1 Introduction

Since it was established in [28], the Ohsawa–Takegoshi extension theorem turned out to be a fundamental tool in complex geometry. As of today, there are uncountable many proofs and refinements of the original result and even more applications to both complex analysis and algebraic geometry. Very roughly, the set-up is as follows: u is a canonical form defined on a sub-variety \(Y\subset X\) with values in a Hermitian bundle\(F\rightarrow X\). We are interested in the following two main questions:

\(Q_1.\):

Does the section u extend to X?

\(Q_2.\):

If the answer to the previous question is “yes,” can one construct an extension whose \(L^2\) norm is bounded by the \(L^2\) norm of u, up to an universal constant?

If Y is non-singular, then the results in e.g., [23] give—practically optimal—curvature conditions for the bundle F such that the answer to both questions above is affirmative. We refer to the articles [2, 3, 5, 8, 11, 13, 14, 17, 24,25,26,27, 29,30,31,32,33,34, 36] for many interesting developments and applications.

The case of a singular sub-variety Y turns out to be significantly more difficult and the most complete results obtained so far only treat the qualitative aspect of the extension problem, that is to say the question \(Q_1\), cf. [7].

In this article we are concerned with the question \(Q_2\). We obtain a few quantitative results for extension of twisted forms defined on sub-varieties Y, which have simple normal crossings. Our main motivation is the Conjecture in [14]. To begin with we fix some notations/conventions.

Let X be a non-singular, projective manifold and let \(Y:= \sum _{i=1}^N Y_i\) be a divisor with simple normal crossings. Let \((L, h_L)\) be a Hermitian line bundle on X, endowed with a metric \(h_L\). The following assumptions will be in force throughout this article:

  1. (a)

    The usual curvature requirements are satisfied

    $$\begin{aligned} \Theta _{h_L}(L)\geqslant 0, \quad \Theta _{h_L}(L)\geqslant \delta \Theta _{h_Y} (Y), \end{aligned}$$

    where \(\delta > 0\) is a positive real number and \(h_Y\) is a smooth metric on the bundle corresponding to \({\mathcal {O}}(Y)\). Let s be the canonical section of \({\mathcal {O}}(Y)\) with the normalization condition \(|s|_{h_Y} ^2 \leqslant e^{-\delta }\).

  2. (b)

    We write locally \(h_L =e^{-\varphi _L}\) and \(h_Y = e^{-\varphi _Y}\). The singularities of the metric \(h_L\) of L are of the following type:

    $$\begin{aligned} \varphi _L= \sum _j r_j\log |f_j|^2+ \tau _L, \end{aligned}$$

    where \(f_j\) are local holomorphic functions such that they are not identically zero when restricted to any of the components of Y and \(r_j>0\) is positive number. Moreover, we assume that \(\tau _L\) is non-singular.

  3. (c)

    Let \(u\in H^0\big (X, (K_X+ Y+ L)\otimes {\mathcal {O}}_X/{\mathcal {O}}_X(-Y)\big )\) be a twisted canonical form defined over Y. There exists a covering \((\Omega _i)\) of X with coordinate sets such that the restriction of the section \(\displaystyle u|_{\Omega _i}\) of u admits an extension \(U_i \in H^0 (\Omega _i, K_X+ Y+ L)\), which belongs to the multiplier ideal of \(h_L\), i.e.,

    $$\begin{aligned} \int _{\Omega _i}|U_i|^2e^{-\varphi _L- \varphi _Y}< + \infty . \end{aligned}$$
    (1.0.1)

We note that near a non-singular point of Y the existence of \(U_i\) follows from the usual \(L^2\) hypothesis of OT theorem provided that u belongs to the multiplier ideal sheaf of \(h_L|_Y\). But this may no longer be true in a neighborhood of singular point of Y.

In addition to the natural hypotheses (a), (b), and (c) above we collect next two other requirements we need to impose for some of our statements to hold.

  1. (i)

    We assume that there exists an open subset \(V_{\textrm{sing}}\) of X containing the singularities of Y such that the following hold.

(i.\(\alpha \)):

There exists a snc divisor \(W= \sum W_j\) on X such that the singularities of the restriction of the metric \(h_L\) of L to \(V_{\textrm{sing}}\) are as follows:

$$\begin{aligned} \varphi _L= \sum _j \left( 1- \frac{1}{k_j}\right) \log |z_j|^2+ \tau _L, \end{aligned}$$

where \(k_j\) are positive integers, and \(z_j\) are the local equations of W. The local weight \(\tau _L\) above is assumed to be bounded.

(i.\(\beta \)):

The curvature of the restriction of \(h_L|_{V_{\textrm{sing}}}\) is greater than \(\displaystyle C_{\textrm{sing}}\omega _{{\mathcal {C}}}|_{V_{\textrm{sing}}}\), where \(C_\textrm{sing}\) is a positive constant, and \(\omega _{{\mathcal {C}}}\) is a fixed Kähler metric with conic singularities on \(\big (X, \sum (1-1/k_i)W_i\big )\), i.e., locally we have

$$\begin{aligned} \omega _{\mathcal {C}}\simeq \sum _{i=1}^r\frac{\sqrt{-1}dz_i\wedge \textrm{d}\overline{z}_i}{|z_i|^{2-\frac{2}{k_i}}} + \sum _{i\geqslant r+1}\sqrt{-1}dz_i\wedge \textrm{d}\overline{z}_i \end{aligned}$$
(1.0.1)

that is to say, \(\omega _{\mathcal {C}}\) is quasi-isometric with the RHS of (1.0.1), where \(z_1\dots z_r=0\) is the local equation of the divisor \(W=\sum _{i=1}^r W_i\).

  1. (ii)

    There exists an open subset \(V_{\textrm{sing}}\) of X containing the singularities of Y such that the curvature of the restriction of \(h_L|_{V_{\textrm{sing}}}\) is identically zero.

In this context our first result states as follows:

Theorem 1.1

We assume that the metric \(h_L= e^{-\varphi _L}\) of L and the section u verifies the requirements (a), (b), (c) as well as (i) above. Then u extends to X, and for each \(1\geqslant \alpha > 0\) there exists a section \(U \in H^0 (X, K_X +Y+L)\) such that \(U |_Y =u\) and we have the estimates

$$\begin{aligned} \frac{1}{C}\int _{X\setminus V_{\textrm{sing}}} |U|^2e^{-\varphi _Y- \varphi _L}\textrm{d}V_{\omega _{{\mathcal {C}}}}&\leqslant \int _{Y\setminus V_\textrm{sing}}\left| \frac{u}{\textrm{d}s}\right| ^2e^{-\varphi _L}\nonumber \\&\quad + \sum _i\left( \int _{Y_i\cap V_{\textrm{sing}}} \left| \frac{u}{\textrm{d}s}\right| _{\omega _{\mathcal C}}^{\frac{2}{1+\alpha }}e^{-\frac{\varphi _L}{1+ \alpha }}\textrm{d}V_{\omega _{{\mathcal {C}}}}\right) ^{1+\alpha } \end{aligned}$$
(1.1.1)

where \(\omega _{{\mathcal {C}}}\) is the reference metric on X and the constant C depends on \(\alpha \), the geometry of \((V_{\textrm{sing}}, \omega _{{\mathcal {C}}})\), the positivity constant and the upper bound for \(\textrm{Tr} \,_{\omega _{{\mathcal {C}}}}dd^c\tau _L\) in (i.\(\alpha \)).

Remark 1.2

The precise dependence of the constant C in (1.1.1) of the quantities mentioned in Theorem 1.1 can be easily extracted from the proof we present in Sect. 5.

Remark 1.3

It is very likely that our arguments work under more general circumstances, e.g., one can probably establish the same result in the absence of the hypothesis (b) (via the regularization procedure due to Demailly, cf. [12]). But so far it is unclear to us how to remove the local strict positivity hypothesis in (i.\(\beta \)), or the fact that the singularities of \(\displaystyle h_L|_{V_{\textrm{sing}}}\) are assumed to be of conic type.

A result pointing in the direction of the preceding Remark 1.3 is the following.

Theorem 1.4

Let X be a projective manifold and let \((L, h_L)\) be a line bundle such that the usual conditions (a)–(c) are satisfied, respectively. We assume moreover that \(\Theta _{h_L} (L) \leqslant C_1 \omega \) on \(V_{\textrm{sing}}\) for some constant \(C_1\). Here \(\omega \) is a non-singular Kähler metric on X. Then there exists a section \(U \in H^0 (X, K_X +Y+L)\) such that \(U |_Y =u\) with the estimates

$$\begin{aligned} \frac{1}{C}\int _{X\setminus V_{\textrm{sing}}} |U|^2e^{-\varphi _Y- \varphi _L} \leqslant \int _{Y\setminus V_{\textrm{sing}}}\left| \frac{u}{\textrm{d}s}\right| ^2e^{-\varphi _L} + \left( \int _{Y\cap V_{\textrm{sing}}} \left| \frac{u}{\textrm{d}s}\right| e^{-\frac{\varphi _L}{2}}\textrm{d}V_{\omega }\right) ^2 \end{aligned}$$

where C depends only on the geometry of \((V_{\textrm{sing}}, \omega )\) and \(C_1\).

Remark 1.5

We stress on the fact that in the statement Theorem 1.4 the strict positivity of the curvature of \((L, h_L)\) on \(V_\textrm{sing}\) is not part of hypothesis.

In conclusion, Theorems 1.1 and 1.4 are providing an extension of u whose \(L^2\) norm is bounded by the usual quantity outside the singularities of Y, and by an ad hoc \(L^{p}\) norm near \(Y_{\textrm{sing}}\), for any \(p\in [1, 2[\). The example proving Claim 4 in Appendix shows that this type of estimates are sharp.

However, the constant \(C_\alpha (V_{\textrm{sing}}, \omega _{{\mathcal {C}}})\) in Theorem 1.1 involves the geometry of the local pair \((V_{\textrm{sing}}, \omega _{{\mathcal {C}}})\), or if one prefers, the restriction of \(h_L\) to \(V_{\textrm{sing}}\). Moreover, we only allow singularities of conic type for \(h_L|_{V_{\textrm{sing}}}\). In order to try to “guess” the type of estimates one could hope for in general, we make the following observation. Let \(\Omega \subset V_{\textrm{sing}}\) be a coordinate open subset. The restriction of the RHS of (1.1.1) to \(\Omega \) is given by the following expression:

$$\begin{aligned} \int _{\Omega \cap Y_i}\frac{1}{\prod _{j\ne i}|f_j|^{\frac{2}{1+\alpha }}}\frac{|f_u|^{\frac{2}{1+\alpha }}\textrm{d}\lambda }{\prod _i |z_i|^{2(1-1/k_i)}} \end{aligned}$$
(1.5.1)

where \(\prod f_j= 0\) is the local equation of \(Y\cap \Omega \) and \(\prod z_i= 0\) is the equation of W. Therefore the second term of RHS of (1.1.1) is equivalent to

$$\begin{aligned} \int _{Y_i\cap \Omega } \left| \frac{u}{\textrm{d}s_Y}\right| _{\omega }^{\frac{2}{1+\alpha }}e^{-{\varphi _L}}\textrm{d} V_\omega , \end{aligned}$$
(1.5.2)

where \(\omega \) is a smooth Kähler metric on X. So from this point of view the following important—and very challenging—problem is natural.

Conjecture 1.6

We assume that the metric \(h_L= e^{-\varphi _L}\) of L and the section u verifies the requirements (a), (b), (c). Then for each \(1\geqslant \alpha > 0\), there exists an section \(U \in H^0 (X, K_X +Y+L)\) such that \(U |_Y =u\) with the estimates

$$\begin{aligned} \frac{1}{C_\alpha (V_{\textrm{sing}})}\int _{X\setminus V_{\textrm{sing}}} |U|^2e^{-\varphi _Y- \varphi _L}\textrm{d}V_{\omega }&\leqslant \int _{Y\setminus V_{\textrm{sing}}}\left| \frac{u}{\textrm{d}s}\right| _\omega ^2e^{-\varphi _L} \textrm{d}V_{\omega }\nonumber \\&\quad + \sum _i\left( \int _{Y_i\cap V_{\textrm{sing}}} \left| \frac{u}{\textrm{d}s}\right| _\omega ^{\frac{2}{1+\alpha }}e^{-{\varphi _L}}\textrm{d}V_{\omega }\right) ^{1+\alpha } \end{aligned}$$
(1.6.1)

where \(\omega \) is a reference smooth Kähler metric on X and \(C_\alpha (V_{\textrm{sing}})\) only depends on \((V_{\textrm{sing}}, \omega )\) and the restriction of the metric \(h_L|_Y\) to Y.

It is our belief that the most subtle part of the previous conjecture would be to have an accurate estimate for the constant \(C_\alpha (V_{\textrm{sing}})\).

Our next two results are of experimental nature and therefore we have decided to formulate them for surfaces only, so that we have \(\dim (X)= 2\). The same type of statements hold in arbitrary dimension, as one can easily convince oneself. The method of proof is completely identical to the case we explain here in detail, so for simplicity’s sake we stick to the case of surfaces.

We fix next few more notations adapted to the pair (XY).

Let \((\Omega _i)_{i\in I}\) be covering of X with open coordinate subsets. By the simple normal crossing hypothesis we can choose coordinates \(z_i= (z_i^1, z_i^2)\) such as

$$\begin{aligned} Y\cap \Omega _i= (z_i^1 \dots z_i^p= 0) \end{aligned}$$
(1.6.2)

for each \(i\in I\) and some p (depending on i). Let \((\theta _i)_{i\in I}\) be a partition of unity subordinate to \((\Omega _i)_{i\in I}\).

Since we assume that X is a complex surface, the components of Y are smooth curves. The singular set of Y (i.e., the mutual intersections of its components) consists of a finite number of points of X, denoted by \(p_1,\dots , p_s\).

We assume that \(\Omega _i\) is refined enough so that the section \(u|_{\Omega _i}\) is given by

$$\begin{aligned} f_idz_i^1\wedge dz_i^2 \end{aligned}$$
(1.6.3)

for some holomorphic function \(f_i\). On overlapping open subsets, different expressions (1.6.3) are gluing only modulo a 2-form divisible by the equation of the divisor Y.

Let p be one of the singular points of Y, assumed to be the center of some \(\Omega _i\). We denote by \(t_i:= z_i^1\cdot z_i^2\); this is—by our previous conventions—the local equation of the cross \(Y\cap \Omega _i\). We can interpret the function (= \(n-2\)-form in general) \(f_i\) as a local section of the bundle \(L|_{\Omega _i}\), and as such we can consider its derivative

$$\begin{aligned} \partial _{\varphi _L}f_i \end{aligned}$$
(1.6.4)

with respect to the Chern connection of L. The result is a 1-form on \(\Omega _i\).

Given the hypothesis in our following statement, it is possible to construct an extension of u by applying the result in [23]. However, here we obtain different types of estimates.

Theorem 1.7

Let X be a smooth projective surface, and let \((L, h_L)\) be a line bundle such that the usual curvature conditions (a) and (c) are satisfied. Assume moreover that \(h_L\) is non-singular and for each \(i\in I_{\textrm{sing}}\) we have \(f_i\in (z_i^1, z_i^2)\), in other words our section vanishes on the set singular points of Y.

Then there exists a section \(U \in H^0 (X, K_X +Y+L)\) enjoying the following properties:

  1. (1)

    \(U|_{Y} =u\).

  2. (2)

    There exists a constant \(C(X, V_{\textrm{sing}})> 0\) such that we have

    $$\begin{aligned} \frac{1}{C(X, V_{\textrm{sing}})}\int _X|U|^2e^{-\varphi _Y-\varphi _L}&\leqslant \int _{Y\setminus V_{\textrm{sing}}}\log ^2(\max |s_{j}|^2)\left| \frac{u}{\textrm{d}s}\right| ^2e^{-\varphi _L}\nonumber \\&\quad \quad + \int _{Y\cap V_{\textrm{sing}}}\log ^2(\max |s_{j}|^2) |\partial _{\varphi _L} u|^2e^{-\varphi _L}. \end{aligned}$$
    (1.7.1)

We obtain the same type of result provided that the bundle \((L, h_L)\) is flat near \(Y_{\textrm{sing}}\), as follows.

Theorem 1.8

Let X be a smooth projective surface, and let \((L, h_L)\) be a line bundle such that the curvature and \(L^2\) conditions (a), (b), (c) as well as the additional property (ii) are satisfied.

Then there exists a section \(U \in H^0 (X, K_X +Y+L)\) enjoying the following properties:

  1. (1)

    \(U|_{Y} =u\).

  2. (2)

    We have

    $$\begin{aligned} \frac{1}{C(X, V_{\textrm{sing}})}\int _{X\setminus V_{\textrm{sing}}}|U|^2e^{-\varphi _Y-\varphi _L}&\leqslant \int _{Y\setminus V_{\textrm{sing}}}\log ^2(\max |s_{j}|^2)\left| \frac{u}{\textrm{d}s}\right| ^2e^{-\varphi _L}\nonumber \\&\quad + \int _{Y\cap V_{\textrm{sing}}}\log ^2(\max |s_{j}|^2) |\partial _{\varphi _L} u|^2e^{-\varphi _L}. \end{aligned}$$
    (1.8.1)

Our next statement is confined to the two-dimensional case.

Theorem 1.9

Let X be a smooth projective surface, and let \((L, h_L)\) be a line bundle such that the usual curvature and \(L^2\) conditions (a), (b), and (c) are satisfied, respectively. We assume moreover that the following hold.

  1. (1)

    There exists a component \(Y_1\) of Y which intersects \(\cup _{j\ne 1} Y_j\) in a unique point \(p_1\), such that \(u(p_1)\ne 0\).

  2. (2)

    The restriction \((L, h_L)|_{Y_1}\) is Hermitian flat.

Then the section u admits an extension U satisfying the same estimates as in Theorem 1.7.

The raison d’être of the previous Theorems 1.7, 1.8, and 1.9 is that the inequality (1.8.1) is meaningful even in the absence of the additional hypothesis these statements contain. Because of the variety of contexts in which an extension of u verifying the estimates of type (2) of Theorem 1.8 can be obtained, it is very tempting to formulate the following.

Conjecture 1.10

Let (XY) be a smooth projective pair, where X is a surface and Y is an snc divisor. Let \((L, h_L)\) be a line bundle such that the properties (a), (b), and (c) are satisfied. Then there exists a section \(U \in H^0 (X, K_X +Y+L)\) enjoying the following properties:

  1. (1)

    The section U is an extension of u.

  2. (2)

    We have

    $$\begin{aligned} \frac{1}{C(X, V_{\textrm{sing}})}\int _{X\setminus V_\textrm{sing}}|U|^2e^{-\varphi _Y-\varphi _L}&\leqslant \int _{Y\setminus V_\textrm{sing}}\log ^2(\max |s_{j}|^2)\left| \frac{u}{\textrm{d}s}\right| ^2e^{-\varphi _L}\nonumber \\&\quad + \lim _{\varepsilon \rightarrow 0}\int _{Y\cap V_{\textrm{sing}}}\log ^2(\max |s_{j}|^2) |\partial _{\varphi _{L, \varepsilon }} u|^2e^{-\varphi _{L, \varepsilon }}, \end{aligned}$$
    (1.10.1)

    where \(\varphi _{L, \varepsilon }:= \log \left( \varepsilon ^2e^{\phi }+ e^{\varphi _L}\right) \) is a non-singular approximation of \(h_L\).

In the sequel of this article we will formulate the precise higher dimensional version of this conjecture, and we will explain its impact on the extension of the pluricanonical forms.

In Appendix A by Bo Berndtsson some examples are given that indicate that the estimates (2) in Conjecture 1.10 are most likely the best one could hope for: without the \(\log \) factor, this conjecture is simply wrong. Moreover, the example given in order to prove Claim 4 shows that the factor \(e^{-\varphi _L}\) in (1.6.1) cannot be replaced by the slightly less singular weight \(e^{-(1-\varepsilon )\varphi _L}\), for any \(\varepsilon > 0\). Finally, Appendix contains a comparison with a one-dimensional problem (the fat point), intended to highlight the origin of the difficulties in a very simple setting.

1.1 Organization of the Paper

In the second section we explain the main ideas involved in the proof of our results. The next section is dedicated to the revision and slight improvement of the usual a-priori inequalities. Our principal contribution to the Ohsawa–Takegoshi artisan industry is in Sect. 4, where the necessary tools from geometric analysis are recalled/developed. The proof of the results stated above is presented in Sect. 5.

1.2 In Memoriam Nessim Sibony

This article is dedicated to a mentor and colleague, Nessim Sibony, who has left us in 2021. His untimely death has affected us all deeply, leaving a void in the complex geometry community. He was a highly accomplished mathematician and a colleague who was inspirational to generations. He will be remembered not only as a revered collaborator with his profound understanding of the field of complex geometry but also as a dear friend with his sharp grasp of a wide range of subjects from literature, history, and to life in general.

2 An Overview of the Arguments

Our results are obtained by combining the method in [28] with the method in [2]. In order to highlight the main arguments as well as some of the difficulties, we only discuss here the case of a non-singular metric \(h_L\) on L. In general the whole scheme of the proof becomes more technical, since the regularization procedure we have to use for the metric is quite tricky to implement in the presence of the singular hypersurface Y.

We start with a quick review of the usual case.

2.1 The Case of a Non-singular Hypersurface Y

Let \(\xi \) be a L-valued form of type (n, 1). We denote by \(\gamma _\xi := \star \xi \) its Hodge dual (induced by an arbitrary Kähler form on X).

Consider the functional

$$\begin{aligned} {{\mathcal {F}}}(\xi )= \int _X{\overline{\partial }} \left( \frac{u}{s_Y}\right) \wedge \overline{\gamma _\xi }e^{-\varphi _L} \end{aligned}$$
(2.0.2)

associated to the current \({\overline{\partial }} \left( \frac{u}{s_Y}\right) \).

We decompose \(\xi = \xi _1+ \xi _2\) according to \(\textrm{Ker}({\overline{\partial }})\) and \(\textrm{Ker}({\overline{\partial }})^\perp \). It turns out that we have the equality

$$\begin{aligned} {{\mathcal {F}}}(\xi )= \int _Y \frac{u}{\textrm{d}s_Y}\wedge \overline{\gamma _{\xi _1}}e^{-\varphi _L} \end{aligned}$$
(2.0.3)

which is not completely obvious, given that the current defining \({\mathcal {F}}\) is not in \(L^2\).

We have \(\frac{u}{\textrm{d}s_Y}\in L^2(e^{-\varphi _L})\), hence it is enough to find an upper bound for

$$\begin{aligned} c_{n-1}\int _Y \gamma _{\xi _1}\wedge \overline{\gamma _{\xi _1}}e^{-\varphi _L}. \end{aligned}$$
(2.0.4)

This is done by the next estimate, which is derived in [4] via the \(\partial {\overline{\partial }}\)—Bochner method due to Siu cf. [39]

$$\begin{aligned} c_{n-1}\int _Y\gamma _{\xi _1}\wedge \overline{\gamma _{\xi _1}}e^{-\varphi _L}\leqslant C\int _X\log ^2(|s_Y|^2)|{\overline{\partial }}^\star {\xi _1}|^2e^{-\varphi _L}\textrm{d}V_\omega . \end{aligned}$$
(2.0.5)

In conclusion, we have

$$\begin{aligned} \left| \int _X{\overline{\partial }} \left( \frac{u}{s_Y}\right) \wedge \overline{\gamma _\xi }e^{-\varphi _L}\right| ^2 \leqslant C\int _X\log ^2(|s_Y|^2)|{\overline{\partial }}^\star {\xi }|^2e^{-\varphi _L}\textrm{d}V_\omega \end{aligned}$$
(2.0.6)

and the “estimable” extension will be obtained by using the solution of the equation \({\overline{\partial }} \left( \frac{u}{s_Y}\right) = {\overline{\partial }} v\). We define \(U:= s_Y v\) and then we have

$$\begin{aligned} U|_{Y}= u, \qquad \int _X\frac{|U|^2}{|s|^2\log ^2|s|^2}e^{-\varphi _Y-\varphi _L}\leqslant C \int _Y|\frac{u}{\textrm{d}s_Y}|^2e^{-\varphi _L}. \end{aligned}$$
(2.0.7)

2.1.1 Difficulties in the Case of an snc Hypersurface Y

In our setting we have \(Y= \bigcup Y_i\), and the difficulty steams from the fact that the functional

$$\begin{aligned} \int _Y \frac{u}{\textrm{d}s_Y}\wedge \overline{\gamma _{\xi _1}}e^{-\varphi _L}= \sum _i \int _{Y_i} \frac{u}{\prod _{j\ne i} s_j \textrm{d}s_{Y_i}}\wedge \overline{\gamma _{\xi _1}}e^{-\varphi _L} \end{aligned}$$
(2.0.8)

becomes a sum of expressions involving forms with log poles. We have

$$\begin{aligned} \frac{u}{\prod _{j\ne i} s_j \textrm{d}s_{Y_i}}\not \in L^2(e^{-\varphi _L}|_{Y_i}) \end{aligned}$$
(2.0.9)

in general, so the previous arguments are breaking down.

Nevertheless we do have

$$\begin{aligned} \left| \frac{u}{\textrm{d}s_Y}\right| _{\omega }^{\frac{2}{1+\alpha }}\in L^1(Y, \omega |_{Y}) \end{aligned}$$
(2.0.10)

near the singularities of Y for any reasonable metric \(\omega \). This means that we have to find an estimate of the \(L^\infty \) norm of \(\gamma _{\xi _1}|_{V_{\textrm{sing}}\cap Y}\) in terms of the RHS of (2.0.5).

To this end, we use a procedure due to Donaldson–Sun in [15]. This consists in the following simple observation. Assume that the support of \(\xi \) is contained in \(X\setminus V_{\textrm{sing}}\). Then we have

$$\begin{aligned} {{\bar{\partial }}}\xi _1= 0,\qquad {{\bar{\partial }}}^\star \xi _1|_{V_{\textrm{sing}}}= 0 \end{aligned}$$
(2.0.11)

in other words, the restriction of \(\xi _1\) to \(V_{\textrm{sing}}\) is harmonic. As we learn from function theory, harmonic functions satisfy the mean value inequality: this is what we implement in our context, and it leads to the proof of Theorem 1.1.

The drawback of this method is that in the end, the constant measuring the \(L^2\) norm of the extension is far from being as universal as in the case \(Y_{\textrm{sing}}= \emptyset \). This is due to the fact that the quantity \(\Delta ''|\xi _1|^2\) has a term with wrong sign involving the trace of the curvature of \((L, h_L)\) with respect to the metric \(\omega \) on X. This trace is not bounded e.g., if \(h_L\) is singular and \(\omega \) is a fixed, non-singular Kähler metric. It is for this reason that the singularities of \(h_L\) and those of \(\omega _{{\mathcal {C}}}\) must be the same in Theorem 1.1.

3 A-Priori Inequalities Revisited

We first recall the following estimate, which is essentially due to [4].

Theorem 3.1

Let \((X, \omega )\) be a Kähler manifold, and Y be simple normal crossing divisor in X. Let L be a line bundle on X with a non-singular metric \(h_L\) such that

$$\begin{aligned} \Theta _{h_L} (L)\geqslant 0,\qquad \Theta _{h_L} (L) \geqslant \delta \Theta _{h_Y} (Y) \end{aligned}$$

for some \(\delta >0\) small enough, where \(h_Y\) is a smooth metric on \({\mathcal {O}}_X (Y)\) such that \(|s_Y|^2_{h_Y}\leqslant e^{-\delta }\). Let \(\xi \) be a smooth (n, 1) form with compact support and with values in L. We denote by \(\gamma _\xi := \star \xi \) the image of \(\xi \) by the Hodge operator. Then we have

$$\begin{aligned} c_{n-1}\int _X&\frac{\tau ^2}{(\tau ^2+ |s_Y|^2)^2}\gamma _\xi \wedge \overline{\gamma _\xi }e^{-\varphi _L}\wedge \sqrt{-1}\partial s_Y\wedge \overline{\partial s_Y} \\&\leqslant C\int _X\log ^2(\tau ^2+ |s_{Y}|^2)\left( |{{\bar{\partial }}}^\star \xi |^2+ |{{\bar{\partial }}}\xi |^2\right) e^{-\varphi _L}\textrm{d}V_\omega \end{aligned}$$

where C is a numerical constant and \(\tau \) is an arbitrary real number.

Before giving the proof of Theorem 3.1 we notice that it implies the following statement:

Theorem 3.2

Let \((X, \omega )\) be a Kähler manifold, and Y be simple normal crossing divisor in X. Let L be a line bundle on X with a non-singular metric \(h_L\) such that

$$\begin{aligned} \Theta _{h_L} (L)\geqslant 0,\qquad \Theta _{h_L} (L) \geqslant \delta \Theta _{h_Y} (Y) \end{aligned}$$

for any \(\delta >0\) small enough, where \(h_Y\) is a smooth metric on \({\mathcal {O}}_X (Y)\). Let \(\xi \) be a smooth (n, 1) form with compact support and with values in L. We denote by \(\gamma _\xi := \star \xi \) the image of \(\xi \) by the Hodge operator. Then we have

$$\begin{aligned} c_{n-1}\int _Y\gamma _\xi \wedge \overline{\gamma _\xi }e^{-\varphi _L}\leqslant C\int _X\log ^2(|s_{Y}|_{h_Y}^2)\left( |{{\bar{\partial }}}^\star \xi |^2+ |{{\bar{\partial }}}\xi |^2\right) e^{-\varphi _L}\textrm{d}V_\omega , \end{aligned}$$
(3.2.1)

where \(s_{Y}\) is the canonical section of \({\mathcal {O}}(Y)\), normalized in a way that works for the proof.

Proof of Theorem 3.1

We note that this improves slightly the estimate of Bo Berndtsson in [4], but the proof is virtually the same. Nevertheless, we will provide a complete treatment for the convenience of the reader.

To start with, we recall the following "\(\partial {\bar{\partial }}\)-Bochner formula.”

Lemma 3.3

([38]) Let \(\xi \) be a (n, 1)-form with values in \((L, h_L)\) and compact support. We denote by \(\gamma _\xi = \star \xi \) the Hodge \(\star \) of \(\xi \) with respect to a Kähler metric \(\omega \). Let

$$\begin{aligned} T_\xi := c_{n-1}\gamma _\xi \wedge \overline{\gamma }_{\xi }e^{-\varphi _L} \end{aligned}$$
(3.3.1)

be the \((n-1, n-1)\)-form on X corresponding to \(\xi \), where \(c_{n-1}= \sqrt{-1}^{(n-1)^2}\) is the usual constant. Then we have the equality

$$\begin{aligned} \sqrt{-1}\partial {\bar{\partial }}T_\xi&= \left( -2\Re \langle {{\bar{\partial }}}{{\bar{\partial }}}^\star _\varphi \xi , \xi \rangle + \Vert {{\bar{\partial }}}\gamma _\xi \Vert ^2+ \Vert {{\bar{\partial }}}^\star _\varphi \xi \Vert ^2- \Vert {{\bar{\partial }}}\xi \Vert ^2 \right) \textrm{d}V_\omega \nonumber \\&\quad + \Theta _{h_L}(L)\wedge T_\xi . \end{aligned}$$
(3.3.2)

We apply this in the following context. Consider the function \(w:= \log \frac{1}{|s_Y|^2+ \tau ^2}\). A quick computation gives

$$\begin{aligned} \sqrt{-1}\partial {\bar{\partial }}w= \frac{|s_Y|^2}{|s_Y|^2+ \tau ^2}\theta _Y- \frac{\tau ^2}{(|s_Y|^2+ \tau ^2)^2}\sqrt{-1}\partial s_Y\wedge \overline{\partial s_Y} \end{aligned}$$
(3.3.3)

where \(\theta _Y:= \Theta _{h_Y}\left( {\mathcal {O}}(Y)\right) \) is the curvature of the bundle \({\mathcal {O}}(Y)\) with respect to the metric \(h_Y\).

We multiply the equality (3.3.2) with w and integrate the resulting top form over X. The left-hand side term is equal to the difference of two terms

$$\begin{aligned} c_{n-1}\int _X\frac{|s_Y|^2}{|s_Y|^2+ \tau ^2}\theta _Y\wedge \gamma _\xi \wedge \overline{\gamma }_{\xi }e^{-\varphi _L} \end{aligned}$$
(3.3.4)

and

$$\begin{aligned} c_{n-1}\int _X \frac{\tau ^2}{(|s_Y|^2+ \tau ^2)^2}\gamma _\xi \wedge \overline{\gamma }_{\xi }\wedge \sqrt{-1}\partial s_Y\wedge \overline{\partial s_Y}. \end{aligned}$$
(3.3.5)

and we see that (3.3.5) is the term we have to estimate.

We drop the positive terms on the right-hand side of (3.3.2) and we therefore get

$$\begin{aligned}&c_{n-1}\int _X \frac{\tau ^2}{(|s_Y|^2+ \tau ^2)^2}\gamma _\xi \wedge \overline{\gamma }_{\xi }\wedge \sqrt{-1}\partial s_Y\wedge \overline{\partial s_Y}\leqslant \int _Xw\vert {{\bar{\partial }}}\xi |^2e^{-\varphi _L}\textrm{d}V_\omega \nonumber \\&\quad + 2\Re \int _Xw\langle \partial _\varphi ^\star \partial _\varphi \gamma _\xi , \gamma _\xi \rangle e^{-\varphi _L}\textrm{d}V_\omega \nonumber \\&\quad - c_{n-1}\int _X\left( w\Theta _{h_L}(L)-\frac{|s_Y|^2}{|s_Y|^2+ \tau ^2}\theta _Y\right) \gamma _\xi \wedge \overline{\gamma }_{\xi }e^{-\varphi _L} \end{aligned}$$
(3.3.6)

A first observation is that the curvature term (3.3.6) is negative, by the hypothesis of Theorem 3.1. Moreover, by Stokes formula we have

$$\begin{aligned} \int _Xw\langle \partial _\varphi ^\star \partial _\varphi \gamma _\xi , \gamma _\xi \rangle e^{-\varphi }\textrm{d}V_\omega= & {} \int _Xw|\partial _\varphi \gamma _\xi |^2e^{-\varphi }\textrm{d}V_\omega \nonumber \\{} & {} + \int _X\langle \partial _\varphi \gamma _\xi , \partial w\wedge \gamma _\xi \rangle e^{-\varphi }\textrm{d}V_\omega \end{aligned}$$
(3.3.7)

so we see that modulo the second term on the RHS of (3.3.7), we are done.

In order to take care of it we use Cauchy–Schwarz inequality and we obtain

$$\begin{aligned} \left| \int _X\langle \partial _\varphi \gamma _\xi , \partial w\wedge \gamma _\xi \rangle e^{-\varphi _L}\textrm{d}V_\omega \right|&\leqslant \int _Xw^2|\partial _\varphi \gamma _\xi |^2e^{-\varphi _L}\textrm{d}V_\omega \nonumber \\&\quad + c_{n-1}\int _X\gamma _\xi \wedge \overline{\gamma }_{\xi }\wedge \frac{\sqrt{-1}\partial w\wedge {{\bar{\partial }}}w}{w^2}e^{-\varphi _L}. \end{aligned}$$
(3.3.8)

Thus the new term to bound is

$$\begin{aligned} c_{n-1}\int _X\gamma _\xi \wedge \overline{\gamma }_{\xi }\wedge \frac{\sqrt{-1}\partial w\wedge {{\bar{\partial }}}w}{w^2}e^{-\varphi _L} \end{aligned}$$
(3.3.9)

and as observed in [4], the quantity (3.3.9) is less singular that the LHS of (3.3.6), which was our initial problem.

In order to obtain a bound for (3.3.9) we consider the function

$$\begin{aligned} w_1:= \log w. \end{aligned}$$

We have

$$\begin{aligned} \sqrt{-1}\partial {\bar{\partial }}w_1= \frac{\sqrt{-1}\partial {\bar{\partial }}w}{w}- \frac{\sqrt{-1}\partial w\wedge {{\bar{\partial }}}w}{w^2} \end{aligned}$$
(3.3.10)

and we use the same procedure as before, but with \(w_1\) instead of w. The analog of (3.3.4) and (3.3.5) read as

$$\begin{aligned} c_{n-1}\int _X\frac{\sqrt{-1}\partial {\bar{\partial }}w}{w}\wedge \gamma _\xi \wedge \overline{\gamma }_{\xi }e^{-\varphi _L}- c_{n-1}\int _X \gamma _\xi \wedge \overline{\gamma }_{\xi }\wedge \frac{\sqrt{-1}\partial w\wedge {{\bar{\partial }}}w}{w^2},\nonumber \\ \end{aligned}$$
(3.3.11)

and this is good, because the second term in (3.3.11) is the one we are now after. We skip some intermediate steps because they are absolutely the same as in the preceding consideration, except that \(w_1\) appears instead of w. After integration by parts, the new “bad term,” i.e., the analog of the RHS of (3.3.7) in our current setting is

$$\begin{aligned} \int _X\langle \partial _\varphi \gamma _\xi , \partial w_1\wedge \gamma _\xi \rangle e^{-\varphi _L}\textrm{d}V_\omega \end{aligned}$$
(3.3.12)

for which we use Cauchy–Schwarz inequality and the observation is that \(\partial w_1\wedge \overline{\partial w_1}\) coincides with \(\frac{\sqrt{-1}\partial w\wedge {{\bar{\partial }}}w}{w^2}\).

As a result of this second part of the proof we infer that we have

$$\begin{aligned}{} & {} c_{n-1}\int _X \gamma _\xi \wedge \overline{\gamma }_{\xi }\wedge \frac{\sqrt{-1}\partial w\wedge {{\bar{\partial }}}w}{w^2}\nonumber \\ {}{} & {} \quad \leqslant C\int _X\log ^2(|s_{Y}|^2+ \tau ^2)\left( |{{\bar{\partial }}}^\star \xi |^2+ |{{\bar{\partial }}}\xi |^2\right) e^{-\varphi _L}\textrm{d}V_\omega \end{aligned}$$
(3.3.13)

Then Theorem 3.1 follows, by combining (3.3.13) with (3.3.8). \(\square \)

Remark 3.4

Actually we can use the second part of the proof of Theorem 3.2 in order to get the estimates

$$\begin{aligned}{} & {} c_{n-1}\int _X\gamma _\xi \wedge \overline{\gamma _\xi }e^{-\varphi _L}\wedge \frac{\partial \sigma \wedge \overline{\partial \sigma }e^{-\varphi _F}}{|\sigma |^2\log ^2|\sigma |^2}\nonumber \\ {}{} & {} \quad \leqslant C\int _X\log ^2(|\sigma |^2)\left( |\xi |^2+ |{{\bar{\partial }}}^\star \xi |^2+ |{{\bar{\partial }}}\xi |^2\right) e^{-\varphi _L}\textrm{d}V_\omega \end{aligned}$$
(3.4.1)

where \(\sigma \) is a holomorphic section of a line bundle \((F, h_F)\) endowed with a non-singular metric \(h_F\). The constant “C” in (3.4.1) depends on the norm of the curvature of \((F, h_F)\). Thus, we obtain an estimate of the norm of \(\gamma _\xi \) in the tangential directions of \(\sigma =0\) with respect to the Poincaré-type measure associated to \(\sigma \). If the curvature of \((L, h_L)\) is greater than the some (small) multiple of \(\Theta _{h_F}(F)\), then we can remove the term \(|\xi |^2\) in the formula (3.4.1).

4 Geometric Analysis Methods and Results

In this section, we follow the notations in Theorem 1.1. Moreover, we ask that the metric \(\varphi _L\) satisfies the requirements (a), (b), (c) and \((i.\alpha )\). In particular, we don’t assume \((i.\beta )\) for \(\varphi _L\).

Let \(\xi \) be a L-valued form of (n, 1) type such that \(\textrm{Supp}(\xi ) \subset X\setminus (V_{\textrm{sing}}\cup |H|)\). We recall that here \(V_{\textrm{sing}}\) is an open subset of X containing the singularities of Y, and H is a hyperplane section containing the singularities of the metric \(h_L\).

We consider the orthogonal decomposition

$$\begin{aligned} \xi = \xi _1+ \xi _2 \end{aligned}$$
(4.0.2)

where \(\xi _1\in \textrm{Ker}({{\bar{\partial }}})\) and \(\xi _2\in \textrm{Ker}({{\bar{\partial }}})^\perp \) with respect to the fixed Kähler metric \(\omega _{{\mathcal {C}}}\) with conic singularities on X and the given metric \(h_L\) on L.

\(\bullet \) The convention during the current section is that we denote by “C” any constant which depends in an explicit way of the quantities we will indicate.

4.1 Orthogonal Decomposition, I: Approximation

In the following sections we will use an approximation statement, for which the context is as follows:

We can write

$$\begin{aligned} X\setminus H= \bigcup \Omega _m \end{aligned}$$
(4.0.3)

where each \(\Omega _m\) is a Stein domain with smooth boundary. Let \(\omega _m\) be a complete metric on \(\Omega _m\). Corresponding to each positive \(\delta \) we introduce

$$\begin{aligned} \omega _{m, \delta }:= \omega _{{\mathcal {C}}}+ \delta \omega _m; \end{aligned}$$
(4.0.4)

it is a complete metric on \(\Omega _m\) such that \(\omega _{m, \delta }> \omega _{{\mathcal {C}}}\) and \(\lim _{\delta \rightarrow 0}\omega _{m, \delta }= \omega _{{\mathcal {C}}}\) for each m.

We remark that the \(L^2\) norm of \(\xi \) with respect to \(\omega _{m, \delta }\) and \(h_L|_{\Omega _m}\) is finite, given the pointwise monotonicity of the norm of (n, 1)-forms. Then we can decompose the restriction of \(\xi \) to each \(\Omega _m\) as follows:

$$\begin{aligned} \xi |_{\Omega _m}= \xi _1^{(m,\delta )}+ \xi _2^{(m, \delta )}. \end{aligned}$$
(4.0.5)

We establish next the following statement.

Lemma 4.1

We have

$$\begin{aligned} \xi _1= \lim _{m,\delta }\xi _1^{(m,\delta )} \end{aligned}$$
(4.1.1)

uniformly on compact sets of \(X\setminus H\).

The proof is based on the monotonicity of the \(L^2\) norms

$$\begin{aligned} \vert \rho \vert _{\omega _{m, \delta }}^2\textrm{d}V_{\omega _{m, \delta }}< \vert \rho \vert _{\omega }^2\textrm{d}V_{\omega } \end{aligned}$$
(4.1.2)

for each \(m, \delta \) and for any form \(\rho \) of type (n, 1) with values in L. The details are as follows:

Proof

Let \(K\subset X\setminus (\varphi _L= -\infty )\) be a compact subset. In what follows we are using the notation “\(\varepsilon \)” to indicate the set of parameters \((m, \delta )\), and we assume that \(m\gg 0\) so that \(K\subset \Omega _m\).

We first notice that for each parameter \(\varepsilon \) the form \(\xi _1^{(\varepsilon )}\) is smooth, and that it verifies the equation

$$\begin{aligned} \Delta _\varepsilon ''(\xi _1^{(\varepsilon )})= {{\bar{\partial }}}{{\bar{\partial }}}^\star (\xi ), \end{aligned}$$
(4.1.3)

where \(\Delta _\varepsilon ''\) is the Laplace operator on (n, 1)-forms with values in \((L, h_L)\) and \((\Omega _m, \omega _{m, \delta })\). We also have

$$\begin{aligned} \int _{\Omega _m}\left| \xi _1^{(\varepsilon )}\right| ^2_{\omega _\varepsilon }e^{-\varphi _L}\textrm{d}V_{\omega _\varepsilon }\leqslant \int _X\left| \xi \right| ^2_{\omega _{{\mathcal {C}}}}e^{-\varphi _L}\textrm{d}V_{\omega _{{\mathcal {C}}}}= \Vert \xi \Vert ^2 \end{aligned}$$
(4.1.4)

given the fact that (4.1.1) is orthogonal. It follows that the family

$$\begin{aligned} \xi _1^{(\varepsilon )}|_K \end{aligned}$$
(4.1.5)

is uniformly bounded in \({\mathcal {C}}^\infty \) norm. We can therefore extract a limit \(\xi _1^{(0)}\) as \(\varepsilon \rightarrow 0\), uniform on compact subsets by the usual diagonal process. We remark that we have

$$\begin{aligned} {{\bar{\partial }}}\xi _1^{(0)}= 0, \qquad \int _X\left| \xi _1^{(0)}\right| ^2_{\omega }e^{-\varphi _L}\textrm{d}V_\omega < \infty \end{aligned}$$
(4.1.6)

given that each form \(\xi _1^{(\varepsilon )}\) is \({{\bar{\partial }}}\)-closed, combined with (4.1.4).

On the other hand, let \(\rho \) be a \({{\bar{\partial }}}\)-closed form of (n, 1)-type with values in L. We assume moreover that \(\rho \) is \(L^2\) with respect to \(\omega _{{\mathcal {C}}}\) and \(h_L\). Then we equally have

$$\begin{aligned} \int _{\Omega _m}\left| \rho \right| ^2_{\omega _\varepsilon }e^{-\varphi _L}\textrm{d}V_{\omega _\varepsilon }< \int _{\Omega _m}\left| \rho \right| ^2_{\omega _{{\mathcal {C}}}}e^{-\varphi _L}\textrm{d}V_{\omega _{{\mathcal {C}}}}< \infty \end{aligned}$$
(4.1.7)

for each \(\varepsilon = (m, \delta )\), and open subset \(\Omega _m\). We infer that

$$\begin{aligned} \int _{\Omega _m}\langle \xi _2^{(\varepsilon )}, \rho \rangle _{\omega _\varepsilon } e^{-\varphi _L}\textrm{d}V_{\omega _\varepsilon }= 0 \end{aligned}$$
(4.1.8)

for each value of m and \(\varepsilon \).

Let \((K_l)\) be an increasing exhaustion of \(X{\setminus } (\varphi _L= -\infty )\) by relatively compact sets. If \(m\gg 0\) (depending on l) then we have

$$\begin{aligned} \left| \int _{\Omega _m\setminus K_l}\langle \xi _2^{(\varepsilon )}, \rho \rangle _\omega e^{-\varphi _L}\textrm{d}V_\omega \right| ^2\leqslant C(\xi )\int _{X\setminus K_l}\left| \rho \right| ^2_{\omega }e^{-\varphi _L}\textrm{d}V_\omega \end{aligned}$$
(4.1.9)

by Cauchy inequality combined with (4.1.7). It follows that

$$\begin{aligned} \left| \int _{K_l}\langle \xi _2^{(\varepsilon )}, \rho \rangle _{\omega _\varepsilon } e^{-\varphi _L}\textrm{d}V_{\omega _\varepsilon }\right| ^2\leqslant C(\xi )\int _{X\setminus K_l}\left| \rho \right| ^2_{\omega _{{\mathcal {C}}}}e^{-\varphi _L}\textrm{d}V_{{\omega _{{\mathcal {C}}}}}. \end{aligned}$$
(4.1.10)

By letting \(\varepsilon \rightarrow 0\) we infer that for each fixed l we have

$$\begin{aligned} \left| \int _{K_l}\langle \xi _2^{(0)}, \rho \rangle _{{\omega _{{\mathcal {C}}}}} e^{-\varphi _L} \textrm{d}V_{{\omega _{{\mathcal {C}}}}}\right| ^2\leqslant C(\xi )\int _{X\setminus K_l}\left| \rho \right| ^2_{\omega _{{\mathcal {C}}}}e^{-\varphi _L}\textrm{d}V_{{\omega _{{\mathcal {C}}}}}. \end{aligned}$$
(4.1.11)

Next, the inequality (4.1.6) shows that \(\xi _2^{(0)}\) is \(L^2\)-integrable with respect to \((L, h_L)\) and \((X, \omega _{{\mathcal {C}}})\). It follows that we have

$$\begin{aligned} \int _{X\setminus K_l}\langle \xi _2^{(0)}, \rho \rangle _{\omega _{{\mathcal {C}}}} e^{-\varphi _L} \textrm{d}V_{\omega _{{\mathcal {C}}}}\rightarrow 0 \end{aligned}$$
(4.1.12)

as \(l\rightarrow \infty \) since both \(\rho \) and \(\xi _2^{(0)}\) are in \(L^2\).

In other words, the form \(\xi _2^{(0)}\) is orthogonal to \(\textrm{Ker} \,{{\bar{\partial }}}\) and since we have

$$\begin{aligned} \xi = \xi _1^{(0)}+ \xi _2^{(0)} \end{aligned}$$
(4.1.13)

our lemma is proved (thanks to the uniqueness of such decomposition). \(\square \)

4.2 Orthogonal Decomposition, II: Mean Value Inequality

We analyze here the behavior of \(\xi _1\) restricted to the set \(V_{\textrm{sing}}\). During the current subsection we make the following conventions.

  1. (i)

    We work with respect to the Kähler metric \(\omega _{{\mathcal {C}}}\) exclusively on \(V_{\textrm{sing}}\subset X\) (this will be understood even if we do not mention it explicitly) and with respect to the Hermitian metric \(h_{L}\) defined in the previous section on L.

  2. (ii)

    We denote by \(\xi \) a (n, 1) form with values in L such that we have

    $$\begin{aligned} \textrm{Supp}(\xi ) \subset X\setminus (V_{\textrm{sing}}\cup |H|). \end{aligned}$$

    We use the notations in (4.0.2) for its orthogonal decomposition with respect to \((\omega _{{\mathcal {C}}}, h_L)\).

  3. (iii)

    We fix \(\frac{1}{2}V_{\textrm{sing}} \Subset V_{\textrm{sing}}\) an open set of compact support in \(V_{\textrm{sing}}\), and \(\frac{1}{2}V_{\textrm{sing}}\) contains the singular locus of Y.

In this subsection we establish the next result.

Theorem 4.2

We have the mean value type inequality

$$\begin{aligned} \sup _{\frac{1}{2}V_{\textrm{sing}}} |\xi _{1}|^{2}e^{-\varphi _L}\leqslant C(V_{\textrm{sing}})\int _{V_\textrm{sing}}|\xi _{1}|^{2}e^{-\varphi _L}\textrm{d}V_{\omega _{{\mathcal {C}}}} \end{aligned}$$
(4.2.1)

where \(C(V_{\textrm{sing}})\) here is a constant which only depends on the allowed quantities, i.e., the geometry of \((V_{\textrm{sing}}, \omega _{{\mathcal {C}}})\) as well as \(\tau _L\) in the assumption \(\mathrm (i.\alpha )\).

The norm of \(\xi _1\) in (4.2.1) is measured with the conic metric \(\omega _{{\mathcal {C}}}\).

The proof of Theorem 4.2 unfolds as follows (cf. [15, 22] for similar computations). In order to simplify the notations, we drop the \(e^{-\varphi _L}\) in (4.2.1), and write \(|\xi _1|^2\) to express the pointwise norm of \(\xi _1\) with respect to \(\omega _{{\mathcal {C}}}\) and \(h_L\). First we show that there exists a constant C such that

$$\begin{aligned} \sup _{V_{\textrm{sing}}\setminus |W|}|\xi _1|^2\leqslant C< \infty \end{aligned}$$
(4.2.2)

where we denote by |W| the support of the divisor W. This is the main reason why we have to assume that the singularities of \(h_L\) and \(\omega \) are “the same” in Theorem 1.1.

After this, we establish a differential inequality satisfied by the function \(|\xi _1|^2\) in the complement of the set

$$\begin{aligned} \left( \varphi _L= -\infty \right) \cap V_{\textrm{sing}}. \end{aligned}$$

This is standard, and it combines nicely with (4.2.2) and Moser iteration process to give (4.2.1).

Proof of (4.2.2)

First we establish the crucial boundedness result (4.2.2). Let \(z_1,\dots , z_n\) be a set of local coordinates defined on a open subset \(\Omega \subset V_{\textrm{sing}}\). We assume that the \((z_i)_{i=1\dots n}\) are adapted to the pair (XW), meaning that the local equation of \(\Omega \cap W\) is

$$\begin{aligned} z_1\dots z_p= 0 \end{aligned}$$

for some \(p\leqslant n\). By hypothesis, the weight of the metric \(h_L\) can be written as

$$\begin{aligned} \varphi _L= \sum _{i=1}^p\Big (1-\frac{1}{k_i}\Big )\log |z_i|^2+ \tau , \end{aligned}$$
(4.2.3)

where \(k_i\) are positive integers and \(\tau \) is a smooth function defined on \(\Omega \).

The restriction of \(\xi _1\) to \(\Omega \) has the following properties:

$$\begin{aligned} \int _{\Omega }|\xi _1|^2_{\omega _{{\mathcal {C}}}}e^{-\varphi _L}\textrm{d}V_{\omega _{{\mathcal {C}}}}< \infty ,\qquad {{\bar{\partial }}}\xi _1=0, \qquad \int _\Omega \langle \xi _1, {{\bar{\partial }}}\phi \rangle _{\omega _{{\mathcal {C}}}}e^{-\varphi _L}\textrm{d}V_{\omega _{{\mathcal {C}}}}= 0\nonumber \\ \end{aligned}$$
(4.2.4)

where \(\phi \) is any smooth (n, 0) form with compact support in \(\Omega \) which is \(L^2\)-integrable, and such that \({{\bar{\partial }}}\phi \) is in \(L^2\) as well.

These properties have a very neat interpretation in terms of ramified covers, as follows. Let

$$\begin{aligned} \pi : \widehat{\Omega }\rightarrow \Omega , \qquad \pi (w_1,\dots , w_n)= (w_1^{k_1},\dots ,w_p^{k_p}, w_{p+1}, \dots , w_n) \end{aligned}$$
(4.2.5)

be the usual local covering map corresponding to the divisor \(\sum _{i=1}^p\Big (1-\frac{1}{k_i}\Big ) W_i\). We define the (n, 1)-form \(\eta \) on \(\widehat{\Omega }\) as follows:

$$\begin{aligned} \eta := \frac{1}{\prod _{j=1}^p w_j^{k_j-1}}\pi ^\star \xi _1, \end{aligned}$$
(4.2.6)

and a first remark is that we have

$$\begin{aligned} \int _{\widehat{\Omega }}|\eta |^2_ge^{-\tau \circ \pi }\textrm{d}V_g< \infty , \end{aligned}$$
(4.2.7)

where g is the inverse image of the conic metric \(\displaystyle g:= \pi ^\star \omega _{{\mathcal {C}}}\). The relation (4.2.7) is an immediate consequence of the change of variables formula, combined with the expression of \(\varphi _L\) in (4.2.3). It follows that

$$\begin{aligned} {{\bar{\partial }}}\eta = 0 \end{aligned}$$
(4.2.8)

on \(\widehat{\Omega }\) (this is true pointwise outside the support of W, and it extends across W by using [12, Chap VIII, Lemma 7.3]).

Let \(\alpha \) be a smooth (n, 0)-form on \(\widehat{\Omega }\) with compact support. We claim that we have

$$\begin{aligned} \int _{\widehat{\Omega }}\langle \eta , {{\bar{\partial }}}\alpha \rangle _g e^{-\tau \circ \pi }\textrm{d}V_g= 0. \end{aligned}$$
(4.2.9)

Indeed this is clear, given the equality

$$\begin{aligned} \int _{\widehat{\Omega }}\langle \eta , {{\bar{\partial }}}\alpha \rangle _g e^{-\tau \circ \pi }\textrm{d}V_g= \int _{\widehat{\Omega }}\langle \pi ^\star \xi _i, {{\bar{\partial }}}\widehat{\alpha }\rangle _g e^{-\varphi _L\circ \pi }\textrm{d}V_g \end{aligned}$$
(4.2.10)

where \(\widehat{\alpha }:= {\prod _{j=1}^p w_j^{k_j-1}}\alpha \). On the right-hand side of (4.2.10) we can assume that \(\widehat{\alpha }\) is the inverse image of a (n, 0) form with compact support on \(\Omega \). Indeed, let f be an element of the group G acting on \(\widehat{\Omega }\). Then we have

$$\begin{aligned} \int _{\widehat{\Omega }}\langle \pi ^\star \xi _i, {{\bar{\partial }}}\widehat{\alpha }\rangle _g e^{-\varphi _L\circ \pi }\textrm{d}V_g = \int _{\widehat{\Omega }}\langle \pi ^\star \xi _i, {{\bar{\partial }}}(f^\star \widehat{\alpha }) \rangle _g e^{-\varphi _L\circ \pi }\textrm{d}V_g \end{aligned}$$
(4.2.11)

since all the other objects involved in the integral in question are invariant by inverse image. It follows that

$$\begin{aligned} \int _{\widehat{\Omega }}\langle \pi ^\star \xi _i, {{\bar{\partial }}}\widehat{\alpha }\rangle _g e^{-\varphi _L\circ \pi }\textrm{d}V_g = \int _{\widehat{\Omega }}\langle \pi ^\star \xi _i, {{\bar{\partial }}}(\pi ^\star \phi ) \rangle _g e^{-\varphi _L\circ \pi }\textrm{d}V_g \end{aligned}$$
(4.2.12)

where \(\pi ^\star \phi := \frac{1}{|G|}\sum _{f\in G}f^\star \widehat{\alpha }\). Then our claim follows by the third property in (4.2.4).

In conclusion, the form \(\eta \) is harmonic on \(\widehat{\Omega }\) with respect to the metric g and the weight \(e^{-\tau \circ \pi }\). It is in particular bounded, and this is equivalent to (4.2.2). \(\square \)

We choose geodesic local coordinates \(\displaystyle (z_i)_{i=1,\dots , n}\) for the Kähler metric \(\omega _{{\mathcal {C}}}\) locally near a point \(x_0\in V_{\textrm{sing}}{\setminus } W\). Let \(e_L\) be a local holomorphic frame of L, such that the induced weight \(\phi \) of \(h_{L}\) verifies the relations

$$\begin{aligned} \phi (x_0)= 0, \qquad \textrm{d}\phi (x_0)= 0. \end{aligned}$$
(4.2.13)

We express \(\xi _1\) locally with respect to these coordinates

$$\begin{aligned} \xi _1= \sum \xi _{\overline{\alpha }}dz\wedge dz^{\overline{\alpha }}\otimes e_L \end{aligned}$$
(4.2.14)

where \(dz:= dz^1\wedge \dots \wedge dz^n\). We then have

$$\begin{aligned} |\xi _1|^2_{\omega _{{\mathcal {C}}}}= \sum _{\alpha , \beta } \xi _{\overline{\alpha }}\overline{\xi _{\overline{\beta }}}g^{\overline{\alpha }\beta }\frac{e^{-\phi }}{\det g}. \end{aligned}$$
(4.2.15)

The formula for the Laplace operator is \(\Delta ''= \textrm{Tr} \,_{\omega _{{\mathcal {C}}}}\sqrt{-1}\partial {\bar{\partial }}\) and so we have

$$\begin{aligned} \Delta ''(|\xi _1|^2)&= |\nabla \xi _1|^2+ 2\sum _{\alpha , \beta } \Re \left( \xi _{\overline{\alpha }, p\overline{q}}g^{\overline{q} p}\overline{\xi _{\overline{\beta }}}g^{\overline{\alpha }\beta }\right) \frac{e^{-\phi }}{\det g} \end{aligned}$$
(4.2.16)
$$\begin{aligned}&\quad + \sum _{\alpha , \beta } \xi _{\overline{\alpha }}\overline{\xi _{\overline{\beta }}}g^{\overline{\alpha }\beta }_{, p\overline{q}}g^{\overline{q} p}\frac{e^{-\phi }}{\det g} \end{aligned}$$
(4.2.17)
$$\begin{aligned}&\quad - \sum _{\alpha , \beta } \xi _{\overline{\alpha }}\overline{\xi _{\overline{\beta }}} (\phi + \log \det g)_{, p\overline{q}}g^{\overline{q} p}g^{\overline{\alpha }\beta }\frac{e^{-\phi }}{\det g} \end{aligned}$$
(4.2.18)

where we denote by \(\big (g_{p\overline{q}}\big )\) the coefficients of the metric \(\omega _{{\mathcal {C}}}\) with respect to the geodesic coordinates \((z_i)\) and by g the corresponding matrix.

In order to obtain an intrinsic expression of the terms containing the second derivative in the RHS of (4.2.16), we recall that we have

$$\begin{aligned} {{\bar{\partial }}}^\star _{\varphi _L}\xi _1= (-1)^n\left( -\frac{\partial \xi _{\overline{\alpha }}}{\partial z_\beta }g^{\overline{\alpha }\beta }- \frac{\partial g^{\overline{\alpha }\beta }}{\partial z_\beta }\xi _{\overline{\alpha }}+ \xi _{\overline{\alpha }}g^{\overline{\alpha }\beta } \frac{\partial \varphi _L}{\partial z_\beta }\right) dz\otimes e_L \end{aligned}$$
(4.2.19)

hence the next equality holds at \(x_0\)

$$\begin{aligned} \langle \square \xi _1, \xi _1\rangle = \left( -\xi _{\overline{\alpha }, p\overline{q}}g^{\overline{q} \beta }\overline{\xi _{\overline{\beta }}}g^{\overline{\alpha }p}- \xi _{\overline{\alpha }}\overline{\xi _{\overline{\beta }}}g^{\overline{\alpha }\delta }_{, \delta \overline{\gamma }}g^{\overline{\gamma }\beta }+ \xi _{\overline{\alpha }}\overline{\xi _{\overline{\beta }}}\varphi _{L, \delta \overline{\gamma }}g^{\overline{\alpha }\delta }g^{\overline{\gamma }\beta }\right) \frac{e^{-\phi }}{\det g}\nonumber \\ \end{aligned}$$
(4.2.20)

where \(\square := [{{\bar{\partial }}}, {{\bar{\partial }}}^\star ]\) is the Laplace operator acting on L-valued forms of (n, 1) type. The formula (4.2.20) is only valid for closed forms, which is the case for \(\xi _1\). Also, we have \(\xi _{\overline{\alpha }, p\overline{q}}= \xi _{\overline{q}, p\overline{\alpha }}\) and therefore (twice the real part of) the first term on the RHS of (4.2.20) coincides with the second one on the RHS of (4.2.16).

Next, since the metric \(\omega _{{\mathcal {C}}}\) is Kähler we have \(g^{\overline{\alpha }\delta }_{, \delta \overline{\gamma }}= -g_{\alpha \overline{\gamma }, \delta \overline{\delta }}\) hence we obtain

$$\begin{aligned} \xi _{\overline{\alpha }}\overline{\xi _{\overline{\beta }}}g^{\overline{\alpha }\delta }_{, \delta \overline{\gamma }}g^{\overline{\gamma }\beta }= {\mathcal {R}} _{\alpha \overline{\beta }}\xi _{\overline{\alpha }}\overline{\xi _{\overline{\beta }}} \end{aligned}$$
(4.2.21)

where \({\mathcal {R}} _{\alpha \overline{\beta }}\) are the coefficients of the Ricci tensor of \(\omega _{{\mathcal {C}}}\).

The last term in (4.2.20) is simply \(\displaystyle \theta _{\alpha \overline{\beta }}\xi _{\overline{\alpha }}\overline{\xi _{\overline{\beta }}}\) where \(\theta _{\alpha \overline{\beta }}\) are the coefficients of \(\Theta _{h_L}(L)\).

Again by the Kähler hypothesis the term (4.2.17) is equal to

$$\begin{aligned} {\mathcal {R}} _{\alpha \overline{\beta }}\xi _{\overline{\alpha }}\overline{\xi _{\overline{\beta }}} \end{aligned}$$
(4.2.22)

and therefore we obtain

$$\begin{aligned} \Delta ''(|\xi _1|^2)&= |\nabla \xi _1|^2- 2\Re \langle \square \xi _1, \xi _1\rangle \end{aligned}$$
(4.2.23)
$$\begin{aligned}&\quad + 2\sum _{\alpha , \beta }\left( \theta _{\alpha \overline{\beta }}- {\mathcal {R}} _{\alpha \overline{\beta }}\right) \xi _{\overline{\alpha }}\overline{\xi _{\overline{\beta }}}+ \sum _{\alpha , \beta } {\mathcal {R}} _{\alpha \overline{\beta }} \xi _{\overline{\alpha }}\overline{\xi _{\overline{\beta }}} \end{aligned}$$
(4.2.24)
$$\begin{aligned}&\quad - \sum _{\alpha , \beta }\left( \theta _{\alpha \overline{\beta }}- {\mathcal {R}} _{\alpha \overline{\beta }}\right) g^{\overline{\beta }\alpha }|\xi _1|^2 \end{aligned}$$
(4.2.25)

by collecting the previous equalities at \(x_0\).

The Ricci curvature of the metric \(\omega _{{\mathcal {C}}}\) is uniformly bounded, so the function

$$\begin{aligned} f^2:= |\xi _1|^2 \end{aligned}$$
(4.2.26)

(where the norm is measured with respect to \(\omega _{{\mathcal {C}}}\) and \(h_L\)) verifies the following properties:

  1. (1)

    We have \(\sup _{V_{\textrm{sing}}\setminus W}f< \infty \), and moreover f is smooth on \(V_{\textrm{sing}}\setminus W\).

  2. (2)

    The following differential inequality holds true

    $$\begin{aligned} \Delta ''f^2\geqslant |\nabla f|^2- Cf^2 \end{aligned}$$
    (4.2.27)

    where C is a constant depending on the Ricci curvature of the metric \(\omega _{{\mathcal {C}}}\) and the trace of \(dd^c\tau \) with respect to it.

Indeed the inequality at the point (2) follows from (4.2.23), since we have

$$\begin{aligned} |\nabla \xi _1|^2= \big |\nabla |\xi _1|\big |^2. \end{aligned}$$

Based on (1) and (2) we can conclude in two ways: either show that Schoen–Yau mean value inequality holds for functions f which verify these properties (the proof would be a simple adaptation of the arguments presented in [37]), or use the Moser iteration procedure. In what follows, we use Moser procedure.

We show next that the following statement holds true.

Lemma 4.3

Let f be the function defined in (4.2.26). Then there exists a constant \(C_1\) depending on C and \((V_{\textrm{sing}}, \omega _{{\mathcal {C}}})\) only such that we have

$$\begin{aligned} \sup _{\frac{1}{2}V_{\textrm{sing}} \setminus W}f^2 \leqslant C_1\int _{V_\textrm{sing}} f^2\textrm{d}V_{\omega _{{\mathcal {C}}}}. \end{aligned}$$
(4.3.1)

Remark that the main point here is that the constant \(C_1\) is independent of the \(\sup \) norm in (1). After establishing this statement we are basically done, i.e., this implies Theorem 4.2 announced at the beginning of the current section.

Proof of Lemma 4.3

Let \(\rho \) be a function which is equal to 1 on \(\displaystyle 1/2V_{\textrm{sing}}\) and whose support is in \(V_{\textrm{sing}}\). Then we have

$$\begin{aligned} |\Delta \rho |\leqslant C, \qquad |\textrm{d}\rho | \leqslant C \end{aligned}$$
(4.3.2)

where the norm of the differential in (4.3.2) is measured with respect to the reference metric \(\omega \) on X.

Following [3], there exists a family of functions \((\Xi _\varepsilon )_{\varepsilon > 0}\) associated to the analytic subset \(W= (h_L= \infty )\cap V_{\textrm{sing}}\) such that \(\textrm{Supp}(\Xi _\varepsilon )\subset V_{\textrm{sing}}{\setminus } W\) and for each compact subset \(K\subset X\setminus W\) we have \(\Xi _\varepsilon |_K= 1\) if \(\varepsilon < \varepsilon (K)\) is small enough. Moreover we have

$$\begin{aligned} \int _X|d (\Xi _\varepsilon )|^2\textrm{d}V_\omega \rightarrow 0, \qquad \int _X|\Delta (\Xi _\varepsilon )|\textrm{d}V_\omega \rightarrow 0 \end{aligned}$$
(4.3.3)

as \(\varepsilon \rightarrow 0\). We recall very briefly the construction: let \(\rho _\varepsilon \) be a function which is equal to one on the interval \([0, \varepsilon ^{-1}]\) and which equals zero on \([1+ \varepsilon ^{-1}, \infty [\). Then we define

$$\begin{aligned} \Xi _\varepsilon := \rho _\varepsilon \left( \log \big (\log \frac{1}{|s_W|^{2}}\big )\right) \end{aligned}$$
(4.3.4)

where \(s_W\) is the sections whose zero set is W. Then with respect to the conic metric \(\omega _{{\mathcal {C}}}\) we have

$$\begin{aligned} |d (\Xi _\varepsilon )|^2_{\omega _{{\mathcal {C}}}} \leqslant \frac{\rho _\varepsilon '}{\log ^2|s_W|^2}\sum \frac{1}{|z_j|^{2/k_j}}, \end{aligned}$$
(4.3.5)

up to a constant, from which (4.3.3) follows (we get a similar inequality for the Laplacian of \(\Xi _\varepsilon \)).

The existence of \((\Xi _\varepsilon )_{\varepsilon > 0}\) combined with the second property in (4.3.2) allows us to deal with the fact that f is not necessarily smooth.

The proof which follows is rather standard, but we will nevertheless provide a complete treatment for convenience. We refer to [16] for a general discussion about Moser iteration method, and more specifically to [22] where this is implemented in a context which is very similar to ours here.

We multiply the differential inequality (4.2.27) with \(\Xi _\varepsilon \rho ^2\) and integrate the result over X; we infer that we have

$$\begin{aligned} \int _X \Xi _\varepsilon \rho ^2\Delta (f^2)\textrm{d}V_{\omega _{{\mathcal {C}}}}\geqslant \int _X \Xi _\varepsilon \rho ^2|d(f)|^2\textrm{d}V_{\omega _{{\mathcal {C}}}}. \end{aligned}$$
(4.3.6)

On the LHS we integrate by parts. The terms containing derivatives of \(\Xi _\varepsilon \) are

$$\begin{aligned} \int _X|\Delta (\Xi _\varepsilon )|\rho ^2 f^2\textrm{d}V_{\omega _{{\mathcal {C}}}}, \int _X\langle \textrm{d}\Xi _\varepsilon , d(\rho ^2)\rangle f^2\textrm{d}V_{\omega _{{\mathcal {C}}}} \end{aligned}$$
(4.3.7)

and they tend to zero precisely because of the uniform boundedness property (1) of f, together with (4.3.3). These terms are vanishing as \(\varepsilon \rightarrow 0\), and the inequality (4.3.6) becomes

$$\begin{aligned} \int _X f^2\Delta (\rho ^2)\textrm{d}V_{\omega _{{\mathcal {C}}}}\geqslant \int _X \rho ^2|d(f)|^2\textrm{d}V_{\omega _{{\mathcal {C}}}}. \end{aligned}$$
(4.3.8)

On the other hand we write

$$\begin{aligned} \int _X \rho ^2|d f|^2\textrm{d}V_{\omega _{{\mathcal {C}}}}\geqslant \frac{1}{2}\int _X \big |d (\rho f)\big |^2\textrm{d}V_{\omega _{{\mathcal {C}}}} - \int _X f^2|d \rho |^2 \textrm{d}V_{\varepsilon }, \end{aligned}$$
(4.3.9)

which combined with (4.3.8) gives

$$\begin{aligned} \int _X \big |\nabla (\rho f)\big |^2\textrm{d}V_{\omega _{{\mathcal {C}}}}\leqslant C\int _{V_\textrm{sing}} f^2 \textrm{d}V_{\omega _{{\mathcal {C}}}}, \end{aligned}$$
(4.3.10)

where the constant C in (4.3.10) only depends on the norm of the first and second derivatives of \(\rho \).

The following version of Sobolev inequality is a direct consequence of [21], page 153.

Theorem 4.4

There exists a constant \(C>0 \) such that the following holds

$$\begin{aligned} \frac{1}{C} \left( \int _X|f|^{\frac{2n}{n-1}}\textrm{d}V_{\omega _{{\mathcal {C}}}}\right) ^{\frac{n-1}{n}}\leqslant \int _X|f|^2\textrm{d}V_{\omega _{{\mathcal {C}}}}+ \int _X|\nabla f|^2\textrm{d}V_{\omega _{{\mathcal {C}}}} \end{aligned}$$
(4.4.1)

for any function f on X.

We therefore infer that we have

$$\begin{aligned} \left( \int _{1/2V_\textrm{sing}}|f|^{\frac{2n}{n-1}}\textrm{d}V_{\omega _{{\mathcal {C}}}}\right) ^{\frac{n-1}{n}} \leqslant C \int _{V_{\textrm{sing}}} |f|^2 \textrm{d}V_{{\omega _{{\mathcal {C}}}}}. \end{aligned}$$
(4.4.2)

In order to obtain estimates for higher norms, we use (4.4.1) for \(f:= \Xi _\varepsilon \rho f^{\frac{p}{2}}\) and we obtain

$$\begin{aligned} \frac{1}{C}\left( \int _X(\Xi _\varepsilon \rho )^{\frac{2n}{n-1}}f^{\frac{pn}{n-1}}\textrm{d}V_{\omega _{{\mathcal {C}}}}\right) ^{\frac{n-1}{n}}\leqslant & {} \int _X(\Xi _\varepsilon \rho )^2f^p\textrm{d}V_{\omega _{{\mathcal {C}}}}\nonumber \\ {}{} & {} + \int _X\left| \nabla \left( \Xi _\varepsilon \rho f^{\frac{p}{2}}\right) \right| ^2\textrm{d}V_{\omega _{{\mathcal {C}}}}. \end{aligned}$$
(4.4.3)

We show now that the second term of the right-hand side of (4.4.3) verifies the inequality

$$\begin{aligned} \int _X\left| \nabla \left( \rho f^{\frac{p}{2}}\right) \right| ^2\textrm{d}V_{\omega _{{\mathcal {C}}}}\leqslant Cp\int _X(\rho ^2+ |\nabla \rho |^2)f^p \textrm{d}V_{\omega _{{\mathcal {C}}}}. \end{aligned}$$
(4.4.4)

This is done using integration by parts: we have

$$\begin{aligned} \nabla \left( \rho \Xi _\varepsilon f^{\frac{p}{2}}\right) = f^{\frac{p}{2}}\nabla (\rho \Xi _\varepsilon )+ \frac{p}{2}\Xi _\varepsilon \rho f^{\frac{p-2}{2}}\nabla f, \end{aligned}$$

so we have to obtain a bound for the term

$$\begin{aligned} \int _X(\rho \Xi _\varepsilon )^2f^{{p-2}}|\nabla f|^2\textrm{d}V_{\omega _{{\mathcal {C}}}}= \frac{1}{2}\int _X(\rho \Xi _\varepsilon )^2f^{{p-3}}\langle \nabla f^2, \nabla f|\rangle \textrm{d}V_{\omega _{{\mathcal {C}}}}. \end{aligned}$$
(4.4.5)

We have

$$\begin{aligned} (p-2)&\int _X(\rho \Xi _\varepsilon )^2f^{{p-3}}\langle \nabla f^2, \nabla f\rangle \textrm{d}V_{\omega _{{\mathcal {C}}}} \nonumber \\&= - \int _X(\rho \Xi _\varepsilon )^2f^{{p-2}}\Delta f^2\textrm{d}V_{\omega _{{\mathcal {C}}}} - 2\int _Xf^{{p-2}}\langle (\rho \Xi _\varepsilon )\nabla f, f\nabla (\rho \Xi _\varepsilon )\rangle \textrm{d}V_{\omega _{{\mathcal {C}}}} \nonumber \\&\leqslant -\int _X(\rho \Xi _\varepsilon )^2f^{{p-2}}|\nabla (f)|^2\textrm{d}V_{\omega _{{\mathcal {C}}}} - 2\int _Xf^{{p-2}}\langle (\rho \Xi _\varepsilon )\nabla f, f\nabla (\rho \Xi _\varepsilon )\rangle \textrm{d}V_{\omega _{{\mathcal {C}}}} \nonumber \\&\quad + C\int _X(\rho \Xi _\varepsilon )^2f^{{p}}\textrm{d}V_{\omega _{{\mathcal {C}}}} \nonumber \\&\leqslant C\int _X((\rho \Xi _\varepsilon )^2+ |\nabla (\rho \Xi _\varepsilon )|^2)f^p\textrm{d}V_{\omega _{{\mathcal {C}}}} \end{aligned}$$
(4.4.6)

and as before, the terms involving \(\nabla (\Xi _\varepsilon )\) tend to zero as \(\varepsilon \rightarrow 0\). We therefore get the inequality (4.4.4). Remark that we are using the inequality (4.2.27) in order to obtain (4.4.6).

We define \(V_i:= (1/2+ 1/2^i)V_{\textrm{sing}}\) and let \(\rho _i\) be a cutoff function such that \(\rho _i= 1\) on \(V_{i+1}\) and such that \(\textrm{Supp}(\rho _i)\subset V_i\). Then we have \(|\nabla \rho _i|\leqslant C2^i\), and by using (4.4.4) combined with the usual iteration process, Lemma 4.3 follows. \(\square \)

Theorem 4.2 is proved. \(\square \)

Remark 4.5

Actually a careful examination of the proof shows that one can obtain a constant “C” in Lemma 4.3 as follows:

$$\begin{aligned} C= \frac{C(X, \omega )}{\textrm{Vol} \,(V_{\textrm{sing}})}. \end{aligned}$$
(4.5.1)

If necessary, this can be obtained by adapting the arguments of Schoen–Yau in [37, p. 75].

5 Proof of the Main Results

5.1 Proof of Theorem 1.1

We consider the “usual” family of cutoff functions

$$\begin{aligned} \rho _\varepsilon : X\rightarrow {\mathbb {R}},\qquad \rho _\varepsilon (z)= \rho \left( \frac{|s_Y|^2}{\varepsilon ^2}\right) \end{aligned}$$
(5.0.2)

where \(\rho \) is a function defined on the set of positive real numbers such that \(\rho = 1\) on [0, 1] and \(\rho = 0\) on \([2, \infty [\).

We will show here that the following a-priori inequality holds

$$\begin{aligned} \left| \int _X{{\bar{\partial }}}\left( \rho _\varepsilon U_0\right) \wedge \overline{\gamma _{\xi }}e^{-\phi _L}\right| ^2 \leqslant C_\varepsilon (U_0) \int _X\prod \log ^2(|s_{Y_j}|^2+ \varepsilon ^2)|{{\bar{\partial }}}^\star \xi |^2e^{-\phi _L}\textrm{d}V_{\omega }\nonumber \\ \end{aligned}$$
(5.0.3)

where \(\xi \) is a (n, 1)-form with values in \(Y+ L\) whose support is contained in the complement of \(V_{\textrm{sing}}\cup H\), and \(U_0\) is an arbitrary holomorphic extension of u, cf. [7]. Also, we denote by \(\phi _L\) the metric

$$\begin{aligned} \phi _L:= \varphi _L+ \log |f_Y|^2 \end{aligned}$$
(5.0.4)

on the bundle \(L+ Y\). We will see that the constant \(C_\varepsilon (U_0)\) in (5.0.3) is explicit, and it converges to the RHS of (1.6.1) as \(\varepsilon \rightarrow 0\). Note that all the integrals above are at least well defined, given the condition we impose on the support of \(\xi \).

The proof of (5.0.3) will be presented along the following line of arguments.

\(\bullet \) Consider a (n, 1)-form \(\xi \) as above together with the orthogonal decomposition \(\frac{\xi }{s_Y}= \xi _1+ \xi _2\) we have already discussed in detail in the previous section. Then we show that we have

$$\begin{aligned}{} & {} c_{n}\int _X\frac{\varepsilon ^2}{(\varepsilon ^2+ |s_Y|^2)^2}\gamma _{\xi _1}\wedge \partial s_Y\wedge \overline{\gamma _{\xi _1}\wedge \partial s_Y}e^{-\varphi _L}\nonumber \\ {}{} & {} \quad \leqslant \int _X\prod \log ^2(|s_{Y_j}|^2+ \varepsilon ^2)|{{\bar{\partial }}}^\star \xi |^2e^{-\phi _L}\textrm{d}V_{\omega _L}. \end{aligned}$$
(5.0.5)

up to a numerical constant. This will be done by an approximation argument, using Lemma 4.1 as well as Theorem 3.1.

\(\bullet \) The norm of the functional on the LHS of (5.0.3) is evaluated in two steps on the set \(V_{\textrm{sing}}\) we use Theorem 4.2, combined with a few simple calculations. In the complement \(X\setminus V_{\textrm{sing}}\) the arguments are rather standard: we will use (5.0.5).

The remaining part of the current section is organized as follows. We first show that (5.0.3) implies the existence of an “estimable extension” of u. Then we prove that the estimate (5.0.3) holds true.

5.1.1 Functional Analysis

Our method relies on the next statement.

Theorem 5.1

Let \(u\in H^0 (Y, (K_X +Y+ L) |_Y)\) be a holomorphic section. We assume that there is a constant \(C_\varepsilon (U)\) such that for any L-valued smooth test form \(\xi \) of type (n, 1) with compact support in \(X\setminus (V_{\textrm{sing}}\cup |W|)\) the a-priori inequality (5.0.3) holds. Then u admits an extension \(U\in H^0 (X, K_X+Y+ L)\) such that

$$\begin{aligned} \int _{X\setminus V_\textrm{sing}}\frac{|U|^2}{|s_Y|^{2}\prod \log ^2(|s_{Y_j}|^2)}e^{-\varphi _L- \varphi _Y}\textrm{d}V_\omega \leqslant \lim _{\varepsilon \rightarrow 0}C_\varepsilon (U). \end{aligned}$$
(5.1.1)

Proof

This is done as in the classical case, by considering the vector subspace

$$\begin{aligned} E:= \big \{{{\bar{\partial }}}^\star \xi : \xi \in C^2_c\big (X\setminus (V_\textrm{sing}\cup H)\big )\big \} \end{aligned}$$
(5.1.2)

of the \(L^2_{n,0}(X, Y+L)\) forms, endowed with the scalar product induced by

$$\begin{aligned} \Vert \rho \Vert ^2:= \int _X|\rho |^2 e^{-\phi _L}\prod \log ^2(|s_{Y_j}|^2+ \varepsilon ^2)\textrm{d}V_{\omega }. \end{aligned}$$
(5.1.3)

The functional

$$\begin{aligned} {{\bar{\partial }}}^\star \xi \rightarrow \int _X{{\bar{\partial }}}\left( \rho _\varepsilon U_0\right) \wedge \overline{\gamma _{\xi }}e^{-\phi _L} \end{aligned}$$
(5.1.4)

is well defined and bounded on E by (5.0.3), hence it extends by Hahn–Banach. The representation theorem of Riesz implies that there exists some form

$$\begin{aligned} v\in L^2_{n,0}(X, Y+ L) \end{aligned}$$

such that we have

$$\begin{aligned} \int _X\left\langle {{\bar{\partial }}}\left( \rho _\varepsilon U_0\right) , {\xi }\right\rangle e^{-\phi _L}= \int _X \left\langle v, {{\bar{\partial }}}^\star \xi \right\rangle e^{-\phi _L}\prod \log ^2(|s_{Y_j}|^2+ \varepsilon ^2)\textrm{d}V_{\omega _L} \end{aligned}$$
(5.1.5)

for all test forms \(\xi \in E\) and such that

$$\begin{aligned} \int _X|v|^2e^{-\phi _L}\prod \log ^2(|s_{Y_j}|^2+ \varepsilon ^2)\textrm{d}V_{\omega _L}\leqslant C_\varepsilon (U_0). \end{aligned}$$
(5.1.6)

Equation (5.1.5) shows that we have

$$\begin{aligned} {{\bar{\partial }}}\left( \rho _\varepsilon U_0\right) = {{\bar{\partial }}}\left( \prod \log ^2(|s_{Y_j}|^2+ \varepsilon ^2)v\right) \end{aligned}$$
(5.1.7)

on \(X\setminus V_{\textrm{sing}}\). On the other hand, the form

$$\begin{aligned} \rho _\varepsilon U_0- \left( \prod \log ^2(|s_{Y_j}|^2+ \varepsilon ^2)\right) v \end{aligned}$$
(5.1.8)

is in \(L^2(X\setminus V_{\textrm{sing}})\): this is clear for the first term, as for the second one it is a consequence of (5.1.6).

We infer that the form

$$\begin{aligned} U_\varepsilon := \rho _\varepsilon U_0- \left( \prod \log ^2(|s_{Y_j}|^2+ \varepsilon )\right) v \end{aligned}$$
(5.1.9)

extends holomorphically on \(X\setminus V_{\textrm{sing}}\). This implies that \(v|_{X\setminus V_{\textrm{sing}}}\) is non-singular, in particular v equal zero when restricted to \(Y{\setminus } V_\textrm{sing}\)—given the estimates in (5.1.6).

Therefore we infer the equality

$$\begin{aligned} U_\varepsilon |_{Y\setminus V_{\textrm{sing}}}= u. \end{aligned}$$
(5.1.10)

We remark that \(U_\varepsilon \) extends to X by theorem of Hartog’s. This is clear if X is a surface cf. e.g., [19, Theorem 2.3.2]. The general case follows as a consequence of this, by a simple argument of slicing which we will not detail here.

Finally, the estimate for the \(L^2\) norm of U is deduced from (5.1.6): we have

$$\begin{aligned} \int _{X\setminus V_\textrm{sing}}\frac{|U_\varepsilon |^2}{|s_Y|^2\prod \log ^2(|s_{Y_j}|^2+ \varepsilon )}e^{-\varphi _L- \varphi _Y} \leqslant C_\varepsilon (U_0) \end{aligned}$$
(5.1.11)

modulo a quantity which tends to zero as \(\varepsilon \rightarrow 0\). The conclusion follows. \(\square \)

5.1.2 End of the Proof

We prove now the inequality (5.0.3). As we have already mentioned, one of the main part of the proof is based on the a-priori estimate (5.0.5) which we derive here from Theorem 3.2 combined with the results established in the first part of Sect. 3.

We start with the following technical result, which plays a key role in the arguments to come. In order to simplify the notations, we write \(\xi \) instead of the quotient \(\displaystyle \frac{1}{s_Y}\xi \).

Proposition 5.2

Consider the orthogonal decomposition \(\xi = \xi _1+ \xi _2\). Then the following hold: for each positive \(\varepsilon \) we have

$$\begin{aligned} \sum _i c_{n-1}\int _X&\frac{\varepsilon ^2}{(\varepsilon ^2+ |s_Y|^2)^2}\gamma _{\xi _1}\wedge \overline{\gamma _{\xi _1}}e^{-\varphi _L}\wedge \sqrt{-1}\partial s_i\wedge \overline{\partial s_i} \nonumber \\&\leqslant C\int _X\prod \log ^2(\varepsilon ^{2/N}+ |s_{j}|^2)\left( |{{\bar{\partial }}}^\star \xi |^2\right) e^{-\varphi _L}\textrm{d}V_{\omega _L} \end{aligned}$$
(5.2.1)

where N is the number of components of Y.

Remark that we have the equality \(\displaystyle |{{\bar{\partial }}}^\star \mu |^2e^{-\varphi _L}= |{{\bar{\partial }}}^\star \xi |^2e^{-\phi _L}\) if \(\mu = \frac{1}{s_Y}\xi \), so the estimate (5.2.1) is precisely what we have to prove.

Proof

We recall the context in Sect. 3: we have considered an exhaustion

$$\begin{aligned} X\setminus H= \bigcup \Omega _m \end{aligned}$$
(5.2.2)

where each \(\Omega _m\) was a Stein domain with smooth boundary, endowed with the family of complete metrics \(\omega _{m, \delta }\) cf. (4.0.4). The restriction of \(\xi \) to each \(\Omega _m\) decomposes as follows:

$$\begin{aligned} \xi |_{\Omega _m}= \xi _1^{(m,\delta )}+ \xi _2^{(m, \delta )} \end{aligned}$$
(5.2.3)

according to \((\Omega _m, \omega _{m, \delta })\) and \((L, h_L)\).

We apply the inequality in Theorem 3.1 for \(\xi _1^{(m,\delta )}\) and we get

$$\begin{aligned} \sum _i c_{n-1}\int _{\Omega _m}&\frac{\varepsilon ^2}{(\varepsilon ^2+ |s_Y|^2)^2} \gamma _{\xi _1^{(m,\delta )}}\wedge \overline{\gamma _{\xi ^{(\varepsilon )}_1}}e^{-\varphi _L}\wedge \sqrt{-1}\partial s_Y\wedge \overline{\partial s_Y} \nonumber \\&\leqslant C\int _{\Omega _m}\prod \log ^2(\varepsilon ^{2/N}+ |s_{j}|^2)\left| {{\bar{\partial }}}^\star (\xi _1^{(m,\delta )})\right| ^2e^{-\varphi _{L}}\textrm{d}V_{\omega _{m, \delta }}. \end{aligned}$$
(5.2.4)

Indeed we can use Theorem 3.1 in this context even if the form does not have compact support because the metric \(\omega _{m, \delta }\) is complete. This has another consequence: we have the equality \({{\bar{\partial }}}^\star (\xi _1^{(m,\delta )})= {{\bar{\partial }}}^\star (\xi )\). For the inequality (5.2.4) we have used the inequality

$$\begin{aligned} \log ^2(\varepsilon ^2+ |s_Y|^2)\leqslant C \prod \log ^2(\varepsilon ^{2/N}+ |s_{j}|^2), \end{aligned}$$

where N is the number of components of Y.

Let \(K\subset X\) be any open set with compact closure in \(X{\setminus } H\); for any \(m\geqslant m_0(K)\) we have \(\overline{K}\subset \Omega _m\) so the inequality (5.2.4) implies

$$\begin{aligned} \sum _i c_{n-1}\int _{K}&\frac{\varepsilon ^2}{(\varepsilon ^2+ |s_Y|^2)^2} \gamma _{\xi _1^{(m,\delta )}}\wedge \overline{\gamma _{\xi _1^{(m,\delta )}}}e^{-\varphi _L}\wedge \sqrt{-1}\partial s_Y\wedge \overline{\partial s_Y} \nonumber \\&\leqslant C\int _{\Omega _m}\prod \log ^2(\varepsilon ^{2/N}+ |s_{j}|^2)\left| {{\bar{\partial }}}^\star \xi \right| _{\omega _{m, \delta }}^2e^{-\varphi _{L}}\textrm{d}V_{\omega _{m, \delta }}. \end{aligned}$$
(5.2.5)

Now the support of \(\xi \) is a compact contained in \(X\setminus H\), so if m is large enough the boundary of \(\Omega _m\) is disjoint from \(\textrm{Supp}(\xi )\). A limit process (i.e., \(\delta \rightarrow 0, m\rightarrow \infty \)), together with Lemma 4.1 implies that we have

$$\begin{aligned}&\sum _i c_{n-1}\int _{K}\frac{\varepsilon ^2}{(\varepsilon ^2+ |s_Y|^2)^2} \gamma _{\xi _1}\wedge \overline{\gamma _{\xi _1}}e^{-\varphi _L}\wedge \sqrt{-1}\partial s_Y\wedge \overline{\partial s_Y} \nonumber \\&\quad \leqslant C\int _X \prod \log ^2(\varepsilon ^{2/N}+ |s_{j}|^2)\left| {{\bar{\partial }}}^\star \xi \right| _{\omega _{{\mathcal {C}}}}^2e^{-\varphi _{L}}\textrm{d}V_{\omega _{{\mathcal {C}}}}. \end{aligned}$$
(5.2.6)

The compact subset K in (5.2.6) is arbitrary, so Proposition 5.2 is proved. \(\square \)

We are now ready to finish the proof of Theorem 1.1. Consider the integral

$$\begin{aligned} \int _X\left\langle {{\bar{\partial }}}\left( \rho _\varepsilon U_0\right) , {\xi }\right\rangle e^{-\phi _L} \textrm{d}V_{\omega _{{\mathcal {C}}}} \end{aligned}$$
(5.2.7)

which up to a sign equals

$$\begin{aligned} \int _X\rho '\left( \frac{|s_Y|^2}{\varepsilon ^2}\right) \left\langle U_0\wedge \frac{\overline{\partial s_Y}}{\varepsilon ^2}, {\frac{\xi }{s_Y}}\right\rangle e^{-\varphi _L} \textrm{d}V_{\omega _{{\mathcal {C}}}}. \end{aligned}$$
(5.2.8)

We decompose as usual \(\frac{\xi }{s_Y}= \xi _1+ \xi _2\) and then (5.2.8) becomes

$$\begin{aligned} \int _X\rho '\left( \frac{|s_Y|^2}{\varepsilon ^2}\right) \left\langle U_0\wedge \frac{\overline{\partial s_Y}}{\varepsilon ^2}, \xi _1\right\rangle e^{-\varphi _L} \textrm{d}V_{\omega _{{\mathcal {C}}}}= \int _X\rho '\left( \frac{|s_Y|^2}{\varepsilon ^2}\right) U_0\wedge \overline{\partial s_Y\wedge \gamma _{\xi _1} }\frac{e^{-\varphi _L}}{\varepsilon ^2}\nonumber \\ \end{aligned}$$
(5.2.9)

We split its evaluation into two parts. The first one is

$$\begin{aligned} \int _{X\setminus V_\textrm{sing}}\rho '\left( \frac{|s_Y|^2}{\varepsilon ^2}\right) U_0\wedge \overline{\partial s_Y\wedge \gamma _{\xi _1} }\frac{e^{-\varphi _L}}{\varepsilon ^2} \end{aligned}$$
(5.2.10)

and by Cauchy–Schwarz inequality the square of its absolute value is smaller than

$$\begin{aligned} \int _{K_\varepsilon } |U_0|^2\frac{e^{-\varphi _L}}{\varepsilon ^2}\textrm{d}V_{\omega _{{\mathcal {C}}}}\cdot \int _{K_{\varepsilon }} |\partial s_Y\wedge \gamma _{\xi _1}|^2_{\omega _{{\mathcal {C}}}}\frac{e^{-\varphi _L}}{\varepsilon ^2}\textrm{d}V_{\omega _{{\mathcal {C}}}}, \end{aligned}$$
(5.2.11)

where \(K_\varepsilon \) is the support of the function \(\rho '\left( \frac{|s_Y|^2}{\varepsilon ^2}\right) \). We remark that we have

$$\begin{aligned} \frac{1}{\varepsilon ^2}\simeq \frac{\varepsilon ^2}{(\varepsilon ^2+ |s_Y|^2)^2} \end{aligned}$$
(5.2.12)

on the set \(K_\varepsilon \). Therefore, the second factor of the product (5.2.11) is smaller than

$$\begin{aligned} C\int _X\prod \log ^2(\varepsilon ^2+ |s_{j}|^2)\left( |{{\bar{\partial }}}^\star \xi |^2\right) e^{-\varphi _L}\textrm{d}V_{\omega _{{\mathcal {C}}}} \end{aligned}$$
(5.2.13)

by Proposition 5.2.

The rest of the integral (5.2.9) is analyzed as follows. For simplicity we assume that \(V_{\textrm{sing}}= \Omega \) is a coordinate subset and the expression we have to evaluate is

$$\begin{aligned} \left| \frac{1}{\varepsilon ^2}\int _{\Omega }\rho '\left( \frac{|s_Y|^2}{\varepsilon ^2}\right) U_0\wedge \overline{\partial s_Y\wedge \gamma _{\xi _1} }e^{-\varphi _L}\textrm{d}V_{{\omega _{{\mathcal {C}}}}}\right| . \end{aligned}$$
(5.2.14)

This is bounded by the quantity

$$\begin{aligned} \sup _\Omega (|\xi _1|^\alpha ) \left| \frac{1}{\varepsilon ^2}\int _{\Omega }\rho '\left( \frac{|s_Y|^2}{\varepsilon ^2}\right) |\partial s_Y|^\alpha |U_0|_{\omega _{{\mathcal {C}}}} |\partial s_Y\wedge \gamma _{\xi _1}|^{1-\alpha }e^{-(1-\alpha /2)\varphi _L}\textrm{d}V_{{\omega _{{\mathcal {C}}}}}\right| \nonumber \\ \end{aligned}$$
(5.2.15)

and Hölder inequality shows that (5.2.15) is smaller than the product of

$$\begin{aligned} \sup _\Omega (|\xi _1|^\alpha ) \left( \int _{\Omega \cap K_\varepsilon } |\partial s_Y\wedge \gamma _{\xi _1}|^{2}\frac{e^{-\varphi _L}}{\varepsilon ^2}\textrm{d}V_{{\omega _{{\mathcal {C}}}}}\right) ^{\frac{1-\alpha }{2}} \end{aligned}$$
(5.2.16)

with

$$\begin{aligned} \left( \int _{\Omega } \rho '\left( \frac{|s_Y|^2}{\varepsilon ^2}\right) |\partial s_Y|^{\frac{2\alpha }{1+\alpha }} |U_0|_{\omega _{{\mathcal {C}}}}^{\frac{2}{1+\alpha }} \frac{e^{-\frac{\varphi _L}{1+\alpha }}}{\varepsilon ^2}\textrm{d}V_{{\omega _{{\mathcal {C}}}}}\right) ^{\frac{1+\alpha }{2}} \end{aligned}$$
(5.2.17)

The limit of the quantity (5.2.17) as \(\varepsilon \rightarrow 0\) is equal to

$$\begin{aligned} \left( \int _{\Omega \cap Y} \left| \frac{u}{\partial s_Y}\right| _{\omega _{{\mathcal {C}}}}^{\frac{2}{1+\alpha }}e^{-\frac{\varphi _L}{1+\alpha }}\textrm{d}V_{{\omega _{{\mathcal {C}}}}}\right) ^{\frac{1+\alpha }{2}}. \end{aligned}$$
(5.2.18)

As for the (5.2.16), we use Theorem 4.2 together with our previous considerations and it follows that it is smaller than

$$\begin{aligned} C\left( \int _\Omega |\xi _1|^2 \textrm{d}V_{\omega _{{\mathcal {C}}}}\right) ^{\alpha /2} \left( \int _X\prod \log ^2(\varepsilon ^{2/N}+ |s_{j}|^2)\left( |{{\bar{\partial }}}^\star \xi |^2\right) e^{-\varphi _L}\textrm{d}V_{{\omega _{{\mathcal {C}}}}}\right) ^{\frac{1-\alpha }{2}}.\nonumber \\ \end{aligned}$$
(5.2.19)

It is at this point that we are using the positivity assumption (i): we have

$$\begin{aligned} \int _{V_{\textrm{sing}}}|\xi _1|^2e^{-\varphi _L}\textrm{d}V_{\omega _{{\mathcal {C}}}}\leqslant \frac{1}{C_{\textrm{sing}}}\int _{V_{\textrm{sing}}} \big \langle [\Theta _{h_L}(L), \Lambda _{\omega _{{\mathcal {C}}}}]\xi _1, \xi _1\big \rangle e^{-\varphi _L}\textrm{d}V_{\omega _{{\mathcal {C}}}},\nonumber \\ \end{aligned}$$
(5.2.20)

where \(C_{\textrm{sing}}\) is the (positive) lower bound for the positivity of \((L, h_L)|_{V_{\textrm{sing}}}\). By Bochner formula we get

$$\begin{aligned} \int _{V_{\textrm{sing}}}|\xi _1|^2e^{-\varphi _L}\textrm{d}V_{{\omega _{{\mathcal {C}}}}}\leqslant \frac{1}{C_{\textrm{sing}}} \int _X\left| {{\bar{\partial }}}^\star \xi _1\right| ^2e^{-\varphi _L}\textrm{d}V_{{\omega _{{\mathcal {C}}}}}. \end{aligned}$$
(5.2.21)

Thus we obtain the expected estimate for the functional (5.2.9) and Theorem 1.1 is proved. We remark that the contribution of the singularities of Y to the estimate in this result is

$$\begin{aligned} C\left( 1+ \frac{1}{C_{\textrm{sing}} ^\alpha }\right) \left( \int _{\Omega \cap Y} \left| \frac{u}{\partial s_Y}\right| _{\omega _{{\mathcal {C}}}}^{\frac{2}{1+\alpha }}e^{-\frac{\varphi _L}{1+\alpha }}\textrm{d}V_{{\omega _{{\mathcal {C}}}}}\right) ^{1+\alpha }, \end{aligned}$$
(5.2.22)

where C is a constant depending on \((X, V_{\textrm{sing}}, {\omega _{{\mathcal {C}}}})\) and \(\alpha \in [0, 1]\) is an arbitrary positive real which is smaller than 1.

Remark 5.3

The quantity (5.2.22) is part of the term estimating

$$\begin{aligned} \int _{X\setminus V_\textrm{sing}}\frac{|U|^2}{|s_Y|^2\prod \log ^2(|s_{j}|^2)}e^{-\varphi _L-\varphi _Y}\textrm{d}V_{\omega _{{\mathcal {C}}}}. \end{aligned}$$
(5.3.1)

A slight modification of the proof shows that we can get a similar estimate for the integral

$$\begin{aligned} \int _{X\setminus V_\textrm{sing}}\frac{|U|^2}{|s_Y|^2\log ^{2+\tau }(1/|s_{Y}|^2)}e^{-\varphi _L-\varphi _Y}\textrm{d}V_{\omega _{{\mathcal {C}}}} \end{aligned}$$
(5.3.2)

for any strictly positive real \(\tau \).

Remark 5.4

Actually one can replace the curvature condition (i) with the following: there exists a constant \(C_{\textrm{sing}}> 0\) such that we have

$$\begin{aligned} \Theta _{h_L}(L)\geqslant \frac{C_\textrm{sing}}{\log \frac{1}{|s_Y|^2}}{\omega _{{\mathcal {C}}}} \end{aligned}$$
(5.4.1)

pointwise on \(V_{\textrm{sing}}\). The estimate for the extension we obtain in the end is the same, but we are using a twisted Bochner formula instead of (5.2.21).

5.2 Proof of Theorem 1.4

We follow the notations in Theorem 1.8. Since \({{\bar{\partial }}}\xi _1 =0\), the Hodge relation \(\partial ^\star _{\varphi _L} = [{{\bar{\partial }}}, \Lambda _{\omega }]\) implies that

$$\begin{aligned} {{\bar{\partial }}}(\gamma _{\xi _1}) = \partial ^\star _{\varphi _L} \xi _1 \qquad \text {on } X. \end{aligned}$$

Let \(\cup \Omega _i\) be a small Stein cover of \(V_{\textrm{sing}}\). The standard \(L^2\)-estimate as well as the Bochner inequality (5.5.7) imply that we can find \(G_i\) satisfying

$$\begin{aligned} {{\bar{\partial }}}G_i = \partial ^\star _{\varphi _L} \xi _1 \text { on } \Omega _i \qquad \text {and} \qquad \int _{ \Omega _i} |G_i|_{\omega , h_L} ^2d V_\omega \leqslant C\int _X |{{\bar{\partial }}}^\star \xi _1|^2_{\omega , h_L}\textrm{d}V_{\omega },\nonumber \\ \end{aligned}$$
(5.4.2)

where C is a uniform constant independent of \(h_L\).

We would like to control the \(C^0\)-norm of \(G_i\). Note first that, as \(i\Theta _{h_L}(L) \) is bounded on \(V_{\textrm{sing}}\), we can found a bounded function \(\varphi _L '\) on \(\Omega _i\) such that

$$\begin{aligned} dd^c\varphi _L = dd^c\varphi _L '. \end{aligned}$$

Therefore \(\varphi _L -\varphi _L'\) is the real part of some holomorphic function \(f_i\) on \(\Omega _i\). Replacing \(\xi _1\) by \(\xi _1 \cdot e^{-\frac{f_L}{2}}\), we can suppose in the beginning that the \(C^2\)-norm of \(\varphi _L\) is bounded by the constant \(C_1\).

Thanks to (5.4.2), \(F_i:= \gamma _{\xi _1} - G_i\) is a holomorphic \((n-1, 0)\)-form on \(\Omega _i\). Recall that \(\xi _1\) is \(\Delta ^{''}\)-harmonic on \(V_{sing}\). Then

$$\begin{aligned} \Delta ^{''} (\omega \wedge G_i) = \Delta ^{''} (\omega \wedge F_i) \qquad \text {on } \Omega _i. \end{aligned}$$
(5.4.3)

Using (5.4.2), the \(H^{-2}\)-norm of \( \Delta ^{''} (\omega \wedge F_i) \) on \(\frac{3}{4}\Omega _i\) is bounded by \(C\int _X |{{\bar{\partial }}}^\star \xi _1|^2_{\omega , h_L}\textrm{d}V_{\omega }\). Since \(F_i\) is holomorphic and the \(C^2\)-norm of \(\varphi _L\) is bounded, \( \Delta ^{''} (\omega \wedge F_i) \) is continuous. Therefore the \(C^0\)-norm of \( \Delta ^{''} (\omega \wedge F_i) \) on \(\frac{3}{4}\Omega _i\) is bounded by \(C\int _X |{{\bar{\partial }}}^\star \xi _1|^2_{\omega , h_L}\textrm{d}V_{\omega }\). As a consequence, the elliptic regularity implies that the

$$\begin{aligned} \sup _{\frac{1}{2}\Omega _i}| G_i|_{\omega , h_L} ^2\leqslant & {} C \left( \int _{ \Omega _i} |G_i|_{\omega , h_L} ^2d V_\omega + \sup _{\frac{3}{4} \Omega _i} | \Delta ^{''} (\omega \wedge F_i) |_{\omega , h_L} ^2\right) \nonumber \\\leqslant & {} C' \int _X |{{\bar{\partial }}}^\star \xi _1|^2_{\omega , h_L}\textrm{d}V_{\omega }, \end{aligned}$$
(5.4.4)

where \(C'\) is a uniform constant.

Now we can prove the theorem. Theorem 3.2 and (5.4.4) imply that

$$\begin{aligned} \int _{Y \cap \frac{1}{2} \Omega _i } F_i \wedge \overline{F_i} e^{-\varphi _L} \leqslant C'\int _X \log ^2 |s_Y|^2 |{{\bar{\partial }}}^\star \xi _1|^2_{\omega , h_L}\textrm{d}V_{\omega }. \end{aligned}$$

Then the mean value inequality for holomorphic functions implies that

$$\begin{aligned} \sup _{Y \cap \frac{1}{3} \Omega _i } |F_i |_Y |_{\omega } ^2 e^{-\varphi _L} \leqslant C'\int _X \log ^2 |s_Y|^2 |{{\bar{\partial }}}^\star \xi _1|^2_{\omega , h_L}\textrm{d}V_{\omega }. \end{aligned}$$

Together with (5.4.4), we have thus

$$\begin{aligned} \sup _{ Y\cap \frac{1}{3} \Omega _i } |\gamma _{\xi _1} |_Y |_{\omega } ^2 e^{-\varphi _L} \leqslant C'\int _X \log ^2 |s_Y|^2 |{{\bar{\partial }}}^\star \xi _1|^2_{\omega , h_L}\textrm{d}V_{\omega }. \end{aligned}$$
(5.4.5)

As a consequence, we have

$$\begin{aligned} \left| \int _{Y\cap V_{\textrm{sing}}}\frac{u}{\textrm{d}s_Y}\wedge \overline{\gamma _{\xi _1}}e^{-\varphi _L}\right|\leqslant & {} C' \int _{Y\cap V_{\textrm{sing}}} |\frac{u}{\textrm{d}s_Y}|_\omega e^{-\frac{\varphi _L}{2}}\textrm{d}V_{\omega }\nonumber \\ {}{} & {} \times \left( \int _X \log ^2 |s_Y|^2 |{{\bar{\partial }}}^\star \xi _1|^2_{\omega , h_L}\textrm{d}V_{\omega }\right) ^{\frac{1}{2}}. \end{aligned}$$

Like the arguments as in the above theorems, this implies the existence of an holomorphic extension U with the estimates

$$\begin{aligned} \frac{1}{C'}\int _{X\setminus V_{\textrm{sing}}} |U|^2e^{-\varphi _Y- \varphi _L} \leqslant \int _{Y\setminus V_{\textrm{sing}}}\left| \frac{u}{\textrm{d}s}\right| ^2e^{-\varphi _L} + \left( \int _{Y\cap V_{\textrm{sing}}} \left| \frac{u}{\textrm{d}s}\right| e^{-\frac{\varphi _L}{2}}\textrm{d}V_{\omega }\right) ^2, \end{aligned}$$
(5.4.6)

where \(C'\) depends only on the geometry of \((V_\textrm{sing}, \omega )\) and \(C_1\).

5.3 Proof of Theorem 1.7

We are using the notations from the previous section, so \(\omega \) is a fixed reference Kähler metric on X.

By hypothesis, the metric \(h_L\) is non-singular and in this case the equality

$$\begin{aligned} \int _Y\frac{u}{\textrm{d}s_Y}\wedge \overline{\gamma _{\xi }}e^{-\varphi _L}= \int _Y\frac{u}{\textrm{d}s_Y}\wedge \overline{\gamma _{\xi _1}}e^{-\varphi _L} \end{aligned}$$
(5.4.7)

is immediate.

We remark that the restriction \(\displaystyle \frac{u}{\textrm{d}s_Y}\Big |_{Y_j}\) is holomorphic, for each component \(Y_j\) of Y. This is where the vanishing of u on the singularities of Y is used. We decompose the restriction of \(\displaystyle \gamma _{\xi _1}\) to \(Y_j\) as follows:

$$\begin{aligned} \gamma _{\xi _1}|_{Y_j}= \alpha _j+ \beta _j \end{aligned}$$
(5.4.8)

where \(\alpha _j\) is holomorphic and \(\beta _j\) is orthogonal to the space of L-valued holomorphic top forms on \(Y_j\). Then we have

$$\begin{aligned} \int _{Y_j}\frac{u}{\textrm{d}s_Y}\wedge \overline{\gamma _{\xi _1}}e^{-\varphi _L}= \int _{Y_j}\frac{u}{\textrm{d}s_Y}\wedge \overline{\alpha _j}e^{-\varphi _L} \end{aligned}$$
(5.4.9)

Let \(x_0\) be a singular point of Y. We then have coordinates \((z_1, z_2)\) defined on a open subset \(x_0\in V\), centered at \(x_0\) and such that \((z_k= 0)= Y_k\cap V\) for each \(k=1, 2\). We equally fix a trivialization of \(L|_V\) and let \(\varphi _L\) be the corresponding weight of the metric \(h_L\). We write

$$\begin{aligned} u|_{V}= f_udz_1\wedge dz_2\otimes e_L \end{aligned}$$
(5.4.10)

and let \(\theta \) be a function which is equal to 1 near \(x_0\) and such that \(\textrm{Supp}(\theta )\subset V\).

Thus we have

$$\begin{aligned} \frac{u}{\textrm{d}s_Y}\big |_{V\cap Y_1}= f_u\frac{dz_{2}}{z_{2}}\otimes e_L \end{aligned}$$
(5.4.11)

together with a similar equality on \(V\cap Y_2\). We can write

$$\begin{aligned} \partial _{\varphi _L}\left( \theta f_u\log |z_2|^2\otimes e_L\right)&= \theta f_u\frac{dz_{2}}{z_{2}}\otimes e_L \nonumber \\&\quad + \theta \log |z_2|^2\partial _{\varphi _L}(f_u\otimes e_L)+ f_u \log |z_2|^2\partial \theta \otimes e_L \end{aligned}$$
(5.4.12)

and then we observe that the left-hand side term of (5.4.12) is \(\partial _{\varphi _L}\)—exact on \(Y_1\). Therefore we have

$$\begin{aligned} \int _{Y_1}\partial _{\varphi _L}\left( \theta f_u\log |z_2|^2\otimes e_L\right) \wedge \overline{\alpha _1}e^{-\varphi _L}= 0 \end{aligned}$$
(5.4.13)

since \(\alpha _1\) is holomorphic. We infer that we have

$$\begin{aligned} -\int _{Y_1}\theta \frac{u}{\textrm{d}s_Y}\wedge \overline{\alpha _1}e^{-\varphi _L}&= \int _{Y_1}\theta \log |z_2|^2\partial _{\varphi _L}(f_u\otimes e_L)\wedge \overline{\alpha _1}e^{-\varphi _L}\\&\quad +\int _{Y_1}f_u \log |z_2|^2\partial \theta \otimes e_L\wedge \overline{\alpha _1}e^{-\varphi _L} \end{aligned}$$

and all that we still have to do is to apply the Cauchy–Schwarz inequality to each of the two terms of the RHS of the inequality above.

A last remark is that we have

$$\begin{aligned} \int _{Y_1}|\alpha _1|^2e^{-\varphi _L}\leqslant \int _{Y_1}|\gamma _{\xi _1}|^2e^{-\varphi _L} \end{aligned}$$
(5.4.14)

by the definition of \(\alpha _1\) and \(\beta _1\). We use the a-priori inequality and we conclude as in Theorem 1.1. \(\square \)

Remark 5.5

In the absence of hypothesis \(u|_{Y_{\textrm{sing}}}\not \equiv 0\) the evaluation of the term (5.4.7) near the singularities of Y is problematic. In the decomposition (5.4.8), we write \(\beta _j= {{\bar{\partial }}}^\star (\tau _j)\), and then the question is to estimate the quotient

$$\begin{aligned} f_j:= \frac{\tau _j}{\omega } \end{aligned}$$

at the points of \(Y_{\textrm{sing}}\). This does not seem to be possible, since we only have the norm \(W^{1,2}\) of \(f_j\) at our disposal. Indeed, the quantity \({{\bar{\partial }}}\beta _j\) is equal to the restriction of the form \({{\bar{\partial }}}\gamma _{\xi _1}\) to \(Y_j\). In Question 5.7 we provide a few more precisions about this matter.

5.4 Proof of Theorem 1.8

This is another set-up in which the considerations above work, as follows. We recall that the metric of \(h_L\) of L satisfies the hypotheses (a) and (b) at the beginning and moreover \((L, h_L)\) is flat near the singularities of Y, i.e.,

$$\begin{aligned} \Theta _{h_L}(L)\big |_{V_{\textrm{sing}}}= 0. \end{aligned}$$
(5.5.1)

Then we get an estimable extension as follows. Let \(\omega \) be a fixed Kähler metric on X. As in the proof of the preceding result Theorem 1.7, we will use the method of Berndtsson [2], so the quantity to be bounded is

$$\begin{aligned} \int _{Y\cap V_{\textrm{sing}}}\frac{u}{\textrm{d}s_Y}\wedge \overline{\gamma _{\xi _1}}e^{-\varphi _L}. \end{aligned}$$
(5.5.2)

Integration by parts shows that it is enough to obtain a mean value inequality for the function

$$\begin{aligned} \sup _{1/2V_{\textrm{sing}}}|{{\bar{\partial }}}\gamma _{\xi _1}|^2= \sup _{1/2V_\textrm{sing}}|\partial ^\star _{\varphi _L}\xi _1|^2. \end{aligned}$$
(5.5.3)

This is done according to the same principle as before. In the first place the differential inequality satisfied by \(\displaystyle |\partial ^\star _{\varphi _L}\xi _1|^2\) is as follows:

$$\begin{aligned} \Delta ''\left( |\partial ^\star _{\varphi _L}\xi _1|^2\right) \geqslant |\nabla \left( \partial ^\star _{\varphi _L}\xi _1\right) |^2- C|\partial ^\star _{\varphi _L}\xi _1|^2 \end{aligned}$$
(5.5.4)

for some constant \(C> 0\) which only depends on (the curvature of) \(\displaystyle \omega |_{V_{\textrm{sing}}}\). We will not detail the calculation here because this is very similar with the one in the proof of Theorem 1.1. However, we highlight next the main differences:

  1. (1)

    It is not necessary to introduce any regularization of the metric, since by hypothesis (5.5.1) the restriction \(h_L|_{V_{\textrm{sing}}}\) is non-singular.

  2. (2)

    Without any additional information about \((L, h_L)\), the term \(\left\langle \partial ^\star _{\varphi _L}\xi _1, \square \partial ^\star _{\varphi _L}\xi _1\right\rangle \) is problematic. Actually (5.5.1) is needed precisely in order to deal with it: the curvature of \((L, h_L)|_{V_{\textrm{sing}}}\) equals zero, then we have \(\square \partial ^\star _{\varphi _L}\xi _1= 0\) pointwise on \(V_{\textrm{sing}}\). In general we have the term

    $$\begin{aligned} \left\langle [{{\bar{\partial }}}, \Lambda _{\Theta _{h_L}(L)}]\xi _1, \partial ^\star _{\varphi _L}\xi _1\right\rangle \end{aligned}$$

    which appears in the computation and seems impossible to manage.

  3. (3)

    In the evaluation of the Laplacian of the norm of a harmonic tensor we have two terms: the gradient of the tensor, and several curvature terms corresponding to the metric on the ambient manifold and to the twisting, respectively. Here we do not have any contribution from L, and the term involving the curvature of \(\omega \) is taken care by the constant \(-C\) in (5.5.4).

Anyway, the inequality (5.5.4) can be re-written as

$$\begin{aligned} \Delta ''\left( |\partial ^\star _{\varphi _L}\xi _1|^2\right) \geqslant \big |\nabla \left| \partial ^\star _{\varphi _L}\xi _1\right| \big |^2- C|\partial ^\star _{\varphi _L}\xi _1|^2 \end{aligned}$$
(5.5.5)

and this combined with Moser iteration procedure shows that we have

$$\begin{aligned} \sup _{1/2V_{\textrm{sing}}}|\partial ^\star _{\varphi _L}\xi _1|_{\omega , h_L}^{\alpha }\leqslant C\int _{V_\textrm{sing}}|\partial ^\star _{\varphi _L}\xi _1|^{\alpha }_{\omega , h_L}\textrm{d}V_\omega . \end{aligned}$$
(5.5.6)

Finally, the term that one (almost) never uses in Bochner formula shows that we have

$$\begin{aligned} \int _X|\partial ^\star _{\varphi _L}\xi _1|^2_{\omega , h_L}\textrm{d}V_\omega \leqslant \int _X|{{\bar{\partial }}}^\star \xi _1|^2_{\omega , h_L}\textrm{d}V_{\omega } \end{aligned}$$
(5.5.7)

and we thus obtain the inequality

$$\begin{aligned} \sup _{1/2V_{\textrm{sing}}}|\partial ^\star _{\varphi _L}\xi _1|_{\omega , h_L}^{\alpha }\leqslant C\int _{V_\textrm{sing}}|{{\bar{\partial }}}^\star _{\varphi _L}\xi _1|^{\alpha }_{\omega , h_L}\textrm{d}V_\omega . \end{aligned}$$
(5.5.8)

Then we conclude as in Theorem 1.7. \(\square \)

Remark 5.6

Actually in the proof of Theorem 1.8 only needs to evaluate the \(L^2\) norm

$$\begin{aligned} \int _{Y}|{{\bar{\partial }}}\gamma _{\xi _1}|^2\textrm{d}V_{\omega } \end{aligned}$$
(5.6.1)

of \({{\bar{\partial }}}\gamma _{\xi _1}|_{Y}\). One might try to use a similar method as the one in Sect. 2, but there are serious difficulties to overcome.

5.5 Proof of Theorem 1.10

By hypothesis we know that Y has one component \(Y_1\) which only intersects \(\cup _{i\ne 1} Y_i\) in a unique point \(p_0\) such that \(u(p_0)\ne 0\). We also assume that \(\displaystyle L|_{Y_1}\) is flat, in the sense that there exists a section \(\tau \) such that \(\tau (p_0)\ne 0\) and \(\partial _{\varphi _L} \tau =0\). Then we argue as follows:

Let \(\omega \) be a fixed, reference metric on X. On each component \(Y_j\) of Y we solve the equation

$$\begin{aligned} \gamma _{\xi _1}|_{Y_j}= \alpha _j+ {{\bar{\partial }}}^\star \beta _j, \end{aligned}$$
(5.6.2)

where \(\alpha _j\) is holomorphic (1, 0) form and \(\beta _j\) is of type (1, 1) on \(Y_j\). We note that by elliptic regularity the form \(\beta _j\) is smooth.

We have \(\beta _j= f_j\omega |_{Y_j}\) and then the equality

$$\begin{aligned} \int _{Y_j}\big \langle \frac{u}{\sigma _j\textrm{d}s_{Y_j}}, {{\bar{\partial }}}^\star \beta _j\big \rangle _{\omega } e^{-\varphi _L}\textrm{d}V_{\omega }= \sum _{x\in Y_{\textrm{sing}}\cap Y_j} f_u(x)\overline{f_j(x)}e^{-\varphi _L(x)} \end{aligned}$$
(5.6.3)

follows by the residues formula. Here we denote by \(\displaystyle \sigma _j:= \prod _{i\ne j} s_{Y_i}\).

In case \(j=1\), the sum above only has one term, by hypothesis. Since we have \({{\bar{\partial }}}^\star (\tau \omega )= 0\), we can modify the solution \(f_1\) so that the global sum of residues is zero.

Question 5.7

Let p be one of the intersection points of two curves \(Y_1\cap Y_2\) in X. The analog of the a-priori inequality in Sect. 2 gives

$$\begin{aligned} |f_j(p)|^2e^{-\varphi _L(p)}\leqslant C \int _X\frac{\log ^2\Vert s\Vert ^2}{\Vert s\Vert ^{2}}|\partial _{\varphi _L} f_j|^2e^{-\varphi _L}\textrm{d}V_\omega \end{aligned}$$
(5.7.1)

provided that the bundle \((L, h_L)\) has the right curvature hypothesis, let us assume this holds for the moment. In (5.7.1) we denote by C a constant which we can compute explicitly. This a-priori inequality is obtained by considering the (2, 2)-form with values in L

$$\begin{aligned} f_j\omega ^2 \end{aligned}$$
(5.7.2)

whose \(\star \) coincides with the section \(f_j\), and use the procedure Theorem 2.1 for the function \(w:= \frac{1}{|s_1|^2+ |s_2|^2}\). The curvature requirements this induces will most likely be

$$\begin{aligned} \frac{1}{\delta }\Theta _{h_L}(L)\wedge \omega \geqslant \frac{|s_1|^2\Theta (Y_1)+ |s_2|^2\Theta (Y_2)}{|s_1|^2+ |s_2|^2}\wedge \omega . \end{aligned}$$
(5.7.3)

We have

$$\begin{aligned} {{\bar{\partial }}}f_j\omega ^2= 0, \qquad {{\bar{\partial }}}^\star f_j\omega ^2= -\star \left( \partial _{\varphi _L}f_j \right) \end{aligned}$$
(5.7.4)

which explains (5.7.1).

Anyway, by equality (5.6.2) we control the norm

$$\begin{aligned} \int _Y|\partial _{\varphi _L} f_j|^2e^{-\varphi _L}\textrm{d}V_\omega . \end{aligned}$$
(5.7.5)

Then the question is: can we find a smooth section \(\widetilde{f}_j\) of L such that it equals \(f_j\) on Y and such that

$$\begin{aligned} \int _X\frac{\log ^2\Vert s\Vert ^2}{\Vert s\Vert ^{2}}|\partial _{\varphi _L} \widetilde{f}_j|^2e^{-\varphi _L}\textrm{d}V_\omega \leqslant C\int _Y|\partial _{\varphi _L} f_j|^2e^{-\varphi _L}\textrm{d}V_\omega , \end{aligned}$$
(5.7.6)

where C in (5.7.6) is universal?