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

Cohomological χ\chi-independence for Higgs bundles and Gopakumar–Vafa invariants

Tasuki Kinjo and Naoki Koseki The University of Liverpool, Mathematical Sciences Building, Liverpool, L69 7ZL, UK. [email protected] graduate school of mathematical science, the university of tokyo, 3-8-1 komaba, meguroku, tokyo 153-8914, japan. [email protected]
Abstract.

The aim of this paper is two-fold: Firstly, we prove Toda’s χ\chi-independence conjecture for Gopakumar–Vafa invariants of arbitrary local curves. Secondly, following Davison’s work, we introduce the BPS cohomology for moduli spaces of Higgs bundles of rank rr and Euler characteristic χ\chi which are not necessary coprime, and show that it does not depend on χ\chi. This result extends the Hausel–Thaddeus conjecture on the χ\chi-independence of E-polynomials proved by Mellit, Groechenig–Wyss–Ziegler and Yu in two ways: we obtain an isomorphism of mixed Hodge modules on the Hitchin base rather than an equality of E-polynomials, and we do not need the coprime assumption.

The proof of these results is based on a description of the moduli stack of one-dimensional coherent sheaves on a local curve as a global critical locus which is obtained in the companion paper by the first author and Naruki Masuda.

1. Introduction

We work over the complex number field \mathbb{C}.

1.1. Motivation and Results

1.1.1. Non-abelian Hodge theory

Let CC be a smooth projective curve with genus g2g\geq 2. By the non-abelian Hodge correspondence due to [Cor88, Don87, Hit87, Sim92], there is a homeomorphism

(1.1) MDol(r,m)MB(r,m),M_{\operatorname{Dol}}(r,m)\simeq M_{B}(r,m),

where MDol(r,m)M_{\operatorname{Dol}}(r,m) is the moduli space of slope semistable Higgs bundles EE on CC with rank(E)=r,χ(E)=m\operatorname{rank}(E)=r,\chi(E)=m, and MB(r,m)M_{B}(r,m) is the twisted character variety, i.e., the quotient variety

MB(r,m){Ai,BiGLr,i=1,,gi[Ai,Bi]=e2mπ1/r}GLrM_{\operatorname{B}}(r,m)\coloneqq\Bigl{\{}A_{i},B_{i}\in\operatorname{GL}_{r},i=1,\ldots,g\mid\prod_{i}[A_{i},B_{i}]=e^{2m\pi\sqrt{-1}/r}\Bigr{\}}\sslash\operatorname{GL}_{r}

by the conjugate GLr\operatorname{GL}_{r} action.

Assume for a while that rr and mm are coprime so that the moduli spaces in (1.1) are smooth. The homeomorphism (1.1) induces an isomorphism

(1.2) H(MDol(r,m))H(MB(r,m))\operatorname{H}^{*}(M_{\operatorname{Dol}}(r,m))\simeq\operatorname{H}^{*}(M_{B}(r,m))

between the singular cohomology groups. However, since (1.1) is only a diffeomorphism, the isomorphism (1.2) is not an isomorphism of mixed Hodge structures. Indeed, the mixed Hodge structure on H(MDol(r,m))\operatorname{H}^{*}(M_{\operatorname{Dol}}(r,m)) is pure, while that on H(MB(r,m))\operatorname{H}^{*}(M_{B}(r,m)) is not pure. Instead, the cohomology group H(MDol(r,m))\operatorname{H}^{*}(M_{\operatorname{Dol}}(r,m)) has the so-called perverse filtration induced by the Hitchin morphism

h:MDol(r,m)B.h\colon M_{\operatorname{Dol}}(r,m)\to B.

De Cataldo–Hausel–Migliorini [dCHM12] conjectured that the perverse filtration on H(MDol(r,m))\operatorname{H}^{*}(M_{\operatorname{Dol}}(r,m)) matches with the weight filtration on H(MB(r,m))\operatorname{H}^{*}(M_{B}(r,m)) via the isomorphism (1.2) (P=W conjecture). This conjecture was recently proved by Hausel–Mellit–Minets–Schiffmann [HMMS] and Maulik–Shen [MSb] independently.

For character varieties, MB(r,m)M_{B}(r,m) and MB(r,m)M_{B}(r,m^{\prime}) are Galois conjugate to each other, for all m,mm,m^{\prime}\in\mathbb{Z} with gcd(r,m)=gcd(r,m)=1\gcd(r,m)=\gcd(r,m^{\prime})=1. In particular, we have an isomorphism

(1.3) H(MB(r,m))H(MB(r,m))\operatorname{H}^{*}(M_{\operatorname{B}}(r,m))\cong\operatorname{H}^{*}(M_{\operatorname{B}}(r,m^{\prime}))

of mixed Hodge structures. According to the P=W conjecture, the perverse filtration on H(MDol(r,m))\operatorname{H}^{*}(M_{\operatorname{Dol}}(r,m)) should be independent of mm\in\mathbb{Z}, as long as we have gcd(r,m)=1\gcd(r,m)=1. We prove this statement using the cohomological Donaldson–Thomas theory.

Theorem 1.1 (Example 5.18).

Let r,m,mr,m,m^{\prime} be integers such that r>0r>0 and gcd(r,m)=gcd(r,m)=1\gcd(r,m)=\gcd(r,m^{\prime})=1 hold. Then there exists an isomorphism

H(MDol(r,m))H(MDol(r,m))\operatorname{H}^{*}(M_{\operatorname{Dol}}(r,m))\cong\operatorname{H}^{*}(M_{\operatorname{Dol}}(r,m^{\prime}))

preserving the Hodge structure and the perverse filtration.

This kind of statement is called a χ\chi-independence phenomenon, as an invariant of the moduli space of Higgs bundles depends only on the rank rr and independent of the choice of the Euler characteristic χ\chi. Note that the above result for the perverse filtration was obtained by [dCMSZ] independently, via a completely different method.

Now assume that (r,m)(r,m) is not coprime. In this case, the moduli spaces MDol(r,m)M_{\operatorname{Dol}}(r,m) and MB(r,m)M_{\operatorname{B}}(r,m) are singular. Hence it is not clear which cohomology theory is a right choice to obtain a P=W type statement. There are two candidates for this:

  1. (1)

    Intersection cohomology ([dCM, FM, Mau, FSY]).

  2. (2)

    BPS cohomology ([CDP14]).

One advantage of using the intersection cohomology is that it is mathematically defined whereas the BPS cohomology is defined in the physical language. Instead of this, BPS cohomology has its own advantage: whereas the χ\chi-independence phenomena for the intersection cohomology is only expected when we have gcd(r,m)=gcd(r,m)\gcd(r,m)=\gcd(r,m^{\prime}), the χ\chi-independence for the BPS cohomology is expected to hold without any assumption. Further, the BPS cohomology groups in both sides are expected to carry a Lie algebra structure (see [Davb]) and the non-abelian Hodge correspondence (1.1) is expected to induce an isomorphism of these Lie algebras [SS20, Conjecture 1.5]. This suggests that we would have a representation theoretic approach to the original P=W conjecture.

Following Davison’s idea [Davc], we propose a definition of the BPS cohomology for the Dolbeault moduli space HBPS(MDol(r,m))\operatorname{H}_{\operatorname{BPS}}^{*}(M_{\operatorname{Dol}}(r,m)) as a cohomology of a pure Hodge module 𝒫𝒮r,m\mathop{\mathcal{BPS}}\nolimits_{r,m} on MDol(r,m)M_{\operatorname{Dol}}(r,m) defined using the cohomological Donaldson–Thomas theory (or refined BPS state counting) for TotC(𝒪CωC)\operatorname{Tot}_{C}({\mathcal{O}}_{C}\oplus\omega_{C}). We have a split injection 𝒞MDol(r,m)𝒫𝒮r,m\mathop{\mathcal{IC}}\nolimits_{M_{\operatorname{Dol}}(r,m)}\hookrightarrow\mathop{\mathcal{BPS}}\nolimits_{r,m} which is an isomorphism when gcd(r,m)=1\gcd(r,m)=1, but not necessarily so for general (r,m)(r,m). We prove the following χ\chi-independence for the BPS cohomology, which is a non-coprime generalization of Theorem 1.1:

Theorem 1.2 (Corollary 5.15).

Let r,m,mr,m,m^{\prime} be integers such that r>0r>0. Then there exists an isomorphism

HBPS(MDol(r,m))HBPS(MDol(r,m))\operatorname{H}_{\operatorname{BPS}}^{*}(M_{\operatorname{Dol}}(r,m))\cong\operatorname{H}_{\operatorname{BPS}}^{*}(M_{\operatorname{Dol}}(r,m^{\prime}))

preserving the Hodge structure and the perverse filtration.

Remark 1.3.

When we have gcd(r,m)=gcd(r,m)\gcd(r,m)=\gcd(r,m^{\prime}), the Betti moduli spaces MB(r,m)M_{\operatorname{B}}(r,m) and MB(r,m)M_{\operatorname{B}}(r,m^{\prime}) are Galois conjugate. Therefore we expect that there exists an isomorphism

(1.4) HBPS(MB(r,m))HBPS(MB(r,m))\operatorname{H}_{\operatorname{BPS}}^{*}(M_{\operatorname{B}}(r,m))\cong\operatorname{H}_{\operatorname{BPS}}^{*}(M_{\operatorname{B}}(r,m^{\prime}))

preserving the mixed Hodge structure, though we do not discuss the definition of the BPS cohomology for the Betti moduli spaces in this paper. Therefore Theorem 1.2 gives an evidence of the P=W conjecture for the BPS cohomology. Conversely, P=W conjecture and Theorem 1.2 suggest that the isomorphism (1.4) holds without the assumption gcd(r,m)=gcd(r,m)\gcd(r,m)=\gcd(r,m^{\prime}), which is of independent interest.

Remark 1.4.

Recently, Davison–Hennecart–Schlegel-Mejia [DHM22] established a theorem relating the BPS cohomology and the intersection cohomology for the moduli space of Higgs bundles and for the character varieties. Their work imply the equivalence of two versions of the P=W conjectures via the BPS cohomology and via the intersection cohomology, and that the χ\chi-independence of the intersection cohomology of the Dolbeault moduli space follows from Theorem 1.2 as long as gcd(r,m)=gcd(r,m)\gcd(r,m)=\gcd(r,m^{\prime}) holds.

We also establish the cohomological integrality theorem for Higgs bundles, which claims the decomposition of the Borel–Moore homology of the moduli stack of Higgs bundles HBM(𝔐Dol(r,m))\operatorname{H}^{\mathrm{BM}}_{*}({\mathfrak{M}}_{\operatorname{Dol}}(r,m)) into tensor products of the BPS cohomology (see Theorem 5.16 for the precise statement). A similar statement was proved for quivers with potentials in [DM20, Theorem A] and for preprojective algebras in [Davc, Theorem D]. As explained in [DM, §6.7], a plethystic computation and the cohomological integrality theorem imply that the Euler characteristic of the BPS cohomology is equal to the genus zero BPS invariant introduced by Joyce–Song [JS12, §6.2]. In particular, cohomological integrality theorem strengthens the integrality conjecture for the genus zero BPS invariants [JS12, Conjecture 6.12].

Combining the cohomological integrality theorem and the χ\chi-independence theorem (Theorem 1.2), we obtain the following χ\chi-independence result for the Borel–Moore homology of the moduli stack 𝔐Dol(r,m){\mathfrak{M}}_{\operatorname{Dol}}(r,m) of Higgs bundles:

Corollary 1.5 (Corollary 5.20).

Let r,m,mr,m,m^{\prime} be integers such that r>0r>0 and gcd(r,m)=gcd(r,m)\gcd(r,m)=\gcd(r,m^{\prime}) hold. Then there exists an isomorphism

HBM(𝔐Dol(r,m))HBM(𝔐Dol(r,m))\operatorname{H}^{\mathrm{BM}}_{*}({\mathfrak{M}}_{\operatorname{Dol}}(r,m))\cong\operatorname{H}^{\mathrm{BM}}_{*}({\mathfrak{M}}_{\operatorname{Dol}}(r,m^{\prime}))

preserving the Hodge structure and the perverse filtration introduced in [Dava, Proposition 7.24].

1.1.2. Gopakumar–Vafa (BPS) invariants

More generally, we investigate the χ\chi-independence phenomena for curve counting theory on a class of Calabi–Yau (CY) 3-folds called local curves. By definition, a local curve is a CY 3-fold of the form TotC(N)\operatorname{Tot}_{C}(N), where CC is a smooth projective curve and NN is a rank 22 vector bundle on CC such that det(N)ωC\det(N)\cong\omega_{C}. To explain our result, we recall some basic background of curve counting theory for CY 3-folds.

There are several ways to count curves in a CY 3-fold XX, and one of them is the Gromov–Witten (GW) theory: For an integer g0g\geq 0 and a homology class βH2(X,)\beta\in H_{2}(X,\mathbb{Z}), denote by Mg,β(X)M_{g,\beta}(X) the moduli space of stable maps f:CXf\colon C\to X with CC nodal curves of arithmetic genus gg and f[C]=βf_{*}[C]=\beta. Then the GW invariant is defined as

GWg,β[Mg,β(X)]vir1,\operatorname{GW}_{g,\beta}\coloneqq\int_{[M_{g,\beta}(X)]^{\operatorname{vir}}}1,

where [Mg,β(X)]vir[M_{g,\beta}(X)]^{\operatorname{vir}} denotes the virtual fundamental cycle. Due to the existence of stacky points in the moduli space Mg,β(X)M_{g,\beta}(X), the GW invariant GWg,β\operatorname{GW}_{g,\beta} is in general a rational number.

Based on string theory, Gopakumar–Vafa [GV] conjectured the existence of integer valued invariants ng,βn_{g,\beta}\in\mathbb{Z} for g0g\geq 0 and βH2(X,)\beta\in H_{2}(X,\mathbb{Z}), satisfying the equation

(1.5) g0,β>0GWg,βλ2g2tβ=g0,β>0,k1ng,βk(2sin(kλ2))2g2tkβ.\sum_{g\geq 0,\beta>0}\operatorname{GW}_{g,\beta}\lambda^{2g-2}t^{\beta}=\sum_{g\geq 0,\beta>0,k\geq 1}\frac{n_{g,\beta}}{k}\left(2\sin\left(\frac{k\lambda}{2}\right)\right)^{2g-2}t^{k\beta}.

We call the invariants ng,βn_{g,\beta} the Gopakumar–Vafa (GV) invariants (also known as the BPS invariants).

Building on the previous works by Hosono–Saito–Takahashi [HST01] and Kiem–Li [KL], Maulik–Toda [MT18] and Toda [Todb] proposed the mathematical definition of the GV invariants. Following the original idea of Gopakumar–Vafa, they consider the moduli space MX(β,m)M_{X}(\beta,m) of slope semistable one-dimensional sheaves EE on XX satisfying [E]=βH2(X,)[E]=\beta\in H_{2}(X,\mathbb{Z}) and χ(E)=m\chi(E)=m\in\mathbb{Z}. The moduli space MX(β,m)M_{X}(\beta,m) admits the Hilbert–Chow morphism

πM:MX(β,m)redChowβ(X),\pi_{M}\colon M_{X}(\beta,m)^{\operatorname{red}}\to\operatorname{Chow}_{\beta}(X),

which sends a sheaf to its fundamental cycle. Maulik–Toda [MT18] and Toda [Todb] defined the generalized GV invariants by the formula

(1.6) iχ(i(πM(φMX(β,m))))yi=g0ng,β,m(y12+y12)2g,\sum_{i\in\mathbb{Z}}\chi\left({\mathcal{H}}^{i}(\pi_{M*}\left(\varphi_{M_{X}(\beta,m)})\right)\right)y^{i}=\sum_{g\geq 0}n_{g,\beta,m}\left(y^{\frac{1}{2}}+y^{-\frac{1}{2}}\right)^{2g},

where φMX(β,m)\varphi_{M_{X}(\beta,m)} is a certain perverse sheaf on MX(β,m)M_{X}(\beta,m), see Sections 2.2 and 2.3 for more detail.

As the GV invariants are conjecturally equivalent to the GW invariants by the formula (1.5), the GV invariants should be independent of the additional choice of the Euler characteristic mm\in\mathbb{Z}:

Conjecture 1.6 ([Todb, Conjecture 1.2]).

The generalized GV invariants are independent of the choice of mm\in\mathbb{Z}, i.e., we have

ng,β,m=ng,β,mn_{g,\beta,m}=n_{g,\beta,m^{\prime}}

for all m,mm,m^{\prime}\in\mathbb{Z}.

We call the above conjecture as χ\chi-independence conjecture for GV invariants. In this paper, we prove it for local curves in full generality:

Theorem 1.7 (Theorem 3.1).

Conjecture 1.6 holds for X=TotC(N)X=\operatorname{Tot}_{C}(N).

1.2. Strategy of the proof

1.2.1. Results on local curves

The key ingredient in our arguments is the main result of the companion paper by the first author and Masuda [KM21] on the construction of a global d-critical chart for the moduli space 𝔐X𝔐X(β,m){\mathfrak{M}}_{X}\coloneqq{\mathfrak{M}}_{X}(\beta,m) of one-dimensional semistable sheaves on a local curve X=TotC(N)X=\operatorname{Tot}_{C}(N), i.e., the description of the moduli space 𝔐X{\mathfrak{M}}_{X} as the critical locus inside a certain smooth space: Take an exact sequence

(1.7) 0L1NL20,0\to L_{1}\to N\to L_{2}\to 0,

where L1,L2L_{1},L_{2} are line bundles with deg(L2)\deg(L_{2}) sufficiently large. We denote by YTotC(L2)Y\coloneqq\operatorname{Tot}_{C}(L_{2}). Then it is shown in [KM21, Theorem 5.6] that there exists a function f:MY𝔸1f\colon M_{Y}\to\mathbb{A}^{1} on the good moduli space of one-dimensional semistable sheaves on YY such that we have an isomorphism

𝔐X{d(fpY)=0}𝔐Y,{\mathfrak{M}}_{X}\cong\{d(f\circ p_{Y})=0\}\subset{\mathfrak{M}}_{Y},

where pY:𝔐YMYp_{Y}\colon{\mathfrak{M}}_{Y}\to M_{Y} is the natural map from the moduli stack of one-dimensional semistable sheaves to its good moduli space.

In this situation, the perverse sheaf appeared in the definition (1.6) of the generalized GV invariants coincides with the vanishing cycle sheaf:

φMXφf(𝒞MY),\varphi_{M_{X}}\cong\varphi_{f}(\mathop{\mathcal{IC}}\nolimits_{M_{Y}}),

and the proof of Theorem 1.7 is reduced to proving the corresponding statement for the intersection complex 𝒞MY\mathop{\mathcal{IC}}\nolimits_{M_{Y}}. The latter is proved in the recent paper by Maulik–Shen [MSa], hence we obtain Theorem 1.7.

1.2.2. Results on Higgs bundles

We define the BPS sheaf 𝒫𝒮r,m\mathop{\mathcal{BPS}}\nolimits_{r,m} on the moduli space MDol(r,m)M_{\operatorname{Dol}}(r,m) using the vanishing cycle complex φMX\varphi_{M_{X}} for X=TotC(𝒪CωC)X=\operatorname{Tot}_{C}({\mathcal{O}}_{C}\oplus\omega_{C}). Then the argument as above also implies Theorem 1.2. The cohomological integrality theorem for Higgs bundles (Theorem 5.16) is obtained by extending the argument for quivers with potentials [DM20, Theorem A] using the global critical locus description of 𝔐X{\mathfrak{M}}_{X} and applying the first author’s dimensional reduction theorem [Kin, Theorem 4.14] which relates the vanishing cycle cohomology for 𝔐X{\mathfrak{M}}_{X} and the Borel–Moore homology for 𝔐Dol{\mathfrak{M}}_{\operatorname{Dol}}.

1.3. Relation with existing works

  1. (1)

    Mellit [Mel20], Groechenig–Wyss–Ziegler [GWZ20], and Yu [Yu] proved that the E-polynomial of MDol(r,m)M_{\operatorname{Dol}}(r,m) is independent of mm\in\mathbb{Z} when gcd(r,m)=1\gcd(r,m)=1. These results were proved via the reduction to the positive characteristics.

    We extended the result to the non-coprime case and further lifted the equality to an isomorphism of Hodge structures via the completely different methods.

  2. (2)

    Recently, de Cataldo–Maulik–Shen–Zhang [dCMSZ] used a positive characteristic method to prove that the isomorphism (1.3) preserves the perverse filtration induced by the non-abelian Hodge theorem.

    At present, we do not know whether our cohomological χ\chi-independence results are compatible with the Galois conjugate.

  3. (3)

    Toda [Todb] proved Conjecture 1.6 for primitive classes βH2(X,)\beta\in H_{2}(X,\mathbb{Z}) (assuming a technical conjecture on orientation data). For non-primitive classes, Maulik–Shen [MSa] proved it for local toric del Pezzo surfaces and recently [Yua] removed the toric assumption from their result. Maulik–Shen [MSa] also proved the conjecture for local curves of the form TotC(𝒪(D)ωC(D))\operatorname{Tot}_{C}(\mathcal{O}(D)\oplus\omega_{C}(-D)) for a divisor DD with deg(D)>2g(C)2\deg(D)>2g(C)-2.

    Our Theorem 1.7 proves Conjecture 1.6 for arbitrary local curves. In particular, the result for X=TotC(N)X=\operatorname{Tot}_{C}(N) with indecomposable NN is completely new.

1.4. Structure of the paper

The paper is organized as follows. In Section 2, we recall Joyce’s theory on d-critical structures. Then we recall the definition of the GV invariants, and introduce the notion of local curves and twisted Higgs bundles.

In Section 3, we prove Theorem 1.7. In Section 4, we prove the cohomological integrality theorem for DD-twisted Higgs bundles where deg(D)>2g2\deg(D)>2g-2, which plays an important role in the proofs of Theorems 1.7 and Corollary 1.5. Finally in Section 5, we discuss applications to Higgs bundles. We prove Theorem 1.2 and the cohomological integrality theorem for Higgs bundles (Theorem 5.16).

In Appendix A, we give a brief overview of the shifted symplectic geometry and prove some technical lemmas that we use in this paper.

In Appendix B, we prove a version of the support lemma of the vanishing cycle complexes which is needed to define the BPS sheaf.

Acknowledgement.

The authors would like to thank Professors Ben Davison, Yukinobu Toda and Junliang Shen for fruitful discussions and for carefully reading the previous version of this article. The first author would like to thank Naruki Masuda for the collaboration on the companion paper [KM21]. The second author would like to thank Professors Arend Bayer and Jim Bryan for related discussions.

T.K. was supported by WINGS-FMSP program at the Graduate School of Mathematical Science, the University of Tokyo and JSPS KAKENHI Grant number JP21J21118. N.K. was supported by ERC Consolidator grant WallCrossAG, no. 819864.

Notation and Convention.

In this paper, we work over the complex number field \mathbb{C}. We use the following notations:

  • We let 𝕊\mathbb{S} denote the \infty-category of spaces (see [Lur09, Definition 1.2.16.1]).

  • We basically write stacks in Fraktur (e.g. 𝔛,𝔜,{\mathfrak{X}},{\mathfrak{Y}},\ldots), and derived schemes, derived stacks and morphisms between derived stacks in bold (e.g. 𝑿,𝖃,𝒇,\boldsymbol{X},\boldsymbol{{\mathfrak{X}}},\boldsymbol{f},\ldots). We will write X=t0(𝑿)X=t_{0}(\boldsymbol{X}), 𝔛=t0(𝖃){\mathfrak{X}}=t_{0}(\boldsymbol{{\mathfrak{X}}}), f=t0(𝒇)f=t_{0}(\boldsymbol{f}) and so on.

  • A derived Artin stack 𝔛\mathbb{{\mathfrak{X}}} is said to be quasi-smooth if the cotangent complex 𝕃𝔛\mathbb{L}_{{\mathfrak{X}}} has Tor-amplitude [1,1][-1,1].

  • All derived/underived Artin stacks are assumed to have quasi-compact and separated diagonals and locally finitely presented over the complex number field. As the fiber product of finite type separated schemes over such a stack is again of finite type and separated, we can consider the category of mixed Hodge modules on such stacks (see §4.1 for the detail).

  • For a separated complex analytic space XX, we let Dcb(X)D_{c}^{b}(X) denote the bounded derived category of complexes of sheaves in \mathbb{Q}-vector spaces on XX with constructible cohomology.

  • For a complex analytic stack 𝔛{\mathfrak{X}}, we let Dcb(𝔛)D^{b}_{c}({\mathfrak{X}}) denote the bounded derived category of sheaves in \mathbb{Q}-vector spaces on 𝔛lis-an{\mathfrak{X}}_{\textrm{lis-an}} with constructible cohomology. Here 𝔛lis-an{\mathfrak{X}}_{\textrm{lis-an}} denote the lisse-analytic topos of 𝔛{\mathfrak{X}} (see [Sun17, §3.2.3]).

  • If there is no confusion, we use expressions such as ff_{*} and f!f_{!} for the derived functors 𝐑f\mathbf{R}f_{*} and 𝐑f!\mathbf{R}f_{!}.

2. Preliminaries

In this section, we collect some basic notions that we use in this paper. Firstly we recall Joyce’s theory of d-critical locus in §2.1. Then we review the construction of vanishing cycle complexes associated with d-critical stacks in §2.2. In §2.3 we review Maulik–Toda’s construction [MT18] of Gopakumar–Vafa invariants based on vanishing cycle complexes. In §2.4 we collect some basic facts on local curves and recall Maulik–Shen’s cohomological χ\chi-independence theorem [MSa].

2.1. D-critical structures

In [Joy15], Joyce introduced the notion of d-critical structures which are classical models of derived critical loci. We now briefly recall it.

Let XX be a complex analytic space. Joyce [Joy15, Theorem 2.1] introduced a sheaf 𝒮X{\mathcal{S}}_{X} on XX with the following property: for an open subset RXR\subset X and an embedding i:RUi\colon R\hookrightarrow U to a complex manifold UU, there exists a short exact sequence

0𝒮X|Ri1𝒪U/IR,U2ddRi1ΩU/(IR,Ui1ΩU),0\to{\mathcal{S}}_{X}|_{R}\to i^{-1}{\mathcal{O}}_{U}/I_{R,U}^{2}\ \xrightarrow[]{d_{\mathrm{dR}}}i^{-1}\Omega_{U}/(I_{R,U}\cdot i^{-1}\Omega_{U}),

where IR,UI_{R,U} is the ideal sheaf of RR in UU. One can show that the natural map

𝒮X|Ri1𝒪U/IR,U2𝒪R{\mathcal{S}}_{X}|_{R}\to i^{-1}{\mathcal{O}}_{U}/I_{R,U}^{2}\twoheadrightarrow{\mathcal{O}}_{R}

glues to define a morphism 𝒮X𝒪X{\mathcal{S}}_{X}\to{\mathcal{O}}_{X}. We define a subsheaf 𝒮X0𝒮X{\mathcal{S}}_{X}^{0}\subset{\mathcal{S}}_{X} by the kernel of the map

𝒮X𝒪X𝒪Xred.{\mathcal{S}}_{X}\to{\mathcal{O}}_{X}\twoheadrightarrow{\mathcal{O}}_{X^{\operatorname{red}}}.

If XX is the critical locus Crit(f)\operatorname{Crit}(f) of a holomorphic function ff on a complex manifold UU such that f|Xred=0f|_{X^{\operatorname{red}}}=0, then f+IX,U2f+I_{X,U}^{2} defines an element of 𝒮X0{\mathcal{S}}_{X}^{0}.

Definition 2.1.

Let XX be a complex analytic space. A section sΓ(X,𝒮X0)s\in\Gamma(X,{\mathcal{S}}_{X}^{0}) is called a d-critical structure if for each point xXx\in X, there exists an open neighborhood RXR\subset X, an embedding i:RUi\colon R\hookrightarrow U into a complex manifold, and a holomorphic function ff on UU with the property f|Rred=0f|_{R^{\operatorname{red}}}=0 such that f+IR,U2=s|Rf+I_{R,U}^{2}=s|_{R}. The quadruple (R,U,f,i)(R,U,f,i) is called a d-critical chart of XX. A complex analytic space equipped with a d-critical structure is called a d-critical analytic space.

The sheaf 𝒮X0{\mathcal{S}}_{X}^{0} has the following functorial property: for a given morphism of complex analytic spaces q:X1X2q\colon X_{1}\to X_{2}, there exist natural morphisms

q:q1𝒮X20𝒮X10.q^{\star}\colon q^{-1}{\mathcal{S}}_{X_{2}}^{0}\to{\mathcal{S}}_{X_{1}}^{0}.

Now assume that qq is smooth surjective and take a section sΓ(X,𝒮X0)s\in\Gamma(X,{\mathcal{S}}_{X}^{0}). Then it is shown in [Joy15, Proposition 2.8] that qsq^{\star}s is a d-critical structure if and only if ss is a d-critical structure.

Now let 𝔛{\mathfrak{X}} be a complex analytic stack. Then it is shown in [Joy15, Corollary 2.52] that there exists a sheaf 𝒮𝔛0{\mathcal{S}}_{{\mathfrak{X}}}^{0} on the lisse-analytic site of 𝔛{\mathfrak{X}} with the following property:

  • For a smooth morphism t:T𝔛t\colon T\to{\mathfrak{X}}, there exists a natural isomorphism ηt:𝒮𝔛0|T𝒮T0\eta_{t}\colon{\mathcal{S}}_{{\mathfrak{X}}}^{0}|_{T}\cong{\mathcal{S}}_{T}^{0}.

  • For a morphism

    q:(t1:T1𝔛)(t2:T2𝔛)q\colon(t_{1}\colon T_{1}\to{\mathfrak{X}})\to(t_{2}\colon T_{2}\to{\mathfrak{X}})

    between complex analytic spaces smooth over 𝔛{\mathfrak{X}}, the natural map q1(𝒮𝔛0|T2)𝒮𝔛0|T1q^{-1}({\mathcal{S}}_{{\mathfrak{X}}}^{0}|_{T_{2}})\to{\mathcal{S}}_{{\mathfrak{X}}}^{0}|_{T_{1}} is identified with qq^{\star}.

For a smooth morphism t:T𝔛t\colon T\to{\mathfrak{X}} from a scheme and a section sΓ(𝔛,𝒮𝔛0)s\in\Gamma({\mathfrak{X}},{\mathcal{S}}_{{\mathfrak{X}}}^{0}), we write tsηt(s|T)Γ(T,𝒮T0)t^{\star}s\coloneqq\eta_{t}(s|_{T})\in\Gamma(T,{\mathcal{S}}_{T}^{0}).

Definition 2.2.

For a complex analytic stack 𝔛{\mathfrak{X}}, a section sΓ(𝔛,𝒮𝔛0)s\in\Gamma({\mathfrak{X}},{\mathcal{S}}_{{\mathfrak{X}}}^{0}) is called a d-critical structure if for any smooth surjective morphism t:T𝔛t\colon T\to{\mathfrak{X}}, the element tst^{\star}s is a d-critical structure on TT. A d-critical stack is a complex analytic stack 𝔛{\mathfrak{X}} equipped with a d-critical structure.

For a complex analytic stack 𝔛{\mathfrak{X}} equipped with a d-critical structure ss, Joyce [Joy15, §2.4, §2.8] defines a line bundle K𝔛,svirK_{{\mathfrak{X}},s}^{\operatorname{vir}} on 𝔛red{\mathfrak{X}}^{\operatorname{red}} called the virtual canonical bundle of (𝔛,s)({\mathfrak{X}},s). If there is no confusion, we simply write K𝔛vir=K𝔛,svirK_{{\mathfrak{X}}}^{\operatorname{vir}}=K_{{\mathfrak{X}},s}^{\operatorname{vir}}. We now recall some of its basic properties. Firstly assume 𝔛{\mathfrak{X}} is a complex analytic space and write 𝔛=X{\mathfrak{X}}=X. Take a d-critical chart =(R,U,f,i){\mathscr{R}}=(R,U,f,i) of (X,s)(X,s). Then there exists a natural isomorphism

ι:KX,svir|RredKU2|Rred.\iota_{{\mathscr{R}}}\colon K_{X,s}^{\operatorname{vir}}|_{R^{\operatorname{red}}}\cong K_{U}^{\otimes 2}|_{R^{\operatorname{red}}}.

Let q:X1X2q\colon X_{1}\to X_{2} be a smooth morphism and s2s_{2} be a d-critical structure on X2X_{2}. Write s1=qs2s_{1}=q^{\star}s_{2}. Then it is shown in [Joy15, Proposition 2.30] that there exists a natural isomorphism

Υq:qred,KX2,s2virdet(ΩX1/X2)|X1red2KX1,s1vir\Upsilon_{q}\colon q^{\operatorname{red},*}K_{X_{2},s_{2}}^{\operatorname{vir}}\otimes\det(\Omega_{X_{1}/X_{2}})|_{X_{1}^{\operatorname{red}}}^{\otimes 2}\cong K_{X_{1},s_{1}}^{\operatorname{vir}}

with the following property: if we are given d-critical charts 1=(R1,U1,f1,i1){\mathscr{R}}_{1}=(R_{1},U_{1},f_{1},i_{1}) of (X1,s1)(X_{1},s_{1}) and 2=(R2,U2,f2,i2){\mathscr{R}}_{2}=(R_{2},U_{2},f_{2},i_{2}) of (X2,s2)(X_{2},s_{2}) such that q(R1)R2q(R_{1})\subset R_{2}, and a smooth morphism q~:U1U2\tilde{q}\colon U_{1}\to U_{2} such that f1=f2q~f_{1}=f_{2}\circ\tilde{q} and i2q|R1=q~i1i_{2}\circ q|_{R_{1}}=\tilde{q}\circ i_{1}, the following diagram of line bundles on R1redR_{1}^{\operatorname{red}} commutes:

qred,KX2,s2vir|R1reddet(ΩX1/X2)|R1red2\textstyle{q^{\operatorname{red},*}K_{X_{2},s_{2}}^{\operatorname{vir}}|_{R_{1}^{\operatorname{red}}}\otimes\det(\Omega_{X_{1}/X_{2}})|_{R_{1}^{\operatorname{red}}}^{\otimes 2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Υq|R1red\scriptstyle{\Upsilon_{q}|_{R_{1}^{\operatorname{red}}}}(q|R1)red,ι2id\scriptstyle{(q|_{R_{1}})^{\operatorname{red},*}\iota_{{\mathscr{R}}_{2}}\otimes\operatorname{id}}KX1,s1vir|R1red\textstyle{K_{X_{1},s_{1}}^{\operatorname{vir}}|_{R_{1}^{\operatorname{red}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι1\scriptstyle{\iota_{{\mathscr{R}}_{1}}}q~KU22|R1reddet(ΩX1/X2)|R1red2\textstyle{\tilde{q}^{*}K_{U_{2}}^{\otimes{2}}|_{R_{1}^{\operatorname{red}}}\otimes\det(\Omega_{X_{1}/X_{2}})|_{R_{1}^{\operatorname{red}}}^{\otimes 2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}KU12|R1red.\textstyle{K_{U_{1}}^{\otimes{2}}|_{R_{1}^{\operatorname{red}}}.}

Here the bottom horizontal arrow is defined by the natural isomorphism ΩX1/X2|R1ΩU1/U2|R1\Omega_{X_{1}/X_{2}}|_{R_{1}}\cong\Omega_{U_{1}/U_{2}}|_{R_{1}}.

Now we treat the stacky case. Let 𝔛{\mathfrak{X}} be a complex analytic space and t:T𝔛t\colon T\to{\mathfrak{X}} be a smooth morphism from an analytic space. Then there exists a natural isomorphism

Υt:tred,K𝔛,svirdet(ΩT/𝔛)|Tred2KT,tsvir,\Upsilon_{t}\colon t^{\operatorname{red},*}K_{{\mathfrak{X}},s}^{\operatorname{vir}}\otimes\det(\Omega_{T/{\mathfrak{X}}})|_{T^{\operatorname{red}}}^{\otimes 2}\cong K_{T,t^{\star}s}^{\operatorname{vir}},

see [Joy15, Theorem 2.56]. For a morphism

q:(t1:T1𝔛)(t2:T2𝔛)q\colon(t_{1}\colon T_{1}\to{\mathfrak{X}})\to(t_{2}\colon T_{2}\to{\mathfrak{X}})

between complex analytic spaces smooth over 𝔛{\mathfrak{X}}, the following diagram commutes:

(2.1) t1red,K𝔛,svirdet(ΩT1/𝔛)|T1red2Υt1qred,(t2red,K𝔛,svirdet(ΩT2/𝔛)|T2red2)det(ΩT1/T2)|T1red2qred,Υt2idqred,KT2,t2svirdet(ΩT1/T2)|T1red2ΥqKT1,t1svir.\begin{split}\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 91.40045pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\\&\\&\crcr}}}\ignorespaces{\hbox{\kern-46.09471pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{t_{1}^{\operatorname{red},*}K_{{\mathfrak{X}},s}^{\operatorname{vir}}\otimes\det(\Omega_{T_{1}/{\mathfrak{X}}})|_{T_{1}^{\operatorname{red}}}^{\otimes 2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 0.0pt\raise-30.5pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}\ignorespaces\ignorespaces{\hbox{\kern 97.946pt\raise-11.43422pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.26103pt\hbox{$\scriptstyle{\Upsilon_{t_{1}}}$}}}\kern 3.0pt}}}}}}\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{\hbox{\kern 128.44745pt\raise-73.9025pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{}{{}}{{}{}{}\lx@xy@spline@}{}}}}\ignorespaces{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}{\hbox{\kern 128.12004pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern-91.40045pt\raise-42.80444pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{q^{\operatorname{red},*}(t_{2}^{\operatorname{red},*}K_{{\mathfrak{X}},s}^{\operatorname{vir}}\otimes\det(\Omega_{T_{2}/{\mathfrak{X}}})|_{T_{2}^{\operatorname{red}}}^{\otimes 2})\otimes\det(\Omega_{T_{1}/T_{2}})|_{T_{1}^{\operatorname{red}}}^{\otimes 2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 0.0pt\raise-61.30444pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.29993pt\hbox{$\scriptstyle{q^{\operatorname{red},*}\Upsilon_{t_{2}}\otimes\operatorname{id}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 0.0pt\raise-73.30444pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 128.12004pt\raise-42.80444pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern-54.63644pt\raise-85.60889pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{q^{\operatorname{red},*}K_{T_{2},t_{2}^{\star}s}^{\operatorname{vir}}\otimes\det(\Omega_{T_{1}/T_{2}})|_{T_{1}^{\operatorname{red}}}^{\otimes 2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 78.33165pt\raise-79.39778pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.57222pt\hbox{$\scriptstyle{\Upsilon_{q}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 115.40045pt\raise-85.60889pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 115.40045pt\raise-85.60889pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{K_{T_{1},t_{1}^{\star}s}^{\operatorname{vir}}.}$}}}}}}}\ignorespaces}}}}\ignorespaces\end{split}

For a d-critical stack (𝔛,s)({\mathfrak{X}},s), an orientation is a choice of a line bundle LL on 𝔛red{\mathfrak{X}}^{\operatorname{red}} and an isomorphism o:L2K𝔛,sviro\colon L^{\otimes{2}}\cong K_{{\mathfrak{X}},s}^{\operatorname{vir}}. For a smooth morphism t:T𝔛t\colon T\to{\mathfrak{X}}, we define an orientation

to:(tred,Ldet(ΩT/𝔛)|Tred)2KT,tsvirt^{\star}o\colon(t^{\operatorname{red},*}L\otimes\det(\Omega_{T/{\mathfrak{X}}})|_{T^{\operatorname{red}}})^{\otimes 2}\cong K_{T,t^{\star}s}^{\operatorname{vir}}

using Υt\Upsilon_{t}. If we are given a smooth morphism q:(t1:T1𝔛)(t2:T2𝔛)q\colon(t_{1}\colon T_{1}\to{\mathfrak{X}})\to(t_{2}\colon T_{2}\to{\mathfrak{X}}) between analytic spaces smooth over 𝔛{\mathfrak{X}}, there exists a natural isomorphism

(2.2) t1oqt2o.t_{1}^{\star}o\cong q^{\star}t_{2}^{\star}o.

2.2. Vanishing cycle complexes on d-critical stacks

In this subsection, we recall some basic properties of the vanishing cycle functors and the vanishing cycle complexes associated with oriented d-critical stacks.

Let UU be a complex manifold and ff be a holomorphic function on UU. Write U0=f1(0)U_{0}=f^{-1}(0). Then the vanishing cycle functor

φf:Dcb(U)Dcb(U0).\varphi_{f}\colon D^{b}_{c}(U)\to D^{b}_{c}(U_{0}).

is defined by the following formula

φf(U0U0)(U0U)!,\varphi_{f}\coloneqq(U_{0}\hookrightarrow U_{\leq 0})^{*}(U_{\leq 0}\hookrightarrow U)^{!},

where U0UU_{\leq 0}\subset U denotes the closed subset Re(f)1(0){\mathrm{Re}(f)}^{-1}(\mathbb{R}_{\leq 0}). It is shown in [KS13, Corollary 10.3.13] that the functor φf\varphi_{f} preserves the perversity. If there is no confusion, we write φfφf(U[dimU])\varphi_{f}\coloneqq\varphi_{f}(\mathbb{Q}_{U}[\dim U]).

Let q:VUq\colon V\to U be a holomorphic map between complex manifolds. Write V0(fq)1(0)V_{0}\coloneqq(f\circ q)^{-1}(0) and we let q0:V0U0q_{0}\colon V_{0}\to U_{0} be the restriction of qq. By the definition of the vanishing cycle functor, we have the following base change morphisms

φfq\displaystyle\varphi_{f}\circ q_{*} q0φfq\displaystyle\to{q_{0}}_{*}\circ\varphi_{f\circ q}
q0φf\displaystyle q_{0}^{*}\varphi_{f} φfqq.\displaystyle\to\varphi_{f\circ q}\circ q^{*}.

The first morphism is an isomorphism if qq is proper and the latter morphism is an isomorphism if qq is smooth. These are direct consequences of the proper/smooth base change theorem.

Now let 𝔘{\mathfrak{U}} be a smooth complex analytic stack and ff be a holomorphic function on 𝔘{\mathfrak{U}}. Write 𝔘0f1(0){\mathfrak{U}}_{0}\coloneqq f^{-1}(0). For a perverse sheaf 𝒫Perv(𝔛){\mathcal{P}}\in\operatorname{Perv}({\mathfrak{X}}), we define the perverse sheaf

φf(𝒫)Perv(𝔘0)\varphi_{f}({\mathcal{P}})\in\operatorname{Perv}({\mathfrak{U}}_{0})

as follows: Take a smooth surjective morphism q:U𝔘q\colon U\to{\mathfrak{U}}. We let pri:U×𝔘UU\operatorname{pr}_{i}\colon U\times_{{\mathfrak{U}}}U\to U denote the ii-th projection and pri,0:(fqpri)1(0)(fq)1(0)\operatorname{pr}_{i,0}\colon(f\circ q\circ\operatorname{pr}_{i})^{-1}(0)\to(f\circ q)^{-1}(0) denote the restriction of pri\operatorname{pr}_{i}. Then we have a natural isomorphism

pr1,0φfq(q𝒫)φfqpr1(pr1q𝒫)φfqpr2(pr2q𝒫)pr2,0φfq(q𝒫).\displaystyle\operatorname{pr}_{1,0}^{*}\varphi_{f\circ q}(q^{*}{\mathcal{P}})\cong\varphi_{f\circ q\circ\operatorname{pr}_{1}}(\operatorname{pr}_{1}^{*}q^{*}{\mathcal{P}})\cong\varphi_{f\circ q\circ\operatorname{pr}_{2}}(\operatorname{pr}_{2}^{*}q^{*}{\mathcal{P}})\cong\operatorname{pr}_{2,0}^{*}\varphi_{f\circ q}(q^{*}{\mathcal{P}}).

This isomorphism satisfies the cocycle condition, hence the shifted perverse sheaf φfq(q𝒫)\varphi_{f\circ q}(q^{*}{\mathcal{P}}) descends to a perverse sheaf φf(𝒫)Perv(𝔘0)\varphi_{f}({\mathcal{P}})\in\operatorname{Perv}({\mathfrak{U}}_{0}). One can show that the construction does not depend on the choice of the smooth morphism qq.

Now we recall the vanishing cycle complex associated with an oriented d-critical stack constructed in [BBBBJ15, Theorem 4.8].

First we treat the non-stacky case. Let (X,s,o)(X,s,o) be an oriented d-critical analytic space. Then it is shown in [BBD+15, Theorem 6.9] that there is a natural perverse sheaf

φX,s,oPerv(X)\varphi_{X,s,o}\in\operatorname{Perv}(X)

called the vanishing cycle complex associated with (X,s,o)(X,s,o). We sometimes omit ss and oo and write φX=φX,s,o\varphi_{X}=\varphi_{X,s,o} if there is no confusion. For a d-critical chart =(R,U,f,i){\mathscr{R}}=(R,U,f,i) of (X,s)(X,s), we have a natural isomorphism

ω:φX,s,o|Riφf/2Qo,\omega_{{\mathscr{R}}}\colon\varphi_{X,s,o}|_{R}\cong i^{*}\varphi_{f}\otimes_{\mathbb{Z}/2\mathbb{Z}}Q_{{\mathscr{R}}}^{o},

where QoQ_{{\mathscr{R}}}^{o} is a /2\mathbb{Z}/2\mathbb{Z}-local system on RR parametrizing local square roots of the isomorphism

L2|Rred𝑜KX,svir|RrediKU2|Rred.L^{\otimes 2}|_{R^{\operatorname{red}}}\xrightarrow[\cong]{o}K_{X,s}^{\operatorname{vir}}|_{R^{\operatorname{red}}}\cong i^{*}K_{U}^{\otimes 2}|_{R^{\operatorname{red}}}.
Example 2.3.

Let UU be a complex manifold and f:U𝔸1f\colon U\to\mathbb{A}^{1} be a holomorphic function such that f|Crit(f)red=0f|_{\operatorname{Crit}(f)^{\operatorname{red}}}=0. Write X=Crit(f)X=\operatorname{Crit}(f) and equip it with the canonical d-critical structure ss and the canonical orientation o:KU|Xred2KX,sviro\colon K_{U}|_{X^{\operatorname{red}}}^{\otimes 2}\cong K_{X,s}^{\operatorname{vir}}. Then (X,U,f,XU){\mathscr{R}}\coloneqq(X,U,f,X\hookrightarrow U) defines a d-critical chart. In this case the local system QoQ_{{\mathscr{R}}}^{o} is trivial. Therefore we have a natural isomorphism

φX,s,oφf|X.\varphi_{X,s,o}\cong\varphi_{f}|_{X}.

Let q:X1X2q\colon X_{1}\to X_{2} be a smooth morphism and equip X2X_{2} with a d-critical structure s2s_{2} and an orientation o2o_{2}. Write s1=qs2s_{1}=q^{\star}s_{2} and o1=qo2o_{1}=q^{\star}o_{2}. Then there exists a natural isomorphism of perverse sheaves

Θq:φX1,s1,o1qφX2,s2,o2[dimq]\Theta_{q}\colon\varphi_{X_{1},s_{1},o_{1}}\cong q^{*}\varphi_{X_{2},s_{2},o_{2}}[\dim q]

with the following property: If we are given d-critical charts 1=(R1,U1,f1,i1){\mathscr{R}}_{1}=(R_{1},U_{1},f_{1},i_{1}) of (X1,s1)(X_{1},s_{1}) and 2=(R2,U2,f2,i2){\mathscr{R}}_{2}=(R_{2},U_{2},f_{2},i_{2}) of (X2,s2)(X_{2},s_{2}) such that q(R1)R2q(R_{1})\subset R_{2}, and a smooth morphism q~:U1U2\tilde{q}\colon U_{1}\to U_{2} such that f1=f2q~f_{1}=f_{2}\circ\tilde{q} and i2q|R=q~i1i_{2}\circ q|_{R}=\tilde{q}\circ i_{1}, the following diagram in Perv(R1)\operatorname{Perv}(R_{1}) commutes:

(2.3) φX1,s1,o1|R1\textstyle{\varphi_{X_{1},s_{1},o_{1}}|_{R_{1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ω1\scriptstyle{\omega_{{\mathscr{R}}_{1}}}Θq|R1\scriptstyle{\Theta_{q}|_{R_{1}}}i1φf1/2Q1o1\textstyle{i_{1}^{*}\varphi_{f_{1}}\otimes_{\mathbb{Z}/2\mathbb{Z}}Q_{{\mathscr{R}}_{1}}^{o_{1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(q|R1)(φX2,s2,o2)[dimq]\textstyle{(q|_{R_{1}})^{*}(\varphi_{X_{2},s_{2},o_{2}})[\dim q]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(q|R1)ω2[dimq]\scriptstyle{(q|_{R_{1}})^{*}\omega_{{\mathscr{R}}_{2}}[\dim q]}(q|R1)(i2φf2/2Q2o2)[dimq],\textstyle{(q|_{R_{1}})^{*}(i_{2}^{*}\varphi_{f_{2}}\otimes_{\mathbb{Z}/2\mathbb{Z}}Q_{{\mathscr{R}}_{2}}^{o_{2}})[\dim q],}

where the right vertical arrow is defined using the natural isomorphisms φf1(q~|(f2q~)1(0))φf2[dimq]\varphi_{f_{1}}\cong(\tilde{q}|_{(f_{2}\circ\tilde{q})^{-1}(0)})^{*}\varphi_{f_{2}}[\dim q] and Q1o1(q|R1)Q2o2Q_{{\mathscr{R}}_{1}}^{o_{1}}\cong(q|_{R_{1}})^{*}Q_{{\mathscr{R}}_{2}}^{o_{2}}.

Now we move to the stacky case. Let (𝔛,s,o)({\mathfrak{X}},s,o) be a d-critical stack. Then it is shown in [BBBBJ15, Theorem 4.8] that there exists a natural perverse sheaf

φ𝔛,s,oPerv(𝔛)\varphi_{{\mathfrak{X}},s,o}\in\operatorname{Perv}({\mathfrak{X}})

with the following property: If we are given a smooth morphism t:T𝔛t\colon T\to{\mathfrak{X}} from a complex analytic space, there exists a natural isomorphism

Θt:φT,to,totφX,s,o[dimt].\Theta_{t}\colon\varphi_{T,t^{\star}o,t^{\star}o}\cong t^{*}\varphi_{X,s,o}[\dim t].

Furthermore, if we are given a smooth morphism q:(t1:T1𝔛)(t2:T2𝔛)q\colon(t_{1}\colon T_{1}\to{\mathfrak{X}})\to(t_{2}\colon T_{2}\to{\mathfrak{X}}) between schemes smooth over 𝔛{\mathfrak{X}}, the following diagram commutes:

(2.4) φT1,t1s,t1o\textstyle{\varphi_{T_{1},t_{1}^{\star}s,t_{1}^{\star}o}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}Θt1\scriptstyle{\Theta_{t_{1}}}φT1,qt2s,qt2o\textstyle{\varphi_{T_{1},q^{\star}t_{2}^{\star}s,q^{\star}t_{2}^{\star}o}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Θq\scriptstyle{\Theta_{q}}qφT2,t2s,t2o[dimq]\textstyle{q^{*}\varphi_{T_{2},t_{2}^{\star}s,t_{2}^{\star}o}\ignorespaces\ignorespaces\ignorespaces\ignorespaces[\dim q]}qΘt2\scriptstyle{q^{*}\Theta_{t_{2}}}t1φ𝔛,s,o[dimt1]\textstyle{t_{1}^{*}\varphi_{{\mathfrak{X}},s,o}[\dim t_{1}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}qt2φ𝔛,s,o[dimt1].\textstyle{q^{*}t_{2}^{*}\varphi_{{\mathfrak{X}},s,o}[\dim t_{1}].}

Let 𝔘{\mathfrak{U}} be a smooth Artin stack and f:𝔘𝔸1f\colon{\mathfrak{U}}\to\mathbb{A}^{1} be a regular function on it. Then it is shown in Example A.7 that 𝐂𝐫𝐢𝐭(f)\mathop{\mathbf{Crit}}\nolimits(f) carries a natural (1)(-1)-shifted symplectic structure hence there exists a natural d-critical structure ss on its classical truncation 𝔛Crit(f){\mathfrak{X}}\coloneqq\operatorname{Crit}(f). We will see in Lemma A.10 that the d-critical analytic stack (𝔛,s)({\mathfrak{X}},s) admits a canonical orientation o:K𝔘2|𝔛redK𝔛,sviro\colon K_{{\mathfrak{U}}}^{\otimes{2}}|_{{\mathfrak{X}}^{\operatorname{red}}}\cong K_{{\mathfrak{X}},s}^{\operatorname{vir}}.

Proposition 2.4.

There exists a natural isomorphism of perverse sheaves:

θ:φ𝔛,s,oφf(𝔘[dim𝔘]).\theta\colon\varphi_{{\mathfrak{X}},s,o}\cong\varphi_{f}(\mathbb{Q}_{{\mathfrak{U}}}[\dim{\mathfrak{U}}]).

We postpone the proof to §A.2.

Remark 2.5.

The argument in [BBBBJ15, Theorem 4.8] shows that the perverse sheaf φ𝔛,s,o\varphi_{{\mathfrak{X}},s,o} naturally extends to a mixed Hodge module φ𝔛,s,omhm\varphi_{{\mathfrak{X}},s,o}^{\mathrm{mhm}} and to a monodromic mixed Hodge module φ𝔛,s,ommhm\varphi_{{\mathfrak{X}},s,o}^{\mathrm{mmhm}}. Proposition 2.4 extends to an isomorphism of monodromic mixed Hodge modules with the same proof. We refer the reader to §4.2 for a brief discussion on monodromic mixed Hodge modules.

2.3. Maulik–Toda’s construction of Gopakumar–Vafa invariants

In this subsection, we recall the definition of generalized Gopakumar–Vafa (GV) invariants following [MT18, Todb]. Let XX be a smooth quasi-projective Calabi–Yau threefold and HH be an ample divisor on XX.

Definition 2.6.

Let EE be a pure one-dimensional coherent sheaf with compact support on XX.

  1. (1)

    We define the HH-slope to be

    μH(E)χ(E)H.[E],\mu_{H}(E)\coloneqq\frac{\chi(E)}{H.[E]},

    where [E]H2(X,)[E]\in H_{2}(X,\mathbb{Z}) denotes the second homology class of EE.

  2. (2)

    The sheaf EE is μH\mu_{H}-semistable (resp. stable) if for any saturated subsheaf 0FE0\neq F\subsetneq E, the inequality

    μH(F)μH(E)(resp. μH(F)<μH(E))\mu_{H}(F)\leq\mu_{H}(E)\quad(\mbox{resp. }\mu_{H}(F)<\mu_{H}(E))

    holds.

For a given element v=(β,m)H2(X,)×v=(\beta,m)\in H_{2}(X,\mathbb{Z})\times\mathbb{Z}, we denote by 𝔐H(v){\mathfrak{M}}_{H}(v) the moduli stack of μH\mu_{H}-semistable one-dimensional sheaves EE satisfying

[E]=β,χ(E)=m.[E]=\beta,\quad\chi(E)=m.

The stack 𝔐H(v){\mathfrak{M}}_{H}(v) admits the good moduli space p:𝔐H(v)MH(v)p\colon{\mathfrak{M}}_{H}(v)\to M_{H}(v), and we have the Hilbert–Chow morphism

(2.5) πM:MH(v)redChowβ(X)\pi_{M}\colon M_{H}(v)^{\operatorname{red}}\to\operatorname{Chow}_{\beta}(X)

sending a sheaf EE to its fundamental one cycle. Here, Chowβ(X)\operatorname{Chow}_{\beta}(X) denotes the Chow variety of compactly supported effective one cycles with homology class β\beta (see [Kol96] for the definition. Note that it is denoted as Chow(X)\operatorname{Chow}^{\prime}(X) in [Kol96]). We denote by π𝔐\pi_{\mathfrak{M}} the composition

π𝔐:𝔐H(v)redMH(v)redChowβ(X).\pi_{\mathfrak{M}}\colon{\mathfrak{M}}_{H}(v)^{\operatorname{red}}\to M_{H}(v)^{\operatorname{red}}\to\operatorname{Chow}_{\beta}(X).

Recall from Example A.2 that the stack 𝔐H(v){\mathfrak{M}}_{H}(v) is the classical truncation of a (1)(-1)-shifted derived Artin stack. In particular, the stack 𝔐H(v){\mathfrak{M}}_{H}(v) carries a natural d-critical structure and (A.2) implies that there exists a natural isomorphism

K𝔐H(v)virdet(𝐑p𝔐𝐑om(,))|𝔐H(v)red,K^{\operatorname{vir}}_{{\mathfrak{M}}_{H}(v)}\cong\det(\mathbf{R}p_{{\mathfrak{M}}*}\mathbf{R}\mathcal{H}om(\mathcal{E},\mathcal{E}))|_{{\mathfrak{M}}_{H}(v)^{\operatorname{red}}},

where p𝔐:𝔐H(v)×X𝔐H(v)p_{\mathfrak{M}}\colon{\mathfrak{M}}_{H}(v)\times X\to{\mathfrak{M}}_{H}(v) denotes the projection and \mathcal{E} denotes the universal sheaf on 𝔐H(v)×X{\mathfrak{M}}_{H}(v)\times X. In order to define the well-defined notion of Gopakumar–Vafa invariants, Maulik–Toda [MT18] and Toda [Todb] proposed the following conjecture on the virtual canonical bundle of the stack 𝔐H(v){\mathfrak{M}}_{H}(v).

Conjecture 2.7 ([Todb, Conjecture 2.10]).

The stack 𝔐H(v){\mathfrak{M}}_{H}(v) is Calabi–Yau (CY) at any point γChowβ(X)\gamma\in\operatorname{Chow}_{\beta}(X), i.e., there exists an analytic open neighborhood γUChowβ(X)\gamma\in U\subset\operatorname{Chow}_{\beta}(X) such that the virtual canonical bundle K𝔐virK^{\operatorname{vir}}_{{\mathfrak{M}}} is trivial on π𝔐1(U)\pi^{-1}_{\mathfrak{M}}(U).

Suppose that Conjecture 2.7 holds. Then we can take an orientation of π𝔐1(U)\pi^{-1}_{\mathfrak{M}}(U) with

(K𝔐vir|π𝔐1(U))1/2𝒪π𝔐1(U),\left(K^{\operatorname{vir}}_{\mathfrak{M}}|_{\pi_{\mathfrak{M}}^{-1}(U)}\right)^{1/2}\cong\mathcal{O}_{\pi^{-1}_{\mathfrak{M}}(U)},

which we call a Calabi–Yau (CY) orientation. As we have seen in §2.2, we have the associated perverse sheaf

φ𝔐H(v)|UPerv(𝔐H(v)|U).\varphi_{{\mathfrak{M}}_{H}(v)|_{U}}\in\operatorname{Perv}({\mathfrak{M}}_{H}(v)|_{U}).

We then define the perverse sheaf on the good moduli space as

(2.6) φMH(v)|U1(pφ𝔐H(v)|U)Perv(MH(v)|U),\varphi_{M_{H}(v)|_{U}}\coloneqq{\mathcal{H}^{1}}(p_{*}\varphi_{{\mathfrak{M}}_{H}(v)|_{U}})\in\operatorname{Perv}(M_{H}(v)|_{U}),

where we denote by 𝔐H(v)|U,MH(v)|U{\mathfrak{M}}_{H}(v)|_{U},M_{H}(v)|_{U} the pull-back of 𝔐H(v),MH(v){\mathfrak{M}}_{H}(v),M_{H}(v) along the open embedding UChowβ(X)U\subset\operatorname{Chow}_{\beta}(X), respectively. Note that we denote by i()\mathcal{H}^{i}(-) the ii-th perverse cohomology.

Definition 2.8.

Suppose Conjecture 2.7 holds. For an element γChowβ(X)\gamma\in\operatorname{Chow}_{\beta}(X), we define a Laurent polynomial ΦH(γ,m)\Phi_{H}(\gamma,m) as follows:

(2.7) ΦH(γ,m)iχ(i(πMφMH(v)|U))yi[y±1],\Phi_{H}(\gamma,m)\coloneqq\sum_{i\in\mathbb{Z}}\chi(\mathcal{H}^{i}(\pi_{M*}\varphi_{M_{H}(v)|_{U}}))y^{i}\in\mathbb{Z}[y^{\pm 1}],

where the perverse sheaf φMH(v)|U\varphi_{M_{H}(v)|_{U}} is defined as in (2.6).

Remark 2.9.
  1. (1)

    By [Todb, Lemma 2.14], the Laurent polynomial (2.7) is independent of the choice of a CY orientation on 𝔐H(v)|U{\mathfrak{M}}_{H}(v)|_{U}.

  2. (2)

    The definition of the perverse sheaf in (2.6) is motivated from the notion of BPS sheaves for quivers with super-potentials introduced by Davison–Meinhardt [DM20]. See [Todb, Section 2.8] for the detailed discussion.

The following χ\chi-independence conjecture is the main subject in this paper:

Conjecture 2.10 ([Todb, Conjecture 2.15]).

The Laurent polynomial (2.7) is independent of mm\in\mathbb{Z}.

At this moment, the above conjecture is known to hold in the following cases:

  • X=TotS(ωS)X=\operatorname{Tot}_{S}(\omega_{S}), where SS is a smooth projective surface, and γ\gamma is primitive [Todb].

  • X=TotS(ωS)X=\operatorname{Tot}_{S}(\omega_{S}), where SS is a del Pezzo surface, and γ\gamma is arbitrary [MSa, Yua].

  • X=TotC(𝒪(D)ωC(D))X=\operatorname{Tot}_{C}(\mathcal{O}(D)\oplus\omega_{C}(-D)), where CC is a smooth projective curve and DD is a divisor with deg(D)>2g(C)2\deg(D)>2g(C)-2, and γ\gamma is arbitrary [MSa].

Remark 2.11.

Suppose that Conjecture 2.10 holds. Then we may write the Laurent polynomial (2.7) as Φ(γ)ΦH(γ,1)=ΦH(γ,m)\Phi(\gamma)\coloneqq\Phi_{H}(\gamma,1)=\Phi_{H}(\gamma,m) for mm\in\mathbb{Z}. Note that we can drop the subscript HH in the notation since for m=1m=1, the moduli space is independent of the choice of an ample divisor HH.

Furthermore, for m=1m=1, we know that the perverse sheaf ϕMH(v)\phi_{M_{H}(v)} is Verdier self-dual. Hence there exist integers ng,γn_{g,\gamma}\in\mathbb{Z} for g0g\geq 0 such that the equation

Φ(γ)=g0ng,γ(y12+y12)2g\Phi(\gamma)=\sum_{g\geq 0}n_{g,\gamma}\left(y^{\frac{1}{2}}+y^{-\frac{1}{2}}\right)^{2g}

holds. Following Maulik–Toda [MT18], we call the integers ng,γn_{g,\gamma} as the GV invariants of XX.

2.4. Local curves and twisted Higgs bundles

In this section, we introduce a class of Calabi–Yau threefolds which we call local curves. Then we review the results on the twisted Higgs bundles due to Maulik–Shen [MSa].

2.4.1. Spectral correspondence for local curves

Let CC be a smooth projective curve and NN be a rank two vector bundle on CC with detNωC\det N\cong\omega_{C}. Then the total space XTotC(N)X\coloneqq\operatorname{Tot}_{C}(N) of the bundle NN gives an example of quasi-projective Calabi–Yau threefolds, which we call a local curve. Denote by p:XCp\colon X\to C the projection.

In this section, we recall the spectral-type correspondence for coherent sheaves on local curves. See e.g. [Sim94] for the details.

Lemma 2.12.

Giving a compactly supported pure one-dimensional coherent sheaf on XX is equivalent to giving a pair (E,ϕ)(E,\phi) of a locally free sheaf EE on CC and a morphism ϕHom(E,EN)\phi\in\operatorname{Hom}(E,E\otimes N) satisfying ϕϕ=0\phi\wedge\phi=0.

We call a pair (E,ϕ)(E,\phi) in the above lemma as an NN-Higgs bundle. We can define the slope semistability for NN-Higgs bundles as in Definition 2.6:

Definition 2.13.

Let (E,ϕ)(E,\phi) be an NN-Higgs bundle.

  1. (1)

    We define the slope of (E,ϕ)(E,\phi) as

    μ(E)χ(E)rk(E).\mu(E)\coloneqq\frac{\chi(E)}{\operatorname{rk}(E)}.
  2. (2)

    The NN-Higgs bundle (E,ϕ)(E,\phi) is μ\mu-semistable (resp. stable) if for any saturated subsheaf 0FE0\neq F\subsetneq E with ϕ(F)FN\phi(F)\subset F\otimes N, the inequality

    μ(F)μ(E)(resp. μ(F)<μ(E))\mu(F)\leq\mu(E)\quad(\mbox{resp. }\mu(F)<\mu(E))

    holds.

Lemma 2.14.

Take an ample divisor HH on CC. Let \mathcal{E} be a pure one-dimensional coherent sheaf on XX and (E,ϕ)(E,\phi) be the corresponding NN-Higgs bundle.

Then the sheaf \mathcal{E} is μpH\mu_{p^{*}H}-(semi)stable if and only if the NN-Higgs bundle (E,ϕ)(E,\phi) is μ\mu-(semi)stable.

Let 𝔐Xss(r,m)𝔐pH(r[C],m){\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m)\coloneqq{\mathfrak{M}}_{p^{*}H}(r[C],m) be the moduli stack of μpH\mu_{p^{*}H}-semistable sheaves \mathcal{E} on XX satisfying []=r[C][\mathcal{E}]=r[C] and χ()=m\chi(\mathcal{E})=m. Let MXss(r,m)M^{\operatorname{ss}}_{X}(r,m) be the good moduli space of 𝔐Xss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m). By the above lemma, \mathbb{C}-valued points of 𝔐Xss(r,m){\mathfrak{M}}^{\mathrm{ss}}_{X}(r,m) correspond to μ\mu-semistable NN-Higgs bundles.

The moduli space MXss(r,m)M^{\operatorname{ss}}_{X}(r,m) admits a Hitchin type morphism: Define a Hitchin base as

BXi=1rH0(C,Symi(N)),B_{X}\coloneqq\bigoplus_{i=1}^{r}\operatorname{H}^{0}(C,\operatorname{Sym}^{i}(N)),

and a Hitchin morphism as follows:

(2.8) hX:MXss(r,m)BX,(E,ϕ)(tr(ϕi))i=1r,h_{X}\colon M^{\operatorname{ss}}_{X}(r,m)\to B_{X},\quad(E,\phi)\mapsto(\operatorname{tr}(\phi^{i}))_{i=1}^{r},

where ϕi:EESymi(N)\phi^{i}\colon E\to E\otimes\operatorname{Sym}^{i}(N) is obtained by the ii-th iteration of ϕ\phi.

Remark 2.15.

We can construct a bijection between the sets of closed points of im(hX)\operatorname{im}(h_{X}) and im(πM)\operatorname{im}(\pi_{M}) by sending a point in im(hX)\operatorname{im}(h_{X}) to its spectral curve, where πM:MXss(r,m)redChowr[C](X)\pi_{M}\colon M^{\operatorname{ss}}_{X}(r,m)^{\operatorname{red}}\to\operatorname{Chow}_{r[C]}(X) denotes the Hilbert–Chow morphism defined as in (2.5). Moreover, by the properness of the morphisms hXh_{X} and πM\pi_{M}, the spaces im(hX)\operatorname{im}(h_{X}) and im(πM)\operatorname{im}(\pi_{M}) are homeomorphic.

As a result, the GV invariants do not change if we replace the Hilbert–Chow morphism with the Hitchin morphism. Hence we use the Hitchin morphism for the GV theory of local curves in this paper.

2.4.2. Twisted Higgs bundles

Let LL be a line bundle on a smooth projective curve CC. Denote by YTotC(L)Y\coloneqq\operatorname{Tot}_{C}(L) the total space of LL.

An LL-Higgs bundle is a pair (E,θ)(E,\theta) consisting of a locally free sheaf EE on CC and a homomorphism θHom(E,EL)\theta\in\operatorname{Hom}(E,E\otimes L). For the canonical divisor L=KCL=K_{C}, the notion of KCK_{C}-Higgs bundles agrees with the usual notion of Higgs bundles.

As in Definition 2.13, we can define the notion of μ\mu-semistability for LL-Higgs bundles. We denote by 𝔐Yss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m) the moduli stack of μ\mu-semistable LL-Higgs bundles (E,θ)(E,\theta) with rk(E)=r,χ(E)=m\operatorname{rk}(E)=r,\chi(E)=m, and MYss(r,m)M^{\operatorname{ss}}_{Y}(r,m) its good moduli space. Similarly to (2.8), we have a Hitchin morphism

(2.9) hY:MYss(r,m)BYi=1rH0(C,Li)h_{Y}\colon M^{\operatorname{ss}}_{Y}(r,m)\to B_{Y}\coloneqq\bigoplus_{i=1}^{r}\operatorname{H}^{0}(C,L^{\otimes i})

sending an LL-Higgs bundle (E,θ)(E,\theta) to (tr(θi))i=1r(\operatorname{tr}(\theta^{i}))_{i=1}^{r}.

We denote by h~Y:𝔐Yss(r,m)BY\widetilde{h}_{Y}\colon{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)\to B_{Y} the composition

(2.10) h~Y:𝔐Yss(r,m)MYss(r,m)BY.\widetilde{h}_{Y}\colon{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)\to M^{\operatorname{ss}}_{Y}(r,m)\to B_{Y}.

Given an element aBYa\in B_{Y}, we denote by CaYC_{a}\subset Y its spectral curve. Define an open dense subset UBYU\subset B_{Y} as

U{aBY:Ca is smooth},U\coloneqq\left\{a\in B_{Y}:C_{a}\mbox{ is smooth}\right\},

and let g:𝒞Ug\colon\mathcal{C}\to U be the universal spectral curve. The following result plays a key role in this paper:

Theorem 2.16 ([MSa, Theorem 0.4]).

Suppose that deg(L)>2g(C)2\deg(L)>2g(C)-2. Then we have an isomorphism

hY𝒞MYss(r,m)i=02d𝒞(i𝐑1g𝒞)[i+d],h_{Y*}\mathop{\mathcal{IC}}\nolimits_{M^{\operatorname{ss}}_{Y}(r,m)}\cong\bigoplus_{i=0}^{2d}\mathop{\mathcal{IC}}\nolimits(\wedge^{i}\mathbf{R}^{1}g_{*}\mathbb{Q}_{\mathcal{C}})[-i+d],

where dd denotes the genus of the fibers of g:𝒞Ug\colon\mathcal{C}\to U.

In particular, we have isomorphisms

hY𝒞MYss(r,m)hY𝒞MYss(r,m)h_{Y*}\mathop{\mathcal{IC}}\nolimits_{M^{\operatorname{ss}}_{Y}(r,m)}\cong h_{Y*}\mathop{\mathcal{IC}}\nolimits_{M^{\operatorname{ss}}_{Y}(r,m^{\prime})}

for all m,mm,m^{\prime}\in\mathbb{Z}.

3. Cohomological χ\chi-independence for local curves

Let CC be a smooth projective curve of genus gg and NN be a rank two vector bundle on CC with detNωC\det N\cong\omega_{C}. We put XTotC(N)X\coloneqq\operatorname{Tot}_{C}(N). The goal of this section is to prove the following theorem:

Theorem 3.1.

Let X=TotC(N)X=\operatorname{Tot}_{C}(N) be a local curve. For every positive integer r>0r\in\mathbb{Z}_{>0} and a class γBX\gamma\in B_{X}, Conjecture 2.10 holds.

3.1. Global d-critical charts for moduli spaces on local curves

We first recall the main result of the companion paper [KM21]:

Theorem 3.2.

[KM21, Theorem 5.6, Proposition 5.7] Let CC be a smooth projective curve and take a short exact sequence

(3.1) 0L1NL200\to L_{1}\to N\to L_{2}\to 0

of locally free sheaves on CC where L1L_{1} and L2L_{2} are rank one. Suppose that there exists an isomorphism det(N)ωC\det(N)\cong\omega_{C}, and the inequality deg(L2)>2g(C)2\deg(L_{2})>2g(C)-2 holds. Write X=TotC(N)X=\operatorname{Tot}_{C}(N) and Y=TotC(L2)Y=\operatorname{Tot}_{C}(L_{2}). Let 𝕸X\boldsymbol{{\mathfrak{M}}}_{X} and 𝕸Y\boldsymbol{{\mathfrak{M}}}_{Y} be the derived moduli stack of compactly supported coherent sheaves on XX and YY respectively.

  1. (i)

    There exists a function ff on 𝕸Y\boldsymbol{{\mathfrak{M}}}_{Y} such that the projection from XX to YY induces an equivalence of (1)(-1)-shifted symplectic derived Artin stacks

    (3.2) 𝕸X𝐂𝐫𝐢𝐭(f).\boldsymbol{{\mathfrak{M}}}_{X}\simeq\mathop{\mathbf{Crit}}\nolimits(f).
  2. (ii)

    Let (E,ϕ)(E,\phi) be an L2L_{2}-Higgs bundle. Then we have an equality

    f([(L2,ϕ)])=1/2α(tr(ϕ2))f([(L_{2},\phi)])=1/2\cdot\alpha(\operatorname{tr}(\phi^{2}))

    where αH0(C,L22)Ext1(L2,L1)\alpha\in\operatorname{H}^{0}(C,L_{2}^{\otimes 2})^{\vee}\cong\operatorname{Ext}^{1}(L_{2},L_{1}) is the class corresponding to the short exact sequence (3.1).

We now want to describe the moduli stack of semistable NN-Higgs bundle as a global critical locus. We begin with the following easy lemma:

Lemma 3.3.

Let CC be a smooth projective curve and NN be a rank two vector bundle on CC. Then we can take the short exact sequence (3.1) so that deg(L2)>2g(C)2\deg(L_{2})>2g(C)-2 holds. More generally, we can take L2L_{2} so that its degree is arbitrarily large.

Proof.

Let 𝒪C(1)\mathcal{O}_{C}(1) be an ample line bundle on CC. Then there exists an integer l0>0l_{0}>0 such that for every integer ll0l\geq l_{0}, the bundle N(l)N^{\vee}(l) is globally generated. Then a general element sHom(N,𝒪C(l))H0(C,N(l))s\in\operatorname{Hom}(N,\mathcal{O}_{C}(l))\cong\operatorname{H}^{0}(C,N^{\vee}(l)) is surjective. Putting L2𝒪C(l)L_{2}\coloneqq\mathcal{O}_{C}(l) and L1Ker(s)L_{1}\coloneqq\operatorname{Ker}(s), we get the desired exact sequence as in (3.1). ∎

Lemma 3.4.

Take integers r,mr,m\in\mathbb{Z} with r>0r>0. Then there exists an integer k(r)>2g(C)2k(r)>2g(C)-2 depending only on rr, such that, for any short exact sequence (3.1) with deg(L2)k(r)\deg(L_{2})\geq k(r), the following statement holds: For every μ\mu-semistable NN-Higgs bundle (E,ϕ)𝔐Xss(r,m)(E,\phi)\in{\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m), the L2L_{2}-Higgs bundle (E,sϕ)(E,s\circ\phi) is μ\mu-semistable.

Proof.

Let (E,ϕ)𝔐Xss(r,m)(E,\phi)\in{\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m) be a μ\mu-semistable NN-Higgs bundle. Suppose that the L2L_{2}-Higgs bundle (E,sϕ)(E,s\circ\phi) is not μ\mu-semistable. We claim that there exists a saturated subsheaf FEF\subset E such that μ(F)>μ(E)\mu(F)>\mu(E) and Hom(F,E/FL1)0\operatorname{Hom}(F,E/F\otimes L_{1})\neq 0. Indeed, let FEF\subset E be the maximal destabilizing subsheaf of (E,sϕ)(E,s\circ\phi). This means that we have μ(F)>μ(E)\mu(F)>\mu(E) and (sϕ)(F)FL2(s\circ\phi)(F)\subset F\otimes L_{2}. The latter condition is equivalent that the composition

FEsϕEL2E/FL2F\hookrightarrow E\xrightarrow{s\circ\phi}E\otimes L_{2}\to E/F\otimes L_{2}

is zero. On the other hand, by the μ\mu-semistability of (E,ϕ)(E,\phi), we have ϕ(F)FN\phi(F)\nsubseteq F\otimes N, i.e., the composition

FEϕENE/FNF\hookrightarrow E\xrightarrow{\phi}E\otimes N\to E/F\otimes N

is non-zero. As a result, we obtain the following diagram

F\textstyle{F\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\scriptstyle{\neq 0}0\scriptstyle{0}E/FL1\textstyle{E/F\otimes L_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E/FN\textstyle{E/F\otimes N\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E/FL2,\textstyle{E/F\otimes L_{2},}

hence we have Hom(F,E/FL1)0\operatorname{Hom}(F,E/F\otimes L_{1})\neq 0.

By Lemma 3.5 below, we can replace an exact sequence (3.1) so that Hom(F,E/FL1)=0\operatorname{Hom}(F,E/F\otimes L_{1})=0 for all μ\mu-semistable NN-Higgs bundles (E,ϕ)𝔐Xss(r,m)(E,\phi)\in{\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m) and all saturated subsheaves FEF\subset E with μ(F)>μ(E)\mu(F)>\mu(E). Hence the above argument shows that (E,sϕ)(E,s\circ\phi) remains μ\mu-semistable for such a choice of the exact sequence (3.1). ∎

Lemma 3.5.

Take integers r,mr,m\in\mathbb{Z} with r>0r>0. Then the following sets are bounded:

𝒮{F,E/F:(E,ϕ)𝔐Xss(r,m),FE is saturated with μ(F)>μ(E)},\displaystyle\mathcal{S}\coloneqq\left\{F,E/F:\begin{aligned} &(E,\phi)\in{\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m),\\ &F\subset E\mbox{ is saturated with }\mu(F)>\mu(E)\end{aligned}\right\},
HN(𝒮){A:A is a Harder–Narasimhan factor of G𝒮}.\displaystyle\operatorname{HN}(\mathcal{S})\coloneqq\left\{A:A\mbox{ is a Harder--Narasimhan factor of }G\in\mathcal{S}\right\}.

Moreover, there exists an integer k(r)k^{\prime}(r)\in\mathbb{Z}, depending only on rr, such that for all line bundles L1L_{1} with deg(L1)k(r)\deg(L_{1})\leq k^{\prime}(r) and for all F,E/F𝒮F,E/F\in\mathcal{S}, we have the vanishing Hom(F,E/FL1)=0\operatorname{Hom}(F,E/F\otimes L_{1})=0.

Proof.

The boundedness of the sets 𝒮,HN(𝒮)\mathcal{S},\operatorname{HN}(\mathcal{S}) follows from the boundedness of 𝔐Xss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m) and Grothendieck’s boundedness lemma (cf. [HL97, Lemma 1.7.9]).

In particular, there exist integers a,ba,b such that the inequalities μmin(F)a\mu_{\min}(F)\geq a and μmax(F/E)b\mu_{\max}(F/E)\leq b hold for all F,E/F𝒮F,E/F\in\mathcal{S}. By setting k(r,m)ab1k^{\prime}(r,m)\coloneqq a-b-1, we obtain the inequality

μmin(F)>μmax(E/FL1)\mu_{\min}(F)>\mu_{\max}(E/F\otimes L_{1})

for all line bundles L1L_{1} with deg(L1)k(r,m)\deg(L_{1})\leq k^{\prime}(r,m) and all F,E/F𝒮F,E/F\in\mathcal{S}.

Finally, observe that we have an isomorphism 𝔐Xss(r,m)𝔐Xss(r,m+r){\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m)\cong{\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m+r) by tensoring with a degree one line bundle. Hence by putting k(r)min{k(r,m):m=0,,r1}k^{\prime}(r)\coloneqq\min\{k^{\prime}(r,m)\colon m=0,\ldots,r-1\}, the second assertion holds. ∎

Proposition 3.6.

Let r,mr,m\in\mathbb{Z} be integers with r>0r>0. Take an exact sequence (3.1) as in Lemma 3.4. Let αH0(C,L22)Ext1(L2,L1)\alpha\in\operatorname{H}^{0}(C,L_{2}^{\otimes 2})^{\vee}\cong\operatorname{Ext}^{1}(L_{2},L_{1}) be the corresponding class. Denote by YTotC(L2)Y\coloneqq\operatorname{Tot}_{C}(L_{2}). Define the function g:BY𝔸1g\colon B_{Y}\to\mathbb{A}^{1} as

(3.3) g:BY=i=1rH0(C,L2i)𝔸1,(ai)i=1r1/2α(a2).g\colon B_{Y}=\bigoplus_{i=1}^{r}\operatorname{H}^{0}(C,L_{2}^{\otimes i})\to\mathbb{A}^{1},\quad(a_{i})_{i=1}^{r}\mapsto 1/2\cdot\alpha(a_{2}).

Then we have an isomorphism

(3.4) 𝔐Xss(r,m){d(gh~Y)=0}𝔐Yss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m)\cong\left\{d\left(g\circ\widetilde{h}_{Y}\right)=0\right\}\subset{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)

which preserves the d-critical structure.

Proof.

By Lemma 3.4, the isomorphism (3.2) restricts to the semistable loci. Then the claim follows from Theorem 3.2 (ii), the fact that the derived moduli stack 𝕸Yss(r,m)\boldsymbol{{\mathfrak{M}}}^{\operatorname{ss}}_{Y}(r,m) is a smooth (classical) stack, and that the classical truncation of the derived critical locus of a function on a smooth stack coincides with the classical critical locus. ∎

For the vanishing cycle sheaves on the good moduli spaces, we have the following result:

Proposition 3.7.

Let r,m,mr,m,m^{\prime}\in\mathbb{Z} be integers with r>0r>0. Let g:BY𝔸1g\colon B_{Y}\to\mathbb{A}^{1} be a function as in Proposition 3.6. Then we have an isomorphism

(3.5) hY(φghY(𝒞MYss(r,m)))hY(φghY(𝒞MYss(r,m))),h_{Y*}\left(\varphi_{g\circ h_{Y}}\left(\mathop{\mathcal{IC}}\nolimits_{M^{\operatorname{ss}}_{Y}(r,m)}\right)\right)\cong h_{Y*}\left(\varphi_{g\circ h_{Y}}\left(\mathop{\mathcal{IC}}\nolimits_{M^{\operatorname{ss}}_{Y}(r,m^{\prime})}\right)\right),

where hY:MYss(r,m)BYh_{Y}\colon M^{\operatorname{ss}}_{Y}(r,m)\to B_{Y}, hY:MYss(r,m)BYh_{Y}\colon M^{\operatorname{ss}}_{Y}(r,m^{\prime})\to B_{Y} denote the Hitchin morphisms (2.9) on the good moduli spaces.

Proof.

Since the Hitchin morphism hY:MYss(r,m)BYh_{Y}\colon M^{\operatorname{ss}}_{Y}(r,m)\to B_{Y} is proper, the result follows from Theorem 2.16 together with the commutativity of the vanishing cycle functors and proper push forwards. ∎

In the following subsections, we will show that the complexes in (3.5) compute the generalized GV invariants for the local curve X=TotC(N)X=\operatorname{Tot}_{C}(N).

3.2. CY property for local curves

In this subsection, we fix integers r,mr,m\in\mathbb{Z} with r>0r>0, and an exact sequence (3.1). We assume that the line bundle L2L_{2} satisfies the following conditions:

  • We have deg(L2)k(m)\deg(L_{2})\geq k(m) (see Lemma 3.4),

  • L2L_{2} is globally generated.

Recall that we denote as XTotC(N)X\coloneqq\operatorname{Tot}_{C}(N) and YTotC(L2)Y\coloneqq\operatorname{Tot}_{C}(L_{2}). By Proposition 3.6, the moduli stack 𝔐Xss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m) is written as the global critical locus inside 𝔐Yss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m).

Proposition 3.8.

The canonical bundle K𝔐Yss(r,m)K_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)} of the stack 𝔐Yss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m) is trivial, and hence so is the virtual canonical bundle K𝔐Xss(r,m)virK^{\operatorname{vir}}_{{\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m)} of the stack 𝔐Xss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m). In particular, the stack 𝔐Xss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m) is CY at any point γBX\gamma\in B_{X}, i.e., Conjecture 2.7 holds for 𝔐Xss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m).

Proof.

A similar argument can be found in [Todb, Theorem 7.1]. Take a morphism T𝔐Yss(r,m)T\to{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m) from a scheme TT. Let Coh(Y×T)\mathcal{E}\in\operatorname{Coh}(Y\times T) be the corresponding family of μ\mu-semistable one-dimensional sheaves on YY. We consider the following diagram:

YTY×T\textstyle{Y_{T}\coloneqq Y\times T\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πT\scriptstyle{\pi_{T}}pT\scriptstyle{p_{T}}T\textstyle{T}CTC×T.\textstyle{C_{T}\coloneqq C\times T.\ignorespaces\ignorespaces\ignorespaces\ignorespaces}qT\scriptstyle{q_{T}}

We need to construct an isomorphism

det𝐑omπT(,)𝒪T,\det\mathbf{R}\mathcal{H}om_{\pi_{T}}(\mathcal{E},\mathcal{E})\cong\mathcal{O}_{T},

which is functorial in TT. We have the following exact sequence

(3.6) 0pT(L21pT)pTpT0.0\to p_{T}^{*}(L_{2}^{-1}\boxtimes p_{T*}\mathcal{E})\to p_{T}^{*}p_{T*}\mathcal{E}\to\mathcal{E}\to 0.

Applying the functor 𝐑omπT(,)\mathbf{R}\mathcal{H}om_{\pi_{T}}(-,\mathcal{E}) to the exact sequence (3.6), we obtain the exact triangle

𝐑omπT(,)𝐑omqT(,)𝐑omqT(L21,),\mathbf{R}\mathcal{H}om_{\pi_{T}}(\mathcal{E},\mathcal{E})\to\mathbf{R}\mathcal{H}om_{q_{T}}(\mathcal{F},\mathcal{F})\to\mathbf{R}\mathcal{H}om_{q_{T}}(\mathcal{F}\boxtimes L_{2}^{-1},\mathcal{F}),

where we put pT\mathcal{F}\coloneqq p_{T*}\mathcal{E}. By taking the determinants, we get

(3.7) det𝐑omπT(,)det𝐑omqT(,)(det𝐑omqT(L21,))1.\det\mathbf{R}\mathcal{H}om_{\pi_{T}}(\mathcal{E},\mathcal{E})\cong\det\mathbf{R}\mathcal{H}om_{q_{T}}(\mathcal{F},\mathcal{F})\otimes(\det\mathbf{R}\mathcal{H}om_{q_{T}}(\mathcal{F}\boxtimes L_{2}^{-1},\mathcal{F}))^{-1}.

On the other hand, we have an exact sequence

0𝒪CL2𝒪Z0,0\to\mathcal{O}_{C}\to L_{2}\to\mathcal{O}_{Z}\to 0,

where Z|L2|Z\in|L_{2}| is a finite set of points. Applying the functor 𝐑omqT(,())\mathbf{R}\mathcal{H}om_{q_{T}}(\mathcal{F},\mathcal{F}\boxtimes(-)) and taking the determinants, we get

(3.8) det𝐑omqT(L21,)\displaystyle\det\mathbf{R}\mathcal{H}om_{q_{T}}(\mathcal{F}\boxtimes L_{2}^{-1},\mathcal{F}) det𝐑omqT(,L2)\displaystyle\cong\det\mathbf{R}\mathcal{H}om_{q_{T}}(\mathcal{F},\mathcal{F}\boxtimes L_{2})
det𝐑omqT(,)det𝐑omqT(,Z),\displaystyle\cong\det\mathbf{R}\mathcal{H}om_{q_{T}}(\mathcal{F},\mathcal{F})\otimes\det\mathbf{R}\mathcal{H}om_{q_{T}}(\mathcal{F},\mathcal{F}_{Z}),

where we put Z|Z×T\mathcal{F}_{Z}\coloneqq\mathcal{F}|_{Z\times T}. Combining the equations (3.7) and (3.8), we obtain the desired isomorphism

det𝐑omπT(,)\displaystyle\det\mathbf{R}\mathcal{H}om_{\pi_{T}}(\mathcal{E},\mathcal{E}) det𝐑omqT(,Z)1\displaystyle\cong\det\mathbf{R}\mathcal{H}om_{q_{T}}(\mathcal{F},\mathcal{F}_{Z})^{-1}
det𝐑omrT(Z,Z)1\displaystyle\cong\det\mathbf{R}\mathcal{H}om_{r_{T}}(\mathcal{F}_{Z},\mathcal{F}_{Z})^{-1}
i=1kdet𝐑om(pi,pi)𝒪T,\displaystyle\cong\bigotimes_{i=1}^{k}\det\mathbf{R}\mathcal{H}om(\mathcal{F}_{p_{i}},\mathcal{F}_{p_{i}})\cong\mathcal{O}_{T},

where we denote by rT:Z×TTr_{T}\colon Z\times T\to T the projection. For the third isomorphism, we put Z={p1,,pk}Z=\{p_{1},\ldots,p_{k}\} and pi|{pi}×T\mathcal{F}_{p_{i}}\coloneqq\mathcal{F}|_{\{p_{i}\}\times T}.

The triviality of the virtual canonical bundle K𝔐Xss(r,m)virK^{\operatorname{vir}}_{{\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m)} now follows from Lemma A.10. ∎

3.3. Proof of Theorem 3.1

In this subsection, we finish the proof of Theorem 3.1.

Consider the following commutative diagram:

(3.9) MXss(r,m)\textstyle{M^{\operatorname{ss}}_{X}(r,m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota}hX\scriptstyle{h_{X}}MYss(r,m)\textstyle{M^{\operatorname{ss}}_{Y}(r,m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}hY\scriptstyle{h_{Y}}BX\textstyle{B_{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}b\scriptstyle{b}BY,\textstyle{B_{Y},}

where the morphism b:BXBYb\colon B_{X}\to B_{Y} is induced by the surjection Symk(N)L2k\operatorname{Sym}^{k}(N)\to L_{2}^{\otimes k} for k=1,,rk=1,\ldots,r.

Lemma 3.9.

The morphism

(3.10) b|im(hX):im(hX)BY,b|_{\operatorname{im}(h_{X})}\colon\operatorname{im}(h_{X})\to B_{Y},

is finite.

Proof.

It is enough to show that the morphism in (3.10) is proper and affine. The composition hYι=bhX:MXss(r,m)BYh_{Y}\circ\iota=b\circ h_{X}\colon M^{\operatorname{ss}}_{X}(r,m)\to B_{Y} is proper as so are hYh_{Y} and ι\iota. Furthermore, the morphism hX:MXss(r,m)im(hX)h_{X}\colon M^{\operatorname{ss}}_{X}(r,m)\to\operatorname{im}(h_{X}) is proper and surjective. Hence the morphism (3.10) is proper.

On the other hand, by the properness of the Hitchin morphism hXh_{X}, the inclusion im(hX)BX\operatorname{im}(h_{X})\hookrightarrow B_{X} is closed. As the morphism b:BXBYb\colon B_{X}\to B_{Y} is just the projection of affine spaces, it is also affine. We conclude that the composition

im(hX)BX𝑏BY\operatorname{im}(h_{X})\hookrightarrow B_{X}\xrightarrow{b}B_{Y}

is affine, as required. ∎

Recall from (2.9) and (2.10) that we denote by hY:MYss(r,m)BYh_{Y}\colon M^{\operatorname{ss}}_{Y}(r,m)\to B_{Y}, h~Y:𝔐Yss(r,m)BY\widetilde{h}_{Y}\colon{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)\to B_{Y} the Hitchin morphisms. Recall also that we have the function g:BY𝔸1g\colon B_{Y}\to\mathbb{A}^{1} defined in Proposition 3.6. We equip 𝔐Xss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m) with the orientation defined by the global critical chart description in Proposition 3.6 and Lemma A.10. We define the vanishing cycle complex φ𝔐Xss(r,m)Perv(𝔐Xss(r,m))\varphi_{{\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m)}\in\operatorname{Perv}({\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m)) using this orientation. We set

(3.11) φMXss(r,m)1(pXφ𝔐Xss(r,m))Perv(MXss(r,m))\varphi_{M^{\operatorname{ss}}_{X}(r,m)}\coloneqq{\mathcal{H}}^{1}(p_{X*}\varphi_{{\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m)})\in\operatorname{Perv}(M^{\operatorname{ss}}_{X}(r,m))

We need the following proposition:

Proposition 3.10.

We have isomorphisms

φMXss(r,m)1(pX(φgh~Y(𝒞𝔐Yss(r,m))))φghY(𝒞MYss(r,m)).\varphi_{M^{\operatorname{ss}}_{X}(r,m)}\cong{\mathcal{H}^{1}}(p_{X*}(\varphi_{g\circ\widetilde{h}_{Y}}(\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)})))\cong\varphi_{g\circ h_{Y}}(\mathop{\mathcal{IC}}\nolimits_{M^{\operatorname{ss}}_{Y}(r,m)}).

We postpone the proof of this proposition until the next section.

Proof of Theorem 3.1.

Let us take integers r,m,mr,m,m^{\prime}\in\mathbb{Z} with r>0r>0. Let k(r)k(r)\in\mathbb{Z} be integers as in Lemma 3.4. We take an exact sequence (3.1) such that L2L_{2} is globally generated and deg(L2)k(r)\deg(L_{2})\geq k(r) holds. Then by Proposition 3.6, the moduli stacks 𝔐Xss(r,m),𝔐Xss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m),{\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m^{\prime}) are written as the global critical loci inside the stacks 𝔐Yss(r,m),𝔐Yss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m),{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m^{\prime}), respectively.

Recall from Proposition 3.8 that the canonical bundles K𝔐Yss(r,m)K_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)} and K𝔐Yss(r,m)K_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m^{\prime})} are trivial, hence the natural orientation data in Lemma A.10 is a CY orientation data.

Let φMXss(r,m),φMXss(r,m)\varphi_{M^{\operatorname{ss}}_{X}(r,m)},\varphi_{M^{\operatorname{ss}}_{X}(r,m^{\prime})} be the associated perverse sheaves on MXss(r,m)M^{\operatorname{ss}}_{X}(r,m), MXss(r,m)M^{\operatorname{ss}}_{X}(r,m^{\prime}), defined as in (3.11). We have an isomorphism

(3.12) hY(φMXss(r,m))hY(φMXss(r,m))h_{Y*}(\varphi_{M^{\operatorname{ss}}_{X}(r,m)})\cong h_{Y*}(\varphi_{M^{\operatorname{ss}}_{X}(r,m^{\prime})})

by Proposition 3.7. By using the commutative diagram (3.9), we can rewrite the left hand side of (3.12) as

hY(φMXss(r,m))bhX(φMXss(r,m)).h_{Y*}(\varphi_{M^{\operatorname{ss}}_{X}(r,m)})\cong b_{*}h_{X*}(\varphi_{M^{\operatorname{ss}}_{X}(r,m)}).

By Lemma 3.9, the map (3.10) is finite. Since the push-forward along a finite morphism preserves the perverse t-structures, we obtain

(3.13) i(hY(φMXss(r,m)))bi(hX(φMXss(r,m))){\mathcal{H}^{i}}(h_{Y*}(\varphi_{M^{\operatorname{ss}}_{X}(r,m)}))\cong b_{*}{\mathcal{H}^{i}}(h_{X*}(\varphi_{M^{\operatorname{ss}}_{X}(r,m)}))

for all ii\in\mathbb{Z}, and we have the same isomorphisms if we replace the integer mm with mm^{\prime}.

Combining the isomorphisms (3.12) and (3.13), and taking the Euler characteristics, we conclude that

χ(i(hX(φMXss(r,m))))=χ(i(hX(φMXss(r,m))))\chi({\mathcal{H}^{i}}(h_{X*}(\varphi_{M^{\operatorname{ss}}_{X}(r,m)})))=\chi({\mathcal{H}^{i}}(h_{X*}(\varphi_{M^{\operatorname{ss}}_{X}(r,m^{\prime})})))

as desired. ∎

4. Cohomological integrality theorem for twisted Higgs bundles

In this section, we prove the cohomological integrality theorem in the sense of [DM20, §1.3] for twisted Higgs bundles. Since semistable twisted Higgs bundles form a category of homological dimension one, we can prove the cohomological integrality theorem using the techniques of [DM20, Mei], which treat the case of quivers.

4.1. Mixed Hodge modules on stacks

Here we give a quick introduction to mixed Hodge modules, which is a sheaf theoretic version of mixed Hodge structures introduced by Morihiko Saito [Sai90]. An advantage of working with mixed Hodge modules (rather than perverse sheaves) is the fact that the category of pure Hodge modules is semi-simple. In particular, an equality of Grothendieck group of the category of pure Hodge modules implies an isomorphism between them. This was used by Davison–Meinhardt [DM20] in their proof of the cohomological integrality theorem for quivers with potentials, and will be used in the proof of Theorem 4.6.

Let XX be a separated scheme locally of finite type over complex number whose connected components are quasi-compact. For such XX, we can define the category of mixed Hodge modules MHM(X)\operatorname{MHM}(X) and its bounded derived category Db(MHM(X))D^{b}(\operatorname{MHM}(X)) which admits a six-functor formalism (see [Sai89] for an overview). There exists an exact functor

rat:Db(MHM(X))Db(Perv(X))\operatorname{rat}\colon D^{b}(\operatorname{MHM}(X))\to D^{b}(\operatorname{Perv}(X))

which restricts to a faithful functor MHM(X)Perv(X)\operatorname{MHM}(X)\to\operatorname{Perv}(X). The functor rat\operatorname{rat} is compatible with all six functors. A mixed Hodge module MM is equipped with an increasing filtration called the weight filtration which we denote by WMW_{\bullet}M. A mixed Hodge module MMHM(X)M\in\operatorname{MHM}(X) is called pure of weight ii if Wi1M=0W_{i-1}M=0 and WiM=MW_{i}M=M holds, and an object MDb(MHM(X))M^{\bullet}\in D^{b}(\operatorname{MHM}(X)) is called pure if the ii-th cohomology mixed Hodge module i(M)\mathcal{H}^{i}(M) is pure of weight ii.

The category of mixed Hodge modules over a point is equivalent to the category of graded polarizable mixed Hodge structures. Let aX:XSpeca_{X}\colon X\to\operatorname{Spec}\mathbb{C} be the constant map to a point. Then we define objects X,𝔻XDb(MHM(X))\mathbb{Q}_{X},\mathbb{D}\mathbb{Q}_{X}\in D^{b}(\operatorname{MHM}(X)) as

XaX,𝔻XaX!,\mathbb{Q}_{X}\coloneqq a_{X}^{*}\mathbb{Q},\ \mathbb{D}\mathbb{Q}_{X}\coloneqq a_{X}^{!}\mathbb{Q},

where \mathbb{Q} denotes the mixed Hodge structure of weight zero and dimension one.

The category of mixed Hodge modules forms a stack in the smooth topology (see [Ach, Theorem 2.3]). This motivates the following definition of the category of mixed Hodge modules on an Artin stack 𝔛{\mathfrak{X}}.

Definition 4.1.

Let 𝔛{\mathfrak{X}} be a complex Artin stack. We let Sch/𝔛sm,sep\operatorname{Sch}^{\operatorname{sm},\operatorname{sep}}_{/{\mathfrak{X}}} denote the category of separated schemes smooth and of finite type over 𝔛{\mathfrak{X}}. A mixed Hodge module on 𝔛{\mathfrak{X}} is a pair consisting of an assignment

Sch/𝔛sm,sep(t:T𝔛)MtMHM(T)\operatorname{Sch}^{\operatorname{sm},\operatorname{sep}}_{/{\mathfrak{X}}}\ni(t\colon T\to{\mathfrak{X}})\mapsto M_{t}\in\operatorname{MHM}(T)

and a choice of an isomorphism

θq:qMt2[dimq]Mt1\theta_{q}\colon q^{*}M_{t_{2}}[\dim q]\cong M_{t_{1}}

for each smooth morphism q:(t1:T1𝔛)(t2:T2𝔛)q\colon(t_{1}\colon T_{1}\to{\mathfrak{X}})\to(t_{2}\colon T_{2}\to{\mathfrak{X}}) in Sch/𝔛sm,sep\operatorname{Sch}^{\operatorname{sm},\operatorname{sep}}_{/{\mathfrak{X}}} satisfying the associativity relation. Mixed Hodge modules on 𝔛{\mathfrak{X}} form a category MHM(𝔛)\operatorname{MHM}({\mathfrak{X}}) in the natural way. We have a natural forgetful functor

rat:MHM(𝔛)Perv(𝔛).\operatorname{rat}\colon\operatorname{MHM}({\mathfrak{X}})\to\operatorname{Perv}({\mathfrak{X}}).

Take a smooth surjective morphism from a separated finite type scheme t:T𝔛t\colon T\to{\mathfrak{X}} and we let pri:T×𝔛TT\operatorname{pr}_{i}\colon T\times_{{\mathfrak{X}}}T\to T denote the ii-th projection. Then we can identify MHM(𝔛)\operatorname{MHM}({\mathfrak{X}}) with the category of pairs (M,σ)(M,\sigma), where MM is a mixed Hodge module on TT and σ\sigma is an isomorphism

σ:pr1Mpr2M\sigma\colon\operatorname{pr}_{1}^{*}M\cong\operatorname{pr}_{2}^{*}M

satisfying the cocycle condition.

At present we do not have a full six-functor formalism for mixed Hodge modules on Artin stacks. However we have some part of it which is sufficient for applications in this paper. Firstly, if we are given a smooth morphism q:𝔛1𝔛2q\colon{\mathfrak{X}}_{1}\to{\mathfrak{X}}_{2} between Artin stacks, we can define a functor

q[dimq]:MHM(𝔛2)MHM(𝔛1)q^{*}[\dim q]\colon\operatorname{MHM}({\mathfrak{X}}_{2})\to\operatorname{MHM}({\mathfrak{X}}_{1})

in the standard way.

Now assume that we are given a finite type morphism p:𝔛Xp\colon{\mathfrak{X}}\to X from an Artin stack to a separated finite type scheme. We want to define the functor

n(p()):MHM(𝔛)MHM(X){\mathcal{H}}^{n}(p_{*}(-))\colon\operatorname{MHM}({\mathfrak{X}})\to\operatorname{MHM}(X)

compatible with the functor rat\operatorname{rat}. Here n{\mathcal{H}}^{n} denotes the nn-th cohomology with respect to the perverse t-structure on D(MHM(X))D(\operatorname{MHM}(X)). We assume that the morphism pp satisfies the following assumption:

()\mathrm{(*)} For each object Dcb(𝔛){\mathcal{F}}\in D^{b}_{c}({\mathfrak{X}}) in the bounded derived category of sheaves on 𝔛{\mathfrak{X}} with constructible cohomology and integer NN, there exists a smooth morphism from a scheme qN:TN𝔛q_{N}\colon T_{N}\to{\mathfrak{X}} such that the natural map

n()n(qNqN)\mathcal{H}^{n}({\mathcal{F}})\to\mathcal{H}^{n}({q_{N}}_{*}q_{N}^{*}{\mathcal{F}})

is isomorphism for each nNn\leq N. Here n\mathcal{H}^{n} denotes the perverse t-structure on Dcb(𝔛)D^{b}_{c}({\mathfrak{X}}).

This assumption is automatically satisfied when 𝔛{\mathfrak{X}} is of the form [Y/G][Y/G] for some scheme YY and a linear algebraic group GG (see [Dava, §2.3.2]). Let p:𝔛Xp\colon{\mathfrak{X}}\to X be a morphism to a scheme. For a mixed Hodge module MMHM(𝔛)M\in\operatorname{MHM}({\mathfrak{X}}) and nn\in\mathbb{Z}, we define a mixed Hodge module

n(pM)n((pqN)qNM)\mathcal{H}^{n}(p_{*}M)\coloneqq\mathcal{H}^{n}((p\circ q_{N})_{*}q_{N}^{*}M)

where NN is a sufficiently large integer. We can show that n(pM)\mathcal{H}^{n}(p_{*}M) is independent of the choice of NN and qNq_{N}. If we take pp as the constant map a𝔛:𝔛Speca_{{\mathfrak{X}}}\colon{\mathfrak{X}}\to\operatorname{Spec}\mathbb{C}, we can construct a mixed Hodge structure Hn(𝔛,M)n(a𝔛M)\operatorname{H}^{n}({\mathfrak{X}},M)\coloneqq\mathcal{H}^{n}({a_{{\mathfrak{X}}}}_{*}M). Similarly, we can extend the perverse sheaves n(p𝔛)\mathcal{H}^{n}(p_{*}\mathbb{Q}_{{\mathfrak{X}}}) and n(p𝔻𝔛)\mathcal{H}^{n}(p_{*}\mathbb{D}_{{\mathfrak{X}}}) to mixed Hodge modules, and the vector spaces Hn(𝔛)\operatorname{H}^{n}({\mathfrak{X}}) and HnBM(𝔛)\operatorname{H}^{\mathrm{BM}}_{n}({\mathfrak{X}}) to mixed Hodge structures.

For a complex of mixed Hodge modules MD(MHM(X))M\in D(\operatorname{MHM}(X)), we define

(M)ii(M)[i].{\mathcal{H}}(M)\coloneqq\bigoplus_{i\in\mathbb{N}}{\mathcal{H}}^{i}(M)[-i].
Lemma 4.2.

Let 𝔛{\mathfrak{X}} be a stack satisfying the condition ()\mathrm{(*)}, p:𝔛Xp\colon{\mathfrak{X}}\to X be a morphism to a separated finite type complex scheme, and h:XBh\colon X\to B be a proper morphism between separated finite type complex schemes. Take MDb(MHM(X))M\in D^{b}(\operatorname{MHM}(X)) and assume that (pM){\mathcal{H}}(p_{*}M) is pure. Then we have an isomorphism

(h(pM))((hp)M).{\mathcal{H}}(h_{*}{\mathcal{H}}(p_{*}M))\cong{\mathcal{H}}((h\circ p)_{*}M).
Proof.

Take an integer nn and an integer NN such that N>n+dimh1(x)N>n+\dim h^{-1}(x) holds for each xBx\in B. Then [Dim04, Corollary 5.2.14] implies that we have an isomorphism

n(h(pM))n(hτN(pM)).{\mathcal{H}}^{n}(h_{*}{\mathcal{H}}(p_{*}M))\cong{\mathcal{H}}^{n}(h_{*}\tau^{\leq N}{\mathcal{H}}(p_{*}M)).

Take a smooth morphism q:TXq\colon T\to X such that we have isomorphisms

τN(pM))\displaystyle\tau^{\leq N}{\mathcal{H}}(p_{*}M)) τN((pq)qM)),\displaystyle\cong\tau^{\leq N}{\mathcal{H}}((p\circ q)_{*}q^{*}M)),
n((hp)M)\displaystyle{\mathcal{H}}^{n}((h\circ p)_{*}M) n((hpq)qM).\displaystyle\cong{\mathcal{H}}^{n}((h\circ p\circ q)_{*}q^{*}M).

Then what it is enough to prove the following isomorphism

n(hτN((pq)qM)))n((hpq)qM).{\mathcal{H}}^{n}(h_{*}\tau^{\leq N}{\mathcal{H}}((p\circ q)_{*}q^{*}M)))\cong{\mathcal{H}}^{n}((h\circ p\circ q)_{*}q^{*}M).

Saito’s decomposition theorem implies an isomorphism

n(hτN((pq)qM)))n(hτN(pq)qM)){\mathcal{H}}^{n}(h_{*}\tau^{\leq N}{\mathcal{H}}((p\circ q)_{*}q^{*}M)))\cong{\mathcal{H}}^{n}(h_{*}\tau^{\leq N}(p\circ q)_{*}q^{*}M))

Then using [Dim04, Corollary 5.2.14] again, we obtain the desired isomorphism. ∎

4.2. Monodromic mixed Hodge modules

Here we recall some basic properties of monodromic mixed Hodge modules. We do not give the precise definition here and refer the reader to [DM20, §2] and [BBBBJ15, §2.9] for the detailed discussion. Let XX be a separated scheme locally of finite type over complex number whose connected components are quasi-compact. Then we can define an abelian category MMHM(X)\operatorname{MMHM}(X) of monodromic mixed Hodge modules on XX. Roughly speaking, a monodromic mixed Hodge module consists of its underlying mixed Hodge module MM and a monodromy operator acting on it. We have a natural functor

MMHM(X)MHM(X)\operatorname{MMHM}(X)\to\operatorname{MHM}(X)

forgetting the monodromy operator and a fully faithful functor

MHM(X)MMHM(X)\operatorname{MHM}(X)\hookrightarrow\operatorname{MMHM}(X)

which associates a mixed Hodge module MM to a monodromic mixed Hodge module whose underlying mixed Hodge module is MM and the monodromy operator is trivial. As similar to the usual mixed Hodge modules, monodromic mixed Hodge modules are also equipped with weight filtrations.

The bounded derived category Db(MMHM(X))D^{b}(\operatorname{MMHM}(X)) admits a six-functor formalism, similarly to the usual mixed Hodge modules. The inclusion functor Db(MHM(X))Db(MMHM(X))D^{b}(\operatorname{MHM}(X))\to D^{b}(\operatorname{MMHM}(X)) is compatible with these six operations. The forgetful functor Db(MMHM(X))Db(MHM(X))D^{b}(\operatorname{MMHM}(X))\to D^{b}(\operatorname{MHM}(X)) is compatible with four operations f,f!,f,f!f_{*},f_{!},f^{*},f^{!} for a morphism ff between separated finite type complex schemes. However, the tensor product of monodromic mixed Hodge modules is not compatible with the tensor product of the underlying mixed Hodge modules.

For a regular function f:X𝔸1f\colon X\to\mathbb{A}^{1}, we can define the monodromic vanishing cycle functor for (possibly unbounded) mixed Hodge module complexes

(4.1) φfmmhm:D(MHM(X))D(MMHM(X)),\varphi^{\mathrm{mmhm}}_{f}\colon D(\operatorname{MHM}(X))\to D(\operatorname{MMHM}(X)),

which enhances the usual vanishing cycle functor by incorporating the monodromy operator acting on it.

The essential difference between monodromic and the usual mixed Hodge modules are the following:

  • Thom–Sebastiani isomorphism holds for the monodromic vanishing cycle functors (4.1). See [DM20, Proposition 2.13] for the precise statement and other basic properties.

  • There exists an object 𝕃1/2Db(MMHM(pt))\mathbb{L}^{1/2}\in D^{b}(\operatorname{MMHM}(\operatorname{pt})) with an isomorphism

    (𝕃1/2)2𝕃Db(MMHM(pt)),(\mathbb{L}^{1/2})^{\otimes 2}\cong\mathbb{L}\in D^{b}(\operatorname{MMHM}(\operatorname{pt})),

    where we put 𝕃Hc(𝔸1,)Db(MHM(pt))Db(MMHM(pt))\mathbb{L}\coloneqq\operatorname{H}_{c}^{*}(\mathbb{A}^{1},\mathbb{Q})\in D^{b}(\operatorname{MHM}(\operatorname{pt}))\subset D^{b}(\operatorname{MMHM}(\operatorname{pt})), which is concentrated in cohomological degree two, and is pure of weight two.

When XX is an irreducible variety, we define the object 𝒞XMMHM(X)\mathop{\mathcal{IC}}\nolimits_{X}\in\operatorname{MMHM}(X) as follows:

𝒞X𝕃dimX/2𝒞~XMMHM(X),\mathop{\mathcal{IC}}\nolimits_{X}\coloneqq\mathbb{L}^{-\dim X/2}\otimes\widetilde{\mathop{\mathcal{IC}}\nolimits}_{X}\in\operatorname{MMHM}(X),

where 𝒞~X\widetilde{\mathop{\mathcal{IC}}\nolimits}_{X} denotes the intermediate extension of Xreg\mathbb{Q}_{X_{\operatorname{reg}}} on the regular locus XregXX_{\operatorname{reg}}\subset X. We will also use the following object:

(4.2) H(B)vir𝕃1/2H(B)D(MMHM(pt)).\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})_{\operatorname{vir}}\coloneqq\mathbb{L}^{1/2}\otimes\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})\in D(\operatorname{MMHM}(\operatorname{pt})).

As in the previous subsection, we can define the notion of monodromic mixed Hodge modules for an Artin stack. In particular, we can define the object 𝒞𝔛MMHM(𝔛)\mathop{\mathcal{IC}}\nolimits_{\mathfrak{X}}\in\operatorname{MMHM}({\mathfrak{X}}) for a smooth Artin stack 𝔛{\mathfrak{X}}. Moreover, we can define the functor

n(p()):MMHM(𝔛)MMHM(X)\mathcal{H}^{n}(p_{*}(-))\colon\operatorname{MMHM}({\mathfrak{X}})\to\operatorname{MMHM}(X)

for a morphism p:𝔛Xp\colon{\mathfrak{X}}\to X from an Artin stack 𝔛{\mathfrak{X}} to a scheme XX satisfying the condition (*) in §4.1.

Let XX be a separated scheme locally of finite type over \mathbb{C} whose connected components are quasi-compact. We say that a (possibly unbounded) complex MD(MMHM(X))M\in D(\operatorname{MMHM}(X)) is locally finite if for each connected component ZXZ\subset X, the following conditions hold:

  • For each nn\in\mathbb{Z}, the set {igrnWi(M)|Z0}\{i\in\mathbb{Z}\mid\mathrm{gr}_{n}^{W}{\mathcal{H}}^{i}(M)|_{Z}\neq 0\} is finite.

  • There exists an integer nn such that Wni(M)|Z=0W_{n}{\mathcal{H}}^{i}(M)|_{Z}=0 for all ii.

We let D,lf(MMHM(X))D(MMHM(X))D^{\geq,\mathrm{lf}}(\operatorname{MMHM}(X))\subset D^{\geq}(\operatorname{MMHM}(X)) denote the full subcategory consisting of locally finite monodromic mixed Hodge complexes. We can see that the Grothendieck group K0(D,lf(MMHM(X)))K_{0}(D^{\geq,\mathrm{lf}}(\operatorname{MMHM}(X))) is isomorphic to the completion of K0(MMHM(X))K_{0}(\operatorname{MMHM}(X)) with respect to ideals {Ii}i\{I_{i}\}_{i\in\mathbb{Z}} where IiI_{i} is generated by objects whose weight is greater than ii.

Let (X,m)(X,m) be a monoid scheme where XX is a separated and locally of finite type over complex number whose connected components are quasi-compact and m:X×XXm\colon X\times X\to X is a finite morphism. For objects M,ND,lf(MMHM(X))M,N\in D^{\geq,\mathrm{lf}}(\operatorname{MMHM}(X)), we define

MmNm(MN)D,lf(MMHM(X)).M\boxtimes_{m}N\coloneqq m_{*}(M\boxtimes N)\in D^{\geq,\mathrm{lf}}(\operatorname{MMHM}(X)).

The functor m\boxtimes_{m} defines a symmetric monoidal structure on the category D,lf(MMHM(X))D^{\geq,\mathrm{lf}}(\operatorname{MMHM}(X)). Therefore for each n>0n\in\mathbb{Z}_{>0}, we can define the symmetric product functor

Symmn:D,lf(MMHM(X))D,lf(MMHM(X)).\operatorname{Sym}_{\boxtimes_{m}}^{n}\colon D^{\geq,\mathrm{lf}}(\operatorname{MMHM}(X))\to D^{\geq,\mathrm{lf}}(\operatorname{MMHM}(X)).

4.3. Approximation by proper morphisms

Let LL be a line bundle on a smooth projective curve CC with deg(L)>2g(C)2\deg(L)>2g(C)-2, and put YTotC(L)Y\coloneqq\operatorname{Tot}_{C}(L). For given integers r,mr,m\in\mathbb{Z} with r>0r>0, we denote by p:𝔐Yss(r,m)MYss(r,m)p\colon{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)\to M_{Y}^{\operatorname{ss}}(r,m) the morphism from the moduli stack to its good moduli space. We fix a regular function F:MYss(r,m)𝔸1F\colon M^{\operatorname{ss}}_{Y}(r,m)\to\mathbb{A}^{1} and denote by F~Fp:𝔐Yss(r,m)𝔸1\widetilde{F}\coloneqq F\circ p\colon{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)\to\mathbb{A}^{1} the composition.

Let us recall the construction of moduli spaces of framed objects following [DM20, MSa, Mei]. We follow the notations in [MSa, §3.3]. By construction, we have 𝔐Yss(r,m)=[Quotss/GLn]{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)=\left[\operatorname{Quot}^{\operatorname{ss}}/\operatorname{GL}_{n}\right], where

QuotssQuot\operatorname{Quot}^{\operatorname{ss}}\subset\operatorname{Quot}

is the GIT semistable locus inside a certain quot scheme Quot\operatorname{Quot} with respect to a certain GLn\operatorname{GL}_{n}-linearization. For a given integer f>nf>n, we put

𝔸Hom(,n)f,Gn×GLn.\mathbb{A}\coloneqq\operatorname{Hom}(\mathbb{C},\mathbb{C}^{n})^{f},\quad G_{n}\coloneqq\mathbb{C}^{*}\times\operatorname{GL}_{n}.

We have a GnG_{n}-action on Quot\operatorname{Quot} which passes through the GLn\operatorname{GL}_{n}-action. We define a GnG_{n}-action on 𝔸\mathbb{A} as follows:

(t,g)(ai)(t1aig),(t,g)Gn,(ai)i=1f𝔸.(t,g)\cdot(a_{i})\coloneqq(t^{-1}a_{i}g),\quad(t,g)\in G_{n},(a_{i})_{i=1}^{f}\in\mathbb{A}.

By choosing certain GnG_{n}-linearizations on 𝔸\mathbb{A} and Quot×𝔸\operatorname{Quot}\times\mathbb{A}, we obtain the diagram

(4.3) Uf\textstyle{U_{f}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\kappa}Mf\textstyle{M_{f}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πf\scriptstyle{\pi_{f}}𝒳f\textstyle{\mathcal{X}_{f}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MYss(r,m),\textstyle{M^{\operatorname{ss}}_{Y}(r,m),}

where we put

Uf(Quotss×𝔸ss)/PGn,Mf(Quot×𝔸)ss/PGn,\displaystyle U_{f}\coloneqq(\operatorname{Quot}^{\operatorname{ss}}\times\mathbb{A}^{\operatorname{ss}})/\operatorname{PG}_{n},\quad M_{f}\coloneqq(\operatorname{Quot}\times\mathbb{A})^{\operatorname{ss}}/\operatorname{PG}_{n},
𝒳f[Quotss×𝔸/PGn].\displaystyle\mathcal{X}_{f}\coloneqq[\operatorname{Quot}^{\operatorname{ss}}\times\mathbb{A}/\operatorname{PG}_{n}].

The diagram (4.3) satisfies the following properties (cf. [MSa, Proposition 3.6]):

  • the horizontal morphisms are open immersions,

  • UfU_{f} and MfM_{f} are smooth schemes and the morphism πf\pi_{f} is projective,

  • we have limfcodim𝒳f(𝒲f)=\lim_{f\to\infty}\operatorname{codim}_{\mathcal{X}_{f}}(\mathcal{W}_{f})=\infty, where 𝒲f𝒳f\mathcal{W}_{f}\subset\mathcal{X}_{f} denotes the complement of UfU_{f}.

By the following proposition, we can compute the cohomology objects n(pφF~mmhm𝒞𝔐Yss(r,m))\mathcal{H}^{n}(p_{*}\varphi^{\mathrm{mmhm}}_{\widetilde{F}}\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)}) using the push-forward along the proper morphism πf\pi_{f}:

Proposition 4.3 (cf. [DM20, Lemma 4.1, Proposition 4.3]).

The following statements hold:

  1. (1)

    For each nn\in\mathbb{Z}, there exists f0f\gg 0 such that

    n(pφF~mmhm𝒞𝔐Yss(r,m))n(πfφFπfmmhm𝒞Mf).\mathcal{H}^{n}(p_{*}\varphi^{\mathrm{mmhm}}_{\widetilde{F}}\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)})\cong\mathcal{H}^{n}(\pi_{f*}\varphi^{\mathrm{mmhm}}_{F\circ\pi_{f}}\mathop{\mathcal{IC}}\nolimits_{M_{f}}).
  2. (2)

    We have an isomorphism

    (pφF~mmhm𝒞𝔐Yss(r,m))φFmmhm(p𝒞𝔐Yss(r,m)).\mathcal{H}(p_{*}\varphi^{\mathrm{mmhm}}_{\widetilde{F}}\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)})\cong\varphi^{\mathrm{mmhm}}_{F}\mathcal{H}(p_{*}\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)}).
Proof.

We just give an outline of the proof. See [DM20, Lemma 4.1, Proposition 4.3] for the details. Using the fact limfcodim𝒳f(𝒲f)=\lim_{f\to\infty}\operatorname{codim}_{\mathcal{X}_{f}}(\mathcal{W}_{f})=\infty, we can check that the morphism πf:MfMYss(r,m)\pi_{f}\colon M_{f}\to M^{\operatorname{ss}}_{Y}(r,m) approximates the map p:𝔐Yss(r,m)MYss(r,m)p\colon{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)\to M^{\operatorname{ss}}_{Y}(r,m) in the sense of (*) in §4.1. Hence the first assertion holds.

The second assertion now follows from the natural isomorphism

φFmmhmπfπfφπfFmmhm\varphi^{\mathrm{mmhm}}_{F}\circ\pi_{f*}\cong\pi_{f*}\circ\varphi^{\mathrm{mmhm}}_{\pi_{f}\circ F}

between functors, which holds since the morphism πf:MfMYss(r,m)\pi_{f}\colon M_{f}\to M^{\operatorname{ss}}_{Y}(r,m) is proper. ∎

The following statement will be used in §5.3.

Proposition 4.4.

Let 𝔐Yss(r,m){\mathfrak{Z}}\subset{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m) be the critical locus of F~\widetilde{F} and ZMYss(r,m)Z\subset M^{\operatorname{ss}}_{Y}(r,m) be its good moduli space. Given a morphism q:ZWq\colon Z\to W between schemes, we have an isomorphism

(q((p|)φF~mmhm𝒞𝔐Yss(r,m)))((qp|)φF~mmhm𝒞𝔐Yss(r,m)){\mathcal{H}}(q_{*}\mathcal{H}((p|_{{\mathfrak{Z}}})_{*}\varphi^{\mathrm{mmhm}}_{\widetilde{F}}\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)}))\cong\mathcal{H}((q\circ p|_{{\mathfrak{Z}}})_{*}\varphi^{\mathrm{mmhm}}_{\widetilde{F}}\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)})
Proof.

Fix an integer nn. We let ZfZ_{f} denote the critical locus of the function FπfF\circ\pi_{f}. Take a sufficiently large integer ff such that the following isomorphism holds:

n((qp|)φF~mmhm𝒞𝔐Yss(r,m))n((qπf|Zf)φFπfmmhm𝒞Mf).\displaystyle\mathcal{H}^{n}((q\circ p|_{{\mathfrak{Z}}})_{*}\varphi^{\mathrm{mmhm}}_{\widetilde{F}}\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)})\cong\mathcal{H}^{n}((q\circ\pi_{f}|_{Z_{f}})_{*}\varphi^{\mathrm{mmhm}}_{F\circ\pi_{f}}\mathop{\mathcal{IC}}\nolimits_{M_{f}}).

We have the following isomorphisms

n((qπf|Zf)φFπfmmhm𝒞Mf)\displaystyle\mathcal{H}^{n}((q\circ\pi_{f}|_{Z_{f}})_{*}\varphi^{\mathrm{mmhm}}_{F\circ\pi_{f}}\mathop{\mathcal{IC}}\nolimits_{M_{f}}) n(qφFπfmmhm(πf𝒞Mf))\displaystyle\cong\mathcal{H}^{n}(q_{*}\varphi^{\mathrm{mmhm}}_{F\circ\pi_{f}}(\pi_{f*}\mathop{\mathcal{IC}}\nolimits_{M_{f}}))
n(qφFπfmmhm((πf𝒞Mf)))\displaystyle\cong\mathcal{H}^{n}(q_{*}\varphi^{\mathrm{mmhm}}_{F\circ\pi_{f}}({\mathcal{H}}(\pi_{f*}\mathop{\mathcal{IC}}\nolimits_{M_{f}})))
n(q(φFπfmmhm(πf𝒞Mf)))\displaystyle\cong\mathcal{H}^{n}(q_{*}{\mathcal{H}}(\varphi^{\mathrm{mmhm}}_{F\circ\pi_{f}}(\pi_{f*}\mathop{\mathcal{IC}}\nolimits_{M_{f}})))

where the second isomorphism follows from Saito’s decomposition theorem. If ff is sufficiently large, we also have an isomorphism

n(q(φFπfmmhm(πf𝒞Mf)))n(q(pφF~mmhm𝒞𝔐Yss(r,m)))\mathcal{H}^{n}(q_{*}{\mathcal{H}}(\varphi^{\mathrm{mmhm}}_{F\circ\pi_{f}}(\pi_{f*}\mathop{\mathcal{IC}}\nolimits_{M_{f}})))\cong{\mathcal{H}}^{n}(q_{*}\mathcal{H}(p_{*}\varphi^{\mathrm{mmhm}}_{\widetilde{F}}\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)}))

so we obtain the claim. ∎

4.4. Cohomological integrality theorem for LL-Higgs bundles

Here we prove the cohomological integrality theorem for YTotC(L)Y\coloneqq\operatorname{Tot}_{C}(L), where LL is a line bundle on a smooth projective curve CC with deg(L)>2g(C)2\deg(L)>2g(C)-2. Since the category of μ\mu-semistable one-dimensional sheaves on YY is homological dimension one, this can be proved in the same manner as [DM20, Theorem A] by applying the main result of [Mei]. However we give a sketch of the proof for reader’s convenience.

For a given rational number μ\mu\in\mathbb{Q}, we set

MYss(μ)mr=μMYss(r,m).M^{\operatorname{ss}}_{Y}(\mu)\coloneqq\coprod_{\frac{m}{r}=\mu}M^{\operatorname{ss}}_{Y}(r,m).

For each positive integer n>0n\in\mathbb{Z}_{>0}, we have the following morphism:

:(MYss(μ))×nMYss(μ),(Ei)i=1niEi,\oplus\colon(M^{\operatorname{ss}}_{Y}(\mu))^{\times n}\to M^{\operatorname{ss}}_{Y}(\mu),\quad(E_{i})_{i=1}^{n}\mapsto\oplus_{i}E_{i},

which is finite (cf. [DM, Examples 2.14 and 2.16]). We define functors

Symn,Sym:D,lf(MMHM(MYss(μ)))D,lf(MMHM(MYss(μ)))\operatorname{Sym}^{n}_{\boxtimes_{\oplus}},\quad\operatorname{Sym}_{\boxtimes_{\oplus}}\colon D^{\geq,lf}(\operatorname{MMHM}(M^{\operatorname{ss}}_{Y}(\mu)))\to D^{\geq,lf}(\operatorname{MMHM}(M^{\operatorname{ss}}_{Y}(\mu)))

as follows:

Symn()(n)𝔖n,Sym()n1Symn().\operatorname{Sym}^{n}_{\boxtimes_{\oplus}}(\mathcal{F})\coloneqq(\oplus_{*}\mathcal{F}^{\boxtimes n})^{{\mathfrak{S}}_{n}},\quad\operatorname{Sym}_{\boxtimes_{\oplus}}(\mathcal{F})\coloneqq\oplus_{n\geq 1}\operatorname{Sym}^{n}_{\boxtimes_{\oplus}}(\mathcal{F}).
Proposition 4.5.

The following statements hold:

  1. (1)

    The functor Sym\operatorname{Sym}_{\boxtimes_{\oplus}} is exact.

  2. (2)

    The functor Sym\operatorname{Sym}_{\boxtimes_{\oplus}} sends pure objects to pure objects.

  3. (3)

    Let Fμ:MYss(μ)𝔸1F_{\mu}\colon M^{\operatorname{ss}}_{Y}(\mu)\to\mathbb{A}^{1} be a regular function satisfying Fμ(AB)=Fμ(A)+Fμ(B)F_{\mu}(A\oplus B)=F_{\mu}(A)+F_{\mu}(B) for all A,BMYss(μ)A,B\in M^{\operatorname{ss}}_{Y}(\mu). Then the functors φFμmmhm\varphi^{\mathrm{mmhm}}_{F_{\mu}} and Sym\operatorname{Sym}_{\boxtimes_{\oplus}} commute.

Proof.

The same proofs as in [DM20, Propositions 3.5, 3.8, 3.11] work by using the finiteness of the morphism :(MYss(μ))×nMYss(μ)\oplus\colon(M^{\operatorname{ss}}_{Y}(\mu))^{\times n}\to M^{\operatorname{ss}}_{Y}(\mu) and Thom–Sebastiani isomorphism for the vanishing cycle functors φ()mmhm\varphi^{\mathrm{mmhm}}_{(-)}. ∎

We use the following notations:

𝒞𝔐Yss(μ)mr=μ𝒞𝔐Yss(r,m),𝒞MYss(μ)mr=μ𝒞MYss(r,m).\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(\mu)}\coloneqq\bigoplus_{\frac{m}{r}=\mu}\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)},\quad\mathop{\mathcal{IC}}\nolimits_{M^{\operatorname{ss}}_{Y}(\mu)}\coloneqq\bigoplus_{\frac{m}{r}=\mu}\mathop{\mathcal{IC}}\nolimits_{M^{\operatorname{ss}}_{Y}(r,m)}.

Recall that we denote by p:𝔐Yss(r,m)MYss(r,m)p\colon{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)\to M^{\mathrm{ss}}_{Y}(r,m) the canonical morphism to the good moduli space. Recall also that the definition of the object H(B)vir\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})_{\operatorname{vir}} from (4.2).

Theorem 4.6.

We have the following isomorphisms in D,lf(MMHM(MYss(μ)))D^{\geq,lf}(\operatorname{MMHM}(M^{\operatorname{ss}}_{Y}(\mu))):

(4.4) (p𝒞𝔐Yss(μ))Sym(H(B)vir𝒞MYss(μ)),\displaystyle\mathcal{H}(p_{*}\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(\mu)})\cong\operatorname{Sym}_{\boxtimes_{\oplus}}\left(\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})_{\operatorname{vir}}\otimes\mathop{\mathcal{IC}}\nolimits_{M^{\operatorname{ss}}_{Y}(\mu)}\right),
(4.5) (pφFμpmmhm𝒞𝔐Yss(μ))Sym(H(B)virφFμmmhm𝒞MYss(μ))\displaystyle\mathcal{H}(p_{*}\varphi^{\mathrm{mmhm}}_{F_{\mu}\circ p}\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(\mu)})\cong\operatorname{Sym}_{\boxtimes_{\oplus}}\left(\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})_{\operatorname{vir}}\otimes\varphi^{\mathrm{mmhm}}_{F_{\mu}}\mathop{\mathcal{IC}}\nolimits_{M^{\operatorname{ss}}_{Y}(\mu)}\right)

for a regular function Fμ:MYss(μ)𝔸1F_{\mu}\colon M^{\operatorname{ss}}_{Y}(\mu)\to\mathbb{A}^{1} satisfying Fμ(AB)=Fμ(A)+Fμ(B)F_{\mu}(A\oplus B)=F_{\mu}(A)+F_{\mu}(B) for all A,BMYss(μ)A,B\in M^{\operatorname{ss}}_{Y}(\mu).

Proof.

We first construct the isomorphism (4.4). By the exactness of the functor Sym\operatorname{Sym}_{\boxtimes_{\oplus}} (see Proposition 4.5 (1)), the right hand side is isomorphic to its total cohomology. Hence it is enough to prove the isomorphism for each cohomology. By Proposition 4.3 (1), for each nn\in\mathbb{Z} and (r,m)>0×(r,m)\in\mathbb{Z}_{>0}\times\mathbb{Z}, there exists f0f\gg 0 such that we have an isomorphism

(4.6) n(p𝒞𝔐Yss(r,m))n(πf𝒞Mf).\mathcal{H}^{n}(p_{*}\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)})\cong\mathcal{H}^{n}(\pi_{f*}\mathop{\mathcal{IC}}\nolimits_{M_{f}}).

Since the morphism πf\pi_{f} is proper and the object 𝒞Mf\mathop{\mathcal{IC}}\nolimits_{M_{f}} is pure, it follows that the object in (4.6) is a pure mixed Hodge module.

On the other hand, since 𝒞MYss(r,m)\mathop{\mathcal{IC}}\nolimits_{M^{\operatorname{ss}}_{Y}(r,m)} is pure, Proposition 4.5 (2) implies that the nn-th cohomology of the right hand side of (4.4) is also pure.

Now nn-th cohomology of both sides of (4.4) are direct sums of simple pure mixed Hodge modules. Hence it is enough to prove the equality in the Grothendieck group K0(D,lf(MYss(μ)))K_{0}(D^{\geq,lf}(M^{\operatorname{ss}}_{Y}(\mu))), which holds by the main theorem of [Mei].

The second isomorphism (4.5) follows by applying the vanishing cycle functor φFμmmhm\varphi^{\mathrm{mmhm}}_{F_{\mu}} to both sides of the isomorphism (4.4) and then using Proposition 4.3 (2) and Proposition 4.5 (3). ∎

We end this section by proving Proposition 3.10 in the previous section:

Proof of Proposition 3.10.

Fix integers r>0r>0 and mm\in\mathbb{Z}, and put μm/r\mu\coloneqq m/r. Let g:BY𝔸1g\colon B_{Y}\to\mathbb{A}^{1} be the function defined in (3.3). Recall from Proposition 3.6 that we have

𝔐Xss(r,m){d(h~Yg)=0}𝔐Yss(r,m){\mathfrak{M}}^{\operatorname{ss}}_{X}(r,m)\cong\{d(\widetilde{h}_{Y}\circ g)=0\}\subset{\mathfrak{M}}^{\operatorname{ss}}_{Y}(r,m)

for a line bundle L2L_{2} with deg(L2)0\deg(L_{2})\gg 0. Hence the first isomorphism in Proposition 3.10 follows from Proposition 2.4.

For the second isomorphism, let us put

FμghY:MYss(μ)𝔸1.F_{\mu}\coloneqq g\circ h_{Y}\colon M^{\operatorname{ss}}_{Y}(\mu)\to\mathbb{A}^{1}.

By taking the first cohomology of the isomorphism (4.5), we obtain

1(pφFμp𝒞𝔐Yss(μ))φFμ𝒞MYss(μ).\mathcal{H}^{1}(p_{*}\varphi_{F_{\mu}\circ p}\mathop{\mathcal{IC}}\nolimits_{{\mathfrak{M}}^{\operatorname{ss}}_{Y}(\mu)})\cong\varphi_{F_{\mu}}\mathop{\mathcal{IC}}\nolimits_{M^{\operatorname{ss}}_{Y}(\mu)}.

Restricting it to the component MYss(r,m)MYss(μ)M^{\operatorname{ss}}_{Y}(r,m)\subset M^{\operatorname{ss}}_{Y}(\mu), we get the second isomorphism in Proposition 3.10. ∎

Remark 4.7.

It is clear from the proof that Proposition 3.10 naturally extends to an isomorphism of monodromic mixed Hodge modules.

5. The case of Higgs bundles

In this section, we prove the cohomological integrality theorem and the cohomological χ\chi-independence theorem for Higgs bundle moduli spaces on curves using the dimensional reduction theorem due to the first author [Kin].

5.1. Dimensional reduction theorem

Let 𝖄\boldsymbol{{\mathfrak{Y}}} be a quasi-smooth derived Artin stack and 𝐓[1]𝖄\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}} be its (1)(-1)-shifted cotangent stack. We write 𝔜t0(𝖄){\mathfrak{Y}}\coloneqq t_{0}(\boldsymbol{{\mathfrak{Y}}}) and 𝔜~t0(𝐓[1]𝖄)\widetilde{{\mathfrak{Y}}}\coloneqq t_{0}(\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}), and π:𝔜~𝔜\pi\colon\widetilde{{\mathfrak{Y}}}\to{\mathfrak{Y}} be the natural projection. As we have seen in Example A.3, 𝐓[1]𝖄\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}} carries a natural (1)(-1)-shifted symplectic structure. Further, as is proved in [Toda, Lemma 3.3.3], there exists a natural orientation

(5.1) o:det𝕃𝐓[1]𝖄|𝔜red(πred)det(𝕃𝖄)2.o\colon\det{\mathbb{L}_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}}}|_{{\mathfrak{Y}}^{\operatorname{red}}}\cong({\pi^{\operatorname{red}}})^{*}\det(\mathbb{L}_{\boldsymbol{{\mathfrak{Y}}}})^{\otimes{2}}.

We let φ𝐓[1]𝖄\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}} denote the perverse sheaf on 𝔜~\widetilde{{\mathfrak{Y}}} recalled in §2.2 with respect to this (1)(-1)-shifted symplectic structure and orientation. The following theorem is called the dimensional reduction theorem.

Theorem 5.1 ([Kin, Theorem 4.14]).

There exists a natural isomorphism in Dcb(𝔜)D^{b}_{c}({\mathfrak{Y}})

(5.2) πφ𝐓[1]𝖄𝔻𝔜[vdim𝖄].\pi_{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}}\cong\mathbb{D}\mathbb{Q}_{{\mathfrak{Y}}}[-\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}].

Here vdim𝖄rank𝕃𝖄\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}\coloneqq\operatorname{rank}\mathbb{L}_{\boldsymbol{{\mathfrak{Y}}}} denotes the virtual dimension of 𝖄\boldsymbol{{\mathfrak{Y}}}.

We now discuss the generalization of this theorem to the level of complexes in mixed Hodge modules. Firstly we discuss the case when 𝖄\boldsymbol{{\mathfrak{Y}}} is a derived scheme. To specify that 𝖄\boldsymbol{{\mathfrak{Y}}} is schematic, we write 𝒀=𝖄\boldsymbol{Y}=\boldsymbol{{\mathfrak{Y}}}, Y=𝔜Y={\mathfrak{Y}} and Y~=𝔜~\widetilde{Y}=\widetilde{{\mathfrak{Y}}}. The following lemma is useful:

Lemma 5.2.

Let XX be an algebraic variety and take complexes of mixed Hodge module M,NDb(MHM(X))M,N\in D^{b}(\mathrm{MHM}(X)) such that there exists an isomorphism η:rat(M)rat(N)\eta\colon\operatorname{rat}(M)\cong\operatorname{rat}(N) in Db(X)D^{b}(X). Assume that for each i<0i<0 the group Exti(rat(M),rat(N))\operatorname{Ext}^{i}(\operatorname{rat}(M),\operatorname{rat}(N)) vanishes and we have an isomorphism of mixed Hodge structures H0(X,𝑜𝑚(M,N))\operatorname{H}^{0}(X,\mathop{\mathcal{H}\!\mathit{om}}\nolimits(M,N))\cong\mathbb{Q}. Then η\eta naturally extends to an isomorphism MNM\cong N in Db(MHM(X))D^{b}(\mathrm{MHM}(X)).

Proof.

Consider the natural map of mixed Hodge complexes

τ0RHom(M,N)RHom(M,N).\tau_{\leq 0}\operatorname{RHom}(M,N)\to\operatorname{RHom}(M,N).

The assumption implies an isomorphism τ0RHom(M,N)\mathbb{Q}\cong\tau_{\leq 0}\operatorname{RHom}(M,N) hence we obtain a map in Db(MMHM(X))D^{b}(\mathrm{MMHM}(X))

X𝑜𝑚(M,N)\mathbb{Q}_{X}\to\mathop{\mathcal{H}\!\mathit{om}}\nolimits(M,N)

by adjunction. Then the following composition of morphisms in Db(MHM(X))D^{b}(\mathrm{MHM}(X))

M=MXM𝑜𝑚(M,N)NM=M\otimes\mathbb{Q}_{X}\to M\otimes\mathop{\mathcal{H}\!\mathit{om}}\nolimits(M,N)\to N

upgrades the isomorphism η\eta up to scalar. ∎

Proposition 5.3.

Assume that the virtual dimension vdim𝐘\operatorname{vdim}\boldsymbol{Y} is even. Then the dimensional reduction isomorphism (5.2) for a quasi-smooth derived scheme 𝐘\boldsymbol{Y} naturally upgrades to an isomorphism in Db(MHM(Y))D^{b}(\operatorname{MHM}(Y)):

(5.3) πφ𝐓[1]𝒀mhm𝕃vdim𝒀/2𝔻Y.\pi_{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{Y}}^{\mathrm{mhm}}\cong\mathbb{L}^{\operatorname{vdim}\boldsymbol{Y}/2}\otimes\mathbb{D}\mathbb{Q}_{Y}.
Proof.

Using Lemma 5.2, we only need to prove that the mixed Hodge structure on

Hom(πφ𝐓[1]𝒀,𝔻Y[vdim𝒀])\operatorname{Hom}(\pi_{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{Y}},\mathbb{D}\mathbb{Q}_{Y}[-\operatorname{vdim}\boldsymbol{Y}])

is pure of weight zero. As this statement can be checked locally, using [BBJ19, Theorem 4.1], we may assume that there exists a smooth scheme UU which admits a global étale coordinate, a trivial vector bundle EE on UU, and a section sΓ(U,E)s\in\Gamma(U,E) such that 𝒀\boldsymbol{Y} is isomorphic to the derived zero locus 𝒁(s)\boldsymbol{Z}(s). In this case the proof of [Kin, Theorem 3.1] shows that the dimensional reduction isomorphism (5.2) can be identified with Davison’s local dimensional reduction theorem [Dav17, Theorem A.1]. As the proof of this theorem works verbatim for complexes of mixed Hodge modules, we conclude that the claim holds. ∎

Remark 5.4.

We expect that the isomorphism (5.3) further upgrades to an isomorphism Db(MMHM(Y))D^{b}(\operatorname{MMHM}(Y))

πφ𝐓[1]𝒀mmhm𝕃vdim𝒀/2𝔻Y.\pi_{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{Y}}^{\mathrm{mmhm}}\cong\mathbb{L}^{\operatorname{vdim}\boldsymbol{Y}/2}\otimes\mathbb{D}\mathbb{Q}_{Y}.

However, we could not prove this since we do not know whether the tensor-hom adjunction holds for monodromic mixed Hodge modules. Instead, we can easily see that we have an isomorphism in Db(MMHM(Y))D^{b}(\operatorname{MMHM}(Y))

(πφ𝐓[1]𝒀mmhm)(𝕃vdim𝒀/2𝔻Y){\mathcal{H}}(\pi_{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{Y}}^{\mathrm{mmhm}})\cong{\mathcal{H}}(\mathbb{L}^{\operatorname{vdim}\boldsymbol{Y}/2}\otimes\mathbb{D}\mathbb{Q}_{Y})

since the monodromy operator acts trivially on both sides (see [Davc, Remark 3.9]). It is enough for our purposes.

Now we discuss the stacky case of this proposition. Let 𝖄\boldsymbol{{\mathfrak{Y}}} be a quasi-smooth derived Artin stack such that its classical truncation 𝔜=t0(𝖄){\mathfrak{Y}}=t_{0}(\boldsymbol{{\mathfrak{Y}}}) is of the form [Y/G][Y/G] for some scheme YY and a linear algebraic group GG. In this case, we can upgrade the dimensional reduction theorem to an isomorphism of mixed Hodge structures.

Proposition 5.5.

Assume that vdim𝖄\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}} is even. Then the dimensional reduction isomorphism H(𝔜~,φ𝐓[1]𝖄)H+vdim𝖄BM(𝔜)\operatorname{H}^{*}(\widetilde{{\mathfrak{Y}}},\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}})\cong\operatorname{H}^{\mathrm{BM}}_{-*+\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}}({\mathfrak{Y}}) upgrades naturally to an isomorphism of mixed Hodge structures

H(𝔜~,φ𝐓[1]𝖄mhm)𝕃vdim𝖄/2HBM(𝔜)\operatorname{H}^{*}(\widetilde{{\mathfrak{Y}}},\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}}^{\mathrm{mhm}})\cong\mathbb{L}^{\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}/2}\otimes\operatorname{H}^{\mathrm{BM}}_{-*}({\mathfrak{Y}})
Proof.

For a fixed ii, take a smooth morphism q:T𝔜q\colon T\to{\mathfrak{Y}} of relative dimension dd such that

q!:Hi+vdim𝖄BM(𝔜)Hi+vdim𝖄+2dBM(T)q^{!}\colon\operatorname{H}^{\mathrm{BM}}_{-i+\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}}({\mathfrak{Y}})\to\operatorname{H}^{\mathrm{BM}}_{-i+\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}+2d}(T)

and the map

q~!:Hi(𝔜~,φ𝐓[1]𝖄)Hi(T~,q~φ𝐓[1]𝖄)\tilde{q}^{!}\colon\operatorname{H}^{i}(\widetilde{{\mathfrak{Y}}},\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}})\to\operatorname{H}^{i}(\widetilde{T},\tilde{q}^{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}})

are isomorphisms. Here q~:T~𝔜~\tilde{q}\colon\widetilde{T}\to\widetilde{{\mathfrak{Y}}} is the base change of qq. Therefore we need to show that the following composition of isomorphism of vector spaces

Hi(T~,q~φ𝐓[1]𝖄))\displaystyle\operatorname{H}^{i}(\widetilde{T},\tilde{q}^{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}})) Hi(𝔜~,φ𝐓[1]𝖄)\displaystyle\cong\operatorname{H}^{i}(\widetilde{{\mathfrak{Y}}},\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}})
Hi+vdim𝖄BM(𝔜)\displaystyle\cong\operatorname{H}^{\mathrm{BM}}_{-i+\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}}({\mathfrak{Y}})
Hi+vdim𝖄+2dBM(T)\displaystyle\cong\operatorname{H}^{\mathrm{BM}}_{-i+\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}+2d}(T)

upgrades to an isomorphism of mixed Hodge structures. To prove this, we will show that the following morphism in Db(T)D^{b}(T)

ηq:πTq~φ𝐓[1]𝖄\displaystyle\eta_{q}\colon{\pi_{T}}_{*}\tilde{q}^{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}} qπ𝔜φ𝐓[1]𝖄\displaystyle\cong q^{*}{\pi_{{\mathfrak{Y}}}}_{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}}
q𝔻𝔜[vdim𝖄]\displaystyle\cong q^{*}\mathbb{D}\mathbb{Q}_{{\mathfrak{Y}}}[-\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}]
𝔻T[vdim𝖄2d]\displaystyle\cong\mathbb{D}\mathbb{Q}_{T}[-\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}-2d]

upgrades an isomorphism in Db(MHM(T))D^{b}(\operatorname{MHM}(T)). Here πT:T~T\pi_{T}\colon\widetilde{T}\to T and π𝔜:𝔜~𝔜\pi_{{\mathfrak{Y}}}\colon\widetilde{{\mathfrak{Y}}}\to{\mathfrak{Y}} are natural projections and the second isomorphism is the dimensional reduction isomorphism. Using Lemma 5.2, we need to show that the mixed Hodge structure of

(5.4) Hom(πTq~φ𝐓[1]𝖄,𝔻T[vdim𝖄2d])\operatorname{Hom}({\pi_{T}}_{*}\tilde{q}^{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}},\mathbb{D}\mathbb{Q}_{T}[-\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}-2d])

is weight zero. To prove this, take a smooth surjective morphism from a derived scheme 𝒉:𝑼𝖄\boldsymbol{h}\colon\boldsymbol{U}\to\boldsymbol{{\mathfrak{Y}}} such that vdim𝑼\operatorname{vdim}\boldsymbol{U} is even. Write h=t0(𝒉)h=t_{0}(\boldsymbol{h}), t0(𝑼)=Ut_{0}(\boldsymbol{U})=U, and X~t0(𝐓[1]𝑼)\widetilde{X}\coloneqq t_{0}(\mathbf{T}^{*}[-1]\boldsymbol{U}). Let T×𝔜U~\widetilde{T\times_{{\mathfrak{Y}}}U} be the fibre product (T×𝔜U)×UU~(T\times_{{\mathfrak{Y}}}U)\times_{U}\widetilde{U}. We let πT×𝔜U:T×𝔜U~T×𝔜U\pi_{T\times_{{\mathfrak{Y}}}U}\colon\widetilde{T\times_{{\mathfrak{Y}}}U}\to T\times_{{\mathfrak{Y}}}U and πU:U~U\pi_{U}\colon\widetilde{U}\to U denote the natural projections and qU:T×𝔜UUq_{U}\colon T\times_{{\mathfrak{Y}}}U\to U and q~U:T×𝔜U~U~\tilde{q}_{U}\colon\widetilde{T\times_{{\mathfrak{Y}}}U}\to\widetilde{U} be the base changes of qq. Then we can construct a natural isomorphism

ηqU:πT×𝔜Uq~Uφ𝐓[1]𝑼𝔻T×𝔜U[vdim𝑼2d]\eta_{q_{U}}\colon{\pi_{T\times_{{\mathfrak{Y}}}U}}_{*}\tilde{q}_{U}^{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{U}}\cong\mathbb{D}\mathbb{Q}_{T\times_{{\mathfrak{Y}}}U}[-\operatorname{vdim}\boldsymbol{U}-2d]

in the same manner as ηq\eta_{q}. As we have seen in Proposition 5.3, the map ηqU\eta_{q_{U}} upgrades to an isomorphism in Db(MHM(T×𝔜U))D^{b}(\operatorname{MHM}(T\times_{{\mathfrak{Y}}}U))

πT×𝔜Uq~Uφ𝐓[1]𝑼mhm𝕃d+vdim𝑼/2𝔻T×𝔜U.{\pi_{T\times_{{\mathfrak{Y}}}U}}_{*}\tilde{q}_{U}^{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{U}}^{\mathrm{mhm}}\cong\mathbb{L}^{d+\operatorname{vdim}\boldsymbol{U}/2}\otimes\mathbb{D}\mathbb{Q}_{T\times_{{\mathfrak{Y}}}U}.

Let h𝔜~:U~𝔜~h_{\widetilde{{\mathfrak{Y}}}}\colon\widetilde{U}\to\widetilde{{\mathfrak{Y}}}, hT:T×𝔜UTh_{T}\colon T\times_{{\mathfrak{Y}}}U\to T and hT×𝔜U:T×𝔜U~T~h_{T\times_{{\mathfrak{Y}}}U}\colon\widetilde{T\times_{{\mathfrak{Y}}}U}\to\widetilde{T} be the base changes of hh. Then we have a natural isomorphism

hTπTq~φ𝐓[1]𝖄mhmπT×𝔜Uq~Uh𝔜~φ𝐓[1]𝖄mhm𝕃dimh/2πT×𝔜Uq~Uφ𝐓[1]𝑼mhmh_{T}^{*}{\pi_{T}}_{*}\tilde{q}^{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}}^{\mathrm{mhm}}\cong{\pi_{T\times_{{\mathfrak{Y}}}U}}_{*}\tilde{q}_{U}^{*}h_{\widetilde{{\mathfrak{Y}}}}^{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}}^{\mathrm{mhm}}\cong\mathbb{L}^{\dim h/2}\otimes{\pi_{T\times_{{\mathfrak{Y}}}U}}_{*}\tilde{q}_{U}^{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{U}}^{\mathrm{mhm}}

where the latter isomorphism follows from [Kin, Proposition 4.10]. We also have a natural isomorphism

𝕃d+vdim𝖄/2hT𝔻T𝕃d+h/2+vdim𝑼/2𝔻T×𝔜U.\mathbb{L}^{d+\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}/2}\otimes h_{T}^{*}\mathbb{D}\mathbb{Q}_{T}\cong\mathbb{L}^{d+h/2+\operatorname{vdim}\boldsymbol{U}/2}\otimes\mathbb{D}\mathbb{Q}_{T\times_{{\mathfrak{Y}}}U}.

Under these identifications, the proof of [Kin, Theorem 4.14] implies that hTηqh_{T}^{*}\eta_{q} is equal to ηqU[dimh]\eta_{q_{U}}[-\dim h] up to a certain choice of the sign. This and the fact that ηqU\eta_{q_{U}} upgrades to an isomorphism in Db(MHM(T×𝔜U))D^{b}(\operatorname{MHM}(T\times_{{\mathfrak{Y}}}U)) imply that the weight of the mixed Hodge structure (5.4) is zero. ∎

The following statement can be proved in the same manner as the previous proposition:

Proposition 5.6.

We keep the notation from the previous proposition. Let p:𝔜Zp\colon{\mathfrak{Y}}\to Z be a morphism to a separated finite type complex scheme. Then we have an isomorphism of mixed Hodge modules

(5.5) ((pπ𝔜)φ𝐓[1]𝖄mhm)𝕃vdim𝖄/2(p𝔻𝔜)\mathcal{H}((p\circ\pi_{{\mathfrak{Y}}})_{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}}^{\operatorname{mhm}})\cong\mathbb{L}^{\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}/2}\otimes\mathcal{H}(p_{*}\mathbb{D}\mathbb{Q}_{{\mathfrak{Y}}})

where π𝔜:𝔜~𝔜\pi_{{\mathfrak{Y}}}\colon\widetilde{{\mathfrak{Y}}}\to{\mathfrak{Y}} is the natural projection.

Remark 5.7.

The argument as in Remark 5.4 implies that the isomorphism (5.5) upgrades to an isomorphism in Db(MMHM(Z))D^{b}(\operatorname{MMHM}(Z))

((pπ𝔜)φ𝐓[1]𝖄mmhm)𝕃vdim𝖄/2(p𝔻𝔜).\mathcal{H}((p\circ\pi_{{\mathfrak{Y}}})_{*}\varphi_{\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{Y}}}}^{\operatorname{mmhm}})\cong\mathbb{L}^{\operatorname{vdim}\boldsymbol{{\mathfrak{Y}}}/2}\otimes\mathcal{H}(p_{*}\mathbb{D}\mathbb{Q}_{{\mathfrak{Y}}}).

5.2. BPS cohomology for Higgs bundles

In [Dav16], Davison defined BPS sheaves and BPS cohomology for preprojective algebras. In this section, we introduce Higgs counterpart of these notions.

Let CC be a smooth projective curve of genus gg. We write S=TotC(ωC)S=\operatorname{Tot}_{C}(\omega_{C}) and X=TotC(𝒪CωC)X=\operatorname{Tot}_{C}({\mathcal{O}}_{C}\oplus\omega_{C}). Recall that 𝔐Xss(r,m){\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m) (resp. 𝔐Sss(r,m){\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)) denotes the moduli stack of one-dimensional semistable sheaves of rank rr and Euler characteristic mm on XX (resp. SS), and MXss(r,m)M_{X}^{\operatorname{ss}}(r,m) (resp. MSss(r,m)M_{S}^{\operatorname{ss}}(r,m)) denotes the good moduli space of 𝔐Xss(r,m){\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m) (resp. 𝔐Sss(r,m){\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)). We have the following commutative diagram:

𝔐Xss(r,m)\textstyle{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pX\scriptstyle{p_{X}}π\scriptstyle{\pi}𝔐Sss(r,m)\textstyle{{\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pS\scriptstyle{p_{S}}MXss(r,m)\textstyle{M_{X}^{\operatorname{ss}}(r,m)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π¯\scriptstyle{\bar{\pi}}MSss(r,m).\textstyle{M_{S}^{\operatorname{ss}}(r,m).}

It is shown in [Kin, Theorem 5.1] that there exists a natural equivalence of (1)(-1)-shifted symplectic derived Artin stacks 𝕸Xss(r,m)𝐓[1]𝕸Sss(r,m)\boldsymbol{{\mathfrak{M}}}_{X}^{\operatorname{ss}}(r,m)\cong\mathbf{T}^{*}[-1]\boldsymbol{{\mathfrak{M}}}_{S}^{\operatorname{ss}}(r,m), where 𝕸Xss(r,m)\boldsymbol{{\mathfrak{M}}}_{X}^{\operatorname{ss}}(r,m) (resp. 𝕸Sss(r,m)\boldsymbol{{\mathfrak{M}}}_{S}^{\operatorname{ss}}(r,m)) denotes the derived enhancement of 𝔐Xss(r,m){\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m) (resp. 𝔐Sss(r,m){\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)). Therefore (5.1) implies that there exists a canonical orientation

o:(πred,det(𝕃𝕸Sss(r,m)|𝔐Xss(r,m)red))2K𝕸Xss(r,m)vir.o\colon(\pi^{\operatorname{red},*}\det(\mathbb{L}_{\boldsymbol{{\mathfrak{M}}}_{S}^{\operatorname{ss}}(r,m)}|_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}}))^{\otimes 2}\cong K_{\boldsymbol{{\mathfrak{M}}}_{X}^{\operatorname{ss}}(r,m)}^{\operatorname{vir}}.

On the other hand, we have seen in Proposition 3.6 that there exist a line bundle LL with degL>2g2\deg L>2g-2 and a function ff on the moduli stack 𝔐TotC(L)ss(r,m){\mathfrak{M}}_{\operatorname{Tot}_{C}(L)}^{\mathrm{ss}}(r,m) of semistable sheaves on TotC(L)\operatorname{Tot}_{C}(L) such that there exists an equivalence of (1)(-1)-shifted symplectic derived Artin stacks 𝕸Xss(r,m)𝐂𝐫𝐢𝐭(f)\boldsymbol{{\mathfrak{M}}}_{X}^{\operatorname{ss}}(r,m)\cong\mathop{\mathbf{Crit}}\nolimits(f). Therefore there exists an orientation

o:K𝔐Tot(L)(r,m)|𝔐Xss(r,m)red2K𝕸Xss(r,m)vir.o^{\prime}\colon K_{{\mathfrak{M}}_{\operatorname{Tot}(L)}(r,m)}|_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}}^{\otimes{2}}\cong K_{\boldsymbol{{\mathfrak{M}}}_{X}^{\operatorname{ss}}(r,m)}^{\operatorname{vir}}.
Proposition 5.8.

There exists an isomorphism of orientations ooo\cong o^{\prime}.

Proof.

We have seen in Proposition 3.8 that there exists a trivialization

K𝔐TotC(L)ss(r,m)𝒪𝔐TotC(L)ss(r,m).K_{{\mathfrak{M}}_{\operatorname{Tot}_{C}(L)}^{\mathrm{ss}}(r,m)}\cong{\mathcal{O}}_{{\mathfrak{M}}_{\operatorname{Tot}_{C}(L)}^{\mathrm{ss}}(r,m)}.

On the other hand, we also have a trivialization

det(𝕃𝕸Sss(r,m)|𝔐Sss(r,m)red))𝒪𝔐Sss(r,m)red,\det(\mathbb{L}_{\boldsymbol{{\mathfrak{M}}}_{S}^{\operatorname{ss}}(r,m)}|_{{\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)^{\operatorname{red}}}))\cong{\mathcal{O}}_{{\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)^{\operatorname{red}}},

since there exists an open immersion

𝕸Sss(r,m)𝐓𝔐C,\boldsymbol{{\mathfrak{M}}}_{S}^{\operatorname{ss}}(r,m)\hookrightarrow\mathbf{T}^{*}{\mathfrak{M}}_{C},

where 𝔐C{\mathfrak{M}}_{C} denotes the moduli stack of coherent sheaves on CC. Therefore we need to show that the following composition

𝒪𝔐Xss(r,m)red2\displaystyle{\mathcal{O}}_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}}^{\otimes{2}} (πred,det(𝕃𝕸Sss(r,m)|𝔐Xss(r,m)red))2\displaystyle\cong(\pi^{\operatorname{red},*}\det(\mathbb{L}_{\boldsymbol{{\mathfrak{M}}}_{S}^{\operatorname{ss}}(r,m)}|_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}}))^{\otimes 2}
K𝕸Xss(r,m)vir\displaystyle\cong K_{\boldsymbol{{\mathfrak{M}}}_{X}^{\operatorname{ss}}(r,m)}^{\operatorname{vir}}
K𝔐TotC(L)ss(r,m)|𝔐Xss(r,m)red2\displaystyle\cong K_{{\mathfrak{M}}_{\operatorname{Tot}_{C}(L)}^{\mathrm{ss}}(r,m)}|_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}}^{\otimes{2}}
𝒪𝔐Xss(r,m)red2\displaystyle\cong{\mathcal{O}}_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}}^{\otimes{2}}

has a square root. More strongly, we will show that any invertible function

fΓ(𝔐Xss(r,m)red,𝒪𝔐Xss(r,m)red×)Γ(MXss(r,m)red,𝒪MXss(r,m)red×)f\in\Gamma({\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}},{\mathcal{O}}_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}}^{\times})\cong\Gamma(M_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}},{\mathcal{O}}_{M_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}}^{\times})

is constant hence admits a square root.

We say that a reduced finite type complex scheme XX satisfies the property (P) if every invertible regular function on XX is locally constant. What we need to prove is that the scheme MXss(r,m)redM_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}} satisfies the property (P). Property (P) satisfies the following:

  • If we are given a surjective morphism between reduced finite type complex schemes XYX\to Y and XX satisfies the property (P), then YY satisfies the property (P).

  • For reduced finite type complex schemes XX and YY satisfying the property (P), X×YX\times Y also satisfies the property (P).

Write k=gcd(r,m)k=\gcd(r,m) and (r,m)=(kr0,km0)(r,m)=(kr_{0},km_{0}). Take a partition λ=(λ1,λ2,,λt)\lambda_{\bullet}=(\lambda_{1},\lambda_{2},\ldots,\lambda_{t}) of kk, i.e., λi\lambda_{i} is a positive integer with λ1λ2λt\lambda_{1}\geq\lambda_{2}\geq\cdots\geq\lambda_{t} such that iλi=k\sum_{i}\lambda_{i}=k holds. Let MXss(r,m)λredM_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}_{\lambda_{\bullet}} be the subscheme consisting of points corresponding to polystable sheaves which can be written as

iEi,\bigoplus_{i}E_{i},

where EiE_{i} is a stable sheaf on XX such that rank(πXEi)=λir0\operatorname{rank}({\pi_{X}}_{*}E_{i})=\lambda_{i}r_{0}, where πX:XC\pi_{X}\colon X\to C is the projection. We let MXss(r,m)λred¯\overline{M_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}_{\lambda_{\bullet}}} denote the closure of MXss(r,m)λredM_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}_{\lambda_{\bullet}}. Since we have an equality

MXss(r,m)red=λMXss(r,m)λred¯,M_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}=\bigcup_{\lambda_{\bullet}}\overline{M_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}_{\lambda_{\bullet}}},

we need to show that the scheme MXss(r,m)λred¯\overline{M_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}_{\lambda_{\bullet}}} satisfies the property (P).

Consider the map

i=1tMXss(λir0,λim0)(λi)red¯MXss(r,m)red\prod_{i=1}^{t}\overline{M_{X}^{\operatorname{ss}}(\lambda_{i}r_{0},\lambda_{i}m_{0})^{\operatorname{red}}_{(\lambda_{i})}}\to M_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}

taking the direct sum. The image of this map is nothing but MXss(r,m)λred¯\overline{M_{X}^{\operatorname{ss}}(r,m)^{\operatorname{red}}_{\lambda_{\bullet}}}. Therefore it is enough to show that MXss(λir0,λim0)(λi)red¯\overline{M_{X}^{\operatorname{ss}}(\lambda_{i}r_{0},\lambda_{i}m_{0})^{\operatorname{red}}_{(\lambda_{i})}} satisfies the property (P). As we have an isomorphism

MXss(λir0,λim0)(λi)red¯MSss(λir0,λim0)×𝔸1,\overline{M_{X}^{\operatorname{ss}}(\lambda_{i}r_{0},\lambda_{i}m_{0})^{\operatorname{red}}_{(\lambda_{i})}}\cong M_{S}^{\operatorname{ss}}(\lambda_{i}r_{0},\lambda_{i}m_{0})\times\mathbb{A}^{1},

we need to prove that MSss(λir0,λim0)M_{S}^{\operatorname{ss}}(\lambda_{i}r_{0},\lambda_{i}m_{0}) satisfies property (P).

Let g:MSss(λir0,λim0)𝔸1g\colon M_{S}^{\operatorname{ss}}(\lambda_{i}r_{0},\lambda_{i}m_{0})\to\mathbb{A}^{1} be an invertible function. We need to prove that gg is constant. Let hS:MSss(λir0,λim0)BSh_{S}\colon M_{S}^{\operatorname{ss}}(\lambda_{i}r_{0},\lambda_{i}m_{0})\to B_{S} be the Hitchin fibration. Since the general fiber of hSh_{S} is connected, we have an isomorphism hS,𝒪MSss(λir0,λim0)𝒪BSh_{S,*}{\mathcal{O}}_{M_{S}^{\operatorname{ss}}(\lambda_{i}r_{0},\lambda_{i}m_{0})}\cong{\mathcal{O}}_{B_{S}}. Therefore there exists an invertible function gg^{\prime} on BSB_{S} such that g=ghSg=g^{\prime}\circ h_{S} holds. Since BSB_{S} is an affine space, gg^{\prime} is a constant function. ∎

Remark 5.9.

The proof shows that any orientation o′′:L2K𝕸Xss(r,m)viro^{\prime\prime}\colon L^{\otimes 2}\cong K_{\boldsymbol{{\mathfrak{M}}}_{X}^{\operatorname{ss}}(r,m)}^{\operatorname{vir}} such that LL is trivial is isomorphic to oo.

From now we always equip 𝕸Xss(r,m)\boldsymbol{{\mathfrak{M}}}_{X}^{\operatorname{ss}}(r,m) with the orientation oo. Define a monodromic mixed Hodge module φMXss(r,m)mmhm\varphi_{M_{X}^{\operatorname{ss}}(r,m)}^{\mathrm{mmhm}} on MXss(r,m)M_{X}^{\operatorname{ss}}(r,m) by

φMXss(r,m)mmhm0(𝕃1/2pXφ𝔐Xss(r,m)mmhm).\varphi_{M_{X}^{\operatorname{ss}}(r,m)}^{\mathrm{mmhm}}\coloneqq{\mathcal{H}}^{0}(\mathbb{L}^{-1/2}\otimes{p_{X}}_{*}\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)}^{\mathrm{mmhm}}).

For a given rational number μ\mu, we define

MXss(μ)mr=μMXss(r,m),φ𝔐Xss(μ)mmhmmr=μφ𝔐Xss(r,m)mmhm,φMXss(μ)mmhm\displaystyle M^{\operatorname{ss}}_{X}(\mu)\coloneqq\coprod_{\frac{m}{r}=\mu}M^{\operatorname{ss}}_{X}(r,m),\ \varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(\mu)}^{\mathrm{mmhm}}\coloneqq\bigoplus_{\frac{m}{r}=\mu}\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)}^{\mathrm{mmhm}},\ \varphi_{M_{X}^{\operatorname{ss}}(\mu)}^{\mathrm{mmhm}} mr=μφMXss(r,m)mmhm.\displaystyle\coloneqq\bigoplus_{\frac{m}{r}=\mu}\varphi_{M_{X}^{\operatorname{ss}}(r,m)}^{\mathrm{mmhm}}.

Recall that we have constructed a symmetric monoidal structure \boxtimes_{\oplus} on D,lf(MMHM(MXss(μ)))D^{\geq,lf}(\operatorname{MMHM}(M^{\mathrm{ss}}_{X}(\mu))) in §4.2. The following proposition is the cohomological integrality theorem (in the sense of [DM20, Theorem A]) for the Calabi–Yau threefold XX:

Proposition 5.10.

We have an isomorphism

(pXφ𝔐Xss(μ)mmhm)Sym(H(B)virφMXss(μ)mmhm){\mathcal{H}}({p_{X}}_{*}\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(\mu)}^{\mathrm{mmhm}})\cong\operatorname{Sym}_{\boxtimes_{\oplus}}\left(\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})_{\operatorname{vir}}\otimes\varphi_{M_{X}^{\operatorname{ss}}(\mu)}^{\mathrm{mmhm}}\right)

in D,lf(MMHM(MXss(μ)))D^{\geq,lf}(\operatorname{MMHM}(M^{\mathrm{ss}}_{X}(\mu))).

Proof.

Using Proposition 5.8, we may use the orientation oo^{\prime} instead of oo. Then the claim follows from Proposition 3.10 and Theorem 4.6. ∎

Now we state the Higgs version of the support lemma [Dav16, Lemma 4.1]:

Proposition 5.11.

Let :MSss(r,m)×𝔸1MXss(r,m)\ell\colon M_{S}^{\operatorname{ss}}(r,m)\times\mathbb{A}^{1}\to M_{X}^{\operatorname{ss}}(r,m) be the map given by

MSss(r,m)×𝔸1([E],t)[itE]MXss(r,m)M_{S}^{\operatorname{ss}}(r,m)\times\mathbb{A}^{1}\ni([E],t)\mapsto[{i_{t}}_{*}E]\in M_{X}^{\operatorname{ss}}(r,m)

where iti_{t} is the composition S=S×{t}XS=S\times\{t\}\hookrightarrow X. Then the support of the perverse sheaf φMXss(r,m)\varphi_{M_{X}^{\operatorname{ss}}(r,m)} is contained in the image of \ell.

The proof will be given in Appendix B.

Proposition 5.12.

The monodromic mixed Hodge module φMXss(r,m)mmhm\varphi_{M_{X}^{\operatorname{ss}}(r,m)}^{\operatorname{mmhm}} is 𝔸1\mathbb{A}^{1}-equivariant with respect to the natural 𝔸1\mathbb{A}^{1}-action on MXss(r,m)M_{X}^{\operatorname{ss}}(r,m). Further there exists a monodromic mixed Hodge module 𝒫𝒮r,mMMHM(MSss(r,m))\mathop{\mathcal{BPS}}\nolimits_{r,m}\in\operatorname{MMHM}(M_{S}^{\operatorname{ss}}(r,m)) such that

φMXss(r,m)mmhm𝕃1/2lpr1𝒫𝒮r,m\varphi_{M_{X}^{\operatorname{ss}}(r,m)}^{\operatorname{mmhm}}\cong\mathbb{L}^{-1/2}\otimes l_{*}\operatorname{pr}_{1}^{*}\mathop{\mathcal{BPS}}\nolimits_{r,m}

holds where pr1:MSss(r,m)×𝔸1MSss(r,m)\operatorname{pr}_{1}\colon M_{S}^{\operatorname{ss}}(r,m)\times\mathbb{A}^{1}\to M_{S}^{\operatorname{ss}}(r,m) is the projection.

Proof.

In general, a monodromic mixed Hodge module MM on T×𝔸1T\times\mathbb{A}^{1} for an algebraic variety TT is 𝔸1\mathbb{A}^{1}-equivariant if and only if the counit map pr1pr1MM\operatorname{pr}_{1}^{*}{\operatorname{pr}_{1}}_{*}M\to M is isomorphic, where pr1:T×𝔸1T\operatorname{pr}_{1}\colon T\times\mathbb{A}^{1}\to T is the first order projection. Therefore the 𝔸1\mathbb{A}^{1}-equivariance of MM is equivalent to the 𝔸1\mathbb{A}^{1}-equivariance of rat(M)\operatorname{rat}(M). This is further equivalent to the condition σ1rat(M)rat(M)\sigma_{1}^{*}\operatorname{rat}(M)\cong\operatorname{rat}(M), where σ1:T×𝔸1T×𝔸1\sigma_{1}\colon T\times\mathbb{A}^{1}\cong T\times\mathbb{A}^{1} is the map translating in the 𝔸1\mathbb{A}^{1}-direction by 1𝔸11\in\mathbb{A}^{1}.

Now we return to the proposition. Let σ1:𝔐Xss(r,m)𝔐Xss(r,m)\sigma_{1}\colon{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)\cong{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m) be the map induced by the translation map on X=S×𝔸1X=S\times\mathbb{A}^{1} in the 𝔸1\mathbb{A}^{1}-direction by 1𝔸11\in\mathbb{A}^{1}. We need to show that there exists an isomorphism of perverse sheaves

σ1φ𝔐Xss(r,m)φ𝔐Xss(r,m).\sigma_{1}^{*}\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)}\cong\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)}.

To do this, it is enough to show that there exists an isomorphism of orientations σ1oo\sigma_{1}^{*}o\cong o where oo is the natural orientation on 𝔐Xss(r,m){\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m). But this is a consequence of Remark 5.9. The latter statement follows from Proposition 5.11. ∎

The object 𝒫𝒮r,mMMHM(MSss(r,m))\mathop{\mathcal{BPS}}\nolimits_{r,m}\in\operatorname{MMHM}(M_{S}^{\operatorname{ss}}(r,m)) is called the BPS sheaf. We will see that it is a pure Hodge module in the next section. We write

BPSr,mH(MSss(r,m),𝒫𝒮r,m)\operatorname{BPS}_{r,m}\coloneqq\operatorname{H}^{*}(M_{S}^{\operatorname{ss}}(r,m),\mathop{\mathcal{BPS}}\nolimits_{r,m})

and it is called the BPS cohomology.

5.3. Cohomological integrality and χ\chi-independence for Higgs bundles

In this section, we prove the χ\chi-independence theorem and cohomological integrality theorem for Higgs bundles using the dimensional reduction theorem.

We first need the following lemma:

Lemma 5.13.

The map b|im(hX):im(hX)BYb|_{\mathrm{im}(h_{X})}\colon\mathrm{im}(h_{X})\to B_{Y} considered in Lemma 3.9 is injective for X=TotC(𝒪CωC)X=\operatorname{Tot}_{C}({\mathcal{O}}_{C}\oplus\omega_{C}).

Proof.

Let γ,γim(hX)\gamma,\gamma^{\prime}\in\operatorname{im}(h_{X}) be cycles on XX such that the pushforward cycles σγ\sigma_{*}\gamma and σγ\sigma_{*}\gamma^{\prime} define the same cycle on Y=TotC(L)Y=\operatorname{Tot}_{C}(L) where σ\sigma is the projection from XX to YY. We want to show γ=γ\gamma=\gamma^{\prime}. Write

γ=iγti,γ=iγti\gamma=\sum_{i}\gamma_{t_{i}},\ \gamma^{\prime}=\sum_{i}\gamma_{t_{i}^{\prime}}^{\prime}

where γti\gamma_{t_{i}} is supported on S×{ti}XS\times\{t_{i}\}\subset X and similarly for γti\gamma^{\prime}_{t_{i}^{\prime}}. Take a point pSupp(coker(ωCL))p\in\operatorname{Supp}(\mathrm{coker}(\omega_{C}\hookrightarrow L)). Then the restriction of σ\sigma at the fiber of pp is given by

𝔸1ωC|p𝔸1Lp\mathbb{A}^{1}\oplus\omega_{C}|_{p}\to\mathbb{A}^{1}\cong L_{p}

where the first map is the projection to the first factor and the latter map is induced from the composition 𝒪C𝒪CωCL{\mathcal{O}}_{C}\hookrightarrow{\mathcal{O}}_{C}\oplus\omega_{C}\to L. Therefore the cycle σγti|Lp\sigma_{*}\gamma_{t_{i}}|_{L_{p}} is concentrated in {ti}𝔸1Lp\{t_{i}\}\subset\mathbb{A}^{1}\cong L_{p}. Therefore we may assume that γ\gamma and γ\gamma^{\prime} are contained in S×{t}S\times\{t\} for some t𝔸1t\in\mathbb{A}^{1}. Then the claim follows since the map S×{t}XYS\times\{t\}\subset X\to Y defines an injection on the set of cycles. ∎

The following corollary is an immediate consequence of the isomorphism (3.12) and the above lemma.

Corollary 5.14.

Let us take integers r,m,mr,m,m^{\prime} such that r>0r>0. Then there exists an isomorphism in Db(MMHM(BX))D^{b}(\operatorname{MMHM}(B_{X})):

hXφMXss(r,m)mmhmhXφMXss(r,m)mmhm.h_{X*}\varphi_{M_{X}^{\operatorname{ss}}(r,m)}^{\operatorname{mmhm}}\cong h_{X*}\varphi_{M_{X}^{\operatorname{ss}}(r,m^{\prime})}^{\operatorname{mmhm}}.
Corollary 5.15.

Let r,m,mr,m,m^{\prime} be as in the previous corollary. Then there exists an isomorphism in Db(MMHM(BS))D^{b}(\operatorname{MMHM}(B_{S})):

hS𝒫𝒮r,mhS𝒫𝒮r,m.h_{S*}\mathop{\mathcal{BPS}}\nolimits_{r,m}\cong h_{S*}\mathop{\mathcal{BPS}}\nolimits_{r,m^{\prime}}.

We now prove the cohomological integrality theorem for Higgs bundles. Recall that we have the following diagram:

𝔐Xss(μ)\textstyle{{\mathfrak{M}}_{X}^{\operatorname{ss}}(\mu)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pX\scriptstyle{p_{X}}π\scriptstyle{\pi}𝔐Sss(μ)\textstyle{{\mathfrak{M}}_{S}^{\operatorname{ss}}(\mu)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pS\scriptstyle{p_{S}}MXss(μ)\textstyle{M_{X}^{\operatorname{ss}}(\mu)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π¯\scriptstyle{\bar{\pi}}MSss(μ).\textstyle{M_{S}^{\operatorname{ss}}(\mu).}

For a rational number μ\mu, we write

𝒫𝒮μmr=μ𝒫𝒮r,m.\mathop{\mathcal{BPS}}\nolimits_{\mu}\coloneqq\bigoplus_{\frac{m}{r}=\mu}\mathop{\mathcal{BPS}}\nolimits_{r,m}.
Theorem 5.16.

The monodromic mixed Hodge module 𝒫𝒮μ\mathop{\mathcal{BPS}}\nolimits_{\mu} is contained in MHM(MSss(μ))\operatorname{MHM}(M^{\mathrm{ss}}_{S}(\mu)), i.e., it has a trivial monodromy operator. Further, we have an isomorphism

(5.6) mr=μ(pS𝔻𝔐Sss(r,m))𝕃r2(g1)Sym(H(B)𝒫𝒮μ)\bigoplus_{\frac{m}{r}=\mu}{\mathcal{H}}({p_{S}}_{*}\mathbb{D}\mathbb{Q}_{{\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)})\otimes\mathbb{L}^{r^{2}(g-1)}\cong\operatorname{Sym}_{\boxtimes_{\oplus}}\left(\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})\otimes\mathop{\mathcal{BPS}}\nolimits_{\mu}\right)

in D,lf(MHM(MSss(μ)))D^{\geq,lf}(\operatorname{MHM}(M^{\mathrm{ss}}_{S}(\mu))).

Proof.

Proposition 4.4 and Proposition 5.10 imply isomorphisms

((pSπ)φ𝔐Xss(μ)mmhm)\displaystyle{\mathcal{H}}((p_{S}\circ\pi)_{*}\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(\mu)}^{\operatorname{mmhm}}) ((π¯pX)φ𝔐Xss(μ)mmhm)\displaystyle\cong{\mathcal{H}}((\bar{\pi}\circ p_{X})_{*}\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(\mu)}^{\operatorname{mmhm}})
(π¯(pXφ𝔐Xss(μ)mmhm))\displaystyle\cong{\mathcal{H}}(\bar{\pi}_{*}{\mathcal{H}}({p_{X}}_{*}\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(\mu)}^{\operatorname{mmhm}}))
(π¯Sym(H(B)virφMXss(μ)mmhm))\displaystyle\cong{\mathcal{H}}\left(\bar{\pi}_{*}\operatorname{Sym}_{\boxtimes_{\oplus}}\left(\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})_{\operatorname{vir}}\otimes\varphi_{M_{X}^{\operatorname{ss}}(\mu)}^{\operatorname{mmhm}}\right)\right)
Sym(H(B)𝒫𝒮μ).\displaystyle\cong\operatorname{Sym}_{\boxtimes_{\oplus}}\left(\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})\otimes\mathop{\mathcal{BPS}}\nolimits_{\mu}\right).

As we have seen in Remark 5.7, the left-hand side is monodromy-free, hence so is the BPS sheaf. The isomorphism (5.6) follows from the above isomorphism and an isomorphism

((pSπ)φ𝔐Xss(μ)mmhm)mr=μ(pS𝔻𝔐Sss(r,m))𝕃r2(g1){\mathcal{H}}((p_{S}\circ\pi)_{*}\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(\mu)}^{\operatorname{mmhm}})\cong\bigoplus_{\frac{m}{r}=\mu}{\mathcal{H}}({p_{S}}_{*}\mathbb{D}\mathbb{Q}_{{\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)})\otimes\mathbb{L}^{r^{2}(g-1)}

which is a consequence of Proposition 5.6 and the equality vdim𝕸Sss(r,m)=2r2(g1)\operatorname{vdim}\boldsymbol{{\mathfrak{M}}}_{S}^{\operatorname{ss}}(r,m)=2r^{2}(g-1). ∎

Corollary 5.17.

The mixed Hodge module 𝒫𝒮μ\mathop{\mathcal{BPS}}\nolimits_{\mu} is pure.

Proof.

The above theorem implies that there exists an embedding

𝒫𝒮r,m(pS𝔻𝔐Sss(r,m))𝕃r2(g1).\mathop{\mathcal{BPS}}\nolimits_{r,m}\hookrightarrow{\mathcal{H}}({p_{S}}_{*}\mathbb{D}\mathbb{Q}_{{\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)})\otimes\mathbb{L}^{r^{2}(g-1)}.

The purity of the right-hand side is proved in [Dava, Proposition 7.20], so we obtain the claim. ∎

Example 5.18.

Assume that (r,m)(r,m) is coprime, in which case 𝔐Sss(r,m){\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m) is smooth and pSp_{S} is a \mathbb{C}^{*}-gerbe. In this case, we have an isomorphism

0(pS𝔻𝔐Sss(r,m))𝕃r2(g1)𝒞MSss(r,m).{\mathcal{H}}^{0}({p_{S}}_{*}\mathbb{D}\mathbb{Q}_{{\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)})\otimes\mathbb{L}^{r^{2}(g-1)}\cong\mathop{\mathcal{IC}}\nolimits_{M_{S}^{\operatorname{ss}}(r,m)}.

Therefore we have an isomorphism

𝒞MSss(r,m)𝒫𝒮r,m.\mathop{\mathcal{IC}}\nolimits_{M_{S}^{\operatorname{ss}}(r,m)}\cong\mathop{\mathcal{BPS}}\nolimits_{r,m}.

In particular, for coprime pairs (r,m)(r,m) and (r,m)(r,m^{\prime}), Corollary 5.15 implies an isomorphism

hS𝒞MSss(r,m)hS𝒞MSss(r,m).h_{S*}\mathop{\mathcal{IC}}\nolimits_{M_{S}^{\operatorname{ss}}(r,m)}\cong h_{S*}\mathop{\mathcal{IC}}\nolimits_{M_{S}^{\operatorname{ss}}(r,m^{\prime})}.

Now let (r,m)(r,m) be a non-coprime pair. It follows from [Sim94, Theorem 11.1] and [Dava, Theorem 5.11] that MSss(r,m)M_{S}^{\operatorname{ss}}(r,m) is normal. The connectedness of MSss(r,m)M_{S}^{\operatorname{ss}}(r,m) is proved in [DP12, Claim 3.5 (iii)]. Therefore the moduli space MSss(r,m)M_{S}^{\operatorname{ss}}(r,m) is irreducible. Then using [Dava, Theorem 6.6], we can construct an inclusion

(5.7) 𝒞MSss(r,m)𝒫𝒮r,m\mathop{\mathcal{IC}}\nolimits_{M_{S}^{\operatorname{ss}}(r,m)}\hookrightarrow\mathop{\mathcal{BPS}}\nolimits_{r,m}

but it is not necessary an isomorphism (see §5.4).

Write BSB_{S}^{*} for the disjoint union of all the Hitchin bases (i.e. BSB_{S}^{*} is the moduli space of all one-dimensional cycles on SS). We let B:BS×BSBS\oplus_{B}\colon B_{S}^{*}\times B_{S}^{*}\to B_{S}^{*} and +:×+\colon\mathbb{N}\times\mathbb{N}\to\mathbb{N} denote the canonical monoid structures. The following statement is a direct consequence of Theorem 5.16 and Lemma 4.2.

Corollary 5.19.

We have isomorphisms

mr=μ((hSpS)𝔻𝔐Sss(r,m))𝕃r2(g1)\displaystyle\bigoplus_{\frac{m}{r}=\mu}{\mathcal{H}}((h_{S}\circ p_{S})_{*}\mathbb{D}\mathbb{Q}_{{\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)})\otimes\mathbb{L}^{r^{2}(g-1)} SymB(H(B)hS𝒫𝒮μ)\displaystyle\cong\operatorname{Sym}_{\boxtimes_{\oplus_{B}}}\left(\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})\otimes{h_{S}}_{*}\mathop{\mathcal{BPS}}\nolimits_{\mu}\right)
mr=μHBM(𝔐Sss(r,m))𝕃r2(g1)\displaystyle\bigoplus_{\frac{m}{r}=\mu}\operatorname{H}^{\mathrm{BM}}({{\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)})\otimes\mathbb{L}^{r^{2}(g-1)} Sym+(H(B)BPSμ)\displaystyle\cong\operatorname{Sym}_{\boxtimes_{+}}\left(\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})\otimes\operatorname{BPS}_{\mu}\right)

in D,lf(MHM(BS))D^{\geq,lf}(\operatorname{MHM}(B_{S}^{*})) and D,lf(MHM())D^{\geq,lf}(\operatorname{MHM}(\mathbb{N})) respectively.

Combining the above corollary and the χ\chi-independence theorem for BPS cohomology (= Corollary 5.14), we obtain the following χ\chi-independence theorem for the Borel–Moore homology:

Corollary 5.20.

Let r,m,mr,m,m^{\prime} be integers such that r>0r>0 and gcd(r,m)=gcd(r,m)\gcd(r,m)=\gcd(r,m^{\prime}). Then there exist isomorphisms

((hSpS)𝔻𝔐Sss(r,m))\displaystyle{\mathcal{H}}((h_{S}\circ p_{S})_{*}\mathbb{D}\mathbb{Q}_{{\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)}) ((hSpS)𝔻𝔐Sss(r,m)),\displaystyle\cong{\mathcal{H}}((h_{S}\circ p_{S})_{*}\mathbb{D}\mathbb{Q}_{{\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m^{\prime})}),
HBM(𝔐Sss(r,m))\displaystyle\operatorname{H}^{\mathrm{BM}}({{\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m)}) HBM(𝔐Sss(r,m))\displaystyle\cong\operatorname{H}^{\mathrm{BM}}({{\mathfrak{M}}_{S}^{\operatorname{ss}}(r,m^{\prime})})

in D,lf(MHM(BS))D^{\geq,lf}(\operatorname{MHM}(B_{S})) and D,lf(MHM(pt))D^{\geq,lf}(\operatorname{MHM}(\operatorname{pt})) respectively.

Remark 5.21.

Based on P = W conjecture, it is conjectured in [FSY] that there exists an isomorphism of intersection cohomology groups

IH(MSss(r,m))IH(MSss(r,m))\operatorname{IH}({M_{S}^{\operatorname{ss}}(r,m)})\cong\operatorname{IH}({M_{S}^{\operatorname{ss}}(r,m^{\prime})})

preserving the perverse filtration for r,m,mr,m,m^{\prime} such that gcd(r,m)=gcd(r,m)\gcd(r,m)=\gcd(r,m^{\prime}). At present we do not know how to prove this conjecture. However, once Davison’s conjecture [Davb, Conjecture 7.7] on the structure of the BPS sheaf is established, it would be possible to deduce the χ\chi-independence for intersection cohomology from the χ\chi-independence for BPS cohomology (= Corollary 5.14).

5.4. An example: g=2,r=2g=2,r=2

Here we give an example where the intersection cohomology and the BPS cohomology are different. Let CC be a smooth projective curve of genus 22, and put STotC(ωC)S\coloneqq\operatorname{Tot}_{C}(\omega_{C}). We consider the moduli space MS(2,0)M_{S}(2,0). By taking the cohomology of the inclusion (5.7), we have an inclusion

(5.8) IH(MS(2,0))BPS2,0.\operatorname{IH}(M_{S}(2,0))\hookrightarrow\operatorname{BPS}_{2,0}.

We will check that the above inclusion is not an isomorphism. Note that by Corollary 5.15 and Example 5.18, we have an isomorphism

(5.9) BPS2,0BPS2,1IH(MS(2,1)).\operatorname{BPS}_{2,0}\cong\operatorname{BPS}_{2,1}\cong\operatorname{IH}(M_{S}(2,1)).

We denote by

ΦIC(t)ihi(IH(MS(2,0)))ti,ΦBPS(t)ihi(BPS2,0)ti.\Phi_{\operatorname{IC}}(t)\coloneqq\sum_{i\in\mathbb{Z}}h^{i}(\operatorname{IH}(M_{S}(2,0)))t^{i},\quad\Phi_{\operatorname{BPS}}(t)\coloneqq\sum_{i\in\mathbb{Z}}h^{i}(\operatorname{BPS}_{2,0})t^{i}.

By [Fel21, Theorem 1.2] and [Ray18, Exercise 4.1], we have

(5.10) ΦIC(t)=t10(t+1)4(2t6+2t4+t2+1),\displaystyle\Phi_{\operatorname{IC}}(t)=t^{-10}(t+1)^{4}(2t^{6}+2t^{4}+t^{2}+1),
ΦBPS(t)=t10(t+1)4(2t6+4t5+2t4+4t3+t2+1).\displaystyle\Phi_{\operatorname{BPS}}(t)=t^{-10}(t+1)^{4}(2t^{6}+4t^{5}+2t^{4}+4t^{3}+t^{2}+1).

For the formula of ΦBPS\Phi_{\operatorname{BPS}}, we used the isomorphisms (5.9). Note that the term t10t^{-10} appears by our shift convention so that the intersection and the BPS complexes are perverse sheaves, together with the fact dimMS(2,0)=10\dim M_{S}(2,0)=10. From the formulas (5.10), it is obvious that the inclusion (5.8) is not an isomorphism.

On the other hand, assume that [Davb, Conjecture 7.7] is true. Then we have an isomorphism

𝒫𝒮2,0𝒞MS(2,0)2(𝒞MS(1,0)),\mathop{\mathcal{BPS}}\nolimits_{2,0}\cong\mathop{\mathcal{IC}}\nolimits_{M_{S}(2,0)}\oplus\wedge_{\oplus}^{2}(\mathop{\mathcal{IC}}\nolimits_{M_{S}(1,0)}),

where 2()\wedge^{2}(-) is the graded wedge product in the category of graded vector spaces. Since the moduli space MS(1,0)M_{S}(1,0) is isomorphic to the cotangent bundle of the Jacobian Jac(C)\operatorname{Jac}(C) of CC, we have

IH(MS(1,0))H(Jac(C))[4]=[4]4[3]6[2]4[1],\operatorname{IH}(M_{S}(1,0))\cong\operatorname{H}^{*}(\operatorname{Jac}(C))[4]=\mathbb{Q}[4]\oplus\mathbb{Q}^{4}[3]\oplus\mathbb{Q}^{6}[2]\oplus\mathbb{Q}^{4}[1]\oplus\mathbb{Q},

and hence we conclude that

ihi(2(IH(MS(1,0))))ti\displaystyle\sum_{i\in\mathbb{Z}}h^{i}(\wedge^{2}(\operatorname{IH}(M_{S}(1,0))))t^{i} =t7(t+1)4(t2+1)\displaystyle=t^{-7}(t+1)^{4}(t^{2}+1)
=ΦBPS(t)ΦIC(t).\displaystyle=\Phi_{\operatorname{BPS}}(t)-\Phi_{\operatorname{IC}}(t).

This computation gives an evidence of [Davb, Conjecture 7.7]. At the same time, we can see that the χ\chi-independence for the intersection cohomology does not necessarily hold when gcd(r,m)gcd(r,m)\gcd(r,m)\neq\gcd(r,m^{\prime}).

Appendix A Shifted symplectic structure and vanishing cycles

In this appendix, we briefly recall the theory of shifted symplectic geometry and prove some technical lemmas including Proposition 2.4.

A.1. Shifted symplectic structures

We recall the notion of shifted symplectic structures introduced in [PTVV13]. Let 𝖃\boldsymbol{{\mathfrak{X}}} be a derived Artin stack. We define the space of nn-shifted pp-forms 𝒜p(𝖃,n)𝕊{\mathcal{A}}^{p}(\boldsymbol{{\mathfrak{X}}},n)\in\mathbb{S} by

𝒜p(𝖃,n)Map(𝒪𝖃,p𝕃𝖃[n]).{\mathcal{A}}^{p}(\boldsymbol{{\mathfrak{X}}},n)\coloneqq\operatorname{Map}({\mathcal{O}}_{\boldsymbol{{\mathfrak{X}}}},\wedge^{p}\mathbb{L}_{\boldsymbol{{\mathfrak{X}}}}[n]).

We can also define the space of nn-shifted closed pp-forms 𝒜p,cl(𝖃,n)𝕊{\mathcal{A}}^{p,\operatorname{cl}}(\boldsymbol{{\mathfrak{X}}},n)\in\mathbb{S} (see [PTVV13, Definition 1.12]). It satisfies the étale descent and for a connective commutative differential graded algebra AA we have an equivalence

𝒜p,cl(𝐒𝐩𝐞𝐜A,n)|i0p+i𝕃A[i+n],d+ddR|,{\mathcal{A}}^{p,\operatorname{cl}}(\mathop{\mathbf{Spec}}\nolimits A,n)\simeq\left|\prod_{i\geq 0}\wedge^{p+i}\mathbb{L}_{A}[-i+n],d+d_{\mathrm{dR}}\right|,

where dd is the internal differential, ddRd_{\mathrm{dR}} is the de Rham differential, and |||-| is the geometric realization functor. The space of nn-shifted pp-forms and the space of nn-shifted closed pp-forms is functorial with respect to morphisms between derived Artin stacks, i.e., if we are given a morphism 𝒇:𝖃𝖄\boldsymbol{f}\colon\boldsymbol{{\mathfrak{X}}}\to\boldsymbol{{\mathfrak{Y}}}, there exist natural maps

𝒇\displaystyle\boldsymbol{f}^{\star} :𝒜p(𝖄,n)𝒜p(𝖃,n),\displaystyle\colon{\mathcal{A}}^{p}(\boldsymbol{{\mathfrak{Y}}},n)\to{\mathcal{A}}^{p}(\boldsymbol{{\mathfrak{X}}},n),
𝒇\displaystyle\boldsymbol{f}^{\star} :𝒜p,cl(𝖄,n)𝒜p,cl(𝖃,n).\displaystyle\colon{\mathcal{A}}^{p,\operatorname{cl}}(\boldsymbol{{\mathfrak{Y}}},n)\to{\mathcal{A}}^{p,\operatorname{cl}}(\boldsymbol{{\mathfrak{X}}},n).

We have a natural forgetful map

π:𝒜p,cl(,n)𝒜p(,n)\pi\colon{\mathcal{A}}^{p,\operatorname{cl}}(-,n)\to{\mathcal{A}}^{p}(-,n)

and the de Rham differential map

ddRcl:𝒜p(,n)𝒜p+1,cl(,n).d_{\mathrm{dR}}^{\,\operatorname{cl}}\colon{\mathcal{A}}^{p}(-,n)\to{\mathcal{A}}^{p+1,\operatorname{cl}}(-,n).
Definition A.1.

An nn-shifted closed 2-form ω𝖃\omega_{\boldsymbol{{\mathfrak{X}}}} is called an nn-shifted symplectic form if its underlying nn-shifted 22-form is non-degenerate, i.e., the natural map

π(ω𝖃):𝕃𝖃𝕃𝖃[n]\pi(\omega_{\boldsymbol{{\mathfrak{X}}}})\cdot\colon\mathbb{L}_{\boldsymbol{{\mathfrak{X}}}}^{\vee}\to\mathbb{L}_{\boldsymbol{{\mathfrak{X}}}}[n]

is an equivalence.

In this paper, we are only interested in (1)(-1)-shifted symplectic structures.

Example A.2.

Let XX be a Calabi–Yau threefold, i.e., a three dimensional smooth variety with trivial canonical bundle. Then the derived moduli stack 𝕸X\boldsymbol{{\mathfrak{M}}}_{X} of compactly supported coherent sheaves on XX carries a canonical (1)(-1)-shifted symplectic structure. See [PTVV13, Theorem 0.1] and [BD, Main Theorem].

Example A.3.

Let 𝖄\boldsymbol{{\mathfrak{Y}}} be a derived Artin stack and

𝐓[n]𝖄𝐒𝐩𝐞𝐜𝖄(Sym(𝕃𝖄[n]))\mathbf{T}^{*}[n]\boldsymbol{{\mathfrak{Y}}}\coloneqq\mathop{\mathbf{Spec}}\nolimits_{\boldsymbol{{\mathfrak{Y}}}}(\operatorname{Sym}(\mathbb{L}_{\boldsymbol{{\mathfrak{Y}}}}^{\vee}[-n]))

be its nn-shifted cotangent stack. Let λ𝒜1(𝐓[n]𝖄,n)\lambda\in{\mathcal{A}}^{1}(\mathbf{T}^{*}[n]\boldsymbol{{\mathfrak{Y}}},n) be the tautological 11-form. Then it is shown in [PTVV13, Proposition 1.21] and [Cal19, Theorem 2.2] that the nn-shifted closed 22-form ddRclλd_{\mathrm{dR}}^{\,\operatorname{cl}}\lambda is shifted symplectic.

Example A.4.

Let U=SpecAU=\operatorname{Spec}A be a smooth affine scheme which admits an étale coordinate (x1,,xn)(x_{1},\ldots,x_{n}), and f:U𝔸1f\colon U\to\mathbb{A}^{1} be a regular function. Let BB be the cdga defined by the Koszul complex

B(ΩAddRfA).B\coloneqq(\cdots\to\Omega_{A}^{\vee}\xrightarrow[]{d_{\mathrm{dR}}f}A).

Then 𝐒𝐩𝐞𝐜B\mathop{\mathbf{Spec}}\nolimits B is equivalent to the derived critical locus 𝐂𝐫𝐢𝐭(f)\mathop{\mathbf{Crit}}\nolimits(f). We let yiB1y_{i}\in B^{-1} be the element of degree 1-1 corresponding to /xi\partial/\partial x_{i}. Then the (1)(-1)-shifted closed 22-form

ω(ddRx1ddRy1++ddRxnddRyn,0,0,)𝒜2,cl(𝐒𝐩𝐞𝐜B,1)\omega\coloneqq(d_{\mathrm{dR}}x_{1}\wedge d_{\mathrm{dR}}y_{1}+\cdots+d_{\mathrm{dR}}x_{n}\wedge d_{\mathrm{dR}}y_{n},0,0,\ldots)\in{\mathcal{A}}^{2,\operatorname{cl}}(\mathop{\mathbf{Spec}}\nolimits B,-1)

defines a (1)(-1)-shifted symplectic structure on 𝐂𝐫𝐢𝐭(f)\mathop{\mathbf{Crit}}\nolimits(f).

It is shown in [BBJ19, Theorem 5.18] that any (1)(-1)-shifted symplectic derived scheme is Zariski locally of this form.

Now we discuss the canonical (1)(-1)-shifted symplectic structure on the derived critical locus of a function on a general derived Artin stack. To do this, we need to recall the notion of Lagrangian structures.

Definition A.5.

Let (𝖃,ω𝖃)(\boldsymbol{{\mathfrak{X}}},\omega_{\boldsymbol{{\mathfrak{X}}}}) be an nn-shifted symplectic derived Artin stack and 𝝉:𝑳𝖃\boldsymbol{\tau}\colon\boldsymbol{L}\to\boldsymbol{{\mathfrak{X}}} be a morphism of derived Artin stacks. An isotropic structure is a path from 0 to 𝝉ω𝖃\boldsymbol{\tau}^{\star}\omega_{\boldsymbol{{\mathfrak{X}}}} in 𝒜2,cl(𝑳,n){\mathcal{A}}^{2,\operatorname{cl}}(\boldsymbol{L},n). An isotropic structure η\eta is called a Lagrangian structure if it induces an equivalence

𝕃𝑳𝕃𝝉[n1].\mathbb{L}_{\boldsymbol{L}}^{\vee}\simeq\mathbb{L}_{\boldsymbol{\tau}}[n-1].

See [PTVV13, §2.2] for the detail.

Example A.6.
  • (1)

    Let 𝖄\boldsymbol{{\mathfrak{Y}}} be a derived Artin stack and λ𝒜1(𝐓[n]𝖄,n)\lambda\in{\mathcal{A}}^{1}(\mathbf{T}^{*}[n]\boldsymbol{{\mathfrak{Y}}},n) be the tautological 11-form. Then λ|𝖄\lambda|_{\boldsymbol{{\mathfrak{Y}}}} is naturally equivalent to zero hence so is ddRclλ|𝖄d_{\mathrm{dR}}^{\,\operatorname{cl}}\lambda|_{\boldsymbol{{\mathfrak{Y}}}}. Therefore the zero section map 𝖄𝐓[n]𝖄\boldsymbol{{\mathfrak{Y}}}\to\mathbf{T}^{*}[n]\boldsymbol{{\mathfrak{Y}}} carries a natural isotropic structure. It is shown in [Cal19, Theorem 2.2] that this isotropic structure is a Lagrangian structure.

  • (2)

    Let 𝖄\boldsymbol{{\mathfrak{Y}}} and λ\lambda be as above, and take a function fΓ(𝖄,𝒪𝖄[n])f\in\Gamma(\boldsymbol{{\mathfrak{Y}}},{\mathcal{O}}_{\boldsymbol{{\mathfrak{Y}}}}[n]) of degree nn. Let ddRf¯:𝖄𝐓[n]𝖄\overline{d_{\mathrm{dR}}f}\colon\boldsymbol{{\mathfrak{Y}}}\to\mathbf{T}^{*}[n]\boldsymbol{{\mathfrak{Y}}} be the map corresponding to the section ddRfΓ(𝖄,𝕃𝖄[n])d_{\mathrm{dR}}f\in\Gamma(\boldsymbol{{\mathfrak{Y}}},\mathbb{L}_{\boldsymbol{{\mathfrak{Y}}}}[n]). Then the natural homotopy

    (ddRf¯)ddRclλddRclddRf0(\overline{d_{\mathrm{dR}}f})^{\star}d_{\mathrm{dR}}^{\,\operatorname{cl}}\lambda\sim d_{\mathrm{dR}}^{\,\operatorname{cl}}\circ d_{\mathrm{dR}}f\sim 0

    defines an isotropic structure on (ddRf¯)(\overline{d_{\mathrm{dR}}f}). It is shown in [Cal19, Theorem 2.15] that this isotropic structure is a Lagrangian structure.

Let 𝖃\boldsymbol{{\mathfrak{X}}} be an nn-shifted symplectic derived Artin stack and 𝝉1:𝑳1𝖃\boldsymbol{\tau}_{1}\colon\boldsymbol{L}_{1}\to\boldsymbol{{\mathfrak{X}}} and 𝝉2:𝑳2𝖃\boldsymbol{\tau}_{2}\colon\boldsymbol{L}_{2}\to\boldsymbol{{\mathfrak{X}}} be Lagrangians. These Lagrangian structures define a loop in 𝒜2,cl(𝑳𝟏×𝖃𝑳2,n){\mathcal{A}}^{2,\operatorname{cl}}(\boldsymbol{L_{1}}\times_{\boldsymbol{{\mathfrak{X}}}}\boldsymbol{L}_{2},n) hence a point in 𝒜2,cl(𝑳𝟏×𝖃𝑳2,n1){\mathcal{A}}^{2,\operatorname{cl}}(\boldsymbol{L_{1}}\times_{\boldsymbol{{\mathfrak{X}}}}\boldsymbol{L}_{2},n-1). It is shown in [PTVV13, Theorem 2.9] that this (n1)(n-1)-shifted closed 22-form is shifted symplectic.

Example A.7.

Let 𝖄\boldsymbol{{\mathfrak{Y}}} be a derived Artin stack and fΓ(𝖄,𝒪𝖄[n])f\in\Gamma(\boldsymbol{{\mathfrak{Y}}},{\mathcal{O}}_{\boldsymbol{{\mathfrak{Y}}}}[n]) be a function of degree nn. The derived critical locus 𝐂𝐫𝐢𝐭(f)\mathop{\mathbf{Crit}}\nolimits(f) is defined to be the intersection

𝐂𝐫𝐢𝐭(f)𝖄×0,𝐓[n]𝖄,ddRf¯𝖄.\mathop{\mathbf{Crit}}\nolimits(f)\coloneqq\boldsymbol{{\mathfrak{Y}}}\times_{0,\,\mathbf{T}^{*}[n]\boldsymbol{{\mathfrak{Y}}},\,\overline{d_{\mathrm{dR}}f}}\boldsymbol{{\mathfrak{Y}}}.

Example A.6 and the above discussion implies that 𝐂𝐫𝐢𝐭(f)\mathop{\mathbf{Crit}}\nolimits(f) carries a canonical (n1)(n-1)-shifted symplectic structure.

The (1)(-1)-shifted symplectic structure constructed in Example A.4 is a special case of the above example:

Lemma A.8.

Let f:U𝔸1f\colon U\to\mathbb{A}^{1} be a regular function on a smooth affine scheme. Assume that UU admits a global étale coordinate. Then the (1)(-1)-shifted symplectic structure on 𝐂𝐫𝐢𝐭(f)\mathop{\mathbf{Crit}}\nolimits(f) constructed in Example A.4 is equivalent to the (1)(-1)-shifted symplectic structure constructed in Example A.7.

Proof.

We write U=SpecAU=\operatorname{Spec}A and take a global étale coordinate x1,x2,,xnAx_{1},x_{2},\ldots,x_{n}\in A. We let BB be the cdga appeared in Example A.4 whose underlying graded algebra is A[y1,,yn]A[y_{1},\ldots,y_{n}]. As we have seen in Example A.4, 𝐒𝐩𝐞𝐜B\mathop{\mathbf{Spec}}\nolimits B gives a model for 𝐂𝐫𝐢𝐭(f)\mathop{\mathbf{Crit}}\nolimits(f). Consider the element

α=iyiddRxiΩB1.\alpha=\sum_{i}y_{i}\cdot d_{\mathrm{dR}}x_{i}\in\Omega_{B}^{-1}.

Then we have an identity dα=ddRfd\alpha=d_{\mathrm{dR}}f which corresponds to the natural homotopy

ddRf|𝐂𝐫𝐢𝐭(f)0d_{\mathrm{dR}}f|_{\mathop{\mathbf{Crit}}\nolimits(f)}\sim 0

in 𝒜1(𝐂𝐫𝐢𝐭(f),0)\mathcal{A}^{1}(\mathop{\mathbf{Crit}}\nolimits(f),0). Therefore the element

ddRclα=(iddRyiddRxi,0,)d_{\mathrm{dR}}^{\,\operatorname{cl}}\alpha=(\sum_{i}d_{\mathrm{dR}}y_{i}\cdot d_{\mathrm{dR}}x_{i},0,\ldots)

corresponds to the (1)(-1)-shifted symplectic structure constructed in Example A.7. ∎

Now we discuss the relation of the (1)(-1)-shifted symplectic structure and the d-critical structure. Let (𝖃,ω)(\boldsymbol{{\mathfrak{X}}},\omega) be a (1)(-1)-shifted symplectic derived Artin stack. Then it is shown in [BBBBJ15, Theorem 3.18(a)] that the classical truncation 𝔛=t0(𝖃){\mathfrak{X}}=t_{0}(\boldsymbol{{\mathfrak{X}}}) carries a natural d-critical structure ss. We now recall some of its basic properties.

Firstly assume that 𝖃\boldsymbol{{\mathfrak{X}}} is a derived scheme and write 𝑿=𝖃\boldsymbol{X}=\boldsymbol{{\mathfrak{X}}} and X=𝔛X={\mathfrak{X}}. Take an open embedding 𝜾:𝐂𝐫𝐢𝐭(f)𝑿\boldsymbol{\iota}\colon\mathop{\mathbf{Crit}}\nolimits(f)\hookrightarrow\boldsymbol{X} where ff is a regular function on a smooth scheme UU such that f|Crit(f)red=0f|_{\operatorname{Crit}(f)^{\operatorname{red}}}=0 and UU has a global étale coordinate. We equip 𝐂𝐫𝐢𝐭(f)\mathop{\mathbf{Crit}}\nolimits(f) with the (1)(-1)-shifted symplectic structure constructed in Example A.4 and assume that 𝜾\boldsymbol{\iota} preserves the (1)(-1)-shifted symplectic structures. We let RR denote the image of t0(𝜾)t_{0}(\boldsymbol{\iota}) and i:RUi\colon R\hookrightarrow U denote the natural inclusion. Then (R,U,f,i)(R,U,f,i) defines a d-critical chart of (X,s)(X,s), see [BBJ19, Theorem 6.6].

Now we remove the assumption that 𝖃\boldsymbol{{\mathfrak{X}}} is a derived scheme. Take a smooth morphism 𝒒:𝑻𝖃\boldsymbol{q}\colon\boldsymbol{T}\to\boldsymbol{{\mathfrak{X}}}. Assume that there exist a morphism between derived schemes 𝝉:𝑻𝑻^\boldsymbol{\tau}\colon\boldsymbol{T}\to\widehat{\boldsymbol{T}}, a (1)(-1)-shifted symplectic structure ω𝑻^\omega_{\widehat{\boldsymbol{T}}} on 𝑻^\widehat{\boldsymbol{T}}, and an equivalence 𝒒ω𝝉ω𝑻^\boldsymbol{q}^{\star}\omega\sim\boldsymbol{\tau}^{\star}\omega_{\widehat{\boldsymbol{T}}}. Then there exists an equality

(A.1) t0(𝒒)s=t0(𝝉)sT^t_{0}(\boldsymbol{q})^{\star}s=t_{0}(\boldsymbol{\tau})^{\star}s_{\widehat{T}}

of d-critical structures, where sT^s_{\widehat{T}} is the d-critical structure on T^=t0(𝑻^)\widehat{T}=t_{0}(\widehat{\boldsymbol{T}}) induced from the (1)(-1)-shifted symplectic structure ω𝑻^\omega_{\widehat{\boldsymbol{T}}}. See [Kin, Theorem 4.6] for the proof.

It is shown in [BBBBJ15, Theorem 3.18(b)] that there exists a natural isomorphism

(A.2) K𝔛,svirdet(𝕃𝖃|𝔛red).K_{{\mathfrak{X}},s}^{\operatorname{vir}}\cong\det(\mathbb{L}_{\boldsymbol{{\mathfrak{X}}}}|_{{\mathfrak{X}}^{\operatorname{red}}}).

A.2. Proof of Proposition 2.4

Here we give the proof of Proposition 2.4.

Lemma A.9.

Let q:U𝔘q\colon U\to{\mathfrak{U}} be a smooth morphism from a smooth scheme. We let sXs_{X} denote the canonical d-critical structure on XCrit(fq)X\coloneqq\operatorname{Crit}(f\circ q). Then there exists an equality (q|X)s=sX(q|_{X})^{\star}s=s_{X}.

Proof.

We let ω𝐂𝐫𝐢𝐭(f)\omega_{\mathop{\mathbf{Crit}}\nolimits(f)} (resp. ω𝐂𝐫𝐢𝐭(fq)\omega_{\mathop{\mathbf{Crit}}\nolimits(f\circ q)}) denote the natural (1)(-1)-shifted symplectic structure on 𝐂𝐫𝐢𝐭(f)\mathop{\mathbf{Crit}}\nolimits(f) (resp. 𝐂𝐫𝐢𝐭(fq)\mathop{\mathbf{Crit}}\nolimits(f\circ q)). Consider the following diagram of derived Artin stacks:

𝐂𝐫𝐢𝐭(f)×𝔘U\textstyle{\mathop{\mathbf{Crit}}\nolimits(f)\times_{{\mathfrak{U}}}U\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝜾\scriptstyle{\boldsymbol{\iota}}𝒒0\scriptstyle{\boldsymbol{q}_{0}}𝐂𝐫𝐢𝐭(fq)\textstyle{\mathop{\mathbf{Crit}}\nolimits(f\circ q)}𝐂𝐫𝐢𝐭(f),\textstyle{\mathop{\mathbf{Crit}}\nolimits(f),}

where the map 𝒒0\boldsymbol{q}_{0} is induced by qq and 𝜾\boldsymbol{\iota} is the natural map which is identity on the truncation. We claim that there is an equivalence of (1)(-1)-shifted symplectic structures:

(A.3) 𝒒0ω𝐂𝐫𝐢𝐭(f)𝜾ω𝐂𝐫𝐢𝐭(fq).\boldsymbol{q}_{0}^{\star}\omega_{\mathop{\mathbf{Crit}}\nolimits(f)}\sim\boldsymbol{\iota}^{\star}\omega_{\mathop{\mathbf{Crit}}\nolimits(f\circ q)}.

We have a natural homotopy

α:ddR(f)|𝐂𝐫𝐢𝐭(f)0\alpha\colon d_{\mathrm{dR}}(f)|_{\mathop{\mathbf{Crit}}\nolimits(f)}\sim 0

in 𝒜1(𝐂𝐫𝐢𝐭(f),0)\mathcal{A}^{1}(\mathop{\mathbf{Crit}}\nolimits(f),0). By definition, the symplectic form ω𝐂𝐫𝐢𝐭(f)\omega_{\mathop{\mathbf{Crit}}\nolimits(f)} corresponds to the loop

0ddRclddR(f)|𝐂𝐫𝐢𝐭(f)00\sim d_{\mathrm{dR}}^{\,\operatorname{cl}}\circ d_{\mathrm{dR}}(f)|_{\mathop{\mathbf{Crit}}\nolimits(f)}\sim 0

in 𝒜2,cl(𝐂𝐫𝐢𝐭(f),0)\mathcal{A}^{2,\operatorname{cl}}(\mathop{\mathbf{Crit}}\nolimits(f),0), where the first homotopy is defined by the equivalence ddRclddR0d_{\mathrm{dR}}^{\,\operatorname{cl}}\circ d_{\mathrm{dR}}\sim 0 and the latter homotopy is defined by ddRclαd_{\mathrm{dR}}^{\,\operatorname{cl}}\alpha. Therefore the closed (1)(-1)-shifted 22-form 𝒒0ω𝐂𝐫𝐢𝐭(f)\boldsymbol{q}_{0}^{\star}\omega_{\mathop{\mathbf{Crit}}\nolimits(f)} corresponds to the loop

0ddRclddR(fq)|𝐂𝐫𝐢𝐭(f)×𝔘U00\sim d_{\mathrm{dR}}^{\,\operatorname{cl}}\circ d_{\mathrm{dR}}(f\circ q)|_{\mathop{\mathbf{Crit}}\nolimits(f)\times_{{\mathfrak{U}}}U}\sim 0

in 𝒜2,cl(𝐂𝐫𝐢𝐭(f)×𝔘U,0)\mathcal{A}^{2,\operatorname{cl}}(\mathop{\mathbf{Crit}}\nolimits(f)\times_{{\mathfrak{U}}}U,0). A similar argument shows that 𝜾ω𝐂𝐫𝐢𝐭(fq)\boldsymbol{\iota}^{\star}\omega_{\mathop{\mathbf{Crit}}\nolimits(f\circ q)} has the same description, hence we obtain the equivalence (A.3).

Combining this equivalence and the equality (A.1), we obtain the desired equality. ∎

Lemma A.10.

There exists a natural orientation of (𝔛,s)({\mathfrak{X}},s)

o:K𝔘2|𝔛redK𝔛,svir.o\colon K_{{\mathfrak{U}}}^{\otimes{2}}|_{{\mathfrak{X}}^{\operatorname{red}}}\cong K_{{\mathfrak{X}},s}^{\operatorname{vir}}.
Proof.

Take a smooth surjective morphism q:U𝔘q\colon U\to{\mathfrak{U}} and write X=Crit(fq)X=\operatorname{Crit}(f\circ q). Consider the following composition

oq:qK𝔘2|Xred\displaystyle o_{q}\colon q^{*}K_{{\mathfrak{U}}}^{\otimes{2}}|_{X^{\operatorname{red}}} (KU2det(ΩU/𝔘)2)|Xred\displaystyle\cong(K_{U}^{\otimes{2}}\otimes\det(\Omega_{U/{\mathfrak{U}}})^{\otimes{-2}})|_{X^{\operatorname{red}}}
KX,sXdet(ΩX/𝔛)2|Xred\displaystyle\cong K_{X,s_{X}}\otimes\det(\Omega_{X/{\mathfrak{X}}})^{\otimes{-2}}|_{X^{\operatorname{red}}}
(q|Xred)K𝔛,s,\displaystyle\cong(q|_{X^{\operatorname{red}}})^{*}K_{{\mathfrak{X}},s},

where we set sX=(q|X)ss_{X}=(q|_{X})^{\star}s. We claim that the isomorphism oqo_{q} descends to an orientation for (𝔛,s)({\mathfrak{X}},s). To do this, take an étale surjective morphism η:VU×𝔘U\eta\colon V\to U\times_{{\mathfrak{U}}}U from a scheme VV. We let pri:U×𝔘U\operatorname{pr}_{i}\colon U\times_{{\mathfrak{U}}}U be the ii-th projection for i=1,2i=1,2. Write Y=Crit(fqpr1η)Y=\operatorname{Crit}(f\circ q\circ\operatorname{pr}_{1}\circ\eta) and define

oqpriη:((qpriη)K𝔘2)|Yred(qpriη|Yred)K𝔛,so_{q\circ\operatorname{pr}_{i}\circ\eta}\colon((q\circ\operatorname{pr}_{i}\circ\eta)^{*}K_{{\mathfrak{U}}}^{\otimes^{2}})|_{Y^{\operatorname{red}}}\to(q\circ\operatorname{pr}_{i}\circ\eta|_{Y^{\operatorname{red}}})^{*}K_{{\mathfrak{X}},s}

in the same manner as oqo_{q}. It is enough to prove the commutativity of the following diagram for each i=1,2i=1,2:

((qpriη)K𝔘2)|Yred\textstyle{((q\circ\operatorname{pr}_{i}\circ\eta)^{*}K_{{\mathfrak{U}}}^{\otimes^{2}})|_{Y^{\operatorname{red}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}oqpriη\scriptstyle{o_{q\circ\operatorname{pr}_{i}\circ\eta}}(qpriη|Yred)K𝔛,s\textstyle{(q\circ\operatorname{pr}_{i}\circ\eta|_{Y^{\operatorname{red}}})^{*}K_{{\mathfrak{X}},s}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(priη)red,(qK𝔘2|Xred)\textstyle{(\operatorname{pr}_{i}\circ\eta)^{\operatorname{red},*}(q^{*}K_{{\mathfrak{U}}}^{\otimes{2}}|_{X^{\operatorname{red}}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(priη)red,oq\scriptstyle{(\operatorname{pr}_{i}\circ\eta)^{\operatorname{red},*}o_{q}}(priη)red,(q|Xred)K𝔛,s.\textstyle{(\operatorname{pr}_{i}\circ\eta)^{\operatorname{red},*}(q|_{X^{\operatorname{red}}})^{*}K_{{\mathfrak{X}},s}.}

This follows from the commutativity of the diagram (2.1).

Proof of Proposition 2.4.

We keep the notation as in the proof of the previous lemma. Let oXo_{X} and oYo_{Y} be natural orientations on (X,qs)(X,q^{\star}s) and (Y,ηpr1qs)(Y,\eta^{\star}\operatorname{pr}_{1}^{\star}q^{\star}s) coming from the descriptions as global critical loci. The construction of the orientation oo in the previous lemma implies that we have the following natural commutative diagram of orientations

(A.4) oY\textstyle{o_{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ηprioX\textstyle{\eta^{\star}\operatorname{pr}_{i}^{\star}o_{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ηpriqo\textstyle{\eta^{\star}\operatorname{pr}_{i}^{\star}q^{\star}o}

for each i=1,2i=1,2.

Now define an isomorphism

θq:qφ𝔛,s,oqφf(𝔘[dim𝔘])\theta_{q}\colon q^{*}\varphi_{{\mathfrak{X}},s,o}\cong q^{*}\varphi_{f}(\mathbb{Q}_{{\mathfrak{U}}}[\dim{\mathfrak{U}}])

by the following composition

qφ𝔛,s,o\displaystyle q^{*}\varphi_{{\mathfrak{X}},s,o} φX,qs,qo[dimq]\displaystyle\cong\varphi_{X,q^{\star}s,q^{\star}o}[-\dim q]
φX,sX,oX[dimq]\displaystyle\cong\varphi_{X,s_{X},o_{X}}[-\dim q]
φfq(U[dimUdimq])\displaystyle\cong\varphi_{f\circ q}(\mathbb{Q}_{U}[\dim U-\dim q])
qφf(𝔘[dim𝔘])\displaystyle\cong q^{*}\varphi_{f}(\mathbb{Q}_{{\mathfrak{U}}}[\dim{\mathfrak{U}}])

for i=1,2i=1,2. Here the third isomorphism follows from Lemma A.8 and Example 2.3. Similarly we can define an isomorphism

θqpriη:(qpriη)φ𝔛,s,o(qpriη)φf(X[dim𝔛]).\theta_{q\circ\operatorname{pr}_{i}\circ\eta}\colon(q\circ\operatorname{pr}_{i}\circ\eta)^{*}\varphi_{{\mathfrak{X}},s,o}\cong(q\circ\operatorname{pr}_{i}\circ\eta)^{*}\varphi_{f}(\mathbb{Q}_{X}[\dim{\mathfrak{X}}]).

The commutativity of the diagrams (A.4), (2.4), and (2.3) implies an equality

(priη)θq=θqpriη({\operatorname{pr}_{i}\circ\eta})^{*}\theta_{q}=\theta_{q\circ\operatorname{pr}_{i}\circ\eta}

hence θq\theta_{q} descends to the desired isomorphism. ∎

Appendix B Proof of the support lemma

We will give the proof of the support lemma (= Proposition 5.11) here.

Fix positive integers r1,r2r_{1},r_{2} and integers m1,m2m_{1},m_{2} such that μ=m1/r1=m2/r2\mu=m_{1}/r_{1}=m_{2}/r_{2} holds. Write r=r1+r2r=r_{1}+r_{2} and m=m1+m2m=m_{1}+m_{2}. Define 𝖂𝕸X(r1,m1)×𝕸X(r2,m2)\boldsymbol{{\mathfrak{W}}}\subset\boldsymbol{{\mathfrak{M}}}_{X}(r_{1},m_{1})\times\boldsymbol{{\mathfrak{M}}}_{X}(r_{2},m_{2}) to be the substack consisting of pairs ([E],[F])([E],[F]) such that p(SuppE)p(SuppF)=p(\operatorname{Supp}E)\cap p(\operatorname{Supp}F)=\emptyset where p:X=S×𝔸1𝔸1p\colon X=S\times\mathbb{A}^{1}\to\mathbb{A}^{1} is the projection. Define a map 𝒘:𝖂𝕸X(r,m)\boldsymbol{w}\colon\boldsymbol{{\mathfrak{W}}}\to\boldsymbol{{\mathfrak{M}}}_{X}(r,m) by taking direct sum. The map 𝒘\boldsymbol{w} is an étale map.

Lemma B.1.

Let ω1\omega_{1} (resp. ω2\omega_{2}, ω\omega) be the (1)(-1)-shifted symplectic structure on 𝕸X(r1,m1)\boldsymbol{{\mathfrak{M}}}_{X}(r_{1},m_{1}) (resp. 𝕸X(r2,m2)\boldsymbol{{\mathfrak{M}}}_{X}(r_{2},m_{2}), 𝕸X(r,m)\boldsymbol{{\mathfrak{M}}}_{X}(r,m)). Then there exists an equivalence

(ω1ω2)|𝖂𝒘ω.(\omega_{1}\boxplus\omega_{2})|_{\boldsymbol{{\mathfrak{W}}}}\simeq\boldsymbol{w}^{\star}\omega.
Proof.

It follows from [BD, Corollary 6.5] that there exists a Lagrangian structure on the morphism

(𝜾,𝒘):𝖂𝕸X(r1,m1)×𝕸X(r2,m2)×𝕸X(r,m)(\boldsymbol{\iota},\boldsymbol{w})\colon\boldsymbol{{\mathfrak{W}}}\to\boldsymbol{{\mathfrak{M}}}_{X}(r_{1},m_{1})\times\boldsymbol{{\mathfrak{M}}}_{X}(r_{2},m_{2})\times\boldsymbol{{\mathfrak{M}}}_{X}(r,m)

where 𝜾\boldsymbol{\iota} is the natural inclusion and we equip 𝕸X(r1,m1)×𝕸X(r2,m2)×𝕸X(r,m)\boldsymbol{{\mathfrak{M}}}_{X}(r_{1},m_{1})\times\boldsymbol{{\mathfrak{M}}}_{X}(r_{2},m_{2})\times\boldsymbol{{\mathfrak{M}}}_{X}(r,m) with the (1)(-1)-shifted symplectic structure ω1ω2(ω)\omega_{1}\boxplus\omega_{2}\boxplus(-\omega). The Lagrangian structure induces the desired equivalence.

For an open subset UU\subset\mathbb{C} in the analytic topology, we define 𝔐Xss(r,m)U𝔐Xss(r,m){\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)^{U}\subset{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m) to be the complex analytic open substack consisting of points corresponding to sheaves whose supports are contained in S×UXS\times U\subset X. The following statement is a straightforward consequence of the above lemma.

Corollary B.2.

Let U1,U2U_{1},U_{2}\subset\mathbb{C} be disjoint open subsets in the analytic topology. Consider the following open immersion

wU1,U2:𝔐Xss(r1,m1)U1×𝔐Xss(r2,m2)U2𝔐Xss(r,m)U1U2w_{U_{1},U_{2}}\colon{\mathfrak{M}}_{X}^{\operatorname{ss}}(r_{1},m_{1})^{U_{1}}\times{\mathfrak{M}}_{X}^{\operatorname{ss}}(r_{2},m_{2})^{U_{2}}\to{\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m)^{U_{1}\coprod U_{2}}

induced from 𝐰\boldsymbol{w}. We let sis_{i} denote the d-critical structure on 𝔐Xss(ri,mi)Ui{\mathfrak{M}}_{X}^{\operatorname{ss}}(r_{i},m_{i})^{U_{i}} and ss denote the d-critical structure on the right-hand side. Then we have an equality wU1,U2(s)=s1s2w_{U_{1},U_{2}}^{\star}(s)=s_{1}\boxplus s_{2}.

We now want to prove that the map wU1,U2w_{U_{1},U_{2}} preserves the canonical orientation following the idea of the proof of Proposition 5.8. Let WW be the image of t0(𝖂)t_{0}(\boldsymbol{{\mathfrak{W}}}) along the map

𝔐Xss(r1,m1)×𝔐Xss(r2,m2)MXss(r1,m1)×MXss(r2,m2).{\mathfrak{M}}_{X}^{\operatorname{ss}}(r_{1},m_{1})\times{\mathfrak{M}}_{X}^{\operatorname{ss}}(r_{2},m_{2})\to M_{X}^{\operatorname{ss}}(r_{1},m_{1})\times M_{X}^{\operatorname{ss}}(r_{2},m_{2}).

Write (r1,m1)=(k1r0,k1m0)(r_{1},m_{1})=(k_{1}r_{0},k_{1}m_{0}) and (r2,m2)=(k2r0,k2m0)(r_{2},m_{2})=(k_{2}r_{0},k_{2}m_{0}) where (r0,m0)(r_{0},m_{0}) is coprime. Define an open subspace

𝔸WSymk1(𝔸1)×Symk2(𝔸1)\mathbb{A}_{W}\subset\operatorname{Sym}^{k_{1}}(\mathbb{A}^{1})\times\operatorname{Sym}^{k_{2}}(\mathbb{A}^{1})

consisting of configurations (P,Q)(P,Q) such that PQ=P\cap Q=\emptyset. There exists a natural map

WredMXss(r1,m1)red×MXss(r2,m2)redSymr1(𝔸1)×Symr2(𝔸1)W^{\operatorname{red}}\hookrightarrow M_{X}^{\operatorname{ss}}(r_{1},m_{1})^{\operatorname{red}}\times M_{X}^{\operatorname{ss}}(r_{2},m_{2})^{\operatorname{red}}\to\operatorname{Sym}^{r_{1}}(\mathbb{A}^{1})\times\operatorname{Sym}^{r_{2}}(\mathbb{A}^{1})

where the latter map is induced by the projection X=S×𝔸1𝔸1X=S\times\mathbb{A}^{1}\to\mathbb{A}^{1}. Note that the above map factors through the inclusion

𝔸WSymk1(𝔸1)×Symk2(𝔸1)Symr1(𝔸1)×Symr2(𝔸1)\mathbb{A}_{W}\hookrightarrow\operatorname{Sym}^{k_{1}}(\mathbb{A}^{1})\times\operatorname{Sym}^{k_{2}}(\mathbb{A}^{1})\hookrightarrow\operatorname{Sym}^{r_{1}}(\mathbb{A}^{1})\times\operatorname{Sym}^{r_{2}}(\mathbb{A}^{1})

where the latter map is the diagonal embedding. Therefore we obtain a surjective map

ηW:Wred𝔸W.\eta_{W}\colon W^{\operatorname{red}}\to\mathbb{A}_{W}.
Lemma B.3.

Let ff be an invertible regular function on WredW^{\operatorname{red}}. Then there exists a regular function gg on 𝔸W\mathbb{A}_{W} such that f=gηWf=g\circ\eta_{W} holds.

Proof.

We first claim that the function ff is constant along the reduced parts of fibers of ηW\eta_{W}. Take a configuration (P,Q)𝔸W(P,Q)\in\mathbb{A}_{W}. Write PQ={p1,pt}P\cup Q=\{p_{1},\ldots p_{t}\} and we let lil_{i} be the multiplicity of PQP\cup Q at pip_{i}. Then the fiber ηW1((P,Q))\eta_{W}^{-1}((P,Q)) is isomorphic to the scheme

i=1tMSss(lir0,lim0).\prod_{i=1}^{t}M_{S}^{\operatorname{ss}}(l_{i}r_{0},l_{i}m_{0}).

Therefore it follows from the proof of Proposition 5.8 that ff is constant along ηW1((P,Q))red\eta_{W}^{-1}((P,Q))^{\operatorname{red}}. Therefore it is enough to prove that the map ηW\eta_{W} admits a section.

Take an arbitrary stable sheaf EMSss(r0,m0)E\in M_{S}^{\operatorname{ss}}(r_{0},m_{0}). Consider the map sW:𝔸WWreds_{W}\colon\mathbb{A}_{W}\to W^{\operatorname{red}} defined by

(P={p1,pt},Q={q1,qs})(j(ipj,E)mult(pj),(j(iqj,E)mult(qj))(P=\{p_{1},\ldots p_{t}\},Q=\{q_{1},\ldots q_{s}\})\mapsto(\bigoplus_{j}(i_{p_{j},*}E)^{\oplus\mathrm{mult}(p_{j})},(\bigoplus_{j}(i_{q_{j},*}E)^{\oplus\mathrm{mult}(q_{j})})

where ipji_{p_{j}} denotes the embedding S×{pj}XS\times\{p_{j}\}\hookrightarrow X and mult(pj)\operatorname{mult}(p_{j}) denotes the multiplicity of pjp_{j} in PP, and similarly for iqji_{q_{j}} and mult(qj)\operatorname{mult}(q_{j}). We can see that sWs_{W} is a section of ηW\eta_{W}. Thus we obtain the claim.

Let s𝔚s_{{\mathfrak{W}}} be the natural d-critical structure on 𝔚=t0(𝖂){\mathfrak{W}}=t_{0}(\boldsymbol{{\mathfrak{W}}}). Let o𝔚:M12K𝔚,s𝔚viro_{{\mathfrak{W}}}\colon M_{1}^{\otimes{2}}\cong K_{{\mathfrak{W}},s_{{\mathfrak{W}}}}^{\operatorname{vir}} be the orientation on 𝔚{\mathfrak{W}} induced from the canonical orientation on 𝔐Xss(r1,m1)×𝔐Xss(r2,m2){\mathfrak{M}}_{X}^{\operatorname{ss}}(r_{1},m_{1})\times{\mathfrak{M}}_{X}^{\operatorname{ss}}(r_{2},m_{2}) and o𝔚:M22K𝔚,s𝔚viro_{{\mathfrak{W}}}^{\prime}\colon M_{2}^{\otimes{2}}\cong K_{{\mathfrak{W}},s_{{\mathfrak{W}}}}^{\operatorname{vir}} be the orientation on 𝔚{\mathfrak{W}} induced from the canonical orientation on 𝔐Xss(r,m){\mathfrak{M}}_{X}^{\operatorname{ss}}(r,m). We have seen in the proof of Proposition 5.8 that there exist trivializations M1𝒪𝔚redM_{1}\cong{\mathcal{O}}_{{\mathfrak{W}}^{\operatorname{red}}} and M2𝒪𝔚redM_{2}\cong{\mathcal{O}}_{{\mathfrak{W}}^{\operatorname{red}}}. Therefore the composition (o𝔚)1o𝔚(o_{{\mathfrak{W}}}^{\prime})^{-1}\circ o_{{\mathfrak{W}}} defines an element

αΓ(𝔚red,𝒪𝔚red)Γ(Wred,𝒪Wred).\alpha\in\Gamma({\mathfrak{W}}^{\operatorname{red}},{\mathcal{O}}_{{\mathfrak{W}}^{\operatorname{red}}})\cong\Gamma(W^{\operatorname{red}},{\mathcal{O}}_{W^{\operatorname{red}}}).
Corollary B.4.

We use the notations as in Corollary B.2. Assume that each connected component of U1U_{1} and U2U_{2} are homeomorphic to the disk. Then there exists an isomorphism of orientations

o𝔚|𝔐Xss(r1,m1)U1×𝔐Xss(r2,m2)U2o𝔚|𝔐Xss(r1,m1)U1×𝔐Xss(r2,m2)U2.o_{{\mathfrak{W}}}|_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r_{1},m_{1})^{U_{1}}\times{\mathfrak{M}}_{X}^{\operatorname{ss}}(r_{2},m_{2})^{U_{2}}}\cong o_{{\mathfrak{W}}}^{\prime}|_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(r_{1},m_{1})^{U_{1}}\times{\mathfrak{M}}_{X}^{\operatorname{ss}}(r_{2},m_{2})^{U_{2}}}.
Proof.

Let MXss(r1,m1)U1MXss(r1,m1)M_{X}^{\operatorname{ss}}(r_{1},m_{1})^{U_{1}}\subset M_{X}^{\operatorname{ss}}(r_{1},m_{1}) (resp. MXss(r2,m2)U2MXss(r2,m2)M_{X}^{\operatorname{ss}}(r_{2},m_{2})^{U_{2}}\subset M_{X}^{\operatorname{ss}}(r_{2},m_{2})) be the image of 𝔐Xss(r1,m1)U1{\mathfrak{M}}_{X}^{\operatorname{ss}}(r_{1},m_{1})^{U_{1}} (resp. 𝔐Xss(r2,m2)U2{\mathfrak{M}}_{X}^{\operatorname{ss}}(r_{2},m_{2})^{U_{2}}) along the map pXp_{X}. We need to show that α|MXss(r1,m1)U1×MXss(r2,m2)U2\alpha|_{M_{X}^{\operatorname{ss}}(r_{1},m_{1})^{U_{1}}\times M_{X}^{\operatorname{ss}}(r_{2},m_{2})^{U_{2}}} admits a square root. Lemma B.3 implies that there exists a regular function gg on 𝔸W\mathbb{A}_{W} such that α=gηW\alpha=g\circ\eta_{W} holds. Let 𝔸U1,U2𝔸W\mathbb{A}_{U_{1},U_{2}}\subset\mathbb{A}_{W} be an open subset consisting of configurations (P,Q)(P,Q) such that PU1P\subset U_{1} and QU2Q\subset U_{2}. Since the image of MXss(r1,m1)U1×MXss(r2,m2)U2M_{X}^{\operatorname{ss}}(r_{1},m_{1})^{U_{1}}\times M_{X}^{\operatorname{ss}}(r_{2},m_{2})^{U_{2}} under the map ηW\eta_{W} is contained in 𝔸U1,U2\mathbb{A}_{U_{1},U_{2}}, we need to show that g|𝔸U1,U2g|_{\mathbb{A}_{U_{1},U_{2}}} admits a square root. But this follows from the simply connectedness of 𝔸U1,U2\mathbb{A}_{U_{1},U_{2}}. ∎

Proof of Proposition 5.11.

The proof is almost identical to the proof of the support lemma for preprojective algebras [Dav16]. Take disjoint open subsets U1,U2𝔸1U_{1},U_{2}\subset\mathbb{A}^{1} whose connected components are homeomorphic to the disk. It follows from Proposition 5.10 that there exists an isomorphism

((pXφ𝔐Xss(μ))|MXss(μ)Ui)Sym(H(B)virφMXss(μ)|MXss(μ)Ui)\displaystyle{\mathcal{H}}\left(({p_{X}}_{*}\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(\mu)})|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{i}}}\right)\cong\operatorname{Sym}_{\boxtimes_{\oplus}}\left(\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})_{\operatorname{vir}}\otimes\varphi_{M_{X}^{\operatorname{ss}}(\mu)|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{i}}}}\right)

for each i=1,2i=1,2. Then Thom–Sebastiani theorem [ABB17, Remark 5.23], Corollary B.2 and Corollary B.4 imply the following isomorphism

((pXφ𝔐Xss(μ))|MXss(μ)U1U2)\displaystyle{\mathcal{H}}\left(({p_{X}}_{*}\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(\mu)})|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{1}\coprod U_{2}}}\right)
\displaystyle\cong ((pXφ𝔐Xss(μ))|MXss(μ)U1)((pXφ𝔐Xss(μ))|MXss(μ)U2)\displaystyle{\mathcal{H}}\left(({p_{X}}_{*}\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(\mu)})|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{1}}}\right)\boxtimes_{\oplus}{\mathcal{H}}\left(({p_{X}}_{*}\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(\mu)})|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{2}}}\right)
\displaystyle\cong Sym(H(B)virφMXss(μ)|MXss(μ)U1)\displaystyle\operatorname{Sym}_{\boxtimes_{\oplus}}\left(\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})_{\operatorname{vir}}\otimes\varphi_{M_{X}^{\operatorname{ss}}(\mu)|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{1}}}}\right)
Sym(H(B)virφMXss(μ)|MXss(μ)U2)\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \boxtimes_{\oplus}\operatorname{Sym}_{\boxtimes_{\oplus}}\left(\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})_{\operatorname{vir}}\otimes\varphi_{M_{X}^{\operatorname{ss}}(\mu)|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{2}}}}\right)
\displaystyle\cong Sym(H(B)vir(φMXss(μ)|MXss(μ)U1φMXss(μ)|MXss(μ)U2))\displaystyle\operatorname{Sym}_{\boxtimes_{\oplus}}\left(\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})_{\operatorname{vir}}\otimes(\varphi_{M_{X}^{\operatorname{ss}}(\mu)|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{1}}}}\oplus\varphi_{M_{X}^{\operatorname{ss}}(\mu)|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{2}}}})\right)

On the other hand, Proposition 5.10 implies an isomorphism

((pXφ𝔐Xss(μ))|MXss(μ)U1U2))Sym(H(B)virφMXss(μ)|MXss(μ)U1U2).\displaystyle{\mathcal{H}}\left(({p_{X}}_{*}\varphi_{{\mathfrak{M}}_{X}^{\operatorname{ss}}(\mu)})|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{1}\coprod U_{2}}})\right)\cong\operatorname{Sym}_{\boxtimes_{\oplus}}\left(\operatorname{H}^{*}(\mathrm{B}\mathbb{C}^{*})_{\operatorname{vir}}\otimes\varphi_{M_{X}^{\operatorname{ss}}(\mu)|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{1}\coprod U_{2}}}}\right).

Therefore we obtain an isomorphism

(B.1) φMXss(μ)|MXss(μ)U1φMXss(μ)|MXss(μ)U2φMXss(μ)|MXss(μ)U1U2.\varphi_{M_{X}^{\operatorname{ss}}(\mu)|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{1}}}}\oplus\varphi_{M_{X}^{\operatorname{ss}}(\mu)|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{2}}}}\cong\varphi_{M_{X}^{\operatorname{ss}}(\mu)|_{{M}_{X}^{\operatorname{ss}}(\mu)^{U_{1}\coprod U_{2}}}}.

Now take a point [E]Supp(φMXss(μ))[E]\in\operatorname{Supp}(\varphi_{M_{X}^{\operatorname{ss}}(\mu)}) where EE is a polystable sheaf on XX. Assume that the support of EE is contained in S×(U1U2)S\times(U_{1}\coprod U_{2}) for some disjoint open subsets U1,U2U_{1},U_{2}\subset\mathbb{C} (or equivalently, [E]MXss(μ)U1U2[E]\in M_{X}^{\operatorname{ss}}(\mu)^{U_{1}\coprod U_{2}}). Then the isomorphism (B.1) implies that the support of EE is contained in either of U1U_{1} or U2U_{2}. Therefore there exists some tt\in\mathbb{C} such that Supp(E)S×{t}\operatorname{Supp}(E)\subset S\times\{t\}, which implies the proposition.

References

  • [ABB17] L. Amorim and O. Ben-Bassat, Perversely categorified Lagrangian correspondences, Adv. Theor. Math. Phys. 21 (2017), no. 2, 289–381. MR 3678002
  • [Ach] P. N Achar, Equivariant mixed Hodge modules, https://www.math.lsu.edu/ pramod/docs/emhm.pdf.
  • [BBBBJ15] O. Ben-Bassat, C. Brav, V. Bussi, and D. Joyce, A ‘Darboux theorem’ for shifted symplectic structures on derived Artin stacks, with applications, Geom. Topol. 19 (2015), no. 3, 1287–1359. MR 3352237
  • [BBD+15] C. Brav, V. Bussi, D. Dupont, D. Joyce, and B. Szendrői, Symmetries and stabilization for sheaves of vanishing cycles, J. Singul. 11 (2015), 85–151, With an appendix by Jörg Schürmann. MR 3353002
  • [BBJ19] C. Brav, V. Bussi, and D. Joyce, A Darboux theorem for derived schemes with shifted symplectic structure, J. Amer. Math. Soc. 32 (2019), no. 2, 399–443. MR 3904157
  • [BD] C. Brav and T. Dyckerhoff, Relative Calabi-Yau structures II: Shifted Lagrangians in the moduli of objects, arXiv preprint arXiv:1812.11913.
  • [Cal19] D. Calaque, Shifted cotangent stacks are shifted symplectic, Ann. Fac. Sci. Toulouse Math. (6) 28 (2019), no. 1, 67–90. MR 3940792
  • [CDP14] W.-Y. Chuang, D.-E. Diaconescu, and G. Pan, BPS states and the P=WP=W conjecture, Moduli spaces, London Math. Soc. Lecture Note Ser., vol. 411, Cambridge Univ. Press, Cambridge, 2014, pp. 132–150. MR 3221294
  • [Cor88] K. Corlette, Flat GG-bundles with canonical metrics, J. Differential Geom. 28 (1988), no. 3, 361–382. MR 965220
  • [Dava] B. Davison, Purity and 2-Calabi-Yau categories, arXiv preprint arXiv:2106.07692.
  • [Davb] Ben Davison, BPS Lie algebras and the less perverse filtration on the preprojective CoHA, arXiv preprint arXiv:2007.03289.
  • [Davc] by same author, The integrality conjecture and the cohomology of preprojective stacks, arXiv preprint arXiv:1602.02110.
  • [Dav16] B. Davison, Cohomological Hall algebras and character varieties, Internat. J. Math. 27 (2016), no. 7, 1640003, 18. MR 3521588
  • [Dav17] by same author, The critical CoHA of a quiver with potential, Q. J. Math. 68 (2017), no. 2, 635–703. MR 3667216
  • [dCHM12] M. A. de Cataldo, T. Hausel, and L. Migliorini, Topology of Hitchin systems and Hodge theory of character varieties: the case A1A_{1}, Ann. of Math. (2) 175 (2012), no. 3, 1329–1407. MR 2912707
  • [dCM] Mark Andrea A de Cataldo and Davesh Maulik, The perverse filtration for the Hitchin fibration is locally constant, arXiv preprint arXiv:1808.02235.
  • [dCMSZ] Mark Andrea de Cataldo, Davesh Maulik, Junliang Shen, and Siqing Zhang, Cohomology of the moduli of Higgs bundles via positive characteristic, arXiv preprint arXiv:2105.03043.
  • [DHM22] Ben Davison, Lucien Hennecart, and Sebastian Schlegel Mejia, BPS Lie algebras for totally negative 2-Calabi-Yau categories and nonabelian Hodge theory for stacks, arXiv preprint arXiv:2212.07668 (2022).
  • [Dim04] Alexandru Dimca, Sheaves in topology, Universitext, Springer-Verlag, Berlin, 2004. MR 2050072
  • [DM] Ben Davison and Sven Meinhardt, Donaldson-thomas theory for categories of homological dimension one with potential, arXiv preprint arXiv:1512.08898.
  • [DM20] B. Davison and S. Meinhardt, Cohomological Donaldson-Thomas theory of a quiver with potential and quantum enveloping algebras, Invent. Math. 221 (2020), no. 3, 777–871. MR 4132957
  • [Don87] S. K. Donaldson, Twisted harmonic maps and the self-duality equations, Proc. London Math. Soc. (3) 55 (1987), no. 1, 127–131. MR 887285
  • [DP12] R. Donagi and T. Pantev, Langlands duality for Hitchin systems, Invent. Math. 189 (2012), no. 3, 653–735. MR 2957305
  • [Fel21] C. Felisetti, Intersection cohomology of the moduli space of Higgs bundles on a genus 2 curve, Journal of the Institute of Mathematics of Jussieu (2021), 1–50.
  • [FM] Camilla Felisetti and Mirko Mauri, P= W conjectures for character varieties with symplectic resolution, arXiv preprint arXiv:2006.08752.
  • [FSY] C. Felisetti, J. Shen, and Q. Yin, On intersection cohomology and Lagrangian fibrations of irreducible symplectic varieties, arXiv preprint arXiv:2108.02464.
  • [GV] R. Gopakumar and C. Vafa, M-Theory and Topological Strings II.
  • [GWZ20] M. Groechenig, D. Wyss, and P. Ziegler, Mirror symmetry for moduli spaces of Higgs bundles via p-adic integration, Invent. Math. 221 (2020), no. 2, 505–596. MR 4121158
  • [Hit87] N. J. Hitchin, The self-duality equations on a Riemann surface, Proc. London Math. Soc. (3) 55 (1987), no. 1, 59–126. MR 887284
  • [HL97] D. Huybrechts and M. Lehn, The geometry of moduli spaces of sheaves, Aspects of Mathematics, E31, Friedr. Vieweg & Sohn, Braunschweig, 1997. MR 1450870
  • [HMMS] T. Hausel, A. Mellit, A. Minets, and O. Schiffmann, P=W via H2H_{2}, arXiv preprint arXiv:2209.05429.
  • [HST01] S. Hosono, M. Saito, and A. Takahashi, Relative Lefschetz action and BPS state counting, Internat. Math. Res. Notices (2001), no. 15, 783–816. MR 1849482
  • [Joy15] D. Joyce, A classical model for derived critical loci, Journal of Differential Geometry 101 (2015), no. 2, 289–367.
  • [JS12] D. Joyce and Y. Song, A theory of generalized Donaldson-Thomas invariants, Mem. Amer. Math. Soc. 217 (2012), no. 1020, iv+199. MR 2951762
  • [Kin] T. Kinjo, Dimensional reduction in cohomological Donaldson-Thomas theory, arXiv preprint arXiv:2102.01568, To appear in Compos. Math.
  • [KL] Young-Hoon Kiem and Jun Li, Categorification of Donaldson-Thomas invariants via perverse sheaves, arXiv preprint arXiv:1212.6444.
  • [KM21] Tasuki Kinjo and Naruki Masuda, Global critical chart for local Calabi-Yau threefolds, arXiv preprint arXiv:2112.10052 (2021).
  • [Kol96] J. Kollár, Rational curves on algebraic varieties, Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics], vol. 32, Springer-Verlag, Berlin, 1996. MR 1440180
  • [KS13] M. Kashiwara and P. Schapira, Sheaves on Manifolds: With a short history.«les débuts de la théorie des faisceaux». by christian houzel, vol. 292, Springer Science & Business Media, 2013.
  • [Lur09] J. Lurie, Higher topos theory, Annals of Mathematics Studies, vol. 170, Princeton University Press, Princeton, NJ, 2009. MR 2522659
  • [Mau] Mirko Mauri, Intersection cohomology of rank two character varieties of surface groups, arXiv preprint arXiv:2101.04628.
  • [Mei] S. Meinhardt, Donaldson-Thomas invariants vs. intersection cohomology for categories of homological dimension one, arXiv preprint arXiv:1512.03343.
  • [Mel20] A. Mellit, Poincaré polynomials of moduli spaces of Higgs bundles and character varieties (no punctures), Invent. Math. 221 (2020), no. 1, 301–327. MR 4105090
  • [MSa] D. Maulik and J. Shen, Cohomological χ\chi-independence for moduli of one-dimensional sheaves and moduli of Higgs bundles, arXiv preprint arXiv:2012.06627.
  • [MSb] by same author, The P=W conjecture for GLn\mathrm{GL}_{n}, arXiv preprint arXiv:2209.02568.
  • [MT18] D. Maulik and Y. Toda, Gopakumar–Vafa invariants via vanishing cycles, Inventiones mathematicae 213 (2018), no. 3, 1017–1097.
  • [PTVV13] T. Pantev, B. Toën, M. Vaquié, and G. Vezzosi, Shifted symplectic structures, Publ. Math. Inst. Hautes Études Sci. 117 (2013), 271–328. MR 3090262
  • [Ray18] S. Rayan, Aspects of the topology and combinatorics of Higgs bundle moduli spaces, SIGMA Symmetry Integrability Geom. Methods Appl. 14 (2018), Paper No. 129, 18. MR 3884746
  • [Sai89] M. Saito, Introduction to mixed Hodge modules.
  • [Sai90] Morihiko Saito, Mixed hodge modules, Publications of the Research Institute for Mathematical Sciences 26 (1990), no. 2, 221–333.
  • [Sim92] C. T. Simpson, Higgs bundles and local systems, Inst. Hautes Études Sci. Publ. Math. (1992), no. 75, 5–95. MR 1179076
  • [Sim94] by same author, Moduli of representations of the fundamental group of a smooth projective variety. II, Inst. Hautes Études Sci. Publ. Math. (1994), no. 80, 5–79 (1995). MR 1320603
  • [SS20] F. Sala and O. Schiffmann, Cohomological Hall algebra of Higgs sheaves on a curve, Algebr. Geom. 7 (2020), no. 3, 346–376. MR 4087863
  • [Sun17] Shenghao Sun, Generic base change, Artin’s comparison theorem, and the decomposition theorem for complex Artin stacks, Journal of Algebraic Geometry 26 (2017), no. 3, 513–555.
  • [Toda] Y. Toda, Categorical Donaldson-Thomas theory for local surfaces, arXiv preprint arXiv:1907.09076.
  • [Todb] by same author, Gopakumar-Vafa invariants and wall-crossing, arXiv preprint arXiv:1710.01843.
  • [Yu] Hongjie Yu, Comptage des systemes locaux \ell-adiques sur une courbe, arXiv preprint arXiv:1807.04659.
  • [Yua] Yao Yuan, Sheaves on non-reduced curves in a projective surface, arXiv preprint arXiv:2111.10071.