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

Positivity vs. slope semistability for bundles with vanishing discriminant

Mihai Fulger Department of Mathematics, University of Connecticut, Storrs, CT 06269-1009, USA Institute of Mathematics of the Romanian Academy, P. O. Box 1-764, RO-014700, Bucharest, Romania [email protected]  and  Adrian Langer Institute of Mathematics, University of Warsaw, ul. Banacha 2, 02-097 Warszawa, Poland [email protected]
Abstract.

It is known that a strongly slope semistable bundle with vanishing discriminant is nef if and only if its determinant is nef. We give an algebraic proof of this result in all characteristics and generalize it to arbitrary proper schemes. We also address a question of S. Misra.

2010 Mathematics Subject Classification:
Primary 14C17, 14C20, 14F06, 14H60, 14N05
The first author was partially supported by the Simons Foundation Collaboration Grant 579353. The second author was partially supported by Polish National Centre (NCN) contract numbers 2018/29/B/ST1/01232 and 2021/41/B/ST1/03741.

1. Introduction

Let \mathcal{E} be a vector bundle of rank rr on a smooth projective variety XX defined over an algebraically closed field. Inspired by the case of line bundles, one might hope that the positivity of \mathcal{E} is determined by the positivity of its characteristic classes. The bundle 𝒪1(n)𝒪1(n)\mathcal{O}_{\mathbb{P}^{1}}(n)\oplus\mathcal{O}_{\mathbb{P}^{1}}(-n) is an easy counterexample. To rectify this, one adds stability assumptions on \mathcal{E}.

Let HH be an ample polarization on XX. In characteristic zero, say that \mathcal{E} is strongly slope semistable with respect to HH if it is slope semistable in the usual sense. In positive characteristic, say that \mathcal{E} is strongly slope semistable with respect to HH if \mathcal{E} and all its iterated Frobenius pullbacks are slope semistable.

On curves the polarization is irrelevant and we just say that \mathcal{E} is strongly semistable. Here the connection between positivity and strong semistability is well known by work of Hartshorne [Har71], Barton [Bar71], and Miyaoka [Miy87]. A strongly semistable bundle \mathcal{E} on a curve is ample (or just nef) if and only if Xc1()>0\int_{X}c_{1}(\mathcal{E})>0 (resp.  0\geq 0). On surfaces, even on 2\mathbb{P}^{2}, it is not sufficient. See Example 2.1. Furthermore, on curves \mathcal{E} is strongly semistable if and only if the twisted normalized bundle 1rdet\mathcal{E}\langle-\frac{1}{r}\det\mathcal{E}\rangle is nef (equivalently nd\mathcal{E}{\rm nd}\,\mathcal{E} is nef).

A link in codimension two between semistability and positivity comes from the famous Bogomolov inequality. The discriminant of \mathcal{E} is Δ()=2rc2()(r1)c12()\Delta(\mathcal{E})=2rc_{2}(\mathcal{E})-(r-1)c_{1}^{2}(\mathcal{E}). Assume that \mathcal{E} is strongly slope semistable with respect to some ample polarization. The classical form of the inequality states that if XX is a surface, then the degree of the discriminant of \mathcal{E} is non-negative. When XX has arbitrary dimension, the Mehta–Ramanathan theorem [MR82] implies that Δ()\Delta(\mathcal{E}) has nonnegative degree with respect to any polarization.

Note that Δ()=0\Delta(\mathcal{E})=0 on curves. This suggests a close connection between (strong) semistability and positivity for vector bundles that are extremal with respect to the Bogomolov inequality, meaning that Δ()\Delta(\mathcal{E}) is numerically trivial. We have the following known results:

Theorem 1.1.

Let XX be a smooth projective variety of dimension nn, and let HH be an ample class on XX. Let \mathcal{E} be a reflexive sheaf of rank rr on XX. The following are equivalent:

  1. (1)

    \mathcal{E} is strongly slope semistable with respect to HH, and Δ()Hn2=0\Delta(\mathcal{E})\cdot H^{n-2}=0.

  2. (2)

    \mathcal{E} is locally free and nd\mathcal{E}{\rm nd}\,\mathcal{E} is nef.

  3. (3)

    \mathcal{E} is universally semistable (see below).

In particular, if \mathcal{E} is strongly slope semistable with respect to HH, and Δ()Hn2=0\Delta(\mathcal{E})\cdot H^{n-2}=0, then \mathcal{E} is a nef (resp. ample) vector bundle if and only if det\det\mathcal{E} is nef (resp. ample).

In the above theorem we say that \mathcal{E} is universally semistable if for every morphism f:YXf:Y\to X from any projective manifold YY, the pullback ff^{*}\mathcal{E} is slope semistable with respect to any ample polarization on YY. It is sufficient to check this condition for morphisms from curves.

When det\det\mathcal{E} is numerically trivial, condition (2)(2) is simply that the vector bundle \mathcal{E} is nef, equivalently \mathcal{E} is numerically flat (\mathcal{E} and \mathcal{E}^{\vee} are nef). Theorem 1.1 takes the following form:

Theorem 1.2.

Let XX be a smooth projective variety of dimension nn over an algebraically closed field, and let HH be an ample polarization. Let \mathcal{E} be a reflexive sheaf on XX. The following are equivalent:

  1. (1)

    \mathcal{E} is strongly slope semistable with respect to HH, and c1()Hn1=ch2()Hn2=0c_{1}(\mathcal{E})\cdot H^{n-1}=\operatorname{ch}_{2}(\mathcal{E})\cdot H^{n-2}=0.

  2. (2)

    \mathcal{E} is universally semistable and all the Chern classes ci()c_{i}(\mathcal{E}) are numerically trivial for i1i\geq 1.

  3. (3)

    \mathcal{E} is locally free and numerically flat.

In particular, if \mathcal{E} is a strongly slope semistable with respect to HH with c1()c_{1}(\mathcal{E}) numerically trivial, then \mathcal{E} is nef if and only if Δ()\Delta(\mathcal{E}) is numerically trivial.

The result fails if \mathcal{E} is only assumed to be torsion free. It also fails if the condition ch2()Hn2=0\operatorname{ch}_{2}(\mathcal{E})\cdot H^{n-2}=0 in (1) is replaced by c2()Hn2=0c_{2}(\mathcal{E})\cdot H^{n-2}=0.

The two theorems are essentially equivalent. As mentioned before, they are not new. In characteristic zero, Theorem 1.1 is proved by [Nak99]. It is proved by [GKP16] for more general polarizations by movable curves on varieties with mild singularities. It is proved by [BB08] for principal GG-bundles, and by [BHR06] as part of their study of Higgs bundles. Note that for Higgs bundles with nontrivial Higgs field it is not known whether an analogue of (3)(1)(3)\Rightarrow(1) holds (this is known as Bruzzo’s conjecture). See [BBG19]. The proofs given in the references above have important transcendental components coming from [Sim92] or [DPS94]. The last statement of Theorem 1.1 has also been observed by [MR21]. In positive characteristic, a version of the result is proved algebraically by the second named author in [Lan11] making crucial use of the Frobenius morphism. In characteristic zero, [Lan19, Corollary 4.10] gives an algebraic proof of the implication (1)(3)(1)\Rightarrow(3) for the more general case of Higgs bundles by reduction to positive characteristic. See also [Lan15, Theorem 12]. In characteristic zero, Theorem 1.2 is proved in [Sim92, Theorem 2]. A positive characteristic version appears in [Lan11, Proposition 5.1].

A conjecture of Bloch predicts that the Chern classes of numerically flat bundles vanish in B(X)B^{*}(X)\otimes\mathbb{Q}, where Bm(X)B^{m}(X) is the group of codimension mm cycles modulo algebraic equivalence (see [Lan21, Conjecture 3.2]). This would strengthen Theorem 1.2. However, this is known only if kk has positive characteristic (see [Lan21, Corollary 3.7]). In this case the result holds also for general proper schemes (see Corollary 3.10). Analogously, one can formulate a similar conjecture and a result for bundles in Theorem 1.1 (see [Lan21, Proposition 3.6]). If k=k=\mathbb{C} one can prove only a weaker version of the vanishing of Chern classes in the rational cohomology H2(Xan,)H^{2*}(X^{\rm an},\mathbb{Q}), where XanX^{\rm an} is the complex manifold underlying variety XX. This result would follow from Bloch’s conjecture by using the cycle map B(X)H2(Xan,)B^{*}(X)\otimes\mathbb{Q}\to H^{2*}(X^{\rm an},\mathbb{Q}).

1.1. Main results

We give algebraic proofs for Theorems 1.1 and 1.2 that are characteristic free. We avoid the application of the non-abelian Hodge theorem [Sim92, Corollary 1.3] or the reduction to positive characteristic techniques of [Lan11, Lan19]. Moreover, we generalize both theorems to general proper schemes over an algebraically closed field (see Theorem 3.13 and Corollary 3.14).

1.2. On a question of S. Misra

On curves CC it follows from [Miy87] that \mathcal{E} is strongly semistable if and only if every effective divisor on C()\mathbb{P}_{C}(\mathcal{E}) is nef. The first named author proved in [Ful11] a generalization of this for cycles of arbitrary (co)dimension in C()\mathbb{P}_{C}(\mathcal{E}). It is interesting to see to what extent does Miyaoka’s result carry over to a projective manifold XX of arbitrary dimension. The equality Eff¯(X)=Nef(X)\operatorname{\overline{Eff}}(X)=\operatorname{{Nef}}(X) of cones of divisors is a necessary condition which held trivially on curves. Here one has the following result:

Theorem 1.3.

Let XX be a smooth projective variety such that Eff¯(X)=Nef(X)\operatorname{\overline{Eff}}(X)=\operatorname{{Nef}}(X). Let \mathcal{E} be a strongly slope semistable bundle with respect to some ample polarization of XX and assume that Δ()0\Delta(\mathcal{E})\equiv 0. Then Eff¯(())=Nef(())\operatorname{\overline{Eff}}(\mathbb{P}(\mathcal{E}))=\operatorname{{Nef}}(\mathbb{P}(\mathcal{E})).

In the characteristic zero case this result was proven by S. Misra in [Mis21, Theorem 1.2] as an application of Theorem 1.1. A similar proof works also in an arbitrary characteristic. Misra [Mis21, Question 3.11] also asks about a possible converse to this result.

Question 1.4.

Let XX be a smooth projective variety. Let \mathcal{E} be a vector bundle on XX such that every effective divisor on ()\mathbb{P}(\mathcal{E}) is nef. Is it true that \mathcal{E} is slope semistable with respect to some (any) polarization of XX and Δ()0\Delta(\mathcal{E})\equiv 0? Or equivalently, is nd\operatorname{{\mathcal{E}}nd\,}\mathcal{E} numerically flat?

For every n2n\geq 2, the tangent bundle TnT_{\mathbb{P}^{n}} is a counterexample to the current phrasing of the question. In Example 5.7 we even give a slope unstable counterexample. However [Mis21] also observes that under the hypothesis of Theorem 1.3 the effective and nef cones of divisors coincide on (Symm)\mathbb{P}(\operatorname{Sym}^{m}\mathcal{E}) for all m0m\geq 0. This motivates the following positive answer to a version of Question 1.4 in arbitrary characteristic.

Theorem 1.5.

Let XX be a smooth projective variety. Let \mathcal{E} be a vector bundle on XX such that every effective divisor on (Symm)\mathbb{P}(\operatorname{Sym}^{m}\mathcal{E}) is nef for all m0m\geq 0. Then \mathcal{E} is strongly slope semistable with respect to any polarization HH and Δ()0\Delta(\mathcal{E})\equiv 0.

Our proof shows that in fact existence of only one positive even value 2m2m for which the cones Eff¯((Sym2m))\operatorname{\overline{Eff}}(\mathbb{P}(\operatorname{Sym}^{2m}\mathcal{E})) and Nef((Sym2m))\operatorname{{Nef}}(\mathbb{P}(\operatorname{Sym}^{2m}\mathcal{E})) coincide is sufficient. The key idea is a result on plethysms which guarantees that the line bundle (det)2m(\det\mathcal{E})^{\otimes 2m} is a subbundle of SymrSym2m\operatorname{Sym}^{r}\operatorname{Sym}^{2m}\mathcal{E}. In characteristic zero, this also follows from [BCI11]. One can also see that equality Eff¯((nd))=Nef((nd))\operatorname{\overline{Eff}}(\mathbb{P}(\operatorname{{\mathcal{E}}nd\,}\mathcal{E}))=\operatorname{{Nef}}(\mathbb{P}(\operatorname{{\mathcal{E}}nd\,}\mathcal{E})) implies that nd\operatorname{{\mathcal{E}}nd\,}\mathcal{E} is numerically flat, which provides a satisfactory answer to the original question.

1.3. Acknowledgments

The authors would like to thank S. Misra, D. S. Nagaraj and J. Weyman for useful comments, suggestions, and references.

2. Preliminaries

Let XX be a connected proper scheme over an algebraically closed field kk and \mathcal{E} be a vector bundle on XX. Note that terms vector bundle and locally free sheaf are used interchangeably (since XX is connected, a locally free sheaf has the same rank at every point of XX).

2.1. Positive bundles

Let ()=X()=Proj𝒪XSym\mathbb{P}(\mathcal{E})=\mathbb{P}_{X}(\mathcal{E})={\rm Proj}_{\mathcal{O}_{X}}\operatorname{Sym}^{\bullet}\mathcal{E} with natural bundle map π:()X\pi:\mathbb{P}(\mathcal{E})\to X, and let ξ=c1(𝒪()(1))\xi=c_{1}(\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)). We say that \mathcal{E} is ample (resp. nef) on XX, if ξ\xi is ample (resp. nef) on ()\mathbb{P}(\mathcal{E}). The definition also makes sense for coherent sheaves.

2.2. Numerically trivial Chern classes

In the following we write A(X)A_{*}(X) for the Chow group of rational equivalence classes on XX and A(X)=A(XidX)A^{*}(X)=A^{*}(X\stackrel{{\scriptstyle\rm id}}{{\longrightarrow}}X) for the operational Chow ring, i.e., the group of bivariant rational equivalence classes (see [Ful98, Chapter 17]). The Chern classes of \mathcal{E} are operations on the Chow group (see [Ful98, Chapter 3]) so they are elements of A(X)A^{*}(X). Following [Ful98, Chapter 19], we say that cj()c_{j}(\mathcal{E}) is numerically trivial if for every proper closed subscheme YXY\subseteq X of dimension jj we have Ycj()=Xcj()[Y]=0\int_{Y}c_{j}(\mathcal{E})=\int_{X}c_{j}(\mathcal{E})\cap[Y]=0, where X:A0(X)\int_{X}:A_{0}(X)\to\mathbb{Z} denotes the natural degree map. Then we write cj()0c_{j}(\mathcal{E})\equiv 0. By additivity it is sufficient to check the condition Xcj()[Y]=0\int_{X}c_{j}(\mathcal{E})\cap[Y]=0 for all YY that are irreducible and reduced. Similarly, we can define numerical triviality for any polynomial in Chern classes of \mathcal{E}, or even in Chern classes of finitely many bundles.

2.3. Positive polynomials

Consider n1n\geq 1 and grade [c1,,cn]\mathbb{Q}[c_{1},\ldots,c_{n}] so that degci=i\deg c_{i}=i. Let P(c1,,cn)P(c_{1},\ldots,c_{n}) be a weighted homogeneous polynomial of degree nn. Fulton and Lazarsfeld proved in [FL83] that XP()0\int_{X}P(\mathcal{E})\geq 0 for every nn-dimensional variety XX and every nef vector bundle \mathcal{E} on XX if and only if PP is a linear combination with nonnegative coefficients of Schur polynomials of degree nn. We call such polynomials positive. For example, the degree 1 positive polynomials are spanned over 0\mathbb{Q}_{\geq 0} by c1c_{1}, while the degree 2 ones are spanned by c2c_{2} and by c12c2c_{1}^{2}-c_{2}. In particular, c12=c2+(c12c2)c_{1}^{2}=c_{2}+(c_{1}^{2}-c_{2}) is positive, but c122c2c_{1}^{2}-2c_{2} is not.

Let XX be a smooth projective surface and let \mathcal{E} be a nef vector bundle of rank rr on it. Then c1()c_{1}(\mathcal{E}) is nef, and Xc2()\int_{X}c_{2}(\mathcal{E}) and X(c12()c2())\int_{X}(c_{1}^{2}(\mathcal{E})-c_{2}(\mathcal{E})) are nonnegative integers. If \mathcal{E} is strongly slope semistable with respect to some polarization, Bogomolov’s inequality gives X(2rc2()(r1)c12())0\int_{X}(2rc_{2}(\mathcal{E})-(r-1)c_{1}^{2}(\mathcal{E}))\geq 0. However, there exist examples of strongly slope semistable vector bundles \mathcal{E} on surfaces such that all the above positivity conditions hold for the characteristic classes of \mathcal{E} without \mathcal{E} being nef.

Example 2.1.

Let =𝒪2(1)3(T2(1))\mathcal{E}=\mathcal{O}_{\mathbb{P}^{2}}(-1)\otimes\bigotimes^{3}(T_{\mathbb{P}^{2}}(-1)). This is strongly slope semistable since T2T_{\mathbb{P}^{2}} is strongly slope semistable. Let h=c1(𝒪2(1))h=c_{1}(\mathcal{O}_{\mathbb{P}^{2}}(1)). Then r=rk=8r={\rm rk}\,\mathcal{E}=8, c1()=4hc_{1}(\mathcal{E})=4h, and c2()=16h2c_{2}(\mathcal{E})=16h^{2}. We compute c12()c2()=0c_{1}^{2}(\mathcal{E})-c_{2}(\mathcal{E})=0. So the characteristic classes of \mathcal{E} suggest that \mathcal{E} might be nef. However, the restriction of \mathcal{E} to every line in 2\mathbb{P}^{2} is 𝒪1(1)3(𝒪1(1)𝒪1)\mathcal{O}_{\mathbb{P}^{1}}(-1)\otimes\bigotimes^{3}(\mathcal{O}_{\mathbb{P}^{1}}(1)\oplus\mathcal{O}_{\mathbb{P}^{1}}) and this has 𝒪1(1)\mathcal{O}_{\mathbb{P}^{1}}(-1) as a summand, hence it is not nef. ∎

See also [BHP14, Section 5] for a related example.

Convention 2.2.

For the rest of this section we assume that XX is a smooth projective variety of dimension nn.

2.4. Positive cones of cycles

Denote by Nc(X)N^{c}(X) the space of numerical classes of cycles of codimension cc with real coefficients. For example N1(X)N^{1}(X) is the real Néron–Severi space spanned by Cartier divisors modulo numerical equivalence. The space Nc(X)N^{c}(X) is finite dimensional. It contains important convex cones. For instance it contains the pseudoeffective cone Eff¯c(X)\operatorname{\overline{Eff}}^{c}(X), the closure of the cone spanned by classes of closed subsets of codimension cc. It also contains the nef cone Nefc(X)\operatorname{{Nef}}^{c}(X), the set of classes that intersect every cc-dimensional subvariety nonnegatively, the dual of the pseudoeffective cone in the complementary codimension. The nef cone and the pseudoeffective cone in codimension cc each span Nc(X)N^{c}(X) and do not contain linear subspaces of Nc(X)N^{c}(X). See [FL17a].

When c=1c=1, we put Eff¯(X)=Eff¯1(X)\operatorname{\overline{Eff}}(X)=\operatorname{\overline{Eff}}^{1}(X) and Nef(X)=Nef1(X)\operatorname{{Nef}}(X)=\operatorname{{Nef}}^{1}(X). In this case, Nef(X)Eff¯(X)\operatorname{{Nef}}(X)\subseteq\operatorname{\overline{Eff}}(X). The interior of the nef cone is the ample cone, and the interior of the pseudoeffective cone is the big cone of numerical classes with positive volume [Laz04a, Chapter 2.2].

2.5. Determinant and discriminant

Let \mathcal{F} be a torsion free sheaf on the smooth projective XX. Then the singular (non-locally free) locus of \mathcal{F} has codimension 2\geq 2. If \mathcal{F} is reflexive, then the singular locus has codimension 3\geq 3.

One can abstractly compute Chern classes of coherent sheaves in A(X)A^{*}(X) by taking locally free resolutions and using the additivity of the Chern character ch{\rm ch}. The determinant line bundle det\det\mathcal{F} of a torsion free sheaf can be defined by extending from the locally free locus, or considering the reflexive hull (rk)(\bigwedge^{\rm rk\,\mathcal{F}}\mathcal{F})^{\vee\vee}. This gives a concrete definition of c1c_{1}. The second Chern class of a reflexive sheaf can be similarly concretely defined.

Remark 2.3.

Let f:YXf:Y\to X be a morphism of nonsingular projective varieties and let \mathcal{E} be a coherent sheaf on XX. Then ch(𝐋f)=fch()\operatorname{ch}({\bf L}f^{*}\mathcal{E})=f^{*}\operatorname{ch}(\mathcal{E}) in the Chow ring of YY, where 𝐋f{\bf L}f^{*}\mathcal{E} is the derived pullback. In particular ci(f)=fci()c_{i}(f^{*}\mathcal{E})=f^{*}c_{i}(\mathcal{E}) for all ii in each of the following cases:

  1. (1)

    \mathcal{E} is locally free.

  2. (2)

    ff is flat.

  3. (3)

    ff is the inclusion of a Cartier divisor and \mathcal{E} is torsion free.

If ı:DX\imath:D\hookrightarrow X is the inclusion of a Cartier divisor and \mathcal{E} is reflexive, then ı\imath^{*}\mathcal{E} is torsion free. If |H||H| is a basepoint free linear system and \mathcal{E} is torsion free (resp. reflexive), then ı\imath^{*}\mathcal{E} is again torsion free (resp. reflexive) for DD general in |H||H|. See [HL10, Corollary 1.1.14].

The discriminant of \mathcal{E} is

Δ()=2rc2()(r1)c12(),\Delta(\mathcal{E})=2rc_{2}(\mathcal{E})-(r-1)c_{1}^{2}(\mathcal{E}),

where r=rkr={\rm rk}\,\mathcal{E}. It is an element of A2(X)A^{2}(X), but we use the same notation for its image in N2(X)N^{2}(X). The class

logr+c1()rΔ()2r2+\log r+\frac{c_{1}(\mathcal{E})}{r}-\frac{\Delta(\mathcal{E})}{2r^{2}}+\ldots

in A(X)A^{*}(X)_{\mathbb{R}} is the formal logarithm of the Chern character ch(){\rm ch}(\mathcal{E}). It follows that Δ()2rk2\frac{\Delta(\mathcal{E})}{2{\rm rk}^{2}\,\mathcal{E}} is additive for tensor products if one of the factors is locally free, just like the slope. In particular, Δ()=Δ(𝒪X(D))\Delta(\mathcal{E})=\Delta(\mathcal{E}\otimes\mathcal{O}_{X}(D)) for every divisor DD on XX. Furthermore, note that if c1()0c_{1}(\mathcal{E})\equiv 0, then Δ()0\Delta(\mathcal{E})\equiv 0 if and only if c2()0c_{2}(\mathcal{E})\equiv 0. The vanishing Δ()0\Delta(\mathcal{E})\equiv 0 is equivalent to the numerical vanishing of the second Chern class c2c_{2} of 1rdet\mathcal{E}\langle-\frac{1}{r}\det\mathcal{E}\rangle, the formal twist of \mathcal{E} in the sense of [Laz04b, Chapter 6.2]. If \mathcal{E} is locally free, then Δ()0\Delta(\mathcal{E})\equiv 0 is also equivalent to the numerical vanishing of c2c_{2} of rdet\mathcal{E}^{\otimes r}\otimes\det\mathcal{E}^{\vee} or nd{\mathcal{E}}{\rm nd}\,\mathcal{E}.

2.6. Semistability

Let HH be an ample (or just nef) divisor on XX. For a nonzero torsion free sheaf \mathcal{E} on XX, we define the slope by μH()=c1()Hn1rk\mu_{H}(\mathcal{E})=\frac{c_{1}(\mathcal{E})\cdot H^{n-1}}{{\rm rk}\,\mathcal{E}}. We say that \mathcal{E} is μH\mu_{H}-semistable (or slope semistable with respect to HH) if no proper subsheaf 00\neq\mathcal{F}\subsetneq\mathcal{E} verifies μH()>μH()\mu_{H}(\mathcal{F})>\mu_{H}(\mathcal{E}). We say that \mathcal{E} is μH\mu_{H}-stable (or slope stable with respect to HH) if no proper subsheaf 00\neq\mathcal{F}\subsetneq\mathcal{E} with rk<rk{\rm rk}\,\mathcal{F}<{\rm rk}\,\mathcal{E} has μH()μH()\mu_{H}(\mathcal{F})\geq\mu_{H}(\mathcal{E}). When XX is a curve and HH is ample, slope semistability is independent of HH. In this case we simply say that \mathcal{E} is semistable. In positive characteristic it is useful to also consider Frobenius pullbacks.

Definition 2.4.

A torsion free sheaf \mathcal{E} is called strongly μH\mu_{H}-(semi)stable if (FXm)(F_{X}^{m})^{*}\mathcal{E} is μH\mu_{H}-(semi)stable for all m0m\geq 0. Here FXF_{X} denotes the Frobenius morphism in positive characteristic, and the identity morphism in characteristic zero.

Examples show that these notions are indeed stronger that μH\mu_{H}-(semi)stability in positive characteristic. The first example showing this is due to J.-P. Serre and it was published in [Gie71]. More precisely, the example shows that there exists a stable bundle of rank 22 and degree 11 on a genus 33 curve in characteristic 33, whose Frobenius pullback splits as direct sum of line bundles of different degrees. Nowadays there are many more examples of stable bundles that are not strongly semistable in an arbitrary positive characteristic. See, e.g., [JP15, Theorem 1.1.3] for recent examples that appear on any smooth projective curve in large characteristic.

2.7. Semistability on curves

In the sequel we will use the following well–known lemma. It goes back to R. Hartshorne [Har70] in the characteristic zero case, with a subsequent algebraic proof in any characteristic due to C. M. Barton [Bar71, Theorem 2.1]. See also [Mor98, Proposition 7.1].

Lemma 2.5.

Let CC be a smooth projective curve defined over some algebraically closed field kk and let \mathcal{E} be a degree 0 vector bundle on CC. Then \mathcal{E} is strongly semistable if and only if \mathcal{E} is nef.

We will also use the following standard lemma (see, e.g., [LP97, Lemma 7.1.2]).

Lemma 2.6.

Let \mathcal{E} be a semistable vector bundle on a smooth projective curve CC defined over some algebraically closed field kk. Then

h0(C,)rk+deg.h^{0}(C,\mathcal{E})\leq{\rm rk}\,\mathcal{E}+\deg\mathcal{E}.

3. Numerically flat vector bundles on proper schemes

Let XX be a proper scheme over a field kk.

Definition 3.1.

A vector bundle \mathcal{E} on XX is called numerically flat if both \mathcal{E} and \mathcal{E}^{\vee} are nef.

Remark 3.2.

A bundle \mathcal{E} is numerically flat if and only if it is nef with c1()0c_{1}(\mathcal{E})\equiv 0. Indeed if \mathcal{E} and \mathcal{E}^{\vee} are nef, then c1(det)=c1()c_{1}(\det\mathcal{E})=c_{1}(\mathcal{E}) is numerically trivial. Conversely, if \mathcal{E} is nef with c1()0c_{1}(\mathcal{E})\equiv 0 then by [Laz04b, Theorem 6.2.12] any exterior power of \mathcal{E} is nef and hence detr1\mathcal{E}^{\vee}\simeq\det\mathcal{E}^{\vee}\otimes\bigwedge^{r-1}\mathcal{E} is also nef.

It follows that if \mathcal{E} is numerically flat, then all tensor functors of \mathcal{E} and their duals are numerically flat, e.g., tensor powers, symmetric powers, divided powers, exterior powers, and all other Schur and co-Schur (Weyl) functors associated to \mathcal{E}.

3.1. Numerical triviality of Chern classes of numerically flat bundles

Proposition 3.3.

Let XX be a proper scheme over a field kk and let \mathcal{E} be a numerically flat vector bundle on XX. Then cj()c_{j}(\mathcal{E}) is numerically trivial for all j>0j>0.

Proof.

The proof is by induction on jj. For j=1j=1 the assertion is clear since by assumption for any proper curve YXY\subseteq X we have Xc1()[Y]0\int_{X}c_{1}(\mathcal{E})\cap[Y]\geq 0 and Xc1()[Y]=Xc1()[Y]0\int_{X}c_{1}(\mathcal{E}^{\vee})\cap[Y]=-\int_{X}c_{1}(\mathcal{E})\cap[Y]\geq 0.

Let YXY\subseteq X be a jj-dimensional subvariety of XX. Let sj()=π(ξj+r1)s_{j}(\mathcal{E}^{\vee})=\pi_{*}(\xi^{j+r-1}) be the jj-th Segre class of the dual of \mathcal{E}. By [Ful98, Chapter 3.2], we have sj()=(1)j+1cj()i=1j1(1)isi()cji()s_{j}(\mathcal{E}^{\vee})=(-1)^{j+1}c_{j}(\mathcal{E})-\sum_{i=1}^{j-1}(-1)^{i}s_{i}(\mathcal{E}^{\vee})c_{j-i}(\mathcal{E}). By our induction hypothesis, we deduce Ysj(Y)=(1)j+1Ycj(Y)\int_{Y}s_{j}(\mathcal{E}_{Y}^{\vee})=(-1)^{j+1}\cdot\int_{Y}c_{j}(\mathcal{E}_{Y}), so is suffices to prove that Ysj(Y)=0\int_{Y}s_{j}(\mathcal{E}^{\vee}_{Y})=0. Set Z(Y)Z\coloneqq{\mathbb{P}(\mathcal{E}_{Y})} and 𝒪(Y)(1)\mathcal{L}\coloneqq\mathcal{O}_{\mathbb{P}(\mathcal{E}_{Y})}(1). Since \mathcal{E} is nef, Y\mathcal{E}_{Y} and \mathcal{L} are also nef. By the asymptotic Riemann–Roch theorems (see [Kol96, Chapter VI, Corollary 2.14 and Theorem 2.15]) we have

h0(Z,m)=χ(Z,m)+O(mj+r2)=Zc1()j+r1(j+r1)!mj+r1+O(mj+r2).h^{0}(Z,\mathcal{L}^{\otimes m})=\chi(Z,\mathcal{L}^{\otimes m})+O(m^{j+r-2})=\frac{\int_{Z}c_{1}(\mathcal{L})^{j+r-1}}{(j+r-1)!}m^{j+r-1}+O(m^{j+r-2}).

Since Y\mathcal{E}_{Y} is numerically flat, SymmY{\rm Sym}^{m}\mathcal{E}_{Y} is also numerically flat. Hence for any ample divisor HH on YY we have h0(Y,(SymmY)(H))=0h^{0}(Y,({\rm Sym}^{m}\mathcal{E}_{Y})(-H))=0 (otherwise SymmY{\rm Sym}^{m}\mathcal{E}_{Y} contains 𝒪(H)\mathcal{O}(H) contradicting the nefness of (SymmY)(\operatorname{Sym}^{m}\mathcal{E}_{Y})^{\vee}). Again using the asymptotic Riemann–Roch theorem we get

h0(Z,m)=h0(Y,SymmY)\displaystyle h^{0}(Z,\mathcal{L}^{\otimes m})=h^{0}(Y,{\rm Sym}^{m}\mathcal{E}_{Y}) h0(Y,(SymmY)(H))+h0(H,SymmH)\displaystyle\leq h^{0}(Y,({\rm Sym}^{m}\mathcal{E}_{Y})(-H))+h^{0}(H,{\rm Sym}^{m}\mathcal{E}_{H})
=h0((H),𝒪(H)(m))=O(mj+r2).\displaystyle=h^{0}(\mathbb{P}(\mathcal{E}_{H}),\mathcal{O}_{\mathbb{P}(\mathcal{E}_{H})}(m))=O(m^{j+r-2}).

Summing up, we have Ysj(Y)=Zc1()j+r1=0\int_{Y}s_{j}(\mathcal{E}^{\vee}_{Y})=\int_{Z}c_{1}(\mathcal{L})^{j+r-1}=0 as required. ∎

Remark 3.4.

The idea of proof of vanishing of highest Segre classes comes from the proof of [Ful20, Proposition 5.1], in turn inspired by the proof of the Bogomolov inequality. We avoid using this result and semistability and give a proof working in an arbitrary characteristic. The proof of an analogue of [Ful20, Proposition 5.1] would require small rewriting and the use of deep results of Ramanan and Ramanathan [RR84] on the behaviour of strong slope semistability.

Remark 3.5.

If XX is a smooth complex manifold the above proposition was proven in [DPS94, Corollary 1.19] using earlier deep analytic results. If XX is a smooth variety and kk has positive characteristic the above proposition follows from [Lan11, Proposition 5.1 and Theorem 4.1]. The proof of [Lan11, Proposition 5.1] cites rather deep results from [FL83] although it uses only a much weaker and easier result of Kleiman [Kle69]. However, it also depends on [Lan04] and the proof above is much more elementary.

Remark 3.6.

From [BG71] or [FL83], the Chern class ck()c_{k}(\mathcal{E}) is nef if \mathcal{E} is a nef bundle. Thus for \mathcal{E} numerically flat the classes ck()c_{k}(\mathcal{E}), ck()c_{k}(\mathcal{E}^{\vee}), and sk()s_{k}(\mathcal{E}^{\vee}) are nef. Using induction and the fact that Nefk(X)\operatorname{{Nef}}^{k}(X) does not contain linear subspaces (cf. [FL17a]), this gives another argument than the one above. Note though that [BG71] and consequently also [FL83] use the hard Lefschetz theorem on cohomology so this proof is much harder.

Together with the main result of [Lan04] the above propostion implies the following corollary:

Corollary 3.7.

Let f:XSf:X\to S be a flat projective morphism of noetherian schemes. Then the set of numerically flat vector bundles of fixed rank rr on the fibers of ff is bounded.

Proof.

Let 𝒪X(1)\mathcal{O}_{X}(1) be an ff-very ample line bundle on XX and let \mathcal{E} be a rank rr numerically flat vector bundle on a geometric fiber XsX_{s} for some geometric point ss of SS. The singular Grothendieck–Riemann–Roch theorem (see [Ful98, Corollary 18.3.1]) and Proposition 3.3 imply that

χ(Xs,(m))=Xsch((m))Td(Xs)=rXsch(𝒪Xs(m))Td(Xs)=rχ(Xs,𝒪Xs(m)).\chi(X_{s},\mathcal{E}(m))=\int_{X_{s}}\operatorname{ch}(\mathcal{E}(m))\cap{\mathrm{Td}}({X_{s}})=r\int_{X_{s}}\operatorname{ch}(\mathcal{O}_{X_{s}}(m))\cap{\mathrm{Td}}({X_{s}})=r\chi({X_{s}},\mathcal{O}_{X_{s}}(m)).

Since ff is flat, for every connected component S0S_{0} of SS the Hilbert polynomial Ps(m)=χ(Xs,𝒪Xs(m))P_{s}(m)=\chi({X_{s}},\mathcal{O}_{X_{s}}(m)) is independent of the geometric point ss of S0S_{0}. Moreover, any numerically flat vector bundle on XsX_{s} is slope 𝒪Xs(1)\mathcal{O}_{X_{s}}(1)-semistable (the general definition of slope semistability in case of singular projective schemes can be found in [HL10, Definition and Corollary 1.6.9]). Therefore the required assertion follows from [Lan04, Theorem 4.4]. ∎

Remark 3.8.

In case XX is a normal variety and S=SpeckS=\operatorname{Spec}k, the above corollary was proved in [Lan12, Theorem 1.1]. If S=SpeckS=\operatorname{Spec}k and kk is a finite field the above corollary was proved in [DW20, Theorem 2.4].

Theorem 3.9.

Let XX be a projective scheme over a perfect field kk of positive characteristic. Let \mathcal{E} be a rank rr vector bundle on XX. Then the following conditions are equivalent:

  1. (1)

    \mathcal{E} is numerically flat.

  2. (2)

    The set {(FXm)}m0\{(F_{X}^{m})^{*}\mathcal{E}\}_{m\in\mathbb{Z}_{\geq 0}} is bounded.

  3. (3)

    There exist m1>m20m_{1}>m_{2}\geq 0 such that (FXm1)(F_{X}^{m_{1}})^{*}\mathcal{E} and (FXm2)(F_{X}^{m_{2}})^{*}\mathcal{E} are algebraically equivalent.

Proof.

If \mathcal{E} is numerically flat then all m=(FXm)\mathcal{E}_{m}=(F_{X}^{m})^{*}\mathcal{E} are numerically flat, so (1)(2)(1)\Rightarrow(2) follows from Corollary 3.7. Assume (2)(2). Then by definition there exists a kk-scheme SS of finite type and an SS-flat coherent sheaf \mathcal{F} on XS:=X×kSX_{S}:=X\times_{k}S such that for every mm\in\mathbb{Z} there exists a geometric kk-point sms_{m} in SS such that Xsmm\mathcal{F}_{X_{s_{m}}}\simeq\mathcal{E}_{m}. Now (3)(3) follows by the pigeonhole principle applied to the finitely many connected components of SS. Assume (3)(3). Then for all m0m\geq 0 the bundles (FXm1+m)(F_{X}^{m_{1}+m})^{*}\mathcal{E} and (FXm2+m)(F_{X}^{m_{2}+m})^{*}\mathcal{E} are algebraically equivalent. This implies that the family {(FXm2+m(m1m2))}m0\{(F_{X}^{m_{2}+m(m_{1}-m_{2})})^{*}\mathcal{E}\}_{m\in\mathbb{Z}_{\geq 0}} is bounded. The implication (3)(1)(3)\Rightarrow(1) follows as in the first part of proof of [Lan11, Proposition 5.1]. ∎

The following corollary can be thought of as a generalization of [DW20, Theorem 2.3] from finite fields to arbitrary perfect fields of positive characteristic. In case XX is smooth the result is contained in [Lan21, Corollary 3.7].

Corollary 3.10.

Let XX be a projective scheme over a perfect field kk of positive characteristic. Let \mathcal{E} be a numerically flat vector bundle on XX. Then for all i>0i>0 the Chern classes ci()c_{i}(\mathcal{E}) are, up to torsion, algebraically equivalent to 0.

Proof.

By the above theorem we know that there exist m1>m20m_{1}>m_{2}\geq 0 such that (FXm1)(F_{X}^{m_{1}})^{*}\mathcal{E} and (FXm2)(F_{X}^{m_{2}})^{*}\mathcal{E} are algebraically equivalent. Since ci((FXm))=pimci()c_{i}((F_{X}^{m})^{*}\mathcal{E})=p^{im}c_{i}(\mathcal{E}) in A(X)A^{*}(X), we get

0=ci((FXm1))ci((FXm2))=(pim1pim2)ci()0=c_{i}((F_{X}^{m_{1}})^{*}\mathcal{E})-c_{i}((F_{X}^{m_{2}})^{*}\mathcal{E})=(p^{im_{1}}-p^{im_{2}})c_{i}(\mathcal{E})

in B(X)B^{*}(X), so ci()=0c_{i}(\mathcal{E})=0 in B(X)B^{*}(X)_{\mathbb{Q}}. ∎

3.2. Characterizations of numerically flat bundles

Let XX be a proper scheme over an algebraically closed field kk.

Definition 3.11.

A vector bundle \mathcal{E} on XX is called universally semistable if for all kk-morphisms f:CXf:C\to X from smooth connected projective curves CC over kk the pullback ff^{*}\mathcal{E} is semistable. We say that \mathcal{E} is Nori semistable if it is universally semistable and c1()c_{1}(\mathcal{E}) is numerically trivial.

To better justify the terminology, note that if \mathcal{E} is universally semistable, and f:YXf:Y\to X is a morphism from a projective manifold YY over kk, then for every polarization HH on YY the pullback ff^{*}\mathcal{E} is μH\mu_{H}-semistable.

If f:YXf:Y\to X is a proper generically finite morphism between smooth projective varieties and \mathcal{E} is a strongly slope HH-semistable bundle on XX then ff^{*}\mathcal{E} is slope fHf^{*}H-semistable bundle. This motivates the following definition:

Definition 3.12.

If XX is irreducible then we say that a vector bundle \mathcal{E} on XX is strongly semistable if there exist a proper generically finite kk-morphism f:YXf:Y\to X from a smooth projective kk-variety YY to XX and an ample divisor HH on YY such that the bundle ff^{*}\mathcal{E} is strongly slope HH-semistable. In general, we say that a vector bundle \mathcal{E} on XX is strongly semistable if its restriction to every irreducible component of XX is strongly semistable.

A line bundle \mathcal{L} is said to be τ\tau-trivial, if m\mathcal{L}^{\otimes m} is algebraically equivalent to 𝒪X\mathcal{O}_{X} for some m1m\geq 1. This notion is equivalent to \mathcal{L} being numerically trivial by [Kle05, Theorem 6.3]. Numerically flat bundles can be seen as a higher rank version of τ\tau-trivial bundles. In the proof we use Theorem 4.8.

Theorem 3.13.

Let XX be a proper scheme over an algebraically closed field kk. Let \mathcal{E} be a rank rr vector bundle on XX. Then the following conditions are equivalent:

  1. (1)

    \mathcal{E} is numerically flat.

  2. (2)

    \mathcal{E} is Nori semistable.

  3. (3)

    \mathcal{E} is strongly semistable and cj()c_{j}(\mathcal{E}) is numerically trivial for all j>0j>0.

  4. (4)

    \mathcal{E} is strongly semistable and both c1()c_{1}(\mathcal{E}) and c2()c_{2}(\mathcal{E}) are numerically trivial.

  5. (5)

    \mathcal{E} is strongly semistable and for every coherent sheaf \mathcal{F} on XX we have χ(X,)=rχ(X,)\chi(X,\mathcal{E}\otimes\mathcal{F})=r\cdot\chi(X,\mathcal{F}).

Proof.

The equivalence of (1)(1) and (2)(2) is well known and it follows from Lemma 2.5 (see, e.g., [Lan11, 1.2]). Assume that \mathcal{E} is numerically flat and let X0X_{0} be some irreducible component of XX. By [dJ96] there exists a proper generically finite kk-morphism f:YX0f:Y\to X_{0} from a smooth projective kk-variety YY. Then ff^{*}\mathcal{E} is numerically flat, so it is strongly μH\mu_{H}-semistable for every ample divisor HH on XX. In particular, \mathcal{E} is strongly semistable. Now implication (1)(3)(1)\Rightarrow(3) follows from Proposition 3.3.

In proof of the implication (3)(5)(3)\Rightarrow(5) we use singular Riemann–Roch [Ful98, Chapter 18]. Let f:XY=Speckf:X\to Y=\operatorname{Spec}k be the structural morphism. We denote by [][\mathcal{E}] the class of \mathcal{E} in K0(X)K^{0}(X) and by [][\mathcal{F}] the class of \mathcal{F} in K0(X)K_{0}(X). By [Ful98, Theorem 18.3] we have canonical maps τX:K0(X)A(X)\tau_{X}:K_{0}(X)\to A_{*}(X)_{\mathbb{Q}} and τY:K0(Y)A(Y)=\tau_{Y}:K_{0}(Y)\to A_{*}(Y)_{\mathbb{Q}}=\mathbb{Q}, which satisfy the following equalities:

χ(X,)=τYf([][])=fτX([][])=f(ch([])τX([]))=rτYf([])=rχ(X,).\chi(X,\mathcal{E}\otimes\mathcal{F})=\tau_{Y}f_{*}([\mathcal{E}]\otimes[\mathcal{F}])=f_{*}\tau_{X}([\mathcal{E}]\otimes[\mathcal{F}])=f_{*}(\operatorname{ch}([\mathcal{E}])\cap\tau_{X}([\mathcal{F}]))=r\tau_{Y}f_{*}([\mathcal{F}])=r\chi(X,\mathcal{F}).

To prove that (5)(5) implies (1)(1) we can assume that XX is irreducible. Then by assumption there exist a proper generically finite kk-morphism f:YXf:Y\to X from a smooth projective kk-scheme YY to XX and an ample divisor HH on YY such that the bundle ff^{*}\mathcal{E} is strongly slope HH-semistable. By the Leray spectral sequence and the projection formula we have

χ(Y,f(mH))\displaystyle\chi(Y,f^{*}\mathcal{E}(mH)) =i(1)iχ(X,Rif(f(mH)))=i(1)iχ(X,Rif𝒪Y(mH))\displaystyle=\sum_{i}(-1)^{i}\chi(X,R^{i}f_{*}(f^{*}\mathcal{E}(mH)))=\sum_{i}(-1)^{i}\chi(X,\mathcal{E}\otimes R^{i}f_{*}\mathcal{O}_{Y}(mH))
=ri(1)iχ(X,Rif𝒪Y(mH))=rχ(Y,𝒪Y(mH)).\displaystyle=r\sum_{i}(-1)^{i}\chi(X,R^{i}f_{*}\mathcal{O}_{Y}(mH))=r\chi(Y,\mathcal{O}_{Y}(mH)).

So by the implication (4)(2)(4)\Rightarrow(2) of Theorem 4.8 we see that ff^{*}\mathcal{E} is numerically flat. Since ff is surjective, this implies that \mathcal{E} is also numerically flat.

The implication (3)(4)(3)\Rightarrow(4) is obvious, so it is sufficient to prove that (4)(1)(4)\Rightarrow(1). Without loss of generality we can assume that XX is irreducible and there exist a proper generically finite f:YXf:Y\to X from a smooth projective kk-variety YY and an ample divisor HH on YY such that the bundle ff^{*}\mathcal{E} is strongly μH\mu_{H}-semistable. Then c1(f)c_{1}(f^{*}\mathcal{E}) and c2(f)c_{2}(f^{*}\mathcal{E}) are numerically trivial. So by the implication (1)(2)(1)\Rightarrow(2) of Theorem 4.8 we see that ff^{*}\mathcal{E} is numerically flat. As before this implies that \mathcal{E} is also numerically flat. ∎

Corollary 3.14.

Let \mathcal{E} be a vector bundle on XX. Then the following conditions are equivalent:

  1. (1)

    \mathcal{E} is strongly semistable and Δ()\Delta(\mathcal{E}) is numerically trivial.

  2. (2)

    nd\mathcal{E}{\rm nd}\,\mathcal{E} is nef.

  3. (3)

    \mathcal{E} is universally semistable.

Proof.

If \mathcal{E} is strongly semistable and Δ()0\Delta(\mathcal{E})\equiv 0 then nd\mathcal{E}{\rm nd}\,\mathcal{E} is strongly semistable and both c1(nd)c_{1}(\mathcal{E}{\rm nd}\,\mathcal{E}) and c2(nd)c_{2}(\mathcal{E}{\rm nd}\,\mathcal{E}) are numerically trivial. So nd\mathcal{E}{\rm nd}\,\mathcal{E} is numerically flat, which proves (1)(2)(1)\Rightarrow(2). If nd\mathcal{E}{\rm nd}\,\mathcal{E} is nef then it is also numerically flat (as it isomorphic to its dual) and hence it is Nori semistable. This implies (3)(3). To prove that (3)(1)(3)\Rightarrow(1) it is sufficient to prove that Δ()0\Delta(\mathcal{E})\equiv 0. But nd\mathcal{E}{\rm nd}\,\mathcal{E} is universally semistable with trivial determinant, so it is Nori semistable. Hence it is numerically flat and Δ()=c2(nd)0\Delta(\mathcal{E})=c_{2}(\mathcal{E}{\rm nd}\,\mathcal{E})\equiv 0. ∎

4. Algebraic proofs of Theorems 1.1 and 1.2

4.1. Restriction theorems

In the sequel we frequently use the following strengthening of the Mehta–Ramanathan theorem for sheaves with vanishing discriminant. The result follows from [Lan04, Theorem 5.2] with a different proof from the Mehta–Ramanathan theorem.

Lemma 4.1.

Let XX be a smooth projective variety defined over an algebraically closed field kk and let HH be an ample divisor class on XX. Let \mathcal{E} be a torsion free sheaf with Δ()0\Delta(\mathcal{E})\equiv 0. Then there exists m0=m0(X,H,rk)1m_{0}=m_{0}(X,H,{\rm rk}\,\mathcal{E})\geq 1 such that for all mm0m\geq m_{0}

  1. (1)

    If \mathcal{E} is strongly μH\mu_{H}-stable and D\mathcal{E}_{D} is torsion free for some normal divisor D|mH|D\in|mH|, then D\mathcal{E}_{D} is strongly μHD\mu_{H_{D}}-stable.

  2. (2)

    If \mathcal{E} is strongly μH\mu_{H}-semistable, then for general D|mH|D\in|mH| the restriction D\mathcal{E}_{D} is strongly μHD\mu_{H_{D}}-semistable.

Proof.

In characteristic zero pick m0m_{0} such that |mH||mH| is basepoint free for all mm0m\geq m_{0}. In positive characteristic we also need to exceed a constant βr\beta_{r} depending on XX, HH, and rk{\rm rk}\,\mathcal{E}. See the inequality in [Lan04, Theorem 5.2]. (1) follows immediately from [Lan04, Theorem 5.2]. We obtain (2) as a consequence of (1) as in [Lan04, Corollary 5.4]. The factors (successive quotients) in any Jordan–Hölder filtration of \mathcal{E} are μH\mu_{H}-stable. As in [Lan04, Theorem 5.4] we observe that they also have numerically trivial discriminant. Their restriction to a general DD in a basepointfree |mH||mH| is again torsion free. In characteristic zero then (2) follows from (1). In positive characteristic strong μH\mu_{H}-semistability also takes into account the countably many Frobenius pullbacks. We remark that there exists some s0s_{0} such that the factors in a Jordan–Hölder filtration of (FXs0)(F_{X}^{s_{0}})^{*}\mathcal{E} are strongly μH\mu_{H}-stable. Then for a general divisor D|mH|D\in|mH| the restrictions of the factors in a Jordan–Hölder filtration of the sheaves {(FXs)}ss0\{(F_{X}^{s})^{*}\mathcal{E}\}_{s\leq s_{0}} to DD are torsion free. The restrictions of the factors in a Jordan–Hölder filtration of (FXs)(F_{X}^{s})^{*}\mathcal{E} for all ss0s\geq s_{0} to such a divisor are also torsion free since they are pullbacks of those of (FXs0)(F_{X}^{s_{0}})^{*}\mathcal{E}. ∎

4.2. The surface case

Proposition 4.2.

Let (X,H)(X,H) be an amply polarized smooth projective surface defined over an algebraically closed field kk. Let \mathcal{E} be a strongly μH\mu_{H}-semistable locally free sheaf of rank rr with ci()0c_{i}(\mathcal{E})\equiv 0 for i=1,2i=1,2. Then

  1. (1)

    For any line bundle LL on XX and any i{0,1,2}i\in\{0,1,2\} we have hi(X,LSymm)=O(mr1)h^{i}(X,L\otimes\operatorname{Sym}^{m}\mathcal{E})=O(m^{r-1}).

  2. (2)

    \mathcal{E} is numerically flat.

Proof.

Without loss of generality we can assume that HH is very ample. Since dim()=r+1\dim\mathbb{P}(\mathcal{E})=r+1, the claimed growth rate in (1)(1) is two degrees lower than expected. The proof is similar to the proof of the Bogomolov inequality in [HL10, Theorem 7.3.1].

By [RR84, Theorem 3.23 and the remark at the end of Section 4] the bundle Symm\operatorname{Sym}^{m}\mathcal{E} is strongly μH\mu_{H}-semistable. Since Δ(Symm)0\Delta(\operatorname{Sym}^{m}\mathcal{E})\equiv 0 Lemma 4.1 implies that if D|H|D\in|H| is a general divisor then (Symm)D(\operatorname{Sym}^{m}\mathcal{E})_{D} is strongly semistable of degree 0. By Lemma 2.6 we have

h0(D,(LSymm)D)(1+degLD)rkSymm=O(mr1)h^{0}(D,(L\otimes\operatorname{Sym}^{m}\mathcal{E})_{D})\leq(1+\deg L_{D})\cdot{\rm rk}\,\operatorname{Sym}^{m}\mathcal{E}=O(m^{r-1})

From the short exact sequence

0(LSymm)(H)LSymm(LSymm)D00\to(L\otimes\operatorname{Sym}^{m}\mathcal{E})(-H)\to L\otimes\operatorname{Sym}^{m}\mathcal{E}\to(L\otimes\operatorname{Sym}^{m}\mathcal{E})_{D}\to 0

we have

h0(X,LSymm)h0(D,(LSymm)D)+h0(X,(LSymm)(H)).h^{0}(X,L\otimes\operatorname{Sym}^{m}\mathcal{E})\leq h^{0}(D,(L\otimes\operatorname{Sym}^{m}\mathcal{E})_{D})+h^{0}(X,(L\otimes\operatorname{Sym}^{m}\mathcal{E})(-H)).

Changing HH by its multiple (which does not depend on mm) if necessary, we can assume that L(H)L(-H) has negative degree with respect to HH (e.g., we can assume that L(H)L^{\vee}(H) is effective). Then the bundle (LSymm)(H)(L\otimes\operatorname{Sym}^{m}\mathcal{E})(-H) is μH\mu_{H}-semistable with negative slope so it does not have any nonzero sections. Therefore h0(X,LSymm)=O(mr1)h^{0}(X,L\otimes\operatorname{Sym}^{m}\mathcal{E})=O(m^{r-1}).

By Serre’s duality, h2(X,LSymm)=h0(X,(Symm)ωXL)h^{2}(X,L\otimes\operatorname{Sym}^{m}\mathcal{E})=h^{0}(X,(\operatorname{Sym}^{m}\mathcal{E})^{\vee}\otimes\omega_{X}\otimes L^{\vee}). The bundle (Symm)(\operatorname{Sym}^{m}\mathcal{E})^{\vee} is also strongly semistable with numerically trivial Chern classes. An analogous proof to the case i=0i=0 gives h2(X,LSymm)=O(mr1)h^{2}(X,L\otimes\operatorname{Sym}^{m}\mathcal{E})=O(m^{r-1}). Note that in positive characteristic this equality does not follow formally from the previous case applied to \mathcal{E}^{\vee}.

Finally, to prove that h1(X,LSymm)h^{1}(X,L\otimes\operatorname{Sym}^{m}\mathcal{E}) grows at most like O(mr1)O(m^{r-1}), it is sufficient to prove that χ(X,LSymm)=O(mr1)\chi(X,L\otimes\operatorname{Sym}^{m}\mathcal{E})=O(m^{r-1}). Let π:()X\pi:\mathbb{P}(\mathcal{E})\to X be the bundle map and let us set ξc1(𝒪()(1))\xi\coloneqq c_{1}(\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1)). Since the Chern classes c1()c_{1}(\mathcal{E}) and c2()c_{2}(\mathcal{E}) are both numerically trivial, we have ξr=0\xi^{r}=0. By the Riemann–Roch theorem

χ(X,LSymm)=χ((),𝒪()(m)πL)=()exp(mξ+πc1(L))Td(),\chi(X,L\otimes\operatorname{Sym}^{m}\mathcal{E})=\chi(\mathbb{P}(\mathcal{E}),\mathcal{O}_{\mathbb{P}(\mathcal{E})}(m)\otimes\pi^{*}L)=\int_{\mathbb{P}(\mathcal{E})}\exp(m\xi+\pi^{*}c_{1}(L))\cap\operatorname{Td\,}\mathbb{P}(\mathcal{E}),

so the coefficients of mr+1m^{r+1} and mrm^{r} in the expression above are 0. The claim is proved.

(2)(2). Let f:CXf:C\to X be a morphism from a smooth projective curve and let CXC^{\prime}\subset X be the (possibly singular) image of ff. Consider the restriction sequence

0(Symm)(C)SymmSymmC0.0\to(\operatorname{Sym}^{m}\mathcal{E})(-C^{\prime})\to\operatorname{Sym}^{m}\mathcal{E}\to\operatorname{Sym}^{m}\mathcal{E}_{C^{\prime}}\to 0.

Let ξC=c1(𝒪(f)(1))\xi_{C}=c_{1}(\mathcal{O}_{\mathbb{P}(f^{*}\mathcal{E})}(1)). If ff^{*}\mathcal{E} is not strongly semistable, then since it has degree 0, it follows from Lemma 2.5 that ξC\xi_{C} is not nef. It is however big since some Frobenius pullback of ff^{*}\mathcal{E} has a strongly semistable subbundle of positive degree, so an ample subbundle. Bigness is invariant under birational pullback (cf. [Laz04a, Chapter 2.2]) and even under dominant generically finite pullback, hence h0(C,SymmC)=h0((C),𝒪(C)(m))h^{0}(C^{\prime},\operatorname{Sym}^{m}\mathcal{E}_{C^{\prime}})=h^{0}(\mathbb{P}(\mathcal{E}_{C^{\prime}}),\mathcal{O}_{\mathbb{P}(\mathcal{E}_{C^{\prime}})}(m)) grows like O(mr)O(m^{r}). However, we have

h0(C,SymmC)h0(X,Symm)+h1(X,(Symm)(C)).h^{0}(C^{\prime},\operatorname{Sym}^{m}\mathcal{E}_{C^{\prime}})\leq h^{0}(X,\operatorname{Sym}^{m}\mathcal{E})+h^{1}(X,(\operatorname{Sym}^{m}\mathcal{E})(-C^{\prime})).

We get a contradiction from part (1)(1). ∎

4.3. Local freeness via the vanishing of the discriminant

Lemma 4.3.

Let (X,H)(X,H) and (Y,A)(Y,A) be amply polarized smooth projective varieties defined over an algebraically closed field kk. Let \mathcal{E} be a μH\mu_{H}-semistable (μH\mu_{H}-stable, strongly μH\mu_{H}-semistable or strongly μH\mu_{H}-stable) torsion free sheaf on XX. Then prX{\rm pr}_{X}^{*}\mathcal{E} is μL\mu_{L}-semistable (respectively μL\mu_{L}-stable, strongly μL\mu_{L}-semistable or strongly μL\mu_{L}-stable) for L=prXH+prYAL={\rm pr}_{X}^{*}H+{\rm pr}_{Y}^{*}A.

Proof.

Let r=rkr={\rm rk}\,\mathcal{E} and denote n=dimXn=\dim X and m=dimYm=\dim Y. Let prX\mathcal{F}\subseteq{\rm pr}_{X}^{*}\mathcal{E} be a subsheaf of rank less than rr. For xXx\in X and yYy\in Y, let x{x}×Y\mathcal{F}_{x}\coloneqq\mathcal{F}_{\{x\}\times Y} and yX×{y}\mathcal{F}_{y}\coloneqq\mathcal{F}_{X\times\{y\}}. For general points xX(k)x\in X(k) and yY(k)y\in Y(k) we have x𝒪Yr\mathcal{F}_{x}\subseteq\mathcal{O}_{Y}^{\oplus r} and y\mathcal{F}_{y}\subseteq\mathcal{E}. Since \mathcal{E} is μH\mu_{H}-semistable and 𝒪Yr\mathcal{O}_{Y}^{\oplus r} is μA\mu_{A}-semistable, we deduce that c1()prXHn1prYAm(Am)rk=c1(y)Hn1rkμH()\frac{c_{1}(\mathcal{F})\cdot{\rm pr}_{X}^{*}H^{n-1}{\rm pr}_{Y}^{*}A^{m}}{(A^{m})\cdot{\rm rk}\,\mathcal{F}}=\frac{c_{1}(\mathcal{F}_{y})\cdot H^{n-1}}{{\rm rk}\,\mathcal{F}}\leq\mu_{H}(\mathcal{E}) and c1()prXHnprYAm1(Hn)rk=c1(x)Am1rk0\frac{c_{1}(\mathcal{F})\cdot{\rm pr}_{X}^{*}H^{n}{\rm pr}_{Y}^{*}A^{m-1}}{(H^{n})\cdot{\rm rk}\,\mathcal{F}}=\frac{c_{1}(\mathcal{F}_{x})\cdot A^{m-1}}{{\rm rk}\,\mathcal{F}}\leq 0. Then

μL()=c1()Ln+m1rk=c1()((n+m1n1)prXHn1prYAm+(n+m1n)prXHnprYAm1)rk(n+m1n1)(Am)μH()=(n+m1n1)(c1()Hn1)(Am)rk=μL(prX).\mu_{L}(\mathcal{F})=\frac{c_{1}(\mathcal{F})\cdot L^{n+m-1}}{{\rm rk}\,\mathcal{F}}=\frac{c_{1}(\mathcal{F})\cdot\left({{n+m-1}\choose{n-1}}{\rm pr}_{X}^{*}H^{n-1}{\rm pr}_{Y}^{*}A^{m}+{{n+m-1}\choose n}{\rm pr}_{X}^{*}H^{n}{\rm pr}_{Y}^{*}A^{m-1}\right)}{{\rm rk}\,\mathcal{F}}\\ \leq{{n+m-1}\choose{n-1}}(A^{m})\cdot\mu_{H}(\mathcal{E})=\frac{{{n+m-1}\choose{n-1}}\cdot(c_{1}(\mathcal{E})\cdot H^{n-1})(A^{m})}{{\rm rk}\,\mathcal{E}}=\mu_{L}({\rm pr}_{X}^{*}\mathcal{E}).

If \mathcal{E} is μH\mu_{H}-stable then c1()prXHn1prYAm(Am)rk=c1(y)Hn1rk<μH()\frac{c_{1}(\mathcal{F})\cdot{\rm pr}_{X}^{*}H^{n-1}{\rm pr}_{Y}^{*}A^{m}}{(A^{m})\cdot{\rm rk}\,\mathcal{F}}=\frac{c_{1}(\mathcal{F}_{y})\cdot H^{n-1}}{{\rm rk}\,\mathcal{F}}<\mu_{H}(\mathcal{E}) and we get μL()<μL(prX)\mu_{L}(\mathcal{F})<\mu_{L}({\rm pr}_{X}^{*}\mathcal{E}) as required.

Applying the above assertions for slope semistability and slope stability to all Frobenius pull-backs gives immediately the assertions for strong slope semistability and strong slope stability. ∎

Proposition 4.4.

Let XX be a smooth projective variety of dimension nn defined over an algebraically closed field kk and let HH be an ample polarization on XX. Let \mathcal{E} be a strongly μH\mu_{H}-stable torsion free sheaf on XX with c1()0c_{1}(\mathcal{E})\equiv 0. Then the following conditions are equivalent:

  1. (1)

    \mathcal{E} is reflexive and c2()Hn2=0c_{2}(\mathcal{E})\cdot H^{n-2}=0.

  2. (2)

    \mathcal{E} is locally free and numerically flat.

  3. (3)

    cj()0c_{j}(\mathcal{E})\equiv 0 for all j1j\geq 1.

  4. (4)

    The normalized Hilbert polynomial 1rkχ(X,(mH))\frac{1}{{\rm rk}\,\mathcal{E}}\cdot\chi(X,\mathcal{E}(mH)) of \mathcal{E} is equal to the Hilbert polynomial of 𝒪X\mathcal{O}_{X}.

Proof.

We argue by induction on nn. If n=1n=1, the equivalence of all four conditions is tautological, with (2)(2) being implied by Lemma 2.5. Let us assume that n=2n=2. Since every reflexive sheaf on a smooth surface is locally free, the implication (1)(2)(1)\Rightarrow(2) follows from Proposition 4.2. The implication (2)(3)(2)\Rightarrow(3) follows from Proposition 3.3 and (3)(4)(3)\Rightarrow(4) follows from the Hirzebruch–Riemann–Roch theorem. To prove (4)(1)(4)\Rightarrow(1), consider the exact sequence 0Q0,0\to\mathcal{E}\to\mathcal{E}^{\vee\vee}\to Q\to 0, where QQ has a finite support. Since c1()=c1()0c_{1}(\mathcal{E}^{\vee\vee})=c_{1}(\mathcal{E})\equiv 0 and \mathcal{E}^{\vee\vee} is strongly μH\mu_{H}-stable, the Bogomolov type inequality (see [Lan04, Theorem 3.2]) gives XΔ()0\int_{X}\Delta(\mathcal{E}^{\vee\vee})\geq 0. Our assumption on the Hilbert polynomial of \mathcal{E} and the Hirzebruch–Riemann–Roch theorem imply that Xc2(E)=0\int_{X}c_{2}(E)=0. Therefore with r=rkr={\rm rk}\,\mathcal{E}

XΔ()=2rXc2()=2rX(c2()+c2(Q))=2rXc2(Q)0,\int_{X}\Delta(\mathcal{E}^{\vee\vee})=2r\int_{X}c_{2}(\mathcal{E}^{\vee\vee})=2r\int_{X}(c_{2}(\mathcal{E})+c_{2}(Q))=2r\int_{X}c_{2}(Q)\leq 0,

which gives Xc2(Q)=0\int_{X}c_{2}(Q)=0. But then Q=0Q=0 and \mathcal{E} is reflexive.

Let us now assume that the result holds for varieties of dimension <n<n, where n3n\geq 3.

(1)(2)(1)\Rightarrow(2)

Let \mathcal{E} be reflexive, strongly μH\mu_{H}-stable, with c1()c_{1}(\mathcal{E}) numerically trivial and c2()Hn2=0c_{2}(\mathcal{E})\cdot H^{n-2}=0. Let ı:DX\imath:D\to X be the inclusion of a general member of |mH||mH| for sufficiently large mm. Then DD is smooth of dimension n1n-1. The restriction ı\imath^{*}\mathcal{E} is still reflexive by the general choice of DD and it is strongly μıH\mu_{\imath^{*}H}-stable by Lemma 4.1. We also have c2(ı)ıHn3=c2()Hn2c_{2}(\imath^{*}\mathcal{E})\cdot\imath^{*}H^{n-3}=c_{2}(\mathcal{E})\cdot H^{n-2} by Remark 2.3. Using the implication (1)(4)(1)\Rightarrow(4) on DD we see that

χ(X,(mH))χ(((m1)H))\displaystyle\chi(X,\mathcal{E}(mH))-\chi(\mathcal{E}((m-1)H)) =χ(D,ı(mH))=rχ(D,𝒪D(mH))\displaystyle=\chi(D,\imath^{*}\mathcal{E}(mH))=r\chi(D,\mathcal{O}_{D}(mH))
=r(χ(X,𝒪X(mH))χ(𝒪X((m1)H))).\displaystyle=r(\chi(X,\mathcal{O}_{X}(mH))-\chi(\mathcal{O}_{X}((m-1)H))).

Now assume that \mathcal{E} is not locally free. For sufficiently large mm, let ı:DX\imath:D\to X be an embedding of a member of |mH||mH| that is general among those that pass through one of the points where \mathcal{E} is not locally free. By [DH91] we know that DD is smooth. We also know that the restriction ı\imath^{*}\mathcal{E} is torsion-free. So Lemma 4.1 implies that ı\imath^{*}\mathcal{E} is strongly μıH\mu_{\imath^{*}H}-stable. By the above, we also know that

χ(D,ı(mH))\displaystyle\chi(D,\imath^{*}\mathcal{E}(mH)) =χ(X,(mH))χ(((m1)H))\displaystyle=\chi(X,\mathcal{E}(mH))-\chi(\mathcal{E}((m-1)H))
=r(χ(X,𝒪X(mH))χ(𝒪X((m1)H)))=rχ(D,𝒪D(mH)).\displaystyle=r(\chi(X,\mathcal{O}_{X}(mH))-\chi(\mathcal{O}_{X}((m-1)H)))=r\chi(D,\mathcal{O}_{D}(mH)).

Then the induction assumption implies that ı\imath^{*}\mathcal{E} is locally free. By [Lan19, Lemma 1.14] we deduce that \mathcal{E} is locally free around DD, a contradiction. Thus \mathcal{E} is locally free and we need to prove that it is numerically flat. Let f:CXf:C\to X be a morphism from a smooth projective curve. Let AA be any ample polarization on CC and let C~X×C\widetilde{C}\subset X\times C be the embedding of the graph of CC. Denote by prX:X×CX{\rm pr}_{X}:X\times C\to X and prC:X×CC{\rm pr}_{C}:X\times C\to C the projections onto the two factors. Let us also set LprXH+prCAL\coloneqq{\rm pr}_{X}^{*}H+{\rm pr}_{C}^{*}A and ~prX\widetilde{\mathcal{E}}\coloneqq{\rm pr}_{X}^{*}\mathcal{E}. Then C~C\widetilde{C}\simeq C and f~C~f^{*}\mathcal{E}\simeq\widetilde{\mathcal{E}}_{\widetilde{C}}, so it is sufficient to check that ~C~\widetilde{\mathcal{E}}_{\widetilde{C}} is semistable of degree 0. We have c2(~)Ln1=j=0n1(n1j)prX(c2(~)Hn1j)prCAjc_{2}(\widetilde{\mathcal{E}})\cdot L^{n-1}=\sum_{j=0}^{n-1}{{n-1}\choose j}{\rm pr}_{X}^{*}(c_{2}(\widetilde{\mathcal{E}})\cdot H^{n-1-j})\cdot{\rm pr}_{C}^{*}A^{j}. For dimension reasons, all terms except possibly j=1j=1 vanish. The term j=1j=1 also vanishes from the assumption c2()Hn2=0c_{2}(\mathcal{E})\cdot H^{n-2}=0. Therefore we have c1(~)0c_{1}(\widetilde{\mathcal{E}})\equiv 0 and c2(~)Ln1=0c_{2}(\widetilde{\mathcal{E}})\cdot L^{n-1}=0. By Lemma 4.3 we also know that ~\widetilde{\mathcal{E}} is strongly μL\mu_{L}-stable. By [DH91, Theorem 3.1] for large mm there exists a chain of smooth varieties S=X2X3Xn=X×CS=X_{2}\subset X_{3}\subset...\subset X_{n}=X\times C containing C~\widetilde{C} such that all Xi|mLXi+1|X_{i}\in|mL_{X_{i+1}}| are smooth. Then Lemma 4.1 implies that ~S\widetilde{\mathcal{E}}_{S} is strongly μLS\mu_{L_{S}}-stable with c1(~S)0c_{1}(\widetilde{\mathcal{E}}_{S})\equiv 0 and Sc2(~S)=0\int_{S}c_{2}(\widetilde{\mathcal{E}}_{S})=0. So the required assertion follows from Proposition 4.2.

(2)(3)(2)\Rightarrow(3) follows from Proposition 3.3.

(3)(4)(3)\Rightarrow(4) follows from the Hirzebruch–Riemann–Roch theorem.

(4)(1)(4)\Rightarrow(1) Let \mathcal{E} be a rank rr torsion free, strongly μH\mu_{H}-stable sheaf with c1()0c_{1}(\mathcal{E})\equiv 0 and χ(X,(mH))=rχ(X,𝒪X(mH))\chi(X,\mathcal{E}(mH))=r\chi(X,\mathcal{O}_{X}(mH)) for all mm\in\mathbb{Z}. Comparing coefficients of these polynomials at mn2m^{n-2} we see that c2()Hn2=0c_{2}(\mathcal{E})\cdot H^{n-2}=0. Consider the exact sequence

0Q0,0\to\mathcal{E}\to\mathcal{E}^{\vee\vee}\to Q\to 0,

where QQ is a torsion sheaf on XX. Then \mathcal{E}^{\vee\vee} is reflexive, strongly μH\mu_{H}-stable sheaf with c1()0c_{1}(\mathcal{E}^{\vee\vee})\equiv 0. As in the surface case, the Bogomolov type inequality (see [Lan04, Theorem 3.2]) gives

Δ()Hn2=2rc2(Q)Hn20,\Delta(\mathcal{E}^{\vee\vee})\cdot H^{n-2}=2rc_{2}(Q)\cdot H^{n-2}\geq 0,

which implies Δ()Hn2=c2(Q)Hn2=0\Delta(\mathcal{E}^{\vee\vee})\cdot H^{n-2}=c_{2}(Q)\cdot H^{n-2}=0. Using already proven implication (1)(4)(1)\Rightarrow(4) we see that Hilbert polynomials of \mathcal{E} and \mathcal{E}^{\vee\vee} coincide. Therefore the Hilbert polynomial of QQ vanishes. This implies that Q=0Q=0 and hence \mathcal{E} is reflexive with c2()Hn2=0c_{2}(\mathcal{E})\cdot H^{n-2}=0. ∎

The following known example shows that reflexivity assumption is necessary in condition (1)(1) of Proposition 4.4.

Example 4.5.

Let XX be a smooth projective variety of dimension n3n\geq 3. Let 𝒪X\mathcal{I}\subset\mathcal{O}_{X} be the ideal sheaf of a nonempty closed subset of codimension j3j\geq 3. Then \mathcal{I} is torsion-free, strongly slope semistable with respect to any polarization, and c1()c_{1}(\mathcal{I}) and c2()c_{2}(\mathcal{I}) are numerically trivial. However cj()c_{j}(\mathcal{I}) is not numerically trivial and of course \mathcal{I} is not locally free.

Corollary 4.6.

Let \mathcal{E} be a reflexive strongly μH\mu_{H}-semistable sheaf with c1()0c_{1}(\mathcal{E})\equiv 0 and c2()Hn2=0c_{2}(\mathcal{E})\cdot H^{n-2}=0. Then \mathcal{E} is locally free and numerically flat. Moreover, every factor in a Jordan–Hölder filtration of \mathcal{E} is also locally free and numerically flat.

This proof is analogous to [Nak99, Proposition 2.5] and [Lan19, Section 2.2].

Proof.

The statement is clear if n=1n=1 so we assume that n2n\geq 2. We perform induction on the rank rr of \mathcal{E}. The case r=1r=1 or more generally \mathcal{E} is strongly μH\mu_{H}-stable is Proposition 4.4. Thus we may assume that \mathcal{E} is not strongly μH\mu_{H}-stable. First, let us assume that there exists an exact sequence

01200\to\mathcal{E}_{1}\to\mathcal{E}\to\mathcal{E}_{2}\to 0

where 1\mathcal{E}_{1} is μH\mu_{H}-stable, strongly μH\mu_{H}-semistable and reflexive and 2\mathcal{E}_{2} is strongly μH\mu_{H}-semistable and torsion free, both 1\mathcal{E}_{1} and 2\mathcal{E}_{2} are nonzero sheaves and μH(1)=μH(2)=0\mu_{H}(\mathcal{E}_{1})=\mu_{H}(\mathcal{E}_{2})=0. By the Hodge index theorem we have

0\displaystyle 0 =Δ()Hn2r=Δ(1)Hn2r1+Δ(2)Hn2r2r1r2r(c1(1)r1c1(2)r2)2Hn2\displaystyle=\frac{\Delta(\mathcal{E})\cdot H^{n-2}}{r}=\frac{\Delta(\mathcal{E}_{1})\cdot H^{n-2}}{r_{1}}+\frac{\Delta(\mathcal{E}_{2})\cdot H^{n-2}}{r_{2}}-\frac{r_{1}r_{2}}{r}\left(\frac{c_{1}(\mathcal{E}_{1})}{r_{1}}-\frac{c_{1}(\mathcal{E}_{2})}{r_{2}}\right)^{2}\cdot H^{n-2}
Δ(1)Hn2r1+Δ(2)Hn2r2.\displaystyle\geq\frac{\Delta(\mathcal{E}_{1})\cdot H^{n-2}}{r_{1}}+\frac{\Delta(\mathcal{E}_{2})\cdot H^{n-2}}{r_{2}}.

Using the Bogomolov type inequality [Lan04, Theorem 3.2] for 1\mathcal{E}_{1} and 2\mathcal{E}_{2}, we see that Δ(1)Hn2=Δ(2)Hn2=0\Delta(\mathcal{E}_{1})\cdot H^{n-2}=\Delta(\mathcal{E}_{2})\cdot H^{n-2}=0. Equality in the Hodge index inequality implies also that c1(1)c_{1}(\mathcal{E}_{1}) and c1(2)c_{1}(\mathcal{E}_{2}) are numerically trivial. The induction hypothesis directly applies only to the reflexive sheaf 1\mathcal{E}_{1} (not the torsion free 2\mathcal{E}_{2}) and we deduce that it is locally free and numerically flat. But we also have an exact sequence

022Q0,0\to\mathcal{E}_{2}\to{\mathcal{E}_{2}}^{\vee\vee}\to Q\to 0,

where QQ is supported in codimension at least 22. Since 2{\mathcal{E}_{2}}^{\vee\vee} is also strongly μH\mu_{H}-semistable, Δ(2)Hn20\Delta({\mathcal{E}_{2}}^{\vee\vee})\cdot H^{n-2}\geq 0. But Δ(2)Hn2=2rc2(Q)Hn20\Delta({\mathcal{E}_{2}}^{\vee\vee})\cdot H^{n-2}=2rc_{2}(Q)\cdot H^{n-2}\leq 0, so Δ(2)Hn2=0\Delta({\mathcal{E}_{2}}^{\vee\vee})\cdot H^{n-2}=0 and QQ is supported in codimension at least 33. By the induction hypothesis, 2{\mathcal{E}_{2}}^{\vee\vee} is locally free and every factor in a Jordan–Hölder filtration of 2{\mathcal{E}_{2}}^{\vee\vee} is also locally free and numerically flat. An Ext computation using that QQ has codimension at least 33 shows that we have a commutative diagram

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\textstyle{\mathcal{E}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\mathcal{E}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2\textstyle{\mathcal{E}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\textstyle{\mathcal{E}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\mathcal{E}^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2\textstyle{{\mathcal{E}_{2}}^{\vee\vee}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

for some sheaf \mathcal{E}^{\prime} that is then necessarily locally free. See [Nak99, Proposition 2.5] or [Lan19, Lemma 1.12] for details. The middle vertical arrow is an isomorphism since \mathcal{E} and \mathcal{E}^{\prime} are both reflexive, and isomorphic on the locally free locus of 2\mathcal{E}_{2}. This implies that \mathcal{E} is locally free and 2\mathcal{E}_{2} is reflexive, so we can apply the induction assumption also to 2\mathcal{E}_{2}.

To finish the proof one needs to deal with the case when \mathcal{E} is μH\mu_{H}-stable but not strongly μH\mu_{H}-stable. Then we can apply the above arguments for some Frobenius pull-back (FXm)(F_{X}^{m})^{*}\mathcal{E}. Since local freeness and numerical flatness of (FXm)(F_{X}^{m})^{*}\mathcal{E} implies local freeness and numerical flatness of \mathcal{E}, we get the required assertion. ∎

Corollary 4.7.

Let \mathcal{E} be a reflexive strongly μH\mu_{H}-semistable sheaf of rank rr with Δ()Hn2=0\Delta(\mathcal{E})\cdot H^{n-2}=0. Then \mathcal{E} is locally free and nd\operatorname{{\mathcal{E}}nd\,}\mathcal{E} is numerically flat. Moreover, every factor in a Jordan–Hölder filtration of \mathcal{E} is also locally free and its endomorphism bundle is numerically flat.

Proof.

We use a finite cover to extract an rr-th root of det\det\mathcal{E} and reduce to the case c1()0c_{1}(\mathcal{E})\equiv 0 where Corollary 4.6 applies. Let φ:YX\varphi:Y\to X be a Bloch–Gieseker cover (see [BG71, Lemma 2.1]), i.e., a finite surjective map from a smooth projective variety YY such that φdet=r\varphi^{*}\det\mathcal{E}=\mathcal{L}^{\otimes r} for some line bundle \mathcal{L} on YY. The map φ\varphi is flat so φ\varphi^{*}\mathcal{E} is reflexive. By Remark 2.3 we also have ci(φ)=φci()c_{i}(\varphi^{*}\mathcal{E})=\varphi^{*}c_{i}(\mathcal{E}) for all ii. In particular, we have det(φ)=𝒪Y\det(\varphi^{*}\mathcal{E}\otimes\mathcal{L}^{\vee})=\mathcal{O}_{Y} and Δ(φ)φHn2=0\Delta(\varphi^{*}\mathcal{E})\cdot\varphi^{*}H^{n-2}=0. Furthermore, φH\varphi^{*}H is ample, and φ\varphi^{*}\mathcal{E} is strongly μφH\mu_{\varphi^{*}H}-semistable. Then φ1\varphi^{*}\mathcal{E}\otimes\mathcal{L}^{-1} is reflexive, strongly μϕH\mu_{\phi^{*}H}-semistable, with trivial determinant, and Δ(φ1)φHn2=Δ(φ)φHn2=0\Delta(\varphi^{*}\mathcal{E}\otimes\mathcal{L}^{-1})\cdot\varphi^{*}H^{n-2}=\Delta(\varphi^{*}\mathcal{E})\cdot\varphi^{*}H^{n-2}=0. From the results above we deduce that φ1\varphi^{*}\mathcal{E}\otimes\mathcal{L}^{-1} is locally free and numerically flat. Hence φnd=nd(φ1)\varphi^{*}\operatorname{{\mathcal{E}}nd\,}\mathcal{E}=\operatorname{{\mathcal{E}}nd\,}(\varphi^{*}\mathcal{E}\otimes\mathcal{L}^{-1}) is also numerically flat. This implies that \mathcal{E} is locally free and nd\operatorname{{\mathcal{E}}nd\,}\mathcal{E} is numerically flat. The last part follows analogously. The only difference is that the pull-back of μH\mu_{H}-stable sheaf need not be μφH\mu_{\varphi^{*}H}-stable and it is only μφH\mu_{\varphi^{*}H}-semistable. So one needs to take a refinement of the pull-back of a Jordan–Hölder filtration of \mathcal{E} to a Jordan–Hölder filtration of φ\varphi^{*}\mathcal{E} and then use Corollary 4.6. ∎

4.4. Main theorems in the smooth case

Theorem 4.8.

Let XX be a smooth projective variety of dimension nn defined over an algebraically closed field kk and let HH be an ample polarization on XX. Let \mathcal{E} be a torsion free sheaf on XX. Then the following conditions are equivalent:

  1. (1)

    \mathcal{E} is reflexive, strongly μH\mu_{H}-semistable and ch1()Hn1=ch2()Hn2=0\operatorname{ch}_{1}(\mathcal{E})\cdot H^{n-1}=\operatorname{ch}_{2}(\mathcal{E})\cdot H^{n-2}=0.

  2. (2)

    \mathcal{E} is locally free and numerically flat.

  3. (3)

    \mathcal{E} is strongly μH\mu_{H}-semistable and cj()0c_{j}(\mathcal{E})\equiv 0 for all j1j\geq 1.

  4. (4)

    \mathcal{E} is strongly μH\mu_{H}-semistable and the normalized Hilbert polynomial of \mathcal{E} equals to the Hilbert polynomial of 𝒪X\mathcal{O}_{X}.

In particular, if \mathcal{E} is a strongly slope semistable vector bundle on XX with c1()0c_{1}(\mathcal{E})\equiv 0, then \mathcal{E} is nef if and only if Δ()0\Delta(\mathcal{E})\equiv 0.

Proof.

The conditions ch1()Hn1=0\operatorname{ch}_{1}(\mathcal{E})\cdot H^{n-1}=0 and ch2()Hn2=0\operatorname{ch}_{2}(\mathcal{E})\cdot H^{n-2}=0 imply that c1()Hn2c_{1}(\mathcal{E})\cdot H^{n-2} is numerically trivial and Δ()Hn2=0\Delta(\mathcal{E})\cdot H^{n-2}=0 by the Bogomolov inequality and by the Hodge index theorem on surfaces (see [Lan11, Lemma 4.2]). The condition c1()Hn2=0c_{1}(\mathcal{E})\cdot H^{n-2}=0 implies that c1()c_{1}(\mathcal{E}) is numerically trivial by [Ful98, Example 19.3.3]. Therefore (1)(2)(1)\Rightarrow(2) follows from Corollary 4.6. The implication (2)(3)(2)\Rightarrow(3) follows from Proposition 3.3. The proofs of implications (3)(4)(3)\Rightarrow(4) and (4)(1)(4)\Rightarrow(1) are analogous to the proofs of the corresponding implications in Proposition 4.4.

If \mathcal{E} is a nef vector bundle with c1()0c_{1}(\mathcal{E})\equiv 0, then \mathcal{E} is numerically flat and in particular Δ()0\Delta(\mathcal{E})\equiv 0. Conversely, if \mathcal{E} is a strongly semistable vector bundle with c1()0c_{1}(\mathcal{E})\equiv 0, and Δ()0\Delta(\mathcal{E})\equiv 0, then ch2()0\operatorname{ch}_{2}(\mathcal{E})\equiv 0 and so \mathcal{E} is numerically flat. ∎

Even in the locally free case, one cannot replace condition (3)(3) in Theorem 4.8 with cj()Hnj=0c_{j}(\mathcal{E})\cdot H^{n-j}=0 for all j1j\geq 1, or the condition ch2()Hn2=0\operatorname{ch}_{2}(\mathcal{E})\cdot H^{n-2}=0 in (1) with c2()Hn2=0c_{2}(\mathcal{E})\cdot H^{n-2}=0.

Example 4.9.

Let XX be a smooth projective surface of Picard rank at least 3. Let H,L,LH,L,L^{\prime} be divisors on XX with HH ample such that the intersection pairing on span(H,L,L)N1(X){\rm span}(H,L,L^{\prime})\subseteq N^{1}(X) has diagonal matrix with respect to the basis (H,L,L)(H,L,L^{\prime}). Let =𝒪X(L)𝒪X(L)\mathcal{E}=\mathcal{O}_{X}(L)\oplus\mathcal{O}_{X}(L^{\prime}). It is strongly μH\mu_{H}-semistable. Furthermore, we have c1()H=0c_{1}(\mathcal{E})\cdot H=0 and Xc2()=0\int_{X}c_{2}(\mathcal{E})=0. However, \mathcal{E} is not numerically flat since c1()=L+Lc_{1}(\mathcal{E})=L+L^{\prime} is not numerically trivial. Note that in this case Xch2()<0\int_{X}\operatorname{ch}_{2}(\mathcal{E})<0.

Theorem 4.10.

Let XX be a smooth projective variety of dimension nn defined over an algebraically closed field kk and let HH be an ample polarization on XX. Let \mathcal{E} be a reflexive sheaf of rank rr on XX. Then the following conditions are equivalent:

  1. (1)

    \mathcal{E} is strongly μH\mu_{H}-semistable and Δ()\Delta(\mathcal{E}) is numerically trivial.

  2. (2)

    \mathcal{E} is strongly μH\mu_{H}-semistable and Δ()Hn2=0\Delta(\mathcal{E})\cdot H^{n-2}=0.

  3. (3)

    \mathcal{E} is locally free and the twisted normalized bundle 1rdet\mathcal{E}\langle-\frac{1}{r}\det\mathcal{E}\rangle is nef.

  4. (4)

    \mathcal{E} is locally free and nd\operatorname{{\mathcal{E}}nd\,}\mathcal{E} is nef.

  5. (5)

    For every morphism f:CXf:C\to X from a smooth projective curve, ff^{*}\mathcal{E} is semistable.

  6. (6)

    For every morphism f:YXf:Y\to X, where YY is a smooth projective variety, ff^{*}\mathcal{E} is strongly slope semistable with respect to any ample polarization on YY.

In particular, if \mathcal{E} is strongly μH\mu_{H}-semistable and Δ()Hn2=0\Delta(\mathcal{E})\cdot H^{n-2}=0, then \mathcal{E} is locally free. Furthermore, it is nef (resp. ample) if and only if det\det\mathcal{E} is nef (resp. ample).

Proof.

(1)(2)(1)\Rightarrow(2) is trivial. For locally free \mathcal{E}, the nefness of 1rdet\mathcal{E}\langle-\frac{1}{r}\det\mathcal{E}\rangle is equivalent to the nefness of φ\varphi^{*}\mathcal{E}\otimes\mathcal{L}^{\vee}, where φ:YX\varphi:Y\to X is a finite cover as in Corollary 4.7. Since this bundle has trivial determinant, its nefness is equivalent to the nefness (equivalently numerical flatness) of ndφ\mathcal{E}{\rm nd}\,\varphi^{*}\mathcal{E} and then to that of nd\mathcal{E}{\rm nd}\,\mathcal{E}. We get the implications (2)(3)(4)(2)\Rightarrow(3)\Leftrightarrow(4) by Corollary 4.7. We also have (4)(1)(4)\Rightarrow(1) by Proposition 3.3.

Numerically flat bundles are universally slope semistable, in fact universally strongly slope semistable. We obtain (4)(6)(5)(4)\Rightarrow(6)\Rightarrow(5). Then non-torsion semistable sheaves on smooth projective curves are torsion free, in particular locally free. General complete intersection curves of high degree passing through a given point xXx\in X are smooth by [DH91]. Assuming (5)(5), we obtain that \mathcal{E} is locally free. By precomposing f:CXf:C\to X with iterates of the Frobenius FCF_{C}, we see that (5)(5) is equivalent to the analogous statement for strong semistability. On CC, the strong semistability of ff^{*}\mathcal{E} is equivalent to the nefness of fndf^{*}\mathcal{E}{\rm nd}\,\mathcal{E}. We deduce that (5)(4)(5)\Rightarrow(4).

For the last statements, if \mathcal{E} is strongly μH\mu_{H}-semistable with Δ()Hn2=0\Delta(\mathcal{E})\cdot H^{n-2}=0, then \mathcal{E} is locally free by Corollary 4.7. Clearly if \mathcal{E} is nef (resp. ample), then det\det\mathcal{E} is nef (resp. ample). The implication (2)(3)(2)\Rightarrow(3) and the identity =(1rdet)1rdet\mathcal{E}=\bigl{(}\mathcal{E}\langle-\frac{1}{r}\det\mathcal{E}\rangle\bigr{)}\langle\frac{1}{r}\det\mathcal{E}\rangle give the converse. ∎

Remark 4.11.

In the above theorem one can also give another condition analogous to (3)(3) of Theorem 4.8. See [Lan19, Theorem 2.2] for the precise formulation. We leave an easy proof of this result along the above lines to the interested reader.

Remark 4.12.

If XX is a complex projective manifold then homological equivalence over \mathbb{Q} implies numerical equivalence. Classically they are known to agree for divisors. Lieberman [Lie68] also proved it for codimension 2 cycles using the hard Lefschetz theorem. So in this case proving that Δ()\Delta(\mathcal{E}) is 0 in H4(X,)H^{4}(X,\mathbb{Q}) is equivalent to proving that it is numerically trivial.

5. On Misra’s question

Definition 5.1.

Let XX be a projective variety defined over an algebraically closed field kk. We say that XX is 11-homogeneous if Nef(X)=Eff¯(X)\operatorname{{Nef}}(X)=\operatorname{\overline{Eff}}(X).

Curves, or more generally varieties of Picard rank 1, and homogeneous spaces are 11-homogeneous. By [Miy87, Theorem 3.1 and remark on p. 464] if \mathcal{E} is a vector bundle on a smooth projective curve then ()\mathbb{P}(\mathcal{E}) is 11-homogeneous if and only if \mathcal{E} is strongly semistable.

Remark 5.2.

  1. (i)

    If XX is a projective variety with Picard number 2 and AA and BB are globally generated line bundles, but not big, and not proportional in N1(X)N^{1}(X), then XX is 1-homogeneous and AA and BB span the boundary rays of Nef(X)=Eff¯(X)\operatorname{{Nef}}(X)=\operatorname{\overline{Eff}}(X).

  2. (ii)

    If XX is projective with Picard number 1 and dimension nn, and \mathcal{E} is a vector bundle of rank rr on XX such that \mathcal{E} can be generated by fewer than n+rn+r global sections, then ()\mathbb{P}(\mathcal{E}) is 11-homogeneous. (We get an induced morphism f:()Nf:\mathbb{P}(\mathcal{E})\to\mathbb{P}^{N} for some N<dim()N<\dim\mathbb{P}(\mathcal{E}). In particular r2r\geq 2 and ()\mathbb{P}(\mathcal{E}) has Picard rank 2. The fibers of π:()X\pi:\mathbb{P}(\mathcal{E})\to X are embedded by ff. Thus if HH is a very ample line bundle on XX, then πH\pi^{*}H and f𝒪N(1)f^{*}\mathcal{O}_{\mathbb{P}^{N}}(1) satisfy the requirements of (i).)

5.1. Misra’s theorem in an arbitrary characteristic

The following theorem generalizes [Mis21, Theorem 1.2] to an arbitrary characteristic:

Theorem 5.3.

Let XX be a smooth projective variety defined over an algebraically closed field kk and let \mathcal{E} be a strongly slope semistable bundle with respect to some ample polarization of XX. Let us also assume that Δ()0\Delta(\mathcal{E})\equiv 0. Then the following conditions are equivalent:

  1. (1)

    XX is 11-homogeneous,

  2. (2)

    c1(π𝒪(E)(D))c_{1}(\pi_{*}{\mathcal{O}}_{\mathbb{P}(E)}(D)) is nef for every effective divisor DD on ()\mathbb{P}(\mathcal{E}),

  3. (3)

    ()\mathbb{P}(\mathcal{E}) is 11-homogeneous.

Proof.

We specify the adjustments needed to port the proof of [Mis21, Theorem 1.2] to positive characteristic. Once can use Theorem 4.10 instead of [Mis21, Theorem 2.1]. Apart from that the implication (1)(2)(1)\Rightarrow(2) uses the fact that symmetric powers of \mathcal{E} are semistable. This follows either from the Ramanan–Ramanathan theorem (see the proof of Proposition 4.2) or one can use Theorem 4.10 and the fact that symmetric powers of nef bundles are nef.

Finally, the proof of (1)(2)(1)\Rightarrow(2) uses the duality between the cone of strongly movable curves and pseudo-effective divisors. This fact also holds in positive characteristic by [FL17b, Theorem 2.22] (see also [Das20, Theorem 1.4]). The rest of the proof is the same as in [Mis21]. ∎

5.2. Syzygy bundle counterexamples to Question 1.4

Definition 5.4.

Let XX be a projective variety. Let 𝒱\mathcal{V} be a globally generated vector bundle on XX. The associated syzygy bundle M𝒱M_{\mathcal{V}} is the kernel of the natural evaluation morphism H0(X,𝒱)𝒪Xev𝒱H^{0}(X,\mathcal{V})\otimes\mathcal{O}_{X}\overset{{\rm ev}}{\twoheadrightarrow}\mathcal{V}. Put E𝒱=M𝒱E_{\mathcal{V}}=M_{\mathcal{V}}^{\vee}. This is a globally generated vector bundle.

For example the Euler sequence on n\mathbb{P}^{n} gives M𝒪n(1)=Ωn(1)M_{\mathcal{O}_{\mathbb{P}^{n}}(1)}=\Omega_{\mathbb{P}^{n}}(1). If \mathcal{L} is a globally generated line bundle, and φ:XN\varphi:X\to\mathbb{P}^{N} is the induced morphism with φ𝒪N(1)=\varphi^{*}\mathcal{O}_{\mathbb{P}^{N}}(1)=\mathcal{L}, then M=φΩN(1)M_{\mathcal{L}}=\varphi^{*}\Omega_{\mathbb{P}^{N}}(1) and E=φTN(1)E_{\mathcal{L}}=\varphi^{*}T_{\mathbb{P}^{N}}(-1).

Remark 5.5.

If 𝒱\mathcal{V} is a globally generated vector bundle on XX with dimX>rk𝒱\dim X>{\rm rk}\,\mathcal{V}, then r=rkE𝒱=h0(X,𝒱)rk𝒱r={\rm rk}\,E_{\mathcal{V}}=h^{0}(X,\mathcal{V})-{\rm rk}\,\mathcal{V} and E𝒱E_{\mathcal{V}} is generated by h0(X,𝒱)=r+rk𝒱<dimX+rh^{0}(X,\mathcal{V})=r+{\rm rk}\,\mathcal{V}<\dim X+r global sections. In particular, if XX has Picard rank 1 then (E𝒱)\mathbb{P}(E_{\mathcal{V}}) is 11-homogeneous by Remark 5.2.

Proposition 5.6.

For n2n\geq 2 we have that

  1. (i)

    (Tn)\mathbb{P}(T_{\mathbb{P}^{n}}) is 11-homogeneous.

  2. (ii)

    TnT_{\mathbb{P}^{n}} is strongly slope semistable with respect to the hyperplane class.

  3. (iii)

    Δ(Tn)0\Delta(T_{\mathbb{P}^{n}})\neq 0.

  4. (iv)

    The restriction of TnT_{\mathbb{P}^{n}} to every line in n\mathbb{P}^{n} is unstable.

Proof.

(i)(i). Apply Remark 5.5 to 𝒱=𝒪n(1)\mathcal{V}=\mathcal{O}_{\mathbb{P}^{n}}(1). (ii)(ii) is classical and it follows, e.g., from the Bott vanishing. (iii)(iii). By direct computation, Δ(Tn)=Δ(Tn(1))=n+1\Delta(T_{\mathbb{P}^{n}})=\Delta(T_{\mathbb{P}^{n}}(-1))=n+1. (iv)(iv). TnT_{\mathbb{P}^{n}} restricts as 𝒪(2)𝒪(1)n1\mathcal{O}(2)\oplus\mathcal{O}(1)^{\oplus n-1} on every line. ∎

The easy counterexample above is slope semistable. We also give a slope unstable example inspired in part by suggestions of S. Misra and D. S. Nagaraj.

Example 5.7.

On 3\mathbb{P}^{3} consider the globally generated bundle 𝒱=𝒪3(1)𝒪3(2)\mathcal{V}=\mathcal{O}_{\mathbb{P}^{3}}(1)\oplus\mathcal{O}_{\mathbb{P}^{3}}(2). Consider the associated syzygy bundle M𝒱M_{\mathcal{V}} and let =E𝒱=M𝒱\mathcal{E}=E_{\mathcal{V}}=M_{\mathcal{V}}^{\vee}. Then \mathcal{E} is slope unstable, has positive discriminant, and ()\mathbb{P}(\mathcal{E}) is 1-homogeneous. The bundle \mathcal{E} has rank 12, c1()=3c_{1}(\mathcal{E})=3 and c2()=7c_{2}(\mathcal{E})=7. It is an immediate computation that Δ()>0\Delta(\mathcal{E})>0. We have that M𝒱=M𝒪3(1)M𝒪3(2)M_{\mathcal{V}}=M_{\mathcal{O}_{\mathbb{P}^{3}}(1)}\oplus M_{\mathcal{O}_{\mathbb{P}^{3}}(2)}. The summands have slopes 1/3-1/3 and respectively 2/9-2/9. Thus M𝒱M_{\mathcal{V}} and its dual \mathcal{E} are unstable. Since rk𝒱=2<dim3{\rm rk}\,\mathcal{V}=2<\dim\mathbb{P}^{3}, we get that ()\mathbb{P}(\mathcal{E}) is 1-homogeneous by Remark 5.5.∎

We list related problems asking if our counterexamples are the simplest/smallest possible.

Question 5.8.
  1. (1)

    Does there exist a complex projective manifold XX of Picard rank 1 and dimension at least 2 supporting an ample and globally generated line bundle LL such that the syzygy bundle MLM_{L} is μL\mu_{L}-unstable, but (ML)\mathbb{P}(M_{L}^{\vee}) is 11-homogeneous? 111[Sch05] constructs an example on curves. The semistability of syzygy bundles is an active topic of research. We refer to [EL92, ELM13, BPMGNO19] and the references therein for a history of the problem.

  2. (2)

    Does there exist a complex projective surface XX supporting a slope unstable \mathcal{E} such that ()\mathbb{P}(\mathcal{E}) is 11-homogeneous?

  3. (3)

    Are there any μH\mu_{H}-unstable bundles \mathcal{E} with Δ()Hn2=0\Delta(\mathcal{E})\cdot H^{n-2}=0 such that ()\mathbb{P}(\mathcal{E}) is 11-homogeneous?

5.3. A positive result

Lemma 5.9.

Let VV be a free module of rank rr over a commutative ring kk. Then for any a,b1a,b\geq 1 there exist

  1. (1)

    a surjection of GL(V){\rm GL}\,(V)-modules

    Syma(SymbV)SymabV,\operatorname{Sym}^{a}(\operatorname{Sym}^{b}V)\twoheadrightarrow\operatorname{Sym}^{ab}V,
  2. (2)

    an inclusion of GL(V){\rm GL}\,(V)-modules

    (rV)2aSymr(Sym2aV).(\bigwedge^{r}V)^{\otimes 2a}\hookrightarrow\operatorname{Sym}^{r}(\operatorname{Sym}^{2a}V).

Moreover, the composition

(rV)2aSymr(Sym2aV)Sym2raV(\bigwedge^{r}V)^{\otimes 2a}\hookrightarrow\operatorname{Sym}^{r}(\operatorname{Sym}^{2a}V)\to\operatorname{Sym}^{2ra}V

is zero.

Over \mathbb{C}, assertion (2)(2) is a particular case of the main result of [BCI11] which also applies to other even partitions than λ=((2a)r)\lambda=((2a)^{r}).

Proof.

We have a canonical surjection i=1aSymbVSymabV\bigotimes_{i=1}^{a}\operatorname{Sym}^{b}V\twoheadrightarrow\operatorname{Sym}^{ab}V coming from the symmetric multiplication. By definition we also have a canonical surjection i=1aSymbVSyma(SymbV)\bigotimes_{i=1}^{a}\operatorname{Sym}^{b}V\twoheadrightarrow\operatorname{Sym}^{a}(\operatorname{Sym}^{b}V). Using the universal property of the symmetric product we get an induced map Syma(SymbV)SymabV\operatorname{Sym}^{a}(\operatorname{Sym}^{b}V)\to\operatorname{Sym}^{ab}V, which is also surjective. This gives the first assertion.

To prove the second assertion we reduce to the case k=k=\mathbb{Z}. For k=k=\mathbb{Z} we construct an explicit non-zero map of GL(V){\rm GL}\,(V)-modules that has an associated matrix with an entry equal to 11. The map is then a split inclusion as a morphism of \mathbb{Z}-modules, hence base changing to any commutative ring kk is still injective.

Let λ=((2a)r)=(2a,,2a)\lambda=((2a)^{r})=(2a,...,2a) be a partition of 2ar2ar, i.e., we have a rectangle of size (r×2a)(r\times 2a). Let Σ\Sigma be the set of all tableaux TT of shape λ\lambda with the entries in [1,r]={1,,r}[1,r]=\{1,...,r\} so that in each column we have a permutation of the set [1,r][1,r] and the first column corresponds to an even permutation. We set sgnT=i=12asgnσi\operatorname{sgn\,}T=\prod_{i=1}^{2a}\operatorname{sgn\,}\sigma_{i}, where σi\sigma_{i} is the permutation of [1,r][1,r] corresponding to the ii-th column of TT and sgnσ\operatorname{sgn\,}\sigma is the sign of permutation σ\sigma. Now we define the map

φ:i=12a(j=1rV)Symr(Sym2aV)\varphi:\prod_{i=1}^{2a}\left(\prod_{j=1}^{r}V\right)\to\operatorname{Sym}^{r}(\operatorname{Sym}^{2a}V)

by setting

φ((v11,,vr1),,(v1,2a,,vr,2a))=TΣsgnTj=1r(i=12avT(j,i),i).\varphi((v_{11},...,v_{r1}),...,(v_{1,2a},...,v_{r,2a}))=\sum_{T\in\Sigma}{\operatorname{sgn\,}T}\prod_{j=1}^{r}\left(\prod_{i=1}^{2a}v_{T(j,i),i}\right).

Since this map is multilinear it factors to the map

φ:(Vr)2a=i=12a(j=1rV)Symr(Sym2aV)\varphi:(V^{\otimes r})^{\otimes 2a}=\bigotimes_{i=1}^{2a}\left(\bigotimes_{j=1}^{r}V\right)\to\operatorname{Sym}^{r}(\operatorname{Sym}^{2a}V)

Note that this last map is alternating in each set of variables (v1m,,vrm)(v_{1m},...,v_{rm}), where m[1,2a]m\in[1,2a]. Since we work over \mathbb{Z}, it is sufficient to check that the corresponding multilinear form is antisymmetric. This is clear for m>1m>1 as exchanging vimv_{im} with vjmv_{jm} defines a bijection on the set Σ\Sigma that replaces the tableau TT with another tableau with exchanged entries between (i,m)(i,m) and (j,m)(j,m) places. For m=1m=1 it follows from the fact that exchanging vi1v_{i1} and vj1v_{j1} defines a bijection on the set Σ\Sigma that replaces the tableau TT with another tableau with the same first column but exchanged ii-th and jj-th entries on all of the remaining 2a12a-1 columns. This changes the sign with which the corresponding product is taken. Therefore φ\varphi is antisymmetric also in the variables (v11,,vr1)(v_{11},...,v_{r1}). This implies that the formula

i=12a(v1ivri)TΣsgnTj=1r(i=12avT(j,i),i)\bigotimes_{i=1}^{2a}(v_{1i}\wedge...\wedge v_{ri})\to\sum_{T\in\Sigma}{\operatorname{sgn\,}T}\prod_{j=1}^{r}\left(\prod_{i=1}^{2a}v_{T(j,i),i}\right)

defines a map of GL(V){\rm GL}\,(V)-modules

(rV)2aSymr(Sym2aV).(\bigwedge^{r}V)^{\otimes 2a}\to\operatorname{Sym}^{r}(\operatorname{Sym}^{2a}V).

If (e1,,er)(e_{1},...,e_{r}) is a basis of VV, the element i=12a(e1er)\bigotimes_{i=1}^{2a}(e_{1}\wedge...\wedge e_{r}) is mapped to

W=TΣsgnTj=1r(i=12aeT(j,i)).W=\sum_{T\in\Sigma}{\operatorname{sgn\,}T}\prod_{j=1}^{r}\left(\prod_{i=1}^{2a}e_{T(j,i)}\right).

Note that Symr(Sym2aV)\operatorname{Sym}^{r}(\operatorname{Sym}^{2a}V) has a standard basis corresponding to

j=1r(i=1reinij),\prod_{j=1}^{r}\left(\prod_{i=1}^{r}e_{i}^{n_{ij}}\right),

where inij=2a\sum_{i}n_{ij}=2a for j=1,,rj=1,...,r. If we write WW in this basis, the coefficient at the element j=1rej2a\prod_{j=1}^{r}e_{j}^{2a} is equal to 11, so the corresponding map is non-zero.

To see the last part of the lemma, it is sufficient to remark that we have

TΣsgnTj=1ri=12aeT(j,i)=TΣsgnTi=12aj=1reT(j,i)=(TΣsgnT)i=12aj=1rej=0\sum_{T\in\Sigma}{\operatorname{sgn\,}T}\prod_{j=1}^{r}\prod_{i=1}^{2a}e_{T(j,i)}=\sum_{T\in\Sigma}{\operatorname{sgn\,}T}\prod_{i=1}^{2a}\prod_{j=1}^{r}e_{T(j,i)}=(\sum_{T\in\Sigma}{\operatorname{sgn\,}T})\prod_{i=1}^{2a}\prod_{j=1}^{r}e_{j}=0

in Sym2raV\operatorname{Sym}^{2ra}V. ∎

Remark 5.10.

[Wei90, Example 1.9] shows that the plethysm Sym5(Sym35)\operatorname{Sym}^{5}(\operatorname{Sym}^{3}\mathbb{C}^{5}) does not contain (55)3(\bigwedge^{5}\mathbb{C}^{5})^{\otimes 3} as a GL(,5){\rm GL}\,(\mathbb{C},5)-submodule. Thus the parity condition in the above lemma is necessary.

Corollary 5.11.

Let \mathcal{E} be a rank rr vector bundle on some scheme XX defined over some commutative ring kk. Then for any a,b1a,b\geq 1 we have

  1. (1)

    a canonical surjection

    Syma(Symb)Symab,\operatorname{Sym}^{a}(\operatorname{Sym}^{b}\mathcal{E})\twoheadrightarrow\operatorname{Sym}^{ab}\mathcal{E},
  2. (2)

    a canonical inclusion

    (det)2aSymr(Sym2a)(\det\mathcal{E})^{\otimes 2a}\hookrightarrow\operatorname{Sym}^{r}(\operatorname{Sym}^{2a}\mathcal{E})

    onto a subbundle.

Moreover, the composition

(det)2aSymr(Sym2a)Sym2ra(\det\mathcal{E})^{\otimes 2a}\hookrightarrow\operatorname{Sym}^{r}(\operatorname{Sym}^{2a}\mathcal{E})\twoheadrightarrow\operatorname{Sym}^{2ra}\mathcal{E}

is zero.

Proof.

The corollary follows immediately from the previous lemma. For the convenience of the reader we recall the idea of proof. Let VV be a free kk-module of rank rr and let PXP\to X be the principal GL(V){\rm GL\,}(V)-bundle associated to \mathcal{E}. Then for any GL(V){\rm GL}\,(V)-module WW we have the associated vector bundle P(W)P(W) and maps of GL(V){\rm GL}\,(V)-modules induce the corresponding maps of vector bundles. Applying this construction to maps from Lemma 5.9, we get the corresponding maps from the corollary. ∎

When \mathcal{E} is a μH\mu_{H}-semistable bundle on (X,H)(X,H) such that Δ()0\Delta(\mathcal{E})\equiv 0, [Mis21, Lemma 2.3] observes that Symm\operatorname{Sym}^{m}\mathcal{E} is also μH\mu_{H}-semistable and Δ(Symm)0\Delta(\operatorname{Sym}^{m}\mathcal{E})\equiv 0 for all m0m\geq 0. If furthermore XX is 11-homogeneous, then it follows from [Mis21, Theorem 1.2] that (Symm)\mathbb{P}(\operatorname{Sym}^{m}\mathcal{E}) is 11-homogeneous for all m0m\geq 0. Question 1.4 should also consider m1m\geq 1.

Example 5.12.

On 2\mathbb{P}^{2} consider =Sym2(T2(1))\mathcal{E}=\operatorname{Sym}^{2}(T_{\mathbb{P}^{2}}(-1)). Then ()Hilb22\mathbb{P}(\mathcal{E})\simeq{\rm Hilb}^{2}\mathbb{P}^{2}. The divisor EE of nonreduced length 2 subschemes of 2\mathbb{P}^{2} is contracted by the birational Hilbert–Chow morphism. In particular, it is effective, but not a nef divisor. If L=𝒪()(1)L=\mathcal{O}_{\mathbb{P}(\mathcal{E})}(1) and HH is the pullback of the class of a line in 2\mathbb{P}^{2} then EE is linearly equivalent to 2(LH)2(L-H). See [FL17b, Section 7.2] for details.

Another perspective at this example is as follows. Since \mathcal{E} has rank 22, Corollary 5.11 and comparison of ranks imply that we have a short exact sequence of vector bundles

0(det)2Sym2(Sym2)Sym40.0\to(\det\mathcal{E})^{\otimes 2}\to\operatorname{Sym}^{2}(\operatorname{Sym}^{2}\mathcal{E})\to\operatorname{Sym}^{4}\mathcal{E}\to 0.

This gives a short exact sequence

0𝒪2(Sym2(Sym2T2(1)))(2)(Sym4(T2(1)))(2)0.0\to\mathcal{O}_{\mathbb{P}^{2}}\to\bigl{(}\operatorname{Sym}^{2}(\operatorname{Sym}^{2}T_{\mathbb{P}^{2}}(-1))\bigr{)}(-2)\to\bigl{(}\operatorname{Sym}^{4}(T_{\mathbb{P}^{2}}(-1))\bigr{)}(-2)\to 0.

In particular, we see that (Sym2(Sym2T2(1)))(2)\bigl{(}\operatorname{Sym}^{2}(\operatorname{Sym}^{2}T_{\mathbb{P}^{2}}(-1))\bigr{)}(-2) is effective. However, it is not nef, e.g., because its quotient (Sym4(T2(1)))(2)\bigl{(}\operatorname{Sym}^{4}(T_{\mathbb{P}^{2}}(-1))\bigr{)}(-2) restricts to 𝒪L(2)𝒪L(1)𝒪L𝒪L(1)𝒪L(2)\mathcal{O}_{L}(-2)\oplus\mathcal{O}_{L}(-1)\oplus\mathcal{O}_{L}\oplus\mathcal{O}_{L}(1)\oplus\mathcal{O}_{L}(2) on every line LL in 2\mathbb{P}^{2}.

Remark 5.13.

Let \mathcal{E} is a vector bundle on XX. Then one can easily see that the following conditions are equivalent:

  1. (1)

    ()\mathbb{P}(\mathcal{E}) is 11-homogeneous

  2. (2)

    If DD is a divisor on XX such that (Symm)(D)\bigl{(}\operatorname{Sym}^{m}\mathcal{E}\bigr{)}(D) is effective for some m0m\geq 0 then (Symm)(D)\bigl{(}\operatorname{Sym}^{m}\mathcal{E}\bigr{)}(D) is nef.

Theorem 5.14.

Let XX be a smooth projective variety defined over an algebraically closed field kk. Let \mathcal{E} be a vector bundle of rank rr on XX. Then \mathcal{E} is strongly slope semistable with respect to any ample polarization and Δ()0\Delta(\mathcal{E})\equiv 0 if any of the following conditions hold.

  1. (1)

    For every m0m\geq 0, we have that (Symm)\mathbb{P}(\operatorname{Sym}^{m}\mathcal{E}) is 11-homogeneous.

  2. (2)

    For every divisor DD on XX and every m,l0m,l\geq 0, we have that (Syml(Symm))(D)(\operatorname{Sym}^{l}(\operatorname{Sym}^{m}\mathcal{E}))(D) is effective if and only if it is nef.

  3. (3)

    There exists m1m\geq 1 such that Symr(Sym2m)(det)2m\operatorname{Sym}^{r}(\operatorname{Sym}^{2m}\mathcal{E})\otimes(\det\mathcal{E}^{\vee})^{\otimes 2m} is nef.

  4. (4)

    (r)\mathbb{P}(\mathcal{E}^{\otimes r}) is 11-homogeneous.

  5. (5)

    (nd)\mathbb{P}(\operatorname{{\mathcal{E}}nd\,}\mathcal{E}) is 11-homogeneous.

Proof.

The equivalence of (1)(1) and (2)(2) follows from Remark 5.13. We focus on (2)(2). By Corollary 5.11 for all m1m\geq 1, the bundle Symr(Sym2m)\operatorname{Sym}^{r}(\operatorname{Sym}^{2m}\mathcal{E}) contains (det)2m(\det\mathcal{E})^{\otimes 2m}. Thus Symr(Sym2m)(det)2m\operatorname{Sym}^{r}(\operatorname{Sym}^{2m}\mathcal{E})\otimes(\det\mathcal{E}^{\vee})^{\otimes 2m} is effective and hence it is also nef. Since Corollary 5.11 implies that Sym2mr(det)2m\operatorname{Sym}^{2mr}\mathcal{E}\otimes(\det\mathcal{E}^{\vee})^{\otimes 2m} is a quotient of Symr(Sym2m)(det)2m\operatorname{Sym}^{r}(\operatorname{Sym}^{2m}\mathcal{E})\otimes(\det\mathcal{E}^{\vee})^{\otimes 2m}, it is also nef. Since nefness for (twisted) vector bundles is homogeneous (cf. [Laz04b, Theorem 6.2.12], or [FM21, Lemma 3.24 and Remark 3.10]), we deduce that 1rdet\mathcal{E}\langle-\frac{1}{r}\det\mathcal{E}^{\vee}\rangle is nef. Conclude by Theorem 1.1. This argument also handled (3)(3).

(4)(4) We have a natural inclusion det=rr\det\mathcal{E}=\bigwedge^{r}\mathcal{E}\hookrightarrow\mathcal{E}^{\otimes r}. It is obtained by dualizing the natural surjection ()rr()(\mathcal{E}^{\vee})^{\otimes r}\twoheadrightarrow\bigwedge^{r}(\mathcal{E}^{\vee}). It shows that rdet\mathcal{E}^{\otimes r}\otimes\det\mathcal{E}^{\vee} is effective. By the assumption on the positive cones, it is then also nef. Hence so is its quotient Symrdet\operatorname{Sym}^{r}\mathcal{E}\otimes\det\mathcal{E}^{\vee}. Argue as above.

(5)(5) We have a natural inclusion 𝒪Xnd\mathcal{O}_{X}\hookrightarrow{\mathcal{E}}{\rm nd}\,\mathcal{E} induced by sending 1𝒪X(U)1\in\mathcal{O}_{X}(U) to id(U){\rm id}_{\mathcal{E}(U)}. Therefore nd{\mathcal{E}}{\rm nd}\,\mathcal{E} is effective and hence our assumption implies that it is nef. Now Theorem 4.10 implies the required assertion. ∎

Together with Theorem 5.3 this implies the following result:

Corollary 5.15.

Let XX be a smooth projective variety defined over an algebraically closed field kk and let \mathcal{E} be a vector bundle of rank rr on XX. Then the following conditions are equivalent:

  1. (1)

    (Symm)\mathbb{P}(\operatorname{Sym}^{m}\mathcal{E}) is 11-homogeneous for every m0m\geq 0.

  2. (2)

    (Sym2m)\mathbb{P}(\operatorname{Sym}^{2m}\mathcal{E}) is 11-homogeneous for some m1m\geq 1.

  3. (3)

    (r)\mathbb{P}(\mathcal{E}^{\otimes r}) is 11-homogeneous.

  4. (4)

    (nd)\mathbb{P}(\operatorname{{\mathcal{E}}nd\,}\mathcal{E}) is 11-homogeneous.

  5. (5)

    The bundle nd\operatorname{{\mathcal{E}}nd\,}\mathcal{E} is nef and XX is 11-homogeneous.

References

  • [Bar71] Charles M. Barton, Tensor products of ample vector bundles in characteristic pp, Amer. J. Math. 93 (1971), 429–438. MR 289525
  • [BB08] Indranil Biswas and Ugo Bruzzo, On semistable principal bundles over a complex projective manifold, Int. Math. Res. Not. IMRN (2008), no. 12, Art. ID rnn035, 28. MR 2426752
  • [BBG19] Indranil Biswas, Ugo Bruzzo, and Sudarshan Gurjar, Higgs bundles and fundamental group schemes, Adv. Geom. 19 (2019), no. 3, 381–388. MR 3982575
  • [BCI11] Peter Bürgisser, Matthias Christandl, and Christian Ikenmeyer, Even partitions in plethysms, J. Algebra 328 (2011), 322–329. MR 2745569
  • [BG71] Spencer Bloch and David Gieseker, The positivity of the Chern classes of an ample vector bundle, Invent. Math. 12 (1971), 112–117. MR 297773
  • [BHP14] Indranil Biswas, Amit Hogadi, and A. J. Parameswaran, Pseudo-effective cone of Grassmann bundles over a curve, Geom. Dedicata 172 (2014), 69–77. MR 3253771
  • [BHR06] U. Bruzzo and D. Hernández Ruipérez, Semistability vs. nefness for (Higgs) vector bundles, Differential Geom. Appl. 24 (2006), no. 4, 403–416. MR 2231055
  • [BPMGNO19] L. Brambila-Paz, O. Mata-Gutiérrez, P. E. Newstead, and Angela Ortega, Generated coherent systems and a conjecture of D. C. Butler, Internat. J. Math. 30 (2019), no. 5, 1950024, 25. MR 3961440
  • [Das20] Omprokash Das, Finiteness of log minimal models and nef curves on 3-folds in characteristic p>5p>5, Nagoya Math. J. 239 (2020), 76–109. MR 4138896
  • [DH91] Steven Diaz and David Harbater, Strong Bertini theorems, Trans. Amer. Math. Soc. 324 (1991), no. 1, 73–86. MR 986689
  • [dJ96] Aise Johan de Jong, Smoothness, semi-stability and alterations, Inst. Hautes Études Sci. Publ. Math. (1996), no. 83, 51–93.
  • [DPS94] Jean-Pierre Demailly, Thomas Peternell, and Michael Schneider, Compact complex manifolds with numerically effective tangent bundles, J. Algebraic Geom. 3 (1994), no. 2, 295–345. MR 1257325
  • [DW20] Christopher Deninger and Annette Werner, Parallel transport for vector bundles on pp-adic varieties, J. Algebraic Geom. 29 (2020), no. 1, 1–52. MR 4028065
  • [EL92] Lawrence Ein and Robert Lazarsfeld, Stability and restrictions of Picard bundles, with an application to the normal bundles of elliptic curves, Complex projective geometry (Trieste, 1989/Bergen, 1989), London Math. Soc. Lecture Note Ser., vol. 179, Cambridge Univ. Press, Cambridge, 1992, pp. 149–156. MR 1201380
  • [ELM13] Lawrence Ein, Robert Lazarsfeld, and Yusuf Mustopa, Stability of syzygy bundles on an algebraic surface, Math. Res. Lett. 20 (2013), no. 1, 73–80. MR 3126723
  • [FL83] William Fulton and Robert Lazarsfeld, Positive polynomials for ample vector bundles, Ann. of Math. (2) 118 (1983), no. 1, 35–60.
  • [FL17a] Mihai Fulger and Brian Lehmann, Positive cones of dual cycle classes, Algebraic Geometry 4 (2017), no. 1, 1–28.
  • [FL17b] by same author, Zariski decompositions of numerical cycle classes, J. Algebraic Geom. 26 (2017), no. 1, 43–106. MR 3570583
  • [FM21] Mihai Fulger and Takumi Murayama, Seshadri constants for vector bundles, J. Pure Appl. Algebra 225 (2021), no. 4, 106559, 35. MR 4158762
  • [Ful98] William Fulton, Intersection theory, second ed., Ergeb. Math. Grenzgeb. (3), vol. 2, Springer-Verlag, Berlin, 1998. MR 1644323
  • [Ful11] Mihai Fulger, Cones of effective cycles on projective bundles over curves, Math. Z. 269 (2011), no. 1-2, 449–459. MR 2836078
  • [Ful20] by same author, Cones of positive vector bundles, Rev. Roumaine Math. Pures Appl. 65 (2020), no. 3, 285–302. MR 4216530
  • [Gie71] David Gieseker, pp-ample bundles and their Chern classes, Nagoya Math. J. 43 (1971), 91–116. MR 296078
  • [GKP16] Daniel Greb, Stefan Kebekus, and Thomas Peternell, Movable curves and semistable sheaves, Int. Math. Res. Not. IMRN (2016), no. 2, 536–570. MR 3493425
  • [Har70] Robin Hartshorne, Ample subvarieties of algebraic varieties, Lecture Notes in Mathematics, Vol. 156, Springer-Verlag, Berlin-New York, 1970, Notes written in collaboration with C. Musili.
  • [Har71] by same author, Ample vector bundles on curves, Nagoya Math. J. 43 (1971), 73–89. MR 292847
  • [HL10] Daniel Huybrechts and Manfred Lehn, The geometry of moduli spaces of sheaves, second ed., Cambridge Mathematical Library, Cambridge University Press, Cambridge, 2010. MR 2665168
  • [JP15] Kirti Joshi and Christian Pauly, Hitchin-Mochizuki morphism, opers and Frobenius-destabilized vector bundles over curves, Adv. Math. 274 (2015), 39–75. MR 3318144
  • [Kle69] Steven L. Kleiman, Ample vector bundles on algebraic surfaces, Proc. Amer. Math. Soc. 21 (1969), 673–676. MR 251044
  • [Kle05] by same author, The Picard scheme, Fundamental algebraic geometry, Math. Surveys Monogr., vol. 123, Amer. Math. Soc., Providence, RI, 2005, pp. 235–321. MR 2223410
  • [Kol96] János Kollár, Rational curves on algebraic varieties, Ergeb. Math. Grenzgeb. (3), vol. 32, Springer-Verlag, Berlin, 1996. MR 1440180
  • [Lan04] Adrian Langer, Semistable sheaves in positive characteristic, Ann. of Math. (2) 159 (2004), no. 1, 251–276. MR 2051393
  • [Lan11] by same author, On the S-fundamental group scheme, Ann. Inst. Fourier (Grenoble) 61 (2011), no. 5, 2077–2119 (2012). MR 2961849
  • [Lan12] by same author, On the S-fundamental group scheme. II, J. Inst. Math. Jussieu 11 (2012), no. 4, 835–854. MR 2979824
  • [Lan15] by same author, Bogomolov’s inequality for Higgs sheaves in positive characteristic, Invent. Math. 199 (2015), no. 3, 889–920. MR 3314517
  • [Lan19] by same author, Nearby cycles and semipositivity in positive characteristic, 2019, to appear in J. Eur. Math. Soc., arXiv:1902.05745v3 [math.AG].
  • [Lan21] by same author, On algebraic chern classes of flat vector bundles, 2021, arXiv:2107.03127 [math.AG].
  • [Laz04a] Robert Lazarsfeld, Positivity in algebraic geometry. I, Ergeb. Math. Grenzgeb. (3), vol. 48, Springer-Verlag, Berlin, 2004, Classical setting: line bundles and linear series. MR 2095472
  • [Laz04b] by same author, Positivity in algebraic geometry. II, Ergeb. Math. Grenzgeb. (3), vol. 49, Springer-Verlag, Berlin, 2004, Positivity for vector bundles, and multiplier ideals. MR 2095472
  • [Lie68] David I. Lieberman, Numerical and homological equivalence of algebraic cycles on Hodge manifolds, Amer. J. Math. 90 (1968), 366–374. MR 230336
  • [LP97] J. Le Potier, Lectures on vector bundles, Cambridge Studies in Advanced Mathematics, vol. 54, Cambridge University Press, Cambridge, 1997, Translated by A. Maciocia. MR 1428426
  • [Mis21] Snehajit Misra, Pseudo-effective cones of projective bundles and weak Zariski decomposition, Eur. J. Math. 7 (2021), no. 4, 1438–1457. MR 4340943
  • [Miy87] Yoichi Miyaoka, The Chern classes and Kodaira dimension of a minimal variety, Algebraic geometry, Sendai, 1985, Adv. Stud. Pure Math., vol. 10, North-Holland, Amsterdam, 1987, pp. 449–476. MR 946247
  • [Mor98] Atsushi Moriwaki, Relative Bogomolov’s inequality and the cone of positive divisors on the moduli space of stable curves, J. Amer. Math. Soc. 11 (1998), no. 3, 569–600. MR 1488349
  • [MR21] Snehajit Misra and Nabanita Ray, On Ampleness of vector bundles, C. R. Math. Acad. Sci. Paris 359 (2021), 763–772. MR 4311802
  • [MR82] V. B. Mehta and A. Ramanathan, Semistable sheaves on projective varieties and their restriction to curves, Math. Ann. 258 (1981/82), no. 3, 213–224. MR 649194
  • [Nak99] Noboru Nakayama, Normalized tautological divisors of semi-stable vector bundles, no. 1078, 1999, Free resolutions of coordinate rings of projective varieties and related topics (Japanese) (Kyoto, 1998), pp. 167–173. MR 1715587
  • [RR84] S. Ramanan and A. Ramanathan, Some remarks on the instability flag, Tohoku Math. J. (2) 36 (1984), no. 2, 269–291. MR 742599
  • [Sch05] Olivier Schneider, Stabilité des fibrés ΛpEL\Lambda^{p}E_{L} et condition de Raynaud, Ann. Fac. Sci. Toulouse Math. (6) 14 (2005), no. 3, 515–525. MR 2172589
  • [Sim92] Carlos T. Simpson, Higgs bundles and local systems, Inst. Hautes Études Sci. Publ. Math. (1992), no. 75, 5–95. MR 1179076
  • [Wei90] Steven H. Weintraub, Some observations on plethysms, J. Algebra 129 (1990), no. 1, 103–114. MR 1037395