Reider-type theorems on normal surfaces via Bridgeland stability
Abstract.
Using Langer’s construction of Bridgeland stability conditions on normal surfaces, we prove Reider-type theorems generalizing the work done by Arcara-Bertram in the smooth case. Our results still hold in positive characteristic or when is not necessarily a line bundle. They also hold when the dualizing sheaf is replaced by a variant arising from the theory of Du Bois complexes. For complex surfaces with at most rational double point singularities, we recover the optimal bounds for global generation and very ampleness as predicted by Fujita’s conjecture.
1. Introduction
In [5] Fujita proposed a fundamental conjecture in algebraic geometry stating that if is an ample line bundle on a smooth projective complex variety of dimension , then is globally generated and is very ample. The conjecture has seen much interest over the years and the global generation part is now known in dimension ([17], [3], [9], [23]).
One strategy that has seen much success (including in giving effective bounds for higher dimensional varieties) involves vanishing theorems. Another approach, which is, however, hard to generalize in higher dimensions, is Reider’s strategy of using vector bundle techniques and Bogomolov’s inequality. We will focus on the latter approach in the context of normal surfaces. Our main source of inspiration is Arcara and Bertram’s reinterpretation of Reider’s theorem in the context of Bridgeland stability conditions [1].
Our work is concerned with giving Reider-type (and generation of higher jets of ) bounds on normal surfaces. For this generalization to singular surfaces, another key ingredient used in our work is Langer’s recent construction of stability conditions on normal surfaces [13].
Remark 1.1.
We now discuss the main results of this article. Let be a projective normal surface defined over an algebraically closed field (of any characteristic), a dimension 0 subscheme of length , and let be an ample line bundle on . The constant will be defined in Section 2 and in characteristic 0, depends only on the singularities of ; roughly speaking, ensures a Bogomolov inequality holds on For concrete examples where this constant is computed see Section 2.2.
Theorem 1.2.
Let be a projective normal surface defined over an algebraically closed field, a zero-dimensional subscheme of length , and an ample line bundle on . Suppose there is a pair of nonnegative integers such that and for any curve . Then we have the vanishing
In particular, this implies separates jets along i.e.
is surjective.
Remark 1.3.
Definition 1.4.
Define a function by
Also define
Note, in particular, that and which will recover the global generation and very ampleness bounds in Fujita’s conjecture for complex surfaces with at worst rational double point singularities.
Corollary 1.5.
Let be a projective normal surface defined over an algebraically closed field, a subscheme of length , and an ample line bundle on . Then
for any integer In particular, this implies separates jets along i.e.
is surjective.
Remark 1.6.
Unlike previous work, which gives similar results for a divisor under the assumption that is Cartier, our theorem shows separation of jets for the tensor product of an ample bundle with the canonical sheaf, even when this is not a line bundle.
We also recover the classical Fujita conjecture for surfaces:
Corollary 1.7.
If is a projective normal surface with (e.g. a complex surface with at most rational double point singularities) and an ample line bundle, then is globally generated for and very ample for .
Remark 1.8.
The proof of the optimal bound for very ampleness in the smooth case is contained in [1], but our argument appears to be the first proof of the correct bound for global generation via Bridgeland stability.
In positive characteristic, we can no longer take for every smooth surface , but there is still an explicit bound:
Corollary 1.9.
Let be a smooth projective surface defined over an algebraically closed field of positive characteristic and a dimension 0 subscheme of length Following Koseki [10], depends only on the birational equivalence class of and is defined as:
-
(1)
If is a minimal surface of general type, then
-
(2)
If and is quasi-elliptic, then
-
(3)
Otherwise,
Then
for any integer In particular, this implies separates jets along i.e.
is surjective.
Remark 1.10.
It is known that Fujita’s conjecture does not hold for smooth surfaces in positive characteristic. In fact, for any power one can find a smooth surface such that is not globally generated [7]. Our results account for this, as the bound on depends on the constant corresponding to the surface
Remark 1.11.
Note that for we get a Kodaira vanishing type result in positive characteristic, again depending on
Now specializing to the case when is complex, we also obtain similar results when the dualizing sheaf is replaced by the dual of where is the -th Du Bois complex of . This line of inquiry was based on the observation that has the shape of a (potentially) Bridgeland stable object (namely, slope stable sheaf in degree and dimension 0 sheaf in degree 0), and follows the philosophy of viewing the Du Bois complexes as better-behaved versions of the sheaves of Kähler differentials on (or in this case, of as a better-behaved version of ). To be more precise, consider
which is in fact a sheaf (see Section 2.3 for details). Then we obtain similar Reider-type results for (It may be helpful to keep in mind that, at least in the Gorenstein case, , where is an ideal sheaf supported exactly on the non-Du Bois locus. Thus, the problem of separating jets along some zero-dimensional subscheme is more or less the problem of separating jets along a larger subscheme containing both and the non-Du Bois locus, and so the bounds are correspondingly worse.)
Theorem 1.12.
Let be a projective complex normal surface, a subscheme of length , and an ample line bundle on . Let . Let be a pair of nonnegative integers such that . Suppose and for every curve . Then we have the vanishing
In particular, this implies separates jets along i.e.
is surjective.
Corollary 1.13.
Let be a projective normal complex surface, a subscheme of length , and an ample line bundle on . Let . Then
for any integer , where is defined in Definition 1.4. In particular, this implies separates jets along i.e.
is surjective.
We now briefly explain the strategy of the proof. First, we note the separation of jets along of is implied by the vanishing which is equivalent to the vanishing of . In the case of , the object is replaced by , and so we consider more generally the spaces of morphisms from objects of the form to a class of objects “of type ” (see Definition 3.1 for details), which includes both and . As in Arcara-Bertram [1], the goal is to find Bridgeland stability conditions with respect to which objects of type (respectively, objects of the form ) are stable, and then use Schur’s lemma to conclude. Refining the argument slightly, we see that in fact the image of a nonzero morphism of this form (appropriately defined) is a torsion sheaf whose support gives an effective divisor satisfying Reider-type conditions.
Outline of the paper. In Section 2, we fix notation and recall the necessary background on Bridgeland stability conditions on a (potentially singular) normal surface. We also briefly introduce Du Bois complexes and discuss some of their properties. In section 3 we find conditions guaranteeing that objects of the form and objects of type are Bridgeland stable, adapting the proof of Arcara and Bertram [1] to the more general setting of normal surfaces. We also explain why these results are stronger than we need for the proof of the main theorem, where it suffices to consider only destabilizing objects of a certain form. In section 4, we deduce the final form of our main technical theorem (Theorem 4.7), from which all our results follow.
Acknowledgements. We are grateful to Mihnea Popa for suggesting this problem and for insightful conversations throughout the project. We would also like to thank Davesh Maulik, Mircea Mustaţă, Sung Gi Park, and Wanchun Shen for valuable discussions.
2. Preliminaries
In this section, we review Langer’s construction of Bridgeland stability conditions on normal proper surfaces [13] (cf. [15]). We also give a brief introduction to Du Bois complexes. In particular, we focus on the zeroth Du Bois complex of a normal surface.
2.1. Bridgeland stability on normal surfaces
Let be a normal surface. Langer constructs Chern character homomorphisms
where and are defined as usual on the smooth locus and is defined so that a Riemann–Roch formula holds (see [14] for details). We will simply denote the homomorphism by in what follows. (This agrees with the standard Chern character on vector bundles, justifying the use of the notation.)
In addition to the Chern character, Langer’s definition uses the Mumford intersection product on defined as follows [6, Ex. 7.1.16]: Let be a minimal resolution of singularities. Then we define a homomorphism by
where is assumed to be an irreducible curve on , is the proper transform of , and is the unique -divisor supported on the exceptional locus of such that for any component of the exceptional locus, we have . For future use, we record the following observations:
Observations 2.1.
-
(1)
Writing (as a normal surface has isolated singularities) and the set of irreducible components, then for and , we have .
-
(2)
If is the class of a Cartier divisor , then .
-
(3)
Assuming from now on that is in addition proper, then the composition of the map with the standard intersection product yields an intersection product on satisfying the Hodge index theorem.
-
(4)
Given with , then , as with . In particular, if is the class of a Cartier divisor, then . (More precisely, is given by composition of the degree map with the first Chern class intersection map of [6, Section 2.5]).
Now, let be a normal proper surface over an algebraically closed field, and choose -divisors such that is numerically ample. Langer defines a Bridgeland stability condition as follows: First, he shows that one can choose a constant satisfying the Bogomolov-type inequality
for any torsion-free, -slope semistable coherent sheaf on . (For example, if is smooth and defined over an algebraically closed field of characteristic 0, then by the ordinary Bogomolov inequality we may choose . For more discussion, see 2.2)
Given the Bogomolov inequality on , one defines the central charge
We note that this matches the standard definition of the central charge in the smooth case, assuming one takes . The heart is defined exactly as in the smooth case, i.e., we define to be the tilt of by the torsion pair
where the -slope of a coherent sheaf is defined by
(setting when ).
Recall that the central charge encodes the Bridgeland rank and degree . For we have , and in the case of equality . We then define the Bridgeland slope
An object of is said to be Bridgeland stable if it has Bridgeland slope greater than that of any proper subobject. Given Bridgeland stable objects with we have the important property that .
In this paper, we only consider the half-plane of stability conditions for a fixed ample and varying . We will use to denote , and similarly for etc. We note that is in fact independent of the choice of . For convenience, we record the following:
2.2. More about the constant
Since the constant appears in our Reider-type bounds, we will briefly recall its definition and compute it in an example.
First, it is not hard to reduce to showing the inequality for an arbitrary reflexive sheaf Let be a minimal resolution of , and set . Then one uses the fact that a Bogomolov inequality is satisfied for , i.e.,
which is classical in the characteristic 0 (with ) and a theorem of Koseki in positive characteristic [10].
Let be the divisor supported on the exceptional locus of uniquely defined by the property that for each irreducible exceptional divisor . Langer calculates in [13] that
and so we get
In [14], Langer shows that
and so
Moreover, he calculates that for a component of the exceptional divisor we have
Putting these together, one sees that we can bound for some independent of (but depending on , and the genera and intersection numbers of the ’s).
Example 2.2.
Let be a degree 3 complex projective plane curve with line bundle Then consider the projectivization of the cone over with conormal bundle , i.e. of the affine cone
Note that by [21, Appendix A] the cone has a Du Bois but not rational singularity. We would like to determine in this case.
First, note that Moreover, by adjunction. Now, suppose Then, by the above discussion, and since by classical Bogomolov, one can choose to be the smallest positive number such that
Then is minimized at so we would like to find such that for any integer This means
Example 2.3.
One can similarly compute a bound for for the cone over a smooth projective degree complex plane curve with conormal bundle This bound grows asymptotically on the order of
2.3. Du Bois complexes
Let be a complex variety. One would like to consider an analogue of the standard de Rham complex on smooth varieties for singular varieties. Fix a hyperresolution In [2], following ideas of Deligne, Du Bois introduced which is an object in the derived category of filtered differential complexes on , and showed that this is independent of the choice of hyperresolution. One can associate a filtration by recalling that is filtered by Consider then the -th Du Bois complex of
For more details on Du Bois complexes, we refer the reader to [2], [8, Chapter V], or [16, Chapter 7.3].
In this paper, we will be concerned with understanding the stability of the 0-th Du Bois complex of A result of Saito and Schwede [18, Proposition 5.2], [20, Lemma 5.6] says that for any complex variety
where is the structure sheaf of the seminormalization of
In our case, since is normal, note that we simply have
Moreover, for each , there is a canonical morphism , which is an isomorphism if is smooth; see [16, Page 175]. In particular, the are supported on the singular locus of for all Hence, in our case, is supported in codimension 2. General vanishing results for the cohomologies of Du Bois complexes show that for
Note that any normal surface is Cohen-Macaulay (since it is ). Let be the dualizing sheaf of , and consider the duality functor on We define
Now, by the injectivity theorem of Kovács-Schwede [11, Theorem 3.3], we have that the map
induced by dualizing the canonical morphism is injective on cohomology, i.e., for all . In other words, is a subsheaf of .
We conclude this section with a result that will be necessary in order to apply our main theorem.
Proposition 2.4.
For any closed point we have .
Proof.
We note first that , as is supported on and so by the local-to-global spectral sequence it suffices to compute . Then
by vanishing of negative Exts since is a sheaf, as discussed above. ∎
We note that this proposition cannot be deduced purely from the cohomology sheaves but requires the additional input that is a sheaf; indeed, the statement would be false in the non-Du Bois case if were replaced by .
3. Bridgeland stability
In this section we prove several lemmas about Bridgeland stability of elements of whose cohomologies match those of the objects we will consider. Namely, we study the stability of a class of objects with properties matching those of and (“type ”) and of for an ample line bundle and a subscheme of dimension 0. We also analyze the image of a morphism from an object to an object of type .
3.1. Conditions for stability
In this section, we derive inequalities that would be satisfied if an object of one of the two forms of interest were not Bridgeland stable, following the proof of [1] with the necessary modifications for the non-smooth case.
3.1.1. Objects of type
The following definition captures the essential properties of and .
Definition 3.1.
We say an object is of type if for every closed point and
Example 3.2.
As explained in section 2.3, the object is of type . Because is also a sheaf, the argument of Proposition 2.4 shows that for any , so is also of type . Another example is the shifted derived dual , where is a 0-dimensional subscheme contained in the Gorenstein locus of ; this is the type of object considered (in the smooth case) in [1].
Fix some numerically ample , and let be an element of type . Our goal will be to find a choice of such that is a Bridgeland stable element of .
Proposition 3.3.
An object of type belongs to the heart if and only if .
Proof.
We note that is torsion, therefore is contained in for any choice of and . So it suffices to see that . Since is of rank 1, and thus is automatically -stable, the necessary condition is that
and since necessarily , this amounts to the condition . ∎
We now explore the implications of being Bridgeland unstable.
Proposition 3.4.
Fix . If is destabilized by some quotient , then for any -slope HN factor of with Bridgeland rank , we have that
where we write and .
Proof.
Suppose we have a destabilizing short exact sequence
in . Then since Bridgeland rank is additive in short exact sequences and nonnegative on , we have that
for as described above, as is a subobject of in and can be built up from a series of extensions of and , where the ’s run over the -slope HN factors of . Thus, using our assumption, we get
∎
Define . We start by noting that , where is a line bundle and therefore Langer’s Chern character is the standard one. On the other hand, we have and so by [14].
In the following proposition, we will restrict attention to one particular stability condition, chosen so that and (taking ).
Proposition 3.5.
Suppose , and fix . Suppose is an -slope semistable sheaf with such that . Then
defining and as before.
Proof.
The slope inequality with denominator implies that the numerator
Now, by Langer’s Bogomolov-type inequality [13, Theorem 0.1], we have
∎
We now combine the previous two propositions to deduce a contradiction if is sufficiently positive.
Lemma 3.6.
Let where is an ample line bundle such that . Given satisfying
we must have either and , or and (so in particular and ). In the case , the same holds when is any numerically ample -divisor.
Proof.
Define . We start by showing that in the case we must have . Indeed, suppose first that Then by our assumptions we have
unless . When , we also use the assumption that is (integral) Cartier: recalling from 2.1(4) that and are integers, we have
which implies and and so as above .
We now show that our assumption on implies that : recalling from 2.1(3) that the intersection product on satisfies the Hodge index theorem, we have
and
If we had , using that we would conclude that
However, this contradicts our assumption on , so we conclude that in fact , which implies by the above that
It follows that
∎
We now describe conditions under which an object is Bridgeland stable.
Proposition 3.7.
Let be of type , and suppose where is an ample line bundle such that and for some . Then is Bridgeland stable for the stability condition with . (If , the same holds with the additional bound .)
Proof.
Suppose is destabilized by a quotient , i.e., . Our strategy will be to find an -slope semistable factor of with and and then to apply Lemma 3.6.
We start by showing that our assumption that for every closed point implies strictly: Suppose , and consider the kernel object . We have the long exact sequence of cohomology
where all the cohomologies are by definition torsion free. This implies immediately that or , and the latter is impossible as in this case
where and . So we conclude , and then implies, by definition of , that is supported in dimension 0 (as this can be decomposed into a torsion sheaf and torsion-free sheaves with , and any torsion components supported in dimension 1 would pair positively with the ample divisor , again giving ). Now, restricting the inclusion to any length one subsheaf gives a nonzero element of , a contradiction. Thus we have
for any semistable factor of .
Now, recall that by definition of and the long exact sequence of cohomology, we have that is a quotient of , and has a filtration whose quotients are torsion-free -slope semistable sheaves with . Then , being supported in dimension 0, has Bridgeland rank 0 and nonnegative degree, and so . Now, is a weighted average of the , and so at least one of these, say, , must have Bridgeland slope , and by the previous paragraph we have .
In fact, to obtain the optimal bounds in our main theorem, we will use the following variant of the above proposition with a slightly weaker assumption on (which can now be an -divisor, and the inequality in the bounds is no longer strict). This small change turns out to be necessary to obtain the optimal Fujita-type bounds, at least in the smooth case.
Proposition 3.8.
Let be of type and let be an numerically ample -divisor such that and for any nonzero effective divisor . Then is not strictly destabilized by an injection from a torsion sheaf in cohomological degree 0 with respect to the Bridgeland stability condition for if (i.e., in any case except ).
Proof.
Suppose is a strictly destabilizing injection in , where is torsion. Let be the quotient object. Then by the long exact sequence of cohomology we conclude that is of rank 1, so in particular this is -slope semistable, and since is a destabilizing quotient, we must have (or else, since is supported in dimension 0, we would conclude that ).
For any define , , and . We note that for all , and for sufficiently small (where here we use that with strict inequality because and thus the same holds for sufficiently small). Thus the exact triangle is contained in and so continues to be a surjection in for sufficiently small. Furthermore, since Bridgeland slope is a continuous function of (since the denominators ), we conclude that must be a strictly destabilizing quotient of for sufficiently small.
Now, we apply the argument of the proof of Proposition 3.7: first, by definition of being of type we know that must be supported in dimension 1 (as this is torsion but not supported in dimension 0), and so in particular has positive Bridgeland rank. Also is supported in dimension 0, so we have
and also
Thus we can apply Propositions 3.4 and 3.5 to end up in the situation of Lemma 3.6 with . Then we conclude , a contradiction since is the effective divisor given by the support of . ∎
3.1.2. Objects of the form
Let be an ample line bundle and be a subscheme of dimension 0. We write and fix . Our goal is to show that is Bridgeland stable for the stability condition considered in the last section (here with replacing ).
We note first of all that is -slope semistable as this is of rank 1, and so exactly when
i.e., when . So in particular, for our choice of .
Proposition 3.9.
Let , , , and be as above, and suppose satisfies and for any nonzero effective divisor . Then is not destabilized by a subsheaf .
Proof.
Let be dimension 1 component of . We note immediately that if , then the sheaf is not destabilizing, as and .
Then since is of rank 1 and thus -slope semistable, we see that if and only if
where the second inequality follows from the fact that is effective.
Now, for to be a destabilizing subsheaf we must have
and since this means
By Langer’s Bogomolov inequality [13, Theorem 0.1] we have
and so this would imply
In other words, to prove our result, it suffices to show
It follows from the Hodge index theorem (see 2.1(3)) and the condition that above that
Now since is effective, by assumption we have
Putting the two together
completing the proof. ∎
Proposition 3.10.
Let , , , be as above, and suppose satisfies and for some . Then is stable with respect to the Bridgeland stability condition .
Proof.
Suppose that is destabilized by some quotient , i.e., that . Using the short exact sequence in given by the exact triangle
we conclude that at least one of and satisfies this inequality as well.
If , then is a destabilizing quotient whose kernel is a subsheaf of , but this is impossible by Proposition 3.9.
On the other hand, if this is not the case, then and , and we proceed as in the proof of Proposition 3.7: we may choose an -slope stable factor of with and , so that Proposition 3.5 applies. Furthermore, since and thus , the same is true of , letting be the kernel object. Then we have by definition of , and thus the inequality is strict. In other words, we get
writing and as before. Then by Lemma 3.6 we conclude that , but by our assumption on this is impossible. (We have unless , in which case the correct bound is .)
∎
3.2. Reduction to the torsion sheaf case
As usual, we fix a projective normal surface , an ample line bundle on , a finite length subscheme of length , and an object of type with . Setting , recall that belong to the heart defined in Section 2. In particular, as this is a full abelian subcategory of , it makes sense to talk about the image of a homomorphism .
Proposition 3.11.
Let be as above, and suppose . Then given we have that is a torsion sheaf.
Proof.
Let . We use the fact that , combined with the definition of .
First, since , the corresponding short exact sequence in gives a long exact sequence of cohomology
which implies that with equality if and only if (as is torsion-free of rank 1). Furthermore, in this case by definition of , so we conclude that either or . We claim the latter is impossible: indeed, since
we conclude that the quotient must satisfy
for all . However, our assumption on implies that this is false for sufficiently small. Thus .
Second, since , we have the long exact sequence
Since and are torsion-free, we conclude that either or . In the latter case, the fact that is torsion-free and injects into , which is of rank 0 by the above, implies that , so in particular is a subsheaf of , therefore of dimension 0. But then we see that
which is impossible by definition of . So in fact . ∎
4. Proof of the main results
We start by translating our separation of jets statement into a form that can be more effectively approached using Bridgeland stability.
Proposition 4.1.
Let be a projective normal surface, and . We have
Recall we are interested in a bound for such that In light of Proposition 4.1, our strategy will be to use Bridgeland stability conditions in order to show More generally, we consider the vanishing of where is an ample line bundle, a subscheme of dimension 0, and is of type . One advantage of this more general setup is that we can obtain better bounds by relating to other spaces given by changing the relative lengths of and . In what follows, we use the notation and , and similarly for and .
Proposition 4.2.
Let be as above, with . Then if we can find a dimension 0 subscheme and of type with , , and .
Proof.
Choose a surjection and some . Then consider the map
Since belong to the heart , we can define ; similarly, we define . It is immediate that is of type with , and as for , we note that either , in which case and we take , or and is thus surjective, in which case for some with by the long exact sequence of cohomology.
We note that the restriction of to is nontrivial (because by assumption) and factors through , and so we find an element . ∎
Proposition 4.3.
Let be as above, with . Then if we can find a dimension 0 subscheme and of type with , , and .
Proof.
Choose a subscheme with , and let . Then
gives a map . Choose , and let be its image. If , then is the image of some , and so for we get .
Otherwise, let . By chasing the diagram
we find a morphism whose restriction to is given by . Furthermore, we can see the cohomology groups of are of the correct form, with , and for every closed point : for this is clear from the fact that , and for it follows from the long exact sequence
Thus is of type . ∎
Remark 4.4.
For the purposes of the main theorem, it is convenient to assume that decreasing by 1 will have the effect of increasing by 1 and vice versa, while the preceding propositions leave open the possibility that in fact may not need to increase if already factors through the chosen , and similarly may not need to increase if can be extended to for the chosen . However, we note that it is always possible to arrange for to remain constant: in the first case, one takes an arbitrary with and restricts to , and in the second case, one chooses an arbitrary point and takes to be the cone of an arbitrary nonzero element of and extends by . (Note that by assumption on , we have .)
Before proving our main results, we note that a priori our methods give the vanishing of where the tensor products are derived. To obtain Reider-type results one would of course like to have the subjectivity of the map This is implied by the following result:
Proposition 4.5.
Let be a normal surface, a coherent sheaf and a 0-dimensional subscheme. Then the vanishing implies the surjectivity of
Proof.
The vanishing implies the map is surjective. Consider the spectral sequence
Note that and is supported in codimension 2 for all . Therefore only the terms are nontrivial, so the spectral sequence degenerates and . ∎
Remark 4.6.
One can similarly show implies since the non-Gorenstein locus is codimension but we will not make use of this fact.
We now come to the main technical theorem of the paper.
Theorem 4.7.
Let be a projective normal surface over an algebraically closed field, a subscheme of length , and an ample line bundle on with . Let be an object of type with . Choose nonnegative integers with . Then
if and for any nonzero effective divisor on .
Proof.
By repeated applications of Propositions 4.2 and 4.3, we see that to show vanishing of , it suffices show that for all subschemes of length and objects of type with .
By Proposition 3.11, if , then is a torsion sheaf. By Proposition 3.9, we see that the quotient is not destabilizing at the stability condition as long as and . So we conclude in particular that . Then we note that is independent of since , so for , which implies in particular that is strictly destabilizing at . But this is impossible if and by Proposition 3.8. ∎
In order to turn the previous theorem into a Fujita-type bound, we use the auxiliary functions defined in Definition 1.4.
Remark 4.8.
We can also describe as
which shows, for example, that for . In the following corollary, we also prove the lower bound .
Corollary 4.9.
Let be a projective normal surface over an algebraically closed field, a subscheme of length , an ample line bundle on , and an object of type with . Let . Then
whenever .
Proof.
Using that , setting clearly accounts for all the bounds in Theorem 4.7 except for . We claim that the only case in which this bound is relevant is : indeed, for any choice of with , we clearly have
and so
for . It therefore suffices to consider the case .
When , we have
with equality exactly when , in which case implies we need . When and we compute
and when and we again have
∎
Finally, we also give a more explicitly “Reider-type” statement relating the existence of a nontrivial homomorphism to the existence of an effective divisor passing through the points of interest with certain intersection numbers.
Theorem 4.10.
Let be a projective normal surface over an algebraically closed field, a subscheme of length , an ample line bundle on , and an object of type with . Define , i.e., if this is even and otherwise. Assume that satisfies . Then, if
there is an effective divisor such that and
Moreover, if one assumes
for every pair of subscheme and subobject of type (i.e., with ) other than , then passes through all the points of
Proof.
Choose . Then by Proposition 3.11 (whose bound is implied by our bound on ) we have that is a torsion sheaf, so in particular is an effective divisor. (Note that does not contain any points outside of by the assumption that there is no nontrivial map from a skyscraper sheaf to ; in particular, by assumption , we have that is nontrivial.) Note as well that if did not pass through some point , then letting be the subscheme supported on , we would have a map
given by on and 0 on , or in other words, we would get that is the restriction of some nonzero element of . Similarly, if did not pass through some point , letting
we see that the composition
must be trivial, and thus factors through , showing that gives a nonzero element of
Now, as before, we note that the Bridgeland degree is independent of , as . So in particular, either , in which case is a destabilizing quotient at , or , in which case is a strictly destabilizing subsheaf at . In the first case, we see as in the proof of Proposition 3.9 that we must have
which, combined with the fact that
(the rank being nonzero because and ), means that we are in the situation of Lemma 3.6, assuming , and thus . Similarly, in the second case, as in the proof of Proposition 3.7 (where here satisfies ) we get
and thus again are in the situation of Lemma 3.6, assuming , and can deduce . So we conclude that we must have
with .
Finally, it remains to note that we can optimize by using Propositions 4.2 and 4.3 to redistribute points between and while preserving , , and (this last point being because the different homomorphisms constructed all agree away from , which is of codimension 2 and so has no effect on the divisor corresponding to ). In particular, we see that the lower bound for is maximized and the upper bound for is minimized when are as close to equal as possible, i.e., either both or . ∎
Our desired applications immediately follow by applying the above to and .
Proof of Theorem 1.12.
By letting in Theorem 4.7 we can similarly deduce Theorem 1.2. In particular, the classical Fujita’s conjecture for surfaces follows:
Proof of Corollary 1.7.
This follows by applying Corollary 1.5 with and 2, respectively. In the first case, we see that for we have , and in the second case, for , we have . ∎
The Reider-type theorem also recovers a version of the classical Reider theorem for very ampleness (differing from [17] only in that is assumed to be ample, not nef):
Corollary 4.11.
Let be a smooth projective normal surface with (e.g. a smooth complex surface) and an ample line bundle such that satisfies . Let be a subscheme of length 2. Then if
is not surjective there exists an effective divisor passing through (i.e., passing through if , and if , then passes through at least one of , and both if neither is a base point of ) such that
Proof.
If this is not a surjection, then we must have , which is equivalent to by Proposition 4.1. Thus, we can apply Theorem 4.10 with and (thus ) to conclude that there is an effective divisor such that and . (For the statement about support, note that by Theorem 4.7, so must at least pass through some point of .) Since was assumed to be smooth, we have (and the same, of course, is true for ), so the statement follows. ∎
References
- [1] Daniele Arcara and Aaron Bertram “Reider’s theorem and Thaddeus pairs revisited” In Grassmannians, moduli spaces and vector bundles 14, Clay Math. Proc. Amer. Math. Soc., 2011, pp. 51–68
- [2] Philippe Du Bois “Complexe de de Rham filtré d’une variété singulière” In Bulletin de la Société mathématique de France 109, 1981, pp. 41–81
- [3] Lawrence Ein and Robert Lazarsfeld “Global generation of pluricanonical and adjoint linear series on smooth projective threefolds” In Journal of the American Mathematical Society 6.4 JSTOR, 1993, pp. 875–903
- [4] Lawrence Ein, Robert Lazarsfeld and Vladimir Maşek “Global Generation Of Linear Series On Terminal Threefolds” In International Journal of Mathematics 6.01 World Scientific, 1995, pp. 1–18
- [5] Takao Fujita “Contribution to birational geometry of algebraic varieties: open problems” In the 23rd International Symposium, Division of Mathematics, the Taniguchi Foundation; August 22-27, 1988, Katata, 1988
- [6] William Fulton “Intersection theory” Springer Science & Business Media, 2013
- [7] Yi Gu, Lei Zhang and Yongming Zhang “Counterexamples to Fujita’s conjecture on surfaces in positive characteristic” In Advances in Mathematics 400 Elsevier, 2022, pp. 108271
- [8] Francisco Guillen, Vincente Navarro Aznar, Pedro Pascual-Gainza and Fernando Puerta “Hyperrésolutions cubiques et descente cohomologique” Springer, 2006
- [9] Yujiro Kawamata “On Fujita’s freeness conjecture for 3-folds and 4-folds” In arXiv preprint alg-geom/9510004, 1995
- [10] Naoki Koseki “On the Bogomolov–Gieseker inequality in positive characteristic” In International Mathematics Research Notices 2023.24 Oxford University Press, 2023, pp. 20784–20811
- [11] Sándor Kovács and Karl Schwede “Du Bois singularities deform” In Minimal Models and Extremal Rays (Kyoto, 2011), 2016, pp. 49–65 Math. Soc. Japan
- [12] Adrian Langer “A note on k-jet ampleness on surfaces” In arXiv preprint math/9802098, 1998
- [13] Adrian Langer “Bridgeland stability conditions on normal surfaces” In Annali di Matematica Pura ed Applicata (1923-) Springer, 2024, pp. 1–12
- [14] Adrian Langer “Intersection theory and Chern classes on normal varieties” In arXiv preprint arXiv:2210.08766, 2022
- [15] Howard Nuer and Alan Sorani “The Bogomolov-Gieseker-Koseki inequality on surfaces with canonical singularities in arbitrary characteristic” In arXiv preprint arXiv:2308.03307, 2023
- [16] Chris AM Peters and Joseph HM Steenbrink “Mixed hodge structures” Springer Science & Business Media, 2008
- [17] Igor Reider “Vector bundles of rank 2 and linear systems on algebraic surfaces” In Annals of Mathematics 127.2 JSTOR, 1988, pp. 309–316
- [18] Morihiko Saito “Mixed Hodge complexes on algebraic varieties” In Mathematische Annalen 316 Springer, 2000, pp. 283–331
- [19] Fumio Sakai “Reider-Serrano’s method on normal surfaces” In Algebraic Geometry: Proceedings of the International Conference held in L’Aquila, Italy, May 30–June 4, 1988, 2006, pp. 301–319 Springer
- [20] Karl Schwede “F-injective singularities are Du Bois” In American journal of mathematics 131.2 Johns Hopkins University Press, 2009, pp. 445–473
- [21] Wanchun Shen, Sridhar Venkatesh and Anh Duc Vo “On -Du Bois and -rational singularities” In arXiv preprint arXiv:2306.03977, 2023
- [22] The Stacks Project Authors “Stacks Project”, https://stacks.math.columbia.edu, 2018
- [23] Fei Ye and Zhixian Zhu “On Fujita’s freeness conjecture in dimension 5” In Advances in Mathematics 371 Elsevier, 2020, pp. 107210
Department of Mathematics, MIT, 77 Massachusetts Avenue, Cambridge, MA 02139, USA
E-mail address: [email protected]
Department of Mathematics, Harvard University, 1 Oxford Street, Cambridge, MA 02138, USA
E-mail address: [email protected]