Skip to main content
Log in

Motivic degree zero Donaldson–Thomas invariants

  • Published:
Inventiones mathematicae Aims and scope

Abstract

Given a smooth complex threefold X, we define the virtual motive \([\operatorname{Hilb}^{n}(X)]_{\operatorname {vir}}\) of the Hilbert scheme of n points on X. In the case when X is Calabi–Yau, \([\operatorname{Hilb}^{n}(X)]_{\operatorname{vir}}\) gives a motivic refinement of the n-point degree zero Donaldson–Thomas invariant of X. The key example is X=ℂ3, where the Hilbert scheme can be expressed as the critical locus of a regular function on a smooth variety, and its virtual motive is defined in terms of the Denef–Loeser motivic nearby fiber. A crucial technical result asserts that if a function is equivariant with respect to a suitable torus action, its motivic nearby fiber is simply given by the motivic class of a general fiber. This allows us to compute the generating function of the virtual motives \([\operatorname{Hilb}^{n} (\mathbb{C}^{3})]_{\operatorname{vir}}\) via a direct computation involving the motivic class of the commuting variety. We then give a formula for the generating function for arbitrary X as a motivic exponential, generalizing known results in lower dimensions. The weight polynomial specialization leads to a product formula in terms of deformed MacMahon functions, analogous to Göttsche’s formula for the Poincaré polynomials of the Hilbert schemes of points on surfaces.

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

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. Note that the operator \(\operatorname{Exp}\) depends on the variable t. In particular, one cannot simply substitute t for T in the above equation for Z X . See Remark 2.8.

  2. One can also consider the rings K 0(Sch ) and K 0(Sp ) generated by schemes or algebraic spaces (of finite type) with the same relations. By [5, Lemma 2.12], they are all the same. In particular, for X a scheme, its class in \(K_{0} (\operatorname {Var}_{\mathbb{C}})\) is given by \([X_{\operatorname{red}}]\), the class of the associated reduced scheme. We will implicitly use this identification throughout the paper without further comment.

  3. We thank Sven Meinhardt for calling our attention to this sign issue.

  4. We thank Sheldon Katz for discussions on this issue.

  5. We thank Patrick Brosnan and Jörg Schürmann for very helpful correspondence on this subject.

  6. We thank Lothar Göttsche, Ezra Getzler and Sven Meinhardt for discussions which led us to this formulation.

References

  1. Behrend, K.: Donaldson-Thomas invariants via microlocal geometry. Ann. Math. (2) 170, 1307–1338 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  2. Behrend, K., Fantechi, B.: Symmetric obstruction theories and Hilbert schemes of points on threefolds. Algebra Number Theory 2, 313–345 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  3. Białynicki-Birula, A.: Some theorems on actions of algebraic groups. Ann. Math. (2) 98, 480–497 (1973)

    Article  MATH  Google Scholar 

  4. Bittner, F.: On motivic zeta functions and the motivic nearby fiber. Math. Z. 249, 63–83 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  5. Bridgeland, T.: An introduction to motivic Hall algebras. Adv. Math. 229, 102–138 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  6. Cheah, J.: On the cohomology of Hilbert schemes of points. J. Algebr. Geom. 5, 479–511 (1996)

    MathSciNet  MATH  Google Scholar 

  7. Davison, B.: Invariance of orientation data for ind-constructible Calabi-Yau A categories under derived equivalence. D.Phil. thesis, University of Oxford (2011). arXiv:1006.5475

  8. Davison, B., Meinhardt, S.: Motivic DT-invariants for the one loop quiver with potential. arXiv:1108.5956

  9. Denef, J., Loeser, F.: Motivic Igusa zeta functions. J. Algebr. Geom. 7, 505–537 (2008)

    MathSciNet  MATH  Google Scholar 

  10. Denef, J., Loeser, F.: Motivic exponential integrals and a motivic Thom-Sebastiani theorem. Duke Math. J. 99, 285–309 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  11. Denef, J., Loeser, F.: Geometry on arc spaces of algebraic varieties. In: European Congress of Mathematics, Vol. I, Barcelona, 2000. Progr. Math., vol. 201, pp. 327–348. Birkhäuser, Basel (2001)

    Chapter  Google Scholar 

  12. Dimca, A., Szendrői, B.: The Milnor fibre of the Pfaffian and the Hilbert scheme of four points on ℂ3. Math. Res. Lett. 16, 1037–1055 (2009)

    MathSciNet  MATH  Google Scholar 

  13. Feit, W., Fine, N.J.: Pairs of commuting matrices over a finite field. Duke Math. J. 27, 91–94 (1960)

    Article  MathSciNet  MATH  Google Scholar 

  14. Getzler, E.: Mixed Hodge structures of configuration spaces. arXiv:math/9510018

  15. Göttsche, L.: The Betti numbers of the Hilbert scheme of points on a smooth projective surface. Math. Ann. 286, 193–207 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  16. Göttsche, L.: On the motive of the Hilbert scheme of points on a surface. Math. Res. Lett. 8, 613–627 (2001)

    MathSciNet  MATH  Google Scholar 

  17. Gusein-Zade, S.M., Luengo, I., Melle-Hernández, A.: A power structure over the Grothendieck ring of varieties. Math. Res. Lett. 11, 49–57 (2004)

    MathSciNet  MATH  Google Scholar 

  18. Gusein-Zade, S.M., Luengo, I., Melle-Hernández, A.: Power structure over the Grothendieck ring of varieties and generating series of Hilbert schemes of points. Mich. Math. J. 54, 353–359 (2006)

    Article  MATH  Google Scholar 

  19. Iqbal, A., Kozcaz, C., Vafa, C.: The refined topological vertex. J. High Energy Phys. 10, 069 (2009)

    Article  MathSciNet  Google Scholar 

  20. Joyce, D., Song, Y.: A Theory of Generalized Donaldson-Thomas Invariants. Memoirs of the AMS, vol. 217 (2012)

    Google Scholar 

  21. Kontsevich, M., Soibelman, Y.: Stability structures, motivic Donaldson-Thomas invariants and cluster transformations. arXiv:0811.2435

  22. Kontsevich, M., Soibelman, Y.: Cohomological Hall algebra, exponential Hodge structures and motivic Donaldson–Thomas invariants. Commun. Number Theory Phys. 5, 231–352 (2011)

    MathSciNet  MATH  Google Scholar 

  23. Levine, M., Pandharipande, R.: Algebraic cobordism revisited. Invent. Math. 176, 63–130 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  24. Li, J.: Zero dimensional Donaldson-Thomas invariants of threefolds. Geom. Topol. 10, 2117–2171 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  25. Looijenga, E.: Motivic measures. Astérisque 276, 267–297 (2002)

    MathSciNet  Google Scholar 

  26. MacMahon, P.A.: Combinatory Analysis. Chelsea Publishing Co., New York (1960)

    MATH  Google Scholar 

  27. Maulik, D., Nekrasov, N., Okounkov, A., Pandharipande, R.: Gromov-Witten theory and Donaldson-Thomas theory. I. Compos. Math. 142, 1263–1285 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  28. Morrison, A.: Computing motivic Donaldson–Thomas invariants. PhD thesis, University of British Columbia (2012)

  29. Mustonen, V., Rajesh, R.: Numerical estimation of the asymptotic behaviour of solid partitions of an integer. J. Phys. A 36, 6651–6659 (2003)

    Article  MathSciNet  Google Scholar 

  30. Nakajima, H.: Lectures on Hilbert Schemes of Points on Surfaces. American Mathematical Society, Providence (1999)

    MATH  Google Scholar 

  31. Okounkov, A., Reshetikhin, N.: Random skew plane partitions and the Pearcey process. Commun. Math. Phys. 269, 571–609 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  32. Reineke, M.: Poisson automorphisms and quiver moduli. J. Inst. Math. Jussieu 9, 653–667 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  33. Saito, M.: Modules de Hodge polarisables. Publ. Res. Inst. Math. Sci. 24, 849–995 (1989)

    Article  Google Scholar 

  34. Saito, M.: Mixed Hodge modules. Publ. Res. Inst. Math. Sci. 26, 221–333 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  35. Szendrői, B.: Non-commutative Donaldson-Thomas invariants and the conifold. Geom. Topol. 12, 1171–1202 (2008)

    Article  MathSciNet  Google Scholar 

  36. Villamayor U, O.E.: Patching local uniformizations. Ann. Sci. Éc. Norm. Super. (4) 25, 629–677 (1992)

    MathSciNet  Google Scholar 

Download references

Acknowledgements

We would like to thank D. Abramovich, T. Bridgeland, P. Brosnan, A. Dimca, B. Fantechi, E. Getzler, L. Göttsche, I. Grojnowski, D. Joyce, T. Hausel, F. Heinloth, S. Katz, M. Kontsevich, S. Kovács, E. Looijenga, S. Meinhardt, G. Moore, A. Morrison, J. Nicaise, R. Pandharipande, A. Rechnitzer, R. Thomas, M. Saito, J. Schürmann, Y. Soibelman and D. van Straten for interest in our work, comments, conversations and helpful correspondence. Some of the ideas of the paper were conceived during our stay at MSRI, Berkeley, during the Jumbo Algebraic Geometry Program in Spring 2009; we would like to thank for the warm hospitality and excellent working conditions there. J.B. thanks the Miller Institute and the Killiam Trust for support during his sabbatical stay in Berkeley. B.S.’s research was partially supported by OTKA grant K61116.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Balázs Szendrői.

Appendices

Appendix A: q-Deformations of the MacMahon function

Let \(\mathcal{P}\) denote the set of all finite 3-dimensional partitions. For a partition \(\alpha\in\mathcal{P}\), let w(α) denote the number of boxes in α. The combinatorial generating series

$$M(t) = \sum_{\alpha\in\mathcal{P}} t^{w(\alpha)} $$

was determined in closed form by MacMahon [26] to be

$$M(t)= \prod_{m=1}^\infty \bigl(1-t^m\bigr)^{-m}. $$

Motivated by work of Okounkov and Reshetikhin [31], in a recent paper [19], Iqbal–Kozçaz–Vafa discussed a family of q-deformations of this formula. Think of a 3-dimensional partition \(\alpha\in\mathcal{P}\) as a subset of the positive octant lattice ℕ3, and break the symmetry by choosing one of the coordinate directions. Define w (α),w 0(α) and w +(α), respectively, as the number of boxes (lattice points) in α∩{xy<0}, α∩{xy=0} and α∩{xy>0}. For a half-integer \(\delta\in\frac {1}{2}\mathbb{Z}\), consider the generating series

$$M_\delta(t_1, t_2) = \sum _{\alpha\in\mathcal{P}} t_1^{w_-(\alpha) + (\frac{1}{2}+\delta )w_0(\alpha)}t_2^{w_+(\alpha) + (\frac{1}{2}-\delta )w_0(\alpha)}. $$

Clearly M δ (t,t)=M(t) for all δ.

Theorem A.1

(Okounkov–Reshetikhin [31, Theorem 2])

The series M δ (t 1,t 2) admits the product form

$$M_\delta(t_1, t_2)=\prod _{i,j=1}^\infty \bigl(1-t_1^{i-\frac {1}{2}+\delta}t_2^{j-\frac{1}{2}-\delta} \bigr)^{-1}. $$

In the main body of the paper, we use a different set of variables. Namely, we set

$$t_{1} = tq^{\frac{1}{2}},\qquad t_{2} = tq^{-\frac{1}{2}}. $$

Then the product formula becomes

$$M_\delta\bigl(t,q^{\frac{1}{2}}\bigr)=\prod _{m=1}^\infty\prod_{k=0}^{m-1} \bigl(1-t^mq^{k+\frac{1}{2}-\frac{m}{2}+\delta} \bigr)^{-1}. $$

The specialization to the MacMahon function is \(M_{\delta}(t,q^{\frac{1}{2}}=1)=M(t)\) for all δ.

Appendix B: The motivic nearby fiber of an equivariant function

In this appendix we prove Proposition 2.12, which asserts that if a regular function f:X→ℂ on a smooth variety is equivariant with respect to a torus action satisfying certain assumptions, then Denef-Loeser’s motivic nearby fiber [ψ f ] is simply equal to the motivic class of the geometric fiber [f −1(1)]. To make this appendix self-contained, we recall the definitions and restate the result below.

Let

$$f\colon X\to\mathbb{C} $$

be a regular function on a smooth quasi-projective variety X, and let X 0=f −1(0) be the central fiber. Denef and Loeser define \([\psi_{f}]\in\mathcal{M}^{{\hat{\mu}}}_{\mathbb{C}}\), the motivic nearby cycle of f using arc spaces and the motivic zeta function [9, 25]. Using motivic integration, they give an explicit formula for [ψ f ] in terms of any embedded resolution which we now recall.

Let h:YX be an embedded resolution of X 0, namely Y is non-singular and

$$\widetilde{f}=h\circ f:Y\to\mathbb{C} $$

has central fiber Y 0 which is a normal crossing divisor with non-singular components {E j :jJ}. For IJ, let

$$E_{I} = \bigcap_{i\in I} E_{i} $$

and let

$$E_{I}^{o} = E_{I} - \bigcup _{j\in I^{c}} (E_{j}\cap E_{I} ). $$

By convention, E =Y and \(E^{o}_{\emptyset}=Y-Y_{0}\).

Let N i be the multiplicity of E i in the divisor \(\widetilde{f}^{-1} (0)\). Letting

$$m_{I} = gcd (N_{i})_{i\in I}, $$

there is a natural etale cyclic \(\mu_{m_{I}}\)-cover

$$\widetilde{E}^{o}_{I}\to E^{o}_{I}. $$

The formula of Denef and Loeser for the (absolute) motivic nearby cycles of f is given by

$$ [\psi_f]= \sum_{I \neq\emptyset} (1-\mathbb{L})^{|I|-1}\bigl[ \widetilde{E}^{o}_{I},\mu_{m_{I}}\bigr]\in \mathcal{M}_{\mathbb{C}}^{{\hat{\mu}}}. $$
(B.1)

Since all the \(E_{I}^{o}\) appearing in the above sum have natural maps to X 0, the above formula determines the relative motivic nearby cycle \([\psi_{f}]_{X_{0}}\in\mathcal{M}^{{\hat{\mu}}}_{X_{0}}\). The relative motivic vanishing cycle is supported on Z={df=0}, the degeneracy locus of f:

$$[\varphi_{f}]_{Z} = [\psi_{f}]_{X_{0}}-[X_{0}]_{X_{0}} \in\mathcal{M}_{Z}^{{\hat{\mu}}} \subset\mathcal{M}_{X_{0}}^{{\hat{\mu}}}. $$

Recall that an action of ℂ on a variety V is circle compact, if the fixed point set \(V^{\mathbb{C}^{*}}\) is compact and moreover, for all vV, the limit lim t→0 ty exists.

The following is a restatement of Proposition 2.12 and Proposition 2.13. It is the main result of this appendix.

Theorem B.1

Let f:X→ℂ be a regular morphism on a smooth quasi-projective complex variety. Let Z={df=0} be the degeneracy locus of f and let \(Z_{\operatorname{aff}}\subset X_{\operatorname{aff}}\) be the affinization of Z and X respectively. Assume that there exists an action of a connected complex torus T on X so that f is T-equivariant with respect to a primitive character χ:T→ℂ, namely f(tx)=χ(t)f(x) for all xX and tT. We further assume that there exists a one parameter subgroupT such that the induced action is circle compact. Then the motivic nearby cycle class [ψ f ] is in \(\mathcal{M}_{\mathbb{C}}\subset\mathcal{M}^{{\hat{\mu}}}_{\mathbb{C}}\) and is equal to [X 1]=[f −1(1)]. Consequently the motivic vanishing cycle class [φ f ] is given by

$$[\varphi_{f}] = \bigl[f^{-1} (1)\bigr] - \bigl[f^{-1} (0)\bigr]. $$

If we further assume that X 0 is reduced then \([\varphi_{f}]_{Z_{\operatorname{aff}}}\), the motivic vanishing cycle, considered as a relative class on \(Z_{\operatorname{aff}}\), lies in the subring \(\mathcal{M}_{Z_{\operatorname{aff} }}\subset\mathcal{M}_{Z_{\operatorname{aff}}}^{{\hat{\mu}}}\).

By equivariant resolution of singularities [36, Corollary 7.6.3], we may assume that h:YX, the embedded resolution of X 0, is T-equivariant. Namely Y is a non-singular T-variety and

$$\widetilde{f}=h\circ f:Y\to\mathbb{C} $$

is T-equivariant with central fiber E which is a normal crossing divisor with non-singular components E j , jJ. Let ℂT be the one-parameter subgroup whose action on X is circle compact. Then the action of ℂ on Y is circle compact (since h is proper) and each E j is invariant (but not necessarily fixed).

We will make use of the Białynicki-Birula decomposition for smooth varieties [3]. This result states that if V is a smooth projective variety with a ℂ-action, then there is a locally closed stratification:

$$V=\bigcup_{F}Z_{F} $$

where the union is over the components of the fixed point locus and Z F F is a Zariski locally trivial affine bundle. The rank of the affine bundle Z F F is given by

$$n (F) = \operatorname{index}(N_{F/V}) $$

where the index of the normal bundle N F/V is the number of positive weights of the fiberwise action of ℂ. The morphisms Z F F are defined by x↦lim t→0 tx and consequently, the above stratification also exists for smooth varieties with a circle compact action. As a corollary of the Białynicki-Birula decomposition, we get the following relation in the ring of motivic weights.

Lemma B.2

Let V be a smooth quasi-projective variety with a circle compact-action. For each component F of the fixed point locus, we define the index of F, denoted by n(F), to be the number of positive weights in the action of on N F/V . Then in \(\mathcal{M}_{\mathbb{C}}\) we have

$$[V] = \sum_{F} \mathbb{L}^{n (F)} [F], $$

where the sum is over the components of the fixed point locus and \(\mathbb{L}=[\mathbb{A}^{1}_{\mathbb{C}}]\) is the Lefschetz motive.

We call this decomposition a BB decomposition.

We begin our proof of Theorem B.1 with a “no monodromy” result.

Lemma B.3

The following equation holds in \(\mathcal{M}_{\mathbb{C}}^{{\hat{\mu}}}\):

$$\bigl[\widetilde{E}^{o}_{I},\mu_{m_{I}}\bigr] = \bigl[E_{I}^{o}\bigr]. $$

Under the further assumption that X 0 is reduced, the above equation holds in \(\mathcal{M}_{X_{\operatorname {aff}}}^{{\hat{\mu}}}\).

An immediate corollary of Lemma B.3 is that [ψ f ] lies in the subring \(\mathcal{M}_{\mathbb{C}}\subset\mathcal {M}_{\mathbb{C}}^{{\hat{\mu}}}\) and that if X 0 is reduced, then \([\varphi_{f}]_{Z_{\operatorname {aff}}}\) lies in the subring \(\mathcal{M}_{Z_{\operatorname{aff}}}\subset\mathcal {M}_{Z_{\operatorname{aff}}}^{{\hat{\mu}}}\).

To prove Lemma B.3, we recall the construction of the \(\mu_{m_{I}}\)-cover \(\widetilde{E}_{I}^{o}\to E^{o}_{I}\) given in [25, §5]. Let N=lcm(N i ) and let \(\widetilde {Y}\to Y\) be the μ N -cover obtained by base change over the Nth power map (⋅)N:ℂ→ℂ followed by normalization. Define \(\widetilde{E}^{o}_{I}\) to be any connected component of the preimage of \(E^{o}_{I}\) in \(\widetilde{Y}\). The component \(\widetilde{E}^{o}_{I}\) is stabilized by \(\mu_{m_{I}}\subset \mu_{N}\) whose action defines the cover \(\widetilde{E}^{o}_{I}\to E^{o}_{I}\).

Observe that the composition

$$E_{i} \to Y \to X \to X_{\operatorname{aff}} $$

contracts E i to a point unless E i is a component of the proper transform of X 0. Thus for these components (and their intersections), the proof given below (stated for absolute classes), applies to relative classes over \(X_{\operatorname{aff}}\) as well. Under the assumption that X 0 is reduced, those E i which are components of the proper transform of X 0 have multiplicity one and so \(\widetilde{E}_{i}=E_{i}\) and there is nothing to prove.

We define \(\widetilde{T}\) by the fibered product

Thus \(\widetilde{T}\) is an extension of T by μ N ⊂ℂ and it has character \(\widetilde{\chi}\) satisfying \(\widetilde{\chi }^{N}=\chi\). Moreover, \(\widetilde{\chi}\) is the identity on the subgroup \(\mu_{N}\subset\widetilde{T}\). The key fact here is that since χ is primitive, \(\widetilde{T}\) is connected.

By construction, \(\widetilde{T}\) acts on the base change of Y over the Nth power map and hence it acts on the normalization \(\widetilde {Y}\). Thus we have obtained an action of a connected torus \(\widetilde {T}\) on \(\widetilde{Y}\) covering the T-action on Y. Since T acts on each \(E^{o}_{I}\), \(\widetilde{T}\) acts on each component of the preimage of \(E^{o}_{I}\) in \(\widetilde{Y}\). Thus we have an action of the connected torus \(\widetilde{T}\) on \(\widetilde{E}^{o}_{I}\) such that the \(\mu_{m_{I}}\)-action is induced by the subgroup

$$\mu_{m_{I}}\subset \mu_{N}\subset\widetilde{T}. $$

Lemma B.3 then follows from the following:

Lemma B.4

Let W be a smooth quasi-projective variety with the action of a connected torus \(\widetilde{T}\). Then for any finite cyclic subgroup \(\mu \subset\widetilde{T}\), the equation

$$[W,\mu] = [W] = [W/\mu] $$

holds in \(\mathcal{M}^{{\hat{\mu}}}_{\mathbb{C}}\).

Proof

Let \(\mathbb{C}^{*}\subset\widetilde{T}\) be the 1-parameter subgroup generated by \(\mu\subset\widetilde{T}\). The ℂ-action on W gives rise to a ℂ-equivariant stratification of W into varieties W i of the form (V−{0})×F where F is fixed and V is a ℂ-representation. This assertion follows from applying the Białynicki-Birula decomposition to \(\overline{W}\), any ℂ-equivariant smooth compactification of W and stratifying further to trivialize all the bundles and to make the zero sections separate strata. The induced stratification of W then is of the desired form. Thus to prove Lemma B.4, it then suffices to prove it for the case of μ acting on V−{0} where V is a μ-representation. By the relation given in (2.4), we have [V,μ]=[V] and hence [V−{0},μ]=[V−{0}]. The equality [V−{0}]=[(V−{0})/μ] follows from [25, Lemma 5.1].

Applying Lemma B.4 to (B.1), we see that in order to prove [ψ f ]=[X 1], we must prove

$$[X_{1}] = \sum_{I \neq\emptyset} (1- \mathbb{L})^{|I|-1} \bigl[E^{o}_{I}\bigr]. $$

As explained in the beginning of Sect. 2.7, there is an isomorphism

$$X_{1}\times\mathbb{C}^{*} \cong X-X_{0}. $$

Consequently we get \((\mathbb{L}-1)[X_{1}] = [X]-[X_{0}]\) or equivalently

$$[Y] = [Y_{0}] + (\mathbb{L}-1) [X_{1}]. $$

Combining this with the previous equation we find that the equation we wish to prove, [ψ f ]=[X 1], is equivalent to

$$[Y] = [Y_{0}] - \sum_{I\neq\emptyset} (1- \mathbb{L})^{|I|} \bigl[E_{I}^{o}\bigr], $$

which can also be written (using the conventions about the empty set) as

$$ 0=\sum_{I} (1- \mathbb{L})^{|I|} \bigl[E_{I}^{o}\bigr]. $$
(B.2)

By the principle of inclusion/exclusion, we can write

thus we have

$$(-1)^{|I|}\bigl[E^{o}_{I}\bigr] = \sum _{A\supset I} (-1)^{|A|}[E_{A}]. $$

Thus the right hand side of (B.2) becomes

$$\sum_{I} (\mathbb{L}-1)^{|I|} \sum _{A\supset I} (-1)^{|A|} [E_{A}] = \sum_{A} (-1)^{|A|} [E_{A}] \sum_{I\subset A} (\mathbb{L}-1)^{|I|}. $$

Since

$$\sum_{I\subset A} (\mathbb{L}-1)^{|I|} =\sum _{n=0}^{|A|} \binom{|A|}{n} (\mathbb{L} -1)^{n} = \mathbb{L}^{|A|}, $$

we can reformulate the equation we need to prove as

$$ 0 = \sum_{A} (- \mathbb{L})^{|A|} [E_{A}]. $$
(B.3)

Note that each E A is smooth and ℂ invariant and that the induced ℂ-action on E A is circle compact, so we have a BB decomposition for each.

Let F denote a component of the fixed point set \(Y^{\mathbb{C}^{*}}\) and let θ F,A denote a component of FE A . Since E A is smooth and ℂ invariant and the induced ℂ-action is circle compact, E A admits a BB decomposition

$$[E_{A}] = \sum_{F}\sum _{\theta_{F,A}} \mathbb{L}^{n (\theta_{F,A})} [\theta_{F,a}] $$

where

$$n (\theta_{F,A}) = \operatorname{index}(N_{\theta_{F,A}/E_{A}}) . $$

Therefore the sum in (B.3) (which we wish to prove is zero) is given by

$$\sum_{A} (-\mathbb{L})^{|A| } [E_{A}] = \sum_{A}\sum _{F}\sum_{\theta _{F,A}} (-1)^{|A|}\mathbb{L}^{|A|+n (\theta_{F,A})} [ \theta_{F,A}]. $$

We define a set I(F) by

$$I (F) = \{i: F\subset E_{i} \}. $$

Then clearly

$$\theta_{F,A} = \theta_{F,A'}\quad\text{if } A\cup I (F) =A'\cup I (F). $$

So writing A=BC where BI(F) and CI(F)c, we can rewrite the above sum as

$$ \sum_{F}\sum _{C\subset I (F)^{c}}\sum_{\theta_{F,C}} (- \mathbb{L} )^{|C|} [\theta_{F,C}]\sum _{B\subset I (F)} (-1)^{|B|} \mathbb{L}^{|B|+n (\theta_{F,B\cup C})}. $$
(B.4)

We will show that the inner most sum is always zero which will prove Theorem B.1.

Let yθ F,A where A=BC. We write the ℂ representation T y Y in two ways:

where restriction to the point y is implicit in the above equation. Counting positive weights on each side, we get

$$n (F) = n (\theta_{F,A}) +\sum_{i\in A} m_{i} (\theta_{F,A}) $$

where

$$m_{i} (\theta_{F,A}) = \operatorname{index}(N_{E_{i}/Y}|_{\theta_{F,A}}). $$

Note that m i (θ F,A ) is 1 or 0 depending on if the weight of the ℂ-action on E i |y is positive or not. Note also that if iI(F) then

$$m_{i} (\theta_{F,A}) = m_{i,F} = \operatorname{index}(N_{E_{i}/Y}|_{F}). $$

Moreover, if iI(F)c, then

$$N_{E_{i}/Y}|_{y} \subset TF|_{y} $$

and so m i (θ F,A )=0.

Thus writing A=BC with BI(F) and CI(F)c, we get

$$n (\theta_{F,A}) = n (F) - \sum_{i\in B} m_{i,F}, $$

and so

$$\sum_{B\subset I (F)} (-1)^{|B|} \mathbb{L}^{|B|+n (\theta_{F,A})} = \mathbb{L}^{n (F)}\sum _{B\subset I (F)} (-1)^{|B|} \mathbb{L}^{\sum_{i\in B} (1-m_{i,F})}. $$

For k=0,1, we define

$$I_{k} (F) = \bigl\{i\in I (F): m_{i,F}=k \bigr\}. $$

Then

$$\sum_{B\subset I (F)} (-1)^{|B|} \mathbb{L}^{\sum_{i\in B} (1-m_{i,F})} = \sum_{B_{0}\subset I_{0} (F)} (- \mathbb{L})^{|B_{0}|} \sum_{B_{1}\subset I_{1} (F)} (-1)^{|B_{1}|}, $$

but

$$\sum_{B_{1}\subset I_{1} (F)} (-1)^{|B_{1}|} = 0 $$

unless I 1(F)=∅. Since the above equation implies that the expression in (B.4) is zero, and that in turn verifies (B.2) and (B.3) which are equivalent to Theorem B.1, it only remains for us to prove that I 1(F)≠∅ for all F.

Let yF. We need to show that for some i, the action of ℂ on \(N_{E_{i}/Y}|_{y}\) has positive weight.

By the Luna slice theorem, there is an etale local neighborhood of yY which is equivariantly isomorphic to T y Y. Over the point y, we have a decomposition

$$TY = TF + N_{F/E_{I (F)}} +\sum_{i\in I (F)} N_{E_{i}/Y}. $$

Let (u 1,…,u s ,v 1,…,v p ,{w i } iI(F)) be linear coordinates on T y Y compatible with the above splitting. The action of t∈ℂ on T y Y is given by

$$t\cdot(u,v,w) = \bigl(u_{1},\dots,u_{s},t^{a_{1}}v_{1}, \dots ,t^{a_{p}}v_{p},\bigl\{t^{b_{i}}w_{i} \bigr\}_{i\in I (F)}\bigr). $$

In these coordinates, the function \(\widetilde{f}\) is given by

$$\widetilde{f} (u,v,w) = g (u) \prod_{i\in I (F)} w_{i}^{N_{i}} $$

where g(u) is a unit. Since the ℂ-action on Y is circle compact, \(\widetilde{f}\) is equivariant with respect to an action on ℂ of positive weight l, that is

$$\widetilde{f} \bigl(t\cdot(u,v,w)\bigr) = t^{l} \widetilde{f} (u,v,w). $$

This implies that

$$l=\sum_{i\in I (F)} b_{i}N_{i}. $$

Then since N i >0 and l>0 we have that b i >0 for some iI(F) and so for this i, we have m i,F =1 which was what we needed to prove. The proof of Theorem B.1 is now complete. □

Rights and permissions

Reprints and permissions

About this article

Cite this article

Behrend, K., Bryan, J. & Szendrői, B. Motivic degree zero Donaldson–Thomas invariants. Invent. math. 192, 111–160 (2013). https://doi.org/10.1007/s00222-012-0408-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00222-012-0408-1

Mathematics Subject Classification

Navigation