯logωnωn0=O(rτ2).\displaystyle\rho=-i\partial\overline{\partial}\log\frac{\omega^{n}}{\omega^{n}_{0}}=O(r^{-\tau-2}).

The adjunction formula tells us that as line bundles over 𝒪(k)\mathcal{O}(-k),

K𝒪(k)=nkk[D0].\displaystyle K_{\mathcal{O}(-k)}=\frac{n-k}{k}[D_{0}].

Since ρ\rho is the curvature form of a Hermitian metric on K𝒪(k)1K_{\mathcal{O}(-k)}^{-1} and ρ0\rho_{0} is the curvature form of a Hermitian metric on [D0][D_{0}], it follows that

ρ+nkkρ0is globally i¯-exact.\displaystyle\rho+\frac{n-k}{k}\rho_{0}\;\,\text{is globally $i\partial\overline{\partial}$-exact}.

By restricting ρ0\rho_{0} in (8.3) to the asymptotic chart of 𝒪(k)\mathcal{O}(-k), we have

ρ0=i¯log(1+r2k).\displaystyle\rho_{0}=-i\partial\overline{\partial}\log(1+r^{-2k}).

Hence, by Theorem 1.1, ρ\rho can be written as

ρ=nkkρ0+i¯f for f𝒞τ(𝒪(k)),τ=min{2k,τ}>0.\displaystyle\rho=-\frac{n-k}{k}\rho_{0}+i\partial\overline{\partial}f\quad\text{ for }f\in\mathcal{C}^{\infty}_{-\tau^{\prime}}(\mathcal{O}(-k)),\quad\tau^{\prime}=\min\{2k,\tau\}>0.

Since ρ\rho cannot be extended smoothly to MkM_{k}, we define a smooth cut-off function χ\chi,

χ(t)={1,0t1,0,t2,smooth,1<t<2,\displaystyle\chi(t)=\begin{cases}1,\quad&0\leq t\leq 1,\\ 0,\quad&t\geq 2,\\ \text{smooth},\quad&1<t<2,\end{cases}

and we define χR(t)=χ(t/R)\chi_{R}(t)=\chi(t/R). Applying the cutoff function, we can extend ρ\rho to be

ρR={nkkρ0+i¯(χRf),in MkD,nkkρ0,on D.\displaystyle\rho_{R}=\begin{cases}\displaystyle-\frac{n-k}{k}\rho_{0}+i\partial\overline{\partial}\big{(}\chi_{R}f\big{)},\quad&\text{in }M_{k}\setminus D_{\infty},\\ \vspace{-2.5mm}\\ \displaystyle-\frac{n-k}{k}\rho_{0},\quad&\text{on }D_{\infty}.\end{cases}

Step 2: Integral argument for n=2n=2.

Recall that the intersection numbers between DD_{\infty}, D0D_{0} and DfD_{f} are given by

(D0)(D0)=k,(D0)(Df)=1,(Df)(Df)=0,(D0)(D)=0.\displaystyle(D_{0})\cdot(D_{0})=-k,\quad(D_{0})\cdot(D_{f})=1,\quad(D_{f})\cdot(D_{f})=0,\quad(D_{0})\cdot(D_{\infty})=0. (8.6)

In particular, if we integrate ρ\rho over D0D_{0}, then

D0ρ=D0ρR=MkρRρ0=2kkMkρ02=2k.\displaystyle\int_{D_{0}}{\rho}=\int_{D_{0}}\rho_{R}=\int_{M_{k}}{\rho_{R}}\wedge\rho_{0}=-\frac{2-k}{k}\int_{M_{k}}\rho_{0}^{2}=2-k. (8.7)

On the other hand, we have ρRρ\rho_{R}\to\rho pointwise and ρR=O(rτ2)\rho_{R}=O(r^{-\tau^{\prime}-2}) uniformly as RR\to\infty. Hence, by the dominated convergence theorem,

{u10=0}ρ=limR{u10=0}ρR=MkρRρf=2kkMkρ0ρf=k2k.\displaystyle\int_{\{u^{1}_{0}=0\}}\rho=\lim_{R\rightarrow\infty}\int_{\{u^{1}_{0}=0\}}\rho_{R}=\int_{M_{k}}{\rho_{R}}\wedge\rho_{f}=-\frac{2-k}{k}\int_{M_{k}}\rho_{0}\wedge\rho_{f}=\frac{k-2}{k}. (8.8)

Now assume that ρ\rho is seminegative (or semi-positive). Then the left-hand sides of both (8.7) and (8.8) are non-positive (or non-negative). However, the right-hand sides have opposite signs because k2k\neq 2. This is a contradiction.

Step 3: Integral argument for n3n\geq 3.

In higher dimension, the difficulty is to calculate the intersection numbers of divisors. However, in the case of MkM_{k}, we can apply the formula of intersection numbers on toric varieties [16, Chapter VII.6], or, more explicitly, take integral of formulas of Poincaré dual (8.3)–(8.5). Notice that

D0ρ0n1=(k)n1.\displaystyle\int_{D_{0}}\rho_{0}^{n-1}=(-k)^{n-1}. (8.9)

Then, we have

D0ρn1=D0ρRn1=MkρRn1ρ0=(nkk)n1(k)n1.\displaystyle\int_{D_{0}}{\rho}^{n-1}=\int_{D_{0}}\rho_{R}^{n-1}=\int_{M_{k}}\rho_{R}^{n-1}\wedge\rho_{0}=\Big{(}{-\frac{n-k}{k}}\Big{)}^{n-1}(-k)^{n-1}.

On the other hand, we have ρRρ\rho_{R}\to\rho pointwise and ρR=O(rτ2)\rho_{R}=O(r^{-\tau^{\prime}-2}) uniformly as RR\to\infty. Hence, by the dominated convergence theorem,

{u10=0}ρn1=limR{u10=0}ρRn1=limRDfρRn1=limRMkρRn1ρf=(nkk)n1Mkρ0n1ρf=(nkk)n1(k)n2,\displaystyle\begin{split}\int_{\{u^{1}_{0}=0\}}\rho^{n-1}&=\lim_{R\rightarrow\infty}\int_{\{u^{1}_{0}=0\}}\rho_{R}^{n-1}=\lim_{R\rightarrow\infty}\int_{D_{f}}\rho_{R}^{n-1}=\lim_{R\to\infty}\int_{M_{k}}\rho_{R}^{n-1}\wedge\rho_{f}\\ &=\Big{(}{-\frac{n-k}{k}\Big{)}^{n-1}}\int_{M_{k}}\rho_{0}^{n-1}\wedge\rho_{f}=\Big{(}{-\frac{n-k}{k}}\Big{)}^{n-1}(-k)^{n-2},\end{split}

where the last equality can be observed from (8.2) and (8.9):

Mkρ0n1ρf=Mkρ0n11k(ρρ0)=01kMkρ0n=(k)n2\displaystyle\int_{M_{k}}\rho_{0}^{n-1}\wedge\rho_{f}=\int_{M_{k}}\rho_{0}^{n-1}\wedge\frac{1}{k}(\rho_{\infty}-\rho_{0})=0-\frac{1}{k}\int_{M_{k}}\rho_{0}^{n}=(-k)^{n-2}

because ρ0|D=0\rho_{0}|_{D_{\infty}}=0. By the same argument as in dimension 22, we complete the proof. ∎

References

  • [1] S. Ali Aleyasin, The space of Kähler potentials on an asymptotically locally euclidean Kähler manifold, preprint, arXiv:1311.0419 (2013).
  • [2] C. Arezzo and F. Pacard, Blowing up and desingularizing constant scalar curvature Kähler manifolds, Acta Math. 196 (2006), 179–228.
  • [3] H. Auvray, The space of Poincaré type Kähler metrics on the complement of a divisor, J. reine angew. Math. 722 (2017), 1–64.
  • [4] R. Bartnik, The mass of an asymptotically flat manifold, Comm. Pure Appl. Math. 39 (1986), 661–693.
  • [5] E. Bedford and B. A. Taylor, The Dirichlet problem for a complex Monge-Ampère equation, Invent. math. 37 (1976), 1–44.
  • [6] Z. Błocki, A gradient estimate in the Calabi-Yau theorem, Math. Ann. 344 (2009), 317–327.
  • [7] S. Boucksom, Monge-Ampère equations on complex manifolds with boundary, Lecture Notes in Mathematics, vol. 2038, Springer, 2012, pp. 257–282.
  • [8] L. Caffarelli, J. Kohn, L. Nirenberg, and J. Spruck, The Dirichlet problem for nonlinear second order elliptic equations, II: Complex Monge-Ampère, and uniformly elliptic equations, Comm. Pure Appl. Math. 38 (1985), 209–252.
  • [9] X. Chen, The space of Kähler metrics, J. Differ. Geom. 56 (2000), 189–234.
  • [10] S.-Y. Cheng and S.-T. Yau, On the existence of a complete Kähler metric on non-compact complex manifolds and the regularity of Fefferman’s equation, Comm. Pure Appl. Math. 33 (1980), 507–544.
  • [11] P. Chruściel, Boundary conditions at spatial infinity from a Hamiltonian point of view, Topological Properties and Global Structure of Space-Time (Erice, 1985), Plenum Press, 1986, pp. 49–59.
  • [12] J. Chu, V. Tosatti, and B. Weinkove, On the regularity of geodesics in the space of Kähler metrics, Ann. PDE 3 (2017), 1–12.
  • [13] R. J. Conlon and H.-J. Hein, Asymptotically conical Calabi–Yau manifolds, I, Duke Math. J. 162 (2013), 2855–2902.
  • [14] T. Darvas and L. Lempert, Weak geodesics in the space of Kähler metrics, Math. Res. Lett. 19 (2012), 1127–1135.
  • [15] S. K. Donaldson, Symmetric spaces, Kahler geometry and Hamiltonian dynamics, Translations of the AMS, vol. 196, American Mathematical Society, 1999, pp. 13–34.
  • [16] G. Ewald, Combinatorial convexity and algebraic geometry, Graduate Texts in Mathematics, vol. 168, Springer, 1996.
  • [17] B. Guan, The Dirichlet problem for complex Monge-Ampère equations and regularity of the pluri-complex Green function, Comm. Anal. Geom. 6 (1998), 687–703.
  • [18] W. He, On the space of Kähler potentials, Comm. Pure Appl. Math. 68 (2015), 332–343.
  • [19] H.-J. Hein and C. LeBrun, Mass in Kähler geometry, Comm. Math. Phys. 347 (2016), 183–221.
  • [20] D. D. Joyce, Compact manifolds with special holonomy, Oxford University Press, 2000.
  • [21] P. B. Kronheimer, The construction of ALE spaces as hyper-Kähler quotients, J. Differ. Geom. 29 (1989), 665–683.
  • [22] P. B. Kronheimer, A Torelli-type theorem for gravitational instantons, J. Differ. Geom. 29 (1989), 685–697.
  • [23] C. LeBrun, Counter-examples to the generalized positive action conjecture, Comm. Math. Phys. 118 (1988), 591–596.
  • [24] L. Lempert and L. Vivas, Geodesics in the space of Kähler metrics, Duke Math. J. 162 (2013), 1369–1381.
  • [25] J. Lohkamp, Metrics of negative ricci curvature, Annals of Mathematics 140 (1994), no. 3, 655–683.
  • [26] S. Semmes, Complex Monge-Ampère and symplectic manifolds, Amer. J. Math. 114 (1992), 495–550.
  • [27] C. van Coevering, Ricci-flat Kähler metrics on crepant resolutions of Kähler cones, Math. Ann. 347 (2010), 581–611.
  • [28] Craig van Coevering, Regularity of asymptotically conical ricci-flat kähler metrics, 2010.
  • [29] D. Wu, Kähler-Einstein metrics of negative Ricci curvature on general quasi-projective manifolds, Comm. Anal. Geom. 16 (2008), 395–435.
  • [30] Q. Yao, Mass and expansion of asymptotically conical Kähler metrics, preprint, arXiv:2206.06505 (2022).
  • [31] S.-T. Yau, On the Ricci curvature of a compact Kähler manifold and the complex Monge-Ampére equation, I, Comm. Pure Appl. Math. 31 (1978), 339–411.