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

Projective embedding of stably degenerating sequence of hyperbolic Riemann surfaces

Jingzhou Sun Department of Mathematics, Shantou University, Shantou City, Guangdong 515063,China [email protected]
Abstract.

Given a sequence of genus g2g\geq 2 curves converging to a punctured Riemann surface with complete metric of constant Gaussian curvature 1-1. we prove that the Kodaira embedding using orthonormal basis of the Bergman space of sections of a pluri-canonical bundle also converges to the embedding of the limit space together with extra complex projective lines.

The author is partially supported by NNSF of China no.11701353 and the STU Scientific Research Foundation for Talents no.130/760181.

1. Introduction

Let g\mathcal{M}_{g} be the moduli of smooth compact Riemann surfaces of genus gg. When g2g\geq 2, the Deligne-Mumford compactification [1] g¯\overline{\mathcal{M}_{g}} is the moduli of stable curves. Each smooth curve of genus gg carries an unique Poincaré metric with constant Gaussian curvature 1-1. If Cg¯C\in\overline{\mathcal{M}_{g}} is a singular stable curve, then by removing the nodes, the smooth part carries an unique complete hyperbolic metric with constant Gaussian curvature 1-1. And if a holomorphic family π:𝒞𝐃\pi:\mathcal{C}\to{\mathbf{D}} of compact smooth curves CtC_{t} degenerate to C=C0C=C_{0}, then the hyperbolic metrics is continuous on the vertical line bundle [12].

In this article, from the point view of the quantization framework by Donaldson [2, 3], we are interested in the convergence of the pluri-canonical Bergman embeddings of the family of hyperbolic surfaces in the complex projective spaces. More precisely, let (Cj,gj)(C_{j},g_{j}) be a sequence of genus g2g\geq 2 Riemann surfaces with Riemannian metric gjg_{j} of constant Gaussian curvature 1-1, that converges, in the topology of pointed Gromov-Hausdorff, to a Punctured Riemann surface (C0,g0)(C_{0},g_{0})(not necessarily connected) with a complete Riemannian metric g0g_{0} of constant Gaussian curvature 1-1. Let KCjK_{C_{j}} denote the canonical bundle of CjC_{j}, then KCjK_{C_{j}} is endowed with a Hermitian metric hjh_{j} defined by the Kähler form ωj\omega_{j} associated to gjg_{j}. We consider the Bergman space j,k\mathcal{H}_{j,k} consisting of L2L^{2}-integrable holomorphic sections of KCjkK_{C_{j}}^{k}. Then j,k\mathcal{H}_{j,k} is a finite-dimensional Hermitian space with the Hermitian product defined by

<s,t>=Cj(s,t)hjωj,<s,t>=\int_{C_{j}}(s,t)_{h_{j}}\omega_{j},

where, by abuse of notation, we still use hjh_{j} to denote the induced Hermitian metric on KCjkK_{C_{j}}^{k}. For kk large enough, a basis of j,k\mathcal{H}_{j,k} will induce a Kodaira embedding of CjC_{j} to Nk,{\mathbb{C}}{\mathbb{P}}^{N_{k}}, where Nk=dimj,k1N_{k}=\dim\mathcal{H}_{j,k}-1 is independent of j1j\geq 1. For j=0j=0, the dimension of j,k\mathcal{H}_{j,k} is smaller than that of j>0j>0. It is natural to consider the embedding induced by an orthonormal basis for j,k\mathcal{H}_{j,k}, which can be considered as a bridge from Kähler geometry to complex geometry [4, 10]. It is worth mentioning that after this article is finished, the author learned that Dong[5] recently proved that if a smooth family of hyperelliptic curves degenerate to a nodal curve, then their Bergman kernels also converges to the Bergman kernel of the nodal curve.

As the Gaussian curvature is 1-1, the degeneration of metrics can only be ”pinching a nontrivial loop”, namely a sequence of surfaces with growingly thiner and longer handles, with the central loops degenerating to points. So C0C_{0} has dd pairs of punctures, which will be called ends. And for kk large enough,the dimension of 0,k\mathcal{H}_{0,k} equals Nk+1dN_{k}+1-d. Now we can state our main theorem.

Theorem 1.1.

For kk large enough, we can choose an orthonormal basis for j,k\mathcal{H}_{j,k} for all j>0j>0, so that as jj\to\infty the image of the embedding

Φj,k:CjNk\Phi_{j,k}:C_{j}\to{\mathbb{C}}{\mathbb{P}}^{N_{k}}

induced by the orthonormal basis converges to the image of C0C_{0} under the embedding

Φ0,k:C0NkdNk,\Phi_{0,k}:C_{0}\to{\mathbb{C}}{\mathbb{P}}^{N_{k}-d}\subset{\mathbb{C}}{\mathbb{P}}^{N_{k}},

attached with dd pairs of linear 1{\mathbb{C}}{\mathbb{P}}^{1}’s. To each pair of the ends (pα,pα+d)(p_{\alpha},p_{\alpha+d}), a pair of linear 1{\mathbb{C}}{\mathbb{P}}^{1}’s are associated, and form a connected chain connecting the images of these two points.

It is interesting to mention that during the process of taking limit, the pair of linear 1{\mathbb{C}}{\mathbb{P}}^{1}’s are developed as a pair of bubbles. Also, we should mention that kk depends only on the geometry of C0C_{0}, and does not need to be too big by the results in [8].

The proof of this theorem makes heavy use of the methods we developed from [10] to [8]. And just as in [3], the main point is basically proving the convergence of the Bergman kernels. And we hope this result may shine a light on the study of the degeneration of higher dimensional projective manifolds [6, 7, 9].

The structure of this article is as follows. We will first quickly recall the necessary background for this article. Then we will calculate in the model for the thin handles, or ”the collar”, of the Riemann surfaces close to the limit. And in the end, we will finish the proof of the convergence of the pluri-canonical Bergman embeddings.

Acknowledgements. The author would like to thank Professor Song Sun for many very helpful discussions.

2. Punctured Model

The model 𝐃{\mathbf{D}}^{*} with the Poincaré metric

ωP=2idzdz¯|z|2(log|z|2)2\omega_{P}=\frac{2idz\wedge d\bar{z}}{|z|^{2}(\log|z|^{2})^{2}}

Take the local section of the canonical bundle e=dzze=\frac{dz}{z}, the local potential is

φP=log|e|2=log(log1|z|2)22\varphi_{P}=-\log|e|^{2}=-\log\frac{(\log\frac{1}{|z|^{2}})^{2}}{2}

We use the notation τ=log|z|\tau=-\log|z|, so φP=log(2τ2)\varphi_{P}=-\log(2\tau^{2}). We are interested in the L2L^{2}-norm of the sections zaek+1z^{a}e^{k+1} of K𝐃k+1K_{{\mathbf{D}}^{*}}^{k+1}, a+a\in{\mathbb{Z}}^{+}. So we have the following integrals

Ya=(2τ2)k+1|z|2aωPY_{a}=\int(2\tau^{2})^{k+1}|z|^{2a}\omega_{P}

We have

Ya=2k+2π0e2aτ+2klogτ𝑑τY_{a}=2^{k+2}\pi\int_{0}^{\infty}e^{-2a\tau+2k\log\tau}d\tau

We denote by ga(τ)=2aτ+2klogτg_{a}(\tau)=-2a\tau+2k\log\tau, then ga′′(t)=2kτ2g_{a}^{\prime\prime}(t)=-\frac{2k}{\tau^{2}}. So ga(τ)g_{a}(\tau) is a concave function which attains its only maximum at τa=ka\tau_{a}=\frac{k}{a}. We will use the following basic lemma from [8].

Lemma 2.1.

Let f(x)f(x) be a concave function. Suppose f(x0)<0f^{\prime}(x_{0})<0, then we have

x0ef(x)𝑑xef(x0)f(x0)\int_{x_{0}}^{\infty}e^{f(x)}dx\leq\frac{e^{f(x_{0})}}{-f^{\prime}(x_{0})}

We can use Laplace’s method and the lemma above to estimate

Ya2k+2πe2k+2klogkakπ2a2Y_{a}\approx 2^{k+2}\pi e^{-2k+2k\log\frac{k}{a}}\sqrt{\frac{k\pi}{2a^{2}}}

Of course, we can directly calculate the integral to get

Ya=2k+2π(2k)!(2a)2k+1,Y_{a}=2^{k+2}\pi\frac{(2k)!}{(2a)^{2k+1}},

But the idea of mass concentration is key to our arguments. The Bergman kernel of 𝐃{\mathbf{D}}^{*} is then

ρ0,k+1=22kτ2k+2π(2k)!(a)2k+1|z|2a\rho_{0,k+1}=\frac{2^{2k}\tau^{2k+2}}{\pi(2k)!}\sum(a)^{2k+1}|z|^{2a}

Let C0C_{0} be a punctured Riemann surface obtained by removing 2d2d points {pα}1α2d\{p_{\alpha}\}_{1\leq\alpha\leq 2d} from a compact Riemann surface. C0C_{0} is endowed with a complete Poincaré metric ω\omega with constant Gaussian curvature 1-1. ω\omega defines a Hermitian metric hh on the canonical bundle KC0K_{C_{0}}. Then, for any positive integer kk, we denote by k\mathcal{H}_{k} the space of holomorphic sections of KC0kK_{C_{0}}^{k} that are L2L^{2}-integrable, namely

C0|s|h2ω<.\int_{C_{0}}|s|_{h}^{2}\omega<\infty.

For each pαp_{\alpha}, there is a neighborhood UαU_{\alpha} with local coordinate zz so that ω=ωP\omega=\omega_{P} on Uα\pαU_{\alpha}\backslash p_{\alpha}. We can assume that UαU_{\alpha} contains the points satisfying |z|Rα|z|\leq R_{\alpha}. We note that the injective radius at the points |z|=Rα|z|=R_{\alpha} is about π4(logRα)2\frac{\pi}{4(\log R_{\alpha})^{2}}. For simplicity, we let RR be the minimum of the RαR_{\alpha}’s, 1α2d1\leq\alpha\leq 2d. Clearly, for the complement of 1α2dUα\cup_{1\leq\alpha\leq 2d}U_{\alpha} in C0C_{0}, there is a positive lower bound λ0\lambda_{0} for the injective radius. The basic conclusion of [8] is that for kk large enough, in the ”inside” of UαU_{\alpha} where τ=log|z|>k+1\tau=-\log|z|>\sqrt{k+1}, the Bergman kernel ρk+1\rho_{k+1} is very much like ρ0,k+1\rho_{0,k+1}, which is dominated by the terms ca|z|2ac_{a}|z|^{2a} for a<k3/4a<k^{3/4}. In particular, when τk\tau\geq k, ρ0,k+1\rho_{0,k+1} is dominated by c1|z|2c_{1}|z|^{2}. The sections for C0C_{0} corresponding to zaz^{a} in the model 𝐃{\mathbf{D}}^{*} is constructed as follows. We let zαz_{\alpha} denote the local coordinate zz on UαU_{\alpha}. Let eα=dzααe_{\alpha}=\frac{dz_{\alpha}}{\alpha} be the local frame of KC0K_{C_{0}}. Then zαaeαk+1z_{\alpha}^{a}e_{\alpha}^{k+1}, a1a\geq 1, are local sections of KC0k+1K_{C_{0}}^{k+1}. We choose and fix a cut-off function χ(r)\chi(r) that equals 11 for r<R/2r<R/2 and that equals 0 for r>2R/3r>2R/3. Then we denote by χα\chi_{\alpha} the function χ(|zα|)\chi(|z_{\alpha}|) defined on C0C_{0}. Then χαzαaeαk+1\chi_{\alpha}z_{\alpha}^{a}e_{\alpha}^{k+1} is a global smooth L2L^{2}-integrable section of KC0k+1K_{C_{0}}^{k+1}. We then take the orthogonal projection of this section into the space k+1\mathcal{H}_{k+1}, and then normalize the holomorphic section to be of norm 1, obtaining a section sα,ak+1s_{\alpha,a}\in\mathcal{H}_{k+1}. We denote by

V0={sα,a,1α2d,1a<k3/4}V_{0}=\{s_{\alpha,a},\quad 1\leq\alpha\leq 2d,\quad 1\leq a<k^{3/4}\}

For kk large enough, within r<R/4r<R/4, the sections sα,acazas_{\alpha,a}\approx\sqrt{c_{a}}z^{a} with relative error less than 1k2\frac{1}{k^{2}}.

We choose and fix an orthonormal basis W0={sj}W_{0}=\{s_{j}\} for the orthogonal complement V0k+1V_{0}^{\perp}\subset\mathcal{H}_{k+1}.

To obtain global sections of LkL^{k} from local ones, we will need to use Hörmander’s L2L^{2} estimate. The following lemma is well-known, see for example [11].

Lemma 2.2.

Suppose (M,g)(M,g) is a complete Kähler manifold of complex dimension nn, \mathcal{L} is a line bundle on MM with hermitian metric hh. If

2πiΘh+Ric(g),vv¯gC|v|g2\langle-2\pi i\Theta_{h}+Ric(g),v\wedge\bar{v}\rangle_{g}\geq C|v|^{2}_{g}

for any tangent vector vv of type (1,0)(1,0) at any point of MM, where C>0C>0 is a constant and Θh\Theta_{h} is the curvature form of hh. Then for any smooth \mathcal{L}-valued (0,1)(0,1)-form α\alpha on MM with ¯α=0\bar{\partial}\alpha=0 and M|α|2𝑑Vg\int_{M}|\alpha|^{2}dV_{g} finite, there exists a smooth \mathcal{L}-valued function β\beta on MM such that ¯β=α\bar{\partial}\beta=\alpha and

M|β|2𝑑Vg1C|α|2𝑑Vg\int_{M}|\beta|^{2}dV_{g}\leq\frac{1}{C}\int|\alpha|^{2}dV_{g}

where dVgdV_{g} is the volume form of gg and the norms are induced by hh and gg.

In the setting of this article, for a curve CjC_{j}, j0j\geq 0, with line bundle KCjk+1K_{C_{j}}^{k+1}, the constant is kk, independent of jj.

3. the Collar model

The model ε=(,ωε){\mathbb{C}}^{*}_{\varepsilon}=({\mathbb{C}}^{*},\omega_{\varepsilon}), where

ωε=fεidzdz¯2|z|2,\omega_{\varepsilon}=\frac{f_{\varepsilon}idz\wedge d\bar{z}}{2|z|^{2}},

where fεf_{\varepsilon} is a function depending only on |z||z|, satisfying the following conditions

  • \bullet

    fε>0f_{\varepsilon}>0;

  • \bullet

    fε(1)=ε2f_{\varepsilon}(1)=\varepsilon^{2};

  • \bullet

    fε(1)=0f^{\prime}_{\varepsilon}(1)=0;

  • \bullet

    Δlogfε=2fε|z|2\Delta\log f_{\varepsilon}=\frac{2f_{\varepsilon}}{|z|^{2}};

Clearly, such fεf_{\varepsilon} exists and is unique. Also, our choice of fεf_{\varepsilon} makes the metric have constant Gaussian Curvature 1-1. We denote by t=log|z|t=\log|z|, then by abuse of the notation, we consider fεf_{\varepsilon} as a function of tt. Then we have

Δzlogfε=d2logfεdt21|z|2\Delta_{z}\log f_{\varepsilon}=\frac{d^{2}\log f_{\varepsilon}}{dt^{2}}\frac{1}{|z|^{2}}

For simplicity, we will use f(t)f(t) to denote fεf_{\varepsilon}. Therefore, we have

(logf(t))′′=2f(t)(\log f(t))^{\prime\prime}=2f(t)

The first fundamental form of the metric is

I=f(t)dt2+f(t)dθ2I=f(t)dt^{2}+f(t)d\theta^{2}

Clearly, by the requirements on ff, the circle {z||z|=1}\{z||z|=1\} is a geodesic. Then we use the arc-length parameter uu for the tt-curves. Then by the curvature condition, we have

f(t)=ε2cosh2u,u(0)=0f(t)=\varepsilon^{2}\cosh^{2}u,\quad u(0)=0

and dtdu=1εcoshu\frac{dt}{du}=\frac{1}{\varepsilon\cosh u}. Therefore, we have

t=2εtan1[tanhu2]t=\frac{2}{\varepsilon}\tan^{-1}[\tanh\frac{u}{2}]

So we have

εt2=tan112coshu+1\frac{\varepsilon t}{2}=\tan^{-1}\sqrt{1-\frac{2}{\cosh u+1}}

Therefore when coshu\cosh u is large, we have the following estimation

t=π2ε1ε(coshu+1)+O(ε(ε(coshu+1))2)t=\frac{\pi}{2\varepsilon}-\frac{1}{\varepsilon(\cosh u+1)}+O(\frac{\varepsilon}{(\varepsilon(\cosh u+1))^{2}}) (1)

So it is natural to use the notations

τε=π2εt.\tau_{\varepsilon}=\frac{\pi}{2\varepsilon}-t.

We also use the frame e=dzze=\frac{dz}{z} for the canonical bundle, so we have

|e|2=2f|e|^{2}=\frac{2}{f}

We are interested in the L2L^{2}-norm of the sections zaek+1z^{a}e^{k+1} of Kk+1K_{{\mathbb{C}}^{*}}^{k+1}, aa\in{\mathbb{Z}}. So we have the following integrals

Iε,a=(2f)k+1|z|2aωεI_{\varepsilon,a}=\int(\frac{2}{f})^{k+1}|z|^{2a}\omega_{\varepsilon}

We have

Iε,a=2k+2πe2atklogf𝑑tI_{\varepsilon,a}=2^{k+2}\pi\int_{-\infty}^{\infty}e^{2at-k\log f}dt

We denote by ga(t)=2atklogfg_{a}(t)=2at-k\log f, then ga′′(t)=2kf(t)g_{a}^{\prime\prime}(t)=-2kf(t). So ga(t)g_{a}(t) is a concave function which attains its only maximum at tat_{a}. Write ua=u(ta)u_{a}=u(t_{a}), we have

εsinhua=ak\varepsilon\sinh u_{a}=\frac{a}{k}

We will assume that ε\varepsilon is very small compared to kkk^{-k}. So sinhua=akε\sinh u_{a}=\frac{a}{k\varepsilon} is very large. So f(ta)>a2k2f(t_{a})>\frac{a^{2}}{k^{2}}, and we can use Laplace’s method and lemma 2.1 to estimate

Iε,a2k+2πe2ataklogf(ta)π2kf(ta)I_{\varepsilon,a}\approx 2^{k+2}\pi e^{2at_{a}-k\log f(t_{a})}\sqrt{\frac{\pi}{2kf(t_{a})}}

And we have that the mass of Iε,aI_{\varepsilon,a} is concentrated within the neiborhood {|tta|<klogka}\{|t-t_{a}|<\frac{\sqrt{k}\log k}{a}\} with relative error less than klogk+3/2k^{-\log k+3/2}. Also, when ε\varepsilon is small,

Iε,a+1Iε,aeπ/εa2k+1(a+1)2k+1\frac{I_{\varepsilon,a+1}}{I_{\varepsilon,a}}\approx e^{\pi/\varepsilon}\frac{a^{2k+1}}{(a+1)^{2k+1}}

Therefore, the power series a>0|z|2aIε,a\sum_{a>0}\frac{|z|^{2a}}{I_{\varepsilon,a}} is very close to a multiple of the power series

a>0a2k+1e2aτε.\sum_{a>0}a^{2k+1}e^{-2a\tau_{\varepsilon}}.

Recall that the power series in the expression of ρ0,k+1\rho_{0,k+1} is also

a2k+1|z|2a=a2k+1e2aτ\sum a^{2k+1}|z|^{2a}=\sum a^{2k+1}e^{-2a\tau}

So by the same argument in [8], for t[0,t1]t\in[0,t_{1}] the Bergman kernel is dominated by [|z|2Iε,1+1Iε,0](2f)k+1[\frac{|z|^{2}}{I_{\varepsilon,1}}+\frac{1}{I_{\varepsilon,0}}](\frac{2}{f})^{k+1}, and, by symmetry, for t[t1,0]t\in[t_{-1},0] the Bergman kernel is dominated by [|z|2Iε,1+1Iε,0](2f)k+1[\frac{|z|^{-2}}{I_{\varepsilon,-1}}+\frac{1}{I_{\varepsilon,0}}](\frac{2}{f})^{k+1}. In particular, we have the following

Lemma 3.1.

For any holomorphic section ss of Kk+1K^{k+1}_{{\mathbb{C}}^{*}}, satisfying s=1\parallel s\parallel=1, we have

|s|2<ε(logεk)2k|s|^{2}<\varepsilon(\frac{\log\varepsilon}{k})^{2k}

when coshu(12εlogε,1εlogε)\cosh u\in(\frac{-1}{2\varepsilon\log\varepsilon},\frac{-1}{\varepsilon\log\varepsilon}).

Proof.

By symmetry, we can assume t>0t>0. For the right end of the interval, we only need to estimate the norms of zIε,1ek+1\frac{z}{I_{\varepsilon,1}}e^{k+1} and 1Iε,0ek+1\frac{1}{I_{\varepsilon,0}}e^{k+1} at tt where coshu=1εlogε\cosh u=\frac{-1}{\varepsilon\log\varepsilon}. For the first one, we have

|zIε,1ek+1|22kε4kπ3/2(logεk)2k|\frac{z}{I_{\varepsilon,1}}e^{k+1}|^{2}\approx\frac{\sqrt{2k}\varepsilon}{4k\pi^{3/2}}(\frac{\log\varepsilon}{k})^{2k}

For the second one, we have

|1Iε,0ek+1|22k4π3/2ε2k+1(logε)2k,|\frac{1}{I_{\varepsilon,0}}e^{k+1}|^{2}\approx\frac{\sqrt{2k}}{4\pi^{3/2}}\varepsilon^{2k+1}(\log\varepsilon)^{2k},

which is much smaller than the first one. For the left end of the interval, we have smaller norm for the section zIε,1ek+1\frac{z}{I_{\varepsilon,1}}e^{k+1}, and still very small norm for the section 1Iε,0ek+1\frac{1}{I_{\varepsilon,0}}e^{k+1}. Combining these estimates, we have proved the lemma. ∎

Assume CjC_{j} converges to C0C_{0} in the pointed Gromov-Hausdorff topology. For jj big enough, CjC_{j} has exactly dd closed geodesics whose arc length is less than λ0/4\lambda_{0}/4. We denote these circles by γj,α\gamma_{j,\alpha}, 1αd1\leq\alpha\leq d, and arc length of γj,α\gamma_{j,\alpha} by εj,α\varepsilon_{j,\alpha}. By rearranging the points pαp_{\alpha}, we can assume that 2πεj,α2\pi\varepsilon_{j,\alpha} converges to the pair (pα,pα+d)(p_{\alpha},p_{\alpha+d}) as jj\to\infty. Also for jj large enough, there a neighborhood Uj,αU_{j,\alpha}, usually referred to as a collar , of each γj,α\gamma_{j,\alpha} which is homeomorphic to an annulus. We define a map

hj,α:Uj,αεh_{j,\alpha}:U_{j,\alpha}\to{\mathbb{C}}^{*}_{\varepsilon}

with ε=εj,α\varepsilon=\varepsilon_{j,\alpha}, as following. Fix an isometry λ\lambda of γj,α\gamma_{j,\alpha} to the circle |z|=1|z|=1 in ε{\mathbb{C}}^{*}_{\varepsilon}. Then passing through each point qq on γj,α\gamma_{j,\alpha}, there is an unique geodesic lql_{q} orthogonal to γj,α\gamma_{j,\alpha}. Then we define hj,αh_{j,\alpha} to be the map that sends each such geodesic lql_{q} to the geodesic passing through λ(q)\lambda(q) and being orthogonal to the unit circle, preserving λ\lambda and the orientation. Since both surfaces have constant Gaussian curvature 1-1, hj,αh_{j,\alpha} is an isometry so long as the geodesics lql_{q} do not intersect each other. But since the curvature is negative, by the Gauss-Bonnet theorem, these geodesics can not intersect within Uj,αU_{j,\alpha}. Therefore, hj,αh_{j,\alpha} is also holomorphic. so we can use the coordinate zz from ε{\mathbb{C}}^{*}_{\varepsilon} as the holomorphic coordinate of Uj,αU_{j,\alpha}. By switching pαp_{\alpha} and pα+dp_{\alpha+d} if necessary, we can assume that the part |z|>1|z|>1 of Uj,αU_{j,\alpha} converges to a neighborhood of pαp_{\alpha} and the part |z|<1|z|<1 to that of pα+dp_{\alpha+d}. We can assume that Uj,α={1/M|z|M}U_{j,\alpha}=\{1/M\leq|z|\leq M\} and for jj large enough, we can assume that the injective radius at |z|=M|z|=M is larger than π4(log3R/4)2\frac{\pi}{4(\log 3R/4)^{2}}. We denote by Uj,α+U_{j,\alpha}^{+} the part of Uj,αU_{j,\alpha} with |z|>1|z|>1, similarly Uj,αU_{j,\alpha}^{-} the part with |z|<1|z|<1. We then define a map

φj,α:Uj,α+Uα\varphi_{j,\alpha}:U_{j,\alpha}^{+}\to U_{\alpha}

by sending εcoshu\varepsilon\cosh u to 12τ\frac{1}{2\tau} while preserving the circles {u=constant}\{u=\textit{constant}\}. Clearly, we are only preserving the length of the circles. By symmetry, we also have

φj,α+d:Uj,αUα+d.\varphi_{j,\alpha+d}:U_{j,\alpha}^{-}\to U_{\alpha+d}.

By our assumption on the injective radius, the image of φj,α\varphi_{j,\alpha} contains the circle |zα|=3R4|z_{\alpha}|=\frac{3R}{4}. On UαU_{\alpha}, the first fundamental form is

I0=1τ2(dτ2+dθ2)I_{0}=\frac{1}{\tau^{2}}(d\tau^{2}+d\theta^{2})

So the pull back

φj,αI0=tanh2udu2+(εcoshu)2dθ2\varphi_{j,\alpha}^{*}I_{0}=\tanh^{2}udu^{2}+(\varepsilon\cosh u)^{2}d\theta^{2}

is almost isometric to the metric

Ij=du2+(εcoshu)2dθ2,I_{j}=du^{2}+(\varepsilon\cosh u)^{2}d\theta^{2},

when uu is large. In particular, for the part where εcoshu1logε\varepsilon\cosh u\geq\frac{-1}{\log\varepsilon}, φj,α\varphi_{j,\alpha} converges to an isometry when jj\to\infty.

Let Uα(r)U_{\alpha}(r) denote the subset of UαU_{\alpha} consists of the points |zα|<r|z_{\alpha}|<r. Let F=C0\1α2dUα(2R3)F=C_{0}\backslash\cup_{1\leq\alpha\leq 2d}U_{\alpha}(\frac{2R}{3}), and let ψj:FCj\psi_{j}:F\to C_{j} be the diffeomorphism with its image. Since ψj\psi_{j} converges to an isometry as jj\to\infty, we can glue ψj1\psi_{j}^{-1} with the φj,α\varphi_{j,\alpha}’s, by rotating φj,α\varphi_{j,\alpha} if necessary, for jj large enough, to get a map

Gj:Cj\γj,αC0,G_{j}:C_{j}\backslash\cup\gamma_{j,\alpha}\to C_{0},

with the following properties:

  • \bullet

    GjG_{j} is a diffeomorphism of Cj\γj,αC_{j}\backslash\cup\gamma_{j,\alpha} with its image.

  • \bullet

    Gj=φj,αG_{j}=\varphi_{j,\alpha} for pφj,α1Uα(2R3)p\in\varphi_{j,\alpha}^{-1}U_{\alpha}(\frac{2R}{3}), 1α2d1\leq\alpha\leq 2d.

  • \bullet

    GjG_{j} is almost an isometry on Cj\1α2dφj,α1Uα(2R3)C_{j}\backslash\cup_{1\leq\alpha\leq 2d}\varphi_{j,\alpha}^{-1}U_{\alpha}(\frac{2R}{3}), and converges to an isometry when jj\to\infty.

For any conformal metric, the compatible complex structure JJ is just a counterclockwise rotation by π2\frac{\pi}{2}. We see that almost isometry implies almost holomorphic. Therefore Gj1KCjG_{j}^{-1*}K_{C_{j}} converges to KC0K_{C_{0}} as subbundles of TC0T_{C_{0}}\otimes{\mathbb{C}}. More precisely, let JjJ_{j} be the complex structure compatible with the Riemannian metric gjg_{j}, if the point-wise norm

supvTp,|v|g=1|gj(v,v)g(v,v)|<δ,\sup_{v\in T_{p},|v|_{g}=1}|g_{j}(v,v)-g(v,v)|<\delta,

then we have

supvTp,|v|g=1|Jj(v)J(v)|g<λδ\sup_{v\in T_{p},|v|_{g}=1}|J_{j}(v)-J(v)|_{g}<\lambda\delta

for some constant λ\lambda independent of pp and gg. We call the supremum above the pointwise distance from JjJ_{j} to JJ. Moreover, if gjg_{j} converges to gg in C2C^{2}-norm, then JjJ_{j} converges to JJ in C2C^{2}-norm also. If we denote by TJT_{J} the holomorphic tangent space with respect to JJ, then the orthogonal projection of TJjT_{J_{j}} to TJT_{J} is close to an isometry if JjJ_{j} is close to JJ. We identify TJjT_{J_{j}} with TjT_{j} via this orthogonal projection, similarly Kj=TJjK_{j}=T_{J_{j}}^{*} with K=TJK=T_{J}^{*}, which we will also call an orthogonal projection, for simplicity. Since the metric on the canonical bundle is defined by the Kähler form ω\omega, and ωj\omega_{j} converges to ω\omega, we have that the Chern connection j\nabla_{j} on KjK_{j} converges to the Chern connection \nabla on KK.

4. convergence of projective embedding

By assigning value 1 on γj,α\gamma_{j,\alpha}, we glue together the pull back GjχαG_{j}^{*}\chi_{\alpha} and Gjχα+dG_{j}^{*}\chi_{\alpha+d} to get a function denoted by χ~α\tilde{\chi}_{\alpha}, for 1αd1\leq\alpha\leq d. On each φj,α\varphi_{j,\alpha}, we also consider χ~αza\tilde{\chi}_{\alpha}z^{a} as global smooth sections of KCjk+1K^{k+1}_{C_{j}}. Then we repeat the construction of V0V_{0} by normalizing the orthogonal projection of χ~αza\tilde{\chi}_{\alpha}z^{a} onto j,k+1\mathcal{H}_{j,k+1}, and denote the result section by sj,α,as_{j,\alpha,a}, |a|<k3/4|a|<k^{3/4}. Then we denote by

Vj={sj,α,a,1αd,|a|<k3/4}V_{j}=\{s_{j,\alpha,a},\quad 1\leq\alpha\leq d,\quad|a|<k^{3/4}\}

We should remark here that the choice of the upper bound k3/4k^{3/4} is not necessary, it is purely a habit from [10]. Notice that the number of sections of VjV_{j} is larger than that of V0V_{0} by the number dd. Those extra sections are {sj,α,0}1αd\{s_{j,\alpha,0}\}_{1\leq\alpha\leq d}. We consider sj,α,as_{j,\alpha,a} as a smooth section of Gj1KCjk+1G_{j}^{-1*}K_{C_{j}}^{k+1} on image(Gj)image(G_{j}). We then define a piecewise smooth section s~j,α,a\tilde{s}_{j,\alpha,a} of KC0k+1K_{C_{0}}^{k+1} which equals the orthogonal projection of sj,α,as_{j,\alpha,a} to KC0k+1K_{C_{0}}^{k+1} on image(Gj)image(G_{j}), and equals 0 in the complement. For simplicity, we will say that sj,α,as_{j,\alpha,a} converges in some topology if s~j,α,a\tilde{s}_{j,\alpha,a} converges in that topology.

Lemma 4.1.

sj,α,as_{j,\alpha,a} converges to sα,as_{\alpha,a} for a>0a>0, to sα+d,as_{\alpha+d,-a} for a<0a<0, in L2L^{2}-norm, as jj\to\infty.

Proof.

By symmetry, we only have to prove for a>0a>0. By taking jj large enough, we can assume that pC0\Uα(ε(j,α))p\in C_{0}\backslash\cup U_{\alpha}(\varepsilon(j,\alpha)). Since when εcoshu1logε\varepsilon\cosh u\geq\frac{-1}{\log\varepsilon}, tanh2u=1(εlogε)2\tanh^{2}u=1-(\varepsilon\log\varepsilon)^{2} is very close to 1. So φj,α\varphi_{j,\alpha} is very close to an isometry. For simplicity, we still use ε\varepsilon for short for ε(j,α)\varepsilon(j,\alpha). Then we look at the integral

Ij,α,a\displaystyle I_{j,\alpha,a} =\displaystyle= 2k+2πχ~α21(εcoshu)2ke2at𝑑t\displaystyle 2^{k+2}\pi\int\tilde{\chi}_{\alpha}^{2}\frac{1}{(\varepsilon\cosh u)^{2k}}e^{2at}dt
=\displaystyle= 2k+2πeπa/εχ~α21(εcoshu)2ke2aτε𝑑τε\displaystyle 2^{k+2}\pi e^{\pi a/\varepsilon}\int\tilde{\chi}_{\alpha}^{2}\frac{1}{(\varepsilon\cosh u)^{2k}}e^{-2a\tau_{\varepsilon}}d\tau_{\varepsilon}

On C0C_{0}, we have

Jα,a=2k+2πχα2τ2ke2aτ𝑑τJ_{\alpha,a}=2^{k+2}\pi\int\chi_{\alpha}^{2}\tau^{2k}e^{-2a\tau}d\tau

For Ij,α,aI_{j,\alpha,a}, by lemma 2.1, we can truncate the part τε>logε\tau_{\varepsilon}>-\log\varepsilon by introducing a relative error <ε<\varepsilon. Also for Jα,aJ_{\alpha,a}, we can truncate the part τ>logε\tau>-\log\varepsilon by introducing a relative error <ε<\varepsilon. Then for the part τεlogε\tau_{\varepsilon}\leq-\log\varepsilon, χ~α21(εcoshu)2ke2aτε\tilde{\chi}_{\alpha}^{2}\frac{1}{(\varepsilon\cosh u)^{2k}}e^{-2a\tau_{\varepsilon}} converges to χα2τ2ke2aτ\chi_{\alpha}^{2}\tau^{2k}e^{-2a\tau} uniformly. Therefore Ij,α,aeπa/εI_{j,\alpha,a}e^{-\pi a/\varepsilon} converges to Jα,aJ_{\alpha,a} as jj\to\infty. Therefore, Ij,α,a1/2zαaI_{j,\alpha,a}^{-1/2}z_{\alpha}^{a} converges to Jα,a1/2zaJ_{\alpha,a}^{-1/2}z^{a} as jj\to\infty. Also for this part, the 1-form dzαzα\frac{dz_{\alpha}}{z_{\alpha}} converges uniformly to dzz\frac{dz}{z}. The way to get orthogonal projection onto holomorphic sections is to find the solutions of

¯v=¯Jα,a1/2za(dzz)k+1\bar{\partial}v=\bar{\partial}J_{\alpha,a}^{-1/2}z^{a}(\frac{dz}{z})^{k+1}

and

¯jvj=¯jIj,α,a1/2zαa(dzαzα)k+1\bar{\partial}_{j}v_{j}=\bar{\partial}_{j}I_{j,\alpha,a}^{-1/2}z_{\alpha}^{a}(\frac{dz_{\alpha}}{z_{\alpha}})^{k+1}

with minimal L2L^{2}-norms, where we denote by ¯j\bar{\partial}_{j} the ¯\bar{\partial} operator on CjC_{j}. And in order to prove the conclusion of the lemma, it suffices to prove that vjv_{j} converges to vv. Notice that ¯v\bar{\partial}v is supported within the annulus 2R3|z|3R4Uα\frac{2R}{3}\leq|z|\leq\frac{3R}{4}\subset U_{\alpha}. And by the mass concentration property of zaz^{a}, the mass

¯v<1k2.\parallel\bar{\partial}v\parallel<\frac{1}{k^{2}}.

Therefore

|v|2ω<1k3,\int|v|^{2}\omega<\frac{1}{k^{3}},

and vv is holomorphic outside the support of ¯v\bar{\partial}v. Since the Bergman kernel is dominated by Y11z\sqrt{Y_{1}^{-1}}z, and

|z|<εY11|z|2ω<ε,\int_{|z|<\varepsilon}Y_{1}^{-1}|z|^{2}\omega<\varepsilon,

we have

|z|<ε|v|2ω<ε\int_{|z|<\varepsilon}|v|^{2}\omega<\varepsilon

in every UαU_{\alpha}. We pull back the restriction vv to C0\Uα(ε(j,α))C_{0}\backslash\cup U_{\alpha}(\varepsilon(j,\alpha)) by GjG_{j}, and use cut-off functions κε\kappa_{\varepsilon} near the edges to get a global smooth section of Gj(KC0k+1)G_{j}^{*}(K_{C_{0}}^{k+1}), which is then projected to a smooth section of KCjk+1K_{C_{j}}^{k+1}, called v~j\tilde{v}_{j}. More explicitly, we define the κε\kappa_{\varepsilon} as a smooth function of τε\tau_{\varepsilon} that equals 1 for τε<logε\tau_{\varepsilon}<-\log\varepsilon and equals 0 for τε>2logε\tau_{\varepsilon}>-2\log\varepsilon. We can also assume that |κε|<2logε|\kappa_{\varepsilon}^{\prime}|<\frac{2}{-\log\varepsilon}. We have

v=2¯¯v+(k+1)v\nabla^{*}\nabla v=2\bar{\partial}^{*}\bar{\partial}v+(k+1)v

So

|v|2ω=2|¯v|2ω+(k+1)|v|2ω<3k2.\int|\nabla v|^{2}\omega=\int 2|\bar{\partial}v|^{2}\omega+(k+1)\int|v|^{2}\omega<\frac{3}{k^{2}}.

Therefore,

|jv~j|2ωj<4k2\int|\nabla_{j}\tilde{v}_{j}|^{2}\omega_{j}<\frac{4}{k^{2}}
Cj|¯jv~j¯jvj|2ωj=δj,\int_{C_{j}}|\bar{\partial}_{j}\tilde{v}_{j}-\bar{\partial}_{j}v_{j}|^{2}\omega_{j}=\delta_{j},

where δj0\delta_{j}\to 0 as jj\to\infty. Then we can solve the equation

¯ju=¯jv~j¯jvj\bar{\partial}_{j}u=\bar{\partial}_{j}\tilde{v}_{j}-\bar{\partial}_{j}v_{j}

with minimal L2L^{2}-norm, so that

Cj|u|2ωj1kδj\int_{C_{j}}|u|^{2}\omega_{j}\leq\frac{1}{k}\delta_{j}

Since vjv_{j} is a minimal solution,

|v~ju|2ωj|vj|2ωj.\int|\tilde{v}_{j}-u|^{2}\omega_{j}\geq\int|v_{j}|^{2}\omega_{j}.

So

|v~j|2ωj|vj|2ωjδjk5.\int|\tilde{v}_{j}|^{2}\omega_{j}\geq\int|v_{j}|^{2}\omega_{j}-\sqrt{\frac{\delta_{j}}{k^{5}}}.

Conversely, for each vjv_{j}, we first use the cut off functions κε\kappa_{\varepsilon} to make it vanish near the edges, then we pull it back by Gj1G_{j}^{-1} to C0C_{0}. By orthogonal projection and extension by 0, we obtain a smooth section v~j\tilde{v}^{j} of KC0k+1K_{C_{0}}^{k+1}. Similar to v~j\tilde{v}_{j}, by lemma 3.1, we have

C0|¯v~j¯v|2ωj=δj,\int_{C_{0}}|\bar{\partial}\tilde{v}^{j}-\bar{\partial}v|^{2}\omega_{j}=\delta_{j}^{\prime},

where δj0\delta_{j}^{\prime}\to 0 as jj\to\infty. Then by solve a ¯\bar{\partial}-equation again, we get that

|v~j|2ω|v|2ωδjk5.\int|\tilde{v}^{j}|^{2}\omega\geq\int|v|^{2}\omega-\sqrt{\frac{\delta_{j}^{\prime}}{k^{5}}}.

Since the L2L^{2} norm of v~j\tilde{v}_{j} is close to that of vv, and

|v~j|2ω|vj|2ωj+δj′′,\int|\tilde{v}^{j}|^{2}\omega\leq\int|v_{j}|^{2}\omega_{j}+\delta_{j}^{\prime\prime},

where δj0\delta_{j}^{\prime}\to 0 as jj\to\infty, we get that

|v~j|2ωj|vj|2ωj0,\int|\tilde{v}_{j}|^{2}\omega_{j}-\int|v_{j}|^{2}\omega_{j}\to 0,

as jj\to\infty. Then by the uniqueness of the minimal solution of the ¯\bar{\partial}-equation, we get that

|v~jvj|2ωj0\int|\tilde{v}_{j}-v_{j}|^{2}\omega_{j}\to 0

as jj\to\infty. Finally, since the L2L^{2}-norm of sj,α,as_{j,\alpha,a} on the area where 1εcoshu<12logε\frac{1}{\varepsilon\cosh u}<\frac{-1}{2\log\varepsilon} also goes to 0 as jj\to\infty, we have proved the theorem.

The same ideas can be used for the sections in W0W_{0}. For each sW0s\in W_{0}, we have

|z|<ε|s|2ω<ε\int_{|z|<\varepsilon}|s|^{2}\omega<\varepsilon

in every UαU_{\alpha}. We pull back the restriction ss to C0\Uα(ε(j,α))C_{0}\backslash\cup U_{\alpha}(\varepsilon(j,\alpha)) by GjG_{j}, and use cut-off functions κε\kappa_{\varepsilon} near the edges to get a global smooth section of Gj(KC0k+1)G_{j}^{*}(K_{C_{0}}^{k+1}), which is then projected to smooth section of KCjk+1K_{C_{j}}^{k+1}, called s~j\tilde{s}_{j}. Then we have

Cj|¯js~j|2ωj0\int_{C_{j}}|\bar{\partial}_{j}\tilde{s}_{j}|^{2}\omega_{j}\to 0

as jj\to\infty. So we can solve the equation

¯juj=¯js~j\bar{\partial}_{j}u_{j}=\bar{\partial}_{j}\tilde{s}_{j}

with minimal L2L^{2}-norm to get a holomorphic section s,j=uj+s~jj,k+1s_{,j}=u_{j}+\tilde{s}_{j}\in\mathcal{H}_{j,k+1} satisfying Cj|s,j|2ωj10\int_{C_{j}}|s_{,j}|^{2}\omega_{j}-1\to 0 and s,jss_{,j}\to s in L2L^{2}-norm, as jj\to\infty. So we have proved the following:

Lemma 4.2.

For each sW0s\in W_{0}, we can find s,jj,k+1s_{,j}\in\mathcal{H}_{j,k+1}, so that s,js_{,j} converges to ss in the L2L^{2}-norm, and s,j1\parallel s_{,j}\parallel\to 1 as jj\to\infty.

We will denote by WjW_{j} the set of sections {s,j|sW0}\{s_{,j}|s\in W_{0}\}. Notice that WjW_{j} is not an orthonormal set, but gets closer as jj gets larger. We fix an order on W0W_{0}, the order each WjW_{j} accordingly. We give each VjV_{j} the dictionary order. Recall that sj,α,as_{j,\alpha,a} corresponds to sα,as_{\alpha,a} for a>0a>0, sα+d,as_{\alpha+d,-a} for a<0a<0. Then we add to V0V_{0} dd 0’s, and order the obtained set V~0\tilde{V}_{0} according to the correspondence to VjV_{j}, where each 0 corresponds to a section sj,α,0s_{j,\alpha,0}, 1αd1\leq\alpha\leq d. Then we define the embedding

Φj:CjNk,\Phi_{j}:C_{j}\to{\mathbb{C}}{\mathbb{P}}^{N_{k}},

where Nk=dimj,k+11N_{k}=\dim\mathcal{H}_{j,k+1}-1, by Φj=[Vj,Wj]\Phi_{j}=[V_{j},W_{j}], where [][\cdots] means the homogeneous coordinates. Similarly we define the embedding

Φ0:C0Nk,\Phi_{0}:C_{0}\to{\mathbb{C}}{\mathbb{P}}^{N_{k}},

by Φ0=[V~0,W0]\Phi_{0}=[\tilde{V}_{0},W_{0}].

Let [Z0,,ZN][Z_{0},\cdots,Z_{N}] be the homogeneous coordinates of N{\mathbb{C}}{\mathbb{P}}^{N}. U0={[1,w],wN}U_{0}=\{[1,w],w\in{\mathbb{C}}^{N}\} is a coordinate patch with wi=ZiZ0w_{i}=\frac{Z_{i}}{Z_{0}}. The ZiZ_{i}’s can be identified as generating sections in H0(N,𝒪(1))H^{0}({\mathbb{C}}{\mathbb{P}}^{N},\mathcal{O}(1)). In particular, Z0Z_{0} is a local frame in U0U_{0}. Then on U0U_{0}, the Fubini-Study form ω=i2¯log(1+|w|2)\omega=\frac{i}{2}\partial\bar{\partial}\log(1+|w|^{2}) has the following explicit form

ω=i2(1+|w|2)dwidw¯i(w¯idwi)(widw¯i)(1+|w|2)2\omega=\frac{i}{2}\frac{(1+|w|^{2})\sum dw^{i}\wedge d\bar{w}_{i}-(\sum\bar{w}_{i}dw_{i})(\sum w_{i}d\bar{w}_{i})}{(1+|w|^{2})^{2}}

On each Uj,αU_{j,\alpha}, within the area 0tt0=π2ε+logε0\leq t\leq t_{0}=\frac{\pi}{2\varepsilon}+\log\varepsilon, the image under Φj\Phi_{j} is dominated by the two sections sj,α,0s_{j,\alpha,0} and sj,α,1s_{j,\alpha,1}, since the contribution of other sections is of relative size <k2ε<k^{2}\varepsilon. We can estimate the map [sj,α,0,sj,α,1,0,][s_{j,\alpha,0},s_{j,\alpha,1},0,\cdots] by the local sections [b0,b1z,0,][b_{0},b_{1}z,0,\cdots], with relative error <1k2<\frac{1}{k^{2}}, where b02=2k+2π3/22kε2k1b_{0}^{-2}=\frac{2^{k+2}\pi^{3/2}}{\sqrt{2k}}\varepsilon^{-2k-1} and b12=2k+2π3/22keπ/εk2k+1b_{1}^{-2}=\frac{2^{k+2}\pi^{3/2}}{\sqrt{2k}}e^{\pi/\varepsilon}k^{2k+1}. So the map is simplified to

[1,eπ2ε(εk)k1/2z,0,][1,e^{-\frac{\pi}{2\varepsilon}}(\varepsilon k)^{-k-1/2}z,0,\cdots]

So we can estimate the length of the image of γj,α\gamma_{j,\alpha}, which is approximately 2πeπ2ε(εk)k1/22\pi e^{-\frac{\pi}{2\varepsilon}}(\varepsilon k)^{-k-1/2}, and goes to 0 as jj\to\infty. So the image of γj,α\gamma_{j,\alpha} converges to [1,0,][1,0,\cdots] in the current ordering of coordinates. Similarly, we can estimate the length of the image of the circle t0=π2ε+logεt_{0}=\frac{\pi}{2\varepsilon}+\log\varepsilon, which is approximately 2πε(εk)k+1/2\frac{2\pi}{\varepsilon}(\varepsilon k)^{k+1/2}, which goes to 0 as jj\to\infty. Therefore, the image of the circle t0=π2ε+logεt_{0}=\frac{\pi}{2\varepsilon}+\log\varepsilon converges to [0,1,][0,1,\cdots] in the current ordering of coordinates. Notice that eπ2ε(εk)k1/2|z|e^{-\frac{\pi}{2\varepsilon}}(\varepsilon k)^{-k-1/2}|z| goes to \infty at t0t_{0}, so the image of the area 0tt0=π2ε+logε0\leq t\leq t_{0}=\frac{\pi}{2\varepsilon}+\log\varepsilon converges to the 1{\mathbb{C}}{\mathbb{P}}^{1} connecting the points [1,0,0,][1,0,0,\cdots] and [0,1,0,][0,1,0,\cdots] in the current ordering of coordinates. This area is the bubble mentioned in the introduction. By symmetry, the image of the area 0tt00\geq t\geq-t_{0} also converges to a 1{\mathbb{C}}{\mathbb{P}}^{1}.

For the part Gj1(C0\Uα(ε(j,α)))G_{j}^{-1}(C_{0}\backslash\cup U_{\alpha}(\varepsilon(j,\alpha))), the sections {sj,α,0}1αd\{s_{j,\alpha,0}\}_{1\leq\alpha d} are negligible. So the convergence of the remaining sections in VjV_{j} and WjW_{j} in the L2L^{2} sense implies that the image of Gj1(C0\Uα(ε(j,α)))G_{j}^{-1}(C_{0}\backslash\cup U_{\alpha}(\varepsilon(j,\alpha))) under Φj\Phi_{j} converges to the image of Φ0\Phi_{0}. To conclude the proof of the main theorem, we only have to notice that although VjWjV_{j}\cup W_{j} is not orthonormal, we can modify them. The sections in WjW_{j} are almost orthonormal, so we can transform them to be orthonormal with a matrix AjA_{j} whose difference from the identity matrix goes to 0 as jj\to\infty. And the modified sections still converge to sections in W0W_{0} in L2L^{2}. We first apply a Gram-Schmidt process to the set V0V_{0}, then we apply a Gram-Schmidt process following the order of V0V_{0} to the set Vj\{sj,α,0}1αdV_{j}\backslash\{s_{j,\alpha,0}\}_{1\leq\alpha d}. Finally, since the sections {sj,α,0}1αd\{s_{j,\alpha,0}\}_{1\leq\alpha d} are almost orthonormal and almost orthogonal to the other section in VjV_{j}, we can modify the new VjV_{j} again with a matrix BjB_{j} whose difference from the identity matrix goes to 0 as jj\to\infty, to get a orthonormal set. And we have proved the main theorem.

References

  • [1] P. Deligne and D. Mumford. The irreducibility of the space of curves of given genus. Inst. Hautes Études Sci. Publ. Math., (36):75–109, 1969.
  • [2] Simon Donaldson. Scalar curvature and projective embeddings, I. Journal of Differential Geometry, 59(3):479–522, 2001.
  • [3] Simon Donaldson and Song Sun. Gromov–hausdorff limits of kähler manifolds and algebraic geometry. Acta Mathematica, 213(1):63–106, 2014.
  • [4] Simon Donaldson and Song Sun. Gromov-Hausdorff limits of Kähler manifolds and algebraic geometry, II. J. Differential Geom., 107(2):327–371, 2017.
  • [5] Robert Xin Dong. Boundary asymptotics of the relative bergman kernel metric for curves. arXiv:2005.11826.
  • [6] Shouhei Honda, Song Sun, and Ruobing Zhang. A note on the collapsing geometry of hyperkähler four manifolds. Sci. China Math., 62(11):2195–2210, 2019.
  • [7] Jian Song. Degeneration of kahler-einstein manifolds of negative scalar curvature. arXiv:1706.01518.
  • [8] Jingzhou Sun. Estimations of the Bergman kernel of the punctured disk. ArXiv e-prints, 2017.
  • [9] Jingzhou Sun. Projective embedding of pairs and logarithmic K-stability. Math. Ann., 375(3-4):1307–1336, 2019.
  • [10] Jingzhou Sun and Song Sun. Projective embedding of log Riemann surfaces and K-stability. ArXiv e-prints, May 2016.
  • [11] Gang Tian. On a set of polarized Kähler metrics on algebraic manifolds. Journal of Differential Geometry, 32(1990):99–130, 1990.
  • [12] Scott A. Wolpert. The hyperbolic metric and the geometry of the universal curve. J. Differential Geom., 31(2):417–472, 1990.