Abstract
In this paper, we establish an injectivity theorem on a weakly pseudoconvex Kähler manifold X with negative sectional curvature. For this purpose, we develop the harmonic theory in this circumstance. The negative sectional curvature condition is usually satisfied by the manifolds with hyperbolicity, such as symmetric spaces, bounded symmetric domains in , hyperconvex bounded domains, and so on.
Keywords:
non-compact Kähler manifold; Hodge decomposition; harmonic differential form; Hilbert space MSC:
Primary 32J25; Secondary 32Q15
1. Introduction
The injectivity theorem was first developed in [1,2] on a (compact) projective manifold X for an ample line bundle L. Then, it is generalized by a series of articles, such as [3,4,5,6,7,8,9], eventually to a compact Kähler manifold X with pseudo-effective line bundle L. After that, it is natural to seek the similar result on a non-compact manifold. To my best acknowledgement, there are only a few results, such as [10,11], in this aspect.
In this paper, we are interested in the manifolds with convexity. More precisely, let be a weakly pseudoconvex Kähler manifold. By this, we mean a Kähler manifold X such that there exists a smooth plurisubharmonic exhaustion function on X ( is said to be an exhaustion if for every the upperlevel set is relatively compact, i.e., tends to when z is taken outside larger and larger compact subsets of X). We prove that
Theorem 1.
Let be a weakly pseudoconvex Kähler manifold such that the sectional curvature
for some positive constant K. Let and be two (singular) Hermitian line bundles on X. Assume the following conditions:
- There exists a closed subvariety Z on X such that and are smooth on ;
- and on X;
- for some positive number δ.
For a (non-zero) section s of H with , the multiplication map induced by the tensor product with s
is (well-defined and) injective for any .
Remark 1.
The assumption (1) can be immediately removed if Demailly’s approximation technique [12] is valid in this situation. However, it seems to me that the compactness of the base manifold is of crucial importance in his original proof. Thus, it is hard to directly apply his argument here. We are interested to know whether such an approximation exists on a non-compact manifold.
We will recall the definition of singular metric and multiplier ideal sheaf in Section 2, and the elementary properties of manifolds with negative sectional curvature in Section 3.
Theorem 1 implies the following -extension theorem concerning the subvariety that is not necessary to be reduced. Such type of extension problem was studied in [10] before.
Corollary 1.
Let be a weakly pseudoconvex Kähler manifold such that
for some positive constant K. Let be a (singular) Hermitian line bundle on X, and let φ be a quasi-plurisubharmonic function on X. Assume the following conditions:
- There exists a closed subvariety Z on X such that is smooth on ;
- ;
- for all non-negative number with .
Then, the natural morphism
is surjective.
Remark 2.
If is smooth, we have and
where Y is the subvariety defined by the ideal sheaf . In particular, Y is not necessary to be reduced. Then, the surjectivity statement can interpret an extension theorem for holomorphic sections, with respect to the restriction morphism
In order to prove Theorem 1, we improve the -Hodge theory introduced in [13], such that it is suitable for the forms taking value in a line bundle. The crucial thing is the Hodge decomposition [14,15] on a non-compact manifold. Since the base manifold has negative sectional curvature, it is Kähler hyperbolic by [13]. We then apply the Kähler hyperbolicity to establish the Hodge decomposition. We leave all the details in the text.
Remark 3.
The Kähler hyperbolic manifold was deeply studied in [13]. In particular, Ref. [13] provides many examples for Kähler hyperbolic manifolds, such as symmetric spaces, bounded symmetric domains in , hyperconvex bounded domains, and so on. Certainly, Theorem 1 is valid on these manifolds.
Remark 4.
All the results are still valid if L is twisted by a Nakano semi-positive [16] vector bundle E. The proof involves nothing new hence we omit it.
The plan of this paper is as follows: we will first recall the background materials in Section 2. The Kähler hyperbolicity is discussed in Section 3. Then, we discuss the Hodge decomposition on a non-compact manifold in Section 4. In Section 5, we prove the injectivity theorem and the extension theorem.
2. Preliminarily
2.1. Singular Metric
Recall that a smooth Hermitian metric h on a line bundle L is given in any trivialization by
where is an arbitrary function, called the weight of the metric with respect to the trivialization . Then, the singular Hermitian metric is defined in [16] as follows:
Definition 1
(Singular metric). A singular Hermitian metric h on a line bundle L is given in any trivialization by
where is an arbitrary function, called the weight of the metric with respect to the trivialization θ.
Sometimes, we will directly say that is a (singular) metric on L if nothing is confused.
2.2. Multiplier Ideal Sheaf
The multiplier ideal sheaf is an important tool in modern complex geometry, which was originally introduced in [16,17].
Definition 2
(Multiplier ideal sheaf). Let L be a line bundle. Let φ be a singular metric on L such that for a smooth real -form γ on X. Then, the multiplier ideal sheaf is defined as
Note that X is non-compact, and in general will not imply that
However, when X is furthermore assumed to be weakly pseudoconvex, we could substitute for . Here, is a convex increasing function of arbitrary fast growth at infinity and is the smooth plurisubharmonic exhaustion function provided by the weak pseudoconvexity of X. This factor can be used to ensure the convergence of integrals at infinity. Moreover, we have
and . Therefore, we can always assume without loss of generality that, for every ,
3. The Kähler Manifold with Negative Curvature
3.1. Negative Curvature
Firstly, let us recall the definition for a manifold with negative sectional curvature.
Definition 3.
Let be a Kähler manifold. Let be the curvature associated with ω. Then, X is said to have negative sectional curvature, if there exists a positive constant K such that, for any non-zero complex vector ,
It is denoted by .
A complete Kähler manifold with negative sectional curvature will be Kähler hyperbolic (see Proposition 1). The Kähler hyperbolicity was first introduced in [13] for a compact Kähler manifold. However, there is no obstacle to extend it to the non-compact case. Firstly, let us recall the d-boundedness of a differential form.
Definition 4.
Let α be a differential form on X. Let be the universal covering of X. Then,
- (i)
- α is called bounded (with respect to ω) if the -norm of α is finite,Here, is the pointwise norm induced by ω.
- (ii)
- α is called d-bounded if there exists a differential form β on X such that and .
- (iii)
- α is called -bounded if is d-bounded on .
Remark 5.
When X is compact, these notions bring nothing new. When X is non-compact, it is easy to verify that d-boundedness implies -boundedness, whereas there is no direct relationship between boundedness and d-bounxdedness.
The Kähler hyperbolic manifold is then defined as
Definition 5.
A Kähler manifold is called Kähler hyperbolic if ω is -bounded.
We list some functionality property of the Kähler hyperbolicity here. They are almost obvious, and one could refer to [13] for more details.
Proposition 1.
- (i)
- Let X be a Kähler hyperbolic manifold. Then, every complex submanifold of X is still Kähler hyperbolic. In fact, if Y is a complex manifold which admits a finite morphism , then Y is Kähler hyperbolic.
- (ii)
- Cartesian product of Kähler hyperbolic manifolds is Kähler hyperbolic.
- (iii)
- A complete Kähler manifold with negative sectional curvature must be Kähler hyperbolic. This fact was pointed out in [13], whose proof can be found in [18]. More precisely, if , there exists a 1-form η on such that and
3.2. Notations and Conventions
We make a brief introduction for the basic notations and conventions in Kähler geometry to finish this section. We recommend readers to see [15] for a sophisticated comprehension.
Let be a Kähler manifold of dimension n, and let be a holomorphic line bundle on X endowed with a smooth metric . The standard operators, such as , * as well as L, , etc., in Kähler geometry are defined locally and thus make sense with or without the compactness or completeness assumptions. For an m-form , we define . Let be the Chern connection on L associated with . Moreover, for an L-valued k-form , we define the operators , , and .
Let be the space of all the smooth L-valued -forms on X. The pointwise inner product on is defined by the equation:
for . The pointwise norm is then induced by . The -inner product is defined by
for , and the norm is induced by .
Let be the space of all the L-valued (not necessary to be smooth) -forms with bounded -norm on X, and it equipped with becomes a Hilbert space. The operators , , and are then the adjoint operators of D, , and with respect to if X is compact. However, when X is non-compact, the situation would be much more complicated. We will deal with it in the next section.
4. The Hodge Decomposition
The Hodge decomposition is the ingredient to study the geometry of a compact Kähler manifold. One can consult [14,15] for a complete survey. In this section, we will discuss the Hodge decomposition on a non-compact manifold. Let be a complete Kähler manifold of dimension n with negative sectional curvature, and let be a holomorphic line bundle on X endowed with a smooth metric .
4.1. Elementary Materials
We collect from [13] some basic properties concerning the Hodge decomposition here. Remember that the adjoint relationship between and in general fails when X is non-compact. In fact, the compactness becomes important when one takes an integral. However, since X is complete here, we still have the Stokes formula as follows:
Lemma 1.
Let η be an L-valued -bounded -form on X, i.e.,
such that is also -bounded. Then,
Essentially, this lemma is not a surprise after applying the cut-off function to reduce it to the case that has the compact support, while the existence of such a cut-off function is guaranteed by the completeness of . For example, we could use the geodesic distance to construct a function on X for every satisfying the following conditions:
- is smooth and takes values in the interval with compact support;
- The subset exhausts X as tends zero, and
- .
Now the proof of Lemma 1 is elementary, and we omit it here. With the help of Lemma 1, most of the canonical identities on compact Kähler manifold extend into this situation. Remember that the Laplacian operators are defined as
Proposition 2.
Let α be an L-valued -bounded form on X. Then,
- Integral identities.
- Bochner–Kodaira–Nakano identity.
In particular, 1. and 2. together give that
Proof.
We only prove that
Recall that, for any differential forms α, β with proper degree, we always have
where the sign on the right-hand side is determined by the degree of α. Therefore,
We apply Lemma 1 to obtain the third equality. Obviously,
Then, we choose such that on X and estimate by Schwarz inequality. This yields
Hence, as ε tends to zero. As a result, we obtain the desired equality.
The other identities are similar. □
There are various quick consequences of this proposition. For example, is □-harmonic, i.e., , if and only if and . The similar conclusion holds for the operators and .
Moreover, with Lemma 1 and Proposition 2, one concludes that the -space of the L-valued k-forms on X admits Hodge decomposition as follows:
Definition 6
(Hodge decomposition, I). For the -space , we have the following orthogonal decomposition:
where
and
Similarly, for the -space of the L-valued -forms, we have
Definition 7
(Hodge decomposition, II).
where
and
4.2. Lower Bound on the Spectrum
In this section, we will show that and in the decomposition (1), , and in the decomposition (2) are actually closed, in which the negative sectional curvature really comes into effect. Remembering that is Kähler hyperbolic by Proposition 1, we have , where is the universal covering and is a bounded form on .
Let , and . The -spaces
and the related subsets such as , are defined in an obvious way.
Proposition 3.
Every with satisfies that
where λ is a strictly positive constant which depends only on k, and . Furthermore, when inequality (3) is satisfied by which is orthogonal to .
Proof.
When , inequality (3) was proved in [19]. According to Proposition 2, it shows that the D-closed -valued k-form α () satisfies
and the -closed -valued k-form α () satisfies
Therefore, as well as is itself closed.
In particular, applying the conclusion above to the -valued -closed -form and -valued D-closed -form respectively, we obtain that
and
are both closed. Therefore, Hodge decomposition is improved as
for .
Now we are able to prove the proposition for the form α of degree m orthogonal to . We have , where is orthogonal to , and the -forms β and γ of degrees and correspondingly satisfy , . This implies
as well as and . On the other hand, apply inequality (3) to β and γ yields, as a consequence of Schwarz inequality that
and
Thus,
□
If the curvature of L is bounded from below, we will have a similar estimate for .
Corollary 2.
Proof.
Since π is locally isometric, . By Proposition 2, we have
Therefore, when ,
The last inequality is due to Proposition 3 and elementary computation. Then, we apply the same argument as Proposition 3 to obtain the desired conclusion for . The proof is complete. □
In particular, if and , the -closed -valued -form satisfies
and the -closed -valued -form satisfies
Here, is a positive constant. Therefore, as well as is itself closed.
Therefore, we see that, when and or , Hodge decomposition (2) can be improved on as:
Thus, we have Hodge’s theorem on X as follows.
Proposition 4.
Assuming that for some ε small enough, then
Proof.
Firstly, we claim that
is closed provided that
is closed. In order to prove this claim, let us review the relationship between the -spaces and shown in [20]. Remember an open subset is called a fundamental domain of the action of the fundamental group on if the following conditions are satisfied:
- ;
- for and
- has zero measure.
We construct a fundamental domain in the following way. Let be a locally finite cover of X with open balls having the property that, for each k, there exists an open set such that is biholomorphic with inverse . Define . Then, is a fundamental domain.
Then, it is easy to see that
A basis of is formed by the functions
Then, for , the above identification is given by
Now let be a sequence in that is convergent in . Fix a . Then, is a sequence in that is convergent in . Hence, the limit is for some due to the closeness of . We define an -bounded form α on X by
for any test form μ with proper degree on X. Consequently, we have
It exactly means that , hence is closed. The claim is proved. Remember the fact that is closed if and only if is closed; we have a similar conclusion between and .
Due to this claim and the discussions after Corollary 2, we only need to prove that
and
are both closed with . However, it is respectively equivalent to the closeness of
and
which has been verified. The proof is complete. □
5. The Injectivity Theorem
Let be a weakly pseudoconvex Kähler manifold with negative sectional curvature. Let be a (singular) Hermitian line bundle on X such that . Moreover, is smooth on with Z a closed subvariety of X. Firstly, we specify Hodge’s theorem (Proposition 4) in this situation.
Since X is weakly pseudoconvex, so will be Y. Then, there exists a smooth plurisubharmonic exhaustion function on Y. Set
which is a complete Kähler metric on Y for every . Obviously,
on Y when . Moreover, when we take large enough l, the sectional curvature of approximates the sectional curvature of and hence is negative. Therefore, is a complete Kähler hyperbolic manifold by Proposition 1.
Now, we apply Proposition 4 to , and obtain that
where
and
We use to denote the adjoint operator of on Y defined through and . Let
Under this circumstance, Hodge’s theorem is formulated as
Proposition 5.
Proof.
Let . Applying Proposition 2 to each , we have
Since , . Therefore,
In particular, . It means that is a holomorphic L-valued -form on Y. On the other hand,
The first inequality is due to the fact that and . By canonical -extension theorem [21], extends to a holomorphic L-valued -form on X, which is denoted by . Fix . Then, for ,
hence is uniformly bounded in -norm . Consequently, it converges to a holomorphic L-valued -form, say γ. Furthermore, as tends to ∞, we obtain that . Now, it is easy to verify that
We denote this morphism by .
Conversely, let . Let be the sheaf of germs of -forms β on X with values in L and with measurable coefficients, such that both and are locally integrable. The operator defines a complex of sheaves , and it is easy to verify that is a resolution of . Each sheaf is a -module, so is a resolution by acyclic sheaves.
Then, we can find a representative of
through this resolution by acyclic sheaves. In other words, α is a -closed L-valued -form on X such that and are locally integrable. Moreover, through the discussions in Section 2.2, we could arrange the things so that
In particular, . Now let be the harmonic representative of in . Equivalently, . Applying the same argument of the first part, we will eventually obtain a sequence of holomorphic L-valued -forms and its limit γ on X. On the other hand,
the sequence is convergent to, say . Since ,
Therefore, by definition. We denote this morphism by .
It is easy to verify that and . The proof is finished. □
Now, we are ready to prove the injectivity theorem on a non-compact manifold. One could consult [3,5,7,8] for a sophisticated comprehension for the injectivity theorem on a compact manifold.
Theorem 2 (=Theorem 1).
Let be a weakly pseudoconvex Kähler manifold such that
for some positive constant K. Let and be two (singular) Hermitian line bundles on X. Assume the following conditions:
- There exists a closed subvariety Z on X such that and are both smooth on ;
- and on X;
- for some positive number δ.
For a (non-zero) section s of H with , the multiplication map induced by the tensor product with s
is (well-defined and) injective for any .
Proof.
By Proposition 5, it is enough to prove that
is well-defined, hence injective. In other words, let , and we should prove that .
In fact, since , there exists and
with . Applying Proposition 2, we obtain that
Notice that , . Hence,
In particular, .
Now, apply Proposition 2 again on and observe that , we obtain that
Since , and
it is easy to see that
In particular, , hence . Obviously, and . Therefore,
by definition. The injectivity is now obvious. □
Proof of Corollary 1.
Consider the short exact sequence
The associated cohomology long exact sequence implies that the surjectivity of
is equivalent to the injectivity of
Applying Proposition 5, it reduces to prove that
is well-defined.
In fact, let and such that
Obviously, . Here, , and are understood in an obvious way. Moreover, applying Proposition 2 with , we obtain that
Applying Proposition 2 one more time with , then
Since , we also have . On the other hand,
The last inequality comes from the assumption that for ,
In summary, . Therefore, . by definition. The proof is complete. □
6. Conclusions
We establish an injectivity theorem on a weakly pseudoconvex Kähler manifold X with negative sectional curvature. In particular, X is not necessary to be compact. As an application, we obtain an -extension theorem concerning the subvariety that is not necessary to be reduced. Such type of extension theorem is of crucial importance in complex geometry.
Funding
This research was funded by China Postdoctoral Science Foundation Grant No. 2019M661328.
Acknowledgments
The author thanks the referees for detailed and constructive criticism of the original manuscript.
Conflicts of Interest
The author declares no conflict of interest.
References
- Kollár, J. Higher direct images of dualizing sheaves I. Ann. Math. 1986, 123, 11–42. [Google Scholar] [CrossRef]
- Kollár, J. Higher direct images of dualizing sheaves II. Ann. Math. 1986, 124, 171–202. [Google Scholar] [CrossRef]
- Ambro, F. An injectivity theorem. Compos. Math. 2014, 150, 999–1023. [Google Scholar] [CrossRef] [Green Version]
- Enoki, I. Kawamata-Viehweg vanishing theorem for compact Kähler manifolds. In Einstein Metrics and Yang-Mills Connections; CRC Press: Boca Raton, FL, USA, 1993; pp. 59–68. [Google Scholar]
- Fujino, O. A transcendental approach to Kollár’s injectivity theorem. Osaka J. Math. 2012, 49, 833–852. [Google Scholar] [CrossRef] [Green Version]
- Fujino, O.; Matsumura, S. Injectivity theorem for pseudo-effective line bundles and its applications. arXiv 2016, arXiv:1605.02284. [Google Scholar] [CrossRef]
- Gongyo, Y.; Matsumura, S. Versions of injectivity and extension theorems. In Annales Scientifiques de l’Ecole Normale Superieure; Societe Mathematique de France: Paris, France, 2017; Volume 50, pp. 479–502. [Google Scholar]
- Matsumura, S. A Nadel vanishing theorem via injective theorems. Math. Ann. 2014, 359, 785–802. [Google Scholar] [CrossRef]
- Matsumura, S. Injectivity theorems with multiplier ideal sheaves and their applications. Complex Anal. Geom. 2015, 144, 241–255. [Google Scholar]
- Cao, J.; Demailly, J.-P.; Matsumura, S. A general extension theorem for cohomology classes on non reduced analytic subspaces. Sci. China Math. 2017, 60, 949–962. [Google Scholar] [CrossRef] [Green Version]
- Zhou, X.; Zhu, R. Extension of cohomology classed and holomorphic sections defined on subvarieties. arXiv 2019, arXiv:1909.08822. [Google Scholar]
- Demailly, J.-P.; Peternell, T.; Schneider, M. Pseudo-effective line bundles on compact Kähler manifolds. Int. J. Math. 2001, 6, 689–741. [Google Scholar] [CrossRef] [Green Version]
- Gromov, M. Kähler hyperbolicity and L2-Hodge theory. J. Differ. Geom. 1991, 33, 263–292. [Google Scholar] [CrossRef]
- Griffiths, P.; Harris, J. Principles of Algebraic Geometry; Wiley: New York, NY, USA, 1978. [Google Scholar]
- Morrow, J.; Kodaira, K. Complex Manifolds; Holt, Rinehart and Winston: New York, USA, 1971. [Google Scholar]
- Demailly, J.-P. Analytic Methods in Algebraic Geometry; International Press: Somerville, MA, USA, 2012. [Google Scholar]
- Nadel, A.M. Multiplier ideal sheaves and Kähler-Einstein metrics of positive scalar curvature. Ann. Math. 1990, 132, 549–596. [Google Scholar] [CrossRef]
- Chen, B.; Yang, X. Compact Kähler manifolds homotopic to negatively curved Riemannian manifolds. Math. Ann. 2018, 370, 1477–1489. [Google Scholar] [CrossRef]
- Huang, T. L2 vanishing theorem on some Kähler manifolds. Isr. J. Math. 2021, 241, 147–186. [Google Scholar] [CrossRef]
- Ma, X.; Marinescu, G. Holomorphic Morse Inequalities and Bergman Kernels; Progress in Mathematics; Springer: Berlin/Heidelberg, Germany, 2007; Volume 254. [Google Scholar]
- Ohsawa, T. Analysis of Several Complex Variables; American Mathematical Soc.: Providence, RI, USA, 2002. [Google Scholar]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations. |
© 2021 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).