Log in

On the Bogomolov–Gieseker inequality for tame Deligne–Mumford surfaces

  • Published:
Mathematische Zeitschrift Aims and scope Submit manuscript

Abstract

We generalize the Bogomolov–Gieseker inequality for semistable coherent sheaves on smooth projective surfaces to smooth Deligne–Mumford surfaces. We work over positive characteristic \(p>0\) and generalize Langer’s method to smooth Deligne–Mumford stacks. As applications we obtain the Bogomolov inequality for semistable coherent sheaves on a Deligne–Mumford surface in characteristic zero, and the Bogomolov inequality for semistable sheaves on a root stack over a smooth surface which is equivalent to the Bogomolov inequality for the rational parabolic sheaves on a smooth surface S. In a joint appendix with Hao Sun, we generalize the Bogomolov inequality formula to Simpson Higgs sheaves on tame Deligne–Mumford stacks.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Germany)

Instant access to the full article PDF.

Similar content being viewed by others

References

  1. Biswas, I.: Parabolic bundles as orbifold bundles. Duke Math. J. 88(2), 305–325 (1997)

    Article  MathSciNet  Google Scholar 

  2. Bloch, S., Gieseker, D.: The positivity of the Chern classes of an ample vector bundle. Invent. Math. 12, 112–117 (1971)

    Article  MathSciNet  ADS  Google Scholar 

  3. Borne, N., Vistoli, A.: Parabolic sheaves on logarithmic schemes. Adv. Math. 231(3–4), 1327–1363 (2012). ar**v:1001.0466

    Article  MathSciNet  Google Scholar 

  4. Chen, J., Tseng, H.: On the Bogomolov–Miyaoka–Yau inequality for stacky surfaces. Taiwan. J. Math. 24(4), 841–853 (2020)

    Article  MathSciNet  Google Scholar 

  5. Coates, T., Iritani, H., Jiang, Y., Segal, E.: K-theoretical and categorical properties of toric Deligne–Mumford stacks. Pure Appl. Math. Q. 11(2), 239–266 (2015). ar**v:1410.0027

    Article  MathSciNet  Google Scholar 

  6. Edidin, D.: Riemann–Roch for Deligne–Mumford Stacks. In: Proceedings of the Harris 60 Conference. ar**v:1205.4742

  7. Gieseker, D.: On a theorem of Bogomolov on Chern classes of stable bundles. Am. J. Math. 101(1), 77–85 (1979)

    Article  MathSciNet  Google Scholar 

  8. Huybrechts, D., Lehn, M.: The geometry of moduli spaces of sheaves, Aspects of Mathematics, E31, Friedr. Vieweg & Sohn, Braunschweig. MR MR1450870 (98g:14012) (1997)

  9. Jiang, Y., Kundu, P.: The Tanaka-Thomas’s Vafa–Witten invariants for surface Deligne–Mumford stacks. Pure Appl. Math. Q. 17(1), 503–573 (2021). ar**v:1903.11477

    Article  MathSciNet  Google Scholar 

  10. Jiang, Y.: The Vafa–Witten invariants for surface Deligne–Mumford stacks and S-duality. In: Proceedings of ICCM-2019. ar**v:1909.03067

  11. Jiang, Y.: Counting twisted sheaves and S-duality. Adv. Math. 400, 108332 (2022). ar**v:1909.04241

  12. Kawamata, Y.: Abundance theorem for minimal threefolds. Invent. Math. 108, 229–246 (1992)

    Article  MathSciNet  ADS  Google Scholar 

  13. Katz, N.: Nilpotent connections and the monodromy theorem: applications of a result of Turrittin. Publ. Math. IHES 39, 175–232 (1970)

    Article  MathSciNet  Google Scholar 

  14. Kresch, A., Vistoli A.: On coverings of Deligne–Mumford stacks and surjectivity of the Brauer map. The Bulletin of the London Mathematical Society, vol. 36, pp. 88–192. ar**v:math/0301249

  15. Langer, A.: Semistable sheaves in positive characteristic. Ann. Math. 159, 251–276 (2004)

    Article  MathSciNet  Google Scholar 

  16. Langer, A.: Bogomolov’s inequality for Higgs sheaves in positive characteristic. Invent. Math. 199, 889–920 (2015)

    Article  MathSciNet  ADS  Google Scholar 

  17. Langer, A.: Semistable modules over Lie algebroids in positive characteristic. Doc. Math. 19, 509–540 (2014)

    Article  MathSciNet  Google Scholar 

  18. Langer, A.: The Bogomolov–Miyaoka–Yau inequality for logarithmic surfaces in positive characteristic. Duke Math. J. 165(14), 2737–2769 (2016)

    Article  MathSciNet  Google Scholar 

  19. Laumon, G., Moret-Bailly, L.: Champs Algebriques. Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge/A Series of Modern Surveys in Mathematics

  20. Maruyama, M., Yokogawa, K.: Moduli of parabolic stable sheaves. Math. Ann. 293(1), 77–99 (1992)

    Article  MathSciNet  Google Scholar 

  21. Matsuki, K., Olsson, M.: Kawamata–Viehweg vanishing as Kodaira vanishing for stacks. Mathematical Research Letters. ar**v:math/0212259

  22. Moriwaki, A.: A note on Bogomolov–Gieseker’s inequality in positive characteristic. Duke Math. J. 64, 361–375 (1991)

    Article  MathSciNet  Google Scholar 

  23. Nironi, F.: Moduli Spaces of Semistable Sheaves on Projective Deligne–Mumford Stacks. ar**v:0811.1949

  24. Ogus, A., Vologodsky, V.: Nonabelian Hodge theory in characteristic \(p\). Publ. Math. Inst. Hautes études Sci. 106, 1–138 (2007)

    Article  MathSciNet  Google Scholar 

  25. Olsson, M., Starr, J.: Quot functors for Deligne-Mumford stacks. Commun. Algebra 31(8), 4069–4096 (2003). Special issue in honor of Steven L. Kleiman

  26. Shepherd-Barron, N.I.: Unstable vector bundles and linear systems on surfaces in characteristic \(p\). Invent. Math. 106, 243–262 (1991)

    Article  MathSciNet  ADS  Google Scholar 

  27. Stack Project. https://stacks.math.columbia.edu/download/stacks-morphisms.pdf

  28. Tanaka, Y., Thomas, R.P.: Vafa–Witten invariants for projective surfaces I: stable case. J. Algebraic Geom. 29, 603–668 (2020). ar**v:1702.08487

    Article  MathSciNet  Google Scholar 

  29. Tanaka, Y., Thomas, R. P.: Vafa–Witten invariants for projective surfaces II: semistable case. Pure Appl. Math. Q. 13, 517–562 (2017). Special Issue in Honor of Simon Donaldson. ar**v:1702.08488

  30. Toen, B.: Théorémes de Riemann–Roch pour les champs de Deligne–Mumford. K-Theory 18(1), 33–76 (1999)

    Article  MathSciNet  Google Scholar 

  31. Vafa, C., Witten, E.: A strong coupling test of S-duality. Nucl. Phys. B 431, 3–77 (1994). hep-th/9408074

    Article  MathSciNet  ADS  Google Scholar 

Download references

Acknowledgements

The authors would like to thank A. Langer for useful email correspondences and valuable discussions. This work is partially supported by NSF and a Simons Foundation Collaboration Grant.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Promit Kundu.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Higher dimension case

In the appendix we generalize Langer’s argument to higher dimension case. For simplicity of the calculation of modified slopes, we restrict to a special case of smooth Deligne–Mumford stacks \(\mathcal {X}=[Z/G]\) which is a quotient stack such that the action of G is diagonalizable. We still let the generating sheaf \(\** \) on \(\mathcal {X}\) satisfying condition Sect. 3.1. Still let \(d=\dim (\mathcal {X})\) be the dimension of \(\mathcal {X}.\)

We state several theorems generalizing Langer [15, §3].

Theorem A.1

Let \(D_1\) be a very ample divisor on X and \({{\mathcal {D}}}_1:=\pi ^{-1}(D_1)\) and \({{\mathcal {D}}}=\pi ^{-1}(D)\) for D a general element \(D\in |D_1|\). If the restriction of a coherent sheaf E on \(\mathcal {X}\) to \({{\mathcal {D}}}\) is not modified slope semistable with respect to \(H|_{D}\) and \(\** |_{D},\) then let \(\mu _{\** ,i}, r_i\) denote the modified slopes and ranks respectively in the Harder–Narasimhan filtration of \(E|_{D},\) we have

$$\begin{aligned}{} & {} \sum _{i<j}r_i r_j(\mu _{\** ,i}-\mu _{\** ,j})^2\leqslant H^{d}\Delta (E)\nonumber \\{} & {} \quad +2\mathop {\text {rk}}(E)^2(L_{\** ,\max }(E)-\mu _{\** }(E)) (\mu _{\** }(E)-L_{\** ,\min }(E)). \end{aligned}$$
(A.1)

(See 2.3 for definitions related to the above theorem.)

Theorem A.2

If a torsion free sheaf E on \(\mathcal {X}\) is strongly modified slope semistable,  we have

$$\begin{aligned} \Delta (E)\cdot H^{d-2}\geqslant 0. \end{aligned}$$

Theorem A.3

If a torsion free sheaf E on \(\mathcal {X}\) is just modified slope semistable,  then we have

$$\begin{aligned} H^{d}\cdot \Delta (E)\cdot H^{d-2}+\mathop {\text {rk}}(\** )^2\beta _{\mathop {\text {rk}}(E)}\geqslant 0. \end{aligned}$$

Before stating the last theorem, we introduce some notations. First for torsion free sheaves \(G^{\prime }, G\) on \(\mathcal {X},\) we set

$$\begin{aligned} \xi ^{\** }_{G^{\prime }, G}=\frac{c_1(G^{\prime })}{\mathop {\text {rk}}(\** )\mathop {\text {rk}}(G^{\prime })}-\frac{c_1(G)}{\mathop {\text {rk}}(\** )\mathop {\text {rk}}(G)}. \end{aligned}$$

We also set

$$\begin{aligned} K^+:=\{D\in \text {Num}(X)| D^2H^{d-2}>0, DH^{d-1}\geqslant 0 \text { for all nef }H\}. \end{aligned}$$

where \(\text {Num}(X)={{\,\textrm{Pic}\,}}(X)\otimes {{\mathbb {R}}}/\sim \) and \(\sim \) is an equivalence relation meaning \(L_1\sim L_2\) if and only if \(L_1AH^{d-2}=L_2AH^{d-2}\) for all divisors A on X.

Theorem A.4

If we have

$$\begin{aligned} H^{d}\cdot \Delta (E)\cdot H^{d-2}+\mathop {\text {rk}}(\** )^2\beta _{\mathop {\text {rk}}(E)}< 0, \end{aligned}$$

then there exists a saturated subsheaf \(E^{\prime }\subset E\) such that \(\xi ^{\** }_{E^{\prime }, E}\in K^{+}.\)

We prove these theorems by induction on the rank \(\mathop {\text {rk}}(E),\) and following Langer’s method. We only state the parts of the proof which are different to Langer’s method in smooth case and refer to [15, §3] for detailed arguments in the proof which is the same as Langer. For the induction process, let \({{\,\textrm{Thm}\,}}^{i}(\mathop {\text {rk}})\) represent the statement that Theorem A.i holds for ranks \(\leqslant \mathop {\text {rk}}\) for \(i=1,2,3,4.\) The rest of the induction chain follows the original proof due to Langer.

1.1 A.1. \({{\,\textrm{Thm}\,}}^{1}(\mathop {\text {rk}})\) implies \({{\,\textrm{Thm}\,}}^{2}(\mathop {\text {rk}})\)

Suppose that the torsion free sheaf E is strongly modified slope semistable with respect to \((H, \** ),\) and \(\Delta (E)\cdot H^{d-2}<0.\) We have \(L_{\** ,\max }(E)=L_{\** ,\min }=\mu _{\** }(E).\) Theorem \({{\,\textrm{Thm}\,}}^{1}(\mathop {\text {rk}})\) implies that the restriction of E to H is still modified slope semistable. Since E is strongly modified slope semistable, \((F^k)^*E\) is also strongly modified slope semistable, and its restriction to a very general element in |H| is strongly modified slope semistable. Therefore by induction the restriction of \((F^k)^*E\) to a very general element in \(H_1\cap \cdots \cap H_{d-1}\) for \(H_1, \ldots , H_{d-1}\in |H|\) is strongly modified slope semistable. Therefore we are reduced to the two dimensional Deligne–Mumford stack case. Then this is Theorem 3.1.

1.2 A.2. \({{\,\textrm{Thm}\,}}^{2}(\mathop {\text {rk}})\) implies \({{\,\textrm{Thm}\,}}^{3}(\mathop {\text {rk}})\)

First note that in this case, from (3.1), we have \(\beta _{\mathop {\text {rk}}}.\) Our polarization is \((H, \** ),\) we first have the following inequality:

$$\begin{aligned}{} & {} H^{d}\cdot \Delta (E)H^{d-2}+\mathop {\text {rk}}(E)^2\mathop {\text {rk}}(\** )^2(L_{\** ,\max }(E)\nonumber \\{} & {} \quad -\mu _{\** }(E))(\mu _{\** }(E)-L_{\** ,\min }(E))\geqslant 0. \end{aligned}$$
(A.2)

To prove this inequality, first from the finite property fdHPN in Proposition 2.11 there exists a positive integer k such that all the quotients in the Harder–Narasimhan filtration of \((F^k)^*E\) are strongly modified slope semistable. Consider the Harder–Narasimhan filtration

$$\begin{aligned} 0=E_0\subset E_1\subset \cdots \subset E_m=(F^k)^*E \end{aligned}$$

and let \(F_i=E_i/E_{i-1},\) \(r_i=\mathop {\text {rk}}(F_i),\) \(\mu _i=\mu _{\** }(F_i).\) The Hodge index theorem (holds for smooth Deligne–Mumford stacks and we give a short proof in Lemma A.5 below) implies that

$$\begin{aligned} \frac{\Delta ((F^k)^*E)H^{d-2}}{\mathop {\text {rk}}(E)}&=\sum _{i} \frac{\Delta (F_i)H^{d-2}}{r_i}-\frac{1}{\mathop {\text {rk}}(E)} \sum _{i<j}r_ir_j\left( \frac{c_1(F_i)}{r_i}-\frac{c_1(F_j)}{r_j}\right) ^2H^{d-2}\\&\geqslant \sum _{i}\frac{\Delta (F_i)H^{d-2}}{r_i}-\frac{\mathop {\text {rk}}(\** )^2}{H^{d}\mathop {\text {rk}}(E)} \sum _{i<j}r_ir_j(\mu _i-\mu _j)^2 \end{aligned}$$

\({{\,\textrm{Thm}\,}}^{5}(\mathop {\text {rk}})\) implies that \(\Delta (F_i)H^{d-2}\geqslant 0.\) Therefore by [15, Lemma 1.4], we have

$$\begin{aligned}{} & {} \frac{H^{d}\cdot \Delta (E)H^{d-2}}{\mathop {\text {rk}}(E)}\\{} & {} \quad \geqslant -\mathop {\text {rk}}(E)\mathop {\text {rk}}(\** )^2\left( \mu _{\** ,\max }((F^k)^*E)- \mu _{\** }((F^k)^*E)\right) \left( \mu _{\** }((F^k)^*E)-\mu _{\** ,\min }((F^k)^*E)\right) \end{aligned}$$

Both sides are divided by \(p^{2k},\) we get:

$$\begin{aligned} H^{d}\cdot \Delta (E)H^{d-2}+\mathop {\text {rk}}(E)^2\mathop {\text {rk}}(\** )^2(L_{\** ,\max }(E)- \mu _{\** }(E))(\mu _{\** }(E)-L_{\** ,\min }(E))\geqslant 0. \end{aligned}$$

It is ready to prove \({{\,\textrm{Thm}\,}}^{3}(\mathop {\text {rk}}).\) Suppose that E is just modified slope semistable. We aim to use (A.2) and Corollary 2.10. The method is the same as in [15, §3.6], and we take an ample divisor D on X and set \(H(t)=H+tD.\) Similar method shows that the Harder–Narasimhan filtration of E with respect to \((H(t), \** )\) is independent of t when t is positively small. Let \(0=E_0\subset E_1\subset \cdots \subset E_m=E\) be the Harder–Narasimhan filtration with respect to \((H(t), \** ).\) We have (since E is modified slope semistable)

$$\begin{aligned} \mu _{\** , H}(E)\geqslant \mu _{\** , H}(E_1)=\lim _{t\rightarrow 0}\mu _{\** , H(t)}(E_1)\geqslant \lim _{t\rightarrow 0}\mu _{\** , H(t)}(E)=\mu _{\** ,H}(E). \end{aligned}$$

Hence

$$\begin{aligned} \lim _{t\rightarrow 0}\mu _{\** ,\max , H(t)}(E)=\lim _{t\rightarrow 0}\mu _{\** ,\max , H(t)}(E_1)=\mu _{\** ,H}(E), \end{aligned}$$

and similarly,

$$\begin{aligned} \lim _{t\rightarrow 0}\mu _{\** ,\min , H(t)}(E)=\mu _{\** ,H}(E). \end{aligned}$$

Thus we can apply (A.2) and Corollary 2.10, we get the result

$$\begin{aligned} H^{d}\cdot \Delta (E)H^{d-2}+\mathop {\text {rk}}(\** )^2\beta _{\mathop {\text {rk}}(E)}\geqslant 0. \end{aligned}$$

Lemma A.5

Let \(\pi : \mathcal {X}\rightarrow X\) be the coarse moduli associated to a tame smooth projective surface Deligne–Mumford stack \(\mathcal {X}\) over an algebraically closed field k. Consider \(\pi ^{*}H\) where H is an ample divisor on X and let \({{\mathcal {D}}}\) be a numerically non trivial divisor with \({{\mathcal {D}}}\cdot \pi ^{*}H=0,\) then \({{\mathcal {D}}}^{2}<0.\)

Proof

First we can assume that \(\mathcal {X}\) is a quotient stack because of projectivity. From [6, Corollary 4.4], there exists a finite surjective schematic cover of degree d given by \(p:Y\rightarrow \mathcal {X}.\) Since \(\pi \) is finite (as the stack has finite diagonal) we have \(f=\pi \circ p\) is a finite morphism to X and hence Y is projective. From the proof of the same corollary we also have Y can be taken to be smooth. Hence Y is a smooth projective surface and Hodge Index Theorem is known for Y. Next we have \(f^{*}H\) is ample and \(p^{*}[{{\mathcal {D}}}]\) is the class of a divisor induced in \(A^{*}(Y).\) Also note \(p^{*}[{{\mathcal {D}}}]\cdot f^{*}H=d [{{\mathcal {D}}}]\cdot \pi ^{*}H=0\) from our assumption. Hence \((p^{*}[{{\mathcal {D}}}])^{2}=d[{{\mathcal {D}}}]^{2}\) which is less than 0 from Hodge Index for smooth projective surfaces and hence we have the desired result. \(\square \)

1.3 A.3. \({{\,\textrm{Thm}\,}}^{3}(\mathop {\text {rk}})\) implies \({{\,\textrm{Thm}\,}}^{4}(\mathop {\text {rk}})\)

If we have the condition in \({{\,\textrm{Thm}\,}}^{4}(\mathop {\text {rk}}),\) i.e., \(H^{d}\cdot \Delta (E)H^{d-2}+\mathop {\text {rk}}(\** )^2\beta _{\mathop {\text {rk}}(E)}< 0.\) Then from \({{\,\textrm{Thm}\,}}^{3}(\mathop {\text {rk}}),\) E is not modified slope semistable. Let \(E^{\prime }\subset E\) be the maximal destabilizing subsheaf of E and set \(E^{\prime \prime }=E/E^{\prime },\) \(r^{\prime }=\mathop {\text {rk}}(E^{\prime }),\) \(r^{\prime \prime }=\mathop {\text {rk}}(E^{\prime \prime }).\) First we calculate:

$$\begin{aligned} \frac{\Delta (E)H^{d-2}}{\mathop {\text {rk}}(E)} +\frac{r r^{\prime }\xi ^2_{E^{\prime }, E}H^{d-2}}{r^{\prime \prime }}=\frac{\Delta (E^{\prime })H^{d-2}}{r^{\prime }}+ \frac{\Delta (E^{\prime \prime })H^{d-2}}{r^{\prime \prime }}. \end{aligned}$$

We also have \(\mathop {\text {rk}}(\** )^2\frac{\beta _{\mathop {\text {rk}}(E)}}{\mathop {\text {rk}}(E)}\geqslant \mathop {\text {rk}}(\** )^2(\frac{\beta _{r^{\prime }}}{r^{\prime }}+\frac{\beta _{r^{\prime \prime }}}{r^{\prime \prime }}).\) Since \(H^{d}\cdot \Delta (E)H^{d-2}+\mathop {\text {rk}}(\** )^2\beta _{\mathop {\text {rk}}(E)}< 0,\) and we require \(H^d>0,\) either \(\xi ^2_{E^{\prime }, E}>0\) or at least one of the \(H^{d}\cdot \Delta (E^{\prime })H^{d-2}+\mathop {\text {rk}}(\** )^2\beta _{r^{\prime }}\) and \(H^{d}\cdot \Delta (E^{\prime \prime })H^{d-2}+\mathop {\text {rk}}(\** )^2\beta _{r^{\prime \prime }}\) is negative. Therefore the same argument as in [8, Theorem 7.3.3] gives the result.

1.4 A.4. \({{\,\textrm{Thm}\,}}^{4}(\mathop {\text {rk}})\) implies \({{\,\textrm{Thm}\,}}^{2}(\mathop {\text {rk}})\)

Suppose that \(\Delta (E)H^{d-2}<0.\) The condition in \({{\,\textrm{Thm}\,}}^{4}(\mathop {\text {rk}}),\) applying to \((F^l)^*E\) (since E is strongly modified slope semistable by the condition in \({{\,\textrm{Thm}\,}}^{2}(\mathop {\text {rk}})\)), is:

$$\begin{aligned} H^{d}\cdot \Delta ((F^l)^*E)H^{d-2}+\mathop {\text {rk}}(\** )^2\beta _{\mathop {\text {rk}}(E)}< 0 \end{aligned}$$

which is equivalent to (logarithm with base p)

$$\begin{aligned} l>\frac{1}{2}\log _{p}\left( -\frac{\mathop {\text {rk}}(\** )^2\beta _{\mathop {\text {rk}}(E)}}{H^{d}\cdot \Delta (E)H^{d-2}}\right) . \end{aligned}$$

Then for large l,  there exists a saturated torsion free subsheaf \(E^{\prime }\subset (F^l)^*E\) such that \(\xi _{E^{\prime }, (F^l)^*E}\in K^+.\) By “self-duality” property of \(K^+\) we have \(\xi _{E^{\prime }, (F^l)^*E}H^{d-1}>0,\) which means that the sheaf E is not strongly modified semistable, a contradiction.

1.5 A.5. \({{\,\textrm{Thm}\,}}^{2}(\mathop {\text {rk}}-1)\) implies \({{\,\textrm{Thm}\,}}^{1}(\mathop {\text {rk}})\)

We use \(\Pi =|H|\) to denote the complete linear system, and let \(Z:=\{(D,x)\in \Pi \times X: x\in D\}\) be the incidence variety. Let

$$\begin{aligned} p: Z\rightarrow \Pi ; \quad q: Z\rightarrow X \end{aligned}$$

be the corresponding projections. For each \(s\in \Pi ,\) let \(Z_s\) be the scheme theoretic fiber of p over the point s. Consider the following cartesian diagram:

Then \({{\mathcal {Z}}}\) is a Deligne–Mumford stack which is given by \(\{(\pi ^{-1}(D),x): (D,x)\in \Pi \times X, x\in D\}.\) The pull-back of the generating sheaf \(\** \) on \(\mathcal {X}\) under q gives the generating sheaf \(q^*\** \) on \({{\mathcal {Z}}}.\)

We work on the sheaf \(q^*E\) for a torsion free sheaf E on \(\mathcal {X},\) and let

$$\begin{aligned} 0=E_0\subset E_1\subset \cdots \subset E_m=q^*E \end{aligned}$$

be the modified relative Harder–Narasimhan filtration with respect to \(p\circ \pi \) and the pull-back generating sheaf. This means that there exists an open subset \(U\subset \Pi \) such that all \(F_i=E_i/E_{i-1}\) are flat over U and for each \(s\in U\) the fibers \((E_{\bullet })_s\) is the modified Harder–Narasimhan filtration of \(E_s=q^*E|_{{{\mathcal {Z}}}_s}\) for \({{\mathcal {Z}}}_s=\pi ^{-1}(Z_s).\) As in the proof of [15, §3.9], the modified relative Harder–Narasimhan filtration is actually the Harder–Narasimhan filtration of \(q^*E\) with respect to

$$\begin{aligned} ((\pi ^*p^*{{\mathcal {O}}}_{\Pi }(1))^{\dim (\Pi )}q^*H, q^*\** ). \end{aligned}$$

By the finite property in Sect. 2.4, for the sheaf \(q^*E,\) there exists a positive integer k such that all the quotients in the Harder–Narasimhan filtration of \((F^k)^*(q^*E)=q^*((F^k)^*E)\) are strongly modified semistable. We will prove the inequality (A.1), and from [15, Lemma 1.5], when applying to the polygons of the Harder–Narasimhan filtration for modified slopes, we just prove the case that all the graded pieces \(F_i\)’s are strongly modified slope semistable with respect to \((p^*{{\mathcal {O}}}_{\Pi }(1)^{\dim (\Pi )}q^*H, q^*\** ).\)

We perform the same argument as in [15, §3.9], and let \(\Lambda \subset \Pi \) be a pencil. Set \(Y=p^{-1}(\Lambda ),\) and \({{\mathcal {Y}}}=(p\circ \pi )^{-1}(\Lambda )\subset {{\mathcal {Z}}}.\) From the same arguments as in [15, §3.9], we have \(q|_{{{\mathcal {Y}}}}: {{\mathcal {Y}}}\rightarrow \mathcal {X}\) to be the stacky blow up of \(\mathcal {X}\) along \({{\mathcal {B}}}\) which is the base locus of the pull back of the general pencil \(\pi ^{-1}(\Lambda ).\) If \(d\geqslant 3,\) then \({{\mathcal {B}}}\) is a smooth connected substack (this follows from the following Lemma A.6 that we prove below), and there is only one exceptional divisor \({{\mathcal {N}}}\) for \(q|_{{{\mathcal {Y}}}}.\) We write down

$$\begin{aligned} c_1(F_i|_{Y})=q|_{{{\mathcal {Y}}}}^*{{\mathcal {M}}}_i+ b_i {{\mathcal {N}}}\end{aligned}$$

where \({{\mathcal {M}}}_i\) are divisors on \(\mathcal {X}\) which are pullbacks of divisors \(M_i\) on X and \(b_i\) are rational numbers. If the dimension \(d=2,\) then B consists of \(N=H^d\) distinct points and \({{\mathcal {B}}}\) consists of N distinct stacky points. Let \({{\mathcal {N}}}_1,\cdots , {{\mathcal {N}}}_{N}\) be the exceptional divisors of \(q|_{{{\mathcal {Y}}}}.\) There exist rational numbers \(b_{ij}\) and divisors \({{\mathcal {M}}}_i\) such that

$$\begin{aligned} c_1(F_i|_{Y})=q|_{{{\mathcal {Y}}}}^*{{\mathcal {M}}}_i+ \sum _{j}b_{ij} {{\mathcal {N}}}_j. \end{aligned}$$

Let \(b_i=(\sum _{j}b_{ij})/N.\) We have

$$\begin{aligned} \mu _{\** ,i}=\frac{c_1(F_i|_{Y})p^*{{\mathcal {O}}}_{\Pi }(1)q^*H^{d-2}}{r_i\mathop {\text {rk}}(\** )}=\frac{{{\mathcal {M}}}_i\cdot H^{d-1}+b_iN}{r_i\mathop {\text {rk}}(\** )}. \end{aligned}$$

\({{\,\textrm{Thm}\,}}^{2}(\mathop {\text {rk}}-1)\) implies that \(\Delta (F_j|_{Y})p^*{{\mathcal {O}}}_{\Pi }(1)q^*H^{d-2}\geqslant 0\) for every j. We calculate

$$\begin{aligned}&\frac{N\Delta (E)q|_{{{\mathcal {Y}}}}^*H^{d-2}}{\mathop {\text {rk}}(E)}=\sum _{i} \frac{N\Delta (F_i|_{{{\mathcal {Y}}}})q|_{{{\mathcal {Y}}}}^*H^{d-2}}{r_i}\\&\qquad -\frac{N}{\mathop {\text {rk}}(E)}\sum _{i<j}r_ir_j\left( \frac{c_1(F_i|_{{{\mathcal {Y}}}})}{r_i}- \frac{c_1(F_j|_{{{\mathcal {Y}}}})}{r_j}\right) ^2\cdot q|_{{{\mathcal {Y}}}}^*H^{d-2}\\&\quad \geqslant \frac{N}{\mathop {\text {rk}}(E)}\sum _{i<j}r_ir_j\left( N\left( \frac{b_i}{r_i}- \frac{b_j}{r_j}\right) ^2-\left( \frac{{{\mathcal {M}}}_i}{r_i}-\frac{{{\mathcal {M}}}_j}{r_j}\right) ^2H^{d-2}\right) \\&\quad \geqslant \frac{1}{\mathop {\text {rk}}(E)}\sum _{i<j}r_ir_j\left( (N)^2\left( \frac{b_i}{r_i}- \frac{b_j}{r_j}\right) ^2-\left( \frac{{{\mathcal {M}}}_iH^{d-1}}{r_i}- \frac{{{\mathcal {M}}}_jH^{d-1}}{r_j}\right) ^2\right) . \end{aligned}$$

The last inequality is from Hodge index theorem Lemma A.5 for smooth Deligne–Mumford stacks, and from the slope \(\mu _{\** ,i},\) the last expression above gives

$$\begin{aligned} 2\sum _{i}N b_i\mu _{\** , i}-\frac{1}{\mathop {\text {rk}}(E)}\sum _{i<j}r_i r_j (\mu _{\** ,i}-\mu _{\** ,j})^2. \end{aligned}$$

To prove the claim, first \((q|_{{{\mathcal {Y}}}})_*(E_i|_{{{\mathcal {Y}}}})\subset E\) implies that

$$\begin{aligned} \frac{\sum _{j\leqslant i}{{\mathcal {M}}}_jH^{d-1}}{\mathop {\text {rk}}(\** )\sum _{j\leqslant i}r_j}\leqslant \mu _{\** , \max }(E) \end{aligned}$$

which gives the inequality:

$$\begin{aligned} \sum _{j\leqslant i}b_j N\geqslant \sum _{j\leqslant i}\mathop {\text {rk}}(\** )r_j(\mu _{\** , j}-\mu _{\** , \max }(E)). \end{aligned}$$
(A.3)

Therefore

$$\begin{aligned}&\sum _{i}Nb_i \mu _{\** ,i}=\sum _{i}\left( \sum _{i<j}Nb_j\right) (\mu _{\** ,i}-\mu _{\** ,i+1})\\&\quad \geqslant \sum _{i}\left( \sum _{j\leqslant i}\mathop {\text {rk}}(\** )r_j(\mu _{\** , j}- \mu _{\** , \max }(E))\right) (\mu _{\** ,i}-\mu _{\** ,i+1})\\&\quad =\mathop {\text {rk}}(\** )\cdot \sum _{i<j}\frac{r_ir_j}{\mathop {\text {rk}}(E)} (\mu _{\** ,i}-\mu _{\** ,j})^2+\mathop {\text {rk}}(E)(\mu _{\** }(E)- \mu _{\** ,\max }(E))(\mu _{\** }(E)-\mu _{\** ,\min }(E)). \end{aligned}$$

So we get:

$$\begin{aligned}&\frac{N\Delta (E)q|_{{{\mathcal {Y}}}}^*H^{d-2}}{\mathop {\text {rk}}(E)}\\&\quad \geqslant \sum _{i<j}\frac{2\mathop {\text {rk}}(\** )-1}{\mathop {\text {rk}}(E)}r_i r_j(\mu _{\** ,i}-\mu _{\** ,j})^2+2\mathop {\text {rk}}(E)(\mu _{\** }(E)- \mu _{\** ,\max }(E))(\mu _{\** }(E)-\mu _{\** ,\min }(E)). \end{aligned}$$

Lemma A.6

Let \(\pi : \mathcal {X}\rightarrow X\) be the coarse moduli associated to a tame smooth projective Deligne–Mumford stack \(\mathcal {X}\) over an algebraically closed field k of dimension greater than 2. Let |D| be the complete linear system associated to a very ample divisor on X,  and let B be the base locus of a general pencil associated to |D|,  then \({{\mathcal {B}}}:=B\times _{X} \mathcal {X}\) is smooth and connected substack of \(\mathcal {X}.\)

Proof

Let \(p:Y\rightarrow \mathcal {X}\) be a projective smooth scheme with p being faithfully flat and finite from [14, Theorem 2.1] and \(f:Y\rightarrow X\) be the composition \(\pi \circ p,\) which is finite. We note that \(f^{*}D\) is a base point free divisor on Y since f is finite. Then if \(\Lambda \) be a sufficiently general pencil associated to |D| with base locus B then \(f^{-1}(\Lambda )\) is a general pencil associated with the base point free linear system \(|f^{*}D|\) with base locus given by \(f^{-1}(B)\) which from Bertini’s Theorem on smooth projective varieties is smooth and connected. Thus \(f:f^{-1}(B)\rightarrow B\) being surjective we obtain B is connected and \({{\mathcal {B}}}\) being the stack with coarse moduli B we have the connectedness of \({{\mathcal {B}}}.\)

To prove smoothness of \({{\mathcal {B}}}\) we note that \(p:f^{-1}(B)\rightarrow {{\mathcal {B}}}\) is a faithfully flat and finite morphism and hence we conclude the smoothness of \({{\mathcal {B}}}\) from the descent of smoothness of \(f^{-1}(B)\) along p, see [27]. \(\square \)

1.6 A.6. Restriction theorems

We generalize the restriction theorem of Langer to smooth Deligne–Mumford stacks in higher dimensions. We give a general statement for the (A.2). Recall from Corollary 2.10,

$$\begin{aligned} \alpha _{\** }(E):=\max (L_{\** ,\max }(E)-\mu _{\** , \max }(E), \mu _{\** , \min }(E)-L_{\** ,\min }(E)). \end{aligned}$$

Let A be a nef divisor for X such that \(\pi _*(T_{\mathcal {X}}\otimes \** )(A)\) is globally generated and a \(\pi \) ample line bundle \({{\mathcal {L}}},\) then we have,

$$\begin{aligned} \alpha _{\** }(E)\leqslant \frac{\mathop {\text {rk}}(E)-1}{p-1}\left( max_{1\leqslant k\leqslant m}\mu _{\** }(A\otimes {{\mathcal {L}}}^{k})\right) . \end{aligned}$$

See Corollary 2.10 for the details of this inequality.

Theorem A.7

Let E be a torsion free sheaf on a smooth Deligne–Mumford stack \(\mathcal {X}.\) Then we have

$$\begin{aligned}{} & {} H^{d}\cdot \Delta (E)H^{d-2}+\mathop {\text {rk}}(E)^2\mathop {\text {rk}}(\** )^2(L_{\** ,\max }(E)\nonumber \\{} & {} \quad -\mu _{\** }(E))(\mu _{\** }(E)-L_{\** ,\min }(E))\geqslant 0 \end{aligned}$$
(A.4)

and

$$\begin{aligned}{} & {} H^{d}\cdot \Delta (E)H^{d-2}+\mathop {\text {rk}}(E)^2\mathop {\text {rk}}(\** )^2(\mu _{\** ,\max }(E)\nonumber \\{} & {} \quad -\mu _{\** }(E))(\mu _{\** }(E)-\mu _{\** ,\min }(E))+\beta _{r}\geqslant 0. \end{aligned}$$
(A.5)

Proof

The proof of (A.4) is the same as in Claim (A.2). The proof of formula (A.5) is the same as [15, Theorem 5.1]. \(\square \)

Theorem A.8

Let E be a torsion free sheaf of rank \(\mathop {\text {rk}}(E)\geqslant 2\) on a smooth Deligne–Mumford stack \(\mathcal {X}.\) Suppose that E is slope modified stable with respect to \((\** , {{\mathcal {O}}}_{X}(1)=H).\) Let \(D\subset |mH|\) be a normal divisor such that \(E|_{{{\mathcal {D}}}}\) has no torsion where \({{\mathcal {D}}}=\pi ^{-1}(D)).\) If

$$\begin{aligned} m>\lfloor \frac{\mathop {\text {rk}}(E)-1}{\mathop {\text {rk}}(E)\mathop {\text {rk}}(\** )}\Delta (E)H^{d-2}+\frac{1}{H^d \mathop {\text {rk}}(E)\mathop {\text {rk}}(\** )(\mathop {\text {rk}}(E)-1)}+\frac{(\mathop {\text {rk}}(E)-1)\beta _{r}}{H^d \mathop {\text {rk}}(E)\mathop {\text {rk}}(\** )} \rfloor \end{aligned}$$

then \(E|_{{{\mathcal {D}}}}\) is slope modified stable with respect to \((\** |_{{{\mathcal {D}}}}, H|_{{{\mathcal {D}}}}).\)

Proof

The proof is the same as [15, Theorem 5.2]. \(\square \)

Appendix B: Bogomolov’s inequality for Higgs sheaves

By Yunfeng Jiang, Promit Kundu and Hao (Max) Sun

Let k be an algebraically closed field of characteristic \(p\geqslant 0\) and \({\mathcal {X}}\) be a smooth tame projective Deligne–Mumford stack of dimension d over k with coarse moduli space \(\pi :{\mathcal {X}}\rightarrow X.\) Let \(\** \) be a generating sheaf on \({\mathcal {X}}\) satisfying Condition \(\star \) in Sect. 3.1, and let H an ample divisor on X.

Definition B.1

A Higgs sheaf \((E, \theta )\) is a pair consisting of a coherent sheaf \(E\in \hbox {Coh}({\mathcal {X}})\) and an \({\mathcal {O}}_{{\mathcal {X}}}\)-homomorphism \(\theta : E\rightarrow E\otimes \Omega _{{\mathcal {X}}}\) satisfying the integrability condition \(\theta \wedge \theta =0.\) We say a Higgs sheaf \((E, \theta )\) a system of Hodge sheaves if there is a decomposition \(E=\oplus E^i\) such that \(\theta : E^i\rightarrow E^{i-1}\otimes _{{\mathcal {O}}_{{\mathcal {X}}}}\Omega _{{\mathcal {X}}}\)

  1. (1)

    We say that \((E, \theta )\) is slope semistable if \(\mu _{\** }(E^{\prime })\leqslant \mu _{\** }(E)\) for every Higgs subsheaf \((E^{\prime },\theta ^{\prime })\) of \((E,\theta ),\) which means that \(E^{\prime }\subset E\) is a subsheaf and \(\theta ^{\prime }: E^{\prime }\rightarrow E^{\prime }\otimes \Omega _{\mathcal {X}}\) is preserved by \(\theta .\)

  2. (2)

    A system of Hodge sheaves \((E, \theta )\) is slope semistable if the inequality \(\mu _{\** }(E^{\prime })\leqslant \mu _{\** }(E)\) is satisfied for every subsystem of Hodge sheaves \((E^{\prime },\theta ^{\prime })\) of \((E,\theta ).\) Here a subsystem means that \(E^{\prime }\subset E\) is a subsheaf, and \(E^{\prime }=\oplus E^{\prime i}\) such that \(\theta ^{\prime }: E^{\prime i}\rightarrow E^{\prime i}\otimes _{\mathcal {X}}\Omega _{\mathcal {X}}\) is preserved by \(\theta .\)

We recall the main results of Ogus and Vologodsky [24], where the theory is for schemes, but in étale topology it works for Deligne–Mumford stacks.

Assume that \(p>0.\) Let S be a scheme over k and \(f:{\mathcal {X}}\rightarrow S\) be a morphism of stacks over k. A lifting of \({\mathcal {X}}/S\) modulo \(p^2\) is a morphism \({\widetilde{f}}: \widetilde{{\mathcal {X}}}\rightarrow {\widetilde{S}}\) of flat \({\mathbb {Z}}/p^2{\mathbb {Z}}\)-stacks such that f is the base change of \({\widetilde{f}}\) by the closed embedding \(S\rightarrow {\widetilde{S}}\) defined by p. Let \(\hbox {MIC}_{p-1}({\mathcal {X}}/S)\) be the category of \({\mathcal {O}}_{{\mathcal {X}}}\)-modules with an integrable connection whose p-curvature is nilpotent of level \(\leqslant p-1.\) Here we say a \({\mathcal {O}}_{\mathcal {X}}\)-modules V with integrable connection such that the p-curvature \(\Psi : T_{\mathcal {X}/S}\rightarrow {\mathcal {E}}nd_S(V)\) is nilpotent of level \(\leqslant p-1\) if for all open subsets \(U\subset \mathcal {X}\) and for all derivations \(D\in T_{\mathcal {X}/S}(U)\) we have \(\Psi (D)^p=0.\)

Let \(\hbox {HIG}_{p-1}({\mathcal {X}}^{(1)}/S)\) denote the category of Higgs \({\mathcal {O}}_{{\mathcal {X}}^{(1)}}\)-modules with a nilpotent Higgs field of level \(\leqslant p-1,\) which means that \(\theta ^p=0.\) We have the following theorem of Ogus and Vologodsky [24, Theorem 2.8].

Theorem B.2

If \(f:{\mathcal {X}}\rightarrow S\) is a smooth morphism with a lifting \(\widetilde{{\mathcal {X}}}^{(1)}\rightarrow {\widetilde{S}}\) of \({\mathcal {X}}^{(1)}\rightarrow S\) modulo \(p^2,\) then the Cartier transform

$$\begin{aligned} C_{{\mathcal {X}}/S}:\hbox {MIC}_{p-1}({\mathcal {X}}/S) \rightarrow \hbox {HIG}_{p-1}({\mathcal {X}}^{(1)}/S) \end{aligned}$$

defines an equivalence of categories with quasi-inverse

$$\begin{aligned} C^{-1}_{{\mathcal {X}}/S}:\hbox {HIG}_{p-1}({\mathcal {X}}^{(1)}/S) \rightarrow \hbox {MIC}_{p-1}({\mathcal {X}}/S). \end{aligned}$$

Lemma B.3

Let \((E,\theta )\in \hbox {HIG}_{p-1}({\mathcal {X}}^{(1)}/S).\) Then we have \([C^{-1}_{{\mathcal {X}}/S}(E)]=F_g^*[E],\) where \([\cdot ]\) denotes the class of a coherent sheaf in the Grothendieck group \(K_0({\mathcal {X}}).\)

Proof

See [16, Lemma 2]. \(\square \)

Corollary B.4

Assume \(S=\mathop {\text {Spec}}\nolimits k,\) and let \((E,\theta )\in \hbox {HIG}_{p-1}({\mathcal {X}}^{(1)}/S).\) Then \((E, \theta )\) is slope semistable with respect to \((H,\** )\) iff \(C^{-1}_{{\mathcal {X}}/S}(E)\) is slope \(\nabla \)-semistable with respect to \((F_g^*H,F_g^*\** ).\)

Proof

The proof is the same as that of [16, Corollary 1]. \(\square \)

Lemma B.5

Let \((E, \theta )\) be a torsion free slope semistable Higgs sheaf on \({\mathcal {X}}.\) Then there exists an \({\mathbb {A}}^1\)-flat family of Higgs sheaves \(({\widetilde{E}}, {\widetilde{\theta }})\) on \({\mathcal {X}}\times {\mathbb {A}}^1\) such that the restriction \(({\widetilde{E}}_t,{\widetilde{\theta }}_t)\) to the fiber over any closed point \(t\in {\mathbb {A}}^1\) is isomorphic to \((E,\theta )\) and \((E_0, \theta _0)\) is a slope semistable system of Hodge sheaves.

Proof

See [17, Corollary 5.7]. \(\square \)

Proposition B.6

Let V be a torsion free sheaf on \({\mathcal {X}},\) then

$$\begin{aligned} H^{d-2}\Delta (V) \geqslant -\frac{(\mathop {\text {rank}}\nolimits V\mathop {\text {rank}}\nolimits \** )^2}{H^d}(L_{\** ,\max }(V)-\mu _{\** }(V))(\mu _{\** }(V)-L_{\** ,\min }(V)). \end{aligned}$$

Proof

The proposition follows from Theorem A.2 by the same arguments as in the proof of Theorem A.3. \(\square \)

Theorem B.7

Assume \(p=0.\) Let \((E, \theta )\) be a slope semistable Higgs sheaf with respect to \((H,\** ).\) Then we have \(H^{d-2}\Delta (E)\geqslant 0.\)

Proof

Deforming \((E,\theta )\) to a system of Hodge sheaves (see Lemma B.5) we can assume that \((E, \theta )\) is nilpotent. This is because that if \((E=\oplus _{j=1}^n E^j, \theta )\) is the Hodge sheaf. From \(\theta : E^j\rightarrow E^{j-1}\otimes \Omega _{\mathcal {X}},\) one sees that \(\theta ^n=0,\) i.e., \((E, \theta )\) is nilpotent. Now we use the standard reduction to positive characteristic technique, which we recall for the convenience of the reader (see [21, section 2]). There exists a finitely generated \({\mathbb {Z}}\)-algebra \(R\subset k\) and a tame smooth Deligne–Mumford stack \(\widetilde{{\mathcal {X}}}\rightarrow S=\mathop {\text {Spec}}\nolimits R\) such that \({\mathcal {X}}=\widetilde{{\mathcal {X}}}\times _S\mathop {\text {Spec}}\nolimits k.\) Let \({\widetilde{\pi }}: \widetilde{{\mathcal {X}}}\rightarrow {\widetilde{X}}\) be its coarse moduli space. We can assume that \(X={\widetilde{X}}\times _S\mathop {\text {Spec}}\nolimits k,\) \(\pi \) is induced by \({\widetilde{\pi }}\) after base change, and there exists an ample divisor \({\widetilde{H}}\) on \({\widetilde{X}}\) extending H and a generating sheaf \({\widetilde{\** }}\) on \(\widetilde{{\mathcal {X}}}\) extending \(\** \) such that it is a direct sum of \(\pi \) ample vector bundles obtained from direct sum of line bundles. Since we have the Higgs sheaf on the generic fiber, which can be extended to a general fiber. Thus, we can also assume there exists an S-flat family of Higgs sheaves \(({\widetilde{E}}, {\widetilde{\theta }})\) on \(\widetilde{{\mathcal {X}}}\) extending \((E, \theta ).\)

Shrinking S,  by openness of semistability we can assume that \(({\widetilde{E}}_s, {\widetilde{\theta }}_s)\) is slope semistable with respect to \(({\widetilde{\** }}_s, {\widetilde{H}}_s)\) for any \(s\in S.\) Choose a closed point \(s\in S\) such that the characteristic q of the residue field k(s) is \(\geqslant \mathop {\text {rank}}\nolimits E.\) We can choose such a closed point from the following reason. For a big enough prime number q,  one takes the prime ideal generated by q in R. Then one sees that the characteristic of the residue field of that ideal is q. Then the stack \(\widetilde{{\mathcal {X}}}\times _S\mathop {\text {Spec}}\nolimits (R/m_s^2)\) is a lifting of \(\widetilde{{\mathcal {X}}}_s\) modulo \(q^2.\) By Corollary B.4, one can associate to \(({\widetilde{E}}_s, {\widetilde{\theta }}_s)\) a slope \(\nabla \)-semistable sheaf with integrable connection \((V_s, \nabla _s)\) with respect to \((F_g^*{\widetilde{H}}_s,F_g^*{\widetilde{\** }}_s).\) By Lemma B.3, one sees that \([V_s]=[C^{-1}_{X/S}(E_s)]=[F^*_g(E_s)].\) Hence we have \(ch_0(V_s)=ch_0(E_s),\) \(ch_1(V_s)=q ch_1(E_s)\) and \(ch_2(V_s)=q^2 ch_2(E_s).\) These imply the following equality,

$$\begin{aligned} {\widetilde{H}}_s^{d-2}\Delta (V_s)=q^2{\widetilde{H}}_s^{d-2}\Delta ({\widetilde{E}}_s). \end{aligned}$$
(B.1)

Let \(0=V_0\subset V_1\subset \cdots \subset V_m=V_s\) be the usual Harder–Narasimhan filtration of \(V_s,\) then by [9, Lemma 2.7], the induced morphisms \(V_i\rightarrow (V_s/V_i)\otimes \Omega _{\widetilde{{\mathcal {X}}}_s}\) are nonzero \({{\mathcal {O}}}_{\widetilde{{\mathcal {X}}}_s}\)-morphisms. Take a nef divisor A on \({\widetilde{X}}_s\) such that \(\pi _*(T_{\widetilde{{\mathcal {X}}}_s}\otimes {\widetilde{\** }}_s)(A)\) is globally generated. From [9, Proposition 2.10 and Corollary 2.11], it follows that

$$\begin{aligned} \mu _{{\widetilde{\** }}_s, \max }(V_s)-\mu _{{\widetilde{\** }}_s, \min }(V_s)\leqslant (\mathop {\text {rank}}\nolimits V_s-1)\left( \frac{{\widetilde{H}}_s^{d-1}A}{{\widetilde{H}}_s^d}+M\right) \end{aligned}$$

and

$$\begin{aligned} \max (L_{{\widetilde{\** }}_s, \max }(V_s)-\mu _{{\widetilde{\** }}_s, \max }(V_s), \mu _{{\widetilde{\** }}_s,\min }(V_s)-L_{{\widetilde{\** }}_s, \min }(V_s)) \leqslant \frac{\mathop {\text {rank}}\nolimits V_s-1}{q-1}\left( \frac{{\widetilde{H}}_s^{d-1}A}{{\widetilde{H}}_s^d}+M\right) , \end{aligned}$$

for some constant M depending on A and \(\** .\) They imply that

$$\begin{aligned} (L_{{\widetilde{\** }}_s,\max }(V_s)-L_{{\widetilde{\** }}_s,\min }(V_s))\leqslant \frac{\mathop {\text {rank}}\nolimits V_s-1}{1-\frac{1}{q}}\left( \frac{{\widetilde{H}}_s^{d-1}A}{{\widetilde{H}}_s^d}+M\right) . \end{aligned}$$

Hence Proposition B.6 gives

$$\begin{aligned} q^2{\widetilde{H}}_s^{d-2}\Delta ({\widetilde{E}}_s)={\widetilde{H}}_s^{d-2}\Delta (V_s) \geqslant -\frac{(\mathop {\text {rank}}\nolimits V_s\mathop {\text {rank}}\nolimits \** )^2}{{\widetilde{H}}_s^d}(\frac{\mathop {\text {rank}}\nolimits V_s-1}{1-\frac{1}{q}})^2\left( \frac{{\widetilde{H}}_s^{d-1}A}{{\widetilde{H}}_s^d}+M\right) ^2. \end{aligned}$$

Taking sufficiently large q,  one obtains

$$\begin{aligned} H^{d-2}\Delta (E)={\widetilde{H}}_s^{d-2}\Delta ({\widetilde{E}}_s)\geqslant 0. \end{aligned}$$

\(\square \)

Using the same argument in [16, section 6], one can recover [4, Theorem 1.1]:

Theorem B.8

Assume that \(p=0,\) \(d=2,\) and the canonical line bundle \(K_{{\mathcal {X}}}\) is nef,  then we have

$$\begin{aligned} c^2_1(T_{{\mathcal {X}}})\leqslant 3c_2(T_{{\mathcal {X}}}). \end{aligned}$$

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jiang, Y., Kundu, P. On the Bogomolov–Gieseker inequality for tame Deligne–Mumford surfaces. Math. Z. 306, 32 (2024). https://doi.org/10.1007/s00209-023-03421-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00209-023-03421-4

Navigation