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

Calabi-Yau metrics of Calabi type with polynomial rate of convergence

Yifan Chen Department of Mathematics
University of California
Berkeley, CA, USA, 94720
[email protected]
Abstract.

We present new complete Calabi-Yau metrics defined on the complement of a smooth anticanonical divisor with ample normal bundle, approaching the Calabi model space at a polynomial rate. Moreover, we establish the uniqueness of this type of Calabi-Yau metric within a fixed cohomology class.

1. Introduction

In 1978, Yau [22] gave the celebrated proof of Calabi conjecture by solving the Monge-Ampère equation for each Kähler class on compact Kähler manifolds. In 1990, Tian and Yau [20] expanded upon this achievement by constructing complete Calabi-Yau metrics on quasi-projective manifolds, extending the techniques introduced in [22] to the non-compact case. Specifically, when MM is a Fano manifold and DD is a smooth irreducible anticanonical divisor, they constructed complete Calabi-Yau metric ωTY\omega_{TY} on MDM\setminus D, called Tian-Yau metric. ωTY\omega_{TY} is exponentially close to the Calabi model space. Here the Calabi model space (𝒞,I𝒞,ω𝒞,Ω𝒞)(\mathcal{C},I_{\mathcal{C}},\omega_{\mathcal{C}},\Omega_{\mathcal{C}}) is the disc bundle over DD within NDN_{D} removing the zero section, with complex structure I𝒞I_{\mathcal{C}} given by NDN_{D}, Kähler metric ω𝒞\omega_{\mathcal{C}} given by Calabi anstaz and a nowhere vanishing (n,0)(n,0)-form Ω𝒞\Omega_{\mathcal{C}}. The strict definition of Calabi model space will be given in Section 2. In fact, for any compact supported class in MDM\setminus D, we can find Calabi-Yau metric exponentially close to the Calabi model space. We call the complete Calabi-Yau manifold with this property asymptotically Calabi, which is defined as follows.

Definition 1.1.

Let XX be an nn-dim complete Kähler manifold with, complex structure II, 2-form ω\omega and (n,0)(n,0)-form Ω\Omega. We say (X,I,ω,Ω)(X,I,\omega,\Omega) is

  1. (1)

    weak asymptotically Calabi with rate κ\kappa if there exists δ¯>0\underline{\delta}>0, κ>0\kappa>0, a Calabi model space (𝒞,I𝒞,ω𝒞,Ω𝒞)(\mathcal{C},I_{\mathcal{C}},\omega_{\mathcal{C}},\Omega_{\mathcal{C}}) with function zz on 𝒞\mathcal{C} defined in Section 2, and a diffeomorphism Φ:𝒞𝒦XK\Phi:\mathcal{C}\setminus\mathcal{K}\rightarrow X\setminus K, where KXK\subset X and 𝒦𝒞\mathcal{K}\subset\mathcal{C} are compact, such that the following hold uniformly as z+z\rightarrow+\infty:

    |ω𝒞k(ΦII𝒞)|ω𝒞+|ω𝒞k(ΦΩΩ𝒞)|ω𝒞=O(eδ¯zn/2),\left|\nabla_{\omega_{\mathcal{C}}}^{k}\left(\Phi^{*}I-I_{\mathcal{C}}\right)\right|_{\omega_{\mathcal{C}}}+\left|\nabla_{\omega_{\mathcal{C}}}^{k}\left(\Phi^{*}\Omega-\Omega_{\mathcal{C}}\right)\right|_{\omega_{\mathcal{C}}}=O(e^{-\underline{\delta}z^{n/2}}),
    |ω𝒞k(Φωω𝒞)|ω𝒞=O(zκ) for all k0.\quad\left|\nabla_{\omega_{\mathcal{C}}}^{k}\left(\Phi^{*}\omega-\omega_{\mathcal{C}}\right)\right|_{\omega_{\mathcal{C}}}=O(z^{-\kappa})\text{ for all }k\in\mathbb{N}_{0}.
  2. (2)

    asymptotically Calabi if it is weak asymptotically Calabi and

    |ω𝒞k(Φωω𝒞)|ω𝒞=O(eδ¯zn/2) for all k0.\quad\left|\nabla_{\omega_{\mathcal{C}}}^{k}\left(\Phi^{*}\omega-\omega_{\mathcal{C}}\right)\right|_{\omega_{\mathcal{C}}}=O(e^{-\underline{\delta}z^{n/2}})\text{ for all }k\in\mathbb{N}_{0}.

The asymptotically Calabi Calabi-Yau manifold is well understood by Hein-Sun-Viaclovsky-Zhang [7], in which they showed that any asymptotically Calabi Calabi-Yau manifold can be compactified complex analytically to a weak Fano manifold, i.e. a smooth projective manifold with nef and big anti-canonical bundle. In this paper, we are going to show the existence and uniqueness of weak asymptotically Calabi Calabi-Yau metrics that is not asymptotically Calabi. Our setting is as follows:

Definition 1.2.

Let MM be a compact Kähler manifold with complex dimension n3n\geqslant 3, D|KM|D\in|-K_{M}| be a smooth divisor with ample normal bundle and X=MDX=M\setminus D. We denote H+2(X)H_{+}^{2}(X) as the subset of H2(X)H^{2}(X) which consists of classes 𝔨\mathfrak{k} such that 𝔨p\mathfrak{k}^{p} is positively paired with any compact analytic subset YY of XX of pure complex dimension pp.

The main result of this paper is the following

Theorem 1.3.

For any class 𝔨\mathfrak{k} in H+2(X)H_{+}^{2}(X), there exists a Calabi-Yau metric ω\omega in the class 𝔨\mathfrak{k} which is weak asymptotically Calabi with rate 11.

To show the existence, we use the method in Tian-Yau [20], which was subsequently generalized and refined by Hein [8]. This Tian-Yau-Hein’s package facilitates the production of complete Calabi-Yau metrics when a suitable model metric at infinity is known. However, we cannot directly apply it here because the 2-form on XX coming from the restriction of 2-forms on MM will only decay at the rate r2n+1r^{-\frac{2}{n+1}}, while Tian-Yau-Hein’s package requires the Ricci potential decays faster than r2r^{-2}. We need to modify our background metric by solving the linearized operator of Monge-Ampère equation to improve the decay.

Remark 1.4.

The metric in Theorem 1.3 is new in the sense that under a fixed diffeomorphism Φ\Phi, ω\omega is weak asymptotically Calabi with rate 11 but not asymptotically Calabi. We can also find Calabi-Yau metrics when dimM=2\mathrm{dim}_{\mathbb{C}}M=2. However, these metrics are not new. In fact, the 2 dimensional case is well understood: Sun-Zhang [17] showed that the ALHALH^{*} gravitational instanton is always asymptotically Calabi.

Besides, we also prove that the constructed metric is unique in the sense that:

Theorem 1.5.

Fix any pXp\in X. Let rr be the distance function towards pp under the metric ω\omega in Theorem 1.3. If we have another Calabi-Yau metric ω~\tilde{\omega} in the same class 𝔨\mathfrak{k} satisfying

|ω~ω|ωCrκ, as r,\left|\,\tilde{\omega}-\omega\right|_{\omega}\leqslant C\cdot r^{-\kappa},\text{ as }r\to\infty,

for some positive constant C,κC,\kappa, then ω~=ω\tilde{\omega}=\omega.

One could always have uniqueness if the metrics in Theorem 1.5 are close with rate O(rN)O(r^{-N}) for some NN large enough. Nevertheless, this rate is much faster than the decay rate of our metric constructed in Theorem 1.3. The proof of uniqueness theorem under any polynomial closeness requires two ingredients: the i¯i\partial\bar{\partial} lemma with the L2L^{2} estimate on XX, and the Liouville type theorem on 𝒞\mathcal{C}.

Outline of the paper

The paper is organized as follows. In Section 2, we provide an introductory overview of the Calabi ansatz, denoted by ω𝒞\omega_{\mathcal{C}}, within an open neighborhood 𝒞\mathcal{C} of the divisor DD in its normal bundle, excluding the zero section. Our exposition primarily adheres to the notation and discussion presented by Hein, Sun, Viaclovsky, and Zhang [10]. Furthermore, we briefly review the analytical framework developed by Tian, Yau, and Hein [8], list some geometric properties of the Calabi model space that facilitates the application of uniform elliptic estimates later.

In section 3, we have the solution of the Poisson equation with appropriate weighted estimates. The method of variable separation, as detailed by Sun and Zhang [16], allows us to simplify the Poisson equation in the model space into a particular form of ordinary differential equation. Leveraging the solutions’ estimates for these ordinary differential equations, as presented in Appendix A, we construct an inversion of the Laplacian operator within suitably weighted spaces. This enables us to initiate the iterative processes detailed in Sections 4 and 5.

In Sections 4 and 5, we construct a good background metric ω\omega within the cohomology class 𝔨\mathfrak{k}. We solve the Poisson equation on the model space iteratively to enhance the decay rate of the Ricci potential of ω\omega, as detailed in Section 4. In Section 5, we refine the Kähler potential by incorporating the harmonic moment map zz alongside other pluri-subharmonic functions. This adjustment ensures not only the positivity of ω\omega but also its compliance with the integral condition outlined in Tian-Yau-Hein’s package [8]. The iterative method adopted here is inspired by Conlon and Hein [5]. The specific technique of modifying the potential via the harmonic moment map zz is equivalent to choosing appropriate scaling of the metric hDh_{D} on the normal bundle in Hein, Sun, Viaclovsky, and Zhang [7].

In Section 6, we deform our good background metric ω\omega to a genuine Calabi-Yau metric on XX by Tian-Yau-Hein’s package. We also show that the perturbed metric is weak asymptotically Calabi with rate 11. This requires a slight generalization of Hein’s decay result in [8].

In Section 7, we discuss the uniqueness with restricted asymptotics of the Calabi-Yau metric in the fixed class. We first prove the i¯i\partial\bar{\partial}-lemma on XX based on the global Hörmander L2L^{2} estimate. Then by our solution of the Poisson equation and the behavior of the harmonic function on the model space we can deduce the global C0C^{0} estimate to do integration by parts.

In Section 8, we present some examples of Calabi-Yau manifolds which are weak asymptotically Calabi but not asymptotically Calabi under a fixed diffeomorphism. We also make some conjectures about the stronger uniqueness theorem and the compactification and classification of weak asymptotically Calabi manifold or under even weaker condition.

In the following sections, CC and ϵ\epsilon will be two uniform constants that may vary.

Other works on complete non-compact Calabi-Yau manifolds

There are many progress in the exploration of new complete Calabi-Yau manifolds, extending the seminal work of Tian and Yau to cases where DD is singular. Collins-Li [2] constructed new complete Calabi-Yau metrics when DD consists of two proportional transversely intersecting smooth divisors. The metric in their construction is i¯i\partial\bar{\partial} exact and polynomially closed to the generalized Calabi anstaz. Later Collins-Tong-Yau [3] solved a certain free boundary Monge-Ampère equation crucial to the inductive strategy proposed in [2] to deal with the general case when DD is simple normal crossing.

There are also many interesting works on this problem based on Tian-Yau-Hein’s package. For example, the non-flat Calabi-Yau metric on n\mathbb{C}^{n} constructed by Li [12], Székelyhidi [18] and Conlon-Rochon [6] with maximal volume growth, and Min [13] with volume growth 2n12n-1 when nn is even. Those works constructed Calabi-Yau with singular tangent cone at infinity.

Recently, Apostolov-Cifarelli [1] constructed new complete Calabi-Yau metrics on n\mathbb{C}^{n} with volume growth 2n12n-1. Their new method using toric geometry and Hamiltonian 2-forms is significantly different from Tian-Yau-Hein package and produces exotic complete Calabi-Yau metrics with interesting behavior at infinity.

Acknowledgement

The author is deeply grateful to Professor Song Sun for suggesting this problem, enlightening discussion and constant support. The author thanks Junsheng Zhang for fruitful conversation and encouragement, and thanks Yueqing Feng and Hongyi Liu for reading the draft of the paper and many helpful comments.The author also thanks IASM for their hospitality during the visit when the research was partially carried out, and the NSF for the generosity of the grant DMS-2304692.

2. Calabi model space

Let us give a brief introduction of Calabi ansatz and some notations we will use later. The notations in this section mainly follow Hein-Sun-Viaclovsky-Zhang [10, Section 3].

2.1. Calabi ansatz

Let MM be an nn-dimensional compact Kähler manifold with n2n\geqslant 2, DD be a smooth anticanonical divisor with ample normal bundle NDN_{D} and let X=MDX=M\setminus D denote the complement of DD in MM. By adjunction formula we know that DD has trivial canonical bundle with a nowhere-vanishing holomorphic volume form ΩD\Omega_{D} such that

(2πc1(ND))n1=Di(n1)2ΩDΩ¯D.(2\pi c_{1}(N_{D}))^{n-1}=\int_{D}i^{(n-1)^{2}}\Omega_{D}\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu_{D}.

Hence by Yau’s theorem, up to scaling we have a unique hermitian metric hDh_{D} on ND=KM|DN_{D}=\left.-K_{M}\right|_{D} with curvature form ωD\omega_{D} on DD in the class 2πc1(D)2\pi c_{1}(D) satisfying

ωDn1=i(n1)2ΩDΩ¯D.\omega_{D}^{n-1}=i^{(n-1)^{2}}\Omega_{D}\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu_{D}.

We fix such an hDh_{D}. For any point ξND\xi\in N_{D}, let t=log|ξ|hD2t=-\log|\xi|^{2}_{h_{D}} and z=t1nz=t^{\frac{1}{n}} be functions on the complement of the zero section in the total space NDDN_{D}\setminus D. Let 𝒞={0<|ξ|hD<1}ND\mathcal{C}=\{0<\left|\xi\right|_{h_{D}}<1\}\subseteq N_{D} be the disc bundle over DD with complex structure I𝒞I_{\mathcal{C}} restricted from NDN_{D}. On 𝒞\mathcal{C}, we have a Calabi-Yau metric given by the Calabi ansatz:

ω𝒞=nn+1i¯(tn+1n)=zi¯t+1nzn1it¯t.\omega_{\mathcal{C}}=\frac{n}{n+1}i\partial\bar{\partial}(t^{\frac{n+1}{n}})=zi\partial\bar{\partial}t+\frac{1}{nz^{n-1}}i\partial t\wedge\bar{\partial}t.

Let Ω𝒞\Omega_{\mathcal{C}} denotes the unique holomorphic (n,0)(n,0)-form on NDN_{D} such that

Z\righthalfcupΩ𝒞=πΩD,Z\righthalfcup\Omega_{\mathcal{C}}=\pi^{*}\Omega_{D},

where ZZ denotes the holomorphic vector field generated by the scalar multiplication along the fiber direction and π:NDD\pi:N_{D}\to D is the projection map. And the data (𝒞,I𝒞,ω𝒞,Ω𝒞)(\mathcal{C},I_{\mathcal{C}},\omega_{\mathcal{C}},\Omega_{\mathcal{C}}) is called Calabi model space.

On MM we can also choose a holomorphic section SS of KM-K_{M} to be the defining function of DD such that S1S^{-1} can be seen as an (n,0)(n,0)-form ΩX\Omega_{X} on XX with a simple pole and residue ΩD\Omega_{D} along DD. We choose a metric hMh_{M} of KM-K_{M} such that hM|D=hD\left.h_{M}\right|_{D}=h_{D}. Then we can construct the following (1,1)(1,1)-form on XX:

ωX=nn+1i¯(log|S|hM2)n+1n.\displaystyle\omega_{X}=\frac{n}{n+1}i\partial\bar{\partial}\left(-\log\left|S\right|_{h_{M}}^{2}\right)^{\tfrac{n+1}{n}}.

As proved in Hein-Sun-Viaclovsky-Zhang [10, Proposition 3.4.], it is asymptotically Calabi in the following way:

Proposition 2.1.

The complex structure IXI_{X} and ICI_{C}, metric ω𝒞\omega_{\mathcal{C}} and ωX\omega_{X} and canonical form ΩX\Omega_{X} and Ω𝒞\Omega_{\mathcal{C}} are exponentially closed. To be more precise: there exists a compact set KK in XX, a compact set 𝒦\mathcal{K} in 𝒞\mathcal{C}, and a diffeomorphism induced by exponential map Φ:𝒞𝒦XK\Phi:\mathcal{C}\setminus\mathcal{K}\to X\setminus K such that for all k0,ϵ>0k\geqslant 0,\epsilon>0:

|g𝒞k(ΦIXI𝒞)|g𝒞+|g𝒞k(ΦΩXΩ𝒞)|g𝒞+|g𝒞k(ΦωXω𝒞)|g𝒞=O(e(12ϵ)zn).\left|\nabla_{g_{\mathcal{C}}}^{k}\left(\Phi^{*}I_{X}-I_{\mathcal{C}}\right)\right|_{g_{\mathcal{C}}}+\left|\nabla_{g_{\mathcal{C}}}^{k}\left(\Phi^{*}\Omega_{X}-\Omega_{\mathcal{C}}\right)\right|_{g_{\mathcal{C}}}+\left|\nabla_{g_{\mathcal{C}}}^{k}\left(\Phi^{*}\omega_{X}-\omega_{\mathcal{C}}\right)\right|_{g_{\mathcal{C}}}=O(e^{-(\frac{1}{2}-\epsilon)z^{n}}).

2.2. Geometric properties and Tian-Yau-Hein’s package

Let us first list some geometry of Calabi model space directly coming from the formula of ω𝒞\omega_{\mathcal{C}}. ω𝒞\omega_{\mathcal{C}} is complete when |ξ|hD0\left|\xi\right|_{h_{D}}\to 0 and incomplete when |ξ|hD1\left|\xi\right|_{h_{D}}\to 1. It has volume growth of order 2nn+1\tfrac{2n}{n+1} and its sectional curvature decays at the rate r2n+1r^{-\frac{2}{n+1}}. Any distance function on 𝒞\mathcal{C} will be comparable to zn+12z^{\frac{n+1}{2}}.

Then we introduce some requirements of the base Riemannian manifold (N,g)(N,g) in Tian-Yau-Hein’s package and show that our model space (𝒞,ω𝒞)(\mathcal{C},\omega_{\mathcal{C}}) satisfies those properties. One may refer to [8] for details and examples.

We begin with the definition of SOB(ν)\mathrm{SOB}(\nu) property:

Definition 2.2.

Let (N,g)(N,g) be a complete noncompact Riemannian manifold with real dimension at least 3. We say (N,g)(N,g) satisfy SOB(ν)\mathrm{SOB}(\nu) condition for some ν>0\nu>0 if and only if there exists a point pNp\in N and a positive constant C>1C>1 such that

  1. (1)

    Volume growth is at most ν\nu, i.e. Volg(B(p,R))CRν\mathrm{Vol}_{g}(B(p,R))\leqslant CR^{\nu} for all R>CR>C.

  2. (2)

    Ric(x)Cdg(x,p)2\mathrm{Ric}(x)\geqslant-Cd_{g}(x,p)^{-2}, xN\forall x\in N.

  3. (3)

    Volg(B(x,(11C)dg(x,p)))1Cdg(x,p)ν\mathrm{Vol}_{g}(B(x,(1-\tfrac{1}{C})d_{g}(x,p)))\geqslant\tfrac{1}{C}d_{g}(x,p)^{\nu}.

  4. (4)

    For any D>CD>C, any two points x,yNx,y\in N with dg(p,x)=dg(p,y)=Dd_{g}(p,x)=d_{g}(p,y)=D can be joined by a curve of length at most CDC\cdot D, lying in the annulus A(p,1CD,CD):={xN|1CD<dg(p,x)<CD}A(p,\tfrac{1}{C}D,CD):=\{x\in N|\tfrac{1}{C}D<d_{g}(p,x)<CD\}.

Remark 2.3.

In the original definition of SOB(ν)\mathrm{SOB}(\nu) property in [9], we need that the annulus A(p,s,t)A(p,s,t) is connected for any t>s>0t>s>0. To apply Theorem 2.8 it suffices to check the RCA property (4) instead of the connectivity of the annulus.

Proposition 2.4.

(𝒞,ω𝒞)(\mathcal{C},\omega_{\mathcal{C}}) has SOB(2nn+1)\mathrm{SOB}(\tfrac{2n}{n+1}) property.

Proof.

(2)(2) follows from the Ricci-flat property of 𝒞\mathcal{C}. For (1)(1), recall that ω𝒞=zi¯t+1nzn1it¯t\omega_{\mathcal{C}}=zi\partial\bar{\partial}t+\frac{1}{nz^{n-1}}i\partial t\wedge\bar{\partial}t and the volume form ω𝒞n=(i¯t)n1it¯t\omega_{\mathcal{C}}^{n}=(i\partial\bar{\partial}t)^{n-1}\wedge i\partial t\wedge\bar{\partial}t. The distance function rω𝒞r_{\omega_{\mathcal{C}}} to some fixed point pp in 𝒞\mathcal{C} is comparable to zn+12z^{\frac{n+1}{2}}. Consequently, one can see that A~(z1,z2)={x𝒞|z1<z(x)<z2}\tilde{A}(z_{1},z_{2})=\{x\in\mathcal{C}\;|\;z_{1}<z(x)<z_{2}\} is comparable to the annulus A(p,R1,R2)A(p,R_{1},R_{2}) in the sense that: for any z2>z1>Cz_{2}>z_{1}>C and R2>R1>CR_{2}>R_{1}>C, we have

A~(z1,z2)\displaystyle\tilde{A}(z_{1},z_{2}) A(p,1Cz1n+12,Cz2n+12),\displaystyle\subseteq A\left(p,\tfrac{1}{C}z_{1}^{\frac{n+1}{2}},Cz_{2}^{\frac{n+1}{2}}\right),
A(p,R1,R2)\displaystyle A(p,R_{1},R_{2}) A~((1CR1)2n+1,(CR2)2n+1).\displaystyle\subseteq\tilde{A}\left(\left(\tfrac{1}{C}R_{1}\right)^{\frac{2}{n+1}},\left(CR_{2}\right)^{\frac{2}{n+1}}\right).

Similarly, for any RCR\leqslant C, we have:

B(p,R)𝒦A~(0,(CR)2n+1).\displaystyle B(p,R)\subseteq\mathcal{K}\cup\tilde{A}\left(0,(CR)^{\frac{2}{n+1}}\right).

Also, we can see from the ansatz that the diameter of {x𝒞|z=z0}\{x\in\mathcal{C}\,|\,z=z_{0}\} is comparable to z0\sqrt{z_{0}}, which shows that

A~(z(x)1C(z(x)z(p)),z(x)+1C(z(x)z(p)))B(x,(11C)dg(x,p)).\displaystyle\tilde{A}\left(z(x)-\tfrac{1}{C}\left(z(x)-z(p)\right),z(x)+\tfrac{1}{C}\left(z(x)-z(p)\right)\right)\subset B\left(x,\left(1-\tfrac{1}{C}\right)d_{g}(x,p)\right).

With all these equivalence, we know that

Volg(B(p,R))C+{(CR)2nn+1t(CR)2nn+1}(i¯t)n1dtdctCR2nn+1.\displaystyle\mathrm{Vol}_{g}(B(p,R))\leqslant C+\int_{\left\{-(CR)^{\frac{2n}{n+1}}\leqslant\,t\,\leqslant(CR)^{\frac{2n}{n+1}}\right\}}(i\partial\bar{\partial}t)^{n-1}\wedge dt\wedge d^{c}t\leqslant CR^{\frac{2n}{n+1}}.

Similarly for (3)(3) we have

Volg(B(x,(11C)dg(x,p)))Cz(x)nCdg(x,p)2nn+1.\displaystyle\mathrm{Vol}_{g}(B\left(x,\left(1-\tfrac{1}{C}\right)d_{g}(x,p)\right))\geqslant C\cdot z(x)^{n}\geqslant C\cdot d_{g}(x,p)^{\frac{2n}{n+1}}.

To show (4)(4), for DD large enough and any two points xx and yy with dg(x,p)=dg(y,p)=Dd_{g}(x,p)=d_{g}(y,p)=D, we have z(x),z(y)(1CD2n+1,CD2n+1)z(x),z(y)\in\left(\tfrac{1}{C}\cdot D^{\frac{2}{n+1}},C\cdot D^{\frac{2}{n+1}}\right). Then by the formula of ω𝒞\omega_{\mathcal{C}} we can join xx and yy by a curve of length at most C(D+D1n+1)C\cdot\left(D+D^{\frac{1}{n+1}}\right) lying in A(1CD2n+1,CD2n+1)A\left(\tfrac{1}{C}\cdot D^{\frac{2}{n+1}},C\cdot D^{\frac{2}{n+1}}\right) and consequently in the annulus A(p,1CD,CD)A(p,\tfrac{1}{C}D,CD).

Hence we have the SOB(2nn+1)\mathrm{SOB}(\tfrac{2n}{n+1}) condition on (𝒞,ω𝒞)(\mathcal{C},\omega_{\mathcal{C}}). ∎

We continue with the definition of HMG(λ,k,α)\operatorname{HMG}(\lambda,k,\alpha) property:

Definition 2.5.

We say that (Nn,g)\left(N^{n},g\right) is HMG(λ,k,α)\operatorname{HMG}(\lambda,k,\alpha), for some λ[0,1]\lambda\in[0,1], k0k\in\mathbb{N}_{0}, α(0,1)\alpha\in(0,1), if there exist x0Nx_{0}\in N and C1C\geqslant 1 such that

  1. (1)

    for every xNx\in N with r(x)Cr(x)\geqslant C there exists a local holomorphic diffeomorphism Φx\Phi_{x} from the unit ball BnB\subset\mathbb{R}^{n} into NN such that Φx(0)=x\Phi_{x}(0)=x and Φx(B)B(x,1Cr(x)λ)\Phi_{x}(B)\supset B\left(x,\tfrac{1}{C}r(x)^{\lambda}\right),

  2. (2)

    h:=r(x)2λΦxgh:=r(x)^{-2\lambda}\Phi_{x}^{*}g satisfies Inj(h)1C,1CgeuchCgeuc\operatorname{Inj}(h)\geqslant\tfrac{1}{C},\tfrac{1}{C}g_{\text{euc}}\leqslant h\leqslant Cg_{\text{euc}}, and hgeucCk,α(B,geuc)C\left\|h-g_{\text{euc}}\right\|_{C^{k,\alpha}\left(B,g_{\text{euc}}\right)}\leqslant C.

Tian-Yau [20, Proposition 1.2.] provides a simple criterion for a complete Kähler manifold (N,ω)(N,\omega) to be HMG(0,k,α)\operatorname{HMG}(0,k,\alpha). We refer to Hein’s thesis [8, Lemma 4.7.] for a slightly generalized statement and sketch of the proof.

Lemma 2.6.

A complete Kähler manifold with |Rm|+i=1kriλ|iScal|Cr2λ\left|\mathrm{Rm}\right|+\sum_{i=1}^{k}r^{i\lambda}\left|\nabla^{i}\mathrm{Scal}\right|\leqslant Cr^{-2\lambda} for some k0k\in\mathbb{N}_{0} and λ[0,1]\lambda\in[0,1] is HMG(λ,k+1,α)\operatorname{HMG}(\lambda,k+1,\alpha) for every α(0,1)\alpha\in(0,1).

Proposition 2.7.

(𝒞,ω𝒞)(\mathcal{C},\omega_{\mathcal{C}}) has HMG(1n+1,k,α)\mathrm{HMG}(\tfrac{1}{n+1},k,\alpha) property, for any k0k\in\mathbb{N}_{0} and α(0,1)\alpha\in(0,1).

Proof.

The proof goes almost verbatim with the proof of Lemma 2.6. The proof of Lemma 2.6 only used the completeness to guarantee that the injectivity radius of local universal cover around x0x_{0} has uniform lower bound independent of x0x_{0}. Since 𝒞\mathcal{C} is the disc bundle over DD, after we do rescaling by ω~𝒞=z(x0)1ω𝒞\tilde{\omega}_{\mathcal{C}}=z(x_{0})^{-1}\omega_{\mathcal{C}}, the S1S^{1} action gives the only collapsing direction which disappears after passing to the local universal cover. So the local universal cover has uniform curvature bound and is volume non-collapsing, which leads to uniform injectivity radius lower bound. ∎

Now we are ready to present the following result taken from Hein’s thesis [8] which is a powerful tool to give the existence of Calabi-Yau metric on complete noncompact Kähler manifold.

Theorem 2.8 (Tian-Yau-Hein’s Package).

Let (Xn,ω)(X^{n},\omega) be a complete noncompact Kähler manifold, which satisfies the condition SOB(ν)\mathrm{SOB}(\nu) and HMG(0,3,α)\mathrm{HMG}(0,3,\alpha) for some ν>0\nu>0, 0<α<10<\alpha<1. Let rr be the distance function to a fixed point pXp\in X with respect to the metric ω\omega. Let fC2,α(X)f\in C^{2,\alpha}(X) satisfy |f|Crμ|f|\leqslant Cr^{-\mu} on {r>1}\{r>1\} for some μ>2\mu>2 and X(ef1)ωn=0\int_{X}\left(e^{f}-1\right)\omega^{n}=0. Then there exist α¯(0,α]\bar{\alpha}\in(0,\alpha] and uC4,α¯(X)u\in C^{4,\bar{\alpha}}(X) such that (ω+i¯u)n=efωn\left(\omega+i\partial\bar{\partial}u\right)^{n}=e^{f}\omega^{n}. Moreover X|u|2ωn<\int_{X}|\nabla u|^{2}\omega^{n}<\infty. If in addition fClock,α¯(X)f\in C_{\mathrm{loc}}^{k,\bar{\alpha}}(X) for some k3k\geqslant 3, then all such solutions uu belong to Clock+2,α¯(X)C_{\mathrm{loc}}^{k+2,\bar{\alpha}}(X).

Remark 2.9.

We can not directly use Theorem 2.8 because in our setting, the decay rate of the class 𝔨\mathfrak{k} is only r2n+1r^{-\frac{2}{n+1}} which is slower than r2r^{-2}. So we need to modify the representative β\beta by i¯ui\partial\bar{\partial}u for some function uu on XX, which comes from the suitable solution of Poisson equation on model space 𝒞\mathcal{C}, which we will discuss in section 3.

2.3. Uniform elliptic estimates

The previous properties are mainly used to guarantee that we have weighted Sobolev inequality, weighted Hölder space, and can do weighted elliptic estimates.

Recall that the Ck,αC^{k,\alpha} norm of a function ff on a ball Bg(x,r)B_{g}(x,r) inside a manifold (X,g)(X,g) is

fCk,α(Bg(x,r))=j=1k1gjfC0(Bg(x,r))+gkfC0,α(Bg(x,r)).\displaystyle\left\|f\right\|_{C^{k,\alpha}(B_{g}(x,r))}=\sum_{j=1}^{k-1}\left\|\nabla_{g}^{j}f\right\|_{C^{0}(B_{g}(x,r))}+\left\|\nabla_{g}^{k}f\right\|_{C^{0,\alpha}(B_{g}(x,r))}.

If we scale the metric by a constant 1z0\frac{1}{z_{0}}, i.e. g~=1z0g\tilde{g}=\frac{1}{z_{0}}g, we then have

|g~jf|g~=z0j2|gjf|g,\displaystyle\left|\nabla_{\tilde{g}}^{j}f\right|_{\tilde{g}}=z_{0}^{\frac{j}{2}}\cdot|\nabla_{g}^{j}f|_{g},
fWk,p(Bg~(x,r))=j=0kz0j2gjfLp(Bg(x,z0r)),\displaystyle\left\|f\right\|_{W^{k,p}(B_{\tilde{g}}(x,r))}=\sum_{j=0}^{k}z_{0}^{\frac{j}{2}}\left\|\nabla_{g}^{j}f\right\|_{L^{p}(B_{g}(x,\sqrt{z_{0}}r))},
fCk,α(Bg~(x,r))=j=0k1z0j2gjfC0(Bg(x,z0r))+z0k+α2gkfC0,α(Bg(x,z0r)).\displaystyle\left\|f\right\|_{C^{k,\alpha}(B_{\tilde{g}}(x,r))}=\sum_{j=0}^{k-1}z_{0}^{\frac{j}{2}}\left\|\nabla_{g}^{j}f\right\|_{C^{0}(B_{g}(x,\sqrt{z_{0}}r))}+z_{0}^{\frac{k+\alpha}{2}}\left\|\nabla_{g}^{k}f\right\|_{C^{0,\alpha}(B_{g}(x,\sqrt{z_{0}}r))}.

Now we can prove the uniform Schauder estimate here for future reference:

Proposition 2.10.

Let (N,g)(N,g) be a manifold satisfying HMG(λ,k,α)\mathrm{HMG}(\lambda,k,\alpha) property. Let uu and vv be smooth functions on MM such that Δgu=v\Delta_{g}u=v. Then there exist some constants CkC_{k} such that

r(x)kλkuC0(Bg(x,r(x)λ))Ck(uC0(Bg(x,r(x)λ))+i=0k1r(x)(i+2)λivC0(Bg(x,r(x)λ))).\displaystyle r(x)^{k\lambda}\left\|\nabla^{k}u\right\|_{C^{0}(B_{g}(x,r(x)^{\lambda}))}\leqslant C_{k}\left(\left\|u\right\|_{C^{0}(B_{g}(x,r(x)^{\lambda}))}+\sum_{i=0}^{k-1}r(x)^{(i+2)\lambda}\left\|\nabla^{i}v\right\|_{C^{0}(B_{g}(x,r(x)^{\lambda}))}\right).
Proof.

Fix any xx in NN. Let h=r(x)2λΦgh=r(x)^{-2\lambda}\Phi^{*}g be the rescaled pull back metric on the unit ball B(0,1)B(0,1) in n\mathbb{R}^{n}, let u~=Φxu\tilde{u}=\Phi_{x}^{*}u, v~=r(x)2λΦxv\tilde{v}=r(x)^{2\lambda}\Phi_{x}^{*}v. Since we use the pull back metric, the Laplacian is preserved. We have Δhu~=v~\Delta_{h}\tilde{u}=\tilde{v}. Write this elliptic equation under the Euclidean coordinate, with khgeucC(k)\|\nabla^{k}h\|_{g_{\text{euc}}}\leqslant C(k) for any integer k0k\geqslant 0, we have u~\tilde{u} satisfies the following elliptic equation

1dethxi(hijdethxju~)=v~.\displaystyle\frac{1}{\sqrt{\operatorname{det}h}}\frac{\partial}{\partial x^{i}}\left(h^{ij}\sqrt{\operatorname{det}h}\frac{\partial}{\partial x^{j}}\tilde{u}\right)=\tilde{v}. (2.1)

By the standard elliptic estimates on the Euclidean space and passing to the original metric gg we get the required estimate. To be more precise, by W2,qW^{2,q} estimates we know that there exists a uniform constant CqC_{q} only depends on qq such that for any 1<q<+1<q<+\infty

u~W2,q(B(0,12))Cq(v~C0(B(0,1))+u~C0(B(0,1))).\displaystyle\left\|\tilde{u}\right\|_{W^{2,q}\left(B(0,\frac{1}{2})\right)}\leqslant C_{q}\cdot\left(\left\|\tilde{v}\right\|_{C^{0}(B(0,1))}+\left\|\tilde{u}\right\|_{C^{0}(B(0,1))}\right).

Consequently by Sobolev embedding, we know

u~C1,12nq(B(0,12))Cqu~W2,q(B(0,12))Cq(v~C0(B(0,1))+u~C0(B(0,1))).\displaystyle\left\|\tilde{u}\right\|_{C^{1,1-\frac{2n}{q}}\left(B(0,\frac{1}{2})\right)}\leqslant C_{q}\left\|\tilde{u}\right\|_{W^{2,q}\left(B(0,\frac{1}{2})\right)}\leqslant C_{q}\left(\left\|\tilde{v}\right\|_{C^{0}(B(0,1))}+\left\|\tilde{u}\right\|_{C^{0}(B(0,1))}\right).

Taking derivative of (2.1) under the Euclidean coordinate, we have:

u~W3,q(B(0,12))\displaystyle\left\|\tilde{u}\right\|_{W^{3,q}\left(B(0,\frac{1}{2})\right)} Cq(v~W1,q(B(0,23))+u~W2,q(B(0,23)))\displaystyle\leqslant C_{q}\cdot\left(\left\|\tilde{v}\right\|_{W^{1,q}\left(B(0,\frac{2}{3})\right)}+\left\|\tilde{u}\right\|_{W^{2,q}\left(B(0,\frac{2}{3})\right)}\right)
Cq(v~C1(B(0,1))+u~C0(B(0,1))).\displaystyle\leqslant C_{q}\cdot\left(\left\|\tilde{v}\right\|_{C^{1}(B(0,1))}+\left\|\tilde{u}\right\|_{C^{0}(B(0,1))}\right).

And by bootstrapping and Sobolev embedding

u~Ck+1,12nq(B(0,12))\displaystyle\left\|\tilde{u}\right\|_{C^{k+1,1-\frac{2n}{q}}\left(B(0,\frac{1}{2})\right)} Cq,ku~Wk+2,q(B(0,12))\displaystyle\leqslant C_{q,k}\left\|\tilde{u}\right\|_{W^{k+2,q}\left(B(0,\frac{1}{2})\right)}
Cq,k(v~Wk,q(B(0,23))+u~Wk+1,q(B(0,23)))\displaystyle\leqslant C_{q,k}\left(\left\|\tilde{v}\right\|_{W^{k,q}\left(B(0,\frac{2}{3})\right)}+\left\|\tilde{u}\right\|_{W^{k+1,q}\left(B(0,\frac{2}{3})\right)}\right)
Cq,k(v~Ck(B(0,1))+u~C0(B(0,1))).\displaystyle\leqslant C_{q,k}\left(\left\|\tilde{v}\right\|_{C^{k}\left(B(0,1)\right)}+\left\|\tilde{u}\right\|_{C^{0}\left(B(0,1)\right)}\right).

Passing to the original metric, we know that for any k1k\geqslant 1

r(x)kλkuC0(Bg(x,r(x)λ))Ck(uC0(Bg(x,r(x)λ))+i=0k1r(x)(i+2)λivC0(Bg(x,r(x)λ))).\displaystyle r(x)^{k\lambda}\left\|\nabla^{k}u\right\|_{C^{0}(B_{g}(x,r(x)^{\lambda}))}\leqslant C_{k}\left(\left\|u\right\|_{C^{0}(B_{g}(x,r(x)^{\lambda}))}+\sum_{i=0}^{k-1}r(x)^{(i+2)\lambda}\left\|\nabla^{i}v\right\|_{C^{0}(B_{g}(x,r(x)^{\lambda}))}\right).

3. Solving Poisson Equation on the Model Space

In this section, following the approach of Sun-Zhang in [16], we will use separation of variables to solve Δω𝒞u=v\Delta_{\omega_{\mathcal{C}}}u=v for uu, vv functions on 𝒞\mathcal{C} and give some uniform estimate of our solution. We use the same notation introduced in section 2.

We first notice that 𝒞\mathcal{C} is diffeomorphic to Y×Y\times\mathbb{R} where the level set Y={ξ𝒞|log|ξ|hD2=z0n}Y=\{\xi\in\mathcal{C}\,|\,-\log|\xi|^{2}_{h_{D}}=z_{0}^{n}\} for a fixed z0>0z_{0}>0 and YY is equipped with an S1S^{1} bundle structure over DD and a metric hYh_{Y} induced by ω𝒞\omega_{\mathcal{C}}. Let 0=Λ0<Λ1<0=\Lambda_{0}<\Lambda_{1}<\cdots be the spectrum of the Laplacian Δ(Y,hY)-\Delta_{(Y,h_{Y})} on YY with respect to the metric hYh_{Y}. Let {ψk}k=0\{\psi_{k}\}_{k=0}^{\infty} be the corresponding eigenfunctions with ψkL2(Y)=1\left\|\psi_{k}\right\|_{L^{2}(Y)}=1. They showed that Λk=z01λk+nz0n1jk2\Lambda_{k}=z_{0}^{-1}\lambda_{k}+nz_{0}^{n-1}j_{k}^{2} for some λk0\lambda_{k}\geqslant 0 and jkj_{k}\in\mathbb{N}. Moreover, {ψk}k=0\{\psi_{k}\}_{k=0}^{\infty} form an orthonormal basis of L2(Y)L^{2}(Y) and each ψk\psi_{k} is homogeneous of degree jkj_{k} under the S1S^{1} action. The product structure allows us to do Fourier expansion on 𝒞\mathcal{C}. In particular, for any smooth function vv on 𝒞\mathcal{C}, if we take Pk(v)(z)=Yv(z,y)ψk(y)P_{k}(v)(z)=\int_{Y}v(z,y)\psi_{k}(y), we can write

v(z,y)=k=0Pk(v)(z)ψk(y),\displaystyle v(z,y)=\sum_{k=0}^{\infty}P_{k}(v)(z)\cdot\psi_{k}(y), (3.1)

which is convergent in L2L^{2} sense. In fact, we will prove later that the convergence is in CkC^{k} if vv has proper higher regularity estimate. For the separated function u(z)ψ(y)u(z)\psi(y) we have

Δω𝒞u(z)ψ(y)=1nzn1(u′′(z)(λ+j2n4zn)nzn2u(z))ψ(y).\displaystyle\Delta_{\omega_{\mathcal{C}}}u(z)\psi(y)=\frac{1}{nz^{n-1}}(u^{\prime\prime}(z)-(\lambda+\frac{j^{2}n}{4}\cdot z^{n})nz^{n-2}u(z))\psi(y).

Moreover, [16] proved the following:

Proposition 3.1.

Let (𝒞,g𝒞)\left(\mathcal{C},g_{\mathcal{C}}\right) be the Calabi model space, and let uu solve the Poisson equation Δ𝒞u=v\Delta_{\mathcal{C}}u=v for some vCK0(𝒞)v\in C^{K_{0}}(\mathcal{C}) and K0K_{0}\in\mathbb{N} sufficiently large. Let uu, vv have ”fiber-wise” expansions as in (3.1). Then for every kk\in\mathbb{N}, the coefficient functions uk(z)u_{k}(z) and vk(z)v_{k}(z) satisfy

uk′′(z)(nλk+jk2n24zn)zn2uk(z)=nzn1vk(z),z1.\displaystyle u_{k}^{\prime\prime}(z)-\left(n\lambda_{k}+\frac{j_{k}^{2}n^{2}}{4}\cdot z^{n}\right)z^{n-2}u_{k}(z)=nz^{n-1}\cdot v_{k}(z),\quad z\geqslant 1. (3.2)

Specifically, they found a solution of Δω𝒞u=v\Delta_{\omega_{\mathcal{C}}}u=v by solving ODE (3.2). With the estimate of the solution of this equation, they showed that the L2L^{2} formal solution given by ukψk\sum u_{k}\psi_{k} is actually a regular solution to the Poisson equation. Similarly but directly via careful estimates, we can construct the solution of Poisson equation with respect to ω𝒞\omega_{\mathcal{C}} with some weighted regularity and finer polynomial growth order:

Proposition 3.2.

Assume that vv is a function on 𝒞\mathcal{C} such that for any kk\in\mathbb{N} there exist constants CkC_{k} and δ\delta such that |zk2kv|ω𝒞Ckzδ|z^{\frac{k}{2}}\nabla^{k}v|_{\omega_{\mathcal{C}}}\leqslant C_{k}z^{\delta} on z>Ckz>C_{k}. Then there exist constants CkC^{\prime}_{k} and a function u:𝒞u:\mathcal{C}\to\mathbb{R} such that Δω𝒞u=v\Delta_{\omega_{\mathcal{C}}}u=v and

|u|C0zδ+n+1+ϵ,|u|ω𝒞C1zδ+n+12+ϵ,|ku|ω𝒞Ckzδk22+ϵ\displaystyle|u|\leqslant C_{0}^{\prime}z^{\delta+n+1+\epsilon},\quad\left|\nabla u\right|_{\omega_{\mathcal{C}}}\leqslant C_{1}^{\prime}z^{\delta+\frac{n+1}{2}+\epsilon},\quad\left|\nabla^{k}u\right|_{\omega_{\mathcal{C}}}\leqslant C^{\prime}_{k}z^{\delta-\frac{k-2}{2}+\epsilon}

on z>Ckz>C_{k}^{\prime}, for any ϵ>0\epsilon>0 and any integer k2k\geqslant 2.

Proof.

Let

u(z,y)=u0(z)ψ0+j=1uj(z)ψj(y)\displaystyle u(z,y)=u_{0}(z)\psi_{0}+\sum_{j=1}^{\infty}u_{j}(z)\psi_{j}(y)

be the formal solution, where u0u_{0} is constructed as follows and uju_{j} is constructed in the Appendix A which mainly follows from Sun-Zhang [16].

Step 1: We first show that the formal solution converges in C0C^{0} sense with the polynomial order depending on the polynomial order of vv.

For u0u_{0}, we have

u0′′(z)=nzn1P0v(z),u0(z)=C2znsn1P0v(s)𝑑s,u0(z)=C1z(C2tnsn1P0v(s)𝑑s)𝑑t.\displaystyle u_{0}^{\prime\prime}(z)=nz^{n-1}P_{0}v(z),\quad u_{0}^{\prime}(z)=\int_{C_{2}}^{z}ns^{n-1}P_{0}v(s)ds,\quad u_{0}(z)=\int_{C_{1}}^{z}\left(\int_{C_{2}}^{t}ns^{n-1}P_{0}v(s)ds\right)dt.

We choose C1C_{1} and C2C_{2} here to be 11 or ++\infty depending on the order of P0vP_{0}v to make u0(z)u_{0}(z) and u0(z)u_{0}^{\prime}(z) finite with proper order. Also, the integration here is the only place that we will lose the rate zϵz^{\epsilon}.

For uju_{j}, we first prove that the projection Pj(v)P_{j}(v) is well-defined and has the following estimate:

|vj(z)|=|Pj(v)(z)|\displaystyle|v_{j}(z)|=\left|P_{j}(v)(z)\right| =|YΔhYK0v(z,y)ψj(Λj)K0dVolY|v(z,)C2K0(Y,hY)(Λj)K0CK0zδΛjK0\displaystyle=\left|\int_{Y}\frac{\Delta^{K_{0}}_{h_{Y}}v(z,y)\psi_{j}}{(\Lambda_{j})^{K_{0}}}\mathrm{dVol_{Y}}\right|\leqslant\frac{\left\|v(z,\cdot)\right\|_{C^{2K_{0}}(Y,h_{Y})}}{(\Lambda_{j})^{K_{0}}}\leqslant C_{K_{0}}\frac{z^{\delta}}{\Lambda_{j}^{K_{0}}} (3.3)

for any K0K_{0} and z>Cz>C. Consequently, the solution uju_{j} constructed in (A.3) and (A.9) for the equation (3.2) has the following C0C^{0} bound for z>Cn,δ,λ1z>C_{n,\delta,\lambda_{1}}:

|uj(z)|1(Λj)K0Cn,δ,λ1,K0zδ+1nλ1.\displaystyle\left|u_{j}(z)\right|\leqslant\frac{1}{(\Lambda_{j})^{K_{0}}}\frac{C_{n,\delta,\lambda_{1},K_{0}}z^{\delta+1}}{n\lambda_{1}}. (3.4)

Here the constant CC is uniform for jj and zz and only depends on K0K_{0} and nn.We also have the uniform estimate of ψj\psi_{j} as the eigenfunctions of ΔωY-\Delta_{\omega_{Y}} on YY by its eigenvalues showed in Sun-Zhang [16, Lemma 5.1.]:

ψjCk(Y)\displaystyle\left\|\psi_{j}\right\|_{C^{k}\left(Y\right)} Ck,Y(Λj)n+k2.\displaystyle\leqslant C_{k,Y}\cdot\left(\Lambda_{j}\right)^{\frac{n+k}{2}}. (3.5)

Combine (3.4) and (3.5), we get

|j=1ujψj|\displaystyle|\sum_{j=1}^{\infty}u_{j}\psi_{j}| j=1|uj(z)||ψj(y)|Cn,δ,λ1,K0,Yj=1zδ+1(Λj)K0n2.\displaystyle\leqslant\sum_{j=1}^{\infty}\left|u_{j}(z)\right|\cdot\left|\psi_{j}(y)\right|\leqslant C_{n,\delta,\lambda_{1},K_{0},Y}\sum_{j=1}^{\infty}\frac{z^{\delta+1}}{\left(\Lambda_{j}\right)^{K_{0}-\frac{n}{2}}}.

Weyl’s law gives the bound of Λj\Lambda_{j} with CY1j22n1|Λj|CYj22n1C_{Y}^{-1}j^{\frac{2}{2n-1}}\leq\left|\Lambda_{j}\right|\leqslant C_{Y}j^{\frac{2}{2n-1}} for some CYC_{Y} only depend on (Y,hY)(Y,h_{Y}). Take K0=2nK_{0}=2n we can conclude that the summation converges.

Step 2: We prove that uu is smooth. We mostly follow the proof in Sun-Zhang [16, Proposition 6.2.]. Let

UN=j=0Nujψj,VN=j=0Nvjψj.\displaystyle U_{N}=\sum_{j=0}^{N}u_{j}\psi_{j},\quad V_{N}=\sum_{j=0}^{N}v_{j}\psi_{j}.

Then we have Δω𝒞UN=VN\Delta_{\omega_{\mathcal{C}}}U_{N}=V_{N}.
We first show that VNV_{N} has CkC^{k} bound independent of NN. In fact, we have the higher regularity estimate for vjv_{j} as in (3.3):

|kvj(z)|\displaystyle|\nabla^{k}v_{j}(z)| =|kPj(v)(z)|=|kYΔhYK0v(z,y)ψj(Λj)K0dVolY|Cz,Y,K0,k1ΛjK0.\displaystyle=\left|\nabla^{k}P_{j}(v)(z)\right|=\left|\nabla^{k}\int_{Y}\frac{\Delta^{K_{0}}_{h_{Y}}v(z,y)\psi_{j}}{(\Lambda_{j})^{K_{0}}}\mathrm{dVol_{Y}}\right|\leqslant C_{z,Y,K_{0},k}\frac{1}{\Lambda_{j}^{K_{0}}}.

Then given by the estimate of ψj\psi_{j} (3.5), for any integer jj, we know that

j=1N|k(vjψj)|Cz,Y,K0,kj=1NΛjn+k2K0.\displaystyle\sum_{j=1}^{N}\left|\nabla^{k}(v_{j}\psi_{j})\right|\leqslant C_{z,Y,K_{0},k}\sum_{j=1}^{N}\Lambda_{j}^{\frac{n+k}{2}-K_{0}}.

So again by Weyl’s law VNV_{N} converges to vv as a CkC^{k} function. Now we can prove that UNU_{N} also have the uniform CkC^{k} bound with respect to NN via local elliptic estimates.

For any fixed point x𝒞x\in\mathcal{C}, we consider the ball Bω𝒞(x,1)B_{\omega_{\mathcal{C}}}(x,1). Then for any p>0p>0, there exists a constant Cp,xC_{p,x} such that

UNW2,p(Bω𝒞(x,12))Cp,x(VNC0(Bω𝒞(x,1))+UNC0(Bω𝒞(x,1)))Cp,x.\displaystyle\left\|U_{N}\right\|_{W^{2,p}(B_{\omega_{\mathcal{C}}}(x,\frac{1}{2}))}\leqslant C_{p,x}\cdot(\left\|V_{N}\right\|_{C^{0}(B_{\omega_{\mathcal{C}}}(x,1))}+\left\|U_{N}\right\|_{C^{0}(B_{\omega_{\mathcal{C}}}(x,1))})\leqslant C_{p,x}.

By bootstrapping and Sobolev embedding, for any k0k\geqslant 0 and p>0p>0, there exists a constant Cp,xC_{p,x} such that

UNCk+1,12np(Bω𝒞(x,12))\displaystyle\left\|U_{N}\right\|_{C^{k+1,1-\frac{2n}{p}}(B_{\omega_{\mathcal{C}}}(x,\frac{1}{2}))} Cp,xUNWk+2,p(Bω𝒞(x,12))\displaystyle\leqslant C_{p,x}\left\|U_{N}\right\|_{W^{k+2,p}(B_{\omega_{\mathcal{C}}}(x,\frac{1}{2}))}
Cp,x(VNWk,p(Bω𝒞(x,1))+UNWk,p(Bω𝒞(x,1)))Cp,k,x.\displaystyle\leqslant C_{p,x}\left(\left\|V_{N}\right\|_{W^{k,p}(B_{\omega_{\mathcal{C}}}(x,1))}+\left\|U_{N}\right\|_{W^{k,p}(B_{\omega_{\mathcal{C}}}(x,1))}\right)\leqslant C_{p,k,x}.

Consequently we have UNU_{N} converges to uu in CkC^{k}. Since xx is arbitrary, we know that uu is smooth on 𝒞\mathcal{C}.

Step 3. We can now give global bound on the C1C^{1}, C2C^{2} and higher regularity of uu. We treat u0u_{0} and uu0u-u_{0} seperately.

For u0u_{0}, we have explicit estimate by computation of the Christoffel symbol under the following holomorphic coordinate: Let π:𝒞D\pi:\mathcal{C}\rightarrow D be the projection map. For any point ξ𝒞\xi\in\mathcal{C} we take the local holomorphic coordinate z¯=(z1,z2,,zn1)\underline{z}=(z_{1},z_{2},\ldots,z_{n-1}) on BDB\subset D around the point π(ξ)\pi(\xi). Take ξ0\xi_{0} be a local holomorphic section of NDN_{D} such that |ξ0|hD=eφ\left|\xi_{0}\right|_{h_{D}}=e^{-\varphi}, where i¯φ=ωDi\partial\bar{\partial}\varphi=\omega_{D} with φ(0)=0\varphi(0)=0, φ(0)=0\nabla\varphi(0)=0. Then ξ=ξ0w\xi=\xi_{0}\cdot w where ww is the fiber coordinate. Recall that z=(log|ξ|hD2)1nz=\left(-\log|\xi|_{h_{D}}^{2}\right)^{\frac{1}{n}}, w=eiθt+φ/2w=e^{i\theta-t+\varphi/2} and

ω𝒞=zπωD+1nzn1i(dwwφ)(dw¯w¯¯φ).\omega_{\mathcal{C}}=z\pi^{*}\omega_{D}+\frac{1}{nz^{n-1}}\cdot i\cdot\left(\frac{dw}{w}-\partial\varphi\right)\wedge\left(\frac{d\bar{w}}{\bar{w}}-\bar{\partial}\varphi\right).

Under this coordinate, we can prove by induction that

|kdzi|ω𝒞Czk+12, for any k0,\displaystyle\left|\nabla^{k}dz_{i}\right|_{\omega_{\mathcal{C}}}\leqslant Cz^{-\frac{k+1}{2}},\text{ for any }k\geqslant 0,
|dww|ω𝒞zn12,|kdww|ω𝒞Czk+12, for any k1,\displaystyle\left|\frac{dw}{w}\right|_{\omega_{\mathcal{C}}}\leqslant z^{\frac{n-1}{2}},\quad\left|\nabla^{k}\frac{dw}{w}\right|_{\omega_{\mathcal{C}}}\leqslant Cz^{-\frac{k+1}{2}},\text{ for any }k\geqslant 1,
|dt|ω𝒞zn12,|kt|ω𝒞zk2,|kta|ω𝒞zn(a1)k2, for any k2.\displaystyle\left|dt\right|_{\omega_{\mathcal{C}}}\leqslant z^{\frac{n-1}{2}},\quad\left|\nabla^{k}t\right|_{\omega_{\mathcal{C}}}\leqslant z^{-\frac{k}{2}},\quad\left|\nabla^{k}t^{a}\right|_{\omega_{\mathcal{C}}}\leqslant z^{n(a-1)-\frac{k}{2}},\text{ for any }k\geqslant 2. (3.6)

Now we can estimate the higher derivative of u0u_{0}:

ku0\displaystyle\nabla^{k}u_{0} =j=1ku0(j)(z)i1++ij=ki1>0,,ij>0Ci1,,iji1zi2zijz,\displaystyle=\sum_{j=1}^{k}u_{0}^{(j)}(z)\sum_{\begin{subarray}{c}i_{1}+\cdots+i_{j}=k\\ i_{1}>0,\cdots,i_{j}>0\end{subarray}}C_{i_{1},\cdots,i_{j}}\nabla^{i_{1}}z\otimes\nabla^{i_{2}}z\otimes\cdots\otimes\nabla^{i_{j}}z,
=u0(z)kz+j=2ku0(j)(z)i1++ij=ki1>0,,ij>0Ci1,,iji1zi2zijz,\displaystyle=u_{0}^{\prime}(z)\nabla^{k}z+\sum_{j=2}^{k}u_{0}^{(j)}(z)\sum_{\begin{subarray}{c}i_{1}+\cdots+i_{j}=k\\ i_{1}>0,\cdots,i_{j}>0\end{subarray}}C_{i_{1},\cdots,i_{j}}\nabla^{i_{1}}z\otimes\nabla^{i_{2}}z\otimes\cdots\otimes\nabla^{i_{j}}z,
=u0(z)kz+j=2k(nzn1P0v(z))(j2)i1++ij=ki1>0,,ij>0Ci1,,iji1zi2zijz,\displaystyle=u_{0}^{\prime}(z)\nabla^{k}z+\sum_{j=2}^{k}\left(nz^{n-1}P_{0}v(z)\right)^{(j-2)}\sum_{\begin{subarray}{c}i_{1}+\cdots+i_{j}=k\\ i_{1}>0,\cdots,i_{j}>0\end{subarray}}C_{i_{1},\cdots,i_{j}}\nabla^{i_{1}}z\otimes\nabla^{i_{2}}z\otimes\cdots\otimes\nabla^{i_{j}}z,

where f(k)f^{(k)} refers to the higher derivative of the function ff. Since

(nzn1P0v(z))(j2)=l=0j2Cl,j,nznj+l+1P0v(l) and |P0v(l)(z)|CY|lv|ω𝒞|dz|ω𝒞lCYzδ+(n2)l2,\displaystyle\left(nz^{n-1}P_{0}v(z)\right)^{(j-2)}=\sum_{l=0}^{j-2}C_{l,j,n}z^{n-j+l+1}P_{0}v^{(l)}\text{ and }\left|P_{0}v^{(l)}(z)\right|\leqslant C_{Y}\left|\nabla^{l}v\right|_{\omega_{\mathcal{C}}}\cdot\left|dz\right|^{-l}_{\omega_{\mathcal{C}}}\leqslant C_{Y}z^{\delta+\frac{(n-2)l}{2}},

we know that

|ku0|ω𝒞\displaystyle\left|\nabla^{k}u_{0}\right|_{\omega_{\mathcal{C}}} Cn,k,Yzδ+1k2+ϵ+Cn,k,Yj=2kl=0j2znj+1+δ+nl2i1++ij=ki1>0,,ij>0|i1z|ω𝒞|i2z|ω𝒞|ijz|ω𝒞\displaystyle\leqslant C_{n,k,Y}z^{\delta+1-\frac{k}{2}+\epsilon}+C_{n,k,Y}\sum_{j=2}^{k}\sum_{l=0}^{j-2}z^{n-j+1+\delta+\frac{nl}{2}}\sum_{\begin{subarray}{c}i_{1}+\cdots+i_{j}=k\\ i_{1}>0,\cdots,i_{j}>0\end{subarray}}\left|\nabla^{i_{1}}z\right|_{\omega_{\mathcal{C}}}\otimes\left|\nabla^{i_{2}}z\right|_{\omega_{\mathcal{C}}}\otimes\cdots\otimes\left|\nabla^{i_{j}}z\right|_{\omega_{\mathcal{C}}}
Cn,k,Yzδ+1k2+ϵ+Cn,k,Yz1+δj=2ki1++ij=ki1>0,,ij>0|zn22i1z|ω𝒞|zn22i2z|ω𝒞|zn22ijz|ω𝒞.\displaystyle\leqslant C_{n,k,Y}z^{\delta+1-\frac{k}{2}+\epsilon}+C_{n,k,Y}z^{1+\delta}\sum_{j=2}^{k}\sum_{\begin{subarray}{c}i_{1}+\cdots+i_{j}=k\\ i_{1}>0,\cdots,i_{j}>0\end{subarray}}\left|z^{\frac{n-2}{2}}\nabla^{i_{1}}z\right|_{\omega_{\mathcal{C}}}\otimes\left|z^{\frac{n-2}{2}}\nabla^{i_{2}}z\right|_{\omega_{\mathcal{C}}}\otimes\cdots\otimes\left|z^{\frac{n-2}{2}}\nabla^{i_{j}}z\right|_{\omega_{\mathcal{C}}}.

By |z|ω𝒞zn12\left|\nabla z\right|_{\omega_{\mathcal{C}}}\leqslant z^{-\frac{n-1}{2}} and |iz|ω𝒞zi2n+1\left|\nabla^{i}z\right|_{\omega_{\mathcal{C}}}\leqslant z^{-\frac{i}{2}-n+1} for i2i\geqslant 2, we have

i1++ij=ki1>0,,ij>0|zn22i1z||zn22i2z||zn22ijz|Ckzk2,\displaystyle\sum_{\begin{subarray}{c}i_{1}+\cdots+i_{j}=k\\ i_{1}>0,\cdots,i_{j}>0\end{subarray}}\left|z^{\frac{n-2}{2}}\nabla^{i_{1}}z\right|\otimes\left|z^{\frac{n-2}{2}}\nabla^{i_{2}}z\right|\otimes\cdots\otimes\left|z^{\frac{n-2}{2}}\nabla^{i_{j}}z\right|\leqslant C_{k}z^{-\frac{k}{2}},

Consequently, we have

|u0|ω𝒞Cn,Yzδ+n+12+ϵ.\displaystyle\left|\nabla u_{0}\right|_{\omega_{\mathcal{C}}}\leqslant C_{n,Y}z^{\delta+\frac{n+1}{2}+\epsilon}.

and for k2k\geqslant 2

|ku0|ω𝒞Cn,k,Yzδ+1k2+ϵ.\displaystyle\left|\nabla^{k}u_{0}\right|_{\omega_{\mathcal{C}}}\leqslant C_{n,k,Y}z^{\delta+1-\frac{k}{2}+\epsilon}.

For uu0u-u_{0} we use our C0C_{0} bound of uu0u-u_{0} and do elliptic estimate around a point x{z=z0}x\in\{z=z_{0}\} by Proposition 2.10. We have the uniform estimate of uu0u-u_{0} on {zC}\{z\geqslant C^{\prime}\}:

zk2k(uu0)C0(Bg(x,z))Ck(uu0C0(Bg(x,z))+i=0k1zi+22ivP0(v)C0(Bg(x,z))),\displaystyle z^{\frac{k}{2}}\left\|\nabla^{k}(u-u_{0})\right\|_{C^{0}(B_{g}(x,\sqrt{z}))}\leqslant C_{k}\left(\left\|u-u_{0}\right\|_{C^{0}(B_{g}(x,\sqrt{z}))}+\sum_{i=0}^{k-1}z^{\frac{i+2}{2}}\left\|\nabla^{i}v-P_{0}(v)\right\|_{C^{0}(B_{g}(x,\sqrt{z}))}\right),

which yields

|zk2k(uu0)|ω𝒞Ckzδ+1.\displaystyle\left|z^{\frac{k}{2}}\nabla^{k}(u-u_{0})\right|_{\omega_{\mathcal{C}}}\leqslant C_{k}^{\prime}z^{\delta+1}.

Together with the estimate of u0u_{0}, we have

|u|C0zδ+n+1+ϵ,|u|ω𝒞C1zδ+n+12+ϵ,|ku|ω𝒞Ckzδk22+ϵ for k2.\displaystyle|u|\leqslant C_{0}^{\prime}z^{\delta+n+1+\epsilon},\quad\left|\nabla u\right|_{\omega_{\mathcal{C}}}\leqslant C_{1}^{\prime}z^{\delta+\frac{n+1}{2}+\epsilon},\quad\left|\nabla^{k}u\right|_{\omega_{\mathcal{C}}}\leqslant C_{k}^{\prime}z^{\delta-\frac{k-2}{2}+\epsilon}\text{ for }k\geqslant 2.

Remark 3.3.

We see from the proof that the main term in C0C^{0} and C1C^{1} estimate of the solution uu is the fiber direction u0u_{0}. However, for the higher estimate CkC^{k} where k3k\geqslant 3, they will give the same order contribution.

4. Improve the Decay of the Ricci potential

Now we look back at the quasi-projective manifold XX and the class 𝔨H2(X)\mathfrak{k}\in H^{2}(X). In this section we are going to find a good representative form β\beta inside the class 𝔨H2(X)\mathfrak{k}\in H^{2}(X). We look at its behavior on the model space and then find some function UU by finite step iteration such that

(Φβ+i¯U)nω𝒞n1\frac{(\Phi^{*}\beta+i\partial\bar{\partial}U)^{n}}{\omega_{\mathcal{C}}^{n}}-1

has faster decay rate.

4.1. A good representative

Lemma 4.1.

For any 𝔨H2(X)\mathfrak{k}\in H^{2}(X), there exists a closed (1,1)(1,1)-form β\beta on MM such that [β|X]=𝔨[\beta|_{X}]=\mathfrak{k}.

Proof.

By discussion in Section 2 of [7] we know MM is weak Fano, and consequently simply connected by [19]. Consider the exact sequence 0𝒪M(D)𝒪M𝒪D00\to\mathcal{O}_{M}(-D)\to\mathcal{O}_{M}\to\mathcal{O}_{D}\to 0, we have long exact sequence

H1(M,𝒪M)H1(D,𝒪D)H2(M,𝒪M(D))\cdots\to H^{1}(M,\mathcal{O}_{M})\to H^{1}(D,\mathcal{O}_{D})\to H^{2}(M,\mathcal{O}_{M}(-D))\to\cdots

On the other hand, by Serre duality we have

H2(M,𝒪M(D))H2(M,KM)Hn,n2(M,KM).H^{2}(M,\mathcal{O}_{M}(-D))\simeq H^{2}(M,K_{M})\simeq H^{n,n-2}(M,-K_{M}).

Apply Kawamata–Viehweg vanishing theorem to the nef and big line bundle KM-K_{M} so we know that Hn,p(M,KM)=H0,p(M)=0H^{n,p}(M,-K_{M})=H^{0,p}(M)=0 for p1p\geqslant 1. So when n3n\geqslant 3, we have H1(D,)=0H^{1}(D,\mathbb{C})=0. The long exact sequence given by the excision theorem and Thom-Gysin sequence

H0(D)H2(M)H2(X)H1(D)H^{0}(D)\to H^{2}(M)\to H^{2}(X)\to H^{1}(D)\to\cdots

yields that the restriction map j:H2(M)H2(X)j^{*}:H^{2}(M)\to H^{2}(X) induced by j:XMj:X\to M is surjective with dimKerj=1\mathrm{dim}\,\mathrm{Ker}j^{*}=1, generated by c1(KM)c_{1}(-K_{M}). Hence there exists a closed (1,1)(1,1)-form β\beta on MM such that [β|X]=𝔨[\beta|_{X}]=\mathfrak{k}. ∎

Remark 4.2.

Here we use the fact that dimX=n3\mathrm{dim}_{\mathbb{C}}X=n\geqslant 3 to apply Kawamata–Viehweg vanishing theorem. When n=2n=2 we have H2(M,KM)H0,0(M)H^{2}(M,K_{M})\simeq H^{0,0}(M)\simeq\mathbb{C}, so H1(D,)H^{1}(D,\mathbb{C}) may not vanish.

The global (1,1)(1,1) form β\beta on MM satisfies the following property on the end:

Proposition 4.3.

Let Φ:𝒞𝒦XK\Phi:\mathcal{C}\setminus\mathcal{K}\to X\setminus K be the fixed diffeomorphism and let p:𝒞Dp:\mathcal{C}\to D be the projection map. Then |Φβp(β|D)|ω𝒞=O(e(12ϵ)zn)\left|\Phi^{*}\beta-p^{*}(\left.\beta\right|_{D})\right|_{\omega_{\mathcal{C}}}=O(e^{-(\frac{1}{2}-\epsilon)z^{n}}).

Proof.

Let pp be a fixed point in DD. Let (w,z¯)=(w,z1,,zn)(w,\underline{z})=(w,z_{1},\cdots,z_{n}) be local holomorphic coordinates around this point such that DD is given by {w=0}\{w=0\}. Then (w,z¯)(w,\underline{z}) can also be seen as a group of local holomorphic coordinates around pp in 𝒞\mathcal{C} where ww represent the fiber direction. We can express β\beta locally around pp on MM as

β=\displaystyle\beta= i,j=1n1fij¯dzidz¯j+i=1n1fi0¯dzidw¯+i=1n1f0i¯dwdz¯i+f00dwdw¯,\displaystyle\sum_{i,j=1}^{n-1}f_{i\,\overline{\!{j}}}dz_{i}\wedge d\bar{z}_{j}+\sum_{i=1}^{n-1}f_{i\bar{0}}dz_{i}\wedge d\bar{w}+\sum_{i=1}^{n-1}f_{0\,\overline{\!{i}}}dw\wedge d\bar{z}_{i}+f_{00}dw\wedge d\bar{w},
Φβ=\displaystyle\Phi^{*}\beta= i,j=1n1fij¯(Φ)dzi(Φ)Φ(JX)dzj(Φ)+i=1n1fi0¯(Φ)dzi(Φ)Φ(JX)dw(Φ)\displaystyle\sum_{i,j=1}^{n-1}f_{i\,\overline{\!{j}}}(\Phi)dz_{i}(\Phi)\wedge\Phi^{*}(J_{X})dz_{j}(\Phi)+\sum_{i=1}^{n-1}f_{i\bar{0}}(\Phi)dz_{i}(\Phi)\wedge\Phi^{*}(J_{X})dw(\Phi)
+i=1n1f0i¯(Φ)dw(Φ)Φ(JX)dzi(Φ)+f00¯(Φ)dw(Φ)Φ(JX)dw(Φ).\displaystyle+\sum_{i=1}^{n-1}f_{0\,\overline{\!{i}}}(\Phi)dw(\Phi)\wedge\Phi^{*}(J_{X})dz_{i}(\Phi)+f_{0\bar{0}}(\Phi)dw(\Phi)\wedge\Phi^{*}(J_{X})dw(\Phi).

Notice that we have the estimate of |g𝒞kw|=O(e(12ϵ)zn)\left|\nabla^{k}_{g_{\mathcal{C}}}w\right|=O(e^{-(\frac{1}{2}-\epsilon)z^{n}}), |g𝒞kzi|=O(1)\left|\nabla^{k}_{g_{\mathcal{C}}}z_{i}\right|=O(1), Φ|D=Id\left.\Phi\right|_{D}=\mathrm{Id}, and the complex structure ΦJXJC\Phi^{*}J_{X}-J_{C} is exponentially decay as in Proposition 2.1, we know that on 𝒞\mathcal{C} we have

(Φβ)|D=\displaystyle\left.(\Phi^{*}\beta)\right|_{D}= i,j=1n1fij¯(0,z¯)dziJ𝒞dzj+O(e(12ϵ)zn).\displaystyle\sum_{i,j=1}^{n-1}f_{i\,\overline{\!{j}}}(0,\underline{z})dz_{i}\wedge J_{\mathcal{C}}dz_{j}+O(e^{-(\frac{1}{2}-\epsilon)z^{n}}).

On the other hand, we know that p(β|D)|D=i,j=1n1fij¯(0,zi)dzidz¯j\left.p^{*}(\left.\beta\right|_{D})\right|_{D}=\sum_{i,j=1}^{n-1}f_{i\,\overline{\!{j}}}(0,z_{i})dz_{i}\wedge d\bar{z}_{j}. So Φβp(β|D)\Phi^{*}\beta-p^{*}(\left.\beta\right|_{D}) extends to a smooth form on NDN_{D} vanishing on the zero section DD, which yields |Φβp(β|D)|ω𝒞=O(e(12ϵ)zn)\left|\Phi^{*}\beta-p^{*}(\left.\beta\right|_{D})\right|_{\omega_{\mathcal{C}}}=O(e^{-(\frac{1}{2}-\epsilon)z^{n}}). ∎

4.2. Iteration process

With this exponential closeness, we can view Φβ\Phi^{*}\beta as a (1,1)(1,1)-form on 𝒞\mathcal{C} with only horizontal direction component. This will greatly simplify our computation below.

Definition 4.4.

Let η\eta be a (1,1)(1,1)-form on 𝒞\mathcal{C}. We define F(η):=1(ω𝒞+η)nω𝒞nF(\eta):=1-\frac{(\omega_{\mathcal{C}}+\eta)^{n}}{\omega_{\mathcal{C}}^{n}}, called ω𝒞\omega_{\mathcal{C}}-potential.

Definition 4.5.

With the same β\beta as before, we define by iteration

F0:=F(p(β|D)),Fj:=F(p(β|D)+i¯Uj)=1(ω𝒞+p(β|D)+i¯Uj)nω𝒞n\displaystyle F_{0}:=F(p^{*}(\left.\beta\right|_{D})),\quad F_{j}:=F(p^{*}(\left.\beta\right|_{D})+i\partial\bar{\partial}U_{j})=1-\frac{(\omega_{\mathcal{C}}+p^{*}(\left.\beta\right|_{D})+i\partial\bar{\partial}U_{j})^{n}}{\omega_{\mathcal{C}}^{n}}

where

U0=0,Uj=Uj1+uj,Δω𝒞uj=Fj1.\displaystyle U_{0}=0,\quad U_{j}=U_{j-1}+u_{j},\quad\Delta_{\omega_{\mathcal{C}}}u_{j}=F_{j-1}.

Here uju_{j} is the solution constructed in Proposition 3.2. We will prove in Proposition 4.6 that the derivative of FjF_{j}’s satisfy the decay condition required for vv in Proposition 3.2 so this iteration process works.

Proposition 4.6.

With UjU_{j}’s and uju_{j}’s defined in 4.5, we have Fn=F(p(β|D)+i¯Un)F_{n}=F(p^{*}(\left.\beta\right|_{D})+i\partial\bar{\partial}U_{n}) decays faster than znz^{-n}. More precisely, for any positive integer jj and kk

zk2kFjω𝒞Ck,jzj1+ϵ.\|z^{\frac{k}{2}}\nabla^{k}F_{j}\|_{\omega_{\mathcal{C}}}\leqslant C_{k,j}z^{-j-1+\epsilon}. (4.1)
Proof.

We prove (4.1) by induction.

For F0F_{0}, we can see this estimate directly follows from computation:

F0=1(ω𝒞+p(β|D))nω𝒞n=j=1nnjnzjp(β|DjωDnj1ωDn1).\displaystyle F_{0}=1-\frac{(\omega_{\mathcal{C}}+p^{*}(\left.\beta\right|_{D}))^{n}}{\omega_{\mathcal{C}}^{n}}=\sum_{j=1}^{n}\frac{n-j}{nz^{j}}\cdot p^{*}(\frac{\left.\beta\right|_{D}^{j}\wedge\omega_{D}^{n-j-1}}{\omega_{D}^{n-1}}).

By (3.6) we know that |zk2kF0|ω𝒞C(k)z1|z^{\frac{k}{2}}\nabla^{k}F_{0}|_{\omega_{\mathcal{C}}}\leqslant C(k)z^{-1}, for any positive integer kk. So when j=0j=0 (4.1) holds.

Assume (4.1) holds for iji\leqslant j, i.e. |zk2kFi|ω𝒞C(k,i)zi1|z^{\frac{k}{2}}\nabla^{k}F_{i}|_{\omega_{\mathcal{C}}}\leqslant C(k,i)z^{-i-1}, for any k0k\geqslant 0. By straightforward computation,

Fj+1\displaystyle F_{j+1} =ni¯uj+1k=1n1(n1k)(p(β|D)+i¯Uj)kω𝒞nk1ω𝒞n\displaystyle=-\frac{ni\partial\bar{\partial}u_{j+1}\wedge\sum_{k=1}^{n-1}{n-1\choose k}(p^{*}(\left.\beta\right|_{D})+i\partial\bar{\partial}U_{j})^{k}\wedge\omega_{\mathcal{C}}^{n-k-1}}{\omega_{\mathcal{C}}^{n}}
k=2n(nk)(i¯uj+1)k(ω𝒞+p(β|D)+i¯Uj)nkω𝒞n.\displaystyle\qquad\qquad\qquad\qquad-\frac{\sum_{k=2}^{n}{n\choose k}\left(i\partial\bar{\partial}u_{j+1}\right)^{k}\wedge\left(\omega_{\mathcal{C}}+p^{*}(\left.\beta\right|_{D})+i\partial\bar{\partial}U_{j}\right)^{n-k}}{\omega_{\mathcal{C}}^{n}}.

Actually, the function in each term is of the from

Ci¯uj+1qQi¯uqp(β|D)mω𝒞nm|Q|1ω𝒞n,\displaystyle\frac{Ci\partial\bar{\partial}u_{j+1}\wedge\bigwedge_{q\in Q}i\partial\bar{\partial}u_{q}\wedge p^{*}(\left.\beta\right|_{D})^{m}\wedge\omega_{\mathcal{C}}^{n-m-|Q|-1}}{\omega_{\mathcal{C}}^{n}}, (4.2)

where QQ is a set with repeated elements from {1,2,,j+1}\{1,2,\cdots,j+1\}, mm is a non-negative integer and positive when j+1Qj+1\notin Q. By Proposition 3.2 we know that |zk2kuj+1|ω𝒞Ck,jzj+ϵ\left|z^{\frac{k}{2}}\nabla^{k}u_{j+1}\right|_{\omega_{\mathcal{C}}}\leqslant C_{k,j}z^{-j+\epsilon}. It is easier to deduce the bound for (4.2) by passing to the rescaled metric ω~𝒞\tilde{\omega}_{\mathcal{C}} on Bω~𝒞(x,1)B_{\tilde{\omega}_{\mathcal{C}}}(x,1), where z(x)=z0z(x)=z_{0}, ω~𝒞=ω𝒞z0\tilde{\omega}_{\mathcal{C}}=\frac{\omega_{\mathcal{C}}}{z_{0}}. We have the uniform weighted bound for each term in the wedge product

ki¯uqC0(Bω~𝒞(x,1))Ck,qz0q+1+ϵ,ki¯uj+1C0(Bω~𝒞(x,1))Ck,jz0j+ϵ,\displaystyle\left\|\nabla^{k}i\partial\bar{\partial}u_{q}\right\|_{C^{0}(B_{\tilde{\omega}_{\mathcal{C}}}(x,1))}\leqslant C_{k,q}z_{0}^{-q+1+\epsilon},\left\|\nabla^{k}i\partial\bar{\partial}u_{j+1}\right\|_{C^{0}(B_{\tilde{\omega}_{\mathcal{C}}}(x,1))}\leqslant C_{k,j}z_{0}^{-j+\epsilon},
kp(β|D)C0(Bω~𝒞(x,1))Ck.\displaystyle\left\|\nabla^{k}p^{*}(\left.\beta\right|_{D})\right\|_{C^{0}(B_{\tilde{\omega}_{\mathcal{C}}}(x,1))}\leqslant C_{k}.

If we consider the scaled metric ω~𝒞\tilde{\omega}_{\mathcal{C}} our function (4.2) becomes

Ci¯uj+1qQi¯uqp(β|D)mω~𝒞nm|Q|1z0m+|Q|+1ω~𝒞n.\displaystyle\frac{Ci\partial\bar{\partial}u_{j+1}\wedge\bigwedge_{q\in Q}i\partial\bar{\partial}u_{q}\wedge p^{*}(\left.\beta\right|_{D})^{m}\wedge\tilde{\omega}_{\mathcal{C}}^{n-m-|Q|-1}}{z_{0}^{m+|Q|+1}\tilde{\omega}_{\mathcal{C}}^{n}}.

So we have

kCi¯uj+1qQi¯uqp(β|D)mω~𝒞nm|Q|1z0m+|Q|+1ω~𝒞nC0(Bω~𝒞(x,1))Ck,j,Q,mz0j2+ϵ.\displaystyle\left\|\nabla^{k}\frac{Ci\partial\bar{\partial}u_{j+1}\wedge\bigwedge_{q\in Q}i\partial\bar{\partial}u_{q}\wedge p^{*}(\left.\beta\right|_{D})^{m}\wedge\tilde{\omega}_{\mathcal{C}}^{n-m-|Q|-1}}{z_{0}^{m+|Q|+1}\tilde{\omega}_{\mathcal{C}}^{n}}\right\|_{C^{0}(B_{\tilde{\omega}_{\mathcal{C}}}(x,1))}\leqslant C_{k,j,Q,m}z_{0}^{-j-2+\epsilon}.

Consequently,

z0k2kFj+1C0(Bω𝒞(x,z0))=kFj+1C0(Bω~𝒞(x,1))Ck,jz0j2+ϵ.\displaystyle\left\|z_{0}^{\frac{k}{2}}\nabla^{k}F_{j+1}\right\|_{C^{0}(B_{\omega_{\mathcal{C}}}(x,\sqrt{z_{0}}))}=\left\|\nabla^{k}F_{j+1}\right\|_{C^{0}(B_{\tilde{\omega}_{\mathcal{C}}}(x,1))}\leqslant C_{k,j}z_{0}^{-j-2+\epsilon}.

So we finish the proof of (4.1). Specially,

|zk2kFn|ω𝒞Ckzn1+ϵ.\displaystyle\left|z^{\frac{k}{2}}\nabla^{k}F_{n}\right|_{\omega_{\mathcal{C}}}\leqslant C_{k}z^{-n-1+\epsilon}.

5. The Integral Condition

For the convenience of statement, let us introduce the following notation.

Definition 5.1.

For an (n,0)(n,0) form Ω\Omega, we say that a (1,1)-form α\alpha is Ω\Omega-compatible if XΩΩ¯αn=0\int_{X}\Omega\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu-\alpha^{n}=0.

Remark 5.2.

Since ΩΩ¯\Omega\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu and αn\alpha^{n} are not integrable for most of the time, this integration identity means that the function f=αnΩΩ¯1f=\frac{\alpha^{n}}{\Omega\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu}-1 satisfies XfΩΩ¯=0\int_{X}f\Omega\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu=0.

In this section, we will show that by adding a suitable potential we can make β+i¯U\beta+i\partial\bar{\partial}U to be ΩX\Omega_{X}-compatible.

Proposition 5.3.

There exists a smooth function U~\tilde{U} on XX such that β+i¯U~\beta+i\partial\bar{\partial}\tilde{U} is an ΩX\Omega_{X}-compatible Kähler form. Meanwhile, we have that

|ΦU~(zn+1+Un+λz)|Ceδzn\displaystyle\left|\Phi^{*}\tilde{U}-(z^{n+1}+U_{n}+\lambda z)\right|\leqslant Ce^{-\delta z^{n}}

when z>Cz>C for some constant C>0C>0 and λ\lambda\in\mathbb{R}.

Proof.

We first show that we can find UU such that (β+i¯U+ωX)n(\beta+i\partial\bar{\partial}U+\omega_{X})^{n} is integrable on the end. Recall that we have Uj=p=1jupU_{j}=\sum_{p=1}^{j}u_{p} and uju_{j} are functions on 𝒞\mathcal{C} such that

Δω𝒞uj+1=1(p(β|D)+ω𝒞+i¯Uj)nω𝒞n=Fj.\Delta_{\omega_{\mathcal{C}}}u_{j+1}=1-\frac{\left(p^{*}(\left.\beta\right|_{D})+\omega_{\mathcal{C}}+i\partial\bar{\partial}U_{j}\right)^{n}}{\omega_{\mathcal{C}}^{n}}=F_{j}.

We know that the following integration is finite since |Fn|Cnzn1+ϵ|F_{n}|\leqslant C_{n}z^{-n-1+\epsilon}:

|𝒞𝒦(Φβ+ω𝒞+i¯Un)nω𝒞n|\displaystyle\quad\left|\int_{\mathcal{C}\setminus\mathcal{K}}(\Phi^{*}\beta+\omega_{\mathcal{C}}+i\partial\bar{\partial}U_{n})^{n}-\omega_{\mathcal{C}}^{n}\right|
𝒞𝒦|Fn|(i¯t)n1dtdct+|𝒞𝒦(Φβ+ω𝒞+i¯Un)n(p(β|D)+ω𝒞+i¯Un)n|\displaystyle\leqslant\int_{\mathcal{C}\setminus\mathcal{K}}|F_{n}|(i\partial\bar{\partial}t)^{n-1}dt\wedge d^{c}t+\left|\int_{\mathcal{C}\setminus\mathcal{K}}(\Phi^{*}\beta+\omega_{\mathcal{C}}+i\partial\bar{\partial}U_{n})^{n}-(p^{*}(\left.\beta\right|_{D})+\omega_{\mathcal{C}}+i\partial\bar{\partial}U_{n})^{n}\right|
={t=T}nT1+ϵn(i¯t)n1dct+C<+.\displaystyle=\int_{\{t=T\}}nT^{\frac{-1+\epsilon}{n}}(i\partial\bar{\partial}t)^{n-1}\wedge d^{c}t+C<+\infty.

Since we have the exponentially closed estimate between XKX\setminus K and 𝒞𝒦\mathcal{C}\setminus\mathcal{K}, we will have

XK(β+i¯(Φ1)Un+ωX)nωXn=𝒞𝒦(Φβ+i¯Un+ω𝒞)nω𝒞n+O(eδzn)<+.\int_{X\setminus K}\left(\beta+i\partial\bar{\partial}(\Phi^{-1})^{*}U_{n}+\omega_{X}\right)^{n}-\omega_{X}^{n}=\int_{\mathcal{C}\setminus\mathcal{K}}(\Phi^{*}\beta+i\partial\bar{\partial}U_{n}+\omega_{\mathcal{C}})^{n}-\omega_{\mathcal{C}}^{n}+O(e^{-\delta z^{n}})<+\infty.

Now we can construct the Kähler potential following the construction in Hein-Sun-Viaclovsky-Zhang [7, Lemma 2.7.]:

The ampleness of NDN_{D} implies that XX is 1-convex. Hence by Remmert reduction we know that KM-K_{M} is semi-ample, we denote its non-ample locus by EE. Recall that [β]pY>0[\beta]^{p}\cdot Y>0 for any pp-dimensional compact subvariety YY in XX, by the generalized Demailly-Păun criterion in [4] we know that there exists a smooth function u0u_{0} on XX such that β+i¯u0\beta+i\partial\bar{\partial}u_{0} is positive on the neighborhood UU of EE. Let χ0\chi_{0} be a smooth function on XX support on UU and χ0=1\chi_{0}=1 on EE. Then β+i¯(χ0u0)\beta+i\partial\bar{\partial}\left(\chi_{0}\cdot u_{0}\right) is positive around EE and i¯(χ0u0)i\partial\bar{\partial}\left(\chi_{0}\cdot u_{0}\right) is supported on UU.

Let 𝔱=log|S|hM2\mathfrak{t}=-\log|S|^{2}_{h_{M}}, then the curvature form i¯𝔱0i\partial\bar{\partial}\mathfrak{t}\geqslant 0 and i¯𝔱>0i\partial\bar{\partial}\mathfrak{t}>0 on XEX\setminus E. Let χ1\chi_{1} be a smooth cutoff function on [0,+)[0,+\infty) such that χ1=1\chi_{1}=1 on [0,1][0,1] and χ1=0\chi_{1}=0 on [2,+)[2,+\infty). Then i¯(χ1(𝔱C2)𝔱)i\partial\bar{\partial}(\chi_{1}(\tfrac{\mathfrak{t}}{C_{2}})\cdot\mathfrak{t}) is positive on {𝔱C2}\{\mathfrak{t}\leqslant C_{2}\} and supported on {𝔱2C2}\{\mathfrak{t}\leqslant 2C_{2}\}.

Let ρA\rho_{A} be a smooth convex function on \mathbb{R} with ρA(x)=2A3\rho_{A}(x)=\tfrac{2A}{3} on (,A2](-\infty,\tfrac{A}{2}] and ρA(x)=x\rho_{A}(x)=x on (A,+)(A,+\infty). Then we can obtain that i¯(ρC3(𝔱n+1n))0i\partial\bar{\partial}(\rho_{C_{3}}(\mathfrak{t}^{\tfrac{n+1}{n}}))\geqslant 0 on XX and i¯(ρC3(nn+1𝔱n+1n))=i¯(nn+1𝔱n+1n)i\partial\bar{\partial}(\rho_{C_{3}}(\tfrac{n}{n+1}\mathfrak{t}^{\tfrac{n+1}{n}}))=i\partial\bar{\partial}(\tfrac{n}{n+1}\mathfrak{t}^{\tfrac{n+1}{n}}) on {𝔱2C3nn+1}\{\mathfrak{t}\geqslant 2C_{3}^{\tfrac{n}{n+1}}\}.

Let

β1=β+i¯(χ0u0+C1χ1(𝔱C2)𝔱+ρC3(nn+1𝔱n+1n)).\displaystyle\beta_{1}=\beta+i\partial\bar{\partial}\left(\chi_{0}\cdot u_{0}+C_{1}\chi_{1}\left(\frac{\mathfrak{t}}{C_{2}}\right)\cdot\mathfrak{t}+\rho_{C_{3}}\left(\tfrac{n}{n+1}\mathfrak{t}^{\frac{n+1}{n}}\right)\right).

By our choice of u0u_{0}, β+i¯u0\beta+i\partial\bar{\partial}u_{0} is positive around EE. By choosing C1C_{1} and C2C_{2} large we can make β1\beta_{1} is positive on UU. Then choosing C2C_{2} large enough depending on C1C_{1} and C3C_{3} we have that β1\beta_{1} is positive on {C2𝔱2C2}\{C_{2}\leqslant\mathfrak{t}\leqslant 2C_{2}\} hence Kähler on XX.

Then we can glue our perturbation function UnU_{n} via a cut-off function χ2\chi_{2} supported outside a compact set KK^{\prime} with χ2=1\chi_{2}=1 outside a open neighborhood UU of KK^{\prime} and let

β2(λ)=β1+i¯(χ2((Φ1)(Un+λz))).\beta_{2}(\lambda)=\beta_{1}+i\partial\bar{\partial}\left(\chi_{2}\cdot\left((\Phi^{-1})^{*}(U_{n}+\lambda z)\right)\right).

Our goal next is to find suitable λ\lambda and KK^{\prime} such that β2(λ)\beta_{2}(\lambda) is Kähler and ΩX\Omega_{X}-compatible.

Let us first show that the ΩX\Omega_{X}-compatible condition is a linear equation of λ\lambda and only the constant term depends on the choice of χ2\chi_{2}. By our previous estimate of UnU_{n} we know as a starting point that Xβ2(0)nΩXΩ¯X=C\int_{X}\beta_{2}(0)^{n}-\Omega_{X}\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu_{X}=C is finite. The ΩX\Omega_{X}-compatible condition becomes

0=\displaystyle 0= Xβ2(λ)nΩXΩ¯X=Xβ2(λ)nβ2(0)n+Xβ2(0)nΩXΩ¯X\displaystyle\int_{X}\beta_{2}(\lambda)^{n}-\Omega_{X}\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu_{X}=\int_{X}\beta_{2}(\lambda)^{n}-\beta_{2}(0)^{n}+\int_{X}\beta_{2}(0)^{n}-\Omega_{X}\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu_{X}
=limε0{(Φ1)tlogε}λi¯(χ2(Φ1)z)k=0n1(nk)β2(0)ki¯(λχ2(Φ1)z)nk1+C\displaystyle=\lim_{\varepsilon\to 0}\int_{\{(\Phi^{-1})^{*}t\leqslant-\log\varepsilon\}}\lambda i\partial\bar{\partial}\left(\chi_{2}\cdot(\Phi^{-1})^{*}z\right)\wedge\sum_{k=0}^{n-1}{n\choose k}\beta_{2}(0)^{k}\wedge i\partial\bar{\partial}\left(\lambda\chi_{2}\cdot(\Phi^{-1})^{*}z\right)^{n-k-1}+C
=limε0{t=logε}λdct1nzn1k=0n1(nk)(ω𝒞+β+i¯Un)k(λi¯z)nk1+C.\displaystyle=\lim_{\varepsilon\to 0}\int_{\{t=-\log\varepsilon\}}\lambda d^{c}t\wedge\frac{1}{nz^{n-1}}\sum_{k=0}^{n-1}{n\choose k}(\omega_{\mathcal{C}}+\beta+i\partial\bar{\partial}U_{n})^{k}\wedge(\lambda i\partial\bar{\partial}z)^{n-k-1}+C.

Expanding the terms in the bracket, we notice that only ω𝒞n1\omega_{\mathcal{C}}^{n-1} remains non-vanishing after we take the limit, so we have the equation

0\displaystyle 0 =limε0{t=logε}λdct(i¯t)n1+C=λDωDn1+C.\displaystyle=\lim_{\varepsilon\to 0}\int_{\{t=-\log\varepsilon\}}\lambda d^{c}t\wedge(i\partial\bar{\partial}t)^{n-1}+C=\lambda\cdot\int_{D}\omega_{D}^{n-1}+C.

This is a linear equation on λ\lambda. On the other hand, we also notice by the previous computation that χ2\chi_{2} does not affect the integral, so we can choose λ0\lambda_{0} first to satisfy the integral condition and then choose KK^{\prime} large enough such that β2(λ0)\beta_{2}(\lambda_{0}) is Kähler. So by choosing

U~=χ0u0+C1χ1(𝔱C2)𝔱+ρC3(nn+1𝔱n+1n)+χ2((Φ1)(Un+λz))\tilde{U}=\chi_{0}\cdot u_{0}+C_{1}\chi_{1}\left(\tfrac{\mathfrak{t}}{C_{2}}\right)\cdot\mathfrak{t}+\rho_{C_{3}}\left(\tfrac{n}{n+1}\mathfrak{t}^{\frac{n+1}{n}}\right)+\chi_{2}\cdot\left(\left(\Phi^{-1}\right)^{*}(U_{n}+\lambda z)\right)

we finish our proof. ∎

In order to apply Tian-Yau-Hein’s package, we need to repeat the iteration process for one more step such that the ω𝒞\omega_{\mathcal{C}}-potential of β+i¯U\beta+i\partial\bar{\partial}U decays faster than r2r^{-2}.

Proposition 5.4.

Furthermore, we can construct a Kähler ΩX\Omega_{X}-compatible form β+i¯U\beta+i\partial\bar{\partial}U on XX such that

|1(β+i¯U)nΩXΩ¯X|Cr2ϵ,|ω𝒞Φ(β+i¯U)|ω𝒞Cz1,\displaystyle\left|1-\frac{(\beta+i\partial\bar{\partial}U)^{n}}{\Omega_{X}\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu_{X}}\right|\leqslant Cr^{-2-\epsilon},\qquad\left|\,\omega_{\mathcal{C}}-\Phi^{*}(\beta+i\partial\bar{\partial}U)\right|_{\omega_{\mathcal{C}}}\leqslant Cz^{-1},

where rr is the distance function to some point pXp\in X under metric β+i¯U\beta+i\partial\bar{\partial}U, C>0C>0.

Proof.

Let un+1=λzu_{n+1}=\lambda z, Un+1=Un+un+1U_{n+1}=U_{n}+u_{n+1}. Let un+2u_{n+2} be the solution of Δω𝒞un+2=Fn+1\Delta_{\omega_{\mathcal{C}}}u_{n+2}=F_{n+1} constructed in Proposition 3.2.

Fn+1\displaystyle F_{n+1} =Fnni¯λzk=1n1(n1k)(p(β|D)+i¯Un)kω𝒞nk1ω𝒞n\displaystyle=F_{n}-\frac{ni\partial\bar{\partial}\lambda z\wedge\sum_{k=1}^{n-1}{n-1\choose k}(p^{*}(\left.\beta\right|_{D})+i\partial\bar{\partial}U_{n})^{k}\wedge\omega_{\mathcal{C}}^{n-k-1}}{\omega_{\mathcal{C}}^{n}}
k=2n(nk)(i¯λz)k(ω𝒞+p(β|D)+i¯Un)nkω𝒞n.\displaystyle\qquad\qquad\qquad\qquad-\frac{\sum_{k=2}^{n}{n\choose k}\left(i\partial\bar{\partial}\lambda z\right)^{k}\wedge\left(\omega_{\mathcal{C}}+p^{*}(\left.\beta\right|_{D})+i\partial\bar{\partial}U_{n}\right)^{n-k}}{\omega_{\mathcal{C}}^{n}}.

By Proposition 3.2, we know that |zk2kFn+1(z,)|ω𝒞CKzn1+ϵ|z^{\frac{k}{2}}\nabla^{k}F_{n+1}(z,\cdot)|_{\omega_{\mathcal{C}}}\leqslant C_{K}z^{-n-1+\epsilon}, which is of the same order of FnF_{n}. Then

Fn+2\displaystyle F_{n+2} =ni¯un+2k=1n1(n1k)(p(β|D)+i¯Un+1)kω𝒞nk1ω𝒞n\displaystyle=-\frac{ni\partial\bar{\partial}u_{n+2}\wedge\sum_{k=1}^{n-1}{n-1\choose k}(p^{*}(\beta|_{D})+i\partial\bar{\partial}U_{n+1})^{k}\wedge\omega_{\mathcal{C}}^{n-k-1}}{\omega_{\mathcal{C}}^{n}}
k=2n(nk)(i¯un+2)k(ω𝒞+p(β|D)+i¯Un+1)nkω𝒞n.\displaystyle\qquad\qquad\qquad-\frac{\sum_{k=2}^{n}{n\choose k}\left(i\partial\bar{\partial}u_{n+2}\right)^{k}\wedge\left(\omega_{\mathcal{C}}+p^{*}(\beta|_{D})+i\partial\bar{\partial}U_{n+1}\right)^{n-k}}{\omega_{\mathcal{C}}^{n}}.

We know from the estimate in Proposition 3.2 that |zk2kun+2|Czn+ϵ\left|z^{\frac{k}{2}}\nabla^{k}u_{n+2}\right|\leqslant Cz^{-n+\epsilon}. So Fn+2Czn2+ϵF_{n+2}\leqslant Cz^{-n-2+\epsilon}.

Let U=U~+χ2(Φ1)un+2U=\tilde{U}+\chi_{2}(\Phi^{-1})^{*}u_{n+2}, we can choose KK^{\prime} large enough such that β+i¯U\beta+i\partial\bar{\partial}U is Kähler on XX. Also we see from the construction of UU that

|ω𝒞Φ(β+i¯U)|ω𝒞=|Φ(β)|ω𝒞+k=1n+2|i¯uk|ω𝒞Cz1.\displaystyle\left|\,\omega_{\mathcal{C}}-\Phi^{*}(\beta+i\partial\bar{\partial}U)\right|_{\omega_{\mathcal{C}}}=\left|\Phi^{*}(\beta)\right|_{\omega_{\mathcal{C}}}+\sum_{k=1}^{n+2}\left|i\partial\bar{\partial}u_{k}\right|_{\omega_{\mathcal{C}}}\leqslant Cz^{-1}. (5.1)

Fix a point pXp\in X. Let r(x)r(x) denote the distance function to pp with the metric β+i¯U\beta+i\partial\bar{\partial}U. With this asymptotic behavior, we know that r(x)r(x) is in the same order of the distance function on 𝒞\mathcal{C}. Then outside a compact set on XX we have the estimate that

|1(β+i¯U)nΩXΩ¯X|=(Φ1)(|1(p(β|D)+ω𝒞+i¯Un+2)nω𝒞n+O(eδzn)|)\displaystyle\left|1-\frac{(\beta+i\partial\bar{\partial}U)^{n}}{\Omega_{X}\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu_{X}}\right|=(\Phi^{-1})^{*}\left(\left|1-\frac{(p^{*}(\left.\beta\right|_{D})+\omega_{\mathcal{C}}+i\partial\bar{\partial}U_{n+2})^{n}}{\omega_{\mathcal{C}}^{n}}+O(e^{-\delta z^{n}})\right|\right)
=\displaystyle= (Φ1)(|Fn+2|+O(eδzn))C(Φ1)(zn2+ϵ)Cr2ϵ.\displaystyle(\Phi^{-1})^{*}\left(|F_{n+2}|+O(e^{-\delta z^{n}})\right)\leqslant C(\Phi^{-1})^{*}\left(z^{-n-2+\epsilon}\right)\leqslant Cr^{-2-\epsilon}.

For the ΩX\Omega_{X}-compatible condition, we notice that the small term un+2u_{n+2} does not affect the integration of the form β+i¯U\beta+i\partial\bar{\partial}U:

X(β+i¯U)nΩXΩ¯X=X(β+i¯U)n(β+i¯U~)n\displaystyle\int_{X}(\beta+i\partial\bar{\partial}U)^{n}-\Omega_{X}\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu_{X}=\int_{X}(\beta+i\partial\bar{\partial}U)^{n}-(\beta+i\partial\bar{\partial}\tilde{U})^{n}
=limε0{(Φ1)tlogε}i¯(ρA3((Φ1)un+2))(k=0n1(nk)(β+i¯U~)ki¯(ρA2((Φ1)un+2))nk1)\displaystyle=\lim_{\varepsilon\to 0}\int_{\{(\Phi^{-1})^{*}t\leqslant-\log\varepsilon\}}i\partial\bar{\partial}(\rho_{A_{3}}((\Phi^{-1})^{*}u_{n+2}))\wedge\left(\sum_{k=0}^{n-1}{n\choose k}(\beta+i\partial\bar{\partial}\tilde{U})^{k}\wedge i\partial\bar{\partial}\left(\rho_{A_{2}}((\Phi^{-1})^{*}u_{n+2})\right)^{n-k-1}\right)
=limε0{t=logε}dcun+21nzn1(k=0n1(nk)(β+ω𝒞+i¯Un+1)k(i¯un+2)nk1)=0.\displaystyle=\lim_{\varepsilon\to 0}\int_{\{t=-\log\varepsilon\}}d^{c}u_{n+2}\wedge\frac{1}{nz^{n-1}}\left(\sum_{k=0}^{n-1}{n\choose k}(\beta+\omega_{\mathcal{C}}+i\partial\bar{\partial}U_{n+1})^{k}\wedge(i\partial\bar{\partial}u_{n+2})^{n-k-1}\right)=0.

So β+i¯U\beta+i\partial\bar{\partial}U is a Kähler form satisfying both ΩX\Omega_{X}-compatible condition and decay condition in Tian-Yau-Hein’s package. ∎

6. Existence and the proof

Now we are ready to apply Tian-Yau-Hein’s package to deform our metric β+i¯U\beta+i\partial\bar{\partial}U to a Calabi-Yau metric.

Theorem 6.1.

For any class 𝔨\mathfrak{k} in H+2(X)H_{+}^{2}(X), there exists a Calabi-Yau metric ω\omega in the class 𝔨\mathfrak{k}.

Proof.

Let β\beta be the good representative we chose in UU be the potential constructed in section 5 with the form β\beta. We know that β+i¯U\beta+i\partial\bar{\partial}U is a Kähler metric on XX such that

F(β+i¯U):=1(β+i¯U)nΩXΩ¯X\displaystyle F(\beta+i\partial\bar{\partial}U):=1-\frac{(\beta+i\partial\bar{\partial}U)^{n}}{\Omega_{X}\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu_{X}}

decays in order r2ϵr^{-2-\epsilon} and X(β+i¯U)n\int_{X}(\beta+i\partial\bar{\partial}U)^{n}, here rr is any distance function under the metric β+i¯U\beta+i\partial\bar{\partial}U. Let

f=log(ΩXΩ¯X(β+i¯U)n).\displaystyle f=\log{\frac{\Omega_{X}\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu_{X}}{(\beta+i\partial\bar{\partial}U)^{n}}}.

We have ff satisfies integral condition X(ef1)(β+i¯U)n=0\int_{X}(e^{f}-1)(\beta+i\partial\bar{\partial}U)^{n}=0 and the decay condition |f|Cr2ϵ\left|f\right|\leqslant Cr^{-2-\epsilon}.

On the other hand, we have higher regularity estimate of uiu_{i}’s:

|zk2k(Φ(β+i¯U)ω𝒞)|ω𝒞=|zk2kΦ(β)+j=1n+2zk2ki¯uj|ω𝒞Cz1\displaystyle|z^{\frac{k}{2}}\nabla^{k}(\Phi^{*}(\beta+i\partial\bar{\partial}U)-\omega_{\mathcal{C}})|_{\omega_{\mathcal{C}}}=|z^{\frac{k}{2}}\nabla^{k}\Phi^{*}(\beta)+\sum_{j=1}^{n+2}z^{\frac{k}{2}}\nabla^{k}i\partial\bar{\partial}u_{j}|_{\omega_{\mathcal{C}}}\leqslant Cz^{-1}

for any z>Cz>C with some C>0C>0. Then we have higher estimate of metric and scalar curvature. So (X,β+i¯U)(X,\beta+i\partial\bar{\partial}U) satisfies the SOB(2nn+1)\mathrm{SOB}(\tfrac{2n}{n+1}) condition by Lemma 2.4 and HMG(1n+1,k,α)\mathrm{HMG}(\tfrac{1}{n+1},k,\alpha) by Lemma 2.6 for any k>0k>0 and 0<α<10<\alpha<1.

So we know that there exists a function ϕ\phi on XX such that

(β+i¯U+i¯ϕ)n=ef(β+i¯U)n=ΩXΩ¯X,\displaystyle(\beta+i\partial\bar{\partial}U+i\partial\bar{\partial}\phi)^{n}=e^{f}(\beta+i\partial\bar{\partial}U)^{n}=\Omega_{X}\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu_{X}, (6.1)

with ϕC4(X)\phi\in C^{4}(X). ∎

The iteration process shows that for any K>0K>0, there exists function UKU_{K} and constant CKC_{K} such that

|fK|:=|log(ΩXΩ¯X(β+i¯UK)n)|CKrK.\displaystyle\left|f_{K}\right|:=|\,\log{\frac{\Omega_{X}\wedge\mskip 1.5mu\overline{\mskip-1.5mu{\Omega}\mskip-1.5mu}\mskip 1.5mu_{X}}{(\beta+i\partial\bar{\partial}U_{K})^{n}}}|\leqslant C_{K}r^{-K}.

If we choose UU such that the ω𝒞\omega_{\mathcal{C}} potential F(β+i¯U)F(\beta+i\partial\bar{\partial}U) decays fast enough, we can show that the solution ϕ\phi of (6.1) also decays fast to a constant. To do this, we first present the following local Poincaré lemma for SOB(ν)\mathrm{SOB}(\nu) manifold with ν(0,2]\nu\in(0,2]:

Lemma 6.2.

Assume (M,ω0)(M,\omega_{0}) is a complete Kähler manifold satisfying SOB(ν)SOB(\nu) condition with ν(0,2]\nu\in(0,2], and r(x)κ|B(x,1)|Cr(x)^{\kappa}|B(x,1)|\leqslant C as r(x)r(x)\to\infty for some fixed κ>0\kappa>0, C>0C>0. Let u,fC(M)u,f\in C^{\infty}(M) such that sup|iu|+sup|if|<\sup|\nabla^{i}u|+\sup|\nabla^{i}f|<\infty for all i0i\in\mathbb{N}_{0}, and (ω0+i¯u)m=efω0m(\omega_{0}+i\partial\bar{\partial}u)^{m}=e^{f}\omega_{0}^{m}. Then for any δ>0\delta>0, there exists Kδ>0K_{\delta}>0 such that if X|u|2ωn<\int_{X}|\nabla u|^{2}\omega^{n}<\infty and |f|CrKδ|f|\leqslant Cr^{-K_{\delta}}, then

supB(x,1)|uuB(x,1)|Cr(x)δ\sup_{B(x,1)}|u-u_{B(x,1)}|\leqslant Cr(x)^{-\delta}

for any xXx\in X.

Remark 6.3.

The proof is entirely same as the proof in [8, Proposition 4.8(ib)]. The only difference is that we choose rir_{i} to be ii. So we omit the proof here.

Then we can improve the C0C^{0} bound of our solution ϕ\phi to get the optimal close rate of our weak asymptotically Calabi metric:

Theorem 6.4.

For any class 𝔨\mathfrak{k} in H+2(X)H_{+}^{2}(X), there is a Calabi-Yau metric ω\omega in 𝔨\mathfrak{k} which is weak asymptotically Calabi with rate 11.

Proof.

With Lemma 6.2, together with [8, Proposition 4.8(ii)], we know that if we choose KK large enough, there exists a constant ϕ¯\bar{\phi} and CC such that the solution ϕ\phi satisfies that

|ϕϕ¯|Crδ+nn+1, for any x such that r(x)>C.\displaystyle|\phi-\bar{\phi}|\leqslant Cr^{-\delta+\frac{n}{n+1}},\text{ for any }x\text{ such that }r(x)>C.

Then we can replace ϕ\phi by ϕϕ¯\phi-\bar{\phi} to get a better candidate for the solution of (6.1), so ϕ\phi could be chosen to decay at any polynomial rate. Repeat our local rescaling and local Schauder estimate, we know that the Calabi-Yau metric β+i¯UK+i¯ϕ\beta+i\partial\bar{\partial}U_{K}+i\partial\bar{\partial}\phi is polynomially closed to the Calabi model space with the leading error term β+i¯UK\beta+i\partial\bar{\partial}U_{K}.

If β|D=0\beta|_{D}=0, the error term is exponentially close to Calabi model space. If β|D\beta|_{D} is nonzero, the decay rate of β\beta is exactly r2n+1r^{-\frac{2}{n+1}}. If we choose β\beta such that β|D\beta|_{D} is primitive with respect to ωD\omega_{D}, the decay of i¯UKi\partial\bar{\partial}U_{K} would be r4n+1+ϵr^{-\frac{4}{n+1}+\epsilon}, which is strictly lower order term compared with β\beta. Thus, the Calabi-Yau metric β+i¯(U+ϕ)\beta+i\partial\bar{\partial}(U+\phi) decays exactly at the rate r2n+1r^{-\frac{2}{n+1}}, which is equivalent to z1z^{-1}. ∎

7. Uniqueness

In this section, we prove that the Calabi-Yau metric asymptotic to ω𝒞\omega_{\mathcal{C}} in the class 𝔨\mathfrak{k} is unique.

Theorem 7.1.

Let (M,D)(M,D) be the pair we considered before. If we have another Calabi-Yau metric ω~\tilde{\omega} in the same class 𝔨\mathfrak{k} satisfying |ω~ω|ωrκ\left|\,\tilde{\omega}-\omega\right|_{\omega}\leqslant r^{-\kappa}, when rr\to\infty, for some distance function rr with respect to ω\omega and some κ>0\kappa>0, then ω~=ω\tilde{\omega}=\omega.

Remark 7.2.

We are also interested in the problem that how different choice of the diffeomorphism Φ\Phi will change our Calabi-Yau metric. For example, the scaling in the fiber direction will change the metric by the rate r2nn+1r^{-\frac{2n}{n+1}} and by our uniqueness theorem, we get the same Calabi-Yau metric.

The proof of the theorem can be sketched as follows. We start with a ¯\partial\bar{\partial}-lemma by solving ¯\bar{\partial} equation via the L2L^{2} method. Then we can write ω~=ω+i¯l\tilde{\omega}=\omega+i\partial\bar{\partial}l with some estimate on ll. By pulling back to 𝒞\mathcal{C}, we construct ff on the model space to solve the Poisson equation Δω𝒞f=Δω𝒞l\Delta_{\mathcal{\omega_{\mathcal{C}}}}f=\Delta_{\mathcal{\omega_{\mathcal{C}}}}l. Via the estimate of harmonic function on 𝒞\mathcal{C} in Sun-Zhang [16], we can use the equation (ω+i¯l)n=ωn(\omega+i\partial\bar{\partial}l)^{n}=\omega^{n} and take integration by parts to deduce that i¯l=0i\partial\bar{\partial}l=0.

Lemma 7.3.

There exists a smooth function ll on XX such that ω~=ω+i¯l\tilde{\omega}=\omega+i\partial\bar{\partial}l with |Φl|<Ceϵt|\Phi^{*}l|<Ce^{\epsilon t} on {tC}\{t\geqslant C\} for any ϵ>0\epsilon>0 and some C>0C>0.

Proof.

We prove the lemma by several steps:

Step 1: We show that there exists a smooth 11-form σ\sigma on XX such that ω~ω=dσ\tilde{\omega}-\omega=d\sigma with |σ|ωCzn+12κ\left|\sigma\right|_{\omega}\leqslant Cz^{n+\frac{1}{2}-\kappa}.

After pulling back to the model space 𝒞\mathcal{C} we have Φ(ω~ω)\Phi^{*}(\tilde{\omega}-\omega) is a closed 2-form with |Φ(ω~ω)|ω𝒞Czκ|\Phi^{*}(\tilde{\omega}-\omega)|_{\omega_{\mathcal{C}}}\leqslant Cz^{-\kappa}. By viewing 𝒞\mathcal{C} as Y×(0,+)Y\times(0,+\infty), we can write it as

Φ(ω~ω)=η+dzγ\displaystyle\Phi^{*}(\tilde{\omega}-\omega)=\eta+dz\wedge\gamma

with z\righthalfcupη=0\partial_{z}\righthalfcup\eta=0, z\righthalfcupγ=0\partial_{z}\righthalfcup\gamma=0. Then the fact that d(η+dzγ)=0d(\eta+dz\wedge\gamma)=0 implies dYη=0d_{Y}\eta=0, zη=dYγ\partial_{z}\eta=d_{Y}\gamma. So we can choose

σ~=1zγ𝑑z\displaystyle\tilde{\sigma}=\int_{1}^{z}\gamma dz

such that

dσ~=η+dzγ=Φ(ω~ω).\displaystyle d\tilde{\sigma}=\eta+dz\wedge\gamma=\Phi^{*}(\tilde{\omega}-\omega).

Since |Φ(ω~ω)|ω𝒞Czκ|\Phi^{*}(\tilde{\omega}-\omega)|_{\omega_{\mathcal{C}}}\leqslant Cz^{-\kappa}, we have the decay of dzγdz\wedge\gamma which implies that

|γ|ω𝒞Czn12κ.\displaystyle|\gamma|_{\omega_{\mathcal{C}}}\leqslant Cz^{\frac{n-1}{2}-\kappa}.

Given the formula of ω𝒞\omega_{\mathcal{C}} we can have an estimate of |σ~|ω𝒞|\tilde{\sigma}|_{\omega_{\mathcal{C}}} at the point (y,z0)𝒞(y,z_{0})\in\mathcal{C}:

|σ~(y,z0)|ω𝒞(y,z0)\displaystyle|\tilde{\sigma}(y,z_{0})|_{\omega_{\mathcal{C}}(y,z_{0})} 1z0|γ(y,z)|ω𝒞(y,z0)𝑑z1z0|γ(y,z)|ω𝒞(y,z)z0n12z12𝑑zCz0n+12κ.\displaystyle\leqslant\int_{1}^{z_{0}}|\gamma(y,z)|_{\omega_{\mathcal{C}}(y,z_{0})}dz\leqslant\int_{1}^{z_{0}}|\gamma(y,z)|_{\omega_{\mathcal{C}}(y,z)}z_{0}^{\frac{n-1}{2}}z^{\frac{1}{2}}dz\leqslant Cz_{0}^{n+\frac{1}{2}-\kappa}.

After extending (Φ1)σ~(\Phi^{-1})^{*}\tilde{\sigma} as a smooth 1-form on XX, we can write ω~ω=d((Φ1)σ~)+θ\tilde{\omega}-\omega=d((\Phi^{-1})^{*}\tilde{\sigma})+\theta for some smooth compact supported closed 2-form θ\theta on X.

Recall that XX is 1-convex. Then by the vanishing theorem for 1-convex manifold from Van Coevering [21] Proposition 4.2., θ=i¯s=ddcs\theta=i\partial\bar{\partial}s=dd^{c}s for some compact supported function ss on XX. Then σ=(Φ1)σ~+dcs\sigma=(\Phi^{-1})^{*}\tilde{\sigma}+d^{c}s is the smooth 1-form that we are looking for.

Step 2: Recall that EE is the non-ample locus of KM-K_{M}. The XEX\setminus E admits a complete Kähler metric by Proposition 4.1 in Ohsawa [15]. So we can use L2L^{2}-estimate on XEX\setminus E to solve the ¯\bar{\partial} equation to construct the potential ll such that ω~ω=i¯l\tilde{\omega}-\omega=i\partial\bar{\partial}l.

Let τ=ϵρB1(𝔱)δρB2(𝔱)1n\tau=\epsilon\cdot\rho_{B_{1}}(\mathfrak{t})-\delta\cdot\rho_{B_{2}}(\mathfrak{t})^{\frac{1}{n}}. Choose z0z_{0}, B1B_{1} and B2B_{2} large, then choose δ\delta small depending on ϵ\epsilon, we can guarantee that the (1,1)(1,1) form

i¯τ=i¯(ϵρB1(𝔱)δρB2(𝔱)1n)\displaystyle i\partial\bar{\partial}\tau=i\partial\bar{\partial}(\epsilon\cdot\rho_{B_{1}}(\mathfrak{t})-\delta\cdot\rho_{B_{2}}(\mathfrak{t})^{\frac{1}{n}})

is a Kähler form on XEX\setminus E. We have i¯τCϵ,δ𝔱1ωi\partial\bar{\partial}\tau\geqslant C_{\epsilon,\delta}\mathfrak{t}^{-1}\omega outside a compact set.

If we take the type decomposition of (Φ1)σ~=((Φ1)σ~)1,0+((Φ1)σ~)0,1\left(\Phi^{-1}\right)^{*}\tilde{\sigma}=\left(\left(\Phi^{-1}\right)^{*}\tilde{\sigma}\right)^{1,0}+\left(\left(\Phi^{-1}\right)^{*}\tilde{\sigma}\right)^{0,1}, we have the estimate of ((Φ1)σ~)0,1\left(\left(\Phi^{-1}\right)^{*}\tilde{\sigma}\right)^{0,1} that |((Φ1)σ~)0,1|ωC𝔱1+12κ2n\left|\left(\left(\Phi^{-1}\right)^{*}\tilde{\sigma}\right)^{0,1}\right|_{\omega}\leqslant C\mathfrak{t}^{1+\frac{1-2\kappa}{2n}} for 𝔱>C\mathfrak{t}>C and (Φ1)σ~\left(\Phi^{-1}\right)^{*}\tilde{\sigma} supported on 𝔱>C\mathfrak{t}>C.

So with the same weighted L2L^{2} estimate in Hein-Sun-Viaclovsky-Zhang [7, Proposition 2.2.], we have

XE𝔱|(Φ1)σ~0,1|ωeτωnCXE𝔱𝔱2n+12κ2neϵ𝔱+δ𝔱1nωn<\displaystyle\int_{X\setminus E}\mathfrak{t}\cdot\left|(\Phi^{-1})^{*}\tilde{\sigma}^{0,1}\right|_{\omega}e^{-\tau}\omega^{n}\leqslant C\int_{X\setminus E}\mathfrak{t}\cdot\mathfrak{t}^{\frac{2n+1-2\kappa}{2n}}e^{-\epsilon\mathfrak{t}+\delta\mathfrak{t}^{\frac{1}{n}}}\omega^{n}<\infty

which yields that we have a solution ι\iota such that ¯ι=σ0,1\bar{\partial}\iota=\sigma^{0,1} with

XE|ι|2eτωnXE𝔱|(Φ1)σ~0,1|ωeτωn.\displaystyle\int_{X\setminus E}|\iota|^{2}e^{-\tau}\omega^{n}\leqslant\int_{X\setminus E}\mathfrak{t}\cdot|(\Phi^{-1})^{*}\tilde{\sigma}^{0,1}|_{\omega}e^{-\tau}\omega^{n}.

Consequently, we have i¯(2Imι)=d(Φ1)σ~i\partial\bar{\partial}(2Im\iota)=d(\Phi^{-1})^{*}\tilde{\sigma}. Set l=2Imι+sl=2Im\iota+s then we have ω~ω=i¯l\tilde{\omega}-\omega=i\partial\bar{\partial}l.

Step 3: We give the C0C^{0} bound and CkC^{k} bound for ll via elliptic estimates on the scaled metric.

Let xx be any point in XKX\setminus K. With the same local elliptic estimate under the scaled metric ω^=𝔱(x)1nω\hat{\omega}=\mathfrak{t}(x)^{-\frac{1}{n}}\omega as in Proposition 3.2, we can give a global C0C^{0} bound of ll. We know that ll satisfies the elliptic equation (ω+i¯l)n=ωn(\omega+i\partial\bar{\partial}l)^{n}=\omega^{n} with

Bω^(x,1)|l|2ωneϵC𝔱(x)XE|l|2eϵρB1(𝔱)+δρB2(𝔱)1nωnCϵeϵC𝔱(x),\displaystyle\int_{B_{\hat{\omega}}(x,1)}|l|^{2}\omega^{n}\leqslant e^{\epsilon\cdot C\mathfrak{t}(x)}\int_{X\setminus E}|l|^{2}e^{-\epsilon\cdot\rho_{B_{1}}(\mathfrak{t})+\delta\cdot\rho_{B_{2}}(\mathfrak{t})^{\frac{1}{n}}}\omega^{n}\leqslant C_{\epsilon}e^{\epsilon\cdot C\mathfrak{t}(x)},

since we have some uniform constant CC such that 𝔱(y)C𝔱(x)\mathfrak{t}(y)\leqslant C\cdot\mathfrak{t}(x) for any yBω^(x,1)y\in B_{\hat{\omega}}(x,1) and any xXKx\in X\setminus K. By adjusting ϵ\epsilon small enough we have

lL2(Bω^(x,1))Cϵeϵ𝔱(x).\displaystyle\|l\|_{L^{2}(B_{\hat{\omega}}(x,1))}\leqslant C_{\epsilon}e^{\epsilon\mathfrak{t}(x)}.

Now we can do local elliptic estimates on the scaled metric after lifting to the universal cover. Since the S1S^{1} direction on 𝒞\mathcal{C} collapsing in polynomial order with respect to zz, we know that

lL2(B~ω^(x,1))C𝔱(x)n12nlL2(Bω^(x,1))Cϵeϵ𝔱(x).\displaystyle\|l\|_{L^{2}(\widetilde{B}_{\hat{\omega}}(x,1))}\leqslant C\cdot\mathfrak{t}(x)^{\frac{n-1}{2n}}\cdot\|l\|_{L^{2}(B_{\hat{\omega}}(x,1))}\leqslant C_{\epsilon}e^{\epsilon\mathfrak{t}(x)}.

We have the global C0C^{0} bound for ll:

|l|(x)lW2,2(B~ω^(x,1))ClL2(B~ω^(x,1))Cϵeϵ𝔱(x)\displaystyle|l|(x)\leqslant\|l\|_{W^{2,2}(\widetilde{B}_{\hat{\omega}}(x,1))}\leqslant C\cdot\|l\|_{L^{2}(\widetilde{B}_{\hat{\omega}}(x,1))}\leqslant C_{\epsilon}e^{\epsilon\mathfrak{t}(x)}

for any ϵ>0\epsilon>0. ∎

Remark 7.4.

In the proof of the i¯i\partial\bar{\partial}-lemma 7.3 we did not use the polynomial decay of ω~ω\tilde{\omega}-\omega. In fact, we can always find ll even when |ω~ω||\tilde{\omega}-\omega| is polynomially growth.

Furthermore, we can prove that ω~ω\tilde{\omega}-\omega has weighted higher regularity bound.

Lemma 7.5.

There exists a constant C>0C>0 such that

|zk2k(ω~ω)|ωCzκ\displaystyle\left|z^{\frac{k}{2}}\nabla^{k}(\tilde{\omega}-\omega)\right|_{\omega}\leqslant Cz^{-\kappa}

for any z>Cz>C.

Proof.

Fix any point xx in XX with 𝔱(x)=z0n\mathfrak{t}(x)=z_{0}^{n}. We still work on the scaled metric ω^=z01ω\hat{\omega}=z_{0}^{-1}\omega with uniform bounded curvature. The injectivity radius of the universal covering around xx is bounded below by a universal constant δ\delta independent of xx. Now we are working on the ball B~(x~,δ)\tilde{B}(\tilde{x},\delta) in the universal cover. Since ω~ω\tilde{\omega}-\omega is dd-exact, locally we can take integration of ω~ω\tilde{\omega}-\omega along the geodesic lines to have 1-form σ\sigma on B~(x~,δ)\tilde{B}(\tilde{x},\delta) such that

dσ=ω~ω,σCω^0(B~(x~,δ))ω~ωCω^0(B~(x~,δ))Cz01κ.\displaystyle d\sigma=\tilde{\omega}-\omega,\qquad\|\sigma\|_{C^{0}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}\leqslant\|\tilde{\omega}-\omega\|_{C^{0}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}\leqslant Cz_{0}^{1-\kappa}.

Consider the type decomposition of σ=σ0,1+σ1,0\sigma=\sigma^{0,1}+\sigma^{1,0}. The operator

N:L2(B~(x~,δ),Ω0,1)L2(B~(x~,δ),Ω0,1)N:L^{2}(\tilde{B}(\tilde{x},\delta),\Omega^{0,1})\to L^{2}(\tilde{B}(\tilde{x},\delta),\Omega^{0,1})

constructed in [11, Theorem 8.9] satisfies that

Δ¯(Nσ0,1)=σ0,1,Nσ0,1Lω^2(B~(x~,δ))Cσ0,1Lω^2(B~(x~,δ))\displaystyle\Delta_{\bar{\partial}}(N\sigma^{0,1})=\sigma^{0,1},\quad\|N\sigma^{0,1}\|_{L^{2}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}\leqslant C\|\sigma^{0,1}\|_{L^{2}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}

and NN commutes with \partial and ¯\bar{\partial}. Then

Nσ0,1Wω^2,2(B~(x~,δ))Cσ0,1Lω^2(B~(x~,δ))Cσ0,1Cω^0(B~(x~,δ)).\displaystyle\|N\sigma^{0,1}\|_{W^{2,2}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}\leqslant C\|\sigma^{0,1}\|_{L^{2}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}\leqslant C\|\sigma^{0,1}\|_{C^{0}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}.

Then we know by Sobolev lemma and iteration process that for any q>1q>1

Nσ0,1Wω^2,q(B~(x~,δ))Cσ0,1Lω^q(B~(x~,δ))Cσ0,1Cω^0(B~(x~,δ)).\displaystyle\|N\sigma^{0,1}\|_{W^{2,q}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}\leqslant C\|\sigma^{0,1}\|_{L^{q}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}\leqslant C\|\sigma^{0,1}\|_{C^{0}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}.

Take q>nq>n, there exists α>0\alpha>0 such that

Nσ0,1Cω^1,α(B~(x~,δ))CNσ0,1Wω^2,q(B~(x~,δ))Cσ0,1Cω^0(B~(x~,δ)).\displaystyle\|N\sigma^{0,1}\|_{C^{1,\alpha}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}\leqslant C\|N\sigma^{0,1}\|_{W^{2,q}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}\leqslant C\|\sigma^{0,1}\|_{C^{0}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}.

Let f=¯Nσ0,1f=\bar{\partial}^{*}N\sigma^{0,1}. We have

fCω^0,α(B~(x~,δ))Cσ0,1Cω^0(B~(x~,δ)),¯f=Δ¯(Nσ0,1)=σ0,1.\displaystyle\|f\|_{C^{0,\alpha}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}\leqslant C\|\sigma^{0,1}\|_{C^{0}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))},\qquad\bar{\partial}f=\Delta_{\bar{\partial}}(N\sigma^{0,1})=\sigma^{0,1}.

Let l^=z012Imf\hat{l}=z_{0}^{-1}\cdot 2\mathrm{Im}f. We have i¯l^=z01(ω~ω)i\partial\bar{\partial}\hat{l}=z_{0}^{-1}(\tilde{\omega}-\omega). Hence (ω^+i¯l^)n=ω^n(\hat{\omega}+i\partial\bar{\partial}\hat{l})^{n}=\hat{\omega}^{n} with

l^Cω^0,α(B~(x~,δ))Cz0κ.\displaystyle\|\hat{l}\|_{C^{0,\alpha}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}\leqslant Cz_{0}^{-\kappa}.

By Schauder estimates we have higher regularity

l^Cω^k,α(B~(x~,δ))Cz0κ\displaystyle\|\hat{l}\|_{C^{k,\alpha}_{\hat{\omega}}(\tilde{B}(\tilde{x},\delta))}\leqslant Cz_{0}^{-\kappa}

which yields

|z0k2k(ω~ω)|ωCz0κ\displaystyle\left|z_{0}^{\frac{k}{2}}\nabla^{k}(\tilde{\omega}-\omega)\right|_{\omega}\leqslant Cz_{0}^{-\kappa}

for any z0>Cz_{0}>C. ∎

Then we are ready to prove the uniqueness:

Proof of Theorem 7.1.

Given by previous estimate, we have ω~ω=i¯l\tilde{\omega}-\omega=i\partial\bar{\partial}l with |l|Cϵeϵ𝔱|l|\leqslant C_{\epsilon}e^{\epsilon\mathfrak{t}}. If we pull back ll to 𝒞\mathcal{C}, by the closeness of complex structure we have

|dJ𝒞dl|ω𝒞|d(J𝒞JX)dl|ω𝒞+|dJXdl|ω𝒞C(e(12ϵ)zn+zκ).\displaystyle|dJ_{\mathcal{C}}dl|_{\omega_{\mathcal{C}}}\leqslant|d(J_{\mathcal{C}}-J_{X})dl|_{\omega_{\mathcal{C}}}+|dJ_{X}dl|_{\omega_{\mathcal{C}}}\leqslant C(e^{-(\frac{1}{2}-\epsilon)z^{n}}+z^{-\kappa}).

The function Fl=Δω𝒞lF_{l}=\Delta_{\omega_{\mathcal{C}}}l has higher regularity bound on 𝒞\mathcal{C}:

|zk2kFl|ω𝒞Czκ for any z>C.\displaystyle\left|z^{\frac{k}{2}}\nabla^{k}F_{l}\right|_{\omega_{\mathcal{C}}}\leqslant Cz^{-\kappa}\text{ for any }z>C.

By Proposition 3.2 there exists a smooth function ff on 𝒞\mathcal{C} such that Δω𝒞f=Fl\Delta_{\omega_{\mathcal{C}}}f=F_{l} with

|i¯f|ω𝒞Czκ+ϵ,|df|ω𝒞Czn+12κ+ϵ,|f|Czn+1κ+ϵ for any ϵ>0.\displaystyle|i\partial\bar{\partial}f|_{\omega_{\mathcal{C}}}\leqslant Cz^{-\kappa+\epsilon},\quad|df|_{\omega_{\mathcal{C}}}\leqslant Cz^{\frac{n+1}{2}-\kappa+\epsilon},\quad|f|\leqslant Cz^{n+1-\kappa+\epsilon}\text{ for any }\epsilon>0.

Since Δω𝒞(lf)=0\Delta_{\omega_{\mathcal{C}}}(l-f)=0 and |lf|eϵt|l-f|\leqslant e^{\epsilon t} for any ϵ>0\epsilon>0, from the behavior of harmonic function [16, Proposition 5.3.] we know that l=f+λz+g+O(eδz)l=f+\lambda z+g+O(e^{-\delta z}) for some λ>0\lambda>0 and some harmonic S1S^{1}-invariant function gg on 𝒞\mathcal{C} with |g|Ceδzn2|g|\leqslant Ce^{\delta z^{\frac{n}{2}}}. Since |zn1gtt(t,q)||i¯g|ω𝒞Czκ+ϵ|z^{n-1}g_{tt}(t,q)|\leqslant|i\partial\bar{\partial}g|_{\omega_{\mathcal{C}}}\leqslant Cz^{-\kappa+\epsilon} holds uniformly for any qDq\in D, integration along the \mathbb{R}-fiber direction shows that gg and hence ll is at most polynomially growth. Again by [16, Proposition 5.3.] we know that l=f+λz+O(eδz)l=f+\lambda z+O(e^{-\delta z}).

Recall that ll satisfies that

Δω𝒞l=Δω𝒞lΔΦωl+k=2n(nk)(dΦJXdl)kΦωnk,\Delta_{\omega_{\mathcal{C}}}l=\Delta_{\omega_{\mathcal{C}}}l-\Delta_{\Phi^{*}\omega}l+\sum_{k=2}^{n}{n\choose k}(d\Phi^{*}J_{X}dl)^{k}\wedge\Phi^{*}\omega^{n-k},

by our previous construction we know that |Φωω𝒞|ω𝒞Cz1|\Phi^{*}\omega-\omega_{\mathcal{C}}|_{\omega_{\mathcal{C}}}\leqslant Cz^{-1}, so |Δω𝒞lΔΦωl|ω𝒞Cz1κ|\Delta_{\omega_{\mathcal{C}}}l-\Delta_{\Phi^{*}\omega}l|_{\omega_{\mathcal{C}}}\leqslant Cz^{-1-\kappa},

Δω𝒞lC(z1|i¯l|ω𝒞+|i¯l|ω𝒞2)Czmin{1+κ,2κ,1+n}.\displaystyle\Delta_{\omega_{\mathcal{C}}}l\leqslant C\left(z^{-1}|i\partial\bar{\partial}l|_{\omega_{\mathcal{C}}}+|i\partial\bar{\partial}l|^{2}_{\omega_{\mathcal{C}}}\right)\leqslant Cz^{-\min\{1+\kappa,2\kappa,1+n\}}. (7.1)

Thus by finite step iteration we can find a better candidate f~\tilde{f} and another constant λ~\tilde{\lambda} such that

l=f~+λ~z+O(eδz),|2f~|ω𝒞Czn1+ϵ,|f~|ω𝒞Czn+12+ϵ,|f~|<Czϵ.\displaystyle l=\tilde{f}+\tilde{\lambda}z+O(e^{-\delta z}),\quad|\nabla^{2}\tilde{f}|_{\omega_{\mathcal{C}}}\leqslant Cz^{-n-1+\epsilon},\quad|\nabla\tilde{f}|_{\omega_{\mathcal{C}}}\leqslant Cz^{-\frac{n+1}{2}+\epsilon},\quad|\tilde{f}|<Cz^{\epsilon}.

On the other hand,

0\displaystyle 0 =Xi¯lk=1n(nk)(i¯l)k1ωnk=limε0{(Φ1)tlogε}i¯lk=1n(nk)(i¯l)k1ωnk\displaystyle=\int_{X}i\partial\bar{\partial}l\wedge\sum_{k=1}^{n}{n\choose k}(i\partial\bar{\partial}l)^{k-1}\wedge\omega^{n-k}=\lim_{\varepsilon\to 0}\int_{\{(\Phi^{-1})^{*}t\leqslant-\log\varepsilon\}}i\partial\bar{\partial}l\wedge\sum_{k=1}^{n}{n\choose k}(i\partial\bar{\partial}l)^{k-1}\wedge\omega^{n-k}
=limε0{t=logε}λ~dczk=1n(nk)(i¯l)k1ωnk=limε0{t=logε}λ~dct(i¯t)n1=λ~Vol(D).\displaystyle=\lim_{\varepsilon\to 0}\int_{\{t=-\log\varepsilon\}}\tilde{\lambda}d^{c}z\wedge\sum_{k=1}^{n}{n\choose k}(i\partial\bar{\partial}l)^{k-1}\wedge\omega^{n-k}=\lim_{\varepsilon\to 0}\int_{\{t=-\log\varepsilon\}}\tilde{\lambda}d^{c}t\wedge(i\partial\bar{\partial}t)^{n-1}=\tilde{\lambda}\mathrm{Vol}(D).

Consequently, λ~=0\tilde{\lambda}=0, lCzϵl\leqslant Cz^{\epsilon} for any ϵ>0\epsilon>0. From the equation of ll we know that

limε0|{(Φ1)t=logε}ldclk=1n(nk)(i¯l)k1ωnk|limε0{t=logε}|Cz1ϵdct(i¯t)n1|=0.\displaystyle\lim_{\varepsilon\to 0}\left|\int_{\{\left(\Phi^{-1}\right)^{*}t=-\log\varepsilon\}}ld^{c}l\wedge\sum_{k=1}^{n}{n\choose k}(i\partial\bar{\partial}l)^{k-1}\wedge\omega^{n-k}\right|\leqslant\lim_{\varepsilon\to 0}\int_{\{t=-\log\varepsilon\}}\left|\frac{C}{z^{1-\epsilon}}\cdot d^{c}t\wedge(i\partial\bar{\partial}t)^{n-1}\right|=0.

Hence by integration by parts and 0=l(ω~nωn)=li¯lk=0n1ωkω~n1k0=l(\tilde{\omega}^{n}-\omega^{n})=l\cdot i\partial\bar{\partial}l\wedge\sum_{k=0}^{n-1}\omega^{k}\wedge\tilde{\omega}^{n-1-k}:

0=Xli¯lk=0n1ωkω~n1k=X𝑑ldclk=0n1ωkω~n1k.\displaystyle 0=-\int_{X}l\cdot i\partial\bar{\partial}l\wedge\sum_{k=0}^{n-1}\omega^{k}\wedge\tilde{\omega}^{n-1-k}=\int_{X}dl\wedge d^{c}l\wedge\sum_{k=0}^{n-1}\omega^{k}\wedge\tilde{\omega}^{n-1-k}.

Since k=0n1ωkω~n1k\sum_{k=0}^{n-1}\omega^{k}\wedge\tilde{\omega}^{n-1-k} is a positive form, we know that dl=dcl=0dl=d^{c}l=0. ∎

8. Discussion and Questions

8.1. Examples

We present examples that (X,ω)(X,\omega) is a Calabi-Yau manifold not asymptotically Calabi but weak asymptotically Calabi under the fixed diffeomorphism Φ\Phi. As discussed in the end of the proof of Theorem 1.3, we have the following:

Claim.

Let (M,D)(M,D) be the pair in Definition 1.2 with X=MDX=M\setminus D. Let H+,c2(X)=Im(Hc2(X)H2(X))H+2(X)H_{+,c}^{2}(X)=\mathrm{Im}(H^{2}_{c}(X)\to H^{2}(X))\cap H_{+}^{2}(X). Fix a diffeomorphism Φ:𝒞𝒦XK\Phi:\mathcal{C}\setminus\mathcal{K}\to X\setminus K. Then for any 𝔨\mathfrak{k} in H+2(X)H^{2}_{+}(X) but not H+,c2(X)H^{2}_{+,c}(X) the metric ω\omega we constructed in Theorem 1.3 is a Calabi-Yau metric not asymptotically Calabi but weak asymptotically Calabi.

Example 8.1.

Let M=1×2M=\mathbb{P}^{1}\times\mathbb{P}^{2} with two projection maps π1:M1\pi_{1}:M\to\mathbb{P}^{1} and π2:M2\pi_{2}:M\to\mathbb{P}^{2}. Then we have D=KM=π1O1(2)π2O2(3)D=-K_{M}=\pi_{1}^{*}O_{\mathbb{P}^{1}}(2)\otimes\pi_{2}^{*}O_{\mathbb{P}^{2}}(3). Pic(M)\mathrm{Pic}(M) is generated by π1O1(1)\pi_{1}^{*}O_{\mathbb{P}^{1}}(1) and π2O2(1)\pi_{2}^{*}O_{\mathbb{P}^{2}}(1) and the image of each of them under the map i:H2(M)H2(D)i^{*}:H^{2}(M)\to H^{2}(D) induced by the inclusion map i:DMi:D\to M is not parallel to [ωD]=c1(ND)[\omega_{D}]=c_{1}(N_{D}). Choose a primitive representative of any of these two classes and apply Theorem 1.3 we will find a Calabi-Yau metric not asymptotically Calabi but weak asymptotically Calabi.

These kind of examples could be found on any Fano manifold MM with dimM3\mathrm{dim}_{\mathbb{C}}M\geqslant 3 and h2(M)2h_{2}(M)\geqslant 2. We can find many examples in Mori-Mukai [14]. Besides, there are also many examples in the weak Fano case but we do not have a simple topological sufficient condition.

8.2. Weaker Decay Condition

In our statement of uniqueness Theorem 1.5, we need the metric ω~\tilde{\omega} to be polynomially closed to ω\omega. The main difficulty to get rid of this condition lies in how to deduce the decomposition of l=f+λz+O(eδz)l=f+\lambda z+O(e^{-\delta z}) with |f|zϵ|f|\leqslant z^{\epsilon} for any ϵ>0\epsilon>0, where we cannot do iteration to improve the decay of ff as in (7.1).

It is natural to ask the following question:

Question 8.2.

Can we prove a stronger uniqueness theorem: If we have another Calabi-Yau metric ω~\tilde{\omega} such that |ω~ω|ω0|\tilde{\omega}-\omega|_{\omega}\to 0 when rr\to\infty for some distance function rr with respect to ω\omega, then ω~=ω\tilde{\omega}=\omega?

One possible obstruction of this stronger uniqueness theorem is that we cannot rule out the possibility that there is a Calabi-Yau metric ω\omega closed to the Calabi model space in a logarithm rate rather than any polynomial rate. The existence of this type of Calabi-Yau metric is also an interesting question to study.

8.3. Compactification and Classification

Hein-Sun-Viaclovsky-Zhang [7] showed that any asymptotically Calabi manifold which is Calabi-Yau can be compactified complex analytically to a weak Fano manifold and the Calabi-Yau comes from the construction by Tian-Yau-Hein’s package.

In the weak asymptotic Calabi manifold case, when we only have the exponential closeness of complex structure, even though the metric is polynomial close, we can still get that any weak asymptotically Calabi manifold which is Calabi-Yau can be compactified complex analytically to a weak Fano manifold by repeating the argument in [7] as one can also do L2L^{2} estimate to construct the holomorphic function on XX from the holomorphic section of NDN_{D}. The key difference is to show that the compactification we get is Kähler by considering the behavior of the class at the end. By our uniqueness Theorem 1.5 this Calabi-Yau metric ω\omega comes from our generalized Tian-Yau construction in Theorem 1.3.

We would like to make the following conjecture to further generalize this into slower decay assumption.

Definition 8.3.

Let XX be a complete Kähler manifold with complex dimension nn, complex structure II, Kähler form ω\omega and (n,0)(n,0)-form Ω\Omega. We say (X,I,ω,Ω)(X,I,\omega,\Omega) is polynomial asymptotically Calabi with rate (κ1,κ2)(\kappa_{1},\kappa_{2}) if:

there exists κ1,κ2>0\kappa_{1},\kappa_{2}>0, a Calabi model space (𝒞,I𝒞,ω𝒞,Ω𝒞)(\mathcal{C},I_{\mathcal{C}},\omega_{\mathcal{C}},\Omega_{\mathcal{C}}), and a diffeomorphism Φ:𝒞𝒦XK\Phi:\mathcal{C}\setminus\mathcal{K}\rightarrow X\setminus K, where KXK\subset X and 𝒦𝒞\mathcal{K}\subset\mathcal{C} are compact, such that the following hold uniformly as z+z\rightarrow+\infty :

|ω𝒞k(ΦIXI𝒞)|ω𝒞+|ω𝒞k(ΦΩΩ𝒞)|ω𝒞=O(zκ1),|ω𝒞k(Φωω𝒞)|ω𝒞=O(zκ2)\left|\nabla_{\omega_{\mathcal{C}}}^{k}\left(\Phi^{*}I_{X}-I_{\mathcal{C}}\right)\right|_{\omega_{\mathcal{C}}}+\left|\nabla_{\omega_{\mathcal{C}}}^{k}\left(\Phi^{*}\Omega-\Omega_{\mathcal{C}}\right)\right|_{\omega_{\mathcal{C}}}=O\left(z^{-\kappa_{1}}\right),\quad\left|\nabla_{\omega_{\mathcal{C}}}^{k}\left(\Phi^{*}\omega-\omega_{\mathcal{C}}\right)\right|_{\omega_{\mathcal{C}}}=O\left(z^{-\kappa_{2}}\right)

for all k0k\in\mathbb{N}_{0}.

Conjecture 8.4.

There are optimal constants λ\lambda and μ\mu such that for any κ1>λ\kappa_{1}>\lambda and κ2>μ\kappa_{2}>\mu, any polynomial asymptotically Calabi Calabi-Yau manifold with rate (κ1,κ2)(\kappa_{1},\kappa_{2}) can be compactified complex analytically to a weak Fano manifold. Furthermore, the Calabi-Yau metric comes from our generalized Tian-Yau construction in Theorem 1.3.

When κ1\kappa_{1} is large enough, we can still use L2L^{2} estimate to construct holomorphic coordinate on the end of X¯\mskip 1.5mu\overline{\mskip-1.5mu{X}\mskip-1.5mu}\mskip 1.5mu. However, the question to find optimal λ\lambda may not be approachable by L2L^{2} method. From the uniqueness theorem 1.5 and the compactification process of weak asymptotically Calabi Calabi-Yau manifold, we expect that μ\mu should be 0.

Appendix A estimate of the solution of ODE

In this section, we will look closely to the solution of the following ordinary differential equation:

u′′(j2n24+nλ)zn2u=nzn1v,\displaystyle u^{\prime\prime}-(\frac{j^{2}n^{2}}{4}+n\lambda)z^{n-2}u=nz^{n-1}v,

where λ>0\lambda>0, n3n\geqslant 3 and n,jn,j\in\mathbb{N}.
By the transformation in [16], we have two cases: zero node case when j=0j=0 and non-zero node case when j>0j>0. We will give a brief summary of the estimate of fundamental solutions and have a estimate of uu with polynomial rate which slightly generalizes the results in [16].

A.1. fundamental solution of zero mode

In this section we focus on the zero mode: the equation

u′′nλzn2u=nzn1v.\displaystyle u^{\prime\prime}-n\lambda z^{n-2}u=nz^{n-1}v. (A.1)

By [16] we have the decay solution 𝒟(z)\mathcal{D}(z) and growth solution 𝒢(z)\mathcal{G}(z) of the homogeneous equation u′′(z)=nzn2λu(z)u^{\prime\prime}(z)=nz^{n-2}\lambda u(z) given by

𝒟(z)=zK1n(2λnzn2),\displaystyle\mathcal{D}(z)=\sqrt{z}K_{\frac{1}{n}}\left(2\sqrt{\tfrac{\lambda}{n}}\cdot z^{\frac{n}{2}}\right), (A.2)
𝒢(z)=zI1n(2λnzn2)\displaystyle\mathcal{G}(z)=\sqrt{z}I_{\frac{1}{n}}\left(2\sqrt{\tfrac{\lambda}{n}}\cdot z^{\frac{n}{2}}\right) (A.3)

where KK and II have the following expression: for ν\nu\in\mathbb{R}

Kν(y)\displaystyle K_{\nu}(y) =0eycoshtcosh(νt)𝑑t,\displaystyle=\int_{0}^{\infty}e^{-y\cosh t}\cosh(\nu t)dt,
Iν(y)\displaystyle I_{\nu}(y) =1π0πeycosθcos(νθ)𝑑θsin(νπ)π0eycoshtνt𝑑t\displaystyle=\frac{1}{\pi}\int_{0}^{\pi}e^{y\cos\theta}\cos(\nu\theta)d\theta-\frac{\sin(\nu\pi)}{\pi}\int_{0}^{\infty}e^{-y\cosh t-\nu t}dt
Lemma A.1.

[16][Proposition 3.3.] We have the following uniform estimate:

  1. (1)

    For all ν\nu\in\mathbb{R}, there is a constant C(ν)>1C(\nu)>1 such that

    C1(ν)eyyKν(y)C(ν)eyy,y1;Iν(y){C(ν)eyy,y1,C(ν)yν,0<y1.\begin{array}[]{l}C^{-1}(\nu)\cdot\frac{e^{-y}}{\sqrt{y}}\leqslant K_{\nu}(y)\leqslant C(\nu)\cdot\frac{e^{-y}}{\sqrt{y}},\quad y\geqslant 1;\\ I_{\nu}(y)\leqslant\begin{cases}C(\nu)\cdot\frac{e^{y}}{\sqrt{y}},&y\geqslant 1,\\ C(\nu)\cdot y^{\nu},&0<y\leqslant 1.\end{cases}\end{array}
  2. (2)

    For all ν>1\nu>-1, we have

    Iν(y){C(ν)1eyy,y1C(ν)1yν,0<y1I_{\nu}(y)\geqslant\begin{cases}C(\nu)^{-1}\cdot\frac{e^{y}}{\sqrt{y}},&y\geqslant 1\\ C(\nu)^{-1}\cdot y^{\nu},&0<y\leqslant 1\end{cases}
Corollary A.2.

[16] For z>n4λ1nz>\sqrt[n]{\frac{n}{4\lambda_{1}}}, there exists a constant CC which only depends on nn such that

1Ce2nλ12zn2λ14zn24<\displaystyle\frac{1}{C}\cdot\frac{e^{-\frac{2}{\sqrt{n}}\lambda^{\frac{1}{2}}\cdot z^{\frac{n}{2}}}}{\lambda^{\frac{1}{4}}\cdot z^{\frac{n-2}{4}}}< 𝒟(z)<Ce2nλ12zn2λ14zn24\displaystyle\;\mathcal{D}(z)<C\cdot\frac{e^{-\frac{2}{\sqrt{n}}\lambda^{\frac{1}{2}}\cdot z^{\frac{n}{2}}}}{\lambda^{\frac{1}{4}}\cdot z^{\frac{n-2}{4}}}
1Ce2nλ12zn2λ14zn24<\displaystyle\frac{1}{C}\cdot\frac{e^{\frac{2}{\sqrt{n}}\lambda^{\frac{1}{2}}\cdot z^{\frac{n}{2}}}}{\lambda^{\frac{1}{4}}\cdot z^{\frac{n-2}{4}}}< 𝒢(z)<Ce2nλ12zn2λ14zn24.\displaystyle\;\mathcal{G}(z)<C\cdot\frac{e^{\frac{2}{\sqrt{n}}\lambda^{\frac{1}{2}}\cdot z^{\frac{n}{2}}}}{\lambda^{\frac{1}{4}}\cdot z^{\frac{n-2}{4}}}.

With those estimates, we can give a C0C^{0} bound of u(z)u(z). By computation in [16] we know that the Wronskian

𝒲(𝒢,𝒟)=𝒢(z)𝒟(z)𝒢(z)𝒟(z)=n2.\mathcal{W}(\mathcal{G},\mathcal{D})=\mathcal{G}(z)\mathcal{D}^{\prime}(z)-\mathcal{G}^{\prime}(z)\mathcal{D}(z)=-\frac{n}{2}.

Hence we have a solution of A.1 as follows:

u(z)\displaystyle u(z) =2(𝒟(z)1z𝒢(s)sn1v(s)𝑑s+𝒢(z)z𝒟(s)sn1v(s)𝑑s)\displaystyle=-2\left(\mathcal{D}(z)\int_{1}^{z}\mathcal{G}(s)s^{n-1}v(s)ds+\mathcal{G}(z)\int_{z}^{\infty}\mathcal{D}(s)s^{n-1}v(s)ds\right) (A.4)

We firstly introduce an estimate of the solution of this ordinary differential equation:

Proposition A.3.

Recall that λ1\lambda_{1} is the first nonzero positive eigenvalue of ΔY-\Delta_{Y}. Let vv be a function such that |v(z)|C0zδ|v(z)|\leqslant C_{0}z^{\delta} for z>1z>1. For any λ\lambda such that λ>λ1>0\lambda>\lambda_{1}>0, we can find solution of equation u′′(z)=nzn2λu(z)+nzn1vu^{\prime\prime}(z)=nz^{n-2}\lambda u(z)+nz^{n-1}v such that |u(z)|CC0zδ+1|u(z)|\leqslant C\cdot C_{0}z^{\delta+1}, |u(z)|CC0zδ+n2|u^{\prime}(z)|\leqslant C\cdot C_{0}z^{\delta+\frac{n}{2}}, |u′′(z)|CC0zδ+n1|u^{\prime\prime}(z)|\leqslant C\cdot C_{0}z^{\delta+n-1} on z>Cz>C for some constant C>1C>1 only depend on nn, λ1\lambda_{1} and δ\delta .

Proof.

Now we can estimate uL((max{1,(n4λ1)1n},))\left\|u\right\|_{L^{\infty}((\max\{1,(\frac{n}{4\lambda_{1}})^{\frac{1}{n}}\},\infty))}. Let μ=2nλ12\mu=\tfrac{2}{\sqrt{n}}\lambda^{\frac{1}{2}}. By integration by parts the first term 𝒟(z)1z𝒢(s)sn1v(s)𝑑s\mathcal{D}(z)\int_{1}^{z}\mathcal{G}(s)s^{n-1}v(s)ds in A.4 is bounded by a constant C(n)C(n) times the following term:

1λ12zn241zeμ(sn2zn2)sδ+3n24𝑑s\displaystyle\frac{1}{\lambda^{\frac{1}{2}}\cdot z^{\frac{n-2}{4}}}\int_{1}^{z}e^{\,\mu\left(s^{\frac{n}{2}}-z^{\frac{n}{2}}\right)}s^{\delta+\frac{3n-2}{4}}ds
\displaystyle\leqslant 1nλzδ+1δ+n+24nλ32zδ+1n2+(δ+n+24)(δ+n+24)nnλ2zδ+1n\displaystyle\,\frac{1}{\sqrt{n}\lambda}z^{\delta+1}-\frac{\delta+\frac{n+2}{4}}{n\lambda^{\frac{3}{2}}}z^{\delta+1-\frac{n}{2}}+\frac{(\delta+\frac{n+2}{4})(\delta+\frac{-n+2}{4})}{n\sqrt{n}\lambda^{2}}z^{\delta+1-n}
(δ+n+24)(δ+n+24)(δ+3n+24)n2λ52zδ+13n2\displaystyle\hskip 71.13188pt-\frac{(\delta+\frac{n+2}{4})(\delta+\frac{-n+2}{4})(\delta+\frac{-3n+2}{4})}{n^{2}\lambda^{\frac{5}{2}}}z^{\delta+1-\frac{3n}{2}}
+max{0,(δ+n+24)(δ+n+24)(δ+3n+24)(δ+5n+24)}n2λ52zδ+13n2\displaystyle\hskip 71.13188pt+\frac{\max\{0,(\delta+\frac{n+2}{4})(\delta+\frac{-n+2}{4})(\delta+\frac{-3n+2}{4})(\delta+\frac{-5n+2}{4})\}}{n^{2}\lambda^{\frac{5}{2}}}z^{\delta+1-\frac{3n}{2}}
(1nλ(δ+n+24)nλ32+(δ+n+24)(δ+n+24)nnλ2+C(n,δ)1λ52)eμeμzn2zn24.\displaystyle\hskip 71.13188pt-\left(\frac{1}{\sqrt{n}\lambda}-\frac{(\delta+\frac{n+2}{4})}{n\lambda^{\frac{3}{2}}}+\frac{(\delta+\frac{n+2}{4})(\delta+\frac{-n+2}{4})}{n\sqrt{n}\lambda^{2}}+C(n,\delta)\frac{1}{\lambda^{\frac{5}{2}}}\right)e^{\mu}\frac{e^{-\mu z^{\frac{n}{2}}}}{z^{\frac{n-2}{4}}}.

Here we also use the following observation: Since μ2nλ112\mu\geqslant\frac{2}{\sqrt{n}}\lambda_{1}^{\frac{1}{2}}, we know that the maximum of sδ+5n24eμ(sn2zn2)s^{\delta+\frac{-5n-2}{4}}e^{\,\mu\left(s^{\frac{n}{2}}-z^{\frac{n}{2}}\right)} on the interval [1,z][1,z] is at zz when zz is larger than a uniform constant which is independent with respect to λ\lambda but only on nn and λ1\lambda_{1}. So we have

1zsδ+5n24eμ(sn2zn2)𝑑s(z1)zδ+5n24<zδ+5n+24.\int_{1}^{z}s^{\delta+\frac{-5n-2}{4}}e^{\,\mu\left(s^{\frac{n}{2}}-z^{\frac{n}{2}}\right)}ds\leqslant(z-1)z^{\delta+\frac{-5n-2}{4}}<z^{\delta+\frac{-5n+2}{4}}.

For the second term 𝒢(z)z𝒟(s)sn1v(s)𝑑s\mathcal{G}(z)\int_{z}^{\infty}\mathcal{D}(s)s^{n-1}v(s)ds, we have similar estimate:

1λ12zn24zeμ(zn2sn2)sδ+3n24𝑑s\displaystyle\frac{1}{\lambda^{\frac{1}{2}}\cdot z^{\frac{n-2}{4}}}\int_{z}^{\infty}e^{\,\mu\left(z^{\frac{n}{2}}-s^{\frac{n}{2}}\right)}s^{\delta+\frac{3n-2}{4}}ds
\displaystyle\leqslant 1nλzδ+1+δ+n+24nλ32zδ+1n2+(δ+n+24)(δ+n+24)nnλ2zδ+1n\displaystyle\,\frac{1}{\sqrt{n}\lambda}z^{\delta+1}+\frac{\delta+\frac{n+2}{4}}{n\lambda^{\frac{3}{2}}}z^{\delta+1-\frac{n}{2}}+\frac{(\delta+\frac{n+2}{4})(\delta+\frac{-n+2}{4})}{n\sqrt{n}\lambda^{2}}z^{\delta+1-n}
+2|(δ+n+24)(δ+n+24)(δ+3n+24)|n2λ52zδ+13n2.\displaystyle\hskip 56.9055pt+\frac{2\left|(\delta+\frac{n+2}{4})(\delta+\frac{-n+2}{4})(\delta+\frac{-3n+2}{4})\right|}{n^{2}\lambda^{\frac{5}{2}}}z^{\delta+1-\frac{3n}{2}}.

So we have the uniform estimate for uu that for any z>C(n,δ,λ1)z>C(n,\delta,\lambda_{1}),

|u(z)|C(n)C0zδ+1λ.\displaystyle|u(z)|\leqslant C(n)\cdot C_{0}\frac{z^{\delta+1}}{\lambda}.

For the derivative uu^{\prime} we can do the same computation as in [16] to estimate 𝒟(z)\mathcal{D}^{\prime}(z) and 𝒢(z)\mathcal{G}^{\prime}(z). In fact, we have the following estimate:

Lemma A.4.
1Ceyy<I1n(y)<Ceyy,1Ceyy<K1n(y)<Ceyy\displaystyle\frac{1}{C}\cdot\frac{e^{y}}{\sqrt{y}}<I_{\frac{1}{n}}^{\prime}(y)<C\cdot\frac{e^{y}}{\sqrt{y}},\quad\frac{1}{C}\cdot\frac{e^{-y}}{\sqrt{y}}<-K_{\frac{1}{n}}^{\prime}(y)<C\cdot\frac{e^{-y}}{\sqrt{y}}

for some fixed constant CC and any y>1y>1.

Proof.

Notice that

I1n(y)\displaystyle I_{\frac{1}{n}}^{\prime}(y) =12π0πeycosθ(cos(n+1)θn+cos(n1)θn)𝑑θ+sin(πn)2π0eycosht(e(n+1)tn+e(n1)tn)𝑑t.\displaystyle=\frac{1}{2\pi}\int_{0}^{\pi}e^{y\cos\theta}\left(\cos\tfrac{(n+1)\theta}{n}+\cos\tfrac{(n-1)\theta}{n}\right)d\theta+\frac{\sin{\frac{\pi}{n}}}{2\pi}\int_{0}^{\infty}e^{-y\cosh t}\left(e^{-\frac{(n+1)t}{n}}+e^{\frac{(n-1)t}{n}}\right)dt.

As in the proof of [16] Prop. 3.3, we know that for any ν\nu\in\mathbb{R}

0eycoshtνt𝑑tey0eyt22νt𝑑tC(ν)eyy,\displaystyle\int_{0}^{\infty}e^{-y\cosh t-\nu t}dt\leqslant e^{-y}\int_{0}^{\infty}e^{-\frac{yt^{2}}{2}-\nu t}dt\leqslant C(\nu)\cdot\frac{e^{-y}}{\sqrt{y}},
|0πeycosθcos(νθ)𝑑θ|ey0π3eyθ24𝑑θ+2ey232eyπy+2ey2310eyy.\displaystyle\left|\int_{0}^{\pi}e^{y\cos\theta}\cos(\nu\theta)d\theta\right|\leqslant e^{y}\int_{0}^{\frac{\pi}{3}}e^{-\frac{y\cdot\theta^{2}}{4}}d\theta+\frac{2e^{\frac{y}{2}}}{3}\leqslant\frac{2e^{y}}{\sqrt{\pi}\cdot\sqrt{y}}+\frac{2e^{\frac{y}{2}}}{3}\leqslant\frac{10e^{y}}{\sqrt{y}}.

For ν>1\nu>-1 and ν0\nu\neq 0, let ην=min(π,π3|ν|)\eta_{\nu}=\min\left(\pi,\frac{\pi}{3|\nu|}\right),

0πeycosθcos(νθ)𝑑θ=0ηνeycosθcos(νθ)𝑑θ+ηνπeycosθcos(νθ)𝑑θ\displaystyle\quad\int_{0}^{\pi}e^{y\cos\theta}\cos(\nu\theta)d\theta=\int_{0}^{\eta_{\nu}}e^{y\cos\theta}\cos(\nu\theta)d\theta+\int_{\eta_{\nu}}^{\pi}e^{y\cos\theta}\cos(\nu\theta)d\theta
12ey0ηνeθ22y𝑑θ|ηνπeycosθcos(νθ)𝑑θ|C(ν)eyyηνπeycosθ𝑑θ\displaystyle\geqslant\frac{1}{2}e^{y}\int_{0}^{\eta_{\nu}}e^{-\frac{\theta^{2}}{2}y}d\theta-\left|\int_{\eta_{\nu}}^{\pi}e^{y\cos\theta}\cos(\nu\theta)d\theta\right|\geqslant C(\nu)\frac{e^{y}}{\sqrt{y}}-\int_{\eta_{\nu}}^{\pi}e^{y\cos\theta}d\theta
C(ν)eyy(πην)ecos(ην)yC(ν)eyy.\displaystyle\geqslant C(\nu)\frac{e^{y}}{\sqrt{y}}-\left(\pi-\eta_{\nu}\right)e^{\cos\left(\eta_{\nu}\right)y}\geqslant C(\nu)\frac{e^{y}}{\sqrt{y}}.

So we get that 1Ceyy<I1n(y)<Ceyy\frac{1}{C}\cdot\frac{e^{y}}{\sqrt{y}}<I_{\frac{1}{n}}^{\prime}(y)<C\cdot\frac{e^{y}}{\sqrt{y}}.

On the other hand,

K1n(y)\displaystyle K_{\frac{1}{n}}^{\prime}(y) =120eycosht(cosh((n+1)tn)+cosh((n1)tn))𝑑t\displaystyle=-\frac{1}{2}\int_{0}^{\infty}e^{-y\cosh t}\left(\cosh{\frac{(n+1)t}{n}}+\cosh{\frac{(n-1)t}{n}}\right)dt
=12(Kn+1n(y)+Kn1n(y)).\displaystyle=-\frac{1}{2}(K_{\frac{n+1}{n}}(y)+K_{\frac{n-1}{n}}(y)).

By the estimate of KνK_{\nu}, we know that 1Ceyy<K1n(y)<Ceyy\frac{1}{C}\cdot\frac{e^{-y}}{\sqrt{y}}<-K_{\frac{1}{n}}^{\prime}(y)<C\cdot\frac{e^{-y}}{\sqrt{y}}. ∎

Corollary A.5.

For z>n4λ1nz>\sqrt[n]{\frac{n}{4\lambda_{1}}}, there exists a constant CC which only depends on nn such that

1Cλ14zn24e2nλ12zn2<\displaystyle\frac{1}{C}\cdot\lambda^{\frac{1}{4}}z^{\frac{n-2}{4}}e^{-\frac{2}{\sqrt{n}}\lambda^{\frac{1}{2}}\cdot z^{\frac{n}{2}}}<- 𝒟(z)<Cλ14zn24e2nλ12zn2,\displaystyle\mathcal{D}^{\prime}(z)<C\cdot\lambda^{\frac{1}{4}}z^{\frac{n-2}{4}}e^{-\frac{2}{\sqrt{n}}\lambda^{\frac{1}{2}}\cdot z^{\frac{n}{2}}},
1Cλ14zn24e2nλ12zn2<\displaystyle\frac{1}{C}\cdot\lambda^{\frac{1}{4}}z^{\frac{n-2}{4}}e^{\frac{2}{\sqrt{n}}\lambda^{\frac{1}{2}}\cdot z^{\frac{n}{2}}}<\quad 𝒢(z)<Cλ14zn24e2nλ12zn2.\displaystyle\mathcal{G}^{\prime}(z)<C\cdot\lambda^{\frac{1}{4}}z^{\frac{n-2}{4}}e^{\frac{2}{\sqrt{n}}\lambda^{\frac{1}{2}}\cdot z^{\frac{n}{2}}}.
Proof.

This can be seen directly from computing 𝒟\mathcal{D}^{\prime} and 𝒢\mathcal{G}^{\prime} with the substitution in (A.2).

𝒟(z)\displaystyle-\mathcal{D}^{\prime}(z) =12zK1n(2λnzn2)nλzn12K1n(2λnzn2),\displaystyle=-\frac{1}{2\sqrt{z}}K_{\frac{1}{n}}\left(2\sqrt{\tfrac{\lambda}{n}}\cdot z^{\frac{n}{2}}\right)-\sqrt{n\lambda}z^{\frac{n-1}{2}}K^{\prime}_{\frac{1}{n}}\left(2\sqrt{\tfrac{\lambda}{n}}\cdot z^{\frac{n}{2}}\right),
𝒢(z)\displaystyle\mathcal{G}^{\prime}(z) =12zI1n(2λnzn2)+nλzn12I1n(2λnzn2).\displaystyle=\frac{1}{2\sqrt{z}}I_{\frac{1}{n}}\left(2\sqrt{\tfrac{\lambda}{n}}\cdot z^{\frac{n}{2}}\right)+\sqrt{n\lambda}z^{\frac{n-1}{2}}I^{\prime}_{\frac{1}{n}}\left(2\sqrt{\tfrac{\lambda}{n}}\cdot z^{\frac{n}{2}}\right).

By Lemma A.4 we get the estimate. ∎

Consequently, by integration by parts as before we have the C1C^{1} estimate of uu:

|u(z)|\displaystyle|u^{\prime}(z)| =|2(𝒟(z)1z𝒢(s)sn1v(s)𝑑s+𝒢(z)z𝒟(s)sn1v(s)𝑑s)|C(n)C0zδ+n2λ12\displaystyle=\left|2\left(\mathcal{D}^{\prime}(z)\int_{1}^{z}\mathcal{G}(s)s^{n-1}v(s)ds+\mathcal{G}^{\prime}(z)\int_{z}^{\infty}\mathcal{D}(s)s^{n-1}v(s)ds\right)\right|\leqslant C(n)\cdot C_{0}\frac{z^{\delta+\frac{n}{2}}}{\lambda^{\frac{1}{2}}}

and

u′′(z)=nλzn2u+nzn1vC(n)C0zδ+n1\displaystyle u^{\prime\prime}(z)=n\lambda z^{n-2}u+nz^{n-1}v\leqslant C(n)\cdot C_{0}z^{\delta+n-1}

for z>C(n,δ,λ1)z>C(n,\delta,\lambda_{1}).
In the end, we get if |v(z)|C0zδ|v(z)|\leqslant C_{0}z^{\delta} on z>C(n,δ,M)z>C(n,\delta,M), then for any z>C(n,δ,M)z>C(n,\delta,M)

|u(z)|C(n)C0zδ+1λ,|u(z)|C(n)C0zδ+n2λ12,|u′′(z)|C(n)C0zδ+n1.\left|u(z)\right|\leqslant C(n)\cdot C_{0}\frac{z^{\delta+1}}{\lambda},\quad\left|u^{\prime}(z)\right|\leqslant C(n)\cdot C_{0}\frac{z^{\delta+\frac{n}{2}}}{\lambda^{\frac{1}{2}}},\quad\left|u^{\prime\prime}(z)\right|\leqslant C(n)\cdot C_{0}z^{\delta+n-1}.

A.2. fundamental solution of non-zero mode

In this section we focus on the non-zero mode: the equation

u′′(j2n24+nλ)zn2u=nzn1v.\displaystyle u^{\prime\prime}-(\frac{j^{2}n^{2}}{4}+n\lambda)z^{n-2}u=nz^{n-1}v. (A.5)

By [16] we have the decay solution 𝒟(z)\mathcal{D}(z) and growth solution 𝒢(z)\mathcal{G}(z) of the homogeneous equation u′′(z)=(j2n24+nλ)zn2uu^{\prime\prime}(z)=(\frac{j^{2}n^{2}}{4}+n\lambda)z^{n-2}u given by

𝒟(z)=ejzn2Ψ(β,α,jzn),\displaystyle\mathcal{D}(z)=e^{\frac{jz^{n}}{2}}\cdot\Psi^{\flat}\left(\beta,\alpha,-jz^{n}\right), (A.6)
𝒢(z)=ejzn2Φ(β,α,jzn),\displaystyle\mathcal{G}(z)=e^{\frac{jz^{n}}{2}}\cdot\Phi^{\sharp}\left(\beta,\alpha,-jz^{n}\right), (A.7)

where α=11n\alpha=1-\frac{1}{n}, β=n12nλnj0\beta=\frac{n-1}{2n}-\frac{\lambda}{nj}\leqslant 0, Φ(β,α,jzn)\Phi^{\sharp}\left(\beta,\alpha,-jz^{n}\right) and Ψ(β,α,jzn)\Psi^{\flat}\left(\beta,\alpha,-jz^{n}\right) have the following expression:

Ψ(β,α,y)\displaystyle\Psi^{\flat}(\beta,\alpha,y) =eyΓ(αβ)0eyssαβ1(1+s)β1𝑑s,\displaystyle=\frac{e^{y}}{\Gamma(\alpha-\beta)}\int_{0}^{\infty}e^{ys}s^{\alpha-\beta-1}(1+s)^{\beta-1}ds, (A.8)
Φ(β,α,y)\displaystyle\Phi^{\sharp}(\beta,\alpha,y) =Γ(α)Γ(αβ)ey(y)βα0esysα12βIα1(2s)𝑑s.\displaystyle=\frac{\Gamma(\alpha)}{\Gamma(\alpha-\beta)}\cdot e^{y}(-y)^{\beta-\alpha}\cdot\int_{0}^{\infty}e^{\frac{s}{y}}\cdot s^{\frac{\alpha-1}{2}-\beta}\cdot I_{\alpha-1}(2\sqrt{s})ds. (A.9)

Let Q=αβ1Q=\alpha-\beta-1, γn=12+1n\gamma_{n}=\frac{1}{2}+\frac{1}{n}. Denote

F(t)=yt+Qlogtt+1.F(t)=yt+Q\log\frac{t}{t+1}.

Then FF is strictly concave in \mathbb{R} if Q>0Q>0. Let t0t_{0} be the only critical point of FF. We have

t0=12(1+1+4Qy).t_{0}=\frac{1}{2}\left(-1+\sqrt{1+\frac{4Q}{-y}}\right).

Denote

G(u)=u2+2(y)12u+(2Q+γn)logu.G(u)=-u^{2}+2(-y)^{\frac{1}{2}}\cdot u+\left(2Q+\gamma_{n}\right)\log u.

Then GG is strictly concave in +\mathbb{R}_{+}. Let u0u_{0} be the only critical point of GG. Then

u0=(y)122(1+1+4Qy+2γny).u_{0}=\frac{(-y)^{\frac{1}{2}}}{2}\cdot\left(1+\sqrt{1+\frac{4Q}{-y}+\frac{2\gamma_{n}}{-y}}\right).

In [16] by Laplace method, we can show the following estimate:

Lemma A.6.

[16] There is a constant CC which only depends on nn such that
when Q1Q\geqslant 1,

Cn1Q1412n(y)1ey+F(t0)Γ(Q+1)\displaystyle C_{n}^{-1}\cdot Q^{-\frac{1}{4}-\frac{1}{2n}}\cdot\frac{(-y)^{-1}\cdot e^{y+F\left(t_{0}\right)}}{\Gamma(Q+1)}\leqslant Ψ(β,α,y)CnQ14ey+F(t0)Γ(Q+1),\displaystyle\Psi^{\flat}(\beta,\alpha,y)\leqslant C_{n}\cdot Q^{\frac{1}{4}}\cdot\frac{e^{y+F\left(t_{0}\right)}}{\Gamma(Q+1)},
Cn1Q14(y)2n4ney+G(u0)Γ(Q+1)\displaystyle C_{n}^{-1}\cdot Q^{-\frac{1}{4}}\cdot\frac{(-y)^{\frac{2-n}{4n}}\cdot e^{y+G\left(u_{0}\right)}}{\Gamma(Q+1)}\leqslant Φ(β,α,y)Cn(y)2n4ney+G(u0)Γ(Q+1);\displaystyle\Phi^{\sharp}(\beta,\alpha,y)\leqslant C_{n}\cdot\frac{(-y)^{\frac{2-n}{4n}}\cdot e^{y+G\left(u_{0}\right)}}{\Gamma(Q+1)};

when Q1Q\leqslant 1,

Cn1ey(y)βα\displaystyle C_{n}^{-1}\cdot e^{y}\cdot(-y)^{\beta-\alpha} Ψ(β,α,y)ey(y)βα,\displaystyle\leqslant\Psi^{\flat}(\beta,\alpha,y)\leqslant e^{y}\cdot(-y)^{\beta-\alpha},
Cn1(y)β\displaystyle C_{n}^{-1}\cdot(-y)^{-\beta} Φ(β,α,y)Cn(y)β\displaystyle\leqslant\Phi^{\sharp}(\beta,\alpha,y)\leqslant C_{n}\cdot(-y)^{-\beta}

for any y1y\leqslant-1.

Corollary A.7.

[16] There is a constant CC which only depends on nn such that
when Q1Q\geqslant 1,

C1Q1412nΓ(Q+1)ejzn2+F(t0(z))(jzn)1\displaystyle C^{-1}\cdot\frac{Q^{-\frac{1}{4}-\frac{1}{2n}}}{\Gamma(Q+1)}\cdot e^{-\frac{jz^{n}}{2}+F\left(t_{0}(z)\right)}\cdot\left(jz^{n}\right)^{-1}\leqslant\; 𝒟(z)CQ14Γ(Q+1)ejzn2+F(t0(z)),\displaystyle\mathcal{D}(z)\leqslant C\cdot\frac{Q^{\frac{1}{4}}}{\Gamma(Q+1)}\cdot e^{-\frac{jz^{n}}{2}+F\left(t_{0}(z)\right)},
C1Q14(jzn)2n4nΓ(Q+1)ejzn2+G(u0(z))\displaystyle C^{-1}\cdot Q^{-\frac{1}{4}}\cdot\frac{\left(jz^{n}\right)^{\frac{2-n}{4n}}}{\Gamma(Q+1)}\cdot e^{-\frac{jz^{n}}{2}+G\left(u_{0}(z)\right)}\leqslant\; 𝒢(z)C(jzn)2n4nΓ(Q+1)ejzn2+G(u0(z));\displaystyle\mathcal{G}(z)\leqslant C\cdot\frac{\left(jz^{n}\right)^{\frac{2-n}{4n}}}{\Gamma(Q+1)}\cdot e^{-\frac{jz^{n}}{2}+G\left(u_{0}(z)\right)};

when Q1Q\leqslant 1,

C1ejzn2(jzn)βα\displaystyle C^{-1}\cdot e^{-\frac{jz^{n}}{2}}\cdot\left(jz^{n}\right)^{\beta-\alpha}\leqslant\; 𝒟(z)Cejzn2(jzn)βα,\displaystyle\mathcal{D}(z)\leqslant C\cdot e^{-\frac{jz^{n}}{2}}\cdot\left(jz^{n}\right)^{\beta-\alpha},
C1ejzn2(jzn)β\displaystyle C^{-1}\cdot e^{\frac{jz^{n}}{2}}\cdot\left(jz^{n}\right)^{-\beta}\leqslant\; 𝒢(z)Cejzn2(jzn)β\displaystyle\mathcal{G}(z)\leqslant C\cdot e^{\frac{jz^{n}}{2}}\cdot\left(jz^{n}\right)^{-\beta}

for any z>1z>1.

We also need the following lemma

Lemma A.8.

For any z1z\geqslant 1, eF(t0(z))+G(u0(z))Cjn+24nzn+24ejzneQQQ+n+24ne^{F\left(t_{0}(z)\right)+G\left(u_{0}(z)\right)}\leqslant C\cdot j^{\frac{n+2}{4n}}z^{\frac{n+2}{4}}e^{jz^{n}}e^{-Q}Q^{Q+\frac{n+2}{4n}}.

Proof.

By similar straight forward computation as in [16]. ∎

With those estimates, we can give a C0C^{0} bound of u(z)u(z). By [16] we know that the Wronskian

𝒲(𝒢,𝒟)=𝒢(z)𝒟(z)𝒢(z)𝒟(z)=Γ(α1)Γ(αβ)j1n.\mathcal{W}(\mathcal{G},\mathcal{D})=\mathcal{G}(z)\mathcal{D}^{\prime}(z)-\mathcal{G}^{\prime}(z)\mathcal{D}(z)=\frac{\Gamma(\alpha-1)}{\Gamma(\alpha-\beta)}j^{\frac{1}{n}}.

Hence we have a solution of (A.5) as follows:

u(z)\displaystyle u(z) =Γ(αβ)nΓ(α1)j1n(𝒟(z)1z𝒢(s)sn1v(s)𝑑s+𝒢(z)z𝒟(s)sn1v(s)𝑑s).\displaystyle=\frac{\Gamma(\alpha-\beta)n}{\Gamma(\alpha-1)j^{\frac{1}{n}}}\left(\mathcal{D}(z)\int_{1}^{z}\mathcal{G}(s)s^{n-1}v(s)ds+\mathcal{G}(z)\int_{z}^{\infty}\mathcal{D}(s)s^{n-1}v(s)ds\right). (A.10)

Then we can have the following estimate of our solution:

Proposition A.9.

Recall that λ1\lambda_{1} is the first nonzero positive eigenvalue of ΔY-\Delta_{Y}. Let vv be a smooth function such that |v(z)|C0zδ|v(z)|\leqslant C_{0}z^{\delta} for z>1z>1. For any λ\lambda such that λ>λ1>0\lambda>\lambda_{1}>0, we can find solution of equation u′′(j2n24+nλ)zn2u=nzn1vu^{\prime\prime}-(\frac{j^{2}n^{2}}{4}+n\lambda)z^{n-2}u=nz^{n-1}v such that |u(z)|CC0zδ+1j2n24zn+nλ|u(z)|\leqslant C\cdot C_{0}\frac{z^{\delta+1}}{\frac{j^{2}n^{2}}{4}z^{n}+n\lambda}, |u′′(z)|CC0zδ+n1|u^{\prime\prime}(z)|\leqslant C\cdot C_{0}z^{\delta+n-1} on z>Cz>C for some constant C>1C>1 only depends on nn, λ1\lambda_{1} and aa.

Proof.

Similar as the zero-mode case, we estimate 𝒟(z)1z𝒢(s)sn1v(s)𝑑s+𝒢(z)z𝒟(s)sn1v(s)𝑑s\mathcal{D}(z)\int_{1}^{z}\mathcal{G}(s)s^{n-1}v(s)ds+\mathcal{G}(z)\int_{z}^{\infty}\mathcal{D}(s)s^{n-1}v(s)ds.
By integration by parts we have

𝒟(z)1z𝒢(s)sδ+n1𝑑s+𝒢(z)z𝒟(s)sδ+n1𝑑s\displaystyle\mathcal{D}(z)\int_{1}^{z}\mathcal{G}(s)s^{\delta+n-1}ds+\mathcal{G}(z)\int_{z}^{\infty}\mathcal{D}(s)s^{\delta+n-1}ds
=\displaystyle= 𝒲(𝒢,𝒟)k=0N1P(TP)k(zδ+n1)+𝒟(z)1z𝒢(s)(TP)N(sδ+n1)𝑑s+𝒢(z)z𝒟(s)(TP)N(sδ+n1)𝑑s,\displaystyle-\mathcal{W}(\mathcal{G},\mathcal{D})\sum_{k=0}^{N-1}P(TP)^{k}(z^{\delta+n-1})+\mathcal{D}(z)\int_{1}^{z}\mathcal{G}(s)(TP)^{N}(s^{\delta+n-1})ds+\mathcal{G}(z)\int_{z}^{\infty}\mathcal{D}(s)(TP)^{N}(s^{\delta+n-1})ds,

where P,T:C(+)C(+)P,T:C^{\infty}(\mathbb{R}_{+})\to C^{\infty}(\mathbb{R}_{+}) are given by

P(f)=fzn2(j2n24zn+nλ),T(f)=f′′.\displaystyle P(f)=\frac{f}{z^{n-2}(\frac{j^{2}n^{2}}{4}z^{n}+n\lambda)},\quad T(f)=f^{\prime\prime}.

By straight forward computation and induction we can see that

(TP)k(zδ+n1)C(k,n,δ)zδ+n12nkj2k.(TP)^{k}(z^{\delta+n-1})\leqslant C(k,n,\delta)\frac{z^{\delta+n-1-2nk}}{j^{2k}}.

So by taking NN large enough we have (TP)N(zδ+n1)C(n,δ,j)zM(TP)^{N}(z^{\delta+n-1})\leqslant C(n,\delta,j)z^{-M} for some M>2M>2 which will be chosen later.
We first consider the case that Q1Q\geqslant 1. By Lemma A.8 the first term becomes

𝒟(z)1z𝒢(s)(TP)N(sδ+n1)𝑑s\displaystyle\mathcal{D}(z)\int_{1}^{z}\mathcal{G}(s)(TP)^{N}(s^{\delta+n-1})ds\leqslant\; Cejzn2+F(t0(z))1z(jsn)2n4nsMejsn2+G(u0(s))𝑑s\displaystyle C\cdot e^{-\frac{jz^{n}}{2}+F\left(t_{0}(z)\right)}\int_{1}^{z}\left(js^{n}\right)^{\frac{2-n}{4n}}s^{-M}\cdot e^{-\frac{js^{n}}{2}+G\left(u_{0}(s)\right)}ds
\displaystyle\;\leqslant\; Cejzn+G(u0(z))+F(t0(z))j2n4nz2n4M+1C(n,Q)j1nz2M.\displaystyle C\cdot e^{-jz^{n}+G\left(u_{0}(z)\right)+F\left(t_{0}(z)\right)}\cdot j^{\frac{2-n}{4n}}\cdot z^{\frac{2-n}{4}-M+1}\leqslant C(n,Q)\cdot j^{\frac{1}{n}}\cdot z^{2-M}.

For the second term, we have similar estimate:

𝒢(z)z𝒟(s)(TP)N(sδ+n1)𝑑s\displaystyle\mathcal{G}(z)\int_{z}^{\infty}\mathcal{D}(s)(TP)^{N}(s^{\delta+n-1})ds\leqslant\; C(jzn)2n4nejzn2+G(t0(z))zsMejsn2+F(u0(s))𝑑s\displaystyle C\cdot\left(jz^{n}\right)^{\frac{2-n}{4n}}\cdot e^{-\frac{jz^{n}}{2}+G\left(t_{0}(z)\right)}\int_{z}^{\infty}s^{-M}e^{-\frac{js^{n}}{2}+F\left(u_{0}(s)\right)}ds
\displaystyle\;\leqslant\; Cejzn+G(u0(z))+F(t0(z))j2n4nz2n4M+1C(n,Q)j1nz2M.\displaystyle C\cdot e^{-jz^{n}+G\left(u_{0}(z)\right)+F\left(t_{0}(z)\right)}\cdot j^{\frac{2-n}{4n}}\cdot z^{\frac{2-n}{4}-M+1}\leqslant C(n,Q)\cdot j^{\frac{1}{n}}\cdot z^{2-M}.

Then we consider the case where Q1Q\leqslant 1

𝒟(z)1z𝒢(s)(TP)N(sδ+n1)𝑑sCejzn2(jzn)βα1zsMejsn2(jsn)β𝑑sCjαz1nαM,\displaystyle\mathcal{D}(z)\int_{1}^{z}\mathcal{G}(s)(TP)^{N}(s^{\delta+n-1})ds\leqslant\;C\cdot e^{-\frac{jz^{n}}{2}}\cdot\left(jz^{n}\right)^{\beta-\alpha}\int_{1}^{z}s^{-M}e^{\frac{js^{n}}{2}}\cdot\left(js^{n}\right)^{-\beta}ds\;\leqslant\;C\cdot j^{-\alpha}z^{1-n\alpha-M},
𝒢(z)z𝒟(s)(TP)N(sδ+n1)𝑑sCejzn2(jzn)βzsMejsn2(jsn)βα𝑑sCjαz1nαM.\displaystyle\mathcal{G}(z)\int_{z}^{\infty}\mathcal{D}(s)(TP)^{N}(s^{\delta+n-1})ds\leqslant\;C\cdot e^{\frac{jz^{n}}{2}}\cdot\left(jz^{n}\right)^{-\beta}\int_{z}^{\infty}s^{-M}e^{-\frac{js^{n}}{2}}\cdot\left(js^{n}\right)^{\beta-\alpha}ds\;\leqslant\;C\cdot j^{-\alpha}z^{1-n\alpha-M}.

So we have the uniform estimate for uu that for any z>C(n,δ)z>C(n,\delta),

|u(z)|Czδ+1j2n24zn+nλ.\displaystyle|u(z)|\leqslant C\frac{z^{\delta+1}}{\frac{j^{2}n^{2}}{4}z^{n}+n\lambda}.

In the end, we get if |v(z)|C0zδ|v(z)|\leqslant C_{0}z^{\delta} on z>C(n,δ,M)z>C(n,\delta,M), then for any z>C(n,δ,M)z>C(n,\delta,M)

|u(z)|C(n)C0zδ+1j2n24zn+nλ,|u(z)|C(n)C0zδ+n,|u′′(z)|C(n)C0zδ+n1.\left|u(z)\right|\leqslant C(n)\cdot C_{0}\frac{z^{\delta+1}}{\frac{j^{2}n^{2}}{4}z^{n}+n\lambda},\quad\left|u^{\prime}(z)\right|\leqslant C(n)\cdot C_{0}z^{\delta+n},\quad\left|u^{\prime\prime}(z)\right|\leqslant C(n)\cdot C_{0}z^{\delta+n-1}.

Remark A.10.

Even though the separation of variable method is very explicit, we can only get a bound of uu with respect to the polynomial growth order of vv rather than the function vv itself. This is mainly because the behavior of operator TT and PP is not clear for general function.

References

  • [1] Vestislav Apostolov and Charles Cifarelli. Hamiltonian 22-forms and new explicit calabi–yau metrics and gradient steady kähler–ricci solitons on n\mathbb{C}^{n}, 2023.
  • [2] Tristan C. Collins and Yang Li. Complete calabi-yau metrics in the complement of two divisors, 2022.
  • [3] Tristan C. Collins, Freid Tong, and Shing-Tung Yau. A free boundary monge-ampère equation and applications to complete calabi-yau metrics, 2024.
  • [4] Tristan C. Collins and Valentino Tosatti. A singular demailly–păun theorem. Comptes Rendus Mathematique, 354(1):91–95, 2016.
  • [5] Ronan J. Conlon and Hans-Joachim Hein. Asymptotically conical Calabi–Yau manifolds, I. Duke Mathematical Journal, 162(15):2855 – 2902, 2013.
  • [6] Ronan J Conlon and Frédéric Rochon. New examples of complete calabi-yau metrics on n\mathbb{C}^{n} for n3n\geqslant 3. arXiv preprint arXiv:1705.08788, 2017.
  • [7] Jeff Viaclovsky Hans-Joachim Hein, Song Sun and Ruobing Zhang. Asymptotically calabi metrics and weak fano manifolds. 2023.
  • [8] Hans-Joachim Hein. On gravitational instantons. 2010.
  • [9] Hans-Joachim Hein. Gravitational instantons from rational elliptic surfaces. Journal of the American Mathematical Society, 25(2):355–393, 2012.
  • [10] Hans-Joachim Hein, Song Sun, Jeff Viaclovsky, and Ruobing Zhang. Nilpotent structures and collapsing ricci-flat metrics on the k3 surface. Journal of the American Mathematical Society, 35:123–209, 2022.
  • [11] J. J. Kohn. Harmonic integrals on strongly pseudo-convex manifolds, i. Annals of Mathematics, 78(1):112–148, 1963.
  • [12] Yang Li. A new complete calabi–yau metric on 3\mathbb{C}^{3}. Inventiones mathematicae, 217(1):1–34, 2019.
  • [13] Daheng Min. Construction of higher dimensional alf calabi-yau metrics, 2023.
  • [14] Shigefumi Mori and Shigeru Mukai. Classification of fano 3-folds with b22b_{2}\geqslant 2. manuscripta mathematica, 36(2):147–162, 1981.
  • [15] Takeo Ohsawa. Vanishing theorems on complete kähler manifolds. Publications of The Research Institute for Mathematical Sciences, 20:21–38, 1984.
  • [16] Song Sun and Ruobing Zhang. A liouville theorem on asymptotically calabi spaces. Calculus of Variations and Partial Differential Equations, 60(3):103, 2021.
  • [17] Song Sun and Ruobing Zhang. Collapsing geometry of hyperkähler 4-manifolds and applications, 2022.
  • [18] Gábor Székelyhidi. Degenerations of n\mathbb{C}^{n} and Calabi–Yau metrics. Duke Mathematical Journal, 168(14):2651 – 2700, 2019.
  • [19] S Takayama. Simple connectedness of weak fano varieties. Journal of algebraic geometry, 9(2):403–407, 2000.
  • [20] Gang Tian and Shing-Tung Yau. Complete kähler manifolds with zero ricci curvature. i. Journal of the American Mathematical Society, 3(3):579–609, 1990.
  • [21] Craig van Coevering. A construction of complete ricci-flat kähler manifolds. arXiv: Differential Geometry, 2008.
  • [22] Shing-Tung Yau. On the ricci curvature of a compact kähler manifold and the complex monge-ampére equation, i. Communications on Pure and Applied Mathematics, 31(3):339–411, 1978.